1 Introduction

Chemotaxis is the directed movement of cells or organisms in response to the gradients of concentration of the chemical stimuli, plays essential roles in various biological process such as aggregative patterns of bacteria, slime mold formation, angiogenesis in tumor progression and wound healing [9, 23, 34]. The classical Keller–Segel model proposed by Keller and Segel [26] to the following chemotaxis model

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{t}=\Delta n-\chi \nabla \cdot (n\nabla v), & x\in \Omega ,t>0,\\ v_{t}=\Delta v-v+n, & x\in \Omega ,t>0,\\ \end{array}\right. } \end{aligned}$$
(1.1)

where \(\Omega \subset {\mathbb {R}}^N~(N\ge 1)\) is a bounded domain with smooth boundary, \(\chi >0\) is the chemotactic sensitivity and the unknown variables n(xt) and v(xt) represent the density of cell and the concentration of chemical signal, respectively. The system (1.1) have received considerable interest of mathematicians to develop a detailed qualitative analysis concentrating on the global existence, boundedness and asymptotic behavior of solutions for the corresponding homogeneous Neumann initial-boundary value problem (IBVP). To be specific, if \(N=1\), the corresponding solution is always globally bounded [33]; if \(N=2\), a critical mass blow-up phenomenon occurs in the radially symmetric setting, namely solution is globally bounded in the case \(\Vert n_{0}\Vert _{L^{1}(\Omega )}<\frac{8\pi }{\chi }\) [31], while the corresponding solution blows up in finite time for the case \(\Vert n_{0}\Vert _{L^{1}(\Omega )}>\frac{8\pi }{\chi }\) [22, 24]; if \(N\ge 3\), the finite-time blow-up of radial solutions may happen with arbitrarily small mass \(\Vert n_{0}\Vert _{L^{1}(\Omega )}\) [46], and the corresponding solution exists globally in time and converges to the constant steady state provided that \(\Vert n_{0}\Vert _{L^{\frac{N}{2}}(\Omega )}+\Vert v_{0}\Vert _{L^{N}(\Omega )}\) is sufficiently small [8, 40].

In the case that the time scale of chemotactic movement is rather long, the proliferation and death of cell should be taken into account. On the basis of this fact, a large body of work has been devoted to the following Keller–Segel-growth model

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{t}=\Delta n-\chi \nabla \cdot (n\nabla v)+rn-\mu n^{\alpha }, & x\in \Omega ,t>0,\\ v_{t}=\Delta v-v+n, & x\in \Omega ,t>0,\\ \end{array}\right. } \end{aligned}$$
(1.2)

where \(r\in {\mathbb {R}}\), \(\mu >0\) and \(\alpha >1\). In contrast to the minimal Keller–Segel system (1.1) in which solutions of problem may blow up in the case \(N\ge 2\) [22, 24, 31, 40, 46], the associated IBVP of (1.2) with quadratic degradation (i.e., \(\alpha =2\)) possesses a global bounded classical solution for the case either \(N=2\) and arbitrary \(\mu >0\) [32], or \(N\ge 3\) and suitably large \(\mu >0\) [45], and this solution was further shown to approach the spatially homogeneous steady state [47]. Of note, only the global existence of weak solutions have been obtained for \(N\ge 3\) and any \(\mu >0\) [27], it is still unknown whether or not explosions may occur for small \(\mu >0\). After all, the above results indicate that the suitably strong logistic-type degradation exerts a somewhat stabilizing influence on the system in the sense of blow-up prevention. Some analytical findings have revealed that some weaker degradation may fail to prevent finite-time blow-up of solution on \(N\ge 2\), but exclude the possibility of collapse into persistent Dirac-type measures. Indeed, it was asserted in [51] that the no-flux IBVP of (1.2) admits at least one global very weak solution for any \(\alpha >1\) and \(\mu >0\), and that these solutions were shown to stabilize toward the nontrivial spatially homogeneous steady state in the large time limit for \(\alpha \ge 2-\frac{2}{N}\) and appropriately large \(\mu >0\) [43]. For more related results on (1.2) and its numerous variants, we refer to [28, 41] and two impressive surveys [1, 2].

In the Keller–Segel model (1.1) and Keller–Segel-growth model (1.2), the chemical signal is directly produced by cells themselves. Nevertheless, the signal may be produced indirectly in numerous biological circumstances such as the Mountain Pine Beetle spread and aggregation in a forest habitat [25], predator–prey interaction in marine ecosystem [36] and tumor invasion inside the body of cancer patient [15]. In recent years, much attention has been focused on the following Keller–Segel-growth system with indirect signal production

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{t}=\Delta n-\chi \nabla \cdot (n\nabla v)+rn-\mu n^{\alpha }, & x\in \Omega ,t>0,\\ v_{t}=\Delta v-v+w, & x\in \Omega ,t>0,\\ w_{t}=\Delta w-w+n, & x\in \Omega ,t>0.\\ \end{array}\right. } \end{aligned}$$
(1.3)

Note that the signal production mechanism in (1.3) is indirect, that is, the chemoattractant v is not produced by cells directly, but is governed by the quantity w arising from n. The prototypical model of (1.3) (i.e., \(r=\mu =0\) in (1.3)) was proposed and studied by Fujie and Senba [16]. Several recent studies indicate that the indirect signal production mechanism can give rise to different interactions of the cross-diffusion and the mass condition of initial data or the logistic source. When \(r=\mu =0\), in the radially symmetric setting, Fujie and Senba [16] proved that the corresponding homogeneous Neumann problem of (1.3) possesses a unique globally bounded classical solution if either \(N\le 3\) or \(N=4\) and \(\Vert n_{0}\Vert _{L^{1}(\Omega )}<\frac{(8\pi )^{2}}{\chi }\), and they claimed that the classical solution in will be blowing up in finite or infinite time if \(N=4\) and \(\Vert n_{0}\Vert _{L^{1}(\Omega )}\in \left( \frac{(8\pi )^{2}}{\chi },\infty \right) \left\{ j\cdot \frac{(8\pi )^{2}}{\chi }|j\in N\right\} \) [17]. These results revealed that the qualities \(\Vert n_{0}\Vert _{L^{1}(\Omega )}=\frac{(8\pi )^{2}}{\chi }\) and \(N=4\) are critical mass and dimension in distinguishing the global boundedness and blow-up of solutions, respectively.

Models (1.1)–(1.3) are supposed that there is no interplay between cells/chemicals and their ambient surroundings. However, some experimental observations indicate that the motion of cells also can be substantially influenced by the surrounding fluid [37]. For instance, populations of aerobic bacteria suspended in sessile drops of water may exhibit quite a complex but structured collective dynamics, inter alia involving the spontaneous formation of plume-like aggregates [10, 12]. Considering the interaction, some researchers lately have investigated the following chemotaxis model

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{t}+u\cdot \nabla n=\Delta n-\nabla \cdot (n\chi (v)\nabla v), & x\in \Omega ,t>0,\\ v_{t}+u\cdot \nabla v=\Delta v-ng(v), & x\in \Omega ,t>0,\\ u_{t}+\kappa (u\cdot \nabla )u=\Delta u+\nabla P+n\nabla \phi ,~~\nabla \cdot u=0,& x\in \Omega ,t>0,\\ \end{array}\right. } \end{aligned}$$
(1.4)

where \(n=n(x,t)\) and \(v=v(x,t)\) are defined as above. Here, \(u=u(x,t)\), \(P=P(x,t)\), \(\phi =\phi (x)\) and \(\kappa \in {\mathbb {R}}\) represent the fluid velocity field, the associated pressure of the fluid, the potential of the gravitational field and the strength of the nonlinear fluid convection correspondingly. The fluid velocity evolution may be characterized by the Stokes equation (\(\kappa =0\)) instead of the full Navier–Stokes equation (\(\kappa =1\)) if the fluid flow is comparatively slow.

In the case of no-flux boundary condition for the signal: There have been quite a comprehensive analytical results on system (1.4) at the level of solvability and stabilization when it is posed on a bounded domain \(\Omega \). For instance, when the fluid flow remains small (i.e., \(\kappa =0\)), Lorz [30] firstly gave the analytical result involving the local existence of weak solution in two or three dimensional setting. However, when the fixed number \(\kappa \in {\mathbb {R}}/\{0\}\), the fluid motion is governed by Navier–Stokes equations with nonlinear convection, and Winkler [48] proved the system admits a unique global classical solution in the case of \(N= 2\). While in the case of \(N=3\), global weak solutions have been established in [49], and the large time asymptotic behavior was discovered in [50]. Based on the literature [48], Winkler [53] showed that in the two-dimensional case such a classical solution will stabilize to a spatially uniform equilibrium.

In the case of realistic boundary condition for the signal: As pointed in Tuval et al. [37], the boundary conditions on n, v and u are central to the global flows and possible singularities. Some recent analytical results intended to introduce suitably nontrivial boundary conditions, in particular, for the signal concentration v. For instance, taking into account for the exchange of oxygen between the drop and its environment, the Robin boundary condition for v was proposed by Braukhoff [6] and then the corresponding solvability was further investigated by [7, 54]. The inhomogeneity of Robin boundary conditions restrained the use of the standard semigroup argument and the maximal Sobolev regularity. However, these obstacles have been overcome in these works by introducing a Lions–Magenes-type transformation, which in particular transformed the Robin signal boundary condition into the usual no-flux boundary value at the cost of some extra terms appeared in the equations. On the other hand, on the water-air surface, the oxygen concentration outside the drop can be assumed equal to its saturation value \(v_{*}\) inside the water to match the experiment in Tuval et al. [37]. This assumption motivated no-flux/saturation/no-slip boundary conditions

$$\begin{aligned} (\nabla n-n\nabla v)\cdot \nu =0,~~~v=v_{*},~~~u=0,~~~~~x\in \partial \Omega ,~t>0. \end{aligned}$$
(1.5)

Under such boundary conditions, Wang et al. [38] developed a local energy method to show the existence of globally defined generalized solution to the three-dimensional system (1.4) with \(\kappa =0\). When the system (1.4) involves logistic source (i.e., \(f(n)=rn-\mu n^{2}\)), Black and Wu [5] showed the Stokes variant of the system (1.4) has at least one global weak solution for any suitably regular triplet of initial data. Very recently, they [4] extended the previously known result on time-global existence of weak solutions for the Stokes variant to the full Navier–Stokes setting and investigated the eventual regularity properties in the slightly more restrictive setting of \(v_{*}\) being also constant in space. They showed that sufficiently strong logistic influence, in the sense that for \(\omega >0\) and \(\mu _{0}>0\), there is some \(\eta =\eta (\omega ,\mu _{0},v_{*})>0\) with the property that whenever

$$\begin{aligned} \mu _{0}\le \mu ~~~and~~~\frac{r}{\min \{\mu ,\mu ^{\omega +\frac{N+6}{6}}\}}<\eta \end{aligned}$$
(1.6)

are satisfied the global weak solution eventually becomes a smooth and classical solution waiting time depending on \(\omega ,\mu _{0},\eta ,v_{*}\) and the initial data. In the nonlinear cell diffusion case (i.e., \(n_{t}+u\cdot \nabla n=\Delta n-\nabla \cdot (n\chi (v)\nabla v)\) is replaced by \(n_{t}+u\cdot \nabla n=\nabla \cdot (n^{m-1}\nabla n)-\nabla \cdot (n\chi (v)\nabla v)\) in (1.4)), the boundary conditions (1.5) becomes the following setting

$$\begin{aligned} (n^{m-1}\nabla n-n\nabla v)\cdot \nu =0,~~~v=v_{*},~~~u=0,~~~~~x\in \partial \Omega ,~t>0. \end{aligned}$$
(1.7)

In the above boundary conditions, Wang et al. [39] constructed the global mass-preserving solutions for system (1.4) with \(\kappa =0\) and \(m>\frac{3N-2}{2N}\). Wu and Xiang [55] established the existence of globally bounded weak solutions in the same setting. Recently, Black and Winkler [3] also investigated such boundary conditions on the full chemotaxis-Navier–Stokes system and obtained global weak solutions which either may be unbounded for \(m>\frac{7}{6}\) and \(N=3\). Very recently, the positive effect of the indirect signal production mechanism on the global solvability of the three-dimensional (3D) chemotaxis-Navier–Stokes system under Dirichlet signal boundary condition has been considered in our work [29]. Nevertheless, the knowledge on regularity properties of solution for the 3D chemotaxis-Navier–Stokes system with indirect signal production mechanism under Dirichlet signal boundary condition has not yet exceeded some information on fairly basic integrability features as derived in [29]. This inspires us to ask the following interesting and significant question: Will the indirect signal production mechanism genuinely contribute to the eventual smoothness and boundedness of solution to the 3D chemotaxis-Navier–Stokes system involving Dirichlet signal boundary condition?

Motivations and main results

Inspired by the outstanding works [3, 11, 42, 49, 52], in this paper, we investigate the following chemotaxis-Navier–Stokes with indirect signal production

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{t}+u\cdot \nabla n=\Delta n-\chi \nabla \cdot (n\nabla v)+rn-\mu n^{\alpha }, & x\in \Omega ,t>0,\\ v_{t}+u\cdot \nabla v=\Delta v-wv, & x\in \Omega ,t>0,\\ w_{t}+u\cdot \nabla w=\Delta w-w+n, & x\in \Omega ,t>0,\\ u_{t}+\kappa (u\cdot \nabla )u=\Delta u+\nabla P+n\nabla \phi ,~~\nabla \cdot u=0,& x\in \Omega ,t>0,\\ (\nabla n-n\nabla v)\cdot \nu =0,~~v=v_{*},~~\nabla w\cdot \nu =0,~~u=0,& x\in \partial \Omega ,t>0,\\ n(x,0)=n_{0}(x),~~v(x,0)=v_{0}(x),~~w(x,0)=w_{0}(x),~~u(x,0)=u_{0}(x),& x\in \Omega \end{array}\right. } \end{aligned}$$
(1.8)

in a bounded domain \(\Omega \subset {\mathbb {R}}^3\) with smooth boundary \(\partial \Omega \), where \(\nu \) denotes the outward normal vector on \(\partial \Omega \), \(\chi >0\), \(r\in {\mathbb {R}}\), \(\mu >0\) and \(\alpha \in (1,2)\). To prepare a precise presentation of our main results, throughout this work, we assume that the given gravitational potential function \(\phi \) fulfills

$$\begin{aligned} \phi \not \equiv 0\;\textrm{in}\;\Omega \quad \textrm{and}\quad \phi \in W^{2,\infty }(\Omega ), \end{aligned}$$
(1.9)

and the time constant function \(v_{*}\) satisfies

$$\begin{aligned} v_{*}\in C^{2}({\bar{\Omega }})~~with~~v_{*}\ge 0, \end{aligned}$$
(1.10)

as well as that the initial data \(n_{0}\), \(v_{0}\), \(w_{0}\), \(u_{0}\) satisfy

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{0}\in C^{\vartheta }({\bar{\Omega }})~~for~some~\vartheta \in (0,1), ~~~ with~n_{0}\ge 0~and~n_{0}\not \equiv 0~in~{\bar{\Omega }},\\ v_{0}\in W^{1,\infty }(\Omega ) ~~~with~v_{0}\ge 0~in~{\bar{\Omega }},\\ w_{0}\in W^{1,\infty }(\Omega ) ~~~with~w_{0}\ge 0~in~{\bar{\Omega }},\\ u_{0}\in D(A^{\beta })~~for~some~\beta \in (\frac{3}{4},1),\\ \end{array}\right. } \end{aligned}$$
(1.11)

where \(A=-P\Delta \) denotes the realization of the Stokes operator in \(L^{2}(\Omega ;{\mathbb {R}}^3)\) defined on its domain \(D(A):=W^{2,2}(\Omega ;{\mathbb {R}}^3)\cap W_{0}^{1,2}(\Omega ;{\mathbb {R}}^3)\cap L_{\sigma }^{2}(\Omega ;{\mathbb {R}}^3)\) with \(L_{\sigma }^{2}(\Omega ;{\mathbb {R}}^3):=\{\varphi \in L^{2}(\Omega ;{\mathbb {R}}^3)|\nabla \cdot \varphi =0\}\), and P represents the Helmholtz projection of \(L^{2}(\Omega ;{\mathbb {R}}^3)\) onto \(L_{\sigma }^{2}(\Omega ;{\mathbb {R}}^3)\). Since \(\Omega \) is bounded, we know that A is self-adjoint and sectorial in \(L_{\sigma }^{2}(\Omega ;{\mathbb {R}}^3)\), and possesses densely defined self-adjoint fractional powers \(A^{\delta }\) for arbitrary \(\delta \in {\mathbb {R}}\) [35].

Within this setting, the following result on global existence of weak solution for problem (1.8) has been established in our recent work [29].

Proposition 1.1

Let \(\Omega \) be a bounded domain in \({\mathbb {R}}^3\) with smooth boundary, and let \(\chi >0\), \(r\in {\mathbb {R}}\), \(\mu >0\), \(\alpha \in \left( \frac{4}{3},2\right) \) and \(\kappa \ne 0\). Then for given \(\phi \) fulfilling (1.9), \(v_{*}\) satisfies (1.10) and initial data \((n_0,v_0,w_0,u_0)\) complying with (1.11), the system (1.8) admits at least one global weak solution (nvwu) in the sense of Definition 2.1 below.

Remark 1.1

In comparison with the existing result on the chemotaxis-Navier–Stokes system with direct signal production,Proposition 1.1 rigorously confirms that the indirect signal production mechanism genuinely facilitates the global solvability of the 3D chemotaxis-Navier–Stokes system. Indeed, the global existence of a weak solution to the 3D chemotaxis-Navier–Stokes system with quadratic degradation (i.e., \(\alpha =2\)) was obtained in [4, Theorem 1.1], while Proposition 1.1 in this work established the global solvability in the same sense as that of [4] to the 3D chemotaxis-Navier–Stokes system (1.8) with suitably weak logistic-type degradation (i.e., \(\alpha \in (\frac{4}{3},2)\)). In addition, Proposition 1.1 also holds for \(N=2\).

Before stating our main result, let us introduce some positive constants which will usually appear in the statement of our main result and its proof. First of all, since \(\alpha \in (\frac{5}{3},2)\), we have \(\frac{3\alpha -4}{4}\in (\frac{1}{4},\frac{1}{2})\) and \(\frac{3\alpha -3}{8}\in (\frac{1}{4},\frac{3}{8})\) as well as \(\frac{3\alpha -4}{4}-\frac{3\alpha -3}{8}=\frac{3\alpha -5}{8}>0\). This enables us to conclude that for some \(\beta _0\in (\frac{3\alpha -3}{8},\frac{3\alpha -4}{4})\subset (\frac{1}{4},\frac{1}{2})\),

$$\begin{aligned} 2\cdot \frac{1-2\beta _0}{2}-\frac{3}{\frac{2}{3-\alpha }}=1-2\beta _0-\frac{9-3\alpha }{2} >1-2\cdot \frac{3\alpha -4}{4}-\frac{9-3\alpha }{2} =-\frac{3}{2} \end{aligned}$$
(1.12)

and

$$\begin{aligned} 2\beta _0-\frac{3}{2}>2\cdot \frac{3\alpha -3}{8}-\frac{3}{2}=-\frac{3}{\frac{4}{3-\alpha }} \end{aligned}$$
(1.13)

as well as

$$\begin{aligned} 2\cdot \frac{1+2\beta _0}{2}-\frac{3}{2}>2\cdot \frac{1+2\cdot \frac{3\alpha -3}{8}}{2}-\frac{3}{2} =1-\frac{3}{\frac{4}{3-\alpha }}. \end{aligned}$$
(1.14)

The inequality (1.12) allows us to use [44, Lemma 3.2] to find \(M_1=M_1(\beta _0,\alpha ,\Omega )>0\) such that

$$\begin{aligned} \Vert A^{-\frac{1-2\beta _0}{2}}\varphi \Vert _{L^2(\Omega )}\le M_1\Vert \varphi \Vert _{L^\frac{2}{3-\alpha }(\Omega )} \qquad \mathrm{for\;all\;}\varphi \in L^\frac{2}{3-\alpha }(\Omega ). \end{aligned}$$
(1.15)

Conjunction with the inequalities (1.13) and (1.14) as well as some embedding properties of the Stokes operator and its fractional powers (see [19, 20]) warrants that \(D(A^{\beta _0}_2)\hookrightarrow L^\frac{4}{3-\alpha }(\Omega )\) and \(D\left( A^{\frac{1+2\beta _0}{2}}_2\right) \hookrightarrow W^{1,\frac{4}{3-\alpha }}(\Omega )\). This amounts to say, we have

$$\begin{aligned} \Vert \varphi \Vert _{L^\frac{4}{3-\alpha }(\Omega )}\le M_2\Vert A^{\beta _0}\varphi \Vert _{L^2(\Omega )} \qquad \mathrm{for\;all\;}\varphi \in D(A^{\beta _0}_2) \end{aligned}$$
(1.16)

and

$$\begin{aligned} \Vert \nabla \varphi \Vert _{L^\frac{4}{3-\alpha }(\Omega )}\le M_3\Vert A^{\frac{1+2\beta _0}{2}}\varphi \Vert _{L^2(\Omega )} \qquad \mathrm{for\;all\;}\varphi \in D(A^{\frac{1+2\beta _0}{2}}_2) \end{aligned}$$
(1.17)

with \(M_2=M_2(\beta _0,\alpha ,\Omega )>0\) and \(M_3=M_3(\beta _0,\alpha ,\Omega )>0\). In addition, for the above \(\alpha \in (\frac{5}{3},2)\) and \(\beta _0\in (\frac{3\alpha -3}{8},\frac{3\alpha -4}{4})\), it is easy to see that \(\frac{1}{4}<\beta _0<\frac{1+2\beta _0}{2}<1\), which enables us to apply the fact that \(D(A^{\alpha _1}_2)\hookrightarrow D(A^{\alpha _2}_2)\) for \(0<\alpha _2\le \alpha _1<1\) (see [21, p.25]) to conclude that \(D\left( A^{\frac{1+2\beta _0}{2}}_2\right) \hookrightarrow D(A^{\beta _0}_2)\). That is to say, there exists \(M_4=M_4(\beta _0,\Omega )>0\) such that

$$\begin{aligned} M_4\Vert A^{\beta _0}\varphi \Vert _{L^2(\Omega )}\le \Vert A^{\frac{1+2\beta _0}{2}}\varphi \Vert _{L^2(\Omega )} \qquad \mathrm{for\;all\;}\varphi \in D(A^{\frac{1+2\beta _0}{2}}_2). \end{aligned}$$
(1.18)

Apart from that, the fact that the Helmholtz projection \({\mathcal {P}}\) is bounded on \(L^p(\Omega ;{\mathbb {R}}^3)\) for \(p>1\) (cf. [18, Theorems 1 and 2]) provides \(M_5=M_5(\alpha ,\Omega )>0\) such that

$$\begin{aligned} \Vert {\mathcal {P}}\varphi \Vert _{L^\frac{2}{3-\alpha }(\Omega )}\le M_5\Vert \varphi \Vert _{L^\frac{2}{3-\alpha }(\Omega )} \qquad \mathrm{for\;all\;}\varphi \in L^\frac{2}{3-\alpha }(\Omega ). \end{aligned}$$
(1.19)

Similarly to (1.18), recalling \(\Vert A^\frac{1}{2}\varphi \Vert _{L^2(\Omega )}=\Vert \nabla \varphi \Vert _{L^2(\Omega )}\), we once more infer from \(\beta _0\in (\frac{3\alpha -3}{8},\frac{3\alpha -4}{4})\subset (\frac{1}{4},\frac{1}{2})\) that there exists \(M_6=M_6(\beta _0,\Omega )>0\) such that

$$\begin{aligned} \Vert A^{\beta _0}\varphi \Vert _{L^2(\Omega )}\le M_6\Vert \nabla \varphi \Vert _{L^2(\Omega )} \qquad \mathrm{for\;all\;}\varphi \in D(A^\frac{1}{2}_2). \end{aligned}$$
(1.20)

Next, relying on (1.9), we let

$$\begin{aligned} M_\phi :=\Vert \nabla \phi \Vert _{L^\infty (\Omega )}. \end{aligned}$$
(1.21)

Finally, an application of the Poincaré inequality and the imbedding \(W_0^{1,2}(\Omega )\hookrightarrow L^6(\Omega )\) warrants the existence of \(M_P=M_P(\Omega )>0\) and \(M_S=M_S(\Omega )>0\) such that

$$\begin{aligned} M_P\Vert \varphi \Vert _{L^2(\Omega )}\le \Vert \nabla \varphi \Vert _{L^2(\Omega )}\qquad \mathrm{for\;all}\;\varphi \in W^{1,2}_0(\Omega ;{\mathbb {R}}^3) \end{aligned}$$
(1.22)

and

$$\begin{aligned} \Vert \varphi \Vert _{L^6(\Omega )}\le M_S\Vert \nabla \varphi \Vert _{L^2(\Omega )}\qquad \mathrm{for\;all}\;\varphi \in W^{1,2}_0(\Omega ;{\mathbb {R}}^3). \end{aligned}$$
(1.23)

In this context, our main result reads as follows.

Theorem 1.1

Let the conditions of Proposition 1.1 hold. Furthermore, if \(v_{*}\ge 0\) is constant, \(\alpha \) and r satisfy

$$\begin{aligned} \alpha \in \left( \frac{5}{3},2\right) ~~and~~r\le \min \{{\tilde{\eta }}^{*},{\tilde{\eta }}^{**}\}\cdot \min \left\{ \mu ,\mu ^\frac{2}{3-\alpha }\right\} \end{aligned}$$

with

$$\begin{aligned} {\tilde{\eta }}^{*}:=\min \left\{ \frac{\left( 1- e^{\frac{M_4^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{2^\frac{5\alpha -3}{3-\alpha }|\Omega |^{\alpha -1}\left( \kappa M_1^2M_2M_3M_5^2M_\phi \right) ^\frac{2(\alpha -1)}{3-\alpha }},1\right\} \end{aligned}$$

and

$$\begin{aligned} {\tilde{\eta }}^{**}:=\min \left\{ \frac{\left( 1-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{2^\frac{6\alpha -4}{3-\alpha }|\Omega |^\frac{5(\alpha -1)}{3(3-\alpha )}(\kappa M_1M_2M_3M_5M_6M_\phi M_S) ^\frac{2(\alpha -1)}{3-\alpha }\left( 2-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }},1\right\} , \end{aligned}$$

where constants \(M_i\;(i=1,2,...,6),M_\phi ,M_P,M_S>0\) are specified in (1.15)–(1.23), then one can find \(t_0=t_0(r,\mu ,\alpha ,n_0,v_0,w_0, u_0,\Omega )>0\) such that the global weak solution (nvwu) of the problem (1.8) provided by Proposition 1.1 enjoys the following properties satisfy

$$\begin{aligned} (n,v,w,u)\in \left[ C^{2,1}({\bar{\Omega }}\times [t_0,\infty ))\right] ^3\times C^{2,1}({\bar{\Omega }}\times [t_0,\infty );{\mathbb {R}}^3), \end{aligned}$$
(1.24)

and such that with some \(P\in C^{1,0}({\bar{\Omega }}\times [t_0,\infty ))\), the quintuple (nvwuP) forms a classical solution of (1.8) in \({\bar{\Omega }}\times [t_0,\infty )\). In addition, the problem (1.8) possesses a bounded absorbing set in \(L^\infty (\Omega )\times W^{1,\infty }(\Omega )\times W^{1,\infty }(\Omega )\times L^\infty (\Omega ;{\mathbb {R}}^3)\) in the sense that there exists \(C=C(\chi ,r,\mu ,\alpha ,\kappa ,\Omega )>0\) such that any such solution fulfills

$$\begin{aligned} \limsup _{t\nearrow \infty } \left\{ \Vert n(\cdot ,t)\Vert _{L^\infty (\Omega )}+\Vert v(\cdot ,t)\Vert _{W^{1,\infty }(\Omega )}+\Vert w(\cdot ,t)\Vert _{W^{1,\infty }(\Omega )} +\Vert u(\cdot ,t)\Vert _{L^\infty (\Omega )}\right\} \le C.~~~~~~~~ \end{aligned}$$
(1.25)

Remark 1.2

  1. (i)

     Apart from the positive effect of indirect signal production mechanism on global solvability, Theorem 1.1 also shows that the indirect signal production mechanism genuinely facilitates the global eventual smoothness and bounedness of solution to the 3D chemotaxis-Navier–Stokes system. To be specific, it was asserted by a recent analytic progress (see [4, Theorem 1.2) that the global weak solution of the 3D Navier–Stokes system (1.4) with quadratic degradation (i.e., \(\alpha =2\)) involving the boundary condition (1.5) constructed in [4, Theorem 1.2] will eventually become smooth if \(\mu \ge \mu _{0}\) and \(r<\eta \cdot \min \{\mu ,\mu ^{\frac{3}{2}+\omega }\}\) for all \(\omega >0\) with suitably small \(\eta =\eta (\omega ,\Omega )>0\) (i.e., (1.6) holds), whereas in Theorem 1.1 of this work we have shown the eventual smoothness and boundedness of the weak solution gained in Theorem 1.1 for the 3D chemotaxis-Navier–Stokes problem (1.8) with sub-quadratic degradation (i.e., \(\alpha \in (\frac{5}{3},2)\)) if \(r\le \min \{{\tilde{\eta }}^{*},{\tilde{\eta }}^{**}\}\cdot \min \{\mu ,\mu ^\frac{2}{3-\alpha }\}\) for some \({\tilde{\eta }}^{*}={\tilde{\eta }}^{*}(\alpha ,\Omega )>0\) and \({\tilde{\eta }}^{**}={\tilde{\eta }}^{**}(\alpha ,\Omega )>0\). Of note, unlike the appropriately small constant \(\eta >0\) and \(\mu \ge \mu _{0}\) required in [4, Theorem 1.2], our result shows that the positive constant \(\min \{{\tilde{\eta }}^{*},{\tilde{\eta }}^{**}\}\) is not necessary to be properly small and \(\mu \ge \mu _{0}\) is not necessary, and that we can determine the explicit expression of \({\tilde{\eta }}^{*}\) and \({\tilde{\eta }}^{**}\).

  2. (ii)

    ) To the best of our knowledge, the global solvability of the 3D chemotaxis-Navier–Stokes system with indirect signal production involving Dirichlet signal boundary condition has never been touched before, thus our work fills this gap to some extent. However, there is a gap between \(\alpha \in (\frac{4}{3},2)\) and \(\alpha \in (1,2)\). Accordingly, we have to leave an open question here whether or not the range of can be relaxed to \(\alpha \in (1,2)\), and we expect that we are able to solve it in future work.

Plan of this paper

In Sect. 2, we introduce a family of regularized problems (2.21) of system (1.8) and state the local-in-time solvability in the framework of classical solution (Lemma 2.1). In addition, some useful lemmas are summarized (Lemmas 2.22.5). Section 3 is dedicated to deriving the eventual smoothness and boundedness of weak solution to the 3D chemotaxis-Navier–Stokes system (1.8).

Notations. Throughout this paper, we abbreviate the integrals \(\int \limits _{\Omega }f(x)dx\) and \(\int \limits _{0}^{t}\int \limits _{\Omega }f(x,s)dxds\) as \(\int \limits _{\Omega }f(x)\) and \(\int \limits _{0}^{t}\int \limits _{\Omega }f(x,s)\) for simplicity, and adapt standard notation by abbreviating \({\bar{\rho }}:=\frac{1}{|\Omega |}\int \limits _{\Omega }\rho \) for all \(\rho \in L^{1}(\Omega )\). Moreover, for the convenience of notation, we use symbols \(M_{i},C,C_{i}\) \((i=1,2,3...)\) to denote some universal positive constants which may vary in context, and denote them by \(M_{i}(a,b...),C(a,b...),C_{i}(a,b...)\) when we need emphasize their dependence on parameters ab...

2 Preliminaries

Firstly, Let us specify the notion of weak solution to which we will refer in the sequel. Throughout the sequel, for vectors \(v\in {\mathbb {R}}^3\) and \(w\in {\mathbb {R}}^3\), we let v \(\otimes \) w denote the matrix \((a_{ij})_{i,j\in \{1,2,3\}}\in {\mathbb {R}}^{3\times 3}\) defined on setting \(a_{ij}:=v_{i}w_{j}\) for \(i,j\in \{1,2,3\}\). In addition, we introduce the divergence-free spaces \(W_{0,\sigma }^{1,1}(\Omega ;{\mathbb {R}}^N):=W_{0}^{1,1}(\Omega ;{\mathbb {R}}^N)\cap L_{\sigma }^{2}(\Omega )\), which appears a few times during our investigations.

Definition 2.1

Let \(\alpha \in (\frac{4}{3},2)\). By a global weak solution of system (1.8) we mean a quadruple (nvwu) of functions

$$\begin{aligned} {\left\{ \begin{array}{ll} n\in L_{loc}^{\alpha }({\bar{\Omega }}\times [0,\infty ))\cap L_{loc}^{1}([0,\infty );W^{1,1}(\Omega )),\\ v\in L_{loc}^{1}([0,\infty );W^{1,1}(\Omega )),\\ w\in L_{loc}^{1}([0,\infty );W^{1,1}(\Omega )),\\ u\in L_{loc}^{1}([0,\infty );W_{0,\sigma }^{1,1}(\Omega ;{\mathbb {R}}^3)) \end{array}\right. } \end{aligned}$$
(2.1)

such that \(n\ge 0\), \(v\ge 0\) and \(w\ge 0\) a.e. in \(\Omega \times (0,\infty )\),

$$\begin{aligned} & vw\in L_{loc}^{1}({\bar{\Omega }}\times [0,\infty )), \end{aligned}$$
(2.2)
$$\begin{aligned} & \{n\nabla v,~nu,~vu,~wu\}\subset L_{loc}^{1}({\bar{\Omega }}\times [0,\infty );{\mathbb {R}}^3),~~~u\otimes u\in L_{loc}^{1}({\bar{\Omega }}\times [0,\infty );{\mathbb {R}}^{3\times 3}), \end{aligned}$$
(2.3)

that \(\nabla \cdot u=0\) a.e. in \(\Omega \times (0,\infty )\), and that

$$\begin{aligned} -\int \limits _{0}^{\infty }\int \limits _{\Omega }n\varphi _{t}-\int \limits _{\Omega }n_{0}\varphi (\cdot ,0)= & -\int \limits _{0}^{\infty }\int \limits _{\Omega } \nabla n\cdot \nabla \varphi +\int \limits _{0}^{\infty }\int \limits _{\Omega }n\nabla v\cdot \nabla \varphi +r\int \limits _{0}^{\infty }\int \limits _{\Omega }n\varphi \nonumber \\ & \quad -\mu \int \limits _{0}^{\infty }\int \limits _{\Omega }n^{\alpha }\varphi +\int \limits _{0}^{\infty }\int \limits _{\Omega }nu\cdot \nabla \varphi \end{aligned}$$
(2.4)

for all \(\varphi \in C_{0}^{\infty }({\bar{\Omega }}\times [0,\infty ))\),

$$\begin{aligned} -\int \limits _{0}^{\infty }\int \limits _{\Omega }v\varphi _{t}-\int \limits _{\Omega }v_{0}\varphi (\cdot ,0)= & -\int \limits _{0}^{\infty }\int \limits _{\Omega } \nabla v\cdot \nabla \varphi -\int \limits _{0}^{\infty }\int \limits _{\Omega }vw\varphi +\int \limits _{0}^{\infty }\int \limits _{\Omega }vu\cdot \nabla \varphi \end{aligned}$$
(2.5)

for all \(\varphi \in C_{0}^{\infty }({\bar{\Omega }}\times [0,\infty ))\), and

$$\begin{aligned} -\int \limits _{0}^{\infty }\int \limits _{\Omega }w\varphi _{t}-\int \limits _{\Omega }w_{0}\varphi (\cdot ,0)= & -\int \limits _{0}^{\infty }\int \limits _{\Omega } \nabla w\cdot \nabla \varphi -\int \limits _{0}^{\infty }\int \limits _{\Omega }w\varphi \nonumber \\ & +\int \limits _{0}^{\infty }\int \limits _{\Omega }n\varphi +\int \limits _{0}^{\infty }\int \limits _{\Omega }wu\cdot \nabla \varphi \end{aligned}$$
(2.6)

for all \(\varphi \in C_{0}^{\infty }({\bar{\Omega }}\times [0,\infty ))\), as well as

$$\begin{aligned} -\int \limits _{0}^{\infty }\int \limits _{\Omega }u\cdot \varphi _{t}-\int \limits _{\Omega }u_{0}\cdot \varphi (\cdot ,0)= & -\int \limits _{0}^{\infty }\int \limits _{\Omega } \nabla u\cdot \nabla \varphi +\int \limits _{0}^{\infty }\int \limits _{\Omega }(u\otimes u)\cdot \nabla \varphi \nonumber \\ & +\int \limits _{0}^{\infty }\int \limits _{\Omega }n\nabla \phi \cdot \psi \end{aligned}$$
(2.7)

for all \(\varphi \in C_{0}^{\infty }(\Omega \times [0,\infty );{\mathbb {R}}^3)\) satisfying \(\nabla \cdot \varphi \equiv 0\).

In order to gain a global weak solutions of (1.8) through a suitable approximation procedure, we employ the approaches used in [38, 44] to regularize the system (1.8). Namely, we fix a family \((\rho _{\varepsilon })_{\varepsilon \in (0,1)}\subset C_{0}^{\infty }(\Omega )\) of smooth cut-off functions satisfying

$$\begin{aligned} 0\le \rho _{\varepsilon }(x)\le 1~~~for~all~x\in \Omega ~~~such~that~\rho _{\varepsilon }\nearrow 1~as~\varepsilon \searrow 0, \end{aligned}$$
(2.8)

and introduce the corresponding family of approximating problems to (1.8) given by

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{\varepsilon t}+u_{\varepsilon }\cdot \nabla n_{\varepsilon }=\Delta n_{\varepsilon }-\nabla \cdot {(\rho _{\varepsilon }f_{\varepsilon }(n_{\varepsilon })n_{\varepsilon }\nabla v_{\varepsilon })}+rn_{\varepsilon }-\mu n_{\varepsilon }^{\alpha }, & x\in \Omega ,t>0,\\ v_{\varepsilon t}+u_{\varepsilon }\cdot \nabla v_{\varepsilon }=\Delta w_{\varepsilon }-g_{\varepsilon }(w_{\varepsilon })v_{\varepsilon }, & x\in \Omega ,t>0,\\ w_{\varepsilon t}+u_{\varepsilon }\cdot \nabla w_{\varepsilon }=\Delta w_{\varepsilon }-w_{\varepsilon }+n_{\varepsilon }, & x\in \Omega ,t>0,\\ u_{\varepsilon t}+(Y_{\varepsilon }u_{\varepsilon }\cdot \nabla )u_{\varepsilon }=\Delta u_{\varepsilon }+\nabla P_{\varepsilon }+n_{\varepsilon }\nabla \phi ,~~\nabla \cdot u_{\varepsilon }=0,& x\in \Omega ,t>0,\\ \nabla n_{\varepsilon }\cdot \nu =0,~~v_{\varepsilon }=v_{*}(x),~~\nabla w_{\varepsilon }\cdot \nu =0,~~u_{\varepsilon }=0,& x\in \partial \Omega ,t>0,\\ n_{\varepsilon }(x,0)=n_{0}(x),~~v_{\varepsilon }(x,0)=v_{0}(x),~~w_{\varepsilon }(x,0)=w_{0}(x),~~u_{\varepsilon }(x,0)=u_{0}(x),& x\in \Omega , \end{array}\right. } \end{aligned}$$
(2.9)

with \(f_{\varepsilon }(s):=\frac{1}{(1+\varepsilon s)^{3}}\) and \(g_{\varepsilon }(s):=\frac{s}{1+\varepsilon s}\) for \(s\ge 0\) and \(\varepsilon \in (0,1)\), where \(Y_{\varepsilon }\) denotes the standard Yosida approximation [35] defined by

$$\begin{aligned} Y_{\varepsilon }\varphi :=(1+\varepsilon A)^{-1}\varphi ~~~for~all~\varphi \in L_{\sigma }^{2}(\Omega ). \end{aligned}$$
(2.10)

Applying the well-established arguments involving the fixed point theorem, the standard regularity theory of parabolic and Stokes equation and the maximum principle, we are able to prove that the regularized system (2.9) is locally solvable in a classical sense for each \(\varepsilon \in (0,1)\). We omit the proof for simplicity and refer the reader to [48, Lemma 2.1] and [5, Lemma 3.1] for its more details.

Lemma 2.1

Let \(\Omega \subset {\mathbb {R}}^3\) be a bounded domain with smooth boundary, \(r\in {\mathbb {R}}\), \(\mu >0\), \(\chi >0\) and \(\kappa \ne 0\). Assume that \(v_{*}\) and \(\phi \) satisfy (1.10) and (1.9), respectively. Then for each \(\varepsilon \in (0,1)\) and \((n_{0},v_{0},w_{0},u_{0})\) fulfilling (1.11), there exists a maximal existence time \(T_{max,\varepsilon }\in (0,\infty ]\) and a uniquely determined quintuple \((n_{\varepsilon },v_{\varepsilon },w_{\varepsilon },u_{\varepsilon },P_{\varepsilon })\) of functions such that

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{\varepsilon }\in C^{0}({\bar{\Omega }}\times [0,T_{max,\varepsilon }))\cap C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon })),\\ v_{\varepsilon }\in C^{0}({\bar{\Omega }}\times [0,T_{max,\varepsilon }))\cap C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon }))\cap L_{loc}^{\infty }([0,T_{max,\varepsilon });W^{1,q}(\Omega )),\\ w_{\varepsilon }\in C^{0}({\bar{\Omega }}\times [0,T_{max,\varepsilon }))\cap C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon }))\cap L_{loc}^{\infty }([0,T_{max,\varepsilon });W^{1,q}(\Omega )),\\ u_{\varepsilon }\in C^{0}({\bar{\Omega }}\times [0,T_{max,\varepsilon });{\mathbb {R}}^3)\cap C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon });{\mathbb {R}}^3),\\ P_{\varepsilon }\in C^{2,1}({\bar{\Omega }}\times (0,T_{max,\varepsilon })), \end{array}\right. } \end{aligned}$$
(2.11)

which classically solves (2.9) and fulfills \(n_{\varepsilon }\ge 0\), \(v_{\varepsilon }\ge 0\) and \(w_{\varepsilon }\ge 0\) in \({\bar{\Omega }}\times (0,T_{max,\varepsilon })\). Moreover, if \(T_{max,\varepsilon }<\infty \), then for all \(\beta \in (\frac{3}{4},1)\), we have

$$\begin{aligned} \limsup _{t\nearrow T_{max}}(\Vert n_{\varepsilon }(\cdot ,t)\Vert _{L^{\infty }(\Omega )}+\Vert v_{\varepsilon }(\cdot ,t)\Vert _{W^{1,q}(\Omega )}+\Vert w_{\varepsilon }(\cdot ,t)\Vert _{W^{1,q}(\Omega )}+\Vert A^{\beta }u_{\varepsilon }(\cdot ,t)\Vert _{L^{2}(\Omega )})=\infty \end{aligned}$$
(2.12)

Then, we recall the following well-known Gagliardo–Nirenberg interpolation inequality which will be frequently used in the forthcoming proofs (see [14]).

Lemma 2.2

Let \(\Omega \subset {\mathbb {R}}^N~(N\ge 1)\) be a bounded domain with smooth boundary. Suppose that \(1\le p,q\le \infty \) satisfying \(p(N-q)<Nq\) and \(r\in (0,p)\). Then, there exists \(C_{GN}=C_{GN}(p,q,r,N,\Omega )>0\) such that

$$\begin{aligned} \Vert u\Vert _{L^{p}(\Omega )}\le C_{GN}(\Vert \nabla u\Vert _{L^{q}(\Omega )}^{\theta }\Vert u\Vert _{L^{r}(\Omega )}^{1-\theta }+\Vert u\Vert _{L^{r}(\Omega )}),~~~\forall u\in W^{1,q}(\Omega )\cap L^{r}(\Omega ), \end{aligned}$$
(2.13)

where \(\theta =\frac{\frac{N}{r}-\frac{N}{p}}{1+\frac{N}{r}-\frac{N}{q}}\in (0,1)\).

In addition, we state the following auxiliary lemma on boundedness in a linear differential inequality which will be utilized in several places below. For its proof, we refer to [42, Lemma 3.4].

Lemma 2.3

Let \(t_{0}\in {\mathbb {R}}\), \(T\in (t_{0},\infty ]\) and assume that \(y\in C^{0}([t_{0},T))\cap C^{1}((t_{0},T))\) fulfills

$$\begin{aligned} y'(t)+ay(t)\le h(t)~~~for~all~t\in (t_{0},T) \end{aligned}$$
(2.14)

with any \(a>0\) and the nonnegative function \(h\in L_{loc}^{1}({\mathbb {R}})\) for which there exist \(\tau \) and \(b>0\) such that

$$\begin{aligned} \frac{1}{\tau }\int \limits _{\tau }^{t+\tau }h(t)\le b~~~for~all~t\in (t_{0},T-\tau ), \end{aligned}$$
(2.15)

then

$$\begin{aligned} y(t)\le e^{-a(t-t_{0})}y(t_{0})+\frac{b\tau }{1-e^{-a\tau }}~~~for~all~t\in [t_{0},T). \end{aligned}$$
(2.16)

Moreover, as a straightforward consequence of the Hölder inequality, the following \(L^{p}((0,T);L^{q}(\Omega ))\) interpolation result can be deduced (see [13]).

Lemma 2.4

Let \(T>0\), and let \(p_{1},p_{2},q_{1},q_{2},p,q\ge 1\) be such that

$$\begin{aligned} \frac{1}{p}=\frac{\theta }{p_{1}}+\frac{1-\theta }{p_{2}}~~and~~\frac{1}{q}=\frac{\theta }{q_{1}}+\frac{1-\theta }{q_{2}} \end{aligned}$$

with \(\theta \in [0,1]\). Then,

$$\begin{aligned} \Vert z\Vert _{L^{p}((0,T);L^{q}(\Omega )}\le \Vert z\Vert _{L^{p_{1}}((0,T);L^{q_{1}}(\Omega )}\Vert z\Vert _{L^{p_{2}}((0,T);L^{q_{2}}(\Omega )} \end{aligned}$$
(2.17)

holds for all \(z\in L^{p_{1}}((0,T);L^{q_{1}}(\Omega )\cap L^{p_{2}}((0,T);L^{q_{2}}(\Omega )\).

Finally, we give the following lemma which plays an important role in later proof. For its detailed proofs, we refer to [44, Lemma 3.8].

Lemma 2.5

Let \(\Omega \subset {\mathbb {R}}^3\) be a bounded domain with smooth boundary. Assume that \(q\ge 1\) and

$$\begin{aligned} \lambda \in [2q+2,4q+1]. \end{aligned}$$

Then, there exists \(C=C(\lambda ,q,\Omega )>0\) such that for all \(\varphi \in C^2({\bar{\Omega }})\) fulfilling \(\varphi \cdot \frac{\partial \varphi }{\partial \nu }=0\) on \(\partial \Omega \), we have

$$\begin{aligned} \Vert \nabla \varphi \Vert _{L^\lambda (\Omega )}\le C\left\| |\nabla \varphi |^{q-1}D^2\varphi \right\| _{L^2(\Omega )}^\frac{2\lambda -6}{(2q-1)\lambda } \Vert \varphi \Vert _{L^\infty (\Omega )}^\frac{6q-\lambda }{(2q-1)\lambda } +C\Vert \varphi \Vert _{L^\infty (\Omega )}. \end{aligned}$$

3 Eventual smoothness

In this section, we devote to improving the regularity properties of the weak solutions after a large waiting time under an explicit smallness condition, in which we assume \(v_{*}\ge 0\) to not only be constant in time, but also constant in space.

First of all, we give the following estimate, which is the cornerstone of the later proof.

Lemma 3.1

Let \(r\in {\mathbb {R}}\), \(\mu >0\), \(\chi >0\), \(\kappa \ne 0\) and \(\alpha \in (\frac{5}{3},2)\). If \({\tilde{r}}>0\) is such that \({\tilde{r}}\ge r\), then we have

$$\begin{aligned} \Vert n_{\varepsilon }(\cdot ,t)\Vert _{L^{1}(\Omega )}\le 2^{\frac{1}{\alpha -1}}|\Omega |\cdot (\frac{{\tilde{r}}}{\mu })^{\frac{1}{\alpha -1}} ~~~for~all~t>\frac{\ln 2}{{\tilde{r}}(\alpha -1)}~and~\varepsilon \in (0,1) \end{aligned}$$
(3.1)

and

$$\begin{aligned} \int \limits _{t}^{t+1}\Vert n_{\varepsilon }\Vert _{L^{\alpha }(\Omega )}^{\alpha }\le 2^{\frac{1}{\alpha -1}}|\Omega |\cdot \frac{{\tilde{r}}^{\frac{1}{\alpha -1}}({\tilde{r}}+1)}{\mu ^{\frac{\alpha }{\alpha -1}}}~~~for~all~t>\frac{\ln 2}{{\tilde{r}}(\alpha -1)}~and~\varepsilon \in (0,1). \end{aligned}$$
(3.2)

Proof

Since the proof is similar to [4, Lemma 5.1] and [52, Lemma 2.1], we omit its details to avoid repetition. \(\square \)

With the help of Lemma 3.1, the following eventual smallness of \(n_{\varepsilon }\) can be obtained by suitable use of Hölder inequality.

Lemma 3.2

Let \(r\in {\mathbb {R}}\), \(\mu >0\), \(\chi >0\), \(\kappa \ne 0\) and \(\alpha \in (\frac{5}{3},2)\). Assume that

$$\begin{aligned} {\tilde{\eta }}_{1}:=\min \{\frac{\delta ^{\frac{\alpha -1}{3-\alpha }}}{2^{\frac{2}{3-\alpha }}|\Omega |^{\alpha -1}},1\},~~for~all~\delta >0, \end{aligned}$$
(3.3)

and

$$\begin{aligned} r\le {\tilde{\eta }}_{1}\cdot \min \{\mu ,\mu ^{\frac{2}{3-\alpha }}\}. \end{aligned}$$
(3.4)

Then, there exists \(t_{0}=t_{0}({\tilde{\eta }}_{1},\delta ,\mu ,\alpha )>0\) such that for all \(\delta >0\) we have

$$\begin{aligned} \int \limits _{t}^{t+1}\Vert n_{\varepsilon }(\cdot ,s)\Vert _{L^{\frac{2}{3-\alpha }}(\Omega )}^{2}ds\le \delta ~~~for~all~t\ge t_{0}~and~\varepsilon \in (0,1). \end{aligned}$$
(3.5)

Proof

For fixed \(\delta >0\), we let

$$\begin{aligned} {\tilde{\eta }}_{1}:=\min \{\frac{\delta ^{\frac{\alpha -1}{3-\alpha }}}{2^{\frac{2}{3-\alpha }}|\Omega |^{\alpha -1}},1\}, \end{aligned}$$
(3.6)

so that we have

$$\begin{aligned} 2^{\frac{2}{\alpha -1}}|\Omega |^{3-\alpha }{\tilde{\eta }}_{1}^{\frac{3-\alpha }{\alpha -1}}\le \delta . \end{aligned}$$
(3.7)

In addition, we assume that

$$\begin{aligned} r\le {\tilde{r}}:={\tilde{\eta }}_{1}\cdot \min \{\mu ,\mu ^{\frac{2}{3-\alpha }}\}. \end{aligned}$$
(3.8)

From Lemma 3.1, we know that

$$\begin{aligned} \Vert n_{\varepsilon }(\cdot ,t)\Vert _{L^{1}(\Omega )}\le 2^{\frac{1}{\alpha -1}}|\Omega |\cdot (\frac{{\tilde{r}}}{\mu })^{\frac{1}{\alpha -1}} ~~~for~all~t\ge t_{0} \end{aligned}$$
(3.9)

and

$$\begin{aligned} \int \limits _{t}^{t+1}\Vert n_{\varepsilon }\Vert _{L^{\alpha }(\Omega )}^{\alpha }\le 2^{\frac{1}{\alpha -1}}|\Omega |\cdot \frac{{\tilde{r}}^{\frac{1}{\alpha -1}}({\tilde{r}}+1)}{\mu ^{\frac{\alpha }{\alpha -1}}}~~~for~all~t\ge t_{0} \end{aligned}$$
(3.10)

with \(t_{0}=t_{0}({\tilde{\eta }}_{1},\delta ,\mu ,\alpha )=\frac{\ln 2}{{\tilde{r}}(\alpha -1)}\). By means of Lemma 2.4, (3.9) and (3.10), we have

$$\begin{aligned} \int \limits _{t}^{t+1}\Vert n_{\varepsilon }(\cdot ,s)\Vert _{L^{\frac{2}{3-\alpha }}(\Omega )}^{2}ds= & \Vert n_{\varepsilon }\Vert _{L^{2}((t,t+1);L^{\frac{2}{3-\alpha }}(\Omega ))}\nonumber \\\le & \Vert n_{\varepsilon }\Vert _{L^{\alpha }((t,t+1);L^{\alpha }(\Omega ))}^{\alpha }\Vert n_{\varepsilon }\Vert _{L^{\infty }((t,t+1);L^{1}(\Omega ))}^{2-\alpha }\nonumber \\\le & 2^{\frac{1}{\alpha -1}}|\Omega |\cdot \frac{{\tilde{r}}^{\frac{1}{\alpha -1}}({\tilde{r}}+1)}{\mu ^{\frac{\alpha }{\alpha -1}}}\cdot \{2^{\frac{1}{\alpha -1}}|\Omega |\cdot (\frac{{\tilde{r}}}{\mu })^{\frac{1}{\alpha -1}}\}^{2-\alpha }\nonumber \\\le & 2^{\frac{3-\alpha }{\alpha -1}}|\Omega |^{3-\alpha }{\tilde{r}}^{\frac{3-\alpha }{\alpha -1}}({\tilde{r}}+1)\mu ^{\frac{2}{1-\alpha }}~~~for~all~t\ge t_{0}. \end{aligned}$$
(3.11)

\(\square \)

Assume that \(\mu \in (0,1]\). From (3.8) and \(0<{\tilde{\eta }}_{1}\le 1\) we know that \({\tilde{r}}={\tilde{\eta }}_{1}\mu ^{\frac{2}{3-\alpha }}\) and \({\tilde{r}}+1={\tilde{\eta }}_{1}\mu ^{\frac{2}{3-\alpha }}+1\le {\tilde{\eta }}_{1}+1\le 2\) and

$$\begin{aligned} 2^{\frac{3-\alpha }{\alpha -1}}|\Omega |^{3-\alpha }{\tilde{r}}^{\frac{3-\alpha }{\alpha -1}}({\tilde{r}}+1)\mu ^{\frac{2}{1-\alpha }}\le & 2^{\frac{3-\alpha }{\alpha -1}}|\Omega |^{3-\alpha }({\tilde{\eta }}_{1}\mu ^{\frac{2}{3-\alpha }})^{\frac{3-\alpha }{\alpha -1}}\cdot 2\cdot \mu ^{\frac{2}{1-\alpha }}\nonumber \\= & 2^{\frac{3-\alpha }{\alpha -1}}|\Omega |^{3-\alpha }{\tilde{\eta }}_{1}^{\frac{3-\alpha }{\alpha -1}}\nonumber \\\le & \delta . \end{aligned}$$
(3.12)

As for the case \(\mu >1\), proceeding in a same as proving Lemma 2.2 in [14], we can achieve it. Here, we omit its details to avoid repetition.

Based on the eventual smallness of \(n_{\varepsilon }\) in Lemma 3.2 and the standard Navier–Stokes energy inequality, we can prove the following inequality.

Lemma 3.3

Let \(r\in {\mathbb {R}}\), \(\mu >0\), \(\chi >0\), \(\kappa \ne 0\) and \(\alpha \in (\frac{5}{3},2)\). Assume that \(M_{\phi },M_{P}\), and \(M_{S}\) satisfy (1.21)–(1.23). For all \(\delta >0\) we let

$$\begin{aligned} {\tilde{\eta }}_{2}:=\min \left\{ \frac{\delta ^\frac{\alpha -1}{3-\alpha } \left( 1-e^{-\frac{ M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{4^\frac{\alpha }{3-\alpha }|\Omega |^\frac{5(\alpha -1)}{3(3-\alpha )}(M_\phi M_S)^\frac{2(\alpha -1)}{3-\alpha } \left( 2-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }},1\right\} , \end{aligned}$$
(3.13)

and

$$\begin{aligned} {\tilde{r}}\le \eta _2\cdot \min \left\{ \mu ,\mu ^\frac{2}{3-\alpha }\right\} . \end{aligned}$$
(3.14)

Thus, we have

$$\begin{aligned} {\int \limits _t^{t+1}\int \limits _\Omega |\nabla u_\varepsilon |^2\le \delta }\qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1). \end{aligned}$$
(3.15)

with \(t_0=t_0({\tilde{\eta }}_{2},\delta ,r,\mu ,\alpha ,n_0,u_0,\Omega )>0\).

Proof

We test the fourth equation of (2.9) by \(2u_\varepsilon \), and use the Hölder inequality, Young’s inequality, (1.21) and (1.23), we have

$$\begin{aligned} \frac{{\text {d}}}{{{\text {d}}t}}\int \limits _{\Omega } | u_{\varepsilon } |^{2} + 2\int \limits _{\Omega } | \nabla u_{\varepsilon } |^{2} \le&\, 2\left\| {\nabla \phi } \right\| _{{L^{\infty } (\Omega )}} \left\| {n_{\varepsilon } } \right\| _{{L^{{\frac{6}{5}}} (\Omega )}} \left\| {u_{\varepsilon } } \right\| _{{L^{6} (\Omega )}} \nonumber \\ \le&\, \frac{1}{{M_{S}^{2} }}\left\| {u_{\varepsilon } } \right\| _{{L^{6} (\Omega )}}^{2} + M_{\phi }^{2} M_{S}^{2} \left\| {n_{\varepsilon } } \right\| _{{L^{{\frac{6}{5}}} (\Omega )}}^{2} \nonumber \\ \le&\, \left\| {\nabla u_{\varepsilon } } \right\| _{{L^{2} (\Omega )}}^{2} + M_{\phi }^{2} M_{S}^{2} \left\| {n_{\varepsilon } } \right\| _{{L^{{\frac{6}{5}}} (\Omega )}}^{2} \qquad {\text {for}}\;{\text {all}}\;t > 0, \end{aligned}$$
(3.16)

which in conjunction with (1.22) directly implies

$$\begin{aligned} \frac{{\text {d}}}{{{\text {d}}t}}\int \limits _{\Omega } | u_{\varepsilon } |^{2} + \frac{{M_{P}^{2} }}{2}\int \limits _{\Omega } | u_{\varepsilon } |^{2} + \frac{1}{2}\int \limits _{\Omega } | \nabla u_{\varepsilon } |^{2} \le M_{\phi }^{2} M_{S}^{2} \left\| {n_{\varepsilon } } \right\| _{{L^{{\frac{6}{5}}} (\Omega )}}^{2} \qquad {\text {for}}\;{\text {all}}\;t > 0. \end{aligned}$$
(3.17)

By means of Lemmas 3.1 and 2.4, we obtain

$$\begin{aligned} \int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2ds \le \frac{(r_++1)\cdot \max \left\{ \Vert n_0\Vert _{L^1(\Omega )}^{3-\alpha },\left( \frac{r_+}{\mu }\right) ^\frac{3-\alpha }{\alpha -1}|\Omega |^{3-\alpha }\right\} }{\mu }\qquad \mathrm{for\;all}\;t>0. \end{aligned}$$
(3.18)

Moreover, since \(\alpha \in (\frac{5}{3},2)\), we have \(\frac{2}{3-\alpha }>\frac{3}{2}>\frac{6}{5}\). Furthermore, we apply the Hölder inequality and (3.18) to find that for all \(t>0\),

$$\begin{aligned} \int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{6}{5}(\Omega )}^2ds\le & |\Omega |^\frac{3\alpha -4}{3}\int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{2}{3-\alpha }(\Omega ))}^2ds\nonumber \\\le & \frac{|\Omega |^\frac{3\alpha -4}{3}(r_++1)\cdot \max \left\{ \Vert n_0\Vert _{L^1(\Omega )}^{3-\alpha },\left( \frac{r_+}{\mu }\right) ^\frac{3-\alpha }{\alpha -1}|\Omega |^{3-\alpha }\right\} }{\mu }. \end{aligned}$$
(3.19)

Thus, applying Lemma 2.3 in (3.17) yields

$$\begin{aligned} \int \limits _\Omega |u_\varepsilon |^2\le & \Vert u_0\Vert _{L^2(\Omega )}^2 +\frac{|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2(r_++1)\cdot \max \left\{ \Vert n_0\Vert _{L^1(\Omega )}^{3-\alpha },\left( \frac{r_+}{\mu }\right) ^\frac{3-\alpha }{\alpha -1}|\Omega |^{3-\alpha }\right\} }{\mu \left( 1-e^{-\frac{M_P^2}{2}}\right) }\nonumber \\=: & C_1\qquad \mathrm{for\;all}\;t>0\;\textrm{and}\;\varepsilon \in (0,1). \end{aligned}$$
(3.20)

For any fixed \(\delta >0\) and \(\alpha \in (\frac{5}{3},2)\) as well as \(C_P,C_S>0\) specified in (1.22) and (1.23), we let

$$\begin{aligned} \delta _1:=\frac{\delta \left( 1-e^{-\frac{ M_P^2}{2}}\right) }{4|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2 \left( 2-e^{-\frac{M_P^2}{2}}\right) }, \end{aligned}$$
(3.21)

and yields we obtain directly

$$\begin{aligned} \frac{|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2\delta _1}{1-e^{-\frac{M_P^2}{2}}} +|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2\delta _1=\frac{\delta }{4}. \end{aligned}$$
(3.22)

Moreover, we set

$$\begin{aligned} \overline{\eta _1}:=\min \left\{ \frac{\delta _1^\frac{\alpha -1}{3-\alpha }}{2^\frac{2}{3-\alpha }|\Omega |^{\alpha -1}},1\right\} \end{aligned}$$

and

$$\begin{aligned} \eta _2:=\overline{\eta _1}=\min \left\{ \frac{\delta _1^\frac{\alpha -1}{3-\alpha }}{2^\frac{2}{3-\alpha }|\Omega |^{\alpha -1}},1\right\} =\min \left\{ \frac{\delta ^\frac{\alpha -1}{3-\alpha } \left( 1-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{4^\frac{\alpha }{3-\alpha }|\Omega |^\frac{5(\alpha -1)}{3(3-\alpha )}(M_\phi M_S)^\frac{2(\alpha -1)}{3-\alpha } \left( 2-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }},1\right\} , \end{aligned}$$

and assume that

$$\begin{aligned} r\le \eta _2\cdot \min \left\{ \mu ,\mu ^\frac{2}{3-\alpha }\right\} . \end{aligned}$$

Noticing that \(\eta _2:=\eta _2(\delta ,\alpha ,\Omega ):=\overline{\eta _1}(\delta _1,\alpha ,\Omega )\), from Lemma 3.2, we can know that there exists \(t_1=t_1(\eta _2,\delta ,\mu ,\alpha ,\Omega )>0\) such that

$$\begin{aligned} \int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2ds\le \delta _1 \qquad \mathrm{for\;all}\;t\ge t_1, \end{aligned}$$
(3.23)

which implies

$$\begin{aligned} \int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{6}{5}(\Omega )}^2ds\le & |\Omega |^\frac{3\alpha -4}{3}\int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2ds\nonumber \\\le & |\Omega |^\frac{3\alpha -4}{3}\delta _1 \qquad \mathrm{for\;all}\;t\ge t_1 \end{aligned}$$
(3.24)

by using the Hölder inequality and \(\alpha \in (\frac{5}{3},2)\). In addition, we can choose \(t_0=t_0(\eta _2,\delta ,r,\mu ,\alpha ,n_0,u_0,\Omega )>t_1\) large enough satisfying

$$\begin{aligned} C_1e^{-\frac{M_P^2}{2}(t_0-t_1)}\le \frac{\delta }{4}. \end{aligned}$$
(3.25)

Based on (3.24) and (3.25), employing Lemma 2.3 again in (3.17) entails

$$\begin{aligned} \int \limits _\Omega |u_\varepsilon (\cdot ,t)|^2\le & e^{-\frac{M_P^2}{2}(t-t_1)}\int \limits _\Omega |u_\varepsilon (\cdot ,t_1)|^2 +\frac{|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2\delta _1}{1-e^{-\frac{M_P^2}{2}}}\nonumber \\\le & e^{-\frac{M_P^2}{2}(t_0-t_1)}\cdot C_1+\frac{|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2\delta _1}{1-e^{-\frac{M_P^2}{2}}}\nonumber \\\le & \frac{\delta }{4}+\frac{|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2\delta _1}{1-e^{-\frac{M_P^2}{2}}} \qquad \mathrm{for\;all}\;t\ge t_0. \end{aligned}$$
(3.26)

Consequently, we integrate (3.17) in time and use (3.26), (3.24) and (3.22), we can conclude that

$$\begin{aligned} \frac{1}{2}\int \limits _t^{t+1}\int \limits _\Omega |\nabla u_\varepsilon |^2\le & \int \limits _\Omega |u_\varepsilon (\cdot ,t)|^2+M_\phi ^2M_S^2\int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{6}{5}(\Omega )}^2ds \nonumber \\\le & \frac{\delta }{4}+\frac{|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2\delta _1}{1-e^{-\frac{M_P^2}{2}}} +|\Omega |^\frac{3\alpha -4}{3}M_\phi ^2M_S^2\delta _1 \nonumber \\= & \frac{\delta }{4}+\frac{\delta }{4} \nonumber \\= & \frac{\delta }{2} \qquad \mathrm{for\;all}\;t\ge t_0\;\textrm{and}\;\varepsilon \in (0,1). \end{aligned}$$
(3.27)

This readily accomplishes the proof of (3.15). \(\square \)

Indeed, the nonlinear convection term involving in the fluid equation is highly intractable due to the typical difficulty in studying 3D Navier–Stokes equation. For the sake of overcoming this obstacle, inspired largely by the innovative and technical approach used in the proof of Lemma 3.3 in [52], we can establish the eventual smallness bound for \(u_{\varepsilon }\) in \(L^{p}(\Omega )\).

Lemma 3.4

Let \(r\in {\mathbb {R}}\), \(\mu >0\), \(\chi >0\), \(\kappa \ne 0\), \(\alpha \in (\frac{5}{3},2)\) and \({\tilde{\beta }}_{0}\in \left( \frac{3\alpha -3}{8},\frac{3\alpha -4}{4}\right) \) and \(\kappa \ne 0\). Assume that

$$\begin{aligned} {\tilde{r}}\le \min \{{\tilde{\eta }}^{*},{\tilde{\eta }}^{**}\}\cdot \min \left\{ \mu ,\mu ^\frac{2}{3-\alpha }\right\} \end{aligned}$$

with

$$\begin{aligned} {\tilde{\eta }}^{*}:=\min \left\{ \frac{\left( 1- e^{\frac{M_4^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{2^\frac{5\alpha -3}{3-\alpha }|\Omega |^{\alpha -1}\left( \kappa M_1^2M_2M_3M_5^2M_\phi \right) ^\frac{2(\alpha -1)}{3-\alpha }},1\right\} \end{aligned}$$

and

$$\begin{aligned} {\tilde{\eta }}^{**}:=\min \left\{ \frac{\left( 1-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{2^\frac{6\alpha -4}{3-\alpha }|\Omega |^\frac{5(\alpha -1)}{3(3-\alpha )}(\kappa M_1M_2M_3M_5M_6M_\phi M_S) ^\frac{2(\alpha -1)}{3-\alpha }\left( 2-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }},1\right\} , \end{aligned}$$

where positive constants \(M_i\;(i=1,2,...,6)\) and \(M_\phi ,M_p,M_S\) are specified in (1.15)–(1.23). Then, there exists \(t_0=t_0(r,\mu ,\alpha ,{\tilde{\beta }}_0,n_0,u_0,\Omega )>0\) such that

$$\begin{aligned} \Vert A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t)\Vert _{L^2(\Omega )}\le C\qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1) \end{aligned}$$
(3.28)

and

$$\begin{aligned} \int \limits _t^{t+1}\left\| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon (\cdot ,s)\right\| _{L^2(\Omega )}^2ds\le C\qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1) \end{aligned}$$
(3.29)

with \(C=C({\tilde{\beta }}_0,\alpha ,\kappa ,\Omega )>0\). In particular, for the above \(t_0\) we have

$$\begin{aligned} \Vert u_\varepsilon (\cdot ,t)\Vert _{L^{\frac{4}{3-\alpha }}(\Omega )}\le {\widetilde{C}}\qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1) \end{aligned}$$
(3.30)

and

$$\begin{aligned} \int \limits _t^{t+1}\Vert \nabla u_\varepsilon (\cdot ,s)\Vert _{L^{\frac{4}{3-\alpha }}(\Omega )}^2ds\le {\widetilde{C}}\qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1) \end{aligned}$$
(3.31)

with \({\widetilde{C}}={\widetilde{C}}(\alpha ,\kappa ,\Omega )>0\).

Proof

Using the Helmholtz projector \({\mathcal {P}}\) to the fourth equation of (2.9), testing the resulting equation \(u_{\varepsilon t}+ Au_\varepsilon ={\mathcal {P}}[-\kappa (Y_\varepsilon u_\varepsilon \cdot \nabla )u_\varepsilon +n_\varepsilon \nabla \phi ]\) against \(A^{2\tilde{\beta _0}}u_\varepsilon \), we have

$$\begin{aligned} & \frac{1}{2}\frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon |^2+\int \limits _\Omega |A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon |^2\nonumber \\ & \quad =-\kappa \int \limits _\Omega A^{2{\tilde{\beta }}_0}u_\varepsilon \cdot {\mathcal {P}}[(Y_\varepsilon u_\varepsilon \cdot \nabla )u_\varepsilon ] +\int \limits _\Omega A^{2{\tilde{\beta }}_0}u_\varepsilon \cdot {\mathcal {P}}[n_\varepsilon \nabla \phi ] \qquad \mathrm{for\;all}\;t>0. \end{aligned}$$
(3.32)

Employing the self-adjointness of fractional powers of A, Young’s inequality, (1.15), (1.19) and (1.21) to obtain

$$\begin{aligned} \int \limits _\Omega A^{2{\tilde{\beta }}_0}u_\varepsilon \cdot {\mathcal {P}}[n_\varepsilon \nabla \phi ]\le & \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +\int \limits _\Omega \left| A^{-\frac{1-2{\tilde{\beta }}_0}{2}}{\mathcal {P}}[n_\varepsilon \nabla \phi ]\right| ^2\nonumber \\\le & \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +M_1^2\left\| {\mathcal {P}}[n_\varepsilon \nabla \phi ]\right\| _{L^\frac{2}{3-\alpha }(\Omega )}^2\nonumber \\\le & \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +M_1^2M_5^2M_\phi ^2\Vert n_\varepsilon \Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2, \end{aligned}$$
(3.33)

where \(M_1, M_5, M_\phi >0\) are determined in (1.15), (1.19) and (1.21), respectively. Utilizing the Cauchy–Schwarz inequality, (1.15), (1.16), (1.17) and (1.19), we obtain that

$$\begin{aligned} & -\kappa \int \limits _\Omega A^{2{\tilde{\beta }}_0}u_\varepsilon \cdot {\mathcal {P}}[(Y_\varepsilon u_\varepsilon \cdot \nabla )u_\varepsilon ] \nonumber \\ & \quad \le \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +\kappa ^2\int \limits _\Omega \left| A^{-\frac{1-2{\tilde{\beta }}_0}{2}}{\mathcal {P}}[(Y_\varepsilon u_\varepsilon \cdot \nabla )u_\varepsilon ]\right| ^2 \nonumber \\ & \quad \le \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +\kappa ^2M_1^2\left\| {\mathcal {P}}[(Y_\varepsilon u_\varepsilon \cdot \nabla )u_\varepsilon ]\right\| _{L^\frac{2}{3-\alpha }(\Omega )}^2 \nonumber \\ & \quad \le \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +\kappa ^2M_1^2M_5^2\Vert (Y_\varepsilon u_\varepsilon \cdot \nabla )u_\varepsilon \Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2 \nonumber \\ & \quad \le \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +\kappa ^2M_1^2M_5^2\Vert Y_\varepsilon u_\varepsilon \Vert _{L^\frac{4}{3-\alpha }(\Omega )}^2 \Vert \nabla u_\varepsilon \Vert _{L^\frac{4}{3-\alpha }(\Omega )}^2 \nonumber \\ & \quad \le \frac{1}{4}\int \limits _\Omega \left| A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \right| ^2 +\kappa ^2M_1^2M_2^2M_3^2M_5^2\Vert A^{{\tilde{\beta }}_0}Y_\varepsilon u_\varepsilon \Vert _{L^2(\Omega )}^2 \Vert A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \Vert _{L^2(\Omega )}^2, \end{aligned}$$
(3.34)

where \(M_2, M_3>0\) are specified in (1.16) and (1.17), respectively. Substituting (3.33) and (3.34) into (3.32) directly shows that for each \(\varepsilon \in (0,1)\) and all \(t>0\),

$$\begin{aligned} & \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon |^2+\int \limits _\Omega |A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon |^2 \nonumber \\ & \quad \le 2\kappa ^2M_1^2M_2^2M_3^2M_5^2\Vert A^{{\tilde{\beta }}_0}Y_\varepsilon u_\varepsilon \Vert _{L^2(\Omega )}^2 \Vert A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon \Vert _{L^2(\Omega )}^2 +2M_1^2M_5^2M_\phi ^2\Vert n_\varepsilon \Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2. \end{aligned}$$
(3.35)

\(\square \)

For \(\kappa \ne 0\) and constants \(M_1,M_2,M_3,M_4,M_5,M_6,M_\phi >0\) determined in (1.15)–(1.21), we define

$$\begin{aligned} \varrho _1:=\frac{1- e^{\frac{M_4^2}{2}}}{32\kappa ^2M_1^4M_2^2M_3^2M_5^4M_\phi ^2} \end{aligned}$$
(3.36)

and

$$\begin{aligned} \varrho _2:=\frac{1}{16\kappa ^2M_1^2M_2^2M_3^2M_5^2M_6^2}. \end{aligned}$$
(3.37)

Furthermore, we choose

$$\begin{aligned} {{\tilde{\eta }}_{1}} =\min \left\{ \frac{\varrho _1^\frac{\alpha -1}{3-\alpha }}{2^\frac{2}{3-\alpha }|\Omega |^{\alpha -1}},1\right\} =\min \left\{ \frac{\left( 1- e^{\frac{M_4^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{2^\frac{5\alpha -3}{3-\alpha }|\Omega |^{\alpha -1}\left( \kappa M_1^2M_2M_3M_5^2M_\phi \right) ^\frac{2(\alpha -1)}{3-\alpha }},1\right\} \end{aligned}$$

and

$$\begin{aligned} {{\tilde{\eta }}_{2}}= & \min \left\{ \frac{\varrho _2^\frac{\alpha -1}{3-\alpha } \left( 1-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{4^\frac{\alpha }{3-\alpha }|\Omega |^\frac{5(\alpha -1)}{3(3-\alpha )}(M_\phi M_S)^\frac{2(\alpha -1)}{3-\alpha } \left( 2-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }},1\right\} \\ = & \min \left\{ \frac{ \left( 1-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }}{2^\frac{6\alpha -4}{3-\alpha }|\Omega |^\frac{5(\alpha -1)}{3(3-\alpha )}(\kappa M_1M_2M_3M_5M_6M_\phi M_S) ^\frac{2(\alpha -1)}{3-\alpha }\left( 2-e^{-\frac{M_P^2}{2}}\right) ^\frac{\alpha -1}{3-\alpha }},1\right\} . \end{aligned}$$

For the above \({\tilde{\eta }}_{1},{\tilde{\eta }}_{2}>0\), we suppose that

$$\begin{aligned} {r\le \min \{{\tilde{\eta }}_{1},{\tilde{\eta }}_{2}\}\cdot \min \left\{ \mu ,\mu ^\frac{2}{3-\alpha }\right\} ,} \end{aligned}$$
(3.38)

so that we derive from Lemma 3.2 that there exists \(t_1=t_1({\tilde{\eta }}_{1},\mu ,\alpha ,\Omega )>0\) satisfying

$$\begin{aligned} \int \limits _t^{t+1}\Vert n_\varepsilon (\cdot ,s)\Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2ds\le \varrho _1 \qquad \mathrm{for\;all}\;t\ge t_1, \end{aligned}$$
(3.39)

and find from Lemma 3.3 that there exists \(t_2=t_2({\tilde{\eta }}_{1},{\tilde{\eta }}_{2},r,\mu ,\alpha ,\tilde{\beta _{0}},n_0,u_0,\Omega )>t_1\) satisfying

$$\begin{aligned} \int \limits _{t_2}^{t_2+1}\Vert \nabla u_\varepsilon (\cdot ,s)\Vert _{L^2(\Omega )}^2ds\le \varrho _2. \end{aligned}$$
(3.40)

Now, we shall claim that (3.28)–(3.31) for \(t_0=t_0({\tilde{\eta }}_{1},{\tilde{\eta }}_{2},r,\mu ,\alpha ,\tilde{\beta _{0}},n_0,u_0,\Omega ):=t_2+1\). We first deduce from (3.40) that for each \(\varepsilon \in (0,1)\) there exists \(t_\varepsilon =t_\varepsilon ({\tilde{\eta }}_{1},{\tilde{\eta }}_{2},r,\mu ,\alpha ,\tilde{\beta _{0}},n_0,u_0,\Omega )\in (t_2,t_2+1)\) such that

$$\begin{aligned} \Vert \nabla u_\varepsilon (\cdot ,t_\varepsilon )\Vert _{L^2(\Omega )}^2\le \varrho _2, \end{aligned}$$
(3.41)

which combined with (1.23) and (3.37) shows

$$\begin{aligned} \int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t_\varepsilon )|^2\le M_6^2\varrho _2=\frac{1}{16\kappa ^2M_1^2M_2^2M_3^2M_5^2}. \end{aligned}$$
(3.42)

Let

$$\begin{aligned} S:=\left\{ T_\varepsilon >t_\varepsilon \Big |\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t)|^2 \le \frac{1}{4\kappa ^2M_1^2M_2^2M_3^2M_5^2}\qquad \mathrm{for\;all\;}t\in [t_\varepsilon ,T_\varepsilon )\right\} . \end{aligned}$$
(3.43)

Here we point out that (3.42) and the continuity of \(\Vert A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t)\Vert _{L^2(\Omega )}\) with respect to t asserted by Lemma 2.1 and Proposition 1.1 guarantee that S is indeed not empty and thus \(T_\varepsilon :=\sup S\in (0,\infty ]\) is well-defined. Clearly, it remains to make sure that \(T_\varepsilon =\infty \). To achieve it, using the facts that \(A^{{\tilde{\beta }}_0}\) commutes with \(Y_\varepsilon \) on \(D(A^{{\tilde{\beta }}_0}_2)\), and that \(\Vert Y_\varepsilon \varphi \Vert _{L^2(\Omega )}\le \Vert \varphi \Vert _{L^2(\Omega )}\) for all \(\varphi \in L^2_\sigma (\Omega )\), we infer from the definition of \(T_\varepsilon \) that for all \(t\in (t_\varepsilon ,T_\varepsilon )\),

$$\begin{aligned} \Vert A^{{\tilde{\beta }}_0}Y_\varepsilon u_\varepsilon \Vert _{L^2(\Omega )}^2 =\Vert Y_\varepsilon A^{{\tilde{\beta }}_0}u_\varepsilon \Vert _{L^2(\Omega )}^2 \le \Vert A^{{\tilde{\beta }}_0}u_\varepsilon \Vert _{L^2(\Omega )}^2 \le \frac{1}{4\kappa ^2M_1^2M_2^2M_3^2M_5^2}, \end{aligned}$$
(3.44)

and hence by (3.35), we have

$$\begin{aligned} \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon |^2+\frac{1}{2}\int \limits _\Omega |A^{\frac{1+2{\tilde{\beta }}_0}{2}}u_\varepsilon |^2 \le 2C_1^2C_5^2M_\phi ^2\Vert n_\varepsilon \Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2 \qquad \mathrm{for\;all}\;t\in (t_\varepsilon ,T_\varepsilon ), \end{aligned}$$
(3.45)

which together with (1.18) gives

$$\begin{aligned} \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon |^2+\frac{K_4^2}{2}\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon |^2 \le 2C_1^2C_5^2M_\phi ^2\Vert n_\varepsilon \Vert _{L^\frac{2}{3-\alpha }(\Omega )}^2 \qquad \mathrm{for\;all}\;t\in (t_\varepsilon ,T_\varepsilon ). \end{aligned}$$
(3.46)

Thus, applying Lemma 2.3 along with (3.39), (3.36) and (3.42) immediately reveals

$$\begin{aligned} \int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon |^2\le & e^{-\frac{M_4^2}{2}(t-t_\varepsilon )}\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t_\varepsilon )|^2 +\frac{2M_1^2M_5^2M_\phi ^2\varrho _1}{1- e^{\frac{ M_4^2}{2}}} \nonumber \\\le & \int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t_\varepsilon )|^2 +\frac{2M_1^2M_5^2M_\phi ^2\varrho _1}{1-e^{\frac{M_4^2}{2}}} \nonumber \\\le & \frac{1}{16\kappa ^2M_1^2M_2^2M_3^2M_5^2} +\frac{1}{16\kappa ^2M_1^2M_2^2M_3^2M_5^2} \nonumber \\= & \frac{1}{8\kappa ^2M_1^2M_2^2M_3^2M_5^2} \qquad \mathrm{for\;all}\;t\in [t_\varepsilon ,T_\varepsilon ). \end{aligned}$$
(3.47)

Suppose that \(T_\varepsilon <\infty \), once more making use of the continuity of \(\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t)|^2\) with respect to t ensured by Lemma 2.1, we infer from (3.47) that we can find appropriately small \(\varrho =\varrho (\kappa ,\alpha ,\Omega )>0\) such that \(\int \limits _\Omega |A^{{\tilde{\beta }}_0}u_\varepsilon (\cdot ,t)|^2\le \frac{1}{4\kappa ^2M_1^2M_2^2M_3^2M_5^2}\) for all \(t\in [t_\varepsilon ,T_\varepsilon +\delta )\), which contradicts our definition of \(T_\varepsilon \) evidently. Therefore, we conclude that we must have \(T_\varepsilon =\infty \), and that (3.28) is valid upon evident choice of \(t_0:=t_2+1>t_\varepsilon \). Moreover, on the basis of (3.28) and (3.39), the estimate (3.29) results from a straightforward time integration of (3.45). Having (3.28) and (3.29) at hand, we easily derive (3.30) and (3.31) through the use of (1.16) and (1.17). Consequently, the proof is completed.

In fact, the Dirichlet signal boundary condition for \(v_{\varepsilon }\) is highly intractable. In order to solve this obstacle, motivated by the innovative and technical approach used in the proof of Lemma 3.4 in [3], we can deal with the normal derivative \(\frac{\partial |\nabla v_{\varepsilon }|^{2}}{\partial \nu }\).

Lemma 3.5

Let \(N\ge 2\) and \(\Omega \subset {\mathbb {R}}^{N}\) be a bounded domain with \(C^{2}\)-boundary and let \(L\in {\mathbb {R}}\) denote the maximum of the curvatures on \(\partial \Omega \). Then, whenever \(\varphi \in C^{2}({\bar{\Omega }})\) and \(\varphi _{*}\in R\) are such that \(\varphi =\varphi _{*}\) on \(\partial \Omega \),

$$\begin{aligned} \frac{\partial |\nabla \varphi |^{2}}{\partial \nu }\le 2\frac{\partial \varphi }{\partial \nu }\Delta \varphi +2L|\frac{\partial \varphi }{\partial \nu }|^{2}~~~on~\partial \Omega . \end{aligned}$$
(3.48)

An application of Lemma 3.5, Hölder inequality and Young’s inequality directly, we can obtain the following inequality, which is taken from [4, Lemma 5.9].

Lemma 3.6

Let \(p\in (1,2)\). There is \(C=C(p)>0\) such that for all \(\gamma >0\), \(\vartheta >0\) and each \(\varphi \in C^{2}({\bar{\Omega }})\) with \(\varphi =0\) on \(\partial \Omega \) the inequality

$$\begin{aligned} \gamma \int \limits _{\Omega }|\nabla \varphi |^{2p}\le \vartheta \int \limits _{\Omega }|\nabla \varphi |^{2p-2}|D^{2}\varphi |^{2}+ \frac{C\gamma ^{p+1}}{\vartheta ^{p}}\int \limits _{\Omega }|\varphi |^{2p} \end{aligned}$$
(3.49)

holds.

Using Young’s inequality, Hölder inequality and Gagliardo–Nirenberg inequality, we can improve the regularity of \(w_{\varepsilon }\).

Lemma 3.7

Let \(r\in {\mathbb {R}}\), \(\mu >0\), \(\chi >0\), \(\kappa \ne 0\) and \(\alpha \in (\frac{5}{3},2)\). Then for all initial data fulfilling (1.11), we have

$$\begin{aligned} \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }\le C \qquad \mathrm{for\;all}\;t>0 \end{aligned}$$
(3.50)

with \(C=C(r,\mu ,\alpha ,n_{0},w_{0},\Omega )>0\).

Proof

Testing the third equation of (2.9) against \(w_\varepsilon ^\frac{5(\alpha -1)}{5-2\alpha }\) with \(\alpha \in (\frac{4}{3},2)\) and use the Hölder inequality to prove that

$$\begin{aligned} & \frac{5-2\alpha }{3\alpha }\frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha } +\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+\frac{5(\alpha -1)}{5-2\alpha }\int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2 \nonumber \\ & \quad \le \left( \int \limits _\Omega n_\varepsilon ^\alpha \right) ^\frac{1}{\alpha }\left( \int \limits _\Omega w_\varepsilon ^\frac{5\alpha }{5-2\alpha }\right) ^\frac{\alpha -1}{\alpha } \end{aligned}$$
(3.51)

for all \(t\in (0,T_{max})\). By means of (4.1), we can know that \(\Vert w_{\varepsilon }\Vert _{L^{1}(\Omega )}\le {\tilde{C}}\) with some constants \({\tilde{C}}>0\). Furthermore, applying the Gagliardo–Nirenberg inequality and Young’s inequality to find that

$$\begin{aligned} \left( \int \limits _\Omega w_\varepsilon ^\frac{5\alpha }{5-2\alpha }\right) ^\frac{\alpha -1}{\alpha }\le & C_1\Vert \nabla w_\varepsilon ^{\frac{3\alpha }{2(5-2\alpha )}}\Vert _{L^2(\Omega )}^\frac{2(\alpha -1)}{\alpha } \Vert w_\varepsilon ^{\frac{3\alpha }{2(5-2\alpha )}}\Vert _{L^2(\Omega )}^{\frac{4(\alpha -1)}{3\alpha }} +C_1\Vert w_\varepsilon ^{\frac{3\alpha }{2(5-2\alpha )}}\Vert _{L^\frac{4(5-2\alpha )}{3\alpha }(\Omega )}^{\frac{10(\alpha -1)}{3\alpha }} \nonumber \\ \le & C_2\left( \int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2\right) ^\frac{\alpha -1}{\alpha } \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }\right) ^\frac{2(\alpha -1)}{3\alpha }+C_3 \end{aligned}$$

with \(C_1=C_1(\Omega )>0\), \(C_2=C_2(\alpha ,\Omega )>0\) and \(C_3=C_3(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\). By means of Young’s inequality implies

$$\begin{aligned} & \left( \int \limits _\Omega n_\varepsilon ^\alpha \right) ^\frac{1}{\alpha }\left( \int \limits _\Omega w_\varepsilon ^\frac{5\alpha }{5-2\alpha }\right) ^\frac{\alpha -1}{\alpha }\nonumber \\ & \quad \le C_2\left( \int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2\right) ^\frac{\alpha -1}{\alpha } \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }\right) ^\frac{2(\alpha -1)}{3\alpha }\left( \int \limits _\Omega n_\varepsilon ^\alpha \right) ^\frac{1}{\alpha } +C_3\left( \int \limits _\Omega n_\varepsilon ^\alpha \right) ^\frac{1}{\alpha } \nonumber \\ & \quad \le \frac{5(\alpha -1)}{2(5-2\alpha )}\int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2+C_4\left( \int \limits _\Omega n_\varepsilon ^\alpha \right) \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }\right) ^\frac{2(\alpha -1)}{3}+C_3\left( \int \limits _\Omega n_\varepsilon ^\alpha +1\right) \nonumber \\ & \quad \le \frac{5(\alpha -1)}{2(5-2\alpha )}\int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2+C_4\left( \int \limits _\Omega n_\varepsilon ^\alpha +1\right) \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) +C_3\left( \int \limits _\Omega n_\varepsilon ^\alpha +1\right) ~~~~~~ \end{aligned}$$
(3.52)

with \(C_4=C_4(\alpha ,\Omega )>0\). Substituting (3.52) into (3.51) and rearranging the resulting inequality provide \(C_5=C_5(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\) such that for all \(t>0\),

$$\begin{aligned} \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha } +\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2}\int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2 \le C_7\left( \int \limits _\Omega n_\varepsilon ^\alpha +1\right) \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) . \end{aligned}$$
(3.53)

\(\square \)

Employing the Gagliardo–Nirenberg inequality and Young’s inequality again to show that there exist \(C_6=C_6(\alpha ,\Omega )>0\) and \(C_7=C_7(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\) fulfilling

$$\begin{aligned} \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\le & C_6\Vert \nabla w_\varepsilon ^{\frac{3\alpha }{2(5-2\alpha )}}\Vert _{L^2(\Omega )}^\frac{30\alpha -30}{11\alpha -5} \Vert w_\varepsilon ^{\frac{3\alpha }{2(5-2\alpha )}}\Vert _{L^\frac{2(5-2\alpha )}{3\alpha }(\Omega )}^{\frac{20-8\alpha }{11\alpha -5}} +C_6\Vert w_\varepsilon ^{\frac{3\alpha }{2(5-2\alpha )}}\Vert _{L^\frac{2(5-2\alpha )}{3\alpha }(\Omega )}^2+1 \nonumber \\ \le & C_7\left( \int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2\right) ^\frac{15\alpha -15}{11\alpha -5} +C_7, \end{aligned}$$

which, by means of an elementary inequality \((a-b)_+^\xi \ge 2^{1-\xi }a^\xi -b^\xi \) with \(a,b\ge 0\) and \(\alpha \ge 1\), enables us to take \(a:=\frac{1}{C_7}\left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) \), \(b:=1\) and \(\xi :=\frac{11\alpha -5}{15\alpha -15}>\frac{17}{15}>1\) because of \(\alpha \in (\frac{4}{3},2)\) and to arrive at

$$\begin{aligned} \int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2\ge & \left\{ \frac{1}{C_7}\left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) -1\right\} ^\frac{11\alpha -5}{15\alpha -15}_+ \nonumber \\ \ge & 2^\frac{4\alpha -10}{15\alpha -15}C_7^\frac{5-11\alpha }{15\alpha -15} \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) ^\frac{11\alpha -5}{15\alpha -15}-1. \end{aligned}$$

This in conjunction with the basic inequality \(y^p\ge ep\ln y\) for all \(y>0\) and \(p>0\) further entails

$$\begin{aligned} & \frac{\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2}\int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2}{\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1} \nonumber \\ \ge & \frac{15\alpha (\alpha -1)2^\frac{4\alpha -10}{15\alpha -15}C_7^\frac{5-11\alpha }{15\alpha -15}}{2(5-2\alpha )^2} \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) ^\frac{10-4\alpha }{15\alpha -15} -\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2\left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) } \nonumber \\ \ge & C_{8}\ln \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) -\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2}. \end{aligned}$$
(3.54)

with \(C_{8}:=\frac{\alpha e(10-4\alpha )(2C_9)^\frac{5-11\alpha }{15\alpha -15}}{(5-2\alpha )^2}>0\). Dividing the both sides of (4.53) by \(\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\) and recalling (3.54) we have

$$\begin{aligned} & \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\ln \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) +C_{8}\ln \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) \nonumber \\ & \quad \le \frac{\frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }}{\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1} +\frac{\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2}\int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2}{\int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1} +\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2} \nonumber \\ & \quad \le C_5\left( \int \limits _\Omega n_\varepsilon ^\alpha +1\right) +\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2}\qquad \mathrm{for\;all}\;t>0. \end{aligned}$$
(3.55)

In view of (4.2), we get

$$\begin{aligned} \int \limits _t^{t+1}\left\{ C_5\left( \int \limits _\Omega n_\varepsilon ^\alpha +1\right) +\frac{15\alpha (\alpha -1)}{2(5-2\alpha )^2}\right\} ds \le C_{9} \qquad \mathrm{for\;all}\;t>0 \end{aligned}$$
(3.56)

with \(C_{9}=C_{9}(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\). Applying Lemma 2.3 to (3.55) and recalling (3.56), we conclude that

$$\begin{aligned} \ln \left( \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }+1\right) \le \ln \left( \int \limits _\Omega w_0^\frac{3\alpha }{5-2\alpha }+1\right) +{\frac{C_{9}}{1-e^{-C_{8}}}} \qquad \mathrm{for\;all}\;t>0. \end{aligned}$$

Applying the exponential function to the above inequality immediately gives rise to

$$\begin{aligned} \int \limits _\Omega w_\varepsilon ^\frac{3\alpha }{5-2\alpha }\le C_{10} \qquad \mathrm{for\;all}\;t>0 \end{aligned}$$
(3.57)

with \(C_{10}=C_{10}(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\). Integrating (3.53) in time and recurring to (3.57), we can find \(C_{11}=C_{11}(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\) fulfilling

$$\begin{aligned} \int \limits _t^{t+1}\int \limits _\Omega w_\varepsilon ^\frac{7\alpha -10}{5-2\alpha }|\nabla w_\varepsilon |^2 \le C_{11}, \end{aligned}$$

and further deduce from the Gagliardo–Nirenberg inequality that for some \(C_{12}=C_{12}(\Omega )>0\) and \(C_{13}=C_{13}(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\),

$$\begin{aligned} \int \limits _t^{t+1}\int \limits _\Omega w_\varepsilon ^\frac{5\alpha }{5-2\alpha }= & \int \limits _t^{t+1}\Vert w_\varepsilon ^\frac{3\alpha }{10-4\alpha }\Vert _{L^\frac{10}{3}(\Omega )}^\frac{10}{3}ds \nonumber \\ \le & C_{12}\int \limits _t^{t+1}\left( \Vert \nabla w_\varepsilon ^\frac{3\alpha }{10-4\alpha }\Vert _{L^2(\Omega )}^2 \Vert w_\varepsilon ^\frac{3\alpha }{10-4\alpha }\Vert _{L^2(\Omega )}^\frac{4}{3} +\Vert w_\varepsilon ^\frac{3\alpha }{10-4\alpha }\Vert _{L^2(\Omega )}^\frac{10}{3}\right) ds \nonumber \\ \le & C_{13}\qquad \mathrm{for\;all}\;t>0. \end{aligned}$$
(3.58)

In order to derive the time-ultimate \(L^{\infty }\) bounds for \(n_{\varepsilon }\), we need to establish suitable regularity properties of \(\nabla v_{\varepsilon }\).

Lemma 3.8

Let \(r\in {\mathbb {R}}\), \(\mu >0\), \(\chi >0\), \(\kappa \ne 0\) and \(\alpha \in (\frac{5}{3},2)\). Suppose that the assumption on r prescribed in Lemma 3.4 remains valid. Then, for all initial data fulfilling (1.11), one can find \(t_0=t_0(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega )>0\) such that

$$\begin{aligned} \int \limits _\Omega |\nabla v_\varepsilon (\cdot ,t)|^\frac{4}{3-\alpha }\le C \qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1) \end{aligned}$$
(3.59)

with \(C=C(r,\mu ,\alpha ,\kappa ,\Omega )>0\).

Proof

From (3.30) and (3.31) we know that there exist \(C_1=C_1(\kappa ,\alpha ,\Omega )>0\) and \(t_1=t_1(r,\mu ,\alpha ,n_0,u_0,\Omega )>0\) satisfying

$$\begin{aligned} \Vert u_\varepsilon (\cdot ,t)\Vert _{L^{\frac{4}{3-\alpha }}(\Omega )}+\int \limits _t^{t+1}\Vert \nabla u_\varepsilon (\cdot ,s)\Vert _{L^{\frac{4}{3-\alpha }}(\Omega )}^2ds\le C_1\qquad \mathrm{for\;all}\;t\ge t_1. \end{aligned}$$
(3.60)

Moreover, by the similar way in [11, Proposition 3.1] and (3.50), we can find \(C_2=C_2(r,\mu ,\alpha ,\Omega )>0\) and \(t_2=t_2(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega )>t_1\) such that

$$\begin{aligned} & \Vert v_\varepsilon (\cdot ,t)\Vert _{L^\infty (\Omega )}+\int \limits _\Omega w_\varepsilon ^{\frac{3\alpha }{5-2\alpha }}(\cdot ,t) +\int \limits _\Omega |\nabla v_\varepsilon (\cdot ,t)|^2 +\int \limits _t^{t+1}\int \limits _\Omega |\nabla v_\varepsilon |^4 \nonumber \\ & \quad \le C_2 \qquad \mathrm{for\;all}\;t\ge t_2>t_1, \end{aligned}$$
(3.61)

which enables us to find \(t_3=t_3(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega )\in (t_2,t_2+1)\) such that

$$\begin{aligned} \int \limits _\Omega |\nabla v_\varepsilon (\cdot ,t_3)|^4 \le C_2. \end{aligned}$$
(3.62)

\(\square \)

In addition, applying \(\nabla \) to the second equation in (2.9), multiplying the resulting identity by \(|\nabla v_\varepsilon |^{2p-2}\nabla v_\varepsilon \) for all \(p\in (\frac{3}{2},2)\) and integrating over \(\Omega \) by parts, we obtain that

$$\begin{aligned} & \frac{1}{2p}\frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p}+\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p}\nonumber \\ & \quad =-\frac{p-1}{2}\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-4}|\nabla |\nabla v_{\varepsilon }|^{2}|^{2}+\frac{1}{2}\int \limits _{\partial \Omega }|\nabla v_{\varepsilon }|^{2p-2}\frac{\partial |\nabla v_{\varepsilon }|^{2}}{\partial \nu }-\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-2}|D^{2}v_{\varepsilon }|^{2}\nonumber \\ & \quad +\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-2}\Delta v_{\varepsilon }(u_{\varepsilon }\cdot \nabla v_{\varepsilon })+(p-1)\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-4}(\nabla |\nabla v_{\varepsilon }|^{2}\cdot \nabla v_{\varepsilon })(u_{\varepsilon }\cdot \nabla v_{\varepsilon })\nonumber \\ & \quad +\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-2}\Delta v_{\varepsilon }g_{\varepsilon }(w_{\varepsilon })v_{\varepsilon }+(p-1)\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-4}(\nabla |\nabla v_{\varepsilon }|^{2}\cdot \nabla v_{\varepsilon })g_{\varepsilon }(w_{\varepsilon })v_{\varepsilon }\nonumber \\ & \quad +\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p}-\int \limits _{\partial \Omega }|\nabla v_{\varepsilon }|^{2p-2}g_{\varepsilon }(w_{\varepsilon })v_{*}\frac{\partial v_{\varepsilon }}{\partial \nu }. \end{aligned}$$
(3.63)

By means of Lemma 3.6, Lemma 3.7 and the similar way in [3, Lemma 3.5], we have

$$\begin{aligned} & \frac{1}{2}\int \limits _{\partial \Omega }|\nabla v_{\varepsilon }|^{2p-2}\frac{\partial |\nabla v_{\varepsilon }|^{2}}{\partial \nu }-\int \limits _{\partial \Omega }|\nabla v_{\varepsilon }|^{2p-2}g_{\varepsilon }(w_{\varepsilon })v_{*}\frac{\partial v_{\varepsilon }}{\partial \nu }\nonumber \\ & \quad \le \frac{p-1}{8}\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-4}|\nabla |\nabla v_{\varepsilon }|^{2}|^{2}+\frac{1}{4}\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-2}|D^{2}v_{\varepsilon }|^{2}+C_{3} \end{aligned}$$
(3.64)

with \(C_{3}=C_{3}(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\). Then, by Young’s inequality, Hölder inequality and the fact that \(|\Delta v_{\varepsilon }|\le \sqrt{3}|D^{2}v_{\varepsilon }|\), we obtain

$$\begin{aligned} & \int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-2}\Delta v_{\varepsilon }(u_{\varepsilon }\cdot \nabla v_{\varepsilon })+\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-2}\Delta v_{\varepsilon }g_{\varepsilon }(w_{\varepsilon })v_{\varepsilon }\nonumber \\ & \quad \le \frac{1}{4}\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-2}|D^{2}v_{\varepsilon }|^{2}+6\int \limits _{\Omega }w_{\varepsilon }^{2}|\nabla v_{\varepsilon }|^{2p-2}+6\int \limits _{\Omega }u_{\varepsilon }^{2}|\nabla v_{\varepsilon }|^{2p} \end{aligned}$$
(3.65)

and

$$\begin{aligned} \int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p}\le |\Omega |^{\frac{1}{p+1}}\cdot \Vert \nabla v_{\varepsilon }\Vert _{L^{2p+2}(\Omega )}^{2p} \end{aligned}$$
(3.66)

as well as

$$\begin{aligned} & (p-1)\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-4}(\nabla |\nabla v_{\varepsilon }|^{2}\cdot \nabla v_{\varepsilon })(u_{\varepsilon }\cdot \nabla v_{\varepsilon })+(p-1)\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-4}(\nabla |\nabla v_{\varepsilon }|^{2}\cdot \nabla v_{\varepsilon })g_{\varepsilon }(w_{\varepsilon })v_{\varepsilon }\nonumber \\ & \quad \le \frac{p-1}{8}\int \limits _{\Omega }|\nabla v_{\varepsilon }|^{2p-4}|\nabla |\nabla v_{\varepsilon }|^{2}|^{2}+C_{4}\int \limits _{\Omega }w_{\varepsilon }^{2}|\nabla v_{\varepsilon }|^{2p-2}+C_{5}\int \limits _{\Omega }u_{\varepsilon }^{2}|\nabla v_{\varepsilon }|^{2p} \end{aligned}$$
(3.67)

with \(C_{4}=C_{4}(p)>0\) and \(C_{5}=C_{5}(p,v_{0},v_{*})>0\). Substituting (3.64)–(3.67) into (3.63) implies

$$\begin{aligned} & \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |\nabla v_\varepsilon |^{2p}+2p\int \limits _\Omega |\nabla v_\varepsilon |^{2p} +\frac{p(p-1)}{2}\int \limits _\Omega |\nabla v_\varepsilon |^{2p-4}\left| \nabla |\nabla v_\varepsilon |^2\right| ^2 +\frac{3p}{2}\int \limits _\Omega |\nabla v_\varepsilon |^{2p-2}|D^2v_\varepsilon |^2 \nonumber \\ & \quad \le C_{6}\int \limits _\Omega w^2|\nabla v_\varepsilon |^{2p-2}+C_{7}\int \limits _\Omega |u_\varepsilon |^2|\nabla v_\varepsilon |^{2p}+C_{6}\Vert \nabla v_{\varepsilon }\Vert _{L^{2p+2}(\Omega )}^{2p}+C_{8} \qquad \mathrm{for\;all}\;t\ge t_3 \end{aligned}$$
(3.68)

with \(C_{6}=C_{6}(p)>0\), \(C_{7}=C_{7}(p,v_{0},v_{*})>0\) and \(C_{8}=C_{8}(r,\mu ,\alpha ,n_0,w_0,\Omega )>0\). In view of \(\alpha \in (\frac{5}{3},2)\), we have \(\frac{3\alpha }{5-2\alpha }>3\). We can apply (3.61), Hölder inequality and Gagliardo–Nirenberg inequality to obtain that there exist \(C_9=C_9(p,r,\mu ,\alpha ,\Omega )>0\) and \(C_{10}=C_{10}(p,r,\mu ,\alpha ,\Omega )>0\) satisfying

$$\begin{aligned} C_{6}\int \limits _\Omega w_\varepsilon ^2|\nabla v_\varepsilon |^{2p-2}\le & (4p^2+2p)\Vert w_\varepsilon \Vert _{L^3(\Omega )}^2\left\| |\nabla v_\varepsilon |^p\right\| _{L^\frac{6(p-1)}{p}(\Omega )}^\frac{2(p-1)}{p} \nonumber \\ \le & C_9\left\| \nabla |\nabla v_\varepsilon |^p\right\| _{L^2(\Omega )}^{\frac{2(p-1)}{p}\theta _2} \left\| |\nabla v_\varepsilon |^p\right\| _{L^\frac{2}{p}(\Omega )}^{\frac{2(p-1)}{p}(1-\theta _2)} +C_9\left\| |\nabla v_\varepsilon |^p\right\| _{L^\frac{2}{p}(\Omega )}^{\frac{2(p-1)}{p}} \nonumber \\ \le & C_{10}\left\| \nabla |\nabla v_\varepsilon |^p\right\| _{L^2(\Omega )}^{\frac{2(p-1)}{p}\theta _2}+C_{10}, \end{aligned}$$
(3.69)

where \(\theta _2:=\frac{3p^2-4p}{3p^2-4p+1}\in (0,1)\) thanks to \(p>\frac{3}{2}>\frac{4}{3}\). This means \(\frac{2(p-1)}{p}\theta _2=\frac{6p-8}{3p-1}\in (0,2)\), which allows us to invoke Young’s inequality in (3.69) to find \(C_{11}=C_{11}(p,r,\mu ,\alpha ,\Omega )>0\) such that

$$\begin{aligned} C_{6}\int \limits _\Omega w_\varepsilon ^2|\nabla v_\varepsilon |^{2p-2} \le \frac{p(p-1)}{4}\int \limits _\Omega |\nabla v_\varepsilon |^{2p-4}\left| \nabla |\nabla v_\varepsilon |^2\right| ^2+C_{11}. \end{aligned}$$
(3.70)

Moreover, by the Hölder inequality, we see that

$$\begin{aligned} C_{7}\int \limits _\Omega |u_\varepsilon |^2|\nabla v_\varepsilon |^{2p} \le C_{7}\Vert u_\varepsilon \Vert _{L^{2(p+1)}(\Omega )}^2\Vert \nabla v_\varepsilon \Vert _{L^{2(p+1)}(\Omega )}^{2p}, \end{aligned}$$
(3.71)

where an invocation of the Gagliardo–Nirenberg inequality warrants the existence of \(C_{12}=C_{12}(p,\Omega )>0\) satisfying

$$\begin{aligned} \Vert u_\varepsilon \Vert _{L^{2(p+1)}(\Omega )}^2&\le C_{12}\Vert \nabla u_\varepsilon \Vert _{L^{2p}(\Omega )}^\frac{3}{p(p+1)}\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^{2-\frac{3}{p(p+1)}} +C_{12}\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^2. \end{aligned}$$
(3.72)

and an application of Lemma 2.5 and (3.61) provides \(C_{13}=C_{13}(p,\Omega )>0\) and \(C_{14}=C_{14}(p,r,\mu ,\alpha ,\Omega )>0\) such that

$$\begin{aligned} \Vert \nabla v_\varepsilon \Vert _{L^{2(p+1)}(\Omega )}^{2p}\le & C_{13}\left\| |\nabla v_\varepsilon |^{p-1}D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{2p}{p+1}\Vert v_\varepsilon \Vert _{L^\infty (\Omega )}^\frac{2p}{p+1} +C_{13}\Vert v_\varepsilon \Vert _{L^\infty (\Omega )}^{2p} \nonumber \\ \le & C_{14}\left( \left\| |\nabla v_\varepsilon |^{p-1}D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{2p}{p+1}+1\right) . \end{aligned}$$
(3.73)

Moreover, we have

$$\begin{aligned} C_{6}\Vert \nabla v_{\varepsilon }\Vert _{L^{2p+2}(\Omega )}^{2p}\le \frac{p}{2}\int \limits _\Omega |\nabla v_\varepsilon |^{2p-2}|D^2v_\varepsilon |^2+C(q,r,\mu ,\alpha ,\Omega ). \end{aligned}$$
(3.74)

Inserting (3.72) and (3.73) into (3.71), and use Young’s inequality to show that

$$\begin{aligned} & C_{7}\int \limits _\Omega |u_\varepsilon |^2|\nabla v_\varepsilon |^{2p} \nonumber \\ & \quad \le C_{7}C_{12}C_{14}\left( \left\| |\nabla v_\varepsilon |^{p-1}D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{2p}{p+1}\Vert \nabla u_\varepsilon \Vert _{L^{2p}(\Omega )}^\frac{3}{p(p+1)}\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^{2-\frac{3}{p(p+1)}} +\Vert \nabla u_\varepsilon \Vert _{L^{2p}(\Omega )}^\frac{3}{p(p+1)}\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^{2-\frac{3}{p(p+1)}}\right. \nonumber \\ & \quad \left. +\left\| |\nabla v|^{p-1}D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{2p}{p+1}\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^2+\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^2\right) . \end{aligned}$$
(3.75)

Substituting (3.69), (3.74) and (3.75) into (3.68), we deduce that for all \(t\ge t_3\),

$$\begin{aligned} & \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |\nabla v_\varepsilon |^{2p}+2p\int \limits _\Omega |\nabla v_\varepsilon |^{2p} +p\int \limits _\Omega |\nabla v_\varepsilon |^{2p-2}|D^2v_\varepsilon |^2 \nonumber \\ & \quad \le C_{7}C_{12}C_{14}\left( \left\| |\nabla v_\varepsilon |^{p-1}D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{2p}{p+1} \Vert \nabla u_\varepsilon \Vert _{L^{2p}(\Omega )}^\frac{3}{p(p+1)}\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^{2-\frac{3}{p(p+1)}} +\Vert \nabla u_\varepsilon \Vert _{L^{2p}(\Omega )}^\frac{3}{p(p+1)}\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^{2-\frac{3}{p(p+1)}}\right. \nonumber \\ & \quad \left. +\left\| |\nabla v_\varepsilon |^{p-1}D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{2p}{p+1}+\Vert u_\varepsilon \Vert _{L^{2p}(\Omega )}^2\right) +C_{11}+C(q,r,\mu ,\alpha ,\Omega ). \end{aligned}$$
(3.76)

Noticing that \(\frac{2}{3-\alpha }\in (\frac{3}{2},2)\) due to \(\alpha \in (\frac{5}{3},2)\), and that (3.75) holds for any \(p>\frac{3}{2}\), we can particularly take \(p=\frac{2}{3-\alpha }\) in (3.75) and apply the first assertion of (3.60) to find \(C_{15}=C_{15}(r,\mu ,\alpha ,\kappa ,\Omega )>0\) such that

$$\begin{aligned} & \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |\nabla v_\varepsilon |^\frac{4}{3-\alpha }+\frac{4}{3-\alpha }\int \limits _\Omega |\nabla v_\varepsilon |^\frac{4}{3-\alpha } +\frac{2}{3-\alpha }\int \limits _\Omega |\nabla v_\varepsilon |^\frac{2(\alpha -1)}{3-\alpha }|D^2v_\varepsilon |^2 \nonumber \\ & \quad \le C_{15}\left( \left\| |\nabla v_\varepsilon |^\frac{\alpha -1}{3-\alpha }D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{4}{5-\alpha } \Vert \nabla u_\varepsilon \Vert _{L^\frac{4}{3-\alpha }(\Omega )}^\frac{3(3-\alpha )^2}{2(5-\alpha )} +\Vert \nabla u_\varepsilon \Vert _{L^\frac{4}{3-\alpha }(\Omega )}^\frac{3(3-\alpha )^2}{2(5-\alpha )} +\left\| |\nabla v_\varepsilon |^\frac{\alpha -1}{3-\alpha }D^2v_\varepsilon \right\| _{L^2(\Omega )}^\frac{4}{5-\alpha }\right) \nonumber \\ & \quad +C_{15}\qquad \mathrm{for\;all}\;t\ge t_3. \end{aligned}$$
(3.77)

Thanks to \(\alpha \in \left( \frac{5}{3},2\right) \), we easily see that

$$\begin{aligned} 3\alpha ^2-18\alpha +35-4(5-\alpha )=3\alpha ^2-14\alpha +15=(3\alpha -5)(\alpha -3)<0, \end{aligned}$$

and thus

$$\begin{aligned} 0<\frac{4}{5-\alpha }+\frac{3(3-\alpha )^2}{2(5-\alpha )}=\frac{3\alpha ^2-18\alpha +35}{2(5-\alpha )}<2. \end{aligned}$$

This enables us to make use of Young’s inequality in (3.77) so as to find \(C_{16}=C_{16}(r,\mu ,\alpha ,\kappa ,\Omega )>0\) with the property that

$$\begin{aligned} \frac{\mathrm{{d}}}{{\mathrm{{d}}}t}\int \limits _\Omega |\nabla v_\varepsilon |^\frac{4}{3-\alpha } +\frac{4}{3-\alpha }\int \limits _\Omega |\nabla v_\varepsilon |^\frac{4}{3-\alpha } \le \Vert \nabla u_\varepsilon \Vert _{L^\frac{4}{3-\alpha }(\Omega )}^2+C_{16}\qquad \mathrm{for\;all}\;t\ge t_3, \end{aligned}$$
(3.78)

and thereby employ Lemma 2.3, the second assertion of (3.60), (3.62), the fact that \(\frac{4}{3-\alpha }\in (3.4)\) due to \(\alpha \in (\frac{5}{3},2)\) and the Hölder inequality to achieve the desired result.

Based on above lemmas, we can obtain eventual \(L^{\infty }\) bound for \(n_{\varepsilon }\).

Lemma 3.9

Under the assumption of Lemma 3.4. Then, for all initial data complying with (1.8), there exists \(t_0=t_0(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega )>0\) such that

$$\begin{aligned} \Vert n_\varepsilon (\cdot ,t)\Vert _{L^\infty (\Omega )}\le C \qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1) \end{aligned}$$
(3.79)

with \(C=C(\chi ,r,\mu ,\alpha ,\kappa ,\Omega )>0\).

Proof

From Lemma 3.1, we know

$$\begin{aligned} \int \limits _\Omega n_\varepsilon (\cdot ,t)\le C_1:=2^\frac{1}{\alpha -1}|\Omega |\cdot \left( \frac{r_++1}{\mu }\right) ^\frac{1}{\alpha -1} \qquad \mathrm{for\;all}\;t\ge \frac{\ln 2}{(r_++1)(\alpha -1)}. \end{aligned}$$
(3.80)

Moreover, based on Lemmata 3.4 and 3.8, we can find \(C_2=C_2(r,\mu ,\alpha ,\kappa ,\Omega )>0\) and \(t_1=t_1(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega )\ge \frac{\ln 2}{(r_++1)(\alpha -1)}\) such that

$$\begin{aligned}&\int \limits _\Omega |\nabla v_\varepsilon |^\frac{4}{3-\alpha }+ \int \limits _\Omega |u_\varepsilon |^\frac{4}{3-\alpha } \le C_2\qquad \mathrm{for\;all}\;t\ge t_1. \end{aligned}$$
(3.81)

Now, we shall prove that (3.79) is true for \(t_0=t_0(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega ):=t_1+1\). For any fixed \(T\in (t_0,\infty )\), we define

$$\begin{aligned} M_\varepsilon (T):=\sup _{t\in [t_0,T)}\Vert n_\varepsilon (\cdot ,t)\Vert _{L^\infty (\Omega )}, \qquad \varepsilon \in (0,1). \end{aligned}$$
(3.82)

Of note, \(M_\varepsilon (T)\) is finite due to Lemma 2.1. By Duhamel’s principle, we express the solution component \(n_\varepsilon \) of (2.9) according to

$$\begin{aligned} & n_\varepsilon (\cdot ,t) \nonumber \\ & \quad =e^{\Delta }n_\varepsilon (\cdot ,t-1)-\int \limits _{t-1}^te^{(t-s)\Delta }\nabla \cdot \{h_\varepsilon (\cdot ,s)n_\varepsilon (\cdot ,s)\}ds +\int \limits _{t-1}^te^{(t-s)\Delta } \left[ rn_\varepsilon (\cdot ,s)-\mu n_\varepsilon ^\alpha (\cdot ,s)\right] ds \nonumber \\ & \quad =:n_{1\varepsilon }(\cdot ,t)+n_{2\varepsilon }(\cdot ,t)+n_{3\varepsilon }(\cdot ,t) \qquad \mathrm{for\;all}\;t\ge t_0\;\mathrm{and\;}\varepsilon \in (0,1), \end{aligned}$$
(3.83)

where \(h_\varepsilon (x,t):=\chi \rho _{\varepsilon }f_{\varepsilon }(n_{\varepsilon })\nabla v_\varepsilon (x,t)+u_\varepsilon (x,t)\). \(\square \)

Firstly, by means of the well-known smoothing estimate for \(\{e^{t\Delta }\}_{0\le t<2}\) (see [40]) and (3.81), we can find \(C_3=C_3(\Omega )>0\) such that

$$\begin{aligned} \Vert n_{1\varepsilon }(\cdot ,t)\Vert _{L^\infty (\Omega )}\le C_3\Vert n_\varepsilon (\cdot ,t-1)\Vert _{L^1(\Omega )}\le C_1C_3 \qquad \mathrm{for\;all}\;t\ge t_0. \end{aligned}$$
(3.84)

Secondly, thanks to Young’s inequality, we have \(rn_\varepsilon -\mu n_\varepsilon ^\alpha \le \frac{(\alpha -1)(r_+)^\frac{\alpha }{\alpha -1}}{\mu ^\frac{1}{\alpha -1}\alpha ^\frac{\alpha }{\alpha -1}}\). This allows us to utilize the maximum principle to arrive at

$$\begin{aligned} n_{3\varepsilon }(x,t)\le \int \limits _{t-1}^t\frac{(\alpha -1)(r_+)^\frac{\alpha }{\alpha -1}}{\mu ^\frac{1}{\alpha -1}\alpha ^\frac{\alpha }{\alpha -1}} \le \frac{(\alpha -1)(r_+)^\frac{\alpha }{\alpha -1}}{\mu ^\frac{1}{\alpha -1}\alpha ^\frac{\alpha }{\alpha -1}} \qquad \mathrm{for\;all}\;(x,t)\in \Omega \times [t_0,\infty ). \end{aligned}$$
(3.85)

Finally, a combination of (3.81) and the fact that \(|sf_{\varepsilon }(s)|\le 1\) for \(x\in \Omega \) and \(s\ge 0\) entails

$$\begin{aligned} \Vert h_\varepsilon (\cdot ,t)\Vert _{L^{\frac{4}{3-\alpha }}(\Omega )}\le C_4 \qquad \mathrm{for\;all}\;t\ge t_1 \end{aligned}$$
(3.86)

with \(C_4=C_4(\chi ,r,\mu ,\alpha ,\kappa ,\Omega )>0\). Since \(\alpha \in (\frac{5}{3},2)\), we see that \(\frac{4}{3-\alpha }\in (3,4)\). This guarantees us to take \(q>3\) such that \(q\in (3,\frac{4}{3-\alpha })\), and take advantage of the smoothing \(L^p\)-\(L^q\) estimates of \(\{e^{t\Delta }\}_{t\ge 0}\) (see [40, Lemma 1.3) and the Hölder inequality as well as the interpolation inequality for \(L^p\)-norms to deduce that

$$\begin{aligned} \Vert n_{2\varepsilon }(\cdot ,t)\Vert _{L^\infty (\Omega )}\le & \int \limits _{t-1}^t\Vert e^{(t-s)\Delta }\nabla \cdot \{h_\varepsilon (\cdot ,s)n_\varepsilon (\cdot ,s)\}\Vert _{L^\infty (\Omega )}ds \nonumber \\ \le & C_5\int \limits _{t-1}^t(t-s)^{-\frac{1}{2}-\frac{3}{2q}}\Vert h_\varepsilon (\cdot ,s) n_\varepsilon (\cdot ,s)\Vert _{L^{q}(\Omega )}ds \nonumber \\ \le & C_5\int \limits _{t-1}^t(t-s)^{-\frac{1}{2}-\frac{3}{2q}}\Vert h_\varepsilon (\cdot ,s)\Vert _{L^{\frac{4}{3-\alpha }}(\Omega )} \Vert n_\varepsilon (\cdot ,s)\Vert _{L^{\frac{4q}{4-(3-\alpha )q}}(\Omega )}ds \nonumber \\ \le & C_4C_5\int \limits _{t-1}^t(t-s)^{-\frac{1}{2}-\frac{3}{2q}} \Vert n_\varepsilon (\cdot ,s)\Vert _{L^\infty (\Omega )}^{\frac{(7-\alpha )q-4}{4q}} \Vert n_\varepsilon (\cdot ,s)\Vert _{L^1(\Omega )}^{\frac{4-(3-\alpha )q}{4q}}ds \nonumber \\ \le & C_1^\frac{4-(3-\alpha )q}{4q}C_4C_5 M_\varepsilon ^{\frac{(7-\alpha )q-4}{4q}}(T)\int \limits _{t-1}^t(t-s)^{-\frac{1}{2}-\frac{3}{2q}}ds \nonumber \\ = & C_6M_\varepsilon ^{\frac{(7-\alpha )q-4}{4q}}(T) \qquad \mathrm{for\;all}\;t\in [t_0,T), \end{aligned}$$
(3.87)

where \(C_5=C_5(q,\Omega )>0\), and \(C_6:=C_1^\frac{4-(3-\alpha )q}{4q}C_4C_5\int \limits _{t-1}^t(t-s)^{-\frac{1}{2}-\frac{3}{2q}}ds =\frac{2qC_1^\frac{4-(3-\alpha )q}{4q}C_4C_5}{q-3}\) due to \(q>3\). Inserting (3.84), (3.85) and (3.87) into (3.83) and recalling (3.82) immediately yield

$$\begin{aligned} M_\varepsilon (T)\le C_6M_\varepsilon ^{\frac{(7-\alpha )q-4}{4q}}(T)+C_7 \qquad \mathrm{for\;all}\;T\in (t_0,\infty ){\;\mathrm and\;}\varepsilon \in (0,1) \end{aligned}$$
(3.88)

with \(C_7:=C_1C_3+ \frac{(\alpha -1)(r_+)^\frac{\alpha }{\alpha -1}}{\mu ^\frac{1}{\alpha -1}\alpha ^\frac{\alpha }{\alpha -1}}\). Therefore, by direct calculation, we have

$$\begin{aligned} M_\varepsilon (T)\le \max \left\{ \left( \frac{C_7}{C_6}\right) ^\frac{4q}{(7-\alpha )q-4},(2C_6)^\frac{4q}{4-(3-\alpha )q}\right\} \qquad \mathrm{for\;all}\;T\in (t_0,\infty ){\;\mathrm and\;}\varepsilon \in (0,1), \end{aligned}$$
(3.89)

which together with (3.82) readily establishes (3.79) by taking \(T\rightarrow \infty \).

With the help of Lemma 3.9, the eventual Hölder regularity of fluid velocity field, the eventual \(W^{1,\infty }(\Omega )\) boundedness of \(v_{\varepsilon }\) and \(w_{\varepsilon }\) and the higher order time-ultimate estimates of solutions can be obtain. The proof is similar to [4, Lemma 6.1, 6.2 and Proposition 6.3], we omit its details to avoid repetition.

Lemma 3.10

Let the assumptions of Lemma 3.4 be true. Then, for all initial data satisfying (1.11), one can find \(t_0=t_0(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega )>0\) such that

$$\begin{aligned} \Vert u_\varepsilon \Vert _{C^{\gamma ,\frac{\gamma }{2}}({\bar{\Omega }}\times [t,t+1])}\le C\qquad for\;all\;t\ge t_0\;and\;\varepsilon \in (0,1) \end{aligned}$$
(3.90)

Furthermore, we have

$$\begin{aligned} \Vert v_\varepsilon (\cdot ,t)\Vert _{W^{1,\infty }(\Omega )}\le C \qquad for\;all\;t\ge t_0\;\mathrm{and\;}\varepsilon \in (0,1) \end{aligned}$$
(3.91)

and

$$\begin{aligned} \Vert w_\varepsilon (\cdot ,t)\Vert _{W^{1,\infty }(\Omega )}\le C \qquad for\;all\;t\ge t_0\;\mathrm{and\;}\varepsilon \in (0,1) \end{aligned}$$
(3.92)

with some \(\gamma \in (0,1)\) and \(C=C(\chi ,r,\mu ,\alpha ,\kappa ,\Omega )>0\).

Lemma 3.11

Under the hypotheses of Lemma 3.4. Then there exist \(\gamma \in (0,1)\), \(t_0=t_0(r,\mu ,\alpha ,n_0, v_0,w_0,u_0,\Omega )>0\) and \(C=C(T,\chi ,r,\mu ,\alpha ,\kappa ,\Omega )>0\) such that

$$\begin{aligned} & \Vert n_\varepsilon \Vert _{C^{2+\gamma ,1+\frac{\gamma }{2}}({\bar{\Omega }}\times [t_0,T])} +\Vert v_\varepsilon \Vert _{C^{2+\gamma ,1+\frac{\gamma }{2}}({\bar{\Omega }}\times [t_0,T])}\nonumber \\ & \quad +\Vert w_\varepsilon \Vert _{C^{2+\gamma ,1+\frac{\gamma }{2}}({\bar{\Omega }}\times [t_0,T])} +\Vert u_\varepsilon \Vert _{C^{2+\gamma ,1+\frac{\gamma }{2}}({\bar{\Omega }}\times [t_0,T])} \le C \end{aligned}$$
(3.93)

for each \(T>t_0\) and \(\varepsilon \in (0,1)\).

Now, we ready to prove Theorem 1.1.

The proof of Theorem 1.1 Firstly, we have known that the global existence of weak solution in the sense of Definition 2.1 for the problem (1.8). From Lemma 3.11, we can make use of the \(Arzel\grave{a}\)-Ascoli theorem to find \(t_0=t_0(r,\mu ,\alpha ,n_0,v_0,w_0,u_0,\Omega )>0\) with the property that for any given \(\{\varepsilon _j\}_{j\in {\mathbb {N}}}\subset (0,1)\) fulfilling \(\varepsilon _j\searrow 0\) as \(j\rightarrow \infty \) one can find \((n,v,w,u,P)\in \left[ C^{2,1}({\bar{\Omega }}\times [t_0,\infty ))\right] ^3\times C^{2,1}({\bar{\Omega }}\times [t_0,\infty );{\mathbb {R}}^3)\times C^{1,0}({\bar{\Omega }}\times [t_0,\infty ))\) and a subsequence \(\{\varepsilon _{j_k}\}_{k\in {\mathbb {N}}}\subset (0,1)\) such that

$$\begin{aligned} {\left\{ \begin{array}{ll} n_{\varepsilon _{j_k}}\rightarrow n ~\mathrm{in\;}C^{2,1}({\bar{\Omega }}\times [t_0,\infty )), \\ v_{\varepsilon _{j_k}}\rightarrow v ~\mathrm{in\;}C^{2,1}({\bar{\Omega }}\times [t_0,\infty )), \\ w_{\varepsilon _{j_k}}\rightarrow w ~\mathrm{in\;}C^{2,1}({\bar{\Omega }}\times [t_0,\infty )), \\ u_{\varepsilon _{j_k}}\rightarrow u ~\mathrm{in\;}C^{2,1}({\bar{\Omega }}\times [t_0,\infty );{\mathbb {R}}^3) \end{array}\right. } \end{aligned}$$
(3.94)

as \(\varepsilon _{j_k}\searrow 0\). Taking \(\varepsilon =\varepsilon _{j_k}\searrow 0\) in (2.9) and constructing the corresponding pressure P as performing in [4], we can conclude that the limit objects (nvwuP) indeed forms the classical solution of (1.8). This together with Proposition 1.1 means that the weak solution (nvwu) actually becomes eventually smooth in the sense that (1.24). Obviously, the boundedness property shown in (1.25) is as a straightforward consequence of Lemmas 3.93.10 and the approximation procedure.