Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-07-02T08:27:26.413Z Has data issue: false hasContentIssue false

Diffusive instabilities and spatial patterning from the coupling of reaction–diffusion processes with Stokes flow in complex domains

Published online by Cambridge University Press:  27 August 2019

Robert A. Van Gorder*
Affiliation:
Department of Mathematics and Statistics, University of Otago, P.O. Box 56, Dunedin 9054, New Zealand
Hyunyeon Kim
Affiliation:
Mathematical Institute, University of Oxford, Andrew Wiles Building, Radcliffe Observatory Quarter, Woodstock Road, Oxford OX2 6GG, UK
Andrew L. Krause
Affiliation:
Mathematical Institute, University of Oxford, Andrew Wiles Building, Radcliffe Observatory Quarter, Woodstock Road, Oxford OX2 6GG, UK
*
Email address for correspondence: rvangorder@maths.otago.ac.nz

Abstract

We study spatial and spatio-temporal pattern formation emergent from reaction–diffusion–advection systems formed by considering reaction–diffusion systems coupled to prescribed fluid flows. While there have been a number of studies on the planar dynamics of such systems and the resulting instabilities and spatio-temporal patterning in the plane, less has been done on complicated flows in complex domains. We consider a general approach for the study of bounded domains in order to model two- and three-dimensional geometries which are more likely to be of relevance for modelling dynamics within fluid vessels used in experiments. Considering a variety of problem geometries with finite cross-sections, such as two-dimensional channels, three-dimensional ducts and three-dimensional pipes, we demonstrate the role cross-section geometry plays in pattern formation under such systems. We find that the generic instability is that of an oscillatory or wave Turing instability, resulting in patterns which change in time, often being advected with the fluid flow. As in previous works, we observe a change in patterns formed when progressing from zero to weak to strong advection for uniform advection across the domain, with particularly strong advection destroying patterns. One novel finding is that heterogeneous fluid flow can induce qualitatively different patterns across the domain. For instance, Poiseuille flow with maximal advection in the centre of a vessel and zero advection at the boundary of a vessel is shown to exhibit patterns in the centre of the vessel which are different from patterns near the boundary, with differences attributed to the differential local advection within each region of the vessel. Additionally, we observe sheared patterns, which appear due to gradients in the fluid velocity, and cannot be obtained via any kind of uniform flow. Finally we also explore flow in more complex domains, including wavy-walled channels, continuous stirred-tank reactors, U-shaped pipes and a toroidal domain, in order to demonstrate behaviours when the flow is both heterogeneous and bidirectional, as well as to demonstrate that our results still apply for complex finite domains. Our analysis suggests that such non-trivial advection results in moving patterns which are more complex than observed in simpler reaction–diffusion–advection, and may be more characteristic of realistic flow regimes in biological media.

Type
JFM Papers
Copyright
© 2019 Cambridge University Press 

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

Abraham, E. R. 1998 The generation of plankton patchiness by turbulent stirring. Nature 391 (6667), 577.10.1038/35361Google Scholar
Abramowitz, M. & Stegun, I. A. 1965 Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables, vol. 55. Courier Corporation.Google Scholar
Almog, Y. 2008 The stability of the normal state of superconductors in the presence of electric currents. SIAM J. Math. Anal. 40 (2), 824850.10.1137/070699755Google Scholar
Andresén, P., Bache, M., Mosekilde, E., Dewel, G. & Borckmanns, P. 1999 Stationary space-periodic structures with equal diffusion coefficients. Phys. Rev. E 60 (1), 297.Google Scholar
Ayodele, S. G., Raabe, D. & Varnik, F. 2015 Shear-flow-controlled mode selection in a nonlinear autocatalytic medium. Phys. Rev. E 91 (2), 022913.Google Scholar
Baldyga, J. & Bourne, J. R. 1984 A fluid mechanical approach to turbulent mixing and chemical reaction part III computational and experimental results for the new micromixing model. Chem. Engng Commun. 28 (4–6), 259281.10.1080/00986448408940137Google Scholar
Balinsky, Y. & Pismen, L. M. 1998 Differential flow induced chemical instability and turing instability for couette flow. Phys. Rev. E 58 (4), 4524.Google Scholar
Barreira, R., Elliott, C. M. & Madzvamuse, A. 2011 The surface finite element method for pattern formation on evolving biological surfaces. J. Math. Biol. 63 (6), 10951119.10.1007/s00285-011-0401-0Google Scholar
Berenstein, I. 2012a Distinguishing similar patterns with different underlying instabilities: effect of advection on systems with Hopf, Turing-Hopf, and wave instabilities. Chaos 22 (4), 043109.10.1063/1.4766591Google Scholar
Berenstein, I. 2012b Pattern formation in a reaction-diffusion-advection system with wave instability. Chaos 22 (2), 023112.10.1063/1.4704809Google Scholar
Berger, S. A., Talbot, L. & Yao, L. S. 1983 Flow in curved pipes. Annu. Rev. Fluid Mech. 15 (1), 461512.10.1146/annurev.fl.15.010183.002333Google Scholar
Bernasconi, G. P. & Boissonade, J. 1997 Phyllotactic order induced by symmetry breaking in advected Turing patterns. Phys. Lett. A 232 (3–4), 224230.10.1016/S0375-9601(97)00361-7Google Scholar
Bois, J. S., Jülicher, F. & Grill, S. W. 2011 Pattern formation in active fluids. Phys. Rev. Lett. 106 (2), 028103.10.1103/PhysRevLett.106.028103Google Scholar
Boissonade, J., Dulos, E. & De Kepper, P. 1995 Turing patterns: from myth to reality. In Chemical Waves and Patterns, pp. 221268. Springer.10.1007/978-94-011-1156-0_7Google Scholar
Boussinesq, J. 1868 Mémoire sur l’influence des frottements dans les mouvements réguliers des fluids. J. Math. Pure Appl. 13 (2), 377424.Google Scholar
Brown, B. M., McCormack, D. K. R., Evans, W. D. & Plum, M. 1999 On the spectrum of second-order differential operators with complex coefficients. Proc. R. Soc. Lond. A 455 (1984), 12351257.10.1098/rspa.1999.0357Google Scholar
Burrows, S. M., Butler, T., Jöckel, P., Tost, H., Kerkweg, A., Pöschl, U. & Lawrence, M. G. 2009a Bacteria in the global atmosphere. Part 2. Modeling of emissions and transport between different ecosystems. Atmos. Chem. Phys. 9 (23), 92819297.10.5194/acp-9-9281-2009Google Scholar
Burrows, S. M., Elbert, W., Lawrence, M. G. & Pöschl, U. 2009b Bacteria in the global atmosphere. Part 1. Review and synthesis of literature data for different ecosystems. Atmos. Chem. Phys. 9 (23), 92639280.10.5194/acp-9-9263-2009Google Scholar
Carballido-Landeira, J., Taboada, P. & Muñuzuri, A. P. 2012 Effect of electric field on turing patterns in a microemulsion. Soft Matt. 8 (10), 29452949.Google Scholar
Cartwright, J. H. E., Piro, O. & Tuval, I. 2009 Fluid dynamics in developmental biology: moving fluids that shape ontogeny. HFSP J. 3 (2), 7793.10.2976/1.3043738Google Scholar
Chadwick, R. S. 1985 Slow viscous flow inside a torus – the resistance of small tortuous blood vessels. Q. Appl. Maths 43 (3), 317323.10.1090/qam/814230Google Scholar
Chandran, K. B. & Yearwood, T. L. 1981 Experimental study of physiological pulsatile flow in a curved tube. J. Fluid Mech. 111, 5985.10.1017/S0022112081002292Google Scholar
Chavel, I. 1984 Eigenvalues in Riemannian Geometry, vol. 115. Academic Press.Google Scholar
Chen, Z.-M. & Price, W. G. 1996 Time dependent periodic Navier–Stokes flows on a two-dimensional torus. Commun. Math. Phys. 179 (3), 577597.10.1007/BF02100098Google Scholar
Crampin, E. J., Gaffney, E. A. & Maini, P. K. 1999 Reaction and diffusion on growing domains: scenarios for robust pattern formation. Bull. Math. Biol. 61 (6), 10931120.10.1006/bulm.1999.0131Google Scholar
Dähmlow, P., Luengviriya, C. & Müller, S. C. 2015 Electric field effects in chemical patterns. In Bottom-Up Self-Organization in Supramolecular Soft Matter, pp. 6582. Springer.10.1007/978-3-319-19410-3_3Google Scholar
Dähmlow, P. & Müller, S. C. 2015 Nonlinear effects of electric fields in the Belousov–Zhabotinsky reaction dissolved in a microemulsion. Chaos 25 (4), 043117.10.1063/1.4919217Google Scholar
Dawson, S. P., Lawniczak, A. & Kapral, R. 1994 Interaction of turing and flow-induced chemical instabilities. J. Chem. Phys. 100 (7), 52115218.10.1063/1.467185Google Scholar
De Kepper, P., Castets, V., Dulos, E. & Boissonade, J. 1991 Turing-type chemical patterns in the chlorite-iodide-malonic acid reaction. Physica D 49 (1–2), 161169.10.1016/0167-2789(91)90204-MGoogle Scholar
Dehghan, M. & Abbaszadeh, M. 2016 Numerical study of three-dimensional turing patterns using a meshless method based on moving Kriging element free galerkin (EFG) approach. Comput. Math Appl. 72 (3), 427454.10.1016/j.camwa.2016.04.038Google Scholar
Del Pino, M., Felmer, P. & Kowalczyk, M. 2002 Boundary spikes in the Gierer–Meinhardt system. Commun. Pure Appl. Anal. 1 (4), 437456.Google Scholar
Dickinson, E. J. F., Ekström, H. & Fontes, E. 2014 Comsol multiphysics®: finite element software for electrochemical analysis. A mini-review. Electrochem. Commun. 40, 7174.10.1016/j.elecom.2013.12.020Google Scholar
Eckstein, T., Vidal-Henriquez, E., Bae, A., Zykov, V., Bodenschatz, E. & Gholami, A. 2018 Influence of fast advective flows on pattern formation of dictyostelium discoideum. PloS One 13 (3), e0194859.10.1371/journal.pone.0194859Google Scholar
Edmunds, D. E., Evans, W. D. & Fleckinger, J. 1983 On the spectrum and the distribution of singular values of Schrödinger operators with a complex potential. Proc. R. Soc. Lond. A 388 (1794), 195218.10.1098/rspa.1983.0078Google Scholar
Edwards, B. F. 2002 Poiseuille advection of chemical reaction fronts. Phys. Rev. Lett. 89 (10), 104501.10.1103/PhysRevLett.89.104501Google Scholar
Elezgaray, J. & Arneodo, A. 1991 Modeling reaction–diffusion pattern formation in the Couette flow reactor. J. Chem. Phys. 95 (1), 323350.10.1063/1.461823Google Scholar
Epstein, I. R. & Lengyel, I. 1995 Turing structures. Progress toward a room temperature, closed system. Physica D 84 (1–2), 111.10.1016/0167-2789(95)00003-MGoogle Scholar
Ermakova, E. A., Shnol, E. E., Panteleev, M. A., Butylin, A. A., Volpert, V. & Ataullakhanov, F. I. 2009 On propagation of excitation waves in moving media: the Fitzhugh–Nagumo model. PloS One 4 (2), e4454.10.1371/journal.pone.0004454Google Scholar
Ermentrout, B. 1991 Stripes or spots? Nonlinear effects in bifurcation of reaction–diffusion equations on the square. Proc. R. Soc. Lond. A 434 (1891), 413417.10.1098/rspa.1991.0100Google Scholar
Evans, J. W. 1975 Nerve axon equations: IV the stable and the unstable impulse. Indiana Univ. Math. J. 24 (12), 11691190.10.1512/iumj.1975.24.24096Google Scholar
Evans, W. D. 1981 On the spectra of Schrödinger operators with a complex potential. Math. Ann. 255 (1), 5776.Google Scholar
Faivre, D. & Schuler, D. 2008 Magnetotactic bacteria and magnetosomes. Chem. Rev. 108 (11), 48754898.10.1021/cr078258wGoogle Scholar
Flach, E. H., Schnell, S. & Norbury, J. 2007 Limit cycles in the presence of convection: a traveling wave analysis. Phys. Rev. E 76 (3), 036216.Google Scholar
Ghosh, S., Paul, S. & Ray, D. S. 2016 Differential-flow-induced transition of traveling wave patterns and wave splitting. Phys. Rev. E 94 (4), 042223.Google Scholar
Golovin, A. A., Matkowsky, B. J. & Volpert, V. A. 2008 Turing pattern formation in the Brusselator model with superdiffusion. SIAM J. Appl. Maths 69 (1), 251272.10.1137/070703454Google Scholar
Greenberg, L. & Marletta, M. 2001 Numerical solution of non–self-adjoint Sturm–Liouville problems and related systems. SIAM J. Numer. Anal. 38 (6), 18001845.10.1137/S0036142999358743Google Scholar
Grindrod, P. 1988 Models of individual aggregation or clustering in single and multi-species communities. J. Math. Biol. 26 (6), 651660.10.1007/BF00276146Google Scholar
Happel, J. & Brenner, H. 2012 Low Reynolds Number Hydrodynamics: With Special Applications to Particulate Media, vol. 1. Springer Science & Business Media.Google Scholar
Helffer, B. 2011 On pseudo-spectral problems related to a time-dependent model in superconductivity with electric current. Confluentes Mathematici 3 (02), 237251.10.1142/S1793744211000308Google Scholar
Hillen, T. & Painter, K. J. 2009 A users guide to PDE models for chemotaxis. J. Math. Biol. 58 (1–2), 183.10.1007/s00285-008-0201-3Google Scholar
Horstmann, D. 2011 Generalizing the Keller–Segel model: Lyapunov functionals, steady state analysis, and blow-up results for multi-species chemotaxis models in the presence of attraction and repulsion between competitive interacting species. J. Nonlinear Sci. 21 (2), 231270.10.1007/s00332-010-9082-xGoogle Scholar
Hu, H. X., Li, X. C. & Li, Q. S. 2009 Flow-induced symmetry reduction in two-dimensional reaction–diffusion system. Chem. Phys. 358 (1–2), 2124.10.1016/j.chemphys.2008.11.026Google Scholar
Jönsson, H., Heisler, M. G., Shapiro, B. E., Meyerowitz, E. M. & Mjolsness, E. 2006 An auxin-driven polarized transport model for phyllotaxis. Proc. Natl Acad. Sci. USA 103 (5), 16331638.10.1073/pnas.0509839103Google Scholar
Kærn, M. & Menzinger, M. 1999 Flow-distributed oscillations: stationary chemical waves in a reacting flow. Phys. Rev. E 60 (4), R3471.Google Scholar
Kærn, M. & Menzinger, M. 2000 Pulsating wave propagation in reactive flows: flow-distributed oscillations. Phys. Rev. E 61 (4), 3334.Google Scholar
Kærn, M. & Menzinger, M. 2002 Experiments on flow-distributed oscillations in the Belousov–Zhabotinsky reaction. J. Phys. Chem. A 106 (19), 48974903.Google Scholar
Kapitula, T. & Sandstede, B. 2004 Eigenvalues and resonances using the Evans function. J. Discrete Continuous Dyn. Syst. 10, 857870.10.3934/dcds.2004.10.857Google Scholar
Kellogg, C. A. & Griffin, D. W. 2006 Aerobiology and the global transport of desert dust. Trends Ecology Evolution 21 (11), 638644.10.1016/j.tree.2006.07.004Google Scholar
Khuri, S. A. & Wazwaz, A. M. 1997 On the solution of a partial differential equation arising in Stokes flow. Appl. Math. Comput. 85 (2–3), 139147.Google Scholar
Kim, T. & Lin, M. 2007 Stable advection-reaction-diffusion with arbitrary anisotropy. Computer Animation Virtual Worlds 18 (4–5), 329338.10.1002/cav.187Google Scholar
Klika, V., Kozaák, M. & Gaffney, E. A. 2018 Domain size driven instability: self-organization in systems with advection. SIAM J. Appl. Maths 78 (5), 22982322.10.1137/17M1138571Google Scholar
Komori, S., Hunt, J. C. R., Kanzaki, T. & Murakami, Y. 1991 The effects of turbulent mixing on the correlation between two species and on concentration fluctuations in non-premixed reacting flows. J. Fluid Mech. 228, 629659.Google Scholar
Kondo, S. & Miura, T. 2010 Reaction-diffusion model as a framework for understanding biological pattern formation. Science 329 (5999), 16161620.10.1126/science.1179047Google Scholar
Koptyug, I. V., Zhivonitko, V. V. & Sagdeev, R. Z. 2008 Advection of chemical reaction fronts in a porous medium. J. Phys. Chem. B 112 (4), 11701176.10.1021/jp077612rGoogle Scholar
Krause, A. L., Burton, A. M., Fadai, N. T. & Van Gorder, R. A. 2018 Emergent structures in reaction-advection-diffusion systems on a sphere. Phys. Rev. E 97 (4), 042215.Google Scholar
Krause, A. L., Ellis, M. A. & Van Gorder, R. A. 2019 Influence of curvature, growth, and anisotropy on the evolution of turing patterns on growing manifolds. Bull. Math. Biol. 81 (3), 759799.10.1007/s11538-018-0535-yGoogle Scholar
Kuptsov, P. V., Satnoianu, R. A. & Daniels, P. G. 2005 Pattern formation in a two-dimensional reaction-diffusion channel with poiseuille flow. Phys. Rev. E 72 (3), 036216.Google Scholar
Kurowski, L., Krause, A. L., Mizuguchi, H., Grindrod, P. & Van Gorder, R. A. 2017 Two-species migration and clustering in two-dimensional domains. Bull. Math. Biol. 79, 23022333.10.1007/s11538-017-0331-0Google Scholar
Ledesma-Durán, A., Juárez-Valencia, L.-H., Morales-Malacara, J.-B. & Santamaría-Holek, I. 2017 The interplay between phenotypic and ontogenetic plasticities can be assessed using reaction-diffusion models. J. Biol. Phys. 43 (2), 247264.10.1007/s10867-017-9450-yGoogle Scholar
Liu, Y., Zhou, W., Zheng, T., Zhao, Y., Gao, Q., Pan, C. & Horvaáth, A. K. 2016 Convection-induced fingering fronts in the chlorite–trithionate reaction. J. Phys. Chem. A 120 (16), 25142520.10.1021/acs.jpca.6b01192Google Scholar
Lu, L.-H., Ryu, K. S. & Liu, C. 2002 A magnetic microstirrer and array for microfluidic mixing. J. Microelectromech. Syst. 11 (5), 462469.Google Scholar
Lyne, W. H. 1971 Unsteady viscous flow in a curved pipe. J. Fluid Mech. 45 (1), 1331.10.1017/S0022112071002970Google Scholar
Malchow, H. 1995 Flow-and locomotion-induced pattern formation in nonlinear population dynamics. Ecol. Model. 82 (3), 257264.10.1016/0304-3800(94)00095-YGoogle Scholar
Malchow, H. 2000 Motional instabilities in prey–predator systems. J. Theor. Biol. 204 (4), 639647.10.1006/jtbi.2000.2074Google Scholar
Marcos, Fu, H. C., Powers, T. R. & Stocker, R. 2012 Bacterial rheotaxis. Proc. Natl Acad. Sci. USA 109 (13), 47804785.10.1073/pnas.1120955109Google Scholar
Marlow, M., Sasaki, Y. & Vasquez, D. A. 1997 Spatiotemporal behavior of convective turing patterns in porous media. J. Chem. Phys. 107 (13), 52055211.10.1063/1.474883Google Scholar
McGraw, P. N. & Menzinger, M. 2003 General theory of nonlinear flow-distributed oscillations. Phys. Rev. E 68 (6), 066122.Google Scholar
Míguez, D. G., Izús, G. G. & Muñuzuri, A. P. 2006a Robustness and stability of flow-and-diffusion structures. Phys. Rev. E 73 (1), 016207.Google Scholar
Míguez, D. G., Satnoianu, R. A. & Muñuzuri, A. P. 2006b Experimental steady pattern formation in reaction-diffusion-advection systems. Phys. Rev. E 73 (2), 025201.Google Scholar
Müller, P., Rogers, K. W., Jordan, B. M., Lee, J. S., Robson, D., Ramanathan, S. & Schier, A. F. 2012 Differential diffusivity of nodal and lefty underlies a reaction–diffusion patterning system. Science 336 (6082), 721724.10.1126/science.1221920Google Scholar
Murray, J. D. 2001 Mathematical Biology. II Spatial Models and Biomedical Applications. Interdisciplinary Applied Mathematics V. 18. Springer.Google Scholar
Nakagaki, T., Yamada, H. & Ito, M. 1999 Reaction–diffusion–advection model for pattern formation of rhythmic contraction in a giant amoeboid cell of thephysarumplasmodium. J. Theor. Biol. 197 (4), 497506.10.1006/jtbi.1998.0890Google Scholar
Nenniger, J. E., Kridiotis, A., Chomiak, J., Longwell, J. P. & Sarofim, A. F. 1985 Characterization of a toroidal well stirred reactor. In Symposium (International) on Combustion, vol. 20, pp. 473479. Elsevier.Google Scholar
Nugent, C. R., Quarles, W. M. & Solomon, T. H. 2004 Experimental studies of pattern formation in a reaction-advection-diffusion system. Phys. Rev. Lett. 93 (21), 218301.10.1103/PhysRevLett.93.218301Google Scholar
Olver, F. W. J., Lozier, D. W., Boisvert, R. F. & Clark, C. W. 2010 NIST Handbook of Mathematical Functions Hardback and CD-ROM. Cambridge University Press.Google Scholar
Perumpanani, A. J., Sherratt, J. A. & Maini, P. K. 1995 Phase differences in reaction-diffusion-advection systems and applications to morphogenesis. IMA J. Appl. Maths 55 (1), 1933.10.1093/imamat/55.1.19Google Scholar
Pierce, C. D. & Moin, P. 2004 Progress-variable approach for large-eddy simulation of non-premixed turbulent combustion. J. Fluid Mech. 504, 7397.10.1017/S0022112004008213Google Scholar
Plaza, R. G., Sanchez-Garduno, F., Padilla, P., Barrio, R. A. & Maini, P. K. 2004 The effect of growth and curvature on pattern formation. J. Dynam. Diff. Equ. 16 (4), 10931121.10.1007/s10884-004-7834-8Google Scholar
Pozrikidis, C. 1987 Creeping flow in two-dimensional channels. J. Fluid Mech. 180, 495514.10.1017/S0022112087001927Google Scholar
Pozrikidis, C. 1988 The flow of a liquid film along a periodic wall. J. Fluid Mech. 188, 275300.10.1017/S0022112088000734Google Scholar
Pudykiewicz, J. A. 2006 Numerical solution of the reaction–advection–diffusion equation on the sphere. J. Comput. Phys. 213 (1), 358390.10.1016/j.jcp.2005.08.021Google Scholar
Qi, J., Zheng, Z. & Sun, H. 2011 Classification of Sturm–Liouville differential equations with complex coefficients and operator realizations. Proc. R. Soc. Lond. A 467 (2131), 18351850.10.1098/rspa.2010.0281Google Scholar
Radszuweit, M., Alonso, S., Engel, H. & Bär, M. 2013 Intracellular mechanochemical waves in an active poroelastic model. Phys. Rev. Lett. 110 (13), 138102.10.1103/PhysRevLett.110.138102Google Scholar
Ralph, M. E. 1986 Oscillatory flows in wavy-walled tubes. J. Fluid Mech. 168, 515540.10.1017/S0022112086000496Google Scholar
Reigada, R., Hillary, R. M., Bees, M. A., Sancho, J. M. & Sagués, F. 2003 Plankton blooms induced by turbulent flows. Proc. R. Soc. Lond. B 270 (1517), 875880.10.1098/rspb.2002.2298Google Scholar
Riaz, S. S., Kar, S. & Ray, D. S. 2004 Mobility-induced instability and pattern formation in a reaction-diffusion system. J. Chem. Phys. 121 (11), 53955399.10.1063/1.1783275Google Scholar
Rickett, L. M., Penfold, R., Blyth, M. G., Purvis, R. & Cooker, M. J. 2015 Incipient mixing by marangoni effects in slow viscous flow of two immiscible fluid layers. IMA J. Appl. Maths 80 (5), 15821618.10.1093/imamat/hxv009Google Scholar
Roberts, E. P. L. & Mackley, M. R. 1996 The development of asymmetry and period doubling for oscillatory flow in baffled channels. J. Fluid Mech. 328, 1948.10.1017/S0022112096008634Google Scholar
Rovinsky, A. B. & Menzinger, M. 1992 Chemical instability induced by a differential flow. Phys. Rev. Lett. 69 (8), 1193.10.1103/PhysRevLett.69.1193Google Scholar
Rovinsky, A. B. & Menzinger, M. 1993 Self-organization induced by the differential flow of activator and inhibitor. Phys. Rev. Lett. 70 (6), 778.10.1103/PhysRevLett.70.778Google Scholar
Rush, T. A., Newell, T. A. & Jacobi, A. M. 1999 An experimental study of flow and heat transfer in sinusoidal wavy passages. Intl J. Heat Mass Transfer 42 (9), 15411553.10.1016/S0017-9310(98)00264-6Google Scholar
Sánchez-Garduño, F., Krause, A. L., Castillo, J. A. & Padilla, P.2018 Turing–Hopf patterns on growing domains: the torus and the sphere. J. Theor. Biol. doi:10.1016/j.jtbi.2018.09.028.Google Scholar
Satnoianu, R. A. 2003 Coexistence of stationary and traveling waves in reaction–diffusion–advection systems. Phys. Rev. E 68 (3), 032101.Google Scholar
Satnoianu, R. A., Maini, P. K. & Menzinger, M. 2001 Parameter space analysis, pattern sensitivity and model comparison for Turing and stationary flow-distributed waves (FDS). Physica D 160 (1–2), 79102.Google Scholar
Satnoianu, R. A. & Menzinger, M. 2002 A general mechanism for ‘inexact’ phase differences in reaction–diffusion–advection systems. Phys. Lett. A 304 (5), 149156.10.1016/S0375-9601(02)01387-7Google Scholar
Schnakenberg, J. 1979 Simple chemical reaction systems with limit cycle behaviour. J. Theor. Biol. 81 (3), 389400.10.1016/0022-5193(79)90042-0Google Scholar
Selvarajan, S., Tulapurkara, E. G. & Ram, V. V. 1999 Stability characteristics of wavy walled channel flows. Phys. Fluids 11 (3), 579589.10.1063/1.869946Google Scholar
Şeref, F. & Veliev, O. A. 2014 On non-self-adjoint sturm–liouville operators in the space of vector functions. Math. Notes 95 (1–2), 180190.10.1134/S0001434614010192Google Scholar
Sherratt, J. A. 2011 Pattern solutions of the Klausmeier model for banded vegetation in semi-arid environments II: patterns with the largest possible propagation speeds. Proc. Math. Phys. Engng Sci. 467 (2135), 32723294.10.1098/rspa.2011.0194Google Scholar
Shoji, H., Iwasa, Y. & Kondo, S. 2003 Stripes, spots, or reversed spots in two-dimensional Turing systems. J. Theor. Biol. 224 (3), 339350.10.1016/S0022-5193(03)00170-XGoogle Scholar
Shoji, H., Yamada, K., Ueyama, D. & Ohta, T. 2007 Turing patterns in three dimensions. Phys. Rev. E 75 (4), 046212.Google Scholar
Siero, E., Doelman, S., Eppinga, M. B., Rademacher, J. D. M., Rietkerk, M. & Siteur, K. 2015 Striped pattern selection by advective reaction–diffusion systems: resilience of banded vegetation on slopes. Chaos 25 (3), 036411.10.1063/1.4914450Google Scholar
Sloman, B. M., Please, C. P. & Van Gorder, R. A. 2018 Asymptotic analysis of a silicon furnace model. SIAM J. Appl. Maths 78 (2), 11741205.10.1137/17M1122335Google Scholar
Smith, R. S., Guyomarc’h, S., Mandel, T., Reinhardt, D., Kuhlemeier, C. & Prusinkiewicz, P. 2006 A plausible model of phyllotaxis. Proc. Natl Acad. Sci. USA 103 (5), 13011306.10.1073/pnas.0510457103Google Scholar
Smith, S. & Dalchau, N. 2018 Model reduction enables turing instability analysis of large reaction–diffusion models. J. R. Soc. Interface 15 (140), 20170805.10.1098/rsif.2017.0805Google Scholar
Smolarkiewicz, P. K. & Rasch, P. J. 1991 Monotone advection on the sphere: an eulerian versus semi-lagrangian approach. J. Atmos. Sci. 48 (6), 793810.10.1175/1520-0469(1991)048<0793:MAOTSA>2.0.CO;22.0.CO;2>Google Scholar
Sobey, I. J. 1985 Dispersion caused by separation during oscillatory flow through a furrowed channel. Chem. Engng Sci. 40 (11), 21292134.10.1016/0009-2509(85)87031-7Google Scholar
Spangler, R. S. & Edwards, B. F. 2003 Poiseuille advection of chemical reaction fronts: Eikonal approximation. J. Chem. Phys. 118 (13), 59115915.10.1063/1.1553752Google Scholar
Straube, A. V. & Pikovsky, A. 2011 Pattern formation induced by time-dependent advection. Math. Model. Natural Phenom. 6 (1), 138148.10.1051/mmnp/20116107Google Scholar
Stucchi, L. & Vasquez, D. A. 2014 Pattern formation induced by a differential poiseuille flow. Eur. Phys. J. Special Topics 223 (13), 30113020.10.1140/epjst/e2014-02314-8Google Scholar
Stucchi, L. & Vasquez, D. A. 2013 Pattern formation induced by a differential shear flow. Phys. Rev. E 87 (2), 024902.Google Scholar
Tan, Z., Chen, S., Peng, X., Zhang, L. & Gao, C. 2018 Polyamide membranes with nanoscale turing structures for water purification. Science 360 (6388), 518521.10.1126/science.aar6308Google Scholar
Tse, W. H., Wei, J. & Winter, M. 2010 The Gierer–Meinhardt system on a compact two-dimensional Riemannian manifold: interaction of Gaussian curvature and Green’s function. J. Math. Pures Appl. 94 (4), 366397.10.1016/j.matpur.2010.03.003Google Scholar
Turing, A. M. 1952 The chemical basis of morphogenesis. Phil. Trans. R. Soc. Lond. B 237 (641), 3772.Google Scholar
Uzunca, M., Küçükseyhan, T., Yücel, H. & Karasözen, B. 2017 Optimal control of convective FitzHugh–Nagumo equation. Comput. Maths Applics. 73 (9), 21512169.10.1016/j.camwa.2017.02.028Google Scholar
Vadivelu, R. K., Kamble, H., Munaz, A. & Nguyen, N.-T. 2017 Liquid marble as bioreactor for engineering three-dimensional toroid tissues. Sci. Rep. 7 (1), 12388.Google Scholar
Vasquez, D. A. 2004 Chemical instability induced by a shear flow. Phys. Rev. Lett. 93 (10), 104501.10.1103/PhysRevLett.93.104501Google Scholar
Vasquez, D. A. & Coroian, D. I. 2010 Stability of convective patterns in reaction fronts: a comparison of three models. Chaos 20 (3), 033109.10.1063/1.3467858Google Scholar
Vasquez, D. A., Meyer, J. & Suedhoff, H. 2008 Chemical pattern formation induced by a shear flow in a two-layer model. Phys. Rev. E 78 (3), 036109.Google Scholar
Veliev, O. A. 2007 Non-self-adjoint sturm-liouville operators with matrix potentials. Math. Notes 81 (3–4), 440448.10.1134/S0001434607030200Google Scholar
Veynante, D. & Vervisch, L. 2002 Turbulent combustion modeling. Prog. Energy Combust. Sci. 28 (3), 193266.10.1016/S0360-1285(01)00017-XGoogle Scholar
Vidal-Henriquez, E., Zykov, V., Bodenschatz, E. & Gholami, A. 2017 Convective instability and boundary driven oscillations in a reaction–diffusion-advection model. Chaos 27 (10), 103110.10.1063/1.4986153Google Scholar
Vilar, J. M. G., Solé, R. V. & Rubı, J. M. 2003 On the origin of plankton patchiness. Physica A 317 (1–2), 239246.10.1016/S0378-4371(02)01322-5Google Scholar
Vilela, P. M. & Vasquez, D. A. 2014 Stability of Kuramoto-Sivashinsky fronts in moving fluid. Eur. Phys. J. Special Topics 223 (13), 30013010.10.1140/epjst/e2014-02313-9Google Scholar
Wang, C. Y. 2006 Stokes flow through a tube with bumpy wall. Phys. Fluids 18 (7), 078101.10.1063/1.2214883Google Scholar
Wei, H. H., Waters, S. L., Liu, S. Q. & Grotberg, J. B. 2003 Flow in a wavy-walled channel lined with a poroelastic layer. J. Fluid Mech. 492, 2345.10.1017/S0022112003005378Google Scholar
Wei, J. & Winter, M. 2012 Flow-distributed spikes for schnakenberg kinetics. J. Math. Biol. 64 (1–2), 211254.10.1007/s00285-011-0412-xGoogle Scholar
Williamson, D. L. & Rasch, P. J. 1989 Two-dimensional semi-Lagrangian transport with shape-preserving interpolation. Mon. Weath. Rev. 117 (1), 102129.10.1175/1520-0493(1989)117<0102:TDSLTW>2.0.CO;22.0.CO;2>Google Scholar
Yamada, H., Nakagaki, T. & Ito, M. 1999 Pattern formation of a reaction-diffusion system with self-consistent flow in the amoeboid organism physarum plasmodium. Phys. Rev. E 59 (1), 1009.Google Scholar
Yang, L., Dolnik, M., Zhabotinsky, A. M. & Epstein, I. R. 2002 Pattern formation arising from interactions between Turing and wave instabilities. J. Chem. Phys. 117 (15), 72597265.10.1063/1.1507110Google Scholar
Yang, L. & Epstein, I. R. 2003 Oscillatory Turing patterns in reaction–diffusion systems with two coupled layers. Phys. Rev. Lett. 90 (17), 178303.10.1103/PhysRevLett.90.178303Google Scholar
Yochelis, A., Bar-On, T. & Gov, N. S. 2016 Reaction–diffusion–advection approach to spatially localized treadmilling aggregates of molecular motors. Physica D 318, 8490.10.1016/j.physd.2015.10.023Google Scholar
Yochelis, A. & Sheintuch, M. 2009a Principal bifurcations and symmetries in the emergence of reaction–diffusion–advection patterns on finite domains. Phys. Rev. E 80 (5), 056201.Google Scholar
Yochelis, A. & Sheintuch, M. 2009b Towards nonlinear selection of reaction–diffusion patterns in presence of advection: a spatial dynamics approach. Phys. Chem. Chemical Phys. 11 (40), 92109223.10.1039/b903266eGoogle Scholar
Yochelis, A. & Sheintuch, M. 2010 Drifting solitary waves in a reaction–diffusion medium with differential advection. Phys. Rev. E 81 (2), 025203.Google Scholar
Yurko, V. 2006 Inverse problems for matrix Sturm–Liouville operators. Russian J. Math. Phys. 13 (1), 111118.10.1134/S1061920806010110Google Scholar
Zhou, P., Wang, J., Li, X. & Jin, Z. 2009 Emergence of traveling pattern in a predator–prey system. Intl J. Mod. Phys. C 20 (11), 18611870.10.1142/S0129183109014783Google Scholar