1 Introduction

Let \(\Omega \subset \mathbf{R} ^3\) be a bounded and convex domain with the sufficiently smooth boundary \(\partial \Omega \) and [0, T] the time interval with some \(T>0\). We will use notations \(Q_T=[0, T]\times \Omega \) and \(\Sigma _T=[0, T]\times \partial \Omega \). In this paper, we consider the 3D incompressible Navier–Stokes equations with mass diffusion (or called the Kazhikhov–Smagulov model) in \(Q_T\), which can be deduced from the following 3D compressible Navier–Stokes equations:

$$\begin{aligned}&\rho _t+\nabla \cdot (\rho \mathbf{v} ) = 0, \end{aligned}$$
(1.1)
$$\begin{aligned}&(\rho \mathbf{v} )_t+\nabla \cdot (\rho \mathbf{v} \otimes \mathbf{v} ) -\mu \Delta \mathbf{v} - (\mu +\widetilde{\lambda })\nabla (\nabla \cdot \mathbf{v} )+\nabla q = \mathbf{f} . \end{aligned}$$
(1.2)

In the above system (1.11.2), the unknows \(\rho : Q_T\rightarrow \mathbf{R} \) is the density of the fluid, \(\mathbf{v} : Q_T\rightarrow \mathbf{R} ^3\) is the velocity of the fluid and \(q: Q_T\rightarrow \mathbf{R} \) is the pressure which depends on the density \(\rho \). \(\mathbf{f} : Q_T\rightarrow \mathbf{R} ^3\) denotes the external force, \(\mu \) and \(\widetilde{\lambda }\) are two constants and present viscosity coefficients which are assumed to satisfy \(\mu >0\) and \(3 \widetilde{\lambda }+ 2\mu >0\). If the mass diffusion process obeys Fick’s law (cf. [12]), the velocity \(\mathbf{v} \) of the fluid can be decomposed into a potential part and an incompressible part:

$$\begin{aligned} \mathbf{v} = \mathbf{u} - \lambda \nabla \ln \rho \quad \text{ with }\quad \nabla \cdot \mathbf{u} =0, \end{aligned}$$

where \(\lambda >0\) is the mass diffusion coefficient. Then the compressible Navier–Stokes equations (1.11.2) can be rewritten as

$$\begin{aligned}&\rho _t-\lambda \Delta \rho +\nabla \rho \cdot \mathbf{u} = 0, \end{aligned}$$
(1.3)
$$\begin{aligned}&(\rho \mathbf{u} )_t+\nabla \cdot \left( (\rho \mathbf{u} -\lambda \nabla \rho )\otimes \mathbf{u} -\lambda \mathbf{u} \otimes \nabla \rho \right) -\mu \Delta \mathbf{u} +\lambda ^2 \nabla \cdot ( \rho ^{-1} \nabla \rho \otimes \nabla \rho )+\nabla P =\mathbf{f} , \end{aligned}$$
(1.4)
$$\begin{aligned}&\nabla \cdot \mathbf{u} = 0 , \end{aligned}$$
(1.5)

where \(P=q-\lambda \rho _t+\lambda (2\mu +\widetilde{\lambda })\Delta \ln \rho \). Eliminating the \(\lambda ^2\)-term in (1.4) and using the following relations:

$$\begin{aligned} (\rho \mathbf{u} )_t+\nabla \cdot \left( (\rho \mathbf{u} -\lambda \nabla \rho )\otimes \mathbf{u} \right)= & {} \rho \mathbf{u} _t+\left( (\rho \mathbf{u} -\lambda \nabla \rho )\cdot \nabla \right) \mathbf{u} , \\ -\lambda \nabla \cdot (\mathbf{u} \otimes \nabla \rho )= & {} -\lambda \nabla (\mathbf{u} \cdot \nabla \rho )+\lambda \nabla \cdot (\rho (\nabla \mathbf{u} )^t), \end{aligned}$$

we get the simplified model of (1.31.5) in \(Q_T\) which is described as

$$\begin{aligned}&\rho _t-\lambda \Delta \rho +\nabla \rho \cdot \mathbf{u} = 0, \end{aligned}$$
(1.6)
$$\begin{aligned}&\rho \mathbf{u} _t+ \left( (\rho \mathbf{u} -\lambda \nabla \rho )\cdot \nabla \right) \mathbf{u} - \nabla \cdot \left( \mu \nabla \mathbf{u} - \lambda \rho (\nabla \mathbf{u} )^t \right) + \nabla p = \mathbf{f} \end{aligned}$$
(1.7)
$$\begin{aligned}&\nabla \cdot \mathbf{u} = 0 , \end{aligned}$$
(1.8)

where \(p=P-\lambda \mathbf{u} \cdot \nabla \rho \). The above coupled system (1.61.8) are the incompressible Navier–Stokes equations with mass diffusion. It is clear that the system (1.61.8) reduce to the incompressible Navier–Stokes equations with variable density if \(\lambda =0\).

We complete (1.61.8) by the following boundary conditions

$$\begin{aligned} \mathbf{u} =0\quad \text{ and }\quad \partial _\mathbf{n} \rho =0 \quad \text{ on } \ \Sigma _T \end{aligned}$$
(1.9)

and the initial conditions

$$\begin{aligned} \rho (0, x)=\rho _0(x) \quad \text{ and } \quad \mathbf{u} (0, x)=\mathbf{u} _0(x)\quad \text{ in } \ \Omega , \end{aligned}$$
(1.10)

where \(\mathbf{n} \) denotes the outwards unit normal vector to \(\partial \Omega \). Furthermore, we assume that there have two positive constants m and M such that

$$\begin{aligned} 0< m \le \rho _0({ x})\le M \quad \text{ in } \ \Omega , \end{aligned}$$
(1.11)

which means that there has no vacuum state in \(\Omega \).

We recall some known results on the incompressible Navier–Stokes equations with mass diffusion. For the full model (1.31.5), Beirão da Veiga in [31] and Secchi in [29] established the local existence of the strong solution in terms of linearization and a fixed point method. Moreover, Secchi in [29] proved the eixstence and uniqueness of a global weak solution to 2D problem by imposing smallness on \(\lambda / \mu \) and established the asymptotic behavior towards a weak solution to the incompressible Navier–Stokes problem with variable density when the mass diffusion coefficient \(\lambda \rightarrow 0\). Guillén-González etc. in [15] proved the global existence of the strong solution for small initial data by means of an iterative method. When the mass diffusion coefficient \(\lambda \rightarrow 0\) and the viscosity coefficient \(\mu \rightarrow 0\), Araruna etc. in [4] studied the asymptotic behavior towards a solution to a inhomogeneous, inviscid and incompressible fluid governed by an Euler type system. For the numerical method of (1.31.5), Cabrales etc. in [6] proposed a fully discrete decoupled scheme by using a first-order time discretization and a \(C^0\) finite element approximation for all unknowns and proved some stability and convergences results.

For the simplified model (1.61.8), Kazhikhov and Smagulov in [21] proved the global existence of the weak solution and the local existence of the strong solution by means of the Galerkin method under the assumptions that the initial density \(\rho _0(x)\) satisfies (1.11) and the viscosity and mass diffusion coefficients satisfy \(\lambda <2\mu / (M-m)\). The global existence of the weak solution in the non-cylindrical domain was derived in [26]. Secchi in [28] studied the 3D Cauchy problem and established the local existence and uniqueness of the strong solution. The global existence of the strong solution to the 2D Cauchy problem and the 2D initial-boundary value problem were studied in [8, 9], respectively. For the numerical methods, there are not many works concerning numerical analysis of the simplified model (1.61.8). By using a first-order time discretization and a \(C^0\) finite element approximation for all unknowns, two decoupled numerical schemes were proposed for solving the 2D problem and the 3D problem in [16] and [17], respectively, where the stabilities of algorithms and the convergences of numerical solutions were investigated. Other numerical schemes can be found in [10] and [11, 27], where an hybrid finite volume-finite element scheme and spectral Galerkin schemes were studied, respectively. Furthermore, the stability and convergence of numerical algorithm were investigated in [10].

To our best knowledge, the first error analysis of finite element fully discrete scheme for the simplified model (1.61.8) was presented by Guillén-González and Gutiérrez-Santacreu in [18]. To describe error estimates derived in [18], we introduce some notations. Let \(0=t_0<t_1<\cdots <t_N=T\) be a uniform partition of the time interval [0, T] with the time step \(\tau =T/N\) and \(t_n=n\tau \). If \(\{\mathbf{v }^n\}_{n=1}^N\) is a given vector sequence with \(\mathbf{v} ^n\in X\) for a Banach space X, we introduce the following notations for the discrete-in-time norms:

$$\begin{aligned} \Vert \mathbf{v} ^n\Vert _{l^2(X)}=\left( \tau \displaystyle \sum _{n=1}^N \Vert \mathbf{v} ^n\Vert _X^2 \right) ^{1/2}\quad \text{ and }\quad \Vert \mathbf{v} ^n\Vert _{l^\infty (X)}= \sup _{1\le n\le N} \Vert \mathbf{v} ^n\Vert _X. \end{aligned}$$

Let \((\mathbf{u} _h^n, \rho _h^n)\) be the finite element approximations of \((\mathbf{u} (t_n), \rho (t_n))\) for \(1\le n\le N\). By using the mini-element (cf. [14]) for the approximation of velocity-pressure pair and the \(P_2\) element for the approximation of density, the authors in [18] proved that

$$\begin{aligned} \Vert \mathbf{u} (t_n)-\mathbf{u} _h^n\Vert _{l^\infty (L^2)} + \Vert \rho (t_n)-\rho _h^n\Vert _{l^\infty (H^1)} \le C (\tau + h) \end{aligned}$$
(1.12)

under the weaker regularity assumptions on the exact solution. Concretely, the authors in [18] avoided using the assumption \(\mathbf{u} _{tt}\in L^2(0,T; \mathbf{L} ^{6/5}(\Omega ))\) which required that the data should satisfy an extra compatibility condition at \(t=0\).

In this paper, a decoupled numerical scheme is proposed by using the mini-element for the velocity-pressure pair and the \(P_2\) element for the density as that in [18]. Inspired by [22], this scheme is slightly different the scheme in [18] by introducing the post-processed velocity in the discretization of the density equation and the stable terms in the discretization of the Navier–Stokes type equation such that the proposed finite element scheme is unconditionally stable. The main result derived in this paper is the following optimal error estimate:

$$\begin{aligned} \Vert \mathbf{u} (t_n)-\mathbf{u} _h^n\Vert _{l^\infty (L^2)} + \Vert \rho (t_n)-\rho _h^n\Vert _{l^\infty (H^1)} \le C (\tau + h^2), \end{aligned}$$
(1.13)

where \(h>0\) is the mesh size and \(C>0\) is some constant independent of h and \(\tau \). However, compared to [18], the higher regularities of the exact solution are assumed in this paper. The method of analysis is based on the technique of error splitting for the nonlinear parabolic problems proposed by Li and Sun in [23,24,25] and further developed in [2, 3, 7, 13, 32].

The rest of this paper is organized as follows. In Sect. 2, we state the proposed linear and decoupled Euler finite element scheme, present the stability of numerical scheme in Theorem 2.3 and the main result in Theorem 2.4. The proof of Theorem 2.4 is given in Sect. 3 by using the technique of error splitting. In particular, we firstly derive temporal error estimates and regularities of solutions to the time discrete scheme in Sect. 3.1, and then prove optimal spatial error estimates in Sect. 3.2.

2 Numerical Scheme and Main Result

2.1 Preliminaries

For the mathematical setting, we introduce the following notations. For \(k\in \mathbb N^+\) and \(1\le p\le +\infty \), we use \(W^{k,p}(\Omega )\) to denote the classical Sobolev space. The norm in \(W^{k,p}(\Omega )\) is denoted by \(\Vert \cdot \Vert _{W^{k,p}}\) defined by a classical way (cf. [1]). Denote \(W_0^{k,p}(\Omega )\) be the subspace of \(W^{k,p}(\Omega )\) where the functions have zero trace on \(\partial \Omega \). Especially, \(W^{0,p}(\Omega )\) is the Lebesgue space \(L^p(\Omega )\) and \(W^{k,2}(\Omega )\) is the Hilbert space which is simply denoted by \(H^k(\Omega )\). The boldface notations \(\mathbf{H} ^k(\Omega ), \mathbf{W} ^{k,p}(\Omega )\) and \(\mathbf{L} ^p(\Omega )\) are used to denote the vector-value Sobolev spaces corresponding to \(H^k(\Omega )^3, W^{k,p}(\Omega )^3\) and \(L^p(\Omega )^3\), respectively. We use \((\cdot ,\cdot )\) to denote the \(L^2\) or \(\mathbf{L} ^2\) inner product.

Introduce the following function spaces:

$$\begin{aligned} \mathbf{H}= & {} \{\mathbf{u }\in \mathbf{L} ^2(\Omega ), \ \nabla \cdot \mathbf{u} =0 \ \text{ in } \ \Omega , \ \mathbf{u} \cdot \mathbf{n} =0 \ \text{ on } \ \partial \Omega \},\\ \mathbf{V}= & {} \mathbf{H} ^1_0(\Omega ), \quad \mathbf{V} _0=\{\mathbf{u }\in \mathbf{V} , \ \nabla \cdot \mathbf{u} =0 \ \text{ in } \ \Omega \}, \\ \mathbf{H} (\text{ div }, \Omega )= & {} \{ \mathbf{u }\in \mathbf{L} ^2(\Omega ), \ \nabla \cdot \mathbf{u} \in L^2(\Omega ) \}, \\ W= & {} \{ r\in H^1(\Omega ), \ \displaystyle \int _\Omega r(x) dx=0\},\\ M= & {} L_0^2(\Omega )=\{ p\in L^2(\Omega ), \ \displaystyle \int _\Omega p(x) dx=0\} \end{aligned}$$

and

$$\begin{aligned} H_N^2(\Omega )= & {} \{ \rho \in H^2(\Omega ), \ \partial _\mathbf{n} \rho =0 \ \text{ on } \ \partial \Omega , \ \displaystyle \int _\Omega \rho (x) dx = \displaystyle \int _\Omega \rho _0(x) dx \}, \\ H_{N,0}^2(\Omega )= & {} \{ \rho \in H^2(\Omega ), \ \partial _\mathbf{n} \rho =0 \ \text{ on } \ \partial \Omega , \ \displaystyle \int _\Omega \rho (x) dx =0\}. \end{aligned}$$

It is known that the norms \(\Vert \nabla \rho \Vert _{H^1}\) and \(\Vert \rho \Vert _{H^2}\) are equivalent to the seminorm \(\Vert \Delta \rho \Vert _{L^2}\) for \(\rho \in H_N^2(\Omega )\) and \(\rho \in H_{N,0}^2(\Omega )\), respectively.

Introduce the trilinear term \(a(\rho ; \mathbf{u} , \mathbf{v} )\) by

$$\begin{aligned} a(\rho ; \mathbf{u} , \mathbf{v} )=\mu (\nabla \mathbf{u} , \nabla \mathbf{v} ) -\lambda \displaystyle \int _\Omega \left( \rho -\displaystyle \frac{ \widetilde{M}+\widetilde{m}}{2} \right) (\nabla \mathbf{u} )^t:\nabla \mathbf{v} dx \end{aligned}$$

with

$$\begin{aligned} \widetilde{M}>M,\quad 0<\widetilde{m}<m\quad \text{ such } \text{ that } \quad \displaystyle \frac{ \lambda (\widetilde{M}-\widetilde{m}) }{2}<\mu \end{aligned}$$
(2.1)

for any \(\rho \in L^\infty (\Omega )\) and \(\mathbf{u} , \mathbf{v} \in \mathbf{V} \). Under the condition (2.1), we can see that if \(\widetilde{m}\le \rho (x)\le \widetilde{M}\), then

$$\begin{aligned} a(\rho ; \mathbf{u} , \mathbf{u} )\ge & {} {\mu _1} \Vert \nabla \mathbf{u} \Vert _{L^2}^2 \ \text{ where } \ \mu _1=\mu - \displaystyle \frac{ \lambda (\widetilde{M}-\widetilde{m}) }{2}>0, \end{aligned}$$
(2.2)
$$\begin{aligned} a(\rho ; \mathbf{u} , \mathbf{v} )\le & {} \mu _2 \Vert \nabla \mathbf{u} \Vert _{L^2}\Vert \nabla \mathbf{v} \Vert _{L^2}. \end{aligned}$$
(2.3)

The existence and uniqueness of weak solution to (1.61.8) are established by Kazhikhov and Smagulov in [21]. We recall it in the following theorem.

Theorem 2.1

Let \(\mathbf{u} _0\in \mathbf{H} \) and \(\rho _0\in W\) satisfying (1.11) and \(\mathbf{f} \in L^2(0,T; \mathbf{L} ^2(\Omega ))\). Suppose that the constants \(\lambda , \mu , m\) and M satisfies

$$\begin{aligned} \lambda < \displaystyle \frac{2\mu }{M-m}. \end{aligned}$$
(2.4)

Then there exists a unique weak solution \((\rho , \mathbf{u} )\) to (1.61.8) such that the solution satisfies

$$\begin{aligned}&\mathbf{u} \in L^\infty (0,T;\mathbf{H} )\cap L^2(0,T;\mathbf{V} _0), \quad \rho \in L^\infty (0,T;W)\cap L^2(0,T;H_N^2(\Omega )), \end{aligned}$$
(2.5)
$$\begin{aligned}&0<m\le \rho (t, x) \le M \quad \text{ in } \ Q_T \end{aligned}$$
(2.6)

and the energy inequalities:

$$\begin{aligned}&\displaystyle \frac{1}{2}\Vert \sigma (t)\mathbf{u} (t)\Vert _{L^2}^2+\left( \mu -\displaystyle \frac{\lambda (M-m)}{2} \right) \displaystyle \int _0^t \Vert \nabla \mathbf{u} (\tau )\Vert _{L^2}^2 d\tau \le \displaystyle \frac{1}{2}\Vert \sigma _0\mathbf{u} _0\Vert _{L^2}^2+\displaystyle \int _0^t (\mathbf{f} (\tau ), \mathbf{u} (\tau )) d\tau , \\&\quad \displaystyle \frac{1}{2}\Vert \rho (t)\Vert _{L^2}^2+ \lambda \displaystyle \int _0^t \Vert \nabla \rho (\tau )\Vert _{L^2}^2 d\tau \le \displaystyle \frac{1}{2}\Vert \rho _0\Vert _{L^2}^2 \end{aligned}$$

for all \(0<t\le T\), where \(\sigma (t)=\sqrt{\rho (t)}\) and \(\sigma _0=\sqrt{\rho _0}\).

Throughout this paper, we make the following assumptions on the prescribed data, the regularity of the solution to (1.61.10) and the domain \(\Omega \).

Assumption (A1): Assume that the prescribed data \(\mathbf{f} \), \(\mathbf{u} _0\) and \(\rho _0\) satisfy

$$\begin{aligned} \mathbf{f} \in L^2(0,T; \mathbf{L} ^4(\Omega )), \quad \mathbf{u} _0\in \mathbf{V} _0\cap \mathbf{H} ^{2}(\Omega ) \quad \text{ and }\quad \rho _0\in H^2_N(\Omega ) \ \text{ with } \text{(1.11) }. \end{aligned}$$

Assumption (A2): Let \(\lambda , \mu , m, M\) satisfy (2.4) and \(\widetilde{m}, \widetilde{M}\) satisfy (2.1).

Assumption (A3): Assume that the solution \((\rho , \mathbf{u} , p)\) satisfies the following regularities:

$$\begin{aligned} \rho\in & {} L^\infty (0, T; H^{3}(\Omega )\cap W), \ \rho _t\in L^2(0,T; H^2(\Omega ))\cap L^\infty (0,T; H^1(\Omega )), \\ \mathbf{u}\in & {} L^\infty (0, T; \mathbf{W} ^{2,4}(\Omega )\cap \mathbf{V} _0), \ \mathbf{u} _t \in L^\infty (0,T; \mathbf{H} ^1(\Omega ))\cap L^2(0,T; \mathbf{H} ^2(\Omega )),\\ \rho _{tt}\in & {} L^2(0,T; L^2(\Omega )), \ \mathbf{u} _{tt} \in L^2(0,T; \mathbf{L} ^2(\Omega )), \ p \in L^\infty (0, T; H^2(\Omega )\cap M). \end{aligned}$$

Assumption (A4): Assume that the boundary \(\partial \Omega \) is sufficiently smooth such that the unique solution \(\phi \) of the Neumann problem

$$\begin{aligned} -\Delta \phi = g \quad \text{ in } \ \Omega , \qquad \partial _\mathbf{n} \phi =0\quad \text{ on } \ \partial \Omega \end{aligned}$$

for prescribed \(g\in M\cap H^{k}(\Omega )\) satisfies

$$\begin{aligned} \Vert \phi \Vert _{H^{2+k}} \le C \Vert g\Vert _{H^{k}},\quad \text{ for } \ k=0,1, \end{aligned}$$

and the unique solution \((\mathbf{v} , q)\) of the Stokes problem

$$\begin{aligned} -\Delta \mathbf{v} + \nabla q = \mathbf{g} \ \text{ in } \ \Omega , \quad \nabla \cdot \mathbf{v} =0 \ \text{ in } \ \Omega , \quad \mathbf{v} =0 \ \text{ on } \ \partial \Omega \end{aligned}$$

for prescribed \(\mathbf{g} \in \mathbf{L} ^p(\Omega )\) with \(1\le p\le 4\) satisfies

$$\begin{aligned} \Vert \mathbf{v} \Vert _{W^{2,p}} + \Vert q\Vert _{W^{1,p}} \le C \Vert \mathbf{g} \Vert _{L^p}. \end{aligned}$$

Remark 2.1

The verification of the regularity assumption \(\mathbf{u} _{tt} \in L^2(0,T; \mathbf{L} ^2(\Omega ))\) should involve an extra compatibility condition on the data at \(t=0\) which is not generally satisfied (see such condition for Navier–Stokes equations in [19]). We make this assumption merely to simplify the presentation. In [18], such assumption was avoided by using the technique of Euler integrator in the consistency error analysis.

2.2 Time Discrete Scheme

We first describe the time discrete scheme based on the backward Euler method. Let \(0=t_0<t_1<\cdots <t_N=T\) be a uniform partition of the time interval [0, T] with the time step \(\tau =T/N\) and \(t_n=n\tau \) with \(0\le n\le N\).

Given \(\rho ^0=\rho _0\) and \(\mathbf{u} ^0=\mathbf{u} _0\), we consider the following first-order Euler time discrete scheme for the simplified system (1.61.10).

Euler time discrete scheme:

Step I: For given \(\rho ^n\) and \(\mathbf{u} ^n\), we find \(\rho ^{n+1}\) by

$$\begin{aligned} D_\tau \rho ^{n+1} -\lambda \Delta \rho ^{n+1} + \nabla \rho ^{n+1}\cdot \mathbf{u} ^n= 0 \end{aligned}$$
(2.7)

with the boundary condition \(\partial _\mathbf{n} \rho ^{n+1}=0\) on \(\partial \Omega \), where

$$\begin{aligned} D_\tau \rho ^{n+1} = \displaystyle \frac{\rho ^{n+1}-\rho ^n}{\tau }. \end{aligned}$$

Step II: For given \(\rho ^n\), \(\mathbf{u} ^n\) and \(\rho ^{n+1}\) derived from (2.7), we find \((\mathbf{u} ^{n+1}, p^{n+1})\) by

$$\begin{aligned}&\rho ^{n} D_\tau \mathbf{u} ^{n+1} - \nabla \cdot \left( \mu \nabla \mathbf{u} ^{n+1}- \lambda \rho ^{n+1}(\nabla \mathbf{u} ^{n+1} )^t \right) + \rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{u} ^{n+1}\nonumber \\&\quad - \lambda (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{u} ^{n+1} +\nabla p^{n+1} = \mathbf{f} ^{n+1},\quad \nabla \cdot \mathbf{u} ^{n+1}=0 \end{aligned}$$
(2.8)

with the boundary condition \(\mathbf{u} ^{n+1}=0\) on \(\partial \Omega \).

The weak form of (2.72.8) are described as follows. Find the weak solutions \(\rho ^{n+1}\in W\) and \((\mathbf{u} ^{n+1}, p^{n+1})\in \mathbf{V} \times M\), respectively, by

$$\begin{aligned} (D_\tau \rho ^{n+1}, r) + \lambda (\nabla \rho ^{n+1}, \nabla r) + (\nabla \rho ^{n+1}\cdot \mathbf{u} ^n, r)= 0, \qquad \forall \ r\in W, \end{aligned}$$
(2.9)

and

$$\begin{aligned}&(\rho ^{n} D_\tau \mathbf{u} ^{n+1}, \mathbf{v} ) + a(\rho ^{n+1}; \mathbf{u} ^{n+1}, \mathbf{v} ) - (\nabla \cdot \mathbf{v} , p^{n+1}) + (\nabla \cdot \mathbf{u} ^{n+1}, q) \nonumber \\&\quad + (\rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{u} ^{n+1}, \mathbf{v} ) - \lambda ( (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{u} ^{n+1}, \mathbf{v} ) =(\mathbf{f} ^{n+1}, \mathbf{v} ), \ \forall (\mathbf{v} , q)\in \mathbf{V} \times M. \end{aligned}$$
(2.10)

In the above form (2.10), we use

$$\begin{aligned} \displaystyle \int _\Omega (\nabla \mathbf{u} )^t:\nabla \mathbf{v} dx =0 \end{aligned}$$

due to \(\nabla \cdot \mathbf{u} =0\) and \(\nabla \cdot ((\nabla \mathbf{u} )^t)=0\) and \(\mathbf{v} =0\) on \(\partial \Omega \).

From the assumption (A4) on the elliptic regularity, the well-posedness of solution to (2.7) was established in [17]. We recall it in the following lemma.

Lemma 2.1

For each \(0\le n\le N-1\), if

$$\begin{aligned} \Vert \nabla \mathbf{u} ^n\Vert _{L^2}\le \kappa _1, \end{aligned}$$
(2.11)

for some \(\kappa _1>0\) being independent of \(\tau \) and n, then for sufficiently small \(\tau \), the solution \(\rho ^{n+1}\) to (2.7) satisfies

$$\begin{aligned}&m\le \rho ^{n+1}(x)\le M ,\quad \forall \ x\in \Omega , \end{aligned}$$
(2.12)
$$\begin{aligned}&\Vert \rho ^{n+1}\Vert _{H^1}^2+\tau \displaystyle \sum _{i=0}^n\Vert \rho ^{i+1}\Vert _{H^2}^2 \le \kappa _2, \end{aligned}$$
(2.13)

for some \(\kappa _2>0\) being independent of \(\tau \) and n.

Remark 2.2

Although \(\mathbf{u} ^n\) in (2.7) replaces \(\mathbf{u} _h^n\) in [17], the proof of Lemma 2.1 follows immediately from the proof of Lemma 3.4 in [17] by noting the fact that \(\tau \displaystyle \sum _{n=1}^N\Vert \nabla \mathbf{u} ^n\Vert _{L^2}^2\le C\). Please see Appendix A in [17].

Next, we discuss the stability of the time discrete scheme (2.72.8). Setting \(\phi =2\tau \rho ^{n+1}\) in (2.9) gives

$$\begin{aligned} \Vert \rho ^{n+1}\Vert _{L^2}^2 - \Vert \rho ^{n}\Vert _{L^2}^2 + \Vert \rho ^{n+1}-\rho ^n\Vert _{L^2}^2 +2\lambda \tau \Vert \nabla \rho ^{n+1}\Vert _{L^2}^2=0 \end{aligned}$$

by using

$$\begin{aligned} 2 \displaystyle \int _\Omega (\nabla \rho ^{n+1}\cdot \mathbf{u} ^{n})\rho ^{n+1}dx= \displaystyle \int _\Omega \nabla |\rho ^{n+1}|^2\cdot \mathbf{u} ^{n} dx =-\displaystyle \int _\Omega |\rho ^{n+1}|^2\nabla \cdot \mathbf{u} ^n=0. \end{aligned}$$

Taking the sum gives

$$\begin{aligned} \Vert \rho ^{n+1}\Vert _{L^2}^2+2\lambda \tau \displaystyle \sum _{i=0}^n\Vert \nabla \rho ^{i+1}\Vert _{L^2}^2 \le \Vert \rho _0\Vert _{L^2}^2 \end{aligned}$$

for all \(0\le n\le N-1\).

Suppose that

$$\begin{aligned} m\le \rho ^{n+1}(x)\le M,\quad \forall \ 0\le n\le N-1. \end{aligned}$$
(2.14)

Setting \((\mathbf{v} , q)=2\tau (\mathbf{u} ^{n+1}, p^{n+1})\) in (2.10) and using (2.2), we have

$$\begin{aligned}&\Vert \sigma ^n\mathbf{u} ^{n+1}\Vert _{L^2}^2 - \Vert \sigma ^n\mathbf{u} ^{n}\Vert _{L^2}^2 + \Vert \sigma ^n(\mathbf{u} ^{n+1}-\mathbf{u} ^n)\Vert _{L^2}^2 + 2\mu _1\tau \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2}^2 \\&\quad + \tau \displaystyle \int _\Omega \rho ^{n+1}\mathbf{u} ^n\cdot \nabla |\mathbf{u} ^{n+1}|^2 dx - \lambda \tau \displaystyle \int _\Omega \nabla \rho ^{n+1}\cdot \nabla |\mathbf{u} ^{n+1}|^2 dx \le 2\tau (\mathbf{f} ^{n+1}, \mathbf{u} ^{n+1}), \end{aligned}$$

where \(\sigma ^{n+1}=\sqrt{\rho ^{n+1}}\). Setting \(\phi =\tau |\mathbf{u} ^{n+1}|^2\) in (2.9) leads to

$$\begin{aligned} \Vert \sigma ^{n+1}\mathbf{u} ^{n+1}\Vert _{L^2}^2 - \Vert \sigma ^n\mathbf{u} ^{n+1}\Vert _{L^2}^2 + \lambda \tau \displaystyle \int _\Omega \nabla \rho ^{n+1}\cdot \nabla |\mathbf{u} ^{n+1}|^2 dx +\tau \displaystyle \int _\Omega (\nabla \rho ^{n+1}\cdot \mathbf{u} ^n)|\mathbf{u} ^{n+1}|^2 dx=0. \end{aligned}$$

Then we obtain

$$\begin{aligned}&\Vert \sigma ^{n+1}\mathbf{u} ^{n+1}\Vert _{L^2}^2 - \Vert \sigma ^n\mathbf{u} ^{n}\Vert _{L^2}^2 + \Vert \sigma ^n(\mathbf{u} ^{n+1}-\mathbf{u} ^n)\Vert _{L^2}^2 + 2\mu _1\tau \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2}^2 \\&\quad = 2\tau (\mathbf{f} ^{n+1}, \mathbf{u} ^{n+1}) \le \mu _1\tau \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2}^2 + \displaystyle \frac{\tau }{\mu _1}\Vert \mathbf{f} ^{n+1}\Vert _{L^2}^2. \end{aligned}$$

Taking the sum gives

$$\begin{aligned} \Vert \sigma ^{n+1}\mathbf{u} ^{n+1}\Vert _{L^2}^2+\mu _1\tau \displaystyle \sum _{i=0}^n\Vert \nabla \mathbf{u} ^{i+1}\Vert _{L^2}^2 \le \Vert \sigma _0\mathbf{u} _0\Vert _{L^2}^2 + \displaystyle \frac{\tau }{\mu _1}\displaystyle \sum _{i=0}^{N-1}\Vert \mathbf{f} ^{i+1}\Vert _{L^2}^2 \end{aligned}$$

for all \(0\le n\le N-1\), where \(\sigma _0=\sqrt{\rho _0}\).

Thus, we get the following stable result of the time discrete scheme (2.72.8).

Theorem 2.2

Under the condition (2.11) and the assumptions on the time step \(\tau \) in Lemma 2.1, the solutions \(\rho ^{n+1}\) and \(\mathbf{u} ^{n+1}\) to (2.7) and (2.8) satisfy the following the discrete energy inequalities:

$$\begin{aligned} \max \limits _{0\le n\le N-1}\left( \Vert \rho ^{n+1}\Vert _{L^2}^2+2\lambda \tau \displaystyle \sum _{i=0}^n\Vert \nabla \rho ^{i+1}\Vert _{L^2}^2\right)\le & {} \Vert \rho _0\Vert _{L^2}^2, \\ \max \limits _{0\le n\le N-1}\left( \Vert \sigma ^{n+1}\mathbf{u} ^{n+1}\Vert _{L^2}^2+\mu _1\tau \displaystyle \sum _{i=0}^n\Vert \nabla \mathbf{u} ^{i+1}\Vert _{L^2}^2\right)\le & {} \Vert \sigma _0\mathbf{u} _0\Vert _{L^2}^2 + \displaystyle \frac{\tau }{\mu _1}\displaystyle \sum _{i=0}^{N-1}\Vert \mathbf{f} ^{i+1}\Vert _{L^2}^2, \end{aligned}$$

Remark 2.3

From the temporal error analysis in next section, we can see that the condition (2.11) holds for any \(0\le n\le N\). Thus, the above energy inequalities imply that the time discrete scheme (2.72.8) is unconditionally stable.

2.3 Finite Element Scheme

We give the finite element fully discretization of (2.72.8). Let \(\mathcal {T}_h = \{K_j\}_{j=1}^L\) be a quasi-uniform tetrahedral partition of \(\Omega \) with the mesh size \(h =\max _{1\le j\le L} \, \{diam \, K_j\}\). When \(\partial \Omega \) has a smooth curve, the element \(K_j\) adjacent to the boundary may represent a curved tetrahedron with a curved face. The definitions of finite element spaces on such a partition with curved elements can be dealt with that in [13, 25]. We use the mini element (\(P_1b-P_1\)) to approximate the velocity field \(\mathbf{u} \) and the pressure p, and use the piecewise quadratic Lagrange element (\(P_2\)) to approximate the density \(\rho \). The finite element spaces of \(\mathbf{V} , M\) and W are denoted by \(\mathbf{V} _h,\) \(M_h\) and \(W_h\), respectively. For this choice, the finite element spaces \(\mathbf{V} _h\) and \(M_h\) satisfy the discrete inf-sup condition. Further, we define the \(\mathbf{H} (\text{ div }, \Omega )\) conforming Raviart-Thomas finite element spaces of order 1 by

$$\begin{aligned} \mathbf{RT} _h= & {} \{ \mathbf{u} _h\in \mathbf{H} (\text{ div }, \Omega ), \ \mathbf{u} _h|_K\in P_1(K)^3 + x P_1(K), \ \forall \ K\in \mathcal T_h \},\\ \mathbf{RT} _{0h}= & {} \{ \mathbf{u} _h\in \mathbf{RT} _h, \ \nabla \cdot \mathbf{u} _h=0 \ \text{ in } \ \Omega \ \text{ and } \ \mathbf{u} _h\cdot \mathbf{n} =0 \ \text{ on } \ \partial \Omega \}. \end{aligned}$$

We denote by \(\mathbf{P} _{0h}\) the \(L^2\)-orthogonal projection operator from \(\mathbf{L} ^2(\Omega )\) to \(\mathbf{RT} _{0h}\) defined by

$$\begin{aligned} (\mathbf{u} -\mathbf{P} _{0h}\mathbf{u} , \mathbf{v} _h)=0,\qquad \forall \ \mathbf{v} _h\in \mathbf{RT} _{0h}, \ \mathbf{u} \in \mathbf{L} ^2(\Omega ). \end{aligned}$$

Start with \(\mathbf{u} _h^0=I_h\mathbf{u} _0\) and \(\rho _h^0=J_h\rho _0\), where \(I_h\) and \(J_h\) are the interpolation operators from \(\mathbf{V} \rightarrow \mathbf{V} _h\) and \(W \rightarrow W_h\), respectively, and satisfy

$$\begin{aligned} \Vert \mathbf{u} _0-\mathbf{u} _h^0\Vert _{L^2}+h\Vert \nabla (\mathbf{u} _0-\mathbf{u} _h^0)\Vert _{L^2}\le & {} Ch^{2}\Vert \mathbf{u} _0\Vert _{H^2}, \end{aligned}$$
(2.15)
$$\begin{aligned} \Vert \rho _0-\rho _h^0\Vert _{L^2}+h\Vert \rho _0-\rho _h^0\Vert _{H^1}\le & {} Ch^{2}\Vert \rho _0\Vert _{H^2}. \end{aligned}$$
(2.16)

For \(1\le n\le N\), the finite element fully discrete approximations of (2.72.8) are described as follows.

Finite element fully discrete scheme:

Step I: For given \(\rho _h^n\in W_h\) and \(\mathbf{u} _h^n\in \mathbf{V} _h\), we find \(\rho ^{n+1}_h\in W_h\) such that

$$\begin{aligned} (D_\tau \rho _h^{n+1}, r_h) +\lambda (\nabla \rho _h^{n+1}, \nabla r_h) + (\nabla \rho _h^{n+1}\cdot \mathbf{P} _{0h} \mathbf{u} _h^n, r_h) = 0 \end{aligned}$$
(2.17)

for all \(r_h\in W_h\).

Step II: For given \(\rho _h^n\in W_h\), \(\mathbf{u} _h^n\in \mathbf{V} _h\) and \(\rho _h^{n+1}\in W_h\) derived from (2.17), we find \((\mathbf{u} _h^{n+1}, p_h^{n+1})\in \mathbf{V} _h\times M_h\) such that

$$\begin{aligned}&(\rho _h^{n} D_\tau \mathbf{u} _h^{n+1}, \mathbf{v} _h) + a(\rho _h^{n+1}; \mathbf{u} _h^{n+1}, \mathbf{v} _h) - (\nabla \cdot \mathbf{v} _h, p_h^{n+1}) + (\nabla \cdot \mathbf{u} _h^{n+1}, q_h) \nonumber \\&\quad + (\rho _h^{n+1}(\mathbf{u} _h^n\cdot \nabla )\mathbf{u} _h^{n+1}, \mathbf{v} _h) +\displaystyle \frac{1}{2}(D_\tau \rho _h^{n+1}, \mathbf{u} _h^{n+1}\cdot \mathbf{v} _h) + \displaystyle \frac{1}{2} (\nabla \cdot (\rho _h^{n+1} \mathbf{u} _h^{n}), \mathbf{u} _h^{n+1}\cdot \mathbf{v} _h)\nonumber \\&\quad + \displaystyle \frac{\lambda }{2}(\nabla \rho _h^{n+1}, \nabla (\mathbf{u} _h^{n+1}\cdot \mathbf{v} _h)) - \lambda ( (\nabla \rho _h^{n+1}\cdot \nabla )\mathbf{u} _h^{n+1}, \mathbf{v} _h) = (\mathbf{f} ^{n+1}, \mathbf{v} _h) \end{aligned}$$
(2.18)

for all \((\mathbf{v} _h, q_h)\in \mathbf{V} _h\times M_h\).

Remark 2.4

In the above algorithm, the post-processed velocity \(\mathbf{P} _{0h}\mathbf{u} _h^n\) in (2.17) and the stabilized terms \( (D_\tau \rho _h^{n+1} , \mathbf{u} _h^{n+1}\cdot \mathbf{v} _h ) + (\nabla \cdot (\rho _h^n \mathbf{u} _h^{n+1}), \mathbf{u} _h^{n+1}\cdot \mathbf{v} _h) + \lambda (\nabla \rho _h^{n+1}, \nabla (\mathbf{u} _h^{n+1}\cdot \mathbf{v} _h))\) in (2.18) are used to preserve the unconditional stability of numerical scheme.

Taking \(r_h=2\tau \rho _h^{n+1}\) in (2.17), we get

$$\begin{aligned} \Vert \rho _h^{n+1}\Vert _{L^2}^2 - \Vert \rho _h^{n}\Vert _{L^2}^2 + \Vert \rho _h^{n+1}-\rho _h^n\Vert _{L^2}^2 + 2\lambda \tau \Vert \nabla \rho _h^{n+1}\Vert _{L^2}^2=0 \end{aligned}$$
(2.19)

by using

$$\begin{aligned} 2 (\nabla \rho _h^{n+1}\cdot \mathbf{P} _{0h} \mathbf{u} _h^n, \rho _h^{n+1}) = \displaystyle \int _\Omega \mathbf{P} _{0h} \mathbf{u} _h^n\cdot \nabla |\rho _h^{n+1}|^2 dx =-\displaystyle \int _\Omega \nabla \cdot (\mathbf{P} _{0h} \mathbf{u} _h^n) |\rho _h^{n+1}|^2 dx=0. \end{aligned}$$

Taking the sum of (2.19) gives

$$\begin{aligned} \Vert \rho _h^{n+1}\Vert _{L^2}^2+2\lambda \tau \displaystyle \sum _{i=0}^n\Vert \nabla \rho _h^{i+1}\Vert _{L^2}^2 \le \Vert \rho _{h}^0\Vert _{L^2}^2 \end{aligned}$$

for all \(0\le n\le N-1\).

Suppose that the following condition holds:

$$\begin{aligned} \widetilde{m}< \rho _h^{n+1}(x) < \widetilde{M},\qquad \forall \ 0\le n\le N-1. \end{aligned}$$
(2.20)

Taking \((\mathbf{v} _h, q_h)=2\tau (\mathbf{u} _h^{n+1}, p_h^{n+1})\) in (2.18) and using (2.2), we have

$$\begin{aligned} \Vert \sigma _h^{n+1}\mathbf{u} _h^{n+1}\Vert _{L^2}^2 - \Vert \sigma _h^{n}\mathbf{u} _h^{n}\Vert _{L^2}^2 + \Vert \sigma _h^{n}(\mathbf{u} _h^{n+1}-\mathbf{u} _h^n)\Vert _{L^2}^2 + 2\mu _1\tau \Vert \nabla \mathbf{u} _h^{n+1}\Vert _{L^2}^2 \le 2\tau (\mathbf{f} ^{n+1}, \mathbf{u} _h^{n+1}) \end{aligned}$$
(2.21)

by using

$$\begin{aligned} 2(\rho _h^{n+1}(\mathbf{u} _h^n\cdot \nabla )\mathbf{u} _h^{n+1}, \mathbf{u} _h^{n+1}) =\displaystyle \int _\Omega \rho _h^{n+1}\mathbf{u} _h^n \cdot \nabla |\mathbf{u} _h^{n+1}|^2 dx =-\displaystyle \int _\Omega \nabla \cdot (\rho _h^{n+1}\mathbf{u} _h^n)|\mathbf{u} _h^{n+1}|^2 dx, \end{aligned}$$

where \(\sigma _h^{n+1}=\sqrt{\rho _h^{n+1}}\). Taking the sum of (2.21), we can get

$$\begin{aligned} \Vert \sigma _h^{n+1}\mathbf{u} _h^{n+1}\Vert _{L^2}^2+\mu _1\tau \displaystyle \sum _{i=0}^n\Vert \nabla \mathbf{u} _h^{i+1}\Vert _{L^2}^2 \le \Vert \sigma _h^0\mathbf{u} _h^0\Vert _{L^2}^2 + \displaystyle \frac{\tau }{\mu _1}\displaystyle \sum _{i=0}^{N-1}\Vert \mathbf{f} ^{i+1}\Vert _{L^2}^2 \end{aligned}$$

for all \(0\le n\le N-1\), where \(\sigma _h^0=\sqrt{\rho _h^0}\).

Like that for the time discrete scheme (2.72.8), we get the following stable result of the fully discrete scheme (2.172.18).

Theorem 2.3

Under the condition (2.20), the solutions \(\rho _h^{n+1}\in W_h\) and \(\mathbf{u} _h^{n+1}\in \mathbf{V} _h\) to (2.17) and (2.18) satisfy the following the discrete energy inequalities:

$$\begin{aligned} \max \limits _{0\le n\le N-1}\left( \Vert \rho _h^{n+1}\Vert _{L^2}^2+2\lambda \tau \displaystyle \sum _{i=0}^n\Vert \nabla \rho _h^{i+1}\Vert _{L^2}^2\right)\le & {} \Vert \rho _h^0\Vert _{L^2}^2, \\ \max \limits _{0\le n\le N-1}\left( \Vert \sigma _h^{n+1}\mathbf{u} _h^{n+1}\Vert _{L^2}^2+\mu _1\tau \displaystyle \sum _{i=0}^n\Vert \nabla \mathbf{u} _h^{i+1}\Vert _{L^2}^2\right)\le & {} \Vert \sigma _h^0\mathbf{u} _h^0\Vert _{L^2}^2 + \displaystyle \frac{\tau }{\mu _1}\displaystyle \sum _{i=0}^{N-1}\Vert \mathbf{f} ^{i+1}\Vert _{L^2}^2. \end{aligned}$$

Remark 2.5

From the temporal-spatial error analysis in next section, we can see that the condition (2.20) holds for sufficiently small h and \(\tau \). Thus, the above energy inequalities imply that the fully discrete scheme (2.172.18) is unconditionally stable. Furthermore, the discrete energy inequalities show the existence and uniqueness of solutions \(\rho _h^{n+1}\in W_h\) and \(\mathbf{u} _h^{n+1}\in \mathbf{V} _h\) when h and \(\tau \) are sufficiently small.

2.4 Main Result

We present the optimal error estimate in the following theorem. The proof will be given in Section 3. In the rest of this paper, we denote by C a generic positive constant, which is independent of n, h and \(\tau \), and C may be different at different places.

Theorem 2.4

Under the assumptions (A1)-(A4), there exist \(\tau _0>0\) and \(h_0>0\) such that when \(\tau <\tau _0\) and \(h<h_0\), the FE solutions \(\rho _h^{n+1}\) and \(\mathbf{u} _h^{n+1}\) to (2.17) and (2.18) satisfy

$$\begin{aligned} \max \limits _{0\le n\le N-1}\left( \Vert \mathbf{u} (t_{n+1})-\mathbf{u} _h^{n+1}\Vert _{L^2} + \Vert \rho (t_{n+1})-\rho _h^{n+1}\Vert _{H^1} \right) \le C (\tau + h^2). \end{aligned}$$
(2.22)

In the proof of Theorem 2.4, the following inverse inequalities and interpolation inequalities are frequently used (cf. [5]):

$$\begin{aligned} \Vert \mathbf{u} _h\Vert _{L^\infty }\le C h^{-3/2}\Vert \mathbf{u} _h\Vert _{L^2} \quad \text{ and }\quad \Vert \rho _h\Vert _{L^\infty }\le C h^{-3/2}\Vert \rho _h\Vert _{L^2} \end{aligned}$$
(2.23)

for any \(\mathbf{u} _h\in \mathbf{V} _h\) and \(\rho _h\in W_h\), and

$$\begin{aligned} \Vert u \Vert _{L^3} \le C \Vert u\Vert _{L^2}^{1/2}\Vert u\Vert _{H^1}^{1/2} \quad \text{ and } \quad \Vert u \Vert _{L^4} \le C \Vert u\Vert _{L^2}^{1/4}\Vert u\Vert _{H^1}^{3/4}, \quad \forall \ u\in H^1(\Omega ). \end{aligned}$$
(2.24)

Finally, we recall the discrete Gronwall’s inequality established in [20].

Lemma 2.2

Let \(a_k, b_k\) and \(\gamma _k\) be the nonnegative numbers such that

$$\begin{aligned} a_n+\tau \displaystyle \sum \limits _{k=0}^n b_k\le \tau \displaystyle \sum \limits _{k=0}^n \gamma _ka_k+ B, \quad \text{ for } \ n\ge 1. \end{aligned}$$
(2.25)

Suppose \(\tau \gamma _k<1\) and set \(\sigma _k=(1-\tau \gamma _k)^{-1}\). Then there holds:

$$\begin{aligned} a_n+\tau \displaystyle \sum \limits _{k=0}^n b_k\le \exp \left( \tau \displaystyle \sum \limits _{k=0}^n \gamma _k \sigma _k\right) B, \quad \text{ for } \ n\ge 1 \end{aligned}$$
(2.26)

Remark 2.6

If the sum on the right-hand side of (2.25) extends only up to \(n-1\), then the estimate (2.26) still holds for all \(k \ge 1\) with \(\sigma _k=1\).

3 Error Analysis

In this section, we will prove Theorem 2.4 by using the technique of error splitting. We first prove temporal errors in Sect. 3.1 and then prove spatial errors in Sect. 3.2. The finite element error estimates can be derived by combining temporal errors, projection errors and spatial errors.

3.1 Temporal Error Analysis

In this subsection, we will prove the optimal temporal errors. For \(0\le n\le N-1\), we take \(t=t_{n+1}\) in (1.61.8) to deduce that

$$\begin{aligned} D_\tau \rho (t_{n+1})-\lambda \Delta \rho (t_{n+1}) + \nabla \rho (t_{n+1})\cdot \mathbf{u} (t_n)=R_\rho ^{n+1} \end{aligned}$$
(3.1)

and

$$\begin{aligned}&\rho (t_{n})D_\tau \mathbf{u} (t_{n+1}) -\nabla \cdot \left( \mu \nabla \mathbf{u} (t_{n+1})-\lambda \rho (t_{n+1})(\nabla \mathbf{u} (t_{n+1}))^t \right) +\nabla p(t_{n+1}) \nonumber \\&\quad +\rho (t_{n+1}) (\mathbf{u} (t_n)\cdot \nabla )\mathbf{u} (t_{n+1}) -\lambda (\nabla \rho (t_{n+1})\cdot \nabla )\mathbf{u} (t_{n+1}) =\mathbf{f} ^{n+1}+R_u^{n+1} , \end{aligned}$$
(3.2)

where the truncation functions \(R_\sigma ^{n+1}\) and \(R_u^{n+1}\) are given by

$$\begin{aligned} R_\rho ^{n+1}= & {} D_\tau \rho (t_{n+1})-\rho _t(t_{n+1})- \nabla \rho (t_{n+1})\cdot \left( \displaystyle \int _{t_n}^{t_{n+1}}\mathbf{u} _t(t)dt \right) ,\\ R_u^{n+1}= & {} (\rho (t_{n})-\rho (t_{n+1}) ) D_\tau \mathbf{u} (t_{n+1}) + \rho (t_{n+1}) \left( D_\tau \mathbf{u} (t_{n+1}) - \mathbf{u} _t(t_{n+1})\right) \\&-\rho (t_{n+1}) \left( \displaystyle \int _{t_n}^{t_{n+1}}\mathbf{u} _t(t)dt\cdot \nabla \right) \mathbf{u} (t_{n+1}). \end{aligned}$$

Under the regularity assumption (A3), we have

$$\begin{aligned} \tau \displaystyle \sum \limits _{n=0}^{N-1}\left( \Vert R_\rho ^{n+1}\Vert _{L^2}^2 + \Vert R_u^{n+1}\Vert _{L^2}^2\right) \le C\tau ^2. \end{aligned}$$
(3.3)

For \(0\le n\le N\), we introduce temporal error functions by

$$\begin{aligned} \eta ^n=\rho (t_n)-\rho ^n,\quad \mathbf{e} ^n=\mathbf{u} (t_n)-\mathbf{u} ^n,\quad \epsilon ^n=p(t_n)-p^n. \end{aligned}$$

Then error equations satisfied by \((\eta ^{n+1}, \mathbf{e} ^{n+1}, \epsilon ^{n+1})\) with \(0\le n\le N-1\) are

$$\begin{aligned} D_\tau \eta ^{n+1} -\lambda \Delta \eta ^{n+1} + \nabla \rho (t_{n+1})\cdot \mathbf{e} ^n + \nabla \eta ^{n+1}\cdot \mathbf{u} ^n = R_\rho ^{n+1}, \end{aligned}$$
(3.4)

and

$$\begin{aligned} \rho ^n D_\tau \mathbf{e} ^{n+1} -\nabla \cdot \left( \mu \nabla \mathbf{e} ^{n+1} -\lambda \rho ^{n+1}(\nabla \mathbf{e} ^{n+1})^t \right) +\nabla \epsilon ^{n+1} + \displaystyle \sum _{i=1}^7 I_i^{n+1} =R_u^{n+1} \end{aligned}$$
(3.5)

with \(\nabla \cdot \mathbf{e} ^{n+1}=0\) in \(\Omega \), where

$$\begin{aligned} I_1^{n+1}= & {} -\lambda (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{e} ^{n+1}, \\ I_2^{n+1}= & {} \rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{e} ^{n+1},\\ I_3^{n+1}= & {} \rho ^{n+1}(\mathbf{e} ^n\cdot \nabla )\mathbf{u} (t_{n+1}),\\ I_4^{n+1}= & {} \eta ^n D_\tau \mathbf{u} (t_{n+1}) ,\\ I_5^{n+1}= & {} \lambda \nabla \cdot (\eta ^{n+1}(\nabla \mathbf{u} (t_{n+1}) )^t),\\ I_6^{n+1}= & {} - \lambda (\nabla \eta ^{n+1}\cdot \nabla )\mathbf{u} (t_{n+1}),\\ I_7^{n+1}= & {} \eta ^{n+1} (\mathbf{u} (t_n)\cdot \nabla )\mathbf{u} (t_{n+1}). \end{aligned}$$

Moreover, the weak formulations of (3.4) and (3.5) can be described as: find \(\eta ^{n+1}\in W \) such that

$$\begin{aligned} \left( D_\tau \eta ^{n+1}, r \right) +\lambda (\nabla \eta ^{n+1}, \nabla r) + \left( \nabla \rho (t_{n+1})\cdot \mathbf{e} ^n, r\right) + \left( \nabla \eta ^{n+1}\cdot \mathbf{u} ^n, r\right) = ( R_\rho ^{n+1}, r) \end{aligned}$$
(3.6)

for all \(r\in W\), and find \((\mathbf{e} ^{n+1}, \epsilon ^{n+1})\in \mathbf{V} \times M\) such that

$$\begin{aligned} \left( \rho ^n D_\tau \mathbf{e} ^{n+1} , \mathbf{v} \right) + a(\rho ^{n+1}; \mathbf{e} ^{n+1}, \mathbf{v} ) - (\nabla \cdot \mathbf{v} , \epsilon ^{n+1}) + (\nabla \cdot \mathbf{e} ^{n+1}, q) + \displaystyle \sum _{i=1}^7 (I_i^{n+1}, \mathbf{v} ) = ( R_u^{n+1}, \mathbf{v} ) \end{aligned}$$
(3.7)

for all \((\mathbf{v} , q)\in \mathbf{V} \times M\).

We estimate \(\eta ^{n+1}\) and \(\mathbf{e} ^{n+1}\) in \(l^\infty (L^2)\)-norm and \(l^2(H^1)\)-norm in the following two lemmas.

Lemma 3.1

Under the regularity assumption (A3), there exists some \(C>0\) such that

$$\begin{aligned} \Vert \eta ^{m+1}\Vert _{L^2}^2+\displaystyle \sum _{n=0}^m\Vert \eta ^{n+1}-\eta ^n\Vert _{L^2}^2 + \lambda \tau \displaystyle \sum _{n=0}^m\Vert \nabla \eta ^{n+1}\Vert _{L^2}^2 \le C\left( \tau ^2+ \tau \displaystyle \sum _{n=0}^m\Vert \mathbf{e} ^{n}\Vert _{L^2}^2 \right) \end{aligned}$$
(3.8)

for all \(0\le m \le N-1\).

Proof

Taking \(r=2\tau \eta ^{n+1}\) in (3.6) and using

$$\begin{aligned} \displaystyle \int _\Omega (\nabla \eta ^{n+1}\cdot \mathbf{u} ^n)\eta ^{n+1} dx = -\frac{1}{2} \displaystyle \int _\Omega |\eta ^{n+1}|^2\nabla \cdot \mathbf{u} ^n dx + \displaystyle \frac{1}{2} \displaystyle \int _{\partial \Omega } |\eta ^{n+1}|^2 \mathbf{u} ^n\cdot \mathbf{n} ds =0, \end{aligned}$$

it is easy to see that

$$\begin{aligned}&\Vert \eta ^{n+1}\Vert _{L^2}^2 - \Vert \eta ^{n}\Vert _{L^2}^2 + \Vert \eta ^{n+1}-\eta ^n\Vert _{L^2}^2 +2 \lambda \tau \Vert \nabla \eta ^{n+1}\Vert _{L^2}^2 \\&\quad \le \displaystyle \frac{\tau }{2}\Vert \eta ^{n+1}\Vert _{L^2}^2 + C\tau \left( \Vert \mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert R_\rho ^{n+1}\Vert _{L^2}^2 \right) . \end{aligned}$$

Summing up the above estimate for n from 0 to m and using (3.3) and the discrete Gronwall’s inequality in Lemma 2.2, we complete the proof of (3.8). \(\square \)

Lemma 3.2

Under the assumptions (A2) and (A3), there exists some small \(\tau _1>0\) such that when \(\tau <\tau _1\), there holds

$$\begin{aligned} \Vert \sigma ^{m+1}\mathbf{e} ^{m+1}\Vert _{L^2}^2+\displaystyle \sum _{n=0}^m \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 + \tau \displaystyle \sum _{n=0}^m\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 \le C \tau ^2 \end{aligned}$$
(3.9)

for all \(0\le m \le N-1\).

Proof

Setting \( (\mathbf{v} , q)=2\tau (\mathbf{e} ^{n+1}, \epsilon ^{n+1})\) in (3.7), we have

$$\begin{aligned}&\Vert \sigma ^n\mathbf{e} ^{n+1}\Vert _{L^2}^2 - \Vert \sigma ^n\mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 + 2 a(\rho ^{n+1}; \mathbf{e} ^{n+1}, \mathbf{e} ^{n+1}) \\&\quad - \lambda \tau (\nabla \rho ^{n+1}, \nabla |\mathbf{e} ^{n+1}|^2) + \tau (\rho ^{n+1}\mathbf{u} ^n, \nabla |\mathbf{e} ^{n+1}|^2) + 2\tau \displaystyle \sum _{i=3}^7 (I_i^{n+1}, \mathbf{e} ^{n+1}) = 2\tau ( R_u^{n+1}, \mathbf{e} ^{n+1}). \end{aligned}$$

Multiplying (2.7) by \(\tau |\mathbf{e} ^{n+1}|^2\) and integrating over \(\Omega \), one has

$$\begin{aligned} \Vert \sigma ^{n+1}\mathbf{e} ^{n+1}\Vert _{L^2}^2-\Vert \sigma ^n\mathbf{e} ^{n+1}\Vert _{L^2}^2 + \lambda \tau (\nabla \rho ^{n+1}, \nabla |\mathbf{e} ^{n+1}|^2) - \tau (\rho ^{n+1}\mathbf{u} ^n, \nabla |\mathbf{e} ^{n+1}|^2)=0 \end{aligned}$$

where the integration by parts is used. Taking the sum of the above formulations, we get

$$\begin{aligned}&\Vert \sigma ^{n+1}\mathbf{e} ^{n+1}\Vert _{L^2}^2 - \Vert \sigma ^n\mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 + 2 a(\rho ^{n+1}; \mathbf{e} ^{n+1}, \mathbf{e} ^{n+1}) \nonumber \\&\quad + 2\tau \displaystyle \sum _{i=3}^7 (I_i^{n+1}, \mathbf{e} ^{n+1}) = 2\tau ( R_u^{n+1}, \mathbf{e} ^{n+1}). \end{aligned}$$
(3.10)

Now, we suppose that

$$\begin{aligned} \Vert \nabla \mathbf{u} ^n\Vert _{L^2}\le 1+\Vert \nabla \mathbf{u} _0\Vert _{L^2}+\Vert \nabla \mathbf{u} \Vert _{L^\infty (0,T; L^2)}:= \kappa _1, \quad \forall \ 0\le n\le N-1. \end{aligned}$$
(3.11)

According to Lemma 2.1, we have

$$\begin{aligned} \widetilde{m}< m\le \rho ^{n+1}(x)\le M < \widetilde{M}, \qquad \forall \ 0\le n\le N-1, \end{aligned}$$
(3.12)

which with (2.2) and (3.10) gives

$$\begin{aligned}&\Vert \sigma ^{n+1}\mathbf{e} ^{n+1}\Vert _{L^2}^2 - \Vert \sigma ^n\mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 + 2\mu _1\tau \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2\nonumber \\&\quad \le 2\tau \left| ( R_u^{n+1}, \mathbf{e} ^{n+1}) \right| + 2\tau \left| \displaystyle \sum _{i=3}^7 (I_i^{n+1}, \mathbf{e} ^{n+1}) \right| . \end{aligned}$$
(3.13)

To close the mathematical induction (3.11), we need to prove that

$$\begin{aligned} \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2}\le \kappa _1, \quad \quad \forall \ 0\le n\le N-1. \end{aligned}$$
(3.14)

The right-hand side of (3.13) can be estimated term by term as follows. It is easy to see that

$$\begin{aligned} 2\tau \left| ( R_u^{n+1}, \mathbf{e} ^{n+1}) \right| \le \tau \left( \Vert R_u^{n+1}\Vert _{L^2}^2 + \Vert \mathbf{e} ^{n+1}\Vert _{L^2}^2 \right) . \end{aligned}$$

By (A3) and (3.12), we have

$$\begin{aligned} 2\tau \left| (I_3^{n+1}, \mathbf{e} ^{n+1}) \right|\le & {} C\tau \Vert \rho ^{n+1}\Vert _{L^\infty } \Vert \mathbf{e} ^n\Vert _{L^2}\Vert \nabla \mathbf{u} (t_{n+1})\Vert _{L^3}\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}\\\le & {} \displaystyle \frac{\mu _1\tau }{5} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 + C\tau \Vert \mathbf{e} ^{n}\Vert _{L^2}^2, \\ 2\tau \left| (I_4^{n+1}, \mathbf{e} ^{n+1}) \right|\le & {} C\tau \Vert \eta ^{n}\Vert _{L^2} \Vert D_\tau \mathbf{u} (t_{n+1})\Vert _{L^3}\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{5} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 + C\tau \Vert \eta ^{n}\Vert _{L^2}^2, \\ 2\tau \left| (I_5^{n+1}, \mathbf{e} ^{n+1}) \right|\le & {} C\tau \Vert \eta ^{n+1}\Vert _{L^6} \Vert \nabla \mathbf{u} (t_{n+1})\Vert _{L^3} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{5} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2+ C\tau \Vert \eta ^{n+1}\Vert _{L^2}^2 + C\tau \Vert \nabla \eta ^{n+1}\Vert _{L^2}^2, \\ 2\tau \left| (I_6^{n+1}, \mathbf{e} ^{n+1}) \right|\le & {} C\tau \Vert \nabla \eta ^{n+1}\Vert _{L^2} \Vert \nabla \mathbf{u} (t_{n+1})\Vert _{L^3}\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{5} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 + C\tau \Vert \nabla \eta ^{n+1}\Vert _{L^2}^2, \end{aligned}$$

and

$$\begin{aligned} 2\tau \left| (I_7^{n+1}, \mathbf{e} ^{n+1}) \right|\le & {} C\tau \Vert \eta ^{n+1}\Vert _{L^6}\Vert \mathbf{u} (t_{n})\Vert _{L^\infty } \Vert \nabla \mathbf{u} (t_{n+1})\Vert _{L^3} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{5} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2+ C\tau \Vert \eta ^{n+1}\Vert _{L^2}^2 + C\tau \Vert \nabla \eta ^{n+1}\Vert _{L^2}^2, \end{aligned}$$

where the Hölder inequality and the Young inequality are used. Taking into account the above estimates, we get from (3.13) that

$$\begin{aligned}&\Vert \sigma ^{n+1}\mathbf{e} ^{n+1}\Vert _{L^2}^2 - \Vert \sigma ^n\mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 + \mu _1\tau \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 \nonumber \\&\quad \le C\tau \left( \Vert R_u^{n+1}\Vert _{L^2}^2 + \Vert \mathbf{e} ^{n+1}\Vert _{L^2}^2 + \Vert \mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert \eta ^{n+1}\Vert _{L^2}^2 + \Vert \eta ^{n}\Vert _{L^2}^2 + \Vert \nabla \eta ^{n+1}\Vert _{L^2}^2\right) \nonumber \\&\quad \le C\tau \left( \Vert R_u^{n+1}\Vert _{L^2}^2 + \Vert \sigma ^{n+1}\mathbf{e} ^{n+1}\Vert _{L^2}^2 + \Vert \sigma ^n\mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert \nabla \eta ^{n+1}\Vert _{L^2}^2\right) \nonumber \\&\qquad + C\tau ^3 + C\tau ^2 \displaystyle \sum _{n=0}^m\Vert \sigma ^n\mathbf{e} ^{n}\Vert _{L^2}^2 , \end{aligned}$$
(3.15)

where (3.8) in Lemma 3.1 and (3.12) are used. Summing up (3.15) for n from 0 to m, using (3.3), (3.8) and the discrete Gronwall’s inequality in Lemma 2.2, we derive (3.9) and complete the mathematical induction (3.14) by taking a sufficiently small \(\tau _1\) such that

$$\begin{aligned} \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2}\le & {} \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2} + \Vert \nabla \mathbf{u} (t_{n+1})\Vert _{L^2} \le \Vert \nabla \mathbf{u} \Vert _{L^\infty (0,T; L^2)} + (C\tau _1)^{1/2} \\\le & {} \Vert \nabla \mathbf{u} \Vert _{L^\infty (0,T; L^2)}+1\le \kappa _1. \end{aligned}$$

\(\square \)

From the proof of Lemma 3.2, we can see that (3.14) holds for all \(0\le n\le N-1\). It follows from Lemma 2.1 that the solutions \(\rho ^{n+1}\) to (2.7) and \(\mathbf{u} ^{n+1}\) to (2.8) satisfy

$$\begin{aligned} \widetilde{m}< m\le \rho ^{n+1}(x)\le M <\widetilde{M}, \end{aligned}$$
(3.16)
$$\begin{aligned} \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2} + \Vert \rho ^{n+1}\Vert _{H^1}^2+\tau \displaystyle \sum _{i=0}^n\Vert \rho ^{i+1}\Vert _{H^2}^2 \le C \end{aligned}$$
(3.17)

for all \(0\le n\le N-1\). By (3.8) and (3.9), we get the following estimate for the density:

$$\begin{aligned} \Vert \eta ^{m+1}\Vert _{L^2}^2+\displaystyle \sum _{n=0}^m\Vert \eta ^{n+1}-\eta ^n\Vert _{L^2}^2 + \lambda \tau \displaystyle \sum _{n=0}^m\Vert \eta ^{n+1}\Vert _{H^1}^2 \le C\tau ^2 \end{aligned}$$
(3.18)

for \(0\le m \le N-1\). Furthermore, we can estimate \(\eta ^{n+1}\) in \(l^\infty (H^1)\)-norm and \(l^2(H^2)\)-norm as follows.

Lemma 3.3

Under the assumptions (A2) and (A3), when \(\tau <\tau _1\), where \(\tau _1\) is from Lemma 3.2, there exists some \(C>0\) such that

$$\begin{aligned} \Vert \eta ^{m+1}\Vert _{H^1}^2+\displaystyle \sum _{n=0}^m\Vert \eta ^{n+1}-\eta ^n\Vert _{H^1}^2 + \lambda \tau \displaystyle \sum _{n=0}^m\Vert \eta ^{n+1}\Vert _{H^2}^2 \le C\tau ^2 \end{aligned}$$
(3.19)

for all \(0\le m \le N-1\).

Proof

Multiplying (3.4) by \(-2\tau \Delta \eta ^{n+1}\) and integrating over \(\Omega \), we can prove that

$$\begin{aligned}&\Vert \nabla \eta ^{n+1}\Vert _{L^2}^2 - \Vert \nabla \eta ^{n}\Vert _{L^2}^2 + \Vert \nabla (\eta ^{n+1}-\eta ^n)\Vert _{L^2}^2 +2 \lambda \tau \Vert \Delta \eta ^{n+1}\Vert _{L^2}^2 \\&\quad \le C\tau \left( \Vert \mathbf{e} ^{n}\Vert _{L^2} + \Vert R_\rho ^{n+1}\Vert _{L^2} \right) \Vert \Delta \eta ^{n+1}\Vert _{L^2} + C\tau \Vert \nabla \eta ^{n+1}\Vert _{L^2}^{1/2}\Vert \mathbf{u} ^n\Vert _{L^6} \Vert \Delta \eta ^{n+1}\Vert _{L^2}^{3/2} \\&\quad \le \lambda \tau \Vert \Delta \eta ^{n+1}\Vert _{L^2}^2 + C\tau \left( \Vert \mathbf{e} ^{n}\Vert _{L^2} + \Vert R_\rho ^{n+1}\Vert _{L^2} + \Vert \nabla \eta ^{n+1}\Vert _{L^2}^2 \right) \end{aligned}$$

Summing up the above estimate for n from 0 to m and using (3.18), we obtain

$$\begin{aligned} \Vert \nabla \eta ^{m+1}\Vert _{L^2}^2+\displaystyle \sum _{n=0}^m\Vert \nabla (\eta ^{n+1}-\eta ^n)\Vert _{L^2}^2 + \lambda \tau \displaystyle \sum _{n=0}^m\Vert \Delta \eta ^{n+1}\Vert _{L^2}^2 \le C\tau ^2 \end{aligned}$$

for \(0\le m \le N-1\). By noticing (3.18), again, we complete the proof of (3.19). \(\square \)

The error estimate (3.19) provides a uniform boundness of \(\rho ^{n+1}\) in \(H^2\)-norm. That is to say that there exists some \(C>0\) such that

$$\begin{aligned} \Vert \rho ^{n+1}\Vert _{H^2} \le C,\qquad \forall \ 0\le n\le N-1. \end{aligned}$$
(3.20)

Next, we estimate \(\mathbf{u} ^{n+1}\) in \(l^2(\mathbf{H} ^{2})\)-norm under the assumption (A4). To do this, we rewrite (2.8) as the Stokes type problem:

$$\begin{aligned} -\mu \Delta \mathbf{u} ^{n+1} +\nabla p^{n+1} = \mathbf{F} ^{n+1}, \end{aligned}$$
(3.21)

where

$$\begin{aligned} \mathbf{F} ^{n+1} = \mathbf{f} ^{n+1} - \rho ^{n} D_\tau \mathbf{u} ^{n+1} - \lambda \nabla \rho ^{n+1}\cdot (\nabla \mathbf{u} ^{n+1} )^t - \rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{u} ^{n+1} + \lambda (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{u} ^{n+1} \end{aligned}$$

by using \(\nabla \cdot (\nabla \mathbf{u} ^{n+1} )^t=0\) due to \(\nabla \cdot \mathbf{u} ^{n+1}=0\). From (3.16) and

$$\begin{aligned} \Vert \rho ^n(\mathbf{u} ^{n+1}-\mathbf{u} ^n)\Vert _{L^2}^2\le & {} 2\Vert \rho ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+2\Vert \rho ^n(\mathbf{u} (t_{n+1})-\mathbf{u} (t_n))\Vert _{L^2}^2\\\le & {} 2\Vert \sigma ^n\Vert _{L^\infty }^2\Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+2\tau \Vert \rho ^n\Vert _{L^\infty }^2 \displaystyle \int ^{t_{n+1}}_{t_n}\Vert \mathbf{u} _t(t)\Vert _{L^2}^2dt \\\le & {} C \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ C\tau \displaystyle \int ^{t_{n+1}}_{t_n}\Vert \mathbf{u} _t(t)\Vert _{L^2}^2dt, \end{aligned}$$

we have

$$\begin{aligned} \tau \displaystyle \sum \limits _{n=0}^{N-1} \Vert \rho ^{n} D_\tau \mathbf{u} ^{n+1} \Vert _{L^2}^2\le C, \end{aligned}$$
(3.22)

where (3.9) is used. From (3.16), (3.17) and (3.20), one has

$$\begin{aligned}&\Vert \rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{u} ^{n+1}\Vert _{L^{3/2}} +\Vert (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{u} ^{n+1}\Vert _{L^{3/2}} +\Vert \nabla \rho ^{n+1}\cdot (\nabla \mathbf{u} ^{n+1} )^t \Vert _{L^{3/2}} \nonumber \\&\quad \le \Vert \rho ^{n+1}\Vert _{L^\infty }\Vert \mathbf{u} ^n\Vert _{L^6}\Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2} +2\Vert \nabla \rho ^{n+1}\Vert _{L^6}\Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2}\nonumber \\&\quad \le C. \end{aligned}$$
(3.23)

Then (3.22) and (3.23) yield

$$\begin{aligned} \tau \displaystyle \sum \limits _{n=0}^{N-1} \Vert \mathbf{F} ^{n+1} \Vert _{L^{3/2}}^2\le C, \end{aligned}$$

which with the assumption (A4) gives

$$\begin{aligned} \tau \displaystyle \sum \limits _{n=0}^{N-1}\left( \Vert \mathbf{u} ^{n+1}\Vert _{W^{2,3/2}}^2+\Vert p^{n+1}\Vert _{W^{1,3/2}}^2\right) \le C. \end{aligned}$$

From the Sobolev imbedding theorem \(\mathbf{W} ^{2, 3/2}(\Omega )\hookrightarrow \mathbf{W} ^{1, 3}(\Omega )\), we have

$$\begin{aligned} \tau \displaystyle \sum \limits _{n=0}^{N-1} \Vert \mathbf{u} ^{n+1}\Vert _{W^{1,3}}^2 \le C. \end{aligned}$$
(3.24)

By (3.16), (3.17) and (3.20), again, we have

$$\begin{aligned}&\Vert \rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{u} ^{n+1}\Vert _{L^{2}} +\Vert (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{u} ^{n+1}\Vert _{L^{2}} +\Vert \nabla \rho ^{n+1}\cdot (\nabla \mathbf{u} ^{n+1} )^t \Vert _{L^{2}} \nonumber \\&\quad \le \Vert \rho ^{n+1}\Vert _{L^\infty }\Vert \mathbf{u} ^n\Vert _{L^6}\Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^3}+2\Vert \nabla \rho ^{n+1}\Vert _{L^6}\Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^3}\nonumber \\&\quad \le C\Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^3}. \end{aligned}$$
(3.25)

Then (3.22) and (3.25) yield

$$\begin{aligned} \tau \displaystyle \sum \limits _{n=0}^{N-1} \Vert \mathbf{F} ^{n+1} \Vert _{L^{2}}^2\le C. \end{aligned}$$

By the assumption (A4), again, we get

$$\begin{aligned} \tau \displaystyle \sum \limits _{n=0}^{N-1}\left( \Vert \mathbf{u} ^{n+1}\Vert _{H^{2}}^2+\Vert p^{n+1}\Vert _{H^{1}}^2\right) \le C. \end{aligned}$$
(3.26)

Thus, the numerical velocity \(\mathbf{u} ^{n+1}\) is uniformly bound in \(l^2(\mathbf{H} ^2)\)-norm. Based on the regularities (3.20) and (3.26), we can obtain the error estimate of \(\mathbf{e} ^{n+1}\) in \(l^\infty (\mathbf{V} )\)-norm and \(l^2(\mathbf{H} ^2)\)-norm stated in Lemma 3.4. To make this, we rewrite (3.5) as

$$\begin{aligned} \rho ^n D_\tau \mathbf{e} ^{n+1} - \mu \Delta \mathbf{e} ^{n+1} + \lambda \nabla \rho ^{n+1}\cdot (\nabla \mathbf{e} ^{n+1})^t +\nabla \epsilon ^{n+1} + \displaystyle \sum _{i=1}^6 I_i^{n+1} =R_u^{n+1} \end{aligned}$$
(3.27)

with \(\nabla \cdot \mathbf{e} ^{n+1}=0\) in \(\Omega \).

Lemma 3.4

Under the assumptions (A2)-(A4), there exists some \(\tau _2<\tau _1\) such that when \(\tau <\tau _2\), there holds

$$\begin{aligned} \Vert \nabla \mathbf{e} ^{m+1}\Vert _{L^2}^2 + \displaystyle \sum _{n=0}^m \Vert \nabla (\mathbf{e} ^{n+1}-\mathbf{e} ^{n})\Vert _{L^2}^2 + \tau \displaystyle \sum _{n=0}^m \left( \Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2 + \Vert \epsilon ^{n+1}\Vert _{H^1}^2 \right) \le C\tau ^2 \end{aligned}$$
(3.28)

for all \(0\le m \le N-1\).

Proof

Testing (3.27) by \(2\tau (\mathbf{e} ^{n+1}-\mathbf{e} ^n)\) leads to

$$\begin{aligned}&2\Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+\mu \tau \left( \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2- \Vert \nabla \mathbf{e} ^{n}\Vert _{L^2}^2+ \Vert \nabla (\mathbf{e} ^{n+1}-\mathbf{e} ^{n})\Vert _{L^2}^2 \right) \nonumber \\&\quad \le 2\tau \left| \left( R_u^{n+1}, \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| + 2\lambda \tau \left| \left( \nabla \rho ^{n+1}\cdot (\nabla \mathbf{e} ^{n+1})^t, \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| \nonumber \\&\qquad + 2\tau \displaystyle \sum _{i=1}^7 \left| \left( I_i^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| . \end{aligned}$$
(3.29)

The right-hand side of (3.29) can be estimated term by term by using the Hölder inequality and the Young inequality. From (3.16), it is easy to show that

$$\begin{aligned} 2\tau \left| \left( R_u^{n+1}, \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| \le \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+C\tau ^2 \Vert R_u^{n+1}\Vert _{L^2}^2, \end{aligned}$$

and

$$\begin{aligned}&2\lambda \tau \left| \left( \nabla \rho ^{n+1}\cdot (\nabla \mathbf{e} ^{n+1})^t, \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| \\&\quad \le 2\lambda \tau \left| \left( \nabla \eta ^{n+1}\cdot (\nabla \mathbf{e} ^{n+1})^t, \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| + 2\lambda \tau \left| \left( \nabla \rho (t_{n+1})\cdot (\nabla \mathbf{e} ^{n+1})^t, \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| \\&\quad \le \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+C\tau ^2\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 + C\tau ^2 \Vert \nabla \eta ^{n+1}\Vert _{L^3}^2\Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2. \end{aligned}$$

A similar argument gives

$$\begin{aligned} 2\tau \left| \left( I_1^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right| \le \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+C\tau ^2\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 + C\tau ^2 \Vert \nabla \eta ^{n+1}\Vert _{L^3}^2\Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2. \end{aligned}$$

Other terms can be bound, respectively, by

$$\begin{aligned} 2\tau \left| \left( I_2^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right|\le & {} \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ C\tau ^2 \Vert \mathbf{u} ^n\Vert _{H^2}^2\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 \\ 2\tau \left| \left( I_3^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right|\le & {} \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ C\tau ^2 \Vert \nabla \mathbf{e} ^{n}\Vert _{L^2}^2, \\ 2\tau \left| \left( I_4^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right|\le & {} \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ C\tau ^2 \Vert \eta ^n\Vert _{H^2}^2, \\ 2\tau \left| \left( I_5^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right|\le & {} \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ C\tau ^2 \Vert \eta ^{n+1}\Vert _{H^2}^2, \\ 2\tau \left| \left( I_6^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right|\le & {} \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ C\tau ^2\Vert \eta ^{n+1}\Vert _{H^2}^2, \\ 2\tau \left| \left( I_7^{n+1} , \mathbf{e} ^{n+1}-\mathbf{e} ^n\right) \right|\le & {} \displaystyle \frac{1}{9} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ C\tau ^2\Vert \eta ^{n+1}\Vert _{H^2}^2, \end{aligned}$$

where we use the regularity assumption (A3) and (3.16). Substituting the above estimates into (3.29) leads to

$$\begin{aligned}&\Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+\mu \tau \left( \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2- \Vert \nabla \mathbf{e} ^{n}\Vert _{L^2}^2+ \Vert \nabla (\mathbf{e} ^{n+1}-\mathbf{e} ^{n})\Vert _{L^2}^2 \right) \nonumber \\&\quad \le C\tau ^2 \left( \Vert R_u^{n+1}\Vert _{L^2}^2+ \Vert \mathbf{u} ^n\Vert _{H^2}^2\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 + \Vert \eta ^{n}\Vert _{H^2}^2 + \Vert \eta ^{n+1}\Vert _{H^2}^2 \right) \nonumber \\&\qquad + C\tau ^2 \left( \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 +\Vert \nabla \mathbf{e} ^{n}\Vert _{L^2}^2 \right) + C\tau ^2 \Vert \nabla \eta ^{n+1}\Vert _{L^3}^2\Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2. \end{aligned}$$
(3.30)

Summing up (3.30) for n from 0 to m and using (3.9) and (3.19), we obtain

$$\begin{aligned}&\mu \tau \Vert \nabla \mathbf{e} ^{m+1}\Vert _{L^2}^2 + \displaystyle \sum _{n=0}^m \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+\mu \tau \displaystyle \sum _{n=0}^m \Vert \nabla (\mathbf{e} ^{n+1}-\mathbf{e} ^{n})\Vert _{L^2}^2 \\&\quad \le C\tau ^3 + C \tau ^2 \displaystyle \sum _{n=0}^m \Vert \mathbf{u} ^n\Vert _{H^2}^2\Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}^2 + C\tau ^2 \displaystyle \sum _{n=0}^m \Vert \nabla \eta ^{n+1}\Vert _{L^3}^2\Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2. \end{aligned}$$

By (3.26) and the discrete Gronwall’s inequality in Lemma 2.2, we get

$$\begin{aligned}&\tau \Vert \nabla \mathbf{e} ^{m+1}\Vert _{L^2}^2 + \displaystyle \sum _{n=0}^m \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ \tau \displaystyle \sum _{n=0}^m \Vert \nabla (\mathbf{e} ^{n+1}-\mathbf{e} ^{n})\Vert _{L^2}^2 \nonumber \\&\quad \le C\tau ^3 + C \tau ^3 \displaystyle \sum _{n=0}^m \Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2 \end{aligned}$$
(3.31)

by using \( \Vert \nabla \eta ^{n+1}\Vert _{L^3}^2 \le C \Vert \eta ^{n+1}\Vert _{H^2}^2 \le C\tau \). On the other hand, from (3.27) and the regularity assumption (A4) of the solution to the Stokes problem, we have

$$\begin{aligned}&\Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2 + \Vert \epsilon ^{n+1}\Vert _{H^1}^2 \nonumber \\&\quad \le C\Vert \rho ^n D_\tau \mathbf{e} ^{n+1}\Vert _{L^2}^2 + C\Vert R_u^{n+1}\Vert _{L^2}^2 + C\Vert \nabla \rho ^{n+1}\cdot (\nabla \mathbf{e} ^{n+1})^t \Vert _{L^2}^2+ C \displaystyle \sum _{i=1}^6 \Vert I_i^{n+1}\Vert _{L^2}^2 \nonumber \\&\quad \le C \tau ^{-2} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 + C \Vert R_u^{n+1}\Vert _{L^2}^2 + C \Vert \nabla \mathbf{e} ^{n+1}\Vert _{L^2}\Vert \mathbf{e} ^{n+1}\Vert _{H^2} \nonumber \\&\qquad + C \Vert \nabla \mathbf{e} ^{n}\Vert _{L^2}^2 + C\Vert \eta ^{n+1}\Vert _{H^2}^2 + C\Vert \eta ^{n}\Vert _{H^2}^2\nonumber \\&\quad \le \displaystyle \frac{1}{2}\Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2 + C \tau ^{-2} \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 \nonumber \\&\qquad + C\left( \Vert R_u^{n+1}\Vert _{L^2}^2 + \Vert \nabla \mathbf{e} ^{n}\Vert _{L^2}^2 + \Vert \eta ^{n+1}\Vert _{H^2}^2 + \Vert \eta ^{n}\Vert _{H^2}^2 \right) . \end{aligned}$$
(3.32)

Summing up (3.32) for n from 0 to m and using (3.19) and (3.31), we obtain

$$\begin{aligned}&\tau \displaystyle \sum _{n=0}^m \left( \Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2 + \Vert \epsilon ^{n+1}\Vert _{H^1}^2 \right) \le C\tau ^2 + C \tau ^{-1} \displaystyle \sum _{n=0}^m \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2 \nonumber \\&\quad \le C\tau ^2 + C \tau ^2 \displaystyle \sum _{n=0}^m \Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2. \end{aligned}$$
(3.33)

Taking a sufficiently small \(\tau _2<\tau _1\) such that \(C\tau _2<1\), we derive

$$\begin{aligned} \tau \displaystyle \sum _{n=0}^m \left( \Vert \mathbf{e} ^{n+1}\Vert _{H^2}^2 + \Vert \epsilon ^{n+1}\Vert _{H^1}^2 \right) \le C\tau ^2, \end{aligned}$$

which with (3.31) leads to

$$\begin{aligned} \tau \Vert \nabla \mathbf{e} ^{m+1}\Vert _{L^2}^2 + \displaystyle \sum _{n=0}^m \Vert \sigma ^n(\mathbf{e} ^{n+1}-\mathbf{e} ^n)\Vert _{L^2}^2+ \tau \displaystyle \sum _{n=0}^m \Vert \nabla (\mathbf{e} ^{n+1}-\mathbf{e} ^{n})\Vert _{L^2}^2 \le C\tau ^3. \end{aligned}$$

Thus, we complete the proof of Lemma 3.4. \(\square \)

From (3.28), we can see that

$$\begin{aligned} \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^2}\le & {} C\tau + \Vert \nabla \mathbf{u} \Vert _{L^\infty (0,T; L^2)} \\\le & {} 1 +\Vert \nabla \mathbf{u} _0\Vert _{L^2}+\Vert \nabla \mathbf{u} \Vert _{L^\infty (0,T; L^2)}=\kappa _1, \quad \forall \ 0\le n\le N-1 \end{aligned}$$

for some small \(\tau >0\). Thus, (3.14) holds and we close the mathematical induction.

The estimate (3.28) provides a uniform boundness of the time discrete solution \((\mathbf{u} ^{n+1}, p^{n+1)}\) in \( l^\infty (\mathbf{H} ^2)\times l^\infty (H^1)\)-norm, which means that there exists some \(C>0\) such that

$$\begin{aligned} \Vert \mathbf{u} ^{n+1}\Vert _{H^2} + \Vert p^{n+1}\Vert _{H^1} \le C,\qquad \forall \ 0\le n\le N-1. \end{aligned}$$
(3.34)

In addition, the estimates (3.19) and (3.28) imply that

$$\begin{aligned} \Vert \nabla (D_\tau \mathbf{u} ^{n+1})\Vert _{L^2}+ \Vert D_\tau \rho ^{n+1}\Vert _{H^1} +\tau \displaystyle \sum _{i=0}^n\left( \Vert D_\tau \mathbf{u} ^{n+1}\Vert _{H^2}+\Vert D_\tau \rho ^{n+1}\Vert _{H^2} \right) \le C \end{aligned}$$
(3.35)

for all \(0\le n\le N-1\), if we notice the regularity assumption (A3).

Next, we estimate the time discrete solutions \((\rho ^{n+1}, \mathbf{u} ^{n+1}, p^{n+1)}\) in \(H^3\times \mathbf{W} ^{2,4}\times W^{1,4}\)-norm. We turn back to (2.7) and (3.21). In terms of (3.20) and (3.34), one has

$$\begin{aligned} \Vert \nabla \rho ^{n+1}\cdot \mathbf{u} ^n\Vert _{L^2}\le & {} C\Vert \rho ^{n+1}\Vert _{H^2}\Vert \nabla \mathbf{u} ^{n}\Vert _{L^2}\le C,\\ \Vert \nabla (\nabla \rho ^{n+1}\cdot \mathbf{u} ^n)\Vert _{L^2}\le & {} C \Vert \rho ^{n+1}\Vert _{H^2}\Vert \mathbf{u} ^{n}\Vert _{H^2} \le C \end{aligned}$$

and

$$\begin{aligned}&\Vert \rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{u} ^{n+1}\Vert _{L^{4}} +\Vert (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{u} ^{n+1}\Vert _{L^{4}} +\Vert \nabla \rho ^{n+1}\cdot (\nabla \mathbf{u} ^{n+1} )^t \Vert _{L^{4}} \\&\quad \le C\Vert \rho ^{n+1}\Vert _{L^\infty }\Vert \mathbf{u} ^n\Vert _{H^2}\Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^4} +C\Vert \nabla \rho ^{n+1}\Vert _{L^\infty }\Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^4}\\&\quad \le C . \end{aligned}$$

From the regularity assumption (A4), we obtain

$$\begin{aligned} \Vert \rho ^{n+1}\Vert _{H^3}+ \Vert \mathbf{u} ^{n+1}\Vert _{W^{2,4}}+\Vert p^{n+1}\Vert _{W^{1,4}}\le C,\quad \forall \ 0\le n\le N-1. \end{aligned}$$
(3.36)

3.2 Spatial Error Analysis

In this subsection, we will prove the optimal spatial error estimate for the velocity in \(l^\infty (\mathbf{L} ^2)\)-norm and the density in \(l^\infty (H^1)\)-norm. The proof is based on the regularities of time discrete solutions derived in Sect. 3.1 and the following new projection operators.

For \(1\le n\le N\), we introduce three new projection operators \((\mathbf{R} _h^n, Q_h^n):\mathbf{V} \times M\rightarrow \mathbf{V} _h\times M_h\) and \(\Pi _h^n: W\rightarrow W_h\) defined by

$$\begin{aligned} a(\rho ^{n}; \mathbf{R} _h^n\mathbf{u} -\mathbf{u} , \mathbf{v} _h) - (\nabla \cdot \mathbf{v} _h, Q_h^n p-p)&=0, \quad \forall \ \mathbf{v} _h\in \mathbf{V} _h,\\ (\nabla \cdot (\mathbf{R} _h^n\mathbf{u} -\mathbf{u} ), q_h)&=0,\quad \forall \ q_h\in M_h, \end{aligned}$$

and

$$\begin{aligned} \lambda (\Pi _h^n \rho -\rho , r_h) + \lambda (\nabla (\Pi _h^n \rho -\rho ),\nabla r_h)+ (\nabla (\Pi _h^{n}\rho -\rho )\cdot \mathbf{u} ^{n-1}, r_h) =0,\quad \forall \ r_h\in W_h, \end{aligned}$$

where \(\rho ^n\) and \(\mathbf{u} ^{n-1}\) are the solutions to (2.72.8) and satisfy the point-wise inequality (3.16) and the regularity (3.34), respectively.

Then from the coercive property (2.2), and using a classical argument (cf.[5, 14]), the following approximations hold:

$$\begin{aligned} \Vert \mathbf{u} -\mathbf{R} _h^n\mathbf{u} \Vert _{L^2}+h\Vert \nabla (\mathbf{u} -\mathbf{R} _h^n\mathbf{u} )\Vert _{L^2}+h\Vert p-Q_h^np\Vert _{L^2}\le & {} Ch^{2}(\Vert \mathbf{u} \Vert _{H^2}+\Vert p\Vert _{H^1}), \end{aligned}$$
(3.37)
$$\begin{aligned} \Vert \rho -\Pi _h^n\rho \Vert _{L^2}+h\Vert \rho -\Pi _h^n\rho \Vert _{H^1}\le & {} Ch^2\Vert \rho \Vert _{H^2}, \end{aligned}$$
(3.38)
$$\begin{aligned} \Vert \rho -\Pi _h^n\rho \Vert _{L^\infty }+\Vert \mathbf{u} -\mathbf{R} _h^n\mathbf{u} \Vert _{L^\infty }\le & {} Ch^{1/2} \end{aligned}$$
(3.39)

for any \((\rho , \mathbf{u} , p)\in H^2(\Omega )\cap \mathbf{H} ^2(\Omega )\cap \mathbf{V} \times H^1(\Omega )\). Furthermore, one has

$$\begin{aligned} \Vert \rho -\Pi _h^n\rho \Vert _{L^4}+\Vert \mathbf{u} -\mathbf{R} _h^n\mathbf{u} \Vert _{L^4}\le & {} Ch^2( \Vert \rho \Vert _{W^{2,4}} + \Vert \mathbf{u} \Vert _{W^{2,4}}+\Vert p\Vert _{W^{1,4}}), \end{aligned}$$
(3.40)
$$\begin{aligned} \Vert \rho -\Pi _h^n\rho \Vert _{W^{1,4}}+\Vert \mathbf{u} -\mathbf{R} _h^n\mathbf{u} \Vert _{W^{1,4}}\le & {} Ch(\Vert \rho \Vert _{W^{2,4}} + \Vert \mathbf{u} \Vert _{W^{2,4}}+\Vert p\Vert _{W^{1,4}}) \end{aligned}$$
(3.41)

if \((\rho , \mathbf{u} , p)\in W^{2,4}(\Omega )\cap \mathbf{W} ^{2,4}(\Omega )\cap \mathbf{V} \times W^{1,4}(\Omega )\).

We denote by \(\mathbf{P} _{1h}\) the standard Raviart-Thomas projection from \(\mathbf{H} (\text{ div },\Omega )\) onto \(\mathbf{RT} _h\), which satisfies the following properties (cf. [30]):

$$\begin{aligned} (\nabla \cdot \mathbf{P} _{1h}\mathbf{u} , v_h)= & {} (\nabla \cdot \mathbf{u} , v_h),\qquad \forall \ v_h\in P_1(\mathcal {T}_h),\\ \Vert \mathbf{u} -\mathbf{P} _{1h}\mathbf{u} \Vert _{L^2}\le & {} C h^l \Vert \mathbf{u} \Vert _{H^{l}} ,\qquad \forall \ \mathbf{u} \in \mathbf{H} ^l(\Omega ), l=1,2, \end{aligned}$$

where \(P_1(\mathcal {T}_h)\subset H^1(\Omega )\) is the finite element space of functions which are the piecewise linear polynomials on each \(K\in \mathcal T_h\). For the time discrete solution \(\mathbf{u} ^n\), since \(\nabla \cdot \mathbf{u} ^n=0\) in \(\Omega \) and \(\mathbf{u} ^n\cdot \mathbf{n} =0 \) on \(\partial \Omega \), then

$$\begin{aligned} \nabla \cdot \mathbf{P} _{1h}\mathbf{u} ^n=0 \ \text{ in } \ \Omega \quad \text{ and } \quad \mathbf{P} _{1h}\mathbf{u} ^n\cdot \mathbf{n} =0 \ \text{ on } \ \partial \Omega , \end{aligned}$$

which implies that \(\mathbf{P} _{1h}\mathbf{u} ^n\in \mathbf{RT} _{0h}\). By noticing the definition of the \(L^2\)-projection \(\mathbf{P} _{0h}\), there holds that

$$\begin{aligned} \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} ^n\Vert _{L^2} \le \Vert \mathbf{u} ^n-\mathbf{P} _{1h}\mathbf{u} ^n\Vert _{L^2} \le Ch^2. \end{aligned}$$
(3.42)

Introduce spatial error functions by

$$\begin{aligned} \eta _h^0&=J_h \rho _0-\rho _h^0=0,\quad \mathbf{e} _h^0=I_h \mathbf{u} _0-\mathbf{u} ^0_h=0,\\ \eta _h^n&=\Pi _h^n\rho ^n-\rho _h^n,\quad \mathbf{e} _h^n=\mathbf{R} _h^n\mathbf{u} ^n-\mathbf{u} ^n_h, \quad \epsilon _h^n=Q_h^n p^n-p_h^n, \quad \forall \ 1\le n\le N, \end{aligned}$$

where \((\rho _h^n, \mathbf{u} _h^n, p_h^n)\) and \((\rho ^n, \mathbf{u} ^n, p^n)\) are numerical solutions to (2.172.18) and (2.72.8), respectively. Moreover, we denote projection error functions by

$$\begin{aligned} \theta ^0&=J_h\rho _0-\rho _0,\quad \mathbf{E} ^0=I_h\mathbf{u} _0-\mathbf{u} _0,\\ \theta ^n&=\Pi _h^n\rho ^n-\rho ^n,\quad \mathbf{E} ^n=\mathbf{R} _h^n\mathbf{u} ^n-\mathbf{u} ^n, \quad \xi ^n=Q_h^n p^n- p^n, \quad \forall \ 1\le n\le N. \end{aligned}$$

From (3.373.41) and the regularities (3.20), (3.34) and (3.35), projection error functions satisfy

$$\begin{aligned} \Vert \mathbf{E} ^n\Vert _{L^2}+h (\Vert \nabla \mathbf{E} ^n\Vert _{L^2}+\Vert \xi ^n\Vert _{L^2})\le & {} Ch^{2}, \end{aligned}$$
(3.43)
$$\begin{aligned} \Vert \mathbf{E} ^n\Vert _{L^\infty }+\Vert \theta ^n\Vert _{L^\infty }\le & {} Ch^{1/2},\end{aligned}$$
(3.44)
$$\begin{aligned} \Vert D_\tau \mathbf{E} ^{n}\Vert _{L^2} + \Vert D_\tau \theta ^{n}\Vert _{L^2}\le & {} Ch^2 (\Vert D_\tau \mathbf{u} ^n\Vert _{H^2}+\Vert D_\tau \rho ^n\Vert _{H^2}) , \end{aligned}$$
(3.45)
$$\begin{aligned} \Vert \mathbf{E} ^n\Vert _{L^4} + h\Vert \mathbf{E} ^n\Vert _{W^{1,4}}\le & {} Ch^{2} ,\end{aligned}$$
(3.46)
$$\begin{aligned} \Vert \theta ^n\Vert _{L^2}+h \Vert \theta ^n\Vert _{H^1}\le & {} Ch^{3}. \end{aligned}$$
(3.47)

For \(0\le n\le N-1\), subtracting (2.172.18) from (2.92.10) with \((r, \mathbf{v} , q)=(r_h, \mathbf{v} _h, q_h)\) and noticing the definitions of projection operators \((\mathbf{R} _h^{n+1}, Q_h^{n+1})\) and \(\Pi _h^{n+1}\), we get the following error equations satisfied by \(\eta _h^{n+1}\) and \((\mathbf{e} _h^{n+1}, \epsilon _h^{n+1})\), respectively,

$$\begin{aligned}&(D_\tau \eta _h^{n+1}, r_h) + \lambda (\nabla \eta _h^{n+1}, \nabla r_h)\nonumber \\&\quad =(D_\tau \theta ^{n+1}, r_h) - \lambda (\theta ^{n+1}, r_h) -(\nabla \theta ^{n+1}\cdot (\mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n), r_h)- (\nabla \rho ^{n+1}\cdot (\mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n), r_h) \nonumber \\&\qquad + (\nabla \eta _h^{n+1}\cdot (\mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n), r_h) -(\nabla \eta _h^{n+1}\cdot \mathbf{u} ^n, r_h) \nonumber \\&\quad := \displaystyle \sum _{i=1}^{6} (I_{ih}^{n+1}, r_h), \quad \forall \ r_h\in W_h, \end{aligned}$$
(3.48)

and

$$\begin{aligned}&(\rho _h^n D_\tau \mathbf{e} _h^{n+1}, \mathbf{v} _h) + \displaystyle \frac{1}{2}\left( D_\tau \rho _h^{n+1}, \mathbf{e} _h^{n+1}\cdot \mathbf{v} _h\right) + a(\rho _h^{n+1}; \mathbf{e} _h^{n+1}, \mathbf{v} _h) \nonumber \\&\quad - (\nabla \cdot \mathbf{v} _h, \epsilon _h^{n+1}) + (\nabla \cdot \mathbf{e} _h^{n+1}, q_h) \nonumber \\&\quad = (\rho _h^n D_\tau \mathbf{E} ^{n+1}, \mathbf{v} _h) - \left( (\rho ^n-\rho _h^n) D_\tau \mathbf{u} ^{n+1}, \mathbf{v} _h\right) + \displaystyle \frac{1}{2}\left( D_\tau \theta ^{n+1}, \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h\right) \nonumber \\&\qquad +\lambda \left( (\rho ^{n+1}-\rho _h^{n+1}), (\nabla \nabla \mathbf{u} ^{n+1})^t: \nabla \mathbf{v} _h \right) -\displaystyle \frac{1}{2} \left( \nabla \cdot ( \rho ^{n+1}\mathbf{u} ^n - \rho _h^{n+1}\mathbf{u} _h^n), \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h \right) \nonumber \\&\qquad -\displaystyle \frac{1}{2} \left( \nabla \cdot ( \rho _h^{n+1}\mathbf{u} _h^n), \mathbf{e} _h^{n+1}\cdot \mathbf{v} _h \right) - \displaystyle \frac{\lambda }{2} \left( \nabla (\eta _h^{n+1}-\theta ^{n+1}), \nabla (\mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h) \right) \nonumber \\&\qquad + \displaystyle \frac{\lambda }{2} \left( \nabla \rho _h^{n+1}, \nabla (\mathbf{e} _h^{n+1}\cdot \mathbf{v} _h) \right) - \lambda \left( (\nabla \rho _h^{n+1}\cdot \nabla )\mathbf{e} _h^{n+1}, \mathbf{v} _h \right) \nonumber \\&\qquad - \lambda \left( (\nabla (\eta _h^{n+1}-\theta ^{n+1})\cdot \nabla )\mathbf{u} ^{n+1}, \mathbf{v} _h \right) + \lambda \left( (\nabla \rho ^{n+1}\cdot \nabla )\mathbf{E} ^{n+1}, \mathbf{v} _h \right) \nonumber \\&\qquad - \lambda \left( (\nabla (\eta _h^{n+1}-\theta ^{n+1})\cdot \nabla )\mathbf{E} ^{n+1}, \mathbf{v} _h \right) + \left( \rho ^{n+1}(\mathbf{u} ^n\cdot \nabla )\mathbf{E} ^{n+1}, \mathbf{v} _h \right) \nonumber \\&\qquad - \left( \rho _h^{n+1}(\mathbf{u} _h^n\cdot \nabla )\mathbf{e} _h^{n+1}, \mathbf{v} _h\right) -\left( (\eta _h^{n+1}-\theta ^{n+1})(\mathbf{u} ^n\cdot \nabla )\mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}, \mathbf{v} _h\right) \nonumber \\&\qquad - \left( \rho _h^{n+1}((\mathbf{u} ^n-\mathbf{u} _h^n)\cdot \nabla )\mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}, \mathbf{v} _h\right) -\displaystyle \frac{1}{2}\left( D_\tau \eta _h^{n+1}, \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h\right) \nonumber \\&\quad := \displaystyle \sum _{i=1}^{17} (J_{ih}^{n+1}, \mathbf{v} _h),\quad \forall \ (\mathbf{v} _h, q_h)\in \mathbf{V} _h\times M_h, \end{aligned}$$
(3.49)

where we have noted \(\nabla \cdot \mathbf{u} ^n=0\) in \(\Omega \) and

$$\begin{aligned} (D_\tau \rho ^{n+1}, \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h) + (\nabla \cdot (\rho ^{n+1} \mathbf{u} ^{n}), \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h) = - \lambda (\nabla \rho ^{n+1}, \nabla (\mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h)) \end{aligned}$$

by taking \(r=\mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{v} _h\in W_h\) in (2.9).

We first estimate \(\eta _h^{n+1}\) and \(\nabla \eta _h^{n+1}\) in \(l^\infty (L^2)\)-norm in the following two lemmas.

Lemma 3.5

Under the assumptions (A1A4), there exists some \(\tau _3<\tau _2\) such that when \(\tau <\tau _3\), there holds

$$\begin{aligned} \Vert \eta _h^{m+1}\Vert _{L^2}^2 + \tau \displaystyle \sum _{n=0}^m\Vert \eta _h^{n+1}\Vert _{H^1}^2 \le Ch^4 +C\tau \displaystyle \sum _{n=0}^m\Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 +C h^2 \tau \displaystyle \sum _{n=0}^m\Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^3}^2 \end{aligned}$$
(3.50)

for all \(0\le m\le N-1\).

Proof

Taking \(r_h=2\tau \eta _h^{n+1}\) in (3.48) leads to

$$\begin{aligned} \Vert \eta _h^{n+1}\Vert _{L^2}^2 -\Vert \eta _h^{n}\Vert _{L^2}^2 +\Vert \eta _h^{n+1}-\eta _h^{n}\Vert _{L^2}^2 +2 \lambda \tau \Vert \nabla \eta _h^{n+1}\Vert _{L^2}^2 = 2\tau \displaystyle \sum _{i=1}^4 (I_{ih}^{n+1}, \eta ^{n+1}_h) \end{aligned}$$
(3.51)

by noticing

$$\begin{aligned} (I_{5h}^{n+1}, \eta ^{n+1}_h)= & {} \displaystyle \frac{1}{2} \displaystyle \int _\Omega \nabla |\eta _h^{n+1}|^2\cdot (\mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n) dx \\= & {} - \displaystyle \frac{1}{2} \displaystyle \int _\Omega |\eta _h^{n+1}|^2\nabla \cdot (\mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n) dx = 0 \end{aligned}$$

and

$$\begin{aligned} (I_{6h}^{n+1}, \eta ^{n+1}_h) = \displaystyle \frac{1}{2} \displaystyle \int _\Omega \nabla |\eta _h^{n+1}|^2\cdot \mathbf{u} ^n dx = -\displaystyle \frac{1}{2} \displaystyle \int _\Omega |\eta _h^{n+1}|^2\nabla \cdot \mathbf{u} ^n dx =0. \end{aligned}$$

The right-hand side of (3.51) can be bound by using the Hölder inequality and the Young inequality. It follows from (3.433.45) that

$$\begin{aligned} 2\tau (I_{1h}^{n+1}, \eta ^{n+1}_h)\le & {} C\tau \Vert \eta _h^{n+1}\Vert _{L^2}^2 + C\tau h^4,\\ 2\tau (I_{2h}^{n+1}, \eta ^{n+1}_h)\le & {} C\tau \Vert \eta _h^{n+1}\Vert _{L^2}^2 + C\tau h^4. \end{aligned}$$

For \(I_3^{n+1}\) and \(I_4^{n+1}\), we can prove that

$$\begin{aligned} 2\tau (I_{3h}^{n+1}, \eta ^{n+1}_h)\le & {} C\tau \Vert \nabla \theta ^n\Vert _{L^2} \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^3}\Vert \eta _h^{n+1}\Vert _{H^1} \\\le & {} \displaystyle \frac{ \lambda \tau }{2}\Vert \eta _h^{n+1}\Vert _{H^1}^2 + C \tau h^2 \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^3}^2, \\ 2\tau (I_{4h}^{n+1}, \eta ^{n+1}_h)\le & {} C\tau \Vert \rho ^{n+1}\Vert _{H^2} \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}\Vert \eta _h^{n+1}\Vert _{H^1} \\\le & {} \displaystyle \frac{ \lambda \tau }{2}\Vert \eta _h^{n+1}\Vert _{H^1}^2 + C \tau \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2. \end{aligned}$$

Substituting the above estimates into (3.51), we obtain

$$\begin{aligned}&\Vert \eta _h^{n+1}\Vert _{L^2}^2 -\Vert \eta _h^{n}\Vert _{L^2}^2 +\Vert \eta _h^{n+1}-\eta _h^{n}\Vert _{L^2}^2 + \lambda \tau \Vert \eta _h^{n+1}\Vert _{H^1}^2 \nonumber \\&\quad \le C\tau h^4 + C\tau \Vert \eta _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 + C\tau h^2 \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^3}^2. \end{aligned}$$
(3.52)

Taking the sum from 0 to m and using the discrete Gronwall’s inequality in Lemma 2.2, we get the desired result (3.50) for some small \(\tau<\tau _3<\tau _2\). \(\square \)

Lemma 3.6

Under the assumptions (A1)-(A4), if

$$\begin{aligned} \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 \le Ch^3,\quad \forall \ 0\le n\le N-1, \end{aligned}$$
(3.53)

then there exists some \(C>0\) such that

$$\begin{aligned} \tau \displaystyle \sum _{n=0}^m \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + \lambda \Vert \nabla \eta _h^{m+1}\Vert _{L^2}^2 \le C h^4 + C\tau \displaystyle \sum _{n=0}^m \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 \end{aligned}$$
(3.54)

for all \(0\le m\le N-1\).

Proof

Taking \(r_h=2 D_\tau \eta _h^{n+1}\) in (3.48) leads to

$$\begin{aligned} 2 \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 +\lambda D_\tau \Vert \nabla \eta _h^{n+1}\Vert _{L^2}^2 +\lambda \tau \Vert \nabla (D_\tau \eta _h^{n+1})\Vert _{L^2}^2 = 2 \displaystyle \sum _{i=1}^{6} (I_{ih}^{n+1}, D_\tau \eta _h^{n+1}). \end{aligned}$$
(3.55)

We estimate the right-hand side of (3.55) term by term according to the regularities derived in (3.20), (3.34), (3.36) and (3.36). From the Hölder inequality and the Young inequality, we have

$$\begin{aligned} 2(I_{1h}^{n+1}, D_\tau \eta _h^{n+1})\le & {} \displaystyle \frac{1}{6} \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + Ch^4 \Vert D_\tau \rho ^{n+1}\Vert _{H^2}^2,\\ 2(I_{2h}^{n+1}, D_\tau \eta _h^{n+1})\le & {} \displaystyle \frac{1}{6} \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + Ch^4 \Vert \rho ^{n+1}\Vert _{H^2}^2, \\ 2(I_{3h}^{n+1}, D_\tau \eta _h^{n+1})\le & {} C \Vert \nabla \theta ^{n+1}\Vert _{L^3} \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2} \Vert D_\tau \eta _h^{n+1}\Vert _{L^6} \\\le & {} C \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}\Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2} \\\le & {} \displaystyle \frac{1}{6} \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + C \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2, \end{aligned}$$

where the inverse inequalities (2.23) is used, and

$$\begin{aligned} 2(I_{4h}^{n+1}, D_\tau \eta _h^{n+1})\le & {} \Vert \nabla \rho ^{n+1}\Vert _{L^\infty } \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2} \Vert D_\tau \eta _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{1}{6} \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + C\Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 , \end{aligned}$$

and

$$\begin{aligned} 2(I_{5h}^{n+1}, D_\tau \eta _h^{n+1})\le & {} \Vert \nabla \eta _h^{n+1}\Vert _{L^\infty } \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2} \Vert D_\tau \eta _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{1}{6} \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + C h^{-3} \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2\Vert \nabla \eta _h^{n+1}\Vert _{L^2}^2, \end{aligned}$$

where the inverse inequalities (2.23) is used, and

$$\begin{aligned} 2(I_{6h}^{n+1}, D_\tau \eta _h^{n+1})\le & {} \displaystyle \frac{1}{6}\Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + C\Vert \nabla \eta _h^{n+1}\Vert _{L^2}^2. \end{aligned}$$

Substituting the above estimates into (3.55) and taking the sum from 0 to m, we get

$$\begin{aligned}&\tau \displaystyle \sum _{n=0}^m \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 + \lambda \Vert \nabla \eta _h^{m+1}\Vert _{L^2}^2 \\&\quad \le C h^4 + C\tau \displaystyle \sum _{n=0}^m \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 + C\tau \displaystyle \sum _{n=0}^m (1+ h^{-3}\Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2) \Vert \nabla \eta _h^{n+1}\Vert _{L^2}^2 . \end{aligned}$$

By the condition (3.53) and using the discrete Gronwall inequality in Lemma 2.2, we get the desired result (3.54). \(\square \)

Next lemma presents the estimate of \(\mathbf{e} _h^{n+1}\) in \(l^\infty (\mathbf{L} ^2)\)-norm and \(l^2(\mathbf{V} )\)-norm.

Lemma 3.7

Under the assumptions (A1A4), there exists sufficiently small \(h_4>0\) and \(\tau _4<\tau _3\) such that when \(h<h_4\) and \(\tau <\tau _4\), the finite element scheme (2.18) admits a unique solution \((\mathbf{u} _h^{n+1}, p_h^{n+1})\in \mathbf{V} _h\times M_h\). Moreover, there holds

$$\begin{aligned} \Vert \mathbf{e} _h^{m+1}\Vert _{L^2}^2 + \tau \displaystyle \sum _{n=0}^m\Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 \le C_0^2 h^4, \qquad \forall \ 0\le m\le N-1, \end{aligned}$$
(3.56)

where \(C_0>0\) is independent of \(\tau \), h and m.

Proof

We will prove (3.56) by the method of mathematical induction.

\(\bullet \) Initialization \((m=0)\)

We first prove that (3.56) is valid for \(m=0\). Taking \(m=0\) in (3.50) and (3.54), we can get

$$\begin{aligned} \tau \Vert D_\tau \eta _h^1\Vert _{L^2}^2 + \Vert \eta _h^{1}\Vert _{L^2}^2 \le C h^4 \end{aligned}$$
(3.57)

by using

$$\begin{aligned} \Vert \mathbf{u} ^0-\mathbf{P} _{0h}\mathbf{u} _h^0\Vert _{L^2}^2\le & {} 2 \Vert \mathbf{u} ^0-\mathbf{P} _{0h}\mathbf{u} ^0\Vert _{L^2}^2 + 2 \Vert \mathbf{P} _{0h}\mathbf{u} ^0 - \mathbf{P} _{0h}\mathbf{u} _h^0\Vert _{L^2}^2 \\\le & {} 2 \Vert \mathbf{u} ^0-\mathbf{P} _{0h}\mathbf{u} ^0\Vert _{L^2}^2 + 2\Vert \mathbf{u} ^0 - \mathbf{u} _h^0\Vert _{L^2}^2 \\\le & {} Ch^4, \\ h^2\Vert \mathbf{u} ^0-\mathbf{P} _{0h}\mathbf{u} _h^0\Vert _{L^3}^2\le & {} 2 h^2 \Vert \mathbf{u} ^0-\mathbf{P} _{0h}\mathbf{u} ^0\Vert _{L^3}^2 + 2 h^2 \Vert \mathbf{P} _{0h}\mathbf{u} ^0 - \mathbf{P} _{0h}\mathbf{u} _h^0\Vert _{L^3}^2 \\\le & {} Ch^4 + C h \Vert \mathbf{P} _{0h}\mathbf{u} ^0 - \mathbf{P} _{0h}\mathbf{u} _h^0\Vert _{L^2}^2 \\\le & {} Ch^4. \end{aligned}$$

Furthermore, we get from the inverse inequality (2.23) and (3.57) that

$$\begin{aligned} \Vert \eta _h^{1}\Vert _{L^\infty }\le C h^{1/2}\quad \text{ and } \quad \Vert \rho ^{1}-\rho _h^1\Vert _{L^\infty }\le C h^{1/2}. \end{aligned}$$
(3.58)

Then there exists some sufficiently small \(h_4\) such that when \(h<h_4\), one has

$$\begin{aligned} \widetilde{m}< m - Ch^{1/2}< \Vert \rho _h^1\Vert _{L^\infty }< M+Ch^{1/2} <\widetilde{M}, \end{aligned}$$
(3.59)

which with (2.2) implies that the numerical scheme (2.18) with \(n=0\) admits a unique solution \((\mathbf{u} _h^1, p_h^1)\in \mathbf{V} _h\times M_h\). Taking \(n=0\) and \((\mathbf{v} _h, q_h)=2\tau (\mathbf{e} _h^1, \epsilon _h^1)\) in (3.49) and using (2.2) and \(\mathbf{e} _h^0=0\), we get

$$\begin{aligned} \Vert \sigma _h^0\mathbf{e} _h^1\Vert _{L^2}^2 + \Vert \sigma _h^1\mathbf{e} _h^1\Vert _{L^2}^2 +2\mu _1\tau \Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 \le 2\tau \displaystyle \sum _{i=1}^{17} (J_{ih}^{1}, \mathbf{e} _h^1). \end{aligned}$$
(3.60)

Due to \(\eta _h^0=0\), then from (3.433.45) and (3.35), one has

$$\begin{aligned} 2\tau \displaystyle \sum _{i=1}^{3} (J_{ih}^{1}, \mathbf{e} _h^1) \le \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4. \end{aligned}$$
(3.61)

By (3.57) and (3.58), we estimate \((J_{4h}^{1}, \mathbf{e} _h^1)\) by

$$\begin{aligned} 2\tau (J_{4h}^{1}, \mathbf{e} _h^1)\le & {} C\tau \Vert \rho ^1-\rho _h^1\Vert _{L^\infty }\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2+ C\tau \Vert \nabla \mathbf{R} _h^1\mathbf{u} ^1\Vert _{L^3}\Vert \rho ^1-\rho _h^1\Vert _{L^2}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4 \end{aligned}$$

for sufficiently small \(h<h_4\). Using the integration by parts, we estimate \((J_{5h}^{1}, \mathbf{e} _h^1)\) by

$$\begin{aligned} 2\tau (J_{5h}^{1}, \mathbf{e} _h^1)= & {} \tau \left( ((\rho ^1-\rho _h^1)\mathbf{u} _0+\rho _h^1(\mathbf{u} _0-\mathbf{u} _h^0))\cdot \nabla (\mathbf{R} _h^1\mathbf{u} ^1), \mathbf{e} _h^1 \right) \\&+ \tau \left( ((\rho ^1-\rho _h^1)\mathbf{u} _0+\rho _h^1(\mathbf{u} _0-\mathbf{u} _h^0))\cdot (\mathbf{R} _h^1\mathbf{u} ^1),\nabla \mathbf{e} _h^1 \right) \\\le & {} C\tau \left( \Vert \rho ^1-\rho _h^1\Vert _{L^2}+\Vert \mathbf{u} _0-\mathbf{u} _h^0\Vert _{L^2} \right) \Vert \nabla \mathbf{e} _h^1\Vert _{L^2}\\\le & {} \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4, \\ 2\tau (J_{6h}^{1}, \mathbf{e} _h^1)= & {} - 2\tau (J_{14h}^{1}, \mathbf{e} _h^1). \end{aligned}$$

Similarly, we can prove

$$\begin{aligned} 2\tau (J_{10h}^{1}, \mathbf{e} _h^1) + 2\tau (J_{12h}^{1}, \mathbf{e} _h^1)\le & {} C\tau \Vert \rho ^1-\rho _h^1\Vert _{L^2} \Vert \nabla \mathbf{e} _h^1\Vert _{L^2}\\\le & {} \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4,\\ 2\tau (J_{8h}^{1}, \mathbf{e} _h^1)= & {} - 2\tau (J_{9h}^{1}, \mathbf{e} _h^1) . \end{aligned}$$

From the definition of \(\Pi _h^1\), one has

$$\begin{aligned} 2\tau (J_{7h}^{1}, \mathbf{e} _h^1)= & {} -\lambda \tau (\nabla \eta _h^1, \nabla (\mathbf{R} _h^1\mathbf{u} ^1\cdot \mathbf{e} _h^1))- \tau (\nabla \theta ^1\cdot \mathbf{u} _0, \mathbf{R} _h^1\mathbf{u} ^1\cdot \mathbf{e} _h^1)-\lambda \tau (\theta ^1, \mathbf{R} _h^1\mathbf{u} ^1\cdot \mathbf{e} _h^1)\\\le & {} C\tau (\Vert \nabla \eta _h^1\Vert _{L^2}+\Vert \theta ^1\Vert _{L^2}) \Vert \nabla \mathbf{e} _h^1\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4, \end{aligned}$$

where (3.57) is used. It is easy to see that

$$\begin{aligned} 2\tau (J_{11h}^{1}, \mathbf{e} _h^1)+2\tau (J_{13h}^{1}, \mathbf{e} _h^1)\le & {} C\tau \Vert \mathbf{E} ^1\Vert _{L^2}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4, \end{aligned}$$

and

$$\begin{aligned} 2\tau (J_{15h}^{1}, \mathbf{e} _h^1)\le & {} 2\tau \Vert \mathbf{u} ^0\Vert _{L^\infty }\Vert \nabla \mathbf{R} _h^1\mathbf{u} ^1\Vert _{L^3}\Vert \rho ^1-\rho _h^1\Vert _{L^2}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}\\\le & {} \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4,\\ 2\tau (J_{16h}^{1}, \mathbf{e} _h^1)\le & {} 2\tau \Vert \rho _h^1\Vert _{L^\infty }\Vert \nabla \mathbf{R} _h^1\mathbf{u} ^1\Vert _{L^3}\Vert \mathbf{E} ^0\Vert _{L^2}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}\\\le & {} \displaystyle \frac{\mu _1\tau }{11}\Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4. \end{aligned}$$

For \(J_{17h}\), we get from (3.57) that

$$\begin{aligned} 2\tau (J_{17h}^{1}, \mathbf{e} _h^1)= & {} -\lambda \tau (D_\tau \eta _h^1, \mathbf{R} _h^1\mathbf{u} ^1\cdot \mathbf{e} _h^1) \\\le & {} C\tau \Vert D_\tau \eta _h^1\Vert _{L^2} \Vert \mathbf{e} _h^1\Vert _{L^2} \\\le & {} \displaystyle \frac{1}{4} \Vert \sigma _h^1 \mathbf{e} _h^1\Vert _{L^2}^2 + C\tau h^4. \end{aligned}$$

Taking into account these estimates with (3.60), there exits some \(C_1>0\) independent of \(C_0\), h and \(\tau \) such that

$$\begin{aligned} \Vert \mathbf{e} _h^1\Vert _{L^2}^2 + \tau \Vert \nabla \mathbf{e} _h^1\Vert _{L^2}^2 \le C_1^2\tau h^4. \end{aligned}$$

Thus, (3.56) is valid for \(m=0\) by taking \(C_0\ge C_1\).

\(\bullet \) General step \((m\ge 1)\)

For \(0\le n\le N-1\), we assume that (3.56) is valid for \(m=n\), i.e.,

$$\begin{aligned} \Vert \mathbf{e} _h^{n}\Vert _{L^2}^2 + \tau \displaystyle \sum _{i=1}^n\Vert \nabla \mathbf{e} _h^{i}\Vert _{L^2}^2 \le C_0^2 h^4. \end{aligned}$$
(3.62)

Then

$$\begin{aligned} \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2\le & {} 2\Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} ^n\Vert _{L^2}^2 + 2 \Vert \mathbf{P} _{0h}( \mathbf{u} ^n-\mathbf{u} _h^n)\Vert _{L^2}^2 \nonumber \\\le & {} Ch^4 + C (\Vert \mathbf{e} _h^n\Vert _{L^2}^2 + \Vert \mathbf{E} ^n\Vert _{L^2}^2) \nonumber \\\le & {} Ch^4 + C \Vert \mathbf{e} _h^n\Vert _{L^2}^2 \nonumber \\\le & {} C(1+C_0)^2 h^4. \end{aligned}$$
(3.63)

Thus, the condition (3.53) is valid and the estimate (3.54) holds in Lemma 3.6.

By the inverse inequality (2.23), we have

$$\begin{aligned} \Vert \mathbf{e} _h^{n}\Vert _{L^\infty }\le Ch^{-3/2}\Vert \mathbf{e} _h^{n}\Vert _{L^2} \le CC_0h^{1/2}, \end{aligned}$$

which implies that

$$\begin{aligned} \Vert \mathbf{u} _h^n\Vert _{L^\infty }\le & {} \Vert \mathbf{u} ^n\Vert _{L^\infty } + \Vert \mathbf{E} ^n\Vert _{L^\infty }+\Vert \mathbf{e} _h^{n}\Vert _{L^\infty }\nonumber \\\le & {} C \Vert \mathbf{u} ^n\Vert _{H^2} + C(1+C_0)h^{1/2} \nonumber \\\le & {} C \end{aligned}$$
(3.64)

for sufficiently small \(h<h_4\) such that \((1+C_0)h^{1/2}_4\le 1\). From (3.50) in Lemma 3.5 and (3.63), one has

$$\begin{aligned} \Vert \eta _h^{n+1}\Vert _{L^2}^2 + \tau \displaystyle \sum _{i=0}^n\Vert \eta _h^{i+1}\Vert _{H^1}^2 \le C(1+C_0)^2 h^4, \end{aligned}$$
(3.65)

where we use the inverse inequality (2.23), (3.36) and

$$\begin{aligned} h^2\Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^3}^2\le & {} Ch^2 \Vert \mathbf{E} ^n\Vert _{L^3}^2 + C h \Vert \mathbf{R} _h^n\mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 \nonumber \\\le & {} Ch^4 + C h ( \Vert \mathbf{E} ^n\Vert _{L^2}^2 + \Vert \mathbf{u} ^n-\mathbf{P} _{0h}\mathbf{u} _h^n\Vert _{L^2}^2 ) \nonumber \\\le & {} Ch^4 \end{aligned}$$
(3.66)

for sufficiently small \(h<h_4\). Then the finite element solution \(\rho _h^{n+1}\) satisfies

$$\begin{aligned} \Vert \rho ^{n+1}-\rho _h^{n+1}\Vert _{L^\infty }\le & {} \Vert \theta ^{n+1}\Vert _{L^\infty }+Ch^{-3/2}\Vert \eta _h^{n+1}\Vert _{L^2}\nonumber \\\le & {} C(1+C_0)h^{1/2}, \end{aligned}$$
(3.67)

which with (3.16) implies that

$$\begin{aligned} \widetilde{m}< m - C(1+C_0)h^{1/2}< \rho _h^{n+1}< M+C(1+C_0)h^{1/2} <\widetilde{M},\quad \forall \ 0\le n\le N-1 \end{aligned}$$
(3.68)

for sufficiently small \(h<h_4\) such that

$$\begin{aligned} C(1+C_0)h^{1/2}_4 < \max \{ m-\widetilde{m}, \ \widetilde{M}-M \}. \end{aligned}$$

According to (2.2), the fully discrete scheme (2.18) admits a unique solution \((\mathbf{u} _h^{n+1}, p_h^{n+1})\in \mathbf{V} _h\times M_h\).

To close the mathematical induction, we need to prove that (3.56) is valid for \(m=n+1\). Setting \((\mathbf{v} _h, q_h)=2\tau (\mathbf{e} _h^{n+1}, \epsilon _h^{n+1})\) in (3.49), we get

$$\begin{aligned} \Vert \sigma _h^{n+1}\mathbf{e} _h^{n+1}\Vert _{L^2}^2 - \Vert \sigma _h^{n}\mathbf{e} _h^{n}\Vert _{L^2}^2 + 2\mu _1\tau \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 \le 2\tau \displaystyle \sum _{i=1}^{17} (J_{ih}^{n+1}, \mathbf{e} _h^{n+1}). \end{aligned}$$
(3.69)

Using the Hölder inequality, the Young inequality, the regularity results (3.343.36) derived in Sect. 3 and the projection approximations (3.433.47) and the induction assumptions (3.623.68), we estimate the right-hand side of (3.69) term by term as follows:

\(\bullet \) Estimate of \(2\tau (J_{1h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{1h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} 2\tau \Vert \rho _h^n\Vert _{L^\infty } \Vert D_\tau \mathbf{E} ^{n+1}\Vert _{L^2}\Vert \mathbf{e} _h^{n+1}\Vert _{L^2}\\\le & {} C \tau \Vert \sigma _h^{n+1}\mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C \tau h^4 \Vert D_\tau \mathbf{u} ^{n+1}\Vert _{H^2}^2. \end{aligned}$$

\(\bullet \) Estimate of \(2\tau (J_{2h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{2h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} 2 \tau \left( \Vert \eta _h^n\Vert _{L^2} + \Vert \theta ^n\Vert _{L^2} \right) \Vert D_\tau \mathbf{u} ^{n+1}\Vert _{L^3}\Vert \mathbf{e} _h^{n+1}\Vert _{L^6}\\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \eta _h^n\Vert _{L^2}^2 + C\tau h^4. \end{aligned}$$

\(\bullet \) Estimate of \(2\tau (J_{3h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{3h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} \tau \Vert D_\tau \theta ^{n+1}\Vert _{L^2} \Vert \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\Vert _{L^3}\Vert \mathbf{e} _h^{n+1}\Vert _{L^6}\\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau h^4\Vert D_\tau \rho ^{n+1}\Vert _{H^2}^2. \end{aligned}$$

\(\bullet \) Estimate of \(2\tau (J_{4h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{4h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} 2\lambda \tau \left( \Vert \eta _h^{n+1}\Vert _{L^\infty } + \Vert \theta ^{n+1}\Vert _{L^\infty } \right) \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 \\&+ 2\lambda \tau \left( \Vert \eta _h^{n+1}\Vert _{L^6} + \Vert \theta ^{n+1}\Vert _{L^6} \right) \Vert \nabla \mathbf{E} ^{n+1}\Vert _{L^3} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\&+ 2\lambda \tau \left( \Vert \eta _h^{n+1}\Vert _{L^2} + \Vert \theta ^{n+1}\Vert _{L^2} \right) \Vert \nabla \mathbf{u} ^{n+1}\Vert _{L^\infty } \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \left( C(1+C_0)h^{1/2} + \displaystyle \frac{\mu _1}{32} \right) \tau \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau h^4 \\&+ C\tau h^2 \Vert \eta _h^{n+1}\Vert _{H^1}^2 + C \tau \Vert \eta _h^{n+1}\Vert _{L^2}^2 \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \eta _h^{n+1}\Vert _{L^2}^2+ C\tau h^4 + C\tau h^2 \Vert \eta _h^{n+1}\Vert _{H^1}^2 \end{aligned}$$

for sufficiently small \(h<h_4\) such that \(C(1+C_0)h_4^{1/2} < \mu _1/32\).

\(\bullet \) Estimate of \(2\tau (J_{5h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{5h}^{n+1}, \mathbf{e} _h^{n+1})= & {} \tau \left( \rho ^{n+1}(\mathbf{u} ^n-\mathbf{u} _h^n)+(\rho ^{n+1}-\rho _h^{n+1})\mathbf{u} _h^n, \nabla \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{e} _h^{n+1} \right) \\&+ \tau \left( \rho ^{n+1}(\mathbf{u} ^n-\mathbf{u} _h^n)+(\rho ^{n+1}-\rho _h^{n+1})\mathbf{u} _h^n, \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \nabla \mathbf{e} _h^{n+1} \right) \\\le & {} C\tau \left( \Vert \mathbf{e} _h^n\Vert _{L^2} + \Vert \mathbf{E} ^n\Vert _{L^2} + \Vert \eta _h^{n+1}\Vert _{L^2} + \Vert \theta ^{n+1}\Vert _{L^2} \right) \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \left( \Vert \sigma _h^n\mathbf{e} _h^n\Vert _{L^2}^2 + \Vert \eta _h^{n+1}\Vert ^2_{L^2} + h^4 \right) , \end{aligned}$$

where the integration by parts is used.

\(\bullet \) Relation of \(2\tau (J_{6h}^{n+1}, \mathbf{e} _h^{n+1})\) and \(2\tau (J_{14h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{6h}^{n+1}, \mathbf{e} _h^{n+1}) = 2 \tau (\rho _h^{n+1}\mathbf{u} ^n_h, \nabla \mathbf{e} _h^{n+1}\cdot \mathbf{e} _h^{n+1})= - 2\tau (J_{14h}^{n+1}, \mathbf{e} _h^{n+1}), \end{aligned}$$

where the integration by parts is used.

\(\bullet \) Estimate of \(2\tau (J_{7h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{7h}^{n+1}, \mathbf{e} _h^{n+1})= & {} -\lambda \tau \left( \nabla \eta _h^{n+1}, \nabla (\mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{e} _h^{n+1}) \right) \\&-\lambda \tau \left( \theta ^{n+1}, \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{e} _h^{n+1} \right) \\&-\tau \left( \nabla \theta ^{n+1}\cdot \mathbf{u} ^n, \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\cdot \mathbf{e} _h^{n+1} \right) \\\le & {} C\tau \Vert \eta _h^{n+1}\Vert _{H^1} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} + C\tau \Vert \theta ^{n+1}\Vert _{L^2} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \eta _h^{n+1}\Vert _{H^1}^2+C\tau h^4, \end{aligned}$$

where the definition of \(\Pi _h^{n+1}\) is used.

\(\bullet \) Relation of \(2\tau (J_{8h}^{n+1}, \mathbf{e} _h^{n+1})\) and \(2\tau (J_{9h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{8h}^{n+1}, \mathbf{e} _h^{n+1})= & {} - 2\tau (J_{9h}^{n+1}, \mathbf{e} _h^{n+1}) \end{aligned}$$

by using the integration by parts.

\(\bullet \) Estimate of \(2\tau (J_{10h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{10h}^{n+1}, \mathbf{e} _h^{n+1})= & {} 2\lambda \tau ((\eta _h^{n+1}-\theta ^{n+1})\Delta \mathbf{u} ^{n+1}, \mathbf{e} _h^{n+1}) \\&+ 2\lambda \tau ((\eta _h^{n+1}-\theta ^{n+1})\nabla \mathbf{u} ^{n+1}, \nabla \mathbf{e} _h^{n+1}) \\\le & {} C\tau \left( \Vert \eta _h^{n+1}\Vert _{L^2}+\Vert \theta ^{n+1}\Vert _{L^2} \right) \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \eta _h^{n+1}\Vert _{L^2}^2 + C\tau h^4 \end{aligned}$$

by using the integration by parts.

\(\bullet \) Estimate of \(2\tau (J_{11h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{11h}^{n+1}, \mathbf{e} _h^{n+1})= & {} -2\lambda \tau ( \Delta \rho ^{n+1}\mathbf{E} ^{n+1}, \mathbf{e} _h^{n+1}) -2\lambda \tau ( \nabla \rho ^{n+1}, \mathbf{E} ^{n+1}\cdot \nabla \mathbf{e} _h^{n+1}) \\\le & {} C\tau \Vert \rho ^{n+1}\Vert _{H^3} \Vert \mathbf{E} ^{n+1}\Vert _{L^2}\Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau h^4 \end{aligned}$$

by using the integration by parts.

\(\bullet \) Estimate of \(2\tau (J_{12h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{12h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} C\tau \left( \Vert \nabla \eta _h^{n+1}\Vert _{L^2}+\Vert \nabla \theta ^{n+1}\Vert _{L^2} \right) \Vert \nabla \mathbf{E} ^{n+1}\Vert _{L^3} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \nabla \eta _h^{n+1}\Vert _{L^2}^2 + C\tau h^4 . \end{aligned}$$

\(\bullet \) Estimate of \(2\tau (J_{13h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{13h}^{n+1}, \mathbf{e} _h^{n+1})= & {} -2\tau \left( \nabla \rho ^{n+1}\cdot \mathbf{u} ^n, \mathbf{E} ^{n+1}\cdot \mathbf{e} _h^{n+1} \right) -2\tau \left( \rho ^{n+1} (\mathbf{u} ^n\cdot \nabla )\mathbf{e} _h^{n+1}, \mathbf{E} ^{n+1} \right) \\\le & {} C\tau \Vert \mathbf{u} ^n\Vert _{H^2}\Vert \rho ^{n+1}\Vert _{H^2} \Vert \mathbf{E} ^{n+1}\Vert _{L^2} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}\\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau h^4 \end{aligned}$$

by using the integration by parts and \(\nabla \cdot \mathbf{u} ^n=0\) in \(\Omega \).

\(\bullet \) Estimate of \(2\tau (J_{15h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{15h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} C\tau \left( \Vert \eta _h^{n+1}\Vert _{L^2}+\Vert \theta ^{n+1}\Vert _{L^2} \right) \Vert \mathbf{u} ^{n}\Vert _{L^\infty } \Vert \nabla \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\Vert _{L^3} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \eta _h^{n+1}\Vert _{L^2}^2 + C\tau h^4. \end{aligned}$$

\(\bullet \) Estimate of \(2\tau (J_{16h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{16h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} C\tau \left( \Vert \mathbf{e} _h^{n}\Vert _{L^2}+\Vert \mathbf{E} ^{n}\Vert _{L^2} \right) \Vert \rho _h^{n+1}\Vert _{L^\infty } \Vert \nabla \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\Vert _{L^3} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert \sigma _h^n\mathbf{e} _h^{n}\Vert _{L^2}^2 + C\tau h^4. \end{aligned}$$

\(\bullet \) Estimate of \(2\tau (J_{17h}^{n+1}, \mathbf{e} _h^{n+1})\)

$$\begin{aligned} 2\tau (J_{17h}^{n+1}, \mathbf{e} _h^{n+1})\le & {} C\tau \Vert D_\tau \eta _h^{n+1}\Vert _{L^2} \Vert \mathbf{R} _h^{n+1}\mathbf{u} ^{n+1}\Vert _{L^3} \Vert \mathbf{e} _h^{n+1}\Vert _{L^6} \\\le & {} \displaystyle \frac{\mu _1\tau }{16} \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + C\tau \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2. \end{aligned}$$

Substituting these estimates for \(J_{1h}^{n+1}\) to \(J_{17h}^{n+1}\) into (3.69), we get

$$\begin{aligned}&\Vert \sigma _h^{n+1}\mathbf{e} _h^{n+1}\Vert _{L^2}^2 - \Vert \sigma _h^{n}\mathbf{e} _h^{n}\Vert _{L^2}^2 + \mu _1\tau \Vert \nabla \mathbf{e} _h^{n+1}\Vert _{L^2}^2 \\&\quad \le C\tau h^4 + C\tau (\Vert \sigma _h^{n+1}\mathbf{e} _h^{n+1}\Vert _{L^2}^2+ \Vert \sigma _h^n\mathbf{e} _h^{n}\Vert _{L^2}^2) + C\tau h^4 ( \Vert D_\tau \mathbf{u} ^{n+1}\Vert _{H^2}^2+\Vert D_\tau \rho ^{n+1}\Vert _{H^2}^2 ) \\&\qquad + C\tau (\Vert \eta _h^{n}\Vert _{L^2}^2 + \Vert \eta _h^{n+1}\Vert _{L^2}^2 + \Vert \nabla \eta _h^{n+1}\Vert _{L^2}^2 + \Vert D_\tau \eta _h^{n+1}\Vert _{L^2}^2 ). \end{aligned}$$

Taking the sum gives

$$\begin{aligned}&\Vert \sigma _h^{n+1}\mathbf{e} _h^{n+1}\Vert _{L^2}^2 + \mu _1\tau \displaystyle \sum _{i=0}^n \Vert \nabla \mathbf{e} _h^{i+1}\Vert _{L^2}^2 \nonumber \\&\quad \le C h^4 + C\tau \displaystyle \sum _{i=0}^n (\Vert \sigma _h^{i+1}\mathbf{e} _h^{i+1}\Vert _{L^2}^2+ \Vert D_\tau \eta _h^{i+1}\Vert _{L^2}^2 + \Vert \eta _h^{i+1}\Vert _{L^2}^2 + \Vert \nabla \eta _h^{i+1}\Vert _{L^2}^2) \nonumber \\&\quad \le C h^4 + C\tau \displaystyle \sum _{i=0}^n \Vert \sigma _h^{i+1}\mathbf{e} _h^{i+1}\Vert _{L^2}^2 + C h^2 \tau \displaystyle \sum _{i=0}^n \Vert \mathbf{u} ^i-\mathbf{P} _{0h}\mathbf{u} _h^i\Vert _{L^3}^2 \nonumber \\&\qquad + C \tau \displaystyle \sum _{i=0}^n \Vert \mathbf{u} ^i-\mathbf{P} _{0h}\mathbf{u} _h^i\Vert _{L^2}^2, \end{aligned}$$
(3.70)

where we used error estimates derived in Lemmas 3.5 and 3.6. From (3.36) (3.63) and (3.66), we have

$$\begin{aligned} h^2 \tau \displaystyle \sum _{i=0}^n \Vert \mathbf{u} ^i-\mathbf{P} _{0h}\mathbf{u} _h^i\Vert _{L^3}^2\le & {} Ch^4 \end{aligned}$$

and

$$\begin{aligned} C \tau \displaystyle \sum _{i=0}^n \Vert \mathbf{u} ^i-\mathbf{P} _{0h}\mathbf{u} _h^i\Vert _{L^2}^2\le & {} C h^4 + C \tau \displaystyle \sum _{i=0}^n \Vert \mathbf{e} _h^i\Vert _{L^2}^2 \\\le & {} C h^4 + C \tau \displaystyle \sum _{i=0}^n \Vert \sigma _h^i \mathbf{e} _h^i\Vert _{L^2}^2. \end{aligned}$$

Then (3.70) reduces to

$$\begin{aligned}&\Vert \sigma _h^{n+1}\mathbf{e} _h^{n+1}\Vert _{L^2}^2 + \mu _1\tau \displaystyle \sum _{i=0}^n \Vert \nabla \mathbf{e} _h^{i+1}\Vert _{L^2}^2 \\&\quad \le C h^4 + C\tau \displaystyle \sum _{i=0}^n \Vert \sigma _h^{i+1}\mathbf{e} _h^{i+1}\Vert _{L^2}^2 + C \tau \displaystyle \sum _{i=0}^n \Vert \sigma _h^i\mathbf{e} _h^i\Vert _{L^2}^2. \end{aligned}$$

Applying the discrete Gronwall’s inequality in Lemma 2.2, we derive

$$\begin{aligned} \Vert \sigma _h^{n+1}\mathbf{e} _h^{n+1}\Vert _{L^2}^2 + \mu _1\tau \displaystyle \sum _{i=0}^n \Vert \nabla \mathbf{e} _h^{i+1}\Vert _{L^2}^2 \le C\exp (CT) h^4 \end{aligned}$$

and

$$\begin{aligned} \Vert \mathbf{e} _h^{n+1}\Vert _{L^2}^2 + \mu _1\tau \displaystyle \sum _{i=0}^n \Vert \nabla \mathbf{e} _h^{i+1}\Vert _{L^2}^2 \le C\exp (CT) h^4 \le C_0^2 h^4 \end{aligned}$$

by using (3.68) and taking \(\sqrt{C\exp (CT)} \le C_0\). Thus, we prove that (3.56) is valid for \(m=n+1\) and finish the mathematical induction. \(\square \)

3.3 Proof of Theorem 2.4

By (3.9) in Lemma 3.2 and (3.56) in Lemma 3.7, it is easy to see that

$$\begin{aligned} \Vert \mathbf{u} (t_{n+1})-\mathbf{u} _h^{n+1}\Vert _{L^2}\le & {} \Vert \mathbf{e} ^{n+1}\Vert _{L^2} + \Vert \mathbf{E} ^{n+1}\Vert _{L^2} + \Vert \mathbf{e} _h^{n+1}\Vert _{L^2} \\\le & {} C (\tau + h^2),\quad \forall \ 0\le n\le N-1, \end{aligned}$$

where the uniform boundness (3.12) of \(\rho ^{n+1}\) is used. Thus, we get the optimal \(\mathbf{L} ^2\) error estimate for the velocity. To establish the optimal \(H^1\) error estimate for the density, we have

$$\begin{aligned} \Vert \eta _h^{n+1}\Vert _{H^1} \le C h^2, \end{aligned}$$

where (3.50) in Lemma 3.5 and (3.54) in Lemma 3.6 are used. Then,

$$\begin{aligned} \Vert \rho (t_{n+1})-\rho _h^{n+1}\Vert _{H^1}\le & {} \Vert \eta ^{n+1}\Vert _{H^1} + \Vert \theta ^{n+1}\Vert _{H^1} + \Vert \eta _h^{n+1}\Vert _{H^1} \\\le & {} C (\tau + h^2), \quad \forall \ 0\le n\le N-1. \end{aligned}$$

Thus, we complete the proof of Theorem 2.4.