Research article Special Issues

Physical vs mathematical origin of the extended KdV and mKdV equations

  • The higher-order Korteweg-de Vries (KdV) and modified KdV (mKdV) equations are derived from a physical model describing a three-component plasma composed of cold fluid ions and two species of Boltzmann electrons at different temperatures. While the higher-order KdV equation is well established, the corresponding mKdV equation is typically derived using the system's integrability properties. In this work, we present the extended mKdV equation, derived directly from the physical system, offering a fundamentally different form from its integrable counterpart. We explore the connections between the two equations via Miura transformations and analyze their solutions within the framework of asymptotic integrability.

    Citation: Saleh Baqer, Theodoros P. Horikis, Dimitrios J. Frantzeskakis. Physical vs mathematical origin of the extended KdV and mKdV equations[J]. AIMS Mathematics, 2025, 10(4): 9295-9309. doi: 10.3934/math.2025427

    Related Papers:

    [1] Musong Gu, Chen Peng, Zhao Li . Traveling wave solution of (3+1)-dimensional negative-order KdV-Calogero-Bogoyavlenskii-Schiff equation. AIMS Mathematics, 2024, 9(3): 6699-6708. doi: 10.3934/math.2024326
    [2] Musawa Yahya Almusawa, Hassan Almusawa . Numerical analysis of the fractional nonlinear waves of fifth-order KdV and Kawahara equations under Caputo operator. AIMS Mathematics, 2024, 9(11): 31898-31925. doi: 10.3934/math.20241533
    [3] Zhi-Ying Feng, Xiang-Hua Meng, Xiao-Ge Xu . The data-driven localized wave solutions of KdV-type equations via physics-informed neural networks with a priori information. AIMS Mathematics, 2024, 9(11): 33263-33285. doi: 10.3934/math.20241587
    [4] Hayman Thabet, Subhash Kendre, James Peters . Travelling wave solutions for fractional Korteweg-de Vries equations via an approximate-analytical method. AIMS Mathematics, 2019, 4(4): 1203-1222. doi: 10.3934/math.2019.4.1203
    [5] 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. AIMS Mathematics, 2025, 10(3): 5859-5879. doi: 10.3934/math.2025269
    [6] M. Mossa Al-Sawalha, Rasool Shah, Adnan Khan, Osama Y. Ababneh, Thongchai Botmart . Fractional view analysis of Kersten-Krasil'shchik coupled KdV-mKdV systems with non-singular kernel derivatives. AIMS Mathematics, 2022, 7(10): 18334-18359. doi: 10.3934/math.20221010
    [7] Yanhua Gu . High-order numerical method for the fractional Korteweg-de Vries equation using the discontinuous Galerkin method. AIMS Mathematics, 2025, 10(1): 1367-1383. doi: 10.3934/math.2025063
    [8] Ihsan Ullah, Aman Ullah, Shabir Ahmad, Hijaz Ahmad, Taher A. Nofal . A survey of KdV-CDG equations via nonsingular fractional operators. AIMS Mathematics, 2023, 8(8): 18964-18981. doi: 10.3934/math.2023966
    [9] Yaya Wang, Md Nurul Raihen, Esin Ilhan, Haci Mehmet Baskonus . On the new sine-Gordon solitons of the generalized Korteweg-de Vries and modified Korteweg-de Vries models via beta operator. AIMS Mathematics, 2025, 10(3): 5456-5479. doi: 10.3934/math.2025252
    [10] Miguel Vivas-Cortez, Yasir Masood, Absar Ul Haq, Imran Abbas Baloch, Abdul Hamid Kara, F. D. Zaman . Symmetry analysis and conservation laws of time fractional Airy type and other KdV type equations. AIMS Mathematics, 2023, 8(12): 29569-29576. doi: 10.3934/math.20231514
  • The higher-order Korteweg-de Vries (KdV) and modified KdV (mKdV) equations are derived from a physical model describing a three-component plasma composed of cold fluid ions and two species of Boltzmann electrons at different temperatures. While the higher-order KdV equation is well established, the corresponding mKdV equation is typically derived using the system's integrability properties. In this work, we present the extended mKdV equation, derived directly from the physical system, offering a fundamentally different form from its integrable counterpart. We explore the connections between the two equations via Miura transformations and analyze their solutions within the framework of asymptotic integrability.



    The study of nonlinear wave phenomena has been a cornerstone of mathematical physics. An important model governing weakly nonlinear and weakly dispersive long waves, described by the field u(x,t), is the Korteweg-de Vries (KdV) equation,

    ut+q1uux+uxxx=0, (1.1)

    where q1 is a real constant (this constant can be trivially scaled out of the equation but is kept here for reasons that will become apparent later), and subscripts denote partial derivatives. This model, originally derived in the context of shallow water waves, has been a benchmark for understanding nonlinear wave phenomena, finding extensive applications in various physical contexts, such as shallow water wave dynamics [1,2], plasma physics [3], nonlinear lattices [4], and so on.

    An important variant of the KdV model is the modified KdV (mKdV) equation, where the quadratic nonlinear term uux is replaced by the cubic nonlinear term u2ux, namely:

    ut+r1u2ux+uxxx=0, (1.2)

    where r1 is, again, a real constant. Much like the KdV equation, the mKdV equation finds many applications in various fields, ranging from fluid mechanics [5,6], plasmas [7,8], and nonlinear optics with few-optical-cycle pulses [9,10] to traffic flow [11].

    Both the KdV and mKdV equations are completely integrable via the Inverse Scattering Transform (IST) [12]. Moreover, the equations are linked through the Miura transformation [5,13], which allows solutions of one equation to be mapped onto the other. This transformation not only highlights the deep connection between these systems but may also provide a pathway for deriving extended versions of the equations—incorporating higher-order dispersive and nonlinear terms—while preserving their integrable structure [14,15]. The relevant extended versions of the KdV and mKdV models are first, the extended KdV (eKdV) equation [16], which reads:

    ut+q1uux+uxxx+ε(q2u2ux+q3uxuxx+q4uuxxx+q5uxxxxx)=0, (1.3)

    where qj (j=2,3,4,5) are constants and ε1. Second, the extended mKdV (emKdV) equation is of the form [17,18]:

    ut+r1u2ux+uxxx+ε(r2u3x+r3u4ux+r4uuxuxx+r5u2uxxx+r6uxxxxx)=0, (1.4)

    where rj (j=2,3,,6) are constants.

    The need for higher-order equations arises from the limitations of simpler systems in accurately capturing the complexities of nonlinear wave phenomena. In physical problems, by systematically retaining higher-order terms in the perturbation expansion, extended equations provide corrections to wave speeds, amplitudes, and stability properties that align more closely with experimental and observational data. In this way, the eKdV equation has been derived from a variety of physical systems, including shallow water waves [16,19], solid mechanics [20], nonlinear optics of nematic liquid crystals [21], and plasma dynamics [22,23]. Importantly, the eKdV retains many properties of its integrable counterpart, such as soliton solutions, modulation theory solutions for dispersive shock waves (also known as "undular bores" in fluid mechanics [24]), and conservation laws. These properties make the eKdV equation a powerful tool for understanding the limitations of simpler models. It is thus not surprising that it has been used to describe complex solitary wave interactions [25,26], distinct regimes of shallow water dispersive shock propagation in the presence of surface tension effects [27,28], the transition from nonlocal to local effects in the evolution of defocusing nematic resonant dispersive shocks [21], and resonant soliton radiation in shallow water and optical media [29]. Recent studies have also explored its asymptotic integrability [30,31,32] and connections to higher-order hierarchies, further emphasizing its significance in both theoretical and applied contexts.

    Our scope in this work is to delve into the physical and mathematical origin of the eKdV and emKdV equations. To be more specific, we aim to investigate whether the emKdV equation (1.4), which is directly connected with the eKdV equation (1.3) via a Miura map as mentioned above, results—as a higher-order correction of the regular mKdV equation (1.2)—in a physical problem. To address this question, we consider a model of a plasma of cold positive ions in the presence of a two-temperature electron population. Employing the reductive perturbation method [33,34], we show that while the eKdV equation (1.3) can indeed result as a higher-order correction of the KdV equation (1.1), this is not the case with the higher-order mKdV equation: indeed, after deriving the regular mKdV equation (1.2), at the next order of approximation, we obtain a higher-order mKdV equation that is not of the form of Eq (1.4). Hence, unlike previous studies where the mKdV equation and its higher-order extensions were primarily derived as mathematical constructs, this study grounds a novel, physically relevant form of an emKdV equation. Thus, our study bridges the gap between physical and mathematical derivations of higher-order evolution equations and reveals nonlinear interactions that are absent in standard formulations.

    The extended KdV and mKdV equations derived in this work not only enhance our understanding of plasma dynamics but also provide a platform for exploring the asymptotic integrability of higher-order systems. Connections between the extended equations via Miura transformations further highlight the intricate relationship between these nonlinear systems and their integrable counterparts. We thus provide a comprehensive framework for analyzing these equations, with implications for both theoretical and applied research.

    The paper is organized as follows: In Section 2, we present the model and derive the KdV and eKdV equations. Section 3 is devoted to the derivation of the mKdV and emKdV equations. In Section 4, we present Miura map connections between the regular and extended KdV and mKdV models, while in Section 5, we establish an asymptotic integrability argument for the emKdV model derived in this work. Finally, in Section 6 we summarize our conclusions and discuss perspectives for future work.

    Our analysis is based on a set of equations describing a three-component plasma comprising cold fluid ions and two Boltzmann electron species at different temperatures [7]. The system consists of the continuity and momentum equations for the cold ions and Poisson's equation coupling the electrostatic potential to the plasma densities. In normalized/dimensionless form, the model under consideration expressed in (1+1)-dimensions reads:

    nt+x(nu)=0, (2.1a)
    ut+uux+φx=0, (2.1b)
    2φx2+nfexp(αcφ)(1f)exp(αhφ)=0. (2.1c)

    Here, n and u refer to the ion density and fluid velocity, respectively; ϕ is the electrostatic potential; and f is the fractional charge density of the cool electrons. The temperatures Tc and Th of the Boltzmann electrons are expressed through αc=Teff/Tc and αh=Teff/Th for the cool and hot species, respectively, whereas the effective temperature is given by Teff=TcTh/[fTh+(1f)Tc], such that fαc+(1f)αh=1.

    Our goal is to reduce Eq (2.1) to a single nonlinear evolution equation whose solutions are known and thus can asymptotically represent the solutions of the original system. To do so, we identify the equilibrium solution n=n0 (where n0 is the equilibrium density), u=0, and ϕ=0, and consider the following asymptotic expansions of the unknown fields:

    n=n0+εn1+ε2n2+ε3n3+, (2.2a)
    u=εu1+ε2u2+ε3u3+, (2.2b)
    φ=εφ1+ε2φ2+ε3φ3+, (2.2c)

    where 0<ε1 is a formal small parameter. Of particular interest is the linearized system occurring at O(ε):

    η1t+η0u1x=0, (2.3a)
    u1t+ϕ1x=0, (2.3b)
    2ϕ1x2[f(αcαh)+αh]ϕ1+η1=0. (2.3c)

    The dispersion relation of the above system can be obtained upon considering plane wave solutions of the form exp[i(kxωt)], where k and ω denote the wavenumber and frequency, respectively. Substituting, we obtain

    ω2=k2k2+[f(αcαh)+αh], (2.4)

    where η0=1, so that O(1) is also satisfied. Focusing our analysis on long waves (i.e., waves with small wave number k), we substitute [3,35] k=εpk (with p>0), where the exponent p is unknown (to be determined). Hence, the phase θ=kxωt of the plane waves is written as

    θ=kxωt=kx(a1k+a2k3)t=εpk(xa1t)ε3pk3a2t, (2.5)

    where a1=[f(αcαh)+αh]1/2 and a2=0.5[f(αcαh)+αh]3/2. As such, a natural rescaling based on Eq (2.5) is

    ξ=εp(xct),τ=ε3pβt, (2.6)

    where c is the velocity and β is an auxiliary parameter. These new variables ξ and τ are "slow", in the sense that it needs a large change in x and t in order to change ξ and τ appreciably. The value of p can be determined upon requiring the leading-order dispersion and nonlinearity terms of the system (2.1a)–(2.1c)—for the considered form of the asymptotic expansions in Eqs (2.2a)–(2.2c)—to be of the same order; this way, the perturbation scheme leads to a reduced model (which turns out to be the KdV equation in this case) that can support soliton solutions. This "maximal balance" condition [1] leads to p=1, and hence, the slow variables become

    ξ=ε1/2(xct),τ=ε3/2βt. (2.7)

    Note that the above slow variables are consistent with the similarity of the asymptotic behavior of the KdV equation, which holds for a coordinate system satisfying ζ(xct)/t1/3=const. [1]; hence, the asymptotic behavior along the direction defined by (2.7) is expected to be the same as that of the KdV equation.

    To proceed, we assume that the perturbations around the equilibrium solution, nj, uj and ϕj (with j=1,2,) in (2.2a)–(2.2c) depend on the slow variables (2.7), with the velocity c to be determined in a self-consistent manner (also, the value of β will be chosen below). Substitute back to Eq (2.1) and collect the different orders of the parameter ε. At the lowest order, O(1), we obtain the value of the equilibrium density, n0=1, while at O(ε) we obtain

    n1ξ[f(αcαh)+αh]φ1ξ=0, (2.8)

    and at O(ε3/2) we obtain

    n0u1ξcn1ξ=0, (2.9a)
    cu1ξ+φ1ξ=0. (2.9b)

    The compatibility of Eqs (2.8) and (2.9) yields the velocity c:

    c2=n0f(αcαh)+αh. (2.10)

    Nonlinear equations arise at the next orders, O(ε2) and O(ε5/2). Differentiating with respect to ξ the equations at order O(ε2), we find

    n2ξ[α2h+f(α2cα2h)]φ1φ1ξ[f(αcαh)+αh]φ2ξ=0, (2.11)

    so that the fields n2, u2, and φ2, are eliminated from the system at O(ε5/2):

    βn0c2φ1τ+n0u2ξcn2ξ+2n0c3φ1φ1ξ=0, (2.12a)
    βcφ1τcu2ξ+φ1φ1ξc2+φ2ξ=0. (2.12b)

    Then, using (2.11) and eliminating the fields n2, u2, and φ2, we obtain the regular KdV equation:

    φ1τ+c1φ1φ1ξ+φ1ξξξ=0, (2.13)

    where we have chosen β=c3/(2n0), and the nonlinearity coefficient c1 is given by

    c1=1n0[3(αcαh)2f2(αcαh)(αh(n06)+αcn0)f(n03)α2h]. (2.14)

    To extend the analysis to higher-order, i.e., derive an extended KdV equation, we take into account the higher-order of approximation, at O(ε7/2), and obtain

    βn2τ+ξ(u1n2+u2n1+n0u3cn3)=0, (2.15a)
    βu2τ+ξ(u2u1cu3+φ3)=0. (2.15b)

    Proceeding as above, we differentiate the equations at O(ε3) with respect to ξ and obtain

    c2u2τ+c22n0n2τ+2c2φ1φ1τ+2n0c3u2φ1ξ+1c2n2φ1ξ(α3h+(α3cα3h)f)c610n02c6φ21φ1ξ[α2h+(α2cα2h)f]φ2φ1ξ[α2h+(α2cα2h)f]c43n0c4φ1φ2ξ[αh+(αcαh)f]c2n0c2φ3ξ+φ2ξξξ=0. (2.16)

    Next, define Φ=φ1+εφ2 and eliminate the fields n3, u3, and φ3 from the higher-order equations to conclude with the eKdV equation:

    Φt+c1ΦΦξ+Φξξξ+ε(c2Φ2Φξ+c3ΦξΦξξ+c4ΦΦξξξ+c5Φξξξξξ)=0, (2.17)

    where

    c2=3[α2h+(α2cα2h)f]2c82n0[α3h+(α3cα3h)f]c6+2n0[α2h+(α2cα2h)f]c43n204n0c6, (2.18a)
    c3=9[α2h+(α2cα2h)f]c4+19n04n0c2, (2.18b)
    c4=3[α2h+(α2cα2h)f]c4+n02n0c2, (2.18c)
    c5=3c24n0. (2.18d)

    The solutions of the regular KdV equation, Eq (2.13), are well known and will be used to construct the solutions of the extended KdV equation, Eq (2.17), above. This will be done in a section below, as we now turn our attention to the derivation of the mKdV and extended mKdV equations.

    As above, the starting system is Eq (2.1). We use the asymptotic expansions:

    n=n0+εn1+ε2n2+ε3n3+, (3.1a)
    u=εu1+ε2u2+ε3u3+, (3.1b)
    φ=εφ1+ε2φ2+ε3φ3+, (3.1c)

    where the perturbations of the equilibrium solution now depend on the new slow variables:

    ξ=ε(xct),τ=ε3βt. (3.2)

    Substitute back to the system, and, similarly to the previous section, we obtain at different orders the following results. At order O(1), we find n0=1. At O(ε), we obtain

    n1ξ[f(αcαh)+αh]φ1ξ=0, (3.3)

    and at O(ε2) we obtain

    u1ξcn1ξ=0, (3.4a)
    cu1ξ+φ1ξ=0. (3.4b)

    The compatibility of the above equations leads again to the velocity c:

    c2=1(αcαh)f+αh, (3.5)

    as before. Continuing along the same lines, we get at O(ε2), after differentiating with respect to ξ,

    n2ξ[f(α2cα2h)+α2h]φ1φ1ξ[f(αcαh)+αh]φ2ξ=0, (3.6)

    and at O(ε3):

    u2ξcn2ξ+ξ(n1u1)=0, (3.7a)
    φ2ξcu2ξ+u1u1ξ=0. (3.7b)

    These lead to the following compatibility condition:

    3(αcαh)2f2(αcαh)(ac5αh)f+2α2h=0. (3.8)

    Contrary to the derivation of the KdV equation, here, and for the sake of consistency, the constants of the original system have to satisfy the above relation, Eq (3.8). As such, obtaining an mKdV equation is more challenging and restrictive than the KdV equation. Note that throughout our analysis, this is found to be true for any of the mKdV properties (solutions and Miura transformations).

    Continuing the analysis, at O(ε3) we obtain—after differentiating with respect to ξ—the following equation:

    n3ξ12[f(α3cα3h)+α3h]φ21φ1ξ[f(α2cα2h)+α2h](φ1ξφ2+φ1φ2ξ)[f(αcαh)+αh]φ3ξ+φ3ξξξ=0, (3.9)

    and at O(ε4) we obtain the system:

    βn1τ+ξ(n2u1+n1u2+u3cn3)=0, (3.10a)
    βu1τ+ξ(u1u2cu3+φ3)=0. (3.10b)

    Eliminating the fields n3, u3, φ3, and using the condition (3.8), we obtain the mKdV equation:

    φ1τc1φ21φ1ξ+φ1ξξξ=0, (3.11)

    where β=c3/2 and the coefficient c1 is given by

    c1=[(α3cα3h)f+α3h]c6152c6. (3.12)

    At the next order of approximation, O(ε5), we obtain the following equations:

    βn2τ+ξ(n3u1+n2u2+n1u3+u4cn4)=0, (3.13a)
    βu2τ+ξ(u1u3+12u2cu4+φ4)=0, (3.13b)

    as well as

    n4(α4cα4h)f+α4h24φ41(α2cα2h)f+α2h2φ22[(α2cα2h)f+α2h]φ3φ1(α3cα3h)f+α3h2φ2φ21[(αcαh)f+αh]φ4+φ2ξξξ=0. (3.14)

    Proceeding as in the case of the eKdV equation, we differentiate the above equation with respect to ξ, define Φ=ϕ1+εϕ2, and eliminate n3, u3, and ϕ3. In this way, we end up with the following extended mKdV equation:

    Φ1τ+c1Φ2Φξ+Φξξξ+ε(c2Φ3Φξ+c3ΦξΦξξ+c4ΦΦξξξ)=0, (3.15)

    where the coefficients c2, c3, and c4 are given by

    c2=16c8{[(c2αc14)α3c(c2αh14)α3h]c6f+(c2αh14)c6α3h+105}, (3.16a)
    c3=2c2, (3.16b)
    c4=4c2. (3.16c)

    This is a rather unexpected result. Indeed, as can be readily seen, the form of the emKdV equation (3.15) differs significantly from Eq (1.4), which can be derived from the eKdV via a Miura map, as will be shown below. In particular, there are no higher-order terms that yield resonant nonlinear dispersive wave solutions, such as resonant solitary and dispersive shock waves. This is due to the pure convexity or concavity of the associated linear dispersion relation [36,37], in contrast to the one governed by Eq (1.4).

    Consider the regular form of the KdV equation:

    Φτ+c1ΦΦξ+Φξξξ=0, (4.1)

    which can be shown to be converted to the relative mKdV equation:

    ˜Φτ+˜c1˜Φ2˜Φξ+˜Φξξξ=0, (4.2)

    under the Miura map:

    Φτ+c1ΦΦξ+Φξξξ=(2A˜Φ+Bξ)(˜Φτ+˜c1˜Φ2˜Φξ+˜Φξξξ)=0, (4.3)

    where Φ=A˜Φ2+B˜Φξ, A=˜c1/c1 and B2=6˜c1/c21.

    However, when the extended systems are considered, while such a map may still exist, certain restrictions have to apply. In particular, the extended KdV (denoted below as eKdV[Φ]), Eq (2.17), can be mapped to an extended mKdV (denoted below as emKdV[˜Φ]) [15],

    emKdV[˜Φ]=˜Φτ+˜c1˜Φ2˜Φξ+˜Φξξξ+ε(˜c2˜Φξ3+˜c3˜Φ4˜Φξ+˜c4˜Φ˜Φξ˜Φξξ+˜c5˜Φ2˜Φξξξ+˜c6˜Φξξξξξ)=0, (4.4)

    which is very different from Eq (3.15), as

    eKdV[Φ]=eKdV[A˜Φ2+B˜Φx]=(2A˜Φ+Bx)(emKdV)[˜Φ], (4.5)

    where A=˜c1/c1, B2=6˜c1/c21, as before. However, certain conditions now apply not only for the constants of the emKdV equation but also for the original eKdV equation; these are

    ˜c2=2˜c1c2c21,˜c3=˜c21c2c21,˜c4=8˜c1c2c21,˜c5=2˜c1c2c21,˜c6=6c25c21, (4.6)

    and for the eKdV equation:

    c3=4c2c1,c4=2c2c1,c5=6c25c21. (4.7)

    Note that these are consistent with the findings of Ref. [15].

    In this section, we will discuss the possibility of connecting the above derived eKdV (2.17) and emKdV (3.15) equations with their regular counterparts, the KdV and mKdV equations, respectively, employing the concept of asymptotic integrability. The latter refers to the transformation of a complicated, higher-order evolution equation (which may not be exactly solvable or integrable in the strict mathematical sense) to a simpler, integrable system [30,31,32]. Importantly, the connection of higher-order nonlinear evolution equations with their lower-order integrable counterparts with the asymptotic integrability argument allows for the derivation of solutions of such higher-order equations—see also Refs. [38] and [18] for asymptotic soliton solutions of the eKdV and emKdV equations, respectively.

    For the eKdV equation, which we write here with general coefficients,

    Φt+c1ΦΦξ+Φξξξ+ε(c2Φ2Φξ+c3ΦξΦξξ+c4ΦΦξξξ+c5Φξξξξξ)=0, (5.1)

    we introduce the transformation:

    Φ=Ψ+ε(λ1Ψ2+λ2Ψξξ+λ3ΨξΨdξ+λ4ξΨξξξ+λ5ξΨΨξ), (5.2)

    where

    λ1=18c2+3c1c4+2c21c518c1,λ2=6c2+c1c3c21c52c21,λ3=3c4+4c1c59,λ4=c53,λ5=c1c53, (5.3)

    so that

    Ψt+c1ΨΨξ+Ψξξξ=0. (5.4)

    The above KdV system is IST integrable, and its solutions may now be used to approximate the solutions of the eKdV equation. For example, consider the single soliton solution of Eq (5.4):

    Ψ(ξ,τ)=12η2c1sech2[η(ξ4η2τ)+ξ0], (5.5)

    where η and ξ0 are O(1) real parameters. Then the O(ε) correction based on Eq (5.2) reads:

    Φ(ξ,τ)=12η2[c1+4η2(λ2c16λ3)ε]c21sech2[η(ξ4η2τ)+ξ0]96η5λ4εc1ξsech2[η(ξ4η2τ)+ξ0]tanh[η(ξ4η2τ)+ξ0]+72η4(2λ1c1λ2+4λ3)εc21sech4[η(ξ4η2τ)+ξ0]+288η5(c1λ4λ5)εc21ξsech4[η(ξ4η2τ)+ξ0]tanh[η(ξ4η2τ)+ξ0]. (5.6)

    Here it should be noted that, in principle, a similar procedure could be used to identify other decaying approximate solutions of the eKdV equation from relevant solutions of the KdV equation. Such solutions include the rational solutions of the KdV (see, e.g., Refs. [39,40,41,42]) and are particularly relevant because they are connected with rogue waves [43]—especially in the context of the complex KdV equation [44,45].

    In a similar manner, consider the extended mKdV equation

    Φ1τ+c1Φ21Φ1ξ+Φ1ξξξ+ε(c2Φ3Φξ+c3ΦξΦξξ+c4ΦΦξξξ)=0, (5.7)

    and introduce the transformation

    Φ=Ψ+ε(λ1Ψ2+λ2ΨξΨdξ), (5.8)

    where now

    λ1=c3+c46,λ2=c43,3c2c1c3+c4/3=0. (5.9)

    This results in the integrable mKdV equation:

    Ψt+c1Ψ2Ψξ+Ψξξξ=0. (5.10)

    Notably, an additional restriction between the equation's coefficients needs to be held in order for the reduction to the simpler equation, also consistent with the method of Ref. [18] to asymptotically approximate the soliton of Eq (1.4).

    As in the case of the KdV, we proceed with the single soliton solution of the mKdV Eq (5.10), which is of the form

    Ψ(ξ,τ)=6c1ηsech[η(ξη2τ)+ξ0]. (5.11)

    Then substituting this solution into Eq (5.8) leads to the O(ε) correction:

    Φ(ξ,τ)=6c1ηsech[η(ξη2τ)+ξ0]+6η2λ1εc1sech2[η(ξη2τ)+ξ0]+6η2λ2εc1cot1(sinh[η(ξη2τ)+ξ0])sech[η(ξη2τ)+ξ0]tanh[η(ξη2τ)+ξ0]. (5.12)

    Similarly to the above discussion, one should expect that rational solutions of the emKdV equation could also be found by pertinent ones existing in the mKdV equation (see, e.g., Refs. [39,46,47,48,49]). Furthermore, relevant considerations could also be extended to the case of rogue waves thanks to the connections of the mKdV model with the nonlinear Schrödinger equation [50,51].

    In this work, we have derived and analyzed extended Korteweg-de Vries (KdV) and modified Korteweg-de Vries (mKdV) equations from a physical model describing a three-component plasma composed of cold fluid ions and two species of Boltzmann electrons at different temperatures. While we manage to recover the "usual" higher-order KdV system, our analysis provides a fundamentally different derivation compared to the conventional mKdV equation obtained through integrability considerations. Through the analysis of this new formulation, we have explored its structural properties and solutions, offering insights into its connection with the extended KdV equation via Miura transformations.

    One of the significant aspects of our study is the broader implication of these equations in the context of nonlinear wave theory. The KdV equation is widely regarded as a universal equation for weakly nonlinear, weakly dispersive wave systems. This universality arises due to its emergence in a diverse range of physical settings, from shallow water waves to plasma dynamics, optical fibers, and liquid crystals. A natural extension of this perspective is to investigate whether the extended mKdV equation we have derived could also share similar universal properties. Given that the standard mKdV equation appears in various contexts, including nonlinear optics and fluid dynamics, its extended version might exhibit a similarly broad applicability. Future work should focus on identifying physical systems where this equation naturally arises and exploring its integrability and solution structures in greater depth. Additionally, further investigations into asymptotic integrability and constructing the higher KdV system (if that exists) that is connected, through a Miura transformation, to the physically relevant mKdV system may provide useful information towards the "universal" character of the equation. Finally, the construction of rational and, when relevant, rogue wave solutions for the extended models considered in this work constitutes a very interesting future direction.

    S. Baqer, T. P. Horikis and D. J. Frantzeskakis: Conceptualization, Methodology, Validation, Writing-original draft, Writing-review & editing. All authors contributed equally to this article.

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

    The authors declare no conflict of interest.



    [1] M. J. Ablowitz, Nonlinear dispersive waves: asymptotic analysis and solitons, Cambridge: Cambridge University Press, 2011. https://doi.org/10.1017/CBO9780511998324
    [2] G. B. Whitham, Linear and nonlinear waves, New York: J. Wiley and Sons, 1974.
    [3] R. K. Dodd, J. C. Eilbeck, J. D. Gibbon, H. C. Morris, Solitons and nonlinear wave equations, London: Academic Press, 1982.
    [4] M. Remoissenet, Waves called solitons: concepts and experiments, Heidelberg: Springer, 1999. https://doi.org/10.1007/978-3-662-03790-4
    [5] R. M. Miura, C. S. Gardner, M. D. Kruskal, Korteweg-de Vries equation and generalizations, Ⅱ: existence of conservation laws and constants of motion, J. Math. Phys., 9 (1968), 1204–1209.
    [6] M. A. Helal, Soliton solution of some nonlinear partial differential equations and its applications in fluid mechanics, Chaos Soliton. Fract., 13 (2002), 1917–1929. https://doi.org/10.1016/S0960-0779(01)00189-8 doi: 10.1016/S0960-0779(01)00189-8
    [7] F. Verheest, C. P. Olivier, W. A. Hereman, Modified korteweg-de vries solitons at supercritical densities in two-electron temperature plasmas, J. Plasma Phys., 82 (2016), 905820208. https://doi.org/10.1017/S0022377816000349 doi: 10.1017/S0022377816000349
    [8] F. Verheest, W. A. Hereman. Collisions of acoustic solitons and their electric fields in plasmas at critical compositions, J. Plasma Phys., 85 (2019), 905850106. https://doi.org/10.1017/S0022377818001368 doi: 10.1017/S0022377818001368
    [9] H. Leblond, D. Mihalache, Few-optical-cycle solitons: modified Korteweg-de Vries sine-Gordon equation versus other non-slowly-varying envelope-approximation models, Phys. Rev. A, 79 (2009), 063835. https://doi.org/10.1103/PhysRevA.79.063835 doi: 10.1103/PhysRevA.79.063835
    [10] H. Leblond, D. Mihalache, Optical solitons in the fewcycle regime: recent theoretical results, Rom. Rep. Phys., 63 (2011), 1254–1266.
    [11] Z. P. Li, Y. C. Liu, Analysis of stability and density waves of traffic flow model in an its environment, Eur. Phys. J. B, 53 (2006), 367–374. https://doi.org/10.1140/epjb/e2006-00382-7 doi: 10.1140/epjb/e2006-00382-7
    [12] M. J. Ablowitz, H. Segur, Solitons and the inverse scattering transform, Philadelphia: Society for Industrial and Applied Mathematics, 1981. https://doi.org/10.1137/1.9781611970883
    [13] R. M. Miura, Korteweg-de Vries equation and generalizations, Ⅰ: a remarkable explicit nonlinear transformation, J. Math. Phys., 9 (1968), 1202–1204. https://doi.org/10.1063/1.1664700 doi: 10.1063/1.1664700
    [14] M. Ito, An extension of nonlinear evolution equations of the K-dV (mK-dV) type to higher orders, J. Phys. Soc. Jpn., 49 (1980), 771–778. https://doi.org/10.1143/JPSJ.49.771 doi: 10.1143/JPSJ.49.771
    [15] J. F. Gomes, G. S. Franca, A. H. Zimerma, Nonvanishing boundary condition for the mKdV hierarchy and the Gardner equation, J. Phys. A: Math. Theor., 45 (2012), 015207. https://doi.org/10.1088/1751-8113/45/1/015207 doi: 10.1088/1751-8113/45/1/015207
    [16] T. R. Marchant, N. F. Smyth, The extended Korteweg-de Vries equation and the resonant flow of a fluid over topography, J. Fluid Mech., 221 (1990), 263–287. https://doi.org/10.1017/S0022112090003561 doi: 10.1017/S0022112090003561
    [17] Y. Matsuno, Bilinearization of nonlinear evolution equations, Ⅱ: higher-order modified Korteweg-de Vries equations, J. Phys. Soc. Jpn., 49 (1980), 787–794. https://doi.org/10.1143/JPSJ.49.787 doi: 10.1143/JPSJ.49.787
    [18] T. R. Marchant, Asymptotic solitons for a higher-order modified Korteweg-de Vries equation, Phys. Rev. E, 66 (2002), 046623. https://doi.org/10.1103/PhysRevE.66.046623 doi: 10.1103/PhysRevE.66.046623
    [19] T. P. Horikis, D. J. Frantzeskakis, N. F. Smyth, Extended shallow water wave equations, Wave Motion, 112 (2022), 102934. https://doi.org/10.1016/j.wavemoti.2022.102934 doi: 10.1016/j.wavemoti.2022.102934
    [20] C. G. Hooper, P. D. Ruiz, J. M. Huntley, K. R. Khusnutdinova, Undular bores generated by fracture, Phys. Rev. E, 104 (2021), 044207. https://doi.org/10.1103/PhysRevE.104.044207 doi: 10.1103/PhysRevE.104.044207
    [21] S. Baqer, D. J. Frantzeskakis, T. P. Horikis, C. Houdeville, T. R. Marchant, N. F. Smyth, Nematic dispersive shock waves from nonlocal to local, Appl. Sci., 11 (2021), 4736. https://doi.org/10.3390/app11114736 doi: 10.3390/app11114736
    [22] P. Chatterjee, B. Das, G. Mondal, S. V. Muniandy, C. S. Wong, Higher-order corrections to dust ion-acoustic soliton in a quantum dusty plasma, Phys. Plasmas, 17 (2010), 103705. https://doi.org/10.1063/1.3491101 doi: 10.1063/1.3491101
    [23] Y. Kodama, T. Taniuti, Higher order approximation in the reductive perturbation method, Ⅰ: the weakly dispersive system, J. Phys. Soc. Jap., 45 (1978), 298–310. https://doi.org/10.1143/JPSJ.45.298 doi: 10.1143/JPSJ.45.298
    [24] G. A. El, M. A. Hoefer, Dispersive shock waves and modulation theory, Physica D, 333 (2016), 11–65. https://doi.org/10.1016/j.physd.2016.04.006 doi: 10.1016/j.physd.2016.04.006
    [25] 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
    [26] T. R. Marchant, N. F. Smyth, Soliton interaction for the extended Korteweg-de Vries equation, IMA J. Appl. Math., 56 (1996), 157–176. https://doi.org/10.1093/imamat/56.2.157 doi: 10.1093/imamat/56.2.157
    [27] T. R. Marchant, N. F. Smyth, An undular bore solution for the higher-order Korteweg-de Vries equation, J. Phys. A: Math. Gen., 39 (2006), L563. https://doi.org/10.1088/0305-4470/39/37/L02 doi: 10.1088/0305-4470/39/37/L02
    [28] S. Baqer, N. F. Smyth, Whitham shocks and resonant dispersive shock waves governed by the higher order korteweg-de vries equation, Proc. R. Soc. A, 479 (2023), 20220580. https://doi.org/10.1098/rspa.2022.0580 doi: 10.1098/rspa.2022.0580
    [29] T. P. Horikis, D. J. Frantzeskakis, T. R. Marchant, N. F. Smyth, Higher-dimensional extended shallow water equations and resonant soliton radiation, Phys. Rev. Fluids, 6 (2021), 104401. https://doi.org/10.1103/PhysRevFluids.6.104401 doi: 10.1103/PhysRevFluids.6.104401
    [30] A. S. Fokas, Q. M. Liu, Asymptotic integrability of water waves, Phys. Rev. Lett., 77 (1996), 2347. https://doi.org/10.1103/PhysRevLett.77.2347 doi: 10.1103/PhysRevLett.77.2347
    [31] A. S. Fokas, R. H. J. Grimshaw, D. E. Pelinovsky, On the asymptotic integrability of a higher‐order evolution equation describing internal waves in a deep fluid, J. Math. Phys., 37 (1996), 3415–3421. https://doi.org/10.1063/1.531572 doi: 10.1063/1.531572
    [32] Y. Kodama, On integrable systems with higher order corrections, Phys. Lett. A, 107 (1985), 245–249. https://doi.org/10.1016/0375-9601(85)90207-5 doi: 10.1016/0375-9601(85)90207-5
    [33] T. Kakutani, H. Ono, T. Taniuti, C. C. Wei, Reductive perturbation method in nonlinear wave propagation Ⅱ: application to hydromagnetic waves in cold plasma, J. Phys. Soc. Jpn., 24 (1968), 1159–1166. https://doi.org/10.1143/JPSJ.24.1159 doi: 10.1143/JPSJ.24.1159
    [34] R. A. Kraenkel, J. G. Pereira, M. A. Manna, The reductive perturbation method and the Korteweg-de Vries hierarchy, Acta Appl. Math., 39 (1995), 389–403. https://doi.org/10.1007/BF00994645 doi: 10.1007/BF00994645
    [35] R. Carretero-González, D. J. Frantzeskakis, P. G. Kevrekidis, Nonlinear waves in Hamiltonian systems: from one to many degrees of freedom, from discrete to continuum, Oxford: Oxford University Press, 2024. https://doi.org/10.1093/oso/9780192843234.001.0001
    [36] G. El, N. F. Smyth, Radiating dispersive shock waves in non-local optical media, Proc. Roy. Soc. Lond. A, 472 (2016), 20150633. https://doi.org/10.1098/rspa.2015.0633 doi: 10.1098/rspa.2015.0633
    [37] P. Sprenger, M. A. Hoefer, Shock waves in dispersive hydrodynamics with nonconvex dispersion, SIAM J. Appl. Math., 77 (2017), 26–50. https://doi.org/10.1137/16M1082196 doi: 10.1137/16M1082196
    [38] T. R. Marchant, Asymptotic solitons of the extended Korteweg-de Vries equation, Phys. Rev. E, 59 (1999), 3745. https://doi.org/10.1103/PhysRevE.59.3745 doi: 10.1103/PhysRevE.59.3745
    [39] M. J. Ablowitz, J. Satsuma, Solitons and rational solutions of nonlinear evolution equations, J. Math. Phys., 19 (1978), 2180–2186. https://doi.org/10.1063/1.523550 doi: 10.1063/1.523550
    [40] H. Airault, H. P. McKean, J. Moser, Rational and elliptic solutions of the kdv equation and related many-body problems, Commun. Pur. Appl. Math., 30 (1977), 95–148. https://doi.org/10.1002/cpa.3160300106 doi: 10.1002/cpa.3160300106
    [41] M. Adler, J. Moser, On a class of polynomials associated with the Korteweg-de Vries equation, Commun. Math. Phys., 61 (1978), 1–30. https://doi.org/10.1007/BF01609465 doi: 10.1007/BF01609465
    [42] P. A. Clarkson, Special polynomials associated with rational solutions of the Painlevé equations and applications to soliton equations, Comput. Methods Funct. Theory, 6 (2006), 329–401. https://doi.org/10.1007/BF03321618 doi: 10.1007/BF03321618
    [43] A. Ankiewicz, M. Bokaeeyan, N. Akhmediev, Shallow-water rogue waves: an approach based on complex solutions of the Korteweg-de Vries equation, Phys. Rev. E, 99 (2019), 050201. https://doi.org/10.1103/PhysRevE.99.050201 doi: 10.1103/PhysRevE.99.050201
    [44] D. Levi, Levi-Civita theory for irrotational water waves in a one-dimensional channel and the complex Korteweg-de Vries equation, Theor. Math. Phys., 99 (1994), 705–709. https://doi.org/10.1007/BF01017056 doi: 10.1007/BF01017056
    [45] D. Levi, M. Sanielevici, Irrotational water waves and the complex Korteweg-de Vries equation, Physica D, 98 (1996), 510–514. https://doi.org/10.1016/0167-2789(96)00109-1 doi: 10.1016/0167-2789(96)00109-1
    [46] Y. Kemataka, On rational similarity solutions of KdV and mKdV equations, Proc. Japan Acad. Ser. A Math. Sci., 59 (1983), 407–409. https://doi.org/10.3792/pjaa.59.407 doi: 10.3792/pjaa.59.407
    [47] Y. Sun, D. Zhang, Rational solutions with non-zero asymptotics of the modified Korteweg-de Vries equation, Commun. Theor. Phys., 57 (2012), 923–929. https://doi.org/10.1088/0253-6102/57/6/03 doi: 10.1088/0253-6102/57/6/03
    [48] A. Chowdury, A. Ankiewicz, N. Akhmediev, Periodic and rational solutions of modified Korteweg-de Vries equation, Eur. Phys. J. D, 70 (2016), 104. https://doi.org/10.1140/epjd/e2016-70033-9 doi: 10.1140/epjd/e2016-70033-9
    [49] J. Chen, D. E. Pelinovsky, Periodic travelling waves of the modified KdV equation and rogue waves on the periodic background, J. Nonlinear Sci., 29 (2019), 2797–2843. https://doi.org/10.1007/s00332-019-09559-y doi: 10.1007/s00332-019-09559-y
    [50] A. V. Slunyaev, E. N. Pelinovsky, Role of multiple soliton interactions in the generation of rogue waves: the modified Korteweg-de Vries equation, Phys. Rev. Lett., 117 (2016), 214501. https://doi.org/10.1103/PhysRevLett.117.214501 doi: 10.1103/PhysRevLett.117.214501
    [51] A. Ankiewicz, N. Akhmediev, Rogue wave-type solutions of the mKdV equation and their relation to known NLSE rogue wave solutions, Nonlinear Dyn., 91 (2018), 1931–1938. https://doi.org/10.1007/s11071-017-3991-2 doi: 10.1007/s11071-017-3991-2
  • 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(426) PDF downloads(33) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog