Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-fwgfc Total loading time: 0 Render date: 2024-07-11T01:21:00.681Z Has data issue: false hasContentIssue false

5 - Multiscale topographical approaches for cell mechanobiology studies

from 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

Human tissues are sophisticated ensembles of various cell types embedded in the complex but defined structures of the extracellular matrix (ECM). ECM is configured in a hierarchical structure from nano- to microscale, with many biological molecules forming large scale configurations and textures with feature sizes up to macroscopic scale (several hundred microns). The physicochemical, biological and mechanostructural properties of native ECM play a critical role in constructing a microenvironment for cells and tissues. In conjunction with the rapid evolution of material science and its fabrication techniques, studies of the topography and elasticity of ECM and other materials have allowed advanced interrogation of cellular mechanotransduction and cellular responses to mechanostructural cues. By learning from and mimicking the highly organized ECM structures found in vivo, topography-guided approaches to regulate cell function and fate have been widely investigated in the last several decades. Here, we review recent efforts in mimicking the micro- and nanotopography of the native ECM in vitro for the regulation of cellular behaviors. We also discuss how these biomimetic topographical surfaces have been applied to fundamental cell mechanobiology studies into cell adhesions, migrations, and differentiation as well as toward efforts in tissue engineering.

Type
Chapter
Information
Integrative Mechanobiology
Micro- and Nano- Techniques in Cell Mechanobiology
, pp. 69 - 89
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

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

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
×