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:

CLASPs link focal-adhesion-associated microtubule capture to localized exocytosis and adhesion site turnover

Abstract

Turnover of integrin-based focal adhesions (FAs) with the extracellular matrix (ECM) is essential for coordinated cell movement. In collectively migrating human keratinocytes, FAs assemble near the leading edge, grow and mature as a result of contractile forces and disassemble underneath the advancing cell body. We report that clustering of microtubule-associated CLASP1 and CLASP2 proteins around FAs temporally correlates with FA turnover. CLASPs and LL5β (also known as PHLDB2), which recruits CLASPs to FAs, facilitate FA disassembly. CLASPs are further required for FA-associated ECM degradation, and matrix metalloprotease inhibition slows FA disassembly similarly to CLASP or PHLDB2 (LL5β) depletion. Finally, CLASP-mediated microtubule tethering at FAs establishes an FA-directed transport pathway for delivery, docking and localized fusion of exocytic vesicles near FAs. We propose that CLASPs couple microtubule organization, vesicle transport and cell interactions with the ECM, establishing a local secretion pathway that facilitates FA turnover by severing cell–matrix connections.

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: Mature FAs recruit CLASP2-decorated microtubules.
Figure 2: CLASPs facilitate FA disassembly in migrating epithelial cells.
Figure 3: CLASP clusters around FAs do not depend on microtubules.
Figure 4: LL5β is required for CLASP-mediated FA turnover.
Figure 5: CLASPS are required for FA-associated ECM degradation.
Figure 6: MT1–MMP dynamics in migrating epithelial cells.
Figure 7: Targeting of RAB6A-mediated exocytosis to FAs depends on CLASPs.

Similar content being viewed by others

References

  1. Parsons, J. T., Horwitz, A. R. & Schwartz, M. A. Cell adhesion: integrating cytoskeletal dynamics and cellular tension. Nat. Rev. Mol Cell Biol. 11, 633–643 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Gardel, M. L., Schneider, I. C., Aratyn-Schaus, Y. & Waterman, C. M. Mechanical integration of actin and adhesion dynamics in cell migration . Annu. Rev. Cell Dev. Biol. 26, 315–333 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Stehbens, S. & Wittmann, T. Targeting and transport: How microtubules control focal adhesion dynamics. J. Cell Biol. 198, 481–489 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Kaverina, I., Krylyshkina, O. & Small, J. V. Microtubule targeting of substrate contacts promotes their relaxation and dissociation. J. Cell Biol. 146, 1033–1044 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Ezratty, E. J., Partridge, M. A. & Gundersen, G. G. Microtubule-induced focal adhesion disassembly is mediated by dynamin and focal adhesion kinase. Nat. Cell Biol. 7, 581–590 (2005).

    Article  CAS  PubMed  Google Scholar 

  6. Ezratty, E. J., Bertaux, C., Marcantonio, E. E. & Gundersen, G. G. Clathrin mediates integrin endocytosis for focal adhesion disassembly in migrating cells. J. Cell Biol. 187, 733–747 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Rooney, C. et al. The Rac activator STEF (Tiam2) regulates cell migration by microtubule-mediated focal adhesion disassembly. EMBO Rep. 11, 292–298 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kumar, P. & Wittmann, T. +TIPs: SxIPping along microtubule ends. Trends Cell Biol. 22, 418–428 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Matsumoto, S., Fumoto, K., Okamoto, T., Kaibuchi, K. & Kikuchi, A. Binding of APC and dishevelled mediates Wnt5a-regulated focal adhesion dynamics in migrating cells. EMBO J. 29, 1192–1204 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Wu, X., Kodama, A. & Fuchs, E. ACF7 regulates cytoskeletal-focal adhesion dynamics and migration and has ATPase activity. Cell 135, 137–148 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Kumar, P. et al. GSK3beta phosphorylation modulates CLASP-microtubule association and lamella microtubule attachment. J. Cell Biol. 184, 895–908 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Mimori-Kiyosue, Y. et al. CLASP1 and CLASP2 bind to EB1 and regulate microtubule plus-end dynamics at the cell cortex. J. Cell Biol. 168, 141–153 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Mimori-Kiyosue, Y. et al. Mammalian CLASPs are required for mitotic spindle organization and kinetochore alignment. Genes Cells 11, 845–857 (2006).

    Article  CAS  PubMed  Google Scholar 

  14. Miller, P. M. et al. Golgi-derived CLASP-dependent microtubules control Golgi organization and polarized trafficking in motile cells. Nat. Cell Biol. 11, 1069–1080 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Paszek, M. J. et al. Scanning angle interference microscopy reveals cell dynamics at the nanoscale. Nat. Methods 9, 825–827 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Krylyshkina, O. et al. Nanometer targeting of microtubules to focal adhesions. J. Cell Biol. 161, 853–859 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Kanchanawong, P. et al. Nanoscale architecture of integrin-based cell adhesions. Nature 468, 580–584 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Meenderink, L. M. et al. P130Cas Src-binding and substrate domains have distinct roles in sustaining focal adhesion disassembly and promoting cell migration. PLoS ONE 5, e13412 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Webb, D. J. et al. FAK-Src signalling through paxillin, ERK and MLCK regulates adhesion disassembly. Nat. Cell Biol. 6, 154–161 (2004).

    Article  CAS  PubMed  Google Scholar 

  20. Stehbens, S. & Wittmann, T. Analysis of focal adhesion turnover: A quantitative live cell imaging example. Methods Cell Biol. (2014)http://dx.doi.org/10.1016/B978-0-12-420138-5.00018-5

  21. Chang, Y. C., Nalbant, P., Birkenfeld, J., Chang, Z. F. & Bokoch, G. M. GEF-H1 couples nocodazole-induced microtubule disassembly to cell contractility via RhoA. Mol Biol. Cell 19, 2147–2153 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Katoh, K., Kano, Y., Amano, M., Kaibuchi, K. & Fujiwara, K. Stress fiber organization regulated by MLCK and Rho-kinase in cultured human fibroblasts. Am. J. Physiol. Cell Physiol. 280, C1669–C1679 (2001).

    Article  CAS  PubMed  Google Scholar 

  23. Lansbergen, G. et al. CLASPs attach microtubule plus ends to the cell cortex through a complex with LL5beta. Dev Cell 11, 21–32 (2006).

    Article  CAS  PubMed  Google Scholar 

  24. Wang, Y. & McNiven, M. A. Invasive matrix degradation at focal adhesions occurs via protease recruitment by a FAK-p130Cas complex. J. Cell Biol. 196, 375–385 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Massimi, P., Zori, P., Roberts, S. & Banks, L. Differential regulation of cell-cell contact, invasion and anoikis by hScrib and hDlg in keratinocytes. PLoS ONE 7, e40279 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Davies, B., Brown, P. D., East, N., Crimmin, M. J. & Balkwill, F. R. A synthetic matrix metalloproteinase inhibitor decreases tumor burden and prolongs survival of mice bearing human ovarian carcinoma xenografts . Cancer Res. 53, 2087–2091 (1993).

    CAS  PubMed  Google Scholar 

  27. Chao, W. T. & Kunz, J. Focal adhesion disassembly requires clathrin-dependent endocytosis of integrins. FEBS Lett. 583, 1337–1343 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Yu, X. et al. N-WASP coordinates the delivery and F-actin-mediated capture of MT1–MMP at invasive pseudopods. J. Cell Biol. 199, 527–544 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Grigoriev, I. et al. Rab6, Rab8, and MICAL3 cooperate in controlling docking and fusion of exocytotic carriers. Curr. Biol. 21, 967–974 (2011).

    Article  CAS  PubMed  Google Scholar 

  30. Grigoriev, I. et al. Rab6 regulates transport and targeting of exocytotic carriers. Dev Cell 13, 305–314 (2007).

    Article  CAS  PubMed  Google Scholar 

  31. White, J. et al. Rab6 coordinates a novel Golgi to ER retrograde transport pathway in live cells. J. Cell Biol. 147, 743–760 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Toomre, D., Steyer, J. A., Keller, P., Almers, W. & Simons, K. Fusion of constitutive membrane traffic with the cell surface observed by evanescent wave microscopy. J. Cell Biol. 149, 33–40 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Paranavitane, V., Coadwell, W. J., Eguinoa, A., Hawkins, P. T. & Stephens, L. LL5beta is a phosphatidylinositol (3,4,5)-trisphosphate sensor that can bind the cytoskeletal adaptor, gamma-filamin. J. Biol. Chem. 278, 1328–1335 (2003).

    Article  CAS  PubMed  Google Scholar 

  34. Takabayashi, T. et al. LL5beta directs the translocation of filamin A and SHIP2 to sites of phosphatidylinositol 3,4,5-triphosphate (PtdIns(3,4,5)P3) accumulation, and PtdIns(3,4,5)P3 localization is mutually modified by co-recruited SHIP2. J. Biol. Chem. 285, 16155–16165 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Nakamura, F., Stossel, T. P. & Hartwig, J. H. The filamins: Organizers of cell structure and function. Cell Adh Migr 5, 160–169 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Carter, W. G., Wayner, E. A., Bouchard, T. S. & Kaur, P. The role of integrins alpha 2 beta 1 and alpha 3 beta 1 in cell-cell and cell-substrate adhesion of human epidermal cells. J. Cell Biol. 110, 1387–1404 (1990).

    Article  CAS  PubMed  Google Scholar 

  37. Hotta, A. et al. Laminin-based cell adhesion anchors microtubule plus ends to the epithelial cell basal cortex through LL5alpha/beta. J. Cell Biol. 189, 901–917 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Franco, S. J. & Huttenlocher, A. Regulating cell migration: Calpains make the cut. J. Cell Sci. 118, 3829–3838 (2005).

    Article  CAS  PubMed  Google Scholar 

  39. Storr, S. J., Carragher, N. O., Frame, M. C., Parr, T. & Martin, S. G. The calpain system and cancer. Nat. Rev. Cancer 11, 364–374 (2011).

    Article  CAS  PubMed  Google Scholar 

  40. Takino, T., Saeki, H., Miyamori, H., Kudo, T. & Sato, H. Inhibition of membrane-type 1 matrix metalloproteinase at cell–matrix adhesions. Cancer Res. 67, 11621–11629 (2007).

    Article  CAS  PubMed  Google Scholar 

  41. Gonzalo, P. et al. MT1–MMP is required for myeloid cell fusion via regulation of Rac1 signaling. Dev. Cell 18, 77–89 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Shi, F. & Sottile, J. MT1–MMP regulates the turnover and endocytosis of extracellular matrix fibronectin. J. Cell Sci. 124, 4039–4050 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Overall, C. M. & Dean, R. A. Degradomics: Systems biology of the protease web Pleiotropic roles of MMPs in cancer. Cancer Metastasis Rev. 25, 69–75 (2006).

    Article  PubMed  Google Scholar 

  44. Mu, D. et al. The integrin α(v)β8 mediates epithelial homeostasis through MT1–MMP-dependent activation of TGF-beta1. J. Cell Biol. 157, 493–507 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Chan, K. M. et al. MT1–MMP inactivates ADAM9 to regulate FGFR2 signaling and calvarial osteogenesis. Dev Cell 22, 1176–1190 (2012).

    Article  CAS  PubMed  Google Scholar 

  46. Rowe, R. G. & Weiss, S. J. Breaching the basement membrane: Who, when and how? Trends Cell Biol. 18, 560–574 (2008).

    Article  CAS  PubMed  Google Scholar 

  47. Friedl, P. & Wolf, K. Tube travel: The role of proteases in individual and collective cancer cell invasion. Cancer Res 68, 7247–7249 (2008).

    Article  CAS  PubMed  Google Scholar 

  48. Bravo-Cordero, J. J. et al. MT1–MMP proinvasive activity is regulated by a novel Rab8-dependent exocytic pathway. EMBO J. 26, 1499–1510 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Wiesner, C., Azzouzi, K. E. & Linder, S. A specific subset of RabGTPases controls cell surface exposure of MT1–MMP, extracellular matrix degradation and 3D invasion of macrophages. J. Cell Sci. 126, 2820–2833 (2013).

    Article  CAS  PubMed  Google Scholar 

  50. Monteiro, P. et al. Endosomal WASH and exocyst complexes control exocytosis of MT1–MMP at invadopodia. J. Cell Biol. 203, 1063–1079 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Suozzi, K. C., Wu, X. & Fuchs, E. Spectraplakins: Master orchestrators of cytoskeletal dynamics. J. Cell Biol. 197, 465–475 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Burgo, A. et al. A molecular network for the transport of the TI-VAMP/VAMP7 vesicles from cell center to periphery. Dev Cell 23, 166–180 (2012).

    Article  CAS  PubMed  Google Scholar 

  53. Gierke, S. & Wittmann, T. EB1-recruited microtubule +TIP complexes coordinate protrusion dynamics during 3D epithelial remodeling. Curr. Biol. 22, 753–762 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Boukamp, P. et al. Normal keratinization in a spontaneously immortalized aneuploid human keratinocyte cell line. J. Cell Biol. 106, 761–771 (1988).

    Article  CAS  PubMed  Google Scholar 

  55. Zhao, M. et al. Assembly and initial characterization of a panel of 85 genomically validated cell lines from diverse head and neck tumor sites. Clin Cancer Res. 17, 7248–7264 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Stehbens, S., Pemble, H., Murrow, L. & Wittmann, T. Imaging intracellular protein dynamics by spinning disk confocal microscopy. Methods Enzymol 504, 293–313 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Gierke, S., Kumar, P. & Wittmann, T. Analysis of microtubule polymerization dynamics in live cells. Methods Cell Biol. 97, 15–33 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Hu, K., Ji, L., Applegate, K. T., Danuser, G. & Waterman-Storer, C. M. Differential transmission of actin motion within focal adhesions. Science 315, 111–115 (2007).

    Article  CAS  PubMed  Google Scholar 

  59. Matanis, T. et al. Bicaudal-D regulates COPI-independent Golgi-ER transport by recruiting the dynein-dynactin motor complex. Nat. Cell Biol. 4, 986–992 (2002).

    Article  CAS  PubMed  Google Scholar 

  60. Bryant, D. M. et al. A molecular network for de novo generation of the apical surface and lumen. Nat. Cell Biol. 12, 1035–1045 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Moffat, J. et al. A lentiviral RNAi library for human and mouse genes applied to an arrayed viral high-content screen. Cell 124, 1283–1298 (2006).

    Article  CAS  PubMed  Google Scholar 

  62. Thompson, O. et al. Dystroglycan, Tks5 and Src mediated assembly of podosomes in myoblasts. PLoS ONE 3, e3638 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by National Institutes of Health grant R01 GM079139 to T.W., and American Heart Association postdoctoral fellowship 10POST3870021 to S.J.S. Research was conducted in a facility constructed with support from the Research Facilities Improvement Program grant C06 RR16490, and on a microscope system funded by shared equipment grant S10 RR26758 from the National Center for Research Resources of the National Institutes of Health. We also thank C. O’Connell at Nikon for imaging on the Nikon structured illumination super-resolution microscopy system, B. Webb for help with the gelatin degradation assay, A. Akhmanova, D. Bryant, K. Hu, M. McNiven and J. Norman for reagents, and D. Barber, D. Bryant, A. Yap and members of the Wittmann and Barber laboratories for discussions and comments on the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

S.J.S. and T.W. conceived the project. S.J.S., M.P., H.P., A.E., S.G. and T.W. generated reagents, conducted experiments and analysed data. M.P. contributed SAIM data and analysis. S.J.S. and T.W. wrote the manuscript and assembled figures.

Corresponding author

Correspondence to Torsten Wittmann.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Validation and phenotypes of CLASP depletion.

(a) Representative images of Golgi fragmentation in CLASP-depleted cells stained for GM130 to identify the Golgi apparatus, and Sytox orange as a nuclear stain. (b) Cytoskeleton phenotypes of CLASP depletion. Peripheral microtubules are sparser and disorganized. F-actin stress fibres anchoring into FAs are more predominant in CLASP-depleted cells. (c) Immunofluorescence of phospho-myosin (ppMLC) and E-cadherin indicate that CLASP depletion increases cell contractility. (d) Representative images of migrating HaCaT cells at the edge of a cell monolayer stained for paxillin (magenta) and F-actin (green), expressing control (non-targeting), CLASP1 or CLASP2 shRNA. Only one of the two shRNA sequences used for each CLASP isoform is shown. Single channels of the indicated regions are displayed with inverted contrast. (e) Immunoblot of lysates from cells expressing different shRNA constructs after 7 days of puromycin selection. Tubulin was used as loading control. Blots were probed with isoform-specific antibodies to either CLASP1 or CLASP2. Uncropped blots are shown in Supplementary Fig. 8. (f) Quantification of the degree of Golgi fragmentation around the nucleus in HaCaT cells migrating at the edge of a cell monolayer. The Golgi fragmentation angle α was defined as the angle through the center of the nucleus that encompasses all Golgi structures. n = 73 (control shRNA); 35 (CLASP1 shRNA 32); 47 (CLASP1 shRNA 33); 65 (CLASP2 shRNA 55); 43 (CLASP2 shRNA 58) cells. Representative data set of three experiments. (g) Quantification of FA size in CLASP-depleted cells. Each data point is the average FA size from one image which contained 3–5 cells. n = 28 (control shRNA); 32 (CLASP1 shRNA 32); 35 (CLASP1 shRNA 33); 27 (CLASP2 shRNA 55); 34 (CLASP2 shRNA 58) images. Representative data set of three experiments. Box-and-whisker plots show median, first and third quartile (box), and 95% confidence intervals (notches) with whiskers extending to the furthest observations within ±1.5 times the interquartile range. Dots are individual data points.

Supplementary Figure 2 CLASPs are required for directional migration.

(a) Example phase images of HaCaT cells migrating at the edge of a cell monolayer expressing control, CLASP1 or CLASP2 shRNAs. (b) Plots of representative cell migration paths in control and CLASP-depleted cells from 20 cells per condition. Data is representative of 3 independent experiments. Phase contrast time lapse sequences were rotated such that the wound edge was aligned with the vertical image axis, and cells are migrating to the right. The positions of dark features in the nucleus (nucleoli) were tracked over time with the ‘Time Measurements’ function in NIS Elements using ‘Auto Detect ROI’. Migration paths were normalized to the starting position.

Supplementary Figure 3 Validation and phenotypes of LL5β depletion.

(a) Localization of endogenous LL5β (green) around paxillin-labeled FAs (magenta) in control HaCaT cells, and cells treated with the indicated drugs. Insets show only the LL5β channel of the indicated regions with inverted contrast. (b) Immunoblot of lysates from cells expressing different shRNA constructs after 7 days of puromycin selection. Tubulin and GAPDH were used as loading controls. Uncropped blots are shown in Supplementary Fig. 8. (c) Immunofluorescence of LL5β and paxillin, in wound-edge HaCaT cells expressing control (non-targeting), or the indicated LL5β shRNAs. (d) Immunofluorescence of CLASP2, in HaCaT cells expressing control or the indicated LL5β shRNAs. LL5β-depletion abolishes peripheral CLASP accumulations. (e) Immunofluorecence of paxillin and LL5β in HaCaT cells expressing control or CLASP2 shRNA. CLASP-depletion in HaCaT cells does not affect LL5β localization to adhesion sites. (f) Immunofluorescence of tubulin or (g) F-actin (phalloidin) in HaCaT cells expressing control or LL5β shRNA. Cytoskeleton phenotypes of LL5β depletion are qualitatively similar to CLASP-depletion. Peripheral microtubules are sparser and disorganized. F-actin stress-fibers anchoring into FAs are more predominant in LL5β-depleted cells.

Supplementary Figure 4 CLASPs do not co-localize with endocytic vesicles.

(a) Nocodazole washout experiment in cells expressing EGFP–CLASP2. Cells were pre-treated for 90 min with 3.3 μM nocodazole prior to washout, and immunostained for endocytic markers clathrin heavy chain (HC) or dynamin II. Images were acquired by spinning disc confocal microscopy. No obvious colocalization is observed between CLASP2 and either clathrin HC or dynamin II. (b) Nocodazole washout experiment in cells immunostained for clathrin HC and paxillin. Cells were pre-treated for 90 min with 3.3 μM nocodazole prior to washout. Samples were imaged using total internal reflection microscopy (TIRF).

Supplementary Figure 5 MT1–MMP–mCherry–TM and ER dynamics in migrating HaCaT cells.

(a) Diagram of fluorescently tagged MT1–MMP constructs used. SP, signal peptide; HPX, hemopexin domain (b) Spinning disk confocal microscopy of a migrating HaCaT epithelial cell at the edge of a cell monolayer expressing MT1–MMP–mCherry–TM. Red arrowhead indicates the evolution of a bright, luminal labelled vesicle through macropinocytosis indicative of cleavage of the mCherry tag from the transmembrane domain. (c) Dynamics of AcGFP-tagged Sec61 as an endoplasmic reticulum marker1 in a migrating HaCaT cell expressing paxillin–mCherry. Although ER tubules are dynamically pulled forward into the leading lamella, there was no obvious correlation between ER dynamics and FA turnover. Elapsed time is in minutes.

Supplementary Figure 6 EGFP–Rab6A vesicle fusion with the plasma membrane.

(a) High speed TIRF imaging of EGFP–Rab6A vesicles. Red arrowheads indicate two independent fusion events associated with lateral spreading of EGFP–Rab6A signal in the plasma membrane. Elapsed time is in seconds. (b) Analysis of fluorescence intensity across EGFP–Rab6A vesicles at different time points during fusion normalized to vesicle intensity before fusion (n = 11 vesicles). Solid lines are Gaussian fits. The integrated intensity remains high while the signal spreads in the TIRF plane demonstrating lateral diffusion of EGFP–Rab6A indicative of fusion with the plasma membrane and thus exocytosis. Error bars are 95% confidence intervals.

Supplementary Figure 7 Workflow of generating kymographs to quantify EGFP–Rab6A vesicle transport.

(a) Rotated raw and filtered data, cropped region around a FA (dashed box), and single x-t kymograph at the orange line. (b) Set of all twenty x-t kymographs for each horizontal row of pixels in the cropped image. (c) Maximum intensity projection of all twenty x-t kymographs. See methods for additional details.

Supplementary Figure 8 Uncropped immunoblots shown in supplementary Figures 1 and 3.

Membranes were cut along the dotted lines before incubation with primary antibodies in order to probe shRNA targets and loading controls (GAPDH and tubulin) on the same blots.

Supplementary information

Supplementary Information

Supplementary Information (PDF 2789 kb)

Supplementary Table 1

Supplementary Information (XLSX 10 kb)

Supplementary Table 2

Supplementary Information (XLSX 10 kb)

Supplementary Table 3

Supplementary Information (XLSX 111 kb)

EGFP–CLASP2 and paxillin–mCherry dynamics in a migrating HaCaT cell.

Spinning disk confocal microscopy time lapse sequence of a migrating HaCaT epithelial cell at the edge of a cell monolayer expressing EGFP–CLASP2 (black) and paxillin–mCherry (magenta). EGFP–CLASP2-decorated microtubules engulf FAs before FA disassembly underneath the advancing cell body. Images were acquired every 2 min. The video plays at 15 frames s−1 and is thus accelerated 1800 times. (MOV 8090 kb)

Focal adhesion turnover dynamics in control and CLASP-depleted cells.

Spinning disk confocal microscopy time lapse sequences of migrating HaCaT epithelial cells at the edge of a cell monolayer expressing paxillin–mCherry. Left panel: Control shRNA; middle panel: CLASP1 shRNA; right panel: CLASP2 shRNA. FA disassembly underneath the advancing cell body is disrupted in CLASP-depleted cells. Images were acquired every 3 min. The video plays at 15 frames s−1 and is thus accelerated 2700 times. (MOV 8312 kb)

Focal adhesion-associated CLASP clusters are microtubule independent.

Spinning disk confocal microscopy time lapse sequence of a HaCaT epithelial cell expressing EGFP–CLASP2 (black) and paxillin–mCherry (magenta). 3.3 μM nocodazole was added after 20 min, when EGFP–CLASP2-labeled growing microtubule plus ends abruptly disappear. Images were acquired every 2 min. The video plays at 15 frames s−1 and is thus accelerated 1800 times. (MOV 12138 kb)

Focal adhesion-associated EGFP–CLASP2 dynamics after nocodazole washout.

Spinning disk confocal microscopy time lapse sequence of a HaCaT epithelial cell expressing EGFP–CLASP2 (black) and paxillin–mCherry (magenta). 3.3 μM nocodazole was washed out at the beginning of the time lapse sequence. Re-growing microtubules associate with CLASP clusters prior to FA disassembly. Note that other FAs that did not have associated CLASP clusters do not disassemble. Images were acquired every 30 s. The video plays at 15 frames s−1 and is thus accelerated 450 times. (MOV 1742 kb)

CLASP clusters depend on focal adhesions.

Spinning disk confocal microscopy time lapse sequence of a HaCaT epithelial cell expressing EGFP–CLASP2 (black) and paxillin–mCherry (magenta). 10 M Y-27632 was added after 20 min, which induces FA and subsequent EGFP–CLASP2 cluster disassembly. Images were acquired every 2 min. The video plays at 15 frames s−1 and is thus accelerated 1800 times. (MOV 9486 kb)

EGFP–LL5β and paxillin–mCherry dynamics in a migrating HaCaT cell.

Spinning disk confocal microscopy time lapse sequence of a migrating HaCaT epithelial cell at the edge of a cell monolayer expressing EGFP–LL5β (black) and paxillin–mCherry (magenta). Similar to CLASPs, EGFP–LL5β punctae surround FAs before FA disassembly. Images were acquired every 3 min. The video plays at 15 frames s−1 and is thus accelerated 2700 times. (MOV 4223 kb)

Focal adhesion-associated matrix degradation in a spreading HaCaT cell.

Spinning disk confocal microscopy time lapse sequence of a paxillin–mCherry (magenta) expressing HaCaT epithelial cell spreading on Alexa 488-labeled gelatin. Images were acquired every 5 min. The video plays at 10 frames s−1 and is thus accelerated 3000 times. (MOV 4920 kb)

MT1–MMP–EGFP vesicle dynamics near focal adhesions.

Total internal reflection microscopy time lapse sequence of HaCaT epithelial cells expressing MT1–MMP–EGFP (black) and paxillin–mCherry (magenta) after initial photobleaching of membrane-bound signal. The same three fusion events as shown in Fig. 6c are highlighted by coloured squares. Images were acquired every 0.3 s. The video plays at 24 frames s−1 and is thus accelerated 7 times. (MOV 6307 kb)

EGFP–Rab6A vesicle dynamics in a migrating HaCaT cell.

Total internal reflection microscopy time lapse sequence of HaCaT epithelial cells expressing EGFP–Rab6A (black) and paxillin–mCherry (magenta). A number of closely FA-associated membrane fusion events are highlighted in the magnified region on the right. Images were acquired every second. The video plays at 30 frames s−1 and is thus accelerated 30 times. (MOV 20372 kb)

EGFP–Rab6A vesicle dynamics in CLASP-depleted cells.

Total internal reflection microscopy time lapse sequence of HaCaT epithelial cells expressing EGFP–Rab6A (black) and paxillin–mCherry (magenta). Left panel: CLASP1 shRNA; right panel: CLASP2 shRNA. Images were acquired every second. The video plays at 30 frames s−1 and is thus accelerated 30 times. (MOV 18926 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Stehbens, S., Paszek, M., Pemble, H. et al. CLASPs link focal-adhesion-associated microtubule capture to localized exocytosis and adhesion site turnover. Nat Cell Biol 16, 558–570 (2014). https://doi.org/10.1038/ncb2975

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb2975

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