DOI QR코드

DOI QR Code

Integrin activation

  • Ginsberg, Mark H. (Department of Medicine, University of California)
  • 투고 : 2014.11.07
  • 발행 : 2014.12.31

초록

Integrin-mediated cell adhesion is important for development, immune responses, hemostasis and wound healing. Integrins also function as signal transducing receptors that can control intracellular pathways that regulate cell survival, proliferation, and cell fate. Conversely, cells can modulate the affinity of integrins for their ligands a process operationally defined as integrin activation. Analysis of activation of integrins has now provided a detailed molecular understanding of this unique form of "inside-out" signal transduction and revealed new paradigms of how transmembrane domains (TMD) can transmit long range allosteric changes in transmembrane proteins. Here, we will review how talin and mediates integrin activation and how the integrin TMD can transmit these inside out signals.

키워드

INTRODUCTION

Integrins, heterodimeric type I transmembrane proteins consisting of α and β subunits, are a major class of receptors involved in adhesive events that control development and lead to pathologies such as cancer and thrombosis. Eighteeen integrin α subunits and 8 β subunits heterodimerize to form 24 different integrins (1). Each subunit contains a single transmembrane domain (TMD) and a short cytoplasmic tail. Besides mediating cell adhesion, integrins transmit signals across the plasma membrane that regulate cell migration, cell survival and growth (2). Conversely, signals from inside cells can increase the binding of integrin extracellular domains to ligands, a process operationally defined as integrin activation. Integrin activation encompasses both changes in affinity of individual integrins due to conformational changes and avidity increases due to integrin clustering (3-5). Precise regulation of integrin activation is particularly important in controlling platelet aggregation through integrin αIIbβ3 (6). Rapid activation of this integrin at the site of a wound is required for hemostasis (7); conversely inappropriate activation of αIIbβ3 can cause a platelet thrombus to occlude a blood vessel resulting in myocardial infarction or stroke. Here we will discuss recent progress in understanding how integrins are activated.

 

INTEGRIN TMD-THE CONDUIT FOR ALLOSTERIC REARRANGEMENTS

Changes in the conformation of integrin extracellular domains are responsible for the changes in integrin monomer affinity. These conformational changes have been the subject of several excellent reviews (3, 8-12) and will not be discussed here. Similarly, clustering of integrins can enchance the binding of multivalent ligands and kindlins have recently emerged as major players in clustering (13). The capacity of intracellular signals to change the conformation of the extracellular domain requires a remarkable transmemebrane allosteric change, a change that must traverse the integrin TMD. Truncation of the integrins at the C-termini of extracellular domains results in constitutively active integrins (14), indicating that TMDs and cytoplasmic tails limit the activation state of integrins. Furthermore, many activating mutations, map to the α or β TMD (15-18). Heterodimeric interactions between α and β TMDs and cytoplasmic tails have been observed by co-immunoprecipitation (19), cysteine crosslinking (20, 21) and by NMR (22) in phospholipid bicelles, but not in detergent micelles (23). Importantly, mutations in TMDs that activate integrins invariably inhibit α and β TMD interactions (19). Thus, physiological integrin activation is likely to require that intracellular signals disrupt integrin αβ TMD interactions.

The structure of the αIIbβ3 TMD complex in a phospholipid bicelle (22) revealed the basis of association of the α and β through two interaction interfaces. The αIIb TMD helix is short, straight and broken at Gly991, the first residue of the highly-conserved Gly-Phe-Phe-Lys-Arg (GFFKR) motif in the membrane proximal region of the α subunits. The two Phe residues of the αIIb GFFKR motif do not form a continuous helix but instead make a sharp turn toward β3 (Fig. 1). In this way, the hydrophobic side chains of those residues reside in the hydrophobic core of the lipid bilayer and stack against hydrophobic residues in the β3 TMD, particularly Trp715 and Ile719. The turning of the membrane-proximal region of αIIb also permits the long-predicted (24) electrostatic interaction between αIIb Arg995 and β3 Asp723 by placing those residues in proximity (Fig. 1). The structure at the inner membrane interface is unique to and likely conserved in integrins and is termed the inner membrane clasp (IMC) (22). The second interface involves helical packing centered on β3 Gly708 and αIIb G972XXXG976 motif at the outer membrane region and is termed the outer membrane clasp (OMC) (Fig. 1). Integrin β3 TMD makes a long and continuous helix with a 25° tilting angle to enable the multipoint interactions with αIIb and accommodate the extra hydrophobic residues in the β3 TMD.

Fig. 1.Structure of integrin αIIbβ3 TMD (ribbon view; αIIb in red and β3 in blue. From PDB 2K9J) showing the two interaction interfaces. Right, outer membrane clasp (OMC) illustrating the helical packing involving αIIb Gly 702 and 706. Left, inner membrane clasp (IMC) showing the electrostatic interaction between αIIb Arg995 and β3 Asp723. Also depicted are the hydrophobic interactions of αIIb Phe992,933 with the β3 TMD. Adapted from reference (22).

As noted above, an important feature of the structure of the αIIbβ3 TMD dimer is that the helical β TMD must be precisely tilted to maintain simultaneous formation of the OMC and IMC. Precise tilt is maintained via β3 Lys716 whose alpha carbon resides in the hydrophobic region of the lipid bilayer but its positively charged ε-NH3+ is predicted to snorkel into the negatively charged phosphate head group region (25). Mutation of Lys716 any residue other than Arg (which also contains a snorkeling basic side chain) reduces α-β TMD interactions and dramatically increases integrin activation (25). The effects of Lys716 mutation can be ameliorated by breaking the continuous β TMD helix into two halves by introduction of a Pro mutation (A711P). The Pro mutation, introduces a flexible hinge that partially decouples the tilting angles of inner and outer helices favoring simultaneous formation of OMC and IMC (25).

 

TALIN “TILTS” THE INTEGRIN β TMD TO INDUCE ACTIVATION

Talin regulates integrin affinity and provides a mechanical link between integrins and the actin cytoskeleton. Talin comprises a 50-kDa N-terminal FERM domain (talin head domain or THD) that contains a high-affinity binding site for integrin β tails and a 220-kDa rod domain that contains multiple binding sites for actin and vinculin (26). The THD is further divided into F0, F1, F2 and F3 subdomains (26, 27). The F3 subdomain, contains the major integrin integrin β tail binding site (28, 29). The essential role of talin in regulating integrin affinity has been well documented in model cells (29-32), transgenic mice (33-36) and reconstituted systems with purified proteins (37). In in vitro systems, recombinant THD alone is sufficient to activate αIIbβ3 reconstituted in either liposomes or phospholipid nanodiscs, and activation is associated with a shift towards an αIIbβ3 extended conformation (37).

We now have considerable insight into how talin induces this allosteric rearrangement in integrins. Talin binds to two sites on integrin β tails: a strong binding site centered around the first NPxY motif that contributes most of the binding free energy and a weaker membrane proximal (MP) binding site (38). In addition, THD also binds to negatively charged phospholipids via positively-charged residues (38-40). The weak interaction with the MP region has two important effects: 1) it brings talin Lys324 close to Asp723 of the β3 tail, thus competing for the Arg995-Asp723 electrostatic interactions in IMC (40); 2) it stabilizes α-helical structure of the β MP region and to form a continuous helix with the β3 TMD (38, 40). As the simultaneous interaction with integrin β tails and phospholipids, can change the tilt angle of the β3 MP tail and thus of the contiguous β3 TMD (Fig. 2) (41). Such talin-induced motion was demonstrated by increased fluorescence of solvatochromic dyes attached to the N- or C-terminii of the β3 TMD in the presence of THD (41) and is further supported by molecular dynamic simulations (42). The change in tilting angle destabilizes α-β TMD interactions and shifts the equilibrium towards an activated integrin conformation. In further support of this model, introducing a flexible proline kink in the middle of the β3 TMD blocks THD-induced tilting of the outer membrane segment without blocking tilting of the inner membrane segment (Fig. 3). Integrins bearing this mutation are remarkably resistant to talin induced integrin activation (41).

Fig. 2.Talin changes the topology of the β3 TMD. A peptide containing the β3 TMD and cytoplasmic domain was labeled with environment sensitive bimanes at the outer edge of the TMD (N terminal labeling, Leu694) or at the TMD cytosol interface (C-terminal labeling, Ile721). The peptides were individually embedded in phospholipid nanodiscs and increasing concentrations of talin head domain (THD) were added and bimanes emission spectra were recorded. The increased fluorescence indicates that both sides of the β3 TMD were in a less polar environment suggesting that THD increased the tilting of the β3 TMD. Adapted from reference (41).

Fig. 3.A proline kink prevents transmission of altered tilt across the β3 TMD. In the left two panels The experimental design was identical to that in Fig. 2 and depicts the talin-induced increased embedding at both the inner (where THD binds) and outer edges of the TMD. Introduction of a flexible kink by β3 (A711P) mutation (right two panels) prevents the transmission of increased embedding of the inner TMD to the outer region. Adapted from reference (41).

 

SUMMARY AND CONCLUSIONS

Integrin activation was first observed in 1978 in integrin αIIbβ3 and it has proved to be a conserved property of β1, β2, and β3 integrins. As summarized here, our understanding of this unique form of transmembrane signal transduction has moved from a black box in which agonists, such as thrombin, caused a change in the affinity of integrin αIIbβ3. Today, cell biological and reverse genetic experiments have verified that talin binding to the integrin β cytoplasmic domain is a final common step in activation. Structural studies have revealed how two binding interfaces of talin with the integrin in combination with talin membrane binding sites can effect this form of transmembrane allostery. Studies have also revealed unique features of the heterodimeric integrin TMD that form a stable yet dynamic αβ TMD interaction that enables transmission of the activation signal across the phospholipid bilayer. The lessons learned in studying integrin transmembrane signaling, such as the importance of snorkeling basic residues in maintaining TMD topology, are likely to pertain to other examples of transmembrane signaling through transmembrane receptors.

참고문헌

  1. Hynes, R. O. (2002) Integrins: bidirectional, allosteric signaling machines. Cell 110, 673-687. https://doi.org/10.1016/S0092-8674(02)00971-6
  2. Giancotti, F. G. and Ruoslahti, E. (1999) Integrin signaling. Science 285, 1028-1032. https://doi.org/10.1126/science.285.5430.1028
  3. Kim, C., Ye, F. and Ginsberg, M. H. (2011) Regulation of integrin activation. Annu. Rev. Cell. Dev. Biol. 27, 321-345. https://doi.org/10.1146/annurev-cellbio-100109-104104
  4. Shattil, S. J., Kim, C. and Ginsberg, M. H. (2010) The final steps of integrin activation: the end game. Nat. Rev. Mol. Cell. Biol. 11, 288-300. https://doi.org/10.1038/nrm2871
  5. Shattil, S. J. and Newman, P. J. (2004) Integrins: dynamic scaffolds for adhesion and signaling in platelets. Blood 104, 1606-1615. https://doi.org/10.1182/blood-2004-04-1257
  6. Wagner, C. L., Mascelli, M. A., Neblock, D. S., Weisman, H. F., Coller, B. S. and Jordan, R. E. (1996) Analysis of GPIIb/IIIa receptor number by quantification of 7E3 binding to human platelets. Blood 88, 907-914.
  7. Shattil, S. J., Kashiwagi, H. and Pampori, N. (1998) Integrin signaling: the platelet paradigm. Blood 91, 2645-2657.
  8. Luo, B. H., Carman, C. V. and Springer, T. A. (2007) Structural basis of integrin regulation and signaling. Annu. Rev. Immunol. 25, 619-647. https://doi.org/10.1146/annurev.immunol.25.022106.141618
  9. Arnaout, M. A., Goodman, S. L. and Xiong, J. P. (2007) Structure and mechanics of integrin-based cell adhesion. Curr. Opin. Cell. Biol. 19, 495-507. https://doi.org/10.1016/j.ceb.2007.08.002
  10. Luo, B. H. and Springer, T. A. (2006) Integrin structures and conformational signaling. Curr. Opin. Cell. Biol. 18, 579-586. https://doi.org/10.1016/j.ceb.2006.08.005
  11. Arnaout, M. A., Mahalingam, B. and Xiong, J. P. (2005) Integrin structure, allostery, and bidirectional signaling. Annu. Rev. Cell. Dev. Biol. 21, 381-410. https://doi.org/10.1146/annurev.cellbio.21.090704.151217
  12. Shimaoka, M., Takagi, J. and Springer, T. A. (2002) Conformational regulation of integrin structure and function. Annu. Rev. Biophys. Biomol. Struct. 31, 485-516. https://doi.org/10.1146/annurev.biophys.31.101101.140922
  13. Ye, F., Petrich, B. G., Anekal, P., Lefort, C. T., Kasirer-Friede, A., Shattil, S. J., Ruppert, R., Moser, M., Fassler, R. and Ginsberg, M. H. (2013) The Mechanism of Kindlin-mediated Activation of Integrin ${\alpha}IIb{\beta}3$. Curr. Biol. 23, 2288-2295. https://doi.org/10.1016/j.cub.2013.09.050
  14. Mehta, R. J., Diefenbach, B., Brown, A., Cullen, E., Jonczyk, A., Gussow, D., Luckenbach, G. A. and Goodman, S. L. (1998) Transmembrane-truncated alphavbeta3 integrin retains high affinity for ligand binding: evidence for an 'inside-out' suppressor? Biochem. J. 330, 861-869. https://doi.org/10.1042/bj3300861
  15. Partridge, A. W., Liu, S., Kim, S., Bowie, J. U. and Ginsberg, M. H. (2005) Transmembrane domain helix packing stabilizes integrin alphaIIbbeta3 in the low affinity state. J. Biol. Chem. 280, 7294-7300. https://doi.org/10.1074/jbc.M412701200
  16. Luo, B. H., Carman, C. V., Takagi, J. and Springer, T. A. (2005) Disrupting integrin transmembrane domain heterodimerization increases ligand binding affinity, not valency or clustering. Proc. Natl. Acad. Sci. U. S. A. 102, 3679-3684. https://doi.org/10.1073/pnas.0409440102
  17. Li, W., Metcalf, D. G., Gorelik, R., Li, R., Mitra, N., Nanda, V., Law, P. B., Lear, J. D., Degrado, W. F. and Bennett, J. S. (2005) A push-pull mechanism for regulating integrin function. Proc. Natl. Acad. Sci. U. S. A. 102, 1424-1429. https://doi.org/10.1073/pnas.0409334102
  18. Li, R., Mitra, N., Gratkowski, H., Vilaire, G., Litvinov, R., Nagasami, C., Weisel, J. W., Lear, J. D., DeGrado, W. F. and Bennett, J. S. (2003) Activation of integrin alphaIIbbeta3 by modulation of transmembrane helix associations. Science 300, 795-798. https://doi.org/10.1126/science.1079441
  19. Kim, C., Lau, T. L., Ulmer, T. S. and Ginsberg, M. H. (2009) Interactions of platelet integrin alphaIIb and beta3 transmembrane domains in mammalian cell membranes and their role in integrin activation. Blood 113, 4747-4753. https://doi.org/10.1182/blood-2008-10-186551
  20. Zhu, J., Luo, B. H., Barth, P., Schonbrun, J., Baker, D. and Springer, T. A. (2009) The structure of a receptor with two associating transmembrane domains on the cell surface: integrin alphaIIbbeta3. Mol. Cell 34, 234-249. https://doi.org/10.1016/j.molcel.2009.02.022
  21. Luo, B. H., Springer, T. A. and Takagi, J. (2004) A specific interface between integrin transmembrane helices and affinity for ligand. PLoS Biol. 2, e153. https://doi.org/10.1371/journal.pbio.0020153
  22. Lau, T. L., Kim, C., Ginsberg, M. H. and Ulmer, T. S. (2009) The structure of the integrin alphaIIbbeta3 transmembrane complex explains integrin transmembrane signalling. EMBO J. 28, 1351-1361. https://doi.org/10.1038/emboj.2009.63
  23. Li, R., Babu, C. R., Lear, J. D., Wand, A. J., Bennett, J. S. and DeGrado, W. F. (2001) Oligomerization of the integrin alphaIIbbeta3: roles of the transmembrane and cytoplasmic domains. Proc. Natl. Acad. Sci. U. S. A. 98, 12462-12467. https://doi.org/10.1073/pnas.221463098
  24. Hughes, P. E., Diaz-Gonzalez, F., Leong, L., Wu, C., McDonald, J. A. and Shattil, S. J., Ginsberg, M. H. (1996) Breaking the integrin hinge: a defined structural constraint regulates integrin signaling. J. Biol. Chem. 271, 6571-6574. https://doi.org/10.1074/jbc.271.12.6571
  25. Kim, C., Schmidt, T., Cho, E. G., Ye, F., Ulmer, T. S. and Ginsberg, M. H. (2012) Basic amino-acid side chains regulate transmembrane integrin signalling. Nature 481, 209-213. https://doi.org/10.1038/nature10697
  26. Critchley, D. R. (2009) Biochemical and structural properties of the integrin-associated cytoskeletal protein talin. Annu. Rev. Biophys 38, 235-254. https://doi.org/10.1146/annurev.biophys.050708.133744
  27. Elliott, P. R., Goult, B. T., Kopp, P. M., Bate, N., Grossmann, J. G., Roberts, G. C., Critchley, D. R. and Barsukov, I. L. (2010) The Structure of the talin head reveals a novel extended conformation of the FERM domain. Structure 18, 1289-1299. https://doi.org/10.1016/j.str.2010.07.011
  28. Calderwood, D. A., Fujioka, Y., de Pereda, J. M., Garcia-Alvarez, B., Nakamoto, T., Margolis, B., McGlade, C. J., Liddington, R. C. and Ginsberg, M. H. (2003) Integrin beta cytoplasmic domain interactions with phosphotyrosine-binding domains: a structural prototype for diversity in integrin signaling. Proc. Natl. Acad. Sci. U. S. A. 100, 2272-2277. https://doi.org/10.1073/pnas.262791999
  29. Calderwood, D. A., Yan, B., de Pereda, J. M., Alvarez, B. G., Fujioka, Y., Liddington, R. C. and Ginsberg, M. H. (2002) The phosphotyrosine binding-like domain of talin activates integrins. J. Biol. Chem. 277, 21749-21758. https://doi.org/10.1074/jbc.M111996200
  30. Lee, H. S., Lim, C. J., Puzon-McLaughlin, W., Shattil, S. J. and Ginsberg, M. H. (2009) RIAM activates integrins by linking talin to ras GTPase membrane-targeting sequences. J. Biol. Chem. 284, 5119-5127. https://doi.org/10.1074/jbc.M807117200
  31. Han, J., Lim, C. J., Watanabe, N., Soriani, A., Ratnikov, B., Calderwood, D. A., Puzon-McLaughlin, W., Lafuente, E. M., Boussiotis, V. A., Shattil, S. J. and Ginsberg, M. H. (2006) Reconstructing and deconstructing agonist-induced activation of integrin alphaIIbbeta3. Curr. Biol. 16, 1796-1806. https://doi.org/10.1016/j.cub.2006.08.035
  32. Calderwood, D. A., Zent, R., Grant, R., Rees, D. J., Hynes, R. O. and Ginsberg, M. H. (1999) The Talin head domain binds to integrin beta subunit cytoplasmic tails and regulates integrin activation. J. Biol. Chem. 274, 28071-28074. https://doi.org/10.1074/jbc.274.40.28071
  33. Petrich, B. G., Marchese, P., Ruggeri, Z. M., Spiess, S., Weichert, R. A., Ye, F., Tiedt, R., Skoda, R. C., Monkley, S. J., Critchley, D. R. and Ginsberg, M. H. (2007) Talin is required for integrin-mediated platelet function in hemostasis and thrombosis. J. Exp. Med. 204, 3103-3111. https://doi.org/10.1084/jem.20071800
  34. Nieswandt, B., Moser, M., Pleines, I., Varga-Szabo, D., Monkley, S., Critchley, D. and Fassler, R. (2007) Loss of talin1 in platelets abrogates integrin activation, platelet aggregation, and thrombus formation in vitro and in vivo. J. Exp. Med. 204, 3113-3118. https://doi.org/10.1084/jem.20071827
  35. Petrich, B. G., Fogelstrand, P., Partridge, A. W., Yousefi, N., Ablooglu, A. J., Shattil, S. J. and Ginsberg, M. H. (2007) The antithrombotic potential of selective blockade of talin-dependent integrin alpha IIb beta 3 (platelet GPIIb-IIIa) activation. J. Clin. Invest. 117, 2250-2259. https://doi.org/10.1172/JCI31024
  36. Haling, J. R., Monkley, S. J., Critchley, D. R. and Petrich, B. G. (2011) Talin-dependent integrin activation is required for fibrin clot retraction by platelets. Blood 117, 1719-1722. https://doi.org/10.1182/blood-2010-09-305433
  37. Ye, F., Hu, G., Taylor, D., Ratnikov, B., Bobkov, A. A., McLean, M. A., Sligar, S. G., Taylor, K. A. and Ginsberg, M. H. (2010) Recreation of the terminal events in physiological integrin activation. J. Cell. Biol. 188, 157-173. https://doi.org/10.1083/jcb.200908045
  38. Wegener, K. L., Partridge, A. W., Han, J., Pickford, A. R., Liddington, R. C., Ginsberg, M. H. and Campbell, I. D. (2007) Structural basis of integrin activation by talin. Cell 128, 171-182. https://doi.org/10.1016/j.cell.2006.10.048
  39. Goult, B. T., Bouaouina, M., Elliott, P. R., Bate, N., Patel, B., Gingras, A. R., Grossmann, J. G., Roberts, G. C., Calderwood, D. A., Critchley, D. R. and Barsukov, I. L. (2010) Structure of a double ubiquitin-like domain in the talin head: a role in integrin activation. EMBO J. 29, 1069-1080. https://doi.org/10.1038/emboj.2010.4
  40. Anthis, N. J., Wegener, K. L., Ye, F., Kim, C., Goult, B. T., Lowe, E. D., Vakonakis, I., Bate, N., Critchley, D. R., Ginsberg, M. H. and Campbell, I. D. (2009) The structure of an integrin/talin complex reveals the basis of inside-out signal transduction. EMBO J. 28, 3623-3632. https://doi.org/10.1038/emboj.2009.287
  41. Kim, C., Ye, F., Hu, X. and Ginsberg, M. H. (2012) Talin activates integrins by altering the topology of the beta transmembrane domain. J. Cell. Biol. 197, 605-611. https://doi.org/10.1083/jcb.201112141
  42. Kalli, A. C., Wegener, K. L., Goult, B. T., Anthis, N. J., Campbell, I. D. and Sansom, M. S. (2010) The structure of the talin/integrin complex at a lipid bilayer: an NMR and MD simulation study. Structure 18, 1280-1288. https://doi.org/10.1016/j.str.2010.07.012

피인용 문헌

  1. Hanging on for the ride: Adhesion to the extracellular matrix mediates cellular responses in skeletal muscle morphogenesis and disease vol.401, pp.1, 2015, https://doi.org/10.1016/j.ydbio.2015.01.002
  2. Proteins involved in focal adhesion signaling pathways are differentially regulated in experimental branch retinal vein occlusion vol.138, 2015, https://doi.org/10.1016/j.exer.2015.06.011
  3. Dynamin2 controls Rap1 activation and integrin clustering in human T lymphocyte adhesion vol.12, pp.3, 2017, https://doi.org/10.1371/journal.pone.0172443
  4. Extracellular matrix component signaling in cancer vol.97, 2016, https://doi.org/10.1016/j.addr.2015.10.013
  5. Regulatory Mechanisms of the Molecular Pathways in Fibrosis Induced by MicroRNAs vol.129, pp.19, 2016, https://doi.org/10.4103/0366-6999.190677
  6. Astrocytes in Migration vol.42, pp.1, 2017, https://doi.org/10.1007/s11064-016-2089-4
  7. Integrins in the Spotlight of Cancer vol.17, pp.12, 2016, https://doi.org/10.3390/ijms17122037
  8. Human IGF-I Eb-peptide induces cell attachment and lamellipodia outspread of metastatic breast carcinoma cells (MDA-MB-231) vol.358, pp.2, 2017, https://doi.org/10.1016/j.yexcr.2017.06.015
  9. Recognizing asymmetry in pseudo-symmetry; structural insights into the interaction between amphipathic α-helices and X-bundle proteins 2017, https://doi.org/10.1016/j.bbapap.2017.06.017
  10. Wdr1-Dependent Actin Reorganization in Platelet Activation vol.11, pp.9, 2016, https://doi.org/10.1371/journal.pone.0162897
  11. Roles of integrin β3 cytoplasmic tail in bidirectional signal transduction in a trans-dominant inhibition model vol.10, pp.3, 2016, https://doi.org/10.1007/s11684-016-0460-0
  12. Focal adhesion kinase signaling in unexpected places vol.45, 2017, https://doi.org/10.1016/j.ceb.2017.01.003
  13. CIB1: a small protein with big ambitions vol.30, pp.8, 2016, https://doi.org/10.1096/fj.201500073R
  14. Phosphatidylinositol 3-Kinase/Akt Mediates Integrin Signaling To Control RNA Polymerase I Transcriptional Activity vol.36, pp.10, 2016, https://doi.org/10.1128/MCB.00004-16
  15. Talin regulates integrin β1-dependent and -independent cell functions in ureteric bud development vol.144, pp.22, 2017, https://doi.org/10.1242/dev.149914
  16. The secret life of ion channels: Kv1.3 potassium channels and proliferation vol.314, pp.1, 2018, https://doi.org/10.1152/ajpcell.00136.2017
  17. Regulation of inside-out β1-integrin activation by CDCP1 vol.37, pp.21, 2018, https://doi.org/10.1038/s41388-018-0142-2
  18. Wdr-1 is essential for F-actin interaction with focal adhesions in platelets vol.29, pp.6, 2018, https://doi.org/10.1097/MBC.0000000000000756