1 Introduction

The reaction-diffusion systems have been advocated to interpret numerous biological phenomena such as wave propagations [13, 41], pattern formation [16, 25], ecological invasions [11, 19], competition of species [18, 23], wound healing [39], and so on (cf. [5, 33]). However, in many biological processes involving directed motions, such as chemotaxis and predators seeking prey (prey-taxis), reaction-diffusion models may not be adequate to describe how organisms move and disperse. For instance, Rowell in [37] explains how models with random diffusion fail to explain certain ecological phenomena and do not accurately reflect the non-Brownian motion of individuals. The Lotka–Volterra type predator–prey system with random diffusion only is unable to produce spatial patterns (cf. [47]) to interpret the spatiotemporal heterogeneity observed in the field experiment [24, 45]. Rational movement along resource gradients has been thought to reduce the diffusive effect and result in clustering and formation of colonies to increase the chances of survival of individuals like human populations forming small nucleated settlements which grow as the population saturates locally. Incorporating both random and rational movements, Grindrod [15] proposed models of individual clustering in single-species and multi-species communities by taking the advective flux to depend directly upon local population densities without requiring intermediate signals like attractants or repellents.

Let us first briefly review the origin of Grindrod clustering models [15]. Classically, models for the spatial dispersion of biological populations have the form

$$\begin{aligned} u_t=\Delta \phi (u)+f(u,x,t), \end{aligned}$$

where u(xt) denotes the population density at location x and time t, and f(uxt) represents population kinetics due to the birth and death; \(\phi \) satisfies \(\phi (0)=0\) and \(\phi '(u)>0\) for \(u>0\). As highlighted in [15], the above model contains no aggregation mechanism such as swarming, herding, and clustering of individuals, which can serve as a balancing factor between death and birth rates and increase survival chances. To incorporate this phenomenon, a modified population balance equation reads

$$\begin{aligned} \partial _t u=-\nabla \cdot (u \varvec{V}(u, t, x))+u E(u, t, x), \end{aligned}$$

where \(\varvec{V}\) and E are the average velocity of individuals and the net rate of reproduction per individual, respectively. E typically has the form

$$\begin{aligned} E(u)= \left\{ \begin{array}{ll} 1-u, &{}\quad \text {monostable case},\\ (1-u)(u-a) \text { for some } a \in (0,1),&{}\quad \text {bistable case}. \end{array}\right. \end{aligned}$$

Considering random diffusion with a probability \(\delta \in (0,1)\), and deterministic dispersion with the probability \(1-\delta \) in an average velocity \(\textbf{w}\) to increase the expected net rate of reproduction, \(\varvec{V}\) responding to u and E is given by \(\varvec{V}=-\delta \frac{\nabla u}{u}+(1-\delta )\textbf{w}\). The former obeys Fickian diffusion \(\frac{\nabla u}{u}\), while the latter is supposed to increase the net rate of reproduction per individual, such as \(\textbf{w}\approx \lambda \nabla E\) with \(\lambda >0\). This leads to the following model [15]

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t u=\delta \Delta u-(1-\delta ) \nabla \cdot (u \textbf{w})+u E(u, t, x),\\ -\varepsilon \Delta \textbf{w}+\textbf{w}=\lambda \nabla E(u), \end{array}\right. } \end{aligned}$$

where \(\varepsilon >0\) is a small constant accounting for the small fluctuation to smooth out any sharp local variations in \(\nabla E\). After some rescalings, and assuming that the environment is homogeneous, the single-species model proposed in [15] reads

$$\begin{aligned} \left\{ \begin{array}{ll} u_t=d \Delta u-\chi \nabla \cdot \left( u\textbf{w}\right) +\gamma u E(u), &{} x\in \Omega ,\ t>0,\\ -\varepsilon \Delta \textbf{w}+\textbf{w}=\nabla E(u), &{} x\in \Omega ,\ t>0,\\ u(x, 0)=u_{ 0}(x), &{} x\in \Omega , \end{array} \right. \end{aligned}$$
(1.1)

where \(\Omega \subset \mathbb {R}^{n}\) \((n \ge 2)\) is a bounded domain with a smooth boundary, Variables u(xt) and E(u), and the parameter \(\varepsilon \) have the same meaning as above, \(\textbf{w}(x,t)\) denotes the average velocity increasing the expected rate of reproduction of individuals up to a rescaling. The parameters \(d, \chi , \gamma \) are all positive.

In the case of multi-species communities, the interspecific interactions (like competition or cooperation) between different species are indispensable. In particular the m-species Grindrod clustering model with competitive interactions reads as (cf. [15])

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u_{i}=d_{i} \Delta u_{i}-\chi _{i} \nabla \cdot \left( u_{i} \textbf{w}_{i}\right) +u_{i} E_{i}, &{} x\in \Omega ,\ t>0,\\ -\varepsilon _{i} \Delta \textbf{w}_{i}+\textbf{w}_{i}-\nabla E_{i}=0, &{} x\in \Omega ,\ t>0,\\ u_{i}(x, 0)=u_{i 0}(x), &{} x\in \Omega , \end{array}\right. \end{aligned}$$
(1.2)

with

$$\begin{aligned} E_{i}:=E_{i}\left( u_{1}, u_{2}, \cdots , u_{m}\right) =a_{i}-\sum _{j=1}^{m} b_{i j} u_{j}, \quad i=1,2, \cdots , m, \end{aligned}$$

where all parameters \(d_{i}, \chi _{i}, \varepsilon _{i}, a_{i}, b_{i j}\) are positive. The original no-flux boundary condition (i.e. no individuals can cross the boundary) proposed in [15, formula (2.4)] for the two-species Grindrod clustering model (i.e., \(m=2\)) is

$$\begin{aligned} \nabla u_{i} \cdot \textbf{n}=\textbf{w}_{i} \cdot \textbf{n}=0, \text{ on } \partial \Omega , \quad i=1,2, \end{aligned}$$

where \(\textbf{n}\) denotes the unit outer normal vector to the boundary \(\partial \Omega \). However, the above boundary condition \(\textbf{w}_i \cdot \textbf{n}\mid _{\partial \Omega }=0\) for \(\textbf{w}_i\) is inadequate to warrant the global well-posedness of the model (1.2) in multi-dimensions (\(n\ge 2\)). This limitation was identified by Nasreddine [34,35,36] for the single-species Grindroid clustering model, where the additional boundary condition \(\frac{\partial \textbf{w}}{\partial \textbf{n}} \times \textbf{n}\mid _{\partial \Omega }=0\) is suggested for the velocity \(\textbf{w}\). Such a kind of boundary condition is not peculiar, see e.g. [12, 38] for other models incorporating this kind of boundary condition. Accordingly, for the m-species Grindroid clustering model (1.2), we incorporate the boundary condition \(\frac{\partial \textbf{w}_i}{\partial \textbf{n}} \times \textbf{n}\mid _{\partial \Omega }=0\) for \(\textbf{w}_i\) (\(i=1,2,\cdots ,m\)). Therefore, the boundary conditions of (1.2) to be considered are

$$\begin{aligned} \nabla u_{i} \cdot \textbf{n}=\textbf{w}_{i} \cdot \textbf{n}=0,\ \partial _{\textbf{n}}\textbf{w}_i\times \textbf{n}=\textbf{0}, \text{ on } \partial \Omega ,\quad i=1,2, \cdots , m. \end{aligned}$$
(1.3)

As usual, for vectors \(\textbf{a}=(a_1,a_2,\cdots ,a_n)\) and \(\textbf{b}=(b_1,b_2,\cdots ,b_n)\), the cross product \(\textbf{a}\times \textbf{b}\) is the number \(a_1b_2-a_2b_1\) if \(n=2\) and the vector \((a_{2} b_{3}-a_{3} b_{2},a_{1} b_{3}-a_{3} b_{1},a_{1} b_{2}-a_{2} b_{1})\) if \(n=3\). In one dimension (\(n=1\)), the condition \(\partial _{\textbf{n}}\textbf{w}_i\times \textbf{n}=\textbf{0}\) is not needed. The Dirchlet boundary condition \(\textbf{w}|_{\partial \Omega }=0\) satisfies the condition for \(\textbf{w}\) in (1.3). Though the Grindrod models were proposed three decades ago, there is no mathematical result available, except some preliminary results obtained for the single-species Grindrod model (1.1) with boundary conditions in (1.3). Nasreddine [34] proved the local-in-time existence of strong solutions of (1.1) with (1.3) in multi-dimensions for \((u, \textbf{w}) \in W^{1, p}(\Omega )\) with \(p>n\) and global existence of strong solutions to (1.1) in one dimension with (1.3) for both monostable and bistable functions E(u) as well as \(L^{2}\) convergence of solutions to constant steady states in the monostable case. The global existence of strong solutions of (1.1) with (1.3) in two dimensions was later established in [36], where the solution bound in \(W^{1, p}(\Omega )\) depends on time and the possibility of blow-up at infinite time was not precluded. The diffusion vanishing problem of (1.1) with (1.3) as \(\varepsilon \rightarrow 0\) in one dimension was investigated in [35], and the existence of traveling wave solutions of (1.1) in \(\mathbb {R}\) was established in [21] for \(E(u)=1-u\). The planar and radial traveling wave solutions of single-species and two-species Grindrod models with monostable function E(u) in \(\mathbb {R}^{n}\) was investigated in [26] alongside numerical simulations and found that directed motion can have substantial impacts not only on wave speed but also on the existence and structure of emergent patterns. The pattern formation of the two-species Grindrod model (i.e. (1.2) with \(m=2)\) with (1.3) in a two-dimensional convex domain was studied in [27], by assuming \(\textbf{w}_{i}=\nabla \phi _{i}\) for some potential functions \(\phi _{i}\) \((i=1,2)\), for three interspecific interactions: competition, generalist predator–prey and predator–prey. In particular, how the advective dispersal of species in heterogeneous resources and hazards leads to asymptotic steady states that retain spatial aggregation or clustering in regions of resource abundance and away from hazards was examined. By investigating pattern formation of the multi-species Grindrod model (1.2) approximated by a non-local cross-diffusion model, the authors of [40] proved that the Turing patterns, which were impossible for the two-species models, may arise for m-species Grindrod models with \(m \ge 3\).

From the above literature review, we see that the qualitative understanding of the Grindrod clustering models remains poorly understood, especially whether the solution blows up in infinite time in multi-dimensions is inconclusive and the large-time behavior of solutions is also unclear. This paper is devoted to exploring these basic questions. Without loss of generality, we consider the two-species Grindrod clustering model with competitions

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u_{1}=d_{1} \Delta u_{1}-\chi _1 \nabla \cdot \left( u_{1} \textbf{w}_{1}\right) + u_{1}E_1(u_1,u_2), &{} x\in \Omega ,\ t>0,\\ \partial _{t} u_{2}=d_{2} \Delta u_{2}-\chi _2 \nabla \cdot \left( u_2 \textbf{w}_{2}\right) +u_{2}E_2(u_1,u_2), &{} x\in \Omega ,\ t>0,\\ -\varepsilon _{1} \Delta \textbf{w}_{1}+\textbf{w}_{1}=\nabla E_1(u_1,u_2), &{} x\in \Omega ,\ t>0,\\ -\varepsilon _{2} \Delta \textbf{w}_{2}+\textbf{w}_{2}=\nabla E_2(u_1,u_2), &{} x\in \Omega ,\ t>0,\\ \nabla u_{i} \cdot \textbf{n}=\textbf{w}_{i} \cdot \textbf{n}=0, \ \partial _{\textbf{n}}\textbf{w}_i\times \textbf{n}=0,\quad i=1,2,\qquad \qquad &{} x\in \partial \Omega ,\ t>0,\\ (u_1,u_2)(x,0)=(u_{10},u_{20})(x), &{} x\in \Omega , \end{array} \right. \end{aligned}$$
(1.4)

with

$$\begin{aligned} E_1(u_1,u_2):=\gamma _1-u_{1}-c u_{2}\quad \text {and}\quad E_2(u_1,u_2):=\gamma _2-b u_{1}-u_{2}, \end{aligned}$$
(1.5)

where \(\Omega \subset \mathbb {R}^{2}\) is a bounded domain with a smooth boundary and \(\textbf{n}\) is the unit outward normal vector of \(\partial \Omega \), and all parameters \(d_{i}, \chi _{i}, \varepsilon _{i}, \gamma _{i}, b, c\), \(i\in \left\{ 1,2\right\} \), are positive. We underline that the boundary conditions in (1.4) (see also (1.3)) means that \(\partial _{\textbf{n}} \textbf{w}_i\) is parallel to \(\textbf{n}\) on \(\partial \Omega \), which along with the boundary condition \(\textbf{w}_{i} \cdot \textbf{n}\mid _{\partial \Omega }=0\) implies

$$\begin{aligned} \textbf{w}_i\cdot \partial _{\textbf{n}}\textbf{w}_j\mid _{\partial \Omega }=0,\quad i,j\in \left\{ 1,2\right\} . \end{aligned}$$
(1.6)

We remark that without advection (i.e. \(\chi _1=\chi _2=0\)), the first two equations of (1.4)–(1.5) have no components \(\textbf{w}_{i} (i=1,2)\) and become the well-known competition-diffusion Lotka–Volterra model which has been well studied (cf. [4, 30]). The competition models with advection have also been widely studied in literatures (cf. [1, 2, 6,7,8,9, 14, 50]). All these works have assumed that the advection is biased to the concentration gradient of given resources. The competition dynamics in advective environments like the river or stream were also studied (cf. [29, 31]). We refer to [48, 49] for the study of global dynamics of diffusion-advection competition models with more general diffusive and/or advective coefficients. When the advection is modeled by the prey-taxis in a predator–prey system, the competition dynamics were investigated in [44]. Evidently, the advection considered in the Grindrod model (1.4)–(1.5) are different from those considered in the existing studies mentioned above.

In this paper, we shall establish the global boundedness and time-asymptotic dynamics of solutions to (1.4). Our first result concerning the global existence and boundedness of classical solutions is stated in the following theorem.

Theorem 1.1

Let \(\Omega \subset \mathbb {R}^{2}\) be a bounded domain with a smooth boundary. Assume \(u_{10}, u_{20} \in W^{1, p}(\Omega )\) with \(p>2\) and \(u_{10}, u_{20}\gvertneqq 0\). Then the system (1.4) admits a unique classical solution \((u_1, u_2, \textbf{w}_1, \textbf{w}_2)\) satisfying

$$\begin{aligned} \left\{ \begin{array}{lcl} u_i \in C^0(\bar{\Omega } \times [0, \infty )) \cap C^{2,1}(\bar{\Omega } \times (0, \infty )),\\ \textbf{w}_i \in \left[ C^{2,1}(\bar{\Omega } \times (0, \infty ))\right] ^2, \end{array}\right. \quad i=1,2, \end{aligned}$$

and \(u_1,u_2>0\) in \({\Omega } \times (0, \infty )\). Moreover, there exists a positive constant C independent of t such that

$$\begin{aligned} \sum _{i=1}^2\left( \Vert u_i(\cdot , t)\Vert _{{W^{1,p}(\Omega )}}+\Vert \textbf{w}_i(\cdot , t)\Vert _{{W^{1,p}(\Omega )}}\right) \le C\quad \text {for all }t>0. \end{aligned}$$
(1.7)

Next, we explore asymptotic dynamics of (1.4). Except the extinction steady state \((0,0,\textbf{0},\textbf{0})\), the system (1.4) has three possible homogeneous steady states \(\left( u_{1s}, u_{2s},\textbf{0},\textbf{0}\right) \) depending on the value of parameters \(b,c,\gamma _1,\gamma _2\). They can be classified into the following three categories similar to the classical Lotka–Volterra competition system (cf. [30]):

  • Case 1. Weak competition: \(c<\frac{\gamma _1}{\gamma _2}<\frac{1}{b}\).

  • Case 2. Competitive exclusion: \(\frac{\gamma _1}{\gamma _2}<\min \{\frac{1}{b},c\}\) (resp. \(\frac{\gamma _1}{\gamma _2}>\max \{\frac{1}{b},c\}\)).

  • Case 3. Strong competition: \(\frac{1}{b}<\frac{\gamma _1}{\gamma _2}<c\).

Then the corresponding homogeneous steady state \(\left( u_{1\,s}, u_{2\,s},\textbf{0},\textbf{0}\right) \) can be solved as follows:

$$\begin{aligned} \left( u_{1s}, u_{2s}\right) =\left\{ \begin{array}{ll} (0,\gamma _2)\text { or }(\gamma _1,0)\text { or }(u_{1}^*,u_{2}^*), &{} \text{ in } {\textbf {Case 1}},\\ (0,\gamma _2)\text { or }(\gamma _1,0), &{} \text{ in } {\textbf {Case 2}}, \end{array}\right. \end{aligned}$$

where

$$\begin{aligned} (u_{1}^*,u_{2}^*):=\left( \frac{\gamma _1-\gamma _2 c}{1-bc},\frac{\gamma _2-\gamma _1 b}{1-bc}\right) . \end{aligned}$$
(1.8)

To state our main results on the large time behavior of solutions, we introduce some notations. Denote the function

$$\begin{aligned} f(x,y):= \frac{1+x^2 y^2}{1-xy}>1\quad \text {for }x,y>0\text { and }xy<1, \end{aligned}$$
(1.9)

and define two positive constants

$$\begin{aligned} K_1:=\frac{16d_1\varepsilon _1}{ \chi _1^2}\quad \text {and}\quad \quad K_2:=\frac{16d_2\varepsilon _2}{\chi _2^2}. \end{aligned}$$
(1.10)

Moreover, when \(c<\frac{\gamma _1}{\gamma _2}<\frac{1}{b}\) (weak competition), define two positive constants

$$\begin{aligned} K_1^*:=\frac{2 {u_1^*}\left( \frac{K_2}{1-b c}-{u_2^*}(1+b c)\right) }{K_2-f(b,c){u_2^*}}\quad \text {if }K_2>f(b,c){u_2^*} \end{aligned}$$
(1.11)

and

$$\begin{aligned} \quad K_2^*:=\frac{2 {u_2^*}\left( \frac{K_1}{1-b c}-{u_1^*}(1+b c)\right) }{K_1-f(b,c){u_1^*}}\quad \text {if }K_1>f(b,c){u_1^*}. \end{aligned}$$
(1.12)

Our second result is stated in the following.

Theorem 1.2

Suppose that the conditions in Theorem 1.1 hold. Then the solution \((u_1, u_2, \textbf{w}_1, \textbf{w}_2)\) of the system (1.4) obtained in Theorem 1.1 has the following convergence properties.

  1. (i)

    Assume \(c<\frac{\gamma _1}{\gamma _2}<\frac{1}{b}\) and \(({u_1^*},{u_2^*})\) is given by (1.8). If \((K_1,K_2)\) defined in (1.10) satisfies

    $$\begin{aligned} K_1> f(b,c){u_1^*}\quad \text {and}\quad K_2>K_2^*, \end{aligned}$$
    (1.13)

    or

    $$\begin{aligned} K_2> f(b,c){u_2^*}\quad \text {and}\quad K_1>K_1^*, \end{aligned}$$
    (1.14)

    then there exist positive constants C and \(\lambda \) independent of t such that

    $$\begin{aligned} \sum _{i=1}^2\left( \left\| u_i(\cdot , t)-u_{i}^{*}\right\| _{{W^{1,\infty }(\Omega )}} +\left\| \textbf{w}_i\right\| _{{W^{1,\infty }(\Omega )}}\right) \le C e^{-\lambda t}\quad \text {as }t\rightarrow \infty . \end{aligned}$$
  2. (ii)

    Assume \(\frac{\gamma _1}{\gamma _2}<\min \{\frac{1}{b},c\}\). If

    $$\begin{aligned} K_2> f\left( b,\frac{\gamma _1}{\gamma _2}\right) \gamma _2=\frac{\gamma _2^2+b^2\gamma _1^2}{\gamma _2-b\gamma _1}, \end{aligned}$$
    (1.15)

    then there exists a positive constant C independent of t such that

    $$\begin{aligned}{} & {} \left\| u_1(\cdot , t)\right\| _{{W^{1,\infty }(\Omega )}}+\left\| u_2(\cdot , t)-\gamma _2\right\| _{{W^{1,\infty }(\Omega )}}\\{} & {} \qquad \quad +\sum _{i=1}^2\left\| \textbf{w}_i\right\| _{{W^{1,\infty }(\Omega )}}\le \frac{C}{1+t}\quad \text {as }t\rightarrow \infty . \end{aligned}$$
  3. (iii)

    Assume \(\frac{\gamma _1}{\gamma _2}>\max \{\frac{1}{b},c\}\). If

    $$\begin{aligned} K_1> f\left( \frac{\gamma _2}{\gamma _1},c\right) \gamma _1=\frac{\gamma _1^2+c^2\gamma _2^2}{\gamma _1-c\gamma _2}, \end{aligned}$$

    then there exists a positive constant C independent of t such that

    $$\begin{aligned}{} & {} \left\| u_1(\cdot , t)-\gamma _1\right\| _{{W^{1,\infty }(\Omega )}}+\left\| u_2(\cdot , t)\right\| _{{W^{1,\infty }(\Omega )}}\\{} & {} +\sum _{i=1}^2\left\| \textbf{w}_i\right\| _{{W^{1,\infty }(\Omega )}}\le \frac{C}{1+t}\quad \text {as }t\rightarrow \infty . \end{aligned}$$

Remark 1.1

It is unclear whether the parameter regime of \((K_1,K_2)\) satisfying (1.13) or (1.14) in Theorem 1.2(i) is admissible. Below we shall confirm this and further show what the admissible regime looks like. First, one can check that \((K_1^*,K_2^*)\) defined by (1.11) and (1.12) satisfies

$$\begin{aligned} K_1^*>f(b,c){u_1^*}>{u_1^*}\quad \text {and}\quad K_2^*>f(b,c){u_2^*}>{u_2^*}. \end{aligned}$$

Viewing \(K_1^*\) and \(K_2^*\) as functions of \(K_2\) and \(K_1\) according to (1.11) and (1.12), respectively, we get

$$\begin{aligned} K_1^*= \frac{2 {u_1^*}}{1-b c}+\frac{I_K}{K_2-f(b,c){u_2^*}}\quad \text {and}\quad K_2^*= \frac{2 {u_2^*}}{1-b c}+\frac{I_K}{K_1-f(b,c){u_1^*}}, \end{aligned}$$

where

$$\begin{aligned} I_K=\frac{2{u_1^*}{u_2^*}( f(b,c)+b^2c^2-1)}{(1-bc)}>0 \end{aligned}$$

due to \(f(b,c)>1\). Therefore, \(K_1^*\) (resp. \(K_2^*\)) decreases monotonically with respect to \(K_2\in (f(b,c){u_2^*},+\infty )\) (resp. \(K_1\in (f(b,c){u_1^*},+\infty )\)). Let

$$\begin{aligned} \Sigma :=\left\{ (K_1,K_2)\mid (K_1,K_2)\text { satisfies }(1.13)\text { or }(1.14)\right\} , \end{aligned}$$
(1.16)

then the region of \(\Sigma \) is showed in Fig. 1.

Fig. 1
figure 1

Illustration of \(\Sigma \) defined by (1.16)

The rest of this paper is organized as follows. In Sect. 2, we shall address the local existence of solutions to (1.4), and then we will use an extension criterion to prove that the local solution is actually uniformly bounded and exists globally in time in Sect. 3. In Sect. 4, we shall prove the global stabilities stated in Theorem 1.2 by constructing Lyapunov functionals along with compactness arguments.

2 Preliminaries: local existence and some inequalities

Before proceeding, we introduce some notations used throughout the paper.

Notations:

  • For brevity, we abbreviate \(\int _{0}^{t} \int _{\Omega } f(\cdot , s) d x d s\) and \(\int _{\Omega } f(\cdot , t) d x\) as \(\int _{0}^{t} \int _{\Omega } f\) and \(\int _{\Omega } f\), respectively. In addition, C and \(C_{i}\) \((i=1,2,3, \cdots )\) stand for generic positive constants which may vary from line to line.

  • \(W^{k,p}(\Omega )=\{u\in L^p(\Omega ): D^\alpha u\in L^p(\Omega ) \ \text {for}\ 0\le |\alpha |\le k\} \) denotes the usual Sobolev space, where \(D^\alpha u\) is the weak partial derivative. If \(p=2\), we write \(H^k(\Omega )=W^{k,2}(\Omega )\).

In this section, we establish the local existence of solutions to the system (1.4) under appropriate initial conditions. Moreover, we shall collect and prove some useful inequalities which will be used in the subsequent sections. To begin with, we consider the regularity of the solution \(\textbf{w}\) to the following system:

$$\begin{aligned} \left\{ \begin{array}{lll} -\Delta \textbf{w}+\mu \textbf{w}=\textbf{f}&{} \text{ in } \Omega , \\ \textbf{w}\cdot \textbf{n}\mid _{\partial \Omega }=0,\ \partial _{\textbf{n}}\textbf{w}\times \textbf{n}\mid _{\partial \Omega }=\textbf{0}, \quad &{} \text{ if } n=2,3, \end{array}\right. \end{aligned}$$
(2.1)

where \(\mu \in \left\{ 0,1\right\} \) and \(\textbf{f} \in \left( L^{p}(\Omega )\right) ^{n}\) for some \(1<p<+\infty \). The system (2.1) has the following properties.

Lemma 2.1

(cf. [38, Theorem 3, Theorem 4, Remark 5]) Let \(\mu =0\) and \(\Omega \subset \mathbb {R}^{n}\), \(n\in \left\{ 2,3\right\} \), be a bounded domain with a smooth boundary.

  1. (i)

    If \(\textbf{f} \in \left( L^{p}(\Omega )\right) ^{n}\) with some \(1<p<+\infty \), then the system (2.1) has a unique solution \(\textbf{w}\in (W^{2, p}(\Omega ))^n\) satisfying

    $$\begin{aligned} \Vert \textbf{w}\Vert _{W^{2, p}(\Omega )} \le C(p,n, \Omega )\Vert \textbf{f}\Vert _{{L^{p}(\Omega )}}, \end{aligned}$$
    (2.2)

    where \(C(p, n,\Omega )\) is a positive constant depending only on p, n and \(\Omega \).

  2. (ii)

    If \(\textbf{f} \in (C^{k,\alpha }(\bar{\Omega }))^n\) with some \(\alpha \in (0,1)\) and \(k\in \mathbb N\), then the system (2.1) has a unique solution \(\textbf{w}\in C^{k+2,\beta }(\bar{\Omega })\) for some \(\beta \in (0,1)\) and

    $$\begin{aligned} \Vert \textbf{w}\Vert _{C^{k+2,\beta }(\bar{\Omega })} \le C(k,\alpha ,\beta ,n, \Omega )\Vert \textbf{f}\Vert _{C^{k,\alpha }(\bar{\Omega })}. \end{aligned}$$
    (2.3)

    where \(C(k,\alpha ,\beta ,n, \Omega )\) is a positive constant depending only on k, \(\alpha \), \(\beta \), n and \(\Omega \).

Now we prove the following results for the case \(\mu =1\).

Lemma 2.2

Let \(\mu =1\) and \(\Omega \subset \mathbb {R}^{n}\), \(n\in \left\{ 2,3\right\} \), be a bounded domain with a smooth boundary.

  1. (i)

    If \(\textbf{f} \in \left( H^{k}(\Omega )\right) ^{n}\) with \(k\in \left\{ 0,1,2\right\} \), then the system (2.1) has a unique solution \(\textbf{w}\in H^{k+2}(\Omega )\) satisfying

    $$\begin{aligned} \Vert \textbf{w}\Vert _{H^{k+2}(\Omega )} \le C(k,n, \Omega )\Vert \textbf{f}\Vert _{H^{k}(\Omega )}, \end{aligned}$$
    (2.4)

    where \(C(k, n,\Omega )\) is a positive constant depending only on k, n and \(\Omega \).

  2. (ii)

    If \(\textbf{f} \in \left( L^{p}(\Omega )\right) ^{n}\) with some \(1<p<+\infty \), then the system (2.1) has a unique solution \(\textbf{w}\in (W^{2, p}(\Omega ))^n\) satisfying

    $$\begin{aligned} \Vert \textbf{w}\Vert _{W^{2, p}(\Omega )} \le C(p,n, \Omega )\Vert \textbf{f}\Vert _{{L^{p}(\Omega )}}, \end{aligned}$$
    (2.5)

    where \(C(p, n,\Omega )\) is a positive constant depending only on p, n and \(\Omega \).

  3. (iii)

    If \(\textbf{f} \in (C^{k,\alpha }(\bar{\Omega }))^n\) with some \(\alpha \in (0,1)\) and \(k\in \left\{ 0,1,2\right\} \), then the system (2.1) has a unique solution \(\textbf{w}\in C^{k+2,\beta }(\bar{\Omega })\) for some \(\beta \in (0,1)\) and

    $$\begin{aligned} \Vert \textbf{w}\Vert _{C^{k+2,\beta }(\bar{\Omega })} \le C(k,\alpha ,\beta ,n, \Omega )\Vert \textbf{f}\Vert _{C^{k,\alpha }(\bar{\Omega })}. \end{aligned}$$
    (2.6)

    where \(C(k,\alpha ,\beta ,n, \Omega )\) is a positive constant depending only on k, \(\alpha \), \(\beta \), n and \(\Omega \).

Proof

The existence and uniqueness of the solution to (2.1) for \(\mu =1\) can be established by similar arguments for (2.1) in the case of \(\mu =0\) (cf. [38]). Below we only prove the regularity properties given in (2.4)–(2.6).

(i) Since the proof of (2.4) involves tedious calculations, we place it in Appendix A (see Lemma A1 and Lemma A2 in Appendix A).

(ii) Next we prove (2.5). The first equation of (2.1) can be rewritten as

$$\begin{aligned} -\Delta \textbf{w}=\textbf{f}-\textbf{w}. \end{aligned}$$
(2.7)

Therefore, in view of (2.2) and (2.7), it is sufficient to prove that for \(p\in (1,\infty )\), there exists a positive constant \(C(p,n,\Omega )\) such that

$$\begin{aligned} \Vert \textbf{f}-\textbf{w}\Vert _{L^{p}(\Omega )}\le C(p,n,\Omega )\Vert \textbf{f}\Vert _{L^{p}(\Omega )}. \end{aligned}$$
(2.8)

We consider three cases: \(p\ge 2\), \(\frac{6}{5}\le p<2\) and \(1<p<\frac{6}{5}\). First, if \(p\ge 2\), it follows from \(\textbf{f}\in ({L^{p}(\Omega )})^n\hookrightarrow ({L^{2}(\Omega )})^n \) for \(p\ge 2\) and (2.4) that \(\textbf{w}\in (H^{2}(\Omega ))^n\) with \(\left\| \textbf{w}\right\| _{H^2(\Omega )}\le C(p,n,\Omega )\Vert \textbf{f}\Vert _{L^{p}(\Omega )}\), which alongside the Sobolev embedding \((H^{2}(\Omega ))^n\hookrightarrow ({L^{p}(\Omega )})^n\) yields

$$\begin{aligned} \Vert \textbf{w}\Vert _{L^{p}(\Omega )}\le C(p,n,\Omega )\Vert \textbf{f}\Vert _{L^{p}(\Omega )} \end{aligned}$$

and hence (2.8) holds for \(p\ge 2\). Secondly, if \(\frac{6}{5}\le p<2\), we claim that

$$\begin{aligned} \text {the solution }\textbf{w}\text { to }(2.1)\text { satisfies } \left\| \textbf{w}\right\| _{H^1(\Omega )}\le C(p,n,\Omega )\Vert \textbf{f}\Vert _{L^{p}(\Omega )}. \end{aligned}$$
(2.9)

To this end, we define the real Hilbert space (cf. [10])

$$\begin{aligned}&X:=\left\{ \textbf{u} \in (H^1(\Omega ))^n \mid \textbf{u} \cdot \textbf{n}=0\text { on }\partial \Omega \right\} \quad \text {with the norm}\\&\quad \left\| \textbf{u} \right\| _{X}:=\left( \int _\Omega |\nabla \textbf{u} |^2\right) ^\frac{1}{2}. \end{aligned}$$

Since \(H^1(\Omega )\hookrightarrow {L^{6}(\Omega )}\hookrightarrow {L^{p^*}(\Omega )}=({L^{p}(\Omega )})^*\) with \(p^*:=\frac{p}{p-1}\), we have \(\textbf{f}\in ({L^{p}(\Omega )})^n\hookrightarrow (H^{1}(\Omega ))^*\). Define the bilinear form \(B[\cdot ,\cdot ]\) on X by

$$\begin{aligned} B[\textbf{u}, \textbf{v}]:=\int _\Omega \nabla \textbf{u} \cdot \nabla \textbf{v}+\int _\Omega \textbf{u} \cdot \textbf{v},\quad \forall \ \textbf{u}, \textbf{v}\in X. \end{aligned}$$

Then we have

$$\begin{aligned} |B[\textbf{u}, \textbf{v}]|&\le \Vert \nabla \textbf{u} \Vert _{L^{2}(\Omega )}\Vert \nabla \textbf{v}\Vert _{L^{2}(\Omega )}+\Vert \textbf{u} \Vert _{L^{2}(\Omega )}\Vert \textbf{v}\Vert _{L^{2}(\Omega )}\\&\le \left\| \textbf{u} \right\| _{H^1(\Omega )}\left\| \textbf{v}\right\| _{H^1(\Omega )}\\&\le C\left\| \textbf{u} \right\| _{X}\left\| \textbf{v}\right\| _{X}\quad \forall \ \textbf{u}, \textbf{v}\in X, \end{aligned}$$

where we have used the fact that the X norm is equivalent to the usual \(H^1(\Omega )\) norm (cf. [12]). Moreover,

$$\begin{aligned} B[\textbf{u}, \textbf{u} ]=\left\| \textbf{u} \right\| _{H^1(\Omega )}^2 \ge C\left\| \textbf{u} \right\| _{X}^2, \quad \forall \ \textbf{u} \in X. \end{aligned}$$
(2.10)

Therefore, by the Lax–Milgram theorem, we obtain that for each \(\textbf{f}\in ({L^{p}(\Omega )})^n\) with \(\frac{6}{5}\le p<2\), there exists a unique \(\textbf{u} _f\in X\) such that

$$\begin{aligned} B[\textbf{u} _f,\textbf{v}]=\int _\Omega \textbf{f}\cdot \textbf{v},\quad \forall \ \textbf{v}\in X. \end{aligned}$$

Therefore, by (2.10), Hölder’s inequality, the Sobolev embedding \(H^1(\Omega )\hookrightarrow {L^{p^*}(\Omega )}\) and the equivalence of X norm and \(H^1(\Omega )\) norm, we have

$$\begin{aligned} \left\| \textbf{u} \right\| _{H^1(\Omega )}^2&\le C\left\| \textbf{u} \right\| _{X}^2 \le CB[\textbf{u}, \textbf{u} ] \le C\int _\Omega \textbf{f}\cdot \textbf{u} \le C\Vert \textbf{f}\Vert _{L^{p}(\Omega )}\Vert \textbf{u} \Vert _{L^{p^*}(\Omega )} \\&\le C\Vert \textbf{f}\Vert _{L^{p}(\Omega )}\left\| \textbf{u} \right\| _{H^1(\Omega )} \end{aligned}$$

for all \(\textbf{u} \in X\), that is

$$\begin{aligned} \left\| \textbf{u} \right\| _{H^1(\Omega )}\le C\Vert \textbf{f}\Vert _{L^{p}(\Omega )},\quad \forall \ \textbf{u} \in X. \end{aligned}$$

Therefore, the claim (2.9) is proved since the solution \(\textbf{w}\) to (2.1) satisfies \(\textbf{w}\in X\), and hence (2.8) holds for \(\frac{6}{5}\le p<2\) due to the Sobolev embedding \(H^1(\Omega )\hookrightarrow {L^{p}(\Omega )}\). It remains to consider the case \(1<p<\frac{6}{5}\). We let \(\alpha :=\frac{p}{3-2p}\in (1,2)\) which satisfies

$$\begin{aligned} 3\alpha =(\alpha -1)p^*. \end{aligned}$$
(2.11)

Then \(|\textbf{w}|^\alpha =\left( \sum \limits _{i=1}^n w_i^2\right) ^\frac{\alpha }{2}\) satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta |\textbf{w}|^\alpha =\alpha (\alpha -2)|\textbf{w}|^{\alpha -4}|\textbf{w}\cdot \nabla \textbf{w}|^2 +\alpha |\textbf{w}|^{\alpha -2}\left( |\nabla \textbf{w}|^2+\textbf{w}\cdot \Delta \textbf{w}\right) &{} \text{ in } \Omega ,\\ \partial _{\textbf{n}}|\textbf{w}|^\alpha =\alpha \sum \limits _{i,j=1}^n|\textbf{w}|^{\alpha -2}w_i\partial _j w_i n_j=\alpha |\textbf{w}|^{\alpha -2}\textbf{w}\cdot \partial _{\textbf{n}}\textbf{w}=0&{} \text{ on } \partial \Omega , \end{array}\right. \end{aligned}$$
(2.12)

where we have used \(\textbf{w}\cdot \partial _{\textbf{n}}\textbf{w}=0\) on \(\partial \Omega \) due to the boundary conditions of \(\textbf{w}\) in (2.1). Integrating the first equation of (2.12) on \(\Omega \) by parts with \(\partial _{\textbf{n}}|\textbf{w}|^\alpha =0\) on \(\partial \Omega \) and using \(-\Delta \textbf{w}=\textbf{f}-\textbf{w}\) in \(\Omega \), one has

$$\begin{aligned}&(\alpha -2)\int _\Omega |\textbf{w}|^{\alpha -4}|\textbf{w}\cdot \nabla \textbf{w}|^2 +\int _\Omega |\textbf{w}|^{\alpha -2}|\nabla \textbf{w}|^2\\&\quad =-\int _\Omega |\textbf{w}|^{\alpha -2}\textbf{w}\cdot \Delta \textbf{w}=\int _\Omega |\textbf{w}|^{\alpha -2}\textbf{w}\cdot \left( \textbf{f}-\textbf{w}\right) , \end{aligned}$$

which together with \(\alpha -2<0\) and \(\int _\Omega |\textbf{w}|^{\alpha -4}|\textbf{w}\cdot \nabla \textbf{w}|^2\le \int _\Omega |\textbf{w}|^{\alpha -2}|\nabla \textbf{w}|^2\) implies

$$\begin{aligned} (\alpha -1)\int _\Omega |\textbf{w}|^{\alpha -2}|\nabla \textbf{w}|^2+\int _\Omega |\textbf{w}|^\alpha \le \int _\Omega |\textbf{w}|^{\alpha -1}|\textbf{f}|. \end{aligned}$$
(2.13)

With (2.11) and (2.13), we are in a position to use the same arguments as in [38, pp. 133–134] to show that

$$\begin{aligned} \Vert \textbf{w}\Vert _{L^{3\alpha }(\Omega )}\le C\Vert \textbf{f}\Vert _{L^{p}(\Omega )}. \end{aligned}$$

By \(3\alpha >3\) and Hölder’s inequality: \({L^{3\alpha }(\Omega )}\hookrightarrow {L^{p}(\Omega )}\) for \(1<p<\frac{6}{5}\), we know that (2.8) holds for \(1<p<\frac{6}{5}\). Therefore, we have proved that (2.8) holds for \(p\in (1,\infty )\) as desired.

(iii) If \(\textbf{f} \in (C^{k,\alpha }(\bar{\Omega }))^n\) with some \(\alpha \in (0,1)\) and \(k\in \left\{ 0,1,2\right\} \), then (2.4) implies

$$\begin{aligned} \Vert \textbf{w}\Vert _{H^{k+2}(\Omega )} \le C(k,n, \Omega )\Vert \textbf{f}\Vert _{H^{k}(\Omega )}\le C(k,n, \Omega ), \end{aligned}$$

which together with the Sobolev embedding \(H^{k+2}(\Omega )\hookrightarrow C^{k,\theta }(\bar{\Omega })\) for any \(\theta \in \left( 0,2-\frac{n}{2}\right) \) shows that

$$\begin{aligned} \textbf{f} -\textbf{w}\in (C^{k,\alpha '}(\bar{\Omega }))^n \end{aligned}$$
(2.14)

for any \(\alpha '\in (0,\min \left\{ \alpha ,\frac{1}{2}\right\} )\). In view of (2.3), (2.7) and (2.14), (2.6) is proved.

\(\square \)

We are now in a position to show the local existence of the unique classical solution to (1.4).

Lemma 2.3

Let \(\Omega \subset \mathbb {R}^2\) be a bounded domain with a smooth boundary. Assume \(u_{10}, u_{20} \in W^{1, p}(\Omega )\) with \(p>2\) and \(u_{10}, u_{20}\gvertneqq 0\). Then there exists \(T_{max} \in (0, \infty ]\) such that the system (1.4) has a unique classical solution \((u_1,u_2,\textbf{w}_1,\textbf{w}_2)\) satisfying

$$\begin{aligned} \left\{ \begin{array}{lcl} u_i \in C^0(\bar{\Omega } \times [0, {T_{max}})) \cap C^{2,1}(\bar{\Omega } \times (0, {T_{max}})),\\ \textbf{w}_i \in \left[ C^{2,1}(\bar{\Omega } \times (0, {T_{max}}))\right] ^2, \end{array}\right. \quad i=1,2, \end{aligned}$$

and \(u_1,u_2>0\) in \(\Omega \times (0, T_{max})\). Moreover,

$$\begin{aligned} \text {if}\quad T_{max}<\infty ,\quad \text {then}\quad \lim _{t\rightarrow T_{max}}\left( \Vert u_1(\cdot ,t)\Vert _{W^{1, p}(\Omega )} +\Vert u_2(\cdot ,t)\Vert _{W^{1, p}(\Omega )}\right) =\infty . \end{aligned}$$
(2.15)

Proof

Fix \(R>0\), and define for \(T \in (0,1)\)

$$\begin{aligned}{} & {} X_R(T):=\left\{ (u_1,u_2) \in C\left( [0, T];W^{1,p}(\Omega )\right) \right. \\{} & {} \qquad \qquad \left. \bigg | \sup \limits _{t \in [0, T]}\Vert u_1(t)\Vert _{W^{1, p}} +\sup \limits _{t \in [0, T]}\Vert u_2(t)\Vert _{W^{1, p}} \le R\right\} , \end{aligned}$$

which is a complete metric space with the metric

$$\begin{aligned} d_{X}(\textbf{u}, \textbf{v})=\sup _{t \in [0, T]}\Vert u_1(t)-v_1(t)\Vert _{W^{1, p}}+\sup _{t \in [0, T]}\Vert u_2(t)-v_2(t)\Vert _{W^{1, p}} \end{aligned}$$

for \(\textbf{u}=(u_1(t),u_2(t))\in X_{R}(T)\) and \(\textbf{v} =(v_1(t),v_2(t))\in X_{R}(T) \). For any \(\textbf{u}=(u_1,u_2) \in X_{R}(T)\) and \(t \in [0, T]\), by \(p>2\) and Lemma 2.2 we know that there exists a unique solution \((\textbf{w}_1,\textbf{w}_2)\in \left( H^{2}(\Omega )\right) ^{2}\times \left( H^{2}(\Omega )\right) ^{2}\) to the following system

$$\begin{aligned} \left\{ \begin{array}{lll} -\varepsilon _i \Delta \textbf{w}_{i}(t)+\textbf{w}_{i}(t) =\nabla E_i(u_1,u_2),\quad &{}i=1,2,&{} \text{ in } \Omega , \\ \textbf{w}_i \cdot \textbf{n}=0,\ \partial _{\textbf{n}}\textbf{w}_i\times \textbf{n}=0, \quad &{}i=1,2,&{} \text { on }\partial \Omega . \\ \end{array}\right. \end{aligned}$$
(2.16)

Letting \(\left( e^{t \Delta }\right) _{t \ge 0}\) denote the Neumann heat semigroup on \(\Omega \), we introduce a mapping \(\varvec{\Phi }=\left( \Phi _{1}, \Phi _{2}\right) \) on \(X_R(T)\) by defining

$$\begin{aligned} \tilde{u}_i(t)= & {} \Phi _{i}(u_1,u_2)(\cdot , t):=e^{t d_i\Delta } u_{i0}\\{} & {} +\int _{0}^{t} e^{(t-s) d_i\Delta }\left\{ -\nabla \cdot \left( \chi _i u_{i} \textbf{w}_{i}\right) + u_{i}E_i(u_1,u_2) \right\} (\cdot , s)d s \end{aligned}$$

for \(i=1,2\), where \((\textbf{w}_1,\textbf{w}_2)\in \left( H^{2}(\Omega )\right) ^{2}\times \left( H^{2}(\Omega )\right) ^{2}\) is the solution of (2.16) uniquely determined by the given \(\textbf{u}=(u_1,u_2)\in X_{R}(T)\) according to Lemma 2.2. Then one can show that \(\varvec{\Phi }\) is a contraction map from \(X_{R}(T)\) into itself if T is sufficiently small by a standard argument (see e.g. [34, Theorem 2.2]). Therefore, for sufficiently small T, by the Banach fixed point theorem, there is a unique \((u_1,u_2) \in X_{R}(T)\) such that

$$\begin{aligned} (u_1,u_2)=(\tilde{u}_1,\tilde{u}_2)=\varvec{\Phi }(u_1,u_2)=\left( \Phi _{1}(u_1,u_2), \Phi _{2}(u_1,u_2)\right) , \end{aligned}$$

and \((u_1,u_2,\textbf{w}_1,\textbf{w}_2)\) is a unique strong solution of the system (1.4) satisfying

$$\begin{aligned} \left\{ \begin{array}{ll} u_i \in C\left( \left[ 0, T\right) , W^{1, p}(\Omega )\right) \cap C\left( \left( 0, T\right) , W^{2, p}(\Omega )\right) ,\\ \textbf{w}_i(t)\in \left( H^2(\Omega )\right) ^2\quad \text {for all }t\in [0, T), \end{array}\right. \quad i=1,2. \end{aligned}$$
(2.17)

Based on a bootstrap argument, we can use Lemma 2.2, the \(L^{p}\)-estimate and the Schauder estimate (cf. [28]) to show that the unique strong solution \((u_1,u_2,\textbf{w}_1,\textbf{w}_2)\) of the system (1.4) satisfying (2.17) is actually a classical solution. Finally, \(u_1\), \(u_2\ge 0\) follows from the maximum principle. To be precise, we rewrite the first equation of system (1.4) as

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u_{1}-d_{1} \Delta u_{1}+\chi _1 \textbf{w}_{1}\cdot \nabla u_1+Q_1(x,t)u_{1}=0, &{} x \in \Omega , t \in \left( 0, T\right) , \\ \frac{\partial u_1}{\partial \textbf{n}}=0, &{} x \in \partial \Omega , t \in \left( 0, T\right) , \\ u_1(x, 0)=u_{10} (x), &{} x \in \Omega , \end{array}\right. \end{aligned}$$
(2.18)

where \(Q_1(x,t):=\chi _1\nabla \cdot \textbf{w}_1-E_1(u_1,u_2)\) for \((x,t)\in \Omega \times (0,T)\). Then one can apply the strong maximum principle to system (2.18) and gets that \(u_1(x,t)>0\) for \((x,t)\in \Omega \times (0,T)\). Similarly, it holds that \(u_2(x,t)>0\) for \((x,t)\in \Omega \times (0,T)\). The proof is completed. \(\square \)

Next we prove a useful Lemma.

Lemma 2.4

Let \(\Omega \subset \mathbb {R}^n\), \(n\ge 1\), be a bounded domain with a smooth boundary, and let \(r,q\ge 1\) be two constants satisfying

$$\begin{aligned} \frac{1}{q}>\frac{1}{2}-\frac{1}{n}\quad \text {and}\quad \frac{1}{r}>\frac{1}{q}-\frac{1}{n}. \end{aligned}$$
(2.19)

Then for any \(\varphi \in H^{2}(\Omega )\) satisfying \(\left. \frac{\partial \varphi }{\partial \textbf{n}}\right| _{\partial \Omega }=0\), there exists a positive constant C depending only on \(\Omega \), n, q and r such that

$$\begin{aligned} \Vert \nabla \varphi \Vert _{{L^{q}(\Omega )}} \le C\left( \Vert \Delta \varphi \Vert _{{L^{2}(\Omega )}}^\theta \Vert \varphi \Vert _{{L^{r}(\Omega )}}^{1-\theta }+\Vert \varphi \Vert _{{{L^{r}(\Omega )}}}\right) , \end{aligned}$$

where

$$\begin{aligned} \theta =\frac{\frac{1}{r}+\frac{1}{n}-\frac{1}{q}}{\frac{1}{r}+\frac{2}{n}-\frac{1}{2}}\in (0,1). \end{aligned}$$
(2.20)

Proof

First, one can use (2.19) to check that \(\theta \) defined by (2.20) satisfies \(\theta \in (0,1)\). Using the Gagliardo–Nirenberg inequality, we have

$$\begin{aligned} \Vert \nabla \varphi \Vert _{L^{q}(\Omega )}\le C\left( \Vert D^2 \varphi \Vert _{L^{2}(\Omega )}^\theta \Vert \varphi \Vert _{L^{r}(\Omega )}^{1-\theta }+\Vert \varphi \Vert _{L^{r}(\Omega )}\right) . \end{aligned}$$
(2.21)

Under the homogeneous Neumann boundary condition \(\left. \frac{\partial \varphi }{\partial \textbf{n}}\right| _{\partial \Omega }=0\), it follows from [3, Lemma 1] that \(\Vert \nabla \varphi \Vert _{H^1(\Omega )} \le C\Vert \Delta \varphi \Vert _{{L^{2}(\Omega )}}\), which implies

$$\begin{aligned} \Vert D^2 \varphi \Vert _{{L^{2}(\Omega )}} \le \Vert \nabla \varphi \Vert _{H^1(\Omega )} \le C\Vert \Delta \varphi \Vert _{{L^{2}(\Omega )}}. \end{aligned}$$
(2.22)

The proof is completed by substituting (2.22) into (2.21). \(\square \)

We recall the following basic result which will be used to investigate the global stability of solutions.

Lemma 2.5

([43, Lemma 1.1]) Let \(\tau \ge 0\) and \(c>0\) be two constants, \(F(t) \ge 0, \int _{\tau }^{\infty } H(t) \textrm{d} t<\infty \). Assume that \(E \in C^{1}([\tau , \infty )), E\) is bounded from below and satisfies

$$\begin{aligned} E^{\prime }(t) \le -c F(t)+H(t) \quad \text{ in } [\tau , \infty ). \end{aligned}$$

If either \(F \in C^{1}([\tau , \infty ))\) and \(F^{\prime }(t) \le k\) in \([\tau , \infty )\) for some \(k>0,\) or \(F \in C^{\alpha }([\tau , \infty ))\) and \(\Vert F\Vert _{C^{\alpha }([\tau , \infty ))} \le k\) for constants \(0<\alpha <1\) and \(k>0\), then \(\lim _{t \rightarrow \infty } F(t)=0.\)

3 Boundedness of solutions

In this section, we focus on the global boundedness of solutions of the system (1.4). Throughout this section, we assume that the conditions in Theorem 1.1 hold and \((u_1, u_2,\textbf{w}_1,\textbf{w}_2)\) be a local-in-time classical solution of the system (1.4) obtained in Lemma 2.3 with a maximal time \(T_{max}\). First of all, we give the following basic property for the solution components \(\textbf{w}_1\) and \(\textbf{w}_2\).

Lemma 3.1

For \(i,j\in \left\{ 1,2\right\} \), the solution of (1.4) satisfies

$$\begin{aligned} \int _\Omega \Delta \textbf{w}_i\cdot \textbf{w}_j=-\int _\Omega \nabla \textbf{w}_i\cdot \nabla \textbf{w}_j. \end{aligned}$$

Proof

Denote \(\textbf{w}_i=({w}_1^{(i)},{w}_2^{(i)})\), \(i=1,2\). Using integration by parts we get

$$\begin{aligned} \int _\Omega \Delta \textbf{w}_i\cdot \textbf{w}_j&=\sum _{k=1}^2 \int _\Omega \Delta {w}^{(i)}_k\cdot {w}^{(j)}_k\nonumber \\&=-\sum _{k=1}^2 \int _\Omega \nabla {w}^{(i)}_k\cdot \nabla {w}^{(j)}_k+\sum _{k=1}^2 \int _{\partial \Omega }\left( \nabla {w}^{(i)}_k\cdot \textbf{n}\right) {w}^{(j)}_kdS. \end{aligned}$$
(3.1)

By (1.6) we have

$$\begin{aligned} \sum _{k=1}^2 \int _{\partial \Omega }\left( \nabla {w}^{(i)}_k\cdot \textbf{n}\right) {w}^{(j)}_kdS=\sum _{k=1}^2 \int _{\partial \Omega }{w}^{(j)}_k\partial _{\textbf{n}}{w}^{(i)}_kdS=\int _{\partial \Omega }\textbf{w}_j\cdot \partial _{\textbf{n}} \textbf{w}_idS=0. \end{aligned}$$

This together with (3.1) completes the proof. \(\square \)

The following result is a basic property about the uniform-in-time boundedness of \(u_1\) and \(u_2\) in \({L^{1}(\Omega )}\).

Lemma 3.2

Suppose that the conditions in Theorem 1.1 hold. Then for \(i=1,2\), it holds that

$$\begin{aligned} \Vert u_i(\cdot ,t)\Vert _{L^{1}(\Omega )}\le \max \left\{ \Vert u_{i0}\Vert _{L^{1}(\Omega )},\frac{(1+\gamma _i)^2}{4}|\Omega |\right\} \quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.2)

Proof

Integrating the first equation of (1.4) with respect to \(x\in \Omega \), and using \(\nabla u_{1} \cdot \textbf{n} \mid _{\partial \Omega }=\textbf{w}_1 \cdot \textbf{n} \mid _{\partial \Omega }=0\), \(u_1\), \(u_2>0\) and the Young’s inequality \(s\le \frac{1}{1+\gamma _1}s^2+\frac{1+\gamma _1}{4}\) for \(s\ge 0\), we have

$$\begin{aligned} \frac{d}{dt}\int _\Omega u_1&\le \gamma _1\int _\Omega u_1-\int _\Omega u_1^2\nonumber \\&\le \gamma _1\int _\Omega u_1-\int _\Omega \left[ \left( 1+\gamma _1\right) u_1-\frac{(1+\gamma _1)^2}{4}\right] \nonumber \\&= -\int _\Omega u_1+\frac{(1+\gamma _1)^2}{4}|\Omega |\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.3)

An application of Grönwall’s inequality to (3.3) yields (3.2) for \(i=1\). Similarly, (3.2) holds for \(i=2\). This completes the proof. \(\square \)

Now we are in a position to derive the following estimates.

Lemma 3.3

Suppose that the conditions in Theorem 1.1 hold. Then there exist two constants \(k>0\) and \(C>0\) independent of t such that for all \(\tau \in [0,T_{max})\) and \(t\in (0,T_{max}-\tau )\), it hold that

$$\begin{aligned} {\int _\Omega u_1\log u_1+\int _\Omega u_2\log u_2}\le C \end{aligned}$$
(3.4)

and

$$\begin{aligned} \sum _{i=1}^2\int _t^{t+\tau }\left( \Vert \nabla \sqrt{u_i}\Vert _{{L^{2}(\Omega )}}^2+\Vert u_i\Vert _{L^{2}(\Omega )}^2+\Vert \textbf{w}_i\Vert _{W^{1,2}(\Omega )}^2+\int _\Omega u_i^2\log u_i\right) ds\le k\tau +C. \end{aligned}$$
(3.5)

Proof

Multiplying the first equation in (1.4) by \((\log u_1+1)\) and integrating the resulting equation by parts along with the boundary conditions in (1.4), we have

$$\begin{aligned} \frac{d}{dt}\int _\Omega u_1\log u_1+d_1\int _\Omega \frac{|\nabla u_1|^2}{u_1}={\chi _1\int _\Omega \textbf{w}_1\cdot \nabla u_1}+\int _\Omega u_1E_1(u_1,u_2)(\log u_1+1) \end{aligned}$$
(3.6)

for all \(t\in (0,T_{max})\). For the first term on the right hand side of (3.6), using integration by parts subject to the boundary conditions \(\textbf{w}_1 \cdot \textbf{n} \mid _{\partial \Omega }=0\), one has

$$\begin{aligned} {\chi _1\int _\Omega \textbf{w}_1\cdot \nabla u_1}&=-\chi _1\int _\Omega u_1\nabla \cdot \textbf{w}_1\le \frac{\varepsilon _1}{4}\Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\\&\quad +\frac{\chi _1^2}{\varepsilon _1}\Vert u_1\Vert _{L^{2}(\Omega )}^2\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$

Multiplying the third equation of (1.4) by \(\textbf{w}_1\), integrating the resulting equation by parts along with the boundary conditions in (1.4), and using Lemma 3.1 and Young’s inequality, for all \(t\in (0,T_{max})\) one has

$$\begin{aligned} \varepsilon _1 \Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+ \Vert \textbf{w}_1\Vert _{L^{2}(\Omega )}^2&= \int _\Omega \left( u_1+cu_2\right) \nabla \cdot \textbf{w}_1\nonumber \\&\le \frac{\varepsilon _1}{4}\Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+\frac{2}{\varepsilon _1}\Vert u_1\Vert _{L^{2}(\Omega )}^2+\frac{2c^2}{\varepsilon _1}\Vert u_2\Vert _{L^{2}(\Omega )}^2. \end{aligned}$$
(3.7)

The combination of (3.6)–(3.7) shows that

$$\begin{aligned}&\frac{d}{dt}\int _\Omega u_1\log u_1+d_1\int _\Omega \frac{|\nabla u_1|^2}{u_1}+\frac{\varepsilon _1}{2} \Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+ \Vert \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\nonumber \\&\quad \le \int _\Omega u_1E_1(u_1,u_2)(\log u_1+1)+\frac{(2+\chi _1^2)}{\varepsilon _1}\Vert u_1\Vert _{L^{2}(\Omega )}^2+\frac{2c^2}{\varepsilon _1}\Vert u_2\Vert _{L^{2}(\Omega )}^2 \end{aligned}$$
(3.8)

for all \(t\in (0,T_{max})\). Similarly, for \(u_2\), it holds that

$$\begin{aligned}&\frac{d}{dt}\int _\Omega u_2\log u_2+d_2\int _\Omega \frac{|\nabla u_2|^2}{u_2}+\frac{\varepsilon _2}{2} \Vert \nabla \textbf{w}_2\Vert _{L^{2}(\Omega )}^2+ \Vert \textbf{w}_2\Vert _{L^{2}(\Omega )}^2\nonumber \\&\quad \le \int _\Omega u_2E_2(u_1,u_2)(\log u_2+1)+\frac{(2+\chi _2^2)}{\varepsilon _2}\Vert u_2\Vert _{L^{2}(\Omega )}^2+\frac{2b^2}{\varepsilon _2}\Vert u_1\Vert _{L^{2}(\Omega )}^2 \end{aligned}$$
(3.9)

for all \(t\in (0,T_{max})\). Using (3.8) and (3.9), for all \(t\in (0,T_{max})\) one has

$$\begin{aligned}&\frac{d}{dt}\int _\Omega \left( u_1\log u_1+u_2\log u_2\right) +\int _\Omega \left( u_1\log u_1+ u_2\log u_2\right) +d_1\int _\Omega \frac{|\nabla u_1|^2}{u_1}\nonumber \\&\qquad +d_2\int _\Omega \frac{|\nabla u_2|^2}{u_2} +\frac{\varepsilon _1}{2} \Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+ \Vert \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+\frac{\varepsilon _2}{2} \Vert \nabla \textbf{w}_2\Vert _{L^{2}(\Omega )}^2\nonumber \\&\qquad + \Vert \textbf{w}_2\Vert _{L^{2}(\Omega )}^2 +\Vert u_1\Vert _{L^{2}(\Omega )}^2+\Vert u_2\Vert _{L^{2}(\Omega )}^2+\frac{1}{2}\int _\Omega {u_1^2\log u_1+ \frac{1}{2}\int _\Omega u_2^2\log u_2}\nonumber \\&\quad \le \underbrace{\int _\Omega u_1E_1(u_1,u_2)(\log u_1+1)+q_1\Vert u_1\Vert _{L^{2}(\Omega )}^2+\int _\Omega u_1\log u_1 +\frac{1}{2}\int _\Omega u_1^2\log u_1}_{=:I_1}\nonumber \\&\qquad +\underbrace{\int _\Omega u_2E_2(u_1,u_2)(\log u_2{+}1){+}q_2\Vert u_2\Vert _{L^{2}(\Omega )}^2{+}\int _\Omega u_2\log u_2 {+}\frac{1}{2}\int _\Omega u_2^2\log u_2}_{=:I_2}, \end{aligned}$$
(3.10)

where \(q_1:=\frac{(2+\chi _1^2)}{\varepsilon _1}+\frac{2b^2}{\varepsilon _2}+1\) and \(q_2:=\frac{(2+\chi _2^2)}{\varepsilon _2}+\frac{2c^2}{\varepsilon _1}+1\). Clearly, the following results hold:

$$\begin{aligned} s\log s\le s^2\quad \text {and}\quad -s(\log s+1)\le -s\log s\le \frac{1}{e}\le 1\quad \text {for all }s\ge 0. \end{aligned}$$
(3.11)

Making use of (1.5), (3.2) and (3.11), we obtain

$$\begin{aligned} I_1&=\int _\Omega u_1\left( \gamma _1-u_{1}-c u_{2}\right) (\log u_1+1){+}q_1\int _\Omega u_1^2{+}\int _\Omega u_1\log u_1{+}\frac{1}{2}\int _\Omega u_1^2\log u_1,\nonumber \\&= \left( 1+\gamma _1\right) \int _\Omega u_1\log u_1 +\gamma _1\int _\Omega u_1- \frac{1}{2}\int _\Omega u_1^2\log u_1\nonumber \\&\quad -c\int _\Omega u_1u_2(\log u_1+1)+(q_1-1)\int _\Omega u_1^2,\nonumber \\&\le \left( 1+\gamma _1\right) \int _\Omega u_1^2+\gamma _1\int _\Omega u_1- \frac{1}{2}\int _\Omega u_1^2\log u_1+c\int _\Omega u_2+(q_1-1)\int _\Omega u_1^2,\nonumber \\&\le - \frac{1}{2} \int _\Omega u_1^2\left[ \log u_1-2(\gamma _1+q_1)\right] +C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.12)

The continuous function \(\phi (s):=s^2\left[ \log s-2(\gamma _1+q_1)\right] \) for \(s>0\) is bounded from below. In fact, \(\phi (s)>0\) for \(s\in (s_*,\infty )\) with \(s_*:=e^{2(\gamma _1+q_1)}>0\) and

$$\begin{aligned} \phi (s)\ge s^2\log s-2(\gamma _1+q_1)s_*^2\ge -\frac{1}{2e}-2(\gamma _1+q_1)s_*^2\quad \text {for }s\in (0,s_*], \end{aligned}$$

where we have used the fact that \(s^2\log s\ge -\frac{1}{2e}\) for \(s\in (0,\infty )\). This along with (3.12) shows that

$$\begin{aligned} I_1\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.13)

Similarly, one can deduce that

$$\begin{aligned} I_2\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.14)

Substituting (3.13) and (3.14) into (3.10), one has

$$\begin{aligned}&\frac{d}{dt}\int _\Omega \left( u_1\log u_1+u_2\log u_2\right) +\int _\Omega \left( u_1\log u_1+ u_2\log u_2\right) +d_1\int _\Omega \frac{|\nabla u_1|^2}{u_1}\nonumber \\&\qquad +d_2\int _\Omega \frac{|\nabla u_2|^2}{u_2} +\frac{\varepsilon _1}{2} \Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+ \Vert \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+\frac{\varepsilon _2}{2} \Vert \nabla \textbf{w}_2\Vert _{L^{2}(\Omega )}^2\nonumber \\&\qquad + \Vert \textbf{w}_2\Vert _{L^{2}(\Omega )}^2 +\Vert u_1\Vert _{L^{2}(\Omega )}^2+\Vert u_2\Vert _{L^{2}(\Omega )}^2+\frac{1}{2}\int _\Omega {u_1^2\log u_1+ \frac{1}{2}\int _\Omega u_2^2\log u_2}\nonumber \\&\quad \le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.15)

Finally, (3.4) is achieved by applying Grönwall’s inequality to (3.15), and (3.5) is obtained by (3.4) and an integration of (3.15) with respect to t. \(\square \)

With (3.5), we can obtain the following uniform-in-time estimates of \(\Vert u_1\Vert _{L^{q}(\Omega )}\) and \(\Vert u_2\Vert _{L^{q}(\Omega )}\) for \(q\in (1,\infty )\).

Lemma 3.4

Suppose that the conditions in Theorem 1.1 hold. For any \(1<q<\infty \), there exists a positive constant C(q) independent of t such that

$$\begin{aligned} \Vert u_1\Vert _{L^{q}(\Omega )}^q+\Vert u_2\Vert _{L^{q}(\Omega )}^q\le C(q)\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.16)

Proof

Multiplying the first equation of (1.4) by \(q u_1^{q-1}\) for \(q>1\) and integrating the resulting equation by parts subject to the boundary condition \(\nabla u_{1} \cdot \textbf{n}\mid _{\partial \Omega }=\textbf{w}_1\cdot \textbf{n}\mid _{\partial \Omega }=0\), we have

$$\begin{aligned}&\frac{d}{dt}\Vert u_1\Vert _{L^{q}(\Omega )}^q+\frac{4(q-1)d_1}{q}\Vert \nabla u_1^\frac{q}{2}\Vert _{L^{2}(\Omega )}^2\nonumber \\&\quad =(q-1)\chi _1\int _\Omega \textbf{w}_1\cdot \nabla u_1^q+ q\int _\Omega u_1^q(\gamma _1-u_1-cu_2)\nonumber \\&\quad = -(q-1)\chi _1\int _\Omega u_1^q\nabla \cdot \textbf{w}_1+q\int _\Omega u_1^q(\gamma _1-u_1)-c q\int _\Omega u_2u_1^q\nonumber \\&\quad =: I_3+I_4-c q\int _\Omega u_2u_1^q\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.17)

By Hölder’s inequality we have

$$\begin{aligned} I_3\le (q-1)\chi _1\Vert u_1^q\Vert _{L^{2}(\Omega )}\Vert \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}\quad \text {for all }t\in (0,T_{max}), \end{aligned}$$
(3.18)

where it follows from the Gagliardo–Nirenberg inequality that

$$\begin{aligned} \Vert u_1^q\Vert _{L^{2}(\Omega )} = \Vert u_1^{\frac{q}{2}}\Vert _{L^{4}(\Omega )}^2 \le C(\Vert \nabla u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}^{\frac{1}{2}}\Vert u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}^{\frac{1}{2}}+\Vert u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )})^2. \end{aligned}$$
(3.19)

Substituting (3.19) into (3.18), and using Young’s inequality and Hölder’s inequality, for all \(t\in (0,T_{max})\), one has

$$\begin{aligned} I_3&\le C(q)\left( \Vert \nabla u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )} \Vert u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}+\Vert u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}^2\right) \Vert \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}\nonumber \\&\le \frac{2(q-1)d_1}{q}\Vert \nabla u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}^2+C(q)\Vert u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}^2\Vert \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\nonumber \\&\quad +C(q)\Vert u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}^2\Vert \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}\nonumber \\&\le \frac{2(q-1)d_1}{q}\Vert \nabla u_1^{\frac{q}{2}}\Vert _{L^{2}(\Omega )}^2+C(q)\Vert u_1\Vert _{L^{q}(\Omega )}^q\left( \Vert \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+1\right) . \end{aligned}$$
(3.20)

Using Hölder’s inequality: \(\Vert u_1\Vert _{L^{q}(\Omega )}\le |\Omega |^{\frac{1}{q(q+1)}}\Vert u_1\Vert _{L^{q+1}(\Omega )}\) and Young’s inequality: \(q\gamma _1 u_1^q\le (q-1)u_1^{q+1}+\gamma _1^{q+1}C(q)\), we have

$$\begin{aligned} I_4&\le -\Vert u_1\Vert _{L^{q+1}(\Omega )}^{q+1}+\gamma _1^{q+1}C(q)|\Omega | \nonumber \\&\le -|\Omega |^{-\frac{1}{q}}\Vert u_1\Vert _{L^{q}(\Omega )}^{q+1}+C(q)\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.21)

Then the combination of (3.17), (3.20) and (3.21) shows that

$$\begin{aligned} \frac{d}{dt}\Vert u_1\Vert _{L^{q}(\Omega )}^q-C(q)\Vert u_1\Vert _{L^{q}(\Omega )}^q\left( \Vert \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+1\right) +|\Omega |^{-\frac{1}{q}}\Vert u_1\Vert _{L^{q}(\Omega )}^{q+1}\le 0 \end{aligned}$$
(3.22)

for all \(t\in (0,T_{max})\). Using (3.5), for all \(\tau \in (0,T_{max})\) and \(t\in (0,T_{max}-\tau )\), we have

$$\begin{aligned} \int _t^{t+\tau } \Vert \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\le \int _t^{t+\tau } \Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\le k\tau +C. \end{aligned}$$
(3.23)

With (3.23), one can apply a nonlinear Gronwall’s inequality shown in [22, Lemma 2.3] to (3.22) to obtain

$$\begin{aligned} \sup _{t\in (0,T_{max})}\Vert u_1\Vert _{L^{q}(\Omega )}^q\le C(q). \end{aligned}$$
(3.24)

Performing the same procedures as \(u_1\), one can deduce the following estimate for \(u_2\) as

$$\begin{aligned} \sup _{t\in (0,T_{max})}\Vert u_2\Vert _{L^{q}(\Omega )}^q\le C(q), \end{aligned}$$

which along with (3.24) completes the proof. \(\square \)

The following uniform-in-time estimates of \(\textbf{w}_1\) and \(\textbf{w}_2\) in \({W^{1,2}(\Omega )}\) can be easily obtained.

Lemma 3.5

Suppose that the conditions in Theorem 1.1 hold. Then there exists a positive constant C independent of t such that

$$\begin{aligned} \Vert \textbf{w}_1\Vert _{{W^{1,2}(\Omega )}}+\Vert \textbf{w}_2\Vert _{{W^{1,2}(\Omega )}}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.25)

Proof

In view of (3.7) and (3.16), we get \(\Vert \textbf{w}_1\Vert _{{W^{1,2}(\Omega )}}\le C\) for all \(t\in (0,T_{max})\). The estimate for \(\textbf{w}_2\) follows similarly. \(\square \)

Now we are in a position to derive the \(L^\infty \)-boundedness of \(u_1\) and \(u_2\) by the \(L^p\)-\(L^q\) estimates of the Neumann heat semigroup.

Lemma 3.6

Suppose that the conditions in Theorem 1.1 hold. Then there exists a positive constant C independent of t such that

$$\begin{aligned} \Vert u_1(\cdot , t)\Vert _{L^{\infty }(\Omega )}+\Vert u_2(\cdot , t)\Vert _{L^{\infty }(\Omega )}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.26)

Proof

It follows from (3.16), (3.25) and the Sobolev embedding \(W^{1,2}(\Omega )\hookrightarrow {L^{6}(\Omega )}\) that

$$\begin{aligned} \Vert u_1\textbf{w}_1\Vert _{L^{4}(\Omega )}\le \Vert u_1\Vert _{L^{12}(\Omega )}\Vert \textbf{w}_1\Vert _{L^{6}(\Omega )}\le C\quad \text {for all }t\in (0,T_{max}), \end{aligned}$$
(3.27)

where Hölder’s inequality was used. Now given \(t\in (0,T_{max})\), we let \(t_0:=(t-1)_+\). Applying the variation-of-constants formula, using \(u_1\), \(u_2\ge 0\) and the comparison principle, for all \(t\in (0,T_{max})\), one has

$$\begin{aligned} u_1(\cdot , t)&\le e^{(t-t_0) d_1 \Delta } u_{1}(\cdot ,t_0)-\chi _1\int _{t_0}^{t} e^{(t-s) d_1\Delta }\nabla \cdot \left( u_{1}(\cdot ,s) \textbf{w}_{1}(\cdot ,s)\right) d s\\&\quad +\gamma _1\int _{t_0}^{t} e^{(t-s) d_1\Delta }u_1(\cdot ,s) d s, \end{aligned}$$

which implies

$$\begin{aligned} \Vert u_1(\cdot , t)\Vert _{L^{\infty }(\Omega )}&\le \Vert e^{(t-t_0) d_1 \Delta } u_{1}(\cdot ,t_0)\Vert _{L^{\infty }(\Omega )}\nonumber \\&\quad +\chi _1\int _{t_0}^{t}\Vert e^{(t-s) d_1\Delta }\nabla \cdot \left( u_{1}(\cdot ,s) \textbf{w}_{1}(\cdot ,s)\right) \Vert _{L^{\infty }(\Omega )}d s\nonumber \\&\quad +\gamma _1\int _{t_0}^{t}\Vert e^{(t-s) d_1\Delta }u_1(\cdot ,s) \Vert _{L^{\infty }(\Omega )}d s\nonumber \\&=: I_5+I_6+I_7\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.28)

It follows from the well-known Neumann heat semigroup (cf. [32, Lemma 2.2], see also [46, formula (1.8)]) that

$$\begin{aligned} \Vert e^{td_1 \Delta } u_{1}(\cdot ,t)\Vert _{L^{\infty }(\Omega )} \le \frac{C}{t} \Vert u_{1}(\cdot ,t)\Vert _{L^{1}(\Omega )} \quad \text{ for } \text{ all } t\in (0,2)\cap (0,T_{max}). \end{aligned}$$

For \(t\ge 1\), \(t_0=t-1\) and hence we have along with (3.2)

$$\begin{aligned} I_5&=\Vert e^{d_1 \Delta } u_{1}(\cdot ,t-1)\Vert _{L^{\infty }(\Omega )}\le C\Vert u_{1}(\cdot ,t-1)\Vert _{L^{1}(\Omega )}\nonumber \\&\le C\quad \text {for all }t\in [1,\infty )\cap (0,T_{max}). \end{aligned}$$
(3.29)

Since \(u_{10}\in W^{1, p}(\Omega )\hookrightarrow {L^{\infty }(\Omega )}\) due to \(p>2\), it follows from the parabolic maximum principle that

$$\begin{aligned} I_5= \Vert e^{t d_1 \Delta } u_{10}\Vert _{L^{\infty }(\Omega )}\le \Vert u_{10}\Vert _{L^{\infty }(\Omega )}\le C\quad \text {for all }t\in (0,1)\cap (0,T_{max}). \end{aligned}$$
(3.30)

For \(I_6\), using (3.27) and the smoothing property of the Neumann heat semigroup [46, Lemma 1.3 (ii)], we get

$$\begin{aligned} I_6&\le C \int _{t_0}^{t} (1+(t-s)^{-\frac{3}{4}})e^{-\lambda _1(t-s)}\Vert u_{1}(\cdot ,s) \textbf{w}_{1}(\cdot ,s)\Vert _{L^{4}(\Omega )} d s\nonumber \\&\le C\quad \text {for all }t\in (0,T_{max}), \end{aligned}$$
(3.31)

where \(\lambda _1\) denotes the smallest positive eigenvalue of \(-\Delta \) in \(\Omega \). Now it remains to estimate the term \(I_7\). Let \(\overline{u}_1(s):=\frac{1}{|\Omega |}\int _\Omega u_1(x,t)\). Then using \(t-t_0\le 1\), the smoothing property of the Neumann heat semigroup [46, Lemma 1.3 (i)], (3.2) and Lemma 3.4 with \(q=2\), we obtain

$$\begin{aligned} I_7&\le \gamma _1\int _{t_0}^{t}\Vert e^{(t-s) d_1\Delta }(u_1(\cdot ,s)-\overline{u}_1(s)) \Vert _{L^{\infty }(\Omega )}d s+\gamma _1\int _{t_0}^{t}\Vert e^{(t-s) d_1\Delta }\overline{u}_1(s) \Vert _{L^{\infty }(\Omega )}d s\nonumber \\&\le C \int _{t_0}^{t} (1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\Vert u_1(\cdot ,s)-\overline{u}_1(s)\Vert _{L^{2}(\Omega )} d s+C\int _{t_0}^{t}\Vert \overline{u}_1(s)\Vert _{L^{\infty }(\Omega )}\nonumber \\&\le C\int _{t_0}^{t} (1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\left( \Vert u_1(\cdot ,s)\Vert _{L^{2}(\Omega )}+\Vert \overline{u}_1(s)\Vert _{L^{2}(\Omega )}\right) +C(t-t_0)\nonumber \\&\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.32)

Now the combination of (3.28)–(3.32) yields

$$\begin{aligned} \Vert u_1(\cdot , t)\Vert _{L^{\infty }(\Omega )}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$

Similar arguments applied to \(u_2\) give

$$\begin{aligned} \Vert u_2(\cdot , t)\Vert _{L^{\infty }(\Omega )}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$

The proof is completed. \(\square \)

We next deduce the uniform-in-time estimates for \(\nabla u_1\) and \(\nabla u_2\) in \({L^{2}(\Omega )}\).

Lemma 3.7

Suppose that the conditions in Theorem 1.1 hold. Then there exists a positive constant C independent of t such that

$$\begin{aligned} \Vert \nabla u_1(\cdot , t)\Vert _{L^{2}(\Omega )}+\Vert \nabla u_2(\cdot , t)\Vert _{L^{2}(\Omega )}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.33)

Proof

Multiplying the first equation of (1.4) by \(-\Delta u_1\) and integrating the resulting equation by parts along with the boundary conditions in (1.4), one has

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2=&-d_1\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2+\chi _1\int _\Omega \Delta u_1\nabla \cdot ( u_1 \textbf{w}_1)\nonumber \\&-\int _\Omega u_{1}E_1(u_1,u_2)\Delta u_1 \end{aligned}$$
(3.34)

for all \(t\in (0,T_{max})\). With the boundary condition \(\nabla u_{1} \cdot \textbf{n}\mid _{\partial \Omega }=0\) and (3.26), we further have

$$\begin{aligned} -\int _\Omega u_{1}E_1(u_1,u_2) \Delta u_1&= \int _\Omega \left( \gamma _1|\nabla u_{1}|^2-2u_1|\nabla u_{1}|^2-c(u_1\nabla u_2+u_{2}\nabla u_{1})\nabla u_{1}\right) \nonumber \\&\le C_1(\Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2)\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.35)

Clearly, for all \(t\in (0,T_{max})\), it holds that

$$\begin{aligned} \int _\Omega \Delta u_1\nabla \cdot ( u_1 \textbf{w}_1)&= \int _\Omega u_1 (\nabla \cdot \textbf{w}_1)\Delta u_1+\int _\Omega (\textbf{w}_1\cdot \nabla u_1)\Delta u_1=:I_8+I_9. \end{aligned}$$
(3.36)

Making use of (3.25) and (3.26), for all \(t\in (0,T_{max})\), we obtain

$$\begin{aligned} I_8\le \frac{d_1}{4\chi _1}\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2+\frac{\chi _1}{d_1}\Vert u_1 \nabla \cdot \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\le \frac{d_1}{4\chi _1}\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2+C_2. \end{aligned}$$
(3.37)

Noticing that \(\nabla |\nabla z|^{2}=2 D^{2} z \cdot \nabla z\) for \(z \in C^{2}(\bar{\Omega })\) with \(D^2z=\nabla (\nabla z)\), we arrive at

$$\begin{aligned} I_9&=-\int _\Omega \left( \textbf{w}_1\cdot D^2u_1+\nabla \textbf{w}_1\cdot \nabla u_1\right) \cdot \nabla u_1\nonumber \\&= -\int _\Omega \textbf{w}_1\cdot (\nabla u_1\cdot D^2u_1)-\int _\Omega \left( \nabla \textbf{w}_1\cdot \nabla u_1\right) \cdot \nabla u_1\nonumber \\&= -\frac{1}{2}\int _\Omega \textbf{w}_1\cdot \nabla |\nabla u_1|^2-\int _\Omega \left( \nabla \textbf{w}_1\cdot \nabla u_1\right) \cdot \nabla u_1\nonumber \\&= \frac{1}{2}\int _\Omega (\nabla \cdot \textbf{w}_1)|\nabla u_1|^2-\int _\Omega \left( \nabla \textbf{w}_1\cdot \nabla u_1\right) \cdot \nabla u_1\nonumber \\&\le \frac{3}{2}\Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}\Vert \nabla u_1\Vert _{L^{4}(\Omega )}^2\nonumber \\&\le C_3\Vert \nabla u_1\Vert _{L^{4}(\Omega )}^2 \quad \text {for all }t\in (0,T_{max}), \end{aligned}$$
(3.38)

where Hölder’s inequality and Lemma 3.5 have been used. Applying Lemma 2.4 (with \(q=4\), \(r=n=2\) and \(\theta =\frac{3}{4}\)) to \(\Vert \nabla u_1\Vert _{L^{4}(\Omega )}\) and using (3.26), (3.38) and Young’s inequality, for all \(t\in (0,T_{max})\), we have

$$\begin{aligned} I_9\le C_4(\Vert \Delta u_1\Vert _{L^{2}}^\frac{3}{2}\Vert u_1\Vert _{{L^{2}(\Omega )}}^{\frac{1}{2}}+\Vert u_1\Vert _{{{L^{2}(\Omega )}}}^2)\le \frac{d_1}{4\chi _1}\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2+C_5. \end{aligned}$$
(3.39)

The substitution of (3.37) and (3.39) into (3.36) yields

$$\begin{aligned} \chi _1\int _\Omega \Delta u_1\nabla \cdot ( u_1 \textbf{w}_1)\le \frac{d_1}{2}\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2+(C_2+C_5)\chi _1\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.40)

Using (3.26), (3.34), (3.35) and (3.40), for all \(t\in (0,T_{max})\), one can see that

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\frac{d_1}{2}\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2\le C_6(\Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2+1). \end{aligned}$$
(3.41)

Similar procedures applied to \(u_2\) yield that

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2+\frac{d_2}{2}\Vert \Delta u_2\Vert _{L^{2}(\Omega )}^2\le C_{7}(\Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2+1) \end{aligned}$$

for all \(t\in (0,T_{max})\). This along with (3.41) shows that

$$\begin{aligned}&\frac{1}{2}\frac{d}{dt}\left( \Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2\right) +{\frac{d_1}{2}\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2+\frac{d_2}{2}\Vert \Delta u_2\Vert _{L^{2}(\Omega )}^2}\nonumber \\&\quad \le C_{8}\left( \Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2+1\right) \quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.42)

Applying Lemma 2.4 (with \(q=r=n=2\) and \(\theta =\frac{1}{2}\)) to \(\Vert \nabla u_i\Vert _{{L^{2}(\Omega )}}\) for \(i=1,2\) and using (3.26) and Young’s inequality, for all \(t\in (0,T_{max})\), we obtain

$$\begin{aligned} \left( C_{8}+\frac{1}{2}\right) \Vert \nabla u_i\Vert _{{L^{2}(\Omega )}}^2&\le C_{9}(\Vert \Delta u_i\Vert _{L^{2}}^\frac{1}{2}\Vert u_i\Vert _{{L^{2}(\Omega )}}^{\frac{1}{2}}+\Vert u_i\Vert _{{{L^{2}(\Omega )}}})^2\\&\le \frac{d_i}{4}\Vert \Delta u_i\Vert _{{L^{2}(\Omega )}}^2+C_{10}. \end{aligned}$$

This together with (3.42) shows that

$$\begin{aligned}&\frac{1}{2}\frac{d}{dt}(\Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2)+\frac{1}{2}(\Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2)\nonumber \\&\qquad +{\frac{d_1}{4}\Vert \Delta u_1\Vert _{L^{2}(\Omega )}^2+\frac{d_2}{4}\Vert \Delta u_2\Vert _{L^{2}(\Omega )}^2}\nonumber \\&\quad \le C_{11}\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.43)

Applying Grönwall’s inequality to (3.43) leads to

$$\begin{aligned} \Vert \nabla u_1\Vert _{L^{2}(\Omega )}^2+\Vert \nabla u_2\Vert _{L^{2}(\Omega )}^2\le C_{12}\quad \text {for all }t\in (0,T_{max}), \end{aligned}$$

which proves (3.33). \(\square \)

With (3.33), we can use Lemma 2.2 to derive the uniform-in-time estimates of \(\textbf{w}_1\) and \(\textbf{w}_2\) in \(H^2(\Omega )\) (Lemma 3.8), which along with the \(L^p\)-\(L^q\)-estimates of the Neumann heat semigroup enables us to further obtain the boundedness of \(\nabla u_1\) and \(\nabla u_2\) as stated in Lemma 3.9.

Lemma 3.8

Suppose that the conditions in Theorem 1.1 hold. Then there exists a positive constant C independent of t such that

$$\begin{aligned} \Vert \textbf{w}_1\Vert _{H^2(\Omega )}+\Vert \textbf{w}_2\Vert _{H^2(\Omega )}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.44)

Proof

From Lemma 2.2 and (3.33), (3.44) follows immediately. \(\square \)

Lemma 3.9

Suppose that the conditions in Theorem 1.1 hold. Then there exists a positive constant C independent of t such that

$$\begin{aligned} \Vert u_1(\cdot , t)\Vert _{{W^{1,p}(\Omega )}}+\Vert u_2(\cdot , t)\Vert _{{W^{1,p}(\Omega )}}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.45)

Proof

Applying the variation-of-constants formula, one has

$$\begin{aligned} u_1(\cdot , t)=e^{t d_1(\Delta -1)} u_{10}+\int _{0}^{t} e^{(t-s) d_1(\Delta -1)}\varphi (u_1,u_2,\textbf{w}_1) (\cdot ,s)d s, \end{aligned}$$

where

$$\begin{aligned} \varphi (u_1,u_2,\textbf{w}_1):={-\nabla \cdot \left( \chi _1 u_{1} \textbf{w}_{1}\right) + u_{1}\left( \gamma _1+d_1-u_{1}-c u_{2}\right) }. \end{aligned}$$

It follows from (3.26), (3.33), (3.44) and the Sobolev embedding \(H^{2}(\Omega )\hookrightarrow {L^{\infty }(\Omega )}\) that

$$\begin{aligned} \Vert -\nabla \cdot (\chi u_{1} \textbf{w}_{1})\Vert _{L^{2}(\Omega )}&\le C\left( \Vert \nabla u_1\Vert _{L^{2}(\Omega )}+\Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}\right) \\&\le C\quad \text {for all }t\in (0,T_{max}), \end{aligned}$$

which along with (3.26) shows that

$$\begin{aligned} \Vert \varphi (u_1,u_2,\textbf{w}_1)\Vert _{L^{2}(\Omega )}\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$

Now using the smoothing property of the Neumann heat semigroup [46, Lemma 1.3 (ii)] again, we obtain

$$\begin{aligned} \Vert \nabla u_1(\cdot , t)\Vert _{L^{p}(\Omega )}&\le C \left\| u_{10}\right\| _{{W^{1,p}(\Omega )}}\nonumber \\&\quad +C\int _0^t\left( 1+(t-s)^{\frac{1}{p}-1}\right) e^{-\lambda _1(t-s)}\nonumber \\&\quad \times \Vert \varphi (u_1,u_2,\textbf{w}_1)(\cdot ,s)\Vert _{L^{2}(\Omega )} ds\nonumber \\&\le C\quad \text {for all }t\in (0,T_{max}). \end{aligned}$$
(3.46)

In a similar manner, we have

$$\begin{aligned} \left\| \nabla u_2(\cdot , t)\right\| _{{W^{1,p}(\Omega )}} \le C\quad \text {for all }t\in (0,T_{max}), \end{aligned}$$

which along with (3.26) and (3.46) completes the proof. \(\square \)

Proof of Theorem 1.1

\(T_{max}=\infty \) is a direct consequence (2.15) and (3.45). By (3.45) and Lemma 2.2, one can obtain (1.7) directly. \(\square \)

4 Global Stability

In this section, we shall investigate the asymptotic behavior of solutions to the system (1.4) and prove Theorem 1.2 by the Lyapunov functional method alongside compactness arguments. To begin with, we derive the following higher-order estimates of solutions when time t is away from 0.

Lemma 4.1

Suppose that the conditions in Theorem 1.1 hold. Then for any \(\theta \in (0,1)\), there exists a positive constant \(C(\theta )\) such that

$$\begin{aligned} \sum _{i=1}^2\left( \Vert u_i\Vert _{C^{2+\theta , 1+\frac{\theta }{2}}{\left( \bar{\Omega } \times [1,\infty )\right) }}+\Vert \textbf{w}_i\Vert _{C^{2+\theta , 1+\frac{\theta }{2}}{\left( \bar{\Omega } \times [1,\infty )\right) }}\right) \le C(\theta ). \end{aligned}$$
(4.1)

Proof

It follows from (1.7) that

$$\begin{aligned} \Vert u_1(\cdot , t)\Vert _{{W^{1,p}(\Omega )}}+\Vert u_2(\cdot , t)\Vert _{{W^{1,p}(\Omega )}}\le C\quad \text {for all }t>0. \end{aligned}$$

Taking \(t_0=\frac{1}{8}\) as the initial time, then \(u_1(\cdot ,t_0),u_2(\cdot ,t_0)\in {W^{1,p}(\Omega )}\). Using a similar argument as in the proof of Lemma 3.9, for any \(q\in (1,\infty )\), one can find a positive constant C(q) such that

$$\begin{aligned} \Vert u_1(\cdot , t)\Vert _{{W^{1,q}(\Omega )}}+\Vert u_2(\cdot , t)\Vert _{{W^{1,q}(\Omega )}}\le C(q)\quad \text {for all }t>t_0. \end{aligned}$$
(4.2)

Then using Lemma 2.2 and (4.2) one has

$$\begin{aligned} \Vert \textbf{w}_1(\cdot , t)\Vert _{{W^{2,q}(\Omega )}}+\Vert \textbf{w}_2(\cdot , t)\Vert _{{W^{2,q}(\Omega )}}\le C(q)\quad \text {for all }t>t_0. \end{aligned}$$
(4.3)

For any \(\theta \in (0,1)\), using (4.3) and the Sobolev embedding \(W^{2,r}(\Omega )\hookrightarrow C^{1,1-\frac{2}{r}}(\bar{\Omega })\hookrightarrow C^{1,\theta }(\bar{\Omega })\) for \(r>\frac{2}{1-\theta }\), we can find some \(r_0>\frac{2}{1-\theta }\) and \(C(\theta )>0\) such that

$$\begin{aligned}&\left\| \textbf{w}_1(\cdot , t)\right\| _{C^{1,\theta }(\bar{\Omega })}+\left\| \textbf{w}_2(\cdot , t)\right\| _{C^{1,\theta }(\bar{\Omega })}\nonumber \\&\quad \le C \left\| \textbf{w}_1(\cdot , t)\right\| _{W^{2,r_0}(\Omega )}+\left\| \textbf{w}_2(\cdot , t)\right\| _{W^{2,r_0}(\Omega )} \le C(\theta ) \end{aligned}$$
(4.4)

for all \(t>t_0\). From (1.4), we know that \(u_1\) satisfies

$$\begin{aligned} \left\{ \begin{array}{lll} \partial _{t} u_{1}-d_{1} \Delta u_{1}+\chi _1 \textbf{w}_{1}\cdot \nabla u_1+Q_1(x,t)u_{1}=0, &{} x \in \Omega ,&{}t>t_0, \\ \frac{\partial {u_1}}{\partial \textbf{n}}=0, &{} x \in \partial \Omega ,&{}t >t_0, \\ u_1(x, t)\mid _{t=t_0}=u_1(x,t_0), &{} x \in \Omega , \end{array}\right. \end{aligned}$$
(4.5)

where \(Q_1(x,t)=\chi _1\nabla \cdot \textbf{w}_1-E_1(u_1,u_2)\) for \((x,t)\in \Omega \times (t_0,\infty )\). Using (4.2) and (4.4), one has

$$\begin{aligned} \left\| \chi _1\textbf{w}_1\right\| _{L^{\infty }\left( \Omega \times \left[ j+\frac{1}{4},j+2\right] \right) }+\left\| Q_1\right\| _{L^{\infty }\left( \Omega \times \left[ j+\frac{1}{4},j+2\right] \right) }\le C\quad \text {for all }j\ge 0. \end{aligned}$$
(4.6)

In view of (4.2) and (4.6), one can apply the interior \(L^{p}\) estimate [28, Theorems 7.22 and 7.35] to (4.5) to obtain

$$\begin{aligned} \Vert u_1\Vert _{W^{2,1}_q\left( \Omega \times \left[ j+\frac{1}{2},j+2\right] \right) }\le C(q)\quad \text {for all }j\ge 0, \end{aligned}$$

where \(W^{2,1}_q\left( \Omega \times \left[ t_1,t_2\right] \right) :=\left\{ u\mid Du,D^2 u,u_t\in L^q\left( \Omega \times \left[ t_1,t_2\right] \right) \right\} \) for \(t_2>t_1>0\). By taking q appropriately large and using the Sobolev embedding theorem we have

$$\begin{aligned} \Vert u_1\Vert _{C^{1+\theta , \frac{1+\theta }{2}}{\left( \bar{\Omega } \times \left[ j+\frac{1}{2}, j+2\right] \right) }} \le C(\theta )\quad \text {for all }j\ge 0. \end{aligned}$$
(4.7)

Similarly, it follows that

$$\begin{aligned} \Vert u_2\Vert _{C^{1+\theta , \frac{1+\theta }{2}}{\left( \bar{\Omega } \times \left[ j+\frac{1}{2}, j+2\right] \right) }} \le C(\theta )\quad \text {for all }j\ge 0. \end{aligned}$$
(4.8)

Using (4.4), (4.7) and (4.8), we obtain that the solution \((\textbf{w}_1,\textbf{w}_2)\) of the elliptic system (2.16) satisfies

$$\begin{aligned} \Vert \textbf{w}_1\Vert _{C^{1+\theta , \frac{1+\theta }{2}}{\left( \bar{\Omega } \times \left[ j+\frac{1}{2}, j+2\right] \right) }}+\Vert \textbf{w}_2\Vert _{C^{1+\theta , \frac{1+\theta }{2}}{\left( \bar{\Omega } \times \left[ j{+}\frac{1}{2}, j+2\right] \right) }} {\le } C(\theta ) \quad \text {for all }j\ge 0. \end{aligned}$$
(4.9)

Clearly, it follows from (4.7)–(4.9) that

$$\begin{aligned} \left\| \chi _1\textbf{w}_1\right\| _{C^{\theta , \frac{\theta }{2}}{\left( \bar{\Omega } \times \left[ j+\frac{1}{2}, j+2\right] \right) }} +\left\| Q_1\right\| _{C^{\theta , \frac{\theta }{2}}{\left( \bar{\Omega } \times \left[ j+\frac{1}{2}, j+2\right] \right) }} \le C(\theta )\quad \text {for all }j\ge 0. \end{aligned}$$

An application of the Schauder estimate to (4.5) shows that

$$\begin{aligned} \Vert u_1\Vert _{C^{2+\theta , 1+\frac{\theta }{2}}{\left( \bar{\Omega } \times \left[ j+1, j+2\right] \right) }} \le C(\theta )\quad \text {for all }j\ge 0. \end{aligned}$$
(4.10)

Similarly, we have

$$\begin{aligned} \Vert u_2\Vert _{C^{2+\theta , 1+\frac{\theta }{2}}{\left( \bar{\Omega } \times \left[ j+1, j+2\right] \right) }} \le C(\theta )\quad \text {for all }j\ge 0. \end{aligned}$$
(4.11)

In view of (4.10) and (4.11), we can apply Lemma 2.2 to (2.16) to obtain

$$\begin{aligned} \Vert \textbf{w}_1\Vert _{C^{2+\theta , 1+\frac{\theta }{2}}{\left( \bar{\Omega } \times \left[ j+1, j+2\right] \right) }} +\Vert \textbf{w}_2\Vert _{C^{2+\theta , 1+\frac{\theta }{2}}{\left( \bar{\Omega } \times \left[ j+1, j+2\right] \right) }}\le C(\theta ) \quad \text {for all }j\ge 0. \end{aligned}$$
(4.12)

Noting that the constant \(C(\theta )\) is independent of \(j\ge 0\), we get (4.1) directly from (4.10)-(4.12). \(\square \)

4.1 Weak competition: \(c<\frac{\gamma _1}{\gamma _2}<\frac{1}{b}\)

To proceed, we define the positive constants

$$\begin{aligned} \eta _0:=\frac{K_1-K_1^*}{2(2{u_1^*}+K_1)} \quad \text {and}\quad \quad \eta :=\frac{1-b c }{2 c^2}\left( 1-\eta _0\right) \left( \frac{K_2}{{u_2^*}}-f(b,c)\right) \end{aligned}$$
(4.13)

under the condition (1.14).

Lemma 4.2

Let \(\eta _0\) and \(\eta \) be defined by (4.13). If \(bc\in (0,1)\) and (1.14) holds, then the positive constants

$$\begin{aligned} \Gamma _1:=\frac{\eta _0}{2}f(b,c)+\left( 1-\frac{\eta _0}{2}\right) \frac{K_2}{{u_2^*}} \quad \text {and}\quad \Gamma _2:=\frac{1}{2}\min \left\{ {\Gamma _2}_*+\frac{K_1}{{u_1^*}}\eta ,{\Gamma _2}_*+\Gamma _2^*\right\} \end{aligned}$$

satisfy

$$\begin{aligned} b^2+\eta<\Gamma _2<\frac{K_1}{{u_1^*}}\eta ,\quad \psi _1(\Gamma _2)>0\quad \text {and}\quad \Gamma _1<\frac{K_2}{{u_2^*}}, \end{aligned}$$
(4.14)

where

$$\begin{aligned} {\Gamma _2}_*=\frac{2\left( \alpha _1-\sqrt{\alpha _1^2+\alpha _2 c^2}\right) }{c^2},\quad \Gamma _2^*:=\frac{2\left( \alpha _1+\sqrt{\alpha _1^2+\alpha _2 c^2}\right) }{c^2}, \end{aligned}$$

and

$$\begin{aligned} \psi _1(s):=-\frac{c^2}{4}s^2+\alpha _1 s+\alpha _2\quad \text {for }s>0, \end{aligned}$$

with

$$\begin{aligned} \alpha _1&:=\left( 1-\frac{b c }{2}\right) \Gamma _1-\eta c^2 -1 \quad \text {and}\\ \alpha _2&:=-\frac{1}{4} b^2 \Gamma _1^2- \left( b^2 +\eta \right) \Gamma _1+b^2 \left( c^2 \eta +1 \right) +c^2 \eta ^2+ \eta . \end{aligned}$$

Proof

This proof is straightforward and tedious, and we give the detailed proof in Appendix B. \(\square \)

Now we are in a position to derive the following result.

Lemma 4.3

Let \((u_1, u_2, \textbf{w}_1, \textbf{w}_2)\) be the global classical solution of (1.4) obtained in Theorem 1.1. Assume \(c<\frac{\gamma _1}{\gamma _2}<\frac{1}{b}\) and let \((u_{1}^*,u_{2}^*)\) be defined in (1.8). If (1.13) or (1.14) holds, then there exist positive constants \(\Gamma _1\) and \(\Gamma _2\) such that the energy functional

$$\begin{aligned} \mathcal E_1(t):=&\Gamma _2 \int _\Omega \left( u_1-u_1^*-{u_1^*}\ln \frac{u_1}{u_1^*}\right) \\&+\Gamma _1\int _\Omega \left( u_2-u_2^*-{u_2^*}\ln \frac{u_2}{u_2^*}\right) \quad \text {for all }t>0\end{aligned}$$

satisfies

$$\begin{aligned} \mathcal E_1(t)\ge 0\quad \text {for all }t\ge 0, \end{aligned}$$
(4.15)

and

$$\begin{aligned} \frac{d}{dt}\mathcal E_1(t)\le -\theta _1\mathcal F_1(t)\quad \text {for all }t>0 \end{aligned}$$
(4.16)

for a positive constant \(\theta _1\), where

$$\begin{aligned} \mathcal F_1(t):=\int _\Omega \left( u_1-{u_1^*}\right) ^2+\int _\Omega \left( u_2-{u_2^*}\right) ^2\quad \text {for all }t>0. \end{aligned}$$
(4.17)

Proof

By the symmetry of the equations satisfied by \(u_1\) and \(u_2\) in (1.4), we only need to prove the conclusion under the condition (1.14). In the rest of this proof, we let the positive constants \(\Gamma _1\) and \(\Gamma _2\) be given by Lemma 4.2.

We first prove (4.15). Define the function \(\psi (s):=s-u_1^*-u_1^*\ln \frac{s}{u_1^*}\) for \(s>0\), then \(\psi '(s)=1-\frac{{u_1^*}}{s}\) with \(\psi '({u_1^*})=0\) and \(\psi ''(s)=\frac{{u_1^*}}{s^2}>0\). Hence we have \(\psi (s)\ge \psi ({u_1^*})=0\) for \(s>0\), which implies \( u_1-u_1^*-u_1^*\ln \frac{u_1}{u_1^*}\ge 0. \) Similar arguments for \(u_2\) yield \( u_2-u_2^*-u_2^*\ln \frac{u_2}{u_2^*}\ge 0. \) Therefore, (4.15) is proved. It remains to prove (4.16). To this end, we multiply the first equation in (1.4) by \({1-\frac{u_1^*}{u_1}}\), integrate the resulting equation by parts along with the boundary conditions in (1.4) and use \(\gamma _1-{u_1^*}-c{u_2^*}=0\) to get

$$\begin{aligned}&\frac{d}{dt}\int _\Omega \left( u_1-{u_1^*}-{u_1^*}\ln \frac{u_1}{u_1^*}\right) \nonumber \\&\quad =\int _\Omega \left( 1-\frac{u_1^*}{u_1}\right) \left( d_{1} \Delta u_{1}-\nabla \cdot \left( \chi _1 u_{1} \textbf{w}_{1}\right) + u_{1}E_1(u_1,u_2) \right) \nonumber \\&\quad =-d_1 {u_1^*}\int _\Omega \frac{|\nabla u_1|^2}{u_1^2}+\chi _1{u_1^*}\int _\Omega {\textbf{w}_1\cdot \frac{\nabla {u_1}}{u_1}}\nonumber \\&\qquad +\int _\Omega ({u_1}-{u_1^*})\left( -u_{1}-c u_{2}+{u_1^*}+c{u_2^*}\right) \nonumber \\&\quad =-d_1 {u_1^*}\int _\Omega \frac{|\nabla u_1|^2}{u_1^2}+\chi _1{u_1^*}\int _\Omega {\textbf{w}_1\cdot \frac{\nabla {u_1}}{u_1}}\nonumber \\&\qquad -\int _\Omega ({u_1}-{u_1^*})^2-c\int _\Omega ({u_1}-{u_1^*})\left( u_{2}-{u_2^*}\right) \end{aligned}$$
(4.18)

for all \(t>0\). Similarly, it holds that

$$\begin{aligned} \frac{d}{dt}\int _\Omega \left( u_2-u_2^*-{u_2^*}\ln \frac{u_2}{u_2^*}\right)&=-d_2 {u_2^*}\int _\Omega \frac{|\nabla u_2|^2}{u_2^2}+\chi _2{u_2^*}\int _\Omega {\textbf{w}_2\cdot \frac{\nabla {u_2}}{u_2}}\nonumber \\&\quad -\int _\Omega ({u_2}-{u_2^*})^2-b\int _\Omega ({u_1}-{u_1^*})\left( u_{2}-{u_2^*}\right) \end{aligned}$$
(4.19)

for all \(t>0\). As deriving (3.7), by Young’s inequality we can obtain

$$\begin{aligned}&\varepsilon _1 \Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+ \Vert \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\\&\quad =\int _\Omega \nabla \left( u_1-u_1^*+c(u_2-u_2^*)\right) \cdot \textbf{w}_1=\int _\Omega \left( u_1-u_1^*+c(u_2-u_2^*)\right) \nabla \cdot \textbf{w}_1\\&\quad \le \varepsilon _1\Vert \nabla \textbf{w}_1\Vert _{L^{2}(\Omega )}^2+\frac{1}{4\varepsilon _1}\left( \Vert u_1-u_1^*\Vert _{L^{2}(\Omega )}^2+c^2\Vert u_2-u_2^*\Vert _{L^{2}(\Omega )}^2\right) , \end{aligned}$$

which implies

$$\begin{aligned} \Vert \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\le \frac{1}{4\varepsilon _1}\left( \Vert u_1-u_1^*\Vert _{L^{2}(\Omega )}^2+c^2\Vert u_2-u_2^*\Vert _{L^{2}(\Omega )}^2\right) \quad \text {for all }t>0. \end{aligned}$$
(4.20)

Similarly, we can obtain

$$\begin{aligned} \Vert \textbf{w}_2\Vert _{L^{2}(\Omega )}^2 \le \frac{1}{4\varepsilon _2}\left( b^2\Vert u_1-u_1^*\Vert _{L^{2}(\Omega )}^2+\Vert u_2-u_2^*\Vert _{L^{2}(\Omega )}^2\right) \quad \text {for all }t>0. \end{aligned}$$
(4.21)

For \(\eta \) given by (4.13), the combination of (4.18)–(4.21) shows that

$$\begin{aligned} \frac{d}{dt}\mathcal E_1(t)&= -d_1\Gamma _2 {u_1^*}\int _\Omega \frac{|\nabla u_1|^2}{u_1^2}-d_2\Gamma _1 {u_2^*}\int _\Omega \frac{|\nabla u_2|^2}{u_2^2}- \Gamma _2 \int _\Omega ({u_1}-{u_1^*})^2\nonumber \\&\quad - \Gamma _1\int _\Omega ({u_2}-{u_2^*})^2 +\chi _1\Gamma _2 {u_1^*}\int _\Omega {\textbf{w}_1\cdot \frac{\nabla {u_1}}{u_1}}+\chi _2\Gamma _1{u_2^*}\int _\Omega {\textbf{w}_2\cdot \frac{\nabla {u_2}}{u_2}}\nonumber \\&\quad -\left( c \Gamma _2+b\Gamma _1\right) \int _\Omega ({u_1}-{u_1^*})\left( u_{2}-{u_2^*}\right) \nonumber \\&\le -d_1\Gamma _2 {u_1^*}\int _\Omega \frac{|\nabla u_1|^2}{u_1^2}-d_2\Gamma _1 {u_2^*}\int _\Omega \frac{|\nabla u_2|^2}{u_2^2}\nonumber \\&\quad -4\varepsilon _2\Vert \textbf{w}_2\Vert _{L^{2}(\Omega )}^2-4\eta \varepsilon _1\Vert \textbf{w}_1\Vert _{L^{2}(\Omega )}^2\nonumber \\&\quad -\left( \Gamma _2 - b^2-\eta \right) \int _\Omega ({u_1}-{u_1^*})^2-\left( \Gamma _1-1-\eta c^2\right) \int _\Omega ({u_2}-{u_2^*})^2\nonumber \\&\quad +\chi _1\Gamma _2 {u_1^*}\int _\Omega {\textbf{w}_1\cdot \frac{\nabla {u_1}}{u_1}}+\chi _2\Gamma _1{u_2^*}\int _\Omega {\textbf{w}_2\cdot \frac{\nabla {u_2}}{u_2}}\nonumber \\&\quad -\left( c \Gamma _2+b\Gamma _1 \right) \int _\Omega ({u_1}-{u_1^*})\left( u_{2}-{u_2^*}\right) \nonumber \\&\le -\int _{\Omega } X_{1} A_{1} X_{1}^{T}-\int _{\Omega } Y_{1} B_{1} Y_{1}^{T}\quad \text {for all }t>0, \end{aligned}$$
(4.22)

where \(X_{1}:=\left( u_1-u_{1}^{*}, u_2-u_{2}^{*}\right) \), \(Y_{1}:=\left( \frac{\nabla u_1}{u_1}, \frac{\nabla u_2}{u_2}, \textbf{w}_1, \textbf{w}_2\right) \) and \(A_{1}, B_{1}\) are matrices denoted by

$$\begin{aligned}{} & {} A_{1}:=\left( \begin{array}{ll} \Gamma _2 - b^2-\eta &{}\quad \frac{c \Gamma _2+b\Gamma _1 }{2} \\ \frac{c \Gamma _2+b\Gamma _1 }{2} &{}\quad \Gamma _1-1 -\eta c^2 \end{array}\right) ,\\{} & {} B_{1}:=\left( \begin{array}{llll} d_1 \Gamma _2 {u_1^*} &{}\quad 0&{}\quad -\frac{\chi _1\Gamma _2 {u_1^*}}{2} &{}\quad 0\\ 0 &{}\quad d_2\Gamma _1 {u_2^*} &{}\quad 0 &{}\quad -\frac{\chi _2 \Gamma _1{u_2^*}}{2}\\ -\frac{\chi _1\Gamma _2 {u_1^*}}{2} &{}\quad 0&{}\quad 4\eta \varepsilon _1&{}\quad 0\\ 0&{}\quad -\frac{\chi _2 \Gamma _1{u_2^*}}{2}&{}\quad 0&{}\quad 4\varepsilon _2 \end{array}\right) . \end{aligned}$$

We next prove that the matrices \(A_1\) and \(B_1\) are both positive definite. Denoting the determinant of a general square matrix X by |X| and denote the upper left k-by-k (\(k=1,2,3\)) corner of \(B_1\) by \(B_1^{(k)}\), then by (4.14) we have \(|B_1^{(1)}|=d_1 \Gamma _2 {u_1^*}>0\), \(|B_1^{(2)}|=d_1 d_2 \Gamma _2\Gamma _1{u_1^*}{u_2^*}>0\),

$$\begin{aligned} |B_1^{(3)}|&=\left| \begin{array}{lll} d_1 \Gamma _2 {u_1^*} &{}\quad 0&{}\quad -\frac{\chi _1\Gamma _2 {u_1^*}}{2} \\ 0 &{}\quad d_2 \Gamma _1 {u_2^*} &{}\quad 0 \\ -\frac{\chi _1\Gamma _2 {u_1^*}}{2} &{}\quad 0&{}\quad 4\eta \varepsilon _1 \end{array}\right| \\&= \frac{1}{4} d_2 \Gamma _2\Gamma _1 \left( \chi _1u_1^*\right) ^2 u_2^* \left( \frac{K_1}{{u_1^*}}\eta -\Gamma _2 \right) >0, \end{aligned}$$

and

$$\begin{aligned} \left| B_1\right| =\frac{1}{16} \Gamma _2\Gamma _1 \left( \chi _1\chi _2 u_1^* u_2^*\right) ^2 \left( \frac{K_1}{{u_1^*}}\eta -\Gamma _2\right) \left( \frac{K_2}{{u_2^*}}-\Gamma _1\right) >0. \end{aligned}$$

Sylvester’s criterion thus entails that the matrix \(B_1\) is positive definite. For the matrix \(A_1\), we know from (4.14) that \(\Gamma _2 - b^2-\eta >0\) and

$$\begin{aligned} \left| A_{1}\right| =\left| \begin{array}{cc} \Gamma _2 - b^2-\eta &{} \frac{c \Gamma _2+b\Gamma _1 }{2} \\ \frac{c \Gamma _2+b\Gamma _1 }{2} &{} \Gamma _1-1 -\eta c^2 \end{array}\right| =\psi _1(\Gamma _2)>0, \end{aligned}$$

where the function \(\psi _1\) is defined in Lemma 4.2. Again, it follows from Sylvester’s criterion that the matrix \(A_1\) is positive definite. Therefore, we can find a positive constant \(\theta _1\) such that

$$\begin{aligned} X_{1} A_{1} X_{1}^{T} \ge \theta _1|X_{1}|^2\quad \text {and}\quad Y_{1} B_{1} Y_{1}^{T} \ge \theta _1|Y_{1}|^2 \quad \text {for all }t>0. \end{aligned}$$
(4.23)

The combination of (4.17), (4.22) and (4.23) proves (4.16). \(\square \)

With Lemmas 2.5, 4.1 and 4.3, we can use a similar argument as in the proof of [42, Lemma 3.4] to prove the following result.

Lemma 4.4

Under the conditions of Lemma 4.3, for any \(\theta \in (0,1)\), it holds that

$$\begin{aligned} \Vert u_1-{{u_1^*}}\Vert _{C^{2+\theta }(\bar{\Omega })}+\Vert u_2-{{u_2^*}}\Vert _{C^{2+\theta }(\bar{\Omega })}+\Vert \textbf{w}_1\Vert _{C^{2+\theta }(\bar{\Omega })}+\Vert \textbf{w}_2\Vert _{C^{2+\theta }(\bar{\Omega })} \rightarrow 0 \text{ as } t \rightarrow \infty . \end{aligned}$$

Proof

Let \(\mathcal E_{1}(t), \mathcal F_{1}(t)\) be given in Lemma 4.3. Clearly, \(\mathcal E_{1}(t)\) is bounded from below according to (4.15). Thanks to Lemma 4.1, it can be seen that \(\mathcal F_{1}(t) \in C^{\theta / 2}([1, \infty ))\) and \(\left\| \mathcal F_{1}\right\| _{C^{\theta / 2}([1, \infty ))} \le k\) in \([1, \infty )\) for some \(k>0\). In view of (4.16), we can apply Lemma 2.5 to deduce \(\lim _{t \rightarrow \infty }\mathcal F_{1}(t)=0\). That is

$$\begin{aligned} \lim _{t \rightarrow \infty }\left( \left\| u_1-{u_1^*}\right\| _{2}+\left\| u_2-{u_2^*}\right\| _{2}\right) =0. \end{aligned}$$

Take \(0<\theta<\theta ^{\prime }<1.\) According to Theorem 2.1, in the space \(C^{2+\theta ^{\prime }}(\bar{\Omega })\), \(u_1(\cdot , t)\) and \(u_2(\cdot , t)\) are bounded for \(t \ge 1.\) Using the compact arguments and the uniqueness of limits (cf. [42, (3.12)], see also [20, Remark 6.2]) we can show that

$$\begin{aligned} \Vert u_1-{{u_1^*}}\Vert _{C^{2+\theta }(\bar{\Omega })}+\Vert u_2-{{u_2^*}}\Vert _{C^{2+\theta }(\bar{\Omega })} \rightarrow 0 \text{ as } t \rightarrow \infty . \end{aligned}$$
(4.24)

Using (4.20), (4.21) and (4.24), one has

$$\begin{aligned} \Vert \textbf{w}_1\Vert _{W^{1,2}(\Omega )}+ \Vert \textbf{w}_2\Vert _{W^{1,2}(\Omega )}\rightarrow 0\quad \text {as }t\rightarrow \infty . \end{aligned}$$

This together with (4.1), the compact arguments and the uniqueness of limits shows that

$$\begin{aligned} \Vert \textbf{w}_1\Vert _{C^{1+\theta }(\bar{\Omega })}+\Vert \textbf{w}_2\Vert _{C^{1+\theta }(\bar{\Omega })}\rightarrow \infty \quad \text {as }t\rightarrow \infty , \end{aligned}$$

which along with (4.24) completes the proof. \(\square \)

We are now in a position to investigate the convergence rate.

Lemma 4.5

Under the conditions of Lemma 4.3. There exist two constants \(\sigma >0\) and \(C>0\) independent of t such that

$$\begin{aligned} \sum _{i=1}^2\left( \left\| u_i(\cdot , t)-u_{i}^{*}\right\| _{{W^{1,\infty }(\Omega )}}+\left\| \textbf{w}_i\right\| _{{W^{1,\infty }(\Omega )}}\right) \le C e^{-\sigma t} \end{aligned}$$
(4.25)

holds for all \(t>T_0\) with some \(T_0>1\), where \(u_{1}^*\) and \(u_{2}^*\) are given by (1.8).

Proof

With Lemmas 4.3 and 4.4, one can use a similar argument as in the proof of [44, Lemma 3.7] (where the \(L^\infty (\Omega )\) decay rate are obtained) to obtain that there exist positive constants \(\sigma _1\) and \(T_0\) such that

$$\begin{aligned} \left\| u_1(\cdot , t)-u_{1}^{*}\right\| _{{W^{1,\infty }(\Omega )}}+\left\| u_2(\cdot , t)-u_{2}^{*}\right\| _{{W^{1,\infty }(\Omega )}}\le C_1 e^{-\sigma _1 t}\quad \text {for all }t>T_0. \end{aligned}$$
(4.26)

For the convenience of readers, we shall sketch the proof of (4.26). In view of Lemma 4.4, we can apply L’Hôpital’s rule to derive that

$$\begin{aligned} \lim _{u_{i} \rightarrow u_{i}^{*}} \frac{u_i-u_{i}^{*}-u_{i}^{*} \ln \frac{u_i}{u_{i}^{*}}}{\left( u_i-u_{i}^{*}\right) ^{2}}=\lim _{u_{i} \rightarrow u_{i}^{*}} \frac{1-\frac{u_{i}^{*}}{u_i}}{2\left( u_i-u_{i}^{*}\right) }=\lim _{u_{i} \rightarrow u_{i}^{*}} \frac{1}{2 u_i}=\frac{1}{2 u_{i}^{*}},\quad i=1,2, \end{aligned}$$

which by the continuity yields a constant \(T_0>1\) such that

$$\begin{aligned} \frac{1}{4 u_{i}^{*}} \int _{\Omega }\left( u_i-u_{i}^{*}\right) ^{2} \le \int _{\Omega }\left( u_i-u_{i}^{*}-u_{i}^{*} \ln \frac{u}{u_{i}^{*}}\right) \le \frac{1}{u_{i}^{*}} \int _{\Omega }\left( u_i-u_{i}^{*}\right) ^{2},\quad i=1,2 \end{aligned}$$
(4.27)

for all \(t\ge T_0\). Then, it follows from the definition of \(\mathcal {E}_{1}(t)\) and \(\mathcal {F}_{1}(t)\) that

$$\begin{aligned} \mathcal {E}_{1}(t) \le C_{2} \mathcal {F}_{1}(t) \quad \text{ for } \text{ all } t\ge T_0, \end{aligned}$$

which together with (4.16) implies

$$\begin{aligned} \frac{d}{d t} \mathcal {E}_{1}(t)+\frac{\theta _{1}}{C_{2}} \mathcal E_{1}(t) \le 0 \quad \text{ for } \text{ all } t\ge T_0. \end{aligned}$$

Solving the above inequality, we obtain

$$\begin{aligned} \mathcal {E}_{1}(t) \le \mathcal {E}_{1}(T_0) e^{-\frac{\theta _{1}}{C_{2}} t} \quad \text{ for } \text{ all } t\ge T_0. \end{aligned}$$
(4.28)

By the definition of \(\mathcal {E}_{1}(t)\) and \(\mathcal {F}_{1}(t)\) and (4.27), one can find a constant \(C_{3}>0\) such that

$$\begin{aligned} \mathcal {F}_{1}(t) \le C_{3} \mathcal {E}_{1}(t) \quad \text{ for } \text{ all } t\ge T_0, \end{aligned}$$

which along with (4.28) shows that

$$\begin{aligned} \left\| u_1(\cdot , t)-u_{1}^{*}\right\| _{L^{2}(\Omega )}+\left\| u_2(\cdot , t)-u_{2}^{*}\right\| _{L^{2}(\Omega )}\le C_{4} e^{-\frac{\theta _{1}}{2 C_{2}} t} \quad \text {for all }t\ge T_0. \end{aligned}$$
(4.29)

The combination of (4.1), (4.29) and the Gagliardo–Nirenberg inequality

$$\begin{aligned} \left\| u_i-u_{i}^{*}\right\| _{{W^{1,\infty }(\Omega )}} \le C_5\left( \Vert D^2 u_i\Vert _{L^{4}}^{\frac{4}{5}}\left\| u_i-u_{i}^{*}\right\| _{L^{2}}^{\frac{1}{5}}+\left\| u_i-u_ {i}^{*}\right\| _{L^{2}}\right) \quad \text {for } i=1,2 \end{aligned}$$
(4.30)

yields (4.26) by choosing \(C_1\) large enough and taking \(\sigma _1=\frac{\theta _{1}}{10 C_{2}}\).

In view of (4.20) and (4.21), it follows from (4.26) immediately that

$$\begin{aligned} \left\| \textbf{w}_1\right\| _{{L^{2}(\Omega )}}+\left\| \textbf{w}_2\right\| _{{L^{2}(\Omega )}}\le C_6 e^{-\sigma _1 t}\quad \text {for all }t>T_0. \end{aligned}$$
(4.31)

With (4.1) and (4.31), we can use a similar argument as deriving (4.30) to show that

$$\begin{aligned} \left\| \textbf{w}_1\right\| _{{W^{1,\infty }(\Omega )}}+\left\| \textbf{w}_2\right\| _{{W^{1,\infty }(\Omega )}}\le C_7 e^{-\sigma _2 t}\quad \text {for all }t>T_0, \end{aligned}$$
(4.32)

where \(\sigma _2=\frac{\sigma _1}{5}\). Therefore, (4.25) is a direct consequence of (4.26) and (4.32) by taking C appropriately large and \(\sigma =\sigma _2\). The proof is completed. \(\square \)

4.2 Competitive exclusion: \(\frac{\gamma _1}{\gamma _2}<\min \{\frac{1}{b},c\}\)

As the results stated for the weak competition case in the above subsection, we have the following conclusions.

Lemma 4.6

Let \((u_1, u_2, \textbf{w}_1, \textbf{w}_2)\) be the global classical solution of (1.4) obtained in Theorem 1.1. Assume \(\frac{\gamma _1}{\gamma _2}<\min \{\frac{1}{b},c\}\) and (1.15) holds. Define two constants

$$\begin{aligned} \Gamma _3:=\frac{1}{2}\left( \frac{K_2}{\gamma _2}+ f\left( b,\frac{\gamma _1}{\gamma _2}\right) \right) \quad \text {and}\quad \Gamma _4:=\frac{\Gamma _3\left( 2\gamma _2^2-b\gamma _1\gamma _2\right) -2\gamma _2^2}{\gamma _1^2}, \end{aligned}$$
(4.33)

where the function f is given by (1.9). Then

$$\begin{aligned} \frac{K_2}{\gamma _2}>\Gamma _3>f\left( b,\frac{\gamma _1}{\gamma _2}\right)>1\quad \text {and}\quad \Gamma _4>b^2. \end{aligned}$$
(4.34)

Moreover, the energy functional

$$\begin{aligned} \mathcal E_2(t):=\Gamma _4\int _\Omega u_1+\Gamma _3\int _\Omega \left( u_2-\gamma _2-\gamma _2\ln \frac{ u_2}{\gamma _2}\right) \quad \text {for all }t>0 \end{aligned}$$
(4.35)

satisfies \(\mathcal E_2(t)\ge 0\) for all \(t>0\), and

$$\begin{aligned} \frac{d}{dt}\mathcal E_2(t)\le -\theta _2\mathcal F_2(t)\quad \text {for all }t>0 \end{aligned}$$
(4.36)

for a positive constant \(\theta _2\), where

$$\begin{aligned} \mathcal F_2(t):=\int _\Omega u_1^2+\int _\Omega \left( u_2-\gamma _2\right) ^2\quad \text {for all }t>0. \end{aligned}$$
(4.37)

Proof

First of all, by a similar argument as proving (4.15), we can obtain \(\mathcal E_2(t)\ge 0\) for all \(t>0\). We then prove (4.34). Using \(\gamma _2-b\gamma _1>0\) and (1.15), we obtain the first inequality in (4.34) directly, moreover, we have

$$\begin{aligned} \Gamma _4-b^2&= \frac{1}{2\gamma _1^2\left( \gamma _2-b\gamma _1\right) }\left[ (2\gamma _2-b\gamma _1)(\gamma _2-b\gamma _1)K_2+b^3\gamma _1^3+3b\gamma _1\gamma _2^2-2\gamma _2^3\right] \\&\ge \frac{1}{2\gamma _1^2\left( \gamma _2-b\gamma _1\right) }\left[ (2\gamma _2-b\gamma _1)\left( \gamma _2^2+b^2\gamma _1^2\right) +b^3\gamma _1^3+3b\gamma _1\gamma _2^2-2\gamma _2^3\right] \\&= \frac{b\gamma _2(\gamma _2+b\gamma _1)}{\gamma _1\left( \gamma _2-b\gamma _1\right) }>0, \end{aligned}$$

which proves the second inequality in (4.34).

It remains to prove (4.36). Integrating the first two equations of (1.4) by parts with \(\nabla u_{1} \cdot \textbf{n} \mid _{\partial \Omega }=\textbf{w}_1 \cdot \textbf{n} \mid _{\partial \Omega }=0\) and using \(\frac{\gamma _1}{\gamma _2}<c\), we have

$$\begin{aligned} \frac{d}{dt}\int _\Omega u_1&=\int _\Omega u_{1}\left( \gamma _1-u_{1}-c u_{2}\right) \nonumber \\&\le \int _\Omega u_{1}\left( \gamma _1-u_{1}-\frac{\gamma _1}{\gamma _2} u_{2} \right) \nonumber \\&= -\int _\Omega u_1^2-\frac{\gamma _1}{\gamma _2}\int _\Omega u_1(u_{2}-\gamma _2)\quad \text {for all }t>0. \end{aligned}$$
(4.38)

As deriving (4.19), for all \(t>0\), we have

$$\begin{aligned}&\frac{d}{dt}\int _\Omega \left( u_2-\gamma _2-\gamma _2\ln \frac{ u_2}{\gamma _2}\right) \nonumber \\&\quad \le -d_2 \gamma _2\int _\Omega \frac{|\nabla u_2|^2}{u_2^2}+\chi _2\gamma _2\int _\Omega {\textbf{w}_2\cdot \frac{\nabla {u_2}}{u_2}}{-}\int _\Omega ({u_2}{-}{\gamma _2})^2-b\int _\Omega {u_1}\left( u_{2}-\gamma _2\right) . \end{aligned}$$
(4.39)

It follows from (4.21), (4.35), (4.38), (4.39) and Young’s inequality that

$$\begin{aligned} \frac{d}{dt}\mathcal E_2(t)&\le -\left( \Gamma _4 - b^2\right) \int _\Omega u_1^2-\left( \Gamma _3-1\right) \int _\Omega ({u_2}-{\gamma _2})^2-\left( \frac{\gamma _1}{\gamma _2}\Gamma _4+b\Gamma _3\right) \nonumber \\&\quad \times \int _\Omega {u_1}\left( u_2-\gamma _2\right) -d_2\gamma _2\Gamma _3 \int _\Omega \frac{|\nabla u_2|^2}{u_2^2}\nonumber \\&\quad +\chi _2\gamma _2\Gamma _3\int _\Omega {\textbf{w}_2\cdot \frac{\nabla {u_2}}{u_2}}-4 \varepsilon _2\Vert \textbf{w}_2\Vert _{L^{2}(\Omega )}^2\nonumber \\&\le -\int _{\Omega } X_{2} A_{2} X_{2}^{T}-\int _{\Omega } Y_{2} B_{2} Y_{2}^{T}\quad \text {for all }t>0, \end{aligned}$$
(4.40)

where \(X_{2}=\left( u_1, u_2-\gamma _2\right) \), \(Y_{2}=\left( \frac{\nabla u_2}{u_2},\textbf{w}_2\right) \) and \(A_{2}, B_{2}\) are matrices denoted by

$$\begin{aligned} A_{2}:=\left( \begin{array}{ll} \Gamma _4 - b^2 &{}\quad \frac{\frac{\gamma _1}{\gamma _2}\Gamma _4+b\Gamma _3 }{2} \\ \frac{\frac{\gamma _1}{\gamma _2}\Gamma _4+b\Gamma _3 }{2} &{}\quad \Gamma _3 -1 \end{array}\right) , \quad B_{2}:= \left( \begin{array}{ll} d_2 \gamma _2 \Gamma _3 &{}\quad -\frac{\chi _2 \gamma _2\Gamma _3}{2} \\ -\frac{\chi _2 \gamma _2\Gamma _3}{2} &{}\quad 4 \varepsilon _2 \end{array}\right) . \end{aligned}$$

Using \(\gamma _2-b\gamma _1>0\), (1.15), (4.33) and (4.34), we obtain \(\Gamma _4 - b^2 >0\) and

$$\begin{aligned} \left| A_{2}\right|&=\frac{\left[ (\gamma _2-b\gamma _1 )K_2+b^2\gamma _1^2+2b\gamma _1\gamma _2-\gamma _2^2 \right] \left[ (\gamma _2-b\gamma _1 )K_2-\left( \gamma _2^2+ b^2\gamma _1^2\right) \right] }{4\gamma _1^2\gamma _2(\gamma _2-b\gamma _1 )}\nonumber \\&=\frac{\left( (\gamma _2-b\gamma _1 )K_2+b^2\gamma _1^2+2b\gamma _1\gamma _2-\gamma _2^2 \right) }{4\gamma _1^2}\left( \frac{K_2}{\gamma _2}-f\left( b,\frac{\gamma _1}{\gamma _2}\right) \right) \nonumber \\&\ge \frac{b(b\gamma _1+\gamma _2 )}{2\gamma _1}\left( \frac{K_2}{\gamma _2}-f\left( b,\frac{\gamma _1}{\gamma _2}\right) \right) >0. \end{aligned}$$
(4.41)

Moreover, it is obvious that \(d_2 \gamma _2 \Gamma _3>0\), and by (4.34) we have

$$\begin{aligned} |B_2|=\frac{\Gamma _3\left( \gamma _2 \chi _2\right) ^2}{4}\left( \frac{K_2}{\gamma _2}-\Gamma _3\right) >0. \end{aligned}$$
(4.42)

In view of Sylvester’s criterion, it follows from (4.41) and (4.42) that the matrices \(A_2\) and \(B_2\) are positive definite, and hence we can find a positive constant \(\theta _2\) such that

$$\begin{aligned} X_{2} A_{2} X_2^{T} \ge \theta _2|X_2|^2\quad \text {and}\quad Y_2 B_2 Y_2^{T} \ge \theta _2|Y_2|^2 \quad \text {for all }t>0, \end{aligned}$$

which along with \(X_{2}=\left( u_1, u_2-\gamma _2\right) \), (4.37) and (4.40) proves (4.36). The proof is completed. \(\square \)

Lemma 4.7

Under the conditions of Lemma 4.6, there exists a constant \(T_1>0\) such that

$$\begin{aligned} \left\| u_1(\cdot , t)\right\| _{{W^{1,\infty }(\Omega )}}+\left\| u_2(\cdot , t)-\gamma _2\right\| _{{W^{1,\infty }(\Omega )}}+\sum _{i=1}^2\left\| \textbf{w}_i\right\| _{{W^{1,\infty }(\Omega )}}\le \frac{C}{1+t} \end{aligned}$$

for all \(t>T_1\), where C is a positive constant independent of t.

Proof

First, by a similar argument as in the proof of Lemma 4.4, one can obtain

$$\begin{aligned} \Vert u_1\Vert _{C^{2+\theta }(\bar{\Omega })}+\Vert u_2-\gamma _2\Vert _{C^{2+\theta }(\bar{\Omega })}+\Vert \textbf{w}_1\Vert _{C^{1+\theta }(\bar{\Omega })}+\Vert \textbf{w}_2\Vert _{C^{1+\theta }(\bar{\Omega })} \rightarrow 0 \text{ as } t \rightarrow \infty . \end{aligned}$$

Recalling the definitions of \(\mathcal E_2(t)\) and \(\mathcal F_2(t)\), and using (4.27) and Hölder’s inequality, we can find some \(T_1>0\) such that

$$\begin{aligned} \mathcal E_2(t)&\le C_1\left( \int _\Omega u_1+\int _\Omega \left( u_2-\gamma _2\right) ^2\right) \\&\le C_2 \left\{ \left( \int _\Omega u_1^2\right) ^\frac{1}{2}+\left( \int _\Omega \left( u_2-\gamma _2\right) ^2\right) ^\frac{1}{2}\right\} \\&\le C_3 \mathcal F_2^\frac{1}{2}(t)\quad \text {for all }t>T_1, \end{aligned}$$

which together with (4.36) gives that

$$\begin{aligned} \frac{d}{d t} \mathcal {E}_{2}(t)+\frac{\theta _{2}}{C_{3}^2} \mathcal E_{2}^2(t) \le 0 \quad \text{ for } \text{ all } t>T_1. \end{aligned}$$

Solving the above ordinary differential inequality, we arrive at

$$\begin{aligned} \mathcal {E}_{2}(t) \le \frac{C_4}{1+t} \quad \text{ for } \text{ all } t>T_1. \end{aligned}$$

The rest of the proof can follow similar arguments as in the proof of Lemma 4.5 and we omit it for brevity. \(\square \)

Proof of Theorem 1.2.

The assertions in (i) and (ii) of Theorem 1.2 result from Lemma 4.5 and Lemma 4.7, respectively. The assertions in (iii) are essentially similar to those in (ii) and can be proved by simply swapping \(u_1, b,\gamma _1\) with \(u_2, c,\gamma _2\), respectively, in the proof of (ii). \(\square \)