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:

Locally excitable Cdc42 signals steer cells during chemotaxis

Abstract

Neutrophils and other amoeboid cells chemotax by steering their front ends towards chemoattractant. Although Ras, Rac, Cdc42 and RhoA small GTPases all regulate chemotaxis, it has been unclear how they spatiotemporally control polarization and steering. Using fluorescence biosensors in neutrophil-like PLB-985 cells and photorelease of chemoattractant, we show that local Cdc42 signals, but not those of Rac, RhoA or Ras, precede cell turning during chemotaxis. Furthermore, pre-existing local Cdc42 signals in morphologically unpolarized cells predict the future direction of movement on uniform stimulation. Moreover, inhibition of actin polymerization uncovers recurring local Cdc42 activity pulses, suggesting that Cdc42 has the excitable characteristic of the compass activity proposed in models of chemotaxis. Globally, Cdc42 antagonizes RhoA, and maintains a steep spatial activity gradient during migration, whereas Ras and Rac form shallow gradients. Thus, chemotactic steering and de novo polarization are both directed by locally excitable Cdc42 signals.

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: Neutrophil chemotaxis controlled by automated photorelease of chemoattractant.
Figure 2: Cdc42 and opposing RhoA activities are steeply polarized at the front.
Figure 3: Endogenous local Cdc42 signal differences between the left- and right-hand sides of the cell front direct turning towards chemoattractant.
Figure 4: Basal Cdc42 activity fluctuations in morphologically unpolarized cells predict the future direction of cell polarization.
Figure 5: Locally pulsatile Cdc42 activity in the absence of actin polymerization argues for the existence of an excitable network.
Figure 6: Cdc42 and RhoA activities polarize before Rac and Ras activities during de novo polarization and induced chemokinetic movement.
Figure 7: Cdc42 and opposing RhoA activities correlate spatiotemporally with local protrusion during cell migration.
Figure 8: Cdc42 antagonizes RhoA activity.

Similar content being viewed by others

References

  1. Ridley, A. J. et al. Cell migration: integrating signals from front to back. Science 302, 1704–1709 (2003).

    Article  CAS  PubMed  Google Scholar 

  2. Niggli, V. Signaling to migration in neutrophils: importance of localized pathways. Int. J. Biochem. Cell Biol. 35, 1619–1638 (2003).

    Article  CAS  PubMed  Google Scholar 

  3. Becker, E. L. Stimulated neutrophil locomotion: chemokinesis and chemotaxis. Arch. Pathol. Lab. Med. 101, 509–513 (1977).

    CAS  PubMed  Google Scholar 

  4. Arrieumerlou, C. & Meyer, T. A local coupling model and compass parameter for eukaryotic chemotaxis. Dev. Cell 8, 215–227 (2005).

    Article  CAS  PubMed  Google Scholar 

  5. Devreotes, P. N. & Zigmond, S. H. Chemotaxis in eukaryotic cells: a focus on leukocytes and Dictyostelium. Annu. Rev. Cell Biol. 4, 649–686 (1988).

    Article  CAS  PubMed  Google Scholar 

  6. Andrew, N. & Insall, R. H. Chemotaxis in shallow gradients is mediated independently of PtdIns 3-kinase by biased choices between random protrusions. Nat. Cell Biol. 9, 193–200 (2007).

    Article  CAS  PubMed  Google Scholar 

  7. Alt, W. Biased random walk models for chemotaxis and related diffusion approximations. J. Math. Biol. 9, 147–177 (1980).

    Article  CAS  PubMed  Google Scholar 

  8. Wells, W. A. Chemotaxis by local steering. J. Cell Biol. 168, 674–675 (2005).

    PubMed Central  Google Scholar 

  9. Iglesias, P. A. & Devreotes, P. N. Biased excitable networks: How cells direct motion in response to gradients. Curr. Opin. Cell Biol. 24, 245–253 (2012).

    Article  CAS  PubMed  Google Scholar 

  10. Tang, M. et al. Evolutionarily conserved coupling of adaptive and excitable networks mediates eukaryotic chemotaxis. Nat. Commun. 5, 5175 (2014).

    Article  CAS  PubMed  Google Scholar 

  11. Rickert, P., Weiner, O. D., Wang, F., Bourne, H. R. & Servant, G. Leukocytes navigate by compass: roles of PI3Kγ and its lipid products. Trends Cell Biol. 10, 466–473 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Bokoch, G. M. Chemoattractant signaling and leukocyte activation. Blood 86, 1649–1660 (1995).

    Article  CAS  PubMed  Google Scholar 

  13. Charest, P. G. & Firtel, R. A. Big roles for small GTPases in the control of directed cell movement. Biochem. J. 401, 377–390 (2007).

    Article  CAS  PubMed  Google Scholar 

  14. Heasman, S. J. & Ridley, A. J. Mammalian Rho GTPases: new insights into their functions from in vivo studies. Nat. Rev. Mol. Cell Biol. 9, 690–701 (2008).

    Article  CAS  PubMed  Google Scholar 

  15. Nobes, C. D. & Hall, A. Rho, Rac, and Cdc42 GTPases regulate the assembly of multimolecular focal complexes associated with actin stress fibers, lamellipodia, and filopodia. Cell 81, 53–62 (1995).

    Article  CAS  PubMed  Google Scholar 

  16. Srinivasan, S. et al. Rac and Cdc42 play distinct roles in regulating PI(3,4,5)P3 and polarity during neutrophil chemotaxis. J. Cell Biol. 160, 375–385 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Weiner, O. D. et al. A PtdInsP(3)- and Rho GTPase-mediated positive feedback loop regulates neutrophil polarity. Nat. Cell Biol. 4, 509–513 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sun, C. X. et al. Rac1 is the small GTPase responsible for regulating the neutrophil chemotaxis compass. Blood 104, 3758–3765 (2004).

    Article  CAS  PubMed  Google Scholar 

  19. Huang, C.-H., Tang, M., Shi, C., Iglesias, P. A. & Devreotes, P. N. An excitable signal integrator couples to an idling cytoskeletal oscillator to drive cell migration. Nat. Cell Biol. 15, 1307–1316 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Li, Z. et al. Directional sensing requires G βγ-mediated PAK1 and PIX α-dependent activation of Cdc42. Cell 114, 215–227 (2003).

    Article  CAS  PubMed  Google Scholar 

  21. Wong, K., Pertz, O., Hahn, K. & Bourne, H. Neutrophil polarization: spatiotemporal dynamics of RhoA activity support a self-organizing mechanism. Proc. Natl Acad. Sci. USA 103, 3639–3644 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Xu, J. et al. Divergent signals and cytoskeletal assemblies regulate self-organizing polarity in neutrophils. Cell 114, 201–214 (2003).

    Article  CAS  PubMed  Google Scholar 

  23. Kitzing, T. M. et al. Positive feedback between Dia1, LARG, and RhoA regulates cell morphology and invasion. Genes Dev. 21, 1478–1483 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Medina, F. et al. Activated RhoA is a positive feedback regulator of the Lbc family of Rho guanine nucleotide exchange factor proteins. J. Biol. Chem. 288, 11325–11333 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Bourne, H. R. & Weiner, O. A chemical compass. Nature 419, 21 (2002).

    Article  CAS  PubMed  Google Scholar 

  26. Wang, F. et al. Lipid products of PI(3)Ks maintain persistent cell polarity and directed motility in neutrophils. Nat. Cell Biol. 4, 513–518 (2002).

    Article  CAS  PubMed  Google Scholar 

  27. Hoeller, O. & Kay, R. R. Chemotaxis in the absence of PIP3 gradients. Curr. Biol. 17, 813–817 (2007).

    Article  CAS  PubMed  Google Scholar 

  28. Afonso, P. V. & Parent, C. A. PI3K and chemotaxis: a priming issue? Sci. Signal. 4, pe22 (2011).

    Article  PubMed  CAS  Google Scholar 

  29. Lämmermann, T. et al. Cdc42-dependent leading edge coordination is essential for interstitial dendritic cell migration. Blood 113, 5703–5710 (2009).

    Article  PubMed  CAS  Google Scholar 

  30. Roberts, A. W. et al. Deficiency of the hematopoietic cell-specific Rho family GTPase Rac2 is characterized by abnormalities in neutrophil function and host defense. Immunity 10, 183–196 (1999).

    Article  CAS  PubMed  Google Scholar 

  31. Glogauer, M. et al. Rac1 deletion in mouse neutrophils has selective effects on neutrophil functions. J. Immunol. 170, 5652–5657 (2003).

    Article  CAS  PubMed  Google Scholar 

  32. Bolourani, P., Spiegelman, G. B. & Weeks, G. Delineation of the roles played by RasG and RasC in cAMP-dependent signal transduction during the early development of Dictyostelium discoideum. Mol. Biol. Cell 17, 4543–4550 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Sasaki, A. T., Chun, C., Takeda, K. & Firtel, R. A. Localized Ras signaling at the leading edge regulates PI3K, cell polarity, and directional cell movement. J. Cell Biol. 167, 505–518 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Wang, Y. et al. Identifying network motifs that buffer front-to-back signaling in polarized neutrophils. Cell Rep. 3, 1607–1616 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Yoo, S. K. et al. Differential regulation of protrusion and polarity by PI3K during neutrophil motility in live zebrafish. Dev. Cell 18, 226–236 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Komatsu, N. et al. Development of an optimized backbone of FRET biosensors for kinases and GTPases. Mol. Biol. Cell 22, 4647–4656 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Collins, S. R. et al. Using light to shape chemical gradients for parallel and automated analysis of chemotaxis. Mol. Syst. Biol. 11, 804 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Friedl, P. & Weigelin, B. Interstitial leukocyte migration and immune function. Nat. Immunol. 9, 960–969 (2008).

    Article  CAS  PubMed  Google Scholar 

  39. Renkawitz, J. & Sixt, M. Mechanisms of force generation and force transmission during interstitial leukocyte migration. EMBO Rep. 11, 744–750 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Pirrung, M. C., Drabik, S. J., Ahamed, J. & Ali, H. Caged chemotactic peptides. Bioconjug. Chem. 11, 679–681 (2000).

    Article  CAS  PubMed  Google Scholar 

  41. Zheng, L., Eckerdal, J., Dimitrijevic, I. & Andersson, T. Chemotactic peptide-induced activation of Ras in human neutrophils is associated with inhibition of p120-GAP activity. J. Biol. Chem. 272, 23448–23454 (1997).

    Article  CAS  PubMed  Google Scholar 

  42. Zigmond, S. H. Cell polarity: an examination of its behavioral expression and its consequences for polymorphonuclear leukocyte chemotaxis. J. Cell Biol. 89, 585–592 (1981).

    Article  CAS  PubMed  Google Scholar 

  43. Ghosh, M. et al. Cofilin promotes actin polymerization and defines the direction of cell motility. Science 304, 743–746 (2004).

    Article  CAS  PubMed  Google Scholar 

  44. Gilbert, S. H., Perry, K. & Fay, F. S. Mediation of chemoattractant-induced changes in [Ca2 + ]i and cell shape, polarity, and locomotion by InsP3, DAG, and protein kinase C in newt eosinophils. J. Cell Biol. 127, 489–503 (1994).

    Article  CAS  PubMed  Google Scholar 

  45. Weiger, M. C., Ahmed, S., Welf, E. S. & Haugh, J. M. Directional persistence of cell migration coincides with stability of asymmetric intracellular signaling. Biophys. J. 98, 67–75 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Chau, A. H., Walter, J. M., Gerardin, J., Tang, C. & Lim, W. A. Designing synthetic regulatory networks capable of self-organizing cell polarization. Cell 151, 320–332 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Altschuler, S. J., Angenent, S. B., Wang, Y. & Wu, L. F. On the spontaneous emergence of cell polarity. Nature 454, 886–889 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Butty, A.-C. et al. A positive feedback loop stabilizes the guanine-nucleotide exchange factor Cdc24 at sites of polarization. EMBO J. 21, 1565–1576 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Wedlich-Soldner, R., Altschuler, S., Wu, L. & Li, R. Spontaneous cell polarization through actomyosin-based delivery of the Cdc42 GTPase. Science 299, 1231–1235 (2003).

    Article  CAS  PubMed  Google Scholar 

  50. Meyer, T. & Stryer, L. Molecular model for receptor-stimulated calcium spiking. Proc. Natl Acad. Sci. USA 85, 5051–5055 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Shankaran, H. & Wiley, H. S. Oscillatory dynamics of the extracellular signal-regulated kinase pathway. Curr. Opin. Genet. Dev. 20, 650–655 (2010).

    Article  CAS  PubMed  Google Scholar 

  52. Machacek, M. et al. Coordination of Rho GTPase activities during cell protrusion. Nature 461, 99–103 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Pertz, O., Hodgson, L., Klemke, R. L. & Hahn, K. M. Spatiotemporal dynamics of RhoA activity in migrating cells. Nature 440, 1069–1072 (2006).

    Article  CAS  PubMed  Google Scholar 

  54. Moreau, V., Tatin, F., Varon, C. & Génot, E. Actin can reorganize into podosomes in aortic endothelial cells, a process controlled by Cdc42 and RhoA. Mol. Cell. Biol. 23, 6809–6822 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Friesland, A. et al. Small molecule targeting Cdc42-intersectin interaction disrupts Golgi organization and suppresses cell motility. Proc. Natl Acad. Sci. USA 110, 1261–1266 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Amano, M. et al. Phosphorylation and Activation of Myosin by Rho-associated Kinase (Rho-kinase). J. Biol. Chem. 271, 20246–20249 (1996).

    Article  CAS  PubMed  Google Scholar 

  57. Howell, A. S. et al. Singularity in polarization: rewiring yeast cells to make two buds. Cell 139, 731–743 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Eichinger, L. et al. The genome of the social amoeba Dictyostelium discoideum. Nature 435, 43–57 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Johnson, J. M., Jin, M. & Lew, D. J. Symmetry breaking and the establishment of cell polarity in budding yeast. Curr. Opin. Genet. Dev. 21, 740–746 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Allen, W. E., Zicha, D., Ridley, A. J. & Jones, G. E. A role for Cdc42 in macrophage chemotaxis. J. Cell Biol. 141, 1147–1157 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Houk, A. R. et al. Membrane tension maintains cell polarity by confining signals to the leading edge during neutrophil migration. Cell 148, 175–188 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Postma, M. & Van Haastert, P. J. A diffusion-translocation model for gradient sensing by chemotactic cells. Biophys. J. 81, 1314–1323 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Jilkine, A. & Edelstein-Keshet, L. A comparison of mathematical models for polarization of single eukaryotic cells in response to guided cues. PLoS Comput. Biol. 7, e1001121 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Glogauer, M., Hartwig, J. & Stossel, T. Two pathways through Cdc42 couple the N-formyl receptor to actin nucleation in permeabilized human neutrophils. J. Cell Biol. 150, 785–796 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Mullins, R. D., Heuser, J. A. & Pollard, T. D. The interaction of Arp2/3 complex with actin: nucleation, high affinity pointed end capping, and formation of branching networks of filaments. Proc. Natl Acad. Sci. USA 95, 6181–6186 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Weiner, O. D., Marganski, W. A., Wu, L. F., Altschuler, S. J. & Kirschner, M. W. An actin-based wave generator organizes cell motility. PLoS Biol. 5, 2053–2063 (2007).

    Article  CAS  Google Scholar 

  67. Peyrollier, K. et al. A role for the actin cytoskeleton in the hormonal and growth-factor-mediated activation of protein kinase B. Biochem. J. 352, 617–622 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Yusa, K., Rad, R., Takeda, J. & Bradley, A. Generation of transgene-free induced pluripotent mouse stem cells by the piggyBac transposon. Nat. Methods 6, 363–369 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Guignet, E. G. & Meyer, T. Suspended-drop electroporation for high-throughput delivery of biomolecules into cells. Nat. Methods 5, 393–395 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank K. Aoki and M. Matsuda for GTPase sensors; A. Hayer, D. Garbett and A. Winans for critical reading of the manuscript and helpful discussions; and the Stanford Shared FACS Facility for cell sorting and the National Institute of General Medical Sciences for funding.

Author information

Authors and Affiliations

Authors

Contributions

H.W.Y., S.R.C. and T.M. designed the experiments. H.W.Y. and S.R.C. carried out the experiments and analysed the data. H.W.Y., S.R.C. and T.M. interpreted the data. H.W.Y., S.R.C. and T.M. wrote the paper.

Corresponding authors

Correspondence to Hee Won Yang, Sean R. Collins or Tobias Meyer.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Systematic analysis of cell motility and chemotaxis for control PLB-985 cells and PLB-985 cells stably expressing the indicated GTPase biosensors.

All cells expressed an H2B-mCherry marker which was used for cell tracking. Chemoattractant gradients were generated at time zero by UV uncaging, cells were tracked from frame-to-frame, and statistics of cell movement were calculated37. (a) Scheme for quantifying speed and direction of movement. For b and c: n = 160 (Control), n = 318 (Rac), n = 309 (Cdc42), n = 136 (Ras), and n = 123 (RhoA) cells. (b) Measurement of mean cell speed as function of time relative to chemoattractant gradient generation. (c) Histograms of instantaneous cell direction relative to the chemoattractant gradient for moving cells (an angle of zero indicates perfect directionality). For the number of cells indicated above, the data includes: n = 19743 (Control), n = 42954 (Rac), n = 39045 (Cdc42), n = 13901 (Ras), and n = 15307 (RhoA) frame-to-frame movement steps. (d) Histograms of instantaneous cell direction relative to the chemoattractant gradient from the steeper gradient condition used for all high resolution imaging in this study. The data includes: n = 2056 (Rac), n = 1307 (Cdc42), n = 1801 (Ras), and n = 2798 (RhoA) frame-to-frame movement steps for 59, 49, 80, and 103 independent cells, respectively. (e) Intensity of the Fluo-3 dye inside PLB-985 cells after chemoattractant photorelease. Scale bar is 50 μm. Color bar indicates relative fluorescence unit. (f) Quantitative analysis of timecourse of fluorescence intensity of Fluo-3 as a function of time relative to Nv-fMLF photorelease. Green dotted line marks the time of chemoattractant release. Data were normalized by the initial fluorescence intensity for each cell. Data represent the mean ± s.e.m. of n = 22 cells.

Supplementary Figure 2 Measurement of relative activities of GTPases in migrating cells as a function of distance backwards from the leading edge.

(a) A protrusion/retraction map computed by overlaying and subtracting cell masks from sequential images (left), a color coded map of the distance from the leading edge with associated color bar (middle), and Cdc42 activity (right). Color bar (right) indicates dynamic range of Cdc42 activity. The direction of the fMLF gradient is indicated with a . Scale bar is 10 μm. (b) Computed intracellular gradients of FRET ratio as a function of distance from the leading edge are shown for multiple time points for the individual cells shown in Fig. 2a–d. Each single timepoint curve is shown in gray. The averaged curve over all time points is shown in red. (c) The time-averaged activity curves for each individual cell used in our analysis for Fig. 2e–h (Control (Gradient)) are shown in gray. The curves for the cells depicted in the images in Fig. 2a–d are shown in red. Each curve is normalized by the levels at the front edge (mean of 5 front pixels).

Supplementary Figure 3 Spatial gradients of GTPase activities in the absence of PI3K activation.

(a) Inhibition of the polarization of the PHAkt domain fused to YFP (a biosensor for PIP3) by LY29 (50 μM). Time relative to stimulation is indicated and the direction of the chemoattractant gradient is indicated with a . Scale bar is 10 μm. (b) Relative intensity of PHAkt as a function of distance from the leading edge. Values are normalized to the levels at the front edge. Error bars indicate ± s.e.m. of n = 15 (control) and n = 20 (LY29) averaged traces from timecourses of independent cells. (c) Spatial gradients of GTPase activity in the presence of LY29 (50 μM). Time relative to stimulation is indicated, and the direction of the chemoattractant gradient is indicated with a . Color bars indicate the range of biosensor FRET ratios. Scale bar is 10 μm.

Supplementary Figure 4 Spatial GTPases activity at the front during cell turning and alternate method to assess correlations between asymmetric GTPase signaling and cell turning towards chemoattractant.

(ac) Rac (a), Ras (b), and RhoA (c) activities in cells responding to a changing gradient generated by photorelease of Nv-fMLF. White line connects the centroid of the cell to the center of the cell front. Yellow arrow shows GTPase activity bias at the cell front. Time relative to stimulation is indicated and the direction of the chemoattractant gradient is indicated with a . Color bars indicate the range of biosensor FRET ratios. Scale bar is 10 μm. (d) Schematic of the correlation analysis between a ‘signaling angle’ and a cell turning angle. Cell centroid positions are used to compute cell movement vectors, and a cell turning angle for each frame. A ‘signaling vector’ is computed as a weighted sum of vectors pointing from the cell centroid to parametrized regions on the cell periphery with greater than average FRET ratio. The signaling angle is computed as the angle between the previous cell movement vector and the signaling vector. (e) Comparison of the temporal cross-correlation analysis between the signaling angle and the turning angle for each GTPase sensor. Negative time offsets indicate that signaling asymmetry precedes turning. Error bars indicate ±s.e.m. of n = 41 (Rac), n = 47 (Cdc42), n = 42 (Ras), and n = 47 (RhoA) cells.

Supplementary Figure 5 Local fluctuations or enrichments of GTPases activity in unpolarized cells.

(a) Schematic of the assay to generate a rapid spatially uniform increase of the chemoattractant fMLF to induce a polarization and chemokinesis migration response. (b) Histogram of the maximum or minimum FRET ratio over the periphery of individual cells. Ratios were computed after smoothing to minimize the effects of imaging noise. Values were normalized by the mean FRET ratio for each cell. The red curves indicate histograms for the observed maximum (left) or minimum (right) FRET ratios. The blue curves show control values computed for the same cells in which the pixel positions in the cell periphery were randomly permuted before smoothing and detection of the maximum and minimum signals. The control curves are intended to simulate the distributions expected for uniform signaling activity with similar levels of imaging noise. n = 146 (Rac), n = 137 (Cdc42), n = 168 (Ras), n = 162 (RhoA) cells.

Supplementary Figure 6 Cdc42 activity in cells having two competitive protrusions and prediction of future direction.

(a) Rose plots showing distributions of the angle between the minimum local GTPase activity before stimulation and the subsequent migration direction. indicates p-value is less than 0.001. P-values were calculated using the sign test applied to the cosine of the angles. n = 68 (Rac), n = 69 (Cdc42), n = 63 (Ras), and n = 65 (RhoA) cells. (b) Two examples of cells which initially generated two active protrusions after chemoattractant stimulation are shown. Cdc42 activity is indicated by the color scale. White arrows indicate the sites of pre-existing Cdc42 activity at cell periphery. Pink arrows mark the presence of two protruding fronts in the same cell. Scale bar is 10 μm.

Supplementary Figure 7 Wave like behavior Cdc42 activity in the absence of PI3K activation.

(a) Examples of autonomous waves of Cdc42 activity in the presence of LY29. Cells were treated with LatA (1 μM) and LY (50 μM). White dots mark the location of maximum Cdc42 activity. Cdc42 activity is indicated by the color scale. The images were taken every 2 s for 120 s. Scale bar is 10 μm. (b) Kymograph of Cdc42 activity for a cell 1 treated with both LatA (1 μM) and LY29 (50 μM). The Cdc42 activity was averaged over the horizontal (x-axis) direction to get a one dimensional profile for each timepoint. For this analysis, images were taken at 2 s intervals. Cdc42 activity is indicated by the color scale. (c) Quantitative measurements of local Cdc42 activity in LatA (1 μM) and LY29 (50 μM)-treated cells. Shown are temporal traces for selected 5 μm square regions within individual cells. For this analysis, images were taken every 5 s for 420 s.

Supplementary Figure 8 Timecourses of directed speed during de novo polarization, effect of ZCL278 on GTPase activities, efficiency and specificity of Cdc42 and RhoA knock down, and effect of ZCL278 on pMLC.

(a) Directed speed was measured as the rate of movement of the cell centroid in the direction of eventual cell polarization. Green dotted line marks the time of chemoattractant release. Error bars indicate ±s.e.m. of n = 113 cells. (b) Mean activity of GTPases in PLB-985 cells treated with different doses of ZCL278. Values were normalized by the control condition. Error bars indicate ±s.e.m. of n = 163 (Ras 0 μM), n = 170 (Ras 10 μM), n = 160 (Ras 50 μM), n = 209 (Ras 100 μM), n = 136 (Rac 0 μM), n = 159 (Rac 10 μM), n = 173 (Rac 50 μM), n = 130 (Rac 100 μM), n = 141 (Cdc42 0 μM), n = 179 (Cdc42 10 μM), n = 144 (Cdc42 50 μMl), and n = 162 (Cdc42 100 μM). indicates p-value is less than 0.01. P-values were calculated using the rank sum test. (c) Western blots of whole cell lysates were performed to assess siRNA-mediated gene knockdown efficiency of Cdc42 and RhoA. Blotting for GAPDH is also shown to demonstrate equal loading. (d) Histogram of pMLC intensities measured in individual cells by immunofluorescence for cells treated with ZCL278 (50 μM) for 30 min. n = 50892 (DMSO), and n = 57584 (ZCL278) cells.

Supplementary information

Supplementary Information

Supplementary Information (PDF 1412 kb)

Rac activity during neutrophil-like PLB-985 cell chemotaxis.

These movie files contains 360 s imaging sequences with 5 s intervals between frames, corresponding to the examples shown in Fig. 2a. Time relative to stimulation is indicated and direction of the chemoattractant gradient and sequential UV photorelease are indicated with UV. (AVI 2172 kb)

Cdc42 activity during neutrophil-like PLB-985 cell chemotaxis.

These movie files contains 360 s imaging sequences with 5 s intervals between frames, corresponding to the examples shown in Fig. 2b. Time relative to stimulation is indicated and direction of the chemoattractant gradient and sequential UV photorelease are indicated with UV. (AVI 4686 kb)

Ras activity during neutrophil-like PLB-985 cell chemotaxis.

These movie files contains 360 s imaging sequences with 5 s intervals between frames, corresponding to the examples shown in Fig. 2c. Time relative to stimulation is indicated and direction of the chemoattractant gradient and sequential UV photorelease are indicated with UV. (AVI 2190 kb)

RhoA activity during neutrophil-like PLB-985 cell chemotaxis.

These movie files contains 360 s imaging sequences with 5 s intervals between frames, corresponding to the examples shown in Fig. 2d. Time relative to stimulation is indicated and direction of the chemoattractant gradient and sequential UV photorelease are indicated with UV. (AVI 2346 kb)

Cdc42 activity bias at the cell front predicts direction of cell turning.

This movie file contains a 380 s imaging sequence with 5 s intervals between frames, corresponding to the example shown in Fig. 3b. Time relative to stimulation is indicated and direction of the chemoattractant gradient and sequential UV photorelease are indicated with UV. (AVI 2356 kb)

Prepolarized Cdc42 predicts cell direction after uniform fMLF stimulation.

This movie file contains a 50 s imaging sequence with 2 s intervals between frames, corresponding to the example shown in Fig. 4a. Time relative to stimulation is indicated. (AVI 1632 kb)

Locally pulsatile activation and propagation of Cdc42 in the absence of actin polymerization.

The original video was incorrect; a new file was uploaded on 5 January 2016. This movie file contains a 120 s imaging sequence with 2 s intervals between frames, corresponding to the example shown in Fig. 5a. (AVI 4198 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, H., Collins, S. & Meyer, T. Locally excitable Cdc42 signals steer cells during chemotaxis. Nat Cell Biol 18, 191–201 (2016). https://doi.org/10.1038/ncb3292

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3292

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