Skip to main content
Log in

A multiscale modeling approach for biomolecular systems

  • Published:
Multibody System Dynamics Aims and scope Submit manuscript

Abstract

This paper presents a new multiscale molecular dynamic model for investigating the effects of external interactions, such as contact and impact, during stepping and docking of motor proteins and other biomolecular systems. The model retains the mass properties ensuring that the result satisfies Newton’s second law. This idea is presented using a simple particle model to facilitate discussion of the rigid body model; however, the particle model does provide insights into particle dynamics at the nanoscale. The resulting three-dimensional model predicts a significant decrease in the effect of the random forces associated with Brownian motion. This conclusion runs contrary to the widely accepted notion that the motor protein’s movements are primarily the result of thermal effects. This work focuses on the mechanical aspects of protein locomotion; the effect ATP hydrolysis is estimated as internal forces acting on the mechanical model. In addition, the proposed model can be numerically integrated in a reasonable amount of time. Herein, the differences between the motion predicted by the old and new modeling approaches are compared using a simplified model of myosin V.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  1. Abraham, F.F., Broughton, J.Q., Bernstein, N., Kaxiras, E.: Spanning the continuum to quantum length scales in a dynamic simulation of brittle fracture. EPL (Europhys. Lett.) 44(6), 783 (1998). http://stacks.iop.org/0295-5075/44/i=6/a=783

    Google Scholar 

  2. Acary, V., Brogliato, B.: Lecture Notes in Applied and Computational Mechanics, 1st edn., vol. 35. Springer, Berlin (2008)

    Google Scholar 

  3. Aksimentiev, A., Balabin, I.A., Fillingame, R.H., Schulten, K.: Insights into the molecular mechanism of rotation in the fo sector of atp synthase. Biophys. J. 86(3), 1332–1344 (2004)

    Google Scholar 

  4. Anderson, K., Poursina, M., Bhalerao, K.D.: On adaptive multiscale modeling of biomolecular systems with application in RNA. In: Proceedings of the Joint International Conference on Multibody System Dynamics. Lappeenranta, Finland, (2010)

    Google Scholar 

  5. Asenjo, A.B., Sosa, H.: A mobile kinesin-head intermediate during the ATP-waiting state. Proc. Natl. Acad. Sci. USA 106(14), 5657–5662 (2009)

    Google Scholar 

  6. Atzberger, P.J., Peskin, C.S.: A Brownian dynamics model of kinesin in three dimensions incorporating the force-extension profile of the coiled-coil cargo tether. Bull. Math. Biol. 68(1), 131–160 (2006)

    MathSciNet  Google Scholar 

  7. Austin, R.H.: Nanoscale hydrodynamics in the cell: balancing motorized transport with diffusion. HFSP J. 2(5), 262–265 (2008)

    Google Scholar 

  8. Aydt, E.M., Wolff, G., Morano, I.: Molecular modeling of the myosin-S1(A1) isoform. J. Struct. Biol. 159(1), 158–163 (2007)

    Google Scholar 

  9. Ayton, G.S., Noid, W.G., Voth, G.A.: Multiscale modeling of biomolecular systems: in serial and in parallel. Curr. Opin. Struct. Biol. 17(2), 192–198 (2007)

    Google Scholar 

  10. Baruh, H.: Analytical Dynamics, 1st edn. WCB McGraw-Hill, New York (1999)

    Google Scholar 

  11. Ben-Ari, I., Boushaba, K., Matzavinos, A., Roitershtein, A.: Stochastic Analysis of the Motion of dna Nanomechanical Bipeds. Bulletin of Mathematical Biology (2010)

    Google Scholar 

  12. Bevan, D.R., Garst, J.F., Osborne, C.K., Sims, A.M.: Application of molecular modeling to analysis of inhibition of kinesin motor proteins of the BimC subfamily by monastrol and related compounds. Chem. Biodivers. 2(11), 1525–1532 (2005)

    Google Scholar 

  13. Bier, M.: Processive motor protein as an overdamped Brownian stepper. Phys. Rev. Lett. 91(14), 148104 (2003)

    Google Scholar 

  14. Bier, M.: Modelling processive motor proteins: moving on two legs in the microscopic realm. Contemp. Phys. 46(1), 41–51 (2005)

    MathSciNet  Google Scholar 

  15. Bierbaum, V., Lipowsky, R.: Chemomechanical coupling and motor cycles of myosin V. Biophys. J. 100(7), 1747–1755 (2011)

    Google Scholar 

  16. Block, S.M.: Kinesin motor mechanics: binding, stepping, tracking, gating, and limping. Biophys. J. 92(9), 2986–2995 (2007)

    Google Scholar 

  17. Bockmann, R.A., Grubmuller, H.: Nanoseconds molecular dynamics simulation of primary mechanical energy transfer steps in F1-ATP synthase. Nat. Struct. Biol. 9(3), 198–202 (2002)

    Google Scholar 

  18. Bolterauer, H., Tuszynski, J.A., Unger, E.: Directed binding—a novel physical mechanism that describes the directional motion of two-headed kinesin motor proteins. Cell Biochem. Biophys. 42(2), 95–119 (2005)

    Google Scholar 

  19. Bouzat, S., Falo, F.: The influence of direct motor–motor interaction in models for cargo transport by a single team of motors. Phys. Biol. 7(4), 046009 (2010)

    Google Scholar 

  20. Bowling, A., Flickinger, D.M., Harmeyer, S.: Energetically consistent simulation of simultaneous impacts and contacts in multibody systems with friction. Multibody Syst. Dyn. 22(1), 27–45 (2009)

    MATH  MathSciNet  Google Scholar 

  21. Bowling, A., Haghshenas-Jaryani, M.: Spatial multibody dynamics of nano-scale motor protein locomotion. In: Proceedings of the 1st International Conference on Bionics and Biomechanics (ICABB) (2010)

    Google Scholar 

  22. Bowling, A., Palmer, A.F.: The small mass assumption applied to the multibody dynamics of motor proteins. J. Biomech. 42(9), 1218–1223 (2009). http://www.jbiomech.com/issues. doi:10.1016/j.jbiomech.2009.03.017

    Google Scholar 

  23. Bowling, A., Palmer, A.F., Wilhelm, L.: Contact and impact in the multibody dynamics of motor protein locomotion. Langmuir 25(22), 12974–12981 (2009). http://pubs.acs.org/toc/langd5/0/0

    Google Scholar 

  24. Bueche, F.J.: Introduction to Physics for Scientists and Engineers, 3rd edn. McGraw-Hill Book Company, New York (1979)

    Google Scholar 

  25. Bulatovic, R.M.: A note on the damped vibrating systems. Theor. Appl. Mech. 33(63), 213–221 (2006)

    MATH  MathSciNet  Google Scholar 

  26. Bustamante, C., Keller, D., Oster, G.: The physics of molecular motors. Acc. Chem. Res. 34(6), 412–420 (2001)

    Google Scholar 

  27. Cappello, G., Pierobon, P., Symonds, C., Busoni, L., Gebhardt, J.C., Rief, M., Prost, J.: Myosin V stepping mechanism. Proc. Natl. Acad. Sci. USA 104(39), 15328–15333 (2007)

    Google Scholar 

  28. Chang, R., Ayton, G.S., Voth, G.A.: Multiscale coupling of mesoscopic- and atomistic-level lipid bilayer simulations. J. Chem. Phys. 122(24), 244716 (2005)

    Google Scholar 

  29. Chen, J.C., Kim, A.S.: Brownian dynamics, molecular dynamics, and Monte Carlo modeling of colloidal systems. Adv. Colloid Interface Sci. 112(1–3), 159–173 (2004)

    Google Scholar 

  30. Chu, J.W., Ayton, G.S., Izvekov, S., Voth, G.A.: Emerging methods for multiscale simulation of biomolecular systems. Mol. Phys. 105(2–3), 167–175 (2007)

    Google Scholar 

  31. Chun, H.M., Padilla, C.E., Chin, D.N., Watanabe, M., Karlov, V.I., Alper, H.E., Soosaar, K., Blair, K.B., Becker, O.M., Caves, L.S.D., Nagle, R., Haney, D.N., Farmer, B.L.: MBO(N)D: a multibody method for long-time molecular dynamics simulations. J. Comput. Chem. 21(3), 159–184 (2000)

    Google Scholar 

  32. Ciudad, A., Sancho, J.M., Tsironis, G.P.: Kinesin as an electrostatic machine. J. Biol. Phys. 32(5), 455–463 (2006)

    Google Scholar 

  33. Clemen, A., Vilfan, M., Jaud, J., Zhang, J., Barmann, M., Rief, M.: Force-dependent stepping kinetics of myosin-V. Biophys. J. 88, 4402–4410 (2005)

    Google Scholar 

  34. Coe, J.D., Levine, B.G., Martinez, T.J.: Ab initio molecular dynamics of excited-state intramolecular proton transfer using multireference perturbation theory. J. Phys. Chem. A 111(44), 11302–11310 (2007)

    Google Scholar 

  35. Cordova, N.J., Ermentrout, B., Oster, G.F.: Dynamics of single-motor molecules: the thermal ratchet model. Proc. Natl. Acad. Sci. USA 89(1), 339–343 (1992)

    Google Scholar 

  36. Craig, E.M., Linke, H.: Mechanochemical model for myosin V. Proc. Natl. Acad. Sci. USA 106(43), 18261–18266 (2009)

    Google Scholar 

  37. Cressman, A., Togashi, Y., Mikhailov, A.S., Kapral, R.: Mesoscale modeling of molecular machines: cyclic dynamics and hydrodynamical fluctuations. Phys. Rev. E, Stat. Nonlinear Soft Matter Phys. 77(5 Pt 1), 050901 (2008)

    Google Scholar 

  38. Currie, I.G.: Foundamental Mechanics of Fluids, 3rd edn. (2007). Accel Developement

    Google Scholar 

  39. Cytrynbaum, E.N., Rodionov, V., Mogilner, A.: Computational model of dynein-dependent self-organization of microtubule asters. J. Cell Sci. 117(Pt 8), 1381–1397 (2004)

    Google Scholar 

  40. Derenyi, I., Vicsek, T.: The kinesin walk: a dynamic model with elastically coupled heads. Proc. Natl. Acad. Sci. USA 93, 6775–6779 (1996)

    Google Scholar 

  41. DeVille, R.E.L., Vanden-Eijnden, E.: Regular gaits and optimal velocities for motor proteins. Biophys. J. 95(6), 2681–2691 (2008)

    Google Scholar 

  42. Duke, T., Leibler, S.: Motor protein mechanics: a stochastic model with minimal mechanochemical coupling. Biophys. J. 71(3), 1235–1247 (1996)

    Google Scholar 

  43. Dunn, A.R., Spudich, J.A.: Dynamics of the unbound head during myosin V processive translocation. Nat. Struct. Mol. Biol. 14(3), 246–248 (2007)

    Google Scholar 

  44. Eisenberg, E., Hill, T.L.: A cross-bridge model of muscle contraction. Prog. Biophys. Mol. Biol. 33(1), 55–82 (1978)

    Google Scholar 

  45. Fan, D., Zheng, W., Hou, R., Li, F., Wang, Z.: Modeling motility of the kinesin dimer from molecular properties of individual monomers. Biochemistry 47(16), 4733–4742 (2008)

    Google Scholar 

  46. Ferreira, A.M., Bashford, D.: Model for proton transport coupled to protein conformational change: application to proton pumping in the bacteriorhodopsin photocycle. J. Am. Chem. Soc. 128(51), 16778–16790 (2006)

    Google Scholar 

  47. Fisher, M.E., Kolomeisky, A.B.: Simple mechanochemistry describes the dynamics of kinesin molecules. Proc. Natl. Acad. Sci. USA 98(14), 7748–7753 (2001)

    Google Scholar 

  48. Fricks, J., Wang, H., Elston, T.C.: A numerical algorithm for investigating the role of the motor-cargo linkage in molecular motor-driven transport. J. Theor. Biol. 239(1), 33–48 (2006)

    MathSciNet  Google Scholar 

  49. Gao, Y.Q., Yang, W., Marcus, R.A., Karplus, M.: A model for the cooperative free energy transduction and kinetics of ATP hydrolysis by F1-atpase. Proc. Natl. Acad. Sci. USA 100(20), 11339–11344 (2003)

    Google Scholar 

  50. Gapinski, J., Szymanski, J., Wilk, A., Kohlbecher, J., Patkowski, A., Holyst, R.: Size and shape of micelles studied by means of SANS, PCS, and FCS. Langmuir 26(12), 9304–9314 (2010)

    Google Scholar 

  51. Gardner, M.K., Odde, D.J., Bloom, K.: Kinesin-8 molecular motors: putting the brakes on chromosome oscillations. Trends Cell Biol. 18(7), 307–310 (2008)

    Google Scholar 

  52. Greenberg, M.J., Moore, J.R.: The molecular basis of frictional loads in the in vitro motility assay with applications to the study of the loaded mechanochemistry of molecular motors. Cytoskeleton (Hoboken, N.J.) 67(5), 273–285 (2010)

    Google Scholar 

  53. Haghshenas-Jaryani, M., Bowling, A.: Multiscale dynamic modeling of processive motor proteins. In: Proceedings of the IEEE International Conference Robotics and Biomimetics (ROBIO), Phuket Island, Thailand, December, pp. 1403–1408 (2011)

    Google Scholar 

  54. Haghshenas-Jaryani, M., Bowling, A.: Spatial multibody dynamics of motor proteins. In: Proceedings of Multibody Dynamics 2011, an ECCOMAS Thematic Conference, Brussels, Belgium, July (2011)

    Google Scholar 

  55. Haghshenas-Jaryani, M., Bowling, A.: Multiscale dynamic modeling of flexibility in myosin V using a planar mechanical model. In: Proceedings of the IEEE International Conference Robotics and Biomimetics (ROBIO), Guangzhou, China, December, pp. 366–371 (2012)

    Google Scholar 

  56. Haghshenas-Jaryani, M., Bowling, A.: A new numerical strategy for handling quaternions in dynamic modeling and simulation of rigid multibody systems. In: Proceedings of the 2nd Joint International Conference on Multibody System Dynamics (IMSD), Stuttgart, Germany, May–June (2012)

    Google Scholar 

  57. Haghshenas-Jaryani, M., Bowling, A.: A new switching strategy for addressing Euler parameters in dynamic modeling and simulation of rigid multibody systems. Multibody Syst. Dyn. 30(2), 185–197 (2013). doi:10.1007/s11044-012-9333-8

    MathSciNet  Google Scholar 

  58. Hancock, W.O., Howard, J.: Kinesin’s processivity results from mechanical and chemical coordination between the ATP hydrolysis cycles of the two motor domains. Proc. Natl. Acad. Sci. 96(23), 13147–13152 (1999)

    Google Scholar 

  59. Hayashi, K., Takano, M.: Violation of the fluctuation–dissipation theorem in a protein system. Biophys. J. 93(3), 895–901 (2007)

    Google Scholar 

  60. Hendricks, A., Epureanu, B., Meyhfer, E.: Mechanistic mathematical model of kinesin under time and space fluctuating loads. Nonlinear Dyn. 53(4), 303–320 (2008)

    MATH  Google Scholar 

  61. Howard, J.: Motor proteins as nanomachines: the role of thermal fluctuations in generating force and motion. In: 12th Poincaré Seminar, pp. 33–44 (2009)

    Google Scholar 

  62. Hwang, W., Lang, M.J.: Mechanical design of translocating motor proteins. Cell Biochem. Biophys. 54(1–3), 11–22 (2009)

    Google Scholar 

  63. Izvekov, S., Voth, G.A.: A multiscale coarse-graining method for biomolecular systems. J. Phys. Chem. B 109(7), 2469–2473 (2005). doi:10.1021/jp044629q

    Google Scholar 

  64. Jain, A., Vaidehi, N., Rodriguez, G.: A fast recursive algorithm for molecular dynamics simulation. J. Comput. Phys. 106(2), 258–268 (1993)

    MATH  Google Scholar 

  65. Jamali, Y., Foulaadvand, M.E., Rafii-Tabar, H.: Computational modelling of the collective stochastic motion of kinesin nano motors. J. Theor. Comput. Nano Sci. 7, 146–152 (2010)

    Google Scholar 

  66. Jamali, Y., Lohrasebi, A., Rafii-Tabar, H.: Computational modelling of the stochastic dynamics of kinesin biomolecular motors. Phys. A, Stat. Mech. Appl. 381, 239–254 (2007)

    Google Scholar 

  67. Julicher, F., Ajdari, A., Prost, J.: Modeling molecular motors. Rev. Mod. Phys. 69(4), 1269–1282 (1997)

    Google Scholar 

  68. Julicher, F., Prost, J.: Spontaneous oscillations of collective molecular motors. Phys. Rev. Lett. 78(23), 4510–4513 (1997)

    Google Scholar 

  69. Karplus, M., McCammon, J.A.: Molecular dynamics simulations of biomolecules. Nat. Struct. Biol. 9(9), 646–652 (2002)

    Google Scholar 

  70. Kim, D.N., Nguyen, C.T., Bathe, M.: Conformational dynamics of supramolecular protein assemblies. J. Struct. Biol. 173(2), 261–270 (2011)

    Google Scholar 

  71. Kim, T., Kao, M.T., Hasselbrink, E.F., Meyhofer, E.: Nanomechanical model of microtubule translocation in the presence of electric fields. Biophys. J. 94(10), 3880–3892 (2008)

    Google Scholar 

  72. Kolomeisky, A.B., Fisher, M.E.: A simple kinetic model describes the processivity of myosin-V. Biophys. J. 84, 1642–1650 (2003)

    Google Scholar 

  73. Kolomeisky, A.B., Fisher, M.E.: Molecular motors: a theorist’s perspective. Annu. Rev. Phys. Chem. 58, 675–695 (2007)

    Google Scholar 

  74. Korn, C.B., Klumpp, S., Lipowsky, R., Schwarz, U.S.: Stochastic simulations of cargo transport by processive molecular motors. J. Chem. Phys. 131(24), 245107 (2009)

    Google Scholar 

  75. Kuznetsov, A.V., Avramenko, A.A., Blinov, D.G.: Numerical modeling of molecular-motor-assisted transport of adenoviral vectors in a spherical cell. Comput. Methods Biomech. Biomed. Eng. 11(3), 215–222 (2008)

    Google Scholar 

  76. Lan, G., Sun, S.X.: Dynamics of myosin-V processivity. Biophys. J. 88(2), 999–1008 (2005)

    Google Scholar 

  77. Lan, G., Sun, S.X.: Flexible light-chain and helical structure of F-actin explain the movement and step size of myosin-VI. Biophys. J. 91, 4002–4013 (2006)

    Google Scholar 

  78. Lei, U., Yang, C.Y., Wu, K.C.: Viscous torque on a sphere under arbitrary rotation. Appl. Phys. Lett. 89(18), 181908 (2006). http://link.aip.org/link/?APL/89/181908/1. doi:10.1063/1.2372704

    Google Scholar 

  79. Leibler, S., Huse, D.A.: Porters versus rowers: a unified stochastic model of motor proteins. J. Cell Biol. 121(6), 1357–1368 (1993)

    Google Scholar 

  80. Levin, Y.: Dynamics of myosin-V processivity. Rep. Prog. Phys. 65(11), 1577–1632 (2002)

    Google Scholar 

  81. Lin, C.T., Meyhofer, E., Kurabayashi, K.: Predicting the stochastic guiding of kinesin-driven microtubules in microfabricated tracks: a statistical-mechanics-based modeling approach. Phys. Rev. E, Stat. Nonlinear Soft Matter Phys. 81(1 Pt 1), 011919 (2010)

    Google Scholar 

  82. Lipowsky, R., Liepelt, S.: Chemomechanical coupling of molecular motors: thermodynamics, network representations, and balance conditions. J. Stat. Phys. 130(1), 39–67 (2008)

    MATH  MathSciNet  Google Scholar 

  83. Liu, J., Taylor, D.W., Krementsova, E.B., Trybus, K.M., Taylor, K.A.: Three-dimensional structure of the myosin V inhibited state by cryoelectron tomography. Nature 442(13), 208–211 (2006)

    Google Scholar 

  84. Lohrasebi, A., Jamali, Y., Rafii-Tabar, H.: Modeling the effect of external electric field and current on the stochastic dynamics of atpase nano-biomolecular motors. Phys. A, Stat. Mech. Appl. 387, 5466–5476 (2007)

    Google Scholar 

  85. Masuda, T.: A simulation model of the conventional kinesin based on the driven-by-detachment mechanism. Biosystems 97(2), 121–126 (2009)

    MathSciNet  Google Scholar 

  86. Mateos, J.L.: Walking on ratchets with two Brownian motors. Fluct. Noise Lett. 4(1), L161–L170 (2004)

    Google Scholar 

  87. Mather, W.H., Fox, R.F.: Kinesin’s biased stepping mechanism: amplification of neck linker zippering. Biophys. J. 91(7), 2416–2426 (2006)

    Google Scholar 

  88. Miller, R., Tadmor, E.: The quasicontinuum method: overview, applications and current directions. J. Comput.-Aided Mater. Des. 9, 203–239 (2002). http://dx.doi.org/10.1023/A:1026098010127

    Google Scholar 

  89. Mukherjee, R.M., Crozier, P.S., Plimpton, S.J., Anderson, K.S.: Substructured molecular dynamics using multibody dynamics algorithms. Int. J. Non-Linear Mech. 43(10), 1040–1055 (2008)

    MATH  Google Scholar 

  90. Mullner, F.E., Syed, S., Selvin, P.R., Sigworth, F.J.: Improved hidden Markov models for molecular motors, part 1: Basic theory. Biophys. J. 99(11), 3684–3695 (2010)

    Google Scholar 

  91. Nayfeh, A.H.: Perturbation Methods. Wiley, New York (1973)

    MATH  Google Scholar 

  92. Neto, N., Bellucci, L.: A new algorithm for rigid body molecular dynamics. Chem. Phys. 328(1–3), 259–268 (2006)

    Google Scholar 

  93. Parker, D., Bryant, Z., Delp, S.L.: Coarse-grained structural modeling of molecular motors using multibody dynamics. Cell. Mol. Bioeng. 2(3), 366–374 (2009)

    Google Scholar 

  94. Pavliotis, G.A., Stuart, A.M.: Periodic homogenization for inertial particles. Phys. D, Nonlinear Phenom. 2004(3–4), 161–187 (2005)

    MathSciNet  Google Scholar 

  95. Peskin, C.S., Odell, G.M., Oster, G.F.: Cellular motions and thermal fluctuations: the Brownian ratchet. Biophys. J. 65(1), 316–324 (1993)

    Google Scholar 

  96. Peskin, C.S., Oster, G.: Coordinated hydrolysis explains the mechanical behavior of kinesin. Biophys. J. 68(4 Suppl), 202S–210S (1995). Discussion, 210S–211S

    Google Scholar 

  97. Ping, X., Shuo-Xing, D., Peng-Ye, W.: A model for processivity of molecular motors. Chin. Phys. 13(9), 1569–2863 (2004)

    Google Scholar 

  98. Poursina, M., Anderson, K.S.: Canonical ensemble simulation of biopolymers using a coarse-grained articulated generalized divide-and-conquer scheme. Comput. Phys. Commun. 184(3), 652–660 (2013)

    MATH  MathSciNet  Google Scholar 

  99. Poursina, M., Anderson, K.S.: Efficient coarse-grained molecular simulations in the multibody dynamics scheme. Multibody Dyn. 28, 147–172 (2013)

    MathSciNet  Google Scholar 

  100. Poursina, M., Bhalerao, K.D., Flores, S.C., Anderson, K.S., Laederach, A.: Strategies for articulated multibody-based adaptive coarse grain simulation of rna. Methods Enzymol. 487, 73–98 (2011)

    Google Scholar 

  101. Praprotnik, M., Site, L.D., Kremer, K.: Adaptive resolution molecular-dynamics simulation: changing the degrees of freedom on the fly. J. Chem. Phys. 123(22) (2005)

  102. Pratt, C., Cornely, K.: Essential Biochemistry. Wiley, New York (2004)

    Google Scholar 

  103. Purcell, T.J., Sweeney, H.L., Spudich, J.A.: A force-dependent state controls the coordination of processive myosin V. Proc. Natl. Acad. Sci. 102(39), 13873–13878 (2005)

    Google Scholar 

  104. Rafii-Tabar, H., Jamali, Y., Lohrasebi, A.: Computational modelling of the stochastic dynamics of kinesin biomolecular motors. Physica A 381, 239–254 (2007)

    Google Scholar 

  105. Reif, F.: Fundamentals of Statistical and Thermal Physics. McGraw Hill, New York (1965)

    Google Scholar 

  106. Reimann, P.: Brownian motors: noisy transport far from equilibrium. Phys. Rep. 361(2–4), 57–265 (2002)

    MATH  MathSciNet  Google Scholar 

  107. Rice, S.E., Purcell, T.J., Spudich, J.A.: Building and using optical traps to study properties of molecular motors. Methods Enzymol. 361, 112–133 (2003)

    Google Scholar 

  108. Rief, M., Rock, R.S., Mehta, A.D., Mooseker, M.S., Cheney, R.E., Spudich, J.A.: Myosin-V stepping kinetics: a molecular model for processivity. Proc. Natl. Acad. Sci. 97(17), 9482–9486 (2000)

    Google Scholar 

  109. Rossi, R., Isorce, M., Morin, S., Flocard, J., Arumugam, K., Crouzy, S., Vivaudou, M., Redon, S.: Adaptive torsion-angle quasi-statics: a general simulation method with applications to protein structure analysis and design. Bioinformatics 23(13), i408–417 (2007)

    Google Scholar 

  110. Rudd, R.E., Broughton, J.Q.: Coarse-grained molecular dynamics and the atomic limit of finite elements. Phys. Rev. B 58(10), R5893–R5896 (1998). doi:10.1103/PhysRevB.58.R5893

    Google Scholar 

  111. Schief, W.R., Howard, J.: Conformational changes during kinesin motility. Curr. Opin. Cell Biol. 13(1), 19–28 (2001)

    Google Scholar 

  112. Schuyler, A.D., Chirikjian, G.S.: Normal mode analysis of proteins: a comparison of rigid cluster modes with c α coarse graining. J. Mol. Graph. Model. 22(3), 183–193 (2004)

    Google Scholar 

  113. Schuyler, A.D., Chirikjian, G.S.: Efficient determination of low-frequency normal modes of large protein structures by cluster-nma. J. Mol. Graph. Model. 24(1), 46–58 (2005)

    Google Scholar 

  114. Schwieters, C.D., Clore, G.M.: A physical picture of atomic motions within the Dickerson DNA dodecamer in solution derived from joint ensemble refinement against NMR and large-angle X-ray scattering data. Biochemistry 46(5), 1152–1166 (2007)

    Google Scholar 

  115. Shao, Q., Gao, Y.Q.: On the hand-over-hand mechanism of kinesin. Proc. Natl. Acad. Sci. USA 103(21), 8072–8077 (2006)

    Google Scholar 

  116. Shiroguchi, K., Kinosita, K.: Myosin V walks by lever action and Brownian motion. Science 316(5828), 1208–1212 (2007)

    Google Scholar 

  117. Simon, S.M., Peskin, C.S., Oster, G.F.: What drives the translocation of proteins? Proc. Natl. Acad. Sci. 89(9), 3770–3774 (1992)

    Google Scholar 

  118. Singh, M.P., Mallik, R., Gross, S.P., Yu, C.C.: Monte Carlo modeling of single-molecule cytoplasmic dynein. Proc. Natl. Acad. Sci. USA 102(34), 12059–12064 (2005)

    Google Scholar 

  119. Skau, K.I., Hoyle, R.B., Turner, M.S.: A kinetic model describing the processivity of myosin-V. Biophys. J. 91, 2475–2489 (2006)

    Google Scholar 

  120. Sosa, H., Peterman, E.J.G., Moerner, W.E., Goldstein, L.S.B.: ADP-induced rocking of the kinesin motor domain revealed by single-molecule fluorescence polarization microscopy. Nat. Struct. Biol. 8(6), 540–544 (2001)

    Google Scholar 

  121. Stratopoulos, G.N., Dialynas, T.E., Tsironis, G.P.: Directional Newtonian motion and reversals of molecular motors. Phys. Lett. A 252(3–4), 151–156 (1999)

    Google Scholar 

  122. Szymanski, J., Patkowski, A., Wilk, A., Garstecki, P., Holyst, R.: Diffusion and viscosity in a crowded environment: from nano- to macroscale. Phys. Chem. Lett. B 110, 25593–25597 (2006)

    Google Scholar 

  123. Tsygankov, D., Fisher, M.E.: Kinetic models for mechanoenzymes: structural aspects under large loads. J. Chem. Phys. 128(1), 015102 (2008)

    Google Scholar 

  124. Vaidehi, N., Jain, A., Goddard, W.A.: Constant temperature constrained molecular dynamics: the Newton–Euler inverse mass operator method. J. Phys. Chem. 100(25), 10508–10517 (1996). doi:10.1021/jp953043o

    Google Scholar 

  125. Vale, R.D.: Myosin V motor proteins: marching stepwise towards a mechanism. J. Cell Biol. 163(3), 445–450 (2003)

    Google Scholar 

  126. Veigel, C., Wang, F., Bartoo, M.L., Sellers, J.R., Molloy, J.E.: The gated gait of the processive molecular motor, myosin V. Nat. Cell Biol. 4(1), 59–65 (2002)

    Google Scholar 

  127. Vilfan, A.: Elastic lever-arm model for myosin V. Biophys. J. 88, 3792–3805 (2005)

    Google Scholar 

  128. Vilfan, A.: Five models for myosin V. Front. Biosci. 14, 2269–2284 (2009)

    Google Scholar 

  129. Wagner, G.J., Liu, W.K.: Coupling of atomistic and continuum simulations using a bridging scale decomposition. J. Comput. Phys. 190(1), 249–274 (2003)

    MATH  Google Scholar 

  130. Walcott, S., Warshaw, D.M.: Modeling smooth muscle myosin’s two heads: long-lived enzymatic roles and phosphorylation-dependent equilibria. Biophys. J. 99(4), 1129–1138 (2010)

    Google Scholar 

  131. Wang, H.: Mathematical theory of molecular motors and a new approach for uncovering motor mechanism. IEE Proc. Nanobiotechnol. 150(3), 127–133 (2003)

    Google Scholar 

  132. Wang, H., Elston, T.C.: Mathematical and computational methods for studying energy transduction in protein motors. J. Stat. Phys. 128(1–2), 35–76 (2007)

    MATH  MathSciNet  Google Scholar 

  133. Warshaw, D.M., Kennedy, G.G., Work, S.S., Krementsova, E.B., Beck, S.: Differential labeling of myosin V heads with quantum dots allows direct visualization of hand-over-hand processivity. Biophys. J. 88(5), L30–L32 (2005)

    Google Scholar 

  134. Wereley, S.T., Meinhart, C.D.: Recent advances in micro-particle image velocimetry. Annu. Rev. Fluid Mech. 42, 557–576 (2010)

    Google Scholar 

  135. Wu, Y., Gao, Y.Q., Karplus, M.: A kinetic model of coordinated myosin V. Biochemistry 46, 6318–6330 (2007)

    Google Scholar 

  136. Xiao, S.P., Belytschko, T.: A bridging domain method for coupling continua with molecular dynamics. Comput. Methods Appl. Mech. Eng. 193(17–20), 1645–1669 (2004)

    MATH  MathSciNet  Google Scholar 

  137. Xie, P.: Stepping behavior of two-headed kinesin motors. Biochim. Biophys. Acta (BBA), Bioenerg. 1777(9), 1195–1202 (2008)

    Google Scholar 

  138. Xing, J., Wang, H., Oster, G.: From continuum Fokker–Planck models to discrete kinetic models. Biophys. J. 89(3), 1551–1563 (2005)

    Google Scholar 

  139. Yamada, M.D., Maruta, S., Yasuda, S., Kondo, K., Maeda, H., Arata, T.: Conformational dynamics of loops l11 and l12 of kinesin as revealed by spin-labeling EPR. Biochem. Biophys. Res. Commun. 364(3), 620–626 (2007)

    Google Scholar 

  140. Yildiz, A., Tomishige, M., Vale, R.D., Selvin, P.R.: Kinesin walks hand-over-hand. Langmuir 20(12), 4892–4897 (2004)

    Google Scholar 

  141. Yu, H., Ma, L., Yang, Y., Cui, Q.: Mechanochemical coupling in the myosin motor domain. I. Insights from equilibrium active-site simulations. PLoS Comput. Biol. 3(2), e21 (2007)

    Google Scholar 

  142. Yu, H., Ma, L., Yang, Y., Cui, Q.: Mechanochemical coupling in the myosin motor domain. II. Analysis of critical residues. PLoS Comput. Biol. 3(2), e23 (2007)

    Google Scholar 

  143. Yu, J., Ha, T., Schulten, K.: Structure-based model of the stepping motor of PcrA helicase. Biophys. J. 91(6), 2097–2114 (2006)

    Google Scholar 

  144. Zeldovich, K.B., Joanny, J.F., Prost, J.: Motor proteins transporting cargos. Eur. Phys. J. E 17(2), 155–163 (2005)

    Google Scholar 

  145. Zhang, J., Li, W., Wang, J., Qin, M., Wu, L., Yan, Z., Xu, W., Zuo, G., Wang, W.: Protein folding simulations: from coarse-grained model to all-atom model. IUBMB Life 61(6), 627–643 (2009)

    Google Scholar 

  146. Zheng, W.: Multiscale modeling of structural dynamics underlying force generation and product release in actomyosin complex. Proteins 78(3), 638–660 (2010)

    Google Scholar 

  147. Zheng, W., Doniach, S.: A comparative study of motor-protein motions by using a simple elastic-network model. Proc. Natl. Acad. Sci. 100(23), 13253–13258 (2003)

    Google Scholar 

Download references

Acknowledgements

This work was supported by National Science Foundation Grant No. MCB-1148541 and funds from the Department of Mechanical and Aerospace Engineering at the University of Texas at Arlington.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alan Bowling.

Appendices

Appendix A: The model

This section discusses in detail the different types of force acting on the model presented in Sect. 2.2. The physical parameters of the myosin V mechanical model are given in Table 2.

Table 2 Physical parameters of the Myosin V mechanical model

1.1 A.1 Viscous friction

Each body has a force and moment applied at and about its mass center to approximate the viscous friction of the fluid through which the motor protein moves, as shown in Fig. 9. In order to truly assess the effects of viscous friction on a rigid body, one should consider drag that depends on its shape and orientation, but here a simple coefficient of viscous friction is used. For example, the friction force and moment on body A are

$$ {\mathbf{f}}_{A_o} = - \beta_2 \boldsymbol {v}_{A_o} \quad \mbox{and}\quad {\mathbf{m}}_{A} = - \beta_2 \bar{L}_A^2 {\boldsymbol{\omega}}_{A} $$
(23)

where \(\mathbf{f}_{i_{o}}\) and m i are the force and moment acting on body i, and \(\mathbf{v}_{i_{o}}\) and ω i are the translational and rotational inertial velocities associated with body i. The term \(\bar{L}_{i}\) is a characteristic length for rotational viscous friction; the characteristic length for a sphere shape body equals its radius, which is the half length of the whole body [78]. Thus, the characteristic length for other shape bodies is approximately considered as the half length of the whole body. The values of these coefficients are listed in Table 2. The viscous friction forces for each body comprise Γ friction.

Fig. 9
figure 9

Viscous friction forces acting on myosin V

1.2 A.2 Conformational changes and external forces

Several forces contribute to protein locomotion, including the conformational changes due to ATP hydrolysis. Regardless of their source and application, their resultant can be transformed into equivalent forces at the protein’s joints. At this point in the development of the multibody model, it is not critical to model all the forces contributing to protein locomotion in detail. The equivalent forces provide a simple means for generating stepping, which allows examination of contact and impact between the heads and substrate.

The viscous friction, binding charge, contact, and random forces are excluded from the equivalent forces because they are modeled explicitly. The equivalent forces are modeled as three moments, one acting between the necks, and two acting between the heads and necks; see Fig. 10. They are loosely associated with the neck linker force and power strokes at each head; however, they may have several sources beyond ATP hydrolysis. These equivalent forces comprise Γ conform in Eq. (15).

Fig. 10
figure 10

Forces emulating conformational changes and external forces

A combination of constant neck linker [66, 87] and power stroke forces is used here to produce locomotion. This is done so that the same forces can be applied to the massless and massive models. The motors are activated when both heads have docked, and only one power stroke is active at any given time. Both forces are deactivated when the stepping head’s binding site is 5 nm from the actin binding site. Allowing the actin binding site to pull the head in to bind and dock yields a more robust control, as opposed to steering the head all the way in. In addition, the power stroke is deactivated when the angle between a head and its neck is 90 assuming that this is the maximum travel of the power stroke. The neck linker force is deactivated if the angle between the necks is greater than 75. The motors apply a constant torque when activated. The equivalent moments acting on the proteins to achieve the step are listed in Table 3. These moments are separated into large and small components as

$$ {\boldsymbol{\varGamma}}_{\mathrm{conform}} = \mathsf{G}^T \left [ \begin{array}{c} \varUpsilon_R \\ \varUpsilon_S \\ \varUpsilon_P \end{array} \right ] ( \beta_1 + 1) = \mathsf{G}^T {\boldsymbol{\varUpsilon}} ( \beta_1 + 1 ) $$
(24)

where ϒ i is the moment associated with point i, and G T transforms the moments into generalized active forces.

Table 3 Equivalent joint moments

1.3 A.3 Binding charges

In the literature, there are charge potentials proposed that attempt to model the repulsion required to keep the head from penetrating the actin filament. Since the repulsion forces are addressed using contact forces, herein the binding potential only models the attraction between the binding sites using a simple Coulomb-like charge model. The attractive forces resulting from these charges are shown in Fig. 11.

Fig. 11
figure 11

Binding charges on myosin V

The interaction of the motor protein with the actin filament involves docking of the heads at the binding sites [13, 14]. Here, this docking is assumed to completely immobilize the bound head and only occurs when the point binding sites on a head and the actin filament closely align. Modeling the head as an ellipse ensures that the head must achieve a particular configuration before docking is allowable, as would occur for the actual protein.

When a head docks, it is assumed that the charges involved sum and are neutralized; the charges at the docking site do not affect the undocked head. This is accomplished by setting the charges involved equal to zero. These charges remain zero until the head detaches, after which only the charge on the binding site is set back to its original value.

After detaching, the head must recharge at some point during its motion in order to achieve the step. The head charge is switched back on when its binding site has passed mid step and is 2/3 of the way to the targeted binding site on the actin filament. The head needs to be closer to the new binding site than to the old since the binding charges are functions of distance. If not, the head will be pulled back to the original binding site.

The position vectors from the heads’ mass centers to their binding sites are expressed as

$$\begin{aligned} {\mathbf{P}}_{\mathrm{C}_{\mathrm{o}} \mathrm{C}_{\mathrm{E}}} = & L_6 \mathbf{C}_1 + L_7 \mathbf{C}_2 \end{aligned}$$
(25)
$$\begin{aligned} {\mathbf{P}}_{\mathrm{D}_\mathrm{o} \mathrm{D}_\mathrm{E}} = & L_8 \mathbf{D}_1 + L_9 \mathbf{D}_2 \end{aligned}$$
(26)

The force on a head’s binding site \(\rm C_{E}\) due to site \(\rm B_{1}\), for example, is

$$ {\mathbf{F}}_{\mathrm{C}_\mathrm{E} \mathrm{B}_{1}} = \mathrm{f}_{\mathrm{C}_\mathrm{E} \mathrm{B}_{1}} \frac{\mathbf{P}_{\mathrm{C}_\mathrm{E} \mathrm{B}_{1}} }{\| {\mathbf{P}}_{\mathrm{C_E B_1}}\|} $$
(27)

There are a number of charge potentials investigated in the literature for defining the attractive force \(\mathrm{f}_{\mathrm{C_{E} B_{1}}}\). These include ratchet potentials [66] and Coulomb-like potentials [32]. A Coulomb-like charge potential [32] is used here:

$$ V(r) = \frac{ C_{\mathrm{C}_\mathrm{E}} C_{\mathrm{B_1}} e^{k(a-r)} }{ 4 \pi\epsilon_0 \epsilon_r (1 + k a) (r+k_o)} $$
(28)

where ϵ 0, ϵ r , k, and a are the permittivity of free space, relative electric permittivity, inverse of the Debye length, and excluded volume radius; the values for these constants are given in Table 2. The term r is the distance between the binding sites, and k o is a penetration depth used to keep the potential from becoming infinite. The terms \(C_{\mathrm{C}_{\mathrm{E}}}\) and \(C_{\mathrm{B}_{1}}\) are the charges on the binding sites of head C and B1 on the actin fiber; these values are given in Table 2. The absolute value of this potential as a function of radial distance is illustrated in Fig. 12.

Fig. 12
figure 12

Absolute value of the charge potential and the associated force

The force can be found as

$$ {\mathrm{f}}_{\mathrm{C}_\mathrm{E} \mathrm{B}_{1}} = -\frac{dV(r)}{dr} = \frac{C_{\mathrm{C}_\mathrm{E}} C_{\mathrm{B_1}} e^{k(a-r)} }{4 \pi\epsilon_0 \epsilon_r (1 + k a) } \frac{ k(r+k_o) + 1}{(r+k_o)^2} $$
(29)

where \(r=\| {\mathbf{P}}_{\mathrm{C}_{\mathrm{E}} \mathrm{B}_{1}} \|\) is the distance between the head and actin binding sites at points CE and B1 in Fig. 2. Similar forces can be defined between the other binding sites. These charge forces comprise Γ charge in (15) and are separated into small and large components:

$$ {\boldsymbol{\varGamma}}_{\mathrm{charge}} = (\beta_1+1) Q \mathbf{f}_{\mathrm{charge}} $$
(30)

where the matrix Q transforms the charges into generalized active forces. The force \(\mathrm{f}_{\mathrm{C}_{\mathrm{E}} \mathrm{B}_{1}}\) generated by the charge potential as a function of radial distance is illustrated in Fig. 12.

These charges are electrostatic and only depend on the relative positions of the binding sites. They are conservative forces whose values are always known. Thus, it is necessary to ensure that their computed values and the decomposed values match up. This is accomplished by defining \(\overline{\mathrm{f}}_{ij}\):

$$ {\mathbf{f}}_{\mathrm{charge}} = [ \overline{\mathrm{f}}_{\mathrm{C}_\mathrm{E} \mathrm{B}_{1}}\quad \overline{\mathrm{f}}_{\mathrm{C}_\mathrm{E} \mathrm{B}_{2}} \quad\cdots\quad\overline{\mathrm{f}}_{\mathrm{D}_\mathrm{E} \mathit{B}_{5} } \quad \overline {\mathrm{f}}_{\mathrm{D}_\mathrm{E} \mathrm{B}_{6}} ]^T $$
(31)

where the \(\overline{\mathrm{f}}_{ij}\) are solved such that, for example,

$$ {\mathrm{f}}_{\mathrm{C_E B_1}} = (\beta_1 + 1) \overline { \mathrm{f}}_{\mathrm{C_E B_1}} $$
(32)

and \(\mathrm{f}_{\mathrm{C_{E} B_{1}}}\) is determined from (29).

1.4 A.4 Contact forces

1.4.1 A.4.1 Contact points

Here it is assumed that contact between the heads and the actin filament occurs in a localized region that can be approximated as point contact. Modeling the heads as ellipsoids aligns with this assumption. In this work the heads and actin are assumed to be frictionless for the sake of simplicity; thus, the horizontal forces illustrated in Fig. 13 do not represent tangential friction. They are used to enforce an immobilization of the head when it docks.

Fig. 13
figure 13

Contact forces on myosin V

As stated, the heads are modeled as ellipsoids:

$$ \frac{ x_C^2 }{ r_{\mathrm{head}_1}^2 } + \frac{ y_C^2 }{r_{\mathrm{head}_2}^2 } + \frac{ z_C^2 }{ r_{\mathrm{head}_3}^2 }= 1 $$
(33)

where the position vector between point Co and any point CC on the surface of the head is expressed as

$$ {\mathbf{P}}_{C_o C_C} = x_C \mathbf{C}_1 + y_C \mathbf{C}_2 + z_C \mathbf{C}_3 $$
(34)

and likewise for head D. Table 2 gives the values of \(r_{\mathrm{head}_{1}}\), \(r_{\mathrm{head}_{2}}\), and \(r_{\mathrm{head}_{3}}\) assuming that both heads are identical.

At each step in the simulation, the contact points between each head and the actin filament are determined. The approach used to determine the contact forces at these points is presented in [20] and will not be duplicated fully here. Tangential forces associated with sliding or sticking friction are omitted.

1.4.2 A.4.2 Complementarity conditions

The complementarity conditions are a well-known means for describing the relationships between friction, contact forces, velocities, and accelerations [2]. Assuming that the distance between the impacting points equals zero, the conditions are dependent on the value of the preimpact normal velocity and acceleration as

$$ \left \{ \begin{array}{l@{\quad}c@{\quad}l} v_{n_i}(t) < 0 & & \mathrm{impact or contact} \\ v_{n_i}(t) = 0 & \mathrm{and} & \left \{ \begin{array}{l@{\quad}l} \dot{v}_{n_i}(t) \leq0 & \mathrm{contact} \\ \dot{v}_{n_i}(t) > 0 & \mathrm{separation} \end{array} \right . \\ v_{n_i}(t) > 0 & & \mathrm{separation} \end{array} \right . $$
(35)

A transition occurs when the preimpact normal velocity equals zero. The preimpact acceleration must be checked to determine whether impact forces will exist.

When the pre-impact normal velocity indicates impact or contact, its postimpact value is

$$ v_{n_i}(t+\epsilon) = -e_{n_i} v_{n_i}(t) $$
(36)

Herein the nonpenetration condition is modeled by \(e_{n_{i}} = 0\).

1.4.3 A.4.3 Contact force calculation

In brief, the method treats contact as a succession of small impacts. Impact is treated as a discrete event with an instantaneous velocity change. At each step in the numerical integration, an impact law is used to determine an instantaneous change in velocity. This law uses a coefficient of restitution, which is set to zero for the motor protein case. This means that the head does not rebound but can remain attached to the substrate after impact. This approach also involves the calculation of the contact and impact forces that comprise Γ contact. These forces are necessary, in addition to the velocity resolution, in order to keep the head from penetrating the actin filament. See [20] for further details.

The contact forces are shown in Fig. 13 and defined as

$$\begin{aligned} \boldsymbol {\varGamma }_{\mathrm{contact}} = J_p^T \mathbf{F}_{\mathrm{contact}} = J_p^T \left [ \begin{array}{c} \mathbf{f}_{C_{C}} \\ \mathbf{m}_{C} \\ \mathbf{f}_{D_{C}} \\ \mathbf{m}_{D} \\ \mathbf{f}_{R} \end{array} \right ] \end{aligned}$$
(37)
$$\begin{aligned} \boldsymbol {v}_p = \left [ \begin{array}{c} \boldsymbol {v}_{C_{C}} \\ \boldsymbol {v}_{D_{C}} \\ \boldsymbol {v}_{R} \end{array} \right ] = J_p \ \dot{\mathbf{q}} \end{aligned}$$
(38)

where v p contains the inertial velocities of the contact points C C and D C , as well as that of point R. Although all possible contacts are shown in (38), the ones that actually contact/impact are selected using the impact Jacobian J p .

The contact forces are calculated as

$$\begin{aligned} {\mathbf{F}}_{\mathrm{contact}} =& - \bigl( J_p A^{-1} J_p^T \bigr)^{-1} \bigl( \dot{J}_p \dot{\mathbf{q}} + J_p A^{-1} ( \boldsymbol {\varGamma }_{\mathit{fccB}} - \mathbf{b} ) \bigr) \end{aligned}$$
(39)
$$\begin{aligned} \boldsymbol {\varGamma }_{\mathit{fccB}} =& \boldsymbol {\varGamma }_{\mathrm{friction}} + \boldsymbol {\varGamma }_{\mathrm{charge}} + \boldsymbol {\varGamma }_{\mathrm{conform}} + \boldsymbol {\varGamma }_{\mathrm{Brown}} \end{aligned}$$
(40)

Since F contact depends on the other forces, its large portion is found by substituting the other large forces into (39) and solving.

1.5 A.5 Brownian motion

Random forces and moments in the model representing Brownian motion are implemented as Gaussian white noise. They act at and about the mass center of each body, as shown in Fig. 14.

Fig. 14
figure 14

Brownian motion for myosin V

The random forces and moments shown in Fig. 14 representing Brownian motion are defined, for example, as

$$ \begin{aligned} {\mathbf{f}}_{Co} & = C_{o1}(t) \mathbf{N}_1 + C_{o2}(t) \mathbf{N}_2 + C_{o3}(t) \mathbf{N}_3 \\ {\mathbf{m}}_C & = \bar{L}_C C_{o1}(t) \mathbf{N}_1 + \bar{L}_C C_{o2}(t) \mathbf{N}_2 + \bar{L}_C C_{o3}(t) \mathbf{N}_3 \end{aligned} $$
(41)

where \(\bar{L}_{C}\) is a characteristic length of body C. Similar forces act on the other bodies. The C oi (t) represent forces produced by randomly fluctuating thermal noise. Each component of the random force and moment is treated independently as a normally distributed random variable [66]. They have the following expectations E[⋅] or weighted average values:

$$ E \bigl[ C_{oi}(t) \bigr] = \bigl\langle C_{oi}(t) \bigr\rangle = 0 = \mu $$
(42)

and are governed by a fluctuation–dissipation relation expressed as

$$ E \bigl[ C_{oi}(t_1) C_{oj}(t_2) \bigr] = 2 \beta k_B \mathrm{T} \delta(t_1-t_2) \delta_{i,j} $$
(43)

where k B and T are the Boltzmann constant and absolute temperature [59, 66]. The relation in (43) implies that there is no time dependency between the random process over time; the random sequence of forces does not repeat regularly.

In addition, (42) and (43) imply

$$ E\bigl[ C_{oi}^2(t) \bigr] = 2 \beta k_B \mathrm{T} = \mathrm {Var}\bigl(C_{oi}(t)\bigr) = \sigma^2 $$
(44)

which is the variance of C oi . Thus, the C oi can be generated using the Matlab function normrnd(μ,σ,…), which generates random variables with normal distribution.

The collection of random forces comprise Γ Brown. These randomly fluctuating discontinuous functions slow numerical integration so that each random variable is held constant during a single integration step; the random variable is updated at the beginning of each step. Thus, the value of each random variable is known before the integration step, and the decomposed value of the random force must equal it. This is accomplished by defining

$$ \boldsymbol {\varGamma }_{\mathrm{Brown}} = (\beta_1 +1) {R}_{nd} \mathbf{r}_{nd} = (\beta_1 +1) {R}_{nd} \left [ \begin{array}{c} \bar{C}_{o1} \\ \bar{C}_{o2} \\ \bar {C}_{o3} \\ \vdots \end{array} \right ] $$
(45)

where R nd transforms the random forces into generalized active forces, and

$$ C_{oi} = (\beta_1+1) \bar{C}_{oi} $$
(46)

and likewise for the other random forces. An example of the random forces used is given in Fig. 15.

Fig. 15
figure 15

Random forces acting on head C in Fig. 8

Appendix B: Simulation algorithm

Here it is assumed that the equations of motion in (15) have been expressed as a system of first-order differential equations in order to integrate them. This can be accomplished using a simple transformation such as

$$ \dot{\mathbf{q}} = \mathbf{u} \qquad \ddot{\mathbf{q}} = \dot{ \mathbf{u}} $$
(47)

such that

$$ \dot{\mathbf{x}}(t) = \left [ \begin{array}{c} \dot{\mathbf{q}}(t) \\ \dot{\mathbf{u}}(t) \end{array} \right ] = \left [ \begin{array}{c} \mathbf{u}(t) \\ A^{-1} ( {\sum}{\boldsymbol{\varGamma}} - \mathbf{b} ) \end{array} \right ] . $$
(48)

Nominally at each time step, \(\dot{\mathbf{x}}(t)\) is calculated and used to increment x(t) yielding x(tt). This calculation is modified here to enforce nonpenetration and docking constraints. At each time step, if a contact or impact is identified, the postimpact state \(\dot{\mathbf{x}}(t+\epsilon)\) is calculated, and the numerical integration is carried on from there; ϵ is considered an infinitesimally small amount of time.

The impact/contact resolution process involves identifying when an impact or contact occurs, resolving the postimpact/contact velocity, and possibly restarting the integration with the updated postimpact velocity. The details involved in addressing nonpenetration include: (1) treating contact as a succession of impacts, (2) using an impact law that addresses simultaneous impacts and contacts, (3) using an adaptive and event-driven numerical integration scheme, and (4) calculating contact forces. The details of this process, discussed in another paper [20], are too lengthy to duplicate here, so only a brief description of them is given below. The parameters for the simulation are given in Table 2.

2.1 B.1 Contact as a succession of impacts

Herein, a distinction is drawn between impact and contact forces, which are of short and long duration, relative to each other and to the time step of the numerical integration. Impact forces correspond to abrupt changes in velocity, as occurs during docking with the actin filament, whereas contact forces do not. Treating contact as a succession of impacts allows them both to be addressed within the same mathematical framework. This facilitates the development of an impact law, discussed in detail in another paper [20], which can address simultaneous contact and impact, as occurs during protein locomotion.

2.2 B.2 Simultaneous impact law

This law is intended to enforce nonpenetration and other motion constraints. It is assumed, for simplicity, that the head and actin filament are lubricated sufficiently by the intracellular fluid to make them frictionless. When undocked, only nonpenetration constraints are enforced on the heads. When docked, a head is immobilized by enforcing nonpenetration, nonsliding, and nonrotation constraints. Point R is prevented from penetrating the filament as well. Again, for simplicity, it is assumed that when a head impacts the actin filament, it does not rebound. The associated constraint forces are calculated for use in determining contact and impact events.

2.3 B.3 Adaptive event-driven simulation

An adaptive numerical integrator is used because it can address phenomena occurring at large and small time scales while maintaining a specified accuracy. Matlab’s ode45.m is used here along with an event function for finding contact and impact events. The simulation is stopped and restarted at each impact event. However, since contact is treated as a succession of impacts, this would yield a very slow numerical integration.

This is avoided by monitoring contact forces in the event function. It is clear when an impact event occurs because the contact forces jump from zero to nonzero values; during contact, they sustain nonzero values. The discontinuities in force and velocity caused by an impact event will create difficulties for an adaptive numerical integrator, so it must be stopped and restarted. During contact, the forces and velocities are more continuous and do not create large errors in the numerical integration, and thus stopping and restarting the simulation is not required. This practice increases the speed of the simulation considerably.

Appendix C: Error analysis

This section discusses in detail the concept of error analysis used in the numerical integration. It is difficult to assess the error in the first-order model because it inherently involves abrupt, discontinuous changes in forces and velocities. This creates difficulties in numerically integrating the checking functions used herein, which increases the run time significantly. Therefore, this error analysis is only applied to the second-order model. The correctness of the second-order model should carry over to the first-order model since they are identical, except that one omits the generalized inertia forces.

Since the motor protein’s environment contains nonconservative forces, such as dissipative viscous friction, the system energy is not conserved. Thus the work–energy theorem [10] is used to account for the effect of nonconservative forces:

$$ T_1 + W_{1 \rightarrow2} = T_2 $$
(49)

where T i is the kinetic energy at state i, W 1→2 is the work done in moving from state 1 to 2 including the work done by nonconservative forces. The total work can be calculated as

$$ W_{1 \rightarrow2} = {\displaystyle\int^2_1} \dot{W}_{1 \rightarrow 2} \, dt , \qquad \dot{W}_{1 \rightarrow2} = {\textstyle\sum} {\mathbf{F}} \cdot \dot{\mathbf{q}} $$
(50)

where ∑F represents the generalized active forces in (15). The time derivative \(\dot{W}_{1 \rightarrow2}\) can be numerically integrated along with the model. Subtracting this work from the kinetic energy at state 2 should yield a constant number to within the specified error tolerances. In order to examine the errors, a checking function is defined as follows:

$$ {\mathrm{check}}_{1\rightarrow i} = T_i - ( T_1 + W_{1 \rightarrow i} ) $$
(51)

The evaluation of the checking functions involves a difficulty due to the calculation of the work done by the contact forces during the stepping. This is because of the discontinuities that contact/impact create in velocities and forces. The work done by contact forces cannot be captured in (50) while the change in velocities appears in the kinetic energy; that leads to a gap (jump) in the checking function (51). In order to address this issue, the difference between the kinetic energy at the moment of contact/impact, which equals to the nonconservative work, is added to the check function when the numerical is stopped and restarted at the impact event, see Appendix B, to compensate for the gap.

The check function is examined for a single deterministic step, corresponding to Fig. 5, of motor protein locomotion as shown in Fig. 16. The deterministic step is chosen because the simulation of Brownian motion is very discontinuous and difficult to integrate along with the check function. However, the deterministic step gives confidence in the results from the simulations with Brownian motion. It involves large numbers, which may give the appearance that the integrator is not maintaining the desired accuracy, even though it may be for the variables in \(\dot{\mathbf {x}}\); see (48). This situation can be remedied by normalizing the checking functions using a characteristic system value. The system energy E i at state i is defined as

$$ E_i = T_i + V_i $$
(52)

where V i is the potential energy. Since energies are being evaluated, as shown in Fig. 17, the system energy E i is used for this purpose. The characteristic energy is taken from Fig. 17 as 105.

$$ {\mathrm{chk}}_{1 \rightarrow i} = \frac{\mathrm{check}_{1 \rightarrow i}}{E_i} $$
(53)
Fig. 16
figure 16

The checking function of myosin V’s locomotion during one deterministic step that includes drifting toward actin filament, docking to binding sites, and taking one step

Fig. 17
figure 17

Total energy of myosin V during one step

The normalized checking function chk1→i is shown in Fig. 18. Notice that the scaled checking function satisfies the specified AbsTol and RelTol, whereas the unscaled one, Fig. 16, does not. The check function in Fig. 18 shows that the desired tolerance is maintained even in the presence of contact and impact forces. This gives confidence in the model’s correctness and the techniques involved in the numerical integration.

Fig. 18
figure 18

The normalized checking function remains beneath the specified AbsTol=10−5 and RelTol=10−6 during the simulation

Appendix D: Units

This appendix gives a brief summary of the units used in the simulation. The unit basis consists of the attogram (ag), the nanometer (nm), the millisecond (ms), and the radian (rad). With this choice the force unit becomes

$$ 1 \mathrm{N} = 1~\frac{\mathrm{kg}\cdot \mathrm{m}}{\mathrm{s}^2} \ \frac{10^{21}\,\mathrm{ag} }{\mathrm{kg}} \ \frac{ 10^9 \,\mathrm{nm} }{\mathrm{m }} \ \frac{\mathrm{s}^2 }{ (10^3~\mathrm{ms} )^2 } = 10^{24}~\frac{\mathrm{ag} \cdot \mathrm{nm} }{\mathrm{ms}^2} $$
(54)

and the unit of work becomes

$$ 1~\mathrm{J} = 1\ \frac{\mathrm{kg \cdot m}^2}{\mathrm{s}^2} = 10^{33}~\frac{\mathrm{ag} \cdot \mathrm{nm}^2 }{\mathrm{ms}^2} $$
(55)

These values are chosen so that the masses have values near unity and to improve the stability of the numerical integration.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bowling, A., Haghshenas-Jaryani, M. A multiscale modeling approach for biomolecular systems. Multibody Syst Dyn 33, 333–365 (2015). https://doi.org/10.1007/s11044-014-9431-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11044-014-9431-x

Keywords

Navigation