Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-18T19:47:07.569Z Has data issue: false hasContentIssue false

Thermodynamically consistent Navier–Stokes–Cahn–Hilliard models with mass transfer and chemotaxis

Published online by Cambridge University Press:  09 October 2017

KEI FONG LAM
Affiliation:
Department of Mathematics, The Chinese University of Hong Kong, Shatin, N.T., Hong Kong email: kflam@math.cuhk.edu.hk
HAO WU
Affiliation:
School of Mathematical Sciences and Shanghai Key Laboratory for Contemporary Applied Mathematics, Fudan University, 220 Han Dan Road, Shanghai 20043, China email: haowufd@fudan.edu.cn, haowufd@yahoo.com

Abstract

We derive a class of Navier–Stokes–Cahn–Hilliard systems that models two-phase flows with mass transfer coupled to the process of chemotaxis. These thermodynamically consistent models can be seen as the natural Navier–Stokes analogues of earlier Cahn–Hilliard–Darcy models proposed for modelling tumour growth, and are derived based on a volume-averaged velocity, which yields simpler expressions compared to models derived based on a mass-averaged velocity. Then, we perform mathematical analysis on a simplified model variant with zero excess of total mass and equal densities. We establish the existence of global weak solutions in two and three dimensions for prescribed mass transfer terms. Under additional assumptions, we prove the global strong well-posedness in two dimensions with variable fluid viscosity and mobilities, which also includes a continuous dependence on initial data and mass transfer terms for the chemical potential and the order parameter in strong norms.

Type
Papers
Copyright
Copyright © Cambridge University Press 2017 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

[1] Abels, H. (2009) Existence of weak solutions for a diffuse interface model for viscous, incompressible fluids with general densities. Commun. Math. Phys. 289 (1), 4573.CrossRefGoogle Scholar
[2] Abels, H. (2009) On a diffuse interface model for two-phase flows of viscous, incompressible fluids with matched densities. Arch. Ration. Mech. Anal. 194 (2), 463506.Google Scholar
[3] Abels, H., Depner, D. & Garcke, H. (2013) Existence of weak solutions for a diffuse interface model for two-phase flows of incompressible fluids with different densities. J. Math. Fluid Mech. 15 (3), 453480.Google Scholar
[4] Abels, H., Depner, D. & Garcke, H. (2013) On an incompressible Navier–Stokes/Cahn–Hilliard system with degenerate mobility. Ann. Inst. H. Poincaré Anal. Non Linéaire 30 (6), 11751190.CrossRefGoogle Scholar
[5] Abels, H., Garcke, H. & Grün, G. (2012) Thermodynamically consistent, frame indifferent diffuse interface models for incompressible two-phase flow with different densities. Math. Models Methods Appl. Sci. 22 (3), 1150013, 40pp.CrossRefGoogle Scholar
[6] Abels, H., Garcke, H. & Weber, J. Existence of weak solutions for a diffuse interface model for two-phase flow with surfactants. In preparation.Google Scholar
[7] Alber, M. S., Kiskowski, M. A., Glazier, J. A. & Jiang, Y. (2003) On cellular automaton approaches to modeling biological cells. In: Rosenthal, J. and Giliam, D. S. (editors), Mathematical Systems Theory in Biology, Communications, Computation, and Finance, Vol. 134, Springer, New York, pp. 139.Google Scholar
[8] Ambrosi, D. & Preziosi, L. (2002) On the closure of mass balance models for tumor growth. Math. Models Methods Appl. Sci. 12 (5), 737754.CrossRefGoogle Scholar
[9] Anderson, D.-M., McFadden, G.-B. & Wheeler, A.-A. (1997) Diffuse-interface methods in fluid mechanics. Annu. Rev. Fluid Mech. 30 (1), 139165.CrossRefGoogle Scholar
[10] Beysens, D. A., Forgacs, G. & Glazier, J. A. (2000) Cell sorting is analogous to phase ordering in fluids. Proc. Natl. Acad. Sci. USA 97 (17), 94679471.Google Scholar
[11] Boyer, F. (1999) Mathematical study of multi-phase flow under shear through order parameter formulation. Asymptot. Anal. 20 (2), 175212.Google Scholar
[12] Brézis, H. and Gallouet, T. (1980) Nonlinear Schrödinger evolution equations. Nonlinear Anal. 4 (4), 677681.Google Scholar
[13] Chae, M., Kang, K. & Lee, J. (2013) Existence of smooth solutions to coupled chemotaxis fluid equations. Discrete Contin. Dyn. Syst. 33 (6), 22712297.Google Scholar
[14] Chae, M., Kang, K. & Lee, J. (2014) Global existence and temporal decay in Keller–Segel models coupled to fluid equations. Commun. Partial Differ. Equ. 39 (7), 12051235.CrossRefGoogle Scholar
[15] Colli, P., Gilardi, G. & Hilhorst, D. (2015) On a Cahn–Hilliard type phase field model related to tumor growth. Discrete Contin. Dyn. Syst. 35 (6), 24232442.CrossRefGoogle Scholar
[16] Cristini, V., Li, X., Lowengrub, J. S. & Wise, S. M. (2009) Nonlinear simulations of solid tumor growth using a mixture model: Invasion and branching. J. Math. Biol. 58, 723763.Google Scholar
[17] Cristini, V. & Lowengrub, J. (2010) Multiscale Modeling of Cancer: An Integrated Experimental and Mathematical Modeling Approach, Cambridge University Press, Cambridge, U.K.Google Scholar
[18] Dedè, L., Garcke, H. & Lam, K. F. (to appear) A Hele–Shaw–Cahn–Hilliard model for incompressible two-phase flows with different densities. J. Math. Fluid Mech. doi:10.1007/s00021-017-0334-5.Google Scholar
[19] Di Francesco, M., Lorz, A. & Markowich, P. A. (2010) Chemotaxis-fluid coupled model for swimming bacteria with nonlinear diffusion: Global existence and asymptotic behavior. Discrete Contin. Dyn. Syst. 28 (4), 14371453.Google Scholar
[20] Duan, R., Lorz, A. & Markowich, P. A. (2010) Global solutions to the coupled chemotaxis-fluid equations. Commun. Partial Differ. Equ. 35 (9), 16351673.Google Scholar
[21] Engler, H. (1989) An alternative proof of the Brezis-Wainger inequality. Comm. Partial Differ. Equ. 14 (4), 541544.Google Scholar
[22] Foty, R. A., Forgacs, G., Pfleger, C. M. & Steinberg, M. S. (1994) Liquid properties of embryonic tissues: Measurement of interfacial tensions. Phys. Rev. Lett. 72 (14), 22982301.Google Scholar
[23] Frieboes, H.B., Fang, J., Chuang, Y.-L., Wise, S. M., Lowengrub, J. S. & Cristini, V. (2010) Three-dimensional multispecies nonlinear tumor growth – II: Tumor invasion and angiogenesis. J. Theor. Biol. 264 (4), 12541278.Google Scholar
[24] Frigeri, S., Gal, C. & Grasselli, M. (2016) On nonlocal Cahn–Hilliard–Navier–Stokes systems in two dimensions. J. Nonlinear Sci. 26 (4), 847893.Google Scholar
[25] Frigeri, S., Grasselli, M. & Rocca, E. (2015) On a diffuse interface model of tumor growth. Eur. J. Appl. Math. 26 (2), 215243.CrossRefGoogle Scholar
[26] Gal, C. & Grasselli, M. (2010) Asymptotic behavior of a Cahn–Hilliard–Navier–Stokes system in 2D. Ann. Inst. H. Poincaré Anal. Non Linéaire 27 (1), 401436.CrossRefGoogle Scholar
[27] Gal, C. & Grasselli, M. (2010) Trajectory attractors for binary fluid mixtures in 3D. Chin. Ann. Math. Ser. B 31 (5), 655678.Google Scholar
[28] Garcke, H. & Lam, K. F. (2016) Global weak solutions and asymptotic limits of a Cahn–Hilliard–Darcy system modelling tumour growth. AIMS Math. 1 (3), 318360.CrossRefGoogle Scholar
[29] Garcke, H. & Lam, K. F. (2016) On a Cahn–Hilliard–Darcy system for tumour growth with solution dependent source terms. Preprint arXiv:1611.00234.Google Scholar
[30] Garcke, H. & Lam, K. F. (2017) Analysis of a Cahn–Hilliard system with non-zero Dirichlet conditions modeling tumor growth with chemotaxis. Discrete Contin. Dyn. Syst. 37 (8), 42774308.Google Scholar
[31] Garcke, H. & Lam, K. F. (2017) Well-posedness of a Cahn–Hilliard–Darcy system modelling tumour growth with chemotaxis and active transport. Eur. J. Appl. Math. 28 (2), 284316.Google Scholar
[32] Garcke, H., Lam, K. F., Sitka, E. & Styles, V. (2016) A Cahn–Hilliard–Darcy model for tumour growth with chemotaxis and active transport. Math. Models Methods Appl. Sci. 26 (6), 10951148.Google Scholar
[33] Giorgini, A., Grasselli, M. & Wu, H. (2017) On the Cahn–Hilliard–Hele–Shaw system with singular potential. Preprint, https://hal.archives-ouvertes.fr/hal-01543386.Google Scholar
[34] Gurtin, M. E. (1989) On a nonequilibrium thermodynamics of capillarity and phase. Quart. Appl. Math. 47 (1), 129145.Google Scholar
[35] Gurtin, M. E. (1996) Generalized Ginzburg–Landau and Cahn–Hilliard equations based on a microforce balance. Physica D 92 (3–4), 178192.Google Scholar
[36] Gurtin, M. E., Fried, E. & Anand, L. (2010) The Mechanics and Thermodynamics of Continua, Cambridge University Press, Cambridge, U.K.Google Scholar
[37] Hawkins-Daarud, A., van der Zee, K. G. & Oden, J. T. (2012) Numerical simulation of a thermodynamically consistent four-species tumor growth model. Int. J. Numer. Methods Biomed. Eng. 28 (1), 324.Google Scholar
[38] Hilhorst, D., Kampmann, J., Nguyen, T.N. & van der Zee, K. G. (2015) Formal asymptotic limit of a diffuse-interface tumor-growth model. Math. Models Methods Appl. Sci. 25 (6), 10111043.CrossRefGoogle Scholar
[39] Jiang, J., Wu, H. & Zheng, S. (2015) Global existence and asymptotic behavior of solutions to a chemotaxis-fluid system on general bounded domains. Asymptot. Anal. 92 (3–4), 249258.Google Scholar
[40] Jiang, J., Wu, H. & Zheng, S. (2015) Well-posedness and long-time behavior of a non-autonomous Cahn–Hilliard–Darcy system with mass source modeling tumor growth. J. Differ. Equ. 259 (7), 30323077.Google Scholar
[41] Keller, E. F. & Segel, L. A. (1971) Model for chemotaxis. J. Theor. Biol. 30 (2), 225234.CrossRefGoogle ScholarPubMed
[42] Li, J. & Wang, Q. (2014) A class of conservative phase field models for multiphase fluid flows. J. Appl. Mech. 81 (2), 021004.Google Scholar
[43] Liu, I.-S. (2002) Continuum Mechanics, Advanced Texts in Physics, Springer–Verlag, Berlin.Google Scholar
[44] Liu, J.-G. & Lorz, A. (2011) A coupled chemotaxis-fluid model: Global existence. Ann. Inst. H. Poincaré Anal. Non Linéaire 28 (5), 643652.Google Scholar
[45] Lorca, S. A. & Boldrini, J. L. (1999) The initial value problem for a generalized Boussinesq model. Nonlinear Anal. 36 (4), 457480.Google Scholar
[46] Lorz, A. (2010) Coupled chemotaxis-fluid model. Math. Models Methods Appl. Sci. 20 (6), 9871004.Google Scholar
[47] Lowengrub, J., Titi, E. & Zhao, K. (2013) Analysis of a mixture model of tumor growth. Eur. J. Appl. Math. 24 (5), 691734.Google Scholar
[48] Lowengrub, J. & Truskinovsky, L. (1998) Quasi-incompressible Cahn–Hilliard fluids and topological transitions. R. Soc. Lond. Proc. Ser. A: Math. Phys. Eng. Sci. 454 (1978), 26172654.Google Scholar
[49] Ockendon, H. & Ockendon, J. R. (1995) Viscous Flow, Cambridge University Press, Cambridge, U.K.Google Scholar
[50] Oden, J. T., Hawkins, A. & Prudhomme, S. (2010) General diffuse-interface theories and an approach to predictive tumor growth modeling. Math. Models Methods Appl. Sci. 58, 723763.Google Scholar
[51] Podio-Guidugli, P. (2006) Models of phase segregation and diffusion of atomic species on a lattice. RIC Mat. 55 (1), 105118.Google Scholar
[52] Ranft, J., Basan, M., Elgeti, J., Joanny, J-F., Prost, J. & Jülicher, F. (2010) Fluidization of tissues by cell division and apoptosis. Proc. Natl. Acad. Sci. USA 107 (49), 2086320868.Google Scholar
[53] Sitka, E. (2013) Modeling Tumor Growth: A Mixture Model with Mass Exchange, Master's thesis, Universität Regensburg.Google Scholar
[54] Sohr, H. (2010) The Navier–Stokes Equations: An Elementary Functional Analytic Approach, Birkhäuser Advanced Texts, Springer, Basel.Google Scholar
[55] Sun, Y.-Z. & Zhang, Z.-F. (2013) Global regularity for the initial boundary value problem of the 2-D Boussinesq system with variable viscosity and thermal diffusivity. J. Differ. Equ. 255 (6), 10691085.CrossRefGoogle Scholar
[56] Tuval, I., Cisneros, L., Dombrowski, C., Wolgemuth, C. W., Kessler, J. O. & Goldstein, R. E. (2005) Bacterial swimming and oxygen transport near contact lines. Proc. Natl. Acad. Sci. USA 102 (7), 22772282.Google Scholar
[57] Wang, X.-M. & Wu, H. (2012) Long-time behavior for the Hele–Shaw–Cahn–Hilliard system. Asymptot. Anal. 78 (4), 217245.Google Scholar
[58] Wang, X.-M. & Zhang, Z.-F. (2013) Well-posedness of the Hele–Shaw–Cahn–Hilliard system. Ann. Inst. H. Poincaré Anal. Non Linéaire 30 (3), 367384.Google Scholar
[59] Weber, J. T. (2016) Analysis of Diffuse Interface Models for Two-Phase Flows with and without Surfactants. PhD thesis, Universität Regensburg.Google Scholar
[60] Winkler, M. (2012) Global large data solutions in a chemotaxis-(Navier-)Stokes system modeling cellular swmiming in fluid drops. Commun. Partial Differ. Equ. 37 (2), 319351.Google Scholar
[61] Winkler, M. (2014) Stabilization in a two-dimensional chemotaxis-Navier–Stokes system. Arch. Ration. Mech. Anal. 211 (2), 455487.CrossRefGoogle Scholar
[62] Winkler, M. (2016) Global weak solutions in a three-dimensional chemotaxis-Navier–Stokes system. Ann. Inst. H. Poincaré Anal. Non Linéaire 33 (5), 13291352.CrossRefGoogle Scholar
[63] Wise, S. M., Lowengrub, J., Frieboes, H. B. & Cristini, V. (2008) Three-dimensional multispecies nonlinear tumor growth - I: Model and numerical method. J. Theor. Biol. 253 (3), 524543.Google Scholar
[64] Zhang, Q. & Li, Y. (2015) Convergence rates of solutions for a two-dimensional chemotaxis–Navier–Stokes system. Discrete Contin. Dyn. Syst. Ser. B 20 (8), 27512759.Google Scholar
[65] Zhang, Q. & Zheng, X. (2014) Global well-posedness for the two-dimensional incompressible chemotaxis–Navier–Stokes equations. SIAM J. Math. Anal. 46 (4), 30783105.Google Scholar
[66] Zhao, L.-Y., Wu, H. & Huang, H.-Y. (2009) Convergence to equilibrium for a phase-field model for the mixture of two incompressible fluids. Commun. Math. Sci. 7 (4), 939962.Google Scholar