1 Introduction

Let \(\varOmega \subset {\mathbb {R}}^d\), \(d=2,3\) be a bounded domain, we consider the following semi-stationary compressible Stokes problem:

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\rho +{\text {div}}(\rho \varvec{u})=0,\;{\text {in}}\;(0,T)\times \varOmega ,\\&-\mu \varDelta \varvec{u}-(\lambda +\mu )\nabla {\text {div}}\varvec{u}+\nabla p(\rho )=\varvec{0},\;{\text {in}}\;(0,T)\times \varOmega , \end{aligned} \right. \end{aligned}$$
(1.1)

where \(\rho \) is the fluid density and \(\varvec{u}\) is the velocity. The parameters coefficients \(\mu \) and \(\lambda \) are assumed to be constant and satisfy \(\mu >0\), \(d\lambda +2\mu >0\). The pressure \(p(\rho )\) is governed by the isentropic equation (or Boyle’s law):

$$\begin{aligned} p(\rho )=a\rho ^\gamma ,\quad a>0, \end{aligned}$$
(1.2)

where \(\gamma >1\) is the adiabatic exponent. The internal energy \({\mathcal {H}}\) is given by \({\mathcal {H}}(\rho )=\frac{p(\rho )}{\gamma -1}\). The system (1.1)–(1.2) is supplemented with initial conditions for the density

$$\begin{aligned} \rho (0,x)=\rho _0,\;{\text {in}}\;\varOmega \end{aligned}$$
(1.3)

Together with the following no-slip boundary condition for the velocity

$$\begin{aligned} \varvec{u}=\varvec{0},\;{\text {on}}\;(0,t)\times \partial \varOmega . \end{aligned}$$
(1.4)

In recent years, numerical methods for compressible Stokes equations have received some attention. In the pioneering work of [20], the authors proposed a low order mixed finite element-finite volume (FE-FV) scheme based on nonconforming \(P_1\) (also called Crouzeix–Raviart) finite element for solving the stationary compressible isothermal Stokes problem and analyzed its convergence to a weak solution of the continuous problem. After that, the convergence of mixed FE-FV scheme to weak solution of the isentropic case under the assumption of \(\gamma >1\) has been established by Eymard et al. [10]. Meanwhile, they generalized the results to the well known Marker-and-Cell (MAC) scheme in [9]. Later, the convergence of mixed FE-FV scheme to weak solution of the general compressible Stokes problem (\(p=\varphi (\rho )\), where \(\varphi \) is a superlinear nondecreasing function from \({\mathbb {R}}\) to \({\mathbb {R}}\)) under the hypothesis \(\gamma >1\) was proved by Fettah and Gallouët in [18]. The models studied in the above mentioned literature are all steady state compressible stokes models. The semi-steady compressible Stokes model is known as a reasonable approximation of the isentropic Navier–Stokes equations when the convective effects can be neglected. The convergence of mixed DG-FEM based on nonconforming \(P_1\) finite element for the semi-steady compressible Stokes flow with a Navier boundary condition was shown by Karlsen and Karper in [26]. Meanwhile, they proposed and analyzed the convergence of a new mixed DG-FEM (here the velocity and vorticity were approximated by the div-conforming and curl-conforming Nédélec finite element spaces) to the semi-stationary compressible Stokes systems in [27]. We also mention that the convergence of the MAC scheme for the semi-stationary compressible Stokes flow with Dirichlet boundary conditions was proved in [21]. Very recently, a mixed FE-FV scheme based on Bernardi–Raugel finite element scheme for the stationary compressible isothermal Stokes system was proposed in [2]. The authors gave a convergence proof for the isothermal Stokes equations and investigated the convergence of numerical solutions to its incompressible limit. The convergence analysis is restricted to the isothermal Stokes equations (the pressure of the form (1.2) with \(\gamma =1\)) and the extension to the case \(\gamma >1\) remains open.

The aim of this paper is to show the convergence and error estimates of a mixed DG-FEM based on Bernardi–Raugel finite element for the semi-stationary (isentropic) compressible Stokes equations. This work consists of two major parts. The first part of this paper is to show the convergence of a mixed DG-FEM to a weak solution of the system (1.1) for any \(\gamma >1\). The convergence result of this paper is nontrivial compared to the existing literature. On the one hand, we see that the function \(\varvec{v}_h=\varPi _h^{{\mathbb {V}}}\nabla \varDelta ^{-1}[\rho _h]\) is not a solution to the div-curl problem

$$\begin{aligned} {\text {div}}\varvec{v}_h=\rho _h,\quad \textbf{curl}\;\varvec{v}_h=0, \end{aligned}$$

where \(\varPi _h^{{\mathbb {V}}}\) is the reconstruction interpolation operator of Bernardi–Raugel finite element space \({\mathbb {V}}_h\). Therefore, it is more difficult to obtain the discrete version of the effective viscous flux compared to [26], which will complicate the convergence analysis in this paper. On the other hand, the convergence analysis of this paper is valid for the semi-stationary (isentropic) compressible Stokes equations for any \(\gamma >1\). Of course, it is also valid for the stationary compressible Stokes equations with a slight modification, which fills the gap in the convergence analysis of [2] for the case \(\gamma >1\). We also want to remark that the \(\varvec{H}^1\)-conforming Bernardi–Raugel finite element has several advantages compared to the nonconforming Crouzeix–Raviart element used in the references [10, 18, 20, 23, 26]. Firstly, the conforming finite element method has less number of degrees of freedom which results in a cheaper computational cost. Secondly, the Korn’s inequality is admissible for the conforming method employed to approximate the velocity unknown. It is well known that the Korn’s inequality does not hold for the nonconforming Crouzeix–Raviart finite element space. Therefore, the conforming setting of this paper is easier to generalize to other viscous stress tensor compared to the nonconforming method in the references. Third, the convergence proofs of the conforming setting is less “structure dependent” than the nonconforming method. In other words, the methodology of the convergence proofs in this paper can be easily generalized to other numerical schemes.

The second part of this paper is to derive an error estimate between the mixed DG-FEM solution of the semi-stationary compressible Stokes system and its strong solution. By a detailed consistency analysis and the relative energy functional method introduced in reference [13], two error estimates for the numerical solutions of problem (1.1) under the hypothesis \(\gamma >\frac{6}{5}\) are proved in this paper. All the error results are unconditional in the sense that we do not require the boundedness of numerical solutions and the CFL like condition on the temporal mesh size. The relative energy method was originally designed to analyze the weak-strong uniqueness property of the compressible Navier–Stokes equations. Recently, this idea has been used to analyze the error estimate of numerical schemes of compressible Navier–Stokes system under the hypothesis \(\gamma >\frac{3}{2}\), such as the mixed DG-FEM based on nonconforming Crouzeix–Raviart finite element [12, 23], the implicit MAC scheme [24] and the finite difference method [31]. The error analysis of this paper uses similar analytical techniques but with some modifications. Firstly, our analysis is based on a detailed consistency analysis and the continuous relative energy functional method, rather than the discrete version used in the above literatures. Secondly, our numerical scheme is different from the above work and it requires to deal with some different technical estimates. Thirdly and more importantly, we derive the unconditional \(L^2\) error estimate of pressure under the assumption of \(\gamma >\frac{6}{5}\). To the best of our knowledge, this is the first unconditional error estimate of pressure for the compressible flows.

A brief overview of this work is provided as follows. In the next section, we introduce some notations and preliminary knowledge for this paper. In Sect. 3, we consider a mixed DG-FEM based on Bernardi–Raugel finite element for the semi-stationary compressible Stokes equations. After that, we deduce the discrete energy law, a priori estimate of pressure, the existence of numerical solutions and some uniform bounds. In Sect. 4, we establish the consistency formulation for the continuity equations. In Sect. 5, we show the boundedness of discrete time derivative and an important priori estimates for the density. The convergence of mixed DG-FEM for the nonlinear, isentropic Stokes equations is proved by compactness arguments and the existence analysis of Lions on the discrete level in Sect. 6. In Sect. 7, an unconditional error estimate for mixed DG-FEM solution of the problem (1.1) under the hypothesis \(\gamma >\frac{6}{5}\) is proved by the relative energy functional method.

2 Notation and Preliminaries

In this section, we introduce some notations and preliminary results used in this paper. For any \(1\le q\le \infty \), \(L^q(\varOmega )\) denotes the usual Lebesgue space on \(\varOmega \). For all non-negative integers k and r, \(W^{k,r}(\varOmega )\) stands for the standard Sobolev spaces. We write \(H^k(\varOmega )=W^{k,2}(\varOmega )\). We define \(H_0^1(\varOmega )\) as the subspace of \(H^1(\varOmega )\), which is zero on \(\partial \varOmega \). The vector-valued quantities will be denoted in boldface notations, such as \(\varvec{u}=(u_i)_{i=1}^d\) and \(\varvec{L}^2(\varOmega )=(L^2(\varOmega ))^d\).

Hypothesis 2.1

The initial data \(\rho _0\) satisfies the following properties:

$$\begin{aligned} \rho _0\in L^\gamma (\varOmega ),\quad \rho _0>0. \end{aligned}$$

Definition 2.1

We say that \((\rho ,\varvec{u})\) is a weak solution of the problem (1.1) if it satisfies the following properties:

(i) The solution \((\rho ,\varvec{u})\) satisfied the regularity requirements

$$\begin{aligned} \rho \in L^\infty (0,T;L^{\gamma }(\varOmega ))\cap L^{2\gamma }((0,T)\times \varOmega ),\quad \varvec{u}\in L^2(0,T;\varvec{H}_0^1(\varOmega )). \end{aligned}$$

(ii) For any test fuctions \((\varphi ,\varvec{v})\in C_0^\infty ((0,T)\times \varOmega )\times \varvec{C}_0^\infty ((0,T)\times \varOmega )\) and \(t_F\in [0,T]\), there holds the weak formulation

$$\begin{aligned}&\int _0^{t_F}{\int _{\varOmega }{[\rho \partial _t\varphi +\rho \varvec{u}\cdot \nabla \varphi ]}dx}dt=\left[ \int _{\varOmega }{\rho \varphi }dx\right] _{t=0}^{t=t_F}, \end{aligned}$$
(2.1)
$$\begin{aligned}&\int _0^{t_F}{\int _{\varOmega }{[\mu \nabla \varvec{u}:\nabla \varvec{v}+(\mu +\lambda ){\text {div}}\varvec{u}{\text {div}}\varvec{v}-p(\rho ){\text {div}}\varvec{v}]}dx}dt=0. \end{aligned}$$
(2.2)

(iii) The solution \((\rho ,\varvec{u})\) satisfies the energy inequality

$$\begin{aligned} \left[ \int _{\varOmega }{{\mathcal {H}}(\rho )}dx\right] _{t=0}^{t=t_F}+\int _0^{t_F}{\int _{\varOmega }{[\mu |\nabla \varvec{u}|^2+(\lambda +\mu )|{\text {div}}\varvec{u}|^2]}dx}dt\le 0. \end{aligned}$$
(2.3)

Next, we recall the following renormalized solution argument introduced by DiPerna and Lions (see e.g., [6]).

Definition 2.2

We say that \((\rho ,\varvec{u})\in L^\infty (0,T;L^\gamma (\varOmega ))\times L^2(0,T;\varvec{H}_0^1(\varOmega ))\) is a renormalized solution of the continuity equation \(\partial _t\rho +{\text {div}}(\rho \varvec{u})=0\) if the identity

$$\begin{aligned} \partial _t\varPhi (\rho )+{\text {div}}(\varPhi (\rho )\varvec{u})+\varPsi (\rho ){\text {div}}\varvec{u}=0, \end{aligned}$$

in \({\mathcal {D}}'((0,T)\times \varOmega )\) holds for any \(\varPhi \in C[0,\infty )\cap C^1(0,\infty )\) with \(\varPhi (0)=0\), \(\varPsi (\rho )=\varPhi '(\rho )\rho -\varPhi (\rho )\) and \(\varPhi (\rho ),\varvec{u}\varPhi (\rho )\in L^1((0,T)\times \varOmega )\).

Finally, we recall the following well-known lemma [30] which says that the weak solution \(\rho \) is a renormalized solution.

Lemma 2.1

Suppose that couple \((\rho ,\varvec{u})\in L^2((0,T)\times \varOmega )\times L^2(0,T;\varvec{H}_0^1(\varOmega ))\) satisfies the continuity equation in the weak sense (2.1). Then \((\rho ,\varvec{u})\) is also renormalized solution according to Definition 2.2.

3 Numerical Method

In this section, we consider a mixed DG-FEM based on Bernardi–Raugel finite element for solving the compressible stokes problem (1.1).

3.1 Finite Dimensional Function Spaces

In order to introduce the mixed DG-FEM scheme, the mesh and some discrete function spaces are defined. Let \({\mathcal {T}}_h\) be a quasi-uniform tetrahedral partition of \(\varOmega \) with \({\overline{\varOmega }}=\cup _{K\in {\mathcal {T}}_h}{\overline{K}}\), \(K_i\cap K_j=\emptyset \) for \(K_i,K_j\in {\mathcal {T}}_h\), \(i\ne j\). The mesh size is defined by \(h=\max _{K\in {\mathcal {T}}_h}h_K\), where \(h_K\) is the mesh size of K. We write \({\mathcal {F}}_{h}\) as the set of faces in \({\mathcal {T}}_{h}\), while F is the face. Furthermore, \({\mathcal {F}}_{h,ext}\) is the set of faces \(F\in \partial \varOmega \), while \({\mathcal {F}}_{h,int}={\mathcal {F}}_h\setminus {\mathcal {F}}_{h,ext}\).

In order to discretize the problem, we introduce two families of finite-dimensional spaces. Before proceeding further, we write \({\mathbb {P}}_n(K)\) as the space of polynomials of degree n, while \({\mathbb {P}}_n^d(K)=[{\mathbb {P}}_n(K)]^d\). We define the space of piecewise constant functions

$$\begin{aligned} {\mathbb {Q}}_h:=\{v\in L^2(\varOmega ):v|_{K}\in {\mathbb {P}}_0(K),\;\forall \; K\in {\mathcal {T}}_h\} \end{aligned}$$

for the approximation of the density. In addition, we introduce the associated projection operator

$$\begin{aligned} \varPi _{h}^{{\mathbb {Q}}}:L^2(\varOmega )\rightarrow {\mathbb {Q}}_{h},\quad \varPi _{h}^{{\mathbb {Q}}}[v]|_{K}=\frac{1}{|K|}\int _{K}{v}dx,\quad \forall \; K\in {\mathcal {T}}_{h}. \end{aligned}$$

By recalling the standard Poincaré and Jensen’s inequalities, we have the following interpolation error estimates

$$\begin{aligned} \left\{ \begin{aligned} \Vert \varPi _{h}^{{\mathbb {Q}}}[\varphi ]\Vert _{L^q(K)}\le&\Vert \varphi \Vert _{\varvec{L}^q(K)},\quad \Vert \varphi -\varPi _{h}^{{\mathbb {Q}}}[\varphi ]\Vert _{L^q(K)}\le Ch\Vert \nabla \varphi \Vert _{L^q(K)},\\ \Vert \varPi _{h}^{{\mathbb {Q}}}[\varphi ]\Vert _{L^q(\varOmega )}\le&\Vert \varphi \Vert _{\varvec{L}^q(\varOmega )},\quad \Vert \varphi -\varPi _{h}^{{\mathbb {Q}}}[\varphi ]\Vert _{L^q(\varOmega )}\le Ch\Vert \nabla \varphi \Vert _{L^q(\varOmega )}, \end{aligned} \right. \end{aligned}$$
(3.1)

for any \(K\in {\mathcal {T}}_h\) and \(1\le q\le \infty \). We define the trace

$$\begin{aligned} v^{+}:=\lim _{\delta \rightarrow {0^{+}}}{v(\varvec{x}+\delta \varvec{n}_{F})},\quad v^{-}:=\lim _{\delta \rightarrow {0^{+}}}{v(\varvec{x}-\delta \varvec{n}_{F})}, \end{aligned}$$

where \(\varvec{n}_{F}\) is the outer normal vector to the face F. Moreover, we define the jumps \(\llbracket {v}\rrbracket :=v^{+}-v^{-}\) for any \(F\in {\mathcal {F}}_{h,int}\). Finally, we introduce the semi-norm of the space \({\mathbb {Q}}_h\)

$$\begin{aligned} \Vert v\Vert _{{\mathbb {Q}}_h}^2:=\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{\frac{\llbracket {v}\rrbracket ^2}{h}}dS},\quad \forall \; v\in {\mathbb {Q}}_h. \end{aligned}$$

We employ the Bernardi–Raugel finite element space (see, e.g., [3, 25])

$$\begin{aligned} {\mathbb {V}}_{h}:=\{\varvec{v}\in \varvec{C}^0({\overline{\varOmega }}):\varvec{v}|_{K}\in {\mathbb{B}\mathbb{R}}(K),\forall \; K\in {\mathcal {T}}_h\}\cap \varvec{H}_0^1(\varOmega ) \end{aligned}$$

for the approximation of the velocity. The local Bernardi–Raugel finite element space \({\mathbb{B}\mathbb{R}}(K)\) is given by

$$\begin{aligned} {\mathbb{B}\mathbb{R}}(K):={\mathbb {P}}_1^d(K)\oplus {\text {Span}}\{\varvec{p}_i,\; 1\le i\le d+1\},\quad \varvec{p}_i:=\prod \nolimits _{j=1,j\ne i}^{d+1}\lambda _j\varvec{n}_i, \end{aligned}$$

where \(\lambda _j\) is the barycentric coordinate of K and \(\varvec{n}_i\) is the unit outward normal to \(F_i\subset \partial K\). We introduce the reconstruction interpolation operator (see, e.g., [25, Chapter II])

$$\begin{aligned} \varPi _{h}^{{\mathbb {V}}}:\varvec{H}_0^1(\varOmega )\rightarrow {\mathbb {V}}_{h},\quad \int _{\varOmega }{{\text {div}}\varPi _{h}^{{\mathbb {V}}}[\varvec{v}]\varphi _h}dS=\int _{\varOmega }{{\text {div}}\varvec{v}\varphi _h}dS,\quad \forall \;\varphi _h\in {\mathbb {Q}}_h. \end{aligned}$$

The interpolation operator \(\varPi _{h}^{{\mathbb {V}}}\) has the following error estimates (see, e.g., [25, Chapter II, Lemma 2.2 and 2.8]):

$$\begin{aligned} |\varPi _{h}^{{\mathbb {V}}}\varvec{v}-\varvec{v}|_{m,\varOmega }\le Ch^{k-m}|\varvec{v}|_{k,\varOmega },\quad \forall \;\varvec{v}\in \varvec{H}^k(\varOmega ), \end{aligned}$$
(3.2)

where \(|\cdot |_{m,\varOmega }\) is the semi-norm of \(\varvec{H}^m(\varOmega )\) and \(m=0,1\), \(k=1,2\). Obviously, taking \(k=m=1\) in (3.2), the interpolation operators \(\varPi _h^{{\mathbb {V}}}\) have the following \(\varvec{H}^1\)-stable

$$\begin{aligned} \Vert \varPi _h^{{\mathbb {V}}}\varvec{v}\Vert _{\varvec{H}^1(\varOmega )}\le C\Vert \varvec{v}\Vert _{\varvec{H}^1(\varOmega )},\quad \forall \;\varvec{v}\in \varvec{H}_0^1(\varOmega ). \end{aligned}$$
(3.3)

Finally, we introduce some basic estimate for finite dimensional function spaces. By recalling the following inverse estimate from [5, Theorem 4.5.11], there holds

$$\begin{aligned} \Vert \varvec{v}\Vert _{\varvec{W}^{r,q_1}(\varOmega )}\le Ch^{m-r+3\min \{\frac{1}{q_1}-\frac{1}{q_2},0\}}\Vert \varvec{v}\Vert _{\varvec{W}^{m,q_2}(\varOmega )}, \end{aligned}$$
(3.4)

for any polynomial functions \(\varvec{v}|_K\in {\mathbb {P}}_n^d(K)\), \(K\in {\mathcal {T}}_h\), where \(C>0\) is a generic constant independent of the mesh-size h, m and r are two real numbers with \(0\le m\le r\), \(q_1\) and \(q_2\) are two integers with \(1\le q_1,q_2\le \infty \). By applying the scaling arguments and the trace theorem, we obtain

$$\begin{aligned} \Vert \varvec{v}\Vert _{\varvec{L}^q(\partial K)}\le Ch^{-\frac{1}{q}}(\Vert \varvec{v}\Vert _{\varvec{L}^q(K)}+h\Vert \nabla \varvec{v}\Vert _{\varvec{L}^q(K)}), \end{aligned}$$
(3.5)

for any \(K\in {\mathcal {T}}_h\) and \(1\le q\le \infty \) and \(\varvec{v}\in \varvec{W}^{1,p}(K)\); see, e.g., [1]. Moreover, we apply the inverse estimate (3.4) and the trace inequality (3.5) to obtain

$$\begin{aligned} \Vert \varvec{v}\Vert _{\varvec{L}^q(\partial K)}\le Ch^{-\frac{1}{q}}\Vert \varvec{v}\Vert _{\varvec{L}^q(K)}, \end{aligned}$$
(3.6)

for any \(K\in {\mathcal {T}}_h\) and \(1\le q\le \infty \), \(\varvec{v}\in {\mathbb {P}}_n^d(K)\).

3.2 The Discretization of the Convection Term

Before introducing the scheme, we discuss the approximation of the convection operators in the continuity equation. To this end, we define the standard upwind operator \({\text {Up}}[r_h,\varvec{v}_h]\) on a face F, which is described by

$$\begin{aligned} {\text {Up}}[r_h,\varvec{v}_h]=r_{h,+}[\varvec{v}_{h,F}\cdot \varvec{n}]^{-}+r_{h,-}[\varvec{v}_{h,F}\cdot \varvec{n}]^{+},\quad \forall \; r_h\in {\mathbb {Q}}_h,\;\varvec{v}_h\in {\mathbb {V}}_h, \end{aligned}$$

where \([\varvec{v}_{h,F}\cdot \varvec{n}]^{+}:=\max \{0,\varvec{v}_{h,F}\cdot \varvec{n}\}\) and \([\varvec{v}_{h,F}\cdot \varvec{n}]^{-}:=\min \{0,\varvec{v}_{h,F}\cdot \varvec{n}\}\), \(\varvec{v}_{h,F}:=\frac{1}{|F|}\int _F{\varvec{v}_h}dS\). By applying the following lemma, we can show the distributional error of the convective term and its numerical analogue.

Lemma 3.1

For all \(r_h\in {\mathbb {Q}}_h\) and \(\varvec{v}_h\in {\mathbb {V}}_h\), \(\varphi \in H_0^1(\varOmega )\), we conclude that

$$\begin{aligned} \int _{\varOmega }{r_h\varvec{v}_h\cdot \nabla \varphi }dx=&\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\text {Up}}[r_h,\varvec{v}_h]\llbracket {\varPi _h^{{\mathbb {Q}}}[\varphi ]}\rrbracket }dS}+\int _{\varOmega }{(\varPi _h^{{\mathbb {Q}}}[\varphi ]-\varphi )r_h{\text {div}}\varvec{v}_h}dx\\&+\sum _{K\in {\mathcal {T}}_h}\sum _{F\subset \partial K}{\int _F{(\varPi _h^{{\mathbb {Q}}}[\varphi ]-\varphi )\llbracket {r_h}\rrbracket [\varvec{v}_{h,F}\cdot \varvec{n}]^{-}}dS}\\&+\sum _{K\in {\mathcal {T}}_h}\int _{\partial K}{r_h(\varphi -\varphi _F)(\varvec{v}_h-\varvec{v}_{h,F})\cdot \varvec{n}}dS. \end{aligned}$$

Proof

By the same procedure as in [14, Section 2.3], we easily see that

$$\begin{aligned}&\int _{\varOmega }{r_h\varvec{v}_h\cdot \nabla \varphi }dx=\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\text {Up}}[r_h,\varvec{v}_h]\llbracket {g_h}\rrbracket }dS}+\int _{\varOmega }{(g_h-\varphi )r_h{\text {div}}\varvec{v}_h}dx\\&+\sum _{K\in {\mathcal {T}}_h}\sum _{F\subset \partial K}{\int _F{(g_h-\varphi )\llbracket {r_h}\rrbracket [\varvec{v}_{h,F}\cdot \varvec{n}]^{-}}dS}+\sum _{K\in {\mathcal {T}}_h}\int _{\partial K}{r_h\varphi (\varvec{v}_h-\varvec{v}_{h,F})\cdot \varvec{n}}dS \end{aligned}$$

for any \(r_h,g_h\in {\mathbb {Q}}_h\), \(\varvec{u}_h\in {\mathbb {V}}_h\) and \(\varphi \in H_0^1(\varOmega )\). It can easily be seen that

$$\begin{aligned} \sum _{K\in {\mathcal {T}}_h}\int _{\partial K}{r_h\varphi (\varvec{v}_h-\varvec{v}_{h,F})\cdot \varvec{n}}dS=\sum _{K\in {\mathcal {T}}_h}\int _{\partial K}{r_h(\varphi -\varphi _F)(\varvec{v}_h-\varvec{v}_{h,F})\cdot \varvec{n}}dS. \end{aligned}$$

Combining the above analysis, the proof is thus complete. \(\square \)

3.3 Numerical Scheme

For the time discretization, let N be a fixed integer and \(0=t_0<t_1<\cdots <t_N=T\) be a uniform partition of [0, T] with time-step size \(\tau =T/N\). Moreover, let \(t_n=n\tau \) be the discrete time points and \(v^n\) is the approximation value of the function v at time \(t_n\) for \(0\le n \le N\). For convenience, we introduce \(d_t\varvec{v}^n=(\varvec{v}^n-\varvec{v}^{n-1})/\tau \) and \(D_t\varvec{v}(t)=(\varvec{v}(t)-\varvec{v}(t-\tau ))/\tau \).

We initialize the scheme \(\rho _h^0:=\varPi _h^{{\mathbb {Q}}}[\rho _0]\). For any \(1\le n\le N\), we compute \((\rho _h^n,\varvec{u}_h^n)\in {\mathbb {Q}}_h\times {\mathbb {V}}_h\) by the following numerical scheme

Scheme 1

Given \(\rho _h^{n-1}\in {\mathbb {Q}}_h\), for any \((\varphi _h,\varvec{v}_h)\in {\mathbb {Q}}_h\times {\mathbb {V}}_h\), find \((\rho _h^n,\varvec{u}_h^n)\in {\mathbb {Q}}_h\times {\mathbb {V}}_h\) such that

$$\begin{aligned} \int _{\varOmega }{d_t\rho _h^n\varphi _h}dx-\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\text {Up}}[\rho _h^n,\varvec{u}_h^n]\llbracket {\varphi _h}\rrbracket }dS}\nonumber \\ +h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{\llbracket {\rho _h^n}\rrbracket \llbracket {\varphi _h}\rrbracket }dx=0, \end{aligned}$$
(3.7)
$$\begin{aligned} \int _{\varOmega }{[\mu \nabla \varvec{u}_h^n:\nabla \varvec{v}_h+(\mu +\lambda ){\text {div}}\varvec{u}_h^n{\text {div}}\varvec{v}_h]}dx-\int _{\varOmega }{p(\rho _h^n){\text {div}}\varvec{v}_h}dx=0. \end{aligned}$$
(3.8)

Remark 3.1

(i) Taking \(\varphi _h=1\) in the discrete continuity equation (3.7), we can show \(\int _{\varOmega }{\rho _h^n}dx=\int _{\varOmega }{\rho _h^{n-1}}dx\). In other words, we have immediately the scheme satisfying the conservation of mass. (ii) The stabilization term in the discrete continuity equation is useful in the convergence analysis. More specifically, it provides control over the discrete semi-norm of \(\rho _h\) by some (negative) power of the mesh size h. We remark that the artificial stabilization term in the convergence analysis of compressible flows is introduced by [10, 18, 20].

The renormalized continuity scheme can derived by the following lemma and the proof can be referred to [14, Section 4.1] for more details.

Lemma 3.2

(Renormalized continuity scheme). For any \(1\le n\le N\), let \((\rho _h^n,\varvec{u}_h^n)\in {\mathbb {Q}}_h\times {\mathbb {V}}_h\) satisfy the continuity scheme (3.7). Then \((\rho _h^n,\varvec{u}_h^n)\) also satisfies the following renormalized continuity scheme

$$\begin{aligned}&\int _{\varOmega }{d_t{\mathcal {B}}(\rho _h^n)\varphi _h}dx-\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\text {Up}}[{\mathcal {B}}(\rho _h^n),\varvec{u}_h^n]\llbracket {\varphi _h}\rrbracket }dS}\nonumber \\&+h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{{\mathcal {B}}'(\rho _{h,+}^n)\llbracket {\varphi _h}\rrbracket \llbracket {\rho _h}\rrbracket }dS+h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{{\mathcal {B}}''({\overline{\eta }}_{\rho ,h}^n)\llbracket {\rho _h}\rrbracket ^2}dS\nonumber \\&+\int _{\varOmega }{\varphi _h({\mathcal {B}}'(\rho _h^n)\rho _h^n-{\mathcal {B}}(\rho _h^n)){\text {div}}\varvec{u}_h^n}dx=-\frac{1}{2\tau }\int _{\varOmega }{{\mathcal {B}}''(\xi _{\rho ,h}^n)|\rho _h^n-\rho _h^{n-1}|^2\varphi _h}dx\nonumber \\&-\frac{1}{2}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{\varphi _{h}{\mathcal {B}}''(\eta _{\rho ,h}^n)\llbracket \rho _h^n\rrbracket ^2|\varvec{u}_{h,F}^n\cdot \varvec{n}|}dS}, \end{aligned}$$
(3.9)

for any \({\mathcal {B}}\in C^2({\mathbb {R}}_{+})\) and \(\varphi _h\in {\mathbb {Q}}_h\), where \(\xi _{\rho ,h}^n\in {\text {co}}\{\rho _h^{n-1},\rho _h^n\}\) on each element \(K\in {\mathcal {T}}_h\) and \({\bar{\eta }}_{\rho ,h}^n,\eta _{\rho ,h}^n\in {\text {co}}\{\rho _h^n,(\rho _h^n)^{+}\}\) on each face \(F\in {\mathcal {F}}_h\), where \({\text {co}}\{a,b\}=[\min \{a,b\},\max \{a,b\}]\).

In the upcoming analysis, the discrete density solution \(\rho _h\) is necessary for positive. For this purpose, we recall the following lemma (see, e.g., [22, 26, 28]).

Lemma 3.3

For any \(1\le n\le N\), we assume that \(\rho _h^{n-1}>0\) in \(\varOmega \) and \(\varvec{u}_h^n\in {\mathbb {V}}_h\) holds. Then the solution \(\rho _h^n\in {\mathbb {Q}}_h\) of the discontinuous Galerkin method (3.7) satisfies

$$\begin{aligned} \rho _h^n\ge \frac{\min _{\varvec{x}\in \varOmega }\rho _h^{n-1}}{1+\tau \Vert {\text {div}}\varvec{u}_h^n\Vert _{L^\infty (\varOmega )}}>0. \end{aligned}$$

3.4 A Priori Estimates

In this subsection, we establish some a priori estimates for the discrete solutions of the scheme (3.7)–(3.8), including the energy estimate and the uniformly boundedness of pressure in \(L^2((0,T)\times \varOmega )\).

Theorem 3.1

(Discrete energy law) For any \(1\le m\le N\), the solution \((\rho _h^n,\varvec{u}_h^n)\) of the scheme (3.7)–(3.8) satisfies the following discrete energy law

$$\begin{aligned} {\mathcal {J}}_h(\rho _h^m)+\tau \sum _{i=1}^3\sum _{n=1}^m{\mathcal {D}}_{i,h}^n+\tau \sum _{n=1}^m{\mathcal {D}}_h(\varvec{u}_h^n)={\mathcal {J}}_h(\rho _h^0), \end{aligned}$$
(3.10)

where the discerte energy \({\mathcal {J}}_h\) and the discrete dissipation \({\mathcal {D}}_h\) are defined by

$$\begin{aligned} {\mathcal {J}}_h(\rho _h^n):=\int _{\varOmega }{{\mathcal {H}}(\rho _h^n)}dx,\quad {\mathcal {D}}_h(\varvec{u}_h^n):=\mu \Vert \nabla \varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}^2+(\lambda +\mu )\Vert {\text {div}}\varvec{u}_h^n\Vert _{L^2(\varOmega )}^2, \end{aligned}$$

and the numerical diffusion terms \({\mathcal {D}}_{i,h}^n\) are given by

$$\begin{aligned} {\mathcal {D}}_{1,h}^n&:=\frac{1}{2\tau }\int _{\varOmega }{{\mathcal {H}}''(\xi _{\rho ,h}^n)\left| \rho _h^n-\rho _h^{n-1}\right| ^2}dx,\\ {\mathcal {D}}_{2,h}^n&:=h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\mathcal {H}}''({\overline{\eta }}_{\rho ,h}^n)\llbracket \rho _h^n\rrbracket ^2dS}},\\ {\mathcal {D}}_{3,h}^n&:=\frac{1}{2}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\mathcal {H}}''(\eta _{\rho ,h}^n)\left| \varvec{u}_{h,F}^n\cdot \varvec{n}\right| \llbracket \rho _h^n\rrbracket ^2dS}}. \end{aligned}$$

Proof

Taking \(({\mathcal {B}},\varphi _h)=({\mathcal {H}},1)\) in the renormalized continuity scheme (3.9) and by applying \({\mathcal {H}}'(\rho )\rho -{\mathcal {H}}(\rho )=p(\rho )\), we can show

$$\begin{aligned} \int _{\varOmega }{p(\rho _h^n){\text {div}}\varvec{u}_h^n}dx=&-d_t\int _{\varOmega }{{\mathcal {H}}(\rho _h^n)}dx-\frac{\tau }{2}\int _{\varOmega }{{\mathcal {H}}''(\xi _{\rho ,h}^n)(d_t\rho _h^n)^2}dx\\&-h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\mathcal {H}}''({\bar{\eta }}_{\rho ,h}^n)\llbracket \rho _h^n\rrbracket ^2}dS}\\&-\frac{1}{2}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\mathcal {H}}''(\eta _{\rho ,h}^n)\llbracket \rho _h^n\rrbracket ^2\left| \varvec{u}_{h,F}^n\cdot \varvec{n}\right| }dS}. \end{aligned}$$

Let \(\varvec{v}_h=\varvec{u}_h\) in (3.8), we conclude that

$$\begin{aligned} \mu \Vert \nabla \varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}^2+(\lambda +\mu )\Vert \nabla \varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}^2-\int _{\varOmega }{p(\rho _h^n){\text {div}}\varvec{u}_h^n}dx=0. \end{aligned}$$

Combining the above analysis implies

$$\begin{aligned} d_t\int _{\varOmega }{{\mathcal {H}}(\rho _h^n)}dx+{\mathcal {D}}_h(\varvec{u}_h^n)+\sum _{i=1}^3{\mathcal {D}}_{i,h}^n=0, \end{aligned}$$
(3.11)

for any \(1\le n\le N\). Summing (3.11) with respect to n from \(n=1\) to \(n=m\), we obtain (3.10). The proof is thus complete. \(\square \)

In order to show the \(L^2(\varOmega )\) estimate of pressure, we introduce an inverse of the divergence operator \({\textbf{B}}\), which satisfies the following result (see [4] and [19, Chapter 3]).

Lemma 3.4

Let \(\varOmega \subset {\mathbb {R}}^d\), \(d=2,3\) be a bounded domain. There exists a linear operator \({\textbf{B}}\) enjoying the properties

$$\begin{aligned} {\textbf{B}}[r]\in \varvec{W}_0^{1,q}(\varOmega ),\quad {\text {div}}{\textbf{B}}[r]=r\quad \forall \; r\in L^q(\varOmega ),\quad \int _{\varOmega }{r}dx=0,\quad \forall \; 1<q<\infty . \end{aligned}$$

Moverever, the linear operator \({\textbf{B}}\) satisfies the following estimate

$$\begin{aligned} \Vert {\textbf{B}}[r]\Vert _{\varvec{W}_0^{1,q}(\varOmega )}\le C\Vert r\Vert _{L^q(\varOmega )},\quad \forall \; 1<q<\infty . \end{aligned}$$
(3.12)

Next, we prove the stability estimate for the discrete pressure.

Theorem 3.2

Suppose that Hypothesis 2.1 is satisfied. For any \(1\le m\le N\), then the pressure \(p(\rho _h)\) satisfies the following estimate

$$\begin{aligned} \tau \sum _{n=1}^m\Vert p(\rho _h^n)\Vert _{L^2(\varOmega )}^2\le C. \end{aligned}$$
(3.13)

Proof

Let \(r_h^n:=p(\rho _h^n)-\frac{1}{|\varOmega |}\int _{\varOmega }{p(\rho _h^n)}dx\) for \(1\le n\le N\). Taking \(\varvec{v}_h^n=\varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_h^n]\) in (3.8) and by the definition of \(\varPi _h^{{\mathbb {V}}}\) and \({\textbf{B}}\), we can show

$$\begin{aligned} \Vert p(\rho _h^n)\Vert _{L^2(\varOmega )}^2=&\frac{1}{|\varOmega |}\Vert p(\rho _h^n)\Vert _{L^1(\varOmega )}^2+\mu \int _{\varOmega }{\nabla \varvec{u}_h^n:\nabla \varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_h^n]}dx\\&+(\lambda +\mu )\int _{\varOmega }{{\text {div}}\varvec{u}_h^n{\text {div}}\varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_h^n]}dx. \end{aligned}$$

By applying Hölder inequality, the estimates (3.3) and (3.12), we obtain

$$\begin{aligned} \left| \int _{\varOmega }{\nabla \varvec{u}_h^n:\nabla \varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_h^n]}dx\right| \le&\Vert \nabla \varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}\Vert \nabla \varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_h^n]\Vert _{\varvec{L}^2(\varOmega )}\\ \le&C\Vert \nabla \varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}\Vert p(\rho _h^n)\Vert _{L^2(\varOmega )},\\ \left| \int _{\varOmega }{{\text {div}}\varvec{u}_h^n{\text {div}}\varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_h^n]}dx\right| \le&\Vert {\text {div}}\varvec{u}_h^n\Vert _{L^2(\varOmega )}\Vert {\text {div}}\varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_h^n]\Vert _{L^2(\varOmega )}\\ \le&C\Vert {\text {div}}\varvec{u}_h^n\Vert _{L^2(\varOmega )}\Vert p(\rho _h^n)\Vert _{L^2(\varOmega )}. \end{aligned}$$

Combining the above analysis, by applying Young inequality, we have

$$\begin{aligned} \Vert p(\rho _h^n)\Vert _{L^2(\varOmega )}^2\le&\frac{1}{|\varOmega |}\Vert p(\rho _h^n)\Vert _{L^1(\varOmega )}^2+C\Vert \nabla \varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}^2\nonumber \\&+C\Vert {\text {div}}\varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}^2+\frac{1}{2}\Vert p(\rho _h^n)\Vert _{L^2(\varOmega )}^2, \end{aligned}$$
(3.14)

for any \(1\le n\le N\).

Summing (3.14) with respect to n from \(n=1\) to \(n=m\) and applying the discrete energy estimate (3.10) implies

$$\begin{aligned} \tau \sum _{n=1}^m\Vert p(\rho _h^n)\Vert _{L^2(\varOmega )}^2\le&C\tau \sum _{n=1}^m\Vert p(\rho _h^n)\Vert _{L^1(\varOmega )}^2+C\tau \sum _{n=1}^m\Vert \nabla \varvec{u}_h^n\Vert _{\varvec{L}^2(\varOmega )}^2\\ \le&C{\mathcal {J}}_h^2(\rho _h^0)+C{\mathcal {J}}_h(\rho _h^0)\le C\Vert \rho _0\Vert _{L^\gamma (\varOmega )}^{2\gamma }+C\Vert \rho _0\Vert _{L^\gamma (\varOmega )}^{\gamma }.\nonumber \end{aligned}$$
(3.15)

By applying Hypothesis 2.1 for the inequality (3.15), we have (3.13). The proof is thus complete. \(\square \)

3.5 Existence of Numerical Solution

By applying Schaeffer’s fixed point theorem, we can show the existence of numerical solutions for the scheme (3.7)–(3.8) in this subsection. Firstly, we recall Schaeffer’s fixed point theory (see, e.g., [8, Theorem 9.2.4]):

Lemma 3.5

Let \({\mathcal {L}}:D\rightarrow D\) be a continuous mapping defined on a finite dimensional normed vector space D. Suppose that the set

$$\begin{aligned} \{z\in D:z=\varLambda {\mathcal {L}}(z),\;\varLambda \in [0,1]\} \end{aligned}$$

is non empty and bounded. Then there exists \(z\in D\) such that \(z={\mathcal {L}}(z)\).

Then we can prove an existence result of numerical solutions for the scheme (3.7)–(3.8).

Theorem 3.3

For any \(1\le n\le N\), let \((\rho _h^{n-1},\varvec{u}_h^{n-1})\in {\mathbb {Q}}_h\times {\mathbb {V}}_h\) and \(\rho _h^{n-1}>0\) be given. Then, for each fixed \(h,\tau >0\), the scheme (3.7)–(3.8) has at least one solution

$$\begin{aligned} (\rho _h^n,\varvec{u}_h^n)\in {\mathbb {Q}}_h\times {\mathbb {V}}_h,\quad \rho _h^n>0. \end{aligned}$$

The proof of Theorem 3.3 can be found in “Appendix A.1”.

3.6 Uniform Bounds

In this subsection, we deduce some priori estimates from the discrete energy law (3.10). To this end, we need to extend the definition of discrete solution for any \(t\le T\). We define the piecewise constant interpolations of \(\rho _h^n\) by

$$\begin{aligned} \rho _h(t,\cdot ):=\left\{ \begin{aligned}&\rho _h^0,{} & {} {\text {for}}\; t\in (-\infty ,0],\\&\rho _h^n,{} & {} {\text {for}}\; t\in (t_{n-1},t_n],\;\forall \;1\le n\le N, \end{aligned} \right. \end{aligned}$$
(3.16)

and the piecewise constant interpolations of \(\varvec{u}_h^n\) by

$$\begin{aligned} \varvec{u}_h(t,\cdot ):=\varvec{u}_h^n,\quad {\text {for}}\; t\in (t_{n-1},t_n],\;\forall \; 1\le n\le N. \end{aligned}$$
(3.17)

The following stable results are proved by the discrete energy law and the \(L^2\) estimate of pressure, which is crucial in both error estimates and convergence analysis.

Lemma 3.6

Suppose that Hypothesis 2.1 is satisfied. Then the family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) satisfies the following estimates:

$$\begin{aligned} \Vert \rho _h\Vert _{L^\infty (0,T;L^\gamma (\varOmega ))}\le&C,\quad \Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\le C,\\ \Vert p(\rho _h)\Vert _{L^2((0,T)\times \varOmega )}\le&C,\quad \Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{L}^6(\varOmega ))}\le C,\\ \Vert {\text {div}}\varvec{u}_h\Vert _{L^2((0,T)\times \varOmega )}\le&C,\quad \Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\le C. \end{aligned}$$

Lemma 3.7

Suppose that Hypothesis 2.1 is satisfied. Then the family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) satisfies the following estimates:

$$\begin{aligned} \int _0^T{\int _{\varOmega }{{\mathcal {H}}''(\xi _{\rho ,h})|\rho _h-\rho _h^{\star }|^2}dx}dt\le&C\tau ,\\ \int _0^T{h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{{\mathcal {H}}''({\overline{\eta }}_{\rho ,h})\llbracket \rho _h\rrbracket ^2}dS}dt\le&C,\\ \int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{{\mathcal {H}}''(\eta _{\rho ,h})|\varvec{u}_{h,F}\cdot \varvec{n}|\llbracket \rho _h\rrbracket ^2}dS}dt\le&C. \end{aligned}$$

Lemma 3.8

Suppose that Hypothesis 2.1 and \(\gamma \ge 2\) are satisfied. Then the family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) satisfies the following estimates:

$$\begin{aligned} \int _0^T{h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{\llbracket \rho _h\rrbracket ^2}dS}dt\le&C,\quad \int _0^T{\int _{\varOmega }{|\rho _h-\rho _h^{\star }|^2}dx}dt\le C\tau ,\\ \int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{|\varvec{u}_{h,F}\cdot \varvec{n}|\llbracket \rho _h\rrbracket ^2}dS}dt\le&C. \end{aligned}$$

Proof

Taking \(({\mathcal {B}}(\rho ),\varphi _h)=(\rho ^2,1)\) in renormalized continuity scheme (3.9) and summing this result with respect to n from \(n=1\) to \(n=N\), we obtain

$$\begin{aligned}&\frac{1}{\tau }\int _0^T{\int _{\varOmega }{|\rho _h-\rho _h^{\star }|^2}dx}dt+\int _0^T{h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{\llbracket \rho _h\rrbracket ^2}dS}dt\\&\qquad +\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{|\varvec{u}_{h,F}\cdot \varvec{n}|\llbracket \rho _h\rrbracket ^2}dS}dt\\&\quad \le -\int _0^T{\int _{\varOmega }{\rho _h^2{\text {div}}\varvec{u}_h}dx}dt-\int _{\varOmega }{\rho _h(T,\cdot )^2}dx+\int _{\varOmega }{\rho _h(0,\cdot )^2}dx:=\sum _{i=1}^3{\mathcal {U}}_i. \end{aligned}$$

By applying Hölder inequality and the embedding \(L^{2\gamma }\hookrightarrow L^4\) and \(L^\gamma \hookrightarrow L^2\) for \(\gamma \ge 2\), we conclude that

$$\begin{aligned} |{\mathcal {U}}_1|\le&\Vert \rho _h\Vert _{L^4((0,T)\times \varOmega )}^2\Vert {\text {div}}\varvec{u}_h\Vert _{L^2((0,T)\times \varOmega )}\\ \le&C\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}^2\Vert {\text {div}}\varvec{u}_h\Vert _{L^2((0,T)\times \varOmega )},\\ |{\mathcal {U}}_2|\le&C\Vert \rho _h\Vert _{L^\infty (0,T; L^\gamma (\varOmega ))}^2,\quad |{\mathcal {U}}_3|\le C\Vert \rho _0\Vert _{L^\gamma (\varOmega )}^2. \end{aligned}$$

Combining the above analysis with Hypothesis 2.1 and Lemma 3.6, we have the required estimates, the proof is thus complete. \(\square \)

4 Consistency Formulation of the Continuity Scheme

In this section, we establish the consistency formulation for the discrete solution of the numerical scheme (3.7)–(3.8). In other words, the discrete solution asymptotically satisfies the weak formulation of continuous problem.

Lemma 4.1

The family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) satisfies the following consistency formulation

$$\begin{aligned} \int _0^T{\int _{\varOmega }{[D_t\rho _h\varphi -\rho _h\varvec{u}_h\cdot \nabla \varphi ]}dx}dt=\int _0^T{\int _{\varOmega }{{\mathscr {R}}_h\cdot \nabla \varphi }dx}dt, \end{aligned}$$
(4.1)

for any \(\varphi \in L^2(0,T;H^1(\varOmega ))\), where the remainder functional \({\mathscr {R}}_h\) is given by

$$\begin{aligned}&\int _0^T{\int _{\varOmega }{{\mathscr {R}}_h\cdot \nabla \varphi }dx}dt=\int _0^T{\sum _{K\in {\mathcal {T}}_h}{\int _{\partial K}{(\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ])\llbracket \rho _h\rrbracket [\varvec{u}_{h,F}\cdot \varvec{n}]^{-}}dS}}dt\\&+\int _0^T{\int _{\varOmega }{(\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ])\rho _h{\text {div}}\varvec{u}_h}dx}dt+h^{\epsilon -1}\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{\llbracket \rho _h\rrbracket \llbracket \varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]\rrbracket }dS}dt\nonumber \\&+\int _0^T{\sum _{K\in {\mathcal {T}}_h}\int _{\partial K}{\rho _h(\varphi -\varphi _F)(\varvec{u}_{h,F}-\varvec{u}_h)\cdot \varvec{n}}dS}dt:=\sum _{i=1}^4{\mathcal {P}}_i(\varphi ).\nonumber \end{aligned}$$
(4.2)

Proof

Taking \(\varphi _h=\varPi _h^{{\mathbb {Q}}}[\varphi ]\) in the continuity method (3.7) and summing this identity with respect to n from \(n=1\) to \(n=N\), we can show

$$\begin{aligned} \int _0^T{\int _{\varOmega }{D_t\rho _h\varPi _h^{{\mathbb {Q}}}[\varphi ]}dx}dt&-\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\text {Up}}[\rho _h,\varvec{u}_h]\llbracket {\varPi _h^{{\mathbb {Q}}}[\varphi ]}\rrbracket }dS}}dt\\&+h^{\epsilon -1}\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{\llbracket \rho _h\rrbracket \llbracket \varPi _h^{{\mathbb {Q}}}[\varphi ]-\varphi \rrbracket }dS}dt=0. \end{aligned}$$

It is easy to check that

$$\begin{aligned} \int _0^T{\int _{\varOmega }{D_t\rho _h\varPi _h^{{\mathbb {Q}}}[\varphi ]}dx}dt=\int _0^T{\int _{\varOmega }{D_t\rho _h\varphi }dx}dt. \end{aligned}$$

By taking \((r_h,\varvec{v}_h)=(\rho _h^n,\varvec{u}_h^n)\) in Lemma 3.1 and summing this idenity with respect to n from \(n=1\) to \(n=N\), we conclude that

$$\begin{aligned}&\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\text {Up}}\left[ \rho _h,\varvec{u}_h\right] \llbracket {\varPi _h^{{\mathbb {Q}}}[\varphi ]}\rrbracket }dS}}dt=\int _0^T{\int _{\varOmega }{(\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ])\rho _h{\text {div}}\varvec{u}_h}dx}dt\\&+\int _0^T{\int _{\varOmega }{\rho _h\varvec{u}_h\cdot \nabla \varphi }dx}dt+\int _0^T{\sum _{K\in {\mathcal {T}}_h}{\int _{\partial K}{(\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ])\llbracket {\rho _h}\rrbracket [\varvec{u}_{h,F}\cdot \varvec{n}]^{-}}dS}}dt\\&+\int _0^T{\sum _{K\in {\mathcal {T}}_h}\int _{\partial K}{\rho _h(\varphi -\varphi _F)(\varvec{u}_{h,F}-\varvec{u}_h)\cdot \varvec{n}}dS}dt. \end{aligned}$$

Combining the above analysis, we obtain (4.1). The proof is thus complete. \(\square \)

Next, the error estimate of the remainder term \({\mathscr {R}}_h\) of Lemma 4.1 is proved in the following lemma.

Lemma 4.2

Suppose that Hypothesis 2.1 is satisfied. There exists a constant \(C>0\) independent of h and \(\tau \), such that the error functional \({\mathscr {R}}_h\) of Lemma 4.1 satisfies the following estimates

$$\begin{aligned} \left| \int _0^T{\int _{\varOmega }{{\mathscr {R}}_h\cdot \nabla \varphi }dx}dt\right| \le Ch^A\Vert \nabla \varphi \Vert _{L^{2m_1}(0,T;L^6(\varOmega ))}, \end{aligned}$$
(4.3)

where the parameters A and \(m_1\) are given by

$$\begin{aligned} A:=\frac{\min \{1,\epsilon \}}{2},\quad m_1:=\frac{2\gamma }{\gamma -1}. \end{aligned}$$

Proof

We show the proof of this Lemma in four steps.

Bound on \({\mathcal {P}}_1\). We estimate this term for \(1<\gamma \le 2\) and \(\gamma >2\) separately. If \(1<\gamma \le 2\), by applying Cauchy–Schwarz inequality, we can show

$$\begin{aligned} |{\mathcal {P}}_1(\varphi )|\le \sqrt{{\mathcal {P}}_{1,1,1}}\times \sqrt{{\mathcal {P}}_{1,1,2}}, \end{aligned}$$
(4.4)

where \({\mathcal {P}}_{1,1,1}\) and \({\mathcal {P}}_{1,1,2}\) are given by

$$\begin{aligned} {\mathcal {P}}_{1,1,1}:=&\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\mathcal {H}}''(\eta _{\rho ,h})|\varvec{u}_{h,F}\cdot \varvec{n}|\llbracket \rho _h\rrbracket ^2dS}}}dt,\\ {\mathcal {P}}_{1,1,2}:=&\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _{F}{({\mathcal {H}}''(\eta _{\rho ,h}))^{-1}|\varvec{u}_{h,F}\cdot \varvec{n}||\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]|^2}dS}dt. \end{aligned}$$

It is easy to check that

$$\begin{aligned} ({\mathcal {H}}''(\eta _{\rho ,h}))^{-1}\le&C(\rho _{h,+}+\rho _{h,-})^{2-\gamma }\nonumber \\ \le&C(1+\rho _{h,+}+\rho _{h,-}),\quad {\text {with}}\; 1<\gamma \le 2. \end{aligned}$$
(4.5)

For the term \({\mathcal {P}}_{1,1,2}\), by applying the inequality (4.5), we obtain

$$\begin{aligned} {\mathcal {P}}_{1,1,2}\le C({\mathcal {P}}_{1,1,2,1}+{\mathcal {P}}_{1,1,2,2}). \end{aligned}$$
(4.6)

where \({\mathcal {P}}_{1,1,2,1}\) and \({\mathcal {P}}_{1,1,2,2}\) are defined by

$$\begin{aligned} {\mathcal {P}}_{1,1,2,1}:=&\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _{F}{|\varvec{u}_{h,F}\cdot \varvec{n}||\rho _{h,+}+\rho _{h,-}||\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]|^2}dS}dt,\\ {\mathcal {P}}_{1,1,2,2}:=&\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _{F}{|\varvec{u}_{h,F}\cdot \varvec{n}||\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]|^2}dS}dt. \end{aligned}$$

Using the trace inequalities (3.5)–(3.6), we conclude that

$$\begin{aligned} {\mathcal {P}}_{1,1,2,1}\le&\int _0^T{\sum _{K\in {\mathcal {T}}_h}\Vert \varvec{u}_h\Vert _{\varvec{L}^6(\partial K)}\Vert \rho _h\Vert _{L^2(\partial K)}\Vert \varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]\Vert _{L^6(\partial K)}^2}dt\\ \le&Ch^{-1}\int _0^T{\sum _{K\in {\mathcal {T}}_h}\Vert \varvec{u}_h\Vert _{\varvec{L}^6(K)}\Vert \rho _h\Vert _{L^2(K)}\Vert \varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]\Vert _{L^6(K)}^2}dt\\&+Ch\int _0^T{\sum _{K\in {\mathcal {T}}_h}\Vert \varvec{u}_h\Vert _{\varvec{L}^6(K)}\Vert \rho _h\Vert _{L^2(K)}\Vert \nabla \varphi \Vert _{L^6(K)}^2}dt. \end{aligned}$$

Therefore, using the Hölder inequality and the interpolation error estimate (3.1), we get that

$$\begin{aligned} {\mathcal {P}}_{1,1,2,1}\le Ch\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{L}^6(\varOmega ))}\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\Vert \nabla \varphi \Vert _{L^{2m_1}(0,T;L^6(\varOmega ))}^2. \end{aligned}$$
(4.7)

By a similar proof to the error estimate of \({\mathcal {P}}_{1,1,2,2}\), we find

$$\begin{aligned} {\mathcal {P}}_{1,1,2,2}\le Ch\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{L}^6(\varOmega ))}\Vert \nabla \varphi \Vert _{L^4(0,T;L^{\frac{12}{5}}(\varOmega ))}^2. \end{aligned}$$
(4.8)

Inserting (4.7) and (4.8) into (4.6), using Lemma 3.6, we have arrived at

$$\begin{aligned} {\mathcal {P}}_{1,1,2}\le Ch\Vert \nabla \varphi \Vert _{L^{2m_1}(0,T;L^6(\varOmega ))}^2. \end{aligned}$$
(4.9)

By applying Lemma 3.7 and (4.9) to (4.4) leads to the bound

$$\begin{aligned} |{\mathcal {P}}_1(\varphi )|\le Ch^{\frac{1}{2}}\Vert \nabla \varphi \Vert _{L^{2m_1}(0,T;L^6(\varOmega ))}. \end{aligned}$$

For the case \(\gamma >2\), by using Cauchy–Schwarz inequality, we obtain

$$\begin{aligned} |{\mathcal {P}}_1(\varphi )|\le \sqrt{{\mathcal {P}}_{1,2,1}}\times \sqrt{{\mathcal {P}}_{1,1,2,2}}, \end{aligned}$$

where \({\mathcal {P}}_{1,2,1}\) is defined by

$$\begin{aligned} {\mathcal {P}}_{1,2,1}:=\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{\llbracket \rho _h\rrbracket ^2|\varvec{u}_{h,F}\cdot \varvec{n}|}dS}dt. \end{aligned}$$

By virtue of the inequality (4.8), Lemmas 3.6 and 3.8, we have

$$\begin{aligned} |{\mathcal {P}}_1(\varphi )|\le Ch^{\frac{1}{2}}\Vert \nabla \varphi \Vert _{L^4(0,T;L^{\frac{12}{5}}(\varOmega ))}. \end{aligned}$$

Bound on \({\mathcal {P}}_2(\varphi )\). By applying the Hölder inequality, the inverse estimate (3.4) and the embedding \(L^{2\gamma }\hookrightarrow L^2\) for \(\gamma >1\), we obtain

$$\begin{aligned} |{\mathcal {P}}_2(\varphi )|\le&\Vert \varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]\Vert _{L^{m_1}(0,T;L^6(\varOmega ))}\Vert \rho _h\Vert _{L^{2\gamma }(0,T;L^3(\varOmega ))}\Vert {\text {div}}\varvec{u}_h\Vert _{L^2(0,T;L^2(\varOmega ))}\\ \le&Ch^{-\frac{1}{2}}\Vert \varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]\Vert _{L^{m_1}(0,T;L^6(\varOmega ))}\Vert \rho _h\Vert _{L^{2\gamma }(0,T;L^2(\varOmega ))}\Vert {\text {div}}\varvec{u}_h\Vert _{L^2(0,T;L^2(\varOmega ))}\\ \le&Ch^{\frac{1}{2}}\Vert \nabla \varphi \Vert _{L^{m_1}(0,T;L^6(\varOmega ))}\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\Vert {\text {div}}\varvec{u}_h\Vert _{L^2(0,T;L^2(\varOmega ))}\\ \le&Ch^{\frac{1}{2}}\Vert \nabla \varphi \Vert _{L^{m_1}(0,T;L^6(\varOmega ))}. \end{aligned}$$

Bound on \({\mathcal {P}}_3(\varphi )\). We shall treat the case \(1<\gamma \le 2\) and \(\gamma >2\) separately. If \(1<\gamma \le 2\), by applying the Cauchy Schwarz inequality, we have

$$\begin{aligned} |{\mathcal {P}}_3(\varphi )|\le \sqrt{{\mathcal {P}}_{3,1,1}}\times \sqrt{{\mathcal {P}}_{3,1,2}}, \end{aligned}$$
(4.10)

where \({\mathcal {P}}_{3,1}\) and \({\mathcal {P}}_{3,2}\) are defined by

$$\begin{aligned} {\mathcal {P}}_{3,1,1}:&=\int _0^T{h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _F{{\mathcal {H}}''({\overline{\eta }}_{\rho ,h})\llbracket \rho _h\rrbracket ^2}dS}}dt,\\ {\mathcal {P}}_{3,1,2}:&=\int _0^T{h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _F{({\mathcal {H}}''({\overline{\eta }}_{\rho ,h}))^{-1}\llbracket \varPi _h^{{\mathbb {Q}}}[\varphi ]-\varphi \rrbracket ^2}dS}}dt. \end{aligned}$$

By employing the inequality (4.5), we can show

$$\begin{aligned} {\mathcal {P}}_{3,1,2}\le C({\mathcal {P}}_{3,1,2,1}+{\mathcal {P}}_{3,1,2,2}), \end{aligned}$$
(4.11)

where \({\mathcal {P}}_{3,1,2,1}\) and \({\mathcal {P}}_{3,1,2,2}\) are given by

$$\begin{aligned} {\mathcal {P}}_{3,1,2,1}:=&h^{\epsilon -1}\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _{F}{|\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]|^2}dS}dt,\\ {\mathcal {P}}_{3,1,2,2}:=&h^{\epsilon -1}\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _{F}{|\rho _{h,+}+\rho _{h,-}||\varphi -\varPi _h^{{\mathbb {Q}}}[\varphi ]|^2}dS}dt. \end{aligned}$$

By applying the trace estimate (3.5)–(3.6) and the Poincaré inequality (3.1), the embedding \(L^{2\gamma }\hookrightarrow L^2\) for \(\gamma >1\), we get

$$\begin{aligned} {\mathcal {P}}_{3,1,2,1}\le&Ch^{\epsilon -2}\Vert \varPi _h^{{\mathbb {Q}}}[\varphi ]-\varphi \Vert _{L^2((0,T)\times \varOmega )}^2+Ch^{\epsilon }\Vert \nabla \varphi \Vert _{L^2((0,T)\times \varOmega )}^2\nonumber \\ \le&Ch^{\epsilon }\Vert \nabla \varphi \Vert _{L^2((0,T)\times \varOmega )}^2, \end{aligned}$$
(4.12)
$$\begin{aligned} {\mathcal {P}}_{3,1,2,2}\le&Ch^{\epsilon -2}\Vert \rho _h\Vert _{L^2((0,T)\times \varOmega )}\Vert \varPi _h^{{\mathbb {Q}}}[\varphi ]-\varphi \Vert _{L^4((0,T)\times \varOmega )}^2\nonumber \\&+Ch^{\epsilon }\Vert \rho _h\Vert _{L^2((0,T)\times \varOmega )}\Vert \nabla \varphi \Vert _{L^4((0,T)\times \varOmega )}^2\nonumber \\ \le&Ch^{\epsilon }\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\Vert \nabla \varphi \Vert _{L^4((0,T)\times \varOmega )}^2. \end{aligned}$$
(4.13)

Inserting (4.12) and (4.13) into (4.11), using Lemma  3.6, we conclude that

$$\begin{aligned} {\mathcal {P}}_{3,1,2}\le Ch^{\epsilon }\Vert \nabla \varphi \Vert _{L^4((0,T)\times \varOmega )}^2. \end{aligned}$$
(4.14)

For the inequality (4.10), by using (4.14) and Lemma 3.7, we easily see that

$$\begin{aligned} |{\mathcal {P}}_3(\varphi )|\le Ch^{\frac{\epsilon }{2}}\Vert \nabla \varphi \Vert _{L^4((0,T)\times \varOmega )}. \end{aligned}$$

For the case \(\gamma >2\), by applying the Cauchy Schwarz inequality, we obtain

$$\begin{aligned} |{\mathcal {P}}_3(\varphi )|\le \sqrt{{\mathcal {P}}_{3,2,1}}\times \sqrt{{\mathcal {P}}_{3,1,2,1}}, \end{aligned}$$

where \({\mathcal {P}}_{3,2,1}\) is given by

$$\begin{aligned} {\mathcal {P}}_{3,2,1}:=\int _0^T{h^{\epsilon -1}\sum _{F\in {\mathcal {F}}_{h,int}}{\int _F{\llbracket \rho _h\rrbracket ^2}dS}}dt. \end{aligned}$$

According to Lemma 3.8 and the estimate (4.12), we have

$$\begin{aligned} |{\mathcal {P}}_3(\varphi )|\le Ch^{\frac{\epsilon }{2}}\Vert \nabla \varphi \Vert _{L^4((0,T)\times \varOmega )}. \end{aligned}$$

Bound on \({\mathcal {P}}_4(\varphi )\). By employing the Hölder inequality, the trace estimates (3.5)–(3.6), the Poincaré and inverse inequalities, the embedding \(L^{2\gamma }\hookrightarrow L^2\) for \(\gamma >1\), we easily establish that

$$\begin{aligned} |{\mathcal {P}}_4(\varphi )|\le&Ch^{\frac{1}{2}}\Vert \nabla \varphi \Vert _{L^{m_1}(0,T;L^6(\varOmega ))}\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\\ \le&Ch^{\frac{1}{2}}\Vert \nabla \varphi \Vert _{L^{m_1}(0,T;L^6(\varOmega ))}. \end{aligned}$$

Combining the above analysis, we have the required estimate (4.3). The proof is thus complete. \(\square \)

5 Basic Estimates

This section establishes the boundedness of discrete time derivative \(D_t\rho _h\) and a priori estimate of discrete density \(\rho _h\) in \(L^2(0,T;{\mathbb {Q}}_h)\).

Lemma 5.1

Suppose that the conditions of Lemma 4.2 are satisfied, then the discrete time derivative \(D_t\rho _h\) satisfies

$$\begin{aligned} \Vert D_t\rho _h\Vert _{L^{m_2}(0,T;W^{-1,\frac{6}{5}}(\varOmega ))}\le C,\quad 1<m_2:=\frac{4\gamma }{3\gamma +1}. \end{aligned}$$
(5.1)

Proof

Let \(\phi \in L^{2m_1}(0,T;W^{1,6}(\varOmega ))\) such that \(\Vert \phi \Vert _{L^{2m_1}(0,T;W^{1,6}(\varOmega ))}=1\). Taking \(\varphi _h=\varPi _h^{{\mathbb {Q}}}[\phi ]\) in (3.7) and summing this result with respect to n from \(n=1\) to \(n=N\), applying the same argument as Lemma 4.1, we infer that

$$\begin{aligned} \int _0^T{\int _{\varOmega }{D_t\rho _h\phi }dx}dt=\int _0^T{\int _{\varOmega }{\rho _h\varvec{u}_h\cdot \nabla \phi }dx}dt+\int _0^T{\int _{\varOmega }{{\mathscr {R}}_h\cdot \nabla \phi }dx}dt. \end{aligned}$$

Using Hölder inequality, Lemmas 4.2 and 3.6, we conclude that

$$\begin{aligned} |\int _0^T{\int _{\varOmega }{D_t\rho _h\phi }dx}dt|\le&C\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{L}^6(\varOmega ))}\Vert \nabla \phi \Vert _{L^{m_1}(0,T;L^3(\varOmega ))}\\&+C\Vert \nabla \phi \Vert _{L^{2m_1}(0,T;L^6(\varOmega ))}\le C\Vert \phi \Vert _{L^{2m_1}(0,T;W^{1,6}(\varOmega ))}. \end{aligned}$$

This inequality immediately implies Lemma 5.1. The proof is thus complete. \(\square \)

Lemma 5.2

Suppose that Hypothesis 2.1 and the CFL condition \(\tau \approx h\) are satisfied, there exists \(\epsilon _0>0\) and \(0<\delta <1\) such that for any \(0<\epsilon <\epsilon _0\),

$$\begin{aligned} \int _0^T{\Vert \rho _h\Vert _{{\mathbb {Q}}_h}^2}dt\le Ch^{-2\delta }. \end{aligned}$$
(5.2)

Proof

We divide our proof in two steps. Firstly, if \(1<\gamma \le 2\), by applying Cauchy–Schwarz inequality, we obtain

$$\begin{aligned} \int _0^T{\Vert \rho _h\Vert _{{\mathbb {Q}}_h}^2}dt\le \sqrt{{\mathcal {P}}_5}\times \sqrt{{\mathcal {P}}_6}, \end{aligned}$$
(5.3)

where \({\mathcal {P}}_5\) and \({\mathcal {P}}_6\) are given by

$$\begin{aligned} {\mathcal {P}}_5:=&h^{\epsilon -1}\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{{\mathcal {H}}''({\overline{\eta }}_{\rho ,h})\llbracket {\rho _h}\rrbracket ^2}dS}}dt,\quad \\ {\mathcal {P}}_6:=&h^{-(\epsilon +1)}\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}{\int _{F}{({\mathcal {H}}''({\overline{\eta }}_{\rho ,h}))^{-1}\llbracket {\rho _h}\rrbracket ^2}dS}}dt. \end{aligned}$$

According to the trace estimate (3.6) and the inequality (4.5), we infer that

$$\begin{aligned} {\mathcal {P}}_6\le Ch^{-(\epsilon +2)}\Vert \rho _h\Vert _{L^{4-\gamma }((0,T)\times \varOmega )}^{4-\gamma }. \end{aligned}$$
(5.4)

On the one hand, for \(1<\gamma <\frac{4}{3}\), it is easy check that \(4-\gamma >2\gamma \). Therefore, by applying the inverse estimate (3.4) and the CFL condition \(\tau \approx h\), we have

$$\begin{aligned} \Vert \rho _h\Vert _{L^{4-\gamma }((0,T)\times \varOmega )}^{4-\gamma }\le Ch^{6-\frac{8}{\gamma }}\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}^{4-\gamma }. \end{aligned}$$
(5.5)

On the other hand, for the case \(\frac{4}{3}\le \gamma <2\), by using the embedding result \(L^{2\gamma }\hookrightarrow L^{4-\gamma }\), we conclude that

$$\begin{aligned} \Vert \rho _h\Vert _{L^{4-\gamma }((0,T)\times \varOmega )}^{4-\gamma }\le C\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}^{4-\gamma }. \end{aligned}$$
(5.6)

Inserting (5.5) and (5.6) into (5.4), using Lemma 3.6, we obtain

$$\begin{aligned} {\mathcal {P}}_6\le \left\{ \begin{aligned}&Ch^{-4(\frac{\epsilon }{4}+\frac{2}{\gamma }-1)},\quad&1<\gamma <\frac{4}{3},\\&Ch^{-4(\frac{\epsilon }{4}+\frac{1}{2})},\quad&\frac{4}{3}\le \gamma \le 2. \end{aligned} \right. \end{aligned}$$
(5.7)

By substituting (5.7) into (5.3), using Lemma 3.7, we get that

$$\begin{aligned} \int _0^T{\Vert \rho _h\Vert _{{\mathbb {Q}}_h}^2}dt\le \left\{ \begin{aligned}&Ch^{-2(\frac{\epsilon }{4}+\frac{2}{\gamma }-1)},\quad&1<\gamma <\frac{4}{3},\\&Ch^{-2(\frac{\epsilon }{4}+\frac{1}{2})},\quad&\frac{4}{3}\le \gamma \le 2. \end{aligned} \right. \end{aligned}$$
(5.8)

Secondly, for the case \(\gamma >2\), by using Lemma  3.8, we easily see that

$$\begin{aligned} \int _0^T{\Vert \rho _h\Vert _{{\mathbb {Q}}_h}^2}dt=h^{-1}\int _0^T{\sum _{F\in {\mathcal {F}}_{h,int}}\int _F{\llbracket {\rho _h}\rrbracket ^2}dS}dt\le Ch^{-\epsilon }. \end{aligned}$$
(5.9)

Combining the inequalities (5.8) and (5.9), we have the required estimate (5.2), where the parameter \(\alpha _0\) and \(\delta \) are given by

$$\begin{aligned} \epsilon _0:=\left\{ \begin{aligned}&8(1-\frac{1}{\gamma }),\quad&1<\gamma<\frac{4}{3},\\&2,\quad&\frac{4}{3}\le \gamma . \end{aligned} \right. ,\quad \delta :=\left\{ \begin{aligned}&\frac{\epsilon }{4}+\frac{2}{\gamma }-1,\quad&1<\gamma<\frac{4}{3},\\&\frac{\epsilon }{4}+\frac{1}{2},\quad&\frac{4}{3}\le \gamma <2,\\&\frac{\epsilon }{2},\quad&\gamma \ge 2. \end{aligned} \right. \end{aligned}$$

It is easy check that \(\delta <1\). The proof is thus complete. \(\square \)

Remark 5.1

(i) In fact, for the case of \(\gamma \ge \frac{4}{3}\), the CFL condition \(\tau \approx h\) is not required for the estimate (5.2). (ii) Lemma 5.2 plays a key role in deriving the discrete version of the effective viscous flux identity. See Lemma 6.7 and Theorem 6.2 for more on why it is needed.

6 Convergence Analysis

In this section, we will prove the family \((\rho _h,\varvec{u}_h,p(\rho _h))\) defined in (3.16)–(3.17) converges to weak solution ( see Definition 2.1). For that purpose, we first need to establish a spatial compactness estimate for Bernardi–Raugel finite element space.

Theorem 6.1

Let q satisfies \(2\le q<6\) and \(\frac{1}{q}=\frac{\theta }{2}+\frac{1-\theta }{6}\), \(\theta \in [0,1]\). For any \(\varvec{v}_h\in {\mathbb {V}}_h\), there exists a constant \(C>0\) such that the following estimate holds

$$\begin{aligned} \Vert \varvec{v}_h(\cdot )-\varvec{v}_h(\cdot -\varvec{\xi })\Vert _{\varvec{L}^q({\mathbb {R}}^d)}\le C|\varvec{\xi }|^{\theta }\Vert \nabla \varvec{v}_h\Vert _{\varvec{L}^2(\varOmega )},\quad \forall \;\varvec{\xi }\in {\mathbb {R}}^d. \end{aligned}$$

The proof of Theorem 6.1 can been found in “Appendix A.2”.

According to Lemma  3.6 and Theorem 3.2, we can assert the existence of functions

$$\begin{aligned} \rho \in L^\infty (0,T; L^\gamma (\varOmega ))\cap L^{2\gamma }((0,T)\times \varOmega ),\quad \varvec{u}\in L^2(0,T;\varvec{H}_0^1(\varOmega )) \end{aligned}$$

such that the family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) exists suitable subsequences satisfy

$$\begin{aligned} \left\{ \begin{aligned}&\rho _h\rightharpoonup ^{*}\rho ,\;{\text {in}}\; L^\infty (0,T; L^\gamma (\varOmega ))\cap L^{2\gamma }((0,T)\times \varOmega ),\\&p(\rho _h)\rightharpoonup p,\;{\text {in}}\; L^2((0,T)\times \varOmega ),\quad \varvec{u}_h\rightharpoonup \varvec{u},\;{\text {in}}\; \varvec{L}^2((0,T)\times \varOmega ),\\&{\text {div}}\varvec{u}_h\rightharpoonup {\text {div}}\varvec{u},\;{\text {in}}\; L^2((0,T)\times \varOmega ),\quad \varvec{u}_h\rightharpoonup \varvec{u},\;{\text {in}}\; L^2(0,T;\varvec{H}^1(\varOmega )). \end{aligned} \right. \end{aligned}$$
(6.1)

The following lemma can be found in [26, Lemma 2.3].

Lemma 6.1

Let \(\{f_h\}_{h>1}^\infty \) and \(\{g_h\}_{h>1}^\infty \) be two function sequences on \((0,T)\times \varOmega \) such that

  1. (i)

    \(f_h\) and \(g_h\) converge weakly to f and g respectively in \(L^{p_1}(0,T;L^{q_1}(\varOmega ))\) and \(L^{p_2}(0,T;L^{q_2}(\varOmega ))\), where \(1\le p_1,q_1\le \infty \), \(\frac{1}{p_1}+\frac{1}{p_2}=\frac{1}{q_1}+\frac{1}{q_2}=1\).

  2. (ii)

    Assume that \(\frac{g_h(t,\varvec{x})-g_h(t-h,\varvec{x})}{h}\) is bounded in \(L^1(0,T;W^{-m,1}(\varOmega ))\), for some \(m\ge 0\) independent of h. And \(\Vert f_h(t,\varvec{x})-f_h(t,\varvec{x}-\varvec{\xi })\Vert _{L^{p_1}(0,T;L^{p_2}(\varOmega ))}\rightarrow 0\) as \(|\varvec{\xi }|\rightarrow 0\) uniformly in h.

Then, \(f_hg_h\) converges to fg in the sense of distributions on \((0,T)\times \varOmega \).

Next, we present a weak convergent results for \(\rho _h\varvec{u}_h\).

Lemma 6.2

Suppose that the condition of Lemma 4.2 are satisfied, then the family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) satisfies

$$\begin{aligned} \rho _h\varvec{u}_h\rightharpoonup \rho \varvec{u},\;{\text {in}}\;{\mathcal {D}}'((0,T)\times \varOmega ). \end{aligned}$$

Proof

From Lemma 5.1, we can show

$$\begin{aligned} D_t\rho _h\in L^1(0,T; W^{-1,1}(\varOmega )). \end{aligned}$$
(6.2)

By applying Theorem 6.1 and Lemma 3.6, we conclude that

$$\begin{aligned} \Vert \varvec{u}_h(t,\varvec{x})-\varvec{u}_h(t,\varvec{x}-\varvec{\xi })\Vert _{L^2(0,T;\varvec{L}^2({\mathbb {R}}^d))}{\mathop {\longrightarrow }\limits ^{|\varvec{\xi }|\rightarrow 0}} 0. \end{aligned}$$
(6.3)

By substituting (6.1)–(6.3) into Lemma 6.1, the proof is thus complete. \(\square \)

6.1 Limit in the Compressible Stokes Equations

In this subsection, we can show the limit \((\rho ,\varvec{u},p)\) constructed in (6.1) is a weak solution of Definition 2.1. The remaining major difficulty is to prove the pressure \(p(\rho _h)\rightarrow p(\rho )\).

Lemma 6.3

Suppose that the condition of Lemma 4.2 is satisfied, then the accumulation point \((\rho ,\varvec{u})\) constructed in (6.1) satisfies the weak formulation (2.1).

Proof

We pass to the limit with \(h,\tau \rightarrow 0\) in the consistency formulation (4.1). Firstly, we rewrite the discrete time derivative term

$$\begin{aligned} \int _0^T{\int _{\varOmega }{D_t\rho _h\varphi }dx}dt=&-\int _0^T{\int _{\varOmega }{\rho _hD_t\varphi (t+\tau ,\cdot )}dx}dt+\frac{1}{\tau }\int _{T-\tau }^{T}{\int _{\varOmega }{\rho _h(t,\cdot )\varphi (t+\tau ,\cdot )}dx}dt\\&-\frac{1}{\tau }\int _{-\tau }^{0}{\int _{\varOmega }{\rho _h(t,\cdot )\varphi (t+\tau ,\cdot )}dx}dt\\ =&-\int _0^T{\int _{\varOmega }{\rho _hD_t\varphi (t+\tau ,\cdot )}dx}dt-\int _{\varOmega }{\rho _h^0\varphi (0,\cdot )}dx\\&-\int _0^{\tau }{\int _{\varOmega }{\rho _h^0\frac{\varphi (t,\cdot )-\varphi (0,\cdot )}{\tau }}dx}dt\\ =&-\int _0^T{\int _{\varOmega }{\rho _h(\partial _t\varphi (t,\cdot )+\frac{\tau }{2}\partial _{tt}\varphi (t_{\dag },\cdot ))}dx}dt-\int _{\varOmega }{\rho _h^0\varphi (0,\cdot )}dx\\&-\int _0^{\tau }{\int _{\varOmega }{\rho _h^0\partial _t\varphi (t_{\ddag },\cdot )}dx}dt, \end{aligned}$$

where \(t_{\dag }\in (t,t+\tau )\) and \(t_{\ddag }\in (0,\tau )\). By applying Lemma  3.6 and the embedding \(L^\gamma \hookrightarrow L^1\), we have

$$\begin{aligned} \frac{\tau }{2}\left| \int _0^T{\int _{\varOmega }{\rho _h\partial _{tt}\varphi (t_{\dag },\cdot )}dx}dt\right| \le&C\tau \Vert \rho _h\Vert _{L^1((0,T)\times \varOmega )} \Vert \partial _{tt}\varphi \Vert _{L^\infty ((0,T)\times \varOmega )}\\ \le&C\tau \Vert \rho _h\Vert _{L^\infty (0,T;L^\gamma (\varOmega ))} \Vert \partial _{tt}\varphi \Vert _{L^\infty ((0,T)\times \varOmega )}{\mathop {\longrightarrow }\limits ^{h,\tau \rightarrow 0}}0,\\ \left| \int _0^{\tau }{\int _{\varOmega }{\rho _h\partial _t\varphi (t_{\ddag },\cdot )}dx}dt\right| \le&C\tau \Vert \rho _h^0\Vert _{L^1(\varOmega )} \Vert \partial _t\varphi \Vert _{L^\infty ((0,T)\times \varOmega )}\\ \le&C\tau \Vert \rho _0\Vert _{L^\gamma (\varOmega )} \Vert \partial _t\varphi \Vert _{L^\infty ((0,T)\times \varOmega )}{\mathop {\longrightarrow }\limits ^{h,\tau \rightarrow 0}}0. \end{aligned}$$

According to (6.1) and \(\varPi _h^{{\mathbb {Q}}}[\rho _0]\rightharpoonup \rho _0\) in \(L^\gamma (\varOmega )\), we obtain

$$\begin{aligned}&-\int _0^T{\int _{\varOmega }{\rho _h\partial _t\varphi (t,\cdot )}dx}dt-\int _{\varOmega }{\rho _h^0\varphi (0,\cdot )}dx\\&{\mathop {\longrightarrow }\limits ^{h,\tau \rightarrow 0}}-\int _0^T{\int _{\varOmega }{\rho \partial _t\varphi }dx}dt-\int _{\varOmega }{\rho _0\varphi (0,\cdot )}dx. \end{aligned}$$

Next, by applying Lemma 6.2, we can show

$$\begin{aligned} \int _0^T{\int _{\varOmega }{\rho _h\varvec{u}_h\cdot \nabla \varphi }dx}dt{\mathop {\longrightarrow }\limits ^{h,\tau \rightarrow 0}}\int _0^T{\int _{\varOmega }{\rho \varvec{u}\cdot \nabla \varphi }dx}dt. \end{aligned}$$

Finally, by employing the inequality (4.3) of Lemma 4.2, we conclude that

$$\begin{aligned} \int _0^T{\int _{\varOmega }{{\mathscr {R}}_h\cdot \nabla \varphi }dx}dt{\mathop {\longrightarrow }\limits ^{h,\tau \rightarrow 0}}0. \end{aligned}$$

Combining the above analysis, the proof is thus complete. \(\square \)

Lemma 6.4

Suppose that Hypothesis 2.1 is satisfied, the accumulation limit \((\rho ,\varvec{u})\) constructed in (6.1) satisfies the following weak formulation:

$$\begin{aligned}&\mu \int _0^T{\int _{\varOmega }{\nabla \varvec{u}:\nabla \varvec{v}}dx}dt+(\lambda +\mu )\int _0^T{\int _{\varOmega }{{\text {div}}\varvec{u}{\text {div}}\varvec{v}}dx}dt\\&-\int _0^T{\int _{\varOmega }{p{\text {div}}\varvec{v}}dx}dt=0,\quad \forall \;\varvec{v}\in \varvec{C}_0^\infty ((0,T)\times \varOmega ).\nonumber \end{aligned}$$
(6.4)

Proof

We define \({\mathcal {F}}_{{\mathbb {V}}_h}\) as the \(\varvec{L}^2\)-orthogonal projection operator from \(\varvec{L}^2(\varOmega )\) into \({\mathbb {V}}_h\). For any \(\varvec{v}\in \varvec{C}_0^\infty ((0,T)\times \varOmega )\), we can choose \(\varvec{v}_h={\mathcal {F}}_{{\mathbb {V}}_h}\varvec{v}\) and \(\varvec{v}_h^n=\frac{1}{\tau }\int _{t_{n-1}}^{t_n}{\varvec{v}_h}dt\) such that

$$\begin{aligned} \varvec{v}_h(t,\cdot ){\mathop {\longrightarrow }\limits ^{h\rightarrow 0}}\varvec{v}(t,\cdot ),\;{\text {in}}\;\varvec{H}_0^1(\varOmega ), \end{aligned}$$
(6.5)

for any \(t\in (0,T)\). Taking \(\varvec{v}_h^n\) in (3.8), multiplying by \(\tau \) and summing the results with respect to n from \(n=1\) to \(n=N\), we conclude that

$$\begin{aligned} \int _0^T{\int _{\varOmega }{[\mu \nabla \varvec{u}_h:\nabla \varvec{v}_h+(\lambda +\mu ){\text {div}}\varvec{u}_h{\text {div}}\varvec{v}_h]}dx}dt\\ -\int _0^T{\int _{\varOmega }{p(\rho _h){\text {div}}\varvec{v}_h}dx}dt=0. \end{aligned}$$

Obviously, by applying (6.1) and (6.5), we have the required weak formulation (6.4). The proof is thus complete. \(\square \)

6.2 Strong Convergence of the Density

The strong convergence of the density is proved by the discrete version of the weak continuity property of the effective viscous flux introduced on the continuous level in [30]. For this purpose, we first introduce the following notation

$$\begin{aligned} {\text {curl}}\varvec{v}=\frac{\partial v_2}{\partial x_1}-\frac{\partial v_1}{\partial x_2},\;{\text {if}}\; d=2,\;{\text {curl}}\varvec{v}=\left[ \frac{\partial v_3}{\partial x_2}-\frac{\partial v_2}{\partial x_3},\;\frac{\partial v_1}{\partial x_3}-\frac{\partial v_3}{\partial x_1},\;\frac{\partial v_2}{\partial x_1}-\frac{\partial v_1}{\partial x_2}\right] ,\;{\text {if}}\; d=3, \end{aligned}$$

where \(\varvec{v}\) is a vector-valued function. Obviously, if \(\varvec{v}\in \varvec{H}^1(\varOmega )\) and \(\varvec{w}\in \varvec{H}_0^1(\varOmega )\), we can show

$$\begin{aligned} \int _{\varOmega }{\nabla \varvec{v}:\nabla \varvec{w}}dx=\int _{\varOmega }{{\text {div}}\varvec{v}{\text {div}}\varvec{w}}dx+\int _{\varOmega }{{\text {curl}}\varvec{v}{\text {curl}}\varvec{w}}dx. \end{aligned}$$
(6.6)

Next, we report the following Lemma, which plays a key role in deriving the discrete version of the effective viscous flux.

Lemma 6.5

Let \(\varOmega \subset {\mathbb {R}}^d\) (\(d=2,3\)) be a bounded open set. For any \(1<r<\infty \) and \(q\in L^r(\varOmega )\), there exists \(\varvec{w}\in \varvec{W}^{1,r}(\varOmega )\) such that

$$\begin{aligned} {\text {div}}\varvec{w}=q,\quad {\text {curl}}\varvec{w}=0,\;{\text {a.e.}}\;{\text {in}}\;\varOmega ,\quad \Vert \varvec{w}\Vert _{\varvec{W}^{1,r}(\varOmega )}\le C\Vert q\Vert _{L^r(\varOmega )}, \end{aligned}$$

where C only depends on \(\varOmega \) and r. Moreover, if \(q\in W^{1,r}(\varOmega )\) (or \(q\in W^{-1,r}(\varOmega )\)), it is possible to have \(\varvec{w}\in \varvec{W}^{2,r}(\varOmega )\) (or \(\varvec{w}\in \varvec{L}^r(\varOmega )\)) such that

$$\begin{aligned} \Vert \varvec{w}\Vert _{\varvec{W}^{2,r}(\varOmega )}\le C\Vert q\Vert _{W^{1,r}(\varOmega )},\;({\text {or}}\;\Vert \varvec{w}\Vert _{\varvec{L}^r(\varOmega )}\le C\Vert q\Vert _{W^{-1,r}(\varOmega )}). \end{aligned}$$

Proof

It is easy to check that \(\nabla \varDelta ^{-1}[q]\) can be served as the desired solution, where \(\varDelta ^{-1}\) is the inverse of the Laplacian on \({\mathbb {R}}^3\), and here we applied to q extended by 0 outside \(\varOmega \). Obviously, \(\nabla \varDelta ^{-1}\) is a continuous linear operator from \(L^r(\varOmega )\) to \(\varvec{W}^{1,r}(\varOmega )\) and from \(W^{1,r}(\varOmega )\) to \(\varvec{W}^{2,r}(\varOmega )\), from \(W^{-1,r}(\varOmega )\) to \(\varvec{L}^r(\varOmega )\) (see e.g., [29, Lemma 8.3]). The proof is thus complete. \(\square \)

In the next step, we introduce the operator \(\varPi _h^{{\mathbb {Y}}}:{\mathbb {Q}}_h\mapsto {\mathbb {Y}}_h\) which interpolates the piecewise constant functions to the space of continuous finite element space \({\mathbb {Y}}_h\),

$$\begin{aligned} \forall \; q_h\in {\mathbb {Q}}_h,\quad \varPi _h^{{\mathbb {Y}}}[q_h](A):=\frac{1}{{\text {card}}(N_A)}\sum _{K\in N_A}q_h|_{K}, \end{aligned}$$

for any vertices A in the discretization, where \(N_A\) is the set of elements \(K\in {\mathcal {T}}_h\) of which takes A as its vertices. The operator \(\varPi _h^{{\mathbb {Y}}}\) satisfies the following results (see e.g., [10, Lemma 5.8]).

Lemma 6.6

For any \(q_h\in {\mathbb {Q}}_h\), there exists a constant \(C>0\), depending only on the shape-regularity of \({\mathcal {T}}_h\) such that

$$\begin{aligned} \Vert \nabla \varPi _h^{{\mathbb {Y}}}[q_h]\Vert _{L^2(\varOmega )}\le C\Vert q_h\Vert _{{\mathbb {Q}}_h},\quad \Vert q_h-\varPi _h^{{\mathbb {Y}}}[q_h]\Vert _{L^2(\varOmega )}\le Ch\Vert q_h\Vert _{{\mathbb {Q}}_h}. \end{aligned}$$

Then we can prove the following estimates.

Lemma 6.7

Suppose that the condition of Lemma 5.2 is satisfied, there exists a constant \(C>0\) such that the following estimates hold

$$\begin{aligned} \int _0^T{\Vert \rho _h-\varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2(\varOmega )}^2}dt\le&Ch^{2(1-\delta )},\quad \int _0^T{\Vert \varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2(\varOmega )}^2}dt\le C,\\ \int _0^T{\Vert \varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{H^1(\varOmega )}^2}dt\le&Ch^{-2\delta }. \end{aligned}$$

Proof

By applying the inequality (5.2) and Lemma 6.6, we can show

$$\begin{aligned} \int _0^T{\Vert \rho _h-\varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2(\varOmega )}^2}dt\le Ch^{2(1-\delta )},\quad \int _0^T{\Vert \nabla \varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2(\varOmega )}^2}dt\le Ch^{-2\delta }. \end{aligned}$$

According to the embedding \(L^{2\gamma }\hookrightarrow L^2\) for \(\gamma >1\), we have

$$\begin{aligned} \int _0^T{\Vert \varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2(\varOmega )}^2}dt\le&2\int _0^T{\Vert \rho _h-\varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2(\varOmega )}^2}dt+2\int _0^T{\Vert \rho _h\Vert _{L^2(\varOmega )}^2}dt \\ \le&Ch^{2(1-\delta )}+C\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}^2\le C. \end{aligned}$$

These inequalities immediately implies

$$\begin{aligned} \int _0^T{\Vert \varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{H^1(\varOmega )}^2}dt\le Ch^{-2\delta }. \end{aligned}$$

Combining the above analysis, the proof is thus complete. \(\square \)

Theorem 6.2

Suppose that the condition of Lemma 5.2 is satisfied. The family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) and the accumulation limit \((\rho ,\varvec{u})\) constructed in (6.1) satisfy the following convergence properties:

$$\begin{aligned} \lim _{h,\tau \rightarrow 0}&\int _0^T{\psi \int _{\varOmega }{((\lambda +2\mu ){\text {div}}\varvec{u}_h-p(\rho _h))\rho _h\varphi }dx}dt\nonumber \\ =&\int _0^T{\psi \int _{\varOmega }{((\lambda +2\mu ){\text {div}}\varvec{u}-p)\rho \varphi }dx}dt. \end{aligned}$$
(6.7)

for any \(\psi \in C_0^\infty ((0,T))\) and \(\varphi \in C_0^\infty (\varOmega )\).

Proof

According to Lemmas 6.56.7 and 3.6, the inequality (5.1), and there exists \(\varvec{w}_{{\mathbb {Y}},h}\in L^2(0,T;\varvec{H}^2(\varOmega ))\) and \(\varvec{w}_h\in L^2(0,T;\varvec{H}^1(\varOmega ))\) such that

$$\begin{aligned} \left\{ \begin{aligned} {\text {div}}\varvec{w}_{{\mathbb {Y}},h}&=\varPi _h^{{\mathbb {Y}}}[\rho _h],\;{\text {in}}\;(0,T)\times \varOmega ,\\ {\text {curl}}\varvec{w}_{{\mathbb {Y}},h}&=0,\;{\text {in}}\;(0,T)\times \varOmega , \end{aligned} \right. \quad \left\{ \begin{aligned} {\text {div}}\varvec{w}_h&=\rho _h,\;{\text {in}}\;(0,T)\times \varOmega ,\\ {\text {curl}}\varvec{w}_h&=0,\;{\text {in}}\;(0,T)\times \varOmega , \end{aligned} \right. \end{aligned}$$
(6.8)

and a generic constant C independent of h and \(\tau \) such that

$$\begin{aligned} \left\{ \begin{aligned} \Vert \varvec{w}_{{\mathbb {Y}},h}\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\le&C,\\ \Vert \varvec{w}_{{\mathbb {Y}},h}\Vert _{L^2(0,T;\varvec{H}^2(\varOmega ))}\le&Ch^{-\delta } \end{aligned} \right. \quad \left\{ \begin{aligned} \Vert \varvec{w}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\le&C,\\ \Vert D_t\varvec{w}_h\Vert _{L^{m_2}(0,T;\varvec{L}^\frac{6}{5}(\varOmega ))}\le&C. \end{aligned} \right. \end{aligned}$$
(6.9)

Subtracting the right side of (6.8) from its left side, we obtain

$$\begin{aligned} {\text {div}}(\varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h)=\varPi _h^{{\mathbb {Y}}}[\rho _h]-\rho _h,\quad \textbf{curl}(\varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h)=0,\;{\text {in}}\;(0,T)\times \varOmega . \end{aligned}$$

By employing Lemmas 6.5 and 6.7, we can show

$$\begin{aligned} \Vert \varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\le C\Vert \rho _h-\varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2((0,T)\times \varOmega )}\le Ch^{1-\delta }. \end{aligned}$$
(6.10)

Using the Lemma A.4 together with the estimates (6.9), we deduce for a suitable subsequence that

$$\begin{aligned} \varvec{w}_h\rightarrow \varvec{w},\;{\text {in}}\;\;\varvec{L}^2((0,T)\times \varOmega ). \end{aligned}$$
(6.11)

In addition, the accumulation limit \(\varvec{w}\) satisfies the following properties

$$\begin{aligned} {\text {div}}\varvec{w}=\rho ,\quad {\text {curl}}\varvec{w}=0,\;{\text {in}}\;(0,T)\times \varOmega . \end{aligned}$$
(6.12)

Taking \(\varvec{v}_h=\varPi _h^{{\mathbb {V}}}[\varphi \varvec{w}_{{\mathbb {Y}},h}]\) in (3.8), multiplying by \(\psi \in C_0^\infty ((0,T))\) and integrating from \(t=0\) to T, we derive

$$\begin{aligned}&\mu \int _0^T{\psi \int _{\varOmega }{\nabla \varvec{u}_h:\nabla (\varphi \varvec{w}_{{\mathbb {Y}},h})}dx}dt+(\lambda +\mu )\int _0^T{\psi \int _{\varOmega }{{\text {div}}\varvec{u}_h{\text {div}}(\varphi \varvec{w}_{{\mathbb {Y}},h})}dx}dt\nonumber \\&-\int _0^T{\psi \int _{\varOmega }{p(\rho _h){\text {div}}(\varphi \varvec{w}_{{\mathbb {Y}},h})}dx}dt={\mathscr {R}}_{1,h}, \end{aligned}$$
(6.13)

where \({\mathscr {R}}_{1,h}\) is given by

$$\begin{aligned} {\mathscr {R}}_{1,h}:=&\mu \int _0^T{\psi \int _{\varOmega }{\nabla \varvec{u}_h:\nabla (\varphi \varvec{w}_{{\mathbb {Y}},h}-\varPi _h^{{\mathbb {V}}}[\varphi \varvec{w}_{{\mathbb {Y}},h}])}dx}dt\\&+(\lambda +\mu )\int _0^T{\psi \int _{\varOmega }{{\text {div}}\varvec{u}_h{\text {div}}(\varphi \varvec{w}_{{\mathbb {Y}},h}-\varPi _h^{{\mathbb {V}}}[\varphi \varvec{w}_{{\mathbb {Y}},h}])}dx}dt. \end{aligned}$$

By applying the inequalities (3.2) and (6.9), we can show

$$\begin{aligned} |{\mathscr {R}}_{1,h}|\le&C\Vert \psi \Vert _{L^\infty ((0,T))}\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\Vert \varphi \varvec{w}_{{\mathbb {Y}},h}-\varPi _h^{{\mathbb {V}}}[\varphi \varvec{w}_{{\mathbb {Y}},h}]\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\nonumber \\ \le&Ch\Vert \psi \Vert _{L^\infty ((0,T))}\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\Vert \varphi \varvec{w}_{{\mathbb {Y}},h}\Vert _{L^2(0,T;\varvec{H}^2(\varOmega ))}\nonumber \\ \le&Ch^{1-\delta }\Vert \psi \Vert _{L^\infty ((0,T))}\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\Vert \varphi \Vert _{W^{2,\infty }(\varOmega )}. \end{aligned}$$
(6.14)

Obviously, we have \({\text {div}}(\varphi \varvec{w}_{{\mathbb {Y}},h})=\varPi _h^{{\mathbb {Y}}}[\rho _h]\varphi +\varvec{w}_{{\mathbb {Y}},h}\cdot \nabla \varphi \) and \({\text {curl}}(\varphi \varvec{w}_{{\mathbb {Y}},h})=J(\varphi )\varvec{w}_{{\mathbb {Y}},h}\), where \(J(\varphi )\) is a matrix with entries involving some first-order derivatives of \(\varphi \). Combining the identities (6.6) and (6.13), we obtain

$$\begin{aligned}&\int _0^T{\psi \int _{\varOmega }{((\lambda +2\mu ){\text {div}}\varvec{u}_h-p(\rho _h))\rho _h\varphi }dx}dt\nonumber \\&\quad =-(\lambda +2\mu )\int _0^T{\psi \int _{\varOmega }{{\text {div}}\varvec{u}_h\varvec{w}_{{\mathbb {Y}},h}\cdot \nabla \varphi }dx}dt-\mu \int _0^T{\psi \int _{\varOmega }{{\text {curl}}\varvec{u}_h\cdot J(\varphi )\varvec{w}_{{\mathbb {Y}},h}}dx}dt\nonumber \\&\qquad +\int _0^T{\psi \int _{\varOmega }{p(\rho _h)\varvec{w}_{{\mathbb {Y}},h}\cdot \nabla \varphi }dx}dt+{\mathscr {R}}_{1,h}+{\mathscr {R}}_{2,h}, \end{aligned}$$
(6.15)

where \({\mathscr {R}}_{2,h}\) is defined by

$$\begin{aligned} {\mathscr {R}}_{2,h}:=\int _0^T{\psi \int _{\varOmega }{((\lambda +2\mu ){\text {div}}\varvec{u}_h-p(\rho _h))(\rho _h\varphi -\varPi _h^{{\mathbb {Y}}}[\rho _h]\varphi )}dx}dt. \end{aligned}$$

Applying the Hölder inequality and Lemma 6.7 implies

$$\begin{aligned} |{\mathscr {R}}_{2,h}|\le&C\Vert {\text {div}}\varvec{u}_h\Vert _{L^2((0,T)\times \varOmega )}\Vert \rho _h-\varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2((0,T)\times \varOmega )}\nonumber \\&+C\Vert p(\rho _h)\Vert _{L^2((0,T)\times \varOmega )}\Vert \rho _h-\varPi _h^{{\mathbb {Y}}}[\rho _h]\Vert _{L^2((0,T)\times \varOmega )}\nonumber \\ \le&Ch^{1-\delta }\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}+Ch^{1-\delta }\Vert p(\rho _h)\Vert _{L^2((0,T)\times \varOmega )}. \end{aligned}$$
(6.16)

The identity (6.15) can be rewritten as

$$\begin{aligned}&\int _0^T{\psi \int _{\varOmega }{((\lambda +2\mu ){\text {div}}\varvec{u}_h-p(\rho _h))\rho _h\varphi }dx}dt\nonumber \\&\quad =-(\lambda +2\mu )\int _0^T{\psi \int _{\varOmega }{{\text {div}}\varvec{u}_h\varvec{w}_h\cdot \nabla \varphi }dx}dt-\mu \int _0^T{\psi \int _{\varOmega }{{\text {curl}}\varvec{u}_h\cdot J(\varphi )\varvec{w}_h}dx}dt\nonumber \\&\qquad +\int _0^T{\psi \int _{\varOmega }{p(\rho _h)\varvec{w}_h\cdot \nabla \varphi }dx}dt+{\mathscr {R}}_{1,h}+{\mathscr {R}}_{2,h}+{\mathscr {R}}_{3,h}. \end{aligned}$$
(6.17)

where \({\mathscr {R}}_{3,h}\) is given by

$$\begin{aligned} {\mathscr {R}}_{3,h}:=&-(\lambda +2\mu )\int _0^T{\psi \int _{\varOmega }{{\text {div}}\varvec{u}_h(\varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h)\cdot \nabla \varphi }dx}dt\\&-\mu \int _0^T{\psi \int _{\varOmega }{{\text {curl}}\varvec{u}_h\cdot J(\varphi )(\varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h)}dx}dt\\&+\int _0^T{\psi \int _{\varOmega }{p(\rho _h)(\varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h)\cdot \nabla \varphi }dx}dt. \end{aligned}$$

Using the Hölder inequality and the estimate (6.10), we conclude that

$$\begin{aligned} |{\mathscr {R}}_{3,h}|\le&C_{\psi ,\varphi }\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\Vert \varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h\Vert _{L^2(0,T;\varvec{L}^2(\varOmega ))}\nonumber \\&+C_{\psi ,\varphi }\Vert p(\rho _h)\Vert _{L^2(0,T;L^2(\varOmega ))}\Vert \varvec{w}_{{\mathbb {Y}},h}-\varvec{w}_h\Vert _{L^2(0,T;\varvec{L}^2(\varOmega ))}\nonumber \\ \le&C_{\psi ,\varphi }h^{1-\delta }\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}+C_{\psi ,\varphi }h^{1-\delta }\Vert p(\rho _h)\Vert _{L^2(0,T;L^2(\varOmega ))}. \end{aligned}$$
(6.18)

Passing to the limit with \(h,\tau \rightarrow 0\) in (6.17), using (6.1) and (6.11), we find

$$\begin{aligned}&\lim _{h,\tau \rightarrow 0}\int _0^T{\psi \int _{\varOmega }{((\lambda +2\mu ){\text {div}}\varvec{u}_h-p(\rho _h))\rho _h\varphi }dx}dt\nonumber \\&\quad =-(\lambda +2\mu )\int _0^T{\psi \int _{\varOmega }{{\text {div}}\varvec{u}\varvec{w}\cdot \nabla \varphi }dx}dt-\mu \int _0^T{\psi \int _{\varOmega }{{\text {curl}}\varvec{u}\cdot J(\varphi )\varvec{w}}dx}dt\nonumber \\&\qquad +\int _0^T{\psi \int _{\varOmega }{p\varvec{w}\cdot \nabla \varphi }dx}dt+\lim _{h,\tau \rightarrow 0}{\mathscr {R}}_{1,h}+\lim _{h,\tau \rightarrow 0}{\mathscr {R}}_{2,h}+\lim _{h,\tau \rightarrow 0}{\mathscr {R}}_{3,h}. \end{aligned}$$
(6.19)

By applying Lemma  3.6, the estimates (6.14), (6.16), (6.18) and \(\delta <1\), we get that

$$\begin{aligned} \lim _{h,\tau \rightarrow 0}{\mathscr {R}}_{1,h}+\lim _{h,\tau \rightarrow 0}{\mathscr {R}}_{2,h}+\lim _{h,\tau \rightarrow 0}{\mathscr {R}}_{3,h}=0. \end{aligned}$$
(6.20)

Taking \(\varvec{v}=\psi \varphi \varvec{w}\) in (6.4) and using the identity (6.12), imply

$$\begin{aligned}&\int _0^T{\psi \int _{\varOmega }{((\lambda +2\mu ){\text {div}}\varvec{u}-p)\rho \varphi }dx}dt=-(\lambda +2\mu )\int _0^T{\psi \int _{\varOmega }{{\text {div}}\varvec{u}\varvec{w}\cdot \nabla \varphi }dx}dt\nonumber \\&-\mu \int _0^T{\psi \int _{\varOmega }{{\text {curl}}\varvec{u}\cdot J(\varphi )\varvec{w}}dx}dt+\int _0^T{\psi \int _{\varOmega }{p\varvec{w}\cdot \nabla \varphi }dx}dt. \end{aligned}$$
(6.21)

Combining the identities (6.19)–(6.21), we have the required discrete effective viscous flux identity (6.7). The proof is thus complete. \(\square \)

Lemma 6.8

(Strong convergence of \(\rho _h\)) Suppose that the condition of Lemma 5.2 is satisfied, then, passing to a subsequence if necessary

$$\begin{aligned} \rho _h\rightarrow \rho \;{\text {in}}\; L^1((0,T)\times \varOmega ). \end{aligned}$$
(6.22)

Proof

Firstly, we can show the sequences \(p(\rho _h)\rho _h\), \(\log (\rho _h)\rho _h\) and \(\rho _h{\text {div}}\varvec{u}_h\) have the following convergent properties:

$$\begin{aligned} p(\rho _h)\rho _h\rightharpoonup \overline{p(\rho )\rho },\quad \log (\rho _h)\rho _h\rightharpoonup \overline{\log (\rho )\rho },\quad \rho _h{\text {div}}\varvec{u}_h\rightharpoonup \overline{\rho {\text {div}}\varvec{u}}, \end{aligned}$$

in a suitable \(L^q((0,T)\times \varOmega )\) space with \(q>1\), where the overbar is used to denote the weak limit of a nonlinear function. According to the notation introduced above, we write \(p=\overline{p(\rho )}\), then it can be easily checked

$$\begin{aligned}&\lim _{h\rightarrow 0}{\int _0^T{\psi \int _{\varOmega }{\phi \left[ (\lambda +2\mu ){\text {div}}_h\varvec{u}_h-p(\rho _h)\right] \rho _h}dx}dt}\nonumber \\&\quad =\int _0^T{\psi \int _{\varOmega }{\phi [(\lambda +2\mu )\overline{\rho {\text {div}}\varvec{u}}-\overline{p(\rho )\rho }]}dx}dt, \end{aligned}$$
(6.23)

for any \(\psi \in C_0^\infty (0,T)\) and \(\phi \in C_0^\infty (\varOmega )\). By applying the discrete effective viscous flux identity (6.7) and the identity (6.23), we conclude that

$$\begin{aligned} \int _0^T{\psi \int _{\varOmega }{\phi (\overline{\rho {\text {div}}\varvec{u}}-\rho {\text {div}}\varvec{u})}dx}dt=\int _0^T{\psi \int _{\varOmega }{\phi \frac{\overline{p(\rho )\rho }-\overline{p(\rho )}\rho }{\lambda +2\mu }}dx}dt. \end{aligned}$$
(6.24)

Take the following functions sequence \(\psi _m\in C_0^\infty ((0,T))\) and \(\phi _n\in C_0^\infty (\varOmega )\) such that

$$\begin{aligned} \psi _m\ge 0;\; \psi _m\rightarrow 1;\; \psi _m=1,\; \frac{1}{m}\le t\le T-\frac{1}{m},\\ \phi _n\ge 0;\;\phi _n\rightarrow 1;\;\phi _n=1,\;{\text {dist}}(x,\partial \varOmega )\ge \frac{1}{n}. \end{aligned}$$

Let \((\psi ,\phi )=(\psi _m,\phi _n)\) in (6.24) and \(m,n\rightarrow +\infty \), by applying Lebesgue’s dominated convergence theorem, we obtain

$$\begin{aligned} \int _0^T{\int _{\varOmega }{(\overline{\rho {\text {div}}\varvec{u}}-\rho {\text {div}}\varvec{u})}dx}dt=\int _0^T{\int _{\varOmega }{\frac{\overline{p(\rho )\rho }-\overline{p(\rho )}\rho }{\lambda +2\mu }}dx}dt. \end{aligned}$$
(6.25)

For the identity (6.25), by employing Lemma A.1, we get that

$$\begin{aligned} \int _0^T{\int _{\varOmega }{(\overline{\rho {\text {div}}\varvec{u}}-\rho {\text {div}}\varvec{u})}dx}dt\ge 0. \end{aligned}$$
(6.26)

According to Lemmas 2.1 and 6.3, we obtain \((\rho ,\varvec{u})\) is a renormalized solution of the continuity equation (2.1). Therefore, taking \(\varPhi (\rho )=\rho \log (\rho )\) in Definition 2.2 and integrating over \([0,t_F]\times \varOmega \) for the results, we can show

$$\begin{aligned} \int _{\varOmega }{\rho \log (\rho )(t_F,\cdot )}dx+\int _0^{t_F}{\int _{\varOmega }{\rho {\text {div}}\varvec{u}}dx}dt=\int _{\varOmega }{\rho \log (\rho )(0,\cdot )}dx, \end{aligned}$$
(6.27)

for any \(t_F\in [0,T]\).

Taking \(({\mathcal {B}}(\rho ),\varphi _h)=(\rho \log (\rho ),1)\) in the discrete renormalized continuity scheme (3.9) and passing to the limit with \(h,\tau \rightarrow 0\), we have

$$\begin{aligned} \int _{\varOmega }{\overline{\rho \log (\rho )}(t_F,\cdot )}dx+\int _0^{t_F}{\int _{\varOmega }{\overline{\rho {\text {div}}\varvec{u}}}dx}dt\le \int _{\varOmega }{\rho \log (\rho )(0,\cdot )}dx, \end{aligned}$$
(6.28)

for any \(t_F\in [0,T]\). Subtracting the identity (6.27) from the inequality (6.28), we can show

$$\begin{aligned} \int _{\varOmega }{(\overline{\rho \log (\rho )}-\rho \log (\rho ))(t_F,\cdot )}dx\le \int _0^{t_F}{\int _{\varOmega }{(\rho {\text {div}}\varvec{u}-\overline{\rho {\text {div}}\varvec{u}})}dx}dt \end{aligned}$$
(6.29)

for any \(t_F\in [0,T]\). Inserting (6.26) into (6.29), we obtain

$$\begin{aligned} \int _{\varOmega }{(\overline{\rho \log (\rho )}-\rho \log (\rho ))(t_F,\cdot )}dx\le 0, \end{aligned}$$
(6.30)

On the other hand, according to Lemma A.2, we have

$$\begin{aligned} \overline{\rho \log \left( \rho \right) }\ge \rho \log \left( \rho \right) ,\;{\text {a.e.}}\;{\text {in}}\;\left( 0,T\right) \times \varOmega . \end{aligned}$$
(6.31)

Combining the inequalities (6.30) and (6.31) implies

$$\begin{aligned} \overline{\rho \log (\rho )}=\rho \log (\rho ),\;{\text {a.e.}}\;{\text {in}}\;(0,T)\times \varOmega . \end{aligned}$$

By applying Lemma A.3, we have the required result (6.22). The proof is thus complete. \(\square \)

Theorem 6.3

Suppose that the condition of Lemma 5.2 is satisfied. For any \(q_1\in [1,2\gamma )\) and \(q_2\in [1,2)\), then, passing to a subsequence if necessary

$$\begin{aligned} \rho _h\rightarrow \rho \;{\text {in}}\; L^{q_1}((0,T)\times \varOmega ),\quad p(\rho _h)\rightarrow p(\rho )\;{\text {in}}\; L^{q_2}((0,T)\times \varOmega ). \end{aligned}$$

Proof

By applying (6.22) and Lemma 3.6, we have

$$\begin{aligned} \rho _h\rightarrow \rho \;{\text {in}}\; L^{q_1}((0,T)\times \varOmega ),\quad q_1\in [1,2\gamma ). \end{aligned}$$
(6.32)

Noticing \(x^{\gamma }\) and \(x^{\frac{1}{\gamma }}\) are increasing functions for \(x\in {\mathbb {R}}_+\) and \(\left( x-y\right) ^\vartheta \le x^\vartheta -y^\vartheta \) for \(x\ge y\ge 0\) and \(\vartheta >0\), we obtain

$$\begin{aligned} F_h:=(p(\rho _h)-p(\rho ))(\rho _h-\rho )\ge a|\rho _h^{\gamma }-\rho ^{\gamma }|^{\frac{1}{\gamma }+1},\;{\text {in}}\;(0,T)\times \varOmega . \end{aligned}$$
(6.33)

By employing the Hölder inequality, (6.32) and Lemma 3.6, we can show

$$\begin{aligned} \int _0^T{\int _{\varOmega }{F_h}dx}dt\le&\Vert p(\rho _h)\Vert _{L^2((0,T)\times \varOmega )}\Vert \rho _h-\rho \Vert _{L^2((0,T)\times \varOmega )}\nonumber \\&+C\Vert \rho \Vert _{L^{2\gamma }((0,T)\times \varOmega )}^\gamma \Vert \rho _h-\rho \Vert _{L^2((0,T)\times \varOmega )}. \end{aligned}$$
(6.34)

Inserting (6.33) into (6.34), using (6.32), we can show

$$\begin{aligned} \lim _{h,\tau \rightarrow 0}\int _0^T{\int _{\varOmega }{|\rho _h^{\gamma }-\rho ^{\gamma }|^{\frac{1}{\gamma }+1}}dx}dt\le \frac{1}{a}\lim _{h,\tau \rightarrow 0}\int _0^T{\int _{\varOmega }{F_h}dx}dt=0, \end{aligned}$$

which implies that

$$\begin{aligned} p(\rho _h)\rightarrow p(\rho ),\;{\text {in}}\; L^1((0,T)\times \varOmega ). \end{aligned}$$
(6.35)

By applying (6.35) and Lemma 3.6, we conclude that

$$\begin{aligned} p(\rho _h)\rightarrow p(\rho )\;{\text {in}}\; L^{q_2}((0,T)\times \varOmega ), \end{aligned}$$

where \(q_2\in [1,2)\). The proof is thus complete. \(\square \)

Combining Lemmas 6.3 and 6.4, and Theorem 6.3, we can obtain the main result of the first part of this paper:

Theorem 6.4

Let \(\varOmega \subset {\mathbb {R}}^d\), \(d=2,3\) be a bounded domain and assume that the viscosity coefficients \(\mu \) and \(\lambda \) satisfy \(\mu >0\) and \(d\lambda +2\mu >0\). Suppose that the pressure \(p=p(\rho )\) satisfies the assumption (1.2) with \(\gamma >1\). Furthermore, the initial values \(\rho _0\) satisfies Hypothesis 2.1. The family \((\rho _h,\varvec{u}_h)\) defined in (3.16)–(3.17) satisfies \(\rho _h>0\) for any \(h,\tau >0\) with \(\tau \approx h\) and \(0<\epsilon <\epsilon _0\). Then we have the following convergent properties:

$$\begin{aligned}&\varvec{u}_h\rightharpoonup \varvec{u}\;{\text {in}}\; L^2(0,T;\varvec{H}^1(\varOmega )),{} & {} \rho _h\rightharpoonup ^{*}\rho \;{\text {in}}\; L^\infty (0,T;L^\gamma (\varOmega )),\\&\rho _h\rightharpoonup \rho \;{\text {in}}\; L^{2\gamma }((0,T)\times \varOmega ),{} & {} p(\rho _h)\rightharpoonup p(\rho )\;{\text {in}}\; L^2((0,T)\times \varOmega ),\\&\rho _h\rightarrow \rho \;{\text {in}}\; L^{q_1}((0,T)\times \varOmega ),{} & {} p(\rho _h)\rightarrow p(\rho )\;{\text {in}}\; L^{q_2}((0,T)\times \varOmega ), \end{aligned}$$

for any \(1\le q_1<2\gamma \) and \(1\le q_2<2\), where \((\rho ,\varvec{u})\) is a weak solution of the semi-stationary compressible Stokes equations (1.1)–(1.4) in the sense of Definition 2.1.

Remark 6.1

(i) Theorem 6.4 provides an alternative proof of existence of weak solutions via a mixed DG-FEM based on Bernardi–Raugel finite element for the problem (1.1) under the hypothesis \(\gamma >1\). (ii) In the case \(\gamma >\frac{4}{3}\), the CFL condition \(\tau \approx h\) is not required for Theorem 6.4. It is worth noting that the values of adiabatic exponent \(\gamma \) in the convergence result without the CFL condition includes the real fluid range of \(\gamma \in [\frac{4}{3},\frac{5}{3}]\), such as the monoatomic gas (\(\gamma \sim \frac{5}{3}\)) and the diatomic gas (\(\gamma \sim \frac{7}{5}\)). (iii) Theorem 6.4 is also true with the external force \(\varvec{f}\ne \varvec{0}\in \varvec{L}^2((0,T)\times \varOmega )\) in the momentum equation.

7 Error Estimate

An unconditional error estimate for the semi-stationary compressible Stokes equations is established in the section. Note that the existence of weak solution to this model under the assumption of \(\gamma >1\) is proved by Theorem 6.4. Now we report the weak-strong uniqueness for this model. To this end, we introduce the following functional \({\mathbb {E}}:[0,\infty )\times (0,\infty )\rightarrow {\mathbb {R}}\), which is given by

$$\begin{aligned} {\mathbb {E}}(\rho \,|\,{\underline{\rho }}):={\mathcal {H}}(\rho )-{\mathcal {H}}'({\underline{\rho }})(\rho -{\underline{\rho }})-{\mathcal {H}}({\underline{\rho }}). \end{aligned}$$

Noticing that the function \({\mathcal {H}}\) is strictly convex in \((0,\infty )\), we obtain

$$\begin{aligned} {\mathbb {E}}(\rho \,|\,{\underline{\rho }})\ge 0,\quad {\text {and}}\quad {\mathbb {E}}(\rho \,|\,{\underline{\rho }})=0\Leftrightarrow \rho ={\underline{\rho }}. \end{aligned}$$
(7.1)

Furthermore, the functional \({\mathbb {E}}(\rho \,|\,{\underline{\rho }})\) satisfies the following estimates (see, e.g., [13, 17] for more details)

$$\begin{aligned} {\mathbb {E}}(\rho \,|\,{\underline{\rho }})\ge C({\underline{\rho }})\left\{ \begin{aligned}&(\rho -{\underline{\rho }})^2,\; \frac{1}{2}{\underline{\rho }}<\rho <2{\underline{\rho }},\\&(1+\rho ^\gamma ),\;{\text {otherwise}}, \end{aligned} \right. , \end{aligned}$$
(7.2)

where \(C({\underline{\rho }})\) is uniformly bounded if \({\underline{\rho }}\) lies in some compact subset of \((0,\infty )\). Finally, we introduce the relative energy functional of the problem (1.1), which is defined by

$$\begin{aligned} {\mathcal {E}}(\rho \,|\,{\underline{\rho }}):=\int _{\varOmega }{{\mathbb {E}}(\rho \,|\,{\underline{\rho }})}dx. \end{aligned}$$

Theorem 7.1

Let \(\varOmega \subset {\mathbb {R}}^3\) be a bounded domain and assume that the viscosity coefficient \(\mu ,\lambda \) satisfies \(\mu >0\) and \(3\lambda +2\mu >0\). Suppose that the pressure \(p=p(\rho )\) satisfies the assumption (1.2) with \(\gamma >1\). Let \((\rho ,\varvec{u})\) be a weak solution to the problem (1.1) emanating from the initial data \((\rho _0,\varvec{u}_0)\) with the finite energy \(E_0:=\int _{\varOmega }{{\mathcal {H}}(\rho _0)}dx\) and finite mass \(M_0:=\int _{\varOmega }{\rho _0}dx\). Let \(({\underline{\rho }},\underline{\varvec{u}})\) be a strong solution of the same problem belonging to the class

$$\begin{aligned} \left\{ \begin{aligned}&\nabla {\underline{\rho }}\in L^2(0,T;L^q(\varOmega )),\quad 0<{\underline{\rho }}_{\min }\le {\underline{\rho }}\le {\underline{\rho }}_{\max },\\&\nabla ^2\underline{\varvec{u}}\in \varvec{L}^2(0,T;\varvec{L}^q(\varOmega )),\quad \underline{\varvec{u}}\in L^2(0,T;\varvec{H}_0^1(\varOmega )),\quad q>\max \{3,\frac{6\gamma }{5\gamma -6}\}, \end{aligned} \right. \end{aligned}$$

emanating from the same initial data. Then

$$\begin{aligned} \rho ={\underline{\rho }},\quad \varvec{u}=\underline{\varvec{u}},\;{\text {in}}\;(0,T)\times \varOmega . \end{aligned}$$

The proof of Theorem 7.1 can been found in [13, Theorem 4.1].

Next, we deduce the discrete version of the relative energy inequality from the scheme (3.7)–(3.8), which will play a key role in the subsequent error estimate. To this end, we first introduce the convenient notations

$$\begin{aligned} \left\{ \begin{aligned} {\underline{\rho }}_h^n:=&\varPi _h^{{\mathbb {Q}}}[{\underline{\rho }}^n],\quad {\underline{\rho }}^n:={\underline{\rho }}(t_n,\cdot ),\quad \forall \; 1\le n\le N,\\ \underline{\varvec{u}}_h^n:=&\varPi _h^{{\mathbb {V}}}[\underline{\varvec{u}}^n],\quad \underline{\varvec{u}}^n:=\underline{\varvec{u}}(t_n,\cdot ),\quad \forall \; 1\le n\le N, \end{aligned} \right. \end{aligned}$$

where \(({\underline{\rho }},\underline{\varvec{u}})\) is a strong solution of the problem (1.1) belonging to the class of \(C^2\) functions such that \(\underline{\varvec{u}}|_{(0,T)\times \partial \varOmega }=\varvec{0}\) and \(0<{\underline{\rho }}_{\min }\le {\underline{\rho }}\le {\underline{\rho }}_{\max }\). Furthermore, we define the piecewise constant temporal interpolations of \(({\underline{\rho }}_h^n,\underline{\varvec{u}}_h^n,{\underline{\rho }}^n,\underline{\varvec{u}}^n)\), \(1\le n\le N\), i.e., for any \(t\in [t_{n-1},t_n]\)

$$\begin{aligned} \left\{ \begin{aligned} {\underline{\rho }}_h(t,\cdot ):={\underline{\rho }}_h^n,\quad {\underline{\rho }}_h^{\star }(t,\cdot ):={\underline{\rho }}_h^{n-1},\quad \underline{\varvec{u}}_h(t,\cdot ):=\underline{\varvec{u}}_h^n,\\ {\underline{\rho }}_{\tau }(t,\cdot ):={\underline{\rho }}^n,\quad {\underline{\rho }}_{\tau }^{\star }(t,\cdot ):={\underline{\rho }}^{n-1},\quad \underline{\varvec{u}}_{\tau }(t,\cdot ):=\underline{\varvec{u}}^n. \end{aligned} \right. \end{aligned}$$
(7.3)

Theorem 7.2

Suppose that Hypothesis 2.1 and \(\gamma >1\) are satisfied. The family \((\rho _h,\varvec{u}_h)\) and \(({\underline{\rho }}_{\tau },\underline{\varvec{u}}_{\tau })\) are defined in (3.16)–(3.17) and (7.3), respectively. Then there exists a constant \(C>0\) independent of h and \(\tau \) such that

$$\begin{aligned} \Vert {\mathcal {E}}(\rho _h \,|\,{\underline{\rho }}_{\tau })\Vert _{L^\infty (0,T)}\le C. \end{aligned}$$
(7.4)

Proof

Using the identity \({\mathcal {H}}'(\rho )\rho -{\mathcal {H}}(\rho )=p(\rho )\) and the Hölder inequality, we can show

$$\begin{aligned} {\mathcal {E}}(\rho _h^n \,|\,{\underline{\rho }}^n)\le&C\Vert \rho _h\Vert _{L^\infty (0,T;L^\gamma (\varOmega ))}^\gamma +C\Vert {\underline{\rho }}\Vert _{L^\infty (0,T;L^\gamma (\varOmega ))}^\gamma \\&+C\Vert {\underline{\rho }}\Vert _{L^\infty (0,T;L^{2(\gamma -1)}(\varOmega ))}^{\gamma -1}\Vert \rho _h\Vert _{L^\infty (0,T;L^\gamma (\varOmega ))},\quad \forall \; 1\le n\le N. \end{aligned}$$

By employing Lemma  3.6, we have the estimate (7.4). This proof is thus complete. \(\square \)

Now we establish the discrete version of the relative energy inequality.

Theorem 7.3

Let the families \((\rho _h,\varvec{u}_h)\) and \(({\underline{\rho }}_{\tau },\underline{\varvec{u}}_{\tau })\) be defined as in (3.16)–(3.17) and (7.3), respectively. Then the discrete relative energy inequality holds, for any \(1\le m\le N\),

$$\begin{aligned}&{\mathcal {E}}(\rho _h^m \,|\,{\underline{\rho }}^m)+\mu \int _0^{t_m}{\int _{\varOmega }{|\nabla (\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2}dx}dt\nonumber \\&+(\lambda +\mu )\int _0^{t_m}{\int _{\varOmega }{|{\text {div}}(\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2}dx}dt\le {\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+\sum _{i=1}^6{\mathcal {R}}_i, \end{aligned}$$
(7.5)

where the remainder terms \({\mathcal {R}}_i\) (\(1\le i\le 5\)) are defined by

$$\begin{aligned} {\mathcal {R}}_1:=&\int _0^{t_m}{\int _{\varOmega }{[\mu \nabla \underline{\varvec{u}}_{\tau }:\nabla (\underline{\varvec{u}}_{\tau }-\varvec{u}_h)+(\lambda +\mu ){\text {div}}\underline{\varvec{u}}_{\tau }{\text {div}}(\underline{\varvec{u}}_{\tau }-\varvec{u}_h)]}dx}dt,\\ {\mathcal {R}}_2:=&\int _0^{t_m}{\int _{\varOmega }{[\mu \nabla \varvec{u}_h:\nabla (\underline{\varvec{u}}_h-\underline{\varvec{u}}_{\tau })+(\lambda +\mu ){\text {div}}\underline{\varvec{u}}_{\tau }{\text {div}}(\underline{\varvec{u}}_h-\underline{\varvec{u}}_{\tau })]}dx}dt,\\ {\mathcal {R}}_3:=&-\int _0^{t_m}{\int _{\varOmega }{p(\rho _h){\text {div}}\underline{\varvec{u}}_{\tau }}dx}dt,\quad {\mathcal {R}}_4:=\int _0^{t_m}{\int _{\varOmega }{({\underline{\rho }}_{\tau }-\rho _h)D_t{\mathcal {H}}'({\underline{\rho }}_{\tau })}dx}dt,\\ {\mathcal {R}}_5:=&-\int _0^{t_m}{\int _{\varOmega }{\rho _h\varvec{u}_h\cdot \nabla {\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })}dx}dt,\quad {\mathcal {R}}_6:=-\int _0^{t_m}{\int _{\varOmega }{{\mathscr {R}}_h\cdot \nabla {\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })}dx}dt. \end{aligned}$$

Proof

First, taking \(\varvec{v}_h=\underline{\varvec{u}}_h^n\) in the discrete momentum equation (3.8), and summing this result with respect to n from \(n=1\) to \(n=m\), we conclude that

$$\begin{aligned}&\int _0^{t_m}{\int _{\varOmega }{[\mu \nabla \varvec{u}_h:\nabla \underline{\varvec{u}}_{\tau }+(\lambda +\mu ){\text {div}}\varvec{u}_h{\text {div}}\underline{\varvec{u}}_{\tau }-p(\rho _h){\text {div}}\underline{\varvec{u}}_{\tau }]}dx}dt\\&+\int _0^{t_m}{\int _{\varOmega }{[\mu \nabla \varvec{u}_h:\nabla (\underline{\varvec{u}}_h-\underline{\varvec{u}}_{\tau })+(\lambda +\mu ){\text {div}}\varvec{u}_h{\text {div}}(\underline{\varvec{u}}_h-\underline{\varvec{u}}_{\tau })]}dx}dt=0.\nonumber \end{aligned}$$
(7.6)

Next, using the same argument as Lemma 4.1 by taking \(\varphi ={\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })\) in Lemma 4.1, we obtain

$$\begin{aligned} -\int _0^{t_m}{\int _{\varOmega }{D_t\rho _h{\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })}dx}dt=&-\int _0^{t_m}{\int _{\varOmega }{\rho _h\varvec{u}_h\cdot \nabla {\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })}dx}dt\nonumber \\&-\int _0^{t_m}{\int _{\varOmega }{{\mathscr {R}}_h\cdot \nabla {\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })}dx}dt. \end{aligned}$$
(7.7)

Note that the numerical diffusion terms \({\mathcal {D}}_{i,h}^n\) (\(1\le i\le 3\)) in the discrete energy identity (3.10) are all positive, we have

$$\begin{aligned} \int _0^{t_m}{\int _{\varOmega }{D_t{\mathcal {H}}(\rho _h)}dx}dt+\int _0^{t_m}{\int _{\varOmega }{[\mu |\nabla \varvec{u}_h|^2+(\lambda +\mu )|{\text {div}}\varvec{u}_h|^2]}dx}dt\le 0. \end{aligned}$$
(7.8)

By applying the identity

$$\begin{aligned} \rho _h{\mathcal {H}}'({\underline{\rho }}_{\tau })-\rho _h^{\star }{\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })=\rho _h({\mathcal {H}}'({\underline{\rho }}_{\tau })-{\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star }))+(\rho _h-\rho _h^{\star }){\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star }), \end{aligned}$$

we rewrite

$$\begin{aligned} \int _0^{t_m}{\int _{\varOmega }{D_t\rho _h{\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })}dx}dt=\int _0^{t_m}{\int _{\varOmega }{D_t(\rho _h{\mathcal {H}}'({\underline{\rho }}_{\tau }))}dx}dt-\int _0^{t_m}{\int _{\varOmega }{\rho _hD_t{\mathcal {H}}'({\underline{\rho }}_{\tau })}dx}dt, \end{aligned}$$

which implies that

$$\begin{aligned}&\int _0^{t_m}{\int _{\varOmega }{D_t{\mathcal {H}}(\rho _h)}dx}dt-\int _0^{t_m}{\int _{\varOmega }{D_t\rho _h{\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })}dx}dt=\int _0^{t_m}{\int _{\varOmega }{D_t{\mathbb {E}}(\rho _h \,|\,{\underline{\rho }}_{\tau })}dx}dt\nonumber \\&+\int _0^{t_m}{\int _{\varOmega }{\rho _hD_t{\mathcal {H}}'({\underline{\rho }}_{\tau })}dx}dt-\int _0^{t_m}{\int _{\varOmega }{D_t({\underline{\rho }}_{\tau }{\mathcal {H}}'({\underline{\rho }}_{\tau })-{\mathcal {H}}({\underline{\rho }}_{\tau }))}dx}dt. \end{aligned}$$
(7.9)

According to the convexity of the function \({\mathcal {H}}\), we obtain

$$\begin{aligned} {\mathcal {H}}({\underline{\rho }}_{\tau })-{\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })({\underline{\rho }}_{\tau }-{\underline{\rho }}_{\tau }^{\star })-{\mathcal {H}}({\underline{\rho }}_{\tau }^{\star })\ge 0. \end{aligned}$$
(7.10)

By using the inequality (7.10), we have

$$\begin{aligned}&\int _0^{t_m}{\int _{\varOmega }{D_t({\underline{\rho }}_{\tau }{\mathcal {H}}'({\underline{\rho }}_{\tau })-{\mathcal {H}}({\underline{\rho }}_{\tau }))}dx}dt=\int _0^{t_m}{\int _{\varOmega }{{\underline{\rho }}_{\tau }D_t{\mathcal {H}}'({\underline{\rho }}_{\tau })}dx}dt\\&-\frac{1}{\tau }\int _0^{t_m}{\int _{\varOmega }{{\mathcal {H}}({\underline{\rho }}_{\tau })-{\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })({\underline{\rho }}_{\tau }-{\underline{\rho }}_{\tau }^{\star })-{\mathcal {H}}({\underline{\rho }}_{\tau }^{\star })}dx}dt\le \int _0^{t_m}{\int _{\varOmega }{{\underline{\rho }}_{\tau }D_t{\mathcal {H}}'({\underline{\rho }}_{\tau })}dx}dt.\nonumber \end{aligned}$$
(7.11)

Combining the inequalities (7.6)–(7.9) and (7.11), we obtain the inequality (7.5). This proof is thus complete. \(\square \)

In the next step, we deduce the approximate version of the relative energy inequality from the estimate (7.5).

Theorem 7.4

Suppose that Hypothesis 2.1 is satisfied and the pressure \(p=p(\rho )\) satisfies the hypothesis (1.2) with \(\gamma >1\). Let the internal energy \({\mathcal {H}}\) be given by \({\mathcal {H}}(\rho )=\frac{p(\rho )}{\gamma -1}\). Let the families \((\rho _h,\varvec{u}_h)\) and \(({\underline{\rho }}_{\tau },\underline{\varvec{u}}_{\tau })\) be defined as in (3.16)–(3.17) and (7.3), respectively. Then there exists

$$\begin{aligned} C:=&C(T,\varOmega ,M_0,E_0,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\\&\Vert (\partial _t{\underline{\rho }},\partial _{tt}{\underline{\rho }},\nabla {\underline{\rho }},\partial _t\nabla {\underline{\rho }})\Vert _{L^\infty ((0,T)\times \varOmega )},\Vert \underline{\varvec{u}}\Vert _{L^\infty (0,T;\varvec{H}^2(\varOmega ))})>0, \end{aligned}$$

such that for any \(1\le m\le N\), we have the approximate relative energy inequality holds,

$$\begin{aligned}&{\mathcal {E}}(\rho _h^m \,|\,{\underline{\rho }}^m)+\int _0^{t_m}{\int _{\varOmega }{[\mu |\nabla (\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2+(\lambda +\mu )|{\text {div}}(\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2]}dx}dt\nonumber \\ \le&{\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+{\mathcal {R}}_1+{\mathcal {R}}_3+\sum _{i=1}^3{\mathcal {L}}_3, \end{aligned}$$
(7.12)

where the remainder terms \({\mathcal {L}}_i\) (\(1\le i\le 3\)) are defined by

$$\begin{aligned} {\mathcal {L}}_1&:=\int _0^{t_m}{\int _{\varOmega }{({\underline{\rho }}_{\tau }-\rho _h)\frac{p'({\underline{\rho }}_{\tau })}{{\underline{\rho }}_{\tau }}[\partial _t{\underline{\rho }}]_{\tau }}dx}dt,\quad {\mathcal {L}}_2:=-\int _0^{t_m}{\int _{\varOmega }{\frac{\rho _h}{{\underline{\rho }}_{\tau }}p'({\underline{\rho }}_{\tau })\varvec{u}_h\cdot \nabla {\underline{\rho }}_\tau }dx}dt,\\ |{\mathcal {L}}_3|&\le C(h^A+\tau ),\quad A:=\frac{\min \{\epsilon ,1\}}{2},\quad [\partial _t{\underline{\rho }}]_{\tau }:=\partial _t{\underline{\rho }}(t_n,\cdot ),\;{\text {for}}\; [t_{n-1},t_n]. \end{aligned}$$

Proof

We start the proof from the discrete version of the relative energy inequality (7.5) derived in the previous Theorem 7.3. The terms \({\mathcal {R}}_i\) (\(i=2,4,5,6\)) will be transformed to a more convenient form, and the other terms \({\mathcal {R}}_i\) (\(i=1,3\)) will remain unchanged.

  • The term \({\mathcal {R}}_2\). By applying the Cauchy–Schwarz inequality and the estimate (3.2), we can show

    $$\begin{aligned} |{\mathcal {R}}_2|\le&C\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\Vert \underline{\varvec{u}}_h-\underline{\varvec{u}}_{\tau }\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\\ \le&C(E_0,\Vert \underline{\varvec{u}}\Vert _{L^\infty (0,T;\varvec{H}^2(\varOmega ))})h. \end{aligned}$$
  • The term \({\mathcal {R}}_4\). Firstly, by applying the Taylor formula, we have

    $$\begin{aligned} {\mathcal {H}}'({\underline{\rho }}^n)-{\mathcal {H}}'({\underline{\rho }}^{n-1})={\mathcal {H}}''({\underline{\rho }}^n)({\underline{\rho }}^n-{\underline{\rho }}^{n-1})-\frac{1}{2}{\mathcal {H}}'''(\xi _{{\underline{\rho }}}^n)({\underline{\rho }}^n-{\underline{\rho }}^{n-1})^2, \end{aligned}$$
    (7.13)

    where \(\xi _{{\underline{\rho }}}^n\in {\text {co}}\{{\underline{\rho }}^n,{\underline{\rho }}^{n-1}\}\). Let \(\xi _{{\underline{\rho }}}(t,\cdot ):=\xi _{{\underline{\rho }}}^n\) for \(t\in [t_{n-1},t_n]\). By applying the identity (7.13), the term \({\mathcal {R}}_4\) can be rewritten as

    $$\begin{aligned} {\mathcal {R}}_4={\mathcal {L}}_1+{\mathcal {L}}_{3,1}+{\mathcal {L}}_{3,2}, \end{aligned}$$

    where the remainder terms \({\mathcal {L}}_{3,i}\) are given by

    $$\begin{aligned} {\mathcal {L}}_{3,1}:=&\int _0^{t_m}{\int _{\varOmega }{({\underline{\rho }}_{\tau }-\rho _h)\frac{p'({\underline{\rho }}_{\tau })}{{\underline{\rho }}_{\tau }}(D_t{\underline{\rho }}_{\tau }-[\partial _t{\underline{\rho }}]_{\tau })}dx}dt,\\ {\mathcal {L}}_{3,2}:=&\frac{1}{2\tau }\int _0^{t_m}{\int _{\varOmega }{(\rho _h-{\underline{\rho }}_{\tau }){\mathcal {H}}'''(\xi _{{\underline{\rho }}})({\underline{\rho }}_{\tau }-{\underline{\rho }}_{\tau }^{\star })^2}dx}dt. \end{aligned}$$

    Using the Taylor formula and the mass conservation (see, Remark 3.1), we obtain

    $$\begin{aligned} |{\mathcal {L}}_{3,1}|\le&\tau C({\underline{\rho }}_{\min },{\underline{\rho }}_{\max })|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])}\Vert {\underline{\rho }}_{\tau }-\rho _h\Vert _{L^1((0,T)\times \varOmega )}\Vert \partial _{tt}{\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )}\\ \le&\tau C(M_0,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \partial _{tt}{\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )}). \end{aligned}$$

    By a similar argument, we conclude that

    $$\begin{aligned} |{\mathcal {L}}_{3,2}|\le \tau C(M_0,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \partial _t{\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )}). \end{aligned}$$
  • The term \({\mathcal {R}}_5\). We may write

    $$\begin{aligned} {\mathcal {R}}_5=&-\int _0^{t_m}{\int _{\varOmega }{\rho _h\varvec{u}_h\cdot ({\mathcal {H}}''({\underline{\rho }}_{\tau }^{\star })\nabla {\underline{\rho }}_{\tau }^{\star }-{\mathcal {H}}''({\underline{\rho }}_{\tau })\nabla {\underline{\rho }}_{\tau })}dx}dt\\&-\int _0^{t_m}{\int _{\varOmega }{\rho _h\varvec{u}_h\cdot {\mathcal {H}}''({\underline{\rho }}_{\tau })\nabla {\underline{\rho }}_{\tau }}dx}dt={\mathcal {L}}_{3,3}+{\mathcal {L}}_2. \end{aligned}$$

    By applying the first-order Taylor formula, we obtain

    $$\begin{aligned}&\Vert p'({\underline{\rho }}_{\tau }^\star )-p'({\underline{\rho }}_{\tau })\Vert _{L^\infty ((0,T)\times \varOmega )}\nonumber \\&\quad \le C(|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \partial _t{\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )})\tau . \end{aligned}$$
    (7.14)

    Using the estimate (7.14) and the Taylor formula, we have

    $$\begin{aligned}&\Vert {\mathcal {H}}''({\underline{\rho }}_{\tau }^{\star })-{\mathcal {H}}''({\underline{\rho }}_{\tau })\Vert _{L^\infty ((0,T)\times \varOmega )}\nonumber \\&\quad \le C({\underline{\rho }}_{\min },{\underline{\rho }}_{\max })\Vert p'({\underline{\rho }}_{\tau }^\star )-p'({\underline{\rho }}_{\tau })\Vert _{L^\infty ((0,T)\times \varOmega )}\nonumber \\&\qquad +C({\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])})\Vert {\underline{\rho }}_{\tau }^\star -{\underline{\rho }}_{\tau }\Vert _{L^\infty ((0,T)\times \varOmega )}\nonumber \\&\quad \le C({\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \partial _t{\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )})\tau . \end{aligned}$$
    (7.15)

    Therefore, by using the Hölder inequality and the estimate (7.15), we obtain

    $$\begin{aligned} |{\mathcal {L}}_{3,3}|\le&C(\varOmega ,T)\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\\&\times \Vert ({\mathcal {H}}''({\underline{\rho }}_{\tau }^{\star })-{\mathcal {H}}''({\underline{\rho }}_{\tau }))\nabla {\underline{\rho }}_{\tau }^{\star }\Vert _{L^\infty ((0,T)\times \varOmega )}\\ \le&C(\varOmega ,T)\Vert \rho _h\Vert _{L^{2\gamma }((0,T)\times \varOmega )}\Vert \varvec{u}_h\Vert _{L^2(0,T;\varvec{H}^1(\varOmega ))}\\&\times \Vert {\mathcal {H}}''({\underline{\rho }}_{\tau })\nabla ({\underline{\rho }}_{\tau }^{\star }-{\underline{\rho }}_{\tau })\Vert _{L^\infty ((0,T)\times \varOmega )}\\ \le&C(\varOmega ,T,E_0,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert (\partial _t{\underline{\rho }},\nabla {\underline{\rho }},\partial _t\nabla {\underline{\rho }})\Vert _{L^\infty ((0,T)\times \varOmega )})\tau . \end{aligned}$$
  • The term \({\mathcal {R}}_6\). By applying the estimate (4.3) of Lemma 4.2, we get that

    $$\begin{aligned} |{\mathcal {R}}_6|\le&C(\varOmega ,T)\Vert \nabla {\mathcal {H}}'({\underline{\rho }}_{\tau }^{\star })\Vert _{L^\infty ((0,T)\times \varOmega )}h^A\\ \le&C(\varOmega ,T,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \nabla {\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )})h^A. \end{aligned}$$

Combining the above analysis, we have \(\sum _{i=1}^6{\mathcal {R}}_i={\mathcal {R}}_1+{\mathcal {R}}_3+\sum _{i=1}^3{\mathcal {L}}_i\). After setting \({\mathcal {L}}_3:=\sum _{i=1}^3{\mathcal {L}}_{3,i}+{\mathcal {R}}_2+{\mathcal {R}}_6\), we deduce the approximate relative energy inequality (7.12) from the estimate (7.5). The proof is thus complete. \(\square \)

Next we derive a discrete identity for the strong solutions.

Theorem 7.5

Suppose that Hypothesis 2.1 is satisfied and the pressure \(p=p(\rho )\) satisfies the hypothesis (1.2) with \(\gamma >1\). Let the internal energy \({\mathcal {H}}\) is given by \({\mathcal {H}}(\rho )=\frac{p(\rho )}{\gamma -1}\). Let the family \(({\underline{\rho }}_{\tau },\underline{\varvec{u}}_{\tau })\) be defined as in (7.3). For any \(1\le m\le N\), then the following identity holds:

$$\begin{aligned} {\mathcal {R}}_1+\sum _{i=4}^5{\mathcal {L}}_i=0, \end{aligned}$$
(7.16)

where the remainder terms \({\mathcal {L}}_i\) are defined by

$$\begin{aligned} {\mathcal {L}}_4:=&-\int _0^{t_m}{\int _{\varOmega }{p'({\underline{\rho }}_{\tau })\varvec{u}_h\cdot \nabla {\underline{\rho }}_\tau }dx}dt,\\ {\mathcal {L}}_5:=&-\int _0^{t_m}{\int _{\varOmega }{p({\underline{\rho }}_{\tau }){\text {div}}\underline{\varvec{u}}_\tau }dx}dt. \end{aligned}$$

Proof

Since \(({\underline{\rho }},\underline{\varvec{u}})\) is a strong solution of the problem (1.1), the second equation of (1.1) can be rewritten in the form

$$\begin{aligned} \mu \varDelta \underline{\varvec{u}}+(\lambda +\mu )\nabla {\text {div}}\underline{\varvec{u}}=\nabla p({\underline{\rho }}). \end{aligned}$$
(7.17)

Taking \(t=t_n\) in (7.17), multiplying this identity by \(\varvec{u}_h^n-\underline{\varvec{u}}^n\) and integration over \(\varOmega \). We get, after summation from \(n=1\) to \(n=m\),

$$\begin{aligned} 0=&-\int _0^{t_m}{\int _{\varOmega }{(\mu \varDelta \underline{\varvec{u}}_{\tau }+(\lambda +\mu )\nabla {\text {div}}\underline{\varvec{u}}_{\tau })\cdot (\varvec{u}_h-\underline{\varvec{u}}_{\tau })}dx}dt\\&+\int _0^{t_m}{\int _{\varOmega }{\nabla p({\underline{\rho }}_{\tau })\cdot \varvec{u}_h}dx}dt-\int _0^{t_m}{\int _{\varOmega }{\nabla p({\underline{\rho }}_{\tau })\cdot \underline{\varvec{u}}_{\tau }}dx}dt. \end{aligned}$$

which implies that \({\mathcal {R}}_1+\sum _{i=4}^5{\mathcal {L}}_i=0\). The proof is thus complete. \(\square \)

Now, we will derive the unconditional error estimate of the problem (1.1) based on the approximate relative energy inequality (7.13) and the discrete identity (7.16).

Theorem 7.6

Let \(\varOmega \subset {\mathbb {R}}^d\), \(d=2,3\) be a bounded domain and assume that the viscosity coefficient \(\mu ,\lambda \) satisfies \(\mu >0\) and \(d\lambda +2\mu >0\). Suppose that the pressure \(p=p(\rho )\) satisfies the assumption (1.2) with \(\gamma >\frac{6}{5}\). The initial values \((\rho _0,\varvec{u}_0)\) satisfies Hypothesis 2.1 with the finite energy \(E_0:=\int _{\varOmega }{{\mathcal {H}}(\rho _0)}dx\) and finite mass \(M_0:=\int _{\varOmega }{\rho _0}dx\). Let \(({\underline{\rho }},\underline{\varvec{u}})\) be a strong solution of the problem (1.1) belonging to the class

$$\begin{aligned} \left\{ \begin{aligned}&{\underline{\rho }}\in C^2([0,T]\times \varOmega ),\quad 0<{\underline{\rho }}_{\min }\le {\underline{\rho }}\le {\underline{\rho }}_{\max },\\&\underline{\varvec{u}}\in \varvec{C}^2([0,T]\times \varOmega ),\quad \underline{\varvec{u}}|_{(0,T)\times \partial \varOmega }=\varvec{0}, \end{aligned} \right. \end{aligned}$$

emanating from the initial data \(({\underline{\rho }}_0,\underline{\varvec{u}}_0)\). Let the families \((\rho _h,\varvec{u}_h)\) and \(({\underline{\rho }}_{\tau },\underline{\varvec{u}}_{\tau })\) be defined as in (3.16)–(3.17) and (7.3), respectively. Then there exists

$$\begin{aligned} C:=&C(T,\varOmega ,M_0,E_0,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\\&\Vert (\partial _t{\underline{\rho }},\partial _{tt}{\underline{\rho }},\nabla {\underline{\rho }},\partial _t\nabla {\underline{\rho }},\nabla ^2\underline{\varvec{u}})\Vert _{L^\infty ((0,T)\times \varOmega )})>0, \end{aligned}$$

such that for any \(1\le m\le N\), then we have

$$\begin{aligned}&{\mathcal {E}}(\rho _h^m \,|\,{\underline{\rho }}^m)+\int _0^{t_m}{\int _{\varOmega }{[\mu |\nabla (\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2+(\lambda +\mu )|{\text {div}}(\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2]}dx}dt\nonumber \\&\quad \le C({\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+h^A+\tau ),\quad A:=\frac{\min \{\epsilon ,1\}}{2}. \end{aligned}$$
(7.18)

Proof

Combining the approximate relative energy inequality (7.12) and the discrete identity (7.16), we can show

$$\begin{aligned}&{\mathcal {E}}(\rho _h^m \,|\,{\underline{\rho }}^m)+\int _0^{t_m}{\int _{\varOmega }{[\mu |\nabla (\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2+(\lambda +\mu )|{\text {div}}(\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2]}dx}dt\\&\quad \le {\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+\sum _{i=6}^7{\mathcal {L}}_i. \end{aligned}$$

where the terms \({\mathcal {L}}_i\) are defined by

$$\begin{aligned} {\mathcal {L}}_6:=&\int _0^{t_m}{\int _{\varOmega }{\frac{{\underline{\rho }}_{\tau }-\rho _h}{{\underline{\rho }}_{\tau }}p'({\underline{\rho }}_{\tau })\varvec{u}_h\cdot \nabla {\underline{\rho }}_{\tau }}dx}dt+\int _0^{t_m}{\int _{\varOmega }{({\underline{\rho }}_{\tau }-\rho _h)\frac{p'({\underline{\rho }}_{\tau })}{{\underline{\rho }}_{\tau }}[\partial _t{\underline{\rho }}]_{\tau }}dx}dt,\\&+\int _0^{t_m}{\int _{\varOmega }{(p({\underline{\rho }}_{\tau })-p(\rho _h)){\text {div}}\varvec{u}_\tau }dx}dt,\quad |{\mathcal {L}}_7|\le C(h^A+\tau ). \end{aligned}$$

We next bound the term \({\mathcal {L}}_6\). Since the pair \(({\underline{\rho }},\underline{\varvec{u}})\) is a strong solution of the problem (1.1), the first equation of (1.1) can be rewritten in the form

$$\begin{aligned}{}[\partial _t{\underline{\rho }}]_\tau =-\underline{\varvec{u}}_\tau \cdot \nabla {\underline{\rho }}_\tau -{\underline{\rho }}_\tau {\text {div}}\underline{\varvec{u}}_\tau . \end{aligned}$$
(7.19)

By the identity (7.19), we write

$$\begin{aligned} {\mathcal {L}}_6={\mathcal {L}}_{6,1}+{\mathcal {L}}_{6,2}, \end{aligned}$$

where

$$\begin{aligned} {\mathcal {L}}_{6,1}:=&-\int _0^{t_m}{\int _{\varOmega }{(p(\rho _h)-p'({\underline{\rho }}_{\tau })(\rho _h-{\underline{\rho }}_{\tau })-p({\underline{\rho }}_{\tau })){\text {div}}\varvec{u}_\tau }dx}dt,\\ {\mathcal {L}}_{6,2}:=&\int _0^{t_m}{\int _{\varOmega }{\frac{{\underline{\rho }}_{\tau }-\rho _h}{{\underline{\rho }}_{\tau }}p'({\underline{\rho }}_{\tau })(\varvec{u}_h-\underline{\varvec{u}}_\tau )\cdot \nabla {\underline{\rho }}_{\tau }}dx}dt. \end{aligned}$$

It is easy to check that

$$\begin{aligned} |{\mathcal {L}}_{6,1}|\le C(T,\varOmega ,\Vert {\text {div}}\underline{\varvec{u}}\Vert _{L^\infty ((0,T)\times \varOmega )})\int _0^{t_m}{{\mathcal {E}}(\rho _h\, |\,{\underline{\rho }}_{\tau })}dt. \end{aligned}$$

Let \(\varOmega _{h,1}:=\{\frac{{\underline{\rho }}_{\tau }}{2}<\rho _h<2{\underline{\rho }}_{\tau }\}\) and \(\varOmega _{h,2}:=\varOmega \setminus \varOmega _{h,1}\). The term \({\mathcal {L}}_{6,2}\) can be rewritten as

$$\begin{aligned} {\mathcal {L}}_{6,2}:={\mathcal {L}}_{6,2,\varOmega _{h,1}}+{\mathcal {L}}_{6,2,\varOmega _{h,2}}, \end{aligned}$$

where

$$\begin{aligned} {\mathcal {L}}_{6,2,\varOmega _{h,i}}:=\int _0^{t_m}{\int _{\varOmega _{h,i}}{\frac{{\underline{\rho }}_{\tau }-\rho _h}{{\underline{\rho }}_{\tau }}p'({\underline{\rho }}_{\tau })(\varvec{u}_h-\underline{\varvec{u}}_\tau )\cdot \nabla {\underline{\rho }}_{\tau }}dx}dt,\quad i=1,2. \end{aligned}$$

By applying the Poincaré and Young inequalities, the estimate (7.2), \({\underline{\rho }}_{\tau }\in ({\underline{\rho }}_{\min },{\underline{\rho }}_{\max })\), we can show

$$\begin{aligned} |{\mathcal {L}}_{6,2,\varOmega _{h,1}}|\le&C(\delta ,\varOmega ,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \nabla {\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )})\int _0^{t_m}{{\mathcal {E}}(\rho _h \,|\, {\underline{\rho }}_{\tau })}dt\\&+\delta \int _0^{t_m}{\int _{\varOmega }{|\nabla (\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2}dx}dt. \end{aligned}$$

By employing the estimate (7.2) and \({\underline{\rho }}_{\tau }\in ({\underline{\rho }}_{\min },{\underline{\rho }}_{\max })\), \(\gamma >\frac{6}{5}\), we have

$$\begin{aligned} {\mathbb {E}}(\rho _h \,|\, {\underline{\rho }}_{\tau })\ge C({\underline{\rho }}_{\min },{\underline{\rho }}_{\max },\gamma )(1+\rho _h^\gamma )\ge |\rho _h-{\underline{\rho }}_{\tau }|^{\frac{6}{5}},\;{\text {in}}\;\varOmega _{h,2}. \end{aligned}$$
(7.20)

Using the Poincaré and Young inequalities, the estimates (7.20) and (7.4), \(\underline{\rho }_\tau \in (\underline{\rho }_{\min },\underline{\rho }_{\max })\), we conclude that

$$\begin{aligned} |{\mathcal {L}}_{6,2,\varOmega _{h,2}}|\le&C(\delta ,\varOmega ,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \nabla {\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )})\int _0^{t_m}{{\mathcal {E}}(\rho _h \,|\, {\underline{\rho }}_{\tau })}dt\\&+\delta \int _0^{t_m}{\int _{\varOmega }{|\nabla (\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2}dx}dt. \end{aligned}$$

Combining the above analysis with \(\delta =\frac{\mu }{4}\), we get that

$$\begin{aligned}&{\mathcal {E}}(\rho _h^m \,|\,{\underline{\rho }}^m)+\int _0^{t_m}{\int _{\varOmega }{[\mu |\nabla (\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2+(\lambda +\mu )|{\text {div}}(\varvec{u}_h-\underline{\varvec{u}}_{\tau })|^2]}dx}dt\\&\quad \le C({\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+h^A+\tau )+C\int _0^{t_m}{{\mathcal {E}}(\rho _h \,|\, {\underline{\rho }}_{\tau })}dt, \end{aligned}$$

where the constant \(C>0\) is given by

$$\begin{aligned} C:=&C(T,\varOmega ,M_0,E_0,\underline{\rho }_{\min },\underline{\rho }_{\max },|p'|_{C^1([\underline{\rho }_{\min },\underline{\rho }_{\max }])},\\&\Vert (\partial _t\underline{\rho },\partial _{tt}\underline{\rho },\nabla {\underline{\rho }},\partial _t\nabla {\underline{\rho }},\nabla ^2\underline{\varvec{u}})\Vert _{L^\infty ((0,T)\times \varOmega )}). \end{aligned}$$

Using the estimate (7.4), we can show

$$\begin{aligned} \int _0^{t_m}{{\mathcal {E}}(\rho _h \,|\, {\underline{\rho }}_{\tau })}dt=\tau \sum _{n=1}^m{\mathcal {E}}(\rho _h^n \,|\, {\underline{\rho }}^n)\le \tau \sum _{n=1}^{m-1}{\mathcal {E}}(\rho _h^n \,|\, {\underline{\rho }}^n)+C\tau , \end{aligned}$$

which implies that

$$\begin{aligned}&{\mathcal {E}}(\rho _h^m \,|\,{\underline{\rho }}^m)+\sum _{n=1}^m\tau \int _{\varOmega }{[\mu |\nabla (\varvec{u}_h^n-\underline{\varvec{u}}_{\tau }^n)|^2+(\lambda +\mu )|{\text {div}}(\varvec{u}_h^n-\underline{\varvec{u}}_{\tau }^n)|^2]}dx\nonumber \\&\quad \le C({\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+h^A+\tau )+C\tau \sum _{n=1}^{m-1}{\mathcal {E}}(\rho _h^n \,|\, {\underline{\rho }}^n). \end{aligned}$$
(7.21)

By applying the standard discrete version of Gronwall’s lemma for the inequality (7.21), the proof is thus complete. \(\square \)

Finally, we will give an error estimate for the discrete \(L^2(L^2)\) norm of \(p(\rho _h)\).

Theorem 7.7

Suppose that the condition of Theorem 7.6 holds. Let the families \((\rho _h,\varvec{u}_h)\) and \(({\underline{\rho }}_{\tau },\underline{\varvec{u}}_{\tau })\) be defined as in (3.16)–(3.17) and (7.3), respectively. Then there exists

$$\begin{aligned} C:=&C(T,\varOmega ,M_0,E_0,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\\&\Vert (\partial _t{\underline{\rho }},\partial _{tt}{\underline{\rho }},\nabla {\underline{\rho }},\partial _t\nabla {\underline{\rho }},\nabla ^2\underline{\varvec{u}})\Vert _{L^\infty ((0,T)\times \varOmega )})>0, \end{aligned}$$

such that for any \(1\le m\le N\), we have

$$\begin{aligned} \tau \sum _{n=1}^m\Vert p(\rho _h^n)-p({\underline{\rho }}^n)\Vert _{L^2(\varOmega )}^2\le C({\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+h^A+\tau ),\quad A:=\frac{\min \{\epsilon ,1\}}{2}. \end{aligned}$$
(7.22)

Proof

Taking \(t=t_n\) in (7.17), multiplying this identity by \(\varvec{v}_h\in {\mathbb {V}}_h\) and integral over \(\varOmega \), we conclude that

$$\begin{aligned} \mu \int _{\varOmega }{\nabla \underline{\varvec{u}}^n:\nabla \varvec{v}_h}dx&+(\lambda +\mu )\int _{\varOmega }{{\text {div}}\underline{\varvec{u}}^n{\text {div}}\varvec{v}_h}dx\nonumber \\&-\int _{\varOmega }{p({\underline{\rho }}^n){\text {div}}\varvec{v}_h}dx=0. \end{aligned}$$
(7.23)

Subtracting (7.23) from (3.8), we can get the error equation

$$\begin{aligned}&\mu \int _{\varOmega }{\nabla (\varvec{u}_h^n-\underline{\varvec{u}}^n):\nabla \varvec{v}_h}dx+(\lambda +\mu )\int _{\varOmega }{{\text {div}}(\varvec{u}_h^n-\underline{\varvec{u}}^n){\text {div}}\varvec{v}_h}dx\nonumber \\&-\int _{\varOmega }{(p(\rho _h^n)-p({\underline{\rho }}^n)){\text {div}}\varvec{v}_h}dx=0,\quad \forall \;\varvec{v}_h\in {\mathbb {V}}_h. \end{aligned}$$
(7.24)

Let \(r_{\rho }^n:=(p(\rho _h^n)-p({\underline{\rho }}_h^n))-\frac{1}{|\varOmega |}\int _{\varOmega }{(p(\rho _h^n)-p({\underline{\rho }}_h^n))}dx\) for \(1\le n\le N\). Taking \(\varvec{v}_h=\varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_{\rho }^n]\) in (7.24), we can show

$$\begin{aligned}&\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^2(\varOmega )}^2=|\varOmega |^{-1}\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^1(\varOmega )}^2+\mu \int _{\varOmega }{\nabla (\varvec{u}_h^n-\underline{\varvec{u}}^n):\nabla \varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_{\rho }^n]}dx\\&+(\lambda +\mu )\int _{\varOmega }{{\text {div}}(\varvec{u}_h^n-\underline{\varvec{u}}^n){\text {div}}\varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_{\rho }^n]}dx-\int _{\varOmega }{(p({\underline{\rho }}_h^n)-p({\underline{\rho }}^n)){\text {div}}\varPi _h^{{\mathbb {V}}}{\textbf{B}}[r_{\rho }^n]}dx. \end{aligned}$$

By applying the Cauchy–Schwarz inequality, the estimates (3.3) and (3.12), we obtain

$$\begin{aligned} \Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^2(\varOmega )}^2\le&C(\varOmega )\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^1(\varOmega )}\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^2(\varOmega )}\\&+C(\varOmega )\Vert \nabla (\varvec{u}_h^n-\underline{\varvec{u}}^n)\Vert _{\varvec{L}^2(\varOmega )}\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^2(\varOmega )}\\&+C(\varOmega )\Vert p({\underline{\rho }}_h^n)-p({\underline{\rho }}^n)\Vert _{L^2(\varOmega )}\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^2(\varOmega )}, \end{aligned}$$

which implies that

$$\begin{aligned} \Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^2(\varOmega )}^2\le&C(\varOmega )\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^1(\varOmega )}^2+C(\varOmega )\Vert \nabla (\varvec{u}_h^n-\underline{\varvec{u}}^n)\Vert _{\varvec{L}^2(\varOmega )}^2\nonumber \\&+C(\varOmega )\Vert p({\underline{\rho }}_h^n)-p({\underline{\rho }}^n)\Vert _{L^2(\varOmega )}^2. \end{aligned}$$
(7.25)

Summing (7.25) from \(n=1\) to \(n=m\) and multiplying the resulting inequality by \(\tau \), we conclude that

$$\begin{aligned} \tau \sum _{n=1}^m\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^2(\varOmega )}^2\le C(\varOmega )\sum _{i=8}^{10}{\mathcal {L}}_i, \end{aligned}$$

where the terms \({\mathcal {L}}_i\) (\(8\le i\le 10\)) are defined by

$$\begin{aligned} {\mathcal {L}}_8:=&\tau \sum _{n=1}^m\Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^1(\varOmega )}^2,\quad {\mathcal {L}}_9:=\tau \sum _{n=1}^m\Vert \nabla (\varvec{u}_h^n-\underline{\varvec{u}}^n)\Vert _{\varvec{L}^2(\varOmega )}^2,\\ {\mathcal {L}}_{10}:=&\tau \sum _{n=1}^m\Vert p({\underline{\rho }}_h^n)-p({\underline{\rho }}^n)\Vert _{L^2(\varOmega )}^2. \end{aligned}$$

Bound on \({\mathcal {L}}_9\). By applying the estimate (3.1), the mean value theorem and \({\underline{\rho }}^n,{\underline{\rho }}_h^n\in [{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }]\), we can show

$$\begin{aligned} \Vert p(\rho _h^n)-p({\underline{\rho }}_h^n)\Vert _{L^1(\varOmega )}\le&\Vert p({\underline{\rho }}^n)-p({\underline{\rho }}_h^n)\Vert _{L^1(\varOmega )}+\Vert p'({\underline{\rho }}^n)(\rho _h^n-{\underline{\rho }}^n)\Vert _{L^1(\varOmega )}\\&+\Vert p(\rho _h^n)-p'({\underline{\rho }}^n)(\rho _h^n-{\underline{\rho }}^n)-p({\underline{\rho }}^n)\Vert _{L^1(\varOmega )}\\ \le&C(\varOmega ,|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \nabla {\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )})h\\&+C(|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])})\Vert \rho _h^n-{\underline{\rho }}^n\Vert _{L^1(\varOmega )}+C{\mathcal {E}}(\rho _h^n \,|\, {\underline{\rho }}^n). \end{aligned}$$

Let \(\varOmega _{h,1}^n:=\{\frac{{\underline{\rho }}^n}{2}<\rho _h^n<2{\underline{\rho }}^n\}\) and \(\varOmega _{h,2}^n:=\varOmega \setminus \varOmega _{h,1,n}\). By applying the estimate (7.2) and (7.20), we obtain

$$\begin{aligned} \Vert \rho _h^n-{\underline{\rho }}^n\Vert _{L^1(\varOmega )}\le&C(\varOmega )\Vert \rho _h^n-{\underline{\rho }}^n\Vert _{L^2(\varOmega _{h,1}^n)}+C(\varOmega )\Vert \rho _h^n-{\underline{\rho }}^n\Vert _{L^{\frac{6}{5}}(\varOmega _{h,2}^n)}\\ \le&C(\varOmega ,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }){\mathcal {E}}^{\frac{1}{2}}(\rho _h^n \,|\, {\underline{\rho }}^n)+C(\varOmega ,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }){\mathcal {E}}^{\frac{5}{6}}(\rho _h^n \,|\, {\underline{\rho }}^n). \end{aligned}$$

Using the estimate (7.4), we get that

$$\begin{aligned} \Vert \rho _h^n-{\underline{\rho }}^n\Vert _{L^1(\varOmega )}\le C(\varOmega ,E_0,{\underline{\rho }}_{\min },{\underline{\rho }}_{\max },|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])}){\mathcal {E}}^{\frac{1}{2}}(\rho _h^n \,|\, {\underline{\rho }}^n). \end{aligned}$$

Combining the error estimate of Theorem 7.6, which implies that

$$\begin{aligned} |{\mathcal {L}}_8|\le C({\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+h^A+\tau ). \end{aligned}$$

By a similar argument, we can show

$$\begin{aligned} |{\mathcal {L}}_9|\le&C({\mathcal {E}}(\rho _h^0 \,|\,{\underline{\rho }}^0)+h^A+\tau ),\\ |{\mathcal {L}}_{10}|\le&C(\varOmega ,T,|p'|_{C^1([{\underline{\rho }}_{\min },{\underline{\rho }}_{\max }])},\Vert \nabla {\underline{\rho }}\Vert _{L^\infty ((0,T)\times \varOmega )})h^2. \end{aligned}$$

Combining the above analysis, the proof is thus complete. \(\square \)