1. Introduction
This paper is concerned with the transonic shock problem in a two-dimensional nozzle with an external force. In 1948, Courant and Friedrichs [Reference Courant and Friedrichs9] had described the following transonic shock phenomena in a de Laval nozzle: given the appropriately large receiver pressure, if the upcoming flow is still supersonic after passing through the throat of the nozzle, to match the prescribed appropriately large exit pressure, a shock front intervenes at some place in the diverging part of the nozzle and the gas is compressed and slowed down to subsonic speed. In this paper, we will investigate such a problem for the two-dimensional compressible Euler flow exerted by proper external force in a two-dimensional flat nozzle.
The studies of transonic shock solutions for inviscid compressible flows in different kinds of nozzles have a long history. A lot of significant results have been achieved over the past two decades. People first used the quasi one-dimensional model to study the transonic shock problem [Reference Bers1, Reference Courant and Friedrichs9, Reference Embid, Goodman and Majda10, Reference Liu14]. In [Reference Liu14], Liu has proven that the flows along the expanding portion of the nozzle are stable, which suggests that the widening nozzle played a stabilizing effect. However, our results imply that the external force also has the effect of stabilizing the shock. Next, the structural stability of multidimensional transonic shocks in flat or diverging nozzles was further investigated in [Reference Chen and Feldman6, Reference Xin and Yin30, Reference Xin and Yin31] using the steady potential flows with different kinds of boundary conditions. In particular, [Reference Xin and Yin30, Reference Xin and Yin31] proved that the stability of transonic shocks for potential flows is usually ill-posed under the perturbations of the exit pressure. The structural stability of the transonic shock problem in two-dimensional divergent nozzles under the perturbations for the exit pressure was first established in [Reference Li, Xin and Yin16] when the opening angle of the nozzle is suitably small. Later on, this restriction was removed in [Reference Li, Xin and Yin17, Reference Li, Xin and Yin20]. Moreover, the transonic shock in general two-dimensional straight divergent nozzles was shown in [Reference Li, Xin and Yin20] to be structurally stable under generic perturbations for both the nozzle shape and the exit pressure. The existence and stability of transonic shock for three-dimensional axisymmetric flows without swirl in a conic straight nozzle were established in [Reference Li, Xin and Yin18, Reference Li, Xin and Yin19] with respect to small perturbations of the exit pressure. Compared with those results, a new decomposition method to the compressible Euler system was applied in this article.
Many researchers also considered the transonic shock problem in the flat or almost flat nozzles with the exit pressure satisfying some special constraint, see [Reference Chen3–Reference Chen5, Reference Liu and Yuan15, Reference Xin, Yan and Yin29] and the references therein. Recently, there has been interesting progress on the stability and existence of transonic shock solutions to the two-dimensional and three-dimensional axisymmetric steady compressible Euler system in an almost flat finite nozzle with the receiver pressure prescribed at the exit of the nozzle (see [Reference Fang and Xin11, Reference Fang and Gao12]), where the shock position was uniquely determined. For the structural stability under the axisymmetric perturbation of the nozzle wall, a modified Lagrangian coordinate was introduced in [Reference Weng, Xie and Xin26] to deal with the corner singularities near the intersection points of the shock surface and nozzle boundary and the artificial singularity near the axis simultaneously. Most recently, the authors in [Reference Weng, Xin and Yuan24, Reference Weng, Xin and Yuan25] studied radially symmetric transonic flow with/without shock in an annulus. Thanks to the effect of angular velocity, it was found in [Reference Weng, Xin and Yuan24] that besides the well-known supersonic-subsonic shock in a divergent nozzle as in the case without angular velocity, there exists a supersonic–supersonic shock solution, where the downstream state may change smoothly from supersonic to subsonic. Furthermore, there exists a supersonic–sonic shock solution where the shock circle and the sonic circle coincide.
The rest of this paper will be organized as follows. In the next section, we formulate the problem investigated in this article and state our main results. In § 3, we reformulate the original 2-D problem by deformation-curl decomposition developed in [Reference Weng and Xin27, Reference Weng28] so that one can rewrite system (2.1) with the velocity and the Bernoulli function. We obtain a $2\times 2$ first-order system for the velocity field, a transport-type equation for the Bernoulli function, and the first-order ordinary differential equation for the shock after linearization. In § 4, we design an elaborate iteration scheme inspired by the works [Reference Li, Xin and Yin17] for the non-linear system. The investigation of well-posedness and regularity for the linear system is given in the remainder part of this section. In § 5, we prove the main existence and uniqueness theorem.
2. Formulation of the problem and main results
In this section, we present the two-dimensional transonic shock problem with external force under suitable perturbation of exit pressure and the main results. The 2-D steady compressible isentropic Euler system with external force is of the form
where $(u_1,\,u_2)={\bf u}:\mathbb {R}^2\rightarrow \mathbb {R}^2$ is the unknown velocity filed and $\rho :\mathbb {R}^2\rightarrow \mathbb {R}$ is the density, and $\Phi (x_1,\,x_2)$ is a given potential function of external force. For the ideal polytropic gas, the equation of state is given by $P(\rho )=A \rho ^{\gamma }$, here $A$ and $\gamma$ ($1<\gamma <\infty$) are positive constants. We take $A=1$ throughout this paper for the convenience.
To this end, let's first focus on the 1-D steady compressible flow with an external force on an interval $I=[L_0,\, L_1]$, which is governed by
where we assume that the flow state at the entrance $x_1=L_0$ is supersonic, meaning that $u_0^2>c^2(\rho _0)=\gamma \rho _0^{\gamma -1}$.
Denote $J= \bar {\rho } \bar {u}=\rho _0 u_0>0$, then it follows from (2.2) that
Also one has
where $\bar {M}(x_1)={\bar {u}(x_1)}/{c(\bar {\rho })}$ is the Mach number.
Since $\bar {M}^2(L_0)>1$, it follows from (2.5) that if the external force satisfies
then problem (2.2) has a global supersonic solution $(\bar {\rho }^-,\, \bar {u}^-)$ on $[L_0,\,L_1]$. If one prescribes a large enough end pressure at $x_1=L_1$, a shock will form at some point $x_1=L_s\in (L_0,\,L_1)$ and the gas is compressed and slowed down to subsonic speed, the gas pressure will increase to match the given end pressure. Mathematically, one looks for a shock $x_1=L_s$ and smooth functions $(\bar {\rho }^{\pm },\, \bar {u}^{\pm },\, \bar {P}^{\pm })$ defined on $I^+=[L_{s},\, L_1]$ and $I^-=[L_0,\,L_s]$ respectively, which solves (2.3) on $I^{\pm }$ with the jump at the shock $x_1=L_{s}\in (L_0,\, L_1)$ satisfying the physical entropy condition $[\bar {P}(L_{s})]=\bar {P}^+(L_s)-\bar {P}^-(L_s)>0$ and the Rankine–Hugoniot conditions
and also the boundary conditions
We will show that there is a unique transonic shock solution to the 1-D Euler system when the end pressure $P_e$ lies in a suitable interval. Such a problem will be solved by a shooting method employing the monotonicity relation between the shock position and the end pressure.
Lemma 2.1 Suppose that the initial state $(u_0,\,\rho _0)$ at $x_1=L_0$ is supersonic and the external force $f$ satisfying (2.6), there exist two positive constants $P_0,\,P_1>0$ such that if the end pressure $P_e\in (P_1,\, P_0)$, there exists a unique transonic shock solution $(\bar {u}^-,\, \bar {\rho }^-)$ and $(\bar {u}^+,\,\bar {\rho }^+)$ defined on $I^-=[L_0,\,L_s)$ and $I^+=(L_s,\, L_1)$ respectively, with a shock located at $x_1=L_{s}\in (L_0,\,L_1)$. In addition, the shock position $x_1=L_s$ increases as the exit pressure $P_{e}$ decreases. Furthermore, the shock position $L_s$ approaches to $L_1$ if $P_{e}$ goes to $P_1$ and $L_s$ tends to $L_0$ if $P_{e}$ goes to $P_0$.
Proof. The existence and uniqueness of smooth supersonic flow $(\bar {u}^-,\, \bar {\rho }^-)$ starting from $(\rho _0,\,u_0)$ on $[L_0,\,L_1]$ is trivial. Suppose the shock occurs at $x_1=L_s\in (L_0,\, L_1)$, then it is well-known that there exists a unique subsonic state $(\bar {u}^+(L_s),\, \bar {\rho }^+(L_s))$ satisfying the Rankine–Hugoniot conditions (2.7) and the entropy condition. With $(\bar {u}^+(L_s),\, \bar {\rho }^+(L_s))$ as the initial data, equation (2.2) has a unique smooth solution $(\bar {u}^+,\, \bar {\rho }^+)$ on $[L_s,\,L_1]$. Denote $P_e= (\bar {\rho }^+(L_1))^{\gamma }$. In the following, we show that the monotonicity between the shock position $x_1=L_s$ and the exit pressure $P_e=(\bar {\rho }^+(L_1))^{\gamma }$. $\bar {\rho }^+(L_1)$ is regarded as a function of $L_s$. Since $(\bar {\rho }^+ \bar {u}^+)(L_s)=(\bar {\rho }^- \bar {u}^-)(L_s)=J=\rho _0 u_0>0$, then
It follows from the second equation in (2.2) that
Differentiating with respect to $L_s$, one deduces that
Also (2.10) yields that
Finally, we conclude that
Since the coefficients
then (2.11) implies that the end density $\bar \rho ^+(L_1)$ is a strictly decreasing function of the shock position $x_1=L_s$. It follows that the end pressure $P_e=(\bar \rho ^+(L_1))^{\gamma }$ is a strictly decreasing and continuous differentiable function on the shock position $x_1=L_s$. In particular, when $L_s=L_0$ and $L_s=L_1$, there are two different end pressure $P_1,\,P_2$ with $P_0>P_1$. Hence, by the monotonicity, one can obtain a transonic shock for the end pressure $P_e\in (P_1,\,P_0)$.
Remark 2.2 Lemma 2.1 shows that the external force helps to stabilize the transonic shock in flat nozzles and the shock position is uniquely determined.
The one-dimensional transonic shock solution $(\bar {u}^{\pm },\, \bar {\rho }^{\pm })$ with a shock occurring at $x_1= L_s$ constructed in lemma 2.1 will be called the background solution in this paper. The extension of the subsonic flow $(\bar u^+(x_1),\,\bar \rho ^+(x_1))$ of the background solution to $L_s-\delta _0< x_1< L_1$ for a small positive number $\delta _0$ will be denoted by $(\hat u^+(x_1),\,\hat \rho ^+(x_1))$.
It is natural to focus on the structural stability of these transonic shock flows. For simplicity, we only investigate the structural stability under suitable small perturbations of the end pressure. Therefore, the supersonic incoming flow is unchanged and remains to be $(\bar {u}^-(x_1),\, 0,\, \bar {\rho }^-(x_1))$.
Assume that the possible shock curve $\Sigma$ and the flow behind the shock are denoted by $x_1=\xi (x_2)$ and $(u_1^+,\,u_2^+,\,P^+)(x)$ respectively (see figure 1). Let $\Omega ^+=\{(x_1,\,x_2): \xi (x_2)< x_1< L_1,\, -1< x_2<1\}$ denotes the subsonic region of the flow. Then the Rankine–Hugoniot conditions on $\Sigma$ gives
In addition, the pressure $P$ satisfies the physical entropy conditions
Since the flow is tangent to the nozzle walls $x_2=\pm 1$, then
The end pressure is perturbed by
due to some technical reasons, we may readily suppose that $P_{ex}(x_2)=P_e^{{1}/{\gamma }}\hat P_{ex}(x_2)\in C^{2,\alpha }([-1,\,1]) (\alpha \in (0,\,1))$ satisfies the compatibility conditions
The following theorem gives the main results of this paper.
Theorem 2.3 Suppose that (2.6), (2.8) and (2.9) hold. Then there exist positive constant $\epsilon _0>0$ such that, for all $\epsilon \in (0,\,\epsilon _0]$, system (2.1) with boundary conditions (2.12)–(2.15) has a unique transonic shock solution $(u_1^+(x),\,u_2^+(x),\, P^+(x);\xi (x_2))$ which admits the following properties:
(i). The shock $x_1=\xi (x_2)\in C^{3,\alpha }([-1,\,1])$, and satisfies
where the positive constant $C$ only depends on the background solution, the exit pressure and $\alpha$.
(ii). The velocity and pressure in subsonic region $(u_1^+,\,u_2^+,\,P^+)(x)\in (C^{2,\alpha }(\bar \Omega ^+))^3$, and there holds
where $\Omega ^+=\{(x_1,\,x_2): \xi (x_2)< x_1< L_1,\, -1< x_2<1\}$ is the subsonic region and $(\hat u,\,0,\,\hat P)=(\hat u(x_1),\,0,\,P(\hat \rho (x_1)))$ is the extended background solution.
Our proof is influenced by the approach developed in [Reference Li, Xin and Yin16, Reference Li, Xin and Yin17, Reference Li, Xin and Yin20], yet the reformulation of the problem is different from there. It is well-known that steady Euler equations are hyperbolic-elliptic coupled in the subsonic region. The entropy and Bernoulli's function are conserved along the particle path, while the pressure and the flow angle satisfy a first-order elliptic system in the subsonic region. These facts are widely used in the structural stability analysis for the transonic shock problems in flat or divergent nozzles, one may refer to [Reference Chen3, Reference Chen, Chen and Song7, Reference Chen, Chen and Feldman8, Reference Li, Xin and Yin16, Reference Li, Xin and Yin17, Reference Li, Xin and Yin20, Reference Serre22, Reference Xin and Yin32, Reference Yuan33] and the references therein. Here we resort to a different decomposition based on the deformation and curl of the velocity developed in [Reference Weng and Xin27, Reference Weng28] for three-dimensional steady Euler and Euler–Poisson systems. The idea in that decomposition is to rewrite the density equation as a Frobenius inner product of a symmetric matrix and the deformation matrix by using Bernoulli's law. The vorticity is resolved by an algebraic equation of Bernoulli's function and the entropy. We should mention that there are several different decompositions to the three-dimensional steady Euler system [Reference Chen and Xie2, Reference Chen4, Reference Chen5, Reference Liu, Xu and Yuan21, Reference Weng23, Reference Xin and Yin32] developed by many researchers for different purposes. An interesting issue that deserves further discussion is when using the deformation-curl decomposition to deal with the transonic shock problem, the end pressure boundary condition becomes non-local since it involves the information from the shock front. However, this non-local boundary condition reduces to be local after introducing the potential function.
3. Reformulation of the problem
Different from previous works on transonic shock problems [Reference Chen3, Reference Chen, Chen and Song7, Reference Chen, Chen and Feldman8, Reference Li, Xin and Yin16, Reference Li, Xin and Yin17, Reference Li, Xin and Yin20], we will use the deformation-curl decomposition developed in [Reference Weng and Xin27, Reference Weng28] for steady Euler system to decompose the original system (2.1) into an equivalent system (3.3), where the hyperbolic quantity $B$ and elliptic quantities $u_1,\,u_2$ are effectively decoupled in subsonic regions. To this end, define the Bernoulli's function
where $h(s)=\frac {\gamma }{\gamma -1}s^{\gamma -1}$ is the enthalpy function. Hence, the density can be expressed by the Bernoulli function and velocity field as
Consequently, the 2-D Euler system (2.1) with unknown function $(u_1,\,u_2,\,P)$ is equivalent to the following system
with unknown function $(u_1,\,u_2,\,B)$.
The shock curve is determined by
Furthermore, it follows from the R-H conditions (2.12) that
A direct computation by using (3.5) shows that on $x_1=\xi (x_2)$
here $h_i(0,\,0,\,0)=0$ for $i=1,\,2$. In addition, we have
and
By substituting (3.6) into (3.1), we conclude that there is a function $h_3$ such that
Thus, theorem 2.3 is established as long as we solve problem (3.3)–(3.4) with boundary conditions (3.5), (2.14)–(2.15). In order to deal with the free boundary value problem (3.3)–(3.4), we introduce the following transformation to reduce it into a fixed boundary value problem. Setting
then, the domain $\Omega ^+=\{(x_1,\,x_2): \xi (x_2)< x_1< L_1,\, -1< x_2<1\}$ is changed into
The inverse change variable gives
We now set for $y\in Q$
Shock equation (3.4) becomes
and system (3.3) is changed into
where
Consider the perturbed functions $v_i(y_1,\,y_2)$, $i=1,\,2,\,3,\,4,$ as
and define the vector functions
It follows from (3.12) that the shock satisfies
Through a direct computation, one can derive from (3.9) and (3.14) that the Bernoulli function satisfies a transport-type equation
where
and $R_3(y_2)=R_3(V(L_s,\,y_2))=O(|V(L_s,\,y_2)|)^2$ is an error term of second order. We may readily drop superscribe $+$ on the background solutions if there is no risk of confusion. And the first-order system for $v_1,\, v_2$ is given by,
where $F_3,\, F_4$ represent the remainder term of the second order with respect to $V$ and $\nabla V$, and
It's obvious that the second formula in (3.6) gives the boundary condition of $v_1$ on the entrance $x_1=L_s$. Meanwhile, formula (3.2) after changing the variable becomes
which together with (2.15) gives the boundary condition of $v_1$ on the exit $x_1=L_1$. Hence, the boundary conditions to system (3.19) read as follow
here
and $R_2(y_2)=R_2(V(L_s,\,y_2))=O(|V(L_s,\,y_2)|)^2$, $R_4(y_2)=R_4(V(L_1,\,y_2))=O(|V (L_1,\,y_2)|)^2$ are error terms of second order. Based on our reformulation, theorem 2.3 follows from the following results.
Theorem 3.1 Under the same assumptions as in theorem 2.3, there exists a positive constant $\epsilon _0>0$ such that for each $\epsilon \in (0,\,\epsilon _0]$, system (3.16)–(3.21) has a unique solution $V\in (C^{2,\alpha }(\bar {Q}))^3\times C^{3,\alpha }([-1,\,1])$ satisfying the following estimate
where the constant $C$ depends only on the background solution, the exit pressure and $\alpha \in (0,\,1)$.
4. Iteration scheme and the linear system
In the first part of this section, we construct an iteration scheme for the non-linear system, and the problem is reduced to the solvability of corresponding linear systems. Indeed, it turns out that the linear system is a non-local elliptic equation of second order with a free parameter denoting the relative location of the shock position on the wall $x_2=-1$. Then we study the existence, uniqueness and regularity for this linear system in the remainder part of this section.
4.1. Iteration scheme
Inspired by [Reference Li, Xin and Yin17], we will develop an iteration scheme to prove theroem 3.1. Consider the Banach space
here $\delta >0$ is a small constant to be determined later. For a fix $\hat V\in {\mathcal {V}}_{\delta }$, equivalently, we have the following quantity
We now define the linearized scheme to problem (3.16)–(3.21) as follows.
Firstly, thanks to (3.16), $v_4$ is determined by
where
hence, one can express the shock as
where $R_5(y_2)=\int _{-1}^{y_2}F_5(\tau ) \,{\rm d}\tau$ is a error term of second order. Due to $\hat V\in {\mathcal {V}}_{\delta }$, we have
Secondly, using (3.17), we get the linearized transport equation for $v_3$:
with initial data
Thus, it can be solved by characteristic methods. Let $y_2(s;\beta )$ be the characteristics going through $(y_1,\, y_2)$ with $y_2(L_s)=\beta$, i.e.
It is noted that $\beta$ can be also regarded as a function of $y=(y_1,\,y_2)$, this is denoted by $\beta =\beta (y)$, which leads to
where
is an error term of second order. Furthermore, we have
It remains to determine the velocity $v_1,\,v_2$ and the shock position difference $v_4(-1)$ on the wall $x_2=-1$. Substituting (4.3) and (4.6) into (3.19) and (3.21), we get the following linearized system for $v_1,\,v_2$ with an unknown parameter $v_4(-1)$:
and the boundary conditions
where
It follows from (4.4), (4.7) and a simple calculation that
and
The second equation in (4.8) implies that there is a potential function $\phi (y)$ that satisfies
it follows that $v_1,\,v_2$ can be represented by
Substituting (4.12) into the first equation in (4.8), we conclude that $\phi$ satisfies the following non-local elliptic equation of the second with an unknown constant $v_4(-1)$
where
Similarly, substituting (4.12) into boundary conditions (4.9), combined with the boundary condition of $\phi$ in (4.11), we have
So far, we have reduced problem (4.8)–(4.9) into a non-local elliptic equation of $\phi$ with an unknown constant $v_4(-1)$. Hence, it is sufficient to establish the solvability and regularity of problem (4.13)–(4.14) to study the original problem. We are going to do it in the next subsection.
4.2. A non-local elliptic equation with a free constant
In this section, we prove the existence, uniqueness and regularity of problem (4.13). To this end, we consider the following more concise form of second-order elliptic system with an unknown constant $\kappa$
where the smooth coefficients $a_i(y_1)$, i=0, 1, 2 and the constant $a_3$ satisfy
and the parameter $\kappa$ is a constant to be determined with the solution itself.
The first lemma implies that the inhomogeneous problem corresponding to system (4.15) without the unknown constant has a unique weak solution.
Lemma 4.1 Suppose that $f\in L^2(Q)$ and $g_i\in L^2(-1,\,1),\, i=1,\,2$, then there exists a suitable large positive constant $K$, such that the following inhomogeneous second-order elliptic equation
admits a unique weak solution $\phi \in H^1(Q)$ satisfying
for some positive constant $C>0$.
Proof. For $\phi,\,\psi \in H^1(Q)$, define the bilinear form
and the linear functional on $H^1(Q)$
It's obviously that the linear functional $l(\psi )$ on $H^1(Q)$ is continuous, i.e.
where we have used the trace theorem. So, what we need to do is just verify that the conditions of the Lax–Milgram Theorem are satisfied for the bilinear form ${\mathcal {B}}$. The boundedness of ${\mathcal {B}}_{K}$ is trivial, we will show that ${\mathcal {B}}_{K}$ is also coercive. Denote $\Lambda =\min \{1,\,C_0\}>0$, then
Cauchy's inequality gives
and
Then, fix $\epsilon _0$ such that $C_1\epsilon _0(1+(L_1-L_s)C_{tr})<{\Lambda }/{2}$, and choosing $K=\max \{\Lambda,\,{C_1}/{\epsilon _0}\}$, thanks to the positivity of $a_3$, we obtain
the unique existence follows from the Lax–Milgram Theorem and (4.19) gives the estimates (4.18). Thus, the proof is complete.
The unique existence of regular solution to non-local system (4.15) is established in the following proposition.
Proposition 4.2 For any $f\in C^{1,\alpha }(\bar Q)$, $g_i\in C^{\alpha }(\bar Q)$, there is a unique weak solution $(\phi,\,\kappa )$, such that $\phi \in H^1(Q)$ and the following estimate holds
Moreover, if the compatibility conditions
are fulfilled, then $\phi \in C^{2,\alpha }(\bar Q)$
and
for some positive constant $C>0$.
Proof. The proof is divided into two steps.
Step 1: Regularity of weak solutions. We will use the symmetric extension methods to exclude the possible singularities that may appear at the corner, which implies that the weak solution $\phi \in H^1(Q)$ to system (4.15) is essentially more regular. To this end, introduce the notation
and define the extended function $\phi ^*(y)$ on $Q^*$ as
Then the extended function $\phi ^*$ satisfies
Using the standard interior and the boundary estimates for the second-order linear elliptic equations in [Reference Gilbarg and Tudinger13], we obtain that $\phi ^*(y)\in C^{1,\alpha }(Q^*)$ and
which implies that $\phi ^*(L_s,\,y_2)\in C^{1,\alpha }[-2,\,2]$. Use estimate (4.26) again to conclude that $\phi ^*(y)\in C^{2,\alpha }(Q^*)$ and
Then (4.22) and (4.23) follow immediately.
Step 2: Existence and uniqueness of weak solutions. Due to the linearity, any solution $\phi$ to problem (4.15) can be decomposed as $\phi =\phi _1+\phi _2$, where $\phi _i$, i=1,2, satisfy the following equation respectively
and
Combing lemma 4.1 with the Fredholm alternative theorem, one can easily derive that (4.28) has a unique $H^1(Q)$ solution $\phi _1$ which satisfies (4.20). On the other hand, one can prove that the only weak solution to (4.29) must be $(\phi _2,\,\kappa )=(-\phi _1(L_s,\,-1),\,\phi _1(L_s,\,-1))$ by applying the maximum principle. For the detailed proof, one can refer to lemma 4.1 in [Reference Li, Xin and Yin17]. Thus, the proof is complete.
At this point, we can easily illustrate the well-posedness to reformulated problem (4.13)–(4.14).
Lemma 4.3 Problem (4.13)–(4.14) has a unique solution $(\phi,\,\kappa )\in C^{2,\alpha }(\bar Q)\times \mathbb {R}$ satisfying
for some positive constant C.
Proof. It suffices to verify solvability condition (4.16) for problem (4.13)–(4.14). A direct but tedious computation shows that
since the background solution is subsonic and smooth, the upper bound is trivial. Hence, proposition 4.2 implies that there exists a unique weak solution $(\phi,\,\kappa )$, and estimate (4.30) follows from (4.22) and (4.23).
In view of the analysis of problem (4.13)–(4.14), the well-posedness of equation (4.8)–(4.9) follows.
Lemma 4.4 Problem (4.8)–(4.9) admits a unique solution $(v_1,\,v_2,\,v_4(-1))\in (C^{2,\alpha }(\bar Q))^2\times \mathbb {R}$ satisfying
and the compatibility conditions
Proof. Combine (4.12) and (4.30), we conclude that there is a unique solution $(v_1,\,v_2,\,v_4(-1))\in (C^{1,\alpha }(\bar Q))^2\times \mathbb {R}$ such that
The similar estimates also hold true for $\|(v_1,\,v_2)\|_{C^{1,\alpha }(\bar Q)}$, but we can derive an even better estimate by rewriting (4.8) into an elliptic equation of $v_1$. To this end, we first rewrite system (4.8) as
where
and
Note that (4.12) and the boundary conditions of $\phi$ imply that
A simple cancellation yields to
Since the boundary conditions in (4.37) are compatible at the corners, we obtain the following estimates for $v_1$ by using the symmetric extension defined by (4.24)
These estimates together with (4.34) also imply that
Hence, (4.31) follows from (4.33) and (4.38)–(4.39). Finally, differentiating the first equation in (4.34) with respect to $x_2$ and combining with $\partial _{y_2}\mathcal {G}_1(y_1,\,\pm 1)=0$ and $\partial _{y_2}v_1(y_1,\,\pm 1)=0$, we obtain
which gives compatibility condition (4.32). Thus, the proof is complete.
5. A priori estimates and proofs of main results
In this section, we will use the Banach contraction mapping theorem to prove theorem 3.1. Given any $\hat {V}\in \mathcal {V}_{\delta }$, we could establish some a priori estimates to the linearized problems defined in subsection 4.1, and construct a contractible mapping from ${\mathcal {V}}_{\delta }$ into itself so that there exists a unique fixed point, which is the solutions obtained in theorem 3.1 and the proof of theorem 3.1 will be finished.
Lemma 4.4 implies that there is a unique solution $(v_1,\,v_2,\,v_4(-1))\in (C^{2,\alpha }(\bar Q))^2\times \mathbb {R}$ to system (4.8)–(4.9) satisfying
and compatibility condition (4.32) holds true.
The shock curve $v_4$ is given by (4.3), which satisfies
Moreover, it follows from (4.32) and (4.4) that
It remains to solve $v_3$. Due to (4.6), combining with (4.7) and (5.3), we obtain the estimate
and the compatibility condition
Taking $\delta =O(1)\epsilon$, then for any given $\hat V\in {\mathcal {V}}_{\delta }$ we can define a continuous mapping $T:{\mathcal {V}}_{\delta } \rightarrow {\mathcal {V}}_{\delta }$ as
due to the iteration scheme introduced in the previous section and estimates (5.1)–(5.5). Finally, we show that the mapping is also contractible in the space $(C^{1,\alpha }(\bar Q))^3\times C^{2,\alpha }[-1,\,1]$.
For arbitrarily given two states $\hat V_i=(\hat v_1^i,\,\hat v_2^i,\,\hat v_3^i,\,\hat v_4^i)\in {\mathcal {V}}_{\delta },\,i=1,\,2$ with the corresponding fluid variable $(\hat u_1^i,\,\hat u_2^i,\,\hat B^i,\,\hat \xi ^i)$, set
where $V_i=(v_1^i,\,v_2^i,\,v_3^i,\,v_4^i)$. For the convenience, we denote $\hat W=\hat V_1-\hat V_2$ and $W=V_1-V_2$, or equivalently,
Equation (4.2) implies that
which yields that
It follows from (4.6) that
where $\beta _i,\,i=1,\,2$ is the initial position such that the corresponding characteristic $y_2^i(s,\,\beta _i)$ going through $(y_1,\,y_2)$ with $y_2^i(L_s)=\beta _i$. It is easy to verify that
thus,
It is straightforward to show that $w_1,\,w_2$ satisfies
and the boundary conditions
where
Then, applying estimate (4.31) to system (5.12)–(5.13) with $k=1$ and together with (5.10), we obtain
Finally, collecting all these estimates above leads to
By (5.15), there is a small constant $\epsilon _0$ such that for all $\epsilon \in (0,\,\epsilon _0]$, the mapping $T$ defined by (5.6) is contractible in the Banach space $(C^{1,\alpha }(\bar Q))^3\times C^{2,\alpha }[-1,\,1]$. Therefore, there exists a unique solution $V$ in $(C^{1,\alpha }(\bar Q))^3\times C^{2,\alpha }[-1,\,1]$. Due to lemma 4.4 and a priori estimates (5.2), (5.4), we know that $V$ also belongs to ${\mathcal {V}}_{\delta }$. It follows that $V$ satisfies estimates (3.23). Thus, the proof of theorem 3.1 is complete. Theorem 2.3 is a direct inference of theorem 3.1. We omit the details.
Data availability statement
Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.
Acknowledgements
Weng is partially supported by the National Natural Science Foundation of China 12071359, 12221001. The authors would like to thank the referees for their useful comments and suggestions.
Competing interest
None.