Skip to main content

Cell-Based, Continuum and Hybrid Models of Tissue Dynamics

  • Chapter
  • First Online:
Mathematical Models and Methods for Living Systems

Part of the book series: Lecture Notes in Mathematics ((LNMCIME,volume 2167))

Abstract

Movement of amoeboid cells is involved in embryonic development, wound repair, the immune response to bacterial invasion, and tumor formation and metastasis. Individual cells detect extracellular chemical and mechanical signals via membrane receptors, and this initiates signal transduction cascades that produce intracellular signals. These signals control the motile machinery of the cell and thereby determine the spatial localization of contact sites with the substrate and the sites of force-generation needed to produce directed motion. The coordination and control of this complex process of direction sensing, amplification of spatial differences in the signal, assembly of the motile machinery, and control of the attachment to the substratum involves numerous molecules whose spatial distribution serves to distinguish the front from the rear of the cell, and whose temporal expression is tightly controlled. How chemical and mechanical signals are integrated, how spatial differences in signals are produced, and how propulsive and adhesive forces are controlled are issues that are amenable to mathematical modeling. An overview of some approaches to these complex problems is the subject of this chapter.

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 64.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 84.99
Price excludes VAT (USA)
  • Compact, lightweight 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.

    Supported in part by NSF Grants DMS-0817529 & -1311974 and by NIH Grant GM 29123.

  2. 2.

    Phalloidin functions by binding and stabilizing F-actin and thus this may not represent the true in vitro distribution, but later we obtain very similar distributions.

  3. 3.

    A brief introduction to this technique is given in Appendix.

  4. 4.

    This may not be true in reality, and we could allow small changes without changing the results. Osmotic fluxes may well change the total volume during deformations, but this has not been characterized experimentally, and incorporating it would add a great deal of complexity for an as yet unquantifiable increase in accuracy.

References

  1. B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts, P. Walter, Molecular Biology of the Cell, 4th edn. (Garland Science, New York, London, 2002)

    Google Scholar 

  2. U. Alon, M.G. Surette, N. Barkai, S. Leibler, Robustness in bacterial chemotaxis. Nature 15, 168–171 (1999)

    Google Scholar 

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

    Article  Google Scholar 

  4. L. Blanchoin, T.D. Pollard, Hydrolysis of ATP by polymerized actin depends on the bound divalent cation but not profilin. Biochemistry 41 (2), 597–602 (2002)

    Article  Google Scholar 

  5. D. Boal, Mechanics of the Cell (Cambridge University Press, Cambridge, 2002)

    Google Scholar 

  6. J.T. Bonner, The Cellular Slime Molds (Princeton University Press, Princeton, NJ, 1967)

    Book  Google Scholar 

  7. J.T. Bonner, A way of following individual cells in the migrating slugs of Dictyostelium discoideum. Proc. Natl. Acad. Sci. 95 (16), 9355–9359 (1998)

    Article  Google Scholar 

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

    Google Scholar 

  9. T. Bretschneider, K. Anderson, M. Ecke, A. Müller-Taubenberger, B. Schroth-Diez, H.C. Ishikawa-Ankerhold, G. Gerisch, The three-dimensional dynamics of actin waves, a model of cytoskeletal self-organization. Biophys. J. 96 (7), 2888–2900 (2009)

    Article  Google Scholar 

  10. J.A. Brzostowski, C.A. Parent, A.R. Kimmel, A Gα-dependent pathway that antagonizes multiple chemoattractant responses that regulate directional cell movement. Genes Dev. 18, 805–15 (2004)

    Article  Google Scholar 

  11. B. Bugyi, M.F. Carlier, Control of actin filament treadmilling in cell motility. Annu. Rev. Biophys. 39, 449–470 (2010)

    Article  Google Scholar 

  12. L. Cai, A.M. Makhov, D.A. Schafer, J.E. Bear, Coronin 1B antagonizes cortactin and remodels arp2/3-containing actin branches in lamellipodia. Cell 134 (5), 828–842 (2008)

    Article  Google Scholar 

  13. Y. Cao, H. Li, L. Petzold, Efficient formulation of the stochastic simulation algorithm for chemically reacting systems. J. Chem. Phys. 121 (9), 4059–67 (2004)

    Article  Google Scholar 

  14. M. Carlier, D. Pantaloni, E. Korn, The effects of Mg2+ at the high-affinity and low-affinity sites on the polymerization of actin and associated ATP hydrolysis. J. Biol. Chem. 261, 10785–10792 (1986)

    Google Scholar 

  15. L.B. Case, C.M. Waterman, Adhesive F-actin waves: a novel integrin-mediated adhesion complex coupled to ventral actin polymerization. PLoS One 6 (11), e26631 (2011)

    Google Scholar 

  16. M.Y. Chen, R.H. Insall, P.N. Devreotes, Signaling through chemoattracnt receptors in Dictyostelium. Trends Genet. 12 (2), 52–57 (1996)

    Google Scholar 

  17. H. Chen, B.W. Bernstein, J.R. Bamburg, Regulating actin-filament dynamics in vivo. Trends Biochem. Sci. 25 (1), 19–23 (2000). Review

    Article  Google Scholar 

  18. Y. Cheng, H.G. Othmer, A model for direction sensing in dictyostelium discoideum: Ras activity and symmetry breaking driven by a g β γ - mediated, g α2-ric8 – dependent signal transduction network. PLoS Comput. Biol. 12, e1004900 (2016)

    Article  Google Scholar 

  19. C.Y. Chung, S. Funamoto, R.A. Firtel, Signaling pathways controlling cell polarity and chemotaxis. Trends Biochem. Sci. 26 (9), 557–566 (2001). Review

    Article  Google Scholar 

  20. J. Condeelis, A. Bresnick, M. Demma, S. Dharmawardhane, R. Eddy, A.L. Hall, R. Sauterer, V. Warren, Mechanisms of amoeboid chemotaxis: an evaluation of the cortical expansion model. Dev. Genet. 11 (5–6), 333–340 (1990)

    Article  Google Scholar 

  21. L.P. Cramer, M. Siebert, T.J. Mitchison, Identification of novel graded polarity actin filament bundles in locomoting heart fibroblasts: implications for the generation of motile force. J. Cell Biol. 136 (6), 1287–1305 (1997)

    Article  Google Scholar 

  22. J.C. Dallon, H.G. Othmer, A discrete cell model with adaptive signalling for aggregation of dictyostelium discoideum. Philos. Trans. R. Soc. Lond. B Biol. Sci. 352 (1351), 391–417 (1997)

    Article  Google Scholar 

  23. J.C. Dallon, H.G. Othmer, A continuum analysis of the chemotactic signal seen by dictyostelium discoideum. J. Theor. Biol. 194 (4), 461–483 (1998)

    Google Scholar 

  24. J.C. Dallon, H.G. Othmer, How cellular movement determines the collective force generated by the dictyostelium discoideum slug. J. Theor. Biol. 231, 203–222 (2004)

    Article  MathSciNet  Google Scholar 

  25. P. Dancker, L. Hess, Phalloidin reduces the release of inorganic phosphate during actin polymerization. Biochim. Biophys. Acta 1035 (2), 197–200 (1990)

    Article  Google Scholar 

  26. L.A. Davidson, R.E. Keller, Neural tube closure in xenopus laevis involves medial migration, directed protrusive activity, cell intercalation and convergent extension. Development 126 (20), 4547–4556 (1999)

    Google Scholar 

  27. P.N. Devreotes, T.L. Steck, Cyclic 3′, 5′ AMP relay in Dictyostelium discoideum II. Requirements for the initiation and termination of the response. J. Cell Biol. 80, 300–309 (1979)

    Google Scholar 

  28. P.N. Devreotes, P.L. Derstine, T.L. Steck, Cyclic 3′, 5′ AMP relay in Dictyostelium discoideum I. A technique to monitor responses to controlled stimuli. J. Cell Biol. 80, 291–299 (1979)

    Google Scholar 

  29. G. DeYoung, P.B. Monk, H.G. Othmer, Pacemakers in aggregation fields of Dictyostelium disc oideum. Does a single cell suffice? J. Math. Biol. 26, 486–517 (1988)

    Google Scholar 

  30. D. Dormann, C.J. Weijer, Propagating chemoattractant waves coordinate periodic cell movement in dictyostelium slugs. Development 128 (22), 4535–4543 (2001)

    Google Scholar 

  31. D. Dormann, G. Weijer, C.A. Parent, P.N. Devreotes, C.J. Weijer, Visualizing PI3 kinase-mediated cell-cell signaling during dictyostelium development. Curr. Biol. 12 (14), 1178–1188 (2002)

    Article  Google Scholar 

  32. R.A. Firtel, R. Meili, Dictyostelium: a model for regulated cell movement during morphogenesis. Curr. Opin. Genet. Dev. 10 (4), 421–427 (2000). Review

    Article  Google Scholar 

  33. P. Friedl, K. Wolf, Tumour-cell invasion and migration: diversity and escape mechanisms. Nat. Rev. Cancer 3 (5), 362–74 (2003)

    Article  Google Scholar 

  34. I. Fujiwara, S. Takahashi, H. Tadakuma, S. Ishiwata, Microscopic analysis of polymerization dynamics with individual actin filaments. Nat. Cell. Biol. 4 (9), 666–673 (2002)

    Article  Google Scholar 

  35. C. Gadgil, C.H. Lee, H.G. Othmer, A stochastic analysis of first-order reaction networks. Bull. Math. Biol. 67, 901–946 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  36. C.W. Gardiner, Handbook of Stochastic Methods (Springer, Berlin, Heildeberg, 1983)

    Book  MATH  Google Scholar 

  37. G. Gerisch, Chemotaxis in dictyostelium. Annu. Rev. Physiol. 44 (1), 535–552 (1982)

    Article  Google Scholar 

  38. G. Gerisch, U. Wick, Intracellular oscillations and release of cyclic AMP from Dictyostelium cells. Biochem. Biophys. Res. Commun. 65, 364–370 (1975)

    Article  Google Scholar 

  39. G. Gerisch, D. Hulser, D. Malchow, U. Wick, Cell communication by periodic cyclic amp pulses. Philos. Trans. R. Soc. Lond. 272, 181–192 (1975)

    Article  Google Scholar 

  40. G. Gerisch, T. Bretschneider, A. Muler-Taubenberger, E. Simmeth, M. Ecke, S. Diez, K. Anderson, Mobile actin clusters and traveling waves in cells recovering from actin depolymerization. Biophys. J. 87 (5), 3493–3503 (2004)

    Article  Google Scholar 

  41. D.T. Gillespie, A general method for numerically simulating the stochastic time evolution of coupled chemical reactions. J. Comput. Phys. 22, 403–434 (1976)

    Article  MathSciNet  Google Scholar 

  42. D.T. Gillespie, Exact stochastic simulation of coupled chemical reactions. J. Phys. Chem. 81 (25), 2340–2361 (1977)

    Article  Google Scholar 

  43. A.R. Gingle, Critical density for relaying in Dictyostelium discoideum and its relation to phosphodiesterase secretion into the extracellular medium. J. Cell Sci. 20, 1–20 (1976)

    Google Scholar 

  44. A.L. Hall, A. Schlein, J. Condeelis, Relationship of pseudopod extension to chemotactic hormone-induced actin polymerization in amoeboid cells. J. Cell Biol. 37 (3), 285–299 (1988)

    Google Scholar 

  45. B. Heit, S. Tavener, E. Raharjo, P. Kubes, An intracellular signaling hierarchy determines direction of migration in opposing chemotactic gradients. J. Cell Biol. 159 (1), 91–102 (2002)

    Article  Google Scholar 

  46. J. Hu, H.G. Othmer, A theoretical analysis of filament length fluctuations in actin and other polymers. J. Math. Biol. 63 (6), 1001–1049 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  47. J. Hu, A. Matzavinos, H.G. Othmer, A theoretical approach to actin filament dynamics. J. Stat. Phys. 128 (1–2), 111–138 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  48. J. Hu, H.W. Kang, H.G. Othmer, Stochastic analysis of reaction–diffusion processes. Bull. Math. Biol. 76, 854–894 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  49. M. Iijima, Y.E. Huang, P. Devreotes, Temporal and spatial regulation of chemotaxis. Dev. Cell 3 (4), 469–478 (2002). Review

    Article  Google Scholar 

  50. K. Inouye, Measurement of the motive force of the migrating slug of dictyostelium discoideum by a centrifuge method. Protoplasma 121, 171–177 (1984)

    Article  Google Scholar 

  51. K. Inouye, I. Takeuchi, Analytical studies on migrating movement of the pseudoplasmodium of Dictyostelium Discoideum. Protoplasma 99, 289–304 (1979)

    Google Scholar 

  52. R.H. Insall, O.D. Weiner, PIP3, PIP2, and cell movement–similar messages, different meanings? Dev. Cell. 1 (6), 743–747 (2001)

    Article  Google Scholar 

  53. C. Janetopoulos, T. Jin, P. Devreotes, Receptor-mediated activation of heterotrimeric G-proteins in living cells. Science 291 (5512), 2408–2411 (2001)

    Article  Google Scholar 

  54. P.A. Janmey, Mechanical properties of cytoskeletal polymers. Curr. Opin. Cell Biol. 2, 4–11 (1991)

    Article  Google Scholar 

  55. T. Jin, N. Zhang, Y. Long, C.A. Parent, P.N. Devreotes, Localization of the G protein β γ complex in living cells during chemotaxis. Science 287 (5455), 1034–1036 (2000)

    Article  Google Scholar 

  56. H.W. Kang, L. Zheng, H.G. Othmer, A new method for choosing the computational cell in stochastic reaction–diffusion systems. J. Math. Biol. 60, 1017–1099 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  57. H.W. Kang, L. Zheng, H.G. Othmer, A new method for choosing the computational cell in stochastic reaction-diffusion systems. J. Math. Biol. 60, 1017–1099 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  58. V. Khamviwath, J. Hu, H.G. Othmer, A continuum model of actin waves in dictyostelium discoideum. PloS One 8 (5), e64272 (2013)

    Google Scholar 

  59. A.R. Kimmel, C.A. Parent, The signal to move: D. discoideum go orienteering. Science 300 (5625), 1525–1527 (2003)

    Google Scholar 

  60. J.R. Kuhn, T.D. Pollard, Real-time measurements of actin filament polymerization by total internal reflection fluorescence microscopy. Biophys. J. 88 (2), 1387–1402 (2005)

    Article  Google Scholar 

  61. D.A. Lauffenburger, A.F. Horwitz, Cell migration: a physically integrated molecular process. Cell 84, 359–369 (1996)

    Article  Google Scholar 

  62. A. Levchenko, P.A. Iglesias, Models of eukaryotic gradient sensing: application to chemotaxis of amoebae and neutrophils. Biophys. J. 82 (1 Pt 1), 50–63 (2002)

    Article  Google Scholar 

  63. L. Limozin, M. Barmann, E. Sackmann, On the organization of self-assembled actin networks in giant vesicles. Eur. Phys. J E 10 (4), 319–330 (2003)

    Article  Google Scholar 

  64. F.C. MacKintosh, Theoretical models of viscoelasticity of actin solutions and the actin cortex. Biol. Bull. 194 (3), 351–353 (1998). No abstract available

    Google Scholar 

  65. A. Matzavinos, H.G. Othmer, A stochastic analysis of actin polymerization in the presence of twinfilin and gelsolin. J. Theor. Biol. 249 (4), 723–736 (2007)

    Article  MathSciNet  Google Scholar 

  66. J.L. McGrath, E.A. Osborn, Y.S. Tardy, C.F. Dewey Jr, J.H. Hartwig, Regulation of the actin cycle in vivo by actin filament severing. Proc. Natl. Acad. Sci. USA 97 (12), 6532–7 (2000)

    Article  Google Scholar 

  67. S. McLaughlin, J. Wang, A. Gambhir, D. Murray, PIP2 and proteins: interactions, organization and information flow. Annu. Rev. Biophys. Biomol. Struct. 31, 151–175 (2002). Review

    Article  Google Scholar 

  68. R. Meili, C. Ellsworth, S. Lee, T.B. Reddy, H. Ma, R.A. Firtel, Chemoattractant-mediated transient activation and membrane localization of akt/PKB is required for efficient chemotaxis to cAMP in dictyostelium. EMBO J. 18 (8), 2092–2105 (1999)

    Article  Google Scholar 

  69. H. Meinhardt, Orientation of chemotactic cells and growth cones: models and mechanisms. J. Cell Sci. 17 (17), 2867–2874 (1999)

    Google Scholar 

  70. R. Melki, S. Fievez, M.F. Carlier, Continuous monitoring of pi release following nucleotide hydrolysis in actin or tubulin assembly using 2-amino-6-mercapto-7-methylpurine ribonucleoside and purine-nucleoside phosphorylase as an enzyme-linked assay. Biochemistry 35 (37), 12038–45 (1996)

    Article  Google Scholar 

  71. T.J. Mitchison, L.P. Cramer, Actin-based cell motility and cell locomotion. Cell 3, 371–379 (1996). Review. No abstract available

    Google Scholar 

  72. F. Oosawa, S. Asakura, Thermodynamics of the Polymerization of Protein (Academic, New York, 1975)

    Google Scholar 

  73. H.G. Othmer, A graph-theoretic analysis of chemical reaction networks (1979). Lecture Notes, Rutgers University. Available at www.math.umn.edu/~othmer/graphrt.pdf

  74. H.G. Othmer, P. Schaap, Oscillatory cAMP signaling in the development of dictyostelium discoideum. Comments Theor. Biol. 5, 175–282 (1998)

    Google Scholar 

  75. E. Palsson, H.G. Othmer, A model for individual and collective cell movement in dictyostelium discoideum. Proc. Natl. Acad. Sci. 97, 11448–11453 (2000)

    Google Scholar 

  76. C.A. Parent, P.N. Devreotes, A cell’s sense of direction. Science 284 (5415), 765–770 (1999). Review

    Article  Google Scholar 

  77. E. Pate, H.G. Othmer, Differentiation, cell sorting and proportion regulation in the slug stage of Dictyostelium discoideum. J. Theor. Biol. 118, 301–319 (1986)

    Google Scholar 

  78. M. Pineda, C. Weijer, R. Eftimie, Modelling cell movement, cell differentiation, cell sorting and proportion regulation in dictyostelium discoideum aggregations. J. Theor. Biol. 370, 135–150 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  79. T.D. Pollard, Regulation of actin filament assembly by arp2/3 complex and formins. Annu. Rev. Biophys. Biomol. Struct. 36, 451–477 (2007)

    Article  Google Scholar 

  80. T.D. Pollard, L. Blanchoin, R.D. Mullins, Molecular mechanisms controlling actin filament dynamics in nonmuscle cells. Annu. Rev. Biophys. Biomol. Struct. 29 (1), 545–76 (2000)

    Article  Google Scholar 

  81. A.Y. Pollitt, R.H. Insall, WASP and SCAR/WAVE proteins: the drivers of actin assembly. J. Cell Sci. 122 (Pt 15), 2575–2578 (2009). doi:10.1242/jcs.023879

    Article  Google Scholar 

  82. R.K. Raman, Y. Hashimoto, M.H. Cohen, A. Robertson, Differentiation for aggregation in the cellular slime molds: the emergence of autonomously signalling cells in Dictyostelium discoideum. J. Cell. Sci. 21, 243–259 (1976)

    Google Scholar 

  83. D. Raucher, T. Stauffer, W. Chen, K. Shen, S. Guo, J.D. York, M.P. Sheetz, T. Meyer, Phosphatidylinositol 4,5-bisphosphate functions as a second messenger that regulates cytoskeleton-plasma membrane adhesion. Cell 100 (2), 221–228 (2000)

    Article  Google Scholar 

  84. J. Rosenblatt, P. Peluso, T.J. Mitchison, The bulk of unpolymerized actin in xenopus egg extracts is ATP-bound. Mol. Biol. Cell 6 (2), 227–36 (1995)

    Article  Google Scholar 

  85. K.R. Sanft, H.G. Othmer, Constant-complexity stochastic simulation algorithm with optimal binning. J. Chem. Phys. 143 (8), 074108 (2015)

    Google Scholar 

  86. P. Schaap, Evolutionary crossroads in developmental biology: dictyostelium discoideum. Development 138 (3), 387–396 (2011)

    Article  Google Scholar 

  87. D. Sept, J. Xu, T.D. Pollard, J.A. McCammon, Annealing accounts for the length of actin filaments formed by spontaneous polymerization. Biophys. J. 77 (6), 2911–2919 (1999)

    Article  Google Scholar 

  88. M.P. Sheetz, D. Felsenfeld, C.G. Galbraith, D. Choquet, Cell migration as a five-step cycle. Biochem. Soc. Symp. 65, 233–43 (1999)

    Google Scholar 

  89. F. Siegert, C.J. Weijer, Analysis of optical density wave propagation and cell movement in the cellular slime mould Dictyostelium discoideum 49, 224–232 (1991)

    Google Scholar 

  90. J.V. Small, Microfilament-based motility in non-muscle cells. Curr. Opin. Cell Biol. 1, 75–79 (1989)

    Article  Google Scholar 

  91. J.V. Small, T. Stradal, E. Vignal, K. Rottner, The lamellipodium: where motility begins. Trends Cell Biol. 12 (3), 112–120 (2002). Review

    Article  Google Scholar 

  92. D.R. Soll, The use of computers in understanding how animal cells crawl, in International Review of Cytology, vol. 163, ed. by K.W. Jeon, J. Jarvik (Academic, New York, 1995), pp. 43–104

    Google Scholar 

  93. P.A. Spiro, J.S. Parkinson, H.G. Othmer, A model of excitation and adaptation in bacterial chemotaxis. Proc. Natl. Acad. Sci. 94 (14), 7263–7268 (1997)

    Article  Google Scholar 

  94. T.M. Svitkina, G.G. Borisy, Arp2/3 complex and actin depolymerizing factor/cofilin in dendritic organization and treadmilling of actin filament array in lamellipodia. J. Cell Sci. 145(5), 1009–26 (1999)

    Article  Google Scholar 

  95. J. Swanson, D.L. Taylor, Local and spatially coordinated movements in D ictyostelium discoideum amoedae during chemotaxis. Cell 28, 225–232 (1982)

    Article  Google Scholar 

  96. Y. Tang, H.G. Othmer, A G protein-based model of adaptation in Dictyostelium discoideum. Math. Biosci. 120 (1), 25–76 (1994)

    Google Scholar 

  97. Y. Tang, H.G. Othmer, Excitation, oscillations and wave propagation in a G-protein-based model of signal transduction in Dictyostelium discoideum. Philos. Trans. R. Soc. Lond. B Biol. Sci. 349 (1328), 179–95 (1995)

    Article  Google Scholar 

  98. J.A. Theriot, T.J. Mitchison, Actin microfilament dynamics in locomoting cells. Nature 352 (6331), 126–131 (1991)

    Article  Google Scholar 

  99. A. Tikhonov, Systems of differential equations containing small parameters multiplying derivatives. Math. Sb. 31, 575–586 (1952)

    Google Scholar 

  100. P.H. Vardy, L.R. Fisher, E. Smith, K.L. Williams, Traction proteins in the extracellular matrix of Dictyostelium discoideum slugs. Nature 320 (6062), 526–529 (1986)

    Article  Google Scholar 

  101. B. Varnum, K.B. Edwards, D.R. Soll, Dictyostelium amebae alter motility differently in response to increasing versus decreasing temporal gradients of cAMP. J Cell Biol. 101, 1–5 (1985)

    Article  Google Scholar 

  102. B. Vasiev, C.J. Weijer, Modelling of dictyostelium discoideum slug migration. J. Theor. Biol. 223, 347–59 (2003)

    Article  MathSciNet  Google Scholar 

  103. J.B. Wallingford, L.A. Niswander, G.M. Shaw, R.H. Finnell, The continuing challenge of understanding, preventing, and treating neural tube defects. Science 339 (6123), 1222002 (2013)

    Google Scholar 

  104. C.J. Weijer, Signalling during dictyostelium development, in Dictyostelids (Springer, Berlin, 2013), pp. 49–70

    Google Scholar 

  105. D. Wessels, J. Murray, D.R. Soll, Behavior of Dictyostelium amoebae is regulated primarily by the temporal dynamic of the natural cAMP wave. Cell Motil. Cytoskeleton 23 (2), 145–156 (1992)

    Article  Google Scholar 

  106. L. Wolpert, R. Beddington, T. Jessel, P. Lawrence, E. Meyerowitz, J. Smith, Principles of Development (Oxford University Press, Oxford, 2002)

    Google Scholar 

  107. B. Wurster, K. Schubiger, U. Wick, G. Gerisch, Cyclic GMP in Dictyostelium discoideum: oscillations and pulses in response to folic acid and cyclic AMP signals. FEBS Lett. 76, 141–144 (1977)

    Article  Google Scholar 

  108. X. Xin, H.G. Othmer, A trimer of dimers- based model for the chemotactic signal transduction network in bacterial chemotaxis. Bull. Math. Biol. 74, 2339–2382 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  109. X.S. Xu, A. Kuspa, D. Fuller, W.F. Loomis, D.A. Knecht, Cell-cell adhesion prevents mutant cells lacking myosin II from penetrating aggregation streams of dictyostelium. Dev. Biol. 175 (2), 218–226 (1996)

    Article  Google Scholar 

  110. J. Xu, Y. Tseng, D. Wirtz, Strain hardening of actin filament networks. regulation by the dynamic cross-linking protein alpha-actinin. J. Biol. Chem. 275 (46), 35886–35892 (2000)

    Google Scholar 

  111. S.H. Zigmond, Recent quantitative studies of actin filament turnover during locomotion. Cell Motil. Cytoskeleton 25, 309–316 (1993)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hans G. Othmer .

Editor information

Editors and Affiliations

Appendices

Appendix: Singular Perturbation Reduction

We know how to express reaction dynamics as the evolution equation

$$\displaystyle{ \frac{dx} {dt} = f(x,p) }$$
(1.27)

where x ∈ M ⊆ R n and p ∈ R p. Suppose there is a separation of variables into those that vary rapidly and those that change slowly on the chosen time scale. This leads to a system of equations of the form

$$\displaystyle\begin{array}{rcl} \dfrac{dx} {dt} = f(x,y,\epsilon,p)& & \\ \epsilon \dfrac{dy} {dt} = g(x,y,\epsilon,p)& &{}\end{array}$$
(1.28)

wherein ε is a small parameter and x (resp., y) is a slow (resp., fast) variable on the t time scale.

  • What are the implications of this separation for understanding the dynamics?

  • More precisely, can we use the separation of time scales to reduce the dimensionality of the problem we have to analyze?

The following simple example will illustrate some of the underlying ideas. Consider the reaction

$$\displaystyle{ A\stackrel{k_{1}}{\longrightarrow }B\stackrel{k_{2}}{\longrightarrow }C, }$$
(1.29)

where A, B, and C are generic chemical species. We assume mass action kinetics and first-order reactions and therefore can write the governing equations as

$$\displaystyle\begin{array}{rcl} \mathop{x}\limits^{\cdot }& =& k_{1}x \\ \mathop{y}\limits^{\cdot }& =& k_{1}x - k_{2}y \\ \mathop{z}\limits^{\cdot }& =& k_{2}y. {}\end{array}$$
(1.30)

where x, y, and z stand for the concentration of A, B, and C, resp. In addition we have the conservation condition

$$\displaystyle{ (x + y + z)(t) = (x + y + z)(0). }$$
(1.31)

Clearly A and B will disappear completely in the steady state, and the exact solution for the transient problem for A and B is

$$\displaystyle\begin{array}{rcl} x& =& x_{0}e^{-k_{1}t} {}\\ y& =& y_{0}e^{-k_{2}t} + \dfrac{k_{1}x_{0}} {k_{2} - k_{1}}[e^{-k_{1}t} - e^{-k_{2}t}]. {}\\ \end{array}$$

If k 2 ≫ k 1, then after a time \(t \sim \mathcal{O}(1/k_{2})\) we have

$$\displaystyle{y(t) \sim \dfrac{k_{1}x_{0}e^{-k_{1}t}} {k_{2}} = \frac{k_{1}} {k_{2}}x(t).}$$

But this is what we would get directly if we set \(\mathop{y}\limits^{\cdot } = k_{1}x - k_{2}y = 0\) since then

$$\displaystyle{y = \frac{k_{1}x} {k_{2}}.}$$

This is what we call the ‘pseudo-steady-state’ or PSS value of y, (or alternatively, the ‘quasi-steady-state’ QSS value of y), and the hypothesis that we can do this is called the ‘pseudo-(or quasi) steady-state hypothesis’—the PSSH or QSSA. Can we justify setting \(\mathop{y}\limits^{\cdot } = 0\) analytically, and how do we understand it geometrically?

To formalize this reduction, consider the system

$$\displaystyle\begin{array}{rcl} \dfrac{dx} {d\tau } & =& f(x,y) \\ \epsilon \;\dfrac{dy} {d\tau } & =& g(x,y),{}\end{array}$$
(1.32)

where x ∈ R n, y ∈ R m, and ε > 0 is a small parameter.

Let y = ϕ(x) be one of the solutions of the system

$$\displaystyle{ g(x,y) = 0 }$$
(1.33)

defined on a closed bounded domain D ⊂ R n. The degenerate or slow system associated with (1.32) and the solution y = ϕ(x) is the system

$$\displaystyle{ \dfrac{dx} {d\tau } = f(x,\phi (x)). }$$
(1.34)

The solution y = ϕ(x) is said to be isolated on D if there exists an η > 0 such that (1.33) has no solution other than ϕ with the property that

$$\displaystyle{\|y -\phi (x)\| <\eta.}$$

The associated adjoined (or fast) system is

$$\displaystyle{ \dfrac{dy} {ds} = g(x^{{\ast}},y), }$$
(1.35)

where x is regarded as a parameter.

  • The isolated solution y = ϕ(x) is said to be positively stable in D if for all x in D the points y = ϕ(x ) are asymptotically stable stationary points, in the sense of Lyapunov, of the adjoined system.

  • The domain of influence of an isolated, positively stable y = ϕ(x)is the set of points (x , y ) such that the solution of the adjoined system with the initial conditions y(0) = y tends to ϕ(x ) as s → .

One of the earliest formal statements of what is now know as a singular perturbation reduction is given by Tikhonov’s theorem [99].

Theorem 2

If some root y = ϕ(x) of g = 0 is an isolated, positively stable root in some bounded closed domain D, if the initial point (x(0), y(0)) belongs to the domain of influence of this root, and if the solution of the degenerate system belongs to D for 0 ≤ t ≤ T, then the solution (x(t,ε),y(t,ε)) of the original system tends to the solution \((\bar{x}(t),\bar{y}(t))\) of the degenerate or slow system, as ε approaches 0, the passage to the limit

$$\displaystyle{\lim _{\epsilon \rightarrow 0}y(t,\epsilon ) =\tilde{ y}(t) =\phi (\bar{x},t)}$$

holding for 0 < t ≤ T 0 < T, and the passage to the limit

$$\displaystyle{\lim _{\epsilon \rightarrow 0}x(t,\epsilon ) =\bar{ x}(t)}$$

holding for 0 ≤ t ≤ T 0 < T.

The geometry behind the theorem is qualitatively as shown in the following sketch.

Remarks

 

  1. (1)

    This can be generalized to allow ε and t dependence in the right-hand side of the DE.

  2. (2)

    If y is sufficiently close to ϕ(x) then on the fast time scale this analysis can be viewed in the framework of center manifold theory.

  3. (3)

    There are many variations on the theme and many expositions of it.

Glossary

Actin is one of the most abundant proteins found in eukaryotic cells and is highly conserved throughout evolution. Actin monomers polymerize in an ATP-dependent process to form polarized actin filaments. The helical actin filament has a barbed end and a pointed end. In non-muscle cells, the dynamic organization of filamentous actin gives rise to various cell structures, such as filopodia, lamellipodia, and stress fibers.

AKT/PKB Cellular homologue of AKT retroviral oncogene protein Ser/Thr kinase. Also called Protein Kinase B because of homology with Protein Kinase A and Protein Kinase C family members. AKT has an amino-terminal pleckstrin homology (PH) domain that binds to the lipid products of phosphoinositide 3-kinase, phosphatidylinositol-3,4-bisphosphate and phosphatidylinositol-3,4,5-trisphosphate. This binding locates it at the plasma membrane where it becomes phosphorylated on Thr-308 (human AKT1) in the activation loop of the catalytic domain by phosphoinositide-dependent kinase 1 (PDK1). This phosphorylation leads to activation. Full activation requires phosphorylation at a second site (Ser-473 of human AKT). Mammals have at least three distinct genes for AKT family members (AKT1, AKT2 and AKT3) and they appear to at least partially redundant in function.

Alpha-actinin is an actin cross-linking protein that belongs to the spectrin superfamily. It forms antiparallel homodimers in a rod-like structure with one actin-binding domain on each side of the rod. It can, therefore, cross-link two filaments of actin.

Arp2/3 (Actin-related protein complex) Extracellular and intracellular signals that initiate actin cytoskeletal rearrangements flow through the Arp2/3 complex. The Arp2/3 complex is composed of seven subunits, two unconventional actin proteins, Arp2 and Arp3, and five additional proteins. One copy of each subunit is present in the complex. In addition to contributing to the nucleation of actin filaments, the Arp2/3 complex caps the slow-growing pointed ends of actin filaments and promotes the elongation of the actin filament at the fast-growing barbed end. Arp2/3 complexes are localized at the branch point of two different filaments, binding to both the side of the mother filament and the pointed end of the daughter filament. These interactions create an approximate 70 angle with the actin barbed end oriented towards the plasma membrane. These observations serve as the basis for the “dendritic nucleation” model in which Arp2/3, through its interaction with F-actin, is a key regulator of the unique meshwork of F-actin in the lamellipodial structure. The Arp2/3 complex influences the strictly regulated reorganization of the actin cytoskeleton in response to various signaling pathways.

B-Raf is a serine/threonine protein kinase that is a member of the Raf family of protein kinases. Raf-B can be activated by the small GTPase Ras as well as Rap1. cAMP activates MAP-Kinase through B-Raf.

Cdc42 is a Rho-family GTPase that was first identified as a regulator of cell polarity and budding in yeast. In mammalian cells, Cdc42 is activated in response to integrins and bradykinin. Activated Cdc42 induces formation of actin-rich structures known as filopodia. Cdc42 affects actin directly by binding to its effector WASP, an activator of Arp2/3 and actin polymerization. During cell migration, Cdc42 is also responsible for polarization of the cell.

Chemotaxis the directed movement of a microorganism or cell in response to a chemical stimulus.

G-proteins G-proteins are heterotrimeric proteins comprising alpha, beta and gamma subunits. When bound to guanine nucleotide diphosphate (GDP), G-proteins are inactive, and are activated when bound to guanine nucleotide triphosphate (GTP). When activated, they split into alpha-GTP and beta-gamma subunits that are separately used for control of downstream processes. The alpha subunit has intrinsic hydrolysis activity that removes a phosphate group, whereupon the alpha-GDP subunit recombines with a beta-gamma subunit.

G-protein coupled receptors (GPCRs) G-protein-coupled receptors are integral membrane proteins with seven membrane-spanning helices. G-protein is bound to the receptor in its inactive state, and when a ligand is bound, the receptor may undergo a conformational change that facilitates the GDP-GTP exchange on a G protein, which then detaches from the receptor. GPCRs can desensitize when exposed to ligand for a sufficient length of time as a result of phosphorylation of a cytoplasmic receptor domain by protein kinases.

In vitro the recreation of biological processes in an artificial laboratory environment.

In vivo biological processes that take place within a living organism or cell.

Kinase an enzyme that catalyzes addition of a phosphoryl group to proteins.

MLCK (Myosin Light Chain kinase) is a serine/threonine kinase that phosphorylates the regulatory light chain (RLC) of the molecular motor myosin, resulting in increased contraction. In nonmuscle cells MLCK is involved in the regulation of cell motility.

MLCP (Myosin Light Chain phosphatase) dephosphorylates the regulatory light chain of molecular motor myosin, thus inactivating it. MLCP is involved in the regulation of the actin cytoskeleton and cell spreading.

PAK (p21-activated kinase) is a serine/threonine kinase that binds to and is activated by Rho family GTPases. The PAK family of kinases is highly conserved among species. PAK family members have an N-terminal kinase domain and a C-terminal p21 Rac and Cdc-42-binding domain. The binding of Rho family GTPases to the C-terminus of PAK causes a conformational change resulting in autophosphorylation and activation of the kinase. PAK family members have been shown to play a role in actin polymerization and cytoskeletal dynamics.

PI3K Phosphoinositide 3-kinase catalyzes the phosphoryation of phosphatidylinositol-4,5-bisphosphate to produce phosphatidylinositol-3,4, 5-trisphosphate. The enzyme is cytosolic and become localized to the plasma membrane, which activates the catalytic subunit. Binding of the catalytic subunit to activated Ras appears to enhance the membrane association and activation.

Phosphatase an enzyme that removes the phosphoryl group from proteins.

Protease (proteinase, proteolytic enzyme) an enzyme that degrades proteins, such as collagen, laminin, and fibrin.

PTEN is a phosphatase that can dephosphorylate the lipid products or phosphoinositide 3-kinase, especially phosphatidylinositol-3,4,5-trisphosphate. By dephosphorylating these lipids, PTEN attenuates signaling via the phosphoinositide 3-kinase pathway.

Rac is a small GTP-binding protein which belongs to the Ras superfamily. Rac activity depends on the ratio of GTP/GDP bound forms. GEFs (guanine nucleotide exchange factors), GAPs (GTPase activating proteins) and GDIs (guanine nucleotide dissociation inhibitors) thus determine its activity by regulating of the GTP/GDP bound form. The downstream effectors of Rac are involved in regulation of the cytoskeleton.

Ras is a small protein that is activated by exchange proteins of the SOS (Son of Sevenless) or Ras-GRF (Ras GDP Releasing Factor) family. The release of GDP and binding of GTP ‘activates’ Ras, allowing it to bind to downstream effectors, including phosphoinositide 3-kinase (PI3K). Ras is anchored at the plasma membrane via a carboxy-terminal farnesyl group. One function of Ras is to facilitate localization of its cytosolic effectors at the plasma membrane. Mutant forms of Ras that stabilize the GTP bound state contribute to tumor formation.

Rho GTPase is a member of Rho family of small GTP binding proteins, which include Rho A, B, G. It is active in the GTP-bound state and inactive when bound to GDP. Activation of Rho results in its association with a wide variety of effector molecules and leads to activation of downstream signaling cascades. Rho proteins exhibit intrinsic GTPase activity Rho that is significantly stimulated by GTPase activating proteins (GAPs). Conversely, GTP exchange factors (GEFs) promote exchange of GDP for GTP, thus activating Rho. Rho GTPase regulates actin cytoskeleton organization and assembly, in particular actin stress fibers and focal adhesion formation. It is involved in cellular processes that depend on actin cytoskeleton such as cell spreading and migration.

RLC (Regulatory Light Chain of myosin) binds to the neck region of Heavy Chain (HC) of myosin and regulates its motor activity. When RLC is phosphorylated by myosin light chain kinase on Ser19 it causes conformational changes in myosin and activates its motor function. Dephosphorylation of RLC by myosin light chain phosphatase negatively regulates myosin activity. Phosphorylation of RLC in smooth muscle cells leads to initiation of contraction, whereas in striated muscles it increases speed and force of contractions. There is also evidence that in nonmuscle cells myosins are participating in cell motility, spreading, cytokinesis and vesicle transport.

ROCK Rho kinase/ROKalpha/ROCK-II and a related protein p160ROCK/ROCK-I are serine/threonine kinases. Integrin-dependent activation of Rho family small GTPases induces Rho-Rho kinase interaction and facilitates translocation of Rho kinase to the cell membrane. Rho kinase facilitates formation of focal adhesions and actin stress fiber assembly. It cooperates with other Rho effectors to regulate the formation of actin structures. Rho kinase associates with and phosphrylates myosin phosphatase. Phosphorylation leads to the inhibition of phosphatase activity and thereby enhances myosin-mediated contractility. Rho kinase has been shown to facilitate tumor cell invasion due to its role in cytoskeletal reorganization. In addition, Rho kinase promotes cytokinesis by phosphorylation of myosin light chain at the cleavage furrow.

Small G proteins A class of monomeric, low molecular weight (20–25 kDa) GTP-binding proteins that regulate a variety of intracellular processes. The GTP bound form of the protein is active and is inactivated by an intrinsic GTPase activity, which is controlled by GTPase activators (GAPs), GDP dissociation inhibitors (GDIs), and guanine nucleotide exchange factors (GEFs).

SOS (Son of Sevenless) is an exchange factor that stimulates GDP release from Ras to allow GTP binding and formation of the active state. SOS has a pleckstrin homology (PH) domain that mediates binding to the membrane.

WASP (The Wiskott-Aldrich Syndrome protein) mediates a variety of signals from kinases, receptors, and small G proteins to the Arp2/3 complex and the actin cytoskeleton. Homologs in the WASP family include SCAR and WAVE. The family members share two similar domains: one that mediates interactions of WASP with adaptor proteins such as profilin, and another involved in interaction with Arp2/3 complex and G-actin. GTP-Cdc42 binding to WASP stimulates the Arp2/3 complex resulting in actin polymerization.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Othmer, H.G. (2016). Cell-Based, Continuum and Hybrid Models of Tissue Dynamics. In: Preziosi, L., Chaplain, M., Pugliese, A. (eds) Mathematical Models and Methods for Living Systems. Lecture Notes in Mathematics(), vol 2167. Springer, Cham. https://doi.org/10.1007/978-3-319-42679-2_1

Download citation

Publish with us

Policies and ethics