1932

Abstract

Single-molecule atomic force microscopy and magnetic tweezers experiments have demonstrated that titin immunoglobulin (Ig) domains are capable of folding against a pulling force, generating mechanical work that exceeds that produced by a myosin motor. We hypothesize that upon muscle activation, formation of actomyosin cross bridges reduces the force on titin, causing entropic recoil of the titin polymer and triggering the folding of the titin Ig domains. In the physiological force range of 4–15 pN under which titin operates in muscle, the folding contraction of a single Ig domain can generate 200% of the work of entropic recoil and occurs at forces that exceed the maximum stalling force of single myosin motors. Thus, titin operates like a mechanical battery, storing elastic energy efficiently by unfolding Ig domains and delivering the charge back by folding when the motors are activated during a contraction. We advance the hypothesis that titin folding and myosin activation act as inextricable partners during muscle contraction.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-physiol-021317-121254
2018-02-10
2024-03-29
Loading full text...

Full text loading...

/deliver/fulltext/physiol/80/1/annurev-physiol-021317-121254.html?itemId=/content/journals/10.1146/annurev-physiol-021317-121254&mimeType=html&fmt=ahah

Literature Cited

  1. Maruyama K, Matsubara S, Natori R, Nonomura Y, Kimura S. 1.  et al. 1977. Connectin, an elastic protein of muscle characterization and function. J. Biochem. 82:2317–37 [Google Scholar]
  2. Polissar MJ.2.  1952. Physical chemistry of contractile process in muscle. I. A physicochemical model of contractile mechanism. Am. J. Physiol. 168:3766–81 [Google Scholar]
  3. Polissar MJ.3.  1952. Physical chemistry of contractile process in muscle. II. Analysis of other mechano-chemical properties of muscle. Am. J. Physiol. 168:3782–92 [Google Scholar]
  4. Polissar MJ.4.  1952. Physical chemistry of contractile process in muscle. III. Interpretation of thermal behavior of stimulated muscle. Am. J. Physiol. 168:3793–804 [Google Scholar]
  5. Polissar MJ.5.  1952. Physical chemistry of contractile process in muscle. IV. Estimates of size of contractile unit. Am. J. Physiol. 168:3805–11 [Google Scholar]
  6. Gordon AM, Huxley AF, Julian FJ. 6.  1966. Tension development in highly stretched vertebrate muscle fibres. J. Physiol. 184:1143–69 [Google Scholar]
  7. Gordon AM, Huxley AF, Julian FJ. 7.  1966. The variation in isometric tension with sarcomere length in vertebrate muscle fibres. J. Physiol. 184:1170–92 [Google Scholar]
  8. Huxley AF, Niedergerke R. 8.  1954. Structural changes in muscle during contraction: interference microscopy of living muscle fibres. Nature 173:4412971–73 [Google Scholar]
  9. Huxley AF, Simmons RM. 9.  1971. Proposed mechanism of force generation in striated muscle. Nature 233:5321533–38 [Google Scholar]
  10. Rivas-Pardo JA, Eckels EC, Popa I, Kosuri P, Linke WA, Fernández JM. 10.  2016. Work done by titin protein folding assists muscle contraction. Cell Rep 14:61339–47 [Google Scholar]
  11. Huxley HE.11.  1964. Structural arrangements and the contraction mechanism in striated muscle. Proc. R. Soc. B 160:981442–48 [Google Scholar]
  12. Maruyama K, Natori R, Nonomura Y. 12.  1976. New elastic protein from muscle. Nature 262:556358–60 [Google Scholar]
  13. Maruyama K, Kimura S, Ohashi K, Kuwano Y. 13.  1981. Connectin, an elastic protein of muscle. Identification of “titin” with connectin. J. Biochem. 89:3701–9 [Google Scholar]
  14. Stull JT.14.  1996. Myosin minireview series. J. Biol. Chem. 271:2715849 [Google Scholar]
  15. Bang M-L, Centner T, Fornoff F, Geach AJ, Gotthardt M. 15.  et al. 2001. The complete gene sequence of titin, expression of an unusual ≈700-kDa titin isoform, and its interaction with obscurin identify a novel Z-line to I-band linking system. Circ. Res. 89:111065–72 [Google Scholar]
  16. Burkholder TJ, Lieber RL. 16.  2001. Sarcomere length operating range of vertebrate muscles during movement. J. Exp. Biol. 204:91529–36 [Google Scholar]
  17. Millman BM.17.  1998. The filament lattice of striated muscle. Physiol. Rev. 78:2359–91 [Google Scholar]
  18. Cazorla O, Freiburg A, Helmes M, Centner T, McNabb M. 18.  et al. 2000. Differential expression of cardiac titin isoforms and modulation of cellular stiffness. Circ. Res. 86:159–67 [Google Scholar]
  19. Bogomolovas J, Gasch A, Simkovic F, Rigden DJ, Labeit S, Mayans O. 19.  2014. Titin kinase is an inactive pseudokinase scaffold that supports MuRF1 recruitment to the sarcomeric M-line. Open Biol 4:5140041 [Google Scholar]
  20. Lange S, Xiang F, Yakovenko A, Vihola A, Hackman P. 20.  et al. 2005. The kinase domain of titin controls muscle gene expression and protein turnover. Science 308:57281599–603 [Google Scholar]
  21. Linke WA, Hamdani N. 21.  2014. Gigantic business titin properties and function through thick and thin. Circ. Res. 114:61052–68 [Google Scholar]
  22. Hessel AL, Lindstedt SL, Nishikawa KC. 22.  2017. Physiological mechanisms of eccentric contraction and its applications: a role for the giant titin protein. Front. Physiol. 8:70 [Google Scholar]
  23. Tskhovrebova L, Trinick J. 23.  2003. Titin: properties and family relationships. Nat. Rev. Mol. Cell Biol. 4:9679–89 [Google Scholar]
  24. Linke WA, Krüger M. 24.  2010. The giant protein titin as an integrator of myocyte signaling pathways. Physiology 25:3186–98 [Google Scholar]
  25. Spudich JA.25.  2014. Hypertrophic and dilated cardiomyopathy: four decades of basic research on muscle lead to potential therapeutic approaches to these devastating genetic diseases. Biophys. J. 106:61236–49 [Google Scholar]
  26. Prado LG, Makarenko I, Andresen C, Krüger M, Opitz CA, Linke WA. 26.  2005. Isoform diversity of giant proteins in relation to passive and active contractile properties of rabbit skeletal muscles. J. Gen. Physiol. 126:5461–80 [Google Scholar]
  27. Finer JT, Simmons RM, Spudich JA. 27.  1994. Single myosin molecule mechanics: piconewton forces and nanometre steps. Nature 368:6467113–19 [Google Scholar]
  28. Wilkie DR.28.  1968. Heat work and phosphorylcreatine break-down in muscle. J. Physiol. 195:1157–83 [Google Scholar]
  29. Kaya M, Higuchi H. 29.  2010. Nonlinear elasticity and an 8-nm working stroke of single myosin molecules in myofilaments. Science 329:5992686–89 [Google Scholar]
  30. Kaya M, Tani Y, Washio T, Hisada T, Higuchi H. 30.  2017. Coordinated force generation of skeletal myosins in myofilaments through motor coupling. Nat. Commun. 8:16036 [Google Scholar]
  31. Herzog W, Schappacher G, DuVall M, Leonard TR, Herzog JA. 31.  2016. Residual force enhancement following eccentric contractions: a new mechanism involving titin. Physiology 31:4300–12 [Google Scholar]
  32. Herzog W, Leonard TR. 32.  2002. Force enhancement following stretching of skeletal muscle. J. Exp. Biol. 205:91275–83 [Google Scholar]
  33. Nishikawa KC, Monroy JA, Uyeno TE, Yeo SH, Pai DK, Lindstedt SL. 33.  2012. Is titin a “winding filament”? A new twist on muscle contraction. Proc. R. Soc. B 279:1730981–90 [Google Scholar]
  34. Carrion-Vazquez M, Oberhauser AF, Fowler SB, Marszalek PE, Broedel SE. 34.  et al. 1999. Mechanical and chemical unfolding of a single protein: a comparison. PNAS 96:73694–99 [Google Scholar]
  35. Fernandez JM, Li H. 35.  2004. Force-clamp spectroscopy monitors the folding trajectory of a single protein. Science 303:56641674–78 [Google Scholar]
  36. Kellermayer MSZ, Smith SB, Granzier HL, Bustamante C. 36.  1997. Folding-unfolding transitions in single titin molecules characterized with laser tweezers. Science 276:53151112–16 [Google Scholar]
  37. Li H, Linke WA, Oberhauser AF, Carrion-Vazquez M, Kerkvliet JG. 37.  et al. 2002. Reverse engineering of the giant muscle protein titin. Nature 418:6901998–1002 [Google Scholar]
  38. Rief M, Gautel M, Oesterhelt F, Fernandez JM, Gaub HE. 38.  1997. Reversible unfolding of individual titin immunoglobulin domains by AFM. Science 276:53151109–12 [Google Scholar]
  39. Schlierf M, Li H, Fernandez JM. 39.  2004. The unfolding kinetics of ubiquitin captured with single-molecule force-clamp techniques. PNAS 101:197299–304 [Google Scholar]
  40. Tskhovrebova L, Trinick J, Sleep JA, Simmons RM. 40.  1997. Elasticity and unfolding of single molecules of the giant muscle protein titin. Nature 387:6630308–12 [Google Scholar]
  41. Popa I, Kosuri P, Alegre-Cebollada J, Garcia-Manyes S, Fernandez JM. 41.  2013. Force dependency of biochemical reactions measured by single-molecule force-clamp spectroscopy. Nat. Protoc. 8:71261–76 [Google Scholar]
  42. Popa I, Berkovich R, Alegre-Cebollada J, Badilla CL, Rivas-Pardo JA. 42.  et al. 2013. Nanomechanics of HaloTag tethers. J. Am. Chem. Soc. 135:3412762–71 [Google Scholar]
  43. Chen H, Fu H, Zhu X, Cong P, Nakamura F, Yan J. 43.  2011. Improved high-force magnetic tweezers for stretching and refolding of proteins and short DNA. Biophys. J. 100:2517–23 [Google Scholar]
  44. Chen H, Yuan G, Winardhi RS, Yao M, Popa I. 44.  et al. 2015. Dynamics of equilibrium folding and unfolding transitions of titin immunoglobulin domain under constant forces. J. Am. Chem. Soc. 137:103540–46 [Google Scholar]
  45. Svoboda K, Schmidt CF, Schnapp BJ, Block SM. 45.  1993. Direct observation of kinesin stepping by optical trapping interferometry. Nature 365:6448721–27 [Google Scholar]
  46. Mártonfalvi Z, Bianco P, Linari M, Caremani M, Nagy A. 46.  et al. 2014. Low-force transitions in single titin molecules reflect a memory of contractile history. J. Cell Sci. 127:4858–70 [Google Scholar]
  47. Neuman KC, Nagy A. 47.  2008. Single-molecule force spectroscopy: optical tweezers, magnetic tweezers and atomic force microscopy. Nat. Methods 5:6491–505 [Google Scholar]
  48. Huhle A, Klaue D, Brutzer H, Daldrop P, Joo S. 48.  et al. 2015. Camera-based three-dimensional real-time particle tracking at kHz rates and Ångström accuracy. Nat. Commun. 6:5885 [Google Scholar]
  49. Dulin D, Cui TJ, Cnossen J, Docter MW, Lipfert J, Dekker NH. 49.  2015. High spatiotemporal-resolution magnetic tweezers: calibration and applications for DNA dynamics. Biophys. J. 109:102113–25 [Google Scholar]
  50. Popa I, Rivas-Pardo JA, Eckels EC, Echelman DJ, Badilla CL. 50.  et al. 2016. A HaloTag anchored ruler for week-long studies of protein dynamics. J. Am. Chem. Soc. 138:3310546–53 [Google Scholar]
  51. Bustamante C, Marko JF, Siggia ED, Smith S. 51.  1994. Entropic elasticity of lambda-phage DNA. Science 265:51781599–600 [Google Scholar]
  52. Rief M, Clausen-Schaumann H, Gaub HE. 52.  1999. Sequence-dependent mechanics of single DNA molecules. Nat. Struct. Mol. Biol. 6:4346–49 [Google Scholar]
  53. Rief M, Oesterhelt F, Heymann B, Gaub HE. 53.  1997. Single molecule force spectroscopy on polysaccharides by atomic force microscopy. Science 275:53041295–97 [Google Scholar]
  54. Marszalek PE, Oberhauser AF, Pang Y-P, Fernandez JM. 54.  1998. Polysaccharide elasticity governed by chair-boat transitions of the glucopyranose ring. Nature 396:6712661–64 [Google Scholar]
  55. Florin EL, Moy VT, Gaub HE. 55.  1994. Adhesion forces between individual ligand-receptor pairs. Science 264:5157415–17 [Google Scholar]
  56. Radmacher M, Fritz M, Hansma HG, Hansma PK. 56.  1994. Direct observation of enzyme activity with the atomic force microscope. Science 265:51781577–79 [Google Scholar]
  57. Carrion-Vazquez M, Marszalek PE, Oberhauser AF, Fernandez JM. 57.  1999. Atomic force microscopy captures length phenotypes in single proteins. PNAS 96:2011288–92 [Google Scholar]
  58. Valle-Orero J, Rivas-Pardo JA, Tapia-Rojo R, Popa I, Echelman DJ. 58.  et al. 2017. Mechanical deformation accelerates protein ageing. Angew. Chem. Int. Ed. 129:339873–78 [Google Scholar]
  59. Oberhauser AF, Hansma PK, Carrion-Vazquez M, Fernandez JM. 59.  2001. Stepwise unfolding of titin under force-clamp atomic force microscopy. PNAS 98:2468–72 [Google Scholar]
  60. Wiita AP, Ainavarapu SRK, Huang HH, Fernandez JM. 60.  2006. Force-dependent chemical kinetics of disulfide bond reduction observed with single-molecule techniques. PNAS 103:197222–27 [Google Scholar]
  61. Wiita AP, Perez-Jimenez R, Walther KA, Gräter F, Berne BJ. 61.  et al. 2007. Probing the chemistry of thioredoxin catalysis with force. Nature 450:7166124–27 [Google Scholar]
  62. Ainavarapu SRK, Brujić J, Huang HH, Wiita AP, Lu H. 62.  et al. 2007. Contour length and refolding rate of a small protein controlled by engineered disulfide bonds. Biophys. J. 92:1225–33 [Google Scholar]
  63. Sotomayor M, Schulten K. 63.  2007. Single-molecule experiments in vitro and in silico. Science 316:58281144–48 [Google Scholar]
  64. Flory PJ.64.  1953. Principles of Polymer Chemistry Ithaca, NY: Cornell Univ. Press
  65. Stacklies W, Vega MC, Wilmanns M, Gräter F. 65.  2009. Mechanical network in titin immunoglobulin from force distribution analysis. PLOS Comput. Biol. 5:3e1000306 [Google Scholar]
  66. Yuan G, Le S, Yao M, Qian H, Zhou X. 66.  et al. 2017. Elasticity of the transition state leading to an unexpected mechanical stabilization of titin immunoglobulin domains. Angew. Chem. Int. Ed. 129:205582–85 [Google Scholar]
  67. Berkovich R, Garcia-Manyes S, Klafter J, Urbakh M, Fernández JM. 67.  2010. Hopping around an entropic barrier created by force. Biochem. Biophys. Res. Commun. 403:1133–37 [Google Scholar]
  68. Garcia-Manyes S, Dougan L, Badilla CL, Brujić J, Fernández JM. 68.  2009. Direct observation of an ensemble of stable collapsed states in the mechanical folding of ubiquitin. PNAS 106:2610534–39 [Google Scholar]
  69. Redfield C, Smith RAG, Dobson CM. 69.  1994. Structural characterization of a highly-ordered “molten globule” at low pH. Nat. Struct. Mol. Biol. 1:123–29 [Google Scholar]
  70. Mártonfalvi Z, Bianco P, Naftz K, Ferenczy GG, Kellermayer M. 70.  2017. Force generation by titin folding. Protein Sci 26:71380–90 [Google Scholar]
  71. Eckels EC, Rivas-Pardo JA, Valle-Orero J, Popa I, Fernandez JM. 71.  2016. The science of stretching: mechanical anisotropy in titin Ig domains. Biophys. J. 110:3393a [Google Scholar]
  72. Li H, Wang H-C, Cao Y, Sharma D, Wang M. 72.  2008. Configurational entropy modulates the mechanical stability of protein GB1. J. Mol. Biol. 379:4871–80 [Google Scholar]
  73. Li H, Carrion-Vazquez M, Oberhauser AF, Marszalek PE, Fernandez JM. 73.  2000. Point mutations alter the mechanical stability of immunoglobulin modules. Nat. Struct. Mol. Biol. 7:121117–20 [Google Scholar]
  74. Berkovich R, Hermans RI, Popa I, Stirnemann G, Garcia-Manyes S. 74.  et al. 2012. Rate limit of protein elastic response is tether dependent. PNAS 109:3614416–21 [Google Scholar]
  75. Berkovich R, Garcia-Manyes S, Urbakh M, Klafter J, Fernandez JM. 75.  2010. Collapse dynamics of single proteins extended by force. Biophys. J. 98:112692–701 [Google Scholar]
  76. Neupane K, Woodside MT. 76.  2016. Quantifying instrumental artifacts in folding kinetics measured by single-molecule force spectroscopy. Biophys. J. 111:2283–86 [Google Scholar]
  77. Brenner H.77.  1961. The slow motion of a sphere through a viscous fluid towards a plane surface. Chem. Eng. Sci. 16:3–4242–51 [Google Scholar]
  78. Valle-Orero J, Eckels EC, Stirnemann G, Popa I, Berkovich R, Fernandez JM. 78.  2015. The elastic free energy of a tandem modular protein under force. Biochem. Biophys. Res. Commun. 460:2434–38 [Google Scholar]
  79. Alegre-Cebollada J, Kosuri P, Giganti D, Eckels E, Rivas-Pardo JA. 79.  et al. 2014. S-glutathionylation of cryptic cysteines enhances titin elasticity by blocking protein folding. Cell 156:61235–46 [Google Scholar]
  80. Li H, Fernandez JM. 80.  2003. Mechanical design of the first proximal Ig domain of human cardiac titin revealed by single molecule force spectroscopy. J. Mol. Biol. 334:175–86 [Google Scholar]
  81. Rivas-Pardo JA, Eckels E, Valle-Orero J, Fernandez JM. 81.  2017. Disulfide bonding in the contractile work of titin. Biophys. J. 112:3456a [Google Scholar]
  82. Hill AV.82.  1938. The heat of shortening and the dynamic constants of muscle. Proc. R. Soc. B 126:843136–95 [Google Scholar]
  83. Debold EP, Patlak JB, Warshaw DM. 83.  2005. Slip sliding away: load-dependence of velocity generated by skeletal muscle myosin molecules in the laser trap. Biophys. J. 89:5L34–36 [Google Scholar]
  84. Visscher K, Schnitzer MJ, Block SM. 84.  1999. Single kinesin molecules studied with a molecular force clamp. Nature 400:6740184–89 [Google Scholar]
  85. Clemen AEM, Vilfan M, Jaud J, Zhang J, Bärmann M, Rief M. 85.  2005. Force-dependent stepping kinetics of myosin-V. Biophys. J. 88:64402–10 [Google Scholar]
  86. Gennerich A, Carter AP, Reck-Peterson SL, Vale RD. 86.  2007. Force-induced bidirectional stepping of cytoplasmic dynein. Cell 131:5952–65 [Google Scholar]
  87. Linke WA, Stockmeier MR, Ivemeyer M, Hosser H, Mundel P. 87.  1998. Characterizing titin's I-band Ig domain region as an entropic spring. J. Cell Sci. 111:111567–74 [Google Scholar]
  88. Campbell KS.88.  2014. Dynamic coupling of regulated binding sites and cycling myosin heads in striated muscle. J. Gen. Physiol. 143:3387–99 [Google Scholar]
  89. Piazzesi G, Reconditi M, Linari M, Lucii L, Bianco P. 89.  et al. 2007. Skeletal muscle performance determined by modulation of number of myosin motors rather than motor force or stroke size. Cell 131:4784–95 [Google Scholar]
  90. Llewellyn ME, Barretto RPJ, Delp SL, Schnitzer MJ. 90.  2008. Minimally invasive high-speed imaging of sarcomere contractile dynamics in mice and humans. Nature 454:7205784–88 [Google Scholar]
  91. Rassier DE, MacIntosh BR, Herzog W. 91.  1999. Length dependence of active force production in skeletal muscle. J. Appl. Physiol. 86:51445–57 [Google Scholar]
  92. Cromie MJ, Sanchez GN, Schnitzer MJ, Delp SL. 92.  2013. Sarcomere lengths in human extensor carpi radialis brevis measured by microendoscopy. Muscle Nerve 48:2286–92 [Google Scholar]
  93. Sanchez GN, Sinha S, Liske H, Chen X, Nguyen V. 93.  et al. 2015. In vivo imaging of human sarcomere twitch dynamics in individual motor units. Neuron 88:61109–20 [Google Scholar]
  94. Gittes F, Mickey B, Nettleton J, Howard J. 94.  1993. Flexural rigidity of microtubules and actin filaments measured from thermal fluctuations in shape. J. Cell Biol. 120:4923–34 [Google Scholar]
  95. Fudge DS, Gardner KH, Forsyth VT, Riekel C, Gosline JM. 95.  2003. The mechanical properties of hydrated intermediate filaments: insights from hagfish slime threads. Biophys. J. 85:32015–27 [Google Scholar]
  96. Lovelady HH, Shashidhara S, Matthews WG. 96.  2014. Solvent specific persistence length of molecular type I collagen. Biopolymers 101:4329–35 [Google Scholar]
  97. Hidalgo C, Hudson B, Bogomolovas J, Zhu Y, Anderson B. 97.  et al. 2009. PKC phosphorylation of titin's PEVK element—a novel and conserved pathway for modulating myocardial stiffness. Circ. Res. 105:7631–38 [Google Scholar]
  98. Kötter S, Gout L, Frieling-Salewsky MV, Müller AE, Helling S. 98.  et al. 2013. Differential changes in titin domain phosphorylation increase myofilament stiffness in failing human hearts. Cardiovasc. Res. 99:4648–56 [Google Scholar]
  99. Perkin J, Slater R, Del Favero G, Lanzicher T, Hidalgo C. 99.  et al. 2015. Phosphorylating titin's cardiac N2B element by ERK2 or CaMKIIδ lowers the single molecule and cardiac muscle force. Biophys. J. 109:122592–601 [Google Scholar]
  100. Avner BS, Shioura KM, Scruggs SB, Grachoff M, Geenen DL. 100.  et al. 2012. Myocardial infarction in mice alters sarcomeric function via post-translational protein modification. Mol. Cell. Biochem. 363:1–2203–15 [Google Scholar]
  101. Li Y, Linke WA. 101.  2017. Mechanically unfolded titin immunoglobulin domains refold faster and more accurately in presence of chaperone alpha-B-crystallin. Biophys. J. 112:342a [Google Scholar]
  102. Kötter S, Unger A, Hamdani N, Lang P, Vorgerd M. 102.  et al. 2014. Human myocytes are protected from titin aggregation-induced stiffening by small heat shock proteins. J. Cell Biol. 204:2187–202 [Google Scholar]
  103. Bullard B, Ferguson C, Minajeva A, Leake MC, Gautel M. 103.  et al. 2004. Association of the chaperone αB-crystallin with titin in heart muscle. J. Biol. Chem. 279:97917–24 [Google Scholar]
  104. Brinkmeier H, Ohlendieck K. 104.  2014. Chaperoning heat shock proteins: proteomic analysis and relevance for normal and dystrophin-deficient muscle. Proteom. Clin. Appl. 8:11–12875–95 [Google Scholar]
  105. Zhu Y, Bogomolovas J, Labeit S, Granzier H. 105.  2009. Single molecule force spectroscopy of the cardiac titin N2B element effects of the molecular chaperone αB-crystallin with disease-causing mutations. J. Biol. Chem. 284:2013914–23 [Google Scholar]
  106. Roach NT, Venkadesan M, Rainbow MJ, Lieberman DE. 106.  2013. Elastic energy storage in the shoulder and the evolution of high-speed throwing in Homo. . Nature 498:7455483–86 [Google Scholar]
  107. Minajeva A, Neagoe C, Kulke M, Linke WA. 107.  2002. Titin-based contribution to shortening velocity of rabbit skeletal myofibrils. J. Physiol. 540:1177–88 [Google Scholar]
  108. Opitz CA, Kulke M, Leake MC, Neagoe C, Hinssen H. 108.  et al. 2003. Damped elastic recoil of the titin spring in myofibrils of human myocardium. PNAS 100:2212688–93 [Google Scholar]
  109. Dugdale TM, Dagastine R, Chiovitti A, Mulvaney P, Wetherbee R. 109.  2005. Single adhesive nanofibers from a live diatom have the signature fingerprint of modular proteins. Biophys. J. 89:64252–60 [Google Scholar]
  110. Sarkar A, Caamano S, Fernandez JM. 110.  2007. The mechanical fingerprint of a parallel polyprotein dimer. Biophys. J. 92:4L36–38 [Google Scholar]
  111. Bianco P, Reconditi M, Piazzesi G, Lombardi V. 111.  2016. Is muscle powered by springs or motors?. J. Muscle Res. Cell Motil. 37:4165–67 [Google Scholar]
  112. Caremani M, Pinzauti F, Reconditi M, Piazzesi G, Stienen GJM. 112.  et al. 2016. Size and speed of the working stroke of cardiac myosin in situ. PNAS 113:133675–80 [Google Scholar]
  113. Takano M, Terada TP, Sasai M. 113.  2010. Unidirectional Brownian motion observed in an in silico single molecule experiment of an actomyosin motor. PNAS 107:177769–74 [Google Scholar]
  114. Freikamp A, Cost A-L, Grashoff C. 114.  2016. The piconewton force awakens: quantifying mechanics in cells. Trends Cell Biol 26:11838–47 [Google Scholar]
  115. Meng F, Sachs F. 115.  2012. Orientation-based FRET sensor for real-time imaging of cellular forces. J. Cell Sci. 125:3743–50 [Google Scholar]
  116. Austen K, Ringer P, Mehlich A, Chrostek-Grashoff A, Kluger C. 116.  et al. 2015. Extracellular rigidity sensing by talin isoform-specific mechanical linkages. Nat. Cell Biol. 17:121597–606 [Google Scholar]
  117. Grashoff C, Hoffman BD, Brenner MD, Zhou R, Parsons M. 117.  et al. 2010. Measuring mechanical tension across vinculin reveals regulation of focal adhesion dynamics. Nature 466:7303263–66 [Google Scholar]
  118. Arsenovic PT, Ramachandran I, Bathula K, Zhu R, Narang JD. 118.  et al. 2016. Nesprin-2G, a component of the nuclear LINC complex, is subject to myosin-dependent tension. Biophys. J. 110:134–43 [Google Scholar]
  119. Brenner MD, Zhou R, Conway DE, Lanzano L, Gratton E. 119.  et al. 2016. Spider silk peptide is a compact, linear nanospring ideal for intracellular tension sensing. Nano Lett 16:32096–102 [Google Scholar]
  120. Prosser BL, Ward CW, Lederer WJ. 120.  2011. X-ROS signaling: rapid mechano-chemo transduction in heart. Science 333:60481440–45 [Google Scholar]
  121. Rivas-Pardo JA, Mártonfalvi Z, Manteca A, Eckels EC, Echelman DJ. 121.  et al. 2017. A multi-tool mouse model to study the elasticity of native titin. Biophys. J. 112:3167a [Google Scholar]
  122. Kohl J, Ng J, Cachero S, Ciabatti E, Dolan M-J. 122.  et al. 2014. Ultrafast tissue staining with chemical tags. PNAS 111:36E3805–14 [Google Scholar]
  123. Stagge F, Mitronova GY, Belov VN, Wurm CA, Jakobs S. 123.  2013. Snap-, CLIP- and Halo-tag labelling of budding yeast cells. PLOS ONE 8:10e78745 [Google Scholar]
  124. Horowits R, Kempner ES, Bisher ME, Podolsky RJ. 124.  1986. A physiological role for titin and nebulin in skeletal muscle. Nature 323:6084160–64 [Google Scholar]
  125. Higuchi H.125.  1992. Changes in contractile properties with selective digestion of connectin (titin) in skinned fibers of frog skeletal muscle. J. Biochem. 111:3291–95 [Google Scholar]
  126. Dobbie I, Linari M, Piazzesi G, Reconditi M, Koubassova N. 126.  et al. 1998. Elastic bending and active tilting of myosin heads during muscle contraction. Nature 396:6709383–87 [Google Scholar]
  127. Piazzesi G, Reconditi M, Linari M, Lucii L, Sun Y-B. 127.  et al. 2002. Mechanism of force generation by myosin heads in skeletal muscle. Nature 415:6872659–62 [Google Scholar]
  128. Reconditi M, Linari M, Lucii L, Stewart A, Sun Y-B. 128.  et al. 2004. The myosin motor in muscle generates a smaller and slower working stroke at higher load. Nature 428:6982578–81 [Google Scholar]
  129. Fenn WO.129.  1923. A quantitative comparison between the energy liberated and the work performed by the isolated sartorius muscle of the frog. J. Physiol. 58:2–3175–203 [Google Scholar]
  130. Haldar S, Tapia-Rojo R, Eckels EC, Valle-Orero J, Fernandez JM.130.  2017. Trigger factor chaperone acts as a mechanical foldase. Nature Commun. 8:668 [Google Scholar]
/content/journals/10.1146/annurev-physiol-021317-121254
Loading
/content/journals/10.1146/annurev-physiol-021317-121254
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error