
The system of equations describing motion of compressible viscoelastic fluids is considered in a one dimensional half space under the outflow boundary condition. We investigate the existence and stability of stationary solutions. It is shown that the stationary solution exists for large Mach number and small number of propagation speed of elastic wave. We next show that the stationary solution is asymptotically stable, provided that the initial perturbation is sufficiently small.
Citation: Yusuke Ishigaki, Yoshihiro Ueda. Stability of stationary solutions to outflow problem for compressible viscoelastic system in one dimensional half space[J]. AIMS Mathematics, 2024, 9(11): 33215-33253. doi: 10.3934/math.20241585
[1] | Tohru Nakamura . Asymptotic stability of degenerate stationary solution to a system of viscousconservation laws in half line. AIMS Mathematics, 2018, 3(1): 35-43. doi: 10.3934/Math.2018.1.35 |
[2] | Yingying Chen, Lan Huang, Jianwei Yang . Large time behavior of the Euler-Poisson system coupled to a magnetic field. AIMS Mathematics, 2023, 8(5): 11460-11479. doi: 10.3934/math.2023580 |
[3] | Quan Zhou, Yang Liu, Dong Yang . Global attractors for a class of viscoelastic plate equations with past history. AIMS Mathematics, 2024, 9(9): 24887-24907. doi: 10.3934/math.20241212 |
[4] | Caifeng Liu . Linear Rayleigh-Taylor instability for compressible viscoelastic fluids. AIMS Mathematics, 2023, 8(7): 14894-14918. doi: 10.3934/math.2023761 |
[5] | Ruonan Liu, Tomás Caraballo . Random dynamics for a stochastic nonlocal reaction-diffusion equation with an energy functional. AIMS Mathematics, 2024, 9(4): 8020-8042. doi: 10.3934/math.2024390 |
[6] | Banan Al-Homidan, Nasser-eddine Tatar . Fractional Timoshenko beam with a viscoelastically damped rotational component. AIMS Mathematics, 2023, 8(10): 24632-24662. doi: 10.3934/math.20231256 |
[7] | Soh E. Mukiawa, Tijani A. Apalara, Salim A. Messaoudi . Stability rate of a thermoelastic laminated beam: Case of equal-wave speed and nonequal-wave speed of propagation. AIMS Mathematics, 2021, 6(1): 333-361. doi: 10.3934/math.2021021 |
[8] | Abdelbaki Choucha, Salah Boulaaras, Asma Alharbi . Global existence and asymptotic behavior for a viscoelastic Kirchhoff equation with a logarithmic nonlinearity, distributed delay and Balakrishnan-Taylor damping terms. AIMS Mathematics, 2022, 7(3): 4517-4539. doi: 10.3934/math.2022252 |
[9] | Abdelbaki Choucha, Sofian Abuelbacher Adam Saad, Rashid Jan, Salah Boulaaras . Decay rate of the solutions to the Lord Shulman thermoelastic Timoshenko model. AIMS Mathematics, 2023, 8(7): 17246-17258. doi: 10.3934/math.2023881 |
[10] | Kaile Chen, Yunyun Liang, Nengqiu Zhang . Global existence of strong solutions to compressible Navier-Stokes-Korteweg equations with external potential force. AIMS Mathematics, 2023, 8(11): 27712-27724. doi: 10.3934/math.20231418 |
The system of equations describing motion of compressible viscoelastic fluids is considered in a one dimensional half space under the outflow boundary condition. We investigate the existence and stability of stationary solutions. It is shown that the stationary solution exists for large Mach number and small number of propagation speed of elastic wave. We next show that the stationary solution is asymptotically stable, provided that the initial perturbation is sufficiently small.
This paper studies the compressible viscoelastic system in the one-dimensional half space R+=(0,∞):
ρt+(ρv)x=0, | (1.1) |
[1.0ex](ρv)t+(ρv2)x−νvxx+P(ρ)x=β2(ρF2)x, | (1.2) |
[1.0ex]Ft+vFx=vxF. | (1.3) |
Here ρ=ρ(t,x), v=v(t,x), and F=F(t,x) are the unknown density, velocity field, and deformation tensor, respectively, at time t≥0 and position x∈R+; P(ρ) stands for the pressure assumed to be a smooth function of ρ satisfying P′(ρ)>0 and P′′(ρ)>0 for ρ>0; ν>0 is the viscosity coefficient; β>0 is the strength of the elasticity. In particular, if we set β=0, the systems (1.1) and (1.2) become the usual compressible Navier-Stokes equation.
We impose the initial condition and boundary conditions at x=∞ and x=0:
(ρ,v,F)|t=0=(ρ0,v0,F0),infx∈R+ρ0(x)>0, | (1.4) |
limx→∞(ρ,v,F)=(ρ+,v+,F+), | (1.5) |
v(t,0)=vb. | (1.6) |
Here, the end states ρ+,v+ and F+ are given constants with ρ+>0, and vb is a given constant assumed to be vb<0 for considering the situation that the fluid flows out from the boundary x=0. Throughout this paper, we consider the initial boundary problems (1.1)–(1.6), called the outflow problem. The aim of this paper is to show the existence and asymptotic stability of stationary solutions for the outflow problems (1.1)–(1.6) and clarify the interaction between the effect of the elastic force β2(ρF2)x and outflow boundary condition (1.6).
We first discuss the existence and properties for the stationary solution (˜ρ,˜v,˜F)(x), called the boundary layer solution, solving the system:
(˜ρ˜v)x=0, | (1.7) |
(˜ρ˜v2)x−ν˜vxx+P(˜ρ)x=β2(˜ρ˜F2)x, | (1.8) |
˜v˜Fx=˜vx˜F, | (1.9) |
with the conditions:
limx→∞(˜ρ,˜v,˜F)=(ρ+,v+,F+),infx∈R+˜ρ(x)>0, | (1.10) |
˜v(0)=vb. | (1.11) |
It is shown that the solution of the problems (1.7)–(1.11) exists uniquely if and only if Mβ≥1 and vb<v∗ hold. Here, Mβ is the modified Mach number given by Mβ:=|v+|/√c2++β2F2+, where c+:=√P′(ρ+) stands for sound speed; v∗ is a certain negative constant determined in Section 2. In addition, the solution satisfies the following estimate:
|(˜ρ−ρ+,˜v−v+,˜F−F+)(x)|≤{Cδe−cx,Mβ>1,Cδ(1+δx)−1,Mβ=1, |
where δ denotes δ:=|vb−v+|. We call (˜ρ,˜v,˜F)(x) the non-degenerate stationary solution tending to the end state exponentially when Mβ>1, while we say (˜ρ,˜v,˜F)(x) by the degenerate stationary solution converging to the end state algebraically when Mβ=1. We also note that if we take β large so that Mβ<1 under the fixed end state, then the stationary solution does not exist. This means that the stationary outflow does not occur due to the recoiling effect of strong elastic force.
We next establish the asymptotic stability of the stationary solution under the small initial perturbation (ρ0,v0,F0)−(˜ρ,˜v,˜F), provided that δ is sufficiently small. This follows from the local-in-time solvability of (1.1)–(1.6) and the a priori estimates for the perturbation in H1(R+). Since the systems (1.1)–(1.3) is classified by a quasilinear parabolic-hyperbolic system, the local-in-time solvability is shown by the iteration method and theory of weak solutions to linear transport equations and parabolic equations, inspired by Kagei and Kawashima's paper [11].
To derive the a priori estimate, we need to deal with the term β2(ρF2−˜ρ˜F2)x when we consider the problem for a usual perturbation (ρ,v,F)−(˜ρ,˜v,˜F). However, it seems to be difficult to control ρF−ρ+F+ appearing from this term. To overcome this difficulty, we assume the following condition for (ρ0,F0):
ρ0F0=ρ+F+ | (1.12) |
for x∈R+. We then see from (1.1), (1.3), (1.6) and (1.12) that ρF=ρ+F+ holds for t≥0 and x≥0. This constraint is a one-dimensional version of the following equality
ρdetF=ρ+detF+,t≥0,x=⊤(x1,x2,…,xn)∈R+×Rn−1, |
which is equivalent to the conservation law of mass in the Lagrange coordinate. Here, n≥2; F+ stands for given n×n matrix-valued constant; F(t,x)=(Fj,k(t,x))1≤j,k≤n denotes an n×n matrix-valued deformation tensor, respectively.
Therefore, rewriting F=ρ+F+ρ−1, the problems (1.1)–(1.6) for (ρ,v,F) is reduced to the system for (ρ,v):
ρt+(ρv)x=0, | (1.13) |
[1.0ex](ρv)t+(ρv2)x−νvxx+P(ρ)x=β2(ρ+F+)2(1ρ)x | (1.14) |
with the initial condition and boundary conditions at x=∞ and x=0:
(ρ,v)|t=0=(ρ0,v0),infx∈R+ρ0(x)>0 | (1.15) |
limx→∞(ρ,v)=(ρ+,v+), | (1.16) |
v(t,0)=vb. | (1.17) |
We then prepare the reduced perturbation (ϕ,ψ):=(ρ−˜ρ,v−˜v) and carry out its estimate as two steps. For (ϕ,ψ) itself, based on the idea of Kawashima, Nishibata and Zhu's paper [13] and the fact that P(ρ)−β2(ρ+F+)2/ρ monotonically increases for ρ>0, we construct a suitable energy form equivalent to |(ϕ,ψ)|2, and use its equation and the properties of the stationary solution. The effect of the term β2(ρ+F+)2(1/ρ)x is mainly involved to the proof of this estimate. Indeed, the convexity of P(ρ)−β2(ρ+F+)2/ρ around ρ=ρ+ requires for the degenerate case Mβ=1. In order to get this convexity, we need to add the condition for the Mach number 1<M+<√ρ+(P′′(ρ+)/2P′(ρ+))+1, where M+:=|v+|/c+. For the 1-st order spatial derivative of (ϕ,ψ), we utilize the structure of (1.13) and (1.14) and monotonicity of P(ρ)−β2(ρ+F+)2/ρ for ρ>0.
We remark the reason why we only assume the regularity condition (ρ0−˜ρ,v0−˜v,F0−˜F)∈H1(R+) in contrast of the case β=0 in [13]. In the argument in [13], we need the equation
∫∞0v(t,x)ϕxx(t,x)ϕx(t,x)dx=−12vb|ϕx(t,0)|2−12∫∞0vx(t,x)|ϕx(t,x)|2dx | (1.18) |
to construct the a priori estimate for ϕx. This means that we need to take care of ϕxx(t,x) and |ϕx(t,0)|. The equation (1.18) is obtained by the integration by parts with (1.6) and makes sense under ψ∈C([0,T];H1(R+)) and ϕ∈C([0,T];Hk(R+)) for T>0 and k≥2 because Hk(R+)⊂C1([0,∞)) holds for k≥2. In our setting, the equation (1.18) is not valid since we restrict the situation that ϕ∈C([0,T];H1(R+)) which is not sufficient to define ϕxx(t,x) and ϕx(t,0). To deal with this difficulty, Kawashima, Nishibata and Zhu in [13] additionally assumed that the initial perturbation belongs to the Hölder space for guaranteeing ϕx(t,0), and then applied the method of difference quotient to avoid appearing higher derivatives such as ϕxx(t,x). Later, Kagei and Kawashima introduced the weak forms of the parabolic equation and first-order transport equation to show the local-in-time existence of the quasilinear parabolic-hyperbolic system in [11], and then used their theory to obtain the estimates for the higher order derivatives without the Hölder regularity of initial perturbation in [12]. Therefore, inspired the idea of [12], we can conclude that it is enough to suppose that initial perturbation is small only in H1(R+) to show the asymptotic stability result.
Known results. The systems (1.1)–(1.3) describing the motion of compressible viscoelastic fluid is governed in the macroscopic scale by the variational modeling. Indeed the second equation (1.2) is treated as the conservation law of momentum following from the energy dissipation law with the free energy induced by elastic solids. Here, the free energy is taken as the derivative of β2ρW′(F)F=β2ρF2, where W(F)=β2F2/2 denotes linear isentropic elasticity. The other equations (1.1) and (1.3) are the kinematic assumptions for ρ and F. For more physical detail, we refer to [3,4,5,17,25]. Starting with Sideris and Thomases [25], the mathematical analysis of the systems (1.1)–(1.3) has been progressed mainly on the stability of the trivial motionless state. In fact, its stability is investigated by [1,7,8,16,24] in the three dimensional whole space and is studied by [2,9,23] in the three-dimensional bounded domain with smooth boundary case under the Dirichlet boundary condition. For the stability of non-trivial flows with non-zero velocity, the dynamics of solutions around them becomes more complicated than the trivial motionless case since the advection terms in (1.1)–(1.3) produce the additional hyperbolic aspect. Therefore, comparing to the the trivial motionless case, there are few results on their stability as follows. Ishigaki [10] and Haruki and Ishigaki [6] investigated the stability of parallel flows in the three dimensional layer, and Morando, Trakhinin and Trebeschi [19] and Trakhinin [26] studied the stability of shock waves in the two-dimensional whole space without the viscous effect.
We next review the mathematical analysis of outflow problem. For the case β=0, it is natural to expect that the behavior of the solution in the half space is closely related to the boundary and the end states vb, v+ and ρ+. Then, Matsumura [18] suggested that the long time asymptotic states will be composed of the rarefaction wave, the viscous shock wave and the boundary layer solution. The stability of stationary solution is related to the case that the asymptotic state is given by the boundary layer only. Kawashima, Nishibata and Zhu [13] characterized the existence of stationary solutions by determining suitable conditions for vb, v+ and ρ+, and then showed its asymptotic stability by the energy method in the Eulerian coordinate. Nakamura, Nishibata and Yuge [21] established the convergence rate toward the stationary solutions as t→∞ under the small initial perturbation belonging to the weighted L2 Sobolev space. Later that, Nakamura, Ueda and Kawashima [22] refined the convergence rate toward the degenerate stationary solution discussed in [13,21]. Furthermore, Kawashima and Kagei [11], and Nakamura and Nishibata [20] extended these stability results to the multidimensional case. On the other hand, if vb, v+ and ρ+ do not satisfy the conditions for the existence of stationary solutions, the asymptotic state of the solution becomes different from the boundary layer solution. For the details, we refer to [14,15,27] when the rarefaction wave involves its time asymptotic state. Concerning the case β>0, as far as the authors know, it remains open.
Outline of this paper. This paper is organized as four sections and one appendix. In Section 2, notations of several function spaces and lemmata are explained. In Section 3, we give the detailed necessary conditions for the existence of a stationary solution and provide its properties. In Section 4, the detail of the main result in this paper is stated. In Section 5, we show the asymptotic stability of the stationary solutions. In Appendix A, we prove the local-in-time existence around the stationary solutions.
In this section, we introduce several function spaces and important lemmata.
For 1≤p≤∞, the symbol Lp stands for the usual Lebesgue space on R+, and its norm is denoted by ‖⋅‖Lp. For a non-negative integer m≥0, we define Hm as the m-th order L2 Sobolev space on R+, and its norm is denoted by ‖⋅‖Hm. For simplicity, we write Lp=Lp×Lp×Lp (resp., Hm=Hm×Hm×Hm). The symbol C10(R+) denotes the set of all C1(R+) functions whose supports are compact in R+. We call H10(R+) the completion of C10(R+) in H1(R+). For T>0, we define C10((0,T)×R+) (resp., C10([0,T)×R+)) as the set of all C1((0,T)×R+) (resp., C1([0,T)×R+)) functions whose supports are compact in (0,T)×R+ (resp., [0,T)×R+).
For 0≤a<b≤∞, a Banach space X endowed with norm ‖⋅‖X and a non-negative integer k, we define Ck([a,b];X) that
Ck([a,b];X):={f:[a,b]→X;fis a C^k function in [a, b] satisfying‖f‖Ck([a,b];X)<∞}, |
where
‖f‖Ck([a,b];X):=k∑l=0supt∈[a,b]‖∂ltf(t)‖X. |
Here, we identify [a,∞]:=[a,∞) in the case b=∞. For simplicity, we write C([a,b];X):=C0([a,b];X).
For 0≤a<b≤∞, a Banach space X endowed with norm ‖⋅‖X and a non-negative integer k, L2(a,b;X) and Hk(a,b;X) denote
L2(a,b;X):={f:[a,b]→X;fis a measurable function in [a, b] satisfying‖f‖L2(a,b;X)<∞}, |
where
‖f‖L2(a,b;X):=(∫ba‖f(t)‖2Xdt)1/2, |
and
Hk(a,b;X):={f∈L2(a,b;X);fis a k -th times weakly differentiable function in (a, b) satisfying‖f‖Hk(a,b;X)<∞}, |
where
‖f‖Hk(a,b;X):=(k∑l=0∫ba‖∂ltf(t)‖2Xdτ)1/2. |
For a real number α, we set [α] as its integer part. Throughout this paper, we simply regard the letters c, C, ˜c and ˜C as positive various constants.
We next state several lemmata to establish our asymptotic stability result. We first introduce the well-known Gagliardo–Nirenberg inequality to obtain a priori estimate in H1(R+).
Lemma 2.1. Let f∈H1(R+). Then, f∈C([0,∞)) and it satisfies
‖f‖L∞≤√2‖f‖1/2L2‖fx‖1/2L2. |
To show the asymptotic stability of the stationary solution, the following lemma plays a role.
Lemma 2.2. [12, Lemma 4.5.] Let T>0 be an arbitrary number, and f=f(t,x) be f∈C([0,T];H10(R+))∩L2(0,T;H2(R+))∩H1(0,T;L2(R+)). Then f satisfies
‖fx(t2)‖2L2≤C(‖fx(t1)‖2L2+∫t2t1‖ft(τ)‖L2‖fx(τ)‖H1dτ) |
for 0≤t1≤t2≤T.
We give definitions of weak solutions to the linear transport equation and the parabolic equation for studying the local-in-time existence of the unique solution to (1.1)–(1.6) around the stationary solution.
For 0≤a<b≤∞ and k=0,1, we define the function spaces Xk(a,b) and Yk(a,b) as
Xk(a,b):=C([a,b];Hk−l(R+)),Yk(a,b):=C([a,b];˜Hk(R+))∩k⋂l=0Hl(a,b;˜Hk+1−2l(R+)), |
respectively. Here, ˜Hk(R+) is given by ˜Hk(R+):=Hk(R+)∩H10(R+) when k=1,2, and ˜H0(R+):=L2(R+). We set the corresponding norms such that
‖f‖Xk(a,b):=k∑l=0‖f‖Cl([a,b];Hk−l(R+)),‖g‖Yk(a,b):=(‖g‖2C([a,b];Hk(R+))+k∑l=0‖g‖2Hl(a,b;Hk+1−2l(R+)))1/2. |
We introduce the function space Zk(a,b) as
Zk(a,b):=Xk(a,b)×Yk(a,b)×Xk(a,b), |
and the norms of u=(ϕ,ψ,ζ) in Zk(a,b) are given by
‖u‖Zk(a,b):=(‖ϕ‖2Xk(a,b)+‖ψ‖2Yk(a,b)+‖ζ‖2Xk(a,b))1/2. |
For simplicity, we write Xk(T):=Xk(0,T), Yk(T):=Yk(0,T) and Zk(T):=Zk(0,T) for T>0. We also set
XkM(T):={f∈Xk(T);‖f‖Xk(T)≤M},YkM(T):={g∈Yk(T);‖g‖Yk(T)≤M},ZkM(T):={u∈Zk(T);‖u‖Zk(T)≤M}. |
for M>0.
We define weak solutions to the linear transport equation and the parabolic equation associated with the stationary solution of (1.1)–(1.6). The stationary solution (˜ρ,˜v,˜F) is the smooth solution to (1.7)–(1.11), and the existence of solutions will be discussed in Section 3. Using this stationary solution, we introduce functions a(ψ):=˜v+ψ and b(ϕ):=˜ρ+ϕ, as well as a linear operator B:H10(R+)→H−1(R+) defined by ⟨Bψ,φ⟩:=ν(ψx,φx)L2 for ψ,φ∈H10(R+). Here, H−1(R+) is the dual space of H10(R+), ⟨⋅,⋅⟩ that stands for the pairing between H−1(R+) and H10(R+), and (⋅,⋅)L2 denotes the usual L2(R+)-inner product. We also note that B is identified as a usual expression Bψ=−νψxx if ψ∈˜H2(R+) holds. Then, we define weak solutions as follows.
Definition 2.3. Let T>0 and let ˜ψ=˜ψ(t,x) be a given function. For given ϕ0∈L2(R+) and f∈L2(0,T;L2(R+)), we call ϕ a weak solution of the initial value problem
ϕt+a(˜ψ)ϕx=f,ϕ|t=0=ϕ0,ϕ|x=∞=0 | (2.1) |
if ϕ belongs to X0(T) and satisfies the weak form
−∫T0(ϕ,φt+(a(˜ψ)φ)x)L2dt=(ϕ0,φ(0))L2+∫T0(f,φ)L2dt | (2.2) |
for all φ∈C10([0,T)×R+).
Definition 2.4. Let T>0 and let ˜ϕ=˜ϕ(t,x) be a given function. For given ψ0∈L2(R+) and g∈L2(0,T;H−1(R+)), we call ψ a weak solution of the initial-boundary value problem
b(˜ϕ)ψt+Bψ=g,ψ|t=0=ψ0,ψ|x=0=ψ|x=∞=0 | (2.3) |
if ψ belongs to Y0(T) and satisfies the weak form
−∫T0(b(˜ϕ)ψ,φ)L2h′dt−∫T0⟨˜ϕtψ,φ⟩hdt+∫T0⟨Bψ,φ⟩hdt=(b(˜ϕ(0,⋅))ψ0,φ)L2h(0)+∫T0⟨g,φ⟩hdt | (2.4) |
for all φ∈H10(R+) and h∈C10([0,T)).
The existence, regularity and estimates of weak solutions to (2.1) and (2.3) are stated as the following lemma.
Lemma 2.5. Let T>0 be a positive constant. Assume that ˜ψ=˜ψ(t,x) satisfies ˜ψ∈Y1(T). Then, for any ϕ0∈Hk(R+) and f∈L2(0,T;Hk(R+)) with k=0,1, there exists a unique weak solution ϕ∈Xk(T) to satisfying
‖ϕ(t2)‖2Hk≤‖ϕ(t1)‖2Hk+k∑l=0∫t2t1{(a(˜ψ)x(τ),|∂lxϕ(τ)|2)L2+2l((a(˜ψ)xϕx)(τ),ϕx(τ))L2}dτ+2k∑l=0∫t2t1(∂lxf(τ),∂lxϕf(τ))L2dτ | (2.5) |
for all 0≤t1≤t2≤T.
In addition, in the case k=1, ϕ belongs to C1([0,T];L2(R+)) and is controlled as
maxt∈[0,T]‖ϕ(t)−ϕ0(˜y(0;t,⋅)))‖L∞≤T12(∫T0‖f(τ)‖2H1dτ)12. | (2.6) |
Here ˜y=˜y(τ;t,x)∈R+ is a unique solution of the problem
d˜ydτ(τ;t,x)=a(˜ψ(τ,˜y(τ;t,x))),0≤τ≤t≤T,˜y(t;t,x)=x. |
Lemma 2.6. Let T, M and m be positive constants. Assume that ˜ϕ=˜ϕ(t,x) satisfies
˜ϕ∈X1M(T),˜ϕ∈C1([0,T];L2(R+)),inf(t,x)∈[0,T]×R+˜ϕ(t,x)≥(m−1)infx∈R+˜ρ(x). |
Then, for any ψ0∈˜Hk(R+) and g∈L2(0,T;Hk−1(R+)) with k=0,1, there exists a unique weak solution ψ∈Yk(T) to (2.3) satisfying
‖ψ(t)‖2Hk+C1(δ,M,m)∫t0(‖ψ(τ)‖2Hk+1+k‖ψt(τ)‖2L2)dτ≤‖ψ0‖2Hk+C2(δ,M,m)∫t0(‖g(τ)‖2Hk−1+‖ψ(τ)‖2L2)dτ | (2.7) |
for all 0≤t≤T. Here C1(δ,M,m) is a positive constant decreasing in δ, M and increasing in m, and C2(δ,M,m) is a positive constant increasing in δ, M and decreasing in m.
Lemmas 2.5 and 2.6 are proved by the same method as [11], so we omit the details.
Remark 2.7 (ⅰ) If the same assumptions as in Lemma 2.5 with k=1 are satisfied, then (2.2) for any φ∈C10([0,T)×R+) becomes equivalent to the first equation of (2.1) in C([0,T];L2(R+)).
(ⅱ) If the same assumptions as in Lemma 2.6 with k=1 hold, then (2.4) for any φ∈H10(R+) and h∈C10([0,T)) becomes equivalent to the first equation of (2.3) in L2(0,T;L2(R+)).
In this section, we discuss the existence and the convergence rate of the stationary solution (˜ρ,˜v,˜F) satisfying the following stationary problems (1.7)–(1.11). To solve this problem, we analyze the properties of the solutions and derive the reduced problem. Integrating (1.7) over [x,∞) for x>0, we have
˜ρ=ρ+v+˜v. | (3.1) |
Letting x→0 in (3.1), we obtain
v+=˜ρ(0)ρ+vb. |
Therefore, vb<0 and (1.10) give the fact that v+<0 is necessary for the existence of the stationary solution to the problems (1.7)–(1.11). Furthermore, because of (1.10) and (3.1), we regard as ˜v<0.
On the other hand, (1.9) gives (˜F/˜v)x=0. Thus, integrating the resultant inequality over [x,∞) for x>0 and employing (1.10), we have
˜F=F+v+˜v=ρ+F+˜ρ. | (3.2) |
This equality means that
˜F(x)>0ifF+>0,˜F(x)<0ifF+<0,˜F(x)=0ifF+=0 |
for any x∈R. Furthermore, we also get
˜F(0)=F+v+vb. |
Namely, in the case F+=0, our problem is reduced to the stationary problem of the compressible Navier-Stokes equation, which problem was studied in [13,21]. Thus, we mainly consider the case F+≠0 in this paper. We will discuss the case F+=0 at the end of this section.
By integrating (1.8) over [x,∞) for x>0 and substituing (3.1) and (3.2) into (1.10), we arrive at the following problem:
ν˜vx=Iβ(˜v), | (3.3) |
[1ex]limx→∞˜v(x)=v+, | (3.4) |
[1ex]˜v(0)=vb, | (3.5) |
where
Iβ(z):=ρ+v+(1−β2F2+|v+|2)(z−v+)+P(ρ+v+z)−P(ρ+). | (3.6) |
Here, we remark that Iβ(v+)=0. Our main purpose of this section is to construct the solution to (3.3)–(3.5). To this end, we analyze the profile of Iβ(z). It is easy to get limz→−0Iβ(z)=∞ and
limz→−∞Iβ(z)={∞if0<β<|v+/F+|,P(0)−P(ρ+)ifβ=|v+/F+|,−∞ifβ>|v+/F+|. | (3.7) |
For the derivative, we calculate
I′β(z)=ρ+v+(1−β2F2+v2+)−P′(ρ+v+z)ρ+v+z2, | (3.8) |
and this function leads to
I′β(v+)=ρ+c2+v+(M2+−1)−β2ρ+F2+v+. | (3.9) |
Furthermore, we also obtain
I′′β(z)={2P′(ρ+v+z)+P′′(ρ+v+z)ρ+v+z}ρ+v+z3>0 |
for z<0, which means that Iβ(z) is a convex function. In the case M+>1, (3.9) gives
I′β(v+)<0if0<β<βc, | (3.10) |
I′β(v+)=0ifβ=βc, | (3.11) |
I′β(v+)>0ifβ>βc, | (3.12) |
where
βc:=c+|F+|√M2+−1. | (3.13) |
In view of the classification of the sign of I′β(v+), we use the modified Mach number Mβ to see that the conditions (3.10)–(3.12) are rewritten as
I′β(v+)<0ifMβ>1,I′β(v+)=0ifMβ=1,I′β(v+)>0ifMβ<1. |
Remark that βc≤|v+/F+|. Especially, in the case 0<β<βc, there exists v∗ such that v+<v∗<0 and Iβ(v∗)=0. Furthermore, we also have Iβ(z)>0 for z<v+ and Iβ(z)<0 for v+<z<v∗. These properties are important to solve the problems (3.3)–(3.5).
Then the existence and property for the solution to (3.3)–(3.5) are described as the following key lemma.
Lemma 3.1. Assume v+<0. Then, the following facts hold true.
(ⅰ) (subsonic case) Assume that Mβ<1 holds. Then, the problems (3.3)–(3.5) with v+≠vb has no solution.
(ⅱ) Assume that Mβ≥1 holds. Then, the following assertions hold.
(ⅱ-ⅰ) (supersonic case) Suppose Mβ>1. Then, there exists a unique solution to (3.3)–(3.5) satisfying the following decay estimates if, and only if, vb<v∗:
Cδe−cx≤|(˜v−v+)(x)|≤˜Cδe−˜cx, | (3.14) |
|∂kx(˜v−v+)(x)|≤Cδe−cx,k=1,2. | (3.15) |
Furthermore, the solution ˜v monotonically increases if, and only if, vb<v+, and monotonically decreases if, and only if, v+<vb<v∗.
(ⅱ-ⅱ) (transonic case) Suppose Mβ=1. Then, there exists a unique solution to (3.3)–(3.5) if, and only if, vb<v+. Furthermore, the solution ˜v monotonically increases and satisfies the following estimates
cδ1+δx≤(v+−˜v)(x)≤Cδ1+δx, | (3.16) |
|∂kx(˜v−v+)(x)|≤Cδk+1(1+δx)k+1,k=1,2. | (3.17) |
Remark 3.2. The generalized Mach number Mβ gives the following characterization. Let Ωsuper, Γtrans and Ωsub be the sets that
Ωsuper:={(M+,β)|Mβ>1}={(M+,β)|M+>1, 0<β<βc},Γtrans:={(M+,β)|Mβ=1}={(M+,β)|M+>1, β=βc},Ωsub:={(M+,β)|Mβ<1}={(M+,β)|M+>1, β>βc}∪{(M+,β)|M+≤1} |
for (M+,β)∈R+×R+. Then, we find that the set {(M+,β)|M+>0, β>0} is separated to the three sets Ωsuper, Ωsub and Γtrans. More precisely, these sets are drawn in the (M+,β)-plain as follows (Figure 1).
Proof. Proof of Lemma 3.1 (ⅰ) Since (3.9), we have I′β(v+)>0 if Mβ<1. Therefore, using (3.3), we immediately conclude that ˜v(x) can not approach to v+ as x→∞, and there does not exist a solution to (3.3)–(3.5) (see also the two graphs of Iβ(z) displayed in the below figures).
(ⅱ-ⅰ) In the case Mβ>1, employing (3.3) and the properties of Iβ(z) mentioned before, we see that there exists a monotonically increasing solution ˜v to (3.3)–(3.5) if vb<v+. Similarly, we also find that there exists a monotonically decreasing solution ˜v to (3.3)–(3.5) if v+<vb<v∗ (see also the graph of Iβ(z) displayed in the below figure).
The uniqueness and smoothness of the solutions are derived from the standard argument for ordinary differential equations.
We shall derive the convergence estimate. Taylor's theorem gives
Iβ(z)=Iβ(v+)+I′β(v+)(z−v+)+12I′′β(v++θ(z−v+))(z−v+)2 | (3.18) |
for some θ∈(0,1). Since Iβ(v+)=0, and I′β(v+)<0 for Mβ>1, there exist positive constants δ∗, c∗ and C∗ such that
c∗(˜v−v+)≤−Iβ(˜v)≤C∗(˜v−v+) | (3.19) |
for |˜v−v+|≤δ∗. Thus, combining (3.3) and (3.19), and solving the resultant problem, we obtain
˜v(x)−v+≤(vb−v+)e−c∗x/ν,˜v(x)−v+≥(vb−v+)e−C∗x/ν | (3.20) |
for |˜v−v+|≤δ∗. Now, we had already obtained the existence of the global solution ˜v which approached to v+. This means that there exists x∗>0 such that |˜v(x)−v+|≤δ∗ for x≥x∗. This fact and (3.20) yields to (3.16). For the higher derivatives of ˜v, we can apply the same argument and omit it in detail.
(ⅱ-ⅱ) In the case Mβ=1, using the same argument as before, we find that there exists a monotonically increasing solution ˜v to (3.3)–(3.5) if, and only if, vb<v+ (see also the graph of Iβ(z) displayed in the below figure).
We consider the convergence estimate. In this case, it is not possible to derive the estimate (3.19), and it should be analyzed more carefully. Using (3.18) with Iβ(v+)=I′β(v+)=0 and the convexity of Iβ(z), we obtain
˜c∗(z−v+)2≤Iβ(z)≤˜C∗(z−v+)2, | (3.21) |
where ˜c∗ and ˜C∗ are positive constants defined by
˜c∗=12infz∈[vb,v+],θ∈[0,1]I′′β(v++θ(z−v+)),˜C∗=12supz∈[vb,v+],θ∈[0,1]I′′β(v++θ(z−v+)). |
Then, combining (3.3) and (3.21), and solving the resultant problem, we also get
v+−˜v(x)≤v+−vb1+˜c∗(v+−vb)x/ν,v+−˜v(x)≥v+−vb1+˜C∗(v+−vb)x/ν, |
which means (3.16). For the higher derivatives of ˜v, we employ (3.3) and (3.21) again and conclude (3.17).
Lemma 3.1 immediately gives the following proposition for (˜ρ,˜v,˜F).
Proposition 3.3. Assume v+<0. Then, the following facts hold true.
(ⅰ) (subsonic case) Assume that Mβ<1 holds. Then, the problems (1.7)–(1.11) with v+≠vb has no solution.
(ⅱ) Assume that Mβ≥1 holds. Then, the following assertions hold.
(ⅱ-ⅰ) (supersonic case)Suppose Mβ>1. Then, there exists a unique solution to (1.7)–(1.11) if, and only if, vb<v∗, and the following decay estimates are satisfied:
|∂kx(˜ρ−ρ+,˜v−v+,˜F−F+)(x)|≤Cδe−cx,k=0,1,2. | (3.22) |
Furthermore, the solution ˜v monotonically increases if, and only if, vb<v+ and monotonically decreases if, and only if, v+<vb<v∗.
(ⅱ-ⅱ) (transonic case) Suppose Mβ=1. Then, there exists a unique solution to (1.7)–(1.11) if, and only if, vb<v+. Furthermore, the solution ˜v monotonically increases and satisfies the following estimates:
|∂kx(˜ρ−ρ+,˜v−v+,˜F−F+)(x)|≤Cδk+1(1+δx)k+1,k=0,1,2. | (3.23) |
Remark 3.4. (ⅰ) In the case F+=0, the solution ˜F(x) to (1.7)–(1.11) becomes ˜F(x)≡0.
(ⅱ) The proof of Lemma 3.1 shows that taking β large so that β≥|v+/F+| at least, the stationary solution does not exist regardless of the Mach number M+ (see for the proof). This indicates that the strong recoiling effect of elastic force disturbs the stationary outflow of the fluid.
(ⅲ) It follows from (3.22) and (3.23) that (˜ρ,˜v,˜F) satisfies
‖(˜ρ,˜v,˜F)‖L∞≤|(ρ+,v+,F+)|+Cδ,‖∂kx(˜ρ,˜v,˜F)‖L∞≤{Cδ,Mβ>1,Cδk+1,Mβ=1,‖∂kx(˜ρ,˜v,˜F)‖L2≤{Cδ,Mβ>1,Cδk+12,Mβ=1 |
for k=1,2. These estimates will be repeatedly used in Section 5 and Appendix A.
Proof of Proposition 3.3. The equalities (3.1) and (3.2) tell us that (˜ρ,˜F) is rewritten by ˜v, which is the solution to (3.3)–(3.5). Then, it is easy to confirm that (˜ρ,˜v,˜F) is a solution to (1.7)–(1.11). Furthermore, employing (3.15), (3.17), and the fact that
˜ρ−ρ+=−ρ+˜v(˜v−v+),˜F−F+=F+v+(˜v−v+), |
the estimates (3.22) and (3.23) are also obtained.
This section is devoted to introducing the main result of this paper.
We first introduce the compatibility condition. We once assume that a smooth solution (ρ,v,F)(t,x) of the problems (1.1)–(1.6) exists and the perturbation (ρ−˜ρ,v−˜v,F−˜F) belongs to Z1(T) with some positive time T>0. Then, since v−˜v∈C([0,T];H10(R+)) holds, (v−˜v)(0,⋅) belongs to H10(R+). Therefore, it is necessary to impose the compatibility condition of 0-th order:
v0−˜v∈H10(R+). | (4.1) |
We state the following theorem related to the asymptotic stability of the stationary solution.
Theorem 4.1. Assume that at least one of the following two cases holds:
(ND) Mβ>1 and vb<v∗.
(D) Mβ=1, 1<M+<√ρ+P′′(ρ+)2P′(ρ+)+1 and vb<v+.
Suppose that the initial data (ρ0,v0,F0) and the boundary data vb satisfy (1.4)–(1.6), compatibility condition (4.1) and (ρ0−˜ρ,v0−˜v)∈H1(R+). Furthermore, assume that the initial data also satisfies (1.12) in H1(R+). Then, there exists a positive small number ε1>0 such that if
‖(ρ0−˜ρ,v0−˜v)‖H1+δ≤ε1,infx∈R+ϕ0(x)≥−12infx∈R+˜ρ(x), |
then the problems (1.1)–(1.6) has a unique solution (ρ,v,F)(t,x) satisfying (ρ−˜ρ,v−˜v,F−˜F)∈C([0,∞);H1). Furthermore, the solution converges to the stationary solution as a time goes to infinity, that is,
limt→∞‖(ρ−˜ρ,v−˜v,F−˜F)(t)‖L∞=0. | (4.2) |
Remark 4.2. (ⅰ) The restriction 1<M+<(ρ+(P′′(ρ+)/2P′(ρ+))+1)1/2 in the condition (D) arises from the convexity of the function P(ρ)−β2cρ2+F2+/ρ around ρ=ρ+ to show the stability of the degenerate stationary solution.
(ⅱ) Under the second condition of (1.4), we can reformulate (1.12) as F0−˜F=−(ρ0−˜ρ)˜F/ρ0, which gives the smallness of F0−˜F in H1(R+), provided that ρ0−˜ρ is small in H1(R+). This fact and Theorem 4.1 mean the smallness of (ρ0−˜ρ,v0−˜v,F0−˜F) and the asymptotic stability of (˜ρ,˜v,˜F). $
Our purpose of this section is to prove Theorem 4.1. The argument is based on the combination of the local existence theory and the corresponding a priori estimate for the solution.
To consider the stability of the stationary solutions, we introduce a new function by
J(t,x):=ρ(t,x)F(t,x). |
Then, the Eqs (1.1) and (1.3) lead to
Jt+vJx=0. | (5.1) |
For the Eq (5.1), we assign the initial data J(0,x)=J0(x) in x∈R+, where J0:=ρ0F0. The Eq (5.1) has useful properties, and we often utilize (5.1) instead of (1.3). Indeed, the stationary solution ˜J(x) of (5.1) satisfies ˜v˜Jx=0 and this gives ˜J(x)=J+, where J+:=ρ+F+. This fact will be used later.
On the other hand, it is useful to employ the different expression for (1.2), that is,
ρvt+ρvvx−νvxx+Pβ(ρ)x−β2Q(ρ,J)x=0, |
where
Pβ(ρ):=P(ρ)−β2J2+ρ,Q(ρ,J):=1ρ(J2−J2+). |
Remark that P′β(ρ)=P′(ρ)+β2J2+/ρ2>0 holds true.
We set a perturbation from the stationary solution as
(ϕ,ψ)(t,x):=(ρ,v)(t,x)−(˜ρ,˜v)(x),ζ(t,x):=J(t,x)−J+. |
Coupling (1.1), (1.2) and (5.1) with (1.4)–(1.11), the perturbation (ϕ,ψ,ζ) satisfies the system
ϕt+vϕx+ρψx=f1,ρψt+ρvψx+P′β(ρ)ϕx−νψxx=f2,ζt+vζx=0 | (5.2) |
with the initial conditions and boundary conditions at x=∞ and x=0:
(ϕ,ψ,ζ)|t=0=(ϕ0,ψ0,ζ0),infx∈R+(˜ρ(x)+ϕ0(x))>0, | (5.3) |
limx→∞(ϕ,ψ,ζ)=(0,0,0), | (5.4) |
ψ(t,0)=0. | (5.5) |
Here, the initial perturbation is defined by
(ϕ0,ψ0)(x):=(ρ0−˜ρ,v0−˜v)(x),ζ0(x):=J0(x)−J+, |
and the functions f1 and f2 are given by
f1:=−(ϕ˜vx+ψ˜ρx),f2:=−(ρv−˜ρ˜v)˜vx−(P′β(ρ)−P′β(˜ρ))˜ρx+β2Q(ρ,J)x. |
Firstly, we mention the local-in-time existence of the solution of (5.2) with boundary conditions (5.4) and (5.5), and initial and compatibility conditions at t=τ:
(ϕ,ψ,ζ)|t=τ=(ϕτ,ψτ,ζτ),infx∈R+(˜ρ(x)+ϕτ(x))>0, | (5.6) |
ψτ∈H10(R+). | (5.7) |
Here, τ≥0 is an arbitrary non-negative number.
Proposition 5.1. There exists a positive small number ε∈(0,1) independent of τ such that the following assertion holds true:
Let τ≥0 and M0∈(0,ε]. If (ϕτ,ψτ,ζτ) satisfies the compatibility condition of 0-th order (5.7) and
‖(ϕτ,ψτ,ζτ)‖H1≤M0,infx∈R+ϕτ(x)≥−12infx∈R+˜ρ(x), |
then there exists T=T(δ,M0)>0 independent of τ such that the problem (5.2) with (5.4)–(5.6) has a unique solution (ϕ,ψ,ζ)∈Z1(τ,τ+T) satisfying
inf(t,x)∈[τ,τ+T]×R+ϕ(t,x)≥−34infx∈R+˜ρ(x), | (5.8) |
(ϕ,ψ,ζ)∈X13M0(τ,τ+T),ϕ,ζ∈C1([τ,τ+T];L2(R+)). | (5.9) |
Remark 5.2. Under (5.9), the first and third equations of (5.2) hold in C([τ,τ+T];L2(R+)), and the second equation of (5.2) makes sense in L2(τ,τ+T;L2(R+)).
Proposition 5.1 will be proved in Appendix A for the case τ=0 by using the iteration argument.
We next focus on the a priori estimate in the Sobolev space. To state this, we introduce the following notations:
N(t)2:=sup0≤τ≤t(‖ϕ(τ)‖2H1+‖ψ(τ)‖2H1),N0(t):=sup0≤τ≤t‖J(τ)‖L∞,M(t)2:=∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2H1+|ϕ(τ,0)|2)dτ. |
Then, the a priori estimate is summarized by the following proposition.
Proposition 5.3. Suppose that the same assumptions in Theorem 4.1 hold true. Let (ϕ,ψ,ζ) be a solution to (5.2)–(5.5) in a time interval [0,T] with (ϕ,ψ,ζ)∈Z1(T). Then there exist positive constants ε0 and C0, such that if N(T)+δ≤ε0, then the following estimates hold uniformly for t∈[0,T]:
N(t)2+M(t)2≤C0(‖ϕ0‖2H1+‖ψ0‖2H1),‖ζ(t)‖H1=0, | (5.10) |
inf(t,x)∈[0,t]×R+ϕ(t,x)≥−12infx∈R+˜ρ(x). | (5.11) |
To construct the a priori estimate, we have to employ the properties for the stationary solutions constructed in Lemma 3.1. Specifically, the following lemma is important to derive the a priori estimate.
Lemma 5.4 ([13]). Suppose that the same assumptions as in Proposition 5.3 hold. Then, the following estimates are obtained.
(ⅰ) Let Mβ>1. Then, the following estimates hold true:
∫t0∫∞0|∂kx˜v|j|ϕ|2dxdτ≤Cδj∫t0(‖ϕx(τ)‖2L2+|ϕ(τ,0)|2)dτ,∫t0∫∞0|∂kx˜v|j|ψ|2dxdτ≤Cδj∫t0‖ψx(τ)‖2L2dτ | (5.12) |
for t∈[0,T] and k,j∈N.
(ⅱ) Let Mβ=1. Then the following estimates hold true:
∫t0∫∞0|∂kx˜v|j|ϕ|2dxdτ≤Cδ(k+1)j−2∫t0(‖ϕx(τ)‖2L2+|ϕ(τ,0)|2)dτ,∫t0∫∞0|∂kx˜v|j|ψ|2dxdτ≤Cδ(k+1)j−2∫t0‖ψx(τ)‖2L2dτ | (5.13) |
for t∈[0,T] and k,j∈N with k+j≥3.
Proof. The proof is based on (3.22) and (3.23). The argument is the same as the one in [13], and we omit it in detail.
Proposition 3.3 and the smallness assumptions on ε0 in Proposition 5.3 ensure that if ε0 is sufficiently small, then (5.11) is obtained, and there exist certain positive constants cρ, Cρ, cv and Cv such that
cρ≤ρ(t,x)≤Cρ,cv≤−v(t,x)≤Cv,˜ρ(x)≥ρ+2 | (5.14) |
for (t,x)∈[0,T]×R+.
Here, we have noticed the facts ρ+−C(δ+‖ϕ‖L∞(R+))≤ρ≤ρ++C(δ+‖ϕ‖L∞(R+)) and |v+|−C(δ+‖ψ‖L∞(R+))≤−v≤|v+|+C(δ+‖ψ‖L∞(R+)) by using Proposition 3.3, and then applying Lemma 2.1. To derive (5.10) in Proposition 5.3, we often utilize this boundedness for ρ and ˜ρ.
To construct the a priori estimate, we need an important property of ζ. More precisely, the following key lemma is shown.
Lemma 5.5. Suppose that the same assumptions as in Proposition 5.3 hold true. Then, ζ(t)≡0 in H1(R+) for all t∈[0,T].
Proof. In view of Remark 5.2, we first show that ζ is the weak solution of
ζt+a(ψ)ζx=0,ζ|t=0=ζ0,ζ|x=∞=0. |
In fact, multiplying third equation of (5.2) by an arbitrary function φ∈C10([0,T)×R+), integrating over (0,T)×R+ and applying integration by parts, we have the following weak form
−∫T0(ζ,φt+(a(ψ)φ)x)L2dτ=(ζ0,φ(0,⋅))L2. |
This gives the desired fact. We then apply Lemma 2.5 with k=1,˜ψ=ψ,ϕ=ζ,t1=0,t2=t and f=0 to obtain
‖ζ(t)‖2H1≤‖ζ0‖2H1+1∑l=0∫t0∫∞0(vx|∂lxζ|2+2lvx|ζx|2)dxdτ≤‖ζ0‖2H1+C∫t0‖vx(τ)‖L∞‖ζ(τ)‖2H1dτ, |
and the Gronwall inequality gives
‖ζ(t)‖2H1≤‖ζ0‖2H1exp(Ct12{∫t0‖vx(τ)‖2L∞dτ}12) |
for t∈[0,T]. Therefore, using the condition (1.12), Remark 3.4 (ⅲ) and (5.9), we obtain ‖ζ(t)‖H1=0 for all t∈[0,T], and this completes the proof.
Remark 5.6. Lemma 5.5 means that the solution F is represented by F=ρ+F+/ρ, once the solution ρ is constructed.
In view of Lemma 5.5, it is enough to concentrate the derivation of a priori estimate for (ϕ,ψ)(t) only. We first show the basic estimate for (ϕ,ψ)(t). The proof is based on its suitable energy form.
Lemma 5.7. Suppose that the same assumptions as in Proposition 5.3 hold true. Then the following estimate holds:
‖ϕ(t)‖2L2+‖ψ(t)‖2L2+∫t0(‖ψx(τ)‖2L2+|ϕ(τ,0)|2)dτ≤C(‖ϕ0‖2L2+‖ψ0‖2L2)+CδM(t)2+β4C(1+N0(t)2)∫t0‖ζ(τ)‖2L2dτ | (5.15) |
for t∈[0,T].
Proof. To derive the desired estimate, we employ the energy form. We introduce the useful energy function as follows:
E:=Φ(ρ,˜ρ)+12ψ2,Φ(ρ,˜ρ):=∫ρ˜ρPβ(η)−Pβ(˜ρ)η2dη. | (5.16) |
Then the function Φ(ρ,˜ρ) has the following expansion.
Φ(ρ,˜ρ)=Φ(ρ)−Φ(˜ρ)−∂˜wΦ(˜ρ)(w−˜w),Φ(ρ):=∫ρPβ(η)η2dη, |
where w=1/ρ and ˜w=1/˜ρ. This expression with ∂wΦ(ρ)=−Pβ(ρ) and ∂2wΦ(ρ)=ρ2P′β(ρ)>0 lead to the fact that ρΦ(ρ,˜ρ) is equivalent to |ρ−˜ρ|2 for small |ρ−˜ρ|, and there exist positive constants c0 and C0 such that
c0(ϕ2+ψ2)≤ρE≤C0(ϕ2+ψ2). | (5.17) |
This energy function satisfies the following energy form:
(ρE)t+Fx+νψ2x=R1+R2−β2ψxQ(ρ,J), | (5.18) |
where
F:=ρvE+(Pβ(ρ)−Pβ(˜ρ))ψ−νψψx−β2ψQ(ρ,J),R1:=−ν˜vxx˜ρϕψ,R2:=−{ρψ2+Pβ(ρ)−Pβ(˜ρ)−P′β(˜ρ)ϕ}˜vx. |
Integrating (5.18) over (0,t)×R+ and employing the boundary conditions (5.4) and (5.5), we get
∫∞0ρ(t,x)E(t,x)dx−∫t0(ρvE)(τ,0)dτ+ν∫t0‖ψx(τ)‖2L2dτ=∫∞0ρ0(x)E(0,x)dx+∫t0∫∞0(R1(τ,x)+R2(τ,x))dxdτ−β2∫t0∫∞0ψx(τ,x)Q(ρ,J)(τ,x)dxdτ. | (5.19) |
Because of vb<0, the second term on the left-hand side of (5.19) is handled as
−(ρvE)(τ,0)=|vb|ρ(τ,0)E(τ,0)≥c0|vb||ϕ(τ,0)|2. |
For the remainder terms R1 and R2, we estimate
|R1|≤C|˜vxx||ϕ||ψ|,|R2|≤C|˜vx|(ψ2+ϕ2). |
Therefore, for the non-degenerate case Mβ>1, we apply (5.12) and obtain
∫t0∫∞0(|R1|+|R2|)dxdτ≤Cδ∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2L2+|ϕ(τ,0)|2)dτ≤CδM(t)2. | (5.20) |
On the other hand, we can not employ the same argument for R2 in the degenerate case Mβ=Mβc=1. To overcome difficulty, we reformulate R2 as
R2=−˜vxρψ2−12P′′β(ρ+)˜vxϕ2−12(P′′βc(˜ρ)−P′′βc(ρ+))˜vxϕ2−(Pβc(ρ)−Pβc(˜ρ)−P′βc(˜ρ)ϕ−12P′′βc(˜ρ)ϕ2)˜vx. | (5.21) |
Due to ˜vx>0, the first term of the right hand side in (5.21) is negative. Furthermore, using the fact that P′′βc(ρ+)>0 is satisfied if, and only if, 0<βc<β∗, where
β∗:=√ρ+P′′(ρ+)2F2+, |
the second term of the right hand side in (5.21) is also negative if 0<βc<β∗. Namely, R2 is estimated as
˜vxρψ2+c˜vxϕ2+R2≤0 | (5.22) |
for suitably small δ and |ϕ|. Therefore, using (5.22), (5.20) also holds for the case Mβ=Mβc=1.
For the last term in (5.19), using the fact that ‖Q(ρ,J)‖L2≤(|J+|+‖J‖L∞)‖ζ‖L2/cρ, we obtain
β2∫∞0ψx(τ,x)Q(ρ,J)(τ,x)dx≤ε‖ψx‖2L2+β4Cε‖Q(ρ,J)‖2L2≤ε‖ψx‖2L2+β4Cε(1+‖J‖2L∞)‖ζ‖2L2. |
Finally, substituting the above estimate into (5.19), we arrive at the desired estimate and complete the proof.
The next goal is to derive the estimate for the first-order derivatives of the solution. To obtain the estimate for ϕx, we need to deal with ϕxx and ϕ(t,0) after differentiating the first equation of (5.2) in x and applying integration by parts, formally. However, since we only treat ϕ(t)∈H1(R+) for 0≤t≤T, ϕxx and ϕ(t,0) do not always exist. Therefore, the formal argument cannot make sense in our setting. To overcome this difficulty, we recall Definition 2.3 and Lemma 2.5 to follow the theory of weak solutions to transport equations. This merit is that ϕxx and ϕ(t,0) are not needed in their statements and the proof of the estimate for ϕx. This argument is inspired in [12].
Lemma 5.8. Suppose that the same assumptions as in Proposition 5.3 hold true. Then the following estimate holds:
‖ϕx(t)‖2L2+∫t0‖ϕx(τ)‖2L2dτ≤C(‖ψ0‖2L2+‖ϕ0,x‖2L2)+C(‖ψ(t)‖2L2+∫t0‖ψx(τ)‖2L2dτ)+C(N(t)+δ)M(t)2+β4C(1+N0(t))2(1+N(t))2∫t0‖ζ(τ)‖2H1dτ | (5.23) |
for t∈[0,T].
Proof. We first claim that ¯ϕ:=ϕx∈C([0,T];L2(R+)) is the weak solution of
¯ϕt+a(ψ)¯ϕx=˜f1,¯ϕ|t=0=ϕ0,x,¯ϕ|x=∞=0, | (5.24) |
where ˜f1:=−vxϕx+(f1−ρψx)x∈L2(0,T;L2(R+)). To show this fact, we notice that the first equation of (5.2) holds for almost every (t,x)∈(0,T)×R+ because of Remark 5.2. Let φ be an arbitrary function belonging to C10([0,T)×R+). Multiplying the first equation of (5.2) by −φx and integrating over (0,T)×R+ we have
−∫T0∫∞0(ϕt+a(ψ)ϕx)φxdxdt=−∫T0∫∞0(−ρψx+f1)φxdxdt. |
Applying integration by parts and using a(ψ)φx=(a(ψ)φ)x−vxφ in we arrive at the weak form
−∫T0(ϕx,φt+(a(ψ)φ)x)L2dτ=(ϕ0,x,φ(0,⋅))L2+∫T0(˜f1,φ)L2dτ, |
which yields the desired fact. Therefore, we are able to apply Lemma 2.5 with k=0,˜ψ=ψ,ϕ=ϕx,t1=0,t2=tt and f=˜f1 to give
‖ϕx(t)‖2L2≤‖ϕ0,x‖2L2−∫t0∫∞0vxϕ2xdxdτ−2∫t0∫∞0ρϕxψxxdxdτ−2∫t0∫∞0ρxψxϕxdxdτ+2∫t0∫∞0∂xf1ϕxdxdτ | (5.25) |
for 0≤t≤T. We next eliminate ρϕxψxx in the right-hand side of (5.25). Multiplying the second equation in (5.2) by ρϕx yields
ρ2ϕxψt+ρP′β(ρ)ϕ2x−νρϕxψxx+ρ2vϕxψx=ρϕxf2. |
Then, integrating above equation over (0,t)×R+ and using the formula
∫t0∫∞0ρ2ϕxψtdxdτ=∫∞0ρ2ϕxψdx−∫∞0ρ20ϕ0,xψ0dx+∫t0∫∞0(2˜ρxψ+ρψx)ρϕtdxdτ, |
we have
∫∞0ρ2ϕxψdx+∫t0∫∞0ρP′β(ρ)ϕ2xdxdτ−ν∫t0∫∞0ρϕxψxxdxdτ=∫∞0ρ20ϕ0,xψ0dx−∫t0∫∞0ρ2vϕxψxdxdτ−∫t0∫∞0(2˜ρxψ+ρψx)ρϕtdxdτ+∫t0∫∞0ρϕxf2dxdτ. | (5.26) |
Here, the integral of ρ2ϕxψt is calculated via mollification with respect to t and integration by parts. This calculation is standard, so we omit the detail. Then, combining (5.25) and (5.26) to eliminate ρϕxψxx, we get
ν2‖ϕx‖2L2+∫∞0ρ2ϕxψdx+∫t0∫∞0ρP′β(ρ)ϕ2xdxdτ≤ν2‖ϕ0,x‖2L2+∫∞0ρ20ϕ0,xψ0dx+∫t0∫∞0R1dxdτ, | (5.27) |
where
R1:=νϕx∂xf1+ρϕxf2−ν2vxϕ2x−(νρx+ρ2v)ϕxψx−(2˜ρxψ+ρψx)ρϕt. |
We estimate the remainder term R1. Since Q(ρ,J)x=2JJx/ρ−(J2−J2+)ρx/ρ2, we have
‖Q(ρ,J)x‖L2≤2cρ‖J‖L∞‖ζx‖L2+1c2ρ(|J+|+‖J‖L∞)(‖˜ρxζ‖L2+‖ϕxζ‖L2)≤C(1+‖J‖L∞)(1+‖ϕx‖L2)‖ζ‖H1. |
Thus, this estimate and
|f1|≤|˜vx||ϕ|+|˜ρx||ψ|,|∂xf1|≤|˜vxx||ϕ|+|˜ρxx||ψ|+|˜vx||ϕx|+|˜ρx||ψx|,|f2|≤|˜vx|(|˜v||ϕ|+|ρ||ψ|)+C|˜ρx||ϕ|+β2|Q(ρ,J)x| |
give
∫t0‖f1‖2L2dτ≤2∫t0∫∞0(˜v2xϕ2+˜ρ2xψ2)dxdτ≤Cδ2∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2L2+|ϕ(τ,0)|2)dτ, | (5.28) |
∫t0‖∂xf1‖2L2dτ≤4∫t0∫∞0(˜v2xxϕ2+˜ρ2xxψ2+˜v2xϕ2x+˜ρ2xψ2x)dxdτ≤Cδ2∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2L2+|ϕ(τ,0)|2)dτ, | (5.29) |
and
∫t0‖f2‖2L2dτ≤C∫t0∫∞0(˜v2x(ϕ2+ψ2)+˜ρ2xϕ2)dxdτ+β4C∫t0‖Q(ρ,J)x‖2L2dτ≤Cδ2∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2L2+|ϕ(τ,0)|2)dτ+β4C(1+N0(t))2(1+N(t))2∫t0‖ζ‖2H1dτ. | (5.30) |
Here, Lemmas 3.1 and 5.4, and (5.14) are also applied. Using the first equation in (5.2), we have
|R1|≤C|ϕx|(|∂xf1|+|f2|)+C|vx|ϕ2x+C(|ρx|+|v|)|ϕx||ψx|+C(|˜ρx||ψ|+|ψx|)(|f1|+|v||ϕx|+|ψx|), |
and this gives
∫∞0|R1|dx≤σ‖ϕx‖2L2+Cσ(‖ψx‖2L2+‖f1‖2H1+‖f2‖2L2)+C∫∞0|˜vx|ϕ2xdx+C∫∞0|ψx|ϕ2xdx+Cσ∫∞0|˜ρx|2|ψ|2dx |
for all σ>0. Thus, the estimates (5.28), –(5.30) and Lemma 5.4 yield
∫t0∫∞0|R1|dxdτ≤(σ+Cδ)∫t0‖ϕx‖2L2dτ+Cσ∫t0‖ψx‖2L2dτ+Cσδ2M(t)2+CN(t)M(t)2+β4Cσ(1+N0(t))2(1+N(t))2∫t0‖ζ‖2H1dτ. | (5.31) |
Here, we used the fact that
∫t0∫∞0|ψx|ϕ2xdxdτ≤∫t0‖ψx‖H1‖ϕx‖2L2dτ≤supτ∈[0,t]‖ϕx(τ)‖L2∫t0(‖ψx(τ)‖2H1+‖ϕx(τ)‖2L2)dτ. |
Therefore, (5.27) with (5.31) leads to
c‖ϕx‖2L2+˜cρ∫t0‖ϕx‖2L2dτ≤C(‖ψ0‖2L2+‖ϕ0,x‖2L2)+C‖ψ‖2L2+(σ+Cδ)∫t0‖ϕx‖2L2dτ+Cσ∫t0‖ψx‖2L2dτ+Cσδ2M(t)2+CN(t)M(t)2+β4Cσ(1+N0(t))2(1+N(t))2∫t0‖ζ‖2H1dτ, |
where ˜cρ:=minρ∈[cρ,Cρ]ρP′β(ρ). Thus, letting σ suitably small, we arrive at the desired estimate (5.23) and the proof is completed.
We next focus on ψx whose estimate is given in the following lemma.
Lemma 5.9. Suppose that the same assumptions as in Proposition 5.3 hold true. Then, the following inequality holds:
‖ψx(t)‖2L2+∫t0‖ψxx(τ)‖2L2dτ≤C‖ψ0,x‖2L2+C∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2L2) dτ+Cδ2M(t)2+β4C(1+N0(t))2(1+N(t))2∫t0‖ζ(τ)‖2H1dτ | (5.32) |
for t∈[0,T].
Proof. Multiplying the second equation in (5.2) by −ψxx/ρ yields
−ψxψxx+νρψ2xx=R2, |
where
R2:=−ψxxρf2+vψxψxx+P′β(ρ)ρϕxψxx. |
Then, integrating this equality over (0,t)×R+ and using the formula
−∫t0∫∞0ψxψxxdxdτ=12(‖ψx‖2L2−‖ψ0,x‖2L2), |
we obtain
12‖ψx‖2L2+νCρ∫t0‖ψxx‖2L2dτ≤12‖ψ0,x‖2L2+∫t0∫∞0|R2|dxdτ. | (5.33) |
Here, the integral of ψxψxx is calculated via mollification with respect to t and integration by parts. This calculation is standard, so we omit the detail. We then estimate the remainder term as |R2|≤C(|ϕx|+|ψx|+|f2|)|ψxx|, and this gives
∫t0∫∞0|R2|dxdτ≤σ∫t0‖ψxx‖2L2dτ+Cσ∫t0(‖ϕx‖2L2+‖ψx‖2L2+‖f2‖2L2)dτ≤σ∫t0‖ψxx‖2L2dτ+Cσ∫t0(‖ϕx‖2L2+‖ψx‖2L2)dτ+Cσδ2∫t0(‖ϕx‖2L2+‖ψx‖2L2+|ϕ(τ,0)|2)dτ+β4Cσ(1+N0(t))2(1+N(t))2∫t0‖ζ‖2H1dτ | (5.34) |
for all σ>0. The estimate (5.33) with (5.34) leads to the desired estimate (5.32). This completes the proof.
Proof of Proposition 5.3. Combining (5.15) and (5.23), we have
‖ϕ(t)‖2H1+‖ψ(t)‖2L2+∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2L2+|ϕ(τ,0)|2)dτ≤C(‖ϕ0‖2H1+‖ψ0‖2L2)+C(N(t)+δ)M(t)2+β4C(1+N0(t))2(1+N(t))2∫t0‖ζ(τ)‖2H1dτ. |
Furthermore, substituting the resultant estimate and (5.32), we obtain
‖ϕ(t)‖2H1+‖ψ(t)‖2H1+∫t0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2H1+|ϕ(τ,0)|2)dτ≤C(‖ϕ0‖2H1+‖ψ0‖2H1)+C(N(t)+δ)M(t)2+β4C(1+N0(t))2(1+N(t))2∫t0‖ζ(τ)‖2H1dτ. |
Namely this gives
N(t)2+M(t)2≤C(‖ϕ0‖2H1+‖ψ0‖2H1)+C(N(t)+δ)M(t)2+β4C(1+N0(t))2(1+N(t))2∫t0‖ζ(τ)‖2H1dτ |
for t∈[0,T]. Consequently, by using Lemma 5.5 and taking ε0 suitably small, we arrive at the desired estimate (5.10). This completes the proof.
In order to complete the proof of Theorem 4.1, it remains to derive the estimate for F−˜F, which is summarized as the following lemma.
Lemma 5.10. Suppose that the same assumptions as in Theorem 4.1 hold. Then, F satisfies
‖(F−˜F)(t)‖H1≤C‖(ϕ0,ψ0)‖H1 | (5.35) |
for all t∈[0,T].
Proof. Using (3.2) and Lemma 5.5, F−˜F is rewritten as
F−˜F=−˜Fρϕ. | (5.36) |
Therefore, we easily see from Proposition 3.3, Proposition 5.3, (5.14) and (5.36) that
‖(F−˜F)(t)‖L2≤C‖ϕ(t)‖L2≤C‖(ϕ0,ψ0)‖H1 | (5.37) |
for all t∈[0,T]. We next study (F−˜F)x. Differentiating (5.36) in x gives
(F−˜F)x=−˜Fxρϕ+˜Fρ2ρxϕ−˜Fρϕx. |
It then follows from Proposition 3.3, Proposition 5.3, (5.14) and (5.36) that
‖(F−˜F)x(t)‖L2≤C‖ϕ(t)‖H1≤C‖(ϕ0,ψ0)‖H1 | (5.38) |
for all t∈[0,T]. Consequently, the desired estimate (5.35) immediately follows from (5.37) and (5.38). This completes the proof.
Proof of Theorem 4.1. To show the global existence result, we use a bootstrap method based on the combination of Proposition 5.1, Proposition 5.3 and Lemma 5.10. Let ε0 in Proposition 5.3 be ε0∈(0,ε], where ε is introduced in Proposition 5.1. We then set
ε1:=ε06min{1,1C0}≤ε. |
We start from the assumptions
‖(ϕ0,ψ0,ζ0)‖H1+δ≤ε1,infx∈R+ϕ0(x)≥−12infx∈R+˜ρ(x). |
Since ‖(ϕ0,ψ0,ζ0)‖H1≤ε0/6≤ε clearly holds, we are able to apply Proposition 5.1 with τ=0, M0=ε0/6(≤ε) and (ϕτ,ψτ,ζτ)=(ϕ0,ψ0,ζ0) to show that (5.2)–(5.5) has a unique solution (ϕ,ψ,ζ)∈Z1(0,T)=Z1(T) satisfying (5.8) and (5.9) with T=T(δ,ε0/6), τ=0, M0=ε0/6 and (ϕτ,ψτ,ζτ)=(ϕ0,ψ0,ζ0). Moreover, it follows from (5.9) and δ≤ε1≤ε0/6 that
supt∈[0,T]‖(ϕ,ψ,ζ)(t)‖H1+δ≤3⋅ε06+ε06≤ε0. | (5.39) |
Therefore, we can apply Proposition 5.3 to obtain
supt∈[0,T]‖(ϕ,ψ,ζ)(t)‖H1≤C0‖(ϕ0,ψ0,ζ0)‖H1≤C0ε1≤ε06,inf(t,x)∈[0,T]×R+ϕ(t,x)≥−12infx∈R+˜ρ(x), |
which means
‖(ϕ,ψ,ζ)(T)‖H1≤ε06(≤ε),infx∈R+ϕ(T,x)≥–12infx∈R+˜ρ(x). |
Here, we use the fact ε1≤ε0/(6C0). Therefore, we are able to construct a unique solution (ϕ1,ψ1,ζ1)∈Z1(T,2T) of (5.2)–(5.5) satisfying (5.8) and (5.9) with τ=T, M0=ε0/6 and (ϕτ,ψτ,ζτ)=(ϕ,ψ,ζ)(T) by applying Proposition 5.1 with τ=T=T(δ,ε0/6), M0=ε0/6 and (ϕτ,ψτ,ζτ)=(ϕ,ψ,ζ)(T). Moreover, we see from (5.9) and δ≤ε1≤ε0/6 that
supt∈[T,2T]‖(ϕ1,ψ1,ζ1)(t)‖H1+δ≤3⋅ε06+ε06≤ε0. | (5.40) |
Therefore, we are able to extend the solution of (5.2)–(5.5) in Z1(T) to that in Z1(2T) by defining (ϕ,ψ,ζ)(t,x):=(ϕ1,ψ1,ζ1)(t,x) for (t,x)∈[T,2T]×R+. Combining (5.39) and (5.40) to confirm
supt∈[0,2T]‖(ϕ,ψ,ζ)(t)‖H1+δ≤ε0, |
we can apply Proposition 5.3 with T=2T and notice the fact ε1≤ε0/(6C0) to obtain
supt∈[0,2T]‖(ϕ,ψ,ζ)(t)‖H1≤C0‖(ϕ0,ψ0,ζ0)‖H1≤C0ε1≤ε06,inf(t,x)∈[0,2T]×R+ϕ(t,x)≥−12infx∈R+˜ρ(x) |
which yields
‖(ϕ,ψ,ζ)(2T)‖H1≤ε06(≤ε),infx∈R+ϕ(2T,x)≥−12infx∈R+˜ρ(x). |
Therefore, it follows from Proposition 5.1 with τ=2T=2T(δ,ε0/6) and (ϕτ,ψτ,ζτ)=(ϕ,ψ,ζ)(2T) that we are able to extend the solution (ϕ,ψ,ζ) of (5.2)–(5.5) in Z1(2T) to that in Z1(3T). Consequently, repeating the above argument we have a global-in-time solution (ϕ,ψ,ζ) of (5.2)–(5.5) in [0,∞)×R+ with the desired properties
(ϕ,ψ,ζ)∈C([0,∞);H1(R+)),supt∈[0,∞)‖(ϕ,ψ,ζ)(t)‖H1≤C0‖(ϕ0,ψ0,ζ0)‖H1,inf(t,x)∈[0,∞)×R+ϕ(t,x)≥−12infx∈R+˜ρ(x). | (5.41) |
We next prove the limit (4.2). We see from (5.41) and the definition of ε1 that ‖(ϕ,ψ,ζ)(t)‖H1+δ≤ε0 holds for all t≥0. Therefore, Proposition 5.3 leads to
M(∞)2:=∫∞0(‖ϕx(τ)‖2L2+‖ψx(τ)‖2H1+|ϕ(τ,0)|2)dτ≤C‖(ϕ0,ψ0,ζ0)‖2H1<∞, | (5.42) |
which implies that there exists a monotone increasing sequence {tm}∞m=1 with tm→∞(m→∞) such that the following limit holds:
limm→∞(‖ϕx(tm)‖2L2+‖ψx(tm)‖2H1)=0. | (5.43) |
In view of Lemma 2.1, it suffices to prove ‖(ϕ,ψ)x(t)‖L2→0 as t→∞. Since ψ belongs to same function space as Lemma 2.2 with an arbitrary T, we have the estimate
‖ψx(t)‖2L2≤C{‖ψx(tm)‖2L2+(∫ttm‖ψt(τ)‖2L2dτ)12(∫ttm‖ψx(τ)‖2H1dτ)12} | (5.44) |
for t≥tm. It is directly seen from the second equation of (5.2) and inequalities (5.14), (5.30) and (5.42) that ψt satisfies
∫∞0‖ψt(τ)‖2L2dτ≤CM(∞)2≤C‖(ϕ0,ψ0,ζ0)‖2H1<∞. |
Therefore, for any η>0, there exists m1∈N such that if t≥tm1 holds, then (5.44) leads to ‖ψx(t)‖2L2<η, which yields ‖ψx(t)‖L2→0 as t→∞. To show ‖ϕx(t)‖L2→0 as t→∞, we make use the fact that ϕx is the weak solution of the problem (5.24) for an arbitrary number T>0. Applying Lemma 2.5 with k=0,˜ψ=ψ,ϕ=ϕx,t1=tm,t2=t and f=˜f1 we obtain
‖ϕx(t)‖2L2≤‖ϕx(tm)‖2L2+∫ttm(‖vx(τ)‖L∞‖ϕx(τ)‖L2+2‖˜f1(τ)‖L2)‖ϕx(τ)‖L2dτ | (5.45) |
for t≥tm. It follows from (3.22), (5.14), (5.28), (5.41) and (5.42) that the integrand in the right-hand side of (5.45) satisfies
∫∞0(‖vx(τ)‖L∞‖ϕx(τ)‖L2+2‖g1(τ)‖L2)‖ϕx(τ)‖L2dτ≤CM(∞)2≤C‖(ϕ0,ψ0,ζ0)‖2H1<∞. |
Therefore, for any η>0, there exists m2∈N such that if t≥tm2 holds, then (5.45) gives ‖ϕx(t)‖2L2<η, which leads to ‖ϕx(t)‖L2→0 as t→∞. Consequently we have ‖(ϕ,ψ)(t)‖L∞→0 as t→∞. The limit ‖(F−˜F)(t)‖L∞→0 as t→∞ and property F−˜F∈C([0,∞);H1(R+)) directly follow from (5.35) and (5.36), and hence we arrive at the desired results (ρ−˜ρ,v−˜v,F−˜F)∈C([0,∞);H1(R+)) and (4.2). This completes the proof of Theorem 4.1.
Yusuke Ishigaki and Yoshihiro Ueda: Conceptualization, Formal analysis, Investigation, Methodology, Validation, Writing-original draft, Writing-review & editing. Both authors have read and approved the final version of the manuscript for publication.
Y. I. is supported by JSPS KAKENHI Grant Number JP20H00118 and JP23KJ1408. Y. U. is supported by JSPS KAKENHI Grant Number 21K03327 and 21KK0243. The authors would like to thank Professor Yoshiyuki Kagei for his valuable discussion for the proof of Proposition 5.1.
The authors declare no conflict of interests.
In this appendix, we show Proposition 5.1 by using standard iteration method.
Without loss of geneality, we only consider the case τ=0. Throughout this appendix, we set u(t,x):=(ϕ,ψ,ζ)(t,x) and u0(x):=(ϕ0,ψ0,ζ0)(x).
In order to perform iteration method, we rewrite (5.2)–(5.5) in the following form
ϕt+a(ψ)ϕx=g1(ϕ,ψ,ψx),b(ϕ)ψt+Bψ=g2(u,ux),ζt+a(ψ)ζx=0, | (A.1) |
u|t=0=u0,u|x=∞=0,ψ|x=0=0, | (A.2) |
infx∈R+(˜ρ(x)+ϕ0(x))>0, | (A.3) |
where the functions a,b and the operator Bψ are introduced in Section 2, and g1 and g2 are defined by
g1(ϕ,ψ,ψx):=−(ϕ+˜ρ)ψx−ϕ˜vx−ψ˜ρx,g2(u,ux):=−(ϕ+˜ρ)(ψ+˜v)ψx−(ϕψ+˜ρψ+˜vϕ)˜vx−(Pβ(ϕ+˜ρ)−Pβ(˜ρ))x+β2(Q(ϕ+˜ρ,ζ+˜J)−Q(˜ρ,˜J))x. |
Our purpose of this appendix is to show the local-in-time solvability of (A.1)–(A.3). We note that the first and third equations of (A.1) are transport equations of ϕ and ζ, and the second equation of (A.1) is a parabolic equation of ψ. In view of this consideration, we construct the sequence of functions {u(n)}∞n=0⊂Z1(T) with some positive T to approximate u by the iteration argument.
We first define u(0)=(ϕ(0),ψ(0),ζ(0)) by the following two steps. The first step is to take ψ(0) by solving the problem
b(ϕ0)ψ(0)t+Bψ(0)=g2(u0,u0,x),ψ(0)|t=0=ψ0,ψ(0)|x=∞=ψ(0)|x=0=0. | (A.4) |
The second step is to determine ϕ(0) and ζ(0) by solving the problem
ϕ(0)t+a(ψ(0))ϕ(0)x=g1(ϕ0,ψ(0),ψ(0)x),ζ(0)t+a(ψ(0))ζ(0)x=0,(ϕ(0),ζ(0))|t=0=(ϕ0,ζ0),(ϕ(0),ζ(0))|x=∞=(0,0) | (A.5) |
by using u0 and ψ(0). We next set u(n)=(ϕ(n),ψ(n),ζ(n)), for n∈N, inductively by the following process. Assuming that u(n−1) is obtained, we define u(n) as the solution of the problem
ϕt+a(ψ(n−1))ϕx=g1(ϕ(n−1),ψ(n−1),ψ(n−1)x),b(ϕ(n−1))ψt+Bψ=g2(u(n−1),u(n−1)x),ζt+a(ψ(n−1))ζx=0 | (A.6) |
with (A.2) and (A.3). If the obtained sequence {u(n)}∞n=0 has a limit u=(ϕ,ψ,ζ) in some function space, then it will be a solution to (A.1)–(A.3). To derive this fact, we prepare the following lemma.
Lemma A.1. Let u0=(ϕ0,ψ0,ζ0) satisfy the assumptions as in Proposition 5.1 with τ=0 except the smallness condition M0≤ε. Then the following assertions hold true.
(ⅰ) The pair of three functions u(0)=(ϕ(0),ψ(0),ζ(0)) solving (A.4) and (A.5) exists uniquely in Z1(T0) with some T0=T0(δ,M0). Furthermore, u(0) satisfies u(0)∈Z1C0M0(T0)∩X13M0(T0)3, (ϕ(0),ζ(0))∈C1([0,T0];L2(R+)), and
ϕ(0)(t,x)≥−34infx∈R+˜ρ(x). |
Here, C0=C0(δ,M0) is a certain positive constant increasing in δ,M0>0.
(ⅱ) Let n∈{0}∪N. Then, there exists T1=T1(δ,M0)∈(0,T0] independent of n such that (A.6) has a unique solution u(n)=(ϕ(n),ψ(n),ζ(n)) in Z1(T1) satisfying u(n)∈Z1C0M0(T1)∩X13M0(T1)3, (ϕ(n),ζ(n))∈C1([0,T1];L2(R+)),
inf(t,x)∈[0,T1]×R+ϕ(n)(t,x)≥−34infx∈R+˜ρ(x). |
Here T0 and C0 are same constants as in (ⅰ).
(ⅲ) Let n,m∈N. Then, there exists T2=T2(δ,M0)∈(0,T1] independent of n and m such that the following estimates hold:
‖u(n+1)−u(n)‖Z0(T2)≤C(M20+T2)‖u(n)−u(n−1)‖Z0(T2), | (A.7) |
‖ϕ(n)t−ϕ(m)t‖C([0,T2];H−1(R+))≤C(‖u(n)−u(m)‖Z0(T2)+‖u(n−1)−u(m−1)‖Z0(T2)). | (A.8) |
Here, T1 is the same constant as in (ⅱ) , and C=C(T2,δ,M0) is a certain positive constant increasing in T2, δ and M0.
Before proving Lemma A.1, we prepare the following lemmata which will play a crutial role.
Lemma A.2. Let T, M and m be positive constants, and let ψ0∈H10(R+). Then, the following facts are obtained.
(ⅰ) Let ¯u=(¯ϕ,¯ψ,¯ζ) be a given pair of functions satisfying
¯u∈X1M(T)3,¯ϕ∈C1([0,T];L2(R+)),¯ψ(t)∈H10(R+),inf(t,x)∈[0,T]×R+¯ϕ(t,x)≥(m−1)infx∈R+˜ρ(x) | (A.9) |
for 0≤t≤T. Then, there exists a unique solution ψ∈Y1(T) of the problem
b(¯ϕ)ψt+Bψ=g2(¯u,¯ux),ψ|t=0=ψ0,ψ|x=0=ψ|x=∞=0. | (A.10) |
Furthermore, ψ satisfies
‖ψ‖2X1(T)+C1(δ,M,m)∫t0(‖ψ(τ)‖2H2+‖ψt(τ)‖2L2)dτ≤eC1(δ,M,m)T(‖ψ0‖2H1+C1(δ,M,m)T). | (A.11) |
Here, C1(δ,M,m) is a positive constant taken from (2.7) in Lemma 2.6, and C1(δ,M,m) is a positive constant increasing in δ,M and decreasing in m.
(ⅱ) For j=1,2, let ¯u(j)=(¯ϕ(j),¯ψ(j),¯ζ(j)) be given pairs of functions satisfying (A.9), and let ψ(j)∈Y1(T) be solutions of (A.10) with ¯u=¯u(j) respectively. Then, ψ(1)−ψ(2) satisfies
‖ψ(1)−ψ(2)‖2Y0(T)≤C2(T,δ,M,m)(‖ψ(2)‖2Y1(T)+T)‖¯u(1)−¯u(2)‖2Z0(T). | (A.12) |
Here, C2(T,δ,M,m) is a positive constant increasing in T,δ,M and decreasing in m.
Lemma A.3. Let T, ˜M and M be positive constants, and let ϕ0,ζ0∈H1(R+). Then, the following facts are obtained.
(ⅰ) Let ¯u=(¯ϕ,¯ψ,¯ζ) be a given pair of functions satisfying ¯u∈Z1˜M(T)∩X1M(T)3. Then, there exists a unique pair of solutions (ϕ,ζ)∈X1(T)2 of the two problems
ϕt+a(¯ψ)ϕx=g1(¯ϕ,¯ψ,¯ψx),ϕ|t=0=ϕ0,ϕ|x=∞=0, | (A.13) |
and
ζt+a(¯ψ)ζx=0,ζ|t=0=ζ0,ζ|x=∞=0. | (A.14) |
Furthermore, ϕ and ζ satisfy ϕ,ζ∈C1([0,T];L2(R+)) and the following estimates
‖ϕ‖2X1(T)+‖ζ‖2X1(T)≤eC3(T,δ,˜M,M)√T(2‖ϕ0‖2H1+2‖ζ0‖2H1+C3(T,δ,˜M,M)√T), | (A.15) |
[2ex]inf(t,x)∈[0,T]×R+ϕ(t,x)≥infx∈R+ϕ0(x)−C3(T,δ,˜M,M)√T. | (A.16) |
Here, C3(T,δ,˜M,M) is a positive constant increasing in T,δ,˜M,M.
(ⅱ) For j=1,2, let ¯u(j)=(¯ϕ(j),¯ψ(j),¯ζ(j)) be given pairs of functions satisfying ¯u(j)∈Z1˜M(T)∩X1M(T)3, and let (ϕ(j),ζ(j))∈(X1(T))2 be pairs of solutions of (A.13) and (A.14) with ¯u=¯u(j) respectively.
Then, ϕ(1)−ϕ(2) and ζ(1)−ζ(2) satisfy
‖ϕ(1)−ϕ(2)‖2X0(T)+‖ζ(1)−ζ(2)‖2X0(T)≤C4(T,δ,˜M,M)T(‖ϕ(2)‖X1(T)+‖ζ(2)‖X1(T)+1)‖¯u(1)−¯u(2)‖2Z0(T), | (A.17) |
‖ϕ(1)t−ϕ(2)t‖2C([0,T];H−1(R+))≤C5(δ,M)‖ϕ(1)−ϕ(2)‖2X0(T)+C5(δ,M)(‖ϕ(2)‖X1(T)+1)‖¯u(1)−¯u(2)‖2Z0(T). | (A.18) |
Here, C4(T,δ,˜M,M) and C5(δ,M) are positive constants increasing in T,δ,˜M,M.
Proof of Lemma A.2. (ⅰ) We first note that g2(¯u,¯ux) belongs to L2(0,T;L2(R+)) by using (A.9). Then, applying Lemma 2.6 with ˜ϕ=¯ϕ, g=g2(¯u,¯ux) and k=1, the problem (A.10) has a unique solution ψ∈Y1(T) satisfying
‖ψ(t)‖2H1+C1(δ,M,m)∫t0(‖ψ(τ)‖2H2+‖ψt(τ)‖2L2)dτ≤‖ψ0‖2H1+C2(δ,M,m)∫t0(‖ψ(τ)‖2L2+‖g2(ˉu,ˉux)(τ)‖2L2)dτ |
for 0≤t≤T. Here C1(δ,M,m) and C2(δ,M,m) are taken in (2.7). We then use the Gronwall inequality to earn
‖ψ‖2X1(T)+C1(δ,M,m)∫T0(‖ψ(τ)‖2H2+‖ψt(τ)‖2L2)dτ≤eC2(δ,M,m)T(‖ψ0‖2H1+∫T0‖g2(ˉu,ˉux)(τ)‖2L2dτ). |
Therefore, since g2(¯u,¯ux) satisfies
∫T0‖g2(¯u,¯ux)(τ)‖2L2dτ≤C1,1(δ,M,m)T |
with some constant C1,1(δ,M,m) increasing in δ,M and decreasing in m, we arrive at the desired inequality (A.11). This completes the proof of (ⅰ).
(ⅱ) We first see from (A.10) that ψ(1)−ψ(2) is a weak solution of
b(¯ϕ(1))(ψ(1)−ψ(2))t+B(ψ(1)−ψ(2))=g2(¯u(1),¯u(1)x)−g2(¯u(2),¯u(2)x)−(¯ϕ(1)−¯ϕ(2))ψ(2)t,(ψ(1)−ψ(2))|t=0=0,(ψ(1)−ψ(2))|x=0=(ψ(1)−ψ(2))|x=∞=0. |
Applying (2.7) in Lemma 2.6 with k=0, ˜ϕ=¯ϕ(1), ψ=ψ(1)−ψ(2), g=g2(¯u(1),¯u(1)x)−g2(¯u(2),¯u(2)x)−(¯ϕ(1)−¯ϕ(2))ψ(2)t, we obtain
‖(ψ(1)−ψ(2))(t)‖2L2+¯c(M,m)∫t0‖(ψ(1)−ψ(2))(τ)‖2H1dτ≤C2,1(M,m)∫t0(‖((¯ϕ(1)−¯ϕ(2))ψ(2)t)(τ)‖2H−1+‖(g2(¯u(1),¯u(1)x)−g2(¯u(2),¯u(2)x))(τ)‖2H−1)dτ+C2,1(δ,M,m)∫t0‖(ψ(1)−ψ(2))(τ)‖2L2dτ, |
for 0≤t≤T, where C2,1(δ,M,m) is a positive constant increasing in δ,M and decreasing in m. We then use the Gronwall inequality to rewrite this inequality as
‖(ψ(1)−ψ(2))(t)‖2X0(T)+∫T0‖(ψ(1)−ψ(2))(τ)‖2H1dτ≤C2,1(δ,M,m)eC2,1(δ,M,m)T∫t0(‖((¯ϕ(1)−¯ϕ(2))ψ(2)t)(τ)‖2L2+‖(g2(¯u(1),¯u(1)x)−g2(¯u(2),¯u(2)x))(τ)‖2H−1)dτ. | (A.19) |
We next estimate the right-hand side of (A.19). The function (ˉϕ(1)−ˉϕ(2))ψ(2)t is controlled as
∫T0‖((¯ϕ(1)−¯ϕ(2))ψ(2)t)(τ)‖2L2dτ≤‖¯ϕ(1)−¯ϕ(2)‖2X0(T)∫t0‖ψ(2)t(τ)‖2L2dτ≤‖¯u(1)−¯u(2)‖2Z0(T)‖ψ(2)t‖2Y1(T). |
Using (A.9) with ¯u=¯u(j) with j=1,2 and integration by parts, g2(¯u(1),¯u(1)x)−g2(¯u(2),¯u(2)x) is estimated as
∫T0‖(g2(¯u(1),¯u(1)x)−g2(¯u(2),¯u(2)x))(τ)‖2H−1dτ≤C2,2(δ,M,m)T‖¯u(1)−¯u(2)‖2Z0(T), |
where C2,2(δ,M,m) is a positive constant increasing in δ,M and decreasing in m. Therefore, together with (A.19) and the estimates for (¯ϕ(1)−¯ϕ(2))ψ(2)t and g2(¯u(1),¯u(1)x)−g2(¯u(2),¯u(2)x), we earn (A.12). This completes the proof of (ⅱ).
Proof of Lemma A.3. (ⅰ) We first note that g1(¯ϕ,¯ψ,¯ψx) belongs to L2(0,T;H1(R+)) and a(¯ψ) satisfies the same assumptions for ˜ψ as in Lemma 2.5 by using the assumptions of ˉu. We then apply Lemma 2.5 with k=1, ˜ψ=¯ψ and f=g1(¯ϕ,¯ψ,¯ψx) to show that (A.13) has a unique solution ϕ∈X1(T) with ϕ∈C1([0,T];L2(R+)). As in this argument, (A.14) has a unique solution ζ∈X1(T) with ζ∈C1([0,T];L2(R+)) by replacing ϕ and g1(¯ϕ,¯ψ,¯ψx) as ζ and 0, respectively.
We next derive (A.15). It follows from Lemma 2.1 and (2.5) in Lemma 2.5 with k=1, ˜ψ=¯ψ, f=g1(¯ϕ,¯ψ,¯ψx), t1=0 and t2=t that ϕ satisfies
‖ϕ(t)‖2H1≤‖ϕ0‖2H1+C3,1(δ)∫T0‖g1(¯ϕ,¯ψ,¯ψx)‖H1‖ϕ(τ)‖H1dτ+C3,1(δ)∫t0(1+‖¯ψx(τ)‖H1)‖ϕ(τ)‖2H1dτ |
for 0≤t≤T, where C3,1(δ) is a positive constant increasing in δ. Applying the Gronwall inequality to the resultant inequality and using the fact ∫t0‖ˉψx(τ)‖H1dτ≤√T‖¯u‖Z1(T)≤√T˜M yields
‖ϕ(t)‖2H1≤eC3,2(T,δ,˜M)√T{‖ϕ0‖2H1+C3,1(δ)∫T0‖g1(¯ϕ,¯ψ,¯ψx)‖H1‖ϕ(τ)‖H1dτ} |
for 0≤t≤T, where C3,2(T,δ,˜M):=C3,1(δ)√T+˜M. Similarly, ζ is estimated as
‖ζ(t)‖2H1≤eC3,2(T,δ,˜M)√T‖ζ0‖2H1 |
for 0≤t≤T. We then add these inequalities to give
‖ϕ(t)‖2H1+‖ζ(t)‖2H1≤eC3,2(T,δ,˜M)√T{‖ϕ0‖2H1+‖ζ0‖2H1+C3,1(δ)∫T0‖g1(¯ϕ,¯ψ,¯ψx)‖H1‖ϕ(τ)‖H1dτ} |
for 0≤t≤T. To complete the derivation of (A.15), it remains to estimate the right-hand side of this inequality. From the assumption of ¯u, g1(¯ϕ,¯ψ,¯ψx) satisfies
∫T0‖g1(¯ϕ,¯ψ,¯ψx)‖H1dτ≤√T(∫T0‖g1(¯ϕ,¯ψ,¯ψx)‖2H1dτ)12≤C3,3(T,δ,˜M,M)√T, | (A.20) |
where C3,3(T,δ,˜M,M) is a positive constant increasing in T,δ,˜M,M>0. Hence (A.20) leads to
C3,1(δ)eC3,2(T,δ,˜M)√T∫T0‖g1(¯ϕ,¯ψ,¯ψx)(τ)‖H1‖ϕ(τ)‖H1dτ≤C3,1(δ)eC3,2(T,δ,˜M)√T(‖ϕ‖2X1(T)+‖ζ‖2X1(T))12∫T0‖g1(¯ϕ,¯ψ,¯ψx)(τ)‖H1dτ≤12(‖ϕ‖2X1(T)+‖ζ‖2X1(T))+C3,4(T,δ,˜M,M)eC3,2(T,δ,˜M)√TT, |
which yields (A.15). Here C3,4(T,δ,˜M,M) is a positive constant increasing in T,δ,˜M,M. We finally show (A.16) via a characteristic method. Let ¯y=¯y(τ;t,x)∈R+ be a unique solution of
d¯ydτ(τ;t,x)=a(¯ψ(τ,¯y(τ;t,x))),0≤τ≤t≤T,ˉy(t;t,x)=x. |
Decomposing ϕ(t,x)=ϕ0(¯y(0;t,x))+(ϕ(t,x)−ϕ0(¯y(0;t,x))) and using (2.6) in Lemma 2.5 and (A.20) immediately yields
inf(t,x)∈[0,T]×R+ϕ(x)≥infx∈R+ϕ0(x)−‖ϕ(t,⋅)−ϕ0(¯y(0;t,⋅))‖L∞≥infx∈R+ϕ0(x)−C3,3(T,δ,˜M,M)√T, |
which leads to (A.16). This completes the proof of (ⅰ).
(ⅱ) We first see from (A.13) that ϕ(1)−ϕ(2) is a weak solution of
(ϕ(1)−ϕ(2))t+a(¯ψ(1))(ϕ(1)−ϕ(2))x=−(¯ψ(1)−¯ψ(2))ϕ(2)x+g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x),(ϕ(1)−ϕ(2))|t=0=0,(ϕ(1)−ϕ(2))|x=∞=0. |
Applying (2.5) in Lemma 2.5 with k=0, ˜ψ=¯ψ, ϕ=ϕ(1)−ϕ(2), f=−(¯ψ(1)−¯ψ(2))ϕ(2)x+g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x), t1=0 and t2=T, we obtain
‖(ϕ(1)−ϕ(2))(t)‖2L2≤C4,1(δ){∫T0‖((¯ψ(1)−¯ψ(2))ϕ(2)x)(τ)‖L2‖(ϕ(1)−ϕ(2))(τ)‖L2dτ+∫T0‖(g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x))(τ)‖L2‖(ϕ(1)−ϕ(2))(τ)‖L2dτ+∫t0(1+‖¯ψ(1)x(τ)‖H1)‖(ϕ(1)−ϕ(2))(τ)‖2L2dτ} |
for 0≤t≤T, where C4,1(δ) is a positive constant depending only on δ. We then use the Gronwall inequality to read the resultant inequality as
‖(ϕ(1)−ϕ(2))(t)‖2L2≤C4,1(δ)eC4,2(T,δ,˜M)√T{∫T0‖((¯ψ(1)−¯ψ(2))ϕ(2)x)(τ)‖L2‖(ϕ(1)−ϕ(2))(τ)‖L2dτ+∫T0‖(g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x))(τ)‖L2‖(ϕ(1)−ϕ(2))(τ)‖L2dτ} | (A.21) |
for 0≤t≤T. Here C4,2(T,δ,˜M) is a positive constant increasing in T,δ,˜M. We next focus on the right-hand side of (A.21). We directly compute the term involving (¯ψ(1)−¯ψ(2))ϕ(2)x as
C4,1(δ)eC4,2(T,˜M)√T∫T0‖((¯ψ(1)−¯ψ(2))ϕ(2)x)(τ)‖L2‖(ϕ(1)−ϕ(2))(τ)‖L2dτ≤C4,1(δ)eC4,2(T,˜M)√T∫T0‖(¯ψ(1)−¯ψ(2))(τ)‖L∞dτ‖ϕ(2)x‖X1(T)‖ϕ(1)−ϕ(2)‖X0(T)≤C4,1(δ)eC4,2(T,˜M)√T√T‖¯ψ(1)−¯ψ(2)‖Y0(T)‖ϕ(2)x‖X1(T)‖ϕ(1)−ϕ(2)‖X0(T)≤14‖ϕ(1)−ϕ(2)‖2X0(T)+C4,3(T,˜M,M)T‖ϕ(2)x‖X1(T)‖¯u(1)−¯u(2)‖Z0(T), |
where C4,3(T,δ,˜M,M) denotes C4,3(T,δ,˜M,M):=2C4,1(δ)e2C4,2(T,δ,˜M)√T. For the term containing g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x), we use the assumptions of ˉu(j)(j=1,2) to deduce
∫T0‖(g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x))(τ)‖L2dτ≤C4,4(T,δ,˜M,M)√T‖¯u(1)−¯u(2)‖Z0(T), |
where C4,4(T,δ,˜M,M) is a positive constant increasing in T,δ,˜M,M. This estimate leads to
C4,1(δ)eC4,2(T,δ,˜M)√T∫T0‖(g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x))(τ)‖L2‖(ϕ(1)−ϕ(2))(τ)‖L2dτ≤C4,5(T,δ,˜M,M)∫T0‖(g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x))(τ)‖L2dτ‖ϕ(1)−ϕ(2)‖X0(T)≤C4,5(T,δ,˜M,M)√T‖¯u(1)−¯u(2)‖Z0(T)‖ϕ(1)−ϕ(2)‖X0(T)≤14‖ϕ(1)−ϕ(2)‖2X0(T)+C4,5(T,δ,˜M,M)T‖¯u(1)−¯u(2)‖Z0(T), |
where C4,5(T,δ,˜M,M) is a positive constant increasing in T,δ,˜M,M. Therefore, combining (A.21) and these estimates, we arrive at
‖ϕ(1)−ϕ(2)‖2X0(T)≤C4,6(T,δ,˜M,M)T(‖ϕ(2)x‖X1(T)+1)‖¯u(1)−¯u(2)‖2Z0(T). |
Here, C4,6(T,δ,˜M,M) is a positive constant increasing in T,δ,˜M,M. Similarly, the following estimate for ζ(1)−ζ(2) is obtained by replacing ϕ(j) and g1(¯ϕ(j),¯ψ(j),¯ψ(j)x) as ζ(j) and 0, respectively for j=1,2 in the above argument:
‖ζ(1)−ζ(2)‖2X0(T)≤C4,6(T,δ,˜M,M)T‖ζ(2)x‖X1(T)‖¯u(1)−¯u(2)‖2Z0(T). |
Hence, combining the above two estimates leads to (A.17). We finally prove (A.18) by using the equality
⟨ϕ(1)t−ϕ(2)t,φ⟩=(ϕ(1)−ϕ(2),(a(¯ψ(1))φ)x)L2−((¯ψ(1)−¯ψ(2))ϕ(2)x,φ)L2+(g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x),φ)L2 |
for any φ∈H10(R+) with ‖φ‖H1≤1. By virtue of ¯u(j)∈X1M(T)3,j=1,2 and integration by parts, each terms in the right-hand side of this equality are controlled as
|(g1(¯ϕ(1),¯ψ(1),¯ψ(1)x)−g1(¯ϕ(2),¯ψ(2),¯ψ(2)x),φ)L2|≤C5(δ,M)‖¯u(1)−¯u(2)‖Z0(T),|(ϕ(1)−ϕ(2),(a(¯ψ(1))φ)x)L2|≤C5(δ,M)‖ϕ(1)−ϕ(2)‖X0(T),|((¯ψ(1)−¯ψ(2))ϕ(2)x,φ)L2|≤C5(δ,M)‖ϕ(2)x‖X1(T2)‖¯u(1)−¯u(2)‖Z0(T) |
for 0≤t≤T, where C5(δ,M) is a positive constant increasing in δ,M. Therefore, using these estimates, we have (A.18) This completes the proof of (ⅱ).
Proof of Lemma A.1. (ⅰ) We first focus on the unique existence and properties of ψ(0). Let T0,1>0 be a constant taken suitably small later. Since u0 satisfies (A.9) in Lemma A.2 with ˉu=u0, T=T0,1, M=M0 and m=1/2, (A.4) has a unique solution ψ(0)∈Y1(T0,1). Furthermore, using (A.11) with T=T0,1, the following estimate for ψ=ψ(0) is obtained:
‖ψ(0)‖2X1(T0,1)+C1(δ,M0,12)∫T0,10(‖ψ(0)(τ)‖2H2+‖ψ(0)t(τ)‖2L2)dτ≤eC1(δ,M0,12)T0,1(M20+C1(δ,M0,12)T0,1). |
Therefore using the properties C1(δ,M0,1/2)≥C1(δ,3M0,1/4) and C1(δ,M0,1/2)≤C1(δ,3M0,1/4), and then letting T0,1=T0,1(δ,M0) small such that eC1(δ,3M0,1/4)T0,1(M20+C1(δ,3M0,1/4)T0,1)≤9M20, we have
‖ψ(0)‖2X1(T0,1)+C1(δ,3M0,14)∫T0,10(‖ψ(0)(τ)‖2H2+‖ψ(0)t(τ)‖2L2)dτ≤9M20, | (A.22) |
which yields ψ(0)∈Y1˜M0(T0,1)∩X13M0(T0,1). Here ˜M0 denotes ˜M0:=3{1+C1(δ,3M0,1/4)−1}1/2M0. We next investigate the existence and properties of (ϕ(0),ζ(0)). Let ˜u(0)=˜u(0)(t,x) be ˜u(0):=(ϕ0,ψ(0),ζ0) and let T0 be a constant T0∈(0,T0,1] determined later. Then in view of (A.22), ˜u(0) satisfies ˜u(0)∈Z1˜M0(T0)∩X13M0(T0)3. We then apply Lemma A.3 (ⅰ) with ¯u=˜u(0), T=T0, ˜M=˜M0 and M=3M0 to show that (A.5) has a unique solution (ϕ(0),ζ(0))∈X1(T0)2 satisfying ϕ(0),ζ(0)∈C1([0,T0];L2(R+)) and
‖ϕ(0)‖2X1(T0)+‖ζ(0)‖2X1(T0)≤e˜C3,0(T0,δ,M0)√T0(4M20+˜C3,0(T0,δ,M0)√T0),inf(t,x)∈[0,T0]×R+ϕ(0)(t,x)≥−12infx∈R+˜ρ(x)−˜C3,0(T0,δ,M0)√T0. |
Here, ˜C3,0(T0,δ,M0) is defined as ˜C3,0(T0,δ,M0):=C3(T0,δ,˜M0,3M0) by using C3(T,δ,˜M,M) defined in Lemma A.3 (ⅱ). We also note that ˜C3,0(T0,δ,M0) increases inT0,δ,M0. Therefore taking T0=T0(δ,M0) small such that e˜C3,0(T0,δ,M0)√T0(4M20+˜C3,0(T0,δ,M0)√T0)≤9M20 and ˜C3,0(T0,δ,M0)√T0≤(1/4)infx∈R+˜ρ(x) in the above inequalities, we have
‖ϕ(0)‖2X1(T0)+‖ζ(0)‖2X1(T0)≤9M20,inf(t,x)∈[0,T0]×R+ϕ(0)(t,x)≥−34infx∈R+˜ρ(x). | (A.23) |
As a result, adding (A.22) to (A.23) imply u(0)∈Z1C0M0(T0)∩X13M0(T0)3. Here C0=C0(δ,M0) is given by C0:=3{3+C1(δ,3M0,1/4)−1}1/2 increasing in δ,M0. This completes the proof of Lemma A.1 (ⅰ).
(ⅱ) We employ the induction argument with respect to n. The case n=0 is already true. We assume that the properties u(n−1)∈Z1C0M0(T1)∩X13M0(T1)3, ϕ(n−1),ζ(n−1)∈C1([0,T1];L2(R+)) and
inf(t,x)∈[0,T1]×R+ϕ(n−1)(t,x)≥−34infx∈R+˜ρ(x) |
hold with some n∈N and T1=T1(δ,M0)∈(0,T0] independent of n. Then using Lemma A.2 (ⅰ) and Lemma A.3 (ⅰ), u(n)=(ϕ(n),ψ(n),ζ(n)) is uniquely determined by solving (A.6) and satisfies u(n)∈Z1(T1) and ϕ(n),ζ(n)∈C1([0,T1];L2(R+)). In addition, it follows (A.11) and (A.15) with u=u(n), ¯u=u(n−1), T=T1, ˜M=C0M0, M=3M0 and m=1/4 that
‖ψ(n)‖2X1(T1)+C1(δ,3M0,14)∫T10(‖ψ(n)(τ)‖2H2+‖ψ(n)t(τ)‖2L2)dτ≤eC1(δ,3M0,14)T1(M20+C1(δ,3M0,14)T1),‖ϕ(n)‖2X1(T1)+‖ζ(n)‖2X1(T1)≤e˜C3(T1,δ,M0)√T1(4M20+˜C3(δ,T1,M0)√T1),inf(t,x)∈[0,T1]×R+ϕ(n)(t,x)≥−12infx∈R+˜ρ(x)−˜C3(T1,δ,M0)√T1. |
Here, ˜C3(T1,δ,M0) is defined as ˜C3(T1,δ,M0):=C3(T1,δ,C0M0,3M0) by using C3(T,δ,˜M,M) defined in Lemma A.3 (ⅰ). Therefore, taking T1=T1(δ,M0) suitably small without the dependence of n, u(n) also belongs to Z1C0M0(T1)∩X13M0(T1)3, and satisfies
inf(t,x)∈[0,T1]×R+ϕ(n)(t,x)≥−34infx∈R+˜ρ(x). |
As a result, the statement for n becomes true. This completes the proof of Lemma A.1 (ⅱ).
(ⅲ) We first estimate ψ(n+1)−ψ(n). Replacing (ψ(1),ψ(2)) and (¯u(1),¯u(2)) as (ψ(n+1),ψ(n)) and (u(n),u(n−1)) in Lemma A.2 (ⅱ) respectively, (A.12) is read as
‖ψ(n+1)−ψ(n)‖Y0(T2)≤˜C2(T2,δ,M0)(‖ψ(n)‖2Y1(T)+T2)‖u(n)−u(n−1)‖Z0(T2). |
Here, we set ˜C2(T2,δ,M0):=C2(T2,δ,3M0,1/4). Then, using u(n)∈Z1C0M0(T2) for all n∈N∪{0} to earn ‖ψ(n)‖Y1(T)≤C0M0, we have
‖ψ(n+1)−ψ(n)‖Y0(T2)≤C6(T2,δ,M0)(M20+T2)‖u(n)−u(n−1)‖Z0(T2), |
where C6(T2,δ,M0) is a positive constant increasing in T2,δ,M0. We next focus on ϕ(n+1)−ϕ(n) and ζ(n+1)−ζ(n). Changing (ϕ(1),ϕ(2)), (ζ(1),ζ(2)) and (ˉu(1),ˉu(2)) as (ϕ(n+1),ϕ(n)), (ζ(n+1),ζ(n)) and (u(n),u(n−1)) in (A.17) in Lemma A.3 (ⅱ), we have
‖ϕ(n+1)−ϕ(n)‖2X0(T2)+‖ζ(n+1)−ζ(n)‖2X0(T2)≤˜C4(T2,δ,M0)T2(‖ϕ(n)‖X1(T)+‖ζ(n)‖X1(T)+1)‖u(n)−u(n−1)‖2Z0(T2). |
Here, we put ˜C4(T2,δ,M0):=C4(T2,δ,C0M0,3M0). Then noticing from (ⅱ) of this lemma that u(n)∈X13M0(T2)3 holds for all n∈N, this inequality is rewritten as
‖ϕ(n+1)−ϕ(n)‖2X0(T2)+‖ζ(n+1)−ζ(n)‖2X0(T2)≤C7(T2,δ,M0)T2‖u(n)−u(n−1)‖Z0(T2), |
where C7(T2,δ,M0) is a positive constant increasing in T2,δ,M0. Consequently, combining the above estimates we arrive at (A.7). Similarly (A.8) is obtained by replacing (ϕ(1),ϕ(2)), (ˉu(1),ˉu(2)) and M as (ϕ(n),ϕ(m)), (u(n−1),u(m−1)) and 3M0, respectively in (A.17) in Lemma A.3 (ⅱ) to derive
‖ϕ(n)t−ϕ(m)t‖2C([0,T2];H−1(R+))≤C5(δ,3M0)‖ϕ(n)−ϕ(m)‖2X0(T2)+C5(δ,3M0)(‖ϕ(m)‖X1(T)+1)‖u(n)−u(m)‖2Z0(T2) |
for any n,m∈N, and then applying the fact u(m)∈X13M0(T2)3 for any m∈N. This completes the proof of Lemma A.1 (ⅲ).
Proof of Proposition 5.1. Let ε and T2 be the constants stated in Proposition 5.1 and Lemma A.1 (ⅲ). For simplicity, we rewrite T2 as T, and let ε and T be ε,T∈(0,1).
We first see from (A.7) in Lemma A.1 (ⅲ) that if u0 satisfies ‖u0‖H1≤M0≤ε, then {u(n)}∞n=0 is a Cauchy sequence in Z0(T), provided that ε and T=T(δ,M0) are chosen as C(1,δ,1)(ε2+T)<1, respectively. Furthermore, owing to this fact and (A.8), {ϕ(n)t}∞n=0 becomes a Cauchy sequence in C([0,T];H−1(R+)). Therefore {u(n)}∞n=0 has a limit u=(ϕ,ψ,ζ) such that
u(n)→uinZ0(T),ϕ(n)t→ϕtinC([0,T];H−1(R+)). | (A.24) |
Moreover, Lemma A.1 (ⅱ) and (A.24) lead to
g1(ϕ(n−1),ψ(n−1),ψ(n−1)x)→g1(ϕ,ψ,ψx)n→∞inL2(0,T;L2(R+)),g2(u(n−1),u(n−1)x)→g2(u,ux)inL2(0,T;H−1(R+)), |
and it follows from Lemma A.1 (ⅱ) that there exists a subsequence {u(nk)}∞k=0⊂{u(n)}∞n=0 satisfying
u(nk)∗⇀uweakly-∗inL∞(0,T;H1(R+)), | (A.25) |
ψ(nk)⇀ψweakly inL2(0,T;H2(R+))∩H1(0,T;L2(R+)). | (A.26) |
We are now going to claim that u is a solution of (A.1) in (0,T)×R+ with u∈Z1(T). We first investigate ϕ and ζ. Since the first equation of (A.6) is satisfied in C[0,T];L2(R+)) for n∈N, the following weak form holds for any φ∈C10([0,T)×R+):
−∫T0(ϕ(n),φt+(a(ψ(n−1))φ)x)L2dt=(ϕ0,φ(0))L2+∫T0(g1(ϕ(n−1),ψ(n−1),ψ(n−1)x),φ)L2dt. |
Therefore, it is not difficult to see from taking the limit n→∞ in this equality and using Lemma A.1 (ⅱ) and (A.24) that ϕ satisfies
−∫T0(ϕ,φt+(a(ψ)φ)x)L2dt=(ϕ0,φ(0))L2+∫T0(g1(ϕ,ψ,ψx),φ)L2dt. | (A.27) |
This means that ϕ is a weak solution of (2.1) with ˜ψ=ψ and f=g1(ϕ,ψ,ψx) in Definition 2.3.
Similarly, since ζ(n) solves the third equation of (A.6) for n∈N in C[0,T];L2(R+)), ζ satisfies
−∫T0(ζ,φt+(a(ψ)φ)x)L2dt=(ζ0,φ(0))L2. | (A.28) |
This implies that ζ is a weak solution of (2.1) with ˜ψ=ψ and f=0 in Definition 2.3.
We next check g1(ϕ,ψ,ψx)∈L2(0,T;H1(R+)) and ψ∈Y1(T), which imply ϕ,ζ∈X1(T) and ϕ,ζ∈C1([0,T];L2(R+)) by applying the regularity properties of weak solutions in Lemma 2.5. From (A.24)–(A.26), g1(ϕ,ψ,ψx)∈L2(0,T;H1(R+)) holds true. Moreover ψ∈Y1(T) is also satisfied by (A.24), (A.26) and C([0,T];L2(R+))∩L2(0,T;˜H2(R+))⊂C([0,T];H10(R+)). Therefore ϕ,ζ∈X1(T) and ϕ,ζ∈C1([0,T];L2(R+)) hold ture, and the properties ϕ,ζ∈X13M0(T) and inf(t,x)∈[0,T]×R+ϕ(t,x)≥−(3/4)infx∈R+˜ρ(x) follow from Lemma A.1 (ⅱ), (A.24) and (A.25). As a result, restricting φ∈C10((0,T)×R+) in (A.27) and (A.28), and then applying integration by parts, we see that u=(ϕ,ψ,ζ) satisfies the first and third equations of (A.1) in C([0,T];L2(R+)).
We next study ψ. Since ψ(n) is the solution of the second equation of (A.6) for n∈N, the following weak form holds for all φ∈H10(R+) and h∈C10([0,T)):
−∫T0(b(ϕ(n−1))ψ(n),φ)L2h′dt−∫T0⟨ϕ(n−1)tψ(n),φ⟩hdt+∫T0⟨Bψ(n),φ⟩hdt=(b(ϕ(0,⋅))ψ0,φ)L2h(0)+∫T0⟨g2(u(n−1),u(n−1)x),φ⟩hdt. |
Therefore taking the limit n→∞ in this equality and using Lemma A.1 (ⅱ) and (A.24), we obtain
−∫T0(b(ϕ)ψ,φ)L2h′dt−∫T0⟨ϕtψ,φ⟩hdt+∫T0⟨Bψ,φ⟩hdt=(b(ϕ(0,⋅))ψ0,φ)L2h(0)+∫T0⟨g2(u,ux),φ⟩hdt. | (A.29) |
This shows that ψ is a weak solution of (2.3) with ˜ϕ=ϕ and g=g2(u,ux) in Definition 2.4. The properties ψ∈Y1C0M0(T)∩X13M0(T) and g2(u,ux)∈L2(0,T;L2(R+)) are confirmed by Lemma A.1 (ⅱ), (A.25) and (A.26). From the above discussion, the assumptions for ˜ϕ=ϕ in Lemma 2.6 with M=3M0 and m=1/4 are also justified. Therefore, we gurantee these properties and then obtain the second equation of (A.1) in L2(0,T;L2(R+)) by restricting h∈C10(0,T) in (A.29) and applying integration by parts.
Consequently, we prove the local-in-time existence of the solution to (5.2) with (5.4)–(5.6) satisfying u∈X13M0(T), (5.8) and (5.9) with τ=0. The uniqueness of the solution is confirmed in a similar manner to the proof of Lemma A.1 (ⅲ). This completes the proof of Proposition 5.1.
[1] |
Y. Bai, T. Zhang, The pointwise estimates of solutions for the 3D compressible viscoelastic fluids, J. Differ. Equ., 356 (2023), 336–374. https://doi.org/10.1016/j.jde.2023.01.048 doi: 10.1016/j.jde.2023.01.048
![]() |
[2] | Q. Chen, G. Wu, The 3D compressible viscoelastic fluid in a bounded domain, Commun. Math. Sci., 16 (2018), 1303–1323. |
[3] | Y. I. Dimitrienko, Tensor analysis and nonlinear tensor functions, Berlin: Springer, 2002. |
[4] | M. Giga, A. Kirshtein, C. Liu, Variational modeling and complex fluids, In : Handbooks of Mathematical Analysis in Mechanics of Viscous Fluids, Berlin: Springer, 2018, 73–113. |
[5] | M. E. Gurtin, An introduction to continuum mechanics, London: Academic Press, 1981. |
[6] | A. Haruki, Y. Ishigaki, Stability of time-periodic parallel flow of compressible viscoelastic system in an infinite layer, Adv. Math. Sci. Appl., 30 (2021), 65–103. |
[7] |
X. Hu, D. Wang, Global existence for the multi-dimensional compressible viscoelastic flows, J. Differ. Equ., 250 (2011), 1200–1231. https://doi.org/10.1016/j.jde.2010.10.017 doi: 10.1016/j.jde.2010.10.017
![]() |
[8] |
X. Hu, G. Wu, Global existence and optimal decay rates for three-dimensional compressible viscoelastic flows, SIAM J. Math. Anal., 45 (2013), 2815–2833. https://doi.org/10.1137/120892350 doi: 10.1137/120892350
![]() |
[9] |
X. Hu, D. Wang, The initial-boundary value problem for the compressible viscoelastic flows, Discrete Contin. Dyn. Syst., 35 (2015), 917–934. https://doi.org/10.3934/dcds.2015.35.917 doi: 10.3934/dcds.2015.35.917
![]() |
[10] |
Y. Ishigaki, Global existence of solutions of the compressible viscoelastic fluid around a parallel flow, J. Math. Fluid Mech., 20 (2018), 2073–2104. https://doi.org/10.1007/s00021-018-0401-6 doi: 10.1007/s00021-018-0401-6
![]() |
[11] |
Y. Kagei, S. Kawashima, Local solvability of an initial boundary value problem for a quasilinear hyperbolic-parabolic system, J. Hyperbol. Differ. Equ., 3 (2006), 195–232. https://doi.org/10.1142/S0219891606000768 doi: 10.1142/S0219891606000768
![]() |
[12] |
Y. Kagei, S. Kawashima, Stability of planar stationary solutions to the compressible Navier-Stokes Equation on the half space, Commun. Math. Phys., 266 (2006), 401–430. https://doi.org/10.1007/s00220-006- doi: 10.1007/s00220-006-
![]() |
[13] |
S. Kawashima, S. Nishibata, P. Zhu, Asymptotic stability of the stationary solution to the compressible Navier-Stokes equations in the half space, Commun. Math. Phys., 240 (2003), 483–500. https://doi.org/10.1007/s00220-003-0909-2 doi: 10.1007/s00220-003-0909-2
![]() |
[14] |
S. Kawashima, P. Zhu, Asymptotic stability of nonlinear wave for the compressible Navier-Stokes equations in the half space, J. Differ. Equ., 244 (2008), 3151–3179. https://doi.org/10.1016/j.jde.2008.01.020 doi: 10.1016/j.jde.2008.01.020
![]() |
[15] |
S. Kawashima, P. Zhu, Asymptotic stability of rarefaction wave for the Navier-Stokes equations for a compressible fluid in the half space, Arch. Ration. Mech. Anal., 194 (2009), 105–132. https://doi.org/10.1007/s00205-008-0191-8 doi: 10.1007/s00205-008-0191-8
![]() |
[16] |
Y. Li, R. Wei, Z. Yao, Optimal decay rates for the compressible viscoelastic flows, J. Math. Phys., 57 (2016), 111506. https://doi.org/10.1063/1.4967975 doi: 10.1063/1.4967975
![]() |
[17] |
F. Lin, C. Liu, P. Zhang, On hydrodynamics of viscoelastic fluids, Commun. Pure Appl. Math., 58 (2005), 1437–1471. https://doi.org/10.1002/cpa.20074 doi: 10.1002/cpa.20074
![]() |
[18] | A. Matsumura, Inflow and outflow problems in the half space for a one-dimensional isentropic model system of compressible viscous gas, Meth. Appl. Anal., 8 (2001), 645–666. |
[19] |
A. Morando, Y. Trakhinin, P. Trebeschi, Structural stability of shock waves in 2D compressible elastodynamics, Math. Ann., 378 (2020), 1471–1504. https://doi.org/10.1007/s00208-019-01920-6 doi: 10.1007/s00208-019-01920-6
![]() |
[20] |
T. Nakamura, S. Nishibata, Convergence rate toward planar stationary waves for compressible viscous fluid in multidimensional half space, SIAM J. Math. Anal., 41 (2009), 1757–1791. https://doi.org/10.1137/090755357 doi: 10.1137/090755357
![]() |
[21] |
T. Nakamura, S. Nishibata, T. Yuge, Convergence rate of solutions toward stationary solutions to the compressible Navier-Stokes equations in a half line, J. Differ. Equ., 241 (2007), 94–111. https://doi.org/10.1016/j.jde.2007.06.016 doi: 10.1016/j.jde.2007.06.016
![]() |
[22] | T. Nakamura, Y. Ueda, S. Kawashima, Convergence rate toward degenerate stationary wave for compressible viscous gases, Yokohama Publ., 2010,239–248. |
[23] |
J. Qian, Initial boundary value problems for the compressible viscoelastic fluid, J. Differ. Equ., 250 (2011), 848–865. https://doi.org/10.1016/j.jde.2010.07.026 doi: 10.1016/j.jde.2010.07.026
![]() |
[24] |
J. Qian, Z. Zhang, Global well-posedness for compressible viscoelastic fluids near equilibrium, Arch. Ration. Mech. Anal., 198 (2010), 835–868. https://doi.org/10.1007/s00205-010-0351-5 doi: 10.1007/s00205-010-0351-5
![]() |
[25] | T. C. Sideris, B. Thomases, Global existence for 3D incompressible isotropic elastodynamics via the incompressible limit, Comm. Pure Appl. Math., 57 (2004), 1–39. |
[26] |
Y. Trakhinin, On weak stability of shock waves in 2D compressible elastodynamics, J. Hyperbol. Differ. Equ., 19 (2022), 157–173. https://doi.org/10.1142/S0219891622500035 doi: 10.1142/S0219891622500035
![]() |
[27] |
T. Wang, On the long-time behavior of solution for compressible Navier-Stokes system with outflow boundary condition, J. Differ. Equ., 323 (2022), 312–358. https://doi.org/10.1016/j.jde.2022.03.033 doi: 10.1016/j.jde.2022.03.033
![]() |