This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\UseRawInputEncoding

Subsonic flows with a contact discontinuity in a finitely long axisymmetric cylinder

Shangkun Weng School of mathematics and statistics, Wuhan University, Wuhan, Hubei Province, 430072, People’s Republic of China. Email: [email protected]    Zihao Zhang School of mathematics and statistics, Wuhan University, Wuhan, Hubei Province, 430072, People’s Republic of China. Email: [email protected]
Abstract

This paper concerns the structural stability of subsonic flows with a contact discontinuity in a finitely long axisymmetric cylinder. We establish the existence and uniqueness of axisymmetric subsonic flows with a contact discontinuity by prescribing the horizontal mass flux distribution, the swirl velocity, the entropy and the Bernoulli’s quantity at the entrance and the radial velocity at the exit. It can be formulated as a free boundary problem with the contact discontinuity to be determined simultaneously with the flows. Compared with the two-dimensional case, a new difficulty arises due to the singularity near the axis. One of the key points in the analysis is the introduction of an invertible modified Lagrangian transformation which can overcome this difficulty and straighten the contact discontinuity. Another one is to utilize the deformation-curl decomposition for the steady Euler system introduced in [17] to effectively decouple the hyperbolic and elliptic modes. Finally, the contact discontinuity will be located by using the implicit function theorem.

Mathematics Subject Classifications 2010: 35J15, 35L65, 76J25, 76N15.

Key words: contact discontinuity, structural stability, the modified Lagrangian transformation, the deformation-curl decomposition.

1 Introduction

In this paper, we are concerned with the structural stability of subsonic flows with a contact discontinuity governed by the three-dimensional steady full Euler system in a finitely long axisymmetric cylinder. The three-dimensional steady full Euler system for compressible inviscid gas is of the following form:

{x1(ρu1)+x2(ρu2)+x3(ρu3)=0,x1(ρu12)+x2(ρu1u2)+x3(ρu1u3)+x1P=0,x1(ρu1u2)+x2(ρu22)+x3(ρu2u3)+x2P=0,x1(ρu1u3)+x2(ρu2u3)+x3(ρu32)+x3P=0,x1(ρu1E+u1P)+x2(ρu2E+u2P)+x3(ρu3E+u3P)=0,\displaystyle\begin{cases}{\partial}_{x_{1}}(\rho u_{1})+{\partial}_{x_{2}}(\rho u_{2})+{\partial}_{x_{3}}(\rho u_{3})=0,\\ {\partial}_{x_{1}}(\rho u_{1}^{2})+{\partial}_{x_{2}}(\rho u_{1}u_{2})+{\partial}_{x_{3}}(\rho u_{1}u_{3})+{\partial}_{x_{1}}P=0,\\ {\partial}_{x_{1}}(\rho u_{1}u_{2})+{\partial}_{x_{2}}(\rho u_{2}^{2})+{\partial}_{x_{3}}(\rho u_{2}u_{3})+{\partial}_{x_{2}}P=0,\\ {\partial}_{x_{1}}(\rho u_{1}u_{3})+{\partial}_{x_{2}}(\rho u_{2}u_{3})+{\partial}_{x_{3}}(\rho u_{3}^{2})+{\partial}_{x_{3}}P=0,\\ {\partial}_{x_{1}}(\rho u_{1}E+u_{1}P)+{\partial}_{x_{2}}(\rho u_{2}E+u_{2}P)+{\partial}_{x_{3}}(\rho u_{3}E+u_{3}P)=0,\\ \end{cases} (1.1)

where 𝒖=(u1,u2,u3)\bm{u}=(u_{1},u_{2},u_{3}) is the velocity, ρ\rho is the density, PP is the pressure, EE is the energy, respectively. For polytropic gas, the equation of state and the energy are of the form

P=A(S)ργ,andE=12|𝒖2|+P(γ1)ρ,P=A(S)\rho^{\gamma},\quad{\rm{and}}\quad E=\frac{1}{2}|{\bm{u}}^{2}|+\frac{P}{(\gamma-1)\rho},

where A(S)=ReSA(S)=Re^{S} and γ(1,+)\gamma\in(1,+\infty), RR are positive constants. Denote the Bernoulli’s function and the local sonic speed by B=12|u|2+γP(γ1)ρB=\frac{1}{2}|{\textbf{u}}|^{2}+\frac{\gamma P}{(\gamma-1)\rho} and c(ρ,A)=Aγργ12c(\rho,A)=\sqrt{A\gamma}\rho^{\frac{\gamma-1}{2}}, respectively. Then the system (1.1) is hyperbolic for supersonic flows (|u|>c(ρ,A)|\textbf{u}|>c(\rho,A)), and hyperbolic-elliptic coupled for subsonic flows (|u|<c(ρ,A)|\textbf{u}|<c(\rho,A)).

To understand the contact discontinuity surface, we first give the definition of steady flows with a contact discontinuity. Let 𝒟3\mathcal{D}\in\mathbb{R}^{3} be an open and connected domain. Suppose that a non-self-intersecting C1C^{1}-curve Γ\Gamma divides 𝒟\mathcal{D} into two disjoint open subsets 𝒟±\mathcal{D}^{\pm} such that 𝒟=𝒟Γ𝒟+\mathcal{D}=\mathcal{D}^{-}\cup\Gamma\cup\mathcal{D}^{+}. Assume that 𝑼=(ρ,u1,u2,u3,P)\bm{U}=(\rho,u_{1},u_{2},u_{3},P) satisfies the following properties:

  1. (1)

    𝑼[L(𝒟)Cloc1(𝒟±)Cloc0(𝒟±Γ)]5\bm{U}\in[L^{\infty}(\mathcal{D})\cap C^{1}_{loc}(\mathcal{D}^{\pm})\cap C^{0}_{loc}(\mathcal{D}^{\pm}\cup\Gamma)]^{5};

  2. (2)

    For any ηC0(𝒟)\eta\in C_{0}^{\infty}(\mathcal{D}),

    {𝒟(ρu1x1η+ρu2x2η+ρu3x3η)d𝐱=0,𝒟((ρu12+P)x1η+ρu1u2x2η+ρu1u3x3η)d𝐱=0,𝒟(ρu1u2x1η+(ρu22+P)x2η+ρu2u3x3η)d𝐱=0,𝒟(ρu1u3x1η+ρu2u3x2η+(ρu32+P)x3η)d𝐱=0,𝒟(ρu1(E+Pρ)x1η+ρu2(E+Pρ)x2η+ρu3(E+Pρ)x3η)d𝐱=0.\begin{cases}\int_{\mathcal{D}}(\rho u_{1}{\partial}_{x_{1}}\eta+\rho u_{2}{\partial}_{x_{2}}\eta+\rho u_{3}{\partial}_{x_{3}}\eta){\mathrm{d}}\mathbf{x}=0,\\ \int_{\mathcal{D}}((\rho u_{1}^{2}+P){\partial}_{x_{1}}\eta+\rho u_{1}u_{2}{\partial}_{x_{2}}\eta+\rho u_{1}u_{3}{\partial}_{x_{3}}\eta){\mathrm{d}}\mathbf{x}=0,\\ \int_{\mathcal{D}}(\rho u_{1}u_{2}{\partial}_{x_{1}}\eta+(\rho u_{2}^{2}+P){\partial}_{x_{2}}\eta+\rho u_{2}u_{3}{\partial}_{x_{3}}\eta){\mathrm{d}}\mathbf{x}=0,\\ \int_{\mathcal{D}}(\rho u_{1}u_{3}{\partial}_{x_{1}}\eta+\rho u_{2}u_{3}{\partial}_{x_{2}}\eta+(\rho u_{3}^{2}+P){\partial}_{x_{3}}\eta){\mathrm{d}}\mathbf{x}=0,\\ \int_{\mathcal{D}}(\rho u_{1}(E+\frac{P}{\rho}){\partial}_{x_{1}}\eta+\rho u_{2}(E+\frac{P}{\rho}){\partial}_{x_{2}}\eta+\rho u_{3}(E+\frac{P}{\rho}){\partial}_{x_{3}}\eta){\mathrm{d}}\mathbf{x}=0.\\ \end{cases} (1.2)

By integration by parts, we get the Rankine-Hugoniot conditions:

{n1[ρu1]+n2[ρu2]+n2[ρu3]=0,n1[ρu12]+n1[P]+n2[ρu1u2]+n3[ρu1u3]=0,n1[ρu1u2]+n2[ρu22]+n2[P]+n3[ρu2u3]=0,n1[ρu1u3]+n2[ρu2u3]+n3[ρu32]+n3[P]=0,n1[ρu1(E+Pρ)]+n2[ρu2(E+Pρ)]+n3[ρu3(E+Pρ)]=0,\begin{cases}n_{1}[\rho u_{1}]+n_{2}[\rho u_{2}]+n_{2}[\rho u_{3}]=0,\\ n_{1}[\rho u_{1}^{2}]+n_{1}[P]+n_{2}[\rho u_{1}u_{2}]+n_{3}[\rho u_{1}u_{3}]=0,\\ n_{1}[\rho u_{1}u_{2}]+n_{2}[\rho u_{2}^{2}]+n_{2}[P]+n_{3}[\rho u_{2}u_{3}]=0,\\ n_{1}[\rho u_{1}u_{3}]+n_{2}[\rho u_{2}u_{3}]+n_{3}[\rho u_{3}^{2}]+n_{3}[P]=0,\\ n_{1}[\rho u_{1}(E+\frac{P}{\rho})]+n_{2}[\rho u_{2}(E+\frac{P}{\rho})]+n_{3}[\rho u_{3}(E+\frac{P}{\rho})]=0,\\ \end{cases} (1.3)

where 𝐧=(n1,n2,n3)\mathbf{n}=(n_{1},n_{2},n_{3}) is the unit normal vector to Γ\Gamma, and [F](𝐱)=F+(𝐱)F(𝐱)[F](\mathbf{x})=F_{+}(\mathbf{x})-F_{-}(\mathbf{x}) denotes the jump across the surface Γ\Gamma for a piecewise smooth function FF.

Let 𝝉1=(τ11,τ21,τ31)\bm{\tau}_{1}=(\tau_{11},\tau_{21},\tau_{31}) and 𝝉2=(τ12,τ22,τ32)\bm{\tau}_{2}=(\tau_{12},\tau_{22},\tau_{32}) as the unit tangential vectors to Γ\Gamma, which means that 𝐧𝝉k=0\mathbf{n}\cdot{\bm{\tau}_{k}}=0. Taking the dot product of ((1.3)2,(1.3)3,(1.3)4)(\eqref{1-3}_{2},\eqref{1-3}_{3},\eqref{1-3}_{4}) with 𝐧\mathbf{n} and 𝝉k\bm{\tau}_{k} respectively, one has

[ρ(𝐮𝐧)2+P]Γ=0,ρ(𝐮𝐧)[𝐮𝝉k]Γ=0,fork=1,2.[\rho(\mathbf{u}\cdot\mathbf{n})^{2}+P]_{\Gamma}=0,\quad\rho(\mathbf{u}\cdot\mathbf{n})[\mathbf{u}\cdot\bm{\tau}_{k}]_{\Gamma}=0,\quad{\rm{for}}\quad k=1,2. (1.4)

Assume that ρ>0\rho>0 in 𝒟¯\bar{\mathcal{D}}, (1.4) implies either 𝐮𝐧=0\mathbf{u}\cdot\mathbf{n}=0 on Γ\Gamma or [𝐮𝝉k]Γ=0[\mathbf{u}\cdot\bm{\tau}_{k}]_{\Gamma}=0. If 𝐮𝐧0\mathbf{u}\cdot\mathbf{n}\neq 0 and [𝐮𝝉k]Γ=0[\mathbf{u}\cdot\bm{\tau}_{k}]_{\Gamma}=0 hold on Γ\Gamma, the surface Γ\Gamma is called a shock; if the flow moves along both sides of Γ\Gamma such that 𝐮𝐧=0\mathbf{u}\cdot\mathbf{n}=0 on Γ\Gamma, the surface Γ\Gamma is called a contact discontinuity. In the latter case, 𝐮𝐧=0\mathbf{u}\cdot\mathbf{n}=0 and the first equation in (1.4) give [P]=0[P]=0. Then we get the R-H conditions corresponding to a contact discontinuity as follows:

𝐮𝐧=0and[P]=0onΓ.\mathbf{u}\cdot\mathbf{n}=0\quad{\rm{and}}\quad[P]=0\quad{\rm{on}}\ \Gamma. (1.5)
Definition 1.1.

We define 𝑼=(ρ,u1,u2,u3,P)\bm{U}=(\rho,u_{1},u_{2},u_{3},P) to be a weak solution of the full Euler system (1.1) in 𝒟\mathcal{D} with a contact discontinuity Γ\Gamma if the the following properties hold:

  1. (i)

    Γ\Gamma is a non-self-intersecting C1C^{1}-curve dividing 𝒟\mathcal{D} into two disjoint open subsets 𝒟±\mathcal{D}^{\pm} such that 𝒟=𝒟Γ𝒟+\mathcal{D}=\mathcal{D}^{-}\cup\Gamma\cup\mathcal{D}^{+};

  2. (ii)

    𝑼\bm{U} satisfies (1)\rm{(1)} and (2)\rm{(2)};

  3. (iii)

    ρ>0\rho>0 in 𝒟¯\bar{\mathcal{D}};

  4. (iv)

    (𝐮|𝒟¯Γ𝐮|𝒟¯+Γ)𝟎(\mathbf{u}|_{\bar{\mathcal{D}}_{-}\cap\Gamma}-\mathbf{u}|_{\bar{\mathcal{D}}_{+}\cap\Gamma})\neq\mathbf{0} holds for all 𝐱Γ\mathbf{x}\in\Gamma;

  5. (v)

    𝐮𝐧=0\mathbf{u}\cdot\mathbf{n}=0 and [P]=0[P]=0 on Γ\Gamma.

This is a continuous work on the study of subsonic Euler flows with a contact discontinuity in a finitely long nozzle. In the previous work [20], we established the existence and uniqueness of subsonic flows with a contact discontinuity in a two-dimensional finitely long slightly curved nozzle. As an attempt to extend the results from [20] to the three dimensional case, this paper investigates the structural stability of subsonic flows with a contact discontinuity in a finitely long axisymmetric cylinder under the suitable axisymmetric perturbations of boundary conditions.

The study on the steady compressible flows with a contact discontinuity is not only of fundamental importance in developing the mathematical theory of partial differential equations arising from fluid dynamics, but also has important applications to engineering designs, such as rocket launching, sharp charged jet and so on. Up to now, there have been many works in the literature on the steady flows with a contact discontinuity. For the subsonic flow, the stability of flat contact discontinuity in infinite nozzles was established in [2] and [3, 4]. The authors in [3, 4] decompose the Rankine-Hugoniot conditions on the contact discontinuity via Helmholtz decomposition so that the compactness of approximated solutions can be achieved. The uniqueness and existence of the contact discontinuity in infinitely long nozzles, which is not a perturbation of the flat contact discontinuity, was obtained in [8]. The existence, uniqueness and stability of subsonic flows past an airfoil with a vortex line were obtained in [5]. The key idea in [5] is to use the implicit function theorem as the framework to solve the problem of subsonic flows past an airfoil with a vortex line. Inspired by [5], the stability of contact discontinuity in a finite nozzle was established in [20] by using the implicit function theorem. For the supersonic flow, the stability of flat contact discontinuity in finitely long nozzles was studied in [11]. The stability of three-dimensional supersonic contact discontinuity was investigated in [15, 16]. Recently, the stability of supersonic contact discontinuity for the two-dimensional steady rotating Euler system in a finitely curved nozzle has been established in [21]. For the transonic flow, the stability of flat contact discontinuity in finitely long nozzles was established in [12]. The stability of two-dimensional transonic contact discontinuity over a solid wedge and three-dimensional transonic contact discontinuity were established in [6, 7] and [14].

We make some comments on the new ingredients in our analysis for the contact discontinuity problem. Note that the contact discontinuity is part of the solution and is unknown, thus this is a free boundary that separates both subsonic flows in the inner and outer layers of the cylinder. The strategy to overcome this difficulty in the two dimensional case [20] is to introduce a Lagrangian transformation to straighten the contact discontinuity. The idea also applies to three dimensional steady axisymmetric Euler system. However, in the three dimensional axisymmetric setting, there is a singular term rr in the density equation. Inspired by [19], the singular term rr in the density equation is of order O(r)O(r) near the axis r=0r=0, hence we can find a simple modified Lagrangian transformation such that it is invertible near the axis and also straightens the contact discontinuity. Another key issue is to decompose the hyperbolic and elliptic modes in the steady axisymmetric Euler system. It is well-known that the steady axisymmetric Euler system is hyperbolic-elliptic mixed in subsonic regions, whose effective decomposition of elliptic and hyperbolic modes is crucial for developing a well-defined iteration. Here we will use the deformation-curl decomposition introduced in [17, 18] to effectively decouple the hyperbolic and elliptic modes in subsonic regions.

The other key ingredient in our analysis is to employ the implicit function theorem to locate the contact discontinuity. The idea is inspired by the discussion of the airfoil problem in [5]. We choose a suitable Hölder space and design a proper map to verify the conditions in the implicit function theorem. However, it seems quite difficult to verify that the isomorphism of the differential of the map for general background flows with a straight contact discontinuity. Here we choose the background outer-layer flow is stagnant and restrict the perturbation only on the entrance of the inner-outer flow. In this case, the outer-layer flow is fixed and one can prove the isomorphism, the contact discontinuity can be located by the implicit function theorem.

This paper will be arranged as follows. In Section 2, we formulate the problem of subsonic flows with a contact discontinuity in a finitely long axisymmetric cylinder and state the main result. In Section 3, the modified Lagrange transformation is employed to straighten the contact discontinuity and reformulate the free boundary value problem 2.2. Then we use the deformation-curl decomposition in [17, 18] to derive an equivalent system. Finally, we state the main steps to solve the free boundary problem 3.1. In Section 4, we first linearize the nonlinear system and solve the linear system in a suitable weighted Ho¨\ddot{\rm{o}}lder space. Then the framework of the contraction mapping theorem can be used to find the solution of the nonlinear system. In Section 5, we choose a suitable Ho¨\ddot{\rm{o}}lder space and design a proper map to verify the conditions in the implicit function theorem. Then by using the implicit function theorem, we locate the contact discontinuity. In Section 6, we finish the proof of the main theorem.

2 Mathematical formulation of the problem

In this section, we first construct a special class of subsonic Euler flows with a straight contact discontinuity in a finitely long axisymmetric cylinder. Then we give a detailed formulation of the stability problem for these background flows with a contact discontinuity and state the main result.

2.1 The background solutions

The axisymmetric cylinder (Fig 1) of the length LL is given by

𝒩:={(x1,x2,x3)3:0<x1<L, 0x22+x32<1}.\mathcal{N}:=\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}:0<x_{1}<L,\ 0\leq x_{2}^{2}+x_{3}^{2}<1\}.
Refer to caption
Fig 1: Subsonic flows with a contact discontinuity in an axisymmetric cylinder

Consider two layers of steady smooth Euler flows separated by the cylindrical surface r=x22+x32=12r=\sqrt{x_{2}^{2}+x_{3}^{2}}=\frac{1}{2} satisfying the following properties:

  1. (i)

    The velocity and density of the outer and inner layers are given by (0,0,0),ρb+(0,0,0),\rho_{b}^{+} and (ub,0,0),ρb(u_{b}^{-},0,0),\rho_{b}^{-}, where ub>0u_{b}^{-}>0 and ρb±>0\rho_{b}^{\pm}>0;

  2. (ii)

    the pressure of both the outer and inner layers is given by the same positive constant PbP_{b};

  3. (iii)

    the flows in the outer and inner layers are subsonic, i.e.,

    (ub)2<γPbρb.(u_{b}^{-})^{2}<\frac{\gamma P_{b}}{\rho_{b}^{-}}.

Then

𝑼b={𝑼b+:=(ρb+,0,0,0,Pb),for12<r<1,𝑼b:=(ρb,ub,0,0,Pb),for0r<12,\bm{U}_{b}=\begin{cases}\bm{U}_{b}^{+}:=(\rho_{b}^{+},0,0,0,P_{b}),\quad&{\rm{for}}\quad\frac{1}{2}<r<1,\\ \bm{U}_{b}^{-}:=(\rho_{b}^{-},u_{b}^{-},0,0,P_{b}),\quad&{\rm{for}}\quad 0\leq r<\frac{1}{2},\\ \end{cases} (2.1)

with a contact discontinuity on the surface r=12r=\frac{1}{2} satisfy the steady Euler system (1.1) in the sense of Definition 1.1, which will be called the background solutions in this paper. This paper is going to establish the structural stability of these background solutions under the suitable axisymmetric perturbations of boundary conditions.

2.2 The stability problem and the main result

Let (x,r,θ)(x,r,\theta) be the cylindrical coordinates of (x1,x2,x3)3(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}, that is

x=x1,r=x22+x32,θ=arctanx3x2.x=x_{1},\ r=\sqrt{x_{2}^{2}+x_{3}^{2}},\ \theta=\arctan\frac{{{x_{3}}}}{{{x_{2}}}}.

Any function f(𝐱)f({\bf x}) can be represented as f(𝐱)=f(x,r,θ)f({\bf x})=f(x,r,\theta), and a vector-valued function 𝐡(𝐱){\bf h}({\bf x}) can be represented as 𝐡(𝐱)=hx(x,r,θ)𝐞x+hr(x,r,θ)𝐞r+hθ(x,r,θ)𝐞θ{\bf h}({\bf x})=h_{x}(x,r,\theta)\,{\bf e}_{x}+h_{r}(x,r,\theta)\,{\bf e}_{r}+h_{\theta}(x,r,\theta)\,{\bf e}_{\theta}, where

𝐞x=(1,0,0),𝐞r=(0,cosθ,sinθ),𝐞θ=(0,sinθ,cosθ).{\bf e}_{x}=(1,0,0),\quad{\bf e}_{r}=(0,\cos\theta,\sin\theta),\quad{\bf e}_{\theta}=(0,-\sin\theta,\cos\theta).

We say that a function f(𝐱)f({\bf x}) is axisymmetric if its value is independent of θ\theta and that a vector-valued function 𝐡=(hx,hr,hθ){\bf h}=(h_{x},h_{r},h_{\theta}) is axisymmetric if each of functions hx(𝐱),hr(𝐱)h_{x}({\bf x}),h_{r}({\bf x}) and hθ(𝐱)h_{\theta}({\bf x}) is axisymmetric.

Assume that

ρ(x)\displaystyle\rho({\textbf{x}}) =ρ(x,r),P(x)=P(x,r),A(x)=A(x,r),\displaystyle=\rho(x,r),\quad P({\textbf{x}})=P(x,r),\quad A({\textbf{x}})=A(x,r),
𝐮(x)\displaystyle{\bf u}({\textbf{x}}) =ux(x,r)𝐞x+ur(x,r)𝐞r+uθ(x,r)𝐞θ,\displaystyle=u_{x}(x,r){\bf e}_{x}+u_{r}(x,r){\bf e}_{r}+u_{\theta}(x,r){\bf e}_{\theta},

then (1.1) can be rewritten as

{x(ρux)+r(ρur)+ρurr=0,ρ(uxx+urr)ux+xP=0,ρ(uxx+urr)urρuθ2r+rP=0,ρ(uxx+urr)(ruθ)=0,ρ(uxx+urr)A=0.\begin{cases}\begin{aligned} &{\partial}_{x}(\rho u_{x})+{\partial}_{r}(\rho u_{r})+\frac{\rho u_{r}}{r}=0,\\ &\rho(u_{x}{\partial}_{x}+u_{r}{\partial}_{r})u_{x}+{\partial}_{x}P=0,\\ &\rho(u_{x}{\partial}_{x}+u_{r}{\partial}_{r})u_{r}-\frac{\rho u_{\theta}^{2}}{r}+{\partial}_{r}P=0,\\ &\rho(u_{x}{\partial}_{x}+u_{r}{\partial}_{r})(ru_{\theta})=0,\\ &\rho(u_{x}{\partial}_{x}+u_{r}{\partial}_{r})A=0.\\ \end{aligned}\end{cases} (2.2)

The axis and boundary of the cylinder are denoted by Γa\Gamma_{a} and Γw\Gamma_{w}, i.e;

Γa:={(x,r)2:0<x<L,r=0},Γw:={(x,r)2:0<x<L,r=1}.\Gamma_{a}:=\{(x,r)\in\mathbb{R}^{2}:0<x<L,r=0\},\quad\Gamma_{w}:=\{(x,r)\in\mathbb{R}^{2}:0<x<L,r=1\}. (2.3)

The exit of the cylinder is denoted by

ΓL:={(x,r):x=L, 0r<1}.\Gamma_{L}:=\{(x,r):x=L,\ 0\leq r<1\}. (2.4)

The entrance of the cylinder is separated into two parts:

Γ0+:={(x,r):x=0,12<r<1},Γ0:={(x,r):x=0, 0r<12}.\displaystyle\Gamma_{0}^{+}:=\{(x,r):x=0,\ \frac{1}{2}<r<1\},\quad\Gamma_{0}^{-}:=\{(x,r):x=0,\ 0\leq r<\frac{1}{2}\}. (2.5)

At the entrance, we prescribe the boundary data for the horizontal mass distribution J=ρuxJ=\rho u_{x}, the swirl velocity uθu_{\theta}, the entropy AA and the Bernoulli’s quantity BB:

(J,uθ,A,B)(0,r)={(0,0,Ab+,Bb+),onΓ0+,(J0,ν0,A0,B0)(r),onΓ0,(J,u_{\theta},A,B)(0,r)=\begin{cases}(0,0,A_{b}^{+},B_{b}^{+}),\quad&{\rm{on}}\quad\Gamma_{0}^{+},\\ (J_{0},\nu_{0},A_{0},B_{0})(r),\quad&{\rm{on}}\quad\Gamma_{0}^{-},\\ \end{cases} (2.6)

where

Ab+=Pb(ρb+)γ,Bb+=γPb(γ1)ρb+,A_{b}^{+}=\frac{P_{b}}{(\rho_{b}^{+})^{\gamma}},\quad B_{b}^{+}=\frac{\gamma P_{b}}{(\gamma-1)\rho_{b}^{+}},

and functions (J0,ν0,A0,B0)(r)(C1,α([0,12]))4(J_{0},\nu_{0},A_{0},B_{0})(r)\in\left(C^{1,\alpha}([0,\frac{1}{2}])\right)^{4} are close to the background solutions in some sense that will be clarified later. Moreover, the compatibility conditions hold:

ν0(0)=r(ν0,A0,B0)(0)=0,\nu_{0}(0)={\partial}_{r}(\nu_{0},A_{0},{B}_{0})(0)=0, (2.7)

since (ν0,A0,B0)(r)(\nu_{0},A_{0},B_{0})(r) are C1,αC^{1,\alpha} in Γ0\Gamma_{0}^{-}. At the exit, the following boundary condition is satisfied:

𝐮𝐞r=0,onΓL.{\bf u}\cdot{\bf e}_{r}=0,\quad{\rm{on}}\quad\Gamma_{L}. (2.8)

We expect the flow in the cylinder will be separated by a contact discontinuity Γ:={r=gcd(x),0<x<L}\Gamma:=\{r=g_{cd}(x),0<x<L\} with gcd(0)=12g_{cd}(0)=\frac{1}{2}, which divides the domain 𝒩\mathcal{N} into the subsonic and subsonic regions:

𝒩+:=𝒩{gcd(x)<r<1},𝒩:=𝒩{0r<gcd(x)}.\mathcal{N}^{+}:=\mathcal{N}\cap\{g_{cd}(x)<r<1\},\quad\mathcal{N}^{-}:=\mathcal{N}\cap\{0\leq r<g_{cd}(x)\}. (2.9)

Let

𝑼(x,r)={𝑼+(x,r):=(ρ+,ux+,ur+,uθ+,P+)(x,r)in𝒩+,𝑼(x,r):=(ρ,ux,ur,uθ,P)(x,r)in𝒩.{\bm{U}}(x,r)=\begin{cases}{\bm{U}}^{+}(x,r):=(\rho^{+},u_{x}^{+},u_{r}^{+},u_{\theta}^{+},P^{+})(x,r)\quad{\rm{in}}\quad\mathcal{N}^{+},\\ {\bm{U}}^{-}(x,r):=(\rho^{-},u_{x}^{-},u_{r}^{-},u_{\theta}^{-},P^{-})(x,r)\quad{\rm{in}}\quad\mathcal{N}^{-}.\\ \end{cases}

Along the contact discontinuity r=gcd(x)r=g_{cd}(x), the following Rankine-Hugoniot conditions hold:

𝐮𝐧cd=0,P+=P,onΓ,{\bf u}\cdot{\bf n}_{cd}=0,\quad P^{+}=P^{-},\quad{\rm{on}}\quad\Gamma, (2.10)

where

𝐧cd=gcd(x)𝐞x+𝐞r1+|gcd(x)|2.{\bf n}_{cd}=\frac{-g_{cd}^{\prime}(x){\bf e}_{x}+{\bf e}_{r}}{\sqrt{1+|g_{cd}^{\prime}(x)|^{2}}}.

On the nozzle wall Γw\Gamma_{w} , the flow satisfies the slip condition u+n+=0\textbf{u}^{+}\cdot\textbf{n}^{+}=0, where n+\textbf{n}^{+} is the outer normal vector of the nozzle wall. Using cylindrical coordinates, the slip boundary condition can be rewritten as

ur+(x,1)=0,onΓw.u_{r}^{+}(x,1)=0,\quad{\rm{on}}\quad\Gamma_{w}. (2.11)

Moreover, since the flow is smooth near the axis Γa\Gamma_{a}, thus we have the following compatibility conditions:

ur(x,0)=uθ(x,0)=0,x[0,L].u_{r}^{-}(x,0)=u_{\theta}^{-}(x,0)=0,\quad\forall x\in[0,L]. (2.12)

In summary, we will investigate the following problem:

Problem 2.1.

Given functions (J0,ν0,A0,B0)(r)(J_{0},\nu_{0},A_{0},B_{0})(r) at the entrance satisfying (2.7), find a unique piecewise smooth axisymmetric subsonic solution (𝐔+,𝐔)(\bm{U}^{+},\bm{U}^{-}) defined on 𝒩+\mathcal{N}^{+} and 𝒩\mathcal{N}^{-} respectively, with the contact discontinuity Γ\Gamma satisfying the axisymmetric Euler system (2.2) in the sense of Definition 1.1 and the Rankine-Hugoniot conditions in (2.10) and the slip boundary condition in (2.11) and the compatibility conditions (2.12).

It is easy to see that 𝑼b+\bm{U}_{b}^{+} satisfies the following properties:

  • ρb+>0\rho_{b}^{+}>0 and 0<γPbρb+0<\sqrt{\frac{\gamma P_{b}}{\rho_{b}^{+}}};

  • Ab+=Pb(ρb+)γA_{b}^{+}=\frac{P_{b}}{(\rho_{b}^{+})^{\gamma}} and Bb+=γPb(γ1)ρb+B_{b}^{+}=\frac{\gamma P_{b}}{(\gamma-1)\rho_{b}^{+}};

  • 𝟎𝒗=0\bm{0}\cdot\bm{v}=0 for any vector 𝒗2\bm{v}\in\mathbb{R}^{2}.

From this observation, we fix 𝑼+(x1,x2)=𝑼b+{\bm{U}}^{+}(x_{1},x_{2})=\bm{U}_{b}^{+} in 𝒩+\mathcal{N}^{+} and solve the following free boundary value problem:

Problem 2.2.

Under the assumptions of Problem 2.1, find a smooth axisymmetric subsonic solution 𝐔\bm{U}^{-} defined on 𝒩\mathcal{N}^{-} with the contact discontinuity Γ:r=gcd(x)\Gamma:r=g_{cd}(x) such that the following hold.

  1. (a)

    gcd(0)=12g_{cd}(0)=\frac{1}{2}.

  2. (b)

    The flow has positive density in the inner cylinder, i.e, ρ>0\rho^{-}>0.

  3. (c)

    Along the contact discontinuity r=gcd(x)r=g_{cd}(x), the following Rankine-Hugoniot conditions hold:

    𝐮𝐧cd=0,P=Pb,onΓ.{\bf u}^{-}\cdot{\bf n}_{cd}=0,\quad P^{-}=P_{b},\quad{\rm{on}}\quad\Gamma. (2.13)
  4. (d)

    On the axis Γa\Gamma_{a}, the following compatibility conditions hold:

    ur(x,0)=uθ(x,0)=0,x[0,L].u_{r}^{-}(x,0)=u_{\theta}^{-}(x,0)=0,\quad\forall x\in[0,L]. (2.14)
  5. (e)

    On the exit, the following boundary condition holds:

    ur(L,r)=0.u_{r}^{-}(L,r)=0. (2.15)
Refer to caption
Fig 2: Problem 2.2

Before we state our main result, some weighted Hölder norms are first introduced: For a bounded connected open set 𝒟3\mathcal{D}\in\mathbb{R}^{3}, let 𝒫\mathcal{P} be a closed portion of 𝒟{\partial}\mathcal{D}. For 𝐱,𝐱~𝒟\mathbf{x},\tilde{\mathbf{x}}\in\mathcal{D}, define

δ𝐱:=dist(𝐱,𝒫),δx,x~:=min{δ𝐱,δ𝐱~},\displaystyle\delta_{\mathbf{x}}:=\rm{dist}(\mathbf{x},\mathcal{P}),\quad\delta_{\textbf{x},\tilde{\textbf{x}}}:=\min\{\delta_{\mathbf{x}},\delta_{\tilde{\mathbf{x}}}\},

Given positive integer mm, α(0,1)\alpha\in(0,1) and kk\in\mathbb{R}, we define

um,0;𝒟(k,𝒫)\displaystyle{\|u\|}_{m,0;\mathcal{D}}^{(k,\mathcal{P})} :=0|β|msup𝒙𝒟δ𝐱max(|β|+k,0)|Dβu(𝒙)|;\displaystyle:=\sum_{0\leq|\beta|\leq m}\sup_{\bm{x}\in\mathcal{D}}\delta_{\mathbf{x}}^{\max(|\beta|+k,0)}|D^{\beta}u(\bm{x})|;
[u]m,α;𝒟(k,𝒫)\displaystyle[u]_{m,\alpha;\mathcal{D}}^{(k,\mathcal{P})} :=|β|=msup𝒙,𝒙~𝒟,𝒙𝒙~δ𝐱,𝐱~max(m+α+k,0)|Dβu(𝒙)Dβu(𝒙~)||𝒙𝒙~|α;\displaystyle:=\sum_{|\beta|=m}\sup_{\bm{x},\tilde{\bm{x}}\in\mathcal{D},\bm{x}\neq\tilde{\bm{x}}}\delta_{\mathbf{x},\tilde{\mathbf{x}}}^{\max(m+\alpha+k,0)}\frac{|D^{\beta}u({\bm{x}})-D^{\beta}u(\tilde{\bm{x}})|}{|\bm{x}-\tilde{\bm{x}}|^{\alpha}};
um,α;𝒟(k,𝒫)\displaystyle\|u\|_{m,\alpha;\mathcal{D}}^{(k,\mathcal{P})} :=um,0;𝒟k,𝒫+[u]m,α;𝒟k,𝒫\displaystyle:=\|u\|_{m,0;\mathcal{D}}^{k,\mathcal{P}}+[u]_{m,\alpha;\mathcal{D}}^{k,\mathcal{P}}

with the corresponding function space defined as

Cm,α(k,𝒫)(𝒟)={u:um,α;𝒟(k,𝒫)<}.C_{m,\alpha}^{(k,\mathcal{P})}(\mathcal{D})=\{u:\|u\|_{m,\alpha;\mathcal{D}}^{(k,\mathcal{P})}<\infty\}.

For a vector function 𝒖=(u1,u2,,un)\bm{u}=(u_{1},u_{2},\cdots,u_{n}), define

𝒖m,α;𝒟(k,𝒫):=i=1nuim,α;𝒟(k,𝒫).\|\bm{u}\|_{m,\alpha;\mathcal{D}}^{(k,\mathcal{P})}:=\sum_{i=1}^{n}\|u_{i}\|_{m,\alpha;\mathcal{D}}^{(k,\mathcal{P})}.

The main theorem of this paper can be stated as follows.

Theorem 2.3.

Given functions (J0,ν0,A0,B0)(r)(J_{0},\nu_{0},A_{0},B_{0})(r) at the entrance satisfying (2.7) and α(12,1)\alpha\in(\frac{1}{2},1), we define

σ(J0,ν0,A0,B0):\displaystyle\sigma(J_{0},\nu_{0},A_{0},B_{0}): =(J0,ν0,A0,B0)(Jb,0,Ab,Bb1,α;[0,12],\displaystyle=\|(J_{0},\nu_{0},A_{0},B_{0})-(J_{b}^{-},0,A_{b}^{-},B_{b}^{-}\|_{1,\alpha;[0,\frac{1}{2}]}, (2.16)

where

Jb=ρbub,Ab=Pb(ρb)γ,Bb=12(ub)2+γPb(γ1)ρb.J_{b}^{-}=\rho_{b}^{-}u_{b}^{-},\quad A_{b}^{-}=\frac{P_{b}}{(\rho_{b}^{-})^{\gamma}},\quad B_{b}^{-}=\frac{1}{2}(u_{b}^{-})^{2}+\frac{\gamma P_{b}}{(\gamma-1)\rho_{b}^{-}}.

There exist positive constants σ1\sigma_{1} and 𝒞\mathcal{C}^{\ast} depending only on (𝐔b,L,α)(\bm{U}_{b}^{-},L,\alpha) such that if

σ(J0,ν0,A0,B0)σ1,\displaystyle\sigma(J_{0},\nu_{0},A_{0},B_{0})\leq\sigma_{1}, (2.17)

𝐏𝐫𝐨𝐛𝐥𝐞𝐦 2.2\mathbf{Problem\ 2.2} has a unique smooth axisymmetric subsonic solution 𝐔\bm{U}^{-} with the contact discontinuity Γ:r=gcd(x)\Gamma:r=g_{cd}(x) satisfying the following properties:

  1. (i)

    The axisymmetric subsonic solution 𝑼C1,α(α,Γ)(𝒩)\bm{U}^{-}\in C_{1,\alpha}^{(-\alpha,\Gamma)}(\mathcal{N}^{-}) satisfies the following estimate:

    𝑼𝑼b1,α;𝒩(α,Γ)𝒞σ(J0,ν0,A0,B0).\|\bm{U}^{-}-\bm{U}_{b}^{-}\|_{1,\alpha;\mathcal{N}^{-}}^{(-\alpha,\Gamma)}\leq\mathcal{C}^{\ast}\sigma(J_{0},\nu_{0},A_{0},B_{0}). (2.18)
  2. (ii)

    The contact discontinuity surface gcd(x)C1,α([0,L])g_{cd}(x)\in C^{1,\alpha}([0,L]) satisfies gcd(0)=12g_{cd}(0)=\frac{1}{2}. Furthermore, it holds that

    gcd121,α;[0,L]𝒞σ(J0,ν0,A0,B0).\|g_{cd}-\frac{1}{2}\|_{1,\alpha;[0,L]}\leq\mathcal{C}^{\ast}\sigma(J_{0},\nu_{0},A_{0},B_{0}). (2.19)
Remark 2.4.

There are several differences between our result and the previous work [4]. The first one is that the boundary conditions imposed on the entrance and exit of the cylinder. We prescribe the boundary data for the horizontal mass distribution, the swirl velocity, the entropy and the Bernoulli’s quantity at the entrance, while [4] prescribes the entropy, the swirl velocity and the radial velocity. The second one is that the decomposition of the axisymmetric Euler system. In [4], the Helmholtz decomposition of the velocity field plays a crucial role. Instead, we utilize the deformation-curl decomposition developed in [17] for steady Euler system to effectively decouple the hyperbolic and elliptic modes. The last one is that the approach to locate contact discontinuity. The contact discontinuity in [4] is determined by an ordinary differential equation arising from the Rankine-Hugoniot conditions. In this paper, we employ the implicit function theorem to locate the contact discontinuity.

Remark 2.5.

In 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 2.2\mathbf{Problem\ 2.2}, by fixing the outer-layer flow in 𝒩+\mathcal{N}^{+} as the background flow (ρb+,0,0,0,Pb)(\rho_{b}^{+},0,0,0,P_{b}), we seek a smooth axisymmetric subsonic solution 𝑼\bm{U}^{-} defined on 𝒩\mathcal{N}^{-} with the contact discontinuity Γ:r=gcd(x)\Gamma:r=g_{cd}(x). One can also find a smooth axisymmetric subsonic solution 𝑼+\bm{U}^{+} defined on 𝒩+\mathcal{N}^{+} with the contact discontinuity Γ:r=gcd(x)\Gamma:r=g_{cd}(x) by fixing the inner-layer flow in 𝒩\mathcal{N}^{-} as the background flow (ρb,0,0,0,Pb)(\rho_{b}^{-},0,0,0,P_{b}). In fact, this case is even simpler than 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 2.2\mathbf{Problem\ 2.2} since the singularity near the axis is not needed to be considered. Thus we can introduce the usual Lagrangian transformation and reduce the axisymmetric Euler system to a second order elliptic equation for the stream function as in [20] to obtain the solution 𝑼+\bm{U}^{+}.

3 The reformulation of Problem 2.2

In this section, we first introduce the modified Lagrange transformation to straighten the contact discontinuity and reformulate the free boundary value problem 2.2. Then the deformation-curl decomposition in [17, 18] is employed to derive an equivalent system. Finally, we state the main steps to solve the free boundary problem 3.1.

3.1 Reformulation by the modified Lagrangian transformation

For steady Euler flows, the main advantage of the Euler-Lagrange coordinate transformation is to straighten the stream lines. However, in the three-dimensional axisymmetric setting, there is a singular term rr in the density equation. We introduce the modified Lagrange transformation to overcome this difficulty and apply this modified Lagrange transformation to straighten the contact discontinuity.

Let (𝑼(x,r),gcd(x))(\bm{U}^{-}(x,r),g_{cd}(x)) be a solution to 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 2.2\mathbf{Problem\ 2.2}. Define

m2=012sJ0(s)ds.m^{2}=\int_{0}^{\frac{1}{2}}sJ_{0}(s){\mathrm{d}}s. (3.1)

For any x(0,L)x\in(0,L), it follows from the first equation in (2.2) that

0gcd(x)sρux(x,s)ds=m2.\int_{0}^{g_{cd}(x)}s\rho^{-}u_{x}^{-}(x,s){\mathrm{d}}s=m^{2}. (3.2)

Define the modified Lagrangian transformation as

y1=x,y2(x,r)=(0rsρux(x,s)d.s)12.y_{1}=x,\quad y_{2}(x,r)=\left(\int_{0}^{r}s\rho^{-}u_{x}^{-}(x,s){\mathrm{d}}.s\right)^{\frac{1}{2}}. (3.3)

Note that if (ρ,ux,ur,uθ)(\rho^{-},u_{x}^{-},u_{r}^{-},u_{\theta}^{-}) is close to the background solutions (ρb,ub,0,0)(\rho_{b}^{-},u_{b}^{-},0,0), there exist two positive constants C1C_{1} and C2C_{2}, depending on the background solution, such that

C1r20rsρux(x,s)dsC2r2.C_{1}r^{2}\leq\int_{0}^{r}s\rho^{-}u_{x}^{-}(x,s){\mathrm{d}}s\leq C_{2}r^{2}.

Hence the Jacobian of the modified Lagrange transformation satisfies

(y1,y2)(x,r)=|10rρur2y2rρux2y2|=rρux2y2C3>0.\frac{{\partial}(y_{1},y_{2})}{{\partial}(x,r)}=\left|\begin{matrix}1&0\\ -\frac{r\rho^{-}u_{r}^{-}}{2y_{2}}&\frac{r\rho^{-}u_{x}^{-}}{2y_{2}}\end{matrix}\right|=\frac{r\rho^{-}u_{x}^{-}}{2y_{2}}\geq C_{3}>0. (3.4)

That is invertible.

Under this transformation, the domain 𝒩\mathcal{N}^{-} becomes

Ω:={(y1,y2):0<y1<L, 0<y2<m}.\Omega:=\{(y_{1},y_{2}):0<y_{1}<L,\ 0<y_{2}<m\}.

The entrance and exit of Ω\Omega are defined as

Σ0:={(y1,y2):y1=0, 0<y2<m},ΣL:={(y1,y2):y1=L, 0<y2<m}.\displaystyle\Sigma_{0}:=\{(y_{1},y_{2}):y_{1}=0,\ 0<y_{2}<m\},\quad\Sigma_{L}:=\{(y_{1},y_{2}):y_{1}=L,\ 0<y_{2}<m\}.

The axis Γa\Gamma_{a} is transformed into

Σa:={(y1,y2):0<y1<L,y2=0}.\displaystyle\Sigma_{a}:=\{(y_{1},y_{2}):0<y_{1}<L,\ y_{2}=0\}.

Moreover, on Γ\Gamma, one has

y2(x,gcd(x))=(0gcd(x)sρux(x,s)ds)12=m.y_{2}(x,g_{cd}(x))=\left(\int_{0}^{g_{cd}(x)}s\rho^{-}u_{x}^{-}(x,s){\mathrm{d}}s\right)^{\frac{1}{2}}=m.

Hence the free boundary Γ\Gamma becomes the following fixed straight line

Σ:={(y1,y2):0<y1<L,y2=m}.\Sigma:=\{(y_{1},y_{2}):0<y_{1}<L,\ y_{2}=m\}. (3.5)

In the following, the superscript “-” in ρ,ux,ur,uθ,P\rho^{-},u_{x}^{-},u_{r}^{-},u_{\theta}^{-},P^{-} will be ignored to simplify the notations. Let

𝑼(y1,y2):=(ρ,ux,ur,uθ,P)(y1,r(y1,y2))inΩ.{\bm{U}}^{-}(y_{1},y_{2}):=(\rho,u_{x},u_{r},u_{\theta},P)(y_{1},r(y_{1},y_{2}))\quad{\rm{in}}\quad\Omega. (3.6)

Then the axisymmetric Euler system (2.2) in the new coordinates can be rewritten as

{y1(2y2rρux)y2(urux)=0,y1(ux+Pρux)r2y2y2(Purux)Purrρux2=0,y1ur+r2y2y2Puθ2rux=0,y1(ruθ)=0,y1A=0.\begin{cases}\begin{aligned} &{\partial}_{y_{1}}\left(\frac{2y_{2}}{r\rho u_{x}}\right)-{\partial}_{y_{2}}\left(\frac{u_{r}}{u_{x}}\right)=0,\\ &{\partial}_{y_{1}}\left(u_{x}+\frac{P}{\rho u_{x}}\right)-\frac{r}{2y_{2}}{\partial}_{y_{2}}\left(\frac{Pu_{r}}{u_{x}}\right)-\frac{Pu_{r}}{r\rho u_{x}^{2}}=0,\\ &{\partial}_{y_{1}}u_{r}+\frac{r}{2y_{2}}{\partial}_{y_{2}}P-\frac{u_{\theta}^{2}}{ru_{x}}=0,\\ &{\partial}_{y_{1}}(ru_{\theta})=0,\\ &{\partial}_{y_{1}}A=0.\\ \end{aligned}\end{cases} (3.7)

The background solutions in the Lagrange coordinates are

𝑼b:=(ρb,ub,0,0,Pb),inΩb,{\bm{U}}_{b}^{-}:=(\rho_{b}^{-},u_{b}^{-},0,0,P_{b}),\quad{\rm{in}}\quad\Omega_{b}, (3.8)

where Ωb:={(y1,y2):0<y1<L, 0<y2<mb}\Omega_{b}:=\{(y_{1},y_{2}):0<y_{1}<L,\ 0<y_{2}<m_{b}^{-}\} and mb=ρbub8m_{b}^{-}=\sqrt{\frac{\rho_{b}^{-}u_{b}^{-}}{8}}. Without loss of generality, we assume that

ρbub=2andm=mb=12.\rho_{b}^{-}u_{b}^{-}=2\quad{\rm{and}}\quad m=m_{b}^{-}=\frac{1}{2}.

Furthermore, under the modified Lagrangian transformation, rr as a function of (y1,y2)(y_{1},y_{2}) becomes nonlinear and nonlocal in the new coordinates. In fact, it follows from the inverse transformation that

ry1=urux,ry2=2y2rρux,r(y1,0)=0.\displaystyle\frac{{\partial}r}{{\partial}y_{1}}=\frac{u_{r}}{u_{x}},\quad\frac{{\partial}r}{{\partial}y_{2}}=\frac{2y_{2}}{r\rho u_{x}},\quad r(y_{1},0)=0.

Thus one derives

r(y1,y2)=(20y22sρux(y1,s)ds)12,inΩ.r(y_{1},y_{2})=\left(2\int_{0}^{y_{2}}\frac{2s}{\rho u_{x}(y_{1},s)}{\mathrm{d}}s\right)^{\frac{1}{2}},\ {\rm{in}}\quad\Omega.\\ (3.9)

In particular, for the background solutions (ρb,ub,0,0)(\rho_{b}^{-},u_{b}^{-},0,0), one has

rb(y2)=2ρbuby2=y2,inΩ.r_{b}(y_{2})=\sqrt{{\frac{2}{\rho_{b}^{-}u_{b}^{-}}}}y_{2}=y_{2},\ {\rm{in}}\quad\Omega. (3.10)

In the new coordinates, the boundary data (2.6) at the entrance is given by

(J,uθ,A,B)(0,y2)=(J~0,ν~0,A~0,B~0)(y2),onΣ0,(J,u_{\theta},A,B)(0,y_{2})=(\tilde{J}_{0},\tilde{\nu}_{0},\tilde{A}_{0},\tilde{B}_{0})(y_{2}),\quad{\rm{on}}\quad\Sigma_{0}, (3.11)

where

(J~0,ν~0,A~0,B~0)(y2)=(J0,ν0,A0,B0)((0y22sJ0(s)ds)12).(\tilde{J}_{0},\tilde{\nu}_{0},\tilde{A}_{0},\tilde{B}_{0})(y_{2})=(J_{0},\nu_{0},A_{0},B_{0})\left(\left(\int_{0}^{y_{2}}\frac{2s}{J_{0}(s)}{\mathrm{d}}s\right)^{\frac{1}{2}}\right).

The Rankine-Hugoniot conditions in (2.13) become

urux(y1,12)=gcd(y1),\frac{u_{r}}{u_{x}}(y_{1},\frac{1}{2})=g_{cd}^{\prime}(y_{1}), (3.12)

and

P(y1,12)=Pb.P(y_{1},\frac{1}{2})=P_{b}. (3.13)

3.2 The deformation-curl decomposition for axisymmetric Euler system

It is well-known that the steady Euler system is elliptic-hyperbolic coupled in subsonic region, to construct a well-defined iteration scheme, one should decompose the hyperbolic and elliptic modes effectively. Different from the pervious decomposition in [20], we will employ the deformation-curl decomposition developed in [17, 18] to deal with the elliptic-hyperbolic coupled structure in the axisymmetric Euler system.

First, using the Bernoulli’s function, the density ρ\rho can be represented as

ρ=H(B,A,|u|2)=(γ1Aγ(B12|u|2))1γ1.\rho=H(B,A,|\textbf{u}|^{2})=\left(\frac{\gamma-1}{A\gamma}(B-\frac{1}{2}|\textbf{u}|^{2})\right)^{\frac{1}{\gamma-1}}. (3.14)

Define the vorticity ω=curlu=ωx𝐞x+ωr𝐞r+ωθ𝐞θ\omega={\rm curl\,}\textbf{u}=\omega_{x}\mathbf{e}_{x}+\omega_{r}\mathbf{e}_{r}+\omega_{\theta}\mathbf{e}_{\theta}, where

ωx=1rr(ruθ),ωr=xuθ,ωθ=xurrux.\omega_{x}=\frac{1}{r}{\partial}_{r}(ru_{\theta}),\ \omega_{r}=-{\partial}_{x}u_{\theta},\ \omega_{\theta}={\partial}_{x}u_{r}-{\partial}_{r}u_{x}.

From the third equation in (2.2) and the Bernoulli’s law, one derives that

ωθ=uθruθ+uθ2r+(B12|𝐮|2)AγrArBux.\omega_{\theta}=\frac{u_{\theta}{\partial}_{r}u_{\theta}+\frac{u_{\theta}^{2}}{r}+\frac{(B-\frac{1}{2}|{\bf u}|^{2})}{A\gamma}{\partial}_{r}A-{\partial}_{r}B}{u_{x}}. (3.15)

Substituting (3.14) into the density equation in (2.2), the axisymmetric Euler system (2.2) is equivalent to the following system:

{(c2(H,A)ux2)xux+(c2(H,A)ur2)ruxuxur(xur+rux)+urc2(H,A)+uθ2r=0,ux(xurrux)=uθruθ+uθ2r+(B12|𝐮|2)AγrArB,(uxx+urr)(ruθ)=0,(uxx+urr)A=0,(uxx+urr)B=0.\begin{cases}\begin{aligned} &(c^{2}(H,A)-u_{x}^{2}){\partial}_{x}u_{x}+(c^{2}(H,A)-u_{r}^{2}){\partial}_{r}u_{x}-u_{x}u_{r}({\partial}_{x}u_{r}+{\partial}_{r}u_{x})+u_{r}\frac{c^{2}(H,A)+u_{\theta}^{2}}{r}=0,\\ &u_{x}({\partial}_{x}u_{r}-{\partial}_{r}u_{x})=u_{\theta}{\partial}_{r}u_{\theta}+\frac{u_{\theta}^{2}}{r}+\frac{(B-\frac{1}{2}|{\bf u}|^{2})}{A\gamma}{\partial}_{r}A-{\partial}_{r}B,\\ &(u_{x}{\partial}_{x}+u_{r}{\partial}_{r})(ru_{\theta})=0,\\ &(u_{x}{\partial}_{x}+u_{r}{\partial}_{r})A=0,\\ &(u_{x}{\partial}_{x}+u_{r}{\partial}_{r})B=0.\\ \end{aligned}\end{cases} (3.16)

Under the transformation (3.3), AA and BB satisfy the following transport equations:

{y1A=0,y1B=0.\begin{cases}\begin{aligned} &{\partial}_{y_{1}}A=0,\\ &{\partial}_{y_{1}}B=0.\\ \end{aligned}\end{cases} (3.17)

Thus one has

A(y1,y2)=A~0(y2),B(y1,y2)=B~0(y2).A(y_{1},y_{2})=\tilde{A}_{0}(y_{2}),\quad B(y_{1},y_{2})=\tilde{B}_{0}(y_{2}). (3.18)

Next, it follows from (3.3) and (3.18) that uxu_{x}, uru_{r} and uθu_{\theta} satisfy the following system:

{(c2(ρ,A~0)ux2)(y1uxrρur2y2y2ux)+(c2(ρ,A~0)ur2)(rρux2y2y2ur)+c2(ρ,A~0)rur+uθ2rur=uxur(y1urrρur2y2y2ur+rρux2y2y2ux),ux(y1urrρur2y2y2urrρux2y2y2ux)=rρux2y2uθy2uθ+uθ2r+rρux2y2ργ1γ1y2A~0rρux2y2y2B~0,y1(ruθ)=0,\begin{cases}\begin{aligned} &(c^{2}(\rho,\tilde{A}_{0})-u_{x}^{2})\left({\partial}_{y_{1}}u_{x}-\frac{r\rho u_{r}}{2y_{2}}{\partial}_{y_{2}}u_{x}\right)+(c^{2}(\rho,\tilde{A}_{0})-u_{r}^{2})\left(\frac{r\rho u_{x}}{2y_{2}}{\partial}_{y_{2}}u_{r}\right)\\ &+\frac{c^{2}(\rho,\tilde{A}_{0})}{r}u_{r}+\frac{u_{\theta}^{2}}{r}u_{r}=u_{x}u_{r}\left({\partial}_{y_{1}}u_{r}-\frac{r\rho u_{r}}{2y_{2}}{\partial}_{y_{2}}u_{r}+\frac{r\rho u_{x}}{2y_{2}}{\partial}_{y_{2}}u_{x}\right),\\ &u_{x}\left({\partial}_{y_{1}}u_{r}-\frac{r\rho u_{r}}{2y_{2}}{\partial}_{y_{2}}u_{r}-\frac{r\rho u_{x}}{2y_{2}}{\partial}_{y_{2}}u_{x}\right)\\ &=\frac{r\rho u_{x}}{2y_{2}}u_{\theta}{\partial}_{y_{2}}u_{\theta}+\frac{u_{\theta}^{2}}{r}+\frac{r\rho u_{x}}{2y_{2}}\frac{\rho^{\gamma-1}}{\gamma-1}{\partial}_{y_{2}}\tilde{A}_{0}-\frac{r\rho u_{x}}{2y_{2}}{\partial}_{y_{2}}\tilde{B}_{0},\\ &{\partial}_{y_{1}}(ru_{\theta})=0,\\ \end{aligned}\end{cases} (3.19)

with the following boundary conditions:

{ρux(0,y2)=J~0(y2),uθ(0,y2)=ν~0(y2),onΣ0,ur(L,y2)=0,onΣL,ur(y1,12)=ux(y1,12)gcd(y1),onΣ,ur(y1,0)=0,onΣa.\begin{cases}\rho u_{x}(0,y_{2})=\tilde{J}_{0}(y_{2}),\ u_{\theta}(0,y_{2})=\tilde{\nu}_{0}(y_{2}),\quad&{\rm{on}}\quad\Sigma_{0},\\ u_{r}(L,y_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L},\\ u_{r}(y_{1},\frac{1}{2})=u_{x}(y_{1},\frac{1}{2})g_{cd}^{\prime}(y_{1}),\quad&{\rm{on}}\quad\Sigma,\\ u_{r}(y_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a}.\\ \end{cases} (3.20)

Furthermore, by (3.13), one obtains

A~0(12)(ρ(ux,ur,uθ,A~0,B~0))γ(y1,12)=Pb.\tilde{A}_{0}\left(\frac{1}{2}\right)\left(\rho(u_{x},u_{r},u_{\theta},\tilde{A}_{0},\tilde{B}_{0})\right)^{\gamma}(y_{1},\frac{1}{2})=P_{b}. (3.21)

Therefore 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 2.2\mathbf{Problem\ 2.2} is reformulated as follows.

Problem 3.1.

Given functions (J0,ν0,A0,B0)(J_{0},\nu_{0},A_{0},B_{0}) at the entrance satisfying (2.7), find a unique smooth subsonic solution (ux,ur,uθ;gcd)(u_{x},u_{r},u_{\theta};g_{cd}) satisfying (3.19) and (3.20) and the Rankine-Hugoniot condition (3.21).

3.3 Solving the free boundary problem 3.1

Note that 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 3.1\mathbf{Problem\ 3.1} is a free boundary problem since the function gcdg_{cd} is unknown, this free boundary problem will be solved by using the implicit function theorem. We follow the steps below to solve 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 3.1\mathbf{Problem\ 3.1}:

  1. (a)

    Given any function gcd(y1)=0y1w(s)ds+12g_{cd}(y_{1})=\int_{0}^{y_{1}}w(s){\mathrm{d}}s+\frac{1}{2} belonging to some suitable function classes, we will solve the nonlinear system (3.19) with mixed boundary condition (3.20) in Ω\Omega. This will be achieved by decomposing the system (3.19) into two boundary value problems with different inhomogeneous terms and employing the standard elliptic theory. The detailed analysis will be given in Section 4.

  2. (b)

    We use the implicit function theorem to locate the contact discontinuity. More precisely, define the map 𝒬(𝝋0,w):=A~0(12)(ρ(ux,ur,uθ,A~0,B~0))γ(y1,12)Pb\mathcal{Q}(\bm{\varphi}_{0},w):=\tilde{A}_{0}\left(\frac{1}{2}\right)\left(\rho(u_{x},u_{r},u_{\theta},\tilde{A}_{0},\tilde{B}_{0})\right)^{\gamma}(y_{1},\frac{1}{2})-P_{b}, we need to compute the Fréchet derivative Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w) of the functional 𝒬(𝝋0,w)\mathcal{Q}(\bm{\varphi}_{0},w) with respect to ww and show that Dw𝒬(𝝋b,0)D_{w}\mathcal{Q}(\bm{\varphi}_{b},0) is an isomorphism. This step will be achieved in Section 5.

4 The solution to a fixed boundary value problem in Ω\Omega

In this section, given any function gcd(y1)=0y1w(s)ds+12C1,α([0,L])g_{cd}(y_{1})=\int_{0}^{y_{1}}w(s){\mathrm{d}}s+\frac{1}{2}\in C^{1,\alpha}([0,L]) satisfying gcd(L)=0g_{cd}^{\prime}(L)=0, we will solve the nonlinear system (3.19) with mixed boundary condition (3.20) in Ω\Omega.

4.1 Linearization

To solve nonlinear system (3.19) in the domain Ω\Omega, we first linearize (3.19) and then solve the linear system in the domain Ω\Omega. Define

W1=uxub,W2=ur,W3=uθ,(J^,A^,B^)=(J~0,A~0,B~0)(Jb,Ab,Bb).W_{1}=u_{x}-u_{b}^{-},\quad W_{2}=u_{r},\quad W_{3}=u_{\theta},\quad(\hat{J},\hat{A},\hat{B})=(\tilde{J}_{0},\tilde{A}_{0},\tilde{B}_{0})-(J_{b}^{-},A_{b}^{-},B_{b}^{-}).

Denoting the solution space by 𝒥(δ1)\mathcal{J}(\delta_{1}), which is defined as

𝒥(δ1)={𝐖=(W1,W2,W3):j=13Wj1,α;Ω(α,Σ)δ1,W2(y1,0)=W2(L,y2)=W3(y1,0)=0}.\mathcal{J}(\delta_{1})=\{{\bf W}=(W_{1},W_{2},W_{3}):\sum_{j=1}^{3}\|W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq\delta_{1},\ W_{2}(y_{1},0)=W_{2}(L,y_{2})=W_{3}(y_{1},0)=0\}. (4.1)

Here δ1\delta_{1} is a positive constant to be determined later.

Given 𝐖^𝒥(δ1)\hat{{\bf W}}\in\mathcal{J}(\delta_{1}), it follows from the third equation in (3.18) that W3W_{3} can be solved as follows

W3(y1,y2)=Λ^(y2)r^(y1,y2),W_{3}(y_{1},y_{2})=\frac{\hat{\Lambda}(y_{2})}{\hat{r}(y_{1},y_{2})}, (4.2)

where

Λ^(y2)=r(0,y2)ν~0(y2)=(20y22sJ~0(s)ds)12ν~0(y2),\displaystyle\hat{\Lambda}(y_{2})=r(0,y_{2})\tilde{\nu}_{0}(y_{2})=\left(2\int_{0}^{y_{2}}\frac{2s}{\tilde{J}_{0}(s)}{\mathrm{d}}s\right)^{\frac{1}{2}}\tilde{\nu}_{0}(y_{2}),
r^(y1,y2)=(20y22sρ^(W^1+ub,W^2,W^3,A~0,B~0)(W^1+ub)(y1,s)ds)12.\displaystyle\hat{r}(y_{1},y_{2})=\left(2\int_{0}^{y_{2}}\frac{2s}{\hat{\rho}(\hat{W}_{1}+u_{b}^{-},\hat{W}_{2},\hat{W}_{3},\tilde{A}_{0},\tilde{B}_{0})(\hat{W}_{1}+u_{b}^{-})(y_{1},s)}{\mathrm{d}}s\right)^{\frac{1}{2}}.

Then one derives that

W31,α;Ω(α,Σ)Cσcd,\|W_{3}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq C\sigma_{cd}, (4.3)

where σcd=σ(J0,ν0,A0,B0)\sigma_{cd}=\sigma(J_{0},\nu_{0},A_{0},B_{0}) is defined in (2.16) and C>0C>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha).

In the following, we turn to concern W1W_{1} and W2W_{2}. It follows from the first and second equations in (3.19) that W1W_{1} and W2W_{2} satisfy the following first order elliptic system:

{(c2(ρb,Ab)(ub)2)y1W1+c2(ρb,Ab)y2W2+c2(ρb,Ab)rbW2=F1(𝐖^,𝐖^,A^,B^),y1W2rbρbub2y2y2W1=F2(𝐖^,𝐖^,A^,B^),\begin{cases}\begin{aligned} &(c^{2}(\rho_{b}^{-},A_{b}^{-})-(u_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}+c^{2}(\rho_{b}^{-},A_{b}^{-}){\partial}_{y_{2}}W_{2}\\ &+\frac{c^{2}(\rho_{b}^{-},A_{b}^{-})}{r_{b}}W_{2}=F_{1}(\hat{{\bf W}},\nabla{\hat{\bf W}},\hat{A},\hat{B}),\\ &{\partial}_{y_{1}}W_{2}-\frac{r_{b}\rho_{b}^{-}u_{b}^{-}}{2y_{2}}{\partial}_{y_{2}}W_{1}=F_{2}(\hat{{\bf W}},\nabla{\hat{\bf W}},\hat{A},\hat{B}),\\ \end{aligned}\end{cases} (4.4)

where

F1(𝐖^,𝐖^,A^,B^)\displaystyle F_{1}(\hat{{\bf W}},\nabla{\hat{\bf W}},\hat{A},\hat{B})
=(c2(ρ^,A~0)(ub+W^1)2)r^ρ^W^22y2y2W^1+W^22r^ρ^(ub+W^1)2y2y2W^2(c2(ρ^,A~0)r^c2(ρb,Ab)rb)W^2\displaystyle=(c^{2}(\hat{\rho},\tilde{A}_{0})-(u_{b}^{-}+\hat{W}_{1})^{2})\frac{\hat{r}\hat{\rho}\hat{W}_{2}}{2y_{2}}{\partial}_{y_{2}}\hat{W}_{1}+\hat{W}_{2}^{2}\frac{\hat{r}\hat{\rho}(u_{b}^{-}+\hat{W}_{1})}{2y_{2}}{\partial}_{y_{2}}\hat{W}_{2}-\left(\frac{c^{2}(\hat{\rho},\tilde{A}_{0})}{\hat{r}}-\frac{c^{2}(\rho_{b}^{-},A_{b}^{-})}{r_{b}}\right)\hat{W}_{2}
((γ1)B^γ+12W^12γ12W^22γ12W^32(γ+1)ubW^1)y1W^1\displaystyle\quad-\left((\gamma-1)\hat{B}-\frac{\gamma+1}{2}\hat{W}_{1}^{2}-\frac{\gamma-1}{2}\hat{W}_{2}^{2}-\frac{\gamma-1}{2}\hat{W}_{3}^{2}-(\gamma+1)u_{b}^{-}\hat{W}_{1}\right){\partial}_{y_{1}}\hat{W}_{1}
((γ1)B^γ12W^12γ12W^22γ12W^32(γ1)ubW^1)y2W^2\displaystyle\quad-\left((\gamma-1)\hat{B}-\frac{\gamma-1}{2}\hat{W}_{1}^{2}-\frac{\gamma-1}{2}\hat{W}_{2}^{2}-\frac{\gamma-1}{2}\hat{W}_{3}^{2}-(\gamma-1)u_{b}^{-}\hat{W}_{1}\right){\partial}_{y_{2}}\hat{W}_{2}
W^32r^W^2+(W^1+ub)W^2(y1W^2r^ρ^W^22y2y2W^2+r^ρ^(W^1+ub)2y2y2W^1),\displaystyle\quad-\frac{\hat{W}_{3}^{2}}{\hat{r}}\hat{W}_{2}+(\hat{W}_{1}+u_{b}^{-})\hat{W}_{2}\left({\partial}_{y_{1}}\hat{W}_{2}-\frac{\hat{r}\hat{\rho}\hat{W}_{2}}{2y_{2}}{\partial}_{y_{2}}\hat{W}_{2}+\frac{\hat{r}\hat{\rho}(\hat{W}_{1}+u_{b}^{-})}{2y_{2}}{\partial}_{y_{2}}\hat{W}_{1}\right),
F2(𝐖^,𝐖^,A^,B^)\displaystyle F_{2}(\hat{{\bf W}},\nabla{\hat{\bf W}},\hat{A},\hat{B})
=1ub+W^1(r^ρ^(W^1+ub)2y2W^3y2W^3+W^32r^+r^ρ^(W^1+ub)2y2ρ^γ1γ1y2A^r^ρ(W^1+ub)2y2y2B^)\displaystyle=\frac{1}{u_{b}^{-}+\hat{W}_{1}}\left(\frac{\hat{r}\hat{\rho}(\hat{W}_{1}+u_{b}^{-})}{2y_{2}}\hat{W}_{3}{\partial}_{y_{2}}\hat{W}_{3}+\frac{\hat{W}_{3}^{2}}{\hat{r}}+\frac{\hat{r}\hat{\rho}(\hat{W}_{1}+u_{b}^{-})}{2y_{2}}\frac{\hat{\rho}^{\gamma-1}}{\gamma-1}{\partial}_{y_{2}}\hat{A}-\frac{\hat{r}\rho(\hat{W}_{1}+u_{b}^{-})}{2y_{2}}{\partial}_{y_{2}}\hat{B}\right)
+1ub+W^1(r^ρ^W^22y2y2W^2).\displaystyle\quad+\frac{1}{u_{b}^{-}+\hat{W}_{1}}\left(\frac{\hat{r}\hat{\rho}\hat{W}_{2}}{2y_{2}}{\partial}_{y_{2}}\hat{W}_{2}\right).

Recalling that r(A0,B0)(0)=0{\partial}_{r}(A_{0},{B}_{0})(0)=0, then one obtains

y2(A^,B^)(0)=y2(A~0,B~0)(0)=0.{\partial}_{y_{2}}(\hat{A},\hat{B})(0)={\partial}_{y_{2}}(\tilde{A}_{0},\tilde{B}_{0})(0)=0. (4.5)

Thus for 𝐖^𝒥(δ1)\hat{{\bf W}}\in\mathcal{J}(\delta_{1}), it is easy to check that F2(y1,0)=0F_{2}(y_{1},0)=0. Furthermore, there exist positive constants κ1\kappa_{1} and κ2\kappa_{2} depending only on the background solutions 𝑼b\bm{U}_{b}^{-} such that

κ1y2r^κ2y2,\kappa_{1}y_{2}\leq\hat{r}\leq\kappa_{2}y_{2},

which implies that

κ1r^y2κ2.\kappa_{1}\leq\frac{\hat{r}}{y_{2}}\leq\kappa_{2}. (4.6)

Hence one can derive

j=12Fj0,α;Ω(1α,Σ)C((j=13Wj^1,α;Ω(α,Σ))2+σcd)C(δ12+σcd),\displaystyle\sum_{j=1}^{2}\|F_{j}\|_{0,\alpha;\Omega}^{(1-\alpha,\Sigma)}\leq C\left(\bigg{(}\sum_{j=1}^{3}\|\hat{W_{j}}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\bigg{)}^{2}+\sigma_{cd}\right)\leq C\left(\delta_{1}^{2}+\sigma_{cd}\right), (4.7)

where C>0C>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha).

Next, we derive the boundary conditions for 𝐖{\bf W}. It follows from (3.19) that

{W1(0,y2)=F3(𝐖^,J^,A^,B^)(0,y2),onΣ0,W2(L,y2)=0,onΣL,W2(y1,12)=F4(𝐖^,gcd)(y1,12),onΣ,W2(y1,0)=0,onΣa,\begin{cases}W_{1}(0,y_{2})=F_{3}(\hat{{\bf W}},\hat{J},\hat{A},\hat{B})(0,y_{2}),\quad&{\rm{on}}\quad\Sigma_{0},\\ W_{2}(L,y_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L},\\ W_{2}(y_{1},\frac{1}{2})=F_{4}(\hat{{\bf W}},g_{cd})(y_{1},\frac{1}{2}),\quad&{\rm{on}}\quad\Sigma,\\ W_{2}(y_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a},\\ \end{cases} (4.8)

where

F3(𝐖^,J^,A^,B^)\displaystyle F_{3}(\hat{{\bf W}},\hat{J},\hat{A},\hat{B})
=J^ρb(1(Mb)2)W^1ρb(1(Mb)2)(H(B^+Bb,A^+Ab,(ub+W^1)2+W^22+W^32)\displaystyle=\frac{\hat{J}}{\rho_{b}^{-}(1-(M_{b}^{-})^{2})}-\frac{\hat{W}_{1}}{\rho_{b}^{-}(1-(M_{b}^{-})^{2})}\bigg{(}H(\hat{B}+B_{b}^{-},\hat{A}+A_{b}^{-},(u_{b}^{-}+\hat{W}_{1})^{2}+\hat{W}_{2}^{2}+\hat{W}_{3}^{2})
H(Bb,Ab,(ub)2))\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad-H(B_{b}^{-},A_{b}^{-},(u_{b}^{-})^{2})\bigg{)}
ubρb(1(Mb)2)(H(B^+Bb,A^+Ab,(ub+W^1)2+W^22+W^32)\displaystyle\quad-\frac{u_{b}^{-}}{\rho_{b}^{-}(1-(M_{b}^{-})^{2})}\left(H(\hat{B}+B_{b}^{-},\hat{A}+A_{b}^{-},(u_{b}^{-}+\hat{W}_{1})^{2}+\hat{W}_{2}^{2}+\hat{W}_{3}^{2})\right.
H(Bb,Ab,(ub)2)+Jbc2(ρb,Ab)W^1),\displaystyle\qquad\qquad\qquad\qquad\qquad\left.-H(B_{b}^{-},A_{b}^{-},(u_{b}^{-})^{2})+\frac{J_{b}^{-}}{c^{2}(\rho_{b}^{-},A_{b}^{-})}\hat{W}_{1}\right),
(Mb)2=(ub)2c2(ρb,Ab),F4(𝐖^,gcd)=(ub+W^1)gcd.\displaystyle(M_{b}^{-})^{2}=\frac{(u_{b}^{-})^{2}}{c^{2}(\rho_{b}^{-},A_{b}^{-})},\quad F_{4}(\hat{{\bf W}},g_{cd})=(u_{b}^{-}+\hat{W}_{1})g_{cd}^{\prime}.

Then a direct computation yields

F31,α;[0,12)(α,{12})+F40,α;[0,L]\displaystyle\|F_{3}\|_{1,\alpha;[0,\frac{1}{2})}^{(-\alpha,\{\frac{1}{2}\})}+\|F_{4}\|_{0,\alpha;[0,L]} (4.9)
C((j=13Wj^1,α;Ω(α,Σ))2+W1^1,α;Ω(α,Σ)gcd0,α;[0,L]+gcd0,α;[0,L]+σcd)\displaystyle\leq C\left(\bigg{(}\sum_{j=1}^{3}\|\hat{W_{j}}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\bigg{)}^{2}+\|\hat{W_{1}}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\|g_{cd}^{\prime}\|_{0,\alpha;[0,L]}+\|g_{cd}^{\prime}\|_{0,\alpha;[0,L]}+\sigma_{cd}\right)
C(δ12+δ1w0,α;[0,L]+w0,α;[0,L]+σcd),\displaystyle\leq C\left(\delta_{1}^{2}+\delta_{1}\|w\|_{0,\alpha;[0,L]}+\|w\|_{0,\alpha;[0,L]}+\sigma_{cd}\right),

where C>0C>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha).

4.2 Solving the linear boundary value problem

In this subsection, we consider the following linear boundary value problem:

{(1(Mb)2)y1W1+y2W2+1y2W2=1(𝐖^,𝐖^,A^,B^),y1W2y2W1=2(𝐖^,𝐖^,A^,B^),W1(0,y2)=3(𝐖^,J^,Λ^,A^,B^)(0,y2),onΣ0,W2(L,y2)=0,onΣL,W2(y1,12)=4(𝐖^,gcd)(y1,12),onΣ,W2(y1,0)=0,onΣa,\begin{cases}\begin{aligned} &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}+{\partial}_{y_{2}}W_{2}+\frac{1}{y_{2}}W_{2}=\mathcal{F}_{1}(\hat{{\bf W}},\nabla{\hat{\bf W}},\hat{A},\hat{B}),\\ &{\partial}_{y_{1}}W_{2}-{\partial}_{y_{2}}W_{1}=\mathcal{F}_{2}(\hat{{\bf W}},\nabla{\hat{\bf W}},\hat{A},\hat{B}),\\ &W_{1}(0,y_{2})=\mathcal{F}_{3}(\hat{{\bf W}},\hat{J},\hat{\Lambda},\hat{A},\hat{B})(0,y_{2}),\quad&{\rm{on}}\quad\Sigma_{0},\\ &W_{2}(L,y_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L},\\ &W_{2}(y_{1},\frac{1}{2})=\mathcal{F}_{4}(\hat{{\bf W}},g_{cd})(y_{1},\frac{1}{2}),\quad&{\rm{on}}\quad\Sigma,\\ &W_{2}(y_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a},\\ \end{aligned}\end{cases} (4.10)

where

1=1c2(ρb,Ab)F1,i=Fi,i=2,3,4.\mathcal{F}_{1}=\frac{1}{c^{2}(\rho_{b}^{-},A_{b}^{-})}F_{1},\quad\mathcal{F}_{i}=F_{i},i=2,3,4.

For the problem (4.10), we have the following conclusion:

Lemma 4.1.

Let α(12,1)\alpha\in(\frac{1}{2},1). For given jC0,α(1α,Σ)(Ω)\mathcal{F}_{j}\in C_{0,\alpha}^{(1-\alpha,\Sigma)}(\Omega), j=1,2j=1,2, 2(y1,0)=0\mathcal{F}_{2}(y_{1},0)=0, 3C1,α(α,{12})([0,12))\mathcal{F}_{3}\in C_{1,\alpha}^{(-\alpha,\{\frac{1}{2}\})}([0,\frac{1}{2})), 4C0,α([0,L])\mathcal{F}_{4}\in C^{0,\alpha}([0,L]), the boundary value problem (4.10) has a unique solution (W1,W2)(C1,α(α,Σ)(Ω))2(W_{1},W_{2})\in\left(C_{1,\alpha}^{(-\alpha,\Sigma)}(\Omega)\right)^{2} satisfying

j=12Wj1,α;Ω(α,Σ)C(j=12j0,α;Ω(1α,Σ)+31,α;[0,12)(α,{12})+40,α;[0,L]),\sum_{j=1}^{2}\|W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq C\left(\sum_{j=1}^{2}\|\mathcal{F}_{j}\|_{0,\alpha;\Omega}^{(1-\alpha,\Sigma)}+\|\mathcal{F}_{3}\|_{1,\alpha;[0,\frac{1}{2})}^{(-\alpha,\{\frac{1}{2}\})}+\|\mathcal{F}_{4}\|_{0,\alpha;[0,L]}\right), (4.11)

where C>0C>0 depends only on (𝐔b,L,α)(\bm{U}_{b}^{-},L,\alpha).

Proof.

We divide the proof into four steps.

Step 1: In this step, in order to solve (4.10), we first introduce a transformation to reduce it to a typical form.

Let

{z1=11(Mb)2y1,z2=y2,and{V1=1(Mb)2W1,V2=W2.\begin{array}[]{ll}\left\{\begin{aligned} &z_{1}=\sqrt{\frac{1}{1-(M_{b}^{-})^{2}}}y_{1},\\ &z_{2}=y_{2},\\ \end{aligned}\right.\quad\text{and}\quad\left\{\begin{aligned} &V_{1}=\sqrt{{1-(M_{b}^{-})^{2}}}W_{1},\\ &V_{2}=W_{2}.\\ \end{aligned}\right.\end{array} (4.12)

The domain Ω\Omega becomes

Ω={(z1,z2):0<z1<L,0<y2<12},\Omega^{\ast}=\{(z_{1},z_{2}):0<z_{1}<L^{\ast},0<y_{2}<\frac{1}{2}\},

where L=11(Mb)2LL^{\ast}=\sqrt{\frac{1}{1-(M_{b}^{-})^{2}}}L, and its boundaries are transformed into

Σ0:={(z1,z2):z1=0, 0z2<12},ΣL:={(z1,z2):z1=L, 0z2<12},\displaystyle\Sigma_{0}^{\ast}:=\{(z_{1},z_{2}):z_{1}=0,\ 0\leq z_{2}<\frac{1}{2}\},\quad\Sigma_{L}^{\ast}:=\{(z_{1},z_{2}):z_{1}=L^{\ast},\ 0\leq z_{2}<\frac{1}{2}\},
Σa:={(z1,z2):0<z1<L,z2=0},Σ:={(z1,z2):0<z1<L,z2=12}.\displaystyle\Sigma_{a}^{\ast}:=\{(z_{1},z_{2}):0<z_{1}<L^{\ast},\ z_{2}=0\},\quad\Sigma^{\ast}:=\{(z_{1},z_{2}):0<z_{1}<L^{\ast},\ z_{2}=\frac{1}{2}\}.

Then the system (4.10) is reformulated as

{z1(z2V1)+z2(z2V2)=z21,z1V2z2V1=1(Mb)22:=~2,V1(0,z2)=1(Mb)23:=~3,onΣ0,V2(L,z2)=0,onΣL,V2(z1,12)=4,onΣ,V2(z1,0)=0,onΣa.\begin{cases}\begin{aligned} &{\partial}_{z_{1}}(z_{2}V_{1})+{\partial}_{z_{2}}(z_{2}V_{2})=z_{2}\mathcal{F}_{1},\\ &{\partial}_{z_{1}}V_{2}-{\partial}_{z_{2}}V_{1}=\sqrt{{1-(M_{b}^{-})^{2}}}\mathcal{F}_{2}:=\tilde{\mathcal{F}}_{2},\\ &V_{1}(0,z_{2})=\sqrt{{1-(M_{b}^{-})^{2}}}\mathcal{F}_{3}:=\tilde{\mathcal{F}}_{3},\quad&{\rm{on}}\quad\Sigma_{0}^{\ast},\\ &V_{2}(L^{\ast},z_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L}^{\ast},\\ &V_{2}(z_{1},\frac{1}{2})=\mathcal{F}_{4},\quad&{\rm{on}}\quad\Sigma^{\ast},\\ &V_{2}(z_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a}^{\ast}.\\ \end{aligned}\end{cases} (4.13)

Next, we decompose the problem (4.13) into two boundary value problems with different inhomogeneous terms as follows. Let (V1,V2)T=(H1,H2)T+(K1,K2)T(V_{1},V_{2})^{T}=(H_{1},H_{2})^{T}+(K_{1},K_{2})^{T}, where (H1,H2)T(H_{1},H_{2})^{T} is the solution to the problem

{z1(z2H1)+z2(z2H2)=0,z1H2z2H1=~2,H1(0,z2)=0,onΣ0,H2(L,z2)=0,onΣL,H2(z1,12)=0,onΣ,H2(z1,0)=0,onΣa,\begin{cases}\begin{aligned} &{\partial}_{z_{1}}(z_{2}H_{1})+{\partial}_{z_{2}}(z_{2}H_{2})=0,\\ &{\partial}_{z_{1}}H_{2}-{\partial}_{z_{2}}H_{1}=\tilde{\mathcal{F}}_{2},\\ &H_{1}(0,z_{2})=0,\quad&{\rm{on}}\quad\Sigma_{0}^{\ast},\\ &H_{2}(L^{\ast},z_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L}^{\ast},\\ &H_{2}(z_{1},\frac{1}{2})=0,\quad&{\rm{on}}\quad\Sigma^{\ast},\\ &H_{2}(z_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a}^{\ast},\\ \end{aligned}\end{cases} (4.14)

and (K1,K2)T(K_{1},K_{2})^{T} satisfies the following problem

{z1(z2K1)+z2(z2K2)=z21,z1K2z2K1=0,K1(0,z2)=~3,onΣ0,K2(L,z2)=0,onΣL,K2(z1,12)=4,onΣ,K2(z1,0)=0,onΣa.\begin{cases}\begin{aligned} &{\partial}_{z_{1}}(z_{2}K_{1})+{\partial}_{z_{2}}(z_{2}K_{2})=z_{2}\mathcal{F}_{1},\\ &{\partial}_{z_{1}}K_{2}-{\partial}_{z_{2}}K_{1}=0,\\ &K_{1}(0,z_{2})=\tilde{\mathcal{F}}_{3},\quad&{\rm{on}}\quad\Sigma_{0}^{\ast},\\ &K_{2}(L^{\ast},z_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L}^{\ast},\\ &K_{2}(z_{1},\frac{1}{2})=\mathcal{F}_{4},\quad&{\rm{on}}\quad\Sigma^{\ast},\\ &K_{2}(z_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a}^{\ast}.\\ \end{aligned}\end{cases} (4.15)

Step 2: In this step, we are going to solve (4.14). The first equation in (4.14) implies that there exists a potential function ϕ1\phi_{1} such that

(z1ϕ1,z2ϕ1)=(z2H2,z2H1).({\partial}_{z_{1}}\phi_{1},{\partial}_{z_{2}}\phi_{1})=(z_{2}H_{2},-z_{2}H_{1}). (4.16)

Let Φ1=ϕ1z2\Phi_{1}=\frac{\phi_{1}}{z_{2}}. Then (4.16) yields that

(H1,H2)=((z2Φ1+Φ1z2),z1Φ1).(H_{1},H_{2})=(-({\partial}_{z_{2}}\Phi_{1}+\frac{\Phi_{1}}{z_{2}}),{\partial}_{z_{1}}\Phi_{1}). (4.17)

Without loss of generality, we assume that Φ1(0,0)=0\Phi_{1}(0,0)=0. Thus (4.14) can be rewritten as the following equation for Φ1\Phi_{1}:

{z12Φ1+z22Φ1+z2Φ1z2Φ1z22=~2,z1Φ1=0,onΣL,Φ1=0,onΣaΣΣ0.\begin{cases}\begin{aligned} &{\partial}_{z_{1}}^{2}\Phi_{1}+{\partial}_{z_{2}}^{2}\Phi_{1}+\frac{{\partial}_{z_{2}}\Phi_{1}}{z_{2}}-\frac{\Phi_{1}}{z_{2}^{2}}=\tilde{\mathcal{F}}_{2},\\ &{\partial}_{z_{1}}\Phi_{1}=0,\quad{\rm{on}}\quad\Sigma_{L}^{\ast},\\ &\Phi_{1}=0,\qquad\ {\rm{on}}\quad\Sigma_{a}^{\ast}\cup\Sigma^{\ast}\cup\Sigma_{0}^{\ast}.\\ \end{aligned}\end{cases} (4.18)

Obviously, the coefficients of equation (4.18) tends to infinity as z2z_{2} goes to 0. By applying the idea of Proposition 3.3 in [1], we rewrite (4.18) as a boundary value problem in 5\mathbb{R}^{5} so that the singular term in (4.18) can be removed from the equation for Φ1\Phi_{1}. Set

Φ=Φ1z2,2=~2z2.\Phi^{\ast}=\frac{\Phi_{1}}{z_{2}},\quad\mathcal{F}_{2}^{\ast}=\frac{\tilde{\mathcal{F}}_{2}}{z_{2}}. (4.19)

We regard Φ\Phi^{\ast} and 2\mathcal{F}_{2}^{\ast} as functions defined in

𝒟:={(z1,𝐳2):z1(0,L),𝐳24:|𝐳2|12}5,\mathcal{D}^{\ast}:=\{(z_{1},\mathbf{z}_{2}):z_{1}\in(0,L^{\ast}),\mathbf{z}_{2}\in\mathbb{R}^{4}:|\mathbf{z}_{2}|\leq\frac{1}{2}\}\subset\mathbb{R}^{5}, (4.20)

where 𝐳2=(z21,z22,z23,z24)\mathbf{z}_{2}=(z_{21},z_{22},z_{23},z_{24}) and j=14z2j2=|𝐳2|2\sum_{j=1}^{4}z_{2j}^{2}=|\mathbf{z}_{2}|^{2}. Define

𝓖(z1,𝐳2)=(0,𝒢(z1,𝐳2)z21,𝒢(z1,𝐳2)z22,𝒢(z1,𝐳2)z23,𝒢(z1,𝐳2)z24),𝐳=(z1,𝐳2)𝒟,\bm{\mathcal{G}}^{\ast}(z_{1},\mathbf{z}_{2})=(0,\mathcal{G}^{\ast}(z_{1},\mathbf{z}_{2})z_{21},\mathcal{G}^{\ast}(z_{1},\mathbf{z}_{2})z_{22},\mathcal{G}^{\ast}(z_{1},\mathbf{z}_{2})z_{23},\mathcal{G}^{\ast}(z_{1},\mathbf{z}_{2})z_{24}),\ \forall\mathbf{z}=(z_{1},\mathbf{z}_{2})\in\mathcal{D}^{\ast},

with

𝒢(z1,𝐳2)=1|𝐳2|40𝐳2s32(z1,s)ds.\mathcal{G}^{\ast}(z_{1},\mathbf{z}_{2})=\frac{1}{|\mathbf{z}_{2}|^{4}}\int_{0}^{\mathbf{z}_{2}}s^{3}\mathcal{F}_{2}^{\ast}(z_{1},s){\mathrm{d}}s.

Then it follows from (4.18) that

{𝐳Φ=div𝐳𝓖,in𝒟,z1Φ(L,𝐳2)=0,onBL:={L}×{𝐳24:|𝐳2|12},Φ(0,𝐳2)=0,onB0:={0}×{𝐳24:|𝐳2|12},Φ(z1,𝐳2)=0,onBw:=[0,L]×{𝐳24:|𝐳2|=12}.\begin{cases}\begin{aligned} &\triangle_{\mathbf{z}}\Phi^{\ast}=\text{div}_{\mathbf{z}}\bm{\mathcal{G}}^{\ast},\qquad{\rm{in}}\quad\mathcal{D}^{\ast},\\ &{\partial}_{z_{1}}\Phi^{\ast}(L^{\ast},\mathbf{z}_{2})=0,\quad{\rm{on}}\quad B_{L^{\ast}}:=\{L^{\ast}\}\times\{\mathbf{z}_{2}\in\mathbb{R}^{4}:|\mathbf{z}_{2}|\leq\frac{1}{2}\},\\ &\Phi^{\ast}(0,\mathbf{z}_{2})=0,\qquad\ \ {\rm{on}}\quad B_{0}:=\{0\}\times\{\mathbf{z}_{2}\in\mathbb{R}^{4}:|\mathbf{z}_{2}|\leq\frac{1}{2}\},\\ &\Phi^{\ast}(z_{1},\mathbf{z}_{2})=0,\qquad\ \ {\rm{on}}\quad B_{w}:=[0,L^{\ast}]\times\{\mathbf{z}_{2}\in\mathbb{R}^{4}:|\mathbf{z}_{2}|=\frac{1}{2}\}.\\ \end{aligned}\end{cases} (4.21)

The standard elliptic theory in [9] yields that (4.21) has a unique weak solution ΦH1(𝒟)\Phi^{\ast}\in H^{1}(\mathcal{D}^{\ast}) satisfying

[Φ,ξ]=(𝓖,ξ)forallξ{ξH1(𝒟):ξ=0onB0Bw},\mathcal{L}[\Phi^{\ast},\xi]=(\bm{\mathcal{G}}^{\ast},\xi)\quad{\rm{for}}\ {\rm{all}}\ \xi\in\{\xi\in H^{1}(\mathcal{D}^{\ast}):\xi=0\quad{\rm{on}}\quad B_{0}\cup B_{w}\}, (4.22)

where

[Φ,ξ]=𝒟Φξd𝐳,(𝓖,ξ)=𝒟𝓖ξd𝐳.\displaystyle\mathcal{L}[\Phi^{\ast},\xi]=\int_{\mathcal{D}^{\ast}}\nabla\Phi^{\ast}\nabla\xi{\mathrm{d}}\mathbf{z},\quad(\bm{\mathcal{G}}^{\ast},\xi)=\int_{\mathcal{D}^{\ast}}\bm{\mathcal{G}}^{\ast}\nabla\xi{\mathrm{d}}\mathbf{z}.

Furthermore, Φ\Phi^{\ast} satisfies

ΦH1(𝒟)C𝓖L2(𝒟).\|\Phi^{\ast}\|_{H^{1}(\mathcal{D}^{\ast})}\leq C\|\bm{\mathcal{G}}^{\ast}\|_{L^{2}(\mathcal{D}^{\ast})}. (4.23)

Next, we prove

𝓖L2(𝒟)C𝓖0,α;𝒟(1α,𝒟).\|\bm{\mathcal{G}}^{\ast}\|_{L^{2}(\mathcal{D}^{\ast})}\leq C\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{(1-\alpha,{\partial}\mathcal{D}^{\ast})}. (4.24)

Note that ~2(z1,0)=0\tilde{\mathcal{F}}_{2}(z_{1},0)=0. Then it holds that

2(z1,z2)=~2(z1,z2)~2(z1,0)z2=~2(z1,z2)~2(z1,0)z2αz2α1.\mathcal{F}_{2}^{\ast}(z_{1},z_{2})=\frac{\tilde{\mathcal{F}}_{2}(z_{1},z_{2})-\tilde{\mathcal{F}}_{2}(z_{1},0)}{z_{2}}=\frac{\tilde{\mathcal{F}}_{2}(z_{1},z_{2})-\tilde{\mathcal{F}}_{2}(z_{1},0)}{z_{2}^{\alpha}}z_{2}^{\alpha-1}. (4.25)

Since ~2C0,α(1α,Σ)(Ω)\tilde{\mathcal{F}}_{2}\in C_{0,\alpha}^{(1-\alpha,\Sigma^{\ast})}(\Omega^{\ast}), it is easy to check that

𝓖0,α;𝒟(1α,𝒟)C~20,α;Ω(1α,Σ).\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{(1-\alpha,{\partial}\mathcal{D}^{\ast})}\leq C\|\tilde{\mathcal{F}}_{2}\|_{0,\alpha;\Omega^{\ast}}^{(1-\alpha,\Sigma^{\ast})}. (4.26)

By the weighted Ho¨\ddot{\rm{o}}lder norm, one derives

|𝓖(𝐳)|δ𝐳α1𝓖0,α;𝒟~(1α;𝒟),|\bm{\mathcal{G}}^{\ast}(\mathbf{z})|\leq\delta_{\mathbf{z}}^{\alpha-1}\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\tilde{\mathcal{D}}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}, (4.27)

with δ𝐳=dist(𝐳,𝒟)\delta_{\mathbf{z}}={\rm{dist}}(\mathbf{z},{\partial}\mathcal{D}^{\ast}). Thus for α(12,1)\alpha\in(\frac{1}{2},1), one has

𝒟|𝓖|2d𝐳C(𝓖0,α;𝒟(1α;𝒟))2.\int_{\mathcal{D}^{\ast}}|\bm{\mathcal{G}}^{\ast}|^{2}{\mathrm{d}}\mathbf{z}\leq C\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}.

Next, we improve the regularity of Φ\Phi^{\ast}. For 𝐳0𝒟\mathbf{z}_{0}\in{\mathcal{D}^{\ast}} and η\eta\in\mathbb{R} with 0<η<1100<\eta<\frac{1}{10}, set

Bη(𝐳0):={𝐳5:|𝐳0𝐳|<η},Dη(𝐳0):=Bη(𝐳0)𝒟,\displaystyle B_{\eta}(\mathbf{z}_{0}):=\{\mathbf{z}\in\mathbb{R}^{5}:|\mathbf{z}_{0}-\mathbf{z}|<\eta\},\quad D_{\eta}(\mathbf{z}_{0}):=B_{\eta}(\mathbf{z}_{0})\cap\mathcal{D}^{\ast},
Φ𝐳0,η:=1|Dη(𝐳0)|Dη(𝐳0)Φd𝐳.\displaystyle\Phi^{\ast}_{\mathbf{z}_{0},\eta}:=\frac{1}{|D_{\eta}(\mathbf{z}_{0})|}\int_{D_{\eta}(\mathbf{z}_{0})}\Phi^{\ast}{\mathrm{d}}\mathbf{z}.

Note that there exists a constant λ0(0,1/10)\lambda_{0}\in(0,1/10) such that

λ0|Dη(𝐳0)||Bη(𝐳0)|1λ0.\lambda_{0}\leq\frac{|D_{\eta}(\mathbf{z}_{0})|}{|B_{\eta}(\mathbf{z}_{0})|}\leq\frac{1}{\lambda_{0}}.

Hence we follow the proof in Theorem 3.8 of [10] to get

Dη(𝐳)|ΦΦ𝐳,η|2d𝐳C(𝓖0,α;𝒟(1α;𝒟))2η5+2α\int_{D_{\eta}(\mathbf{z})}|\Phi^{\ast}-\Phi^{\ast}_{\mathbf{z},\eta}|^{2}{\mathrm{d}}\mathbf{z}\leq C\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}\eta^{5+2\alpha} (4.28)

for any 𝐳𝒟¯\mathbf{z}\in\overline{\mathcal{D}^{\ast}}. Once (4.28) is obtained, it follows from Theorem 3.1 in [10] that

Φ0,α;𝒟C𝓖0,α;𝒟(1α;𝒟).\|\Phi^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}\leq C\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}. (4.29)

We proof (4.28) only for the case 𝐳B0Bw\mathbf{z}\in B_{0}\cap B_{w}, since the other cases can be treated similarly. Fix 𝐳0B0Bw\mathbf{z}_{0}\in B_{0}\cap B_{w} and χ\chi\in\mathbb{R} with 0<χ<1100<\chi<\frac{1}{10}. Let Φh\Phi^{\ast}_{h} be a weak solution of the following problem:

{𝐳Φh=0,inDχ(𝐳0),Φh=Φ,inDχ(𝐳0)𝒟.\begin{cases}\begin{aligned} &\triangle_{\mathbf{z}}\Phi_{h}^{\ast}=0,\qquad{\rm{in}}\quad D_{\chi}(\mathbf{z}_{0}),\\ &\Phi_{h}^{\ast}=\Phi^{\ast},\qquad\ \ {\rm{in}}\quad{\partial}D_{\chi}(\mathbf{z}_{0})\cap\mathcal{D}^{\ast}.\\ \end{aligned}\end{cases} (4.30)

Then Φ~h=ΦΦh\tilde{\Phi}_{h}^{\ast}=\Phi^{\ast}-\Phi_{h}^{\ast} satisfies

Dχ(𝐳0)Φ~hξd𝐳=Dχ(𝐳0)𝓖ξd𝐳,\int_{D_{\chi}(\mathbf{z}_{0})}\nabla\tilde{\Phi}_{h}^{\ast}\nabla\xi{\mathrm{d}}\mathbf{z}=\int_{D_{\chi}(\mathbf{z}_{0})}\bm{\mathcal{G}}^{\ast}\nabla\xi{\mathrm{d}}\mathbf{z},

for any ξ{ξH1(Dχ(𝐳0):ξ=0onDχ(𝐳0)(B0Bw)}\xi\in\{\xi\in H^{1}(D_{\chi}(\mathbf{z}_{0}):\xi=0\ {\rm{on}}\ {\partial}D_{\chi}(\mathbf{z}_{0})\cap(B_{0}\cup B_{w})\}. Taking the test function ξ=Φ~h\xi=\tilde{\Phi}_{h}^{\ast} and using the Hölder inequality to yield that

Dχ(𝐳0)|Φ~h|2d𝐳Dχ(𝐳0)|𝓖|2d𝐳.\int_{D_{\chi}(\mathbf{z}_{0})}|\nabla\tilde{\Phi}_{h}^{\ast}|^{2}{\mathrm{d}}\mathbf{z}\leq\int_{D_{\chi}(\mathbf{z}_{0})}|\bm{\mathcal{G}}^{\ast}|^{2}{\mathrm{d}}\mathbf{z}. (4.31)

Due to 𝓖C0,α(1α;𝒟)(𝒟)\bm{\mathcal{G}}^{\ast}\in C_{0,\alpha}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}(\mathcal{D}^{\ast}), one gets

Dχ(𝐳0)|Φ~h|2d𝐳(𝓖0,α;𝒟(1α;𝒟))2Dχ(𝐳0)δ𝐳2(α1)d𝐳χ3+2α(𝓖0,α;𝒟(1α;𝒟))2.\int_{D_{\chi}(\mathbf{z}_{0})}|\nabla\tilde{\Phi}_{h}^{\ast}|^{2}{\mathrm{d}}\mathbf{z}\leq\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}\int_{D_{\chi}(\mathbf{z}_{0})}\delta_{\mathbf{z}}^{2(\alpha-1)}{\mathrm{d}}\mathbf{z}\leq\chi^{3+2\alpha}\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}. (4.32)

By Corollary 3.11 in [10], for 0<η<χ0<\eta<\chi, one has

Dη(𝐳0)|Φ|2d𝐳\displaystyle\int_{D_{\eta}(\mathbf{z}_{0})}|\nabla\Phi^{\ast}|^{2}{\mathrm{d}}\mathbf{z} C(ηχ)5Dχ(𝐳0)|Φ|2d𝐳+C(𝓖0,α;𝒟(1α;𝒟))2χ3+2α.\displaystyle\leq C\left(\frac{\eta}{\chi}\right)^{5}\int_{D_{\chi}(\mathbf{z}_{0})}|\nabla\Phi^{\ast}|^{2}{\mathrm{d}}\mathbf{z}+C\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}\chi^{3+2\alpha}. (4.33)

Then it follows from Lemma 3.4 in [10] that one obtains

Dη(𝐳0)|Φ|2d𝐳\displaystyle\int_{D_{\eta}(\mathbf{z}_{0})}|\nabla\Phi^{\ast}|^{2}{\mathrm{d}}\mathbf{z} C(1χ3+2α𝒟|Φ|2d𝐳+(𝓖0,α;𝒟(1α;𝒟))2)η3+2α.\displaystyle\leq C\left(\frac{1}{\chi^{3+2\alpha}}\int_{\mathcal{D}^{\ast}}|\nabla\Phi^{\ast}|^{2}{\mathrm{d}}\mathbf{z}+\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}\right)\eta^{3+2\alpha}. (4.34)

By applying Poincaré inequality, one can derive

Dη(𝐳0)|ΦΦ𝐳,η|2d𝐳C(1χ3+2α𝒟|Φ|2d𝐳+(𝓖0,α;𝒟(1α;𝒟))2)η5+2α.\int_{D_{\eta}(\mathbf{z}_{0})}|\Phi^{\ast}-\Phi^{\ast}_{\mathbf{z},\eta}|^{2}{\mathrm{d}}\mathbf{z}\leq C\left(\frac{1}{\chi^{3+2\alpha}}\int_{\mathcal{D}^{\ast}}|\nabla\Phi^{\ast}|^{2}{\mathrm{d}}\mathbf{z}+\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}\right)\eta^{5+2\alpha}. (4.35)

Then it follows from (4.23) and (4.24) that

Dη(𝐳0)|ΦΦ𝐳,η|2d𝐳C(𝓖0,α;𝒟(1α;𝒟))2η5+2α.\int_{D_{\eta}(\mathbf{z}_{0})}|\Phi^{\ast}-\Phi^{\ast}_{\mathbf{z},\eta}|^{2}{\mathrm{d}}\mathbf{z}\leq C\left(\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{({1-\alpha;{\partial}\mathcal{D}^{\ast}})}\right)^{2}\eta^{5+2\alpha}.

Hence the proof of (4.29) is completed.

Therefore, by the scaling argument and the Schauder estimate in [9], one has

Φ1,α;𝒟(α,𝒟)C𝓖0,α;𝒟(1α,𝒟)C~20,α;Ω(1α,Σ).\|\Phi^{\ast}\|_{1,\alpha;\mathcal{D}^{\ast}}^{(-\alpha,{\partial}\mathcal{D}^{\ast})}\leq C\|\bm{\mathcal{G}}^{\ast}\|_{0,\alpha;\mathcal{D}^{\ast}}^{(1-\alpha,{\partial}\mathcal{D}^{\ast})}\leq C\|\tilde{\mathcal{F}}_{2}\|_{0,\alpha;\Omega^{\ast}}^{(1-\alpha,\Sigma^{\ast})}. (4.36)

Furthermore, the rotational invariance of the boundary value problem (4.21) and the uniqueness of the solution Φ\Phi^{\ast} and the estimate (4.36) imply that

Φ1,α;Ω(α,Σ)C~20,α;Ω(1α,Σ).\|\Phi^{\ast}\|_{1,\alpha;\Omega^{\ast}}^{(-\alpha,\Sigma^{\ast})}\leq C\|\tilde{\mathcal{F}}_{2}\|_{0,\alpha;\Omega^{\ast}}^{(1-\alpha,\Sigma^{\ast})}.

Using the first equation in (4.18), one can verify

z12Φ1+z22Φ1=~2z2Φ1z2+Φ1z22=~2z2ΦH0,α(1α,Σ)(Ω).{\partial}_{z_{1}}^{2}\Phi_{1}+{\partial}_{z_{2}}^{2}\Phi_{1}=\tilde{\mathcal{F}}_{2}-\frac{{\partial}_{z_{2}}\Phi_{1}}{z_{2}}+\frac{\Phi_{1}}{z_{2}^{2}}=\tilde{\mathcal{F}}_{2}-{\partial}_{z_{2}}\Phi^{\ast}\in H_{0,\alpha}^{(1-\alpha,\Sigma^{\ast})}(\Omega^{\ast}). (4.37)

By the Schauder estimate in Theorem 4.6 of [13], we obtain

Φ12,α;Ω(1α,Σ)C~20,α;Ω(1α,Σ).\|\Phi_{1}\|_{2,\alpha;\Omega^{\ast}}^{(-1-\alpha,\Sigma^{\ast})}\leq C\|\tilde{\mathcal{F}}_{2}\|_{0,\alpha;\Omega^{\ast}}^{(1-\alpha,\Sigma^{\ast})}. (4.38)

Finally, by the definition of Φ1\Phi_{1}, one has

j=12Hj1,α;Ω(α,Σ)C~20,α;Ω(1α,Σ).\sum_{j=1}^{2}\|H_{j}\|_{1,\alpha;\Omega^{\ast}}^{(-\alpha,\Sigma^{\ast})}\leq C\|\tilde{\mathcal{F}}_{2}\|_{0,\alpha;\Omega^{\ast}}^{(1-\alpha,\Sigma^{\ast})}. (4.39)

Step 3: In this step, we are going to solve (4.15). The second equation in (4.15) implies that there exists a potential function ϕ2\phi_{2} such that

(z1ϕ2,z2ϕ2)=(K1,K2),ϕ2(L,0)=0.({\partial}_{z_{1}}\phi_{2},{\partial}_{z_{2}}\phi_{2})=(K_{1},K_{2}),\quad\phi_{2}(L^{\ast},0)=0. (4.40)

Then (4.15) can be rewritten as the following equation for ϕ2\phi_{2}:

{z1(z2z1ϕ2)+z2(z2z2ϕ2)=z21,z1ϕ2(0,z2)=~3,onΣ0,ϕ2(L,z2)=0,onΣL,z2ϕ2(z1,12)=4,onΣ,z2ϕ2(z1,0)=0,onΣa.\begin{cases}\begin{aligned} &{\partial}_{z_{1}}(z_{2}{\partial}_{z_{1}}\phi_{2})+{\partial}_{z_{2}}(z_{2}{\partial}_{z_{2}}\phi_{2})=z_{2}\mathcal{F}_{1},\\ &{\partial}_{z_{1}}\phi_{2}(0,z_{2})=\tilde{\mathcal{F}}_{3},\quad&{\rm{on}}\quad\Sigma_{0}^{\ast},\\ &\phi_{2}(L^{\ast},z_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L}^{\ast},\\ &{\partial}_{z_{2}}\phi_{2}(z_{1},\frac{1}{2})=\mathcal{F}_{4},\quad&{\rm{on}}\quad\Sigma^{\ast},\\ &{\partial}_{z_{2}}\phi_{2}(z_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a}^{\ast}.\\ \end{aligned}\end{cases} (4.41)

In order to deal with the singularity near z2=0z_{2}=0, we rewrite the problem (4.41) in the three dimensional setting. Define

ζ1=z1,ζ2=z2cosτ,ζ3=z2sinτ,τ[0,2π],\zeta_{1}=z_{1},\ \zeta_{2}=z_{2}\cos\tau,\ \zeta_{3}=z_{2}\sin\tau,\ \tau\in[0,2\pi],

and

E1={(ζ1,ζ2,ζ3):0<ζ1<L,ζ22+ζ3212},E2={(ζ2,ζ3):ζ22+ζ3212},\displaystyle E_{1}=\{(\zeta_{1},\zeta_{2},\zeta_{3}):0<\zeta_{1}<L^{\ast},\zeta_{2}^{2}+\zeta_{3}^{2}\leq\frac{1}{2}\},\quad E_{2}=\{(\zeta_{2},\zeta_{3}):\zeta_{2}^{2}+\zeta_{3}^{2}\leq\frac{1}{2}\},
Γw,ζ=[0,L]×{(ζ2,ζ3):ζ22+ζ32=12},\displaystyle\Gamma_{w,\zeta}=[0,L^{\ast}]\times\{(\zeta_{2},\zeta_{3}):\zeta_{2}^{2}+\zeta_{3}^{2}=\frac{1}{2}\},
Γ0,ζ={0}×{(ζ2,ζ3):ζ22+ζ3212},ΓL,ζ={L}×{(ζ2,ζ3):ζ22+ζ3212},\displaystyle\Gamma_{0,\zeta}=\{0\}\times\{(\zeta_{2},\zeta_{3}):\zeta_{2}^{2}+\zeta_{3}^{2}\leq\frac{1}{2}\},\quad\Gamma_{L^{\ast},\zeta}=\{L^{\ast}\}\times\{(\zeta_{2},\zeta_{3}):\zeta_{2}^{2}+\zeta_{3}^{2}\leq\frac{1}{2}\},
Ψ(𝜻)=ϕ2(ζ1,ζ22+ζ32)=Ψ(ζ1,|ζ|).\displaystyle\Psi(\bm{\zeta})=\phi_{2}(\zeta_{1},\sqrt{\zeta_{2}^{2}+\zeta_{3}^{2}})=\Psi(\zeta_{1},|\zeta^{\prime}|).

Then Ψ\Psi solves the following problem

{ΔΨ=1(ζ1,|ζ|),ζ1Ψ(0,|ζ|)=~3(|ζ|),onΓ0,ζ,Ψ(L,|ζ|)=0,onΓL,ζ,(ζ2ζ2+ζ3ζ3)Ψ(ζ1,|ζ|)=124(ζ1),onΓw,ζ.\begin{cases}\begin{aligned} &\Delta\Psi=\mathcal{F}_{1}(\zeta_{1},|\zeta^{\prime}|),\\ &{\partial}_{\zeta_{1}}\Psi(0,|\zeta^{\prime}|)=\tilde{\mathcal{F}}_{3}(|\zeta^{\prime}|),\quad&{\rm{on}}\quad\Gamma_{0,\zeta},\\ &\Psi(L^{\ast},|\zeta^{\prime}|)=0,\quad&{\rm{on}}\quad\Gamma_{L^{\ast},\zeta},\\ &(\zeta_{2}{\partial}_{\zeta_{2}}+\zeta_{3}{\partial}_{\zeta_{3}})\Psi(\zeta_{1},|\zeta^{\prime}|)=\frac{1}{2}\mathcal{F}_{4}(\zeta_{1}),\quad&{\rm{on}}\quad\Gamma_{w,\zeta}.\\ \end{aligned}\end{cases} (4.42)

First, the weak solution to (4.42) can be obtained as follows. ΨH1(E1)\Psi\in H^{1}(E_{1}) is said to be a weak solution to (4.42) if the following holds

(Ψ,ψ)=𝒫(ψ),forallψ{ψH1(E1):ψ=0onΓL,ζ},\mathcal{B}(\Psi,\psi)=\mathcal{P}(\psi),\quad{\rm{for}}\ {\rm{all}}\ \psi\in\{\psi\in H^{1}(E_{1}):\psi=0\quad{\rm{on}}\quad\Gamma_{L^{\ast},\zeta}\}, (4.43)

where

(Ψ,ψ)=E1Ψψd𝜻,\displaystyle\mathcal{B}(\Psi,\psi)=\int_{E_{1}}\nabla\Psi\nabla\psi{\mathrm{d}}\bm{\zeta},
𝒫(ψ)=E11ψd𝜻+E2~3ψdζ2dζ3+0L124(s)ψ(s,12)ds.\displaystyle\mathcal{P}(\psi)=-\int_{E_{1}}\mathcal{F}_{1}\psi{\mathrm{d}}\bm{\zeta}+\int_{E_{2}}\tilde{\mathcal{F}}_{3}\psi{\mathrm{d}}\zeta_{2}{\mathrm{d}}\zeta_{3}+\int_{0}^{L^{\ast}}\frac{1}{2}\mathcal{F}_{4}(s)\psi(s,\frac{1}{2}){\mathrm{d}}s.

By the Lax-Milgram theorem, there exists a unique weak solution ΨH1(E)\Psi\in H^{1}(E). Then multiplying Ψ\Psi on the sides of the equation (4.42) and integrating over E1E_{1} yield that

ΨL2(E1)2C10,α;E1(1α;Γw,ζ)ΨL2(E1)+C~3L2(E2)ΨL2(E2)+C4L2[0,L]ΨL2[0,L].\|\nabla\Psi\|_{L^{2}(E_{1})}^{2}\leq C\|\mathcal{F}_{1}\|_{0,\alpha;E_{1}}^{({1-\alpha;\Gamma_{w,\zeta}})}\|\Psi\|_{L^{2}(E_{1})}+C\|\tilde{\mathcal{F}}_{3}\|_{L^{2}(E_{2})}\|\Psi\|_{L^{2}(E_{2})}+C\|\mathcal{F}_{4}\|_{L^{2}[0,L^{\ast}]}\|\Psi\|_{{L^{2}[0,L^{\ast}]}}. (4.44)

Then it follows from Poincaré inequality and the trace theorem that one obtains

ΨH1(E1)C(10,α;E1(1α;Γw,ζ)+~31,α;E2(α,Γw,ζ)+40,α;[0,L]).\|\Psi\|_{H^{1}(E_{1})}\leq C\left(\|\mathcal{F}_{1}\|_{0,\alpha;E_{1}}^{({1-\alpha;\Gamma_{w,\zeta}})}+\|\tilde{\mathcal{F}}_{3}\|_{1,\alpha;E_{2}}^{(-\alpha,\Gamma_{w,\zeta})}+\|\mathcal{F}_{4}\|_{0,\alpha;[0,L^{\ast}]}\right). (4.45)

Next, we can follow the analogous argument as in Step 2 to obtain C0,αC^{0,\alpha} estimate for Ψ\Psi. Then the Schauder estimate in Theorem 4.6 of [13] implies that

Ψ2,α;E1(1α,Γw,ζ)C(10,α;E1(1α;Γw,ζ)+~31,α;E2(α,Γw,ζ)+40,α;[0,L]).\|\Psi\|_{2,\alpha;E_{1}}^{(-1-\alpha,\Gamma_{w,\zeta})}\leq C\left(\|\mathcal{F}_{1}\|_{0,\alpha;E_{1}}^{({1-\alpha;\Gamma_{w,\zeta}})}+\|\tilde{\mathcal{F}}_{3}\|_{1,\alpha;E_{2}}^{(-\alpha,\Gamma_{w,\zeta})}+\|\mathcal{F}_{4}\|_{0,\alpha;[0,L^{\ast}]}\right). (4.46)

Finally, it follows from the definition of ϕ2\phi_{2} that

j=12Kj1,α;Ω(α,Σ)C(10,α;Ω(1α;Σ)+~31,α;[0,12)(α;{12})+40,α;[0,L]).\sum_{j=1}^{2}\|K_{j}\|_{1,\alpha;\Omega^{\ast}}^{(-\alpha,\Sigma^{\ast})}\leq C\left(\|\mathcal{F}_{1}\|_{0,\alpha;\Omega^{\ast}}^{({1-\alpha;\Sigma^{\ast}})}+\|\tilde{\mathcal{F}}_{3}\|_{1,\alpha;[0,\frac{1}{2})}^{(-\alpha;\{\frac{1}{2}\})}+\|\mathcal{F}_{4}\|_{0,\alpha;[0,L^{\ast}]}\right). (4.47)

Step 4: By recalling the transformation (4.12) and combining the estimates (4.39) and (4.47), we conclude that the boundary value problem (4.10) has a unique solution (W1,W2)(C1,α(α,Σ)(Ω))2(W_{1},W_{2})\in\left(C_{1,\alpha}^{(-\alpha,\Sigma)}(\Omega)\right)^{2} satisfying

j=12Wj1,α;Ω(α,Σ)C(j=12j0,α;Ω(1α,Σ)+31,α;[0,12)(α;{12})+40,α;[0,L]).\sum_{j=1}^{2}\|W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq C\left(\sum_{j=1}^{2}\|\mathcal{F}_{j}\|_{0,\alpha;\Omega}^{(1-\alpha,\Sigma)}+\|\mathcal{F}_{3}\|_{1,\alpha;[0,\frac{1}{2})}^{(-\alpha;\{\frac{1}{2}\})}+\|\mathcal{F}_{4}\|_{0,\alpha;[0,L]}\right).

Thus, the proof of Lemma 4.1 is completed. ∎

4.3 Solving the nonlinear boundary value problem

For a given 𝐖^𝒥(δ1)\hat{{\bf W}}\in\mathcal{J}(\delta_{1}), it follows from Lemma 4.1 that the problem (4.10) has a unique solution (W1,W2)(C1,α(α,Σ)(Ω))2(W_{1},W_{2})\in\left(C_{1,\alpha}^{(-\alpha,\Sigma)}(\Omega)\right)^{2} satisfying the estimate (4.11). Define a map 𝒯\mathcal{T} as follows

𝒯(𝐖^)=(𝐖),for𝐖^𝒥(δ1).\mathcal{T}(\hat{{\bf W}})=({{\bf W}}),\quad{\rm{for}}\ \hat{{\bf W}}\in\mathcal{J}(\delta_{1}). (4.48)

The estimate (4.11), together with (4.3), (4.7) and (4.9), yields

j=13Wj1,α;Ω(α,Σ)𝒞1(δ12+δ1w0,α;[0,L]+w0,α;[0,L]+σcd),\displaystyle\sum_{j=1}^{3}\|W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq\mathcal{C}_{1}\left(\delta_{1}^{2}+\delta_{1}\|w\|_{0,\alpha;[0,L]}+\|w\|_{0,\alpha;[0,L]}+\sigma_{cd}\right), (4.49)

where 𝒞1>0\mathcal{C}_{1}>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha).

We assume that

w0,α;[0,L]δ2,\|w\|_{0,\alpha;[0,L]}\leq\delta_{2}, (4.50)

where 𝒞1δ2δ14\mathcal{C}_{1}\delta_{2}\leq\frac{\delta_{1}}{4} with 𝒞1\mathcal{C}_{1} given in (4.49). Let σ2=14(𝒞12+𝒞1)\sigma_{2}=\frac{1}{4(\mathcal{C}_{1}^{2}+\mathcal{C}_{1})} and choose δ1=4𝒞1σcd\delta_{1}=4\mathcal{C}_{1}\sigma_{cd}. Then if σcdσ2\sigma_{cd}\leq\sigma_{2}, one has

j=13Wj1,α;Ω(α,Σ)δ1.\sum_{j=1}^{3}\|W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq\delta_{1}. (4.51)

Hence 𝒯\mathcal{T} maps 𝒥(δ1)\mathcal{J}(\delta_{1}) into itself.

In the following, we will show that 𝒯\mathcal{T} is a contraction in 𝒥(δ1)\mathcal{J}(\delta_{1}). Let 𝐖^k𝒥(δ1),k=1,2\hat{{\bf W}}^{k}\in\mathcal{J}(\delta_{1}),k=1,2, one has 𝐖k=𝒯(𝐖^k){\bf W}^{k}=\mathcal{T}(\hat{{\bf W}}^{k}) for k=1,2k=1,2. Define

𝐘=𝐖1𝐖2,𝐘^=𝐖^1𝐖^2.{\bf Y}={\bf W}^{1}-{\bf W}^{2},\quad\quad\hat{{\bf Y}}=\hat{{\bf W}}^{1}-\hat{{\bf W}}^{2}.

Then it follows from (4.2) that

Y3(y1,y2)=Λ^(y2)(1r^1(y1,y2)1r^2(y1,y2)),Y_{3}(y_{1},y_{2})={\hat{\Lambda}(y_{2})}\left(\frac{1}{\hat{r}^{1}(y_{1},y_{2})}-\frac{1}{\hat{r}^{2}(y_{1},y_{2})}\right), (4.52)

where

r^i(y1,y2)=(20y22sρ^(W^1i+ub,W^2i,W^3i,A~0,B~0)(W^1i+ub)(y1,s)ds)12.\hat{r}^{i}(y_{1},y_{2})=\left(2\int_{0}^{y_{2}}\frac{2s}{\hat{\rho}(\hat{W}_{1}^{i}+u_{b}^{-},\hat{W}_{2}^{i},\hat{W}_{3}^{i},\tilde{A}_{0},\tilde{B}_{0})(\hat{W}_{1}^{i}+u_{b}^{-})(y_{1},s)}{\mathrm{d}}s\right)^{\frac{1}{2}}.

Next, we obtain that (Y1,Y2)(Y_{1},Y_{2}) satisfies

{(1(Mb)2)y1Y1+y2Y2+1y2Y2=1(𝐖^1,𝐖^1,A^,B^)1(𝐖^2,𝐖^2,A^,B^),y1Y2y2Y1=2(𝐖^1,𝐖^1,A^,B^)2(𝐖^2,𝐖^2,A^,B^),Y1(0,y2)=(3(𝐖^1,J^,A^,B^)3(𝐖^2,J^,A^,B^))(0,y2),Y2(L,y2)=0,Y2(y1,12)=(4(𝐖^1,gcd)4(𝐖^2,gcd))(y1,12),Y2(y1,0)=0.\begin{cases}\begin{aligned} &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}Y_{1}+{\partial}_{y_{2}}Y_{2}+\frac{1}{y_{2}}Y_{2}=\mathcal{F}_{1}(\hat{{\bf W}}^{1},\nabla{\hat{\bf W}}^{1},\hat{A},\hat{B})-\mathcal{F}_{1}(\hat{{\bf W}}^{2},\nabla{\hat{\bf W}}^{2},\hat{A},\hat{B}),\\ &{\partial}_{y_{1}}Y_{2}-{\partial}_{y_{2}}Y_{1}=\mathcal{F}_{2}(\hat{{\bf W}}^{1},\nabla{\hat{\bf W}}^{1},\hat{A},\hat{B})-\mathcal{F}_{2}(\hat{{\bf W}}^{2},\nabla{\hat{\bf W}}^{2},\hat{A},\hat{B}),\\ &Y_{1}(0,y_{2})=(\mathcal{F}_{3}(\hat{{\bf W}}^{1},\hat{J},\hat{A},\hat{B})-\mathcal{F}_{3}(\hat{{\bf W}}^{2},\hat{J},\hat{A},\hat{B}))(0,y_{2}),\\ &Y_{2}(L,y_{2})=0,\\ &Y_{2}(y_{1},\frac{1}{2})=(\mathcal{F}_{4}(\hat{{\bf W}}^{1},g_{cd})-\mathcal{F}_{4}(\hat{{\bf W}}^{2},g_{cd}))(y_{1},\frac{1}{2}),\\ &Y_{2}(y_{1},0)=0.\\ \end{aligned}\end{cases} (4.53)

where j(𝐖^k),j=1,2,3,4,k=1,2,3\mathcal{F}_{j}(\hat{{\bf W}}^{k}),j=1,2,3,4,k=1,2,3 are functions defined in (4.10) by replacing 𝐖^\hat{{\bf W}} with 𝐖^k\hat{{\bf W}}^{k} respectively. Then it follows from Lemma 4.1 that one can derive

j=13Yj1,α;Ω(α,Σ)\displaystyle\sum_{j=1}^{3}\|Y_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)} 𝒞2(j=12j(𝐖^1,𝐖^1,A^,B^)j(𝐖^2,𝐖^2,A^,B^)0,α;Ω(1α,Σ)\displaystyle\leq\mathcal{C}_{2}\left(\sum_{j=1}^{2}\|\mathcal{F}_{j}(\hat{{\bf W}}^{1},\nabla{\hat{\bf W}}^{1},\hat{A},\hat{B})-\mathcal{F}_{j}(\hat{{\bf W}}^{2},\nabla{\hat{\bf W}}^{2},\hat{A},\hat{B})\|_{0,\alpha;\Omega}^{(1-\alpha,\Sigma)}\right. (4.54)
+3(𝐖^1,J^,A^,B^)3(𝐖^2,J^,A^,B^)1,α;[0,12)(α;{12})\displaystyle\qquad\quad\left.+\|\mathcal{F}_{3}(\hat{{\bf W}}^{1},\hat{J},\hat{A},\hat{B})-\mathcal{F}_{3}(\hat{{\bf W}}^{2},\hat{J},\hat{A},\hat{B})\|_{1,\alpha;[0,\frac{1}{2})}^{(-\alpha;\{\frac{1}{2}\})}\right.
+4(𝐖^1,gcd)4(𝐖^2,gcd)0,α;[0,L]+Λ^(1r^11r^2)1,α;Ω(α,Σ))\displaystyle\qquad\quad\left.+\|\mathcal{F}_{4}(\hat{{\bf W}}^{1},g_{cd})-\mathcal{F}_{4}(\hat{{\bf W}}^{2},g_{cd})\|_{0,\alpha;[0,L]}+\left\|{\hat{\Lambda}}\left(\frac{1}{\hat{r}^{1}}-\frac{1}{\hat{r}^{2}}\right)\right\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\right)
𝒞2(δ1+δ2+σcd)j=13Y^j1,α;Ω(α,Σ),\displaystyle\leq\mathcal{C}_{2}(\delta_{1}+\delta_{2}+\sigma_{cd})\sum_{j=1}^{3}\|\hat{Y}_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)},

where 𝒞2>0\mathcal{C}_{2}>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha). Setting

σ3=min{σ2,14𝒞2(4𝒞1+2)}.\sigma_{3}=\min\left\{\sigma_{2},\frac{1}{4\mathcal{C}_{2}(4\mathcal{C}_{1}+2)}\right\}. (4.55)

Then for σcdσ3\sigma_{cd}\leq\sigma_{3}, one has 𝒞2(δ1+δ2+σcd)𝒞2(4𝒞1+2)σcd14\mathcal{C}_{2}(\delta_{1}+\delta_{2}+\sigma_{cd})\leq\mathcal{C}_{2}(4\mathcal{C}_{1}+2)\sigma_{cd}\leq\frac{1}{4}. Hence the mapping 𝒯\mathcal{T} is a contraction mapping so that 𝒯\mathcal{T} has a unique fixed point in 𝒥(δ1)\mathcal{J}(\delta_{1}).

5 The construction of the contact discontinuity surface

Up to now, for a given function gcd(y1)=0y1w(s)ds+12g_{cd}(y_{1})=\int_{0}^{y_{1}}w(s){\mathrm{d}}s+\frac{1}{2} satisfying gcd(L)=0g_{cd}^{\prime}(L)=0, we have obtained the solution (ux,ur,uθ)(u_{x},u_{r},u_{\theta}) for the nonlinear boundary value problem (3.19)-(3.20). To complete the proof of Theorem 2.3, we will use the implicit function theorem to find the contact discontinuity gcd(y1)g_{cd}(y_{1}) such that (3.21) is satisfied.

First, define a Banach space

𝒱={w:w0,α;[0,L]<}.\mathcal{V}=\{w:\|w\|_{0,\alpha;[0,L]}<\infty\}.

Set

𝒱1={w:w(L)=0,w0,α;[0,L]<}\mathcal{V}_{1}=\{w:w(L)=0,\|w\|_{0,\alpha;[0,L]}<\infty\} (5.1)

and

𝒱1(δ2)={w𝒱1:w0,α;[0,L]δ2},\mathcal{V}_{1}(\delta_{2})=\{w\in\mathcal{V}_{1}:\|w\|_{0,\alpha;[0,L]}\leq\delta_{2}\}, (5.2)

where δ2\delta_{2} is defined in (4.50). Then for any w𝒱1(δ2)w\in\mathcal{V}_{1}(\delta_{2}), the nonlinear boundary value problem (3.19)-(3.20) has a unique solution (ux,ur,uθ)(u_{x},u_{r},u_{\theta}) satisfying

uxub1,α;Ω(α,Σ)+ur1,α;Ω(α,Σ)+uθ1,α;Ω(α,Σ)4𝒞1σcd.\|u_{x}-u_{b}^{-}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|u_{r}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|u_{\theta}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq 4\mathcal{C}_{1}\sigma_{cd}. (5.3)

Let

𝒱0=C1,α([0,1/2])×C1,α([0,1/2])×C1,α([0,1/2])×C1,α([0,1/2]).\displaystyle\mathcal{V}_{0}=C^{1,\alpha}([0,1/2])\times C^{1,\alpha}([0,1/2])\times C^{1,\alpha}([0,1/2])\times C^{1,\alpha}([0,1/2]).

Then we set

𝒱2(δ3)={𝝋0𝒱0:𝝋0𝝋b𝒱0δ3},\mathcal{V}_{2}(\delta_{3})=\{\bm{\varphi}_{0}\in\mathcal{V}_{0}:\|\bm{\varphi}_{0}-\bm{\varphi}_{b}\|_{\mathcal{V}_{0}}\leq\delta_{3}\}, (5.4)

where

𝝋0=(J0,ν0,A0,B0)and𝝋b=(Jb,0,Ab,Bb).\bm{\varphi}_{0}=(J_{0},\nu_{0},A_{0},B_{0})\quad{\rm{and}}\quad\bm{\varphi}_{b}=(J_{b}^{-},0,A_{b}^{-},B_{b}^{-}).

Define a map 𝒬:𝒱2(δ3)×𝒱1(δ2)𝒱\mathcal{Q}:\mathcal{V}_{2}(\delta_{3})\times\mathcal{V}_{1}(\delta_{2})\rightarrow\mathcal{V} by

𝒬(𝝋0,w):=N(W1,W2,W3,A~0,B~0)(y1,12)Pb,\mathcal{Q}(\bm{\varphi}_{0},w):=N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})(y_{1},\frac{1}{2})-P_{b}, (5.5)

where

N(W1,W2,W3,A~0,B~0)(y1,12)=A~0(12)(ρ(W1+ub,W2,W3,A~0,B~0))γ(y1,12).N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})(y_{1},\frac{1}{2})=\tilde{A}_{0}\left(\frac{1}{2}\right)\left(\rho(W_{1}+u_{b}^{-},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})\right)^{\gamma}(y_{1},\frac{1}{2}).

Hence (3.21) can be written as the equation

𝒬(𝝋0,w)=0,\mathcal{Q}(\bm{\varphi}_{0},w)=0, (5.6)

which will be solved (5.6) by employing the implicit function theorem. For the precise statement of the implicit function theorem, one can see Theorem 3.3 in [20]. We will verify the conditions (i)\rm(i), (ii)\rm(ii) and (iii)\rm(iii) in Theorem 3.3 of [20].

Obviously,

𝒬(𝝋b,0)=0.\mathcal{Q}(\bm{\varphi}_{b},0)=0.

Next, the proof is divided into two steps.

Step 1. Differentiability of 𝒬\mathcal{Q}.

Given any w𝒱1(δ2),w1𝒱1w\in\mathcal{V}_{1}(\delta_{2}),w_{1}\in\mathcal{V}_{1}, and τ>0\tau>0, let (ux,ur,uθ)=(W1+ub,W2,W3)(u_{x},u_{r},u_{\theta})=(W_{1}+u_{b}^{-},W_{2},W_{3}) be the solution of (3.19) with the following boundary conditions:

{ρux(0,y2)=J~0(y2),uθ(0,y2)=ν~0(y2),onΣ0,ur(L,y2)=0,onΣL,ur(y1,12)=u~x(y1,12)w(y1),onΣ,ur(y1,0)=0,onΣa,\begin{cases}\rho u_{x}(0,y_{2})=\tilde{J}_{0}(y_{2}),\ u_{\theta}(0,y_{2})=\tilde{\nu}_{0}(y_{2}),\quad&{\rm{on}}\quad\Sigma_{0},\\ u_{r}(L,y_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L},\\ u_{r}(y_{1},\frac{1}{2})=\tilde{u}_{x}(y_{1},\frac{1}{2})w(y_{1}),\quad&{\rm{on}}\quad\Sigma,\\ u_{r}(y_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a},\\ \end{cases}

and (u~x,u~r,u~θ)=(W~1+ub,W~2,W~3)(\tilde{u}_{x},\tilde{u}_{r},\tilde{u}_{\theta})=(\tilde{W}_{1}+u_{b}^{-},\tilde{W}_{2},\tilde{W}_{3}) be the solution of (3.19) with the following boundary conditions:

{ρ~u~x(0,y2)=J~0(y2),u~θ(0,y2)=ν~0(y2),onΣ0,u~r(L,y2)=0,onΣL,u~r(y1,12)=u~x(y1,12)(w+τw1)(y1),onΣ,u~r(y1,0)=0,onΣa.\begin{cases}\tilde{\rho}\tilde{u}_{x}(0,y_{2})=\tilde{J}_{0}(y_{2}),\ \tilde{u}_{\theta}(0,y_{2})=\tilde{\nu}_{0}(y_{2}),\quad&{\rm{on}}\quad\Sigma_{0},\\ \tilde{u}_{r}(L,y_{2})=0,\quad&{\rm{on}}\quad\Sigma_{L},\\ \tilde{u}_{r}(y_{1},\frac{1}{2})=\tilde{u}_{x}(y_{1},\frac{1}{2})(w+\tau w_{1})(y_{1}),\quad&{\rm{on}}\quad\Sigma,\\ \tilde{u}_{r}(y_{1},0)=0,\quad&{\rm{on}}\quad\Sigma_{a}.\\ \end{cases}

Then it follows from Section 4 that

W~3=0(𝐖~,Λ^,A^,B^),andW3=0(𝐖,Λ^,A^,B^),\tilde{W}_{3}=\mathcal{F}_{0}(\tilde{\bf W},\hat{\Lambda},\hat{A},\hat{B}),\quad{\rm{and}}\quad\ W_{3}=\mathcal{F}_{0}({\bf W},\hat{\Lambda},\hat{A},\hat{B}), (5.7)

where

0(𝐖~,Λ^,A^,B^):=Λ^r~=Λ^(20y22sρ(W~1+ub,W~2,W~3,A~0,B~0)(W~1+ub)(y1,s)ds)12,\displaystyle\mathcal{F}_{0}(\tilde{\bf W},\hat{\Lambda},\hat{A},\hat{B}):=\frac{\hat{\Lambda}}{\tilde{r}}=\frac{\hat{\Lambda}}{\left(2\int_{0}^{y_{2}}\frac{2s}{\rho(\tilde{W}_{1}+u_{b}^{-},\tilde{W}_{2},\tilde{W}_{3},\tilde{A}_{0},\tilde{B}_{0})(\tilde{W}_{1}+u_{b}^{-})(y_{1},s)}{\mathrm{d}}s\right)^{\frac{1}{2}}},
0(𝐖,Λ^,A^,B^):=Λ^r=Λ^(20y22sρ(W1+ub,W2,W3,A~0,B~0)(W1+ub)(y1,s)ds)12.\displaystyle\mathcal{F}_{0}({\bf W},\hat{\Lambda},\hat{A},\hat{B}):=\frac{\hat{\Lambda}}{r}=\frac{\hat{\Lambda}}{\left(2\int_{0}^{y_{2}}\frac{2s}{\rho(W_{1}+u_{b}^{-},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})(W_{1}+u_{b}^{-})(y_{1},s)}{\mathrm{d}}s\right)^{\frac{1}{2}}}.

Furthermore, (W1,W2)(W_{1},W_{2}) and (W~1,W~2)(\tilde{W}_{1},\tilde{W}_{2}) satisfy

{(1(Mb)2)y1W1+y2W2+1y2W2=1(𝐖,𝐖,A^,B^),y1W2y2W1=2(𝐖,𝐖,A^,B^),W1(0,y2)=3(𝐖,J^,A^,B^)(0,y2),W2(L,y2)=0,W2(y1,12)=(ub+W1(y1,12))w(y1),W2(y1,0)=0,\begin{cases}\begin{aligned} &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}+{\partial}_{y_{2}}W_{2}+\frac{1}{y_{2}}W_{2}=\mathcal{F}_{1}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B}),\\ &{\partial}_{y_{1}}W_{2}-{\partial}_{y_{2}}W_{1}=\mathcal{F}_{2}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B}),\\ &W_{1}(0,y_{2})=\mathcal{F}_{3}({{\bf W}},\hat{J},\hat{A},\hat{B})(0,y_{2}),\\ &W_{2}(L,y_{2})=0,\\ &W_{2}(y_{1},\frac{1}{2})=\left(u_{b}^{-}+W_{1}(y_{1},\frac{1}{2})\right)w(y_{1}),\\ &W_{2}(y_{1},0)=0,\\ \end{aligned}\end{cases} (5.8)

and

{(1(Mb)2)y1W~1+y2W~2+1y2W~2=1(𝐖~,𝐖~,A^,B^),y1W~2y2W~1=2(𝐖~,𝐖~,A^,B^),W~1(0,y2)=3(𝐖~,J^,A^,B^)(0,y2),W~2(L,y2)=0,W~2(y1,12)=(ub+W~1(y1,12))(w+τw1)(y1),W~2(y1,0)=0.\begin{cases}\begin{aligned} &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}\tilde{W}_{1}+{\partial}_{y_{2}}\tilde{W}_{2}+\frac{1}{y_{2}}\tilde{W}_{2}=\mathcal{F}_{1}(\tilde{\bf W},\nabla{\tilde{\bf W}},\hat{A},\hat{B}),\\ &{\partial}_{y_{1}}\tilde{W}_{2}-{\partial}_{y_{2}}\tilde{W}_{1}=\mathcal{F}_{2}(\tilde{{\bf W}},\nabla{\tilde{\bf W}},\hat{A},\hat{B}),\\ &\tilde{W}_{1}(0,y_{2})=\mathcal{F}_{3}(\tilde{{\bf W}},\hat{J},\hat{A},\hat{B})(0,y_{2}),\\ &\tilde{W}_{2}(L,y_{2})=0,\\ &\tilde{W}_{2}(y_{1},\frac{1}{2})=\left(u_{b}^{-}+\tilde{W}_{1}(y_{1},\frac{1}{2})\right)(w+\tau w_{1})(y_{1}),\\ &\tilde{W}_{2}(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.9)

Choosing τ1>0\tau_{1}>0 such that τ1w10,α;[0,L]δ2\tau_{1}\|w_{1}\|_{0,\alpha;[0,L]}\leq\delta_{2}. Then for τ(0,τ1)\tau\in(0,\tau_{1}), it follows from Section 4 that one gets

j=13W~j1,α;Ω(α,Σ)4𝒞1σcd.\sum_{j=1}^{3}\|\tilde{W}_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq 4\mathcal{C}_{1}\sigma_{cd}. (5.10)

Denote

𝐖τ=𝐖~𝐖τ.{\bf W}^{\tau}=\frac{\tilde{{\bf W}}-{{\bf W}}}{\tau}.

It follows from (5.7)-(5.8) that one obtains

{W3τ=~0(𝐖~,Λ^,A^,B^)~0(𝐖,Λ^,A^,B^)τ,(1(Mb)2)y1W1τ+y2W2τ+1y2W2τ=1(𝐖~,𝐖~,A^,B^)1(𝐖,𝐖,A^,B^)τ,y1W2τy2W1τ=2(𝐖~,𝐖~,A^,B^)2(𝐖,𝐖,A^,B^)τ,W1τ(0,y2)=3(𝐖~,J^,A^,B^)3(𝐖,J^,A^,B^)τ(0,y2),W2τ(L,y2)=0,W2τ(y1,12)=(ub+W~1(y1,12))w1(y1)+W1τ(y1,12)w(y1),W2τ(y1,0)=0.\begin{cases}\begin{aligned} &W_{3}^{\tau}=\frac{\tilde{\mathcal{F}}_{0}(\tilde{\bf W},\hat{\Lambda},\hat{A},\hat{B})-\tilde{\mathcal{F}}_{0}({\bf W},\hat{\Lambda},\hat{A},\hat{B})}{\tau},\\ &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}^{\tau}+{\partial}_{y_{2}}W_{2}^{\tau}+\frac{1}{y_{2}}W_{2}^{\tau}=\frac{\mathcal{F}_{1}(\tilde{{\bf W}},\nabla{\tilde{\bf W}},\hat{A},\hat{B})-\mathcal{F}_{1}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})}{\tau},\\ &{\partial}_{y_{1}}W_{2}^{\tau}-{\partial}_{y_{2}}W_{1}^{\tau}=\frac{\mathcal{F}_{2}(\tilde{{\bf W}},\nabla{\tilde{\bf W}},\hat{A},\hat{B})-\mathcal{F}_{2}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})}{\tau},\\ &W_{1}^{\tau}(0,y_{2})=\frac{\mathcal{F}_{3}(\tilde{{\bf W}},\hat{J},\hat{A},\hat{B})-\mathcal{F}_{3}({{\bf W}},\hat{J},\hat{A},\hat{B})}{\tau}(0,y_{2}),\\ &W_{2}^{\tau}(L,y_{2})=0,\\ &W_{2}^{\tau}(y_{1},\frac{1}{2})=\left(u_{b}^{-}+\tilde{W}_{1}(y_{1},\frac{1}{2})\right)w_{1}(y_{1})+W_{1}^{\tau}(y_{1},\frac{1}{2})w(y_{1}),\\ &W_{2}^{\tau}(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.11)

Then for τ(0,τ1)\tau\in(0,\tau_{1}), one can apply (5.10) and (4.11) to obtain the following estimate:

j=13Wjτ1,α;Ω(α,Σ)\displaystyle\sum_{j=1}^{3}\|W_{j}^{\tau}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)} 𝒞30(𝐖~,Λ^,A^,B^)0(𝐖,Λ^,A^,B^)τ1,α;Ω(α,Σ)\displaystyle\leq\mathcal{C}_{3}\left\|\frac{\mathcal{F}_{0}(\tilde{{\bf W}},\hat{\Lambda},\hat{A},\hat{B})-\mathcal{F}_{0}({{\bf W}},\hat{\Lambda},\hat{A},\hat{B})}{\tau}\right\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)} (5.12)
+𝒞3(j=12j(𝐖~,𝐖~,A^,B^)j(𝐖,𝐖,A^,B^)τ0,α;Ω(1α,Σ))\displaystyle\quad+\mathcal{C}_{3}\left(\sum_{j=1}^{2}\left\|\frac{\mathcal{F}_{j}(\tilde{{\bf W}},\nabla{\tilde{\bf W}},\hat{A},\hat{B})-\mathcal{F}_{j}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})}{\tau}\right\|_{0,\alpha;\Omega}^{(1-\alpha,\Sigma)}\right)
+𝒞33(𝐖~,J^,A^,B^)3(𝐖,J^,A^,B^)τ1,α;[0,12)(α;{12})\displaystyle\quad+\mathcal{C}_{3}\left\|\frac{\mathcal{F}_{3}(\tilde{{\bf W}},\hat{J},\hat{A},\hat{B})-\mathcal{F}_{3}({{\bf W}},\hat{J},\hat{A},\hat{B})}{\tau}\right\|_{1,\alpha;[0,\frac{1}{2})}^{(-\alpha;\{\frac{1}{2}\})}
+𝒞3(ub+W~1)w1+W1τw0,α;[0,L]\displaystyle\quad+\mathcal{C}_{3}\|(u_{b}^{-}+\tilde{W}_{1})w_{1}+W_{1}^{\tau}w\|_{0,\alpha;[0,L]}
𝒞3(δ2+4𝒞1σcd)j=13Wjτ1,α;Ω(α,Σ)+4𝒞3𝒞1σcdw10,α;[0,L]\displaystyle\leq\mathcal{C}_{3}\left(\delta_{2}+4\mathcal{C}_{1}\sigma_{cd}\right)\sum_{j=1}^{3}\|W_{j}^{\tau}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+4\mathcal{C}_{3}\mathcal{C}_{1}\sigma_{cd}\|w_{1}\|_{0,\alpha;[0,L]}
+𝒞3w10,α;[0,L]\displaystyle\quad+\mathcal{C}_{3}\|w_{1}\|_{0,\alpha;[0,L]}
𝒞3((4𝒞1+1)σcd)j=13Wjτ1,α;Ω(α,Σ)+𝒞3((4𝒞1+1)σcd)w10,α;[0,L]\displaystyle\leq\mathcal{C}_{3}((4\mathcal{C}_{1}+1)\sigma_{cd})\sum_{j=1}^{3}\|W_{j}^{\tau}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\mathcal{C}_{3}((4\mathcal{C}_{1}+1)\sigma_{cd})\|w_{1}\|_{0,\alpha;[0,L]}
+𝒞3w10,α;[0,L],\displaystyle\quad+\mathcal{C}_{3}\|w_{1}\|_{0,\alpha;[0,L]},

where 𝒞3>0\mathcal{C}_{3}>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha). Setting

σ4=min{σ3,14𝒞3(1+4𝒞1)},\sigma_{4}=\min\left\{\sigma_{3},\frac{1}{4\mathcal{C}_{3}(1+4\mathcal{C}_{1})}\right\}, (5.13)

where σ3\sigma_{3} is defined in (4.55). Then for τ(0,τ1)\tau\in(0,\tau_{1}) and σcdσ4\sigma_{cd}\leq\sigma_{4}, one has

j=13Wjτ1,α;Ω(α,Σ)(2𝒞3+2)w10,α;[0,L].\sum_{j=1}^{3}\|W_{j}^{\tau}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq(2\mathcal{C}_{3}+2)\|w_{1}\|_{0,\alpha;[0,L]}. (5.14)

Hence there exists a subsequence {τk}k=1\{\tau_{k}\}_{k=1}^{\infty} such that (W1τk,W2τk,W3τk)(W_{1}^{\tau_{k}},W_{2}^{\tau_{k}},W_{3}^{\tau_{k}}) converges to(W10,W20,W30)(W_{1}^{0},W_{2}^{0},W_{3}^{0}) in C1,α(α,Σ)(Ω)C_{1,\alpha^{\prime}}^{(-\alpha^{\prime},\Sigma)}(\Omega) as τk0\tau_{k}\rightarrow 0 for some 0<α<α0<\alpha^{\prime}<\alpha. The estimate (5.14) also implies that (W10,W20,W30)(C1,α(α,Σ)(Ω))3(W_{1}^{0},W_{2}^{0},W_{3}^{0})\in\left(C_{1,\alpha}^{(-\alpha,\Sigma)}(\Omega)\right)^{3} and

j=13Wj01,α;Ω(α,Σ)(2𝒞3+2)w10,α;[0,L].\sum_{j=1}^{3}\|W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq(2\mathcal{C}_{3}+2)\|w_{1}\|_{0,\alpha;[0,L]}. (5.15)

Define a map Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w) by

Dw𝒬(𝝋0,w)(w1)=j=13WjN(W1,W2,W3,A~0,B~0)Wj0(y1,12).\displaystyle D_{w}\mathcal{Q}(\bm{\varphi}_{0},w)(w_{1})=\sum_{j=1}^{3}{\partial}_{W_{j}}N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})W_{j}^{0}(y_{1},\frac{1}{2}). (5.16)

Then (5.15) implies that Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w) is a linear mapping from 𝒱1\mathcal{V}_{1} to 𝒱\mathcal{V}. Next, we need to show that Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w) is the Fréchet derivative of the functional 𝒬(𝝋0,w)\mathcal{Q}(\bm{\varphi}_{0},w) with respect to ww. To this end, It follows from (5.11) that 𝐖0{\bf W}^{0} satisfies

{W30=j=13Wj0(𝐖,Λ^,A^,B^)Wj0,(1(Mb)2)y1W10+y2W20+1y2W20=j=13(Wj1(𝐖,𝐖,A^,B^)Wj0+Wj1(𝐖,𝐖,A^,B^)Wj0),y1W20y2W10=j=13(Wj2(𝐖,𝐖,A^,B^)Wj0+Wj2(𝐖,𝐖,A^,B^)Wj0),W10(0,y2)=(j=13Wj3(𝐖,J^,A^,B^)Wj0)(0,y2),W20(L,y2)=0,W20(y1,12)=(ub+W1(y1,12))w1(y1)+W10(y1,12)w(y1),W20(y1,0)=0.\begin{cases}\begin{aligned} &W_{3}^{0}=\sum_{j=1}^{3}{\partial}_{W_{j}}\mathcal{F}_{0}({{\bf W}},\hat{\Lambda},\hat{A},\hat{B})W_{j}^{0},\\ &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}^{0}+{\partial}_{y_{2}}W_{2}^{0}+\frac{1}{y_{2}}W_{2}^{0}\\ &=\sum_{j=1}^{3}\left({\partial}_{W_{j}}\mathcal{F}_{1}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})W_{j}^{0}+{\partial}_{\nabla W_{j}}\mathcal{F}_{1}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})\nabla W_{j}^{0}\right),\\ &{\partial}_{y_{1}}W_{2}^{0}-{\partial}_{y_{2}}W_{1}^{0}\\ &=\sum_{j=1}^{3}\left({\partial}_{W_{j}}\mathcal{F}_{2}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})W_{j}^{0}+{\partial}_{\nabla W_{j}}\mathcal{F}_{2}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})\nabla W_{j}^{0}\right),\\ &W_{1}^{0}(0,y_{2})=\left(\sum_{j=1}^{3}{\partial}_{W_{j}}\mathcal{F}_{3}({{\bf W}},\hat{J},\hat{A},\hat{B})W_{j}^{0}\right)(0,y_{2}),\\ &W_{2}^{0}(L,y_{2})=0,\ W_{2}^{0}(y_{1},\frac{1}{2})=\left(u_{b}^{-}+W_{1}(y_{1},\frac{1}{2})\right)w_{1}(y_{1})+W_{1}^{0}(y_{1},\frac{1}{2})w(y_{1}),\ W_{2}^{0}(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.17)

By taking difference of (5.11) and (5.17), the following estimate can be derived:

j=13WjτWj01,α;Ω(α,Σ)\displaystyle\sum_{j=1}^{3}\|W_{j}^{\tau}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)} (5.18)
𝒞40(𝐖~,Λ^,A^,B^)0(𝐖,Λ^,A^,B^)τj=13Wj0(𝐖,Λ^,A^,B^)Wj01,α;Ω(α,Σ)\displaystyle\leq\mathcal{C}_{4}\left\|\frac{\mathcal{F}_{0}(\tilde{\bf W},\hat{\Lambda},\hat{A},\hat{B})-\mathcal{F}_{0}({\bf W},\hat{\Lambda},\hat{A},\hat{B})}{\tau}-\sum_{j=1}^{3}{\partial}_{W_{j}}\mathcal{F}_{0}({{\bf W}},\hat{\Lambda},\hat{A},\hat{B})W_{j}^{0}\right\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}
+𝒞4i=12i(𝐖~,𝐖~,A^,B^)i(𝐖,𝐖,A^,B^)τ\displaystyle\quad+\mathcal{C}_{4}\sum_{i=1}^{2}\left\|\frac{\mathcal{F}_{i}(\tilde{{\bf W}},\nabla{\tilde{{\bf W}}},\hat{A},\hat{B})-\mathcal{F}_{i}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})}{\tau}\right.
j=13(Wji(𝐖,𝐖,A^,B^)Wj0+Wji(𝐖,𝐖,A^,B^)Wj0)0,α;Ω(1α,Σ)\displaystyle\qquad\qquad\quad\left.-\sum_{j=1}^{3}\left({\partial}_{W_{j}}\mathcal{F}_{i}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})W_{j}^{0}+{\partial}_{\nabla W_{j}}\mathcal{F}_{i}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B})\nabla W_{j}^{0}\right)\right\|_{0,\alpha;\Omega}^{(1-\alpha,\Sigma)}
+𝒞43(𝐖~,J^,A^,B^)3(𝐖,J^,A^,B^)τj=13Wj3(𝐖,J^,A^,B^)Wj01,α;[0,12)(α;{12})\displaystyle\quad+\mathcal{C}_{4}\left\|\frac{\mathcal{F}_{3}(\tilde{{\bf W}},\hat{J},\hat{A},\hat{B})-\mathcal{F}_{3}({{\bf W}},\hat{J},\hat{A},\hat{B})}{\tau}-\sum_{j=1}^{3}{\partial}_{W_{j}}\mathcal{F}_{3}({{\bf W}},\hat{J},\hat{A},\hat{B})W_{j}^{0}\right\|_{1,\alpha;[0,\frac{1}{2})}^{(-\alpha;\{\frac{1}{2}\})}
+𝒞4τW1τw1+(W1τW10)w0,α;[0,L]\displaystyle\quad+\mathcal{C}_{4}\|\tau W_{1}^{\tau}w_{1}+(W_{1}^{\tau}-W_{1}^{0})w\|_{0,\alpha;[0,L]}
𝒞4(δ2+4𝒞1σcd)j=13WjτWj01,α;Ω(α,Σ)+𝒞4(2𝒞3+2)τ(w10,α;[0,L])2\displaystyle\leq\mathcal{C}_{4}\left(\delta_{2}+4\mathcal{C}_{1}\sigma_{cd}\right)\sum_{j=1}^{3}\|W_{j}^{\tau}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\mathcal{C}_{4}(2\mathcal{C}_{3}+2)\tau(\|w_{1}\|_{0,\alpha;[0,L]})^{2}
𝒞4((4𝒞1+1)σcd)j=13WjτWj01,α;Ω(α,Σ)+𝒞4(2𝒞3+2)τ(w10,α;[0,L])2,\displaystyle\leq\mathcal{C}_{4}\left((4\mathcal{C}_{1}+1)\sigma_{cd}\right)\sum_{j=1}^{3}\|W_{j}^{\tau}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\mathcal{C}_{4}(2\mathcal{C}_{3}+2)\tau(\|w_{1}\|_{0,\alpha;[0,L]})^{2},

where 𝒞4>0\mathcal{C}_{4}>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha). Setting

σ5=min{σ4,18𝒞4(1+4𝒞1)}.\sigma_{5}=\min\left\{\sigma_{4},\frac{1}{8\mathcal{C}_{4}(1+4\mathcal{C}_{1})}\right\}. (5.19)

Then for σcdσ5\sigma_{cd}\leq\sigma_{5}, one has

j=13WjτWj01,α;Ω(α,Σ)4𝒞4(2𝒞3+2)τ(w10,α;[0,L])2.\sum_{j=1}^{3}\|W_{j}^{\tau}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq 4\mathcal{C}_{4}(2\mathcal{C}_{3}+2)\tau\left(\|w_{1}\|_{0,\alpha;[0,L]}\right)^{2}. (5.20)

By the definition of 𝒬(𝝋0,w)\mathcal{Q}(\bm{\varphi}_{0},w) and Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w), one gets

𝒬(𝝋0,w+τw1)𝒬(𝝋0,w)Dw𝒬(𝝋0,w)(τw1)0,α;[0,L]\displaystyle\left\|\mathcal{Q}(\bm{\varphi}_{0},w+\tau w_{1})-\mathcal{Q}(\bm{\varphi}_{0},w)-D_{w}\mathcal{Q}(\bm{\varphi}_{0},w)(\tau w_{1})\right\|_{0,\alpha;[0,L]} (5.21)
=N(W~1,W~2,W~3,A~0,B~0)(y1,12)N(W1,W2,W3,A~0,B~0)(y1,12)\displaystyle=\left\|N(\tilde{W}_{1},\tilde{W}_{2},\tilde{W}_{3},\tilde{A}_{0},\tilde{B}_{0})(y_{1},\frac{1}{2})-N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})(y_{1},\frac{1}{2})\right.
τ(j=13WjN(W1,W2,W3,A~0,B~0)Wj0)(y1,12)0,α;[0,L]\displaystyle\quad\quad-\left.\tau\left(\sum_{j=1}^{3}{\partial}_{W_{j}}N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})W_{j}^{0}\right)(y_{1},\frac{1}{2})\right\|_{0,\alpha;[0,L]}
N(W~1,W~2,W~3,A~0,B~0)(y1,12)N(W1,W2,W3,A~0,B~0)(y1,12)\displaystyle\leq\left\|N(\tilde{W}_{1},\tilde{W}_{2},\tilde{W}_{3},\tilde{A}_{0},\tilde{B}_{0})(y_{1},\frac{1}{2})-N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})(y_{1},\frac{1}{2})\right.
τ(j=13WjN(W1,W2,W3,A~0,B~0)Wjτ)(y1,12)0,α;[0,L]\displaystyle\quad\quad-\left.\tau\left(\sum_{j=1}^{3}{\partial}_{W_{j}}N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})W_{j}^{\tau}\right)(y_{1},\frac{1}{2})\right\|_{0,\alpha;[0,L]}
+τ(j=13WjN(W1,W2,W3,A~0,B~0)(WjτWj0))(y1,12)0,α;[0,L]\displaystyle\quad+\left\|\tau\left(\sum_{j=1}^{3}{\partial}_{W_{j}}N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})(W_{j}^{\tau}-W_{j}^{0})\right)(y_{1},\frac{1}{2})\right\|_{0,\alpha;[0,L]}
𝒞τ2(j=13Wjτ1,α;Ω(α,Σ))2+𝒞τj=13WjτWj01,α;Ω(α,Σ)\displaystyle\leq\mathcal{C}\tau^{2}\left(\sum_{j=1}^{3}\|W_{j}^{\tau}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\right)^{2}+\mathcal{C}\tau\sum_{j=1}^{3}\|W_{j}^{\tau}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}
𝒞τ2(w10,α;[0,L])2,\displaystyle\leq\mathcal{C}\tau^{2}\left(\|w_{1}\|_{0,\alpha;[0,L]}\right)^{2},

where 𝒞>0\mathcal{C}>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha). Thus it holds that

𝒬(𝝋0,w+τw1)𝒬(𝝋0,w)Dw𝒬(𝝋0,w)(τw1)0,α;[0,L]τw10,α;[0,L]0\frac{\left\|\mathcal{Q}(\bm{\varphi}_{0},w+\tau w_{1})-\mathcal{Q}(\bm{\varphi}_{0},w)-D_{w}\mathcal{Q}(\bm{\varphi}_{0},w)(\tau w_{1})\right\|_{0,\alpha;[0,L]}}{\tau\|w_{1}\|_{0,\alpha;[0,L]}}\rightarrow 0

as τ0\tau\rightarrow 0. Hence Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w) is the Fréchet derivative of the functional 𝒬(𝝋0,w)\mathcal{Q}(\bm{\varphi}_{0},w) with respect to ww.

It remains to prove the continuity of the map 𝒬(𝝋0,w)\mathcal{Q}(\bm{\varphi}_{0},w) and Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w). For any fixed (𝝋0,w)𝒱0×𝒱1(δ2)(\bm{\varphi}_{0},w)\in\mathcal{V}_{0}\times\mathcal{V}_{1}(\delta_{2}), we assume that (𝝋0k,wk)(𝝋0,w)(\bm{\varphi}_{0}^{k},w^{k})\rightarrow(\bm{\varphi}_{0},w) in 𝒱0×𝒱1(δ2)\mathcal{V}_{0}\times\mathcal{V}_{1}(\delta_{2}) as kk\rightarrow\infty. Then we first show that as kk\rightarrow\infty,

𝒬(𝝋0k,wk)𝒬(𝝋0,w),in𝒱.\mathcal{Q}(\bm{\varphi}_{0}^{k},w^{k})\rightarrow\mathcal{Q}(\bm{\varphi}_{0},w),\quad{\rm{in}}\quad\mathcal{V}. (5.22)

By (5.7) and (5.8), 𝐖k{\bf W}^{k} satisfies the following problem:

{W3k=0(𝐖k,Λ^k,A^k,B^k),(1(Mb)2)y1W1k+y2W2k+1y2W2k=1(𝐖k,𝐖k,A^k,B^k),y1W2ky2W1k=2(𝐖k,𝐖k,A^k,B^k),W1k(0,y2)=3(𝐖k,J^k,A^k,B^k)(0,y2),W2k(L,y2)=0,W2k(y1,12)=(ub+W1k(y1,12))wk(y1),W2k(y1,0)=0,\begin{cases}\begin{aligned} &W_{3}^{k}=\mathcal{F}_{0}({{\bf W}}^{k},\hat{\Lambda}^{k},\hat{A}^{k},\hat{B}^{k}),\\ &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}^{k}+{\partial}_{y_{2}}W_{2}^{k}+\frac{1}{y_{2}}W_{2}^{k}=\mathcal{F}_{1}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k}),\\ &{\partial}_{y_{1}}W_{2}^{k}-{\partial}_{y_{2}}W_{1}^{k}=\mathcal{F}_{2}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k}),\\ &W_{1}^{k}(0,y_{2})=\mathcal{F}_{3}({{\bf W}}^{k},\hat{J}^{k},\hat{A}^{k},\hat{B}^{k})(0,y_{2}),\\ &W_{2}^{k}(L,y_{2})=0,\\ &W_{2}^{k}(y_{1},\frac{1}{2})=\left(u_{b}^{-}+W_{1}^{k}(y_{1},\frac{1}{2})\right)w^{k}(y_{1}),\\ &W_{2}^{k}(y_{1},0)=0,\\ \end{aligned}\end{cases} (5.23)

where

(J^k,Λ^k,A^k,B^k)=(J~0k,ν~0k,A~0k,B~0k)(Jb,0,Ab,Bb).(\hat{J}^{k},\hat{\Lambda}^{k},\hat{A}^{k},\hat{B}^{k})=(\tilde{J}_{0}^{k},\tilde{\nu}_{0}^{k},\tilde{A}_{0}^{k},\tilde{B}_{0}^{k})-(J_{b}^{-},0,A_{b}^{-},B_{b}^{-}).

Taking the difference of (5.8) and (5.23), one obtains that

{W3kW30=0(𝐖k,Λ^k,A^k,B^k)0(𝐖,Λ^,A^,B^),(1(Mb)2)y1(W1kW1)+y2(W2kW2)+1y2(W2kW2)=1(𝐖k,𝐖k,A^k,B^k)1(𝐖,𝐖,A^,B^),y1(W2kW2)y2(W1kW1)=2(𝐖k,𝐖k,A^k,B^k)2(𝐖,𝐖,A^,B^),(W1kW1)(0,y2)=(3(𝐖k,J^k,A^k,B^k)3(𝐖,J^,A^,B^))(0,y2),(W2kW2)(L,y2)=0,(W2kW2)(y1,12)=ub(wkw)(y1)+W1k(y1,12)wk(y1)W1(y1,12)w(y1),(W2kW2)(y1,0)=0.\begin{cases}\begin{aligned} &W_{3}^{k}-W_{3}^{0}=\mathcal{F}_{0}({{\bf W}}^{k},\hat{\Lambda}^{k},\hat{A}^{k},\hat{B}^{k})-\mathcal{F}_{0}({{\bf W}},\hat{\Lambda},\hat{A},\hat{B}),\\ &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}(W_{1}^{k}-W_{1})+{\partial}_{y_{2}}(W_{2}^{k}-W_{2})+\frac{1}{y_{2}}(W_{2}^{k}-W_{2})\\ &=\mathcal{F}_{1}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k})-\mathcal{F}_{1}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B}),\\ &{\partial}_{y_{1}}(W_{2}^{k}-W_{2})-{\partial}_{y_{2}}(W_{1}^{k}-W_{1})\\ &=\mathcal{F}_{2}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k})-\mathcal{F}_{2}({{\bf W}},\nabla{{\bf W}},\hat{A},\hat{B}),\\ &(W_{1}^{k}-W_{1})(0,y_{2})=\left(\mathcal{F}_{3}({{\bf W}}^{k},\hat{J}^{k},\hat{A}^{k},\hat{B}^{k})-\mathcal{F}_{3}({{\bf W}},\hat{J},\hat{A},\hat{B})\right)(0,y_{2}),\\ &(W_{2}^{k}-W_{2})(L,y_{2})=0,\\ &(W_{2}^{k}-W_{2})(y_{1},\frac{1}{2})=u_{b}^{-}(w^{k}-w)(y_{1})+W_{1}^{k}(y_{1},\frac{1}{2})w^{k}(y_{1})-W_{1}(y_{1},\frac{1}{2})w(y_{1}),\\ &(W_{2}^{k}-W_{2})(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.24)

By similar estimate in (5.12), one can infer that

j=13WjkWj1,α;Ω(α,Σ)𝒞(𝝋0k𝝋0𝒱0+wkw0,α;[0,L]).\sum_{j=1}^{3}\|W_{j}^{k}-W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq\mathcal{C}\left(\|\bm{\varphi}_{0}^{k}-\bm{\varphi}_{0}\|_{\mathcal{V}_{0}}+\|w^{k}-{w}\|_{0,\alpha;[0,L]}\right). (5.25)

Therefore,

𝒬(𝝋0k,wk)𝒬(𝝋0,w)0,α;[0,L]\displaystyle\|\mathcal{Q}(\bm{\varphi}_{0}^{k},w^{k})-\mathcal{Q}(\bm{\varphi}_{0},w)\|_{0,\alpha;[0,L]} (5.26)
=N(W1k,W2k,W3k,A~0k,B~0k)N(W1,W2,W3,A~0,B~0)0,α;[0,L]\displaystyle=\left\|N(W_{1}^{k},W_{2}^{k},W_{3}^{k},\tilde{A}_{0}^{k},\tilde{B}_{0}^{k})-N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})\right\|_{0,\alpha;[0,L]}
𝒞(j=13WjkWj1,α;Ω(α,Σ)+𝝋0k𝝋0𝒱0)\displaystyle\leq\mathcal{C}\left(\sum_{j=1}^{3}\|W_{j}^{k}-W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|\bm{\varphi}_{0}^{k}-\bm{\varphi}_{0}\|_{\mathcal{V}_{0}}\right)
𝒞(𝝋0k𝝋0𝒱0+wkw0,α;[0,L]),\displaystyle\leq\mathcal{C}\left(\|\bm{\varphi}_{0}^{k}-\bm{\varphi}_{0}\|_{\mathcal{V}_{0}}+\|w^{k}-{w}\|_{0,\alpha;[0,L]}\right),

which yields (5.22).

Next, we prove the continuity of the map Dw𝒬(𝝋0,w)D_{w}\mathcal{Q}(\bm{\varphi}_{0},w), i.e. to show that as kk\rightarrow\infty,

Dw𝒬(𝝋0k,wk)Dw𝒬(𝝋0,w),in𝒱.D_{w}\mathcal{Q}(\bm{\varphi}_{0}^{k},w^{k})\rightarrow D_{w}\mathcal{Q}(\bm{\varphi}_{0},w),\quad{\rm{in}}\quad\mathcal{V}. (5.27)

It follows from (5.17) that 𝐖0,k{\bf W}^{0,k} satisfies the following problem:

{W30,k=j=13Wjk0(𝐖k,Λ^k,A^k,B^k)Wj0,k,(1(Mb)2)y1W10,k+y2W20,k+1y2W20,k=j=13(Wjk1(𝐖k,𝐖k,A^k,B^k)Wj0,kWjk1(𝐖k,𝐖k,A^k,B^k)Wj0,k),y1W20,ky2W10,k=j=13(Wjk2(𝐖k,𝐖k,A^k,B^k)Wj0,kWjk2(𝐖k,𝐖k,A^k,B^k)Wj0,k),W10,k(0,y2)=(j=12Wjk3(𝐖k,J^k,A^k,B^k)Wj0,k)(0,y2),W20,k(L,y2)=0,W20,k(y1,12)=(ub+W1k(y1,12))w1(y1)+W10,k(y1,12)wk(y1),W20,k(y1,0)=0.\begin{cases}\begin{aligned} &W_{3}^{0,k}=\sum_{j=1}^{3}{\partial}_{W_{j}^{k}}\mathcal{F}_{0}({{\bf W}}^{k},\hat{\Lambda}^{k},\hat{A}^{k},\hat{B}^{k})W_{j}^{0,k},\\ &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}^{0,k}+{\partial}_{y_{2}}W_{2}^{0,k}+\frac{1}{y_{2}}W_{2}^{0,k}\\ &=\sum_{j=1}^{3}\left({\partial}_{W_{j}^{k}}\mathcal{F}_{1}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k})W_{j}^{0,k}-{\partial}_{\nabla W_{j}^{k}}\mathcal{F}_{1}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k})\nabla W_{j}^{0,k}\right),\\ &{\partial}_{y_{1}}W_{2}^{0,k}-{\partial}_{y_{2}}W_{1}^{0,k}\\ &=\sum_{j=1}^{3}\left({\partial}_{W_{j}^{k}}\mathcal{F}_{2}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k})W_{j}^{0,k}-{\partial}_{\nabla W_{j}^{k}}\mathcal{F}_{2}({{\bf W}}^{k},\nabla{{\bf W}}^{k},\hat{A}^{k},\hat{B}^{k})\nabla W_{j}^{0,k}\right),\\ &W_{1}^{0,k}(0,y_{2})=\left(\sum_{j=1}^{2}{\partial}_{W_{j}^{k}}\mathcal{F}_{3}({{\bf W}}^{k},\hat{J}^{k},\hat{A}^{k},\hat{B}^{k})W_{j}^{0,k}\right)(0,y_{2}),\\ &W_{2}^{0,k}(L,y_{2})=0,\\ &W_{2}^{0,k}(y_{1},\frac{1}{2})=\left(u_{b}^{-}+W_{1}^{k}(y_{1},\frac{1}{2})\right)w_{1}(y_{1})+W_{1}^{0,k}(y_{1},\frac{1}{2})w^{k}(y_{1}),\\ &W_{2}^{0,k}(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.28)

Taking the difference of (5.17) and (5.28), one has

j=13Wj0,kWj01,α;Ω(α,Σ)𝒞(j=13WjkWj1,α;Ω(α,Σ)+𝝋0k𝝋0𝒱0+wkw0,α;[0,L]).\displaystyle\sum_{j=1}^{3}\|W_{j}^{0,k}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq\mathcal{C}\left(\sum_{j=1}^{3}\|W_{j}^{k}-W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|\bm{\varphi}_{0}^{k}-\bm{\varphi}_{0}\|_{\mathcal{V}_{0}}+\|w^{k}-{w}\|_{0,\alpha;[0,L]}\right).

Combining (5.25) yields that

j=13Wj0,kWj01,α;Ω(α,Σ)𝒞(𝝋0k𝝋0𝒱0+wkw0,α;[0,L]).\displaystyle\sum_{j=1}^{3}\|W_{j}^{0,k}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq\mathcal{C}\left(\|\bm{\varphi}_{0}^{k}-\bm{\varphi}_{0}\|_{\mathcal{V}_{0}}+\|w^{k}-{w}\|_{0,\alpha;[0,L]}\right). (5.29)

Then it holds that

Dw𝒬(𝝋0k,wk)(w1)Dw𝒬(𝝋0,w)(w1)0,α;[0,L]\displaystyle\|D_{w}\mathcal{Q}(\bm{\varphi}_{0}^{k},w^{k})(w_{1})-D_{w}\mathcal{Q}(\bm{\varphi}_{0},w)(w_{1})\|_{0,\alpha;[0,L]} (5.30)
=j=13(WjkN(W1k,W2k,W3k,A~0k,B~0k)Wj0,kWjN(W1,W2,W3,A~0,B~0)Wj0)0,α;[0,L]\displaystyle=\left\|\sum_{j=1}^{3}\left({\partial}_{W_{j}^{k}}N(W_{1}^{k},W_{2}^{k},W_{3}^{k},\tilde{A}_{0}^{k},\tilde{B}_{0}^{k})W_{j}^{0,k}-{\partial}_{W_{j}}N(W_{1},W_{2},W_{3},\tilde{A}_{0},\tilde{B}_{0})W_{j}^{0}\right)\right\|_{0,\alpha;[0,L]}
𝒞(j=13Wj0,kWj01,α;Ω(α,Σ)+j=13WjkWj1,α;Ω(α,Σ)+𝝋0k𝝋0𝒱0)\displaystyle\leq\mathcal{C}\left(\sum_{j=1}^{3}\|W_{j}^{0,k}-W_{j}^{0}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\sum_{j=1}^{3}\|W_{j}^{k}-W_{j}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|\bm{\varphi}_{0}^{k}-\bm{\varphi}_{0}\|_{\mathcal{V}_{0}}\right)
𝒞(𝝋0k𝝋0𝒱0+wkw0,α;[0,L]),\displaystyle\leq\mathcal{C}\left(\|\bm{\varphi}_{0}^{k}-\bm{\varphi}_{0}\|_{\mathcal{V}_{0}}+\|w^{k}-{w}\|_{0,\alpha;[0,L]}\right),

which implies that (5.27) holds.

In particular, at the background state,

Dw𝒬(𝝋b,0)(w1)=JbW1b(y1,12)=2W1b(y1,12),\displaystyle D_{w}\mathcal{Q}(\bm{\varphi}_{b},0)(w_{1})=-J_{b}^{-}W_{1}^{b}(y_{1},\frac{1}{2})=-2W_{1}^{b}(y_{1},\frac{1}{2}), (5.31)

where 𝐖b{\bf W}^{b} satisfies

{(1(Mb)2)y1W1b+y2W2b+1y2W2b=0,y1W2by2W1b=0,W3b=0,W1b(0,y2)=0,W2b(L,y2)=0,W2b(y1,12)=ubw1(y1),W2b(y1,0)=0.\begin{cases}\begin{aligned} &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}^{b}+{\partial}_{y_{2}}W_{2}^{b}+\frac{1}{y_{2}}W_{2}^{b}=0,\\ &{\partial}_{y_{1}}W_{2}^{b}-{\partial}_{y_{2}}W_{1}^{b}=0,\\ &W_{3}^{b}=0,\\ &W_{1}^{b}(0,y_{2})=0,\\ &W_{2}^{b}(L,y_{2})=0,\\ &W_{2}^{b}(y_{1},\frac{1}{2})=u_{b}^{-}w_{1}(y_{1}),\\ &W_{2}^{b}(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.32)

Define

𝒱b={v𝒱:v(0)=0}.\mathcal{V}_{b}=\{v\in\mathcal{V}:v(0)=0\}. (5.33)

Then Dw𝒬(𝝋b,0)D_{w}\mathcal{Q}(\bm{\varphi}_{b},0) is a continuous mapping from 𝒱1\mathcal{V}_{1} to 𝒱b𝒱\mathcal{V}_{b}\subset\mathcal{V}.

Step 2. The isomorphism of Dw𝒬(φb,0)D_{w}\mathcal{Q}(\bm{\varphi}_{b},0).

To prove the isomorphism of Dw𝒬(𝝋b,0)D_{w}\mathcal{Q}(\bm{\varphi}_{b},0), we need to show that for any given function P𝒱bP^{\ast}\in\mathcal{V}_{b}, there exists a unique w𝒱1w^{\ast}\in\mathcal{V}_{1} such that Dw𝒬(𝝋b,0)(w)=PD_{w}\mathcal{Q}(\bm{\varphi}_{b},0)(w^{\ast})=P^{\ast}, i.e.,

P(y1)=2W1(y1,12).\displaystyle P^{\ast}(y_{1})=-2W_{1}^{\ast}(y_{1},\frac{1}{2}). (5.34)

It follows from (5.32) that the solution (W1,W2,0)(W_{1}^{\ast},W_{2}^{\ast},0) satisfies

{(1(Mb)2)y1W1+y2W2+1y2W2=0,y1W2y2W1=0,W1(0,y2)=0,W2(L,y2)=0,W1(y1,12)=P(y1)2,W2(y1,0)=0.\begin{cases}\begin{aligned} &(1-(M_{b}^{-})^{2}){\partial}_{y_{1}}W_{1}^{\ast}+{\partial}_{y_{2}}W_{2}^{\ast}+\frac{1}{y_{2}}W_{2}^{\ast}=0,\\ &{\partial}_{y_{1}}W_{2}^{\ast}-{\partial}_{y_{2}}W_{1}^{\ast}=0,\\ &W_{1}^{\ast}(0,y_{2})=0,\\ &W_{2}^{\ast}(L,y_{2})=0,\\ &W_{1}^{\ast}(y_{1},\frac{1}{2})=-\frac{P^{\ast}(y_{1})}{2},\\ &W_{2}^{\ast}(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.35)

The second equation in (5.35) implies that there exists a potential function ϕ\phi^{\ast} such that

(y1ϕ,y2ϕ)=(W1,W2),ϕ(L,0)=0.({\partial}_{y_{1}}\phi^{\ast},{\partial}_{y_{2}}\phi^{\ast})=(W_{1}^{\ast},W_{2}^{\ast}),\quad\phi^{\ast}(L,0)=0. (5.36)

Then (5.35) can be written as

{y1((1(Mb)2)y2y1ϕ)+y2(y2y2ϕ)=0,y1ϕ(0,y2)=0,ϕ(L,y2)=0,ϕ(y1,12)=Ly1P(s)2ds,y2ϕ(y1,0)=0.\begin{cases}\begin{aligned} &{\partial}_{y_{1}}((1-(M_{b}^{-})^{2})y_{2}{\partial}_{y_{1}}\phi^{\ast})+{\partial}_{y_{2}}(y_{2}{\partial}_{y_{2}}\phi^{\ast})=0,\\ &{\partial}_{y_{1}}\phi^{\ast}(0,y_{2})=0,\\ &\phi^{\ast}(L,y_{2})=0,\\ &\phi^{\ast}(y_{1},\frac{1}{2})=\int_{L}^{y_{1}}-\frac{P^{\ast}(s)}{2}{\mathrm{d}}s,\\ &{\partial}_{y_{2}}\phi^{\ast}(y_{1},0)=0.\\ \end{aligned}\end{cases} (5.37)

To deal with the singularity near y2=0y_{2}=0, we rewrite the problem (5.37) by using the cylindrical coordinate transformation again. Define

λ1=y1,λ2=y2cosτ,λ3=y2sinτ,τ[0,2π],\lambda_{1}=y_{1},\ \lambda_{2}=y_{2}\cos\tau,\ \lambda_{3}=y_{2}\sin\tau,\ \tau\in[0,2\pi],

and

D1={(λ1,λ2,λ3):0<λ1<L,λ22+λ3212},D2={(λ2,λ3):λ22+λ3212},\displaystyle D_{1}=\{(\lambda_{1},\lambda_{2},\lambda_{3}):0<\lambda_{1}<L,\lambda_{2}^{2}+\lambda_{3}^{2}\leq\frac{1}{2}\},\quad D_{2}=\{(\lambda_{2},\lambda_{3}):\lambda_{2}^{2}+\lambda_{3}^{2}\leq\frac{1}{2}\},
Γw,λ=[0,L]×{(λ2,λ3):λ22+λ32=12},\displaystyle\Gamma_{w,\lambda}=[0,L]\times\{(\lambda_{2},\lambda_{3}):\lambda_{2}^{2}+\lambda_{3}^{2}=\frac{1}{2}\},
Γ0,λ={0}×{(λ2,λ3):λ22+λ3212},ΓL,λ={L}×{(λ2,λ3):λ22+λ3212},\displaystyle\Gamma_{0,\lambda}=\{0\}\times\{(\lambda_{2},\lambda_{3}):\lambda_{2}^{2}+\lambda_{3}^{2}\leq\frac{1}{2}\},\quad\Gamma_{L,\lambda}=\{L\}\times\{(\lambda_{2},\lambda_{3}):\lambda_{2}^{2}+\lambda_{3}^{2}\leq\frac{1}{2}\},
Ψ(𝝀)=ϕ(λ1,λ22+λ32)=Ψ(λ1,|λ|).\displaystyle\Psi^{\ast}(\bm{\lambda})=\phi^{\ast}(\lambda_{1},\sqrt{\lambda_{2}^{2}+\lambda_{3}^{2}})=\Psi^{\ast}(\lambda_{1},|\lambda^{\prime}|).

Then Ψ\Psi^{\ast} solves the following problem

{(1(Mb)2)λ12Ψ+λ22Ψ+λ32Ψ=0,λ1Ψ(0,|λ|)=0,onΓ0,λ,Ψ(L,|λ|)=0,onΓL,λ,Ψ(λ1,|λ|)=Lλ1P(s)2ds,onΓw,λ.\begin{cases}\begin{aligned} &(1-(M_{b}^{-})^{2}){\partial}_{\lambda_{1}}^{2}\Psi^{\ast}+{\partial}_{\lambda_{2}}^{2}\Psi^{\ast}+{\partial}_{\lambda_{3}}^{2}\Psi^{\ast}=0,\\ &{\partial}_{\lambda_{1}}\Psi^{\ast}(0,|\lambda^{\prime}|)=0,\quad&{\rm{on}}\quad\Gamma_{0,\lambda},\\ &\Psi^{\ast}(L,|\lambda^{\prime}|)=0,\quad&{\rm{on}}\quad\Gamma_{L,\lambda},\\ &\Psi^{\ast}(\lambda_{1},|\lambda^{\prime}|)=\int_{L}^{\lambda_{1}}-\frac{P^{\ast}(s)}{2}{\mathrm{d}}s,\quad&{\rm{on}}\quad\Gamma_{w,\lambda}.\\ \end{aligned}\end{cases} (5.38)

By similar arguments as in the Step 3 of Lemma 4.1, (5.38) has a unique solution ΨC2,α(1α,Γw,λ)(D1)\Psi^{\ast}\in C_{2,\alpha}^{(-1-\alpha,\Gamma_{w,\lambda})}(D_{1}) satisfying

Ψ2,α;D1(1α,Γw,λ)𝒞P0,α;[0,L].\|\Psi^{\ast}\|_{2,\alpha;D_{1}}^{(-1-\alpha,\Gamma_{w,\lambda})}\leq\mathcal{C}\|P^{\ast}\|_{0,\alpha;[0,L]}. (5.39)

Thus

i=12Wi1,α;Ω(α,Σ)𝒞P0,α;[0,L].\sum_{i=1}^{2}\|W_{i}^{\ast}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}\leq\mathcal{C}\|P^{\ast}\|_{0,\alpha;[0,L]}. (5.40)

Set w(y1)=ubW2(y1,12)w^{\ast}(y_{1})=u_{b}^{-}W_{2}^{\ast}(y_{1},\frac{1}{2}), then (5.40) shows that w𝒱1w^{\ast}\in\mathcal{V}_{1}. Hence we have shown there exists a unique w𝒱1w^{\ast}\in\mathcal{V}_{1} such that Dw𝒬(𝝋b,0)(w)=PD_{w}\mathcal{Q}(\bm{\varphi}_{b},0)(w^{\ast})=P^{\ast}. The proof of the isomorphism of Dw𝒬(𝝋b,0)D_{w}\mathcal{Q}(\bm{\varphi}_{b},0) is completed.

6 Proof of Theorem 2.3

Now, by the implicit function theorem, there exist positive constants σ6\sigma_{6} and 𝒞\mathcal{C} depending only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha) such that for δ3σ6\delta_{3}\leq\sigma_{6}, the equation 𝒬(𝝋0,w)=0\mathcal{Q}(\bm{\varphi}_{0},w)=0 has a unique solution w=w(𝝋0)w=w(\bm{\varphi}_{0}) satisfying

w0,α;[0,L]𝒞𝝋0𝝋b𝒱0=𝒞σcd.\|w\|_{0,\alpha;[0,L]}\leq\mathcal{C}\|\bm{\varphi}_{0}-\bm{\varphi}_{b}\|_{\mathcal{V}_{0}}=\mathcal{C}\sigma_{cd}. (6.1)

Here δ3\delta_{3} is defined in (5.4). Hence the contact discontinuity gcd(y1)=0y1w(s)ds+12g_{cd}(y_{1})=\int_{0}^{y_{1}}w(s){\mathrm{d}}s+\frac{1}{2} is determined and (6.1) implies that

gcd121,α;[0,L]𝒞5𝝋0𝝋b𝒱0=𝒞5σcd,\|g_{cd}-\frac{1}{2}\|_{1,\alpha;[0,L]}\leq\mathcal{C}_{5}\|\bm{\varphi}_{0}-\bm{\varphi}_{b}\|_{\mathcal{V}_{0}}=\mathcal{C}_{5}\sigma_{cd}, (6.2)

where 𝒞5>0\mathcal{C}_{5}>0 depends only on (𝑼b,L,α)(\bm{U}_{b}^{-},L,\alpha) .

We choose σ1\sigma_{1} and 𝒞\mathcal{C}^{\ast} as

σ1=min{σ5,σ6}and𝒞=4𝒞1+𝒞5,\sigma_{1}=\min\{\sigma_{5},\sigma_{6}\}\quad{\rm{and}}\quad\mathcal{C}^{\ast}=4\mathcal{C}_{1}+\mathcal{C}_{5}, (6.3)

where σ5\sigma_{5} is defined in (5.19) and 𝒞1\mathcal{C}_{1} is defined in (4.49). Then if σcdσ1\sigma_{cd}\leq\sigma_{1}, 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 3.1\mathbf{Problem\ 3.1} has a unique smooth subsonic solution (ux,ur,uθ;gcd)(u_{x},u_{r},u_{\theta};g_{cd}) satisfying

uxub1,α;Ω(α,Σ)+ur1,α;Ω(α,Σ)+uθ1,α;Ω(α,Σ)+gcd121,α;[0,L]𝒞σcd.\|u_{x}-u_{b}^{-}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|u_{r}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|u_{\theta}\|_{1,\alpha;\Omega}^{(-\alpha,\Sigma)}+\|g_{cd}-\frac{1}{2}\|_{1,\alpha;[0,L]}\leq\mathcal{C}^{\ast}\sigma_{cd}. (6.4)

Furthermore, one has

(A,B)(Ab,Bb)1,α;Ω¯=(A~0,B~0)(Ab,Bb)1,α;Ω¯𝒞σcd.\|(A,B)-(A_{b}^{-},B_{b}^{-})\|_{1,\alpha;\overline{\Omega}}=\|(\tilde{A}_{0},\tilde{B}_{0})-(A_{b}^{-},B_{b}^{-})\|_{1,\alpha;\overline{\Omega}}\leq\mathcal{C}^{\ast}\sigma_{cd}. (6.5)

Since the modified Lagrangian transformation (3.3) is invertible, thus the solution transformed back in (x,r)(x,r)-coordinates solves 𝐏𝐫𝐨𝐛𝐥𝐞𝐦 2.2\mathbf{Problem\ 2.2} and the estimates (6.4) and (6.5) imply that the estimates (2.18) and (2.19) in Theorem 2.3 hold. Thus the proof of Theorem 2.3 is completed.

Acknowledgement. Weng is partially supported by National Natural Science Foundation of China 11971307, 12071359, 12221001.

References

  • [1] Bae, M., Weng, S.: 3-D axisymmetric subsonic flows with nonzero swirl for the compressible Euler-Poisson system, Ann. Inst. H. Poincare Anal. Non Lin eaire 35 (2018), no. 1, 161-186.
  • [2] Bae, M.: Stability of contact discontinuity for steady Euler system in the infinite duct. Z. Angew Math. Phys. 64, 917-936 (2013).
  • [3] Bae, M., Park, H.: Contact discontinuity for 2-D inviscid compressible flows in infinitely long nozzles. SIAM J. Math. Anal. 51, 1730-1760 (2019).
  • [4] Bae, M., Park, H.: Contact discontinuity for 3-D axisymmetric inviscid compressible flows in infinitely long cylinders. J. Differential Equations 267, 2824-2873 (2019).
  • [5] Chen, J; Xin, Z; Zang, A.: Subsonic flows past a profile with a vortex line at the trailing edge. SIAM J. Math. Anal. 54 (2022), no. 1, 912-939.
  • [6] Chen, G.-Q., Kukreja, V., Yuan, H.: Well-posedness of transonic characteristic discontinuities in two-dimensional steady compressible Euler flows. Z. Angew. Math. Phys. 64 (2013), no. 6, 1711-1727.
  • [7] Chen, G.-Q., Kukreja, V., Yuan, H.: Stability of transonic characteristic discontinuities in two-dimensional steady compressible Euler flows. J. Math. Phys. 54 (2013), no. 2, 021506, 24 pp.
  • [8] Chen, G.-Q., Huang, F., Wang, T., Xiang, W.: Steady Euler flows with large vorticity and characteristic discontinuities in arbitrary infinitely long nozzles. Adv. Math. 346 (2019), 946-1008.
  • [9] Gilbarg D., Tudinger N. S.: Elliptic Partial Differential Equations of Second Order, 2nd ed. Grundlehren Math. Wiss. 224, Springer, Berlin, 1998.
  • [10] Han, Q., Lin, F.: Elliptic Partial Differential Equations, vol. 1, American Mathematical Soc., 2011.
  • [11] Huang, F., Kuang, J., Wang, D., Xiang,W.: Stability of supersonic contact discontinuity for 2-D steady compressible Euler flows in a finitely long nozzle. J. Differential Equations 266, 4337-4376 (2019).
  • [12] Huang, F., Kuang, J.,Wang, D., Xiang,W.: Stability of transonic contact discontinuity for two-dimensional steady compressible Euler flows in a finitely long nozzle. Ann. PDE 7 (2021), no. 2, Paper No. 23, 96 pp.
  • [13] Lieberman, G. M.: Oblique Derivative Problems for Elliptic Equations, World Scientific, Hackensack, NJ, 2013.
  • [14] Wang, Y., Yuan, H.: Weak stability of transonic contact discontinuities in three dimensional steady non-isentropic compressible Euler flows. Z. Angew. Math. Phys. 66, 341-388 (2015).
  • [15] Wang, Y., Yu, F.: Stability of contact discontinuities in three dimensional compressible steady flows. J. Differential Equations 255, 1278-1356 (2013).
  • [16] Wang, Y., Yu, F.: Structural stability of supersonic contact discontinuities in three-dimensional compressible steady flows. SIAM J. Math. Anal. 47 (2015), no. 2, 1291-1329.
  • [17] Weng, S., Xin, Z.: A deformation-curl decomposition for three dimensional steady Euler equations (in Chinese), Sci Sin Math 49, 307-320, (2019) doi: 10.1360/N012018-00125.
  • [18] Weng, S.: A deformation-curl-Poisson decomposition to the three dimensional steady Euler-Poisson system with applications, J. Differential Equations 267, 6574-6603 (2019)
  • [19] Weng, S., Xie, C., Xin, Z.: Structural stability of the transonic shock problem in a divergent three-dimensional axisymmetric perturbed nozzle, SIAM J. Math. Anal. 53, 279-308 (2021)
  • [20] Weng, S., Zhang, Z.: Subsonic flows with a contact discontinuity in a two-dimensional finitely long curved nozzle. arXiv:2303.15096.
  • [21] Weng, S., Zhang, Z.: Supersonic flows with a contact discontinuity to the two-dimensional steady rotating Euler system. arXiv:2307.00199.