Skip to main content

Macroscopic Model of Substrate-Based Cell Motility

  • Chapter
  • First Online:
Physical Models of Cell Motility

Abstract

Eukaryotic cells moving in response to chemical or mechanical stimuli play a fundamental role in tissue growth, wound healing and the immune response. In addition, cell migration is essential for understanding several life-threatening pathologies. At the developmental stage, dysfunction of motility can result in certain disabilities, while in the developed organism it may result in cardiovascular diseases. Motility dysfunction is also involved in cancer growth, especially during metastasis. Beyond such obvious biological and medical relevance, cell motility is also a fascinating example of a self-organized and self-propelled system within the realm of physics.

After a brief overview of the processes involved in cellular motility, the experimental facts, and previous modeling approaches, we survey recent advances in model development that enabled capturing numerous features observed for moving cells. The model is based on the phase field method, which is especially suited to treat the moving and deformable boundaries involved in both individual and collective cell motility. We begin with a didactic introduction to a simple model for the movement of an individual cell and then describe how incrementally complexity can be added: the main components involved in cell motility—the deformability of the cell, the acto-myosin dynamics, the adhesion both to the substrate and to other cells, as well as the substrate deformations—can all be accounted for. Finally we discuss how the model can be generalized to describe the interactions and the collective movement of many, self-organized cells.

Throughout we discuss how the model, at the different stages, can describe a large part of the known features of cell motility: the subcritical onset of motion, the shape diversity of moving cells, and the occurrence of steady motion, stick-slip motion with concomitant shape oscillations, as well as of more complex dynamical modes like bipedal motion. We also investigate how cells navigate on substrates with patterned adhesion properties and modulated stiffness, situations currently under technological development to collect or sort cells. For multiple cells, the model is able to predict that collective cell migration emerges spontaneously as a result of inelastic collision-type interactions of cells. Finally, we discuss possible future extensions of the modeling framework.

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

Access this chapter

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

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Here, the interface position x 0 corresponds to \(\rho = \frac{1} {2}\). Alternatively, the interface is also often defined as the maximum of | ∇ρ | .

  2. 2.

    Defining the surface tension as the excess energy due to the interface, one obtains \(\varSigma \propto \sqrt{D_{\rho }}\).

  3. 3.

    Actually, already in Eq. (1.1), a time scale for the change in ρ should be included, which was set to one.

  4. 4.

    Since the cell is self-propelled, the net force exerted by the cell on the substrate, however, is zero.

  5. 5.

    Fibronectin is the protein of the ECM responsible for the binding to the integrin adhesion receptors. RGD is a peptide sequence of fibronectin, implicated in the attachment.

  6. 6.

    The Laplace–Beltrami operator should be used instead of the simple Laplace operator, because the cell membrane is a curved surface.

References

  1. M. Abercrombie, The crawling movement of metazoan cells. Proc. R. Soc. Lond. B 207, 129–147 (1980)

    Article  ADS  Google Scholar 

  2. S. Aland, S. Egerer, J. Lowengrub, A. Voigt, Diffuse interface models of locally inextensible vesicles in a viscous fluid. J. Comp. Phys. 277, 32 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. A.Y. Alexandrova, K. Arnold, S. Schaub, J.M. Vasiliev, J.J. Meister, A.D. Bershadsky, A.B. Verkhovsky, Comparative dynamics of retrograde actin flow and focal adhesions: formation of nascent adhesions triggers transition from fast to slow flow. PLoS ONE 3, e3234 (2008)

    Article  ADS  Google Scholar 

  4. D. Ambrosi, Cellular traction as an inverse problem. SIAM J. Appl. Math. 66, 2049 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  5. R. Ananthakrishnan, A. Ehrlicher, The forces behind cell movement. Int. J. Biol. Sci. 3, 303–17 (2007)

    Article  Google Scholar 

  6. K.I. Anderson, R. Cross, Contact dynamics during keratocyte motility. Curr. Biol. 10, 253 (2000)

    Article  Google Scholar 

  7. T.E. Angelini, E. Hannezo, X. Trepat, J.J. Fredberg, D.A. Weitz, Cell migration driven by cooperative substrate deformation patterns. Phys. Rev. Lett. 104(16), 168104 (2010)

    Google Scholar 

  8. T.E. Angelini, E. Hannezo, X. Trepat, M. Marquez, J.J. Fredberg, D.A. Weitz, Glass-like dynamics of collective cell migration. Proc. Natl. Acad. Sci. U.S.A. 108(12), 4714–4719 (2011)

    Article  ADS  Google Scholar 

  9. I.S. Aranson, L.S. Tsimring, Continuum description of avalanches in granular media. Phys. Rev. E 64, 020301 (2001)

    Article  ADS  Google Scholar 

  10. I.S. Aranson, L.S. Tsimring, Patterns and collective behavior in granular media: theoretical concepts. Rev. Mod. Phys. 78, 641 (2006)

    Article  ADS  Google Scholar 

  11. I.S. Aranson, L.S. Tsimring, Granular Patterns (Oxford University Press, Oxford, 2009)

    MATH  Google Scholar 

  12. I.S. Aranson, V.A. Kalatsky, V.M. Vinokur, Continuum field description of crack propagation. Phys. Rev. Lett. 85, 118 (2000)

    Article  ADS  Google Scholar 

  13. Y. Aratyn-Schaus, M.L. Gardel, Transient frictional slip between integrin and the ECM in focal adhesions under myosin-II tension. Curr. Biol. 20, 1145 (2010)

    Article  Google Scholar 

  14. L. Bai, D. Breen, Calculating center of mass in an unbounded 2D environment. J. Graph. GPU Game Tools 13(4), 53–60 (2008)

    Article  Google Scholar 

  15. E. Barnhart, G. Allen, F. Jülicher, J. Theriot, Bipedal locomotion in crawling cells. Biophys. J. 98(6), 933–942 (2010)

    Article  ADS  Google Scholar 

  16. E.L. Barnhart, K.-C. Lee, K. Keren, A. Mogilner, J.A. Theriot, An adhesion-dependent switch between mechanisms that determine motile cell shape. PLoS Biol. 9, e1001059 (2011)

    Article  Google Scholar 

  17. G.I. Bell, Models for the specific adhesion of cells to cells. Science 200, 618 (1978)

    Article  ADS  Google Scholar 

  18. T. Biben, C. Misbah, Tumbling of vesicles under shear flow within an advected-field approach. Phys. Rev. E 67, 031908 (2003)

    Article  ADS  Google Scholar 

  19. T. Biben, K. Kassner, C. Misbah, Phase-field approach to three-dimensional vesicle dynamics. Phys. Rev. E 72(4), 041921 (2005)

    Google Scholar 

  20. I. Bischofs, S. Schmidt, U. Schwarz, Effect of adhesion geometry and rigidity on cellular force distributions. Phys. Rev. Lett. 103, 048101 (2009)

    Article  ADS  Google Scholar 

  21. C. Blanch-Mercader, J. Casademunt, Spontaneous motility of actin lamellar fragments. Phys. Rev. Lett. 110(7), 078102 (2013)

    Google Scholar 

  22. D.C. Bottino, L.J. Fauci, A computational model of ameboid deformation and locomotion. Eur. Biophys. J. 27, 532 (1998)

    Article  Google Scholar 

  23. J.P. Boyd, Chebyshev and Fourier spectral methods, revised 2nd edn. (Dover Publications, Mineola, 2001)

    MATH  Google Scholar 

  24. H.R. Brand, H. Pleiner, F. Ziebert, Macroscopic dynamics of polar nematic liquid crystals. Phys. Rev. E 74, 021713 (2006)

    Article  ADS  Google Scholar 

  25. D. Bray, Cell Movements: From Molecules to Motility (Garland Publishing, New York, 2001)

    Google Scholar 

  26. F. Brochard-Wyart, P.G. de Gennes, Adhesion induced by mobile binders: dynamics. Proc. Natl. Acad. Sci. U.S.A. 99, 7854–7859 (2002)

    Article  ADS  Google Scholar 

  27. K. Burridge, M. Chrzanowska-Wodnicka, Focal adhesions, contractility, and signaling. Ann. Rev. Cell Dev. Biol. 12, 463 (1996)

    Article  Google Scholar 

  28. M.F. Carlier, D. Pantaloni, Control of actin assembly dynamics in cell motility. J. Biol. Chem. 282, 23005–23009 (2007)

    Article  Google Scholar 

  29. A.E. Carlsson, Mechanisms of cell propulsion by active stresses. New J. Phys. 13, 073009 (2011)

    Article  ADS  Google Scholar 

  30. C.E. Chan, D.J. Odde, Traction dynamics of Filopodia on compliant substrates. Science 322, 1687 (2008)

    Article  ADS  Google Scholar 

  31. S.-Y. Chou, C.-M. Cheng, P.R. LeDuc, Composite polymer systems with control of local substrate elasticity and their effect on cytoskeletal and morphological characteristics of adherent cells. Biomaterials 30, 3136 (2009)

    Article  Google Scholar 

  32. K.U. Claussen, R. Giesa, T. Scheibel, H.-W. Schmidt, Learning from nature: synthesis and characterization of longitudinal polymer gradient materials inspired by mussel byssus threads. Macromol. Rapid Commun. 33, 206 (2012)

    Article  Google Scholar 

  33. S. Cox, P. Matthews, Exponential time differencing for stiff systems. J. Comp. Phys. 176(2), 430–455 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. M. Cross, H. Greenside, Pattern Formation and Dynamics in Nonequilibrium Systems (Cambridge University Press, Cambridge, 2009)

    Book  MATH  Google Scholar 

  35. G. Csucs, K. Quirin, G. Danuser, Locomotion of fish epidermal keratocytes on spatially selective adhesion patterns. Cell Motil. Cytoskeleton 64(11), 856–867 (2007)

    Article  Google Scholar 

  36. G. Danuser, J. Allard, A. Mogilner, Mathematical modeling of eukaryotic cell migration: insights beyond experiments. Annu. Rev. Cell Dev. Biol. 29, 501–528 (2013)

    Article  Google Scholar 

  37. T. Das, T.K. Maiti, S. Chakraborty, Traction force microscopy on-chip: shear deformation of fibroblast cells. Lab Chip 8, 1308 (2008)

    Article  Google Scholar 

  38. M. Deforet, V. Hakim, H. Yevick, G. Duclos, P. Silberzan, Emergence of collective modes and tri-dimensional structures from epithelial confinement. Nat. Commun. 5, 3747 (2014)

    Article  ADS  Google Scholar 

  39. M. Dembo, Y.L. Wang, Stresses at the cell-to-substrate interface during locomotion of fibroblasts. Biophys. J. 76, 2307 (1999)

    Article  ADS  Google Scholar 

  40. E.W. Dent, F.B. Gertler, Cytoskeletal dynamics and transport in growth cone motility and axon guidance. Neuron 40, 209–227 (2003)

    Article  Google Scholar 

  41. P.A. DiMilla, K. Barbee, D.A. Lauffenburger, Mathematical model for the effects of adhesion and mechanics on cell migration speed. Biophys. J. 60, 15–37 (1991)

    Article  Google Scholar 

  42. K. Doubrovinski, K. Kruse, Cytoskeletal waves in the absence of molecular motors. Europhys. Lett. 83(1), 18003 (2008)

    Google Scholar 

  43. K. Doxzen, S.R.K. Vedula, M.C. Leong, H. Hirata, N.S. Gov, A.J. Kabla, B. Ladoux, C.T. Lim, Guidance of collective cell migration by substrate geometry. Integr. Biol. 5(8), 1026–1035 (2013)

    Article  Google Scholar 

  44. A. Dreher, I.S. Aranson, K. Kruse, Spiral actin-polymerization waves can generate amoeboidal cell crawling. New J. Phys. 16(5), 055007 (2014)

    Google Scholar 

  45. Q. Du, X.Q. Wang, Simulating the deformation of vesicle membranes under elastic bending energy in three dimensions. J. Comp. Phys. 212, 757 (2006)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. H. Emmerich, Advances of and by phase-field modelling in condensed-matter physics. Adv. Phys. 57, 1 (2008)

    Article  ADS  Google Scholar 

  47. M. Enculescu, M. Falcke, Modeling morphodynamic phenotypes and dynamic regimes of cell motion. Adv. Exp. Med. Biol. 736, 337–358 (2012)

    Article  Google Scholar 

  48. T. Erdmann, U.S. Schwarz, Stochastic dynamics of adhesion clusters under shared constant force and with rebinding. J. Chem. Phys. 121, 8997 (2004)

    Article  ADS  Google Scholar 

  49. J. Etienne, J. Fouchard, D. Mitrossilis, N. Bufi, P. Durand-Smet, A. Asnacios, Cells as liquid motors. Mechanosensitivity emerges from collective dynamics of actomyosin cortex. Proc. Natl. Acad. Sci. U.S.A. 106, 2740–2745 (2015)

    Article  Google Scholar 

  50. S. Etienne-Manneville, Microtubules in cell migration. Annu. Rev. Cell Dev. Biol. 29, 471–499 (2013)

    Article  Google Scholar 

  51. U. Euteneuer, M. Schliwa, Persistent, directional motility of cells and cytoplasmic fragments in the absence of microtubules. Nature 310, 58 (1984)

    Article  ADS  Google Scholar 

  52. G. Fix, Phase field methods for free boundary problems, in Free Boundary Problems: Theory and Applications, ed. by A. Fasano, M. Primicerio (Pitman, Boston, 1983), p. 580

    Google Scholar 

  53. R. Folch, J. Casademunt, A. Hernandez-Machado, L. Ramirez-Piscina, Phase-field model for Hele-Shaw flows with arbitrary viscosity contrast. I. Theoretical approach. Phys. Rev. E 60, 1724 (1999)

    Article  Google Scholar 

  54. R. Folch, J. Casademunt, A. Hernandez-Machado, L. Ramirez-Piscina, Phase-field model for Hele-Shaw flows with arbitrary viscosity contrast. II. Numerical study. Phys. Rev. E 60, 1734 (1999)

    Article  ADS  Google Scholar 

  55. M.F. Fournier, R. Sauser, D. Ambrosi, J.-J. Meister, A.B. Verkhovsky, Force transmission in migrating cells. J. Cell Biol. 188, 287 (2010)

    Article  Google Scholar 

  56. C. Franck, S.A. Maskarinec, D.A. Tirrell, G. Ravichandran, Three-dimensional traction force microscopy: a new tool for quantifying cell-matrix interactions. PLOS One 6, e17833 (2011)

    Article  ADS  Google Scholar 

  57. T. Fuhs, M. Goegler, C.A. Brunner, C.W. Wolgemuth, J.A. Kaes, Causes of retrograde flow in fish keratocytes. Cytoskeleton 71, 24–35 (2014)

    Article  Google Scholar 

  58. M.L. Gardel, B. Sabass, L. Ji, G. Danuser, U.S. Schwarz, C.M. Waterman, Traction stress in focal adhesions correlates biphasically with actin retrograde flow speed. J. Cell Biol. 183, 999 (2008)

    Article  Google Scholar 

  59. G. Giannone, B.J. Dubin-Thaler, H.-G. Döbereiner, N. Kieffer, A.R. Bresnick, M.P. Sheetz, Periodic lamellopodial contractions correlate with rearward actin waves. Cell 116, 431 (2004)

    Article  Google Scholar 

  60. L. Giomi, A. DeSimone, Spontaneous division and motility in active nematic droplets. Phys. Rev. Lett. 112, 147802 (2014)

    Article  ADS  Google Scholar 

  61. A. Gopinathan, K.-C. Lee, J.M. Schwarz, A.J. Liu, Branching, capping, and severing in dynamic actin structures. Phys. Rev. Lett. 99, 058103 (2007)

    Article  ADS  Google Scholar 

  62. M.E. Gracheva, H.G. Othmer, A continuum model of motility in ameboid cells. Bull. Math. Biol. 66, 167–193 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  63. G. Grégoire, H. Chaté, Y. Tu, Moving and staying together without a leader. Physica D 181(34), 157–170 (2003)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  64. H. Grimm, A. Verkhovsky, A. Mogilner, J. Meister, Analysis of actin dynamics at the leading edge of crawling cells: implications for the shape of keratocyte lamellipodia. Eur. Biophys. J. 32, 563 (2003)

    Article  Google Scholar 

  65. D. Grossman, I. Aranson, E.B. Jacob, Emergence of agent swarm migration and vortex formation through inelastic collisions. New J. Phys. 10(2), 023036 (2008)

    Google Scholar 

  66. M. Herant, M. Dembo, Form and function in cell motility: from fibroblasts to keratocytes. Biophys. J. 98, 1408 (2010)

    Article  ADS  Google Scholar 

  67. W.R. Holmes, L. Edelstein-Keshet, A comparison of computational models for eukaryotic cell shape and motility. PLoS Comput. Biol. 8, e1002793 (2012)

    Article  ADS  Google Scholar 

  68. J. Howard, Mechanics of Motor Proteins and the Cytoskeleton (Sinauer, Sunderland, 2001)

    Google Scholar 

  69. S. Huang, C.P. Brangwynne, K.K. Parker, D.E. Ingber, Symmetry-breaking in mammalian cell cohort migration during tissue pattern formation: role of random-walk persistence. Cell Motil. Cytoskeleton 61, 201–213 (2005)

    Article  Google Scholar 

  70. R.O. Hynes, Integrins: versatility, modulation, and signaling in cell adhesion. Cell 69(1), 11–25 (1992)

    Article  Google Scholar 

  71. P.A. Iglesias, P.N. Devreotes, Biased excitable networks: how cells direct motion in response to gradients. Curr. Opin. Cell Biol. 24, 245–253 (2012)

    Article  Google Scholar 

  72. P.Y. Jay, P.A. Pham, S.A. Wong, E.L. Elson, A mechanical function of myosin II in cell motility. J. Cell Sci. 108, 387–393 (1995)

    Google Scholar 

  73. L. Ji, J. Lim, G. Danuser, Fluctuations of intracellular forces during cell protrusion. Nat. Cell Biol. 10, 1393–1400 (2008)

    Article  Google Scholar 

  74. J.F. Joanny, K. Kruse, F. Jülicher, J. Prost, Hydrodynamic theory for multi-component active polar gels. New J. Phys. 9, 422 (2007)

    Article  ADS  Google Scholar 

  75. K. John, P. Peyla, K. Kassner, J. Prost, C. Misbah, Nonlinear study of symmetry breaking in actin gels: implications for cellular motility. Phys. Rev. Lett. 100, 068101 (2008)

    Article  ADS  Google Scholar 

  76. F. Jülicher, K. Kruse, J. Prost, J.F. Joanny, Active behavior of the cytoskeleton. Phys. Rep. 449, 3 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  77. A. Karma, W.-J. Rappel, Quantitative phase-field modeling of dendritic growth in two and three dimensions. Phys. Rev. E 57, 4323 (1998)

    Article  ADS  MATH  Google Scholar 

  78. K. Kassner, C. Misbah, A phase-field approach for stress-induced instabilities. Europhys. Lett. 46, 217 (1999)

    Article  ADS  Google Scholar 

  79. K. Keren, Z. Pincus, G.M. Allen, E.L. Barnhart, G. Marriott, A. Mogilner, J.A. Theriot, Mechanism of shape determination in motile cells. Nature 453(7194), 475–480 (2008)

    Article  ADS  Google Scholar 

  80. H. Kitahata, K. Iida, M. Nagayama, Spontaneous motion of an elliptic camphor particle. Phys. Rev. E 87(1), 010901 (2013)

    Google Scholar 

  81. M.H. Köpf, L.M. Pismen, A continuum model of epithelial spreading. Soft Matter 9(14), 3727–3734 (2013)

    Article  ADS  Google Scholar 

  82. M.M. Kozlov, A. Mogilner, Model of polarization and bistability of cell fragments. Biophys. J. 93, 3811 (2007)

    Article  ADS  Google Scholar 

  83. K. Kruse, F. Jülicher, Self-organization and mechanical properties of active filament bundles. Phys. Rev. E 67, 051913 (2003)

    Article  ADS  Google Scholar 

  84. K. Kruse, J.F. Joanny, F. Jülicher, J. Prost, K. Sekimoto, Asters, vortices, and rotating spirals in active gels of polar filaments. Phys. Rev. Lett. 92, 078101 (2004)

    Article  ADS  Google Scholar 

  85. 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, 5 (2005)

    Article  Google Scholar 

  86. K. Kruse, J.F. Joanny, F. Jülicher, J. Prost, Contractility and retrograde flow in lamellipodium motion. Phys. Biol. 3, 130 (2006)

    Article  ADS  Google Scholar 

  87. E. Kuusela, W. Alt, Continuum model of cell adhesion and migration. J. Math. Biol. 58, 135 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  88. C.I. Lacayo, Z. Pincus, M.M. Van Duijn, C.A. Wilson, D.A. Fletcher, F.B. Gertler, A. Mogilner, J.A. Theriot, Emergence of large-scale cell morphology and movement from local actin filament growth dynamics. PLoS Biol. 5, 2035 (2007)

    Article  Google Scholar 

  89. T. Lai, K.-H. Chiam, Mechanochemical model of cell migration on substrates of varying stiffness. Phys. Rev. E 84, 061907 (2011)

    Article  ADS  Google Scholar 

  90. L. Landau, E. Lifshitz, Theory of Elasticity. Course of Theoretical Physics, vol. 7 (Elsevier, Oxford, 1986)

    Google Scholar 

  91. J. Langer, Models of pattern formation in first-order phase transitions, in Directions in Condensed Matter Physics, ed. by G. Grinstein, G. Mazenko (World Scientific, Singapore, 1986), p. 165

    Google Scholar 

  92. J. Lee, K. Jacobson, The composition and dynamics of cell-substratum adhesions in locomoting fish keratocytes. J. Cell Sci. 110, 2833 (1997)

    Google Scholar 

  93. P. Lee, C.W. Wolgemuth, Crawling cells can close wounds without purse strings or signaling. PLoS Comput. Biol. 7(3), e1002007 (2011)

    Google Scholar 

  94. J. Lee, A. Ishihara, J.A. Theriot, K. Jacobson, Principles of locomotion for simple-shaped cells. Nature 362, 167 (1993)

    Article  ADS  Google Scholar 

  95. M. Lenz, T. Thoresen, M.L. Gardel, A.R. Dinner, Contractile units in disordered actomyosin bundles arise from F-actin buckling. Phys. Rev. Lett. 108, 238107 (2012)

    Article  ADS  Google Scholar 

  96. Y. Li, P. Bhimalapuram, A.R. Dinner, Model for how retrograde actin flow regulates adhesion traction stresses. J. Phys. Condens. Matter 22, 194113 (2010)

    Article  ADS  Google Scholar 

  97. A.D. Lieber, S. Yehudai-Resheff, E.L. Barnhart, J.A. Theriot, K. Keren, Membrane tension in rapidly moving cells is determined by cytoskeletal forces. Curr. Biol. 23, 1409–1417 (2013)

    Article  Google Scholar 

  98. O. Lieleg, M.M.A.E. Claessens, A.R. Bausch, Structure and dynamics of cross-linked actin networks. Soft Matter 6, 218 (2010)

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

  100. C.M. Lo, H.B. Wang, M. Dembo, Y.L. Wang, Cell movement is guided by the rigidity of the substrate. Biophys. J. 79, 144 (2000)

    Article  Google Scholar 

  101. J. Löber, F. Ziebert, I.S. Aranson, Modeling crawling cell movement on soft engineered substrates. Soft Matter 10(9), 1365–1373 (2014)

    Article  ADS  Google Scholar 

  102. J. Löber, F. Ziebert, I.S. Aranson, Collisions of deformable cells lead to collective migration. Sci. Rep. 5, 9172 (2015)

    Article  Google Scholar 

  103. A.J. Loosley, J.X. Tang, Stick-slip motion and elastic coupling in crawling cells. Phys. Rev. E 86(3), 031908 (2012)

    Google Scholar 

  104. M. Machacek, G. Danuser, Morphodynamic profiling of protrusion phenotypes. Biophys. J. 90, 1439 (2006)

    Article  ADS  Google Scholar 

  105. E. Maitrea, C. Misbah, P. Peyla, A. Raoult, Comparison between advected-field and level-set methods in the study of vesicle dynamics. Physica D 241, 11461157 (2012)

    MATH  Google Scholar 

  106. W. Marth, A. Voigt, Signaling networks and cell motility: a computational approach using a phase field description. J. Math. Biol. 69, 91–112 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  107. E. Meyhöfer, J. Howard, The force generated by a single kinesin molecule against an elastic load. Proc. Natl. Acad. Sci. U.S.A. 92, 574 (1995)

    Article  ADS  Google Scholar 

  108. A. Mogilner, Mathematics of cell motility: have we got its number? J. Math. Biol. 58, 105134 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  109. A. Mogilner, K. Keren, The shape of motile cells. Curr. Biol. 19, R762 (2009)

    Article  Google Scholar 

  110. A. Mogilner, G. Oster, Cell motility driven by actin polymerization. Biophys. J. 71, 3030–3045 (1996)

    Article  ADS  Google Scholar 

  111. A. Mogilner, G. Oster, Force generation by actin polymerization II: the elastic ratchet and tethered filaments. Biophys. J. 84, 1591–1605 (2003)

    Article  ADS  Google Scholar 

  112. A. Mogilner, J. Allard, R. Wollman, Cell polarity: quantitative modeling as a tool in cell biology. Science 336, 175 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  113. S. Munevar, Y.L. Wang, M. Dembo, Traction force microscopy of migrating normal and H-ras transformed 3T3 fibroblasts. Biophys. J. 80, 1744 (2001)

    Article  Google Scholar 

  114. H. Nakazawa, K. Sekimoto, Polarity sorting in a bundle of actin filaments by two-headed myosins. J. Phys. Soc. Jpn. 65, 2404 (1996)

    Article  ADS  Google Scholar 

  115. M.P. Neilson, J.A. Mackenzie, S.D. Webb, R.H. Insall, Modeling cell movement and chemotaxis using pseudopod-based feedback. SIAM J. Sci. Comput. 33, 1035 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  116. S. Nourshargh, P. Hordijk, M. Sixt, Breaching multiple barriers: leukocyte motility through venular walls and the interstitium. Nat. Rev. Mol. Cell Biol. 11, 366–378 (2010)

    Article  Google Scholar 

  117. T. Ohta, T. Ohkuma, Deformable self-propelled particles. Phys. Rev. Lett. 102(15), 154101 (2009)

    Google Scholar 

  118. S. Osher, R. Fedkiw, Level Set Methods and Dynamic Implicit Surfaces (Springer, New York, 2003)

    Book  MATH  Google Scholar 

  119. S.P. Palecek, J.C. Loftus, M.H. Ginsberg, D.A. Lauffenburger, A.F. Horwitz, Integrin-ligand binding properties govern cell migration speed through cell-substratum adhesiveness. Nature 385, 537 (1997)

    Article  ADS  Google Scholar 

  120. S.H. Parekh, O. Chaudhuri, J.A. Theriot, D.A. Fletcher, Loading history determines the velocity of actin-network growth. Nat. Cell Biol. 7, 1219 (2005)

    Article  Google Scholar 

  121. R.J. Pelham, Y.-L. Wang, Cell locomotion and focal adhesions are regulated by substrate flexibility. Proc. Natl. Acad. Sci. U.S.A. 94, 13661–13665 (1997)

    Article  ADS  Google Scholar 

  122. B.N.J. Persson, Sliding Friction (Springer, Berlin, 2000)

    Book  MATH  Google Scholar 

  123. F. Peruani, A. Deutsch, M. Bär, Nonequilibrium clustering of self-propelled rods. Phys. Rev. E 74(3), 030904 (2006)

    Google Scholar 

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

    Article  ADS  Google Scholar 

  125. R. Peter, V. Schaller, F. Ziebert, W. Zimmermann, Pattern formation in active cytoskeletal networks. New J. Phys. 10, 035002 (2008)

    Article  ADS  Google Scholar 

  126. L. Petitjean, M. Reffay, E. Grasland-Mongrain, M. Poujade, B. Ladoux, A. Buguin, P. Silberzan, Velocity fields in a collectively migrating epithelium. Biophys. J. 98, 1790–1800 (2010)

    Article  ADS  Google Scholar 

  127. H. Pleiner, H.R. Brand, Spontaneous splay phases in polar nematic liquid crystals. Europhys. Lett. 9, 243 (1989)

    Article  ADS  Google Scholar 

  128. T.D. Pollard, J.A. Cooper, Actin, a central player in cell shape and movement. Science 326, 1208 (2009)

    Article  ADS  Google Scholar 

  129. M. Poujade, E. Grasland-Mongrain, A. Hertzog, J. Jouanneau, P. Chavrier, B. Ladoux, A. Buguin, P. Silberzan, Collective migration of an epithelial monolayer in response to a model wound. Proc. Natl. Acad. Sci. U.S.A. 104, 15988–15993 (2007)

    Article  ADS  Google Scholar 

  130. M. Prass, K. Jacobson, A. Mogilner, M. Radmacher, Direct measurement of the lamellipodial protrusive force in a migrating cell. J. Cell Biol. 174, 767 (2006)

    Article  Google Scholar 

  131. W.H. Press, S.A. Teukolsky, W.T. Vetterling, B.P. Flannery, Numerical Recipes 3rd Edition: The Art of Scientific Computing (Cambridge University Press, New York, 2007)

    MATH  Google Scholar 

  132. D. Raucher, M.P. Sheetz, Cell spreading and lamellipodial extension rate is regulated by membrane tension. J. Cell Biol. 148, 127–136 (2000)

    Article  Google Scholar 

  133. P. Recho, L. Truskinovsky, Asymmetry between pushing and pulling for crawling cells. Phys. Rev. E 87(2), 022720 (2013)

    Google Scholar 

  134. P. Recho, T. Putelat, L. Truskinovsky, Contraction-driven cell motility. Phys. Rev. Lett. 111, 108102 (2013)

    Article  ADS  Google Scholar 

  135. J. Renkawitz, K. Schumann, M. Weber, T. Lämmermann, H. Pflicke, M. Piel, J. Polleux, J. Spatz, M. Sixt, Adaptive force transmission in amoeboid cell migration. Nat. Cell Biol. 11, 1438 (2009)

    Article  Google Scholar 

  136. C.G. Rolli, H. Nakayama, K. Yamaguchi, J.P. Spatz, R. Kemkemer, J. Nakanishi, Switchable adhesive substrates: revealing geometry dependence in collective cell behavior. Biomaterials 33, 2409–2418 (2012)

    Article  Google Scholar 

  137. B. Sabass, U.S. Schwarz, Modelling cytoskeletal flow over adhesion sites: competition between stochastic bond dynamics and intracellular relaxation. J. Phys. Condens. Matt. 22, 194112 (2010)

    Article  ADS  Google Scholar 

  138. A. Sacan, H. Ferhatosmanoglu, H. Coskun, CellTrack: an open-source software for cell tracking and motility analysis. Bioinformatics 24, 1647 (2008)

    Article  Google Scholar 

  139. E. Sackmann, R. Bruinsma, Cell adhesion as wetting transition? Chem. Phys. Chem. 3, 262 (2002)

    Article  Google Scholar 

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

    Article  ADS  Google Scholar 

  141. C.H. Schreiber, M. Stewart, T. Duke, Simulation of cell motility that reproduces the force-velocity relationship. Proc. Natl. Acad. Sci. U.S.A. 107, 9141 (2010)

    Article  ADS  Google Scholar 

  142. U.S. Schwarz, M.L. Gardel, United we stand - integrating the actin cytoskeleton and cell-matrix adhesions in cellular mechanotransduction. J. Cell Sci. 125, 3051 (2012)

    Article  Google Scholar 

  143. U.S. Schwarz, S.A. Safran, Physics of adherent cells. Rev. Mod. Phys. 85, 1327 (2013)

    Article  ADS  Google Scholar 

  144. U.S. Schwarz, N.Q. Balaban, D. Riveline, A. Bershadsky, B. Geiger, S.A. Safran, Calculation of forces at focal adhesions from elastic substrate data: the effect of localized force and the need for regularization. Biophys. J. 83, 1380 (2002)

    Article  ADS  Google Scholar 

  145. D. Shao, W.J. Rappel, H. Levine, Computational model for cell morphodynamics. Phys. Rev. Lett. 105, 108104 (2010)

    Article  ADS  Google Scholar 

  146. D. Shao, H. Levine, W.-J. Rappel, Coupling actin flow, adhesion, and morphology in a computational cell motility model. Proc. Natl. Acad. Sci. U.S.A. 109, 6851 (2012)

    Article  ADS  Google Scholar 

  147. M.P. Sheetz, D. Felsenfeld, C.G. Galbraith, D. Choquet, Cell migration as a five-step cycle. Biochem. Soc. Symp. 65, 233–243 (1999)

    Google Scholar 

  148. C. Shi, C.-H. Huang, P. Devreotes, P. Iglesias, Interaction of motility, directional sensing, and polarity modules recreates the behaviors of chemotaxing cells. PLoS Cell Biol. 9, e1003122 (2013)

    MathSciNet  Google Scholar 

  149. I. Singer-Loginova, H.M. Singer, The phase field technique for modeling multiphase materials. Rep. Prog. Phys. 71, 106501 (2008)

    Article  ADS  Google Scholar 

  150. A. Smith, U. Seifert, Effective adhesion strength of specifically bound vesicles. Phys. Rev. E 71, 061902 (2005)

    Article  ADS  Google Scholar 

  151. L.A. Smith, H. Aranda-Espinoza, J.B. Haun, M. Dembo, D.A. Hammer, Neutrophil traction stresses are concentrated in the uropod during migration. Biophys. J. 92, L58 (2007)

    Article  Google Scholar 

  152. J. Solon, I. Levental, K. Sengupta, P.C. Georges, P.A. Janmey, Fibroblast adaptation and stiffness matching to soft elastic substrates. Biophys. J. 93, 4453 (2007)

    Article  ADS  Google Scholar 

  153. A. Stéphanou, E. Mylona, M. Chaplain, P. Tracqui, A computational model of cell migration coupling the growth of focal adhesions with oscillatory cell protrusions. J. Theor. Biol. 253, 701–716 (2008)

    Article  MathSciNet  Google Scholar 

  154. M. Sussman, P. Smereka, S. Osher, A level set approach for computing solutions to incompressible two-phase flow. J. Comp. Phys. 114, 146 (1994)

    Article  ADS  MATH  Google Scholar 

  155. T.M. Svitkina, A.B. Verkhovsky, K.M. McQuade, G.G. Borisy, Analysis of the actin-myosin II system in fish epidermal keratocytes: mechanism of cell body translocation. J. Cell Biol. 139, 397 (1997)

    Article  Google Scholar 

  156. B. Szabó, G.J. Szöllösi, B. Gönci, Z. Jurányi, D. Selmeczi, T. Vicsek, Phase transition in the collective migration of tissue cells: experiment and model. Phys. Rev. E 74(6), 061908 (2006)

    Google Scholar 

  157. M. Takeichi, Cadherins: a molecular family important in selective cell-cell adhesion. Annu. Rev. Biochem. 59, 237 (1990)

    Article  Google Scholar 

  158. J.L. Tan, J. Tien, D.M. Pirone, D.S. Gray, K. Bhadriraju, C.S. Chen, Cells lying on a bed of microneedles: an approach to isolate mechanical force. Proc. Natl. Acad. Sci. U.S.A. 100, 1484 (2003)

    Article  ADS  Google Scholar 

  159. M. Tarama, T. Ohta, Oscillatory motions of an active deformable particle. Phys. Rev. E 87(6), 062912 (2013)

    Google Scholar 

  160. M. Thery, V. Racine, M. Piel, A. Pepin, A. Dimitrov, Y. Chen, J.B. Sibarita, M. Bornens, Anisotropy of cell adhesive microenvironment governs cell internal organization and orientation of polarity. Proc. Natl. Acad. Sci. U.S.A. 103, 19771 (2006)

    Article  ADS  Google Scholar 

  161. E. Tjhung, D. Marenduzzo, M.E. Cates, Spontaneous symmetry breaking in active droplets provides a generic route to motility. Proc. Natl. Acad. Sci. U.S.A. 109(31), 12381–12386 (2012)

    Article  ADS  Google Scholar 

  162. E. Tjhung, A. Tiribocchi, D. Marenduzzo, M.E. Cates, A minimal physical model captures the shapes of crawling cells. Nat. Commun. 6, 5420 (2015)

    Article  ADS  Google Scholar 

  163. L. Trichet, J. Le Digabel, R. Hawkins, S.R. Vedula, M. Gupta, C. Ribrault, P. Hersen, R. Voituriez, B. Ladoux, Evidence of a large-scale mechanosensing mechanism for cellular adaptation to substrate stiffness. Proc. Natl. Acad. Sci. U.S.A. 109, 6933 (2012)

    Article  ADS  Google Scholar 

  164. T.A. Ulrich, E.M. de Juan Pardo, S. Kumar, The mechanical rigidity of the extracellular matrix regulates the structure, motility, and proliferation of glioma cells. Cancer Res. 69, 4167 (2009)

    Article  Google Scholar 

  165. P. Vallotton, G. Danuser, S. Bohnet, J.-J. Meister, A.B. Verkhovsky, Tracking retrograde flow in keratocytes: news from the front. Mol. Biol. Cell 16, 1223 (2005)

    Article  Google Scholar 

  166. S.R.K. Vedula, M.C. Leong, T.L. Lai, P. Hersen, A.J. Kabla, C.T. Lim, B. Ladoux, Emerging modes of collective cell migration induced by geometrical constraints. Proc. Natl. Acad. Sci. U.S.A. 109(32), 12974–12979 (2012)

    Article  ADS  Google Scholar 

  167. A.B. Verkhovsky, T.M. Svitkina, G.G. Borisy, Self-polarization and directional motility of cytoplasm. Curr. Biol. 9, 11–20 (1999)

    Article  Google Scholar 

  168. D. Volfson, L.S. Tsimring, I.S. Aranson, Stick-slip dynamics of a granular layer under shear. Phys. Rev. E 69(3), 031302 (2004)

    Google Scholar 

  169. J.H.-C. Wang, J.-S. Lin, Cell traction force and measurement methods. Biomech. Model. Mechanobiol. 6, 361 (2007)

    Article  Google Scholar 

  170. H.-B. Wang, M. Dembo, S.K. Hanks, Y.-L. Wang, Focal adhesion kinase is involved in mechanosensing during fibroblast migration. Proc. Natl. Acad. Sci. U.S.A. 98, 11295–11300 (2001)

    Article  ADS  Google Scholar 

  171. M.F. Ware, A. Wells, D.A. Lauffenburger, Epidermal growth factor alters fibroblast migration speed and directional persistence reciprocally and in a matrix-dependent manner. J. Cell Sci. 111, 2423–2432 (1998)

    Google Scholar 

  172. J. Weichsel, U.S. Schwarz, Two competing orientation patterns explain experimentally observed anomalies in growing actin networks. Proc. Natl. Acad. Sci. U.S.A. 107, 6304 (2010)

    Article  ADS  Google Scholar 

  173. O.D. Weiner, W.A. Marganski, L.F. Wu, S.J. Altschuler, M.W. Kirschner, An actin-based wave generator organizes cell motility. PLoS Biol. 5, e221 (2007)

    Article  Google Scholar 

  174. R. Wittkowski, H. Löwen, H.R. Brand, Microscopic and macroscopic theories for the dynamics of polar liquid crystals. Phys. Rev. E 84, 041708 (2011)

    Article  ADS  Google Scholar 

  175. C.W. Wolgemuth, Lamellipodial contractions during crawling and spreading. Biophys. J. 89, 1643 (2005)

    Article  ADS  Google Scholar 

  176. C.W. Wolgemuth, M. Zajac, The moving boundary node method: a level set-based, finite volume algorithm with applications to cell motility. J. Comp. Phys. 229, 7287 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  177. C.W. Wolgemuth, J. Stajic, A. Mogilner, Redundant mechanisms for stable cell locomotion revealed by minimal models. Biophys. J. 101, 545 (2011)

    Article  ADS  Google Scholar 

  178. J. Xu, D. Wirtz, T. Pollard, Dynamic cross-linking by α-actinin determines the mechanical properties of actin filament networks. J. Biol. Chem. 273, 9570 (1998)

    Article  Google Scholar 

  179. H. Xu, R. Matkar, T. Kyu, Phase-field modeling on morphological landscape of isotactic polystyrene single crystals. Phys. Rev. E 72, 011804 (2005)

    Article  ADS  Google Scholar 

  180. P.T. Yam, C.A. Wilson, L. Ji, B. Hebert, E.L. Barnhart, N.A. Dye, P.W. Wiseman, G. Danuser, J.A. Theriot, Actin-myosin network reorganisation breaks symmetry at the cell rear to spontaneously initiate polarized cell motility. J. Cell Biol. 178, 1207 (2007)

    Article  Google Scholar 

  181. L. Yang, J.C. Effler, B.L. Kutscher, S.E. Sullivan, D.N. Robinson, P.A. Iglesias, Modeling cellular deformations using the level set formalism. BMC Sys. Biol. 2, 68 (2008)

    Article  Google Scholar 

  182. F. Ziebert, I.S. Aranson, Effects of adhesion dynamics and substrate compliance on the shape and motility of crawling cells. PLoS ONE 8(5), e64511 (2013)

    Google Scholar 

  183. F. Ziebert, I.S. Aranson, Modular approach for modeling cell motility. Eur. Phys. J. Spec. Top. 223, 1265–1277 (2014)

    Article  Google Scholar 

  184. F. Ziebert, I.S. Aranson, L.S. Tsimring, Effects of cross-links on motor-mediated filament organization. New J. Phys. 9, 421 (2007)

    Article  ADS  Google Scholar 

  185. F. Ziebert, S. Swaminathan, I.S. Aranson, Model for self-polarization and motility of keratocyte fragments. J. R. Soc. Interface 9, 1084 (2012)

    Article  Google Scholar 

  186. J. Zimmermann, C. Brunner, M. Enculescu, M. Goegler, A. Ehrlicher, J. Käs, M. Falcke, Actin filament elasticity and retrograde flow shape the force-velocity relation of motile cells. Biophys. J. 102, 287 (2012)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We thank Sumanth Swaminathan (now with W.L. Gore & Associates) for his valuable contribution to the earlier stage of this project. In the course of the model development reviewed here, we enjoyed stimulating discussions with J. Oliver, P. Sens, E. Raphaël, J. Prost, J.-F. Joanny, A. Verkhovsky, H. Chaté and B. Winkler. F.Z. acknowledges funding from the German Science Foundation (DFG) via project ZI 1232/2-1. J.L. acknowledges funding from the German Science Foundation (DFG) within the GRK 1558. I.S.A. was supported by the US Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), Materials Science and Engineering Division. Part of the numerical work was performed on Northern Illinois University’s GPU cluster GAEA.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Falko Ziebert .

Editor information

Editors and Affiliations

1.1 Electronic Supplementary material

Appendix

Appendix

1.1.1 Numerical Methods

To solve the phase field model numerically, especially in the multiple cell case, we developed a highly parallel algorithm implemented on GPUs using CUDA. To avoid (slow) copying between GPU and CPU memory, the algorithm exclusively runs on the GPU (except for the output of data). The code can handle an arbitrary number of cells in single or double precision. Since the phase field keeps track of the cells’ boundaries, the problem can be solved on a square domain with periodic boundary conditions. In this case, a pseudo-spectral approach based on the fast Fourier transform (FFT) is the most efficient [131]. Pseudo-Spectral Code To simplify the notations, we illustrate the pseudo-spectral algorithm for an example reaction-diffusion system in one spatial dimension. An extension of the algorithm to higher spatial dimensions is straightforward. We assume a finite domain 0 ≤ x < L and a system of two coupled equations

$$ \displaystyle\begin{array}{rcl} \partial _{t}a\left (t,x\right )& =& D_{a}\partial _{x}^{2}a\left (t,x\right ) + f\left (a\left (t,x\right ),b\left (t,x\right )\right ),{}\end{array}$$
(1.46)
$$\displaystyle\begin{array}{rcl} \partial _{t}b\left (t,x\right )& =& D_{b}\partial _{x}^{2}b\left (t,x\right ) + g\left (a\left (t,x\right ),b\left (t,x\right )\right ),{}\end{array}$$
(1.47)

with periodic boundary conditions

$$\displaystyle\begin{array}{rcl} a\left (t,0\right )& =& a\left (t,L\right ),\quad b\left (t,0\right ) = b\left (t,L\right ).{}\end{array}$$
(1.48)

The time domain is discretized with time step Δ t and the spatial domain is discretized with step size \(\varDelta x = L/N\), where N is the number of grid points. The discretized fields a j, k and b j, k are then defined as

$$\displaystyle\begin{array}{rcl} a_{j,k}& =& a\left (t_{0} + j\varDelta t,k\varDelta x\right ),\quad b_{j,k} = b\left (t_{0} + j\varDelta t,k\varDelta x\right ),{}\end{array}$$
(1.49)

where j enumerates time steps (with t 0 the initial time) and k enumerates the spatial grid points.

In Fourier space (k-space), the wave vectors are then given by k n  = n Δ k with \(\varDelta k = \frac{2\pi } {L}\), and the complex exponential in the Fourier transform becomes

$$\displaystyle\begin{array}{rcl} \exp \left (i\,k_{n}\,k\varDelta x\right ) =\exp \left (2\pi i\frac{nk} {N} \right ).& &{}\end{array}$$
(1.50)

The FFT of a j, k with respect to the spatial grid points k is defined as

$$\displaystyle\begin{array}{rcl} \hat{a}_{j,n} = \mathcal{F}\left [a_{j,k}\right ] =\sum _{ k=0}^{N-1}\exp \left (2\pi i\frac{nk} {N} \right )a_{j,k},& &{}\end{array}$$
(1.51)

and the inverse FFT is given by

$$\displaystyle\begin{array}{rcl} a_{j,k} = \mathcal{F}^{-1}\left [\hat{a}_{ j,n}\right ] = \frac{1} {N}\sum _{n=0}^{N-1}\exp \left (-2\pi i\frac{nk} {N} \right )\hat{a}_{j,n}.& &{}\end{array}$$
(1.52)

In spectral methods, it is especially easy to solve the linear part of the equations. In the given example, the discrete equivalent of the second order space derivative \(\partial _{x}^{2}a\left (t,x\right )\) is easily computed: first, a j, k is transformed to k-space by the forward FFT, then multiplied by − k n 2, and finally transformed back to coordinate space [23]

$$\displaystyle\begin{array}{rcl} \partial _{x}^{2}a\left (t,x\right ) \Leftrightarrow \mathcal{F}^{-1}\left [-k_{ n}^{2}\hat{a}_{ j,n}\right ] = -\frac{1} {N}\sum _{n=0}^{N-1}\frac{4\pi ^{2}n^{2}} {L^{2}} \exp \left (-2\pi i\frac{nk} {N} \right )\hat{a}_{j,n}.& &{}\end{array}$$
(1.53)

For the time stepping of the reaction-diffusion system Eqs. (1.46) and (1.47), we employ the operator-split method [34]:

$$\displaystyle\begin{array}{rcl} a_{j+1,k}& =& \mathcal{F}^{-1}\left \{\exp \left (-\varDelta tD_{ a}k_{n}^{2}\right )\left [\mathcal{F}\left (a_{ j,n} +\varDelta tf\left (a_{j,n},b_{j,n}\right )\right )\right ]\right \},{}\end{array}$$
(1.54)
$$\displaystyle\begin{array}{rcl} b_{j+1,k}& =& \mathcal{F}^{-1}\left \{\exp \left (-\varDelta tD_{ b}k_{n}^{2}\right )\left [\mathcal{F}\left (b_{ j,n} +\varDelta tg\left (a_{j,k},b_{j,n}\right )\right )\right ]\right \},{}\end{array}$$
(1.55)

where we used the reversed version of \(\exp \left (\varDelta t\alpha \right ) = 1 +\varDelta t\alpha + \mathcal{O}\left (\varDelta t^{2}\right )\) to improve numerical stability [33]. Technically, one calculates an Euler step in real space, transforms to k-space, multiplies by the time evolution operator, and transforms back. The method just described is called pseudo-spectral, since the nonlinear part of the equations (contained in f, g) is evaluated in coordinate space, i.e. not in Fourier space where it would give rise to computationally expensive convolutions.

Combining Scalar to Complex Fields In case of dealing with many equations, the performance of the algorithm can be further improved by exploiting the fact that the FFT of two real fields can be computed more efficiently as a single FFT of a complex field [131]. We hence combine the real fields a j, n and b j, n in a single complex field z via

$$\displaystyle\begin{array}{rcl} z_{j,n}& =& a_{j,n} + ib_{j,n}.{}\end{array}$$
(1.56)

Note that the properties of the FFT imply \(\hat{z}_{j,n+N} =\hat{ z}_{j,n}\), and especially

$$\displaystyle\begin{array}{rcl} \hat{z}_{j,n}^{{\ast}} =\sum _{ k=0}^{N-1}\exp \left (-2\pi i\frac{nk} {N} \right )\left (a_{j,k} - ib_{j,k}\right ),& &{}\end{array}$$
(1.57)

where z denotes the complex conjugate of z. In turn, Eq. (1.57) implies

$$\displaystyle\begin{array}{rcl} \hat{z}_{j,N-n}^{{\ast}} =\sum _{ k=0}^{N-1}\exp \left (2\pi i\frac{nk} {N} \right )\left (a_{j,k} - ib_{j,k}\right ).& &{}\end{array}$$
(1.58)

Consequently, using Eqs. (1.57) and (1.58), we can obtain the FFT of a j, n as

$$\displaystyle\begin{array}{rcl} \frac{1} {2}\left (\hat{z}_{j,n} +\hat{ z}_{j,N-n}^{{\ast}}\right ) =\sum _{ k=0}^{N-1}\exp \left (2\pi i\frac{nk} {N} \right )a_{j,k} =\hat{ a}_{j,k},& &{}\end{array}$$
(1.59)

and analogously, the FFT of b j, n as

$$\displaystyle\begin{array}{rcl} \frac{1} {2}\left (\hat{z}_{j,n} -\hat{ z}_{j,N-n}^{{\ast}}\right ) = i\sum _{ k=0}^{N-1}\exp \left (2\pi i\frac{nk} {N} \right )b_{j,k} = i\hat{b}_{j,k}.& &{}\end{array}$$
(1.60)

We combine the equations for a j+1, k and b j+1, k , Eqs. (1.54) and (1.55), respectively, into a single equation for \(z_{j+1,k} = a_{j+1,k} + ib_{j+1,k}\). Using the properties (1.59) and (1.60), this equation can be entirely expressed in terms of z:

$$\displaystyle\begin{array}{rcl} z_{j+1,k}& =& \mathcal{F}^{-1}\left [c_{ n}\hat{w}_{j,n} + d_{n}\hat{w}_{j,N-n}^{{\ast}}\right ],{}\end{array}$$
(1.61)

where we introduced the abbreviations

$$\displaystyle\begin{array}{rcl} \hat{w}_{j,k}& =& \mathcal{F}\left [z_{j,n} +\varDelta tf\left (\mathrm{Re}\left (z_{j,n}\right ),\mathrm{Im}\left (z_{j,n}\right )\right ) + i\varDelta tg\left (\mathrm{Re}\left (z_{j,n}\right ),\mathrm{Im}\left (z_{j,n}\right )\right )\right ],{}\end{array}$$
(1.62)

and

$$\displaystyle\begin{array}{rcl} c_{n}& =& \frac{1} {2}\left (\exp \left (-\varDelta tD_{a}k_{n}^{2}\right ) +\exp \left (-\varDelta tD_{ b}k_{n}^{2}\right )\right ),{}\end{array}$$
(1.63)
$$\displaystyle\begin{array}{rcl} d_{n}& =& \frac{1} {2}\left (\exp \left (-\varDelta tD_{a}k_{n}^{2}\right ) -\exp \left (-\varDelta tD_{ b}k_{n}^{2}\right )\right ).{}\end{array}$$
(1.64)

Compared to Eqs. (1.54)–(1.55), which require four real FFTs (two forwards and two backwards), having expressed the algorithm in terms of a single complex field, Eq. (1.61), requires only two complex FFTs (one forwards and one backwards).

Additional Details on the Implementation of the Algorithm In case of the single cell model, the phase field ρ and the concentration of the adhesive bonds A, as well as the two components of the actin polarization, p x and p y , are combined in two complex fields, respectively. For simulations of multiple cells, additional phase fields are combined as complex fields in pairs. For maximum performance, all arrays except the substrate displacement field u were merged in a single array. Using the CUDA library FFT (CUFFT), this allows to compute all FFTs with a single customized batch FFT. The substrate displacement field u is treated by a separate CUDA kernel, which calculates the spatial derivatives of the traction field in finite difference approximation.

As already discussed in Sect. 1.4.1.1, sums over phase fields, as e.g. \(\sum _{j=1}^{N}\rho _{j}^{2}\) and \(\sum _{j=1}^{N}\nabla \rho _{j}\), have to be calculated for the cells’ interactions. These, as well as the cells’ volumes and centers of mass are computed with customized CUDA kernels. The center of mass of each cell is determined as described in [14] to track the position of each cell over time, from which its velocity is readily determined in finite difference approximation. Mean velocities as well as order parameters over time are obtained by summing over all cells. The time-averaged mean velocity was determined by a subsequent averaging over all time steps for sufficiently long time intervals to diminish the contribution of initial transients. We typically use a resolution of 512 × 512 up to 2048 × 2048 Fourier modes for a square periodic domain size of \(L = \text{100\textendash 200}\). For comparison, the initial cell radii typically used were \(r_{0} = \text{10\textendash 15}\).

1.1.2 Derivation of the Equation for the Elastic Displacements

The stress tensor for an isotropic homogeneous incompressible visco-elastic solid (often called Kelvin–Voigt material) is given by [90]

$$\displaystyle\begin{array}{rcl} \sigma _{ik} =\tilde{ G}\left (u_{i,k} + u_{k,i}\right ) +\tilde{\eta } \left (\dot{u}_{i,k} +\dot{ u}_{k,i}\right ) - p\delta _{ik}\,,& &{}\end{array}$$
(1.65)

where \(u_{i} = u_{i}\left (x,y,z;t\right ),\,i \in \left \{x,y,z\right \}\) are the components of the displacement field. \(p = p\left (x,y,z;t\right )\) is the pressure field ensuring the incompressibility. \(\tilde{G}\), \(\tilde{\eta }\) are the shear modulus and viscosity, respectively. Assuming overdamped motion (\(\ddot{u}_{i} = 0\)) the force balance \(\nabla \cdot \sigma\), i.e. \(\sigma _{ik,k} = 0\), yields

$$\displaystyle\begin{array}{rcl} \tilde{G}\nabla ^{2}\mathbf{u} +\tilde{\eta } \nabla ^{2}\dot{\mathbf{u}}& =& \nabla p\,, \\ \nabla \cdot \mathbf{u}& =& 0\,.{}\end{array}$$
(1.66)

We assume the deformable elastic body to be periodic in the x- and y-direction (with period L). At the lower boundary (z = 0), we assume vanishing displacements, corresponding to the elastic body sticking on a non-deformable surface,

$$\displaystyle\begin{array}{rcl} \mathbf{u}\left (x,y,z = 0,t\right ) = 0\,.& &{}\end{array}$$
(1.67)

At the upper boundary (z = H), the cell exerts a two-dimensional traction force T = (T x , T y , 0), but zero normal force on the elastic body

$$\displaystyle\begin{array}{rcl} \sigma _{xz}\left (x,y,z = H,t\right )& =& T_{x}\left (x,y,t\right )\,, \\ \sigma _{yz}\left (x,y,z = H,t\right )& =& T_{y}\left (x,y,t\right )\,, \\ \sigma _{zz}\left (x,y,z = H,t\right )& =& 0\,. {}\end{array}$$
(1.68)

Nonlinearities arising from the free boundary at z = H are neglected. Eqs. (1.66) are equivalent to a biharmonic equation for \(\mathbf{w} =\tilde{ G}\mathbf{u} +\tilde{\eta }\dot{ \mathbf{u}}\) and Laplace’s equation for p. After Fourier transforming in x- and y-direction and introducing the wavenumber \(k^{2} = k_{x}^{2} + k_{y}^{2}\), these equations become

$$\displaystyle\begin{array}{rcl} \partial _{z}^{4}\mathbf{w} - 2k^{2}\partial _{ z}^{2}\mathbf{w} + k^{4}\mathbf{w}& =& 0\,,{}\end{array}$$
(1.69)
$$\displaystyle\begin{array}{rcl} \partial _{z}^{2}p - k^{2}p& =& 0\,.{}\end{array}$$
(1.70)

Six out of the necessary 14 boundary conditions for Eq. (1.69) are given as before by \(\mathbf{w}\left (x,y,z = 0,t\right ) = 0\) and Eq. (1.68). The remaining eight boundary conditions are obtained by evaluating Eq. (1.66) at the boundaries.

The assumption of a vertical substrate layer height H much smaller than its horizontal extension L, H ≪ L, allows a long wavelength expansion (k x , k y  ≪ 1∕H) of the solution to Eq. (1.69). We will keep terms up to second order in k x , k y , corresponding to retaining spatial derivatives up to second order of the traction force T. Finally, integrating the result over z from z = 0 to z = H leads to Eq. (1.18) given in Sect. 1.2.4.

In case of an inhomogeneous substrate stiffness, \(\tilde{G} =\tilde{ G}\left (x,y,z\right )\), one can easily see that the force balance, \(\sigma _{ik,k} = 0\), creates a plethora of additional terms in Eq. (1.66), involving all kinds of first order derivatives of \(\tilde{G}\). However, we can neglect these terms if the long wavelength expansion is truncated at the lowest order and assume no dependence on the vertical direction, \(\tilde{G} =\tilde{ G}\left (x,y\right )\). Hence for simplicity, in all computations involving a space-dependent substrate stiffness G we used Eq. (1.18) in the limit h → 0 [note that the parameters in that equation are related to the ones introduced here via \(h = \frac{H^{2}} {12}\), \(G = \frac{2\tilde{G}} {\xi H}\) and \(\eta = \frac{2\tilde{\eta }} {\xi H}\) with \(\xi\) the efficiency of traction force transmission, see Eq. (1.19)] and substituted \(G \rightarrow G\left (x,y\right )\).

1.1.3 Characterizing the Cell’s Velocity and Shape

To extract the velocity, the center-of-mass position x c is determined according to

$$\displaystyle{ x_{i}^{c} =\int x_{ i}\,\rho (x,y)dxdy,\,\,\,i = 1,2\,. }$$
(1.71)

The aspect ratio was determined via the corresponding 2×2 variance matrix I ij

$$\displaystyle{ I_{ij} =\int (x_{i} - x_{i}^{c})(x_{ j} - x_{j}^{c})\rho (x,y)dxdy\,. }$$
(1.72)

Its eigenvalues \(\lambda _{1,2}\) were calculated and their ratio

$$\displaystyle{ h = \sqrt{\lambda _{1 } /\lambda _{2}} }$$
(1.73)

is a measure for the aspect ratio of the cell. For a cell moving in x-direction (i = 1), the off-diagonal elements vanish, \(I_{12} = I_{21} = 0\), and the aspect ratio is simply given by \(h = \sqrt{I_{22 } /I_{11}}\). Since a circular shape has aspect ratio h = 1, in Fig. 1.7 we trace h − 1.

The asymmetry of moving cells, namely the deviation from reflection symmetry, can be described via the skewness tensor G ijk

$$\displaystyle{ G_{ijk} =\int (x_{i} - x_{i}^{c})(x_{ j} - x_{j}^{c})(x_{ k} - x_{k}^{c})\rho (x,y)dxdy\,. }$$
(1.74)

Obviously, for an ellipse G ijk  = 0. For a cell with an asymmetric shape moving in x-direction (i, j, k = 1), only 4 elements of the skewness tensor are non-zero: G 111 ≠ 0, and \(G_{122} = G_{212} = G_{221}\neq 0\). Hence one can define the following relative asymmetry measures

$$\displaystyle\begin{array}{rcl} \zeta & =& \frac{\vert G_{111}\vert ^{1/3}} {\sqrt{I_{11 } + I_{22}}}\,,{}\end{array}$$
(1.75)
$$\displaystyle\begin{array}{rcl} \eta & =& \frac{\vert G_{122}\vert ^{1/3}} {\sqrt{I_{11 } + I_{22}}}\,.{}\end{array}$$
(1.76)

For an ellipse, ζ, η = 0; for the asymmetric moving cells obtained by our simulations one has ζ, η ≠ 0, see Fig. 1.7.

1.1.4 Supplementary Movies

  1. 1.

    Bistability I (ch1_video1.mpg) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video1.mpg. The movie shows a round spreading cell that is perturbed by adding a small polarization pointing to the right. The perturbation is insufficient to set the cell into motion. Cf. also Fig. 1.4.

  2. 2.

    Bistability II (ch1_video2.mpg) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video2.mpg. The movie shows a round spreading cell that is perturbed by adding a small polarization pointing to the right. This perturbation leads to the onset of motion. Cf. also Fig. 1.4.

  3. 3.

    Bipedal motion (ch1_video3.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video3.avi. Bipedal motion is an overall straight motion concomitant with out-of-phase oscillations of the lower and upper halves of the cell. Shown is the shape of the cell and the substrate displacement field (Fig. 1.12a–d).

  4. 4.

    A cell bounces off a step in adhesion (ch1_video4.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video4.avi. The motion of a cell on a substrate where the adhesive strength is modulated by a step in the rate of adhesion formation a 0, corresponding to a varying density of the adhesive ligands (a 0 = 0. 2 in the blue region, a 0 = 0. 01 in the black one; G = 0. 15 everywhere). The cell prefers to stay on the highly adhesive region and is reflected from the low-adhesion region (Fig. 1.13a).

  5. 5.

    Motion of cells on striped adhesive patterns I (ch1_video5.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video5.avi. The alternating stripes have a value of high (a 0 = 0. 15, blue) and no adhesiveness (a 0 = 0, black); G = 0. 2. The cell positions itself symmetrically and moves parallel to the stripes in a steady fashion (Fig. 1.13b top).

  6. 6.

    Motion of cells on striped adhesive patterns II (ch1_video6.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video6.avi. Overall low adhesiveness (a 0 = 0. 0015, blue) and no adhesiveness (a 0 = 0, black); G = 0. 1. After moving initially along the stripes, the cell turns and moves perpendicular to the stripes in a stick-slip fashion (Fig. 1.13b bottom).

  7. 7.

    Durotaxis I (ch1_video7.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video7.avi. A cell moving in a linear gradient in the substrate’s modulus G

    [varying along the y-direction from G = 0 (black) at the bottom to G = 0. 4 (blue) at the top]. Independently of the initial conditions, the cell approaches a trajectory with an optimal value of G (Fig. 1.14e).

  8. 8.

    Durotaxis II (ch1_video8.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video8.avi. A cell moving in a linear gradient in the substrate’s modulus G [varying along the y-direction from G = 0 (black) at the bottom to G = 0. 4 (blue) at the top]. Independently of the initial conditions, the cell approaches a trajectory with an optimal value of G (Fig. 1.14e).

  9. 9.

    Inelastic collision of cells (ch1_video9.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video9.avi. A strongly inelastic collision of two canoe-shaped cells, leading to an effective alignment of the directions of motion (Fig. 1.15a).

  10. 10.

    Elastic collision of cells (ch1_video10.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video10.avi. An almost elastic collision of two fan-shaped cells (Fig. 1.15b).

  11. 11.

    Transition from stationary to moving cells (ch1_video11.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video11.avi. Initially, only few cells move, while cells which adhere strongly to the substrate (those with green spots inside) are stationary. Repeated collisions between moving and stationary cells set all cells into motion (Fig. 1.16a–d).

  12. 12.

    Transition from moving to stationary cells (ch1_video12.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video12.avi. Initially moving cells come to rest and collect in stationary clusters (Fig. 1.16e–h).

  13. 13.

    Translational collective migration (ch1_video13.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video13.avi. Alignment of propagation directions due to collisions between cells in a domain with periodic boundary conditions. Cells do not adhere to each other (Fig. 1.17a–d).

  14. 14.

    Rotational collective migration (ch1_video14.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video14.avi. Emergence of rotational collective motion in a circular confined domain (in the red region, the rate of nonlinear adhesive bond formation to the substrate is reduced by a factor of 9) (Fig. 1.17e–h).

  15. 15.

    Suppression of rotational collective migration by cell–cell adhesion (ch1_video15.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video15.avi. Adhesion between cells prevents the emergence of collective rotational motion (Fig. 1.17j–l).

  16. 16.

    Clustering of cells (ch1_video16.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video16.avi. Clustering of cells due to strong cell–cell adhesion forces. A few cells are motile and the cluster changes its shape in time. Cells leave and join the cluster (Fig. 1.18a).

  17. 17.

    Traveling band of cells (ch1_video17.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video17.avi. A band of cells strongly adhering to each other is moving in a domain with periodic boundary conditions (Fig. 1.18b).

  18. 18.

    Cells competing for voids (ch1_video18.avi) http://www.physik.tu-berlin.de/~jakob/movies_small//ch1_video18.avi. Confined high-density state without cell–cell adhesion. Cells compete for voids, thereby moving slowly through a “crowded environment” in a random walk fashion (Fig. 1.18d).

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Ziebert, F., Löber, J., Aranson, I.S. (2016). Macroscopic Model of Substrate-Based Cell Motility. In: Aranson, I. (eds) Physical Models of Cell Motility. Biological and Medical Physics, Biomedical Engineering. Springer, Cham. https://doi.org/10.1007/978-3-319-24448-8_1

Download citation

Publish with us

Policies and ethics