6.1 Introduction

Multi-taxis appears in society interactions and cancer treatment. Society interactions can lead to the complex dynamical behavior in biology and even in criminology (Eftimie et al. 2007; Guttal and Couzin 2010; Short et al. 2008). A particular example in this direction is mixed-species foraging flocks, such as the formation of Alaska’s shearwater flocks through attraction to kittiwake foragers (Hoffman et al. 1981). Oncolytic viruses (OV) are a kind of viruses that preferentially infect and destroy cancer cells. Oncolytic viruses can be engineered by some of the less virulent viruses in nature and be readily combined with other agents. A diverse range of viruses has been investigated as potential cancer therapeutics, such as herpesvirus, adenovirus, vaccinia virus measles virus and polio virus, and oncolytic virotherapy offers a novel promising cancer treatment modality (Breitbach and Parato 2015; Goldsmith et al. 1998; Msaouel et al. 2013).

This chapter is concerned with the multi-taxis diffusion systems modeling foraging–scrounging interplay or oncolytic virotherapy. Section 6.3 is concerned with the asymptotic behavior in a doubly tactic resource consumption model with proliferation. Toward better understanding of the effect of foraging–scrounging interplay on spatio-temporal dynamics, the authors of Tania et al. (2012) proposed the forager–scrounger system given by

$$\begin{aligned} \left\{ \begin{aligned}&u_t= \varDelta u-\chi _1\nabla \cdot (u\nabla w), \\&v_t=\varDelta v-\chi _2\nabla \cdot (v\nabla u), \\&w_t=d\varDelta w-\lambda (u+v)w-\mu w+r \end{aligned} \right. \end{aligned}$$
(6.1.1)

with positive parameters \(d, \chi _1,\chi _2,\lambda \) and nonnegative parameters \(\mu \) and r, for the unknown population densities \(u=u(x,t)\) and \(v=v(x,t)\) of foragers and scroungers and nutrient concentration \(w=w(x,t)\), respectively. The term \(-\nabla \cdot (u\nabla w)\) accounts for the tendency of foragers moving toward the increasing resource concentration, and \(-\nabla \cdot (v\nabla u)\) models the movement of scroungers following the actively searching foragers rather the resource. Due to the sequential taxis-type cross-diffusion mechanisms in (6.1.1), the considerable extra difficulties seem to be expected when compared to the corresponding scrounger-free system

$$\begin{aligned} \left\{ \begin{aligned}&u_t= \varDelta u-\chi _1\nabla \cdot (u\nabla w), \\&w_t=d\varDelta w-\lambda uw-\mu w+r. \end{aligned} \right. \end{aligned}$$
(6.1.2)

Indeed, in the prototypical case \(\mu =r=0\), the two-dimensional version of (6.1.2) exhibits a substantially stronger tendency toward spatial homogeneous equilibria, which is also valid for its 3D analog at least after some waiting times (Tao and Winkler 2012c). This result implies that any destabilization of the taxis mechanism in (6.1.2) can be suppressed by the relaxation of the diffusion process together with nutrient consumption and thereby allows for a certain entropy-like structure. The feature of (6.1.1) is the sequential taxis, that is, the nutrient-taxis mechanism from (6.1.2) coupled with forager-taxis mechanism. In this situation, the mild relaxation of foragers may not suppress the potential of destabilization driven by the forager-taxis mechanism and thus limits the accessibility of energy-like techniques from the mathematical point of view. Accordingly, to the best of our knowledge, the analytical results in the literature are available only for the low dimensions or certain generalized solutions, and thereby, the comprehensive understanding of (6.1.1) is still far from complete (Black 2020; Cao 2020; Cao and Tao 2021; Liu 2019; Liu and Zhuang 2020; Tao and Winkler 2019b; Wang and Wang 2020; Winkler 2019c). For example, Tao and Winkler (2019b) established the existence of global classical solutions to the corresponding Neumann initial-boundary value problem of (6.1.1) in the one-dimensional setting for suitably regular initial data, as well as an exponential stabilization provided that the initial masses of either u or v are suitably small. As for the higher dimensional model (6.1.1), only generalized solutions are considered in Winkler (2019c) under an explicit condition on the initial datum for w and r, and moreover, they can approach spatially homogeneous equilibria in the large time limit if r decays sufficiently fast. For more related works on smooth properties of solutions to the variants of (6.1.1), inter alia accounting for the superlinear degradation mechanisms of two populations, we refer the readers to Black (2020) and Wang and Wang (2020).

On the other hand, (6.1.2) may be viewed as a kind of the predator–prey system with prey-taxis:

$$\begin{aligned} \left\{ \begin{aligned}&u_t= \varDelta u-\chi _1\nabla \cdot (u\nabla w)+ uf(w,u)+h(u), \\&w_t=d\varDelta w-\lambda uf(w,u)+g(w), \end{aligned} \right. \end{aligned}$$
(6.1.3)

where u(x, t) and w(x, t) are predator density and prey density, respectively; \(\chi _1\nabla w \) is the velocity of predators pursuing preys (i.e., prey-taxis); h(u) and g(w) represent the intra-specific interaction of predators and preys, while f(w, u) is the functional response, and its typical form in the literature is \(f(w,u)=w\) (Lotka–Volterra type) and \(\frac{1}{\lambda }\) is the biomass conversion rate from the prey loss to predator gain. In contrast to the attractive Keller–Segel model, prey-taxis in most cases of (6.1.3) tends to stabilize the predator–prey interactions and may actually lead to the lack of pattern formation, which contradicts intuitive assumptions (Chakraborty et al. 2007; Lee et al. 2008, 2009; Lewis 1994). It also has been recognized that the possibility of spatial pattern formation in (6.1.3) crucially depends on the death rate of predators, the prey growth kinetics g(w) and inter alia functional forms of functional response f(w, u) (Cai et al. 2022; Lee et al. 2009; Wang et al. 2015). In addition to the pattern formation in (6.1.3), the question of which extent the intrinsic predator–prey interaction may preclude the population overcrowding has received considerable attention (see Jin and Wang 2017; Wang and Wang 2019b; Wu et al. 2018; Xiang 2018 and references therein).

In synopsis of the above results, it is natural to consider the dynamical behavior of (6.1.1) when the proliferation of foragers and scroungers is taken into account, which thus indicates that the population proliferation essentially relies on the availability of nutrient resources. Specially, this work will be concerned with the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{aligned}&u_t=\varDelta u-\chi _u\nabla \cdot (u\nabla w)+uw,&x\in \varOmega , t>0, \\&v_t=\varDelta v-\chi _v\nabla \cdot (v\nabla u)+vw,&x\in \varOmega , t>0, \\&w_t=\varDelta w-\lambda (u+v)w-\mu w,&x\in \varOmega , t>0, \\&\nabla u\cdot \nu =\nabla v\cdot \nu =\nabla w\cdot \nu =0,&x\in \partial \varOmega , t>0, \\&u(x,0)=u_0(x), \;v(x,0)=v_0(x),\;w(x,0)=w_0(x),&x\in \varOmega \end{aligned}\right. \end{aligned}$$
(6.1.4)

in a smoothly bounded domain \(\varOmega \subset \mathbb R^N\), \(N\ge 1\), where \(\nu \) denotes the outward normal vector field on \(\partial \varOmega \).

It is worthwhile to mention that (6.1.4) can be regarded as a relative of

$$\begin{aligned} \left\{ \begin{aligned}&u_t=\varDelta u-\chi \nabla \cdot (u\nabla w)+ uw,&x\in \varOmega , t>0, \\&v_t= \alpha vw,&x\in \varOmega , t>0, \\&w_t=\varDelta w-\beta uw -\gamma vw,&x\in \varOmega , t>0, \\&\nabla u\cdot \nu =\nabla w\cdot \nu =0,&x\in \partial \varOmega , t>0, \\&u(x,0)=u_0(x), \;v(x,0)=v_0(x),\;w(x,0)=w_0(x),&x\in \varOmega , \end{aligned}\right. \end{aligned}$$
(6.1.5)

which describes the competition between the populations u and v feeding on a common single non-renewable resource w. The authors of Krzyżanowski et al. (2019) asserted global solvability of problem (6.1.5) within a natural weak solution concept and moreover provided an analytical evidence which indicates that under suitably small initial nutrient distributions, in the long time perspective, the motility ability of population u will turn out to be a competitive advantage irrespectively of the competitive kinetics thereof. It should be remarked that the structure of (6.1.5) is comparatively simple enough to allow for the quasi-dissipative property, which seems to be lost due to the taxis-type cross-diffusive term in the second equation of (6.1.4). Inspired by Cao and Lankeit (2016), Myowin et al. (2020) and Li et al. (2019b), we shall consider the asymptotic behavior of (6.1.4) under suitably small initial data. Our standing assumptions on the initial data herein will be that

$$\begin{aligned} \left\{ \begin{aligned}&v_0\in W^{1,\infty }(\varOmega )~\hbox {is nonnegative}~ \hbox {with}~v_0\not \equiv 0~ \hbox {and that}\\&(u_0,w_0)\in (W^{2,\infty }(\varOmega ))^2~\hbox {is nonnegative}~\hbox {with}~\displaystyle \frac{\partial u_0}{\partial \nu }=0,~\frac{\partial w_0}{\partial \nu }=0. \end{aligned} \right. \end{aligned}$$
(6.1.6)

In this setting, all of the solutions of (6.1.4) approach spatially homogeneous profiles in the large time limit when suitably regular initial data satisfy a certain small condition, which reads as follows (Li and Wang 2021a). It is remarked that in comparison with the relative results of Wang and Wang (2020), the small restriction on initial data does not involve \(v_0\) herein.

Theorem 6.1

Let \(\varOmega \subset \mathbb R^N\) (\(N\ge 1\)) be a bounded domain with smooth boundary and \(m_\infty =\frac{1}{|\varOmega |}\int _\varOmega (\lambda u_0+\lambda v_0+w_0)\). Then there exists \(\varepsilon _0>0\) such that for all \(\varepsilon <\varepsilon _0\) and

$$\begin{aligned}&\Vert (\lambda u_0+\lambda v_0+w_0)(\cdot )-m_\infty \Vert _{L^\infty (\varOmega )}\le \varepsilon ,\quad \Vert \nabla u_0\Vert _{L^{2p_0}(\varOmega )}\le \varepsilon , \\&\Vert w_0\Vert _{L^\infty (\varOmega )}\le \varepsilon ,\quad \Vert \nabla w_0\Vert _{L^{2p_0}(\varOmega )}\le \varepsilon ,\quad \Vert \varDelta w_0\Vert _{L^{p_0}(\varOmega )}\le \varepsilon \end{aligned}$$

with some \(p_0\in \mathbb N\) satisfying \(p_0>1+\frac{N}{2}\), the problem (6.1.4) admits a unique nonnegative global classical solution \((u,v,w)\in (C(\overline{\varOmega }\times [0,\infty ))\cap C^{2,1}(\overline{\varOmega }\times (0,\infty )))^3\). Moreover, there exist constants \(u^*\in (0,\frac{m_\infty }{\lambda })\), \(v^*\in (0,\frac{m_\infty }{\lambda })\) and \(K_i>0\) \( (i=1,2,3)\) such that for all \(t\in (0,\infty )\), we have

$$\begin{aligned}&\Vert u(\cdot ,t)-u^*\Vert _{L^\infty (\varOmega )}\le K_1\varepsilon e^{-\alpha t}, \\&\Vert v(\cdot ,t)-v^*\Vert _{L^\infty (\varOmega )}\le K_2\varepsilon e^{-\alpha t}, \\&\Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le K_3\varepsilon e^{-\alpha t}, \end{aligned}$$

where \(\alpha =\min \{\lambda _1,\mu \}\) for \(\mu >0\) and \(\alpha =\min \{\lambda _1,m_\infty \}\) for \(\mu =0\) with \(\lambda _1>0\) the first nonzero eigenvalue of \(-\varDelta \) in \(\varOmega \) under the Neumann boundary condition.

It is noted that the \(L^p-L^q\) estimate for the Neumann heat semigroup \(e^{t\varDelta }\): there exists \(C>0\) such for all \(\omega \in L^q(\varOmega )\) with \(\int _\varOmega \omega =0\),

$$\Vert e^{t\varDelta }\omega \Vert _{L^p(\varOmega )}\le C\left( 1+t^{-\frac{N}{2}(\frac{1}{q}-\frac{1}{p})}\right) e^{-\lambda _1t}\Vert \omega \Vert _{L^q(\varOmega )} $$

plays an important role in the derivation of decay estimations in Cao and Lankeit (2016), Myowin et al. (2020) and Li et al. (2019b). However, for the doubly tactic model (6.1.4) with \(\mu >0\), despite its dissipative feature, a more subtle effort is required in rigorous analysis due to the invalid of the mass conservation of \(\lambda u(\cdot ,t)+ \lambda v(\cdot ,t)+w(\cdot ,t)\). Indeed, the core of our argument is to verify that the interval (0, T) on which solutions enjoy some exponential decay properties can be extended to \((0,\infty )\) in which the application of above \(L^p-L^q\) estimate seems to be necessary. To this end, a nonnegative auxiliary quantity \(z(\cdot ,t)=\mu \int _0^te^{(t-s)\varDelta } w(\cdot ,s)ds\) is introduced and accordingly allows us to apply \(L^p-L^q\) estimate in our argument since the mass of \(\lambda u(\cdot ,t)+ \lambda v(\cdot ,t)+w(\cdot ,t)+z(\cdot ,t)\) is conserved now. It should be remarked that our approach is also valid when system (6.1.4) takes into account nutrient renewal r(x, t) with a certain temporal decay.

Oncolytic virotherapy offers a novel promising cancer treatment modality and currently has some limitations in the oncolytic efficacy, which might be the result of virus clearance and the physical barriers inside tumors such as the interstitial fluid pressure, extracellular matrix (ECM) deposits and tight inter-cellular junctions. The next two sections of this chapter focus on the boundedness and asymptotic behavior for solutions to oncolytic virotherapy models involving triply haptotactic terms. To better understand the physical barriers that limit virus spread, the authors of Alzahrani et al. (2019) recently proposed the PDE-ODE system of the form

$$\begin{aligned} \left\{ \begin{aligned}&u_t=D_u\varDelta u-\xi _u\nabla \cdot (u\nabla v)+\mu _u u(1-u)-\displaystyle \rho _u uz,&\\&\displaystyle {w_t=D_w\varDelta w-\xi _w\nabla \cdot (w\nabla v)- \delta _w w+\rho _w uz},&\\&\displaystyle { v_t=- (\alpha _u u+\alpha _w w)v+\mu _v v(1-v)},&\\&\displaystyle {z_t=D_z\varDelta z-\xi _z\nabla \cdot (z\nabla v)-\delta _z z-\displaystyle \rho _z uz+\beta w}&\\ \end{aligned}\right. \end{aligned}$$
(6.1.7)

to describe the coupled dynamics of uninfected cancer cells u, OV-infected cancer cells w, ECM v and oncolytic viruses (OV) z. Herein, the underlying modeling hypotheses are that in addition to random diffusion with the respective motility coefficient \(D_u\) and \(D_w\), cancer cells can direct their movement toward regions of higher ECM densities with the haptotactic coefficient \(\xi _u, \xi _w\), respectively, and that uninfected cells, apart from proliferating logistically at rate \(\mu _u\), are converted into an infected state upon contact with virus particles, whereas infected cells die owing to lysis at a rate \(\delta _w\). It is assumed that the static ECM is degraded by both types of cancer cells, possibly remodeled with rate \(\mu _v\) in the sense of spontaneous renewal of healthy tissue. Finally, it is also supposed that besides the random motion with \(D_z\) the random motility coefficient, virus particles move up the gradient of ECM with the ECM-OV-taxis rate \(\xi _z\), increase at a rate \(\beta \) due to the release of free virus particles through infected cells and undergo decay at the rate \(\delta _z\) accounting for the natural virions’ death as well as the trapping of these virus particles into the cancer cells.

From a mathematical perspective, model (6.1.7) on the one hand involves three simultaneous haptotaxis processes, but on the other hand contains the production term \(\rho uz\) in \(w-\)equation which distinguishes (6.1.7) from the most of the previous haptotaxis (Fontelos et al. 2002; Marciniak-Czochra and Ptashnyk 2010; Liţcanu and Morales-Rodrigo 2010b; Tao 2011; Winkler 2018b) and chemotaxis–haptotaxis models (Pang and Wang 2018; Stinner et al. 2014; Tao and Winkler 2019a). In fact, the haptotactic migration of u, z toward higher densities v simultaneously, in which no smoothing action on the spatial regularity of v can be expected, renders us unable to apply smoothing estimates for the Neumann heat semigroup to gain a priori boundedness information on u and z beyond the norm in \(L^1 (\varOmega )\). Accordingly, this superlinear production term \(\rho uz\) in (6.1.7) seems likely to increase the destabilizing potential in the sense of enhancing the tendency toward blow-up of solutions and thus becomes the key contributor to mathematical challenges already given in the derivation of global solvability theory of (6.1.7), which is also indicated in the qualitative analysis of chemotaxis-May–Nowak model (Bellomo and Tao 2020; Bellomo et al. 2019; Hu and Lankeit 2018; Winkler 2019d).

Though the methodological limitations seem to widely restrict the theoretical understanding of the full model (6.1.7), some analytical works on simplifications of the latter have recently been achieved in Tao and Winkler (2020a, 2020b, 2021). Indeed, upon neglecting haptotactic migration processes of infected tumor cells and oncolytic viroses, renewal of ECM as well as proliferation of infected tumor cells, Tao and Winkler considered the corresponding Neumann initial-boundary value problem for

$$\begin{aligned} \left\{ \begin{aligned}&u_{t}=\varDelta u-\nabla \cdot \left( u\nabla v \right) -\rho uz,\\&v_{t}=-\left( u+w \right) v,\\&w_{t}=D_{w}\varDelta w-w+uz ,\\&z_{t}=D_{z}\varDelta z-z-uz+\beta w \end{aligned} \right. \end{aligned}$$
(6.1.8)

in a bounded domain \(\varOmega \subset \mathbb {R}^2\) and obtained that the globally defined classical solution is bounded if \(0<\beta <1\), \(\rho \ge 0\) (Tao and Winkler 2020a), whereas for \(\beta >1\) and \(\int _{\varOmega }u(\cdot ,0)>|\varOmega |/(\beta -1)\), infinite-time blow-up occurs at least in the particular case when \(\rho =0\) (Tao and Winkler 2021). In order to provide an complement to this, the study in Tao and Winkler (2022) reveals that for any \(\rho \ge 0\) and arbitrary \(\beta >0\), at each prescribed level \(\gamma \in (0,1/(\beta -1)_+)\), one can identify an \(L^\infty \)-neighborhood of the homogeneous distribution \((u,v,w,z)\equiv (\gamma ,0,0,0)\) within which all initial data lead to globally bounded solutions that stabilize toward the constant equilibrium \((u_\infty ,0,0,0)\) with some \(u_\infty >0\). On the other hand, in Tao and Winkler (2020c), it is proved that if \(\beta \in (0,1)\), for any choice of \(M>0\), one can find initial data such that the globally defined classical solution satisfies \(u\ge M\) in \(\varOmega \times (0,\infty )\).

Moreover, for the doubly haptotactic version of (6.1.7) with \(\xi _z=0\), the global classical solvability to the corresponding initial-boundary value problem for more comprehensive systems of the form

$$\begin{aligned} \left\{ \begin{aligned}&u_t=D_u\varDelta u-\xi _u\nabla \cdot (u\nabla v)+\mu _u u(1-u)-\displaystyle \rho _u uz, \\&\displaystyle {w_t=D_w\varDelta w-\xi _w\nabla \cdot (w\nabla v)- \delta _w w+\rho _w uz}, \\&\displaystyle { v_t=- (\alpha _u u+\alpha _w w)v+\mu _v v(1-v)}, \\&\displaystyle {z_t=D_z\varDelta z-\delta _z z-\displaystyle \rho _z uz+\beta w}. \\ \end{aligned}\right. \end{aligned}$$
(6.1.9)

is proved in Tao and Winkler (2020b). This is achieved by discovering a quasi-Lyapunov functional structure that allows to appropriately cope with the presence of nonlinear zero-order interaction terms which apparently form the most significant additional mathematical challenge of the considered system in comparison to previously studied haptotaxis models.

The purpose of Sect. 6.4 is to a more comprehensive understanding of model (6.1.7) in the biologically most relevant constellation in which the haptotactic motion of virus particles is taken into account particularly, and either the production term uz or proliferating term \(\mu _u u(1-u)\) is adjusted to \(\displaystyle \frac{uz}{k_u + \theta u}\) of the Beddington–deAngelis type with positive parameters \(k_u,\theta \) (Bellomo and Tao 2020) or \(\mu _u u(1-u^r)\) of superquadratic type, respectively. We are concerned with the PDE-ODE system given by

$$\begin{aligned} \left\{ \begin{aligned}&u_t=D_u\varDelta u-\xi _u\nabla \cdot (u\nabla v)+\mu _u u(1-u^r)-\displaystyle \frac{\rho uz}{k_u + \theta u},&x\in \varOmega , t>0,\\&\displaystyle {w_t=D_w\varDelta w-\xi _w\nabla \cdot (w\nabla v)- \delta _w w+\displaystyle \frac{\rho uz}{k_u +\theta u}},&x\in \varOmega , t>0,\\&\displaystyle { v_t=- (\alpha _u u+\alpha _w w)v+\mu _v v(1-v)},&x\in \varOmega , t>0,\\&\displaystyle {z_t=D_z\varDelta z-\xi _z\nabla \cdot (z\nabla v)-\delta _z z-\displaystyle \frac{\rho uz}{k_u + \theta u}+\beta w},&x\in \varOmega , t>0,\\&\displaystyle {(D_u\nabla u-\xi _u u\nabla v)\cdot \nu =0},&x\in \partial \varOmega , t>0,\\&\displaystyle {(D_w\nabla w-\xi _w w\nabla v)\cdot \nu =0},&x\in \partial \varOmega , t>0,\\&\displaystyle {(D_w\nabla z-\xi _z z\nabla v)\cdot \nu =0},&x\in \partial \varOmega , t>0,\\&\displaystyle {u(x,0)=u_0(x)},w(x,0)=w_0(x), v(x,0)=v_0(x),z(x,0)=z_0(x),&x\in \varOmega \end{aligned}\right. \end{aligned}$$
(6.1.10)

in a bounded domain \(\varOmega \subset \mathbb {R}^2\) with smooth boundary, where for the initial data \((u_0,w_0,v_0, z_0)\), we suppose throughout Sect. 6.4 that

$$\begin{aligned} \left\{ \begin{aligned}&\displaystyle {u_0,w_0,z_0~ \hbox {and}\, v_0 ~\hbox {are nonnegative functions from}~C^{2+\vartheta }(\bar{\varOmega })~ \hbox {for some}~\vartheta \in (0,1),} \\&\displaystyle {\text{ with }~u_0\not \equiv 0,~w_0\not \equiv 0,~z_0\not \equiv 0,~v_0\not \equiv 0~\hbox {and}~\frac{\partial w_0}{\partial \nu }=0~~\text{ on }~~\partial \varOmega .} \\ \end{aligned} \right. \end{aligned}$$
(6.1.11)

Beyond the global classical solvability, in Sect. 6.4, we focus on the global boundedness of classical solutions to (6.1.10)–(6.1.11) stated as follows, which can be regarded as a first step toward the qualitative comprehension of (6.1.10) (Li and Wang 2021b).

Theorem 6.2

Let \(\varOmega \subset \mathbb {R}^2\) be a bounded domain with smooth boundary, \(D_u,D_w\), \(D_z\), \(\xi _u,\xi _w, \xi _z,\mu _u,\mu _v,\rho ,k_u,\alpha _u,\alpha _w,\beta ,\delta _w\) and \(\delta _z \) are positive parameters. Suppose that \(r=1,\theta >0\) or \(r>1,\theta \ge 0\). Then for any choice of \((u_0, w_0, v_0, z_0)\) fulfilling (6.1.11), there exists \(C>0\) such that if \(\xi _w \alpha _w<C\), (6.1.10) admits a unique global classical solution (u, w, v, z), where \(\Vert u(\cdot ,t)\Vert _{L^\infty {(\varOmega )}}\),\(\Vert w(\cdot ,t)\Vert _{L^\infty {(\varOmega )}}\) and \(\Vert z(\cdot ,t)\Vert _{L^\infty {(\varOmega )}}\) are uniformly bounded for \(t\in (0,\infty )\).

Remark 6.1

In line with the above discussion, the boundedness result on (6.1.10) with \(r=1,\theta =0\) is also valid when \(\xi _z=0\).

Remark 6.2

When \(\mu _v=0\), one can see that the restriction on \(\xi _w \alpha _w\) in Theorem 6.2 can be replaced by a certain small condition on \(v_0\).

A cornerstone of our analysis is to show that for the suitably small \(\xi _w \alpha _w\), the functional

$$\begin{aligned}&\mathscr {F}(t) \\ :=&\,\,\displaystyle A\int _\varOmega e^{\chi _w v(\cdot ,t)}b(\cdot ,t)\ln b(\cdot ,t) +\int _\varOmega e^{\chi _u v(\cdot ,t)}a(\cdot ,t)\ln a(\cdot ,t)+\int _\varOmega e^{\chi _z v(\cdot ,t)}c(\cdot ,t)\ln c(\cdot ,t) \end{aligned}$$

with \(a= ue^{-\chi _u v}, b= we^{-\chi _w v}\) and \(c= ze^{-\chi _z v}\) enjoys a certain quasi-dissipative property under appropriate choice of the positive constant A (of (6.4.32)). As the first step in this direction, we perform the variable change used in several precedents, by which the crucial haptotactic contribution to the equations in (6.1.10) is reduced to zero-order terms \(\chi _u a(\alpha _u u+\alpha _w w)v-\chi _u\mu _v av(1-v)\), \(\chi _w b(\alpha _u u+\alpha _w w)v-\chi _w\mu _v bv(1-v)\) and \( \chi _z c(\alpha _u u+\alpha _w w)v-\chi _z\mu _v cv(1-v)\), respectively (see (6.4.1) below). Thanks to a variant of the Gagliardo–Nirenberg inequality involving certain \(L\log L\)-type norms, the latter offers the sufficient regularity so as to allow for the \(L^\infty \)-bounds of solutions in the present two-dimensional setting.

Section 6.5 is devoted to understand the dynamics behavior of (6.1.7) to a considerable extent in higher dimensional settings in light of the above-mentioned results and a recent consideration of global classical solutions to the one-dimensional (6.1.7) in Tao (2021). Specially, taking into account the linear degradation instead of the renewal of ECM and neglecting the proliferation of uninfected tumor cells in (6.1.7), we are concerned with the following Neumann initial-boundary problem in \(\varOmega \subset \mathbb R^N(N\ge 1)\):

$$\begin{aligned} \left\{ \begin{aligned}&u_{t}=\varDelta u-\xi _{u}\nabla \cdot \left( u\nabla v \right) -\rho _{u}uz,\quad \quad&x\in \varOmega ,t> 0,\\&w_{t}=\varDelta w-\xi _{w}\nabla \cdot \left( w\nabla v \right) -\delta _{w}w+\rho _{w}uz,&x\in \varOmega ,t> 0,\\&v_{t}=-(\alpha _{u}u+\alpha _{w}w)v-\delta _v v,&x\in \varOmega ,t> 0,\\&z_{t}=\varDelta z-\xi _{z}\nabla \cdot \left( z\nabla v \right) -\delta _{z}z-\rho _{z}uz+\beta w,&x\in \varOmega ,t> 0,\\&(\nabla u-\xi _{u}u\nabla v)\cdot \nu =(\nabla w-\xi _{w}w\nabla v)\cdot \nu =(\nabla z-\xi _{z}z\nabla v)\cdot \nu =0,&x\in \partial \varOmega ,t> 0,\\&u(x,0)=u_{0}(x),w(x,0)=w_{0}(x),v(x,0)=v_{0}(x),z(x,0)=z_{0}(x),&x\in \varOmega , \end{aligned} \right. \end{aligned}$$
(6.1.12)

where \(\xi _u\), \(\xi _w\), \(\xi _z\), \(\rho _u\), \(\rho _w\), \(\rho _z\), \(\delta _w\), \(\delta _v\), \(\delta _z\), \(\alpha _u\), \(\alpha _w\) and \(\beta \) are positive parameters, for the initial data \((u_0,w_0,v_0, z_0)\), we suppose throughout the third part of Chap. 6 that

$$\begin{aligned} \left\{ \begin{aligned} \displaystyle {u_0,w_0,v_0,z_0 ~\hbox {are nonnegative functions from} ~ C^{2+\nu }(\overline{\varOmega })~\hbox {for some}~\nu \in (0,1)}, \\ \displaystyle {~u_0\not \equiv 0,~w_0\not \equiv 0,~v_0\not \equiv 0,~z_0\not \equiv 0~\hbox {and}~\frac{\partial u_0}{\partial \nu }=\frac{\partial w_0}{\partial \nu }=\frac{\partial v_0}{\partial \nu }=\frac{\partial z_0}{\partial \nu }=0~~\text{ on }~~\partial \varOmega .} \\ \end{aligned} \right. \end{aligned}$$
(6.1.13)

Our main result makes sure that for suitably small initial data, these solutions will be globally bounded and approach some constant profiles asymptotically (Wei et al. 2022).

Theorem 6.3

Let \(\varOmega \subset \mathbb {R}^{N}\left( N\ge 1 \right) \) be a bounded domain with smooth boundary. Assume (6.1.13) holds and \(\beta \le \frac{\rho _{u}+\rho _{z}}{\rho _{w}}\delta _{w}\). Then if for some \(p_0>\max \{1,\frac{N}{2}\}\), there exists \(\varepsilon >0\) which depends on \(\xi _u, \xi _w, \xi _z, \rho _u, \rho _w, \rho _z, \alpha _y, \alpha _w\) such that

$$\left\| u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0} \right\| _{L^{\infty }\left( \varOmega \right) }\le \varepsilon ,$$
$$\left\| \nabla u_{0}(\cdot ) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \varepsilon , \;\left\| \nabla w_{0}(\cdot ) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \varepsilon ,$$
$$\left\| \nabla v_{0}(\cdot ) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \varepsilon , \;\left\| \nabla z_{0}(\cdot ) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \varepsilon , $$
$$\left\| \varDelta u_{0}(\cdot ) \right\| _{L^{p_{0}}\left( \varOmega \right) }\le \varepsilon ,\quad \quad \left\| \varDelta w_{0}(\cdot ) \right\| _{L^{p_{0}}\left( \varOmega \right) }\le \varepsilon ,\quad \left\| \varDelta v_{0}(\cdot ) \right\| _{L^{p_{0}}\left( \varOmega \right) }\le \varepsilon , $$

the problem (6.1.12) has a unique nonnegative global classical solution

$$\left( u,v,w,z \right) \in \left( C\left( \bar{\varOmega }\times [0,\infty ) \right) \cap C^{2,1}\left( \bar{\varOmega }\times (0,\infty ) \right) \right) ^{4}.$$

Moreover, there exists a nonnegative constant \(u^{*}\) such that

$$\Vert u(\cdot ,t)-u^{*}\Vert _{L^\infty (\varOmega )}\rightarrow 0, $$
$$\Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\rightarrow 0, $$
$$\Vert z(\cdot ,t)\Vert _{L^\infty (\varOmega )}\rightarrow 0, $$
$$\Vert v(\cdot ,t)\Vert _{L^\infty (\varOmega )}\rightarrow 0$$

as \(t\rightarrow \infty \).

Remark 6.3

Our result indicates that the infected cancer cells and virus particle population can become extinct asymptotically and the density of uninfected cancer cells tends to a nonnegative constant \(u^*\) which is less than \(\overline{u}_0\). This result implies that the oncolytic virotherapy is effective. Unfortunately, the condition under which \(u^*\) equals to zero is left as an open problem.

Same to the analysis in Sect. 6.3, a more subtle effort seems to be required for our analysis in Sect. 6.5 due to the decreasing of \(\int _{\varOmega }\left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z\right) (\cdot ,s)ds\). To this end, a nonnegative auxiliary quantity

$$Q(\cdot ,t)=\int _0^te^{(t-s)\varDelta } \left[ -\left( \beta -\frac{\rho _{u}+\rho _{z}}{\rho _{w}}\delta _{w} \right) w+\delta _{z}z\right] (\cdot ,s)ds$$

is introduced and accordingly allows us to apply \(L^p-L^q\) estimate in our argument since the mass of \(u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q\) is conserved now.

6.2 Preliminaries

In this section, we provide some preliminary results that will be used in the subsequent sections.

By applying the maximal Sobolev regularity (Theorem 3.1 of Hieber and Prüss 1997), we can obtain the following lemma, which together with Lemmas 1.1, 3.2 and 4.3 will play an important role in the proof of our main results in Sects. 6.3 and 6.5.

Lemma 6.1

(Ishida et al. 2014, Lemma 2.1; Yang et al. 2015, Lemma 2.2) Let \(r\in (1,\infty )\) and consider the following evolution equation:

$$\begin{aligned} \left\{ \begin{aligned}&\displaystyle h_t=\varDelta h+f,\quad&(x,t)\in \varOmega \times (0,T), \\&\displaystyle \nabla h\cdot \nu =0,\quad&(x,t)\in \partial \varOmega \times (0,T), \\&h(x,0)=h_0(x),\quad&x\in \varOmega . \end{aligned} \right. \end{aligned}$$
(6.2.1)

Then for each \(h_0\in W^{2,r}(\varOmega )\) with \(\nabla h_0\cdot \nu =0\) on \(\partial \varOmega \) and any \(f\in L^r((0,T),L^r(\varOmega ))\), (6.2.1) admits a unique mild solution \(h\in W^{1,r}((0,T);L^r(\varOmega ))\cap L^r((0,T);W^{2,r}(\varOmega ))\). Moreover, there exists \(C_r>0\), such that

$$\begin{aligned} \int _{0}^T\int _\varOmega |\varDelta h|^r\le C_r\int _{0}^T\int _\varOmega |f|^r+C_r(\Vert h_0\Vert _{L^r(\varOmega )}^r+\Vert \varDelta h_0\Vert _{L^r(\varOmega )}^r). \end{aligned}$$
(6.2.2)

The following lemma is a special case of Lemma A.5 in Tao and Winkler (2014b) and can be regarded as a variant of a Gagliardo–Nirenberg inequality originally derived in Biler et al. (1994), which will be of importance in the later analysis in Sect. 6.4.

Lemma 6.2

Let \(\varOmega \subset \mathbb {R}^2\) be a bounded domain with smooth boundary, and let \(p \in (1,\infty )\) and \(\varepsilon > 0\). Then there exists \(K(p,\varepsilon ) >0\) such that

$$ \Vert \varphi \Vert ^{\frac{2(p+1)}{p}}_ {L^{\frac{2(p+1)}{p}}(\varOmega )} \le \varepsilon \Vert \nabla \varphi \Vert ^2_{L^2(\varOmega )}\cdot \int _\varOmega |\varphi |^{\frac{2}{p}}|\ln |\varphi || + K(p, \varepsilon ) \left( \Vert \varphi \Vert ^{\frac{2(p+1)}{p}}_ {L^{\frac{2}{p}}(\varOmega )}+1\right) $$

holds for all \(\varphi \in W ^{1,2}(\varOmega )\).

The following Lemma is based on simple calculations on the maximal value of the function \(f(t)=t^ae^{-b t}\) for \(t\ge 0\), which is used in the analysis in Sect. 6.5.

Lemma 6.3

For all \(a>0\), \(b>0\), we have

$$t^a e^{-b t}\le \left( \frac{a}{be}\right) ^a$$

holds for all \(t\ge 0\).

6.3 Asymptotic Behavior of Solutions to a Doubly Tactic Resource Consumption Model

At the beginning of this subsection, we provide a basic state on local existence and extensibility of solutions to (6.1.4), which can be readily proved by the Amann theory. Similar proof thereof can be found in Tao and Winkler (2019b, Lemma 2.1) and hence we omit the detail here.

Lemma 6.4

There exist a maximal existence time \(T_{max}\in (0,\infty ]\) and \((u,v,w)\in C(\overline{\varOmega }\times [0,T_{max}))\cap C^{2,1}(\overline{\varOmega }\times (0,T_{max}))\) such that (u, v, w) is the unique nonnegative solution of (6.1.4) in \([0,T_{max})\). Furthermore, if \(T_{max}<+\infty \), then

$$\lim _{t\rightarrow T_{max}}\left( \Vert u(\cdot ,t)\Vert _{W^{1,q}(\varOmega )}+\Vert v(\cdot ,t)\Vert _{W^{1,q}( \varOmega )}+\Vert w(\cdot ,t)\Vert _{W^{1,q}(\varOmega )}\right) =\infty $$

for all \(q>N\).

Theorem 6.1 is the consequence of the following lemmas. In the proof of these lemmas, the constants \(c_i>0\), \(i=0,1,\ldots ,4\), refer to those in Lemmas 1.1 and 4.3, respectively.

We first collect some easily verifiable observations in the following lemma:

Lemma 6.5

Under the assumptions of Theorem 6.1, there exist \(M_i>1(i=1,2,3),\) and \(\varepsilon >0\) such that

$$\begin{aligned}&2c_{10}c_4(1+m_\infty )(\chi _uM_4+\chi _vM_3)\le \frac{M_1}{2}, \end{aligned}$$
(6.3.1)
$$\begin{aligned}&e^{\frac{2}{\alpha }}\le \frac{M_2}{2}, \end{aligned}$$
(6.3.2)
$$\begin{aligned}&2c_3+2c_{10}c_2\left( (\chi _uM_6+M_2|\varOmega |^{\frac{1}{2p_0}})\frac{1+m_\infty }{\lambda }+1\right) \le \frac{M_3}{2}, \end{aligned}$$
(6.3.3)
$$\begin{aligned}&M_2\varepsilon \le 1,\;M_5\varepsilon \le 1,\;M_3\varepsilon \le 1,\;\chi _uM_4M_3\varepsilon \le 1,\; 4\chi _vc_2M_3\varepsilon \le 1, \end{aligned}$$
(6.3.4)

where

$$M_4:=2 c_3+2|\varOmega |^{\frac{1}{2p_0}}\left( 1+m_\infty +\mu \right) c_{10} c_2M_2,\quad M_5:=M_1+2c_1,$$
$$M_6:=2 c_1+2c_4M_2c_{10}\lambda \left( |\varOmega |^{\frac{1}{2p_0}} +C_5|\varOmega |^{\frac{p_0-2}{2p_0(p_0-1)}}\right) +2c_4M_4c_{10}\left( 1 +m_\infty +\mu \right) |\varOmega |^{\frac{1}{2p_0}}$$

and constant \(C_5>0\) is given in Lemma 6.10 below which is independent of \(M_i(i=1,2,3)\).

To obtain a conservation law of mass, we introduce an nonnegative variable z satisfying

$$\begin{aligned} \left\{ \begin{aligned}&z_t=\varDelta z+\mu w,\quad (x,t)\in \varOmega \times (0,T), \\&\nabla z\cdot \nu =0,\quad (x,t)\in \partial \varOmega \times (0,T), \\&z(x,0)=z_0(x)\equiv 0,\quad x\in \varOmega , \end{aligned} \right. \end{aligned}$$
(6.3.5)

then it is easy to see that for any \(t\in [0,T)\),

$$\begin{aligned} \int _\varOmega (\lambda u+\lambda v+w+z)(\cdot ,t)=\int _\varOmega (\lambda u_0+\lambda v_0+w_0). \end{aligned}$$
(6.3.6)

Let

$$\begin{aligned} \!\!T\!\triangleq \!\sup \!\left\{ \!\widetilde{T}\!\in \!(0,T_{max})\!\left| \begin{aligned}&\Vert (\lambda u+\lambda v+w+z)(\cdot ,t)-\!e^{t\varDelta }(\lambda u_0\!+\lambda v_0+\!w_0)(\cdot )\Vert _{L^{\infty }(\varOmega )} \\&\qquad \le M_1\varepsilon e^{-\alpha t}\quad \hbox {for all}~t\in [0,\widetilde{T}); \\&\Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le M_2 \varepsilon e^{-\alpha t}\quad \hbox {for all}~ t\in [0,\widetilde{T}); \\&\Vert \nabla u(\cdot ,t)\Vert _{L^{2p_0}(\varOmega )}\le M_3\varepsilon e^{-\alpha t}~\hbox {for all}~t\in [0,\widetilde{T}). \end{aligned} \!\!\right. \right\} \end{aligned}$$
(6.3.7)

By Lemma 6.4 and the smallness condition on initial data in Theorem 6.1, \(T>0\) is well-defined. We first show \(T=T_{max}\). To this end, we will show that all of the estimates mentioned in (6.3.7) are valid with even smaller coefficients on the right-hand side. The derivation of these estimates will mainly rely on \(L^p-L^q\) estimates for the Neumann heat semigroup and the fact that the classical solutions on \((0,T_{max})\) can be represented as

$$\begin{aligned}&(\lambda u+\lambda v+w+z)(\cdot ,t)=e^{t\varDelta }(\lambda u_0+\lambda v_0+w_0)(\cdot ) \end{aligned}$$
(6.3.8)
$$\begin{aligned}&\qquad \qquad \qquad \qquad \qquad -\lambda \int _0^te^{(t-s)\varDelta }(\chi _u\nabla \cdot (u \nabla w)+\chi _v\nabla \cdot (v \nabla u))(\cdot ,s)ds,\nonumber \\&u(\cdot ,t)=e^{t\varDelta }u_0(\cdot )-\int _0^te^{(t-s)\varDelta }(\chi _u\nabla \cdot (u\nabla w)-uw)(\cdot ,s)ds, \end{aligned}$$
(6.3.9)
$$\begin{aligned}&v(\cdot ,t)=e^{t\varDelta }v_0(\cdot )-\int _0^te^{(t-s)\varDelta }(\chi _v\nabla \cdot (v\nabla u)-vw)(\cdot ,s)ds, \end{aligned}$$
(6.3.10)
$$\begin{aligned}&w(\cdot ,t)=e^{t\varDelta }w_0(\cdot )-\int _0^te^{(t-s)\varDelta }(\lambda (u+v)w+\mu w)(\cdot ,s)ds, \end{aligned}$$
(6.3.11)
$$\begin{aligned}&z(\cdot ,t)=\mu \int _0^te^{(t-s)\varDelta } w(\cdot ,s)ds \end{aligned}$$
(6.3.12)

for all \(t\in (0,T_{max})\) as per the variation-of-constants formula.

Lemma 6.6

Suppose that the assumptions from Theorem 6.1 hold. Then for all \(t\in (0,T)\), we have

$$\begin{aligned}&\Vert (\lambda u+\lambda v+w+z)(\cdot ,t)-m_\infty \Vert _{L^\infty (\varOmega )}\le M_5\varepsilon e^{-\alpha t} \end{aligned}$$
(6.3.13)

with \(M_5:=M_1+2c_1\).

Proof

Due to

$$ e^{t\varDelta }m_\infty =m_\infty , \quad \int _\varOmega [(\lambda u_0+\lambda v_0+w_0)(\cdot )-m_\infty ]=0, $$
$$ \Vert (\lambda u_0+\lambda v_0+w_0)(\cdot )-m_\infty \Vert _{L^{\infty }(\varOmega )}\le \varepsilon , $$

and the assumption of Theorem 6.1, the definition of T along with Lemma 1.1(i) implies that for all \(t\in (0,T)\),

$$\begin{aligned}&\Vert (\lambda u+\lambda v+w+z)(\cdot ,t)-m_\infty \Vert _{L^\infty (\varOmega )} \\ \le&\,\Vert (\lambda u+\lambda v+w+z)(\cdot ,t)-e^{t\varDelta }(\lambda u_0+\lambda v_0+w_0)\Vert _{L^\infty (\varOmega )} \\&+\Vert e^{t\varDelta }[(\lambda u_0+\lambda v_0+w_0)(\cdot )-m_\infty ]\Vert _{L^\infty (\varOmega )} \\ \le&\,\, M_1\varepsilon e^{-\alpha t} +2c_1\Vert (\lambda u_0+\lambda v_0+w_0)(\cdot )-m_\infty \Vert _{L^{\infty }(\varOmega )}e^{-\lambda _1t} \\ \le&\,\,(M_1+2c_1)\varepsilon e^{-\alpha t} \\ =&M_5\varepsilon e^{-\alpha t}. \end{aligned}$$

Lemma 6.7

Under the assumptions of Theorem 6.1, we have

$$\begin{aligned} \Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le \frac{M_2}{2} \varepsilon e^{-\alpha t}\quad \hbox {for all}~ t\in (0,T). \end{aligned}$$
(6.3.14)

Proof

Multiplying the third equation in (6.1.4) by \(kw^{k-1}\) and integrating the result over \(\varOmega \), we get \(\frac{d}{dt}\int _\varOmega w^k\le -k\int _\varOmega (\lambda u+\lambda v+\mu ) w^k\) on (0, T).

In what follows, we shall show (6.3.14) in two cases: \(\mu >0\) and \(\mu =0\).

(I) The case \(\mu >0\). Since \(-(\lambda u+\lambda v+\mu )\le -\mu \), we have

$$\begin{aligned} \quad \frac{d}{dt}\int _\varOmega w^k(\cdot ,t)\le -\mu k\int _\varOmega w^k(\cdot ,t), \end{aligned}$$

and thus

$$\begin{aligned} \int _\varOmega w^k(\cdot ,t) \le&\,\,\int _\varOmega w_0^k(\cdot ) e^{-\mu kt}, \end{aligned}$$

which implies that

$$\begin{aligned} \Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le \varepsilon e^{-\mu t}\le \frac{1}{2} M_2\varepsilon e^{-\mu t} \end{aligned}$$

for any \(t\in (0,T)\), where we have used (6.3.2).

(II) The case \(\mu =0\). Note that \(z\equiv 0\) for all \((x,t)\in \varOmega \times [0,T)\), and thus

$$-(\lambda u+\lambda v)\le |\lambda u+\lambda v+w+z-m_\infty |+w-m_\infty . $$

From the definition of T and Lemma 6.6, it follows that for any \(t\in (0,T)\),

$$\begin{aligned}&\quad \frac{d}{dt}\int _\varOmega w^k(\cdot ,t) \\&\le k\int _\varOmega w^k(\cdot ,t)|(\lambda u+\lambda v+w+z)(\cdot ,t)-m_\infty |+k\int _\varOmega w^{k+1}(\cdot ,t) -km_\infty \int _\varOmega w^k(\cdot ,t) \\&\le k\Vert (\lambda u+\lambda v+w+z)(\cdot ,t)-m_\infty \Vert _{L^\infty (\varOmega )}\int _\varOmega w^k(\cdot ,t)+k\Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\int _\varOmega w^{k}(\cdot ,t) \\&\quad -km_\infty \int _\varOmega w^k(\cdot ,t) \\&\le k((M_5+M_2)\varepsilon e^{-\alpha t}-m_\infty )\int _\varOmega w^k(\cdot ,t) \end{aligned}$$

and hence

$$\begin{aligned} \int _\varOmega w^k(\cdot ,t) \le&\,\,\int _\varOmega w_0^k(\cdot ) e^{k\left( (M_5+M_2)\varepsilon \int _0^te^{-\alpha s}ds -m_\infty t\right) } \\ \le&\,\Vert w_0(\cdot )\Vert _{L^k(\varOmega )}^ke^{k\left( \frac{M_5+M_2}{\alpha }\varepsilon -m_\infty t\right) }, \end{aligned}$$

which implies that for any \(t\in (0,T)\),

$$\begin{aligned} \Vert w(\cdot ,t)\Vert _{L^k(\varOmega )} \le \Vert w_0(\cdot )\Vert _{L^k(\varOmega )}e^{\frac{(M_5+M_2)\varepsilon }{\alpha }-m_\infty t}. \end{aligned}$$
(6.3.15)

Thanks to \(\Vert w_0\Vert _{L^\infty (\varOmega )}\le \varepsilon \) and (6.3.2), (6.3.4), we obtain that for any \(t\in (0,T)\),

$$\begin{aligned} \Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le \frac{1}{2}M_2 \varepsilon e^{-m_\infty t} \end{aligned}$$

by letting \(k\rightarrow \infty \) in (6.3.15).

Lemma 6.8

Let the conditions from Theorem 6.1 be fulfilled. Then for all \(t\in (0,T)\) and \(p_0>\frac{N}{2}\), we have

$$\begin{aligned} \Vert \nabla w(\cdot ,t)\Vert _{L^{2p_0}(\varOmega )}\le M_4\varepsilon e^{-\alpha t} \end{aligned}$$

with \(M_4:=2 c_3+2|\varOmega |^{\frac{1}{2p_0}}\left( 1+m_\infty +\mu \right) c_{10}c_2M_2\).

Proof

By (6.3.11) and Lemma 1.1(iii), noticing \(\Vert \nabla w_0(\cdot )\Vert _{L^{2p_0}(\varOmega )}\le \varepsilon \), we have

$$\begin{aligned}&\quad \Vert \nabla w(\cdot ,t)\Vert _{L^{2p_0}(\varOmega )} \\&\le \Vert \nabla e^{t\varDelta }w_0(\cdot )\Vert _{L^{2p_0}(\varOmega )}+\int _0^t\Vert \nabla e^{(t-s)\varDelta }((\lambda u+\lambda v+\mu )w)(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds\nonumber \\&\le 2 c_3\varepsilon e^{-\lambda _1t}+\int _0^t\Vert \nabla e^{(t-s)\varDelta }((\lambda u+\lambda v+\mu )w)(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds.\nonumber \end{aligned}$$
(6.3.16)

Now, we estimate the last two integrals on the right-hand side of the above inequality. From the definition of T, Lemmas 1.1(ii), 4.3 and 6.6, it follows that

$$\begin{aligned}&\int _0^t\Vert \nabla e^{(t-s)\varDelta }((\lambda u+\lambda v+\mu )w)(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds\nonumber \\ \le&\,\,c_2\int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\Vert (\lambda u+\lambda v+\mu )w(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )} \\ \le&\,\,c_2|\varOmega |^{\frac{1}{2p_0}}\int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\Vert w(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}\Vert (\lambda u+\lambda v+\mu )(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}ds\nonumber \\ \le&\,\,c_2|\varOmega |^{\frac{1}{2p_0}} M_2\varepsilon \int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)} e^{-\alpha s}\left( 1+m_\infty +\mu \right) ds \nonumber \\ \le&\,\, 2|\varOmega |^{\frac{1}{2p_0}}\left( 1+m_\infty +\mu \right) c_{10}c_2M_2\varepsilon e^{-\alpha t}.\nonumber \end{aligned}$$
(6.3.17)

Inserting (6.3.17) into (6.3.16), we get

$$\begin{aligned} \Vert \nabla w\Vert _{L^{2p_0}(\varOmega )}&\le \left( 2c_3+2|\varOmega |^{\frac{1}{2p_0}}\left( 1+m_\infty +\mu \right) c_{10}c_2M_2\right) \varepsilon e^{-\alpha t} \\&=\,\,M_4\varepsilon e^{-\alpha t}, \end{aligned}$$

and thereby complete the proof.

Lemma 6.9

Under the assumptions of Theorem 6.1, for all \(p_0>\frac{N}{2}\), there exists a constant \(C>0\) independent of T such that

$$\begin{aligned} \int _0^{T}\int _\varOmega |\varDelta w(x,s)|^{p_0}dxds\le C, \end{aligned}$$
(6.3.18)
$$\begin{aligned} \int _0^{T}\int _\varOmega |\varDelta u(x,s)|^{p_0}dxds\le C. \end{aligned}$$
(6.3.19)

Proof

Noticing that w satisfies

$$\begin{aligned} \left\{ \begin{aligned}&w_t=\varDelta w+F(x,t),\quad (x,t)\in \varOmega \times (0,T), \\&\nabla w\cdot \nu =0,\quad (x,t)\in \partial \varOmega \times (0,T), \\&w(x,0)=w_0,\quad x\in \varOmega \end{aligned} \right. \end{aligned}$$
(6.3.20)

and u satisfies

$$\begin{aligned} \left\{ \begin{aligned}&u_t=\varDelta u+G(x,t),\quad (x,t)\in \varOmega \times (0,T), \\&\nabla u\cdot \nu =0,\quad (x,t)\in \partial \varOmega \times (0,T), \\&u(x,0)=u_0,\quad x\in \varOmega , \end{aligned} \right. \end{aligned}$$
(6.3.21)

with \(F(x,t)=-\lambda (u+v)w-\mu w\) and \(G(x,t)=-\chi _u(\nabla u\nabla w+u\varDelta w)+uw\), respectively. By Lemmas 6.6 and 6.7, we can see that

$$\begin{aligned} \int _0^{T}\int _\varOmega |F(x,s)|^{p_0}dxds\le \displaystyle \frac{M_2^{p_0}((1+m_\infty )^{p_0}+\mu ^{p_0})}{\alpha p_0}|\varOmega |. \end{aligned}$$

Hence, thanks to Lemma 6.1, we can find \(C_1>0\) independent of T such that

$$\begin{aligned} \int _0^{T}\int _\varOmega |\varDelta w(x,s)|^{p_0}dxds\le C_1. \end{aligned}$$

Similarly, by the definition of T, (6.3.18) and Lemma 6.8, there exists \(C_2>0\) independent of T fulfilling

$$\begin{aligned} \int _0^{T}\int _\varOmega |G(x,s)|^{p_0}dxds\le C_2. \end{aligned}$$

Applying Lemma 6.1 to (6.3.21) once more, we have

$$\begin{aligned} \int _0^{T}\int _\varOmega |\varDelta u(x,s)|^{p_0}dxds\le C_3 \end{aligned}$$

for some \(C_3>0\) independent of T and thereby complete the proof.

Thanks to the decay property of \(w,v, \nabla u\) and space–time \(L^{p_0}\)-estimate for \(\triangle u\), we can establish an \(L^{2(p_0-1)}\) bound for \(\nabla v\) based on Lemmas 1.1 and 4.3.

Lemma 6.10

Suppose that the requirements from Theorem 6.1 are met. Then for all \(p_0>1+\frac{N}{2}\), there exists \(C_5>0\) independent of T and \(M_i(i=1,2,3)\) such that

$$\begin{aligned} \Vert \nabla v(\cdot ,t)\Vert _{L^{2(p_0-1)}(\varOmega )}\le C_5\quad \hbox {for all}~~ t\in (0,T). \end{aligned}$$
(6.3.22)

Proof

By (6.3.10), we have

$$\begin{aligned}&\quad \Vert \nabla v(\cdot ,t)\Vert _{L^{2(p_0-1)}(\varOmega )} \\&\le \Vert \nabla e^{t\varDelta }v_0(\cdot )\Vert _{L^{2(p_0-1)}(\varOmega )}+\chi _v\int _0^t\Vert \nabla e^{(t-s)\varDelta }\nabla \cdot (v\nabla u)(\cdot ,s)\Vert _{L^{2(p_0-1)}(\varOmega )}ds\nonumber \\&\quad +\int _0^t\Vert \nabla e^{(t-s)\varDelta }(vw)(\cdot ,s)\Vert _{L^{2(p_0-1)}(\varOmega )}ds.\nonumber \end{aligned}$$
(6.3.23)

From Lemma 1.1(iii), we obtain

$$\begin{aligned} \Vert \nabla e^{t\varDelta }v_0(\cdot )\Vert _{L^{2(p_0-1)}(\varOmega )}\le 2 c_3\Vert \nabla v_0(\cdot )\Vert _{L^{2(p_0-1)}(\varOmega )} e^{-\lambda _1t}. \end{aligned}$$
(6.3.24)

Now, we estimate the last two integrals on the right-hand side of (6.3.23). From the definition of T, Lemmas 1.1(ii), 4.3, 6.6 and 6.9, it follows that

$$\begin{aligned}&\chi _v\int _0^t\Vert \nabla e^{(t-s)\varDelta }\nabla \cdot (v\nabla u)(\cdot ,s)\Vert _{L^{2(p_0-1)}(\varOmega )}ds\nonumber \\ \le&\,\,c_2\chi _v\int _0^te^{-(t-s)\lambda _1}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{2p_0}+\frac{N}{4(p_0-1)}}\right) \Vert v(\cdot ,s)\varDelta u(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds\nonumber \\&+c_2\chi _v\int _0^te^{-(t-s)\lambda _1}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N(2p_0-1)}{4p_0(p_0-1)}+\frac{N}{4(p_0-1)}}\right) \nonumber \\&\cdot \Vert \nabla v(\cdot ,s)\nabla u(\cdot ,s)\Vert _{L^{\frac{2p_0(p_0-1)}{2p_0-1}}(\varOmega )}ds\nonumber \\ \le&\,\,c_2\chi _v\frac{1+m_\infty }{\lambda } \int _0^t\Vert \varDelta u(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}^{p_0}ds \\&+c_2\chi _v\frac{1+m_\infty }{\lambda } \int _0^te^{-(t-s)\frac{\lambda _1p_0}{p_0-1}}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{2p_0}+\frac{N}{4(p_0-1)}}\right) ^{\frac{p_0}{p_0-1}}ds\nonumber \\&+c_2\chi _v\sup _{t\in (0,T)}\Vert \nabla v(\cdot ,s)\Vert _{L^{2(p_0-1)}(\varOmega )}\nonumber \\&\cdot \int _0^te^{-(t-s)\lambda _1}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N(2p_0-1)}{4p_0(p_0-1)}+\frac{N}{4(p_0-1)}}\right) M_3\varepsilon e^{-\alpha s}ds\nonumber \\ \le&\,\, c_2 \chi _vC_1\frac{1+m_\infty }{\lambda }+c_2\chi _vC_2\frac{1+m_\infty }{\lambda }+2c_2c_{10}\chi _vM_3 \varepsilon e^{-\alpha t}\sup _{t\in (0,T)}\Vert \nabla v(\cdot ,s)\Vert _{L^{2(p_0-1)}(\varOmega )} \nonumber \end{aligned}$$
(6.3.25)

where \(C_1:=\int _0^t\Vert \varDelta u(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}^{p_0}ds\) is bounded by Lemma 6.9 and

$$C_2:=\int _0^te^{-(t-s)\frac{\lambda _1p_0}{p_0-1}}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{2p_0}+\frac{N}{4(p_0-1)}}\right) ^{\frac{p_0}{p_0-1}}ds<+\infty $$

for \(p_0>1+\frac{N}{2}.\)

Next, by Lemmas 1.1(ii), 4.3, 6.6 and 6.7, we get

$$\begin{aligned}&\int _0^t\Vert \nabla e^{(t-s)\varDelta }vw\Vert _{L^{2(p_0-1)}(\varOmega )}ds\nonumber \\ \le&\,\,c_2|\varOmega |^{\frac{1}{2(p_0-1)}}\int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\Vert w\Vert _{L^{\infty }(\varOmega )}\Vert v\Vert _{L^{\infty }(\varOmega )}ds \\ \le&\,\,c_2|\varOmega |^{\frac{1}{2(p_0-1)}}M_2\varepsilon \int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)} e^{-\alpha s}\frac{1+m_\infty }{\lambda }ds \nonumber \\ \le&\,\, 2|\varOmega |^{\frac{1}{2(p_0-1)}} c_{10}c_2 M_2\frac{1+m_\infty }{\lambda } \varepsilon e^{-\alpha t}.\nonumber \end{aligned}$$
(6.3.26)

Inserting (6.3.24)–(6.3.26) into (6.3.23) and using (6.3.4), we readily get

$$\begin{aligned} \sup _{t\in (0,T)}\Vert \nabla v\Vert _{L^{2(p_0-1)}(\varOmega )} \le&\,\,2c_3\Vert \nabla v_0(\cdot )\Vert _{L^{2(p_0-1)}(\varOmega )}+2c_2 \chi _vC_1\frac{1+m_\infty }{\lambda }+2c_2\chi _vC_2\frac{1+m_\infty }{\lambda } \\&+4|\varOmega |^{\frac{1}{2(p_0-1)}}c_{10}c_2 \frac{1+m_\infty }{\lambda }, \end{aligned}$$

and thereby from Lemma 6.5, we arrive at

$$\begin{aligned} \int _\varOmega |\nabla v(x,t)|^{2(p_0-1)}dx\le C_5,\quad t\in (0,T) \end{aligned}$$
(6.3.27)

with some \(C_5>0\) independent of T, \(M_i(i=1,2,3)\) and hence complete the proof.

Beyond the weak information of \(\triangle w\) in Lemma 6.9, we now turn the boundedness of \(\nabla v\) into a statement on decay of \(\triangle w\).

Lemma 6.11

Under the assumptions of Theorem 6.1, for all \(p_0>\frac{N}{2}\),

$$\begin{aligned} \Vert \varDelta w(\cdot ,t)\Vert _{L^{p_0}(\varOmega )}&\le M_6\varepsilon e^{-\alpha t} \quad \hbox {for all}~~ t\in (0,T) \end{aligned}$$
(6.3.28)

with

$$M_6:=2 c_1+2c_4M_2c_{10}\lambda \left( |\varOmega |^{\frac{1}{2p_0}} +C_5|\varOmega |^{\frac{p_0-2}{2p_0(p_0-1)}}\right) +2c_4M_4c_{10}\left( 1 +m_\infty +\mu \right) |\varOmega |^{\frac{1}{2p_0}}.$$

Proof

From (6.3.8), we have

$$\begin{aligned}&\quad \Vert \varDelta w(\cdot ,t)\Vert _{L^{p_0}(\varOmega )}\nonumber \\&\le \Vert e^{t\varDelta }\varDelta w_0\Vert _{L^{p_0}(\varOmega )}+\int _0^t\Vert e^{(t-s)\varDelta }\varDelta ((\lambda (u+v)+\mu )w)(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds \\&\le 2 c_1e^{-\lambda _1t}\Vert \varDelta w_0\Vert _{L^{p_0}(\varOmega )}\nonumber \\&\quad +\int _0^t\Vert e^{(t-s)\varDelta }\nabla \cdot (\lambda (\nabla u+\nabla v)w+(\lambda (u+v)+\mu )\nabla w)(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds\nonumber \\&\le 2 c_1e^{-\lambda _1t}\Vert \varDelta w_0\Vert _{L^{p_0}(\varOmega )}+\lambda \int _0^t\Vert e^{(t-s)\varDelta }\nabla \cdot ((\nabla u+\nabla v)w)(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds\nonumber \\&\quad +\int _0^t\Vert e^{(t-s)\varDelta }\nabla \cdot ((\lambda (u+v)+\mu )\nabla w)(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds.\nonumber \end{aligned}$$
(6.3.29)

From Lemma 1.1(i) and the fact that \(\Vert \varDelta w_0\Vert _{L^{p_0}(\varOmega )}\le \varepsilon \), we obtain that

$$\begin{aligned} \Vert e^{t\varDelta }\varDelta w_0\Vert _{L^{p_0}(\varOmega )}\le 2 c_1\varepsilon e^{-\lambda _1t}. \end{aligned}$$
(6.3.30)

Now, we estimate the last two integrals on the right-hand side of the above inequality. From the definition of T, Lemmas 1.1(iv), 4.3, 6.6, 6.8, 6.10 and (6.3.4), it follows that

$$\begin{aligned}&\lambda \int _0^t\Vert e^{(t-s)\varDelta }\nabla \cdot ((\nabla u(\cdot ,s)+\nabla v(\cdot ,s)) w(\cdot ,s))\Vert _{L^{p_0}(\varOmega )}ds\nonumber \\ \le&\,\,c_4\lambda \int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\Vert w(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}\Vert \nabla u(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds \\&+c_4\lambda \int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\Vert w(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}\Vert \nabla v(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds\nonumber \\ \le&\,\,c_4 M_2\varepsilon \lambda |\varOmega |^{\frac{1}{2p_0}} \int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)} e^{-2\alpha s}M_3\varepsilon ds\nonumber \\&+c_4 M_2\varepsilon \lambda |\varOmega |^{\frac{p_0-2}{2p_0(p_0-1)}} \int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)} e^{-\alpha s}C_5ds \nonumber \\ \le&\,\,2c_4M_2c_{10}\lambda \left( |\varOmega |^{\frac{1}{2p_0}} +C_5|\varOmega |^{\frac{p_0-2}{2p_0(p_0-1)}}\right) \varepsilon e^{-\alpha t}\nonumber \end{aligned}$$
(6.3.31)

and

$$\begin{aligned}&\int _0^t\Vert e^{(t-s)\varDelta }\nabla \cdot ((\lambda (u+v)+\mu )(\cdot ,s)\nabla w(\cdot ,s))\Vert _{L^{p_0}(\varOmega )}ds\nonumber \\ \le&\,\,c_4\int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)}\Vert \nabla w(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}\Vert (\lambda (u+v)+\mu )(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}ds\nonumber \\ \le&\,\,c_4 M_4\varepsilon \left( 1+m_\infty +\mu \right) |\varOmega |^{\frac{1}{2p_0}} \int _0^t(1+(t-s)^{-\frac{1}{2}})e^{-\lambda _1(t-s)} e^{-\alpha s}ds \\ \le&\,\, 2c_4M_4c_{10}\left( 1+m_\infty +\mu \right) |\varOmega |^{\frac{1}{2p_0}}\varepsilon e^{-\alpha t}.\nonumber \end{aligned}$$
(6.3.32)

Inserting (6.3.30), (6.3.31) and (6.3.32) into (6.3.29), we obtain

$$\begin{aligned} \Vert \varDelta w(\cdot ,t)\Vert _{L^{p_0}(\varOmega )}&\le M_6\varepsilon e^{-\alpha t} \end{aligned}$$

and thereby complete the proof.

Lemma 6.12

Under the assumptions of Theorem 6.1, for all \(p_0>\frac{N}{2}\),

$$\begin{aligned} \Vert \nabla u(\cdot ,t)\Vert _{L^{2p_0}(\varOmega )}\le \frac{M_3}{2}\varepsilon e^{-\alpha t}\quad \hbox {for all}~~ t\in (0,T). \end{aligned}$$
(6.3.33)

Proof

By (6.3.9), we have

$$\begin{aligned} \Vert \nabla u(\cdot ,t)\Vert _{L^{2p_0}(\varOmega )}&\le \Vert \nabla e^{t\varDelta }u_0(\cdot )\Vert _{L^{2p_0}(\varOmega )}+\chi _u\int _0^t\Vert \nabla e^{(t-s)\varDelta }\nabla \cdot (u\nabla w)(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds \\&\quad +\int _0^t\Vert \nabla e^{(t-s)\varDelta }(uw)(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds.\nonumber \end{aligned}$$
(6.3.34)

From Lemma 1.1(iii) and the fact that \(\Vert \nabla u_0(\cdot )\Vert _{L^{2p_0}(\varOmega )}\le \varepsilon \), we obtain

$$\begin{aligned} \Vert \nabla e^{t\varDelta }u_0(\cdot )\Vert _{L^{2p_0}(\varOmega )}\le 2c_3\varepsilon e^{-\lambda _1t}. \end{aligned}$$
(6.3.35)

Now, we estimate the last two integrals on the right-hand side of (6.3.34). From the definition of T, Lemmas 1.1(ii), 4.3, 6.6, 6.8 and 6.11, it follows that

$$\begin{aligned} \chi _u\int _0^t\Vert \nabla&e^{(t-s)\varDelta }\nabla \cdot (u\nabla w)(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds\nonumber \\ \le&\,\,c_2\chi _u\int _0^te^{-(t-s)\lambda _1}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) \Vert u(\cdot ,s)\varDelta w(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds \\&+c_2\chi _u\int _0^te^{-(t-s)\lambda _1}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) \Vert \nabla u(\cdot ,s)\nabla w(\cdot ,s)\Vert _{L^{p_0}(\varOmega )}ds\nonumber \\ \le&\,\,c_2\chi _u\int _0^te^{-(t-s)\lambda _1}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) M_6\varepsilon e^{-\alpha s}\frac{1+m_\infty }{\lambda }ds\nonumber \\&+c_2\chi _u\int _0^te^{-(t-s)\lambda _1}\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) M_4M_3\varepsilon ^2e^{-2\alpha s}ds\nonumber \\ \le&\,\, 2c_{10}c_2\chi _u\left( M_6\frac{1+m_\infty }{\lambda }+M_4M_3\varepsilon \right) \varepsilon e^{-\alpha t}\nonumber \end{aligned}$$
(6.3.36)

and

$$\begin{aligned} \int _0^t\Vert \nabla&e^{(t-s)\varDelta }uw\Vert _{L^{2p_0}(\varOmega )}ds\nonumber \\ \le&\,\,c_2|\varOmega |^{\frac{1}{2p_0}}\int _0^t\left( 1+(t-s)^{-\frac{1}{2}}\right) e^{-\lambda _1(t-s)}\Vert w\Vert _{L^{\infty }(\varOmega )}\Vert u\Vert _{L^{\infty }(\varOmega )}ds \\ \le&\,\,c_2 M_2\varepsilon |\varOmega |^{\frac{1}{2p_0}} \int _0^t\left( 1+(t-s)^{-\frac{1}{2}}\right) e^{-\lambda _1(t-s)} e^{-\alpha s}\frac{1+m_\infty }{\lambda }ds \nonumber \\ \le&\,\, 2c_{10}c_2 M_2\frac{1+m_\infty }{\lambda } \varepsilon |\varOmega |^{\frac{1}{2p_0}}e^{-\alpha t}.\nonumber \end{aligned}$$
(6.3.37)

Inserting (6.3.35)–(6.3.37) into (6.3.18) and using (6.3.3), we readily get

$$\begin{aligned} \Vert \nabla u\Vert _{L^{2p_0}(\varOmega )} \le&\,\,2c_3\varepsilon e^{-\lambda _1t}+2c_{10}c_2\left( \left( \chi _uM_6+M_2|\varOmega |^{\frac{1}{2p_0}}\right) \frac{1+m_\infty }{\lambda }+\chi _uM_4M_3\varepsilon \right) \varepsilon e^{-\alpha t} \\ \le&\,\,\frac{M_3}{2}\varepsilon e^{-\alpha t} \end{aligned}$$

and thereby complete the proof.

Lemma 6.13

Under the assumptions of Theorem 6.1, for all \(t\in (0,T)\),

$$\begin{aligned} \Vert (\lambda (u+v)+w+z)(\cdot ,t)-e^{t\varDelta }(\lambda (u_0+v_0)+w_0)\Vert _{L^\infty (\varOmega )}\le \frac{M_1}{2}\varepsilon e^{-\alpha t}. \end{aligned}$$

Proof

According to (6.3.7) and Lemma 1.1(iv), we have

$$\begin{aligned}&\Vert (\lambda (u+v)+w+z)(\cdot ,t)-e^{t\varDelta }(\lambda (u_0+v_0)+w_0)\Vert _{L^\infty (\varOmega )}\nonumber \\ \le&\,\,\lambda \int _0^t\Vert e^{(t-s)\varDelta }(\nabla \cdot (\chi _uu\nabla w+\chi _vv\nabla u))(\cdot ,s)\Vert _{L^\infty (\varOmega )}ds \\ \le&\,\,\lambda \chi _u\int _0^t\Vert e^{(t-s)\varDelta }\nabla \cdot (u\nabla w)(\cdot ,s)\Vert _{L^\infty (\varOmega )}ds +\lambda \chi _v\int _0^t\Vert e^{(t-s)\varDelta }\nabla \cdot (v\nabla u)(\cdot ,s)\Vert _{L^\infty (\varOmega )}ds\nonumber \\ \le&\,\,c_4\lambda \chi _u\int _0^t\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) e^{-\lambda _1(t-s)}\Vert u(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}\Vert \nabla w(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds\nonumber \\&+c_4\lambda \chi _v\int _0^t\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) e^{-\lambda _1(t-s)}\Vert v(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}\Vert \nabla u(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds\nonumber \\ =:&I_1+I_2.\nonumber \end{aligned}$$
(6.3.38)

Now, we need to estimate \(I_1\) and \(I_2\). Firstly, from the definition of T, Lemmas 4.3, 6.6 and 6.8, we obtain

$$\begin{aligned} I_1&\le c_4\chi _u(1+m_\infty )M_4\varepsilon \int _0^t\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) e^{-\lambda _1(t-s)} e^{-\alpha s}ds \\&\le 2c_{10}c_4\chi _u(1+m_\infty )M_4\varepsilon e^{-\alpha t}\nonumber \end{aligned}$$
(6.3.39)

and

$$\begin{aligned} I_2&\le c_4\lambda \chi _v\left( 1+m_\infty \right) M_3\varepsilon \int _0^t\left( 1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_0}}\right) e^{-\lambda _1(t-s)} e^{-\alpha s}ds \nonumber \\&\le 2c_{10}c_4\chi _v(1+m_\infty )M_3\varepsilon e^{-\alpha t}. \end{aligned}$$
(6.3.40)

Combining (6.3.38)–(6.3.40) along with (6.3.1) leads to

$$\begin{aligned}&\Vert (\lambda (u+v)+w+z)(\cdot ,t)-e^{t\varDelta }(\lambda (u_0+v_0)+w_0)\Vert _{L^\infty (\varOmega )} \\ \le&\,\,2c_{10}c_4(1+m_\infty )(\chi _uM_4+\chi _vM_3)\varepsilon e^{-\alpha t} \\ \le&\,\,\frac{1}{2}M_1\varepsilon e^{-\alpha t} \end{aligned}$$

and hence ends the proof.

Now, we have prepared the major parts of the proof of Theorem 6.1 and thus can verify asymptotic properties stated there.

Proof of the Theorem 6.1.  First we claim that \(T=T_{max}\). In fact, if \(T<T_{max}\), then by Lemmas 6.7, 6.12 and 6.13, we have

$$\begin{aligned}&\Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le \frac{M_2}{2} \varepsilon e^{-\alpha t},\\&\Vert \nabla u(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le \frac{M_3}{2}\varepsilon e^{-\alpha t}, \\&\Vert (\lambda (u+v)+w+z)(\cdot ,t)-e^{t\varDelta }(\lambda (u_0+v_0)+w_0)\Vert _{L^\infty (\varOmega )}\le \frac{M_1}{2}\varepsilon e^{-\alpha t} \end{aligned}$$

for all \(t\in (0,T)\), which contradicts the definition of T in (6.3.7).

Next, we show that \(T_{max}=\infty \). In fact, if \(T_{max}<\infty \), then in view of the definition of T, Lemmas 6.6, 6.8 and 6.10, we obtain that for any \(p_0>1+\frac{N}{2}\),

$$\lim _{t\rightarrow T_{max}}\left( \Vert u(\cdot ,t)\Vert _{W^{1,2p_0}( \varOmega )}+\Vert v(\cdot ,t)\Vert _{W^{1,2(p_0-1)}(\varOmega )}+\Vert w(\cdot ,t)\Vert _{W^{1,2p_0}( \varOmega )}\right) <\infty ,$$

which contradicts with Lemma 6.4. Therefore, we have \(T_{max}=\infty \).

Integrating the first equation in (6.1.4) over \(\varOmega \), we have

$$\int _\varOmega u(x,t)dx=\int _\varOmega u_0(x)dx+\int _0^t\int _\varOmega (uw)(x,s)dxds,$$

which, along with the nonnegative property of u, w and the fact that

$$\Vert u(\cdot ,t)\Vert _{L^\infty (\varOmega )}<\frac{1+m_\infty }{\lambda }, \quad \Vert w(\cdot ,t)\Vert _{L^\infty (\varOmega )}<M_2\varepsilon e^{-\alpha t},$$

warrants

$$\begin{aligned} \lim _{t\rightarrow \infty }\int _\varOmega u(x,t)dx=\int _\varOmega u_0(x)dx+\int _0^{+\infty }\int _\varOmega (uw)(x,s)dxds, \end{aligned}$$
(6.3.41)

as well as

$$\begin{aligned} \lim _{t\rightarrow \infty }\overline{u}(t)=\frac{1}{|\varOmega |}\left( \int _\varOmega u_0(x)dx+\int _0^{+\infty }\int _\varOmega (uw)(x,s)dxds\right) :=u^*. \end{aligned}$$
(6.3.42)

As a consequence of the latter, we immediately have

$$\begin{aligned} 0\le u^*-\overline{u}(t)&=\,\,\frac{1}{|\varOmega |}\int _t^{+\infty }\int _\varOmega (uw)(x,s)dxds\nonumber \\&\le \frac{1+m_\infty }{\lambda |\varOmega |}\int _t^{+\infty }\int _\varOmega w(x,s)dxds \\&\le \frac{1+m_\infty }{\alpha \lambda } M_2\varepsilon e^{-\alpha t}.\nonumber \end{aligned}$$
(6.3.43)

On the other hand, by Poincare’s inequality,

$$\begin{aligned} \Vert u-\overline{u}\Vert _{L^{2p_0}(\varOmega )}\le C_{1}\Vert \nabla u\Vert _{L^{2p_0}(\varOmega )}, \end{aligned}$$

and thanks to \(W^{1,2p_0}(\varOmega )\hookrightarrow L^{\infty }(\varOmega )\) for \(p_0>\frac{N}{2}\), we can find \(C_2>0\) such that

$$\begin{aligned} \Vert u-\overline{u}\Vert _{L^{\infty }(\varOmega )}\le C_2\Vert u-\overline{u}\Vert _{W^{1,2p_0}(\varOmega )}\le C_2(1+C_{1})\Vert \nabla u\Vert _{L^{2p_0}(\varOmega )}. \end{aligned}$$
(6.3.44)

Therefore, by (6.3.43) and the fact that \(\Vert \nabla u\Vert _{L^{2p_0}(\varOmega )}\le M_3\varepsilon e^{-\alpha t}\), we can pick \(K_1>0\) such that

$$\begin{aligned} \Vert u-u^*\Vert _{L^{\infty }(\varOmega )}\le&\,\,\Vert u-\overline{u}\Vert _{L^{\infty }(\varOmega )}+\Vert \overline{u}-u^*\Vert _{L^{\infty }(\varOmega )}\nonumber \\ \le&\,\,C_2(1+C_1)\Vert \nabla u\Vert _{L^{2p_0}(\varOmega )}+|\overline{u}(t)-u^*| \\ \le&\,\,K_1\varepsilon e^{-\alpha t}.\nonumber \end{aligned}$$
(6.3.45)

On the other hand, from (6.3.12) and Lemma 1.1(ii), we infer that

$$\begin{aligned} \Vert \nabla z(\cdot ,t)\Vert _{L^{2p_0}}&=\,\,\mu \int _0^t\Vert \nabla e^{(t-s)\varDelta } w(\cdot ,s)\Vert _{L^{2p_0}(\varOmega )}ds\nonumber \\&\le \mu c_2|\varOmega |^{\frac{1}{2p_0}}\int _0^te^{-(t-s)\lambda _1} (1+t^{-\frac{1}{2}})\Vert w(\cdot ,s)\Vert _{L^{\infty }(\varOmega )}ds \\&\le 2\mu c_2c_{10}M_2|\varOmega |^{\frac{1}{2p_0}}\varepsilon e^{-\alpha t}.\nonumber \end{aligned}$$
(6.3.46)

By similar procedure as that in the derivation of (6.3.45), there exists constant \(K_2>0\) such that

$$\begin{aligned} \Vert z-z^*\Vert _{L^{\infty }(\varOmega )}\le K_2\varepsilon e^{-\alpha t} \end{aligned}$$
(6.3.47)

with

$$\begin{aligned} z^*:=\frac{\mu }{{|\varOmega |}}\int _0^{+\infty }\int _\varOmega w(x,s)dxds. \end{aligned}$$
(6.3.48)

Then from the fact that

$$\begin{aligned} \lambda \Vert v-v^*\Vert _{L^\infty (\varOmega )}\le&\,\,\Vert (\lambda u+\lambda v+w+z)-m_\infty \Vert _{L^\infty (\varOmega )}+\lambda \Vert u-u^*\Vert _{L^\infty (\varOmega )} \\&+\Vert w\Vert _{L^\infty (\varOmega )}+\Vert z-z^*\Vert _{L^\infty (\varOmega )} \end{aligned}$$

with

$$\begin{aligned} v^*:=\frac{1}{\lambda }\left( m_{\infty }-z^*\right) -u^*, \end{aligned}$$
(6.3.49)

using (6.3.13), (6.3.14), (6.3.45) and (6.3.47), there exists \(K_3>0\) such that

$$\begin{aligned} \Vert v-v^*\Vert _{L^\infty (\varOmega )}\le K_3\varepsilon e^{-\alpha t}. \end{aligned}$$

The decay estimates claimed in Theorem 6.1 readily follow and the proof of this theorem is thus completed.

6.4 Boundedness of Solutions to an Oncolytic Virotherapy Model

6.4.1 Some Basic a Prior Estimates

For the convenience in our subsequent estimation procedure, we let

$$\chi _u:=\frac{\xi _u}{D_u}, \quad \chi _w:=\frac{\xi _w}{D_w}~~ \hbox {and}~~ \chi _z:=\frac{\xi _z}{D_z} $$

and introduce the variable change used in several precedents (Fontelos et al. 2002; Pang and Wang 2018; Tao and Winkler 2014b)

$$a= ue^{-\chi _u v}\quad b= we^{-\chi _w v} ~~\hbox {and} ~~c= ze^{-\chi _z v} $$

upon which (6.1.10) takes the following form:

$$\begin{aligned} \left\{ \begin{aligned}&a_t=D_u e^{-\chi _u v}\nabla \cdot (e^{\chi _u v}\nabla a)+f(a,b,v,c),&x\in \varOmega ,t>0,\\&b_t=D_w e^{-\chi _w v}\nabla \cdot (e^{\chi _w v}\nabla b)+g(a,b,v,c),&x\in \varOmega ,t>0,\\&v_t=- (\alpha _u ae^{\chi _u v} + \alpha _w be^{\chi _w v})v+\mu _v v(1-v),&x\in \varOmega ,t>0,\\&c_t=D_z e^{-\chi _z v}\nabla \cdot (e^{\chi _z v}\nabla c)+h(a,b,v,c),&x\in \varOmega ,t>0,\\&\displaystyle \frac{ \partial a}{\partial \nu }=\frac{ \partial b}{\partial \nu }=\frac{ \partial c}{\partial \nu }=0,&x\in \partial \varOmega ,t>0,\\&a(x,0)=u_0(x) e^{-\chi _u v_0(x)}, ~b(x,0)=w_0(x) e^{-\chi _w v_0(x)},&x\in \varOmega ,\\&v_0(x,0)= v_0(x),~~ c(x,0)=z_0(x) e^{-\chi _z v_0(x)},&x\in \varOmega \end{aligned} \right. \end{aligned}$$
(6.4.1)

with

$$\begin{aligned} f(a,b,v,c):=&\,\,\mu _u a(1-a^re^{r\chi _u v})-\displaystyle \frac{\rho ace^{\chi _z v}}{k_u +\theta ae^{\chi _u v}}+\chi _u a(\alpha _u ae^{\chi _u v} + \alpha _w be^{\chi _w v})v \\&-\chi _u\mu _v av(1-v) \\ g(a,b,v,c):=&\,\,-\delta _w b+\displaystyle \frac{\rho ace^{(\chi _u+\chi _z-\chi _w)v}}{k_u +\theta ae^{\chi _u v}}+ \chi _w b(\alpha _u ae^{\chi _u v} + \alpha _w be^{\chi _w v})v \\&- \chi _w\mu _v bv(1-v) \\ h(a,b,v,c):=&\,\,-\delta _z c-\displaystyle \frac{\rho ace^{\chi _u v}}{k_u + \theta ae^{\chi _u v}} +\beta be^{(\chi _w-\chi _z)v} +\chi _z c(\alpha _u ae^{\chi _u v} + \alpha _w be^{\chi _w v})v \\&-\chi _z\mu _v cv(1-v). \end{aligned}$$

It is noted that (6.1.10) and (6.4.1) are equivalent in this framework of classical solutions. The following basic statement on the local existence and extensibility criterion of classical solutions to (6.4.1) can be proved by a straightforward adaptation of the reasoning in Pang and Wang (2018) and Tao and Winkler (2020b).

Lemma 6.14

Let \(D_u,D_w\), \(D_z\), \(\xi _u,\xi _w, \xi _z,\mu _u,\mu _v,\rho ,k_u,\alpha _u,\alpha _w,\beta ,\delta _w\) and \(\delta _z \) are positive parameters, and assume that \(r\ge 1,\theta \ge 0\). Then there exist \(T_{max}\in (0,\infty ]\) and a uniquely determined quadruple \((a,b,v,c)\in (C^{2,1}(\overline{\varOmega } \times [0,T_{max})))^4\) which solves (6.4.1) in the classical sense and \(a>0,b>0,c>0\) and \(v>0\) in \(\varOmega \times (0,T_{max})\), and that if \(T_{max}<+\infty \), then

$$\begin{aligned} \Vert a(\cdot , t)\Vert _{L^\infty (\varOmega )}+\Vert b(\cdot , t)\Vert _{L^\infty (\varOmega )}+\Vert \nabla v(\cdot ,t)\Vert _{L^4(\varOmega )} +\Vert c(\cdot , t)\Vert _{L^\infty (\varOmega )}\rightarrow \infty ~~ \text{ as }~~ t\nearrow T_{max}. \end{aligned}$$
(6.4.2)

Proof

Invoking well-established fixed point arguments and applying the standard parabolic regularity theory, one can readily verify the local existence and uniqueness of classical solutions, as well as the extensibility criterion (6.4.2) (cf. Pang and Wang 2018; Tao and Winkler 2014b for instance). With the help of the maximum principle, we can also verify the asserted positivity of the solutions.

From now on without any further explicit mentioning, we shall suppose that the assumptions of Theorem 6.2 are satisfied, and let (a, b, v, c) and \(T_{max}\in (0,\infty ]\) be as provided by Lemma 6.14. Moreover, we may tacitly switch between these variables and the quadruple (u, w, v, z) if necessary.

The following important properties of solutions of (6.1.10) can be easily checked.

Lemma 6.15

Let \(T>0\). Then solution (u, w, v, z) of (6.1.10) satisfies

$$\begin{aligned} \displaystyle \Vert u(\cdot ,t)\Vert _{L^1(\varOmega )}\le u^*:=\max \{|\varOmega |,\Vert u_0\Vert _{L^1(\varOmega )}\} ~~\quad \hbox {for all}~ t\in (0, \hat{T}), \end{aligned}$$
(6.4.3)
$$\begin{aligned} \displaystyle \Vert v(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le v^*:=\max \{1,\Vert v_0\Vert _{L^\infty (\varOmega )}\} ~~\quad \hbox {for all}~ t\in (0, \hat{T}) \end{aligned}$$
(6.4.4)

and

$$\begin{aligned} \Vert w(\cdot ,t)\Vert _{L^1(\varOmega )} \le w^*:= \max \left\{ \Vert u_0\Vert _{L^1(\varOmega )}+\Vert w_0\Vert _{L^1(\varOmega )}, \frac{ 4 \mu _u |\varOmega |}{\min \{\mu _u, \delta _w\}}\right\} ~ \end{aligned}$$
(6.4.5)

for all \(t\in (0, \hat{T})\) as well as

$$\begin{aligned} \Vert z(\cdot ,t)\Vert _{L^1(\varOmega )}\le z^*:= \max \left\{ \Vert z_0\Vert _{L^1(\varOmega )}, \frac{\beta w^*}{\delta _z}\right\} ~~~~~~\hbox {for all}~ t\in (0, \hat{T}) \end{aligned}$$
(6.4.6)

where \(\hat{T}:=\min \{T, T_{max}\}\).

Proof

Integrating the first equation in (6.1.10) over \(\varOmega \) yields

$$\begin{aligned} \displaystyle {\frac{d}{dt}} \int _{\varOmega }u \le \mu _u \int _{\varOmega }u-\mu _u\int _{\varOmega }u^{r+1} \end{aligned}$$
(6.4.7)

due to \(z\ge 0\). Since \( (\int _{\varOmega }u)^{r+1}\le |\varOmega |^{r}\int _{\varOmega }u^{r+1}\) by the Cauchy–Schwartz inequality, (6.4.7) implies that \(y(t):=\int _{\varOmega }u(\cdot ,t)\) satisfies

$$ y'(t)\le \mu _u y(t)- \frac{\mu _u}{|\varOmega |^r }y^{1+r}(t)~~\hbox {for all}~(0, \hat{T}), $$

from which (6.4.3) follows by the Bernoulli inequality. On the other hand, due to the nonnegativity of u, w and v in \(\overline{\varOmega }\times (0, \hat{T})\), the comparison principle entails that \(v_t\le \mu _v v(1-v)\) and thus the estimate in (6.4.4) follows similarly.

Once more integrating the equations in (6.1.10) over \(\varOmega \) and using the fact that \(2u\le u^{r+1}+4 \), we can see that

$$\begin{aligned} \displaystyle {\frac{d}{dt}} \int _{\varOmega }u+\mu _u \int _{\varOmega }u \le 4\mu _u |\varOmega |- \rho \int _{\varOmega } \displaystyle \frac{ uz}{k_u +\theta u} \end{aligned}$$
(6.4.8)

and

$$\begin{aligned} \displaystyle {\frac{d}{dt}} \int _{\varOmega }w+\delta _w \int _{\varOmega } w \le \rho \int _{\varOmega } \displaystyle \frac{ uz}{k_u +\theta u} \end{aligned}$$
(6.4.9)

as well as

$$\begin{aligned} \displaystyle {\frac{d}{dt}} \int _{\varOmega }z+\delta _z \int _{\varOmega } z \le -\rho \int _{\varOmega } \displaystyle \frac{ uz}{k_u +\theta u}+\beta \int _{\varOmega }w. \end{aligned}$$
(6.4.10)

Combining (6.4.8)–(6.4.9), we obtain that

$$\begin{aligned} \displaystyle {\frac{d}{dt}} \left( \int _{\varOmega }u+\int _{\varOmega } w\right) + \mu _u \int _{\varOmega }u + \delta _w\int _{\varOmega }w \le 4 \mu _u |\varOmega |, \end{aligned}$$

which entails that \(y(t):=\int _{\varOmega }u+ \int _{\varOmega } w\) satisfies

$$ y'(t)+\min \{\mu _u, \delta _w\}y(t)\le 4 \mu _u |\varOmega |. $$

Hence, using the Bernoulli inequality to the above inequality, we get the estimate in (6.4.5). Further, it follows from (6.4.10) that

$$\begin{aligned} \displaystyle {\frac{d}{dt}} \int _{\varOmega }z + \delta _z\int _{\varOmega }z\le \beta w^* \end{aligned}$$

and thereby derive (6.4.6) by an ODE comparison argument.

6.4.2 Bounds for a, b and c in LlogL

This section aims to construct an Lyapunov-like functional involving the logarithmic entropy of a, b and c, rather than that of u, w and z, which provides some regularity information of solutions that forms the crucial step in establishing \(L^\infty \) bounds for u, w and z in the present spatially two-dimensional setting. It should be mentioned that upon the special structure of (6.1.9), inter alia neglecting haptotactic migration processes of oncolytic viruses z, the energy-like functional \(\mathscr {F}\) in Tao and Winkler (2020b) can be achieved by appropriately combining the logarithmic entropy of u, w, Dirichlet integral of \(\sqrt{v}\) and integral of \(z^2\) in line with some precedent studies (see Tao and Winkler 2014b; Winkler 2018b).

The first step of our approaches consists in testing the first equation of (6.4.1) against \(\ln a\).

Lemma 6.16

For any \(\varepsilon \in (0,1)\), there exists \(K_1(\varepsilon )>0\) such that

$$\begin{aligned} \displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _u v}a\ln a + D_u\displaystyle \int _\varOmega e^{\chi _u v}\displaystyle \frac{|\nabla a|^2}{a}+ \frac{\mu _u}{2} \int _\varOmega (\ln a+1)u^{r+1} \le \varepsilon \int _\varOmega w^2 +K_1(\varepsilon ). \end{aligned}$$
(6.4.11)

Proof

From the first equation in (6.4.1), it follows

$$\begin{aligned} (ae ^{\chi _u v})_t=D_u\nabla \cdot (e^{\chi _u v}\nabla a)+ \mu _u u(1-u^r)-\displaystyle \frac{\rho uz}{k_u +\theta u}. \end{aligned}$$

By the positivity of a in \(\overline{\varOmega }\times (0,\infty )\), testing the first equation in (6.4.1) by \(\ln a\) then shows that

$$\begin{aligned}&\quad \displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _u v}a\ln a\nonumber \\&=\,\, \displaystyle \int _\varOmega (e^{\chi _u v}a)_t\ln a+\int _\varOmega e^{\chi _u v}a_t \nonumber \\&=\,\,- D_u\displaystyle \int _\varOmega e^{\chi _u v}\displaystyle \frac{|\nabla a|^2}{a}+ \int _\varOmega \left( \mu _u u(1-u^r)-\displaystyle \frac{\rho uz}{k_u +\theta u}\right) \ln a+ \int _\varOmega f(a,b,v,z) e^{\chi _u v} \nonumber \\[1mm]&=\,\,- D_u\displaystyle \int _\varOmega e^{\chi _u v}\displaystyle \frac{|\nabla a|^2}{a}+ \int _\varOmega (\ln a+1) \left( \mu _u u(1-u^r)-\displaystyle \frac{\rho uz}{k_u + \theta u}\right) \\[1mm]&\qquad +\chi _u \displaystyle \int _\varOmega u(\alpha _u u + \alpha _w w )v- \chi _u\mu _v \displaystyle \int _\varOmega u v(1-v)\nonumber \\[1mm]&\le - D_u\displaystyle \int _\varOmega e^{\chi _u v}\displaystyle \frac{|\nabla a|^2}{a}- \mu _u\int _\varOmega (\ln a+1)(u^{r+1}-u) -\rho \int _\varOmega \displaystyle \frac{ uz}{k_u +\theta u}\ln a\nonumber \\[1mm]&\qquad +\chi _u \displaystyle \int _\varOmega u(\alpha _u u + \alpha _w w )v+ \chi _u\mu _v \displaystyle \int _\varOmega u v^2. \nonumber \end{aligned}$$
(6.4.12)

By (6.4.6), we see that

$$ -\rho \int _\varOmega \displaystyle \frac{ uz}{k_u + \theta u}\ln a= -\rho \int _\varOmega \displaystyle \frac{z e^{\chi _u v} }{k_u + \theta u} a \ln a\le \frac{\rho e^{\chi _u v^*}}{k_u e} \int _\varOmega z\le \frac{\rho e^{\chi _u v^*}}{k_u e}z^*, $$

since \(a\ln a\ge -e^{-1}\) for all \(a>0\) and \(v(x,t)\le v^*\) for all \(x\in \varOmega ,t>0\) by (6.4.4). Apart from that, for any \(\varepsilon \in (0,1)\), there exists \(C_1(\varepsilon )>0\) such that

$$\begin{aligned} \begin{aligned}&\mu _u\displaystyle \int _\varOmega (\ln a+1)u+\chi _u \displaystyle \int _\varOmega u(\alpha _u u + \alpha _w w )v+ \chi _u\mu _v \displaystyle \int _\varOmega u v^2\\ \le&\,\, \displaystyle \frac{\mu _u}{2} \int _\varOmega (\ln a+1)u^2 + \varepsilon \int _\varOmega w^2 +C_1(\varepsilon ) \end{aligned} \end{aligned}$$

due to \(a^2\le \varepsilon _1 a^2\ln a+e^{\frac{2}{\varepsilon _1}}\), \(a\ln a\le \varepsilon _1 a^2\ln a- \varepsilon _1^{-1} \ln \varepsilon _1 \) and \(a\le \varepsilon _1 a^2\ln a+2 e^{\frac{2}{\varepsilon _1}}\) for any \(\varepsilon _1\in (0,1)\). Therefore, inserting above two inequalities into (6.4.12), we arrive at (6.4.11).

Lemma 6.17

There exists \(c^*>0\) with the property that if \(\xi _w\alpha _w<c^*\), then one can find \(\varepsilon _0\in (0,1)\) and \(K_2>0\) such that for all \(\varepsilon \in (0,\varepsilon _0)\),

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _w v}b\ln b +\frac{D_w}{4}\displaystyle \int _\varOmega e^{\chi _w v} \displaystyle \frac{|\nabla b|^2}{b}+ \delta _w \int _\varOmega e^{\chi _w v}b\ln b\\ \le&\,\,\displaystyle \varepsilon \Vert c\Vert ^2_{L^2(\varOmega )} +\frac{K_2}{\varepsilon }\Vert a\Vert ^2_{L^{r+1}(\varOmega )}+\frac{K_2}{\varepsilon }, \end{aligned} \end{aligned}$$
(6.4.13)

where \(r=1\) if \(\theta >0\) and \(r>1\) if \(\theta \ge 0\).

Proof

From the second equation in (6.4.1), it follows that

$$\begin{aligned} (be ^{\chi _w v})_t=D_w\nabla \cdot (e^{\chi _w v}\nabla b)- \delta _w w+\displaystyle \frac{\rho uz}{k_u + \theta u}. \end{aligned}$$

By straightforward calculation relying on \(0\le v\le 1\) in \(\overline{\varOmega }\times (0,\infty )\) and the Young inequality, we then see that for any \(\varepsilon >0\),

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _w v}b\ln b+\delta _w \int _\varOmega e^{\chi _w v}b\ln b+ D_w\displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b} \\ =&\,\,\int _\varOmega (\ln b+1) \left( \displaystyle \frac{\rho uz}{k_u + \theta u}-\delta _w w\right) +\chi _w \displaystyle \int _\varOmega w(\alpha _u u + \alpha _w w )v- \chi _w\mu _v \displaystyle \int _\varOmega w v(1-v)\nonumber \\&+\delta _w \int _\varOmega w\ln b \nonumber \\[1mm] \le&\,\, \rho \int _\varOmega \displaystyle \frac{ uz}{k_u + \theta u} \ln b +\rho \int _\varOmega \displaystyle \frac{ uz}{k_u + \theta u} + \chi _w\mu _v v^*\int _\varOmega w+\chi _w\alpha _u v^*\int _\varOmega wu +\chi _w\alpha _w v^*\int _\varOmega w^2. \nonumber \end{aligned}$$
(6.4.14)

The first summand on the right-hand side of (6.4.14) will be estimated in the case \(r=1, \theta >0\) and \(r>1, \theta \ge 0\), respectively.

For \(r=1\) and \(\theta > 0\), we have

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _w v}b\ln b+\delta _w \int _\varOmega e^{\chi _w v}b\ln b+ D_w\displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b} \\ \le&\,\, \frac{\rho }{\theta } \int _{b>1} z \ln b +(\varepsilon +\chi _w\alpha _w v^*)\int _\varOmega w^2+ \frac{\rho }{\theta }\int _\varOmega z+\frac{\chi _w^2\alpha _u^2v^{*2}}{\varepsilon }\int _\varOmega u^2 + \chi _w\mu _v v^*\int _\varOmega w. \nonumber \end{aligned}$$
(6.4.15)

Since \(\ln ^{2} s\le \frac{4}{ e^2}s \) for all \(s>1\), an application of the Hölder inequality leads to

$$\begin{aligned} \displaystyle \frac{\rho }{\theta } \int _{b>1} z \ln b \le \displaystyle \varepsilon \Vert z\Vert ^2_{L^2(\varOmega )}+\displaystyle \frac{\rho ^2}{\varepsilon \theta ^2} \displaystyle \int _\varOmega b. \end{aligned}$$

In conjunction with (6.4.15), we can see that

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _w v}b\ln b+ D_w\displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b}+\delta _w \int _\varOmega e^{\chi _w v}b\ln b \\ \le&\,\,(\varepsilon +\chi _w\alpha _w v^*)\Vert w\Vert ^2_{L^2(\varOmega )}+ \varepsilon \Vert z\Vert ^2_{L^2(\varOmega )} + \frac{\chi _w^2\alpha _u^2v^{*2}}{\varepsilon }\Vert u\Vert ^2_{L^2(\varOmega )}\nonumber \\&+ (\chi _w\mu _v v^*+\displaystyle \frac{\rho ^2}{\varepsilon \theta ^2}) \Vert w\Vert _{L^1(\varOmega )}+\frac{\rho }{\theta }\Vert z\Vert _{L^1(\varOmega )}. \nonumber \end{aligned}$$
(6.4.16)

To estimate the first term on the right-hand side of (6.4.16), by means of the two-dimensional Gagliardo–Nirenberg inequalities, we can find \(K_g>0\) such that

$$\begin{aligned} \Vert \varphi \Vert ^4_{L^4(\varOmega )} \le K_{g}\Vert \nabla \varphi \Vert _{L^2(\varOmega )}^2\Vert \varphi \Vert ^2_{L^2(\varOmega )}+K_g\Vert \varphi \Vert ^4_{L^2(\varOmega )} ~~\hbox {for all}~ \varphi \in W^{1,2}(\varOmega ). \end{aligned}$$
(6.4.17)

Thereby thanks to (6.4.4) and (6.4.5), there exists \(C_1>0\) such that

$$\begin{aligned} (\varepsilon +\chi _w\alpha _wv^*)\Vert w\Vert ^2_{L^2(\varOmega )}\le&e^{2\chi _w v^*}(\varepsilon +\chi _w\alpha _w v^*)\Vert \sqrt{b}\Vert ^4_{L^4(\varOmega )} \nonumber \\ \le&\,\, \displaystyle \frac{e^{2v^*\chi _w }K_g(\varepsilon +\chi _w\alpha _w v^*)}{4}\displaystyle \int _\varOmega b \displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b}+C_1 \nonumber \\ \le&\,\, \displaystyle \frac{e^{2\chi _w v^*}K_g(\varepsilon +\chi _w\alpha _w v^*)}{4} w^* \cdot \displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b}+C_1 \\ \le&\,\, \displaystyle \frac{3 D_w}{4}\displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b}+C_1, \nonumber \end{aligned}$$
(6.4.18)

provided that \(\xi _w\alpha _w<c^*:=\frac{2 D_w}{e^{2\chi _w v^* }K_g w^*v^*}\) and any \(0<\varepsilon <\varepsilon _{0}:=\min \{1, \frac{D_w }{e^{2\chi _w v^* }K_g w^*v^*}\}\). Therefore, along with Lemma 6.15 and the Hölder inequality, we insert (6.4.18) into (6.4.16) to arrive at (6.4.13).

While for \(r>1, \theta \ge 0\), we have

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _w v}b\ln b+\delta _w \int _\varOmega e^{\chi _w v}b\ln b+ D_w\displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b} \\ \le&\,\, \frac{\rho }{k_u}\int _\varOmega z u\ln b +(\varepsilon +\chi _w\alpha _w v^*)\int _\varOmega w^2+ \frac{\rho }{k_u} \int _\varOmega u z+\frac{\chi _w^2\alpha _u^2 v^{*2}}{\varepsilon }\int _\varOmega u^2 + \chi _w\mu _v v^*\int _\varOmega w. \nonumber \end{aligned}$$
(6.4.19)

Here, an apparently challenging issue is to estimate \( \int _\varOmega z u\ln b\) appropriately in terms of expression which can be controlled by the dissipation terms in (6.4.11). Since there exists \(C_2>0\) such that \(\ln ^\frac{2(r+1)}{r-1} s\le s+C_2\) for all \(s>1\), we can infer from (6.4.6) and the Hölder inequality that

$$\begin{aligned} \frac{\rho }{k_u} \displaystyle \int _\varOmega z u\ln b \le&\,\, \frac{\rho }{k_u} \displaystyle \int _{\{b>1\}} z u\ln b \\&\le \frac{\rho }{k_u} \Vert u\Vert _{L^{r+1}(\varOmega )} \Vert z\Vert _{L^2(\varOmega )} \left\{ \displaystyle \int _{\{b>1\}} (\ln b)^{\frac{2(r+1)}{r-1}} \right\} ^{\frac{r-1}{2(r+1)}} \nonumber \\&\le \displaystyle \frac{\rho }{k_u} \Vert u\Vert _{L^{r+1}(\varOmega )} \Vert z\Vert _{L^2(\varOmega )}(\Vert b \Vert _{L^1(\varOmega )}+C_2|\varOmega |)^{\frac{r-1}{2(r+1)}} \nonumber \\&\le \displaystyle \frac{\varepsilon }{2} \Vert z\Vert ^2_{L^2(\varOmega )}+\displaystyle \frac{C_3}{\varepsilon } \Vert u\Vert ^2_{L^{r+1}(\varOmega )} \nonumber \end{aligned}$$

with \(C_3=\frac{\rho ^2}{k_u^2} (\Vert b \Vert _{L^1(\varOmega )}+C_2|\varOmega |)^{\frac{r-1}{r+1}}\).

Apart from that, by the Hölder inequality and the Young inequality, it is easy to see that

$$\begin{aligned} \frac{\rho }{k_u} \displaystyle \int _\varOmega z u&\le \frac{\rho }{k_u} \Vert u\Vert _{L^{2}(\varOmega )} \Vert z\Vert _{L^2(\varOmega )} \nonumber \\&\le \displaystyle \frac{\varepsilon }{2} \Vert z\Vert ^2_{L^2(\varOmega )}+\displaystyle \displaystyle \frac{\rho ^2}{2 k_u^2\varepsilon } \Vert u\Vert ^2_{L^{2}(\varOmega )}. \nonumber \end{aligned}$$

In conjunction with (6.4.19), we get

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _w v}b\ln b+ D_w\displaystyle \int _\varOmega e^{\chi _w v}\displaystyle \frac{|\nabla b|^2}{b}+\delta _w \int _\varOmega e^{\chi _w v}b\ln b \\ \le&\,\,(\varepsilon +\chi _w\alpha _w v^*)\Vert w\Vert ^2_{L^2(\varOmega )}+ \varepsilon \Vert z\Vert ^2_{L^2(\varOmega )}+\displaystyle \frac{C_3}{\varepsilon } \Vert u\Vert ^2_{L^{r+1}(\varOmega )} \\&+ \left( \frac{\chi _w^2\alpha _u^2v^{*2}}{\varepsilon } +\frac{\rho ^2}{2 k_u^2\varepsilon }\right) \Vert u\Vert ^2_{L^{2}(\varOmega )}+ \chi _w\mu _vv^*\Vert w\Vert _{L^1(\varOmega )} \\ \le&\,\,(\varepsilon +\chi _w\alpha _wv^*)\Vert w\Vert ^2_{L^2(\varOmega )}+ \varepsilon \Vert z\Vert ^2_{L^2(\varOmega )} \\&+\left( \displaystyle \frac{C_3+\chi _w^2\alpha _u^2v^{*2}}{\varepsilon }+\frac{\rho ^2}{2 k_u^2\varepsilon }\right) \Vert u\Vert ^2_{L^{r+1}(\varOmega )} + \chi _w\mu _vv^*\Vert w\Vert _{L^1(\varOmega )} +C_4 \end{aligned}$$

for some \(C_4>0\), which together with (6.4.18) and (6.4.5) implies that (6.4.13) holds.

Lemma 6.18

There exists \(K_3>0\) such that

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _z v}c\ln c +\frac{ 3 D_z}{4}\displaystyle \int _\varOmega e^{\chi _z v} \displaystyle \frac{|\nabla c|^2}{c}+ \delta _z \int _\varOmega e^{\chi _z v}c\ln c\\ \le&\,\,\displaystyle K_3\Vert b\Vert ^2_{L^2(\varOmega )} +K_3\Vert a\Vert ^2_{L^2(\varOmega )}+K_3. \end{aligned} \end{aligned}$$
(6.4.20)

Proof

By the fourth equation in (6.4.1), we can see that

$$\begin{aligned} (c e ^{\chi _z v})_t=D_z\nabla \cdot (e^{\chi _z v}\nabla c)- \delta _z z-\displaystyle \frac{\rho uz}{k_u + \theta u}+ \beta w. \end{aligned}$$
(6.4.21)

Proceeding as above, we test (6.4.21) by \(\ln c\) and integrate by parts to see that for \(\varepsilon >0\),

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _z v}c\ln c+\delta _z \int _\varOmega e^{\chi _z v}c\ln c+ D_z\displaystyle \int _\varOmega e^{\chi _z v}\displaystyle \frac{|\nabla c|^2}{c} \\ \le&\,\,\int _\varOmega (\ln c+1) \left( \beta w-\displaystyle \frac{\rho uz}{k_u + \theta u}\right) +\chi _z \displaystyle \int _\varOmega z(\alpha _u u + \alpha _w w )v- \chi _z\mu _v \displaystyle \int _\varOmega c v(1-v) \nonumber \\[1mm] \le&\,\, -\rho \int _\varOmega \displaystyle \frac{ uz}{k_u + \theta u} \ln c + \beta \int _\varOmega w+ \beta \int _\varOmega w \ln c+\chi _z\alpha _u v^*\int _\varOmega cu \nonumber \\&+\chi _z\alpha _w v^*\int _\varOmega cw+\chi _z\mu _v \displaystyle \int _\varOmega c \nonumber \\ \le&\,\, \frac{\rho e^{\chi _z v^*}}{k_u} u^* +\varepsilon \int _\varOmega c^2+\frac{\chi _z^2\alpha _u^2v^{*2}}{\varepsilon }\int _\varOmega u^2 +\left( \frac{\chi _z^2\alpha _w^2v^{*2}}{\varepsilon }+1\right) \int _\varOmega w^2 \nonumber \\&+ (\chi _z\mu _v+\beta ^2)\int _\varOmega c+ \beta \int _\varOmega w \nonumber \end{aligned}$$
(6.4.22)

due to

$$\begin{aligned} -\rho \int _\varOmega \displaystyle \frac{ uz}{k_u + \theta u}\ln c&=\,\, -\rho \int _\varOmega \displaystyle \frac{u e^{\chi _z v} }{k_u + \theta u} c \ln c\\&\le -\rho \int _{\{c<1\}}\displaystyle \frac{u e^{\chi _z v} }{k_u + \theta u} c \ln c\\&\le \frac{\rho e^{\chi _z v^*}}{k_u}u^* \end{aligned}$$

and

$$\begin{aligned} \displaystyle \beta \int _{\{c>1\}} w \ln c&\le \displaystyle \Vert w\Vert ^2_{L^2(\varOmega )}+\frac{ \beta ^2}{4} \displaystyle \int _{\{c>1\}} \ln ^2c\\&\le \displaystyle \Vert w\Vert ^2_{L^2(\varOmega )}+ \beta ^2 \displaystyle \int _\varOmega c. \end{aligned}$$

Now according to the two-dimensional Gagliardo–Nirenberg inequality (6.4.17) and Lemma 6.15, we pick \(\varepsilon =\frac{D_z}{8K_gz^*}\) and thereafter obtain some \(C_1>0\) such that

$$\begin{aligned} \begin{aligned} \varepsilon \Vert c\Vert ^2_{L^2(\varOmega )}=&\varepsilon \Vert \sqrt{c}\Vert ^4_{L^4(\varOmega )}\\ \le&\,\, \left( \displaystyle K_g\varepsilon \displaystyle \int _\varOmega c\right) \displaystyle \int _\varOmega \displaystyle \frac{|\nabla c|^2}{c}+\varepsilon K_g \left( \int _\varOmega c\right) ^2\\ \le&\,\, \displaystyle \frac{D_z}{4}\displaystyle \int _\varOmega e^{\chi _z v}\displaystyle \frac{|\nabla c|^2}{c}+C_1. \end{aligned} \end{aligned}$$
(6.4.23)

Therefore, along with Lemma 6.15, in conjunction with (6.4.22) and (6.4.23), we readily arrive at (6.4.20).

We are now ready to obtain the bounds for a, b and c in LlogL by taking suitable linear combinations of the inequalities provided by Lemmata 6.16–6.18, stated as follows.

Lemma 6.19

Let \(T>0\). Then there exists \(K_4> 0\) such that

$$\begin{aligned} \displaystyle \int _\varOmega a(\cdot ,t)|\ln a(\cdot ,t)|\le K_4, \end{aligned}$$
(6.4.24)
$$\begin{aligned} \displaystyle \int _\varOmega b(\cdot ,t)|\ln b(\cdot ,t)|\le K_4 \end{aligned}$$
(6.4.25)

and

$$\begin{aligned} \displaystyle \int _\varOmega c(\cdot ,t)|\ln c(\cdot ,t)|\le K_4 \end{aligned}$$
(6.4.26)

for all \(t\in (0, \hat{T})\) with \(\hat{T}:=\min \{T, T_{max}\}\).

Proof

From (6.4.18) and (6.4.23), it follows that there exists \(C_1>0\) such that

$$\begin{aligned} \Vert b\Vert ^2_{L^2(\varOmega )}\le C_1\displaystyle \int _\varOmega \displaystyle \frac{|\nabla b|^2}{b}+C_1 \end{aligned}$$
(6.4.27)
$$\begin{aligned} \Vert c\Vert ^2_{L^2(\varOmega )}\le C_1\displaystyle \int _\varOmega \displaystyle \frac{|\nabla c|^2}{c}+C_1 . \end{aligned}$$
(6.4.28)

Multiplying (6.4.13) by \(A:=\frac{8K_3C_1}{D_w}\) and adding the resulting inequality to (6.4.20), using (6.4.27) and (6.4.28), we have

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega \left( A e^{\chi _w v}b\ln b +\int _\varOmega e^{\chi _z v}c\ln c\right) +\frac{D_w}{8}\displaystyle \int _\varOmega e^{\chi _w v} \displaystyle \frac{|\nabla b|^2}{b}+ A\delta _w \int _\varOmega e^{\chi _w v}b\ln b \nonumber \\&+\displaystyle \frac{ 3 D_z}{4}\displaystyle \int _\varOmega e^{\chi _z v} \displaystyle \frac{|\nabla c|^2}{c}+ \delta _z \int _\varOmega e^{\chi _z v}c\ln c \\ \le&\,\,\displaystyle A\varepsilon \Vert c\Vert ^2_{L^2(\varOmega )} +\frac{A K_2}{\varepsilon }\Vert a\Vert ^2_{L^{r+1}(\varOmega )}+\frac{A K_2}{\varepsilon }+ K_3\Vert a\Vert ^2_{L^2(\varOmega )}+K_3 \nonumber \\ \le&\,\,\displaystyle A\varepsilon C_1\displaystyle \int _\varOmega e^{\chi _z v} \displaystyle \frac{|\nabla c|^2}{c} +\frac{A K_2}{\varepsilon }\Vert a\Vert ^2_{L^{r+1}(\varOmega )}+\frac{A K_2}{\varepsilon }+ K_3\Vert a\Vert ^2_{L^2(\varOmega )}+A\varepsilon C_1+K_3. \nonumber \end{aligned}$$
(6.4.29)

Taking \(\varepsilon =\frac{3D_z}{8AC_1}\) in (6.4.29), we can find \(C_2>0\) such that

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega \left( A e^{\chi _w v}b\ln b +\int _\varOmega e^{\chi _z v}c\ln c\right) +\frac{D_w}{8}\displaystyle \int _\varOmega e^{\chi _w v} \displaystyle \frac{|\nabla b|^2}{b}+ A\delta _w \int _\varOmega e^{\chi _w v}b\ln b\\&+\displaystyle \frac{ 3 D_z}{8}\displaystyle \int _\varOmega e^{\chi _z v} \displaystyle \frac{|\nabla c|^2}{c}+ \delta _z \int _\varOmega e^{\chi _z v}c\ln c\\ \le&\,\, C_2\Vert a\Vert ^2_{L^{r+1}(\varOmega )}+ C_2. \end{aligned} \end{aligned}$$
(6.4.30)

Combining (6.4.11) with (6.4.30) and using (6.4.18), we can pick \(\varepsilon >0\) in (6.4.11) appropriately small to derive that for some \(C_3>0\)

$$\begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega \left( A e^{\chi _w v}b\ln b +\int _\varOmega e^{\chi _u v}a\ln a+\int _\varOmega e^{\chi _z v}c\ln c\right) +\frac{D_w}{9}\displaystyle \int _\varOmega e^{\chi _w v} \displaystyle \frac{|\nabla b|^2}{b}\nonumber \\&+ A\delta _w \int _\varOmega e^{\chi _w v}b\ln b\\&+\displaystyle \frac{ 3 D_z}{8}\displaystyle \int _\varOmega e^{\chi _z v} \displaystyle \frac{|\nabla c|^2}{c}+ \delta _z \int _\varOmega e^{\chi _z v}c\ln c + D_u\displaystyle \int _\varOmega e^{\chi _u v}\displaystyle \frac{|\nabla a|^2}{a}+ \frac{\mu _u}{2} \int _\varOmega (\ln a+1)u^{r+1} \nonumber \\ \le&\,\, C_3\Vert a\Vert ^2_{L^{r+1}(\varOmega )}+ C_3,\nonumber \end{aligned}$$
(6.4.31)

which, along with \(a^{r+1}\le \varepsilon a^{r+1}\ln a+c(\varepsilon )\) for some \(c(\varepsilon )>0\), implies that there exist \(C_4>0\) and \(C_5>0\) fulfilling

$$\begin{aligned} \displaystyle \frac{d}{dt}\mathscr {F}(t) +C_4 \mathscr {F}(t)\le C_5 \end{aligned}$$
(6.4.32)

with

$$\begin{aligned} \mathscr {F}(t):=&\,\,\displaystyle A\int _\varOmega e^{\chi _w v(\cdot ,t)}b(\cdot ,t)\ln b(\cdot ,t) +\int _\varOmega e^{\chi _u v(\cdot ,t)}a(\cdot ,t)\ln a(\cdot ,t) \\&+\int _\varOmega e^{\chi _z v(\cdot ,t)}c(\cdot ,t)\ln c(\cdot ,t), \end{aligned}$$

and thereby

$$\begin{aligned} \mathscr {F}(t)\le C_6 \end{aligned}$$
(6.4.33)

is valid for some \(C_6>0\).

Now, by the inequality \(a\ln a\ge -e^{-1}\) for all \(a>0\),

$$\begin{aligned} \displaystyle \int _\varOmega a(\cdot ,t)|\ln a(\cdot ,t)|&=\,\, \displaystyle \int _\varOmega a(\cdot ,t)\ln a(\cdot ,t)-2\int _{a<1} a(\cdot ,t)\ln a(\cdot ,t) \\&\le \displaystyle \int _\varOmega a(\cdot ,t)\ln a(\cdot ,t)+2|\varOmega |, \end{aligned}$$

and similarly,

$$\begin{aligned} \displaystyle \int _\varOmega b(\cdot ,t)|\ln b(\cdot ,t)| \le \displaystyle \int _\varOmega b(\cdot ,t)\ln b(\cdot ,t)+2|\varOmega | \end{aligned}$$

as well as

$$\begin{aligned} \displaystyle \int _\varOmega c(\cdot ,t)|\ln c(\cdot ,t)| \le \displaystyle \int _\varOmega c(\cdot ,t)\ln c(\cdot ,t)+2|\varOmega |. \end{aligned}$$

Hence, (6.4.24)–(6.4.26) result readily from (6.4.33).

6.4.3 \(L^\infty \)-Bounds for a, b and c

By means of some quite straightforward \(L^p\) testing procedures, combining Lemma 2.1 with appropriate interpolation, we can now proceed to turn the outcome of Lemma 6.19 into the \(L^\infty \)-bounds for a, b and c.

Lemma 6.20

Let (a, b, v, c) be the classical solution of (6.4.1) in \(\varOmega \times [0,T_{max})\). Then one can find \(C>0\) fulfilling

$$\begin{aligned} \Vert a(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le C \end{aligned}$$
(6.4.34)

and

$$\begin{aligned} \Vert b(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le C \end{aligned}$$
(6.4.35)

as well as

$$\begin{aligned} \Vert c(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le C \end{aligned}$$
(6.4.36)

for all \(t\in (0,T_{max})\).

Proof

Testing the first equation in (6.4.1) by \( e^{\xi _u v} a^{p-1}\) with \(p>4\), integrating by parts and using the Young inequality, we can find \( C_1>0\) and \(C_2:=C_2(p)>0\) such that

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\chi _u v}a^p+\int _\varOmega e^{\chi _u v}a^p\\ =&\,\, p\displaystyle \int _\varOmega e^{\chi _u v} a^{p-1}a_t+\chi _u\int _\varOmega e^{\chi _u v}a^pv_t +\int _\varOmega e^{\chi _u v}a^p\\ =&\,\, p\displaystyle \int _\varOmega e^{\chi _u v} a^{p-1}\{ D_u e^{-\chi _u v}\nabla \cdot (e^{\chi _u v}\nabla a)+f(a,b,v,c) \}+\chi _u\int _\varOmega e^{\chi _u v}a^pv_t +\int _\varOmega e^{\chi _u v}a^p\\ \le&-\displaystyle \frac{4D_u (p-1)}{p} \int _\varOmega |\nabla a^{\frac{p}{2} }|^2+p\chi _u\alpha _u \displaystyle \int _\varOmega a^{p+1} e^{2\chi _u v} +C_1p\displaystyle \int _\varOmega a^{p} +C_1p\displaystyle \int _\varOmega a^{p}b\\ \le&-\displaystyle \frac{4D_u (p-1)}{p} \int _\varOmega |\nabla a^{\frac{p}{2} }|^2+C_2 \displaystyle \int _\varOmega a^{p+1} +C_2 \displaystyle \int _\varOmega b^{p+1}+C_2.\\ \end{aligned} \end{aligned}$$
(6.4.37)

Similarly, based on the other equations in (6.4.1), we infer the existence of \(C_3 > 0\) such that

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\xi _w v}b^p+\int _\varOmega e^{\xi _w v}b^p\\ \le&-\displaystyle \frac{4D_w (p-1)}{p} \int _\varOmega |\nabla b^{\frac{p}{2} }|^2+C_3 \displaystyle \int _\varOmega a^{p+1} +C_3 \displaystyle \int _\varOmega b^{p+1}+C_3 \displaystyle \int _\varOmega c^{p+1}+C_3\\ \end{aligned} \end{aligned}$$
(6.4.38)

as well as

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\int _\varOmega e^{\xi _z v} c^p+\int _\varOmega e^{\xi _z v}c^p\\ \le&-\displaystyle \frac{4D_z (p-1)}{p} \int _\varOmega |\nabla c^{\frac{p}{2} }|^2+C_3 \displaystyle \int _\varOmega a^{p+1}+C_3\displaystyle \int _\varOmega b^{p+1}+C_3 \displaystyle \int _\varOmega c^{p+1}+ C_3.\\ \end{aligned} \end{aligned}$$
(6.4.39)

Collecting (6.4.37)–(6.4.39), we then have

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\left\{ \int _\varOmega e^{\xi _u v}a^p+\int _\varOmega e^{\xi _w v}b^p+\int _\varOmega e^{\xi _z v} c^p\right\} +\int _\varOmega e^{\xi _u v}a^p+\int _\varOmega e^{\xi _w v}b^p+\int _\varOmega e^{\xi _z v} c^p\\ \le&-\displaystyle \frac{4 (p-1)}{p} \left( D_u \int _\varOmega |\nabla a^{\frac{p}{2} }|^2+ D_w \int _\varOmega |\nabla b^{\frac{p}{2} }|^2+ D_z \int _\varOmega |\nabla c^{\frac{p}{2} }|^2 \right) +(C_2+2C_3) \displaystyle \int _\varOmega a^{p+1}\\&+(C_2+2C_3)\displaystyle \int _\varOmega b^{p+1}+(C_2+2C_3) \displaystyle \int _\varOmega c^{p+1}+ C_2+2C_3. \end{aligned} \end{aligned}$$
(6.4.40)

Now on the basis of Lemma 6.19, we employ Lemma 2.1 to estimate \( \int _\varOmega a^{p+1}\), \( \int _\varOmega b^{p+1}\) and \(\int _\varOmega c^{p+1}\) in term of \(\int _\varOmega |\nabla a^{\frac{p}{2} }|^2\), \(\int _\varOmega |\nabla b^{\frac{p}{2} }|^2\) and \(\int _\varOmega |\nabla c^{\frac{p}{2} }|^2\), respectively.

Indeed, applying Lemma 2.1 to \(\varphi =a^{\frac{p}{2}}\), we have

$$\begin{aligned}&\quad (C_2+2C_3)\displaystyle \int _\varOmega a^{p+1}\nonumber \\ =&\,\,(C_2+2C_3) \Vert a^{\frac{p}{2}}\Vert ^{\frac{2(p+1)}{p}}_ {L^{\frac{2(p+1)}{p}}(\varOmega )}\\ \le&\,\, \displaystyle (C_2+2C_3)\varepsilon \Vert \nabla a^{\frac{p}{2}}\Vert ^2_{L^2(\varOmega )}\cdot \int _\varOmega a |\ln a^{\frac{p}{2}}| + (C_2+2C_3)K(p, \varepsilon )\left( \Vert a^{\frac{p}{2}}\Vert ^{\frac{2(p+1)}{p}}_ {L^{\frac{2}{p}}(\varOmega )}+1\right) \nonumber \\ =&\,\,\displaystyle \frac{p(C_2+2C_3)\varepsilon }{2}\Vert \nabla a^{\frac{p}{2}}\Vert ^2_{L^2(\varOmega )}\cdot \int _\varOmega a |\ln a| + (C_2+2C_3)K(p, \varepsilon )\left( \int _\varOmega a)^{p+1} +1\right) \nonumber \end{aligned}$$
(6.4.41)

which along with (6.4.24) and the appropriate choice of \(\varepsilon \) readily shows that for \(C_4(p)>0\)

$$(C_2+2C_3)\displaystyle \int _\varOmega a^{p+1}\le \displaystyle \frac{4 (p-1)D_u}{p} \int _\varOmega |\nabla a^{\frac{p}{2} }|^2+C_4(p). $$

Similarly,

$$(C_2+2C_3)\displaystyle \int _\varOmega b^{p+1}\le \displaystyle \frac{4 (p-1)D_w}{p} \int _\varOmega |\nabla b^{\frac{p}{2} }|^2+C_5(p) $$

as well as

$$(C_2+2C_3)\displaystyle \int _\varOmega c^{p+1}\le \displaystyle \frac{4 (p-1)D_z}{p} \int _\varOmega |\nabla c^{\frac{p}{2} }|^2+C_6(p). $$

Therefore, (6.4.40) shows that

$$\begin{aligned}&\displaystyle \frac{d}{dt}\left\{ \int _\varOmega e^{\xi _u v}a^p+\int _\varOmega e^{\xi _w v}b^p+\int _\varOmega e^{\xi _z v} c^p\right\} +\int _\varOmega e^{\xi _u v}a^p+\int _\varOmega e^{\xi _w v}b^p+\int _\varOmega e^{\xi _z v} c^p\\ \le&\,\, C_7(p), \end{aligned}$$

which entails that for all \(p\ge 2\) there exists \(C_8(p) > 0\) such that

$$\begin{aligned} \int _\varOmega a^p(\cdot ,t)+ \int _\varOmega b^p(\cdot ,t)+\int _\varOmega c^p(\cdot ,t)\le C_8(p) \end{aligned}$$
(6.4.42)

for all \(t\in (0,T_{max})\).

Furthermore, by adapting a well-established Moser-type iteration, one can readily turn the latter into the \(L^\infty \) bounds for a, b, c. However, since the procedure is rather standard (see Tao and Winkler 2014b, 2020b for example), we give the details only in places which are characteristic of the present setting.

By a straightforward calculation and three integrations by parts, we get

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\left\{ \int _\varOmega e^{\xi _u v}a^p+e^{\xi _w v}b^p+ e^{\xi _z v} c^p\right\} + \int _\varOmega \left\{ e^{\xi _u v}a^p+e^{\xi _w v}b^p+ e^{\xi _z v} c^p\right\} \\ \le&\,\,-2\displaystyle \min \{D_u,D_w,D_z\} \int _\varOmega \left\{ |\nabla a^{\frac{p}{2} }|^2 + |\nabla b^{\frac{p}{2} }|^2+ |\nabla c^{\frac{p}{2} }|^2\right\} \\&+C_9p \int _\varOmega \left\{ a^{p+1}+ b^{p+1} +c^{p+1}\right\} +C_9 \end{aligned} \end{aligned}$$
(6.4.43)

where \(C_9>0\) as all subsequently appearing constants \(C_{10},C_{11},\dots \) is independent of \(p \ge 4\).

It is observed that by the Gagliardo–Nirenberg inequality, due to \(2 \le \frac{2(p+1)}{p}\le 2.5\) for \(p \ge 4\), one can pick \(C_{10}>1\) such that for all \(p\ge 4\),

$$\begin{aligned} \Vert \varphi \Vert _{L^{\frac{2(p+1)}{p}}(\varOmega )} \le C_{10}\Vert \nabla \varphi \Vert _{L^2(\varOmega )}^{\frac{p+2}{2(p+1)}} \Vert \varphi \Vert ^{\frac{p}{2(p+1)}}_{L^1(\varOmega )}+C_{10}\Vert \varphi \Vert _{L^1(\varOmega )} ~~\hbox {for all}~ \varphi \in W^{1,2}(\varOmega ). \end{aligned}$$

Applying this together with the Young inequality, we obtain that for some \(C_{11}>0\),

$$\begin{aligned} C_9 p \displaystyle \int _\varOmega a^{p+1}&=\,\, C_9 p\Vert a^\frac{p}{2}\Vert ^{\frac{2(p+1)}{p}}_{L^{\frac{2(p+1)}{p}}(\varOmega )}\\&\le C_9 p \left\{ C_{10}\Vert \nabla a^\frac{p}{2} \Vert _{L^2(\varOmega )}^{\frac{p+2}{2(p+1)}} \cdot \Vert a^\frac{p}{2}\Vert ^{\frac{p}{2(p+1)}}_{L^1(\varOmega )} +C_{10}\Vert a^\frac{p}{2}\Vert _{L^1(\varOmega )}\right\} ^{\frac{2(p+1)}{p}} \\&\le 8 C_9 C_{10}^3p \Vert \nabla a^\frac{p}{2} \Vert _{L^2(\varOmega )}^{\frac{p+2}{p}} \cdot \Vert a^\frac{p}{2}\Vert _{L^1(\varOmega )} +8 C_9 C_{10}^3p\Vert a^\frac{p}{2}\Vert ^{\frac{2(p+1)}{p}}_{L^1(\varOmega )}\\&\le \displaystyle \min \{D_u,D_w,D_z\} \Vert \nabla a^\frac{p}{2} \Vert _{L^2(\varOmega )}^2 +C_{11} p^4\max \{1,\Vert a^\frac{p}{2}\Vert _{L^1(\varOmega )}\}^{\frac{2p}{p-2}}, \end{aligned}$$

where the fact that \(\frac{2(p+1)}{p}\le \frac{2p}{p-2}\le 4\) for any \(p\ge 4\) is used.

Similarly, we have

$$ C_9 p \displaystyle \int _\varOmega b^{p+1}\le \displaystyle \min \{D_u,D_w,D_z\} \Vert \nabla b^\frac{p}{2} \Vert _{L^2(\varOmega )}^2 +C_{11} p^4\max \{1,\Vert b^\frac{p}{2}\Vert _{L^1(\varOmega )}\}^{\frac{2p}{p-2}} $$

as well as

$$ C_9 p \displaystyle \int _\varOmega c^{p+1}\le \displaystyle \min \{D_u,D_w,D_z\} \Vert \nabla c^\frac{p}{2} \Vert _{L^2(\varOmega )}^2 +C_{11} p^4\max \{1,\Vert c^\frac{p}{2}\Vert _{L^1(\varOmega )}\}^{\frac{2p}{p-2}}. $$

Consequently, inserting the above inequalities into (6.4.43) yields the existence of \(C_{12}>0\) such that

$$\begin{aligned} \begin{aligned}&\displaystyle \frac{d}{dt}\left\{ \int _\varOmega e^{\xi _u v}a^p+e^{\xi _w v}b^p+ e^{\xi _z v} c^p\right\} + \int _\varOmega \left\{ e^{\xi _u v}a^p+e^{\xi _w v}b^p+ e^{\xi _z v} c^p\right\} \\ \le&\,\, C_{12} p^4\max \{1,\Vert a^\frac{p}{2}\Vert _{L^1(\varOmega )} +\Vert b^\frac{p}{2}\Vert _{L^1(\varOmega )}+\Vert c^\frac{p}{2}\Vert _{L^1(\varOmega )}\}^{\frac{2p}{p-2}}. \end{aligned} \end{aligned}$$
(6.4.44)

Now let \(p_k=4\cdot 2^k \) and \(M_k=\max \{1,\displaystyle \sup _{t\in (0,T_{max})}\int _\varOmega a^{p_k}(\cdot ,t)+b^{p_k}(\cdot ,t)+c^{p_k}(\cdot ,t)\}\) for \(k=0,1,2,\ldots \). Then (6.4.44) implies that for \(k=1,2,\ldots \)

$$\begin{aligned}&\displaystyle \frac{d}{dt}\left\{ \int _\varOmega e^{\xi _u v}a^{p_k}+e^{\xi _w v}b^{p_k}+ e^{\xi _z v} c^{p_k}\right\} + \int _\varOmega \left\{ e^{\xi _u v}a^{p_k}+e^{\xi _w v}b^{p_k}+ e^{\xi _z v} c^{p_k}\right\} \\ \le&\,\, C_{12}{p_k}^4 M_{k-1}^4, \end{aligned}$$

which entails the existence of \(L>1\) independent of k such that

$$ M_k\le \max \{L^k M^4_{k-1}, |\varOmega |(\Vert u_0\Vert _{L^\infty (\varOmega )}^{p_k}+\Vert w_0\Vert ^{p_k}_{L^\infty (\varOmega )}+\Vert z_0\Vert ^{p_k}_{L^\infty (\varOmega )}) \} ~~\hbox {for all}~ k\ge 1.$$

Therefore, by means of a standard recursive argument (see Pang and Wang 2018; Tao and Winkler 2014b for example), both when \(L^k M^4_{k-1}\le |\varOmega |(\Vert u_0\Vert _{L^\infty (\varOmega )}^{p_k}+\Vert w_0\Vert ^{p_k}_{L^\infty (\varOmega )}+\Vert z_0\Vert ^{p_k}_{L^\infty (\varOmega )}) \) for infinitely many \(k \ge 1\), and as well in the opposite case, we can obtain some \(C_{13}>0\) such that for all \(k\ge 1\)

$$ M_k^{\frac{1}{p_k}}\le C_{13}, $$

from which, after taking \(k\rightarrow \infty \), the claims (6.4.34)–(6.4.36) readily follow.

According to Lemma 6.14, it remains for us to establish a priori estimates for \(\Vert \nabla v(\cdot ,t)\Vert _{L^4(\varOmega )}\).

Lemma 6.21

Let \(T>0\). Then there exists \(C(\hat{T})>0\) such that \(\Vert \nabla v(\cdot ,t)\Vert _{L^4(\varOmega )} \le C(\hat{T})\) for all \(t<\hat{T}\), where \(\hat{T}:=\min \{T, T_{max}\}\).

Proof

This can be achieved through an appropriate combination of three further testing processes, essentially relying on the \(L^\infty \)-estimates for a, b and c just asserted. We refrain from giving the proof and refer to Tao and Winkler (2020b) or Tao and Winkler (2014b) for details in a closely related setting.

We are now in the position to prove Theorem 6.2.

Proof of Theorem 6.2.   Thanks to the equivalence of (6.1.10) and (6.4.1) in the considered framework of classical solutions and in particular the extensibility criterion provided by Lemma 6.14, the proof is an evident consequence of Lemmas 6.20 and 6.21.

6.5 Asymptotic Behavior of Solutions to an Oncolytic Virotherapy Model

At the beginning of this subsection, in light of the method used in Horstmann and Winkler (2005) and Pang and Wang (2017), we provide the following statement on the local existence and extensibility of solutions to (6.1.12) as below.

Lemma 6.22

Suppose that \(\varOmega \subset \mathbb {R}^{N}\left( N\ge 1 \right) \) be a bounded domain with smooth boundary. Then one can find \(T_{max}\in (0,\infty ] \) and a unique quadruple of nonnegative functions \(\left( u,v,w,z \right) \in \left( C\left( \bar{\varOmega }\times [0,T_{max})\right) \bigcap C^{2,1}\left( \bar{\varOmega }\times (0,T_{max}) \right) \right) ^4\) which solves (6.1.12) classically in \(\varOmega \times (0,T_{max})\). Moreover, if \(T_{max}< +\infty \), then

$$\begin{aligned} \lim _{t\nearrow T_{max}}\!\left( \!\left\| u\left( \cdot ,t \right) \right\| _{W^{1,2q}\left( \varOmega \right) }\!+\!\left\| v\left( \cdot ,t \right) \right\| _{W^{2,q}\left( \varOmega \right) }\!+\!\left\| w\left( \cdot ,t \right) \right\| _{W^{1,2q}\left( \varOmega \right) }\!+\!\left\| z\left( \cdot ,t \right) \right\| _{W^{1,2q}\left( \varOmega \right) }\!\right) \!=\!\infty \end{aligned}$$
(6.5.1)

for all \(q> \frac{N}{2}\).

To make the system mass-conserved, we introduce a nonnegative variable Q satisfying

$$\begin{aligned} \left\{ \begin{aligned}&Q_{t}=\varDelta Q+\left( \frac{\rho _{u}+\rho _{z}}{\rho _{w}}\delta _{w}-\beta \right) w+\delta _{z}z,&\left( x,t \right) \in \varOmega \times \left( 0,T \right) ,\\&\nabla Q\cdot \nu =0,&\left( x,t \right) \in \partial \varOmega \times \left( 0,T \right) ,\\&Q\left( x,0 \right) =0,&x\in \varOmega ,\\ \end{aligned} \right. \end{aligned}$$
(6.5.2)

where \(\frac{\rho _{u}+\rho _{z}}{\rho _{w}}\delta _{w}-\beta \ge 0\). Then it is easy to see that for all \(t\in (0,T)\),

$$\int _{\varOmega } \left[ \left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q \right) \left( \cdot ,t \right) \right] _{t}= 0,$$

which means

$$\begin{aligned} \int _{\varOmega }\left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q \right) \left( \cdot ,t \right) =\int _{\varOmega }\left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w_{0}+z_{0}\right) . \end{aligned}$$
(6.5.3)

We first collect some easily verifiable observations in the following lemma. The constants \(c_{i},(i=1,2,3,4)\), \(c_{10}\) and \(C_{p_0}\) refer to Lemmas 1.1, 4.3 and 3.2 respectively.

Lemma 6.23

Under the assumptions of Theorem 6.3, there exist \(M_{i}> 1\) \(\left( i=0,1,\right. \left. \cdot \cdot \cdot ,6 \right) \) and \(\varepsilon (\xi _u,\xi _w,\xi _z,\rho _u,\rho _w,\rho _z,\alpha _y,\alpha _w) > 0\) such that

$$\begin{aligned} 2c_3+2c_{10}c_2\xi _{u}\varepsilon ( M_{1}M_{4}+M_{0}M_{5} )+ 2c_{10}c_3\rho _{u}M_{0}\varepsilon (M_{1}+M_{3}) \le \frac{M_{1}}{2}, \end{aligned}$$
(6.5.4)
$$\begin{aligned} 2c_3+2c_{10}c_2\xi _{w}\varepsilon \left( M_{2}M_{4}+M_{0}M_{5} \right) +2c_{10}c_3\rho _{w} M_{0}\varepsilon (M_{1}+M_{3} )\le \frac{M_{2}}{2}, \end{aligned}$$
(6.5.5)
$$\begin{aligned} 2c_3+2c_{10}c_2\xi _{z}\varepsilon \left( M_{3}M_{4}+M_{0}M_{5}\right) +2c_{10}c_3\rho _{z}M_{0}\varepsilon ( M_{1}+ M_{3})+2c_{10}c_3\beta M_{2} \le \frac{M_{3}}{2}, \end{aligned}$$
(6.5.6)
$$\begin{aligned} 1+\frac{(\alpha _{u}M_{1}+\alpha _{w}M_{2})\left\| v_{0}\left( \cdot \right) \right\| _{L^{\infty }\left( \varOmega \right) } \left( \frac{2p_0-1}{2p_0e(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{2p_0}}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{2p_{0}}}} \le \frac{M_{4}}{2}, \end{aligned}$$
(6.5.7)
$$\begin{aligned}&1+\varepsilon \Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )}\frac{(\alpha _{u}M_{1}+ \alpha _{w}M_{2})^{2}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{p_{0}}}}\left( \frac{2p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{p_0}}\nonumber \\&+2\varepsilon \frac{\alpha _uM_1+\alpha _wM_2}{(2\alpha p_0)^{\frac{1}{2p_0}}}\left( \frac{2p_0-1}{2p_0e(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{2p_0}}\nonumber \\&+\Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )}C_{p_0}\alpha _u\left( \frac{p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{p_0-1}{p_0}}\nonumber \\&\cdot \left( \frac{\varepsilon \xi _{u}M_1M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{u}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \\&+\Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )}C_{p_0}\alpha _w\left( \frac{p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{p_0-1}{p_0}} \nonumber \\&\cdot \left( \frac{\varepsilon \xi _{w}M_2M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{w}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \nonumber \\&+\frac{\Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )}C_{p_0}M_0|\varOmega |^{\frac{1}{p_0}}}{\left( k p_{0}\right) ^{\frac{1}{p_0}}}\left( \frac{p_0-1}{p_0e(\delta _v-\alpha -k)}\right) ^{\frac{p_0-1}{p_0}} \nonumber \\&\cdot \left( \varepsilon \alpha _u\rho _uM_0+\frac{1}{2}\alpha _u(\delta _v-\alpha ) +\varepsilon \alpha _w\rho _wM_0\right) \le \frac{M_{5}}{2},\nonumber \end{aligned}$$
(6.5.8)
$$\begin{aligned} 2c_{10}c_4M_{0}M_{4}\varepsilon \left( \xi _{u}+ \frac{\rho _{u}+\rho _{z}}{\rho _{w}}\xi _{w}+\xi _{z}\right) \le \frac{M_{6}}{2}, \end{aligned}$$
(6.5.9)

where

$$M_0=1+M_6+2c_1, \quad k=\min \left\{ \frac{1}{2}(\delta _v-\alpha ),\delta _w\right\} \in (0,\alpha ),$$

with \(\alpha \in (0,\min \left\{ \lambda _{1}, \delta _v\right\} )\) and \(\lambda _{1}> 0\) the first nonzero eigenvalue of \(-\varDelta \) in \(\varOmega \) under the Neumann condition.

For constants \(\alpha \in (0,\min \left\{ \lambda _{1}, \delta _v\right\} )\) and \(M_i>1( i=0,1,\ldots ,6 )\) referring to Lemma 6.23, let

(6.5.10)

By Lemma 6.22 and the smallness condition on the initial data in Theorem 6.3, \(T>0\) is well-defined. We first show \(T=T_{max}\). To this end, we will show that all of the estimates mentioned in (6.5.10) are valid with even smaller coefficients on the right-hand side. The derivation of these estimates will mainly rely on \(L^p-L^q\) estimates for the Neumann heat semigroup and the fact that the classical solutions on \((0,T_{max})\) can be represented as

$$\begin{aligned}&\left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q \right) \left( \cdot ,t \right) \nonumber \\ =&\,\,e^{t\varDelta } \left( u_0+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w_0+z_0 \right) \left( \cdot \right) \nonumber \\&-\int _{0}^{t}e^{\left( t-s \right) \varDelta }\left[ \xi _{u}\nabla \cdot \left( u\nabla v \right) + \frac{\rho _{u}+\rho _{z}}{\rho _{w}}\xi _{w}\nabla \cdot \left( w\nabla v \right) +\xi _{z}\nabla \cdot \left( z\nabla v \right) \right] (\cdot ,s)ds, \end{aligned}$$
(6.5.11)
$$\begin{aligned} u(\cdot ,t)=e^{t\varDelta }u_{0}\left( \cdot \right) +\int _{0}^{t}e^{\left( t-s \right) \varDelta }\left[ -\xi _{u}\nabla \cdot \left( u\nabla v \right) -\rho _{u}uz \right] (\cdot ,s)ds, \end{aligned}$$
(6.5.12)
$$\begin{aligned} w(\cdot ,t)=e^{t(\varDelta -\delta _w)}w_{0}\left( \cdot \right) +\int _{0}^{t}e^{\left( t-s \right) (\varDelta -\delta _w) }\left[ -\xi _{w}\nabla \cdot \left( w\nabla v \right) +\rho _{w}uz\right] (\cdot ,s)ds, \end{aligned}$$
(6.5.13)
$$\begin{aligned} z(\cdot ,t)=e^{t(\varDelta -\delta _z) }z_{0}\left( \cdot \right) +\int _{0}^{t}e^{\left( t-s \right) (\varDelta -\delta _z) }\left[ -\xi _{z}\nabla \cdot \left( z\nabla v \right) -\rho _{z}uz+\beta w \right] (\cdot ,s)ds \end{aligned}$$
(6.5.14)

for all \(t\in \left( 0,T_{max} \right) \) as per the variation-of-constants formula.

The global boundedness for solutions of (6.1.12) can be obtained directly from the following lemmas.

Lemma 6.24

Under the assumptions of Theorem 6.3, for all \(t\in \left( 0,T\right) \), we have

$$\begin{aligned} \left\| \left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q\right) \left( \cdot ,t \right) \right\| _{L^{\infty }\left( \varOmega \right) }\le M_{0}\varepsilon , \end{aligned}$$
(6.5.15)

where \(M_0=1+M_6+2c_1\) with \(c_1\) defined in Lemma 1.1.

Proof

Set \(m_{\infty }=\frac{1}{\left| \varOmega \right| }\int _{\varOmega }\left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w_{0}+z_{0} \right) \), then \(m_\infty \le \varepsilon \). It is obvious that

$$e^{t\varDelta }m_{\infty }=m_{\infty },\int _{\varOmega }\left[ \left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0} \right) \left( \cdot \right) -m_{\infty }\right] =0, $$
$$\left\| \left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0} \right) \left( \cdot \right) -m_{\infty } \right\| _{L^{\infty }\left( \varOmega \right) }\le \varepsilon . $$

According to the Lemma 1.1(i), we know for all \(t\in \left( 0,T \right) \),

$$\left\| e^{t\varDelta }\left[ \left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0} \right) \left( \cdot \right) -m_{\infty }\right] \right\| _{L^{\infty }\left( \varOmega \right) }\le 2c_1\varepsilon e^{-\lambda _{1}t}.$$

Then due to the definition of T and Lemma 1.1(i), we have

$$\begin{aligned}&\left\| \left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q-m_{\infty } \right) \left( \cdot ,t \right) \right\| _{L^{\infty }\left( \varOmega \right) }\nonumber \\ =&\,\left\| \left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q \right) \left( \cdot ,t \right) -e^{t\varDelta }\left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0}\right) \left( \cdot \right) \right\| _{L^{\infty }\left( \varOmega \right) }\nonumber \\&+\left\| e^{t\varDelta }\left[ \left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0}\right) \left( \cdot \right) -m_{\infty } \right] \right\| _{L^{\infty }\left( \varOmega \right) }\nonumber \\ \le&\,\, M_{6}\varepsilon e^{-\alpha t} +2c_1\varepsilon e^{-\lambda _1 t}\nonumber \\ \le&\,\left( M_{6}+2c_1 \right) \varepsilon e^{-\alpha t}\nonumber \end{aligned}$$

and hence end the proof.

Lemma 6.25

Under the assumptions of Theorem 6.3, for all \(t\in (0,T)\), we have

$$\left\| \nabla u\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{1}}{2}\varepsilon e^{-\alpha t}.$$

Proof

Applying (6.5.12), Lemma 1.1 (iii), we have

$$\begin{aligned}&\left\| \nabla u\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\, \left\| \nabla e^{t\varDelta }u_{0}\left( \cdot \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }+\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) \varDelta }\left[ -\xi _{u}\nabla \cdot \left( u\nabla v \right) -\rho _{u}uz \right] \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\,\ 2c_3e^{-\lambda _{1}t} \left\| \nabla u_{0}\left( \cdot \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }+\xi _{u} \int _{0}^{t}\left\| \nabla e^{\left( t-s \right) \varDelta }\nabla \cdot \left( u\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\\&+\rho _{u}\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) \varDelta }\left( uz\right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds.\nonumber \end{aligned}$$
(6.5.16)

From Lemmas 1.1(ii) and 4.3, we obtain

$$\begin{aligned}&\xi _{u}\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) \varDelta }\nabla \cdot \left( u\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\, \xi _{u}\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) \varDelta }\left( \nabla u\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds+\xi _{u}\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) \varDelta }\left( u\varDelta v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\,\xi _{u}c_2\int _{0}^{t}\left[ 1+\left( t-s \right) ^{-\frac{1}{2}-\frac{N}{4p_{0}}} \right] e^{-\lambda _{1}\left( t-s \right) }\left\| \nabla u\nabla v\left( \cdot ,s \right) \right\| _{L^{p_{0}}\left( \varOmega \right) }ds\nonumber \\&+\xi _{u}c_2\int _{0}^{t}\left[ 1+\left( t-s \right) ^{-\frac{1}{2}-\frac{N}{4p_{0}}} \right] e^{-\lambda _{1}\left( t-s \right) }\left\| u\varDelta v\left( \cdot ,s \right) \right\| _{L^{p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\,\xi _{u}c_2\int _{0}^{t}\left[ 1+\left( t-s \right) ^{-\frac{1}{2}-\frac{N}{4p_{0}}} \right] e^{-\lambda _{1}\left( t-s \right) }\left( \left\| \nabla u\left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\cdot \left\| \nabla v \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) } \right) ds\nonumber \\&+\xi _{u}c_2\int _{0}^{t}\left[ 1+\left( t-s \right) ^{-\frac{1}{2}-\frac{N}{4p_{0}}} \right] e^{-\lambda _{1}\left( t-s \right) }\left( \left\| u\left( \cdot ,s \right) \right\| _{L^{\infty }\left( \varOmega \right) }\cdot \left\| \varDelta v\left( \cdot ,s \right) \right\| _{L^{p_{0}}\left( \varOmega \right) } \right) ds\nonumber \\ \le&\,\, 2c_{10}c_2\xi _{u}( M_{1}M_{4}\varepsilon ^{2} e^{-\min \{2\alpha ,\lambda _1\} t}+M_{0}M_{5}\varepsilon ^{2}e^{-\alpha t} )\nonumber \\ \le&\,\,2c_{10}c_2\xi _{u}\varepsilon ^{2} e^{-\alpha t}( M_{1}M_{4}+M_{0}M_{5} ).\nonumber \end{aligned}$$

From Hölder’s inequality, using Lemmas 1.1(iii) and 4.3, we get

$$\begin{aligned}&\rho _{u}\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) \varDelta }\left( uz\right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\,2\rho _{u}c_3\int _{0}^{t} e^{-\lambda _{1}(t-s)}\left( \left\| z \nabla u(\cdot ,s) \right\| _{L^{2p_{0}}(\varOmega )}+\left\| u\nabla z(\cdot ,s) \right\| _{L^{2p_{0}}(\varOmega )} \right) ds\nonumber \\ \le&\,\,2\rho _{u}c_3\int _{0}^{t} e^{-\lambda _{1}(t-s)}\nonumber \\&\cdot \left( \left\| \nabla u\left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\cdot \left\| z\left( \cdot ,s \right) \right\| _{L^{\infty }\left( \varOmega \right) }+\left\| u\left( \cdot ,s \right) \right\| _{L^{\infty }\left( \varOmega \right) }\cdot \left\| \nabla z\left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) } \right) ds\nonumber \\ \le&\,\,2c_{10}c_3\rho _{u} (M_{0}M_{1}\varepsilon ^{2} e^{-\alpha t}+M_{0}M_{3}\varepsilon ^{2} e^{-\alpha t})\nonumber \\ =&\,\, 2c_{10}c_3\rho _{u}M_{0}\varepsilon ^{2} e^{-\alpha t}(M_{1}+M_{3}).\nonumber \end{aligned}$$

Therefore, inserting the above two results into (6.5.16), we arrive at

$$\begin{aligned}&\left\| \nabla u\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\, 2c_3\varepsilon e^{-\alpha t}+2c_{10}c_2\xi _{u}\varepsilon ^{2} e^{-\alpha t}( M_{1}M_{4}+M_{0}M_{5} )+ 2c_{10}c_3\rho _{u}M_{0}\varepsilon ^{2} e^{-\alpha t}(M_{1}+M_{3})\nonumber \\ \le&\,\left[ 2c_3+2c_{10}c_2\xi _{u}\varepsilon ( M_{1}M_{4}+M_{0}M_{5} )+ 2c_{10}c_3\rho _{u}M_{0}\varepsilon (M_{1}+M_{3}) \right] \varepsilon e^{-\alpha t}.\nonumber \end{aligned}$$

According to (6.5.4), we thereby complete the proof.

Lemma 6.26

Under the assumptions of Theorem 6.3, for all \(t\in (0,T)\), we have

$$\begin{aligned} \left\| \nabla w\left( \cdot ,t \right) \right\| _{L^{2p_{0}}(\varOmega )}\le \frac{M_{2}}{2}\varepsilon e^{-\alpha t}. \end{aligned}$$
(6.5.17)

Proof

From (6.5.13), using Lemmas 1.1(iii) and 4.3, we have

$$\begin{aligned}&\left\| \nabla w\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\, \left\| \nabla e^{t(\varDelta -\delta _w)}w_{0}(\cdot ) \right\| _{L^{2p_{0}}(\varOmega )}+\xi _{w} \int _{0}^{t}\left\| \nabla e^{\left( t-s \right) (\varDelta -\delta _w)}\nabla \cdot \left( w\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\&+\rho _{w}\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) (\varDelta -\delta _w) }\left( uz\right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\,2c_3\varepsilon e^{-(\lambda _{1}+\delta _{w}) t}+\xi _{w}c_2\int _{0}^{t}[1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_{0}}}]e^{-(\lambda _{1}+\delta _{w})(t-s)}\nonumber \\&\cdot \left( \left\| \nabla w (\cdot ,s)\right\| _{L^{2p_{0}}(\varOmega )} \left\| \nabla v (\cdot ,s)\right\| _{L^{2p_{0}}(\varOmega )}\right) ds\nonumber \\&+\xi _{w}c_2\int _{0}^{t}[1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_{0}}}]e^{-(\lambda _{1}+\delta _{w})(t-s)}\left( \left\| w (\cdot ,s)\right\| _{L^{\infty }(\varOmega )}\left\| \varDelta v (\cdot ,s)\right\| _{L^{p_{0}}(\varOmega )}\right) ds\nonumber \\&+2\rho _{w}c_3\int _{0}^{t}e^{-(\lambda _{1}+\delta _{w})(t-s)}\left( \left\| \nabla u (\cdot ,s)\right\| _{L^{2p_{0}}(\varOmega )} \left\| z (\cdot ,s)\right\| _{L^{\infty }(\varOmega )}\right) ds\nonumber \\&+2\rho _{w}c_3\int _{0}^{t}e^{-(\lambda _{1}+\delta _{w})(t-s)}\left( \left\| u (\cdot ,s)\right\| _{L^{\infty }(\varOmega )} \left\| \nabla z (\cdot ,s)\right\| _{L^{2p_{0}}(\varOmega )}\right) ds\nonumber \\ \le&\,\left[ 2c_3+2c_{10}c_2\xi _{w}\varepsilon \left( M_{2}M_{4}+M_{0}M_{5} \right) +2c_{10}c_3\rho _{w} M_{0}\varepsilon (M_{1}+M_{3} ) \right] \varepsilon e^{-\alpha t},\nonumber \end{aligned}$$

which along with (6.5.5) implies that (6.5.17) is valid.

Similar as done in Lemma 6.26, we also have the following.

Lemma 6.27

Under the assumptions of Theorem 6.3, for all \(t\in (0,T)\), we have

$$\begin{aligned} \left\| \nabla z\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{3}}{2}\varepsilon e^{-\alpha t}. \end{aligned}$$
(6.5.18)

Proof

From (6.5.14), using Lemmas 1.1(iii) and 4.3, we have

$$\begin{aligned}&\left\| \nabla z\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\, \left\| \nabla e^{t(\varDelta -\delta _z)}z_{0}(\cdot ) \right\| _{L^{2p_{0}}(\varOmega )}+\xi _{z} \int _{0}^{t}\left\| \nabla e^{\left( t-s \right) (\varDelta -\delta _z)}\nabla \cdot \left( z\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\&+\int _{0}^{t}\left\| \nabla e^{\left( t-s \right) (\varDelta -\delta _z) }(-\rho _{z}uz +\beta w)\left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\, 2c_3\varepsilon e^{-(\lambda _{1}+\delta _{z})t}+\xi _{z}c_2\int _{0}^{t}[1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_{0}}}]e^{-(\lambda _{1}+\delta _{z})(t-s)}\nonumber \\&\cdot \left\| \nabla z (\cdot ,s)\right\| _{L^{2p_{0}}(\varOmega )} \left\| \nabla v (\cdot ,s)\right\| _{L^{2p_{0}}(\varOmega )}ds\nonumber \\&+\xi _{z}c_2\int _{0}^{t}[1+(t-s)^{-\frac{1}{2}-\frac{N}{4p_{0}}}]e^{-(\lambda _{1}+\delta _{z})(t-s)}\left\| z (\cdot ,s)\right\| _{L^{\infty }(\varOmega )}\left\| \varDelta v (\cdot ,s)\right\| _{L^{p_{0}}(\varOmega )}ds\nonumber \\&+\rho _{z}c_3\int _{0}^{t}2e^{-(\lambda _{1}+\delta _{z})(t-s)}\nonumber \\&\cdot \left( \left\| \nabla u\left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\cdot \left\| z\left( \cdot ,s \right) \right\| _{L^{\infty }\left( \varOmega \right) }+\left\| u\left( \cdot ,s \right) \right\| _{L^{\infty }\left( \varOmega \right) }\cdot \left\| \nabla z\left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) } \right) ds\nonumber \\&+2\beta c_3\int _{0}^{t}e^{-(\lambda _{1}+\delta _{z})(t-s)}\left\| \nabla w(\cdot ,s) \right\| _{L^{2p_{0}}(\varOmega )}ds\nonumber \\ \le&\,\left[ 2c_3+2c_{10}c_2\xi _{z}\varepsilon \left( M_{3}M_{4}+M_{0}M_{5}\right) +2c_{10}c_3\rho _{z}M_{0}\varepsilon ( M_{1}+ M_{3})+2c_{10}c_3\beta M_{2} \right] \varepsilon e^{-\alpha t},\nonumber \end{aligned}$$

which together with (6.5.6) already implies that (6.5.18) holds.

Lemma 6.28

Under the assumptions of Theorem 6.3, for all \(t\in (0,T)\), we have

$$\begin{aligned} \left\| \nabla v\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{4}}{2}\varepsilon e^{-\alpha t}. \end{aligned}$$
(6.5.19)

Proof

We know that

$$v_{t}+(\alpha _{u}u+\alpha _{w}w+\delta _v )v=0,$$

from which we obtain

$$\begin{aligned} v\left( \cdot , t \right) =v_{0}\left( \cdot \right) e^{-\int _{0}^{t}\left( \alpha _{u}u+\alpha _{w}w+\delta _v \right) ds }. \end{aligned}$$
(6.5.20)

It follows that

$$\begin{aligned}&\left\| \nabla v\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\nonumber \\ =&\,\left\| \nabla \left[ v_{0}( \cdot ) e^{-\int _{0}^{t}\left( \alpha _{u}u+\alpha _{w}w +\delta _v \right) ds }\right] \right\| _{L^{2p_{0}}\left( \varOmega \right) } \\ \le&\,\left\| \nabla v_{0}\left( \cdot \right) e^{-\int _{0}^{t}\left( \alpha _{u}u+\alpha _{w}w +\delta _v \right) ds } \right\| _{L^{2p_{0}}\left( \varOmega \right) }+\left\| v_{0}\left( \cdot \right) \nabla e^{-\int _{0}^{t}\left( \alpha _{u}u+\alpha _{w}w+\delta _v \right) ds } \right\| _{L^{2p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\left\| \nabla v_{0}\left( \cdot \right) e^{-\delta _v t} \right\| _{L^{2p_{0}}\left( \varOmega \right) }+\left\| v_{0}\left( \cdot \right) \right\| _{L^{\infty }\left( \varOmega \right) }\left\| e^{-\delta _v t} \int _{0}^{t} \left( \alpha _{u}\nabla u+\alpha _{w}\nabla w \right) \left( \cdot ,s \right) ds\right\| _{L^{2p_{0}}\left( \varOmega \right) }.\nonumber \end{aligned}$$
(6.5.21)

Noticing that

$$\begin{aligned} \left( \int _{0}^{t}\left| \nabla u\left( \cdot ,s \right) \right| ds \right) ^{2p_{0}}\le&\,\, \left[ \left( \int _{0}^{t}\left| \nabla u\left( \cdot ,s \right) \right| ^{2p_{0}}ds \right) ^{\frac{1}{2p_{0}}}\left( \int _{0}^{t}1ds \right) ^{\frac{2p_{0}-1}{2p_{0}}} \right] ^{2p_{0}}\nonumber \\ \le&\,\, t^{2p_{0}-1}\int _{0}^{t}\left| \nabla u\left( \cdot ,s \right) \right| ^{2p_{0}}ds,\nonumber \end{aligned}$$

then

$$\begin{aligned}&\left\| \int _{0}^{t}\nabla u(\cdot ,s) ds \right\| _{L^{2p_{0}}\left( \varOmega \right) }\nonumber \\ =&\,\left[ \int _{\varOmega }\left| \int _{0}^{t} \nabla u \left( \cdot ,s \right) ds \right| ^{2p_{0}}dx \right] ^{\frac{1}{2p_{0}}}\nonumber \\ \le&\,\left[ t^{2p_{0}-1}\int _{0}^{t}\int _{\varOmega }\left| \nabla u\left( \cdot ,s \right) \right| ^{2p_{0}}dxds \right] ^{\frac{1}{2p_{0}}} \\ \le&\,\left( \int _{0}^{t}\int _{\varOmega }\left| \nabla u\left( \cdot ,s \right) \right| ^{2p_{0}}dxds\right) ^{\frac{1}{2p_{0}}}t^{\frac{2p_{0}-1}{2p_{0}}}\nonumber \\ \le&\,\left( \int _{0}^{t}(M_{1}\varepsilon e^{-\alpha s})^{2p_{0}}ds\right) ^{\frac{1}{2p_{0}}}t^{\frac{2p_{0}-1}{2p_{0}}}\nonumber \\ \le&\,\,\frac{M_{1}\varepsilon t^{\frac{2p_{0}-1}{2p_{0}}}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{2p_{0}}}}.\nonumber \end{aligned}$$
(6.5.22)

Similarly, we have

$$\begin{aligned} \left\| \int _{0}^{t}\nabla w (\cdot ,s)ds \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{2}\varepsilon t^{\frac{2p_{0}-1}{2p_{0}}}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{2p_{0}}}}. \end{aligned}$$
(6.5.23)

From Lemma 6.3, noticing \(p_0>1\), \(\alpha -\delta _v<0\), for all \(t\ge 0\), we obtain

$$\begin{aligned} t^{\frac{2p_0-1}{2p_0}}e^{(\alpha -\delta _v)t}\le \left( \frac{2p_0-1}{2ep_0(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{2p_0}}. \end{aligned}$$
(6.5.24)

Inserting (6.5.22) and (6.5.23) into (6.5.21) and using (6.5.24), we obtain

$$\begin{aligned}&\left\| \nabla v\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\\ \le&\left[ 1+\frac{(\alpha _{u}M_{1}+\alpha _{w}M_{2})\left\| v_{0}\left( \cdot \right) \right\| _{L^{\infty }\left( \varOmega \right) } }{\left( 2\alpha p_{0} \right) ^{\frac{1}{2p_{0}}}}\left( \frac{2p_0-1}{2ep_0(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{2p_0}} \right] \varepsilon e^{-\alpha t}. \end{aligned}$$

Therefore, (6.5.19) results from (6.5.7).

To obtain the estimate of \(\Vert \varDelta v(\cdot ,t)\Vert _{L^{p_0}(\varOmega )}\) for \(t\ge 0\), we need the following lemma.

Lemma 6.29

Under the assumptions of Theorem 6.3, for all \(t\in (0,T)\), we have

$$\begin{aligned} \left( \int _{0}^{t}\int _{\varOmega }\left| \varDelta u\left( x,s \right) \right| ^{p_{0}}dxds\right) ^{\frac{1}{p_0}}\le K_{1}(t) \end{aligned}$$
(6.5.25)

and

$$\begin{aligned} \left( \int _{0}^{t}\int _{\varOmega }\left| \varDelta w\left( x,s \right) \right| ^{p_{0}}dxds\right) ^{\frac{1}{p_0}}\le K_{2}(t), \end{aligned}$$
(6.5.26)

where

$$\begin{aligned} K_1(t)=&\,\,C_{p_0}\left( \frac{\varepsilon ^2\xi _{u}M_1M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon ^2\xi _{u}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon M_0|\varOmega |^{\frac{1}{p_0}}e^{k t}\left( \varepsilon \rho _uM_0+\frac{1}{2}(\delta _v-\alpha )\right) }{\left( kp_{0}\right) ^{\frac{1}{p_0}}}\right. \\&\left. +\varepsilon (1+|\varOmega |^{\frac{1}{p_0}})\right) \end{aligned}$$

and

$$K_2(t)=C_{p_0}\left( \frac{\varepsilon ^2\xi _{w}M_2M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon ^2\xi _{w}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon ^2\rho _wM_0^2|\varOmega |^{\frac{1}{p_0}}e^{k t}}{\left( kp_{0}\right) ^{\frac{1}{p_0}}}+\varepsilon (1+|\varOmega |^{\frac{1}{p_0}})\right) $$

with \(k=\min \left\{ \frac{1}{2}(\delta _v-\alpha ),\delta _w\right\} \in (0,\alpha )\).

Proof

Denote \(G\left( x,t \right) =-\xi _{u}\nabla \cdot \left( u\nabla v \right) -\rho _{u}uz+\frac{1}{2}(\delta _v-\alpha ) u\), then \(u_t=\varDelta u-\frac{1}{2}(\delta _v-\alpha ) u+G(x,t)\) and

$$\begin{aligned}&\left( \int _{0}^{t}e^{kp_0s}\int _{\varOmega }\left| G\left( x,s \right) \right| ^{p_{0}}dxds\right) ^{\frac{1}{p_0}} \nonumber \\ \le&\,\left( \int _{0}^{t}e^{kp_0s}\left\| (- \xi _{u}\nabla u\nabla v -\xi _{u} u\varDelta v-\rho _{u}uz+\frac{1}{2}(\delta _v-\alpha ) u)(\cdot ,s)\right\| _{L^{p_{0}}(\varOmega )}^{p_{0}}ds\right) ^{\frac{1}{p_0}}\nonumber \\ \le&\,\left( \int _{0}^{t}e^{kp_0s}\left[ \xi _{u}M_{1} M_{4}\varepsilon ^{2}e^{-2\alpha s}+ \xi _{u}M_{0}M_{5}\varepsilon ^{2}e^{-\alpha s}+\rho _uM_0^2\varepsilon ^2|\varOmega |^{\frac{1}{p_0}}\right. \right. \\&\left. \left. +\frac{1}{2}(\delta _v-\alpha ) M_0\varepsilon |\varOmega |^{\frac{1}{p_0}}\right] ^{p_0}ds\right) ^{\frac{1}{p_0}}\nonumber \\ \le&\,\,\frac{\varepsilon ^2\xi _{u}M_1M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon ^2\xi _{u}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon M_0|\varOmega |^{\frac{1}{p_0}}e^{k t}\left( \varepsilon \rho _uM_0+\frac{1}{2}(\delta _v-\alpha )\right) }{\left( kp_{0}\right) ^{\frac{1}{p_0}}}.\nonumber \end{aligned}$$
(6.5.27)

According to Lemma 3.2, we obtain

$$\begin{aligned}&\left( \int _{0}^{t}\int _{\varOmega }\left| \varDelta u\left( x,s \right) \right| ^{p_{0}}dxds\right) ^{\frac{1}{p_0}} \le K_{1}(t).\nonumber \end{aligned}$$

Similarly, denote \(F\left( x,t \right) =-\xi _{w}\nabla \cdot \left( w\nabla v \right) +\rho _{w}uz,\) then \(w_t=\varDelta w-\delta _ww+F(x,t)\) and

$$\begin{aligned}&\left( \int _{0}^{t}e^{kp_0s}\int _{\varOmega }\left| F\left( x,s \right) \right| ^{p_{0}}dxds\right) ^{\frac{1}{p_0}}\nonumber \\ \le&\,\left( \int _{0}^{t}e^{kp_0s}\left\| (- \xi _{w}\nabla w\nabla v -\xi _{w} w\varDelta v+\rho _{w}uz)(\cdot ,s)\right\| _{L^{p_{0}}(\varOmega )}^{p_{0}}ds\right) ^{\frac{1}{p_0}}\nonumber \\ \le&\,\left( \int _{0}^{t}e^{kp_0s}\left[ \xi _{w}M_{2} M_{4}\varepsilon ^{2}e^{-2\alpha s}+\xi _{w} M_{0}M_{5}\varepsilon ^{2}e^{-\alpha s}+\rho _{w}M_{0}^{2}\varepsilon ^{2} |\varOmega |^{\frac{1}{p_{0}}}\right] ^{p_0}ds\right) ^{\frac{1}{p_{0}}}\nonumber \\ \le&\,\,\frac{\varepsilon ^2\xi _{w}M_2M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon ^2\xi _{w}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon ^2\rho _wM_0^2|\varOmega |^{\frac{1}{p_0}}e^{k t}}{\left( kp_{0}\right) ^{\frac{1}{p_0}}},\nonumber \end{aligned}$$

thus from Lemma 3.2, we obtain

$$\left( \int _{0}^{t}\int _{\varOmega }\left| \bigtriangleup w\left( x,s \right) \right| ^{p_{0}}dxds\right) ^{\frac{1}{p_0}}\le K_{2}(t)$$

and thereby complete the proof.

Lemma 6.30

Under the assumptions of Theorem 6.3, for all \(t\in (0,T)\),

$$\begin{aligned} \left\| \varDelta v\left( \cdot ,t \right) \right\| _{L^{p_{0}}(\varOmega )}\le \frac{M_{5}}{2}\varepsilon e^{-\alpha t}. \end{aligned}$$
(6.5.28)

Proof

By \(v\left( \cdot , t \right) =v_{0}\left( \cdot \right) e^{-\int _{0}^{t}\left( \alpha _{u}u+\alpha _{w}w+\delta _v \right) ds }\), we get

$$\begin{aligned}&\left\| \varDelta v\left( \cdot ,s \right) \right\| _{L^{p_{0}}(\varOmega )}\nonumber \\ \le&\,\left\| div\left( \nabla v_{0}(\cdot )e^{-\int _{0}^{t}\left( \alpha _{u}u+\alpha _{w}w +\delta _v \right) (\cdot ,s)ds } \right) \right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\&+\left\| div\left( v_{0}(\cdot )\nabla e^{-\int _{0}^{t}\left( \alpha _{u}u+\alpha _{w}w +\delta _v \right) (\cdot ,s)ds } \right) \right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\, \left\| \varDelta v_{0}(\cdot )e^{-\int _{0}^{t} \left( \alpha _{u}u+\alpha _{w} w +\delta _v \right) (\cdot ,s)ds }\right\| _{L^{p_{0}}\left( \varOmega \right) }+\left\| v_{0}(\cdot )\varDelta e^{-\int _{0}^{t} \left( \alpha _{u}u+\alpha _{w} w +\delta _v \right) (\cdot ,s)ds}\right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\&+2\left\| \nabla v_{0} (\cdot )e^{-\int _{0}^{t} \left( \alpha _{u} u+\alpha _{w}w +\delta _v \right) (\cdot ,s)ds} \int _{0}^{t} \left( \alpha _{u}\nabla u+\alpha _{w}\nabla w \right) (\cdot ,s)ds \right\| _{L^{p_{0}}\left( \varOmega \right) } \\ \le&\,\left\| \varDelta v_{0}(\cdot )e^{-\delta _v t} \right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\&+\left\| v_{0} (\cdot )e^{-\int _{0}^{t} \left( \alpha _{u}u+\alpha _{w} w +\delta _v \right) (\cdot ,s)ds }\left[ \int _{0}^{t} \left( \alpha _{u}\nabla u+\alpha _{w}\nabla w \right) (\cdot ,s)ds \right] ^{2}\right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\&+\left\| v_{0} (\cdot )e^{-\int _{0}^{t} \left( \alpha _{u} u+\alpha _{w}w +\delta _v \right) (\cdot ,s) ds }\int _{0}^{t} \left( \alpha _{u}\varDelta u+\alpha _{w}\varDelta w \right) \left( \cdot ,s \right) ds\right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\&+2\left\| e^{-\delta _v t} \nabla v_{0}(\cdot )\int _{0}^{t} \left( \alpha _{u}\nabla u+\alpha _{w}\nabla w \right) \left( \cdot ,s \right) ds \right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\left\| \varDelta v_{0}(\cdot )e^{-\delta _v t} \right\| _{L^{p_{0}}\left( \varOmega \right) }+\Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )}e^{-\delta _v t}\left\| \left[ \int _{0}^{t} \left( \alpha _{u}\nabla u+\alpha _{w}\nabla w \right) (\cdot ,s)\right] ^{2} \right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\&+2\left\| \nabla v_{0}(\cdot ) \right\| _{L^{2p_{0}}(\varOmega )}e^{-\delta _v t}\left\| \int _{0}^{t}(\alpha _{u}\nabla u+\alpha _{w}\nabla w)(\cdot ,s)ds \right\| _{L^{2p_{0}}(\varOmega )}\nonumber \\&+\Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )}e^{-\delta _v t}\left\| \int _{0}^{t} \left( \alpha _{u}\varDelta u+\alpha _{w}\varDelta w \right) \left( \cdot ,s \right) ds \right\| _{L^{p_{0}}\left( \varOmega \right) }.\nonumber \end{aligned}$$
(6.5.29)

From (6.5.22) and (6.5.23), we obtain

$$\begin{aligned}&e^{-(\delta _v-\alpha ) t}\left\| \left[ \int _{0}^{t}(\alpha _{u}\nabla u+\alpha _{w}\nabla w )\left( \cdot ,s \right) ds \right] ^{2} \right\| _{L^{p_{0}}(\varOmega )}\nonumber \\ \le&\,\,\alpha _{u}^2e^{-(\delta _v-\alpha ) t}\left\| \left( \int _{0}^{t}\nabla u\left( \cdot ,s \right) ds \right) ^{2} \right\| _{L^{p_{0}}(\varOmega )}+\alpha _{w}^{2}e^{-(\delta _v-\alpha ) t}\left\| \left( \int _{0}^{t}\nabla w\left( \cdot ,s \right) ds \right) ^{2} \right\| _{L^{p_{0}}(\varOmega )}\nonumber \\&+2\alpha _{u}\alpha _{w}e^{-(\delta _v-\alpha ) t}\left\| \int _{0}^{t}\nabla u\left( \cdot ,s \right) ds \int _{0}^{t}\nabla w\left( \cdot ,s \right) ds\right\| _{L^{p_{0}}(\varOmega )}\nonumber \\ \le&\,\, \alpha _{u}^2e^{-(\delta _v-\alpha ) t}\left\| \int _{0}^{t}\nabla u\left( \cdot ,s \right) ds \right\| _{L^{2p_{0}}(\varOmega )}^2+\alpha _{w}^2e^{-(\delta _v-\alpha ) t}\left\| \int _{0}^{t}\nabla w\left( \cdot ,s \right) ds \right\| _{L^{2p_{0}}(\varOmega )}^2\nonumber \\&+2\alpha _{u}\alpha _{w}e^{-(\delta _v-\alpha ) t}\left\| \int _{0}^{t}\nabla u\left( \cdot ,s \right) ds \right\| _{L^{2p_{0}}(\varOmega )}\left\| \int _{0}^{t}\nabla w\left( \cdot ,s \right) ds \right\| _{L^{2p_{0}}(\varOmega )} \\ \le&\,\, \frac{(\alpha _{u}M_{1}\varepsilon )^{2}t^{\frac{2p_{0}-1}{p_{0}}}e^{-(\delta _v-\alpha ) t}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{p_{0}}}}+\frac{2\alpha _{u}\alpha _{w}M_{1}M_{2}\varepsilon ^{2} t^{\frac{2p_{0}-1}{p_{0}}}e^{-(\delta _v-\alpha ) t}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{p_{0}}}}\nonumber \\&+\frac{(\alpha _{w}M_{2}\varepsilon )^{2}t^{\frac{2p_{0}-1}{p_{0}}}e^{-(\delta _v-\alpha ) t}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{p_{0}}}}\nonumber \\ \le&\,\,\frac{(\alpha _{u}M_{1}+ \alpha _{w}M_{2})^{2}\varepsilon ^{2}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{p_{0}}}}\left( \frac{2p_0-1}{ep_0(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{p_0}}\nonumber \end{aligned}$$
(6.5.30)

and

$$\begin{aligned}&2\left\| \nabla v_{0}(\cdot ) \right\| _{L^{2p_{0}}(\varOmega )}e^{-\delta _v t}\left\| \int _{0}^{t}(\alpha _{u}\nabla u+\alpha _{w}\nabla w)(\cdot ,s)ds \right\| _{L^{2p_{0}}(\varOmega )}\nonumber \\ \le&\,\,2\varepsilon ^2e^{-\alpha t}\frac{\alpha _uM_1+\alpha _wM_2}{(2\alpha p_0)^{\frac{1}{2p_0}}}\left( \frac{2p_0-1}{2p_0e(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{2p_0}}. \end{aligned}$$
(6.5.31)

From Lemma 6.3, noticing \(0<k<\delta _v-\alpha \), \(p_0>1\), we obtain

$$t^{\frac{p_0-1}{p_0}}e^{-(\delta _v-\alpha ) t}\le \left( \frac{p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{p_0-1}{p_0}}$$

and

$$t^{\frac{p_0-1}{p_0}}e^{-(\delta _v-\alpha -k) t}\le \left( \frac{p_0-1}{p_0e(\delta _v-\alpha -k)}\right) ^{\frac{p_0-1}{p_0}}$$

for all \(t\ge 0\), which together with Lemma 6.28 implies

$$\begin{aligned}&e^{-(\delta _v-\alpha ) t}\left\| \int _{0}^{t} \left( \alpha _{u}\varDelta u+\alpha _{w}\varDelta w \right) \left( \cdot \right) d\tau \right\| _{L^{p_{0}}\left( \varOmega \right) }\nonumber \\ \le&\,\left( \int _0^t\int _{\varOmega }|\alpha _u \varDelta u+\alpha _w\varDelta w|^{p_0}(x,s)dxds\right) ^{\frac{1}{p_0}}t^{\frac{p_0-1}{p_0}}e^{-(\delta _v-\alpha ) t}\nonumber \\ \le&\,\left( \alpha _u K_1(t)+\alpha _w K_2(t)\right) t^{\frac{p_0-1}{p_0}}e^{-(\delta _v-\alpha ) t} \\ =&\,\,\varepsilon C_{p_0}\alpha _ut^{\frac{p_0-1}{p_0}}e^{-(\delta _v-\alpha ) t}\left( \frac{\varepsilon \xi _{u}M_1M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{u}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \nonumber \\&+\varepsilon C_{p_0}\alpha _wt^{\frac{p_0-1}{p_0}}e^{-(\delta _v-\alpha ) t} \left( \frac{\varepsilon \xi _{w}M_2M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{w}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \nonumber \\&+\varepsilon \frac{C_{p_0}M_0|\varOmega |^{\frac{1}{p_0}}}{\left( k p_{0}\right) ^{\frac{1}{p_0}}}t^{\frac{p_0-1}{p_0}}e^{-(\delta _v-\alpha -k) t} \left( \varepsilon \alpha _u\rho _uM_0+\frac{1}{2}\alpha _u(\delta _v-\alpha ) +\varepsilon \rho _w\alpha _wM_0\right) \nonumber \\ \le&\,\, \varepsilon C_{p_0}\alpha _u\left( \frac{p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{p_0-1}{p_0}}\left( \frac{\varepsilon \xi _{u}M_1M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{u}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \nonumber \\&+\varepsilon C_{p_0}\alpha _w\left( \frac{p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{p_0-1}{p_0}} \left( \frac{\varepsilon \xi _{w}M_2M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{w}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \nonumber \\&+\varepsilon \frac{C_{p_0}M_0|\varOmega |^{\frac{1}{p_0}}}{\left( k p_{0}\right) ^{\frac{1}{p_0}}}\left( \frac{p_0-1}{p_0e(\delta _v-\alpha -k)}\right) ^{\frac{p_0-1}{p_0}} \left( \varepsilon \alpha _u\rho _uM_0+\frac{1}{2}\alpha _u(\delta _v-\alpha ) +\varepsilon \rho _w\alpha _wM_0\right) \nonumber . \end{aligned}$$
(6.5.32)

Inserting (6.5.30), (6.5.31) and (6.5.32) into (6.5.29), we get

$$\begin{aligned}&\left\| \varDelta v\left( \cdot ,t \right) \right\| _{L^{p_{0}}(\varOmega )}\nonumber \\ \le&\,\,\varepsilon e^{-\alpha t}+\varepsilon ^2\Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )} e^{-\alpha t}\frac{(\alpha _{u}M_{1}+ \alpha _{w}M_{2})^{2}}{\left( 2\alpha p_{0} \right) ^{\frac{1}{p_{0}}}}\left( \frac{2p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{p_0}}\nonumber \\&+2\varepsilon ^2e^{-\alpha t}\frac{\alpha _uM_1+\alpha _wM_2}{(2\alpha p_0)^{\frac{1}{2p_0}}}\left( \frac{2p_0-1}{2p_0e(\delta _v-\alpha )}\right) ^{\frac{2p_0-1}{2p_0}} \\&+\varepsilon \Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )} e^{-\alpha t}C_{p_0}\alpha _u\left( \frac{p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{p_0-1}{p_0}}\nonumber \\&\cdot \left( \frac{\varepsilon \xi _{u}M_1M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{u}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \nonumber \\&+\varepsilon \Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )} e^{-\alpha t}C_{p_0}\alpha _w\left( \frac{p_0-1}{p_0e(\delta _v-\alpha )}\right) ^{\frac{p_0-1}{p_0}} \nonumber \\&\cdot \left( \frac{\varepsilon \xi _{w}M_2M_4}{\left( (2\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+\frac{\varepsilon \xi _{w}M_{0}M_{5}}{\left( (\alpha -k)p_{0}\right) ^{\frac{1}{p_0}}}+1+|\varOmega |^{\frac{1}{p_0}}\right) \nonumber \\&+\frac{\varepsilon \Vert v_0(\cdot )\Vert _{L^\infty (\varOmega )} e^{-\alpha t}C_{p_0}M_0|\varOmega |^{\frac{1}{p_0}}}{\left( k p_{0}\right) ^{\frac{1}{p_0}}}\left( \frac{p_0-1}{p_0e(\delta _v-\alpha -k)}\right) ^{\frac{p_0-1}{p_0}}\nonumber \\&\cdot \left( \varepsilon \alpha _u\rho _uM_0+\frac{1}{2}\alpha _u(\delta _v-\alpha ) +\varepsilon \rho _w\alpha _wM_0\right) \nonumber . \end{aligned}$$
(6.5.33)

Therefore, (6.5.28) follows from (6.5.8).

Lemma 6.31

Under the assumptions of Theorem 6.3, for all \(t\in (0,T)\),

$$\begin{aligned}&\left\| \left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q \right) \left( \cdot ,t \right) -e^{t\varDelta }\left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0}\right) \left( \cdot \right) \right\| _{L^{\infty }\left( \varOmega \right) }\nonumber \\ \le&\,\, \frac{M_{6}}{2} \varepsilon e^{-\alpha t}. \end{aligned}$$
(6.5.34)

Proof

From Lemma 1.1(iv) and (6.5.11), using Lemma 4.3, it follows that

$$\begin{aligned}&\left\| \left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q \right) \left( \cdot ,t \right) -e^{t\varDelta }\left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0}\right) \left( \cdot \right) \right\| _{L^{\infty }\left( \varOmega \right) }\nonumber \\ \le&\,\, \int _{0}^{t}\left\| e^{\left( t-s \right) \varDelta } \left[ \xi _{u}\nabla \cdot \left( u\nabla v \right) +\frac{\rho _{u}+\rho _{z}}{\rho _{w}}\xi _{w}\nabla \cdot \left( w\nabla v \right) +\xi _{z}\nabla \cdot \left( z\nabla v \right) \right] \left( \cdot ,s \right) \right\| _{L^{\infty }\left( \varOmega \right) }ds\nonumber \\ \le&\,\,\xi _{u}c_4\int _{0}^{t}\left[ 1+\left( t-s \right) ^{-\frac{1}{2}-\frac{N}{4p_{0}}}\right] e^{-\lambda _{1}\left( t-s \right) }\left\| \left( u\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\&+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}\xi _{w}c_4\int _{0}^{t}\left[ 1+\left( t-s \right) ^{-\frac{1}{2}-\frac{N}{4p_{0}}}\right] e^{-\lambda _{1}\left( t-s \right) }\left\| \left( w\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\&+\xi _{z}c_4\int _{0}^{t}\left[ 1+\left( t-s \right) ^{-\frac{1}{2}-\frac{N}{4p_{0}} }\right] e^{-\lambda _{1}\left( t-s \right) }\left\| \left( z\nabla v \right) \left( \cdot ,s \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }ds\nonumber \\ \le&\,\, 2c_{10}c_4M_{0}M_{4}\varepsilon ^{2} e^{-\alpha t}\left( \xi _{u}+ \frac{\rho _{u}+\rho _{z}}{\rho _{w}}\xi _{w}+\xi _{z} \right) ,\nonumber \end{aligned}$$

and in view of (6.5.9), we already arrive at (6.5.34) and complete the proof.

Proof of Theorem 6.3.  First let us verify \(T=T_{max}\) by contraction. In fact, suppose that \(T< T_{max}\), then from Lemmas 6.25–6.31, it follows

$$\left\| \nabla u\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{1}}{2}\varepsilon e^{-\alpha t}, $$
$$\left\| \nabla w\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{2}}{2}\varepsilon e^{-\alpha t}, $$
$$\left\| \nabla z\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{3}}{2}\varepsilon e^{-\alpha t}, $$
$$\left\| \nabla v\left( \cdot ,t \right) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le \frac{M_{4}}{2}\varepsilon e^{-\alpha t},$$
$$\left\| \varDelta v\left( \cdot ,t \right) \right\| _{L^{p_{0}}\left( \varOmega \right) }\le \frac{M_{5}}{2}\varepsilon e^{-\alpha t} ,$$
$$\left\| \left( u+\frac{\rho _{u}+\rho _{z}}{\rho _{w}}w+z+Q \right) \left( \cdot ,t \right) -e^{t\varDelta }\left( u_{0}+\frac{\rho _{u}+\rho _{z}}{\rho _{w}} w_{0}+z_{0}\right) \left( \cdot \right) \right\| _{L^{\infty }\left( \varOmega \right) }\le \frac{M_{6}}{2}\varepsilon e^{-\alpha t} ,$$

for all \(t\in \left( 0,T \right) \), which contradicts the definition of T.

Next, we show that \(T_{max}=\infty \). In fact, if \(T_{max}<\infty \), then in view of the definition of T, we obtain

$$\begin{aligned}&\lim _{t\nearrow T_{max}}\left( \left\| u\left( \cdot ,t \right) \right\| _{W^{1,2p_0}\left( \varOmega \right) }+\left\| v\left( \cdot ,t \right) \right\| _{W^{2,p_0}\left( \varOmega \right) }+\left\| w\left( \cdot ,t \right) \right\| _{W^{1,2p_0}\left( \varOmega \right) }+\left\| z\left( \cdot ,t \right) \right\| _{W^{1,2p_0}\left( \varOmega \right) }\right) \\&\quad < \infty , \end{aligned}$$

which contradicts with (6.5.1) in Lemma 6.22. Therefore, we have \(T_{max}=\infty \).

Integrating the equation of u in (6.1.12) over \(\varOmega \), we have

$$\int _{\varOmega }u\left( x,t \right) dx=\int _{\varOmega }u_{0}(x)dx-\rho _{u}\int _{0}^{t}\int _{\varOmega }(uz)\left( x,s\right) dxds,$$

which along with the nonnegative property of u, z and the fact that \(\Vert u(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le M_0\varepsilon \) warrants that \(\bar{u}\left( t \right) :=\frac{1}{|\varOmega |}\int _{\varOmega }u\left( x,t \right) dx\) is noncreasing with respect to time t and and its limit \(t\rightarrow \infty \) exists, that is,

$$\begin{aligned} \lim _{t\rightarrow \infty }\int _{\varOmega }u\left( x,t \right) dx=\int _{\varOmega }u_{0}\left( x \right) dx-\rho _{u}\int _{0}^{+\infty }\int _{\varOmega }(uz)\left( x,s \right) dxds \end{aligned}$$

as well as

$$\begin{aligned} \lim _{t\rightarrow \infty }\bar{u}\left( t \right) =\frac{1}{\left| \varOmega \right| }\left( \int _{\varOmega }u_{0}\left( x \right) dx-\rho _{u}\int _{0}^{+\infty }\int _{\varOmega }(uz)\left( x,s \right) dxds \right) :=u^{*}, \end{aligned}$$

which implies

$$ 0\le \frac{\rho _{u}}{\left| \varOmega \right| }\int _{t}^{\infty }\int _{\varOmega } (uz)\left( x,s \right) dxds= \bar{u}(t)-u^{*}\rightarrow 0$$

as \(t\rightarrow \infty \). On the other hand, by Poincare’s inequality, there exists \(k_1>0\) such that

$$\left\| u(\cdot ,t)-\bar{u}(t) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\le k_{1}\left\| \nabla u(\cdot ,t)\right\| _{L^{2p_{0}}\left( \varOmega \right) }.$$

By Embedding theorem, we know \(W^{1,2p_{0}}(\varOmega ) \hookrightarrow C^{1-\frac{N}{p_0}}(\overline{\varOmega })\), for \(p_{0}>\max \{1,\frac{N}{2}\}\). There exists \(k_{2}> 0\), such that

$$\begin{aligned} \left\| u(\cdot ,t)-\bar{u}(t)\right\| _{L^{\infty }\left( \varOmega \right) }\le&\,\,k_{2}\left\| u (\cdot ,t)-\bar{u}(t)\right\| _{W^{1,2p_{0}}\left( \varOmega \right) } \\ \le&\,\,k_{2}\left( 1+k_{1} \right) \left\| \nabla u(\cdot ,t) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\rightarrow 0 \end{aligned}$$

as \(t\rightarrow \infty \). Thus,

$$\begin{aligned} \left\| u(\cdot ,t)-u^{*}\right\| _{L^{\infty }\left( \varOmega \right) }\nonumber \le&\,\left\| u(\cdot ,t)-\bar{u}(t) \right\| _{L^{\infty }\left( \varOmega \right) }+\left\| \bar{u}(t)-u^{*} \right\| _{L^{\infty }\left( \varOmega \right) }\rightarrow 0. \end{aligned}$$

Now, we consider a linear combination of u and w

$$H:=\rho _{w}u+\rho _{u}w,$$

then

$$H_{t}=\rho _{w}u_{t}+\rho _{u}w_{t}=\varDelta H-\rho _{w}\xi _{u}\nabla \cdot \left( u\nabla v \right) -\rho _{u}\xi _{w}\nabla \cdot \left( w\nabla v \right) -\rho _{u}\delta _{w}w.$$

Accordingly,

$$\int _{\varOmega }H\left( x,t \right) dx=\int _{\varOmega }H(x,0)dx-\rho _{u}\delta _w\int _{0}^{t}\int _{\varOmega } w\left( x,s \right) dxds.$$

Similarly, we obtain

$$\begin{aligned} \lim _{t\rightarrow \infty }\overline{H}\left( t \right) =\frac{1}{\left| \varOmega \right| }\left( \int _{\varOmega }H\left( x,0 \right) dx-\rho _{u}\delta _{w}\int _{0}^{+\infty }\int _{\varOmega }w\left( x,s \right) dxds \right) :=H^{*}, \end{aligned}$$

and

$$\begin{aligned} \left\| H(\cdot ,t)-\bar{H}(t)\right\| _{L^{\infty }\left( \varOmega \right) }\le&\,\,k_{2}\left\| H (\cdot ,t)-\bar{H}(t)\right\| _{W^{1,2p_{0}}\left( \varOmega \right) }\\ \le&\,\,k_{2}\left( 1+k_{1} \right) \left\| \nabla H(\cdot ,t) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\rightarrow 0. \end{aligned}$$

Then

$$\Vert H(\cdot ,t)-H^{*}\Vert _{L^\infty (\varOmega )}\le \Vert H(\cdot ,t)-\overline{H}(t)\Vert _{L^\infty (\varOmega )}+\Vert \overline{H}(t)-H^{*}\Vert _{L^\infty (\varOmega )}\rightarrow 0$$

as \(t\rightarrow \infty \). Then denote \(w^*:=\frac{1}{\rho _u}(H^*-\rho _w u^*)\), as \(t\rightarrow \infty \), we obtain

$$\begin{aligned}&\rho _{u}\left\| w(\cdot ,t)-w^{*} \right\| _{L^{\infty }(\varOmega )}\\ =&\,\left\| \rho _{u}w(\cdot ,t)-(H^{*}-\rho _{w}u^{*}) \right\| _{L^{\infty }(\varOmega )}\\ \le&\,\left\| H(\cdot ,t)-H^{*} \right\| _{L^{\infty }(\varOmega )}+\rho _{w}\left\| u(\cdot ,t)-u^{*} \right\| _{L^{\infty }(\varOmega )}\rightarrow 0. \end{aligned}$$

Next, we consider a linear combination of u, w and z. Let

$$I=\rho _{z}(u+w)+(\rho _u+\rho _w)z.$$

Then

$$\begin{aligned} I_{t} =&\,\,\varDelta I-\rho _{z}\xi _{u}\nabla \cdot \left( u\nabla v \right) -\rho _{z}\xi _{w}\nabla \cdot \left( w\nabla v \right) -(\rho _u+\rho _w)\xi _{z}\nabla \cdot \left( z\nabla v \right) \\&-\left( \rho _{z}\delta _{w}-(\rho _u+\rho _w)\beta \right) w-(\rho _u+\rho _w)\delta _{z}z. \end{aligned}$$

Accordingly,

$$\begin{aligned}&\int _{\varOmega }I\left( x,t \right) dx \\ =&\,\,\int _{\varOmega }I(x,0)dx -\int _{0}^{t}\int _{\varOmega }[\left( \rho _{z}\delta _{w}-(\rho _u+\rho _w)\beta \right) w(x,s)+(\rho _u+\rho _w)\delta _{z}z(x,s)]dxds, \end{aligned}$$

where \(\rho _{z}\delta _{w}-(\rho _u+\rho _w)\beta >0\).

Similarly, we obtain

$$\begin{aligned} \quad \lim _{t\rightarrow \infty }\overline{I}\left( t \right) =&\,\,\frac{1}{\left| \varOmega \right| }\left( \int _{\varOmega }I\left( x,0 \right) dx\right. \\&\left. -\int _{0}^{+\infty }\int _{\varOmega }[\left( \rho _{z}\delta _{w}-(\rho _u+\rho _w)\beta \right) w(x,s)+(\rho _u+\rho _w)\delta _{z}z(x,s)]dxds \right) \\&:=I^{*}, \end{aligned}$$

and

$$\begin{aligned} \left\| I(\cdot ,t)-\bar{I}(t)\right\| _{L^{\infty }\left( \varOmega \right) }\le&\,\,k_{2}\left\| I (\cdot ,t)-\bar{I}(t)\right\| _{W^{1,2p_{0}}\left( \varOmega \right) }\\ \le&\,\,k_{2}\left( 1+k_{1} \right) \left\| \nabla I(\cdot ,t) \right\| _{L^{2p_{0}}\left( \varOmega \right) }\rightarrow 0. \end{aligned}$$

Then

$$\Vert I(\cdot ,t)-I^{*}\Vert _{L^\infty (\varOmega )}\le \Vert I(\cdot ,t)-\overline{I}(t)\Vert _{L^\infty (\varOmega )}+\Vert \overline{I}(t)-I^{*}\Vert _{L^\infty (\varOmega )}\rightarrow 0$$

as \(t\rightarrow \infty \). Then denote \(z^*:=\frac{1}{\rho _u+\rho _w}(I^*-\rho _z(u^*+w^*))\), as \(t\rightarrow \infty \), we obtain

$$\begin{aligned}&(\rho _u+\rho _w)\left\| z(\cdot ,t)-z^{*} \right\| _{L^{\infty }(\varOmega )}\\ =&\,\left\| (\rho _u+\rho _w)z(\cdot ,t)-(I^*-\rho _z(u^*+w^*)) \right\| _{L^{\infty }(\varOmega )}\\ \le&\,\left\| I(\cdot ,t)-I^{*} \right\| _{L^{\infty }(\varOmega )}+\rho _{z}\Vert u(\cdot ,t)-u^{*}\Vert _{L^{\infty }(\varOmega )}+\rho _{z}\Vert w(\cdot ,t)-w^{*}\Vert _{L^{\infty }(\varOmega )}\rightarrow 0. \end{aligned}$$

By contradiction, if \(w^*>0\) or \(z^*>0\), then there exists \(t^*>0\) such that for all \(t>t^*\),

$$\begin{aligned} \frac{d}{dt}\int _{\varOmega }Q(x,t)dx=&\,\,\int _{\varOmega }\left( \left( \frac{\rho _{u}+\rho _{z}}{\rho _{w}}\delta _{w}-\beta \right) w(x,t)+\delta _{z}z(x,t)\right) dx \\ \ge&\,\,\frac{|\varOmega |}{2}\left( \left( \frac{\rho _{u}+\rho _{z}}{\rho _{w}}\delta _{w}-\beta \right) w^*+\delta _{z}z^*\right) >0, \end{aligned}$$

which implies that \(\int _{\varOmega }Q(x,t)dx\rightarrow \infty \) as \(t\rightarrow \infty \) and thus contradicts with that \(\Vert Q(\cdot ,t)\Vert _{L^\infty (\varOmega )}\le M_0\varepsilon \). Hence, we have \(w^*=z^*=0\). On the other hand, by (6.5.20), it is easy to see that

$$\Vert v(\cdot ,t)\Vert _{L^\infty (\varOmega )}\rightarrow 0.$$

So the proof of Theorem 6.3 is complete.