Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-dnltx Total loading time: 0 Render date: 2024-04-19T21:21:17.626Z Has data issue: false hasContentIssue false

Part I - Micro-nano techniques 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

Live cells can sense the mechanical characteristics of the microenvironment and translate the mechanical cues to intracellular biochemical signals in physiology and disease. To investigate intracellular signaling transduction during mechanosensing, nanotechnologies, and FRET live-cell imaging technologies have been developed to visualize the output signals in real time, such as intracellular molecular activity. Meanwhile, micropatterned technologies have been applied to modulate the physical and mechanical environment surrounding the cell to fine-tune the input signals in cellular mechanosensing. These advanced technologies can join forces and shed new light into the molecular networks that control mechanotransduction in normal conditions and disease.

Type
Chapter
Information
Integrative Mechanobiology
Micro- and Nano- Techniques in Cell Mechanobiology
, pp. 1 - 202
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

Akerman, M. E., Chan, W. C., Laakkonen, P., Bhatia, S. N., and Ruoslahti, E.. (2002). “Nanocrystal targeting in vivo.” Proc Natl Acad Sci USA 99(20): 1261712621.CrossRefGoogle ScholarPubMed
Alivisatos, A. P., Gu, W., and Larabell, C.. (2005). “Quantum dots as cellular probes.” Annu Rev Biomed Eng 7: 5576.CrossRefGoogle ScholarPubMed
Arias-Salgado, E. G., Lizano, S., Sarkar, S., Brugge, J. S., Ginsburg, M. H., and Shattil, S. J.. (2003). “Src kinase activation by direct interaction with the integrin beta cytoplasmic domain.” Proc Natl Acad Sci USA 100(23): 1329813302.CrossRefGoogle ScholarPubMed
Ballou, B., Lagerholm, B. C., Ernst, L. A., Bruchez, M. P., and Waggoner, A. S.. (2004). “Noninvasive imaging of quantum dots in mice.” Bioconjug Chem 15(1): 7986.CrossRefGoogle ScholarPubMed
Bao, X., Clark, C. B., and Frangos, J. A.. (2000). “Temporal gradient in shear-induced signaling pathway: involvement of MAP kinase, c-fos, and connexin43.” Am J Physiol Heart Circ Physiol 278(5): H15981605.CrossRefGoogle ScholarPubMed
Bao, X., Lu, C., and Frangos, J. A.. (1999). “Temporal gradient in shear but not steady shear stress induces PDGF-A and MCP-1 expression in endothelial cells: role of NO, NF kappa B, and egr-1.” Arterioscler Thromb Vasc Biol 19(4): 9961003.CrossRefGoogle Scholar
Bershadsky, A. D., Balaban, N. Q., and Geiger, B.. (2003). “Adhesion-dependent cell mechanosensitivity.” Annu Rev Cell Dev Biol 19: 677695.CrossRefGoogle ScholarPubMed
Bissell, M. J. and Hines, W. C.. (2011). “Why don’t we get more cancer? A proposed role of the microenvironment in restraining cancer progression.” Nat Med 17(3): 320329.CrossRefGoogle Scholar
Buranachai, C. and Clegg, R. M.. (2008). “Fluorescence lifetime imaging in living cells.” In Methods in Molecular Biology Fluorescent Proteins: Methods and Applications, edited by Rothnagel, J. (Totowa, NJ: Humana Press).Google Scholar
Butler, P. J., Weinbaum, S., Chien, S., and Lemons, D. E.. (2000). “Endothelium-dependent, shear-induced vasodilation is rate-sensitive.” Microcirculation 7(1): 5365.Google ScholarPubMed
Davies, P. F. (1997). “Overview: temporal and spatial relationships in shear stress-mediated endothelial signalling.” J Vasc Res 34(3): 208211.CrossRefGoogle ScholarPubMed
DePaola, N., Gimbrone, M. A. Jr., Davies, P.F., and Dewey, C. F.. (1992). “Vascular endothelium responds to fluid shear stress gradients.” Arterioscler Thromb 12(11): 12541257.CrossRefGoogle ScholarPubMed
Dubertret, B., Skourides, P., Norris, D. J., Noireaux, V., Brivanlou, A. H., and Libchaber, A.. (2002). “In vivo imaging of quantum dots encapsulated in phospholipid micelles.” Science 298(5599): 17591762.CrossRefGoogle ScholarPubMed
Frangos, J. A., Huang, T. Y., and Clark, C. B.. (1996). “Steady shear and step changes in shear stimulate endothelium via independent mechanisms–superposition of transient and sustained nitric oxide production.” Biochem Biophys Res Commun 224(3): 660665.CrossRefGoogle ScholarPubMed
Geiger, B., Spatz, J. P., and Bershadsky, A. D.. (2009). “Environmental sensing through focal adhesions.” Nat Rev Mol Cell Biol 10(1): 2133.CrossRefGoogle ScholarPubMed
Gimbrone, M. A. Jr. and Garcia-Cardena, G.. (2013). “Vascular endothelium, hemodynamics, and the pathobiology of atherosclerosis.” Cardiovasc Pathol 22(1): 915.CrossRefGoogle ScholarPubMed
Kim, T. J., Xu, J., Dong, R., Lu, S., Nuzzo, R., and Wang, Y.. (2009). “Visualizing the effect of microenvironment on the spatiotemporal RhoA and Src activities in living cells by FRET.” Small 5(12): 14531459.CrossRefGoogle ScholarPubMed
Kunkel, M. T., Ni, Q., Tsien, R. Y., Zhang, J., and Newton, A. C.. (2005). “Spatio-temporal dynamics of protein kinase B/Akt signaling revealed by a genetically encoded fluorescent reporter.” J Biol Chem 280(7): 55815587.CrossRefGoogle ScholarPubMed
Lei, L., Lu, S., Wang, Y., Kim, T., Mehta, D., and Wang, Y.. (2014). “The role of mechanical tension on lipid raft dependent PDGF-induced TRPC6 activation.” Biomaterials 35(9): 28682877.CrossRefGoogle ScholarPubMed
Li, S., Butler, P., Wang, Y., Hu, Y., Han, D. C., Unami, S., Guan, J.-L., et al. (2002). “The role of the dynamics of focal adhesion kinase in the mechanotaxis of endothelial cells.” Proc Natl Acad Sci USA 99(6): 35463551.CrossRefGoogle ScholarPubMed
Liao, X., Lu, S., Zhuo, Y., Winter, C., Xu, W., and Wang, Y.. (2012). “Visualization of Src and FAK activity during the differentiation process from hMSCs to osteoblasts.” PLoS One 7(8): e42709.CrossRefGoogle ScholarPubMed
Lu, S., Kim, T. J., Chen, C. E., Ouyang, M., Seong, J., Liao, X., and Wang, Y.. (2011). “Computational analysis of the spatiotemporal coordination of polarized PI3 K and Rac1 activities in micro-patterned live cells.” PLoS One 6(6): e21293.CrossRefGoogle Scholar
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
Makowski, L. and Hotamisligil, G. S.. (2004). “Fatty acid binding proteins–the evolutionary crossroads of inflammatory and metabolic responses.” J Nutr 134(9): 2464S2468S.CrossRefGoogle ScholarPubMed
Medintz, I. L., Clapp, A. R., Mattoussi, H., Goldman, E. R., Fisher, B., and Mauro, J. M.. (2003). “Self-assembled nanoscale biosensors based on quantum dot FRET donors.” Nat Mater 2(9): 630638.CrossRefGoogle ScholarPubMed
Michalet, X., Pinaud, F. F., Bentolilia, L. A., Tsay, J. M., Doose, S., Li, J. J., et al. (2005). “Quantum dots for live cells, in vivo imaging, and diagnostics.” Science 307(5709): 538544.CrossRefGoogle ScholarPubMed
Mitra, S. K., Hanson, D. A., and Schlaepfer, D. D.. (2005). “Focal adhesion kinase: in command and control of cell motility.” Nat Rev Mol Cell Biol 6(1): 5668.CrossRefGoogle ScholarPubMed
Mitra, S. K. and Schlaepfer, D. D.. (2006). “Integrin-regulated FAK-Src signaling in normal and cancer cells.” Curr Opin Cell Biol 18(5): 516523.CrossRefGoogle ScholarPubMed
Miyamoto, S., Teramoto, H., Coso, O. A., Gutkind, J. S., Akiyama, S. K., and Yamada, K. M.. (1995). “Integrin function: molecular hierarchies of cytoskeletal and signaling molecules.” J Cell Biol 131(3): 791805.CrossRefGoogle ScholarPubMed
Miyawaki, A., Llopis, J., Helm, R., McCaffery, J. M., Adams, J. A., Ikura, M., and Tsien, R. Y.. (1997). “Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin.” Nature 388(6645): 882887.CrossRefGoogle ScholarPubMed
Mochizuki, N., Yamashita, S., Kurokawa, K., Obha, Y., Nagai, T., Miyawaki, A., and Matsuda, A.. (2001). “Spatio-temporal images of growth-factor-induced activation of Ras and Rap1.” Nature 411(6841): 10651068.CrossRefGoogle ScholarPubMed
Nagel, T., Resnick, N., Dewey, C. F., and Gimbrone, M. A.. (1999). “Vascular endothelial cells respond to spatial gradients in fluid shear stress by enhanced activation of transcription factors.” Arterioscler Thromb Vasc Biol 19(8): 18251834.CrossRefGoogle ScholarPubMed
Ouyang, M., Huang, H., Shaner, N. C., Remacle, A. G., Shiryaev, S. A., Strongin, A. Y., Tsien, R. Y., et al. (2010). “Simultaneous visualization of protumorigenic Src and MT1-MMP activities with fluorescence resonance energy transfer.” Cancer Res 70(6): 22042212.CrossRefGoogle ScholarPubMed
Ouyang, M., Lu, S., Kim, T., Chen, C. E., Seong, J., Leckband, D. E., Wang, F., et al. (2013). “N-cadherin regulates spatially polarized signals through distinct p120ctn and beta-catenin-dependent signalling pathways.” Nat Commun 4: 1589.CrossRefGoogle ScholarPubMed
Ouyang, M., Sun, J., Chien, S., and Wang, Y.. (2008). “Determination of hierarchical relationship of Src and Rac at subcellular locations with FRET biosensors.” Proc Natl Acad Sci USA 105(38): 1435314358.CrossRefGoogle ScholarPubMed
Pertz, O., Hodgson, L., Klemke, R. L., and Hahn, K. M.. (2006). “Spatiotemporal dynamics of RhoA activity in migrating cells.” Nature 440(7087): 10691072.CrossRefGoogle ScholarPubMed
Prasuhn, D. E., Feltz, A., Blanco-Canosa, J., Susumo, K., Stewart, M. H., Mei, B. C., Yakoviev, A. V., et al. (2010). “Quantum dot peptide biosensors for monitoring caspase 3 proteolysis and calcium ions.” ACS Nano 4(9): 54875497.CrossRefGoogle ScholarPubMed
Reiss, P., Bleuse, J., and Pron, A.. (2002). “Highly Luminescent CdSe/ZnSe Core/Shell Nanocrystals of Low Size Dispersion.” Nano Lett 2(7): 781784.CrossRefGoogle Scholar
Resch-Genger, U., Grabolle, M., Cavaliere-Jaricot, S., Nitschke, R., and Nunn, T.. (2008). “Quantum dots versus organic dyes as fluorescent labels.” Nat Methods 5(9): 763775.CrossRefGoogle ScholarPubMed
Seong, J., Ouyang, M., Kim, T., Sun, J., Wen, P.-C., et al. (2011). “Detection of focal adhesion kinase activation at membrane microdomains by fluorescence resonance energy transfer.” Nat Commun 2: 406.CrossRefGoogle ScholarPubMed
Seong, J., Tajik, A., Sun, J., Guan, J.-L., Humphries, M. J., Craig, S. E., Shekaran, A., et al. (2013). “Distinct biophysical mechanisms of focal adhesion kinase mechanoactivation by different extracellular matrix proteins.” Proc Natl Acad Sci USA 110(48): 1937219377.CrossRefGoogle ScholarPubMed
Smith, A. M., Duan, H., Mohs, A. M., and Nie, S.. (2008). “Bioconjugated quantum dots for in vivo molecular and cellular imaging.” Adv Drug Deliv Rev 60(11): 12261240.CrossRefGoogle ScholarPubMed
Song, Y., Madahar, V., and Liao, J.. (2011). “Development of FRET assay into quantitative and high-throughput screening technology platforms for protein-protein interactions.” Ann Biomed Eng 39(4): 12241234.CrossRefGoogle ScholarPubMed
Sukhanova, A., Devy, J., Venteo, L., Kaplan, H., Artemyev, M., Oleinikov, V., Klinov, D., et al. (2004). “Biocompatible fluorescent nanocrystals for immunolabeling of membrane proteins and cells.” Anal Biochem 324(1): 6067.CrossRefGoogle ScholarPubMed
Tardy, Y., Resnick, N., Gimbone, M. A., and Dewey, C. F.. (1997). “Shear stress gradients remodel endothelial monolayers in vitro via a cell proliferation-migration-loss cycle.” Arterioscler Thromb Vasc Biol 17(11): 31023106.CrossRefGoogle Scholar
Thomas, S. M. and Brugge, J. S.. (1997). “Cellular functions regulated by Src family kinases.” Annu Rev Cell Dev Biol 13: 513609.CrossRefGoogle ScholarPubMed
Ting, A. Y., Kain, K. H., Klemke, R. L., and Tsien, R. Y.. (2001). “Genetically encoded fluorescent reporters of protein tyrosine kinase activities in living cells.” Proc Natl Acad Sci USA 98(26): 1500315008.CrossRefGoogle ScholarPubMed
Violin, J. D., Zhang, J., Tsien, R. Y., and Newton, A. C.. (2003). “A genetically encoded fluorescent reporter reveals oscillatory phosphorylation by protein kinase C.” J Cell Biol 161(5): 899909.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(7036): 10401045.CrossRefGoogle ScholarPubMed
Wang, Y., Shyy, J. Y., and Chien, S.. (2008). “Fluorescence proteins, live-cell imaging, and mechanobiology: seeing is believing.” Annu Rev Biomed Eng 10: 138.CrossRefGoogle ScholarPubMed
Zacharias, D. A., Violin, J. D., Newton, A. C., and Tsien, R. Y.. (2002). “Partitioning of lipid-modified monomeric GFPs into membrane microdomains of live cells.” Science 296(5569): 913916.CrossRefGoogle ScholarPubMed
Zhang, J., Hupfeld, C. J., Taylor, S. S., Olefsky, J. M., and Tsien, R. Y.. (2005). “Insulin disrupts beta-adrenergic signalling to protein kinase A in adipocytes.” Nature 437(7058): 569573.CrossRefGoogle ScholarPubMed
Zhang, J., Ma, Y., Taylor, S. S., and Tsien, R. Y.. (2001). “Genetically encoded reporters of protein kinase A activity reveal impact of substrate tethering.” Proc Natl Acad Sci USA 98(26): 1499715002.CrossRefGoogle ScholarPubMed

References

Anthis, N. J., Wegener, K. L., Ye, F., Kim, C., Goult, B. T., Lowe, E. D., Vakonakis, I., et al. (2009). “The structure of an integrin/talin complex reveals the basis of inside-out signal transduction.” EMBO J 28: 36233632.CrossRefGoogle ScholarPubMed
Bai, X. C., Fernandez, I. S., McMullan, G., and Scheres, S. H.. (2013). “Ribosome structures to near-atomic resolution from thirty thousand cryo-EM particles.” elife 2: e00461.CrossRefGoogle ScholarPubMed
Bakolitsa, C., Cohen, D. M., Bankston, L. A., Bobkov, A. A., Cadwell, G. W., Jennings, L., Critchley, D. R., et al. (2004). “Structural basis for vinculin activation at sites of cell adhesion.” Nature 430: 583586.CrossRefGoogle ScholarPubMed
Bakolitsa, C., de Pereda, J. M., Bagshaw, C. R., Critchley, D. R., and Liddington, R. C.. (1999). “Crystal structure of the vinculin tail suggests a pathway for activation.” Cell 99: 603613.CrossRefGoogle ScholarPubMed
Barstead, R. J., and Waterston, R. H.. (1991). “Vinculin is essential for muscle function in the nematode.” J Cell Biol 114: 715724.CrossRefGoogle ScholarPubMed
Bartesaghi, A., Lecumberry, F., Sapiro, G., and Subramaniam, S.. (2012). “Protein secondary structure determination by constrained single-particle cryo-electron tomography.” Structure 20: 20032013.CrossRefGoogle ScholarPubMed
Blanchoin, L., Boujemaa-Paterski, R., Sykes, C., and Plastino, J.. (2014). “Actin dynamics, architecture, and mechanics in cell motility.” Physiol Rev 94: 235263.CrossRefGoogle ScholarPubMed
Burridge, K., and Mangeat, P.. (1984). “An interaction between vinculin and talin.” Nature 308: 744746.CrossRefGoogle ScholarPubMed
Chafel, M. M., Shen, W., and Matsudaira, P.. (1995). “Sequential expression and differential localization of I-, L-, and T-fimbrin during differentiation of the mouse intestine and yolk sac.” Dev Dyn 203: 141151.CrossRefGoogle Scholar
Chen, H., Cohen, D. M., Choudhury, D. M., Kioka, N., and Craig, S. W.. (2005). “Spatial distribution and functional significance of activated vinculin in living cells.” J Cell Biol 169: 459470.CrossRefGoogle ScholarPubMed
Critchley, D. R. (2009). “Biochemical and structural properties of the integrin-associated cytoskeletal protein talin.” Annu Rev Biophys 38: 235254.CrossRefGoogle ScholarPubMed
Critchley, D. R., and Gingras, A. R.. (2008). “Talin at a glance.” J Cell Sci 121: 13451347.CrossRefGoogle ScholarPubMed
Debrand, E., El Jai, Y., Spence, L., Bate, N., Praekelt, U., Pritchard, C. A., Monkley, S. J., and Critchley, D. R.. (2009). “Talin 2 is a large and complex gene encoding multiple transcripts and protein isoforms.” FEBS J 276: 16101628.CrossRefGoogle ScholarPubMed
DeMali, K. A., Barlow, C. A., and Burridge, K.. (2002). “Recruitment of the Arp2/3 complex to vinculin: coupling membrane protrusion to matrix adhesion.” J Cell Biol 159: 881891.CrossRefGoogle ScholarPubMed
DeRosier, D. J., and Moore, P. B.. (1970). “Reconstruction of three-dimensional images from electron micrographs of structures with helical symmetry.” J Mol Biol 52: 355369.CrossRefGoogle ScholarPubMed
Dubochet, J., Adrian, M., Chang, J.-J., Homo, J.-C., Lepault, J., McDowall, A. W., and Schultz, P.. (1988). “Cryo-electron microscopy of vitrified specimens.” Q Rev Biophys 21: 129228.CrossRefGoogle ScholarPubMed
Egelman, E. H. (2000). “A robust algorithm for the reconstruction of helical filaments using single-particle methods.” Ultramicroscopy 85: 225–34.CrossRefGoogle ScholarPubMed
Egelman, E. H. (2007). “Single-particle reconstruction from EM images of helical filaments.” Curr Opin Struct Biol 17: 556561.CrossRefGoogle ScholarPubMed
Elliott, P. R., Goult, B. T., Kopp, P. M., Bate, N., Grossmann, J. G., Roberts, G. C., Critchley, D. R., et al. (2010). “The structure of the talin head reveals a novel extended conformation of the FERM domain.” Structure 18: 12891299.CrossRefGoogle ScholarPubMed
Frank, J. (2006a). Electron tomography: methods for three-dimensional visualization of structures in the cell. New York: Springer.CrossRefGoogle Scholar
Frank, J. (2006b). Three-dimensional electron microscopy of macromolecular assemblies: visualization of biological molecules in their native state. Oxford: Oxford University Press.CrossRefGoogle Scholar
Frank, J. (2013). “Story in a sample – the potential (and limitations) of cryo-electron microscopy applied to molecular machines.” Biopolymers 99: 832836.CrossRefGoogle Scholar
Galkin, V. E., Orlova, A., Cherepanova, O., Lebart, M. C., and Egelman, E. H.. (2008). “High-resolution cryo-EM structure of the F-actin-fimbrin/plastin ABD2 complex.” Proc Natl Acad Sci USA 105: 14941498.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
Giannone, G., Jiang, G., Sutton, D. H., Critchley, D. R., and Sheetz, M. P.. (2003). “Talin1 is critical for force-dependent reinforcement of initial integrin-cytoskeleton bonds but not tyrosine kinase activation.” J Cell Biol 163: 409419.CrossRefGoogle Scholar
Gingras, A. R., Bate, N., Goult, B. T., Hazelwood, L., Canestrelli, I., Grossmann, J. G., Liu, H., et al. (2007). “The structure of the C-terminal actin-binding domain of talin.” EMBO J 27: 458469.CrossRefGoogle ScholarPubMed
Gingras, A. R., Ziegler, W. H., Frank, R., Barsukov, I. L., Roberts, G. C., Critchley, D. R., and Emsley, J.. (2005). “Mapping and consensus sequence identification for multiple vinculin binding sites within the talin rod.” J Biol Chem 280: 3721737224.CrossRefGoogle ScholarPubMed
Goult, B. T., Xu, X. P., Gingras, A. R., Swift, M., Patel, B., Bate, N., Kopp, P. M., et al. (2013). “Structural studies on full-length talin1 reveal a compact auto-inhibited dimer: Implications for talin activation.” J Struc Biol 184: 2132.CrossRefGoogle ScholarPubMed
Grashoff, C., Hoffman, B. D., Brenner, M. D., Zhou, R., Parsons, M., Yang, M. T., McLean, M. A., et al. (2010). “Measuring mechanical tension across vinculin reveals regulation of focal adhesion dynamics.” Nature 466: 263266.CrossRefGoogle ScholarPubMed
Hanein, D., Matsudaira, P., and DeRosier, D. J.. (1997). “Evidence for a conformational change in actin induced by fimbrin (N375) binding.” J Cell Biol 139: 387396.CrossRefGoogle ScholarPubMed
Hanein, D., and Volkmann, N.. (2011). “Correlative light-electron microscopy.” Adv Protein Chem Struct Biol 82: 9199.CrossRefGoogle ScholarPubMed
Hanein, D., Volkmann, N., Goldsmith, S., Michon, A. M., Lehman, W., Craig, R., DeRosier, D., et al. (1998). “An atomic model of fimbrin binding to F-actin and its implications for filament crosslinking and regulation.” Nat Struct Biol 5: 787792.CrossRefGoogle ScholarPubMed
Hemmings, L., Rees, D. J., Ohanian, V., Bolton, S. J., Gilmore, A. P., Patel, B., Priddle, H., et al. (1996). “Talin contains three actin-binding sites each of which is adjacent to a vinculin-binding site.” J Cell Sci 109: 27152726.CrossRefGoogle ScholarPubMed
Higgs, H. N., and Pollard, T. D.. (1999). “Regulation of actin polymerization by Arp2/3 complex and WASp/Scar proteins.” J Biol Chem 274: 325313254.CrossRefGoogle ScholarPubMed
Holmes, K. C., Angert, I., Kull, F. J., Jahn, W., and Schroder, R. R.. (2003). “Electron cryo-microscopy shows how strong binding of myosin to actin releases nucleotide.” Nature 425: 423427.CrossRefGoogle ScholarPubMed
Janssen, M. E., Kim, E., Liu, H., Fujimoto, L. M., Bobkov, A., Volkmann, N., and Hanein, D.. (2006). “Three-dimensional structure of vinculin bound to actin filaments.” Mol Cell 21: 271281.CrossRefGoogle ScholarPubMed
Janssen, M. E., Liu, H., Volkmann, N., and Hanein, D.. (2012). “The C-terminal tail domain of metavinculin, vinculin’s splice variant, severs actin filaments.” J Cell Biol 197: 585593.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
Jockusch, B. M., and Isenberg, G.. (1981). “Interaction of alpha-actinin and vinculin with actin: opposite effects on filament network formation.” Proc Natl Acad Sci USA 78: 30053009.CrossRefGoogle ScholarPubMed
Jockusch, B. M., and Rudiger, M.. (1996). “Crosstalk between cell adhesion molecules: vinculin as a paradigm for regulation by conformation.” Trends Cell Biol 6: 311315.CrossRefGoogle ScholarPubMed
Johnson, R. P., and Craig, S. W.. (1995). “F-actin binding site masked by the intramolecular association of vinculin head and tail domains.” Nature 373: 261264.CrossRefGoogle ScholarPubMed
Johnson, R. P., and Craig, S. W.. (2000). “Actin activates a cryptic dimerization potential of the vinculin tail domain.” J Biol Chem 275: 95105.CrossRefGoogle ScholarPubMed
Kelleher, J. F., Atkinson, S. J., and Pollard, T. D.. (1995). “Sequences, structural models, and cellular localization of the actin– related proteins Arp2 and Arp3 from Acanthamoeba.” J Cell Biol 131: 385397.CrossRefGoogle ScholarPubMed
Klein, M. G., Shi, W., Ramagopal, U., Tseng, Y., Wirtz, D., Kovar, D. R., Staiger, C. J., et al. (2004). “Structure of the actin crosslinking core of fimbrin.” Structure 12: 9991013.CrossRefGoogle ScholarPubMed
Koster, A. J., Grimm, R., Typke, D., Hegerl, R., Stoschek, A., Walz, J., and Baumeister, W.. (1997). “Perspectives of molecular and cellular electron tomography.” J Struct Biol 120: 276308.CrossRefGoogle ScholarPubMed
Kühlbrandt, W. (2014). “Cryo-EM enters a new era.” elife 3: e03678.CrossRefGoogle ScholarPubMed
Le Clainche, C., and Carlier, M. F.. (2008). “Regulation of actin assembly associated with protrusion and adhesion in cell migration.” Physiol Rev 88: 489513.CrossRefGoogle ScholarPubMed
Lee, H.-S., Bellin, R. M., Walker, D. L., Patel, B., Powers, P., Liu, H., Garcia-Alvarez, B., et al. (2004). “Characterization of an actin-binding site within the talin FERM domain.” J Mol Biol 343: 771784.CrossRefGoogle ScholarPubMed
Luan, Q., and Nolen, B. J.. (2013). “Structural basis for regulation of Arp2/3 complex by GMF.” Nat Struct Mol Biol 20: 10621068.CrossRefGoogle ScholarPubMed
Lucic, V., Yang, T., Schweikert, G., Forster, F., and Baumeister, W.. (2005). “Morphological characterization of molecular complexes present in the synaptic cleft.” Structure 13: 423434.Google ScholarPubMed
Marchand, J. B., Kaiser, D. A., Pollard, T. D., and Higgs, H. N.. (2001). “Interaction of WASP/Scar proteins with actin and vertebrate Arp2/3 complex.” Nat Cell Biol 3: 7682.CrossRefGoogle ScholarPubMed
Martin, A. C., Xu, X. P., Rouiller, I., Kaksonen, M., Sun, Y., Belmont, L., Volkmann, N., et al. (2005). “Effects of Arp2 and Arp3 nucleotide-binding pocket mutations on Arp2/3 complex function.” J Cell Biol 168: 315328.CrossRefGoogle ScholarPubMed
Matsudaira, P. (1991). “Modular organization of actin crosslinking proteins.” Trends Biochem Sci 16: 8792.CrossRefGoogle ScholarPubMed
Nolen, B. J., and Pollard, T. D.. (2007). “Insights into the influence of nucleotides on actin family proteins from seven structures of Arp2/3 complex.” Mol Cell 26: 449457.CrossRefGoogle ScholarPubMed
Nolen, B. J., Littlefield, R. S., and Pollard, T. D.. (2004). “Crystal structures of actin-related protein 2/3 complex with bound ATP or ADP.” Proc Natl Acad Sci USA 101: 1562715632.CrossRefGoogle ScholarPubMed
Nolen, B. J., Tomasevic, N., Russell, A., Pierce, D. W., Jia, Z., McCormick, C. D., Hartman, J., et al. (2009). “Characterization of two classes of small molecule inhibitors of Arp2/3 complex.” Nature 460: 10311034.CrossRefGoogle ScholarPubMed
Parsons, J. T., Horwitz, A. R., and Schwartz, M. A.. (2010). “Cell adhesion: integrating cytoskeletal dynamics and cellular tension.” Nat Rev Mol Cell Biol 11: 633643.CrossRefGoogle ScholarPubMed
Penczek, P. A. (2010). “Image restoration in cryo-electron microscopy.” Methods Enzymol 482: 3572.CrossRefGoogle ScholarPubMed
Peng, X., Nelson, E. S., Maiers, J. L., and Demali, K. A.. (2011). “New insights into vinculin function and regulation.” Int Rev Cell Mol Biol 287: 191231.CrossRefGoogle ScholarPubMed
Pfaendtner, J., Volkmann, N., Hanein, D., Dalhaimer, P., Pollard, T. D., and Voth, G. A.. (2012). “Key structural features of the actin filament Arp2/3 complex branch junction revealed by molecular simulation.” J Mol Biol 416: 148161.CrossRefGoogle ScholarPubMed
Pollard, T. D. (2007). “Regulation of actin filament assembly by Arp2/3 complex and formins.” Annu Rev Biophys Biomol Struct 36: 451477.CrossRefGoogle ScholarPubMed
Radermacher, M. (1988). “Three-dimensional reconstruction of single particles from random and nonrandom tilt series.” J Electron Microsc Tech 9: 359–94.CrossRefGoogle ScholarPubMed
Robinson, R. C., Turbedsky, K., Kaiser, D. A., Marchand, J. B., Higgs, H. N., Choe, S., and Pollard, T. D.. (2001). “Crystal structure of Arp2/3 complex.” Science 294: 1679–184.CrossRefGoogle ScholarPubMed
Rouiller, I., Xu, X. P., Amann, K. J., Egile, C., Nickell, S., Nicastro, D., Li, R., et al. (2008). “The structural basis of actin filament branching by Arp2/3 complex.” J Cell Biol 180: 887895.CrossRefGoogle ScholarPubMed
Saunders, R. M., Holt, M. R., Jennings, L., Sutton, D. H., Barsukov, I. L., Bobkov, A., Liddington, R. C., Adamson, E. A., Dunn, G. A., and Critchley, D. R.. (2006). “Role of vinculin in regulating focal adhesion turnover.” Eur J Cell Biol 85: 487500.CrossRefGoogle ScholarPubMed
Scheres, S. H. (2012). “A Bayesian view on cryo-EM structure determination.” J Mol Biol 415: 406418.CrossRefGoogle ScholarPubMed
Schur, F. K., Hagen, W. J., de Marco, A., and Briggs, J. A.. (2013). “Determination of protein structure at 8.5 Å resolution using cryo-electron tomography and sub-tomogram averaging.” J Struc Biol 184: 394400.CrossRefGoogle ScholarPubMed
Spahn, C. M., and Penczek, P. A.. (2009). “Exploring conformational modes of macromolecular assemblies by multiparticle cryo-EM.” Curr Opin Struct Biol 19: 623631.CrossRefGoogle ScholarPubMed
Subauste, M. C., Pertz, O., Adamson, E. D., Turner, C. E., Junger, S., and Hahn, K. M.. (2004). “Vinculin modulation of paxillin-FAK interactions regulates ERK to control survival and motility.” J Cell Biol 165: 371381.CrossRefGoogle ScholarPubMed
Thievessen, I., Thompson, P. M., Berlemont, S., Plevock, K. M., Plotnikov, S. V., Zemljic-Harpf, A., 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
Ti, S. C., Jurgenson, C. T., Nolen, B. J., and Pollard, T. D.. (2011). “Structural and biochemical characterization of two binding sites for nucleation-promoting factor WASp-VCA on Arp2/3 complex.” Proc Natl Acad Sci USA 108: E463E471.CrossRefGoogle ScholarPubMed
Tilney, L. G., Egelman, E. H., DeRosier, D. J., and Saunder, J. C.. (1983). “Actin filaments, stereocilia, and hair cells of the bird cochlea. II. Packing of actin filaments in the stereocilia and in the cuticular plate and what happens to the organization when the stereocilia are bent.” J Cell Biol 96: 822834.CrossRefGoogle Scholar
Tilney, L. G., Tilney, M. S., and Guild, G. M.. (1995). “F actin bundles in Drosophila bristles. I. Two filament cross-links are involved in bundling.” J Cell Biol 130: 629638.CrossRefGoogle ScholarPubMed
van Heel, M. (1987). “Angular reconstitution: a posteriori assignment of projection directions for 3D reconstruction.” Ultramicroscopy 21: 111124.CrossRefGoogle Scholar
Volkmann, N. (2009). “Confidence intervals for fitting of atomic models into low-resolution densities.” Acta Crystallogr D Biol Crystallogr 65: 679689.CrossRefGoogle ScholarPubMed
Volkmann, N. (2012). “Putting structure into context: fitting of atomic models into electron microscopic and electron tomographic reconstructions.” Curr Opin Cell Biol 24: 141147.CrossRefGoogle ScholarPubMed
Volkmann, N., and Hanein, D.. (1999). “Quantitative fitting of atomic models into observed densities derived by electron microscopy.” J Struc Biol 125: 176184.CrossRefGoogle ScholarPubMed
Volkmann, N., and Hanein, D.. (2009). “Electron microscopy in the context of systems biology.” In Structural Bioinformatics, Gu, J. and Bourne, P. E., eds. New York: Wiley-Blackwell, 143170.Google Scholar
Volkmann, N., Page, C., Li, R., and Hanein, D.. (2014). “Three-dimensional reconstructions of actin filaments capped by Arp2/3 complex.” Eur J Cell Biol 93: 179183.CrossRefGoogle ScholarPubMed
Volkmann, N., Amann, K. J., Stoilova-McPhie, S., Egile, C., Winter, D. C., Hazelwood, L., Heuser, J. E., et al. (2001). “Structure of Arp2/3 complex in its activated state and in actin filament branch junctions.” Science 293: 24562459.CrossRefGoogle ScholarPubMed
Wegener, K. L., Partridge, A. W., Han, J., Pickford, A. R., Liddington, R. C., Ginsberg, M. H., and Campbell, I.D.. (2007). “Structural basis of integrin activation by talin.” Cell 128: 171182.CrossRefGoogle ScholarPubMed
Winkler, H., and Taylor, K. A.. (1996). “Three-dimensional distortion correction applied to tomographic reconstructions of sectioned crystals.” Ultramicroscopy 63: 125132.CrossRefGoogle ScholarPubMed
Xu, W., Baribault, H., and Adamson, E. D.. (1998). “Vinculin knockout results in heart and brain defects during embryonic development.” Development 125: 327337.CrossRefGoogle ScholarPubMed
Xu, X. P., Rouiller, I., Slaughter, B. D., Egile, C., Kim, E., Unruh, J. R., Fan, X., et al. (2011). “Three-dimensional reconstructions of Arp2/3 complex with bound nucleation promoting factors.” EMBO J 31: 236247.CrossRefGoogle ScholarPubMed
Zamir, E., Katz, M., Posen, Y., Erez, N., Yamada, K. M., Katz, B. Z., Lin, S., et al. (2000). “Dynamics and segregation of cell-matrix adhesions in cultured fibroblasts.” Nat Cell Biol 2: 191196.CrossRefGoogle ScholarPubMed

References

Aragona, M., et al. (2013). “A mechanical checkpoint controls multicellular growth through yap/taz regulation by actin-processing factors.” Cell 154: 10471059.CrossRefGoogle ScholarPubMed
Bao, G. and Suresh, S.. (2003). “Cell and molecular mechanics of biological materials.” Nature Mater 2: 715725.CrossRefGoogle ScholarPubMed
Butcher, D., et al. (2009). “A tense situation: Forcing tumour progression.” Nat Rev Cancer 9: 108122.CrossRefGoogle ScholarPubMed
Chen, C. S. (2008). “Mechanotransduction – a field pulling together?J Cell Sc 121: 32853292.CrossRefGoogle ScholarPubMed
Chen, C. S., et al. (2004). “Mechanotransduction at cell-matrix and cell-cell contacts.” Annu Rev Biomed Eng 6: 275302.CrossRefGoogle ScholarPubMed
Chiang, M., et al. (2013). “Relationships among cell morphology, intrinsic cell stiffness and cell-substrate interactions.” Biomaterials 34: 97549762.CrossRefGoogle ScholarPubMed
Chien, S. (2007). “Mechanotransduction and endothelial cell homeostasis: the wisdom of the cell.”Am J Physiol Heart Circ Physiol 292:H1209H1224.CrossRefGoogle ScholarPubMed
Chowdhury, F., et al. (2010). “Material properties of the cell dictate stress-induced spreading and differentiation in embryonic stem cells.” Nature Mater 9: 8288.CrossRefGoogle ScholarPubMed
du Roure, O., et al. (2005). “Force mapping in epithelial cell migration.” Proc Natl Acad Sci USA 102: 23902395.CrossRefGoogle ScholarPubMed
DuFort, C. C., et al. (2011). “Balancing forces: architectural control of mechanotransduction.” Nat RevMol Cell Biol 12: 308319.CrossRefGoogle ScholarPubMed
Engler, A., et al. (2006). “Matrix elasticity directs stem cell lineage specification.” Cell 126: 677689.CrossRefGoogle ScholarPubMed
Fenno, L., et al. (2011). “The development and application of optogenetics.” Annu Rev Neurosci 34: 389412.CrossRefGoogle ScholarPubMed
Fu, J., et al. (2010). “Mechanical regulation of cell function with geometrically modulated elastomeric substrates.” Nat Methods 7: 733736.CrossRefGoogle ScholarPubMed
Hahn, C. and Schwartz, M.. (2009). “Mechanotransduction in vascular physiology and atherogenesis.” Nat Rev Mol Cell Biol 10: 5362.CrossRefGoogle ScholarPubMed
Harris, A. K., et al. (1980). “Silicone-rubber substrata – new wrinkle in the study of cell locomotion.” Science 208: 177179.CrossRefGoogle Scholar
Hoffman, B., et al. (2011). “Dynamic molecular processes mediate cellular mechanotransduction.” Nature 475: 316323.CrossRefGoogle ScholarPubMed
Khetan, S., et al. (2013). “Degradation-mediated cellular traction directs stem cell fate in covalently crosslinked three-dimensional hydrogels.” Nature Mater 12: 458465.CrossRefGoogle ScholarPubMed
Kilian, K., et al. (2010). “Geometric cues for directing the differentiation of mesenchymal stem cells.” Proc Natl Acad Sci USA 107: 48724877.CrossRefGoogle ScholarPubMed
Kim, D.-H., et al. (2009). “Microengineered platforms for cell mechanobiology.” Annu Rev Biomed Eng 11: 203233.CrossRefGoogle ScholarPubMed
Lam, R., et al. (2012). “Live-cell subcellular measurement of cell stiffness using a microengineered stretchable micropost array membrane.” Integr Biol 4: 12891298.CrossRefGoogle ScholarPubMed
Levental, K., et al. (2009). “Matrix crosslinking forces tumor progression by enhancing integrin signaling.” Cell 139: 891906.CrossRefGoogle ScholarPubMed
Liu, Z., et al. (2010). “Mechanical tugging force regulates the size of cell-cell junctions.” Proc Natl Acad Sci USA 107: 99449949.CrossRefGoogle ScholarPubMed
Mammoto, T., et al. (2011). “Mechanochemical control of mesenchymal condensation and embryonic tooth organ formation.” Dev Cell 21: 758769.CrossRefGoogle ScholarPubMed
Mann, J., et al. (2012). “A silicone-based stretchable micropost array membrane for monitoring live-cell subcellular cytoskeletal response.” Lab Chip 12: 731740.CrossRefGoogle ScholarPubMed
Maruthamuthu, V., et al. (2011). “Cell-ECM traction force modulates endogenous tension at cell-cell contacts.” Proc Natl Acad Sci USA 108: 47084713.CrossRefGoogle ScholarPubMed
McBeath, R., et al. (2004). “Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineage commitment.” Dev Cell 6: 483495.CrossRefGoogle ScholarPubMed
Moore, S., et al. (2010). “Stretchy proteins on stretchy substrates: The important elements of integrin-mediated rigidity sensing.” Dev Cell 19: 194206.CrossRefGoogle ScholarPubMed
Nishimura, T., et al. (2012). “Planar cell polarity links axes of spatial dynamics in neural-tube closure.” Cell 149: 10841097.CrossRefGoogle ScholarPubMed
Oakes, P. and Gardel, M.. (2014). “Stressing the limits of focal adhesion mechanosensitivity.” Curr Opin Cell Biol 30C: 6873.CrossRefGoogle Scholar
Ono, A., et al. (2008). “Identification of a fibrin-independent platelet contractile mechanism regulating primary hemostasis and thrombus growth.” Blood 112: 9099.CrossRefGoogle ScholarPubMed
Ruiz, S. and Chen, C.. (2008). “Emergence of patterned stem cell differentiation within multicellular structures.” Stem Cells 26: 29212927.CrossRefGoogle ScholarPubMed
Shao, Y. and Fu, J. P.. (2014). “Integrated micro/nanoengineered functional biomaterials for cell mechanics and mechanobiology: a materials perspective.” Adv Mater 26, 14941533.CrossRefGoogle ScholarPubMed
Shao, Y., et al. (2014). “Global architecture of the F-actin cytoskeleton regulates cell shape-dependent endothelial mechanotransduction.” Integr Biol 6: 300311.CrossRefGoogle ScholarPubMed
Sun, Y. B., et al. (2014). “Hippo/YAP-mediated rigidity-dependent motor neuron differentiation of human pluripotent stem cells.” Nature Mater 13: 599604.CrossRefGoogle ScholarPubMed
Suresh, S. (2007). “Biomechanics and biophysics of cancer cells.” Acta Biomater 3: 413438.CrossRefGoogle ScholarPubMed
Suresh, S., et al. (2005). “Connections between single-cell biomechanics and human disease states: gastrointestinal cancer and malaria.” Acta Biomater 1: 1530.CrossRefGoogle ScholarPubMed
Tan, J., 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
Tee, S.-Y., et al. (2011). “Cell shape and substrate rigidity both regulate cell stiffness.” Biophys J 100: L25L27.CrossRefGoogle ScholarPubMed
Wang, N., et al. (1993). “Mechanotransduction across the cell-surface and through the cytoskeleton.” Science 260: 11241127.CrossRefGoogle ScholarPubMed
Wang, N., et al. (2009). “Mechanotransduction at a distance: mechanically coupling the extracellular matrix with the nucleus.” Nat Rev Mol Cell Biol 10: 7582.CrossRefGoogle Scholar
Wirtz, D., et al. (2011). “The physics of cancer: the role of physical interactions and mechanical forces in metastasis.” Nat Rev Cancer 11: 512522.CrossRefGoogle ScholarPubMed
Wozniak, M. A. and Chen, C. S.. (2009). “Mechanotransduction in development: a growing role for contractility.” Nat Rev Mol Cell Biol 10: 3443.CrossRefGoogle ScholarPubMed
Zhang, H. M., et al. (2011). “A tension-induced mechanotransduction pathway promotes epithelial morphogenesis.” Nature 471: 99103.CrossRefGoogle ScholarPubMed

References

Anderson, D. G. et al. (2005). “Biomaterial microarrays: rapid, microscale screening of polymer–cell interaction.” Biomaterials 26(23): 48924897.CrossRefGoogle ScholarPubMed
Anderson, J. R. et al. (2000). “Fabrication of topologically complex three-dimensional microfluidic systems in PDMS by rapid prototyping.” Analytical Chemistry 72(14): 31583164.CrossRefGoogle ScholarPubMed
Bellin, R. M. et al. (2009). “Defining the role of syndecan-4 in mechanotransduction using surface-modification approaches.” Proceedings of the National Academy of Sciences 106(52): 2210222107.CrossRefGoogle ScholarPubMed
Berthier, E., Young, E. W. K., and Beebe, D.. (2012). “Engineers are from PDMS-land, Biologists are from Polystyrenia.” Lab on a Chip 12(7): 12241237.CrossRefGoogle ScholarPubMed
Bongaerts, J. H. H., Fourtouni, K., and Stokes, J. R.. (2007). “Soft-tribology: lubrication in a compliant PDMS–PDMS contact.” Tribology International 40(10–12): 15311542.CrossRefGoogle Scholar
Brown, T. D. (2000). “Techniques for mechanical stimulation of cells in vitro: a review.” J Biomech 33(1): 314.CrossRefGoogle ScholarPubMed
Chen, J. H. and Simmons, C. A.. (2011). “Cell-matrix interactions in the pathobiology of calcific aortic valve disease: critical roles for matricellular, matricrine, and matrix mechanics cues.” Circ Res 108(12): 15101524.CrossRefGoogle ScholarPubMed
Dennes, T. J. and Schwartz, J.. (2008). “Controlling cell adhesion on polyurethanes.” Soft Matter 4(1): 8689.CrossRefGoogle Scholar
Duffy, D. C. et al. (1998). “Rapid prototyping of microfluidic systems in poly(dimethylsiloxane).” Analytical Chemistry 70(23): 49744984.CrossRefGoogle ScholarPubMed
Figallo, E. et al. (2007). “Micro-bioreactor array for controlling cellular microenvironments.” Lab on a Chip 7(6): 710719.CrossRefGoogle ScholarPubMed
Flaim, C. J., Chien, S., and Bhatia, S. N.. (2005). “An extracellular matrix microarray for probing cellular differentiation.” Nature Methods 2(2): 119125.CrossRefGoogle ScholarPubMed
Gómez-Sjöberg, R. et al. (2007). “Versatile, fully automated, microfluidic cell culture system.” Analytical Chemistry 79(22): 85578563.CrossRefGoogle ScholarPubMed
Gopalan, S. M. et al. (2003). “Anisotropic stretch-induced hypertrophy in neonatal ventricular myocytes micropatterned on deformable elastomers.” Biotechnology and Bioengineering 81(5): 578587.CrossRefGoogle ScholarPubMed
Hinz, B. (2010). “The myofibroblast: paradigm for a mechanically active cell.” Journal of Biomechanics 43(1): 146–55.CrossRefGoogle ScholarPubMed
Holick, M. F. (2000). “Microgravity-induced bone loss—will it limit human space exploration?The Lancet 355(9215): 15691570.CrossRefGoogle ScholarPubMed
Huang, J.-W. et al. (2013). “Interaction between lung cancer cell and myofibroblast influenced by cyclic tensile strain.” Lab on a Chip 13(6): 11141120.CrossRefGoogle ScholarPubMed
Huebsch, N. et al. (2010). “Harnessing traction-mediated manipulation of the cell/matrix interface to control stem-cell fate.” Nature Materials 9: 518526.CrossRefGoogle ScholarPubMed
Huh, D. et al. (2010). “Reconstituting organ-level lung functions on a chip.” Science 328(5986): 16621668.CrossRefGoogle ScholarPubMed
Hwang, N. S. et al. (2006). “Effects of three-dimensional culture and growth factors on the chondrogenic differentiation of murine embryonic stem cells.” Stem Cells 24: 284291.CrossRefGoogle ScholarPubMed
Jo, B. H. et al. (2000). “Three-dimensional micro-channel fabrication in polydimethylsiloxane (PDMS) elastomer.” Journal of Microelectromechanical Systems 9(1): 7681.CrossRefGoogle Scholar
Kim, T. K. and Jeong, O. C.. (2012). “Fabrication of a pneumatically-driven tensile stimulator.” Microelectronic Engineering 98: 715719.CrossRefGoogle Scholar
Kong, H. J. et al. (2005). “Non-viral gene delivery regulated by stiffness of cell adhesion substrates.” Nature Materials 4(6): 460464.CrossRefGoogle ScholarPubMed
Lee, J. N., Park, C. and Whitesides, G. M.. (2003). “Solvent compatibility of poly(dimethylsiloxane)-based microfluidic devices.” Analytical Chemistry 75(23): 65446554.CrossRefGoogle ScholarPubMed
Lee, S. W. and Lee, S. S.. (2008). “Shrinkage ratio of PDMS and its alignment method for the wafer level process.” Microsystem Technologies 14(2): 205208.CrossRefGoogle Scholar
MacKenna, D. A. et al. (1998). “Extracellular signal-regulated kinase and c-Jun NH2-terminal kinase activation by mechanical stretch is integrin-dependent and matrix-specific in rat cardiac fibroblasts.” Journal of Clinical Investigation 101(2): 301310.CrossRefGoogle ScholarPubMed
MacQueen, L. et al. (2012). “Miniaturized platform with on-chip strain sensors for compression testing of arrayed materials.” Lab on a Chip 12(20): 41784184.CrossRefGoogle ScholarPubMed
McBeath, R. et al. (2004). “Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineage commitment.” Developmental Cell 6(4): 483495.CrossRefGoogle ScholarPubMed
Moraes, C. et al. (2011). “Semi-confined compression of microfabricated polymerized biomaterial constructs.” Journal of Micromechanics and Microengineering 21, 054014.CrossRefGoogle Scholar
Moraes, C. et al. (2012). “Organs-on-a-Chip: A Focus on Compartmentalized Microdevices.” Annals of Biomedical Engineering 40(6): 12111227.CrossRefGoogle Scholar
Moraes, C. et al. (2013). “Microdevice array-based identification of distinct mechanobiological response profiles in layer-specific valve interstitial cells.” Integrative Biology 5, 673680.CrossRefGoogle ScholarPubMed
Moraes, C., Chen, J.-H., et al. (2010). “Microfabricated arrays for high-throughput screening of cellular response to cyclic substrate deformation.” Lab on a Chip 10, 227.CrossRefGoogle ScholarPubMed
Moraes, C., Kagoma, Y. K., et al. (2009). “Integrating polyurethane culture substrates into poly(dimethylsiloxane) microdevices.” Biomaterials 30(28): 52415250.CrossRefGoogle ScholarPubMed
Moraes, C., Wang, G. H., et al. (2010). A microfabricated platform for high-throughput unconfined compression of micropatterned biomaterial arrays. Biomaterials 31(3): 577584.CrossRefGoogle ScholarPubMed
Moraes, C., Sun, Y., and Simmons, C. A. (2009). “Solving the shrinkage-induced PDMS alignment registration issue in multilayer soft lithography.” Journal of Micromechanics and Microengineering 19, 065015.CrossRefGoogle Scholar
Moraes, C., Sun, Y., and Simmons, C. A.. (2010). “Microfabricated platforms for mechanically dynamic cell culture. journal of visualized experiments.” J Vis Exp 46: e2224.Google Scholar
Moraes, C., Sun, Y., and Simmons, C. A.. (2011). “(Micro)managing the mechanical microenvironment.” Integrative Biology 3: 959971.CrossRefGoogle ScholarPubMed
Mukundan, V., Nelson, W. J., and Pruitt, B. L.. (2012). “Microactuator device for integrated measurement of epithelium mechanics.” Biomedical Microdevices 15(1): 117123.CrossRefGoogle Scholar
Regehr, K. J. et al. (2009). “Biological implications of polydimethylsiloxane-based microfluidic cell culture.” Lab on a Chip 9: 2132.CrossRefGoogle ScholarPubMed
Scuor, N. et al. (2006). “Design of a novel MEMS platform for the biaxial stimulation of living cells.” Biomedical Microdevices 8(3): 239246.CrossRefGoogle ScholarPubMed
Shimizu, K. et al. (2011). “Development of a biochip with serially connected pneumatic balloons for cell-stretching culture.” Sensors and Actuators B: Chemical 156(1): 486493.CrossRefGoogle Scholar
Simmons, C. S. et al. (2011). “Integrated strain array for cellular mechanobiology studies.” J Micromech Microeng 21(5): 5401654025.CrossRefGoogle ScholarPubMed
Simon, C. G. and Lin-Gibson, S.. (2011). “Combinatorial and high-throughput screening of biomaterials.” Advanced Materials 23(3): 369387.CrossRefGoogle ScholarPubMed
Smalley, K. S. M., Lioni, M., and Herlyn, M.. (2006). “Life isn’t flat: taking cancer biology to the next dimension.” In Vitro Cellular & Developmental Biology – Animal 42: 242.CrossRefGoogle Scholar
Sniadecki, N. J. (2010). “Minireview: a tiny touch: activation of cell signaling pathways with magnetic nanoparticles.” Endocrinology 151(2): 451457.CrossRefGoogle Scholar
Suhir, E., 1991. Structural analysis in microelectronic and fiber-optic systems. New York: Van Nostrand Reinhold.CrossRefGoogle Scholar
Tan, W. et al. (2008). “Development and evaluation of microdevices for studying anisotropic biaxial cyclic stretch on cells.” Biomed Microdevices 10(6): 869882.CrossRefGoogle ScholarPubMed
Tomasek, J. J. et al. (2002). “Myofibroblasts and mechano-regulation of connective tissue remodelling.” Nat Rev Mol Cell Biol 3(5): 349363.CrossRefGoogle ScholarPubMed
Unger, M. A. et al. (2000). “Monolithic microfabricated valves and pumps by multilayer soft lithography.” Science 288(5463): 113116.CrossRefGoogle ScholarPubMed
Volder, M. D. and Reynaerts, D.. (2010). “Pneumatic and hydraulic microactuators: a review.” Journal of Micromechanics and Microengineering 20(4): 043001.CrossRefGoogle Scholar
Weaver, V. M. et al. (1997). “Reversion of the malignant phenotype of human breast cells in three-dimensional culture and in vivo by integrin blocking antibodies.” The Journal of Cell Biology 137(1): 231245.CrossRefGoogle ScholarPubMed
Wipff, P.-J. et al. (2009). “The covalent attachment of adhesion molecules to silicone membranes for cell stretching applications.” Biomaterials 30(9): 17811789.CrossRefGoogle ScholarPubMed
Wolff, J. (1986). Law of Bone Remodelling. New York: Springer.CrossRefGoogle Scholar
Yip, C. Y. Y. and Simmons, C. A.. (2011). “The aortic valve microenvironment and its role in calcific aortic valve disease.” Cardiovascular Pathology 20(3): 177182.CrossRefGoogle ScholarPubMed
Zandstra, P. W. (2004). “The opportunity of stem cell bioengineering.” Biotechnology and Bioengineering 88(3): 263263.CrossRefGoogle ScholarPubMed
Zhou, J. and Niklason, L. E.. (2012). “Microfluidic artificial ‘vessels’ for dynamic mechanical stimulation of mesenchymal stem cells.” Integrative Biology 4(12): 14871497.CrossRefGoogle ScholarPubMed

References

Ahn, E. H., Kim, Y., Kshitiz, , An, S. S., Afzal, J., Lee, S., et al. (2014). “Spatial control of adult stem cell fate using nanotopographic cues.” Biomaterials 35(8): 24012410.CrossRefGoogle ScholarPubMed
Aznavoorian, S, Stracke, M. L., Krutzsch, H., Schiffmann, E., and Liotta, L. A.. (1990). “Signal transduction for chemotaxis and haptotaxis by matrix molecules in tumor cells.” J Cell Biol 110(4): 14271438.CrossRefGoogle ScholarPubMed
Burdick, J. A., and Murphy, W. L.. (2012). “Moving from static to dynamic complexity in hydrogel design.” Nat Commun 3: 1269.CrossRefGoogle ScholarPubMed
Bursac, N., Parker, K. K., Iravanian, S., and Tung, L.. (2002). “Cardiomyocyte cultures with controlled macroscopic anisotropy: a model for functional electrophysiological studies of cardiac muscle.” Circ Res 91(12): e4554.CrossRefGoogle Scholar
Calderwood, D. A., and Ginsberg, M. H.. (2003). “Talin forges the links between integrins and actin.” Nat Cell Biol 5(8): 694697.CrossRefGoogle ScholarPubMed
Carter, S. B. (1965). “Principles of cell motility: the direction of cell movement and cancer invasion.” Nature 208(5016): 11831187.CrossRefGoogle ScholarPubMed
Carter, S. B. (1967). “Haptotaxis and the mechanism of cell motility.” Nature 213(5073): 256260.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(5317): 14251428.CrossRefGoogle ScholarPubMed
Chen, W., Villa-Diaz, L. G., Sun, Y., Weng, S., Kim, J. K., Lam, R. H. W., et al. (2012). “Nanotopography influences adhesion, spreading, and self-renewal of human embryonic stem cells.” ACS Nano 6(5): 40944103.CrossRefGoogle ScholarPubMed
Chien, K. R., Domian, I. J., and Parker, K. K.. (2008). “Cardiogenesis and the complex biology of regenerative cardiovascular medicine.” Science 322(5907): 14941497.CrossRefGoogle ScholarPubMed
Clark, P., Connolly, P., Curtis, A. S., Dow, J. A., and Wilkinson, C. D.. (1991). “Cell guidance by ultrafine topography in vitro.” J Cell Sci 9(Pt 1): 7377.CrossRefGoogle Scholar
Cukierman, E., Pankov, R., Stevens, D. R., and Yamada, K. M. (2001). “Taking cell-matrix adhesions to the third dimension.” Science 294(5547): 17081712.CrossRefGoogle Scholar
Dalby, M. J., Gadegaard, N., and Curtis, G.. (2007). “Nanotopographical control of human osteoprogenitor differentiation.” Curr Stem Cell Res Ther 2(2): 129138.CrossRefGoogle ScholarPubMed
Dalby, M. J., Gadegaard, N., and Oreffo, R. O. C.. (2014). “Harnessing nanotopography and integrin-matrix interactions to influence stem cell fate.” Nat Mater 13(6): 558569.CrossRefGoogle ScholarPubMed
Dalby, M. J., McCloy, D., Robertson, M., Wilkinson, C. D. W., and Oreffo, R. O. C.. (2006). “Osteoprogenitor response to defined topographies with nanoscale depths.” Biomaterials 27(8): 13061315.CrossRefGoogle ScholarPubMed
DeMali, K. A., Barlow, C. A., and Burridge, K.. (2002). “Recruitment of the Arp2/3 complex to vinculin: coupling membrane protrusion to matrix adhesion.” J Cell Biol 159(5): 881891.CrossRefGoogle ScholarPubMed
Diehl, K. A., Foley, J. D., Nealey, P. F., and Murphy, C. J.. (2005). “Nanoscale topography modulates corneal epithelial cell migration.” J Biomed Mater 75(3): 603611.CrossRefGoogle ScholarPubMed
Doyle, A. D., Wang, F. W., Matsumoto, K., and Yamada, K. M.. (2009). “One-dimensional topography underlies three-dimensional fibrillar cell migration.” J Cell Biol 184(4): 481–90.CrossRefGoogle ScholarPubMed
Dunn, G. A., and Ebendal, T.. (1978). “Contact guidance on oriented collagen gels.” Exp Cell Res 111(2): 475479.CrossRefGoogle ScholarPubMed
Dupont, S., Morsut, L., Aragona, M., Enzo, E., Giulitti, S., Cordenonsi, M., et al. (2011). “Role of YAP/TAZ in mechanotransduction.” Nature 474(7350): 179183.CrossRefGoogle ScholarPubMed
Engler, A. J., Sen, S., Sweeney, H. L., and Discher, D. E.. (2006). “Matrix elasticity directs stem cell lineage specification.” Cell 126(4): 677689.CrossRefGoogle ScholarPubMed
Erickson, C. A., and Nuccitelli, R.. (1984). “Embryonic fibroblast motility and orientation can be influenced by physiological electric fields.” J Cell Biol 98(1): 296307.CrossRefGoogle ScholarPubMed
Even-Ram, S., and Yamada, K. M.. (2005). “Cell migration in 3D matrix.” Curr Opin Cell Biol 17(5): 524532.CrossRefGoogle Scholar
Fast, V. G., Darrow, B. J., Saffitz, J. E., and Kléber, A. G.. (1996). “Anisotropic activation spread in heart cell monolayers assessed by high-resolution optical mapping. Role of tissue discontinuities.” Circ Res 79(1): 115127.CrossRefGoogle ScholarPubMed
Feng, J., Chan-Park, M. B., Shen, J., and Chan, V.. (2007). “Quick layer-by-layer assembly of aligned multilayers of vascular smooth muscle cells in deep microchannels.” Tissue Eng 13(5): 10031012.CrossRefGoogle ScholarPubMed
Fu, J., Wang, Y.-K., Yang, M. T., Desai, R. A., Yu, X., Liu, Z., et al. (2010). “Mechanical regulation of cell function with geometrically modulated elastomeric substrates.” Nat Methods 7(9): 733736.CrossRefGoogle ScholarPubMed
Gopalan, S. M., Flaim, C., Bhatia, S. N., Hoshijima, M., Knoell, R., Chien, K. R., et al. (2003). “Anisotropic stretch-induced hypertrophy in neonatal ventricular myocytes micropatterned on deformable elastomers.” Biotechnol Bioeng 81(5): 578587.CrossRefGoogle ScholarPubMed
Guilak, F., Cohen, D. M., Estes, B. T., Gimble, J. M., Liedtke, W., and Chen, C. S.. (2009). “Control of stem cell fate by physical interactions with the extracellular matrix.” Cell Stem Cell 5(1): 1726.CrossRefGoogle ScholarPubMed
Hoffman, B. D., Grashoff, C., and Schwartz, M. A.. (2011). “Dynamic molecular processes mediate cellular mechanotransduction.” Nature 475(7356): 316323.CrossRefGoogle ScholarPubMed
Hu, W., Yim, E. K. F., Reano, R. M., Leong, K. W., and Pang, S. W.. (2005). “Effects of nanoimprinted patterns in tissue-culture polystyrene on cell behavior.” J Vac Sci Technol 23(6): 29842989.CrossRefGoogle ScholarPubMed
Huang, J., Grater, S. V., Corbellini, F., Rinck, S., Bock, E., Kemkemer, R., et al. (2009). “Impact of order and disorder in RGD nanopatterns on cell adhesion.” Nano Lett 9(3): 11111116.CrossRefGoogle ScholarPubMed
Hwang, S. Y., Kwon, K. W., Jang, K.-J., Park, M. C., Lee, J. S., and Suh, K. Y.. (2010). “Adhesion assays of endothelial cells on nanopatterned surfaces within a microfluidic channel.” Anal Chem 82(7): 30163022.CrossRefGoogle ScholarPubMed
Isenberg, B. C., Backman, D. E., Kinahan, M. E., Jesudason, R., Suki, B., Stone, P. J., et al. (2012). “Micropatterned cell sheets with defined cell and extracellular matrix orientation exhibit anisotropic mechanical properties.” J Biomech 45(5): 756–61.CrossRefGoogle ScholarPubMed
Izaguirre, G., Aguirre, L., Hu, Y. P., Lee, H. Y., Schlaepfer, D. D., Aneskievich, B. J., et al. (2001). “The cytoskeletal/non-muscle isoform of alpha-actinin is phosphorylated on its actin-binding domain by the focal adhesion kinase.” J Biol Chem 276(31): 2867628685.CrossRefGoogle ScholarPubMed
Jagodzinski, M., Drescher, M., Zeichen, J., Hankemeier, S., Krettek, C., Bosch, U., et al. (2004). “Effects of cyclic longitudinal mechanical strain and dexamethasone on osteogenic differentiation of human bone marrow stromal cells.” Eur Cell Mater 7: 3541, 41.CrossRefGoogle ScholarPubMed
Jiao, A., Trosper, N. E., Yang, H. S., Kim, J., Tsui, J.H., Frankel, S. D., et al. (2014). “Thermoresponsive nanofabricated substratum for the engineering of three-dimensional tissues with layer-by-layer architectural control.” ACS Nano 8(5): 44304439.CrossRefGoogle ScholarPubMed
Kim, D.-H., Han, K., Gupta, K., Kwon, K. W., Suh, K.-Y., and Levchenko, A.. (2009). “Mechanosensitivity of fibroblast cell shape and movement to anisotropic substratum topography gradients.” Biomaterials 30(29): 54335444.CrossRefGoogle ScholarPubMed
Kim, D.-H., Kshitiz, R. R. Smith, P. Kim, E. H. Ahn, H. N. Kim, et al. (2012). “Nanopatterned cardiac cell patches promote stem cell niche formation and myocardial regeneration.” Integr Biol 4(9): 10191033.CrossRefGoogle ScholarPubMed
Kim, D.-H., Lipke, E. A., Kim, P., Cheong, R., Thompson, S., Delannoy, M., et al. (2010). “Nanoscale cues regulate the structure and function of macroscopic cardiac tissue constructs.” Proc Natl Acad Sci USA 107(2): 565570.CrossRefGoogle ScholarPubMed
Kim, D.-H., Seo, C. H., Han, K., Kwon, K. W., Levchenko, A., and Suh, K.-Y.. (2009). “Guided cell migration on microtextured substrates with variable local density and anisotropy.” Adv Funct Mater 19(10): 15791586.CrossRefGoogle ScholarPubMed
Kim, H. N., Jiao, A., Hwang, N. S., Kim, M. S., Kang, D. H., Kim, D.-H., et al. (2013). “Nanotopography-guided tissue engineering and regenerative medicine.” Adv Drug Deliv Rev 65(4): 536558.CrossRefGoogle ScholarPubMed
Kim, J., and Hayward, R. C.. (2012). “Mimicking dynamic in vivo environments with stimuli-responsive materials for cell culture.” Trends Biotechnol 30(8): 426439.CrossRefGoogle ScholarPubMed
Kim, J., Kim, H. N., Lim, K.-T., Kim, Y., Pandey, S., Garg, P., et al. (2013). “Synergistic effects of nanotopography and co-culture with endothelial cells on osteogenesis of mesenchymal stem cells.” Biomaterials 34(30): 72577268.CrossRefGoogle ScholarPubMed
Kim, J., Kim, H. N., Lim, K.-T., Kim, Y., Seonwoo, H., Park, S. H., et al. (2013). “Designing nanotopographical density of extracellular matrix for controlled morphology and function of human mesenchymal stem cells.” Sci Rep 3: 3552.CrossRefGoogle ScholarPubMed
Kim, D.-H., Provenzano, P. P., Smith, C. L., and Levchenko, A.. (2012). “Matrix nanotopography as a regulator of cell function.” J Cell Biol 197: 351360.CrossRefGoogle ScholarPubMed
Kshitiz, D.-H. Kim, D. J. Beebe, and A. Levchenko. (2011). “Micro- and nanoengineering for stem cell biology: the promise with a caution.” Trends Biotechnol 29(8): 399408.CrossRefGoogle ScholarPubMed
Kshitiz, J. Park, P. Kim, W. Helen, A. J. Engler, A. Levchenko, et al. (2012). “Control of stem cell fate and function by engineering physical microenvironments.” Integr Biol 4(9): 10081018.CrossRefGoogle ScholarPubMed
Kwan, A. P., Cummings, C. E., Chapman, J. A., and Grant, M. E.. (1991). “Macromolecular organization of chicken type X collagen in vitro.” J Cell Biol 114(3): 597604.CrossRefGoogle ScholarPubMed
Laflamme, M. A., and Murry, C. E.. (2011). “Heart regeneration.” Nature 473(7347): 326335.CrossRefGoogle ScholarPubMed
Lee, M. R., Kwon, K. W., Jung, H., Kim, H. N., Suh, K. Y., Kim, K., et al. (2010). “Direct differentiation of human embryonic stem cells into selective neurons on nanoscale ridge/groove pattern arrays.” Biomaterials 31(15): 43604366.CrossRefGoogle ScholarPubMed
Liliensiek, S. J., Wood, J. A., Yong, J., Auerbach, R. Nealey, P. F., and Murphy, C. J.. (2010). “Modulation of human vascular endothelial cell behaviors by nanotopographic cues.” Biomaterials 31(20): 54185426.CrossRefGoogle ScholarPubMed
Lin, Y.-D., Luo, C.-Y., Hu, Y.-N., Yeh, M.-L., Hsueh, Y.-C., Chang, M.-Y., et al. (2012). “Instructive nanofiber scaffolds with VEGF create a microenvironment for arteriogenesis and cardiac repair.” Sci Transl Med 4(146): 146ra109.CrossRefGoogle ScholarPubMed
Lo, C. M., Wang, H. B., Dembo, M., and Wang, Y. L.. (2000). “Cell movement is guided by the rigidity of the substrate.” Biophys J 79(1): 144152.CrossRefGoogle ScholarPubMed
Lowe, B. (1997). “The role of Ca2+ in deflection-induced excitation of motile, mechanoresponsive balancer cilia in the ctenophore statocyst.” J Exp Biol 200(Pt 11): 15931606.CrossRefGoogle ScholarPubMed
Malmström, J., Lovmand, J., Kristensen, S., Sundh, M., Duch, M., and Sutherland, D. S.. (2011). “Focal complex maturation and bridging on 200 nm vitronectin but not fibronectin patches reveal different mechanisms of focal adhesion formation.” Nano Lett 11(6): 22642271.CrossRefGoogle Scholar
McBeath, R., Pirone, D. M., Nelson, C. M., and Bhadriraju, K.. (2004). “Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineage commitment.” Dev Cell 6(4): 483495.CrossRefGoogle ScholarPubMed
McMurray, R. J., Gadegaard, N., Tsimbouri, P. M., Burgess, K. V. McNamara, L. E., Tare, R., et al. (2011). “Nanoscale surfaces for the long-term maintenance of mesenchymal stem cell phenotype and multipotency.” Nat Mater 10(8): 637644.CrossRefGoogle ScholarPubMed
Min, B.-M., Lee, G., Kim, S. H., Nam, Y. S., Lee, T. S., and Park, W. H.. (2004). “Electrospinning of silk fibroin nanofibers and its effect on the adhesion and spreading of normal human keratinocytes and fibroblasts in vitro.” Biomaterials 25(7–8): 12891297.CrossRefGoogle ScholarPubMed
Moore, S. W., and Sheetz, M.P.. (2011). “Biophysics of substrate interaction: influence on neural motility, differentiation, and repair.” Dev Neurobiol 7(11): 10901101CrossRefGoogle Scholar
Noh, H. K., Lee, S. W., Kim, J.-M., Oh, J.-E., Kim, K.-H., Chung, C.-P., et al. (2006). “Electrospinning of chitin nanofibers: degradation behavior and cellular response to normal human keratinocytes and fibroblasts.” Biomaterials 27(21): 39343944.CrossRefGoogle ScholarPubMed
Oh, S., Brammer, K. S., Li, Y. S. J., Teng, D., Engler, A. J., Chien, S., et al. (2009). “Stem cell fate dictated solely by altered nanotube dimension.” Proc Natl Acad Sci USA 106(7): 21302135.CrossRefGoogle ScholarPubMed
Petrie, R. J., Doyle, A. D., and Yamada, K. M.. (2009). “Random versus directionally persistent cell migration.” Nat Rev Mol Cell Biol 10(8): 538549.CrossRefGoogle ScholarPubMed
Pope, A. J., Sands, G. B., Smaill, B. H., and LeGrice, I. J.. (2008). “Three-dimensional transmural organization of perimysial collagen in the heart.” Am J Physiol Heart Circ Physiol 295(3): H1243H1252.CrossRefGoogle ScholarPubMed
Provenzano, P. P., Eliceiri, K. W., Campbell, J. M., Inman, D. R., White, J. G., and Keely, P. J.. (2006). “Collagen reorganization at the tumor-stromal interface facilitates local invasion.” BMC Med 4(1): 38.CrossRefGoogle ScholarPubMed
Provenzano, P. P., Inman, D. R, Eliceiri, K. W., Trier, S. M., Keely, P. J.. (2008). “Contact guidance mediated three-dimensional cell migration is regulated by Rho/ROCK-dependent matrix reorganization.” Biophys J 95(11): 53745384.CrossRefGoogle ScholarPubMed
Recknor, J. B., Sakaguchi, D. S., Mallapragada, S. K.. (2006). “Directed growth and selective differentiation of neural progenitor cells on micropatterned polymer substrates.” Biomaterials 27(22): 40984108.CrossRefGoogle ScholarPubMed
Riveline, D., Zamir, E., Balaban, N. Q., Schwarz, U. S., Ishizaki, T., Narumiya, S., 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(6): 11751186.CrossRefGoogle ScholarPubMed
Sakaguchi, K., Shimizu, T., Horaguchi, S., Sekine, H., Yamato, M., Umezu, M., et al. (2013). “In vitro engineering of vascularized tissue surrogates.” Sci Rep 3: 1316.CrossRefGoogle ScholarPubMed
Saranak, J., and Foster, K. W. (1997). “Rhodopsin guides fungal phototaxis.” Nature 387(6628): 465466.CrossRefGoogle ScholarPubMed
Sarkar, S., Dadhania, M., Rourke, P., Desai, T. A., and Wong, J. Y.. (2005). “Vascular tissue engineering: microtextured scaffold templates to control organization of vascular smooth muscle cells and extracellular matrix.” Acta Biomater 1(1): 93100.CrossRefGoogle ScholarPubMed
Sarkar, S., Isenberg, B. C., Hodis, E., Leach, J. B. Desai, T. A., and Wong, J. Y.. (2008). “Fabrication of a layered microstructured polycaprolactone construct for 3-D tissue engineering.” J Biomater Sci Polym Ed 19(10): 13471362.CrossRefGoogle ScholarPubMed
Sekine, H., Shimizu, T., Sakaguchi, K., Dobashi, I., Wada, M., Yamato, M., et al. (2013). “In vitro fabrication of functional three-dimensional tissues with perfusable blood vessels.” Nat Commun 4: 1399.CrossRefGoogle ScholarPubMed
Seo, C. H., Jeong, H., Furukawa, K. S., Suzuki, Y., and Ushida, T. 2013. “The switching of focal adhesion maturation sites and actin filament activation for MSCs by topography of well-defined micropatterned surfaces.” Biomaterials 34(7): 17641771.CrossRefGoogle ScholarPubMed
Seong, J., Tajik, A., Sun, J. Guan, J.-L., Humphries, M. J., Craig, S. E., et al. (2013). “Distinct biophysical mechanisms of focal adhesion kinase mechanoactivation by different extracellular matrix proteins.” Proc Natl Acad Sci USA 110(48): 1937219377.CrossRefGoogle ScholarPubMed
Shimizu, T., Yamato, M., and Akutsu, T.. (2002). “Electrically communicating three‐dimensional cardiac tissue mimic fabricated by layered cultured cardiomyocyte sheets.” J Biomed Mater Res 60(10): 110117.CrossRefGoogle ScholarPubMed
Stephens, M., Kwan, A. P., Bayliss, M. T., and Archer, C. W.. (1992). “Human articular surface chondrocytes initiate alkaline phosphatase and type X collagen synthesis in suspension culture.” J Cell Sci 103(Pt 4): 11111116.CrossRefGoogle ScholarPubMed
Suh, K.-Y., Park, M. C., and Kim, P.. (2009). “Capillary force lithography: a versatile tool for structured biomaterials interface towards cell and tissue engineering.” Adv Funct Mater 19(17): 2699–712.CrossRefGoogle Scholar
Tan, J., and Saltzman, W. M.. (2002). “Topographical control of human neutrophil motility on micropatterned materials with various surface chemistry.” Biomaterials 23(15): 32153225.CrossRefGoogle ScholarPubMed
Tan, W., and Desai, T. A.. (2005). “Microscale multilayer cocultures for biomimetic blood vessels.” J Biomed Mater Res 72(2): 146160.CrossRefGoogle ScholarPubMed
Tanaka, S. M., Sun, H. B., Roeder, R. K., Burr, D. B., Turner, C. H., and Yokota, H.. (2005). “Osteoblast responses one hour after load-induced fluid flow in a three-dimensional porous matrix.” Calcif Tissue Int 76(4): 261271.CrossRefGoogle Scholar
Teixeira, A. I., Abrams, G. A., Bertics, P. J., Murphy, C. J., and Nealey, P. F.. (2003). “Epithelial contact guidance on well-defined micro- and nanostructured substrates.” J Cell Sci 116(Pt 10): 18811892.CrossRefGoogle ScholarPubMed
Teixeira, A. I., McKie, G. A., Foley, J. D., Bertics, P. J., Nealey, P. F., and Murphy, C. J.. (2006). “The effect of environmental factors on the response of human corneal epithelial cells to nanoscale substrate topography.” Biomaterials 27(21): 39453954.CrossRefGoogle ScholarPubMed
Tsiper, M. V., and Yurchenco, P. D.. (2002). “Laminin assembles into separate basement membrane and fibrillar matrices in Schwann cells.” J Cell Sci 115(5): 10051015.CrossRefGoogle ScholarPubMed
Venugopal, J., and Ramakrishna, S.Biocompatible nanofiber matrices for the engineering of a dermal substitute for skin regeneration.” Tissue Eng 11(5–6): 847–54.Google Scholar
Venugopal, J. R., Zhang, Y., and Ramakrishna, S.. (2006). “In vitro culture of human dermal fibroblasts on electrospun polycaprolactone collagen nanofibrous membrane.” Artif Organs 30(6): 440446.CrossRefGoogle ScholarPubMed
Yang, H. S., Ieronimakis, N., Tsui, J. H., Kim, H. N., Suh, K.-Y., Reyes, M., et al. (2014). “Nanopatterned muscle cell patches for enhanced myogenesis and dystrophin expression in a mouse model of muscular dystrophy.” Biomaterials 35(5): 1478–86.CrossRefGoogle Scholar
Yim, E. K. F., Pang, S. W., and Leong, K. W.. (2007). “Synthetic nanostructures inducing differentiation of human mesenchymal stem cells into neuronal lineage.” Exp Cell Res 313(9): 18201829.CrossRefGoogle ScholarPubMed
Yurchenco, P. D., and Wadsworth, W. G.. (2004) “Assembly and tissue functions of early embryonic laminins and netrins.” Curr Opin Cell Biol 16(5): 572579.CrossRefGoogle ScholarPubMed
Zorlutuna, P., Elsheikh, A., Hasirci, V.. (2009). “Nanopatterning of collagen scaffolds improve the mechanical properties of tissue engineered vascular grafts.” Biomacromolecules 10(4): 814821.CrossRefGoogle ScholarPubMed

References

Anseth, K. S., Metters, A. T., Bryant, S. J., Martens, P. J., Elisseeff, J. H. and Bowman, C. N. (2002). “In situ forming degradable networks and their application in tissue engineering and drug delivery.” Journal of Controlled Release 78: 199209.CrossRefGoogle ScholarPubMed
Barradas, A. M. C., Fernandes, H. A. M., Groen, N., Chai, Y. C., Schrooten, J., van de Peppel, J., van Leeuwen, J., van Blitterswijk, C. A. and de Boer, J. (2012). “A calcium-induced signaling cascade leading to osteogenic differentiation of human bone marrow-derived mesenchymal stromal cells.” Biomaterials 33: 32053215.CrossRefGoogle ScholarPubMed
Bath, J. and Turberfield, A. J. (2007). “DNA nanomachines.” Nature Nanotechnology 2: 275284.CrossRefGoogle ScholarPubMed
Burdick, J. A. and Prestwich, G. D. (2011). “Hyaluronic acid hydrogels for biomedical applications.” Advanced Materials 23: H41H56.CrossRefGoogle ScholarPubMed
Cheng, E. J., Xing, Y. Z., Chen, P., Yang, Y., Sun, Y. W., Zhou, D. J., Xu, L. J., Fan, Q. H. and Liu, D. S. (2009). “A pH-triggered, fast-responding DNA hydrogel.” Angewandte Chemie-International Edition 48: 76607663.CrossRefGoogle ScholarPubMed
Choi, S., Yu, X. H., Jongpaiboonkit, L., Hollister, S. J. and Murphy, W. L. (2013). “Inorganic coatings for optimized non-viral transfection of stem cells.” Scientific Reports 3: 1587.CrossRefGoogle ScholarPubMed
Daley, W. P., Peters, S. B. and Larsen, M. (2008). “Extracellular matrix dynamics in development and regenerative medicine.” Journal of Cell Science 121: 255264.CrossRefGoogle ScholarPubMed
de Groot, C. J., van Luyn, M. J. A., van Dijk-Wolthuis, W. N. E., Cadee, J. A., Plantinga, J. A., Den Otter, W. and Hennink, W. E. (2001). “In vitro biocompatibility of biodegradable dextran-based hydrogels tested with human fibroblasts.” Biomaterials 22: 11971203.CrossRefGoogle ScholarPubMed
DeForest, C. A. and Anseth, K. S. (2011). “Cytocompatible click-based hydrogels with dynamically tunable properties through orthogonal photoconjugation and photocleavage reactions.” Nature Chemistry 3: 925931.CrossRefGoogle ScholarPubMed
DeForest, C. A., Polizzotti, B. D. and Anseth, K. S. (2009). “Sequential click reactions for synthesizing and patterning three-dimensional cell microenvironments.” Nature Materials 8: 659664.CrossRefGoogle ScholarPubMed
Dietz, H., Douglas, S. M. and Shih, W. M. (2009). “Folding DNA into twisted and curved nanoscale shapes.” Science 325: 725730.CrossRefGoogle ScholarPubMed
Discher, D. E., Mooney, D. J. and Zandstra, P. W. (2009). “Growth factors, matrices, and forces combine and control stem cells.” Science 324: 16731677.CrossRefGoogle ScholarPubMed
Dixon, J. E., Shah, D. A., Rogers, C., Hall, S., Weston, N., Parmenter, C. D. J., McNally, D., et al. (2014). “Combined hydrogels that switch human pluripotent stem cells from self-renewal to differentiation.” Proceedings of the National Academy of Sciences USA 111: 55805585.CrossRefGoogle ScholarPubMed
Ebara, M., Yamato, M., Aoyagi, T., Kikuchi, A., Sakai, K. and Okano, T. (2008). “The effect of extensible PEG tethers on shielding between grafted thermo-responsive polymer chains and integrin-RGD binding.” Biomaterials 29: 36503655.CrossRefGoogle ScholarPubMed
Ellis-Davies, G. C. R. (2007). “Caged compounds: photorelease technology for control of cellular chemistry and physiology.” Nature Methods 4: 619628.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
Gawel, K. and Stokke, B. T. (2011). “Logic swelling response of DNA-polymer hybrid hydrogel.” Soft Matter 7: 46154618.CrossRefGoogle Scholar
Georges, P. C., Hui, J. J., Gombos, Z., Mccormick, M. E., Wang, A. Y., Uemura, M., Mick, R., et al. (2007). “Increased stiffness of the rat liver precedes matrix deposition: implications for fibrosis.” American Journal of Physiology-Gastrointestinal and Liver Physiology 293: G1147G1154.CrossRefGoogle ScholarPubMed
Gillette, B. M., Jensen, J. A., Wang, M. X., Tchao, J. and Sia, S. K. (2010). “Dynamic hydrogels: switching of 3D microenvironments using two-component naturally derived extracellular matrices.” Advanced Materials 22: 686691.CrossRefGoogle ScholarPubMed
Gobin, A. S. and West, J. L. (2002). “Cell migration through defined, synthetic extracellular matrix analogues.” Faseb Journal 16: 751753.CrossRefGoogle Scholar
Grieshaber, S. E., Jha, A. K., Farran, A. J. E. and Jia, X. Q. (2011). “Hydrogels in tissue engineering.” In Biomaterials for Tissue Engineering Applications: A Review of the Past and Future Trends, Burdick, J. A. and Mauck, R. L., eds. New York: Springer, 946.CrossRefGoogle Scholar
Griffin, D. R. and Kasko, A. M. (2012). “Photodegradable macromers and hydrogels for live cell encapsulation and release.” Journal of the American Chemical Society 134: 1310313107.CrossRefGoogle ScholarPubMed
Guvendiren, M. and Burdick, J. A. (2012). “Stiffening hydrogels to probe short– and long-term cellular responses to dynamic mechanics.” Nature Communications 3: 792.CrossRefGoogle Scholar
Guvendiren, M., Perepelyuk, M., Wells, R. G. and Burdick, J. A. (2014). “Hydrogels with differential and patterned mechanics to study stiffness-mediated myofibroblastic differentiation of hepatic stellate cells.” Journal of the Mechanical Behavior of Biomedical Materials 38: 198208.CrossRefGoogle ScholarPubMed
Guvendiren, M., Purcell, B. and Burdick, J. A. (2012).“Photopolymerizable systems.” In Polymer Science: A Comprehensive Reference, vol. 9, Krzysztof, M. and Martin, M., eds. Amsterdam: Elsevier, 413438.CrossRefGoogle Scholar
Hadjipanayi, E., Mudera, V. and Brown, R. A. (2009). “Guiding cell migration in 3D: a collagen matrix with graded directional stiffness.” Cell Motility and the Cytoskeleton 66: 121128.CrossRefGoogle ScholarPubMed
Hahn, M. S., Miller, J. S. and West, J. L. (2006). “Three-dimensional biochemical and biomechanical patterning of hydrogels for guiding cell behavior.” Advanced Materials 18: 26792684.CrossRefGoogle Scholar
Han, D. R., Pal, S., Nangreave, J., Deng, Z. T., Liu, Y. and Yan, H. (2011). “DNA origami with complex curvatures in three-dimensional space.” Science 332: 342346.CrossRefGoogle ScholarPubMed
He, X. J., Weiz, B. and Mi, Y. L. (2010). “Aptamer based reversible DNA induced hydrogel system for molecular recognition and separation.” Chemical Communications 46: 63086310.CrossRefGoogle ScholarPubMed
Hoffman, A. S. and Stayton, P. S. (2007). “Conjugates of stimuli-responsive polymers and proteins.” Progress in Polymer Science 32: 922932.CrossRefGoogle Scholar
Hoffman, A. S., Stayton, P. S., Shimoboji, T., Chen, G. H., Ding, Z. L., Chilkoti, A., Long, C., et al. (1997). “Conjugates of stimuli-responsive polymers and biomolecules: Random and site-specific conjugates of temperature-sensitive polymers and proteins.” Macromolecular Symposia 118: 553563.CrossRefGoogle Scholar
Hou, X. and Jiang, L. (2009). “Learning from nature: building bio-inspired smart nanochannels.” Acs Nano 3: 33393342.CrossRefGoogle ScholarPubMed
Huang, S. and Ingber, D. E. (2005). “Cell tension, matrix mechanics, and cancer development.” Cancer Cell 8: 175176.CrossRefGoogle ScholarPubMed
Ifkovits, J. L. and Burdick, J. A. (2007). “Review: photopolymerizable and degradable biomaterials for tissue engineering applications.” Tissue Engineering 13: 23692385.CrossRefGoogle ScholarPubMed
Jiang, F. X., et al. (2010a). “Effect of dynamic stiffness of the substrates on neurite outgrowth by using a DNA-crosslinked hydrogel.” Tissue Engineering Part A 16: 18731889.CrossRefGoogle ScholarPubMed
Jiang, F. X., et al. (2010b). “The relationship between fibroblast growth and the dynamic stiffnesses of a DNA crosslinked hydrogel.” Biomaterials 31: 11991212.CrossRefGoogle ScholarPubMed
Jo, S., Shin, H. and Mikos, A. G. (2001). “Modification of oligo(poly(ethylene glycol) fumarate) macromer with a GRGD peptide for the preparation of functionalized polymer networks.” Biomacromolecules 2: 255261.CrossRefGoogle ScholarPubMed
Kang, H. Z., Liu, H. P., Zhang, X. L., Yan, J. L., Zhu, Z., Peng, L., Yang, H. H., et al. (2011). “Photoresponsive DNA-cross-linked hydrogels for controllable release and cancer therapy.” Langmuir 27: 399408.CrossRefGoogle ScholarPubMed
Khetan, S. and Burdick, J. A. (2010). “Patterning network structure to spatially control cellular remodeling and stem cell fate within 3-dimensional hydrogels.” Biomaterials 31: 82288234.CrossRefGoogle ScholarPubMed
Khetan, S., Guvendiren, M., Legant, W. R., Cohen, D. M., Chen, C. S. and Burdick, J. A. (2013). “Degradation-mediated cellular traction directs stem cell fate in covalently crosslinked three-dimensional hydrogels.” Nature Materials 12: 458465.CrossRefGoogle ScholarPubMed
Khetan, S., Katz, J. S. and Burdick, J. A. (2009). “Sequential crosslinking to control cellular spreading in 3-dimensional hydrogels.” Soft Matter 5: 16011606.CrossRefGoogle Scholar
Kim, H. S. and Yoo, H. S. (2010). “MMPs-responsive release of DNA from electrospun nanofibrous matrix for local gene therapy: in vitro and in vivo evaluation.” Journal of Controlled Release 145: 264271.CrossRefGoogle ScholarPubMed
Kim, S. and Healy, K. E. (2003). “Synthesis and characterization of injectable poly(N-isopropylacrylamide-co-acrylic acid) hydrogels with proteolytically degradable cross-links.” Biomacromolecules 4: 12141223.CrossRefGoogle ScholarPubMed
Klouda, L. and Mikos, A. G. (2008). “Thermoresponsive hydrogels in biomedical applications.” European Journal of Pharmaceutics and Biopharmaceutics 68: 3445.CrossRefGoogle ScholarPubMed
Kloxin, A. M., Benton, J. A. and Anseth, K. S. (2010a). “In situ elasticity modulation with dynamic substrates to direct cell phenotype.” Biomaterials 31: 18.CrossRefGoogle ScholarPubMed
Kloxin, A. M., Tibbett, M. W. and Anseth, K. S. (2010b). “Synthesis of photodegradable hydrogels as dynamically tunable cell culture platforms.” Nature Protocols 5: 18671887.CrossRefGoogle ScholarPubMed
Kloxin, A. M., Kasko, A. M., Salinas, C. N. and Anseth, K. S. (2009). “Photodegradable hydrogels for dynamic tuning of physical and chemical properties.” Science 324: 5963.CrossRefGoogle ScholarPubMed
Krishnan, Y. and Simmel, F. C. (2011). “Nucleic acid based molecular devices.” Angewandte Chemie-International Edition 50: 31243156.CrossRefGoogle ScholarPubMed
Liedl, T., et al. (2007a). “Controlled trapping and release of quantum dots in a DNA-switchable hydrogel.” Small 3: 16881693.CrossRefGoogle Scholar
Liedl, T., Sobey, T. L. and Simmel, F. C. (2007b). “DNA-based nanodevices.” Nano Today 2: 3641.CrossRefGoogle Scholar
Lin, D. C., Yurke, B. and Langrana, N. A. (2004). “Mechanical properties of a reversible, DNA-crosslinked polyacrylamide hydrogel.” Journal of Biomechanical Engineering 126: 104110.CrossRefGoogle ScholarPubMed
Lin, D. C., Yurke, B. and Langrana, N. A. (2005). “Inducing reversible stiffness changes in DNA-crosslinked gels.” Journal of Materials Research 20: 14561464.CrossRefGoogle Scholar
Liu, D. S. and Balasubramanian, S. (2003). “A proton-fuelled DNA nanomachine.” Angewandte Chemie-International Edition 42: 57345736.CrossRefGoogle ScholarPubMed
Liu, D. S., Cheng, E. J. and Yang, Z. Q. (2011). “DNA-based switchable devices and materials.” NPG Asia Materials 3: 109114.CrossRefGoogle Scholar
Luo, Y. and Shoichet, M. S. (2004). “A photolabile hydrogel for guided three-dimensional cell growth and migration.” Nature Materials 3: 249253.CrossRefGoogle ScholarPubMed
Lutolf, M. P. and Hubbell, J. A. (2005). “Synthetic biomaterials as instructive extracellular microenvironments for morphogenesis in tissue engineering.” Nature Biotechnology 23: 4755.CrossRefGoogle ScholarPubMed
Lutolf, M. P., et al. (2003a). “Synthetic matrix metalloproteinase-sensitive hydrogels for the conduction of tissue regeneration: engineering cell-invasion characteristics.” Proceedings of the National Academy of Sciences USA 100: 54135418.CrossRefGoogle ScholarPubMed
Lutolf, M. P., et al. (2003b). “Cell-responsive synthetic hydrogels.” Advanced Materials 15: 888892.CrossRefGoogle Scholar
Marklein, R. A. and Burdick, J. A. (2010). “Spatially controlled hydrogel mechanics to modulate stem cell interactions.” Soft Matter 6: 136143.CrossRefGoogle Scholar
Marklein, R. A., Soranno, D. E. and Burdick, J. A. (2012). “Magnitude and presentation of mechanical signals influence adult stem cell behavior in 3-dimensional macroporous hydrogels.” Soft Matter 8: 81138120.CrossRefGoogle Scholar
Martens, P., Holland, T. and Anseth, K. S. (2002). “Synthesis and characterization of degradable hydrogels formed from acrylate modified poly(vinyl alcohol) macromers.” Polymer 43: 60936100.CrossRefGoogle Scholar
Mather, B. D., Viswanathan, K., Miller, K. M. and Long, T. E. (2006). “Michael addition reactions in macromolecular design for emerging technologies.” Progress in Polymer Science 31: 487531.CrossRefGoogle Scholar
Metters, A. and Hubbell, J. (2005). “Network formation and degradation behavior of hydrogels formed by Michael-type addition reactions.” Biomacromolecules 6: 290301.CrossRefGoogle ScholarPubMed
Metters, A. T., Anseth, K. S. and Bowman, C. N. (2000). “Fundamental studies of a novel, biodegradable PEG-b-PLA hydrogel.” Polymer 41: 39934004.CrossRefGoogle Scholar
Mosiewicz, K. A., Kolb, L., van der Vlies, A. J., Martino, M. M., Lienemann, P. S., Hubbell, J. A., Ehrbar, M., et al. (2013). “In situ cell manipulation through enzymatic hydrogel photopatterning.” Nature Materials 12: 10711077.CrossRefGoogle ScholarPubMed
Murakami, Y. and Maeda, M. (2005). “DNA-responsive hydrogels that can shrink or swell.” Biomacromolecules 6: 29272929.CrossRefGoogle ScholarPubMed
Nagahara, S. and Matsuda, T. (1996). “Hydrogel formation via hybridization of oligonucleotides derivatized in water-soluble vinyl polymers.” Polymer Gels and Networks 4: 111127.CrossRefGoogle Scholar
Nichol, J. W., Koshy, S. T., Bae, H., Hwang, C. M., Yamanlar, S. and Khademhosseini, A. (2010). “Cell-laden microengineered gelatin methacrylate hydrogels.” Biomaterials 31: 55365544.CrossRefGoogle ScholarPubMed
Patterson, J. and Hubbell, J. A. (2010). “Enhanced proteolytic degradation of molecularly engineered PEG hydrogels in response to MMP-1 and MMP-2.” Biomaterials 31: 78367845.CrossRefGoogle ScholarPubMed
Phelps, E. A., Enemchukwu, N. O., Fiore, V. F., Sy, J. C., Murthy, N., Sulchek, T. A., Barker, T. H., et al. (2012). “Maleimide cross-linked bioactive PEG hydrogel exhibits improved reaction kinetics and cross-linking for cell encapsulation and in situ delivery.” Advanced Materials 24: 6470.CrossRefGoogle ScholarPubMed
Pratt, A. B., Weber, F. E., Schmoekel, H. G., Muller, R. and Hubbell, J. A. (2004). “Synthetic extracellular matrices for in situ tissue engineering.” Biotechnology and Bioengineering 86: 2736.CrossRefGoogle ScholarPubMed
Purcell, B. P., Lobb, D., Charati, M. B., Dorsey, S. M., Wade, R. J., Zellars, K. N., Doviak, H., et al. (2014). “Injectable and bioresponsive hydrogels for on-demand matrix metalloproteinase inhibition.” Nature Materials 13: 653661.CrossRefGoogle ScholarPubMed
Raeber, G. P., Lutolf, M. P. and Hubbell, J. A. (2005). “Molecularly engineered PEG hydrogels: a novel model system for proteolytically mediated cell migration.” Biophysical Journal 89: 13741388.CrossRefGoogle ScholarPubMed
Sahoo, S., Chung, C., Khetan, S. and Burdick, J. A. (2008). “Hydrolytically degradable hyaluronic acid hydrogels with controlled temporal structures.” Biomacromolecules 9: 10881092.CrossRefGoogle ScholarPubMed
Salinas, C. N. and Anseth, K. S. (2008). “The enhancement of chondrogenic differentiation of human mesenchymal stem cells by enzymatically regulated RGD functionalities.” Biomaterials 29: 23702377.CrossRefGoogle ScholarPubMed
Seliktar, D., Zisch, A. H., Lutolf, M. P., Wrana, J. L. and Hubbell, J. A. (2008). “MMP-2 sensitive, VEGF-bearing bioactive hydrogels for promotion of vascular healing.” Journal of Biomedical Materials Research 68A: 704716.CrossRefGoogle Scholar
Shikanov, A., Smith, R. M., Xu, M., Woodruff, T. K. and Shea, L. D. (2011). “Hydrogel network design using multifunctional macromers to coordinate tissue maturation in ovarian follicle culture.” Biomaterials 32: 25242531.CrossRefGoogle ScholarPubMed
Simmel, F. C. and Dittmer, W. U. (2005). “DNA nanodevices.” Small 1: 284299.CrossRefGoogle ScholarPubMed
Sugiura, S., Cha, J. M., Yanagawa, F., Zorlutuna, P., Bae, H. and Khademhosseini, A. (2013). “Dynamic three-dimensional micropatterned cell co-cultures within photocurable and chemically degradable hydrogels.” Journal of Tissue Engineering and Regenerative Medicine: n/a-n/a.Google Scholar
Sun, J., Xiao, W. Q., Tang, Y. J., Li, K. F. and Fan, H. S. (2012). “Biomimetic interpenetrating polymer network hydrogels based on methacrylated alginate and collagen for 3D pre-osteoblast spreading and osteogenic differentiation.” Soft Matter 8: 23982404.CrossRefGoogle Scholar
Tauro, J. R. and Gemeinhart, R. A. (2005). “Matrix metalloprotease triggered delivery of cancer chemotherapeutics from hydrogel matrixes.” Bioconjugate Chemistry 16: 11331139.CrossRefGoogle ScholarPubMed
Wang, H., Haeger, S. M., Kloxin, A. M., Leinwand, L. A. and Anseth, K. S. (2012). “Redirecting valvular myofibroblasts into dormant fibroblasts through light-mediated reduction in substrate modulus.” PloS One 7(7): e39969.CrossRefGoogle ScholarPubMed
Wang, N., Butler, J. P. and Ingber, D. E. (1993). “Mechanotrunsduction across the cell surface and through the cytoskeleton.” Science 260: 11241127.CrossRefGoogle ScholarPubMed
Watson, K. J., Park, S. J., Im, J. H., Nguyen, S. T. and Mirkin, C. A. (2001). “DNA-block copolymer conjugates.” Journal of the American Chemical Society 123: 55925593.CrossRefGoogle ScholarPubMed
Wosnick, J. H. and Shoichet, M. S. (2008). “Three-dimensional chemical patterning of transparent hydrogels.” Chemistry of Materials 20: 5560.CrossRefGoogle Scholar
Wylie, R. G., Ahsan, S., Aizawa, Y., Maxwell, K. L., Morshead, C. M. and Shoichet, M. S. (2011). “Spatially controlled simultaneous patterning of multiple growth factors in three-dimensional hydrogels.” Nature Materials 10: 799806.CrossRefGoogle ScholarPubMed
Xing, Y. Z., Cheng, E. J., Yang, Y., Chen, P., Zhang, T., Sun, Y. W., Yang, Z. Q., et al. (2011). “Self-assembled DNA hydrogels with designable thermal and enzymatic responsiveness.” Advanced Materials 23: 11171121.CrossRefGoogle ScholarPubMed
Yamato, M. and Okano, T. (2004). “Cell sheet engineering.” Materials Today 7: 4247.CrossRefGoogle Scholar
Yoshikawa, H. Y., Rossetti, F. F., Kaufmann, S., Kaindl, T., Madsen, J., Engel, U., Lewis, A. L., et al. (2011). “Quantitative evaluation of mechanosensing of cells on dynamically tunable hydrogels.” Journal of the American Chemical Society 133: 13671374.CrossRefGoogle ScholarPubMed
Yurke, B., Turberfield, A. J., Mills, A. P., Simmel, F. C. and Neumann, J. L. (2000). “A DNA-fuelled molecular machine made of DNA.” Nature 406: 605608.CrossRefGoogle ScholarPubMed
Zhang, R., et al. (2013a). “A thermoresponsive and chemically defined hydrogel for long-term culture of human embryonic stem cells.” Nature Communications 4.Google ScholarPubMed
Zhang, W. J., et al. (2013b). “The synergistic effect of hierarchical micro/nano-topography and bioactive ions for enhanced osseointegration.” Biomaterials 34: 31843195.CrossRefGoogle ScholarPubMed
Zisch, A. H., Lutolf, M. P., Ehrbar, M., Raeber, G. P., Rizzi, S. C., Davies, N., Schmokel, H., et al. (2003). “Cell-demanded release of VEGF from synthetic, biointeractive cell-ingrowth matrices for vascularized tissue growth.” FASEB Journal 17: 22602262.CrossRefGoogle ScholarPubMed

References

Abhyankar, V. V., Toepke, M. W., Cortesio, C. L. Lokuta, M. A., Huttenlocher, A. and Beebe, D. J. (2008). “A platform for assessing chemotactic migration within a spatiotemporally defined 3D microenvironment.” Lab on a Chip 8: 15071515.CrossRefGoogle ScholarPubMed
Amadi, O., Steinhauser, M., Nishi, Y., Chung, S., Kamm, R., Mcmahon, A. and Lee, R. (2010). “A low resistance microfluidic system for the creation of stable concentration gradients in a defined 3D microenvironment.” Biomedical Microdevices 12: 10271041.CrossRefGoogle Scholar
Baggiolini, M. (1998). “Chemokines and leukocyte traffic.” Nature 392: 565568.CrossRefGoogle ScholarPubMed
Boyden, S. (2008). “The chemotactic effect of mixtures of antibody and antigen on polymorphonuclear leucocytes.” J Exp Med 115: 453466.CrossRefGoogle Scholar
Campbell, J. and Butcher, E. (2000). “Chemokines in tissue-specific and microenvironment-specific lymphocyte homing.” Curr Opin Immunol 12: 336–41.CrossRefGoogle ScholarPubMed
Cortese, B., Palamà, I. E., D’amone, S. and Gigli, G. (2014). “Influence of electrotaxis on cell behaviour.” Integrative Biology 6: 817830.CrossRefGoogle ScholarPubMed
Delong, E. F., Frankel, R. B. and Bazylinski, D. A. (1993). “Multiple evolutionary origins of magnetotaxis in bacteria.” Science 259: 803.CrossRefGoogle ScholarPubMed
Djamgoz, Mba, Mycielska, M., Madeja, Z., Fraser, S. and Korohoda, W. (2001). “Directional movement of rat prostate cancer cells in direct-current electric field: involvement of voltagegated Na+ channel activity.” J Cell Sci 114: 26972705.CrossRefGoogle Scholar
Gamboa, O. L. Pu, J., Townend, J., Forrester, J. V., Zhao, M., McCaig, C. and Lois, N. (2010). “Electrical estimulation of retinal pigment epithelial cells.” Exp Eye Res 91: 195204.CrossRefGoogle ScholarPubMed
Guo, A., Song, B., Reid, B., Gu, Y., Forrester, J. V., Jahoda, C. A. B. and Zhao, M. (2010). “Effects of physiological electric fields on migration of human dermal fibroblasts.” Journal of Investigative Dermatology 130: 23202327.CrossRefGoogle ScholarPubMed
Haessler, U., Kalinin, Y., Swartz, M. A. and Wu, M. (2009). “An agarose-based microfluidic platform with a gradient buffer for 3D chemotaxis studies.” Biomed Microdevices 11: 827835.CrossRefGoogle ScholarPubMed
Han, B., Kim, D., Hyun Ko, U. and Shin, J. H. (2012). “A sorting strategy for C. elegans based on size-dependent motility and electrotaxis in a micro-structured channel.” Lab on a Chip 12: 41284134.CrossRefGoogle Scholar
Hsu, S., Thakar, R., Liepmann, D. and Li, S. (2005). “Effects of shear stress on endothelial cell haptotaxis on micropatterned surfaces.” Biochemical and Biophysical Research Communications 337: 401409.CrossRefGoogle ScholarPubMed
Huang, C., Cheng, J., Yen, M. and Young, T. (2009). “Electrotaxis of lung cancer cells in a multiple-electric-field chip.” Biosens Bioelectron 24: 35103516.CrossRefGoogle Scholar
Huang, Y.-J., Samorajski, J., Kreimer, R. and Searson, P. C. (2013). “The influence of electric field and confinement on cell motility.” PLoS One 8: e59447.CrossRefGoogle ScholarPubMed
Li, J., et al. (2012a). “Microfluidic device for studying cell migration in single or co-existing chemical gradients and electric fields.” Biomicrofluidics 6: 024121.CrossRefGoogle ScholarPubMed
Li, L., et al. (2012b). “E-cadherin plays an essential role in collective directional migration of large epithelial sheets.” Cellular and Molecular Life Sciences 69: 27792789.CrossRefGoogle ScholarPubMed
Li, J., and Lin, F. (2011). “Microfluidic devices for studying chemotaxis and electrotaxis.” Trends in Cell Biology 21: 489497.CrossRefGoogle ScholarPubMed
Li, J., Nandagopal, S., Wu, D., Romanuik, S. F., Paul, K., Thomson, D. J. and Lin, F. (2011). “Activated T lymphocytes migrate toward the cathode of DC electric fields in microfluidic devices.” Lab on a Chip 11: 12981304.CrossRefGoogle ScholarPubMed
Lin, F., Baldessari, F., Gyenge, C. C., Sato, T., Chambers, R. D., Santiago, J. G. and Butcher, E. C. (2008). “Lymphocyte electrotaxis in vitro and in vivo.” The Journal of Immunology 181: 24652471.CrossRefGoogle ScholarPubMed
Lohof, A., Quillan, M., Dan, Y. and Poo, M. (1992). “Asymmetric modulation of cytosolic cAMP activity induces growth cone turning.” J. Neurosci 12: 12531261.CrossRefGoogle ScholarPubMed
Luster, A., Alon, R. and von Andrian, U. (2005). “Immune cell migration in inflammation: present and future therapeutic targets.” Nat Immunol 6: 1182–90.CrossRefGoogle ScholarPubMed
Manière, X., Lebois, F., Matic, I., Ladoux, B., Di Meglio, J.-M. and Hersen, P. (2011). “Running worms: C. elegans self-sorting by electrotaxis.” PLoS One 6: e16637.CrossRefGoogle Scholar
McCaig, C. D., Allan, D. W., Erskine, L. Rajnicek, A. M. and Stewart, R. (1994). “Growing nerves in an electric field.” Neuroprotocols 4: 134141.CrossRefGoogle Scholar
McCaig, C. D., Rajnicek, A., Song, B. and Zhao, M. (2005). “Controlling cell behavior electrically: current views and future potential.” Physiol Rev 85: 943978.CrossRefGoogle ScholarPubMed
Mccormick, K. E., Gaertner, B. E., Sottile, M., Phillips, P. C. and Lockery, S. R. (2011). “Microfluidic devices for analysis of spatial orientation behaviors in semi-restrained Caenorhabditis elegans.” PLoS One 6: e25710.CrossRefGoogle ScholarPubMed
Minc, N. and Chang, F. (2010). “Electrical control of cell polarization in the fission yeast Schizosaccharomyces pombe.” Current Biology 20: 710716.CrossRefGoogle ScholarPubMed
Mosadegh, B., Saadi, W., Wang, S. J. and Jeon, N. L. (2008). “Epidermal growth factor promotes breast cancer cell chemotaxis in CXCL12 gradients.” Biotechnol Bioeng 100: 12051213.CrossRefGoogle ScholarPubMed
Muller, A., Homey, B., Soto, H., Ge, N., Catron, D., Buchanan, M. E., Mcclanahan, T., et al. (2011). “Involvement of chemokine receptors in breast cancer metastasis.” Nature 410: 5056.CrossRefGoogle Scholar
Mycielska, M. and Djamgoz, M. (2004). “Cellular mechanisms of direct-current electric field effects: galvanotaxis and metastatic disease.” J Cell Sci 117: 16311639.CrossRefGoogle ScholarPubMed
Nelson, R. D., Quie, P. G. and Simmons, R. L. (1975). “Chemotaxis under agarose: a new and simple method for measuring chemotaxis and spontaneous migration of human polymorphonuclear leukocytes and monocytes.” J Immunol 115: 16501656.CrossRefGoogle ScholarPubMed
Pu, J., McCaig, C. D., Cao, L. Zhao, Z., Segall, J. E. and Zhao, M. (2007). “EGF receptor signalling is essential for electric-field-directed migration of breast cancer cells.” J Cell Sci 120: 33953403.CrossRefGoogle ScholarPubMed
Rezai, P., Salam, S., Selvaganapathy, P. R. and Gupta, B. P. (2011). “Effect of pulse direct current signals on electrotactic movement of nematodes Caenorhabditis elegans and Caenorhabditis briggsae.” Biomicrofluidics 5: 44116441169.CrossRefGoogle ScholarPubMed
Rezai, P., Salam, S., Selvaganapathy, P. R. and Gupta, B. P. (2012). “Electrical sorting of Caenorhabditis elegans.” Lab on a Chip 12: 18311840.CrossRefGoogle ScholarPubMed
Rezai, P., Siddiqui, A., Selvaganapathy, P. R. and Gupta, B. (2010a). “Electrotaxis of Caenorhabditis elegans in a microfluidic environment.” Lab Chip 10: 220226.CrossRefGoogle Scholar
Rezai, P., Siddiqui, A., Selvaganapathy, P. R. and Gupta, B. P. (2010b). “Behavior of Caenorhabditis elegans in alternating electric field and its application to their localization and control.” Applied Physics Letters 96: 153702153703.CrossRefGoogle Scholar
Saadi, W., Rhee, S. W., Lin, F., Vahidi, B., Chung, B. G. and Jeon, N. L. (2007). “Generation of stable concentration gradients in 2D and 3D environments using a microfluidic ladder chamber.” Biomed Microdevices 9: 627635.CrossRefGoogle ScholarPubMed
Sato, M. J., Ueda, M., Takagi, H., Watanabe, T. M. and Yanagida, T. (2007). “Input-output relationship in galvanotactic response of Dictyostelium cells.” Biosystems 88: 261272.CrossRefGoogle ScholarPubMed
Song, B., Gu, Y., Pu, J., Reid, B., Zhao, Z. and Zhao, M. (2007). “Application of direct current electric fields to cells and tissues in vitro and modulation of wound electric field in vivo.” Nat Protoc 2: 14791489.CrossRefGoogle ScholarPubMed
Song, S., Han, H., Ko, U. H., Kim, J. and Shin, J. H. (2013). “Collaborative effects of electric field and fluid shear stress on fibroblast migration.” Lab on a Chip 13(8): 16021611.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 3D epithelial coculture.” The FASEB Journal 23: 21552164.CrossRefGoogle ScholarPubMed
Sun, Y.–S., et al. (2012a). “In vitro electrical-stimulated wound-healing chip for studying electric field-assisted wound-healing process.” Biomicrofluidics 6: 034117.CrossRefGoogle ScholarPubMed
Sun, Y. S., et al. (2012b). “Electrotaxis of lung cancer cells in ordered three-dimensional scaffolds.” Biomicrofluidics 6: 1410214114.CrossRefGoogle ScholarPubMed
Tai, G., Reid, B., Cao, L. and Zhao, M. (2009). “Electrotaxis and wound healing: experimental methods to study electric fields as a directional signal for cell migration.” Methods Mol Biol 571: 7797.CrossRefGoogle Scholar
Tsai, H.-F., Huang, C.-W., Chang, H.-F., Chen, J. J., Lee, C.-H. and Cheng, J.-Y. (2013). “Evaluation of EGFR and RTK signaling in the electrotaxis of lung adenocarcinoma cells under direct-current electric field stimulation.” PLoS One 8: e73418.CrossRefGoogle ScholarPubMed
Wadhawan, N., Kalkat, H., Natarajan, K., Ma, X., Gajjeraman, S., Nandagopal, S., Hao, N., et al. (2012). “Growth and positioning of adipose-derived stem cells in microfluidic devices.” Lab on a Chip 12: 48294834.CrossRefGoogle ScholarPubMed
Wang, C.-C., Kao, Y.-C., Chi, P.-Y., Huang, C.-W., Lin, J.-Y., Chou, C.-F., Cheng, J.-Y., et al. (2011). “Asymmetric cancer-cell filopodium growth induced by electric-fields in a microfluidic culture chip.” Lab on a Chip 11: 695699.CrossRefGoogle Scholar
Wang, S.-J., Saadi, W., Lin, F., Minh-Canh Nguyen, C. and Li Jeon, N. (2004). “Differential effects of EGF gradient profiles on MDA-MB-231 breast cancer cell chemotaxis.” Experimental Cell Research, 300, 180189.CrossRefGoogle ScholarPubMed
Whitesides, G. M. (2006). “The origins and the future of microfluidics.” Nature 442: 368373.CrossRefGoogle ScholarPubMed
Wu, D., et al. (2013a). “DC electric fields direct breast cancer cell migration, induce EGFR polarization, and increase the intracellular level of calcium ions.” Cell Biochemistry and Biophysics 67: 11151125.CrossRefGoogle ScholarPubMed
Wu, D. and Lin, F. (2011). “A receptor-electromigration-based model for cellular electrotactic sensing and migration.” Biochemical and Biophysical Research Communications 411: 695701.CrossRefGoogle ScholarPubMed
Wu, J., et al. (2013b). “Recent developments in microfluidics-based chemotaxis studies.” Lab on a Chip 13: 24842499.CrossRefGoogle ScholarPubMed
Wu, J. and Lin, F. (2014). “Recent developments in electrotaxis assays.” Advances in Wound Care 3: 149155.CrossRefGoogle ScholarPubMed
Zervantonakis, I., Chung, S., Sudo, R., Zhang, M., Charest, J. and Kamm, R. (2010). “Concentration gradients in microfluidic 3D matrix cell culture systems.” International Journal of Micro-Nano Scale Transport 1: 2736.CrossRefGoogle Scholar
Zhao, M. (2009). “Electrical fields in wound healing-An overriding signal that directs cell migration.” Semin Cell Dev Biol 20: 674682.CrossRefGoogle ScholarPubMed
Zhao, M., Agius-Fernandez, A., Forrester, J. and McCaig, C. (1996). “Directed migration of corneal epithelial sheets in physiological electric fields.” Invest Ophthalmol Vis Sci 37: 25482558.Google ScholarPubMed
Zhao, M., Dick, A., Forrester, J. and McCaig, C. (1999). “Electric field-directed cell motility involves up-regulated expression and asymmetric redistribution of the epidermal growth factor receptors and is enhanced by fibronectin and laminin.” Mol Biol Cell 10: 12591276.CrossRefGoogle ScholarPubMed
Zhao, M., Song, B., Pu, J., Wada, T., Reid, B., Tai, G., Wang, F., et al. (2006). “Electrical signals control wound healing through phosphatidylinositol-3-OH kinase-gamma and PTEN.” Nature 442: 457460.CrossRefGoogle ScholarPubMed
Zhao, S., Gao, R., Devreotes, P. N., Mogilner, A. and Zhao, M. (2013). “3D arrays for high throughput assay of cell migration and electrotaxis.” Cell biology international 37: 9951002.CrossRefGoogle ScholarPubMed
Zigmond, S. 1977. “Ability of polymorphonuclear leukocytes to orient in gradients of chemotactic factors.” J Cell Biol 75: 606616.CrossRefGoogle ScholarPubMed

References

Aist, J. R. and Berns, M. W. (1981). “Mechanics of chromosome separation during mitosis in Fusarium (Fungi imperfecti): new evidence from ultrastructural and laser microbeam experiments.” J Cell Biol 91: 446458.CrossRefGoogle ScholarPubMed
Aist, J. R., Liang, H. and Berns, M. W. (1993). “Astral and spindle forces in PtK2 cells during anaphase B: a laser microbeam study.” J Cell Sci 104(4): 12071216.CrossRefGoogle Scholar
Aliee, M., Roper, J. C., Landsberg, K. P., Pentzold, C., Widmann, T. J., Julicher, F. and Dahmann, C. (2012). “Physical mechanisms shaping the Drosophila dorsoventral compartment boundary.” Curr Biol 22: 967976.CrossRefGoogle ScholarPubMed
Bambardekar, K., Clement, R., Blanc, O., Chardes, C. and Lenne, P. F. (2015). “Direct laser manipulation reveals the mechanics of cell contacts in vivo.” Proc Natl Acad Sci USA 112(5): 14161421.CrossRefGoogle ScholarPubMed
Behrndt, M., Salbreux, G., Campinho, P., Hauschild, R., Oswald, F., Roensch, J., Grill, S. W., et al. (2012). “Forces driving epithelial spreading in zebrafish gastrulation.” Science 338: 257260.CrossRefGoogle ScholarPubMed
Bertet, C., Sulak, L. and Lecuit, T. (2004). “Myosin-dependent junction remodelling controls planar cell intercalation and axis elongation.” Nature 10: 667671.CrossRefGoogle Scholar
Blanchard, G. B., Murugesu, S., Adams, R. J., Martinez-Arias, A. and Gorfinkiel, N. (2010). “Cytoskeletal dynamics and supracellular organisation of cell shape fluctuations during dorsal closure.” Development 137: 27432752.CrossRefGoogle ScholarPubMed
Blankenship, J. T., Backovic, S. T., Sanny, J. S. P., Weitz, O. and Zallen, J. A. (2006). “Multicellular rosette formation links planar cell polarity to tissue morphogenesis.” Dev Cell 11: 459470.CrossRefGoogle ScholarPubMed
Bonnet, I., Marcq, P., Bosveld, F., Fetler, L., Bellaiche, Y. and Graner, F. (2012). “Mechanical state, material properties and continuous description of an epithelial tissue.” J R Soc Interface 9: 26142623.CrossRefGoogle ScholarPubMed
Brugues, J., Nuzzo, V., Mazur, E. and Needleman, D. J. (2012). “Nucleation and transport organize microtubules in metaphase spindles.” Cell 149: 554564.CrossRefGoogle ScholarPubMed
Campas, O., Mammoto, T., Hasso, S., Sperling, R. A., O’Connell, D., Bischof, A. G., Maas, R., et al. (2014). “Quantifying cell-generated mechanical forces within living embryonic tissues.” Nat Methods 11: 183189.CrossRefGoogle ScholarPubMed
Campinho, P., Behrndt, M., Ranft, J., Risler, T., Minc, N. and Heisenberg, C. P. (2013). “Tension-oriented cell divisions limit anisotropic tissue tension in epithelial spreading during zebrafish epiboly.” Nat Cell Biol 15: 14051414.CrossRefGoogle ScholarPubMed
Classen, A. K., Anderson, K. I., Marois, E. and Eaton, S. (2005). “Hexagonal packing of Drosophila wing epithelial cells by the planar cell polarity pathway.” Dev Cell 9: 805817.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: 47124719.CrossRefGoogle ScholarPubMed
David, D. J., Tishkina, A. and Harris, T. J. (2010). “The PAR complex regulates pulsed actomyosin contractions during amnioserosa apical constriction in Drosophila.” Development 137: 16451655.CrossRefGoogle ScholarPubMed
del Alamo, J. C., Norwich, G. N., Li, Y. S., Lasheras, J. C. and Chien, S. (2008). “Anisotropic rheology and directional mechanotransduction in vascular endothelial cells.” Proc Natl Acad Sci USA 105: 1541115416.CrossRefGoogle ScholarPubMed
Farhadifar, R., Röper, J.-C., Aigouy, B., Eaton, S. and Jülicher, F. (2007). “The influence of cell mechanics, cell-cell interactions, and proliferation on epithelial packing.” Curr Biol 17(24): 20952104.CrossRefGoogle ScholarPubMed
Fernandez-Gonzalez, R., Simoes Sde, M., Roper, J. C., Eaton, S. and Zallen, J. A. (2009). “Myosin II dynamics are regulated by tension in intercalating cells.” Dev Cell 17: 736743.CrossRefGoogle ScholarPubMed
Fernandez-Gonzalez, R. and Zallen, J. A. (2013). “Wounded cells drive rapid epidermal repair in the early Drosophila embryo.” Mol Biol Cell 24: 32273237.CrossRefGoogle ScholarPubMed
Fischer, S. C., Blanchard, G. B., Duque, J., Adams, R. J., Arias, A. M., Guest, S. D. and Gorfinkiel, N. (2014). “Contractile and mechanical properties of epithelia with perturbed actomyosin dynamics.” PLoS One 9: e95695.CrossRefGoogle ScholarPubMed
Forgacs, G., Foty, R. A., Shafrir, Y. and Steinberg, M. S. (1998). “Viscoelastic properties of living embryonic tissues: a quantitative study.” Biophys J 74: 22272234.CrossRefGoogle ScholarPubMed
Grill, S. W., Gonczy, P., Stelzer, E. H. and Hyman, A. A. (2001). “Polarity controls forces governing asymmetric spindle positioning in the Caenorhabditis elegans embryo.” Nature 409: 630633.CrossRefGoogle ScholarPubMed
Grill, S. W., Howard, J., Schaffer, E., Stelzer, E. H. and Hyman, A. A. (2003). “The distribution of active force generators controls mitotic spindle position.” Science 301: 518521.CrossRefGoogle ScholarPubMed
Heisterkamp, A., Maxwell, I. Z., Mazur, E., Underwood, J. M., Nickerson, J. A., Kumar, S. and Ingber, D. E. (2005). “Pulse energy dependence of subcellular dissection by femtosecond laser pulses.” Opt Express 13: 36903696.CrossRefGoogle ScholarPubMed
Heller, E., Kumar, K. V., Grill, S. W. and Fuchs, E. (2014). “Forces generated by cell intercalation tow epidermal sheets in mammalian tissue morphogenesis.” Dev Cell 28: 617632.CrossRefGoogle ScholarPubMed
Herszterg, S., Leibfried, A., Bosveld, F., Martin, C. and Bellaiche, Y. (2013). “Interplay between the dividing cell and its neighbors regulates adherens junction formation during cytokinesis in epithelial tissue.” Dev Cell 24: 256270.CrossRefGoogle ScholarPubMed
Hunter, G. L., Crawford, J. M., Genkins, J. Z. and Kiehart, D. P. (2014). “Ion channels contribute to the regulation of cell sheet forces during Drosophila dorsal closure.” Development 141: 325334.CrossRefGoogle Scholar
Hutson, M. S. and Ma, X. (2007). “Plasma and cavitation dynamics during pulsed laser microsurgery in vivo.” Phys Rev Lett 99: 158104.CrossRefGoogle ScholarPubMed
Hutson, M. S., Tokutake, Y., Chang, M.-S., Bloor, J. W., Venakides, S., Kiehart, D. P. and Edwards, G. S. (2003). “Forces for morphogenesis investigated with laser microsurgery and quantitative modeling.” Science 300: 145149.CrossRefGoogle ScholarPubMed
Irvine, K. D. and Wieschaus, E. (1994). “Cell intercalation during Drosophila germband extension and its regulation by pair-rule segmentation genes.” Development 120: 827841.CrossRefGoogle ScholarPubMed
Kasza, K. E., Farrell, D. L. and Zallen, J. A. (2014). “Spatiotemporal control of epithelial remodeling by regulated myosin phosphorylation.” Proc Natl Acad Sci USA 111: 1173211737.CrossRefGoogle ScholarPubMed
Kiehart, D. P., Galbraith, C. G., Edwards, K. A., Rickoll, W. L. and Montague, R. A. (2000). “Multiple forces contribute to cell sheet morphogenesisfor dorsal closure in Drosophila.” J Cell Biol 149: 471490.CrossRefGoogle ScholarPubMed
Kumar, S., Maxwell, I. Z., Heisterkamp, A., Polte, T. R., Lele, T. P., Salanga, M., Mazur, E., et al. (2006). “Viscoelastic retraction of single living stress fibers and its impact on cell shape, cytoskeletal organization, and extracellular matrix mechanics.” Biophys J 90: 37623773.CrossRefGoogle ScholarPubMed
Landsberg, K. P., Farhadifar, R., Ranft, J., Umetsu, D., Widmann, T. J., Bittig, T., Said, A., et al. (2009). “Increased cell bond tension governs cell sorting at the Drosophila anteroposterior compartment boundary.” Curr Biol 19: 19501955.CrossRefGoogle ScholarPubMed
Lau, K., Tao, H., Liu, H., Wen, J., Sturgeon, K., Sorfazlian, N., Lazic, S., et al. (2015). “Anisotropic stress orients remodelling of mammalian limb bud ectoderm.” Nat Cell Biol 17: 569579.CrossRefGoogle ScholarPubMed
Ma, X., Lynch, H. E., Scully, P. C. and Hutson, M. S. (2009). “Probing embryonic tissue mechanics with laser hole drilling.” Phys Biol 6: 036004.CrossRefGoogle ScholarPubMed
Mainardi, F. and Spada, G. (2011). “Creep, relaxation and viscosity properties for basic fractional models in rheology.” EPJ-Special Topics 193: 133160.CrossRefGoogle Scholar
Major, R. J. and Irvine, K. D. (2005). “Influence of Notch on dorsoventral compartmentalization and actin organization in the Drosophila wing.” Development 132: 38233833.CrossRefGoogle ScholarPubMed
Major, R. J. and Irvine, K. D. (2006). “Localization and requirement for Myosin II at the dorsal-ventral compartment boundary of the Drosophila wing.” Dev Dyn 235: 30513058.CrossRefGoogle ScholarPubMed
Marinari, E., Mehonic, A., Curran, S., Gale, J., Duke, T. and Baum, B. (2012). “Live-cell delamination counterbalances epithelial growth to limit tissue overcrowding.” Nature 484(7395): 542545.CrossRefGoogle ScholarPubMed
Martin, A. C., Gelbart, M., Fernandez-Gonzalez, R., Kaschube, M. and Wieschaus, E. F. (2010). “Integration of contractile forces during tissue invagination.” J Cell Biol 188: 735749.CrossRefGoogle ScholarPubMed
Martin, A. C., Kaschube, M. and Wieschaus, E. F. (2009). “Pulsed contractions of an actin-myosin network drive apical constriction.” Nature 457: 495499.CrossRefGoogle ScholarPubMed
Martin, P. and Lewis, J. (1992). “Actin cables and epidermal movement in embryonic wound healing.” Nature 360: 179183.CrossRefGoogle ScholarPubMed
Mayer, M., Depken, M., Bois, J. S., Julicher, F. and Grill, S. W. (2010). “Anisotropies in cortical tension reveal the physical basis of polarizing cortical flows.” Nature 467: 617621.CrossRefGoogle ScholarPubMed
McNeill, H. (2000). “Sticking together and sorting things out: adhesion as a force in development.” Nat Rev Genet 1: 100108.CrossRefGoogle ScholarPubMed
Niemz, M. H. (2004). Laser-Tissue Interactions: Fundamentals of Microscopy. Berlin: Springer.Google Scholar
Purcell, E. M. (1977). “Life at low Reynolds number.” Am J Phys 45: 311.CrossRefGoogle Scholar
Quinto-Su, P. A. and Venugopalan, V. (2007). “Mechanisms of laser cellular microsurgery.” Methods Cell Biol 82: 113151.Google ScholarPubMed
Raffel, M., Willert, C. E. and Kompenhans, J. (1998). Particle Image Velocimetry: a Practical Guide. Berlin and New York: Springer.CrossRefGoogle Scholar
Rauzi, M. and Lenne, P. F. (2011). “Cortical forces in cell shape changes and tissue morphogenesis.” Curr Top Dev Biol 95: 93144.CrossRefGoogle ScholarPubMed
Rauzi, M., Verant, P., Lecuit, T. and Lenne, P. F. (2008). “Nature and anisotropy of cortical forces orienting Drosophila tissue morphogenesis.” Nat Cell Biol 10: 14011410.CrossRefGoogle ScholarPubMed
Schaffer, C. B., Brodeur, A. and Mazur, E. (2001). “Laser-induced breakdown and damage in bulk transparent materials induced by tightly focused femtosecond laser pulses.” Meas Sci Technol 12: 17841794.CrossRefGoogle Scholar
Schotz, E. M., Burdine, R. D., Julicher, F., Steinberg, M. S., Heisenberg, C. P. and Foty, R. A. (2008). “Quantitative differences in tissue surface tension influence zebrafish germ layer positioning.” HFSP J 2: 4256.CrossRefGoogle ScholarPubMed
Shindo, A. and Wallingford, J. B. (2014). “PCP and septins compartmentalize cortical actomyosin to direct collective cell movement.” Science 343: 649652.CrossRefGoogle ScholarPubMed
Simoes, S., Mainieri, A. and Zallen, J. A. (2014). “Rho GTPase and Shroom direct planar polarized actomyosin contractility during convergent extension.” J Cell Biol 204: 575589.CrossRefGoogle ScholarPubMed
Sokolow, A., Toyama, Y., Kiehart, D. P. and Edwards, G. S. (2012). “Cell ingression and apical shape oscillations during dorsal closure in Drosophila.” Biophys J 102: 969979.CrossRefGoogle ScholarPubMed
Solon, J., Kaya-Copur, A., Colombelli, J. and Brunner, D. (2009). “Pulsed forces timed by a ratchet-like mechanism drive directed tissue movement during dorsal closure.” Cell 137: 13311342.CrossRefGoogle ScholarPubMed
Stephens, R. E. (1965). “Analysis of muscle contraction by ultraviolet microbeam disruption of sarcomere structure.” J Cell Biol 25: 129139.CrossRefGoogle ScholarPubMed
Tepass, U., Godt, D. and Winklbauer, R. (2002). “Cell sorting in animal development: signalling and adhesive mechanisms in the formation of tissue boundaries.” Curr Opin Genet Dev 12: 572582.CrossRefGoogle ScholarPubMed
Tseng, Y., Kole, T. P. and Wirtz, D. (2002). “Micromechanical mapping of live cells by multiple-particle-tracking microrheology.” Biophys J 83(6): 31623176:CrossRefGoogle ScholarPubMed
Umetsu, D., Aigouy, B., Aliee, M., Sui, L., Eaton, S., Julicher, F. and Dahmann, C. (2014). “Local increases in mechanical tension shape compartment boundaries by biasing cell intercalations.” Curr Biol 24: 17981805.CrossRefGoogle ScholarPubMed
Vogel, A. and Venugopalan, V. (2003). “Mechanisms of pulsed laser ablation of biological tissues.” Chem Rev 103: 577644.CrossRefGoogle ScholarPubMed
Zallen, J. A. and Wieschaus, E. (2004). “Patterned gene expression directs bipolar planar polarity in Drosophila.” Dev Cell 6: 343355.CrossRefGoogle ScholarPubMed

References

Bar-Shalom, Y., Li, X. R. and Kirubarajan, T. (2001). Estimation with Applications to Tracking and Navigation. New York: Wiley-Interscience.Google Scholar
Barron, J. L., Fleet, D. J. and Beauchemin, S. S. (1994). “Performance of optical flow techniques.” Int J Comp Vis 12: 4377.CrossRefGoogle Scholar
Betzig, E., Patterson, G. H., Sougrat, R., Lindwasser, O. W., Olenych, S., Bonifacino, J. S., Davidson, M. W., et al. (2006). “Imaging intracellular fluorescent proteins at nanometer resolution.” Science 313: 16421645.CrossRefGoogle ScholarPubMed
Blackman, S. and Popoli, R. (1999). Design and Analysis of Modern Tracking Systems. Norwood, MA: Artech House.Google Scholar
Born, M. and Wolf, E. (1999). Principles of Optics. Cambridge University Press.CrossRefGoogle Scholar
Burkard, R., Dell’amico, M. and Martello, S. (2009). Assignment Problems. Philadelphia: Society for Industrial and Applied Mathematics.CrossRefGoogle Scholar
Cheezum, M. K., Walker, W. F. and Guilford, W. H. (2001). “Quantitative comparison of algorithms for tracking single fluorescent particles.” Biophys J 81: 23782388.CrossRefGoogle ScholarPubMed
Chen, K. C., Yang, G. and Kovacevic, J. (2014). “Spatial density estimation based segmentation of super-resolution localization microscopy images.” Proc 2014 IEEE Int Conf Image Proc (ICIP): 867871.Google Scholar
Chen, K. C. J., Yiyi, Y., Ruiqin, L., Hao-Chih, L., Ge, Y. and Kovacevic, J. (2012). “Adaptive active-mask image segmentation for quantitative characterization of mitochondrial morphology.” Proc 2012 IEEE Int Conf Image Proc (ICIP): 20332036.Google Scholar
Chenouard, N., Smal, I., de Chaumont, F., Maska, M., Sbalzarini, I. F., Gong, Y., Cardinale, J., et al. (2014). “Objective comparison of particle tracking methods.” Nat Meth 11: 281289.CrossRefGoogle ScholarPubMed
Cox, I. (1993). “A review of statistical data association techniques for motion correspondence.” Int J Comp Vis 10: 5366.CrossRefGoogle Scholar
Crocker, J. C. and Hoffman, B. D. (2007). “Multiple-particle tracking and two-point microrheology in cells.” Meth in Cell Biol 83: 141178.CrossRefGoogle ScholarPubMed
Crum, W. R., Hartkens, T. and Hill, D. L. G. (2004). “Non-rigid image registration: theory and practice.” British J Radiology 77: S140S153.CrossRefGoogle ScholarPubMed
Danuser, G. and Waterman-Storer, C. M. (2006). “Quantitative fluorescent speckle microscopy of cytoskeleton dynamics.” Annu Rev Biophys Biomol Struct 35: 361387.CrossRefGoogle ScholarPubMed
Das, R., Cairo, C. W. and Coombs, D. (2009). “A hidden Markov model for single particle tracks quantifies dynamic interactions between LFA-1 and the actin cytoskeleton.” PLoS Comp Biol 5: e1000556.CrossRefGoogle ScholarPubMed
De Chaumont, F., Dallongeville, S., Chenouard, N., Herve, N., Pop, S., Provoost, T., Meas-Yedid, V., et al. (2012). “Icy: an open bioimage informatics platform for extended reproducible research.” Nat Meth 9: 690696.CrossRefGoogle ScholarPubMed
De Vlaminck, I. and Dekker, C. (2012). “Recent advances in magnetic tweezers.” Annu Rev Biophys 41: 453472.CrossRefGoogle ScholarPubMed
Dhawan, A. P. (2011). Medical Image Analysis. New York: Wiley-IEEE Press.CrossRefGoogle Scholar
Dima, A. A., Elliott, J. T., Filliben, J. J., Halter, M., Peskin, A., Bernal, J., Kociolek, M., et al. (2011). “Comparison of segmentation algorithms for fluorescence microscopy images of cells.” Cytometry Part A 79A: 545559.CrossRefGoogle Scholar
Dorn, J. F., Danuser, G., and Yang, G. (2008). “Computational processing and analysis of dynamic fluorescence image data.” Meth Cell Biol 85: 497538.CrossRefGoogle ScholarPubMed
Eliceiri, K. W., Berthold, M. R., Goldberg, I. G., Ibanez, L., Manjunath, B. S., Martone, M. E., Murphy, R. F., et al. (2012). “Biological imaging software tools.” Nat Meth 9: 697710.CrossRefGoogle ScholarPubMed
Fleet, D. J. and Weiss, Y. (2005). Optical flow estimation. In Paragios, N., Chen, Y. and Faugeras, O. (eds.), Mathematical Models in Computer Vision. New York: Springer.Google Scholar
Gao, Y. and Kilfoil, M. L. (2009). “Accurate detection and complete tracking of large populations of features in three dimensions.” Opt Exp 17: 46854704.CrossRefGoogle ScholarPubMed
Gelles, J., Schnapp, B. J. and Sheetz, M. P. (1988). “Tracking kinesin-driven movements with nanometre-scale precision.” Nature 331: 450453.CrossRefGoogle ScholarPubMed
Genovesio, A., Liedl, T., Emiliani, V., Parak, W. J., Coppey-Moisan, M. and Olivo-Marin, J. C. (2006). “Multiple particle tracking in 3-D+t microscopy: method and application to the tracking of endocytosed quantum dots.” IEEE Trans Image Proc 15: 10621070.CrossRefGoogle Scholar
Giepmans, B. N. G., Adams, S. R., Ellisman, M. H. and Tsien, R. Y. (2006). “The fluorescent toolbox for assessing protein location and function.” Science 312: 217224.CrossRefGoogle ScholarPubMed
Grashoff, C., Hoffman, B. D., Brenner, M. D., Zhou, R., Parsons, M., Yang, M. T., Mclean, M. A., et al. (2010). “Measuring mechanical tension across vinculin reveals regulation of focal adhesion dynamics.” Nature 466: 263266.CrossRefGoogle ScholarPubMed
Heimann, T., van Ginneken, B., Styner, M. A., Arzhaeva, Y., Aurich, V., Bauer, C., Beck, A., et al. (2009). “Comparison and evaluation of methods for liver segmentation from CT datasets.” IEEE Trans Med Imaging 28: 12511265.CrossRefGoogle ScholarPubMed
Herbert, K. M., Greenleaf, W. J. and Block, S. M. (2008). “Single-molecule studies of RNA polymerase: motoring along.” Annu Rev Biochem 77: 149176.CrossRefGoogle ScholarPubMed
Hochmuth, R. M. (2000). “Micropipette aspiration of living cells.” J Biomechanics 33: 1522.CrossRefGoogle ScholarPubMed
Huang, B., Bates, M. and Zhuang, X. (2009). “Super-resolution fluorescence microscopy.” Annu Rev Biochem 78: 9931016.CrossRefGoogle ScholarPubMed
Jaqaman, K., Loerke, D., Mettlen, M., Kuwata, H., Grinstein, S., Schmid, S. L. and Danuser, G. (2008). “Robust single-particle tracking in live-cell time-lapse sequences.” Nat Meth 5: 695702.CrossRefGoogle ScholarPubMed
Ji, L. and Danuser, G. (2005). “Tracking quasi-stationary flow of weak fluorescent signals by adaptive multi-frame correlation.” J Microscopy 220: 150167.CrossRefGoogle ScholarPubMed
Ji, L., Lim, J. and Danuser, G. (2008). “Fluctuations of intracellular forces during cell protrusion.” Nat Cell Biol 10: 13931400.CrossRefGoogle ScholarPubMed
Ji, L., Loerke, D., Gardel, M., Danuser, G. (2007). “Probing intracellular force distributions by high-resolution live cell imaging and inverse dynamics.” Meth Cell Biol 83: 199235.CrossRefGoogle ScholarPubMed
Kraning-Rush, C. M., Carey, S. P., Califano, J. P. and Reinhart-King, C. A. (2012). “Quantifying traction stresses in adherent cells.” Methods in Cell Biology 110: 139178.CrossRefGoogle ScholarPubMed
Legant, W. R., Miller, J. S., Blakely, B. L., Cohen, D. M., Genin, G. M. and Chen, C. S. (2010). “Measurement of mechanical tractions exerted by cells in three-dimensional matrices.” Nat Meth 7: 969971.CrossRefGoogle ScholarPubMed
Lei, Y., Zhen, Q., Greenaway, A. H. and Weiping, L. (2012). “A new framework for particle detection in low-SNR fluorescence live-cell images and its application for improved particle tracking.” IEEE Trans Biomed Eng 59: 20402050.Google Scholar
Li, K., Miller, E. D., Chen, M., Kanade, T., Weiss, L. E. and Campbell, P. G. (2008). “Cell population tracking and lineage construction with spatiotemporal context.” Med Image Analy 12: 546566.CrossRefGoogle ScholarPubMed
Lippincott-Schwartz, J., Snapp, E. and Kenworthy, A. (2001). “Studying protein dynamics in living cells.” Nat Rev Mol Cell Biol 2: 444456.CrossRefGoogle ScholarPubMed
Machacek, M. and Danuser, G. (2006). “Morphodynamic profiling of protrusion phenotypes.” Biophys J 90: 14391452.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
Maska, M., Ulman, V., Svoboda, D., Matula, P., Matula, P., Ederra, C., Urbiola, A., et al. (2014). “A benchmark for comparison of cell tracking algorithms.” Bioinformatics 30: 16091617.CrossRefGoogle ScholarPubMed
Matov, A., Edvall, M. M., Yang, G. and Danuser, G. (2011). “Optimal-flow minimum-cost correspondence assignment in particle flow tracking.” Comput Vis Image Underst 115: 531540.CrossRefGoogle ScholarPubMed
Meijering, E., Dzyubachyk, O. and Smal, I. (2012). “Methods for cell and particle tracking.” Meth Enzymol 504: 183200.CrossRefGoogle ScholarPubMed
Meijering, E., Dzyubachyk, O., Smal, I. and van Cappellen, W. A. (2009). “Tracking in cell and developmental biology.” Semi Cell Dev Biol 20: 894902.CrossRefGoogle ScholarPubMed
Mitchison, T. J. (2005). “Mechanism and function of poleward flux in Xenopus extract meiotic spindles.” PhiloTrans Royal Soc B: BiolSci 360: 623629.CrossRefGoogle ScholarPubMed
Moffitt, J. R., Chemla, Y. R., Smith, S. B. and Bustamante, C. (2008). “Recent advances in optical tweezers.” Annu Rev Biochem 77: 205228.CrossRefGoogle ScholarPubMed
Nixon, M. and Aguado, A. (2012). Feature Extraction and Image Processing for Computer Vision. Waltham, MA: Academic Press.Google Scholar
Padfield, D., Rittscher, J. and Roysam, B. (2011). “Coupled minimum-cost flow cell tracking for high-throughput quantitative analysis.” Med Image Analy 15: 650668.CrossRefGoogle ScholarPubMed
Patterson, G., Davidson, M., Manley, S. and Lippincott-Schwartz, J. (2010). “Superresolution imaging using single-molecule localization.” Annu Rev Phys Chem 61: 345367.CrossRefGoogle ScholarPubMed
Peng, H. (2008). “Bioimage informatics: a new area of engineering biology.” Bioinformatics 24: 18271836.CrossRefGoogle ScholarPubMed
Pham, D. L., Xu, C. and Prince, J. L. (2000). “Current methods in medical image segmentation.” Annu Rev Biomed Eng 2: 315337.CrossRefGoogle ScholarPubMed
Planchon, T. A., Gao, L., Milkie, D. E., Davidson, M. W., Galbraith, J. A., Galbraith, C. G. and Betzig, E. (2011). “Rapid three-dimensional isotropic imaging of living cells using Bessel beam plane illumination.” Nat Meth 8: 417423.CrossRefGoogle ScholarPubMed
Ponti, A., Vallotton, P., Salmon, W. C., Waterman-Storer, C. M. and Danuser, G. (2003). “Computational analysis of F-actin turnover in cortical actin meshworks using fluorescent speckle microscopy.” Biophys J 84: 33363352.CrossRefGoogle ScholarPubMed
Qian, H., Sheetz, M. P. and Elson, E. L. (1991). “Single particle tracking: analysis of diffusion and flow in two-dimensional systems.” Biophys J 60: 910921.CrossRefGoogle ScholarPubMed
Qiu, M., Lee, H.-C. and Yang, G. (2012). “Nanometer resolution tracking and modeling of bidirectional axonal cargo transport.” Proc 2012 IEEE Int Symp Biomed Imaging (ISBI): 992995.Google Scholar
Reis, G. F., Yang, G., Szpankowski, L., Weaver, C., Shah, S. B., Robinson, J. T., Hays, T. S., et al. (2012). “Molecular motor function in axonal transport in vivo probed by genetic and computational analysis in Drosophila.” Mol Biol Cell 23: 17001714.CrossRefGoogle ScholarPubMed
Rust, M. J., Bates, M. and Zhuang, X. (2006). “Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM).” Nat Meth 3: 793795.CrossRefGoogle ScholarPubMed
Saxton, M. J. (2007). “Modeling 2D and 3D diffusion.” Meth Mol Biol 400: 295321.CrossRefGoogle Scholar
Saxton, M. J. and Jacobson, K. (1997). “Single-particle tracking: applications to membrane dynamics.” Annu Rev Biophys Biomol Struct 26: 373399.CrossRefGoogle ScholarPubMed
Schindelin, J., Arganda-Carreras, I., Frise, E., Kaynig, V., Longair, M., Pietzsch, T., Preibisch, S., et al. (2012). “Fiji: an open-source platform for biological-image analysis.” Nat Meth 9: 676682.CrossRefGoogle ScholarPubMed
Selvin, P. and Ha, T. (2008). Single-Molecule Techniques. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Smal, I., Loog, M., Niessen, W. and Meijering, E. (2010). “Quantitative comparison of spot detection methods in fluorescence microscopy.” IEEE Trans Med Img 29: 282301.CrossRefGoogle ScholarPubMed
Smal, I., Niessen, W. and Meijering, E. (2008). “A new detection scheme for multiple object tracking in fluorescence microscopy by joint probabilistic data association filtering.” Proc 2008 IEEE Int Symp Biomed Imaging (ISBI): 264267.Google Scholar
Sonka, M., Hlavac, V. and Boyle, R. (2007). Image Processing, Analysis, and Machine Vision. Toronto: Thomson Learning.Google Scholar
Srinivasa, G., Fickus, M. C., Yusong, G., Linstedt, A. D. and Kovacevic, J. (2009). “Active mask segmentation of fluorescence microscope images.” IEEE Trans Image Proc 18: 18171829.CrossRefGoogle ScholarPubMed
Style, R. W., Boltyanskiy, R., German, G. K., Hyland, C., Macminn, C. W., Mertz, A. F., Wilen, , et al. (2014). “Traction force microscopy in physics and biology.” Soft Mat 10: 40474055.CrossRefGoogle ScholarPubMed
Sun, D., Roth, S. and Black, M. J. (2014). “A quantitative analysis of current practices in optical flow estimation and the principles behind them.” Int J Comp Vis 106: 115137.CrossRefGoogle Scholar
Svoboda, K. and Block, S. M. (1994). “Biological applications of optical forces.” Annu Rev Biophy Biomol Struc 23: 247285.CrossRefGoogle ScholarPubMed
Szeliski, R. (2010). Computer Vision: Algorithms and Applications. New York: Springer.Google Scholar
Thompson, R. E., Larson, D. R. and Webb, W. W. (2002). “Precise nanometer localization analysis for individual fluorescent probes.” Biophys J 82: 27752783.CrossRefGoogle ScholarPubMed
Ulrich, T. A., Jain, A., Tanner, K., Mackay, J. L. and Kumar, S. (2010). “Probing cellular mechanobiology in three-dimensional culture with collagen-agarose matrices.” Biomaterials 31: 18751884.CrossRefGoogle ScholarPubMed
Vaziri, A. and Mofrad, M. R. K. (2007). “Mechanics and deformation of the nucleus in micropipette aspiration experiment.” J Biomechanics 40: 20532062.CrossRefGoogle ScholarPubMed
Veenman, C. J., Reinders, M. J. T. and Backer, E. (2001). “Resolving motion correspondence for densely moving points.” IEEE Trans Patt Analy Mach Intel 23: 5472.CrossRefGoogle Scholar
Wang, N., Hu, S., and Butler, J. P. (2007). “Imaging stress propagation in the cytoplasm of a living cell.” Methods in Cell Biology 83: 179198.CrossRefGoogle ScholarPubMed
Waterman-Storer, C. M. and Salmon, E. D. (1998). “How microtubules get fluorescent speckles.” Biophys J 75: 20592069.CrossRefGoogle ScholarPubMed
Wilson, C. A., Tsuchida, M. A., Allen, G. M., Barnhart, E. L., Applegate, K. T., Yam, P. T., Ji, L., et al. (2010). “Myosin II contributes to cell-scale actin network treadmilling through network disassembly.” Nature 465: 373377.CrossRefGoogle ScholarPubMed
Wirtz, D. (2009). “Particle-tracking microrheology of living cells: principles and applications.” Annu Rev Biophys 38: 301326.CrossRefGoogle Scholar
Yang, G. (2013). “Bioimage informatics for understanding spatiotemporal dynamics of cellular processes.” Wiley Inter Rev Sys Biol Med 5: 367380.CrossRefGoogle ScholarPubMed
Yang, G., Cameron, L. A., Maddox, P. S., Salmon, E. D. and Danuser, G. (2008). “Regional variation of microtubule flux reveals microtubule organization in the metaphase meiotic spindle.” J Cell Biol 182: 631639.CrossRefGoogle ScholarPubMed
Yang, G., Houghtaling, B. R., Gaetz, J., Liu, J. Z., Danuser, G. and Kapoor, T. M. (2007). “Architectural dynamics of the meiotic spindle revealed by single-fluorophore imaging.” Nat Cell Biol 9: 12331242.CrossRefGoogle ScholarPubMed
Yildiz, A. and Selvin, P. R. (2005). “Fluorescence imaging with one nanometer accuracy: application to molecular motors.” Acc Chem Res 38: 574582.CrossRefGoogle ScholarPubMed

References

Abdolahad, M., Sanaee, Z., Janmaleki, M., et al. (2012). “Vertically aligned multiwall-carbon nanotubes to preferentially entrap highly metastatic cancerous cells.” Carbon 50(5): 20102017.CrossRefGoogle Scholar
Adamo, A., Sharei, A., Adamo, L., et al. (2012). “Microfluidics-based assessment of cell deformability.” Anal Chem 84(15): 64386443.CrossRefGoogle ScholarPubMed
Ashkin, A. and Dziedzic, J. M.. (1987). “Optical trapping and manipulation of viruses and bacteria.” Science 235(4795): 15171520.CrossRefGoogle ScholarPubMed
Babahosseini, H., Srinivasaraghavan, V. and Agah, M.. (2012). “Microfluidic chip bio-sensor for detection of cancer cells.” Sensors. Taipei, IEEE 14, 28–31 Oct. 2012.Google Scholar
Badique, F., Stamov, D. R., Davidson, P. M., et al. (2013). “Directing nuclear deformation on micropillared surfaces by substrate geometry and cytoskeleton organization.” Biomaterials 34(12): 29913001.CrossRefGoogle ScholarPubMed
Beebe, D. J., Mensing, G. A. and Walker, G. M.. (2002). “Physics and applications of microfluidics in biology.” Annu Rev Biomed Eng 4: 261286.CrossRefGoogle Scholar
Bellini, N., Bragheri, F., Cristiani, I., et al. (2012). “Validation and perspectives of a femtosecond laser fabricated monolithic optical stretcher.” Biomed Opt Express 3(10): 26582668.CrossRefGoogle ScholarPubMed
Binnig, G., Quate, C. F. and Gerber, C.. (1986). “Atomic force microscope.” Phys Rev Lett 56(9): 930933.CrossRefGoogle ScholarPubMed
Byun, S., Son, S., Amodei, D., et al. (2013). “Characterizing deformability and surface friction of cancer cells.” Proc Natl Acad Sci USA 110(19): 6.CrossRefGoogle ScholarPubMed
Chen, C. L., Mahalingam, D., Osmulski, P., et al. (2013). “Single-cell analysis of circulating tumor cells identifies cumulative expression patterns of EMT-related genes in metastatic prostate cancer.” Prostate 73(8): 813826.CrossRefGoogle ScholarPubMed
Cheng, G., Tse, J., Jain, R. K., et al. (2009). “Micro-environmental mechanical stress controls tumor spheroid size and morphology by suppressing proliferation and inducing apoptosis in cancer cells.” PLoS One 4(2): e4632.CrossRefGoogle ScholarPubMed
Cheung, L. S., Zheng, X., Stopa, A., et al. (2009). “Detachment of captured cancer cells under flow acceleration in a bio-functionalized microchannel.” Lab Chip 9(12): 17211731.CrossRefGoogle Scholar
Cree, I. A. (2011). “Principles of cancer cell culture.” Methods Mol Biol 731: 1326.CrossRefGoogle ScholarPubMed
Crick, F. H. C., and Hughes, A. F. W. (1950). “The physical properties of cytoplasm: a study by means of the magnetic particle method.” Experimental Cell Research 1(1): 44.Google Scholar
Cross, S. E., Jin, Y. S., Rao, J., et al. (2007). “Nanomechanical analysis of cells from cancer patients.” Nat Nanotechnol 2(12): 780783.CrossRefGoogle ScholarPubMed
Darling, E. M., Zauscher, S., Block, J. A., et al. (2007). “A thin-layer model for viscoelastic, stress-relaxation testing of cells using atomic force microscopy: do cell properties reflect metastatic potential?Biophys J 92(5): 17841791.CrossRefGoogle ScholarPubMed
Davidson, P. M. and Lammerding, J.. (2014). “Broken nuclei–lamins, nuclear mechanics, and disease.” Trends Cell Biol 24(4): 247256.CrossRefGoogle ScholarPubMed
Dudani, J. S., Gossett, D. R., Tse, H. T., et al. (2013). “Pinched-flow hydrodynamic stretching of single-cells.” Lab Chip 13(18): 37283734.CrossRefGoogle ScholarPubMed
Dufrene, Y. F. and Pelling, A. E.. (2013). “Force nanoscopy of cell mechanics and cell adhesion.” Nanoscale 5(10): 40944104.CrossRefGoogle ScholarPubMed
Faria, E. C., Ma, N., Gazi, E., et al. (2008). “Measurement of elastic properties of prostate cancer cells using AFM.” Analyst 133(11): 14981500.CrossRefGoogle ScholarPubMed
Faustino, V., Pinho, D., Yaginuma, T., et al. (2014). “Extensional flow-based microfluidic device: deformability assessment of red blood cells in contact with tumor cells.” BioChip Journal 8(1): 4247.CrossRefGoogle Scholar
Fu, C., Han, C., Cheng, C., et al. (2012). “Bio-mechanical properties of human renal cancer cells probed by magneto-optical tweezers.” Sensors. Taipei, IEEE, 14, 28–31 Oct. 2012.Google Scholar
Fu, Y., Vandongen, A. M. J., Bourouina, T., et al. (2012). “A study of cancer cell metastasis using microfluidic transmigration device.” MEMS. Paris, IEEE: 773776, 29 Jan. 2012–2 Feb. 2012.Google Scholar
Giverso, C., Grillo, A. and Preziosi, L.. (2014). “Influence of nucleus deformability on cell entry into cylindrical structures.” Biomech Model Mechanobiol 13(3): 481502.CrossRefGoogle ScholarPubMed
Gossett, D. R., Tse, H. T., Lee, S. A., et al. (2012). “Hydrodynamic stretching of single cells for large population mechanical phenotyping.” Proc Natl Acad Sci USA 109(20): 76307635.CrossRefGoogle ScholarPubMed
Guck, J., Ananthakrishnan, R., Mahmood, H., et al. (2001). “The optical stretcher: a novel laser tool to micromanipulate cells.” Biophys J 81(2): 767784.CrossRefGoogle ScholarPubMed
Guck, J., Schinkinger, S., Lincoln, B., et al. (2005). “Optical deformability as an inherent cell marker for testing malignant transformation and metastatic competence.” Biophys J 88(5): 36893698.CrossRefGoogle ScholarPubMed
Guo, Q., Park, S. and Ma, H.. (2012). “Microfluidic micropipette aspiration for measuring the deformability of single cells.” Lab Chip 12(15): 26872695.CrossRefGoogle ScholarPubMed
Hou, H. W., Li, Q. S., Lee, G. Y., et al. (2009). “Deformability study of breast cancer cells using microfluidics.” Biomed Microdevices 11(3): 557564.CrossRefGoogle ScholarPubMed
Jaeger, A. A., Das, C. K., Morgan, N. Y., et al. (2013). “Microfabricated polymeric vessel mimetics for 3-D cancer cell culture.” Biomaterials 34(33): 83018313.CrossRefGoogle ScholarPubMed
Ketene, A. N., Schmelz, E. M., Roberts, P. C., et al. (2012). “The effects of cancer progression on the viscoelasticity of ovarian cell cytoskeleton structures.” Nanomedicine 8(1): 93102.CrossRefGoogle ScholarPubMed
Kirmizis, D. and Logothetidis, S.. (2010). “Atomic force microscopy probing in the measurement of cell mechanics.” Int J Nanomedicine 5: 137145.CrossRefGoogle ScholarPubMed
Kittur, H., Weaver, W. and Di Carlo, D.. (2014). “Well-plate mechanical confinement platform for studies of mechanical mutagenesis.” Biomed Microdevices 16(3): 439447.CrossRefGoogle ScholarPubMed
Kozissnik, B. and Dobson, J.. (2013). “Biomedical applications of mesoscale magnetic particles.” MRS Bulletin 38(11): 927932.CrossRefGoogle Scholar
Krause, M., Te Riet, J. and Wolf, K.. (2013). “Probing the compressibility of tumor cell nuclei by combined atomic force-confocal microscopy.” Phys Biol 10(6): 065002.CrossRefGoogle ScholarPubMed
Lee, M. H., Wu, P. H., Staunton, J. R., et al. (2012). “Mismatch in mechanical and adhesive properties induces pulsating cancer cell migration in epithelial monolayer.” Biophys J 102(12): 27312741.CrossRefGoogle ScholarPubMed
Lekka, M., Laidler, P., Gil, D., et al. (1999). “Elasticity of normal and cancerous human bladder cells studied by scanning force microscopy.” Eur Biophys J 28(4): 312316.CrossRefGoogle ScholarPubMed
Lekka, M., Pogoda, K., Gostek, J., et al. (2012). “Cancer cell recognition–mechanical phenotype.” Micron 43(12): 12591266.CrossRefGoogle ScholarPubMed
Leong, F. Y., Li, Q., Lim, C. T., et al. (2011). “Modeling cell entry into a micro-channel.” Biomech Model Mechanobiol 10(5): 755766.CrossRefGoogle ScholarPubMed
Li, Q. S., Lee, G. Y., Ong, C. N., et al. (2008). “AFM indentation study of breast cancer cells.” Biochem Biophys Res Commun 374(4): 609613.CrossRefGoogle ScholarPubMed
Lim, C. T. (2006). “Single Cell Mechanics Study of the Human Disease Malaria.” Journal of Biomechanical Science and Engineering 1(1): 8292.CrossRefGoogle Scholar
Lim, C. T. and Hoon, S. B. (2014). “Circulating tumor cells: Cancer’s deadly couriers.” Physics Today 67(2): 5.CrossRefGoogle Scholar
Lim, C. T., Zhou, E. H., Li, A., et al. (2006). “Experimental techniques for single cell and single molecule biomechanics.” Materials Science and Engineering C 26(8): 12781288.CrossRefGoogle Scholar
Liu, A., Liu, W., Wang, Y., et al. (2012). “Microvalve and liquid membrane double-controlled integrated microfluidics for observing the interaction of breast cancer cells.” Microfluidics and Nanofluidics 14(3–4): 515526.CrossRefGoogle Scholar
Liu, H., Tan, Q., Geddie, W. R., et al. (2014). “Biophysical characterization of bladder cancer cells with different metastatic potential.” Cell Biochem Biophys 68(2): 241246.CrossRefGoogle ScholarPubMed
Liu, J., Tan, Y., Zhang, H., et al. (2012). “Soft fibrin gels promote selection and growth of tumorigenic cells.” Nat Mater 11(8): 734741.CrossRefGoogle ScholarPubMed
Mak, M. and Erickson, D.. (2013). “A serial micropipette microfluidic device with applications to cancer cell repeated deformation studies.” Integr Biol (Camb) 5(11): 13741384.CrossRefGoogle ScholarPubMed
Mak, M., Reinhart-King, C. A. and Erickson, D.. (2013). “Elucidating mechanical transition effects of invading cancer cells with a subnucleus-scaled microfluidic serial dimensional modulation device.” Lab Chip 13(3): 340348.CrossRefGoogle ScholarPubMed
Muller, D. J., Kim, K. S., Cho, C. H., et al. (2012). “AFM-detected apoptotic changes in morphology and biophysical property caused by paclitaxel in Ishikawa and HeLa cells.” PLoS One 7(1): e30066.Google Scholar
Nguyen, T. A., Yin, T. I., Reyes, D., et al. (2013). “Microfluidic chip with integrated electrical cell-impedance sensing for monitoring single cancer cell migration in three-dimensional matrixes.” Anal Chem 85(22): 1106811076.CrossRefGoogle ScholarPubMed
Nikkhah, M., Strobl, J. S., De Vita, R., et al. (2010). “The cytoskeletal organization of breast carcinoma and fibroblast cells inside three dimensional. (3-D) isotropic silicon microstructures.” Biomaterials 31(16): 45524561.CrossRefGoogle ScholarPubMed
Park, S., Ang, R. R., Duffy, S. P., et al. (2014). “Morphological differences between circulating tumor cells from prostate cancer patients and cultured prostate cancer cells.” PLoS One 9(1): e85264.CrossRefGoogle ScholarPubMed
Paszek, M. J., Zahir, N., Johnson, K. R., et al. (2005). “Tensional homeostasis and the malignant phenotype.” Cancer Cell 8(3): 241254.CrossRefGoogle ScholarPubMed
Plodinec, M., Loparic, M., Monnier, C. A., et al. (2012). “The nanomechanical signature of breast cancer.” Nat Nanotechnol 7(11): 757765.CrossRefGoogle ScholarPubMed
Prabhune, M., Belge, G., Dotzauer, A., et al. (2012). “Comparison of mechanical properties of normal and malignant thyroid cells.” Micron 43(12): 12671272.CrossRefGoogle ScholarPubMed
Preira, P., Grandne, V., Forel, J. M., et al. (2013). “Passive circulating cell sorting by deformability using a microfluidic gradual filter.” Lab Chip 13(1): 161170.CrossRefGoogle ScholarPubMed
Rabineau, M., Kocgozlu, L., Dujardin, D., et al. (2013). “Contribution of soft substrates to malignancy and tumor suppression during colon cancer cell division.” PLoS One 8(10): e78468.CrossRefGoogle ScholarPubMed
Ramos, J. R., Pabijan, J., Garcia, R., et al. (2014). “The softening of human bladder cancer cells happens at an early stage of the malignancy process.” Beilstein J Nanotechnol 5: 447457.CrossRefGoogle ScholarPubMed
Rebelo, L. M., de Sousa, J. S., Filho, J. Mendes, et al. (2013). “Comparison of the viscoelastic properties of cells from different kidney cancer phenotypes measured with atomic force microscopy.” Nanotechnology 24(5): 055102.CrossRefGoogle ScholarPubMed
Sawetzki, T., Eggleton, C. D., Desai, S. A., et al. (2013). “Viscoelasticity as a biomarker for high-throughput flow cytometry.” Biophys J 105(10): 22812288.CrossRefGoogle ScholarPubMed
Scianna, M. and Preziosi, L.. (2013). “Modeling the influence of nucleus elasticity on cell invasion in fiber networks and microchannels.” J Theor Biol 317: 394406.CrossRefGoogle ScholarPubMed
Sharma, S., Santiskulvong, C., Bentolila, L. A., et al. (2012). “Correlative nanomechanical profiling with super-resolution F-actin imaging reveals novel insights into mechanisms of cisplatin resistance in ovarian cancer cells.” Nanomedicine 8(5): 757766.CrossRefGoogle ScholarPubMed
Shojaei-Baghini, E., Zheng, Y. and Sun, Y.. (2013). “Automated micropipette aspiration of single cells.” Ann Biomed Eng 41(6): 12081216.CrossRefGoogle ScholarPubMed
Sun, W., Kurniawan, N. A., Kumar, A. P., et al. (2014). “Effects of migrating cell-induced matrix reorganization on 3d cancer cell migration.” Cellular and Molecular Bioengineering 7(2): 205217.CrossRefGoogle Scholar
Sun, W., Lim, C. T. and Kurniawan, N. A.. (2014). “Mechanistic adaptability of cancer cells strongly affects anti-migratory drug efficacy.” J R Soc Interface 11(99).CrossRefGoogle ScholarPubMed
Suresh, S., Spatz, J., Mills, J. P., et al. (2005). “Connections between single-cell biomechanics and human disease states: gastrointestinal cancer and malaria.” Acta Biomater 1(1): 1530.CrossRefGoogle ScholarPubMed
Tan, Y., Tajik, A., 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
Thiery, J. P. and Lim, C. T.. (2013). “Tumor dissemination: an EMT affair.” Cancer Cell 23(3): 272273.CrossRefGoogle ScholarPubMed
Tilghman, R. W., Cowan, C. R., Mih, J. D., et al. (2010). “Matrix rigidity regulates cancer cell growth and cellular phenotype.” PLoS One 5(9): e12905.CrossRefGoogle ScholarPubMed
Tsai, C. H., Sakuma, S., Arai, F., et al. (2014). “A new dimensionless index for evaluating cell stiffness-based deformability in microchannel.” IEEE Trans Biomed Eng 61(4): 11871195.CrossRefGoogle ScholarPubMed
Tse, H. T., Weaver, W. M. and Di Carlo, D.. (2012). “Increased asymmetric and multi-daughter cell division in mechanically confined microenvironments.” PLoS One 7(6): e38986.CrossRefGoogle ScholarPubMed
Tseng, P., Judy, J. W. and Di Carlo, D.. (2012). “Magnetic nanoparticle–mediated massively parallel mechanical modulation of single-cell behavior.” Nat Methods 09(11): 9.CrossRefGoogle Scholar
Tzvetkova-Chevolleau, T., Stephanou, A., Fuard, D., et al. (2008). “The motility of normal and cancer cells in response to the combined influence of the substrate rigidity and anisotropic microstructure.” Biomaterials 29(10): 15411551.CrossRefGoogle Scholar
Weder, G., Hendriks-Balk, M. C., Smajda, R., et al. (2014). “Increased plasticity of the stiffness of melanoma cells correlates with their acquisition of metastatic properties.” Nanomedicine 10(1): 141148.CrossRefGoogle ScholarPubMed
Werfel, J., Krause, S., Bischof, A. G., et al. (2013). “How changes in extracellular matrix mechanics and gene expression variability might combine to drive cancer progression.” PLoS One 8(10): e76122.CrossRefGoogle ScholarPubMed
Wu, Y., McEwen, G. D., Harihar, S., et al. (2010). “BRMS1 expression alters the ultrastructural, biomechanical and biochemical properties of MDA-MB-435 human breast carcinoma cells: an AFM and Raman microspectroscopy study.” Cancer Lett 293(1): 8291.CrossRefGoogle ScholarPubMed
Wuang, S. C., Ladoux, B. and Lim, C. T.. (2011). “Probing the chemo-mechanical effects of an anti-cancer drug Emodin on breast cancer cells.” Cellular and Molecular Bioengineering 4(3): 466475.CrossRefGoogle Scholar
Xu, W., Mezencev, R., Kim, B., et al. (2012). “Cell stiffness is a biomarker of the metastatic potential of ovarian cancer cells.” PLoS One 7(10): e46609.CrossRefGoogle ScholarPubMed

References

Aoyagi, T., Ebara, M., Sakai, K., Sakurai, Y. and Okano, T. (2000). “Novel bifunctional polymer with reactivity and temperature sensitivityJournal of Biomaterials Science, Polymer Edition 11: 101110.CrossRefGoogle ScholarPubMed
Auernheimer, J., Dahmen, C., Hersel, U., Bausch, A. and Kessler, H. (2005). “Photoswitched cell adhesion on surfaces with RGD peptides.” Journal of the American Chemical Society 127: 1610716110.CrossRefGoogle ScholarPubMed
Davis, K. A., Burke, K. A., Mather, P. T. and Henderson, J. H. (2011). “Dynamic cell behavior on shape memory polymer substrates.” Biomaterials 32: 22852293.CrossRefGoogle ScholarPubMed
Deforest, C. A. and Anseth, K. S. (2011). “Cytocompatible click-based hydrogels with dynamically tunable properties through orthogonal photoconjugation and photocleavage reactions.” Nature Chemistry 3: 925931.CrossRefGoogle ScholarPubMed
Ebara, M., Akimoto, M., Uto, K., et al. (2014a). “Focus on the interlude between topographic transition and cell response on shape-memory surfaces.” Polymer 55: 59615968.CrossRefGoogle Scholar
Ebara, M., Uto, K., Idota, N., Hoffman, J. M. and Aoyagi, T. (2012). “Shape-memory surface with dynamically tunable nano-geometry activated by body heat.” Advanced Materials 24: 273278.CrossRefGoogle ScholarPubMed
Ebara, M., Uto, K., Idota, N., Hoffman, J. M. and Aoyagi, T. (2014b). “The taming of the cell: shape-memory nanopatterns direct cell orientation.” International Journal of Nanomedicine 9(Supplement 1): 117126.CrossRefGoogle ScholarPubMed
Ebara, M., Yamato, M., Aoyagi, T., et al. (2004a). “Immobilization of cell-adhesive peptides to temperature-responsive surfaces facilitates both serum-free cell adhesion and noninvasive cell harvest.” Tissue Engineering 10: 11251135.CrossRefGoogle ScholarPubMed
Ebara, M., Yamato, M., Aoyagi, T., et al. (2004b). “Temperature-responsive cell culture surfaces enable ‘on−off’ affinity control between cell integrins and RGDS ligands.” Biomacromolecules 5: 505510.CrossRefGoogle ScholarPubMed
Ebara, M., Yamato, M., Aoyagi, T., et al. (2008). “The effect of extensible PEG tethers on shielding between grafted thermo-responsive polymer chains and integrin–RGD binding.” Biomaterials 29: 36503655.CrossRefGoogle ScholarPubMed
Ebara, M., Yamato, M., Hirose, M., et al. (2003). “Copolymerization of 2-carboxyisopropylacrylamide with N-isopropylacrylamide accelerates cell detachment from grafted surfaces by reducing temperature.” Biomacromolecules 4: 344349.CrossRefGoogle ScholarPubMed
Edahiro, J.-I., Sumaru, K., Tada, Y., et al. (2005). “In situ control of cell adhesion using photoresponsive culture surface.” Biomacromolecules 6: 970974.CrossRefGoogle ScholarPubMed
Egami, M., Haraguchi, Y., Shimizu, T., Yamato, M. and Okano, T. (2014). “Latest status of the clinical and industrial applications of cell sheet engineering and regenerative medicine.” Archives of Pharmacal Research 37: 96106.CrossRefGoogle ScholarPubMed
Frey, M. T. and Wang, Y.-L. (2009). “A photo-modulatable material for probing cellular responses to substrate rigidity.” Soft Matter 5: 19181924.CrossRefGoogle ScholarPubMed
Georges, P. C., Hui, J.-J., Gombos, Z., et al. (2007). “Increased stiffness of the rat liver precedes matrix deposition: implications for fibrosis.” American Journal of Physiology-Gastrointestinal and Liver Physiology 296(6): G1147G1156.CrossRefGoogle Scholar
Guvendiren, M. and Burdick, J. A. (2012). “Stiffening hydrogels to probe short- and long-term cellular responses to dynamic mechanics.” Nat Commun 3: 792.CrossRefGoogle Scholar
Idota, N., Ebara, M., Kotsuchibashi, Y., Narain, R. and Aoyagi, T. (2012). “Novel temperature-responsive polymer brushes with carbohydrate residues facilitate selective adhesion and collection of hepatocytes.” Science and Technology of Advanced Materials 13: 064206.CrossRefGoogle ScholarPubMed
Jiang, F. X., Yurke, B., Schloss, R. S., Firestein, B. L. and Langrana, N. A. (2010a). “Effect of dynamic stiffness of the substrates on neurite outgrowth by using a DNA-crosslinked hydrogel.” Tissue Eng Part A 16: 18731889.CrossRefGoogle ScholarPubMed
Jiang, F. X., Yurke, B., Schloss, R. S., Firestein, B. L. and Langrana, N. A. (2010b). “The relationship between fibroblast growth and the dynamic stiffnesses of a DNA crosslinked hydrogel.” Biomaterials 31: 11991212.CrossRefGoogle ScholarPubMed
Kakegawa, T., Mochizuki, N., Sadr, N., Suzuki, H. and Fukuda, J. (2014). “Cell-adhesive and cell-repulsive zwitterionic oligopeptides for micropatterning and rapid electrochemical detachment of cells.” Tissue Eng Part A 19: 290298.CrossRefGoogle Scholar
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
Kiang, J. D., Wen, J. H., Del Álamo, J. C. and Engler, A. J. (2013). “Dynamic and reversible surface topography influences cell morphology.” Journal of Biomedical Materials Research Part A 101A: 23132321.CrossRefGoogle Scholar
Kirschner, C. M. and Anseth, K. S. (2013). “In situ control of cell substrate microtopographies using photolabile hydrogels.” Small 9: 578584.Google ScholarPubMed
Kloxin, A. M., Kasko, A. M., Salinas, C. N. and Anseth, K. S. (2009). “Photodegradable hydrogels for dynamic tuning of physical and chemical properties.” Science 324: 5963.CrossRefGoogle ScholarPubMed
Kushida, A., Yamato, M., Konno, C., et al. (1999). “Decrease in culture temperature releases monolayer endothelial cell sheets together with deposited fibronectin matrix from temperature-responsive culture surfaces.” Journal of Biomedical Materials Research 45: 355362.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
Lam, M. T., Clem, W. C. and Takayama, S. (2008). “Reversible on-demand cell alignment using reconfigurable microtopography.” Biomaterials 29: 17051712.CrossRefGoogle ScholarPubMed
Lamb, B. M. and Yousaf, M. N. (2011). “Redox-switchable surface for controlling peptide structure.” Journal of the American Chemical Society 133 88708873.CrossRefGoogle ScholarPubMed
Le, D. M., Kulangara, K., Adler, A. F., Leong, K. W. and Ashby, V. S. (2011). “Dynamic topographical control of mesenchymal stem cells by culture on responsive poly(ε-caprolactone) surfaces.” Advanced Materials 23: 32783283.CrossRefGoogle ScholarPubMed
Luo, W. and Yousaf, M. N. (2011). “Tissue morphing control on dynamic gradient surfaces.” Journal of the American Chemical Society 133: 1078010783.CrossRefGoogle ScholarPubMed
Mosqueira, D., Pagliari, S., Uto, K., et al. (2014). “Hippo pathway effectors control cardiac progenitor cell fate by acting as dynamic sensors of substrate mechanics and nanostructure.” Acs Nano 8: 20332047.CrossRefGoogle ScholarPubMed
Nakanishi, J., Kikuchi, Y., Inoue, S., et al. (2007). “Spatiotemporal control of migration of single cells on a photoactivatable cell microarray.” Journal of the American Chemical Society 129: 66946695.CrossRefGoogle ScholarPubMed
Nakanishi, J., Kikuchi, Y., Takarada, T., et al. (2006). “Spatiotemporal control of cell adhesion on a self-assembled monolayer having a photocleavable protecting group.” Analytica Chimica Acta 578: 100104.CrossRefGoogle ScholarPubMed
Rolli, C. G., Nakayama, H., Yamaguchi, K., et al. (2012). “Switchable adhesive substrates: Revealing geometry dependence in collective cell behavior.” Biomaterials 33: 24092418.CrossRefGoogle ScholarPubMed
Uto, K., Ebara, M. and Aoyagi, T. (2014). “Temperature-responsive poly(ε-caprolactone) cell culture platform with dynamically tunable nano-roughness and elasticity for control of myoblast morphology.” International Journal of Molecular Sciences 15: 15111524.CrossRefGoogle ScholarPubMed
Yamada, N., Okano, T., Sakai, H., et al. (1990). “Thermo-responsive polymeric surfaces; control of attachment and detachment of cultured cells.” Makromolekulare Chemie, Rapid Communications 11: 571576.CrossRefGoogle Scholar
Yang, C., Tibbitt, M. W., Basta, L. and Anseth, K. S. (2014). “Mechanical memory and dosing influence stem cell fate.” Nat Mater 13: 645652.CrossRefGoogle ScholarPubMed
Yeo, W.-S., Yousaf, M. N. and Mrksich, M. (2003). “Dynamic interfaces between cells and surfaces: electroactive substrates that sequentially release and attach cells.” Journal of the American Chemical Society 125: 1499414995.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
Yuan, H., Kononov, S., Cavalcante, F. S. A., et al. 2000. “Effects of collagenase and elastase on the mechanical properties of lung tissue strips.” Journal of Applied Physiology 89: 314.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
×