Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-28T22:08:16.887Z Has data issue: false hasContentIssue false

16 - Micro- and Nanoscale Force Techniques for Mechanotransduction

Published online by Cambridge University Press:  05 July 2014

Nathan J. Sniadecki
Affiliation:
University of Pennsylvania
Wesley R. Legant
Affiliation:
University of Pennsylvania
Christopher S. Chen
Affiliation:
University of Pennsylvania
Mohammad R. K. Mofrad
Affiliation:
University of California, Berkeley
Roger D. Kamm
Affiliation:
Massachusetts Institute of Technology
Get access

Summary

Introduction

Mechanical forces can act as insoluble cues that affect cellular events such as migration, differentiation, growth, and apoptosis. The response to mechanical stimuli leads to adaptive and functional changes in tissue that contribute to physiological homeostasis (Hughes-Fulford 2004; Ingber 2006). Since many diseases occur in a setting where cells are exposed to abnormal forces, it is now evident that alterations in the mechanical context of healthy tissue contributes to pathological responses, such as in hypertension, asthma, and cancer (Ingber 2003; Huang and Ingber 2005). Mechanical forces that affect cellular responses also arise from within cells. Cells generate traction forces through myosin motors and cytoskeletal filaments that are essential for their locomotion and contraction (Lauffenburger and Horwitz 1996; Ridley et al. 2003). These traction forces appear to regulate the same cellular events that are observed with external forces, suggesting a common mechanism for transducing forces into biochemical responses (Chen et al. 2004). For these reasons, identifying the underlying principles in mechanotransduction has been an active area of research.

Depending on the tissue system, cells experience different kinds of external forces. Impulsive forces occur in the musculoskeletal system where strains of 3000–4000 με are common in bone and forces up to 9 kN have been reported in tendons during physical exertion (Lanyon and Smith 1969; Wang 2006). Rhythmic mechanical forces are pervasive in the normal physiology of the vascular or pulmonary systems. Cardiac or ventilatory cycles produce a combination of shear, tensile, and compressive stresses as blood or air flows across the cell surface and pressure levels rise and fall (Davies 1995; Waters et al. 2002). These forces act locally at the site of force but are also dispersed through viscoelastic tissues. These forces propagate along a network of macromolecules that composes the extracellular matrix (ECM), which surrounds the cells, as well as through cell–cell contacts that link adjacent cells. Because these forces are distributed throughout the tissue, the magnitudes of forces acting at the cellular level are not as large as their tissue-level counterparts and range from pico- to nano-Newtons. Yet, even these small forces are able to elicit mechanotransductive responses from cells. Normal physiological processes expose cells to a variety of mechanical stimuli that differ in magnitude, frequency, and direction, but how cells sense and respond to forces at the molecular level to produce orchestrated responses is currently under investigation.

Type
Chapter
Information
Cellular Mechanotransduction
Diverse Perspectives from Molecules to Tissues
, pp. 377 - 402
Publisher: Cambridge University Press
Print publication year: 2009

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Alcaraz, J., Buscemi, L., Grabulosa, M., Trepat, X., Fabry, B., Farre, R. and Navajas, D. (2003). “Microrheology of human lung epithelial cells measured by atomic force microscopy.” Biophys J 84(3): 2071–9.CrossRefGoogle ScholarPubMed
Alenghat, F. J., Fabry, B., Tsai, K. Y., Goldmann, W. H. and Ingber, D. E. (2000). “Analysis of cell mechanics in single vinculin-deficient cells using a magnetic tweezer.” Biochem Biophys Res Commun 277(1): 93–9.CrossRefGoogle ScholarPubMed
Alenghat, F. J. and Ingber, D. E. (2002). “Mechanotransduction: all signals point to cytoskeleton, matrix, and integrins.” Sci STKE 2002(119): PE6.Google ScholarPubMed
Ashkin, A. (1992). “Forces of a single-beam gradient laser trap on a dielectric sphere in the ray optics regime.” Biophys J 61(2): 569–82.CrossRefGoogle ScholarPubMed
Balaban, N. Q., Schwarz, U. S., Riveline, D., Goichberg, P., Tzur, G., Sabanay, I., Mahalu, D., Safran, S., Bershadsky, A., Addadi, L. and Geiger, B. (2001). “Force and focal adhesion assembly: A close relationship studied using elastic micropatterned substrates.” Nat Cell Biol 3(5): 466–72.CrossRefGoogle ScholarPubMed
Ballestrem, C., Erez, N., Kirchner, J., Kam, Z., Bershadsky, A. and Geiger, B. (2006). “Molecular mapping of tyrosine-phosphorylated proteins in focal adhesions using fluorescence resonance energy transfer.” J Cell Sci 119(Pt 5): 866–75.CrossRefGoogle ScholarPubMed
Beningo, K. A., Dembo, M., Kaverina, I., Small, J. V. and Wang, Y. L. (2001). “Nascent focal adhesions are responsible for the generation of strong propulsive forces in migrating fibroblasts.” J Cell Biol 153(4): 881–8.CrossRefGoogle ScholarPubMed
Beningo, K. A., Dembo, M. and Wang, Y. L. (2004). “Responses of fibroblasts to anchorage of dorsal extracellular matrix receptors.” Proc Natl Acad Sci USA 101(52): 18024–9.CrossRefGoogle ScholarPubMed
Beningo, K. A., Hamao, K., Dembo, M., Wang, Y. L. and Hosoya, H. (2006). “Traction forces of fibroblasts are regulated by the Rho-dependent kinase but not by the myosin light chain kinase.” Arch Biochem Biophys 456(2): 224–31.CrossRefGoogle Scholar
Bray, D. (1984). “Axonal growth in response to experimentally applied mechanical tension.” Dev Biol 102(2): 379–89.CrossRefGoogle ScholarPubMed
Bursac, P., Lenormand, G., Fabry, B., Oliver, M., Weitz, D. A., Viasnoff, V., Butler, J. P. and Fredberg, J. J. (2005). “Cytoskeletal remodelling and slow dynamics in the living cell.” Nat Mater 4(7): 557–61.CrossRefGoogle ScholarPubMed
Burton, K., Park, J. H. and Taylor, D. L. (1999). “Keratocytes generate traction forces in two phases.” Mol Biol Cell 10(11): 3745–69.CrossRefGoogle ScholarPubMed
Burton, K. and Taylor, D. L. (1997). “Traction forces of cytokinesis measured with optically modified elastic substrata.” Nature 385(6615): 450–4.CrossRefGoogle ScholarPubMed
Butler, J. P., Tolic-Norrelykke, I. M., Fabry, B. and Fredberg, J. J. (2002). “Traction fields, moments, and strain energy that cells exert on their surroundings.” Am J Physiol Cell Physiol 282(3): C595–605.CrossRefGoogle ScholarPubMed
Cai, Y., Biais, N., Giannone, G., Tanase, M., Jiang, G., Hofman, J. M., Wiggins, C. H., Silberzan, P., Buguin, A., Ladoux, B. and Sheetz, M. P. (2006). “Nonmuscle myosin IIA-dependent force inhibits cell spreading and drives F-actin flow.” Biophys J 91(10): 3907–20.CrossRefGoogle ScholarPubMed
Carter, S. B. (1967). “Haptotaxis and the mechanism of cell motility.” Nature 213(73): 256–60.CrossRefGoogle ScholarPubMed
Charras, G. T. and Horton, M. A. (2002). “Single cell mechanotransduction and its modulation analyzed by atomic force microscope indentation.” Biophys J 82(6): 2970–81.CrossRefGoogle ScholarPubMed
Chen, C. S., Tan, J. and Tien, J. (2004). “Mechanotransduction at cell-matrix and cell-cell contacts.” Annu Rev Biomed Eng 6: 275–302.CrossRefGoogle ScholarPubMed
Chicurel, M. E., Singer, R. H., Meyer, C. J. and Ingber, D. E. (1998). “Integrin binding and mechanical tension induce movement of mRNA and ribosomes to focal adhesions.” Nature 392(6677): 730–3.CrossRefGoogle ScholarPubMed
Choquet, D., Felsenfeld, D. P. and Sheetz, M. P. (1997). “Extracellular matrix rigidity causes strengthening of integrin-cytoskeleton linkages.” Cell 88(1): 39–48.CrossRefGoogle ScholarPubMed
Chrzanowska-Wodnicka, M. and Burridge, K. (1996). “Rho-stimulated contractility drives the formation of stress fibers and focal adhesions.” J Cell Biol 133(6): 1403–15.CrossRefGoogle ScholarPubMed
Cukierman, E., Pankov, R., Stevens, D. R. and Yamada, K. M. (2001). “Taking cell-matrix adhesions to the third dimension.” Science 294(5547): 1708–12.CrossRefGoogle ScholarPubMed
D’Addario, M., Arora, P. D., Ellen, R. P. and McCulloch, C. A. (2002). “Interaction of p38 and Sp1 in a mechanical force-induced, beta 1 integrin-mediated transcriptional circuit that regulates the actin-binding protein filamin-A.” J Biol Chem 277(49): 47541–50.CrossRefGoogle Scholar
D’Addario, M., Arora, P. D., Ellen, R. P. and McCulloch, C. A. (2003). “Regulation of tension-induced mechanotranscriptional signals by the microtubule network in fibroblasts.” J Biol Chem 278(52): 53090–7.CrossRefGoogle ScholarPubMed
D’Addario, M., Arora, P. D., Fan, J., Ganss, B., Ellen, R. P. and McCulloch, C. A. (2001). “Cytoprotection against mechanical forces delivered through beta 1 integrins requires induction of filamin A.” J Biol Chem 276(34): 31969–77.CrossRefGoogle Scholar
D’Addario, M., Arora, P. D. and McCulloch, C. A. (2006). “Role of p38 in stress activation of Sp1.” Gene 379: 51–61.CrossRefGoogle ScholarPubMed
Dai, G., Kaazempur-Mofrad, M. R., Natarajan, S., Zhang, Y., Vaughn, S., Blackman, B. R., Kamm, R. D., Garcia-Cardena, G. and Gimbrone, Jr M. A.. (2004). “Distinct endothelial phenotypes evoked by arterial waveforms derived from atherosclerosis-susceptible and -resistant regions of human vasculature.” Proc Natl Acad Sci USA 101(41): 14871–6.CrossRefGoogle ScholarPubMed
Danowski, B. A., Imanaka-Yoshida, K., Sanger, J. M. and Sanger, J. W. (1992). “Costameres are sites of force transmission to the substratum in adult rat cardiomyocytes.” J Cell Biol 118(6): 1411–20.CrossRefGoogle ScholarPubMed
Davies, P. F. (1995). “Flow-mediated endothelial mechanotransduction.” Physiolog Rev 75(3): 519–60.CrossRefGoogle ScholarPubMed
Davies, P. F., Robotewskyj, A. and Griem, M. L. (1994). “Quantitative studies of endothelial cell adhesion. Directional remodeling of focal adhesion sites in response to flow forces.” J Clin Invest 93(5): 2031–8.CrossRefGoogle ScholarPubMed
Dembo, M., Oliver, T., Ishihara, A. and Jacobson, K. (1996). “Imaging the traction stresses exerted by locomoting cells with the elastic substratum method.” Biophys J 70(4): 2008–22.CrossRefGoogle ScholarPubMed
Dembo, M. and Wang, Y. L. (1999). “Stresses at the cell-to-substrate interface during locomotion of fibroblasts.” Biophys J 76(4): 2307–16.CrossRefGoogle ScholarPubMed
du Roure, O., Saez, A., Buguin, A., Austin, R. H., Chavrier, P., Siberzan, P. and Ladoux, B. (2005). “Force mapping in epithelial cell migration.” Proc Natl Acad Sci USA 102(7): 2390–5.CrossRefGoogle ScholarPubMed
Engler, A. J., Sen, S., Sweeney, H. L. and Discher, D. E. (2006). “Matrix elasticity directs stem cell lineage specification.” Cell 126(4): 677–89.CrossRefGoogle ScholarPubMed
Fabry, B., Maksym, G. N., Butler, J. P., Glogauer, M., Navajas, D. and Fredberg, J. J. (2001). “Scaling the microrheology of living cells.” Phys Rev Lett 87(14): 148102.CrossRefGoogle ScholarPubMed
Fabry, B., Maksym, G. N., Butler, J. P., Glogauer, M., Navajas, D., Taback, N. A., Millet, E. J. and Fredberg, J. J. (2003). “Time scale and other invariants of integrative mechanical behavior in living cells.” Phys Rev E 68(4, Pt. 1): 041914.CrossRefGoogle ScholarPubMed
Fabry, B., Maksym, G. N., Shore, S. A., Moore, P. E., Panettieri, Jr. R. A., Butler, J. P. and Fredberg, J. J. (2001). “Selected contribution: time course and heterogeneity of contractile responses in cultured human airway smooth muscle cells.” J Appl Physiol 91(2): 986–94.CrossRefGoogle ScholarPubMed
Fass, J. N. and Odde, D. J. (2003). “Tensile force-dependent neurite elicitation via anti-beta1 integrin antibody-coated magnetic beads.” Biophys J 85(1): 623–36.CrossRefGoogle ScholarPubMed
Felsenfeld, D. P., Choquet, D. and Sheetz, M. P. (1996). “Ligand binding regulates the directed movement of beta1 integrins on fibroblasts.” Nature 383(6599): 438–40.CrossRefGoogle ScholarPubMed
Fung, Y. C. (1967). “Elasticity of soft tissues in simple elongation.” Am J Physiol 213(6): 1532–44.Google ScholarPubMed
Galbraith, C. G. and Sheetz, M. P. (1997). “A micromachined device provides a new bend on fibroblast traction forces.” Proc Natl Acad Sci USA 94(17): 9114–8.CrossRefGoogle ScholarPubMed
Galbraith, C. G. and Sheetz, M. P. (1999). “Keratocytes pull with similar forces on their dorsal and ventral surfaces.” J Cell Biol 147(6): 1313–24.CrossRefGoogle ScholarPubMed
Galbraith, C. G., Yamada, K. M. and Sheetz, M. P. (2002). “The relationship between force and focal complex development.” J Cell Biol 159(4): 695–705.CrossRefGoogle ScholarPubMed
Garcia-Cardena, G., Comander, J., Anderson, K. R., Blackman, B. R. and Gimbrone, Jr M. A.. (2001). “Biomechanical activation of vascular endothelium as a determinant of its functional phenotype.” Proc Natl Acad Sci USA 98(8): 4478–85.CrossRefGoogle ScholarPubMed
Gaudet, C., Marganski, W. A., Kim, S., Brown, C. T., Gunderia, V., Dembo, M. and Wong, J. Y. (2003). “Influence of type I collagen surface density on fibroblast spreading, motility, and contractility.” Biophys J 85(5): 3329–35.CrossRefGoogle ScholarPubMed
Geiger, B., Bershadsky, A., Pankov, R. and Yamada, K. M. (2001). “Transmembrane extracellular matrix–cytoskeleton crosstalk.” Nat Rev Mol Cell Biol 2(11): 793–805.CrossRefGoogle ScholarPubMed
Giannone, G., Jiang, G., Sutton, D. H., Critchley, D. R. and Sheetz, M. P. (2003). “Talin1 is critical for force-dependent reinforcement of initial integrin-cytoskeleton bonds but not tyrosine kinase activation.” Journal of Cell Biology 163(2): 409–19.CrossRefGoogle Scholar
Glogauer, M., Arora, P., Chou, D., Janmey, P. A., Downey, G. P. and McCulloch, C. A. (1998). “The role of actin-binding protein 280 in integrin-dependent mechanoprotection.” J Biol Chem 273(3): 1689–98.CrossRefGoogle ScholarPubMed
Glogauer, M., Arora, P., Yao, G., Sokholov, I., Ferrier, J. and McCulloch, C. A. (1997). “Calcium ions and tyrosine phosphorylation interact coordinately with actin to regulate cytoprotective responses to stretching.” J Cell Sci 110(Pt 1): 11–21.Google ScholarPubMed
Glogauer, M., Ferrier, J. and McCulloch, C. A. (1995). “Magnetic fields applied to collagen-coated ferric oxide beads induce stretch-activated Ca2+ flux in fibroblasts.” Am J Physiol 269(5 Pt 1): C1093–104.CrossRefGoogle ScholarPubMed
Harris, A. K.. (1984). “Tissue culture cells on deformable substrata: biomechanical implications.” J Biomech Eng 106(1): 19–24.CrossRefGoogle ScholarPubMed
Harris, A. K., Stopak, D. and Wild, P. (1981). “Fibroblast traction as a mechanism for collagen morphogenesis.” Nature 290(5803): 249–51.CrossRefGoogle ScholarPubMed
Harris, A. K., Wild, P. and Stopak, D. (1980). “Silicone rubber substrata: a new wrinkle in the study of cell locomotion.” Science 208(4440): 177–9.CrossRefGoogle Scholar
Heidemann, S. R., Kaech, S., Buxbaum, R. E. and Matus, A. (1999). “Direct observations of the mechanical behaviors of the cytoskeleton in living fibroblasts.” J Cell Biol 145(1): 109–22.CrossRefGoogle ScholarPubMed
Helfman, D. M., Levy, E. T., Berthier, C., Shtutman, M., Riveline, D., Grosheva, I., Lachish-Zalait, A., Elbaum, M. and Bershadsky, A. D. (1999). “Caldesmon inhibits nonmuscle cell contractility and interferes with the formation of focal adhesions.” Mol Biol Cell 10(10): 3097–112.CrossRefGoogle ScholarPubMed
Hill, A. V. (1922). “The maximum work and mechanical efficiency of human muscle, and their most economical speed.” J Physiol 56: 19–41.CrossRefGoogle ScholarPubMed
Hill, A. V. (1938). “The heat of shortening and the dynamic constants of muscle.” Proc Roy Soc B 141: 104–117.CrossRefGoogle Scholar
Hoffman, B. D., Massiera, G., Van Citters, K. M. and Crocker, J. C. (2006). “The consensus mechanics of cultured mammalian cells.” Proc Natl Acad Sci USA 103(27): 10259–64.CrossRefGoogle ScholarPubMed
Hu, S., Chen, J., Fabry, B., Numaguchi, Y., Gouldstone, A., Ingber, D. E., Fredberg, J. J., Butler, J. P. and Wang, N. (2003). “Intracellular stress tomography reveals stress focusing and structural anisotropy in cytoskeleton of living cells.” Am J Physiol Cell Physiol 285(5): C1082–90.CrossRefGoogle ScholarPubMed
Huang, S. and Ingber, D. E. (2005). “Cell tension, matrix mechanics, and cancer development.” Cancer Cell 8(3): 175–6.CrossRefGoogle ScholarPubMed
Hubmayr, R. D., Shore, S. A., Fredberg, J. J., Planus, E., Panettieri, Jr R. A.., Moller, W., Heyder, J. and Wang, N. (1996). “Pharmacological activation changes stiffness of cultured human airway smooth muscle cells.” Am J Physiol 271(5 Pt 1): C1660–8.CrossRefGoogle ScholarPubMed
Hughes-Fulford, M. (2004). “Signal transduction and mechanical stress.” Sci STKE 2004(249): RE12.Google ScholarPubMed
Ingber, D. E. (2003). “Mechanobiology and diseases of mechanotransduction.” Ann Med 35(8): 564–77.CrossRefGoogle ScholarPubMed
Ingber, D. E. (2005). “Mechanical control of tissue growth: Function follows form.” Proc Natl Acad Sci USA 102(33): 11571–2.CrossRefGoogle ScholarPubMed
Ingber, D. E. (2006). “Mechanical control of tissue morphogenesis during embryological development.” Int J Dev Biol 50(2–3): 255–66.CrossRefGoogle ScholarPubMed
James, D. W. and Taylor, J. F. (1969). “The stress developed by sheets of chick fibroblasts in vitro.” Exp Cell Res 54(1): 107–10.CrossRefGoogle ScholarPubMed
Jiang, G. Y., Giannone, G., Critchley, D. R., Fukumoto, E. and Sheetz, M. P. (2003). “Two-piconewton slip bond between fibronectin and the cytoskeleton depends on talin.” Nature 424(6946): 334–7.CrossRefGoogle ScholarPubMed
Kaverina, I., Krylyshkina, O., Beningo, K., Anderson, K., Wang, Y. L. and Small, J. V. (2002). “Tensile stress stimulates microtubule outgrowth in living cells.” J Cell Sci 115(Pt 11): 2283–91.Google ScholarPubMed
Kelley, C., D’Amore, P., Hechtman, H. B. and Shepro, D. (1987). “Microvascular pericyte contractility in vitro: comparison with other cells of the vascular wall.” J Cell Biol 104(3): 483–90.CrossRefGoogle ScholarPubMed
Kiosses, W. B., Daniels, R. H., Otey, C., Bokoch, G. M. and Schwartz, M. A. (1999). “A role for p21-activated kinase in endothelial cell migration.” J Cell Biol 147(4): 831–44.CrossRefGoogle ScholarPubMed
Kolodney, M. S. and Elson, E. L. (1993). “Correlation of myosin light chain phosphorylation with isometric contraction of fibroblasts.” J Biol Chem 268(32): 23850–5.Google ScholarPubMed
Kolodney, M. S. and Elson, E. L. (1995). “Contraction due to microtubule disruption is associated with increased phosphorylation of myosin regulatory light chain.” Proc Natl Acad Sci USA 92(22): 10252–6.CrossRefGoogle ScholarPubMed
Kolodney, M. S. and Wysolmerski, R. B. (1992). “Isometric contraction by fibroblasts and endothelial cells in tissue culture: a quantitative study.” J Cell Biol 117(1): 73–82.CrossRefGoogle ScholarPubMed
Lamoureux, P., Ruthel, G., Buxbaum, R. E. and Heidemann, S. R. (2002). “Mechanical tension can specify axonal fate in hippocampal neurons.” J Cell Biol 159(3): 499–508.CrossRefGoogle ScholarPubMed
Lanyon, L. E. and Smith, R. N. (1969). “Measurements of bone strain in walking animal.” Res Vet Sci 10(1): 93.Google ScholarPubMed
Lardner, T. J. and Archer, R. R. (1994). Mechanics of Solids: an Introduction. New York, McGraw-Hill.Google Scholar
Lauffenburger, D. A. and Horwitz, A. F. (1996). “Cell migration: a physically integrated molecular process.” Cell 84(3): 359–69.CrossRefGoogle Scholar
Laurent, V. M., Kasas, S., Yersin, A., Schaffer, T. E., Catsicas, S., Dietler, G., Verkhovsky, A. B. and Meister, J. J. (2005). “Gradient of rigidity in the lamellipodia of migrating cells revealed by atomic force microscopy.” Biophys J 89(1): 667–75.CrossRefGoogle ScholarPubMed
Lee, J., Leonard, M., Oliver, T., Ishihara, A. and Jacobson, K. (1994). “Traction forces generated by locomoting keratocytes.” J Cell Biol 127(6 Pt 2): 1957–64.CrossRefGoogle ScholarPubMed
Lo, C. M., Wang, H. B., Dembo, M. and Wang, Y. L. (2000). “Cell movement is guided by the rigidity of the substrate.” Biophys J 79(1): 144–52.CrossRefGoogle Scholar
Mack, P. J., Kaazempur-Mofrad, M. R., Karcher, H., Lee, R. T. and Kamm, R. D. (2004). “Force-induced focal adhesion translocation: effects of force amplitude and frequency.” Am J Physiol Cell Physiol 287(4): C954–62.CrossRefGoogle ScholarPubMed
Maniotis, A. J., Chen, C. S. and Ingber, D. E. (1997). “Demonstration of mechanical connections between integrins, cytoskeletal filaments, and nucleoplasm that stabilize nuclear structure.” Proc Natl Acad Sci USA 94(3): 849–54.CrossRefGoogle ScholarPubMed
Mathur, A. B., Collinsworth, A. M., Reichert, W. M., Kraus, W. E. and Truskey, G. A. (2001). “Endothelial, cardiac muscle and skeletal muscle exhibit different viscous and elastic properties as determined by atomic force microscopy.” Biomech 34(12): 1545–53.CrossRefGoogle ScholarPubMed
Mathur, A. B., Truskey, G. A. and Reichert, W. M. (2000). “Atomic force and total internal reflection fluorescence microscopy for the study of force transmission in endothelial cells.” Biophys J 78(4): 1725–35.CrossRefGoogle Scholar
Matthews, B. D., Overby, D. R., Alenghat, F. J., Karavitis, J., Numaguchi, Y., Allen, P. G. and Ingber, D. E. (2004). “Mechanical properties of individual focal adhesions probed with a magnetic microneedle.” Biochem Biophys Res Commun 313(3): 758–64.CrossRefGoogle ScholarPubMed
Matthews, B. D., Overby, D. R., Mannix, R. and Ingber, D. E. (2006). “Cellular adaptation to mechanical stress: Role of integrins, Rho, cytoskeletal tension and mechanosensitive ion channels.” J Cell Sci 119(Pt 3): 508–18.CrossRefGoogle Scholar
Meyer, C. J., Alenghat, F. J., Rim, P., Fong, J. H., Fabry, B. and Ingber, D. E. (2000). “Mechanical control of cyclic AMP signalling and gene transcription through integrins.” Nat Cell Biol 2(9): 666–8.CrossRefGoogle ScholarPubMed
Nelson, C. M., Jean, R. P., Tan, J. L., Liu, W. F., Sniadecki, N. J., Spector, A. A. and Chen, C. S. (2005). “Emergent patterns of growth controlled by multicellular form and mechanics.” Proc Natl Acad Sci USA 102(33): 11594–9.CrossRefGoogle ScholarPubMed
Oliver, T., Dembo, M. and Jacobson, K. (1995). “Traction forces in locomoting cells.” Cell Motil Cytoskeleton 31(3): 225–40.CrossRefGoogle ScholarPubMed
Orr, A. W., Helmke, B. P., Blackman, B. R. and Schwartz, M. A. (2006). “Mechanisms of mechanotransduction.” Dev Cell 10(1): 11–20.CrossRefGoogle ScholarPubMed
Paszek, M. J., Zahir, N., Johnson, K. R., Lakins, J. N., Rozenberg, G. I., Gefen, A., Reinhart-King, C. A., Margulies, S. S., Dembo, M., Boettiger, D., Hammer, D. A. and Weaver, V. M. (2005). “Tensional homeostasis and the malignant phenotype.” Cancer Cell 8(3): 241–54.CrossRefGoogle ScholarPubMed
Pertz, O., Hodgson, L., Klemke, R. L. and Hahn, K. M. (2006). “Spatiotemporal dynamics of RhoA activity in migrating cells.” Nature 440(7087): 1069–72.CrossRefGoogle ScholarPubMed
Pirone, D. M., Liu, W. F., Ruiz, S. A., Gao, L., Raghavan, S., Lemmon, C. A., Romer, L. H. and Chen, C. S. (2006). “An inhibitory role for FAK in regulating proliferation: A link between limited adhesion and RhoA-ROCK signaling.” J Cell Biol 174(2): 277–88.CrossRefGoogle ScholarPubMed
Radmacher, M. (1997). “Measuring the elastic properties of biological samples with the AFM.” IEEE Engin Med Biol Mag 16(2): 47–57.CrossRefGoogle ScholarPubMed
Rajagopalan, P., Marganski, W. A., Brown, X. Q. and Wong, J. Y. (2004). “Direct comparison of the spread area, contractility, and migration of balb/c 3T3 fibroblasts adhered to fibronectin- and RGD-modified substrata.” Biophys J 87(4): 2818–27.CrossRefGoogle ScholarPubMed
Reinhart-King, C. A., Dembo, M. and Hammer, D. A. (2003). “Endothelial cell traction forces on RGD-derivatized polyacrylamide substrata.” Langmuir 19(5): 1573–9.CrossRefGoogle Scholar
Reinhart-King, C. A., Dembo, M. and Hammer, D. A. (2005). “The dynamics and mechanics of endothelial cell spreading.” Biophys J 89(1): 676–89.CrossRefGoogle ScholarPubMed
Ridley, A. J., Schwartz, M. A., Burridge, K., Firtel, R. A., Ginsberg, M. H., Borisy, G., Parsons, J. T. and Horwitz, A. R. (2003). “Cell migration: integrating signals from front to back.” Science 302(5651): 1704–9.CrossRefGoogle ScholarPubMed
Riveline, D., Zamir, E., Balaban, N. Q., Schwarz, U. S., Ishizaki, T., Narumiya, S., Kam, Z., Geiger, B. and Bershadsky, A. D. (2001). “Focal contacts as mechanosensors: externally applied local mechanical force induces growth of focal contacts by an mDia1-dependent and ROCK-independent mechanism.” J Cell Biol 153(6): 1175–86.CrossRefGoogle ScholarPubMed
Rotsch, C. and Radmacher, M. (2000). “Drug-induced changes of cytoskeletal structure and mechanics in fibroblasts: An atomic force microscopy study.” Biophy J 78(1): 520–35.CrossRefGoogle Scholar
Saez, A., Buguin, A., Silberzan, P. and Ladoux, B. (2005). “Is the mechanical activity of epithelial cells controlled by deformations or forces?Biophys J 89(6): L52–4.CrossRefGoogle ScholarPubMed
Sawada, Y. and Sheetz, M. P. (2002). “Force transduction by Triton cytoskeletons.” J Cell Biol 156(4): 609–15.CrossRefGoogle ScholarPubMed
Sawada, Y., Tamada, M., Dubin-Thaler, B. J., Cherniavskaya, O., Sakai, R., Tanaka, S. and Sheetz, M. P. (2006). “Force sensing by mechanical extension of the Src family kinase substrate p130Cas.” Cell 127(5): 1015–26.CrossRefGoogle ScholarPubMed
Schmidt, C. E., Horwitz, A. F., Lauffenburger, D. A. and Sheetz, M. P. (1993). “Integrin-cytoskeletal interactions in migrating fibroblasts are dynamic, asymmetric, and regulated.” J Cell Biol 123(4): 977–91.CrossRefGoogle ScholarPubMed
Schwarz, U. S., Balaban, N. Q., Riveline, D., Bershadsky, A., Geiger, B. and Safran, S. A. (2002). “Calculation of forces at focal adhesions from elastic substrate data: The effect of localized force and the need for regularization.” Biophys J 83: 1380–94.CrossRefGoogle ScholarPubMed
Sun, Z., Martinez-Lemus, L. A., Trache, A., Trzeciakowski, J. P., Davis, G. E., Pohl, U. and Meininger, G. A. (2005). “Mechanical properties of the interaction between fibronectin and alpha(5)beta(1)-integrin on vascular smooth muscle cells studied using atomic force microscopy.” Am J Physiol Heart Circul Physiol 289(6): H2526–35.CrossRefGoogle Scholar
Tan, J. L., Tien, J., Pirone, D. M., Gray, D. S., Bhadriraju, K. and Chen, C. S. (2003). “Cells lying on a bed of microneedles: An approach to isolate mechanical force.” Proc Natl Acad Sci USA 100(4): 1484–9.CrossRefGoogle ScholarPubMed
Trache, A. and Meininger, G. A. (2005). “Atomic force-multi-optical imaging integrated microscope for monitoring molecular dynamics in live cells.” J Biomed Opt 10(6): 064023.CrossRefGoogle ScholarPubMed
von Wichert, G., Haimovich, B., Feng, G. S. and Sheetz, M. P. (2003a). “Force-dependent integrin-cytoskeleton linkage formation requires downregulation of focal complex dynamics by Shp2.” Embo J 22(19): 5023–35.CrossRefGoogle ScholarPubMed
von Wichert, G., Jiang, G., Kostic, A., De Vos, K., Sap, J. and Sheetz, M. P. (2003b). “RPTP-alpha acts as a transducer of mechanical force on alphav/beta3-integrin-cytoskeleton linkages.” J Cell Biol 161(1): 143–53.CrossRefGoogle ScholarPubMed
Walpita, D. and Hay, E. (2002). “Studying actin-dependent processes in tissue culture.” Nat Rev Mol Cell Biol 3(2): 137–41.CrossRefGoogle ScholarPubMed
Wang, H. B., Dembo, M., Hanks, S. K. and Wang, Y. (2001). “Focal adhesion kinase is involved in mechanosensing during fibroblast migration.” Proc Natl Acad Sci USA 98(20): 11295–300.CrossRefGoogle ScholarPubMed
Wang, J. H. (2006). “Mechanobiology of tendon.” J Biomech 39(9): 1563–82.CrossRefGoogle ScholarPubMed
Wang, N., Butler, J. P. and Ingber, D. E. (1993). “Mechanotransduction across the cell surface and through the cytoskeleton.” Science 260(5111): 1124–7.CrossRefGoogle ScholarPubMed
Wang, N. and Ingber, D. E. (1994). “Control of cytoskeletal mechanics by extracellular matrix, cell shape, and mechanical tension.” Biophys J 66(6): 2181–9.CrossRefGoogle ScholarPubMed
Wang, N. and Suo, Z. (2005). “Long-distance propagation of forces in a cell.” Biochem Biophys Res Commun 328(4): 1133–8.CrossRefGoogle Scholar
Wang, Y., Botvinick, E. L., Zhao, Y., Berns, M. W., Usami, S., Tsien, R. Y. and Chien, S. (2005). “Visualizing the mechanical activation of Src.” Nature 434(7036): 1040–5.CrossRefGoogle ScholarPubMed
Waters, C. M., Sporn, P. H., Liu, M. and Fredberg, J. J. (2002). “Cellular biomechanics in the lung.” Am J Physiol Lung Cell Mol Physiol 283(3): L503–9.CrossRefGoogle ScholarPubMed
Wozniak, M. A., Desai, R., Solski, P. A., Der, C. J. and Keely, P. J. (2003). “ROCK-generated contractility regulates breast epithelial cell differentiation in response to the physical properties of a three-dimensional collagen matrix.” J Cell Biol 163(3): 583–95.CrossRefGoogle ScholarPubMed
Yang, Z., Lin, J. S., Chen, J. and Wang, J. H. (2006). “Determining substrate displacement and cell traction fields – a new approach.” J Theor Biol 242(3): 607–16.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×