Loading [MathJax]/jax/output/SVG/jax.js
Research article

Pattern dynamics and Turing instability induced by self-super-cross-diffusive predator-prey model via amplitude equations

  • Received: 14 September 2022 Revised: 22 October 2022 Accepted: 01 November 2022 Published: 14 November 2022
  • MSC : 34C23, 34K18, 35B36, 37G15, 37L10, 49N75, 60J60, 65L12, 70K50

  • Incorporating self-diffusion and super-cross diffusion factors into the modeling approach enhances efficiency and realism by having a substantial impact on the scenario of pattern formation. Accordingly, this work analyzes self and super-cross diffusion for a predator-prey model. First, the stability of equilibrium points is explored. Utilizing stability analysis of local equilibrium points, we stabilize the properties that guarantee the emergence of the Turing instability. Weakly nonlinear analysis is used to get the amplitude equations at the Turing bifurcation point (WNA). The stability analysis of the amplitude equations establishes the conditions for the formation of small spots, hexagons, huge spots, squares, labyrinthine, and stripe patterns. Analytical findings have been validated using numerical simulations. Extensive data that may be used analytically and numerically to assess the effect of self-super-cross diffusion on a variety of predator-prey systems.

    Citation: Naveed Iqbal, Ranchao Wu, Yeliz Karaca, Rasool Shah, Wajaree Weera. Pattern dynamics and Turing instability induced by self-super-cross-diffusive predator-prey model via amplitude equations[J]. AIMS Mathematics, 2023, 8(2): 2940-2960. doi: 10.3934/math.2023153

    Related Papers:

    [1] Zuojun Ma . Pattern formation of a volume-filling chemotaxis model with a bistable source. AIMS Mathematics, 2024, 9(11): 30816-30837. doi: 10.3934/math.20241488
    [2] Rongjie Yu, Hengguo Yu, Min Zhao . Steady states and spatiotemporal dynamics of a diffusive predator-prey system with predator harvesting. AIMS Mathematics, 2024, 9(9): 24058-24088. doi: 10.3934/math.20241170
    [3] Wei Li, Qingkai Xu, Xingjian Wang, Chunrui Zhang . Dynamics analysis of spatiotemporal discrete predator-prey model based on coupled map lattices. AIMS Mathematics, 2025, 10(1): 1248-1299. doi: 10.3934/math.2025059
    [4] Xiao-Long Gao, Hao-Lu Zhang, Xiao-Yu Li . Research on pattern dynamics of a class of predator-prey model with interval biological coefficients for capture. AIMS Mathematics, 2024, 9(7): 18506-18527. doi: 10.3934/math.2024901
    [5] Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra . Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045
    [6] Weiyu Li, Hongyan Wang . The instability of periodic solutions for a population model with cross-diffusion. AIMS Mathematics, 2023, 8(12): 29910-29924. doi: 10.3934/math.20231529
    [7] Lingling Li, Xuechen Li . The spatiotemporal dynamics of a diffusive predator-prey model with double Allee effect. AIMS Mathematics, 2024, 9(10): 26902-26915. doi: 10.3934/math.20241309
    [8] Jing Zhang, Shengmao Fu . Hopf bifurcation and Turing pattern of a diffusive Rosenzweig-MacArthur model with fear factor. AIMS Mathematics, 2024, 9(11): 32514-32551. doi: 10.3934/math.20241558
    [9] Anna Sun, Ranchao Wu, Mengxin Chen . Turing-Hopf bifurcation analysis in a diffusive Gierer-Meinhardt model. AIMS Mathematics, 2021, 6(2): 1920-1942. doi: 10.3934/math.2021117
    [10] Heping Jiang . Complex dynamics induced by harvesting rate and delay in a diffusive Leslie-Gower predator-prey model. AIMS Mathematics, 2023, 8(9): 20718-20730. doi: 10.3934/math.20231056
  • Incorporating self-diffusion and super-cross diffusion factors into the modeling approach enhances efficiency and realism by having a substantial impact on the scenario of pattern formation. Accordingly, this work analyzes self and super-cross diffusion for a predator-prey model. First, the stability of equilibrium points is explored. Utilizing stability analysis of local equilibrium points, we stabilize the properties that guarantee the emergence of the Turing instability. Weakly nonlinear analysis is used to get the amplitude equations at the Turing bifurcation point (WNA). The stability analysis of the amplitude equations establishes the conditions for the formation of small spots, hexagons, huge spots, squares, labyrinthine, and stripe patterns. Analytical findings have been validated using numerical simulations. Extensive data that may be used analytically and numerically to assess the effect of self-super-cross diffusion on a variety of predator-prey systems.



    Pattern formation has captivated scholars in recent decades as a way to better understand natural processes, and Turing's key work on reaction-diffusion systems gave rise to the concept of using it with reaction-diffusion systems [36]. Turing instability causes a variety of spatial structures to arise, including spots, spirals, squares, hexagons, combinations of spots and stripes, and so on [2,3,14,28]. Non-Turing patterns, such as traveling wave, periodic traveling wave, spatiotemporal chaos, and so on, can be detected using diffusion concepts [9,25]. Many of these patterns in the spatio-temporal expansion of interactive population systems, which include competing species and predator-prey systems, were discovered using reaction-diffusion equations [2,3,25].

    Cross-diffusion is vital in the emergence of patterns in the disciplines of chemistry, biology, and ecology. In ecology, the concentration gradient of one species influences the flow of another; this phenomena is known as cross-diffusion in the perspective of the reaction-diffusion model, as detailed in [30,37]. Positive cross-diffusion distributes the respective components, whilst negative cross-diffusion causes them to accumulate locally.The influence of cross-diffusion on excitatory chemical systems has already been discussed extensively for Chemical self-replication according to the Templetor model [7], the chlorite-iodide-malonic acid-starch autocatalytic reaction system [35], the Fitzhugh-Nagumo type reaction kinetics [40], and Stokes equations [5]. Vanag and Epstein [37] investigated the effects of cross-diffusion in cognitive and various chemical systems in particular. Cross-diffusion terms can be used to characterize prey's tendency to avoid predators and also the predator population's weakening in the presence of high-density prey regions [18,19,22]. The predator-prey pursuit-evasion technique is another cross-diffusion paradigm wherein waves propagate via taxis described by nonlinear cross-diffusion parameters.

    Super-diffusion may be seen all over the place in nature. The limiting output of Lˊevyflight is represented at the molecular level, with the jumping length dispersion containing infinite intervals during the relative motion of the individual particles. The particle conducts large hops at this moment. Surface diffusion, turbulence, animal and plasma motion, and other processes can all be affected by these phenomena. They might also have a big influence on pattern creation. The influence of the super diffusion implications on the several models, like the activator-inhibitor model [43], the Fitzhugh-Nagumo model [15], the Lengyel-Epstein model [23], and the complex fractional-order HIV diffusion model [16]] has lately been noted by many investigators. In a predator-prey model [24] and a three-species food chain model with harvesting [17], only a few outcomes have been reported around super-cross-diffusion. As stated in the introduction, we will investigate the Hopf bifurcation along with the Turing patterns described for the predator-prey system with in the self-super-cross diffusion in this research.

    For decades, Turing instability [36] has piqued the curiosity of scientists as a frequent obstacle in the development of morphological patterns. In reaction-diffusion systems, mathematical analysis and numerical simulations have given a range of results on pattern formation. Song et al. [31] investigated the generation of Turing patterns via the Gierer-Meinhardt model in two spatial dimensions with a saturation term. They obtained the spotting, striping, cohabitation, and labyrinth patterns. The results showed the mechanism of morphogenetic processes in mesenchymal cells that have reached maturity. In both the supercritical and subcritical circumstances, the amplitude equations of the stationary pattern were found in the Turing domain.

    Regarding the parameter values close to the Turing bifurcation curve [4,9,21], amplitude equation is critical in explaining the behavior of a reaction-diffusion system. The system's homogeneous steady state loses its stability due to tiny amplitude heterogeneous disturbances [25]. The system exhibits a critical slowdown at a bifurcation threshold, indicating that it takes a substantial amount of time to switch from a stable steady-state homogeneous condition to an inhomogeneous configuration [39].

    As a result, multi-scale perturbation analysis may be used for deriving amplitude equations for the studying of the dynamics of active slow modes. The amplitude equations' stability analysis offers information on possible patterns that would occur near the boundary of Turing bifurcation when the amplitude equations are extracted out from the reaction-diffusion equations original system. This technique may also be used to learn about pattern selection and stability around the onset [41]. Zemskov et al. [40] investigated pattern formation in the Oregonator and Brusselator models involving cross-diffusion via WNA. Topaz and Catla [33] and Chen [6] employed this approach to explore patterns created around a Turing-Hopf bifurcation point, which occurs whenever the system seems unstable because both temporal and spatial disturbances are considered. The amplitude equations were established to analyze the production of spatial patterns utilizing WNA, such as rhomboid, squares and rolls, in an epithelial model theory for insect, mammal, and fish skin pattern generation. The approach was further modified for the spatial and temporal expansion of interactive population models, including the cross-diffusion augmented competitive model [12], the hyperbolic mortality model [42], Beddington-DeAngelis functional responsiveness in the predator-prey model [41], and also the herd behavior model [39].

    Numerical simulations might be used to better understand the influence of diffusion, as a consequence of which represent intriguing phenomena like stationary, pulses, fronts, and periodic patterns, among others [7]. Cross-diffusion and self-diffusion are put into consideration in interactive population modeling including prey-predator models and competitive models [32], interesting spatiotemporal patterns emerge. Numerical simulations will be used to validate the analytical results.

    The following is a breakdown of how this article is structured. The temporal and spatiotemporal predator-prey model will be introduced in the next section. We present the conditions that ensure that the Turing instability arises in Section 3 by examining the local equilibrium points' stability. In Section 4, we use WNA to find the amplitude equations, and in Section 5, we use numerical simulations to test the validity of the analytical findings. Finally, we present conclusions derived from our paper.

    Differential equations are used in predation theory when it can be assumed that generations overlap and populations fluctuate continuously over time. For this method to be fair, the time scale must be set in accordance with the species being investigated. For the majority of mammals, variation should be represented on an annual basis, but for many planktonic creatures, a daily scale is necessary. In situations when the dynamics of a population cannot be approximated by continuous functions, difference equations may be preferable. The Lotka-Volterra model, from which the idea of predation evolved, contains a number of generally acknowledged flaws. In spite of this, differential equations have remained an essential tool in the following development of the theory, and this method penetrates a significant portion of contemporary ecological thought.

    The fundamental model for predator-prey interaction may be expressed using nonlinear coupled ordinary differential equations [1,25]. Corresponding to Gause [10] for predator–prey interaction subject to nonnegative initial conditions is defined as:

    dUdt=Uf(U)Vg(U)dVdt=Bg(U)VμV, (2.1)

    where UU(t) and VV(t) are the density of prey and predator at time t, respectively. μ, supposed to be [1] constant, is the intrinsic mortality rate of predators and B(0<B<1) is the food conversion efficiency of the predator. Additionally, g(U) is the prey based functional response[11] and f(U) is the per capita rate to grow the prey (when the predation is absent). In this paper, we assume that f(U) follows logistic growth law and takes the form

    f(U)=A(1Ub),

    where b is the environmental carrying capacity for the prey population and A is the intrinsic density–independent growth rate. The notion of employing the prey dependent functional response was suggested by Feng and Kang [11] and Tian et al. [32] that is

    g(U)=μUU+H,

    where "gamma(μ)" indicates the predator's attack rate and H denotes the half-saturating constant. Combining f(U) and g(U), hence the system (2.1) takes the form

    dUdt=AU(1Ub)μUVU+HdVdt=BμUVU+HμV, (2.2)

    with regard to the initial conditions U(0),V(0)>0 for (U,V)(0,0). When (U,V)=(0,0). The system described above has been simplified to dUdt=dVdt=0.

    ˜t=At,˜V=μVbA and ˜U=Ub is taken for the non–dimensional time, predator and prey population density, respectively. For the sake of simplicity, we drop the over–bars to get the following non–dimensional model

    dUdt=U(1U)UVU+hdVdt=C(UU+h)VDV, (2.3)

    where C=BμA, D=μA, and h=Hb are all positive parameters. The nonnegative solutions of

    U(1U)UVU+h=β(U,V)=0C(UU+h)VDV=γ(U,V)=0, (2.4)

    are the steady state solutions of (2.3). Aside from the general solution (0,0) (complete extinction), (2.4) yields the predator-free steady state point (1,0) as well as the coexisting equilibrium point L(U=DhCD,V=(1U)(U+h)) in the first quadrant's interior.

    The temporal model (2.3) fits the spatiotemporal model having self-super-cross-diffusion terms. It is necessary to understand how the pattern formation phenomenon takes place in a diffusive model that is anomalous. The unsustainably diffusive operator [13,29,38] is expressed by (using the Fourier transform) (Δ)μ2 with the following definition:

    F{(Δ)μ2}(ν)=|ν|μFu(ν),νRn,μ>0.

    When 1<μ<2, the above operator is known as super-diffusion. The associated random walk's displacement moment develops slower (sub-diffusion) and faster (super-diffusion) than in regular diffusion. For μ=2, we have the Laplacian operator Δ as a special form of the above operator. Here, the predator-prey system is investigated on the basis of self-super-cross-diffusion in a bounded domain ΩR2, with a smooth boundary Ω, as follows:

    Utd12Ud2μUV=U(1U)UVU+h,(x,y)Ω,t>0,Vtd3μUV=CUVU+hDV,(x,y)Ω,t>0,Un=Vn=0,(x,y)ΩU(x,y,0)>0,V(x,y,0)>0,(x,y)Ω, (2.5)

    d1 is indeed the prey's diffusion coefficient, d2 is a coefficient of the cross diffusion that anticipates the movement of prey species based on predator population [19,26,38], and d3 is a coefficient of the cross diffusion in which the predator species movement is influenced by prey population. The unit normal vector along Ω in outward direction is denoted by n. There is no flux via the boundary when the Neumann boundary conditions are homogeneous.

    The alternative obtaining of the fractional operator μ for (1<μ<2) can be done as follows:

    μa=μξμ=μa|y|μ+μa|x|μ=12cos(πμ/2)(RLDμ,ya+RLDμy,+a)12cos(πμ/2)(RLDμ,xa+RLDμx,+a),

    where RLDμ,y and RLDμy,+, , in respective order, as given below:

    RLDμ,ya=1Γ(2μ)2y2y(ys)1μa(s,x,t)ds,RLDμy,+a=1Γ(2μ)2y2+y(sy)1μa(s,x,t)ds,

    with Γ(.) refers to the Gamma function.

    Turing instability takes place as a stable homogeneous steady-state turns out to be unstable in the presence of tiny amplitude heterogeneous disturbances surrounding it [25,36]. With the cross-diffusion term d3, the equilibrium point in (2.5), which itself is stable like a solution of the problem (2.3) in the absence of the cross-diffusion term d3, becomes unstable. Suppose U(x,y,t) and V(x,y,t) denote the homogeneous steady states that satisfy these equations for (2.5). Suppose that under temporal perturbation for Turing instability, asymptotically, the homogeneous steady state is stable. This means that the requirements j11j22j12j21>0 and j11+j22<0 are met. Linear stability analysis for the spatiotemporal model (2.5) at L(U,V) is used to identify the constraints of the Turing instability [25].

    Theorem 3.1. System (2.3) has a single positive equilibrium L(U,V) if DC, and L is asymptotically locally stable.

    Proof. The Jacobian J0 of the system (2.3) at the coexisting equilibrium point L(U,V), is given by

    J0=(DC(1h(C+D)CD)DCCDDh0)(j11j12j21j22). (3.1)

    Assume that h is the model's Hopf bifurcation parameter (2.3). By solving

    j11+j22=0, (3.2)

    we obtain

    hH=CDC+D. (3.3)

    The equilibrium point L(U,V) that co-exists in relation with the parameter h is loses stability by Hopf bifurcation at h=hH and locally asymptotically stable for h>hH.

    We perturb the system (2.3) around L(U,V) as

    [U(x,y,t)V(x,y,t)]=[UV]exp(λt). (3.4)

    The characteristic equation for the growth rate λ is as follows, where λ refers to the rate of perturbation growth. As a result of the replacement of (3.4) into (2.3), the growth rate's characteristic equation "(λ)" is as follows:

    det(J1)=|j11λj12j21j22λ|=0. (3.5)

    Hence

    λ2λj11j12j21=0. (3.6)

    Both j12j21 and j11 may be easily verified as negative. As a result, the two roots λ1,2 have their own negative real components. The completion of the proof is done accordingly.

    In the following, the stability of L(U,V) is to be examined in system (2.5) with no cross–diffusion term d3, having the following,

    Utd12Ud2μUV=U(1U)UVU+h,(x,y)Ω,t>0,Vt=CUVU+hDV,(x,y)Ω,t>0,Un=Vn=0,(x,y)ΩU(x,y,0)>0,V(x,y,0)>0,(x,y)Ω. (3.7)

    Theorem 3.2. Let's say the wave number vector is νν=(ν1,ν2), and the wave number is ν=|νν|. If DC, then L(U,V) of system (3.7) is asymptotically and locally stable without the cross-diffusion factor d3.

    Proof. The system (3.7) is perturbed about L(U,V) as follows:

    [U(x,y,t)V(x,y,t)]=[UV]exp(λt+i(ν1x+ν2y)), (3.8)

    in which λ refers to the perturbation growth rate. By the substitution of (3.8) into (3.7), we get

    det(J2)=|j11λd1k2d2|ν|μVj12d2|ν|μUj21j22λ|=0. (3.9)

    The formulae for j11,j12,j21 and j22 may be discovered in (3.1). As a result, the characteristic equation becomes

    λ2+λTν+hν=0, (3.10)

    in which

    Tν=(j11ν2d1|ν|μ(d2V))hν=(j21d2U)|ν|μj12j21.

    Clearly, hν and Tν are both positive. As a result, the two roots λ1,2 have their own negative real components. The completion of the proof is done accordingly.

    The perturbation all around homogeneous steady state L(U,V) for system (2.5) as

    [U(x,y,t)V(x,y,t)]=[UV]exp(λt+i(ν1x+ν2y)). (3.11)

    By the substitution of (3.11) into (2.5), we get

    det(J3)=|j11λd1k2d2|ν|μVj12d2|ν|μUj21d3|ν|μVj22λd3|ν|μU|=0. (3.12)

    Hence, the characteristic equation is

    λ2+λTν+hν=0, (3.13)

    in which

    Tν=(j11ν2d1|ν|μ(d2V+d3U)),hν=(d1d3U)ν2|ν|μ+(j12d3Vj11d3U+j21d2U)|ν|μj12j21.

    Checking that Tν is positive is easy. Regarding d3 as the bifurcation parameter, we investigate the Turing bifurcation threshold for ννT at d3=dT3 such that hTν=0. The hν-derivative is assessed in regard to νμ at ν=νT, and using dhνdνμ|ν=νT=0, one can obtain νT=μ2+μ(j11d3Uj12d3Vj21d2Ud1d3U). Through the substitution of the value of νT in hν it is possible to find the value of d3 such that hν<0 suggesting that the cross–diffusion term d3 alters an equilibrium point's stability in the way presented in Figure 1.

    Figure 1.  Illustration of hν versus d3, where μ=1.5, h=0.45, D=0.8, C=2, d1=0.1, d2=0.1.

    When j11+j22<0, j11j22j12j21>0 and d3<dT3, the homogeneous steady state L(U,V) is stable in the presence of spatiotemporal or heterogeneous perturbation. When j11j22j12j21>0, j11+j22<0 and d3>dT3, the homogeneous steady state L becomes unstable.

    For the model parameters μ=1.5, h=0.45, D=0.8, d1=0.1, d2=0.1, C=2 and dT3>d3=1 the real component of the eigenvalue "λ" becomes negative (Figure 2). For heterogeneous disturbances, this implies the stability of a homogeneous stable state. When dT3<d3 is taken into account, the matching curves (black, pink, blue, green) in Figure 2 show that inside the interval of |νν|, where the system appears unstable to diversified perturbations and creates Turing patterns, the largest real fraction of the eigenvalue is positive. However, we have no way of knowing which Turing patterns were chosen. The stability of various forms of Turing patterns and also structural transformations among them will be interpreted using the amplitude equations of Turing patterns at inception d3=dT3.

    Figure 2.  The system's dispersion relations (2.5) in numerous forms.

    The system's dynamics change very slowly near the Turing bifurcation threshold. The slow modes get implicated in this condition, and the amplitude equations [39,40,41] may be used to analyze pattern creation. We create amplitude equations and analyze the stability selection of various patterns by employing multiple-scale perturbation assessment such as labyrinthine, squares, spots, and mixes of stripes and spots. We take into consideration three pairs of active resonant modes (ννj,ννj)(j=1,2,3) and making angles of 2π/3 with |ννj|=νT [39,41]. First, by putting the perturbations a=UU and b=VV about L(U,V), we may construct the linearized version of model (2.5) as follow:

    t(ab)=L(ab)+12(βaaa2+2βabab+βbbb2γaaa2+2γabab+γbbb2)+(d2μabd3μab)+16(βaaaa3+3βaaba2b+3βabbab2+βbbbb3γaaaa3+3γaaba2b+3γabbab2+γbbbb3), (4.1)

    in which L is the linear operator that is give as

    L=(j11+d12+d2μVj12+d2μUj21+d3μVj22+d3μU). (4.2)

    The results of (2.5) can be expanded as follows (at the initiation of Turing instability):

    (ab)=3j=1[Xjexp(iνjr)+¯Xjexp(iνjr)], (4.3)

    where Xj, Xj are the relative amplitudes of the modes ννj, ννj.

    We alter the bifurcation parameter d3 with a,b,t as close to the Turing bifurcation threshold as possible by writing

    a=ϵa1+ϵ2a2+ϵa3+,b=ϵb1+ϵ2b2+ϵb3+,t=t0+ϵt1+ϵ2t2+,d3=dT3+ϵd(1)3+ϵ2d(2)3+. (4.4)

    This yields

    L=LT+ϵ(00μVμU)d(1)3+ϵ2(00μVμU)d(2)3+, (4.5)

    where

    LT=(j11+d12+d2μVj12+d2μUj21+dT3μVj22+dT3μU). (4.6)

    The amplitude, indicated by Xj(j=1,2,3), is treated as a variable which shows slow change with respect to time Xjt=0 [39,41]. Thus,

    Xjt=ϵXjt+ϵ2Xjt+O(ϵ3). (4.7)

    To derive the amplitude equations: Substituting (4.4) into (4.1) and collecting the like power coefficients of "ϵ". The linear system at O(ϵ) is obtained as follows:

    LT(a1b1)=0. (4.8)

    Due to the fact that (a1,b1)t is the linear combination of the eigenvectors correlating to the eigenvalue 0, and LT is the system's linear operator at the Turing bifurcation threshold. After solving (4.8), one can get

    (a1b1)=3j=1(N1)(Yjexpiνjr+c.c), (4.9)

    where N=(d3νμTU)j21d3νμTV and amplitude Yj of mode expiνjr. At O(ϵ2), we obain

    LT(a2b2)=t1(a1b1)d(1)3(00VU)μ(a1b1)(d2dT3)μ(a1b1)12(βaaa21+2βaba1b1+βbbb21γaaa21+2γaba1b1+γbbb21)=(FxFy). (4.10)

    The Fredholm solvability criterion stipulates that to assure the nontrivial solution existence (4.10), on the right side of Eq (4.10), the vector function ought to be orthogonal to the eigenvectors of the zero eigenvalue of LT+ (LT+ stands for adjoint of LT). The eigenvectors of the operator LT+ are (1M)expiνjr+c.c.(j=1,2,3), in which M=j11d1ν2Td2νμTVdT3νμTj12. The condition of orthogonality is (1,M)(FjxFjy)=0, in which Fjx and Fjy denote the coefficients of expiνjr term, in respective order. For example, by the substitution of (4.9) into (4.10) and equation of the coefficient of expiν1r, we get

    (F1xF1y)=(N1)Y1t1d(1)3(00VU)(νμT)(N1)Y1N(νμT)(d2dT3)2¯Y2¯Y312(β1γ1)2¯Y2¯Y3, (4.11)

    where

    (β1γ1)=(βaaN2+2βabN+βbbγaaN2+2γabN+γbb).

    We obtain the following result when we use the solvability criterion.

    (M+N)(Y1t1)=νμTd(1)3M(VN+U)Y1+((β1+Mγ1)(2NνμT)(d2+MdT3))¯Y2¯Y3. (4.12)

    Taking the coefficients of expiν2r and expiν3r, we obtain the following relationships.

    (M+N)(Y2t1)=νμTd(1)3M(VN+U)Y2+((β1+Mγ1)(2NνμT)(d2+MdT3))¯Y3¯Y1, (4.13)
    (M+N)(Y3t1)=νμTd(1)3M(VN+U)Y3+((β1+Mγ1)(2NνμT)(d2+MdT3))¯Y1¯Y2. (4.14)

    Equation (4.10) has the following solution:

    (a2b2)=(κ00)+3m=1((κmm)expiνmr+(κmmmm)expi2νmr)+(κ1212)expi(ν1ν2)r+(κ2323)expi(ν2ν3)r+(κ3131)expi(ν3ν1)r+c.c.. (4.15)

    The substitution of Eq (4.15) into Eq (4.10) and collection of the coefficients of exp2iνjr gives

    (j1122ν2Td12μνμTd2Vj122μνμTd2Uj212μνμTd3Vj222μνμTd3U)(κ1111)=(d2dT3)(2μνμT)NY2112(β1γ1)Y21, (4.16)

    that provides

    (κ1111)=(j1122ν2Td12μνμTd2Vj122μνμTd2Uj212μνμTd3Vj222μνμTd3U)1×(2μνμTd2Nβ122μνμTdT3Nγ12)Y1(ξx1ξy1)Y21. (4.17)

    The coefficients of components of (4.15) involving exp0, expiνmr and expi(ν1ν2)r are calculated similarly and given as

    (a0b0)=(j11j12j21j22)1(νμTd2Nβ12νμTdT3Nγ12)(|Y1|2+|Y2|2+|Y3|2)(ξx0ξy0)(|Y1|2+|Y2|2+|Y3|2), (4.18)
    κm=Nnm=1,2,3, (4.19)
    (κ1212)=(j11(3νT)2d1(3νT)μd2Vj12(3νT)μd2Uj21(3νT)μd3Vj22(3νT)μd3U)1×((3νT)μd2Vd2Nβ12(3νT)μd2VdT3Nγ12)2Y1¯Y2(ξx2ξy2)2Y1¯Y2. (4.20)

    The coefficients of the components of (4.15) corresponding to expi(ν2ν3)r and expi(ν3ν1)r are obtained by permuting the suffixes. At O(ϵ3), we obtain

    LT(a3b3)=(GxGy), (4.21)

    in which

    (GxGy)=(a2t1+a1t2b2t1+b1t2)d(1)3(00VU)μ(a2b2)d(2)3(00VU)μ(a1b1)(d2dT3)μ(a1b2+a2b1)(0d(1)3)μ(a1b1)12(βaa(a1a2)+2βab(a1b2+a2b1)+βbb(b1b2)γaa(a1a2)+2γab(a1b2+a2b1)+γbb(b1b2))16(βaaaa31+3βaaba21b1+3βabba1b21+βbbbb31γaaaa31+3γaaba21b1+3γabba1b21+γbbbb31). (4.22)

    The coefficients for expiν1r can be found by collecting them from (4.22).

    (G1xG1y)=(N(B1t1+Y1t2)B1t1+Y1t2)+d(1)3νμT(00VU)(NB1B1)+d(2)3νμT(00VU)×(NY1Y1)+νμT(d2dT3)((Nξy0+ξx0+Nξy1+ξx1)|Y1|2Y1+(Nξy0+ξX0+Nξy2+ξx2)(|Y2|2+|Y3|2)Y1+2N(¯Y2¯b3¯Y3¯N2))+2νμT(0d(1)3)¯Y2¯Y3(((βaaN+βab)(ξx0+ξx1)+(βabN+βbb)(ξy0+ξy1))|Y1|2+((βaaN+βab)×(ξx0+ξx2)+(βabN+βbb)(ξy0+ξy2))(|Y2|2+|Y3|2)|Y1|+β1(¯Y2¯b3+¯Y3¯B2)((γaaN+γab)(ξx0+ξx1)+(γabN+γbb)(ξy0+ξy1))|Y1|2+((γaaN+γab)×(ξx0+ξx2)+(γabN+γbb)(ξy0+ξy2))(|Y2|2+|Y3|2)|Y1|+γ1(¯Y2¯b3+¯Y3¯B2))((|Y1|2+|Y2|2+|Y3|2)(βaaaN3+3βaabN2+3βabbN+βbbb)(|Y1|2+|Y2|2+|Y3|2)(γaaaN3+3γaabN2+3γabbN+γbbb))Y1. (4.23)

    In the case of O(ϵ2), the Fredholm solvability criterion is used, we have (1,M)(GjxGjy)=0. In simplified form, it gives

    (M+N)(B1t1+Y1t2)=νμTM(VN+U)(d(1)3B1+d(2)3Y1)+E1(¯Y2¯b3+¯Y3¯B2)+E2¯Y2Y3(B1|Y1|2+B2(|Y2|2+|Y3|2))Y1, (4.24)

    in which

    β2=βaaaN3+3βaabN2+3βabbN+βbbb,γ2=γaaaN3+3γaabN2+3γabbN+γbbb,E1=(β1+Mγ1)2NνμT(d2+MdT3),E2=2MNνμTd(1)3,B1=(νμT(d2+MdT3)(Nξy0+ξx0+Nξy1+ξx1)+(βaaN+βab+M(γaaN+γab))×(ξx0+ξx1)+(βabN+βbb+M(γabN+γbb))(ξy0+ξy1)+(β2+Mγ2)),B2=(νμT(d2+MdT3)(Nξy0+ξx0+Nξy2+ξx2)+(βaaN+βab+M(γaaN+γab))×(ξx0+ξx2)+(βabN+βbb+M(γabN+γbb))(ξy0+ξy2)+(β2+Mγ2)).

    The permutation of Y's subscript can be used to determine the other two equations. Xm=ϵYm+ϵ2bm+O(ϵ3) can be used to extend the amplitude Xm(m=1,2,3).

    We can correlate the amplitude equation to X1 by using Xm expression and (4.7).

    0X1t=ηX1+E3¯X2¯X3(G1|X1|2+G2(|X2|2+|X3|2))X1, (4.25)

    in which

    η=d3dT3dT3,0=M+NdT3νμTM(VN+U),E3=E1+μ¯E2dT3νμTM(VN+U),¯E2=2MNνμTd(1)3,G1=B1dT3νμTM(VN+U),G2=B2dT3νμTM(VN+U).

    The other two equations may be obtained in a similar way by permuting the subscript of X:

    0X2t=ηX2+E3¯X3¯X1(G1|X2|2+G2(|X3|2+|X1|2))X2, (4.26)
    0X3t=ηX3+E3¯X1¯X2(G1|X3|2+G2(|X1|2+|X2|2))X3. (4.27)

    Stability analysis of amplitude equation

    There are two parts to the amplitude Eq (4.25): a mode fm=|Xm| and a phase Θm. By separating the imaginary and real components and putting Xm=fmexpiΘm into (4.25) the four real-valued differential equations that proceed are generated.

    0Θt=E3f21f22+f22f23+f23f21f1f2f3sinΘ,0f1t=ηf1+E3f2f3cosΘG1f31G2(f22+f23)f1,0f2t=ηf2+E3f3f1cosΘG1f32G2(f23+f21)f2,0f3t=ηf3+E3f1f2cosΘG1f33G2(f21+f22)f3, (4.28)

    in which Θ1+Θ2+Θ3=Θ.

    Properties

    The solution of the dynamical system (4.28) is found as follows:

    (1) f1=f2=f3=0 is the stationary state, which is stable for η<η2=0. As a consequence, for η>η2 we get a spatial pattern and for η<η2 we get a stable homogeneous steady state.

    (2) The striped pattern produced by f1=ηG10, f2=f3=0 is stable for η>η3=E22G1(G2G1)2 and instability occurs for η<η3.

    (3) The mixed state is given by f1=|E2|G1G2, f2=f3=η+G1f12(G1+G2), is always unstable.

    (4) Hexagon pattern represented by f1=f2=f3=|E2|±E22+4(G1+2G2)η2(G1+2G2) exists when η>η1=E224(G1+2G2). The solution ρ=|E2|+E22+4(G1+2G2)η2(G1+2G2) is stable for η<η4=E222G1+G2(G2G1)2, otherwise unstable.

    Therefore, the amplitude equations show how distinct patterns emerge when different thresholds of η are achieved. At this point, we will look at the numerical simulation results and make a comparison with the underlying theory in Section 4.

    For the purpose of demonstrating the theoretical analysis, a numerical simulation is presented. To verify our numerical approach with the space fractional reaction-diffusion model, we have used fractional exponential integrator methodology [8,20,27,34]. Furthermore, in our simulation, boundary constraints in the spatial domain are imposed to verify that Turing patterns are produced. All the simulations were run at N=60. The numerical graphs were made using the Matlab software. The 2019b version of the Matlab was used for the numerical simulation.

    Existence of hexagonal patterns can be seen from (4) where C=3, D=1, h=0.45, μ=1.5, d3=6, d2=1 and d1=1 initially perturbed as a=0.1+0.4cos(y)sin(x), b=0.5+0.1cos(y)sin(x) with iterations (ⅰ) i = 50, (ⅱ) i = 60, (ⅲ) i = 70 and (ⅳ) i = 80 (see Figure 3).

    Figure 3.  Illustrations of hexagonal pattern.

    Existence of big spots can be seen from (4) where C=8, D=2, h=0.45, μ=1.5, d3=10, d2=1 and d1=1 initially perturbed as a=0.1+0.5cos(y)cos(x), b=0.1+0.5cos(y2)sin(x2) with iterations (ⅰ) i = 60, (ⅱ) i = 100, (ⅲ) i = 120 and (ⅳ) i = 140 (refers to Figure 4).

    Figure 4.  Illustrations for big spot patterns.

    Existence of small spots can be seen from (4) where C=2, D=0.8, h=0.45, μ=1.5, d3=4, d2=0.1 and d1=0.1 initially perturbed as a=0.5+0.1cos(y)cos(x), b=0.7+0.1cos(y)cos(x) with iterations (ⅰ) i = 100, (ⅱ) i = 110, (ⅲ) i = 200 and (ⅳ) i = 350 (refers to Figure 5).

    Figure 5.  Illustrations for small spot patterns.

    Existence of square patterns can be seen from (4) where C=4, D=0.5, h=0.45, μ=1.5, d3=10, d2=1 and d1=1 initially perturbed as a=0.1+cos(y2)cos(x2), b=0.1+cos(y2)cos(x2) with iterations (ⅰ) i = 100, (ⅱ) i = 110, (ⅲ) i = 120 and (ⅳ) i = 140 (refers to Figure 6).

    Figure 6.  Illustrations for square patterns.

    Existence of stripe patterns can be seen from (4) where C=6, D=2, h=0.45, μ=1.5, d3=2, d2=0.1 and d1=0.1 initially perturbed as a=0.1+0.1sin2(x), b=0.5+0.1cos2(x) with iterations (ⅰ) i = 64, (ⅱ) i = 128, (ⅲ) i = 256 and (ⅳ) i = 512 (refers to Figure 7).

    Figure 7.  Illustrations for stripe patterns.

    Cross-diffusion plays a crucial role in determining and selecting patterns in the spatio-temporal extension of the model under consideration. In addition to random diffusion (known as self-diffusion) of both species, cross-diffusion components are introduced to account for the influence of the population density of one species on the movement of other species. Due to the effect of cross-diffusion factors, the homogeneous stable state of the self-diffusion model, which is stationary under heterogeneous perturbation, loses its stability and generates diverse patterns, such as big spot, mixture of spots and stripes, and labyrinthine. Cross-diffusion changes the big spot pattern created by the self-diffusion model into a homogeneous steady state or other patterns such as a mixture of big spot and stripes as well as labyrinthine. This clarifies both the stabilizing and destabilizing effects of the nonlinear cross-diffusion factors. In addition, when the self-diffusion coefficients are assumed to be equal, the homogeneous stable state of the self-diffusion model loses its stability and creates spatial patterns consisting of small spots, followed by a mixture of small spots and stripes, and finally labyrinthine. As a result of cross-diffusion, the prey and predator populations display an inverse connection, which is another significant finding in this instance.

    Using amplitude equations, the impacts of super-cross diffusion on a system containing a self-diffusive term have been explored through our study. Until linear or nonlinear super-cross-diffusion terms are utilized in the systems, the self-diffusion approach does not yield significant Turing pattern whenever the prey dependent functional responses are addressed. By using the stability analysis, several criteria have been specified to assure the system's Turing instability. Under such conditions, the system will accommodate the Turing instabilities while producing the related specified configurations. Multiple-scale perturbation analysis has also been done all around Turing bifurcation boundary to obtain the amplitude equations used to analyze the stability of several Turing patterns. These findings are mathematically verified and appear to be in good order. Hexagons, huge spots, tiny spots, squares, and stripes are all produced with the inclusion of self and super-cross-diffusion terms.

    This research received funding support from the NSRF via the Program Management Unit for Human Resources & Institutional Development, Research and Innovation, (grant number~B05F650018).

    The authors declare that they have no conflicts of interest.



    [1] 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
    [2] M. Banerjee, S. Abbas, Existence and non-existence of spatial patterns in a ratio-dependent predator-prey model, Ecol. Complex., 21 (2015), 199–214. https://doi.org/10.1016/j.ecocom.2014.05.005 doi: 10.1016/j.ecocom.2014.05.005
    [3] M. Banerjee, S. Ghorai, N. Mukherjee, Study of cross-diffusion induced Turing patterns in a ratio-dependent prey-predator model via amplitude equations, Appl. Math. Model., 55 (2018), 383–399. https://doi.org/10.1016/j.apm.2017.11.005 doi: 10.1016/j.apm.2017.11.005
    [4] M. Banerjee, N. Mukherjee, V. Volpert, Prey-predator model with a nonlocal bistable dynamics of prey, Mathematics, 6 (2018), 41. https://doi.org/10.3390/math6030041 doi: 10.3390/math6030041
    [5] M. A. Budroni, Cross-diffusion-driven hydrodynamic instabilities in a double-layer system: general classification and nonlinear simulations, Phys. Rev. E, 92 (2015), 063007. https://doi.org/10.1103/PhysRevE.92.063007 doi: 10.1103/PhysRevE.92.063007
    [6] M. Chen, R. Wu, L. Chen, Spatiotemporal patterns induced by Turing and Turing-Hopf bifurcations in a predator-prey system, Appl. Math. Comput., 380 (2020), 125300. https://doi.org/10.1016/j.amc.2020.125300 doi: 10.1016/j.amc.2020.125300
    [7] J. M. Chung, E. Peacock-Lopez, Cross-diffusion in the Templetor model of chemical self-replication, Phys. Lett. A, 371 (2007), 41–47. https://doi.org/10.1016/j.physleta.2007.04.114 doi: 10.1016/j.physleta.2007.04.114
    [8] S. M. Cox, P. C. Matthews, Exponential time differencing for stiff-systems, J. Comp. Phys., 176 (2002), 430–455. https://doi.org/10.1006/jcph.2002.6995 doi: 10.1006/jcph.2002.6995
    [9] M. C. Cross, P. C. Hohenberg, Pattern formation outside of equilibrium, Rev. Mod. Phys., 65 (1993), 851. https://doi.org/10.1103/RevModPhys.65.851 doi: 10.1103/RevModPhys.65.851
    [10] H. I. Freedman, Deterministic mathematical models in population ecology, New York: Marcel Dekker Incorporated, 1980.
    [11] P. Feng, Y. Kang, Dynamics of a modified Leslie-Gower model with double Allee effects, Nonlinear Dyn., 80 (2015), 1051–1062. https://doi.org/10.1007/s11071-015-1927-2 doi: 10.1007/s11071-015-1927-2
    [12] G. Gambino, M. C. Lombardo, M. Sammartino, Pattern formation driven by cross–diffusion in a 2D domain, Nonlinear Anal. Real World Appl., 14 (2013), 1755–1779.
    [13] R. Gorenflo, F. Mainardi, Random walk models for space–fractional diffusion processes, Fract. Calc. Appl. Anal., 1 (1998), 167–191.
    [14] G. Hu, X. Li, Y. Wang, Pattern formation and spatiotemporal chaos in a reaction-diffusion predator–prey system, Nonlinear Dyn., 81 (2015), 265–275. https://doi.org/10.1007/s11071-015-1988-2 doi: 10.1007/s11071-015-1988-2
    [15] N. Iqbal, R. Wu, B. Liu, Pattern formation by super–diffusion in FitzHugh–Nagumo model, Appl. Math. Comput., 313 (2017), 245–258. https://doi.org/10.1016/j.amc.2017.05.072 doi: 10.1016/j.amc.2017.05.072
    [16] N. Iqbal, Y. Karaca, Complex fractional-order HIV diffusion model based on amplitude equations with turing patterns and turing instability, Fractals, 29 (2021), 2140013. https://doi.org/10.1142/S0218348X21400132 doi: 10.1142/S0218348X21400132
    [17] N. Iqbal, R. Wu, W. W. Mohammed, Pattern formation induced by fractional cross-diffusion in a 3-species food chain model with harvesting, Math. Comput. Simul., 188 (2021), 102–119. https://doi.org/10.1016/j.matcom.2021.03.041 doi: 10.1016/j.matcom.2021.03.041
    [18] Y. Jia, P. Xue, Effects of the self- and cross-diffusion on positive steady states for a generalized predator-prey system, Nonlinear Anal. Real World Appl., 32 (2016), 229–241. https://doi.org/10.1016/j.nonrwa.2016.04.012 doi: 10.1016/j.nonrwa.2016.04.012
    [19] T. Kadota, K. Kuto, Positive steady states for a prey-predator model with some nonlinear diffusion terms, J. Math. Anal. Appl., 323 (2006), 1387–1401. https://doi.org/10.1016/j.jmaa.2005.11.065 doi: 10.1016/j.jmaa.2005.11.065
    [20] A. K. Kassam, L. N. Trefethen, Fourth-order time-stepping for stiff PDEs, SIAM J. Sci. Comp., 26 (2005), 1212–1233. https://doi.org/10.1137/S1064827502410633 doi: 10.1137/S1064827502410633
    [21] E. Knobloch, J.D. Luca, Amplitude equations for travelling wave convection, Nonlinearity, 3 (1990), 975–980.
    [22] K. Kuto, Y. Yamada, Multiple coexistence states for a prey-predator system with cross-diffusion, J. Differ. Eq., 197 (2004), 315–348. https://doi.org/10.1016/j.jde.2003.08.003 doi: 10.1016/j.jde.2003.08.003
    [23] B. Liu, R. Wu, N. Iqbal, L. Chen, Turing patterns in the Lengyel-Epstein system with superdiffusion, Int. J. Bifurcat. Chaos, 27 (2017), 1730026. https://doi.org/10.1142/S0218127417300269 doi: 10.1142/S0218127417300269
    [24] B. Liu, R. Wu, L. Chen, Patterns induced by super cross–diffusion in a predator-prey system with Michaelis-Menten type harvesting, Math. Biosci., 298 (2018), 71–79. https://doi.org/10.1016/j.mbs.2018.02.002 doi: 10.1016/j.mbs.2018.02.002
    [25] J. D. Murray, Mathematical biology, Heidelberg: Springer, 1989.
    [26] K. Oeda, Effect of cross-diffusion on the stationary problem of prey-predator model with a protection zone, J. Differ. Eq., 250 (2011), 3988–4009. https://doi.org/10.1016/j.jde.2011.01.026 doi: 10.1016/j.jde.2011.01.026
    [27] K. M. Owolabi, K. C. patidar, Numerical simulations for multicomponent ecological models with adaptive methods, Theor. Biol. Med. Model., 13 (2016), 1–25. https://doi.org/10.1186/s12976-016-0027-4 doi: 10.1186/s12976-016-0027-4
    [28] S. Pal, S. Ghorai, M. Banerjee, Analysis of a prey-predator model with non-local interaction in the prey population, Bull. Math. Biol., 80 (2018), 906–925. https://doi.org/10.1007/s11538-018-0410-x doi: 10.1007/s11538-018-0410-x
    [29] S. G. Samko, A. A. Kilbas, O. I. Marichev, Fractional integrals and derivatives: theory and applications, New York: Gordon and Breach Science Publishers, 1993.
    [30] N. Shigesada, K. Kawasaki, E. Teramoto, Spatial segregation of interacting species, J. Theor. Biol., 79 (1979), 83–99. https://doi.org/10.1016/0022-5193(79)90258-3 doi: 10.1016/0022-5193(79)90258-3
    [31] Y. L Song, R. Yang, G. Q Sun, Pattern dynamics in a Gierer-Meinhardt model with a saturating term, Appl. Math. Model., 46 (2017), 476–491. https://doi.org/10.1016/j.apm.2017.01.081 doi: 10.1016/j.apm.2017.01.081
    [32] C. Tian, L. Zhang, Z. Lin, Pattern formation for a model of plankton allelopathy with cross-diffusion, J. Franklin I., 348 (2011), 1947–1964. https://doi.org/10.1016/j.jfranklin.2011.05.013 doi: 10.1016/j.jfranklin.2011.05.013
    [33] C. M. Topaz, A. J. Catla, Forced patterns near a Turing-Hopf bifurcation, Phys. Rev. E, 81 (2010), 026213. https://doi.org/10.1103/PhysRevE.81.026213 doi: 10.1103/PhysRevE.81.026213
    [34] L. N. Trefethen, Spectral methods in MATLAB, Philadelphia: SIAM Society for industrial and applied mathematics, 2000.
    [35] M. A. Tsyganov, V. N. Biktashev, Classification of wave regimes in excitable systems with linear cross diffusion, Phys. Rev. E, 90 (2014), 062912. https://doi.org/10.1103/PhysRevE.90.062912 doi: 10.1103/PhysRevE.90.062912
    [36] A. M. Turing, The chemical basis of morphogenesis, Phil. Trans. R. Soc. Lond. B, 237 (1952), 37–72.
    [37] V. K. Vanag, I. R. Epstein, Cross-diffusion and pattern formation in reaction-diffusion systems, Phys. Chem. Chem. Phys., 11 (2009), 897–912. https://doi.org/10.1039/B813825G doi: 10.1039/B813825G
    [38] X. Ni, R. Yang, W. Wang, Y. Lai, C. Grebogi, Cyclic competition of mobile species on continuous space: Pattern formation and coexistence, Phys. Rev. E, 82 (2010), 066211. https://doi.org/10.1103/PhysRevE.82.066211 doi: 10.1103/PhysRevE.82.066211
    [39] S. Yuan, C. Xu, T. Zhang, Spatial dynamics in a predator-prey model with herd behavior, Chaos, 23 (2013), 033102. https://doi.org/10.1063/1.4812724 doi: 10.1063/1.4812724
    [40] E. P. Zemskov, K. Kassner, M. J. B. Hauser, W. Horsthemke, Turing space in reaction-diffusion systems with density-dependent cross diffusion, Phys. Rev. E, 87 (2013), 032906. https://doi.org/10.1103/PhysRevE.87.032906 doi: 10.1103/PhysRevE.87.032906
    [41] X. Zhang, G. Sun, Z. Jin, Spatial dynamics in a predator-prey model with Beddington-Deangelis functional response, Phys. Rev. E, 85 (2012), 021924. https://doi.org/10.1103/PhysRevE.85.021924 doi: 10.1103/PhysRevE.85.021924
    [42] T. Zhang, Y. Xing, H. Zang, M. Han, Spatio–temporal dynamics of a reaction–diffusion system for a predator–prey model with hyperbolic mortality, Nonlinear Dyn., 78 (2014), 265–277. https://doi.org/10.1007/s11071-014-1438-6 doi: 10.1007/s11071-014-1438-6
    [43] L. Zhang, C. R. Tian, Turing pattern dynamics in an activator-inhibitor system with super diffusion, Phys. Rev. E, 90 (2014), 062915. https://doi.org/10.1103/PhysRevE.90.062915 doi: 10.1103/PhysRevE.90.062915
  • This article has been cited by:

    1. Rongjie Yu, Hengguo Yu, Min Zhao, Steady states and spatiotemporal dynamics of a diffusive predator-prey system with predator harvesting, 2024, 9, 2473-6988, 24058, 10.3934/math.20241170
    2. Soo-Oh Yang, Jea-Hyun Park, Analysis for the hierarchical architecture of the heterogeneous FitzHugh-Nagumo network inducing synchronization, 2023, 8, 2473-6988, 22385, 10.3934/math.20231142
    3. Vinoth Seralan, R. Vadivel, Dimplekumar Chalishajar, Nallappan Gunasekaran, Dynamical complexities and chaos control in a Ricker type predator-prey model with additive Allee effect, 2023, 8, 2473-6988, 22896, 10.3934/math.20231165
    4. Mohammad Sharif Ullah, K. M. Ariful Kabir, Md. Abdul Hakim Khan, A non-singular fractional-order logistic growth model with multi-scaling effects to analyze and forecast population growth in Bangladesh, 2023, 13, 2045-2322, 10.1038/s41598-023-45773-1
  • Reader Comments
  • © 2023 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(1879) PDF downloads(143) Cited by(4)

Figures and Tables

Figures(7)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog