Skip to main content
Log in

Can large inhomogeneities generate target patterns?

  • Published:
Zeitschrift für angewandte Mathematik und Physik Aims and scope Submit manuscript

Abstract

We study the existence of target patterns in oscillatory media with weak local coupling and in the presence of an impurity, or defect. We model these systems using a viscous eikonal equation posed on the plane and represent the defect as a perturbation. In contrast to previous results, we consider large defects, which we describe using a function with slow algebraic decay, i.e., \(g \sim \textrm{O}(1/|x|^m)\) for \(m \in (1,2]\). We prove that these defects are able to generate target patterns and that, just as in the case of strongly localized impurities, their frequency is small beyond all orders of the small parameter describing their strength. Our analysis consists of finding two approximations to target pattern solutions, one which is valid at intermediate scales and a second one which is valid in the far field. This is done using weighted Sobolev spaces, which allow us to recover Fredholm properties of the relevant linear operators, as well as the implicit function theorem, which is then used to prove existence. By matching the intermediate and far-field approximations, we then determine the frequency of the pattern that is selected by the system.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

References

  1. Abramowitz, M., Stegun, I.A., Romer, R.H.: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables (1988)

  2. Adams, R.A., Fournier, J.J.F.: Sobolev Spaces. Elsevier (2003)

  3. Alcantara, F., Monk, M.: Signal propagation during aggregation in the slime mould dictyostelium discoideum. Microbiology 85(2), 321–334 (1974)

    Google Scholar 

  4. Doelman, A., Sandstede, B., Scheel, A., Schneider, G.: The Dynamics of Modulated Wave Trains. American Mathematical Society, Philadelphia (2009)

  5. Durston, A.J.: Pacemaker activity during aggregation in dictyostelium discoideum. Dev. Biol. 37(2), 225–235 (1974)

    Article  Google Scholar 

  6. Folland, G.B.: Real Analysis: Modern Techniques and Their Applications, vol. 40. Wiley, New York (1999)

    MATH  Google Scholar 

  7. Hagan, P.S.: Target patterns in reaction–diffusion systems. Adv. Appl. Math. 2(4), 400–416 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  8. Hendrey, M., Nam, K., Guzdar, P., Ott, E.: Target waves in the complex Ginzburg–Landau equation. Phys. Rev. E 62, 7627–7631 (2000)

    Article  Google Scholar 

  9. Jaramillo, G.: Inhomogeneities in 3 dimensional oscillatory media. Netw. Heterog. Media 10(2), 387–399 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  10. Jaramillo, G.: Rotating spirals in oscillatory media with nonlocal interactions and their normal form. Discrete Continu. Dyn. Syst. 6, 66 (2022)

    MATH  Google Scholar 

  11. Jaramillo, G., Scheel, A.: Pacemakers in large arrays of oscillators with nonlocal coupling. J. Differ. Equ. 260(3), 2060–2090 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  12. Jaramillo, G., Scheel, A., Qiliang, W.: The effect of impurities on striped phases. Proc. R. Soc. Edinb. Sect. A Math. 149(1), 131–168 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  13. Jaramillo, G., Venkataramani, S.C.: Target patterns in a 2d array of oscillators with nonlocal coupling. Nonlinearity 31(9), 4162–4201 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  14. Kassam, A.-K., Trefethen, L.N.: Fourth-order time-stepping for stiff PDEs. SIAM J. Sci. Comput. 26(4), 1214–1233 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  15. Kollár, R., Scheel, A.: Coherent structures generated by inhomogeneities in oscillatory media. SIAM J. Appl. Dyn. Syst. 6(1), 236–262 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  16. Kopell, N., Howard, L.N.: Target pattern and spiral solutions to reaction–diffusion equations with more than one space dimension. Adv. Appl. Math. 2(4), 417–449 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  17. Kopell, N.: Target pattern solutions to reaction–diffusion equations in the presence of impurities. Adv. Appl. Math. 2(4), 389–399 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  18. Kuramoto, Y.: Chemical Oscillations, Waves, and Turbulence. Courier Corporation (2003)

  19. McOwen, R.C.: The behavior of the Laplacian on weighted Sobolev spaces. Commun. Pure Appl. Math. 32(6), 783–795 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  20. Scheel, A.: Bifurcation to spiral waves in reaction–diffusion systems. SIAM J. Math. Anal. 29(6), 1399–1418 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  21. Simon, B.: The bound state of weakly coupled Schrödinger operators in one and two dimensions. Ann. Phys. 97(2), 279–288 (1976)

    Article  MATH  Google Scholar 

  22. Stich, M., Mikhailov, A.S.: Complex pacemakers and wave sinks in eterogeneous oscillatory. Chem. Syst. 216(4), 521 (2002)

    Google Scholar 

  23. Stich, M., Mikhailov, A.S.: Target patterns in two-dimensional heterogeneous oscillatory reaction–diffusion systems. Phys. D Nonlinear Phenom. 215(1), 38–45 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  24. Townsend, R.G., Solomon, S.S., Chen, S.C., Pietersen, A.N.J., Martin, P.R., Solomon, S.G., Gong, P.: Emergence of complex wave patterns in primate cerebral cortex. J. Neurosci. 35(11), 4657–4662 (2015)

    Article  Google Scholar 

  25. Tyson, J.J., Fife, P.C.: Target patterns in a realistic model of the Belousov–Zhabotinskii reaction. J. Chem. Phys. 73(5), 2224–2237 (1980)

    Article  MathSciNet  Google Scholar 

  26. Wolff, J., Stich, M., Beta, C., Rotermund, H.H.: Laser-induced target patterns in the oscillatory CO oxidation on Pt(110). J. Phys. Chem. B 108(38), 14282–14291 (2004). (09)

    Article  Google Scholar 

  27. Zaikin, A.N., Zhabotinsky, A.M.: Concentration wave propagation in two-dimensional liquid-phase self-oscillating system. Nature 225(5232), 535–537 (1970)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gabriela Jaramillo.

Ethics declarations

Conflict of interest

The author declares that she has no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This work is supported by NSF DMS-1911742.

Appendix

Appendix

In [10], it was shown that the following amplitude equation governs the dynamics of one-armed spiral waves in nonlocal oscillatory media,

$$\begin{aligned} 0 = \beta \Delta _1 w + \lambda w + \alpha |w|^2w + N(w,\varepsilon ), \quad r\in [0,\infty ).\end{aligned}$$

Here, w is a radial and complex-valued function, and

$$\begin{aligned} \beta = ( \sigma - \varepsilon \lambda ), \quad \lambda , \alpha \in \mathbb {C},\quad N \sim \textrm{O}(|\varepsilon ||w|^4). \end{aligned}$$

It was also established in [10] that the constant \(\lambda _I\) is an unknown parameter that needs to be determined when solving the equation.

In this section, a multiple-scale analysis is used to derive a steady-state viscous eikonal equation from the above expression. We will see that this eikonal equation is of the form considered in this paper and that it involves an inhomogeneity that decays at order \(\textrm{O}(1/|x|^2)\).

To accomplish this task, we first let \(w = A {\tilde{w}}\), with \(A^2 =- \lambda _R/\alpha _R\). This change of variables is done for convenience and leads to the following equation,

$$\begin{aligned} 0 = \beta \Delta _1 {\tilde{w}} + \lambda {\tilde{w}} + (-\lambda _R + \textrm{i}{\tilde{\alpha }}_I) |{\tilde{w}}|^2{\tilde{w}} + N({\tilde{w}},\varepsilon ), \qquad {\tilde{\alpha }}_I = - \alpha \lambda _R/\alpha _R.\end{aligned}$$

Letting \({\tilde{w}} = \rho \textrm{e}^{\textrm{i}\phi }\) and separating the real and imaginary parts of the above expression, one finally obtains the system

$$\begin{aligned} 0= & {} \beta _R \left[ \Delta _1 \rho - (\partial _r \phi )^2 \rho \right] - \beta _I \left[ \Delta _0 \phi \rho + 2\partial _r \phi \partial _r \rho \right] + \lambda _R \rho - \lambda _R \rho ^3 + \textrm{Re}\left[ N({\tilde{w}}; \varepsilon ) \textrm{e}^{-\textrm{i}\phi } \right] \end{aligned}$$
(20)
$$\begin{aligned} 0= & {} \beta _R \left[ \Delta _0 \phi \rho + 2\partial _r \phi \partial _r \rho \right] + \beta _I \left[ \Delta _1 \rho - (\partial \phi )^2 \rho \right] + \lambda _I \rho + {\tilde{\alpha }}_I \rho ^3 + \textrm{Im}\left[ N({\tilde{w}}; \varepsilon ) \textrm{e}^{-\textrm{i}\phi } \right] . \end{aligned}$$
(21)

Next, we proceed with a perturbation analysis following [4]. We rescale the variable r by defining \(S = \delta r\), where \(\delta \) is assumed to be a small positive parameter. We also use the following expressions for the unknown functions:

$$\begin{aligned}\begin{array}{r l c l } \rho = &{} \rho _0 + \delta ^2 (R_0 + \delta R_1), &{} \rho _0= \rho _0(r),&{} R_i = R_i(\delta r) \quad i=0,1\\ \phi = &{} \phi _0 + \delta \phi _1, &{} &{} \phi _i = \phi _i(\delta r) i =0,1. \end{array}\end{aligned}$$

And for the parameter, we choose \( \lambda _I = -{\tilde{\alpha }}_I + \delta ^2 {\tilde{\lambda }}_I,\) with \({\tilde{\alpha }}\) as above and \({\tilde{\lambda }}_I\) a free parameter.

Inserting the above ansatz into Eqs. (20) and (21), we obtain a set of equations in powers of \(\delta \). To write this equations more compactly, we use the subscript S to distinguish operators that are applied to functions that depend on this variable, i.e., \(\Delta _{0,S}\). The absence of this subscript indicates that the operator is applied to a function of the original variable r.

At order \(\textrm{O}(1)\), we find that \(\rho _0\) must satisfy,

$$\begin{aligned} 0&= \beta _R \Delta _1\rho _0 + \lambda _R \rho _0 - \lambda _R \rho _0^3,\\ 0&= \beta _I \Delta _1 \rho _0 - {\tilde{\alpha }}_I \rho _0 + {\tilde{\alpha }}_I \rho _0^3. \end{aligned}$$

At the next order, \(\textrm{O}(\delta ^2)\), we find two equations involving \(R_0\) and \(\phi _0\),

$$\begin{aligned} 0&= - \beta _I \rho _0 \Delta _{0,S} \phi _0 - 2\beta _I \partial _S \phi _0 \partial _S \rho _0 - \beta _R \rho _0 (\partial _S \phi _0)^2 + \lambda _R R_0( 1- 3 \rho _0^2),\\ 0&= \beta _R \rho _0 \Delta _{0,S} \phi _0 + 2\beta _R \partial _S \phi _0 \partial _S \rho _0 -\beta _I \rho _0 (\partial _S \phi _0)^2 + {\tilde{\alpha }}_I R_0 ( 3 \rho _0^2-1) + {\tilde{\lambda }}_I \rho _0 . \end{aligned}$$

For our purposes, it is enough to stop at this stage and not list higher-order terms.

We first focus on the order \(\textrm{O}(1)\) system. The first equation can be solved, provided \(\beta _R, \lambda _R>0\). This equation falls into a broader family of o.d.e. which were solved in [16]. In this reference, the authors showed that such equations posses a unique solution \(\rho _*\) satisfying

$$\begin{aligned} \rho _* \rightarrow 1 \quad \text{ as } \quad r\rightarrow \infty , \qquad \rho _*(r) \sim br \quad \text{ when } \quad r \sim 0\end{aligned}$$

Of course, the solution \(\rho _*\) would not satisfy the second equation in the system. So, we let

$$\begin{aligned} G = \beta _I \Delta _1 \rho _* - {\tilde{\alpha }}_I \rho _* + {\tilde{\alpha }}_I \rho _*^3 = \left( \frac{\beta _I}{\beta _R} \lambda _R + {\tilde{\alpha }}_I \right) \rho _* ( \rho _*^2-1),\end{aligned}$$

and add these terms to the order \(\textrm{O}(\delta ^2)\) system.

Fig. 3
figure 3

Solution to the boundary value problem (22)

Going back to the order \(\textrm{O}(\delta ^2)\) system, we first notice that because \(\rho _0 =\rho _* \sim br = bS/\delta \) near the origin, then the terms that involve this variable are in fact ‘large’ when compared to the terms that do not. Concentrating only on these large terms, we find that in the first equation we can solve for \(R_0\) in terms of the variable \(\phi _0\). Inserting this result into the second equation gives us the viscous eikonal equation,

$$\begin{aligned} \Delta _{0,S} \phi _0 - b (\partial _S \phi _0)^2 +\Omega - c g=0\end{aligned}$$

as expected, where

$$\begin{aligned} b = \frac{\beta _I \lambda _R - \beta _R {\tilde{\alpha }}_I}{{\tilde{\alpha }}_I \beta _I + \lambda _R \beta _R},\qquad \Omega = \frac{{\tilde{\lambda }}_I\lambda _R}{{\tilde{\alpha }}_I \beta _I + \lambda _R \beta _R}, \qquad c = -\frac{ \beta _I \lambda _R + {\tilde{\alpha }}_I \beta _R}{ \beta _R( {\tilde{\alpha }}_I \beta _I + \lambda _R \beta _R)},\qquad g = (1 - \rho _*^2 ).\end{aligned}$$

Numerical simulations show that the perturbation g decays at order \(\textrm{O}(1/r^2)\) as r goes to infinity, see Fig. 3. To obtain these results, we solved the boundary value problem

$$\begin{aligned} 0 = \partial _{rr} \rho + \frac{1}{r} \partial _r \rho - \frac{1}{r^2} \rho + \rho - \rho ^3,\qquad \rho (\infty ) = 1, \quad \rho (0) =0, \end{aligned}$$
(22)

treating the equation as a system of o.d.e. and using a shooting method with condition

$$\begin{aligned} \rho (r) \sim br \quad \text{ when } \quad r \sim 0.\end{aligned}$$

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jaramillo, G. Can large inhomogeneities generate target patterns?. Z. Angew. Math. Phys. 74, 134 (2023). https://doi.org/10.1007/s00033-023-02027-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00033-023-02027-4

Keywords

Mathematics Subject Classification

Navigation