Skip to main content
Log in

Turing Pattern Formation Under Heterogeneous Distributions of Parameters for an Activator-Depleted Reaction Model

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

The purpose of this article is to study Turing pattern formation in one- and two-dimensional domains under heterogeneous distributions of the parameters for an activator-depleted model. Unlike previous studies of this nature, the choice of the heterogeneous distributions of the parameters is closely linked and estimated by use of rigorous wave mode selection in order to excite different modes in different subsets of the domains. This allows us to relate the numerical solutions with theoretical linear stability analytical results. Our most revealing results show that the wave modes of adjacent subsets evolve locally and yet possess continuity across the interface. These local patches of the solutions result in a globally heterogeneous solution stable only in the presence of heterogeneous distributions of the parameters. Furthermore, our results show that initial conditions continue to play a crucial role in the selection of excitable wave modes and consequently the formation of the inhomogeneous patterns formed. In particular, initial conditions influence pattern orientation and polarity, and yet with a prepattern, the patterns conserved orientation and polarity. Numerical solutions are obtained by the use of the finite element method and the backward Euler scheme to deal with the spatial and the time discretisations, respectively.

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
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. In cases were \(r=0\) and \(r=0.5\) the distribution of parameters is homogeneous inside the whole domain, for the one-dimensional case.

  2. The \(^\star \)’s indicate the use of \(\gamma =276\) and \(d=11.7\) and \(\gamma =1860\) and \(d=8.9\) instead of the values in Table 1.

References

  • Bird, R.B., Stewart, W.E., Lightfoot, E.N.: Transport Phenomena. Wiley, Hoboken (2002)

    Google Scholar 

  • Diamnod, P.H., Ji, X.: Lecture 16: Spatial Pattern Formation by Turing Instability (2017)

  • Duque-Daza, C.A., Ramirez, A.M., Garzón-Alvarado, D.A.: Patrones de turing sobre superficies sometidas a deformación: un acercamiento desde el método lagrangiano total. Revista Internacional de Metodos Numericos para Calculo y Diseño en Ingeniería 28(4), 198–203 (2012)

    Article  MathSciNet  Google Scholar 

  • Garzón-Alvarado, D.A.: Simulación de procesos de reacción-difusión: Aplicación a la morfogénesis de tejido óseo. Ph.D. thesis, Universidad de Zaragoza (2007)

  • Garzón-Alvarado, D.A., Galeano, C.H., Mantilla, J.M.: Turing pattern formation for reaction–convection–diffusion systems in fixed domains submitted to toroidal velocity fields. Appl. Math. Modell. 35, 4913–4925 (2011)

    Article  MathSciNet  Google Scholar 

  • Garzón-Alvarado, D.A., Galeano, C.H., Mantilla, J.M.: Computational examples of reaction–convection–diffusion equations solution under the influence of fluid flow: first example. Appl. Math. Model. 36, 5029–5045 (2012a)

    Article  MathSciNet  Google Scholar 

  • Garzón-Alvarado, D.A., Galeano, C.H., Mantilla, J.M.: Numerical tests on pattern formation in 2D heterogeneous muediums: an approach using the Schnakenberg model. Dyna 79(172), 56–66 (2012b)

    Google Scholar 

  • Gierer, A., Meinhardt, H.: A theory of biological pattern formation. Kybernetik 12, 30–39 (1972)

    Article  Google Scholar 

  • Klika, V., Baker, R.E., Headon, D., Gaffney, E.A.: The influence of receptor–mediated interactions on reaction–diffusion mechanisms of cellular self-organisation. Bull. Math. Biol. 74, 935–957 (2012)

    Article  MathSciNet  Google Scholar 

  • Klika, V., Gaffney, E.A.: History dependence and the continuum approximation breakdown: the impact of domain growth on Turing’s instability. Proc. R. Soc. A. 473, 20160744 (2017)

    Article  MathSciNet  Google Scholar 

  • Kozák, M., Gaffney, E.A., Klika, V.: Pattern formation in reaction–diffusion systems with piecewise kinetic modulation: an example study of heterogeneous kinetics. Phys. Rev. E 100, 042220 (2019)

    Article  Google Scholar 

  • Krause, A.L., Klika, V., Woolley, T.E., Gaffney, E.A.: Heterogeneity induces spatiotemporal oscillations in reaction–diffusion systems. Phys. Rev. E 97, 052206 (2018)

    Article  MathSciNet  Google Scholar 

  • Krause, A.L., Klika, V., Woolley, T.E., Gaffney, E.A.: From one pattern into another: analysis of Turing patterns in heterogeneous domains via WKBJ. J. R. Soc. Interface 17, 20190621 (2020)

    Article  Google Scholar 

  • Madzvamuse, A.: A numerical approach to the study of spatial pattern formation. Ph.D. thesis, Exeter college, University of Oxford (2000)

  • Madzvamuse, A., Chung, A.H.W.: Fully implicit time-stepping schemes and non-linear solvers for systems of reaction–diffusion equations. Appl. Math. Comput. 244, 361–374 (2014)

    MathSciNet  MATH  Google Scholar 

  • Madzvamuse, A., Gaffney, E.A., Maini, P.K.: Stability analysis of non-autonomous reaction–diffusion systems: the effects of growing domains. J. Math. Biol. 61, 133–164 (2010)

    Article  MathSciNet  Google Scholar 

  • Madzvamuse, A., Thomas, R.D.K., Maini, P.K., Wathen, A.J.: A numerical approach to the study of spatial pattern formation in the ligaments of arcoid bivalves. Bull. Math. Biol. 64, 501–530 (2002)

    Article  Google Scholar 

  • Maini, P.K., Benson, D.L., Sherratt, J.A.: Pattern formation in reaction–diffusion models with spatially inhomogeneoos diffusion coefficients. IMA J. Math. Appl. Med. Biol. 9, 197–213 (1992)

    Article  MathSciNet  Google Scholar 

  • May, A., Firby, P.A., Bassom, A.P.: Diffusion driven instability in an inhomogeneous circular domain. Math. Comput. Modell. 29, 53–66 (1999)

    Article  Google Scholar 

  • Murray, J.: Mathematical Biology II: Spatial Models and Biomedical Applications, 3rd Edition, vol. 18. Springer, New York (2003)

    Book  Google Scholar 

  • Nijhout, H.F., Maini, P.K., Madzvamuse, A., Wathen, A.J., Sekimura, T.: Pigmentation pattern formation in butterflies: experiments and models. Comptes Rendus Biologies 326, 717–727 (2003)

    Article  Google Scholar 

  • Page, K., Maini, P.K., Monk, N.A.M.: Pattern formation in spatially heterogeneous Turing reaction–diffusion models. Phys. D 181, 80–101 (2003)

    Article  MathSciNet  Google Scholar 

  • Page, K.M., Maini, P.K., Monk, N.A.M.: Complex pattern formation in reaction–diffusion systems with spatially varying parameters. Phys. D 202, 95–115 (2005)

    Article  MathSciNet  Google Scholar 

  • Perthame, B.: Linear instability, turing instability and pattern formation. In: Parabolic Equations in Biology, Chapter Linear Ins, pp. 117–143. Springer (2015)

  • Rodrigues, D., Barra, L.P., Lobosco, M., Bastos, F.: Analysis of Turing Instability in Biological Models, pp. 576–591. In ICCSA, Part VI (2014)

  • Rueda-Contreras, M.D., Aragón, J.L.: Alan Turing’s chemical theory of phyllotaxis. Revista Mexicana de Física 60, 1–12 (2014)

    MathSciNet  Google Scholar 

  • Sarfaraz, W., Madzvamuse, A.: Classification of parameter spaces for a reaction–diffusion model on stationary domains. Chaos Solitons Fractals 103, 1339–1351 (2017)

    Article  MathSciNet  Google Scholar 

  • Schnakenberg, J.: Simple chemical reaction systems with limit cycle behaviour. J. Theor. Biol. 81, 389–400 (1979)

    Article  MathSciNet  Google Scholar 

  • Sel’kov, E.E.: Self-oscillations in glycolysis. Eur. J. Biochem. 4, 79–86 (1968)

    Article  Google Scholar 

  • Turing, A.M.: The chemical basis of morphogenesis. Philos. Trans. R. Soc. Lond. B Biol. Sci. 237(641), 37–72 (1952)

    Article  MathSciNet  Google Scholar 

  • Wei, M., Wu, J., Guo, G.: Steady state bifurcations for a glycolysis model in biochemical reaction. Nonlinear Anal. Real World Appl. 22, 155–175 (2015)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

DHA was supported by the Universidad Nacional de Colombia through resolutions: 405 of 2019 and 051 and 0354 of 2020. This work was carried out when AM was visiting the Universidad Nacional de Colombia and DHA thanks the University of Sussex for its hospitality during his one-month research visit to the UK. AM is partly supported by the EPSRC Grant Number EP/J016780/1, the European Union Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie Grant Agreement No. 642866, the Commission for Developing Countries and the Simons Foundation. AM is a Royal Society Wolfson Research Merit Award Holder funded generously by the Wolfson Foundation. AM is a Distinguished Visiting Scholar to the University of Johannesburg, Department of Mathematics, South Africa, and the Università degli Studi di Bari Aldo Moro, Bari, Italy.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Diego A. Garzón-Alvarado.

Ethics declarations

Conflict of interest

The author declare that they have no conflict of interest.

Additional information

Communicated by Jeff Moehlis.

Publisher's Note

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

Turing pattern heterogeneous parameters activator-depleted model.

Appendices

Appendix A: Numerical Solution

1.1 Appendix A.1: The Finite Element Weak Formulation

Let \(w\in H^1(\Omega )\) be a test function. Integrating over \(\Omega \) the product of Eq. (1) and w, the weak formulation following integration by parts reads: find \(u,v\in L^2\big ([0,T],H^1(\Omega )\big )\) such that:

$$\begin{aligned} \begin{aligned}&\begin{aligned} \int _\Omega w\frac{\partial u}{\partial t}d\Omega +\int _\Omega \nabla w\cdot \nabla u d\Omega -\int _\Omega w\gamma f(u,v)d\Omega -\int _\Gamma (\nabla u.{\mathbf {n}})w d\Gamma =0, \end{aligned} \\&\begin{aligned} \int _\Omega w\frac{\partial v}{\partial t}d\Omega +\int _\Omega \nabla w\cdot d\nabla vd\Omega -\int _\Omega w\gamma g(u,v)d\Omega -\int _\Gamma (\nabla v.{\mathbf {n}})w d\Gamma =0, \end{aligned} \end{aligned} \end{aligned}$$
(A.1)

where \({\mathbf {n}}\) is the normal vector to \(\Gamma \). Bringing in the zero Neumann boundary conditions, the last term of Eq. (A.1) vanishes.

1.2 Appendix A.2: Spatial Discretisation

To build the finite element approximation, let us consider \(\Omega ^h\subset \Omega \) a discretisation of \(\Omega \) with vertices \(\varvec{x}^h_j\in \Omega \), \(j=1,\ldots ,N_n\) defined by: \(\Omega ^h=\bigcup _{e=1}^{N_e}\Omega _e^h\). In one-dimensional domains, \(\Omega _e^h\), \(e=1,\ldots ,N_e\), are segments limited by two vertex \(\varvec{x}^h_j\), while in two-dimensional domains are quadrilaterals defined by four vertex \(\varvec{x}^h_j\). \(\Omega _e^h\) are defined such that: \( \bigcap _{e=1}^{N_e}\text {Int}\Omega _e^h=\emptyset \), where \(\text {Int}\Omega _e^h\) is the interior of \(\Omega _e^h\) and \(\emptyset \) denotes the empty set.

Let us now define the finite element space as:

$$\begin{aligned} S^h(\Omega ^h)=\big \{\varphi \in&C^0(\Gamma ^h):\varphi \big |_{\Omega _e^h} \text { is linear affine for each }\Omega _e^h\in \Omega ^h\big \}, \end{aligned}$$

with basis \(\{\chi _i\}\) \(i=1,\ldots ,N_n\), such that \(\chi _i(\varvec{x}_j^h)=\delta _{ij}\) (\(\delta \) the Kronecker delta).

Consider the piece-wise linear approximations of \(u(\cdot ,t)\) and \(v(\cdot ,t)\), respectively, \(u^h(\cdot ,t)\in S^h(\Omega ^h)\) and \(v^h(\cdot ,t)\in S^h(\Omega ^h)\) defined by:

$$\begin{aligned} u^h(\cdot ,t)=\sum _{j=1}^{N_n}\chi _j(\cdot )U_j^h(t),\quad \text {and} \quad v^h(\cdot ,t)=\sum _{j=1}^{N_n}\chi _j(\cdot )V_j^h(t) \end{aligned}$$
(A.2)

and the general form of \(\varphi \) as:

$$\begin{aligned} \varphi (\cdot ,t)=\sum _{i=1}^{N_n}\chi _i(\cdot )\eta _i(t), \end{aligned}$$
(A.3)

where \(\eta _i(t)\) are arbitrary bounded values.

Then, the discrete form of the weak problem Eq. (A.1) is: find \(u^h,v^h\in S^h(\Omega ^h)\) such that

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial U_j}{\partial t}\int _{\Omega ^h}\chi _i\chi _j d{\Omega ^h} +U_j\int _{\Omega ^h} \nabla \chi _i\cdot \nabla \chi _j d{\Omega ^h} =\int _{\Omega ^h} \chi _i\gamma f(\chi _jU_j,\chi _jV_j)d{\Omega ^h}, \\ \\ \frac{\partial V_j}{\partial t}\int _{\Omega ^h} \chi _i\chi _jd{\Omega ^h} +V_j\int _{\Omega ^h} \nabla \chi _i\cdot d\nabla \chi _jd{\Omega ^h} =\int _{\Omega ^h} \chi _i\gamma g(\chi _jU_j,\chi _jV_j)d{\Omega ^h}, \end{array}\right. } \end{aligned}$$
(A.4)

notice that the sums for \(u^h\), \(v^h\) and \(\varphi \), the spatial dependency of \(\chi _i\) and \(\chi _j\), and the time dependency of \(U_j(t)\) and \(V_j(t)\) are omitted for notation simplicity, and that since \(\eta _i(t)\) are arbitrary they cancel out.

Let us rewrite Eq. (A.4) as:

$$\begin{aligned} \begin{aligned}&\sum _{e=1}^{N_e}\left( {\mathbb {M}}_e\cdot \frac{\partial {\mathbf {U}}_e}{\partial t}+{\mathbb {K}}_e\cdot {\mathbf {U}}_e-\gamma {\mathbf {F}}_e \right) ={\mathbf {0}},\\&\sum _{e=1}^{N_e}\left( {\mathbb {M}}_e\cdot \frac{\partial {\mathbf {V}}_e}{\partial t}+d{\mathbb {K}}_e\cdot {\mathbf {V}}_e-\gamma {\mathbf {G}}_e \right) ={\mathbf {0}}, \end{aligned} \end{aligned}$$
(A.5)

where \({\mathbf {U}}_e\) and \({\mathbf {V}}_e\) are local nodal values of \(\Omega ^h_e\), \({\mathbb {M}}_e\) are local mass matrices:

$$\begin{aligned} {\mathbb {M}}_e=\int _{\Omega ^h_e} \chi _i\chi _jd{\Omega ^h_e}, \end{aligned}$$
(A.6)

\({\mathbb {K}}_e\) are local stiffness matrices:

$$\begin{aligned} {\mathbb {K}}_e=\int _{\Omega ^h_e} \nabla \chi _i\cdot d\nabla \chi _jd{\Omega ^h_e}, \end{aligned}$$
(A.7)

and \({\mathbf {F}}_e\) and \({\mathbf {G}}_e\) are local reaction vectors:

$$\begin{aligned} {\mathbf {F}}_e=\int _{\Omega ^h_e} \chi _i f(\chi _jU_j,\chi _jV_j)d{\Omega ^h_e},\quad \text {and} \quad {\mathbf {G}}_e=\int _{\Omega ^h_e} \chi _i g(\chi _jU_j,\chi _jV_j)d{\Omega ^h_e}.\qquad \end{aligned}$$
(A.8)

1.3 Appendix A.3: The Temporal Discretisation Element-Wise

So far the system is completely discrete in space. Now, let us deal with the time-dependent terms. Consider \(\tau >0\) a time step such that \(t^k=k\tau \) with \(k=1,\ldots ,N_T\) and \(N_T\) the number of steps to discretise [0, T]. Applying the backward Euler scheme on Eq. (A.4) the fully discrete system of equations is given by:

$$\begin{aligned} {\left\{ \begin{array}{ll} \sum _{e=1}^{N_e}\left( {\mathbb {M}}_e\cdot \frac{{\mathbf {U}}^{k+1}_e-{\mathbf {U}}^{k}_e}{\tau }+{\mathbb {K}}_e\cdot {\mathbf {U}}^{k+1}_e-\gamma {\mathbf {F}}^{k+1}_e \right) ={\mathbf {0}},\\ \\ \sum _{e=1}^{N_e}\left( {\mathbb {M}}_e\cdot \frac{{\mathbf {V}}^{k+1}_e-{\mathbf {V}}^{k}_e}{\tau }+d{\mathbb {K}}_e\cdot {\mathbf {V}}^{k+1}_e-\gamma {\mathbf {G}}^{k+1}_e \right) ={\mathbf {0}}. \end{array}\right. } \end{aligned}$$
(A.9)

Notice that since the nonlinear terms \({\mathbf {F}}_e\) and \({\mathbf {G}}_e\) implicitly depend on time they are computed at \(t^{k+1}\).

1.4 Appendix A.4: Fully Discrete Formulation

Let us now describe the numerical scheme to solve the nonlinear system of equations in Eq. (A.9). Let us rewrite Eq. (A.9) as:

$$\begin{aligned} {\left\{ \begin{array}{ll} \sum _{e=1}^{N_e}\left( {\mathbb {M}}_e\cdot \frac{{\mathbf {U}}^{k+1}_e-{\mathbf {U}}^{k}_e}{\tau }+{\mathbb {K}}_e\cdot {\mathbf {U}}^{k+1}_e-\gamma {\mathbf {F}}^{k+1}_e \right) ={\mathbf {R}}_U,\\ \\ \sum _{e=1}^{N_e}\left( {\mathbb {M}}_e\cdot \frac{{\mathbf {V}}^{k+1}_e-{\mathbf {V}}^{k}_e}{\tau }+d{\mathbb {K}}_e\cdot {\mathbf {V}}^{k+1}_e-\gamma {\mathbf {G}}^{k+1}_e \right) ={\mathbf {R}}_V, \end{array}\right. } \end{aligned}$$
(A.10)

where \({\mathbf {R}}_U\) and \({\mathbf {R}}_V\) are the residuals of the approximation. The Newton–Raphson scheme is then given by:

$$\begin{aligned} \begin{bmatrix} \dfrac{\partial {\mathbf {R}}_U}{\partial {\mathbf {U}}^{k+1}}&{}\dfrac{\partial {\mathbf {R}}_U}{\partial {\mathbf {V}}^{k+1}}\\ \dfrac{\partial {\mathbf {R}}_V}{\partial {\mathbf {U}}^{k+1}}&{}\dfrac{\partial {\mathbf {R}}_V}{\partial {\mathbf {V}}^{k+1}} \end{bmatrix} \begin{bmatrix} \Delta {\mathbf {U}}^{k+1}\\ \Delta {\mathbf {V}}^{k+1} \end{bmatrix} =- \begin{bmatrix} {\mathbf {R}}_U\\ {\mathbf {R}}_V \end{bmatrix}, \end{aligned}$$
(A.11)

where \(\Delta {\mathbf {U}}^{k+1}\) and \(\Delta {\mathbf {V}}^{k+1}\) are the differences of the approximations of \({\mathbf {U}}^{k+1}\) and \({\mathbf {V}}^{k+1}\), respectively, between solver iterations. The partial derivatives in Eq. (A.11) are given by:

$$\begin{aligned} \begin{aligned}&\dfrac{\partial {\mathbf {R}}_U}{\partial {\mathbf {U}}^{k+1}}=\sum _{e=1}^{N_e}\left( \frac{1}{\tau }{\mathbb {M}}_e+{\mathbb {K}}_e-\gamma \int _{\Omega ^h_e}\chi _i\chi _j \frac{\partial f}{\partial u^h}^{k+1} d{\Omega ^h_e} \right) ,\\&\dfrac{\partial {\mathbf {R}}_U}{\partial {\mathbf {V}}^{k+1}}=\sum _{e=1}^{N_e}\left( -\gamma \int _{\Omega ^h_e}\chi _i\chi _j \frac{\partial f}{\partial v^h}^{k+1} d{\Omega ^h_e} \right) ,\\&\dfrac{\partial {\mathbf {R}}_V}{\partial {\mathbf {U}}^{k+1}}=\sum _{e=1}^{N_e}\left( -\gamma \int _{\Omega ^h_e}\chi _i\chi _j \frac{\partial g}{\partial u^h}^{k+1} d{\Omega ^h_e} \right) ,\\&\dfrac{\partial {\mathbf {R}}_V}{\partial {\mathbf {V}}^{k+1}}=\sum _{e=1}^{N_e}\left( \frac{1}{\tau }{\mathbb {M}}_e+d{\mathbb {K}}_e-\gamma \int _{\Omega ^h_e}\chi _i\chi _j \frac{\partial g}{\partial v^h}^{k+1} d{\Omega ^h_e} \right) .\\ \end{aligned} \end{aligned}$$
(A.12)

Appendix B: Further Details, Figures and Tables

1.1 Appendix B.1: Wave Number Tables

See Tables 1 and 2.

Table 1 Combinations of d and \(\gamma \) for the Schnakenberg reaction model with \(a=0.1\) and \(b=0.9\) for some one-dimensional wave modes
Table 2 Combinations of d and \(\gamma \) for the Schnakenberg reaction model with \(a=0.1\) and \(b=0.9\) for some two-dimensional wave modes (Garzón-Alvarado 2007)

1.2 Appendix B.2: Changes in Polarity and Orientation

See Fig. 14.

Fig. 14
figure 14

Schematic diagrams of pattern polarity and orientation differences. a A reference pattern with a change of polarity in b and a change in orientation in c

1.3 Appendix B.3: Schematic Representation of \(c(k_2)\)

See Fig. 15.

Fig. 15
figure 15

Schematic representation of \(c(k_2)\)

1.4 Appendix B.4: Pattern Normalisation

$$\begin{aligned} v_i^*=\frac{v_i-\min (v)}{\max (v)-\min (v)}, \end{aligned}$$
(B.1)

where \(v_i\) is the nodal value of v at the ith node (given by the finite element method discretisation).

1.5 Appendix B.5: Discretisation and Convergence Criteria

Regarding the discretisation, one-dimensional simulations were performed with uniform meshes of 1001 nodes and 1000 elements, and the two-dimensional simulations with uniform meshes of 10,000 elements and 10,201 nodes. Additionally, a time step of 0.01 was defined.

The discrete \(L^2\)-norm time derivative was calculated globally, as presented in Sarfaraz and Madzvamuse (2017):

$$\begin{aligned} \frac{||\mathbf {u}^{\mathbf {t}+{\varvec{\Delta }} \mathbf {t}}-\mathbf {u}^{\mathbf {t}||}}{\Delta t},\quad \text {and} \quad \frac{||\mathbf {v}^{\mathbf {t}+{\varvec{\Delta }} \mathbf {t}}-\mathbf {v}^{\mathbf {t}||}}{\Delta t}, \end{aligned}$$
(B.2)

and the simulations were stopped after a given tolerance, say \(\varepsilon =10^{-6}\), was reached for both species.

1.6 Appendix B.6: Standard Convergence Graphics

See Fig. 16.

Fig. 16
figure 16

Convergence for the unit one-dimensional domain. The bold line ( ) and the dashed line ( ) show the solutions of u and v, respectively. a Convergence to the heterogeneous solution with \((n_{\text {in}}=2,n_{\text {out}}=1)\) and \(r=0.25\) and b convergence to the homogeneous solution with \((n_{\text {in}}=2,n_{\text {out}}=1)\) and \(r=0.4\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hernandez-Aristizabal, D., Garzón-Alvarado, D.A. & Madzvamuse, A. Turing Pattern Formation Under Heterogeneous Distributions of Parameters for an Activator-Depleted Reaction Model. J Nonlinear Sci 31, 34 (2021). https://doi.org/10.1007/s00332-021-09685-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-021-09685-6

Keywords

Navigation