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:

Intraflagellar transport dynein is autoinhibited by trapping of its mechanical and track-binding elements

Abstract

Cilia are multifunctional organelles that are constructed using intraflagellar transport (IFT) of cargo to and from their tip. It is widely held that the retrograde IFT motor, dynein-2, must be controlled in order to reach the ciliary tip and then unleashed to power the return journey. However, the mechanism is unknown. Here, we systematically define the mechanochemistry of human dynein-2 motors as monomers, dimers, and multimotor assemblies with kinesin-II. Combining these data with insights from single-particle EM, we discover that dynein-2 dimers are intrinsically autoinhibited. Inhibition is mediated by trapping dynein-2's mechanical 'linker' and 'stalk' domains within a novel motor–motor interface. We find that linker-mediated inhibition enables efficient transport of dynein-2 by kinesin-II in vitro. These results suggest a conserved mechanism for autoregulation among dimeric dyneins, which is exploited as a switch for dynein-2's recycling activity during IFT.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Intraflagellar transport motors and constructs used in this study.
Figure 2: Monomeric dynein-2 motor domains power fast microtubule gliding.
Figure 3: Dimerization inhibits dynein-2 motor domains.
Figure 4: Dynein-2's linker and stalk are trapped within a novel motor–motor interface.
Figure 5: Untrapping dynein-2 dimers rescues their motility.
Figure 6: Assembly and motility of dynein-2 and kinesin Kif3 in multimotor arrays.
Figure 7: Model for dynein-2 regulation during IFT.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

Referenced accessions

Protein Data Bank

References

  1. Rosenbaum, J.L. & Witman, G.B. Intraflagellar transport. Nat. Rev. Mol. Cell Biol. 3, 813–825 (2002).

    Article  CAS  PubMed  Google Scholar 

  2. Lechtreck, K.F. IFT-Cargo interactions and protein transport in cilia. Trends Biochem. Sci. 40, 765–778 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Fliegauf, M., Benzing, T. & Omran, H. When cilia go bad: cilia defects and ciliopathies. Nat. Rev. Mol. Cell Biol. 8, 880–893 (2007).

    Article  CAS  PubMed  Google Scholar 

  4. Ishikawa, H. & Marshall, W.F. Ciliogenesis: building the cell's antenna. Nat. Rev. Mol. Cell Biol. 12, 222–234 (2011).

    Article  CAS  PubMed  Google Scholar 

  5. Kozminski, K.G., Johnson, K.A., Forscher, P. & Rosenbaum, J.L. A motility in the eukaryotic flagellum unrelated to flagellar beating. Proc. Natl. Acad. Sci. USA 90, 5519–5523 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Johnson, K.A. & Rosenbaum, J.L. Polarity of flagellar assembly in Chlamydomonas. J. Cell Biol. 119, 1605–1611 (1992).

    Article  CAS  PubMed  Google Scholar 

  7. Hao, L. et al. Intraflagellar transport delivers tubulin isotypes to sensory cilium middle and distal segments. Nat. Cell Biol. 13, 790–798 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Bhogaraju, S. et al. Molecular basis of tubulin transport within the cilium by IFT74 and IFT81. Science 341, 1009–1012 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Kubo, T. et al. Together, the IFT81 and IFT74 N-termini form the main module for intraflagellar transport of tubulin. J. Cell Sci. 129, 2106–2119 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Scholey, J.M. Kinesin-2: a family of heterotrimeric and homodimeric motors with diverse intracellular transport functions. Annu. Rev. Cell Dev. Biol. 29, 443–469 (2013).

    Article  CAS  PubMed  Google Scholar 

  11. Andreasson, J.O., Shastry, S., Hancock, W.O. & Block, S.M. The mechanochemical cycle of mammalian kinesin-2 KIF3A/B under load. Curr. Biol. 25, 1166–1175 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Albracht, C.D., Guzik-Lendrum, S., Rayment, I. & Gilbert, S.P. Heterodimerization of kinesin-2 KIF3AB modulates entry into the processive run. J. Biol. Chem. 291, 23248–23256 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Prevo, B., Mangeol, P., Oswald, F., Scholey, J.M. & Peterman, E.J. Functional differentiation of cooperating kinesin-2 motors orchestrates cargo import and transport in C. elegans cilia. Nat. Cell Biol. 17, 1536–1545 (2015).

    Article  CAS  PubMed  Google Scholar 

  14. Gibbons, B.H., Asai, D.J., Tang, W.J., Hays, T.S. & Gibbons, I.R. Phylogeny and expression of axonemal and cytoplasmic dynein genes in sea urchins. Mol. Biol. Cell 5, 57–70 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Pazour, G.J., Wilkerson, C.G. & Witman, G.B. A dynein light chain is essential for the retrograde particle movement of intraflagellar transport (IFT). J. Cell Biol. 141, 979–992 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Pazour, G.J., Dickert, B.L. & Witman, G.B. The DHC1b (DHC2) isoform of cytoplasmic dynein is required for flagellar assembly. J. Cell Biol. 144, 473–481 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Porter, M.E., Bower, R., Knott, J.A., Byrd, P. & Dentler, W. Cytoplasmic dynein heavy chain 1b is required for flagellar assembly in Chlamydomonas. Mol. Biol. Cell 10, 693–712 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Signor, D. et al. Role of a class DHC1b dynein in retrograde transport of IFT motors and IFT raft particles along cilia, but not dendrites, in chemosensory neurons of living Caenorhabditis elegans . J. Cell Biol. 147, 519–530 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Hou, Y. & Witman, G.B. Dynein and intraflagellar transport. Exp. Cell Res. 334, 26–34 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Roberts, A.J., Kon, T., Knight, P.J., Sutoh, K. & Burgess, S.A. Functions and mechanics of dynein motor proteins. Nat. Rev. Mol. Cell Biol. 14, 713–726 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Carter, A.P., Diamant, A.G. & Urnavicius, L. How dynein and dynactin transport cargos: a structural perspective. Curr. Opin. Struct. Biol. 37, 62–70 (2016).

    Article  CAS  PubMed  Google Scholar 

  22. Cianfrocco, M.A., DeSantis, M.E., Leschziner, A.E. & Reck-Peterson, S.L. Mechanism and regulation of cytoplasmic dynein. Annu. Rev. Cell Dev. Biol. 31, 83–108 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Patel-King, R.S., Gilberti, R.M., Hom, E.F. & King, S.M. WD60/FAP163 is a dynein intermediate chain required for retrograde intraflagellar transport in cilia. Mol. Biol. Cell 24, 2668–2677 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Asante, D., Stevenson, N.L. & Stephens, D.J. Subunit composition of the human cytoplasmic dynein-2 complex. J. Cell Sci. 127, 4774–4787 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Schmidts, M. et al. TCTEX1D2 mutations underlie Jeune asphyxiating thoracic dystrophy with impaired retrograde intraflagellar transport. Nat. Commun. 6, 7074 (2015).

    Article  CAS  PubMed  Google Scholar 

  26. Mencarelli, C., Mitchell, A., Leoncini, R., Rosenbaum, J. & Lupetti, P. Isolation of intraflagellar transport trains. Cytoskeleton 70, 439–452 (2013).

    Article  CAS  PubMed  Google Scholar 

  27. Stepanek, L. & Pigino, G. Microtubule doublets are double-track railways for intraflagellar transport trains. Science 352, 721–724 (2016).

    Article  CAS  PubMed  Google Scholar 

  28. Hancock, W.O. Bidirectional cargo transport: moving beyond tug of war. Nat. Rev. Mol. Cell Biol. 15, 615–628 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Williams, C.L. et al. Direct evidence for BBSome-associated intraflagellar transport reveals distinct properties of native mammalian cilia. Nat. Commun. 5, 5813 2014).

    Article  CAS  PubMed  Google Scholar 

  30. Ichikawa, M., Watanabe, Y., Murayama, T. & Toyoshima, Y.Y. Recombinant human cytoplasmic dynein heavy chain 1 and 2: observation of dynein-2 motor activity in vitro. FEBS Lett. 585, 2419–2423 (2011).

    Article  CAS  PubMed  Google Scholar 

  31. Schmidt, H., Zalyte, R., Urnavicius, L. & Carter, A.P. Structure of human cytoplasmic dynein-2 primed for its power stroke. Nature 518, 435–438 (2015).

    Article  CAS  PubMed  Google Scholar 

  32. Burgess, S.A., Walker, M.L., Sakakibara, H., Knight, P.J. & Oiwa, K. Dynein structure and power stroke. Nature 421, 715–718 (2003).

    Article  CAS  PubMed  Google Scholar 

  33. Kon, T., Mogami, T., Ohkura, R., Nishiura, M. & Sutoh, K. ATP hydrolysis cycle-dependent tail motions in cytoplasmic dynein. Nat. Struct. Mol. Biol. 12, 513–519 (2005).

    Article  CAS  PubMed  Google Scholar 

  34. Roberts, A.J. et al. AAA+ ring and linker swing mechanism in the dynein motor. Cell 136, 485–495 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Kon, T. et al. The 2.8 Å crystal structure of the dynein motor domain. Nature 484, 345–350 (2012).

    Article  CAS  PubMed  Google Scholar 

  36. Schmidt, H., Gleave, E.S. & Carter, A.P. Insights into dynein motor domain function from a 3.3-Å crystal structure. Nat. Struct. Mol. Biol. 19, 492–497, S1 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Roberts, A.J. et al. ATP-driven remodeling of the linker domain in the dynein motor. Structure 20, 1670–1680 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Torisawa, T. et al. Autoinhibition and cooperative activation mechanisms of cytoplasmic dynein. Nat. Cell Biol. 16, 1118–1124 (2014).

    Article  CAS  PubMed  Google Scholar 

  39. Reck-Peterson, S.L. et al. Single-molecule analysis of dynein processivity and stepping behavior. Cell 126, 335–348 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Shima, T., Imamula, K., Kon, T., Ohkura, R. & Sutoh, K. Head-head coordination is required for the processive motion of cytoplasmic dynein, an AAA+ molecular motor. J. Struct. Biol. 156, 182–189 (2006).

    Article  CAS  PubMed  Google Scholar 

  41. Qiu, W. et al. Dynein achieves processive motion using both stochastic and coordinated stepping. Nat. Struct. Mol. Biol. 19, 193–200 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Trokter, M., Mücke, N. & Surrey, T. Reconstitution of the human cytoplasmic dynein complex. Proc. Natl. Acad. Sci. USA 109, 20895–20900 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Imai, H. et al. Direct observation shows superposition and large scale flexibility within cytoplasmic dynein motors moving along microtubules. Nat. Commun. 6, 8179 (2015).

    Article  CAS  PubMed  Google Scholar 

  44. Nicholas, M.P. et al. Control of cytoplasmic dynein force production and processivity by its C-terminal domain. Nat. Commun. 6, 6206 (2015).

    Article  CAS  PubMed  Google Scholar 

  45. McKenney, R.J., Huynh, W., Tanenbaum, M.E., Bhabha, G. & Vale, R.D. Activation of cytoplasmic dynein motility by dynactin-cargo adapter complexes. Science 345, 337–341 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Schlager, M.A., Hoang, H.T., Urnavicius, L., Bullock, S.L. & Carter, A.P. In vitro reconstitution of a highly processive recombinant human dynein complex. EMBO J. 33, 1855–1868 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Amos, L.A. Brain dynein crossbridges microtubules into bundles. J. Cell Sci. 93, 19–28 (1989).

    CAS  PubMed  Google Scholar 

  48. Gibbons, I.R. et al. The affinity of the dynein microtubule-binding domain is modulated by the conformation of its coiled-coil stalk. J. Biol. Chem. 280, 23960–23965 (2005).

    Article  CAS  PubMed  Google Scholar 

  49. Kon, T. et al. Helix sliding in the stalk coiled coil of dynein couples ATPase and microtubule binding. Nat. Struct. Mol. Biol. 16, 325–333 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Cole, D.G. et al. Chlamydomonas kinesin-II-dependent intraflagellar transport (IFT): IFT particles contain proteins required for ciliary assembly in Caenorhabditis elegans sensory neurons. J. Cell Biol. 141, 993–1008 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Piperno, G. & Mead, K. Transport of a novel complex in the cytoplasmic matrix of Chlamydomonas flagella. Proc. Natl. Acad. Sci. USA 94, 4457–4462 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Taschner, M. & Lorentzen, E. The intraflagellar transport machinery. Cold Spring Harb. Perspect. Biol. 8, a028092 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Taschner, M. et al. Intraflagellar transport proteins 172, 80, 57, 54, 38, and 20 form a stable tubulin-binding IFT-B2 complex. EMBO J. 35, 773–790 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Derr, N.D. et al. Tug-of-war in motor protein ensembles revealed with a programmable DNA origami scaffold. Science 338, 662–665 (2012).

    Article  CAS  PubMed  Google Scholar 

  55. Engel, B.D., Ludington, W.B. & Marshall, W.F. Intraflagellar transport particle size scales inversely with flagellar length: revisiting the balance-point length control model. J. Cell Biol. 187, 81–89 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Shih, S.M. et al. Intraflagellar transport drives flagellar surface motility. eLife 2, e00744 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Li, W., Yi, P. & Ou, G. Somatic CRISPR-Cas9-induced mutations reveal roles of embryonically essential dynein chains in Caenorhabditis elegans cilia. J. Cell Biol. 208, 683–692 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Verhey, K.J. & Hammond, J.W. Traffic control: regulation of kinesin motors. Nat. Rev. Mol. Cell Biol. 10, 765–777 (2009).

    Article  CAS  PubMed  Google Scholar 

  59. Belyy, V. et al. The mammalian dynein-dynactin complex is a strong opponent to kinesin in a tug-of-war competition. Nat. Cell Biol. 18, 1018–1024 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Yu, I., Garnham, C.P. & Roll-Mecak, A. Writing and reading the tubulin code. J. Biol. Chem. 290, 17163–17172 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Brunnbauer, M. et al. Regulation of a heterodimeric kinesin-2 through an unprocessive motor domain that is turned processive by its partner. Proc. Natl. Acad. Sci. USA 107, 10460–10465 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Imanishi, M., Endres, N.F., Gennerich, A. & Vale, R.D. Autoinhibition regulates the motility of the C. elegans intraflagellar transport motor OSM-3. J. Cell Biol. 174, 931–937 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Liang, Y. et al. FLA8/KIF3B phosphorylation regulates kinesin-II interaction with IFT-B to control IFT entry and turnaround. Dev. Cell 30, 585–597 (2014).

    Article  CAS  PubMed  Google Scholar 

  64. Pedersen, L.B., Geimer, S. & Rosenbaum, J.L. Dissecting the molecular mechanisms of intraflagellar transport in Chlamydomonas . Curr. Biol. 16, 450–459 (2006).

    Article  CAS  PubMed  Google Scholar 

  65. Urnavicius, L. et al. The structure of the dynactin complex and its interaction with dynein. Science 347, 1441–1446 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Chowdhury, S., Ketcham, S.A., Schroer, T.A. & Lander, G.C. Structural organization of the dynein-dynactin complex bound to microtubules. Nat. Struct. Mol. Biol. 22, 345–347 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Goodman, B.S. & Reck-Peterson, S.L. Engineering defined motor ensembles with DNA origami. Methods Enzymol. 540, 169–188 (2014).

    Article  CAS  PubMed  Google Scholar 

  68. Castoldi, M. & Popov, A.V. Purification of brain tubulin through two cycles of polymerization-depolymerization in a high-molarity buffer. Protein Expr. Purif. 32, 83–88 (2003).

    Article  CAS  PubMed  Google Scholar 

  69. Hyman, A. et al. Preparation of modified tubulins. Methods Enzymol. 196, 478–485 (1991).

    Article  CAS  PubMed  Google Scholar 

  70. Kon, T., Nishiura, M., Ohkura, R., Toyoshima, Y.Y. & Sutoh, K. Distinct functions of nucleotide-binding/hydrolysis sites in the four AAA modules of cytoplasmic dynein. Biochemistry 43, 11266–11274 (2004).

    Article  CAS  PubMed  Google Scholar 

  71. Cho, C., Reck-Peterson, S.L. & Vale, R.D. Regulatory ATPase sites of cytoplasmic dynein affect processivity and force generation. J. Biol. Chem. 283, 25839–25845 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  73. Uyeda, T.Q., Kron, S.J. & Spudich, J.A. Myosin step size: estimation from slow sliding movement of actin over low densities of heavy meromyosin. J. Mol. Biol. 214, 699–710 (1990).

    Article  CAS  PubMed  Google Scholar 

  74. Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer Science & Business Media, 2009).

  75. Scheres, S.H. RELION: implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol. 180, 519–530 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Mindell, J.A. & Grigorieff, N. Accurate determination of local defocus and specimen tilt in electron microscopy. J. Struct. Biol. 142, 334–347 (2003).

    Article  PubMed  Google Scholar 

  77. van Heel, M., Harauz, G., Orlova, E.V., Schmidt, R. & Schatz, M. A new generation of the IMAGIC image processing system. J. Struct. Biol. 116, 17–24 (1996).

    Article  CAS  PubMed  Google Scholar 

  78. Ludtke, S.J., Baldwin, P.R. & Chiu, W. EMAN: semiautomated software for high-resolution single-particle reconstructions. J. Struct. Biol. 128, 82–97 (1999).

    Article  CAS  PubMed  Google Scholar 

  79. Frank, J. et al. SPIDER and WEB: processing and visualization of images in 3D electron microscopy and related fields. J. Struct. Biol. 116, 190–199 (1996).

    Article  CAS  PubMed  Google Scholar 

  80. Pettersen, E.F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank C. Moores, G. Zanetti and A. Osborne for critical comments on the manuscript; M. Williams, S. Reck-Peterson and members of the Birkbeck EM group for advice; S. Nofal, S. Miah, and L. Stejskal for initial experiments; and A. Carter (MRC-LMB, Cambridge) for plasmids. This work was supported by a Sir Henry Dale Fellowship to A.J.R. from the Wellcome Trust and Royal Society [104196/Z/14/Z].

Author information

Authors and Affiliations

Authors

Contributions

K.T. and A.J.R. conducted biochemical and TIRF experiments. K.T. performed electron microscopy experiments. M.M., K.T., and A.J.R. generated and purified constructs. K.T. and A.J.R. analyzed data and wrote the paper.

Corresponding author

Correspondence to Anthony J Roberts.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Influence of microtubule length and buffer conditions on dynein-2 microtubule gliding.

(a) Kymographs of microtubule gliding at 4 nM Dyn2motor. Microtubule lengths are shown below. (b) Plot of gliding velocity as a function of microtubule length at 4 nM input concentration of Dyn2motor. N = 63 microtubules. Microtubule length and velocity are correlated with a Spearman coefficient of 0.56 (p < 0.0001). (c) Plot of mean microtubule gliding velocity (± s.d.) at different dynein-2 motor input concentrations in buffer lacking 50 mM KCl. Slower microtubule gliding velocities were seen for all three dynein-2 constructs in the absence of KCl (compare with Fig. 2e, 3d and 5e). Number of microtubules analyzed per concentration as follows: Dyn2motor 0.2 nM (32), 0.5 nM (46), 2 nM (47), 5 nM (43), 20 nM (48), 200 nM (90), GST-Dyn2motor 0.2 nM (70), 2 nM (32), 10 nM (42), 20 nM (51), 100 nM (46), 200 nM (51), GST-Dyn2(DQR)motor 0.2 nM (35), 1 nM (56), 2 nM (52), 10 nM (51), 20 nM (52), 200 nM (58). Fitted values (± standard error of the fit): Dyn2motor Vmax = 219.3 ± 3.3 nm/s, f = 0.2 ± 0.01, GST-Dyn2motor Vmax = 168.0 ± 2.1 nm/s, f = 0.1 ± 0.004, GST-Dyn2(DQR)motor Vmax = 259.0 ± 2.7 nm/s, f = 0.3 ± 0.01.

Supplementary Figure 2 Impact of nucleotide, salt, and mutagenesis on dynein-2 dimer architecture.

(a) EM micrographs GST-Dyn2motor dimers in different nucleotide and buffer conditions. High salt, 500 mM KCl. GST-Dyn2motor domains appear stacked in ADP.Vi and separated in other conditions. GST-Dyn2motor domains are also stacked in ATP conditions (micrograph shown in Fig. 3c). Arrow, particularly well-resolved GST-Dyn2motor dimer in the separated configuration in which stalks are visible in the raw micrograph. (b) EM micrograph GST-Dyn2(DQR)motor dimers in ATP. (c) Histograms showing the distribution of 2D motor–motor distances in EM class averages of GST-Dyn2motor and GST-Dyn2(DQR)motor. Total number of molecules analyzed: GST-Dyn2motor ATP (917), ADP.Vi (1160), no-nucleotide (890), ADP (853), ATP High salt (846), GST-Dyn2(DQR)motor ATP (891). The GST-Dyn2motor distributions in ATP and ADP.Vi conditions show a peak at low motor–motor separation indicative of stacking. Stacking is largely abolished in GST-Dyn2(DQR)motor.

Supplementary Figure 3 Linker-mediated stacking model and consistency with dynein-1 and dynein-2 data.

(a) Upper panel: Model of the autoinhibited state of dynein-2, derived from PDB 4RH7 (Schmidt, H. et al., Nature. 518, 435-8, 2015). The linker domains (magenta) are trapped in the motor–motor interface, while the C-terminal domains (CTDs) are on the periphery of the dimer and do not interact. All motor–motor interfaces are labeled. Lower panel: previous model of the cytoplasmic cytoplasmic dynein-1 phi particle (Torisawa, T. et al., Nature Cell Biology. 16, 1118-24, 2014), derived from PDB 3VKG (Kon, T. et al., Nature. 484, 345-50, 2012). The motor domains interact via their CTDs and the linker domains are free to move on the periphery of the dimer. (b) Enlarged views of dynein-2 interfaces in the new model. Amino acids involved in inter-motor domain interactions are labeled. (c) Top row, EM averages of GST-Dyn2motor in ATP and ADP.Vi conditions. Bottom row, highest scoring projections of the new model based on cross correlation. GST-Dyn2motor shows preferred orientations on the EM support, which are matched by the new model with Euler angles shown (ϕ, θ; SPIDER convention). GST/SNAPf density is labeled with an arrowhead in the EM average and is absent in projections. (d) Equivalent analysis carried out for human cytoplasmic dynein-1 holoenzyme in the phi particle configuration. The tail and associated subunits of the holoenzyme are labeled with an arrow in the EM average. Scale bar, 10 nm. The linker-mediated stacking model is consistent with class averages of both dynein-1 and dynein-2.

Supplementary Figure 4 DQR mutations have little or no impact on dynein-2 monomer activity.

(a) Sequence diagrams of dimeric GST-Dyn2(DQR)motor, and the Dyn2(DQR)motor construct used to assess if the DQR mutations impact activity in the context of a dynein-2 monomer. (b) Size-exclusion chromatograms. Dyn2(DQR)motor was normalized to the peak value of GST-Dyn2(DQR)motor. (c) Velocity of microtubule gliding at 1 and 20 nM concentrations of Dyn2motor and Dyn2(DQR)motor. Black and pink lines show mean ± s.d. Number of microtubules analyzed per concentration: Dyn2motor 1 nM (39), 20 nM (56), Dyn2(DQR)motor 1 nM (29), 20 nM (49). (d) Microtubule-stimulated ATPase activity of Dyn2(DQR)motor (Dyn2motor values from Fig. 2f shown in gray for comparison). Experiments were carried out in triplicate, mean values ± s.d. are shown. Fitted values (± standard error of the fit): kcat = 4.7 ± 0.3 s-1, kbasal = 1.6 ± 0.1 s-1, Km(MT) = 8.0 ± 2.2 μM.

Supplementary Figure 5 Purification and motility of Kif3 Δtail.

(a) Sequence diagrams of Kif3 and Kif3 Δtail, in which mutations that prevent autoinhibition are introduced and putatively disordered C-terminal regions are removed. Yellow box, SNAPf tag. (b) Size-exclusion chromatogram of Kif3 Δtail and schematic of the construct. (c) SDS-PAGE of Kif3 ∆tail after the final purification step. (d) Plot of mean microtubule gliding velocity (± s.d.) at different Kif3 Δtail input concentrations. Number of microtubules analyzed per concentration: 0.5 nM (50), 0.7 nM (48), 2 nM (43), 5 nM (48), 20 nM (50), 60 nM (47). Fitted values (± standard error of the fit): Vmax = 537.0 ± 2.0 nm/s, f = 0.99 ± 0.002. (e) Kymograph showing single-molecule motility of Kif3 Δtail labeled with Alexa647 via its SNAPf tag. (+) and (–) indicate microtubule polarity. (f) Velocity histogram of Kif3 Δtail single molecules (N = 311 molecules).

Supplementary Figure 6 Assembly and motility of synthetic trains bound with dynein-2.

(a) Gel shift assay showing migration of DNA origami chassis samples with three (3x) or seven (7x) attachment sites and no dynein (‘none’), GST-Dyn2motor (‘WT’), or GST-Dyn2(DQR)motor (‘DQR’) bound. (b,c) Kymographs showing behavior of DNA origami assemblies with (b) seven GST-Dyn2motor and (c) seven GST-Dyn2(DQR)motor sites. (+) and (–) indicate microtubule polarity.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 (PDF 1380 kb)

Supplementary Data : Model of dynein-2 in its autoinhibited state.

Coordinates for the linker-stacking model of dynein-2 autoinhibition, derived from PDB 4RH7. (TXT 3609 kb)

Impact of motor concentration on dynein-2 microtubule gliding.

Microtubule gliding powered by Dyn2motor at different concentrations. See also Fig. 2d. Videos are shown at 24X real time. (MOV 4884 kb)

Dynein-2 dimer architecture in different nucleotide conditions.

Class averages of GST-Dyn2motor in no nucleotide and 1 mM ATP conditions. See also Fig. 4b. Videos show 38 class averages in each condition, looped 4 times. (MOV 3134 kb)

Pairs of dynein-2 motor domains in crystallo match the architecture of isolated dimers.

Analysis of PDB 4RH7 crystal lattice reveals pairs of monomeric dynein-2 motor domains matching the stacked architecture of Dyn2motor dimers observed by single-particle electron microscopy. (MOV 8736 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Toropova, K., Mladenov, M. & Roberts, A. Intraflagellar transport dynein is autoinhibited by trapping of its mechanical and track-binding elements. Nat Struct Mol Biol 24, 461–468 (2017). https://doi.org/10.1038/nsmb.3391

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.3391

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