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

Predator and n-classes-of-prey model incorporating extended Holling type Ⅱ functional response for n different prey species

  • In this paper, the Holling type Ⅱ functional response extended for n different species of prey and the dynamics of interactions between one predator species and its n different classes of prey are modeled. Positivity, boundedness and permanence of all solutions of the model are proved. An ecological threshold parameter for the predator free equilibrium point of the model is established. Local stability and global stability of the predator free equilibrium point are discussed. Furthermore, we also studied that the reproduction number R0 determines whether the equilibrium points are asymptotically stable or unstable. In addition, the model was solved numerically to confirm the analytical results.

    Citation: Saiwan Fatah, Arkan Mustafa, Shilan Amin. Predator and n-classes-of-prey model incorporating extended Holling type Ⅱ functional response for n different prey species[J]. AIMS Mathematics, 2023, 8(3): 5779-5788. doi: 10.3934/math.2023291

    Related Papers:

    [1] Agus Suryanto, Isnani Darti . On the nonstandard numerical discretization of SIR epidemic model with a saturated incidence rate and vaccination. AIMS Mathematics, 2021, 6(1): 141-155. doi: 10.3934/math.2021010
    [2] Xiaolong Chen, Hongfeng Zhang, Cora Un In Wong . Optimization study of tourism total revenue prediction model based on the Grey Markov chain: a case study of Macau. AIMS Mathematics, 2024, 9(6): 16187-16202. doi: 10.3934/math.2024783
    [3] Hassan Ranjbar, Leila Torkzadeh, Dumitru Baleanu, Kazem Nouri . Simulating systems of Itô SDEs with split-step (α,β)-Milstein scheme. AIMS Mathematics, 2023, 8(2): 2576-2590. doi: 10.3934/math.2023133
    [4] Jinran Yao, Zhengwei Yin . Numerical contractivity preserving implicit balanced Milstein-type schemes for SDEs with non-global Lipschitz coefficients. AIMS Mathematics, 2024, 9(2): 2766-2780. doi: 10.3934/math.2024137
    [5] Saima Rashid, Fahd Jarad, Sobhy A. A. El-Marouf, Sayed K. Elagan . Global dynamics of deterministic-stochastic dengue infection model including multi specific receptors via crossover effects. AIMS Mathematics, 2023, 8(3): 6466-6503. doi: 10.3934/math.2023327
    [6] Reem K. Alhefthi, Pakeeza Ashraf, Ayesha Abid, Shahram Rezapour, Abdul Ghaffar, Mustafa Inc . Exploring the flexibility of m-point quaternary approximating subdivision schemes with free parameter. AIMS Mathematics, 2024, 9(11): 33185-33214. doi: 10.3934/math.20241584
    [7] Zafar Iqbal, Muhammad Aziz-ur Rehman, Muhammad Imran, Nauman Ahmed, Umbreen Fatima, Ali Akgül, Muhammad Rafiq, Ali Raza, Ali Asrorovich Djuraev, Fahd Jarad . A finite difference scheme to solve a fractional order epidemic model of computer virus. AIMS Mathematics, 2023, 8(1): 2337-2359. doi: 10.3934/math.2023121
    [8] Dan-Dan Dai, Wei Zhang, Yu-Lan Wang . Numerical simulation of the space fractional (3+1)-dimensional Gray-Scott models with the Riesz fractional derivative. AIMS Mathematics, 2022, 7(6): 10234-10244. doi: 10.3934/math.2022569
    [9] Ishtiaq Ali, Sami Ullah Khan . Dynamics and simulations of stochastic COVID-19 epidemic model using Legendre spectral collocation method. AIMS Mathematics, 2023, 8(2): 4220-4236. doi: 10.3934/math.2023210
    [10] Thanasak Mouktonglang, Zulqurnain Sabir, Muhammad Asif Zahoor Raja, Saira Bhatti, Thongchai Botmart, Wajaree Weera, Chantapish Zamart . Designing Meyer wavelet neural networks for the three-species food chain model. AIMS Mathematics, 2023, 8(1): 61-75. doi: 10.3934/math.2023003
  • In this paper, the Holling type Ⅱ functional response extended for n different species of prey and the dynamics of interactions between one predator species and its n different classes of prey are modeled. Positivity, boundedness and permanence of all solutions of the model are proved. An ecological threshold parameter for the predator free equilibrium point of the model is established. Local stability and global stability of the predator free equilibrium point are discussed. Furthermore, we also studied that the reproduction number R0 determines whether the equilibrium points are asymptotically stable or unstable. In addition, the model was solved numerically to confirm the analytical results.



    This article investigates the stability of the nonlinear magnetic diffusion equation and its fully implicit discrete scheme for the following equation system in [1]:

    {tB(x,t)=x(η(e)μ0xB(x,t)),t(e(x,t)+12μ0B2(x,t))=x(η(e)B(x,t)μ20xB(x,t)), (1.1)

    where B is the magnetic field, e is the internal energy density (that is, internal energy per volume), μ0 is the vacuum permeability constant (μ0=4π×107N/A2), and η(e) is the resistivity in the material. The relationship between η(e) and e results in nonlinearity of the diffusion term x(η(e)μ0xB(x,t)) in (1.1).

    The resistivity η(e) in the equation system (1.1) is a step-function, as shown in Figure 1a:

    η(e)=η(x,t)={ηS=9.7×105,e[0,ec],ηL=9.7×103,e(ec,+), (1.2)

    ec=0.11084958, representing the critical value of internal energy density.

    Figure 1.  Comparison between step-function and smoothed step-function resistivity.

    In electromagnetic loading experiments[2], when the magnetic field outside the metal wall is relatively small (below 10 T), the driving current is also very small, the heating in the metal is weak, the temperature rise is slow, and the change in metal resistivity is not significant. At this point, the diffusion of the magnetic field exhibits behavioral characteristics similar to common diffusion phenomena such as thermal diffusion and concentration diffusion. When the magnetic field outside the metal wall reaches a strong magnetic field level of 100 T, the diffusion of the magnetic field in the metal will exhibit a nonlinear magnetic diffusion wave phenomenon. Compared to ordinary magnetic diffusion in metals, nonlinear magnetic diffusion waves have higher penetration rates and velocities, which can cause rapid magnetic flux leakage and device load erosion in high-energy-density physical experiments. Although nonlinear magnetic diffusion waves in metals with strong magnetic fields were proposed as early as 1970, it was not until after 2000 that phenomena related to nonlinear magnetic diffusion waves gradually attracted people's attention with the widespread development of electromagnetic driven high-energy-density physics experiments. The fundamental reason for the formation of nonlinear magnetic diffusion waves is that during the process of metal temperature rise caused by magnetic diffusion, the metal resistivity also changes accordingly[3]. Before the metal forms a highly conductive plasma, the overall resistivity shows an upward trend. After metal gasification, as the temperature increases, the degree of metal vapor ionization increases, and the resistivity gradually decreases. In [4,5], authors such as B. Xiao assume that the electrical resistivity of metals undergoes a sudden change of several orders of magnitude after reaching a critical temperature, while the electrical resistivity before and after the sudden change is independent of temperature. They consider an approximate theoretical analytical solution for one-dimensional steep-gradient surface magnetic diffusion waves under the step-function resistivity model. In [1], C. H. Yan et al. designed an explicit finite volume discretization scheme for one-dimensional magnetic field diffusion problems based on the step resistivity model. By relaxing the time step, the formulas for excessive magnetic flux transport and total internal energy transport were truncated when solving strong magnetic diffusion problems. On the basis of using the truncated magnetic flux transport capacity and total internal energy transport capacity, the program can allow for larger time steps without causing oscillation dispersion. In addition, there are also some studies on magnetic diffusion problems, such as [6,7,8].

    The stability of solutions is an important issue in the study of differential equations. Stability generally refers to the behavior of the solution remaining unchanged or tending to a certain equilibrium state when there is a small disturbance in the initial or boundary conditions of the equation. In [9], Y. L. Zhou et al. studied a class of parallel nature difference schemes for the initial boundary value problem of quasi-linear parabolic systems, and proved the unconditional stability of the constructed parallel nature difference scheme solutions under the discrete W(2,1)2 norm. In [10], author G. W. Yuan proved the uniqueness and stability of the obtained difference solution under the general non-uniform grid difference scheme. In [11,12], based on the non-uniform grid difference scheme, the authors constructed and developed an implicit discrete scheme that maintains the conservation of the implicit scheme while maintaining the required accuracy and unconditional stability for parallel computing through various methods such as estimation correction, to meet the needs of large-scale numerical solutions to radiation fluid dynamics problems.

    The magnetic diffusion problem studied in this article is also based on the step-function resistivity model. We first reproduced the results of equation system (1.1) in [1] (under explicit finite volume discretization scheme):

    Figure 2.  Magnetic field and internal energy density with step-function resistivity.
    Note: The c in the above figure is the time step influence factor. In this paper, the solution under the explicit finite volume scheme at c=0.4 is considered as the true solution of the problem.

    Next, in the magnetic diffusion equation system (1.1), the smoothed step-function resistivity ηδ(e) is used, where δ is used to describe the distance from the smooth curve inflection point to ec, as shown in Figure 1b. The experimental results of the implicit finite volume method are as follows:

    Figure 3.  Magnetic field and internal energy density with smoothed step-function resistivity.

    The above experiment indicates that by replacing the step-function resistivity η(e) in equation system (1.1) and using the smoothed step-function resistivity model ηδ(e), the experimental results in [1] can be well reproduced. Can the modified resistivity maintain the stability of the solution to the magnetic diffusion equation? What are the advantages of the corrected resistivity compared to the step-function resistivity? These are the starting points of this study and will be answered one by one in the following text. Below, we will first theoretically prove that the solution of the one-dimensional nonlinear magnetic diffusion equation and its fully implicit scheme under the smoothed step-function resistivity are stable with initial values. Then, the correctness and stability of the magnetic diffusion model under the smoothed step-function resistivity in the implicit finite volume discrete scheme are further verified through comparative experiments of explicit and implicit schemes[13].

    A measurable function u[0,T]X that satisfies the following conditions:

    (1)uLp(0,T;X):=(T0u(t)pdt)1/p<,1p<,(2)uL(0,T;X):=esssup0<tTu(t)<,

    forms the Lp(0,T;X) space.

    The polishing function J(x) satisfies

    J(x)={ce1x21,|x|<1,0,|x|1,c=111e1x21dt, (2.1)

    and the conclusions are as follows.

    Lemma 1 ([14]). For any ϵ>0, when taking Jϵ(x)=1ϵJ(xϵ), J(x) and Jϵ(x) satisfy the following properties:

    (1) J(x)C(R), and when |x|1,kN, J(k)(x)=0;

    (2) RJ(x)dx=RJϵ(x)dx=1.

    Then, the step-function resistivity (1.2) can be smoothed to the following continuous differentiable function:

    ηδ(e)=ηδ(x,t)={ηϵ(e),e[ecϵ,ec+ϵ],η(e),else, (2.2)

    where,

    ηϵ(e)=ηϵ(x,t)=Iη(y)Jϵ(xy)dy,I=[ecϵ,ec+ϵ]. (2.3)

    According to [14], it is easy to know that the resistivity (2.3) after the effect of the polishing function (2.1) satisfies the following properties:

    ηϵC(R),andηϵ(x)=Rη(y)Jϵ(xy)dy,x,yI. (2.4)

    Thereby, ηδ(e) in (2.2) is continuously differentiable across all real number fields. Further, ηδ(e) converges to η(e): η(e) is integrable on Iϵ=[ecϵ,ec+ϵ]. By the Lemma 1, for each xIϵ=[ecϵ,ec+ϵ], there is RJϵ(xy)dy=1, and for any ϵ>0, it is easily available that

    |ηδ(x)η(x)|=|Rη(y)Jϵ(xy)dyRη(x)Jϵ(xy)dy|R|η(y)η(x)|Jϵ(xy)dyCec+ϵecϵ|η(y)η(x)|dy=Cececϵ|ηSη(x)|dy+ec+ϵec|ηLη(x)|dy2C(ηLηS)ϵ, (2.5)

    where C=max|Jϵ(xy)|, and ηS,ηL are the minimum and maximum values of η(e). Thus, it can be concluded that limϵ0ηδ(x)=η(x).

    Remark: This provides us with a theoretical basis for a stability proof by replacing the step-function resistivity η(e) in equation system (1.1) with the smoothed resistivity ηδ(e) (continuous, differentiable).

    Lemma 2. (Young's inequality) If p>1,q>1, such that 1p+1q=1, then a,b0, the following inequality holds:

    abapp+bqq,

    and specifically, when a=εu,b=v2ε, Young's inequality can be expressed as

    uvεu2+14εv2. (2.6)

    Lemma 3. (Continuous Gronwall' inequality) Let g(t) and h(t) be non negative integrable functions, and satisfy f(t)f(t)g(t)+h(t). The following inequality holds:

    f(t)et0g(s)ds(f(0)+t0h(s)ds).

    Lemma 4. (Discrete Gronwall' inequality)[15] Let {fn}, {gn}, and {hn} be sequences of non-negative functions satisfying, fn+1fnΔtfn+1gn+1+hn+1, for α>1, such that

    Δtmax1nNgnα1α,

    where Δt>0. Then, the following inequality holds:

    fnCe3τmax1nNgn(f0+nk=0hkΔt), (2.7)

    where C and τ are constants that depend on the initial conditions.

    Lemma 5. (Abel's identity) Let {an} and {bn} be sequences of real or complex functions. If Qn=ni=1bi, the following identity holds:

    S=ni=1aibi=Qnann1i=1Qi(ai+1ai). (2.8)

    Lemma 6. (Embedding inequality)

    w(,s)2 (2.9)

    derived from the Poincaré inequality, w(\cdot, t) \in H_0^1(0, l) . Then, the inequality \|w(\cdot, t)\|_{2} \lesssim \|w_x(\cdot, t)\|_{2} holds. Substituting this inequality into (2.9) yields \|w(\cdot, t)\|_{\infty}^2 \lesssim \|w_x(\cdot, t)\|_{2}^2 .

    Lemma 7. (Discrete embedding inequality) (space direction)

    \begin{align} &||w_{h}^{k}||_{\infty}^{2}\lesssim\varepsilon \|\delta w_{h}^{k}\|_{2}^{2}+\frac{1}{\varepsilon}\| w_{h}^{k}\|_{2}^{2}, \end{align} (2.10)

    and similar to Lemma 6, it is easy to derive \|w_{h}^{k}(\cdot, t)\|_{\infty}^2 \lesssim \|\delta w{_{h}^{k}}(\cdot, t)\|_{2}^2 .

    Remark: The conclusions of Lemmas 6 and 7 hold only in one dimension.

    Lemma 8. (Discrete embedding inequality) (time direction)

    \begin{equation} \begin{aligned} &||w_{h}^{k}||_{2}^{2}\leq\sum\limits_{m = 1}^{k}(\varepsilon\|\Delta_\tau w_{h}^{m}\|_2^2+\frac{1}{2\varepsilon}\|w_{h}^{m}\|^2)+\|w_{h}^{0}\|_2^{2}. \end{aligned} \end{equation} (2.11)

    This section proves that the equation is stable with initial values. Now, we will transform the non-zero boundary value problem into a non-zero initial value problem.

    \begin{equation} \begin{cases} \dfrac{\partial }{{\partial t}}B(x,t) = \dfrac{\partial }{{\partial x}}\left( {\dfrac{{\eta_\epsilon (e)}}{{{\mu _0}}}\dfrac{\partial }{{\partial x}}B(x,t)} \right),\\ B(0,t) = 0.2, \quad B(0.5,t) = 0, t\in (0,1],\\ B(x,0) = 0, x\in (0,0.5]. \end{cases} \end{equation} (3.1)

    Let B({x}, t) = u({x}, t)+v({x}, t) , and v({x}, t) satisfies the boundary conditions in (3.1), that is,

    v({x},t)|_{x = 0} = 0.2,\qquad v({x},t)|_{x = 0.5} = 0.

    Construct auxiliary functions v({x}, t) = 0.2-\dfrac{2}{5}{x} , x\in [0, 0.5] , based on boundary conditions. Then, the equation satisfying the definite solution condition in (3.1) with respect to u({x}, t) is

    \begin{equation} \begin{cases} \dfrac{\partial }{{\partial t}}u(x,t) = \dfrac{\partial }{{\partial x}}\left( {\dfrac{{\eta_\epsilon (e)}}{{{\mu _0}}}\dfrac{\partial }{{\partial x}}u(x,t)} \right), \\ u({x},0) = -(0.2-\dfrac{2}{5}{x}), \quad {x}\in (0,0.5],\\ u({x},t)|_{x = 0} = 0, u({x},t)|_{x = 0.5} = 0, \quad {t}\in (0,1]. \end{cases} \end{equation} (3.2)

    Remark: The equation to be proved below indicates the stability with respect to initial values, which also implies the stability of the original equation with respect to boundary values.

    Consider the one-dimensional magnetic diffusion equation with Dirichlet boundary as follows:

    \begin{equation} \begin{cases} \dfrac{\partial }{{\partial t}}B(x,t) = \dfrac{\partial }{{\partial x}}\left( {\dfrac{{\eta (e(B))}}{{{\mu _0}}}\dfrac{\partial }{{\partial x}}B(x,t)} \right),\\ B(0,t) = B(l,t) = 0, t\in (0,T],\\ B(x,0) = \varphi,{x}\in (0,l], \end{cases} \end{equation} (3.3)
    \begin{equation} \begin{cases} \dfrac{\partial }{{\partial t}}\widetilde{B}(x,t) = \dfrac{\partial }{{\partial x}}\left( {\dfrac{{{\eta} (e(\widetilde{B}))}}{{{\mu _0}}}\dfrac{\partial }{{\partial x}}\widetilde{B}(x,t)} \right),\\ \widetilde{B}(0,t) = \widetilde{B}(l,t) = 0, t\in (0,T],\\ \widetilde{B}(x,0) = \widetilde{\varphi},{x}\in (0,l], \end{cases} \end{equation} (3.4)

    where, B and \widetilde{B} are the magnetic field, and e and \widetilde{e} are the internal energy density ( e = e(B), \; \widetilde{e} = e(\widetilde{B}) ).

    The solutions of Eqs (3.3) and (3.4) belong to L^{\infty}(0, T;H_0^1(0, l))\cap L^{2}(0, T;H^2(0, l)) , and the following is an inequality for energy estimation:

    \begin{equation} \sup\limits_{0\leq t \leq T}\|B_x(\cdot, t)\|_2^2 + \displaystyle {\int}_0^{T} \|B_{xx}(\cdot,t)\|_2^2 dt \leq C\|B_x(\cdot,0)\|_2^2. \end{equation} (3.5)

    On the premise of not causing misunderstandings, for the convenience of labeling and calculation, in this section, we still use \eta(e) to represent the step-function resistivity after polishing. Let w(x, t) = B(x, t)-\widetilde{B}(x, t) , and based on the differentiability of the smoothed resistivity \eta(e) , it can be assumed that the derivatives of \eta and e satisfy the following relationship:

    \begin{equation} |\eta_x+\eta_e|\leq c_1, \quad e_B\leq c_2. \end{equation} (3.6)

    Remark: The c_1 in (3.6) depends on the value of \epsilon in (2.3).

    Subtract the first equation in (3.3) from the first equation in (3.4) to obtain

    \begin{align} B_{t}-\widetilde{B}_{t} = \frac{1}{\mu_{0}}[\eta(e(B))_x B_{x}-\eta(e(\widetilde{B}))_x\widetilde{B}_{x}+\eta(e(B))B_{xx}-{\eta(e(\widetilde{B}))}\widetilde{B}_{xx}]. \end{align} (3.7)

    The above equation can be changed to

    \begin{align} \mu_{0}w_t = \eta(e(B))_xw_x+\widetilde{B}_x(\eta({e(B)})_x-\eta(e(\widetilde{B}))_x)+\eta(e(B))w_{xx}+\widetilde{B}_{xx}(\eta(e(B))-\eta(e(\widetilde{B}))). \end{align} (3.8)

    According to the Lagrange mean value theorem, \eta(e(B))-\eta(e(\widetilde{B})) in (3.8) can be resolved as

    \begin{align} &\eta(e(B))-\eta(e(\widetilde{B})) = \eta_e(\zeta)(e(B)-e(\widetilde{B})) = \eta_e(\zeta)e^\prime{(\xi)}(B-\widetilde{B}) = \eta_e(\zeta)e^\prime{(\xi)}w, \end{align} (3.9)

    where, \eta_e(\zeta) represents the first derivative of \eta with respect to e , \zeta is the value between e(B) and e(\widetilde{B}) , e^\prime{(\xi)} represents the first derivative of e with respect to B and \widetilde{B} , and \xi is the value between B and \widetilde{B} .

    Substituting (3.9) into (3.8) yields

    \begin{align} \mu_{0}w_t = &\eta(e(B))_xw_x+\widetilde{B}_x(\eta({e(B)})_x-\eta(e(\widetilde{B}))_x)+\eta(e(B))w_{xx}+\widetilde{B}_{xx}\eta_e(\zeta)e^\prime{(\xi)}w. \end{align} (3.10)

    Multiply w_{xx} on both sides of (3.10) and integrate on x\in(0, l) to obtain

    \begin{align} &\mu_0\displaystyle {\int}_0^{l}w_tw_{xx}dx-\displaystyle {\int}_0^{l}\eta(e(B)) w_{xx}^2dx \\ = &\displaystyle {\int}_0^{l}\eta (e(B))_x w_xw_{xx}dx+\displaystyle {\int}_{0}^{l}\widetilde{B}_x(\eta({e(B)})_x-\eta(e(\widetilde{B}))_x)w_{xx}dx+\displaystyle {\int}_{0}^{l}\widetilde{B}_{xx}\eta_e(\zeta)e^\prime{(\xi)}ww_{xx}dx. \end{align} (3.11)

    The first term at the left end of the above equation can be written by the partial integration method as follows:

    \begin{equation} \displaystyle {\int}_0^l w_tw_{xx}dx = w_{t}w_{x}|_0^{l}-\displaystyle {\int}_{0}^{l}w_{tx}w_{x}dx = -\frac{1}{2}\displaystyle {\int}_{0}^{l}\frac{d}{dt}w^2_{x}dx = -\frac{1}{2}\frac{d}{dt}\|w_{x} \|_2^2. \end{equation} (3.12)

    By combining (1.2) \eta(e)\geq \eta_S > 0 with (3.12), the left end of (3.11) can be simplified as

    \begin{align} \mu_0\displaystyle {\int}_0^{l}w_tw_{xx}dx-\displaystyle {\int}_0^{l}\eta(e(B)) w_{xx}^2dx &\leq-\frac{\mu_0}{2}\frac{d}{dt}\|w_{x} \|_2^2- \eta_S\displaystyle {\int}_0^l w^2_{xx} dx \\ & = -\frac{\mu_0}{2}\frac{d}{dt}\|w_{x} \|_2^2- \eta_S\| w^2_{xx}\|_2^2, \end{align}

    that is,

    \begin{align} \frac{\mu_0}{2}\frac{d}{dt}\|w_{x} \|_2^2+ \eta_{S}\| w^2_{xx}\|_2^2\leq -(\mu_0\displaystyle {\int}_0^{l}w_tw_{xx}dx-\displaystyle {\int}_0^{l}\eta(e(B)) w_{xx}^2dx). \end{align} (3.13)

    Take absolute values on both sides of (3.13) and obtain from (3.11)

    \begin{align} &\frac{\mu_0}{2}\frac{d}{dt}\|w_{x} \|_2^2+ \eta_{S}\| w^2_{xx}\|_2^2\\ \leq &|\mu_0\displaystyle {\int}_0^{l}w_tw_{xx}dx-\displaystyle {\int}_0^{l}\eta(e(B)) w_{xx}^2dx|\\ \leq&\displaystyle {\int}_0^{l}|\eta (e(B))_x w_xw_{xx}|dx+\displaystyle {\int}_{0}^{l}|\widetilde{B}_x(\eta({e(B)})_x-\eta(e(\widetilde{B}))_x)w_{xx}|dx+\displaystyle {\int}_{0}^{l}|\widetilde{B}_{xx}\eta_e(\zeta)e^\prime{(\xi)}ww_{xx}|dx. \end{align} (3.14)

    On the basis of the assumption (3.6), (3.14) can be resolved as

    \begin{align} &\frac{\mu_0}{2}\frac{d}{dt}\|w_{x} \|_2^2+\eta_{S}\|w_{xx} \|_2^2 \\ \leq&c_1\displaystyle {\int}_0^{l}| w_xw_{xx}|dx+ 2c_1\displaystyle {\int}_{0}^{l}|\widetilde{B}_{x}w_{xx}|dx+c_1c_2\displaystyle {\int}_{0}^{l}|\widetilde{B}_{xx}ww_{xx}|dx, \end{align} (3.15)

    and according to the Lemma 2 (Young's inequality), (3.15) can be transformed into

    \begin{align} &\frac{\mu_0}{2}\frac{d}{dt}\|w_{x} \|_2^2+\eta_{S}\|w_{xx} \|_2^2 \\ \leq & c_1\displaystyle {\int}_0^l\left( \varepsilon_1|{w_{xx}}|^2 +\frac{1}{4\varepsilon_1} |{w_x}|^2\right)dx+ 2c_1\displaystyle {\int}_{0}^{l}\left( \varepsilon_2|{w_{xx}}|^2 + \frac{1}{4\varepsilon_2} |\widetilde{B}_x|^2\right)dx \\ +&c_1c_2\displaystyle {\int}_{0}^{l}\left( \varepsilon_3|{w}_{xx}|^2 +\frac{1}{4\varepsilon_3} |{w}|^2|\widetilde{B}_{xx}|^2\right)dx \\ \leq & c_3\|{w_{xx}}\|_2^2+ c_4\|{w_{x}}\|_2^2+c_5\|\widetilde{B}_x\|_2^2dx+c_6\displaystyle {\int}_{0}^{l}|w|^2|\widetilde{B}_{xx}|^2dx, \end{align} (3.16)

    where c_3 = c_1\varepsilon_1+2c_1\varepsilon_2+c_1c_2\varepsilon_3 .

    By the Lemma 6 (embedding inequality)

    \begin{align} c_6\displaystyle {\int}_{0}^{l}|w|^2|\widetilde{B}_{xx}|^2dx\leq c_6 \sup\limits_{0\leq x\leq l}|w|^2\|\widetilde{B}_{xx}\|_2^2 \leq c_{6} \|w\|_\infty^2\|\widetilde{B}_{xx}\|_2^2 \leq c_{7} \|w_x\|_2^2\|\widetilde{B}_{xx}\|_2^2. \end{align} (3.17)

    Remark: The conclusion of (3.17) only holds for one-dimensional cases.

    Substitute (3.17) into (3.16), and from the energy estimation inequality (3.5), obtain

    \begin{align} \frac{\mu_0}{2}\frac{d}{dt}\|w_{x} \|_2^2+\eta_{S}\|w_{xx} \|_2^2\leq & c_3\|{w_{xx}}\|_2^2+ \|{w_{x}}\|_2^2 (c_4+c_7\|\widetilde{B}_{xx}\|_2^2)+c_5\|\widetilde{B}_x\|_2^2 \\ \leq& c_3\|{w_{xx}}\|_2^2+ c_8\|{w_{x}}\|_2^2 (1+\|\widetilde{B}_{xx}\|_2^2)+c_5\sup\limits_{0\leq t \leq T}\|\widetilde{B}_x\|_2^2 \\ \leq & c_3\|{w_{xx}}\|_2^2+ c_8\|{w_{x}}\|_2^2 (1+\|\widetilde{B}_{xx}\|_2^2)+c_9, \end{align} (3.18)

    where c_9 = c_5C\|B_x(\cdot, 0)\|_2^2 .

    According to (3.18)

    \begin{align} \eta_{S}\|w_{xx} \|_2^2\leq c_3\|{w_{xx}}\|_2^2+ c_8\|{w_{x}}\|_2^2 (1+\|\widetilde{B}_{xx}\|_2^2)+c_9, \end{align} (3.19)

    so

    \begin{equation} \begin{aligned} c_{3}\|w_{xx} \|_2^2\leq \frac{c_{3}c_8}{\eta_S-c_{3}} \|{w_x}\|_2^2(1+\|\widetilde{B}_{xx}\|_2^2)+\frac{c_{3}c_9}{\eta_S-c_{3}}. \end{aligned} \end{equation} (3.20)

    Remark: c_3 = c_1\varepsilon_1+2c_1\varepsilon_2+c_1c_2\varepsilon_3 can ensure that \eta_S-c_{3} > 0 .

    Substitute (3.20) into the right-hand end of (3.18), and organize it to obtain

    \begin{equation} \begin{aligned} \frac{d}{dt}\|w_{x} \|_2^2+ \frac{2\eta_{S}}{\mu_0}\|w_{xx} \|_2^2 \leq c_{10} \|{w_x}\|_2^2(1+\|\widetilde{B}_{xx}\|_2^2)+c_{11}. \end{aligned} \end{equation} (3.21)

    In (3.21), on the one hand

    \begin{equation} \begin{aligned} \frac{d}{dt}\|w_{x} \|_2^2 \leq c_{10} \|{w_x}\|_2^2(1+\|\widetilde{B}_{xx}\|_2^2)+c_{11}. \end{aligned} \end{equation} (3.22)

    In (3.22), take f^\prime(t) = \dfrac{d}{dt}\|w_{x} \|_2^2 , f(t) = \|w_{x} \|_2^2 , g(t) = c_{10}(1+\|\widetilde{B}_{xx}\|_2^2) , h(t) = c_{11} , by using the Lemma 3 (continuous Gronwall' inequality), it can be concluded that

    \begin{equation} \begin{aligned} \|w_{x}(\cdot,t) \|_2^2 \leq e^{c_{10} {\int}_{0}^T(1+\|\widetilde{B}_{xx}\|_2^2)dt} \|w_{x}(\cdot,0) \|_2^2 = c_{12} \|\varphi_x-\widetilde{\varphi}_x \|_2^2+c_{13}. \end{aligned} \end{equation} (3.23)

    On the other hand, as can be seen from (3.21).

    \begin{equation} \begin{aligned} \frac{2\eta_{S}}{\mu_0}\|w_{xx}(\cdot,t) \|_2^2 \leq c_{10} \|{w_x}\|_2^2(1+\|\widetilde{B}_{xx}\|_2^2)+c_{11}. \end{aligned} \end{equation} (3.24)

    Integrate the two sides of Eq (3.24) in the time direction on [0, T] and obtain from (3.23)

    \begin{equation} \begin{aligned} \frac{2\eta_{S}}{\mu_0} \displaystyle {\int}_{0}^T\|w_{xx}(\cdot,t) \|_2^2 ds \leq \displaystyle {\int}_{0}^T c_{10}\|{w_x}\|^2_2(1+\|\widetilde{B}_{xx}\|_2^2)ds + \displaystyle {\int}_{0}^T c_{11}dt\leq c_{14} \|\varphi_x-\widetilde{\varphi}_x \|_2^2+c_{11}T. \end{aligned} \end{equation} (3.25)

    From (3.23) and (3.25), it can be concluded that

    \begin{align} \|w_x(x,t) \|_2^2+a\displaystyle {\int}_{0}^t\|w_{xx}(x,t) \|_2^2 ds \leq &C \|\varphi_x-\widetilde{\varphi}_x \|_2^2+c. \end{align} (3.26)

    Thus, the stability of the magnetic diffusion equation is proven.

    The following proves the stability of the fully implicit scheme corresponding to Eq (3.3) or (3.4):

    \begin{equation} \begin{cases} \mu_{0}\Delta_\tau B_{j}^{n+1} = \eta(e(B_{j}^{n+1}))\delta^{2}B_{j}^{n+1}+\delta\eta(e(B_{j}^{n+1}))\delta B_{j}^{n+1},\\ B_{0}^{n} = B_{J}^{n} = 0,\\ B_{j}^{0} = \varphi(x_{j}), \end{cases} \end{equation} (3.27)

    taking

    \begin{equation} \begin{cases} \mu_{0}\Delta_\tau {\widetilde{B}}_{j}^{n+1} = \eta({e}( {\widetilde{B}}_{j}^{n+1}))\delta^{2} {\widetilde{B}}_{j}^{n+1}+\delta\eta({e}( {\widetilde{B}}_{j}^{n+1}))\delta {\widetilde{B}}_{j}^{n+1},\\ {\widetilde{B}}_{0}^{n} = {\widetilde{B}}_{J}^{n} = 0,\\ {\widetilde{B}}_{j}^{0} = \widetilde{\varphi}(x_{j}). \end{cases} \end{equation} (3.28)

    The energy estimation in the discrete scheme is

    \begin{equation} \sup\limits_{0\leq n \leq N}\|\delta B_j^{n+1}\|_2^2 + \sum\limits_{n = 0}^{N} \|\delta^2 B_j^{n+1}\|_2^2 dt \leq C\|\delta B_j^0\|_2^2. \end{equation} (3.29)

    Let w_j^{n+1} = B_j^{n+1}- {\widetilde{B}}_j^{n+1} , and subtract the first equation in (3.27) from the first equation in (3.28) to obtain:

    \begin{equation} \begin{aligned} &\mu_{0}\Delta_{\tau}\left(B_{j}^{n+1}-{\widetilde{B}}_{j}^{n+1}\right) = \left(\eta(e(B_{j}^{n+1}))\delta^{2}B_{j}^{n+1}-\eta({e}( {\widetilde{B}}_{j}^{n+1}))\delta^{2} {\widetilde{B}}_{j}^{n+1}\right)\nonumber\\ +&\delta\eta(e(B_{j}^{n+1}))\delta B_{j}^{n+1}-\delta\eta({e}( {\widetilde{B}}_{j}^{n+1}))\delta {\widetilde{B}}_{j}^{n+1}.\nonumber\\ \end{aligned} \end{equation}

    After applying the Lagrange mean value theorem to the above equation, the following is obtained:

    \begin{equation} \begin{aligned} &\mu_{0}\Delta_{\tau}w_j^{n+1}-\left(\eta(e(B_{j}^{n+1}))\delta^{2}w_{j}^{n+1}+\delta^{2}{\widetilde{B}_{j}^{n+1}}\delta \eta(\zeta_{j}^{n+1}) \delta e(\xi_{j}^{n+1}) w_{j}^{n+1}\right)\\ = &\delta\eta(e(B_{j}^{n+1}))\delta w_{j}^{n+1} +\delta {\widetilde{B}_{j}^{n+1}}(\delta\eta(e(B_{j}^{n+1}))-\delta\eta({e}({\widetilde{B}_{j}^{n+1}}))), \end{aligned} \end{equation} (3.30)

    where \delta\eta(\zeta_{j}^{n+1}) represents the first derivative of \eta with respect to e , \zeta_{j}^{n+1} is the value between e(B_{j}^{n+1}) and e(\widetilde{B}_{j}^{n+1}) , \delta e{(\xi_{j}^{n+1})} represents the first derivative of e with respect to B_{j}^{n+1} or \widetilde{B}_{j}^{n+1} , and \xi_{j}^{n+1} is the value between B_{j}^{n+1} and \widetilde{B}_{j}^{n+1} .

    Multiply \delta^{2}w_j^{n+1} on both sides of (3.30), and sum j = 1, 2, \ldots, J-1 to obtain

    \begin{equation} \begin{aligned} &\mu_{0}\sum\limits_{j = 1}^{J-1}\delta^{2}w_j^{n+1}\Delta_{\tau}w_j^{n+1}-\sum\limits_{j = 1}^{J-1}\eta(e(B_{j}^{n+1}))(\delta^{2}w_j^{n+1})^2\\ = &\sum\limits_{j = 1}^{J-1}\delta^{2}{\widetilde{B}_{j}^{n+1}}\delta \eta(\zeta_{j}^{n+1}) \delta e(\xi_{j}^{n+1}) w_{j}^{n+1}\delta^2w_j^{n+1}+\sum\limits_{j = 1}^{J-1}\delta\eta(e(B_{j}^{n+1}))\delta w_{j}^{n+1}\delta^2w_j^{n+1}\\ +&\sum\limits_{j = 1}^{J-1}\delta {\widetilde{B}_{j}^{n+1}}(\delta\eta(e(B_{j}^{n+1}))-\delta\eta({e}({\widetilde{B}_{j}^{n+1}})))\delta^2w_j^{n+1}. \end{aligned} \end{equation} (3.31)

    We can consider the first term in equation (3.31)

    \sum\limits_{j = 1}^{J-1}\delta^2w_j^{n+1}\Delta_\tau w_j^{n+1}.

    Let

    \begin{equation} \begin{aligned} &a_{j} = \frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h},\\ &Q_{j} = \frac{w_{j}^{n+1}-w_{j}^{n}}{\Delta t},\\ &b_{j} = \frac{w_{j}^{n+1}-w_{j}^{n}}{\Delta t}-\frac{w_{j-1}^{n+1}-w_{j-1}^{n}}{\Delta t}. \end{aligned} \end{equation} (3.32)

    Using Lemma 5 (Abel's identity), it can be obtained that

    \begin{align} \sum\limits_{j = 1}^{J-1}\delta^2w_j^{n+1}\Delta_\tau w_j^{n+1} = &\sum\limits_{j = 1}^{J-1}\frac{w_{j}^{n+1}-w_{j}^{n}}{\Delta t}\frac{(a_{j+1}-a_j)}{h}\\ = &\frac{1}{h}\sum\limits_{j = 1}^{J-1}Q_j(a_{j+1}-a_j)\\ = &\frac{1}{h}(a_JQ_{J}-\sum\limits_{j = 1}^{J}a_jb_j) \\ = &-\frac{1}{h}\sum\limits_{j = 1}^{J}\left(\frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h}\right)\left(\frac{w_{j}^{n+1}-w_{j}^{n}}{\Delta t}-\frac{w_{j-1}^{n+1}-w_{j-1}^{n}}{\Delta t}\right)\\ = &-\frac{1}{\Delta t}\sum\limits_{i = j}^{J}\left(\frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h}\right)\left(\frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h}-\frac{w_{j}^{n}-w_{j-1}^{n}}{h}\right). \end{align} (3.33)

    According to the inequality

    \begin{equation} \begin{aligned} &u(u-v) = \frac{1}{2}\Big(u^2-u^2+(u-v)^2\Big)\geq\frac12\Bigl(u^2-v^2\Bigr). \end{aligned} \end{equation} (3.34)

    Let u = \frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h} , v = \frac{w_{j}^{n}-w_{j-1}^{n}}{h} , and from (3.34), (3.33) can be changed to

    \begin{align} &\frac{1}{\Delta t}\sum\limits_{j = 1}^{J}\left(\frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h}\right)\left(\frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h}-\frac{w_{j}^{n}-w_{j-1}^{n}}{h}\right) \\ \geq &\frac{1}{2}\frac{1}{\Delta t}\sum\limits_{j = 1}^{J}\left [(\frac{w_{j}^{n+1}-w_{j-1}^{n+1}}{h})^{2} -(\frac{w_{j}^{n}-w_{j-1}^{n}}{h})^{2} \right ] \\ = &\frac{1}{2\Delta t}(||\delta w_{h}^{n+1}||_{2}^{2}-||\delta w_{h}^{n}||_{2}^{2}), \end{align} (3.35)

    that is

    \begin{align} \mu_{0}\sum\limits_{j = 1}^{J-1}\delta^{2}w_j^{n+1}\Delta_{\tau}w_j^{n+1} \leq -\frac{\mu_{0}}{2\Delta t}(||\delta w_{h}^{n+1}||_{2}^{2}-||\delta w_{h}^{n}||_{2}^{2}). \end{align} (3.36)

    From \eta > \eta_S and substituting (3.36) into (3.31), it can be concluded that

    \begin{align} &\frac{\mu_0}{2\Delta t}\left(\left\|\delta w_h^{n+1}\right\|_2^2-\left\|\delta w_h^n\right\|_2^2\right)+\eta_S\left\|\delta^2w_h^{n+1}\right\|_2^2 \\ = &\sum\limits_{j = 1}^{J-1}|\delta^{2}{\widetilde{B}_{j}^{n+1}}\delta \eta(\zeta_{j}^{n+1}) \delta e(\xi_{j}^{n+1}) w_{j}^{n+1}\delta^2w_j^{n+1}|+\sum\limits_{j = 1}^{J-1}|\delta\eta(e(B_{j}^{n+1}))\delta w_{j}^{n+1}\delta^2w_j^{n+1}|\\ +&\sum\limits_{j = 1}^{J-1}|\delta {\widetilde{B}_{j}^{n+1}}(\delta\eta(e(B_{j}^{n+1}))-\delta\eta({e}({\widetilde{B}_{j}^{n+1}})))\delta^2w_j^{n+1}|. \end{align} (3.37)

    By the assumption (3.6)

    \begin{align} &\frac{\mu_0}{2\Delta t}\left(\left\|\delta w_h^{n+1}\right\|_2^2-\left\|\delta w_h^n\right\|_2^2\right)+\eta_S\left\|\delta^2w_h^{n+1}\right\|_2^2 \\ \leq&c_1c_2\sum\limits_{j = 1}^{J-1}|\delta^{2}{\widetilde{B}_{j}^{n+1}} w_j^{n+1}\delta^2w_j^{n+1}|+c_1\sum\limits_{j = 1}^{J-1}|\delta w_{j}^{n+1}\delta^2w_j^{n+1}|+2c_1\sum\limits_{j = 1}^{J-1}|\delta {\widetilde{B}_{j}^{n+1}}\delta^2w_j^{n+1}|. \end{align} (3.38)

    Applying Lemma 2 (Young's inequality) to Eq (3.38) yields

    \begin{align} &\frac{\mu_0}{2\Delta t}\left(\left\|\delta w_h^{n+1}\right\|_2^2-\left\|\delta w_h^n\right\|_2^2\right)+\eta_S\left\|\delta^2w_h^{n+1}\right\|_2^2 \\\ \leq&c_1c_2\left({\varepsilon_1 }\|\delta^2w_h^{n+1}\|_2^2+\frac{1}{4\varepsilon_1}{\|\delta^2\widetilde{B}_h^{n+1}\|_2^2\sup\limits_{0\leq h\leq J-1}|w_h^{n+1}|_2^2}\right)+c_1\left({ \varepsilon_2 }\|\delta^2w_h^{n+1}\|_2^2 +\frac{1}{4\varepsilon_2}\|\delta w_h^{n+1}\|_2^2\right)\\ +&2c_1\left(\varepsilon_3\|\delta^2w_h^{n+1}\|_2^2+\frac{1}{4\varepsilon_3}\|\delta \widetilde{B}_h^{n+1}\|_2^2\right). \end{align} (3.39)

    From the energy estimation inequality (3.29) and Lemma 7 (discrete embedding inequality, space direction), (3.39) can be expressed as

    \begin{align} &\frac{1}{\Delta t}\left(\left\|\delta w_h^{n+1}\right\|_2^2-\left\|\delta w_h^n\right\|_2^2\right)+\frac{2 \eta_S}{\mu_0}\left\|\delta^2w_h^{n+1}\right\|_2^2 \\ \leq & c_4\|\delta^2w_h^{n+1}\|_2^2+c_5 \|w_h^{n+1}(\cdot,t)\|_\infty^2\|\delta^2\widetilde{B}_h^{n+1}\|_2^2+c_6\|\delta w_h^{n+1}\|_2^2+c_7\sup\limits_{0\leq n\leq N}\|\delta B\widetilde{}_h^{n+1}\|_2^2 \\ \leq & c_4\|\delta^2w_h^{n+1}\|_2^2+c_5 \|w_h^{n+1}(\cdot,t)\|_\infty^2\|\delta^2\widetilde{B}_h^{n+1}\|_2^2 +c_6\|\delta w_h^{n+1}\|_2^2+c_7\sup\limits_{0\leq n\leq N}\|\delta \widetilde{B}_h^{n+1}\|_2^2. \end{align} (3.40)

    Let a = \frac{2 \eta_S}{\mu_0} , and according to (3.40)

    \begin{align} a\left\|\delta^2w_h^{n+1}\right\|_2^2\leq c_4\|\delta^2w_h^{n+1}\|_2^2+c_5 \|w_h^{n+1}(\cdot,t)\|_\infty^2\|\delta^2\widetilde{B}_h^{n+1}\|_2^2 +c_6\|\delta w_h^{n+1}\|_2^2+c_7\sup\limits_{0\leq n\leq N}\|\delta \widetilde{B}_h^{n+1}\|_2^2, \end{align} (3.41)

    so

    \begin{align} c_4\left\|\delta^2w_h^{n+1}\right\|_2^2\leq \frac{c_4c_5}{a-c_4}\|w_h^{n+1}(\cdot,t)\|_\infty^2\|\delta^2 \widetilde{B}_h^{n+1}\|_2^2+\frac{c_4c_6}{a-c_4}\|\delta w_h^{n+1}\|_2^2+\frac{c_4c_7}{a-c_4}\sup\limits_{0\leq n\leq N}\|\delta \widetilde{B}_h^{n+1}\|_2^2. \end{align} (3.42)

    Substituting (3.42) into the right-hand side of (3.40) yields

    \begin{align} \frac{1}{\Delta t}\left(\left\|\delta w_h^{n+1}\right\|_2^2-\left\|\delta w_h^n\right\|_2^2\right)+a\left\|\delta^2w_h^{n+1}\right\|_2^2\leq c_9\|\delta w_h^{n+1}\|_2^2 + c_{10}\|w_h^{n+1}(\cdot,t)\|_\infty^2\|\delta^2\widetilde{B}_h^{n+1}\|_2^2+c_8. \end{align} (3.43)

    Summing the two sides of (3.43) with respect to n , from inequality (3.29), it can be concluded that:

    \begin{align} &\left\|\delta w_h^{n+1}\right\|_2^2+a\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{k+1}\right\|_2^2\Delta t\\\ \leq& c_{11} \sup\limits_{1\leq k\leq {n}}\|w_h^{k+1}\|_\infty^2 \sum\limits_{k = 0}^{n}\|\delta^2\widetilde{B}_h^{n+1}\|_2^2\Delta t +c_9\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2 \Delta t+\|\delta w_h^{0}\|_2^2 +c_8\\ \leq& c_{12} \sup\limits_{1\leq k\leq {n}}\|w_h^{k+1}\|_\infty^2 +c_9\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2 \Delta t+\|\delta w_h^{0}\|_2^2 + c_8. \end{align} (3.44)

    According to Lemma 7, the right-hand side of the (3.44) inequality can be written as

    \begin{align} &\left\|\delta w_h^{n+1}\right\|_2^2+a\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{n+1}\right\|_2^2\Delta t\\\ \leq& c_{12}\sup\limits_{1\leq k\leq {n}}(\varepsilon\|\delta w_h^{k+1}\|_2^2+\frac{1}{\varepsilon}\|w_h^{k+1}\|_2^2) +c_9\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2 \Delta t+\|\delta w_h^{0}\|_2^2 +c_8. \end{align} (3.45)

    From (3.45),

    \begin{align} \sup\limits_{1\leq k\leq {n}}\left\|\delta w_h^{k+1}\right\|_2^2 \leq c_{12}\sup\limits_{1\leq k\leq {n}}(\varepsilon\|\delta w_h^{k+1}\|_2^2+\frac{1}{\varepsilon}\|w_h^{k+1}\|_2^2) +c_9\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2 \Delta t+\|\delta w_h^{0}\|_2^2 +c_8, \end{align} (3.46)

    and then,

    \begin{align} & c_{12}\varepsilon\sup\limits_{1\leq k\leq {n}}\left\|\delta w_h^{n+1}\right\|_2^2\\ \leq &\sup\limits_{1\leq k\leq {n}}\frac{(c_{12})^2}{1-c_{12}\varepsilon}\|w_h^{k+1}\|_2^2 +\frac{c_9c_{12}\varepsilon}{1-c_{12}\varepsilon}\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t+\frac{c_{12}\varepsilon}{1-c_{12}\varepsilon}\|\delta w_h^{0}\|_2^2+\frac{c_8c_{12}\varepsilon}{1-c_{12}\varepsilon}. \end{align} (3.47)

    Substituting (3.47) into the right-hand of (3.45) yields

    \begin{align} &\left\|\delta w_h^{n+1}\right\|_2^2+a\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{n+1}\right\|_2^2\Delta t\leq c_{13} \sup\limits_{1\leq k\leq {n}}\|w_h^{k+1}\|_2^2+c_{14}\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t+c_{15}\|\delta w_h^{0}\|_2^2 + c_{16}. \end{align} (3.48)

    On the basis of Lemma 8 (discrete embedding inequality, time direction), \|w_h^{k+1}\|_2^2 on the right hand of (3.48) can be expressed as

    \begin{equation} \begin{aligned} \|w_{h}^{k+1}\|_{2}^{2}\leq\sum\limits_{m = 0}^{k+1}(\varepsilon\|\Delta_\tau w_{h}^{m}\|_2^2+\frac{1}{2\varepsilon}\|w_{h}^{m}\|^2)+\|w_{h}^{0}\|_2^{2}. \end{aligned} \end{equation} (3.49)

    Using the bootstrapping and fully utilizing the properties of the format itself, it can be concluded from (3.30) that

    \begin{equation} \begin{aligned}\nonumber \Delta_{\tau}w_j^{n+1} = &\frac{1}{\mu_0} \left(\eta(e(B_{j}^{n+1}))\delta^{2}w_{j}^{n+1}+\delta^{2}{\widetilde{B}_{j}^{n+1}}\delta \eta(\zeta_{j}^{n+1}) \delta e(\xi_{j}^{n+1}) w_{j}^{n+1}\right)\\ +&\frac{1}{\mu_0}\left(\delta\eta(e(B_{j}^{n+1}))\delta w_{j}^{n+1} +\delta {\widetilde{B}_{j}^{n+1}}(\delta\eta(e(B_{j}^{n+1}))-\delta\eta({e}({\widetilde{B}_{j}^{n+1}})))\right). \end{aligned} \end{equation}

    Substitute the above equation into (3.49), use the energy estimation inequality (3.29) and the assumption condition (3.6), and repeat the above steps to obtain

    \begin{align} \|w_{h}^{k}\|_{2}^{2}\leq&\sum\limits_{m = 0}^{k}\varepsilon(c_{15}\|\delta^2w_h^{m}\|_2^2+c_{16}\|\delta w_h^{m}\|_2^2+ c_{17}\| w_h^{m}\|_2^2 )+\frac{1}{2\varepsilon}\sum\limits_{m = 0}^{k}\|w_{h}^{m}\|_2^2+\|w_{h}^{0}\|_2^{2}, \end{align} (3.50)

    that is,

    \begin{align} \|w_{h}^{k}\|_{2}^{2}\leq c_{18}\sum\limits_{m = 0}^{k}\|\delta^2w_h^{m}\|_2^2+c_{19}\sum\limits_{m = 0}^{k}\|\delta w_h^{m}\|_2^2+c_{20}\sum\limits_{m = 0}^{k}\| w_h^{m}\|_2^2+\|w_{h}^{0}\|_2^{2}. \end{align} (3.51)

    For the \|w_h^{m}\|_2^2 in the above equation, the embedding theorem is applied to obtain: \| w_h^{m}\|_2^2\leq\| w_h^{m}\|_\infty^2 \leq \|\delta w_h^{m}\|_2^2 . Thus(3.51) can be simplified as

    \begin{align} \|w_{h}^{k}\|_{2}^{2}\leq c_{18}\sum\limits_{m = 0}^{k}\|\delta^2w_h^{m}\|_2^2+c_{21}\sum\limits_{m = 0}^{k}\|\delta w_h^{m}\|_2^2+\|w_{h}^{0}\|_2^{2}. \end{align} (3.52)

    Substituting (3.52) into (3.48) yields

    \begin{align} &\left\|\delta w_h^{n+1}\right\|_2^2+a\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{k+1}\right\|_2^2\Delta t \\ \leq & c_{13}\sup\limits_{1\leq k\leq {n+1}}(c_{18}\sum\limits_{m = 0}^{k}\|\delta^2w_h^{m}\|_2^2+c_{21}\sum\limits_{m = 0}^{k}\|\delta w_h^{m}\|_2^2+\|w_{h}^{0}\|_2^{2})+ c_{14}\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t+c_{15}\|\delta w_h^{0}\|_2^2+c_{16}. \end{align} (3.53)

    Similarly, by

    \begin{align} &a\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{k+1}\right\|_2^2\Delta t \\ \leq & c_{13}\sup\limits_{1\leq k\leq {n+1}}(c_{18}\sum\limits_{m = 0}^{k}\|\delta^2w_h^{m}\|_2^2+c_{21}\sum\limits_{m = 0}^{k}\|\delta w_h^{m}\|_2^2+\|w_{h}^{0}\|_2^{2})+ c_{14}\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t+c_{15}\|\delta w_h^{0}\|_2^2, \end{align} (3.54)

    it can be inferred that

    \begin{align} &a\sup\limits_{1\leq k\leq {n+1}}\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{k+1}\right\|_2^2\Delta t \\ \leq & c_{13}\sup\limits_{1\leq k\leq {n+1}}(c_{18}\sum\limits_{m = 0}^{k}\|\delta^2w_h^{m}\|_2^2+c_{21}\sum\limits_{m = 0}^{k}\|\delta w_h^{m}\|_2^2+\|w_{h}^{0}\|_2^{2})+ c_{14}\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t+c_{15}\|\delta w_h^{0}\|_2^2. \end{align} (3.55)

    Thereby it can be deduced that

    \begin{align} c_{13}c_{18}\sup\limits_{1\leq k\leq {n+1}}\sum\limits_{m = 0}^{k}\|\delta^2w_h^{m}\|_2^2 \leq c_{22}\sum\limits_{m = 0}^{k}(\|\delta w_h^{m}\|_2^2+\|w_{h}^{0}\|_2^{2})+ c_{23}\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t+c_{24}\|\delta w_h^{0}\|_2^2. \end{align} (3.56)

    Substituting (3.56) into the right end of (3.53) yields

    \begin{align} &\left\|\delta w_h^{n+1}\right\|_2^2+a\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{k}\right\|_2^2\Delta t\leq c_{25}\sum\limits_{k = 0}^{n}\|\delta w_h^{k}\|_2^2\Delta t+c_{26}(\|w_{h}^{0}\|_2^{2}+\|\delta w_h^{0}\|_2^2). \end{align} (3.57)

    By Lemma 2.7 (discrete Gronwall inequality), we have

    \begin{align} \frac{f^{n}-f^{n-1}}{\Delta t} = \frac{\sum\nolimits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t-\sum\nolimits_{k = 0}^{n-1}\|\delta w_h^{k+1}\|_2^2\Delta t}{\Delta t} = \|\delta w_h^{n+1}\|_2^{2}, \end{align} (3.58)

    and from (3.57),

    \begin{align} &\left\|\delta w_h^{n+1}\right\|_2^2\leq c_{25}\sum\limits_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t+c_{26}(\|w_{h}^{0}\|_2^{2}+\|\delta w_h^{0}\|_2^2). \end{align} (3.59)

    Then, take f^n = \sum_{k = 0}^{n}\|\delta w_h^{k+1}\|_2^2\Delta t , g^{n+1} = c_{25} , h^{n+1} = c_{26}(\|w_{h}^{0}\|_2^{2}+\|\delta w_h^{0}\|_2^2) , and we can obtain

    \begin{align} &\sum\limits_{k = 0}^{n}\left\|\delta w_h^{k+1} \right\|_2^2\Delta t\leq c_{27}(\|w_{h}^{0}\|_2^{2}+\|\delta w_h^{0}\|_2^2). \end{align} (3.60)

    Substituting (3.60) into (3.57) yields

    \begin{align} &\left\|\delta w_h^{n+1}\right\|_2^2+a\sum\limits_{k = 0}^{n}\left\|\delta^2w_h^{k+1}\right\|_2^2\Delta t\leq C(\|w_{h}^{0}\|_2^{2}+\|\delta w_h^{0}\|_2^2). \end{align} (3.61)

    Thus the stability of the discrete scheme is demonstrated.

    Next, we will verify the correctness of the implicit finite volume discretization scheme for the magnetic diffusion equation with constant resistivity, and consider the following magnetic diffusion equation:

    \begin{align} \frac{\partial}{\partial t}B(x,t)-\frac{\partial}{\partial x}(\frac{\eta(e)}{\mu_{0}}\frac{\partial}{\partial x}B(x,t)) = 2t + \frac{2cos(x)\eta(e)}{\mu_0}. \end{align} (4.1)

    The solution interval is (x, t)\in[0, 0.5]\times[0, 1] . Take \mu_0 = 4\pi . The number of mesh segments is, respectively N = 40, 80,160 , and the number of nodes is, respectively, N_1 = 41, 81,161 . The resistivity \eta = 9.7\times10-3 . The length of the line segment L = 0.5 , the average length of the grid dx = L/N , T = 1 , and the time step dt = dx\ast dx . It is easy to know that the true solution to this problem is B(x, t) = 2cos(x)+t^2 . The error used in the experiment is L_2 , that is, Error_{L_2} = \sqrt{\frac{\Sigma_{n = 1}^{N_1}(B-B_{exact})^2}{N_1}} .

    Table 1.  Error order test (space).
    dt N Error_{L^2} Error ratio
    40 1.44E-04
    dx*dx 80 3.64E-05 3.97
    160 9.12E-06 3.99

     | Show Table
    DownLoad: CSV
    Figure 4.  Comparison of different space steps.
    Table 2.  Error Order test (time).
    N dt Error_{L^2} Error ratio
    40 0.01 9.20E-03
    40 0.005 4.60E-03 2.00
    40 0.0025 2.30E-03 2.00

     | Show Table
    DownLoad: CSV
    Figure 5.  Comparison of different time steps.

    Conclusions: From the above comparative experiments, it can be seen that when the grid size increases by 2 times with a fixed time scale, the error ratio between the experimental results and the true solution is close to 4. When the grid size is fixed, and the time scale increases by 2 times, the error ratio between the experimental results and the true solution is equal to 2, which conforms to the expected experimental errors of o(h^2) and o(t) , thus verifying the correctness of the implicit finite volume discretization scheme in this experiment.

    In this experiment, the true solution is denoted as B , and the perturbation solution is denoted as B_\varepsilon . B^{100} represents the true solution at time step dt = 0.01 , and B_\varepsilon^{100} represents the perturbation solution at time step dt = 0.01 . The error ratio still uses L_2 .

    Experiment (1): Step-function resistivity

    \begin{align} \left.\eta(e) = \eta(x,t) = \left\{\begin{array}{ll}\eta_S = 9.7\times10^{-5},&e\in[0,e_c],\\\eta_L = 9.7\times10^{-3},&e\in(e_c,+\infty).\end{array}\right.\right. \end{align}
    Table 3.  Error ratio of magnetic field under step resistivity.
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|B^{100}-B_{\varepsilon}^{100}\| ratio \|B^{1000}-B_{\varepsilon}^{1000}\| ratio \|B^{10000}-B_{\varepsilon}^{10000}\| ratio
    0.1 0.1779 0.2364 0.2409
    0.01 0.0191 9.31 0.0309 7.65 0.0328 7.34
    0.001 8.00E-04 23.87 0.0049 6.31 0.0032 10.25
    0.0001 8.02E-05 9.98 0.0015 3.27 1.77E-04 18.10

     | Show Table
    DownLoad: CSV
    Table 4.  Error ratio of internal energy density under step resistivity.
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|e^{100}-e_{\varepsilon}^{100}\| ratio \|e^{1000}-e_{\varepsilon}^{1000}\| ratio \|e^{10000}-e_{\varepsilon}^{10000}\| ratio
    0.1 1.96E-02 0.0224 0.0231
    0.01 6.60E-03 2.97 0.0039 5.74 0.0033 7.00
    0.001 2.10E-03 3.14 8.56E-04 4.55 6.36E-04 5.19
    0.0001 7.78E-06 269.95 2.69E-04 3.18 5.33E-05 11.93

     | Show Table
    DownLoad: CSV

    Conclusions: Under the same time step dt , when there is a small disturbance in the initial value, the error ratio changes significantly, indicating that the solution of the magnetic diffusion equation under step-function resistivity cannot be stable based on the initial value.

    Experiment (2): Constant resistivity \eta(e) = 9.7e-3 .

    Table 5.  Error ratio of magnetic field under constant resistivity \eta(e) = 9.7e-3 .
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|B^{100}-B_{\varepsilon}^{100}\| ratio \|B^{1000}-B_{\varepsilon}^{1000}\| ratio \|B^{10000}-B_{\varepsilon}^{10000}\| ratio
    0.1 0.0895 0.0895 0.0895
    0.01 0.009 9.94 0.009 9.94 0.009 9.94
    0.001 8.95E-04 10.05 8.95E-04 10.05 8.95E-04 10.05
    0.0001 8.95E-05 10.00 8.95E-05 10.00 8.95E-05 10.00

     | Show Table
    DownLoad: CSV
    Table 6.  Error ratio of internal energy density under constant resistivity.
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|e^{100}-e_{\varepsilon}^{100}\| ratio \|e^{1000}-e_{\varepsilon}^{1000}\| ratio \|e^{10000}-e_{\varepsilon}^{10000}\| ratio
    0.1 0.0186 0.0191 0.0197
    0.01 0.00185 10.05 0.002 9.55 0.002 9.85
    0.001 1.82E-04 10.16 1.9703E-04 10.15 2.0298E-04 9.85
    0.0001 1.82E-05 10.00 1.9708E-05 10.00 2.0303E-05 10.00

     | Show Table
    DownLoad: CSV

    Experiment (3): Linear resistivity \eta(e) = 9.7e-3 .

    \eta(e) = \frac{\eta_L-\eta_S}{2e_c}e+\eta_S,e\in[0,2e_c],

    where, e_c = 0.11084958 .

    Table 7.  Error ratio of magnetic field under linear resistivity \eta(e) = 9.7e-3 .
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|B^{100}-B_{\varepsilon}^{100}\| ratio \|B^{1000}-B_{\varepsilon}^{1000}\| ratio \|B^{10000}-B_{\varepsilon}^{10000}\| ratio
    0.1 4.37E-04 9.80E-05 8.08E-05
    0.01 4.05E-05 10.78 8.68E-06 11.29 7.30E-06 11.06
    0.001 4.02E-06 10.09 8.57E-07 10.13 7.22E-07 10.12
    0.0001 4.01E-07 10.01 8.56E-08 10.01 7.26E-08 9.94

     | Show Table
    DownLoad: CSV
    Table 8.  Error ratio of internal energy density under linear resistivity.
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|e^{100}-e_{\varepsilon}^{100}\| ratio \|e^{1000}-e_{\varepsilon}^{1000}\| ratio \|e^{10000}-e_{\varepsilon}^{10000}\| ratio
    0.1 0.1821 0.4073 0.4729
    0.01 0.0202 9.01 0.0443 9.19 0.0514 9.20
    0.001 0.002 10.10 0.0045 9.84 0.0052 9.88
    0.0001 2.04E-04 9.79 4.47E-04 10.07 5.18E-04 10.03

     | Show Table
    DownLoad: CSV

    Experiment (4): The step-function resistivity after polishing.

    Table 9.  Error ratio of magnetic field under the step-function resistivity after polishing.
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|B^{100}-B_{\varepsilon}^{100}\| ratio \|B^{1000}-B_{\varepsilon}^{1000}\| ratio \|B^{10000}-B_{\varepsilon}^{10000}\| ratio
    0.1 9.89E-02 9.93E-02 9.94E-02
    0.01 9.90E-03 9.99 9.90E-03 10.03 9.98E-03 9.96
    0.001 9.89E-04 10.01 9.98E-04 9.92 9.94E-04 10.04
    0.0001 9.87E-05 10.02 9.88E-05 10.11 9.87E-05 10.07

     | Show Table
    DownLoad: CSV
    Table 10.  Error ratio of internal energy density under the step-function resistivity after polishing.
    dt=0.01 dt=0.001 dt=0.0001
    \varepsilon \|e^{100}-e_{\varepsilon}^{100}\| ratio \|e^{1000}-e_{\varepsilon}^{1000}\| ratio \|e^{10000}-e_{\varepsilon}^{10000}\| ratio
    0.1 0.0012 0.01891 0.0189
    0.01 1.19E-04 10.08 0.0019 9.95 0.00198 9.55
    0.001 1.19E-05 9.98 1.8903E-04 10.05 1.9998E-04 9.90
    0.0001 1.19E-06 10.03 1.8708E-05 10.10 1.9803E-05 10.10

     | Show Table
    DownLoad: CSV

    Conclusion: From experiments (2), (3), and (4), it can be seen that under the same time step dt , when there is a small disturbance in the initial value, the error ratio does not change much. Especially, the solution of the magnetic diffusion equation under the smoothed step-function resistivity model has good stability. This is also the advantage of smoothed step-function resistivity \eta_\delta(e) compared to step-function resistivity \eta(e) .

    In the comparison experiment between explicit and implicit schemes, we take dt = \frac{c\mu_0*(dx)^2}{\eta_L} , where, \mu_0 = 4\pi , dx = L/N , \eta_L = 100\times9.7\times 10-3 . Therefore, c is the factor that affects dt , and the larger c is, the larger the time step dt .

    Conclusions: From the comparison experiment in the figure above, it is evident that when we use the curve at c = 0.4 as the true solution graph, as the value of c increases (that is, as the time step increases), the explicit solution gradually diverges from the true solution. In contrast, the implicit solutions remain nearly identical to the true solution, with differences only noticeable upon close inspection. This observation further demonstrates the strong stability and weak time step constraints of the fully implicit method. These characteristics particularly underscore the superiority of the fully implicit method, especially when dealing with models exhibiting strong nonlinearity.

    Figure 6.  Comparison of magnetic field and internal energy density.

    Gao Chang: Conceptualization, Writing–original draft, Data curation, Software, Investigation; Chunsheng Feng: Conceptualization, Software, Writing–review and editing, Methodology; Jianmeng He: Conceptualization, Writing–review and editing, Validation, Investigation; Shi Shu: Conceptualization, Supervision, Formal analysis, Methodology, Writing–review and editing. All authors have read and agreed to the published version of the manuscript.

    The authors declare that they have not used artificial intelligence (AI) tools in the creation of this article.

    The authors would like to express their gratitude to Professor Guangwei Yuan and Associate researcher Bo Xiao for their valuable contributions and support during the course of this research. This work was partially supported by the National Science Foundation of China (Grant No. 12371373), the National Key Research and Development Program of China (Grant No. 2023YFB3001604), the Hunan innovative province construction special project (Grant No. 2022XK2301), and the Postgraduate Scientific Research Innovation Project of Hunan Province.

    The authors declare that they have no conflicts of interest.



    [1] X. Y. Meng, N. N. Qin, H. F. Huo, Dynamics analysis of a predator-prey system with harvesting prey and disease in prey species, J. Biol. Dynam., 12 (2018), 342–374. https://doi.org/10.1080/17513758.2018.1454515 doi: 10.1080/17513758.2018.1454515
    [2] A. F. Nindjin, M. A. Aziz-Alaoui, M. Cadivel, Analysis of a predator-prey model with modified Leslie–Gower and Holling-type Ⅱ schemes with time delay, Nonlinear Anal.: Real World Appl., 7 (2006), 1104–1118. https://doi.org/10.1016/j.nonrwa.2005.10.003 doi: 10.1016/j.nonrwa.2005.10.003
    [3] K. Hassan, A. Mustafa, M. Hama, An eco-epidemiological model incorporating harvesting factors, Symmetry, 13 (2021), 2179. https://doi.org/10.3390/sym13112179 doi: 10.3390/sym13112179
    [4] I. Al-Darabsah, X. Tang, Y. Yuan, A prey-predator model with migrations and delays, Discrete Cont. Dyn. Syst.-B, 21 (2016), 737–761. https://doi.org/10.3934/dcdsb.2016.21.737 doi: 10.3934/dcdsb.2016.21.737
    [5] P. Kumar, S. Raj, Modelling and analysis of prey-predator model involving predation of mature prey using delay differential equations, Numer. Algebra, Control Optim., 12 (2022), 783–791. https://doi.org/10.3934/naco.2021035 doi: 10.3934/naco.2021035
    [6] R. K. Naji, A. N. Mustafa, The dynamics of an eco-epidemiological model with nonlinear incidence rate, J. Appl. Math., 2012 (2012), 1–24. https://doi.org/10.1155/2012/852631 doi: 10.1155/2012/852631
    [7] E, van Leeuwen, Å. Brännström, V. A. A. Jansen, U. Dieckmann, A. G. Rossberg, A generalized functional response for predators that switch between multiple prey species, J. Theor. Biol., 328 (2013), 89–98. https://doi.org/10.1016/j.jtbi.2013.02.003 doi: 10.1016/j.jtbi.2013.02.003
    [8] T. K. Kar, Stability analysis of a prey-predator model incorporating a prey refuge, Commun. Nonlinear Sci. Numer. Simul., 10 (2005), 681–691. https://doi.org/10.1016/j.cnsns.2003.08.006 doi: 10.1016/j.cnsns.2003.08.006
    [9] 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
    [10] H. Molla, M. S. Rahman, S. Sarwardi, Dynamics of a predator-prey model with Holling type Ⅱ functional response incorporating a prey refuge depending on both the species, Int. J. Nonlinear Sci. Numer. Simul., 20 (2019), 89–104. https://doi.org/10.1515/ijnsns-2017-0224 doi: 10.1515/ijnsns-2017-0224
    [11] M. F. Elettreby, Two-prey one-predator model, Chaos, Solitons Fract., 39 (2009), 2018–2027. https://doi.org/10.1016/j.chaos.2007.06.058 doi: 10.1016/j.chaos.2007.06.058
    [12] S. Sirisubtawee, N. Khansai, A. Charoenloedmongkhon, Investigation on dynamics of an impulsive predator–prey system with generalized Holling type Ⅳ functional response and anti-predator behavior, Adv. Differ. Equ., 2021 (2021), 1–26. https://doi.org/10.1186/s13662-021-03324-w doi: 10.1186/s13662-020-03162-2
    [13] K. Das, A mathematical study of a predator-prey dynamics with disease in predator, Int. Scholarly Res. Notices, 2011 (2011), 669–684. https://doi.org/10.5402/2011/807486 doi: 10.5402/2011/807486
    [14] H. W. Hethcote, W. Wang, L. Han, Z. Ma, A predator–prey model with infected prey, Theor. Popul. Biol., 66 (2004), 259–268. https://doi.org/10.1016/j.tpb.2004.06.010 doi: 10.1016/j.tpb.2004.06.010
  • This article has been cited by:

    1. Tahira Sumbal Shaikh, Muhammad Zafarullah Baber, Nauman Ahmed, Naveed Shahid, Ali Akgül, Manuel De la Sen, On the Soliton Solutions for the Stochastic Konno–Oono System in Magnetic Field with the Presence of Noise, 2023, 11, 2227-7390, 1472, 10.3390/math11061472
    2. H. G. Abdelwahed, A. F. Alsarhana, E. K. El-Shewy, Mahmoud A. E. Abdelrahman, The Stochastic Structural Modulations in Collapsing Maccari’s Model Solitons, 2023, 7, 2504-3110, 290, 10.3390/fractalfract7040290
    3. Yinlin Ye, Hongtao Fan, Yajing Li, Xinyi Liu, Hongbing Zhang, Conformable bilinear neural network method: a novel method for time-fractional nonlinear partial differential equations in the sense of conformable derivative, 2024, 0924-090X, 10.1007/s11071-024-10495-0
    4. Ateq Alsaadi, Fazal Dayan, Nauman Ahmed, Dumitru Baleanu, Muhammad Rafiq, Ali Raza, A novel method for the dynamics of worms in wireless sensor networks with fuzzy partition, 2023, 13, 2158-3226, 10.1063/5.0165342
    5. E. S. Aly, M. A. Sohaly, S. Z. Hassan, Noorjahan Abdul Azees, M. Daher Albalwi, Leema Aliyarukunju, Nadia A. Askar, The new wave structures to the perturbed NLSE via Wiener process with its wide-ranging applications, 2023, 13, 2158-3226, 10.1063/5.0174693
    6. Nauman Ahmed, Muhammad Waqas Yasin, Muhammad Sajid Iqbal, Ali Raza, Muhammad Rafiq, Mustafa Inc, A dynamical study on stochastic reaction diffusion epidemic model with nonlinear incidence rate, 2023, 138, 2190-5444, 10.1140/epjp/s13360-023-03936-z
    7. Shewafera Wondimagegnhu Teklu, Belela Samuel Kotola, Mathematical model and backward bifurcation analysis of pneumonia infection with intervention measures, 2024, 11, 2768-4830, 10.1080/27684830.2024.2419462
    8. Muhammad Waqas Yasin, Syed Muhammad Hamza Ashfaq, Nauman Ahmed, Ali Raza, Muhammad Rafiq, Ali Akgül, Numerical modeling of reaction–diffusion e-epidemic dynamics, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-05209-9
    9. Amani S. Baazeem, Yasir Nawaz, Muhammad Shoaib Arif, Kamaleldin Abodayeh, Mae Ahmed AlHamrani, Modelling Infectious Disease Dynamics: A Robust Computational Approach for Stochastic SIRS with Partial Immunity and an Incidence Rate, 2023, 11, 2227-7390, 4794, 10.3390/math11234794
    10. Muhammad Zafarullah Baber, Nauman Ahmed, Muhammad Waqas Yasin, Syed Mansoor Ali, Mubasher Ali, Ali Akgül, Murad Khan Hassani, Abundant soliton solution for the time-fractional stochastic Gray-Scot model under the influence of noise and M-truncated derivative, 2024, 6, 3004-9261, 10.1007/s42452-024-05759-8
    11. Mustafa Inc, Rubayyi T. Alqahtani, Muhammad Sajid Iqbal, Stability analysis and consistent solitary wave solutions for the reaction–diffusion regularized nonlinear model, 2023, 54, 22113797, 107053, 10.1016/j.rinp.2023.107053
    12. Muhammad Zafarullah Baber, Muhammad Waqas Yasin, Changjin Xu, Nauman Ahmed, Muhammad Sajid Iqbal, Numerical and Analytical Study for the Stochastic Spatial Dependent Prey–Predator Dynamical System, 2024, 19, 1555-1415, 10.1115/1.4066038
    13. Muhammad Zafarullah Baber, Nauman Ahmed, Changjin Xu, Muhammad Sajid Iqbal, Tukur Abdulkadir Sulaiman, A computational scheme and its comparison with optical soliton solutions for the stochastic Chen–Lee–Liu equation with sensitivity analysis, 2024, 0217-9849, 10.1142/S0217984924503767
  • 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(1728) PDF downloads(84) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog