1 Introduction

Chemotaxis, the biased movement of cells (or organisms) in response to chemical gradients, plays an important role in coordinating cell migration in many biological phenomena (see Hillen and Painter [8]). Let n denote the density of the cells and c present the concentration of the chemical signal. In the 1970s, Keller and Segel [12] proposed a mathematical system for chemotaxis through a system of parabolic equations. The mathematical model reads as

$$\begin{aligned} \left\{ \begin{array}{ll} n_t=\Delta n-\nabla \cdot ( nS(x,n,c)\nabla c), &{}\quad x\in \Omega ,\;\; t>0,\\ \displaystyle { c_t=\Delta c- c +n,}&{}\quad x\in \Omega ,\;\; t>0,\\ \end{array} \right. \end{aligned}$$
(1.1)

where S is a given chemotactic sensitivity function, which can either be a scalar function or, more generally, a tensor-valued function (see, e.g., Xue and Othmer [42]). During the past four decades, Keller–Segel models (1.1) and their variants have attracted extensive attention, withthe main issue of investigation focusing on whether the solutions of the models are bounded or blow up (see Winkler et al. [1], Hillen and Painter [8] and Horstmann [9]). For instance, if \(S:=S(n)\) is a scalar function satisfying \(S(s)\le C(1 + s)^{-\alpha }\) for all \(s\ge 1\), \(\alpha > 1 -\frac{2}{N}\), and \(C>0\), then all solutions to the corresponding Neumann problem are global and uniformly bounded (see Horstmann and Winkler [10]). However, if \(N\ge 2\), \(\Omega \) (a ball) \(\subset R^N\), and \(S(s) > cs^{-\alpha }\) for some \(\alpha < 1-\frac{2}{N}\) and \(c>0\), then the solution to problem (1.1) may blow up (see Horstmann and Winkler [10]). Therefore,

$$\begin{aligned} \alpha =1-\frac{2}{N} \end{aligned}$$
(1.2)

is the critical blow-up exponent, which is related to the presence of a so-called volume-filling effect. For related works in this direction, we mention that a corresponding quasilinear version, the logistic damping or the signal consumed by the cells, has been deeply investigated by Cieślak and Stinner [4, 5], Tao and Winkler [20, 31, 41] and Zheng et al. [44,45,46, 50, 51].

As in the classical Keller–Segel model where the chemoattractant is produced by bacteria, the corresponding chemotaxis–fluid model then becomes the following Keller–Segel(–Navier)–Stokes system:

$$\begin{aligned} \left\{ \begin{array}{ll} n_t+u\cdot \nabla n=\Delta n-\nabla \cdot (nS(x, n, c)\nabla c),&{}\quad x\in \Omega ,\;\; t>0,\\ c_t+u\cdot \nabla c=\Delta c-c+n,&{}\quad x\in \Omega ,\;\; t>0,\\ u_t+\kappa (u\cdot \nabla )u+\nabla P=\Delta u+n\nabla \phi ,&{}\quad x\in \Omega ,\;\; t>0,\\ \nabla \cdot u=0,&{}\quad x\in \Omega ,\;\; t>0,\\ \displaystyle {(\nabla n-nS(x, n, c))\cdot \nu =\nabla c\cdot \nu =0,\;u=0,}&{}\quad x\in \partial \Omega ,\;\; t>0,\\ \displaystyle {n(x,0)=n_0(x),c(x,0)=c_0(x),\;u(x,0)=u_0(x),}&{}\quad x\in \Omega ,\\ \end{array}\right. \end{aligned}$$
(1.3)

where n and c are defined as before and \(\Omega \subset {\mathbb {R}}^3\) is a bounded domain with a smooth boundary. Here u, \(P,\phi \), and \(\kappa \in {\mathbb {R}}\) denote, respectively, the velocity field, the associated pressure of the fluid, the potential of the gravitational field, and the strength of nonlinear fluid convection. S(xnc) is a chemotactic sensitivity tensor satisfying

$$\begin{aligned} S\in C^2({\bar{\Omega }}\times [0,\infty )^2;{\mathbb {R}}^{3\times 3}) \end{aligned}$$
(1.4)

and

$$\begin{aligned} |S(x, n, c)|\le C_S(1 + n)^{-\alpha }\quad \text{ for } \text{ all }\;\; (x, n, c)\in \Omega \times [0,\infty )^2 \end{aligned}$$
(1.5)

with some \(C_S > 0\) and \(\alpha > 0\). Problem (1.3) is proposed to describe the chemotaxis-fluid interaction in cases when the evolution of the chemoattractant is essentially dominated by production through cells (see Winkler et al. [1] and Hillen and Painter [8]).

Before delving into our mathematical analysis, we recall some important progress on system (1.3) and its variants. The following chemotaxis-fluid model,which is closely related to the variation of (1.3), was proposed by Tuval et al. [24]:

$$\begin{aligned} \left\{ \begin{array}{ll} n_t+u\cdot \nabla n=\Delta n-\nabla \cdot ( nS(x,n,c)\nabla c),&{}\quad x\in \Omega ,\;\; t>0,\\ c_t+u\cdot \nabla c=\Delta c-nf(c),&{}\quad x\in \Omega ,\;\; t>0,\\ u_t+\kappa (u\cdot \nabla )u+\nabla P=\Delta u+n\nabla \phi ,&{}\quad x\in \Omega ,\;\; t>0,\\ \nabla \cdot u=0,&{}\quad x\in \Omega ,\;\; t>0,\\ \end{array}\right. \end{aligned}$$
(1.6)

where f(c) is the consumption rate of oxygen by the cells. In the past few years, by making use of the energy-type functionals, system (1.6) and its variants have attracted extensive attention (see, e.g., Chae et al. [3], Duan et al. [6], Liu and Lorz [13, 15], Tao and Winkler [23, 33, 34, 37], Zhang and Zheng [43] and references therein). For example, Winkler [37] established the global existence of a weak solution in a three-dimensional domain when \(S(x, n, c) \equiv 1\) and \(\kappa \ne 0\). Recently, if \(S(x,n,c):=S(c),\) the long-term behavior of eventual smoothness of the weak solution was investigated by Winkler [38], in which the weak solution became smooth on some interval \([T,\infty )\) and uniformly converged in the large-time limit. For more literature related to this model, we can refer to Tao and Winkler [21, 22, 39] and the references therein. For example, Winkler [39] proved that the chemotaxis–Stokes system (with nonlinear diffusion) admits a global bounded weak solution under the assumption \(m > \frac{9}{8}\). Furthermore, he also showed that the obtained solution approached the spatially homogeneous steady state in the large-time limit.

If the chemotactic sensitivity S(xnc) is regarded as a tensor rather than a scalar one (see Xue and Othmer [42]), (1.6) turns into a chemotaxis(–Navier)–Stokes system with rotational flux. Owing to the presence of the tensor-valued sensitivity, the corresponding chemotaxis–Stokes system loses some energy structure, which has played a key role in previous studies for the scalar sensitivity case (see Cao [2] and Winkler [36]). Therefore, very few results appear to be available on chemotaxis-Stokes systems with such tensor-valued sensitivities (see, e.g., Ishida [11], Wang et al. [26, 28] and Winkler [36]). In fact, assuming that \(f(c)=c\) and that (1.4) and (1.5) hold, Ishida [11] proved that (1.6) admits a bounded global weak solution in two dimensions with nonlinear diffusion, whereas, in three dimensions, Winkler [36] showed that the chemotaxis-Stokes system [\(\kappa =0\) in the first equation of (1.6)] with nonlinear diffusion (where the coefficient of diffusion satisfies \(m > \frac{7}{6}\)) possesses at least one bounded weak solution that stabilizes to a spatially homogeneous equilibrium \((\frac{1}{|\Omega |}\int _{\Omega }n_0, 0, 0)\).

In contrast to the large number of existing results ofor (1.6), the mathematical analysis of (1.3) with regard to global and bounded solutions is far from trivial. On the one hand, as its subsystem, the Navier–Stokes system lacks a complete existence theory (see Wiegner [30]). On the other hand, the previously mentioned properties for the Keller–Segel system can still emerge (see Wang et al. [17, 25, 27,28,29] and Zheng [48, 49]). In fact, in two dimensions, if \(S=S(x, n, c)\) is a tensor-valued sensitivity fulfilling (1.4) and (1.5), Wang and Xiang [28] proved that the Stokes version [\(\kappa =0\) in the first equation of (1.3)] of system (1.3) admits a unique global classical solution that is bounded. Recently, Wang et al. [27] extended the above result [28] to the Navier–Stokes version (\(\kappa \ne 0\) in the first equation of (1.3)). In both papers [27] and [28], the condition \(\alpha >0\), corresponding to the condition (1.2) with \(n=2\), is optimal for the existence of the solution. Furthermore, similar results are also valid for the three-dimensional Stokes version [\(\kappa =0\) in the first equation of (1.3)] of system (1.3) with \(\alpha >\frac{1}{2}\) (see Wang and Xiang [29]). In the three dimensional case, Wang and Liu [14] showed that the Keller–Segel–Navier–Stokes [\(\kappa \ne 0\) in the first equation of (1.3)) system (1.3] admits a global weak solution for tensor-valued sensitivity S(xnc) satisfying (1.4) and (1.5) with \(\alpha > \frac{3}{7}\). Recently, because of the lack of enough regularity and compactness properties for the first equation, by using the idea proposed by Winkler [35], Wang [25] presented the existence of global very weak solutions for the system (1.3) under the assumption that S satisfies (1.4) and (1.5) with \(\alpha > \frac{1}{3}\), which, in light of the known results for the fluid-free system mentioned above, is an optimal restriction on \(\alpha \) [see (1.2)]. However, the existence of global (stronger than the result of [25]) weak solutions is still open. In this paper, we try to obtain enough regularity and compactness properties (see Lemmas 3.4, 5.1, and 5.2), then show that system (1.3) possesses a globally defined weak solution (see Definition 2.1), which improves the result of [25].

Throughout this paper, we assume that

$$\begin{aligned} \phi \in W^{2,\infty }(\Omega ) \end{aligned}$$
(1.7)

and that the initial data \((n_0, c_0, u_0)\) fulfill

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle {n_0\in C^\kappa ({\bar{\Omega }})~~\text{ for } \text{ certain }~~ \kappa > 0~~ \text{ with }~~ n_0\ge 0 ~~\text{ in }~~\Omega },\\ \displaystyle {c_0\in W^{1,\infty }(\Omega )~~\text{ with }~~c_0\ge 0~~\text{ in }~~{\bar{\Omega }},}\\ \displaystyle {u_0\in D(A^\gamma _{r})~~\text{ for } \text{ some }~~\gamma \in ( {3}/{4}, 1)~~\text{ and } \text{ any }~~ {r}\in (1,\infty ),}\\ \end{array} \right. \end{aligned}$$
(1.8)

where \(A_{r}\) denotes the Stokes operator with domain \(D(A_{r}) := W^{2,{r}}(\Omega )\cap W^{1,{r}}_0(\Omega ) \cap L^{r}_{\sigma }(\Omega )\) and \(L^{r}_{\sigma }(\Omega ) := \{\varphi \in L^{r}(\Omega )|\nabla \cdot \varphi = 0\}\) for \({r}\in (1,\infty )\) (similar to that in [19]).

Our main result assert the existence of the global weak solution for system (1.3).

Theorem 1.1

Let \(\Omega \subset {\mathbb {R}}^3\) be a bounded domain with a smooth boundary. (1.7) and (1.8) hold, and suppose that S satisfies (1.4) and (1.5) with some

$$\begin{aligned} \alpha >\frac{1}{3}. \end{aligned}$$

Then problem (1.3) possesses at least one global weak solution (ncuP) in the sense of Definition 2.1.

Remark 1.1

(i) From Theorem 1.1, we conclude that \(\alpha > \frac{1}{3}\) is sufficient to guarantee the existence of global (weak) solutions. Compared with the results (1.2), we know such a restriction on \(\alpha \) seems to be optimal.

(ii) Obviously, \( \frac{3}{7}>\frac{1}{3}\), so Theorem 1.1 improves the results of Liu and Wang [14], which showed the global weak existence of solutions in cases S(xnc) satisfying (1.4) and (1.5) with \(\alpha > \frac{3}{7}\).

(iii) If \(S:=S(n)=C_S(1+n)^{-\alpha }\) is a scalar function which satisfies that \(\alpha >\frac{1}{3},\) the boundedness of solution to Keller–Segel–Stokes [\(\kappa =0\) in the first equation of (1.3)] system (1.3) is obtained by Winkler (see [40]). Recalling the condition (1.2) for global existence in the fluid-free setting, as implied by the previously mentioned studied (see Horstmann and Winkler [10]), this result appears to be optimal with respect to \(\alpha \).

This paper is organized as followed. In Sect. 2, we give the definition of weak solutions to (1.3), the regularized problems of (1.3), and some preliminary properties. Sections 3 and 4 will be devoted to an analysis of regularized problems of (1.3). Next, on the basis of the compactness properties thereby implied, in Sects. 5 and 6, we can pass to the limit along with an adequate sequence of numbers \(\varepsilon = \varepsilon _j{\searrow }0\) and thereby verify Theorem 1.1.

2 Preliminaries

In light of the strong nonlinear term \((u \cdot \nabla )u\), problem (1.3) has no classical solutions in general, thus we consider its weak solutions.

Definition 2.1

Let \(T > 0\) and assume that \((n_0, c_0, u_0)\) fulfills (1.8). Then a triple of functions (ncu) is called a weak solution of (1.3) if the following conditions are satisfied:

$$\begin{aligned} \left\{ \begin{array}{ll} n\in L_{loc}^1({\bar{\Omega }}\times [0,T)),\\ c \in L_{loc}^1([0,T); W^{1,1}(\Omega )),\\ u \in L_{loc}^1([0,T); W^{1,1}(\Omega );{\mathbb {R}}^{3}), \end{array}\right. \end{aligned}$$
(2.1)

where \(n\ge 0\) and \(c\ge 0\) in \(\Omega \times (0, T)\) as well as \(\nabla \cdot u = 0\) in the distributional sense in \(\Omega \times (0, T)\). Moreover,

$$\begin{aligned}&u\otimes u \in L^1_{loc}({\bar{\Omega }}\times [0, \infty );{\mathbb {R}}^{3\times 3})~\text{ and }~ n~\text{ belongs } \text{ to }~ L^1_{loc}({\bar{\Omega }}\times [0, \infty )),\nonumber \\&cu,~ nu, ~\text{ and }~nS(x,n,c)\nabla c~ \text{ belong } \text{ to }~ L^1_{loc}({\bar{\Omega }}\times [0, \infty );{\mathbb {R}}^{3}) \end{aligned}$$
(2.2)

and

$$\begin{aligned}&\displaystyle {-\,\int _0^{T}\int _{\Omega }n\varphi _t-\int _{\Omega }n_0\varphi (\cdot ,0)}\nonumber \\&\quad = \displaystyle {- \int _0^T\int _{\Omega }\nabla n\cdot \nabla \varphi +\int _0^T\int _{\Omega }n S(x,n,c)\nabla c\cdot \nabla \varphi } +\displaystyle {\int _0^T\int _{\Omega }nu\cdot \nabla \varphi } \end{aligned}$$
(2.3)

for any \(\varphi \in C_0^{\infty } ({\bar{\Omega }}\times [0, T))\) satisfying \(\frac{\partial \varphi }{\partial \nu }= 0\) on \(\partial \Omega \times (0, T)\), as well as

$$\begin{aligned}&\displaystyle {-\,\int _0^{T}\int _{\Omega }c\varphi _t-\int _{\Omega }c_0\varphi (\cdot ,0)}\nonumber \\&\quad = \displaystyle {- \int _0^T\int _{\Omega }\nabla c\cdot \nabla \varphi -\int _0^T\int _{\Omega }c\varphi +\int _0^T\int _{\Omega }n\varphi + \int _0^T\int _{\Omega }cu\cdot \nabla \varphi } \end{aligned}$$
(2.4)

for any \(\varphi \in C_0^{\infty } ({\bar{\Omega }}\times [0, T))\) and

$$\begin{aligned}&\displaystyle {-\,\int _0^{T}\int _{\Omega }u\varphi _t-\int _{\Omega }u_0\varphi (\cdot ,0) -\kappa \int _0^T\int _{\Omega } u\otimes u\cdot \nabla \varphi }\nonumber \\&\quad = \displaystyle {- \int _0^T\int _{\Omega }\nabla u\cdot \nabla \varphi - \int _0^T\int _{\Omega }n\nabla \phi \cdot \varphi } \end{aligned}$$
(2.5)

for any \(\varphi \in C_0^{\infty } ({\bar{\Omega }}\times [0, T);{\mathbb {R}}^3)\) fulfilling \(\nabla \varphi \equiv 0\) in \(\Omega \times (0, T)\).

If (ncu) :  \(\Omega \times (0,\infty )\longrightarrow {\mathbb {R}}^5\) is a weak solution of (1.3) in \(\Omega \times (0, T)\) for all \(T > 0\), then (ncu) is called a global weak solution of (1.3).

To obtain the solution of system (1.3), we first consider the following approximate system of (1.3):

$$\begin{aligned} \left\{ \begin{array}{ll} n_{\varepsilon t}+u_{\varepsilon }\cdot \nabla n_{\varepsilon }=\Delta n_{\varepsilon }-\nabla \cdot (n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_{\varepsilon }),&{}\quad x\in \Omega ,\;\; t>0,\\ c_{\varepsilon t}+u_{\varepsilon }\cdot \nabla c_{\varepsilon }=\Delta c_{\varepsilon }-c_{\varepsilon }+F_{\varepsilon }(n_{\varepsilon }),&{}\quad x\in \Omega ,\;\; t>0,\\ u_{\varepsilon t}+\nabla P_{\varepsilon }=\Delta u_{\varepsilon }-\kappa (Y_{\varepsilon }u_{\varepsilon } \cdot \nabla )u_{\varepsilon }+n_{\varepsilon }\nabla \phi ,&{}\quad x\in \Omega ,\;\; t>0,\\ \nabla \cdot u_{\varepsilon }=0,&{}\quad x\in \Omega ,\;\; t>0,\\ \displaystyle {\nabla n_{\varepsilon }\cdot \nu =\nabla c_{\varepsilon }\cdot \nu =0,u_{\varepsilon }=0,}&{}\quad x\in \partial \Omega ,\;\; t>0,\\ \displaystyle {n_{\varepsilon }(x,0)=n_0(x),c_{\varepsilon }(x,0)=c_0(x),\;u_{\varepsilon }(x,0)=u_0(x)},&{}\quad x\in \Omega ,\\ \end{array}\right. \end{aligned}$$
(2.6)

where

$$\begin{aligned} F_{\varepsilon }(s):=\frac{1}{\varepsilon }\ln (1+\varepsilon s)\quad \text{ for } \text{ all }\;\; s \ge 0\;\; \text{ and }\;\; \varepsilon > 0, \end{aligned}$$
(2.7)

as well as

$$\begin{aligned} \begin{array}{ll} S_\varepsilon (x, n, c) := \rho _\varepsilon (x)S(x, n, c),\quad x\in {\bar{\Omega }},\;\; n\ge 0,\;\; c\ge 0 \end{array} \end{aligned}$$
(2.8)

and

$$\begin{aligned} Y_{\varepsilon }w := (1 + \varepsilon A)^{-1}w \quad ~\text{ for } \text{ all }\;\; w\in L^2_{\sigma }(\Omega ) \end{aligned}$$

is a standard Yosida approximation and A is the realization of the Stokes operator (see [19]). Here, \((\rho _\varepsilon )_{\varepsilon \in (0,1)} \in C^\infty _0 (\Omega )\) is a family of standard cutoff functions satisfying \(0\le \rho _\varepsilon \le 1\) in \(\Omega \) and \(\rho _\varepsilon {\nearrow }1\) in \(\Omega \) as \(\varepsilon {\searrow }0\).

The local solvability of (2.6) can be derived by a suitable extensibility criterion and a slight modification of the well-established fixed-point arguments in Lemma 2.1 of [37] (see also [36] and Lemma 2.1 of [16]), so here we omit the proof.

Lemma 2.1

Assume that \(\varepsilon \in (0,1).\) Then there exist \(T_{max,\varepsilon }\in (0,\infty ]\) and a classical solution \((n_\varepsilon , c_\varepsilon , u_\varepsilon , P_\varepsilon )\) of (2.6) in \(\Omega \times (0, T_{max,\varepsilon })\) such that

$$\begin{aligned} \left\{ \begin{array}{ll} n_\varepsilon \in C^0({\bar{\Omega }}\times [0,T_{max,\varepsilon }))\cap C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon })),\\ c_\varepsilon \in C^0({\bar{\Omega }}\times [0,T_{max,\varepsilon }))\cap C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon })),\\ u_\varepsilon \in C^0({\bar{\Omega }}\times [0,T_{max,\varepsilon }); {\mathbb {R}}^3)\cap C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon }); {\mathbb {R}}^3),\\ P_\varepsilon \in C^{1,0}({\bar{\Omega }}\times (0,T_{max,\varepsilon })), \end{array}\right. \end{aligned}$$

classically solving (2.6) in \(\Omega \times [0,T_{max,\varepsilon })\). Moreover, \(n_\varepsilon \) and \(c_\varepsilon \) are nonnegative in \(\Omega \times (0, T_{max,\varepsilon })\), and

$$\begin{aligned} \Vert n_\varepsilon (\cdot , t)\Vert _{L^\infty (\Omega )}+\Vert c_\varepsilon (\cdot , t)\Vert _{W^{1,\infty }(\Omega )}+\Vert A^\gamma u_\varepsilon (\cdot , t)\Vert _{L^{2}(\Omega )}\rightarrow \infty ~~ \text{ as }~~ t\rightarrow T_{max,\varepsilon }, \end{aligned}$$

where \(\gamma \) is given by (1.8).

Lemma 2.2

[32, 47] Let \((e^{\tau \Delta })_{\tau \ge 0}\) be the Neumann heat semigroup in \(\Omega \) and \(p>3\). Then there exist positive constants \(c_1:=c_1(\Omega ),\) \(c_2:=c_2(\Omega ),\) and \(c_3:=c_3(\Omega )\) such that for all \(\tau >0\) and any \(\varphi \in W^{1,p}(\Omega )\),

$$\begin{aligned} \Vert \nabla e^{\tau \Delta }\varphi \Vert _{L^p(\Omega )} \le c_1(\Omega )\Vert \nabla \varphi \Vert _{L^p(\Omega )}, \end{aligned}$$

and for all \(\tau > 0\) and each \(\varphi \in L^\infty (\Omega )\)

$$\begin{aligned} \Vert \nabla e^{\tau \Delta }\varphi \Vert _{L^p(\Omega )} \le c_2(1+\tau ^{-\frac{1}{2}})\Vert \varphi \Vert _{L^\infty (\Omega )}, \end{aligned}$$

as well as for all \(\tau > 0\) and all \(\varphi \in C^{1}({\bar{\Omega }}; {\mathbb {R}}^3)\) fulfilling \(\varphi \cdot \nu =0~\text{ on }~\partial \Omega \)

$$\begin{aligned} \Vert e^{\tau \Delta }\nabla \cdot \varphi \Vert _{L^\infty (\Omega )}\le c_3(1+\tau ^{-\frac{1}{2}-\frac{3}{2p}})\Vert \varphi \Vert _{L^p(\Omega )}. \end{aligned}$$

3 Some a priori estimates for the regularized problem (2.6) that is independent of \(\varepsilon \)

In this section, we are going to establish an iteration step to develop the main ingredients of our result. The iteration depends on a series of a priori estimates. To proceed, first, we recall some properties of \(F_\varepsilon \) and \(F'_\varepsilon \), which play an important role in demonstrating Theorem 1.1.

Lemma 3.1

Assume \(F_{\varepsilon }\) is given by (2.7). Then

$$\begin{aligned} 0\le F'_{\varepsilon }(s)=\frac{1}{1+\varepsilon s}\le 1\quad \text{ for } \text{ all }\;\; s \ge 0\;\; \text{ and }\;\; \varepsilon > 0 \end{aligned}$$
(3.1)

as well as

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0^{+}}F_{\varepsilon }(s)=s,\quad \lim _{\varepsilon \rightarrow 0^{+}}F'_{\varepsilon }(s)=1\quad \text{ for } \text{ all }\;\; s \ge 0 \end{aligned}$$
(3.2)

and

$$\begin{aligned} 0\le F_{\varepsilon }(s)\le s\quad \text{ for } \text{ all }\;\; s \ge 0. \end{aligned}$$
(3.3)

Proof

Recalling (2.7), by tedious and simple calculations, we can derive (3.1)–(3.3). \(\square \)

The proof of this lemma is very similar to that of Lemmas 2.2 and 2.6 of [23] (see also Lemma 3.2 of [25]), so we omit it here.

Lemma 3.2

There exists \(\lambda > 0\) independent of \(\varepsilon \) such that the solution of (2.6) satisfies

$$\begin{aligned} \int _{\Omega }{n_{\varepsilon }}+\int _{\Omega }{c_{\varepsilon }}\le \lambda \quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }). \end{aligned}$$
(3.4)

Lemma 3.3

Let \(\alpha >\frac{1}{3}\). Then there exists \(C>0\) independent of \(\varepsilon \) such that the solution of (2.6) satisfies

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega } n_{\varepsilon }^{2\alpha }+\int _{\Omega } c_{\varepsilon }^2+\int _{\Omega } | {u_{\varepsilon }}|^2\le C\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }).}\\ \end{array} \end{aligned}$$
(3.5)

Moreover, for \(T\in (0, T_{max,\varepsilon })\), one can find a constant \(C > 0\) independent of \(\varepsilon \) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\int _{\Omega } \left[ n_{\varepsilon }^{2\alpha -2} |\nabla {n_{\varepsilon }}|^2+ |\nabla {c_{\varepsilon }}|^2+ |\nabla {u_{\varepsilon }}|^2\right] \le C.}\\ \end{array} \end{aligned}$$
(3.6)

Proof

The proof consists of two cases.

Case \(2\alpha \ne 1\): We first obtain from \(\nabla \cdot u_\varepsilon =0\) in \(\Omega \times (0, T_{max,\varepsilon })\) and straightforward calculations that

$$\begin{aligned}&\displaystyle {\mathrm{sign}(2\alpha -1)\frac{1}{{2\alpha }}\frac{d}{dt}\Vert n_{\varepsilon } \Vert ^{{{2\alpha }}}_{L^{{2\alpha }}(\Omega )}}\displaystyle {+ \mathrm{sign}(2\alpha -1)(2\alpha -1)\int _{\Omega } n_{\varepsilon }^{2\alpha -2} |\nabla n_{\varepsilon }|^2}\nonumber \\&\quad = \displaystyle {- \int _{\Omega }\mathrm{sign}(2\alpha -1) n_{\varepsilon }^{2\alpha -1}\nabla \cdot (n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\cdot \nabla c_{\varepsilon })}\nonumber \\&\quad \le \displaystyle {\mathrm{sign}(2\alpha -1)(2\alpha -1) \int _{\Omega } n_{\varepsilon }^{2\alpha -2}n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })|S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })||\nabla n_{\varepsilon }||\nabla c_{\varepsilon }|} \end{aligned}$$
(3.7)

for all \(t\in (0, T_{max,\varepsilon }).\) Therefore, from (3.1), in light of (1.5) and (2.7), we can estimate the right-hand side of (3.7) as follows:

$$\begin{aligned}&\displaystyle {\mathrm{sign}(2\alpha -1)(2\alpha -1) \int _{\Omega } n_{\varepsilon }^{2\alpha -2}n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })|S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })||\nabla n_{\varepsilon }||\nabla c_{\varepsilon }|}\nonumber \\&\quad \le \displaystyle {\mathrm{sign}(2\alpha -1)(2\alpha -1) \int _{\Omega } n_{\varepsilon }^{2\alpha -2}n_{\varepsilon }C_S(1 + n_{\varepsilon })^{-\alpha }|\nabla n_{\varepsilon }||\nabla c_{\varepsilon }|}\nonumber \\&\quad \le \displaystyle {\mathrm{sign}(2\alpha -1)\frac{2\alpha -1}{2}\int _{\Omega } n_{\varepsilon }^{2\alpha -2} |\nabla n_{\varepsilon }|^2}\nonumber \\&\qquad \displaystyle {+\,\frac{|2\alpha -1|}{2}C_S^2\int _{\Omega } n_{\varepsilon }^{2\alpha -2}n_{\varepsilon }^2(1 + n_{\varepsilon })^{-2\alpha }|\nabla c_{\varepsilon }|^2}\nonumber \\&\quad \le \displaystyle {\mathrm{sign}(2\alpha -1)\frac{2\alpha -1}{2}\int _{\Omega } n_{\varepsilon }^{2\alpha -2} |\nabla n_{\varepsilon }|^2}\nonumber \\&\qquad \displaystyle {+\,\frac{|2\alpha -1|}{2}C_S^2\int _{\Omega }|\nabla c_{\varepsilon }|^2~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon })} \end{aligned}$$
(3.8)

by using Young’s inequality, where in the last inequality we have used the fact that \( n_{\varepsilon }^{2\alpha -2}n_{\varepsilon }^2(1 + n_{\varepsilon })^{-2\alpha }\le 1\) for all \(\varepsilon \ge 0\) and \((x,t)\in \Omega \times (0, T_{max,\varepsilon }).\) Inserting (3.8) into (3.7), we conclude that

$$\begin{aligned}&\displaystyle {\mathrm{sign}(2\alpha -1)\frac{1}{{2\alpha }}\frac{d}{dt}\Vert {n_{\varepsilon }}\Vert ^{{{2\alpha }}}_{L^{{2\alpha }}(\Omega )}+ \mathrm{sign}(2\alpha -1)\frac{2\alpha -1}{2}\int _{\Omega } n_{\varepsilon }^{2\alpha -2} |\nabla n_{\varepsilon }|^2}\nonumber \\&\quad \le \displaystyle {\frac{|2\alpha -1|}{2}C_S^2\int _{\Omega }|\nabla c_{\varepsilon }|^2 \quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(3.9)

To track the time evolution of \(c_\varepsilon \), taking \({c_{\varepsilon }}\) as the test function for the second equation of (2.6) and using \(\nabla \cdot u_\varepsilon =0\) and (3.3) together with Hölder’s inequality yields

$$\begin{aligned}&\displaystyle \frac{1}{{2}}\displaystyle \frac{d}{dt}\Vert {c_{\varepsilon }}\Vert ^{{{2}}}_{L^{{2}}(\Omega )}+ \int _{\Omega } |\nabla c_{\varepsilon }|^2+ \int _{\Omega } | c_{\varepsilon }|^2= \displaystyle {\int _{\Omega } F_{\varepsilon }(n_{\varepsilon })c_{\varepsilon }}\nonumber \\&\quad \le \displaystyle {\int _{\Omega } n_{\varepsilon }c_{\varepsilon }} \le \displaystyle {\Vert n_{\varepsilon }\Vert _{L^{\frac{6}{5}}(\Omega )}\Vert c_{\varepsilon }\Vert _{L^{6}(\Omega )}~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(3.10)

By applying Sobolev embedding \(W^{1,2}(\Omega ){\hookrightarrow }L^6(\Omega )\) in the three-dimensional setting, in view of (3.4), there exist positive constants \(C_1\) and \(C_2\) such that

$$\begin{aligned} \displaystyle \Vert c_{\varepsilon }\Vert _{L^{6}(\Omega )}^2&\le \displaystyle {C_1\Vert \nabla c_{\varepsilon }\Vert _{L^{2}(\Omega )}^2+C_1\Vert c_{\varepsilon }\Vert _{L^{1}(\Omega )}^2}\nonumber \\&\le \displaystyle {C_1\Vert \nabla c_{\varepsilon }\Vert _{L^{2}(\Omega )}^2+C_2~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(3.11)

Thus, by means of Young’s inequality and (3.11), we proceed to estimate

$$\begin{aligned}&\displaystyle \frac{1}{{2}}\displaystyle \frac{d}{dt}\Vert {c_{\varepsilon }}\Vert ^{{{2}}}_{L^{{2}}(\Omega )}+ \int _{\Omega } |\nabla c_{\varepsilon }|^2+ \int _{\Omega } | c_{\varepsilon }|^2 \le \displaystyle {\frac{1}{2C_1}\Vert c_{\varepsilon }\Vert _{L^{6}(\Omega )}^2+\frac{C_1}{2}\Vert n_{\varepsilon } \Vert _{L^{\frac{6}{5}}(\Omega )}^2}\nonumber \\&\quad \le \displaystyle {\frac{1}{2}\Vert \nabla c_{\varepsilon }\Vert _{L^{2}(\Omega )}^2+\frac{C_1}{2}\Vert n_{\varepsilon } \Vert _{L^{\frac{6}{5}}(\Omega )}^2+C_3 ~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon })} \end{aligned}$$
(3.12)

and some positive constant \(C_3\) independent of \(\varepsilon \). Therefore,

$$\begin{aligned} \begin{array}{rl} \displaystyle \frac{1}{{2}}\displaystyle \frac{d}{dt}\Vert {c_{\varepsilon }}\Vert ^{{{2}}}_{L^{{2}}(\Omega )}+\frac{1}{2} \int _{\Omega } |\nabla c_{\varepsilon }|^2+ \int _{\Omega } | c_{\varepsilon }|^2&{} \le \displaystyle {\frac{C_1}{2}\Vert n_{\varepsilon } \Vert _{L^{\frac{6}{5}}(\Omega )}^2+C_3 \quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }).}\\ \end{array}\nonumber \\ \end{aligned}$$
(3.13)

To estimate \(\Vert n_{\varepsilon }\Vert _{L^{\frac{6}{5}(\Omega )}}\) for all \(t\in (0, T_{max,\varepsilon })\), we should notice that \( \alpha >\frac{1}{3}\) ensures that \(\frac{2}{6\alpha -1}<2\), so that, in light of (3.4), the Gagliardo–Nirenberg inequality and Young’s inequality allow us to estimate that

$$\begin{aligned} \displaystyle \Vert n_{\varepsilon } \Vert _{L^{\frac{6}{5}}(\Omega )}^2&= \displaystyle {\Vert n_{\varepsilon }^{\alpha }\Vert _{L^{\frac{6}{5\alpha }}(\Omega )}^{\frac{2}{\alpha }}} \le \displaystyle {C_4\left( \Vert \nabla n_{\varepsilon }^{\alpha }\Vert _{L^{2}(\Omega )}^{\frac{2}{6\alpha -1}}\Vert n_{\varepsilon }^{\alpha }\Vert _{L^{\frac{1}{\alpha }}(\Omega )}^{\frac{2}{\alpha }-\frac{2}{6\alpha -1}}+\Vert n_{\varepsilon }^{\alpha }\Vert _{L^{\frac{1}{\alpha }}(\Omega )}^{\frac{2}{\alpha }}\right) }\nonumber \\&\le \displaystyle {\frac{1}{4}\frac{1}{C_1\alpha ^2C_S^2}\Vert \nabla n_{\varepsilon }^{\alpha }\Vert _{L^{2}(\Omega )}^{2}+C_5~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon })} \end{aligned}$$
(3.14)

with some positive constants \(C_4\) and \(C_5\) independent of \(\varepsilon \). This together with (3.13) contributes to

$$\begin{aligned}&\displaystyle \frac{1}{{2}}\displaystyle \frac{d}{dt}\Vert {c_{\varepsilon }}\Vert ^{{{2}}}_{L^{{2}}(\Omega )}+\frac{1}{2} \int _{\Omega } |\nabla c_{\varepsilon }|^2+ \int _{\Omega } | c_{\varepsilon }|^2\nonumber \\&\quad \le \displaystyle {\frac{1}{8}\frac{1}{\alpha ^2C_S^2}\Vert \nabla n_{\varepsilon }^{\alpha }\Vert _{L^{2}(\Omega )}^{2}+C_6 \quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })} \end{aligned}$$
(3.15)

and some positive constant \(C_6\). Taking an evident linear combination of the inequalities provided by (3.9) and (3.15), one can obtain

$$\begin{aligned}&\displaystyle \mathrm{sign}(2\alpha -1)\displaystyle \frac{1}{{2\alpha }}\displaystyle \frac{d}{dt}\Vert { n_{\varepsilon } }\Vert ^{{{2\alpha }}}_{L^{{2\alpha }}(\Omega )}+ |2\alpha -1|C_S^2\displaystyle \frac{d}{dt}\Vert {c_{\varepsilon }}\Vert ^{{{2}}}_{L^{{2}}(\Omega )}\nonumber \\&\qquad \displaystyle +\,\displaystyle \frac{|2\alpha -1|}{2}C_S^2 \int _{\Omega } |\nabla c_{\varepsilon }|^2+ 2|2\alpha -1|C_S^2\int _{\Omega } | c_{\varepsilon }|^2\nonumber \\&\qquad \displaystyle +\,\left( \mathrm{sign}(2\alpha -1)\displaystyle \frac{2\alpha -1}{2}-\frac{1}{4}|2\alpha -1|\right) \displaystyle \int _{\Omega } n_{\varepsilon }^{2\alpha -2} |\nabla n_{\varepsilon }|^2\nonumber \\&\quad \le \displaystyle {C_7 ~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon })} \end{aligned}$$
(3.16)

and some positive constant \(C_7.\) Since \(\mathrm{sign}(2\alpha -1)\displaystyle \frac{2\alpha -1}{2}=\displaystyle \frac{|2\alpha -1|}{2},\) (3.16) implies that

$$\begin{aligned} \displaystyle&\mathrm{sign}(2\alpha -1)\displaystyle \frac{1}{{2\alpha }}\displaystyle \frac{d}{dt}\Vert { n_{\varepsilon } }\Vert ^{{{2\alpha }}}_{L^{{2\alpha }}(\Omega )}+ |2\alpha -1|C_S^2\displaystyle \frac{d}{dt}\Vert {c_{\varepsilon }}\Vert ^{{{2}}}_{L^{{2}}(\Omega )}\nonumber \\&\qquad \displaystyle +\,\displaystyle \frac{|2\alpha -1|}{2}C_S^2 \int _{\Omega } |\nabla c_{\varepsilon }|^2+ 2|2\alpha -1|C_S^2\int _{\Omega } | c_{\varepsilon }|^2\nonumber \\&\qquad +\,\displaystyle \frac{|2\alpha -1|}{4}\displaystyle \int _{\Omega } n_{\varepsilon }^{2\alpha -2} |\nabla n_{\varepsilon }|^2\nonumber \\&\quad \le \displaystyle {C_7 ~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(3.17)

If \(2\alpha >1\), then \(\mathrm{sign}(2\alpha -1)=1>0,\) thus, integrating (3.17) over time, we can obtain

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega } n_{\varepsilon }^{2\alpha }+\int _{\Omega } c_{\varepsilon }^2\le C_8\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.18)

and

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\int _{\Omega } \left[ n_{\varepsilon }^{2\alpha -2} |\nabla {n_{\varepsilon }}|^2+ |\nabla {c_{\varepsilon }}|^2\right] \le C_8(T+1)\quad \text{ for } \text{ all }\;\; T\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.19)

and some positive constant \(C_8.\) If \(2\alpha <1\), then \(\mathrm{sign}(2\alpha -1)=-1<0\); hence, in view of (3.4), integrating (3.17) over time and employing Hölder’s inequality, we also conclude that there exists a positive constant \(C_9\) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega } n_{\varepsilon }^{2\alpha }+\int _{\Omega } c_{\varepsilon }^2\le C_9\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.20)

and

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\int _{\Omega } \left[ n_{\varepsilon }^{2\alpha -2} |\nabla {n_{\varepsilon }}|^2+ |\nabla {c_{\varepsilon }}|^2\right] \le C_9(T+1)\quad \text{ for } \text{ all }\;\; T\in (0, T_{max,\varepsilon }).}\\ \end{array} \end{aligned}$$
(3.21)

Case \(2\alpha =1\): Using the first equation of (2.6) and (2.7), integrating by parts, and applying (1.5) and (3.1), we obtain

$$\begin{aligned} \begin{array}{rl} \frac{d}{dt}\int _{\Omega } n_{\varepsilon } \ln n_{\varepsilon }&{}= \displaystyle {\int _{\Omega }n_{\varepsilon t} \ln n_{\varepsilon }+\int _{\Omega }n_{\varepsilon t}}\\ &{}= \displaystyle {\int _{\Omega }\Delta n_{\varepsilon } \ln n_{\varepsilon }- \int _{\Omega }\ln n_{\varepsilon }\nabla \cdot (n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon }) \cdot \nabla c_{\varepsilon })}\\ &{}\le -\int _{\Omega } \frac{|\nabla n_{\varepsilon }|^2}{n_{\varepsilon }}+ \int _{\Omega } C_S(1 + n_{\varepsilon })^{-\alpha }\frac{n_{\varepsilon }}{ n_{\varepsilon } }|\nabla n_{\varepsilon }||\nabla c_{\varepsilon }|\quad \text{ for } \text{ all }\;\;t\in (0, T_{max,\varepsilon }), \end{array} \end{aligned}$$

which combined with Young’s inequality and \(2\alpha =1\) implies that

$$\begin{aligned} \frac{d}{dt}\int _{\Omega } n_{\varepsilon } \ln n_{\varepsilon } +\frac{1}{2}\int _{\Omega } \frac{|\nabla n_{\varepsilon }|^2}{n_{\varepsilon }} \le \frac{1}{2}C_S^2\int _{\Omega }|\nabla c_{\varepsilon }|^2\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }). \end{aligned}$$

However, since \(2\alpha =1\) yields \(\alpha >\frac{1}{3}\), by employing almost exactly the same arguments as in the proof of (3.10)–(3.16) (with the minor necessary changes being left as an easy exercise to the reader), we conclude an estimate of

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega } n_{\varepsilon } \ln n_{\varepsilon } +\int _{\Omega } c_{\varepsilon }^2\le C_{10}\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.22)

and

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\int _{\Omega } \left[ \frac{|\nabla n_{\varepsilon }|^2}{n_{\varepsilon }}+ |\nabla {c_{\varepsilon }}|^2\right] \le C_{10}(T+1)\quad \text{ for } \text{ all }\;\; T\in (0, T_{max,\varepsilon }).}\\ \end{array} \end{aligned}$$
(3.23)

Now, multiplying the third equation of (2.6) by \(u_\varepsilon \), integrating by parts, and using \(\nabla \cdot u_{\varepsilon }=0\) give

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\int _{\Omega }{|u_{\varepsilon }|^2}+\int _{\Omega }{|\nabla u_{\varepsilon }|^2}= \int _{\Omega }n_{\varepsilon }u_{\varepsilon }\cdot \nabla \phi \quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }). \end{aligned}$$
(3.24)

Here we use Hölder’s inequality, Young’s inequality, (1.7), and the continuity of the embedding \(W^{1,2}(\Omega )\hookrightarrow L^6(\Omega )\) to find \(C_{11} \) and \(C_{12}> 0\) such that

$$\begin{aligned} \displaystyle \int _{\Omega }n_{\varepsilon }u_{\varepsilon }\cdot \nabla \phi&\le \displaystyle {\Vert \nabla \phi \Vert _{L^\infty (\Omega )}\Vert n_{\varepsilon } \Vert _{L^{\frac{6}{5}}(\Omega )}\Vert u_{\varepsilon }\Vert _{L^{6}(\Omega )}}\nonumber \\&\le \displaystyle {C_{11}\Vert \nabla \phi \Vert _{L^\infty (\Omega )}\Vert n_{\varepsilon } \Vert _{L^{\frac{6}{5}}(\Omega )}\Vert \nabla u_{\varepsilon }\Vert _{L^{2}(\Omega )}}\nonumber \\&\le \displaystyle {\frac{1}{2}\Vert \nabla u_{\varepsilon }\Vert _{L^{2}(\Omega )}^2+C_{12}\Vert n_{\varepsilon } \Vert _{L^{\frac{6}{5}}(\Omega )}^2\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(3.25)

Next, in view of (3.4) and \(\alpha >\frac{1}{3}\), (3.14) and Young’s inequality along with the Gagliardo–Nirenberg inequality yields

$$\begin{aligned} \begin{array}{rl} \displaystyle {\int _{\Omega }n_{\varepsilon }u_{\varepsilon }\cdot \nabla \phi } &{} \le \displaystyle {\frac{1}{2}\Vert \nabla u_{\varepsilon }\Vert _{L^{2}(\Omega )}^2+C_8\Vert \nabla n_{\varepsilon }^{\alpha }\Vert _{L^{2}(\Omega )}^{\frac{2}{6\alpha -1}}\Vert n_{\varepsilon }^{\alpha }\Vert _{L^{\frac{1}{\alpha }}(\Omega )}^{\frac{2}{\alpha }-\frac{2}{6\alpha -1}}}\\ &{} \le \displaystyle {\frac{1}{2}\Vert \nabla u_{\varepsilon }\Vert _{L^{2}(\Omega )}^2+\Vert \nabla n_{\varepsilon }^{\alpha }\Vert _{L^{2}(\Omega )}^{2}+C_{13}\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.26)

and some positive constant \(C_{13}.\) Now, inserting (3.25) and (3.26) into (3.24) and using (3.21) and (3.23), one has

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega } |u_{\varepsilon }|^2\le C_{14}\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.27)

and

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\int _{\Omega } |\nabla {u_{\varepsilon }}|^2\le C_{14}(T+1)\quad \text{ for } \text{ all }\;\; T\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.28)

and some positive constant \(C_{14}.\) Finally, collecting (3.20)–(3.21), (3.22)–(3.23), and (3.27)–(3.28), we can get (3.5) and (3.6). \(\square \)

With the help of Lemma 3.3, based on the Gagliardo–Nirenberg inequality and an application of well-known arguments from parabolic regularity theory, we can derive the following lemmas:

Lemma 3.4

Let \(\alpha >\frac{1}{3}\). Then there exists \(C>0\) independent of \(\varepsilon \) such that, for each \(T\in (0, T_{max,\varepsilon })\), the solution of (2.6) satisfies

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\int _{\Omega }\left[ |\nabla n_{\varepsilon }|^{\frac{3\alpha +1}{2}}+ n_{\varepsilon }^{\frac{6\alpha +2}{3}}\right] \le C(T+1)\quad \text{ if }\;\;\frac{1}{3}<\alpha \le \frac{1}{2},}\\ \end{array} \end{aligned}$$
(3.29)
$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\int _{\Omega }\left[ |\nabla n_{\varepsilon }|^{\frac{10\alpha }{3+2\alpha }}+ n_{\varepsilon }^{\frac{10\alpha }{3}}\right] \le C(T+1)\quad \text{ if }\;\;\frac{1}{2}<\alpha <1,}\\ \end{array} \end{aligned}$$
(3.30)

as well as

$$\begin{aligned} \int _{0}^T\int _{\Omega } \left[ |\nabla {n_{\varepsilon }}|^2+ n_{\varepsilon }^{\frac{10}{3}}\right] \le C(T+1)\quad \text{ if }\;\;\alpha \ge 1 \end{aligned}$$
(3.31)

and

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\left\{ \int _{\Omega }[c_{\varepsilon }^{\frac{10}{3}}+ |u_{\varepsilon }|^{\frac{10}{3}}]+\Vert u_{\varepsilon }\Vert ^2_{L^6(\Omega )}\right\} \le C(T+1).}\\ \end{array} \end{aligned}$$
(3.32)

Proof

Case \(\frac{1}{3}<\alpha \le \frac{1}{2}\): From (3.4), (3.5), and (3.6), in light of the Gagliardo–Nirenberg inequality, for some \(C_1\) and \(C_2> 0\) that are independent of \(\varepsilon \), one may verify that

$$\begin{aligned} \displaystyle \int _{0}^T\displaystyle \int _{\Omega } n_{\varepsilon }^{\frac{6\alpha +2}{3}}&= \displaystyle {\int _{0}^T\Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{6\alpha +2}{3\alpha }}}_{L^{\frac{6\alpha +2}{3\alpha }}(\Omega )}}\nonumber \\&\le \displaystyle {C_{1}\int _{0}^T\left( \Vert \nabla { n_{\varepsilon }^{\alpha }}\Vert ^{2}_{L^{2}(\Omega )}\Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{2}{3\alpha }}}_{L^{\frac{1}{\alpha }}(\Omega )}+ \Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{6\alpha +2}{3\alpha }}}_{L^{\frac{1}{\alpha }}(\Omega )}\right) }\nonumber \\&\le \displaystyle {C_{2}(T+1)\quad \text{ for } \text{ all }\;\; T > 0.} \end{aligned}$$
(3.33)

Therefore, employing Hölder’s inequality (with two exponents \(\frac{4}{3\alpha +1}\) and \(\frac{4}{3-3\alpha }\)), we conclude that there exists a positive constant \(C_3\) such that

$$\begin{aligned} \displaystyle \int _{0}^T\displaystyle \int _{\Omega }|\nabla n_{\varepsilon }|^{\frac{3\alpha +1}{2}}&\le \displaystyle {\left[ \int _{0}^T\displaystyle \int _{\Omega } n_{\varepsilon }^{2\alpha -2}|\nabla n_{\varepsilon }|^2\right] ^{\frac{3\alpha +1}{4}} \left[ \int _{0}^T\displaystyle \int _{\Omega } n_{\varepsilon }^{\frac{6\alpha +2}{3}}\right] ^{\frac{3-3\alpha }{4}} }\nonumber \\&\le \displaystyle {C_{3}(T+1)\quad \text{ for } \text{ all }\;\; T > 0.} \end{aligned}$$
(3.34)

Case \(\frac{1}{2}<\alpha <1\): Again by (3.4), (3.5), and (3.6) and the Gagliardo–Nirenberg inequality and Hölder’s inequality (with two exponents \(\frac{3+2\alpha }{5\alpha }\) and \(\frac{3+2\alpha }{3-3\alpha }\)), we derive that there exist positive constants \(C_4\), \(C_5\), and \(C_6\) such that

$$\begin{aligned} \displaystyle \int _{0}^T\displaystyle \int _{\Omega } n_{\varepsilon }^{\frac{10\alpha }{3}}&= \displaystyle {\int _{0}^T\Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{10}{3 }}}_{L^{\frac{10}{3 }}(\Omega )}}\nonumber \\&\le \displaystyle {C_{4}\int _{0}^T\left( \Vert \nabla { n_{\varepsilon }^{\alpha }}\Vert ^{2}_{L^{2}(\Omega )}\Vert { n_{\varepsilon }^{\alpha }} \Vert ^{{\frac{4 }{3}}}_{L^{2}(\Omega )}+ \Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{10\alpha }{3}}}_{L^{2}(\Omega )}\right) }\nonumber \\&\le \displaystyle {C_{5}(T+1)\quad \text{ for } \text{ all }\;\; T > 0} \end{aligned}$$
(3.35)

and

$$\begin{aligned} \displaystyle \int _{0}^T\displaystyle \int _{\Omega }|\nabla n_{\varepsilon }|^{\frac{10\alpha }{3+2\alpha }}&\le \displaystyle {\left[ \int _{0}^T\displaystyle \int _{\Omega } n_{\varepsilon }^{2\alpha -2}|\nabla n_{\varepsilon }|^2\right] ^{\frac{5\alpha }{3+2\alpha }} \left[ \int _{0}^T\displaystyle \int _{\Omega } n_{\varepsilon }^{\frac{10\alpha }{3}}\right] ^{\frac{3-3\alpha }{3+2\alpha }}}\nonumber \\&\le \displaystyle {C_{6}(T+1)\quad \text{ for } \text{ all }\;\; T > 0.} \end{aligned}$$
(3.36)

Case \(\alpha \ge 1\): Multiplying the first equation in (2.6) by \( n_{\varepsilon }\), in view of (2.7) and using \(\nabla \cdot u_\varepsilon =0\), we derive

$$\begin{aligned}&\displaystyle {\frac{1}{{2}}\frac{d}{dt}\Vert { n_{\varepsilon } }\Vert ^{{{{2}}}}_{L^{{{2}}}(\Omega )}+ \int _{\Omega } |\nabla n_{\varepsilon }|^2}= \displaystyle {- \int _{\Omega } n_{\varepsilon }\nabla \cdot (n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\cdot \nabla c_{\varepsilon })}\nonumber \\&\quad \le \displaystyle { \int _{\Omega } n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })|S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })||\nabla n_{\varepsilon }||\nabla c_{\varepsilon }|\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(3.37)

Recalling (1.5) and (2.7) and using \(\alpha \ge 1\), via Young’s inequality, we derive

$$\begin{aligned}&\displaystyle \int _{\Omega } n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })|S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })||\nabla n_{\varepsilon }||\nabla c_{\varepsilon }|\le \displaystyle {C_S\int _{\Omega } |\nabla n_{\varepsilon }||\nabla c_{\varepsilon }|}\nonumber \\&\quad \le \displaystyle {\frac{1}{2}\int _{\Omega } |\nabla n_{\varepsilon }|^2+\frac{C_S^2}{2}\int _{\Omega } |\nabla c_{\varepsilon }|^2\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(3.38)

Here we have used the fact that

$$\begin{aligned} n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })|S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })|\le C_S n_{\varepsilon }(1 + n_{\varepsilon })^{-1}\le C_S \end{aligned}$$

by using (1.5). Therefore, collecting (3.37) and (3.38) and using (3.6), we conclude that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega } n_{\varepsilon }^{2}\le C_{7}\quad \text{ for } \text{ all }\;\;t\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(3.39)

and

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^{ T}\int _{\Omega } |\nabla {n_{\varepsilon }}|^2\le C_{7}(T+1).}\\ \end{array} \end{aligned}$$
(3.40)

Hence, from (3.39)–(3.40) and (3.5)–(3.6), in light of the Gagliardo–Nirenberg inequality, we derive that there exist positive constants \(C_{8}\), \(C_{9}\), \(C_{10}\), \(C_{11}\), \(C_{12}\), \(C_{13}\), \(C_{14}\), \(C_{15}\), \(C_{16}\) and \(C_{17}\) such that

$$\begin{aligned} \displaystyle \int _{0}^T\displaystyle \int _{\Omega } n_{\varepsilon }^{\frac{10}{3}}&\le \displaystyle {C_{8}\int _{0}^T\left( \Vert \nabla { n_{\varepsilon }}\Vert ^{2}_{L^{2}(\Omega )}\Vert { n_{\varepsilon } }\Vert ^{{\frac{4}{3}}}_{L^{2}(\Omega )}+ \Vert { n_{\varepsilon } }\Vert ^{{\frac{10}{3}}}_{L^{2}(\Omega )}\right) }\nonumber \\&\le \displaystyle {C_{9}(T+1)\quad \text{ for } \text{ all }\;\; T > 0,} \end{aligned}$$
(3.41)
$$\begin{aligned} \displaystyle \int _{0}^T\displaystyle \int _{\Omega } c_{\varepsilon }^{\frac{10}{3}}&\le \displaystyle {C_{10}\int _{0}^T\left( \Vert \nabla { c_{\varepsilon }}\Vert ^{2}_{L^{2}(\Omega )}\Vert { c_{\varepsilon }}\Vert ^{{\frac{4}{3}}}_{L^{2}(\Omega )}+ \Vert { c_{\varepsilon }}\Vert ^{{\frac{10}{3}}}_{L^{2}(\Omega )}\right) }\nonumber \\&\le \displaystyle {C_{11}(T+1)\quad \text{ for } \text{ all }\;\; T > 0} \end{aligned}$$
(3.42)

as well as

$$\begin{aligned} \displaystyle \int _{0}^T\displaystyle \int _{\Omega } |u_{\varepsilon }|^{\frac{10}{3}}&\le \displaystyle {C_{14}\int _{0}^T\left( \Vert \nabla { u_{\varepsilon }}\Vert ^{2}_{L^{2}(\Omega )}\Vert { u_{\varepsilon }}\Vert ^{{\frac{4}{3}}}_{L^{2}(\Omega )}+ \Vert { u_{\varepsilon }}\Vert ^{{\frac{10}{3}}}_{L^{2}(\Omega )}\right) }\nonumber \\&\le \displaystyle {C_{15}(T+1)\quad \text{ for } \text{ all }\;\; T > 0} \end{aligned}$$
(3.43)

and

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle \int _{0}^T\displaystyle \Vert u_{\varepsilon }\Vert ^{2}_{L^6(\Omega )} \le \displaystyle {C_{16}\int _{0}^T\Vert \nabla { u_{\varepsilon }}\Vert ^{2}_{L^{2}(\Omega )}}\le \displaystyle {C_{17}(T+1)~\text{ for } \text{ all }~ T > 0,}\\ \end{array} \end{aligned}$$
(3.44)

where the last inequality we have used the embedding \(W^{1,2}_{0,\sigma } (\Omega ) {\hookrightarrow } L^6 (\Omega )\) and the Poincaré inequality. Finally, combining (3.33)–(3.36) with (3.40)–(3.44), we can obtain the results. \(\square \)

Lemma 3.5

Let \(\frac{1}{3}<\alpha \le \frac{8}{21}\). Then there exist \(\gamma =\frac{2\alpha +\frac{2}{3}}{\alpha +1}\in (1,2)\) and \(C > 0\) independent of \(\varepsilon \) such that, for each \(T\in (0, T_{max,\varepsilon })\), the solution of (2.6) satisfies

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\Vert n_{\varepsilon }\Vert ^{\frac{2\gamma }{2-\gamma }}_{L^\frac{6\gamma }{6-\gamma }(\Omega )}\le C(T+1).}\\ \end{array} \end{aligned}$$
(3.45)

Proof

To this end, we first prove that for all \(p\in (1,6\alpha ) \), then there exists a positive constant \(C_1\) independent of \(\varepsilon \) such that, for each \(T\in (0, T_{max,\varepsilon })\), the solution of (2.6) satisfies

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^T\Vert n_{\varepsilon }\Vert ^{\frac{2p(\alpha -\frac{1}{6})}{p-1}}_{L^p(\Omega )}\le C_1(T+1).}\\ \end{array} \end{aligned}$$
(3.46)

In fact, by (3.4) and (3.6), we derive that for some positive constants \(C_{2}\) and \(C_3\) independent of \(\varepsilon \) such that

$$\begin{aligned} \begin{array}{rl} \displaystyle \int _{0}^T\Vert n_{\varepsilon }\Vert ^{\frac{2p(\alpha -\frac{1}{6})}{p-1}}_{L^p(\Omega )} &{}= \displaystyle {\int _{0}^T\Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{2p}{p-1 }\cdot \frac{6\alpha -1}{6\alpha }}}_{L^{\frac{p}{\alpha }}(\Omega )}}\\ &{}\le \displaystyle {C_{2}\int _{0}^T\left( \Vert \nabla { n_{\varepsilon }^{\alpha }}\Vert ^{2}_{L^{2}(\Omega )}\Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{2p}{p-1 }\cdot \frac{6\alpha -1}{6\alpha }-2}}_{L^{\frac{1}{\alpha }}(\Omega )}+ \Vert { n_{\varepsilon }^{\alpha }}\Vert ^{{\frac{2p}{p-1 }\cdot \frac{6\alpha -1}{6\alpha }}}_{L^{\frac{1}{\alpha }}(\Omega )}\right) }\\ &{}\le \displaystyle {C_{3}(T+1)\quad \text{ for } \text{ all }\;\; T > 0.}\\ \end{array} \end{aligned}$$

Therefore, (3.46) holds. Next, by \(\alpha \in (\frac{1}{3},\frac{8}{21}],\) we may choose \(\gamma =\frac{2\alpha +\frac{2}{3}}{\alpha +1}\) such that

$$\begin{aligned} 1<\gamma <\min \left\{ \frac{6\alpha }{\alpha +1},2\right\} \end{aligned}$$
(3.47)

as well as

$$\begin{aligned} p:=\frac{6\gamma }{6-\gamma }\in (1,6\alpha ) \end{aligned}$$
(3.48)

and

$$\begin{aligned} \frac{2p(\alpha -\frac{1}{6})}{p-1}=\frac{12\gamma (\alpha -\frac{1}{6})}{7\gamma -6}>\frac{2\gamma }{2-\gamma }. \end{aligned}$$
(3.49)

Collecting (3.46)–(3.49), one can derive (3.45) by using the Young inequality. \(\square \)

4 Global solvability of the regularized problem (2.6)

The main task of this section is to prove the global solvability of the regularized problem (2.6). To this end, first, we need to establish some \(\varepsilon \)-dependent estimates for \(n_{\varepsilon }\), \(c_{\varepsilon },\) and \(u_{\varepsilon }\).

4.1 A priori estimates for the regularized problem (2.6) that depend on \(\varepsilon \)

In this subsection, we obtain some regularity properties for \(n_\varepsilon ,\) \(c_\varepsilon ,\) and \(u_\varepsilon \) in the following form on the basis of Lemma 3.3.

Lemma 4.1

Let \(\alpha >\frac{1}{3}\). Then there exists \(C=C(\varepsilon )>0\) depending on \(\varepsilon \) such that the solution of (2.6) satisfies

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega }n^{2\alpha +2}_{\varepsilon }+\int _{\Omega } | \nabla {u_{\varepsilon }}|^2\le C\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }).}\\ \end{array} \end{aligned}$$
(4.1)

In addition, for each \(T\in (0, T_{max,\varepsilon }]\) with \(T < \infty \), one can find a constant \(C > 0\) depending on \(\varepsilon \) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^{ T}\int _{\Omega } \left[ n_{\varepsilon }^{2\alpha } |\nabla {n_{\varepsilon }}|^2+ |\Delta {u_{\varepsilon }}|^2\right] \le C.}\\ \end{array} \end{aligned}$$
(4.2)

Proof

In view of (2.7), we derive

$$\begin{aligned} F'_{\varepsilon }(n_{\varepsilon })\le \displaystyle \frac{1}{\varepsilon n_{\varepsilon }}, \end{aligned}$$

so that, by multiplying the first equation in (2.6) by \( n_{\varepsilon }^{1+2\alpha }\), using \(\nabla \cdot u_\varepsilon =0\), and applying the same argument as in the proof of (3.7)–(3.20), one can obtain that there exist positive constants \(C_1\) and \(C_2\) depending on \(\varepsilon \) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{\Omega }n^{2\alpha +2}_{\varepsilon }\le C_1\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(4.3)

and

$$\begin{aligned} \int _{0}^{ T}\int _{\Omega } n_{\varepsilon }^{2\alpha } |\nabla {n_{\varepsilon }}|^2\le C_2\quad \text{ for } \text{ all }\;\;T\in (0, T_{max,\varepsilon }]\;\;\text{ with }\;\;T < \infty . \end{aligned}$$

Now, from \(D(1 + \varepsilon A) :=W^{2,2}(\Omega ) \cap W_{0,\sigma }^{1,2}(\Omega )\hookrightarrow L^\infty (\Omega )\) and (3.5), it follows that, for some \(C_3> 0\) and \(C_4 > 0\),

$$\begin{aligned} \Vert Y_{\varepsilon }u_{\varepsilon }\Vert _{L^\infty (\Omega )}= & {} \Vert (I+\varepsilon A)^{-1}u_{\varepsilon }\Vert _{L^\infty (\Omega )}\nonumber \\\le & {} C_3\Vert u_{\varepsilon }(\cdot ,t)\Vert _{L^2(\Omega )}\le C_4 \quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }). \end{aligned}$$
(4.4)

Next, testing the projected Stokes equation \(u_{\varepsilon t} +Au_{\varepsilon } = {\mathcal {P}}[-\kappa (Y_{\varepsilon }u_{\varepsilon } \cdot \nabla )u_{\varepsilon }+n_{\varepsilon }\nabla \phi ]\) by \(Au_{\varepsilon }\), we derive

$$\begin{aligned}&\displaystyle {\frac{1}{{2}}\frac{d}{dt}\Vert A^{\frac{1}{2}}u_{\varepsilon }\Vert ^{{{2}}}_{L^{{2}}(\Omega )}+ \int _{\Omega }|Au_{\varepsilon }|^2 }= \displaystyle { \int _{\Omega }Au_{\varepsilon }{\mathcal {P}}(-\kappa (Y_{\varepsilon }u_{\varepsilon } \cdot \nabla )u_{\varepsilon })+ \int _{\Omega }{\mathcal {P}}(n_{\varepsilon }\nabla \phi ) Au_{\varepsilon }}\nonumber \\&\quad \le \displaystyle { \frac{1}{2}\int _{\Omega }|Au_{\varepsilon }|^2+\kappa ^2\int _{\Omega } |(Y_{\varepsilon }u_{\varepsilon } \cdot \nabla )u_{\varepsilon }|^2+ \Vert \nabla \phi \Vert ^2_{L^\infty (\Omega )}\int _{\Omega }n_{\varepsilon }^2\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).} \nonumber \\ \end{aligned}$$
(4.5)

However, in light of the Gagliardo–Nirenberg inequality, Young’s inequality, and (4.4), there exists a positive constant \(C_5\) such that

$$\begin{aligned} \kappa ^2\displaystyle \int _{\Omega } |(Y_{\varepsilon }u_{\varepsilon } \cdot \nabla )u_{\varepsilon }|^2&\le \displaystyle { \kappa ^2\Vert Y_{\varepsilon }u_{\varepsilon }\Vert ^2_{L^\infty (\Omega )}\int _{\Omega }|\nabla u_{\varepsilon }|^2}\nonumber \\&\le \displaystyle { \kappa ^2\Vert Y_{\varepsilon }u_{\varepsilon }\Vert ^2_{L^\infty (\Omega )}\int _{\Omega }|\nabla u_{\varepsilon }|^2}\nonumber \\&\le \displaystyle { C_5\int _{\Omega }|\nabla u_{\varepsilon }|^2\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).} \end{aligned}$$
(4.6)

Here we have used the well-known fact that \(\Vert A(\cdot )\Vert _{L^{2}(\Omega )}\) defines a norm equivalent to \(\Vert \cdot \Vert _{W^{2,2}(\Omega )}\) on D(A) (see Theorem 2.1.1 of [19]). Now, recall that \(\Vert A^{\frac{1}{2}}u_{\varepsilon }\Vert ^{{{2}}}_{L^{{2}}(\Omega )} = \Vert \nabla u_{\varepsilon }\Vert ^{{{2}}}_{L^{{2}}(\Omega )}.\) Substituting (4.6) into (4.5) yields

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\frac{1}{{2}}\frac{d}{dt}\Vert \nabla u_{\varepsilon }\Vert ^{{{2}}}_{L^{{2}}(\Omega )}+ \int _{\Omega }|\Delta u_{\varepsilon }|^2 \le C_6\int _{\Omega }|\nabla u_{\varepsilon }|^2+ \Vert \nabla \phi \Vert ^2_{L^\infty (\Omega )}\int _{\Omega } n_{\varepsilon }^2\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).}\\ \end{array}\nonumber \\ \end{aligned}$$
(4.7)

Since \(\alpha >\frac{1}{3}\) yields \(2\alpha +2>\frac{8}{3}>2,\) by collecting (4.3) and (4.7) and performing some basic calculations, we can get the results. \(\square \)

Lemma 4.2

Under the assumptions of Theorem 1.1, one can find that there exists \(C=C(\varepsilon )> 0\) depending on \(\varepsilon \) such that

$$\begin{aligned} \int _{\Omega }{|\nabla c_{\varepsilon }(\cdot ,t)|^2}\le C\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon }) \end{aligned}$$
(4.8)

and

$$\begin{aligned} \int _0^{T}\int _{\Omega }{|\Delta c_{\varepsilon }|^2}\le C\quad \text{ for } \text{ all }\;\; T\in (0, T_{max,\varepsilon }]\;\;\text{ with }\;\; T < \infty . \end{aligned}$$
(4.9)

Proof

First, testing the second equation in (2.6) against \(-\Delta c_{\varepsilon }\), employing Young’s inequality, and using (3.3) yields

$$\begin{aligned} \displaystyle {\frac{1}{{2}}\frac{d}{dt} \Vert \nabla c_{\varepsilon }\Vert ^{{{2}}}_{L^{{2}}(\Omega )}}&= \displaystyle {\int _{\Omega } -\Delta c_{\varepsilon }(\Delta c_{\varepsilon }-c_{\varepsilon }+F_{\varepsilon }(n_{\varepsilon })-u_{\varepsilon }\cdot \nabla c_{\varepsilon })}\nonumber \\&= \displaystyle {-\int _{\Omega } |\Delta c_{\varepsilon }|^2-\int _{\Omega } |\nabla c_{\varepsilon }|^{2}-\int _\Omega F_{\varepsilon }(n_{\varepsilon })\Delta c_{\varepsilon }-\int _\Omega (u_{\varepsilon }\cdot \nabla c_{\varepsilon })\Delta c_{\varepsilon }}\nonumber \\&\le \displaystyle {-\frac{1}{4}\int _{\Omega } |\Delta c_{\varepsilon }|^2-\int _{\Omega } |\nabla c_{\varepsilon }|^{2}+ \int _\Omega n_{\varepsilon }^2+\int _\Omega |u_{\varepsilon }\cdot \nabla c_{\varepsilon }||\Delta c_{\varepsilon }|} \end{aligned}$$
(4.10)

for all \(t\in (0,T_{max,\varepsilon })\). Next, one needs to estimate the last term on the right-hand side of (4.10). Indeed, in view of Sobolev’s embedding theorem (\(W^{1,2}(\Omega )\hookrightarrow L^6(\Omega )\)) and applying (4.1) and (3.5), we derive from Hölder’s inequality, the Gagliardo–Nirenberg inequality, and Young’s inequality that there exist positive constants \(C_1\), \(C_2\), \(C_3\), and \(C_4\) such that

$$\begin{aligned} \displaystyle {\int _\Omega |u_{\varepsilon }\cdot \nabla c_{\varepsilon }||\Delta c_{\varepsilon }|}&\le \displaystyle {\Vert u_{\varepsilon }\Vert _{L^{6}(\Omega )}\Vert \nabla c_{\varepsilon }\Vert _{L^{3}(\Omega )}}\Vert \Delta c_{\varepsilon }\Vert _{L^{2}(\Omega )}\nonumber \\&\le \displaystyle {C_1\Vert \nabla c_{\varepsilon }\Vert _{L^{3}(\Omega )}}\Vert \Delta c_{\varepsilon }\Vert _{L^{2}(\Omega )}\nonumber \\&\le \displaystyle {C_2(\Vert \Delta c_{\varepsilon }\Vert ^{\frac{3}{4}}_{L^2(\Omega )}\Vert c_{\varepsilon }\Vert ^{\frac{1}{4}}_{L^2(\Omega )}+\Vert c_{\varepsilon }\Vert ^{2}_{L^2(\Omega )})\Vert \Delta c_{\varepsilon }\Vert _{L^{2}(\Omega )}}\nonumber \\&\le \displaystyle {C_3(\Vert \Delta c_{\varepsilon }\Vert ^{\frac{7}{4}}_{L^2(\Omega )}+\Vert \Delta c_{\varepsilon }\Vert _{L^{2}(\Omega )})}\nonumber \\&\le \displaystyle {\frac{1}{4}\Vert \Delta c_{\varepsilon }\Vert ^{2}_{L^2(\Omega )}+C_4~\text{ for } \text{ all }~ t\in (0, T_{max,\varepsilon }).} \end{aligned}$$
(4.11)

Inserting (4.11) into (4.10) and using (4.1), one obtains (4.8) and (4.9). This completes the proof of Lemma 4.2. \(\square \)

Lemma 4.3

Let \(\alpha >\frac{1}{3}\). Assume that the hypothesis of Theorem 1.1 holds. Then there exists a positive constant \(C=C(\varepsilon )\) depending on \(\varepsilon \) such that, for any \(3<q<6,\) the solution of (2.6) from Lemma 2.1 satisfies

$$\begin{aligned} \begin{array}{rl} \Vert A^\gamma u_{\varepsilon }(\cdot , t)\Vert _{L^2(\Omega )}&{} \le \displaystyle {C \quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(4.12)

as well as

$$\begin{aligned} \begin{array}{rl} \Vert u_{\varepsilon }(\cdot , t)\Vert _{L^{\infty }(\Omega )}&{} \le \displaystyle {C \quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon })}\\ \end{array} \end{aligned}$$
(4.13)

and

$$\begin{aligned} \begin{array}{rl} \Vert \nabla c_{\varepsilon }(\cdot , t)\Vert _{L^{q}(\Omega )}&{} \le \displaystyle {C\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }),}\\ \end{array} \end{aligned}$$
(4.14)

where \(\gamma \) is the same as in (1.8).

Proof

Let \(h_{\varepsilon }(x,t)={\mathcal {P}}[n_{\varepsilon }\nabla \phi -\kappa (Y_{\varepsilon }u_{\varepsilon } \cdot \nabla )u_{\varepsilon } ]\). Because \(\alpha >\frac{1}{3}\), then , along with (4.1), (1.7), and (4.4), there exist positive constants \(q_0>\frac{3}{2}\) and \(C_{1}\) such that

$$\begin{aligned} \Vert n_{\varepsilon }(\cdot ,t)\Vert _{L^{q_0}(\Omega )}\le C_{1} \quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }) \end{aligned}$$
(4.15)

and

$$\begin{aligned} \Vert h_{\varepsilon }(\cdot ,t)\Vert _{L^{q_0}(\Omega )}\le C_{1} \quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }). \end{aligned}$$
(4.16)

Hence, because \(q_0>\frac{3}{2}\), we pick an arbitrary \(\gamma \in (\frac{3}{4}, 1)\) and, then, \(-\gamma -\frac{3}{2}(\frac{1}{q_0}-\frac{1}{2})>-1\). Therefore, in view of the smoothing properties of the Stokes semigroup [7], we derive that, for some \(\lambda \), \(C_{2} > 0,\) and \(C_{3} > 0\),

$$\begin{aligned} \Vert A^\gamma u_{\varepsilon }(\cdot , t)\Vert _{L^2(\Omega )}&\le \displaystyle {\Vert A^\gamma e^{-tA}u_0\Vert _{L^2(\Omega )} +\int _0^t\Vert A^\gamma e^{-(t-\tau )A}h_{\varepsilon }(\cdot ,\tau )d\tau \Vert _{L^2(\Omega )}d\tau }\nonumber \\&\le \displaystyle {\Vert A^\gamma u_0\Vert _{L^2(\Omega )} {+}C_{2}\int _0^t(t-\tau )^{-\gamma -\frac{3}{2}(\frac{1}{q_0}{-}\frac{1}{2})}e^{-\lambda (t-\tau )}\Vert h_{\varepsilon }(\cdot ,\tau )\Vert _{L^{q_0}(\Omega )}d\tau }\nonumber \\&\le \displaystyle {C_{3}\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).} \end{aligned}$$
(4.17)

Observe that \(\gamma >\frac{3}{4},\) \(D(A^\gamma )\) is continuously embedded into \(L^\infty (\Omega )\). Therefore, we derive that there exists a positive constant \(C_{4}\) such that

$$\begin{aligned} \begin{array}{rl} \Vert u_{\varepsilon }(\cdot , t)\Vert _{L^\infty (\Omega )}\le C_{4}\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon })\\ \end{array} \end{aligned}$$
(4.18)

from (4.17). However, from (4.8), with the help of Sobolev’s imbedding theorem, it follows that, for any fixed \({\tilde{q}}\in (3,6)\),

$$\begin{aligned} \begin{array}{rl} \Vert c_{\varepsilon }(\cdot , t)\Vert _{L^{{\tilde{q}}}(\Omega )}\le C_{5}\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).\\ \end{array} \end{aligned}$$
(4.19)

Now, involving the variation-of-constants formula for \(c_{\varepsilon }\) and applying \(\nabla \cdot u_{\varepsilon }=0\) in \(x\in \Omega , t>0\), we have

$$\begin{aligned} c_{\varepsilon }(t)=e^{t(\Delta -1)}c_0 +\int _{0}^{t}e^{(t-s)(\Delta -1)}(F_\varepsilon (n_{\varepsilon }(s))+\nabla \cdot (u_{\varepsilon }(s) c_{\varepsilon }(s)) ds,~ t\in (0, T_{max,\varepsilon }),\nonumber \\ \end{aligned}$$
(4.20)

so that, for any \(3<q <\min \{\frac{3q_0}{(3-q_0)_{+}},{\tilde{q}}\}\), we have

$$\begin{aligned} \displaystyle {\Vert \nabla c_{\varepsilon }(\cdot , t)\Vert _{L^{q}(\Omega )}}&\le \displaystyle {\Vert \nabla e^{t(\Delta -1)} c_0\Vert _{L^{q}(\Omega )}+ \int _{0}^t\Vert \nabla e^{(t-s)(\Delta -1)}F_\varepsilon (n_{\varepsilon }(s))\Vert _{L^q(\Omega )}ds}\nonumber \\&\quad \displaystyle {+\,\int _{0}^t\Vert \nabla e^{(t-s)(\Delta -1)}\nabla \cdot (u_{\varepsilon }(s) c_{\varepsilon }(s))\Vert _{L^q(\Omega )}ds.} \end{aligned}$$
(4.21)

To address the right-hand side of (4.21), in view of (1.8), we first use Lemma 2.2 to get

$$\begin{aligned} \begin{array}{rl} \Vert \nabla e^{t(\Delta -1)} c_0\Vert _{L^{q}(\Omega )}&{} \le \displaystyle {C_{6}\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).}\\ \end{array} \end{aligned}$$
(4.22)

Since (4.15) and (4.19) yields

$$\begin{aligned} -\,\frac{1}{2}-\frac{3}{2}\left( \frac{1}{q_0}-\frac{1}{q}\right) >-1, \end{aligned}$$

together with this and (3.3), by using Lemma 2.2 again, the second term of the right-hand side is estimated as

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle {\int _{0}^t\Vert \nabla e^{(t-s)(\Delta -1)}F_\varepsilon (n_{\varepsilon }(s))\Vert _{L^{q}(\Omega )}ds}\\ &{}\quad \le \displaystyle {C_{7}\int _{0}^t[1+(t-s)^{-\frac{1}{2}-\frac{3}{2}(\frac{1}{q_0}-\frac{1}{q})}] e^{-(t-s)}\Vert n_{\varepsilon }(s)\Vert _{L^{q_0}(\Omega )}ds}\\ &{}\quad \le \displaystyle {C_{8}~ \text{ for } \text{ all }~ t\in (0,T_{max,\varepsilon }).}\\ \end{array} \end{aligned}$$
(4.23)

Finally, we will address the third term on the right-hand side of (4.21). To this end, we choose \(0< \iota < \frac{1}{2}\) satisfying \(\frac{1}{2} + \frac{3}{2}(\frac{1}{{\tilde{q}}}-\frac{1}{q}) <\iota \) and \({\tilde{\kappa }}\in (0, \frac{1}{2}-\iota )\). In view of Hölder’s inequality, we derive from Lemma 2.2, (4.19), and (4.18) that there exist constants \(C_{9}\), \(C_{10}\), \(C_{11}\), and \(C_{12}\) such that

$$\begin{aligned}&\displaystyle {\int _{0}^t\Vert \nabla e^{(t-s)(\Delta -1)}\nabla \cdot (u_{\varepsilon }(s) c_{\varepsilon }(s))\Vert _{L^{{\tilde{q}}}(\Omega )}ds}\nonumber \\&\quad \le \displaystyle {C_{9}\int _{0}^t\Vert (-\Delta +1)^\iota e^{(t-s)(\Delta -1)}\nabla \cdot (u_\varepsilon (s) c_\varepsilon (s))\Vert _{L^{q}(\Omega )}ds}\nonumber \\&\quad \le \displaystyle {C_{10}\int _{0}^t(t-s)^{-\iota -\frac{1}{2}-{\tilde{\kappa }}} e^{-\lambda (t-s)}\Vert u_\varepsilon (s) c_\varepsilon (s)\Vert _{L^{{\tilde{q}}}(\Omega )}ds}\nonumber \\&\quad \le \displaystyle {C_{11}\int _{0}^t(t-s)^{-\iota -\frac{1}{2}-{\tilde{\kappa }}} e^{-\lambda (t-s)}\Vert u_\varepsilon (s)\Vert _{L^{\infty }(\Omega )}\Vert c_\varepsilon (s)\Vert _{L^{{\tilde{q}}}(\Omega )}ds}\nonumber \\&\quad \le \displaystyle {C_{12} \quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).} \end{aligned}$$
(4.24)

Here we have used the fact that

$$\begin{aligned} \begin{array}{rl}\displaystyle \int _{0}^t(t-s)^{-\iota -\frac{1}{2}-{\tilde{\kappa }}} e^{-\lambda (t-s)}ds&{} \le \displaystyle {\int _{0}^{\infty }\sigma ^{-\iota -\frac{1}{2}-{\tilde{\kappa }}} e^{-\lambda \sigma }d\sigma <+\infty .}\\ \end{array} \end{aligned}$$

Finally, collecting (4.21)–(4.24), we can obtain that there exists a positive constant \(C_{13}\) such that

$$\begin{aligned} \int _{\Omega }|\nabla {c_{\varepsilon }}(t)|^{q}\le C_{13}\quad \text{ for } \text{ all }\;\; t\in (0, T_{max,\varepsilon })\;\; \text{ and } \text{ some }\;\; q\in \left( 3,\min \left\{ \frac{3q_0}{(3-q_0)_{+}},{\tilde{q}}\right\} \right) .\nonumber \\ \end{aligned}$$
(4.25)

The proof of Lemma 4.3 is complete. \(\square \)

Then we can establish global existence in the approximate problem (2.6) by using Lemmas 4.1 and 4.2 .

Lemma 4.4

Let \(\alpha >\frac{1}{3}\). Then, for all \(\varepsilon \in (0,1),\) the solution of (2.6) is global in time.

Proof

Assume that \(T_{max,\varepsilon }\) is finite for some \(\varepsilon \in (0,1)\). Fix \(T\in (0, T_{max,\varepsilon })\), and let \(M(T):=\sup _{t\in (0,T)}\Vert n_{\varepsilon }(\cdot ,t)\Vert _{L^\infty (\Omega )}\) and \({\tilde{h}}_{\varepsilon }:=F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_{\varepsilon }+u_\varepsilon \). Then, by Lemma 4.3, (1.5), and (3.1), there exists \(C_1 > 0\) such that

$$\begin{aligned} \begin{array}{rl} \Vert {\tilde{h}}_{\varepsilon }(\cdot , t)\Vert _{L^{q}(\Omega )}&{} \le \displaystyle {C_1\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon })\;\;\text{ and } \text{ some } \;\; 3<q<6.}\\ \end{array} \end{aligned}$$
(4.26)

Hence, because \(\nabla \cdot u_{\varepsilon }=0\), we can derive

$$\begin{aligned} n_{\varepsilon }(t)=e^{(t-t_0)\Delta }n_{\varepsilon }(\cdot ,t_0)-\int _{t_0}^{t}e^{(t-s)\Delta }\nabla \cdot (n_{\varepsilon }(\cdot ,s){\tilde{h}}_{\varepsilon }(\cdot ,s)) ds,\quad t\in (t_0, T) \end{aligned}$$
(4.27)

by means of an associate variation-of-constants formula for n, where \(t_0 := (t-1)_{+}\). If \(t\in (0,1]\), by virtue of the maximum principle, we can derive

$$\begin{aligned} \begin{array}{rl} \Vert e^{(t-t_0)\Delta }n_{\varepsilon }(\cdot ,t_0)\Vert _{L^{\infty }(\Omega )}&{} \le \displaystyle {\Vert n_0\Vert _{L^{\infty }(\Omega )},}\\ \end{array} \end{aligned}$$
(4.28)

while if \(t > 1\) then, with the help of the \(L^p\)\(L^q\) estimates for the Neumann heat semigroup and Lemma 3.2, we conclude that

$$\begin{aligned} \begin{array}{rl} \Vert e^{(t-t_0)\Delta }n_{\varepsilon }(\cdot ,t_0)\Vert _{L^{\infty }(\Omega )}&{} \le \displaystyle {C_2(t-t_0)^{-\frac{3}{2}}\Vert n_{\varepsilon }(\cdot ,t_0)\Vert _{L^{1}(\Omega )}\le C_3.}\\ \end{array} \end{aligned}$$
(4.29)

Finally, we fix an arbitrary \(p\in (3,q)\) and then once more invoke known smoothing properties of the Stokes semigroup (see Page 201 of [7]) and Hölder’s inequality to find \(C_4 > 0\) such that

$$\begin{aligned}&\displaystyle \int _{t_0}^t\Vert e^{(t-s)\Delta }\nabla \cdot (n_{\varepsilon }(\cdot ,s){\tilde{h}}_{\varepsilon }(\cdot ,s)\Vert _{L^\infty (\Omega )}ds\nonumber \\&\quad \le \displaystyle C_4\int _{t_0}^t(t-s)^{-\frac{1}{2}-\frac{3}{2p}}\Vert n_{\varepsilon }(\cdot ,s){\tilde{h}}_{\varepsilon }(\cdot ,s)\Vert _{L^p(\Omega )}ds\nonumber \\&\quad \le \displaystyle C_4\int _{t_0}^t(t-s)^{-\frac{1}{2}-\frac{3}{2p}}\Vert n_{\varepsilon }(\cdot ,s)\Vert _{L^{\frac{pq}{q-p}}(\Omega )}\Vert {\tilde{h}}_{\varepsilon }(\cdot ,s)\Vert _{L^{q}(\Omega )}ds\nonumber \\&\quad \le \displaystyle C_4\int _{t_0}^t(t-s)^{-\frac{1}{2}-\frac{3}{2p}}\Vert u_{\varepsilon }(\cdot ,s)\Vert _{L^{\infty }(\Omega )}^b\Vert u_{\varepsilon }(\cdot ,s)\Vert |_{L^1(\Omega )}^{1-b}\Vert {\tilde{h}}_{\varepsilon }(\cdot ,s)\Vert _{L^{q}(\Omega )}ds\nonumber \\&\quad \le \displaystyle C_5M^b(T)\quad \text{ for } \text{ all }\;\; t\in (0, T), \end{aligned}$$
(4.30)

where \(b:=\frac{pq-q+p}{pq}\in (0,1)\) and

$$\begin{aligned} C_5:=C_4C_1^{2-b}\int _{0}^{1}\sigma ^{-\frac{1}{2}-\frac{3}{2p}}d\sigma . \end{aligned}$$

Since \(p>3\), we conclude that \(-\frac{1}{2}-\frac{3}{2p}>-1\). In combination with (4.27)–(4.30) and using the definition of M(T),  we obtain \(C_6 > 0\) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle M(T)\le C_6+C_6M^b(T)\quad \text{ for } \text{ all }\;\; T\in (0, T_{max,\varepsilon }).\\ \end{array} \end{aligned}$$
(4.31)

Hence, in view of \(b<1\), with some basic calculation, since \(T\in (0, T_{max,\varepsilon })\) was arbitrary, we can obtain there exists a positive constant \(C_7\) such that

$$\begin{aligned} \begin{array}{rl} \Vert n_{\varepsilon }(\cdot , t)\Vert _{L^{\infty }(\Omega )}&{} \le \displaystyle {C_7\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }).}\\ \end{array} \end{aligned}$$
(4.32)

To prove the boundedness of \(\Vert \nabla c_{\varepsilon }(\cdot , t)\Vert _{L^\infty (\Omega )}\), we rewrite the variation-of-constants formula for \(c_{\varepsilon }\) in the form

$$\begin{aligned} c_{\varepsilon }(\cdot , t) = e^{t(\Delta -1) }c_0 +\int _{0}^te^{(t-s)(\Delta -1)}[F_{\varepsilon }(n_{\varepsilon })(s)-u_{\varepsilon }(s)\cdot \nabla c_{\varepsilon }(s)]ds\quad \text{ for } \text{ all }\;\; t\in (0,T_{max,\varepsilon }). \end{aligned}$$

Now, we choose \(\theta \in (\frac{1}{2}+\frac{3}{2q},1),\) where \(3<q<6\) [see (4.25)], then the domain of the fractional power \(D((-\Delta + 1)^\theta )\hookrightarrow W^{1,\infty }(\Omega )\) (see [10]). Hence, in view of \(L^p\)\(L^q\) estimates associated with the heat semigroup, (4.13), (4.14), and (3.3), we derive that there exist positive constants \(\lambda \), \(C_{8}\), \(C_{9}\), \(C_{10}\), and \(C_{11}\) such that

$$\begin{aligned}&\Vert c_{\varepsilon }(\cdot , t)\Vert _{W^{1,\infty }(\Omega )}\le \displaystyle {C_{8}\Vert (-\Delta +1)^\theta c_{\varepsilon }(\cdot , t)\Vert _{L^{q}(\Omega )}}\nonumber \\&\quad \le \displaystyle {C_{9}t^{-\theta }e^{-\lambda t}\Vert c_0\Vert _{L^{q}(\Omega )}+C_{9}\int _{0}^t(t-s)^{-\theta }e^{-\lambda (t-s)} \Vert (F_{\varepsilon }(n_{\varepsilon })-u_{\varepsilon } \cdot \nabla c_{\varepsilon })(s)\Vert _{L^{q}(\Omega )}ds}\nonumber \\&\quad \le \displaystyle {C_{10}+C_{10}\int _{0}^t(t-s)^{-\theta }e^{-\lambda (t-s)}[\Vert n_{\varepsilon }(s)\Vert _{L^q(\Omega )}+\Vert u_{\varepsilon }(s)\Vert _{L^\infty (\Omega )} \Vert \nabla c_{\varepsilon }(s)\Vert _{L^q(\Omega )}]ds}\nonumber \\&\quad \le \displaystyle {C_{11}~ \text{ for } \text{ all }~ t\in (0,T_{max,\varepsilon }).} \end{aligned}$$
(4.33)

Here we have used Hölder’s inequality as well as

$$\begin{aligned} \int _{0}^t(t-s)^{-\theta }e^{-\lambda (t-s)}\le \int _{0}^{\infty }\sigma ^{-\theta }e^{-\lambda \sigma }d\sigma <+\infty . \end{aligned}$$

In view of (4.12), (4.33), and (4.32), we apply Lemma 2.1 to reach a contradiction. \(\square \)

5 Regularity properties of time derivatives

In preparation of an Aubin–Lions type compactness argument, we will rely on an additional regularity estimate for \(n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon \), \(u_\varepsilon \cdot \nabla c_\varepsilon \), \(n_\varepsilon u_\varepsilon \), and \(c_\varepsilon u_\varepsilon \).

Lemma 5.1

Let \(\alpha >\frac{1}{3}\), and assume that (1.7) and (1.8) hold. Then one can find \(C > 0\) independent of \(\varepsilon \) such that, for all \(T\in (0,\infty )\),

$$\begin{aligned}&\begin{array}{ll} \displaystyle \int _0^T\int _{\Omega }\left[ |n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon |^{\frac{3\alpha +1}{2}} +|n_\varepsilon u_\varepsilon |^{\frac{2\alpha +\frac{2}{3}}{\alpha +1}}\right] \le C(T+1),\quad \text{ if }\;\;\frac{1}{3}<\alpha \le \frac{8}{21},\\ \end{array}\nonumber \\ \end{aligned}$$
(5.1)
$$\begin{aligned}&\begin{array}{ll} \displaystyle \int _0^T\int _{\Omega }\left[ |n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon |^{\frac{3\alpha +1}{2}} +|n_\varepsilon u_\varepsilon |^{\frac{10(3\alpha +1)}{9(\alpha +2)}}\right] \le C(T+1),\quad \text{ if }\;\;\frac{8}{21}<\alpha \le \frac{1}{2},\\ \end{array}\nonumber \\ \end{aligned}$$
(5.2)
$$\begin{aligned}&\begin{array}{ll} \displaystyle \int _0^T\int _{\Omega }\left[ |n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon |^{\frac{10\alpha }{3+2\alpha }}+|n_\varepsilon u_\varepsilon |^{\frac{10\alpha }{3(\alpha +1)}} \right] \le C(T+1),\quad \text{ if }\;\;\frac{1}{2}<\alpha <1\\ \end{array}\nonumber \\ \end{aligned}$$
(5.3)

as well as

$$\begin{aligned} \begin{array}{ll} \displaystyle \int _0^T\int _{\Omega }\left[ |n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon |^{2}+|n_\varepsilon u_\varepsilon |^{\frac{5}{3}}\right] \le C(T+1),\quad \text{ if }\;\;\alpha \ge 1\\ \end{array} \end{aligned}$$
(5.4)

and

$$\begin{aligned} \begin{array}{ll} \displaystyle \int _0^T\int _{\Omega }\left[ |u_\varepsilon \cdot \nabla c_\varepsilon |^{\frac{5}{4}} +|c_\varepsilon u_\varepsilon |^{\frac{5}{3}}\right] \le C(T+1).\\ \end{array} \end{aligned}$$
(5.5)

Proof

First, by (1.5), (3.1), and (2.8), we derive

$$\begin{aligned} n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\le C_Sn_{\varepsilon }^{(1-\alpha )_{+}} \end{aligned}$$

with \((1-\alpha )_{+}=\max \{0,1-\alpha \}.\) Case \(\frac{8}{21}<\alpha \le \frac{1}{2}\): It is not difficult to verify that

$$\begin{aligned} \frac{2}{3\alpha +1}=\frac{1}{2}+\frac{3}{6\alpha +2}(1-\alpha ) \end{aligned}$$

and

$$\begin{aligned} \frac{9(\alpha +2)}{10(3\alpha +1)}=\frac{3}{10}+\frac{3}{6\alpha +2}, \end{aligned}$$

so that, recalling (3.29), (3.44), and Hölder’s inequality, we can obtain (5.2). While if \(\frac{1}{3}<\alpha \le \frac{8}{21}\), in light of (3.6), (3.29), (3.32), (3.45), an employment of the Hölder and Young inequalities to shows that

$$\begin{aligned} \begin{array}{ll} &{}\displaystyle \int _0^T\int _{\Omega }\left[ |n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon |^{\frac{3\alpha +1}{2}} +|n_\varepsilon u_\varepsilon |^{\gamma }\right] \\ &{}\quad \le \displaystyle C_1\left[ \int _{0}^T\displaystyle \int _{\Omega } n_\varepsilon ^{\frac{6\alpha +2}{3}}\right] ^{\frac{3-3\alpha }{4}} \left[ \int _{0}^T\displaystyle \int _{\Omega } |\nabla c_\varepsilon |^{2}\right] ^{\frac{3\alpha +1}{4}}\\ &{}\qquad \displaystyle +\,C_1 \int _0^T\Vert n_\varepsilon \Vert ^{\gamma }_{L^{\frac{6\gamma }{6-\gamma }}(\Omega )}\Vert u_\varepsilon \Vert ^{\gamma }_{L^6(\Omega )}\le C_2(T+1), \end{array} \end{aligned}$$

where \(\gamma =\frac{2\alpha +\frac{2}{3}}{\alpha +1}\) is given by Lemma 3.5.

Other cases can be proved very similarly. Therefore, we omit their proofs. \(\square \)

To prepare our subsequent compactness properties of \((n_\varepsilon , c_\varepsilon ,u_\varepsilon )\) by means of the Aubin–Lions lemma (see Simon [18]), we use Lemmas 3.23.4 to obtain the following regularity property with respect to the time variable.

Lemma 5.2

Let \(\alpha >\frac{1}{3}\), and assume that (1.7) and (1.8) hold. Then there exists \(C>0\) independent of \(\varepsilon \) such that

$$\begin{aligned} \begin{array}{ll} \displaystyle \int _0^T\Vert \partial _tn_\varepsilon (\cdot ,t)\Vert _{(W^{2,4}(\Omega ))^*}dt \le C(T+1)\quad \text{ for } \text{ all }\;\; T\in (0,\infty )\\ \end{array} \end{aligned}$$
(5.6)

as well as

$$\begin{aligned} \begin{array}{ll} \displaystyle \int _0^T\Vert \partial _tc_\varepsilon (\cdot ,t)\Vert _{(W^{1,5}(\Omega ))^*}^{\frac{5}{4}}dt \le C(T+1)\quad \text{ for } \text{ all }\;\; T\in (0,\infty )\\ \end{array} \end{aligned}$$
(5.7)

and

$$\begin{aligned} \begin{array}{ll} \displaystyle \int _0^T\Vert \partial _tu_\varepsilon (\cdot ,t)\Vert _{(W^{1,5}_{0,\sigma }(\Omega ))^*}^{\frac{5}{4}}dt \le C(T+1)\quad \text{ for } \text{ all }\;\; T\in (0,\infty ).\\ \end{array} \end{aligned}$$
(5.8)

Proof

Firstly, testing the first equation of (2.6) by certain \(\varphi \in C^{\infty }({\bar{\Omega }})\), we have

$$\begin{aligned} \begin{array}{rl} \displaystyle \left| \int _{\Omega }(n_{\varepsilon ,t})\varphi \right| &{} = \displaystyle {\left| \int _{\Omega }\left[ \Delta n_{\varepsilon }-\nabla \cdot (n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_{\varepsilon })-u_{\varepsilon }\cdot \nabla n_{\varepsilon }\right] \varphi \right| }\\ &{}= \displaystyle {\left| \int _\Omega \left[ -\nabla n_{\varepsilon }\cdot \nabla \varphi +n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_{\varepsilon }\cdot \nabla \varphi + n_{\varepsilon }u_{\varepsilon }\cdot \nabla \varphi \right] \right| }\\ &{}\le \displaystyle {\left| \int _\Omega \left[ |\nabla n_{\varepsilon }|+|n_{\varepsilon }F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_{\varepsilon }|+ |n_{\varepsilon }u_{\varepsilon }|\right] \right| \Vert \varphi \Vert _{W^{1,\infty }(\Omega )}} \end{array} \end{aligned}$$

for all \(t>0\).

Observe that the embedding \(W^{2,4 }(\Omega )\hookrightarrow W^{1,\infty }(\Omega )\), so that, in view of \(\alpha >\frac{1}{3}\), Lemmas 3.4 and 5.1, we deduce from the Young inequality that for some \(C_1\) and \(C_2\) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle \int _0^T\Vert \partial _{t}n_{\varepsilon }(\cdot ,t)\Vert _{(W^{2,4 }(\Omega ))^*}dt\\ &{}\quad \le \displaystyle {C_1\left\{ \int _0^T\int _{\Omega } |\nabla n_{\varepsilon }|^{r_1}+\int _0^T\int _{\Omega }|n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon |^{r_1}{+}\int _0^T\int _{\Omega }|n_\varepsilon u_\varepsilon |^{r_2}{+}T\right\} }\\ &{}\quad \le \displaystyle {C_2(T+1)~~\text{ for } \text{ all }~~ T > 0, }\\ \end{array}\nonumber \\ \end{aligned}$$
(5.9)

where

$$\begin{aligned} r_1= \left\{ \begin{array}{ll} \frac{3\alpha +1}{2}&{}\quad \text{ if }\;\; \frac{1}{3}<\alpha \le \frac{1}{2},\\ \frac{10\alpha }{3+2\alpha }&{}\quad \text{ if }\;\;\frac{1}{2}<\alpha <1,\\ 2&{}\quad \text{ if }\;\;\alpha \ge 1\\ \end{array}\right. \end{aligned}$$

and

$$\begin{aligned} r_2= \left\{ \begin{array}{ll} \frac{2\alpha +\frac{2}{3}}{\alpha +1}&{}\quad \text{ if }\;\;\frac{1}{3}<\alpha \le \frac{8}{21},\\ \frac{10(3\alpha +1)}{9(\alpha +2)}&{}\quad \text{ if }\;\;\frac{8}{21}<\alpha \le \frac{1}{2},\\ \frac{10\alpha }{3(\alpha +1)}&{}\quad \text{ if }\;\;\frac{1}{2}<\alpha <1,\\ \frac{5}{3}&{}\quad \text{ if }\;\;\alpha \ge 1,\\ \end{array}\right. \end{aligned}$$

Likewise, given any \(\varphi \in C^\infty ({\bar{\Omega }})\), we may test the second equation in (2.6) against \(\varphi \) to conclude that

$$\begin{aligned} \begin{array}{rl} \displaystyle \left| \int _{\Omega }\partial _{t}c_{\varepsilon }(\cdot ,t)\varphi \right| &{} = \displaystyle {\left| \int _{\Omega }\left[ \Delta c_{\varepsilon }-c_{\varepsilon }+n_{\varepsilon }-u_{\varepsilon }\cdot \nabla c_{\varepsilon }\right] \cdot \varphi \right| }\\ &{} = \displaystyle {\left| -\int _\Omega \nabla c_{\varepsilon }\cdot \nabla \varphi -\int _\Omega c_{\varepsilon }\varphi +\int _\Omega n_{\varepsilon } \varphi +\int _\Omega c_{\varepsilon }u_\varepsilon \cdot \nabla \varphi \right| }\\ &{} \le \displaystyle {\left\{ \Vert \nabla c_{\varepsilon }\Vert _{L^{{\frac{5}{4}}}(\Omega )}+\Vert c_{\varepsilon } \Vert _{L^{\frac{5}{4}}(\Omega )}+\Vert n_{\varepsilon } \Vert _{L^{\frac{5}{4}}(\Omega )}+\Vert c_{\varepsilon }u_\varepsilon \Vert _{L^{\frac{5}{4}}(\Omega )}\right\} \Vert \varphi \Vert _{W^{1,5}(\Omega )}}\\ \end{array} \end{aligned}$$

for all \(t>0\). Thus, from Lemmas 3.4 and 5.1 again, in light of \(\alpha >\frac{1}{3}\), we invoke the Young inequality again and obtain that there exist positive constant \(C_{3}\) and \(C_{4}\) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle \int _0^T\Vert \partial _{t}c_{\varepsilon }(\cdot ,t)\Vert ^{\frac{5}{4}}_{(W^{1,5}(\Omega ))^*}dt\\ &{}\quad \le \displaystyle {C_{3}\left( \int _0^T\int _\Omega |\nabla c_{\varepsilon }|^{2}+\int _0^T\int _\Omega n_{\varepsilon }^{r_3}+\int _0^T\int _\Omega c_\varepsilon ^{\frac{10}{3}}+\int _0^T\int _\Omega |u_\varepsilon |^{\frac{10}{3}}+T\right) }\\ &{}\quad \le \displaystyle {C_{4}(T+1) \quad \text{ for } \text{ all }\;\; T>0}\\ \end{array} \end{aligned}$$

with

$$\begin{aligned} r_3= \left\{ \begin{array}{ll} \frac{6\alpha +2}{3}&{}\quad \text{ if }\;\; \frac{1}{3}<\alpha \le \frac{1}{2},\\ \frac{10\alpha }{3}&{}\quad \text{ if }\;\; \frac{1}{2}<\alpha <1,\\ \frac{10}{3}&{}\quad \text{ if }\;\; \alpha \ge 1.\\ \end{array}\right. \end{aligned}$$
(5.10)

Finally, for any given \(\varphi \in C^{\infty }_{0,\sigma } (\Omega ;{\mathbb {R}}^3)\), we infer from the third equation in (2.6) that

$$\begin{aligned} \begin{array}{rl} \displaystyle \left| \int _{\Omega }\partial _{t}u_{\varepsilon }(\cdot ,t)\varphi \right|&= \displaystyle {\left| -\int _\Omega \nabla u_{\varepsilon }\cdot \nabla \varphi -\kappa \int _\Omega (Y_{\varepsilon }u_{\varepsilon }\otimes u_{\varepsilon })\cdot \nabla \varphi +\int _\Omega n_{\varepsilon }\nabla \phi \cdot \varphi \right| ~~\text{ for } \text{ all }~~ t>0.} \end{array} \end{aligned}$$

Now, by virtue of (3.6), Lemmas 3.4 and 5.1 , we also get that there exist positive constants \(C_{5},C_{6}\) and \(C_{7}\) such that

$$\begin{aligned} \begin{array}{rl} &{}\displaystyle \int _0^T\Vert \partial _{t}u_{\varepsilon }(\cdot ,t)\Vert _{(W^{1,5}_{0,\sigma }(\Omega ))^*}^{\frac{5}{4}}dt\\ &{}\quad \le \displaystyle {C_{5}\left( \int _0^T\int _\Omega |\nabla u_{\varepsilon }|^{\frac{5}{4}}+\int _0^T\int _\Omega |Y_{\varepsilon }u_{\varepsilon }\otimes u_{\varepsilon }|^{\frac{5}{4}}+\int _0^T\int _\Omega n_\varepsilon ^{\frac{5}{4}}\right) }\\ &{}\quad \le \displaystyle {C_{6}\left( \int _0^T\int _\Omega |\nabla u_{\varepsilon }|^{2}+\int _0^T\int _\Omega |Y_{\varepsilon }u_\varepsilon |^{2}+\int _0^T\int _\Omega |u_\varepsilon |^{\frac{10}{3}}+\int _0^T\int _\Omega n_{\varepsilon }^{r_3}+T\right) }\\ &{}\quad \le \displaystyle {C_{7}(T+1)\quad \text{ for } \text{ all }\;\; T>0,} \end{array} \end{aligned}$$

which implies (5.8). Here \(r_3\) is the same as (5.10). \(\square \)

6 Passing to the limit: Proof of Theorem  1.1

Based on the above lemmas and by extracting suitable subsequences in a standard way, we can prove Theorem  1.1.

Lemma 6.1

Let (1.4), (1.5), (1.7) and (1.8) hold, and suppose that \(\alpha >\frac{1}{3}.\) There exists \((\varepsilon _j)_{j\in {\mathbb {N}}}\subset (0, 1)\) such that \(\varepsilon _j{\searrow } 0\) as \(j{\rightarrow }\infty \) and such that as \(\varepsilon = \varepsilon _j{\searrow } 0\) we have

$$\begin{aligned}&n_\varepsilon \rightarrow n ~\text{ a.e. }~\text{ in }~\Omega \times (0,\infty )~\text{ and } \text{ in }~ L_{loc}^{r}({\bar{\Omega }}\times [0,\infty ))~\text{ with }~r= \left\{ \begin{array}{ll} \frac{3\alpha +1}{2}&{}\quad \text{ if }\;\; \frac{1}{3}<\alpha \le \frac{1}{2},\\ \frac{10\alpha }{3+2\alpha }&{}\quad \text{ if }\;\; \frac{1}{2}<\alpha <1,\\ 2&{}\quad \text{ if }\;\;\alpha \ge 1,\\ \end{array}\right. \nonumber \\ \end{aligned}$$
(6.1)
$$\begin{aligned}&\nabla n_\varepsilon \rightharpoonup \nabla n ~\text{ in }~ L_{loc}^{r}({\bar{\Omega }}\times [0,\infty ))~\text{ with }~r= \left\{ \begin{array}{ll} \frac{3\alpha +1}{2}&{}\quad \text{ if }\;\;\frac{1}{3}<\alpha \le \frac{1}{2},\\ \frac{10\alpha }{3+2\alpha }&{}\quad \text{ if }\;\;\frac{1}{2}<\alpha <1,\\ 2&{}\quad \text{ if }\;\;\alpha \ge 1,\\ \end{array}\right. \end{aligned}$$
(6.2)
$$\begin{aligned}&c_\varepsilon \rightarrow c ~\text{ in }~ L^{2}_{loc}({\bar{\Omega }}\times [0,\infty ))~\text{ and }~\text{ a.e. }~\text{ in }~\Omega \times (0,\infty ), \end{aligned}$$
(6.3)
$$\begin{aligned}&\nabla c_\varepsilon \rightarrow \nabla c ~\text{ a.e. }~\text{ in }~\Omega \times (0,\infty ), \end{aligned}$$
(6.4)
$$\begin{aligned}&u_\varepsilon \rightarrow u~\text{ in }~ L_{loc}^2({\bar{\Omega }}\times [0,\infty ))~\text{ and }~\text{ a.e. }~\text{ in }~\Omega \times (0,\infty ) \end{aligned}$$
(6.5)

as well as

$$\begin{aligned} \nabla c_\varepsilon \rightharpoonup \nabla c~\begin{array}{ll} \text{ in }~ L_{loc}^{2}({\bar{\Omega }}\times [0,\infty )) \end{array} \end{aligned}$$
(6.6)

and

$$\begin{aligned} \nabla u_\varepsilon \rightharpoonup \nabla u ~ \text{ in }~L^{2}_{loc}({\bar{\Omega }}\times [0,\infty )) \end{aligned}$$
(6.7)

and

$$\begin{aligned} u_\varepsilon \rightharpoonup u ~ \text{ in }~L^{\frac{10}{3}}_{loc}({\bar{\Omega }}\times [0,\infty )) \end{aligned}$$
(6.8)

with some triple (ncu) that is a global weak solution of (1.3) in the sense of Definition 2.1.

Proof

First, from Lemma 3.4 and (5.6), we derive that there exists a positive constant \(C_0\) such that

$$\begin{aligned} \Vert n_{\varepsilon }\Vert _{L^{r}_{loc}([0,\infty ); W^{1,r}(\Omega ))}\le C_0(T+1)~\text{ and }~\Vert \partial _{t}n_{\varepsilon }\Vert _{L^{1}_{loc}([0,\infty ); (W^{2,4}(\Omega ))^*)}\le C_0(T+1),\nonumber \\ \end{aligned}$$
(6.9)

where r is given by (6.1). Hence, from (6.9) and the Aubin–Lions lemma (see, e.g., [18]), we conclude that

$$\begin{aligned} \begin{array}{rl} (n_{\varepsilon })_{\varepsilon \in (0,1)}~\text{ is } \text{ strongly } \text{ precompact } \text{ in }~L^{r}_{loc}({\bar{\Omega }}\times [0,\infty )), \end{array} \end{aligned}$$
(6.10)

so that, there exists a sequence \((\varepsilon _j)_{j\in {\mathbb {N}}}\subset (0, 1)\) such that \(\varepsilon =\varepsilon _j{\searrow }0\) as \(j{\rightarrow }\infty \) and

$$\begin{aligned} n_\varepsilon \rightarrow n ~\text{ a.e. }~\text{ in }~\Omega \times (0,\infty )~\text{ and } \text{ in }~ L_{loc}^{r}({\bar{\Omega }}\times [0,\infty ))\quad \text{ as }\;\;\varepsilon =\varepsilon _j{\searrow }0, \end{aligned}$$
(6.11)

where r is the same as (6.1). Now, in view of Lemmas 3.3, 3.4, 5.1, and 5.2, employing the same arguments as in the proof of (6.9)–(6.11), we can derive (6.1)–(6.3) and (6.5)–(6.8) holds. Next, let \(g_\varepsilon (x, t) := -c_\varepsilon +F_{\varepsilon }(n_{\varepsilon })-u_{\varepsilon }\cdot \nabla c_{\varepsilon }.\) With this notation, the second equation of (2.6) can be rewritten in component form as

$$\begin{aligned} c_{\varepsilon t}-\Delta c_{\varepsilon } = g_\varepsilon . \end{aligned}$$
(6.12)

Case \(\frac{1}{3}<\alpha \le \frac{1}{2}\): Observe that

$$\begin{aligned} \frac{5}{4}<\frac{4}{3}<\min \left\{ \frac{6\alpha +2}{3},\frac{10}{3}\right\} ~\text{ for }~\frac{1}{3}<\alpha \le \frac{1}{2}. \end{aligned}$$

Thus, recalling (3.29), (3.32), and (5.5) and applying Hölder’s inequality, we conclude that, for any \(\varepsilon \in (0,1)\), \( g_\varepsilon \) is bounded in \(L^{\frac{5}{4}} (\Omega \times (0, T))\), and we may invoke the standard parabolic regularity theory to (6.12) and infer that \((c_{\varepsilon })_{\varepsilon \in (0,1)}\) is bounded in \(L^{\frac{5}{4}} ((0, T); W^{2,\frac{5}{4}}(\Omega ))\). Hence, by virtue of (5.7) and the Aubin–Lions lemma, we derive the relative compactness of \((c_{\varepsilon })_{\varepsilon \in (0,1)}\) in \(L^{\frac{5}{4}} ((0, T); W^{1,\frac{5}{4}}(\Omega ))\). We can pick an appropriate subsequence that is still written as \((\varepsilon _j )_{j\in {\mathbb {N}}}\) such that \(\nabla c_{\varepsilon _j} \rightarrow z_1\) in \(L^{\frac{5}{4}} (\Omega \times (0, T))\) for all \(T\in (0, \infty )\) and some \(z_1\in L^{\frac{5}{4}} (\Omega \times (0, T))\) as \(j\rightarrow \infty \). Therefore, by (5.7), we can also derive that \(\nabla c_{\varepsilon _j} \rightarrow z_1\) a.e. in \(\Omega \times (0, \infty )\) as \(j \rightarrow \infty \). In view of (6.6) and Egorov’s theorem, we conclude that \(z_1=\nabla c\) and hence (6.4) holds. Next, we pay attention to the case \(\frac{1}{2}<\alpha <1\): By straightforward calculations, and using relation \(\frac{1}{2}<\alpha <1\), one has

$$\begin{aligned} \frac{5}{4}<\frac{5}{3}<\min \left\{ \frac{10\alpha }{3},\frac{10}{3}\right\} . \end{aligned}$$

Consequently, based on (3.30), (3.32), and (5.5), it follows from Hölder’s inequality that

$$\begin{aligned} c_{\varepsilon t}-\Delta c_{\varepsilon }=g_\varepsilon ~~\text{ is } \text{ bounded } \text{ in } ~L^{\frac{5}{4}} (\Omega \times (0, T))~~\text{ for } \text{ any }~~\varepsilon \in (0,1). \end{aligned}$$
(6.13)

Employing almost exactly the same arguments as in the proof of the case \(\frac{1}{3}<\alpha \le \frac{1}{2}\), and taking advantage of (6.13), we conclude the estimate (6.6). The proof of case \(\alpha \ge 1\) is similar to that of case \(\frac{1}{3}<\alpha \le \frac{1}{2}\), so we omit it.

In the following proof, we shall prove that (ncu) is a weak solution of problem (1.3) in Definition 2.1. In fact, by \(\alpha >\frac{1}{3}\), we conclude that

$$\begin{aligned} r>1, \end{aligned}$$

where r is given by (6.1). Therefore, with the help of (6.1)–(6.3) and (6.5)–(6.7), we can derive (2.1). Now, by the nonnegativity of \(n_\varepsilon \) and \(c_\varepsilon \), we obtain \(n \ge 0\) and \(c\ge 0\). Next, from (6.7) and \(\nabla \cdot u_{\varepsilon } = 0\), we conclude that \(\nabla \cdot u = 0\) a.e. in \(\Omega \times (0, \infty )\). However, in view of (5.2), (5.3), and (5.4), we conclude that

$$\begin{aligned} n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon \rightharpoonup z_2 ~\text{ in }~ L^{r}(\Omega \times (0,T))~\text{ as }~\varepsilon = \varepsilon _j{\searrow } 0~\text{ for } \text{ each }~ T\in (0,\infty ),\nonumber \\ \end{aligned}$$
(6.14)

where r is given by (6.1). However, it follows from (1.4), (2.8), (3.2), (6.1), (6.3), and (6.4) that

$$\begin{aligned} n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon \rightarrow nS(x, n, c)\nabla c~\text{ a.e. }~\text{ in }~\Omega \times (0,\infty )~\text{ as }~\varepsilon = \varepsilon _j{\searrow } 0.\qquad \end{aligned}$$
(6.15)

Again by Egorov’s theorem, we gain \(z_2=nS(x, n, c)\nabla c\), and therefore (6.14) can be rewritten as

$$\begin{aligned}&n_\varepsilon F'_{\varepsilon }(n_{\varepsilon })S_\varepsilon (x, n_{\varepsilon }, c_{\varepsilon })\nabla c_\varepsilon \rightharpoonup nS(x, n, c)\nabla c ~\text{ in }~ L^{r}(\Omega \times (0,T))~\text{ as }~\varepsilon \nonumber \\&\quad = \varepsilon _j{\searrow } 0~\text{ for } \text{ each }~ T\in (0,\infty ), \end{aligned}$$
(6.16)

which together with \(r>1\) implies the integrability of \(nS(x, n, c)\nabla c\) in (2.2) as well. It is not difficult to check that

$$\begin{aligned}&{\frac{2\alpha +\frac{2}{3}}{\alpha +1}}>1~\text{ if }~\frac{1}{3}<\alpha \le \frac{8}{21},{\frac{10(3\alpha +1)}{9(\alpha +2)}}>1~\text{ if }~\frac{8}{21}\\&\quad<\alpha \le \frac{1}{2}~\text{ and }~{\frac{10\alpha }{3(\alpha +1)}} >1~\text{ if }~\frac{1}{2}<\alpha <1. \end{aligned}$$

Thereupon, recalling (5.2), (5.3), and (5.4), we infer that, for each \(T\in (0, \infty ),\)

$$\begin{aligned} n_\varepsilon u_{\varepsilon }\rightharpoonup z_3 ~\text{ in }~ L^{{\tilde{r}}}(\Omega \times (0,T))~\text{ with }~{\tilde{r}}= \left\{ \begin{array}{ll} {\frac{2\alpha +\frac{2}{3}}{\alpha +1}}&{}\quad \text{ if }\;\; \frac{1}{3}<\alpha \le \frac{8}{21},\\ \frac{10(3\alpha +1)}{9(\alpha +2)}&{}\quad \text{ if }\;\; \frac{8}{21}<\alpha \le \frac{1}{2},\\ \frac{10\alpha }{3(\alpha +1)}&{}\quad \text{ if }\;\;\frac{1}{2}<\alpha <1,\\ \frac{5}{3}&{}\quad \text{ if }\;\;\alpha \ge 1.\\ \end{array}\right. ~\text{ as }~\varepsilon = \varepsilon _j{\searrow } 0, \nonumber \\ \end{aligned}$$
(6.17)

(6.17) together with (6.1) and (6.5) implies

$$\begin{aligned} n_\varepsilon u_\varepsilon \rightarrow nu~\text{ a.e. }~\text{ in }~\Omega \times (0,\infty )~\text{ as }~\varepsilon = \varepsilon _j{\searrow } 0. \end{aligned}$$
(6.18)

(6.17) along with (6.18) and Egorov’s theorem guarantees that \(z_3=nu\), whereupon we derive from (6.17) that

$$\begin{aligned} n_\varepsilon u_{\varepsilon }\rightharpoonup nu ~\text{ in }~ L^{{\tilde{r}}}(\Omega \times (0,T))~\text{ with }~{\tilde{r}}= \left\{ \begin{array}{ll} {\frac{2\alpha +\frac{2}{3}}{\alpha +1}}&{}\quad \text{ if }\;\; \frac{1}{3}<\alpha \le \frac{8}{21},\\ \frac{10(3\alpha +1)}{9(\alpha +2)}&{}\quad \text{ if }\;\; \frac{8}{21}<\alpha \le \frac{1}{2},\\ \frac{10\alpha }{3(\alpha +1)}&{}\quad \text{ if }\;\;\frac{1}{2}<\alpha <1,\\ \frac{5}{3}&{}\quad \text{ if }\;\;\alpha \ge 1\\ \end{array}\right. ~\text{ as }~\varepsilon = \varepsilon _j{\searrow } 0,\nonumber \\ \end{aligned}$$
(6.19)

for each \(T\in (0, \infty )\).

As a straightforward consequence of (6.3) and (6.5), it holds that

$$\begin{aligned} c_\varepsilon u_\varepsilon \rightarrow cu ~ \text{ in }~ L^{1}_{loc}({\bar{\Omega }}\times (0,\infty ))~\text{ as }~\varepsilon =\varepsilon _j{\searrow }0. \end{aligned}$$
(6.20)

Thus, the integrability of nu and cu in (2.2) is verified by (6.3) and (6.5).

Next, by (6.5) and the fact that \(\Vert Y_{\varepsilon }\varphi \Vert _{L^2(\Omega )} \le \Vert \varphi \Vert _{L^2(\Omega )}(\varphi \in L^2_{\sigma }(\Omega ))\) and \(Y_{\varepsilon }\varphi {\rightarrow } \varphi \) in \(L^2(\Omega )\) as \(\varepsilon {\searrow }0\), we can get that there exists a positive constant \(C_1\) such that, for any \(\varepsilon \in (0,1)\),

$$\begin{aligned} \begin{array}{rl} \left\| Y_{\varepsilon }u_{\varepsilon }(\cdot ,t)-u(\cdot ,t)\right\| _{L^2(\Omega )} &{} \le \displaystyle {\left\| Y_{\varepsilon }[u_{\varepsilon }(\cdot ,t)-u(\cdot ,t)]\right\| _{L^2(\Omega )}+ \left\| Y_{\varepsilon }u(\cdot ,t)-u(\cdot ,t)\right\| _{L^2(\Omega )}}\\ &{} \le \displaystyle {\left\| u_{\varepsilon }(\cdot ,t)-u(\cdot ,t)\right\| _{L^2(\Omega )}+ \left\| Y_{\varepsilon }u(\cdot ,t)-u(\cdot ,t)\right\| _{L^2(\Omega )}}\\ &{} \rightarrow \displaystyle {0~\text{ as }~\varepsilon =\varepsilon _j{\searrow }0}\\ \end{array} \end{aligned}$$

and

$$\begin{aligned} \begin{array}{rl} \left\| Y_{\varepsilon }u_{\varepsilon }(\cdot ,t)-u(\cdot ,t)\right\| _{L^2(\Omega )}^2 &{} \le \displaystyle {\left( \Vert Y_{\varepsilon }u_{\varepsilon }(\cdot ,t)|\Vert _{L^2(\Omega )}+\Vert u(\cdot ,t)|\Vert _{L^2(\Omega )}\right) ^2}\\ &{} \le \displaystyle {\left( \Vert u_{\varepsilon }(\cdot ,t)|\Vert _{L^2(\Omega )}+\Vert u(\cdot ,t)|\Vert _{L^2(\Omega )}\right) ^2}\\ &{} \le \displaystyle {C_1~\text{ for } \text{ all }~t\in (0,\infty )/ N~\text{ with } \text{ some } \text{ null } \text{ set }~ N\subset (0,\infty ), }\\ \end{array} \end{aligned}$$

so that, by the dominated convergence theorem, we also find that

$$\begin{aligned} \begin{array}{rl} \displaystyle \int _{0}^T\Vert Y_{\varepsilon }u_{\varepsilon }(\cdot ,t)-u(\cdot ,t)\Vert _{L^2(\Omega )}^2dt\rightarrow 0 ~\text{ as }~\varepsilon =\varepsilon _j{\searrow }0 \quad \text{ for } \text{ all }\;\; T>0. \end{array} \end{aligned}$$

Therefore,

$$\begin{aligned} Y_\varepsilon u_\varepsilon \rightarrow u ~\text{ in }~ L_{loc}^2([0,\infty ); L^2(\Omega )). \end{aligned}$$
(6.21)

Now, combining (6.5) with (6.21), we derive

$$\begin{aligned} \begin{array}{rl} Y_{\varepsilon }u_{\varepsilon }\otimes u_{\varepsilon }\rightarrow u \otimes u ~\text{ in }~L^1_{loc}({\bar{\Omega }}\times [0,\infty ))~\text{ as }~\varepsilon =\varepsilon _j{\searrow }0. \end{array} \end{aligned}$$
(6.22)

Therefore, the integrability of \(nS(x,n,c)\nabla c\), nu, cu, and \(u\otimes u\) in (2.2) is verified by (6.16), (6.19), (6.20) and (6.22). Finally, for any fixed \(T\in (0, \infty )\), applying (6.1), one can get

$$\begin{aligned}&\displaystyle \int _0^T\left\| F_{\varepsilon }(n_{\varepsilon }(\cdot ,t))-n(\cdot ,t)\right\| _{L^r(\Omega )}^rdt\nonumber \\&\quad \le \displaystyle {\displaystyle \int _0^T\left\| F_{\varepsilon }(n_{\varepsilon }(\cdot ,t))-F_{\varepsilon }(n(\cdot ,t))\right\| _{L^r(\Omega )}^rdt+ \displaystyle \int _0^T\left\| F_{\varepsilon }(n(\cdot ,t))-n(\cdot ,t)\right\| _{L^r(\Omega )}^rdt}\nonumber \\&\quad \le \displaystyle {\Vert F'_{\varepsilon }\Vert _{L^\infty (\Omega \times (0,\infty ))}\displaystyle \int _0^T\left\| n_{\varepsilon }(\cdot ,t)-n(\cdot ,t)\right\| _{L^r(\Omega )}^rdt+ \displaystyle \int _0^T\left\| F_{\varepsilon }(n(\cdot ,t))-n(\cdot ,t)\right\| _{L^r(\Omega )}^rdt,} \nonumber \\ \end{aligned}$$
(6.23)

where r is the same as in (6.1). Besides that, we also deduce from (3.3) and \(r>1\) that

$$\begin{aligned} \begin{array}{rl} \left\| F_{\varepsilon }(n(\cdot ,t))-n(\cdot ,t)\right\| _{L^r(\Omega \times (0,T))}^r &{} \le \displaystyle {2^r\Vert n(\cdot ,t)\Vert }\\ \end{array} \end{aligned}$$

for each \(t\in (0, T)\), which together with (6.1) shows the integrability of \(\left\| F_{\varepsilon }(n(\cdot ,t))\right. \left. -n(\cdot ,t)\right\| _{L^r(\Omega )}^r \) on (0, T). Thereupon, by virtue of (3.2), we infer from the dominated convergence theorem that

$$\begin{aligned} \begin{array}{rl} \displaystyle \int _0^T\left\| F_{\varepsilon }(n)-n\right\| _{L^r(\Omega )}^rdt\rightarrow 0 ~\text{ as }~\varepsilon =\varepsilon _j{\searrow }0 \end{array} \end{aligned}$$
(6.24)

for each \(T\in (0, \infty )\). Inserting (6.24) into (6.23) and using (6.1) and (3.1), we can see clearly that

$$\begin{aligned} \begin{array}{rl} F_{\varepsilon }(n)\rightarrow n ~\text{ in }~L^r_{loc}({\bar{\Omega }}\times [0,\infty ))~\text{ as }~\varepsilon =\varepsilon _j{\searrow }0. \end{array} \end{aligned}$$
(6.25)

Finally, according to (6.1)–(6.3), (6.5)–(6.7), (6.16), (6.19), (6.20), (6.21), (6.22), and (6.25), we may pass to the limit in the respective weak formulations associated with the regularized system (2.6) and obtain the integral identities (2.3)–(2.5). \(\square \)