Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-23T08:20:28.341Z Has data issue: false hasContentIssue false

Part II - Recent progress in cell mechanobiology

Published online by Cambridge University Press:  05 November 2015

Yu Sun
Affiliation:
University of Toronto
Deok-Ho Kim
Affiliation:
University of Washington
Craig A. Simmons
Affiliation:
University of Toronto
Get access

Summary

Tumor angiogenesis is a key regulator of tumor growth and metastasis. Assays allowing the analysis of tumor angiogenesis are an essential tool to elucidate the role played by the tumor microenvironment in regulating tumor angiogenesis. The assays should also be capable of systematically investigating the effects of physiologically relevant, mechanical and chemical stimuli and their synergistic interactions. The high optical resolution of microfluidic assays facilitates three-dimensional studies of cellular morphogenesis. Their versatility can be applied to study the multi-parameter control of angiogenic factors.

Type
Chapter
Information
Integrative Mechanobiology
Micro- and Nano- Techniques in Cell Mechanobiology
, pp. 203 - 367
Publisher: Cambridge University Press
Print publication year: 2015

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

References

Abrams, G. A., Goodman, S. L., Nealey, P. F., Franco, M. and Murphy, C. J. (2000). “Nanoscale topography of the basement membrane underlying the corneal epithelium of the rhesus macaque.” Cell Tissue Res 299: 3946.CrossRefGoogle ScholarPubMed
Adamo, L., Naveiras, O., Wenzel, P. L., et al. (2009). “Biomechanical forces promote embryonic haematopoiesis.” Nature 459: 11311135.CrossRefGoogle ScholarPubMed
Adams, D. S., Keller, R., and Koehl, M. A. (1990). “The mechanics of notochord elongation, straightening and stiffening in the embryo of Xenopus laevis.” Development 110: 115130.CrossRefGoogle ScholarPubMed
Anava, S., Greenbaum, A., Ben Jacob, E., Hanein, Y., and Ayali, A. (2009). “The regulative role of neurite mechanical tension in network development.” Biophys J 96: 16611670.CrossRefGoogle ScholarPubMed
Arnold, M., Cavalcanti-Adam, E. A., Glass, R., et al. (2004). “Activation of integrin function by nanopatterned adhesive interfaces.” Chemphyschem 5: 383388.CrossRefGoogle ScholarPubMed
Banerjee, A., Arha, M., Choudhary, S., et al. (2009). “The influence of hydrogel modulus on the proliferation and differentiation of encapsulated neural stem cells.” Biomaterials 30: 46954699.CrossRefGoogle ScholarPubMed
Beech, D. J. (2005). “TRPC1: store-operated channel and more.” Pflugers Arch 451: 5360.CrossRefGoogle ScholarPubMed
Berry, M. F., Engler, A. J., Woo, Y. J., et al. (2006). “Mesenchymal stem cell injection after myocardial infarction improves myocardial compliance.” Am J Physiol Heart Circ Physiol 290: H2196H2203.CrossRefGoogle ScholarPubMed
Bertet, C., Sulak, L. and Lecuit, T. (2004). “Myosin-dependent junction remodelling controls planar cell intercalation and axis elongation.” Nature 429: 667671.CrossRefGoogle ScholarPubMed
Blanchard, G. B., Kabla, A. J., Schultz, N. L., et al. (2009). “Tissue tectonics: morphogenetic strain rates, cell shape change and intercalation.” Nat Methods 6: 458464.CrossRefGoogle ScholarPubMed
Brehm, P., Kullberg, R. and Moody-Corbett, F. (1984). “Properties of non-junctional acetylcholine receptor channels on innervated muscle of Xenopus laevis.” J Physiol 350: 631648.CrossRefGoogle ScholarPubMed
Breitbach, M., Bostani, T., Roell, W., et al. (2007). “Potential risks of bone marrow cell transplantation into infarcted hearts.” Blood 110: 13621369.CrossRefGoogle ScholarPubMed
Brody, S., Anilkumar, T., Liliensiek, S., et al. (2006). “Characterizing nanoscale topography of the aortic heart valve basement membrane for tissue engineering heart valve scaffold design.” Tissue Eng 12: 413421.CrossRefGoogle ScholarPubMed
Candiello, J., Singh, S. S., Task, K., Kumta, P. N. and Banerjee, I. (2013). “Early differentiation patterning of mouse embryonic stem cells in response to variations in alginate substrate stiffness.” J Biol Eng 7: 9.CrossRefGoogle ScholarPubMed
Chalfie, M. (2009). “Neurosensory mechanotransduction.” Nat Rev Mol Cell Biol 10: 4452.CrossRefGoogle ScholarPubMed
Chen, C. S., Mrksich, M., Huang, S., Whitesides, G. M. and Ingber, D. E. (1997). “Geometric control of cell life and death.” Science 276: 14251428.CrossRefGoogle ScholarPubMed
Chen, W., Villa-Diaz, L. G., Sun, Y., et al. (2012). “Nanotopography influences adhesion, spreading, and self-renewal of human embryonic stem cells.” ACS Nano 6: 40944103.CrossRefGoogle ScholarPubMed
Choi, Y. S., Vincent, L. G., Lee, A. R., et al. (2012a). “Mechanical derivation of functional myotubes from adipose-derived stem cells.” Biomaterials 33: 24822491.CrossRefGoogle ScholarPubMed
Choi, Y. S., Vincent, L. G., Lee, A. R., et al. (2012b). “The alignment and fusion assembly of adipose-derived stem cells on mechanically patterned matrices.” Biomaterials 33: 69436951.CrossRefGoogle ScholarPubMed
Chowdhury, F., et al. (2010a). “Soft substrates promote homogeneous self-renewal of embryonic stem cells via downregulating cell-matrix tractions.” PLoS One 5: e15655.CrossRefGoogle ScholarPubMed
Chowdhury, F., et al. (2010b). “Material properties of the cell dictate stress-induced spreading and differentiation in embryonic stem cells.” Nat Mater 9: 8288.CrossRefGoogle ScholarPubMed
Christopherson, G. T., Song, H. and Mao, H. Q. (2009). “The influence of fiber diameter of electrospun substrates on neural stem cell differentiation and proliferation.” Biomaterials 30: 556564.CrossRefGoogle ScholarPubMed
Connelly, J. T., Gautrot, J. E., Trappmann, B., et al. (2010). “Actin and serum response factor transduce physical cues from the microenvironment to regulate epidermal stem cell fate decisions.” Nat Cell Biol 12: 711718.CrossRefGoogle ScholarPubMed
D’Amour, K. A., Agulnick, A. D., Eliazer, S., et al. (2005). “Efficient differentiation of human embryonic stem cells to definitive endoderm.” Nat Biotechnol 23: 15341541.CrossRefGoogle ScholarPubMed
Dado-Rosenfeld, D., Tzchori, I., Fine, A., Chen-Konak, L. and Levenberg, S. (2014). “Tensile forces applied on a cell-embedded three-dimensional scaffold can direct early differentiation of embryonic stem cells toward the mesoderm germ layer.” Tissue Eng Part A 21(1–2): 124143.CrossRefGoogle ScholarPubMed
Dalby, M. J., Gadegaard, N., Tare, R., et al. (2007). “The control of human mesenchymal cell differentiation using nanoscale symmetry and disorder.” Nat Mater 6: 9971003.CrossRefGoogle ScholarPubMed
Daniels, B. R., Masi, B. C. and Wirtz, D. (2006). “Probing single-cell micromechanics in vivo: the microrheology of C. elegans developing embryos.” Biophys J 90: 47126719.CrossRefGoogle ScholarPubMed
del Rio, A., Perez-Jimenez, R., Liu, R., et al. (2009). “Stretching single talin rod molecules activates vinculin binding.” Science 323: 638641.CrossRefGoogle ScholarPubMed
Deng, J., Petersen, B. E., Steindler, D. A., Jorgensen, M. L. and Laywell, E. D. (2006). “Mesenchymal stem cells spontaneously express neural proteins in culture and are neurogenic after transplantation.” Stem Cells 24: 10541064.CrossRefGoogle ScholarPubMed
Desprat, N., Supatto, W., Pouille, P. A., Beaurepaire, E. and Farge, E. (2008). “Tissue deformation modulates twist expression to determine anterior midgut differentiation in Drosophila embryos.” Dev Cell 15: 470477.CrossRefGoogle ScholarPubMed
Discher, D. E., Janmey, P., and Wang, Y. L. (2005). “Tissue cells feel and respond to the stiffness of their substrate.” Science 310: 11391143.CrossRefGoogle Scholar
Dobson, J., Cartmell, S. H., Keramane, A. and El Haj, A. J. (2006). “Principles and design of a novel magnetic force mechanical conditioning bioreactor for tissue engineering, stem cell conditioning, and dynamic in vitro screening.” IEEE Trans Nanobioscience 5: 173177.CrossRefGoogle ScholarPubMed
Downing, T. L., Soto, J., Morez, C., et al. (2013). “Biophysical regulation of epigenetic state and cell reprogramming.” Nat Mater 12: 11541162.CrossRefGoogle ScholarPubMed
Dupont, S., Morsut, L., Aragona, M., et al. (2011). “Role of YAP/TAZ in mechanotransduction.” Nature 474: 179183.CrossRefGoogle ScholarPubMed
Engler, A. J., Carag-Krieger, C., Johnson, C. P., et al. (2008). “Embryonic cardiomyocytes beat best on a matrix with heart-like elasticity: scar-like rigidity inhibits beating.” J Cell Sci 121: 37943802.CrossRefGoogle Scholar
Engler, A. J., Sen, S., Sweeney, H. L. and Discher, D. E. (2006). “Matrix elasticity directs stem cell lineage specification.” Cell 126: 677689.CrossRefGoogle ScholarPubMed
Eroshenko, N., Ramachandran, R., Yadavalli, V. K. and Rao, R. R. (2013). “Effect of substrate stiffness on early human embryonic stem cell differentiation.” J Biol Eng 7: 7.CrossRefGoogle ScholarPubMed
Evans, N. D., Minelli, C., Gentleman, E., et al. (2009). “Substrate stiffness affects early differentiation events in embryonic stem cells.” Eur Cell Mater 18: 113, discussion 1314.CrossRefGoogle ScholarPubMed
Farge, E. (2003). “Mechanical induction of Twist in the Drosophila foregut/stomodeal primordium.” Curr Biol 13: 13651377.CrossRefGoogle ScholarPubMed
Finley, M. F., Devata, S. and Huettner, J. E. (1999). “BMP-4 inhibits neural differentiation of murine embryonic stem cells.” J Neurobiol 40: 271287.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Fiorio Pla, A., Maric, D., Brazer, S. C., et al. (2005). “Canonical transient receptor potential 1 plays a role in basic fibroblast growth factor (bFGF)/FGF receptor-1-induced Ca2+ entry and embryonic rat neural stem cell proliferation.” J Neurosci 25: 26872701.Google Scholar
Flanagan, L. A., Ju, Y. E., Marg, B., Osterfield, M., and Janmey, P. A. (2002). “Neurite branching on deformable substrates.” Neuroreport 13: 24112415.CrossRefGoogle ScholarPubMed
Folkman, J., and Moscona, A. (1978). “Role of cell shape in growth control.” Nature 273: 345349.CrossRefGoogle ScholarPubMed
Forouhar, A. S., Liebling, M., Hickerson, A., et al. (2006). “The embryonic vertebrate heart tube is a dynamic suction pump.” Science 312: 751753.CrossRefGoogle ScholarPubMed
Forte, G., Carotenuto, F., Pagliari, F., et al. (2008). “Criticality of the biological and physical stimuli array inducing resident cardiac stem cell determination.” Stem Cells 26: 20932103.CrossRefGoogle ScholarPubMed
Foty, R. A. and Steinberg, M. S. (2005). “The differential adhesion hypothesis: a direct evaluation.” Dev Biol 278: 255263.CrossRefGoogle ScholarPubMed
Gang, E. J., Jeong, J. A., Hong, S. H., et al. (2004). “Skeletal myogenic differentiation of mesenchymal stem cells isolated from human umbilical cord blood.” Stem Cells 22: 617624.CrossRefGoogle ScholarPubMed
Gerecht, S., Bettinger, C. J., Zhang, Z., et al. (2007). “ The effect of actin disrupting agents on contact guidance of human embryonic stem cells.” Biomaterials 28: 40684077.CrossRefGoogle ScholarPubMed
Gilbert, P. M., Havenstrite, K. L., Magnusson, K. E., et al. (2010). “Substrate elasticity regulates skeletal muscle stem cell self-renewal in culture.” Science 329: 10781081.CrossRefGoogle ScholarPubMed
Glukhova, M. A. and Thiery, J. P. (1993). “Fibronectin and integrins in development.” Semin Cancer Biol 4: 241249.Google ScholarPubMed
Golji, J., Lam, J. and Mofrad, M. R. (2011). “Vinculin activation is necessary for complete talin binding.” Biophys J 100: 332340.CrossRefGoogle ScholarPubMed
Guharay, F. and Sachs, F. (1984). “Stretch-activated single ion channel currents in tissue-cultured embryonic chick skeletal muscle.” J Physiol 352: 685701.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: 177179.CrossRefGoogle Scholar
Hazeltine, L. B., Badur, M. G., Lian, X., et al. (2014). “Temporal impact of substrate mechanics on differentiation of human embryonic stem cells to cardiomyocytes.” Acta Biomater 10: 604612.CrossRefGoogle ScholarPubMed
Heydemann, A. and McNally, E. M. (2007). “Consequences of disrupting the dystrophin-sarcoglycan complex in cardiac and skeletal myopathy.” Trends Cardiovasc Med 17: 5559.CrossRefGoogle ScholarPubMed
Holle, A. W. and Engler, A. J. (2010). “Cell rheology: Stressed-out stem cells.” Nat Mater 9: 46.CrossRefGoogle ScholarPubMed
Holle, A. W., Tang, X., Vijayraghavan, D., et al. (2013). “In situ mechanotransduction via vinculin regulates stem cell differentiation.” Stem Cells 31: 24672477.CrossRefGoogle ScholarPubMed
Holly, S. P., Larson, M. K. and Parise, L. V. (2000). “Multiple roles of integrins in cell motility.” Exp Cell Res 261: 6974.CrossRefGoogle ScholarPubMed
Horner, V. L. and Wolfner, M. F. (2008). “Mechanical stimulation by osmotic and hydrostatic pressure activates Drosophila oocytes in vitro in a calcium-dependent manner.” Dev Biol 316: 100109.CrossRefGoogle Scholar
Hove, J. R., Koster, R. W., Forouhar, A. S., et al. (2003). “Intracardiac fluid forces are an essential epigenetic factor for embryonic cardiogenesis.” Nature 421: 172177.CrossRefGoogle ScholarPubMed
Huang, C. Y., Hagar, K. L., Frost, L. E., Sun, Y. and Cheung, H. S. (2004). “Effects of cyclic compressive loading on chondrogenesis of rabbit bone-marrow derived mesenchymal stem cells.” Stem Cells 22: 313323.CrossRefGoogle ScholarPubMed
Huang, S. and Ingber, D. E. (2005). “Cell tension, matrix mechanics, and cancer development.” Cancer Cell 8: 175176.CrossRefGoogle ScholarPubMed
Huebsch, N., Arany, P. R., Mao, A. S., et al. (2010). “Harnessing traction-mediated manipulation of the cell/matrix interface to control stem-cell fate.” Nat Mater 9: 518526.CrossRefGoogle ScholarPubMed
Ingber, D. (1991). “Extracellular matrix and cell shape: potential control points for inhibition of angiogenesis.” J Cell Biochem 47: 236241.CrossRefGoogle ScholarPubMed
Ingber, D. E. (2006). “Cellular mechanotransduction: putting all the pieces together again.” FASEB J 20: 811827.CrossRefGoogle Scholar
Jaalouk, D. E. and Lammerding, J. (2009). “Mechanotransduction gone awry.” Nat Rev Mol Cell Biol 10: 6373.CrossRefGoogle ScholarPubMed
Kahn, J., Shwartz, Y., Blitz, E., et al. (2009). “Muscle contraction is necessary to maintain joint progenitor cell fate.” Dev Cell 16: 734743.CrossRefGoogle ScholarPubMed
Kanczler, J. M., Sura, H. S., Magnay, J., et al. (2010). “Controlled differentiation of human bone marrow stromal cells using magnetic nanoparticle technology.” Tissue Eng Part A 16: 32413250.CrossRefGoogle ScholarPubMed
Katayama, Y., Battista, M., Kao, W. M., et al. (2006). “Signals from the sympathetic nervous system regulate hematopoietic stem cell egress from bone marrow.” Cell 124: 407421.CrossRefGoogle ScholarPubMed
Khetan, S., Guvendiren, M., Legant, W. R., et al. (2013). “Degradation-mediated cellular traction directs stem cell fate in covalently crosslinked three-dimensional hydrogels.” Nat Mater 12: 458465.CrossRefGoogle ScholarPubMed
Kinney, M. A., Saeed, R. and McDevitt, T. C. (2014). “Mesenchymal morphogenesis of embryonic stem cells dynamically modulates the biophysical microtissue niche.” Sci Rep 4: 4290.CrossRefGoogle ScholarPubMed
Kong, Y. P., Tu, C. H., Donovan, P. J. and Yee, A. F. (2013). “Expression of Oct4 in human embryonic stem cells is dependent on nanotopographical configuration.” Acta Biomater 9: 63696380.CrossRefGoogle ScholarPubMed
Krieg, M., Arboleda-Estudillo, Y., Puech, P. H., et al. (2008). “Tensile forces govern germ-layer organization in zebrafish.” Nat Cell Biol 10: 429636.CrossRefGoogle ScholarPubMed
Kurpinski, K., Chu, J., Hashi, C., and Li, S. (2006). “Anisotropic mechanosensing by mesenchymal stem cells.” Proc Natl Acad Sci USA 103: 1609516100.CrossRefGoogle ScholarPubMed
Lammerding, J., Kamm, R. D. and Lee, R. T. (2004). “Mechanotransduction in cardiac myocytes.” Ann N Y Acad Sci 1015: 5370.CrossRefGoogle ScholarPubMed
le Noble, F., Moyon, D., Pardanaud, L., et al. (2004). “Flow regulates arterial-venous differentiation in the chick embryo yolk sac.” Development 131: 361375.CrossRefGoogle ScholarPubMed
Lee, M. R., Kwon, K. W., Jung, H., et al. (2010). “Direct differentiation of human embryonic stem cells into selective neurons on nanoscale ridge/groove pattern arrays.” Biomaterials 31: 43604366.CrossRefGoogle ScholarPubMed
Leipzig, N. D. and Shoichet, M. S. (2009). “The effect of substrate stiffness on adult neural stem cell behavior.” Biomaterials 30: 68676878.CrossRefGoogle ScholarPubMed
Leucht, P., Kim, J. B., Currey, J. A., Brunski, J. and Helms, J. A. (2007). “FAK-Mediated mechanotransduction in skeletal regeneration.” PLoS One 2: e390.CrossRefGoogle ScholarPubMed
Li, Y., Chu, J. S., Kurpinski, K., et al. (2011). “Biophysical regulation of histone acetylation in mesenchymal stem cells.” Biophys J 100: 19021909.CrossRefGoogle ScholarPubMed
Majkut, S., Idema, T., Swift, J., et al. (2013). “Heart-specific stiffening in early embryos parallels matrix and myosin expression to optimize beating.” Curr Biol 23: 24342439.CrossRefGoogle ScholarPubMed
Mammoto, T. and Ingber, D. E. (2010). “Mechanical control of tissue and organ development.” Development 137: 14071420.CrossRefGoogle ScholarPubMed
Manasek, F. J., Burnside, M. B. and Waterman, R. E. (1972). “Myocardial cell shape change as a mechanism of embryonic heart looping.” Dev Biol 29: 349371.CrossRefGoogle ScholarPubMed
Mauck, R. L., Byers, B. A., Yuan, X. and Tuan, R. S. (2007). “Regulation of cartilaginous ECM gene transcription by chondrocytes and MSCs in 3D culture in response to dynamic loading.” Biomech Model Mechanobiol 6: 113125.CrossRefGoogle ScholarPubMed
McBeath, R., Pirone, D. M., Nelson, C. M., Bhadriraju, K. and Chen, C. S. (2004). “Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineage commitment.” Dev Cell 6: 483495.CrossRefGoogle ScholarPubMed
McMurray, R. J., Gadegaard, N., Tsimbouri, P. M., et al. (2011). “Nanoscale surfaces for the long-term maintenance of mesenchymal stem cell phenotype and multipotency.” Nat Mater 10: 637644.CrossRefGoogle ScholarPubMed
Montell, D. J. (2003). “Border-cell migration: the race is on.” Nat Rev Mol Cell Biol 4: 1324.CrossRefGoogle ScholarPubMed
Moore, K. A., Polte, T., Huang, S., et al. (2005). “Control of basement membrane remodeling and epithelial branching morphogenesis in embryonic lung by Rho and cytoskeletal tension.” Dev Dyn 232: 268281.CrossRefGoogle Scholar
Moore, S. W., Biais, N., and Sheetz, M. P. (2009). “Traction on immobilized netrin-1 is sufficient to reorient axons.” Science 325: 166.CrossRefGoogle ScholarPubMed
Moore, S. W., Keller, R. E., and Koehl, M. A. (1995). “The dorsal involuting marginal zone stiffens anisotropically during its convergent extension in the gastrula of Xenopus laevis.” Development 121: 31313140.CrossRefGoogle ScholarPubMed
Muramatsu, S., Wakabayashi, M., Ohno, T., et al. (2007). “Functional gene screening system identified TRPV4 as a regulator of chondrogenic differentiation.” J Biol Chem 282: 3215832167.CrossRefGoogle ScholarPubMed
Murphy, W. L., McDevitt, T. C. and Engler, A. J. (2014). “Materials as stem cell regulators.” Nat Mater 13(3): 547557.CrossRefGoogle ScholarPubMed
North, T. E., Goessling, W., Peeters, M., et al. (2009). “Hematopoietic stem cell development is dependent on blood flow.” Cell 137: 736748.CrossRefGoogle ScholarPubMed
Nostro, M. C., Cheng, X., Keller, G. M., and Gadue, P. (2008). “Wnt, activin, and BMP signaling regulate distinct stages in the developmental pathway from embryonic stem cells to blood.” Cell Stem Cell 2: 6071.CrossRefGoogle ScholarPubMed
Nur, E. K. A., Ahmed, I., Kamal, J., Schindler, M. and Meiners, S. (2006). “Three-dimensional nanofibrillar surfaces promote self-renewal in mouse embryonic stem cells.” Stem Cells 24: 426433.CrossRefGoogle Scholar
Oh, S., Brammer, K. S., Li, Y. S., et al. (2009). “Stem cell fate dictated solely by altered nanotube dimension.” Proc Natl Acad Sci USA 106: 21302135.CrossRefGoogle ScholarPubMed
Ohashi, N., Robling, A. G., Burr, D. B. and Turner, C. H. (2002). “The effects of dynamic axial loading on the rat growth plate.” J Bone Miner Res 17: 284292.CrossRefGoogle ScholarPubMed
Park, J., Bauer, S., von der Mark, K. and Schmuki, P. (2007). “Nanosize and vitality: TiO2 nanotube diameter directs cell fate.” Nano Lett 7: 16861691.CrossRefGoogle ScholarPubMed
Park, J. S., Chu, J. S., Cheng, C., et al. (2004). “Differential effects of equiaxial and uniaxial strain on mesenchymal stem cells.” Biotechnol Bioeng 88: 359868.CrossRefGoogle ScholarPubMed
Pasapera, A. M., Schneider, I. C., Rericha, E., Schlaepfer, D. D. and Waterman, C. M. (2010). “Myosin II activity regulates vinculin recruitment to focal adhesions through FAK-mediated paxillin phosphorylation.” J Cell Biol 188: 877890.CrossRefGoogle ScholarPubMed
Paszek, M. J., Zahir, N., Johnson, K. R., et al. (2005). “Tensional homeostasis and the malignant phenotype.” Cancer Cell 8: 241254.CrossRefGoogle ScholarPubMed
Pittenger, M. F., Mackay, A. M., Beck, S. C., et al. (1999). “ Multilineage potential of adult human mesenchymal stem cells.” Science 284: 143147.CrossRefGoogle ScholarPubMed
Poteser, M., Graziani, A., Eder, P., et al. (2008). “Identification of a rare subset of adipose tissue-resident progenitor cells, which express CD133 and TRPC3 as a VEGF-regulated Ca2+ entry channel.” FEBS Lett 582: 26962702.CrossRefGoogle ScholarPubMed
Roux, P. P. and Blenis, J. (2004). “ERK and p38 MAPK-activated protein kinases: a family of protein kinases with diverse biological functions.” Microbiol Mol Biol Rev 68: 320344.CrossRefGoogle ScholarPubMed
Sachs, F. (2010). “Stretch-activated ion channels: what are they?Physiology (Bethesda) 25: 5056.Google Scholar
Saha, K., Keung, A. J., Irwin, E. F., et al. (2008). “Substrate modulus directs neural stem cell behavior.” Biophys J 95: 44264438.CrossRefGoogle ScholarPubMed
Saha, S., Ji, L., de Pablo, J. J. and Palecek, S. P. (2006). “Inhibition of human embryonic stem cell differentiation by mechanical strain.” J Cell Physiol 206: 126137.CrossRefGoogle ScholarPubMed
Sanders, M. C., Way, M., Sakai, J. and Matsudaira, P. (1996). “Characterization of the actin cross-linking properties of the scruin-calmodulin complex from the acrosomal process of Limulus sperm.” J Biol Chem 271: 26512657.CrossRefGoogle ScholarPubMed
Sawada, Y., Tamada, M., Dubin-Thaler, B. J., et al. (2006). “Force sensing by mechanical extension of the Src family kinase substrate p130Cas.” Cell 127: 10151026.CrossRefGoogle ScholarPubMed
Schindler, M., Ahmed, I., Kamal, J., et al. (2005). “A synthetic nanofibrillar matrix promotes in vivo-like organization and morphogenesis for cells in culture.” Biomaterials 26: 56245631.CrossRefGoogle ScholarPubMed
Schwartz, M. A. (2010). “Integrins and extracellular matrix in mechanotransduction.” Cold Spring Harb Perspect Biol 2: a005066.CrossRefGoogle ScholarPubMed
Shih, Y. R., Tseng, K. F., Lai, H. Y., Lin, C. H. and Lee, O. K. (2005). “Matrix stiffness regulation of integrin-mediated mechanotransduction during osteogenic differentiation of human mesenchymal stem cells.” J Bone Miner Res 26: 730738.CrossRefGoogle Scholar
Shimizu, N., Yamamoto, K., Obi, S., et al. (2008). “Cyclic strain induces mouse embryonic stem cell differentiation into vascular smooth muscle cells by activating PDGF receptor beta.” J Appl Physiol (1985) 104: 766772.CrossRefGoogle ScholarPubMed
Shin, J. H., Tam, B. K., Brau, R. R., et al. (2007). “Force of an actin spring.” Biophys J 92: 37298733.CrossRefGoogle ScholarPubMed
Smith, L. A., Liu, X., Hu, J. and Ma, P. X. (2009). “The influence of three-dimensional nanofibrous scaffolds on the osteogenic differentiation of embryonic stem cells.” Biomaterials 30: 25162522.CrossRefGoogle ScholarPubMed
Smith, L. A., Liu, X., Hu, J. and Ma, P. X. (2010). “The enhancement of human embryonic stem cell osteogenic differentiation with nano-fibrous scaffolding.” Biomaterials 31: 55265535.CrossRefGoogle ScholarPubMed
Steward, A. J., Thorpe, S. D., Vinardell, T., et al. (2012). “Cell-matrix interactions regulate mesenchymal stem cell response to hydrostatic pressure.” Acta Biomater 8: 21532159.CrossRefGoogle ScholarPubMed
Stokes, I. A., Mente, P. L., Iatridis, J. C., Farnum, C. E. and Aronsson, D. D. (2002). “Growth plate chondrocyte enlargement modulated by mechanical loading.” Stud Health Technol Inform 88: 378381.Google ScholarPubMed
Stolberg, S. and McCloskey, K. E. (2009). “Can shear stress direct stem cell fate?Biotechnol Prog 25: 1019.CrossRefGoogle ScholarPubMed
Suresh, S. (2007). “Biomechanics and biophysics of cancer cells.” Acta Biomater 3: 413438.CrossRefGoogle ScholarPubMed
Swift, J., Ivanovska, I. L., Buxboim, A., et al. (2013). “Nuclear lamin-A scales with tissue stiffness and enhances matrix-directed differentiation.” Science 341: 1240104.CrossRefGoogle ScholarPubMed
Tai, Y., Feng, S., Du, W. and Wang, Y. (2009). “Functional roles of TRPC channels in the developing brain.” Pflugers Arch 458: 239283.CrossRefGoogle ScholarPubMed
Takito, J. and Al-Awqati, Q. (2009). “Conversion of ES cells to columnar epithelia by hensin and to squamous epithelia by laminin.” J Cell Biol 166: 10931102.CrossRefGoogle Scholar
Tay, C. Y., Yu, H., Pal, M., et al. (2010). “Micropatterned matrix directs differentiation of human mesenchymal stem cells towards myocardial lineage.” Exp Cell Res 316: 11591168.CrossRefGoogle ScholarPubMed
Taylor-Weiner, H., Schwarzbauer, J. E. and Engler, A. J. (2013).“Defined extracellular matrix components are necessary for definitive endoderm induction.” Stem Cells 31: 20842094.CrossRefGoogle ScholarPubMed
Teo, B. K., Wong, S. T., Lim, C. K., et al. (2013). “Nanotopography modulates mechanotransduction of stem cells and induces differentiation through focal adhesion kinase.” ACS Nano 7: 47854798.CrossRefGoogle ScholarPubMed
Tornillo, G., Elia, A. R., Castellano, I., et al. (2013). “p130Cas alters the differentiation potential of mammary luminal progenitors by deregulating c-Kit activity.” Stem Cells 31: 14221433.CrossRefGoogle ScholarPubMed
Toyama, Y., Peralta, X. G., Wells, A. R., Kiehart, D. P. and Edwards, G. S. (2008). “Apoptotic force and tissue dynamics during Drosophila embryogenesis.” Science 321: 16831686.CrossRefGoogle ScholarPubMed
Trappmann, B., Gautrot, J. E., Connelly, J. T., et al. (2012). “Extracellular-matrix tethering regulates stem-cell fate.” Nat Mater 11: 642649.CrossRefGoogle ScholarPubMed
Tse, J. R. and Engler, A. J. (2010). “Preparation of hydrogel substrates with tunable mechanical properties.” Curr Protoc Cell Biol Chapter 10: Unit 10. 16.Google Scholar
Tse, J. R. and Engler, A. J. (2011). “Stiffness gradients mimicking in vivo tissue variation regulate mesenchymal stem cell fate.” PLoS One 6: e15978.CrossRefGoogle ScholarPubMed
Voronov, D. A., Alford, P. W., Xu, G. and Taber, L. A. (2004). “The role of mechanical forces in dextral rotation during cardiac looping in the chick embryo.” Dev Biol 272: 339350.CrossRefGoogle ScholarPubMed
Wang, Y. and Riechmann, V. (2007). “The role of the actomyosin cytoskeleton in coordination of tissue growth during Drosophila oogenesis.” Curr Biol 17: 13491355.CrossRefGoogle ScholarPubMed
Wen, J. H., Vincent, L. G., Fuhrmann, A., et al. (2014). “Interplay of matrix stiffness and protein tethering in stem cell differentiation.” Nat Mater 13(10): 979987.CrossRefGoogle ScholarPubMed
Wilson, N. R., Ty, M. T., Ingber, D. E., Sur, M. and Liu, G. (2007). “Synaptic reorganization in scaled networks of controlled size.” J Neurosci 27: 1358113589.CrossRefGoogle ScholarPubMed
Wozniak, M. A. and Chen, C. S. (2014). “Mechanotransduction in development: a growing role for contractility.” Nat Rev Mol Cell Biol 10: 3443.CrossRefGoogle Scholar
Yang, Y., Beqaj, S., Kemp, P., Ariel, I. and Schuger, L. (2000). “Stretch-induced alternative splicing of serum response factor promotes bronchial myogenesis and is defective in lung hypoplasia.” J Clin Invest 106: 13211330.CrossRefGoogle ScholarPubMed
Yim, E. K., Pang, S. W. and Leong, K. W. (2007). “Synthetic nanostructures inducing differentiation of human mesenchymal stem cells into neuronal lineage.” Exp Cell Res 313: 18201829.CrossRefGoogle ScholarPubMed
Young, D. A., Choi, Y. S., Engler, A. J. and Christman, K. L. (2013). “Stimulation of adipogenesis of adult adipose-derived stem cells using substrates that mimic the stiffness of adipose tissue.” Biomaterials 34: 85818588.CrossRefGoogle ScholarPubMed
Young, J. L. and Engler, A. J. (2011). “Hydrogels with time-dependent material properties enhance cardiomyocyte differentiation in vitro.” Biomaterials 32: 10021009.CrossRefGoogle ScholarPubMed
Zamir, E. A. and Taber, L. A. (2004a). “Material properties and residual stress in the stage 12 chick heart during cardiac looping.” J Biomech Eng 126: 823830.CrossRefGoogle ScholarPubMed
Zamir, E. A. and Taber, L. A. (2004b). “On the effects of residual stress in microindentation tests of soft tissue structures.” J Biomech Eng 126: 276283.CrossRefGoogle ScholarPubMed
Zhao, L., Liu, L., Wu, Z., Zhang, Y. and Chu, P. K. (2012). “Effects of micropitted/nanotubular titania topographies on bone mesenchymal stem cell osteogenic differentiation.” Biomaterials 33: 26292641.CrossRefGoogle ScholarPubMed
Zoldan, J., Karagiannis, E. D., Lee, C. Y., et al. (2011). “The influence of scaffold elasticity on germ layer specification of human embryonic stem cells.” Biomaterials 32: 96129621.CrossRefGoogle ScholarPubMed

References

Abaci, H. E., Devendra, R., Soman, R., Drazer, G. and Gerecht, S. (2012). “Microbioreactors to manipulate oxygen tension and shear stress in the microenvironment of vascular stem and progenitor cells.” Biotechnol Appl Biochem 59: 97105.CrossRefGoogle ScholarPubMed
Ballotta, V., Driessen-Mol, A., Bouten, C. V. and Baaijens, F. P. (2014). “Strain-dependent modulation of macrophage polarization within scaffolds.” Biomaterials 35: 49194928.CrossRefGoogle ScholarPubMed
Battiston, K. G. (2015). Evaluating the Use of Monocytes with a Degradable Polyurethane for Vascular Tissue Regeneration. PhD dissertation, University of Toronto.Google Scholar
Brown, X. Q., Bartolak-Suki, E., Williams, C., Walker, M. L., Weaver, V. M. and Wong, J. Y. (2010). “Effect of substrate stiffness and PDGF on the behavior of vascular smooth muscle cells: implications for atherosclerosis.” J Cell Physiol 225: 115122.CrossRefGoogle ScholarPubMed
Brown, X. Q., Ookawa, K. and Wong, J. Y. (2005). “Evaluation of polydimethylsiloxane scaffolds with physiologically-relevant elastic moduli: interplay of substrate mechanics and surface chemistry effects on vascular smooth muscle cell response.” Biomaterials 26: 31233129.CrossRefGoogle ScholarPubMed
Crapo, P. M. and Wang, Y. (2011). “Hydrostatic pressure independently increases elastin and collagen co-expression in small-diameter engineered arterial constructs.” J Biomed Mater Res A 96: 673681.CrossRefGoogle ScholarPubMed
Engbers-Buijtenhuijs, P., Buttafoco, L., Poot, A. A., Dijkstra, P. J., de Vos, R. A., Sterk, L. M., Geelkerken, R. H., et al. (2006). “Biological characterisation of vascular grafts cultured in a bioreactor.” Biomaterials 27: 23902397.CrossRefGoogle Scholar
Fitzgerald, T. N., Shepherd, B. R., Asada, H., Teso, D., Muto, A., Fancher, T., Pimiento, J. M., et al. (2008). “Laminar shear stress stimulates vascular smooth muscle cell apoptosis via the Akt pathway.” J Cell Physiol 216: 389395.CrossRefGoogle ScholarPubMed
Hagerty, R. D., Salzmann, D. L., Kleinert, L. B. and Williams, S. K. (2010). “Cellular proliferation and macrophage populations associated with implanted expanded polytetrafluoroethylene and polyethyleneterephthalate.” J Biomed Mater Res 49: 489497.3.0.CO;2-2>CrossRefGoogle Scholar
Hahn, M. S., Mchale, M. K., Wang, E., Schmedlen, R. H. and West, J. L. (2007). “Physiologic pulsatile flow bioreactor conditioning of poly(ethylene glycol)-based tissue engineered vascular grafts.” Ann Biomed Eng 35: 190200.CrossRefGoogle ScholarPubMed
Hibino, N., Yi, T., Duncan, D. R., Rathore, A., Dean, E., Naito, Y., Dardik, A., et al. (2011). “A critical role for macrophages in neovessel formation and the development of stenosis in tissue-engineered vascular grafts.” FASEB J 25: 42534263.CrossRefGoogle ScholarPubMed
Huang, A. H. and Niklason, L. E. (2014). “Engineering of arteries in vitro.” Cell Mol Life Sci 71: 21032118.CrossRefGoogle ScholarPubMed
Isenberg, B. C. and Tranquillo, R. T. (2003). “Long-term cyclic distention enhances the mechanical properties of collagen-based media-equivalents.” Ann Biomed Eng 31: 937949.CrossRefGoogle ScholarPubMed
Isenberg, B. C. Williams, C. and Tranquillo, R. T. (2006). “Endothelialization and flow conditioning of fibrin-based media-equivalents.” Ann Biomed Eng 34: 971985.CrossRefGoogle ScholarPubMed
Jeong, S. I., Kwon, J. H., Lim, J. I., Cho, S. W., Jung, Y., Sung, W. J., Kim, S. H., et al. (2005). “Mechano-active tissue engineering of vascular smooth muscle using pulsatile perfusion bioreactors and elastic PLCL scaffolds.” Biomaterials 26: 14051411.CrossRefGoogle ScholarPubMed
Laflamme, K., Roberge, C. J., Pouliot, S., D’orleans-Juste, P., Auger, F. A. and Germain, L. (2006). “Tissue-engineered human vascular media produced in vitro by the self-assembly approach present functional properties similar to those of their native blood vessels.” Tissue Eng 12: 22752281.CrossRefGoogle ScholarPubMed
Lehoux, S., Castier, Y. and Tedgui, A. (2006). “Molecular mechanisms of the vascular responses to haemodynamic forces.” J Intern Med 259: 381392.CrossRefGoogle ScholarPubMed
Lewis, J. S., Dolgova, N. V., Chancellor, T. J., Acharya, A. P., Karpiak, J. V., Lele, T. P. and Keselowsky, B. G. (2013). “The effect of cyclic mechanical strain on activation of dendritic cells cultured on adhesive substrates.” Biomaterials 34: 90639070.CrossRefGoogle ScholarPubMed
Li, J., Zhang, K., Yang, P., Liao, Y., Wu, L., Chen, J., Zhao, A., et al. (2013). “Research of smooth muscle cells response to fluid flow shear stress by hyaluronic acid micro-pattern on a titanium surface.” Exp Cell Res 319: 26632672.CrossRefGoogle ScholarPubMed
Lin, Y. C., Kramer, C. M., Chen, C. S. and Reich, D. H. (2012). “Probing cellular traction forces with magnetic nanowires and microfabricated force sensor arrays.” Nanotechnology 23: 075101.CrossRefGoogle ScholarPubMed
Liu, H., Usprech, J., Sun, Y. and Simmons, C. A. (in press). “A microfabricated platform with hydrogel arrays for 3D mechanical stimulation of cells.” Acta Biomaterialia.Google Scholar
Mann, J. M., Lam, R. H., Weng, S., Sun, Y. and Fu, J. (2012). “A silicone-based stretchable micropost array membrane for monitoring live-cell subcellular cytoskeletal response.” Lab Chip 12: 731740.CrossRefGoogle ScholarPubMed
Matheson, L. A., Maksym, G. N., Santerre, J. P. and Labow, R. S. (2006). “The functional response of U937 macrophage-like cells is modulated by extracellular matrix proteins and mechanical strain.” Biochem Cell Biol 84: 763773.CrossRefGoogle ScholarPubMed
Matheson, L. A., Maksym, G. N., Santerre, J. P. and Labow, R. S. (2007). “Differential effects of uniaxial and biaxial strain on U937 macrophage-like cell morphology: influence of extracellular matrix type proteins.” J Biomed Mater Res A 81: 971981.CrossRefGoogle ScholarPubMed
Mcallister, T. N., Maruszewski, M., Garrido, S. A., Wystrychowski, W., Dusserre, N., Marini, A., Zagalski, K., et al. (2009). “Effectiveness of haemodialysis access with an autologous tissue-engineered vascular graft: a multicentre cohort study.” Lancet 373: 14401446.CrossRefGoogle ScholarPubMed
Miyazaki, H. and Hayashi, K. (2001). “Effects of cyclic strain on the morphology and phagocytosis of macrophages.” Biomed Mater Eng 11: 301309.Google ScholarPubMed
Moraes, C., Likhitpanichkul, M., Lam, C. J., Beca, B. M., Sun, Y. and Simmons, C. A. (2013). “Microdevice array-based identification of distinct mechanobiological response profiles in layer-specific valve interstitial cells.” Integr Biol (Camb) 5(4): 673680.CrossRefGoogle ScholarPubMed
Niklason, L. E., Gao, J., Abbott, W. M., Hirschi, K. K., Houser, S., Marini, R. and Langer, R. (1999). “Functional arteries grown in vitro.” Science 284: 489493.CrossRefGoogle ScholarPubMed
Opitz, F., Schenke-Layland, K., Richter, W., Martin, D. P., Degenkolbe, I., Wahlers, T. and Stock, U. A. (2004). “Tissue engineering of ovine aortic blood vessel substitutes using applied shear stress and enzymatically derived vascular smooth muscle cells.” Ann Biomed Eng 32: 212222.CrossRefGoogle ScholarPubMed
Peyton, S. R. and Putnam, A. J. (2005). “Extracellular matrix rigidity governs smooth muscle cell motility in a biphasic fashion.” J Cell Physiol 204: 198209.CrossRefGoogle Scholar
Peyton, S. R., Raub, C. B., Keschrumrus, V. P. and Putnam, A. J. (2006). “The use of poly(ethylene glycol) hydrogels to investigate the impact of ECM chemistry and mechanics on smooth muscle cells.” Biomaterials 27: 48814893.CrossRefGoogle ScholarPubMed
Redmond, E. M., Cullen, J. P., Cahill, P. A., Sitzmann, J. V., Stefansson, S., Lawrence, D. A. and Okada, S. S. (2001). “Endothelial cells inhibit flow-induced smooth muscle cell migration: role of plasminogen activator inhibitor-1.” Circulation 103: 597603.CrossRefGoogle ScholarPubMed
Riehl, B. D., Park, J. H., Kwon, I. K. and Lim, J. Y. (2012). “Mechanical stretching for tissue engineering: two-dimensional and three-dimensional constructs.” Tissue Eng Part B Rev 18: 288300.CrossRefGoogle ScholarPubMed
Robinson, K. G., Nie, T., Baldwin, A. D., Yang, E. C., Kiick, K. L. and Akins, R. E. Jr. (2012). “Differential effects of substrate modulus on human vascular endothelial, smooth muscle, and fibroblastic cells.” J Biomed Mater Res A 100: 13561367.CrossRefGoogle ScholarPubMed
Rosati, C. and Garay, R. (1991). “Flow-dependent stimulation of sodium and cholesterol uptake and cell growth in cultured vascular smooth muscle.” J Hypertens 9: 10291033.CrossRefGoogle ScholarPubMed
Sakamoto, H., Aikawa, M., Hill, C. C., Weiss, D., Taylor, W. R., Libby, P. and Lee, R. T. (2011). “Biomechanical strain induces class a scavenger receptor expression in human monocyte/macrophages and THP-1 cells: a potential mechanism of increased atherosclerosis in hypertension.” Circulation 104: 109114.CrossRefGoogle Scholar
Sazonova, O. V., Lee, K. L., Isenberg, B. C., Rich, C. B., Nugent, M. A. and Wong, J. Y. (2011). “Cell-cell interactions mediate the response of vascular smooth muscle cells to substrate stiffness.” Biophys J 101: 622630.CrossRefGoogle ScholarPubMed
Seifu, D. G., Purnama, A., Mequanint, K. and Mantovani, D. (2013). “Small-diameter vascular tissue engineering.” Nat Rev Cardiol 10: 410421.CrossRefGoogle ScholarPubMed
Seliktar, D., Black, R. A., Vito, R. P. and Nerem, R. M. (2000). “Dynamic mechanical conditioning of collagen-gel blood vessel constructs induces remodeling in vitro.” Ann Biomed Eng 28: 351362.CrossRefGoogle ScholarPubMed
Shi, Z. D. and Tarbell, J. M. (2011). “Fluid flow mechanotransduction in vascular smooth muscle cells and fibroblasts.” Ann Biomed Eng 39: 16081619.CrossRefGoogle Scholar
Shigematsu, K., Yasuhara, H., Shigematsu, H. and Muto, T. (2000). “Direct and indirect effects of pulsatile shear stress on the smooth muscle cell.” Int Angiol 19: 3946.Google ScholarPubMed
Song, Y., Wennink, J. W., Kamphuis, M. M., Sterk, L. M., Vermes, I., Poot, A. A., Feijen, J. et al. (2011). “Dynamic culturing of smooth muscle cells in tubular poly(trimethylene carbonate) scaffolds for vascular tissue engineering.” Tissue Eng Part A 17: 381387.CrossRefGoogle ScholarPubMed
Stegemann, J. P. and Nerem, R. M. (2003). “Phenotype modulation in vascular tissue engineering using biochemical and mechanical stimulation.” Ann Biomed Eng 31: 391402.CrossRefGoogle ScholarPubMed
Syedain, Z. H. and Tranquillo, R. T. (2011). “TGF-beta1 diminishes collagen production during long-term cyclic stretching of engineered connective tissue: implication of decreased ERK signaling.” J Biomech 44: 848855.CrossRefGoogle ScholarPubMed
Syedain, Z. H., Weinberg, J. S. and Tranquillo, R. T. (2008). “Cyclic distension of fibrin-based tissue constructs: evidence of adaptation during growth of engineered connective tissue.” Proc Natl Acad Sci USA 105: 65376542.CrossRefGoogle ScholarPubMed
Tan, W., Scott, D., Belchenko, D., Qi, H. J. and Xiao, L. (2008). “Development and evaluation of microdevices for studying anisotropic biaxial cyclic stretch on cells.” Biomedical Microdevices 10: 869882.CrossRefGoogle ScholarPubMed
Vara, D. S., Punshon, G., Sales, K. M., Hamilton, G. and Seifalian, A. M. (2011). “Haemodynamic regulation of gene expression in vascular tissue engineering.” Curr Vasc Pharmacol 9: 167187.CrossRefGoogle ScholarPubMed
Wagenseil, J. E. and Mecham, R. P. (2009). “Vascular extracellular matrix and arterial mechanics.” Physiol Rev 89: 957989.CrossRefGoogle ScholarPubMed
Wang, C., Cen, L., Yin, S., Liu, Q., Liu, W., Cao, Y. and Cui, L. (2010). “A small diameter elastic blood vessel wall prepared under pulsatile conditions from polyglycolic acid mesh and smooth muscle cells differentiated from adipose-derived stem cells.” Biomaterials 31: 621630.CrossRefGoogle ScholarPubMed
Wayman, B. H., Taylor, W. R., Rachev, A. and Vito, R. P. (2008). “Arteries respond to independent control of circumferential and shear stress in organ culture.” Ann Biomed Eng 36: 673684.CrossRefGoogle ScholarPubMed
Wilson, E., Sudhir, K. and Ives, H. E. (1995). “Mechanical strain of rat vascular smooth muscle cells is sensed by specific extracellular matrix/integrin interactions.” Journal of Clinical Investigation 96: 23642372.CrossRefGoogle ScholarPubMed
Xu, J., Ge, H., Zhou, X., Yang, D., Guo, T., He, J., Li, Q., et al. (2005). “Tissue-engineered vessel strengthens quickly under physiological deformation: application of a new perfusion bioreactor with machine vision.” J Vasc Res 42: 503508.CrossRefGoogle ScholarPubMed
Yeh, C. H., Tsai, S. H., Wu, L. W. and Lin, Y. C. (2011). “Using a co-culture microsystem for cell migration under fluid shear stress.” Lab Chip 11: 25832590.CrossRefGoogle ScholarPubMed
Zhang, X., Wang, X., Keshav, V., Johanas, J. T., Leisk, G. G. and Kaplan, D. L. (2009). “Dynamic culture conditions to generate silk-based tissue-engineered vascular grafts.” Biomaterials 30: 32133223.CrossRefGoogle ScholarPubMed
Zilla, P., Bezuidenhout, D. and Human, P. (2007). “Prosthetic vascular grafts: wrong models, wrong questions and no healing.” Biomaterials 28: 50095027.CrossRefGoogle ScholarPubMed

References

Ajubi, N. E., Klein-Nulend, J., Nijweide, P. J., Vrijheid-Lammers, T., Alblas, M. J. and Burger, E. H. 1996. “Pulsating fluid flow increases prostaglandin production by cultured chicken osteocytes–a cytoskeleton-dependent process.” Biochemical and Biophysical Research Communications 225: 6268.CrossRefGoogle ScholarPubMed
Bergmann, P. and Schoutens, A. 1995. “Prostaglandins and bone.” Bone 16: 485488.Google ScholarPubMed
Bonewald, L. F. and Johnson, M. L. 2008. “Osteocytes, mechanosensing and Wnt signaling.” Bone 42: 606–15.CrossRefGoogle ScholarPubMed
Boyle, W. J., Simonet, W. S. and Lacey, D. L. 2003. “Osteoclast differentiation and activation.” Nature 423: 337342.CrossRefGoogle ScholarPubMed
Burr, D. 2002. “Targeted and nontargeted remodeling.” Bone 30: 24.CrossRefGoogle ScholarPubMed
Burr, D. B., Forwood, M. R., Fyhrie, D. P., Martin, R. B., Schaffler, M. B. and Turner, C. H. 1997. “Bone microdamage and skeletal fragility in osteoporotic and stress fractures.” Journal of Bone and Mineral Research 12: 615.CrossRefGoogle ScholarPubMed
Cane, V., Marotti, G., Volpi, G., Zaffe, D., Palazzini, S., Remaggi, F. and Muglia, M. A. 1982. “Size and density of osteocyte lacunae in different regions of long bones.” Calcified Tissue International 34: 558563.CrossRefGoogle ScholarPubMed
Chen, J.-H., Liu, C., You, L. and Simmons, C. A. 2010. “Boning up on Wolff’s Law: mechanical regulation of the cells that make and maintain bone.” Journal of Biomechanics 43: 108118.CrossRefGoogle ScholarPubMed
Cheung, W. Y., Liu, C., Tonelli-Zasarsky, R. M. L., Simmons, C. A. and You, L. 2011. “Osteocyte apoptosis is mechanically regulated and induces angiogenesis in vitro.” Journal of Orthopaedic Research 29: 523530.CrossRefGoogle ScholarPubMed
Cowin, S. C., Gailani, G. and Benalla, M. 2009. “Hierarchical poroelasticity: movement of interstitial fluid between porosity levels in bones.” Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 367: 34013444.CrossRefGoogle ScholarPubMed
Doty, S. B. 1981. “Morphological evidence of gap junctions between bone cells.” Calcified Tissue International 33: 509512.CrossRefGoogle ScholarPubMed
Feng, X., Du, W., Luo, Q. and Liu, B. F. 2009. “Microfluidic chip: next-generation platform for systems biology.” Anal Chim Acta 650: 8397.CrossRefGoogle ScholarPubMed
Fox, S., Chambers, T. and Chow, J. 1996. “Nitric oxide is an early mediator of the increase in bone formation by mechanical stimulation.” American Journal of Physiology-Endocrinology and Metabolism 33: E955.CrossRefGoogle Scholar
Fratzl, P., Gupta, H. S., Paschalis, E. P. and Roschger, P. 2004. “Structure and mechanical quality of the collagen-mineral nano-composite in bone.” Journal of Materials Chemistry 14: 21152123.CrossRefGoogle Scholar
Fritton, S. P., McLeod, K. J. and Rubin, C. T. 2000. “Quantifying the strain history of bone: spatial uniformity and self-similarity of low-magnitude strains.” Journal of Biomechanics 33: 317325.CrossRefGoogle ScholarPubMed
Frost, H. M. 1990. “Skeletal structural adaptations to mechanical usage (SATMU): 1. Redefining Wolff’s law: the bone modeling problem.” Anatomical Record 226: 403413.CrossRefGoogle ScholarPubMed
Fukumoto, S. and Martin, T. J. 2009. “Bone as an endocrine organ.” Trends in Endocrinology and Metabolism 20: 230236.CrossRefGoogle ScholarPubMed
Gailani, G. B., Benalla, M., Mahamud, R., Cowin, S. C. and Cardoso, L. L. 2009. “Experimental protocol for the measurement of the permeability of a single osteon.” J Biomech Eng 131: 101007.CrossRefGoogle Scholar
Gardinier, J. D., Majumdar, S., Duncan, R. L. and Wang, L. 2009a. “Cyclic hydraulic pressure and fluid flow differentially modulate cytoskeleton re-organization in MC3T3 osteoblasts.” Cellular and Molecular Bioengineering 2: 133143.CrossRefGoogle ScholarPubMed
Gardinier, J. D., Majumdar, S., Duncan, R. L. and Wang, L. 2009b. “Cyclic hydraulic pressure and fluid flow differentially modulate cytoskeleton re-organization in MC3T3 osteoblasts.” Cell Mol Bioeng 2: 133143.CrossRefGoogle ScholarPubMed
Gardinier, J. D., Townend, C. W., Jen, K.-P., Wu, Q., Duncan, R. L. and Wang, L. 2010. “In situ permeability measurement of the mammalian lacunar-canalicular system.” Bone 46: 10751081.CrossRefGoogle ScholarPubMed
Genetos, D. C., Geist, D. J., Liu, D., Donahue, H. J. and Duncan, R. L. 2005. “Fluid shear-Induced ATP secretion mediates prostaglandin release in MC3T3-E1 osteoblasts.” Journal of Bone and Mineral Research 20.CrossRefGoogle ScholarPubMed
Genetos, D. C., Kephart, C. J., Zhang, Y., Yellowley, C. E. and Donahue, H. J. 2007. “Oscillating fluid flow activation of gap junction hemichannels induces ATP release from MLO-Y4 osteocytes.” Journal of Cellular Physiology 212: 207214.CrossRefGoogle ScholarPubMed
Guo, X. E., Takai, E., Jiang, X., Xu, Q., Whitesides, G. M., Yardley, J. T., Hung, C. T., et al. 2006. “Intracellular calcium waves in bone cell networks under single cell nanoindentation.” MCB Molecular and Cellular Biomechanics 3: 95107.Google ScholarPubMed
Hauge, E. M., Qvesel, D., Eriksen, E. F., Mosekilde, L. and Melsen, F. 2001. “Cancellous bone remodeling occurs in specialized compartments lined by cells expressing osteoblastic markers.” Journal of Bone and Mineral Research 16: 15751582.CrossRefGoogle ScholarPubMed
Hofbauer, L. C., Khosla, S., Dunstan, C. R., Lacey, D. L., Boyle, W. J. and Riggs, B. L. 2000. “The roles of osteoprotegerin and osteoprotegerin ligand in the paracrine regulation of bone resorption.” Journal of Bone and Mineral Research 15: 212.CrossRefGoogle ScholarPubMed
Hung, C. T., Pollack, S. R., Reilly, T. M. and Brighton, C. T. 1995. “Real-time calcium response of cultured bone cells to fluid flow.” Clinical Orthopaedics and Related Research: 256269.Google Scholar
Huo, B., Lu, X. L., Costa, K. D., Xu, Q. and Guo, X. E. 2010. “An ATP-dependent mechanism mediates intercellular calcium signaling in bone cell network under single cell nanoindentation.” Cell Calcium 47: 234–41.CrossRefGoogle ScholarPubMed
Jacobs, C. R., Yellowley, C. E., Davis, B. R., Zhou, Z., Cimbala, J. M. and Donahue, H. J. 1998. “Differential effect of steady versus oscillating flow on bone cells.” Journal of Biomechanics 31: 969976.CrossRefGoogle ScholarPubMed
Johnson, D. L., Mcallister, T. N. and Frangos, J. A. 1996. “Fluid flow stimulates rapid and continuous release of nitric oxide in osteoblasts.” American Journal of Physiology-Endocrinology And Metabolism 34: E205.CrossRefGoogle Scholar
Kamel, M. A., Picconi, J. L., Lara-Castillo, N. and Johnson, M. L. 2010. “Activation of β-catenin signaling in MLO-Y4 osteocytic cells versus 2T3 osteoblastic cells by fluid flow shear stress and PGE2: Implications for the study of mechanosensation in bone.” Bone 47: 872881.CrossRefGoogle Scholar
Kamioka, H., Sugawara, Y., Murshid, S. A., Ishihara, Y., Honjo, T. and Takano-Yamamoto, T. 2006. “Fluid shear stress induces less calcium response in a single primary osteocyte than in a single osteoblast: implication of different focal adhesion formation.” Journal of Bone and Mineral Research 21: 10121021.CrossRefGoogle Scholar
Kato, Y., Windle, J. J., Koop, B. A., Mundy, G. R. and Bonewald, L. F. 1997. “Establishment of an osteocyte-like cell line, MLO-Y4.” Journal of Bone and Mineral Research 12: 20142023.CrossRefGoogle ScholarPubMed
Kennedy, O. D., Herman, B. C., Laudier, D. M., Majeska, R. J., Sun, H. B. and Schaffler, M. B. 2012. “Activation of resorption in fatigue-loaded bone involves both apoptosis and active pro-osteoclastogenic signaling by distinct osteocyte populations.” Bone 50: 11151122.CrossRefGoogle ScholarPubMed
Kidd, L. J., Stephens, A. S., Kuliwaba, J. S., Fazzalari, N. L., Wu, A. C. K. and Forwood, M. R. 2010. “Temporal pattern of gene expression and histology of stress fracture healing.” Bone 46: 369378.CrossRefGoogle ScholarPubMed
King, G. J. and Holtrop, M. E. 1975. “Actin-like filaments in bone cells of cultured mouse calvaria as demonstrated by binding to heavy meromyosin.” Journal of Cell Biology 66: 445451.CrossRefGoogle ScholarPubMed
Klein-Nulend, J., van der Plas, A., Semeins, C. M., Ajubi, N. E., Frangos, J. A., Nijweide, P. J. and Burger, E. H. 1995. “Sensitivity of osteocytes to biomechanical stress in vitro.” FASEB Journal 9: 441445.CrossRefGoogle ScholarPubMed
Kleinnulend, J., Semeins, C., Ajubi, N., Nijweide, P. and Burger, E. 1995. “Pulsating fluid flow increases nitric oxide (NO) synthesis by osteocytes but not periosteal fibroblasts-correlation with prostaglandin upregulation.” Biochemical and Biophysical Research Communications 217: 640648.CrossRefGoogle Scholar
Knothe Tate, M. L. and Knothe, U. 2000. “An ex vivo model to study transport processes and fluid flow in loaded bone.” Journal of Biomechanics 33: 247254.CrossRefGoogle Scholar
Knothe Tate, M. L., Steck, R., Forwood, M. R. and Niederer, P. 2000. “In vivo demonstration of load-induced fluid flow in the rat tibia and its potential implications for processes associated with functional adaptation.” Journal of Experimental Biology 203: 27372745.CrossRefGoogle ScholarPubMed
Kou, S., Pan, L., van Noort, D., Meng, G., Wu, X., Sun, H., Xu, J. and Lee, I. 2011. “A multishear microfluidic device for quantitative analysis of calcium dynamics in osteoblasts.” Biochem Biophys Res Commun 408: 350355.CrossRefGoogle ScholarPubMed
Lacey, D., Timms, E., Tan, H.-L., Kelley, M., Dunstan, C., Burgess, T., Elliott, R., et al. 1998. “Osteoprotegerin ligand is a cytokine that regulates osteoclast differentiation and activation.” Cell 93: 165176.CrossRefGoogle ScholarPubMed
Lau, E., Al-Dujaili, S., Guenther, A., Liu, D., Wang, L. and You, L. 2010. “Effect of low-magnitude, high-frequency vibration on osteocytes in the regulation of osteoclasts.” Bone 46: 15081515.CrossRefGoogle ScholarPubMed
Liu, C., Zhao, Y., Cheung, W. Y., Gandhi, R., Wang, L. and You, L. 2010. “Effects of cyclic hydraulic pressure on osteocytes.” Bone 46: 14491456.CrossRefGoogle ScholarPubMed
Lu, X. L., Huo, B., Chiang, V. and Guo, X. E. 2012. “Osteocytic network is more responsive in calcium signaling than osteoblastic network under fluid flow.” J Bone Miner Res 27: 563574.CrossRefGoogle ScholarPubMed
Mak, A. F. T., Qin, L., Hung, L. K., Cheng, C. W. and Tin, C. F. 2000. “A histomorphometric observation of flows in cortical bone under dynamic loading.” Microvascular Research 59: 290300.CrossRefGoogle ScholarPubMed
Malone, A. M. D., Anderson, C. T., Tummala, P., Kwon, R. Y., Johnston, T. R., Stearns, T. and Jacobs, C. R. 2007. “Primary cilia mediate mechanosensing in bone cells by a calcium-independent mechanism.” Proceedings of the National Academy of Sciences of the United States of America 104: 1332513330.CrossRefGoogle ScholarPubMed
Martin, R. 2007. “Targeted bone remodeling involves BMU steering as well as activation.” Bone 40: 15741580.CrossRefGoogle ScholarPubMed
Matsuo, K. and Irie, N. 2008. “Osteoclast-osteoblast communication.” Archives of Biochemistry and Biophysics 473: 201209.CrossRefGoogle ScholarPubMed
McGarry, J. G., Klein-Nulend, J. and Prendergast, P. J. 2005. “The effect of cytoskeletal disruption on pulsatile fluid flow-induced nitric oxide and prostaglandin E2 release in osteocytes and osteoblasts.” Biochemical and Biophysical Research Communications 330: 341348.CrossRefGoogle ScholarPubMed
McNamara, L., Majeska, R., Weinbaum, S., Friedrich, V. and Schaffler, M. 2009. “Attachment of osteocyte cell processes to the bone matrix.” The Anatomical Record 292: 355363.CrossRefGoogle Scholar
Mi, L. Y., Basu, M., Fritton, S. P. and Cowin, S. C. 2005a. “Analysis of avian bone response to mechanical loading. Part two: Development of a computational connected cellular network to study bone intercellular communication.” Biomechanics and Modeling in Mechanobiology 4: 132146.CrossRefGoogle ScholarPubMed
Mi, L. Y., Fritton, S. P., Basu, M. and Cowin, S. C. 2005b. “Analysis of avian bone response to mechanical loading-Part one: Distribution of bone fluid shear stress induced by bending and axial loading.” Biomechanics and Modeling in Mechanobiology 4: 118131.CrossRefGoogle ScholarPubMed
Mullender, M., El Haj, A. J., Yang, Y., Van Duin, M. A., Burger, E. H. and Klein-Nulend, J. 2004. “Mechanotransduction of bone cells in vitro: mechanobiology of bone tissue.” Medical and Biological Engineering and Computing 42: 1421.CrossRefGoogle ScholarPubMed
Murray, D. and Rushton, N. 1990. “The effect of strain on bone cell prostaglandin E2 release: a new experimental method.” Calcified Tissue International 47: 3539.CrossRefGoogle Scholar
Neuman, W. F., Neuman, M. W., Diamond, A. G., Menanteau, J. and Gibbons, W. S. 1982. “Blood: bone disequilibrium. VI. Studies of the solubility characteristics of brushite: apatite mixtures and their stabilization by noncollagenous proteins of bone.” Calcified Tissue International 34: 149157.CrossRefGoogle ScholarPubMed
Nicolella, P., Moravits, D. M., Lankford, J. and Bonewald, L. F. 2004. “Bone matrix strain is amplified at osteocyte lacunae in cortical bone.” Journal of Bone and Mineral Research 19: S72S72.Google Scholar
Owen, M. and Triffitt, J. 1976. “Extravascular albumin in bone tissue.” The Journal of Physiology 257: 293307.CrossRefGoogle ScholarPubMed
Pederson, L., Ruan, M., Westendorf, J. J., Khosla, S. and Oursler, M. J. 2008. “Regulation of bone formation by osteoclasts involves Wnt/BMP signaling and the chemokine sphingosine-1-phosphate.” Proceedings of the National Academy of Sciences 105: 2076420769.CrossRefGoogle ScholarPubMed
Perez-Amodio, S., Beertsen, W. and Everts, V. 2004. “(Pre-)osteoclasts induce retraction of osteoblasts before their fusion to osteoclasts.” Journal of Bone and Mineral Research 19: 17221731.CrossRefGoogle ScholarPubMed
Ponik, S. M., Triplett, J. W. and Pavalko, F. M. 2007. “Osteoblasts and osteocytes respond differently to oscillatory and unidirectional fluid flow profiles.” Journal of Cellular Biochemistry 100: 794807.CrossRefGoogle ScholarPubMed
Prendergast, P. J. and Huiskes, R. 1996. “Microdamage and osteocyte-lacuna strain in bone: a microstructural finite element analysis.” Journal of Biomechanical Engineering-Transactions of the Asme 118: 240246.CrossRefGoogle Scholar
Price, C., Zhou, X., Li, W. and Wang, L. 2011. “Real-time measurement of solute transport within the lacunar-canalicular system of mechanically loaded bone: direct evidence for load-induced fluid flow.” Journal of Bone and Mineral Research 26: 277285.CrossRefGoogle ScholarPubMed
Raisz, L. G. 1999. “Physiology and pathophysiology of bone remodeling.” Clinical Chemistry 45: 13531358.Google ScholarPubMed
Raisz, L. G., Pilbeam, C. C. and Fall, P. M. 1993. “Prostaglandins: mechanisms of action and regulation of production in bone.” Osteoporosis International 3: 136140.CrossRefGoogle ScholarPubMed
Rath Stern, A., Stern, M., Van Dyke, M., Jähn, K., Prideaux, M. and Bonewald, L. 2012. “Isolation and culture of primary osteocytes from the long bones of skeletally mature and aged mice.” BioTechniques: 52.CrossRefGoogle Scholar
Rawlinson, S. C. F., Pitsillides, A. A. and Lanyon, L. E. 1996. “Involvement of different ion channels in osteoblasts’ and osteocytes’ early responses to mechanical strain.” Bone 19: 609614.CrossRefGoogle ScholarPubMed
Rho, J.-Y., Tsui, T. Y. and Pharr, G. M. 1997. “Elastic properties of human cortical and trabecular lamellar bone measured by nanoindentation.” Biomaterials 18: 13251330.CrossRefGoogle ScholarPubMed
Roach, H. 1994. “Why does bone matrix contain non-collagenous proteins? The possible roles of osteocalcin, osteonectin, osteopontin and bone sialoprotein in bone mineralisation and resorption.” Cell Biology International 18: 617628.CrossRefGoogle ScholarPubMed
Robinson, R. A. 1952. “An electron-microscopic study of the crystalline inorganic component of bone and its relationship to the organic matrix.” The Journal of Bone and Joint Surgery 34: 389476.CrossRefGoogle Scholar
Robling, A. G., Bellido, T. M. and Turner, C. H. 2006. “Mechanical loading reduces osteocyte expression of sclerostin protein.” Journal of Bone and Mineral Research 21: S72S72.Google Scholar
Rubin, C. T. and Lanyon, L. E. 1982. “Limb mechanics as a function of speed and gait: a study of functional strains in the radius and tibia of horse and dog.” Journal of Experimental Biology 101: 187211.CrossRefGoogle Scholar
Ryser, M. D., Nigam, N. and Sr. Komarova, S. V. 2009. “Mathematical modeling of spatio-temporal dynamics of a single bone multicellular unit.” Journal of Bone and Mineral Research 24: 860870.CrossRefGoogle ScholarPubMed
Sauren, Y. M., Mieremet, R. H., Groot, C. G. and Scherft, J. P. 1992. “An electron microscopic study on the presence of proteoglycans in the mineralized matrix of rat and human compact lamellar bone.” The Anatomical Record 232: 3644.CrossRefGoogle ScholarPubMed
Shapiro, F., Cahill, C., Malatantis, G. and Nayak, R. C. 1995. “Transmission electron-microscopic demonstration of vimentin in rat osteoblast and osteocyte cell-bodies and processes using the immunogold technique.” Anatomical Record 241: 3948.CrossRefGoogle ScholarPubMed
Sims, N. A. and Martin, T. J. 2014. “Coupling the activities of bone formation and resorption: a multitude of signals within the basic multicellular unit.” BoneKey Reports: 3.CrossRefGoogle Scholar
Singhvi, R., Kumar, A., Lopez, G. P., Stephanopoulos, G. N., Wang, D. I. C., Whitesides, G. M. and Ingber, D. E. 1994. “Engineering cell shape and function.” Science 264: 696698.CrossRefGoogle ScholarPubMed
Song, S. H., Choi, J. and Jung, H. I. 2010. “A microfluidic magnetic bead impact generator for physical stimulation of osteoblast cell.” Electrophoresis 31: 27622770.CrossRefGoogle ScholarPubMed
Tami, A. E., Schaffler, M. B. and Knothe Tate, M. L. 2003. “Probing the tissue to subcellular level structure underlying bone’s molecular sieving function.” Biorheology 40: 577590.Google ScholarPubMed
Tasevski, V., Sorbetti, J. M., Chiu, S. S., Shrive, N. G. and Hart, D. A. 2005. “Influence of mechanical and biological signals on gene expression in human MG-63 cells: evidence for a complex interplay between hydrostatic compression and vitamin D3 or TGF-beta 1 on MMP-1 and MMP-3 mRNA level.” Biochemistry and Cell Biology 83: 96.CrossRefGoogle ScholarPubMed
Taylor, D., Hazenberg, J. G. and Lee, T. C. 2007. “Living with cracks: damage and repair in human bone.” Nature Materials 6: 263268.CrossRefGoogle ScholarPubMed
Teti, A. and Zallone, A. 2009. “Do osteocytes contribute to bone mineral homeostasis? Osteocytic osteolysis revisited.” Bone 44: 1116.CrossRefGoogle ScholarPubMed
Udagawa, N., Takahashi, N., Jimi, E., Matsuzaki, K., Tsurukai, T., Itoh, K., Nakagawa, N., et al. 1999. “Osteoblasts/stromal cells stimulate osteoclast activation through expression of osteoclast differentiation factor/RANKL but not macrophage colony-stimulating factor.” Bone 25: 517523.CrossRefGoogle Scholar
Van’t Hof, R. J. and Ralston, S. H. 2001. “Nitric oxide and bone.” Immunology 103: 255261.CrossRefGoogle ScholarPubMed
Van Bezooijen, R. L., Winkler, D., Hayes, T., Karperien, M., Visser, A., van der Wee-Pals, L., Hamersma, H., et al. 2002. “Sclerostin: an osteocyte-expressed BMP antagonist the inhibits bone formation by mature osteoblasts.” Journal of Bone and Mineral Research 17: S144S144.Google Scholar
Verborgt, O., Gibson, G. J. and Schaffler, M. B. 2000. “Loss of osteocyte integrity in association with microdamage and bone remodeling after fatigue in vivo.” Journal of Bone and Mineral Research 15: 6067.CrossRefGoogle ScholarPubMed
Wang, L., Fritton, S. P., Cowin, S. C. and Weinbaum, S. 1999. “Fluid pressure relaxation depends upon osteonal microstructure: modeling an oscillatory bending experiment.” Journal of Biomechanics 32: 663672.CrossRefGoogle ScholarPubMed
Wang, L., Wang, Y., Han, Y., Henderson, S. C., Majeska, R. J., Weinbaum, S. and Schaffler, M. B. 2005. “In situ measurement of solute transport in the bone lacunar-canalicular system.” Proceedings of the National Academy of Sciences of the United States of America 102: 1191111916.CrossRefGoogle ScholarPubMed
Wang, Y., McNamara, L. M., Schaffler, M. B. and Weinbaum, S. 2007. “A model for the role of integrins in flow induced mechanotransduction in osteocytes.” Proceedings of the National Academy of Sciences of the United States of America 104: 1594115946.CrossRefGoogle Scholar
Weinbaum, S., Cowin, S. C. and Zeng, Y. 1994. “A model for the excitation of osteocytes by mechanical loading-induced bone fluid shear stresses.” J Biomech 27: 339360.CrossRefGoogle Scholar
Weinger, J. M. and Holtrop, M. E. 1974. “An ultrastructural study of bone cells: the occurrence of microtubules, microfilaments and tight junctions.” Calcified Tissue Research 14: 1529.CrossRefGoogle ScholarPubMed
Xiong, J. and O’Brien, C. A. 2012. “Osteocyte RANKL: new insights into the control of bone remodeling.” Journal of Bone and Mineral Research 27: 499505.CrossRefGoogle ScholarPubMed
You, J., Yellowley, C. E., Donahue, H. J., Zhang, Y., Chen, Q. and Jacobs, C. R. 2000. “Substrate deformation levels associated with routine physical activity are less stimulatory to bone cells relative to loading-induced oscillatory fluid flow.” Journal of Biomechanical Engineering 122: 387393.CrossRefGoogle ScholarPubMed
You, L., Cowin, S. C., Schaffler, M. B. and Weinbaum, S. 2001. “A model for strain amplification in the actin cytoskeleton of osteocytes due to fluid drag on pericellular matrix.” Journal of Biomechanics 34: 13751386.CrossRefGoogle Scholar
You, L., Temiyasathit, S., Coyer, S. R., García, A. J. and Jacobs, C. R. 2008a. “Bone cells grown on micropatterned surfaces are more sensitive to fluid shear stress.” Cellular and Molecular Bioengineering 1: 182188.CrossRefGoogle Scholar
You, L., Temiyasathit, S., Lee, P., Kim, C. H., Tummala, P., Yao, W., Kingery, W., et al. 2008b. “Osteocytes as mechanosensors in the inhibition of bone resorption due to mechanical loading.” Bone 42: 172179.CrossRefGoogle ScholarPubMed
You, L., Temiyasathit, S., Tao, E., Prinz, F. and Jacobs, C. R. 2008c. “3D microfluidic approach to mechanical stimulation of osteocyte processes.” Cellular and Molecular Bioengineering 1: 103107.CrossRefGoogle Scholar
You, L. D., Weinbaum, S., Cowin, S. C. and Schaffler, M. B. 2004. “Ultrastructure of the osteocyte process and its pericellular matrix.” Anatomical Record Part A-Discoveries in Molecular Cellular and Evolutionary Biology 278: 505513.CrossRefGoogle ScholarPubMed
Zhang, D., Weinbaum, S. and Cowin, S. 1998. “On the calculation of bone pore water pressure due to mechanical loading.” International Journal of Solids and Structures 35: 4981.CrossRefGoogle Scholar
Zhang, K., Barragan-Adjemian, C., Ye, L., Kotha, S., Dallas, M., Lu, Y., Zhao, S., et al. 2006. “E11/gp38 selective expression in osteocytes: regulation by mechanical strain and role in dendrite elongation.” Molecular and Cellular Biology 26: 45394552.CrossRefGoogle ScholarPubMed
Zhang, X., Liu, X., Sun, J., He, S., Lee, I. and Pak, H. K. 2008. “Real-time observations of mechanical stimulus-induced enhancements of mechanical properties in osteoblast cells.” Ultramicroscopy 108: 13381341.CrossRefGoogle ScholarPubMed
Zhao, S., Kato, Y., Zhang, Y., Harris, S., Ahuja, S. S. and Bonewald, L. F. 2002. “MLO-Y4 osteocyte-like cells support osteoclast formation and activation.” Journal of Bone and Mineral Research 17: 20682079.CrossRefGoogle ScholarPubMed

References

Alegre-Cebollada, J., Kosuri, P., Giganti, D., Eckels, E., Rivas-Pardo, J. A., Hamdani, N., Warren, C. M., et al. (2014). “S-glutathionylation of cryptic cysteines enhances titin elasticity by blocking protein folding.” Cell 156: 12351246.CrossRefGoogle ScholarPubMed
Altroff, H., Choulier, L. and Mardon, H. J. (2003). “Synergistic activity of the ninth and tenth FIII domains of human fibronectin depends upon structural stability.” J Biol Chem 278: 491497.CrossRefGoogle ScholarPubMed
Ananthanarayanan, B., Kim, Y. and Kumar, S. (2011). “Elucidating the mechanobiology of malignant brain tumors using a brain matrix-mimetic hyaluronic acid hydrogel platform.” Biomaterials 32: 79137923.CrossRefGoogle ScholarPubMed
Aratyn-Schaus, Y. and Gardel, M. L. (2010). “Transient frictional slip between integrin and the ECM in focal adhesions under myosin II tension.” Curr Biol 20: 11451153.CrossRefGoogle ScholarPubMed
Aratyn-Schaus, Y., Oakes, P. W. and Gardel, M. L. (2011). “Dynamic and structural signatures of lamellar actomyosin force generation.” Mol Biol Cell 22: 13301339.CrossRefGoogle ScholarPubMed
Arora, P. D., Narani, N. and Mcculloch, C. A. (1999). “The compliance of collagen gels regulates transforming growth factor-beta induction of alpha-smooth muscle actin in fibroblasts.” Am J Pathol 154: 871882.CrossRefGoogle ScholarPubMed
Baird, A., Schubert, D., Ling, N. and Guillemin, R. (1988). “Receptor- and heparin-binding domains of basic fibroblast growth factor.” Proc Natl Acad Sci USA 85: 23242328.CrossRefGoogle ScholarPubMed
Balaban, N. Q., Schwarz, U. S., Riveline, D., Goichberg, P., Tzur, G., Sabanay, I., Mahalu, D., et al. (2001). “Force and focal adhesion assembly: a close relationship studied using elastic micropatterned substrates.” Nat Cell Biol 3: 466472.CrossRefGoogle ScholarPubMed
Balestrini, J. L., Chaudhry, S., Sarrazy, V., Koehler, A. and Hinz, B. (2012). “The mechanical memory of lung myofibroblasts.” Integr Biol (Camb) 4: 410421.CrossRefGoogle ScholarPubMed
Barker, T. H., et al. (2004a). “Thy-1 regulates fibroblast focal adhesions, cytoskeletal organization and migration through modulation of p190 RhoGAP and Rho GTPase activity.” Exp Cell Res 295: 488496.CrossRefGoogle ScholarPubMed
Barker, T. H., et al. (2004b). “Thrombospondin-1-induced focal adhesion disassembly in fibroblasts requires Thy-1 surface expression, lipid raft integrity, and Src activation.” J Biol Chem 279: 2351023516.CrossRefGoogle ScholarPubMed
Barry-Hamilton, V., Spangler, R., Marshall, D., McCauley, S., Rodriguez, H. M., Oyasu, M., Mikels, A., et al. (2010). “Allosteric inhibition of lysyl oxidase-like-2 impedes the development of a pathologic microenvironment.” Nat Med 16: 10091017.CrossRefGoogle ScholarPubMed
Betz, P., Nerlich, A., Wilske, J., Tubel, J., Penning, R. and Eisenmenger, W. (1993). “Immunohistochemical localization of collagen types I and VI in human skin wounds.” Int J Legal Med 106: 3134.CrossRefGoogle ScholarPubMed
Bhadriraju, K., Yang, M., Alom Ruiz, S., Pirone, D., Tan, J. and Chen, C. S. (2007). “Activation of ROCK by RhoA is regulated by cell adhesion, shape, and cytoskeletal tension.” Exp Cell Res 313: 36163623.CrossRefGoogle Scholar
Boettiger, D. (2012). “Mechanical control of integrin-mediated adhesion and signaling.” Curr Opin Cell Biol 24: 592599.CrossRefGoogle ScholarPubMed
Booth, A. J., Hadley, R., Cornett, A. M., Dreffs, A. A., Matthes, S. A., Tsui, J. L., Weiss, K., et al. (2012). “Acellular normal and fibrotic human lung matrices as a culture system for in vitro investigation.” Am J Respir Crit Care Med 186: 866876.CrossRefGoogle ScholarPubMed
Brown, A. C., Baker, S. R., Douglas, A. M., Keating, M., Alvarez-Elizondo, M. B., Botvinick, E. L., Guthold, M., Barker, T. H. (2015). “Molecular interference of fibron’s divalent polymerization mechanism enables modulation of multiscale material properties.” Biomaterials 49: 2736.CrossRefGoogle ScholarPubMed
Brown, A. C., Fiore, V. F., Sulchek, T. A. and Barker, T. H. (2013). “Physical and chemical microenvironmental cues orthogonally control the degree and duration of fibrosis-associated epithelial-to-mesenchymal transitions.” J Pathol 229: 2535.CrossRefGoogle ScholarPubMed
Brown, A. C., Rowe, J. A. and Barker, T. H. (2011). “Guiding epithelial cell phenotypes with engineered integrin-specific recombinant fibronectin fragments.” Tissue Eng Part A 17: 139150.CrossRefGoogle ScholarPubMed
Brown, E. J. and Frazier, W. A. (2001). “Integrin-associated protein (CD47) and its ligands.” Trends Cell Biol 11: 130135.CrossRefGoogle ScholarPubMed
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: 2807128074.CrossRefGoogle ScholarPubMed
Chang, H. Y., Chi, J. T., Dudoit, S., Bondre, C., Van De Rijn, M., Botstein, D. and Brown, P. O. (2002). “Diversity, topographic differentiation, and positional memory in human fibroblasts.” Proc Natl Acad Sci USA 99: 1287712882.CrossRefGoogle ScholarPubMed
Chaudhuri, O., Koshy, S. T., Branco Da Cunha, C., Shin, J. W., Verbeke, C. S., Allison, K. H. and Mooney, D. J. (2014). “Extracellular matrix stiffness and composition jointly regulate the induction of malignant phenotypes in mammary epithelium.” Nat Mater 13: 970978.CrossRefGoogle ScholarPubMed
Chen, W., Lou, J., Evans, E. A. and Zhu, C. (2012). “Observing force-regulated conformational changes and ligand dissociation from a single integrin on cells.” J Cell Biol 199: 497512.CrossRefGoogle ScholarPubMed
Choi, C. K., Vicente-Manzanares, M., Zareno, J., Whitmore, L. A., Mogilner, A. and Horwitz, A. R. (2008). “Actin and alpha-actinin orchestrate the assembly and maturation of nascent adhesions in a myosin II motor-independent manner.” Nat Cell Biol 10: 10391050.CrossRefGoogle Scholar
Choquet, D., Felsenfeld, D. P. and Sheetz, M. P. (1997). “Extracellular matrix rigidity causes strengthening of integrin-cytoskeleton linkages.” Cell 88: 3948.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: 14031415.CrossRefGoogle ScholarPubMed
Del Pozo, M. A., Alderson, N. B., Kiosses, W. B., Chiang, H. H., Anderson, R. G. and Schwartz, M. A. (2004). “Integrins regulate Rac targeting by internalization of membrane domains.” Science 303: 839842.CrossRefGoogle ScholarPubMed
Del Rio, A., Perez-Jimenez, R., Liu, R., Roca-Cusachs, P., Fernandez, J. M. and Sheetz, M. P. (2009). “Stretching single talin rod molecules activates vinculin binding.” Science 323: 638641.CrossRefGoogle ScholarPubMed
Discher, D. E., Janmey, P. and Wang, Y. L. (2005). “Tissue cells feel and respond to the stiffness of their substrate.” Science 310: 11391143.CrossRefGoogle Scholar
Dolhnikoff, M., Mauad, T. and Ludwig, M. S. (1999). “Extracellular matrix and oscillatory mechanics of rat lung parenchyma in bleomycin-induced fibrosis.” Am J Respir Crit Care Med 160: 17501757.CrossRefGoogle ScholarPubMed
Driskell, R. R., Lichtenberger, B. M., Hoste, E., Kretzschmar, K., Simons, B. D., Charalambous, M., Ferron, S. R., et al. (2013). “Distinct fibroblast lineages determine dermal architecture in skin development and repair.” Nature 504: 277281.CrossRefGoogle ScholarPubMed
Dumbauld, D. W., Lee, T. T., Singh, A., Scrimgeour, J., Gersbach, C. A., Zamir, E. A., Fu, J., et al. (2013). “How vinculin regulates force transmission.” Proc Natl Acad Sci USA 110: 97889793.CrossRefGoogle ScholarPubMed
Ebihara, T., Venkatesan, N., Tanaka, R. and Ludwig, M. S. (2000). “Changes in extracellular matrix and tissue viscoelasticity in bleomycin-induced lung fibrosis. Temporal aspects.” Am J Respir Crit Care Med 162: 15691576.CrossRefGoogle ScholarPubMed
Eggeling, C., Ringemann, C., Medda, R., Schwarzmann, G., Sandhoff, K., Polyakova, S., Belov, V. N., et al. (2009). “Direct observation of the nanoscale dynamics of membrane lipids in a living cell.” Nature 457: 11591162.CrossRefGoogle Scholar
Ehrlicher, A. J., Nakamura, F., Hartwig, J. H., Weitz, D. A. and Stossel, T. P. (2011). “Mechanical strain in actin networks regulates FilGAP and integrin binding to filamin A.” Nature 478: 260263.CrossRefGoogle ScholarPubMed
Engler, A. J., Griffin, M. A., Sen, S., Bonnemann, C. G., Sweeney, H. L. and Discher, D. E. (2004). “Myotubes differentiate optimally on substrates with tissue-like stiffness: pathological implications for soft or stiff microenvironments.” J Cell Biol 166: 877887.CrossRefGoogle ScholarPubMed
Engler, A. J., Sen, S., Sweeney, H. L. and Discher, D. E. (2006). “Matrix elasticity directs stem cell lineage specification.” Cell 126: 677689.CrossRefGoogle ScholarPubMed
Evans, E. A. and Calderwood, D. A. (2007). “Forces and bond dynamics in cell adhesion.” Science 316: 11481153.CrossRefGoogle ScholarPubMed
Fiore, V. F., Ju, L., Chen, Y., Zhu, C. and Barker, T. H. (2014). “Dynamic catch of a Thy-1-alpha5beta1+syndecan-4 trimolecular complex.” Nat Commun 5: 4886.CrossRefGoogle ScholarPubMed
Friedland, J. C., Lee, M. H. and Boettiger, D. (2009). “Mechanically activated integrin switch controls alpha5beta1 function.” Science 323: 642644.CrossRefGoogle ScholarPubMed
Fritsch, C., Orian-Rousseaul, V., Lefebvre, O., Simon-Assmann, P., Reimund, J. M., Duclos, B. and Kedinger, M. (1999). “Characterization of human intestinal stromal cell lines: response to cytokines and interactions with epithelial cells.” Exp Cell Res 248: 391406.CrossRefGoogle ScholarPubMed
Galbraith, C. G., Yamada, K. M. and Galbraith, J. A. (2007). “Polymerizing actin fibers position integrins primed to probe for adhesion sites.” Science 315: 992995.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: 695705.CrossRefGoogle ScholarPubMed
Garcia-Alvarez, B., De Pereda, J. M., Calderwood, D. A., Ulmer, T. S., Critchley, D., Campbell, I. D., Ginsberg, M. H. et al. (2003). “Structural determinants of integrin recognition by talin.” Mol Cell 11: 4958.CrossRefGoogle ScholarPubMed
Gardel, M. L., Sabass, B., Ji, L., Danuser, G., Schwarz, U. S. and Waterman, C. M. (2008). “Traction stress in focal adhesions correlates biphasically with actin retrograde flow speed.” J Cell Biol 183: 9991005.CrossRefGoogle ScholarPubMed
Gardel, M. L., Schneider, I. C., Aratyn-Schaus, Y. and Waterman, C. M. (2010). “Mechanical integration of actin and adhesion dynamics in cell migration.” Annu Rev Cell Dev Biol 26: 315333.CrossRefGoogle ScholarPubMed
Gaus, K., Le Lay, S., Balasubramanian, N. and Schwartz, M. A. (2006). “Integrin-mediated adhesion regulates membrane order.” J Cell Biol 174: 725734.CrossRefGoogle ScholarPubMed
Geiger, B., Spatz, J. P. and Bershadsky, A. D. (2009). “Environmental sensing through focal adhesions.” Nat Rev Mol Cell Biol 10: 2133.CrossRefGoogle ScholarPubMed
Geiger, B. and Yamada, K. M. (2011). “Molecular architecture and function of matrix adhesions.” Cold Spring Harb Perspect Biol 3.CrossRefGoogle ScholarPubMed
Goetz, J. G., Minguet, S., Navarro-Lerida, I., Lazcano, J. J., Samaniego, R., Calvo, E., Tello, M., et al. (2011). “Biomechanical remodeling of the microenvironment by stromal caveolin-1 favors tumor invasion and metastasis.” Cell 146: 148163.CrossRefGoogle ScholarPubMed
Goffin, J. M., Pittet, P., Csucs, G., Lussi, J. W., Meister, J. J. and Hinz, B. (2006). “Focal adhesion size controls tension-dependent recruitment of alpha-smooth muscle actin to stress fibers.” J Cell Biol 172: 259268.CrossRefGoogle ScholarPubMed
Goswami, D., Gowrishankar, K., Bilgrami, S., Ghosh, S., Raghupathy, R., Chadda, R., Vishwakarma, R., et al. (2008). “Nanoclusters of GPI-anchored proteins are formed by cortical actin-driven activity.” Cell 135: 10851097.CrossRefGoogle ScholarPubMed
Grashoff, C., Hoffman, B. D., Brenner, M. D., Zhou, R., Parsons, M., Yang, M. T., Mclean, M. A. (2010). “Measuring mechanical tension across vinculin reveals regulation of focal adhesion dynamics.” Nature 466: 263266.CrossRefGoogle ScholarPubMed
Guilluy, C., Swaminathan, V., Garcia-Mata, R., O’Brien, E. T., Superfine, R. and Burridge, K. (2011). “The Rho GEFs LARG and GEF-H1 regulate the mechanical response to force on integrins.” Nat Cell Biol 13: 722727.CrossRefGoogle ScholarPubMed
Gurtner, G. C., Werner, S., Barrandon, Y. and Longaker, M. T. (2008). “Wound repair and regeneration.” Nature 453: 314321.CrossRefGoogle ScholarPubMed
Hagood, J. S., Lasky, J. A., Nesbitt, J. E. and Segarini, P. (2001). “Differential expression, surface binding, and response to connective tissue growth factor in lung fibroblast subpopulations.” Chest 120: 64S66S.CrossRefGoogle ScholarPubMed
Hagood, J. S., Mangalwadi, A., Guo, B., Macewen, M. W., Salazar, L. and Fuller, G. M. (2002). “Concordant and discordant interleukin-1-mediated signaling in lung fibroblast thy-1 subpopulations.” Am J Respir Cell Mol Biol 26: 702708.CrossRefGoogle ScholarPubMed
Hagood, J. S., Miller, P. J., Lasky, J. A., Tousson, A., Guo, B., Fuller, G. M. and Mcintosh, J. C. (1999). “Differential expression of platelet-derived growth factor-alpha receptor by Thy-1(-) and Thy-1(+) lung fibroblasts.” Am J Physiol 277: L218L224.Google ScholarPubMed
Hagood, J. S., Prabhakaran, P., Kumbla, P., Salazar, L., Macewen, M. W., Barker, T. H., Ortiz, L. A., et al. (2005). “Loss of fibroblast Thy-1 expression correlates with lung fibrogenesis.” Am J Pathol 167: 365379.CrossRefGoogle ScholarPubMed
Hattori, N., Mochizuki, S., Kishi, K., Nakajima, T., Takaishi, H., D’armiento, J. and Okada, Y. (2009). “MMP-13 plays a role in keratinocyte migration, angiogenesis, and contraction in mouse skin wound healing.” Am J Pathol 175: 533546.CrossRefGoogle Scholar
Hinz, B., Celetta, G., Tomasek, J. J., Gabbiani, G. and Chaponnier, C. (2001). “Alpha-smooth muscle actin expression upregulates fibroblast contractile activity.” Mol Biol Cell 12: 27302741.CrossRefGoogle ScholarPubMed
Horowitz, J. C., Rogers, D. S., Simon, R. H., Sisson, T. H. and Thannickal, V. J. (2008). “Plasminogen activation induced pericellular fibronectin proteolysis promotes fibroblast apoptosis.” Am J Respir Cell Mol Biol 38: 7887.CrossRefGoogle ScholarPubMed
Hu, K., Ji, L., Applegate, K. T., Danuser, G. and Waterman-Storer, C. M. (2007). “Differential transmission of actin motion within focal adhesions.” Science 315: 111115.CrossRefGoogle ScholarPubMed
Huang, X., Yang, N., Fiore, V. F., Barker, T. H., Sun, Y., Morris, S. W., Ding, Q., et al. (2012). “Matrix stiffness-induced myofibroblast differentiation is mediated by intrinsic mechanotransduction.” Am J Respir Cell Mol Biol 47: 340348.CrossRefGoogle ScholarPubMed
Humphries, J. D., Byron, A., Bass, M. D., Craig, S. E., Pinney, J. W., Knight, D. and Humphries, M. J. (2009). “Proteomic analysis of integrin-associated complexes identifies RCC2 as a dual regulator of Rac1 and Arf6.” Sci Signal 2: ra51.CrossRefGoogle ScholarPubMed
Humphries, J. D., Wang, P., Streuli, C., Geiger, B., Humphries, M. J. and Ballestrem, C. (2007). “Vinculin controls focal adhesion formation by direct interactions with talin and actin.” J Cell Biol 179: 10431057.CrossRefGoogle ScholarPubMed
Hynes, R. O. (2002). “Integrins: bidirectional, allosteric signaling machines.” Cell 110: 673687.CrossRefGoogle ScholarPubMed
Ingber, D. E. (2003). “Tensegrity I. Cell structure and hierarchical systems biology.” J Cell Sci 116: 11571173.CrossRefGoogle ScholarPubMed
Ingber, D. E., Madri, J. A. and Folkman, J. (1987). “Endothelial growth factors and extracellular matrix regulate DNA synthesis through modulation of cell and nuclear expansion.” In Vitro Cell Dev Biol 23: 387394.CrossRefGoogle ScholarPubMed
Jiang, G., 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: 334337.CrossRefGoogle ScholarPubMed
Jiang, G., Huang, A. H., Cai, Y., Tanase, M. and Sheetz, M. P. (2006). “Rigidity sensing at the leading edge through alphavbeta3 integrins and RPTPalpha.” Biophys J 90: 18041809.CrossRefGoogle Scholar
Johnson, C. P., Tang, H. Y., Carag, C., Speicher, D. W. and Discher, D. E. (2007). “Forced unfolding of proteins within cells.” Science 317: 663666.CrossRefGoogle ScholarPubMed
Jurchenko, C., Chang, Y., Narui, Y., Zhang, Y. and Salaita, K. S. (2014). “Integrin-generated forces lead to streptavidin-biotin unbinding in cellular adhesions.” Biophys J 106: 14361446.CrossRefGoogle ScholarPubMed
Kanchanawong, P., Shtengel, G., Pasapera, A. M., Ramko, E. B., Davidson, M. W., Hess, H. F. and Waterman, C. M. (2010). “Nanoscale architecture of integrin-based cell adhesions.” Nature 468: 580584.CrossRefGoogle ScholarPubMed
Kapanci, Y., Desmouliere, A., Pache, J. C., Redard, M. and Gabbiani, G. (1995). “Cytoskeletal protein modulation in pulmonary alveolar myofibroblasts during idiopathic pulmonary fibrosis. Possible role of transforming growth factor beta and tumor necrosis factor alpha.” Am J Respir Crit Care Med 152: 21632169.CrossRefGoogle ScholarPubMed
Kendall, R. T. and Feghali-Bostwick, C. A. (2011). “Fibroblasts in fibrosis: novel roles and mediators.” Front Pharmacol 5: 123.Google Scholar
Kim, C., Ye, F. and Ginsberg, M. H. (2011). “Regulation of integrin activation.” Annu Rev Cell Dev Biol 27: 321345.CrossRefGoogle ScholarPubMed
Klingberg, F., Hinz, B. and White, E. S. (2013). “The myofibroblast matrix: implications for tissue repair and fibrosis.” J Pathol 229: 298309.CrossRefGoogle ScholarPubMed
Kong, F., Garcia, A. J., Mould, A. P., Humphries, M. J. and Zhu, C. (2009). “Demonstration of catch bonds between an integrin and its ligand.” J Cell Biol 185: 12751284.CrossRefGoogle ScholarPubMed
Kong, F., Li, Z., Parks, W. M., Dumbauld, D. W., Garcia, A. J., Mould, A. P., Humphries, M. J. and Zhu, C. (2013). “Cyclic mechanical reinforcement of integrin-ligand interactions.” Mol Cell 49: 10601068.CrossRefGoogle ScholarPubMed
Kostic, A. and Sheetz, M. P. (2006). “Fibronectin rigidity response through Fyn and p130Cas recruitment to the leading edge.” Mol Biol Cell 17: 26842695.CrossRefGoogle Scholar
Krammer, A., Lu, H., Isralewitz, B., Schulten, K. and Vogel, V. (1999). “Forced unfolding of the fibronectin type III module reveals a tensile molecular recognition switch.” Proc Natl Acad Sci USA 96: 13511356.CrossRefGoogle ScholarPubMed
Kulasekaran, P., Scavone, C. A., Rogers, D. S., Arenberg, D. A., Thannickal, V. J. and Horowitz, J. C. (2009). “Endothelin-1 and transforming growth factor-beta1 independently induce fibroblast resistance to apoptosis via AKT activation.” Am J Respir Cell Mol Biol 4: 484493.CrossRefGoogle Scholar
Kuo, J. C., Han, X., Hsiao, C. T., Yates, J. R., 3rd, and Waterman, C. M. (2011). “Analysis of the myosin-II-responsive focal adhesion proteome reveals a role for β-Pix in negative regulation of focal adhesion maturation.” Nat Cell Biol 13: 383393.CrossRefGoogle ScholarPubMed
Leahy, D. J., Aukhil, I. and Erickson, H. P. (1996). “2.0 A crystal structure of a four-domain segment of human fibronectin encompassing the RGD loop and synergy region.” Cell 84: 155164.CrossRefGoogle ScholarPubMed
Leitinger, B. and Hogg, N. (2002). “The involvement of lipid rafts in the regulation of integrin function.” J Cell Sci 115: 963972.CrossRefGoogle ScholarPubMed
Levental, K. R., Yu, H., Kass, L., Lakins, J. N., Egeblad, M., Erler, J. T., Fong, S. F., et al. (2009). “Matrix crosslinking forces tumor progression by enhancing integrin signaling.” Cell 139: 891906.CrossRefGoogle ScholarPubMed
Li, L., Huang, H. H., Badilla, C. L. and Fernandez, J. M. (2005). “Mechanical unfolding intermediates observed by single-molecule force spectroscopy in a fibronectin type III module.” J Mol Biol 345: 817826.CrossRefGoogle Scholar
Lingwood, D. and Simons, K. (2007). “Detergent resistance as a tool in membrane research.” Nat Protoc 2: 21592165.CrossRefGoogle ScholarPubMed
Lingwood, D. and Simons, K. (2010). “Lipid rafts as a membrane-organizing principle.” Science 327: 4650.CrossRefGoogle ScholarPubMed
Little, W. C., Schwartlander, R., Smith, M. L., Gourdon, D. and Vogel, V. (2009). “Stretched extracellular matrix proteins turn fouling and are functionally rescued by the chaperones albumin and casein.” Nano Lett 9: 41584167.CrossRefGoogle ScholarPubMed
Liu, F., Mih, J. D., Shea, B. S., Kho, A. T., Sharif, A. S., Tager, A. M. and Tschumperlin, D. J. (2010). “Feedback amplification of fibrosis through matrix stiffening and COX-2 suppression.” J Cell Biol 190: 693706.CrossRefGoogle ScholarPubMed
Liu, Y., Medda, R., Liu, Z., Galior, K., Yehl, K., Spatz, J. P., Cavalcanti-Adam, E. A. et al. (2014). “Nanoparticle tension probes patterned at the nanoscale: impact of integrin clustering on force transmission.” Nano Lett 14: 55395546.CrossRefGoogle ScholarPubMed
Liu, Y., Yehl, K., Narui, Y. and Salaita, K. (2013). “Tension sensing nanoparticles for mechano-imaging at the living/nonliving interface.” J Am Chem Soc 135: 53205323.CrossRefGoogle ScholarPubMed
Luo, B. H., Carman, C. V. and Springer, T. A. (2007). “Structural basis of integrin regulation and signaling.” Annu Rev Immunol 25: 619647.CrossRefGoogle ScholarPubMed
Machacek, M., Hodgson, L., Welch, C., Elliott, H., Pertz, O., Nalbant, P., Abell, A., et al. (2009). “Coordination of Rho GTPase activities during cell protrusion.” Nature 461: 99103.CrossRefGoogle ScholarPubMed
Margadant, F., Chew, L. L., Hu, X., Yu, H., Bate, N., Zhang, X. and Sheetz, M. (2011). “Mechanotransduction in vivo by repeated talin stretch-relaxation events depends upon vinculin.” PLoS Biol 9: e1001223.CrossRefGoogle ScholarPubMed
Markowski, M. C., Brown, A. C. and Barker, T. H. (2012). “Directing epithelial to mesenchymal transition through engineered microenvironments displaying orthogonal adhesive and mechanical cues.” J Biomed Mater Res A 100: 21192127.CrossRefGoogle ScholarPubMed
Martino, M. M., Briquez, P. S., Guc, E., Tortelli, F., Kilarski, W. W., Metzger, S., Rice, J. J., (2011). “Growth factors engineered for super-affinity to the extracellular matrix enhance tissue healing.” Science 343: 885888.CrossRefGoogle Scholar
Martino, M. M., Briquez, P. S., Ranga, A., Lutolf, M. P. and Hubbell, J. A. (2013). “Heparin-binding domain of fibrin(ogen) binds growth factors and promotes tissue repair when incorporated within a synthetic matrix.” Proc Natl Acad Sci USA 110: 45634568.CrossRefGoogle ScholarPubMed
Martino, M. M., Mochizuki, M., Rothenfluh, D. A., Rempel, S. A., Hubbell, J. A. and Barker, T. H. (2009). “Controlling integrin specificity and stem cell differentiation in 2D and 3D environments through regulation of fibronectin domain stability.” Biomaterials 30: 10891097.CrossRefGoogle ScholarPubMed
Miroshnikova, Y. A., Jorgens, D. M., Spirio, L., Auer, M., Sarang-Sieminski, A. L. and Weaver, V. M. (2011). “Engineering strategies to recapitulate epithelial morphogenesis within synthetic three-dimensional extracellular matrix with tunable mechanical properties.” Phys Biol 8: 026013.CrossRefGoogle ScholarPubMed
Moore, S. W., Roca-Cusachs, P. and Sheetz, M. P. (2010). “Stretchy proteins on stretchy substrates: the important elements of integrin-mediated rigidity sensing.” Dev Cell 19: 194206.CrossRefGoogle ScholarPubMed
Na, S., Collin, O., Chowdhury, F., Tay, B., Ouyang, M., Wang, Y. and Wang, N. (2008). “Rapid signal transduction in living cells is a unique feature of mechanotransduction.” Proc Natl Acad Sci USA, 105: 66266631.CrossRefGoogle ScholarPubMed
Nakamura, F., Song, M., Hartwig, J. H. and Stossel, T. P. (2014). “Documentation and localization of force-mediated filamin A domain perturbations in moving cells.” Nat Commun 5: 4656.CrossRefGoogle ScholarPubMed
Ng, S. P., Billings, K. S., Randles, L. G. and Clarke, J. (2008). “Manipulating the stability of fibronectin type III domains by protein engineering.” Nanotechnology 19: 384023.CrossRefGoogle ScholarPubMed
Nobes, C. D. and Hall, A. (1995). “Rho, rac, and cdc42 GTPases regulate the assembly of multimolecular focal complexes associated with actin stress fibers, lamellipodia, and filopodia.” Cell 81: 5362.CrossRefGoogle ScholarPubMed
Nurden, A. T., Nurden, P., Sanchez, M., Andia, I. and Anitua, E. (2008). “Platelets and wound healing.” Front Biosci 13: 35323548.Google ScholarPubMed
Oakes, P. W., Beckham, Y., Stricker, J. and Gardel, M. L. (2012). “Tension is required but not sufficient for focal adhesion maturation without a stress fiber template.” J Cell Biol 196: 363374.CrossRefGoogle Scholar
Oberhauser, A. F., Badilla-Fernandez, C., Carrion-Vazquez, M. and Fernandez, J. M. (2002). “The mechanical hierarchies of fibronectin observed with single-molecule AFM.” J Mol Biol 319: 433447.CrossRefGoogle ScholarPubMed
Palazzo, A. F., Eng, C. H., Schlaepfer, D. D., Marcantonio, E. E. and Gundersen, G. G. (2004). “Localized stabilization of microtubules by integrin- and FAK-facilitated Rho signaling.” Science 303: 836839.CrossRefGoogle ScholarPubMed
Parker, M. W., Rossi, D., Peterson, M., Smith, K., Sikstrom, K., White, E. S., Connett, J. E. (2014). “Fibrotic extracellular matrix activates a profibrotic positive feedback loop.” J Clin Invest 124: 16221635.CrossRefGoogle ScholarPubMed
Pasapera, A. M., Schneider, I. C., Rericha, E., Schlaepfer, D. D. and Waterman, C. M. (2010). “Myosin II activity regulates vinculin recruitment to focal adhesions through FAK-mediated paxillin phosphorylation.” J Cell Biol 188: 877890.CrossRefGoogle ScholarPubMed
Paszek, M. J., Zahir, N., Johnson, K. R., Lakins, J. N., Rozenberg, G. I., Gefen, A., Reinhart-King, C. A., et al. (2005). “Tensional homeostasis and the malignant phenotype.” Cancer Cell 8: 241254.CrossRefGoogle ScholarPubMed
Pelham, R. J. Jr. and Wang, Y. (1997). “Cell locomotion and focal adhesions are regulated by substrate flexibility.” Proc Natl Acad Sci USA 94: 1366113665.CrossRefGoogle ScholarPubMed
Ponti, A., Machacek, M., Gupton, S. L., Waterman-Storer, C. M. and Danuser, G. (2004). “Two distinct actin networks drive the protrusion of migrating cells.” Science 305: 17821786.CrossRefGoogle ScholarPubMed
Popova, A. P., Bozyk, P. D., Goldsmith, A. M., Linn, M. J., Lei, J., Bentley, J. K. and Hershenson, M. B. (2010). “Autocrine production of TGF-beta1 promotes myofibroblastic differentiation of neonatal lung mesenchymal stem cells.” Am J Physiol Lung Cell Mol Physiol 298: L735L743.CrossRefGoogle ScholarPubMed
Ridley, A. J. and Hall, A. (1992). “The small GTP-binding protein rho regulates the assembly of focal adhesions and actin stress fibers in response to growth factors.” Cell 70: 389399.CrossRefGoogle ScholarPubMed
Riveline, D., Zamir, E., Balaban, N. Q., Schwarz, U. S., Ishizaki, T., Narumiya, S., Kam, Z., et al. (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: 11751186.CrossRefGoogle ScholarPubMed
Rossier, O., Octeau, V., Sibarita, J. B., Leduc, C., Tessier, B., Nair, D., Gatterdam, V., (2012). “Integrins beta1 and beta3 exhibit distinct dynamic nanoscale organizations inside focal adhesions.” Nat Cell Biol 14: 10571067.CrossRefGoogle ScholarPubMed
Rubashkin, M. G., Cassereau, L., Bainer, R., Dufort, C. C., Yui, Y., Ou, G., Paszek, M. J. (2014). “Force engages vinculin and promotes tumor progression by enhancing PI3K activation of phosphatidylinositol (3,4,5)-triphosphate.” Cancer Res 74: 45974611.CrossRefGoogle ScholarPubMed
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: L5254.CrossRefGoogle ScholarPubMed
Sanders, Y. Y., Kumbla, P. and Hagood, J. S. (2007). “Enhanced myofibroblastic differentiation and survival in Thy-1(-) lung fibroblasts.” Am J Respir Cell Mol Biol 36: 226235.CrossRefGoogle ScholarPubMed
Sanders, Y. Y., Pardo, A., Selman, M., Nuovo, G. J., Tollefsbol, T. O., Siegal, G. P. and Hagood, J. S. (2008). “Thy-1 promoter hypermethylation: a novel epigenetic pathogenic mechanism in pulmonary fibrosis.” Am J Respir Cell Mol Biol 39: 610618.CrossRefGoogle ScholarPubMed
Sansores, R. H., Ramirez-Venegas, A., Perez-Padilla, R., Montano, M., Ramos, C., Becerril, C., Gaxiola, M., et al. (1996). “Correlation between pulmonary fibrosis and the lung pressure-volume curve.” Lung 174: 315323.CrossRefGoogle ScholarPubMed
Santhanam, L., Tuday, E. C., Webb, A. K., Dowzicky, P., Kim, J. H., Oh, Y. J., Sikka, G., et al. (2010). “Decreased S-nitrosylation of tissue transglutaminase contributes to age-related increases in vascular stiffness.” Circ Res 107: 117125.CrossRefGoogle ScholarPubMed
Schiller, H. B., Hermann, M. R., Polleux, J., Vignaud, T., Zanivan, S., Friedel, C. C., Sun, Z., et al. 2013. “Beta1– and alphav-class integrins cooperate to regulate myosin II during rigidity sensing of fibronectin-based microenvironments.” Nat Cell Biol 15: 625636.CrossRefGoogle ScholarPubMed
Sharma, P., Varma, R., Sarasij, R. C., Ira, , Gousset, K., Krishnamoorthy, G., Rao, M. et al. (2004). “Nanoscale organization of multiple GPI-anchored proteins in living cell membranes.” Cell 116: 577589.CrossRefGoogle ScholarPubMed
Shattil, S. J. (2005). “Integrins and Src: dynamic duo of adhesion signaling.” Trends Cell Biol 15: 399403.CrossRefGoogle ScholarPubMed
Shima, T., Nada, S. and Okada, M. (2003). “Transmembrane phosphoprotein Cbp senses cell adhesion signaling mediated by Src family kinase in lipid rafts.” Proc Natl Acad Sci USA 100: 1489714902.CrossRefGoogle ScholarPubMed
Simons, K. and Gerl, M. J. (2010). “Revitalizing membrane rafts: new tools and insights.” Nat Rev Mol Cell Biol 11: 688699.CrossRefGoogle ScholarPubMed
Simons, K. and Ikonen, E. (1997). “Functional rafts in cell membranes.” Nature 387: 569572.CrossRefGoogle ScholarPubMed
Smilenov, L. B., Mikhailov, A., Pelham, R. J., Marcantonio, E. E. and Gundersen, G. G. (1999). “Focal adhesion motility revealed in stationary fibroblasts.” Science 286: 11721174.CrossRefGoogle ScholarPubMed
Smith, M. L., Gourdon, D., Little, W. C., Kubow, K. E., Eguiluz, R. A., Luna-Morris, S. and Vogel, V. (2007). “Force-induced unfolding of fibronectin in the extracellular matrix of living cells.” PLoS Biol: e268.CrossRefGoogle ScholarPubMed
Sohier, J., Carubelli, I., Sarathchandra, P., Latif, N., Chester, A. H. and Yacoub, M. H. (2014). “The potential of anisotropic matrices as substrate for heart valve engineering.” Biomaterials 35: 18331844.CrossRefGoogle ScholarPubMed
Solon, J., Levental, I., Sengupta, K., Georges, P. C. and Janmey, P. A. (2007). “Fibroblast adaptation and stiffness matching to soft elastic substrates.” Biophys J 93: 44534461.CrossRefGoogle ScholarPubMed
Sorrell, J. M. and Caplan, A. I. (2009). “Fibroblasts-a diverse population at the center of it all.” Int Rev Cell Mol Biol 276: 161214.CrossRefGoogle Scholar
Specks, U., Nerlich, A., Colby, T. V., Wiest, I. and Timpl, R. (1995). “Increased expression of type VI collagen in lung fibrosis.” Am J Respir Crit Care Med 151: 19561964.CrossRefGoogle ScholarPubMed
Stabenfeldt, S. E., Gourley, M., Krishnan, L., Hoying, J. B., Barker, T. H. (2012). “Engineering fobrin polymers through engagement of alternative polymerization mechanisms.” Biomaterials 33: 535544.CrossRefGoogle ScholarPubMed
Stricker, J., Aratyn-Schaus, Y., Oakes, P. W. and Gardel, M. L. (2011). “Spatiotemporal constraints on the force-dependent growth of focal adhesions.” Biophys J 100: 28832893.CrossRefGoogle ScholarPubMed
Swift, J., Ivanovska, I. L., Buxboim, A., Harada, T., Dingal, P. C., Pinter, J., Pajerowski, J. D., et al. (2013). “Nuclear lamin-A scales with tissue stiffness and enhances matrix-directed differentiation.” Science 341: 1240104.CrossRefGoogle ScholarPubMed
Takagi, J., Petre, B. M., Walz, T. and Springer, T. A. (2002). “Global conformational rearrangements in integrin extracellular domains in outside-in and inside-out signaling.” Cell 110: 599611.CrossRefGoogle ScholarPubMed
Tan, Y., Tajik, A., Chen, J., Jia, Q., Chowdhury, F., Wang, L., Chen, J., et al. (2014). “Matrix softness regulates plasticity of tumour-repopulating cells via H3K9 demethylation and Sox2 expression.” Nat Commun 5: 4619.CrossRefGoogle ScholarPubMed
Tee, S. Y., Fu, J., Chen, C. S. and Janmey, P. A. (2011). “Cell shape and substrate rigidity both regulate cell stiffness.” Biophys J 100: L25L27.CrossRefGoogle ScholarPubMed
Thievessen, I., Thompson, P. M., Berlemont, S., Plevock, K. M., Plotnikov, S. V., Zemljic-Harpf, A., Ross, R. S., et al. (2013). “Vinculin-actin interaction couples actin retrograde flow to focal adhesions, but is dispensable for focal adhesion growth.” J Cell Biol 202: 163177.CrossRefGoogle ScholarPubMed
Thomas, S. M. and Brugge, J. S. (1997). “Cellular functions regulated by Src family kinases.” Annu Rev Cell Dev Biol 13: 513609.CrossRefGoogle ScholarPubMed
Tomasek, J. J., Gabbiani, G., Hinz, B., Chaponnier, C. and Brown, R. A. (2002). “Myofibroblasts and mechano-regulation of connective tissue remodelling.” Nat Rev Mol Cell Biol 3: 349363.CrossRefGoogle ScholarPubMed
Trappmann, B. and Chen, C. S. (2013). “How cells sense extracellular matrix stiffness: a material’s perspective.” Curr Opin Biotechnol 24: 948953.CrossRefGoogle ScholarPubMed
Tse, J. R. and Engler, A. J. (2010). “Preparation of hydrogel substrates with tunable mechanical properties.” Curr Protoc Cell Biol Chapter 10: Unit 10.16.Google Scholar
Vadillo-Rodriguez, V., Bruque, J. M., Gallardo-Moreno, A. M. and Gonzalez-Martin, M. L. (2013). “Surface-dependent mechanical stability of adsorbed human plasma fibronectin on Ti6Al4V: domain unfolding and stepwise unraveling of single compact molecules.” Langmuir 29: 85548560.CrossRefGoogle ScholarPubMed
van den Bogaart, G., Meyenberg, K., Risselada, H. J., Amin, H., Willig, K. I., Hubrich, B. E., Dier, M., et al. (2011). “Membrane protein sequestering by ionic protein-lipid interactions.” Nature 479: 552555.CrossRefGoogle ScholarPubMed
van Zanten, T. S., Cambi, A., Koopman, M., Joosten, B., Figdor, C. G. and Garcia-Parajo, M. F. (2009). “Hotspots of GPI-anchored proteins and integrin nanoclusters function as nucleation sites for cell adhesion.” Proc Natl Acad Sci USA 106: 1855718562.CrossRefGoogle ScholarPubMed
Velasquez, L. S., Sutherland, L. B., Liu, Z., Grinnell, F., Kamm, K. E., Schneider, J. W., Olson, E. N. et al. (2013). “Activation of MRTF-A-dependent gene expression with a small molecule promotes myofibroblast differentiation and wound healing.” Proc Natl Acad Sci USA 110: 1685016855.CrossRefGoogle ScholarPubMed
Vinogradova, O., Velyvis, A., Velyviene, A., Hu, B., Haas, T., Plow, E. and Qin, J. (2002). “A structural mechanism of integrin alpha(IIb)beta(3) ‘inside-out’ activation as regulated by its cytoplasmic face.” Cell 110: 587597.CrossRefGoogle Scholar
von Wichert, G., Jiang, G., Kostic, A., De Vos, K., Sap, J. and Sheetz, M. P. (2003). “RPTP-alpha acts as a transducer of mechanical force on alphav/beta3-integrin-cytoskeleton linkages.” J Cell Biol 161: 143153.CrossRefGoogle ScholarPubMed
Wang, N., Butler, J. P. and Ingber, D. E. (1993). “Mechanotransduction across the cell surface and through the cytoskeleton.” Science 260: 11241127.CrossRefGoogle ScholarPubMed
Wang, X. and Ha, T. (2013). “Defining single molecular forces required to activate integrin and notch signaling.” Science 340: 991994.CrossRefGoogle ScholarPubMed
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: 10401045.CrossRefGoogle ScholarPubMed
Watanabe, N., Madaule, P., Reid, T., Ishizaki, T., Watanabe, G., Kakizuka, A., Saito, Y., et al. (1997). “p140mDia, a mammalian homolog of Drosophila diaphanous, is a target protein for Rho small GTPase and is a ligand for profilin.” EMBO J 16: 30443056.CrossRefGoogle ScholarPubMed
Wei, Y., Czekay, R. P., Robillard, L., Kugler, M. C., Zhang, F., Kim, K. K., Xiong, J. P., (2005). “Regulation of alpha5beta1 integrin conformation and function by urokinase receptor binding.” J Cell Biol 168: 501511.CrossRefGoogle ScholarPubMed
Wei, Y., Lukashev, M., Simon, D. I., Bodary, S. C., Rosenberg, S., Doyle, M. V. and Chapman, H. A. (1996). “Regulation of integrin function by the urokinase receptor.” Science 273: 15511555.CrossRefGoogle ScholarPubMed
Winer, J. P., Oake, S. and Janmey, P. A. (2009). “Non-linear elasticity of extracellular matrices enables contractile cells to communicate local position and orientation.” PLoS One 4: e6382.CrossRefGoogle ScholarPubMed
Wipff, P. J., Rifkin, D. B., Meister, J. J. and Hinz, B. (2007). “Myofibroblast contraction activates latent TGF-beta1 from the extracellular matrix.” J Cell Biol 179: 13111323.CrossRefGoogle ScholarPubMed
Wiseman, P. W., Brown, C. M., Webb, D. J., Hebert, B., Johnson, N. L., Squier, J. A., Ellisman, M. H. et al. (2004). “Spatial mapping of integrin interactions and dynamics during cell migration by image correlation microscopy.” J Cell Sci 117: 55215534.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: 583595.CrossRefGoogle Scholar
Wynn, T. A. (2007). “Common and unique mechanisms regulate fibrosis in various fibroproliferative diseases.” J Clin Invest 117: 524529.CrossRefGoogle ScholarPubMed
Yeung, T., Georges, P. C., Flanagan, L. A., Marg, B., Ortiz, M., Funaki, M., Zahir, N., et al. (2005). “Effects of substrate stiffness on cell morphology, cytoskeletal structure, and adhesion.” Cell Motil Cytoskeleton 60: 2434.CrossRefGoogle ScholarPubMed
Yoshitake, H., Takeda, Y., Nitto, T., Sendo, F. and Araki, Y. (2003). “GPI-80, a beta2 integrin associated glycosylphosphatidylinositol-anchored protein, concentrates on pseudopodia without association with beta2 integrin during neutrophil migration.” Immunobiology 208: 391399.CrossRefGoogle ScholarPubMed
Zhang, Y., Fan, W., Ma, Z., Wu, C., Fang, W., Liu, G. and Xiao, Y. (2010). “The effects of pore architecture in silk fibroin scaffolds on the growth and differentiation of mesenchymal stem cells expressing BMP7.” Acta Biomater 6: 30213028.CrossRefGoogle ScholarPubMed
Zhou, Y., Hagood, J. S. and Murphy-Ullrich, J. E. (2004). “Thy-1 expression regulates the ability of rat lung fibroblasts to activate transforming growth factor-beta in response to fibrogenic stimuli.” Am J Pathol 165: 659669.CrossRefGoogle ScholarPubMed

References

Araujo, A., and Walker, J. W. (1994). Kinetics of tension development in skinned cardiac myocytes measured by photorelease of Ca2+. Am J Physiol 267: H1643H1653.Google ScholarPubMed
Azeloglu, E. U., and Costa, K. D. (2010). Cross-bridge cycling gives rise to spatiotemporal heterogeneity of dynamic subcellular mechanics in cardiac myocytes probed with atomic force microscopy. Am J Physiol Heart Circ Physiol 298: H853H860.CrossRefGoogle ScholarPubMed
Balaban, N. Q., Schwarz, U. S., Riveline, D., et al. (2001). Force and focal adhesion assembly: a close relationship studied using elastic micropatterned substrates. Nature Cell Biology 3: 466472.CrossRefGoogle ScholarPubMed
Bluhm, W. F., Mcculloch, A. D., and Lew, W. Y. (1995). Active force in rabbit ventricular myocytes. J Biomech 28: 11191122.CrossRefGoogle ScholarPubMed
Borg, T. K., Rubin, K., Lundgren, E., Borg, K., and Obrink, B. (1984). Recognition of extracellular matrix components by neonatal and adult cardiac myocytes. Dev Biol 104: 8696.CrossRefGoogle ScholarPubMed
Boudou, T., Legant, W. R., Mu, A. B., et al. (2012). A microfabricated platform to measure and manipulate the mechanics of engineered cardiac microtissues. Tissue Engineering Part A 18: 910919.CrossRefGoogle ScholarPubMed
Brady, A. J. (1991). Mechanical properties of isolated cardiac myocytes. Physiol Rev 71: 413428.CrossRefGoogle ScholarPubMed
Brady, A. J., Tan, S. T., and Ricchiuti, N. V. (1979). Contractile force measured in unskinned isolated adult rat heart fibres. Nature 282: 728729.CrossRefGoogle ScholarPubMed
Braunwald, E., and Bonow, R. O. (2012). Braunwald’s heart disease: a textbook of cardiovascular medicine. Philadelphia: Saunders.Google Scholar
Bray, M. A., Sheehy, S. P., and Parker, K. K. (2008). Sarcomere alignment is regulated by myocyte shape. Cell Motility and the Cytoskeleton 65: 641651.CrossRefGoogle ScholarPubMed
Burridge, P. W., Keller, G., Gold, J. D., and Wu, J. C. (2012). Production of de novo cardiomyocytes: human pluripotent stem cell differentiation and direct reprogramming. Cell Stem Cell 10: 1628.CrossRefGoogle ScholarPubMed
Chang, W. T., Yu, D., Lai, Y. C., Lin, K. Y., and Liau, I. (2013). Characterization of the mechanodynamic response of cardiomyocytes with atomic force microscopy. Anal Chem 85: 13951400.CrossRefGoogle ScholarPubMed
Chattergoon, N. N., Giraud, G. D., Louey, S., et al. (2012). Thyroid hormone drives fetal cardiomyocyte maturation. FASEB J 26: 397408.CrossRefGoogle ScholarPubMed
Cheng, Q., Sun, Z., Meininger, G. A., and Almasri, M. (2010). Note: Mechanical study of micromachined polydimethylsiloxane elastic microposts. Rev Sci Instrum 81: 106104.CrossRefGoogle ScholarPubMed
Copelas, L., Briggs, M., Grossman, W., and Morgan, J. P. (1987). A method for recording isometric tension development by isolated cardiac myocytes: transducer attachment with fibrin glue. Pflugers Arch 408: 315317.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: 14111420.CrossRefGoogle Scholar
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: 20082022.CrossRefGoogle ScholarPubMed
Dillmann, W. H. (2002). Cellular action of thyroid hormone on the heart. Thyroid 12: 447452.CrossRefGoogle ScholarPubMed
Domke, J., Parak, W. J., George, M., Gaub, H. E., and Radmacher, M. (1999). Mapping the mechanical pulse of single cardiomyocytes with the atomic force microscope. Eur Biophys J 28: 179186.CrossRefGoogle ScholarPubMed
Engler, A. J., Carag-Krieger, C., Johnson, C. P., et al. (2008). Embryonic cardiomyocytes beat best on a matrix with heart-like elasticity: scar-like rigidity inhibits beating. J Cell Sci 121: 37943802.CrossRefGoogle Scholar
Eschenhagen, T., Fink, C., Remmers, U., et al. (1997). Three-dimensional reconstitution of embryonic cardiomyocytes in a collagen matrix: a new heart muscle model system. FASEB J 11: 683694.CrossRefGoogle Scholar
Fabiato, A., and Fabiato, F. (1975). Contractions induced by a calcium-triggered release of calcium from the sarcoplasmic reticulum of single skinned cardiac cells. J Physiol 249: 469495.CrossRefGoogle ScholarPubMed
Forte, G., Pagliari, S., Ebara, M., et al. (2012). Substrate stiffness modulates gene expression and phenotype in neonatal cardiomyocytes in vitro. Tissue Eng Part A 18: 18371848.CrossRefGoogle ScholarPubMed
Garcia-Webb, M. G., Taberner, A. J., Hogan, N. C., and Hunter, I. W. (2007). A modular instrument for exploring the mechanics of cardiac myocytes. Am J Physiol Heart Circ Physiol 293: H866H874.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: 177179.CrossRefGoogle Scholar
Hasenfuss, G., Mulieri, L. A., Blanchard, E. M., et al. (1991). Energetics of isometric force development in control and volume-overload human myocardium. Comparison with animal species. Circulation Research 68: 836846.CrossRefGoogle ScholarPubMed
Hazeltine, L. B., Simmons, C. S., Salick, M. R., et al. (2012). Effects of substrate mechanics on contractility of cardiomyocytes generated from human pluripotent stem cells. Int J Cell Biol 2012: 508294.CrossRefGoogle ScholarPubMed
Hersch, N., Wolters, B., Dreissen, G., et al. (2013). The constant beat: cardiomyocytes adapt their forces by equal contraction upon environmental stiffening. Biol Open 2: 351361.CrossRefGoogle ScholarPubMed
Iribe, G., Helmes, M., and Kohl, P. (2007). Force-length relations in isolated intact cardiomyocytes subjected to dynamic changes in mechanical load. Am J Physiol Heart Circ Physiol 292: H1487H1497.CrossRefGoogle ScholarPubMed
Ivashchenko, C. Y., Pipes, G. C., Lozinskaya, I. M., et al. (2013). Human-induced pluripotent stem cell-derived cardiomyocytes exhibit temporal changes in phenotype. Am J Physiol Heart Circ Physiol 305: H913H922.CrossRefGoogle ScholarPubMed
Jacot, J. G., Martin, J. C., and Hunt, D. L. (2010). Mechanobiology of cardiomyocyte development. J Biomech 43: 9398.CrossRefGoogle ScholarPubMed
Jacot, J. G., Mcculloch, A. D., and Omens, J. H. (2008). Substrate stiffness affects the functional maturation of neonatal rat ventricular myocytes. Biophys J 95: 34793487.CrossRefGoogle ScholarPubMed
Kajzar, A., Cesa, C. M., Kirchgessner, N., Hoffmann, B., and Merkel, R. (2008). Toward physiological conditions for cell analyses: forces of heart muscle cells suspended between elastic micropillars. Biophys J 94: 18541866.CrossRefGoogle ScholarPubMed
Kim, J., Park, J., Na, K., et al. (2008). Quantitative evaluation of cardiomyocyte contractility in a 3D microenvironment. J Biomech 41: 23962401.CrossRefGoogle Scholar
Kim, J., Park, J., Ryu, S. K., et al. (2006). Realistic computational modeling for hybrid biopolymer microcantilevers. Conf Proc IEEE Eng Med Biol Soc 1: 21022105.CrossRefGoogle Scholar
Kim, K., Taylor, R., Sim, J. Y., et al. (2011). Calibrated micropost arrays for biomechanical characterisation of cardiomyocytes. Micro and Nano Letters 6: 317.CrossRefGoogle Scholar
Klein, I., Daood, M., and Whiteside, T. (1985). Development of heart cells in culture: studies using an affinity purified antibody to a myosin light chain. J Cell Physiol 124: 4953.CrossRefGoogle ScholarPubMed
Klein, I., and Ojamaa, K. (2001). Thyroid hormone and the cardiovascular system. N Engl J Med 344: 501509.CrossRefGoogle ScholarPubMed
Korte, F. S., Dai, J., Buckley, K., et al. (2011). Upregulation of cardiomyocyte ribonucleotide reductase increases intracellular 2 deoxy-ATP, contractility, and relaxation. J Mol Cell Cardiol 51: 894901.CrossRefGoogle ScholarPubMed
Kraft, T., and Brenner, B. (1997). Force enhancement without changes in cross-bridge turnover kinetics: the effect of EMD 57033. Biophys J 72: 272281.CrossRefGoogle ScholarPubMed
Kresh, J. Y., and Chopra, A. (2011). Intercellular and extracellular mechanotransduction in cardiac myocytes. Pflugers Arch 462: 7587.CrossRefGoogle ScholarPubMed
Kruger, M., Sachse, C., Zimmermann, W. H., et al. (2008). Thyroid hormone regulates developmental titin isoform transitions via the phosphatidylinositol-3-kinase/ AKT pathway. Circ Res 102: 439447.CrossRefGoogle ScholarPubMed
Le Guennec, J. Y., Peineau, N., Argibay, J. A., Mongo, K. G., and Garnier, D. (1990). A new method of attachment of isolated mammalian ventricular myocytes for tension recording: length dependence of passive and active tension. J Mol Cell Cardiol 22: 10831093.CrossRefGoogle ScholarPubMed
Lee, J., Leonard, M., Oliver, T., Ishihara, A., and Jacobson, K. (1994). Traction forces generated by locomoting keratocytes. J Cell Biol 127: 19571964.CrossRefGoogle ScholarPubMed
Lee, Y. K., Ng, K. M., Chan, Y. C., et al. (2010). Triiodothyronine promotes cardiac differentiation and maturation of embryonic stem cells via the classical genomic pathway. Mol Endocrinol 24: 17281736.CrossRefGoogle ScholarPubMed
Lin, G., Palmer, R. E., Pister, K. S., and Roos, K. P. (2001). Miniature heart cell force transducer system implemented in MEMS technology. IEEE Trans Biomed Eng 48: 9961006.Google ScholarPubMed
Lin, G., Pister, K. S. J., and Roos, K. P. (1995). Microscale force-transducer system to quantify isolated heart cell contractile characteristics. Sensors and Actuators A: Physical 46: 233236.CrossRefGoogle Scholar
Lin, G., Pister, K. S. J., and Roos, K. P. (2000). Surface micromachined polysilicon heart cell force transducer. Journal of Microelectromechanical Systems 9: 917.CrossRefGoogle Scholar
Lipscomb-Allhouse, S., Mulligan, I. P., and Ashley, C. C. (2001). The effects of the inotropic agent EMD 57033 on activation and relaxation kinetics in frog skinned skeletal muscle. Pflugers Arch 442: 171177.Google ScholarPubMed
Liu, J., Sun, N., Bruce, M. A., Wu, J. C., and Butte, M. J. (2012a). Atomic force mechanobiology of pluripotent stem cell-derived cardiomyocytes. PLoS One 7: e37559.CrossRefGoogle ScholarPubMed
Liu, Y., Feng, J., Shi, L., et al. (2012b). In situ mechanical analysis of cardiomyocytes at nano scales. Nanoscale 4: 99102.CrossRefGoogle ScholarPubMed
Lundgren, E., Terracio, L., and Borg, T. K. (1985a). Adhesion of cardiac myocytes to extracellular matrix components. Basic Res Cardiol 80(Suppl 1): 6974.Google ScholarPubMed
Lundgren, E., Terracio, L., Mardh, S. and Borg, T. K. (1985b). Extracellular matrix components influence the survival of adult cardiac myocytes in vitro. Exp Cell Res 158: 371381.CrossRefGoogle ScholarPubMed
Luo, C. H., and Tung, L. (1991). Null-balance transducer for isometric force measurements and length control of single heart cells. IEEE Trans Biomed Eng 38: 11651174.CrossRefGoogle ScholarPubMed
Majkut, S. F., and Discher, D. E. (2012). Cardiomyocytes from late embryos and neonates do optimal work and striate best on substrates with tissue-level elasticity: metrics and mathematics. Biomech Model Mechanobiol 11: 12191225.CrossRefGoogle Scholar
Mann, J. M., Lam, R. H., Weng, S., Sun, Y., and Fu, J. (2012). A silicone-based stretchable micropost array membrane for monitoring live-cell subcellular cytoskeletal response. Lab Chip 12: 731740.CrossRefGoogle ScholarPubMed
McDermott, P. J., and Morgan, H. E. (1989). Contraction modulates the capacity for protein synthesis during growth of neonatal heart cells in culture. Circ Res 64: 542553.CrossRefGoogle ScholarPubMed
McDonald, K. S., Leiden, J. M., Field, L. J., et al. (1993). Length-dependence of Ca-2+ sensitivity of tension in transgenic mouse myocytes expressing skeletal troponin-C. Circulation 88: 8686.Google Scholar
Metzger, J. M. (1995). Myosin binding-induced cooperative activation of the thin filament in cardiac myocytes and skeletal muscle fibers. Biophys J 68: 14301442.CrossRefGoogle ScholarPubMed
Nishimura, S., Yasuda, S., Katoh, M., et al. (2004). Single cell mechanics of rat cardiomyocytes under isometric, unloaded, and physiologically loaded conditions. Am J Physiol Heart Circ Physiol 287: H196H202.CrossRefGoogle ScholarPubMed
Nowakowski, S. G., Kolwicz, S. C., Korte, F. S., et al. (2013). Transgenic overexpression of ribonucleotide reductase improves cardiac performance. Proc Natl Acad Sci USA 110: 61876192.CrossRefGoogle ScholarPubMed
Oliver, T., Lee, J., and Jacobson, K. 1994. Forces exerted by locomoting cells. Semin Cell Biol 5: 139147.CrossRefGoogle ScholarPubMed
Park, J., Kim, I. C., Cha, J., et al. (2007). Mechanotransduction of cardiomyocytes interacting with a thin membrane transducer. Journal of Micromechanics and Microengineering 17: 11621167.CrossRefGoogle Scholar
Park, J., Kim, J., Roh, D., et al. (2006). Fabrication of complex 3D polymer structures for cell-polymer hybrid systems. Journal of Micromechanics and Microengineering 16: 16141619.CrossRefGoogle Scholar
Park, J., Ryu, J., Choi, S. K., et al. (2005). Real-time measurement of the contractile forces of self-organized cardiomyocytes on hybrid biopolymer microcantilevers. Anal Chem 77: 65716580.CrossRefGoogle ScholarPubMed
Parker, K. K., and Ingber, D. E. (2007). Extracellular matrix, mechanotransduction and structural hierarchies in heart tissue engineering. Philos Trans R Soc Lond B Biol Sci 362: 12671279.CrossRefGoogle ScholarPubMed
Pillekamp, F., Halbach, M., Reppel, M., et al. (2007a). Neonatal murine heart slices. A robust model to study ventricular isometric contractions. Cell Physiol Biochem 20: 837846.CrossRefGoogle ScholarPubMed
Pillekamp, F., Reppel, M., Rubenchyk, O., et al. (2007b). Force measurements of human embryonic stem cell-derived cardiomyocytes in an in vitro transplantation model. Stem Cells 25: 174180.CrossRefGoogle Scholar
Puceat, M., Clement, O., Lechene, P., et al. (1990). Neurohormonal control of calcium sensitivity of myofilaments in rat single heart cells. Circ Res 67: 517524.CrossRefGoogle ScholarPubMed
Qin, L., Huang, J. Y., Xiong, C. Y., Zhang, Y. Y., and Fang, J. (2007). Dynamical stress characterization and energy evaluation of single cardiac myocyte actuating on flexible substrate. Biochem Biophys Res Commun 360: 352356.CrossRefGoogle ScholarPubMed
Regnier, M., and Homsher, E. (1998a). The effect of ATP analogs on posthydrolytic and force development steps in skinned skeletal muscle fibers. Biophys J 74: 30593071.CrossRefGoogle ScholarPubMed
Regnier, M., Lee, D. M. and Homsher, E. (1998b). ATP analogs and muscle contraction: mechanics and kinetics of nucleoside triphosphate binding and hydrolysis. Biophys J 74: 30443058.CrossRefGoogle ScholarPubMed
Regnier, M., Martyn, D. A. and Chase, P. B. (1998c). Calcium regulation of tension redevelopment kinetics with 2-deoxy-ATP or low [ATP] in rabbit skeletal muscle. Biophys J 74: 20052015.CrossRefGoogle ScholarPubMed
Regnier, M., Rivera, A. J., Chen, Y., and Chase, P. B. (2000). 2-deoxy-ATP enhances contractility of rat cardiac muscle. Circ Res 86: 12111217.CrossRefGoogle ScholarPubMed
Rodriguez, A. G., Han, S. J., Regnier, M., and Sniadecki, N. J. (2011). Substrate stiffness increases twitch power of neonatal cardiomyocytes in correlation with changes in myofibril structure and intracellular calcium. Biophys J 101: 24552464.CrossRefGoogle ScholarPubMed
Rodriguez, A. G., Rodriguez, M. L., Han, S. J., Sniadecki, N. J., and Regnier, M. (2013). Enhanced contractility with 2-deoxy-ATP and EMD 57033 is correlated with reduced myofibril structure and twitch power in neonatal cardiomyocytes. Integr Biol (Camb) 5: 13661373.CrossRefGoogle ScholarPubMed
Rodriguez, M., Graham, B. T., Pabon, L. M., et al. (2014). Measuring the contractile forces of human induced pluripotent stem cell-derived cardiomyocytes with arrays of microposts. J Biomech Eng 136: 051005.CrossRefGoogle ScholarPubMed
Schaaf, S., Shibamiya, A., Mewe, M., et al. (2011). Human engineered heart tissue as a versatile tool in basic research and preclinical toxicology. Plos One 6.CrossRefGoogle ScholarPubMed
Schoen, I., Hu, W., Klotzsch, E., and Vogel, V. (2010). Probing cellular traction forces by micropillar arrays: contribution of substrate warping to pillar deflection. Nano Lett 10 182318182330.CrossRefGoogle ScholarPubMed
Shepherd, N., Vornanen, M., and Isenberg, G. (1990). Force measurements from voltage-clamped guinea pig ventricular myocytes. Am J Physiol 258: H452H459.Google ScholarPubMed
Shroff, S. G., Saner, D. R., and Lal, R. (1995). Dynamic micromechanical properties of cultured rat atrial myocytes measured by atomic force microscopy. Am J Physiol 269: C286C292.CrossRefGoogle ScholarPubMed
Simpson, D. G., Majeski, M., Borg, T. K., and Terracio, L. (1999). Regulation of cardiac myocyte protein turnover and myofibrillar structure in vitro by specific directions of stretch. Circ Res 85: e59e69.CrossRefGoogle ScholarPubMed
Simpson, D. G., Sharp, W. W., Borg, T. K., et al. (1996). Mechanical regulation of cardiac myocyte protein turnover and myofibrillar structure. Am J Physiol 270: C1075C1087.CrossRefGoogle ScholarPubMed
Sniadecki, N. J., et al. (2007a). Magnetic microposts as an approach to apply forces to living cells. Proc Natl Acad Sci USA, 104: 1455314558.CrossRefGoogle ScholarPubMed
Sniadecki, N. J. et al. (2007b). Microfabricated silicone elastomeric post arrays for measuring traction forces of adherent cells. Methods in Cell Biology 83: 313328.CrossRefGoogle ScholarPubMed
Solaro, R. J., Gambassi, G., Warshaw, D. M., et al. (1993). Stereoselective actions of thiadiazinones on canine cardiac myocytes and myofilaments. Circ Res 73: 981990.CrossRefGoogle ScholarPubMed
Sonnenblick, E. H. (1962). Implications of muscle mechanics in the heart. Fed Proc 21: 975990.Google ScholarPubMed
Soufivand, A. A., Soleimani, M., and Navidbakhsh, M. (2014). Is it appropriate to apply Hertz model to describe cardiac myocytes’ mechanical properties by atomic force microscopy nanoindentation? Micro & Nano Letters 9: 153156.CrossRefGoogle Scholar
Strang, K. T., Sweitzer, N. K., Greaser, M. L., and Moss, R. L. (1994). Beta-adrenergic receptor stimulation increases unloaded shortening velocity of skinned single ventricular myocytes from rats. Circ Res 74: 542549.CrossRefGoogle ScholarPubMed
Strauss, J. D., Bletz, C., and Ruegg, J. C. (1994). The calcium sensitizer EMD 53998 antagonizes phosphate-induced increases in energy cost of isometric tension in cardiac skinned fibres. Eur J Pharmacol 252: 219224.CrossRefGoogle ScholarPubMed
Tan, J. L., Tien, J., Pirone, D. M., et al. (2003). Cells lying on a bed of microneedles: an approach to isolate mechanical force. Proc Natl Acad Sci USA 100: 14841489.CrossRefGoogle Scholar
Tanaka, Y., Morishima, K., Shimizu, T., et al. (2006). Demonstration of a PDMS-based bio-microactuator using cultured cardiomyocytes to drive polymer micropillars. Lab Chip 6: 230235.CrossRefGoogle ScholarPubMed
Tarr, M., Trank, J. W., Leiffer, P., and Shepherd, N. (1981). Evidence that the velocity of sarcomere shortening in single frog atrial cardiac cells is load dependent. Circ Res 48: 200206.CrossRefGoogle ScholarPubMed
Tasche, C., Meyhofer, E., and Brenner, B. (1999). A force transducer for measuring mechanical properties of single cardiac myocytes. Am J Physiol 277: H2400H4008.Google ScholarPubMed
Taylor, R. E., Kim, K., Sun, N., et al. (2013). Sacrificial layer technique for axial force post assay of immature cardiomyocytes. Biomed Microdevices 15: 171181.CrossRefGoogle ScholarPubMed
Terracio, L., Rubin, K., Gullberg, D., et al. (1991). Expression of collagen binding integrins during cardiac development and hypertrophy. Circ Res 68: 734744.CrossRefGoogle ScholarPubMed
Tulloch, N. L., Muskheli, V., Razumova, M. V., et al. (2011). Growth of engineered human myocardium with mechanical loading and vascular coculture. Circ Res 109: 4759.CrossRefGoogle ScholarPubMed
Tung, L. (1986). An ultrasensitive transducer for measurement of isometric contractile force from single heart cells. Pflugers Arch 407: 109115.CrossRefGoogle ScholarPubMed
van der Velden, J., Klein, L. J., Van Der Bijl, M., et al. (1998). Force production in mechanically isolated cardiac myocytes from human ventricular muscle tissue. Cardiovasc Res 38: 414423.CrossRefGoogle ScholarPubMed
Vannier, C., Chevassus, H., and Vassort, G. (1996). Ca-dependence of isometric force kinetics in single skinned ventricular cardiomyocytes from rats. Cardiovasc Res 32: 580586.CrossRefGoogle ScholarPubMed
Wang, I. N., Wang, X., Ge, X., et al. (2012). Apelin enhances directed cardiac differentiation of mouse and human embryonic stem cells. PLoS One 7: e38328.CrossRefGoogle ScholarPubMed
Xi, J., Khalil, M., Shishechian, N., et al. (2010). Comparison of contractile behavior of native murine ventricular tissue and cardiomyocytes derived from embryonic or induced pluripotent stem cells. FASEB J 24: 27392751.CrossRefGoogle ScholarPubMed
Xing, R., Li, S., Liu, K., et al. (2014). HIP-55 negatively regulates myocardial contractility at the single-cell level. J Biomech 47: 27152720.CrossRefGoogle ScholarPubMed
Yang, X., Pabon, L. and Murry, C. E. (2014a). Engineering adolescence: maturation of human pluripotent stem cell-derived cardiomyocytes. Circ Res 114: 511523.CrossRefGoogle ScholarPubMed
Yang, X., Rodriguez, M., Pabon, L., et al. (2014b). Tri-iodo-l-thyronine promotes the maturation of human cardiomyocytes-derived from induced pluripotent stem cells. J Mol Cell Cardiol 72: 296304.CrossRefGoogle ScholarPubMed
Yasuda, S. I., Sugiura, S., Kobayakawa, N., et al. (2001). A novel method to study contraction characteristics of a single cardiac myocyte using carbon fibers. Am J Physiol Heart Circ Physiol 281: H1442H1446.CrossRefGoogle ScholarPubMed
Yin, S., Zhang, X., Zhan, C., et al. (2005). Measuring single cardiac myocyte contractile force via moving a magnetic bead. Biophys J 88: 14891495.CrossRefGoogle ScholarPubMed
Zhao, Y., and Zhang, X. (2005). Contraction force measurements in cardiac myocytes using PDMS pillar arrays. 18th IEEE International Conference on Micro Electro Mechanical Systems: 834837.Google Scholar
Zhao, Y., and Zhang, X. (2006). Cellular mechanics study in cardiac myocytes using PDMS pillars array. Sensors and Actuators A: Physical 125: 398404.CrossRefGoogle Scholar

References

Babich, A., Li, S., O’Connor, R. S., Milone, M. C., Freedman, B. D. and Burkhardt, J. K. (2012). “F-actin polymerization and retrograde flow drive sustained PLCγ1 signaling during T cell activation.” The Journal of Cell Biology 197: 775787.CrossRefGoogle ScholarPubMed
Banchereau, J. and Steinman, R. M. (1998). “Dendritic cells and the control of immunity.” Nature 392: 245252.CrossRefGoogle ScholarPubMed
Bashour, K. T., Gondarenko, A., Chen, H., Shen, K., Liu, X., Huse, M., Hone, J. C. and Kam, L. C. (2014a). “CD28 and CD3 have complementary roles in T-cell traction forces.” Proceedings of the National Academy of Sciences 111: 22412246.CrossRefGoogle ScholarPubMed
Bashour, K. T., Tsai, J., Shen, K., Lee, J.-H., Sun, E., Milone, M. C., Dustin, M. L. and Kam, L. C. (2014b). “Cross talk between CD3 and CD28 is spatially modulated by protein lateral mobility.” Molecular and Cellular Biology 34: 955964.CrossRefGoogle ScholarPubMed
Brossard, C., Feuillet, V., Schmitt, A., Randriamampita, C., Romao, M., Raposo, G. and Trautmann, A. (2005). “Multifocal structure of the T cell – dendritic cell synapse.” European Journal of Immunology 35: 17411753.CrossRefGoogle Scholar
Campi, G., Varma, R. and Dustin, M. L. (2005). “Actin and agonist MHC-peptide complex-dependent T cell receptor microclusters as scaffolds for signaling.” Journal of Experimental Medicine 202: 10311036.CrossRefGoogle ScholarPubMed
Cemerski, S. and Shaw, A. (2006). “Immune synapses in T-cell activation.” Current Opinion in Immunology 18: 298304.CrossRefGoogle ScholarPubMed
Chang, J. T., Ciocca, M. L., Kinjyo, I., Palanivel, V. R., Mcclurkin, C. E., Dejong, C. S., Mooney, E. C., et al. (2011). “Asymmetric proteasome segregation as a mechanism for unequal partitioning of the transcription factor T-bet during T lymphocyte division.” Immunity 34: 492504.CrossRefGoogle ScholarPubMed
Chang, J. T., Palanivel, V. R., Kinjyo, I., Schambach, F., Intlekofer, A. M., Banerjee, A., Longworth, S. A., et al. (2007). “Asymmetric T lymphocyte division in the initiation of adaptive immune responses.” Science 315: 16871691.CrossRefGoogle ScholarPubMed
Deeg, J., Axmann, M., Matic, J., Liapis, A., Depoil, D., Afrose, J., Curado, S., et al. (2013). “T cell activation is determined by the number of presented antigens.” Nano Letters 13: 56195626.CrossRefGoogle ScholarPubMed
Delcassian, D., Depoil, D., Rudnicka, D., Liu, M., Davis, D. M., Dustin, M. L. and Dunlop, I. E. (2013). “Nanoscale ligand spacing influences receptor triggering in T cells and NK cells.” Nano Letters 13: 56085614.CrossRefGoogle Scholar
Demond, A. L., Mossman, K. D., Starr, T., Dustin, M. L. and Groves, J. T. (2008). “T cell receptor microcluster transport through molecular mazes reveals mechanism of translocation.” Biophysical Journal 94: 32863292.CrossRefGoogle ScholarPubMed
Doh, J. and Irvine, D. J. (2006). “Immunological synapse arrays: patterned protein surfaces that modulate immunological synapse structure formation in T cells.” Proceedings of the National Academy of Sciences 103: 57005705.CrossRefGoogle ScholarPubMed
Dustin, M. L. and Groves, J. T. (2012). “Receptor signaling clusters in the immune synapse.” Annual Review of Biophysics 41: 543.CrossRefGoogle ScholarPubMed
Freiberg, B. A., Kupfer, H., Maslanik, W., Delli, J., Kappler, J., Zaller, D. M. and Kupfer, A. (2002). “Staging and resetting T cell activation in SMACs.” Nature Immunology 3: 911917.CrossRefGoogle ScholarPubMed
Fu, J., Wang, Y. K., Yang, M. T., Desai, R. A., Yu, X., Liu, Z. and Chen, C. S. (2010). “Mechanical regulation of cell function with geometrically modulated elastomeric substrates.” Nature Methods 7: 733736.CrossRefGoogle ScholarPubMed
Glass, R., Möller, M. and Spatz, J. P. (2003). “Block copolymer micelle nanolithography.” Nanotechnology 14: 1153.CrossRefGoogle Scholar
Grakoui, A., Bromley, S. K., Sumen, C., Davis, M. M., Shaw, A. S., Allen, P. M. and Dustin, M. L. (1999). “The immunological synapse: a molecular machine controlling T cell activation.” Science 285: 221227.CrossRefGoogle ScholarPubMed
Groves, J. T. and Boxer, S. G. (2002). “Micropattern formation in supported lipid membranes.” Accounts of Chemical Research 35: 149157.CrossRefGoogle ScholarPubMed
Groves, J. T. and Dustin, M. L. (2003). “Supported planar bilayers in studies on immune cell adhesion and communication.” Journal of Immunological Methods 278: 1932.CrossRefGoogle ScholarPubMed
Guermonprez, P., Valladeau, J., Zitvogel, L., Th Ry, C. and Amigorena, S. (2002). “Antigen presentation and T cell stimulation by dendritic cells.” Annual Review of Immunology 20: 621667.CrossRefGoogle Scholar
Hallman, E., Burack, W. R., Shaw, A. S., Dustin, M. L. and Allen, P. M. (2002). “Immature CD4(+)CD8(+) thymocytes form a multifocal immunological synapse with sustained tyrosine phosphorylation.” Immunity 16: 839848.CrossRefGoogle Scholar
Huppa, J. B. and Davis, M. M. (2003). “T-cell-antigen recognition and the immunological synapse.” Nature Reviews Immunology 3: 973983.CrossRefGoogle ScholarPubMed
Irvine, D. J. and Doh, J. (2007). “Synthetic surfaces as artificial antigen presenting cells in the study of T cell receptor triggering and immunological synapse formation.” Seminars in Immunology 19: 245254.CrossRefGoogle Scholar
Jung, H.-R., Song, K. H., Chang, J. T. and Doh, J. (2014). “Geometrically controlled asymmetric division of CD4+ T cells studied by immunological synapse arrays.” PloS One 9: e91926.CrossRefGoogle ScholarPubMed
Jung, H.-R., Choi, J. C., Cho, W. and Doh, J. (2013). “Microfabricated platforms to modulate and monitor T cell synapse assembly.” Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology 5: 6774.Google ScholarPubMed
Krummel, M. F., Sjaastad, M. D., Wulfing, C. and Davis, M. M. (2000). “Differential clustering of CD4 and CD3 zeta during T cell recognition.” Science 289: 13491352.CrossRefGoogle ScholarPubMed
Luther, S. A. and Cyster, J. G. (2001). “Chemokines as regulators of T cell differentiation.” Nature Immunology 2: 102107.CrossRefGoogle ScholarPubMed
Manz, B. N., Jackson, B. L., Petit, R. S., Dustin, M. L. and Groves, J. (2011). “T-cell triggering thresholds are modulated by the number of antigen within individual T-cell receptor clusters.” Proceedings of the National Academy of Sciences 108: 90899094.CrossRefGoogle ScholarPubMed
Matic, J., Deeg, J., Scheffold, A., Goldstein, I. and Spatz, J. P. (2013). “Fine tuning and efficient T cell activation with stimulatory aCD3 nanoarrays.” Nano Letters 13: 50905097.CrossRefGoogle ScholarPubMed
Mempel, T. R., Henrickson, S. E. and von Andrian, U. H. (2004). “T-cell priming by dendritic cells in lymph nodes occurs in three distinct phases.” Nature 427 154159.CrossRefGoogle ScholarPubMed
Miller, M. J., Safrina, O., Parker, I. and Cahalan, M. D. (2004). “Imaging the single cell dynamics of CD4+ T cell activation by dendritic cells in lymph nodes.” The Journal of Experimental Medicine 200: 847856.CrossRefGoogle ScholarPubMed
Milone, M. C. and Kam, L. C. (2013). “Investigative and clinical applications of synthetic immune synapses.” Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology 5: 7585.Google ScholarPubMed
Monks, C. R. F., Freiberg, B. A., Kupfer, H., Sciaky, N. and Kupfer, A. (1998). “Three-dimensional segregation of supramolecular activation clusters in T cells.” Nature 395: 8286.CrossRefGoogle ScholarPubMed
Mossman, K. D., Campi, G., Groves, J. T. and Dustin, M. L. (2005). “Altered TCR signaling from geometrically repatterned immunological synapses.” Science 310: 11911193.CrossRefGoogle ScholarPubMed
Murphy, K. M. and Stockinger, B. (2010). “Effector T cell plasticity: flexibility in the face of changing circumstances.” Nature Immunology 11: 674680.CrossRefGoogle ScholarPubMed
O’Keefe, J. P., Blaine, K., Alegre, M. L. and Gajewski, T. F. (2004). “Formation of a central supramolecular activation cluster is not required for activation of naive CD8(+) T cells.” Proceedings of the National Academy of Sciences of the United States of America 101: 93519356.CrossRefGoogle Scholar
Purtic, B., Pitcher, L. A., van Oers, N. S. C. and Wulfing, C. (2005). “T cell receptor (TCR) clustering in the immunological synapse integrates TCR and costimulatory signaling in selected T cells.” Proceedings of the National Academy of Sciences of the United States of America 102: 29042909.CrossRefGoogle ScholarPubMed
Ricart, B. G., Yang, M. T., Hunter, C. A., Chen, C. S. and Hammer, D. A. (2013). “Measuring traction forces of motile dendritic cells on micropost arrays.” Biophysical Journal 101: 26202628.CrossRefGoogle Scholar
Sackmann, E. (1996). “Supported membranes: scientific and practical applications.” Science 271: 4348.CrossRefGoogle ScholarPubMed
Scholer, A., Hugues, S., Boissonnas, A., Fetler, L. and Amigorena, S. (2008). “Intercellular adhesion molecule-1-dependent stable interactions between T cells and dendritic cells determine CD8+ T cell memory.” Immunity 28: 258270.CrossRefGoogle ScholarPubMed
Sharpe, A. H. and Freeman, G. J. (2002). “The B7–CD28 superfamily.” Nature Reviews Immunology 2: 116126.CrossRefGoogle ScholarPubMed
Shen, K., Thomas, V. K., Dustin, M. L. and Kam, L. C. (2008). “Micropatterning of costimulatory ligands enhances CD4+ T cell function.” Proceedings of the National Academy of Sciences 105: 77917796.CrossRefGoogle ScholarPubMed
Smith-Garvin, J. E., Koretzky, G. A. and Jordan, M. S. (2009). “T cell activation.” Annual Review of Immunology 27: 591.CrossRefGoogle ScholarPubMed
Spatz, J. P., M Ssmer, S., Hartmann, C., M Ller, M., Herzog, T., Krieger, M., Boyen, H.-G., et al. (2000). “Ordered deposition of inorganic clusters from micellar block copolymer films.” Langmuir 16: 407415.CrossRefGoogle Scholar
Thauland, T. J., Koguchi, Y., Wetzel, S. A., Dustin, M. L. and Parker, D. C. (2008). “Th1 and Th2 cells form morphologically distinct immunological synapses.” The Journal of Immunology 181: 393399.CrossRefGoogle ScholarPubMed
Tseng, S. Y., Liu, M. and Dustin, M. L. (2005). “CD80 cytoplasmic domain controls localization of CD28, CTLA-4, and protein kinase C theta in the immunological synapse.” Journal of Immunology 175: 78297836.CrossRefGoogle ScholarPubMed
Varma, R., Campi, G., Yokosuka, T., Saito, T. and Dustin, M. L. (2006). “T cell receptor-proximal signals are sustained in peripheral microclusters and terminated in the central supramolecular activation cluster.” Immunity 25: 117127.CrossRefGoogle ScholarPubMed
Williams, M. A. and Bevan, M. J. (2007). “Effector and memory CTL differentiation.” Annual Review of Immunology 25: 171192.CrossRefGoogle ScholarPubMed
Yokosuka, T., Kobayashi, W., Sakata-Sogawa, K., Takamatsu, M., Hashimoto-Tane, A., Dustin, M. L., Tokunaga, M. and Saito, T. (2008). “Spatiotemporal regulation of T cell costimulation by TCR-CD28 microclusters and protein kinase C theta translocation.” Immunity 29: 589601.CrossRefGoogle ScholarPubMed
Yokosuka, T., Sakata-Sogawa, K., Kobayashi, W., Hiroshima, M., Hashimoto-Tane, A., Tokunaga, M., Dustin, M. L. and Saito, T. (2005). “Newly generated T cell receptor microclusters initiate and sustain T cell activation by recruitment of Zap70 and SLP-76.” Nature Immunology 6: 12531262.CrossRefGoogle ScholarPubMed
Yu, C.-H., Wu, H.-J., Kaizuka, Y., Vale, R. D. and Groves, J. T. (2010). “Altered actin centripetal retrograde flow in physically restricted immunological synapses.” PloS One 5: e11878.CrossRefGoogle ScholarPubMed
Zhu, J., Yamane, H. and Paul, W. E. (2010). “Differentiation of effector CD4+ T cell populations.” Annual Review of Immunology 28: 445489.CrossRefGoogle ScholarPubMed

References

Barkefors, I., Thorslund, S., Nikolajeff, F. and Kreuger, J. (2009). “A fluidic device to study directional angiogenesis in complex tissue and organ culture models.” Lab on a Chip 9: 529535.CrossRefGoogle ScholarPubMed
Bayless, K. J., Salazar, R. and Davis, G. E. (2000). “RGD-dependent vacuolation and lumen formation observed during endothelial cell morphogenesis in three-dimensional fibrin matrices involves the α v β3 and α 5 β 1 integrins.” The American Journal of Pathology 156: 16731683.CrossRefGoogle Scholar
Bersini, S., Jeon, J. S., Dubini, G., Arrigoni, C., Chung, S., Charest, J. L., Moretti, M. and Kamm, R. D. (2014). “A microfluidic 3-D in vitro model for specificity of breast cancer metastasis to bone.” Biomaterials 35: 24542461.CrossRefGoogle Scholar
Bettinger, C., Borenstein, J. T. and Tao, S. L. (2012). Microfluidic Cell Culture Systems. Oxford: William Andrew.Google Scholar
Bischel, L. L., Young, E. W., Mader, B. R. and Beebe, D. J. (2013). “Tubeless microfluidic angiogenesis assay with three-dimensional endothelial-lined microvessels. Biomaterials 34: 14711477.CrossRefGoogle ScholarPubMed
Carmeliet, P. (2000). “Mechanisms of angiogenesis and arteriogenesis.” Nature Medicine 6: 389396.CrossRefGoogle ScholarPubMed
Carmeliet, P. and Jain, R. K. (2000). “Angiogenesis in cancer and other diseases.” Nature 407: 249257.CrossRefGoogle ScholarPubMed
Carrion, B., Huang, C. P., Ghajar, C. M., Kachgal, S., Kniazeva, E., Jeon, N. L. and Putnam, A. J. (2010). “Recreating the perivascular niche ex vivo using a microfluidic approach.” Biotechnology and Bioengineering 107: 10201028.CrossRefGoogle ScholarPubMed
Chalupowicz, D. G., Chowdhury, Z. A., Bach, T. L., Barsigian, C. and Martinez, J. (1995). “Fibrin II induces endothelial cell capillary tube formation.” The Journal of cell Biology 130: 207215.CrossRefGoogle ScholarPubMed
Chaw, K., Manimaran, M., Tay, E. and Swaminathan, S. (2007). “Multi-step microfluidic device for studying cancer metastasis.” Lab on a Chip 7: 10411047.CrossRefGoogle ScholarPubMed
Chen, M. B., Whisler, J. A., Jeon, J. S. and Kamm, R. D. (2013). “Mechanisms of tumor cell extravasation in an in vitro microvascular network platform.” Integrative Biology 5: 12621271.CrossRefGoogle Scholar
Chrobak, K. M., Potter, D. R. and Tien, J. (2006). “Formation of perfused, functional microvascular tubes in vitro.” Microvascular Research 71: 185196.CrossRefGoogle ScholarPubMed
Chung, S., Sudo, R., Mack, P. J., Wan, C.-R., Vickerman, V. and Kamm, R. D. (2009a). “Cell migration into scaffolds under co-culture conditions in a microfluidic platform.” Lab on a Chip 9: 269275.CrossRefGoogle Scholar
Chung, S., Sudo, R., Vickerman, V., Zervantonakis, I. K. and Kamm, R. D. (2010). “Microfluidic platforms for studies of angiogenesis, cell migration, and cell-cell interactions.” Annals of Biomedical Engineering 38: 11641177.CrossRefGoogle ScholarPubMed
Chung, S., Sudo, R., Zervantonakis, I. K., Rimchala, T. and Kamm, R. D. (2009b). “Surface‐treatment‐induced three‐dimensional capillary morphogenesis in a microfluidic platform.” Advanced Materials 21: 48634867.CrossRefGoogle Scholar
Dai, X., Cai, S., Ye, Q., Jiang, J., Yan, X., Xiong, X., Jiang, Q., et al. (2011). “A novel in vitro angiogenesis model based on a microfluidic device.” Chinese Science Bulletin 56: 33013309.CrossRefGoogle Scholar
Eilken, H. M. and Adams, R. H. (2010). “Dynamics of endothelial cell behavior in sprouting angiogenesis.” Current Opinion in Cell Biology 22: 617625.CrossRefGoogle ScholarPubMed
Farahat, W. A., Wood, L. B., Zervantonakis, I. K., Schor, A., Ong, S., Neal, D., Kamm, R. D. et al. (2012). “Ensemble analysis of angiogenic growth in three-dimensional microfluidic cell cultures.” PloS One 7: e37333.CrossRefGoogle ScholarPubMed
Ferrari, G., Pintucci, G., Seghezzi, G., Hyman, K., Galloway, A. C. and Mignatti, P. (2006). “VEGF, a prosurvival factor, acts in concert with TGF-β1 to induce endothelial cell apoptosis.” Proceedings of the National Academy of Sciences 103: 1726017265.CrossRefGoogle ScholarPubMed
Figg, W. and Folkman, J. (2008). Angiogenesis: An Integrative Approach from Science to Medicine. New York: Springer.CrossRefGoogle Scholar
Folkman, J. (1995). “Angiogenesis in cancer, vascular, rheumatoid and other disease.” Nature Medicine 1: 2730.CrossRefGoogle ScholarPubMed
George, E. L., Baldwin, H. S. and Hynes, R. O. (1997). “Fibronectins are essential for heart and blood vessel morphogenesis but are dispensable for initial specification of precursor cells.” Blood 90: 30733081.CrossRefGoogle ScholarPubMed
Ghajar, C. M., Blevins, K. S., Hughes, C. C., George, S. C. and Putnam, A. J. (2006). “Mesenchymal stem cells enhance angiogenesis in mechanically viable prevascularized tissues via early matrix metalloproteinase upregulation.” Tissue Engineering 12: 28752888.CrossRefGoogle ScholarPubMed
Grant, D. S., Yenisey, C., Rose, R. W., Tootell, M., Santra, M. and Iozzo, R. V. (2002). “Decorin suppresses tumor cell-mediated angiogenesis.” Oncogene 21: 47654777.CrossRefGoogle ScholarPubMed
Helm, C.-L. E., Fleury, M. E., Zisch, A. H., Boschetti, F. and Swartz, M. A. (2005). “Synergy between interstitial flow and VEGF directs capillary morphogenesis in vitro through a gradient amplification mechanism.” Proceedings of the National Academy of Sciences of the United States of America 102: 1577915784.CrossRefGoogle ScholarPubMed
Huang, C. P., Lu, J., Seon, H., Lee, A. P., Flanagan, L. A., Kim, H.-Y., Putnam, A. J. et al. (2009). “Engineering microscale cellular niches for three-dimensional multicellular co-cultures.” Lab on a Chip 9: 17401748.CrossRefGoogle ScholarPubMed
Jain, R. K., di Tomaso, E., Duda, D. G., Loeffler, J. S., Sorensen, A. G. and Batchelor, T. T. (2007). “Angiogenesis in brain tumours.” Nature Reviews Neuroscience 8: 610622.CrossRefGoogle ScholarPubMed
Järveläinen, H., Sainio, A., Koulu, M., Wight, T. N. and Penttinen, R. (2009). “Extracellular matrix molecules: potential targets in pharmacotherapy.” Pharmacological Reviews 61: 198223.CrossRefGoogle ScholarPubMed
Jeon, J. S., Bersini, S., Gilardi, M., Dubini, G., Charest, J. L., Moretti, M. and Kamm, R. D. (2014a). “Human 3-D vascularized organotypic microfluidic assays to study breast cancer cell extravasation.” Proceedings of the National Academy of Sciences: 201417115.CrossRefGoogle Scholar
Jeon, J. S., Bersini, S., Whisler, J. A., Chen, M. B., Dubini, G., Charest, J. L., Moretti, M. et al. (2014b). “Generation of 3-D functional microvascular networks with human mesenchymal stem cells in microfluidic systems.” Integrative Biology 6: 555563.CrossRefGoogle Scholar
Jeon, J. S., Zervantonakis, I. K., Chung, S., Kamm, R. D. and Charest, J. L. (2013). “In vitro model of tumor cell extravasation.” PloS One 8: e56910.CrossRefGoogle ScholarPubMed
Jeong, G. S., Han, S., Shin, Y., Kwon, G. H., Kamm, R. D., Lee, S.-H. and Chung, S. (2011a). “Sprouting angiogenesis under a chemical gradient regulated by interactions with an endothelial monolayer in a microfluidic platform.” Analytical Chemistry 83: 84548459.CrossRefGoogle Scholar
Jeong, G. S., Kwon, G. H., Kang, A. R., Jung, B. Y., Park, Y., Chung, S. and Lee, S.-H. (2011b). “Microfluidic assay of endothelial cell migration in 3D interpenetrating polymer semi-network HA-Collagen hydrogel.” Biomedical Microdevices 13: 717723.CrossRefGoogle ScholarPubMed
Kim, C., Chung, S., Yuchun, L., Kim, M.-C., Chan, J. K., Asada, H. H. and Kamm, R. D. (2012). “In vitro angiogenesis assay for the study of cell-encapsulation therapy.” Lab on a Chip 12: 29422950.CrossRefGoogle Scholar
Kim, C., Kasuya, J., Jeon, J., Chung, S. and Kamm, R. D. (2015). “A quantitative microfluidic angiogenesis screen for studying anti-angiogenic therapeutic drugs.” Lab on a Chip 15: 301310.CrossRefGoogle ScholarPubMed
Kim, S., Lee, H., Chung, M. and Jeon, N. L. (2013). “Engineering of functional. perfusable 3-D microvascular networks on a chip.” Lab on a Chip 13: 14891500.CrossRefGoogle Scholar
Mack, P. J., Zhang, Y., Chung, S., Vickerman, V., Kamm, R. D. and García-Cardeña, G. (2009). “Biomechanical regulation of endothelium-dependent events critical for adaptive remodeling.” Journal of Biological Chemistry 284: 84128420.CrossRefGoogle ScholarPubMed
Maeshima, Y., Colorado, P. C. and Kalluri, R. (2000). “Two RGD-independent αvβ3 integrin binding sites on tumstatin regulate distinct anti-tumor properties.” Journal of Biological Chemistry 275: 2374523750.CrossRefGoogle ScholarPubMed
Metheny-Barlow, L. J. and Li, L. Y. (2003). “The enigmatic role of angiopoietin-1 in tumor angiogenesis.” Cell Research 13: 309317.CrossRefGoogle ScholarPubMed
Miller, J. S., Stevens, K. R., Yang, M. T., Baker, B. M., Nguyen, D.-H. T., Cohen, D. M., Toro, E., et al. (2012). “Rapid casting of patterned vascular networks for perfusable engineered three-dimensional tissues.” Nature Materials 11: 768774.CrossRefGoogle ScholarPubMed
Moya, M. L., Hsu, Y.-H., Lee, A. P., Hughes, C. C. and George, S. C. (2013). “In vitro perfused human capillary networks.” Tissue Engineering Part C: Methods 19: 730737.CrossRefGoogle ScholarPubMed
mustonen, T. and Alitalo, K. (1995). “Endothelial receptor tyrosine kinases involved in angiogenesis.” The Journal of Cell Biology 129: 895898.CrossRefGoogle ScholarPubMed
Nakayasu, K., Hayashi, N., Okisaka, S. and Sato, N. (1992). “Formation of capillary-like tubes by vascular endothelial cells cocultivated with keratocytes.” Investigative Ophthalmology & Visual Science 33: 30503057.Google ScholarPubMed
Ng, C. P., Helm, C.-L. E. and Swartz, M. A. (2004). “Interstitial flow differentially stimulates blood and lymphatic endothelial cell morphogenesis in vitro.” Microvascular Research 68: 258264.CrossRefGoogle ScholarPubMed
Nguyen, D.-H. T., Stapleton, S. C., Yang, M. T., Cha, S. S., Choi, C. K., Galie, P. A. and Chen, C. S. (2013). “Biomimetic model to reconstitute angiogenic sprouting morphogenesis in vitro.” Proceedings of the National Academy of Sciences 110: 67126717.CrossRefGoogle ScholarPubMed
Nissen, L. J., Cao, R., Hedlund, E.-M., Wang, Z., Zhao, X., Wetterskog, D., Funa, K., et al. (2007). “Angiogenic factors FGF2 and PDGF-BB synergistically promote murine tumor neovascularization and metastasis.” The Journal of Clinical Investigation 117: 27662777.CrossRefGoogle ScholarPubMed
Rodríguez-Manzaneque, J. C., Lane, T. F., Ortega, M. A., Hynes, R. O., Lawler, J. and Iruela-Arispe, M. L. (2001). “Thrombospondin-1 suppresses spontaneous tumor growth and inhibits activation of matrix metalloproteinase-9 and mobilization of vascular endothelial growth factor.” Proceedings of the National Academy of Sciences 98: 1248512490.CrossRefGoogle ScholarPubMed
Rørth, P. (2009). “Collective cell migration.” Annual Review of Cell and Developmental 25: 407429.CrossRefGoogle ScholarPubMed
Semenza, G. L. (1998). “Hypoxia-inducible factor 1: master regulator of O2 homeostasis.” Current Opinion in Genetics & Development 8: 588594.CrossRefGoogle ScholarPubMed
Shamloo, A., Xu, H. and Heilshorn, S. (2011). “Mechanisms of vascular endothelial growth factor-induced pathfinding by endothelial sprouts in biomaterials.” Tissue Engineering Part A 18: 320330.CrossRefGoogle ScholarPubMed
Shin, Y., Han, S., Jeon, J. S., Yamamoto, K., Zervantonakis, I. K., Sudo, R., Kamm, R. D. et al. (2012). “Microfluidic assay for simultaneous culture of multiple cell types on surfaces or within hydrogels.” Nature Protocols 7: 12471259.CrossRefGoogle ScholarPubMed
Shin, Y., Jeon, J. S., Han, S., Jung, G.-S., Shin, S., Lee, S.-H., Sudo, R., et al. (2011). “In vitro 3D collective sprouting angiogenesis under orchestrated ANG-1 and VEGF gradients.” Lab on a Chip 11: 21752181.CrossRefGoogle ScholarPubMed
Shin, Y., Kim, H., Han, S., Won, J., Jeong, H. E., Lee, E. S., Kamm, R. D., et al. (2013). “Extracellular matrix heterogeneity regulates three‐dimensional morphologies of breast adenocarcinoma cell invasion.” Advanced Healthcare Materials 2: 790794.CrossRefGoogle ScholarPubMed
Shin, Y., Yang, K., Han, S., Park, H. J., Seok Heo, Y., Cho, S. W. and Chung, S. (2014). “Reconstituting vascular microenvironment of neural stem cell niche in three‐dimensional extracellular matrix.” Advanced Healthcare Materials 3: 14571464.CrossRefGoogle ScholarPubMed
Shiu, Y.-T., Weiss, J. A., Hoying, J. B., Iwamoto, M. N., Joung, I. S. and Quam, C. T. (2004). “The role of mechanical stresses in angiogenesis.” Critical Reviews in Biomedical Engineering 33: 431510.Google Scholar
Song, J. W. and Munn, L. L. (2011). “Fluid forces control endothelial sprouting.” Proceedings of the National Academy of Sciences 108: 1534215347.CrossRefGoogle ScholarPubMed
Steeg, P. S. (2006). “Tumor metastasis: mechanistic insights and clinical challenges.” Nature Medicine 12: 895904.CrossRefGoogle ScholarPubMed
Stroock, A. D. and Fischbach, C. (2010). “Microfluidic culture models of tumor angiogenesis.” Tissue Engineering Part A 16: 21432146.CrossRefGoogle ScholarPubMed
Sudo, R., Chung, S., Zervantonakis, I. K., Vickerman, V., Toshimitsu, Y., Griffith, L. G. and Kamm, R. D. (2009). “Transport-mediated angiogenesis in 3-D epithelial coculture.” The FASEB Journal 23: 21552164.CrossRefGoogle Scholar
Swartz, M. A. and Lund, A. W. (2012). “Lymphatic and interstitial flow in the tumour microenvironment: linking mechanobiology with immunity.” Nature Reviews Cancer 12: 210219.CrossRefGoogle ScholarPubMed
Thyboll, J., Kortesmaa, J., Cao, R., Soininen, R., Wang, L., Iivanainen, A., Sorokin, L., et al. (2002). “Deletion of the laminin α4 chain leads to impaired microvessel maturation.” Molecular and Cellular Biology 22: 11941202.CrossRefGoogle ScholarPubMed
Verbridge, S. S., Chakrabarti, A., Delnero, P., Kwee, B., Varner, J. D., Stroock, A. D. and Fischbach, C. (2013). “Physicochemical regulation of endothelial sprouting in a 3-D microfluidic angiogenesis model.” Journal of Biomedical Materials Research Part A 101: 29482956.CrossRefGoogle Scholar
Vickerman, V., Blundo, J., Chung, S. and Kamm, R. (2008). “Design, fabrication and implementation of a novel multi-parameter control microfluidic platform for three-dimensional cell culture and real-time imaging.” Lab on a Chip 8: 14681477.CrossRefGoogle ScholarPubMed
Vickerman, V. and Kamm, R. D. (2012). “Mechanism of a flow-gated angiogenesis switch: early signaling events at cell-matrix and cell-cell junctions.” Integrative Biology 4: 863874.CrossRefGoogle ScholarPubMed
Weis, S. M. and Cheresh, D. A. (2011). “Tumor angiogenesis: molecular pathways and therapeutic targets.” Nature Medicine 17: 13591370.CrossRefGoogle ScholarPubMed
Whisler, J. A., Chen, M. B. and Kamm, R. D. (2012). “Control of perfusable microvascular network morphology using a multiculture microfluidic system.” Tissue Engineering Part C: Methods 7: 543552.Google Scholar
Wong, A. P., Perez-Castillejos, R., Christopher Love, J. and Whitesides, G. M. (2008). “Partitioning microfluidic channels with hydrogel to construct tunable 3-D cellular microenvironments.” Biomaterials 29: 18531861.CrossRefGoogle ScholarPubMed
Woyach, J. A. and Shah, M. H. (2009). “New therapeutic advances in the management of progressive thyroid cancer.” Endocrine-Related Cancer 16: 715731.CrossRefGoogle ScholarPubMed
Xu, X., Yang, G., Zhang, H. and Prestwich, G. D. (2009). “Evaluating dual activity LPA receptor pan-antagonist/autotaxin inhibitors as anti-cancer agents in vivo using engineered human tumors.” Prostaglandins & Other Lipid Mediators 89: 140146.CrossRefGoogle ScholarPubMed
Yang, K., Han, S., Shin, Y., Ko, E., Kim, J., Park, K. I., Chung, S. et al. (2013). “A microfluidic array for quantitative analysis of human neural stem cell self-renewal and differentiation in three-dimensional hypoxic microenvironment.” Biomaterials 34: 66076614.CrossRefGoogle ScholarPubMed
Yeon, J. H., Ryu, H. R., Chung, M., Hu, Q. P. and Jeon, N. L. (2012). “In vitro formation and characterization of a perfusable three-dimensional tubular capillary network in microfluidic devices.” Lab on a Chip 12: 28152822.CrossRefGoogle ScholarPubMed
Yoshida, A., Anand-Apte, B. and Zetter, B. R. (1996). “Differential endothelial migration and proliferation to basic fibroblast growth factor and vascular endothelial growth factor.” Growth Factors 13: 5764.CrossRefGoogle ScholarPubMed
Young, E. W., Wheeler, A. R. and Simmons, C. A. (2007). “Matrix-dependent adhesion of vascular and valvular endothelial cells in microfluidic channels.” Lab on a Chip 7: 17591766.CrossRefGoogle ScholarPubMed
Zervantonakis, I. K., Hughes-Alford, S. K., Charest, J. L., Condeelis, J. S., Gertler, F. B. and Kamm, R. D. (2012). “Three-dimensional microfluidic model for tumor cell intravasation and endothelial barrier function.” Proceedings of the National Academy of Sciences 109: 1351513520.CrossRefGoogle ScholarPubMed
Zetter, P., and Bruce, R. (1998). “Angiogenesis and tumor metastasis.” Annual Review of Medicine 49: 407424.CrossRefGoogle ScholarPubMed
Zheng, Y., Chen, J., Craven, M., Choi, N. W., Totorica, S., Diaz-Santana, A., Kermani, P., et al. (2012). “In vitro microvessels for the study of angiogenesis and thrombosis.” Proceedings of the National Academy of Sciences 109: 93429347.CrossRefGoogle Scholar

References

Abbott, B. C., Hill, A. V., and Howarth, J. V.. (1958). “The positive and negative heat production associated with a nerve impulse.” Proc R Soc Lond B Biol Sci 148(931): 149187.Google ScholarPubMed
Ahmed, W. W., et al. (2012). “Mechanical tension modulates local and global vesicle dynamics in neurons.” Cell Mol Bioeng 5(2): 155164.CrossRefGoogle ScholarPubMed
Almeida, P. F. F. and Vaz, W. L. C.. (1995). “Lateral diffusion in membranes.” In Handbook of Biological Physics: Structure and Dynamics of Membranes – From Cells to Vesicles, Lipowsky, R. and Sackmann, E., eds. Amsterdam: North-Holland, 305357.CrossRefGoogle Scholar
Anava, S., et al. (2009). “The regulative role of neurite mechanical tension in network development.” Biophys J 96(4): 16611670.CrossRefGoogle ScholarPubMed
Arnadottir, J. and Chalfie, M.. (2010). “Eukaryotic mechanosensitive channels.” Annu Rev Biophys 39: 111137.CrossRefGoogle ScholarPubMed
Balice-Gordon, R.J. and Lichtman, J.W.. (1990). “In vivo visualization of the growth of pre– and postsynaptic elements of neuromuscular junctions in the mouse.” J Neurosci 10(3): 894908.CrossRefGoogle ScholarPubMed
Betz, T., et al. (2011). “Growth cones as soft and weak force generators.” Proc Natl Acad Sci USA 108(33): 1342013425.CrossRefGoogle ScholarPubMed
Bray, D. (1984). “Axonal growth in response to experimentally applied mechanical tension.” Dev Biol 102(2): 379–89.CrossRefGoogle ScholarPubMed
Bridgman, P. C., et al. (2001). “Myosin IIB is required for growth cone motility.” Journal of Neuroscience 21(16): 61596169.CrossRefGoogle ScholarPubMed
Cantor, R. S. (1997). “The lateral pressure profile in membranes: a physical mechanism of general anesthesia.” Biochemistry 36(9): 23392344.CrossRefGoogle ScholarPubMed
Chan, C. E. and Odde, D. J.. (2008). “Traction dynamics of filopodia on compliant substrates.” Science 322(5908): 16871691.CrossRefGoogle ScholarPubMed
Chen, B. M. and Grinnell, A. D.. (1995). “Integrins and modulation of transmitter release from motor nerve terminals by stretch.” Science 269(5230): 15781580.CrossRefGoogle ScholarPubMed
Cole, K. S. (1941). “Rectification and inductance in the squid giant axon.” J Gen Physiol 25(1): 2951.CrossRefGoogle ScholarPubMed
Crawford, G. E. and Earnshaw, J. C.. (1987). “Viscoelastic relaxation of bilayer lipid membranes. Frequency-dependent tension and membrane viscosity.” Biophys J 52(1): 8794.CrossRefGoogle ScholarPubMed
Elkin, B. S., et al. (2007). “Mechanical heterogeneity of the rat hippocampus measured by atomic force microscope indentation.” J Neurotrauma 24(5): 812–22.CrossRefGoogle ScholarPubMed
Evans, E.A. and Hochmuth, R.M., Current Topics in Membranes and Transport. 1978. 164.CrossRefGoogle Scholar
Eyckmans, J., et al. (2011). “A hitchhiker’s guide to mechanobiology.” Dev Cell 21(1): 3547.CrossRefGoogle ScholarPubMed
Farkas, O., Lifshitz, J., and Povlishock, J.T.. (2006). “Mechanoporation induced by diffuse traumatic brain injury: an irreversible or reversible response to injury?J Neurosci 26(12): 31303140.CrossRefGoogle ScholarPubMed
Farkas, O. and Povlishock, J. T.. (2007). “Cellular and subcellular change evoked by diffuse traumatic brain injury: a complex web of change extending far beyond focal damage.” Prog Brain Res 161: 4359.CrossRefGoogle ScholarPubMed
Fass, J. N. and Odde, D. J.. (2003). “Tensile force-dependent neurite elicitation via anti-beta 1 integrin antibody-coated magnetic beads.” Biophysical Journal 85(1): 623636.CrossRefGoogle Scholar
Fatt, P. and Katz, B.. (1952). “Spontaneous subthreshold activity at motor nerve endings.” J Physiol 117: 109128.CrossRefGoogle ScholarPubMed
Feynman, R. P., Leighton, R. B., and Sands, M., The Feynman Lectures on Physics. 1963. 4649.CrossRefGoogle Scholar
Fischer, M., et al. (1998). “Rapid actin-based plasticity in dendritic spines.” Neuron 20(5): 847854.CrossRefGoogle ScholarPubMed
Flanagan, L. A., et al. (2002). “Neurite branching on deformable substrates.” Neuroreport 13(18): 24112415.CrossRefGoogle ScholarPubMed
Footer, M. J., et al. (2007). “Direct measurement of force generation by actin filament polymerization using an optical trap.” Proceedings of the National Academy of Sciences of the United States of America 104(7): 21812186.CrossRefGoogle ScholarPubMed
Furukawa, K., et al. (1997). “The actin-severing protein gelsolin modulates calcium channel and NMDA receptor activities and vulnerability to excitotoxicity in hippocampal neurons.” J Neurosci 17(21): 81788186.CrossRefGoogle ScholarPubMed
Gefen, A., et al. (2003). “Age-dependent changes in material properties of the brain and braincase of the rat.” J Neurotrauma 20(11): 11631177.CrossRefGoogle ScholarPubMed
Griesbauer, J., Wixforth, A., and Schneider, M. F.. (2009). “Wave propagation in lipid monolayers.” Biophys J 97(10): 27102716.CrossRefGoogle ScholarPubMed
Halpain, S., Hipolito, A., and Saffer, L.. (1998). “Regulation of F-actin stability in dendritic spines by glutamate receptors and calcineurin.” J Neurosci 18(23): 98359844.CrossRefGoogle ScholarPubMed
Hamill, O. P. and Martinac, B.. (2001). “Molecular basis of mechanotransduction in living cells.” Physiol Rev 81(2): 685740.CrossRefGoogle ScholarPubMed
Heimburg, T. (2010). “Lipid ion channels.” Biophys Chem 150(1–3): 222.CrossRefGoogle ScholarPubMed
Heimburg, T. and Jackson, A. D.. (2005). “On soliton propagation in biomembranes and nerves.” Proc Natl Acad Sci USA 102(28): 97909795.CrossRefGoogle ScholarPubMed
Heimburg, T. and Jackson, A. D.. (2007). “On the action potential as a propagating density pulse and the role of anesthetics.” Biophysical Reviews and Letters 02(01): 5778.CrossRefGoogle Scholar
Hemphill, M. A., et al. (2011). “A possible role for integrin signaling in diffuse axonal injury.” PLoS One 6(7): e22899.CrossRefGoogle ScholarPubMed
Hill, T. L. and Kirschner, M. W.. (1982). “Subunit treadmilling of microtubules or actin in the presence of cellular barriers – possible conversion of chemical free energy into mechanical work.” Proc Natl Acad Sci USA - Biological Sciences 79(2): 490494.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Huxley, A. F.. (1952). “A quantitative description of membrane current and its application to conduction and excitation in nerve.” J Physiol 117(4): 500544.CrossRefGoogle ScholarPubMed
Hoge, C. W., et al. (2008). “Mild traumatic brain injury in U.S. Soldiers returning from Iraq.” N Engl J Med 358(5): 453463.CrossRefGoogle ScholarPubMed
Howarth, J. V., et al. (1975). “Heat production associated with passage of a single impulse in pike olfactory nerve-fibers.” Journal of Physiology-London 249(2): 349368.CrossRefGoogle Scholar
Ingber, D. E. (2003). “Tensegrity I. Cell structure and hierarchical systems biology.” J Cell Sci 116(Pt 7): 11571173.CrossRefGoogle ScholarPubMed
Iwasa, K. and Tasaki, I.. (1980). “Mechanical changes in squid giant axons associated with production of action potentials.” Biochem Biophys Res Commun 95(3): 13281331.CrossRefGoogle ScholarPubMed
Jeon, J. and Voth, G. A.. (2005). “The dynamic stress responses to area change in planar lipid bilayer membranes.” Biophys J 88(2): 11041119.CrossRefGoogle ScholarPubMed
Jerabek, H., et al. (2010). “Membrane-mediated effect on ion channels induced by the anesthetic drug ketamine.” J Am Chem Soc 132(23): 79907997.CrossRefGoogle ScholarPubMed
Kilinc, D., Gallo, G., and Barbee, K. A.. (2008). “Mechanically-induced membrane poration causes axonal beading and localized cytoskeletal damage.” Exp Neurol 212(2): 422430.CrossRefGoogle ScholarPubMed
Kim, C. H. and Lisman, J. E.. (1999). “A role of actin filament in synaptic transmission and long-term potentiation.” J Neurosci 19(11): 43144324.CrossRefGoogle ScholarPubMed
Kim, G. H., et al. (2007). “A mechanical spike accompanies the action potential in Mammalian nerve terminals.” Biophys J 92(9): 31223129.CrossRefGoogle ScholarPubMed
Koch, D., et al. (2012). “Strength in the periphery: growth cone biomechanics and substrate rigidity response in peripheral and central nervous system neurons.” Biophys J 102(3): 452460.CrossRefGoogle ScholarPubMed
Kostic, A., Sap, J., and Sheetz, M. P.. (2007). “RPTPalpha is required for rigidity-dependent inhibition of extension and differentiation of hippocampal neurons.” J Cell Sci 120(Pt 21): 38953904.CrossRefGoogle ScholarPubMed
Kosztin, I., et al. (2002). “Mechanical force generation by G proteins.” Proc Natl Acad Sci USA 99(6): 35753580.CrossRefGoogle ScholarPubMed
Krucker, T., Siggins, G. R., and Halpain, S.. (2000). “Dynamic actin filaments are required for stable long-term potentiation (LTP) in area CA1 of the hippocampus.” Proc Natl Acad Sci USA 97(12): 68566861.CrossRefGoogle ScholarPubMed
Kruse, S. A., et al. (2008). “Magnetic resonance elastography of the brain.” Neuroimage 39(1): 231237.CrossRefGoogle ScholarPubMed
Lamoureux, P., et al. (2002). “Mechanical tension can specify axonal fate in hippocampal neurons.” J Cell Biol 159(3): 499508.CrossRefGoogle ScholarPubMed
Lo, E. H., Wang, X., and Cuzner, M. L.. (2002). “Extracellular proteolysis in brain injury and inflammation: role for plasminogen activators and matrix metalloproteinases.” J Neurosci Res 69(1): 19.CrossRefGoogle ScholarPubMed
Lundstrom, I. (1974). “Mechanical wave propagation on nerve axons.” J Theor Biol 45(2): 487499.CrossRefGoogle ScholarPubMed
Matus, A. (2000). “Actin-based plasticity in dendritic spines.” Science 290(5492): 754758.CrossRefGoogle ScholarPubMed
Matus, A. (2000). “Actin-based plasticity in dendritic spines.” Science 290: 754758.CrossRefGoogle ScholarPubMed
McCracken, P. J., et al. (2005). “Mechanical transient-based magnetic resonance elastography.” Magn Reson Med 53(3): 628639.CrossRefGoogle ScholarPubMed
Meaney, D. F. and Smith, D. H.. (2011). “Biomechanics of concussion.” Clin Sports Med 30(1): 1931, vii.CrossRefGoogle ScholarPubMed
Moore, S. W., Roca-Cusachs, P., and Sheetz, M. P.. (2010). “Stretchy proteins on stretchy substrates: the important elements of integrin-mediated rigidity sensing.” Dev Cell 19(2): 194206.CrossRefGoogle ScholarPubMed
Morris, C. E. (2011). “Voltage-gated channel mechanosensitivity: fact or friction?Front Physiol 2: 25.CrossRefGoogle ScholarPubMed
Murphy, M. C., et al. (2011). “Decreased brain stiffness in Alzheimer’s disease determined by magnetic resonance elastography.” J Magn Reson Imaging 34(3): 494498.CrossRefGoogle ScholarPubMed
Murthy, V. N., et al. (2001). “Inactivity produces increases in neurotransmitter release and synapse size.” Neuron 32(4): 673682.CrossRefGoogle ScholarPubMed
Nguyen, T. D., et al. (2012). “Piezoelectric nanoribbons for monitoring cellular deformations.” Nat Nanotechnol 7(9): 587593.CrossRefGoogle ScholarPubMed
Nordahl, C. W., et al. (2007). “Cortical folding abnormalities in autism revealed by surface-based morphometry.” J Neurosci 27(43): 1172511735.CrossRefGoogle ScholarPubMed
Oh, S., et al. (2012). “Label-free imaging of membrane potential using membrane electromotility.” Biophys J 103(1): 1118.CrossRefGoogle ScholarPubMed
Parekh, S. H., et al. (2005). “Loading history determines the velocity of actin-network growth.” Nat Cell Biol 7(12): 12191223.CrossRefGoogle ScholarPubMed
Pastor, R. W. and Feller, S. E.. (1996). “Time scales of lipid dynamics and molecular dynamics.” Biol Membr 1: 429.Google Scholar
Pekny, M. and Lane, E. B.. (2007). “Intermediate filaments and stress.” Exp Cell Res 313(10): 22442254.CrossRefGoogle ScholarPubMed
Peskin, C. S., Odell, G. M., and Oster, G. F.. (1993). “Cellular motions and thermal fluctuations: the Brownian ratchet.” Biophys J 65(1): 316–24.CrossRefGoogle ScholarPubMed
Petrov, A. G. (2002). “Flexoelectricity of model and living membranes.” Biochim Biophys Acta 1561(1): 125.CrossRefGoogle ScholarPubMed
Petrov, A. G. (2006). “Electricity and mechanics of biomembrane systems: flexoelectricity in living membranes.” Anal Chim Acta 568(1–2): 7083.CrossRefGoogle ScholarPubMed
Pfister, B. J., et al. (2004). “Extreme stretch growth of integrated axons.” J Neurosci 24(36): 79787983.CrossRefGoogle ScholarPubMed
Prange, M. T. and Margulies, S. S.. (2002). “Regional, directional, and age-dependent properties of the brain undergoing large deformation.” J Biomech Eng 124(2): 244252.CrossRefGoogle ScholarPubMed
Reeves, D., et al. (2008). “Membrane mechanics as a probe of ion-channel gating mechanisms.” Physical Review E 78(4): 041901.CrossRefGoogle ScholarPubMed
Rodriguez, O. C., et al. (2003). “Conserved microtubule-actin interactions in cell movement and morphogenesis.” Nat Cell Biol 5(7): 599609.CrossRefGoogle ScholarPubMed
Ronan, L., et al. (2013). “Differential tangential expansion as a mechanism for cortical gyrification.” Cereb Cortex 24(8): 22192228.CrossRefGoogle ScholarPubMed
Roth, S., et al. (2012). “How morphological constraints affect axonal polarity in mouse neurons.” PLoS One 7(3): e33623.CrossRefGoogle ScholarPubMed
Sack, I., et al. (2011). “The influence of physiological aging and atrophy on brain viscoelastic properties in humans.” PLoS One 6(9): e23451.CrossRefGoogle ScholarPubMed
Schikorski, T. and Stevens, C. F.. (1997). “Quantitative ultrastructural analysis of hippocampal excitatory synapses.” J Neurosci 17(15): 58585867.CrossRefGoogle ScholarPubMed
Schliwa, M. and Woehlke, G.. (2003). “Molecular motors.” Nature 422(6933): 759765.CrossRefGoogle ScholarPubMed
Siechen, S., et al. (2009). “Mechanical tension contributes to clustering of neurotransmitter vesicles at presynaptic terminals.” Proc Natl Acad Sci USA 106(31): 1261112616.CrossRefGoogle ScholarPubMed
Smith, B. A., et al. (2007). “Dendritic spine viscoelasticity and soft-glassy nature: balancing dynamic remodeling with structural stability.” Biophys J 92(4): 14191430.CrossRefGoogle ScholarPubMed
Star, E. N., Kwiatkowski, D. J., and Murthy, V. N.. (2002). “Rapid turnover of actin in dendritic spines and its regulation by activity.” Nat Neurosci 5(3): 239246.CrossRefGoogle ScholarPubMed
Sukharev, S. and Corey, D. P.. (2004). “Mechanosensitive channels: multiplicity of families and gating paradigms.” Sci STKE 2004(219): re4.CrossRefGoogle ScholarPubMed
Tasaki, I. and Byrne, P. M.. (1990). “Volume expansion of nonmyelinated nerve fibers during impulse conduction.” Biophys J 57(3): 633635.CrossRefGoogle ScholarPubMed
Tasaki, I. and Byrne, P. M.. (1992). “Heat production associated with a propagated impulse in bullfrog myelinated nerve fibers.” Jpn J Physiol 42(5): 805813.CrossRefGoogle ScholarPubMed
Tasaki, I., Iwasa, K., and Gibbons, R. C.. (1980). “Mechanical changes in crab nerve fibers during action potentials.” Jpn J Physiol 30(6): 897905.CrossRefGoogle ScholarPubMed
Tasaki, I., Kusano, K., and Byrne, P. M.. (1969). “Rapid mechanical and thermal changes in the garfish olfactory nerve associated with a propagated impulse.” Biophys J 55(6): 10331040.CrossRefGoogle Scholar
Thibault, K. L. and Margulies, S. S.. (1998). “Age-dependent material properties of the porcine cerebrum: effect on pediatric inertial head injury criteria.” J Biomech 31(12): 11191126.CrossRefGoogle ScholarPubMed
Umeda, T., Ebihara, T., and Okabe, S.. (2005). “Simultaneous observation of stably associated presynaptic varicosities and postsynaptic spines: morphological alterations of CA3-CA1 synapses in hippocampal slice cultures.” Mol Cell Neurosci 28(2): 264274.CrossRefGoogle ScholarPubMed
VanEssen, D. C. (1997). “A tension-based theory of morphogenesis and compact wiring in the central nervous system.” Nature 385(6614): 313318.Google Scholar
White, T. and Hilgetag, C. C.. (2011). “Gyrification and neural connectivity in schizophrenia.” Dev Psychopathol 23(1): 339352.CrossRefGoogle ScholarPubMed
Wolf, J. A., et al. (2001). “Traumatic axonal injury induces calcium influx modulated by tetrodotoxin-sensitive sodium channels.” J Neurosci 21(6): 19231930.CrossRefGoogle ScholarPubMed
Wuerfel, J., et al. (2010). “MR-elastography reveals degradation of tissue integrity in multiple sclerosis.” Neuroimage 49(3): 25202525.CrossRefGoogle ScholarPubMed
Xu, G. et al. (2010). “Axons pull on the brain, but tension does not drive cortical folding.” J Biomech Eng 132(7): 071013.CrossRefGoogle Scholar
Xu, G., Bayly, P. V., and Taber, L. A.. (2009). “Residual stress in the adult mouse brain.” Biomech Model Mechanobiol 8(4): 253262.CrossRefGoogle ScholarPubMed
Zhang, J., et al. (2011). “Viscoelastic properties of human cerebellum using magnetic resonance elastography.” Journal of Biomechanics 44(10): 19091913.CrossRefGoogle ScholarPubMed
Zhang, P. C., Keleshian, A. M., and Sachs, F.. (2001). “Voltage-induced membrane movement.” Nature 413(6854): 428432.CrossRefGoogle ScholarPubMed
Zheng, J., et al. (1991). “Tensile regulation of axonal elongation and initiation.” J Neurosci 11(4): 11171125.CrossRefGoogle ScholarPubMed
Zito, K., et al. (2004). “Induction of spine growth and synapse formation by regulation of the spine actin cytoskeleton.” Neuron 44(2): 321334.CrossRefGoogle ScholarPubMed
Zohar, O., et al. (1998). “Thermal imaging of receptor-activated heat production in single cells.” Biophys J 74(1): 8289.CrossRefGoogle ScholarPubMed

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
×