1 Introduction

In this paper, we are focusing on two-dimensional ideal compressible magnetohydrodynamics (MHD)(see [15, 18, 19])

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t\rho +\textrm{div}(\rho {\textbf{u}})=0,\\ \partial _t(\rho {\textbf{u}})+\textrm{div}(\rho {\textbf{u}}\otimes {\textbf{u}}-{\textbf{H}}\otimes {\textbf{H}})+\nabla (p+\frac{1}{2}|{\textbf{H}}|^2)=0,\\ \partial _t{\textbf{H}}-\nabla \times ({\textbf{u}}\times {\textbf{H}})=0,\\ \partial _t(\rho e+\frac{1}{2}|{\textbf{H}}|^2)+ \textrm{div}((\rho e+p){\textbf{u}}+{\textbf{H}}\times ({\textbf{u}}\times {\textbf{H}}))=0,\\ \end{array}\right. } \end{aligned}$$
(1.1)

supplemented with

$$\begin{aligned} \textrm{div}{\textbf{H}}=0 \end{aligned}$$
(1.2)

on the initial data for the Cauchy problem. Here density \(\rho ,\) velocity \({\textbf{u}}=(u_1,u_2)\), magnetic field \({\textbf{H}}=(H_1,H_2)\) and pressure p are unknown functions of time t and spacial variables \({\textbf{x}}=(x_1,x_2)\). \(p=p(\rho ,S)\) stands for the pressure, S is the entropy, \(e=E+\frac{1}{2}|{\textbf{u}}|^2\), where \(E=E(\rho ,S)\) stands for the internal energy. By using the state equation of gas, \(\rho =\rho (p,S),\) and the first principle of thermodynamics, we have that (1.1) is a closed system. We write \(\mathbf{{U}}=\mathbf{{U}}(t,{\textbf{x}})=(p,{\textbf{u}},{\textbf{H}},S)^T,\) with initial data \(\mathbf{{U}}(0,{\textbf{x}})=\mathbf{{U}}_0({\textbf{x}}).\) By (1.2), we can rewrite (1.1) into the form

$$\begin{aligned} {\left\{ \begin{array}{ll} (\partial _t+{\textbf{u}}\cdot \nabla )p+\rho c^2\textrm{div}{\textbf{u}}=0,\\ \rho (\partial _t+{\textbf{u}}\cdot \nabla ){\textbf{u}}-({\textbf{H}}\cdot \nabla ){\textbf{H}}+\nabla q=0,\\ (\partial _t+{\textbf{u}}\cdot \nabla ) {\textbf{H}}-({\textbf{H}}\cdot \nabla ){\textbf{u}}+{\textbf{H}} \textrm{div}{\textbf{u}}=0,\\ (\partial _t+{\textbf{u}}\cdot \nabla )S=0,\\ \end{array}\right. } \end{aligned}$$
(1.3)

where \(q=p+\frac{1}{2}|{\textbf{H}}|^2\) represents the total pressure, c denotes the speed of sound and \(\frac{1}{c^2}:=\frac{\partial \rho }{\partial p}(\rho ,S)=\rho _p(p,S)\). (1.3) can be written in the matrix form as

$$\begin{aligned} A_0(\mathbf{{U}})\partial _t\mathbf{{U}}+A_1(\mathbf{{U}})\partial _1\mathbf{{U}}+A_2(\mathbf{{U}})\partial _2\mathbf{{U}}=0, \end{aligned}$$
(1.4)

where

$$\begin{aligned}{} & {} A_0(\mathbf{{U}}):=\textrm{diag}\Big \{\frac{1}{\rho c^2},\rho ,\rho ,1,1,1\Big \}, \nonumber \\{} & {} \begin{aligned}&A_1(\mathbf{{U}}):=\begin{bmatrix} \frac{u_1}{\rho c^2} &{} 1 &{} 0 &{} 0 &{} 0 &{} 0 \\ 1 &{} \rho u_1 &{} 0 &{} 0 &{} H_2 &{} 0 \\ 0 &{} 0 &{} \rho u_1 &{} 0 &{} -H_1 &{} 0 \\ 0 &{} 0 &{} 0 &{} u_1 &{} 0 &{} 0 \\ 0 &{} H_2 &{} -H_1 &{} 0 &{} u_1 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} u_1 \\ \end{bmatrix},\\&A_2(\mathbf{{U}}):=\begin{bmatrix} \frac{u_2}{\rho c^2} &{} 0 &{} 1 &{} 0 &{} 0 &{} 0 \\ 0 &{} \rho u_2 &{} 0 &{} -H_2 &{} 0 &{} 0 \\ 1 &{} 0 &{} \rho u_2 &{} H_1 &{} 0 &{} 0 \\ 0 &{} -H_2 &{} H_1 &{} u_2 &{} 0 &{} 0 \\ 0 &{} 0&{}0 &{} 0 &{} u_2 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} u_2 \\ \end{bmatrix}.\\ \end{aligned} \end{aligned}$$
(1.5)

The quasilinear system (1.4) is symmetric hyperbolic if the state equation \(\rho =\rho (p,S)\) satisfies the hyperbolicity condition \(A_0>0:\)

$$\begin{aligned} \rho (p,S)>0,\quad \rho _p(\rho ,S)>0. \end{aligned}$$
(1.6)

We suppose that \(\Gamma :=\bigcup \limits _{t\in [0,T]}\Gamma (t)\), where \(\Gamma (t):=\{\textbf{x}\in {\mathbb {R}}^2\,:\,\,x_1-\varphi (t,x_2)=0\}\), is a smooth hypersurface in \([0,T]\times {{\mathbb {R}}}^2.\) The weak solutions of (1.1) satisfy the following Rankine–Hugoniot jump conditions on \(\Gamma (t)\):

$$\begin{aligned} {\left\{ \begin{array}{ll} &{}[j]=0,\quad [H_N]=0,\\ &{} j[u_N]+\vert N\vert ^2[q]=0,\quad j[u_{\tau }]=H_N[H_{\tau }],\\ &{}H_N[u_{\tau }]=j[\frac{H_{\tau }}{\rho }],\\ &{}j[e+\frac{1}{2}\frac{|{\textbf{H}}|^2}{\rho }]+[qu_N-H_N({\textbf{H}}\cdot {\textbf{u}})]=0. \end{array}\right. } \end{aligned}$$
(1.7)

Here we denote \([\upsilon ]=(\upsilon ^+-\upsilon ^-)|_{\Gamma }\) the jump of \(\upsilon \), with \(\upsilon ^{\pm }:=\upsilon \) in \(\Omega ^{\pm }(t)=\{\pm (x_1-\varphi (t,x_2))>0\},\) and \(j=\rho (u_N-\partial _t\varphi )\) is the mass transfer flux across the discontinuity surface. We also denote the tangential and normal components of velocity and magnetic fields \(u_{\tau }={\textbf{u}}\cdot \tau ,H_{\tau }={\textbf{H}}\cdot \tau ,\) \(u_N={\textbf{u}}\cdot N, H_N={\textbf{H}}\cdot N,\) where \(N=(1,-\partial _2\varphi ),\tau =(\partial _2\varphi ,1).\)

We are focusing on the current-vortex sheets solutions, which obey the following additional conditions along the interface \(\Gamma \):

$$\begin{aligned} j^{\pm }|_{\Gamma }=0,\quad H^{\pm }_N|_{\Gamma }=0. \end{aligned}$$

Then, the Rankine–Hugoniot conditions reduce to the boundary conditions

$$\begin{aligned} \partial _t\varphi =u^{\pm }_{N},\quad H^{\pm }_{N}=0,\quad [q]=0 \end{aligned}$$
(1.8)

on \(\Gamma (t).\)

The tangential components of the velocity and the magnetic field may undergo any jumps: \([u_{\tau }]\ne 0,\quad [H_{\tau }]\ne 0.\) The initial data are given as follows:

$$\begin{aligned} {\textbf{U}}^{\pm }(0,{\textbf{x}})={\textbf{U}}^{\pm }_0({\textbf{x}}),\quad {\textbf{x}}\in \Omega ^{\pm }(0),\quad \varphi (0,x_2)=\varphi _0(x_2),\quad x_2\in {{\mathbb {R}}}. \end{aligned}$$
(1.9)

It is obvious that there exist trivial vortex sheets (contact discontinuity) solutions consisting of two constant states separated by a flat surface as

$$\begin{aligned} \bar{{\textbf{U}}}({\textbf{x}}):= {\left\{ \begin{array}{ll} \bar{{\textbf{U}}}^+:=({\bar{p}}^+,0,{\bar{u}}^+_2,0,{\bar{H}}^+_2,{\bar{S}}^+), &{} \text { if } x_1>0,\\ \bar{{\textbf{U}}}^-:=({\bar{p}}^-,0,{\bar{u}}^-_2,0,{\bar{H}}^-_2,{\bar{S}}^-), &{} \text { if } x_1<0,\\ \end{array}\right. } \end{aligned}$$
(1.10)

where on account of (1.8), we require that

$$\begin{aligned} {}{} & {} {} {\bar{p}}^+\ne {\bar{p}}^-,\quad {\bar{u}}^+_2\ne {\bar{u}}^-_2,\quad {\bar{H}}^+_2\ne {\bar{H}}^-_2,\quad {\bar{S}}^+_2\ne {\bar{S}}^-_2,\\{} & {} {} {\bar{p}}^++\frac{1}{2}|{\bar{H}}^+_2|^2={\bar{p}}^-+\frac{1}{2}|{\bar{H}}^-_2|^2. \end{aligned}$$

Compressible vortex sheets are fundamental waves in the study of entropy solutions to multidimensional hyperbolic conservation laws, arising in many important physical phenomena. For the two-dimensional compressible flows governed by the Euler equations, Fejer and Miles [16, 22, 23], see also [12, 35], proved that vortex sheets are unstable when \(M<\sqrt{2}\), while the vortex sheets in three-dimensional Euler flows are always violently unstable (the violent instability is the analogue of the Kelvin-Helmholtz instability for incompressible fluids).

In their pioneering works [13, 14] Coulombel and Secchi proved the nonlinear stability and existence of two-dimensional vortex sheets when the Mach number \(M>\sqrt{2}\). The linear and nonlinear stability of vortex sheets has also been established in [25, 26] for two-dimensional nonisentropic Euler flows, in [4] for two-dimensional relativistic fluids, in [41, 42] for three-dimensional steady Euler flows.

For the three-dimensional compressible flows, various stabilizing effects on vortex sheets have been considered. Taking into account the effect of magnetic fields, Chen–Wang [5, 6] and Trakhinin [38, 39] proved that large non-parallel magnetic fields stabilize the motion of three dimensional current-vortex sheets. Elasticity can also provide stabilization of vortex sheets: Chen–Hu–Wang [7, 8] and Chen–Hu–Wang–Wang–Yuan [9] successfully proved the linear stability and nonlinear stability, respectively, of two-dimensional compressible vortex sheets in elastodynamics by introducing the upper-triangularization method. More recently, Chen–Huang–Wang–Yuan [10] confirmed the stabilizing effect of elasticity also on three dimensional compressible vortex sheets. Another stabilizing effect on vortex sheets is provided by surface tension: for the three-dimensional compressible Euler flows, the local existence and structural stability were proved in Stevens [36].

The analysis of three-dimensional current-vortex sheets in [5, 6, 38, 39] does not cover the two dimensional case. In fact, such a case can be considered as the case when the third component of magnetic field \({\textbf{H}}=(H_1,H_2,H_3)\) is zero, i.e. \(H_3=0,\) and therefore the non-collinear stability conditions in [5, 38, 39] fail. As was shown in Trakhinin [38], the case when either tangential magnetic fields are collinear or one of them is zero corresponds to the transition to violent instability.

For the two-dimensional compressible flows, Wang and Yu [40] proved the linear stability of rectilinear current-vortex sheets under suitable stability conditions by the spectral analysis technique, through the computation of the roots of the Lopatinski determinant and construction of a Kreiss symmetrizer.

In the present paper we investigate the nonlinear stability and existence of two dimensional current-vortex sheets. From the mathematical point of view, this is a nonlinear hyperbolic free boundary problem. Since the Kreiss–Lopatinski condition does not hold uniformly, there is a loss of tangential derivatives in the estimates of the solution. The free boundary is characteristic, which yields a possible loss of regularity in the normal direction to the boundary, and the loss of control of the traces of the characteristic part of the solution.

Differently from [40], we consider the nonisentropic flows and for the analysis of linear stability, instead of spectral analysis, we use a direct energy estimate argument, adapting the dissipative symmetrizer technique introduced by Chen–Wang [5, 6] and Trakhinin [38, 39] for the three-dimensional problem. Moreover, in our linear stability result we study a general case of 2D current-vortex sheets, while in [40] both states of the background magnetic field are assumed to have the same strength, see the subsequent Remark 5.2.

First, we introduce a secondary symmetrization of the system of equations by multiplying by a suitable “secondary generalized Friedrichs symmetrizer” and impose the hyperbolicity of the new system of equations. Then, we identify a sufficient stability condition that makes the boundary conditions dissipative for the new symmetrized system. The new stability condition on the boundary takes the form

$$\begin{aligned} \frac{c_A^+}{\sqrt{1+\left( \frac{c^{+}_A}{c^{+}}\right) ^2}} +\frac{c_A^-}{\sqrt{1+\left( \frac{c^{-}_A}{c^{-}}\right) ^2}}-|[ {\textbf{u}}\cdot \tau ]|>0, \end{aligned}$$
(1.11)

where \(c^{\pm }_A={\vert {\textbf{H}}^{\pm }\vert }/{\sqrt{\rho ^{\pm }}}\) stands for the Alfvén speed. This condition indicates that larger magnetic fields than the jump of tangential velocity play a stabilization effect; in some sense this corresponds to the “subsonic” bubble in linear stability result of [40], see Remark 5.2. Condition (1.11) is in agreement with the stability result found for the three-dimensional current-vortex sheets of [5, 38, 39] that only holds in the subsonic regime.

We observe that condition (1.11) has a strong similarity with the Syrovatskij stability condition [3, 21, 37] that is necessary and sufficient for the linear stability of the two-dimensional incompressible current-vortex sheets. For our problem we show that condition (1.11) is sufficient for the linear stability and optimal with respect to the specific dissipative symmetrizer technique that we use in our proof. On the other hand, it is likely that (1.11) is not necessary for the linear and nonlinear stability. Indeed, by taking the incompressible limit as \(c^\pm \rightarrow +\infty \) in (1.11), we formally get the inequality

$$\begin{aligned} \vert {{\textbf{H}}}^+\vert /\sqrt{\bar{\rho }}+\vert {{\textbf{H}}}^-\vert /\sqrt{\bar{\rho }}-\vert [{{\textbf{u}}}\cdot \tau ]\vert >0, \end{aligned}$$

where, because the incompressible flow has uniform density, \(\rho ^\pm \) have been replaced by a constant \(\bar{\rho }>0\). This inequality describes somehow the “half” of the whole 2D neutral stability domain from [21]. Moreover, in Wang-Yu [40] even a “supersonic” linear stability domain is found for the studied case of particular piece-wise constant background states; the same kind of “supersonic” region is therefore expected to appear for a general case of 2D current-vortex sheet; see again Remark 5.2.

Under condition (1.11) for the background state, we show that the linearized problem obeys an energy estimate in anisotropic weighted Sobolev spaces (see [11, 28, 30, 31]) with a loss of derivatives. The energy estimate for the linearized problem takes the form of a \({ tame}\) estimate, since it exhibits a fixed loss of derivatives from the basic state to the solutions. In order to compensate the loss of derivatives, for the proof of the existence of the solution to the nonlinear problem, we apply a modified Nash–Moser iteration scheme. For an introduction to the Nash–Moser technique, refer to [2, 33].

Compared to [5, 39] for the three-dimensional problem, our existence result shows a lower loss of regularity from the initial data to the solution. This is mainly due to the use of finer Moser-type and imbedding estimates in anisotropic Sobolev spaces.

The rest of the paper is organized as follows: In Section 2 we formulate the nonlinear problem for the current-vortex sheets in a fixed domain. In Section 3 we introduce the definition of anisotropic Sobolev spaces and then state our main Theorem 3.1. In Section 4 we linearize the nonlinear problem with respect to the basic state. Then, we introduce dissipative Friedrichs symmetrizer for two dimensional MHD equations. In Section 5 we study the well-posedness of the linearized problems (4.20), (4.31) and determine the stability condition. In Sects. 6 and 7 we prove the tame estimate in anisotropic Sobolev spaces for the linearized problems (4.31) and (4.20), respectively. In Section 8 we formulate the compatibility conditions for the initial data and construct the approximate solution. In Section 9 we introduce the Nash–Moser iteration scheme and in Section 10 we prove the existence of the solution to the nonlinear problem. In this section one important step is the new construction of the modified state, notably the modified magnetic field, and the delicate derivation of its estimates. In Appendix A1 we recall the trace theorem in anisotropic Sobolev space \(H^m_{*}\) and in Appendix A2 we give a detailed proof of the well-posedness of the homogeneous linearized problem (4.31) stated in Theorem 5.1.

2 Reformulate Current-Vortex Sheets Problem in a Fixed Domain

Let us reformulate the current-vortex sheets problem into an equivalent one posed in a fixed domain. Motivated by Métivier [20], we introduce the functions

$$\begin{aligned} \Phi ^{\pm }(t,{\textbf{x}}):=\Phi (t,\pm x_1,x_2)=\pm x_1+\Psi ^{\pm }(t,{\textbf{x}}),\quad \Psi ^{\pm }(t,{\textbf{x}}):=\chi (\pm x_1)\varphi (t,x_2), \end{aligned}$$

where \(\chi \in C^{\infty }_0({{\mathbb {R}}}),\quad \chi \equiv 1 \text { on } [-1,1],\) and \(||\chi '||_{L^{\infty }({{\mathbb {R}}})}\le \frac{1}{2}\). Here, as in [20], we use the cut-off function \(\chi \) to avoid assumptions about compact support of the initial data in our subsequent nonlinear existence Theorem 3.1. The unknowns \({{\textbf{U}}}^{\pm }\) are smooth in \(\Omega ^{\pm }(t)\) and can be replaced by

$$\begin{aligned} {{\textbf{U}}}^{\pm }_{\natural }(t,{\textbf{x}}):={{\textbf{U}}}^{\pm }(t,\Phi (t,\pm x_1,x_2),x_2), \end{aligned}$$

after changing the variables, which are smooth in the fixed domain \(\Omega ={{\mathbb {R}}}^2_{+}=\{x_1>0, x_2\in {{\mathbb {R}}}\}\). Dropping \(\natural \) in \({{\textbf{U}}}^{\pm }_{\natural }\) for convenience, we reduce (1.1), (1.8) and (1.9) into the initial boundary value problem (IBVP)

$$\begin{aligned} {\left\{ \begin{array}{ll} {\mathbb {L}}({{\textbf{U}}}^{\pm },\Psi ^{\pm })=0 &{} \text {in } [0,T]\times {{\mathbb {R}}}^2_+,\\ {\mathbb {B}}({{\textbf{U}}}^{+},{{\textbf{U}}}^{-},\varphi )=0 &{} \text {on }[0,T]\times \Gamma ,\\ {{\textbf{U}}}^{\pm }(0,{\textbf{x}})={{\textbf{U}}}^{\pm }_0 &{} \text { in } {{\mathbb {R}}}^2_+, \\ \varphi (0,x_2)=\varphi _0 &{} \text { in }{{\mathbb {R}}}, \end{array}\right. } \end{aligned}$$
(2.1)

where \(\Gamma :=\{x_1=0\}\times {{\mathbb {R}}},\) \({\mathbb {L}}({{\textbf{U}}}^\pm ,\Psi ^\pm )=L({{\textbf{U}}}^\pm ,\Psi ^\pm ){{\textbf{U}}}^\pm \),

$$\begin{aligned} L({{\textbf{U}}}^\pm ,\Psi ^\pm )=A_0({{\textbf{U}}}^\pm )\partial _t+{\tilde{A}}_1({{\textbf{U}}}^\pm ,\Psi ^\pm )\partial _1+A_2({{\textbf{U}}}^\pm )\partial _2, \end{aligned}$$
(2.2)

with

$$\begin{aligned} {\tilde{A}}_1({{\textbf{U}}}^{\pm },\Psi ^{\pm })=\frac{1}{\partial _1\Phi ^{\pm }}\Big (A_1({{\textbf{U}}}^{\pm })-A_0({{\textbf{U}}}^{\pm })\partial _t\Psi ^{\pm }-A_2({{\textbf{U}}}^{\pm })\partial _2\Psi ^{\pm }\Big ), \end{aligned}$$
(2.3)

where \(\partial _1\Phi ^{\pm }=\pm 1+\partial _1\Psi ^{\pm }.\) The boundary condition in (2.1) takes the form

$$\begin{aligned} \partial _t\varphi =u^{\pm }_{N},\quad [q]=q^+-q^-=0 \text { on } [0,T]\times \{x_1=0\}\times {{\mathbb {R}}}, \end{aligned}$$

where \(u^{\pm }_{N}=u^{\pm }_1-u^{\pm }_2\partial _2\varphi \).

The following Lemma 2.1 yields that the divergence constraints (1.2) and the boundary conditions \(H^{\pm }_N|_{x_1=0}=0\) on \(\Gamma \) (that is not included in (2.1)) can be regarded as the conditions on the initial data. The proof follows by similar calculations as to those in [39, Proposition 1].

Lemma 2.1

Let the initial data in (2.1) satisfy

$$\begin{aligned} \textrm{div}{\textbf{h}}^{\pm }=0,\quad \text{ in }\,\,\,{\mathbb {R}}^2_+ \end{aligned}$$
(2.4)

and the boundary conditions

$$\begin{aligned} H^{\pm }_{N}|_{x_1=0}=0,\quad \text{ on }\,\,\,\{x_1=0\}\times {\mathbb {R}}, \end{aligned}$$
(2.5)

where

$$\begin{aligned} {\textbf{h}}^{\pm }=(H^{\pm }_n, H^{\pm }_2\partial _1\Phi ^{\pm }),\quad H^{\pm }_n=H^{\pm }_1-H^{\pm }_2\partial _2\Psi ^{\pm },\quad H^{\pm }_N|_{x_1=0}=H^{\pm }_n|_{x_1=0}. \end{aligned}$$

If the IBVP (2.1) admits a solution \(({{\textbf{U}}}^{\pm },\varphi ),\) then this solution satisfies (2.4) and (2.5) for all \(t\in [0,T].\)

3 Properties of Function Spaces and Main Theorem

In this section, we first introduce some notations, then anisotropic Sobolev spaces are defined. At the end, we are ready to state the main result of this paper.

3.1 Notations

Let us denote \(\Omega _T:=(-\infty ,T)\times \Omega \) and \(\Gamma _T:=(-\infty ,T)\times \Gamma \) for \(T>0.\) We write \(\partial _t=\frac{\partial }{\partial t}, \partial _i=\frac{\partial }{\partial x_i},i=1,2, \nabla _{t,{\textbf{x}}}=(\partial _t,\nabla ).\) \(D^{\alpha }_{*}:=\partial ^{\alpha _0}_t(\sigma \partial _1)^{\alpha _1}\partial ^{\alpha _2}_2\partial ^{\alpha _3}_1, \alpha :=(\alpha _0,\alpha _1,\alpha _2,\alpha _3),\quad |\alpha |=\alpha _0+\alpha _1+\alpha _2+\alpha _3.\) Here \(\sigma \) is an increasing smooth function, which satisfies \(\sigma (x_1)=x_1\) for \(0\le x_1\le \frac{1}{2}\) and \(\sigma (x_1)=1\) for \(x_1\ge 1.\) The symbol \(A\lesssim B\) represents that \(A\le CB\) holds uniformly for some universal positive constant C.

3.2 Anisotropic Sobolev spaces

For any integer \(m\in {\mathbb {N}}\) and the interval \(I\subseteq {{\mathbb {R}}},\) function spaces \(H^m_{*}(\Omega )\) and \(H^{m}_{*}(I\times \Omega )\) are defined by

$$\begin{aligned}{} & {} H^m_{*}(\Omega ):=\{u\in L^2(\Omega ): D^{\alpha }_{*}u\in L^2(\Omega ) \text { for } \langle \alpha \rangle :=|\alpha |+\alpha _3\le m, \text { with } \alpha _0=0\}, \\{} & {} H^m_{*}(I\times \Omega ):=\{u\in L^2(I\times \Omega ): D^\alpha _{*}u\in L^2(I\times \Omega ) \text { for }\langle \alpha \rangle :=|\alpha |+\alpha _3\le m\}, \end{aligned}$$

and equipped with the norm \(||\cdot ||_{H^m_{*}(\Omega )}\) and \(||\cdot ||_{H^m_{*}(I\times \Omega )}\) respectively, where

$$\begin{aligned}{} & {} ||u||^2_{H^m_{*}(\Omega )}:=\sum _{\langle \alpha \rangle \le m,\alpha _0=0}||D^{\alpha }_{*}u||^2_{L^2(\Omega )}, \end{aligned}$$
(3.1)
$$\begin{aligned}{} & {} ||u||^2_{H^m_{*}(I\times \Omega )}:=\sum _{\langle \alpha \rangle \le m}||D^{\alpha }_{*}u||^2_{L^2(I\times \Omega )}. \end{aligned}$$
(3.2)

Define the norm

$$\begin{aligned} |||u(t)|||^2_{m,*}:=\sum ^m_{j=0}||\partial ^j_tu(t)||^2_{H^{m-j}_{*}(\Omega )}. \end{aligned}$$
(3.3)

We also write \(||\cdot ||_{m,*,t}:=||u||_{H^m_{*}(\Omega _t)}\) for convenience. Then, from (3.2), we have

$$\begin{aligned} ||u||^2_{m,*,t}=\int ^t_{-\infty }|||u(s)|||^2_{m,*}\textrm{d}s. \end{aligned}$$
(3.4)

3.3 Moser-type calculus inequalities

Now, we introduce two lemmata, which will be useful in the proof of tame estimates in \(H^m_{*}(\Omega _T)\) for the problem (4.20) when m is large enough. We first introduce the Moser-type calculus inequalities in \(H^m,\) see [2, Propositions 2.1.2 and 2.2].

Lemma 3.1

Let \(m\in {\mathbb {N}}_+.\) Let \({\mathcal {O}}\) be an open subset of \({{\mathbb {R}}}^n\) with Lipschitz boundary. Assume that \(F\in C^{\infty }\) in a neighbourhood of the origin with \(F(0)=0\) and that \(u,v\in H^m({\mathcal {O}})\cap L^\infty ({\mathcal {O}})\). Then,

$$\begin{aligned}{} & {} ||\partial ^{\alpha }u\,\partial ^{\beta }v||_{L^2}\lesssim ||u||_{H^m}||v||_{L^{\infty }}+||u||_{L^{\infty }}||v||_{H^m}, \\{} & {} ||F(u)||_{H^m}\le C||u||_{H^m}, \end{aligned}$$

for all \(\alpha ,\beta \in {\mathbb {N}}^n\) with \(|\alpha |+|\beta |\le m\) and where C depends only on F and \(\Vert u\Vert _{L^\infty }\).

Next, we introduce the Moser-type calculus inequalities for \(H^m_{*}\).

Let us define the space

$$\begin{aligned} W^{1,\infty }_*(\Omega _T):=\{u\in L^\infty (\Omega _T):\,\,\,D^\alpha _*u\in L^\infty (\Omega _T),\,\,\langle \alpha \rangle \le 1\}, \end{aligned}$$

equipped with the natural norm

$$\begin{aligned} ||u||_{W^{1,\infty }_{*}(\Omega _T)}:=\sum _{\langle \alpha \rangle \le 1}||D^{\alpha }_{*}u||_{L^{\infty }(\Omega _T)}, \end{aligned}$$
(3.5)

where \(D^{\alpha }_{*}\) and \(\langle \alpha \rangle \) are defined in Section 3.1.

Lemma 3.2

([24, 39]) Let \(m\in {\mathbb {N}}_+.\) Assume that F is a \(C^{\infty }-\)function and \(u,v\in H^m_{*}(\Omega _T)\cap W^{1,\infty }(\Omega _T)\). Then, hold that

$$\begin{aligned}{} & {} ||D^{\alpha }_{*}uD^{\beta }_{*}v||_{L^2(\Omega _T)}\lesssim ||u||_{m,*,T}||v||_{W^{1,\infty }_{*}(\Omega _T)}+||v||_{m,*,T}||u||_{W^{1,\infty }_{*}(\Omega _T)}, \end{aligned}$$
(3.6)
$$\begin{aligned}{} & {} ||uv||_{m,*,T}\lesssim ||u||_{m,*,T}||v||_{W^{1,\infty }_{*}(\Omega _T)}+||v||_{m,*,T}||u||_{W^{1,\infty }_{*}(\Omega _T)}, \end{aligned}$$
(3.7)

for any multi-index \(\alpha ,\beta \in {\mathbb {N}}^4\) with \(\langle \alpha \rangle +\langle \beta \rangle \le m.\) Let \(M_*\) be a positive constant such that

$$\begin{aligned} ||u||_{W^{1,\infty }_{*}(\Omega _T)}\le M_*. \end{aligned}$$

Moreover, if we assume that \(F(0)=0\), then holds that

$$\begin{aligned} ||F(u)||_{m,*,T}\le C(M_{*})||u||_{m,*,T}. \end{aligned}$$
(3.8)

For the proof of (3.6) and (3.7), one can check [24]. For (3.8) one can check [39].

3.4 Embedding and trace theorem

Now, we introduce the Sobolev embedding theorem for \(H^m_{*}(\Omega _T)\).

Lemma 3.3

[24] The following inequalities hold:

$$\begin{aligned}{} & {} ||u||_{L^{\infty }(\Omega _T)}\lesssim ||u||_{3,*,T},\quad ||u||_{W^{1,\infty }_{*}(\Omega _T)}\lesssim ||u||_{4,*,T}, \end{aligned}$$
(3.9)
$$\begin{aligned}{} & {} {} ||u||_{W^{1,\infty }(\Omega _T)}\lesssim ||u||_{5,*,T},\quad ||u||_{W^{2,\infty }_{*}(\Omega _T)}\lesssim ||u||_{6,*,T}, \end{aligned}$$
(3.10)

where \(||u||_{W^{1,\infty }_{*}(\Omega _T)}\) is defined by (3.5) and

$$\begin{aligned} ||u||_{W^{2,\infty }_{*}(\Omega _T)}:=\sum _{\langle \alpha \rangle \le 1}||D^{\alpha }_{*}u||_{W^{1,\infty }(\Omega _T)}. \end{aligned}$$

Proof

Using [24, Theorem B.4], we obtain the first inequality in (3.9) in \(\Omega _T\subseteq {{\mathbb {R}}}^3.\) Then, the second one in (3.9) can be obtained by definition. Observing that

$$\begin{aligned} ||u||_{W^{1,\infty }(\Omega _T)}\lesssim \sum _{\langle \alpha \rangle \le 2}||D^{\alpha }_{*}u||_{L^{\infty }(\Omega _T)}, \end{aligned}$$

we can obtain the first inequality in (3.10) from the first inequality in (3.9). Similarly the second one in (3.10) can be obtained by definition. \(\square \)

For higher order energy estimate, we also need to use the following trace theorem by Ohno, Shizuta, Yanagisawa [27] for the anisotropic Sobolev spaces \(H^m_{*}\)(see also Appendix A1).

Lemma 3.4

[27] Let \(m\ge 1\) be an integer. Then, the following arguments hold:

(i) If \(u\in H^{m+1}_{*}(\Omega _T),\) then its trace \(u|_{x_1=0}\) belongs to \(H^m(\Gamma _T),\) and it satisfies

$$\begin{aligned} ||u|_{x_1=0}||_{H^m(\Gamma _T)}\lesssim ||u||_{m+1,*,T}. \end{aligned}$$

(ii) There exists a continuous lifting operator \({\mathcal {R}}_T:\)

$$\begin{aligned} H^m(\Gamma _T)\rightarrow H^{m+1}_{*}(\Omega _T), \end{aligned}$$

such that \(({\mathcal {R}}_Tu)|_{x_1=0}=u \text { and } ||{\mathcal {R}}_Tu||_{m+1,*,T}\lesssim ||u||_{H^m({\Gamma _T})}.\)

3.5 Main theorem

Theorem 3.1

Let \(m\in {\mathbb {N}}\) and \(m\ge 15,\) and let \(\bar{{\textbf{U}}}^{\pm }\) be defined in (1.10). Suppose that the initial data in (2.1) satisfy

$$\begin{aligned} ({{\textbf{U}}}^{\pm }_0-\bar{{\textbf{U}}}^{\pm },\varphi _0)\in H^{m+11.5}({{\mathbb {R}}}^2_+)\times H^{m+11.5}({{\mathbb {R}}}), \end{aligned}$$

and also satisfy the hyperbolicity condition (1.6), the divergence-free constraint (2.4) for all \({\textbf{x}}\in {{\mathbb {R}}}^2_+.\) Let the initial data at \(x_1=0\) satisfy the stability condition (1.11). The hyperbolicity condition (1.6) and the stability condition (1.11) have to be satisfied uniformly in the sense of (4.2) and (4.4), for suitable \(k>0\). Assume that the initial data are compatible up to order m+10 in the sense of Definition 8.1 and satisfy the boundary constraints (2.5). Assume also that

$$\begin{aligned} \Vert \varphi _0\Vert _{L^\infty ({\mathbb {R}})}< \frac{1}{2}. \end{aligned}$$

Then, there exists a sufficiently short time \(T>0\) such that problem (2.1) has a unique solution on the time interval [0, T] satisfying

$$\begin{aligned} ({{\textbf{U}}}^{\pm }-\bar{{\textbf{U}}}^{\pm },\varphi )\in H^{m}_{*}([0,T]\times {{\mathbb {R}}}^2_+)\times H^m([0,T]\times {{\mathbb {R}}}). \end{aligned}$$

Remark 3.1

Since the initial data \(\textbf{U}_0^\pm \) have the form \(\textbf{U}_0^\pm =\overline{\textbf{U}}^\pm +\tilde{\textbf{U}}_0^\pm \), with \(\tilde{\textbf{U}}_0^\pm \in H^m({\mathbb {R}}^2_+)\) vanishing at infinity (as \(\vert \textbf{x}\vert \rightarrow +\infty \)), the hyperbolicity and stability conditions satisfied in the sense of (4.2), (4.4) (see Remark 4.1) yield that the same conditions hold for the constant states (1.10).

Remark 3.2

We note that Theorem 3.1 implies corresponding results for the original free boundary problem (1.1), (1.8) and (1.9), posed in the moving domain, because Lemma 2.1 and the relation \(\partial _1\Phi ^+\ge \frac{1}{2}\) and \(\partial _1\Phi ^-\le -\frac{1}{2}\) hold in \([0,T]\times {{\mathbb {R}}}^2_+\) for sufficiently small \(T>0.\)

Remark 3.3

Compared to [5, 39], in Theorem 3.1 there is less loss of regularity from the initial data to the solution. This is mainly due to the use of finer Moser-type and imbedding estimates in anisotropic Sobolev spaces.

Remark 3.4

The analysis in this paper could also be applied to prove the nonlinear stability and existence of two dimensional relativistic current-vortex sheets; see [4, 17].

4 Linearized Problem

4.1 The basic state

Let the basic state

$$\begin{aligned} (\hat{{{\textbf{U}}}}^{\pm }(t,{\textbf{x}}),{\hat{\varphi }}(t,x_2)) \end{aligned}$$
(4.1)

be a given vector-valued and sufficiently smooth function, where \(\hat{{{\textbf{U}}}}^{\pm }=({\hat{p}}^{\pm },\hat{{\textbf{u}}}^{\pm }, \hat{{\textbf{H}}}^{\pm },{\hat{S}}^{\pm })^T\) are defined in \(\Omega _T.\) We assume that we shall linearize the problem (2.1) around the basic state (4.1), which satisfy the hyperbolicity condition (1.6) in \(\Omega _T\)

$$\begin{aligned} \rho ({\hat{p}}^{\pm },{\hat{S}}^{\pm })\ge k>0,\quad \rho _{p}({\hat{p}}^{\pm },{\hat{S}}^{\pm })\ge k>0, \end{aligned}$$
(4.2)

the Rankine–Hugoniot jump condition

$$\begin{aligned} \partial _t{\hat{\varphi }}={\hat{u}}^{\pm }_N|_{x_1=0}, \end{aligned}$$
(4.3)

where \({\hat{u}}^{\pm }_{N}={\hat{u}}^{\pm }_1-{\hat{u}}^{\pm }_2\partial _2{\hat{\varphi }}\), and the stability condition on the boundary

$$\begin{aligned} {{\hat{a}}}^+|{{\hat{H}}}^+_2| + {{\hat{a}}}^-|{{\hat{H}}}^-_2| -|[{{\hat{u}}}_2]|\ge k>0, \end{aligned}$$
(4.4)

for a suitable constant k, where \({\hat{c}}^{\pm }_A=\frac{\vert \hat{{\textbf{H}}}^{\pm }\vert }{\sqrt{{\hat{\rho }}^{\pm }}}\) denotes the Alfvén speed and

$$\begin{aligned} {{\hat{a}}}^\pm :=\sqrt{\frac{1}{{\hat{\rho }}^{\pm }(1+\frac{({\hat{c}}^{\pm }_A)^2}{({\hat{c}}^{\pm })^2})}}. \end{aligned}$$
(4.5)

Remark 4.1

The stability condition (1.11) can be written in uniform form as (4.4) by making use of (4.3) and the following boundary constraint in (4.10).

Remark 4.2

Let us observe that, differently from [39], in the Rankine–Hugoniot conditions we dot not require that the basic state satisfies \([{{\hat{q}}}]=0\). It seems that this condition could not be helpful to simplify the boundary quadratic form which appear from the application of the energy method to the linearized problem.

Remark 4.3

Estimate (4.4) implies in particular that at least one among \({{\hat{H}}}^+_2\) and \({{\hat{H}}}^-_2\) must be nonzero along the boundary.

Remark 4.4

The presence of the positive constant k in the hyperbolicity and stability assumptions (4.2), (4.4) is needed in order to ensure the boundedness of all coefficients, nonlinearly depending on the basic state \(\hat{{\textbf{U}}}^\pm \), \({{\hat{\varphi }}}\), appearing in the arguments of the energy method developed in the sequel.

Let us assume

$$\begin{aligned} \begin{aligned}&\hat{{{\textbf{U}}}}^{\pm }\in W^{3,\infty }(\Omega _T),\quad {\hat{\varphi }}\in W^{4,\infty }(\Gamma _T),\\&||\hat{{{\textbf{U}}}}^{\pm }||_{W^{3,\infty }(\Omega _T)}+||{\hat{\varphi }}||_{W^{4,\infty }(\Gamma _T)}\le K, \end{aligned} \end{aligned}$$
(4.6)

where \(K>0\) is a constant. Moreover, without loss of generality, we assume

$$\begin{aligned} ||{\hat{\varphi }}||_{L^{\infty }(\Gamma _T)}<\frac{1}{2}. \end{aligned}$$
(4.7)

This implies

$$\begin{aligned} \partial _1{\hat{\Phi }}^+\ge \frac{1}{2},\quad \partial _1{\hat{\Phi }}^-\le -\frac{1}{2}, \end{aligned}$$

with

$$\begin{aligned}{\hat{\Phi }}^{\pm }(t,{\textbf{x}}):=\pm x_1+{\hat{\Psi }}^{\pm }(t,{\textbf{x}}),\quad {\hat{\Psi }}^{\pm }(t,{\textbf{x}}):=\chi (\pm x_1){\hat{\varphi }}(t,x_2).\end{aligned}$$

We also assume the following nonlinear constraints on the background states:

$$\begin{aligned} \partial _t\hat{{\textbf{H}}}^{\pm }+\frac{1}{\partial _1{\hat{\Phi }}^{\pm }}\Big ((\hat{{\textbf{w}}}^{\pm }\cdot \nabla )\hat{{\textbf{H}}}^{\pm }-(\hat{{\textbf{h}}}\cdot \nabla )\hat{{\textbf{u}}}^{\pm }+\hat{{\textbf{H}}}^{\pm }\textrm{div}{\hat{\mathrm {{\textbf{v}}}}^{\pm }}\Big )=0, \end{aligned}$$
(4.8)

where

$$\begin{aligned} \begin{aligned}&\hat{{\textbf{v}}}^{\pm }=({\hat{u}}^{\pm }_n,{\hat{u}}^{\pm }_2\partial _1{\hat{\Phi }}^{\pm }),\quad {\hat{u}}^{\pm }_n={\hat{u}}^{\pm }_1-{\hat{u}}^{\pm }_2\partial _2{\hat{\Psi }}^{\pm },\\&\hat{{\textbf{h}}}^{\pm }=({\hat{H}}^{\pm }_n,{\hat{H}}^{\pm }_2\partial _1{\hat{\Phi }}^{\pm }),\quad {\hat{H}}^{\pm }_n={\hat{H}}^{\pm }_1-{\hat{H}}^{\pm }_2\partial _2{\hat{\Psi }}^{\pm },\\&\hat{{\textbf{w}}}^{\pm }=\hat{{\textbf{v}}}^{\pm }-(\partial _t{\hat{\Psi }}^{\pm },0). \end{aligned} \end{aligned}$$
(4.9)

Then, we can obtain that the constraints (2.4) and (2.5), that is,

$$\begin{aligned} \textrm{div}\hat{{\textbf{h}}}^{\pm }=0,\quad {\hat{H}}^{\pm }_N|_{x_1=0}=0 \end{aligned}$$
(4.10)

hold for all times \(t>0\) if they hold initially (see Appendix A in [39]), where \({\hat{H}}^{\pm }_N={\hat{H}}^{\pm }_1-{\hat{H}}^{\pm }_2\partial _2{\hat{\varphi }}.\)

4.2 The linearized equations

The linearized equations for (2.1) around the basic state (4.1) can be defined as

$$\begin{aligned}{} & {} \mathbb {L'}(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })(\delta {{\textbf{U}}}^{\pm },\delta \Psi ^{\pm }):=\frac{\textrm{d}}{\textrm{d}\varepsilon }{\mathbb {L}}({{\textbf{U}}}^{\pm }_{\varepsilon },\Psi ^{\pm }_{\varepsilon })|_{\varepsilon =0}={\textbf{f}}^{\pm } \text { in } \Omega _T, \\{} & {} \mathbb {B'}(\hat{{{\textbf{U}}}}^{+}, \hat{{{\textbf{U}}}}^{-}, {\hat{\varphi }})(\delta {{\textbf{U}}}^{+}, \delta {{\textbf{U}}}^{-}, \delta \varphi )=\frac{\textrm{d}}{\textrm{d}\varepsilon }{\mathbb {B}}({{\textbf{U}}}^{+}_{\varepsilon },{{\textbf{U}}}^{-}_{\varepsilon },\varphi _{\varepsilon })|_{\varepsilon =0}={\textbf{g}} \text { on }\Gamma _T, \end{aligned}$$

where \({{\textbf{U}}}^{\pm }_{\varepsilon }=\hat{{{\textbf{U}}}}^{\pm }+\varepsilon \delta {{\textbf{U}}}^{\pm },\quad \varphi _{\varepsilon }={\hat{\varphi }}+\varepsilon \delta \varphi \) and

$$\begin{aligned}{} & {} \Psi ^{\pm }_{\varepsilon }(t,{\textbf{x}}):=\chi (\pm x_1)\varphi _{\varepsilon }(t,x_2),\quad \Phi ^{\pm }_{\varepsilon }(t,{\textbf{x}}):=\pm x_1+\Psi ^{\pm }_{\varepsilon }(t,{\textbf{x}}),\\{} & {} \delta \Psi ^{\pm }(t,{\textbf{x}}):=\chi (\pm x_1)\delta \varphi (t,x_2).\end{aligned}$$

In the following argument, we shall drop \(\delta \) for simplicity. The linearized operators have the following form:

$$\begin{aligned}{} & {} {\mathbb {L}}'(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })({{\textbf{U}}}^{\pm },\Psi ^{\pm }) \nonumber \\{} & {} \quad =L(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm }){{\textbf{U}}}^{\pm }+C(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm }){{\textbf{U}}}^{\pm }-\{L(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\Psi ^{\pm }\}\frac{\partial _1\hat{{{\textbf{U}}}}^{\pm }}{\partial _1{\hat{\Phi }}^{\pm }}, \end{aligned}$$
(4.11)
$$\begin{aligned}{} & {} {\mathbb {B}}'(\hat{{{\textbf{U}}}}^{+},\hat{{{\textbf{U}}}}^{-},{\hat{\varphi }})({{\textbf{U}}}^{+},{{\textbf{U}}}^{-},\varphi )=\left[ \begin{array}{c} \partial _t\varphi +{\hat{u}}^+_2\partial _2\varphi -u^+_N\\ \partial _t\varphi +{\hat{u}}^-_2\partial _2\varphi -u^-_N\\ q^+-q^- \end{array}\right] , \end{aligned}$$
(4.12)

where the operator \(L(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\) is defined in (2.2), \(q^{\pm }=p^{\pm }+\hat{{\textbf{H}}}^{\pm }\cdot {\textbf{H}}^{\pm },u^{\pm }_N=u^{\pm }_1-u^{\pm }_2\partial _2{\hat{\varphi }},\) and the matrix \(C(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\) is defined as follows:

$$\begin{aligned} C(\hat{{{{\textbf {U}}}}}^{\pm },{\hat{\Psi }}^{\pm })Y&=(Y,\nabla _yA_0(\hat{{{{\textbf {U}}}}}^{\pm }))\partial _t\hat{{{{\textbf {U}}}}}^{\pm }+(Y,\nabla _y{\tilde{A}}_1(\hat{{{{\textbf {U}}}}}^{\pm },{\hat{\Psi }}^{\pm }))\partial _1\hat{{{{\textbf {U}}}}}^{\pm }\\&\quad +(Y,\nabla _yA_2(\hat{{{{\textbf {U}}}}}^{\pm }))\partial _2\hat{{{{\textbf {U}}}}}^{\pm },\\ {}&(Y,\nabla _yA(\hat{{{{\textbf {U}}}}}^{\pm }))=\sum _{i=1}^6y_i(\frac{\partial A(Y)}{\partial y_i}\Big |_{Y=\hat{{{{\textbf {U}}}}}^{\pm }}),\quad Y=(y_1,\cdots ,y_6). \end{aligned}$$

We introduce the Alinhac’s good unknown [1]

$$\begin{aligned} \dot{{{\textbf{U}}}}^{\pm }:={{\textbf{U}}}^{\pm }-\frac{\partial _1\hat{{{\textbf{U}}}}^{\pm }}{\partial _1{\hat{\Phi }}^{\pm }}\Psi ^{\pm }. \end{aligned}$$
(4.13)

In terms of (4.13), the linearized interior equations have the following form:

$$\begin{aligned} L(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\dot{{{\textbf{U}}}}^{\pm }+C(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\dot{{{\textbf{U}}}}^{\pm }-\frac{\Psi ^{\pm }}{\partial _1{\hat{\Phi }}^{\pm }}\partial _1\{{\mathbb {L}}(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\}=\mathbf {f^{\pm }}. \end{aligned}$$
(4.14)

Since the zero order terms in \(\Psi ^{\pm }\) can be regarded as the error terms in the Nash–Moser iteration, we drop these terms and consider the effective linear operators

$$\begin{aligned} \begin{aligned} {\mathbb {L}}'_e(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\dot{{{\textbf{U}}}}^{\pm }&:=L(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\dot{{{\textbf{U}}}}^{\pm }+C(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\dot{{{\textbf{U}}}}^{\pm }\\&=A_0(\hat{{{\textbf{U}}}}^{\pm })\partial _t\dot{{{\textbf{U}}}}^{\pm }+{\tilde{A}}_1(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\partial _1\dot{{{\textbf{U}}}}^{\pm }\\&\quad +A_2(\hat{{{\textbf{U}}}}^{\pm })\partial _2\dot{{{\textbf{U}}}}^{\pm }+C(\hat{{{\textbf{U}}}}^{\pm },{\hat{\Psi }}^{\pm })\dot{{{\textbf{U}}}}^{\pm }. \end{aligned} \end{aligned}$$
(4.15)

Concerning the boundary differential operator \({\mathbb {B}}'\), we rewrite (4.12) in terms of the Alinhac’s good unknowns (4.13) to get

$$\begin{aligned} {\mathbb {B}}'_e(\hat{{{\textbf{U}}}},{\hat{\varphi }})(\dot{{{\textbf{U}}}},\varphi )=\left[ \begin{array}{c} \partial _t\varphi +{\hat{u}}^+_2\partial _2\varphi -{\dot{u}}^+_N-\varphi \partial _1{\hat{u}}^+_N\\ \partial _t\varphi +{\hat{u}}^-_2\partial _2\varphi -{\dot{u}}^-_N+\varphi \partial _1{\hat{u}}^-_N\\ {\dot{q}}^+-{\dot{q}}^-+\varphi (\partial _1{\hat{q}}^++\partial _1{\hat{q}}^-) \end{array}\right] , \end{aligned}$$
(4.16)

where \({\dot{u}}^{\pm }_N={\dot{u}}^{\pm }_1-{\dot{u}}^{\pm }_2\partial _2{\hat{\varphi }}\) (recall also that \(\partial _1{{\hat{\Phi }}}^\pm \vert _{x_1=0}=\pm 1\)).

Remark 4.5

Notice that, due to the fact that we have transformed the domains \(\Omega ^\pm (t)\) into the same half-space \({\mathbb {R}}^2_+\), the jump on the boundary of a normal derivative of a function \(a=a(t,\textbf{x})\) is defined as follows:

$$\begin{aligned}{}[\partial _1a]:=\partial _1a^+_{|x_1=0}+\partial _1a^-_{|x_1=0}. \end{aligned}$$
(4.17)

This is why the jump of the total pressure q in the last row of (4.16) reduces under Alinhac’s change of unknowns to

$$\begin{aligned} {\dot{q}}^+-{\dot{q}}^-+\varphi (\partial _1{\hat{q}}^++\partial _1{\hat{q}}^-). \end{aligned}$$

In the following, according to (4.17), we will set

$$\begin{aligned}{}[\partial _1{{\hat{q}}}]:=\partial _1{\hat{q}}^+\vert _{x_1=0}+\partial _1{\hat{q}}^-\vert _{x_1=0}. \end{aligned}$$
(4.18)

Denote the operator

$$\begin{aligned} {\mathbb {L}}'_e(\hat{{{{\textbf {U}}}}},{\hat{\Psi }})\dot{{{{\textbf {U}}}}}:= \left[ \begin{array}{c} {\mathbb {L}}'_e(\hat{{{{\textbf {U}}}}}^+,{\hat{\Psi }}^+)\dot{{{{\textbf {U}}}}}^+\\ {\mathbb {L}}'_e(\hat{{{{\textbf {U}}}}}^-,{\hat{\Psi }}^-)\dot{{{{\textbf {U}}}}}^- \end{array}\right] . \end{aligned}$$
(4.19)

Now, we are focusing on the following linear problem for \((\dot{{{\textbf{U}}}}^{\pm },\varphi ):\)

$$\begin{aligned} {\left\{ \begin{array}{ll} {\mathbb {L}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})\dot{{{\textbf{U}}}}={\textbf{f}} &{} \text {in } \Omega _T,\\ {\mathbb {B}}'_e(\hat{{{\textbf{U}}}},{\hat{\varphi }})(\dot{{{\textbf{U}}}},\varphi )={\textbf{g}} &{} \text {on } \Gamma _T,\\ (\dot{{{\textbf{U}}}},\varphi )=0 &{} \text {for } t<0, \\ \end{array}\right. } \end{aligned}$$
(4.20)

where \({\textbf{f}}=({\textbf{f}}^+,{\textbf{f}}^-)=(f^+_1,\cdots ,f^+_6,f^-_1,\cdots ,f^-_6),\) and \({\textbf{g}}=(g^+_1,g^-_1,g_2)\) vanish in the past. In order to prove the well-posedness of the linearized problem (4.20), we can state the following Lemma 4.1, that can be proven as in [39, Proposition 2].

Lemma 4.1

Let the basic state (4.1) satisfy the assumption (4.8) and (4.10). Then, the solutions of the problem (4.20) satisfy

$$\begin{aligned}{} & {} \textrm{div} \dot{{\textbf{h}}}^{\pm }=r^{\pm } \text { in } \Omega _T, \end{aligned}$$
(4.21)
$$\begin{aligned}{} & {} {\hat{H}}_2^{\pm }\partial _2\varphi -{\dot{H}}^{\pm }_N\mp \varphi \partial _1{\hat{H}}^{\pm }_N=g^{\pm }_3 \text { on } \Gamma _{T}. \end{aligned}$$
(4.22)

Here

$$\begin{aligned} \dot{{\textbf{h}}}^{\pm }=({\dot{H}}^{\pm }_n,{\dot{H}}^{\pm }_2\partial _1{\hat{\Phi }}^{\pm }),\quad {\dot{H}}^{\pm }_n={\dot{H}}^{\pm }_1-{\dot{H}}^{\pm }_2\partial _2{\hat{\Psi }}^{\pm }, \quad {\dot{H}}^{\pm }_N |_{x_1=0}={\dot{H}}^{\pm }_n |_{x_1=0}, \end{aligned}$$

where \(r^{\pm }=r^{\pm }(t,{\textbf{x}}),\quad g^{\pm }_3=g^{\pm }_3(t,x_2),\) which vanish in the past, are determined by the source terms and the basic state as solutions to the linear equations

$$\begin{aligned}{} & {} \partial _tR^{\pm }+\frac{1}{\partial _1{\hat{\Phi }}^{\pm }}\Big \{(\hat{{\textbf{w}}}^{\pm }\cdot \nabla R^{\pm })+R^{\pm }\textrm{div}\hat{{\textbf{v}}}^{\pm }\Big \}={\mathcal {F}}^{\pm } \text { in } \Omega _T, \end{aligned}$$
(4.23)
$$\begin{aligned}{} & {} \partial _tg^{\pm }_3+{\hat{u}}^{\pm }_2\partial _2g^{\pm }_3+\partial _2{\hat{u}}^{\pm }_2g^{\pm }_3={\mathcal {G}}^{\pm } \text { on } \Gamma _T, \end{aligned}$$
(4.24)

where \(R^{\pm }=\frac{r^{\pm }}{\partial _1{\hat{\Phi }}^{\pm }},\quad {\mathcal {F}}^{\pm }=\frac{\textrm{div} {\textbf{f}}^{\pm }_h}{\partial _1{\hat{\Phi }}^{\pm }}, \quad {\textbf{f}}^{\pm }_h=(f^{\pm }_n,f^{\pm }_5),\quad f^{\pm }_n=f^{\pm }_4-f^{\pm }_5\partial _2{\hat{\Psi }}^{\pm }\),

\({\mathcal {G}}^{\pm }=\{\partial _2({\hat{H}}^{\pm }_2g^{\pm }_1)-f^{\pm }_n\}|_{x_1=0}\), with \(\hat{{\textbf{w}}}^\pm \) and \(\hat{{\textbf{v}}}^\pm \) defined in (4.9).

4.3 Reduction to homogeneous boundary conditions

In this section, we reduce the inhomogeneous boundary condition in (4.20) to the homogeneous one. We follow the same ideas in Trakhinin [39], that for reader’s convenience, we recall here. Suppose there exists a solution \((\dot{{{\textbf{U}}}},\varphi )\in H^s_{*}(\Omega _T)\times H^s({\Gamma _T})\) to problem (4.20), with a given \(s\in {\mathbb {N}}.\) We define a vector-valued function

$$\begin{aligned} \tilde{{{\textbf{U}}}}=({\tilde{p}}^+,\tilde{{\textbf{u}}}^+,\tilde{{\textbf{H}}}^+,{\tilde{S}}^+,{\tilde{p}}^-,\tilde{{\textbf{u}}}^-,\tilde{{\textbf{H}}}^-,{\tilde{S}}^-)\in H^{s+2}_{*}(\Omega _T) \end{aligned}$$

that vanishes in the past and such that it is a “suitable lifting" of the boundary data \(({{\textbf{g}}}, g^+_3, g^-_3)\in H^{s+1}(\Gamma _T)\). We choose \(\tilde{{{\textbf{U}}}}\) such that on the boundary \(\Gamma _T\), it satisfies the boundary conditions in (4.20) with \(\varphi =0\), i.e.

$$\begin{aligned} {\mathbb {B}}'_e(\hat{{{\textbf{U}}}},{\hat{\varphi }})(\tilde{{{\textbf{U}}}},0)={\textbf{g}}, \end{aligned}$$

and both conditions (4.22) and (4.24) with \(\varphi =0\) (where \({{\tilde{q}}}\), \({{\tilde{u}}}^\pm _N\), \({{\tilde{H}}}_N^\pm \) are defined similarly to \(\dot{q}\), \(\dot{u}_N^\pm \) and \(\dot{H}_N^\pm \)). More explicitly we require

$$\begin{aligned}{} & {} {\tilde{u}}^{\pm }_N |_{x_1=0}=-g^{\pm }_{1}, \\{} & {} {{\tilde{q}}}^+ |_{x_1=0} - {{\tilde{q}}}^- |_{x_1=0}= g_2, \\{} & {} {\tilde{H}}^{\pm }_N |_{x_1=0}=-g^{\pm }_{3} \end{aligned}$$

with \(g^\pm _3\) solution of (4.24). Then, we define \(\tilde{q}^\pm \), \({{\tilde{u}}}_n^\pm \) and \({{\tilde{H}}}_n^\pm \) in the interior domain \(\Omega _T\) by using the lifting operator (that exists thanks to the trace theorem in anisotropic Sobolev spaces \(H^s_*\), see [27] and Appendix A1)

$$\begin{aligned} {\mathcal {R}}_T: H^{s+1}(\Gamma _T) \rightarrow H^{s+2}_*(\Omega _T) \end{aligned}$$

which gives

$$\begin{aligned} {{\tilde{q}}}^\pm ={\mathcal {R}}_T({{\tilde{q}}}^\pm |_{x_1=0}), \quad \tilde{u}^\pm _n={\mathcal {R}}_T({{\tilde{u}}}^\pm _N |_{x_1=0}), \quad \tilde{H}^\pm _n={\mathcal {R}}_T({{\tilde{H}}}^\pm _N |_{x_1=0}). \end{aligned}$$

Let also \({{\tilde{u}}}_1^\pm \) and \({{\tilde{H}}}^\pm _1\) be such that

$$\begin{aligned} {{\tilde{u}}}_n^\pm = {{\tilde{u}}}_1^\pm - {{\tilde{u}}}_2^\pm \partial _2 {{\hat{\Psi }}}^\pm , \quad {{\tilde{H}}}_n^\pm = {{\tilde{H}}}_1^\pm - {{\tilde{H}}}_2^\pm \partial _2 {{\hat{\Psi }}}^\pm , \end{aligned}$$

where \({{\tilde{u}}}_2^\pm \) is arbitrary and can be taken for instance as zero, and where we can define \({{\tilde{H}}}^\pm _2\) in such a way that it satisfies equation (4.23) for \(R^\pm =\frac{\text {div} \tilde{{\textbf{h}}}^\pm }{\partial _1{{\hat{\Phi }}}^\pm }\), where \(\tilde{{\textbf{h}}}^\pm =({{\tilde{H}}}_n^\pm , \tilde{H}_2^\pm \partial _1{{\hat{\Phi }}}^\pm )\) (this is possible since we have a freedom in the choice of “characteristic unknown" \({{\tilde{H}}}^\pm _2\) ). The last components \({{\tilde{S}}}^\pm \) of \(\tilde{{\textbf{U}}}\) can again be taken as zero. To sum up, the vector \(\tilde{{\textbf{U}}}\) is defined as

$$\begin{aligned} \tilde{{{\textbf{U}}}}=({\tilde{p}}^+,{\tilde{u}}^+_n, 0,{\tilde{H}}^+_n+{\tilde{H}}^+_2\partial _2{\hat{\Psi }}^+,{\tilde{H}}^+_2,0,{\tilde{p}}^-,{\tilde{u}}^-_n, 0,{\tilde{H}}^-_n+{\tilde{H}}^-_2\partial _2{\hat{\Psi }}^-,{\tilde{H}}^-_2,0), \end{aligned}$$

where \({{\tilde{H}}}^\pm _2\) satisfies equation (4.23) for \(R^\pm =\frac{\text {div} \tilde{{\textbf{h}}}^\pm }{\partial _1{{\hat{\Phi }}}^\pm }\) and \(\tilde{{\textbf{h}}}^\pm =(\tilde{H}_n^\pm , {{\tilde{H}}}_2^\pm \partial _1{{\hat{\Phi }}}^\pm )\).

We define \(\dot{{{\textbf{U}}}}^{\natural }=\dot{{{\textbf{U}}}}-\tilde{{{\textbf{U}}}},\) then \(\dot{{{\textbf{U}}}}^{\natural }\) satisfies

$$\begin{aligned} {\left\{ \begin{array}{ll} {\mathbb {L}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})\dot{{{\textbf{U}}}}^{\natural }={\textbf{F}}={\textbf{f}}-{\mathbb {L}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})\tilde{{{\textbf{U}}}} &{} \text { in } \Omega _T,\\ {\mathbb {B}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})(\dot{{{\textbf{U}}}}^{\natural },\varphi )=0 &{} \text { on } \Gamma _T. \end{array}\right. } \end{aligned}$$
(4.25)

Here \({\textbf{F}}=(F^+_1,\cdots ,F^+_6,F^-_1,\cdots ,F^-_6).\) Moreover, in view of equations (4.23) for \(R^{\pm }=\frac{\textrm{div}\tilde{{\textbf{h}}}^{\pm }}{\partial _1{\hat{\Phi }}^{\pm }},\) condition (4.22) for \(\tilde{{\textbf{U}}}\) with \(\varphi =0,\) and (4.24), we have, from (4.25), that (4.23) and (4.24) are satisfied for \(R^{\pm }=\frac{\textrm{div}\dot{{\textbf{h}}}^{\natural \pm }}{\partial _1{\hat{\Phi }}^{\pm }},\) \(g^{\pm }_3=({\hat{H}}^{\pm }_2\partial _2\varphi -{\dot{H}}^{\natural \pm }_N\mp \varphi \partial _1{\hat{H}}^{\pm }_N)|_{x_1=0}\) with right-hand sides \({\mathcal {F}}^{\pm }={\mathcal {G}}^{\pm }=0\). Here \(\dot{{\textbf{h}}}^{\natural \pm }\) and \({\dot{H}}^{\natural \pm }_N\) are defined similarly to \(\dot{{\textbf{h}}}^{\pm }\) and \({\dot{H}}^{\pm }_N.\) Notice that \(\dot{{{\textbf{U}}}}^{\natural }=0, \text { for } t<0.\) Hence, the conditions

$$\begin{aligned} \textrm{div}\dot{{\textbf{h}}}^{\natural \pm }=0,\quad ({\hat{H}}^{\pm }_2\partial _2\varphi -{\dot{H}}^{\natural \pm }_N{\mp }\varphi \partial _1{\hat{H}}^{\pm }_N)|_{x_1=0}=0 \end{aligned}$$
(4.26)

hold for \(t<0\). Then, by standard method of characteristic curves, we get that equations (4.26) are satisfied for all \(t\in (-\infty ,T].\) Notice that from

$$\begin{aligned} ||\partial _1(\cdot )||_{s,*,T}\le ||\cdot ||_{s+2,*,T}, \end{aligned}$$

we have that

$$\begin{aligned} ||{\mathbb {L}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})\tilde{{{\textbf{U}}}}||_{s,*,T}\le C||\tilde{{{\textbf{U}}}}||_{s+2,*,T}. \end{aligned}$$
(4.27)

By the definition of \({{\textbf{F}}}\), using (4.27) and the definition of \(\tilde{{\textbf{U}}}\) as a lifting of the boundary data \(({{\textbf{g}}},g_3^+, g_3^-)\), we obtain that

$$\begin{aligned} ||{\textbf{F}}||_{s,*,T}\le C\Big (||{\textbf{f}}||_{s,*,T}+||{\textbf{g}}||_{H^{s+1}(\Gamma _T)} + ||(g_3^+,g_3^-)||_{H^{s+1}(\Gamma _T)}\Big ). \end{aligned}$$
(4.28)

Using (4.24) and the trace Theorem in the anisotropic Sobolev spaces to estimate

$$\begin{aligned} \Vert f_n\vert _{x_1=0}\Vert _{H^{s+1}(\Gamma _T)}\le C\Vert {\textbf{f}}\Vert _{s+2,*,T}, \end{aligned}$$

we get

$$\begin{aligned} ||(g^{+}_3,g^{-}_3) ||_{H^{s+1}(\Gamma _T)}\le & {} C||({\mathcal {G}}^{+},{\mathcal {G}}^{-}) ||_{H^{s+1}(\Gamma _T)}\nonumber \\\le & {} C\Big (||{\textbf{f}}||_{s+2,*,T}+||(g^{+}_1, g^{-}_1) ||_{H^{s+2}(\Gamma _T)}\Big ). \end{aligned}$$
(4.29)

Then, from (4.28) and (4.29), we derive

$$\begin{aligned} ||{\textbf{F}}||_{s,*,T}\le C\Big (||{\textbf{f}}||_{s+2,*,T}+||{\textbf{g}}||_{H^{s+2}(\Gamma _T)}\Big ). \end{aligned}$$
(4.30)

We obtain that \(\dot{{{\textbf{U}}}}^{\natural }\) solves the problem

$$\begin{aligned} {\left\{ \begin{array}{ll} {\mathbb {L}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})\dot{{{\textbf{U}}}^\natural }={\textbf{F}} &{} \text { in } \Omega _T,\\ {\mathbb {B}}'_e(\hat{{{\textbf{U}}}},{\hat{\varphi }})(\dot{{{\textbf{U}}}^\natural },\varphi )=0 &{} \text { on } \Gamma _T,\\ (\dot{{{\textbf{U}}}^\natural },\varphi )=0 &{} \text { for } t<0, \end{array}\right. } \end{aligned}$$
(4.31)

where the source term \({\textbf{F}}\) satisfies the estimate (4.30) and solutions of (4.31) satisfy the constraints

$$\begin{aligned}{} & {} \textrm{div}\dot{{\textbf{h}}^\natural }^{\pm }=0 \text { in }\Omega _T, \end{aligned}$$
(4.32)
$$\begin{aligned}{} & {} {\hat{H}}^{\pm }_2\partial _2\varphi -{\dot{H}}^{\natural \,\pm }_{N}{\mp }\varphi \partial _1{\hat{H}}^{\pm }_{N}=0 \text { on }\Gamma _T. \end{aligned}$$
(4.33)

We have thus proved the following result:

Lemma 4.2

Let problem (4.31) be well-posed and its unique solution \((\dot{{{\textbf{U}}}^\natural },\varphi )\) belongs to \(H^s_{*}(\Omega _T)\times H^s({\Gamma _T})\) for \({\textbf{F}}\in H^s_{*}(\Omega _T),\) where \(s\in {\mathbb {N}}\) is a given number. Then problem (4.20) is well-posed, namely it admits a unique solution \((\dot{{\textbf{U}}},\varphi ) \in H^s_{*}(\Omega _T)\times H^s({\Gamma _T})\) for data \(({\textbf{f}},{\textbf{g}})\in H^{s+2}_{*}(\Omega _T)\times H^{s+2}(\Gamma _T).\)

Remark 4.6

Let us observe that loss of regularity from the data to the solution in the inhomogeneous problem (4.20) is due to the introduction of lifting function \(\tilde{{{\textbf{U}}}}\).

4.4 New Friedrichs Symmetrizer for 2D MHD equations

Motivated by the idea of Trakhinin [39], in the following we will make use of a new symmetric form of the MHD system. This symmetric form is the result of the application of a “secondary generalized Friedrichs symmetrizer” \({\mathbb {S}}=(S({{\textbf{U}}}), {\mathcal {T}}({{\textbf{U}}}))\) to system (1.4):

$$\begin{aligned} \begin{aligned}&{\mathcal {S}}({{\textbf{U}}})A_0({{\textbf{U}}})\partial _t{{\textbf{U}}}+{\mathcal {S}}({{\textbf{U}}})A_1({{\textbf{U}}})\partial _1{{\textbf{U}}}+{\mathcal {S}}({{\textbf{U}}})A_2({{\textbf{U}}})\partial _2{{\textbf{U}}}+{\mathcal {T}}({{\textbf{U}}})\textrm{div} {\textbf{H}}\\&\quad :=B_0({{\textbf{U}}})\partial _t{{\textbf{U}}}+B_1({{\textbf{U}}})\partial _1{{\textbf{U}}}+B_2({{\textbf{U}}})\partial _2{{\textbf{U}}}=0. \end{aligned} \end{aligned}$$
(4.34)

The Friedrichs symmetrizer can be written as (see [38])

$$\begin{aligned} {\mathcal {S}}({{\textbf{U}}})= & {} \left[ \begin{array}{cccccc} 1 &{} \frac{\lambda H_1}{\rho c^2} &{} \frac{\lambda H_2}{\rho c^2} &{} 0 &{} 0 &{} 0 \\ \lambda H_1\rho &{} 1 &{} 0 &{} -\rho \lambda &{} 0 &{} 0\\ \lambda H_2\rho &{} 0 &{} 1 &{} 0 &{} -\rho \lambda &{} 0\\ 0 &{} -\lambda &{} 0 &{} 1 &{} 0 &{} 0 \\ 0 &{} 0 &{} -\lambda &{} 0 &{} 1 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 1 \end{array}\right] ,\quad {\mathcal {T}}({{\textbf{U}}})=-\lambda \left[ \begin{array}{c} 1 \\ 0\\ 0\\ H_1 \\ H_2\\ 0 \end{array}\right] , \end{aligned}$$
(4.35)
$$\begin{aligned} B_0= & {} {\mathcal {S}}A_0=\left[ \begin{array}{cccccc} \frac{1}{\rho c^2} &{} \frac{\lambda H_1}{ c^2} &{} \frac{\lambda H_2}{ c^2} &{} 0 &{} 0 &{} 0 \\ \frac{\lambda H_1}{c^2} &{} \rho &{} 0 &{} -\rho \lambda &{} 0 &{} 0\\ \frac{\lambda H_2}{c^2} &{} 0 &{} \rho &{} 0 &{} -\rho \lambda &{} 0\\ 0 &{} -\rho \lambda &{} 0 &{} 1 &{} 0 &{} 0 \\ 0 &{} 0 &{} -\rho \lambda &{} 0 &{} 1 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 1 \end{array}\right] , \end{aligned}$$
(4.36)

where \(\lambda =\lambda ({\textbf{U}})\) is an arbitrary function.

In order to make system (4.34) symmetric hyperbolic, we need \(B_0>0,\) i.e.,

$$\begin{aligned} \rho \lambda ^2<\frac{1}{1+\frac{c^2_A}{c^2}}. \end{aligned}$$
(4.37)

Condition (4.37) ensures the equivalence of system (1.4) and (4.34) on smooth solutions if \(\lambda ({{\textbf{U}}})\) is a smooth function of \({{\textbf{U}}}\) (see Trakhinin [38] and [39]).

Now let us apply the new symmetrization to the homogeneous linearized problem (4.31). From now on we drop the index \(\natural \) from the unknown \(\dot{{\textbf{U}}}^\natural \) of system (4.31).

Multiplying (4.31) on the left by \({\mathcal {S}}(\hat{{{\textbf{U}}}})\) and adding to the result the vector

$$\begin{aligned} \left[ \begin{array}{c} \frac{\textrm{div} \dot{{\textbf{h}}}^{+}}{\partial _1{\hat{\Phi }}^{+}}{\mathcal {T}}(\hat{{{\textbf{U}}}}^+) \\ \frac{\textrm{div} \dot{{\textbf{h}}}^{-}}{\partial _1{\hat{\Phi }}^{-}}{\mathcal {T}}(\hat{{{\textbf{U}}}}^-) \\ \end{array}\right] , \end{aligned}$$
(4.38)

we obtain

$$\begin{aligned} B_0(\hat{{{\textbf{U}}}})\partial _t\dot{{{\textbf{U}}}}+{\tilde{B}}_1(\hat{{{\textbf{U}}}},{\hat{\Psi }})\partial _1\dot{{{\textbf{U}}}}+B_2(\hat{{{\textbf{U}}}})\partial _2\dot{{{\textbf{U}}}}+{\tilde{C}}(\hat{{{\textbf{U}}}},{\hat{\Psi }})\dot{{{\textbf{U}}}}=\tilde{{\textbf{F}}}(\hat{{{\textbf{U}}}}), \end{aligned}$$
(4.39)

where \({\tilde{C}}(\hat{{{\textbf{U}}}},{\hat{\Psi }})={\mathcal {S}}(\hat{{{\textbf{U}}}})C(\hat{{{\textbf{U}}}},{\hat{\Psi }}),\quad \tilde{{\textbf{F}}}(\hat{{{\textbf{U}}}})={\mathcal {S}}(\hat{{{\textbf{U}}}}){{\textbf{F}}}\),

$$\begin{aligned}{} & {} S(\hat{{{\textbf{U}}}})=\textrm{diag}({\mathcal {S}}(\hat{{{\textbf{U}}}}^+),S(\hat{{{\textbf{U}}}}^-)),\quad B_{\alpha }(\hat{{{\textbf{U}}}})=\textrm{diag}( B_{\alpha }(\hat{{{\textbf{U}}}}^+), B_{\alpha }(\hat{{{\textbf{U}}}}^-)),\,\,\,\alpha =0,2, \\{} & {} C(\hat{{{\textbf{U}}}},{\hat{\Psi }})=\textrm{diag}(C(\hat{{{\textbf{U}}}}^+,{\hat{\Psi }}^+),C(\hat{{{\textbf{U}}}}^-,{\hat{\Psi }}^-)),\\{} & {} {\tilde{B}}_1(\hat{{{\textbf{U}}}},{\hat{\Psi }})=\textrm{diag}({\tilde{B}}_{1}(\hat{{{\textbf{U}}}}^+,{\hat{\Psi }}^+), {\tilde{B}}_{1}(\hat{{{\textbf{U}}}}^-,{\hat{\Psi }}^-)), \\{} & {} {\tilde{B}}_1({{\textbf{U}}}^{\pm },\Psi ^{\pm })=\frac{1}{\partial _1\Phi ^{\pm }}\Big (B_1({{\textbf{U}}}^{\pm })-B_0({{\textbf{U}}}^{\pm })\partial _t\Psi ^{\pm }-B_2({{\textbf{U}}}^{\pm })\partial _2\Psi ^{\pm }\Big ). \end{aligned}$$

5 Stability condition and well-posedness of the linearized problem

In this section, let us introduce the new unknown \({\textbf{V}}=({\textbf{V}}^+,{\textbf{V}}^-),\) where

$$\begin{aligned} {\textbf{V}}^{\pm }=({\dot{q}}^{\pm },{\dot{u}}^{\pm }_n,{\dot{u}}^{\pm }_2,{\dot{H}}^{\pm }_n,{\dot{H}}^{\pm }_2,{\dot{S}}^{\pm }). \end{aligned}$$

Rewriting system (4.31) in terms of \({\textbf{V}}\) shows in clear way that the boundary matrix of the resulting system has constant rank at the boundary (see (5.6)), i.e. the system is symmetric hyperbolic with characteristic boundary of constant multiplicity (in the sense of Rauch [29]). Indeed, we obtain that

$$\begin{aligned} \dot{{{\textbf{U}}}}=J{{\textbf{V}}}, \end{aligned}$$
(5.1)

where \(J=\textrm{diag}(J^+,J^-),\)

$$\begin{aligned} J^{\pm }=\left[ \begin{array}{cccccc} 1 &{} 0 &{} 0 &{} -{\hat{H}}^{\pm }_1 &{} -{\hat{H}}^{\pm }_{\tau } &{} 0 \\ 0 &{} 1 &{} \partial _2{\hat{\Psi }}^{\pm } &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 1 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 1 &{} \partial _2{\hat{\Psi }}^{\pm } &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 1 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 1 \end{array}\right] , \end{aligned}$$
(5.2)

with \({\hat{H}}^{\pm }_{\tau }={\hat{H}}\cdot \tau ^{\pm },\quad \tau ^{\pm }=(\partial _2{\hat{\Psi }}^{\pm },1).\) In terms of \({\textbf{V}}\), systems (4.31) and (4.39) can be equivalently rewritten as

$$\begin{aligned}{} & {} {\mathcal {A}}_0(\hat{{{\textbf{U}}}},{\hat{\Psi }})\partial _t{{\textbf{V}}}+{\mathcal {A}}_1(\hat{{{\textbf{U}}}},{\hat{\Psi }})\partial _1{{\textbf{V}}}+{\mathcal {A}}_2(\hat{{{\textbf{U}}}},{\hat{\Psi }})\partial _2{{\textbf{V}}}+{\mathcal {A}}_3(\hat{{{\textbf{U}}}},{\hat{\Psi }}){{\textbf{V}}}={\mathcal {F}}(\hat{{{\textbf{U}}}},{\hat{\Psi }}), \nonumber \\ \end{aligned}$$
(5.3)
$$\begin{aligned}{} & {} {\mathcal {B}}_0(\hat{{{\textbf{U}}}},{\hat{\Psi }})\partial _t{{\textbf{V}}}+{\mathcal {B}}_1(\hat{{{\textbf{U}}}},{\hat{\Psi }})\partial _1{{\textbf{V}}}+{\mathcal {B}}_2(\hat{{{\textbf{U}}}},{\hat{\Psi }})\partial _2{{\textbf{V}}}+{\mathcal {B}}_3(\hat{{{\textbf{U}}}},{\hat{\Psi }}){{\textbf{V}}}=\tilde{{\mathcal {F}}}(\hat{{{\textbf{U}}}},{\hat{\Psi }}), \nonumber \\ \end{aligned}$$
(5.4)

where \({\mathcal {A}}_{\alpha }=J^TA_{\alpha }J,\quad {\mathcal {B}}_{\alpha }=J^TB_{\alpha }J,\quad \alpha =0,2,\)

$$\begin{aligned}{} & {} {\mathcal {A}}_1=J^T{\tilde{A}}_1J,\quad {\mathcal {B}}_1=J^T{\tilde{B}}_1J,\quad {\mathcal {F}}=J^T{\textbf{F}},\quad \tilde{{\mathcal {F}}}=J^T\tilde{{\textbf{F}}}, \\{} & {} {\mathcal {A}}_3=J^T\Big (CJ+A_0\partial _tJ+{\tilde{A}}_1\partial _1J+A_2\partial _2J\Big ), \\{} & {} {\mathcal {B}}_3=J^T\Big (CJ+B_0\partial _tJ+{\tilde{B}}_1\partial _1J+B_2\partial _2J\Big ). \end{aligned}$$

In view of constraints (4.3) and (4.10) on the basic state \(\hat{{\textbf{U}}}\), the boundary matrix in (5.3) has the following form:

$$\begin{aligned} {\mathcal {A}}_1={\mathcal {A}}+{\mathcal {A}}_{(0)},\quad {\mathcal {A}}=\textrm{diag}\Big (\frac{1}{\partial _1{\hat{\Phi }}^+}{\mathcal {E}}_{12},\frac{1}{\partial _1{\hat{\Phi }}^-}{\mathcal {E}}_{12}\Big ),\quad {\mathcal {A}}_{(0)}|_{x_1=0}=0. \end{aligned}$$
(5.5)

Here, \({\mathcal {E}}_{ij}\) is a \(6\times 6\) symmetric matrix, in which (ij)th and (ji)th elements are 1 and the remaining elements are 0. The explicit form of \({\mathcal {A}}_{(0)}\) is of no interest and it is only important that the all non zero elements of \({\mathcal {A}}_{(0)}\) are multiplied either by the function \({\hat{u}}_n -\partial _t {\hat{\Psi }}\) or by the function \({\hat{H}}_n\). Therefore, the boundary matrix

$$\begin{aligned} {\mathcal {A}}_1|_{x_1=0}=\textrm{diag}({\mathcal {E}}_{12},-{\mathcal {E}}_{12}). \end{aligned}$$
(5.6)

This is a matrix of constant rank 4 and has two positive and two negative eigenvalues.

Concerning system (5.4), note that \({\mathcal {B}}_0>0\) because of the hyperbolicity condition (4.37) satisfied for the basic state, hence the symmetric system (5.4) is hyperbolic. Moreover, considering the boundary matrix \({\mathcal {B}}_1\) in (5.4), we will only need its explicit form on the boundary

$$\begin{aligned} {\mathcal {B}}_1|_{x_1=0}=\textrm{diag}\Big ({\mathcal {B}}(\hat{{{\textbf{U}}}}^+|_{x_1=0}),-{\mathcal {B}}(\hat{{{\textbf{U}}}}^-|_{x_1=0})\Big ),\quad {\mathcal {B}}(\hat{{\textbf{U}}}^\pm )={\mathcal {E}}_{12}-\lambda (\hat{{\textbf{U}}}^\pm ){\mathcal {E}}_{14}, \end{aligned}$$

which gives

$$\begin{aligned} ({\mathcal {B}}_1{{\textbf{V}}},{{\textbf{V}}})|_{x_1=0}=2[{\dot{q}}({\dot{u}}_N-{\hat{\lambda }}{\dot{H}}_N)], \end{aligned}$$
(5.7)

with \({{\hat{\lambda }}}:= \lambda (\hat{{\textbf{U}}})\).

Using (4.31) and (4.33) (see also (4.16)) for \({\dot{u}}^{\pm }_{N}\) and \({\dot{H}}^{\pm }_N,\) (recall that we have dropped the index \(\natural \) from the unknown) and the boundary constraint \([{\dot{q}}]=-\varphi [\partial _{1}{\hat{q}}]\) (recall the definition of \([\partial _1{{\hat{q}}}]\) in (4.18)) we obtain that

$$\begin{aligned} ({\mathcal {B}}_1{{\textbf{V}}}\cdot {{\textbf{V}}})=2[{{\hat{u}}}_2-{{\hat{\lambda }}}{{\hat{H}}}_2]\dot{q}^+\partial _2\varphi +\text{ l.o.t },\quad \text{ on }\,\,\,\{x_1=0\}, \end{aligned}$$
(5.8)

with

$$\begin{aligned} \text{ l.o.t. }:= & {} -2[\partial _1{{\hat{u}}}_N-{{\hat{\lambda }}}\partial _1{{\hat{H}}}_N] \dot{q}^+\varphi -2[\partial _1{{\hat{q}}}]\varphi \partial _t\varphi -2[\partial _1{{\hat{q}}}]({{\hat{u}}}^-_2-{{\hat{\lambda }}}{{\hat{H}}}^-_2)\varphi \partial _2\varphi \nonumber \\{} & {} -2[\partial _1{{\hat{q}}}](\partial _1{{\hat{u}}}^-_N-{{\hat{\lambda }}}^-\partial _1{{\hat{H}}}^-_N)\varphi ^2. \end{aligned}$$
(5.9)

We use “l.o.t." to mean boundary terms that can be manipulated in the energy estimate by passing to volume integral and using integration by parts, so that they do not give any trouble in the derivation of the energy estimate, see Appendix A2.

Let us now make a suitable choice of the function \({\hat{\lambda }}=\lambda (\hat{{\textbf{U}}})\). We first take \({\hat{\lambda }}\) as, (see [39])

$$\begin{aligned} {\hat{\lambda }}^\pm :=\lambda (\hat{{{\textbf{U}}}}^{\pm }):=\eta (x_1)\lambda ^{\pm }(t,x_2), \end{aligned}$$
(5.10)

with \(\eta (x_1)\) a smooth monotone decreasing function satisfying \(\eta (0)=1\) and \(\eta (x_1)=0,\) for \(x_1>\varepsilon \), with \(\varepsilon >0\) sufficiently small. The functions \(\lambda ^{\pm }\) will be chosen below.

Remark 5.1

The motivation of the definition of \({\hat{\lambda }}\) by using the cut-off function \(\eta =\eta (x_1)\) is that it guarantees that the hyperbolicity condition (4.37), that trivially holds for \({{\hat{\lambda }}}=0\) remains true for the basic state in a small neighborhood \(\{0<x_1<\varepsilon \}\) of the boundary, thanks to the continuity of the basic state. Hence, by definition of \(\eta \), the hyperbolicity condition still holds in the whole domain \(\Omega _T\); see [39] for more details.

The functions \(\lambda ^{\pm }(t,x_2)\) are chosen in this way: if the jump \([{\hat{u}}_2](t,x_2)=0,\) we define \(\lambda ^{\pm }(t,x_2)=0\); otherwise, if the jump \([{\hat{u}}_2](t,x_2)\ne 0,\) we choose \(\lambda ^{+}(t,x_2)\) and \(\lambda ^{-}(t,x_2)\) satisfying the following relation:

$$\begin{aligned}{}[{\hat{u}}_2-\lambda {\hat{H}}_2]=0. \end{aligned}$$

The following Lemma ensures the existence of such functions \(\lambda ^{+}(t,x_2)\), \(\lambda ^{-}(t,x_2)\) satisfying the above equation and the hyperbolicity assumption (4.37) written for the basic state, see (5.12):

Lemma 5.1

For all \((t,x_2)\in \Gamma _T\), there exist \(\lambda ^{\pm }(t,x_2)\) satisfying

$$\begin{aligned}{}[{\hat{u}}_2-\lambda {\hat{H}}_2]=0, \end{aligned}$$
(5.11)

and the hyperbolicity conditions

$$\begin{aligned} |{\lambda }^{\pm }|< {{\hat{a}}}^\pm \end{aligned}$$
(5.12)

if and only if the basic states \(\hat{{\textbf{U}}}^\pm \) obey the stability estimate

$$\begin{aligned} |[{{\hat{u}}}_2]|<|{{\hat{H}}}^+_2| {{\hat{a}}}^++ |{{\hat{H}}}^-_2|{{\hat{a}}}^-, \end{aligned}$$
(5.13)

where \({{\hat{a}}}^\pm \) are defined in (4.5). In particular, under (5.13) we can set

$$\begin{aligned} \begin{aligned} \lambda ^+= sgn({{\hat{H}}}^+_2) \frac{{{\hat{a}}}^+ [{{\hat{u}}}_2]}{{{\hat{a}}}^+ |{{\hat{H}}}^+_2| + {{\hat{a}}}^- |{{\hat{H}}}^-_2|},\\ \lambda ^-=- sgn({{\hat{H}}}^-_2) \frac{{{\hat{a}}}^- [{{\hat{u}}}_2]}{{{\hat{a}}}^+ |{{\hat{H}}}^+_2| + {{\hat{a}}}^- |{{\hat{H}}}^-_2|}. \end{aligned} \end{aligned}$$
(5.14)

Proof

For shortness in the proof we drop the hat and the subscript 2 on the variables, moreover we do not write the considered variables are restricted along the boundary \(\{x_1=0\}\), that is, we write

$$\begin{aligned} u={{\hat{u}}}^\pm _2\vert _{x_1=0}, \quad H^\pm = {{\hat{H}}}^\pm _2\vert _{x_1=0},\quad \text {etc.}. \end{aligned}$$

Let us first note that equation (5.11) can be restated as

$$\begin{aligned}{}[u]= \lambda ^+ H^+ -\lambda ^- H^-. \end{aligned}$$
(5.15)

Let us consider the simple case \(H^+\ne 0\) and \(H^-=0\). Then, from equation (5.15) we immediately get

$$\begin{aligned} \lambda ^+= \frac{[u]}{H^+}, \end{aligned}$$

and plugging the above into the constraint in (5.12) for \(\lambda ^+\) gives

$$\begin{aligned} |[u]| < a^+ |H^+|, \end{aligned}$$
(5.16)

that is, (5.13). In this case \(\lambda ^-\) can be whatever function satisfying the corresponding constraint (5.12)

$$\begin{aligned} |\lambda ^-|<a^- \end{aligned}$$

(for instance \(\lambda ^-\equiv 0\)); it does not yield any condition on the background state in addition to (5.16). Of course the symmetric case \(H^+=0\) and \(H^-\ne 0\) is treated similarly, by choosing

$$\begin{aligned} \lambda ^- =- \frac{[u]}{H^-}, \quad \lambda ^+\equiv 0 \end{aligned}$$

and making the condition

$$\begin{aligned} |[u]| < a^- |H^-| \end{aligned}$$
(5.17)

to be satisfied on the background state, which is again (5.13).

Let us consider now the case \(H^+\ne 0\) and \(H^-\ne 0\). From (5.15), we write \(\lambda ^+\) as a known function of \(\lambda ^-\) as

$$\begin{aligned} \lambda ^+= \frac{[u] + \lambda ^- H^-}{H^+} \end{aligned}$$
(5.18)

and plug the above into \(|\lambda ^+| < a^+\) to get, after simple manipulations,

$$\begin{aligned} -a^+ - \frac{[u]}{H^+}< \frac{H^-}{H^+}\lambda ^- < a^+ - \frac{[u]}{H^+}. \end{aligned}$$
(5.19)

We have that \(\lambda ^-\) must obey simultaneously the two constraints (5.19) and

$$\begin{aligned} -a^-<\lambda ^- < a^-. \end{aligned}$$
(5.20)

Let assume now that \(\frac{H^-}{H^+}>0\). Then equation (5.19) becomes

$$\begin{aligned} b_2< \lambda ^- < b_1, \end{aligned}$$
(5.21)

where

$$\begin{aligned} b_1= \frac{H^+}{H^-} \left( a^+ - \frac{[u]}{H^+} \right) \quad \text{ and }\quad b_2= \frac{H^+}{H^-} \left( - a^+ - \frac{[u]}{H^+} \right) . \end{aligned}$$
(5.22)

A necessary condition for (5.20) and (5.21) to hold simultaneously is that

$$\begin{aligned} b_1> - a^- \quad \text {and} \quad a^- > b_2. \end{aligned}$$
(5.23)

After some calculations, in view of the definition (5.22), the above becomes equivalent to

$$\begin{aligned} - \left( \frac{H^+}{H^-} a^+ + a^-\right)<\frac{[u]}{H_-}< \frac{H^+}{H^-} a^+ + a^-. \end{aligned}$$

Assuming \(H^- <0\) (hence \(H^+<0\)) the above is equivalent to

$$\begin{aligned} |[u]| < - \left( H^+ a^+ + H^- a^-\right) = |H^+| a^+ + |H^-| a^-, \end{aligned}$$

which is the estimate (5.13). Similar calculation in the case \(H^- >0\) and \(H^+ >0\) still yield to (5.13).

To sum up if \(\frac{H^-}{H^+} >0\) we get that condition (5.13) is at least necessary in order to find \(\lambda ^-\) satisfying (5.20) and (5.21), that is,

$$\begin{aligned} \max \{ -a^-, b_2\}< \lambda ^- < \min \{a^-, b_1 \}. \end{aligned}$$

If such \(\lambda ^-\) actually exists, then \(\lambda ^+\) will be defined by (5.18).

From (5.19), arguing in the same way as above when \(\frac{H^-}{H^+}<0\) we obtain once again that (5.13) is at least necessary for the existence of such a \(\lambda ^-\) satisfying both (5.19) and (5.20), that in this case are equivalent to

$$\begin{aligned} \max \{ -a^-, b_1\}< \lambda ^- < \min \{a^-, b_2 \}, \end{aligned}$$

where \(b_1\) and \(b_2\) are defined in (5.22).

To complete the proof, it remains to show that condition (5.13) is also sufficient for the existence of functions \(\lambda ^\pm \) satisfying (5.11) and (5.12).

We already proved it in the simplest cases \(H^+\ne 0\) and \(H^-=0\) or viceversa. Now let us do the same in the case \(H^+\ne 0\) and \(H^-\ne 0\). Thus let us assume that the background state satisfies condition (5.13). If we assume for a while that \(\lambda ^\pm \) exist and satisfy (5.11) and (5.12), we derive that

$$\begin{aligned} \vert [u]\vert \le \vert \lambda ^+\vert \vert H^+\vert +\vert \lambda ^-\vert \vert H^-\vert < a^+\vert H^+\vert +a^-\vert H^-\vert , \end{aligned}$$

that is

$$\begin{aligned} \frac{\vert [u]\vert }{a^+\vert H^+\vert +a^-\vert H^-\vert }<1. \end{aligned}$$
(5.24)

The last inequality (5.24) suggests to take \(\lambda ^\pm \) such that

$$\begin{aligned} \vert \lambda ^\pm \vert =\frac{a^\pm \vert [u]\vert }{a^+\vert H^+\vert +a^-\vert H^-\vert }. \end{aligned}$$
(5.25)

Indeed from (5.24) it immediately follows that \(\lambda ^\pm \) as above satisfy the constraint (5.12). Formula (5.25) defines \(\lambda ^\pm \) up to the sign. One can directly check that taking

$$\begin{aligned} \lambda ^+=sgn(H^+)\frac{a^+ [u]}{a^+\vert H^+\vert +a^-\vert H^-\vert }\quad \text{ and }\quad \lambda ^-=-sgn(H^-)\frac{a^- [u]}{a^+\vert H^+\vert +a^-\vert H^-\vert } \nonumber \\ \end{aligned}$$
(5.26)

makes the equation (5.11) to be satisfied. The above definitions of \(\lambda ^\pm \) are just equal to (5.14). This ends the proof. \(\square \)

Remark 5.2

In order to make a comparison between our stability condition (5.13) and the subsonic part of the stability condition found in [40], let us consider the case of background piece-wise constant state

$$\begin{aligned} \hat{{\textbf{U}}}^\pm :=({\hat{\rho }}^\pm , 0, {\hat{u}}^\pm _2, 0, {\hat{H}}_2^\pm , {{\hat{S}}}^\pm )^T. \end{aligned}$$
(5.27)

Then the right-hand side of estimate (5.13) can be restated in terms of sound speeds \({{\hat{c}}}^\pm \) and the Alfvén speeds \({{\hat{c}}}^\pm _A=\frac{\vert \hat{{\textbf{H}}}^\pm \vert }{\sqrt{\rho ^\pm }}=\frac{\vert {{\hat{H}}}^\pm _2\vert }{\sqrt{\rho ^\pm }}\) as

$$\begin{aligned} {{\hat{a}}}^+\vert {\hat{H}}_2^+\vert +{{\hat{a}}}^-\vert {\hat{H}}_2^-\vert =\frac{{{\hat{c}}}^+{{\hat{c}}}^+_A}{(({{\hat{c}}}^+)^2+({{\hat{c}}}^+_A)^2)^{1/2}}+\frac{{{\hat{c}}}^-{{\hat{c}}}^-_A}{(({{\hat{c}}}^-)^2+({{\hat{c}}}^-_A)^2)^{1/2}} \end{aligned}$$

and (5.13) becomes

$$\begin{aligned} \vert [{{\hat{u}}}_2]\vert <\frac{{{\hat{c}}}^+{{\hat{c}}}^+_A}{(({{\hat{c}}}^+)^2+({{\hat{c}}}^+_A)^2)^{1/2}}+\frac{{{\hat{c}}}^-{{\hat{c}}}^-_A}{(({{\hat{c}}}^-)^2+({{\hat{c}}}^-_A)^2)^{1/2}}. \end{aligned}$$
(5.28)

In their two dimensional linear stability analysis [40], Wang and Yu are able to perform a complete normal modes analysis of the linearized problem for an isentropic flow, when the planar piece-wise constant basic state (5.27) (without \({{\hat{S}}}^\pm \)) satisfies suitable technical restrictions. Precisely, the constant components of \(\hat{{{\textbf{U}}}}^\pm \) are required to satisfy, besides the Rankine–Hugoniot conditions

$$\begin{aligned} {\hat{u}}_{2}^++{\hat{u}}_{2}^-=0,\quad {\hat{u}}_{2}^+>0,\quad {\hat{p}}^++\frac{({\hat{H}}^+_2)^2}{2}={\hat{p}}^-+\frac{({\hat{H}}^+_{2})^2}{2}, \end{aligned}$$
(5.29)

the additional assumptions

$$\begin{aligned} |{\hat{H}}_{2}^+|=|{\hat{H}}_{2}^-|=:|{\hat{H}}_2|,\quad {\hat{c}}>{\hat{c}}_A=\sqrt{\frac{|{\hat{H}}_2|^2}{{{\hat{\rho }}}}} \end{aligned}$$
(5.30)

(here the notation of [40] is adapted to our current setting). Notice in particular that, since the flow is isentropic, the assumption \(|{{\hat{H}}}_{2}^+|=|{{\hat{H}}}_{2}^-|\) and the last Rankine–Hugoniot condition in (5.29) imply \({{\hat{\rho }}}^-={{\hat{\rho }}}^+=:{{\hat{\rho }}}\) so that the sound speed and the Alfvén speed take the same value \({{\hat{c}}}=c({{\hat{\rho }}}):=\sqrt{p^\prime ({{\hat{\rho }}})}\) and \({{\hat{c}}}_A:=\sqrt{\frac{\vert {{\hat{H}}}_2\vert ^2}{{{\hat{\rho }}}}}\) on both ±-states of the flow.

The linear stability conditions by Wang-Yu [40] read as follows:

$$\begin{aligned} ({\hat{u}}_{2}^+)^2<{{\hat{c}}}^2-{{\hat{c}}}^2\sqrt{\frac{{{\hat{c}}}^2-{{\hat{c}}}^2_A}{{{\hat{c}}}^2+{{\hat{c}}}^2_A}}\qquad \text {or}\qquad ({\hat{u}}^+_2)^2>{{\hat{c}}}^2+{{\hat{c}}}^2\sqrt{\frac{{{\hat{c}}}^2-{{\hat{c}}}^2_A}{{{\hat{c}}}^2+{{\hat{c}}}^2_A}} . \end{aligned}$$
(5.31)

The left inequality in (5.31) identifies a “subsonic” weak stability region, becoming empty in the absence of a magnetic field (namely for compressible Euler equations), whereas the right inequality corresponds to the “supersonic” weak stability region of 2D compressible vortex sheets (to which it reduces, when formally setting \({{\hat{c}}}_A=0\)). With respect to (5.31), our stability condition in (5.28) provides only a subsonic stability domain. For a planar isentropic background state obeying (5.30), condition (5.28) reduces to

$$\begin{aligned} ({{\hat{u}}}_2^+)^2<\frac{{{\hat{c}}}^2_A{{\hat{c}}}^2}{({{\hat{c}}})^2+({{\hat{c}}}_A)^2}. \end{aligned}$$
(5.32)

To compare the left inequality in (5.31) and (5.32), one can easily check that

$$\begin{aligned} \frac{{{\hat{c}}}^2_A{{\hat{c}}}^2}{({{\hat{c}}})^2+({{\hat{c}}}_A)^2}<{{\hat{c}}}^2-{{\hat{c}}}^2\sqrt{\frac{{{\hat{c}}}^2-{{\hat{c}}}_A^2}{{{\hat{c}}}^2+{{\hat{c}}}_A^2}}, \end{aligned}$$

so that our subsonic condition (5.32) is more restrictive than the one in [40].

Let us also note that we are not able, by our approach, to obtain a counterpart of the supersonic condition in [40] for isentropic flows in the constant coefficients case (that is the right inequality in (5.31)). On the other hand, our stability condition applies also to nonisentropic flows and to general (non piece-wise constant) background states.

We are now in the position to state the well-posedness of the “homogeneous” linearized problem (4.31).

Theorem 5.1

Let all assumptions (4.2)–(4.10) be satisfied for the basic state (4.1) (note that (4.4) implies the stability condition (5.13)). Then, for all \({\textbf{F}}\in H^1_{*}(\Omega _T)\) that vanish in the past, problem (4.31) has a unique solution \((\dot{{{\textbf{U}}}}^\natural ,\varphi )\in H^{1}_{*}(\Omega _T)\times H^1(\Gamma _T).\) The solution satisfies the a priori estimate

$$\begin{aligned} ||\dot{{{\textbf{U}}}}^\natural ||_{1,*,T}+||\varphi ||_{H^1(\Gamma _T)}\le C ||{\textbf{F}}||_{1,*,T}, \end{aligned}$$
(5.33)

where \(C=C(K,T)>0\) is a constant independent of the data \({\textbf{F}}\).

The proof of Theorem 5.1 is given in Appendix A2.

As a consequence of Lemma 4.2 we have the following well-posedness result for the nonhomogeneous problem (4.20):

Theorem 5.2

Let all assumptions of Theorem 5.1 be satisfied for the basic state (4.1). Then, for all \(({\textbf{f}},{\textbf{g}})\in H^3_{*}(\Omega _T)\times H^{3}(\Gamma _T)\) that vanish in the past, problem (4.20) has a unique solution \((\dot{{{\textbf{U}}}},\varphi )\in H^{1}_{*}(\Omega _T)\times H^1(\Gamma _T)\). The solution satisfies the a priori estimate

$$\begin{aligned} ||\dot{{{\textbf{U}}}}||_{1,*,T}+||\varphi ||_{H^1(\Gamma _T)}\le C\Big (||{\textbf{f}}||_{3,*,T}+||{\textbf{g}}||_{H^3(\Gamma _T)}\Big ), \end{aligned}$$
(5.34)

where \(C=C(K,T)>0\) is a constant independent of the data \({\textbf{f}},{\textbf{g}}.\)

6 Higher-order energy estimates for the homogeneous problem (4.31)

In order to get an a priori tame estimate in \(H^s_{*}\) for solution of problem (4.20) with s sufficiently large, as a preliminary step we derive a tame estimate for the homogenous problem (4.31), that we state in the that theorem.

For shortness, in all this section we write \(\dot{{{\textbf{U}}}}^\natural =\dot{{{\textbf{U}}}}\).

Theorem 6.1

Let \(T>0\) and s be positive integer, \(s\ge 6.\) Assume that the basic state \((\breve{{\textbf{U}}}^{\pm },{\hat{\varphi }})\in H^{s+4}_{*}(\Omega _T)\times H^{s+5}(\Gamma _T),\)

$$\begin{aligned} ||\breve{{\textbf{U}}}^{\pm }||_{9,*,T}+||{\hat{\varphi }}||_{H^{10}(\Gamma _T)}\le {\hat{K}}, \text { where } \breve{{\textbf{U}}}^{\pm }:=\hat{\textbf{U}}^{\pm }-\bar{{\textbf{U}}}. \end{aligned}$$
(6.1)

Assume that \({\textbf{F}}\in H^s_{*}(\Omega _T)\) vanishes in the past. Then, there exists a positive constant \(K_0,\) that does not depend on s and T,  and there exists a constant \(C(K_0)>0\) such that if \({\hat{K}}\le K_0,\) then there exists a unique solution \((\dot{{{\textbf{U}}}},\varphi )\in H^s_{*}(\Omega _T)\times H^s(\Gamma _T)\) to homogeneous problem (4.31) satisfying the estimate

$$\begin{aligned} \begin{aligned} ||\dot{{{\textbf{U}}}}||_{s,*,T}+||\varphi ||_{H^s(\Gamma _T)}\le C(K_0)\Big (||{\textbf{F}}||_{s,*,T}+||{\textbf{F}}||_{6,*,T}||{\hat{W}}||_{s+4,*,T}\Big )\ \end{aligned} \end{aligned}$$
(6.2)

for T small enough, where \({\hat{W}}=(\breve{{\textbf{U}}},\nabla _{t,\textbf{x}}{\hat{\Psi }}).\)

To prove the above theorem we need to obtain several higher-order energy estimates.

6.1 Estimate of the normal derivative of the “non-characteristic” unknown

In this section we will prove the estimate of the normal derivative of the “non-characteristic” unknown

$$\begin{aligned} {{\textbf{V}}}_n=({{\textbf{V}}}^+_n,{{\textbf{V}}}^-_n),\quad {{\textbf{V}}}^{\pm }_n=({\dot{q}}^{\pm },{\dot{u}}^{\pm }_n,{\dot{H}}^{\pm }_n). \end{aligned}$$

Let s be positive integer, we need to estimate \(\partial _1{{\textbf{V}}}_n\) in \(H^{s-1}_{*},\) obtained from (5.3) and the divergence constraint (4.32) as follows:

$$\begin{aligned} \partial _1{{\textbf{V}}}^+_n=\left[ \begin{array}{c} {\mathcal {K}}_{1}\\ {\mathcal {K}}_{2}\\ -\partial _2({\dot{H}}^+_2\partial _1{\hat{\Phi }}^+) \end{array}\right] ,\quad \partial _1{{\textbf{V}}}^-_n=\left[ \begin{array}{c} {\mathcal {K}}_{7}\\ {\mathcal {K}}_{8}\\ -\partial _2({\dot{H}}^-_2\partial _1{\hat{\Phi }}^-) \end{array}\right] . \end{aligned}$$
(6.3)

Here \({\mathcal {K}}_{i}\in {{\mathbb {R}}}\) is the \(i-\)th scalar component of the vector

$$\begin{aligned} {\mathcal {K}}:=\tilde{{\mathcal {A}}}\Big ({\mathcal {F}}-{\mathcal {A}}_0\partial _t{{\textbf{V}}}-{\mathcal {A}}_2\partial _2{{\textbf{V}}}-{\mathcal {A}}_3{{\textbf{V}}}-{\mathcal {A}}_{(0)}\partial _1{{\textbf{V}}}\Big ), \end{aligned}$$
(6.4)

where \(\tilde{{\mathcal {A}}}=\textrm{diag}\Big ((\partial _1{\hat{\Phi }}^+){\mathcal {E}}_{12},(\partial _1{\hat{\Phi }}^-){\mathcal {E}}_{12}\Big ),\) \(\partial _1{\hat{\Phi }}^{\pm }=\pm 1+\partial _1{\hat{\Psi }}^{\pm },\) \({\mathcal {F}}=J^T{\textbf{F}}.\)

Now, we are ready to prove the following Lemma, which is needed in the proof of weighted normal derivatives and non-weighted tangential derivatives; see Sections 6.2 and 6.4.

Lemma 6.1

The estimate

$$\begin{aligned} ||\partial _1{{\textbf{V}}}_n||^2_{s-1,*,t}\le C(K){\mathcal {M}}(t), \end{aligned}$$
(6.5)

with

$$\begin{aligned} \begin{aligned} {\mathcal {M}}(t)&=||({{\textbf{V}}},{\textbf{F}})||^2_{s,*,t}+(||\dot{{{\textbf{U}}}}||^2_{W^{2,\infty }_{*}(\Omega _t)}+||{\textbf{F}}||^2_{W^{1,\infty }_{*}(\Omega _t)})||{\hat{W}}||^2_{s+2,*,t},\\ \end{aligned} \end{aligned}$$
(6.6)

holds for problem (4.31) for all \(t\le T.\)

Proof

Using Moser-type calculus inequalities (3.7) and (3.8), we can estimate the right-hand side of (6.4) as

$$\begin{aligned} \begin{aligned} ||\tilde{{\mathcal {A}}}{\mathcal {F}}||^2_{s-1,*,t}\le C(K)\Big (||{\textbf{F}}||^2_{s-1,*,t}+||{\textbf{F}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s-1,*,t}\Big ). \end{aligned} \end{aligned}$$
(6.7)

For \(j=0,2,\)

$$\begin{aligned}{} & {} \begin{aligned} ||\tilde{{\mathcal {A}}}{\mathcal {A}}_j\partial _j{{\textbf{V}}}||^2_{s-1,*,t}&\lesssim ||\tilde{{\mathcal {A}}}{\mathcal {A}}_j{{\textbf{V}}}||^2_{s,*,t}\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big ), \end{aligned} \end{aligned}$$
(6.8)
$$\begin{aligned}{} & {} \begin{aligned} ||\tilde{{\mathcal {A}}}{\mathcal {A}}_3{{\textbf{V}}}||^2_{s-1,*,t}\le C(K)\Big ( ||{{\textbf{V}}}||^2_{s-1,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s+1,*,t}\Big ), \end{aligned} \end{aligned}$$
(6.9)

where C(K) stands for positive constants that depends on K and we used that \(\tilde{{\mathcal {A}}}J^T\), \(\tilde{{\mathcal {A}}}{\mathcal {A}}_j\) for \(j=0,2\) and \(\tilde{{\mathcal {A}}}{\mathcal {A}}_3\) are all nonlinear smooth functions of the basic state \({\hat{W}}\).

Now, we estimate the last term in (6.4), where for simplicity we denote \({\mathcal {A}}=\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}\). We get

$$\begin{aligned} \begin{aligned} ||{\mathcal {A}}\partial _1{{\textbf{V}}}||_{s-1,*,t}&=||\frac{{\mathcal {A}}}{\sigma }\sigma \partial _1{{\textbf{V}}}||_{s-1,*,t}\\&\le ||\frac{{\mathcal {A}}}{\sigma }||_{W^{1,\infty }_*(\Omega _t)}||\sigma \partial _1{{\textbf{V}}}||_{s-1,*,t}+||\frac{{\mathcal {A}}}{\sigma }||_{s-1,*,t}||\sigma \partial _1{{\textbf{V}}}||_{W^{1,\infty }_*(\Omega _t)}\\&\le ||{\mathcal {A}}||_{W^{2,\infty }_*(\Omega _t)}||{{\textbf{V}}}||_{s,*,t}+||{\mathcal {A}}||_{s+1,*,t}||{{\textbf{V}}}||_{W^{2,\infty }_*(\Omega _t)}\\&\le ||{\hat{W}}||_{W^{2,\infty }_*(\Omega _t)}||{{\textbf{V}}}||_{s,*,t}+||{\hat{W}}||_{s+1,*,t}||{{\textbf{V}}}||_{W^{2,\infty }_*(\Omega _t)}, \end{aligned} \nonumber \\ \end{aligned}$$
(6.10)

where we used that \({\mathcal {A}}\vert _{x_1=0}=0\), thus the \(L^\infty -\)norm of \({\mathcal {A}}/\sigma \) can be estimated by the \(L^\infty \)-norms of \({\mathcal {A}}\) and \(\partial _1{\mathcal {A}}\), see [24, Lemma B.9], [32].

For the rest of the term in (6.3), using (3.7) and (3.8), we obtain that

$$\begin{aligned} \begin{aligned} ||\partial _2({\dot{H}}^+_2\partial _1{\hat{\Phi }}^{+})||^2_{s-1,*,t}&\le ||{\dot{H}}^+_2\partial _1{\hat{\Phi }}^{+}||^2_{s,*,t}\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big ). \end{aligned} \nonumber \\ \end{aligned}$$
(6.11)

The second term in (6.3) can be controlled similarly. Summarizing all the estimates (6.7)–(6.11) for the terms in (6.3) and (6.4), Lemma 6.1 is concluded. \(\square \)

We also need the following estimate, still for \(\partial _1{{\textbf{V}}}_n\), which is also essential in the proof of non-weighted tangential derivatives, see (6.47) in Section 6.4:

Lemma 6.2

The estimate

$$\begin{aligned} |||\partial _1{{\textbf{V}}}_n(t)|||^2_{s-1,*}\le C(K)\Big (|||{{\textbf{V}}}(t)|||^2_{s,*}+{\mathcal {M}}(t)\Big ) \end{aligned}$$
(6.12)

holds for problem (4.31) for all \(t\le T\), where s is a positive integer and \({\mathcal {M}}(t)\) is defined in (6.6).

Proof

Denote the differential operator \(D^{\alpha }_{*}=\partial ^{\alpha _0}_t(\sigma \partial _1)^{\alpha _1}\partial ^{\alpha _2}_2\partial ^{\alpha _3}_1\), \(\langle \alpha \rangle :=|\alpha |+\alpha _3.\) We estimate the right-hand side of (6.4). Using the elementary estimate

$$\begin{aligned} |||u(t)|||^2_{s-1,*}\lesssim ||u||^2_{s,*,t}, \end{aligned}$$
(6.13)

we get

$$\begin{aligned} \begin{aligned} |||(\tilde{{\mathcal {A}}}{\mathcal {F}})(t)|||^2_{s-1,*}&\lesssim ||\tilde{{\mathcal {A}}}J^T{\textbf{F}}||^2_{s,*,t}\\&\le C(K)\Big (||{\textbf{F}}||^2_{s,*,t}+||{\textbf{F}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big )\\&\le C(K){\mathcal {M}}(t). \end{aligned} \end{aligned}$$
(6.14)

The second term of (6.4) can be controlled as follows:

$$\begin{aligned} \begin{aligned} |||\tilde{{\mathcal {A}}}{\mathcal {A}}_{0}&\partial _t{{\textbf{V}}}(t)|||^2_{s-1,*}\lesssim \sum \limits _{\langle \alpha \rangle \le s-1}||\tilde{{\mathcal {A}}}{\mathcal {A}}_{0}D^\alpha _*\partial _t{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\\&+\sum \limits _{\langle \alpha \rangle \le s-1}||D_*(\tilde{{\mathcal {A}}}{\mathcal {A}}_{0})D^{\alpha -1}_*\partial _t{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}+\sum \limits _{\langle \alpha \rangle =2}||D^\alpha _*(\tilde{{\mathcal {A}}}{\mathcal {A}}_{0})\partial _t{{\textbf{V}}}(t)||^2_{s-3,*}\\&=\Sigma _1^\prime +\Sigma ^\prime _2+\Sigma _3^\prime , \end{aligned} \end{aligned}$$

where \(D_*\) denotes any tangential derivative in t or \(\textbf{x}\) of order one, and \(D^{\alpha -1}_*\) denotes the derivative of order \(\langle \alpha \rangle -1\) obtained “subtracting" \(D_*\) from \(D_{*}^\alpha \).

We estimate separately each \(\Sigma ^\prime _i\), \(i=1,2,3\) as follows:

$$\begin{aligned} \Sigma _1^\prime&\lesssim ||{{\hat{W}}}||^2_{L^\infty (\Omega _t)}|||{\textbf{V}}(t)|||^2_{s,*}\le C(K)|||{\textbf{V}}(t)|||^2_{s,*}; \\ \Sigma _2^\prime&\lesssim ||{{\hat{W}}}||^2_{W^{1,\infty }_*(\Omega _t)}|||{\textbf{V}}(t)|||^2_{s-1,*}\le C(K)|||{\textbf{V}}(t)|||^2_{s,*}; \\ \Sigma _3^\prime&\lesssim \sum \limits _{\langle \alpha \rangle =2}||D^\alpha _*({{\tilde{A}}}{\mathcal {A}}_0)\partial _t{\textbf{V}}|||^2_{s-2,*,t}\\&\lesssim \sum \limits _{\langle \alpha \rangle =2}\left( ||D^\alpha _*({{\tilde{A}}}{\mathcal {A}}_0)||^2_{W^{1,\infty }_*(\Omega _t)}||\partial _t{\textbf{V}}||^2_{s-2,*,t}\right. \\&\quad \left. +||D^\alpha _*(\tilde{{\mathcal {A}}}{\mathcal {A}}_0)||^2_{s-2,*,t}||\partial _t{\textbf{V}}||^2_{W^{1,\infty }_*(\Omega _t)}\right) \\&\le C(K)||{\textbf{V}}||^2_{s-1,*,t}+||{{\hat{W}}}||^2_{s,*,t}||{\textbf{V}}||^2_{W^{2,\infty }_*(\Omega _t)}. \end{aligned}$$

Adding the above inequalities, we get

$$\begin{aligned} |||\tilde{{\mathcal {A}}}{\mathcal {A}}_{0}\partial _t{{\textbf{V}}}(t)|||^2_{s-1,*}\le C(K)|||{{\textbf{V}}}(t)|||^2_{s,*}+C(K)||{\textbf{V}}||^2_{s-1,*,t}+||{{\hat{W}}}||^2_{s,*,t}||{\textbf{V}}||^2_{W^{2,\infty }_*(\Omega _t)}. \nonumber \\ \end{aligned}$$
(6.15)

Similarly, the third term can be estimated by

$$\begin{aligned} |||\tilde{{\mathcal {A}}}{\mathcal {A}}_2\partial _2{{\textbf{V}}}(t)|||^2_{s-1,*}\le C(K)|||{{\textbf{V}}}(t)|||^2_{s,*}+C(K)||{\textbf{V}}||^2_{s-1,*,t}+||{{\hat{W}}}||^2_{s,*,t}||{\textbf{V}}||^2_{W^{2,\infty }_*(\Omega _t)}. \nonumber \\ \end{aligned}$$
(6.16)

Using Moser-type calculus inequalities (3.7) and (3.8), we can estimate

$$\begin{aligned} \begin{aligned} |||(\tilde{{\mathcal {A}}}{\mathcal {A}}_3{\textbf{V}})(t)|||^2_{s-1,*}&\lesssim ||\tilde{{\mathcal {A}}}{\mathcal {A}}_3{\textbf{V}}||^2_{s,*,t}\\&\le C(K)\Big (||{\textbf{V}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s+2,*,t}\Big )\\&\le C(K){\mathcal {M}}(t). \end{aligned}\nonumber \\ \end{aligned}$$
(6.17)

Now, we estimate the last term \(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}\partial _1{{\textbf{V}}}\) as

$$\begin{aligned} |||(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}\partial _1{{\textbf{V}}})(t)|||^2_{s-1,*}\lesssim \Sigma _1+\Sigma _2+\Sigma _3, \end{aligned}$$

where

$$\begin{aligned}{} & {} \begin{aligned} \Sigma _1&=\sum _{\langle \alpha \rangle \le s-1}||\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}D^{\alpha }_{*}\partial _1{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\\&=\sum _{\langle \alpha \rangle \le s-1}||\frac{\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}}{\sigma }\sigma D^{\alpha }_{*}\partial _1{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\\&\lesssim ||\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}||^2_{W^{1,\infty }(\Omega _t)}|||{\textbf{V}}(t)|||^2_{s,*}\lesssim ||{{\hat{W}}}||_{W^{1,\infty }(\Omega _t)}^2|||{\textbf{V}}(t)|||^2_{s,*}; \end{aligned} \\{} & {} \begin{aligned} \Sigma _2&=\sum \limits _{\langle \alpha \rangle \le s-1}||\frac{D_*(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)})}{\sigma }\sigma D^{\alpha -1}_{*}\partial _1{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\\&\lesssim ||D_*(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)})||^2_{W^{1,\infty }(\Omega _t)}|||{\textbf{V}}(t)|||^2_{s-1,*}\lesssim ||{{\hat{W}}}||^2_{W^{2,\infty }_*(\Omega _t)}|||{\textbf{V}}(t)|||^2_{s-1,*}; \end{aligned} \\{} & {} \begin{aligned} \Sigma _3&=\sum \limits _{\langle \alpha \rangle =2}|||D^\alpha _*(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)})\partial _1{\textbf{V}}(t)|||^2_{s-3,*}\lesssim \sum \limits _{\langle \alpha \rangle =2}|||D^\alpha _*(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)})\partial _1{\textbf{V}}|||^2_{s-2,*,t}\\&\lesssim ||{{\hat{W}}}||^2_{W^{3,\infty }(\Omega _t)}||{\textbf{V}}||^2_{s,*,t}+||{{\hat{W}}}||^2_{s,*,t}||{\textbf{V}}||^2_{W^{2,\infty }_*(\Omega _t)}, \end{aligned} \end{aligned}$$

and where we exploit again the vanishing of \(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}\) and \(D_*(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)})\) along the boundary \(\{x_1=0\}\) in the estimates of \(\Sigma _1\) and \(\Sigma _2\), see [24, Lemma B.9], [32].

Adding the estimates of \(\Sigma _i\), \(i=1,2,3\) above, we get

$$\begin{aligned} |||(\tilde{{\mathcal {A}}}{\mathcal {A}}_{(0)}\partial _1{{\textbf{V}}})(t)|||^2_{s-1,*}\le C(K)|||{{\textbf{V}}}(t)|||^2_{s,*}+C(K)||{\textbf{V}}||^2_{s,*,t}+||{{\hat{W}}}||^2_{s,*,t}||{\textbf{V}}||^2_{W^{2,\infty }_*(\Omega _t)}. \nonumber \\ \end{aligned}$$
(6.18)

For the rest of the term in (6.3), using (3.7) and (3.8), we obtain that

$$\begin{aligned} \begin{aligned} |||\partial _2({\dot{H}}^+_2\partial _1{\hat{\Phi }}^{+})(t)|||^2_{s-1,*,t}&\le ||{\dot{H}}^+_2\partial _1{\hat{\Phi }}^{+}||^2_{s,*,t}\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s+2,*,t}\Big )\\&\le C(K){\mathcal {M}}(t). \end{aligned}\nonumber \\ \end{aligned}$$
(6.19)

The second term in (6.3) can be controlled similarly. Summarizing from (6.14)–(6.19), we conclude that (6.12) holds. \(\square \)

The following lemma gives the estimate of the normal derivative of the entropy, which is treated differently from the other components of the vector \({{\textbf{V}}}\):

Lemma 6.3

The estimate

$$\begin{aligned} |||S^\pm (t)|||^2_{s,*}\le C(K){\mathcal {M}}(t) \end{aligned}$$
(6.20)

holds for problem (4.31) for all \(t\le T\), where s is a positive integer and \({\mathcal {M}}(t)\) is defined in (6.6).

Proof

The linearized equation for the entropy \( S^\pm \) is an evolution-like equation because the coefficient of the normal derivative of the entropy vanishes on the boundary; this yields that no boundary conditions are needed to be coupled to the equation in order to derive an a priori energy estimate. Thus, to estimate \( S^\pm \), we just handle the equation of \( S^\pm \) alone by the standard energy method tools. The details of the proof are similar to those of the following Lemma 6.4. \(\square \)

Now, we derive weighted derivative estimates.

6.2 Estimate of weighted derivatives

Since the differential operators \((\sigma \partial _1)^{\alpha _1}\) and \(\sigma ^{\alpha _1}\partial ^{\alpha _1}_1\) are equivalent, see [28], in the following we discuss the term \(D^{\alpha }_{*}{{\textbf{V}}}\), with \(\alpha _1>0\) and \(\langle \alpha \rangle =|\alpha |+\alpha _3\le s\), in its equivalent form \(\sigma ^{\alpha _1}D^{\alpha ^\prime }_{t,x}\partial ^{\alpha _3}_1{\textbf{V}}\), where \(D^{\alpha ^\prime }_{t,x}:=\partial ^{\alpha _0}_t\partial ^{\alpha _1}_{1}\partial ^{\alpha _2}_{2}\) (\(\alpha ^\prime =(\alpha _0,\alpha _1,\alpha _2)\)).

Lemma 6.4

The following estimate holds for (4.31) for all \(t\le T:\)

$$\begin{aligned} \sum _{\langle \alpha \rangle \le s,\alpha _1>0}||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^2(\Omega )}\le C(K){\mathcal {M}}(t) \end{aligned}$$
(6.21)

there s is a positive integer and \({\mathcal {M}}(t)\) is defined in (6.6).

Proof

It is obvious that when \(\alpha _1>0,\) \(D^{\alpha }_{*}{{\textbf{V}}}|_{x_1=0}=0.\) Note that

$$\begin{aligned} \sigma ^{\alpha _1}\partial ^{m+1}_1=\partial _1(\sigma ^{\alpha _1}\partial ^m_1)-\alpha _1\sigma '\sigma ^{\alpha _1-1}\partial ^m_1, \end{aligned}$$

for some nonnegative integer m. Applying \(D^{\alpha }_{*}\) to (5.3) and using the standard energy method, we obtain that

$$\begin{aligned} \int _{{{\mathbb {R}}}^2_+}({\mathcal {A}}_0D^{\alpha }_{*}{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}})(t)\textrm{d}{\textbf{x}}=\int _{\Omega _t}\Big (((\partial _t{\mathcal {A}}_0+\partial _1{\mathcal {A}}_1+\partial _2{\mathcal {A}}_2)D^{\alpha }_{*}{{\textbf{V}}}+2{\mathcal {R}})\cdot D^{\alpha }_{*}{{\textbf{V}}}\Big )\textrm{d}{\textbf{x}}\textrm{d}\tau , \end{aligned}$$

where

$$\begin{aligned}{} & {} {\mathcal {R}}=D^{\alpha }_{*}{\mathcal {F}}+{\mathcal {R}}_0+{\mathcal {R}}_1,\quad {\mathcal {R}}_0=\alpha _1\sigma '{\mathcal {A}}_1\sigma ^{\alpha _1-1}D^{\alpha '}_{t,x}\partial ^{\alpha _3}_1{{\textbf{V}}}, \\{} & {} {\mathcal {R}}_1=-[D^{\alpha }_{*},{\mathcal {A}}_0]\partial _t{{\textbf{V}}}-\sum ^2_{j=1}[D^{\alpha }_{*},{\mathcal {A}}_j]\partial _j{{\textbf{V}}}-D^{\alpha }_{*}({\mathcal {A}}_3{{\textbf{V}}}) \end{aligned}$$

(the brackets \([\cdot ,\cdot ]\) here denotes the commutator between the operators). Notice that \({\mathcal {A}}_0\) is positive definite, then

$$\begin{aligned} \int _{{{\mathbb {R}}}^2_+}({\mathcal {A}}_0D^{\alpha }_{*}{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}})(t)\textrm{d}{\textbf{x}}\ge c_0||D^{\alpha }_{*}{\textbf{V}}(t)||^2_{L^2({\mathbb {R}}^2_+)}, \end{aligned}$$

where \(c_0\) depends on the number k in (4.2) and (4.4). Hence, we obtain that

$$\begin{aligned} ||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^{2}({{\mathbb {R}}}^2_+)}\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||{\mathcal {R}}||^2_{L^2(\Omega _t)}\Big ). \end{aligned}$$
(6.22)

Now, we estimate \(||{\mathcal {R}}||^2_{L^2(\Omega _t)}\) in (6.22). Recall that \({\mathcal {R}}=D^{\alpha }_{*}{\mathcal {F}}+{\mathcal {R}}_0+{\mathcal {R}}_1.\) Using Moser-type calculus inequalities (3.7) and (3.8), we can prove that

$$\begin{aligned} \begin{aligned} ||D^{\alpha }_{*}{\mathcal {F}}||^2_{L^2(\Omega _t)}&\lesssim ||J^T{\textbf{F}}||^2_{s,*,t}\\&\le C(K)\Big (||{\textbf{F}}||^2_{s,*,t}+||{\textbf{F}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big ). \end{aligned} \end{aligned}$$
(6.23)

Using the decomposition of boundary matrix \({\mathcal {A}}_1\) in (5.5), \({\mathcal {A}}_1={\mathcal {A}}+{\mathcal {A}}_{(0)}\), following arguments similar to those used in Lemma 6.1, we obtain that

$$\begin{aligned} \begin{aligned} ||{\mathcal {R}}_0||^2_{L^2(\Omega _t)}&\lesssim ||{\mathcal {A}}\sigma ^{\alpha _1-1}D^{\alpha '}_{t,x}\partial ^{\alpha _3}_1{{\textbf{V}}}||^2_{L^2(\Omega _t)}+||\frac{{\mathcal {A}}_{(0)}}{\sigma }\sigma ^{\alpha _1}D^{\alpha '}_{t,x}\partial ^{\alpha _3}_1{{\textbf{V}}}||^2_{L^2(\Omega _t)}\\&\lesssim C(K)\left( ||\sigma ^{\alpha _1-1}D^{\alpha '}_{t,x}\partial ^{\alpha _3}_1{{\textbf{V}}}_n||^2_{L^2(\Omega _t)}+ ||\sigma ^{\alpha _1}D^{\alpha '}_{t,x}\partial ^{\alpha _3}_1{{\textbf{V}}}||^2_{L^2(\Omega _t)}\right) \\&\le C(K)\Big (||\partial _1{{\textbf{V}}}_n||^2_{s-1,*,t}+||{{\textbf{V}}}||^2_{s,*,t}\Big )\\&\le C(K){\mathcal {M}}(t), \end{aligned}\nonumber \\ \end{aligned}$$
(6.24)

recall here above that the matrix \({\mathcal {A}}\) acts only on the noncharacteristic part \({{\textbf{V}}}_n\) of the unknown \({\textbf{V}}\).

Now, we estimate the commutators in \({\mathcal {R}}_1:\) For \(j=0,2,\) we obtain that

$$\begin{aligned} \begin{aligned} ||[D^{\alpha }_{*},{\mathcal {A}}_j]\partial _j{{\textbf{V}}}||^2_{L^2(\Omega _t)}\le&\sum _{0<\beta \le \alpha }||D^{\beta }_{*}{\mathcal {A}}_jD^{\alpha -\beta }_{*}\partial _j{{\textbf{V}}}||^2_{L^2(\Omega _t)}. \end{aligned} \end{aligned}$$
(6.25)

For \(\langle \beta \rangle =1\), we get

$$\begin{aligned} \sum _{\langle \beta \rangle =1,\,\beta \le \alpha }||D^{\beta }_{*}{\mathcal {A}}_jD^{\alpha -\beta }_{*}\partial _j{{\textbf{V}}}||^2_{L^2(\Omega _t)}\le C(K)||{{\textbf{V}}}||^2_{s,*,t}. \end{aligned}$$
(6.26)

For \(\langle \beta \rangle \ge 2,\) we obtain that

$$\begin{aligned}{} & {} \sum _{\langle \beta \rangle \ge 2,\,\beta \le \alpha }||D^{\beta }_{*}{\mathcal {A}}_jD^{\alpha -\beta }_{*}\partial _j{{\textbf{V}}}||^2_{L^2(\Omega _t)}\nonumber \\{} & {} \quad \lesssim \sum _{\langle \beta '\rangle =2,\beta '\le \beta \le \alpha }||D^{\beta -\beta '}_{*}(D^{\beta '}_{*}{\mathcal {A}}_j)D^{\alpha -\beta }_{*}(\partial _j{{\textbf{V}}})||^2_{L^2(\Omega _t)}\nonumber \\{} & {} \quad \le C(K)\Big (||{{\textbf{V}}}||^2_{s-1,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{2,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big ). \end{aligned}$$
(6.27)

For \(j=1,\) we need to be very careful. We have

$$\begin{aligned} \begin{aligned} ||[D^{\alpha }_{*},{\mathcal {A}}_1]\partial _1{{\textbf{V}}}||^2_{L^2(\Omega _t)}&\lesssim \sum _{0<\beta \le \alpha }||D^{\beta }_{*}{\mathcal {A}}_1D^{\alpha -\beta }_{*}(\partial _1{{\textbf{V}}})||^2_{L^2(\Omega _t)}. \end{aligned} \end{aligned}$$
(6.28)

For \(\langle \beta \rangle =1\) we get \(\beta _3=0.\) Therefore, from (5.5) it follows that \(D^{\beta }_{*}{\mathcal {A}}_1|_{x_1=0}=0\) and we use Moser-type calculus inequalities (3.7), (3.8) to obtain

$$\begin{aligned} \sum _{\langle \beta \rangle =1,\,\beta \le \alpha }||D^{\beta }_{*}{\mathcal {A}}_1D^{\alpha -\beta }_{*}(\partial _1{{\textbf{V}}})||^2_{L^2(\Omega _t)}\le C(K)||{{\textbf{V}}}||^2_{s,*,t}. \end{aligned}$$
(6.29)

For \(\langle \beta \rangle \ge 2\), we have

$$\begin{aligned}{} & {} \sum _{\langle \beta \rangle \ge 2,\,\beta \le \alpha }||D^{\beta }_{*}{\mathcal {A}}_1D^{\alpha -\beta }_{*}(\partial _1{{\textbf{V}}})||^2_{L^2(\Omega _t)} \nonumber \\{} & {} \quad \lesssim \sum _{\langle \beta '\rangle =2,\beta '\le \beta \le \alpha }||D^{\beta -\beta '}_{*}(D^{\beta '}_{*}{\mathcal {A}}_1)D^{\alpha -\beta }_{*}(\partial _1{{\textbf{V}}})||^2_{L^2(\Omega _t)} \nonumber \\{} & {} \quad \le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{2,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big ). \end{aligned}$$
(6.30)

Using Moser-type calculus inequalities (3.7), (3.8), we can prove that

$$\begin{aligned} \begin{aligned} ||D^{\alpha }_{*}({\mathcal {A}}_3{{\textbf{V}}})||^2_{L^2(\Omega _t)}&\lesssim ||{\mathcal {A}}_3{{\textbf{V}}}||^2_{s,*,t}\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s+2,*,t}\Big ). \end{aligned} \end{aligned}$$
(6.31)

Notice that summing up (6.25)–(6.31) gives

$$\begin{aligned} ||{\mathcal {R}}_1||^2_{L^2(\Omega _t)}\le C(K){\mathcal {M}}(t). \end{aligned}$$
(6.32)

Hence, using (6.23), (6.24) and (6.32), we obtain (6.21). Lemma 6.4 is concluded. \(\square \)

Now, we are going to estimate the non-weighted normal derivatives.

6.3 Estimate of non-weighted normal derivatives

Now we perform the differential operator \(D^{\alpha }_{*}\), in the case \(\alpha _1=0, \alpha _3\ge 1\), that is \(D^\alpha _*=\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3}_1\) with \(\langle \alpha \rangle \le s\). Now, we are ready to prove the following:

Lemma 6.5

The estimate

$$\begin{aligned} \sum _{\langle \alpha \rangle \le s, \alpha _1=0, \alpha _3\ge 1}||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\le C(K){\mathcal {M}}(t) \end{aligned}$$
(6.33)

holds for problem (4.31) for all \(t\le T,\) where s is a positive integer, \({\mathcal {M}}(t)\) is given in (6.6).

Proof

Similar as in Lemma 6.4, applying the operator \(D^{\alpha }_{*}\) on (5.3) and using the standard energy method, we obtain that

$$\begin{aligned}{} & {} \int _{{{\mathbb {R}}}^2_+}({\mathcal {A}}_0D^{\alpha }_{*}{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}})(t)\textrm{d}{\textbf{x}}\\{} & {} \quad =\int _{\Omega _t}\Big (((\partial _t{\mathcal {A}}_0+\partial _1{\mathcal {A}}_1+\partial _2{\mathcal {A}}_2)D^{\alpha }_{*}{{\textbf{V}}}+2{\mathcal {R}})\cdot D^{\alpha }_{*}{{\textbf{V}}}\Big )\textrm{d}{\textbf{x}}\textrm{d}\tau \\{} & {} \qquad +\int _{\Gamma _t}({\mathcal {A}}_1D^{\alpha }_{*}{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}})|_{x_1=0}\textrm{d}x_2\textrm{d}\tau , \end{aligned}$$

where \({\mathcal {R}}\) is defined as in the proof of Lemma 6.4. Thus

$$\begin{aligned} ||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||{\mathcal {R}}||^2_{L^2(\Omega _t)}+\tilde{{\mathcal {M}}}(t)\Big ), \end{aligned}$$

where

$$\begin{aligned} \tilde{{\mathcal {M}}}(t)=\Big |\int _{\Gamma _t}({\mathcal {A}}_1D^{\alpha }_{*}{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}})|_{x_1=0}\textrm{d}x_2\textrm{d}\tau \Big |=\Big |\int _{\Gamma _t}(D^{\alpha }_{*}{\dot{q}}D^{\alpha }_{*}{\dot{u}}_n)|_{x_1=0}\textrm{d}x_2\textrm{d}\tau \Big |. \end{aligned}$$

When \(\alpha _1=0,\) by definition, \({\mathcal {R}}_0=0.\) Hence, we obtain the estimate

$$\begin{aligned} ||{\mathcal {R}}||^2_{L^2(\Omega _t)}\le C(K){\mathcal {M}}(t). \end{aligned}$$

This yields that

$$\begin{aligned} ||D^{\alpha }_*{{\textbf{V}}}(t)||^2_{L^2({{\mathbb {R}}}^2_+)}\le C(K)({\mathcal {M}}(t)+\tilde{{\mathcal {M}}}(t)). \end{aligned}$$
(6.34)

Using (6.3) and (6.4), we obtain that

$$\begin{aligned} \tilde{{\mathcal {M}}}(t)\lesssim ||\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3-1}_1{\mathcal {K}}||^2_{L^2(\Gamma _t)}, \end{aligned}$$

where \({\mathcal {K}}\) is defined in (6.4). Moreover, recalling that \(\tilde{{\mathcal {A}}}|_{x_1=0}=\textrm{diag}\Big ({\mathcal {E}}_{12},-{\mathcal {E}}_{12}\Big )\) because \(\partial _1{\hat{\Phi }}^{\pm }|_{x_1=0}=\pm 1,\) we notice that \(||\tilde{{\mathcal {A}}}|_{x_1=0}||_{L^{\infty }(\Gamma _t)}=1.\) Using \({\mathcal {A}}_{(0)}|_{x_1=0}=0\) we obtain that

$$\begin{aligned} \tilde{{\mathcal {M}}}(t)\le & {} C\Bigg (\Sigma _1(t)+\Sigma _2(t)+\Sigma _3(t) +||\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3-1}_1({\mathcal {A}}_3{{\textbf{V}}})||^2_{L^2(\Gamma _t)} \nonumber \\{} & {} +||\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3-1}_1{\mathcal {F}}||^2_{L^2(\Gamma _t)}\Bigg ), \end{aligned}$$
(6.35)

where

$$\begin{aligned} \Sigma _1(t)= & {} \sum _{j=0,2}||{\mathcal {A}}_j\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}}||^2_{L^2(\Gamma _t)}, \\ \Sigma _2(t)= & {} \sum _{j=0,2}\sum _{\langle \beta '\rangle \ge 1,\beta '+\beta ''=(\alpha _0,\alpha _2,\alpha _3-1)}||D^{\beta '}_{*}{\mathcal {A}}_jD^{\beta ''}_{*}\partial _j{{\textbf{V}}}||^2_{L^2(\Gamma _t)}; \end{aligned}$$

notice that in spite of the notation here \(D^{\beta '}_*\) and \(D^{\beta ''}_*\) dot not involve any weighted derivative \(\sigma \partial _1\) (see the beginning of this section); notice also that \(\beta '+\beta ''=(\alpha _0,\alpha _2,\alpha _3-1)\) implies that \(\langle \beta '\rangle +\langle \beta ''\rangle \le s-2\);

$$\begin{aligned} \Sigma _3(t)= & {} \sum _{\alpha _3'+\alpha _3''\le \alpha _3,\alpha _3',\alpha _3''\ge 1}||\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2(\partial ^{\alpha _3'}_1{\mathcal {A}}_{(0)}\partial ^{\alpha _3''}_1{{\textbf{V}}})||^2_{L^2(\Gamma _t)}. \end{aligned}$$

Now, we estimate \(\Sigma _2(t)\) and \(\Sigma _3(t)\), which can be controlled by using Lemma 3.4 (i) and Moser-type calculus inequalities (3.7) and (3.8) as follows:

$$\begin{aligned} \begin{aligned} \Sigma _2(t)&\lesssim \sum _{j=0,2}\sum _{\langle \beta '\rangle =1}\Big (||D^{\beta '}_{*}{\mathcal {A}}_j\partial _j{{\textbf{V}}}||^2_{s-3,*,t}+||\partial _1(D^{\beta '}_{*}{\mathcal {A}}_j\partial _j{{\textbf{V}}})||^2_{s-3,*,t}\Big )\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{2,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big ). \end{aligned} \end{aligned}$$
(6.36)

In the above estimate it is noted that \(\langle \beta ''\rangle \le s-3\).Then

$$\begin{aligned} \begin{aligned} \Sigma _3(t)&\lesssim ||\partial _1{\mathcal {A}}_{(0)}\partial _1{{\textbf{V}}}||^2_{s-4,*,t}+||\partial _1(\partial _1{\mathcal {A}}_{(0)}\partial _1{{\textbf{V}}})||^2_{s-4,*,t}\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\dot{{{\textbf{U}}}}||^2_{W^{2,\infty }_{*}(\Omega _t)}||{\hat{W}}||^2_{s,*,t}\Big ). \end{aligned} \end{aligned}$$
(6.37)

For the term \(\Sigma _1(t),\) passing to the volume integral then using Leibniz’s rule, we get (for shortness, in the sequel we denote \(D^\alpha :=\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\) whereas \(D^\alpha _*:=\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3}_1\))

$$\begin{aligned} \begin{aligned} \Sigma _1(t)&=-\sum _{j=0,2}\int _{\Omega _t}\partial _1|{\mathcal {A}}_jD^{\alpha }\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}}|^2\textrm{d}{\textbf{x}}\textrm{d}\tau \\&=-2\sum _{j=0,2}\int _{\Omega _t}\Big (({\mathcal {A}}^2_jD^{\alpha }\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}}\cdot D^{\alpha }_{*}\partial _j{{\textbf{V}}})\\&\quad +({\mathcal {A}}_jD^{\alpha }\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}}\cdot \partial _1{\mathcal {A}}_jD^{\alpha }\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}})\Big )\textrm{d}{\textbf{x}}\textrm{d}\tau . \end{aligned} \end{aligned}$$

Then integration by parts with respect to \(\partial _j\) (for \(j=0,2)\) gives

$$\begin{aligned} \Sigma _1(t)={\tilde{\Sigma }}_1(t)+J_0(t), \end{aligned}$$

where

$$\begin{aligned}{} & {} \begin{aligned} {\tilde{\Sigma }}_1(t)&=2\sum _{j=0,2}\int _{\Omega _t}\Big ({\mathcal {A}}^2_jD^{\alpha }\partial ^{\alpha _3-1}_1\partial ^2_j{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}}\Big )\\&\quad +\Big (({\mathcal {A}}_j\partial _j{\mathcal {A}}_j+\partial _j{\mathcal {A}}_j{\mathcal {A}}_j)D^{\alpha }\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}}\Big )\\&\quad -\Big ({\mathcal {A}}_jD^{\alpha }\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}}\cdot \partial _1{\mathcal {A}}_jD^{\alpha }\partial ^{\alpha _3-1}_1\partial _j{{\textbf{V}}}\Big )\textrm{d}{\textbf{x}}\textrm{d}\tau , \end{aligned} \\{} & {} J_0(t)=-2\int _{{\mathbb {R}}^2_+}\Big ({\mathcal {A}}^2_0D^{\alpha }\partial ^{\alpha _3-1}_1\partial _t{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}}\Big )(t)\textrm{d}{\textbf{x}}. \end{aligned}$$

Since \(|\alpha |+2+2(\alpha _3-1)\le s\) we obtain that

$$\begin{aligned} {\tilde{\Sigma }}_1(t)\le C(K)||{{\textbf{V}}}||^2_{s,*,t}. \end{aligned}$$
(6.38)

Using Young’s inequality, we obtain that

$$\begin{aligned} J_0(t)\le C(K)\Big (\varepsilon ||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}+\frac{1}{\varepsilon }|||{{\textbf{V}}}(t)||||^2_{s-1,*}\Big ), \end{aligned}$$
(6.39)

for small \(\varepsilon .\) Therefore, we conclude from (6.38), (6.39) and elementary inequalities (6.13) that

$$\begin{aligned} \Sigma _1(t)\le C(K)\Big (\varepsilon ||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}+\frac{1}{\varepsilon }||{{\textbf{V}}}||^2_{s,*,t}\Big ). \end{aligned}$$
(6.40)

The last two terms in (6.35) can be estimated by using Lemma 3.4 (i), Moser-type calculus inequalities (3.7) and (3.8) as

$$\begin{aligned} ||\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3-1}_1({\mathcal {A}}_3{{\textbf{V}}})||^2_{L^2(\Gamma _t)}+||\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\partial ^{\alpha _3-1}_1{\mathcal {F}}||^2_{L^2(\Gamma _t)}\le C(K){\mathcal {M}}(t). \end{aligned}$$
(6.41)

Summarizing (6.34)–(6.37) and using (6.40) and (6.41), taking \(\varepsilon \) sufficiently small, we get the estimate (6.33). Therefore, Lemma 6.5 is proven. \(\square \)

6.4 Estimate of non-weighted tangential derivatives

Now, we are going to obtain estimates of non-weighted tangential derivatives, i.e. \(\alpha _1=\alpha _3=0\), that is \(D^\alpha _*=\partial ^{\alpha _0}_t\partial ^{\alpha _2}_2\) with \(|\alpha |\le s\). This is the most important case because we shall use the boundary conditions. This gives the loss of two additional derivatives that imply that in final tame estimate we will have the “\(s+4, *, t\)" loss of derivatives from the coefficients, see Theorem 6.1. This loss is caused by the presence of zero order terms in \(\varphi \) (see (4.16)).

Lemma 6.6

The following estimate holds for (4.31) for all \(t\le T:\)

$$\begin{aligned} \begin{aligned} \sum _{|\alpha |\le s,\,\alpha _1=\alpha _3=0}||D^{\alpha }_{*}&{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\le \varepsilon C(K)\Big (|||{{\textbf{V}}}(t)|||^2_{s,*}+|||\varphi (t)|||^2_{H^{s-1}({{\mathbb {R}}})}\Big ) \\&+\frac{C(K)}{\varepsilon }\Big ({\mathcal {M}}(t)+||\varphi ||^2_{H^{s-1}(\Gamma _t)}+||\varphi ||^2_{W^{2,\infty }_{*}(\Gamma _t)}||{\hat{W}}||^2_{s+4,*,t}\Big ), \end{aligned}\nonumber \\ \end{aligned}$$
(6.42)

where s is positive integer, \(\varepsilon \) is a positive constant and \({\mathcal {M}}(t)\) is defined in (6.6).

Proof

We only need to estimate the highest-order tangential derivatives with \(|\alpha |=s, \) since the lower order terms can be controlled by definition of the anisotropic Sobolev norm, through the following estimate:

$$\begin{aligned} \sum _{|\alpha |\le s-1}||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\le C||{{\textbf{V}}}||^2_{s,*,t}. \end{aligned}$$
(6.43)

Therefore, applying same argument as in Lemmata 6.4 and 6.5, we obtain that

$$\begin{aligned} ||D^{\alpha }_{*}{{\textbf{V}}}(t)||^2_{L^2({\mathbb {R}}^2_+)}\le C(K)({\mathcal {M}}(t)+{\mathcal {J}}(t)) \end{aligned}$$
(6.44)

for \(\alpha =(\alpha _0,\alpha _2),\) with \(|\alpha |=s,\) where

$$\begin{aligned} {\mathcal {J}}(t)=\Big |\int _{\Gamma _t}({\mathcal {B}}_1D^{\alpha }_{*}{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}})|_{x_1=0}\textrm{d}x_2\textrm{d}\tau \Big |. \end{aligned}$$

Taking into account the boundary conditions (4.31) and (4.33), and also the important dissipative structure (5.8), the explicit quadratic boundary term in the integral can be rewritten in a suitable form: let us denote

$$\begin{aligned} c^{\pm }=\sum _{|\alpha '|+|\alpha ''|=s,|\alpha '|\ge 1}D^{\alpha '}_{*}{\hat{\lambda }}^{\pm }D^{\alpha ''}_{*}{\dot{H}}^{\pm }_n,\quad {\dot{w}}^{\pm }={\dot{u}}^{\pm }_n-{\hat{\lambda }}^{\pm }{\dot{H}}^{\pm }_n, \end{aligned}$$

where

$$\begin{aligned}{}[\partial _1{\hat{q}}]=(\partial _1{\hat{q}}^++\partial _1{\hat{q}}^-)|_{x_1=0},\,\,{\dot{w}}^{\pm }|_{x_1=0}=({\dot{u}}^{\pm }_N-{\hat{\lambda }}^{\pm }{\dot{H}}^{\pm }_N)|_{x_1=0} \end{aligned}$$

and \({{\hat{\lambda }}}^\pm \) were defined in (5.10); see also Lemma 5.1.

The boundary quadratic form becomes

$$\begin{aligned} \begin{aligned} ({\mathcal {B}}_1D^{\alpha }_{*}{{\textbf{V}}}\cdot D^{\alpha }_{*}{{\textbf{V}}})|_{x_1=0}&=2[(D^{\alpha }_{*}{\dot{u}}_N-{\hat{\lambda }} D^{\alpha }_{*}{\dot{H}}_N)D^{\alpha }_{*}{\dot{q}}]=2[cD^{\alpha }_{*}{\dot{q}}]+2[D^{\alpha }_{*}{\dot{w}}D^{\alpha }_{*}{\dot{q}}]\\&=2[cD^{\alpha }_{*}{\dot{q}}]+2D^{\alpha }_{*}{\dot{w}}^-|_{x_1=0}[D^{\alpha }_{*}{\dot{q}}]+2D^{\alpha }_{*}{\dot{q}}^+|_{x_1=0}[D^{\alpha }_{*}{\dot{w}}]\\&=2D^{\alpha }_{*}{\dot{q}}^+|_{x_1=0}D^{\alpha }_{*}(\partial _2\varphi [{\hat{u}}_2-{\hat{\lambda }}{\hat{H}}_2])+\text {l.o.t}, \end{aligned} \nonumber \\ \end{aligned}$$
(6.45)

where lower order term l.o.t can be expressed by

$$\begin{aligned} 2[cD^{\alpha }_{*}{\dot{q}}]-2D^{\alpha }_{*}{\dot{w}}^-|_{x_1=0}D^{\alpha }_{*}([\partial _1{\hat{q}}]\varphi )-2D^{\alpha }_{*}{\dot{q}}^+|_{x_1=0}D^{\alpha }_{*}([\partial _1{\hat{u}}_N-{\hat{\lambda }} \partial _1{\hat{H}}_N]\varphi ). \end{aligned}$$

Since the boundary conditions are dissipative, it is noted that first term on the right-hand side of (6.45) vanish by Lemma 5.1. The boundary terms can be estimated separately,as

$$\begin{aligned} {\mathcal {J}}(t)\le \sum _{\pm }\sum ^4_{i=1}{\mathcal {J}}^{\pm }_i(t), \end{aligned}$$
(6.46)

where

$$\begin{aligned} {\mathcal {J}}^{\pm }_1(t)= & {} \Big |\int _{\Gamma _t}(c^{\pm }D^{\alpha }_{*}{\dot{q}}^{\pm })|_{x_1=0}d{x_2}\textrm{d}\tau \Big |, \\ {\mathcal {J}}^{\pm }_2(t)= & {} \Big |\int _{\Gamma _t}(D^{\alpha }_{*}{\dot{w}}^-D^{\alpha }_{*}(\varphi \partial _1{\hat{q}}^{\pm }))|_{x_1=0}d{x_2}\textrm{d}\tau \Big |, \\ {\mathcal {J}}^{\pm }_3(t)= & {} \Big |\int _{\Gamma _t}(D^{\alpha }_{*}{\dot{q}}^+D^{\alpha }_{*}(\varphi \partial _1{\hat{u}}^{\pm }_n))|_{x_1=0}d{x_2}\textrm{d}\tau \Big |, \\ {\mathcal {J}}^{\pm }_4(t)= & {} \Big |\int _{\Gamma _t}(D^{\alpha }_{*}{\dot{q}}^+D^{\alpha }_{*}(\varphi {\hat{\lambda }}^{\pm }\partial _1{\hat{H}}^{\pm }_n))|_{x_1=0}d{x_2}\textrm{d}\tau \Big |. \end{aligned}$$

Recall that \(|\alpha |=s\ge 2.\) We denote \(D^{\alpha }_{*}=\partial _lD^{\gamma }\), \(D^\gamma :=\partial ^{\gamma _0}_t\partial ^{\gamma _2}_2\) for \(\gamma =(\gamma _0,\gamma _2),|\gamma |=s-1\ge 1\), where

$$\begin{aligned} l= {\left\{ \begin{array}{ll} 2, &{} \text { if } \alpha _0\ne s,\\ 0, &{} \text { if } \alpha _0=s. \end{array}\right. } \end{aligned}$$

In the ensuring estimate of \({\mathcal {J}}^{+}_1(t)\), we separate the analysis into two cases.

Case A: When \(\alpha _0=s\), then \(l=0, D^{\alpha }_{*}=\partial _tD^{\gamma }.\) Using the normal derivative estimates (6.5), (6.12) for non-characteristic variables, the elementary inequality (6.13) and integration by parts, we obtain that

$$\begin{aligned} \begin{aligned} {\mathcal {J}}^{+}_1(t)&=\Big |\int _{\Omega _t}(\partial _tc^{+}\partial _1D^{\gamma }{\dot{q}}^{+}-\partial _1c^{+}D^{\alpha }_{*}{\dot{q}}^{+})\textrm{d}{\textbf{x}}\textrm{d}\tau -\int _{{\mathbb {R}}^2_+}c^{+}\partial _1D^{\gamma }{\dot{q}}^{+}\textrm{d}{\textbf{x}}\Big |\\&\le C\Big (||{\dot{q}}^{+}||^2_{s,*,t}+||\partial _1{\dot{q}}^{+}||^2_{s-1,*,t}+||\partial _tc^{+}||^2_{L^2(\Omega _t)}\\&\quad +||\partial _1c^{+}||^2_{L^2(\Omega _t)}+\frac{1}{\varepsilon }||c^{+}(t)||^2_{L^2({\mathbb {R}}^2_+)}+\varepsilon ||\partial _1{\dot{q}}^{+}||^2_{H^{s-1}_{*}({\mathbb {R}}^2_+)} \Big )\\&\le C\Big (||{\dot{q}}^{+}||^2_{s,*,t}+||\partial _1{\dot{q}}^{+}||^2_{s-1,*,t}+||\partial _tc^{+}||^2_{L^2(\Omega _t)}\\&\quad +||\partial _1c^{+}||^2_{L^2(\Omega _t)}+\frac{1}{\varepsilon }||c^{+}(t)||^2_{L^2({\mathbb {R}}^2_+)}+\varepsilon |||\partial _1{\dot{q}}^{+}(t)|||^2_{s-1,*} \Big )\\&\le \varepsilon C(K)|||{{\textbf{V}}}(t)|||^2_{s,*}+\frac{C(K)}{\varepsilon }{\mathcal {M}}(t). \end{aligned}\nonumber \\ \end{aligned}$$
(6.47)

Here, \(\varepsilon \) is an arbitrary fixed constant. Similar argument also holds for \({\mathcal {J}}^-_1(t).\) Therefore, we obtain that

$$\begin{aligned}{\mathcal {J}}^{+}_1(t)+{\mathcal {J}}^{-}_1(t)\le \varepsilon C(K)|||{{\textbf{V}}}(t)|||^2_{s,*}+\frac{C(K)}{\varepsilon }{\mathcal {M}}(t).\end{aligned}$$

Case B: When \(\alpha _0\ne s\), then \(l=2, D^{\alpha }_{*}=\partial _2D^{\gamma }.\) Using integration by parts, we could deduce that

$$\begin{aligned} |{\mathcal {J}}^{\pm }_1(t)|\le C(K){\mathcal {M}}(t). \end{aligned}$$
(6.48)

Next, we estimate \({\mathcal {J}}^+_3\) and separate this term into two parts,

$$\begin{aligned} {\mathcal {J}}^+_3\le {\mathcal {J}}_0(t)+\Sigma _4(t), \end{aligned}$$

where

$$\begin{aligned} {\mathcal {J}}_0(t)= & {} \Big |\int _{\Gamma _t}(\varphi D^{\alpha }_{*}{\dot{q}}^+D^{\alpha }_{*}\partial _1{\hat{u}}^+_n)|_{x_1=0}dx_2\textrm{d}\tau \Big |, \\ \Sigma _4(t)= & {} \sum _{|\alpha '|+|\alpha ''|=s,|\alpha '|\ge 1}\Big |\int _{\Gamma _t}(D^{\alpha }_{*}{\dot{q}}^+D^{\alpha '}_{*}\varphi D^{\alpha ''}_{*}\partial _1{\hat{u}}^+_n)\Big |_{x_1=0}dx_2\textrm{d}\tau \Big |. \end{aligned}$$

For the term \({\mathcal {J}}_0(t),\) after integrating by parts, we obtain that

$$\begin{aligned} \begin{aligned} {\mathcal {J}}_0(t)&=\Big |\int _{\Omega _t}(\partial _1\partial _lD^{\alpha }_{*}{\hat{u}}^+_n)\varphi \partial _1D^{\gamma }{\dot{q}}^+-(\partial ^2_1D^{\alpha }_{*}{\hat{u}}^+_n)\varphi D^{\alpha }_{*}{\dot{q}}^+\\&\quad +\partial _1D^{\alpha }_{*}{\hat{u}}^+_n\partial _l\varphi \partial _1D^{\gamma }{\dot{q}}^+\textrm{d}{\textbf{x}}\textrm{d}\tau -\frac{l-2}{2}\int _{{\mathbb {R}}^2_+}(\partial _1D^{\alpha }_{*}{\hat{u}}^+_{n})\varphi \partial _1D^{\gamma }{\dot{q}}^+\textrm{d}{\textbf{x}}\Big |. \end{aligned} \end{aligned}$$

We note that to estimate \(\partial ^2_1D^{\alpha }_{*}{\hat{u}}^+_n\), we need \({\hat{W}}\) with regularity \(s+4.\) For the term \({\mathcal {J}}_0(t),\) using Moser-type calculus inequalities (3.7), (3.8), (6.5) and (6.12) we get

$$\begin{aligned} {\mathcal {J}}_0(t)\le \varepsilon C(K)|||{{\textbf{V}}}(t)|||^2_{s,*}+\frac{C(K)}{\varepsilon }\Big ({\mathcal {M}}(t)+||\varphi ||^2_{W^{1,\infty }{(\Gamma _t)}}||{\hat{W}}||^2_{s+4,*,t}\Big ). \end{aligned}$$

Now, we start to estimate \(\Sigma _4.\) It is noted that for \(|\alpha '|\ge 1,\) we can isolate one tangential derivative in the differential operator

$$\begin{aligned} D^{\alpha '}_{*}\varphi =D^{\gamma '}(\partial _j\varphi ),\quad |\gamma '|\le s-1,\quad j=0 \text { or } j=2. \end{aligned}$$

Using (A.54) and (A.42) we obtain that

$$\begin{aligned} \nabla _{t,x_2}\varphi ={\textbf{H}}(\hat{{{\textbf{U}}}},{\hat{\varphi }}){{\textbf{V}}}_n|_{x_1=0}+{\textbf{G}}(\hat{{{\textbf{U}}}},{\hat{\varphi }})\varphi . \end{aligned}$$
(6.49)

Here \({\textbf{H}}(\hat{{{\textbf{U}}}},{\hat{\varphi }}),{\textbf{G}}(\hat{{{\textbf{U}}}},{\hat{\varphi }})\) depend on \(\hat{{{\textbf{U}}}}|_{x_1=0},\partial _1\hat{{{\textbf{U}}}}|_{x_1=0}\) and second order derivatives of \({\hat{\varphi }}.\)

Using (6.49), we can write the derivatives of \(\varphi \) by using the non-characteristic unknown \({{\textbf{V}}}_n\). We insert these derivatives \(D^{\alpha '}_{*}\varphi =D^{\gamma '}(\cdots )\), with \(|\gamma '|\le s-1\), into \(\Sigma _4\) to obtain

$$\begin{aligned} \Sigma _4(t)=\sum _{|\alpha '|+|\alpha ''|=s,|\alpha '|\ge 1}\Big |\int _{\Gamma _t}(D^{\alpha }_{*}{\dot{q}}^+D^{\gamma '}({\textbf{H}}{{\textbf{V}}}_n+{\textbf{G}}\varphi )D^{\alpha ''}_{*}\partial _1{\hat{u}}^+_n)\Big |_{x_1=0}dx_2\textrm{d}\tau \Big |, \end{aligned}$$

which can be controlled by the right-hand side of (6.42). Indeed, by passing to the volume integral on \(\Omega _t\), we note that the highest order of regularity for \(D^{\alpha ''}_{*}\partial _1^2{\hat{u}}^+_n\) is \(s+3\) because \(|\gamma '|+|\alpha ''|=s-1\) (since \(|\alpha '|= |\gamma '|+1\) and \(|\alpha '|+|\alpha ''|=s\)). Similar estimates also hold for \({\mathcal {J}}^-_3,{\mathcal {J}}^{\pm }_2,{\mathcal {J}}^{\pm }_4.\) Using (6.43), (6.44), (6.46), (6.47), (6.48), we obtain the estimate (6.42). Therefore, Lemma 6.6 is concluded. \(\square \)

6.5 Estimate of front

Now, we are going to estimate the front \(\varphi (t)\) in \(H^{s-1}({{\mathbb {R}}})\) and its tangential derivatives \(\nabla _{t,x_2}\varphi \) in \(H^{s-1}(\Gamma _t).\)

Lemma 6.7

Given the solution \(\varphi \) of (4.31), for all \(t\le T\) and positive integer s,  the following estimate holds:

$$\begin{aligned} \begin{aligned}&|||\varphi (t)|||^2_{H^{s-1}({{\mathbb {R}}})}+||\nabla _{t,x_2}\varphi ||^2_{H^{s-1}(\Gamma _t)} \\&\quad \le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\varphi ||^2_{H^{s-1}(\Gamma _t)}+(||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}+||\varphi ||^2_{W^{1,\infty }_{*}(\Gamma _t)})||{\hat{W}}||^2_{s+2,*,t}\Big ). \end{aligned} \nonumber \\ \end{aligned}$$
(6.50)

Proof

Applying tangential derivatives on the first boundary conditions in (4.31), we obtain that

$$\begin{aligned} \begin{aligned} |||\varphi (t)|||^2_{H^{s-1}({{\mathbb {R}}})}&\le C(K)\Big (||{\dot{u}}_n|_{x_1=0}||^2_{H^{s-1}(\Gamma _t)}+||\varphi ||^2_{H^{s-1}(\Gamma _t)}+\sum _{|\alpha |\le s-1}||{\mathcal {G}}_{\alpha }||^2_{L^2(\Gamma _t)}\Big ), \end{aligned}\nonumber \\ \end{aligned}$$
(6.51)

where

$$\begin{aligned} {\mathcal {G}}_{\alpha }=D^{\alpha }_{*}(\varphi \partial _1{\hat{u}}^+_N)-([D^{\alpha }_{*},{\hat{u}}^+_2]\partial _2\varphi +\frac{1}{2}\partial _2{\hat{u}}^+_2D^{\alpha }_{*}\varphi ). \end{aligned}$$

For the first term on the right-hand side of (6.51), using trace Theorem in Lemma 3.4, we obtain that

$$\begin{aligned} ||{\dot{u}}_n|_{x_1=0}||^2_{H^{s-1}(\Gamma _t)}\lesssim ||{{\textbf{V}}}||^2_{s,*,t}. \end{aligned}$$
(6.52)

For the third term on the right-hand side of (6.51), when \(|\alpha |\le s-1,\) we obtain that

$$\begin{aligned} ||{\mathcal {G}}_{\alpha }||^2_{L^2(\Gamma _t)}{} & {} \lesssim ||\varphi \partial _1{\hat{u}}^+_N||^2_{H^{s-1}(\Gamma _t)}+||[D^{\alpha }_{*},{\hat{u}}^+_2]\partial _2\varphi ||^2_{L^2(\Gamma _t)}+||\partial _2{\hat{u}}^+_2D^{\alpha }_{*}\varphi ||^2_{L^2(\Gamma _t)} \nonumber \\{} & {} \le C(K)\Big (||\varphi ||^2_{H^{s-1}(\Gamma _t)}+||\varphi ||^2_{W^{1,\infty }(\Gamma _t)}||{\hat{W}}||^2_{s+2,*,t}\Big ). \end{aligned}$$
(6.53)

Hence, we obtain the estimate of the first term in the left-hand side of (6.50).

For the estimate of the second term in the left-hand side of (6.50), we use the relation (6.49), \(||{\hat{\varphi }}||_{H^s(\Gamma _t)}\lesssim ||{\hat{\Psi }}||_{s,*,t}\le ||{\hat{\varphi }}||_{H^s(\Gamma _t)},\) and Moser-type calculus inequalities (3.7) and (3.8) to obtain that

$$\begin{aligned} \begin{aligned}&||\nabla _{t,x_2}\varphi ||^2_{H^{s-1}(\Gamma _t)}\le ||{\textbf{H}}(\hat{{\textbf{U}}},{\hat{\varphi }}){{\textbf{V}}}_n|_{x_1=0}||^2_{H^{s-1}(\Gamma _t)}+||{\textbf{G}}(\hat{{\textbf{U}}},{\hat{\varphi }})\varphi ||^2_{H^{s-1}(\Gamma _t)}\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,t}+||\varphi ||^2_{H^{s-1}(\Gamma _t)}+(||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _t)}+||\varphi ||^2_{W^{1,\infty }(\Gamma _t)})||{\hat{W}}||^2_{s+2,*,t}\Big ). \end{aligned} \end{aligned}$$

Therefore, Lemma 6.7 is concluded. \(\square \)

Collecting all the previously established higher order estimates, we can prove the following lemma:

Lemma 6.8

The solution of the homogeneous problem (4.31) satisfies the following a priori estimate

$$\begin{aligned} ||\dot{{{\textbf{U}}}}||^2_{s,*,T}+||\varphi ||^2_{H^s(\Gamma _T)}\le C(K)Te^{C(K)T}{\mathcal {N}}(T) \end{aligned}$$
(6.54)

for positive integer s,  where

$$\begin{aligned} \begin{aligned} {\mathcal {N}}(T)&=||{\textbf{F}}||^2_{s,*,T}+(||\dot{{{\textbf{U}}}}||^2_{W^{2,\infty }_{*}(\Omega _T)}+||\varphi ||^2_{W^{2,\infty }(\Gamma _T)}+||{\textbf{F}}||^2_{W^{1,\infty }_{*}(\Omega _T)})||{\hat{W}}||^2_{s+4,*,T}. \end{aligned} \end{aligned}$$

Proof

Combining the estimates (6.21), (6.33), (6.42) choosing \(\varepsilon \) small enough and (6.50), we obtain that

$$\begin{aligned} {\mathcal {I}}(t)\le C(K)\Big ({\mathcal {N}}(T)+\int ^t_0{\mathcal {I}}(\tau )\textrm{d}\tau \Big ), \end{aligned}$$
(6.55)

where

$$\begin{aligned} {\mathcal {I}}(t)=|||{{\textbf{V}}}(t)|||^2_{s,*}+|||\varphi (t)|||^2_{H^{s-1}({{\mathbb {R}}})}. \end{aligned}$$

Notice that

$$\begin{aligned} {\mathcal {I}}(0)=0. \end{aligned}$$

Then, using Grönwall’s lemma to (6.55), we obtain

$$\begin{aligned} {\mathcal {I}}(t)\le C(K)e^{C(K)T}{\mathcal {N}}(T), \text { for } 0\le t\le T. \end{aligned}$$

Integrating (6.55) with respect to \(t\in (-\infty ,T],\) we get

$$\begin{aligned} ||{{\textbf{V}}}||^2_{s,*,T}+||\varphi ||^2_{H^{s-1}(\Gamma _T)}\le C(K)Te^{C(K)T}{\mathcal {N}}(T). \end{aligned}$$
(6.56)

Notice that \(\dot{{{\textbf{U}}}}=J{{\textbf{V}}}.\) Then, using the decomposition \(J=J({\hat{W}})=I+J_0({\hat{W}}),\) where \(J_0\) satisfies \(J_0(0)=0,\) we apply the Moser-type calculus inequalities (3.7) and (3.8) and derive

$$\begin{aligned} \begin{aligned} ||\dot{{{\textbf{U}}}}||^2_{s,*,T}&=||{{\textbf{V}}}+J_0{{\textbf{V}}}||^2_{s,*,T}\\&\le C(K)\Big (||{{\textbf{V}}}||^2_{s,*,T}+||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_ {*}(\Omega _T)}||{\hat{W}}||^2_{s,*,T}\Big )\\&\le C(K)||{{\textbf{V}}}||^2_{s,*,T}+TC(K)||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _T)}||{\hat{W}}||^2_{s+1,*,T}, \end{aligned} \end{aligned}$$
(6.57)

and, using (6.50) with \(t=T\),

$$\begin{aligned} \begin{aligned}&||\nabla _{t,x_2}\varphi ||^2_{H^{s-1}(\Gamma _T)}\\&\quad \lesssim \Big (||{{\textbf{V}}}||^2_{s,*,T}+||\varphi ||^2_{H^{s-1}(\Gamma _T)}+(||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _T)}+||\varphi ||^2_{W^{1,\infty }(\Gamma _T)})||{\hat{W}}||^2_{s+2,*,T}\Big )\\&\quad \lesssim \Big (||{{\textbf{V}}}||^2_{s,*,T}+||\varphi ||^2_{H^{s-1}(\Gamma _T)}+TC(K)(||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _T)}+||\varphi ||^2_{W^{1,\infty }(\Gamma _T)})||{\hat{W}}||^2_{s+3,*,T}\Big ). \end{aligned} \nonumber \\ \end{aligned}$$
(6.58)

In the proof of last inequalities in (6.57), (6.58), we have used the following relation by applying Sobolev imbedding:

$$\begin{aligned} \begin{aligned} ||{\hat{W}}||^2_{s,*,T}&=\sum ^s_{j=0}\int ^T_0||\partial ^j_t{\hat{W}}(t)||^2_{s-j,*}dt \le \sum ^s_{j=0}T\max _{t\in [0,T]}||\partial ^j_t{\hat{W}}(t)||^2_{s-j,*}\\&\lesssim T\sum ^s_{j=0}\Big \{\int ^T_0||\partial ^j_t{\hat{W}}(t)||^2_{s-j,*}dt+\int ^T_0||\partial ^{j+1}_t{\hat{W}}(t)||^2_{s-j,*}dt\Big \}\\&\lesssim T\int ^T_0|||\partial ^j_t{\hat{W}}(t)|||^2_{s+1-j,*}dt \lesssim T||{\hat{W}}||^2_{s+1,*,T}. \end{aligned} \end{aligned}$$

Using (6.56) and (6.57), we obtain

$$\begin{aligned} ||\dot{{{\textbf{U}}}}||^2_{s,*,T}+||\varphi ||^2_{H^{s-1}(\Gamma _T)}\le C(K)Te^{C(K)T}{\mathcal {N}}(T). \end{aligned}$$
(6.59)

Similar to (6.57), we can get

$$\begin{aligned} ||{{\textbf{V}}}||^2_{s,*,T}\le C(K)||\dot{{{\textbf{U}}}}||^2_{s,*,T}+TC(K)||\dot{{{\textbf{U}}}}||^2_{W^{1,\infty }_{*}(\Omega _T)}||{\hat{W}}||^2_{s+1,*,T}. \end{aligned}$$
(6.60)

Adding (6.58) and (6.59), and using (6.60), we conclude Lemma 6.8. \(\square \)

Using (6.54), we are ready to prove the tame estimate for the homogeneous problem (4.31).

Proof of Theorem 6.1

Using Lemma 6.8, the Sobolev inequalities (3.9), (3.10) for \(s\ge 6,\) we obtain

$$\begin{aligned} \begin{aligned} ||\dot{{{\textbf{U}}}}||_{s,*,T}+||\varphi ||_{H^s(\Gamma _T)}&\le C(K)T^{\frac{1}{2}}e^{C(K)T}\Big (||{\textbf{F}}||_{s,*,T}+(||\dot{{{\textbf{U}}}}||_{6,*,T}\\&\quad +||\varphi ||_{H^6(\Gamma _T)}+||{\textbf{F}}||_{4,*,T})||{\hat{W}}||_{s+4,*,T}\Big ). \end{aligned} \end{aligned}$$
(6.61)

Taking T sufficiently small and \(s=6,\) using (6.1), we obtain that

$$\begin{aligned} ||\dot{{{\textbf{U}}}}||_{6,*,T}+||\varphi ||_{H^6(\Gamma _T)}\le C(K_0)||{\textbf{F}}||_{6,*,T}. \end{aligned}$$
(6.62)

Hence, (6.61) and (6.62) implies (6.2). The existence and uniqueness of the solution comes from Theorem 5.1. The proof of Theorem 6.1 is complete.

7 Higher order energy estimate for problem (4.20)

Now, we are ready to obtain an a priori tame estimate in \(H^s_{*}\) for the nonhomogeneous problem (4.20).

Theorem 7.1

Let \(T>0\) and s be an integer, \(s\ge 6.\) Assume that the basic state \((\hat{{{\textbf{U}}}},{\hat{\varphi }})\) satisfies (4.2)–(4.10), and \((\breve{{\textbf{U}}}^{\pm },{\hat{\varphi }})\in H^{s+4}_{*}(\Omega _T)\times H^{s+5}(\Gamma _T)\) satisfies (6.1). Assume that \({\textbf{f}}\in H^{s+2}_{*}(\Omega _T)\), \({\textbf{g}}\in H^{s+2}(\Gamma _T)\) vanish in the past. Then, there exists a positive constant \(K_0,\) that does not depend on s and T,  and there exists a constant \(C(K_0)>0\) such that if \({\hat{K}}\le K_0\), then there exists a unique solution \((\dot{{{\textbf{U}}}},\varphi )\in H^{s}_{*}(\Omega _T)\times H^s({\Gamma _T})\) to the problem (4.20) that allows the tame estimate

$$\begin{aligned} \begin{aligned}&||\dot{{{\textbf{U}}}}||_{s,*,T}+||\varphi ||_{H^s(\Gamma _T)}\\&\quad \le C(K_0)\Big (||{\textbf{f}}||_{s+2,*,T}+||{\textbf{g}}||_{H^{s+2}(\Gamma _T)}+(||{\textbf{f}}||_{8,*,T}+||{\textbf{g}}||_{H^{8}(\Gamma _T)})||{\hat{W}}||_{s+4,*,T}\Big ), \end{aligned} \nonumber \\ \end{aligned}$$
(7.1)

for T small enough, where \({\hat{W}}=(\breve{{\textbf{U}}},\nabla _{t, \textbf{x}}{\hat{\Psi }}).\)

Remark 7.1

The lower regularity in (6.1) and low norms in (6.2) and (7.1), for both the even and odd case, differ from Trakhinin [39], see Theorem 3 and Theorem 4, due to finer Sobolev imbeddings (3.9), (3.10).

Proof

Using the Moser-type calculus inequalities (3.7) and (3.8), we obtain a refined version of estimate (4.27) in tame form

$$\begin{aligned} ||{\mathbb {L}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})\tilde{{{\textbf{U}}}}||_{s,*,T}\le C(K)\Big (||\tilde{{{\textbf{U}}}}||_{s+2,*,T}+||\tilde{{{\textbf{U}}}}||_{W^{2,\infty }_{*}(\Omega _T)}||\hat{{\textbf{U}}},{\hat{\Psi }}||_{s+2,*,T}\Big ). \end{aligned}$$

Then, using Sobolev embedding inequalities (3.10), we get

$$\begin{aligned} ||{\mathbb {L}}'_e(\hat{{{\textbf{U}}}},{\hat{\Psi }})\tilde{{{\textbf{U}}}}||_{s,*,T}\le C(K)\Big (||\tilde{{{\textbf{U}}}}||_{s+2,*,T}+||\tilde{{{\textbf{U}}}}||_{6,*,T}||\hat{{\textbf{U}}},{\hat{\Psi }}||_{s+2,*,T}\Big ). \end{aligned}$$

Using the above estimate, (4.30) and recalling the definition of \(\tilde{{\textbf{U}}}\), see Section 4.3, it holds that

$$\begin{aligned}{} & {} ||{\textbf{F}}||_{s,*,T} \nonumber \\{} & {} \quad \le C(K)\Big (||{\textbf{f}}||_{s+2,*,T}+||{\textbf{g}}||_{H^{s+2}(\Gamma _T)}+(||{\textbf{f}}||_{8,*,T}+||{\textbf{g}}||_{H^8(\Gamma _T)})||\hat{{\textbf{U}}},{\hat{\Psi }}||_{s+2,*,T}\Big ).\nonumber \\ \end{aligned}$$
(7.2)

Using the assumption (6.1) and (7.2) with \(s=6\), we get

$$\begin{aligned} ||{\textbf{F}}||_{6,*,T}\le C(K_0)\Big ({||{\textbf{f}}||_{8,*,T}+||{\textbf{g}}||_{H^8(\Gamma _T)}}\Big ). \end{aligned}$$
(7.3)

Combining the estimates (6.61), (7.2) and (7.3), we obtain the tame estimate (7.1). \(\square \)

8 Construction of Approximate Solutions

Suppose the initial data

$$\begin{aligned} ({{\textbf{U}}}^{\pm }_0,\varphi _0)=(\bar{{\textbf{U}}}^{\pm }+\tilde{{\textbf{U}}}^{\pm }_0,\varphi _0)=(p^{\pm }_0, u^{\pm }_{1,0}, u^{\pm }_{2,0}, H^{\pm }_{1,0},H^{\pm }_{2,0}, S^{\pm }_0, \varphi _0) \end{aligned}$$

satisfy the stability condition (4.4) and restriction (2.5) at \(x_1=0\) for \(x_2\in {{\mathbb {R}}}\). Since \({H}^+_{2,0}\ne 0\) or \({H}^-_{2,0}\ne 0\) at \({x_1=0}\), see also Remark 4.3, from (2.5) we can solve \(\partial _2\varphi \) as follows (we drop the sub-index 0 for simplicity):

$$\begin{aligned} \partial _2\varphi =\mu ({{{\textbf {U}}}})|_{x_1=0}=\frac{H^+_1H^+_2+H^-_1H^-_2}{(H^{+}_2)^2+(H^-_2)^2}\Big |_{x_1=0}, \end{aligned}$$
(8.1)

where \({{\textbf{U}}}:=({{\textbf{U}}}^+,{{\textbf{U}}}^-).\) Then, using the boundary condition (4.3), we have

$$\begin{aligned} \partial _t\varphi =\eta ({{\textbf{U}}})|_{x_1=0}, \end{aligned}$$
(8.2)

with

$$\begin{aligned} \eta ({{\textbf{U}}})=u^+_1-u^+_2\mu ({{\textbf{U}}}). \end{aligned}$$

By using the hyperbolicity condition (1.6), we can write the system in (2.1) as

$$\begin{aligned} \partial _t{{\textbf{U}}}=-(A_0({{\textbf{U}}}))^{-1}\Big ({\tilde{A}}_1({{\textbf{U}}},\Psi )\partial _1{{\textbf{U}}}+A_2({{\textbf{U}}})\partial _2{{\textbf{U}}}\Big ), \end{aligned}$$
(8.3)

where \(\Psi :=(\Psi ^+,\Psi ^-)\), and the matrices \(A_0,A_2,{\tilde{A}}_1\) are defined by (1.5) and (2.3). The traces

$$\begin{aligned} {{\textbf{U}}}_j=(p^+_j,u^+_{1,j},u^+_{2,j},H^+_{1,j},H^+_{2,j},S^+_j,p^-_j,u^-_{1,j},u^-_{2,j},H^-_{1,j},H^-_{2,j},S^-_j)=\partial ^j_t{{\textbf{U}}}|_{t=0} \end{aligned}$$

and

$$\begin{aligned} \varphi _j=\partial ^j_t\varphi |_{t=0},\quad j\ge 1 \end{aligned}$$

can be defined step by step by applying operator \(\partial ^{j-1}_t\) to (8.2) and (8.3), for \(j\ge 1\) and evaluating \(\partial ^j_t{{\textbf{U}}}\) and \(\partial ^j_t\varphi \) at \(t=0\) in terms of the initial data. Notice that

$$\begin{aligned} \Psi ^{\pm }_j=\partial ^j_t\Psi ^{\pm }|_{t=0}=\chi (\pm x_1)\varphi _j. \end{aligned}$$

Define the zero-th order compatibility condition as

$$\begin{aligned}{}[u_{1,0}]-[u_{2,0}]\partial _2\varphi _0=0,\quad [p_0+\frac{|H_0|^2}{2}]=0. \end{aligned}$$
(8.4)

Taking (8.1), (8.2) evaluated at \(t=0,\) and using (8.4), we obtain that

$$\begin{aligned} \varphi _1=u^{\pm }_{1,0}-u^{\pm }_{2,0}\partial _2\varphi _0\Big |_{x_1=0}. \end{aligned}$$
(8.5)

Denote \((H^{\pm }_N)_j=\partial ^j_tH^{\pm }_N\Big |_{t=0}.\) Using (8.3) and (8.5), taking \(t=0,\) we obtain that

$$\begin{aligned} (H^{\pm }_N)_1=-\Big (u^{\pm }_{2,0}\partial _2(H^{\pm }_N)_0+\partial _2u^{\pm }_{2,0}(H^{\pm }_N)_0\Big |_{x_1=0}\Big ). \end{aligned}$$

Therefore, \((H^{\pm }_N)_0|_{x_1=0}=0\) implies \((H^{\pm }_N)_1|_{x_1=0}=0.\) Once we have defined \({{\textbf{U}}}_1,\varphi _1,\) we can deduce \({{\textbf{U}}}_2,\varphi _2\) and so on. Moreover, at each step, we can prove that

$$\begin{aligned} (H^{\pm }_N)_j|_{x_1=0}=0,\quad j\ge 2, \end{aligned}$$

provided that \({{\textbf{U}}}_j\), \(\varphi _j\) satisfy the compatibility condition (see Definition 8.1).

The following Lemma 8.1 is necessary for the approximate solutions; we refer to [14] and [20, Lemma 4.2.1]. Differently from [39], we take the initial data in the standard Sobolev spaces.

Lemma 8.1

Let \(\mu \in {\mathbb {N}},\mu \ge 3,\) \(\tilde{{{\textbf{U}}}}_0:= {\textbf{U}}_0-\bar{{\textbf{U}}}\in H^{\mu +1.5}({\mathbb {R}}^2_+)\) and \(\varphi _0\in H^{\mu +1.5}({{\mathbb {R}}})\). Then, we can determine \(\tilde{{{\textbf{U}}}}_j\in H^{\mu +1.5-j}({\mathbb {R}}^2_+)\) and \(\varphi _j\in H^{\mu +1.5-j}({{\mathbb {R}}})\) by induction and set \({\textbf{U}}_j=\tilde{{\textbf{U}}}_j +\bar{{\textbf{U}}},\) for \(j=1,\cdots ,\mu \). In addition, we prove

$$\begin{aligned} \sum ^{\mu }_{j=0}(||\tilde{{{\textbf{U}}}}_j||_{H^{\mu +1.5-j}({\mathbb {R}}^2_+)}+||\varphi _j||_{H^{\mu +1.5-j}({{\mathbb {R}}})})\le C(M_0), \end{aligned}$$
(8.6)

where \(C>0\) depends only on \(\mu \), \(||\tilde{{{\textbf{U}}}}_0||_{W^{1,\infty }({\mathbb {R}}^2_+)}\) and \(||\varphi _0||_{W^{1,\infty }({{\mathbb {R}}})}\), and

$$\begin{aligned} M_0:=||\tilde{{{\textbf{U}}}}_0||_{{H^{\mu +1.5}}({\mathbb {R}}^2_+)}+||\varphi _0||_{H^{\mu +1.5}({{\mathbb {R}}})}. \end{aligned}$$
(8.7)

Definition 8.1

Let \(\mu \in {\mathbb {N}},\mu \ge 3\). The initial data \((\tilde{{{\textbf{U}}}}_0,\varphi _0)\in H^{\mu +1.5}({\mathbb {R}}^2_+)\times H^{\mu +1.5}({{\mathbb {R}}})\) are defined to be compatible up to order \(\mu \) if \((\tilde{{{\textbf{U}}}}_j,\varphi _j)\) satisfy (8.4) for \(j=0\) and

$$\begin{aligned}&{} \sum ^j_{l=0}([u_{1,j-l}]-[u_{2,j-l}]\partial _2\varphi _l)=0,\\{}&{} [p_j]+\sum ^{j-1}_{l=0}C_{l,j-1}[(H_l,H_{j-l})]=0,\quad \text { on }\,\,\{x_1=0\}, \end{aligned}$$

for \(j=1,\cdots ,\mu \), where \(C_{l,j-1}\) are suitable constants.

To use the tame estimate for the proof of convergence of the Nash–Moser iteration, we should reduce our nonlinear problem to that whose solution vanishes in the past. This is achieved by the construction of the so-called “approximate solution" that allows to “absorb" the initial data into the interior equation. The “approximate solution" is in the sense of Taylor’s series at \(t=0\).

Below, we will use the notation

$$\begin{aligned} {\mathbb {L}}({{\textbf{U}}},\Psi ):= \left[ \begin{array}{c} {\mathbb {L}}({{\textbf{U}}}^+,\Psi ^+)\\ {\mathbb {L}}({{\textbf{U}}}^-,\Psi ^-) \end{array}\right] .\quad \end{aligned}$$

Lemma 8.2

Let \(\mu \in {\mathbb {N}}, \mu \ge 3\) and let \(\delta >0\). Suppose the initial data \((\tilde{{{\textbf{U}}}}_0,\varphi _0)\in H^{\mu +1.5}({\mathbb {R}}^2_+)\times H^{\mu +1.5}({{\mathbb {R}}})\) are compatible up to order \(\mu \) and satisfy the assumptions (1.6), (2.4), (2.5), (4.4). Then, there exist \(T>0\) and \(({\tilde{{\textbf{U}}}}^a,\varphi ^a)\in H^{\mu +2}(\Omega _T)\times H^{\mu +2}(\Gamma _T)\) such that

$$\begin{aligned} \partial ^j_t{\mathbb {L}}({{\textbf{U}}}^a,\Psi ^a)|_{t=0}=0 \text { in } \Omega , \text { for } j\in \{0,\cdots , \mu -1\}, \end{aligned}$$
(8.8)

where

$$\begin{aligned} {\textbf{U}}^a:=\tilde{{\textbf{U}}}^a+\overline{{\textbf{U}}},\quad \Psi ^{a\,\pm }=\chi (\pm x_1)\varphi ^a. \end{aligned}$$

We call \(({\textbf{U}}^a, \varphi ^a)\) the approximate solution to problem (4.20). Moreover the approximate solution satisfies the estimate

$$\begin{aligned} ||\tilde{{{\textbf{U}}}}^a||_{H^{\mu +2}(\Omega _T)}+||\varphi ^a||_{H^{\mu +2}(\Gamma _T)}<\delta , \end{aligned}$$
(8.9)

the stability conditions (4.4) on \(\Gamma _T\), the hyperbolicity condition (4.2) on \(\Omega _T\).

Proof

Let us first denote \(\Phi ^{a\,\pm }=\pm x_1+\Psi ^{a\,\pm }\), \(\tilde{{\textbf{U}}}^a=(\tilde{{\textbf{U}}}^{a+},\tilde{{\textbf{U}}}^{a-})^T\), \({{\tilde{p}}}^a=({{\tilde{p}}}^{a+}, {\tilde{p}}^{a-})^T\), \({\tilde{u}}^a=({\tilde{u}}^{a+},{\tilde{u}}^{a-})^T\), \({\tilde{H}}^a=({\tilde{H}}^{a+},{\tilde{H}}^{a-})^T\), \({\tilde{S}}^a=({\tilde{S}}^{a+},{\tilde{S}}^{a-})^T\). Consider \(\tilde{{\textbf{U}}}^a\in H^{\mu +2}({{\mathbb {R}}}\times {\mathbb {R}}^2_+),\varphi ^a\in H^{\mu +2}({{\mathbb {R}}}^2)\), such that

$$\begin{aligned}{} & {} \partial ^j_t\tilde{{\textbf{U}}}^a|_{t=0}=\tilde{{\textbf{U}}}_j\in H^{\mu -j+2}({\mathbb {R}}^2_+), \text { for } j=0,\cdots ,\mu ,\\{} & {} \partial ^j_t\varphi ^a|_{t=0}=\varphi _j\in H^{\mu -j+2}({{\mathbb {R}}}),\quad \text { for } j=0,\cdots ,\mu , \end{aligned}$$

where \(\tilde{{\textbf{U}}}_j\) and \(\varphi _j\) are given by Lemma 8.1. Since \((\tilde{{\textbf{U}}}^a,\varphi ^a)\) satisfies the hyperbolicity condition (4.2) and the stability condition (4.4) at \(t=0\), by continuity \((\tilde{{\textbf{U}}}^a,\varphi ^a)\) satisfy (4.4) at \(x_1=0\) and (4.2) for small times. By multiplication of \((\tilde{{\textbf{U}}}^a,\varphi ^a)\) by a cut-off function in time supported on \([-T,T]\) we can assume that (4.2), (4.4) hold for all times (in this regard, recall Remark 3.1). Given any \(\delta >0\), by taking \(T>0\) sufficiently small, we can assume that \(\tilde{{\textbf{U}}}^a\), \(\varphi ^a\) are small in the sense of (8.9). \(\square \)

Remark 8.1

Let us remark that we do not require any constraint (that is interior equations or boundary condition) to be satisfied by the approximate solution constructed above. This allows us to the use of cut-off argument making the hyperbolicity condition (4.2) and the stability condition (4.4) to be satisfied globally in time, without any trouble.

Remark 8.2

In the sequel, in the proof of the main Theorem 3.1, estimate (8.9) will be used with \(\mu =m+10\), being m an integer as in the statement of that theorem.

We assume that

$$\begin{aligned} ||\varphi _0||_{L^{\infty }({{\mathbb {R}}})}<\frac{1}{2}, \end{aligned}$$

then we fix \(T>0\) sufficiently small so that \(||\varphi ^a||_{L^{\infty }([0,T]\times {{\mathbb {R}}})}\le \frac{1}{2}.\) Hence, we get

$$\begin{aligned} \partial _1\Phi ^{a+}\ge \frac{1}{2},\quad \partial _1\Phi ^{a-}\le -\frac{1}{2} \end{aligned}$$

(recall that \(||\chi ^\prime ||_{L^\infty ({{\mathbb {R}}})}\le 1/2\), see Section 2).

The approximate solution \(({{\textbf{U}}}^a,\varphi ^a)\) enables us to reformulate the original problem (2.1) as a nonlinear problem with zero initial data. Set

$$\begin{aligned} {\mathcal {F}}^a:= {\left\{ \begin{array}{ll} -{\mathbb {L}}({{\textbf{U}}}^a,\Psi ^a),\;&{}t>0,\\ ~~~~0,\;&{}t<0.\\ \end{array}\right. } \end{aligned}$$
(8.10)

From \(\tilde{{{\textbf{U}}}}^a\in H^{\mu +2}(\Omega _T)\) and \(\varphi ^a\in H^{\mu +2}(\Gamma _T),\) we have \({\mathcal {F}}^a\in H^{\mu +1}(\Omega _T)\).

Given the approximate solution \((\tilde{{{\textbf{U}}}}^a,\varphi ^a)\) of Lemma 8.2 and \({\mathcal {F}}^a\) defined in (8.10), we see that \(({{\textbf{U}}},\varphi )=({{\textbf{U}}}^a,\varphi ^a)+({{\textbf{V}}},\psi )\) is a solution of the original problem (2.1) if \({{\textbf{V}}}=({{\textbf{V}}}^+,{{\textbf{V}}}^-)^T,\Psi =(\Psi ^+,\Psi ^-)^T,\) \(\Psi |_{x_1=0}:=\psi \) satisfy the following problem:

$$\begin{aligned} {\left\{ \begin{array}{ll} {\mathcal {L}}({{\textbf{V}}},\Psi ):={\mathbb {L}}({{\textbf{U}}}^a+{{\textbf{V}}},\Psi ^a+\Psi )-{\mathbb {L}}({{\textbf{U}}}^a,\Psi ^a)={\mathcal {F}}^a, &{} \text {in } \Omega _T,\\ {\mathcal {B}}({{\textbf{V}}}|_{x_1=0},\psi ):={\mathbb {B}}({{\textbf{U}}}^a|_{x_1=0}+{{\textbf{V}}}|_{x_1=0},\varphi ^a+\psi )=0,&{} \text {on } \Gamma _T,\\ ({{\textbf{V}}},\psi )=0, &{} \text {for } t<0. \end{array}\right. } \end{aligned}$$
(8.11)

The original nonlinear problem on \([0,T]\times {\mathbb {R}}^2_+\) is thus reformulated as a problem on \(\Omega _T\) whose solutions vanish in the past.

9 Nash–Moser Iteration

In this section, we recall the Nash–Moser iteration for reader’s convenience. First, we introduce the smoothing operators \(S_{\theta }\) and describe the iterative scheme for problem (8.11). For more details refer to [5, 14, 39].

Lemma 9.1

Let \(\mu \in {\mathbb {N}},\) with \(\mu \ge 4\). \({\mathcal {F}}^s_{*}(\Omega _T):=\{u\in H^s_{*}(\Omega _T): u=0 \text { for }t<0\}.\) Define a family of smoothing operators \(\{S_{\theta }\}_{{\theta \ge 1}}\) on the anisotropic Sobolev space from \({\mathcal {F}}^{3}_{*}(\Omega _T)\) to \(\bigcap _{s\ge 3}{\mathcal {F}}^{s}_{*}(\Omega _T)\), such that

$$\begin{aligned}{} & {} ||S_{\theta }u||_{k,*,T}\le C\theta ^{(k-j)_+}||u||_{j,*,T}, \text { for all } k,j\in \{1,\cdots ,\mu \}, \end{aligned}$$
(9.1)
$$\begin{aligned}{} & {} ||S_{\theta }u-u||_{k,*,T}\le C\theta ^{k-j}||u||_{j,*,T}, \text { for all } 1\le k\le j\le \mu , \end{aligned}$$
(9.2)
$$\begin{aligned}{} & {} ||\frac{d}{d\theta }S_{\theta }u||_{k,*,T}\le C\theta ^{k-j-1}||u||_{j,*,T}, \text { for all } k,j\in \{1,\cdots ,\mu \}, \end{aligned}$$
(9.3)

where C is positive constant and \(k,j\in {\mathbb {N}},(k-j)_+:=\max \{0,k-j\}.\) In particular, if \(u=v\) on \(\Gamma _T,\) then \(S_{\theta }u=S_{\theta }v\) on \(\Gamma _T\). The definition of \({\mathcal {F}}^s(\Gamma _T)\) is entirely similar.

Now, we begin to formulate the Nash–Moser iteration scheme.

The iteration scheme starts from \(({{\textbf{V}}}_0,\Psi _0,\psi _0)=(0,0,0),\) and \(({{\textbf{V}}}_i,\Psi _i,\psi _i)\) is given such that

$$\begin{aligned} ({{\textbf{V}}}_i,\Psi _i,\psi _i)|_{t<0}=0,\quad \Psi ^+_{i}|_{x_1=0}=\Psi ^-_{i}|_{x_1=0}=\psi _i. \end{aligned}$$
(9.4)

Let us consider

$$\begin{aligned} {{\textbf{V}}}_{i+1}={{\textbf{V}}}_i+\delta {{\textbf{V}}}_i, \; \Psi _{i+1}=\Psi _i+\delta \Psi _i,\; \psi _{i+1}=\psi _i+\delta \psi _i, \end{aligned}$$
(9.5)

where the differences \((\delta {{{\textbf{V}}}}_i,\delta \psi _i)\) will be determined below. First, we can obtain \((\delta \dot{{{\textbf{V}}}}_i,\delta \psi _i)\) by solving the effective linear problem

$$\begin{aligned} {\left\{ \begin{array}{ll} \displaystyle {\mathbb {L}}_e'({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}},\Psi ^a+\Psi _{i+\frac{1}{2}})\delta \dot{{{\textbf{V}}}}_i=f_i &{}\text { in } \Omega _T,\\ \displaystyle {\mathbb {B}}_e'({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}},\Psi ^a+\Psi _{i+\frac{1}{2}})(\delta \dot{ {{\textbf{V}}}}_i,\delta \psi _i)=g_i &{}\text { on }\Gamma _T,\\ \displaystyle (\delta \dot{{{\textbf{V}}}}_i,\delta \psi _i)=0 &{}\text { for } t<0, \end{array}\right. } \end{aligned}$$
(9.6)

where operators \({\mathbb {L}}_e', {\mathbb {B}}_e'\) are defined in (4.15) and (4.16),

$$\begin{aligned} \delta \dot{{{\textbf{V}}}}_i:=\delta {{\textbf{V}}}_i-\frac{\partial _1({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}})}{\partial _1(\Phi ^a+\Psi _{i+\frac{1}{2}})}\delta \Psi _i \end{aligned}$$
(9.7)

is the Alinhac “good unknown" and \(({{\textbf{V}}}_{i+\frac{1}{2}},\Psi _{i+\frac{1}{2}})\) is a smooth modified state such that \(({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}},\Psi ^a+\Psi _{i+\frac{1}{2}})\) satisfies (4.2)–(4.8) and (4.10). The source terms \((f_i,g_i)\) will be defined through the accumulated errors at step i. \(S_{\theta _i}\) is the smoothing operator with \({\theta _i}\) defined by

$$\begin{aligned} \theta _0\ge 1, \; \theta _i=\sqrt{\theta ^2_0+i}. \end{aligned}$$
(9.8)

The errors at step i can be defined from the decompositions

$$\begin{aligned} \begin{aligned}&{\mathcal {L}}({{\textbf{V}}}_{i+1},\Psi _{i+1})-{\mathcal {L}}({{\textbf{V}}}_i,\Psi _i)\\&\quad ={\mathbb {L}}'({{\textbf{U}}}^a+{{\textbf{V}}}_i,\Psi ^a+\Psi _i)(\delta {{\textbf{V}}}_i,\delta \Psi _i)+e'_i\\&\quad ={\mathbb {L}}'({{\textbf{U}}}^a+S_{\theta _i}{{\textbf{V}}}_{i},\Psi ^a+S_{\theta _i}\Psi _i)(\delta {{\textbf{V}}}_i,\delta \Psi _i)+e'_i+e''_i\\&\quad ={\mathbb {L}}'({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}},\Psi ^a+\Psi _{i+\frac{1}{2}})(\delta {{\textbf{V}}}_i,\delta \Psi _i)+e'_i+e''_i+e'''_i\\&\quad ={\mathbb {L}}'_e({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}},\Psi ^a+\Psi _{i+\frac{1}{2}})\delta \dot{{{\textbf{V}}}}_i+e'_i+e''_i+e'''_i+D_{i+\frac{1}{2}}\delta \Psi _i\\ \end{aligned} \nonumber \\ \end{aligned}$$
(9.9)

and

$$\begin{aligned} \begin{aligned}&{\mathcal {B}}({{\textbf{V}}}_{i+1}|_{x_1=0},\psi _{i+1})-{\mathcal {B}}({{\textbf{V}}}_i|_{x_1=0},\psi _i)\\&\quad ={\mathbb {B}}'(({{\textbf{U}}}^a+{{\textbf{V}}}_i)|_{x_1=0},\varphi ^a+\psi _i)(\delta {{\textbf{V}}}_i|_{x_1=0},\delta \psi _i)+{\tilde{e}}'_i\\&\quad ={\mathbb {B}}'(({{\textbf{U}}}^a+S_{\theta _i}{{\textbf{V}}}_{i})|_{x_1=0},\varphi ^a+S_{\theta _i}\Psi _i|_{x_1=0})(\delta {{\textbf{V}}}_i|_{x_1=0},\delta \psi _i)+{\tilde{e}}'_i+{\tilde{e}}''_i\\&\quad ={\mathbb {B}}'_e(({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}})|_{x_1=0},\varphi ^a+\psi _{i+\frac{1}{2}})(\delta \dot{{{\textbf{V}}}}_i|_{x_1=0},\delta \psi _i)+{\tilde{e}}'_i+{\tilde{e}}''_i+{\tilde{e}}'''_i,\\ \end{aligned} \nonumber \\ \end{aligned}$$
(9.10)

where we write

$$\begin{aligned} D_{i+\frac{1}{2}}:=\frac{1}{\partial _1(\Phi ^a+\Psi _{i+\frac{1}{2}})}\partial _1{\mathbb {L}}({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}},\Psi ^a+\Psi _{i+\frac{1}{2}}), \end{aligned}$$
(9.11)

and have used (4.14) to get the last identity in (9.9). Denote

$$\begin{aligned} e_i:=e'_i+e''_i+e'''_i+D_{i+\frac{1}{2}}\delta \Psi _i, \quad {\tilde{e}}_i:={\tilde{e}}'_i+{\tilde{e}}''_i+{\tilde{e}}'''_i. \end{aligned}$$
(9.12)

We assume \(f_0:=S_{\theta _0}{\mathcal {F}}^a,(E_0,{\tilde{E}}_0,g_0):=(0,0,0)\) and \((f_k,g_k,e_k,{\tilde{e}}_k)\) are already given and vanish in the past for \(k\in \{0,\cdots ,i-1\}.\) We can calculate the accumulated errors at step \(i,i\ge 1\), by

$$\begin{aligned} E_i:=\sum ^{i-1}_{k=0}e_k,\quad {\tilde{E}}_i:=\sum ^{i-1}_{k=0}{\tilde{e}}_k. \end{aligned}$$
(9.13)

Then, we obtain \(f_i\) and \(g_i\) for \(i\ge 1\) from the equations

$$\begin{aligned} \sum ^{i}_{k=0}f_k+S_{\theta _i}E_i=S_{\theta _i}{\mathcal {F}}^a,\quad \sum ^{i}_{k=0}g_k+S_{\theta _i}{\tilde{E}}_i=0. \end{aligned}$$
(9.14)

Then, given suitable \(({{\textbf{V}}}_{i+\frac{1}{2}},\Psi _{i+\frac{1}{2}})\), we can obtain \((\delta \dot{{{\textbf{V}}}}_i,\delta \psi _i)\) as the solutions of the linear problem (9.6), \(\delta {{\textbf{V}}}_i\) from (9.7), \(({{\textbf{V}}}_{i+1},\Psi _{i+1}, \psi _{i+1})\) from (9.5). Since \(S_{\theta _i}\rightarrow I \) as \(i\rightarrow \infty ,\) we can formally obtain the solution to problem (8.11) from \({\mathcal {L}}({{\textbf{V}}}_{i},\Psi _{i})\rightarrow {\mathcal {F}}^a,{\mathcal {B}}({{\textbf{V}}}_{i}|_{x_1=0},\psi _{i})\rightarrow 0,\) as error terms \((e_i,{\tilde{e}}_i)\rightarrow 0.\)

10 Proof of the Main Result

Now, we prove the local existence of solutions to (8.11) by a modified iteration scheme of Nash–Moser type. From the sequence \(\{\theta _i\}\) defined in (9.8), we set \(\Delta _i:=\theta _{i+1}-\theta _i.\) Then, the sequence \(\{\Delta _i\}\) is decreasing and tends to 0 as i goes to infinity. Moreover, we have

$$\begin{aligned} \frac{1}{3\theta _i}\le \Delta _i=\sqrt{\theta ^2_i+1}-\theta _i\le \frac{1}{2\theta _i}, \; \forall \, i\in {\mathbb {N}}. \end{aligned}$$

10.1 Inductive analysis

Given a small fixed \(\delta >0\), and an integer \({{\tilde{\alpha }}}\) that will be chosen later on, we assume that the following estimate holds:

$$\begin{aligned} ||\tilde{{{\textbf{U}}}}^a||_{{{\tilde{\alpha }}}+6,*,T}+||\varphi ^a||_{H^{{{\tilde{\alpha }}}+6}(\Gamma _T)}+||{\mathcal {F}}^a||_{{{\tilde{\alpha }}}+4,*,T}\le \delta . \end{aligned}$$
(10.1)

We may assume that (10.1) holds, by taking \(T>0\) sufficiently small.

Given the integer \(\alpha \), our inductive assumptions read as

$$\begin{aligned} (H_{i-1}) \left\{ \begin{aligned}&\displaystyle (a)\quad ||(\delta {{\textbf{V}}}_k,\delta \Psi _k)||_{s,*,T}+||\delta \psi _k||_{H^{s}(\Gamma _T)}\le \delta \theta ^{s-\alpha -1}_{k}\Delta _k, \\&\qquad \forall k\in \{0,\cdots , i-1\},\quad \forall s\in \{6,\cdots ,{\tilde{\alpha }}\}.\\&\displaystyle (b)\quad ||{\mathcal {L}}({{\textbf{V}}}_k,\Psi _k)-{\mathcal {F}}^a||_{s,*,T}\le 2\delta \theta ^{s-\alpha -1}_k,\\&\qquad \forall k\in \{0,\cdots ,i-1\},\quad \forall s\in \{6,\cdots ,{\tilde{\alpha }}-2\}.\\&\displaystyle (c)\quad ||{\mathcal {B}}({{\textbf{V}}}_k|_{x_1=0},\psi _k)||_{H^{s}(\Gamma _T)}\le \delta \theta ^{s-\alpha -1}_k,\\&\qquad \forall k\in \{0,\cdots ,i-1\},\quad \forall s\in \{6,\cdots ,{\tilde{\alpha }}-2\}.\\ \end{aligned} \right. \end{aligned}$$
(10.2)

Our goal is to show that \((H_0)\) holds and \((H_{i-1})\) implies \((H_i)\), for a suitable choice of the parameters \(\alpha \), \({{\tilde{\alpha }}}\), for \(\delta >0\) and \(T>0\) sufficiently small, for \(\theta _0\ge 1\) sufficiently large. Then, we conclude that \((H_i)\) holds for all \(i\in {\mathbb {N}}.\)

Lemma 10.1

If \(T>0\) is sufficiently small, then \((H_0)\) holds.

Proof

The proof follows as in [39, Lemma 17]. \(\square \)

Now we prove that \((H_{i-1})\) implies \((H_i)\). The hypothesis \((H_{i-1})\) yields the following lemma:

Lemma 10.2

[39, Lemma 7], [14, Lemma 7] If \(\theta _0\) is large enough, then, for each \(k\in \{0,\cdots ,i\}\), and each integer \(s\in \{6,\cdots ,{\tilde{\alpha }}\},\)

$$\begin{aligned}{} & {} ||({{\textbf{V}}}_k,\Psi _k)||_{s,*,T}+||\psi _k||_{H^{s}(\Gamma _T)}\lesssim {\left\{ \begin{array}{ll} \delta \theta ^{(s-\alpha )_+}_k, &{}\text {if }s\ne \alpha ,\\ \delta \log \theta _k, &{}\text {if } s=\alpha ,\\ \end{array}\right. } \end{aligned}$$
(10.3)
$$\begin{aligned}{} & {} ||(I-S_{\theta _k})({{\textbf{V}}}_k,\Psi _k)||_{s,*,T}+||(I-S_{\theta _k})\psi _k||_{H^s(\Gamma _T)}\lesssim \delta \theta ^{s-\alpha }_k. \end{aligned}$$
(10.4)

Furthermore, for each \(k\in \{0,\cdots ,i\}\), and each integer \(s\in \{6,\cdots ,{\tilde{\alpha }}+8\},\)

$$\begin{aligned} ||(S_{\theta _k}{{\textbf{V}}}_k,S_{\theta _k}\Psi _k)||_{s,*,T}+||S_{\theta _k}\psi _k||_{H^s(\Gamma _T)}\lesssim {\left\{ \begin{array}{ll} \delta \theta ^{(s-\alpha )_+}_k, &{}\text{ if } s\ne \alpha ,\\ \delta \log \theta _k, &{}\text{ if } s=\alpha .\\ \end{array}\right. } \end{aligned}$$
(10.5)

10.2 Estimate of the error terms

To derive \((H_{i})\) from \((H_{i-1})\), we need to estimate the quadratic error terms \(e'_k\) and \({\tilde{e}}'_k,\) the first substitution error terms \(e''_k\) and \({\tilde{e}}''_k,\) the second substitution error terms \(e'''_k\) and \({\tilde{e}}'''_k\) and the last error term \(D_{k+\frac{1}{2}}\delta \Psi _k\) (cf. (9.9)–(9.11)).

First, we denote the quadratic error terms by

$$\begin{aligned}{} & {} e'_k:={\mathcal {L}}({{\textbf{V}}}_{k+1},\Psi _{k+1})-{\mathcal {L}}({{\textbf{V}}}_{k},\Psi _{k})-{\mathcal {L}}'({{\textbf{V}}}_{k},\Psi _{k})(\delta {{\textbf{V}}}_k,\delta \Psi _k), \end{aligned}$$
(10.6)
$$\begin{aligned}{} & {} {\tilde{e}}'_k:={\mathcal {B}}({{\textbf{V}}}_{k+1}|_{x_1=0},\psi _{k+1})-{\mathcal {B}}({{\textbf{V}}}_{k}|_{x_1=0},\psi _{k})-{\mathcal {B}}'({{\textbf{V}}}_{k}|_{x_1=0},\psi _{k})(\delta {{\textbf{V}}}_k|_{x_1=0},\delta \psi _k). \nonumber \\ \end{aligned}$$
(10.7)

Then, we get

$$\begin{aligned} e'_k={} & {} {} \int ^1_0{\mathbb {L}}''({{{\textbf {U}}}}^a+{{{\textbf {V}}}}_k+\tau \delta {{{\textbf {V}}}}_k,\Psi ^a+\Psi _k+\tau \delta \Psi _k)((\delta {{{\textbf {V}}}}_k,\delta \Psi _k),\\{} & {} (\delta {{{\textbf {V}}}}_k,\delta \Psi _k))(1-\tau )\text {d}\tau ,\\ {\tilde{e}}'_k={} & {} {} \frac{1}{2}{\mathbb {B}}''((\delta {{{\textbf {V}}}}_k,\delta \psi _k),(\delta {{{\textbf {V}}}}_k,\delta \psi _k)), \end{aligned}$$

where \({\mathbb {L}}'',{\mathbb {B}}''\) denote the second order derivatives of the operators \({\mathbb {L}}\) and \({\mathbb {B}}.\) To be more precise, we define

$$\begin{aligned}{} & {} {\mathbb {L}}''(\hat{{\textbf{U}}},{\hat{\Psi }})(({{\textbf{V}}},\Psi ),(\tilde{{{\textbf{V}}}},{\tilde{\Psi }})):=\frac{\textrm{d}}{\textrm{d}\varepsilon }{\mathbb {L}}'(\hat{{\textbf{U}}}+\varepsilon \tilde{{{\textbf{V}}}},{\hat{\Psi }}+\varepsilon {\tilde{\Psi }})({{\textbf{V}}},\Psi )\Big |_{\varepsilon =0}, \\{} & {} {\mathbb {B}}''(({{\textbf{V}}},\psi ),(\tilde{{{\textbf{V}}}},{\tilde{\psi }})):=\frac{\textrm{d}}{\textrm{d}\varepsilon }{\mathbb {B}}'(\hat{{\textbf{U}}}+\varepsilon \tilde{{{\textbf{V}}}},{\hat{\varphi }}+\varepsilon {\tilde{\psi }})({{\textbf{V}}},\psi )\Big |_{\varepsilon =0}, \end{aligned}$$

where \({\mathbb {L}}'\) and \({\mathbb {B}}'\) are defined in (4.11) and (4.12). Simple calculations yield that

$$\begin{aligned} {\mathbb {B}}''(({{\textbf{V}}},\psi ),(\tilde{{{\textbf{V}}}},{\tilde{\psi }})):=\left[ \begin{array}{c} {\tilde{u}}^+_2\partial _2\psi +\partial _2{\tilde{\psi }}u^+_2\\ {\tilde{u}}^-_2\partial _2\psi +\partial _2{\tilde{\psi }}u^-_2\\ {\textbf{H}}^+\cdot \tilde{{\textbf{H}}}^+-{\textbf{H}}^-\cdot \tilde{{\textbf{H}}}^- \end{array}\right] . \end{aligned}$$
(10.8)

To estimate the error terms, we need to estimate the operators \({\mathbb {L}}''\) and \({\mathbb {B}}''.\) Applying the Moser-type calculus inequalities in Lemma 3.1 and Lemma 3.2 and the explicit forms of \({\mathbb {L}}''\) and \({\mathbb {B}}''\), we can obtain the necessary estimates. Omitting the detailed calculation, we have the following Lemma 10.3:

Lemma 10.3

Let \(T>0,\) and \(s\in {\mathbb {N}}\) with \(s\ge 6.\) Assume that \((\breve{{\textbf{U}}},{\hat{\Psi }})\in H^{s+2}_{*}(\Omega _T)\) satisfies

$$\begin{aligned} ||(\hat{{{\textbf{U}}}},{\hat{\Psi }})||_{W^{2,\infty }_{*}(\Omega _T)}\le {\tilde{K}}, \end{aligned}$$

(recall that \(\hat{{{\textbf{U}}}}=\bar{{{\textbf{U}}}}+\breve{{{\textbf{U}}}}\)) for some constant \({\tilde{K}}>0.\) Then, there exists a positive C depending on \({\tilde{K}},\) but not on T,  such that if \(({{\textbf{V}}}_i,\Psi _i)\in H^{s+2}_{*}(\Omega _T)\) and \((W_i,\psi _i)\in H^{s}(\Gamma _T)\times H^{s+1}(\Gamma _T),\) for \(i=1,2,\) then

$$\begin{aligned} \begin{aligned}&||{\mathbb {L}}''(\hat{{\textbf{U}}},{\hat{\Psi }})(({{\textbf{V}}}_1,\Psi _1),({{\textbf{V}}}_2,\Psi _2))||_{s,*,T}\\&\quad \le C||(\breve{{{\textbf{U}}}},{\hat{\Psi }})||_{s+2,*,T}||({{\textbf{V}}}_1,\Psi _1)||_{W^{2,\infty }_{*}(\Omega _T)}||({{\textbf{V}}}_2,\Psi _2)||_{W^{2,\infty }_{*}(\Omega _T)}\\&\quad \quad +C\sum _{i\ne j}||({{\textbf{V}}}_i,\Psi _i)||_{s+2,*,T}||({{\textbf{V}}}_j,\Psi _j)||_{W^{2,\infty }_{*}(\Omega _T)}, \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned}&||{\mathbb {B}}''((W_1,\psi _1),(W_2,\psi _2))||_{H^s(\Gamma _T)}\\&\quad \le C\sum _{i\ne j}\Big (||W_i||_{H^s(\Gamma _T)}||\psi _j||_{W^{1,\infty }(\Gamma _T)} +||W_i||_{L^{\infty }(\Gamma _T)}||\psi _j||_{H^{s+1}(\Gamma _T)}\\&\quad \quad + ||W_i||_{H^s(\Gamma _T)}||W_j||_{L^{\infty }(\Gamma _T)}\Big ). \end{aligned} \end{aligned}$$

10.2.1 Estimate of the quadratic errors

We now apply Lemma 10.3 to prove the following estimate for the quadratic error terms:

Lemma 10.4

Let \(\alpha \ge 7.\) There exist \(\delta >0\) sufficiently small and \(\theta _0\ge 1\) sufficiently large such that, for all \(k\in \{0,\cdots ,i-1\},\) and all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\},\) we have

$$\begin{aligned}{} & {} ||e'_k||_{s,*,T}\lesssim \delta ^2\theta ^{L_1(s)-1}_k\Delta _k, \end{aligned}$$
(10.9)
$$\begin{aligned}{} & {} ||{\tilde{e}}'_k||_{H^{s}(\Gamma _T)}\lesssim \delta ^2\theta ^{L_1(s)-1}_k\Delta _k, \end{aligned}$$
(10.10)

where \(L_1(s):=\max \{(s+2-\alpha )_++10-2\alpha ;s+6-2\alpha \}.\)

Proof

Using (10.1), the hypothesis \((H_{i-1})\) and the estimate (10.3), we use the Sobolev inequalities (3.10) to get

$$\begin{aligned} ||({\textbf{U}}^a,{{\textbf{V}}}_k,\delta {{\textbf{V}}}_k,\Psi ^a,\Psi _k,\delta \Psi _k)||_{W^{2,\infty }_{*}(\Omega _T)}\lesssim 1. \end{aligned}$$

Then, we apply Lemma 10.3 and use Sobolev inequalities (3.10), the assumption (10.1) and the hypothesis \((H_{i-1})\) to give

$$\begin{aligned} \begin{aligned} ||e'_k||_{s,*,T}&\lesssim \delta ^2\theta ^{10-2\alpha }_k\Delta ^2_k(1+||({{\textbf{V}}}_k,\Psi _k)||_{s+2,*,T}+\delta \theta ^{s+1-\alpha }_k\Delta _k)\\&\quad \quad +||(\delta {{\textbf{V}}}_k,\delta \Psi _k)||_{s+2,*,T}||(\delta {{\textbf{V}}}_k,\delta \Psi _k)||_{6,*,T}\\&\lesssim \delta ^2\theta ^{10-2\alpha }_k\Delta ^2_k(1+||({{\textbf{V}}}_k,\Psi _k)||_{s+2,*,T})+\delta ^2\theta ^{s+6-2\alpha }_k\Delta ^2_k, \end{aligned} \end{aligned}$$

for all \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\}.\) If \(s+2\ne \alpha ,\) then it follows from (10.3) and \(2\theta _k\Delta _k\le 1,\) that

$$\begin{aligned} \begin{aligned} ||e'_k||_{s,*,T}&\lesssim \delta ^2\Delta ^2_k(\theta ^{(s+2-\alpha )_++10-2\alpha }_k+\theta ^{s+6-2\alpha }_k)\\&\lesssim \delta ^2\theta ^{L_1(s)-1}_k\Delta _k. \end{aligned} \end{aligned}$$

If \(s+2=\alpha ,\) then it follows from (10.3) and \(\alpha \ge 7,\) that

$$\begin{aligned} \begin{aligned} ||e'_k||_{s,*,T}&\lesssim \delta ^2\Delta ^2_k(\theta ^{11-2\alpha }_k+\theta ^{4-\alpha }_k)\\&\lesssim \delta ^2\theta ^{L_1(\alpha -2)-1}_k\Delta _k. \end{aligned} \end{aligned}$$

Therefore, we obtain (10.9). Now, we prove (10.10). Using Lemma 10.3 and trace Theorem A1.1, we obtain

$$\begin{aligned} \begin{aligned} ||{\tilde{e}}'_k||_{H^s(\Gamma _T)}&\lesssim ||\delta {{\textbf{V}}}_k||_{H^s(\Gamma _T)}||\delta \psi _k||_{W^{1,\infty }(\Gamma _T)}+||\delta {{\textbf{V}}}_k||_{L^{\infty }(\Gamma _T)}||\delta \psi _k||_{H^{s+1}(\Gamma _T)}\\&\quad + ||\delta {{\textbf{V}}}_k||_{H^s(\Gamma _T)}||\delta {{\textbf{V}}}_k||_{L^{\infty }(\Gamma _T)}\\&\lesssim \delta ^2\theta ^{L_1(s)-1}_k\Delta _k. \end{aligned} \end{aligned}$$

This completes the proof of Lemma 10.4. \(\square \)

10.2.2 Estimate of the first substitution errors

We can estimate the first substitution errors \(e''_k, {\tilde{e}}''_k\) of the iteration scheme, defined in (9.9) and (9.10). We rewrite

$$\begin{aligned}{} & {} e''_k:={\mathcal {L}}'({{\textbf{V}}}_k,\Psi _k)(\delta {{\textbf{V}}}_k,\delta \Psi _k)-{\mathcal {L}}'(S_{\theta _k}{{\textbf{V}}}_k,S_{\theta _k}\Psi _k)(\delta {{\textbf{V}}}_k,\delta \Psi _k), \end{aligned}$$
(10.11)
$$\begin{aligned}{} & {} {\tilde{e}}''_k:={\mathcal {B}}'({{\textbf{V}}}_k|_{x_1=0},\psi _k)(\delta {{\textbf{V}}}_k|_{x_1=0},\delta \psi _k)-{\mathcal {B}}'(S_{\theta _k}{{\textbf{V}}}_k|_{x_1=0},S_{\theta _k}\psi _k)(\delta {{\textbf{V}}}_k|_{x_1=0},\delta \psi _k). \nonumber \\ \end{aligned}$$
(10.12)

Lemma 10.5

Let \(\alpha \ge 7.\) There exist \(\delta >0\) sufficiently small and \(\theta _0\ge 1\) sufficiently large, such that for all \(k\in \{0,\cdots ,i-1\}\) and for all integer \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\},\) we have

$$\begin{aligned} ||e''_k||_{s,*,T}\lesssim \delta ^2\theta ^{L_2(s)-1}_k\Delta _k, \end{aligned}$$
(10.13)
$$\begin{aligned} ||{\tilde{e}}''_k||_{H^{s}(\Gamma _T)}\lesssim \delta ^2\theta ^{L_2(s)-1}_k\Delta _k, \end{aligned}$$
(10.14)

where \(L_2(s):=\max \{(s+2-\alpha )_++12-2\alpha ;s+8-2\alpha \}.\)

Proof

In view of (10.11) and (10.12) we have

$$\begin{aligned} \begin{aligned} e''_k&=\int ^1_0{\mathbb {L}}''({\textbf{U}}^a+S_{\theta _k}{{\textbf{V}}}_k+\tau (I-S_{\theta _k}){{\textbf{V}}}_k,\Psi ^a+S_{\theta _k}\Psi _k+\tau (I-S_{\theta _k})\Psi _k)\\&\quad \quad ((\delta {{\textbf{V}}}_k,\delta \Psi _k),((I-S_{\theta _k}){{\textbf{V}}}_k,(I-S_{\theta _k})\Psi _k))\textrm{d}\tau ,\\ {\tilde{e}}''_k&={\mathbb {B}}''((\delta {{\textbf{V}}}_k|_{x_1=0},\delta \psi _k),((I-S_{\theta _k}){{\textbf{V}}}_k|_{x_1=0},(I-S_{\theta _k})\psi _k)). \end{aligned} \end{aligned}$$

Using (10.4) and (10.5), we have

$$\begin{aligned} ||(S_{\theta _k}{{\textbf{V}}}_k,{{\textbf{V}}}_k,S_{\theta _k}\Psi _k,\Psi _k)||_{W^{2,\infty }_{*}(\Omega _T)}\lesssim ||(S_{\theta _k}{{\textbf{V}}}_k,{{\textbf{V}}}_k,S_{\theta _k}\Psi _k,\Psi _k)||_{6,*,T}\lesssim 1. \end{aligned}$$

Next, we apply Lemma 10.3, use Sobolev inequalities (3.10), (10.1), the Hypothesis \((H_{i-1})\) and (10.4) to get that

$$\begin{aligned} ||e''_k||_{s,*,T}\lesssim \delta ^2\theta ^{11-2\alpha }_k\Delta _k(1+||(S_{\theta _k}{{\textbf{V}}}_k,S_{\theta _k}\Psi _k)||_{s+2,*,T})+\delta ^2\theta ^{s+7-2\alpha }_k\Delta _k, \end{aligned}$$

for all \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\}.\) Similar to the proof of Lemma 10.4, we can discuss \(s+2\ne \alpha \) and \(s+2=\alpha \) separately. Hence, using (10.5), we can obtain (10.13) and (10.14). The proof of Lemma 10.5 is completed. \(\square \)

10.2.3 Estimate of the modified state

We need to construct a smooth modified state \(({{\textbf{V}}}_{i+\frac{1}{2}},\psi _{i+\frac{1}{2}})\) such that \(({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}},\varphi ^a+\psi _{i+\frac{1}{2}})\) satisfies the nonlinear constraints (4.2)–(4.8), (4.10) and (6.1). In this regard, we remark it is crucial that the sum of the approximate solution \(({{\textbf{U}}}^a,\varphi ^a)\) and the modified state \(({{\textbf{V}}}_{i+\frac{1}{2}},\psi _{i+\frac{1}{2}})\), instead of the latter two separately, satisfies the aforementioned nonlinear constraints; indeed it is just this sum which plays the role of basic state around which we need to linearize problem (2.1) in the iteration scheme leading to its solution, see problem (9.6). In the construction of some components of the modified state we follow an approach similar to that of [1, 14, 34], while for the magnetic field we are inspired by [34, 39].

Lemma 10.6

Let \(\alpha \ge 10\). There exist some functions \({{\textbf{V}}}_{i+\frac{1}{2}},\Psi _{i+\frac{1}{2}},\psi _{i+\frac{1}{2}}\) vanishing in the past, such that \(({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \Psi ^a+\Psi _{i+\frac{1}{2}}, \varphi ^a+\psi _{i+\frac{1}{2}})\) satisfy the constraints (4.2)– (4.8), (4.10) and (6.1); moreover,

$$\begin{aligned}{} & {} \Psi _{i+\frac{1}{2}}^\pm =S_{\theta _i}\Psi ^{\pm }_i,\quad \psi _{i+\frac{1}{2}}=(S_{\theta _i}\Psi ^{\pm }_i)|_{x_1=0}, \end{aligned}$$
(10.15)
$$\begin{aligned}{} & {} p^{\pm }_{i+\frac{1}{2}}=S_{\theta _i}p^\pm _i,\quad u^{\pm }_{2,i+\frac{1}{2}}=S_{\theta _i}u^{\pm }_{2,i},\quad S^{\pm }_{i+\frac{1}{2}}=S_{\theta _i}S^{\pm }_i, \end{aligned}$$
(10.16)
$$\begin{aligned}{} & {} {} ||{{{\textbf {V}}}}_{i+\frac{1}{2}}-S_{\theta _i}{{{\textbf {V}}}}_i||_{s,*,T}\lesssim \delta \theta ^{s+4-\alpha }_i, \text{ for } s\in \{6,\cdots ,{\tilde{\alpha }}+4\}, \end{aligned}$$
(10.17)

for sufficiently small \(\delta >0\) and \(T>0,\) and sufficiently large \(\theta _0\ge 1.\)

Proof

To shortcut notation, in the proof the ± indices are omitted. Let us define \(\Psi _{i+1/2}\), \(\psi _{i+1/2}\), \(p_{i+1/2}\), \(S_{i+1/2}\) and the tangential component \(u_{2,i+1/2}\) of the velocity as in (10.15), (10.16); this is similar as in [1, 14, Proposition 7], [34, Proposition 28]. It is easily checked that all these functions vanish in the past.

Construction of the modified normal velocity.

In order to construct the normal component \(u_{1,i+1/2}\) of the velocity, we follow the idea of [1, 14]. We first introduce the following function \({\mathcal {G}}\):

$$\begin{aligned} \begin{aligned} {\mathcal {G}}:=&\partial _t(\varphi ^a+\psi _{i+1/2})-(u^a_1+S_{\theta _i}u_{1,i})\vert _{x_1=0}+u^a_2\vert _{x_1=0}\partial _2\varphi ^a\\&+(u^a_2+u_{2,i+1/2})\vert _{x_1=0}\partial _2\psi _{i+1/2}+u_{2,i+1/2}\vert _{x_1=0}\partial _2\varphi ^a. \end{aligned} \end{aligned}$$

The normal component of the velocity \(u_{1,i+1/2}\) is defined by

$$\begin{aligned} u_{1,i+1/2}:=S_{\theta _i}u_{1,i}+{\mathcal {R}}_T{\mathcal {G}}, \end{aligned}$$
(10.18)

where \({\mathcal {R}}_T\) is the lifting operator \(H^{s-1}(\Gamma _T)\rightarrow H^s_*(\Omega _T)\), \(s>1\), see [27]. It is easily checked that \(u_{1,i+1/2}\) vanishes in the past.

Let us note that, by construction, \(u_{1,i+1/2}\) satisfies the following equation on the boundary:

$$\begin{aligned}{} & {} \partial _t(\varphi ^a+\psi _{i+1/2})-(u_1^a+u_{1,i+1/2})+(u_2^a+u_{2,i+1/2})\partial _2(\varphi ^a+\psi _{i+1/2})\nonumber \\{} & {} \quad =0,\quad \text{ on }\,\,\,\{x_1=0\}. \end{aligned}$$
(10.19)

We prove the estimate (10.17) for the part regarding \({{\textbf{u}}}_{i+1/2}\). We have

$$\begin{aligned} ||{\textbf{u}}_{i+1/2}-S_{\theta _i}{{\textbf{u}}}_i||_{s,*,T}=||{\mathcal {R}}_T{\mathcal {G}}||_{s,*,T}\lesssim ||{\mathcal {G}}||_{H^{s-1}(\Gamma _T)}. \end{aligned}$$

Now we rewrite \({\mathcal {G}}\) in a more convenient form, by using the error \(\varepsilon ^i\) defined by

$$\begin{aligned} \begin{aligned} \varepsilon ^i:=&\partial _t\psi _i-u_{1,i}\vert _{x_1=0}+(\partial _t\varphi ^a-u_1^a\vert _{x_1=0}+u_2^a\vert _{x_1=0}\partial _2\varphi ^a)+(u^a_2+u_{2,i})\vert _{x_1=0}\partial _2\psi _i\\&+u_{2,i}\vert _{x_1=0}\partial _2\varphi ^a={\mathcal {B}}({{\textbf{V}}}_i\vert _{x_1=0},\psi _i)_1. \end{aligned} \nonumber \\ \end{aligned}$$
(10.20)

In view of (10.16), by means of \(\varepsilon ^i\) we may rewrite \({\mathcal {G}}\) as

$$\begin{aligned} \begin{aligned} {\mathcal {G}}&= S_{\theta _i}\varepsilon ^i-[S_{\theta _i},\partial _t]\psi _i+(I-S_{\theta _i})(\partial _t\varphi ^a-u^a_1+u^a_2\partial _2\varphi ^a)-S_{\theta _i}((u^a_2+u_{2,i})\partial _2\psi _i)\\&\quad -S_{\theta _i}(u_{2,i}\partial _2\varphi ^a)+(u^a_2+S_{\theta _i}u_{2,i})\partial _2\psi _{i+1/2}+S_{\theta _i}u_{2,i}\partial _2\varphi ^a\\&=S_{\theta _i}\varepsilon ^i-[S_{\theta _i},\partial _t]\psi _i+(I-S_{\theta _i})(\partial _t\varphi ^a-u^a_1+u^a_2\partial _2\varphi ^a))\\&\quad -\left( S_{\theta _i}(u^a_2\partial _2\psi _i)-u_2^a\partial _2 S_{\theta _i}\psi _i\right) -\left( S_{\theta _i}u_{2,i}\partial _2 S_{\theta _i}\psi _i-S_{\theta _i}(u_{2,i}\partial _2\psi _i)\right) \\&\quad -\left( S_{\theta _i}u_{2,i}\partial _2\varphi ^a-S_{\theta _i}(u_{2,i}\partial _2\varphi ^a)\right) . \end{aligned}\nonumber \\ \end{aligned}$$
(10.21)

To estimate the first term \(S_{\theta _i}\varepsilon ^i\) on the right-hand side, we use the decomposition

$$\begin{aligned} \begin{aligned} \varepsilon ^i&=\left( {\mathcal {B}}({{\textbf{V}}}_i,\psi _i)_1-{\mathcal {B}}({{\textbf{V}}}_{i-1},\psi _{i-1})_1\right) +{\mathcal {B}}({{\textbf{V}}}_{i-1},\psi _{i-1})_1\\&=\partial _t\delta \psi _{i-1}-\delta u_{1,i-1}+(u_2^a+u_{2,i-1})\partial _2\delta \psi _{i-1}+\delta u_{2,i-1}\partial _2(\varphi ^a+\delta \psi _{i-1}+\psi _{i-1})\\&\quad +{\mathcal {B}}({{\textbf{V}}}_{i-1},\psi _{i-1})_1. \end{aligned} \end{aligned}$$

Then we exploit point (c) of (\(H_{i-1}\)) and the properties of smoothing operators, to get

$$\begin{aligned} ||S_{\theta _i}\varepsilon ^i||_{H^{s-1}(\Gamma _T)}\lesssim ||\varepsilon ^i||_{H^{s-1}(\Gamma _T)}\lesssim \delta \theta _i^{s-\alpha -1}. \end{aligned}$$
(10.22)

In order to make an estimate of the commutator term \([S_{\theta _i},\partial _t]\psi _i\), we use different arguments for large and small orders s.

For all \(s\in \{6,\dots ,\alpha \}\) we write the commutator as

$$\begin{aligned}{}[\partial _t,S_{\theta _i}]\psi _i=\partial _t(S_{\theta _i}-I)\psi _i+(I-S_{\theta _i})\partial _t\psi _i, \end{aligned}$$

then we use estimates (9.2), (10.3) and (10.4) to get

$$\begin{aligned} ||[\partial _t,S_{\theta _i}]\psi _i||_{s,*,T}\le C\delta \theta _i^{s-\alpha },\quad s\in \{6,\dots ,\alpha \}. \end{aligned}$$
(10.23)

In order to get the similar estimate as (10.23) for \(s\in \{\alpha +1,\dots ,{{\tilde{\alpha }}}+6\}\), we directly estimate the two terms of the commutator \([\partial _t,S_{\theta _i}]\psi _i=\partial _tS_{\theta _i}\psi _i-S_{\theta _i}\partial _t\psi _i\), using (9.1) and (10.5).

To estimate the third term in the right-hand side of (10.21), we proceed as above by applying different arguments to small and large orders s.

For integers \(s\in \{6,\dots ,\alpha \}\), we apply (9.2) to get

$$\begin{aligned} \begin{aligned} ||(I-S_{\theta _i})&(\partial _t\varphi ^a-u^a_1+u_2^a\partial _2\varphi ^a)||_{H^{s-1}(\Gamma _T)}\le C\theta _i^{s-\alpha }||\partial _t\varphi ^a-u^a_1+u_2^a\partial _2\varphi ^a||_{H^{\alpha -1}(\Gamma _T)}\\&\le C\theta _i^{s-\alpha }(||\varphi ^a||_{H^{\alpha -1}(\Gamma _T)}+||u^a||_{\alpha ,*,T})\le C\delta \theta _i^{s-\alpha }, \end{aligned} \end{aligned}$$

in view of (10.1).

For integers \(s\in \{\alpha +1,\dots .{{\tilde{\alpha }}}+6\}\), we use (9.1) and (10.1) to estimate directly

$$\begin{aligned} \begin{aligned} ||(I-S_{\theta _i})&(\partial _t\varphi ^a-u^a_1+u_2^a\partial _2\varphi ^a)||_{H^{s-1}(\Gamma _T)}\le ||E^a||_{H^{s-1}(\Gamma _T)}+||S_{\theta _i}E^a||_{H^{s-1}(\Gamma _T)}\\&\le ||E^a||_{H^{s-1}(\Gamma _T)}+C\theta _i^{(s-1-(\alpha -1))_+}||E^a||_{H^{\alpha -1}(\Gamma _T)}\le C\theta _i^{s-\alpha }||E^a||_{H^{s-1}(\Gamma _T)}\\&\le C\theta _i^{s-\alpha }(||\varphi ^a||_{H^{s}(\Gamma _T)}+||u^a||_{s,*,T})\le C\delta \theta _i^{s-\alpha }, \end{aligned} \end{aligned}$$

where \(E^a:=\partial _t\varphi ^a-u^a_1+u_2^a\partial _2\varphi ^a\) has been set.

Let us now estimate the fourth term \(-\left( S_{\theta _i}(u^a_2\partial _2\psi _i)-u_2^a\partial _2 S_{\theta _i}\psi _i\right) \); once again we need to argue separately on different values of s.

For small integers \(s\in \{6,\dots ,{{\tilde{\alpha }}}\}\) we rewrite the above in the form

$$\begin{aligned} u_2^a\partial _2 S_{\theta _i}\psi _i-S_{\theta _i}(u^a_2\partial _2\psi _i)=u_2^a\partial _2(S_{\theta _i}-I)\psi _i+(I-S_{\theta _i})(u_2^a\partial _2\psi _i), \end{aligned}$$
(10.24)

which takes advantage of the appearing of the difference \(I-S_{\theta _i}\); thus estimate (10.4) and Moser-type calculus inequalities of Lemma 3.1 yield

$$\begin{aligned} \begin{aligned} ||u_2^a\partial _2(S_{\theta _i}-I)&\psi _i||_{H^{s-1}(\Gamma _T)}\lesssim ||u^a_2||_{L^\infty (\Gamma _T)}||(I-S_{\theta _i})\psi _i||_{H^s(\Gamma _T)}\\&+||u^a||_{s,*,T}||(I-S_{\theta _i})\psi _i||_{H^2(\Gamma _T)}\le C\delta \theta _i^{s-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,{{\tilde{\alpha }}}\}. \end{aligned}\nonumber \\ \end{aligned}$$
(10.25)

As for the second term in the right-hand side of (10.24), a further splitting of the range of s covered by estimate (10.25) is required. For integers s such that \(6\le s\le \alpha +1\), we apply (9.2), Moser-type calculus inequalities of Lemma 3.1 and (10.3) to obtain

$$\begin{aligned} \begin{aligned} ||(I-S_{\theta _i})&(u_2^a\partial _2\psi _i)||_{H^{s-1}(\Gamma _T)}\le C\theta _i^{s-1-(\alpha +1)}||u_2^a\partial _2\psi _i||_{H^{\alpha +1}(\Gamma _T)}\\&\le C\theta _i^{s-\alpha -2}\left\{ ||u^a||_{L^\infty (\Gamma _T)}||\psi _i||_{H^{\alpha +1}(\Gamma _T)}+||u^a||_{\alpha +2,*,T}||\psi _i||_{H^3(\Gamma _T)}\right\} \\&\le C\theta _i^{s-\alpha -2}(\delta \theta _i+\delta )\le C\delta \theta _i^{s-\alpha -1},\quad \text{ for }\,\,\,6\le s\le \alpha +1. \end{aligned}\nonumber \\ \end{aligned}$$
(10.26)

On the other hand for integers s such that \(\alpha +1<s\le {{\tilde{\alpha }}}\), we use (9.1), Moser-type calculus inequalities and (10.3) to get

$$\begin{aligned} \begin{aligned} ||(I-S_{\theta _i})&(u_2^a\partial _2\psi _i)||_{H^{s-1}(\Gamma _T)}\le C||u_2^a\partial _2\psi _i||_{H^{s-1}(\Gamma _T)}\\&\le C\left\{ ||u^a||_{L^\infty (\Gamma _T)}||\psi _i||_{H^{s}(\Gamma _T)}+||u^a||_{s,*,T}||\psi _i||_{H^3(\Gamma _T)}\right\} \\&\le C\theta _i^{(s-\alpha )_+}+C\delta \le C\delta \theta _i^{s-\alpha } ,\quad \text{ for }\,\,\,\alpha +1<s\le {{\tilde{\alpha }}}. \end{aligned}\nonumber \\ \end{aligned}$$
(10.27)

Gathering (10.25)–(10.27) we end up with

$$\begin{aligned} ||u_2^a\partial _2 S_{\theta _i}\psi _i-S_{\theta _i}(u^a_2\partial _2\psi _i)||_{H^{s-1}(\Gamma _T)}\le C\delta \theta _i^{s-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,{{\tilde{\alpha }}}\}. \nonumber \\ \end{aligned}$$
(10.28)

For higher integers \(s\in \{{{\tilde{\alpha }}}+1,\dots ,{{\tilde{\alpha }}}+6\}\), we estimate separately the two terms of the difference \(u_2^a\partial _2 S_{\theta _i}\psi _i-S_{\theta _i}(u^a_2\partial _2\psi _i)\); using again estimates (9.1), (10.3), (10.5) and Moser-type calculus inequalities, we obtain

$$\begin{aligned} ||u^a_2\partial _2 S_{\theta _i}\psi _i||_{H^{s-1}(\Gamma _T)}\lesssim & {} ||u^a_2||_{L^\infty (\Gamma _T)}||S_{\theta _i}\psi _i||_{H^s(\Gamma _T)}+||u^a_2||_{s,*,T}||S_{\theta _i}\psi _i||_{H^3(\Gamma _T)}\\\le & {} C\delta \theta _i^{s-\alpha }; \\ ||S_{\theta _i}(u^a_2\partial _2\psi _i)||_{H^{s-1}(\Gamma _T)}\le & {} C\theta _i^{s-1-\alpha }||u^a_2\partial _2\psi _i||_{H^{\alpha }(\Gamma _T)}\\\le & {} C\theta _i^{s-1-\alpha }\Big \{ ||u^a_2||_{L^\infty (\Gamma _T)}||\psi _i||_{H^{\alpha +1}(\Gamma _T)}\\{} & {} +||u^a_2||_{\alpha +1,*,T}||\psi _i||_{H^3(\Gamma _T)}\Big \}\\\le & {} C\theta _i^{s-1-\alpha }(\delta \theta _i+C\delta )\le C\delta \theta _i^{s-\alpha },\\{} & {} \text{ for }\,\,\,{{\tilde{\alpha }}}+1\le s\le {{\tilde{\alpha }}}+6. \end{aligned}$$

Adding the last two inequalities we end up with

$$\begin{aligned} ||u_2^a\partial _2 S_{\theta _i}\psi _i-S_{\theta _i}(u^a_2\partial _2\psi _i)||_{H^{s-1}(\Gamma _T)}\le C\delta \theta _i^{s-\alpha },\quad \text{ for }\,\,\,s\in \{{{\tilde{\alpha }}}+1,\dots ,{{\tilde{\alpha }}}+6\}. \nonumber \\ \end{aligned}$$
(10.29)

Gathering estimates (10.28) and (10.29) provide the estimate for all integers \(s\in \{6,\dots ,{{\tilde{\alpha }}}+6\}\).

For the last two terms in the right-hand side of (10.21) we use the same arguments as above, where we still separate small and large orders s; in the case of small s we manage to rewrite the expression in order to make advantage from the boundedness properties (9.2), (10.4) of \(I-S_{\theta _i}\); for large s we estimate separately each term of the difference using Moser-type calculus inequalities and (9.1), (10.3), (10.5). Doing so, we derive the following:

$$\begin{aligned}{} & {} ||S_{\theta _i}u_{2,i}\partial _2S_{\theta _i}\psi _i-S_{\theta _i}(u_{2,i}\partial _2\psi _i)||_{H^{s-1}(\Gamma _T)}\le C\delta \theta _i^{s-\alpha +1},\quad \text{ for }\,\,\,s\in \{6,\dots ,{{\tilde{\alpha }}}+6\}; \\{} & {} ||S_{\theta _i}u_{2,i}\partial _2\varphi ^a-S_{\theta _i}(u_{2,i}\partial _2\varphi ^a)||_{H^{s-1}(\Gamma _T)}\le C\delta \theta _i^{s-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,{{\tilde{\alpha }}}+6\}. \end{aligned}$$

Gathering all the previously found estimates we end up with

$$\begin{aligned} ||{\textbf{u}}_{i+1/2}-S_{\theta _i}{{\textbf{u}}}_i||_{s,*,T}\lesssim ||{\mathcal {G}}||_{H^{s-1}(\Gamma _T)}\le C\delta \theta _i^{s-\alpha +1},\quad \text{ for }\,\,\,s\in \{6,\dots ,{{\tilde{\alpha }}}+6\}. \nonumber \\ \end{aligned}$$
(10.30)

In the above estimate it is fundamental that \(s\le {\tilde{\alpha }}+6\) in order to prove the following (10.40). Let us recall that the above estimate holds under the smallness assumption (10.1).

Construction of the modified magnetic field.

Let us see now how to define the modified magnetic field \({\textbf{H}}_{i+1/2}\), following [34, Proposition 28], [39, Proposition 12].

Let us denote the nonlinear equation satisfied by the magnetic field in (2.1) by

$$\begin{aligned} {\mathbb {L}}_H({{\textbf{u}}},{\textbf{H}},\Psi )=0,\quad \text{ in }\,\,\,\Omega _T. \end{aligned}$$

The field \({\textbf{H}}_{i+1/2}\) should be such that \(\textbf{H}^a+{\textbf{H}}_{i+1/2}\) satisfies (4.8), that is,

$$\begin{aligned} {\mathbb {L}}_H({{\textbf{u}}}^a+{\textbf{u}}_{i+1/2},{\textbf{H}}^a+{\textbf{H}}_{i+1/2},\Psi ^a+\Psi _{i+1/2})=0,\quad \text{ in }\,\,\,\Omega _T. \end{aligned}$$
(10.31)

We note that equation (10.31) is linear in \({\textbf{H}}^a+{\textbf{H}}_{i+1/2}\) and does not need to be supplemented with any boundary condition; in fact, the coefficient of \(\partial _1({\textbf{H}}^a+{\textbf{H}}_{i+1/2})\) is zero along the boundary because of (10.19) (the left-hand side of (10.19) is nothing but \(w_1\vert _{x_1=0}\), computed for \({{\textbf{u}}}^a+{{\textbf{u}}}_{i+1/2}\) and \(\Psi ^a+\Psi _{i+1/2}\) instead of \({\textbf{u}}\) and \(\Psi \) respectively).

Therefore, for given \({{\textbf{u}}}_{i+1/2}\), \(\Psi _{i+1/2}\), \({{\textbf{U}}}^a\) and \(\Psi ^a\), (10.31) has a unique solution \({\textbf{H}}'\), from which we derive the existence of a unique \({\textbf{H}}_{i+1/2}={\textbf{H}}'-{\textbf{H}}^a\), vanishing in the past.

In order to estimate \({{\textbf{H}}}_{i+1/2}-S_{\theta _i}{{\textbf{H}}}_i\), we first observe that (10.31) yields

$$\begin{aligned} \begin{aligned} {\mathbb {L}}_H&({{\textbf{u}}}^a+{\textbf{u}}_{i+1/2},{{\textbf{H}}}_{i+1/2}-S_{\theta _i}{{\textbf{H}}}_i,\Psi ^a+\Psi _{i+1/2})\\&={\mathbb {L}}_H({{\textbf{u}}}^a+{\textbf{u}}_{i+1/2},{{\textbf{H}}}^a+{{\textbf{H}}}_{i+1/2}-S_{\theta _i}{{\textbf{H}}}_i,\Psi ^a+\Psi _{i+1/2})\\&\quad \quad -{\mathbb {L}}_H({{\textbf{u}}}^a+{\textbf{u}}_{i+1/2},{{\textbf{H}}}^a,\Psi ^a+\Psi _{i+1/2})\\&=-{\mathbb {L}}_H({{\textbf{u}}}^a+{\textbf{u}}_{i+1/2},{{\textbf{H}}}^a+S_{\theta _i}{{\textbf{H}}}_i,\Psi ^a+S_{\theta _i}\Psi _{i}). \end{aligned} \end{aligned}$$

Then \({\textbf{H}}_{i+1/2}-S_{\theta _i}{\textbf{H}}_i\) solves the equation

$$\begin{aligned} {\mathbb {L}}_H({{\textbf{u}}}^a+{\textbf{u}}_{i+1/2},{{\textbf{H}}}_{i+1/2}-S_{\theta _i}{{\textbf{H}}}_i,\Psi ^a+\Psi _{i+1/2})=F^{i+1/2}_H, \end{aligned}$$
(10.32)

where

$$\begin{aligned} F_H^{i+1/2}{}{} & {} {} :=\Delta _1+\Delta _2-S_{\theta _i}{\mathbb {L}}_H({{{\textbf {u}}}}^a+{{\textbf {u}}}_{i},{{{\textbf {H}}}}^a+{{{\textbf {H}}}}_i,\Psi ^a+\Psi _{i});\nonumber \\ \Delta _1{}{} & {} {} :=S_{\theta _i}{\mathbb {L}}_H({{{\textbf {u}}}}^a+{{\textbf {u}}}_{i},{{{\textbf {H}}}}^a+{{{\textbf {H}}}}_i,\Psi ^a+\Psi _{i})\\ {}{} & {} \quad -{\mathbb {L}}_H({{{\textbf {u}}}}^a+S_{\theta _i}{{\textbf {u}}}_{i},{{{\textbf {H}}}}^a\nonumber +S_{\theta _i}{{{\textbf {H}}}}_i,\Psi ^a+S_{\theta _i}\Psi _{i});\nonumber \\ \Delta _2{}{} & {} {} :={\mathbb {L}}_H({{{\textbf {u}}}}^a+S_{\theta _i}{{\textbf {u}}}_{i},{{{\textbf {H}}}}^a+S_{\theta _i}{{{\textbf {H}}}}_i,\Psi ^a+S_{\theta _i}\Psi _{i})\nonumber \\{}{} & {} {} \quad -{\mathbb {L}}_H({{{\textbf {u}}}}^a+{{\textbf {u}}}_{i+1/2},{{{\textbf {H}}}}^a+S_{\theta _i}{{{\textbf {H}}}}_i,\Psi ^a+S_{\theta _i}\Psi _{i}).\nonumber \end{aligned}$$
(10.33)

Let us first write the explicit form of \(\Delta _1\). Using the definition of the nonlinear operator \({\mathbb {L}}_H(\textbf{u},\textbf{H},\Psi )\) we have

$$\begin{aligned} \begin{aligned} \Delta _1=&S_{\theta _i}\Big \{\partial _t(\textbf{H}^a+\textbf{H}_i)+\frac{1}{\partial _1(\Phi ^a+\Psi _i)}\big ((\textbf{w}[\textbf{u}^a+\textbf{u}_i]\cdot \nabla )(\textbf{H}^a+\textbf{H}_i)\\&-(\textbf{h}[\textbf{H}^a+\textbf{H}_i]\cdot \nabla )(\textbf{u}^a+\textbf{u}_i)+(\textbf{H}^a+\textbf{H}_i)\textrm{div}(\textbf{v}[\textbf{u}^a+\textbf{u}_i]))\big )\Big \}\\&-\Big \{\partial _t(\textbf{H}^a+S_{\theta _i}{} \textbf{H}_i)+\frac{1}{\partial _1(\Phi ^a+S_{\theta _i}\Psi _{i})}\left( \textbf{w}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_i]\cdot \nabla \right) (\textbf{H}^a+S_{\theta _i}{} \textbf{H}_i)\\&-(\textbf{h}[\textbf{H}^a+S_{\theta _i}{} \textbf{H}_i]\cdot \nabla )(\textbf{u}^a+S_{\theta _i}{} \textbf{u}_i)+(\textbf{H}^a+S_{\theta _i}{} \textbf{H}_i)\textrm{div}(\textbf{v}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_i]))\Big \}, \end{aligned} \end{aligned}$$

where

$$\begin{aligned} \begin{aligned}&\textbf{v}[\textbf{u}^a+\textbf{u}_i]:=\Bigg ((\textbf{u}^a+\mathbf{u_i})\cdot \Big (1,-\partial _2(\Psi ^a+\Psi _i)\Big ),(\textbf{u}^a+\textbf{u}_i)_2\partial _1(\Phi ^a+\Psi _i)\Bigg ),\\&\textbf{w}[\textbf{u}^a+\textbf{u}_i]:=\textbf{v}[\textbf{u}^a+\textbf{u}_i]-\Big (\partial _t(\Psi ^a+\Psi _i),0\Big ),\\&\textbf{h}[\textbf{H}^a+\textbf{H}_i]:=\Bigg ((\textbf{H}^a+\mathbf{H_i})\cdot \Big (1,-\partial _2(\Psi ^a+\Psi _i)\Big ),(\textbf{H}^a+\textbf{H}_i)_2\partial _1(\Phi ^a+\Psi _i)\Bigg ).\\ \end{aligned} \end{aligned}$$

The vectors \(\textbf{v}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_i]\), \(\textbf{w}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_i]\), \(\textbf{h}[\textbf{H}^a+S_{\theta _i}{} \textbf{H}_i]\) are defined by completely similar expressions with \(S_{\theta _i}{} \textbf{u}_i\), \(S_{\theta _i}{} \textbf{H}_i\), \(S_{\theta _i}\Psi _i\) instead of \(\textbf{u}_i\), \(\textbf{H}_i\), \(\Psi _i\).

We split the range of s into small and large values (in order to take advantage of the continuity estimates of \(I-S_{\theta _i}\), see (9.2) and (10.4)). For small values of s, by using (10.3), (10.4), the smallness assumption (10.1) and Moser-type calculus inequalities we get

$$\begin{aligned}{} & {} ||\textbf{v}[\textbf{u}^a+\textbf{u}_{i}]||_{s,*,T}\le C\delta ^2 {\left\{ \begin{array}{ll} \theta _i^{(s+2-\alpha )_+}\quad &{}s+2\not =\alpha ,\\ \theta _i \quad &{}s+2=\alpha , \end{array}\right. }\qquad s\in \{6,\dots ,{\tilde{\alpha }}-2\}, \nonumber \\ \end{aligned}$$
(10.34)
$$\begin{aligned}{} & {} ||\textbf{v}[\textbf{u}^a+S_{\theta _i}\textbf{u}_{i}]-\textbf{v}[\textbf{u}^a+\textbf{u}_{i}]||_{s,*,T}\le C\delta \theta _i^{s+2-\alpha },\qquad s\in \{6,\dots ,{\tilde{\alpha }}-2\},\,\, \alpha \ge 9.\nonumber \\ \end{aligned}$$
(10.35)

Adding (10.34), (10.35) gives an estimate of \(\textbf{v}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_{i}]\). We obtain similar estimates for \(\textbf{w}[\textbf{u}^a+\textbf{u}_{i}],\textbf{h}[\textbf{H}^a+\textbf{H}_{i}], \textbf{w}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_{i}]\) and \(\textbf{h}[\textbf{H}^a+S_{\theta _i}{} \textbf{H}_{i}]\).

We decompose \(\Delta _1\) as sum of terms with differences \(I-S_{\theta _i}\) put in evidence. Making repeated use of estimates (9.1), (9.2), (10.3)–(10.5), the smallness assumption (10.1), Moser-type calculus inequalities of Lemmata 3.13.2, we get, for \(\alpha \ge 9\),

$$\begin{aligned} ||\Delta _1||_{s,*,T}\le C\delta \theta _i^{s+4-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,\alpha -3\}. \end{aligned}$$
(10.36)

For large values \(s\ge {\alpha }-3\) we use (9.1), (10.5) to obtain

$$\begin{aligned} ||{\textbf {v}}[{\textbf {u}}^a+S_{\theta _i}{} {\textbf {u}}_{i}]||_{s,*,T}\le C\delta ^2 {\left\{ \begin{array}{ll} \theta _i^{(s+2-\alpha )_+}\quad &{}{}s+2\not =\alpha ,\\ \theta _i \quad &{}{}s+2=\alpha , \end{array}\right. }\qquad s\in \{6,\dots ,{\tilde{\alpha }}+6\},\nonumber \\ \end{aligned}$$
(10.37)

with similar estimates for \(\textbf{w}[\textbf{u}^a+S_{\theta _i}\textbf{u}_{i}]\) and \(\textbf{h}[\textbf{H}^a+S_{\theta _i}{} \textbf{H}_{i}]\). With a direct estimate of the terms in \(\Delta _1\) we extend (10.36) to the cases \(s\ge {\alpha }-2\) and finally obtain

$$\begin{aligned} ||\Delta _1||_{s,*,T}\le C\delta \theta _i^{s+4-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,{\tilde{\alpha }}+4\}. \end{aligned}$$
(10.38)

In the above estimate (10.37) it is fundamental that \(s\le {\tilde{\alpha }}+6\). This is used for the estimate (10.38) of \(\Delta _1\), where the estimate of the normal derivative of \({\textbf{v}}[\cdot ]\) in the anisotropic space \(H^s_*\), for \(s\le {\tilde{\alpha }}+4\), requires an estimate of \({\textbf{v}}[\cdot ]\) in \(H^{s+2}_*\).

As for \(\Delta _2\), we have the explicit expression

$$\begin{aligned} \begin{aligned} \Delta _2=&\frac{1}{\partial _1(\Phi ^a+\Psi _{i+1/2})}\left\{ ({\textbf {w}}^{i+1/2}\cdot \nabla )({\textbf {H}}^a+S_{\theta _i}{} {\textbf {H}}_i)\right. \\ {}&\qquad -({\textbf {h}}[{\textbf {H}}^a+S_{\theta _i}{} {\textbf {H}}_i]\cdot \nabla )(S_{\theta _i}{} {\textbf {u}}_i-{\textbf {u}}_{i+1/2})\\&\qquad \left. +({\textbf {H}}^a+S_{\theta _i}{} {\textbf {H}}_i)\text {div}({\textbf {v}}[S_{\theta _i}{} {\textbf {u}}_i-{\textbf {u}}_{i+1/2}])\right\} \end{aligned} \end{aligned}$$

where

$$\begin{aligned}&\textbf{v}[S_{\theta _i}{} \textbf{u}_i-\textbf{u}_{i+1/2}]:=\textbf{v}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_i]-\textbf{v}[\textbf{u}^a+\textbf{u}_{i+1/2}]\nonumber \\&=\Bigg ((S_{\theta _i}{} \mathbf{u_i}-\textbf{u}_{i+1/2})\cdot \Big (1,-\partial _2(\Psi ^a+\Psi _{i+1/2})\Big ),\nonumber \\&\quad (S_{\theta _i}{} \textbf{u}_i-\textbf{u}_{i+1/2})_2\partial _1(\Phi ^a+\Psi _{i+1/2})\Bigg ),\nonumber \\&\textbf{w}^{i+1/2}:=\textbf{w}[\textbf{u}^a+S_{\theta _i}{} \textbf{u}_i]-\textbf{w}[\textbf{u}^a+\textbf{u}_{i+1/2}]=\textbf{v}[S_{\theta _i}{} \textbf{u}_i-\textbf{u}_{i+1/2}]. \end{aligned}$$
(10.39)

From the definition (10.39), using Moser-type calculus inequalities, (10.1), (10.5), (10.15), and the estimate (10.30), we obtain

$$\begin{aligned} ||{\textbf {v}}[S_{\theta _i}{} {\textbf {u}}_i-{\textbf {u}}_{i+1/2}]||_{s,*,T}\le C\delta \theta _i^{s+1-\alpha },\,\,\,\, s\in \{6,\dots ,{\tilde{\alpha }}+6\},\,\, \alpha \ge 9.\qquad \end{aligned}$$
(10.40)

Again in (10.40) we need \(s\le {\tilde{\alpha }}+6\) in order to get the following estimate (10.41) of \(\Delta _2\), where the estimate of the normal derivative of \({\textbf{v}}[\cdot ]\) in the anisotropic space \(H^s_*\), for \(s\le {\tilde{\alpha }}+4\), requires an estimate of \({\textbf{v}}[\cdot ]\) in \(H^{s+2}_*\).

Making repeated use of Moser-type calculus inequalities, (10.1), (10.5), (10.15), the estimates (10.30) and (10.40), we get for \(\alpha \ge 9\)

$$\begin{aligned} ||\Delta _2||_{s,*,T}\le C\delta \theta _i^{s+3-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,{{\tilde{\alpha }}}+4\}. \end{aligned}$$
(10.41)

To estimate the last term of \(F_H^{i+1/2}\) we write

$$\begin{aligned} \begin{aligned}&S_{\theta _i}{\mathbb {L}}_H({{\textbf{u}}}^a+{\textbf{u}}_{i},{{\textbf{H}}}^a+{{\textbf{H}}}_i,\Psi ^a+\Psi _{i}) =S_{\theta _i}\big ({\mathcal {L}}_H({\textbf{u}}_{i},{{\textbf{H}}}_i,\Psi _{i})+{\mathbb {L}}_H({{\textbf{u}}}^a,{{\textbf{H}}}^a,\Psi ^a)\big )\\&=S_{\theta _i}\big ({\mathcal {L}}_H({\textbf{u}}_{i},{{\textbf{H}}}_i,\Psi _{i})-{\mathcal {F}}^a_H\big )\\&=S_{\theta _i}\big ({\mathcal {L}}_H({\textbf{u}}_{i-1}+\delta {\textbf{u}}_{i-1},{{\textbf{H}}}_{i-1}+\delta {{\textbf{H}}}_{i-1},\Psi _{i-1}+\delta \Psi _{i-1})-{\mathcal {L}}_H({\textbf{u}}_{i-1},{{\textbf{H}}}_{i-1},\Psi _{i-1})\big )\\&\quad +S_{\theta _i}\big ({\mathcal {L}}_H({\textbf{u}}_{i-1},{{\textbf{H}}}_{i-1},\Psi _{i-1})-{\mathcal {F}}^a_H\big ), \end{aligned}\nonumber \\ \end{aligned}$$
(10.42)

where \({\mathcal {L}}_H,{\mathcal {F}}^a_H\) are the \({\textbf{H}}\)-component in (8.11), (8.10), respectively. From (9.1), \((H_{i-1})(b)\) we readily obtain

$$\begin{aligned} ||S_{\theta _i}\big ({\mathcal {L}}_H({\textbf{u}}_{i-1},{{\textbf{H}}}_{i-1},\Psi _{i-1})-{\mathcal {F}}^a_H\big )||_{s,*,T}\le C\delta \theta _i^{s-\alpha -1} \quad \text{ for }\,\,\,s\in \{6,\dots ,{{\tilde{\alpha }}}-2\}. \nonumber \\ \end{aligned}$$
(10.43)

For the first difference in (10.42) let us denote

$$\begin{aligned} \begin{aligned} \Delta _3:=&{\mathcal {L}}_H({\textbf{u}}_{i-1}+\delta {\textbf{u}}_{i-1},{{\textbf{H}}}_{i-1}+\delta {{\textbf{H}}}_{i-1},\Psi _{i-1}+\delta \Psi _{i-1})\\&-{\mathcal {L}}_H({\textbf{u}}_{i-1},{{\textbf{H}}}_{i-1},\Psi _{i-1}). \end{aligned} \end{aligned}$$

Hence we have

$$\begin{aligned} \begin{aligned} ||S_{\theta _i}\Delta _3||_{s,*,T} \le C\theta _i^{s-6}||\Delta _3||_{6,*,T} \le C\theta _i^{s-6}\cdot \delta \theta _i^{8-\alpha }=C\delta \theta _i^{s+2-\alpha }, \quad \text{ for }\,\,\,s\ge 6, \end{aligned} \nonumber \\ \end{aligned}$$
(10.44)

where, in particular, for the estimate of \(||\Delta _3||_{6,*,T}\), we have used \((H_{i-1})(a)\), (10.3), (10.34) and

$$\begin{aligned} ||\textbf{v}[\textbf{u}^a+{\textbf{u}}_{i-1}+\delta {\textbf{u}}_{i-1}]-\textbf{v}[\textbf{u}^a+\textbf{u}_{i-1}]||_{s,*,T}\le C\delta \theta _i^{s-\alpha },\qquad s\in \{6,\dots ,{{\tilde{\alpha }}}-2\}, \end{aligned}$$

used in the cases \(s=6\), \(s=8\) and a similar estimate for \(\textbf{w}[\textbf{u}^a+{\textbf{u}}_{i-1}+\delta {\textbf{u}}_{i-1}]-\textbf{w}[\textbf{u}^a+\textbf{u}_{i-1}]\) in case \(s=6\). Collecting (10.38), (10.41), (10.43), (10.44) we get

$$\begin{aligned} ||F_H^{i+1/2}||_{s,*,T}\le C\delta \theta _i^{s+4-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,{\tilde{\alpha }}+4\}. \end{aligned}$$
(10.45)

Equation (10.32) solved by \({\textbf{H}}_{i+1/2}-S_{\theta _i}{\textbf{H}}_i\) has the form

$$\begin{aligned} \partial _t Y+\sum _{j=1}^2{\mathcal {D}}_j(b)\partial _j Y +{\mathcal {Q}}(b)Y=F_H^{i+1/2} \end{aligned}$$
(10.46)

for \(Y={\textbf{H}}_{i+1/2}-S_{\theta _i}{\textbf{H}}_i\), \(b=({{\textbf{u}}}^a+{\textbf{u}}_{i+1/2},\Psi ^a+\Psi _{i+1/2})\), and where \({\mathcal {D}}_j\) and \({\mathcal {Q}}\) are some matrices. The matrices \({\mathcal {D}}_j\) are diagonal and, more important, \({\mathcal {D}}_1\) vanishes at the boundary. This yelds that system (10.46) does not need any boundary condition. A standard energy argument applied to (10.46) gives, in view of (10.45),

$$\begin{aligned} ||{\textbf{H}}_{i+1/2}-S_{\theta _i}{\textbf{H}}_i||_{s,*,T} \le C||F_H^{i+1/2}||_{s,*,T}\le C\delta \theta _i^{s+4-\alpha },\quad \text{ for }\,\,\,s\in \{6,\dots ,{\tilde{\alpha }}+4\}. \nonumber \\ \end{aligned}$$
(10.47)

Collecting (10.30), (10.47) gives (10.17).

Using (8.9), (10.5), (10.17) for \(s=6\), and taking \(\delta >0\) sufficiently small we have that \(\varphi ^a+\psi _{i+\frac{1}{2}}\) and \(\tilde{{{\textbf{U}}}}^a+{\textbf{V}}_{i+\frac{1}{2}}\) are sufficiently small. Then, \(\varphi ^a+\psi _{i+\frac{1}{2}}\) satisfies (4.7) and, recalling Remark 3.1, \({{{\textbf{U}}}}^a+{\textbf{V}}_{i+\frac{1}{2}}\) satisfies (4.2), (4.4), (4.6), (6.1). \({{{\textbf{H}}}}^a+{\textbf{H}}_{i+\frac{1}{2}}\) satifies (4.8) by construction, see (10.31); the initial value at \(t=0\) of \({{{\textbf{H}}}}^a+{\textbf{H}}_{i+\frac{1}{2}}\) satisfies (4.10) since \({{{\textbf{H}}}}^a_{|t=0}={{\textbf{H}}}_0\) satisfies (4.10) by assumption, and \({{{\textbf{H}}}}_{i+\frac{1}{2}}=0\) for \(t\le 0\) by continuity. In conclusion, \({{{\textbf{U}}}}^a+{\textbf{V}}_{i+\frac{1}{2}}\) satisfies all the constraints (4.2)–(4.8), (4.10) and (6.1) for the background state. \(\square \)

10.2.4 Estimate of the second substitution errors

In the following Lemma, we can estimate the second substitution errors \(e'''_k,{\tilde{e}}'''_k\) of the iterative scheme. We define

$$\begin{aligned}{} & {} e'''_k:={\mathcal {L}}'(S_{\theta _k}{{\textbf{V}}}_k,S_{\theta _k}\Psi _k)(\delta {{\textbf{V}}}_k,\delta \Psi _k)-{\mathcal {L}}'({{\textbf{V}}}_{k+\frac{1}{2}},\Psi _{k+\frac{1}{2}})(\delta {{\textbf{V}}}_k,\delta \Psi _k), \nonumber \\ \end{aligned}$$
(10.48)
$$\begin{aligned}{} & {} \begin{aligned} {\tilde{e}}'''_k:=&{\mathcal {B}}'(S_{\theta _k}{{\textbf{V}}}_k|_{x_1=0},S_{\theta _k}\psi _k)((\delta {{\textbf{V}}}_k)|_{x_1=0},\delta \psi _k)\\&-{\mathcal {B}}'({{\textbf{V}}}_{k+\frac{1}{2}}|_{x_1=0},\psi _{k+\frac{1}{2}})((\delta {{\textbf{V}}}_k)|_{x_1=0},\delta \psi _k).\\ \end{aligned} \end{aligned}$$
(10.49)

We can write (10.48) and (10.49) as follows:

$$\begin{aligned} \begin{aligned} e'''_k&=\int ^1_0{\mathbb {L}}''({\textbf{U}}^a+{{\textbf{V}}}_{k+\frac{1}{2}}+\tau (S_{\theta _k}{{\textbf{V}}}_k-{{\textbf{V}}}_{k+\frac{1}{2}}),\Psi ^a+S_{\theta _k}\Psi _k)\\&\quad ((\delta {{\textbf{V}}}_k,\delta \Psi _k),(S_{\theta _k}{{\textbf{V}}}_k-{{\textbf{V}}}_{k+\frac{1}{2}},0))\textrm{d}\tau ,\\&{\tilde{e}}'''_k={\mathbb {B}}''((\delta {{\textbf{V}}}_k|_{x_1=0},\delta \psi _k),((S_{\theta _k}{{\textbf{V}}}_k-{{\textbf{V}}}_{k+\frac{1}{2}})|_{x_1=0},0)). \end{aligned} \end{aligned}$$

Lemma 10.7

Let \(\alpha \ge 10\). There exist \(\delta >0\) sufficiently small and \(\theta _0\ge 1\) sufficiently large such that, for all \(k\in \{0,\cdots ,i-1\}\) and for all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\},\) we have

$$\begin{aligned}{} & {} ||e'''_k||_{s,*,T}\lesssim \delta ^2\theta ^{L_3(s)-1}_k\Delta _k, \end{aligned}$$
(10.50)
$$\begin{aligned}{} & {} ||{\tilde{e}}'''_k||_{H^s(\Gamma _T)}\lesssim \delta ^2\theta ^{L_3(s)-1}_k\Delta _k, \end{aligned}$$
(10.51)

where \(L_3(s):=\max \{(s+2-\alpha )_++16-2\alpha ;s+12-2\alpha \}.\)

Proof

Using (10.17) and Lemma 10.3, similar to the proof of Lemmata 10.4 and 10.5, we obtain (10.50) and (10.51). Here, we can calculate the explicit form of \({\tilde{e}}'''_k\) as

$$\begin{aligned} {\tilde{e}}'''_k:=\left[ \begin{array}{c} 0\\ 0\\ \delta H^+_k\cdot (S_{\theta _k}H^+_k-H^+_{k+\frac{1}{2}})-\delta H^-_k\cdot (S_{\theta _k}H^-_k-H^-_{k+\frac{1}{2}}) \end{array}\right] . \end{aligned}$$
(10.52)

\(\square \)

10.2.5 Estimate of the last error term

We now estimate the last error term (9.11):

$$\begin{aligned} D_{k+\frac{1}{2}}\delta \Psi _k=\frac{\delta \Psi _k}{\partial _1(\Phi ^a+\Psi _{k+\frac{1}{2}})}R_k, \end{aligned}$$
(10.53)

where \(R_k:=\partial _1[{\mathbb {L}}({{\textbf{U}}}^a+{{\textbf{V}}}_{k+\frac{1}{2}},\Psi ^a+\Psi _{k+\frac{1}{2}})].\) It is noted that

$$\begin{aligned} |\partial _1(\Phi ^a+\Psi _{k+\frac{1}{2}})|=|\pm 1+\partial _1(\Psi ^a+\Psi _{k+\frac{1}{2}})|\ge \frac{1}{2}, \end{aligned}$$

for \(\delta >0\) sufficiently small.

The following Lemma 10.8 can be proved by direct calculations:

Lemma 10.8

Let \(\alpha \ge 10,{\tilde{\alpha }}\ge \alpha -4.\) There exist \(\delta >0\) sufficiently small and \(\theta _0\ge 1\) sufficiently large, such that, for all \(k\in \{0,\cdots ,i-1\}\) and for all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\}\), we have

$$\begin{aligned} ||D_{k+\frac{1}{2}}\delta \Psi _k||_{s,*,T}\lesssim \delta ^2\theta ^{L_4(s)-1}_k\Delta _k, \end{aligned}$$
(10.54)

where \(L_4(s):=s+14-2\alpha \).

Proof

Using (3.7) and (3.9), we obtain that

$$\begin{aligned} \begin{aligned} ||D_{k+\frac{1}{2}}\delta \Psi _k||_{s,*,T}&\lesssim ||\delta \Psi _k||_{4,*,T}||R_k||_{s,*,T}+||\delta \Psi _k||_{s,*,T}||R_k||_{4,*,T}\\&\quad +||\delta \Psi _k||_{4,*,T}||R_k||_{4,*,T}||\Psi ^a+\Psi _{k+\frac{1}{2}}||_{s+2,*,T}. \end{aligned} \nonumber \\ \end{aligned}$$
(10.55)

Using (8.10) and (8.11), we can write

$$\begin{aligned} R_k=\partial _1({\mathcal {L}}({{\textbf{V}}}_k,\Psi _k)-{\mathcal {F}}^a+{\mathcal {I}}), \text { for }t>0, \end{aligned}$$
(10.56)

where

$$\begin{aligned} {\mathcal {I}}:= & {} \int ^1_0{\mathbb {L}}'({\textbf{U}}^a+{{\textbf{V}}}_k+\tau ({{\textbf{V}}}_{k+\frac{1}{2}}-{{\textbf{V}}}_k),\Psi ^a+\Psi _k+\tau (\Psi _{k+\frac{1}{2}}-\Psi _k))\\ {}{} & {} ({{\textbf{V}}}_{k+\frac{1}{2}}-{{\textbf{V}}}_k,\Psi _{k+\frac{1}{2}}-\Psi _k)\textrm{d}\tau . \end{aligned}$$

If \(s\in \{4,\cdots ,{\tilde{\alpha }}-4\},\) then using Hypothesis \((H_{i-1}),\) we obtain that

$$\begin{aligned} ||{\mathcal {L}}({{\textbf{V}}}_k,\Psi _k)-{\mathcal {F}}^a||_{s+2,*,T}\le 2\delta \theta ^{s+1-\alpha }_k. \end{aligned}$$
(10.57)

Using (10.3), (10.4), (10.15)–(10.17), (10.56),(10.57), Sobolev inequalities (3.10) and Moser-type calculus inequalities in Lemma 3.2, we have

$$\begin{aligned} ||R_k||_{s,*,T}\lesssim \delta (\theta ^{s+8-\alpha }_k+\theta ^{(s+4-\alpha )_++10-\alpha }_k), \text { for } s\in \{4,\cdots , {\tilde{\alpha }}-4\}.\nonumber \\ \end{aligned}$$
(10.58)

If \(s\in \{{\tilde{\alpha }}-3,{\tilde{\alpha }}-2\}\), then, for \({{\tilde{\alpha }}}\ge \alpha -4\), using (10.3), (10.5) and (10.15)–(10.17), we can deduce directly from (10.56) that

$$\begin{aligned} ||R_k||_{s,*,T}\lesssim \delta \theta ^{s+8-\alpha }_k. \end{aligned}$$

Therefore, (10.58) holds for all \(s\in \{4,\cdots ,{\tilde{\alpha }}-2\}.\) Using Hypothesis \((H_{i-1}),\) (10.58), (10.5), (10.3) and (10.15)–(10.17) into (10.55), we can obtain (10.54). \(\square \)

Using Lemmata 10.410.8, we can conclude the following estimates of the error terms \(e_k, {\tilde{e}}_k,\) defined by (9.12):

Lemma 10.9

Let \(\alpha \ge 10,{\tilde{\alpha }}\ge \alpha -4.\) There exist \(\delta >0\) sufficiently small and \(\theta _0\ge 1\) sufficiently large, such that for all \(k\in \{0,\cdots ,i-1\}\) and for all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\}\), we have

$$\begin{aligned} ||e_k||_{s,*,T}+||{\tilde{e}}_k||_{H^s(\Gamma _T)}\lesssim \delta ^2\theta ^{L_4(s)-1}_k\Delta _k, \end{aligned}$$
(10.59)

where \(L_4(s)\) is defined in Lemma 10.8.

From Lemma 10.9, we obtain the estimate of the accumulated errors \(E_i,{\tilde{E}}_i,{\hat{E}}_i,\) which are defined in (9.13).

Lemma 10.10

Let \(\alpha \ge 16,{\tilde{\alpha }}=\alpha +5.\) There exist \(\delta >0\) sufficiently small and \(\theta _0\ge 1\) sufficiently large, such that

$$\begin{aligned} ||E_i||_{\alpha +3,*,T}+||{\tilde{E}}_i||_{H^{\alpha +3}(\Gamma _T)}\lesssim \delta ^2\theta _i. \end{aligned}$$
(10.60)

Proof

Using \(L_4(\alpha +3)\le 1\) if \(\alpha \ge 16\), it follows from (10.59) that

$$\begin{aligned} ||E_i||_{\alpha +3,*,T}\lesssim \sum ^{i-1}_{k=0}||e_k||_{\alpha +3,*,T}\lesssim \sum ^{i-1}_{k=0}\delta ^2\Delta _k\lesssim \delta ^2\theta _i, \end{aligned}$$

if \(\alpha +3\le {\tilde{\alpha }}-2\). Similar arguments also hold for \(||{\tilde{E}}_i||_{H^{\alpha +3}(\Gamma _T)}\). The minimal possible \({\tilde{\alpha }}\) is \(\alpha +5\). \(\square \)

10.3 Convergence of the iteration scheme

We still need to estimate the source terms \(f_{i},g_{i}.\)

Lemma 10.11

Let \(\alpha \ge 16\) and \({\tilde{\alpha }}=\alpha +5\). There exist \(\delta >0\) sufficiently small and \(\theta _0\ge 1\) sufficiently large, such that for all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}+2\},\)

$$\begin{aligned}{} & {} ||f_i||_{s,*,T}\lesssim \Delta _i\Big (\theta ^{s-\alpha -3}_i(||{\mathcal {F}}^a||_{\alpha +2,*,T}+\delta ^2)+\delta ^2\theta ^{L_4(s)-1}_i\Big ), \end{aligned}$$
(10.61)
$$\begin{aligned}{} & {} ||g_i||_{H^{s}(\Gamma _T)}\lesssim \delta ^2\Delta _i\Big (\theta ^{s-\alpha -3}_i+\theta ^{L_4(s)-1}_i\Big ). \end{aligned}$$
(10.62)

In the above inequalities we need the exponent \(s-\alpha -3\) of \(\theta _i\) to compensate the loss of 2 derivatives for the data in (10.64), in order to recover the exponent \(s-\alpha -1\) in the corresponding terms of (10.68).

Proof

As in [14, 34, 39], using (9.1)–(9.3), (9.14), (10.59), (10.60), we obtain that

$$\begin{aligned} \begin{aligned} ||f_i||_{s,*,T}&\le ||(S_{\theta _i}-S_{\theta _{i-1}}){\mathcal {F}}^a-(S_{\theta _i}-S_{\theta _{i-1}})E_{i-1}-S_{\theta _i}e_{i-1}||_{s,*,T}\\&\lesssim \Delta _i\theta ^{s-\alpha -3}_i(||{\mathcal {F}}^a||_{\alpha +2,*,T}+\theta ^{-1}_i||E_{i-1}||_{\alpha +3,*,T})+||S_{\theta _i}e_{i-1}||_{s,*,T}\\&\lesssim \Delta _i\{\theta ^{s-\alpha -3}_i(||{\mathcal {F}}^a||_{\alpha +2,*,T}+\delta ^2)+\delta ^2\theta ^{L_4(s)-1}_i\}. \end{aligned} \end{aligned}$$

Using (10.59) and (10.60) we can obtain (10.62). \(\square \)

Similar to the proof of [14, 34, 39], we can obtain the estimate of \((\delta {{\textbf{V}}}_i,\delta \Psi _i)\) by (10.17) and the tame estimate (7.1) applied to problem (9.6).

Lemma 10.12

Let \(\alpha \ge 16\) and \({\tilde{\alpha }}=\alpha +5\). If \(\delta >0\) and \(||{\mathcal {F}}^a||_{\alpha +2,*,T}/\delta \) are sufficiently small and \(\theta _0\ge 1\) is sufficiently large, then for all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}\}\),

$$\begin{aligned} ||(\delta {{\textbf{V}}}_i,\delta \Psi _i)||_{s,*,T}+||\delta \psi _i||_{H^{s}(\Gamma _T)}\le \delta \theta ^{s-\alpha -1}_i\Delta _i. \end{aligned}$$
(10.63)

Proof

Let us consider problem (9.6) that will be solved by applying Theorem 7.1. We first notice that \(({{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \Psi ^a+\Psi _{i+\frac{1}{2}}, \varphi ^a+\psi _{i+\frac{1}{2}})\) satisfy the constraints (4.2)–(4.8), (4.10), (6.1). Thus we may apply our tame estimate (7.1) and obtain

$$\begin{aligned} \begin{aligned}&||\delta \dot{{{\textbf{V}}}_i}||_{s,*,T}+||\delta \psi _i||_{H^s(\Gamma _T)}\\&\le C\Big (||{f}_i||_{s+2,*,T}+||{g}_i||_{H^{s+2}(\Gamma _T)}\\&\quad +(||{f}_i||_{8,*,T}+||{g}_i||_{H^{8}(\Gamma _T)})||(\tilde{{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \nabla (\Psi ^a+\Psi _{i+\frac{1}{2}}))||_{s+4,*,T}\Big ). \end{aligned} \nonumber \\ \end{aligned}$$
(10.64)

On the other hand, from (9.7) it follows that

$$\begin{aligned} \begin{aligned}&||\delta {{{\textbf{V}}}_i}||_{s,*,T} \le ||\delta \dot{{{\textbf{V}}}_i}||_{s,*,T}+ C||\delta {{\mathbf \Psi }_i}||_{s,*,T}||(\tilde{{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \Psi ^a+\Psi _{i+\frac{1}{2}})||_{6,*,T}\\&\quad \qquad + C||\delta {{\mathbf \Psi }_i}||_{4,*,T}||(\tilde{{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \Psi ^a+\Psi _{i+\frac{1}{2}})||_{s+2,*,T}. \end{aligned} \nonumber \\ \end{aligned}$$
(10.65)

From

$$\begin{aligned} ||\delta {{\mathbf \Psi }_i}||_{s,*,T}\le ||\delta \psi _i||_{H^s(\Gamma _T)}, \end{aligned}$$
(10.66)

(10.64) for \(s=4\) and (6.1) we have

$$\begin{aligned} \begin{aligned}&||\delta {{\mathbf \Psi }_i}||_{4,*,T}\le ||\delta \psi _i||_{H^4(\Gamma _T)} \le C\Big (||{f}_i||_{6,*,T}+||{g}_i||_{H^{6}(\Gamma _T)}\\&\quad +(||{f}_i||_{8,*,T}+||{g}_i||_{H^{8}(\Gamma _T)})||(\tilde{{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \nabla (\Psi ^a+\Psi _{i+\frac{1}{2}}))||_{8,*,T}\Big )\\&\le C (||{f}_i||_{8,*,T}+||{g}_i||_{H^{8}(\Gamma _T)}). \end{aligned} \end{aligned}$$

Then, from (10.64) and (10.65) we obtain

$$\begin{aligned} \begin{aligned}&||\delta {{{\textbf{V}}}_i}||_{s,*,T}+||\delta \psi _i||_{H^s(\Gamma _T)}\\&\le C\Big (||{f}_i||_{s+2,*,T}+||{g}_i||_{H^{s+2}(\Gamma _T)}\\&\quad +(||{f}_i||_{8,*,T}+||{g}_i||_{H^{8}(\Gamma _T)})||(\tilde{{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \nabla (\Psi ^a+\Psi _{i+\frac{1}{2}}))||_{s+4,*,T}\Big )\\&\quad +||\delta \psi _i||_{H^s(\Gamma _T)}||(\tilde{{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \Psi ^a+\Psi _{i+\frac{1}{2}})||_{6,*,T}. \end{aligned} \end{aligned}$$

Using (8.9), (10.5) and (10.17) for \(s=6\), taking \(\delta >0\) sufficiently small, we can absorb the last term in the right-hand side above into the left-hand side to get

$$\begin{aligned} \begin{aligned}&||\delta {{{\textbf{V}}}_i}||_{s,*,T}+||\delta \psi _i||_{H^s(\Gamma _T)}\\&\le C\Big (||{f}_i||_{s+2,*,T}+||{g}_i||_{H^{s+2}(\Gamma _T)}\\&\quad +(||{f}_i||_{8,*,T}+||{g}_i||_{H^{8}(\Gamma _T)})||(\tilde{{\textbf{U}}}^a+{{\textbf{V}}}_{i+\frac{1}{2}}, \nabla (\Psi ^a+\Psi _{i+\frac{1}{2}}))||_{s+4,*,T}\Big ). \end{aligned} \nonumber \\ \end{aligned}$$
(10.67)

The remaining part of the work is to estimate the right-hand side of (10.67).

Using Lemma 10.11, (10.5), Lemma 10.6 and (10.66), (10.67) becomes

$$\begin{aligned} \begin{aligned}&||(\delta {{{\textbf{V}}}_i},\delta {{\mathbf \Psi }_i})||_{s,*,T}+||\delta \psi _i||_{H^s(\Gamma _T)}\\&\le C\Delta _i\Big (\theta _i^{s-\alpha -1}(||{\mathcal {F}}^a||_{\alpha +2,*,T}+\delta ^2)+\delta ^2\theta _i^{L_4(s+2)-1}\Big )\\&\quad +C\Delta _i\Big (\theta _i^{5-\alpha }(||{\mathcal {F}}^a||_{\alpha +2,*,T}+\delta ^2)+\delta ^2\theta _i^{21-2\alpha }\Big )\Big (\delta \theta _i^{s+8-\alpha }+\delta \theta _i^{(s+6-\alpha )_+}\Big ). \end{aligned}\nonumber \\ \end{aligned}$$
(10.68)

One checks that, for \(\alpha \ge 16\) and \(6\le s\le {\tilde{\alpha }}\), the following inequalities hold true:

$$\begin{aligned} \begin{aligned}&L_4(s+2)\le s-\alpha ,\qquad 5-\alpha +(s+6-\alpha )_+\le s-\alpha -1, \\&s+13-2\alpha \le s-\alpha -1,\qquad 21-2\alpha +s+8-\alpha \le s-\alpha -1,\\&21-2\alpha +(s+6-\alpha )_+\le s-\alpha -1. \end{aligned} \end{aligned}$$

From (10.68) we thus obtain (10.63), provided \(\delta >0\) and \(||{\mathcal {F}}^a||_{\alpha +2,*,T}/\delta \) are sufficiently small and \(\theta _0\ge 1\) is sufficiently large. \(\square \)

Finally, similar to the proof of [14, 34, 39], we can obtain the remaining inequalities in \((H_i).\)

Lemma 10.13

Let \(\alpha \ge 16\). If \(\delta >0\) and \(||{\mathcal {F}}^a||_{\alpha +2,*,T}/\delta \) are sufficiently small and if \(\theta _0\ge 1\) is sufficiently large, then for all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\}\)

$$\begin{aligned} ||{\mathcal {L}}({{\textbf{V}}}_i,\Psi _i)-{\mathcal {F}}^a||_{s,*,T}\le 2\delta \theta ^{s-\alpha -1}_i. \end{aligned}$$
(10.69)

Moreover, for all integers \(s\in \{6,\cdots ,{\tilde{\alpha }}-2\}\)

$$\begin{aligned} ||{\mathcal {B}}({{\textbf{V}}}_i|_{x_1=0},\Psi _i)||_{H^{s}(\Gamma _T)}\le \delta \theta ^{s-\alpha -1}_i. \end{aligned}$$
(10.70)

Proof

Recall that, by summing the relations (9.9), we have

$$\begin{aligned} {\mathcal {L}}({{\textbf{V}}}_i,\Psi _i)-{\mathcal {F}}^a=(S_{\theta _{i-1}}-I){\mathcal {F}}^a+(I-S_{\theta _{i-1}})E_{i-1}+e_{i-1}. \end{aligned}$$
(10.71)

The proof of (10.69) then follows by applying (9.2), (10.59), (10.60), provided that \(\delta >0\) and \(||{\mathcal {F}}^a||_{\alpha +2,*,T}/\delta \) are sufficiently small and \(\theta _0\ge 1\) is sufficiently large. The proof of (10.70) is similar. \(\square \)

We are now in the position to prove the main theorem for the existence of the solution to the nonlinear problem (2.1).

Proof of Theorem 3.1

Let the initial data \(({{\textbf{U}}}^{\pm }_0,\varphi _0)\) satisfy all the assumptions of Theorem 3.1. Let \(\alpha =m+1\ge 16\), \({\tilde{\alpha }}=\alpha +5\), \(\mu =m+10\). Then the initial data \({{\textbf{U}}}^{\pm }_0\in H^{\mu +1.5}({\mathbb {R}}^2_+)\) and \(\varphi _0\in H^{\mu +1.5}({\mathbb {R}})\) are compatible up to order \(\mu \) and there exists an approximate solution \((\tilde{{\textbf{U}}}^a,\varphi ^a)\in H^{\mu +2}(\Omega _T)\times H^{\mu +2}(\Gamma _T)\) to problem (2.1). Observe that \(\mu +2={{\tilde{\alpha }}}+6\) as required in (10.1). We choose \(\delta >0\), \(T>0\) sufficiently small, \(\theta _0\ge 1\) sufficiently large as in the previous lemmata. We also assume \(T>0\) small enough so that \(||{\mathcal {F}}^a||_{\alpha +2,*,T}/\delta \) is sufficiently small. Then in view of Lemmata 10.110.1210.13, property (\(H_i\)) holds for all integers i. In particular, we have

$$\begin{aligned} \sum ^{\infty }_{i=0}(||(\delta {{\textbf{V}}}_i,\delta \Psi _i)||_{s,*,T}+||\delta \psi _i||_{H^{s}(\Gamma _T)})\lesssim \sum ^{\infty }_{i=0}\theta ^{s-\alpha -2}_i<\infty ,\quad \text {for}\,\,\,6\le s\le \alpha -1. \end{aligned}$$

Thus, the sequence \(({{\textbf{V}}}_i,\Psi _i)\) converges to some limit \(({{\textbf{V}}},\Psi )\) in \(H^{\alpha -1}_{*}(\Omega _T),\) and sequence \(\psi _i\) converges to some limit \(\psi \) in \(H^{\alpha -1}(\Gamma _T).\) Passing to the limit in (10.69) and (10.70) for \(s=m=\alpha -1\), we obtain (8.11). Therefore, \(({{\textbf{U}}},\Phi )=({{\textbf{U}}}^a+{{\textbf{V}}},\Phi ^a+\Psi )\) is a solution on \(\Omega _T\) of nonlinear system (2.1).