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.

  • Letter
  • Published:

The mammalian dynein–dynactin complex is a strong opponent to kinesin in a tug-of-war competition

Abstract

Kinesin and dynein motors transport intracellular cargos bidirectionally by pulling them in opposite directions along microtubules, through a process frequently described as a ‘tug of war’1. While kinesin produces 6 pN of force, mammalian dynein was found to be a surprisingly weak motor (0.5–1.5 pN) in vitro, suggesting that many dyneins are required to counteract the pull of a single kinesin2. Mammalian dynein’s association with dynactin and Bicaudal-D2 (BICD2) activates its processive motility3,4,5,6, but it was unknown how this affects dynein’s force output. Here, we show that formation of the dynein–dynactin–BICD2 (DDB) complex increases human dynein’s force production to 4.3 pN. An in vitro tug-of-war assay revealed that a single DDB successfully resists a single kinesin. Contrary to previous reports, the clustering of many dyneins is not required to win the tug of war. Our work reveals the key role of dynactin and a cargo adaptor protein in shifting the balance of forces between dynein and kinesin motors during intracellular transport.

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

Figure 1: Activation of human dynein motility by dynactin, BICD2N and artificial cargos.
Figure 2: Effect of dynactin and BICD2N on dynein force production.
Figure 3: Processive motility of DDB complexes is driven by single dynein dimers.
Figure 4: In vitro competition experiment between human dynein and human kinesin.

Similar content being viewed by others

References

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

    Article  CAS  Google Scholar 

  2. Kunwar, A. et al. Mechanical stochastic tug-of-war models cannot explain bidirectional lipid-droplet transport. Proc. Natl Acad. Sci. USA 108, 18960–18965 (2011).

    Article  CAS  Google Scholar 

  3. 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  Google Scholar 

  4. 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  Google Scholar 

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

    Article  CAS  Google Scholar 

  6. Splinter, D. et al. BICD2, dynactin, and LIS1 cooperate in regulating dynein recruitment to cellular structures. Mol. Biol. Cell 23, 4226–4241 (2012).

    Article  CAS  Google Scholar 

  7. 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  Google Scholar 

  8. 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  Google Scholar 

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

    Article  CAS  Google Scholar 

  10. Moughamian, A. J., Osborn, G. E., Lazarus, J. E., Maday, S. & Holzbaur, E. L. F. Ordered recruitment of dynactin to the microtubule plus-end is required for efficient initiation of retrograde axonal transport. J. Neurosci. 33, 13190–13203 (2013).

    Article  CAS  Google Scholar 

  11. Wang, Z. & Sheetz, M. P. One-dimensional diffusion on microtubules of particles coated with cytoplasmic dynein and immunoglobulins. Cell Struct. Funct. 24, 373–383 (1999).

    Article  CAS  Google Scholar 

  12. Ayloo, S. et al. Dynactin functions as both a dynamic tether and brake during dynein-driven motility. Nat. Commun. 5, 4807 (2014).

    Article  CAS  Google Scholar 

  13. McKenney, R. J., Vershinin, M., Kunwar, A., Vallee, R. B. & Gross, S. P. LIS1 and NudE induce a persistent dynein force-producing state. Cell 141, 304–314 (2010).

    Article  CAS  Google Scholar 

  14. Rai, A. K., Ramaiya, A. J., Jha, R. & Mallik, R. Molecular adaptations allow dynein to generate large collective forces inside cells. Cell 152, 172–182 (2013).

    Article  CAS  Google Scholar 

  15. Mallik, R., Carter, B. C., Lex, S. A., King, S. J. & Gross, S. P. Cytoplasmic dynein functions as a gear in response to load. Nature 427, 649–652 (2004).

    Article  CAS  Google Scholar 

  16. Tripathy, S. K. et al. Autoregulatory mechanism for dynactin control of processive and diffusive dynein transport. Nat. Cell Biol. 16, 1192–1201 (2014).

    Article  CAS  Google Scholar 

  17. Ori-McKenney, K. M., Xu, J., Gross, S. P. & Vallee, R. B. A cytoplasmic dynein tail mutation impairs motor processivity. Nat. Cell Biol. 12, 1228–1234 (2010).

    Article  CAS  Google Scholar 

  18. Visscher, K., Schnitzer, M. J. & Block, S. M. Single kinesin molecules studied with a molecular force clamp. Nature 400, 184–189 (1999).

    Article  CAS  Google Scholar 

  19. Rai, A. et al. Dynein clusters into lipid microdomains on phagosomes to drive rapid transport toward lysosomes. Cell 164, 722–734 (2016).

    Article  CAS  Google Scholar 

  20. Hendricks, A. G. et al. Motor coordination via a tug-of-war mechanism drives bidirectional vesicle transport. Curr. Biol. 20, 697–702 (2010).

    Article  CAS  Google Scholar 

  21. Thirumurugan, K., Sakamoto, T., Hammer, J. A. III, Sellers, J. R. & Knight, P. J. The cargo-binding domain regulates structure and activity of myosin 5. Nature 442, 212–215 (2006).

    Article  CAS  Google Scholar 

  22. Kaan, H. Y. K., Hackney, D. D. & Kozielski, F. The structure of the kinesin-1 motor-tail complex reveals the mechanism of autoinhibition. Science 333, 883–885 (2011).

    Article  CAS  Google Scholar 

  23. King, S. J. & Schroer, T. Dynactin increases the processivity of the cytoplasmic dynein motor. Nat. Cell Biol. 2, 20–24 (2000).

    Article  CAS  Google Scholar 

  24. 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  Google Scholar 

  25. 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  Google Scholar 

  26. DeWitt, M. A., Chang, A. Y., Combs, P. A. & Yildiz, A. Cytoplasmic dynein moves through uncoordinated stepping of the AAA + ring domains. Science 335, 221–225 (2012).

    Article  CAS  Google Scholar 

  27. Cleary, F. B. et al. Tension on the linker gates the ATP-dependent release of dynein from microtubules. Nat. Commun. 5, 4587 (2014).

    Article  CAS  Google Scholar 

  28. Belyy, V., Hendel, N. L., Chien, A. & Yildiz, A. Cytoplasmic dynein transports cargos via load-sharing between the heads. Nat. Commun. 5, 5544 (2014).

    Article  CAS  Google Scholar 

  29. Gennerich, A., Carter, A. P., Reck-Peterson, S. L. & Vale, R. D. Force-induced bidirectional stepping of cytoplasmic dynein. Cell 131, 952–965 (2007).

    Article  CAS  Google Scholar 

  30. Svoboda, K. & Block, S. M. Force and velocity measured for single kinesin molecules. Cell 77, 773–784 (1994).

    Article  CAS  Google Scholar 

  31. 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  Google Scholar 

  32. Diehl, M. R., Zhang, K., Lee, H. J. & Tirrell, D. A. Engineering cooperativity in biomotor-protein assemblies. Science 311, 1468–1471 (2006).

    Article  CAS  Google Scholar 

  33. Fu, M. M. & Holzbaur, E. L. F. JIP1 regulates the directionality of APP axonal transport by coordinating kinesin and dynein motors. J. Cell Biol. 202, 495–508 (2013).

    Article  CAS  Google Scholar 

  34. Soppina, V., Rai, A. K., Ramaiya, A. J., Barak, P. & Mallik, R. Tug-of-war between dissimilar teams of microtubule motors regulates transport and fission of endosomes. Proc. Natl Acad. Sci. USA 106, 19381–19386 (2009).

    Article  CAS  Google Scholar 

  35. Leidel, C., Longoria, R. a., Gutierrez, F. M. & Shubeita, G. T. Measuring molecular motor forces in vivo: implications for tug-of-war models of bidirectional transport. Biophys. J. 103, 492–500 (2012).

    Article  CAS  Google Scholar 

  36. Shubeita, G. T. et al. Consequences of motor copy number on the intracellular transport of kinesin-1-driven lipid droplets. Cell 135, 1098–1107 (2008).

    Article  CAS  Google Scholar 

  37. Fu, M. & Holzbaur, E. L. F. Integrated regulation of motor-driven organelle transport by scaffolding proteins. Trends Cell Biol. 24, 564–574 (2014).

    Article  CAS  Google Scholar 

  38. Klumpp, S. & Lipowsky, R. Cooperative cargo transport by several molecular motors. Proc. Natl Acad. Sci. USA 102, 17284–17289 (2005).

    Article  CAS  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  Google Scholar 

  40. 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  Google Scholar 

Download references

Acknowledgements

We are grateful to S. Can, L. S. Ferro, A. Chien and J. Bandaria for helpful discussions. This work was supported by the NIH (GM094522 (A.Y.)), an NSF CAREER Award (MCB-1055017 (A.Y.)), the Medical Research Council, UK (MC_UP_A025_1011 (A.P.C.)), a Wellcome Trust New Investigator Award (WT100387) and an EMBO Young Investigator Award (both to A.P.C.), a Marie Curie Intra European Fellowship (M.A.S.), and an NSF Graduate Research Fellowship (DGE 1106400 (V.B.)).

Author information

Authors and Affiliations

Authors

Contributions

V.B., A.P.C. and A.Y. designed the study. M.A.S. and H.F. prepared the constructs and purified the protein. V.B. and A.E.R. performed the optical trapping experiments. V.B. labelled dynein and kinesin with DNA, performed the fluorescence motility experiments, and analysed the data. V.B., A.P.C. and A.Y. wrote the manuscript.

Corresponding author

Correspondence to Ahmet Yildiz.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Temporally color-coded projections of free dyneins and DDB complexes on microtubules.

(a) (Top) Interactions of single free dyneins with microtubules are visualized by color-coding a 2D projection of Supplementary Video 1. Scale bar: 5 μm. (Bottom) Fluorescence kymograph along one of the three visible axonemes in Supplementary Video 1. Motile dyneins walked at 79 ± 11 nm s−1 (s.e.m.,n = 15). At 1 nM motor, the density of fluorescent spots on the axoneme was 0.19 per micron length of axoneme per minute, n = 93 (b). (Top) Processive motility of DDB complexes on microtubules is visualized by color-coding a 2D projection of Supplementary Video 2. (Bottom) Fluorescence kymograph along the axoneme in Supplementary Video 3. At 1 nM motor, the density of fluorescent spots on the axoneme was 0.66 per micron length of axoneme per minute, n = 98.

Supplementary Figure 2 Individual steps in trapping traces and Gaussian mixture model fitting of stall distributions.

The kinesin (a) and dynein (b) stall traces are zoomed-in looks on the traces shown in Fig. 2e and Fig. 2c, respectively. Red arrowheads represent the detachment of the motor from the microtubule after the stall. (c,d) Comparison of 2-component and 3-component Gaussian mixture model fits to the DDB stall force distribution shown in Fig. 2e. The fitting is carried out on the underlying distribution and is thus independent of the binning of the histogram. Blue curves are the scaled results of 2-component (c) and 3-component (d) Gaussian fits. Numbers represent center values ± standard errors of each peak. The Bayes Information Criterion (BIC) is lower for the 2-Gaussian fit, indicating that adding a third Gaussian peak is not statistically warranted.

Supplementary Figure 3 Stall force measurements of full-length yeast cytoplasmic dynein.

(a) Full-length yeast cytoplasmic dynein with an N-terminal GFP (GFP-Dyn471kD-DHA) is attached to an optically trapped bead via anti-GFP antibody. (b) Trace showing a typical stall of a bead driven by full-length yeast dynein in a fixed trap assay. Red arrowhead represents the detachment of the motor from the microtubule after the stall. (c) The histogram of observed stalls reveals the mean stall force (mean ± s.e.m.,n = 21 stalls from 5 beads).

Supplementary Figure 4 Examples of dynein-kinesin colocalizers in the presence of dynactin and BICD2N.

Dynein is tagged with Alexa647 (red) and kinesin is tagged with TMR (cyan). The red and cyan channels are laterally offset by five pixels to enhance the visibility of the colocalizers. Individual channels are shown in grayscale. All microtubules are arranged with the plus ends facing towards the right. Scale bars are identical in all images.

Supplementary Figure 5 Full images of all gels shown in main figures.

(a) Denaturing SDS–PAGE gel of purified dynein fractions, stained with Instant Blue. Bands corresponding to all dynein subunits can be observed. (b) Denaturing SDS–PAGE gel of purified BICD2N, stained with Instant Blue. (c) Denaturing SDS-PAGE gel of kinesin labeled with DNA. Excess DNA is removed from labeled kinesin by microtubule bind and release (MTBR). The MTBR fractions shown in lanes 10 through 14 are Pre (kinesin pre-mixed with microtubules), SN1 (supernatant from the first spin), P1 (pellet from the first spin), SN2 (supernatant from the second spin), and P2 (pellet from the second spin). (d,e) Denaturing SDS-PAGE gel of dynein labeled with DNA and Alexa647. Dynein was first labeled with ssDNA-BG-GLA-NHS at the indicated molar ratios, then labeled with an excess of Alexa647-BG-GLA-NHS. (d) Silver stain reveals the total amount of dynein, regardless of what fraction of it is labeled with ssDNA. (e) When imaged specifically for Alexa647 fluorescence, the intensities of individual bands on the same gel reveal the fraction of dyneins whose SNAPf sites are occupied by ssDNA and therefore not accessible to labeling by Alexa647-BG-GLA-NHS.

Supplementary information

Supplementary Information

Supplementary Information (PDF 880 kb)

Human dynein exhibits minimal motility in the absence of tail-binding proteins.

GFP-labeled recombinant dynein motors interacting with surface-immobilized axonemes in the presence of 2 mM ATP. Scale bar: 5 μm. (AVI 8996 kb)

Addition of dynactin and BICD2N to dynein leads to fast processive movement.

GFP-labeled dynein motors walking on surface-immobilized axonemes in the presence of a 5× molar excess of dynactin, 2× molar excess of BICD2N, and 2 mM ATP. Scale bar: 5 μm. (AVI 8996 kb)

Fast motility is activated by the binding of large cargos to dynein’s tail.

Bright-field recording of dynein-driven motility of a 200 nm polystyrene bead on a surface-immobilized axoneme in the presence of 2 mM ATP. The position of the motile bead is marked with a yellow arrow for clarity. Scale bar: 5 μm. (AVI 8996 kb)

Motility of a dynein-kinesin colocalizer in the absence of dynactin and BICD2N.

Two-color TIRF imaging of kinesin-TMR (shown in cyan) and dynein-Alexa647 (shown in red) on a microtubule. Colocalizers are marked with white arrows. (AVI 2354 kb)

Motility of a dynein-kinesin colocalizer in the presence of dynactin and BICD2N.

Two-color TIRF imaging of kinesin-TMR (shown in cyan) and dynein-Alexa647 (shown in red) on a microtubule. The colocalizer is marked with white arrows. (AVI 769 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Belyy, V., Schlager, M., Foster, H. 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). https://doi.org/10.1038/ncb3393

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3393

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