1. Introduction
When the well mixed suspension of Bacillus subtilis cells is placed in a chamber with the upper surface open to the atmosphere, this kind of aerobic bacteria consume oxygen, swim toward the direction of sufficient oxygen, that is, the surface of the water layer. Then they form a thin boundary layer with dense cells upstream. Below this layer, the cells in the suspension were severely depleted. Because that bacteria are about 10% denser than water, thus, the density of the mixed suspension becomes larger near the water surface than at the bottom. When the density of the upper boundary layer is too high, it becomes unstable and forms a descending bacterial plume. And finally evolved into various patterns [Reference Hillesdon, Pedley and Kessler10, Reference Hillesdon and Pedley11, Reference Kessler14]. Based on these experimental observations, Goldstein et al. [Reference Tuval, Cisneros, Dombrowski, Wolgemuth, Kessler and Goldstein19] proposed the following chemotaxis-fluid model
where $Q=\Omega \times \mathbb {R}^+$, $n$, $c$ represent the bacteria cell density, the oxygen concentration respectively, $\chi >0$ is the sensitivity coefficient of aggregation induced by the concentration changes of oxygen, $J=n\nabla c$ is the chemotactic flux, $-cn$ is the consumption term of oxygen, that is more bacteria, more oxygen are consumed, ${\bf u}$, $\pi$ are the fluid velocity and the associated pressure, the fluid couples to $n$ and $c$ through transports ${\bf u}\cdot \nabla n$, ${\bf u}\cdot \nabla c$ and the gravitational potential $n\nabla \varphi$.
In recent years, this kind of models has been widely studied by many authors. For the studies of Cauchy problem in $\mathbb {R}^N$, we refer to [Reference Duan, Lorz and Markowich9, Reference Kozono, Miura and Sugiyama15, Reference Zhang and Zheng22], and for the bounded domain with zero flux boundary value conditions for $n, c$, and no-slip boundary value condition for $u$, please refer to [Reference Di Francesco, Lorz and Markowich8, Reference Winkler20, Reference Winkler21] or the references therein. From these results, one see that the solutions will convergence to the constant steady states, and there is no pattern formation. Matching the experiment descriptions, the following mixed boundary conditions is proposed [Reference Chertock, Fellner, Kurganov, Lorz and Markowich6, Reference Lee and Kim16]: the boundary condition at the top $\Gamma _{top}$ describes the fluid-air surface, where there is no cell flux, the oxygen will be saturated with the air oxygen concentration $c_{air}$ and the vertical fluid velocity and the tangential fluid stress are supposed to be zero
where $\nu$ denotes the outward unit normal vector of the boundary,$\tau$ denotes unit tangent vector of the boundary, and
At the bottom of the domain $\Gamma _{B}$, the cell and oxygen fluxes and the fluid velocity are assumed to be zero:
Finally, periodic boundary conditions at the sides of the domain are imposed in order to avoid any impact of these boundaries. And some numerical results are given in [Reference Chertock, Fellner, Kurganov, Lorz and Markowich6, Reference Lee and Kim16]. However, very little theoretical research in this regard has been carried out. Peng, Xiang [Reference Peng and Xiang18] considered mixed boundary value problem in an unbounded strip domain of $\mathbb {R}^3$ with the consumption term $cn$ being replaced by $cn^{\gamma }$ ($\gamma \ge 2$) , and established the global existence and convergence of small strong solutions around an equilibrium state. Besides the above boundary value problem, such kind of mixed non-homogeneous boundary value problem
also be considered [Reference Braukhoff4, Reference Braukhoff and Tang5, Reference Jin13], in which, the global classical solution in two dimensional space [Reference Braukhoff4, Reference Braukhoff and Tang5] and time periodic solution in two and three dimensional space [Reference Jin13] are established respectively.
Adopting the realistic boundary conditions mentioned above, in the present paper, we consider the model (1.1) in a two-dimensional strip periodic domain
with period $l$ and the following boundary conditions, that is
or
and periodic boundary conditions at the left and right sides of the domain $\Omega _l$ are imposed, where ${\bf u}=(u, v)$, $\Gamma _{T}$ is the upper boundary of the rectangular area, $\Gamma _{B}$ denotes the lower boundary, $c_{air}$ is a positive constant, $\tau =0$ or $1$. It is easy to see that when $\tau =0$, it is corresponding to the Dirichlet boundary condition, when $\tau =1$, it is corresponding to the Robin boundary condition.
We also give the initial value as follows
Throughout this paper, we assume that
We give the global existence theorems as follows.
Theorem 1.1 Assume (1.5) holds. Then the problem (1.1), (1.2), and (1.4) admits a unique global bounded classical solution $(n, c, {\bf u}, \pi )$ ($\pi$ is unique up to a constant) with $(n, c, {\bf u})\in C^{2+\alpha, 1+\frac {\alpha }2}(\overline \Omega _l\times (0,+\infty ))\cap C^{\alpha, \frac {\alpha }2}(\overline \Omega _l\times [0,+\infty ))$, $\nabla \pi \in C^{\alpha, \frac {\alpha }2}(\overline \Omega _l\times (0,+\infty ))$, $n, c\ge 0$,
where the constant $C$ depends only on $n_0$, $c_0$, ${\bf u}_0$, $\alpha$, $l$, $\chi$.
Theorem 1.2 Assume (1.5) holds. When (i) $\tau =0$ with appropriately small $\chi$; or (ii)$\tau =1$ with any large $\chi$, the problem (1.1), (1.3), and (1.4) admits a unique global bounded classical solution $(n, c, {\bf u}, \pi )$ ($\pi$ is unique up to a constant) with $(n, c, {\bf u})\in C^{2+\alpha, 1+\frac {\alpha }2}(\overline \Omega _l\times (0,+\infty ))\cap C^{\alpha, \frac {\alpha }2}(\overline \Omega _l\times [0,+\infty ))$, $\nabla \pi \in C^{\alpha, \frac {\alpha }2}(\overline \Omega _l\times (0,+\infty ))$, ${n, c\ge 0}$,
where the constant $C$ depends only on $n_0$, $c_0$, ${\bf u}_0$, $\alpha$, $l$, $\chi$.
2. Preliminaries
Based on Gagliardo-Nirenberg interpolation inequality and Sobolev trace embedding inequality, we infer the following trace interpolation inequality.
Lemma 2.1 For functions $u: \Omega \to \mathbb {R}$ defined on a bounded Lipschitz domain ${\Omega \in \mathbb {R}^N}$, we have
where $0< r\le p\le q\le \frac {(N-1)p}{(N-p)_+}$, $p\ge 1$, $\beta =\frac {r(N(p-q)+p(q-1))}{q(N(p-r)+pr)}$.
Proof.
(i) We first consider the case $p< N$. By Sobolev trace embedding inequality [Reference Adams1, Reference Biezuner3], we have
\[ \|u\|_{L^{p^*}(\partial\Omega)}\le C_1\|Du\|_{L^p(\Omega)}+ C_2\|u\|_{L^p(\partial\Omega)}, \]where $p^*=\frac {(N-1)p}{N-p}$. On the other hand, for any $p\le q\le p^*$,\[ \|u\|_{L^{q}(\partial\Omega)}\le \|u\|_{L^{p^*}(\partial\Omega)}^{1-\alpha}\|u\|_{L^{p}(\partial\Omega)}^{\alpha} \]with $\alpha =\frac {p(p^*-q)}{q(p^*-p)}=\frac {(N-1)p-(N-p)q}{q(p-1)}$ for $p< N$. Combining the above two inequalities, and noticing that $(a+b)^{\alpha }\le a^\alpha +b^{\alpha }$ for any $a, b>0$, $0<\alpha \le 1$, then we have(2.2)\begin{align} \|u\|_{L^{q}(\partial\Omega)}& \le \left( C_1\|Du\|_{L^p(\Omega)}+ C_2\|u\|_{L^p(\partial\Omega)}\right)^{1-\alpha}\|u\|_{L^{p}(\partial\Omega)}^{\alpha} \nonumber\\ & \le\left( C_1^{1-\alpha}\|Du\|_{L^p(\Omega)}^{1-\alpha}+ C_2^{1-\alpha}\|u\|_{L^p(\partial\Omega)}^{1-\alpha}\right)\|u\|_{L^{p}(\partial\Omega)}^{\alpha} \nonumber\\ & \le C_1^{1-\alpha}\|Du\|_{L^p(\Omega)}^{1-\alpha}\|u\|_{L^{p}(\partial\Omega)}^{\alpha} +C_2^{1-\alpha}\|u\|_{L^p(\partial\Omega)}. \end{align}By [Reference Diaz and Veron7], for any $1\le r\le p<+\infty$,(2.3)\begin{align} \|u\|_{L^{p}(\partial\Omega)}& \le C_3\left(\|Du\|_{L^p(\Omega)}+ \|u\|_{L^r(\Omega)}\right)^{\theta}\|u\|_{L^{r}(\Omega)}^{1-\theta} \nonumber\\ & \le C_3\|Du\|_{L^p(\Omega)}^{\theta}\|u\|_{L^{r}(\Omega)}^{1-\theta}+\|u\|_{L^r(\Omega)} \end{align}with $\theta =\frac {N(p-r)+r}{N(p-r)+pr}$. Combining (2.2) and (2.3), we get that\begin{align*} \|u\|_{L^{q}(\partial\Omega)}& \le C_1^{1-\alpha}\|Du\|_{L^p(\Omega)}^{1-\alpha}\left(C_3\|Du\|_{L^p(\Omega)}^{\theta}\|u\|_{L^{r}(\Omega)}^{1-\theta}+\|u\|_{L^r(\Omega)}\right)^{\alpha} \\ & \quad +C_2^{1-\alpha}\left(C_3\|Du\|_{L^p(\Omega)}^{\theta}\|u\|_{L^{r}(\Omega)}^{1-\theta}+\|u\|_{L^r(\Omega)}\right) \\ & \le C_1^{1-\alpha}C_3^{\alpha}\|Du\|_{L^p(\Omega)}^{1-\alpha+\theta\alpha}\|u\|_{L^{r}(\Omega)}^{\alpha(1-\theta)} +C_1^{1-\alpha}\|Du\|_{L^p(\Omega)}^{1-\alpha}\|u\|_{L^r(\Omega)}^{\alpha} \\ & \quad +C_2^{1-\alpha}C_3\|Du\|_{L^p(\Omega)}^{\theta}\|u\|_{L^{r}(\Omega)}^{1-\theta} +C_2^{1-\alpha}\|u\|_{L^r(\Omega)} \\ & = C_1^{1-\alpha}C_3^{\alpha}\|Du\|_{L^p(\Omega)}^{1-\alpha+\theta\alpha}\|u\|_{L^{r}(\Omega)}^{\alpha(1-\theta)} \\ & \quad +C_1^{1-\alpha}\|Du\|_{L^p(\Omega)}^{1-\alpha}\|u\|_{L^r(\Omega)}^{\frac{\alpha(1-\alpha)(1-\theta)}{1-\alpha+\theta\alpha}} \|u\|_{L^r(\Omega)}^{\frac{\alpha\theta}{1-\alpha+\theta\alpha}} \\ & \quad+C_2^{1-\alpha}C_3\|Du\|_{L^p(\Omega)}^{\theta}\|u\|_{L^{r}(\Omega)}^{\alpha(1-\theta)\frac{\theta}{1-\alpha+\theta\alpha}} \|u\|_{L^{r}(\Omega)}^{\frac{(1-\theta)(1-\alpha)}{1-\alpha+\theta\alpha}} \\ & \quad +C_2^{1-\alpha}\|u\|_{L^r(\Omega)} \\ & \le C_4\|Du\|_{L^p(\Omega)}^{1-\alpha+\theta\alpha}\|u\|_{L^{r}(\Omega)}^{\alpha(1-\theta)}+C_5\|u\|_{L^r(\Omega)}. \end{align*}Let $\beta =\alpha (1-\theta )$, by a direct calculation, we see that\[ \beta=\frac{r(N(p-q)+p(q-1))}{q(N(p-r)+pr)},\quad \text{for}\ p< N. \]Then (2.1) is proved for $p< N$, $r\ge 1$.Next, we show that (2.1) also holds for any $r>0$. By the prove above, we see that
(2.4)\begin{equation} \|u\|_{L^{q}(\partial\Omega)}\le C\|Du\|_{L^p(\Omega)}^{1-\tilde\beta}\|u\|_{L^{\tilde r}(\Omega)}^{\tilde \beta}+C\|u\|_{L^{\tilde r}(\Omega)}, \end{equation}for $1\le \tilde r\le p\le q$, $\tilde \beta =\frac {\tilde r(N(p-q)+p(q-1))}{q(N(p-\tilde r)+p\tilde r)}$ for $p< N$. While by Gagliardo-Nirenberg inequality [Reference Li and Lankeit17], for any $0< r<1$,\[ \|u\|_{L^{\tilde r}(\Omega)}\le C_6\|D u\|_{L^p(\Omega)}^{\alpha_1}\|u\|^{1-\alpha_1}_{L^r(\Omega)}+C_7\|u\|_{L^r(\Omega)}, \]with $\frac 1{\tilde r}=(\frac 1p-\frac 1N)\alpha _1+\frac {1-\alpha _1}{r}$. Substituting it into (2.4), we arrive at(2.5)\begin{align} \|u\|_{L^{q}(\partial\Omega)}& \le C\|Du\|_{L^p(\Omega)}^{1-\tilde\beta}\left(C_6\|D u\|_{L^p(\Omega)}^{\alpha_1}\|u\|^{1-\alpha_1}_{L^r(\Omega)}+C_7\|u\|_{L^r(\Omega)}\right)^{\tilde\beta}\nonumber\\ & \quad +C\left(C_6\|D u\|_{L^p(\Omega)}^{\alpha_1}\|u\|^{1-\alpha_1}_{L^r(\Omega)}+C_7\|u\|_{L^r(\Omega)}\right) \nonumber\\ & \le CC_6^{\tilde\beta}\|Du\|_{L^p(\Omega)}^{1-\tilde\beta+\alpha_1\tilde\beta}\|u\|^{\tilde\beta(1-\alpha_1)}_{L^r(\Omega)} +CC_7^{\tilde\beta}\|Du\|_{L^p(\Omega)}^{1-\tilde\beta}\|u\|_{L^r(\Omega)}^{\tilde\beta}\nonumber\\ & \quad +CC_6\|D u\|_{L^p(\Omega)}^{\alpha_1}\|u\|^{1-\alpha_1}_{L^r(\Omega)}+CC_7\|u\|_{L^r(\Omega)}\nonumber\\ & \le C_8\|Du\|_{L^p(\Omega)}^{1-\tilde\beta+\alpha_1\tilde\beta}\|u\|^{\tilde\beta(1-\alpha_1)}_{L^r(\Omega)}+C_9\|u\|_{L^r(\Omega)}. \end{align}By a direct calculation, we see that $\tilde \beta (1-\alpha _1)=\frac {r(N(p-q)+p(q-1))}{q(N(p-r)+pr)}$. Combining with (2.4), we prove (2.1) for any $r>0$.(ii) Next, we prove the case $p\ge N$. For any $q\ge p$, there exists $\tilde p< N$ such that $q<\frac {(N-1)p}{N-\tilde p}$, namely $\frac {q\tilde p}{p}<\frac {(N-1)\tilde p}{N-\tilde p}$, then by the result of (i), we have
\begin{align*} \|u\|_{L^q(\partial\Omega)}& =\|u^{\frac{p}{\tilde p}}\|_{L^{\frac{q\tilde p}{p}}(\partial\Omega)}^{\frac{\tilde p}{p}}\le C_{10}\|Du^{\frac{p}{\tilde p}}\|_{L^{\tilde p}(\Omega)}^{\frac{\tilde p}{p}(1-\gamma)}\|u^{\frac{p}{\tilde p}}\|_{L^{\tilde p}(\Omega)}^{\frac{\tilde p}{p}\gamma}+C_{11}\|u^{\frac{p}{\tilde p}}\|_{L^{\tilde p}(\Omega)}^{\frac{\tilde p}{p}} \\ & \le C_{10}\left\|\frac{p}{\tilde p}u^{\frac{p-\tilde p}{\tilde p}}Du\right\|_{L^{\tilde p}(\Omega)}^{\frac{\tilde p}{p}(1-\gamma)}\|u\|_{L^{p}(\Omega)}^{\gamma}+C_{11}\|u\|_{L^{p}(\Omega)} \\ & \le C_{12}\|u\|_{L^p(\Omega)}^{\frac{p-\tilde p}{p}(1-\gamma)}\|Du\|_{L^{p}(\Omega)}^{\frac{\tilde p}{p}(1-\gamma)}\|u\|_{L^{p}(\Omega)}^{\gamma}+C_{11}\|u\|_{L^{p}(\Omega)} \\ & = C_{12}\|u\|_{L^p(\Omega)}^{\frac{p-\tilde p}{p}(1-\gamma)+\gamma}\|Du\|_{L^{p}(\Omega)}^{\frac{\tilde p}{p}(1-\gamma)}+C_{11}\|u\|_{L^{p}(\Omega)} \end{align*}where $\gamma =\frac {(N-1)p-q(N-\tilde p)}{q\tilde p}$, it is easy to see that $\frac {p-\tilde p}{p}(1-\gamma )+\gamma =\frac {p(q-1)+N(p-q)}{pq}$. We denote $\tilde \beta =\frac {p(q-1)+N(p-q)}{pq}$. The above inequality is equivalent to(2.6)\begin{equation} \|u\|_{L^q(\partial\Omega)}\le C_{12}\|u\|_{L^p(\Omega)}^{\tilde\beta}\|Du\|_{L^{p}(\Omega)}^{1-\tilde\beta}+C_{11}\|u\|_{L^{p}(\Omega)} \end{equation}Similar to the proof of (2.5), using Gagliardo-Nirenberg inequality [Reference Li and Lankeit17], for any $0< r< p$,\[ \|u\|_{L^{p}(\Omega)}\le C_{13}\|D u\|_{L^p(\Omega)}^{\beta_1}\|u\|^{1-\beta_1}_{L^r(\Omega)}+C_{14}\|u\|_{L^r(\Omega)}, \]with $\frac {1}{p}=\left (\frac {1}{p}-\frac {1}{N}\right )\beta _1+\frac {1-\beta _1}r$. Substituting it into (2.6), we arrive at\begin{align*} \|u\|_{L^q(\partial\Omega)}& \le C_{12}\left(C_{13}\|D u\|_{L^p(\Omega)}^{\beta_1}\|u\|^{1-\beta_1}_{L^r(\Omega)}+C_{14}\|u\|_{L^r(\Omega)}\right)^{\tilde\beta}\|Du\|_{L^{p}(\Omega)}^{1-\tilde\beta}\\ & \quad +C_{11}\left(C_{13}\|D u\|_{L^p(\Omega)}^{\beta_1}\|u\|^{1-\beta_1}_{L^r(\Omega)}+C_{14}\|u\|_{L^r(\Omega)}\right) \\ & \le C_{12}C_{13}^{\tilde\beta}\|D u\|_{L^p(\Omega)}^{\beta_1\tilde\beta+1-\tilde\beta}\|u\|^{\tilde\beta(1-\beta_1)}_{L^r(\Omega)} +C_{12}C_{14}^{\tilde\beta}\|Du\|_{L^{p}(\Omega)}^{1-\tilde\beta}\|u\|_{L^r(\Omega)}^{\tilde\beta}\\ & \quad +C_{11}C_{13}\|D u\|_{L^p(\Omega)}^{\beta_1}\|u\|^{1-\beta_1}_{L^r(\Omega)}+C_{11}C_{14}\|u\|_{L^r(\Omega)} \\ & \le C_{15}\|D u\|_{L^p(\Omega)}^{\beta_1\tilde\beta+1-\tilde\beta}\|u\|^{\tilde\beta(1-\beta_1)}_{L^r(\Omega)}+C_{16}\|u\|_{L^r(\Omega)}. \end{align*}By a direct calculation, we get that $\tilde \beta (1-\beta _1)=\frac {r(N(p-q)+p(q-1))}{q(N(p-r)+pr)}$. We complete the proof.
By [Reference Jin12], we have the following lemma.
Lemma 2.2 Let $T>0$, $\tau \in (0, T)$, $\alpha >0$, $\beta >0$, and suppose that $f: [0, T)\to [0, \infty )$ is absolutely continuous, and satisfies
where $\sigma >0$ is a constant, $g(t), h(t)\ge 0$ with $g(t), h(t)\in L^1_{loc}([0, T))$ and
Then for any $t>t_0$, we have
and
where
3. Global classical solution: Neumann-Dirichlet-Navier slip boundary value condition
We first give some notations, which will be used throughout this paper.
Notations: $Q^l=\Omega _l\times \mathbb {R}^+$, $Q_T^l=\Omega _l\times (0, T)$, $\|f\|_{L^q}:=\|f\|_{L^q(\Omega _l)}$.
In this section, we pay our attention to the global existence of classical solutions to the problem (1.1), (1.2) and (1.4). We use Leray-Schauder's fixed point framework to show the local existence of classical solutions. For this purpose, let's consider the following linear problem
for any $T>1$ and for any given $\tilde {\bf u}\in C^{\alpha,\frac {\alpha }2}(\overline Q_T^l)\cap W_2^{1,0}(Q_T^l), \tilde n\in C^{\alpha,\frac {\alpha }2}(\overline Q_T^l)$ with $\nabla \cdot \tilde {\bf u}=0$ and $\tilde {\bf u}\cdot \nu |_{\Gamma _{T}}=0$. By classical theory of linear parabolic equations, we have ${\bf u}\in C^{2+\alpha, 1+\frac {\alpha }2}(\overline \Omega ^l\times (0, T])\cap C^{\alpha, \frac {\alpha }2}(\overline Q_T^l), \nabla \pi \in C^{\alpha, \frac {\alpha }2}(\overline \Omega ^l\times (0, T])$, and we further have $c\in C^{2+\alpha, 1+\frac {\alpha }2}(\overline \Omega ^l\times (0, T])\cap C^{\alpha, \frac {\alpha }2}(\overline Q_T^l)$, and $n\in C^{2+\alpha, 1+\frac {\alpha }2}(\overline \Omega ^l\times (0, T])\cap C^{\alpha, \frac {\alpha }2}(\overline Q_T^l)$.
On the other hands, by comparison lemma, it is easy to obtain $c\ge 0$. Next, we show that $n\ge 0$. Let $n_-=\min \{0, n\}$. Multiplying the first equation of (3.1) by $n_1$, and integration it over $\Omega _l\times (0,t)$ yields
where $J(t)=\{x\in \Omega _l; n(x,t)\le 0\}$, which implies that $\int _{\Omega _l}|n_-(x,t)|^2\,{\rm d}x\,{\rm d}y=0$, that is $n\ge 0$.
We define the mapping
From the above analysis, we see that ${\bf u}, n\in C^{2+\alpha, 1+\frac {\alpha }2}(\overline \Omega ^l\times (0, T])\cap C^{\alpha, \frac {\alpha }2}(\overline Q_T^l)$, and noticing that $C^{2+\alpha, 1+\frac {\alpha }2}(\overline Q_T^l)\cap C^{\alpha, \frac {\alpha }2}(\overline Q_T^l)\hookrightarrow C^{\alpha,\frac {\alpha }2}(\overline Q_T^l)\cap W_2^{1,0}(Q_T^l)$, then the operator $\mathcal {T}$ is completely continuous. It is easy to verify that
In fact, it is easy to see that when $\sigma =0$, it follows $c=0$, ${\bf u}=0$, then one can further derive that $n=0$.
By Leray-Schauder's fixed point theorem, to show the local existence of classical solutions, we only need to show that $\|{\bf u}, n\|_{C^{\alpha, \frac {\alpha }2}}+\|{\bf u}\|_{W_2^{1,0}(Q_T^l} \le C$ if $({\bf u}, n, \sigma )$ is a classical solution of $\mathcal {T}: ({\bf u}, n, \sigma )= ({\bf u}, n)$. For this purpose, in what follows, we pay our attention to the energy estimates.
Lemma 3.1 Let $\mathcal {T}: ({\bf u}, n, \sigma )=({\bf u}, n)$. Then we have
and
where $C$ depends only on $n_0, c_0, {\bf u}_0$, $\chi$, and $c_{air}$, and it is independent of $T$.
Proof. By comparison lemma, it is easy to obtain that
The positivity of $n$ has been established in the above analysis. By a direct integration, we see that
which implies (3.2) since $\partial _y c\ge 0$ on $\Gamma _T$. By the second equation, and using a direct calculation, we see that
Then
Noticing that $c=\sigma c_{air}$ on $\Gamma _T$, then $c_t\Big |_{\Gamma _T}=0$, $\partial _x c\Big |_{\Gamma _T}=0$, $\partial _{xx} c\Big |_{\Gamma _T}=0$, $\partial _y c\Big |_{\Gamma _T}\ge 0$. Recalling the equation $c$ satisfied, it gives
which implies that
Therefore, we have
Noticing that $\partial _y c=0$ on $\Gamma _B$, it implies that $\partial _{xy} c=0$ on $\Gamma _B$. Thus, we have
Noting that $c=\sigma c_{air}$ on $\Gamma _T$, substituting (3.6), (3.7) into (3.5) yields
for any small $\eta >0$. By a direct calculation, we see that
that is
Taking $\eta =\frac {1}{4(6+4\sqrt 2)}$ in (3.8), and combining with (3.9), we arrive at
Multiplying the first equation of (3.1) by $1+\ln n$, and integrating the resulting equation over $\Omega _l$ gives,
Multiplying the third equation of (3.1) by $u$, and integrating the resulting equation over $\Omega _l$, using the boundary conditions and Poincaré inequality, we see that for any $1< p<2$,
since ${\bf u}\Big |_{\Gamma _B}=0$, which implies that
Combining (3.10), (3.11) and (3.12), we arrive at
Noticing that $\partial _y c\ge 0$ on $\Gamma _T$, then
From Gagliardo-Nirenberg interpolation inequality, and noticing that $4-\frac 4p<2$ we infer that
Substituting (3.14) and (3.15) into (3.13) yields
which implies (3.3).
Lemma 3.2 Assume that $\mathcal {T}: ({\bf u}, n, \sigma )=({\bf u}, n)$. Then
where the constants $C$ depend only on $n_0, c_0, {\bf u}_0$, $\chi$, and $c_{air}$, and they are independent of $T$.
Proof. By Gagliardo-Nirenberg interpolation inequality, we get that
Applying $\nabla$ to the second equation of (3.1), and multiplying the resulting equation by $\nabla c$, noticing that $c_y=c_{xy}=0$ on $\Gamma _B$, and $c_x=c_{xx}=0$, $c_{yy}=nc$ on $\Gamma _T$, and using (3.19)–(3.21), we conclude that
which implies
Multiplying the second equation of (3.1) by $c_t$, then we obtain
Then (3.17) is proved. Multiplying the third equation of (3.1) by ${\bf u}_t$, and integrating it over $\Omega _l$ yields
Noticing that
then by $L^2$ theory of Stokes operator [Reference Acevedo, Amrouche, Conca and Ghosh2], we have
Combining the above two inequalities, and using (3.3) we arrive at
that is
Recalling (3.3), we see that
by mean value theorem of integrals, for any $t\in (0, T-1)$ there exists $t_0\in (t, t+1)$ such that
From (3.22), and by a direct calculation, we derive that for any $t_0$,
Combining the above two inequalities, we get that
Using this inequality, and integrating (3.22) from $t$ to $t+1$ yields
Then (3.18) is proved.
Lemma 3.3 Assume that $\mathcal {T} ({\bf u}, n, \sigma )=({\bf u}, n)$. Then for any $r>0$
where $C$ depends only on $n_0, c_0, {\bf u}_0$, $\chi$, $r$ and $c_{air}$, and it is independent of $T$.
Proof. Multiplying the first equation of (3.1) by $n^r$ for $r>0$, and integrating the resulting equation over $\Omega$ yields
By Gagliardo-Nirenberg interpolation inequality, we get that
Combining the above two inequalities, there exists a small constant $\eta >0$ such that
Lemma 3.4 Assume that $\mathcal {T} ({\bf u}, n, \sigma )=({\bf u}, n)$. Then for any $r>0$
where $C_r$ depends only on $n_0, c_0, {\bf u}_0$, $\chi$, $r$ and $c_{air}$, and it is independent of $T$.
Proof. Using the boundary value conditions, we see that
and
Applying $\nabla$ to the second equation of (3.1), and multiplying the resulting equation by $|\nabla c|^{r-2}\nabla c$ yields
and noticing that
then there exists a constant $\eta >0$ such that
Recalling (3.18) and (3.23), we infer that
Thus (3.25) is proved.
Lemma 3.5 Assume that $\mathcal {T} ({\bf u}, n, \sigma )=({\bf u}, n)$. Then for any $r>0$
and for any $\beta \in \left (\frac 12, 1\right )$,
where $A=P\Delta$, $P$ is Helmholtz projection, $C$, $\tilde C$ depend only on $n_0, c_0, {\bf u}_0$, $\chi$, $c_{air}$ and $\beta$, and they are independent of $T$.
Proof. Multiplying the first equation of (3.1) by $u^r$ for $r>1$, and integrating the resulting equation over $\Omega$ yields
which implies that
Taking $r_j=2r_{j-1}=2^j r_0$, $r_0=1$, $M_j=\max \left \{1, \|n_0\|_{L^\infty }, \sup _{t\in (0,T)}\|n\|_{L^{r_j}}\right \}$, then we have
where $\tilde C$ is independent of $j$, letting $j\to \infty$, we obtain the $L^\infty$ estimate of $n$. Recalling (3.26), using (3.18) and (3.23),for any $r\ge 3$, we see that
that is
Then, similar to above, we obtain the $L^\infty$ estimates of $\nabla c$, and we complete the proof.
By semigroup theory of Stokes operator, we see that
Similarly, for any $\beta \in \left (\frac 12, 1\right )$, we also have
which implies that
Noticing that $\frac 1{2\beta }<1$, then
□
Lemma 3.6 Let $\mathcal {T}: ({\bf u}, n, \sigma )=({\bf u}, n)$. Then for any $p>1$,
where $Q_1^l(t)=\Omega _l\times (t, t+1)$, $C_p$ is independent of $T$, it depends only on $n_0, c_0, {\bf u}_0$, $\chi$, $c_{air}$ and $p$.
Proof. Noticing that
and (3.28), using the $L^p$ theory of linear parabolic equations, it is easy to obtain that
For ${\bf u}$, we see that
then
Recalling that
Then for any $q>1$,
which implies that
Thus, (3.30) is proved.
Proof Proof of theorem 1.1
Noticing that $W_p^{2,1}(Q_T^l)\hookrightarrow C^{\beta, \frac {\beta }2}(\overline Q_T^l)$, for any $\beta < 2-\frac {4}p$, then by lemma 3.6, we derive that
Using Leray-Schauder fixed point theorem, the problem (1.1), (1.2), and (1.4) admits a classical solution $(n, c, {\bf u}, \pi )$ in $Q_T^l$. From the above lemmas, we see that all these estimates are independent of $T$, which implies that the solution $(n, c, {\bf u}, \pi )$ is a global classical solution, and we conclude that
Combining (3.33) and lemma 3.6, and using the classical theory of linear parabolic equations, we have
Using (3.34), we further have
The global existence in theorem 1.1 is proved.
The proof of uniqueness is standard, for the completeness of the paper, in what follows, we still give the proof. Suppose the contrary. Let $(n_1, c_1, {\bf u_1}, \pi _1)$, $(n_2, c_2, {\bf u_2}, \pi _2)$ be two solutions of (1.1), (1.2), and (1.4). Denote $\tilde n=n_1-n_2$, $\tilde c=c_1-c_2$, $\tilde {\bf u}={\bf u_1}-{\bf u_2}=(\tilde u, \tilde v)$, $\tilde \pi =\pi _1-\pi _2$. Then
Multiplying the first equation of (3.36) by $\tilde n$, and integrating it over $\Omega$ yields
that is
Multiplying the second equation of (3.36) by $\tilde c$, and integrating it over $\Omega$ yields
namely,
Applying $\nabla$ to the second equation of (3.36), and multiplying the resulting equation by $\nabla \tilde c$, and integrating it over $\Omega$ yields
noticing that ${\tilde c}_y={\tilde c}_{xy}=0$ on $\Gamma _B$, $\tilde c={\tilde c}_x={\tilde c}_{xx}=0$ on $\Gamma _T$, and ${\tilde c}_{yy}=\tilde n c_1+n_2\tilde c$ on $\Gamma _T$, it implies that
Then
That is
Similarly, we also have
Combining (3.37)–(3.40) yields
Using Sobolev trace embedding inequality, we see that
Substituting it into (3.42) gives
which implies that
and the uniqueness is proved.
4. Global classical solution: Zero Flux-Dirichlet(Robin)-Navier slip boundary conditions
In this section, we consider the global classical solution of the problem (1.1), (1.3) and (1.4).
Consider the following linear problem
Similar to §2, for any $T>1$ and for any given $\tilde {\bf u}\in C^{\alpha,\frac {\alpha }2}(\overline Q_T^l)\cap W_2^{1,0}(Q_T^l), \tilde n\in C^{\alpha,\frac {\alpha }2}(\overline Q_T^l)$ with $\nabla \cdot \tilde {\bf u}=0$ and $\tilde {\bf u}\cdot \nu |_{\Gamma _{T}}=0$. The above problem admits a classical solution in $Q_T^l$. By comparison lemma, we also have $0\le c\le \sigma c_{air}$. Next, we show that $n\ge 0$. Let $n_-=\min \{0, n\}$. Multiplying the first equation of (4.1) by $n_1$, and integrating it over $\Omega _l\times (0,t)$ yields
where $J(t)=\{x\in \Omega _l; n(x,t)\le 0\}$, which implies that $\int _{\Omega _l}|n_-(x,t)|^2\,{\rm d}x\,{\rm d}y=0$, that is $n\ge 0$.
We define the mapping
From the above analysis, we see that ${\bf u}, n\in C^{2+\alpha, 1+\frac {\alpha }2}(\overline \Omega ^l\times (0, T])\cap C^{\alpha, \frac {\alpha }2}(\overline Q_T^l)$, and noticing that $C^{2+\alpha, 1+\frac {\alpha }2}(\overline Q_T^l)\cap C^{\alpha, \frac {\alpha }2}(\overline Q_T^l)\hookrightarrow C^{\alpha,\frac {\alpha }2}(\overline Q_T^l)\cap W_2^{1,0}(Q_T^l)$, then the operator $\mathcal {T}$ is completely continuous. It is easy to verify that
Next, we use Leray-Schauder's fixed point theorem, to show the existence of classical solutions. For this purpose, some a prior energy estimates are necessary.
Lemma 4.1 Let $\mathcal {T}: ({\bf u}, n, \sigma )=({\bf u}, n)$. Then we have
and
Moreover, when $\tau =0$, for appropriately small $\chi >0$, we have
When $\tau =1$,
Here $C_1, C_2$ depend only on $n_0, c_0, {\bf u}_0$, $\chi$, and $c_{air}$, and they are independent of $T$.
Proof. From the above analysis, it is easy to obtain (4.2). And (4.3) is derived from a direct integration as follows
Multiplying the first equation of (4.1) by $1+\ln n$, and integrating the resulting equation over $\Omega _l$ gives,
For ${\bf } u$ it is completely similar to the proof of (3.12), we conclude that
From (3.5) and (3.7), we infer that
(i) When $\tau =0$, completely similar to the proof of (3.10), we have
\begin{align*} & \frac{{\rm d}}{{\rm d}t}\int_{\Omega_l}|\nabla\sqrt c|^2\,{\rm d}x\,{\rm d}y+\frac14 \int_{\Omega_l} c|D^2\ln c|^2\,{\rm d}x\,{\rm d}y+ \int_{\Omega_l} n|\nabla\sqrt c|^2\,{\rm d}x\,{\rm d}y\\ & \quad +\frac{\sqrt 2-1}{2\sigma^2c_{air}^2}\int_{\Gamma_T} |\partial_y c|^3\,{\rm d}S \\ & \le \frac{1}2\int_{\Gamma_T} nc_y\,{\rm d}S-\frac{1}2\int_{\Omega_l} \nabla n\nabla c\,{\rm d}x\,{\rm d}y+\frac {(3+2\sqrt 2)c_{air}}{2}\int_{\Omega_l}|\nabla{\bf u}|^2\,{\rm d}x\,{\rm d}y. \end{align*}Combining with (4.6), (4.7), and using (3.15), we arrive at\begin{align*} & \frac{{\rm d}}{{\rm d}t}\left(\int_{\Omega_l}|\nabla\sqrt c|^2\,{\rm d}x\,{\rm d}y+\frac1{2\chi}\int_{\Omega_l} n\ln n\,{\rm d}x\,{\rm d}y+(3+2\sqrt 2)c_{air}\int_{\Omega_l}|{\bf u}|^2\,{\rm d}x\,{\rm d}y\right)\\ & \quad +\,\frac {(3+2\sqrt 2)c_{air}}{2}\int_{\Omega_l}|\nabla{\bf u}|^2\,{\rm d}x\,{\rm d}y \\ & \quad+\frac1{2\chi}\int_{\Omega_l} \frac{|\nabla n|^2}{n}\,{\rm d}x\,{\rm d}y+\frac14 \int_{\Omega_l} c|D^2\ln c|^2\,{\rm d}x\,{\rm d}y+ \int_{\Omega_l} n|\nabla\sqrt c|^2\,{\rm d}x\,{\rm d}y\\ & \quad +\frac{\sqrt 2-1}{2\sigma^2c_{air}^2}\int_{\Gamma_T} |\partial_y c|^3\,{\rm d}S \\ & \le \frac{1}2\int_{\Gamma_T}nc_y\,{\rm d}S+ C_p(3+2\sqrt 2)c_{air}\|n\|_{L^p}^2 \\ & \le \frac{\sqrt 2-1}{4\sigma^2c_{air}^2}\int_{\Gamma_T} |c_y|^3\,{\rm d}S+\frac{\sigma c_{air}}{\sqrt{2\sqrt2-2}}\int_{\Gamma_T}n^{\frac32}\,{\rm d}S+ \frac1{8\chi}\int_{\Omega_l} \frac{|\nabla n|^2}{n}\,{\rm d}x\,{\rm d}y+C. \end{align*}By lemma 2.1, we have\begin{align*} & \|n\|_{L^{\frac32}(\Gamma_T)}^{\frac32}=\|\sqrt n\|_{L^3(\Gamma_T)}^3\le C_1\|\nabla\sqrt n\|_{L^2}^2\|\sqrt n\|_{L^2}+C_2\|n\|_{L^1}\\ & \quad \le C_3\int_{\Omega_l} \frac{|\nabla n|^2}{n}\,{\rm d}x\,{\rm d}y+C_4. \end{align*}Combining the above two inequalities, then when $\chi$ is appropriately small, such that $\frac {C_3\sigma c_{air}}{\sqrt {2\sqrt 2-2}}\le \frac 1{8\chi }$, we arrive at(4.9)\begin{align} & \frac{{\rm d}}{{\rm d}t}\left(\int_{\Omega_l}|\nabla\sqrt c|^2\,{\rm d}x\,{\rm d}y+\frac1{2\chi}\int_{\Omega_l} n\ln n\,{\rm d}x\,{\rm d}y+(3+2\sqrt 2)c_{air}\int_{\Omega_l}|{\bf u}|^2\,{\rm d}x\,{\rm d}y\right)\nonumber\\ & \quad +\frac {(3+2\sqrt 2)c_{air}}{2}\int_{\Omega_l}|\nabla{\bf u}|^2\,{\rm d}x\,{\rm d}y \nonumber\\ & \quad+\frac1{4\chi}\int_{\Omega_l} \frac{|\nabla n|^2}{n}\,{\rm d}x\,{\rm d}y+\frac14 \int_{\Omega_l} c|D^2\ln c|^2\,{\rm d}x\,{\rm d}y+ \int_{\Omega_l} n|\nabla\sqrt c|^2\,{\rm d}x\,{\rm d}y\nonumber\\ & \quad +\frac{\sqrt 2-1}{4\sigma^2c_{air}^2}\int_{\Gamma_T} |\partial_y c|^3\,{\rm d}S\le C. \end{align}Noticing that\[ \|n\ln n\|_{L^1}\le \|n\|_{L^2}^2+\|n\|_{L^1}\le C\|\nabla\sqrt n\|_{L^2}^2+C, \]and\[ |\nabla\sqrt c|^2\le \frac{|\nabla c|^4}{c^3}+c, \]combining (3.9) and Poincaré inequality, we see that(4.10)\begin{align} & \int_{\Omega_l}\left(|\nabla\sqrt c|^2+n\ln n+|{\bf u}|^2+c \ln c\right) {\rm d}x\,{\rm d}y \nonumber\\ & \le C\int_{\Omega_l}\left(\frac{|\nabla n|^2}{n} +c|D^2\ln c|^2+|\nabla{\bf u}|^2\right){\rm d}x\,{\rm d}y+C\int_{\Gamma_T} |c_y|^3\,{\rm d}S. \end{align}Combining (4.9), (4.11), and (4.4) is derived.(ii) When $\tau =1$, noticing that $c_y=-c+\sigma c_{air}$ on $\Gamma _T$, we take the derivative in the tangential direction and get that $c_{xy}=-c_x$ on $\Gamma _T$, then
Noticing that $c_t+uc_x=c_{xx}+c_{yy}-nc$ on $\Gamma _T$, multiplying this equation by $\frac 12\frac {(\sigma c_{air}-c)}{c}$, and integrating it on $\Gamma _T$ yields
Combining (4.8), (4.4), (4.11) and (4.12), we arrive at
which implies that
Noticing that
from Hölder's inequality, we infer that
then
By a direct calculation, we see that
that is
Taking $\eta =\frac {1}{4(6+4\sqrt 2)}$ in (4.13), and combining with (4.2), (4.14) and (4.15), we arrive at
Similar to (3.15), we have
Combining (4.6), (4.7) , (4.16), (4.17), and using Sobolev interpolation inequality, we arrive at
that is
Multiplying the second equation of (4.1) by $1+\ln c$, and integrating it over $\Omega _l$ yields
Combining (4.18) and (4.19), we arrive that
Noticing that
and
combining (4.15) and Poincaré inequality, we see that
Lemma 4.2 Assume that $\mathcal {T}: ({\bf u}, n, \sigma )=({\bf u}, n)$. Then for $\tau =0$ with small $\chi >0$, or $\tau =1$, we have
where the constants $C$ depend only on $n_0, c_0, {\bf u}_0$, $\chi$, and $c_{air}$, and they are independent of $T$.
Proof. When $\tau =0$, the proof is completely similar to lemma 3.2. We omit it. In what follows, we consider the case $\tau =1$. Applying $\nabla$ to the second equation of (3.1), and multiplying the resulting equation by $\nabla c$, noticing that $c_{xy}=-c_x$ on $\Gamma _T$, and using (3.19)–(3.21) yields
Noticing that $c_t+uc_x=c_{xx}+c_{yy}-nc$ on $\Gamma _T$, multiplying this equation by $(\sigma c_{air}-c)$, integrating it on $\Gamma _T$, and using (4.14) yields
Combining (4.24) with (4.25) gives
Noticing that $c$ is bounded uniformly, using (4.5), we arrive at
Multiplying the second equation of (4.8) by $c_t$, then we obtain
Then (4.22) is proved. The proof of (4.23) is also completely similar to (3.18), we omit it.
Then similar to the proof of lemma 3.3 by deleting the term $\chi \int _{\Gamma _T}n^{r+1}c_y\,{\rm d}S$, and lemma 3.5, we also have
Lemma 4.3 Assume that $\mathcal {T} ({\bf u}, n, \sigma )=({\bf u}, n)$. For $\tau =0$ with small $\chi >0$, or $\tau =1$, we have
and
where $C_r$, $C$ depend on $n_0, c_0, {\bf u}_0$, $\chi$, $r$ and $c_{air}$, and $C_r$ also depends on $r$, and all of them are independent of $T$.
Lemma 4.4 Assume that $\mathcal {T} ({\bf u}, n, \sigma )=({\bf u}, n)$. When $\tau =0$ with small $\chi >0$, or $\tau =1$, for any $r>0$,
where $C_r$ depends only on $n_0, c_0, {\bf u}_0$, $\chi$, $r$ and $c_{air}$, and it is independent of $T$.
Proof. When $\tau =0$, the proof is completely similar to (3.25). In what follows, we only consider the case $\tau =1$. Let
It is easy to see that $\tilde c(x,y,t)$ is periodic on $x$ with period $l$, $-\sigma c_{air}\,{\rm e}^{\frac {l^2}2}\le \tilde c\le 0$, and
From a direct calculation, we derive that
Applying $\nabla$ to the first equation of (4.30), multiplying the resulting equation by $|\nabla \tilde c|^{r-2}\nabla \tilde c$, and using (3.27), (4.22), (4.27), (4.28), it yields
then
which implies (4.29).
Then similar to the proof of lemmas 3.5 and 3.6, we conclude that
Lemma 4.5 Assume that $\mathcal {T} ({\bf u}, n, \sigma )=({\bf u}, n)$. For $\tau =0$ with small $\chi >0$, or $\tau =1$,
where $C$, $C_p$ depend only on $n_0, c_0, {\bf u}_0$, $\chi$, $c_{air}$, $\beta$ and $p$, and they are independent of $T$.
Next, we show theorem 1.2.
Proof Proof of theorem 1.2
Using lemma 4.5, completely similar to the proof theorem 1.1, we complete the proof of global existence in theorem 1.2.
Next, we show the uniqueness. Suppose the contrary. Let $(n_1, c_1, {\bf u_1}, \pi _1)$, $(n_2, c_2, {\bf u_2}, \pi _2)$ be two solutions of (1.1), (1.3), and (1.4). Denote $\tilde n=n_1-n_2$, $\tilde c=c_1-c_2$, $\tilde {\bf u}={\bf u_1}-{\bf u_2}=(\tilde u, \tilde v)$, $\tilde \pi =\pi _1-\pi _2$. Then
Multiplying the first equation of (4.34) by $\tilde n$, and integrating it over $\Omega$ yields
that is
Multiplying the second equation of (3.36) by $\tilde c$, and integrating it over $\Omega$ yields
namely,
Similarly, we also have
Combining (4.35)–(4.37) yields
which implies that
and the uniqueness is proved.
Acknowledgements
This work is supported by Guangdong Basic and Applied Basic Research Foundation (2021A1515010336), NSFC (12271186, 12171166).