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:

Mechanism of IRSp53 inhibition and combinatorial activation by Cdc42 and downstream effectors

Abstract

The Rho family GTPase effector IRSp53 has essential roles in filopodia formation and neuronal development, but its regulatory mechanism is poorly understood. IRSp53 contains a membrane-binding BAR domain followed by an unconventional CRIB motif that overlaps with a proline-rich region (CRIB–PR) and an SH3 domain that recruits actin cytoskeleton effectors. Using a fluorescence reporter assay, we show that human IRSp53 adopts a closed inactive conformation that opens synergistically with the binding of human Cdc42 to the CRIB–PR and effector proteins, such as the tumor-promoting factor Eps8, to the SH3 domain. The crystal structure of Cdc42 bound to the CRIB–PR reveals a new mode of effector binding to Rho family GTPases. Structure-inspired mutations disrupt autoinhibition and Cdc42 binding in vitro and decouple Cdc42- and IRSp53-dependent filopodia formation in cells. The data support a combinatorial mechanism of IRSp53 activation.

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: Autoinhibition and activation of IRSp53 by Cdc42.
Figure 2: Formation of a heterohexameric complex between Cdc42, IRSp53 and Eps8.
Figure 3: Structure of the CRIB–PR bound to Cdc42.
Figure 4: Effect of IRSp53 mutations on filopodia formation in B16F1 melanoma cells as a function of Cdc42 coexpression.
Figure 5: Autoinhibitory interactions involve the CRIB–PR domain.
Figure 6: Autoinhibitory interactions involve the SH3 domain.
Figure 7

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

Referenced accessions

Protein Data Bank

References

  1. Insall, R.H. & Machesky, L.M. Actin dynamics at the leading edge: from simple machinery to complex networks. Dev. Cell 17, 310–322 (2009).

    Article  CAS  PubMed  Google Scholar 

  2. Frost, A., Unger, V.M. & De Camilli, P. The BAR domain superfamily: membrane-molding macromolecules. Cell 137, 191–196 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Suetsugu, S. & Gautreau, A. Synergistic BAR-NPF interactions in actin-driven membrane remodeling. Trends Cell Biol. 22, 141–150 (2012).

    CAS  PubMed  Google Scholar 

  4. Scita, G., Confalonieri, S., Lappalainen, P. & Suetsugu, S. IRSp53: crossing the road of membrane and actin dynamics in the formation of membrane protrusions. Trends Cell Biol. 18, 52–60 (2008).

    CAS  PubMed  Google Scholar 

  5. Ahmed, S., Goh, W.I. & Bu, W. I-BAR domains, IRSp53 and filopodium formation. Semin. Cell Dev. Biol. 21, 350–356 (2010).

    CAS  PubMed  Google Scholar 

  6. Millard, T.H. et al. Structural basis of filopodia formation induced by the IRSp53/MIM homology domain of human IRSp53. EMBO J. 24, 240–250 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Lee, S.H. et al. Structural basis for the actin-binding function of missing-in-metastasis. Structure 15, 145–155 (2007).

    PubMed  PubMed Central  Google Scholar 

  8. Mattila, P.K., Salminen, M., Yamashiro, T. & Lappalainen, P. Mouse MIM, a tissue-specific regulator of cytoskeletal dynamics, interacts with ATP-actin monomers through its C-terminal WH2 domain. J. Biol. Chem. 278, 8452–8459 (2003).

    CAS  PubMed  Google Scholar 

  9. Hori, K., Konno, D., Maruoka, H. & Sobue, K. MALS is a binding partner of IRSp53 at cell-cell contacts. FEBS Lett. 554, 30–34 (2003).

    CAS  PubMed  Google Scholar 

  10. Connolly, B.A., Rice, J., Feig, L.A. & Buchsbaum, R.J. Tiam1-IRSp53 complex formation directs specificity of rac-mediated actin cytoskeleton regulation. Mol. Cell. Biol. 25, 4602–4614 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Mackie, S. & Aitken, A. Novel brain 14–3-3 interacting proteins involved in neurodegenerative disease. FEBS J. 272, 4202–4210 (2005).

    CAS  PubMed  Google Scholar 

  12. Cohen, D., Fernandez, D., Lazaro-Dieguez, F. & Musch, A. The serine/threonine kinase Par1b regulates epithelial lumen polarity via IRSp53-mediated cell-ECM signaling. J. Cell Biol. 192, 525–540 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Robens, J.M., Yeow-Fong, L., Ng, E., Hall, C. & Manser, E. Regulation of IRSp53-dependent filopodial dynamics by antagonism between 14–3-3 binding and SH3-mediated localization. Mol. Cell. Biol. 30, 829–844 (2010).

    CAS  PubMed  Google Scholar 

  14. Funato, Y. et al. IRSp53/Eps8 complex is important for positive regulation of Rac and cancer cell motility/invasiveness. Cancer Res. 64, 5237–5244 (2004).

    CAS  PubMed  Google Scholar 

  15. Disanza, A. et al. Regulation of cell shape by Cdc42 is mediated by the synergic actin-bundling activity of the Eps8–IRSp53 complex. Nat. Cell Biol. 8, 1337–1347 (2006).

    CAS  PubMed  Google Scholar 

  16. Krugmann, S. et al. Cdc42 induces filopodia by promoting the formation of an IRSp53:Mena complex. Curr. Biol. 11, 1645–1655 (2001).

    CAS  PubMed  Google Scholar 

  17. Boczkowska, M., Rebowski, G. & Dominguez, R. Glia maturation factor (GMF) interacts with Arp2/3 complex in a nucleotide state-dependent manner. J. Biol. Chem. 288, 25683–25688 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Lim, K.B. et al. The Cdc42 effector IRSp53 generates filopodia by coupling membrane protrusion with actin dynamics. J. Biol. Chem. 283, 20454–20472 (2008).

    CAS  PubMed  Google Scholar 

  19. Miki, H., Yamaguchi, H., Suetsugu, S. & Takenawa, T. IRSp53 is an essential intermediate between Rac and WAVE in the regulation of membrane ruffling. Nature 408, 732–735 (2000).

    CAS  PubMed  Google Scholar 

  20. Goh, W.I. et al. mDia1 and WAVE2 proteins interact directly with IRSp53 in filopodia and are involved in filopodium formation. J. Biol. Chem. 287, 4702–4714 (2012).

    CAS  PubMed  Google Scholar 

  21. Fujiwara, T., Mammoto, A., Kim, Y. & Takai, Y. Rho small G-protein-dependent binding of mDia to an Src homology 3 domain-containing IRSp53/BAIAP2. Biochem. Biophys. Res. Commun. 271, 626–629 (2000).

    CAS  PubMed  Google Scholar 

  22. Sekerková, G. et al. Novel espin actin-bundling proteins are localized to Purkinje cell dendritic spines and bind the Src homology 3 adapter protein insulin receptor substrate p53. J. Neurosci. 23, 1310–1319 (2003).

    PubMed  PubMed Central  Google Scholar 

  23. Soltau, M. et al. Insulin receptor substrate of 53 kDa links postsynaptic shank to PSD-95. J. Neurochem. 90, 659–665 (2004).

    CAS  PubMed  Google Scholar 

  24. Soltau, M., Richter, D. & Kreienkamp, H.J. The insulin receptor substrate IRSp53 links postsynaptic shank1 to the small G-protein cdc42. Mol. Cell. Neurosci. 21, 575–583 (2002).

    CAS  PubMed  Google Scholar 

  25. Sanda, M. et al. The postsynaptic density protein, IQ-ArfGEF/BRAG1, can interact with IRSp53 through its proline-rich sequence. Brain Res. 1251, 7–15 (2009).

    CAS  PubMed  Google Scholar 

  26. Okamura-Oho, Y., Miyashita, T., Ohmi, K. & Yamada, M. Dentatorubral-pallidoluysian atrophy protein interacts through a proline-rich region near polyglutamine with the SH3 domain of an insulin receptor tyrosine kinase substrate. Hum. Mol. Genet. 8, 947–957 (1999).

    CAS  PubMed  Google Scholar 

  27. Gruenheid, S. & Finlay, B.B. Microbial pathogenesis and cytoskeletal function. Nature 422, 775–781 (2003).

    CAS  PubMed  Google Scholar 

  28. Day, B., Henty, J.L., Porter, K.J. & Staiger, C.J. The pathogen-actin connection: a platform for defense signaling in plants. Annu. Rev. Phytopathol. 49, 483–506 (2011).

    CAS  PubMed  Google Scholar 

  29. de Groot, J.C. et al. Structural basis for complex formation between human IRSp53 and the translocated intimin receptor Tir of enterohemorrhagic E. coli. Structure 19, 1294–1306 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Ruetz, T.J., Lin, A.E. & Guttman, J.A. Enterohaemorrhagic Escherichia coli requires the spectrin cytoskeleton for efficient attachment and pedestal formation on host cells. Microb. Pathog. 52, 149–156 (2012).

    CAS  PubMed  Google Scholar 

  31. Govind, S., Kozma, R., Monfries, C., Lim, L. & Ahmed, S. Cdc42Hs facilitates cytoskeletal reorganization and neurite outgrowth by localizing the 58-kD insulin receptor substrate to filamentous actin. J. Cell Biol. 152, 579–594 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Abbott, M.A., Wells, D.G. & Fallon, J.R. The insulin receptor tyrosine kinase substrate p58/53 and the insulin receptor are components of CNS synapses. J. Neurosci. 19, 7300–7308 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Choi, J. et al. Regulation of dendritic spine morphogenesis by insulin receptor substrate 53, a downstream effector of Rac1 and Cdc42 small GTPases. J. Neurosci. 25, 869–879 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Liu, P.S., Jong, T.H., Maa, M.C. & Leu, T.H. The interplay between Eps8 and IRSp53 contributes to Src-mediated transformation. Oncogene 29, 3977–3989 (2010).

    CAS  PubMed  Google Scholar 

  35. Suetsugu, S. et al. The RAC binding domain/IRSp53-MIM homology domain of IRSp53 induces RAC-dependent membrane deformation. J. Biol. Chem. 281, 35347–35358 (2006).

    CAS  PubMed  Google Scholar 

  36. Chauhan, B.K. et al. Cdc42- and IRSp53-dependent contractile filopodia tether presumptive lens and retina to coordinate epithelial invagination. Development 136, 3657–3667 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Misra, A. et al. Insulin receptor substrate protein 53kDa (IRSp53) is a negative regulator of myogenic differentiation. Int. J. Biochem. Cell Biol. 44, 928–941 (2012).

    CAS  PubMed  Google Scholar 

  38. Etienne-Manneville, S. & Hall, A. Rho GTPases in cell biology. Nature 420, 629–635 (2002).

    CAS  PubMed  Google Scholar 

  39. 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).

    CAS  PubMed  Google Scholar 

  40. Symons, M. & Settleman, J. Rho family GTPases: more than simple switches. Trends Cell Biol. 10, 415–419 (2000).

    CAS  PubMed  Google Scholar 

  41. Abdul-Manan, N. et al. Structure of Cdc42 in complex with the GTPase-binding domain of the 'Wiskott–Aldrich syndrome' protein. Nature 399, 379–383 (1999).

    CAS  PubMed  Google Scholar 

  42. Kim, A.S., Kakalis, L.T., Abdul-Manan, N., Liu, G.A. & Rosen, M.K. Autoinhibition and activation mechanisms of the Wiskott–Aldrich syndrome protein. Nature 404, 151–158 (2000).

    CAS  PubMed  Google Scholar 

  43. Lammers, M., Rose, R., Scrima, A. & Wittinghofer, A. The regulation of mDia1 by autoinhibition and its release by Rho•GTP. EMBO J. 24, 4176–4187 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Rose, R. et al. Structural and mechanistic insights into the interaction between Rho and mammalian Dia. Nature 435, 513–518 (2005).

    CAS  PubMed  Google Scholar 

  45. Lebensohn, A.M. & Kirschner, M.W. Activation of the WAVE complex by coincident signals controls actin assembly. Mol. Cell 36, 512–524 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Chen, Z. et al. Structure and control of the actin regulatory WAVE complex. Nature 468, 533–538 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Alvarez, C.E., Sutcliffe, J.G. & Thomas, E.A. Novel isoform of insulin receptor substrate p53/p58 is generated by alternative splicing in the CRIB/SH3-binding region. J. Biol. Chem. 277, 24728–24734 (2002).

    CAS  PubMed  Google Scholar 

  48. Smith, S.J. & Rittinger, K. Preparation of GTPases for structural and biophysical analysis. Methods Mol. Biol. 189, 13–24 (2002).

    CAS  PubMed  Google Scholar 

  49. Mott, H.R. et al. Structure of the small G protein Cdc42 bound to the GTPase-binding domain of ACK. Nature 399, 384–388 (1999).

    CAS  PubMed  Google Scholar 

  50. Garrard, S.M. et al. Structure of Cdc42 in a complex with the GTPase-binding domain of the cell polarity protein, Par6. EMBO J. 22, 1125–1133 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Hertzog, M. et al. Molecular basis for the dual function of Eps8 on actin dynamics: bundling and capping. PLoS Biol. 8, e1000387 (2010).

    PubMed  PubMed Central  Google Scholar 

  52. Yang, C., Hoelzle, M., Disanza, A., Scita, G. & Svitkina, T. Coordination of membrane and actin cytoskeleton dynamics during filopodia protrusion. PLoS ONE 4, e5678 (2009).

    PubMed  PubMed Central  Google Scholar 

  53. Peter, B.J. et al. BAR domains as sensors of membrane curvature: the amphiphysin BAR structure. Science 303, 495–499 (2004).

    CAS  PubMed  Google Scholar 

  54. Morreale, A. et al. Structure of Cdc42 bound to the GTPase binding domain of PAK. Nat. Struct. Biol. 7, 384–388 (2000).

    CAS  PubMed  Google Scholar 

  55. Ashkenazy, H., Erez, E., Martz, E., Pupko, T. & Ben-Tal, N. ConSurf 2010: calculating evolutionary conservation in sequence and structure of proteins and nucleic acids. Nucleic Acids Res. 38 (suppl.), W529–W533 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Watanabe, N., Kato, T., Fujita, A., Ishizaki, T. & Narumiya, S. Cooperation between mDia1 and ROCK in Rho-induced actin reorganization. Nat. Cell Biol. 1, 136–143 (1999).

    CAS  PubMed  Google Scholar 

  57. Ramalingam, N. et al. Phospholipids regulate localization and activity of mDia1 formin. Eur. J. Cell Biol. 89, 723–732 (2010).

    CAS  PubMed  Google Scholar 

  58. Gorelik, R., Yang, C., Kameswaran, V., Dominguez, R. & Svitkina, T. Mechanisms of plasma membrane targeting of formin mDia2 through its amino terminal domains. Mol. Biol. Cell 22, 189–201 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Saarikangas, J. et al. Molecular mechanisms of membrane deformation by I-BAR domain proteins. Curr. Biol. 19, 95–107 (2009).

    CAS  PubMed  Google Scholar 

  60. Minor, W., Cymborowski, M., Otwinowski, Z. & Chruszcz, M. HKL-3000: the integration of data reduction and structure solution–from diffraction images to an initial model in minutes. Acta Crystallogr. D Biol. Crystallogr. 62, 859–866 (2006).

    PubMed  Google Scholar 

  61. Otwinowski, Z. & Minor, W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307–326 (1997).

    CAS  PubMed  Google Scholar 

  62. Adams, P.D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66, 213–221 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Emsley, P., Lohkamp, B., Scott, W.G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Luan, Q. & Nolen, B.J. Structural basis for regulation of Arp2/3 complex by GMF. Nat. Struct. Mol. Biol. 20, 1062–1068 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by the US National Institutes of Health (NIH) grant R01 MH087950 to R.D. D.J.K. was supported by NIH grant T32 AR053461 and American Cancer Society grant PF-13-033-01-DMC. T.S. and C.Y. were supported by NIH grant GM095977. G.S. and A.D. were supported by the Associazione Italiana per la Ricerca sul Cancro grant IG-2013-14104. Use of IMCA-CAT beamline 17-ID was supported by the Industrial Macromolecular Crystallography Association through a contract with the Hauptman-Woodward Medical Research Institute. The Advanced Photon Source was supported by the US Department of Energy Contract DE-AC02-06CH11357. We thank P. Leavis (Tufts University) for the synthesis of the CRIB–PR peptide.

Author information

Authors and Affiliations

Authors

Contributions

D.J.K. designed and purified proteins, determined crystal structure, designed and performed FRET and ITC experiments, analyzed filopodia from cell images and participated in the writing of the manuscript. C.Y. transfected and imaged cells. A.D. prepared Eps8 and performed cosedimentation assays. M.B. and Y.M. participated in construct design and protein preparation. G.S. and T.S. participated in project design and data analysis. R.D. was responsible for the overall design of the project and participated in structure determination, data analysis and preparation of the manuscript.

Corresponding author

Correspondence to Roberto Dominguez.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Purification of IRSp53 constructs and design of FRET reporters.

(a) SDS–PAGE showing purified IRSp53 fragments and GTPases (see also Supplementary Fig. 2 for gels of IRSp53 FL and Eps8). (b) Design of the FL and BAR–SH3 FRET reporters. Endogenous Cys230 and a C-terminal cysteine added by mutagenesis (Cys519 or Cys452) were first under-labeled with the donor probe fluorescein maleimide and then labeled to saturation with the acceptor probe tetramethylrhodamine maleimide (Methods). In the structure of the BAR domain (PDB ID: 1WDZ) Cys230 falls 53 Å apart from its counterpart in the second molecule of the dimer. The BAR domain has a rigid structure that is not expected to change much upon binding to targets or the membrane. Only reporters containing a donor-acceptor pair will show a change in energy transfer. Molecules that compete with autoinhibitory interactions (GTPases, effectors and the SH3 domain of IRSp53 itself) would be expected to induce a conformational change in the reporter, resulting in a decrease in energy transfer. (c) Peptide mass fingerprinting analysis of tryptic peptides derived from wild type IRSp53 after labeling with fluorescein maleimide. Based on a fluorescein-to-protein ratio of 1:1, IRSp53 was found to be labeled at a single cysteine residue. MS/MS fragmentation of a tryptic peptide with m/z 635.74 containing Cys230 (229QCAVAKNSAAYHSK242) produces a series of b-ions. The masses of all the b-ions are 427.36 Da higher than their theoretical masses, corresponding to the mass of the fluorescein maleimide probe.

Supplementary Figure 2 IRSp53 and Eps8 form a 2:2 complex.

(a) Size exclusion chromatography elution profiles of IRSp53, Eps8, IRSp53–Eps8 complex compared to protein standards with known Stokes radii (1, thyroglobulin; 2, aldolase; 3, albumin; 4, ribonuclease-A). IRSp53 and Eps8 are both elongated in solution, and migrate with apparent MW higher than expected from their theoretical masses. SDS–PAGE analysis identifies the fraction volume corresponding to the peak for each protein. The gel filtration fraction volumes of IRSp53, Eps8, and IRSp53–Eps8 are plotted against those of the protein standards to determine their Stokes radii. (b) Glycerol gradient (10–40%) sedimentation of IRSp53, Eps8, and IRSp53–Eps8 compared to protein standards (1', chymotrypsinogen; 2', albumin; 3', aldolase; 4', catalase) with known Svedberg coefficients. The sedimentation fraction volumes of IRSp53, Eps8, and IRSp53–Eps8 are plotted against those of the protein standards to determine their sedimentation velocities. The buffer used in both experiments was 100 mM Tris-HCl pH 8, 500 mM NaCl, 1 mM dithiothreitol, 1 mM EDTA. The table compares the theoretical masses of the samples to the experimentally determined values. The Stokes radii and the sedimentation velocities were used to calculate the MW of each sample according to the equation1: MW = α × SR × Sv.

Supplementary Figure 3 Localization of IRSp53 and Cdc42G12V in B16F1 mouse melanoma cells.

(a–d) Cells coexpressing GFP and RFP, GFP and mCherry–Cdc42G12V, FL–GFP and RFP, FL–GFP and mCherry–Cdc42G12V. The intensity profiles of actin, GFP (or FL–GFP) and RFP (or mCherry–Cdc42G12V) along line-scans (indicated by a white lines) are shown on the right. Note that when expressed alone, IRSp53 localizes in patches along the leading edge, whereas its plasma membrane localization becomes more uniform when coexpressed with Cdc42G12V.

Supplementary Figure 4 Mutations of CRIB–PR residues disrupt binding to Cdc42G12V.

ITC titration of 400 μM GMPPNP–Cdc42G12V into 15 μM BAR–SH3 constructs carrying mutations within the CRIB and PR regions that disrupt binding.

Supplementary Figure 5 Mutations that disrupt autoinhibitory interactions.

(a) In the structure of the CRIB–PR bound to Cdc42G12V, the canonical 278PxxP281 site is partially exposed, potentially allowing for competitive binding of the SH3 domain of IRSp53 itself or a binding partner. Such mechanism of competitive binding is illustrated here by superimposing a SH3–peptide complex (PDB ID: 1W702) onto the structure of the CRIB–PR bound to Cdc42G12V. SH3 residue P428 and CRIB–PR residues P278 and P281 were mutated to test this model. (b) Cells coexpressing FL–GFP mutant P428L with RFP or with mCherry–Cdc42G12V, and quantification of the number of filopodia per cell, filopodia length and percentage of filopodia filled with actin (as described in Fig. 4, main text). The data for filopodia per cell are presented as mean ± SEM. The total number of cells quantified is indicated inside each bar. *P < 0.005 by two-tailed unpaired Student's t test. Zoomed-in regions (white box) are shown separately for Cdc42G12V, IRSp53 and actin. Scale bars: 10 μm.

Supplementary Figure 6 Nucleotide and GTPase specificity of the CRIB–PR domain.

(a) The CRIB–PR can sense the nucleotide state of Cdc42. Comparison of GMPPNP–Cdc42G12V (gray) and GDP–Cdc42 (slate blue; PDB ID: 1A4R3) showing that switch I (residues 30–38) is open (green) in the GDP-bound state and closed (yellow) in the GMPPNP- and CRIB–PR-bound state. The conformation of switch II (residues 57–72) remains mostly unchanged in the GDP-bound (magenta) and GMPPNP-bound (purple) states. (b) Specificity of the CRIB–PR for Cdc42. Residue conservation scores were calculated with the programs ConSurf4 based on the sequences of the two human isoforms of Cdc42 (UniProt ID: P60953-1 and -2) and the three human isoforms of Rac (Rac1, P63000; Rac2, P15153; Rac3, P60763). The conservation scores were displayed on the surface of the structure of Cdc42 using the program Pymol (Schrödinger). The close-up on the bottom (boxed region) highlights sequence variation between Cdc42 and Rac3 (PDB ID: 2QME) at the CRIB–PR interface. Side chains that differ between the two GTPases are highlighted in the structure and in a sequence alignment (red/salmon).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Cite this article

Kast, D., Yang, C., Disanza, A. et al. Mechanism of IRSp53 inhibition and combinatorial activation by Cdc42 and downstream effectors. Nat Struct Mol Biol 21, 413–422 (2014). https://doi.org/10.1038/nsmb.2781

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.2781

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