Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article Special Issues

The effect of landscape fragmentation on Turing-pattern formation


  • Diffusion-driven instability and Turing pattern formation are a well-known mechanism by which the local interaction of species, combined with random spatial movement, can generate stable patterns of population densities in the absence of spatial heterogeneity of the underlying medium. Some examples of such patterns exist in ecological interactions between predator and prey, but the conditions required for these patterns are not easily satisfied in ecological systems. At the same time, most ecological systems exist in heterogeneous landscapes, and landscape heterogeneity can affect species interactions and individual movement behavior. In this work, we explore whether and how landscape heterogeneity might facilitate Turing pattern formation in predator–prey interactions. We formulate reaction-diffusion equations for two interacting species on an infinite patchy landscape, consisting of two types of periodically alternating patches. Population dynamics and movement behavior differ between patch types, and individuals may have a preference for one of the two habitat types. We apply homogenization theory to derive an appropriately averaged model, to which we apply stability analysis for Turing patterns. We then study three scenarios in detail and find mechanisms by which diffusion-driven instabilities may arise even if the local interaction and movement rates do not indicate it.

    Citation: Nazanin Zaker, Christina A. Cobbold, Frithjof Lutscher. The effect of landscape fragmentation on Turing-pattern formation[J]. Mathematical Biosciences and Engineering, 2022, 19(3): 2506-2537. doi: 10.3934/mbe.2022116

    Related Papers:

    [1] Alessandro Venditti, Claudio Frezza, Giacomo Rossi, Mirella Di Cecco, Giampiero Ciaschetti, Mauro Serafini, Armandodoriano Bianco . Secondary metabolites with ecologic and medicinal implications in Anthemis cretica subsp. petraea from Majella National Park. AIMS Molecular Science, 2016, 3(4): 648-660. doi: 10.3934/molsci.2016.4.648
    [2] Francisco Les, Zsuzsa Iffiú-Soltész, Josep Mercarder, Christian Carpéné . Tyramine activates lipid accumulation in rat adipocytes: influences of in vitro and in vivo administration. AIMS Molecular Science, 2017, 4(3): 339-351. doi: 10.3934/molsci.2017.3.339
    [3] Jin-Yih Low, Helen D. Nicholson . The up-stream regulation of polymerase-1 and transcript release factor(PTRF/Cavin-1) in prostate cancer: an epigenetic analysis. AIMS Molecular Science, 2016, 3(3): 466-478. doi: 10.3934/molsci.2016.3.466
    [4] Harkirat S. Sethi, Jessica L. Osier, Geordan L. Burks, Jennifer F. Lamar, Hana McFeeters, Robert L. McFeeters . Expedited isolation of natural product peptidyl-tRNA hydrolase inhibitors from a Pth1 affinity column. AIMS Molecular Science, 2017, 4(2): 175-184. doi: 10.3934/molsci.2017.2.175
    [5] Rachel L. Flores, Kayla Gonzales, Ryan W. Seaver, Brian T. Livingston . The skeletal proteome of the brittle star Ophiothrix spiculata identifies C-type lectins and other proteins conserved in echinoderm skeleton formation. AIMS Molecular Science, 2016, 3(3): 357-367. doi: 10.3934/molsci.2016.3.357
    [6] Charles A. Downs, Abdel A. Alli, Nicholle M. Johnson, My N. Helms . Cigarette smoke extract is a Nox agonist and regulates ENaC in alveolar type 2 cells. AIMS Molecular Science, 2016, 3(3): 439-453. doi: 10.3934/molsci.2016.3.439
    [7] Genu George, Sumita K. Chellappan, Mandagini Geetha, Padinjaradath S. Appukuttan . Possible molecular basis for macromolecular antigen attachment to host cells: their immune complex with plasma antibodies have unoccupied binding sites enabling binding to smaller ligands. AIMS Molecular Science, 2017, 4(1): 91-102. doi: 10.3934/molsci.2017.1.91
    [8] Sarika Amdekar, Avnish Kumar, Vinod Singh . Transcriptional activity of cytokines induced by oral administration of Lactobacillus casei and Lactobacillus acidophilus in experimental model of arthritis. AIMS Molecular Science, 2017, 4(2): 164-174. doi: 10.3934/molsci.2017.2.164
    [9] Noor Riyadh Thiab, Nicola King, Mary McMillan, Amer Almashhadany, Graham L Jones . Age-related protein and mRNA expression of glutathione peroxidases (GPx) and Hsp-70 in different regions of rat kidney with and without stressor. AIMS Molecular Science, 2016, 3(2): 125-137. doi: 10.3934/molsci.2016.2.125
    [10] Jin-Yih Low, Helen D. Nicholson . Investigating the role of caveolin-2 in prostate cancer cell line. AIMS Molecular Science, 2017, 4(1): 82-88. doi: 10.3934/molsci.2017.1.82
  • Diffusion-driven instability and Turing pattern formation are a well-known mechanism by which the local interaction of species, combined with random spatial movement, can generate stable patterns of population densities in the absence of spatial heterogeneity of the underlying medium. Some examples of such patterns exist in ecological interactions between predator and prey, but the conditions required for these patterns are not easily satisfied in ecological systems. At the same time, most ecological systems exist in heterogeneous landscapes, and landscape heterogeneity can affect species interactions and individual movement behavior. In this work, we explore whether and how landscape heterogeneity might facilitate Turing pattern formation in predator–prey interactions. We formulate reaction-diffusion equations for two interacting species on an infinite patchy landscape, consisting of two types of periodically alternating patches. Population dynamics and movement behavior differ between patch types, and individuals may have a preference for one of the two habitat types. We apply homogenization theory to derive an appropriately averaged model, to which we apply stability analysis for Turing patterns. We then study three scenarios in detail and find mechanisms by which diffusion-driven instabilities may arise even if the local interaction and movement rates do not indicate it.



    The following two-dimensional nonlinear multi-term time-fractional subdiffusion equations with initial and boundary conditions are considered in this paper:

    Dαtu=Δu+f(u), (x,y)Ω, 0<tT, (1.1)
    u(x,y,0)=g(x,y), (x,y)Ω, (1.2)
    u(x,y,t)=0, (x,y)Ω, 0tT, (1.3)

    where Ω=(0,L1)×(0,L2) with the boundary Ω, fC1(R), and g is a continuous function in Ω. The operator Dαt in (1.1) is defined by

    Dαt=Jl=1blDαlt, 0<αJ<αJ1<<α1<1, and bl>0,

    where Dαlt denotes the Caputo derivative with respect to t, i.e.,

    Dαltu=1Γ(1αl)t0(ts)αlu(x,y,s)sds.

    In order to describe the physical process better, some scholars use multi-term time fractional differential equations for modeling, such as the behavior of viscoelastic fluids[1,2], the dispersion of pollutants[3], and magnetic resonance imaging[4]. The finite difference method[5,6,7,8], the finite volume method [9,10], the fractional predictor-corrector method[11], the finite element method[12], the collocation method [13,14], and the spectral method[15] have been developed for solving the multi-term time fractional equations. The work [16] proposed a fast linearized finite difference method for the nonlinear multi-term time-fractional wave equation using SOE approximation. Recently, a lot of work about the time fractional nonlinear subdiffusion equation has been published. For example, the paper [17] considered a Newton linearized Galerkin finite element method to solve the problem with non-smooth solutions in the time direction and provided an optimal error estimate by using the discrete fractional Gronwall-type inequality on the graded meshes. The paper [18] provided an efficient systematic framework for solving nonlinear fractional partial differential equations on unbounded domains and obtained an error estimate. Jiang et al. [19] proposed an efficient ADI scheme for the nonlinear subdiffusion equation with a weakly singular solution and obtained the pointwise-in-time error estimate. Li et al. [20] proposed a new tool, the refined discrete fractional-type Gr¨onwall inequality, which is used to derive a sharp pointwise-in-time error estimate of the L1 scheme for the problem. However, as far as the authors know, at present there is no work dedicated to the pointwise-in-time error estimate of the L1 scheme for the nonlinear multi-term time-fractional subdiffusion equation with weakly singular solutions. The present work is designed to fill this gap. Following [21,22], in the remainder of our paper, we make the following assumption:

    Assumption 1.1. For all (x,y,t)ˉΩ×(0,T], we assume the solution satisfies

    |lu(x,y,t)xpyq|C, l=0,1,2,3,4,5, l=p+q, (1.4)
    |ku(x,y,t)tk|C(1+tα1k), k=0,1,2, (1.5)

    where C is a positive constant. C in this paper represents a constant independent of time and space step, and C in different positions represents different values.

    This paper consists of the following sections. In Section 2, we construct a fully discrete scheme for the problem (1.1)–(1.3). In Section 3, some lemmas are introduced that will be used in the subsequent analysis. In Section 4, convergence and stability of the fully discrete scheme proposed in Section 2 are given, and we obtain the sharp pointwise-in-time error estimate. In Section 5, the theoretical analysis is verified by four numerical examples.

    Let M1, M2, and N be three positive integers. Divide space and time uniformly into M1×M2 and N parts, respectively. Let {tn|tn=nτ, 0nN} be a uniform partition of [0,T] with the time step τ=T/N. Let h1=L1/M1 and h2=L2/M2 be the spatial steps. So spatial grid points consist of xi=ih1 and yj=jh2, where i=0,1,,M1 and j=0,1,,M2. The spatial grid is represented by Ω={(xi,yj)|i=0,1,,M1,j=0,1,,M2}. Let Ωh=ΩΩ, and Ωh=ΩΩ. Defining uni,j=u(xi,yj,tn), the previous equations (1.1)–(1.3) at the grid point (xi,yj,tn) can be transformed into

    Dατuni,j=δ2xuni,j+δ2yuni,j+f(un1i,j)+Jl=1bl(rl)ni,j+Rni,j+(ˉr)ni,j, (xi,yj)Ωh, 0<tnT, (2.1)
    u0i,j=g(xi,yj), (xi,yj)Ωh, (2.2)
    uni,j=0, (xi,yj)Ωh, 0tnT, (2.3)

    where the notations in (2.1) will be given afterwards.

    Applying an L1 formula for the multi-term time Caputo derivatives, i.e., set

    Dαlτuni,j=dαl1uni,jdαlnu0i,jn1k=1(dαlkdαlk+1)unki,j,

    where dαlk=ταl[k1αl(k1)1αl]Γ(2αl). For simplicity, defining dk=Jl=1bldαlk, it yields that

    Dατuni,j:=Jl=1blDαlτuni,j=d1uni,jdnu0i,jn1k=1(dkdk+1)unki,j. (2.4)

    A standard second-order approximation is used to discretize Δuni,j:

    Δuni,jδ2xuni,j+δ2yuni,j,δ2xuni,j=uni+1,j2uni,j+uni1,jh12,δ2yuni,j=uni,j+12uni,j+uni,j1h22.

    To approximate the nonlinear term f(uni,j), we utilize the backward formula, expressed as

    f(uni,j)=f(un1i,j)+(ˉr)ni,j.

    In (2.1), (rl)ni,j and Rni,j are truncation errors, i.e.,

    (rl)ni,j=Dαlτuni,jDαltuni,j,Rni,j=Δuni,j(δ2xuni,j+δ2yuni,j).

    Leaving out the truncation errors in (2.1) and replacing uni,j by Uni,j, the fully discrete scheme is obtained:

    DατUni,j=δ2xUni,j+δ2yUni,j+f(Un1i,j), (xi,yj)Ωh, 0<tnT, (2.5)
    U0i,j=g(xi,yj), (xi,yj)Ωh, (2.6)
    Uni,j=0, (xi,yj)Ωh, 0tnT. (2.7)

    Lemma 3.1. For (xi,yj)Ωh, 1nN, it holds that

    |(ˉr)ni,j|Cτtα11n.

    Proof. Due to the continuity of f and the boundedness of u, f(ζ) is bounded, where ζ is between uni,j and un1i,j. We obtain

    |(ˉr)ni,j|=|f(uni,j)f(un1i,j)|C|uni,jun1i,j|.

    Then we estimate the truncation error in two cases ([19], page 6). Note the condition (1.5).

    Case A: n2

    |(ˉr)ni,j|C|uni,jun1i,j|=Cτ|u(ih1,jh2,t)t|t=ξ (tn1<ξ<tn)Cτ(1+ξα11)Cτ(1+tα11n1)Cτ(1+tα11n)Cτtα11n,

    where we have used tn1tn for n2 in the penultimate inequality.

    Case B: n=1

    |(ˉr)1i,j|C|t10u(ih1,jh2,s)sds|Ct10|u(ih1,jh2,s)s|dsCt10(1+sα11)ds=C(τ+1α1τα1)Cτα1=Cτtα111.

    The proof is completed.

    Lemma 3.2. For (xi,yj)Ωh, 1nN, we have |Rni,j|C(h21+h22).

    Proof. According to the Taylor expansion, we can see that

    |Rni,j|=|Δuni,j(δ2xuni,j+δ2yuni,j)|=|(uxx(xi,yj,tn)δ2xuni,j)+(uyy(xi,yj,tn)δ2yuni,j)|h2112|4u(x,yj,tn)x4|x=ξ1|+h2212|4u(xi,y,tn)y4|y=ξ2|C(h21+h22),

    where xi1<ξ1<xi+1, and yj1<ξ2<yj+1.

    Lemma 3.3. [23, Lemma 6] |Jl=1bl(rl)ni,j|Cnmin{2α1,α1+1}, for (xi,yj)Ωh, 1nN.

    Define the positive stability multipliers θ0=μ, θn=μnk=1(dkdk+1)θnk, where μ=d11.

    Lemma 3.4. ([24], Lemma 5.1) θn is a monotonically decreasing sequence with respect to n. For n=0,1,,N,

    θnb11Γ(2α1)τα1(n+1)α11.

    Notation. θn here has the following relationship with σαn in Literature [24]: θn=σαn.

    Lemma 3.5. ([24], Lemma 5.2, Corollary 5.1) For n=1,2,,N,

    nj=1jβθnjb11Γ(2α1)τα1[κβ,n(n2)α11+1α1(n2)α1β],

    where β0 and

    κβ,n={1+1n1ββ1,  for  β1,1+lnn,  for  β=1.

    If there exists α_, which satisfies 0<α_α1<1, we have

    nj=1jβθnj{ˆC for β=0,ˆCKβ,nτtα11n for β>1,

    where constant ˆC depends on α_ and T.

    Lemma 3.6. ([25], Lemma 5.1) If {yn} is a non-negative sequence, and it satisfies yna1tη1n+a2tη2n+bτn1j=1tα1njyj, 1nN, we get ynC(a1tη1n+a2tη2n), 1nN, where 0η1,η2<1, a1,a2,b>0, 0<α<1 and N is a positive integer.

    Suppose gj is an arbitrary mesh function; we define an integral operator Bαt, which satisfies  Bαt(g0)=0, Bαt(gn)=nj=1θnjgj, for n=1,2,,N.

    Lemma 3.7. ([23], Lemma 2) For any mesh function {Uj}Nj=0, the following formula holds.

    Bαt(DατUn)=UnU0, for n=1,2,,N.

    Theorem 4.1. If Uni,j is the solution of (2.5)(2.7) at point (xi,yj,tn), and suppose |Uni0,j0|=, it holds that

    \begin{align} \left| U_{{{i}_{0}},{{j}_{0}}}^{n} \right|\le \left| U_{{{i}_{0}},{{j}_{0}}}^{0} \right|+ \sum\limits_{k = 1}^{n}{{{\theta }_{n-k}}\left| f(U_{{{i}_{0}},{{j}_{0}}}^{k-1}) \right|},\ for\ n\ge 0. \end{align} (4.1)

    Proof. From (2.5), it is easy to see that U_{i, j}^{n} satisfies

    \begin{align*} &(\sum\limits_{l = 1}^{J}{{{b}_{l}}d_{1}^{{{\alpha }_{l}}}}+\frac{2}{{h_1^{2}}}+\frac{2}{{h_2^{2}}})\left| U_{i,j}^{n} \right| \\ = &\left| \frac{U_{i+1,j}^{n}+U_{i-1,j}^{n}}{{h_1^{2}}}+\frac{U_{i,j+1}^{n}+U_{i,j-1}^{n}}{{h_2^{2}}}+\sum\limits_{l = 1}^{J}{\sum\limits_{k = 2}^{n}{{{b}_{l}}(d_{k-1}^{{{\alpha }_{l}}}-d_{k}^{{{\alpha }_{l}}})U_{i.j}^{n-k+1}}}+d_nU_{i,j}^0+f(U_{{i},{j}}^{n-1}) \right|. \end{align*}

    Based on (2.7) and the fact that d_{k-1}^{\alpha_l} > d_k^{\alpha_l} , we have

    \begin{align*} &(\sum\limits_{l = 1}^{J}{{{b}_{l}}d_{1}^{{{\alpha }_{l}}}}+\frac{2}{{h_1^{2}}}+\frac{2}{{h_2^{2}}})\left| U_{i_0,j_0}^{n} \right| \\ \le & \left| \frac{2U_{{{i}_{0}},{{j}_{0}}}^{n}}{{h}_{1}^{2}} \right|+\left| \frac{2U_{{{i}_{0}},{{j}_{0}}}^{n}}{{h}_{2}^{2}} \right|+\sum\limits_{l = 1}^{J}{\sum\limits_{k = 2}^{n}{{{b}_{l}}(d_{k-1}^{{{\alpha }_{l}}}-d_{k}^{{{\alpha }_{l}}})\left| U_{{{i}_{0}},{{j}_{0}}}^{n-k+1} \right|}}+d_n\left|U_{i_0,j_0}^0\right|+\left| f(U_{{{i}_{0}},{{j}_{0}}}^{n-1}) \right|. \end{align*}

    Then, we obtain

    \begin{equation*} (\sum\limits_{l = 1}^{J}{{{b}_{l}}d_{1}^{{{\alpha }_{l}}}})\left| U_{i_0,j_0}^{n} \right|\le \sum\limits_{l = 1}^{J}{\sum\limits_{k = 2}^{n}{{{b}_{l}}(d_{k-1}^{{{\alpha }_{l}}}-d_{k}^{{{\alpha }_{l}}})\left| U_{{{i}_{0}},{{j}_{0}}}^{n-k+1} \right|}}+d_n\left|U_{i_0,j_0}^0\right|+\left| f(U_{{{i}_{0}},{{j}_{0}}}^{n-1}) \right|, \end{equation*}

    which, according to the definition of D_{\tau }^{\boldsymbol{\alpha} } , is equivalent to the following expression:

    \begin{align} D_{\tau }^{\boldsymbol{\alpha}}\left| U_{i_0,j_0}^{n} \right|\le \left| f(U_{i_0,j_0}^{n-1}) \right|,\ 1\le n\le N. \end{align} (4.2)

    By applying the integral operator to both sides of the above formula, we obtain

    \begin{equation*} B_{t}^{\boldsymbol{\alpha}}(D_{\tau }^{\boldsymbol{\alpha} }\left| U_{{{i}_{0}},{{j}_{0}}}^{n} \right|)\le B_{t}^{\boldsymbol{\alpha} }\left| f(U_{{{i}_{0}},{{j}_{0}}}^{n-1}) \right|. \end{equation*}

    Due to Lemma 3.7 and the above definition of B_{t}^{\boldsymbol{\alpha} } , (4.1) holds for n\ge 1 . Obviously, (4.1) also holds for n = 0 .

    Theorem 4.2. Suppose that u is the solution of (1.1)(1.3) and satisfies Assumption 1.1, and that U is the solution of (2.5)(2.7). There exist positive constants \tau_0 and h_0 . When \tau < \tau_0 and h_1, h_2 < h_0 , it holds for m > 0 that

    \begin{align} {{\left\| {{u}^{m}}-{{U}^{m}} \right\|}_{\infty }}\le C(\tau t_{m}^{{\alpha }_{1}-1}+h_1^2+h_2^2). \end{align} (4.3)

    Proof. For m = 0 , (4.3) holds obviously. Assuming that (4.3) holds for m = 0, 1, 2\cdots n-1 ( n\ge 1 ), then for sufficiently small \tau , h_1 , h_2 , and 1\leq k \leq n , we obtain

    \begin{equation*} \begin{split} {{\left\| {{U}^{k-1}} \right\|}_{\infty }}&\le {{\left\| {{u}^{k-1}} \right\|}_{\infty }}+{{\left\| {{u}^{k-1}}-{{U}^{k-1}} \right\|}_{\infty }} \\ & \le \left\| {{u}^{k-1}} \right\|_{\infty }+1. \end{split} \end{equation*}

    Let us discuss whether the inequality (4.3) holds when m = n . Let e_{i, j}^m = u_{i, j}^m-U_{i, j}^m for 0\leq m\leq n . Subtracting (2.5)–(2.7) from (2.1)–(2.3), we obtain

    \begin{equation*} \begin{split} &D_{\tau }^{\boldsymbol{\alpha}}e_{i,j}^{m} = \delta _{x}^{2}e_{i,j}^{m}+\delta _{y}^{2}e_{i,j}^{m}+({{R}_{f}})_{i,j}^{m}+\sum\limits_{l = 1}^{J}{{{b}_{l}}({{r}_{l}})_{i,j}^{m}+R_{i,j}^{m}+(\bar{r})_{i,j}^{m}},\ (x_i,y_j)\in \Omega_h,\ 0 < t_m\le T, \\ & e_{i,j}^{0} = 0,\ (x_i,y_j)\in \Omega_h,\\ & e_{i,j}^{m} = 0,\ (x_i,y_j)\in \partial \Omega_h,\ 0\le t_m\le T, \end{split} \end{equation*}

    where ({{R}_{f}})_{i, j}^{m} = f(u_{i, j}^{m-1})-f(U_{i, j}^{m-1}) . Considering the continuity of f' , the boundedness of u_{i, j}^{m-1} and U_{i, j}^{m-1} , we have

    \begin{align} \left| ({{R}_{f}})_{i,j}^{m} \right| = \left| f(u_{i,j}^{m-1})-f(U_{i,j}^{m-1}) \right|\le C\left| e_{i,j}^{m-1} \right|. \end{align} (4.4)

    Let \left|e_{{{i}_{0}}, {{j}_{0}}}^{m}\right| = {{\left\| {{e}^{m}} \right\|}_{\infty }} . Using Lemmas 3.1–3.3 and inequality (4.4), similar to (4.2), it can be obtained that for 1\leq m\leq n ,

    \begin{align} D_{\tau }^{\boldsymbol{\alpha}}\left| e_{{{i}_{0},j_0}}^{m} \right|\le {{C}_{1}}\left| e_{{{i}_{0},j_0}}^{m-1} \right|+{{C}_{2}}{{m}^{-\min \{2-{{\alpha }_{1}},{{\alpha }_{1}}+1\}}}+{{C}_{3}}(h_1^2+h_2^2)+{{C}_{4}}\tau t_{m}^{{{\alpha }_{1}}-1}. \end{align} (4.5)

    Applying the definition of B_{t }^{\boldsymbol{\alpha}} again, and Lemmas 3.4 and 3.5, we can further obtain

    \begin{equation*} \begin{split} \left| e_{{{i}_{0},j_0}}^{m} \right|&\le C_1 \sum\limits_{k = 1}^{m}{{{\theta }_{m-k}}\left| e_{{{i}_{0},j_0}}^{k-1} \right|}+{{C}_{2}} \sum\limits_{k = 1}^{m}{{{\theta }_{m-k}}{{k}^{-\min \{2-{{\alpha }_{1}},1+{{\alpha }_{1}}\}}}}\\ &+C_3 \sum\limits_{k = 1}^{m}{{\theta }_{m-k}}{(h_1^2+h_2^2)}+{{C}_{4}}{{\tau }^{{{\alpha }_{1}}}} \sum\limits_{k = 1}^{m}{{{\theta }_{m-k}}{{k}^{{{\alpha }_{1}}-1}}} \\ & = {{C}_{1}} \sum\limits_{k = 1}^{m-1}{{{\theta }_{m-k-1}}\left| e_{{{i}_{0},j_0}}^{k} \right|}+{{C}_{2}} \sum\limits_{k = 1}^{m}{{{\theta }_{m-k}}{{k}^{-\min \{2-{{\alpha }_{1}},1+{{\alpha }_{1}}\}}}} \\ & +{{C}_{3}}\hat{C}(h_1^2+h_2^2)+{{C}_{4}}{{\tau }^{{{\alpha }_{1}}}} \sum\limits_{k = 1}^{m}{{{\theta }_{m-k}}{{k}^{{{\alpha }_{1}}-1}}} \\ & \le {{C}_{1}}b_{1}^{-1}\Gamma (2-{{\alpha }_{1}}){{\tau }^{{{\alpha }_{1}}}}\sum\limits_{k = 1}^{m-1}{{{(m-k)}^{{{\alpha }_{1}}-1}}\left| e_{{{i}_{0},j_0}}^{k} \right|} \\ & +{{C}_{2}}\hat{C}{{\kappa }_{{{\sigma }_{1}},m}}\tau t_{m}^{{{\alpha }_{1}}-1} +{{C}_{3}}\hat{C}(h_1^2+h_2^2) \\ & +{{C}_{4}}{{\tau }^{2{{\alpha }_{1}}}}b_{1}^{-1}\Gamma (2-{{\alpha }_{1}})\left[\left(1-\frac{1}{\alpha_1}\right)\left(\frac{m}{2}\right)^{\alpha_1-1}+\frac{m^{2\alpha_1-1}}{\alpha_12^{\alpha_1-1}}+\frac{1}{\alpha_1}\left(\frac{m}{2}\right)^{2\alpha_1-1}\right]\\ &\le C\left({{\tau }^{{{\alpha }_{1}}}}\sum\limits_{k = 1}^{m-1}{{{(m-k)}^{{{\alpha }_{1}}-1}}\left| e_{{{i}_{0},j_0}}^{k} \right|} +\tau t_{m}^{{{\alpha }_{1}}-1} +(h_1^2+h_2^2)\right), \end{split} \end{equation*}

    where \sigma_1 = \min \{2-{{\alpha }_{1}}, {{\alpha }_{1}}+1\} . From Lemma 3.6, we have

    \begin{align} \left| e_{{{i}_{0},j_0}}^{m} \right| \le C(\tau t_{m}^{{{\alpha }_{1}}-1}+h_1^2+h_2^2),\ for\ 1\leq m\leq n. \end{align} (4.6)

    In summary, the inequality (4.3) holds for m = n , which finishes the mathematical induction. The proof is complete.

    Remark 4.1. The global maximum error of the numerical solution is

    \begin{equation*} \underset{1\le n\le N}{\mathop{\max\limits}}{{\left\| {{u}^{n}}-{{U}^{n}} \right\|}_{\infty }}\le C(\tau^{\alpha_1}+h_1^2+h_2^2). \end{equation*}

    When t_n is away from 0, the local maximum error is

    \begin{equation*} {{\left\| {{u}^{n}}-{{U}^{n}} \right\|}_{\infty }}\le C(\tau+h_1^2+h_2^2). \end{equation*}

    Theorem 4.3. The fully discrete scheme (2.5)(2.7) is stable with respect to the initial value. If \hat{U}_{i, j}^n satisfies the following equations,

    \begin{align} & \sum\limits_{l = 1}^{J}{{{b}_{l}}D_{\tau }^{{{\alpha }_{l}}}\hat{U}_{i,j}^{n}} = \delta _{x}^{2}\hat{U}_{i,j}^{n}+\delta _{y}^{2}\hat{U}_{i,j}^{n}+f(\hat{U}_{i,j}^{n-1}),\ (x_i,y_j)\in \Omega_h,\ 0 < t_n\le T, \end{align} (4.7)
    \begin{align} & \hat{U}_{i,j}^{0} = \hat{g}(x_i,y_j),\ (x_i,y_j)\in \Omega_h, \end{align} (4.8)
    \begin{align} & \hat{U}_{i,j}^{n} = 0,\ (x_i,y_j)\in \partial \Omega_h,\ 0\le t_n\le T, \end{align} (4.9)

    and {{\left\| g- \hat{g}\right\|}_{\infty }} is sufficiently small, then

    \begin{align} {{\left\| \hat{e}^n \right\|}_{\infty }}\le C{{\left\| \hat{e}^0 \right\|}_{\infty }},\ for\ n\ge 0, \end{align} (4.10)

    where \hat{e}_{i, j}^n = U_{i, j}^n-\hat{U}_{i, j}^n .

    Proof. Subtracting (4.7)–(4.9) from (2.5)–(2.7) and according to Theorem 4.1, we have

    \begin{equation*} \left| \hat{e}_{i_0,j_0}^n \right|\le \left| \hat{e}_{i_0,j_0}^0 \right|+ \sum\limits_{k = 1}^{n}{{{\theta }_{n-k}}\left| f(U_{i_0,j_0}^{k-1})-f(\hat{U}_{i_0,j_0}^{k-1}) \right|}, \end{equation*}

    where \hat{e}_{i_0, j_0}^n = {{\left\| \hat{e}^n \right\|}_{\infty }} . (4.10) holds for n = 0 obviously. Suppose that (4.10) holds when n = 0, 1, 2, \cdots, m-1(m\ge 1) , we have {{\left\| \hat{U}^r \right\|}_{\infty }}\le {{\left\| \hat{e}^r \right\|}_{\infty }}+{{\left\| U^r \right\|}_{\infty }}\le C{{\left\| \hat{e}^0 \right\|}_{\infty }}+{{\left\| U^r \right\|}_{\infty }}\le 1+{{\left\| U^r \right\|}_{\infty }} on the condition that {{\left\| \hat{e}^0 \right\|}_{\infty }} is small, for r\le m-1 . According to (4.3), U_{i, j}^r is bounded. Then, combining the continuity of f' and the boundedness of \hat{U}_{i, j}^r , it holds that

    \begin{equation*} \begin{split} \left| \hat{e}_{i_0,j_0}^m \right|&\le \left| \hat{e}_{i_0,j_0}^0 \right|+ \sum\limits_{k = 1}^{m}{{{\theta }_{m-k}}\left| f(U_{i_0,j_0}^{k-1})-f(\hat{U}_{i_0,j_0}^{k-1}) \right|}\\ &\le \left| \hat{e}_{i_0,j_0}^0 \right|+C \sum\limits_{k = 1}^{m}{{{\theta }_{m-k}}\left| \hat{e}_{i_0,j_0}^{k-1} \right|}\\ &\le C\left| \hat{e}_{i_0,j_0}^0 \right|+C \sum\limits_{k = 1}^{m-1}{{{\theta }_{m-k-1}}\left| \hat{e}_{i_0,j_0}^{k} \right|}\\ &\le C\left| \hat{e}_{i_0,j_0}^0 \right|+C\tau \sum\limits_{k = 1}^{m-1}{t_{m-k}^{\alpha_1-1}\left| \hat{e}_{i_0,j_0}^{k} \right|}.\\ \end{split} \end{equation*}

    Due to Lemma 3.6, we have

    \begin{equation*} {{\left\| \hat{e}^m \right\|}_{\infty }}\le C{{\left\| \hat{e}^0 \right\|}_{\infty }}. \end{equation*}

    Therefore, the mathematical induction ends, and the proof is complete.

    Example 5.1. We first consider the following two-term time-fractional nonlinear subdiffusion equation with b_1 = b_2 = 1 .

    \begin{align} &D_{t}^{\alpha_1}u+D_{t}^{\alpha_2}u = \Delta u+u(1-u^2)+h(x,y,t),\ (x,y)\in \Omega,\ 0 < t\le T, \end{align} (5.1)
    \begin{align} &u(x,y,0) = g(x,y),\ (x,y)\in \Omega, \end{align} (5.2)
    \begin{align} &u(x,y,t) = 0,\ (x,y)\in \partial \Omega,\ 0\le t\le T, \end{align} (5.3)

    where h(x, y, t) and g(x, y) are up to the exact solution. We set the exact solution to t^{\alpha_1}sin(\pi x)sin(\pi y) , which satisfies Assumption 1.1. We consider the spatial domain \Omega = (0, 1)\times (0, 1) and set T = 1 .

    Above all, we compute the convergence order in spatial direction. We set N = 1000 so that the influence of errors in temporal direction can be ignored compared with errors in spatial direction. The maximum errors at t_n = 1 and rates, when \alpha_1 = 0.4 and \alpha_2 = 0.3 , are presented in Table 5.1. Numerical results show that the spatial accuracy is O(h_1^2+h_2^2) . For studying temporal convergence rates, we define global errors E_G and local errors E_L by

    \begin{equation*} E_G = \underset{1\le n\le N}{\mathop{\max\limits}}{{\left\| U^n-u^n \right\|}_{\infty }},\ E_L = {{\left\| U^N-u^N \right\|}_{\infty }}. \end{equation*}
    Table 5.1.  maximum errors at t_n = 1 and spatial convergence rates for Example 5.1.
    M_1=M_2 4 8 16 32 64
    E 4.5780e-02 1.1335e-02 2.8363e-03 7.1957e-04 1.9089e-04
    rate 2.0140 1.9987 1.9788 1.9144 *

     | Show Table
    DownLoad: CSV

    Global errors and rates in Table 5.2 show that the global temporal convergence order is \alpha_1 . When t_n is far away from 0, results are shown in Table 5.3, and we get the local temporal accuracy O(\tau) . All in all, numerical results are consistent with the theoretical analysis in Remark 1.

    Table 5.2.  global maximum errors and temporal convergence rates for Example 5.1.
    N=M_1^2=M_2^2 \alpha_1=0.3, \alpha_2=0.1 \alpha_1=0.5, \alpha_2=0.3 \alpha_1=0.7, \alpha_2=0.5
    E_G rate E_G rate E_G rate
    128 1.6333e-02 0.1690 1.0008e-02 0.3361 5.2012e-03 0.8184
    256 1.4528e-02 0.1931 7.9278e-03 0.3660 2.9494e-03 0.6130
    512 1.2708e-02 0.1926 6.1514e-03 0.3744 1.9284e-03 0.6260
    1024 1.1120e-02 0.2038 4.7453e-03 0.4003 1.2496e-03 0.6518
    2048 9.6547e-03 * 3.5956e-03 * 7.9532e-04 *

     | Show Table
    DownLoad: CSV
    Table 5.3.  local maximum errors at t_n = 1 and temporal convergence rates for Example 5.1.
    N=M_1^2=M_2^2 \alpha_1=0.3, \alpha_2=0.1 \alpha_1=0.5, \alpha_2=0.3 \alpha_1=0.7, \alpha_2=0.5
    E_L rate E_L rate E_L rate
    128 5.0943e-03 0.8339 5.1615e-03 0.8359 5.2012e-03 0.8374
    256 2.8579e-03 1.0536 2.8916e-03 1.0534 2.9109e-03 1.0531
    512 1.3768e-03 0.9473 1.3933e-03 0.9477 1.4028e-03 0.9480
    1024 7.1403e-04 1.0464 7.2235e-04 1.0460 7.2716e-04 1.0456
    2048 3.4571e-04 * 3.4984e-04 * 3.5227e-04 *

     | Show Table
    DownLoad: CSV

    Example 5.2. Secondly, we consider a two-dimensional three-term time-fractional nonlinear subdiffusion equation with b_1 = b_2 = b_3 = 1 .

    \begin{align} &D_{t}^{\alpha_1}u+D_{t}^{\alpha_2}u+D_{t}^{\alpha_3}u = \Delta u+u(1-u^2)+h(x,y,t),\ (x,y)\in \Omega,\ 0 < t\le T, \end{align} (5.4)
    \begin{align} &u(x,y,0) = g(x,y),\ (x,y)\in \Omega, \end{align} (5.5)
    \begin{align} &u(x,y,t) = 0,\ (x,y)\in \partial \Omega,\ 0\le t\le T, \end{align} (5.6)

    where \Omega = (0, 1)\times (0, 1) and T = 1 . We calculate h(x, y, t) and g(x, y) based on the exact solution

    \begin{equation*} t^{\alpha_1}sin(\pi x)sin(\pi y). \end{equation*}

    Numerical results are shown in Tables 5.4 and 5.5, which verify the theoretical analysis as well.

    Table 5.4.  global maximum errors and temporal convergence rates for Example 5.2.
    N=M_1^2=M_2^2 \alpha_1=0.3, \alpha_2=0.2, \alpha_3=0.1 \alpha_1=0.5, \alpha_2=0.4, \alpha_3=0.3 \alpha_1=0.7, \alpha_2=0.6, \alpha_2=0.5
    E_G rate E_G rate E_G rate
    128 1.7505e-02 0.1822 1.0842e-02 0.3657 4.9545e-03 0.7240
    256 1.5428e-02 0.2052 8.4150e-03 0.3942 2.9994e-03 0.6375
    512 1.3382e-02 0.2040 6.4030e-03 0.4000 1.9282e-03 0.6436
    1024 1.1618e-02 0.2159 4.8524e-03 0.4227 1.2342e-03 0.6639
    2048 1.0003e-02 * 3.6200e-03 * 7.7897e-04 *

     | Show Table
    DownLoad: CSV
    Table 5.5.  local maximum errors at t_n = 1 and temporal convergence rates for Example 5.2.
    N=M_1^2=M_2^2 \alpha_1=0.3, \alpha_2=0.2, \alpha_3=0.1 \alpha_1=0.5, \alpha_2=0.4, \alpha_3=0.3 \alpha_1=0.7, \alpha_2=0.6, \alpha_2=0.5
    E_L rate E_L rate E_L rate
    128 4.8786e-03 0.8333 4.9371e-03 0.8351 4.9545e-03 0.8362
    256 2.7380e-03 1.0534 2.7675e-03 1.0532 2.7751e-03 1.0529
    512 1.3193e-03 0.9470 1.3337e-03 0.9474 1.3376e-03 0.9476
    1024 6.8430e-04 1.0463 6.9158e-04 1.0460 6.9355e-04 1.0455
    2048 3.3134e-04 * 3.3495e-04 * 3.3600e-04 *

     | Show Table
    DownLoad: CSV

    Example 5.3. Thirdly, we investigate the scenario in which the nonlinear term is represented by f(u) = sin(u) , which satisfies Lipschitz condition. The corresponding equation is formulated as follows:

    \begin{align} &D_{t}^{\alpha_1}u+D_{t}^{\alpha_2}u+D_{t}^{\alpha_3}u = \Delta u+sin(u)+h(x,y,t),\ (x,y)\in \Omega,\ 0 < t\le T, \end{align} (5.7)
    \begin{align} &u(x,y,0) = g(x,y),\ (x,y)\in \Omega, \end{align} (5.8)
    \begin{align} &u(x,y,t) = 0,\ (x,y)\in \partial \Omega,\ 0\le t\le T, \end{align} (5.9)

    where \Omega = (0, 1)\times (0, 1) and T = 1 . Similarly, we derive h(x, y, t) and g(x, y) based on the exact solution

    \begin{equation*} t^{\alpha_1}sin(\pi x)sin(\pi y). \end{equation*}

    The corresponding numerical results are presented in Tables 5.6 and 5.7. The global convergence order is \alpha_1 , and the local convergence order is 1 in the temporal direction.

    Table 5.6.  global maximum errors and temporal convergence rates for Example 5.3.
    N=M_1^2=M_2^2 \alpha_1=0.3, \alpha_2=0.2, \alpha_3=0.1 \alpha_1=0.5, \alpha_2=0.4, \alpha_3=0.3 \alpha_1=0.7, \alpha_2=0.6, \alpha_2=0.5
    E_G rate E_G rate E_G rate
    128 1.7753e-02 0.1908 1.0852e-02 0.3664 4.8924e-03 0.7058
    256 1.5554e-02 0.2101 8.4178e-03 0.3945 2.9995e-03 0.6375
    1024 1.1650e-02 0.2175 4.8527e-03 0.4227 1.2342e-03 0.6639
    2048 1.0019e-02 * 3.6201e-03 * 7.7897e-04 *

     | Show Table
    DownLoad: CSV
    Table 5.7.  local maximum errors at t_n = 1 and temporal convergence rates for Example 5.3.
    N=M_1^2=M_2^2 \alpha_1=0.3, \alpha_2=0.2, \alpha_3=0.1 \alpha_1=0.5, \alpha_2=0.4, \alpha_3=0.3 \alpha_1=0.7, \alpha_2=0.6, \alpha_2=0.5
    E_L rate E_L rate E_L rate
    128 5.0640e-03 0.8278 4.9983e-03 0.8256 4.8924e-03 0.8229
    256 2.8530e-03 1.0552 2.8204e-03 1.0557 2.7656e-03 1.0561
    512 1.3730e-03 0.9454 1.3568e-03 0.9449 1.3301e-03 0.9441
    1024 7.1295e-04 1.0478 7.0484e-04 1.0484 6.9132e-04 1.0489
    2048 3.4485e-04 * 3.4079e-04 * 3.3415e-04 *

     | Show Table
    DownLoad: CSV

    Example 5.4. Finally, we consider a two-dimensional two-term time-fractional nonlinear subdiffusion equation with b_1 = b_2 = 1 , whose exact solution is unknown.

    \begin{align} &D_{t}^{\alpha_1}u+D_{t}^{\alpha_2}u = \Delta u+u(1-u),\ (x,y)\in \Omega,\ 0 < t\le T, \end{align} (5.10)
    \begin{align} &u(x,y,0) = \frac{1}{2}sin(\pi x)sin(\pi y),\ (x,y)\in \Omega, \end{align} (5.11)
    \begin{align} &u(x,y,t) = 0,\ (x,y)\in \partial \Omega,\ 0\le t\le T, \end{align} (5.12)

    where \Omega = (0, 1)\times (0, 1) and T = 1 .

    The two-mesh method[26] is applied to compute errors and convergence rates. We take M_1 = M_2 = 60 . E_L is redefined by

    \begin{equation*} E_L = {{\left\| U^N-W^{2N} \right\|}_{\infty }}, \end{equation*}

    in which W^n is the numerical solution of Example 5.4 with \tau = T/2N . The local errors are shown in Table 5.8. The local temporal convergence rate O(\tau) is consistent with Remark 1.

    Table 5.8.  local maximum errors at t_n = 1 and temporal convergence rates for Example 5.4.
    N \alpha_1=0.3, \alpha_2=0.1 \alpha_1=0.5, \alpha_2=0.3 \alpha_1=0.7, \alpha_2=0.5
    E_L rate E_L rate E_L rate
    32 6.8318e-05 1.0203 1.2027e-04 1.0285 1.4785e-04 1.0455
    64 3.3682e-05 1.0103 5.8957e-05 1.0154 7.1633e-05 1.0270
    128 1.6722e-05 1.0053 2.9166e-05 1.0086 3.5152e-05 1.0173
    256 8.3302e-06 1.0027 1.4496e-05 1.0050 1.7367e-05 1.0117
    512 4.1572e-06 * 7.2231e-06 * 8.6132e-06 *

     | Show Table
    DownLoad: CSV

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The research is supported in part by the Natural Science Foundation of Shandong Province under Grant ZR2023MA077, Fundamental Research Funds for the Central Universities (No. 202264006), and the National Natural Science Foundation of China under Grant 11801026.

    The authors declare there are no conflicts of interest.



    [1] A. M. Turing, The chemical basis of morphogenesis, Philos. Trans. R. Soc., B, 237 (1952), 37–72. https://doi.org/10.1098/rstb.1952.0012. doi: 10.1098/rstb.1952.0012
    [2] J. D. Murray, Mathematical Biology II: Spatial Models and Biomedical Applications, Springer, New York, 2001.
    [3] L. E. Keshet, Mathematical Models in Biology. SIAM: Society for Industrial and Applied Mathematics, Philadelphia, 2005. https://doi.org/10.1137/1.9780898719147.
    [4] A. Okubo, S. A. Levin, Diffusion and Ecological Problems: Modern Perspectives, Springer, New York, 2001. https://doi.org/10.1007/978-1-4757-4978-6.
    [5] M. Rietkerk, S. C. Dekker, P. C. D. Ruiter, J. V. D. Koppel, Self-organized patchiness and catastrophic shifts in ecosystems, Science, 305 (2004), 1926–1929. https://doi.org/10.1126/science.1101867. doi: 10.1126/science.1101867
    [6] M. Rietkerk, J. V. D. Koppel, Regular pattern formation in real ecosystems, Trends Ecol. Evol., 23 (2008), 169–175. https://doi.org/10.1016/j.tree.2007.10.013. doi: 10.1016/j.tree.2007.10.013
    [7] L. Segel, J. Jackson, Dissipative structure: an explanation and an ecological example, J. Theor. Biol., 37 (1972), 545–559. https://doi.org/10.1016/0022-5193(72)90090-2. doi: 10.1016/0022-5193(72)90090-2
    [8] S. Levin, L. Segel, Hypothesis for origin of planktonic patchiness, Nature, 259 (1976). https://doi.org/10.1038/259659a0. doi: 10.1038/259659a0
    [9] J. R. Beddington, Mutual interference between parasites or predators and its effect on searching efficiency, J. Anim. Ecol., 44 (1975), 331–340. https://doi.org/10.2307/3866. doi: 10.2307/3866
    [10] R. Arditi, L. R. Ginzburg, Coupling in predator–prey dynamics: ratio dependence, J. Theor, Biol., 139 (1989), 311–326. https://doi.org/10.1016/S0022-5193(89)80211-5. doi: 10.1016/S0022-5193(89)80211-5
    [11] R. M. May, Stability and Complexity in Model Ecosystems, Princeton University Press, 1974.
    [12] D. Alonso, F. Bartumeus, J. Catalan, Mutual interference between predators can give rise to Turing spatial patterns, Ecology, 83 (2002), 28–34. https://doi.org/10.1890/0012-9658(2002)083[0028:MIBPCG]2.0.CO;2. doi: 10.1890/0012-9658(2002)083[0028:MIBPCG]2.0.CO;2
    [13] B. Mukhopadhyay, R. Bhattacharyya, Modeling the role of diffusion coefficients on Turing instability in a reaction-diffusion predator–prey system, Bull. Math. Biol., 68 (2006), 293–313. https://doi.org/10.1007/s11538-005-9007-2. doi: 10.1007/s11538-005-9007-2
    [14] W. Wang, L. Zhang, Y. Xue, Z. Jin, Spatiotemporal pattern formation of Beddington–Deangelis-type predator–prey model, preprint, arXiv: 0801.0797v1.
    [15] S. Fasani, S. Rinaldi, Factors promoting or inhibiting Turing instability in spatially extended prey–predator systems, Ecol. Modell., 222 (2011), 3449–3452. https://doi.org/10.1016/j.ecolmodel.2011.07.002. doi: 10.1016/j.ecolmodel.2011.07.002
    [16] C. S. Holling, The components of predation as revealed by a study of small-mammal predation of the European pine sawfly, Can. Entomol., 91 (1959), 293–320. https://doi.org/10.4039/Ent91293-5. doi: 10.4039/Ent91293-5
    [17] C. S. Holling, Some characteristics of simple types of predation and parasitism, Can. Entomol., 91 (1959), 385–398. https://doi.org/10.4039/Ent91385-7. doi: 10.4039/Ent91385-7
    [18] C. A. Cobbold, F. Lutscher, J. A. Sherratt, Diffusion-driven instabilities and emerging spatial patterns in patchy landscapes, Ecol. Complexity, 24 (2015), 69–81. https://doi.org/10.1016/j.ecocom.2015.10.001. doi: 10.1016/j.ecocom.2015.10.001
    [19] N. Zaker, Population dynamics in patchy landscapes: steady states and pattern formation, PhD thesis, University of Ottawa, 2021.
    [20] D. L. Benson, P. K. Maini, J. A. Sherratt, Diffusion driven instability in an inhomogeneous domain, Bull. Math. Biol., 55 (1993), 365–384. https://doi.org/10.1016/S0092-8240(05)80270-8. doi: 10.1016/S0092-8240(05)80270-8
    [21] D. L. Benson, P. K. Maini, J. A. Sherratt, Analysis of pattern formation in reaction-diffusion models with spatially inhomogeneous coefficients, Math. Comput. Modell., 17 (1993), 29–34. https://doi.org/10.1016/0895-7177(93)90025-T. doi: 10.1016/0895-7177(93)90025-T
    [22] K. Page, P. K. Maini, N. A. M. Monk, Pattern formation in spatially heterogeneous Turing reaction-diffusion models, Phys. D, 181 (2003), 80–101. https://doi.org/10.1016/S0167-2789(03)00068-X. doi: 10.1016/S0167-2789(03)00068-X
    [23] K. Page, P. K. Maini, N. A. M. Monk, Complex pattern formation in reaction-diffusion systems with spatially varying patterns, Phys. D, 202 (2005), 95–115. https://doi.org/10.1016/j.physd.2005.01.022. doi: 10.1016/j.physd.2005.01.022
    [24] M. Kozák, E. A. Gaffney, V. Klika, Pattern formation in reaction-diffusion systems with piecewise kinetic modulation: an example study of heterogeneous kinetics, Phys. Rev. E, 100 (2019), 042220. https://doi.org/10.1103/PhysRevE.100.042220. doi: 10.1103/PhysRevE.100.042220
    [25] E. Sheffer, J. V. Hardenberg, H. Yizhaq, M. Shachak, E. Meron, Emerged or imposed: a theory on the role of physical templates and self-organisation for vegetation patchiness, Ecol. Lett., 16 (2013), 127–139. https://doi.org/10.1111/ele.12027. doi: 10.1111/ele.12027
    [26] O. Ovaskainen, S. Cornell, Biased movement at a boundary and conditional occupancy times for diffusion processes, J. Appl. Probab., 40 (2003), 557–580. https://doi.org/10.1239/jap/1059060888. doi: 10.1239/jap/1059060888
    [27] G. A. Maciel, F. Lutscher, How individual movement response to habitat edge affects population persistence and spatial spread, Am. Nat., 182 (2013), 42–52. https://doi.org/10.1086/670661. doi: 10.1086/670661
    [28] B. Yurk, C. A. Cobbold, Homogenization techniques for population dynamics in strongly heterogeneous landscapes, J. Biol. Dyn., 12 (2018), 171–193. https://doi.org/10.1080/17513758.2017.1410238. doi: 10.1080/17513758.2017.1410238
    [29] C. A. Cobbold, F. Lutscher, B. Yurk, Bridging the scale gap: predicting large-scale population dynamics from small-scale variation in strongly heterogeneous landscapes, Methods Ecol. Evol., 2021. https://doi.org/10.1111/2041-210X.13799. doi: 10.1111/2041-210X.13799
    [30] G. Maciel, C. Cosner, R. S. Cantrell, F. Lutscher, Evolutionarily stable movement strategies in reaction–diffusion models with edge behavior, J. Math. Biol., 80 (2020), 61–92. https://doi.org/10.1007/s00285-019-01339-2. doi: 10.1007/s00285-019-01339-2
    [31] Y. Alqawasmeh, F. Lutscher, Persistence and spread of stage-structured populations in heterogeneous landscapes, J. Math. Biol., 78 (2019), 1485–1527. https://doi.org/10.1007/s00285-018-1317-8. doi: 10.1007/s00285-018-1317-8
    [32] P. Turchin, Quantitative Analysis of Movement: Measuring and Modeling Population Redistribution of Plants and Animals, Sinauer Associates, Sunderland, 1998.
    [33] P. Turchin, Complex Population Dynamics, Princeton University Press, 2001.
    [34] E. E. Crone, L. M. Brown, J. A. Hodgson, F. Lutscher, C. B. Schultz, Faster movement in nonhabitat matrix promotes range shifts in heterogeneous landscapes, Ecology, 100 (2019), e02701. https://doi.org/10.1002/ecy.2701. doi: 10.1002/ecy.2701
    [35] J. C. Strikwerda, Finite difference schemes and partial differential equations, SIAM: Society for Industrial and Applied Mathematics, Philadelphia, 2004. https://doi.org/10.1137/1.9780898717938.
    [36] A. L. Krause, V. Klika, T. E. Woolley, E. A. Gaffney, From one pattern into another: analysis of Turing patterns in heterogeneous domains via WKBJ, J. R. Soc. Interface, 17 (2020), 20190621. https://doi.org/10.1098/rsif.2019.0621. doi: 10.1098/rsif.2019.0621
    [37] M. J. Garlick, J. A. Powell, M. B. Hooten, L. R. McFarlane, Homogenization of large-scale movement models in ecology, Bull. Math. Biol., 73 (2011), 2088–2108. https://doi.org/10.1007/s11538-010-9612-6. doi: 10.1007/s11538-010-9612-6
    [38] B. Yurk, Homogenization of a directed dispersal model for animal movement in a heterogeneous environment, Bull. Math. Biol., 78 (2016), 2034–2056. https://doi.org/10.1007/s11538-016-0210-0. doi: 10.1007/s11538-016-0210-0
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4138) PDF downloads(140) Cited by(1)

Figures and Tables

Figures(16)  /  Tables(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog