Skip to main content
Log in

Comparison of explicit and mean-field models of cytoskeletal filaments with crosslinking motors

  • Regular Article - Living Systems
  • Published:
The European Physical Journal E Aims and scope Submit manuscript

Abstract

In cells, cytoskeletal filament networks are responsible for cell movement, growth, and division. Filaments in the cytoskeleton are driven and organized by crosslinking molecular motors. In reconstituted cytoskeletal systems, motor activity is responsible for far-from-equilibrium phenomena such as active stress, self-organized flow, and spontaneous nematic defect generation. How microscopic interactions between motors and filaments lead to larger-scale dynamics remains incompletely understood. To build from motor–filament interactions to predict bulk behavior of cytoskeletal systems, more computationally efficient techniques for modeling motor–filament interactions are needed. Here, we derive a coarse-graining hierarchy of explicit and continuum models for crosslinking motors that bind to and walk on filament pairs. We compare the steady-state motor distribution and motor-induced filament motion for the different models and analyze their computational cost. All three models agree well in the limit of fast motor binding kinetics. Evolving a truncated moment expansion of motor density speeds the computation by \(10^3\)\(10^6\) compared to the explicit or continuous-density simulations, suggesting an approach for more efficient simulation of large networks. These tools facilitate further study of motor–filament networks on micrometer to millimeter length scales.

Graphic abstract

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

Similar content being viewed by others

Data Availability Statement

This manuscript has associated data in a data repository. [Authors’ comment: The datasets generated and analysed for this manuscript are stored in a data repository. Data is available from the corresponding author upon request.]

References

  1. Jonathon Howard, Mechanics of Motor Proteins and the Cytoskeleton (Sinauer Associates Publishers, Sunderland, Mass, 2001)

    Google Scholar 

  2. Dennis Bray, Cell Movements: From Molecules to Motility, 2nd edn. (Garland Pub, New York, 2001)

    Google Scholar 

  3. Laurent Blanchoin, Rajaa Boujemaa-Paterski, Cécile Sykes, Julie Plastino, Actin dynamics, architecture, and mechanics in cell motility. Physiol. Rev. 94(1), 235–263 (2014)

    Article  Google Scholar 

  4. Thomas D. Pollard, John A. Cooper, Actin, a central player in cell shape and movement. Science 326(5957), 1208–1212 (2009)

    Article  ADS  Google Scholar 

  5. J.Richard McIntosh, Maxim I Molodtsov, Fazly I Ataullakhanov, Biophysics of mitosis. Q. Rev. Biophys. 45(2), 147–207 (2012)

    Article  Google Scholar 

  6. A.F. Huxley, Muscle structure and theories of contraction. Prog. Biophys. Biophys. Chem 7, 255–318 (1957)

    Article  Google Scholar 

  7. A.F. Huxley, S. Tideswell, Filament compliance and tension transients in muscle. J. Muscle Res. Cell Motil. 17(4), 507–511 (1996)

    Article  Google Scholar 

  8. A.F. Huxley, S. Tidswell, Rapid regeneration of power stroke in contracting muscle by attachment of second myosin head. J. Muscle Res. Cell Motil. 18(1), 111–114 (1997)

    Article  Google Scholar 

  9. Stephanie L. Gupton, Clare M. Waterman-Storer, Spatiotemporal feedback between actomyosin and Focal-Adhesion systems optimizes rapid cell migration. Cell 125(7), 1361–1374 (2006)

    Article  Google Scholar 

  10. Maxime F. Fournier, Roger Sauser, Davide Ambrosi, Jean-Jacques Meister, Alexander B. Verkhovsky, Force transmission in migrating cells. J. Cell Biol. 188(2), 287–297 (2010)

    Article  Google Scholar 

  11. Erin L. Barnhart, Kun-Chun Lee, Kinneret Keren, Alex Mogilner, Julie A. Theriot, An adhesion-dependent switch between mechanisms that determine motile cell shape. PLOS Biol. 9(5), e1001059 (2011)

    Article  Google Scholar 

  12. G. Laevsky, Cross-linking of actin filaments by myosin II is a major contributor to cortical integrity and cell motility in restrictive environments. J. Cell Sci. 116(18), 3761–3770 (2003)

    Article  Google Scholar 

  13. Iain Hagan, Mitsuhiro Yanagida, Kinesin-related cut 7 protein associates with mitotic and meiotic spindles in fission yeast. Nature 356(6364), 74 (1992)

    Article  ADS  Google Scholar 

  14. W.S. Saunders, D. Koshland, D. Eshel, I.R. Gibbons, M.A. Hoyt, Saccharomyces cerevisiae kinesin- and dynein-related proteins required for anaphase chromosome segregation. J. Cell Biol. 128(4), 617–624 (1995)

    Article  Google Scholar 

  15. Tarun M. Kapoor, Thomas U. Mayer, Margaret L. Coughlin, Timothy J. Mitchison, Probing spindle assembly mechanisms with monastrol, a small molecule inhibitor of the mitotic kinesin, Eg5. J. Cell Biol. 150(5), 975–988 (2000)

    Article  Google Scholar 

  16. Shang Cai, Lesley N. Weaver, Stephanie C. Ems-McClung, Claire E. Walczak, Kinesin-14 family proteins HSET/XCTK2 control spindle length by cross-linking and sliding microtubules. Mol. Biol. Cell 20(5), 1348–1359 (2008)

    Article  Google Scholar 

  17. R. Wollman, G. Civelekoglu-Scholey, J.M. Scholey, A. Mogilner, Reverse engineering of force integration during mitosis in the Drosophila embryo. Mol. Syst. Biol. 4(1), 195 (2008)

    Article  Google Scholar 

  18. Gul Civelekoglu-Scholey, Jonathan M. Scholey, Mitotic force generators and chromosome segregation. Cell. Mol. Life Sci. 67(13), 2231–2250 (2010)

    Article  Google Scholar 

  19. Zhen-Yu. She, Wan-Xi Yang, Molecular mechanisms of kinesin-14 motors in spindle assembly and chromosome segregation. J. Cell Sci. 130(13), 2097–2110 (2017)

    Article  Google Scholar 

  20. Kruno Vukušić, Renata Buda, Agneza Bosilj, Ana Milas, Nenad Pavin, Iva M. Tolić, Microtubule sliding within the bridging fiber pushes kinetochore fibers apart to segregate chromosomes. Dev. Cell 43(1), 11–23.e6 (2017)

    Article  Google Scholar 

  21. Sujoy Ganguly, Lucy S. Williams, Isabel M. Palacios, Raymond E. Goldstein, Cytoplasmic streaming in Drosophila oocytes varies with kinesin activity and correlates with the microtubule cytoskeleton architecture. Proc. Nat. Acad. Sci. 109(38), 15109–15114 (2012)

    Article  ADS  Google Scholar 

  22. Peter Satir, Studies on cilia. J. Cell Biol. 39(1), 77–94 (1968)

    Article  Google Scholar 

  23. Keith E. Summers, I.R. Gibbons, Adenosine triphosphate-induced sliding of tubules in trypsin-treated flagella of sea-urchin sperm. Proc. Nat. Acad. Sci. U.S. A. 68(12), 3092–3096 (1971)

    Article  ADS  Google Scholar 

  24. Stephen M. King, Turning dyneins off bends cilia. Cytoskeleton 75(8), 372–381 (2018)

    Article  Google Scholar 

  25. F.J. Nedelec, T. Surrey, A.C. Maggs, S. Leibler, Self-organization of microtubules and motors. Nature 389(6648), 305–308 (1997)

    Article  ADS  Google Scholar 

  26. Thomas Surrey, François Nédélec, Stanislas Leibler, Eric Karsenti, Physical properties determining self-organization of motors and microtubules. Science 292(5519), 1167–1171 (2001)

    Article  ADS  Google Scholar 

  27. F. Backouche, L. Haviv, D. Groswasser, A. Bernheim-Groswasser, Active gels: dynamics of patterning and self-organization. Phys. Biol. 3(4), 264–273 (2006)

    Article  ADS  Google Scholar 

  28. T. Sanchez, D.T.N. Chen, S.J. DeCamp, M. Heymann, Z. Dogic, Spontaneous motion in hierarchically assembled active matter. Nature 491(7424), 431–434 (2012)

    Article  ADS  Google Scholar 

  29. Amin Doostmohammadi, Jordi Ignés-Mullol, Julia M. Yeomans, Francesc Sagués, Active nematic. Nat. Commun. 9(1), 3246 (2018)

    Article  ADS  Google Scholar 

  30. Linnea M. Lemma, Stephen J. DeCamp, Zhihong You, Luca Giomi, Zvonimir Dogic, Statistical properties of autonomous flows in 2D active nematics. Soft Matter 15(15), 3264–3272 (2019)

    Article  ADS  Google Scholar 

  31. G. Duclos, R. Adkins, D. Banerjee, M.S.E. Peterson, M. Varghese, I. Kolvin, A. Baskaran, R.A. Pelcovits, T.R. Powers, A. Baskaran, F. Toschi, M.F. Hagan, S.J. Streichan, V. Vitelli, D.A. Beller, Z. Dogic, Topological structure and dynamics of three-dimensional active nematics. Science 367(6482), 1120–1124 (2020)

    Article  ADS  Google Scholar 

  32. Jan Brugués, Valeria Nuzzo, Eric Mazur, Daniel J. Needleman, Nucleation and transport organize microtubules in metaphase spindles. Cell 149(3), 554–564 (2012)

    Article  Google Scholar 

  33. Johanna Roostalu, Jamie Rickman, Claire Thomas, François Nédélec, Thomas Surrey, Determinants of Polar versus nematic organization in networks of dynamic microtubules and mitotic motors. Cell 175(3), 796–808.e14 (2018)

    Article  Google Scholar 

  34. Kimberly L. Weirich, Kinjal Dasbiswas, Thomas A. Witten, Suriyanarayanan Vaikuntanathan, Margaret L. Gardel, Self-organizing motors divide active liquid droplets. Proc. Nat. Acad. Sci. 116(23), 11125–11130 (2019)

    Article  ADS  Google Scholar 

  35. Francois Nedelec, Dietrich Foethke, Collective langevin dynamics of flexible cytoskeletal fibers. New J. Phys. 9(11), 427 (2007)

    Article  ADS  Google Scholar 

  36. Konstantin Popov, James Komianos, Garegin A. Papoian, MEDYAN: mechanochemical simulations of contraction and polarity alignment in actomyosin networks. PLOS Comput. Biol. 12(4), e1004877 (2016)

    Article  ADS  Google Scholar 

  37. Simon L. Freedman, Shiladitya Banerjee, Glen M. Hocky, Aaron R. Dinner, A versatile framework for simulating the dynamic mechanical structure of cytoskeletal networks. Biophys. J. 113(2), 448–460 (2017)

    Article  ADS  Google Scholar 

  38. D.A. Head, W.J. Briels, Gerhard Gompper, Nonequilibrium structure and dynamics in a microscopic model of thin-film active gels. Phys. Rev. E 89(3), 032705 (2014)

    Article  ADS  Google Scholar 

  39. Igor S. Aranson, Lev S. Tsimring, Pattern formation of microtubules and motors: inelastic interaction of polar rods. Phys. Rev. E 71(5), 050901 (2005)

    Article  ADS  Google Scholar 

  40. K. Kruse, J.F. Joanny, F. Jülicher, J. Prost, K. Sekimoto, Generic theory of active polar gels: a paradigm for cytoskeletal dynamics. Eur. Phys. J. E 16(1), 5–16 (2005)

    Article  Google Scholar 

  41. David Saintillan, Michael J. Shelley, Instabilities and pattern formation in active particle suspensions: kinetic theory and continuum simulations. Phys. Rev. Lett. 100(17), 178103 (2008)

    Article  ADS  Google Scholar 

  42. Luca Giomi, Mark J. Bowick, Xu Ma, M. Cristina Marchetti, Defect annihilation and proliferation in active nematics. Phys. Rev. Lett. 110(22), 228101 (2013)

    Article  ADS  Google Scholar 

  43. Tong Gao, Robert Blackwell, Matthew A. Glaser, M.D. Betterton, Michael J. Shelley, Multiscale modeling and simulation of microtubule-motor-protein assemblies. Phys. Rev. E 92(6), 062709 (2015)

    Article  ADS  Google Scholar 

  44. D. White, G. de Vries, J. Martin, A. Dawes, Microtubule patterning in the presence of moving motor proteins. J. Theor. Biol. 382, 81–90 (2015)

    Article  MATH  Google Scholar 

  45. Ivan Maryshev, Davide Marenduzzo, Andrew B. Goryachev, Alexander Morozov, Kinetic theory of pattern formation in mixtures of microtubules and molecular motors. Phys. Rev. E 97(2), 022412 (2018)

    Article  ADS  Google Scholar 

  46. Sebastian Fürthauer, Bezia Lemma, Peter J. Foster, Stephanie C. Ems-McClung, Yu. Che-Hang, Claire E. Walczak, Zvonimir Dogic, Daniel J. Needleman, Michael J. Shelley, Self-straining of actively crosslinked microtubule networks. Nat. Phys. 15(12), 1295–1300 (2019)

    Article  Google Scholar 

  47. Falko Ziebert, Igor S. Aranson, Lev S. Tsimring, Effects of cross-links on motor-mediated filament organization. J. Phys. 9(11), 421 (2007)

    Google Scholar 

  48. L. Giomi, M.J. Bowick, X. Ma, M.C. Marchetti, Defect annihilation and proliferation in active nematics. Phys. Rev. Lett. 110(22), 228101 (2013)

    Article  ADS  Google Scholar 

  49. Martin Lenz, Geometrical origins of contractility in disordered actomyosin networks. Phys. Rev. X 4(4), 041002 (2014)

    Google Scholar 

  50. K Kruse and F Jülicher. Actively Contracting Bundles of Polar Filaments. Technical report, 2000

  51. K. Kruse, J.F. Joanny, F. Jülicher, J. Prost, K. Sekimoto, Generic theory of active polargels: a paradigm for cytoskeletal dynamics. Eur. Phys. J. E 16(1), 5–16 (2005)

    Article  Google Scholar 

  52. A. Ahmadi, T.B. Liverpool, M.C. Marchetti, Nematic and polar order in active filament solutions. Phys. Rev. E 72(6), 060901 (2005)

    Article  ADS  Google Scholar 

  53. M.C. Ahmadi, A. Marchetti, T.B. Liverpool, Hydrodynamics of isotropic and liquid crystalline active polymer solutions. Phys. Rev. E 74(6), 061913 (2006)

    Article  ADS  Google Scholar 

  54. T.B. Liverpool, M.C. Marchetti, Bridging the microscopic and the hydrodynamic in active filament solutions. Europhys. Lett. 69(5), 846–852 (2005)

    Article  ADS  Google Scholar 

  55. S. Swaminathan, F. Ziebert, I.S. Aranson, D. Karpeev, Motor-mediated microtubule self-organization in dilute and semi-dilute filament solutions. Math. Model. Nat. Phenom. 6(1), 119–137 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  56. Tong Gao, Robert Blackwell, Matthew A. Glaser, M.D. Betterton, Michael J. Shelley, Multiscale polar theory of microtubule and motor-protein assemblies. Phys. Rev. Lett. 114(4), 048101 (2015)

    Article  ADS  Google Scholar 

  57. Tong Gao, Meredith D. Betterton, An-Sheng Jhang, Michael J. Shelley, Analytical structure, dynamics, and coarse graining of a kinetic model of an active fluid. Phys. Rev. Fluids 2(9), 093302 (2017)

    Article  ADS  Google Scholar 

  58. Robert Blackwell, Oliver Sweezy-Schindler, Christopher Baldwin, Loren E. Hough, Matthew A. Glaser, M.D. Betterton, Microscopic origins of anisotropic active stress in motor-driven nematic liquid crystals. Soft Matter 12(10), 2676–2687 (2016)

    Article  ADS  Google Scholar 

  59. Robert Blackwell, Christopher Edelmaier, Oliver Sweezy-Schindler, Adam Lamson, Zachary R. Gergely, Eileen O’Toole, Ammon Crapo, Loren E. Hough, J. Richard McIntosh, Matthew A. Glaser, Meredith D. Betterton, Physical determinants of bipolar mitotic spindle assembly and stability in fission yeast. Sci. Adv. 3(1), e1601603 (2017)

    Article  ADS  Google Scholar 

  60. Sergio A. Rincon, Adam Lamson, Robert Blackwell, Viktoriya Syrovatkina, Vincent Fraisier, Anne Paoletti, Meredith D. Betterton, Phong T. Tran, Kinesin-5-independent mitotic spindle assembly requires the antiparallel microtubule crosslinker Ase1 in fission yeast. Nat. Commun. 8, 15286 (2017)

    Article  ADS  Google Scholar 

  61. Adam R. Lamson, Christopher J. Edelmaier, Matthew A. Glaser, Meredith D. Betterton, Theory of cytoskeletal reorganization during cross-linker-mediated mitotic spindle assembly. Biophys. J. 116(9), 1719–1731 (2019)

    Article  ADS  Google Scholar 

  62. C. Edelmaier, A.R. Lamson, Z.R. Gergely, S. Ansari, R. Blackwell, J.R. McIntosh, M.A. Glaser, M.D. Betterton, Mechanisms of chromosome biorientation and bipolar spindle assembly analyzed by computational modeling. eLife 9, e48787 (2020)

    Article  Google Scholar 

  63. Torsten Wittmann, Anthony Hyman, Arshad Desai, The spindle: a dynamic assembly of microtubules and motors. Nat. Cell Biol. 3(1), E28–E34 (2001)

    Article  Google Scholar 

  64. Viktoria Wollrab, Julio M. Belmonte, Lucia Baldauf, Maria Leptin, François Nédeléc, Gijsje H. Koenderink, Polarity sorting drives remodeling of actin-myosin networks. J. Cell Sci. 132, jcs219717. https://doi.org/10.1242/jcs.219717

  65. Clifford P. Brangwynne, Gijsje H. Koenderink, Frederick C. MacKintosh, David A. Weitz, Cytoplasmic diffusion: molecular motors mix it up. J. Cell Biol. 183(4), 583–587 (2008)

    Article  Google Scholar 

  66. W.K. den Otter, T. Yu-Guo, J.T. Padding, J.K.G. Dhont, W.J. Briels, Brownian dynamics simulations of the self- and collective rotational diffusion coefficients of rigid long thin rods. J. Chem. Phys. 122(24), 244903 (2005)

    Article  ADS  Google Scholar 

  67. Koen Visscher, Mark J. Schnitzer, Steven M. Block, Single kinesin molecules studied with a molecular force clamp. Nature 400(6740), 184–189 (1999)

    Article  ADS  Google Scholar 

  68. Stefan Klumpp, Reinhard Lipowsky, Cooperative cargo transport by several molecular motors. Proc. Nat. Acad. Sci. U. S. A. 102(48), 17284–17289 (2005)

    Article  ADS  Google Scholar 

  69. Melanie J.I. Muller, Stefan Klumpp, Reinhard Lipowsky, Tug-of-war as a cooperative mechanism for bidirectional cargo transport by molecular motors. Proc. Nat. Acad. Sci. U. S. A. 105(12), 4609–4614 (2008)

    Article  ADS  Google Scholar 

  70. Ambarish Kunwar, Suvranta K. Tripathy, Xu Jing, Michelle K. Mattson, Preetha Anand, Roby Sigua, Michael Vershinin, Richard J. McKenney, Clare C. Yu, Alexander Mogilner, Steven P. Gross, Mechanical stochastic tug-of-war models cannot explain bidirectional lipid-droplet transport. Proc. Nat. Acad. Sci. U. S. A. 108(47), 18960–18965 (2011)

    Article  ADS  Google Scholar 

  71. Sebastián Bouzat, Models for microtubule cargo transport coupling the Langevin equation to stochastic stepping motor dynamics: caring about fluctuations. Phys. Rev. E 93, 12401 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  72. Si Kao Guo, Xiao Xuan Shi, Peng Ye Wang, Ping Xie, Force dependence of unbinding rate of kinesin motor during its processive movement on microtubule. Biophys. Chem. 253, 106216 (2019)

    Article  Google Scholar 

  73. Göker Arpağ, Stephen R. Norris, S. Iman Mousavi, Virupakshi Soppina, Kristen J. Verhey, William O. Hancock, Erkan Tüzel, Motor dynamics underlying cargo transport by pairs of kinesin-1 and kinesin-3 motors. Biophys. J. 116(6), 1115–1126 (2019)

    Article  ADS  Google Scholar 

  74. Stephan W. Grill, Karsten Kruse, Frank Jülicher, Theory of mitotic spindle oscillations. Phys. Rev. Lett. 94(10), 108104 (2005)

    Article  ADS  Google Scholar 

  75. Robert Blackwell, Oliver Sweezy-Schindler, Christopher Edelmaier, Zachary R. Gergely, Patrick J. Flynn, Salvador Montes, Ammon Crapo, J. Alireza Doostan, Richard McIntosh, Matthew A. Glaser, Meredith D. Betterton, Contributions of microtubule dynamic instability and rotational diffusion to kinetochore capture. Biophys. J. 112(3), 552–563 (2017)

    Article  ADS  Google Scholar 

  76. Shenshen Wang, Peter G. Wolynes, On the spontaneous collective motion of active matter. Proc. Nat. Acad. Sci. 108(37), 15184–15189 (2011)

    Article  ADS  Google Scholar 

  77. A.J.T.M. Mathijssen, A. Doostmohammadi, J.M. Yeomans, T.N. Shendruk, Hydrodynamics of micro-swimmers in films. J. Fluid Mech. 806, 35–70 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. Fundamental algorithms for scientific computing in python, P Virtanen, R Gommers, T E. Oliphant, M Haberland, T Reddy, D Cournapeau, E Burovski, P Peterson, W Weckesser, J Bright, S J. van der Walt, M Brett, J Wilson, K. J Millman, N Mayorov, A R. J. Nelson, E Jones, R Kern, E Larson, CJ Carey, l Polat, Yu Feng, E W. Moore, J V erPlas, D Laxalde, J Perktold, R Cimrman, I Henriksen, E. A. Quintero, C R Harris, A M. Archibald, A H. Ribeiro, F Pedregosa, P van Mulbregt, and SciPy 1. 0 Contributors. Scipy 1.0. Nature Methods 17, 261–272 (2020)

  79. C. Alan, Hindmarsh, ODEPACK, a Systematized Collection of ODE Solvers, in Scientific Computing, ed. by R.S. Stepleman, et al. (Science and Education Publishing, North-Holland, Amsterdam, 1983), pp. 55–64

    Google Scholar 

  80. Alberts. Molecular biology of the cell, 5th edition by B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts, and P. Walter, Biochem. Mol. Biol. Edu. 36(4), 317–318 (2008)

    Article  Google Scholar 

  81. K. Kawaguchi, S. Ishiwata, Nucleotide-dependent single- to double-headed binding of kinesin. Science 291(5504), 667–669 (2001)

    Article  ADS  Google Scholar 

  82. A.S. Kashina, R.J. Baskin, D.G. Cole, K.P. Wedaman, W.M. Saxton, J.M. Scholey, A bipolar kinesin. Nature 379(6562), 270–272 (1996)

    Article  ADS  Google Scholar 

  83. Adina Gerson-Gurwitz, Christina Thiede, Natalia Movshovich, Vladimir Fridman, Maria Podolskaya, Tsafi Danieli, Stefan Lakämper, Dieter R. Klopfenstein, Christoph F. Schmidt, Larisa Gheber, Directionality of individual kinesin-5 Cin8 motors is modulated by loop 8, ionic strength and microtubule geometry. EMBO J. 30(24), 4942–4954 (2011)

    Article  Google Scholar 

  84. Megan T. Valentine, Polly M. Fordyce, Troy C. Krzysiak, Susan P. Gilbert, Steven M. Block, Individual dimers of the mitotic kinesin motor Eg5 step processively and support substantial loads in vitro. Nat. Cell Biol. 8(5), 470–476 (2006)

    Article  Google Scholar 

  85. J.C. Cochran, Kinesin motor enzymology: chemistry, structure, and physics of nanoscale molecular machines. Biophys. Rev. 7(3), 269–299 (2015)

    Article  Google Scholar 

  86. Yuta Shimamoto, Scott Forth, Tarun M. Kapoor, Measuring pushing and braking forces generated by ensembles of kinesin-5 crosslinking two microtubules. Dev. Cell 34(6), 669–681 (2015)

    Article  Google Scholar 

  87. H.I. Siyyam, M.I. Syam, The modified trapezoidal rule for line integrals. J. Comput. Appl. Math. 84(1), 1–14 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  88. Randall J. LeVeque, Finite Difference Methods for Ordinary and Partial Differential Equations: Steady-State and Time-Dependent Problems (Society for Industrial and Applied Mathematics, Philadelphia, PA, 2007)

    Book  MATH  Google Scholar 

Download references

Acknowledgements

This work was supported by NSF grants DMR-1725065 (MDB), DMS-1620003 (MAG and MDB), DMS-1620331(MJS), DMR-1420736 (MAG and ARL), DMS-1463962(MJS), and DMR-1420073 (MJS); NIH grant R01GM124371 (MDB); and a fellowship provided by matching funds from the NIH/University of Colorado Biophysics Training Program (ARL). Simulations used the Summit supercomputer, supported by NSF grants ACI-1532235 and ACI-1532236.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Adam R. Lamson.

Ethics declarations

Author contribution statement

ARL, MAG, and MDB formulated the explicit motor binding model and kinetic Monte-Carlo algorithm. ARL and JMM wrote the Brownian dynamics simulation code. ARL, FF, MAG, MJS, and MDB derived the MFMD model and numerical integration scheme. ARL, MAG, and MDB derived the moment expansion model. ARL ran the simulations. ARL, MAG, and MDB analyzed and compared results. ARL and MDB drafted the manuscript. ARL, MAG, MJS, and MDB edited the manuscript.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary material 1 (mp4 292 KB)

Supplementary material 2 (mp4 296 KB)

Supplementary material 3 (mp4 233 KB)

Supplementary material 4 (mp4 242 KB)

Supplementary material 5 (mp4 296 KB)

Supplementary material 6 (mp4 281 KB)

Appendices

Appendices

A Determining the time-step for binding

Our kinetic Monte Carlo algorithm assumes that multiple binding/unbinding events do not occur in the same time step \(\Delta t\). As \(\Delta t\) becomes large relative to the kinetic rates, this approximation fails. A time step is appropriate if the maximum probability of two events occurring in \(\Delta t\) satisfies

$$\begin{aligned} \max \{P(C(\Delta t) \cup B(t')|A(0)) \} < \delta \end{aligned}$$
(42)

for a tolerance \(\delta \), where AB, and C denote motor bound states (including unbound, single head bound, and crosslinking) at time \(\Delta t> t' > 0\). \(P(C(\Delta t) \cup B(t')|A(0)) = P(C(\Delta t)|B(t'))P(B(t')|A(0))\) and each individual state change follows a single event Poisson process with \(P(B(t)|A(0)) = 1 - \exp [-k_{A \rightarrow B} t]\). The maximum probability for a double event occurs at \(t'=t'_\mathrm{{max}}\) found by solving

$$\begin{aligned} \begin{aligned}&\frac{dP(C(\Delta t) \cup B(t')|A(0))}{dt'}\Bigg |_{t'_{max}} = 0\\&k_{A \rightarrow B} \left( e^{k_{B \rightarrow C}(\Delta t-t')} - 1 \right) - k_{B \rightarrow C} \left( e^{k_{A \rightarrow B} t'} - 1 \right) = 0. \end{aligned} \end{aligned}$$
(43)

While no analytic solution exists, \(t'_\mathrm{{max}}\) can be numerically computed.

There are four unique processes that must be considered with a two-step binding process with unbound (U), single head bound (S), and crosslinking (C) states: \(U \rightarrow S \rightarrow U\), \(U \rightarrow S \rightarrow C\), \(S \rightarrow C \rightarrow S\), and \(C \rightarrow S \rightarrow U\). The process \(C \rightarrow S \rightarrow C\) has the same probability as \(S \rightarrow C \rightarrow S\), similarly, \(S \rightarrow U \rightarrow S\) has the same probability as \(U \rightarrow S \rightarrow U\). If modeling filament motion with some force- or energy-dependent unbinding, may be large. This means that in the limit of large unbinding rate the probabilities \(P(C \rightarrow S \rightarrow C) \rightarrow P(S \rightarrow C)\) and \(P(C \rightarrow S \rightarrow U) \rightarrow P(S \rightarrow U)\).

B Lookup table for kinetic Monte Carlo binding

Equation (11) gives the transition probability of a singly bound motor crosslinking as an integral of a Boltzmann factor. If \(h_\mathrm{cl}=0\), \(k_{\mathrm{on},C}\) is functionally similar to an error function. However, to model non-zero-length tethers, we numerically integrate Eq. (11). Rather than directly numerically integrating at each time step, we precompute a lookup table.

The cumulative distribution function (CDF) of Eq. (11), is a function \(h_{i,j}\). All other variables in the integral are constant for a given motor species. We reduce the CDF dimensionality by considering the lab position of each bound motor head and an infinite carrier line defined by the position and orientation of the unbound filament. Binding is then determined by the minimum distance \(r_{\perp }\) between the bound motor head position and the filament ends \([s_-, s_+]\) on the carrier line.

The carrier line CDF is

$$\begin{aligned} {{\,\mathrm{\mathrm{CDF}}\,}}(r_{\perp },s) = \int _{-\infty }^{s} e^{-\beta U(r_{\perp }, s')}ds', \end{aligned}$$
(44)

allowing us to write the crosslinking rate as

(45)

We notice that \(e^{-\beta U_{i,j}}\) is symmetric in s, so \({{\,\mathrm{\mathrm{CDF}}\,}}(r_{\perp },s) - {{\,\mathrm{\mathrm{CDF}}\,}}(r_{\perp },0)\) is anti-symmetric. Therefore, instead of integrating from negative infinity, we use

$$\begin{aligned} {{\,\mathrm{\mathrm{CDF}}\,}}'(r_{\perp },s) = {{\,\mathrm{sgn}\,}}(s)\int _{0}^{s} e^{-\beta U(r_{\perp }, s')}ds' \end{aligned}$$
(46)

and (45) to find the crosslinking rate.

We find the values of Eq. (46) by Gauss–Konrad integration. The accuracy desired sets the maximum values for s and \(r_{\perp }\). The integrand is always positive for real values of s and \(r_{\perp }\), so the CDF asymptotes for large values of either variable. The maximum of the integral is the point when the Boltzmann factor drops to the accuracy limit \(\delta \). Therefore, the lookup table domain is

$$\begin{aligned} s, r_{\perp } \in \left[ 0, \sqrt{-\frac{2\ln (\delta )}{\beta k_\mathrm{cl}}} + h_\mathrm{cl}\right] . \end{aligned}$$
(47)

Given a specified grid spacing \(\Delta s, \Delta r\), the memory required for the lookup table scales as \((s_{\max }/\Delta s) \times (r_{\perp , \max }/\Delta r)\).

Fig. 7
figure 7

Visual representation of the lookup table showing CDF values as a function of distance s along the filament for \(h_\mathrm{cl}= 32\) nm, \(k_\mathrm{cl}= .3\) pN/nm, \(\beta = 1./4.11\) (pN\(\cdot \)nm)\(^{-1}\), and \(\delta = 10^{-5}\)

1.1 B.1 Interpolation of lookup table values

Since the lookup table is not a continuous function, we interpolate values between discrete grid points. The 2D linear interpolation for input values of \(r_{\perp }\) and s is

$$\begin{aligned} \begin{aligned} {{\,\mathrm{\mathrm{CDF}}\,}}(r_{\perp }, s) \approx&\left( 1+m-\frac{r_{\perp }}{\Delta r} \right) \left( 1+n-\frac{s}{\Delta s} \right) \\&{{\,\mathrm{\mathrm{CDF}}\,}}_{m,n} + \left( \frac{r_{\perp }}{\Delta r} - m \right) \left( 1+n-\frac{s}{\Delta s} \right) {{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n}\\&+ \left( 1+m-\frac{r_{\perp }}{\Delta r} \right) \left( \frac{s}{\Delta s} - n \right) {{\,\mathrm{\mathrm{CDF}}\,}}_{m,n+1} \\&+ \left( \frac{r_{\perp }}{\Delta r} - m \right) \left( \frac{s}{\Delta s} - n \right) {{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n+1}, \end{aligned}\nonumber \\ \end{aligned}$$
(48)

where \({{\,\mathrm{\mathrm{CDF}}\,}}_{m,n}={{\,\mathrm{\mathrm{CDF}}\,}}(m\Delta r, n\Delta s)\) are the lookup table values at m and n if \(r_{\perp }\) lies within \(m\Delta r\) and \((m+1)\Delta r\) and s lies within \(n\Delta s\) and \((n+1)\Delta s\).

1.2 B.2 Reverse lookup algorithm

When a motor head binds to a filament, the binding position probability distribution function (PDF) is defined by the Boltzmann factor. We sample the PDF by using the lookup table. To transform a uniform random variable X to random variable Y with an arbitrary \({{\,\mathrm{\mathrm{PDF}}\,}}_Y\), X is inserted into the inverted CDF of Y

$$\begin{aligned} Y = {{\,\mathrm{\mathrm{CDF}}\,}}_Y^{-1}(X). \end{aligned}$$
(49)

Since the lookup table holds the CDF values and given a random number from a uniform distribution, we apply a combination of search and interpolation to quickly find the corresponding random number from the PDF. The algorithm is as follows

  1. 1.

    Sample a uniform random number \(X \in [0, {{\,\mathrm{\mathrm{CDF}}\,}}_{\max }]\). Note that the maximum value does not need to be 1.

  2. 2.

    Given \(r_{\perp }\), locate index m such that \(m\Delta r \le r_{\perp } \le (m +1) \Delta r\)

  3. 3.

    Use m to find the set of indices \(\{n_-, n_+\}\) such that \({{\,\mathrm{\mathrm{CDF}}\,}}_{m,n_-} \le X \le {{\,\mathrm{\mathrm{CDF}}\,}}_{m,n_-+1}\) and \({{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n_+} \le X \le {{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n_++1}\).

  4. 4.

    Use the CDF values to interpolate the binding locations \(s_-, s_+\) corresponding to the perpendicular distances \(r_- = m\Delta r\) and \(r_+ = (m+1) \Delta r\). For example,

    $$\begin{aligned} s_-&= \Delta s \frac{X - {{\,\mathrm{\mathrm{CDF}}\,}}_{m,n_-}}{{{\,\mathrm{\mathrm{CDF}}\,}}_{m,n_-+1} - {{\,\mathrm{\mathrm{CDF}}\,}}_{m,n_-}} + \Delta s n_- \end{aligned}$$
    (50)
    $$\begin{aligned} s_+&= \Delta s \frac{X - {{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n_+}}{{{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n_++1} - {{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n_+}} + \Delta s n_+ \end{aligned}$$
    (51)

    Note that \(s_-\) is not necessarily less than \(s_+\).

  5. 5.

    Find s by interpolating the across the lookup table grid with respect to \(r_{\perp }\)

    $$\begin{aligned} s \approx (s_+ - s_-)\frac{r_{\perp }-r_-}{\Delta r} + s_- \end{aligned}$$
    (52)

While this algorithm succeeds in most circumstance, the low slope of the CDF at large values of s can cause errors. For example, if the lookup table has the form of Fig. 7 and a protein is located at a perpendicular distance of \(r_{\perp }=35\) nm, given a random number of \(X=10^3\), no value for \(s_-\) will be found since \({{\,\mathrm{\mathrm{CDF}}\,}}(30, s_\mathrm{max}) < 10^3\). To correct for this, we solve for s using a binary search algorithm.

The binary search algorithm is as follows

  1. 1.

    Determine if \({{\,\mathrm{\mathrm{CDF}}\,}}_{m,n_\mathrm{max}}\) or \({{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n_\mathrm{max}}\) is less than X. If \({{\,\mathrm{\mathrm{CDF}}\,}}_{m,n_\mathrm{max}} < X\), set \(s_- =s_\mathrm{max}\). If \({{\,\mathrm{\mathrm{CDF}}\,}}_{m+1,n_\mathrm{max}} < X\), set \(s_+ =s_\mathrm{max}\).

  2. 2.

    Find other \(s_\pm \) using the inverted lookup table and Eq. (50) or (51).

  3. 3.

    Find the average of \(s_-\) and \(s_+\).

  4. 4.

    Use the lookup table interpolation algorithm to find the \({{\,\mathrm{\mathrm{CDF}}\,}}(r_{\perp }, s_\mathrm{avg})\).

  5. 5.

    If \({{\,\mathrm{\mathrm{CDF}}\,}}(r_{\perp }, s_\mathrm{avg}) > X\) set the larger of the two \(s_{\pm }\) values to \(s_\mathrm{avg}\). Otherwise, set the smaller of the two to \(s_\mathrm{avg}\).

  6. 6.

    Repeat steps 3-5 until \(|{{\,\mathrm{\mathrm{CDF}}\,}}(r_{\perp }, s_\mathrm{avg}) - X| < \delta \) for some desired tolerance \(\delta \).

This process converges at a rate \(O(\log _2(\delta s_{\max }))\).

C Numerical integration of the MFMD equation

We approximate the solution \(\psi _{i,j}(s_i, s_j, t)\) by discretizing the solution in time and space

(53)

for , where \(M_i\) is the number of discretized points along filament i. Additional boundary points for \(m,n=0\) are added.

We use forward Euler time-stepping so our discrete differential operator for time is

(54)

To solve the hyperbolic FPE (17), we use a first-order accurate upwind method [88]. The differential operator for \(s_i\) becomes

(55)

Note this only holds for the indices \(0<m\) and \(0<n\). The matrix representation for Eq. (55) is

(56)

where \(c_m\) and \(d_m\) are chosen to satisfy the boundary conditions. We choose the notation \(\rhd ^{m,n}\) for this matrix. To differentiate along \(s_j\), we use the identity , apply \(\rhd ^{m,n}\) on the matrix, and then convert back,

$$\begin{aligned} \frac{\partial \psi _{i,j}}{\partial s_i} \rightarrow \left( \sum _a \rhd ^{n,a} \psi _{j,i}^{a,m} \right) ^T, \end{aligned}$$
(57)

which in index notation is \(\psi _{i,j}^{m,a}(\rhd ^T)^{n,a}\). For brevity, we use the notation \(\psi _{i,j}^{m,a}(\rhd ^T)^{n,a}=\psi _{i,j}^{m,a}\lhd ^{a,n}\).

The discretized Fokker–Planck Eq. (17) is then

(58)

where \(U_{i,j}^{m,n,k}\) and \(v_{i,j}^{m,n,k}\) are the discretized potential and velocity at time \(k\Delta t\). Note that \(U_{i,j}^{m,n,k} = U_{j,i}^{n,m,k}\), but \(v_{i,j}^{m,n,k} \ne v_{j,i}^{n,m,k}\).

In cases where the flux of the motors \(\frac{\partial \left( v_{i,j}\psi _{i,j}\right) }{\partial s_i}\) is known at the boundaries, we construct \(\rhd \) to satisfy the requirements. When filaments are in solution, there is zero flux from the minus ends, so all \(c_m = 0\). In our simulations, motors walk of filament ends with out pausing, so \(d_{M_i-1}=-1\) and \(d_{M_i}=1\) with all other \(d_m=0\). Although not modeled in this paper, some biological motors end pause at filament plus ends. To model this, \(d_{M_i-1}=-1\) and every other \(d_m=0\).

D Conversion of binding parameters from an explicit to mean-field motor density model

To relate binding parameters of the one-step and multi-step binding models, we use that at steady state, the motor distribution \(\psi _{i,j}\) should be equivalent for both models. Since we only compare binding kinetics, we simplify the Fokker–Planck equation to keep only the binding terms: in Eq. (17), we set \(v_{i,j} = v_{j,i} = 0\),

(59)

The steady-state solution is

$$\begin{aligned} \psi _{i,j}= c e^{-\beta U_{i,j}}, \end{aligned}$$
(60)

which is a Boltzmann factor multiplied by an effective concentration.

The multi-step binding model can be written

(61)
(62)
(63)

where \(\chi _i\) is the mean-field density of motors with one head bound to filament i (cf. Eq. 16). We define \(K_\mathrm{E}' = K_\mathrm{E}/V_\mathrm{bind}\) and solve for the steady state, giving

(64)
(65)
(66)

The equations for \(\chi _i\) and \(\chi _j\) have the forms

$$\begin{aligned} X(s)&= C\int _{a}^{b} \left( Y(t)-X(s)\right) K(s,t)dt + D, \end{aligned}$$
(67)
$$\begin{aligned} Y(t)&= C\int _{c}^{d} \left( X(s)-Y(t)\right) K(s,t)ds + D, \end{aligned}$$
(68)

where \(t \in [a,b]\) and \(s \in [c,d]\). Distributing the integrals, we can rewrite

$$\begin{aligned} X(s)&= C\int _{a}^{b} Y(t)K(s,t)dt - CX(s)F(s) + D, \end{aligned}$$
(69)
$$\begin{aligned} Y(t)&= C\int _{c}^{d} X(s) K(s,t)ds - CY(t)G(t) + D, \end{aligned}$$
(70)

where \(F(s) = \int _{a}^{b}K(s,t)dt\) and \(G(t) = \int _{c}^{d}K(s,t)ds\). Solving for X(s) and Y(t) gives

$$\begin{aligned} X(s)&= \frac{D}{1+CF(s)} + \frac{C}{1+CF(s)}\int _{a}^{b} Y(t) K(s,t)dt, \end{aligned}$$
(71)
$$\begin{aligned} Y(t)&= \frac{D}{1+CG(t)} + \frac{C}{1+CG(t)}\int _{c}^{d} X(s) K(s,t)ds, \end{aligned}$$
(72)

After plugging Eq. (71) into (72), we find

$$\begin{aligned} Y(t)= & {} \frac{D}{1+CG(t)} + \frac{CD}{1+CG(t)}\int _{c}^{d}\frac{K(s,t)}{1+CF(s)}ds \nonumber \\&+ C^2\int _{a}^{b}\int _{c}^{d}Y(t') \frac{K(s,t)K(s,t')}{(1+CF(s))(1+CG(t'))} ds dt'.\nonumber \\ \end{aligned}$$
(73)

This can be rearranged into the form

$$\begin{aligned} Y(t) = A(t) + \int _{a}^{b}Y(t') B(t,t') dt', \end{aligned}$$
(74)

which implies that Y(t) and X(s) each satisfy a Fredholm equation of the second kind. Both A and B are continuous given \(K(s,t)= e^{-\beta U_{i,j}(s,t)}\), so the Fredholm equations of the second kind have unique solutions. By inspection, the solution to Eqs. (67) and (68) is \(X(s)=Y(t)=D\). When we substitute this solution in Eqs. (65) and (66), we find \(\chi _i = \chi _j = \epsilon K_a c_o\) and

$$\begin{aligned} \psi _{i,j}= \epsilon ^2 K_a K_\mathrm{E}'c_o e^{-\beta U_{i,j}}. \end{aligned}$$
(75)

Setting Eq. (60) equal to (75) gives

$$\begin{aligned} c = \frac{\epsilon ^2 K_a K_\mathrm{E}}{V_\mathrm{bind}}c_o. \end{aligned}$$
(76)

E Calculating binding parameters from experiments

The experimental parameters for motor binding are not always independently measured. If all but one binding parameters are known, then the unknown parameter can be found from Eq. (18) and the ratio of the number of motors with one head bound and number of motors crosslinking.

As an example, suppose we wish to find \(K_\mathrm{E}\). The number of motors with one head bound is \(N_S = c_o K_a \epsilon L\), where L is the filament length. In vitro experiments [86] can measure the crosslinking motors number \(N_d\). Integrating Eq. (75), we obtain the model prediction for the number of crosslinking motors as

$$\begin{aligned} N_C = c_o \epsilon ^2 K_a K_\mathrm{E}' \int _{L_{i}} \int _{L_{j}} e^{-\beta U_{i,j}} ds_i ds_j. \end{aligned}$$
(77)

For fully parallel or antiparallel filaments of the same length with adjacent centers, the total number of motors in Eq. (77) is proportional to L. If \(L \gg \sqrt{2/\beta k_\mathrm{cl}}\), the Gaussian integral \(\approx L\sqrt{2\pi /\beta k_\mathrm{cl}}e^{-\beta k_\mathrm{cl}r_{\perp }^2}\), where \(r_{\perp }\) is the center-to-center separation between filaments. The ratio of the number of crosslinking motors relative to the number motors with one head bound is

$$\begin{aligned} \rho = \frac{N_C}{N_S} = \epsilon K_\mathrm{E}' \sqrt{\frac{2\pi }{\beta k_\mathrm{cl}}}e^{-\beta k_\mathrm{cl}r_{\perp }^2}, \end{aligned}$$
(78)

allowing us to estimate \(K_\mathrm{E}' = \frac{\rho }{\epsilon }\sqrt{\frac{\beta k_\mathrm{cl}}{2\pi }}e^{\beta k_\mathrm{cl}r_{\perp }^2}\).

F Gaussian integrals in the moment expansion

The source terms in the moment expansion require a double integral over two filaments. To lower the numerical integration’s computational cost, we find an analytic solution for either the semi-integrated term \(Q_j^l(s_i)\) or the fully integrated term \(q_{i,j}^{k,l}\).

The integrated source terms are

(79)

where \(\alpha = \sqrt{\frac{2}{\beta k_\mathrm{cl}}}\). We define the quantity so that the integral over \(s_j\) becomes

$$\begin{aligned} {\bar{Q}}^l_j(s_i) = \int _{L_j} s_j^l e^{-{\left( \frac{s_j+A}{\alpha }\right) }^2}ds_j. \end{aligned}$$
(80)

This integral has an analytic form in terms of error functions, which can be rapidly computed. For \(l={0,1,2,3}\), we find

$$\begin{aligned} {\bar{Q}}_j^0(s_i)= & {} \frac{\alpha \sqrt{\pi }}{2} \left[ {{\,\mathrm{erf}\,}}\left( \frac{s_j+A}{\alpha } \right) \right] _{\partial L_j} \end{aligned}$$
(81)
$$\begin{aligned} {\bar{Q}}_j^1(s_i)= & {} -\frac{\alpha }{2}\left[ \alpha e^{-\left( \frac{s_j+A}{\alpha }\right) ^2} + A\sqrt{\pi }{{\,\mathrm{erf}\,}}\left( \frac{s_j+A}{\alpha } \right) \right] _{\partial L_j} \end{aligned}$$
(82)
$$\begin{aligned} {\bar{Q}}_j^2(s_i)= & {} \frac{\alpha }{4}\left[ 2\alpha (A-s_j) e^{-\left( \frac{s_j+A}{\alpha }\right) ^2} \right. \nonumber \\&\left. + (2A^2+\alpha ^2)\sqrt{\pi }{{\,\mathrm{erf}\,}}\left( \frac{s_j+A}{\alpha }\right) \right] _{\partial L_j} \end{aligned}$$
(83)
$$\begin{aligned} {\bar{Q}}_j^3(s_i)= & {} \frac{-\alpha }{4}\left[ 2\alpha (A^2 - A s_j + s_j^2 + \alpha ^2) e^{-\left( \frac{s_j+A}{\alpha }\right) ^2} \right. \nonumber \\&\left. + (2A^2 + 3\alpha ^2)A\sqrt{\pi }{{\,\mathrm{erf}\,}}\left( \frac{s_j+A}{\alpha }\right) \right] _{\partial L_j} \end{aligned}$$
(84)

G Moment expansion boundary terms

To generally define boundary conditions, instead of integrating over both \(s_i\) and \(s_i\), we integrate over just one variable. This makes the boundary condition a function of a single filament attachment position. For example, the boundary terms for the first filament are

(85)

These boundary terms are evaluated at \(-L_i/2\) and \(L_i/2\). We derive a recursion relation by integrating Eq. (85) over \(s_j\) and using the definition in Eq. (37)

(86)

Solving this equation requires finding the time evolution of the boundary term spatial derivatives, which solve

(87)

This shows that the boundary terms do not close. However, if the higher-order terms or their coefficients are small compared to the moments \(\mu _{i,j}^{k,l}\), we may take a zeroth-order approximation. We consider this approximation in Sect. 6.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lamson, A.R., Moore, J.M., Fang, F. et al. Comparison of explicit and mean-field models of cytoskeletal filaments with crosslinking motors. Eur. Phys. J. E 44, 45 (2021). https://doi.org/10.1140/epje/s10189-021-00042-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epje/s10189-021-00042-9

Navigation