Show simple item record

Fibronectin splice variants: Understanding their multiple roles in health and disease using engineered mouse models

dc.contributor.authorOvádi, Juditen_US
dc.date.accessioned2011-11-10T15:35:31Z
dc.date.available2012-09-04T15:27:43Zen_US
dc.date.issued2011-07en_US
dc.identifier.citationOvádi, Judit (2011). "Fibronectin splice variants: Understanding their multiple roles in health and disease using engineered mouse models." IUBMB Life 63(7): 538-546. <http://hdl.handle.net/2027.42/86988>en_US
dc.identifier.issn1521-6543en_US
dc.identifier.issn1521-6551en_US
dc.identifier.urihttps://hdl.handle.net/2027.42/86988
dc.description.abstractThe extracellular matrix (ECM) is a highly dynamic network of proteins, glycoproteins, and proteoglycans. Numerous diseases result from mutation in genes coding for ECM proteins, but only recently it has been reported that mutations in the fibronectin (FN) gene were associated with a human disorder. FN is one of the main components of the ECM. It generates protein diversity through alternative splicing of a single pre‐mRNA, having at least 20 different isoforms in humans. The precise function of these protein isoforms has remained obscure in most cases. Only in the recent few years, it was possible to shed light on the multiple roles of the alternatively spliced FN isoforms. This substantial progress was achieved basically with the knowledge derived from engineered mouse models bearing subtle mutations in specific FN domains. These data, together with a recent report associating mutations in the FN gene to a form of glomerulopathy, clearly show that mutations in constitutive exons or misregulation of alternatively spliced domains of the FN gene may have nonlethal pathological consequences. In this review, we focus on the pathological consequences of mutations in the FN gene, by connecting the function of alternatively spliced isoforms of fibronectin to human diseases. © 2011 IUBMB IUBMB Life, 63(7): 538–546, 2011en_US
dc.publisherWiley Subscription Services, Inc., a Wiley companyen_US
dc.subject.otherDisease Modelsen_US
dc.subject.otherGenetic Modelsen_US
dc.subject.otherAlternative Splicingen_US
dc.subject.otherProtein Functionen_US
dc.subject.otherComplex Diseasesen_US
dc.titleFibronectin splice variants: Understanding their multiple roles in health and disease using engineered mouse modelsen_US
dc.typeArticleen_US
dc.rights.robotsIndexNoFollowen_US
dc.subject.hlbsecondlevelEcology and Evolutionary Biologyen_US
dc.subject.hlbtoplevelScienceen_US
dc.description.peerreviewedPeer Revieweden_US
dc.contributor.affiliationumDepartment of Internal Medicine, Division of Pulmonary and Critical Care Medicine, University of Michigan Medical School, Ann Arbor, MIen_US
dc.contributor.affiliationotherMouse Molecular Genetics Laboratory, International Centre for Genetic Engineering and Biotechnology, Trieste, Italyen_US
dc.contributor.affiliationotherInternational Centre for Genetic Engineering and Biotechnology, Padriciano 99, I‐34149 Trieste, Italyen_US
dc.description.bitstreamurlhttp://deepblue.lib.umich.edu/bitstream/2027.42/86988/1/493_ftp.pdf
dc.identifier.doi10.1002/iub.493en_US
dc.identifier.sourceIUBMB Lifeen_US
dc.identifier.citedreferenceDischer, D. E., Mooney, D. J., and Zandstra, P. W. ( 2009 ) Growth factors, matrices, and forces combine and control stem cells. Science 324, 1673 – 1677.en_US
dc.identifier.citedreferenceGeiger, B., Spatz, J. P., and Bershadsky, A. D. ( 2009 ) Environmental sensing through focal adhesions. Nat. Rev. Mol. Cell Biol. 10, 21 – 33.en_US
dc.identifier.citedreferenceBateman, J. F., Boot‐Handford, R. P., and Lamande, S. R. ( 2009 ) Genetic diseases of connective tissues: cellular and extracellular effects of ECM mutations. Nat. Rev. Genet. 10, 173 – 183.en_US
dc.identifier.citedreferenceCastelletti, F., Donadelli, R., Banterla, F., Hildebrandt, F., Zipfel, P. F., et al. ( 2008 ) Mutations in FN1 cause glomerulopathy with fibronectin deposits. Proc. Natl. Acad. Sci. USA 105, 2538 – 2543.en_US
dc.identifier.citedreferenceAstrof, S., Crowley, D., and Hynes, R. O. ( 2007 ) Multiple cardiovascular defects caused by the absence of alternatively spliced segments of fibronectin. Dev. Biol. 311, 11 – 24.en_US
dc.identifier.citedreferenceFukuda, T., Yoshida, N., Kataoka, Y., Manabe, R., Mizuno‐Horikawa, Y., et al. ( 2002 ) Mice lacking the EDB segment of fibronectin develop normally but exhibit reduced cell growth and fibronectin matrix assembly in vitro. Cancer Res. 62, 5603 – 5610.en_US
dc.identifier.citedreferenceGeorge, E. L., Georges‐Labouesse, E. N., Patel‐King, R. S., Rayburn, H., and Hynes, R. O. ( 1993 ) Defects in mesoderm, neural tube and vascular development in mouse embryos lacking fibronectin. Development 119, 1079 – 1091.en_US
dc.identifier.citedreferenceGeorges‐Labouesse, E. N., George, E. L., Rayburn, H., and Hynes, R. O. ( 1996 ) Mesodermal development in mouse embryos mutant for fibronectin. Dev. Dyn. 207, 145 – 156.en_US
dc.identifier.citedreferenceMuro, A. F., Chauhan, A. K., Gajovic, S., Iaconcig, A., Porro, F., et al. ( 2003 ) Regulated splicing of the fibronectin EDA exon is essential for proper skin wound healing and normal lifespan. J. Cell Biol. 162, 149 – 160.en_US
dc.identifier.citedreferenceSakai, T., Johnson, K. J., Murozono, M., Sakai, K., Magnuson, M. A., et al. ( 2001 ) Plasma fibronectin supports neuronal survival and reduces brain injury following transient focal cerebral ischemia but is not essential for skin‐wound healing and hemostasis. Nat. Med. 7, 324 – 330.en_US
dc.identifier.citedreferenceTakahashi, M., Saito, H., Higashimoto, M., and Hibi, T. ( 2007 ) Possible inhibitory effect of oral zinc supplementation on hepatic fibrosis through downregulation of TIMP‐1: a pilot study. Hepatol. Res. 37, 405 – 409.en_US
dc.identifier.citedreferenceTan, M. H., Sun, Z., Opitz, S. L., Schmidt, T. E., Peters, J. H., et al. ( 2004 ) Deletion of the alternatively spliced fibronectin EIIIA domain in mice reduces atherosclerosis. Blood 104, 11 – 18.en_US
dc.identifier.citedreferenceHynes, R. O. ( 1990 ) Fibronectins. Springer‐Verlag, New York.en_US
dc.identifier.citedreferenceWhite, E. S., Baralle, F. E., and Muro, A. F. ( 2008 ) New insights into form and function of fibronectin splice variants. J. Pathol. 216, 1 – 14.en_US
dc.identifier.citedreferenceGilbert, W. ( 1978 ) Why genes in pieces? Nature 271, 501.en_US
dc.identifier.citedreferenceHynes, R. O. ( 2007 ) Cell–matrix adhesion in vascular development. J. Thromb. Haemost. 5 ( Suppl 1 ), 32 – 40.en_US
dc.identifier.citedreferenceChauhan, A. K., Iaconcig, A., Baralle, F. E., and Muro, A. F. ( 2004 ) Alternative splicing of fibronectin: a mouse model demonstrates the identity of in vitro and in vivo systems and the processing autonomy of regulated exons in adult mice. Gene 324, 55 – 63.en_US
dc.identifier.citedreferencePeters, J. H. and Hynes, R. O. ( 1996 ) Fibronectin isoform distribution in the mouse. I. The alternatively spliced EIIIB, EIIIA, and V segments show widespread codistribution in the developing mouse embryo. Cell Adhes. Commun. 4, 103 – 125.en_US
dc.identifier.citedreferenceMagnuson, V. L., Young, M., Schattenberg, D. G., Mancini, M. A., Chen, D. L., et al. ( 1991 ) The alternative splicing of fibronectin pre‐mRNA is altered during aging and in response to growth factors. J. Biol. Chem. 266, 14654 – 14662.en_US
dc.identifier.citedreferencePagani, F., Zagato, L., Vergani, C., Casari, G., Sidoli, A., and Baralle, F. E. ( 1991 ) Tissue‐specific splicing pattern of fibronectin messenger RNA precursor during development and aging in rat. J. Cell Biol. 113, 1223 – 1229.en_US
dc.identifier.citedreferenceMoretti, F. A., Chauhan, A. K., Iaconcig, A., Porro, F., Baralle, F. E., and Muro, A. F. ( 2007 ) A major fraction of fibronectin present in the extracellular matrix of tissues is plasma‐derived. J. Biol. Chem. 282, 28057 – 28062.en_US
dc.identifier.citedreferenceFfrench‐Constant, C., Van de Water, L., Dvorak, H. F., and Hynes, R. O. ( 1989 ) Reappearance of an embryonic pattern of fibronectin splicing during wound healing in the adult rat. J. Cell Biol. 109, 903 – 914.en_US
dc.identifier.citedreferencePeters, J. H., Chen, G. E., and Hynes, R. O. ( 1996 ) Fibronectin isoform distribution in the mouse. II. Differential distribution of the alternatively spliced EIIIB, EIIIA, and V segments in the adult mouse. Cell Adhes. Commun. 4, 127 – 148.en_US
dc.identifier.citedreferenceCarnemolla, B., Balza, E., Siri, A., Zardi, L., Nicotra, M. R., et al. ( 1989 ) A tumor‐associated fibronectin isoform generated by alternative splicing of messenger RNA precursors. J. Cell Biol. 108, 1139 – 1148.en_US
dc.identifier.citedreferenceRybak, J. N., Roesli, C., Kaspar, M., Villa, A., and Neri, D. ( 2007 ) The extra‐domain A of fibronectin is a vascular marker of solid tumors and metastases. Cancer Res. 67, 10948 – 10957.en_US
dc.identifier.citedreferenceCastellani, P., Viale, G., Dorcaratto, A., Nicolo, G., Kaczmarek, J., et al. ( 1994 ) The fibronectin isoform containing the ED‐B oncofetal domain: a marker of angiogenesis. Int. J. Cancer 59, 612 – 618.en_US
dc.identifier.citedreferenceAstrof, S., Crowley, D., George, E. L., Fukuda, T., Sekiguchi, K., et al. ( 2004 ) Direct test of potential roles of EIIIA and EIIIB alternatively spliced segments of fibronectin in physiological and tumor angiogenesis. Mol. Cell. Biol. 24, 8662 – 8670.en_US
dc.identifier.citedreferenceAstrof, S. and Hynes, R. O. ( 2009 ) Fibronectins in vascular morphogenesis. Angiogenesis 12, 165 – 175.en_US
dc.identifier.citedreferenceBazigou, E., Xie, S., Chen, C., Weston, A., Miura, N., et al. ( 2009 ) Integrin‐alpha9 is required for fibronectin matrix assembly during lymphatic valve morphogenesis. Dev. Cell 17, 175 – 186.en_US
dc.identifier.citedreferenceHuang, X. Z., Wu, J. F., Ferrando, R., Lee, J. H., Wang, Y. L., et al. ( 2000 ) Fatal bilateral chylothorax in mice lacking the integrin alpha9beta1. Mol. Cell. Biol. 20, 5208 – 5215.en_US
dc.identifier.citedreferencePaloschi, V., Kurtovic, S., Folkersen, L., Gomez, D., Wagsater, D., et al. ( 2011 ) Impaired splicing of fibronectin is associated with thoracic aortic aneurysm formation in patients with bicuspid aortic valve. Arterioscler. Thromb. Vasc. Biol. 31, 691 – 697.en_US
dc.identifier.citedreferenceBabaev, V. R., Porro, F., Linton, M. F., Fazio, S., Baralle, F. E., and Muro, A. F. ( 2008 ) Absence of regulated splicing of fibronectin EDA exon reduces atherosclerosis in mice. Atherosclerosis 197, 534 – 540.en_US
dc.identifier.citedreferenceHernnas, J., Nettelbladt, O., Bjermer, L., Sarnstrand, B., Malmstrom, A., and Hallgren, R. ( 1992 ) Alveolar accumulation of fibronectin and hyaluronan precedes bleomycin‐induced pulmonary fibrosis in the rat. Eur. Respir. J. 5, 404 – 410.en_US
dc.identifier.citedreferenceMuro, A. F., Moretti, F. A., Moore, B. B., Yan, M., Atrasz, R. G., et al. ( 2008 ) An essential role for fibronectin extra type III domain A in pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 177, 638 – 645.en_US
dc.identifier.citedreferenceJarnagin, W. R., Rockey, D. C., Koteliansky, V. E., Wang, S. S., and Bissell, D. M. ( 1994 ) Expression of variant fibronectins in wound healing: cellular source and biological activity of the EIIIA segment in rat hepatic fibrogenesis. J. Cell Biol. 127, 2037 – 2048.en_US
dc.identifier.citedreferenceSerini, G., Bochaton‐Piallat, M. L., Ropraz, P., Geinoz, A., Borsi, L., et al. ( 1998 ) The fibronectin domain ED‐A is crucial for myofibroblastic phenotype induction by transforming growth factor‐beta1. J. Cell Biol. 142, 873 – 881.en_US
dc.identifier.citedreferenceKohan, M., Muro, A. F., Bader, R., and Berkman, N. ( 2011 ) The extra domain A of fibronectin is essential for allergen‐induced airway fibrosis and hyperresponsiveness in mice. J. Allergy Clin. Immunol. 127, 439 – 446 e431 – 435.en_US
dc.identifier.citedreferenceArslan, F., Smeets, M. B., Riem Vis, P. W., Karper, J. C., Quax, P. H., et al. ( 2011 ) Lack of Fibronectin‐EDA Promotes Survival and Prevents Adverse Remodeling and Heart Function Deterioration After Myocardial Infarction. Circ. Res. 108, 582 – 592.en_US
dc.identifier.citedreferenceBalza, E., Borsi, L., Allemanni, G., and Zardi, L. ( 1988 ) Transforming growth factor beta regulates the levels of different fibronectin isoforms in normal human cultured fibroblasts. FEBS Lett. 228, 42 – 44.en_US
dc.identifier.citedreferenceMurphy‐Ullrich, J. E. and Poczatek, M. ( 2000 ) Activation of latent TGF‐beta by thrombospondin‐1: mechanisms and physiology. Cytokine Growth Factor Rev. 11, 59 – 69.en_US
dc.identifier.citedreferenceMunger, J. S., Huang, X., Kawakatsu, H., Griffiths, M. J., Dalton, S. L., et al. ( 1999 ) The integrin alpha v beta 6 binds and activates latent TGF beta 1: a mechanism for regulating pulmonary inflammation and fibrosis. Cell 96, 319 – 328.en_US
dc.identifier.citedreferenceBlakytny, R., Ludlow, A., Martin, G. E., Ireland, G., Lund, L. R., et al. ( 2004 ) Latent TGF‐beta1 activation by platelets. J. Cell. Physiol. 199, 67 – 76.en_US
dc.identifier.citedreferenceFontana, L., Chen, Y., Prijatelj, P., Sakai, T., Fassler, R., et al. ( 2005 ) Fibronectin is required for integrin alphavbeta6‐mediated activation of latent TGF‐beta complexes containing LTBP‐1. FASEB J. 19, 1798 – 1808.en_US
dc.identifier.citedreferencePeters, J. H., Maunder, R. J., Woolf, A. D., Cochrane, C. G., and Ginsberg, M. H. ( 1989 ) Elevated plasma levels of ED1+ (“cellular”) fibronectin in patients with vascular injury. J. Lab. Clin. Med. 113, 586 – 597.en_US
dc.identifier.citedreferenceCho, J. and Mosher, D. F. ( 2006 ) Enhancement of thrombogenesis by plasma fibronectin cross‐linked to fibrin and assembled in platelet thrombi. Blood 107, 3555 – 3563.en_US
dc.identifier.citedreferenceNi, H., Papalia, J. M., Degen, J. L., and Wagner, D. D. ( 2003 ) Control of thrombus embolization and fibronectin internalization by integrin alpha IIb beta 3 engagement of the fibrinogen gamma chain. Blood 102, 3609 – 3614.en_US
dc.identifier.citedreferenceMatuskova, J., Chauhan, A. K., Cambien, B., Astrof, S., Dole, V. S., et al. ( 2006 ) Decreased plasma fibronectin leads to delayed thrombus growth in injured arterioles. Arterioscler. Thromb. Vasc. Biol. 26, 1391 – 1396.en_US
dc.identifier.citedreferenceCastellanos, M., Leira, R., Serena, J., Blanco, M., Pedraza, S., et al. ( 2004 ) Plasma cellular‐fibronectin concentration predicts hemorrhagic transformation after thrombolytic therapy in acute ischemic stroke. Stroke 35, 1671 – 1676.en_US
dc.identifier.citedreferencevan der Plas, R. M., Schiphorst, M. E., Huizinga, E. G., Hene, R. J., Verdonck, L. F., et al. ( 1999 ) von Willebrand factor proteolysis is deficient in classic, but not in bone marrow transplantation‐associated, thrombotic thrombocytopenic purpura. Blood 93, 3798 – 3802.en_US
dc.identifier.citedreferenceVincent, P. A., Rebres, R. A., Lewis, E. P., Hurst, V. t., and Saba, T. M. ( 1993 ) Release of ED1 fibronectin from matrix of perfused lungs after vascular injury is independent of protein synthesis. Am. J. Physiol. 265, L485 – 492.en_US
dc.identifier.citedreferenceKanters, S. D., Banga, J. D., Algra, A., Frijns, R. C., Beutler, J. J., and Fijnheer, R. ( 2001 ) Plasma levels of cellular fibronectin in diabetes. Diabetes Care 24, 323 – 327.en_US
dc.identifier.citedreferenceChauhan, A. K., Kisucka, J., Cozzi, M. R., Walsh, M. T., Moretti, F. A., et al. ( 2008 ) Prothrombotic effects of fibronectin isoforms containing the EDA domain. Arterioscler. Thromb. Vasc. Biol. 28, 296 – 301.en_US
dc.identifier.citedreferenceTomasini‐Johansson, B. and Mosher, D. F. ( 2009 ) Plasma fibronectin concentration in inbred mouse strains. Thromb. Haemost. 102, 1278 – 1280.en_US
dc.identifier.citedreferencePecheniuk, N. M., Elias, D. J., Deguchi, H., Averell, P. M., and Griffin, J. H. ( 2008 ) Elevated plasma fibronectin levels associated with venous thromboembolism. Thromb. Haemost. 100, 224 – 228.en_US
dc.identifier.citedreferenceHynes, R. O. ( 2002 ) Integrins: bidirectional, allosteric signaling machines. Cell 110, 673 – 687.en_US
dc.identifier.citedreferenceChi‐Rosso, G., Gotwals, P. J., Yang, J., Ling, L., Jiang, K., et al. ( 1997 ) Fibronectin type III repeats mediate RGD‐independent adhesion and signaling through activated beta1 integrins. J. Biol. Chem. 272, 31447 – 31452.en_US
dc.identifier.citedreferenceKorom, S., Hancock, W. W., Coito, A. J., and Kupiec‐Weglinski, J. W. ( 1998 ) Blockade of very late antigen‐4 integrin binding to fibronectin in allograft recipients. II. Treatment with connecting segment‐1 peptides prevents chronic rejection by attenuating arteriosclerotic development and suppressing intragraft T cell and macrophage activation. Transplantation 65, 854 – 859.en_US
dc.identifier.citedreferenceLiao, Y. F., Gotwals, P. J., Koteliansky, V. E., Sheppard, D., and Van De Water, L. ( 2002 ) The EIIIA segment of fibronectin is a ligand for integrins alpha 9beta 1 and alpha 4beta 1 providing a novel mechanism for regulating cell adhesion by alternative splicing. J. Biol. Chem. 277, 14467 – 14474.en_US
dc.identifier.citedreferenceAmersi, F., Shen, X. D., Moore, C., Melinek, J., Busuttil, R. W., et al. ( 2003 ) Fibronectin‐alpha 4 beta 1 integrin‐mediated blockade protects genetically fat Zucker rat livers from ischemia/reperfusion injury. Am. J. Pathol. 162, 1229 – 1239.en_US
dc.identifier.citedreferenceJohansson, S., Svineng, G., Wennerberg, K., Armulik, A., and Lohikangas, L. ( 1997 ) Fibronectin–integrin interactions. Front. Biosci. 2, d126 – d146.en_US
dc.identifier.citedreferencePetrovic, A., Alpdogan, O., Willis, L. M., Eng, J. M., Greenberg, A. S., et al. ( 2004 ) LPAM (alpha 4 beta 7 integrin) is an important homing integrin on alloreactive T cells in the development of intestinal graft‐versus‐host disease. Blood 103, 1542 – 1547.en_US
dc.identifier.citedreferenceKohan, M., Muro, A. F., White, E. S., and Berkman, N. ( 2010 ) EDA‐containing cellular fibronectin induces fibroblast differentiation through binding to {alpha}4{beta}7 integrin receptor and MAPK/Erk 1/2‐dependent signaling. FASEB J. 24, 4503 – 4512.en_US
dc.identifier.citedreferenceSingh, P., Reimer, C. L., Peters, J. H., Stepp, M. A., Hynes, R. O., and Van De Water, L. ( 2004 ) The spatial and temporal expression patterns of integrin alpha9beta1 and one of its ligands, the EIIIA segment of fibronectin, in cutaneous wound healing. J. Invest. Dermatol. 123, 1176 – 1181.en_US
dc.identifier.citedreferenceOkamura, Y., Watari, M., Jerud, E. S., Young, D. W., Ishizaka, S. T., et al. ( 2001 ) The extra domain A of fibronectin activates Toll‐like receptor 4. J. Biol. Chem. 276, 10229 – 10233.en_US
dc.identifier.citedreferenceHollestelle, S. C., De Vries, M. R., Van Keulen, J. K., Schoneveld, A. H., Vink, A., et al. ( 2004 ) Toll‐like receptor 4 is involved in outward arterial remodeling. Circulation 109, 393 – 398.en_US
dc.identifier.citedreferenceLasarte, J. J., Casares, N., Gorraiz, M., Hervas‐Stubbs, S., Arribillaga, L., et al. ( 2007 ) The extra domain A from fibronectin targets antigens to TLR4‐expressing cells and induces cytotoxic T cell responses in vivo. J. Immunol. 178, 748 – 756.en_US
dc.identifier.citedreferenceGondokaryono, S. P., Ushio, H., Niyonsaba, F., Hara, M., Takenaka, H., et al. ( 2007 ) The extra domain A of fibronectin stimulates murine mast cells via toll‐like receptor 4. J. Leukoc. Biol. 82, 657 – 665.en_US
dc.identifier.citedreferenceLefebvre, J. S., Levesque, T., Picard, S., Pare, G., Gravel, A., et al. ( 2011 ) The extra domain a of fibronectin primes leukotriene biosynthesis and stimulates neutrophil migration through toll‐like receptor 4 activation. Arthritis Rheum.en_US
dc.identifier.citedreferenceDesmouliere, A., Chaponnier, C., and Gabbiani, G. ( 2005 ) Tissue repair,contraction, and the myofibroblast. Wound Repair Regen. 13, 7 – 12.en_US
dc.identifier.citedreferenceSerini, G., Bochaton‐Piallat, M. L., Ropraz, P., Geinoz, A., Borsi, L., et al. ( 1998 ) The fibronectin domain ED‐A is crucial for myofibroblastic phenotype induction by transforming growth factor‐beta1. J. Cell Biol. 142, 873 – 881.en_US
dc.identifier.citedreferenceThannickal, V. J., Lee, D. Y., White, E. S., Cui, Z., Larios, J. M., et al. ( 2003 ) Myofibroblast differentiation by transforming growth factor‐beta1 is dependent on cell adhesion and integrin signaling via focal adhesion kinase. J. Biol. Chem. 278, 12384 – 12389.en_US
dc.identifier.citedreferenceDing, Q., Gladson, C. L., Wu, H., Hayasaka, H., and Olman, M. A. ( 2008 ) Focal adhesion kinase (FAK)‐related non‐kinase inhibits myofibroblast differentiation through differential MAPK activation in a FAK‐dependent manner. J. Biol. Chem. 283, 26839 – 26849.en_US
dc.identifier.citedreferenceGreenberg, R. S., Bernstein, A. M., Benezra, M., Gelman, I. H., Taliana, L., and Masur, S. K. ( 2006 ) FAK‐dependent regulation of myofibroblast differentiation. FASEB J. 20, 1006 – 1008.en_US
dc.identifier.citedreferenceWhite, E. S., Atrasz, R. G., Hu, B., Phan, S. H., Stambolic, V., et al. ( 2006 ) Negative regulation of myofibroblast differentiation by PTEN (Phosphatase and Tensin Homolog Deleted on chromosome 10). Am. J. Respir. Crit. Care Med. 173, 112 – 121.en_US
dc.identifier.citedreferenceChan, M. W., Arora, P. D., Bozavikov, P., and McCulloch, C. A. ( 2009 ) FAK, PIP5KIgamma and gelsolin cooperatively mediate force‐induced expression of alpha‐smooth muscle actin. J. Cell Sci. 122, 2769 – 2781.en_US
dc.owningcollnameInterdisciplinary and Peer-Reviewed


Files in this item

Show simple item record

Remediation of Harmful Language

The University of Michigan Library aims to describe library materials in a way that respects the people and communities who create, use, and are represented in our collections. Report harmful or offensive language in catalog records, finding aids, or elsewhere in our collections anonymously through our metadata feedback form. More information at Remediation of Harmful Language.

Accessibility

If you are unable to use this file in its current format, please select the Contact Us link and we can modify it to make it more accessible to you.