Loading [MathJax]/jax/output/SVG/jax.js
Research article

On a non-Newtonian fluid type equation with variable diffusion coefficient

  • Received: 03 June 2022 Revised: 19 July 2022 Accepted: 21 July 2022 Published: 03 August 2022
  • MSC : 35B35, 35K20, 35K55

  • Since the non-Newtonian fluid type equations arise from a broad and in-depth background, many research achievements have been gained from 1980s. Different from the usual non-Newtonian fluid equation, there is a nonnegative variable diffusion in the equations considered in this paper. Such a variable diffusion reflects the characteristic of the medium which may not be homogenous. By giving a generalization of the Gronwall inequality, the stability and the uniqueness of weak solutions to the non-Newtonian fluid equation with variable diffusion are studied. Since the variable diffusion may be degenerate on the boundary Ω, it is found that a partial boundary value condition imposed on a submanifold of Ω×(0,T) is enough to ensure the well-posedness of weak solutions. The novelty is that the concept of the trace of u(x,t) is generalized by a special way.

    Citation: Huashui Zhan, Yuan Zhi, Xiaohua Niu. On a non-Newtonian fluid type equation with variable diffusion coefficient[J]. AIMS Mathematics, 2022, 7(10): 17747-17766. doi: 10.3934/math.2022977

    Related Papers:

    [1] Muhammad Tahir, Yasir Khan, Adeel Ahmad . Impact of pseudoplastic and dilatants behavior of Reiner-Philippoff nanofluid on peristaltic motion with heat and mass transfer analysis in a tapered channel. AIMS Mathematics, 2023, 8(3): 7115-7141. doi: 10.3934/math.2023359
    [2] Khalil Ur Rehman, Nosheen Fatima, Wasfi Shatanawi, Nabeela Kousar . Mathematical solutions for coupled nonlinear equations based on bioconvection in MHD Casson nanofluid flow. AIMS Mathematics, 2025, 10(1): 598-633. doi: 10.3934/math.2025027
    [3] Özge Arıbaş, İsmet Gölgeleyen, Mustafa Yıldız . On the solvability of an initial-boundary value problem for a non-linear fractional diffusion equation. AIMS Mathematics, 2023, 8(3): 5432-5444. doi: 10.3934/math.2023273
    [4] Khalil Ur Rehman, Wasfi Shatanawi, Zeeshan Asghar, Haitham M. S. Bahaidarah . Neural networking analysis for MHD mixed convection Casson flow past a multiple surfaces: A numerical solution. AIMS Mathematics, 2023, 8(7): 15805-15823. doi: 10.3934/math.2023807
    [5] W. Abbas, Ahmed M. Megahed, Osama M. Morsy, M. A. Ibrahim, Ahmed A. M. Said . Dissipative Williamson fluid flow with double diffusive Cattaneo-Christov model due to a slippery stretching sheet embedded in a porous medium. AIMS Mathematics, 2022, 7(12): 20781-20796. doi: 10.3934/math.20221139
    [6] Hana Taklit Lahlah, Hamid Benseridi, Bahri Cherif, Mourad Dilmi, Salah Boulaaras, Rabab Alharbi . On the strong convergence of the solution of a generalized non-Newtonian fluid with Coulomb law in a thin film. AIMS Mathematics, 2023, 8(6): 12637-12656. doi: 10.3934/math.2023635
    [7] Wei Li, Guangying Lv . A fully-decoupled energy stable scheme for the phase-field model of non-Newtonian two-phase flows. AIMS Mathematics, 2024, 9(7): 19385-19396. doi: 10.3934/math.2024944
    [8] Amal Al-Hanaya, Munirah Alotaibi, Mohammed Shqair, Ahmed Eissa Hagag . MHD effects on Casson fluid flow squeezing between parallel plates. AIMS Mathematics, 2023, 8(12): 29440-29452. doi: 10.3934/math.20231507
    [9] Asifa, Poom Kumam, Talha Anwar, Zahir Shah, Wiboonsak Watthayu . Analysis and modeling of fractional electro-osmotic ramped flow of chemically reactive and heat absorptive/generative Walters'B fluid with ramped heat and mass transfer rates. AIMS Mathematics, 2021, 6(6): 5942-5976. doi: 10.3934/math.2021352
    [10] K. Veera Reddy, G. Venkata Ramana Reddy, Ali Akgül, Rabab Jarrar, Hussein Shanak, Jihad Asad . Numerical solution of MHD Casson fluid flow with variable properties across an inclined porous stretching sheet. AIMS Mathematics, 2022, 7(12): 20524-20542. doi: 10.3934/math.20221124
  • Since the non-Newtonian fluid type equations arise from a broad and in-depth background, many research achievements have been gained from 1980s. Different from the usual non-Newtonian fluid equation, there is a nonnegative variable diffusion in the equations considered in this paper. Such a variable diffusion reflects the characteristic of the medium which may not be homogenous. By giving a generalization of the Gronwall inequality, the stability and the uniqueness of weak solutions to the non-Newtonian fluid equation with variable diffusion are studied. Since the variable diffusion may be degenerate on the boundary Ω, it is found that a partial boundary value condition imposed on a submanifold of Ω×(0,T) is enough to ensure the well-posedness of weak solutions. The novelty is that the concept of the trace of u(x,t) is generalized by a special way.



    The mathematical modelling of various physical processes, where spatial heterogeneity has a primary role, usually results in the derivation of nonlinear evolution equations with variable diffusion, or dispersion. As pointed out by Karachalios-Zographopoulos in [18], to name but a few, equations of such a type have been successfully applied to the heat propagation in heterogeneous materials [5,12,16,17], the study of transport of electron temperature in a confined plasma [7], the propagation of varying amplitude waves in a nonlinear medium [24], the study of electromagnetic phenomena in nonhomogeneous superconductors [3,13,14,15] and the dynamics of Josephson junctions [8,9], the epidemiology and the growth and control of brain tumors [21]. It is not possible that all the characteristics of these applications explained in one equation due to its adaptability. In this paper, we focus on the non-Newtonian fluids equations. Applications of non-Newtonian fluids in wide range in many fields like fiber coating and crude oil extraction and many more have fascinated many researchers, one can refer to [22,25] etc. In Newtonian fluids, the viscosity does not change, while in non-Newtonian fluids, the viscosity changes when under force/stress to either more liquid or more solid, and so the non-Newtonian fluids are fluids that describe the relationship between deformation rates and stress none linearly, and they do not follow Newton's law of viscosity.

    Let us give three explicit equations derived from Newtonian flow or non-Newtonian flow. The first one is in the study of water infiltration through porous media, Darcy's linear relation

    V=K(θ)ϕ, (1.1)

    satisfactorily describes the flow conduction provided that the velocities are small. Here V represents the seepage velocity of water, θ is the volumetric moisture content, K(θ) is the hydraulic conductivity and ϕ is the total potential, which can be expressed as the sum of a hydrostatic potential ψ(θ) and a gravitational potential z

    ϕ=ψ(θ)+z.

    If it is assumed that infiltration takes place in a horizontal column of the medium, then the continuity equation has the form

    θt+Vx=0.

    Then we have

    θt=x(D(θ)|θx|θx), (1.2)

    with D(θ)=K(θ)ψ(θ). Certainly, water is the most usual Newtonian fluid. Also, Eq (1.5) is called as a porous medium equation.

    The second one is to consider a compressible fluid flow in a homogeneous isotropic rigid porous medium. Then the volumetric moisture content θ, the seepage velocity V and the density of the fluid are governed by the continuity equation

    θρt+div(ρV)=0. (1.3)

    For non-Newtonian fluid, according to Chapter 2 of [27], the linear Darcy's law is not longer valid, because the influence of many factors such as the molecular and ion effects needs to be concerned. Instead, one has the following nonlinear relation

    ρV=λ|P|α1P, (1.4)

    where ρV and P denote the momentum velocity and pressure respectively, λ>0 and α>0 are some physical constants. Combing (1.3) with (1.4), one has

    θρtλdiv(|P|α1P)=0. (1.5)

    The third one is to consider the flows in fractured media [20]. Let ε be the size ratio of the matrix blocks to the whole medium and let the width of the fracture planes and the porous block diameter be in the same order. If the permeability ratio of matrix blocks to fracture planes is of order εpε, where pε is a positive oscillating constant, then the nonlinear Darcy law combined with the continuity equation leads to the following equation

    ωεuεtdiv(kε(x)|uε|pε2uε)=0, (1.6)

    where uε is the density of the fluid, ωε,kε are the porosity and the permeability of the medium.

    Equations (1.4)–(1.6) can be abstracted as

    utdiv(a(x)|u|p2u)=0, (1.7)

    and the quantity p>1 is a characteristic of the medium, the media with p>2 are called dilatant fluids and those with p<2 are called pseudoplastics; if p=2 they are Newtonian fluids. While a(x) is only a nonnegative function, it reflects the characteristic of the medium which may not be homogenous. The most achievements before are focused on the case a(x)=1,

    utdiv(|u|p2u)=0 (1.8)

    which is simply called as the Non-Newtonian fluid equation usually. The existence, the uniqueness, the regularity and the long-time behaviors of weak solutions to this equation have been studied in [1,2,4,19,27,34] etc.

    In this paper, we consider the following non-Newtonian fluid type equation with a variable, nonnegative diffusion coefficient a(x):

    utdiv(a(x)|u|p2u)Ni=1bi(x,t)Diu=f(u,x,t), (x,t)QT=Ω×(0,T), (1.9)

    where Di=xi, 0a(x)C(¯Ω), bi(x,t)C1(¯QT), and f(u,x,t)C1(RׯQT), ΩRN is a bounded domain with a smooth boundary Ω. The most difference between Eq (1.9) and (1.8) is that the diffusion a(x) in Eq (1.9) may degenerate at some points of ¯Ω.

    A special case of (1.9) is that bi(x,t)=bi(x) and a(x)C1(¯Ω) satisfies

    a(x)|xΩ>0 and a(x)|xΩ=0. (1.10)

    In this case, the corresponding well-posed problem has been considered in [28,31,32] recently. In general, the initial value condition

    u(x,0)=u0(x), xΩ, (1.11)

    is always needed, but instead of the usual boundary value condition

    u(x,t)=0, (x,t)Ω×(0,T),

    only a partial boundary value condition

    u(x,t)=0, (x,t)ΣpΩ×(0,T), (1.12)

    is imposed, where Σp=Σ1×(0,T) and Σ1 is a relatively open subset of Ω in [28], even it can be an empty set sometime in [29,30,31]. In this paper, different [6,10,23,28,29,30,31], since bi(x,t) depends on the time variable t, we find that Σp can not be expressed as a cylindrical space as Σ1×(0,T). Instead, the partial boundary value condition is only imposed on a submanifold of ΣpΩ×(0,T) (the details are given in (1.16) or (3.2) below).

    Now, let us give the definition of weak solution and supply some other related backgrounds.

    Definition 1.1. A function u(x,t) is said to be a weak solution of the initial-boundary value problem of Eq (1.9), if

    uL(QT), utL2(QT), a(x)|u|pL1(QT), (1.13)

    and for any function g(s)C1(R) with g(0)=0, φ1C10(Ω) and φ2L(0,T;W1,ploc(Ω)), there holds

    QT[utg(φ1φ2)+a(x)|u|p2ug(φ1φ2)]dxdt+Ni=1QTu[bixi(x,t)g(φ1φ2)+bi(x,t)gxi(φ1φ2)]dxdt=QTf(u,x,t)g(φ1φ2)dxdt. (1.14)

    The initial value condition is satisfied in the sense of

    limt0Ω|u(x,t)u0(x)|dx=0. (1.15)

    Moreover, the partial boundary value condition is imposed as

    u(x,t)=0, (x,t)Σ={Ω×(0,T):Ni=1bi(x,t)axi(x)<0}. (1.16)

    If bi(x,t)=bi(x), f(u,x,t)=f(x,t)c(x,t)u, the existence of weak solution has been proved in [28]. In addition, if there is

    Ωa(x)1p1dx<, (1.17)

    then for a weak solution of Eq (1.9), we have

    Ω|u|dx<, (1.18)

    and the trace of u(x,t) on the boundary Ω can be defined in the classical sense [28,29]. If the inequality (1.17) is true and bi(x,t)=bi(x), then a similar partial boundary value condition as (1.16) has been imposed in [28]. In this paper, we mainly consider the case of that the inequality (1.17) is not true, i.e.,

    Ωa(x)1p1dx=, (1.19)

    then we can not define the trace of u(x,t) in the classical sense. So, the first dedication of this paper lies in that we give a generalization of the trace of u(x,t) on the boundary Ω in a special way. In details, inspired by [28,32], we can define the trace of u(x,t) on the boundary Ω as

    esssuplimε01ε(ΩεΩ2ε)Ω1tu2Ni=1bi(x,t)axi(x)dx=0, (1.20)

    where

    Ωε={xΩ:a(x)>ε}, Ω1t={xΩ:Ni=1bi(x,t)axi(x)0},

    and

    esssuplimλ0f(λ)=infδ>0{esssup{f(λ):|λ|<δ}}

    is the super limit. In what follows, we only simply denote esssuplimλ0f(λ) as limλ0f(λ). The rationality of such a generalization of the classical trace will be specified later in this paper.

    If f(u,x,t) is a continuous function and is Lipchitz continuous about the variable u, then the existence of weak solutions of the initial-boundary value problem of Eq (1.9) can be proved in a similar way as those [28,32], we don't prepare to prove the existence of weak solutions again. We mainly pay attention to the stability or the uniqueness of weak solutions by a generalized Gronwall inequality.

    Theorem 1.2. Let u(x,t) and v(x,t) be two weak solutions of the initial-boundary value problem of Eq (1.9), and with the same homogeneous the partial boundary condition

    u(x,t)=0=v(x,t), (x,t)Σ.

    Here Σ has the form (1.16). If 2>p>1, a(x) satisfies (1.10), and

    Ωa(x)2p1|Ni=1bi(x,t)|2pp1dx<, t[0,T], (1.21)

    then

    Ω|u(x,t)v(x,t)|2dxcΩ|u0(x)v0(x)|2dx, t[0,T). (1.22)

    Moreover, we can obtain a local stability of weak solutions as follows.

    Theorem 1.3. Let u(x,t) and v(x,t) be two solutions of Eq (1.9) with the differential initial values u0(x) and v0(x) respectively, but no any boundary value condition is required. if p>1, a(x) satisfies (1.10), whether (1.17) or (1.19) is true, and

    Ω|Ni=1axi(x)bi(x,t)|2a(x)dxc, t[0,T], (1.23)

    then

    Ωa(x)|u(x,t)v(x,t)|2dxcΩa(x)|u0(x)v0(x)|2dx, t[0,T). (1.24)

    Different from Theorem 1.2, in this theorem, there is not any boundary value condition imposed. Actually, the uniqueness of weak solution to Eq (1.9) can be obtained without conditions (1.21) and (1.23).

    Theorem 1.4. Let a(x)0 satisfy (1.10), p>1, u(x,t) and v(x,t) be two weak solutions of Eq (1.9) with the initial values u0(x)=v0(x). If a(x) satisfies (1.10) and one of the following assumptions is true.

    ⅰ) a(x) satisfies (1.17), u(x,t) and v(x,t) are with the same partial boundary value condition

    u(x,t)=v(x,t)=0,(x,t)Σ={(x,t)Ω×(0,T):div(b(x,t))0}, (1.25)

    in the sense of the classical trace. Here b={bi}, div(b(x,t))=Ni=1bi(x,t)xi.

    ⅱ) Whether a(x) satisfies (1.17) or (1.19), but

    div(b(x,t))=0, (x,t)Ω×(0,T). (1.26)

    Then

    u(x,t)=v(x,t), (x,t)QT. (1.27)

    One can see that all theorems above admit the case of (1.19), so the generalization of classical trace to the general form (1.20) is the most novelty of this paper. One can refer to the appendix for more details. Certainly, some other restrictions on the convective coefficient bi(x,t), i.e., the inequalities (1.21), (1.23) and (3.1), are imposed. How to relieve these restrictions to obtain the same conclusions? This is a question worth discussing thoroughly.

    Let us review the classical Gronwall inequality.

    Gronwall inequality: Let x(t) and c(t) be two nonnegative integral functions and a(t) be a bounded function on [0,T]. If

    x(t)t0c(τ)x(τ)dτ+a(t), t[0,T], (2.1)

    then

    x(t)sup0tT|a(t)|et0c(τ)dτ. (2.2)

    It is well-known that there are many applications of the Gronwall inequality in PDE, one can refer to [11,27] etc. In this paper, we find a generalization of the Gronwall inequality, and use it to prove the stability theorems of the degenerate parabolic equation (1.9).

    Lemma 2.1. Let x(t) and c(t) be two nonnegative integral functions on t[0,T], a(t) be a bounded function. If there is a constant 0<l1 such that

    x(t)(t0c(τ)x(τ)dτ)l+a(t), (2.3)

    then

    x(t)sup0tT|a(t)|ecτ0c(τ)dτ, (2.4)

    where c is a constant depending on T0x(τ)dτ.

    Proof. If l=1, there is nothing to be proved. When l<1, by (2.3), using the Young inequality, we have

    x(t)(t0c(τ)x(τ)dτ)l+a(t)lt0c(τ)x(τ)dτ+1l+a(t).

    By (2.1) and (2.2), we have (2.4).

    For small η>0, let

    Sη(s)=s0hη(τ)dτ,  hη(s)=2η(1sη)+.

    Obviously, hη(s)C(R), and

    limη0Sη(s)=sgns,  limη0shη(s)=0. (3.1)

    Proof of Theorem 1.2. From the definition of weak solution, if g(s)=s, for any φ1C10(Ω) and φ2L(0,T;W1,ploc(Ω)) we have

    QTφ1φ2(uv)tdxdt=QTa(x)(|u|p2u|v|p2v)(φ1φ2)dxdtNi=1QT(uv)[bixi(x,t)φ1φ2+bi(x,t)(φ1φ2)xi]dxdt+QT[f(u,x,t)v(v,x,t)]φ1φ2dxdt. (3.2)

    Denote Ωε={xΩ:a(x)>ε}. Let ξ be

    ξε(x)={1,  if  xΩ2ε,1ε[a(x)ε], if  xΩεΩ2ε,0,  if  xΩΩε.

    By a process of limit, we can choose

    φ1=ξε and φ2=χ[τ,s](uv)

    in (3.2), where, χ[τ,s] is the characteristic function on [τ,s](0,T). Then

    12Ω[u(x,s)v(x,s]2ξεdx=12Ω[u(x,τ)v(x,τ)]2ξεdxQτsξεa(x)(|u|p2u|v|p2v)(uv)dxdtQτs(uv)a(x)(|u|p2u|v|p2v)ξεdxdtNi=1Qτs(uv){bixi(x,t)(uv)ξε+bi(x,t)[(uv)ξε]xi}dxdt+Qτs[f(u,x,t)f(v,x,t)](uv)ξεdxdt, (3.3)

    where Qτs=Ω×[τ,s].

    A straightforward calculation leads to

    |Qτs(uv)a(x)(|u|p2u|v|p2v)ξεdxdt|Qτs|uv|a(x)(|u|p1+|v|p1)|ξε|dxdtcsτΩεΩ2ε[p1pa(x)(|u|p+|v|p)+1pa(x)|ξε|p]dxdtcsτΩεΩ2ε[p1pa(x)(|u|p+|v|p)+1pa(x)εp]dxdt. (3.4)

    Since 1<p<2, we have

    limε0ΩεΩ2εa(x)εpdx=limε01εΩεΩ2εa(x)ε(p1)dxlimε02εΩεΩ2εa(x)2pdx=2Ωa(x)2pdΣ=0.

    By this inequality, one can see that the right-hand side of (3.4) tends to 0 as ε0.

    Noticing that

    Ni=1Qτs(uv){bixi(x,t)(uv)ξε+bi(x,t)[(uv)ξε]xi}dxdt=Ni=1Qτs(uv)2bixi(x,t)ξεdxdt+Ni=1Qτs(uv)2bi(x,t)ξεxidxdt+Ni=1Qτs(uv)bi(x,t)ξε(uv)xidxdt, (3.5)

    if denoting

    Ω1t={xΩ:Ni=1bi(x,t)axi(x)0} and Ω2t={xΩ:Ni=1bi(x,t)axi(x)>0},

    by the partial boundary value condition (1.16), we have

    limε0sτΩ(uv)2Ni=1bi(x,t)ξεxidxdt=sτlimε01εΩεΩ2ε(uv)2Ni=1bi(x,t)axi(x)dxdtsτlimε01ε(ΩεΩ2ε)Ω1t(uv)2Ni=1bi(x,t)axi(x)dxdt=sτΣp(uv)2Ni=1bi(x,t)axi(x)dΣdt=0. (3.6)

    Meanwhile, since 1<p<2, we have pp1>2. Due to that u(x,t),v(x,t)L(QT), by (1.17), we have

    |Ω(uv)Ni=1bi(x,t)ξε(uv)xidx|=|Ω(uv)Ni=1bi(x,t)ξεa(x)1pa(x)1p(uv)xidx|(Ω|(uv)Ni=1bi(x,t)a(x)1p|pp1dx)p1p(Ωa(x)(|u|p+|v|p)dx)1pc(Ω|(uv)Ni=1bi(x,t)a(x)1p|pp1dx)p1p=c(Ω|uv||uv|pp11|Ni=1bi(x,t)a(x)1p|pp1dx)p1pc(Ω|uv||Ni=1bi(x,t)a(x)1p|pp1dx)p1pc(Ω|uv|2dx)p12p(Ω|Ni=1bi(x,t)a(x)1p|2pp1dx)p12pc(Ω|uv|2dx)p12p. (3.7)

    Naturally, we have

    Qτsξεa(x)(|u|p2u|v|p2v)(uv)dxdt0. (3.8)

    Since u(x,t),v(x,t)L(QT) and f(u,x,t)C1(RׯQT), by (3.4), (3.6)–(3.8), from (3.3), we easily obtain

    Ω[u(x,s)v(x,s)]2dxΩ[u(x,τ)v(x,τ)]2dx+c(sτΩ|u(x,t)v(x,t)|2dxdt)l,

    where l1.

    By virtue of the generalized Gronwall Lemma 2.1, choosing x(s)=Ω[u(x,s)v(x,s)]2dx, we have

    Ω[u(x,s)v(x,s)]2dxc(T)Ω[u(x,τ)v(x,τ)]2dx,

    and letting τ0, we arrive at the desire.

    Proof of Theorem 1.3. Let u(x,t) and v(x,t) be two solutions of Eq (1.9) with the initial values u0(x) and v0(x) respectively, but without any boundary value condition. From the definition of the weak solution, we can choose χ[τ,s]a(x)(uv) as a test function, where χ[τ,s] is the characteristic function on [τ,s][0,T). Denoting Qτs=Ω×[τ,s], then we have

    Qτsa(x)(uv)(uv)tdxdt=Qτsa(x)(|u|p2u|v|p2v)[a(x)(uv)]dxdtNi=1Qτs(uv)[bixi(x,t)a(x)(uv)+bi(x,t)(a(x)(uv))xi]dxdt+Qτs[f(u,x,t)f(v,x,t)]a(x)(uv)dxdt. (3.9)

    In particular,

    Qτsa(x)(|u|p2u|v|p2v)[a(x)(uv)]dxdt=Qτsa(x)2(|u|p2u|v|p2v)(uv)dxdt+Qτsa(x)(|u|p2u|v|p2v)(uv)adxdt. (3.10)

    Clearly,

    Qτsa(x)2(|u|p2u|v|p2v)(uv)dxdt0. (3.11)

    For the second term on the right hand side of (3.10), we have

    |Qτs(uv)a(x)(|u|p2u|v|p2v)adxdt|Qτs|uv|a(x)(|u|p1+|v|p1)|a|dxdtc(sτΩa(x)(|u|p+|v|p)dxdt)p1p(sτΩa(x)|a|p|uv|pdxdt)1pc(sτΩa(x)|a|p|uv|pdxdt)1pc(sτΩa(x)|uv|pdxdt)1p. (3.12)

    If p2,

    (sτΩa(x)|uv|pdxdt)1pc(sτΩa(x)|uv|2dxdt)1p. (3.13)

    If 1<p<2, by then the Hölder inequality

    (sτΩa(x)|uv|pdxdt)1pc(sτΩa(x)|uv|2dxdt)12. (3.14)

    Meanwhile, by (1.23),

    Ω|Ni=1axi(x)bi(x,t)|2a(x)dxc(T),

    since u(x,t),v(x,t)L(QT), we easily get that

    |QτsNi=1bi(x,t)(uv)[(uv)a]xidxdt|=|QτsNi=1bi(x,t)(uv)2axidxdt+QτsNi=1bi(x,t)(uv)(uv)xia(x)dxdt|csτΩ|uv||Ni=1axibi(x,t)|dx+(sτΩa(1p)p(a(x)Ni=1bi(x,t)|uv|)pdxdt)1p(sτΩa(x)(|u|p+|v|p)dxdt)1pc(sτΩ|Ni=1axibi(x,t)a(x)|2dxdt)12(sτΩa(x)|uv|2dxdt)12+c(sτΩ|Ni=1bi(x,t)|pp1dxdt)1p(sτΩa(x)(|u|p+|v|p)dxdt)1pc(sτΩa(x)|uv|2dxdt)12+c(sτΩa(x)|uv|pdxdt)1p. (3.15)

    Here, p=pp1 as usual.

    At the same time, we have

    Qτs|Ni=1bixi(x,t)(uv)2|a(x)dxdtc(sτΩa(x)|uv|2dxdt)12, (3.16)

    and since |f(u,x,t)f(v,x,t)|c, by that u(x,t),v(x,t)L(QT), using the Hölder inequality, we have

    Qτs|[f(u,x,t)f(v,x,t)](uv)|a(x)dxdtc(sτΩa(x)|uv|2dxdt)12. (3.17)

    Also,

    Qτs(uv)a(x)(uv)tdxdt=Ωa(x)[u(x,s)v(x,s)]2dxΩa(x)[u(x,τ)v(x,τ)]2dx. (3.18)

    At last, by (3.10)–(3.18), we let λ0 in (3.9). Then

    Ωa(x)[u(x,s)v(x,s)]2dxΩa(x)[u(x,τ)v(x,τ)]2dxc(s0Ωa(x)|u(x,t)v(x,t)|2dxdt)q, (3.19)

    where q<1. By (3.19), by the generalized Gronwall Lemma 2.1, choosing

    x(s)=Ωa(x)u(x,s)v(x,s)2dx,

    we have

    Ωa(x)u(x,s)v(x,s)2dxΩa(x)u(x,τ)v(x,τ)2dx. (3.20)

    Thus, by the arbitrary of τ, we have

    Ωa(x)u(x,s)v(x,s)2dxΩa(x)u0(x)v0(x)2dx.

    The proof is complete.

    In what follows, we study the uniqueness of weak solution to Eq (1.9).

    Proof of Theorem 1.4. Let u(x,t) and v(x,t) be two weak solutions of Eq (2.1) with the same initial value u0(x)=v0(x).

    For any given small δ>0, we denote Dδ={xΩ:|w|=|uv|>δ}. Without loss of the generality, we may assume that there is a δ>0 such that the measure μ(Dδ)>0. Let φλ(ξ) be a even function. When ξ0, it its defined as

    φλ(ξ)={11βλβ111βξβ1,  if  ξ>λ,0,   if ξλ, (3.21)

    where λ is small enough satisfying δ>2λ>0, 1>β>0, and we define that

    Φλ(ξ)=ξ0φλ(ς)dς.

    By multiplying Eq (1.9) by φλ(w)=φλ(uv), integrating over Qt=Ω×(0,t), since a(x)=0 when xΩ, using the condition (3.2) in ⅰ) or the condition (3.1) in ⅱ), we have

    0=t0Ω[wtφλ(w)+a(x)(|u|p2u|v|p2v)φλ(w)]dxdt+Ni=1t0Ωbi(x,t)xiΦλ(uv)dxdtNi=1t0Ω(f(u,x,t)f(v,x,t))φλ(uv)dxdt. (3.22)

    Let us analyse the three terms in the righthand side of this equality. First, the monotone inequality of the operator Δ yields

    t0Ωa(x)(|u|p2u|v|p2v)(uv)φλ(uv)dxdtt0Dλa(x)(uv)2β2p|w|pφλ(uv)dxdt0. (3.23)

    Secondly, since u(x,t),v(x,t)L(QT),

    |Ni=1t0Ωbi(x,t)xiΦλ(uv)dxdt|+|Ni=1t0Ω(f(u,x,t)f(v,x,t))φλ(uv)dxdt|c. (3.24)

    Thirdly, let t0=inf{τ(0,t]:w>λ}. Then

    t0Dλwtφλ(w)dxdt=Dλ(t00wtφλ(w)dt+tt0wtφλ(w)dt)dxDλw(x,t)λφλ(s)dsdxDλ(w2λ)φλ(2λ)dx(δ2λ)φλ(2λ)μ(Dλ). (3.25)

    From (3.23)–(3.25), we have

    (δ2λ)12β11βλβ1c.

    Letting λ0, we get the contradiction.

    At the end of this section, we would like to point that, φλ(w)=φλ(uv) does not satisfy the request of the test function g(φ1φ2), so in the proof of Theorem 1.4, we do not use the equality (1.14) in Definition 1.1. However, since the condition (3.2) in ⅰ) or the condition (3.1) in ⅱ), all the boundary integrals disappears in (3.22), and the proof of Theorem 1.4 is true.

    The non-Newtonian fluid equation with a variable diffusion is more applicable than the usual fluid equation. In the PDE theory, since the variable diffusion a(x) may be degenerate on some points of ¯Ω, one can not deduce the inequality

    QT|u|pdxdt<,

    and the weak solution matching up with the equation does not belong to Lp(0,T;W1,p0(x)). Such a fact makes the boundary value condition can not be imposed in the classical way. In this paper, a new kind of weak solution introduced, the trace is generalized in accordance with the weak solution defined, and it is found that a partial boundary value condition on a submanifold of Ω×(0,T) is enough to ensure the weak solution well-posedness. In fact, the definitions, the theorems and the methods used in this paper are able to be generalized to study the well-posed problem of the other degenerate parabolic equations such as

    utdiv(a(x,t)|u|p(x,t)2u)Ni=1bi(x,t)Diu=f(u,x,t), (x,t)QT,

    and

    vtdiv(|v|α|v|p2v)Ni=1gi(x,t,v)vxi=d(x,t,v), (x,t)QT,

    where α>0 is a constant, p(x,t)>1 is a continuous function.

    This work is partially supported by Natural Science Foundation of Fujian Province (no: 2021J011193) and supported by NSFC-52171308, China.

    The authors declare no conflict of interest.

    In this appendix, we give a simple discussion of the trace. Without loss the generality, we assume that f(x)0. It is well-known that, C0(Ω) is dense in W1,p0(Ω), and so the trace of f(x)W1,p(Ω) on the boundary Ω

    f(x)=0, xΩ

    is defined as the limit of a sequence fε(x)C0(Ω) as

    0=limε0fε(x)|xΩ. (A.1)

    As above in this paper, we denote Ωλ={xΩ:d(x)>λ} for a small enough positive number λ, and denote by B the closure of the set C0(QT) with respect to the norm

    uB=QTa(x)(|u(x,t)|p+|u(x,t)|p)dxdt, uB.

    Yin-Wang [28] defined the trace of uB, u(x,t)=0 on Σ2 as

    esslimλ0{xΩλ:Ni=1bi(x)ni(x)<0}u2Ni=1bi(x)ni(x)dσ=0. (A.2)

    One can see that if a(x)=d(x), then dxi=ni with that n={ni} is the inner normal. Then the Definition (A.2) is just the same as that of (1.20). The reminder is to show that when a(x) satisfies (1.10), a(x)d(x), the Definition (A.2) is equivalent to that of (1.20).

    At first, since a(x) satisfies (1.10) and Ωa(x)|u|pdx<, we know

    Ωλ|u|pdx<

    and uBV(Ωλ), the traces of

    u2Ni=1bi(x)ni(x), xΩλ

    and

    u2Ni=1bi(x)axi(x), xΩλ

    can be well defined in the sense of (A.1). Thus, the definition of (1.20) or the definition of (A.2) itself has a explicit meaning. In fact, recalling the BV function space BV(Ω), i.e. |fxi| is a regular measure, and

    BV(Ω)={f(x):Ω|fxi|<c,i=1,2,,N}.

    then BV(Ω) is a Banach space under the norm

    fBV=fL1+Ω|Df|.

    By the trace f+(y) of f(x)BV(Ω) on the boundary Ω is defined as the limit of f(x) along the normal (Lemma 1 of [26]). Thus, if we denote that

    Dλ={xΩ:d(x)=dist(x,Ω)>λ},

    then for a f(x)BV(Ω)L(Ω), we have

    limλ01λΩDλf(x)dx=Ωf+(x)dΣ. (A.3)

    Secondly, for a smooth bounded domain Ω, there is a finite open cover such that

    Ω=αBα,  ΩαΣα,

    where Σα=BαΩ. If Σα, after using the flattening technique, we may assume that

    B1={xRN:|x|<1},  BαΩ={xRN:|x|<1,xN>0}=B+(1),
    Σα={xRN:|x|<1,xN=0},

    and

    d(x,Ω)=xN, xBαΩ,

    and

    d(x,Ω)=x2, xBα.

    Let a(x)C1(¯Ω) satisfy

    a(x)>0, xΩ, a(x)=0, xΩ.

    Since a(x)>0 when xB+1, and there is a small enough δ0 such that when 0<x2<δ

    a(x)x2c(δ0)>0, x{B+1:0<x2<δ0}. (A.4)

    For small enough λ<δ0, we set

    Exλ={xB+1:a(x)<λ},

    and denote the local coordinate representation Bα and Σα as above. By the coordinate transformation

    y1=x1,y2=x2,,yN1=xN1,yN=a(x) (A.5)

    the domain Exλ is transformed to Eyλ=Σα×(0,λ).

    In Eyλ, it is clear of that d(y,Σα)=yN. Since fBV(Exλ) or fBV(Eyλ), as (A.3), we have

    Ωf+(y)dΣ=limλ01λEyλf(y)dy=limλ01λExλf(x)axNdx. (A.6)

    When f is nonnegative and f|Ω=0, then the equality (A.6) can be re-calculated as

    0=Ωf+(y)dΣ=limλ01λEyλf(y)dy=limλ01λExλf(x)axNdxM(axN)limλ01λExλf(x)dx=M(axN)Σαf+(x)dΣ=0, (A.7)

    where M(axN)=supxΣαaxN. In other words, when fBV(Ω) is nonnegative and f|Ω=0, (A.7) implies that

    limλ01λExλf(x)dx=Σαf+(x)dΣ=0. (A.8)

    Thirdly, by (A.4), the coordinate transformation (A.5) has a inverse transformation

    x1=y1,x2=y2,,xN1=yN1,xN=a1(y). (A.9)

    Recalling the definition of (1.20), i.e.,

    esssuplimε01ε(ΩεΩ2ε)Ω1tu2Ni=1bi(x,t)axi(x)dx=0,

    without loss the generality, under the inverse transformation (A.9), we may assume that the domain (ΩεΩ2ε)Ω1t is transformed to a domain DεyΩ1t

    Then, similar as the discussion of (A.7) and (A.8), the definition of (1.20) is equivalent to that

    esssuplimε01εDεyΩ1tu2Ni=1bi(y,t)nyidy=0.

    From these point, one can see that the definition of (1.20) is just a reasonable version the definition of (A.2), which is the generalized definition of the trace of uB with degeneracy on the boundary. Actually, such a local analysis of the definition of the trace in BV(Ω) has appeared in our previous work [33]. Since [33] is unpublished till now, we re-give the details here.



    [1] G. Akagi, Local existence of solutions to some degenerate parabolic equation associated with the p-Laplacian, J. Differ. Equ., 241 (2007), 359–385. https://doi.org/10.1016/j.jde.2007.05.009 doi: 10.1016/j.jde.2007.05.009
    [2] G. Rosen, The mathematical theory of diffusion and reaction in permeable catalysts, Bull. Math. Biol., 38 (1976), 95–96. https://doi.org/10.1007/BF02459545 doi: 10.1007/BF02459545
    [3] S. J. Chapman, G. Ridhardson, Vortex pining by inhomogeneities in type-Ⅱ superconductors, Physica D, 108 (1997), 397–407. https://doi.org/10.1016/S0167-2789(97)00053-5 doi: 10.1016/S0167-2789(97)00053-5
    [4] E. DiBenedetto, Degenerate parabolic equations, New York: Spring-Verlag, 1993. https://doi.org/10.1007/978-1-4612-0895-2
    [5] R. Dautray, J. L. Lions, Mathematical analysis and numerical methods for science and technology, Volume Ⅰ: Physical origins and classical methods, Berlin: Springer-Verlag, 1985.
    [6] J. Droniou, R. Eymard, K. S. Talbot, Convergence in C([0,T];L2(Ω)) of weak solutions to perturbed doubly degenerate parabolic equations, J. Differ. Equ., 260 (2016), 7821–7860. https://doi.org/10.1016/j.jde.2016.02.004 doi: 10.1016/j.jde.2016.02.004
    [7] D. Eidus, S. Kamin, The fifiltration equation in a class of functions decreasing at infifinity, Proc. Amer. Math. Soc., 120 (1994), 825–830. https://doi.org/10.1090/S0002-9939-1994-1169025-2 doi: 10.1090/S0002-9939-1994-1169025-2
    [8] Y. Gaididei, N. Lazarides, N.Flytzanis, Fluxons in a superlattice of Josephson junctions: Dynamics and radiation, J. Phys. A: Math. Gen., 36 (2003), 2423–2441. https://doi.org/10.1088/0305-4470/36/10/304 doi: 10.1088/0305-4470/36/10/304
    [9] Y. Gaididei, N. Lazarides, N. Flytzanis, Static flfluxons in a superlattice of Josephson junctions, J. Phys. A: Math. Gen., 35 (2002), 10409–10427. https://doi.org/10.1088/0305-4470/35/48/313 doi: 10.1088/0305-4470/35/48/313
    [10] R. Gianni, A. Tedeev, V. Vespri, Asymptotic expansion of solutions to the Cauchy problem for doubly degenerate parabolic c equations with measurable coefficients, Nonlinear Anal., 138 (2016), 111–126. https://doi.org/10.1016/j.na.2015.09.006 doi: 10.1016/j.na.2015.09.006
    [11] L. Gu, Second order parabolic partial differential equations (in Chinese), Xiamen University Press, Xiamen, 2002.
    [12] J. K. Hale, C. Rocha, Interaction of diffusion and boundary conditions, Nonlinear Anal.-Theor., 11 (1987), 633–649. https://doi.org/10.1016/0362-546X(87)90078-2 doi: 10.1016/0362-546X(87)90078-2
    [13] J. K. Hale, G. Raugel, Reaction-diffusion equation on thin domains, J. Math. Pures Appl., 71 (1992), 33–95.
    [14] S. Jimbo, Y. Morita, Stable vortex solutions to the Ginzburg-Landau equation with a variable coefficient in a disk, J. Differ. Equ., 155 (1999), 153–176. https://doi.org/10.1006/jdeq.1998.3580 doi: 10.1006/jdeq.1998.3580
    [15] H. Y. Jian, B. H. Song, Vortex dynamics of Ginzburg-Landau equations in inhomogeneous superconductors, J. Differ. Equ., 170 (2001), 123–141. https://doi.org/10.1006/jdeq.2000.3822 doi: 10.1006/jdeq.2000.3822
    [16] S. Kamin, P. Rosenau, Propagation of thermal waves in an inhomogeneous medium, Commun. Pure Appl. Math., 34 (1981), 831–852. https://doi.org/10.1002/cpa.3160340605 doi: 10.1002/cpa.3160340605
    [17] S. Kamin, P. Rosenau, Nonlinear thermal evolution in an inhomogeneous medium, J. Math. Phys., 23 (1982), 1385. https://doi.org/10.1063/1.525506 doi: 10.1063/1.525506
    [18] N. I. Karachalios, N. B. Zographopoulos, On the dynamics of a degenerate parabolic equation: global bifurcation of stationary states and convergence, Cala. Var. Partial Differ. Equ., 25 (2006), 361–393. https://doi.org/10.1007/s00526-005-0347-4 doi: 10.1007/s00526-005-0347-4
    [19] K. Lee, A. Petrosyan, J. L. Vazquez, Large time geometric properties of solutions of the evolution pLaplacian equation, J. Differ. Equ., 229 (2006), 389–411. https://doi.org/10.1016/j.jde.2005.07.028 doi: 10.1016/j.jde.2005.07.028
    [20] G. F. Lu, Nonlinear degenerate parabolic equations in infiltration through a porous medium, Commun. Nonlinear Sci., 3 (1998), 97–100. https://doi.org/10.1016/S1007-5704(98)90071-5 doi: 10.1016/S1007-5704(98)90071-5
    [21] J. D. Murray, Mathematical biology Ⅱ: Spatial models and biomedical applications, New York: Springer-Verlag, 2003. https://doi.org/10.1007/b98869
    [22] M. B. Riaz, M. A. Imran, K. Shabbir, Analytic solutions of Oldroyd-B fluid with fractional derivatives in a circular duct that applies a constant couple, Alex. Eng. J., 55 (2016), 3267–3275. https://doi.org/10.1016/j.aej.2016.07.032 doi: 10.1016/j.aej.2016.07.032
    [23] H. F. Shang, J. X. Cheng, Cauchy problem for doubly degenerate parabolic equation with gradient source, Nonlinear Anal.-Theor, 113 (2015), 323–338. https://doi.org/10.1016/j.na.2014.10.006 doi: 10.1016/j.na.2014.10.006
    [24] C. Sulem, P. L. Sulem, The nonlinear Schrödinger equation: Self-focusing and wave collapse, New York: Springer, 1999.
    [25] F. Z. Wang, M. I. Asjad, M. Zahid, A. Iqbal, H. Ahmad, M. D. Alsulami, Unsteady thermal transport flow of Casson nanofluids with generalized Mittage-Lefflfler kernel of Prabhakar's type, J. Mater. Res. Technol., 14 (2021), 1292–1300. https://doi.org/10.1016/j.jmrt.2021.07.029 doi: 10.1016/j.jmrt.2021.07.029
    [26] Z. Q. Wu, J. N. Zhao, The first boundary value problem for quasilinear degenerate parabolic equations of second order in several variables, Chinese Anal. Math., 1 (1983), 319–358.
    [27] Z. Q. Wu, J. X. Zhao, J. Yun, F. H. Li, Nonlinear diffusion equations, World Scientific Publishing Company, 2001. https://doi.org/10.1142/4782
    [28] J. X. Yin, C. P. Wang, Evolutionary weighted pLaplacian with boundary degeneracy, J. Differ. Equ., 237 (2007), 421–445. https://doi.org/10.1016/j.jde.2007.03.012 doi: 10.1016/j.jde.2007.03.012
    [29] H. S. Zhan, Infiltration equation with degeneracy on the boundary, Acta. Appl. Math., 153 (2018), 147–161. https://doi.org/10.1007/s10440-017-0124-3 doi: 10.1007/s10440-017-0124-3
    [30] H. S. Zhan, The uniqueness of the solution to the diffusion equation with a damping term, Appl. Anal., 9 (2019), 1333–1346. https://doi.org/10.1080/00036811.2017.1422725 doi: 10.1080/00036811.2017.1422725
    [31] H. S. Zhan, Z. S. Feng, Degenerate non-Newtonian fluid equation on the half space, Dyn. Partial Differ. Equ., 15 (2018), 215–233.
    [32] H. S. Zhan, Z. S. Feng, Optimal partial boundary condition for degenerate parabolic equations, J. Differ. Equ., 284 (2021), 156–182. https://doi.org/10.1016/j.jde.2021.02.053 doi: 10.1016/j.jde.2021.02.053
    [33] H. S. Zhan, Z. S. Feng, The local stablity of a Kolmogorov equation in financial mathematics, 2022. Preprint.
    [34] J. N. Zhao, Existence and nonexistence of solutions for utdiv(|u|p2u)=f(u,u,x,t), J. Math. Anal. Appl., 172 (1993), 130–146. https://doi.org/10.1006/jmaa.1993.1012 doi: 10.1006/jmaa.1993.1012
  • Reader Comments
  • © 2022 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(1662) PDF downloads(98) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog