Processing math: 100%
Research article

Stochastic solutions of the geophysical KdV equation: Numerical simulations and white noise impact

  • Received: 09 January 2025 Revised: 27 February 2025 Accepted: 04 March 2025 Published: 17 March 2025
  • MSC : 35Q53, 60H15, 37K40

  • This research explored stochastic soliton and periodic wave (SPW) solutions for the geophysical Korteweg-de Vries (KdV) equation with variable coefficients, incorporating the effects of Earth's rotation, fluid stratification, and topographical variations. The classical KdV equation, widely used to model nonlinear wave propagation, was extended to describe geophysical wave dynamics in atmospheric and oceanic systems. Exact solutions for both the deterministic and Wick-type stochastic (W-TS) forms of the geophysical KdV equation were obtained using white noise (WN) theory, the Hermite transform (HT), and the exp-function method. By employing the HT, the stochastic equation was transformed into a deterministic counterpart, facilitating the derivation of novel SPW solutions expressed as rational functions involving exponential terms. The inverse HT is then applied to retrieve stochastic SPW solutions under Gaussian WN conditions. Numerical analysis highlights the influence of Brownian motion (B-M) on the formation and behavior of SPWs in geophysical settings. Additionally, numerical simulations illustrate how random fluctuations affect wave stability and evolution, offering deeper insights into nonlinear wave interactions in oceanic and atmospheric environments.

    Citation: Areej A. Almoneef, Abd-Allah Hyder, Mohamed A. Barakat, Abdelrheem M. Aly. Stochastic solutions of the geophysical KdV equation: Numerical simulations and white noise impact[J]. AIMS Mathematics, 2025, 10(3): 5859-5879. doi: 10.3934/math.2025269

    Related Papers:

    [1] Nafissa T. Trouba, Huiying Xu, Mohamed E. M. Alngar, Reham M. A. Shohib, Haitham A. Mahmoud, Xinzhong Zhu . Soliton solutions and stability analysis of the stochastic nonlinear reaction-diffusion equation with multiplicative white noise in soliton dynamics and optical physics. AIMS Mathematics, 2025, 10(1): 1859-1881. doi: 10.3934/math.2025086
    [2] Zeliha Korpinar, Mustafa Inc, Ali S. Alshomrani, Dumitru Baleanu . The deterministic and stochastic solutions of the Schrodinger equation with time conformable derivative in birefrigent fibers. AIMS Mathematics, 2020, 5(3): 2326-2345. doi: 10.3934/math.2020154
    [3] Hamood Ur Rehman, Aziz Ullah Awan, Sayed M. Eldin, Ifrah Iqbal . Study of optical stochastic solitons of Biswas-Arshed equation with multiplicative noise. AIMS Mathematics, 2023, 8(9): 21606-21621. doi: 10.3934/math.20231101
    [4] Dan Chen, Da Shi, Feng Chen . Qualitative analysis and new traveling wave solutions for the stochastic Biswas-Milovic equation. AIMS Mathematics, 2025, 10(2): 4092-4119. doi: 10.3934/math.2025190
    [5] Ahmed M. Elsherbeny, Taher A. Nofal, Yakup Yıldırım, Ahmed H. Arnous . New solitary waveforms and their dynamics in the stochastic generalized Chen–Lee–Liu model. AIMS Mathematics, 2025, 10(3): 5197-5235. doi: 10.3934/math.2025239
    [6] Da Shi, Zhao Li, Dan Chen . New traveling wave solutions, phase portrait and chaotic patterns for the dispersive concatenation model with spatio-temporal dispersion having multiplicative white noise. AIMS Mathematics, 2024, 9(9): 25732-25751. doi: 10.3934/math.20241257
    [7] Shami A. M. Alsallami . Investigating exact solutions for the (3+1)-dimensional KdV-CBS equation: A non-traveling wave approach. AIMS Mathematics, 2025, 10(3): 6853-6872. doi: 10.3934/math.2025314
    [8] F. A. Mohammed . Soliton solutions for some nonlinear models in mathematical physics via conservation laws. AIMS Mathematics, 2022, 7(8): 15075-15093. doi: 10.3934/math.2022826
    [9] Yazid Alhojilan, Islam Samir . Investigating stochastic solutions for fourth order dispersive NLSE with quantic nonlinearity. AIMS Mathematics, 2023, 8(7): 15201-15213. doi: 10.3934/math.2023776
    [10] Shabir Ahmad, Saud Fahad Aldosary, Meraj Ali Khan . Stochastic solitons of a short-wave intermediate dispersive variable (SIdV) equation. AIMS Mathematics, 2024, 9(5): 10717-10733. doi: 10.3934/math.2024523
  • This research explored stochastic soliton and periodic wave (SPW) solutions for the geophysical Korteweg-de Vries (KdV) equation with variable coefficients, incorporating the effects of Earth's rotation, fluid stratification, and topographical variations. The classical KdV equation, widely used to model nonlinear wave propagation, was extended to describe geophysical wave dynamics in atmospheric and oceanic systems. Exact solutions for both the deterministic and Wick-type stochastic (W-TS) forms of the geophysical KdV equation were obtained using white noise (WN) theory, the Hermite transform (HT), and the exp-function method. By employing the HT, the stochastic equation was transformed into a deterministic counterpart, facilitating the derivation of novel SPW solutions expressed as rational functions involving exponential terms. The inverse HT is then applied to retrieve stochastic SPW solutions under Gaussian WN conditions. Numerical analysis highlights the influence of Brownian motion (B-M) on the formation and behavior of SPWs in geophysical settings. Additionally, numerical simulations illustrate how random fluctuations affect wave stability and evolution, offering deeper insights into nonlinear wave interactions in oceanic and atmospheric environments.



    Nonlinear partial differential equations (PDEs) arise from the mathematical modeling of complex physical systems. Investigating these nonlinear physical models through the analysis of their wave solutions plays a crucial role in applied sciences. The Korteweg-de Vries (KdV) equations, in particular, serve as models for nonlinear wave propagation across a wide range of material science applications. Initially developed to describe long-wavelength, small-amplitude shallow water waves, the KdV equation has proven useful in numerous physical scenarios. Its soliton and periodic exact solutions are relevant to phenomena such as collisionless hydromagnetic waves, internal stratified waves, particle acoustic waves, and plasma dynamics. For more details, see references [1,2,3,4,5]. In geophysical fluid dynamics, the KdV equation is extended to account for more complex phenomena relevant to large-scale waves in the atmosphere and ocean. These geophysical KdV equations incorporate the effects of Coriolis forces (due to Earth's rotation), stratification (layering of fluids with different densities), and variable topography (changing ocean depth), making them crucial for understanding wave propagation in natural environments [6,7,8,9]. For example, the KdV equation is used to model shallow water waves and internal waves in the ocean, as well as atmospheric gravity waves. Internal waves, which occur in stratified fluids like the ocean, propagate along interfaces between layers of different densities and are essential for mixing and energy transfer in the ocean [10,11]. In [12], internal solitons have been observed and simulated, particularly in areas like the South China Sea, confirming the accuracy of KdV-based models in practical environments. The rotational KdV equation incorporates Earth's rotational effects and has been employed to examine equatorial waves and other large-scale wave patterns [13]. Different forms of the KdV equation, including the variable-coefficient KdV and the KdV-Burgers equation (which accounts for dissipative processes), are utilized to represent wave behavior over varying depths and under conditions involving damping, see [14,15]. The geophysical KdV equations provide a robust framework for understanding nonlinear wave dynamics in both oceanographic and atmospheric contexts. These equations offer insight into a wide range of phenomena, from coastal wave propagation to the behavior of atmospheric disturbances, making them a key tool in geophysical fluid dynamics, see [16,17,18].

    This work is devoted to investigating the stochastic geophysical KdV equation:

    νZ(μ,ν)+Ψ(ν)μZ(μ,ν)+Φ(ν)Z(μ,ν)μZ(μ,ν)+Ξ(ν)3μ3Z(μ,ν)=0, (1.1)

    where (μ,ν)R×R+ and Ψ, Φ, and Ξ are non-zero integrable functions from R+ to the Kondrative distribution space (T)1 which was defined by Holden et al. in [20] as a Banach algebra with the Wick-product "". Equation (1.1) is the perturbation of the variable coefficients geophysical KdV equation:

    νz(μ,ν)+ψ(ν)μz(μ,ν)+ϕ(ν)z(μ,ν)μz(μ,ν)+ξ(ν)3μ3z(μ,ν)=0. (1.2)

    The geophysical KdV equation (1.2) represents an extension of the classical KdV equation, adjusted to describe wave phenomena in geophysical contexts, such as the oceans and atmosphere. This formulation takes into account the Coriolis effect, which models the deflection due to the Earth's rotation [19]. In Eq (1.2), the term ψ represents the Coriolis effect, which is responsible for the deflection of objects in motion, such as air or water, caused by the Earth's rotation. The coefficient ϕ captures the nonlinearity, while ξ characterizes the dispersion, illustrating how a complex wave breaks down into its component wavelengths. Additionally, if Eq (1.2) is studied in a stochastic framework, it leads to the stochastic version of the geophysical KdV equation. In this case, the functions Ψ, Φ, and Ξ serve as the stochastic counterparts of ψ, ϕ, and ξ, respectively, thereby accounting for random perturbations within the geophysical wave system. To obtain precise solutions to this stochastic geophysical KdV equation, we confine our analysis to a WN setting, which leads us to the W-TS geophysical KdV equation (1.1). Recent studies have integrated both numerical and analytical approaches to further understand the dynamics of the KdV equation.

    In recent years, a significant number of studies have focused on partial differential equations (PDEs) and the exact solutions of such equations using various methodologies. In [21], the first integral method was applied to determine exact solutions for nonlinear PDEs involving the beta-derivative. Moreover, solutions in the form of optical, dark, complex, and singular solitons have been found for certain nonlinear PDEs using the M-derivative approach [22,23]. The examination of exact and approximate solutions for nonlinear evolution equations is essential for comprehending nonlinear physical phenomena. Several efficacious methodologies have been suggested, including the bilinear transformation method [24], the modified Clarkson and Kruskal (CK) direct method [25], the multiscale expansion approach [26], the binary Bell polynomials technique [27], the Riemann-Hilbert method [28], and the approximate symmetry method [29]. The exp-function method, presented by He and Wu [30], offers a direct approach for deriving generalized soliton, periodic, and compacton-like solutions for several nonlinear partial differential equations. This approach expresses solutions as rational functions of the exponential form, with both the numerator and denominator being polynomials of exponential functions. The primary benefit of the exp-function method is its capacity to produce a diverse array of exact solutions for nonlinear partial differential equations by varying parameters and polynomial degrees in the solutions. A fundamental characteristic of solutions derived from this method is their reducibility [31]. Furthermore, the exp-function method produces both generalized soliton and periodic solutions. The procedure is streamlined through the utilization of Mathematica and can be efficiently implemented for many nonlinear equations. Recent studies have further explored the KdV equation using the exp-function method, providing new perspectives on soliton and periodic wave solutions. For instance, recent work [32] has investigated these techniques in depth, which reinforces the relevance of our approach in addressing stochastic influences in geophysical settings. Modern studies have examined the strain wave equation in micro-structured solids, deriving various analytical solutions through advanced mathematical methods [33].

    While existing studies, such as Zhang [34], have investigated exact solutions for Wick-type stochastic KdV equations, this paper presents several key advancements. We derive new stochastic SPW solutions for the variable coefficient geophysical KdV equation and the W-TS geophysical KdV equation. Through the use of WN theory and the HT, the W-TS geophysical KdV equation is converted into its deterministic counterpart. By applying the exp-function method, we construct exact solutions for the deterministic geophysical KdV equation, represented as an exponential-type rational function, with polynomials of exponential functions in both the numerator and denominator. The highest-order linear term and the highest-order nonlinear term of the deterministic geophysical KdV equation are balanced to determine the polynomial degrees in both parts. With symbolic computation in Mathematica, SPW solutions for the variable coefficient geophysical KdV equation are found. Under certain conditions, we apply the inverse HT to extract stochastic SPW solutions for the W-TS geophysical KdV equation. Additionally, an example with numerical simulations is presented to analyze the impact of Gaussian WN on SPW dynamics, providing deeper insight into how random perturbations influence wave behavior.

    The structure of the paper is as follows: Section 2 reviews key definitions and properties relevant to Gaussian WN analysis and outlines the primary steps in solving nonlinear PDEs. In Section 3, we use the exp-function method, WN theory, and the HT to derive deterministic solutions for the W-TS geophysical KdV equation. Section 4 applies the inverse HT to obtain stochastic SPW solutions for the W-TS geophysical KdV equation. In Section 5, we provide an example with numerical simulations, demonstrating that stochastic solutions can be expressed as B-M functionals. Finally, Section 6 concludes with a summary of the results and findings.

    The foundational study of WN structure begins with the inclusion chain T(Rn)L2(Rn)T(Rn). Here, T(Rn) denotes the space of test functions, which are infinitely differentiable and rapidly vanishing on Rn, while T(Rn) refers to the space of tempered distributions. By the Bochner-Minlos theorem [20], there exists a unique WN measure ν on (T(Rn),θ(T(Rn))).

    Consider the Hermite functions ζm(x)=π1/4((m1)!)1/2ex2/2pm1(2x) for mN, where pm(x) are Hermite polynomials. It is well known that the family {ζm}mN forms an orthonormal basis for L2(R). Let β=(β1,,βn) represent a multi-index of dimension n, where β1,,βnN. The tensor product family ζβ:=ζ(β1,,βn)=ζβ1ζβn, βNn, serves as an orthonormal basis for L2(Rn). To further proceed, we establish an ordering on Nn as follows:

    r<snj=1β(r)jnj=1β(s)j,whereβ(r)=(β(r)j)nj=1, β(s)=(β(s)j)nj=1Nn.

    Using this ordering, define κr:=ζβ(r)=ζβ(r)1ζβ(r)n, rN. Let K=(NN0)c denote the set of sequences β=(βi)iN where βiN0 and has compact support. For βK, define

    Hβ(ω)=i=1pβi(ω,κi),ωT(Rn).

    Let mN, the Kondratiev space of stochastic test functions (T)m1 is defined as:

    (T)m1=(f=βdβHβmj=1L2(ν): dβRm and f21,j:=βd2β(β!)2(2N)jβ<, jN),

    and the Kondratiev space of stochastic distributions (T)m1 is defined as:

    (T)m1=(F=βeβHβ: eβRm and F21,j:=βe2β(2N)pβ< for some pN).

    The family of seminorms f1,j, jN, induces a topology on (T)m1, and (T)m1 is the dual space of (T)m1 with the pairing

    F,f=β(eβ,dβ)β!,

    where F=βeβHβ(T)m1, f=βdβHβ(T)m1, and (eβ,dβ) represents the usual scalar product in Rm.

    The Wick product of two distributions F=βuβHβ, G=γvγHγ(T)m1 with uβ,vγRm is defined by:

    FG=β,γ(uβ,vγ)Hβ+γ.

    Let F=βuβHβ(T)m1 with uβRm. The HT of F is defined as:

    GF(w)=˜F(w)=βuβwβCm(when convergent),

    where w=(w1,w2,)CN and wβ=i=1wβii, with β=(β1,β2,)K and w0i=1.

    For F,G(T)m1, the HT gives:

    ~FG(w)=˜F(w)˜G(w),

    for all w where ˜F(w) and ˜G(w) are defined. The multiplication on the right-hand side corresponds to the standard bilinear multiplication in Cm, defined by (w11,,w1m)(w21,,w2m)=mi=1w1iw2i, where wkiC. Consequently, the HT translates the Wick product into regular multiplication, and convergence in (T)m1 is equivalent to pointwise convergence within a neighborhood of zero in Cm. For additional details on Kondratiev spaces, the Wick product, and the HT, please refer to [20].

    One of the most important tools for obtaining our conclusions is the following theorem.

    Theorem 2.1. [20] Let z(μ,ν,w) be a solution to the following equation:

    ˜Ω(μ,ν,ν,μ1,,μn,z,w)=0, (2.1)

    for (μ,ν) in some bounded open region ORn×R+, and for all wNα(u), for some α<,u>0. Further assume that z(μ,ν,w), along with all of its derivatives involved in (2.1), are uniformly bounded for (μ,ν,w)O×Nα(u), continuous with respect to (μ,ν)O for each wNα(u), and analytic in wNα(u) for all (μ,ν)O. Then there exists a stochastic distribution Z(μ,ν)(T)1 such that z(μ,ν,w)=˜Z(μ,ν)(w) for all (μ,ν,w)O×Nα(u), and Z(μ,ν) satisfies the equation:

    Ω(μ,ν,ν,μ1,,μn,Z)=0, (2.2)

    in the strong sense in (T)1.

    This section utilizes the exp-function method [30] alongside WN theory to derive a family of exact stochastic solutions for the (W-TS) geophysical KdV equation (1.1). By applying the HT to Eq (1.1), we obtain the corresponding deterministic equation:

    ~Zν(μ,ν,w)+˜Ψ(ν,w)~Zμ(μ,ν,w)+˜Φ(ν,w)˜Z(μ,ν,w)~Zμ(μ,ν,w)+˜Ξ(ν,w)~Zμμμ(μ,ν,w)=0, (3.1)

    where w=(w1,w2,...)(CN)c. To find traveling wave solutions for Eq (3.1), we introduce the following transformations:

    {˜Ψ(ν,w):=ψ(ν,w),˜Φ(ν,w):=ϕ(ν,w),˜Ξ(ν,w):=ξ(ν,w),˜Z(μ,ν,w):=z(μ,ν,w)=z(χ(μ,ν,w)), (3.2)

    where

    χ(μ,ν,w)=ςμ+λν0ϑ(τ,w)dτ, (3.3)

    with ς and λ as arbitrary constants, and ϑ as a non-zero function to be determined. This transforms Eq (3.1) into the following NODE:

    (λϑ+ςψ+ςϕz)dzdχ+ς3ξd3zdχ3=0. (3.4)

    Using the exp-function method [30], the general solution of Eq (3.4) can be expressed as an expansion in terms of exponential functions:

    z(χ)=hn=gηn(ν,w)exp(nχ)jm=iρm(ν,w)exp(mχ), (3.5)

    where g,h,i, and j are freely chosen positive integers, and ηn and ρm are functions to be determined. Equation (3.5) may also be written in the form:

    z(χ)=ηgexp(gχ)++ηhexp(hχ)ρiexp(iχ)++ρjexp(jχ). (3.6)

    To determine appropriate values for g and i, we balance the highest-order linear term in Eq (3.4) with the highest-order nonlinear term. A straightforward computation gives:

    d3zdχ3=ϵ1exp[(7i+g)χ]+ϵ2exp[8iχ]+, (3.7)

    and

    zdzdχ=ϵ3exp[(i+2g)χ]+ϵ4exp[3iχ]+=ϵ3exp[(6i+2g)χ]+ϵ4exp[8iχ]+, (3.8)

    where ϵi are coefficients assigned solely for simplicity. By equating the highest-order term of the exponential function in Eqs (3.7) and (3.8), we obtain 7i+g=6i+2g, which simplifies to i=g. Similarly, to find the values of h and j, we balance the lowest-order linear term in Eq (3.4):

    d3zdχ3=+υ1exp[(7j+h)χ]+υ2exp[8jχ], (3.9)

    with the term

    zd3zdχ3=+υ3exp[(j+2h)χ]+υ4exp[3jχ]=+υ3exp[(6j+2h)χ]+υ4exp[8jχ], (3.10)

    where υi are coefficients chosen only for simplicity. By equating the lowest-order term of the exponential function in Eqs (3.9) and (3.10), we obtain (7j+h)=(6j+2h), which gives j=h.

    Case 1. Choosing i=g=1 and h=j=1, Eq (3.4) takes the form:

    z(χ)=η1exp(χ)+η0+η1exp(χ)ρ1exp(χ)+ρ0+ρ1exp(χ). (3.11)

    Substituting Eq (3.11) into Eq (3.4), and setting the coefficients of exp(χ) to zero, results in a system of algebraic equations in η0, η1, η1, ρ0, ρ1, ρ1, and ϑ. Solving this system yields:

    {η1=ρ0(6ς2ξρ0η0ϕ)4ρ1ϕ,   η1=ρ1(6ς2ξρ0η0ϕ)ρ0ϕ,ρ1=ρ204ρ1,   ϑ=ς(5ς3ξρ0η0ϕρ0ψ)λρ0, (3.12)

    where η0, ρ0, and ρ1 are free functions in (ν,w)R+×(CN)c.

    By substituting the solutions from Eq (3.12) into Eq (3.11) and applying Eq (3.3), a soliton wave solution for Eq (3.1) can be obtained:

    z1(μ,ν,w)=[(η0(ν,w)ϕ(ν,w)6ς2ξ(ν,w)ρ0(ν,w))(ρ20(ν,w)A11(μ,ν,w)+4ρ1(ν,w)A12(μ,ν,w))+4ρ0(ν,w)ρ1(ν,w)η0(ν,w)ϕ(ν,w)]×[ρ0(ν,w)ϕ(ν,w)(ρ20(ν,w)A11(μ,ν,w)+4ρ0(ν,w)ρ1(ν,w)+4A12(μ,ν,w))]1, (3.13)

    where

    A11(μ,ν,w)=exp(ςμ+ςν05ς2ξ(τ,w)ρ0(τ,w)η0(τ,w)ϕ(τ,w)ρ0(τ,w)ψ(τ,w)ρ0(τ,w)dτ), (3.14)

    and

    A12(μ,ν,w)=exp([ςμ+ςν05ς2ξ(τ,w)ρ0(τ,w)η0(τ,w)ϕ(τ,w)ρ0(τ,w)ψ(τ,w)ρ0(τ,w)dτ]). (3.15)

    Case 2. If we consider i=g=2 and j=h=2, the proposed solution in Eq (3.6) becomes:

    z(χ)=η2exp(2χ)+η1exp(χ)+η0+η1exp(χ)+η2exp(2χ)ρ2exp(2χ)+ρ1exp(χ)+ρ0+ρ1exp(χ)+ρ2exp(2χ). (3.16)

    To simplify the calculations, we set ρ1=ρ1=0 and ρ2=0. As a result, Eq (3.16) takes the form:

    z(χ)=η2exp(2χ)+η1exp(χ)+η0+η1exp(χ)+η2exp(2χ)exp(2χ)+ρ0+ρ2exp(2χ). (3.17)

    Upon inserting Eq (3.17) into Eq (3.4), collecting the coefficients of exp(χ), and equating them to zero, a system of algebraic equations is formed with variables η0,η1,η2,η1,η2,ρ0,ρ2, and ϑ. By solving this system with Mathematica, we find the following solutions:

    {η0=ρ0ϕ(12ς2ξ+η2ϕ),  η1=6m ς2ϕ(ξ2ρ0),  η1=3m ς2ϕ(ξ2ρ30),η2=14η22ρ20,  ρ2=14ρ20,  ϑ=ςλ(ς2ξ+η2ϕ+ψ), (3.18)

    where ρ0 and η2 are arbitrary functions on R+×(CN)c and m2=1.

    Substituting the expressions from (3.18) into Eq (3.17) and applying Eq (3.3) leads to a soliton wave solution for Eq (3.1), presented below.

    z2(μ,ν,w)=[ϕ(ν,w)(4A23(μ,ν,w)+η22(ν,w)ρ20(ν,w)A24(μ,ν,w))+4ρ0(ν,w)(12ς2ξ(ν,w)+η2(ν,w)ϕ(ν,w))+12m ς2ξ(ν,w)2ρ0(ν,w)(ρ0(ν,w)A22(μ,ν,w)A21(μ,ν,w))]×[4ϕ(ν,w)(A23(μ,ν,w)+ρ0(ν,w)+14ρ20(ν,w)A24(μ,ν,w))]1, (3.19)

    where

    A21(μ,ν,w)=exp(ςμςν0(ς2ξ(τ,w)+η2(τ,w)ϕ(τ,w)+ψ(τ,w))dτ), (3.20)
    A22(μ,ν,w)=exp([ςμςν0(ς2ξ(τ,w)+η2(τ,w)ϕ(τ,w)+ψ(τ,w))dτ]), (3.21)
    A23(μ,ν,w)=exp(2[ςμςν0(ς2ξ(τ,w)+η2(τ,w)ϕ(τ,w)+ψ(τ,w))dτ]), (3.22)

    and

    A24(μ,ν,w)=exp(2[ςμςν0(ς2ξ(τ,w)+η2(τ,w)ϕ(τ,w)+ψ(τ,w))dτ]). (3.23)

    Case 3. If we set i=g=2 and j=h=1, then Eq (3.6) becomes:

    z(χ)=η2exp(2χ)+η1exp(χ)+η0+η1exp(χ)ρ2exp(2χ)+ρ1exp(χ)+ρ0+ρ1exp(χ). (3.24)

    To simplify the calculations, we set ρ2=1. Using the procedures described above, we get:

    {η0=η2ρ213,    η1=η2ρ1,    η1=η2ρ3127,ρ0=ρ213,    ρ1=ρ3127,    ϑ=ξς3η2ςϕςψλ, (3.25)

    where ρ1 and η2 are arbitrary functions on R+×(CN)c.

    Substituting the expressions from (3.21) into Eq (3.20) and applying Eq (3.3) leads to a soliton wave solution for Eq (3.1), presented below.

    z3(μ,ν,w)=[27η2(ν,w)A23(μ,ν,w)+27η2(ν,w)ρ1(ν,w)A21(μ,ν,w)+9η2(ν,w)ρ21(ν,w)+η2(ν,w)ρ31(ν,w)A22(μ,ν,w)]×[27A32(μ,ν,w)+27ρ1(ν,w)A21(μ,ν,w)+9ρ21(ν,w)+ρ31(ν,w)A22(μ,ν,w)]1, (3.26)

    where A21,A22, and A23 are defined in Eqs (3.20), (3.21), and (3.22), respectively.

    Let ς be defined as an imaginary number. Under this assumption, the previously obtained soliton wave solutions can be converted into periodic forms. The corresponding transformations are:

    {ς=mB; BR and m2=1,e±mχ=cos(χ)±msin(χ). (3.27)

    Consequently, the soliton solution presented in (3.13) is modified to:

    z4(μ,ν,w)=((η0(ν,w)ϕ(ν,w)+6B2ξ(ν,w)ρ0(ν,w))[(ρ20(ν,w)+4ρ1(ν,w))cos(BS(μ,ν,τ))+m(ρ20(ν,w)4ρ1(ν,w))sin(BS(μ,ν,τ))]+4ρ0(ν,w)ρ1(ν,w)η0(ν,w)ϕ(ν,w))×(ρ0(ν,w)ϕ(ν,w)[(ρ20(ν,w)+4ρ0(ν,w)ρ1(ν,w))cos(BS(μ,ν,τ))+m(ρ20(ν,w)4)sin(BS(μ,ν,τ))])1, (3.28)

    where

    S(μ,ν,w)=μ+ν05B2ξ(ν,w)ρ0(ν,w)η0(ν,w)ϕ(ν,w)ρ0(ν,w)ψ(ν,w)ρ0(ν,w)dτ. (3.29)

    To achieve real periodic solutions, the imaginary component in Eq (3.28) must equal zero. To accomplish this, we assign the values ρ0(ν,w)=±2 and ρ1(ν,w)=1. Consequently, we derive the following periodic solutions for Eq (3.1):

    z41(μ,ν,w)=η0(ν,w)ϕ(ν,w)+12B2ξ(ν,w)3ϕ(ν,w)+η03 sec(B[μ+12ν0(10B2ξ(τ,w)η0(τ,w)ϕ(τ,w)2ψ(τ,w))dτ]), (3.30)

    and

    z42(μ,ν,w)=η0(ν,w)ϕ(ν,w)12B2ξ(ν,w)ϕ(ν,w)η0 sec(B[μ12ν0(10B2ξ(τ,w)η0(τ,w)ϕ(τ,w)+2ψ(τ,w))dτ]), (3.31)

    Using the same approach as for the solution in Eq (3.13), the soliton wave solutions presented in (3.19) and (3.26) can be transformed into periodic wave solutions. Different parameter choices for i, g, h, and j enable the derivation of a variety of SPW solutions for Eq (3.1).

    It is noteworthy that the construction of these soliton solutions imposes inherent relationships among the variable coefficients. These conditions ensure a balance between the linear and nonlinear terms, which is crucial for the existence of soliton structures. Similar intrinsic constraints have been reported in earlier works [35]. Although we do not impose explicit constraints beforehand, our method naturally satisfies these requirements.

    The properties of exponential and trigonometric functions imply the existence of a bounded, open domain OR×R+, where the solution z(μ,ν,w) to Eq (3.1) and all derivatives in Eq (3.1) are uniformly continuous for (μ,ν,w)O×Nα(u), with α< and u>0. This solution remains continuous at each (μ,ν)O for any wNα(u) and is analytic in wNα(u) for fixed (μ,ν)O. As stated in Theorem 2.1, there exists a stochastic distribution Z(μ,ν)(T)1, satisfying z(μ,ν,w)=˜Z(μ,ν)(w) for all (μ,ν,w)O×Nα(u), and Z(μ,ν) solves Eq (1.1) in (T)1. Consequently, applying the inverse HT to Eqs (3.13), (3.19), and (3.26) provides the exact soliton wave solutions for Eq (1.1) as follows:

    Z1(μ,ν)=[(η0(ν)Φ(ν)6ς2Ξ(ν)ρ0(ν))((ρ0)2(ν)D11(μ,ν)+4ρ1(ν)D12(μ,ν))+4ρ0(ν)ρ1(ν)η0(ν)Φ(ν)][ρ0(ν)Φ(ν)((ρ0)2(ν)D11(μ,ν)+4ρ0(ν)ρ1(ν)+4D12(μ,ν))](1), (4.1)

    where

    D11(μ,ν)=exp(ςμ+ςν05ς2Ξ(τ)ρ0(τ)η0(τ)Φ(τ)ρ0(τ)Ψ(τ)ρ0(τ)dτ), (4.2)
    D12(μ,ν)=exp([ςμ+ςν05ς2Ξ(τ)ρ0(τ)η0(τ)Φ(τ)ρ0(τ)Ψ(τ)ρ0(τ)dτ]), (4.3)

    and

    Z2(μ,ν)=[Φ(ν)(4D23(μ,ν)+(η2)2(ν)(ρ0)2(ν)D24(μ,ν))+4ρ0(ν)(12ς2Ξ(ν)+η2(ν)Φ(ν))+12m ς2Ξ(ν)2ρ0(ν)(ρ0(ν)D22(μ,ν)D21(μ,ν))][4Φ(ν)(D23(μ,ν)+ρ0(ν)+14(ρ0)2(ν)D24(μ,ν))](1), (4.4)

    where

    D21(μ,ν)=exp(ςμςν0(ς2Ξ(τ)+η2(τ)Φ(τ)+Ψ(τ))dτ), (4.5)
    D22(μ,ν)=exp([ςμςν0(ς2Ξ(τ)+η2(τ)Φ(τ)+Ψ(τ))dτ]), (4.6)
    D23(μ,ν)=exp(2[ςμςν0(ς2Ξ(τ)+η2(τ)Φ(τ)+Ψ(τ))dτ]), (4.7)

    and

    D24(μ,ν)=exp(2[ςμςν0(ς2Ξ(τ)+η2(τ)Φ(τ)+Ψ(τ))dτ]), (4.8)
    Z3(μ,ν)=[27η2(ν)D23(μ,ν)+27η2(ν)ρ1(ν)D21(μ,ν)+9η2(ν)(ρ1)2(ν)+η2(ν)(ρ1)3(ν)D22(μ,ν)][27D23(μ,ν)+27ρ1(ν)D21(μ,ν)+9(ρ1)2(ν)+(ρ1)3(ν)D22(μ,ν)](1), (4.9)

    where D21,D22, and D23 are defined in Eqs (4.5), (4.6), and (4.7), respectively.

    Also, by performing the inverse HT on Eqs (3.30) and (3.31), the exact periodic wave solutions for Eq (1.1) are obtained as follows:

    Z41(μ,ν)=η0(ν)Φ(ν)+12B2Ξ(ν)3Φ(ν)+η03 sec(B[μ+12ν0(10B2Ξ(τ)η0(τ)Φ(τ)2Ψ(τ))dτ]), (4.10)

    and

    Z42(μ,ν)=η0(ν)Φ(ν)12B2Ξ(ν)Φ(ν)η0 sec(B[μ12ν0(10B2Ξ(τ)η0(τ)Φ(τ)+2Ψ(τ))dτ]). (4.11)

    Here, ρ0,ρ1,ρ1,η0, and η2 are arbitrary stochastic distributions in (T)1.

    Moreover, the stochastic solutions derived in this work offer significant advantages in modeling real geophysical phenomena. For example, the inclusion of random perturbations allows the model to capture the complex variability observed in internal ocean waves and atmospheric gravity waves. This improved framework can more accurately represent the influence of environmental fluctuations on wave amplitude, energy distribution, and propagation dynamics, thereby enhancing its practical applicability in geophysical studies.

    This section presents a specific example accompanied by numerical simulations to validate the significance of our findings and illustrate their practical applications. The solutions obtained for Eq (1.1) are significantly influenced by the configuration of the functions Φ(ν), Ψ(ν), and Ξ(ν). Consequently, modifying the forms of these functions produces several solutions to Eq (1.1), as demonstrated in Eqs (4.1)–(4.11). This example demonstrates this behavior.

    Example 5.1 Consider the following setup: Let Ψ(ν)=δ1Φ(ν), Ξ(ν)=δ2Φ(ν), Φ(ν)=ϖ(ν)+δ3W(ν), ρ0(ν)=δ4, ρ1=δ5, ρ1=δ6, η0=δ7, and η2=δ8, where the constants δk(k=1,2,,8) are arbitrary, ϖ(ν) is a measurable function that is bounded on R+, and W(ν) represents Gaussian WN, defined as the time derivative (in the strong sense within (T)1) of B-M B(ν). The HT of W(ν) is given by ˜W(ν,w)=r=1wrν0κr(τ)dτ [20]. Using the form of ˜W(ν,w), Eqs (4.1)–(4.11) yield the WN functional solution for Eq (1.1) as follows:

    ZB1(μ,ν)=(δ76ς2δ2)(δ24exp(¯D1(μ,ν))+4δ5exp(¯D1(μ,ν))+4δ4δ5δ7δ34exp(¯D1(μ,ν))+4δ4δ5+4exp(¯D1(μ,ν)), (5.1)

    where

    ¯D1(μ,ν)=ς[μ+5ςδ4δ2δ7δ4δ1δ4[ν0ϖ(τ)dτ+δ3(B(ν)ν22)]], (5.2)
    ZB2(μ,ν)=14(exp(2¯D2(μ,ν))+δ4+14δ24exp(2¯D1(μ,ν)))[4exp(2¯D2(μ,ν))+δ28δ24exp(2¯D2(μ,ν))+4δ4(12ς2δ2+δ8)+12mς2δ22δ4(δ4exp(¯D2(μ,ν))exp(¯D2(μ,ν)))], (5.3)

    where

    ¯D2(μ,ν)=ς[μ(ς2δ2+δ8+δ1)[ν0ϖ(τ)dτ+δ3(B(ν)ν22)]], (5.4)
    ZB3(μ,ν)=9δ8(3exp(2¯D2(μ,ν))+3δ6exp(¯D2(μ,ν))+δ26)+δ8δ36exp(¯D2(μ,ν))27exp(2¯D2(μ,ν))+9δ6(3exp(¯D2(μ,ν))+δ6)+δ36exp(¯D2(μ,ν)), (5.5)

    where ¯D2 is defined in Eq (5.4),

    ZB4=4B2δ2+δ73{1+sec(B(μ+10B2δ2δ72δ12[ν0ϖ(τ)dτ+δ3(B(ν)ν22)]))}, (5.6)

    and

    ZB5=12B2δ2+δ7{1sec(B(μ+10B2δ2+δ72δ12[ν0ϖ(τ)dτ+δ3(B(ν)ν22)]))}. (5.7)

    The numerical simulation of the soliton solution in (5.1) is shown in Figures 1 and 2, when δ1=1, δ2=2, δ3=0.5, δ4=3, δ5=1.5, δ7=1.2, ς=0.7, and ϖ(ν)=tanh(2ν). Figure 1 illustrates the behavior of the soliton solution (5.1) with stochastic influence, where the B-M B(ν) is simulated as random noise with increments dB=ν(2)ν(1). RAND[0,1]. Figure 2 represents the evolution of the the soliton solution (5.1) without stochastic influence, where the B-M B(ν)=0. From Figures 1 and 2, it can be observed that the absence of B-M results in smooth, deterministic wave behavior. The inclusion of noise introduces stochastic fluctuations that influence the overall amplitude and structure of the soliton solution. Additionally, the stochastic forcing term in Figure 1 causes uncertainty in the wave amplitudes and disrupts the traveling wave patterns, leading to random variations in the soliton's behavior.

    Figure 1.  Visualization of the soliton solution in (5.1) including the effects of B-M: (a) 3D plot, (b) contour plot, and (c) 2D plot.
    Figure 2.  Visualization of the soliton solution in (5.1) excluding the effects of B-M: (a) 3D plot, (b) contour plot, and (c) 2D plot.

    The numerical simulation of the periodic solution in (5.6) is shown in Figures 3 and 4, when δ1=0.8, δ2=2, δ3=0.4, δ7=2, and B=0.001. The function ϖ(ν) is modified to exhibit stronger periodic behavior, defined as ϖ(ν)=sin(2πν/max(ν))+cos(3πν/max(ν)), contributing to the oscillations in the periodic solution (5.6). Figure 3 illustrates the periodic solution (5.6) with stochastic influence, where the B-M B(ν) is simulated as amplified random noise with increments dB=ν(2)ν(1)×20RAND[0,1]. Figure 4 shows the behavior of the periodic solution (5.6) without stochastic influence, where B(ν)=0. From Figures 3 and 4, we observe that the inclusion of B-M introduces significant stochastic fluctuations in the periodic solution. These fluctuations affect the overall amplitude and structure, leading to random variations in the system's behavior. The periodic function ϖ(ν) further amplifies the oscillations, creating a more complex wave pattern. In contrast, the absence of stochastic influence in Figure 4 results in smooth, deterministic wave behavior. The stochastic forcing term in Figure 3 creates uncertainty in the wave amplitudes, disrupting the periodic traveling wave patterns and leading to random distortions in the periodicity behavior.

    Figure 3.  Visualization of the periodic solution in (5.6) with the effects of B-M: (a) 3D plot, (b) contour plot, and (c) 2D plot.
    Figure 4.  Visualization of the periodic solution in (5.6) without the effects of B-M: (a) 3D plot, (b) contour plot, and (c) 2D plot.

    This research has derived precise SPW solutions for the stochastic geophysical KdV equation with variable coefficients, highlighting its importance for modeling wave phenomena in the atmosphere and oceans. By employing WN theory along with the HT, the stochastic equation was transformed into a deterministic form, enabling the application of the exp-function method. The solutions revealed that stochastic influences, such as B-M, introduce variability in wave forms, affecting both SPW dynamics. Numerical simulations demonstrated how these random fluctuations disrupt the smooth progression of waves, mirroring the complex dynamics observed in natural geophysical systems. The study offers a solid framework for exploring the interaction between deterministic and stochastic factors in large-scale wave propagation, providing valuable insights for future work on nonlinear wave equations with applications in fluid dynamics, oceanography, and atmospheric science. This study provides theoretical insights into geophysical wave dynamics that align with observed phenomena in the ocean and atmosphere. The deterministic solutions represent internal ocean waves and atmospheric gravity waves, both of which play crucial roles in energy transport and fluid mixing. The stochastic solutions incorporate environmental fluctuations such as wind stress variations and oceanic turbulence, which introduce uncertainties in wave behavior. Numerical simulations demonstrate how these perturbations influence wave stability and structure, reinforcing the real-world applicability of our findings. Future research could validate these results through experimental and observational studies in oceanography and atmospheric sciences.

    Areej A. Almoneef, Abd-Allah Hyder, Mohamed A. Barakat, Abdelrheem M. Aly: Methodology, Conceptualization, Data curation, Writing-original draft, Investigation, Visualization, Validation, Writing-review and editing. All authors have read and approved the final version of the manuscript for publication.

    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 Graduate Studies at King Khalid University for funding this work through the Research Groups Program under grant (RGP.2/82/45). The authors would like to acknowledge the Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2025R337).

    The authors declare that there are no conflicts of interest regarding the publication of this article.



    [1] J. W. Miles, The Korteweg-de Vries equation: A historical essay, J. Fluid Mech., 106 (1981), 131–147.
    [2] K. R. Khusnutdinova, Y. A. Stepanyants, M. R. Tranter, Soliton solutions to the fifth-order Korteweg-de Vries equation and their applications to surface and internal water waves, Phys. Fluids, 30 (2018), 022104. https://doi.org/10.1063/1.5009965 doi: 10.1063/1.5009965
    [3] A. A. Hyder, A. H. Soliman, C. Cesarano, M. A. Barakat, Solving Schrödinger-Hirota equation in a stochastic environment and utilizing generalized derivatives of the conformable type, Mathematics, 9 (2021), 2760. https://doi.org/10.3390/math9212760 doi: 10.3390/math9212760
    [4] A. A. Hyder, M. A. Barakat, General improved Kudryashov method for exact solutions of nonlinear evolution equations in mathematical physics, Phys. Scripta, 95 (2020), 045212. https://doi.org/10.1088/1402-4896/ab6526 doi: 10.1088/1402-4896/ab6526
    [5] Attaullah, M. Shakeel, M. K. Alaoui, A. M. Zidan, N. A. Shah, Closed form solutions for the generalized fifth-order KDV equation by using the modified exp-function method, J. Ocean Eng. Sci., In Press, 2022. https://doi.org/10.1016/j.joes.2022.06.037
    [6] K. R. Helfrich, W. K. Melville, Long nonlinear internal waves, Ann. Rev. Fluid Mech., 38 (2006), 395–425. https://doi.org/10.1146/annurev.fluid.38.050304.092129 doi: 10.1146/annurev.fluid.38.050304.092129
    [7] R. Grimshaw, E. Pelinovsky, T. Talipova, The modified Korteweg-de Vries equation in the theory of large-amplitude internal waves, Nonlinear Proce. Geoph., 4 (1997), 237–250. https://doi.org/10.5194/npg-4-237-1997 doi: 10.5194/npg-4-237-1997
    [8] S. T. R. Rizvi, A. R. Seadawy, F. Ashraf, M. Younis, H. Iqbal, D. Baleanu, Lump and interaction solutions of a geophysical Korteweg-de Vries equation, Results Phys., 19 (2020), 103661. https://doi.org/10.1016/j.rinp.2020.103661 doi: 10.1016/j.rinp.2020.103661
    [9] E. H. M. Zahran, A. Bekir, New unexpected behavior to the soliton arising from the geophysical Korteweg-de Vries equation, Mod. Phys. Lett. B, 36 (2022), 2150623. https://doi.org/10.1142/s0217984921506235 doi: 10.1142/s0217984921506235
    [10] P. L. F. Liu, X. M. Wang, A multi-layer model for nonlinear internal wave propagation in shallow water, J. Fluid Mech., 695 (2012), 341–365. https://doi.org/10.1017/jfm.2012.24 doi: 10.1017/jfm.2012.24
    [11] W. Hereman, Shallow water waves and solitary waves, Solitons, 2022,203–220. https://doi.org/10.1007/978-1-0716-2457-9_480
    [12] Q. Li, Nonlinear internal waves in the South China Sea, University of Rhode Island, 2010. https://doi.org/10.23860/diss-2353
    [13] J. P. Boyd, Nonlinear equatorial waves, Dynamics of the equatorial ocean, Heidelberg: Springer, 2017,329–404. https://doi.org/10.1007/978-3-662-55476-0_16
    [14] H. F. Ismael, M. A. S. Murad, H. Bulut, Various exact wave solutions for KdV equation with time-variable coefficients, J. Ocean Eng. Sci., 7 (2022), 409–418. https://doi.org/10.1016/j.joes.2021.09.014 doi: 10.1016/j.joes.2021.09.014
    [15] A. K. Gupta, S. S. Ray, On the solution of time-fractional KdV-Burgers equation using Petrov-Galerkin method for propagation of long wave in shallow water, Chaos, Soliton. Fract., 116 (2018), 376–380. https://doi.org/10.1016/j.chaos.2018.09.046 doi: 10.1016/j.chaos.2018.09.046
    [16] R. Grimshaw, Environmental stratified flows, Springer Science & Business Media, 2002. https://doi.org/10.1007/b100815
    [17] A. R. Osborne, Nonlinear ocean wave and the inverse scattering transform, Scattering, 2002,637–666. https://doi.org/10.1016/b978-012613760-6/50033-4
    [18] D. Kaya, Partial differential equations that lead to solitons, Solitons, 2022,193–201. https://doi.org/10.1007/978-1-0716-2457-9_380
    [19] A. Geyer, R. Quirchmayr, Shallow water equations for equatorial tsunami waves, Philos. T. Roy. Soc. A: Math. Phys. Eng. Sci., 376 (2018), 20170100. https://doi.org/10.1098/rsta.2017.0100 doi: 10.1098/rsta.2017.0100
    [20] H. Holden, B. Øksendal, J. Ubøe, T. Zhang, Stochastic partial differential equations, Springer, 2010. https://doi.org/10.1007/978-0-387-89488-1_5
    [21] H. Yépez-Martínez, J. F. Gómez-Aguilar, D. Baleanu, Beta-derivative and sub-equation method applied to the optical solitons in medium with parabolic law nonlinearity and higher order dispersion, Optik, 155 (2018), 357–365. https://doi.org/10.1016/j.ijleo.2017.10.104 doi: 10.1016/j.ijleo.2017.10.104
    [22] H. M. Baskonus, J. F. Gómez-Aguilar, New singular soliton solutions to the longitudinal wave equation in a magneto-electro-elastic circular rod with M-derivative, Mod. Phys. Lett. B, 33 (2019), 1950251. https://doi.org/10.1142/s0217984919502518 doi: 10.1142/s0217984919502518
    [23] B. Ghanbari, J. F. Gómez-Aguilar, Optical soliton solutions for the nonlinear Radhakrishnan-Kundu-Lakshmanan equation, Mod. Phys. Lett. B, 33 (2019), 1950402. https://doi.org/10.1142/s0217984919504025 doi: 10.1142/s0217984919504025
    [24] C. Qian, J. G. Rao, Y. B. Liu, J. S. He, Rogue waves in the three-dimensional Kadomtsev-Petviashvili equation, Chinese Phys. Lett., 33 (2016), 110201. https://doi.org/10.1088/0256-307x/33/11/110201 doi: 10.1088/0256-307x/33/11/110201
    [25] X. R. Hu, J. C. Chen, Y. Chen, Groups analysis and localized solutions of the (2+1)-dimensional Ito equation, Chinese Phys. Lett., 32 (2015), 070201. https://doi.org/10.1088/0256-307x/32/7/070201 doi: 10.1088/0256-307x/32/7/070201
    [26] X. Z. Liu, J. Yu, Z. M. Lou, X. M. Qian, A nonlocal Burgers equation in atmospheric dynamical system and its exact solutions, Chinese Phys. B, 28 (2019), 010201. https://doi.org/10.1088/1674-1056/28/1/010201 doi: 10.1088/1674-1056/28/1/010201
    [27] P. F. Zheng, M. Jia, A more general form of lump solution, lumpoff, and instanton/rogue wave solutions of a reduced (3+1)-dimensional nonlinear evolution equation, Chinese Phys. B, 27 (2018), 120201. https://doi.org/10.1088/1674-1056/27/12/120201 doi: 10.1088/1674-1056/27/12/120201
    [28] S. Q. Xu, X. G. Geng, N-soliton solutions for the nonlocal two-wave interaction system via the Riemann-Hilbert method, Chinese Phys. B, 27 (2018), 120202. https://doi.org/10.1088/1674-1056/27/12/120202 doi: 10.1088/1674-1056/27/12/120202
    [29] X. Y. Jiao, Truncated series solutions to the (2+1)-dimensional perturbed Boussinesq equation by using the approximate symmetry method, Chinese Phys. B, 27 (2018), 100202. https://doi.org/10.1088/1674-1056/27/10/100202 doi: 10.1088/1674-1056/27/10/100202
    [30] J. H. He, X. H. Wu, Exp-function method for nonlinear wave equations, Chaos, Soliton. Fract., 30 (2006), 700–708. https://doi.org/10.1016/j.chaos.2006.03.020 doi: 10.1016/j.chaos.2006.03.020
    [31] N. A. Kudryashov, N. B. Loguinova, Be careful with the exp-function method, Commun. Nonlinear Sci. Numer. Simul., 14 (2009), 1881–1890. https://doi.org/10.1016/j.cnsns.2008.07.021. doi: 10.1016/j.cnsns.2008.07.021
    [32] N. A. Shah, Y. S. Hamed, K. M. Abualnaja, J. D. Chung, R. Shah, A. Khan, A Comparative analysis of fractional-order Kaup-Kupershmidt equation within different operators, Symmetry, 14 (2022), 986. https://doi.org/10.3390/sym14050986 doi: 10.3390/sym14050986
    [33] M. Shakeel, Attaullah, N. A. Shah, J. D. Chung, Application of modified exp-function method for strain wave equation for finding analytical solutions, Ain Shams Eng. J., 14 (2023), 101883. https://doi.org/10.1016/j.asej.2022.101883 doi: 10.1016/j.asej.2022.101883
    [34] S. Zhang, Exact solutions of Wick-type stochastic KdV equation, Can. J. Phys., 90 (2012), 181–186. https://doi.org/10.1139/p2012-002 doi: 10.1139/p2012-002
    [35] S. Zhang, H. W. Li, B. Xu, Inverse scattering integrability and fractional soliton solutions of a variable-coefficient fractional-order KdV-type equation, Fractal Fract., 8 (2024), 520. https://doi.org/10.3390/fractalfract8090520 doi: 10.3390/fractalfract8090520
  • Reader Comments
  • © 2025 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(143) PDF downloads(33) Cited by(0)

Figures and Tables

Figures(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog