Research article

A logarithmically improved regularity criterion for the 3D MHD equations in Morrey-Campanato space

  • In this paper, we will establish a sufficient condition for the regularity criterion to the 3D MHD equation in terms of the derivative of the pressure in one direction. It is shown that if the partial derivative of the pressure 3π satisfies the logarithmical Serrin type condition 3π satisfies the logarithmical Serrin type condition T03π(s)22rM2,3r1+ln(1+b(s)L4)ds for 01,
    then the solution (u,b) remains smooth on [0,T]. Compared to the Navier-Stokes result, there is a logarithmic correction involving b in the denominator.

    Citation: Sadek Gala, Maria Alessandra Ragusa. A logarithmically improved regularity criterion for the 3D MHD equations in Morrey-Campanato space[J]. AIMS Mathematics, 2017, 2(1): 16-23. doi: 10.3934/Math.2017.1.16

    Related Papers:

    [1] Songxiao Li, Jizhen Zhou . Essential norm of generalized Hilbert matrix from Bloch type spaces to BMOA and Bloch space. AIMS Mathematics, 2021, 6(4): 3305-3318. doi: 10.3934/math.2021198
    [2] Keun Young Lee, Gwanghyun Jo . The dual of a space of compact operators. AIMS Mathematics, 2024, 9(4): 9682-9691. doi: 10.3934/math.2024473
    [3] Babar Sultan, Mehvish Sultan, Aziz Khan, Thabet Abdeljawad . Boundedness of an intrinsic square function on grand p-adic Herz-Morrey spaces. AIMS Mathematics, 2023, 8(11): 26484-26497. doi: 10.3934/math.20231352
    [4] Gang Wang . Some properties of weaving K-frames in n-Hilbert space. AIMS Mathematics, 2024, 9(9): 25438-25456. doi: 10.3934/math.20241242
    [5] Liu Yang, Ruishen Qian . Volterra integral operator and essential norm on Dirichlet type spaces. AIMS Mathematics, 2021, 6(9): 10092-10104. doi: 10.3934/math.2021586
    [6] Kifayat Ullah, Junaid Ahmad, Hasanen A. Hammad, Reny George . Iterative schemes for numerical reckoning of fixed points of new nonexpansive mappings with an application. AIMS Mathematics, 2023, 8(5): 10711-10727. doi: 10.3934/math.2023543
    [7] Rana Safdar Ali, Saba Batool, Shahid Mubeen, Asad Ali, Gauhar Rahman, Muhammad Samraiz, Kottakkaran Sooppy Nisar, Roshan Noor Mohamed . On generalized fractional integral operator associated with generalized Bessel-Maitland function. AIMS Mathematics, 2022, 7(2): 3027-3046. doi: 10.3934/math.2022167
    [8] Bo Xu . Bilinear θ-type Calderón-Zygmund operators and its commutators on generalized variable exponent Morrey spaces. AIMS Mathematics, 2022, 7(7): 12123-12143. doi: 10.3934/math.2022674
    [9] Shahid Mubeen, Rana Safdar Ali, Iqra Nayab, Gauhar Rahman, Thabet Abdeljawad, Kottakkaran Sooppy Nisar . Integral transforms of an extended generalized multi-index Bessel function. AIMS Mathematics, 2020, 5(6): 7531-7547. doi: 10.3934/math.2020482
    [10] Suixin He, Shuangping Tao . Boundedness of some operators on grand generalized Morrey spaces over non-homogeneous spaces. AIMS Mathematics, 2022, 7(1): 1000-1014. doi: 10.3934/math.2022060
  • In this paper, we will establish a sufficient condition for the regularity criterion to the 3D MHD equation in terms of the derivative of the pressure in one direction. It is shown that if the partial derivative of the pressure 3π satisfies the logarithmical Serrin type condition 3π satisfies the logarithmical Serrin type condition T03π(s)22rM2,3r1+ln(1+b(s)L4)ds for 01,
    then the solution (u,b) remains smooth on [0,T]. Compared to the Navier-Stokes result, there is a logarithmic correction involving b in the denominator.


    Nonlinear partial differential equations (PDEs) are crucial for studying nonlinear physical phenomena in various models across different branches. Analyzing qualitative parameters is necessary to comprehend the underlying facts of a model. Nonlinear evolution equations have significant implications in applied sciences, such as optics, crystal lattice theory, fluid dynamics, and plasma physics, making them an excellent source of motivation in recent decades. There are several analytical methods for solving nonlinear partial differential equations (NLPDEs), including the Jacobi elliptic expansion [1,2,3], the tanh-function method [4,5,6], the improved Q-expansion technique [7], the generalized direct algebraic method [8], the extended generalized (G/G)-expansion method [9,10], and others [11,12,13,14,15,16,17,18,19,20,21]. In 1995, Gilson and Pickering presented a third-order nonlinear differential equation explaining wave propagation in plasma physics[22]. This equation is widely recognized:

    utαuxxt+2γuxuuxxxμ(u22)xβuxuxx=0. (1.1)

    There are four constants, α,μ,γ,β, in the Gilson-Pickering equation, all of which are nonzero. The equation has three different types: Rosenau-Hyman equation when α=0, μ=1, γ=0, β=3; Fuchssteiner-Fokas-Camassa-Holm equation when α=1, μ=3, β=2; and Fornberg-Whitham equation when α=1, μ=1, β=3, γ=0.5. Recently, there have been several numerical schemes developed for solving NLPDEs. These include finite element, finite differences [23,24,25,26,27], the adaptive moving mesh technique [28,29,30,31,32,33], and the parabolic Monge-Ampère (PMA) method [34,35,36,37]. Lately, various computational methods have been employed to explore the Gilson-Pickering equation, including Chen et al. [38] who performed a qualitative analysis of this equation. Clarkson et al. investigated the symmetry reductions of Eq (1.1) using both the classical Lie method and the non-classical method [39], the (1/G) method [40], and the (G/G2) method [41]. Additionally, Baskonus provides several complex soliton solutions using the Bernoulli sub-equation function method [42]. There are many researchers who investigated the Gilson-Pickering equation by different methods [43,44,45,46,47]. The PMA approach mentioned in this context has yet to be used for Eq (1.1). This motivated us significantly to apply it. It is widely acknowledged that the wave solutions in this situation have areas with rapid spatial variations, such as steep front structures. However, it is important to note that this method's effectiveness in minimizing errors is specifically limited to regions with rapid spatial variations, such as steep front structures. However, this method is effective in minimizing errors in these specific regions.

    The following organization is presented in this article: Section 2 provides exact solutions to the Gilson-Pickering equation using the F-expansion method and the (1/G)-expansion method. Section 3 introduces a numerical solution for the same equation on the uniform mesh utilizing implicit finite difference with a study of the stability, error analysis, and convergence of the numerical scheme. Section 4 on the moving mesh method applies the PMA technique to the Gilson-Pickering equation, while the results and discussion are in Section 5. Ultimately, the conclusions are in Section 6.

    We consider a nonlinear evolution equation with some physical fields u(x,t) in two variables x and t as follows:

    Ω1(u,ut,ux,uxx,uxxt,uxxx,)=0. (2.1)

    Step 1. We look for the traveling-wave solutions of Eq (2.1) that are formed as follows:

    u(x,t)=U(ξ),ξ=kxct, (2.2)

    where k,c are constant, and c is the wave speed.

    Step 2. Using Eqs (2.1) and (2.2) we directly reduce to

    Ω2(U(ξ),U(ξ)ξ,U(ξ)ξξ,U(ξ)ξξξ,)=0. (2.3)

    To convert Eq (1.1) into an ordinary differential equation (ODE), substitute the wave transformation Eq (2.2) into it.

    cU+ck2αU+2kγUk3UUkμUUk3βUU=0, (2.4)

    where denotes the derivative with respect to ξ. We can obtain the following result by integrating once for ξ and putting the constant of integration to zero:

    cU+ck2αU+2kγUk3UU+k3(U)22kμ(U22)k3β(U)22=0. (2.5)

    The F-expansion procedure [7] introduces the traveling wave solution of Eq (2.3) of the form

    UN(ξ)=a0+Nj=1ajF(ξ)j+Nj=1bjF(ξ)j, (2.6)

    where a0,aj, and bj are constants to be determined. The function F(ξ) represents a solution of the following differential equation:

    F(ξ)=A0+A1F(ξ)+A2F(ξ)2, (2.7)

    where A0, A1, and A2 are constant [7]. By balancing the second derivative of U with the nonlinear term U2 in Eq (2.5) we required that N+2=2N. Then, we get the value of N as N=2, and Eq (2.6) becomes

    U(ξ)=a0+a1F(ξ)+b1F(ξ)+a2F(ξ)2+b2F(ξ)2. (2.8)

    To solve for the values of a0, a1,a2, b1,b2, k, and c, we can substitute Eq (2.8) into Eq (2.4) and use Eq (2.7). To create a system of algebraic equations, we collect coefficients of F(ξ)l, l=(6,5,,5,6). Then, solving this system using Mathematica 13.2 software will give us the desired values as follows:

    Case I:a0=b1,  k=ϵβ1,  c=±2β1γϵαϵ+β+1,  ba1=a2=b2=0;Case II:a0=4αγ+αb1ϵβb1+b1αϵ+β1,  c=±2(β1)γϵβ1(αϵ+β1),k=±ϵβ1,  a1=a2=b2=0;Case III:a0=b2=b12,  k=±ϵ2β1,  c=β1γϵαϵ+β+1,  a1=a2=0;Case IV:a0=8αγ+b1(αϵ+β1)2αϵ2β+2,  c=±(β1)γϵβ1(αϵ+β1),b2=b12,  k=±ϵ2β1,  a1=a2=0;

    when A0=0, A1=1, A2=1.

    The solution to Eq (1.1) can be determined through Case I and is as follows:

    u1(x,t)=b1(2tanh(ϵ(±2(β+1)γtx(αϵ+β+1))2β1(αϵ+β+1))+11). (2.9)

    Based on Case II, the solution to Eq (1.1) can be determined:

    u2(x,t)=4αγ+b1(αϵ+β1)αϵβ+12b1tanh(ϵ(2(β1)γt+x(αϵβ+1))2β1(αϵ+β1))1. (2.10)

    Using Case III, we can determine the solution for Eq (1.1):

    u3(x,t)=12b1(sinh(ϵ(±x(αϵ+β+1)2(β+1)γt)β1(αϵ+β+1))cosh(ϵ(±x(αϵ+β+1)2(β+1)γt)β1(αϵ+β+1))). (2.11)

    According to Case IV, we can find the solution to Eq (1.1):

    u4(x,t)=4αγαϵβ+112b1(sinh(ϵ(2(β1)γt+x(αϵβ+1))β1(αϵ+β1))±cosh(ϵ(2(β1)γt+x(αϵβ+1))β1(αϵ+β1))), (2.12)

    when A0=0,A1=1,A2=1.

    The solution to Eq (1.1) can be determined through Case I and is as follows:

    u5(x,t)=b1(2coth(ϵ(±2(β+1)γtx(αϵ+β+1))2β1(αϵ+β+1))11). (2.13)

    Based on Case II, the solution to Eq (1.1) can be determined:

    u6(x,t)=4αγ+b1(αϵ+β1)αϵβ+1±2b1coth(ϵ(2(β1)γt+x(αϵβ+1))2β1(αϵ+β1))±1. (2.14)

    Using Case III, we can determine the solution for Eq (1.1):

    u7(x,t)=12b1(sinh(ϵ(±x(αϵ+β+1)2(β+1)γt)β1(αϵ+β+1))+cosh(ϵ(±x(αϵ+β+1)2(β+1)γt)β1(αϵ+β+1))). (2.15)

    According to Case IV, we can find the solution to Eq (1.1):

    u8(x,t)=4αγαϵβ+1±12b1(sinh(ϵ(2(β1)γt+x(αϵβ+1))β1(αϵ+β1))cosh(ϵ(2(β1)γt+x(αϵβ+1))β1(αϵ+β1))). (2.16)

    This subsection presents the important steps of the (1/G)-expansion method. Assume that Eq (2.3) has the solution of the form

    UN(ξ)=Nj=0aj(1G(ξ))j. (2.17)

    The scalars a0,a1,,aN need to be determined, and N represents the balance terms. Additionally, G=G(ξ) must satisfy the given second-order linear ordinary differential equation

    G(ξ)+ϑG(ξ)+ν=0, (2.18)

    where ϑ and ν are constants. The solution of Eq (2.18) is given by

    G(ξ)=νϑ+beϑξ. (2.19)

    We can use the following format to express the algebraic equation provided in Eq (2.19) as a trigonometric function

    G(ξ)=νϑ+(bcosh(ϑξ)bsinh(ϑξ)). (2.20)

    After applying the balance principle to Eq (2.5) we required that N+2=2N, and then we obtain N=2. Equation (2.17) takes the following form:

    U(ξ)=a0+a1(1G(ξ))+a2(1G(ξ))2. (2.21)

    By inserting Eq (2.21) and its necessary derivatives, along with Eq (2.18) into Eq (2.3), we can obtain the polynomial of (1G(ξ))l, l0. By summing the coefficients of (1G(ξ))l with the same power and setting the resulting summation to zero, we can derive a system of algebraic equations for a0,a1 and a2. Solving this system using Mathematica 13.2 software yet gives us the values of the variables, which in turn provides the solutions to Eq (1.1).

    Family 1:

    a0=0,1=24αγk2νϑ(αk2ϑ21)(k2ϑ2ϵ),  a2=24αγk2ν2(αk2ϑ21)(ϵk2ϑ2),β=2,  c=2γkαk2ϑ21.

    We can obtain the solution for Eq (1.1) as follows:

    u9(x,t)=24αbγk2νϑ3exp(kϑ(2γtαk2ϑ21+x))(αk2ϑ21)(k2ϑ2ϵ)(νbϑexp(kϑ(2γtαk2ϑ21+x)))2. (2.22)

    In Figures 13, we show the exact solutions.

    Figure 1.  The 3D graph presents the analytic (left) and 2D plot (right) solutions for u9(x,t). The used parameters are assumed as follows: α=0.1,γ=0.2,k=1,ν=0.09,ϑ=1.2,ϵ=1.1,b=0.01, t=010 and x=1010.
    Figure 2.  The analytical solution of u10(x,t) is displayed in the left 3D surface, while the 2D plot is in the right figure. The parameters are taken as follows: α=0.1,γ=1/7,k=1.6,ν=0.03,ϑ=1.01,ϵ=0.8,b=0.02,t=010 and x=1010.
    Figure 3.  The left plot illustrates the 3D of u11, while the right graph shows the 2D plot. The parameter values are α=1,γ=1/8,ν=0.05,ϑ=1,ϵ=1.5,b=0.02,t=020 and x=1010.

    Family 2:

    a0=4αγk2ϑ2αk2ϵϑ2+2k2ϑ2+ϵ,   a1=24αγk2νϵϑ(ϵk2ϑ2)(αk2ϵϑ2+2k2ϑ2+ϵ),a2=24αγk2ν2ϵ(ϵk2ϑ2)(αk2ϵϑ2+2k2ϑ2+ϵ),   β=2,   c=2(2γk3ϑ2+γkϵ)αk2ϵϑ2+2k2ϑ2+ϵ.

    The solution for the traveling wave of Eq (1.1) can be expressed as

    u10(x,t)=24αγk2ν2ϵS(ϵk2ϑ2)(bexp(ϑ(kx2t(2γk3ϑ2+γkϵ)S))νϑ)2+24αγk2νϵϑS(ϵk2ϑ2)(bexp(ϑ(kx2t(2γk3ϑ2+γkϵ)S))νϑ)+4αγk2ϑ2S, (2.23)

    where

    S=αk2ϵϑ2+2k2ϑ2+ϵ.

    Family 3:

    a0=4αγαϵ+1,   a1=4αγνϑ(αϵ+1),   a2=0,k=ϵϑ,   β=3,   c=2γϵϑ(αϵ+1),   ϵ<0.

    The solution of Eq (1.1) can be determined as follows:

    u11(x,t)=4αbγϑ(αϵ+1)(bϑνexp(ϵ(2γt+αxϵ+x)αϵ+1)). (2.24)

    Figures 4 and 5 show the wave behavior of the solutions u9(x,t) and u10(x,t), respectively, when changing specific parameters and fixing others. The amplitude increases as α and γ increase. Figure 4 shows that when α is positive, the wave is in the positive direction, while in Figure 5, the wave is in the negative direction. However, parameter γ has the opposite effect of parameter α on the wave direction between the solutions u9(x,t) and u10(x,t).

    Figure 4.  These figures show the wave behavior of changing a specific parameter value while fixing the others. The left figure plots the exact solution u9(x,t) as γ increases and α=106, while the right plot shows the behavior when α increases and γ is fixed at 1.
    Figure 5.  The wave behavior of changing a specific parameter while fixing the others is depicted in these figures. The left plot explicitly demonstrates the behavior of the exact solution u10(x,t) as γ increases and α=104, while the right plot highlights the behavior when α increases and γ remains at 17.

    To obtain the numerical solutions of Eq (1.1) in this section, over the domain axb, 0ttf, where tf is a certain time, these coordinates are covered by a rectangular grid of points

    x=xm=a+(m1)Δx,m=1,2,,Nx+1,t=tn=nΔt,n=0,1,2,,

    where Nx represents the nummber of grid points, Δx=(ba)/Nx represents the distance between any two consecutive points, and Δt denote the temporal increases.

    To start, we write the variable v in the following form:

    v=uαuxx. (3.1)

    Thus, Eq (1.1) is reformed as

    vt+2γuxuuxxxμ(u22)xβuxuxx=0. (3.2)

    Then, we use the approximation solutions Unm to the exact solution u(xm,tn)=u. To calculate spatial derivatives, we utilize finite difference operators while keeping the temporal differentiation continuous. The weighted average or (θ-method) approximation [48] for Eq (3.2) is as follows:

    (Vt)nm+γΔxδx(θUn+1m+(1θ)Unm)12Δ3x(θUn+1m+(1θ)Unm)δ3x(θUn+1m+(1θ)Unm)μ4Δxδx(θ(Un+1m)2+(1θ)(Unm)2)β2Δ3xδx(θUn+1m+(1θ)Unm)δ2x(θUn+1m+(1θ)Unm)=0, (3.3)

    where

    δ3xUnm=(Unm+22Unm+1+2Unm1Unm2),  δ2xUnm=(Unm+12Unm+Unm1),  δxUnm=(Unm+1Unm1).

    We will assume we are using an average with nonnegative weights so that 0θ1. We will analyze two implicit methods: the fully implicit scheme when θ=1 and the well-known and popular Crank-Nicolson scheme when θ=12. The boundary conditions for Eq (3.2) are ut,a=ut,b=0. We set t=0 in Eq (2.22) to obtain the initial condition. The above system is solved utilizing an ODE solver in FORTRAN 95 called the DASPK solver [49,50]. To determine the accuracy of the numerical scheme (3.3), we utilize Taylor expansions to analyze the truncation error. Specifically, we apply Taylor's series expansion for all terms in Eq (3.3) around Unm, where u represents the exact solution, then we get

    [utαuxxt+2γ(θux+(1θ)ux)(θu+(1θ)u)(θuxxx+(1θ)uxxx)μ2(θu2x+(1θ)u2x)β((θux+(1θ)ux)(θuxx+(1θ)uxx))]+[Δt2uttαΔt2uxxtt+2γ(θΔtuxt)(θ2Δtuuxxxt+θ(1θ)Δtuuxxxt+θ2Δtutuxxx+θ(1θ)Δtutuxxx)μ2(θΔtu2xt)β(θ2Δtuxuxxt+θ(1θ)Δtuxuxxt+θ2Δtuxtuxx+θ(1θ)Δtuxtuxx)]+[Δ2t6uttαΔ2x12uxxxt+2γ(θΔ2x6uxxx+(1θ)Δ2x6uxxx)(θΔ2t2utt(θuxxx+(1θ)uxxx))μ2(θΔ2x6u2xxx+(1θ)Δ2x6u2xxx)β(θ2Δ2tuxtuxx+(1θ)Δ2xΔt12uxtuxxxx)]+. (3.4)

    Since u is the exact solution of Eq (1.1), for Crank-Nicolson scheme θ=12, the first and second brackets are equal to zero. As a result, in both time and space, the scheme is second-order. Therefore, the truncation error of the numerical method becomes O(Δ2x,Δ2t). With the fully implicit scheme of θ=1, the first bracket equals zero, and the truncation error becomes O(Δ2x,Δt). We will examine the stability of the numerical solution. To begin, we will express Eq (1.1) in the following form:

    utαuxxt+2γuxuuxxxμ(u22)xβ(uuxx)x+βuuxxx=0. (3.5)

    Our numerical scheme's stability is displayed by using von-Neumann stability analysis. The von-Neumann analysis, further called Fourier analysis, is only used when the scheme is linear. The differential Eq (3.5) must be written in the linearized form:

    utαuxxt+2γuxˆwuxxxμˆwux=0, (3.6)

    where ˆw=max|u|. The general form of the implicit scheme for Eq (3.6) can be written as:

    1Δt(Un+1mUnm)αΔtΔ2xδ2x(Un+1mUnm)+γΔxδx(θUn+1m+(1θ)Unm)ˆw2Δ3xδ3x(θUn+1m+(1θ)Unm)μˆw2Δxδx(θUn+1m+(1θ)Unm)=0. (3.7)

    Assume that

    Unm=λneιωmΔx,ι=1. (3.8)

    When using von-Neumann, |λ|1 is required for stability. The following result was obtained by substituting Eq (3.8) into Eq (3.7), and canceling the common factor Unm, and doing some operations

    (λ1)+αΔ2x(4sin2(ωΔx2))(λ1)+2ιγΔtΔxsin(ωΔx)(θλ+(1θ))+3ιˆwΔtΔ3xsin(ωΔx)sin2(ωΔx2)(θλ+(1θ))ιμˆwΔtΔxsin(ωΔx)(θλ+(1θ))=0. (3.9)

    Let

    1+4αΔ2xsin2(ωΔx2)=φ1,sin(ωΔx)(2γΔtΔx+3ˆwΔtΔ3xsin2(wΔx2)μˆwΔtΔx)=φ2.

    Therefore, Eq (3.9) can be written as

    (λ1)φ1+ιφ2(θλ+(1θ))=0. (3.10)

    Hence,

    λ(φ1+ιφ2θ)=φ1ιφ2(1θ). (3.11)

    As a result, λ can be expressed explicitly as

    λ=φ1ιφ2(1θ)φ1+ιφ2θ. (3.12)

    Let us analyze the stability of two schemes: the Crank-Nicolson scheme and the fully implicit scheme. Starting with the Crank-Nicolson scheme, we can determine its stability by plugging in θ=12 into Eq (3.12)

    λ=φ1ι12φ2φ1+ι12φ2. (3.13)

    Then, using Eq (3.13), we find that |λ|=1, meaning that the Crank-Nicolson scheme is linearly unconditionally stable. Moving on to the fully implicit scheme, we substitute the value of θ=1 into Eq (3.12) to get Eq (3.14)

    λ=φ1φ1+ιφ2. (3.14)

    By analyzing Eq (3.14), we find that |λ|<1, indicating that the fully implicit scheme is also linearly unconditionally stable. To discuss the convergence of our numerical method, we have some initial data and perform a sequence of computations using two refined meshes, where Δx and Δt both approach zero. If, for a given point in the rectangular domain [a,b]×[0,tf], with coordinates (x,t), we have xm and tn approaching x and t, respectively, this means that the numerical scheme is convergent. In other words, Unm=u(x,t). Assume that

    en+1m=Un+1mu(xm,tn+1) (3.15)

    at the point (xm,tn+1), the approximation solution is represented by Un+1m, and the analytical solution is represented by u(xm,tn+1). The error at this point is denoted by en+1m

    En+1=max1mNx{|en+1m|},n0. (3.16)

    We insert Eq (3.16) into Eq (3.3) to obtain

    1Δt(En+1En)=Tnm+αΔtΔ2xδ2x(En+1En)γΔxδx(θEn+1+(1θ)En)+12Δ3x(θEn+1+(1θ)En)δ3x(θEn+1+(1θ)En)+μ4Δxδx(θ(En+1)2+(1θ)(En)2)+β2Δ3xδx(θEn+1+(1θ)En)δ2x(θEn+1+(1θ)En). (3.17)

    We have

    En+1En+Δt(Tnm). (3.18)

    We can identify using the provided initial data E0=0. As a result, the inequality is expressed as

    Enn×Δt(Tnm). (3.19)

    Then,

    Enn×Δt(Tnm)0,asΔt0. (3.20)

    Hence, Eq (3.3) is convergent in both the Crank-Nicolson scheme and in the fully implicit method, as Δx,Δt0.

    Figures 6 and 7 illustrate the behavior of the analytical and numerical solutions of Eq (2.22). The parameter values are taken as α=1.2, γ=1.2, k=0.5, ν=1.8, ϑ=1.2, ϵ=0.1, b=1.5, and x=5020. The numerical solution developed for Eq (2.23) aligns with the exact solution of the equation. Figures 8 and 9 demonstrate that the exact and numerical solutions for Eq (2.24) behave similarly when α=1, γ=2, ν=0.05, ϑ=1, ϵ=1.5, b=0.2, and x=5010.

    Figure 6.  The left figure displays the analytical solution for Eq (2.22), whereas the right surface depicts the numerical solution's behavior.
    Figure 7.  Alteration in the time of the numerical solutions with Nx=30000 for t=0:2:8. The waveform at t=8 demonstrates a notable similarity between the numerical and analytical solutions.
    Figure 8.  The figures compare the analytical solution (left) and the numerical solution of u11(x,t) (right). These 3D figures assess the performance of the numerical method with the analytical solution.
    Figure 9.  The numerical results for u11(x,t) have been displayed at time t=0:1:5. The wave at time t=5 demonstrates the coincidence of the numerical and analytic solutions.

    To find an approximate solution to a PDE numerically, we divide the domain into small parts or nodes, creating a mesh. This mesh should be fine enough to capture areas of the solution where the values change rapidly, such as steep fronts or shock-like structures. The step size, Δx, representing the distance between adjacent nodes, should be small enough to resolve these features accurately across the entire domain. While this method is effective, it comes at the cost of intensive computation and high expense. Additionally, it only applies to problems that do not change over time. If the solution changes with time, the mesh must constantly be redistributed. Therefore, a more efficient and alternative approach is needed to reduce the error in regions where the solution has a high curvature. In the PMA method, mesh points are redistributed in time to adapt to changes in the solution. This involves generating moving mesh nodes using the potential function's gradient as the mesh generator. The PMA equation has the advantage of requiring one fewer equation to solve for the mesh in two dimensions, and the resulting meshes are typically regular and free of tangles compared with the moving mesh partial differential equation (MMPDE) method. The MMPDE method is another technique for generating meshes in numerical simulations, particularly for problems involving moving interfaces or regions with high solution gradients. To examine the numerical solution of Eq (3.2) on the moving mesh, the PMA approach is applied. We will now explore a new transformation

    x=x(η,t):[0,1][a,b],t>0. (4.1)

    The physical and computational coordinates are represented by x and η, respectively. The solution for u is derived from this:

    u(x,t)=u(x(η),t). (4.2)

    We now divide the physical domain into equal subintervals as follows:

    x1<x2<<xNx<xNx+1. (4.3)

    Therefore, the moving mesh is rewritten as

    xm=x(ηm,t),whereηm=(m1)Nx,m=1,,Nx+1. (4.4)

    To obtain the physical coordinate x, we can use the gradient of the mesh potential P(η,t), which is achieved using the one-dimensional PMA mesh equation:

    τ(1ρηη)˙P=Λ(Pη,t)Pηη,    x=Pη. (4.5)

    The monitor function is known as Λ(Pη,t). Additionally, τ indicates a relaxation parameter. When the value of τ is smaller, the mesh responds more quickly to changes in Λ(x,t). Conversely, the mesh moves more slowly when the value of τ is extensive. ρR is a positive smoothing parameter. It is subject to the boundary condition given by

    Pη,1=a,Pη,Nx+1=b, (4.6)

    and the initial condition is chosen to be

    P(ηm,t=0)=(ηm)22,m=1,,Nx+1, (4.7)

    where ηm[0,1] is the computational coordinate. We get the following using the chain rule:

    ux=uηPηη,ut=˙uuηPηη˙Pη. (4.8)

    Thus, the expression of Eq (3.2) is given by

    ˙v(vηPηη)˙Pη=2γ(uηPηη)+u(1Pηη(1Pηη(uηPηη)η)η)+μ2(u2ηPηη)+β(uηPηη)(1Pηη(uηPηη)η),v=uαPηη(uηPηη)η, (4.9)

    where m=1,,Nx+1. The boundary conditions

    ut,1=ut,Nx+1=0,

    and the initial condition are determined based on the solution of Eq (1.1) at t=0. To obtain the coordinate transformation x(η), we use the 1D PMA mesh equation

    ˙PρΔ2η(˙Pm+12˙Pm˙Pm1)=1τΛ(Pη,t)1Δ2η(Pm+12PmPm1), (4.10)

    where Δη represents the step size of the computational coordinate.

    The computational coordinate η, in this problem, is determined by

    ηm=a+(m1)Δη,m=1,,Nx+1, (4.11)

    where

    Δη=baNx.

    Then, the physical coordinate x is defined by xm=x(ηm,t), where the boundary grids are forced to be x1=a and xNx+1=b. Thus, the location of the grids xm is determined as follows:

    xm=Pm+1Pm12Δη,m=2,,Nx. (4.12)

    Pm=P(ηm,t) is the mesh potential obtained by solving Eq (4.10). Therefore, the semi-discretisation of Eq (4.9) is defined by

    ˙V=(Vm+1Vm1Pm+22Pm+Pm2)(˙Pm+1˙Pm1)4γΔη(Um+1Um1Pm+22Pm+Pm2)+2Δη(Um((Uxx)m+1(Uxx)m1)Pm+22Pm+Pm2)+μΔη((U2)m+1(U2)m1Pm+22Pm+Pm2)+2βΔη(Um+1Um1Pm+22Pm+Pm2)(Uxx)m, (4.13)
    Uxx=8Δ2ηPm+22Pm+Pm2(Um+1UmPm+2Pm+1Pm+Pm+1UmUm1Pm+1PmPm1+Pm2). (4.14)

    For a successful moving mesh method, it is essential to select a monitor function carefully. This function helps increase mesh density in areas of high interest in the PDE solution, such as propagating fronts or singularities. The monitor function should be stable under numerical discretizations, meaning that the numerical solution should not exhibit excessive oscillations or other problems. The monitor function should be computationally efficient to evaluate, as it is typically called at each iteration of the PMA algorithm. Furthermore, using two monitor functions together is common when employing adaptive methods to solve NPDEs [29].

    For example, one monitor function could track the error in the solution, while another could track the smoothness of the solution. The grid would then be refined in areas where the error is significant or the smoothness is poor. In general, using two monitor functions can be a valuable tool for improving the performance of adaptive methods. By tracking different aspects of the solution, the adaptive method can ensure that the grid is refined in the areas most needed, leading to more accurate and efficient solutions.

    Curvature monitor function:

    Λ(x,t)=1+ˆρu2xx, (4.15)

    where ˆρ is the user-specified parameter. If u is not smooth, the discretized monitor function may change abruptly and cause slow computation. It is common practice in most standard moving mesh methods to smooth the monitor function to achieve a smoother mesh. A simple yet effective smoothing scheme involves weighted averaging as its essential ingredient, as cited in [29,51]. Weighted averaging can help to smooth out the monitor function and reduce oscillations in the mesh. This can lead to a more accurate and stable solution to the underlying PDE. In addition to weighted averaging, other smoothing schemes can be used for monitor functions, such as filtering and interpolation. The choice of smoothing scheme depends on the specific application and the desired properties of the mesh [32].

    From Figures 1012, adaptive methods are crucial in resolving intricate spatial scales and rapidly changing temporal behaviors in various applications. The fundamental idea behind the PMA equation method is to increase the number of points in regions with higher curvature. An important aspect of implementing this method is the transformation between the computational and physical domains.

    Figure 10.  The analytical solution of u9(x,t) obtained using the (1/G)-expansion technique and the numerical solution obtained by the PMA method are compatible. This figure is plotted at t=5.
    Figure 11.  The curvature-based density function.
    Figure 12.  The associated time development of x(η,t).

    Finding a suitable monitor function that works well with the underlying PDE is a crucial aspect of the success of a PMA method. Additionally, it is important to mention that the mesh redistribution is carried out by utilizing the monitor function, which oversees the mesh evaluation. The choice of monitor function can also significantly impact the PMA algorithm's convergence rate. Sometimes, it may be necessary to experiment with different monitor functions to find one that works well for a particular problem.

    The choice of monitor function is ultimately a problem-specific decision. However, by understanding the key properties of a good monitor function, researchers can make informed choices that lead to more accurate and efficient solutions to their problems. It is helpful to understand the concept of a non-uniform mesh and how it can be used to improve the accuracy of numerical simulations. Figure 12 shows the associated time development of a non-uniform mesh. The mesh is initially uniform, but it becomes non-uniform over time. The solution changing rapidly between 30 and 10 highlights the importance of refining the mesh in areas with significant variation. This refinement allows the simulation to capture the details of the solution more accurately, which can be crucial for problems with steep gradients or sharp transitions.

    We have successfully applied two analytical methods to extract traveling wave solutions of the Gilson-Pickering equation. We use the F-expansion method and the (1/G)-expansion strategy. Figure 4 presents the behavior of u9(x,t) when we change the values of γ and α, while the other parameters take fixed values k=1, ϑ=0.2, ν=0.01, ϵ=3.1, and b=1. Figure 5 illustrates the wave behavior of u10(x,t) with the values of parameters k=0.1, ϑ=1.01, ν=0.3, b=2, and ϵ=0.8. As a result, these plots show that the value of α,γ affect the direction and the amplitude of the wave. Figures 4 and 5 are displayed with Nx=1000 and at time t=1. We verify the solutions with numerical results using two numerical techniques. First, we convert the underlying problems into a system of ODEs using the finite difference method while keeping time derivatives continuous. Then, we solve the resulting system of ODEs using the differential-algebraic system problem solver kit (DASPK) and the PMA approach, where the PMA technique reduces the error in the region where the solution has a significant variation. This method gives reliable and robust results. This can be seen in the graphical comparisons in the figures mentioned above.

    In Figure 13, the numerical solutions for u9(x,t) differ from the exact solution for different values of Nx. It is evident from the figure that as Nx becomes large, the numerical solutions of u9(x,t) converge towards the exact solutions. Specifically, there is a notable increase in error when employing Nx=200. However, the numerical solutions closely approximate the exact solution (depicted by the blue line) as Nx is increased to 1600.

    Figure 13.  Convergence of the numerical results to the exact solution at different values of the number of points.

    From Figure 14 and Table 1, the error of the numerical solution for u11(x,t) decreases as the value of Δx approaches zero, indicating a convergence towards zero error when we increase the number of points Nx.

    Figure 14.  Describing the L2 errors depicted in Table 1.
    Table 1.  The relative error with L2 norm at time t=5.
    Nx The relative error
    200 2.17×104
    400 3.81×105
    800 7.03×106
    1600 1.6×106
    3200 6.73×107
    6400 4.27×107
    12800 2.96×107
    20000 2.36×107

     | Show Table
    DownLoad: CSV

    This article has concentrated on developing the exact and numerical solutions of the Gilson-Pickering equation, utilizing the F-expansion and the (1/G)-expansion methods to extract the analytical solutions. Moreover, we used the moving mesh method, the PMA approach for the Gilson-Pickering equation, and the implicit finite difference approach on a uniform mesh with the examination of the stability for two implicit scheme: the fully implicit method and the Crank-Nicolson method. We provided that the numerical schemes are unconditionally stable. We have successfully computed both the analytical and numerical solutions. In this work, it has been demonstrated that the solutions generated using the (1/G)-expansion method are novel compared to the ones determined in [40]. Additionally, the solutions derived through the F-expansion method are demonstrated to be more inclusive than previously noted in the literature. For example, β can take any value and is not specified by a value. We utilized Mathematica 13.2 software to validate the correctness of our solutions by plugging them back into the original equation. One of the most excellent ways to ensure accurate solutions is to check the conformity of exact and numerical solutions. While some experts focus exclusively on obtaining exact solutions, this study employs a comparative approach, examining exact and numerical solutions to guarantee accuracy and correctness. The methodologies employed in this work can be applied to other NLPDEs encountered in the natural sciences.

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

    The authors extend their appreciation to the Deanship of Research and Innovation at the Ministry of Education in Saudi Arabia for funding this research through project number 445-9-746.

    The authors declare that they have no potential conflicts of interest in this article.

    [1] L. Berselli and G. Galdi, Regularity criteria involving the pressure for the weak solutions to the Navier-Stokes equations, Proc. Amer. Math. Soc., 130 (2002), 3585-3595.
    [2] S. Benbernou, M Terbeche, and M.A. Ragusa, A logarithmically improved regularity criterion for the MHD equations in terms of one directional derivative of the pressure, Applicable Analysis, http://dx.doi.org/10.1080/00036811.2016.1207246.
    [3] C. Cao and J. Wu, Two regularity criteria for the 3D MHD equations, J. Differential Equations, 248 (2010), 2263-2274.
    [4] Q. Chen, C. Miao, and Z. Zhang, On the regularity criterion of weak solution for the 3D viscous magneto-hydrodynamics equations, Comm. Math. Phys., 284 (2008), 919-930.
    [5] H. Duan, On regularity criteria in terms of pressure for the 3D viscous MHD equations, Appl.Anal., 91 (2012), 947-952.
    [6] G. Duvaut and J.-L. Lions, Inéquations en thermoélasticité et magnétohydrodynamique, Arch.Ration. Mech. Anal., 46 (1972), 241-279.
    [7] S. Gala, Extension criterion on regularity for weak solutions to the 3D MHD equations, Math.Meth. Appl. Sci., 33 (2010), 1496-1503.
    [8] C. He and Y. Wang, Remark on the regularity for weak solutions to the magnetohydrodynamic equations, Math. Methods Appl. Sci., 31 (2008), 1667-1684.
    [9] L. Ni, Z. Guo, and Y. Zhou, Some new regularity criteria for the 3D MHD equations, J. Math.Anal. Appl., 396 (2012), 108-118.
    [10] C. He and Z. Xin, On the regularity of weak solutions to the magnetohydrodynamic equations, J.Differential Equations, 213 (2005), 235-254.
    [11] E. Ji and J. Lee, Some regularity criteria for the 3D incompressible magnetohydrodynamics, J.Math. Anal. Appl., 369 (2010), 317-322.
    [12] X. Jia and Y. Zhou, Regularity criteria for the 3D MHD equations via partial derivatives, Kinet.Relat. Models, 5 (2012), 505-516.
    [13] X. Jia and Y. Zhou, Regularity criteria for the 3D MHD equations via partial derivatives, II. Kinet.Relat. Models, 7 (2014), no. 2, 291-304.
    [14] X. Jia and Y. Zhou, Ladyzhenskaya-Prodi-Serrin type regularity criteria for the 3D incompressible MHD equations in terms of 3 X 3 mixture matrices, Nonlinearity, 28 (2015), 3289-3307.
    [15] X. Jia and Y. Zhou, A new regularity criterion for the 3D incompressible MHD equations in terms of one component of the gradient of pressure, J. Math. Anal. Appl., 396 (2012), 345-350.
    [16] P.G. Lemarié-Rieusset, The Navier-Stokes equations in the critical Morrey-Campanato space, Rev.Mat. Iberoam., 23 (2007), no. 3, 897-930.
    [17] H. Lin and L. Du, Regularity criteria for incompressible magnetohydrodynamics equations in three dimensions, Nonlinearity, 26 (2013), 219-239.
    [18] S. Machihara and T. Ozawa, Interpolation inequalities in Besov spaces, Proc. Amer. Math. Soc., 131 (2003), 1553-1556.
    [19] M. Sermange and R. Temam, Some mathematical questions related to the MHD equations, Comm.Pure Appl. Math., 36 (1983), 635-664.
    [20] J.Wu, Viscous and inviscid magnetohydrodynamics equations, J. Anal. Math., 73 (1997), 251-265.
    [21] J. Wu, Bounds and new approaches for the 3D MHD equations, J. Nonlinear Sci., 12 (2002), 395-413.
    [22] J.Wu, Regularity results for weak solutions of the 3D MHD equations, Discrete Contin. Dyn. Syst., 10 (2004), 543-556.
    [23] K. Yamazaki, Remarks on the regularity criteria of generalized MHD and Navier-Stokes systems, J. Math. Phys., 54 (2013), 011502, 16pp.
    [24] Z. Zhang, P. Li, and G. Yu, Regularity criteria for the 3D MHD equations via one directional derivative of the pressure, J. Math. Anal. Appl., 401 (2013), 66-71.
    [25] Y. Zhou, Remarks on regularities for the 3D MHD equations, Discrete Contin. Dyn. Syst., 12(2005), 881-886.
    [26] Y. Zhou, Regularity criteria for the 3D MHD equations in terms of the pressure, Int. J. Non-Linear Mech., 41 (2006), 1174-1180.
    [27] Y. Zhou and S. Gala, Regularity criteria for the solutions to the 3D MHD equations in multiplier space, Z. Angrew. Math. Phys., 61 (2010), 193-199.
    [28] Y. Zhou and S. Gala, Regularity Criteria in Terms of the Pressure for the Navier-Stokes Equations in the Critical Morrey-Campanato Space, Z. Anal. Anwend., 30 (2011), 83-93.
  • Reader Comments
  • © 2017 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(4840) PDF downloads(1003) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog