Skip to main content

Biophysical Forces Modulate the Costamere and Z-Disc for Sarcomere Remodeling in Heart Failure

  • Chapter
  • First Online:
Biophysics of the Failing Heart

Abstract

Cardiac remodeling at the macroscopic level occurs as the heart fails and is accompanied by changes in size, shape, and performance of its cardiomyocytes. Therefore, an understanding of the mechanisms by which cardiomyocytes sense and respond to biomechanical stress is of prime importance to the development of a molecular basis of heart failure. The cardiomyocyte experiences changes in stress and strain throughout every cardiac cycle, which is generated both internally by contractile proteins and externally through cell–cell and cell–matrix interactions. Cells within the heart are constantly remodeling through processes of gene transcription, protein translation, posttranslational modification, and the assembly of complex organelles. Because the heart is constantly challenged mechanically, it relies upon the near instantaneous posttranslational modifications of existing proteins to sense and respond rapidly to acute external stressors. Transcription and translation takes longer but are important to the overall response to chronic stress and strain. However, the mechanisms by which cardiomyocytes sense and respond to chronic mechanical strains remain largely unknown. Our team and many others are testing hypotheses that mechanical forces drive the growth of heart cells in healthy exercise but become maladaptive in disease. In this chapter we review how the biophysical forces in normal cardiomyocytes drive sarcomere remodeling in heart failure via two key structural components—namely the costamere at the cell membrane that is connected to the Z-disc of the myofibril. Each of these structural complexes has numerous proteins, many of which sense mechanical forces and are also involved in filament assembly. A diagram of the costamere and Z-disc indicates many of the key proteins (Fig. 1). In this chapter, we review how the costamere and specific components of the Z-disc may accomplish both functions of stress detection and sarcomere remodeling during adaptation to hemodynamic overload.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Samarel, A. M. (2005). Costameres, focal adhesions, and cardiomyocyte mechanotransduction. American Journal of Physiology. Heart and Circulatory Physiology, 289, H2291–H2301.

    Google Scholar 

  2. Ganote, C. E., & Vander Heide, R. S. (1987). Cytoskeletal lesions in anoxic myocardial injury. A conventional and high-voltage electron-microscopic and immunofluorescence study. The American Journal of Pathology, 129, 327–344.

    Google Scholar 

  3. Ganote, C., & Armstrong, S. (1993). Ischaemia and the myocyte cytoskeleton: Review and speculation. Cardiovascular Research, 27, 1387–1403.

    Google Scholar 

  4. Hakim, Z. S., DiMichele, L. A., Rojas, M., Meredith, D., Mack, C. P., & Taylor, J. M. (2009). FAK regulates cardiomyocyte survival following ischemia/reperfusion. Journal of Molecular and Cellular Cardiology, 46, 241–248.

    Google Scholar 

  5. Wei, H., & Vander Heide, R. S. (2008). Heat stress activates AKT via focal adhesion kinase-mediated pathway in neonatal rat ventricular myocytes. American Journal of Physiology. Heart and Circulatory Physiology, 295, H561–H568.

    Google Scholar 

  6. Wei, H., & Vander Heide, R. S. (2010). Ischemic preconditioning and heat shock activate Akt via a focal adhesion kinase-mediated pathway in Langendorff-perfused adult rat hearts. American Journal of Physiology. Heart and Circulatory Physiology, 298, H152–H157.

    Google Scholar 

  7. 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. The Journal of Cell Biology, 153, 1175–1186.

    Google Scholar 

  8. Vogel, V., & Sheetz, M. P. (2009). Cell fate regulation by coupling mechanical cycles to biochemical signaling pathways. Current Opinion in Cell Biology, 21, 38–46.

    Google Scholar 

  9. Lammerding, J., Kamm, R. D., & Lee, R. T. (2004). Mechanotransduction in cardiac myocytes. Annals of the New York Academy of Sciences, 1015, 53–70.

    ADS  Google Scholar 

  10. Hoshijima, M. (2006). Mechanical stress–strain sensors embedded in cardiac cytoskeleton: Z disk, titin, and associated structures. American Journal of Physiology. Heart and Circulatory Physiology, 290, H1313–H1325.

    Google Scholar 

  11. Sharif-Naeini, R., Folgering, J. H., Bichet, D., Duprat, F., Delmas, P., Patel, A., et al. (2010). Sensing pressure in the cardiovascular system: Gq-coupled mechanoreceptors and TRP channels. Journal of Molecular and Cellular Cardiology, 48, 83–89.

    Google Scholar 

  12. Torsoni, A. S., Constancio, S. S., Nadruz, W., Jr., Hanks, S. K., & Franchini, K. G. (2003). Focal adhesion kinase is activated and mediates the early hypertrophic response to stretch in cardiac myocytes. Circulation Research, 93, 140–147.

    Google Scholar 

  13. Street, S. F. (1983). Lateral transmission of tension in frog myofibers: A myofibrillar network and transverse cytoskeletal connections are possible transmitters. Journal of Cellular Physiology, 114, 346–364.

    Google Scholar 

  14. Bloch, R. J., & Gonzalez-Serratos, H. (2003). Lateral force transmission across costameres in skeletal muscle. Exercise and Sport Sciences Reviews, 31, 73–78.

    Google Scholar 

  15. Collinsworth, A. M., Zhang, S., Kraus, W. E., & Truskey, G. A. (2002). Apparent elastic modulus and hysteresis of skeletal muscle cells throughout differentiation. American Journal of Physiology. Cell Physiology, 283, C1219–C1227.

    Google Scholar 

  16. Goldstein, M. A., Schroeter, J. P., & Michael, L. H. (1991). Role of the Z band in the mechanical properties of the heart. The FASEB Journal, 5, 2167–2174.

    Google Scholar 

  17. Kumar, A., Chaudhry, I., Reid, M. B., & Boriek, A. M. (2002). Distinct signaling pathways are activated in response to mechanical stress applied axially and transversely to skeletal muscle fibers. The Journal of Biological Chemistry, 277, 46493–46503.

    Google Scholar 

  18. Senyo, S. E., Koshman, Y. E., & Russell, B. (2007). Stimulus interval, rate and direction differentially regulate phosphorylation for mechanotransduction in neonatal cardiac myocytes. FEBS Letters, 581, 4241–4247.

    Google Scholar 

  19. Pardo, J. V., Siliciano, J. D., & Craig, S. W. (1983). Vinculin is a component of an extensive network of myofibril-sarcolemma attachment regions in cardiac muscle fibers. The Journal of Cell Biology, 97, 1081–1088.

    Google Scholar 

  20. Danowski, B. A., Imanaka-Yoshida, K., Sanger, J. M., & Sanger, J. W. (1992). Costameres are sites of force transmission to the substratum in adult rat cardiomyocytes. The Journal of Cell Biology, 118, 1411–1420.

    Google Scholar 

  21. Mansour, H., de Tombe, P. P., Samarel, A. M., & Russell, B. (2004). Restoration of resting sarcomere length after uniaxial static strain is regulated by protein kinase Cε and focal adhesion kinase. Circulation Research, 94, 642–649.

    Google Scholar 

  22. Michele, D. E., Kabaeva, Z., Davis, S. L., Weiss, R. M., & Campbell, K. P. (2009). Dystroglycan matrix receptor function in cardiac myocytes is important for limiting activity-induced myocardial damage. Circulation Research, 105, 984–993.

    Google Scholar 

  23. Li, R., Wu, Y., Manso, A. M., Gu, Y., Liao, P., Israeli, S., et al. (2012). β1 integrin gene excision in the adult murine cardiac myocyte causes defective mechanical and signaling responses. The American Journal of Pathology, 180, 952–962.

    Google Scholar 

  24. Ervasti, J. M. (2003). Costameres: The Achilles’ heel of Herculean muscle. The Journal of Biological Chemistry, 278, 13591–13594.

    Google Scholar 

  25. Elsherif, L., Huang, M. S., Shai, S. Y., Yang, Y., Li, R. Y., Chun, J., et al. (2008). Combined deficiency of dystrophin and β1 integrin in the cardiac myocyte causes myocardial dysfunction, fibrosis and calcification. Circulation Research, 102, 1109–1117.

    Google Scholar 

  26. Sussman, M. A., McCulloch, A., & Borg, T. K. (2002). Dance band on the Titanic: Biomechanical signaling in cardiac hypertrophy. Circulation Research, 91, 888–898.

    Google Scholar 

  27. Bershadsky, A. D., Balaban, N. Q., & Geiger, B. (2003). Adhesion-dependent cell mechanosensitivity. Annual Review of Cell and Developmental Biology, 19, 677–695.

    Google Scholar 

  28. Hynes, R. O. (2002). Integrins: Bidirectional, allosteric signaling machines. Cell, 110, 673–687.

    Google Scholar 

  29. Keller, R. S., Shai, S. Y., Babbitt, C. J., Pham, C. G., Solaro, R. J., Valencik, M. L., et al. (2001). Disruption of integrin function in the murine myocardium leads to perinatal lethality, fibrosis, and abnormal cardiac performance. The American Journal of Pathology, 158, 1079–1090.

    Google Scholar 

  30. Terracio, L., Rubin, K., Gullberg, D., Balog, E., Carver, W., Jyring, R., et al. (1991). Expression of collagen binding integrins during cardiac development and hypertrophy. Circulation Research, 68, 734–744.

    Google Scholar 

  31. Ross, R. S., & Borg, T. K. (2001). Integrins and the myocardium. Circulation Research, 88, 1112–1119.

    Google Scholar 

  32. Schwartz, M. A. (2009). Cell biology. The force is with us. Science, 323, 588–589.

    Google Scholar 

  33. Campbell, I. D., & Ginsberg, M. H. (2004). The talin-tail interaction places integrin activation on FERM ground. Trends in Biochemical Sciences, 29, 429–435.

    Google Scholar 

  34. Anthis, N. J., Wegener, K. L., Ye, F., Kim, C., Goult, B. T., Lowe, E. D., et al. (2009). The structure of an integrin/talin complex reveals the basis of inside-out signal transduction. The EMBO Journal, 28, 3623–3632.

    Google Scholar 

  35. Zemljic-Harpf, A., Manso, A. M., & Ross, R. S. (2009). Vinculin and talin: Focus on the myocardium. Journal of Investigative Medicine, 57, 849–855.

    Google Scholar 

  36. Senetar, M. A., Moncman, C. L., & McCann, R. O. (2007). Talin2 is induced during striated muscle differentiation and is targeted to stable adhesion complexes in mature muscle. Cell Motility and the Cytoskeleton, 64, 157–173.

    Google Scholar 

  37. Conti, F. J., Monkley, S. J., Wood, M. R., Critchley, D. R., & Muller, U. (2009). Talin 1 and 2 are required for myoblast fusion, sarcomere assembly and the maintenance of myotendinous junctions. Development (Cambridge, England), 136, 3597–3606.

    Google Scholar 

  38. Tadokoro, S., Shattil, S. J., Eto, K., Tai, V., Liddington, R. C., de Pereda, J. M., et al. (2003). Talin binding to integrin β tails: A final common step in integrin activation. Science, 302, 103–106.

    ADS  Google Scholar 

  39. Ye, F., Hu, G., Taylor, D., Ratnikov, B., Bobkov, A. A., McLean, M. A., et al. (2010). Recreation of the terminal events in physiological integrin activation. The Journal of Cell Biology, 188, 157–175.

    Google Scholar 

  40. Anthis, N. J., Haling, J. R., Oxley, C. L., Memo, M., Wegener, K. L., Lim, C. J., et al. (2009). β-integrin tyrosine phosphorylation is a conserved mechanism for regulating talin-induced integrin activation. The Journal of Biological Chemistry, 284, 36700–36710.

    Google Scholar 

  41. Dabiri, G. A., Turnacioglu, K. K., Sanger, J. M., & Sanger, J. W. (1997). Myofibrillogenesis visualized in living embryonic cardiomyocytes. Proceedings of the National Academy of Sciences of the United States of America, 94, 9493–9498.

    ADS  Google Scholar 

  42. Stout, A. L., Wang, J., Sanger, J. M., & Sanger, J. W. (2008). Tracking changes in Z-band organization during myofibrillogenesis with FRET imaging. Cell Motility and the Cytoskeleton, 65, 353–367.

    Google Scholar 

  43. Dumbauld, D. W., Michael, K. E., Hanks, S. K., & Garcia, A. J. (2011). Focal adhesion kinase-dependent regulation of adhesive forces involves vinculin recruitment to focal adhesions. Biology of the Cell, 102, 203–213.

    Google Scholar 

  44. Humphries, J. D., Wang, P., Streuli, C., Geiger, B., Humphries, M. J., & Ballestrem, C. (2007). Vinculin controls focal adhesion formation by direct interactions with talin and actin. The Journal of Cell Biology, 179, 1043–1057.

    Google Scholar 

  45. Brown, M. C., Perrotta, J. A., & Turner, C. E. (1996). Identification of LIM3 as the principal determinant of paxillin focal adhesion localization and characterization of a novel motif on paxillin directing vinculin and focal adhesion kinase binding. The Journal of Cell Biology, 135, 1109–1123.

    Google Scholar 

  46. Hayashi, I., Vuori, K., & Liddington, R. C. (2002). The focal adhesion targeting (FAT) region of focal adhesion kinase is a four-helix bundle that binds paxillin. Nature Structural Biology, 9, 101–106.

    Google Scholar 

  47. Kanchanawong, P., Shtengel, G., Pasapera, A. M., Ramko, E. B., Davidson, M. W., Hess, H. F., et al. (2010). Nanoscale architecture of integrin-based cell adhesions. Nature, 468, 580–584.

    ADS  Google Scholar 

  48. Kuppuswamy, D., Kerr, C., Narishige, T., Kasi, V. S., Menick, D. R., & Cooper, G. T. (1997). Association of tyrosine-phosphorylated c-Src with the cytoskeleton of hypertrophying myocardium. The Journal of Biological Chemistry, 272, 4500–4508.

    Google Scholar 

  49. Laser, M., Willey, C. D., Jiang, W., Cooper, G., Menick, D. R., Zile, M. R., et al. (2000). Integrin activation and focal complex formation in cardiac hypertrophy. The Journal of Biological Chemistry, 275, 35624–35630.

    Google Scholar 

  50. Domingos, P. P., Fonseca, P. M., Nadruz, W., Jr., & Franchini, K. G. (2002). Load-induced focal adhesion kinase activation in the myocardium: Role of stretch and contractile activity. American Journal of Physiology. Heart and Circulatory Physiology, 282, H556–H564.

    Google Scholar 

  51. Bayer, A. L., Heidkamp, M. C., Patel, N., Porter, M. J., Engman, S. J., & Samarel, A. M. (2002). PYK2 expression and phosphorylation increases in pressure overload-induced left ventricular hypertrophy. American Journal of Physiology. Heart and Circulatory Physiology, 283, H695–H706.

    Google Scholar 

  52. Torsoni, A. S., Fonseca, P. M., Crosara-Alberto, D. P., & Franchini, K. G. (2003). Early activation of p160ROCK by pressure overload in rat heart. American Journal of Physiology. Cell Physiology, 284, C1411–C1419.

    Google Scholar 

  53. Ross, R. S., Pham, C., Shai, S. Y., Goldhaber, J. I., Fenczik, C., Glembotski, C. C., et al. (1998). β1 Integrins participate in the hypertrophic response of rat ventricular myocytes. Circulation Research, 82, 1160–1172.

    Google Scholar 

  54. Eble, D. M., Strait, J. B., Govindarajan, G., Lou, J., Byron, K. L., & Samarel, A. M. (2000). Endothelin-induced cardiac myocyte hypertrophy: Role for focal adhesion kinase. American Journal of Physiology. Heart and Circulatory Physiology, 278, H1695–H1707.

    Google Scholar 

  55. Taylor, J. M., Rovin, J. D., & Parsons, J. T. (2000). A role for focal adhesion kinase in phenylephrine-induced hypertrophy of rat ventricular cardiomyocytes. The Journal of Biological Chemistry, 275, 19250–19257.

    Google Scholar 

  56. Pham, C. G., Harpf, A. E., Keller, R. S., Vu, H. T., Shai, S. Y., Loftus, J. C., et al. (2000). Striated muscle-specific β1D-integrin and FAK are involved in cardiac myocyte hypertrophic response pathway. American Journal of Physiology. Heart and Circulatory Physiology, 279, H2916–H2926.

    Google Scholar 

  57. Kovacic-Milivojevic, B., Roediger, F., Almeida, E. A., Damsky, C. H., Gardner, D. G., & Ilic, D. (2001). Focal adhesion kinase and p130Cas mediate both sarcomeric organization and activation of genes associated with cardiac myocyte hypertrophy. Molecular Biology of the Cell, 12, 2290–2307.

    Google Scholar 

  58. Heidkamp, M. C., Bayer, A. L., Kalina, J. A., Eble, D. M., & Samarel, A. M. (2002). GFP-FRNK disrupts focal adhesions and induces anoikis in neonatal rat ventricular myocytes. Circulation Research, 90, 1282–1289.

    Google Scholar 

  59. Heidkamp, M. C., Bayer, A. L., Scully, B. T., Eble, D. M., & Samarel, A. M. (2003). Activation of focal adhesion kinase by protein kinase Cε in neonatal rat ventricular myocytes. American Journal of Physiology. Heart and Circulatory Physiology, 285, H1684–H1696.

    Google Scholar 

  60. Wang, N., Butler, J. P., & Ingber, D. E. (1993). Mechanotransduction across the cell surface and through the cytoskeleton. Science, 260, 1124–1127.

    ADS  Google Scholar 

  61. Schlaepfer, D. D., Mitra, S. K., & Ilic, D. (2004). Control of motile and invasive cell phenotypes by focal adhesion kinase. Biochimica et Biophysica Acta, 1692, 77–102.

    Google Scholar 

  62. Schaller, M. D., Hildebrand, J. D., Shannon, J. D., Fox, J. W., Vines, R. R., & Parsons, J. T. (1994). Autophosphorylation of the focal adhesion kinase, pp125FAK, directs SH2-dependent binding of pp60src. Molecular and Cellular Biology, 14, 1680–1688.

    Google Scholar 

  63. Frame, M. C., Patel, H., Serrels, B., Lietha, D., & Eck, M. J. (2010). The FERM domain: Organizing the structure and function of FAK. Nature Reviews. Molecular Cell Biology, 11, 802–814.

    Google Scholar 

  64. Lim, Y., Han, I., Jeon, J., Park, H., Bahk, Y. Y., & Oh, E. S. (2004). Phosphorylation of focal adhesion kinase at tyrosine 861 is crucial for Ras transformation of fibroblasts. The Journal of Biological Chemistry, 279, 29060–29065.

    Google Scholar 

  65. Yi, X. P., Zhou, J., Huber, L., Qu, J., Wang, X., Gerdes, A. M., et al. (2006). Nuclear compartmentalization of FAK and FRNK in cardiac myocytes. American Journal of Physiology. Heart and Circulatory Physiology, 290, H2509–H2515.

    Google Scholar 

  66. Chu, M., Iyengar, R., Koshman, Y. E., Kim, T., Russell, B., Martin, J. L., et al. (2011). Serine-910 phosphorylation of focal adhesion kinase is critical for sarcomere reorganization in cardiomyocyte hypertrophy. Cardiovascular Research, 92, 409–419.

    Google Scholar 

  67. Seko, Y., Takahashi, N., Tobe, K., Kadowaki, T., & Yazaki, Y. (1999). Pulsatile stretch activates mitogen-activated protein kinase (MAPK) family members and focal adhesion kinase (p125FAK) in cultured rat cardiac myocytes. Biochemical and Biophysical Research Communications, 259, 8–14.

    Google Scholar 

  68. Eble, D. M., Qi, M., Strait, J., & Samarel, A. M. (2000). Contraction-dependent hypertrophy of neonatal rat ventricular myocytes: Potential role for focal adhesion kinase. In N. Takeda & N. S. Dhalla (Eds.), The hypertrophied heart (pp. 91–107). Boston: Kluwer.

    Google Scholar 

  69. Bayer, A. L., Ferguson, A. G., Lucchesi, P. A., & Samarel, A. M. (2001). Pyk2 expression and phosphorylation in neonatal and adult cardiomyocytes. Journal of Molecular and Cellular Cardiology, 33, 1017–1030.

    Google Scholar 

  70. Ilic, D., Furuta, Y., Kanazawa, S., Takeda, N., Sobue, K., Nakatsuji, N., et al. (1995). Reduced cell motility and enhanced focal adhesion contact formation in cells from FAK-deficient mice. Nature, 377, 539–544.

    ADS  Google Scholar 

  71. Ilic, D., Kovacic, B., McDonagh, S., Jin, F., Baumbusch, C., Gardner, D. G., et al. (2003). Focal adhesion kinase is required for blood vessel morphogenesis. Circulation Research, 92, 300–307.

    Google Scholar 

  72. Hakim, Z. S., DiMichele, L. A., Doherty, J. T., Homeister, J. W., Beggs, H. E., Reichardt, L. F., et al. (2007). Conditional deletion of focal adhesion kinase leads to defects in ventricular septation and outflow tract alignment. Molecular and Cellular Biology, 27, 5352–5364.

    Google Scholar 

  73. Peng, X., Kraus, M. S., Wei, H., Shen, T. L., Pariaut, R., Alcaraz, A., et al. (2006). Inactivation of focal adhesion kinase in cardiomyocytes promotes eccentric cardiac hypertrophy and fibrosis in mice. The Journal of Clinical Investigation, 116, 217–227.

    Google Scholar 

  74. DiMichele, L. A., Doherty, J. T., Rojas, M., Beggs, H. E., Reichardt, L. F., Mack, C. P., et al. (2006). Myocyte-restricted focal adhesion kinase deletion attenuates pressure overload-induced hypertrophy. Circulation Research, 99, 636–645.

    Google Scholar 

  75. Nadruz, W., Jr., Corat, M. A., Marin, T. M., Guimaraes Pereira, G. A., & Franchini, K. G. (2005). Focal adhesion kinase mediates MEF2 and c-Jun activation by stretch: Role in the activation of the cardiac hypertrophic genetic program. Cardiovascular Research, 68, 87–97.

    Google Scholar 

  76. Avraham, H., Park, S. Y., Schinkmann, K., & Avraham, S. (2000). RAFTK/Pyk2-mediated cellular signalling. Cellular Signalling, 12, 123–133.

    Google Scholar 

  77. Park, S. Y., Avraham, H. K., & Avraham, S. (2004). RAFTK/Pyk2 activation is mediated by trans-acting autophosphorylation in a Src-independent manner. The Journal of Biological Chemistry, 279, 33315–33322.

    Google Scholar 

  78. Melendez, J., Welch, S., Schaefer, E., Moravec, C. S., Avraham, S., Avraham, H., et al. (2002). Activation of Pyk2/related focal adhesion tyrosine kinase and focal adhesion kinase in cardiac remodeling. The Journal of Biological Chemistry, 277, 45203–45210.

    Google Scholar 

  79. Kodama, H., Fukuda, K., Takahashi, T., Sano, M., Kato, T., Tahara, S., et al. (2002). Role of EGF receptor and Pyk2 in endothelin-1-induced ERK activation in rat cardiomyocytes. Journal of Molecular and Cellular Cardiology, 34, 139–150.

    Google Scholar 

  80. Bayer, A. L., Heidkamp, M. C., Howes, A. L., Heller Brown, J., Byron, K. L., & Samarel, A. M. (2003). Protein kinase Cε-dependent activation of proline-rich tyrosine kinase 2 in neonatal rat ventricular myocytes. Journal of Molecular and Cellular Cardiology, 35, 1121–1133.

    Google Scholar 

  81. Melendez, J., Turner, C., Avraham, H., Steinberg, S. F., Schaefer, E., & Sussman, M. A. (2004). Cardiomyocyte apoptosis triggered by RAFTK/pyk2 via Src kinase is antagonized by paxillin. The Journal of Biological Chemistry, 279, 53516–53523.

    Google Scholar 

  82. Hirotani, S., Higuchi, Y., Nishida, K., Nakayama, H., Yamaguchi, O., Hikoso, S., et al. (2004). Ca(2+)-sensitive tyrosine kinase Pyk2/CAKβ-dependent signaling is essential for G-protein-coupled receptor agonist-induced hypertrophy. Journal of Molecular and Cellular Cardiology, 36, 799–807.

    Google Scholar 

  83. Heidkamp, M. C., Scully, B. T., Vijayan, K., Engman, S. J., Szotek, E. L., & Samarel, A. M. (2005). PYK2 regulates SERCA2 gene expression in neonatal rat ventricular myocytes. American Journal of Physiology. Cell Physiology, 289, C471–C482.

    Google Scholar 

  84. Blaukat, A., Ivankovic-Dikic, I., Gronroos, E., Dolfi, F., Tokiwa, G., Vuori, K., et al. (1999). Adaptor proteins Grb2 and Crk couple Pyk2 with activation of specific mitogen-activated protein kinase cascades. The Journal of Biological Chemistry, 274, 14893–14901.

    Google Scholar 

  85. Pandey, P., Avraham, S., Kumar, S., Nakazawa, A., Place, A., Ghanem, L., et al. (1999). Activation of p38 mitogen-activated protein kinase by PYK2/related adhesion focal tyrosine kinase-dependent mechanism. The Journal of Biological Chemistry, 274, 10140–10144.

    Google Scholar 

  86. Tokiwa, G., Dikic, I., Lev, S., & Schlessinger, J. (1996). Activation of Pyk2 by stress signals and coupling with JNK signaling pathway. Science, 273, 792–794.

    ADS  Google Scholar 

  87. Hart, D. L., Heidkamp, M. C., Iyengar, R., Vijayan, K., Szotek, E. L., Barakat, J. A., et al. (2008). CRNK gene transfer improves function and reverses the myosin heavy chain isoenzyme switch during post-myocardial infarction left ventricular remodeling. Journal of Molecular and Cellular Cardiology, 45, 93–105.

    Google Scholar 

  88. Fukuda, K., Gupta, S., Chen, K., Wu, C., & Qin, J. (2009). The pseudoactive site of ILK is essential for its binding to α-parvin and localization to focal adhesions. Molecular Cell, 36, 819–830.

    Google Scholar 

  89. Wickstrom, S. A., Lange, A., Montanez, E., & Fassler, R. (2010). The ILK/PINCH/parvin complex: The kinase is dead, long live the pseudokinase! The EMBO Journal, 29, 281–291.

    Google Scholar 

  90. Hannigan, G. E., Leung-Hagesteijn, C., Fitz-Gibbon, L., Coppolino, M. G., Radeva, G., Filmus, J., et al. (1996). Regulation of cell adhesion and anchorage-dependent growth by a new β1-integrin-linked protein kinase. Nature, 379, 91–96.

    ADS  Google Scholar 

  91. Wu, C., & Dedhar, S. (2001). Integrin-linked kinase (ILK) and its interactors: A new paradigm for the coupling of extracellular matrix to actin cytoskeleton and signaling complexes. The Journal of Cell Biology, 155, 505–510.

    Google Scholar 

  92. Liang, X., Zhou, Q., Li, X., Sun, Y., Lu, M., Dalton, N., et al. (2005). PINCH1 plays an essential role in early murine embryonic development but is dispensable in ventricular cardiomyocytes. Molecular and Cellular Biology, 25, 3056–3062.

    Google Scholar 

  93. White, D. E., Coutu, P., Shi, Y. F., Tardif, J. C., Nattel, S., St Arnaud, R., et al. (2006). Targeted ablation of ILK from the murine heart results in dilated cardiomyopathy and spontaneous heart failure. Genes & Development, 20, 2355–2360.

    Google Scholar 

  94. Persad, S., Attwell, S., Gray, V., Mawji, N., Deng, J. T., Leung, D., et al. (2001). Regulation of protein kinase B/Akt-serine 473 phosphorylation by integrin-linked kinase: Critical roles for kinase activity and amino acids arginine 211 and serine 343. The Journal of Biological Chemistry, 276, 27462–27469.

    Google Scholar 

  95. Attwell, S., Roskelley, C., & Dedhar, S. (2000). The integrin-linked kinase (ILK) suppresses anoikis. Oncogene, 19, 3811–3815.

    Google Scholar 

  96. Bock-Marquette, I., Saxena, A., White, M. D., Dimaio, J. M., & Srivastava, D. (2004). Thymosin β4 activates integrin-linked kinase and promotes cardiac cell migration, survival and cardiac repair. Nature, 432, 466–472.

    ADS  Google Scholar 

  97. Disatnik, M. H., Buraggi, G., & Mochly-Rosen, D. (1994). Localization of protein kinase C isozymes in cardiac myocytes. Experimental Cell Research, 210, 287–297.

    Google Scholar 

  98. Huang, X. P., Pi, Y., Lokuta, A. J., Greaser, M. L., & Walker, J. W. (1997). Arachidonic acid stimulates protein kinase C-ε redistribution in heart cells. Journal of Cell Science, 110(Pt 14), 1625–1634.

    Google Scholar 

  99. Borg, T. K., Goldsmith, E. C., Price, R., Carver, W., Terracio, L., & Samarel, A. M. (2000). Specialization at the Z line of cardiac myocytes. Cardiovascular Research, 46, 277–285.

    Google Scholar 

  100. Ping, P., Zhang, J., Pierce, W. M., Jr., & Bolli, R. (2001). Functional proteomic analysis of protein kinase Cε signaling complexes in the normal heart and during cardioprotection. Circulation Research, 88, 59–62.

    Google Scholar 

  101. Vondriska, T. M., Zhang, J., Song, C., Tang, X. L., Cao, X., Baines, C. P., et al. (2001). Protein kinase Cε-Src modules direct signal transduction in nitric oxide-induced cardioprotection: Complex formation as a means for cardioprotective signaling. Circulation Research, 88, 1306–1313.

    Google Scholar 

  102. Ping, P., Song, C., Zhang, J., Guo, Y., Cao, X., Li, R. C., et al. (2002). Formation of protein kinase Cε-Lck signaling modules confers cardioprotection. The Journal of Clinical Investigation, 109, 499–507.

    Google Scholar 

  103. Song, C., Vondriska, T. M., Wang, G. W., Klein, J. B., Cao, X., Zhang, J., et al. (2002). Molecular conformation dictates signaling module formation: Example of PKCε and Src tyrosine kinase. American Journal of Physiology. Heart and Circulatory Physiology, 282, H1166–H1171.

    Google Scholar 

  104. Heidkamp, M. C., Iyengar, R., Szotek, E. L., Cribbs, L. L., & Samarel, A. M. (2007). Protein kinase Cε-dependent MARCKS phosphorylation in neonatal and adult rat ventricular myocytes. Journal of Molecular and Cellular Cardiology, 42, 422–431.

    Google Scholar 

  105. Strait, J. B., Martin, J. L., Bayer, A., Mestril, R., Eble, D. M., & Samarel, A. M. (2001). Role of protein kinase Cε in hypertrophy of cultured neonatal rat ventricular myocytes. American Journal of Physiology. Heart and Circulatory Physiology, 280, H756–H766.

    Google Scholar 

  106. Liliental, J., & Chang, D. D. (1998). RACK1, a receptor for activated protein kinase C, interacts with integrin β subunit. The Journal of Biological Chemistry, 273, 2379–2383.

    Google Scholar 

  107. Mochly-Rosen, D., Khaner, H., & Lopez, J. (1991). Identification of intracellular receptor proteins for activated protein kinase C. Proceedings of the National Academy of Sciences of the United States of America, 88, 3997–4000.

    ADS  Google Scholar 

  108. Chang, B. Y., Conroy, K. B., Machleder, E. M., & Cartwright, C. A. (1998). RACK1, a receptor for activated C kinase and a homolog of the β subunit of G proteins, inhibits activity of Src tyrosine kinases and growth of NIH 3T3 cells. Molecular and Cellular Biology, 18, 3245–3256.

    Google Scholar 

  109. Schechtman, D., & Mochly-Rosen, D. (2001). Adaptor proteins in protein kinase C-mediated signal transduction. Oncogene, 20, 6339–6347.

    Google Scholar 

  110. Besson, A., Wilson, T. L., & Yong, V. W. (2002). The anchoring protein RACK1 links protein kinase Cε to integrin β chains. Requirements for adhesion and motility. The Journal of Biological Chemistry, 277, 22073–22084.

    Google Scholar 

  111. Berrier, A. L., Mastrangelo, A. M., Downward, J., Ginsberg, M., & LaFlamme, S. E. (2000). Activated R-Ras, Rac1, PI 3-kinase and PKCε can each restore cell spreading inhibited by isolated integrin β1 cytoplasmic domains. The Journal of Cell Biology, 151, 1549–1560.

    Google Scholar 

  112. Robia, S. L., Ghanta, J., Robu, V. G., & Walker, J. W. (2001). Localization and kinetics of protein kinase Cε anchoring in cardiac myocytes. Biophysical Journal, 80, 2140–2151.

    ADS  Google Scholar 

  113. Huang, X., & Walker, J. W. (2004). Myofilament anchoring of protein kinase C-ε in cardiac myocytes. Journal of Cell Science, 117, 1971–1978.

    Google Scholar 

  114. Vuori, K., & Ruoslahti, E. (1993). Activation of protein kinase C precedes α5β1 integrin-mediated cell spreading on fibronectin. The Journal of Biological Chemistry, 268, 21459–21462.

    Google Scholar 

  115. Parsons, J. T. (2003). Focal adhesion kinase: The first ten years. Journal of Cell Science, 116, 1409–1416.

    ADS  Google Scholar 

  116. Ruwhof, C., van Wamel, J. T., Noordzij, L. A., Aydin, S., Harper, J. C., & van der Laarse, A. (2001). Mechanical stress stimulates phospholipase C activity and intracellular calcium ion levels in neonatal rat cardiomyocytes. Cell Calcium, 29, 73–83.

    Google Scholar 

  117. Decker, M. L., Simpson, D. G., Behnke, M., Cook, M. G., & Decker, R. S. (1990). Morphological analysis of contracting and quiescent adult rabbit cardiac myocytes in long-term culture. The Anatomical Record, 227, 285–299.

    Google Scholar 

  118. Sharp, W. W., Simpson, D. G., Borg, T. K., Samarel, A. M., & Terracio, L. (1997). Mechanical forces regulate focal adhesion and costamere assembly in cardiac myocytes. The American Journal of Physiology, 273, H546–H556.

    Google Scholar 

  119. Peng, X., Wu, X., Druso, J. E., Wei, H., Park, A. Y., Kraus, M. S., et al. (2008). Cardiac developmental defects and eccentric right ventricular hypertrophy in cardiomyocyte focal adhesion kinase (FAK) conditional knockout mice. Proceedings of the National Academy of Sciences of the United States of America, 105, 6638–6643.

    ADS  Google Scholar 

  120. DiMichele, L. A., Hakim, Z. S., Sayers, R. L., Rojas, M., Schwartz, R. J., Mack, C. P., et al. (2009). Transient expression of FRNK reveals stage-specific requirement for focal adhesion kinase activity in cardiac growth. Circulation Research, 104, 1201–1208.

    Google Scholar 

  121. Quach, N. L., & Rando, T. A. (2006). Focal adhesion kinase is essential for costamerogenesis in cultured skeletal muscle cells. Developmental Biology, 293, 38–52.

    Google Scholar 

  122. Hunger-Glaser, I., Salazar, E. P., Sinnett-Smith, J., & Rozengurt, E. (2003). Bombesin, lysophosphatidic acid, and epidermal growth factor rapidly stimulate focal adhesion kinase phosphorylation at Ser-910: Requirement for ERK activation. The Journal of Biological Chemistry, 278, 22631–22643.

    Google Scholar 

  123. Subauste, M. C., Pertz, O., Adamson, E. D., Turner, C. E., Junger, S., & Hahn, K. M. (2004). Vinculin modulation of paxillin-FAK interactions regulates ERK to control survival and motility. The Journal of Cell Biology, 165, 371–381.

    Google Scholar 

  124. Zemljic-Harpf, A. E., Miller, J. C., Henderson, S. A., Wright, A. T., Manso, A. M., Elsherif, L., et al. (2007). Cardiac-myocyte-specific excision of the vinculin gene disrupts cellular junctions, causing sudden death or dilated cardiomyopathy. Molecular and Cellular Biology, 27, 7522–7537.

    Google Scholar 

  125. Hagel, M., George, E. L., Kim, A., Tamimi, R., Opitz, S. L., Turner, C. E., et al. (2002). The adaptor protein paxillin is essential for normal development in the mouse and is a critical transducer of fibronectin signaling. Molecular and Cellular Biology, 22, 901–915.

    Google Scholar 

  126. Honda, H., Oda, H., Nakamoto, T., Honda, Z., Sakai, R., Suzuki, T., et al. (1998). Cardiovascular anomaly, impaired actin bundling and resistance to Src-induced transformation in mice lacking p130Cas. Nature Genetics, 19, 361–365.

    Google Scholar 

  127. Goldstein, M. A., Michael, L. H., Schroeter, J. P., & Sass, R. L. (1988). Structural states in the Z band of skeletal muscle correlate with states of active and passive tension. The Journal of General Physiology, 92, 113–119.

    Google Scholar 

  128. Knoll, R., Hoshijima, M., & Chien, K. (2003). Cardiac mechanotransduction and implications for heart disease. Journal of Molecular Medicine, 81, 750–756.

    Google Scholar 

  129. Pyle, W. G., & Solaro, R. J. (2004). At the crossroads of myocardial signaling: The role of Z-discs in intracellular signaling and cardiac function. Circulation Research, 94, 296–305.

    Google Scholar 

  130. Knoll, R., Buyandelger, B., & Lab, M. (2011). The sarcomeric Z-disc and Z-discopathies. Journal of Biomedicine & Biotechnology, 2011, 569628.

    Google Scholar 

  131. Buyandelger, B., Ng, K. E., Miocic, S., Piotrowska, I., Gunkel, S., Ku, C. H., et al. (2011). MLP (muscle LIM protein) as a stress sensor in the heart. Pflügers Archiv, 462, 135–142.

    Google Scholar 

  132. Boateng, S. Y., Senyo, S. E., Qi, L., Goldspink, P. H., & Russell, B. (2009). Myocyte remodeling in response to hypertrophic stimuli requires nucleocytoplasmic shuttling of muscle LIM protein. Journal of Molecular and Cellular Cardiology, 47, 426–435.

    Google Scholar 

  133. Sanger, J. W., Wang, J., Fan, Y., White, J., & Sanger, J. M. (2010). Assembly and dynamics of myofibrils. Journal of Biomedicine & Biotechnology, 2010, 858606.

    Google Scholar 

  134. Zhu, W., Zou, Y., Shiojima, I., Kudoh, S., Aikawa, R., Hayashi, D., et al. (2000). Ca2+/calmodulin-dependent kinase II and calcineurin play critical roles in endothelin-1-induced cardiomyocyte hypertrophy. The Journal of Biological Chemistry, 275, 15239–15245.

    Google Scholar 

  135. Pyle, W. G., Hart, M. C., Cooper, J. A., Sumandea, M. P., de Tombe, P. P., & Solaro, R. J. (2002). Actin capping protein: An essential element in protein kinase signaling to the myofilaments. Circulation Research, 90, 1299–1306.

    Google Scholar 

  136. Yu, J. G., & Russell, B. (2005). Cardiomyocyte remodeling and sarcomere addition after uniaxial static strain in vitro. The Journal of Histochemistry and Cytochemistry, 53, 839–844.

    Google Scholar 

  137. Hartman, T. J., Martin, J. L., Solaro, R. J., Samarel, A. M., & Russell, B. (2009). CapZ dynamics are altered by endothelin-1 and phenylephrine via PIP2- and PKC-dependent mechanisms. American Journal of Physiology. Cell Physiology, 296, C1034–C1039.

    Google Scholar 

  138. Maruyama, K., & Obinata, T. (1965). Presence of β-actinin in the soluble fraction of the muscle cells of the chick embryo. Journal of Biochemistry, 57, 575–577.

    Google Scholar 

  139. Maruyama, K. (1966). Effect of β-actinin on the particle length of F-actin. Biochimica et Biophysica Acta, 126, 389–398.

    Google Scholar 

  140. Maruyama, K., Kimura, S., Ishi, T., Kuroda, M., & Ohashi, K. (1977). β-actinin, a regulatory protein of muscle. Purification, characterization and function. Journal of Biochemistry, 81, 215–232.

    Google Scholar 

  141. Asakura, S. (1961). F-actin adenosine triphosphatase activated under sonic vibration. Biochimica et Biophysica Acta, 52, 65–75.

    Google Scholar 

  142. Isenberg, G., Aebi, U., & Pollard, T. D. (1980). An actin-binding protein from Acanthamoeba regulates actin filament polymerization and interactions. Nature, 288, 455–459.

    ADS  Google Scholar 

  143. Hart, M. C., Korshunova, Y. O., & Cooper, J. A. (1997). Vertebrates have conserved capping protein α isoforms with specific expression patterns. Cell Motility and the Cytoskeleton, 38, 120–132.

    Google Scholar 

  144. Hug, C., Miller, T. M., Torres, M. A., Casella, J. F., & Cooper, J. A. (1992). Identification and characterization of an actin-binding site of CapZ. The Journal of Cell Biology, 116, 923–931.

    Google Scholar 

  145. Schafer, D. A., Korshunova, Y. O., Schroer, T. A., & Cooper, J. A. (1994). Differential localization and sequence analysis of capping protein β-subunit isoforms of vertebrates. The Journal of Cell Biology, 127, 453–465.

    Google Scholar 

  146. von Bulow, M., Rackwitz, H. R., Zimbelmann, R., & Franke, W. W. (1997). CP β3, a novel isoform of an actin-binding protein, is a component of the cytoskeletal calyx of the mammalian sperm head. Experimental Cell Research, 233, 216–224.

    Google Scholar 

  147. Hart, M. C., & Cooper, J. A. (1999). Vertebrate isoforms of actin capping protein β have distinct functions in vivo. The Journal of Cell Biology, 147, 1287–1298.

    Google Scholar 

  148. Yamashita, A., Maeda, K., & Maeda, Y. (2003). Crystal structure of CapZ: Structural basis for actin filament barbed end capping. The EMBO Journal, 22, 1529–1538.

    Google Scholar 

  149. Barron-Casella, E. A., Torres, M. A., Scherer, S. W., Heng, H. H., Tsui, L. C., & Casella, J. F. (1995). Sequence analysis and chromosomal localization of human Cap Z. Conserved residues within the actin-binding domain may link Cap Z to gelsolin/severin and profilin protein families. The Journal of Biological Chemistry, 270, 21472–21479.

    Google Scholar 

  150. Wear, M. A., Yamashita, A., Kim, K., Maeda, Y., & Cooper, J. A. (2003). How capping protein binds the barbed end of the actin filament. Current Biology, 13, 1531–1537.

    Google Scholar 

  151. Kim, K., McCully, M. E., Bhattacharya, N., Butler, B., Sept, D., & Cooper, J. A. (2007). Structure/function analysis of the interaction of phosphatidylinositol 4,5-bisphosphate with actin-capping protein: Implications for how capping protein binds the actin filament. The Journal of Biological Chemistry, 282, 5871–5879.

    Google Scholar 

  152. Wear, M. A., & Cooper, J. A. (2004). Capping protein binding to S100B: Implications for the tentacle model for capping the actin filament barbed end. The Journal of Biological Chemistry, 279, 14382–14390.

    Google Scholar 

  153. Takeda, S., Minakata, S., Koike, R., Kawahata, I., Narita, A., Kitazawa, M., et al. (2010). Two distinct mechanisms for actin capping protein regulation-steric and allosteric inhibition. PLoS Biology, 8, e1000416.

    Google Scholar 

  154. Kim, T., Cooper, J. A., & Sept, D. (2010). The interaction of capping protein with the barbed end of the actin filament. Journal of Molecular Biology, 404, 794–802.

    Google Scholar 

  155. Narita, A., Takeda, S., Yamashita, A., & Maeda, Y. (2006). Structural basis of actin filament capping at the barbed-end: A cryo-electron microscopy study. The EMBO Journal, 25, 5626–5633.

    Google Scholar 

  156. Dorn, G. W., Souroujon, M. C., Liron, T., Chen, C. H., Gray, M. O., Zhou, H. Z., et al. (1999). Sustained in vivo cardiac protection by a rationally designed peptide that causes epsilon protein kinase C translocation. Proceedings of the National Academy of Sciences of the United States of America, 96, 12798–12803.

    ADS  Google Scholar 

  157. Swanson, E. R., Warren, C. M., Solaro, R. J., Samarel, A. M., & Russell, B. (2011). Cyclic mechanical strain alters CapZ β1 but not CapZ β2 dynamics and phosphorylation via PKCε-dependent mechanisms. Journal of Molecular and Cellular Cardiology, 51, S23 (abstract).

    Google Scholar 

  158. Rybin, V. O., & Steinberg, S. F. (1994). Protein kinase C isoform expression and regulation in the developing rat heart. Circulation Research, 74, 299–309.

    Google Scholar 

  159. Prekeris, R., Mayhew, M. W., Cooper, J. B., & Terrian, D. M. (1996). Identification and localization of an actin-binding motif that is unique to the epsilon isoform of protein kinase C and participates in the regulation of synaptic function. The Journal of Cell Biology, 132, 77–90.

    Google Scholar 

  160. Mochly-Rosen, D., Wu, G., Hahn, H., Osinska, H., Liron, T., Lorenz, J. N., et al. (2000). Cardiotrophic effects of protein kinase Cε: Analysis by in vivo modulation of PKCε translocation. Circulation Research, 86, 1173–1179.

    Google Scholar 

  161. Gupta, M. P., Samant, S. A., Smith, S. H., & Shroff, S. G. (2008). HDAC4 and PCAF bind to cardiac sarcomeres and play a role in regulating myofilament contractile activity. The Journal of Biological Chemistry, 283, 10135–10146.

    Google Scholar 

  162. Takeishi, Y., Ping, P., Bolli, R., Kirkpatrick, D. L., Hoit, B. D., & Walsh, R. A. (2000). Transgenic overexpression of constitutively active protein kinase Cε causes concentric cardiac hypertrophy. Circulation Research, 86, 1218–1223.

    Google Scholar 

  163. Sanger, J. M., & Sanger, J. W. (2008). The dynamic Z bands of striated muscle cells. Science Signaling, 1, pe37.

    Google Scholar 

  164. Lemmon, M. A. (2008). Membrane recognition by phospholipid-binding domains. Nature Reviews. Molecular Cell Biology, 9, 99–111.

    Google Scholar 

  165. Huang, S., Gao, L., Blanchoin, L., & Staiger, C. J. (2006). Heterodimeric capping protein from Arabidopsis is regulated by phosphatidic acid. Molecular Biology of the Cell, 17, 1946–1958.

    Google Scholar 

  166. Kilimann, M. W., & Isenberg, G. (1982). Actin filament capping protein from bovine brain. The EMBO Journal, 1, 889–894.

    Google Scholar 

  167. Hartmann, H., Noegel, A. A., Eckerskorn, C., Rapp, S., & Schleicher, M. (1989). Ca2+-independent F-actin capping proteins. Cap 32/34, a capping protein from Dictyostelium discoideum, does not share sequence homologies with known actin-binding proteins. The Journal of Biological Chemistry, 264, 12639–12647.

    Google Scholar 

  168. Nachmias, V. T., Golla, R., Casella, J. F., & Barron-Casella, E. (1996). Cap Z, a calcium insensitive capping protein in resting and activated platelets. FEBS Letters, 378, 258–262.

    Google Scholar 

  169. Janmey, P. A., Iida, K., Yin, H. L., & Stossel, T. P. (1987). Polyphosphoinositide micelles and polyphosphoinositide-containing vesicles dissociate endogenous gelsolin-actin complexes and promote actin assembly from the fast-growing end of actin filaments blocked by gelsolin. The Journal of Biological Chemistry, 262, 12228–12236.

    Google Scholar 

  170. Janmey, P. A., & Stossel, T. P. (1987). Modulation of gelsolin function by phosphatidylinositol 4,5-bisphosphate. Nature, 325, 362–364.

    ADS  Google Scholar 

  171. Papa, I., Astier, C., Kwiatek, O., Raynaud, F., Bonnal, C., Lebart, M. C., et al. (1999). α-Actinin-CapZ, an anchoring complex for thin filaments in Z-line. Journal of Muscle Research and Cell Motility, 20, 187–197.

    Google Scholar 

  172. Schafer, D. A., Jennings, P. B., & Cooper, J. A. (1996). Dynamics of capping protein and actin assembly in vitro: Uncapping barbed ends by polyphosphoinositides. The Journal of Cell Biology, 135, 169–179.

    Google Scholar 

  173. Visser, M. B., Koh, A., Glogauer, M., & Ellen, R. P. (2011). Treponema denticola major outer sheath protein induces actin assembly at free barbed ends by a PIP2-dependent uncapping mechanism in fibroblasts. PLoS One, 6, e23736.

    ADS  Google Scholar 

  174. Smith, J., Diez, G., Klemm, A. H., Schewkunow, V., & Goldmann, W. H. (2006). CapZ-lipid membrane interactions: A computer analysis. Theoretical Biology & Medical Modelling, 3, 30.

    Google Scholar 

  175. Pyle, W. G., La Rotta, G., de Tombe, P. P., Sumandea, M. P., & Solaro, R. J. (2006). Control of cardiac myofilament activation and PKC-βII signaling through the actin capping protein, CapZ. Journal of Molecular and Cellular Cardiology, 41, 537–543.

    Google Scholar 

  176. Dellefave, L., & McNally, E. M. (2010). The genetics of dilated cardiomyopathy. Current Opinion in Cardiology, 25, 198–204.

    Google Scholar 

  177. Seidman, C. E., & Seidman, J. G. (2011). Identifying sarcomere gene mutations in hypertrophic cardiomyopathy: A personal history. Circulation Research, 108, 743–750.

    Google Scholar 

  178. Moore, J. R., Leinwand, L., & Warshaw, D. M. (2012). Understanding cardiomyopathy phenotypes based on the functional impact of mutations in the myosin motor. Circulation Research, 111, 375–385.

    Google Scholar 

  179. Dorn, G. W. (2012). Decoding the cardiac message: The 2011 Thomas W. Smith memorial lecture. Circulation Research, 110, 755–763.

    Google Scholar 

  180. Arimura, T., Hayashi, T., Terada, H., Lee, S. Y., Zhou, Q., Takahashi, M., et al. (2004). A Cypher/ZASP mutation associated with dilated cardiomyopathy alters the binding affinity to protein kinase C. The Journal of Biological Chemistry, 279, 6746–6752.

    Google Scholar 

  181. Arimura, T., Inagaki, N., Hayashi, T., Shichi, D., Sato, A., Hinohara, K., et al. (2009). Impaired binding of ZASP/Cypher with phosphoglucomutase 1 is associated with dilated cardiomyopathy. Cardiovascular Research, 83, 80–88.

    Google Scholar 

  182. Moric-Janiszewska, E., & Markiewicz-Loskot, G. (2008). Genetic heterogeneity of left-ventricular noncompaction cardiomyopathy. Clinical Cardiology, 31, 201–204.

    Google Scholar 

  183. Zhou, Q., Chu, P. H., Huang, C., Cheng, C. F., Martone, M. E., Knoll, G., et al. (2001). Ablation of Cypher, a PDZ-LIM domain Z-line protein, causes a severe form of congenital myopathy. The Journal of Cell Biology, 155, 605–612.

    Google Scholar 

  184. Zheng, M., Cheng, H., Li, X., Zhang, J., Cui, L., Ouyang, K., et al. (2009). Cardiac-specific ablation of Cypher leads to a severe form of dilated cardiomyopathy with premature death. Human Molecular Genetics, 18, 701–713.

    Google Scholar 

  185. Mochly-Rosen, D., Smith, B. L., Chen, C. H., Disatnik, M. H., & Ron, D. (1995). Interaction of protein kinase C with RACK1, a receptor for activated C-kinase: A role in β protein kinase C mediated signal transduction. Biochemical Society Transactions, 23, 596–600.

    Google Scholar 

  186. Zhou, Q., Ruiz-Lozano, P., Martone, M. E., & Chen, J. (1999). Cypher, a striated muscle-restricted PDZ and LIM domain-containing protein, binds to α-actinin-2 and protein kinase C. The Journal of Biological Chemistry, 274, 19807–19813.

    Google Scholar 

  187. Nakagawa, N., Hoshijima, M., Oyasu, M., Saito, N., Tanizawa, K., & Kuroda, S. (2000). ENH, containing PDZ and LIM domains, heart/skeletal muscle-specific protein, associates with cytoskeletal proteins through the PDZ domain. Biochemical and Biophysical Research Communications, 272, 505–512.

    Google Scholar 

  188. Taniguchi, K., Takeya, R., Suetsugu, S., Kan, O. M., Narusawa, M., Shiose, A., et al. (2009). Mammalian formin fhod3 regulates actin assembly and sarcomere organization in striated muscles. The Journal of Biological Chemistry, 284, 29873–29881.

    Google Scholar 

  189. Iskratsch, T., Lange, S., Dwyer, J., Kho, A. L., dos Remedios, C., & Ehler, E. (2010). Formin follows function: A muscle-specific isoform of FHOD3 is regulated by CK2 phosphorylation and promotes myofibril maintenance. The Journal of Cell Biology, 191, 1159–1172.

    Google Scholar 

  190. Iskratsch, T., & Ehler, E. (2011). Formin-g muscle cytoarchitecture. BioArchitecture, 1, 66–68.

    Google Scholar 

  191. Kan-o, M., Takeya, R., Taniguchi, K., Tanoue, Y., Tominaga, R., & Sumimoto, H. (2012). Expression and subcellular localization of mammalian formin Fhod3 in the embryonic and adult heart. PLoS One, 7, e34765.

    ADS  Google Scholar 

  192. Bai, J., Hartwig, J. H., & Perrimon, N. (2007). SALS, a WH2-domain-containing protein, promotes sarcomeric actin filament elongation from pointed ends during Drosophila muscle growth. Developmental Cell, 13, 828–842.

    Google Scholar 

  193. Pruyne, D., Evangelista, M., Yang, C., Bi, E., Zigmond, S., Bretscher, A., et al. (2002). Role of formins in actin assembly: Nucleation and barbed-end association. Science, 297, 612–615.

    ADS  Google Scholar 

  194. Kovar, D. R., Bestul, A. J., Li, Y. J., & Scott, B. J. (2010). Formin-mediated actin assembly. In M.-F. Carlier (Ed.), Actin-based motility: Cellular, molecular and physical aspects (pp. 279–316). New York: Springer.

    Google Scholar 

  195. Li, F., & Higgs, H. N. (2005). Dissecting requirements for auto-inhibition of actin nucleation by the formin, mDia1. The Journal of Biological Chemistry, 280, 6986–6992.

    Google Scholar 

  196. Westendorf, J. J. (2001). The formin/diaphanous-related protein, FHOS, interacts with Rac1 and activates transcription from the serum response element. The Journal of Biological Chemistry, 276, 46453–46459.

    Google Scholar 

  197. Katoh, M., & Katoh, M. (2004). Identification and characterization of human FHOD3 gene in silico. International Journal of Molecular Medicine, 13, 615–620.

    Google Scholar 

  198. Kanaya, H., Takeya, R., Takeuchi, K., Watanabe, N., Jing, N., & Sumimoto, H. (2005). Fhos2, a novel formin-related actin-organizing protein, probably associates with the nestin intermediate filament. Genes to Cells, 10, 665–678.

    Google Scholar 

  199. Takeya, R., Taniguchi, K., Narumiya, S., & Sumimoto, H. (2008). The mammalian formin FHOD1 is activated through phosphorylation by ROCK and mediates thrombin-induced stress fibre formation in endothelial cells. The EMBO Journal, 27, 618–628.

    Google Scholar 

  200. Zigmond, S. H., Evangelista, M., Boone, C., Yang, C., Dar, A. C., Sicheri, F., et al. (2003). Formin leaky cap allows elongation in the presence of tight capping proteins. Current Biology, 13, 1820–1823.

    Google Scholar 

  201. Cooper, J. A., & Sept, D. (2008). New insights into mechanism and regulation of actin capping protein. International Review of Cell and Molecular Biology, 267, 183–206.

    Google Scholar 

Download references

Acknowledgments

Supported in part by NIH P01 HL62426, NIH 1F32 HL096143, and a grant from the Dr. Ralph and Marian Falk Medical Research Trust.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Allen M. Samarel M.D. .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer Science+Business Media, LLC

About this chapter

Cite this chapter

Samarel, A.M., Koshman, Y., Swanson, E.R., Russell, B. (2013). Biophysical Forces Modulate the Costamere and Z-Disc for Sarcomere Remodeling in Heart Failure. In: Solaro, R., Tardiff, J. (eds) Biophysics of the Failing Heart. Biological and Medical Physics, Biomedical Engineering. Springer, New York, NY. https://doi.org/10.1007/978-1-4614-7678-8_7

Download citation

Publish with us

Policies and ethics