
This paper investigates a delayed shallow water fluid model that has not been studied in previous literature. Applying geometric singular perturbation theory, we prove the existence of traveling wave solutions for the model with a nonlocal weak delay kernel and local strong delay convolution kernel, respectively. When the convection term contains a nonlocal weak generic delay kernel, the desired heteroclinic orbit is obtained by using Fredholm theory and linear chain trick to prove the existence of two kink wave solutions under certain parametric conditions. When the model contains local strong delay convolution kernel and weak backward diffusion, under the same parametric conditions to the previous case, the corresponding traveling wave system can be reduced to a near-Hamiltonian system. The existence of a unique periodic wave solution is established by proving the uniqueness of zero of the Melnikov function. Uniqueness is proved by utilizing the monotonicity of the ratio of two Abelian integrals.
Citation: Minzhi Wei. Existence of traveling waves in a delayed convecting shallow water fluid model[J]. Electronic Research Archive, 2023, 31(11): 6803-6819. doi: 10.3934/era.2023343
[1] | Dong Li, Xiaxia Wu, Shuling Yan . Periodic traveling wave solutions of the Nicholson's blowflies model with delay and advection. Electronic Research Archive, 2023, 31(5): 2568-2579. doi: 10.3934/era.2023130 |
[2] | Yang Yang, Yun-Rui Yang, Xin-Jun Jiao . Traveling waves for a nonlocal dispersal SIR model equipped delay and generalized incidence. Electronic Research Archive, 2020, 28(1): 1-13. doi: 10.3934/era.2020001 |
[3] | Meng Wang, Naiwei Liu . Qualitative analysis and traveling wave solutions of a predator-prey model with time delay and stage structure. Electronic Research Archive, 2024, 32(4): 2665-2698. doi: 10.3934/era.2024121 |
[4] | Jun Meng, Shaoyong Lai . $ L^1 $ local stability to a nonlinear shallow water wave model. Electronic Research Archive, 2024, 32(9): 5409-5423. doi: 10.3934/era.2024251 |
[5] | Léo Girardin, Danielle Hilhorst . Spatial segregation limit of traveling wave solutions for a fully nonlinear strongly coupled competitive system. Electronic Research Archive, 2022, 30(5): 1748-1773. doi: 10.3934/era.2022088 |
[6] | Hami Gündoğdu . Impact of damping coefficients on nonlinear wave dynamics in shallow water with dual damping mechanisms. Electronic Research Archive, 2025, 33(4): 2567-2576. doi: 10.3934/era.2025114 |
[7] | Cui-Ping Cheng, Ruo-Fan An . Global stability of traveling wave fronts in a two-dimensional lattice dynamical system with global interaction. Electronic Research Archive, 2021, 29(5): 3535-3550. doi: 10.3934/era.2021051 |
[8] | Hao Wen, Yantao Luo, Jianhua Huang, Yuhong Li . Stochastic travelling wave solution of the $ N $-species cooperative systems with multiplicative noise. Electronic Research Archive, 2023, 31(8): 4406-4426. doi: 10.3934/era.2023225 |
[9] | Shan Zhao, Jun Feng . Qualitative analysis and traveling wave solutions of the (3+1)-dimensional Kadomtsev-Petviashvili-Sawada-Kotera-Ramani equation with conformable fractional derivative. Electronic Research Archive, 2025, 33(6): 3716-3732. doi: 10.3934/era.2025165 |
[10] | Shao-Xia Qiao, Li-Jun Du . Propagation dynamics of nonlocal dispersal equations with inhomogeneous bistable nonlinearity. Electronic Research Archive, 2021, 29(3): 2269-2291. doi: 10.3934/era.2020116 |
This paper investigates a delayed shallow water fluid model that has not been studied in previous literature. Applying geometric singular perturbation theory, we prove the existence of traveling wave solutions for the model with a nonlocal weak delay kernel and local strong delay convolution kernel, respectively. When the convection term contains a nonlocal weak generic delay kernel, the desired heteroclinic orbit is obtained by using Fredholm theory and linear chain trick to prove the existence of two kink wave solutions under certain parametric conditions. When the model contains local strong delay convolution kernel and weak backward diffusion, under the same parametric conditions to the previous case, the corresponding traveling wave system can be reduced to a near-Hamiltonian system. The existence of a unique periodic wave solution is established by proving the uniqueness of zero of the Melnikov function. Uniqueness is proved by utilizing the monotonicity of the ratio of two Abelian integrals.
Traveling waves in nonlinear wave equations can explain nonlinear complex phenomena in many subjects, such as chemistry, physics, biology, optics and mechanics. The well-known KdV equation is extremely important in modeling the motion of shallow water, which is given by
ut+αuux+βuxxx=0. | (1.1) |
It was first proposed by Korteweg and de Vries in 1895 and is usually used as a model to govern the one-dimensional propagation of small-amplitude, weakly dispersive waves [1]. In (1.1), the first two terms cause the classic overtaking phenomenon, while the last term prevents the formation of discontinuities. It is worth mentioning that the balance between the nonlinear convection term uux and the dispersion effect term uxxx in Eq (1.1) gives rise to solitons [2,3]. Some unusual nonlinear interactions among solitary wave pulses propagating in nonlinear dispersive media were observed in the numerical solutions. According to the important role in nonlinear models, there are a lot of investigations on finding the traveling wave solutions for KdV (1.1) and its generalized forms. In 1993, Derks and Gils [4] discussed the uniqueness of traveling waves in a perturbed KdV equation
ut+uux+uxxx+ε(uxx+uxxxx)=0, | (1.2) |
where ε is a positive parameter. Ogawa [5] studied the existence of solitary waves and periodic waves of (1.2) and gave the relationship between the amplitude and the wavelength. With a higher degree in convection term and by using the geometric singular perturbation theory, Yan et al. [6] proved the existence of solitary wave solutions and periodic wave solutions for a perturbed modified KdV equation
ut+unux+uxxx+ε(uxx+uxxxx)=0. |
Moreover, the KdV-mKdV equation
ut+uux±u2ux+uxxx=0 |
describes the internal solitary waves in shallow seas [7], which have been studied by the various methods [8,9,10,11,12,13,14,15,16,17]. Some new exact explicit solutions for a combined KdV-mKdV equation were obtained by means of the Bäcklund transformation [18] and the exact solutions for a new fractal unsteady KdV model with the non-smooth boundary by means of the sub-equation method were studied [19]. More precisely, Song [20] considered the diffusive single species model with Allee effect and distributed delay time, proving the existence of traveling wavefront solutions for the model with local strong and nonlocal weak generic delay kernels. Sun[21,22] studied a dispersive-dissipative solid model with weakly external dissipation and provided a rigorous proof for the existence of a unique periodic wave as well as investigated the following KdV equation with three perturbed terms
ut+λ1uqux+λ3uxxx+ε(λ2uxx+λ4uxxxx+λ5(uux)x)=0 |
with q=1,2. They proved the model possesses periodic waves with a range of wave speed and gave the explicit amplitude. Du [23] studied the existence of solitary wave solutions for the following generalized KdV-mKdV equation with local weak generic kernel delay
ut+αux+β(f∗u)up−1ux+uxxx+γuxx=0 | (1.3) |
by applying the geometric singular perturbation theory. Here, f∗u represents a convolution as a spatial-temporal variable. When τ→0, (1.3) reduces to a non-delayed model
ut+αux+βupux+uxxx+γuxx=0. | (1.4) |
Xu [24] established the existence of traveling wave solutions for (1.3)p=1. The parametric condition on the traveling wave fronts persisted was given. Now, we are interested in the wave motion model containing a special generic delay kernel in convection term. Consequently, in this paper, we investigate the following delay convecting shallow water fluid model
ut+αux+β((f∗u)u)x+u2ux+((1−q)τ−q)uxx+uxxx=0, | (1.5) |
where (α,β)∈R2, q={0,1}, τ>0 is a small parameter and f∗u represents a convolution in the spatial-temporal variable in (1.5); that is, there is a time delay in the lower order convection term. uxx is backward diffusion effect. When q=0, τ→0, (1.5) reduces to (1.4)p=1,γ=1. To our knowledge, no literature has considered the traveling wave solutions for when q=1. Therefore, the existence of traveling waves for the (1.5) is unknown. When the model contains different delay convolution kernels, are the traveling wave solutions persisted or vanished? If traveling wave solutions persisted, what is the type? What is the number? To solve these questions, we discuss the corresponding ordinary differential equation for (1.5) with a nonlocal weak and local strong delay convolution kernel, respectively. Geometric singular perturbation theory is utilized to reduce the singular perturbed system to regular perturbed system. The existence of traveling wave solutions is proved by different techniques in two cases: q=1 and q=0.
The rest of this paper is organized as follows.In section two, we introduce the geometric singular perturbation theory, which is a key to deal with the delayed equations. In section three, the delay convecting shallow water fluid model (1.5) in the case q=1 without delay is analyzed by qualitative theory. We prove that there are two heteroclinic orbits between the unstable node and saddles. For (1.5) in the case q=1 with a nonlocal weak generic delay kernel, the existence of locally invariant manifold in a small neighborhood of critical manifold is obtained, which reducing the singular perturbed system into a regular perturbed system. The existence of kink wave solutions for (1.5) in the case of q=1 is proved by the Fredholm theory and the linear chain trick. In Section 4, (1.5) in the case q=0 with local strong delay and the weak backward diffusion effect is considered. The singular perturbed system is reduced into regular perturbed system, which is a near-Hamiltonian system. We discuss the existence of periodic waves on certain parametric conditions by analyzing the monotonicity of ratio of two Abelian integrals in the Melnikov function. Section five is a simplified conclusion.
We first introduce the following results on invariant manifolds according to [25,26]. The basic equations considered are of the form
{dxdt=f(x,y,ε),dydt=εg(x,y,ε), | (2.1) |
where x=(x1,x2,…,xk)T∈Rk, y=(y1,y2,…,yl)T∈Rl and 0<ε≪1 is a real parameter. Functions f,g are C∞ on the set U×V, where U⊂Rk+l and V is an open interval containing zero.
With a change of time scaling z=εt, (2.1) can be written as
{εdxdz=f(x,y,ε),dydz=g(x,y,ε), | (2.2) |
then z is called the slow time scale and t is the fast time scale. Clearly, when ε≠0, (2.1) and (2.2) are equivalent. System (2.1) is called the fast system, while (2.2) is called the slow system. At the limit ε→0, system (2.1) reduces to a layer system
{x′(t)=f(x,y,0),y′(t)=0, | (2.3) |
and x is called the fast variable, whereas y is called the slow variable. When ε→0, the limit system of (2.2) is given by
{f(x,y,0)=0,˙y=g(x,y,0), | (2.4) |
which is called a reduced system. Assume that for ε=0, the system has a compact, normally hyperbolic manifold of critical manifold M0, which is contained in the set {f(x,y,0)=0}.
Definition 2.1. The manifold M0 is normally hyperbolic if the linearization of (2.1) at each point in M0 has exactly l eigenvalues with zero real part, where l is the dimension of the slow variable y.
Definition 2.2. A set M is locally invariant under the flow from (2.1) if it has neighborhood V so that no trajectory can leave M without also leaving V. In other words, it is locally invariant if for all x∈M, x⋅[0,t]⊂V implies that x⋅[0,t]⊂M, where the notation x⋅t is used to denote the application of a flow after time t to the initial condition x. Similarly with [0,t] replaced by [t,0], when t<0.
Under the previous hypotheses, the statement holds.
Lemma 2.1. If M0 is compact and normally hyperbolic, then, for any 0<r<+∞, if ε>0 is sufficiently small, there exists a manifold Mε, satisfying
(i) which is locally invariant under the flow of (2.1);
(ii) which is Cr in x,y and ε;
(iii)Mε={(x,y):x=hε(y)} for some Cr function hε(y) and y in some compact K;
(iv) there exists locally invariant stable and unstable manifolds Ws(Mε) and Wu(Mε) that lie within O(ε), and are diffeomorphic to Ws(M0) and Wu(M0), respectively.
Geometric singular perturbation theory is a powerful tool for analyzing high-dimensional systems and exploiting a differential equation's geometric structures, such as its slow (center) manifolds and their fast stable and unstable fibers [27,28,29,30,31].
In this section, the delayed convecting shallow water fluid model (1.5) in the case q=1 is analyzed. We discuss the existence of heteroclinic orbits connecting an unstable node to a saddle when (1.5) is without delay. For (1.5) in the case q=1 with a nonlocal weak generic delay kernel, a locally invariant manifold in a small neighborhood of a normal hyperbolic critical manifold is established, then a singular perturbed system is reduced to a regular perturbed system. The existence of kink wave solutions is proved by Fredholm theory and the linear chain trick.
According to the property of f when τ→0, (1.5) is the non-delay. In the case q=1, (1.5) reduces to
ut+αux+βu2x+u2ux−uxx+uxxx=0. | (3.1) |
For a given wave speed c>0, substituting u(x,t)=u(x+ct)=ϕ(ξ) into (3.1), the following traveling wave equation is obtained
(c+α)ϕ′+β(ϕ2)′+ϕ2ϕ′−ϕ″+ϕ‴=0, | (3.2) |
where ′=ddξ. Integrating (3.2) once and neglecting the integration constant, it can be simplified to
(c+α)ϕ+βϕ2+ϕ33−ϕ′+ϕ″=0, | (3.3) |
which is equivalent to a two-dimensional first-order system
{ϕ′=y,y′=−(c+α)ϕ−βϕ2−ϕ33+y. | (3.4) |
Clearly, (3.4) is a non-Hamiltonian system. Assume that 0<β<√33 and 0<c+α<3β24. Denote Δ:=β2−4(c+α)3. When Δ>0, it is easy to find that (3.4) has three equilibria, E0(0,0), E1(32(−β+√Δ),0) and E2(32(−β−√Δ),0). E0 is an unstable node and E1 and E2 are saddles. Now the existence of heteroclinic orbits between E0 and E1 is discussed. For a suitable value δ>0, there is a negative invariant triangular set
D:={(ϕ,y):0≤ϕ≤3(−β+√Δ)2,0≤y≤δϕ}. |
Let →m be the vector defined by the righthand side of (3.4) and →n=(−δ,1) be the outward normal vector on the boundary of D. On the side of y=δϕ, we have
→m⋅→n=(y,−(c+α)ϕ−βϕ2−ϕ33+y)⋅(−δ,1)∣(ϕ,δϕ)=−δ2ϕ−(c+α)ϕ−βϕ2−ϕ33+δϕ≤ϕ(−δ2+δ−(c+α)). | (3.5) |
It is clear that −δ2+δ−(c+α)=0 has two positive roots, δ1=1−√1−4(c+α)2 and δ2=1+√1−4(c+α)2. Therefore, when choosing δ≤δ1 or δ≥δ2, it has →m⋅→n≤0. Thus, one branch of the unstable manifold at E0(0,0) always stays in the region D and joins the saddle E1(3(−β+√Δ)2,0), which deduces the desired heteroclinic orbit that exists.
Similarly, the existence of a heteroclinic orbit between E0 and E2 can be proved. Therefore, from the relation between heteroclinic orbit and kink wave solution, the following statement holds.
Theorem 3.1. When 0<β<√33 and 0<c+α<3β24, there is a heteroclinic orbit connecting the unstable node E0(0,0) to the saddle E1(3(−β+√Δ)2,0) for (3.4). There is another heteroclinic orbit connecting the unstable node E0(0,0) to the saddle E2(3(−β−√Δ)2,0). Further, there are two kink wave solutions, u1(x+ct)=ϕ1(ξ) and u2(x+ct)=ϕ2(ξ), which satisfy that ϕ1(−∞)=0,ϕ1(+∞)=3(−β+√Δ)2 and ϕ2(−∞)=0,ϕ2(+∞)=3(−β−√Δ)2 with c as the wave speed.
From Section 3.1, when 0<β<√33 and 0<c+α<3β24, there are two heteroclinic orbits connecting the unstable node E0(0,0) to the saddle E1(3(−β+√Δ)2,0) and connecting E0(0,0) to E2(3(−β−√Δ)2,0), respectively, so we shall verify the heteroclinic orbit persists when the model contains the nonlocal delay. Due to the diffusion, the delay needs to be incorporated in a way that allows for associated spatial averaging. Based on the idea first introduced by Britton [32], the system is changed into a slow system. By geometric singular perturbation theory, the existence of locally invariant manifold in a small neighborhood of critical manifold is obtained, which reduces the singular perturbed system to a regular perturbed system. The existence of kink wave solutions for (1.5) is proved by the Fredholm theory and linear chain trick. The convolution f∗u is denoted by
(f∗u)(x,t)=∫t−∞∫∞−∞f(x−y,t−s)u(y,s)dyds. |
The kernel function f(x,t) satisfies the normalization condition
f:[0,+∞)×[0,+∞)→[0,+∞)and∫∞0∫∞−∞f(x,t)dxdt=1, |
so that the kernel does not affect the spatially uniform steady-state. Particularly, the nonlocal weak generic delay kernel is defined as follows
f(x,t)=1√4πte−x24t1τe−tτ, |
where the parameter τ>0 measures the average time delay. Denote that
η(x,t)=(f∗u)(x,t)=∫t−∞∫∞−∞1√4π(t−s)e−(x−y)24(t−s)1τe−t−sτu(y,s)dyds. |
By direct computation, we obtain
ηt=ηxx+1τ(u−η). |
Thus, (1.5) in the case q=1 is equivalent to a two-dimensional system as the form
{ut+αux+β(ηu)x+u2ux−uxx+uxxx=0,ηt=ηxx+1τ(u−η). | (3.6) |
To find the traveling wave solution of (3.6), the transformations u(x,t)=ϕ(ξ),η(x,t)=φ(ξ),ξ=x+ct are taken and we obtain a traveling wave system satisfying the boundary conditions ϕ(−∞)=0, ϕ(+∞)=3(−β+√Δ)2 and ϕ′(±∞)=0, which is given by
{(c+α)ϕ′+β(φϕ)′+ϕ2ϕ′−ϕ″+ϕ‴=0,cφ′−φ″−1τ(ϕ−φ)=0, | (3.7) |
where ′=ddξ. Integrating the first equation of (3.7) once, it obtains
(c+α)ϕ+βφϕ+ϕ33−ϕ′+ϕ″=0, |
then (3.7) changes to the following second order ordinary differential equation
{(c+α)ϕ+βφϕ+ϕ33−ϕ′+ϕ″=0,cφ′−φ′′−1τ(ϕ−φ)=0. | (3.8) |
The small parameter τ>0 represents the delay in the original system, which is regarded as the perturbed parameter. By defining new variables ϕ′=y,φ′=ω, (3.8) is reformulated as a four-dimensional system
{ϕ′=y,y′=−(c+α)ϕ−βφϕ−ϕ33+y,φ′=ω,ω′=cω−1τ(ϕ−φ). | (3.9) |
Setting that τ=ε2 and defining a new variable μ=εφ′, (3.9) is rewritten as a four-dimensional singular perturbed system
{ϕ′=y,y′=−(c+α)ϕ−βφϕ−ϕ33+y,εφ′=μ,εμ′=cεμ−ϕ+φ. | (3.10) |
Undoubtedly, (3.10) is a slow system. When ε→0, (3.9) reduces to (3.4). From Theorem 3.1, we know that (3.10) possesses a heteroclinic orbit connecting E0 to E1. Notice that when ε≠0, it does not define a dynamic in R4. Therefore, by the transformation ξ=εz, we change (3.10) into the form
{˙ϕ=εy,˙y=ε(−(c+α)ϕ−βφϕ−ϕ33+y),˙φ=μ,˙μ=εcμ−ϕ+φ, | (3.11) |
where ˙ is the derivative respect to z. System (3.11) is the fast system. Systems (3.10) and (3.11) are equivalent when ε>0. When ε=0, the slow system defines a set
M0={(ϕ,y,φ,μ)∈R4:μ=0,φ=ϕ}, |
which is an invariant manifold of (3.10) with ε=0. Since the linearized matrix of (3.11) restricted to M0 is
( 0 0 0 0 0 0 0 0 0 0 0 1−1 0 1 0), |
it is easy to obtain that the eigenvalues are 0, 0, 1, 1, the number of the eigenvalues with a zero real part are equal to dimM0 and the other eigenvalues are hyperbolic. Thus, the slow manifold M0 is normally hyperbolic. From geometric singular perturbation theory presented in section two, for sufficiently small ε>0, there exists a locally invariant manifold Mε in a small neighborhood of M0 of the perturbed system (3.10), which is expressed as
Mε={(ϕ,y,φ,μ)∈R4:μ=g(ϕ,y,ε),φ=ϕ+h(ϕ,y,ε)}, |
where g(ϕ,y,ε),h(ϕ,y,ε) are smooth functions and satisfy g(ϕ,y,0)=0,h(ϕ,y,0)=0. Thus the functions g(ϕ,y,ε) and h(ϕ,y,ε) can be expanded into a Taylor series as follows
g(ϕ,y,ε)=εg1(ϕ,y)+ε2g2(ϕ,y)+O(ε3),h(ϕ,y,ε)=εh1(ϕ,y)+ε2h2(ϕ,y)+O(ε3). |
Substituting φ=ϕ+h(ϕ,y,ε), μ=g(ϕ,y,ε) into the slow system (3.10), we have
cε{∂g1∂ϕy+∂g2∂y(−(c+α)ϕ−βφϕ−ϕ33+y)}+O(ε3)=cε2g1+εh1+ε2h2+O(ε3),cε{y+ε(∂h1∂ϕy+∂h1∂y(−(c+α)ϕ−βφϕ−ϕ33+y))}+O(ε3)=εg1+ε2g2++O(ε3). |
By comparing coefficients of ε and ε2, we obtain
g1(ϕ,y)=y,g2(ϕ,y)=0,h1(ϕ,y)=0,h2(ϕ,y)=−(c+α)ϕ−βϕ2−ϕ33−(c−1)y. |
Thus, the dynamics of (3.10) on Mε is determined by the following regular perturbed system
{ϕ′=y,y′=−(c+α)ϕ−βϕ2−ϕ33+y+ε2K(ϕ,y)+O(ε3), | (3.12) |
where K(ϕ,y)=β(c+α)ϕ2+β2ϕ3+βϕ43+(c−1)βϕy. Clearly, when ε=0, (3.12) reduces to (3.4). Denote the equilibria of (3.12) are Eε0, Eε1 and Eε2, which lying in a small neighborhood of E0, E1 and E2, respectively. In order to prove the existence of kink wave solutions of (1.5), we aim to establish the two heteroclinic orbits connecting Eε0 to Eε1, and another connecting Eε0 to Eε2, respectively. From Lemma 2.1, we know that such two heteroclinic orbits exist when ε=0.
Let (ϕ,y) and (u0,v0) be the solutions of (3.12) and (3.4), respectively. For ε>0, note that
ϕ=u0+ε2u1+O(ε3),y=v0+ε2v1+O(ε3). | (3.13) |
Substitute ϕ and y in (3.13) into (3.12) and compare the coefficients of ε and ε2, then u1 and v1 satisfy the following differential equation system
ddξ(u1v1)+(0−1 c+α+2βu0+u20−1)(u1v1)=( 0β(c+α)u20+β2u30+βu403). | (3.14) |
Notice that our goal is finding the traveling wave solution satisfying (3.14) and u1(±∞)=0,v1(±∞)=0. Denote L2 as the space of square integrable functions with inner production, that is
⟨u1(ξ),v1(ξ)⟩=∫+∞−∞(u1(ξ),v1(ξ))dξ, |
where ⟨⋅,⋅⟩ is the Euclidean inner product on R2. From the Fredholm theory, 3.14 has a solution if and only if the following integral equation is satisfied
∫+∞−∞(u1(ξ),(0β(c+α)u20+β2u30+βu403))dξ=0, |
for all functions u1(ξ) in the kernel of the adjoint of operator L defined by the left-hand side of (3.14). Denote L∗ as the adjoint of operator L, then
L∗=−ddξ+(0c+α+2βu0+u20−1−1). |
Implying that for all u1(ξ)∈KerL∗, it has
du1(ξ)dξ=(0c+α+2βu0+u20−1−1)u1(ξ). | (3.15) |
Since the matrix in (3.15) is a variable coefficient matrix, the general solution is difficult to derive. Therefore, we aim to prove that only the zero solution satisfies u0(±∞)=0 and we deduce the existence of homoclinic orbit. Even if the exact expression can not be found, u0(ξ) is a solution for the unperturbed system and satisfies the boundary condition u0(−∞)=0. Thus on the limit status ξ→−∞, the matrix in (3.15) approaches to a constant coefficient matrix
(0c+α−1−1). |
Clearly, the corresponding eigenvalues are determined by λ2+λ+c+α=0. Since 0<c+α<14, there are two real negative eigenvalues λ1,2=−1±√1−4(c+α)2<0. Hence, when ξ→−∞, the solution of (3.15) must be decreasing exponentially with respect to ξ, except for the zero solution. Therefore, the solution satisfying u1(±∞)=0 must be a zero solution, then the Fredholm orthogonality condition holds trivially, implying that such solutions of (3.15) exist and satisfy ϕ(−∞)=0 and y(±∞)=0. Consequently, we conclude that for sufficiently small ε>0, there exists two heteroclinic orbits of (3.15): One connects Eε0 to Eε1, and the other connects Eε0 to Eε2.
Theorem 3.2. In the case q=1, when 0<β<√33 and 0<c+α<3β24, for τ>0 is sufficiently small, the delayed convecting shallow water fluid model (1.5) with the nonlocal weak generic kernel
(f∗u)(x,t)=∫t−∞∫∞−∞1√4π(t−s)e−(x−y)24(t−s)1τe−t−sτu(y,s)dyds |
possesses two kink wave solutions u1(x,t)=ϕ1(ξ) and u2(x,t)=ϕ2(ξ), where ϕ1,2(ξ) satisfy ϕ1(−∞)=0, ϕ1(+∞)=3(−β+√Δ)2 and ϕ2(−∞)=0, ϕ2(+∞)=3(−β−√Δ)2. Here, c is the wave speed.
Remark 3.1. In the previous references [20,24], only one heteroclinic orbit was obtained. In our results, two heteroclinic orbits are proved under certain parametric conditions since there are three equilibria for the system.
In this section, we consider the traveling wave solution for Eq (1.5) in the case q=0 with local delay, that is, f(t)=tτ2e−tτ,t∈[0,+∞). Similar to the case q=1, making a traveling wave transformation ξ=x+ct to (1.5) in the case q=0 and integrating once, we obtain the traveling wave system
(c+α)ϕ+βϕω+ϕ33+ϕ″+τϕ′=0, | (4.1) |
where
ω=∫+∞0sτ2e−sτϕ(ξ−cs)ds. |
By direct calculation, we obtain that
dωdξ=1cτ(ζ−ω),dζdξ=1cτ(ϕ−ζ), | (4.2) |
where
ζ=∫+∞01τe−sτϕ(ξ−cs)ds. |
Introducing new variable ϕ′=y and combining with (4.2), (4.1) is changed to the following four-dimensional system
{ϕ′=y,y′=−((c+α)ϕ+βϕω+ϕ33+τy),cτω′=ζ−ω,cτζ′=ϕ−ζ, | (4.3) |
where ′ is derivative respect to ξ. System (4.3) is the slow system. When τ≠0, a time scale transformation ξ=τs is considered to change the slow system (4.3) into a fast system
{˙ϕ=τy,˙y=−τ((c+α)ϕ+βϕω+ϕ33+τy),c˙ω=ζ−ω,c˙ζ=ϕ−ζ, | (4.4) |
where ˙ is derivative respect to s. Systems (4.3) and (4.4) are equivalent when τ>0. The two different time scales correspond to two different limiting systems. When τ→0, (4.4) tends to the layer system
{˙ϕ=0,˙y=0,c˙ω=ζ−ω,c˙ζ=ϕ−ζ, | (4.5) |
and (4.3) tends to the reduced system
{ϕ′=y,y′=−((c+α)ϕ+βϕω+ϕ33+τy),0=ζ−ω,0=ϕ−ζ. | (4.6) |
Similarly, the critical manifold is given by
M0={(ϕ,y,ω,φ)∈R4:ω=ϕ,ζ=ϕ}, |
which is a slow invariant manifold. The linearized matrix of (4.5) is given as the form
( 0 0 0 0 0 0 0 0 0 0 −1c 1c 1c 0 0 −1c). |
It is not difficult to verify that the number of the eigenvalues with zero real part equals to dimM0 and the other eigenvalues are hyperbolic, then M0 is normally hyperbolic. Similarly, there exists a manifold Mτ for (4.3) with sufficiently small τ>0, which is locally invariant and diffeomorphic to M0 under the flow of (4.3). Then, Mτ can be expressed by
Mτ={(ϕ,y,ω,ζ)∈R4:ω=ϕ+k(ϕ,y,τ),ζ=ϕ+l(ϕ,y,τ)}, |
where k(ϕ,y,τ),l(ϕ,y,τ) are smooth functions and satisfy k(ϕ,y,0)=0, l(ϕ,y,0)=0. Thus k(ϕ,y,τ),l(ϕ,y,τ) can be expanded into Taylor series
k(ϕ,y,τ)=τk1(ϕ,y)+O(τ2),l(ϕ,y,τ)=τl1(ϕ,y)+O(τ2). |
Substituting ω=ϕ+k(ϕ,y,τ), ζ=ϕ+l(ϕ,y,τ) into the last equation of slow system (4.3), we have
cτ(y+O(τ))=τ(l1−k1)+O(τ2),cτ(y+O(τ))=−τl1+O(τ2). |
By comparing the coefficients of τ, we get k1(ϕ,y)=−2cy,l1(ϕ,y)=−cy. Thus, the slow system (4.3) restricted on Mτ reduces into a regular perturbed system
{ϕ′=y,y′=−(c+α)ϕ−βϕ2−ϕ33+τ(2βcϕy−y)+O(τ2), | (4.7) |
which is a near-Hamiltonian system. When τ→0, (4.7) reduces to a Hamiltonian system
{ϕ′=y,y′=−(c+α)ϕ−βϕ2−ϕ33, | (4.8) |
with the Hamiltonian function is given by
H(ϕ,y)=y22+c+α2ϕ2+β3ϕ3+112ϕ4. | (4.9) |
Notice that Δ=β2−4(c+α)3. When Δ>0, there are three equilibria E0(0,0), E1(32(−β+√Δ),0) and E2(32(−β−√Δ),0). When c+α>0, E0 is a center, E1 and E2 are saddles. The corresponding energy function values are H(0,0)=0, h1:=H(32(−β+√Δ),0) and h2:=H(32(−β−√Δ),0), respectively. Since we discuss the traveling wave for two models under the same parametric condition from the analysis in Section 3.1, we do not consider the case Δ=0. With the help of the energy function H(ϕ,y) on parametric conditions c+α>0, β2>43(c+α), we give the phase portrait of (4.8) in Figure 1.
Suppose that there exists a closed orbit Γh of (4.8) surrounding E0. A(h)∈Γh is the rightmost point on the positive ϕ-axis. For 0<|hτ−h|≪1, let Γhτ de a piece of the orbit of the perturbed (4.7) starting from A(h) to the next intersection point B(hτ) with the positive ϕ-axis. Then, the displacement function [33] is given by
d(h,τ)=∫^ABdH=τ(M(h)+O(τ)), |
where
M(h)=∮Γh(2βcϕy−y)dϕ=2βcJ1(h)−J0(h)=2βcJ0(h)(J1(h)J0(h)−12βc), |
which is called the Melnikov function with J1(h)=∮Γhϕydϕ and J0(h)=∮Γhydϕ=∬intΓhdϕdy>0. We shall show that the Abelian integral ratio P(h):=J1(h)J0(h) is strictly monotonic with respect to h, and further prove there exists a unique periodic wave solution for (1.5) in the case q=0. The following lemma provides a simple criterion to verify the monotonic of P(h).
Lemma 4.1. ([34]) Assume that the Hamiltonian function H(ϕ,y) can be written as y22+Φ(ϕ), satisfying
Φ′(ϕ)(ϕ−a)>0,forϕ∈(γ,A), |
then U′(h)>0 (or U′(h)<0) in (h1,h2) implies P′(h)>0 (or P′(h)<0) in (h1,h2). Here,
U(h):=μ(h)+ν(h),P(h):=∮Γhϕydϕ∮Γhydϕ, |
μ(h) and ν(h) are the inverse functions of the corresponding maps Φ: (γ,a)↦(h1,h2) and (a,A)↦(h1,h2), then it has γ<μ(h)<a<ν(h)<A and
Φ(μ(h))≡Φ(ν(h))≡h,h1<h<h2. |
For (1.5) in the case q=0, we have the following results.
Theorem 4.1. In the case q=0, for any sufficient small τ>0, there exist some suitable c,α,β that satisfy 0<β<√33 and 0<c+α<3β24, such that (1.5) has two isolated periodic wave solutions locatd at two sides of ϕ=32(−β+√β2−4(c+α)3) with c>0 as the wave speed.
Proof. According to the previous analysis, we discuss existence of periodic orbit near the family of closed orbits surrounding E0. Existence of periodic orbit near the family of closed orbits surrounding E2 can be proved similarly. Let Γh:={(ϕ,y):H(ϕ,y)=h}, which corresponds closed orbits of (4.8) for each h∈(0,h1) and bounded in a homoclinic loop connecting the saddle point E1. Then, Φ(ϕ):=c+α2ϕ2+β3ϕ3+112ϕ4 is analytic in the interval (0,A) and satisfying that Φ(0)=Φ(A), where A is the rightmost intersection point between the homoclinic loop and positive ϕ-axis. For c+α>0 and 32(−β+√Δ)<ϕ<A, it has
Φ′(ϕ)ϕ=ϕ23(3(c+α)+3βϕ+ϕ2)>0 |
implying that Φ(ϕ) has a minimum at ϕ=0 and is strictly monotonic on (32(−β+√Δ),0) and (0,A), respectively. Let μ(h) and ν(h) be inverse functions of Φ(ϕ) on these two intervals, respectively, and 32(−β+√Δ)<μ(h)<0<ν(h)<A. Define two functions
w(h):=μ(h)+ν(h)2,z(h):=ν(h)−μ(h)2. |
Then, the criterion function in s∈[0,z(h)] is given by
G(s):=Φ(w(h)+s)−Φ(w(h)−s)=2s3((β+w)s2+3βw2+w3+3(c+α)w). |
Since 0, z(h) and −z(h) are the real roots of G(s), we can rewrite G(s) as
G(s)=2s(β+w)3(s2−z(h)2)<0 |
for s∈(0,z(h)).
On the following, we prove U(h) is monotonic for h∈(0,h1) by contradiction argument. Assume that there exists ˜h and ˉh in (0,h1), ˜h<ˉh, such that U(˜h)=U(ˉh), then it has w(˜h)=w(ˉh) and z(˜h)<z(ˉh). Setting that h=ˉh, it yields
G(s)=Φ(w(ˉh)+s)−Φ(w(ˉh)−s)<0,s∈(0,z(h)). |
Letting s=z(˜h) and h=ˉh in G(s), we have
G(s)=Φ(w(ˉh)+z(˜h))−Φ(w(ˉh)−z(˜h))=Φ(w(˜h)+z(˜h))−Φ(w(˜h)−z(˜h))=Φ(μ(˜h))−Φ(ν(˜h))=0, |
which contradicts to G(s)<0 for all s∈(0,z(h)). Therefore, U(h) is strictly monotonic for h∈(0,h1). From Lemma 4.1, it has P(h) as strictly monotonic, which means there exists at most one h∗∈(0,h1) such that M(h∗)=0 and M′(h∗)≠0. By the implicit function theorem for sufficiently small τ>0, there exists at most one h=h∗+O(τ) such that d(h,τ)=0, then there exists at most one periodic wave for (1.5) in the case q=0.
Similarly, the existence of a unique periodic waves near the family of closed orbits surrounding E2 can be proved. The proof of Theorem 4.1 is completed.
This paper mainly discussed a convecting a shallow water fluid model in two cases with different generic delay kernels under ceratin parametric conditions. The existence of traveling waves for the model were given by different techniques. By applying the geometric singular perturbation theory, the existence of locally invariant manifold in a small neighborhood of critical manifold was obtained and the desired orbit was established. According to the relationship between traveling wave solution and orbit on a phase plane of the associated ordinary differential equation, the existence of traveling wave solution was proved. For the model in the case q=1 with nonlocal weak delay kernel, the heteroclinic orbit was established by the Fredholm theory and linear chain trick, which was an effective method to deal with physical models of delay. If the nonlocal weak delay kernel in presented paper was replaced by another delay kernel, the Fredholm theory and linear chain trick was also valid to establish the desired orbits for the corresponding traveling wave system. For the case q=0, (1.5) contained a local strong delay convolution kernel and a weak backward diffusion effect. It can be reduced to a near-Hamiltonian system, then to the existence of periodic wave solutions by investigating the monotonicity of ratio of two Abelian integrals in the Melnikov function. It is worth pointing out that no literature has considered both near-Hamiltonian and non-near-Hamiltonian cases of a delayed model. Consequently, it is an interesting work to be further researched in the future.
The author declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This research was supported by the National Natural Science Foundation of China (2001121) and Guangxi First-class Discipline statistics Construction Project Fund.
The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
The original contributions presented in the study are included in the article/Supplementary Material, and further inquiries can be directed to the corresponding author.
[1] | D. J. Korteweg, G. de Vries, On the change of form of long waves advancing in a rectangular channel and on a new type of long stationary waves, Philos. Mag. R Soc., 39 (1895), 422–413. |
[2] |
Z. Feng, On traveling wave solutions of the Burgers-Korteweg-de Vries equation, Nonlinearity, 20 (2007), 343–356. https://doi.org/10.1088/0951-7715/20/2/006 doi: 10.1088/0951-7715/20/2/006
![]() |
[3] |
N. J. Zabusky, M. D. Kruskal, Interaction of "solitons" in a collisionless plasma and the recurrence of initial states, Phys. Rev. Lett., 15 (1965), 240–243. https://doi.org/10.1103/PhysRevLett.15.240 doi: 10.1103/PhysRevLett.15.240
![]() |
[4] |
G. Derks, S. Gils, On the uniqueness of traveling waves in perturbed Korteweg-de Vries equations, Jpn. J. Ind. Appl. Math., 10 (1993), 413–430. https://doi.org/10.1007/BF03167282 doi: 10.1007/BF03167282
![]() |
[5] |
T. Ogama, Travelling wave solutions to a perturbed Korteweg-de Vries equation, Hiroshima Math. J., 24 (1994), 401–422. https://doi.org/10.32917/hmj/1206128032 doi: 10.32917/hmj/1206128032
![]() |
[6] |
W. Yan, Z. Liu, Y. Liang, Existence of solitary waves and periodic waves to a perturbed generalized KdV equation, Math. Model. Anal., 19 (2014), 537–555. https://doi.org/10.3846/13926292.2014.960016 doi: 10.3846/13926292.2014.960016
![]() |
[7] |
P. E. P. Holloway, E. Pelinovsky, T. Talipova, B. Barnes, A nonlinear model of internal tide transformation on the australian north west shelf, J. Phys. Ocean., 27 (1997), 871–896. https://doi.org/10.1175/1520-0485(1997)027<0871:ANMOIT>2.0.CO;2 doi: 10.1175/1520-0485(1997)027<0871:ANMOIT>2.0.CO;2
![]() |
[8] |
Z. Li, Constructing of new exact solutions to the GKdV-mKdV equation with any-order nonlinear terms by (G′/G)-expansion method, Appl. Math. Comput., 217 (2010), 1398–1403. https://doi.org/10.1016/j.amc.2009.05.034 doi: 10.1016/j.amc.2009.05.034
![]() |
[9] |
X. Li, Z. Du, S. Ji, Existence results of solitary wave solutions for a delayed Camassa-Holm-KP equation, Commun. Pure Appl. Anal., 18 (2019), 2961–2981. https://doi.org/10.3934/cpaa.2019152 doi: 10.3934/cpaa.2019152
![]() |
[10] |
X. Li, M. Wang, A sub-ODE method for finding exact solutions of a generalized KdV-mKdV equation with high-order nonlinear terms, Phys. Lett. A, 361 (2007), 115–118. https://doi.org/10.1016/j.physleta.2006.09.022 doi: 10.1016/j.physleta.2006.09.022
![]() |
[11] |
M. Song, X. Hou, J. Cao, Solitary wave solutions and kink wave solutions for a generalized KDV-mKDV equation, Appl. Math. Comput., 217 (2011), 5942–5948. https://doi.org/10.1016/j.amc.2010.12.109 doi: 10.1016/j.amc.2010.12.109
![]() |
[12] | K. Wang, G. Wang, F. Shi, Diverse optical solitons to the Radhakrishnan-Kundu-Lakshmanan equation for the light pulses, J. Nonlinear Opt. Phys. Mater., https://doi.org/10.1142/S0218863523500741 |
[13] |
K. Wang, J. Si, G. Wang, F. Shi, A new fractal modified Benjamin-Bona-Mahony equation: its genneralized variational principle and abundant exact solutions, Fractals, 31 (2023), 2350047. https://doi.org/10.1142/S0218348X23500470 doi: 10.1142/S0218348X23500470
![]() |
[14] |
H. Wu, Y. Zeng, T. Fan, On the extended KdV equation with self-consistent sources, Phys. Lett. A, 370 (2007), 477–484. https://doi.org/10.1016/j.physleta.2007.06.045 doi: 10.1016/j.physleta.2007.06.045
![]() |
[15] |
G. Xu, Y. Zhang, On the existence of solitary wave solutions for perturbed Degasperis-Procesi equation, Qual. Theory. Dyn. Syst., 20 (2021), 80. https://doi.org/10.1007/s12346-021-00519-0 doi: 10.1007/s12346-021-00519-0
![]() |
[16] |
J. Zhang, F. Wu, J. Shi, Simple soliton solution method for the combined KdV and MKdV equation, Int. J. Theor. Phys., 39 (2000), 1697–1702. https://doi.org/10.1023/A:1003648715053 doi: 10.1023/A:1003648715053
![]() |
[17] |
M. Antonova, A. Biswas, Adiabatic parameter dynamics of perturbed solitary waves, Commun. Non. Sci. Numer. Simul., 14 (2009), 734–748. https://doi.org/10.1016/j.cnsns.2007.12.004 doi: 10.1016/j.cnsns.2007.12.004
![]() |
[18] |
K. Wang, Bäcklund transformation and diverse exact explicit solutions of the fractal combined KdV-mKdV equation, Fractals, 30 (2022), 2250189. https://doi.org/10.1142/S0218348X22501894 doi: 10.1142/S0218348X22501894
![]() |
[19] |
K. Wang, A fractal modification of the unsteady Korteweg-de Vries model and its generalized fractal variational principle and diverse exact solutions, Fractals, 30 (2022), 2250192. https://doi.org/10.1142/S0218348X22501924 doi: 10.1142/S0218348X22501924
![]() |
[20] |
Y. Song, Y. Peng, M. Han, Travelling wavefronts in the diffusive single species model with Allee effect and distributed delay, Appl. Math. Comput., 152 (2004), 483–497. https://doi.org/10.1016/S0096-3003(03)00571-X doi: 10.1016/S0096-3003(03)00571-X
![]() |
[21] |
X. Sun, Y. Zeng, P. Yu, Analysis and simulation of periodic and solitary waves in nonlinear dispersive-dissipative solids, Commun. Non. Sci. Numer. Simul., 102 (2021), 105921. https://doi.org/10.1016/j.cnsns.2021.105921 doi: 10.1016/j.cnsns.2021.105921
![]() |
[22] |
X. Sun, W. Huang, J. Cai, Coexistence of the solitary and periodic waves in convecting shallow water fluid, Non. Anal.: RWA, 53 (2020), 103067. https://doi.org/10.1016/j.nonrwa.2019.103067 doi: 10.1016/j.nonrwa.2019.103067
![]() |
[23] |
Z. Du, D. Wei, Y. Xu, Solitary wave solutions for a generalized KdV-mKdV equation with distributed delays, Nonlinear Anal.-Model. Control, 19 (2014), 551–564. https://doi.org/10.15388/NA.2014.4.2 doi: 10.15388/NA.2014.4.2
![]() |
[24] |
Y. Xu, Z. Du, Existence of traveling wave fronts for a generalized KdV-mKdV equation, Math. Model. Anal., 19 (2014), 509–523. https://doi.org/10.3846/13926292.2014.956827 doi: 10.3846/13926292.2014.956827
![]() |
[25] | C. K. R. T. Jones, Geometric singular perturbation theory, in Dynamical Systems, Springer-Verlag, 1609 (1994), 45–118. https://doi.org/10.1007/BFb0095239 |
[26] |
N. Fenichel, Geometric singular perturbation theory for ordinary differential equations, J. Differ. Equations, 31 (1979), 53–98. https://doi.org/10.1016/0022-0396(79)90152-9 doi: 10.1016/0022-0396(79)90152-9
![]() |
[27] |
Z. Du, J. Liu, Y. Ren, Traveling pulse solutions of a generalized Keller-Segel system with small cell diffusion via a geometric approach, J. Differ. Equations, 270 (2021), 1019–1042. https://doi.org/10.1016/j.jde.2020.09.009 doi: 10.1016/j.jde.2020.09.009
![]() |
[28] |
J. Ge, Z. Du, The solitary wave solutions of the nonlinear perturbed shallow water wave model, Appl. Math. Lett., 103 (2020), 106202. https://doi.org/10.1016/j.aml.2019.106202 doi: 10.1016/j.aml.2019.106202
![]() |
[29] |
S. Ji, X. Li, Solitary wave solutions of delayed coupled Higgs Field equation, Acta Math. Sin. (English Series), 38 (2022), 97–106. https://doi.org/10.1007/s10114-022-0268-6 doi: 10.1007/s10114-022-0268-6
![]() |
[30] |
X. Li, Z. Du, J. Liu, Existence of solitary wave solutions for a nonlinear fifth-order KdV equation, Qual. Theory Dyn. Syst., 19 (2020), 24. https://doi.org/10.1007/s12346-020-00366-5 doi: 10.1007/s12346-020-00366-5
![]() |
[31] |
K. Zhuang, Z. Du, X. Lin, Solitary waves solutions of singularly perturbed higher-order KdV equation via geometric singular perturbation method, Nonlinear Dyn., 80 (2015), 629–635. https://doi.org/10.1007/s11071-015-1894-7 doi: 10.1007/s11071-015-1894-7
![]() |
[32] |
N. F. Britton, Spatial structures and periodic travelling waves in an integro-differential reaction-diffusion population model, SIAM J. Appl. Math., 50 (1990), 1663–1688. https://doi.org/10.1137/0150099 doi: 10.1137/0150099
![]() |
[33] | M. Han, Bifurcation Theory and Periodical Solution of Dynamic System, Science Press, Beijing, 2002. |
[34] |
C. Liu, D. Xiao, The monotonicity of the ratio of two Abelian integrals, Trans. Am. Math. Soc., 365 (2013), 5525–5544. https://doi.org/10.1090/S0002-9947-2013-05934-X doi: 10.1090/S0002-9947-2013-05934-X
![]() |
1. | Jun Meng, Shaoyong Lai, $ L^1 $ local stability to a nonlinear shallow water wave model, 2024, 32, 2688-1594, 5409, 10.3934/era.2024251 |