Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-05-22T20:39:45.687Z Has data issue: false hasContentIssue false

The fully parabolic multi-species chemotaxis system in $\mathbb{R}^{2}$

Published online by Cambridge University Press:  19 January 2024

Ke Lin*
Affiliation:
School of Mathematics, Southwestern University of Economics and Finance, Chengdu, 610074, P.R. China
Rights & Permissions [Opens in a new window]

Abstract

This article is devoted to the analysis of the parabolic–parabolic chemotaxis system with multi-components over $\mathbb{R}^2$. The optimal small initial condition on the global existence of solutions for multi-species chemotaxis model in the fully parabolic situation had not been attained as far as the author knows. In this paper, we prove that under the sub-critical mass condition, any solutions to conflict-free system exist globally. Moreover, the global existence of solutions to system with strong self-repelling effect has been discussed even for large initial data. The proof is based on the modified free energy functional and the Moser–Trudinger inequality for system.

Type
Papers
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - ND
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives licence (http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided that no alterations are made and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use and/or adaptation of the article.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

The well-known classical parabolic–parabolic Keller–Segel model reads as [Reference Keller and Segel24]

(1.1) \begin{equation} \begin{cases} \partial _t u=\Delta u-\alpha \nabla \cdot (u\nabla v), &\quad x\in \mathbb{R}^2,\quad t\gt 0,\\[5pt] \tau \partial _t v=\Delta v-\beta v+\gamma u, &\quad x\in \mathbb{R}^2,\quad t\gt 0,\\[5pt] \end{cases} \end{equation}

where $u=u(x,t)$ and $v=v(x,t)$ denote the cell density and the concentration of the chemical substance, respectively. $\alpha$ and $\gamma$ are positive constants. The constants $\tau$ and $\beta$ are non-negative. The system (1.1) can be regarded as one of the simplest models to describe the overall behaviour of cells under the influence of chemotaxis, that is the motion of cells partially orient their movement towards higher concentration of a certain chemical substance produced by cells themselves. A striking feature of the Keller–Segel system is that the behaviour of solutions is determined by the total mass of cells which remains constant over time, see [Reference Calvez and Corrias5, Reference Herrero and Velázquez16, Reference Mizoguchi31, Reference Nagai and Ogawa34] for instance. Namely, given a non-negative and suitable smooth initial data $u_0$ , any solution with $m=\|u_0\|_{L^1(\mathbb{R}^2)}\lt 8\pi/(\alpha \gamma )$ exists globally, while blow-up solution appears if $m\gt 8\pi/(\alpha \gamma )$ . Note that the main idea to prove the global existence is based on the following free energy functional,

\begin{align*} \mathcal{F}_{KS}=\int _{\mathbb{R}^2}u\log udx+\frac{\alpha }{2\gamma }\int _{\mathbb{R}^2}\left(|\nabla v|^2+\beta v^2\right)dx-\alpha \int _{\mathbb{R}^2}uvdx, \end{align*}

which is a monotonic non-increasing function with respect to time variable. In view of this fact, Calvez and Corrias use a minimisation principle for entropy functionals and Onofri’s inequality to derive a priori estimates under the sub-critical mass $m\lt 8\pi/(\alpha \gamma )$ , where the assumptions $u_0\log\! \big(1+|x|^2\big)\in L^1(\mathbb{R}^2)$ and $u_0\log u_0\in L^1(\mathbb{R}^2)$ are necessary [Reference Calvez and Corrias5], while these extra assumptions have been removed by applying a modified free energy functional with the Moser–Trudinger inequality in unbounded domain [Reference Mizoguchi31].

For the parabolic–elliptic Keller–Segel system (i.e. taking $\tau =0$ in (1.1)2)

\begin{align*} 0=\Delta v-\beta v+\gamma u, \end{align*}

the above two-dimensional mass threshold phenomenon also exists. See [Reference Blanchet, Dolbeault and Perthame4, Reference Diaz, Nagai and Rakotoson8, Reference Nagai33] for the global well-posedness results and [Reference Biler and Nadzieja3, Reference Blanchet, Dolbeault and Perthame4] for the blow-up arguments. The main feature to prove the global existence of solutions in this simplified chemotaxis system over (1.1) is that $v$ could be expressed by the fundamental solution of the elliptic equation, then it leads to a single parabolic problem for $u$ . For example, if $\beta =0$ , an explicit expression for $v$ takes form like $v=\gamma K\ast u$ , so (1.1) $_1$ becomes

\begin{align*} \partial _t u=\Delta u-\alpha \gamma \nabla \cdot (u\nabla K\ast u), \quad x\in \mathbb{R}^2,\quad t\gt 0, \end{align*}

where $K=-(1/2\pi )\log |\cdot |$ . A direct application of the logarithmic Hardy–Littlewood–Sobolev inequality (see [Reference Beckner2]) on the corresponding free energy yields the global existence of solutions if $m\lt 8\pi/(\alpha \gamma )$ [Reference Blanchet, Dolbeault and Perthame4].

Compared with the one-population chemotaxis system (1.1), an interesting and complex question is to derive sharp conditions to recognise global existence and blow-up of solutions for the following multi-species chemotaxis model in $\mathbb{R}^2$ ,

(1.2) \begin{equation} \begin{cases} \begin{split} \partial _t u_i=&\Delta u_i-\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \cdot (u_i\nabla v_j), &i\in \mathcal{I}=\{1,\cdots,n\}, \\ \tau _j\partial _t v_j=&\Delta v_j-\beta _jv_j+\sum \limits ^n_{i=1}\gamma _{i,j}u_i, &j\in \mathcal{J}=\{1,\cdots,m\}, \end{split} \end{cases} \end{equation}

where $\tau _j\geq 0$ , $j\in \mathcal{J}$ . This model was first proposed by Wolansky in [Reference Wolansky42] to describe the chemotactic movement of $n$ populations with respect to $m$ chemical substances. Here, $u_i=u_i(x,t)$ denotes the density of $i$ -th population, and $v_j=v_j(x,t)$ represents the concentration of $j$ -th chemical signal. The total number of species $|\mathcal I|=n\geq 1$ is assumed to be finite. $\boldsymbol{\alpha }=(\alpha _{i,j})_{n\times m}$ and $\boldsymbol{\gamma }=(\gamma _{i,j})_{n\times m}$ define a pair of $n\times m$ matrices for the sensitivity parameter and the production/consumption rate, respectively. $\beta _j\in \mathbb{R}$ , $j\in \mathcal{J}$ , presents the growth/degradation rate for chemical substance. We introduce $\boldsymbol{\beta }=(\beta _{j,l})_{m\times m}$ with $\beta _{j,l}=\beta _j\delta _{j,l}$ as a $m\times m$ diagonal matrix for convenience, where $\delta _{j,l}$ satisfies

\begin{align*} \delta _{j,l}=\begin{cases} 1,&\text{if}\quad j=l,\\[5pt] 0,&\text{if}\quad j\neq l. \end{cases} \end{align*}

It is very important to understand the multi-species chemotaxis in biology, and this phenomenon has been observed in numerous experiments. We take the following two examples. First one is that a system with two different species, reacting on one common chemical, appears in the cell sorting process during the early aggregation state of mound formation [Reference Weijer40]. And a two-species chemotaxis system with two chemicals has been proposed in [Reference Knútsdóttir, Pálsson and Edelstein-Keshet25] to imitate the breast cancer metastatic process. The readers could see [Reference Horstmann20, Reference Horstmann and Lucia21] for other biological motivations.

Just recently, some authors have started to look more closely at the parabolic–elliptic case of (1.2) (i.e. $\tau _j=0$ ) for $n$ -populations interacting via a self-produced chemical agent. Consider (1.2) with $|\mathcal{I}|=|\mathcal{J}|=n$ is subject to symmetric sensitivity coefficients matrix $\boldsymbol{\alpha }=(\alpha _{i,j})_{n\times n}$ with non-negative entries, zero matrix $\boldsymbol{\beta }$ and unit matrix $\boldsymbol{\gamma }$ , that is,

(1.3) \begin{equation} \begin{cases} \partial _t u_i=\Delta u_i-\sum \limits ^n_{j=1}\alpha _{i,j}\nabla \cdot (u_i\nabla v_j), \\[9pt] -\Delta v_j=u_j, \quad i,j \in \mathcal{I}=\{1,\cdots,n\}. \\[5pt] \end{cases} \end{equation}

Karmakar and Wolansky [Reference Karmakar and Wolansky22] had derived the global well-posedness of weak solutions with respect to time in the sub-critical regime

(1.4) \begin{equation} \begin{split} \Lambda _{\mathcal{K}}(\boldsymbol{{m}})\,:\!=\,8\pi \sum \limits _{i\in \mathcal{K}}m_i-\sum \limits _{i,k\in \mathcal{K}}\alpha _{i,k}m_im_k\gt 0, \quad \forall \,\,\emptyset \neq \mathcal{K} \subset \mathcal{I}, \end{split} \end{equation}

where $m_i=\|u_0\|_{L^1(\mathbb{R}^2)}$ . Furthermore, the borderline case of critical mass $\Lambda _{\mathcal{I}}(\boldsymbol{{m}})=0$ , and $\Lambda _{\mathcal{K}}(\boldsymbol{{m}})\gt 0$ , $\forall \,\emptyset \neq \mathcal{K} \subsetneq \mathcal{I}$ , has been considered in [Reference Karmakar and Wolansky23]. It shows that a free energy solution exists globally in time. According to analogous results about (1.1) mentioned above, it is expected that if the condition $\Lambda _{\mathcal{K}}(\boldsymbol{{m}})\geq 0$ for some $\emptyset \neq \mathcal{K} \subset \mathcal{I}$ is violated, a finite-time blow-up solution appears. Using the second-moment techniques in [Reference He and Tadmor15], some solutions of (1.3) blow-up in finite time provided $\Lambda _{\mathcal{I}}(\boldsymbol{{m}})\lt 0$ . While the basic strategy to prove global existence is the logarithmic Hardy–Littlewood–Sobolev inequality for system, see [Reference Shafrir and Wolansky38] for details. In particular, in the case of parabolic–elliptic system (1.2) with $|\mathcal{I}|=2$ , $|\mathcal{J}|=1$ , Espejo et al. [Reference Conca, Espejo and Vilches7, Reference Espejo, Vilches and Conca12] discovered a threshold curve to help us to determine whether the solutions of system are blow-up or global in time. See related works for parabolic–elliptic system (1.2) with $|\mathcal{I}|=|\mathcal{J}|=2$ in [Reference Hong, Wang and Wang18, Reference Hong, Wang, Yu and Zhang19]. Moreover, [Reference Espejo, Stevens and Suzuki9– Reference Espejo, Stevens and Velázquez11, Reference Espejo, Vilches and Conca13] could be refereed to characterise the simultaneous or non-simultaneous blow-up results in two-species model.

However, it should be noted that fewer papers have been considered on Cauchy problem of the fully parabolic multi-species (i.e. $\tau _j\gt 0$ in (1.2)) than the parabolic–elliptic case. For a two-dimensional bounded domain, the author and coauthors have researched the initial boundary problems of (1.2), and we have tried to find optimal conditions on the global existence or blow-up in [Reference Lin27Reference Lin and Zhong30]. In this article, under a conflict-free situation given by Definition 1 $(ii)$ , a sufficient (or possibly optimal) condition on the global solvability of the Cauchy problem for parabolic–parabolic system (1.2) with arbitrary $|\mathcal{I}|=n\geq 1$ and $|\mathcal{J}|=m\geq 1$ has been obtained. For simplicity, we assume $\tau _j=1$ for all $j\in \mathcal{J}$ in (1.2) and consider

(1.5) \begin{equation} \begin{cases} \partial _t u_i=\Delta u_i-\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \cdot (u_i\nabla v_j), &i\in \mathcal{I}, \\[5pt] \partial _t v_j=\Delta v_j-\beta _jv_j+\sum \limits ^n_{i=1}\gamma _{i,j}u_i, &j\in \mathcal{J}, \\[5pt] u_{i0}(x)=u_i(x,0), v_{j0}(x)=v_j(x,0),&i\in \mathcal{I},\quad j\in \mathcal{J}, \end{cases} \end{equation}

with initial data $\boldsymbol{{u}}_0=(u_{10},\cdots,u_{n0})$ and $ \boldsymbol{{v}}_0=(v_{10},\cdots,v_{m0})$ satisfying

(1.6) \begin{equation} \begin{cases} &u_{i0}(x)\in C^0(\mathbb{R}^2)\cap L^1(\mathbb{R}^2)\cap L^1\!\left(\mathbb{R}^2,\log \!\left(1+|x|^2\right)dx\right)\cap L^{\infty }(\mathbb{R}^2),\\[5pt] &u_{i0}\geq 0\quad \text{and}\quad u_{i0}\not \equiv 0,\quad i\in \mathcal{I},\\[5pt] &v_{j0}(x)\in W^{1,p}(\mathbb{R}^2)\cap W^{1,1}(\mathbb{R}^2)\quad \text{with some}\,\,p\gt 2,\quad j\in \mathcal{J}. \end{cases} \end{equation}

Before stating our main results, let us go over some notations and definitions in [Reference Wolansky42].

\begin{align*} \lambda _{i,k}\,:\!=\,\sum ^{m}_{j=1}\alpha _{i,j}\gamma _{k,j}=\boldsymbol{\alpha }^T_i\cdot \boldsymbol{\gamma }_k,\quad i,k\in \mathcal{I},\end{align*}

is the number to quantify the tendency of a population $i$ towards a population $k$ by taking an accounting of the action of all the agents, where $\boldsymbol{\alpha }_i=(\alpha _{i,1},\cdots,\alpha _{i,m})^T$ and $\boldsymbol{\gamma }_i=(\gamma _{i,1},\cdots,\gamma _{i,m})^T$ . The condition $\lambda _{i,k}\gt 0$ means that a population $i$ is attracted by a population $k$ ; otherwise, the population $i$ is repelled from the population $k$ if $\lambda _{i,k}\lt 0$ . Especially, a population is self-attracting (resp. self-repelling) if $\lambda _{i,i}\gt 0$ (resp. $\lambda _{i,i}\lt 0$ ). A pair $(i,k)$ of populations $i,k\in I$ is said to be in a conflict if $\lambda _{i,k}\times \lambda _{k,i}\lt 0$ . In general, $\boldsymbol{\lambda }=(\lambda _{i,k})_{i,k\in \mathcal{I}}$ is not symmetric. We assume that there exist nonzero constants $a_1,\cdots,a_n$ satisfying

(1.7) \begin{equation} a_i\lambda _{i,k}=a_{k}\lambda _{k,i},\quad i,k\in \mathcal{I}, \end{equation}

then $D_{\boldsymbol{{a}}}\boldsymbol{\lambda }$ is symmetric, where $D_{\boldsymbol{{a}}}=\text{Diag}\{a_1,\cdots,a_n\}$ . If $\boldsymbol{\lambda }$ is non-singular, there exists a $m\times m$ symmetric matrix $\boldsymbol{{B}}$ which transforms $\boldsymbol{\gamma }_i$ into $a_i\boldsymbol{\alpha }_i$ for all $i\in \mathcal{I}$ , i.e.

(1.8) \begin{equation} \boldsymbol{{B}}\boldsymbol{\gamma }_i=a_i\boldsymbol{\alpha }_i,\quad i\in \mathcal{I}. \end{equation}

In fact, denote

\begin{align*} \boldsymbol{\alpha }=(\alpha _{i,j})_{n\times m}=\left [ \begin{array}{c} \boldsymbol{\alpha }^T_1\\[5pt] \vdots \\[5pt] \boldsymbol{\alpha }^T_n \end{array} \right ],\,\, \boldsymbol{\gamma }=(\gamma _{i,j})_{n\times m}=\left [ \begin{array}{c} \boldsymbol{\gamma }^T_1\\[5pt] \vdots \\[5pt] \boldsymbol{\gamma }^T_n \end{array} \right ]. \end{align*}

First, we observe that the ranks $R(\boldsymbol{\alpha })=R(\boldsymbol{\gamma })=n$ due to $\boldsymbol{\lambda }=\boldsymbol{\alpha }\boldsymbol{\gamma }^T$ is non-singular. Then, there exists a solution $X=(x_{i,j})_{m\times m}$ to a linear system of equations $ \boldsymbol{\gamma }X=D_{\boldsymbol{{a}}}\boldsymbol{\alpha }$ since both the ranks of its coefficient matrix and augmented matrix equal to $n$ . Finally, the choice $\boldsymbol{{B}}=X^T$ fulfils (1.8). Moreover, using the symmetry of $D_{\boldsymbol{{a}}}\boldsymbol{\lambda }$ , one is able to show that $ \boldsymbol{\gamma }\boldsymbol{{B}}^T\boldsymbol{\gamma }^T=\boldsymbol{\gamma }\boldsymbol{{B}}\boldsymbol{\gamma }^T$ . This implies that $\boldsymbol{{B}}$ can be symmetric.

Now we give the following definitions throughout this paper.

Definition 1. $(i)$ A pair $(i,k)$ of populations $i,k\in \mathcal{I}$ is said to be in a conflict if $\lambda _{i,k}\times \lambda _{k,i}\lt 0$ . All populations are mutually attractive if $\lambda _{i,k}\gt 0$ , $\forall \,i,k\in \mathcal{I}$ .

$(ii)$ System (1.5) is called a conflict-free system if $\lambda _{i,k}\times \lambda _{k,i}\gt 0$ , $\forall \,i,k\in \mathcal{I}$ , and if there exist positive constants $a_1,\cdots,a_n$ such that (1.7) is valid.

The main result of this article is stated as follows.

Theorem 1.1. Let $\boldsymbol{\gamma }=(\gamma _{i,j})_{n\times m}$ , $\boldsymbol{\lambda }=(\lambda _{i,k})_{n\times n}$ , $\boldsymbol{\alpha }=(\alpha _{i,j})_{n\times m}$ with full column rank $R(\boldsymbol{\alpha })=m$ , and $\boldsymbol{\beta }=(\beta _{j,l})_{m\times m}$ with $\beta _{j,l}=\beta _j\delta _{j,l}$ , $\beta _{j}\in \mathbb{R}$ , $j,l\in \mathcal{J}$ . Assume that (1.5) is a conflict-free system with initial data $(\boldsymbol{{u}}_0,\boldsymbol{{v}}_0)$ satisfying (1.6). Suppose that there exist positive constants $a_1,\cdots,a_n$ and a positive definite matrix $\boldsymbol{{R}}=(r_{i,k})_{n\times n}$ with $r_{i,k}\geq 0$ , $i,k\in \mathcal{I}$ , such that

(1.9) \begin{equation} \boldsymbol{\alpha }^T\boldsymbol{{R}}^{-1}\boldsymbol{\alpha }\boldsymbol{\gamma }_i=a_i\boldsymbol{\alpha }_i,\quad \forall \,\,i\in \mathcal{I}. \end{equation}

Then for any initial data satisfying

(1.10) \begin{equation} \Lambda ^{\boldsymbol{{a}}}_{\mathcal{K}}(\boldsymbol{{m}})\,=\!:\, 8\pi \sum \limits _{i\in \mathcal{K}}a_im_i-\sum \limits _{i,k\in \mathcal{K}}a_ia_kr_{i,k}m_im_k\gt 0, \quad \forall \,\,\emptyset \neq \mathcal{K} \subset \mathcal{I}, \end{equation}

the corresponding initial boundary value problem (1.5) possesses a unique smooth global solution.

We would like to give an explanation for assumptions in Theorem 1.1. First, since $\boldsymbol{\alpha }=(\alpha _{i,j})_{n\times m}$ is required to be full column rank, it ensures that $\boldsymbol{{B}}=(b_{j,l})_{m\times m}=\boldsymbol{\alpha }^T\boldsymbol{{R}}^{-1}\boldsymbol{\alpha }$ is a positive definite matrix if $\boldsymbol{{R}}$ is chosen to be positive definition. Condition (1.9) can probably be viewed as one of necessary condition for the existence of energy functional for the conflict-free system (1.5) (see [Reference Horstmann20, Reference Lin and Zhong30, Reference Wolansky42]). In order to handle the whole domain case better in this paper, we use a modified free energy functional $\mathcal{F}$ as

\begin{align*} \begin{split} \mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]=&\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx\\[5pt] &-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx. \end{split} \end{align*}

Second, condition (1.10) can be regarded as an optimality condition to guarantee the global existence of solution to (1.5). This is because if $D_{\boldsymbol{{a}}}\boldsymbol{\lambda }$ is a positive definite matrix with $\lambda _{i,k}\geq 0$ , $\forall \,i,k\in \mathcal{I}$ , then (1.10) is actually equivalent to the following sub-critical mass condition obtained in [Reference Lin and Zhong30, Reference Wolansky42]

\begin{align*} 8\pi \sum \limits _{i\in \mathcal{K}}a_im_i-\sum \limits _{i,k\in \mathcal{K}}a_i\lambda _{i,k}m_im_k\gt 0, \quad \forall \,\,\emptyset \neq \mathcal{K} \subset \mathcal{I}. \end{align*}

Indeed, let $\boldsymbol{{R}}=(r_{i,k})_{n\times n}=D^{-1}_{\boldsymbol{{a}}}\boldsymbol{\lambda }^T$ . Then in terms of $\lambda _{i,k}=\boldsymbol{\alpha }^T_i\cdot \boldsymbol{\gamma }_k$ and $R(\boldsymbol{\alpha })=m$ , $\boldsymbol{{B}}=\boldsymbol{\alpha }^{T}{\boldsymbol{{R}}}^{-1}\boldsymbol{\alpha }$ is a positive definite matrix which satisfies $\boldsymbol{{B}}\boldsymbol{\gamma }_i=a_i\boldsymbol{\alpha }_i$ , $\forall \,\,i\in \mathcal{I}$ . Moreover, we obtain

\begin{align*} 8\pi &\sum \limits _{i\in \mathcal{K}}a_im_i-\sum \limits _{i,k\in \mathcal{K}}a_i\lambda _{i,k}m_im_k= 8\pi \sum \limits _{i\in \mathcal{K}}a_im_i-\sum \limits _{i,k\in \mathcal{K}}a_ia_kr_{i,k}m_im_k\gt 0,\quad \forall \,\,\emptyset \neq \mathcal{K} \subset \mathcal{I}, \end{align*}

from $r_{i,k}={\lambda _{k,i}}/a_i\geq 0$ , $i,k\in \mathcal{I}$ .

Theorem 1.1 gives a sharp criterion on the global existence of the general chemotaxis system (1.5). Hence, a large amount of known global existence results are particular cases in our paper. We give several typical examples here. It is obvious that the sub-critical mass condition (1.10) recovers the threshold condition, i.e. $m\lt 8\pi/(\alpha \gamma )$ , for global regularity of the Keller–Segel model (1.1). When $|\mathcal{I}|=2$ , $|\mathcal{J}|=1$ , consider a chemotaxis system involving two species that interact via one-single chemical signal [Reference Weijer40]

(1.11) \begin{equation} \left\{\begin{array}{l} \partial _t u_i=\Delta u_i-\alpha _{i,1}\nabla \cdot (u_i\nabla v_1),\quad i\in \mathcal{I}=\{1,2\}, \\[5pt] \partial _tv_{1}=\Delta v_1-v_1+\gamma _{1,1}u_1+\gamma _{2,1}u_2, \end{array}\right. \end{equation}

with $\alpha _{i,1}\gt 0$ , $\gamma _{i,1}=1$ , $i=1,2$ . Note that $\lambda _{i,k}=\alpha _{i,1}\gt 0$ , $i,k=1,2$ . Then, (1.9) can be satisfied if one takes $a_i=1/\alpha _{i,1}$ and chooses a positive definite matrix

\begin{align*} \boldsymbol{{R}}=\left [ \begin{array}{c@{\quad}c} \alpha ^2_{1,1}(1+\epsilon )&\alpha _{1,1}\alpha _{2,1}(1-\epsilon )\\[5pt] \alpha _{1,1}\alpha _{2,1}(1-\epsilon )&\alpha ^2_{2,1}(1+\epsilon ) \end{array} \right ] \end{align*}

for some small $\epsilon \in (0,1)$ . Then, (1.10) reads as

(1.12) \begin{equation} \begin{cases} 8\pi \gt \alpha _{1,1}(1+\epsilon )m_1,\\[5pt] 8\pi \gt \alpha _{2,1}(1+\epsilon )m_2,\\[5pt] 8\pi \!\left (\dfrac{m_1}{\alpha _{1,1}}+\dfrac{m_2}{\alpha _{2,1}}\right )\gt (m_1+m_2)^2+\epsilon (m_1-m_2)^2. \end{cases} \end{equation}

However, (1.12) can be simplified as $m_1\lt 8\pi/\alpha _{1,1}$ , $m_2\lt 8\pi/\alpha _{2,1}$ and $(m_1+m_2)^2\lt 8\pi \!\left (m_1/\alpha _{1,1}+m_2/\alpha _{2,1}\right )$ by letting $\epsilon \rightarrow 0$ , which coincides with global existence condition for (1.11) in a bounded domain ([Reference Lin27, Theorem 1.1]).

Now suppose that $|\mathcal{I}|=|\mathcal{J}|=n$ , $\boldsymbol{\beta }=\boldsymbol{0}$ , $\boldsymbol{\alpha }=(\alpha _{i,j})_{n\times n}$ with $\alpha _{i,j}\geq 0$ is positive definite, and $\boldsymbol{\gamma }$ is an unit matrix. Then, (1.5) becomes

(1.13) \begin{equation} \begin{cases} \partial _t u_i=\Delta u_i-\sum \limits ^n_{j=1}\alpha _{i,j}\nabla \cdot (u_i\nabla v_j), \\[8pt] \partial _tv_{j}=\Delta v_j+u_j, \quad i,j \in \mathcal{I}=\{1,\cdots,n\}. \\[5pt] \end{cases} \end{equation}

Taking $a_i=1$ and $\lambda _{i,k}=\alpha _{i,k}$ , $i,k\in \mathcal{I}$ , one can find that Cauchy problem (1.13) has a global solution if (1.4) is valid. This global result is similar to that of the parabolic–elliptic system (1.3) [Reference He and Tadmor15, Reference Karmakar and Wolansky22].

The idea to show the global existence is to derive an a prior estimate for modified total entropy

\begin{align*} \mathcal{S}[\boldsymbol{{u}}]=\sum \limits ^n_{i=1}\|(u_i+1)\log\! (u_i+1)\|_{L^1(\mathbb{R}^2)}. \end{align*}

For this purpose, we need to give a lower bound for $\mathcal{F}$ . In this situation, the last term consisting of $u_i$ and $v_j$ in $\mathcal{F}$ could be controlled by $\mathcal{S}$ and the last second term under (1.10). For the case of single variable, a common approach to achieve this goal is to use the well-known Moser–Trudinger inequality [Reference Moser32, Reference Trudinger39]

(1.14) \begin{equation} \frac{1}{2}\int _{\Omega }|\nabla \rho |^2dx-8\pi \log \!\left (\int _{\Omega }e^{\rho }dx\right )\geq -C,\quad \forall \,\,\rho \in H^1_0(\Omega ), \end{equation}

where $\Omega \subset \mathbb{R}^2$ is a bounded domain. For initial Neumann boundary value problem (1.1), Nagai et al. [Reference Nagai, Senba and Yoshida36] had used a version of (1.14) in the Sobolev space $W^{1,2}$ (see also [Reference Chang and Yang6]) to obtain the global existence of solution if $m\lt 8\pi/(\alpha \gamma )$ . Later, Mizoguchi [Reference Mizoguchi31] applies (1.14) to get a similar global result for Cauchy problem (1.1) in $\mathbb{R}^2$ . Hence, we expect the Moser–Trudinger inequality for vector is valid for our problem. Shafrir and Wolansky [Reference Shafrir and Wolansky38] had proved the following Moser–Trudinger inequality for system. For $\forall \,\,\rho _i\in H^1(\mathbb{S}^2)$ satisfying $\int _{\mathbb{S}^2}\rho _i=0,\,\,i\in \mathcal{I}$ , there exists a constant $C\gt 0$ such that

\begin{align*} \Phi _{\mathbb{S}^2}(\boldsymbol{\rho })= \frac{1}{2}\sum _{i,k\in \mathcal{I}}s_{i,k}\int _{\mathbb{S}^2}\nabla \rho _i\cdot \nabla \rho _k -\sum _{i\in \mathcal{I}}M_i\log \!\left (\frac{1}{4\pi }\int _{\mathbb{S}^2}\exp \!\left(\sum \limits _{k\in \mathcal{I}}s_{i,k}\rho _k\right)\right )\geq -C\end{align*}

if and only if

\begin{align*} \left \{ \begin{array}{l} \Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}}(\boldsymbol{{M}})\geq 0,\quad \forall \,\emptyset \neq \mathcal{K}\subset \mathcal{I},\\[9pt] \text{if}\,\,\Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}}(\boldsymbol{{M}})=0\,\,\text{for some}\,\,\mathcal{K},\,\,\text{then}\,\,s_{i,i}+\Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}\setminus \{i\}}(\boldsymbol{{M}})\gt 0,\,\,\forall \,{i\in \mathcal{K}}, \end{array} \right . \end{align*}

where $\mathbb{S}^2\subset \mathbb{R}^3$ is the unit sphere, $\mathcal{I}=\{1,2,\cdots,n\}$ , $\boldsymbol{{M}}\,:\!=\,\{M_1,\cdots,M_n\}\in (\mathbb{R}_{+})^n$ , $\boldsymbol{{S}}\,:\!=\,(s_{i,k})_{n\times n}$ is a positive definite matrix with $s_{i,k}\geq 0$ , $i,k\in \mathcal{I}$ , and $\Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}}$ is given by

\begin{align*} \Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}}(\boldsymbol{{M}})=8\pi \sum \limits _{i\in \mathcal{K}}M_i-\sum \limits _{i,k\in \mathcal{K}}s_{i,k}M_iM_k,\quad \forall \,\,\emptyset \neq \mathcal{K}\subset \mathcal{I}. \end{align*}

We will firstly transform the above Moser–Trudinger inequality for system to $\mathbb{R}^2$ via the stereographic projection and next use it to show that $\mathcal{S}$ is bounded under the assumption (1.10). Then, one invokes the Moser iterative to obtain the global existence of solutions to (1.5). Finally, we should point out that such idea has been used to establish the global well-posedness of solutions to initial Neumann boundary value problem for multi-species and chemicals in a two-dimensional bounded domain [Reference Lin and Zhong30].

Our second object is to show certain conflict-free parabolic system admits a global solution for any initial data. More precisely,

Theorem 1.2. Let $\boldsymbol{\alpha }=(\alpha _{i,j})_{n\times m}$ , $\boldsymbol{\gamma }=(\gamma _{i,j})_{n\times m}$ , $\boldsymbol{\lambda }=(\lambda _{i,k})_{n\times n}$ and $\boldsymbol{\beta }=(\beta _{j,l})_{m\times m}$ with $\beta _{j,l}=\beta _j\delta _{j,l}$ , $\beta _{j}\in \mathbb{R}$ , $j,l\in \mathcal{J}$ . Assume that (1.5) is a conflict-free system with initial data $(\boldsymbol{{u}}_0,\boldsymbol{{v}}_0)$ satisfying (1.6). Suppose that there exist positive constants $a_1,\cdots,a_n$ and a positive definite matrix $\boldsymbol{{B}}=(b_{j,l})_{m\times m}$ such that

\begin{align*} \boldsymbol{{B}}\boldsymbol{\gamma }_i=-a_i\boldsymbol{\alpha }_i,\quad \forall \,\,i\in \mathcal{I}, \end{align*}

then Cauchy problem (1.5) possesses a unique smooth global solution.

Remark 1. In addition, if $\beta _j\gt 0$ for all $j\in \mathcal{J}$ in Theorem 1.2, then the solution to (1.5) is uniformly bounded with respect to time.

Remark 2. As mentioned above, the existence of $\boldsymbol{{B}}$ in Theorem 1.2 can be ensured if $D_{\boldsymbol{{a}}}\boldsymbol{\lambda }$ is negative definite and $R(\boldsymbol{\alpha })=m$ . Hence, one may assert that there exists a unique global smooth solution, provided that the self-repelling effects are strong enough in the sense that $\lambda _{ii}\lt 0$ , $i\in \mathcal{I}$ , is negative sufficiently large. Following are two prototypical examples. Consider (1.1) with $\alpha \lt 0,\gamma \gt 0$ , the local solution can in fact be extended for all times. As an application of Theorem 1.2 on two-species system (1.11) with $\alpha _{i,1}\lt 0$ , $\gamma _{i,1}\gt 0$ , $i=1,2$ , one would derive the global stability by choosing $a_1=-1/(\alpha _{1,1}\gamma _{2,1})$ , $a_2=-1/(\alpha _{2,1}\gamma _{1,1})$ and $\boldsymbol{{B}}=(1/(\gamma _{1,1}\gamma _{2,1}))_{1\times 1}$ .

Compared with the proof of Theorem 1.1, the main difference to prove Theorem 1.2 is to derive a prior estimates for the modified total entropy $\mathcal{S}$ through the following modified free energy functional

\begin{align*} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]=&\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx. \end{align*}

This paper is organised as follows. In Section 2, we would like to establish the local existence of smooth solutions and present some basic inequalities. Theorem 1.1 will be proved in Section 3. The existence of modified free energy functional $\mathcal{F}$ will be shown firstly, then the stepwise bounds on the total entropy, $L^2$ and $L^{\infty }$ norms under the condition (1.10) will end the proof of this theorem. In Section 4, we will prove Theorem 1.2 by making use of $\mathcal{G}$ . In appendix, the proof of Lemma 2.1 will be contained.

We introduce some notations which will be used later. Let $|\mathcal I|=n\geq 1$ denote the total number of species, and $|\mathcal{J}|=m\geq 1$ represent the total number of chemical substances. $|\Omega |$ is the Lebesgue measure of $\Omega \subset \mathbb{R}^2$ . For $\alpha _{i,j}$ , $\gamma _{i,j}$ , $b_{i,j}$ , $r_{i,k}$ , $t_{i,j}$ , $a_i$ , $\beta _{j}\in \mathbb{R}$ , $i,k\in \mathcal{I}$ , $j\in \mathcal{J}$ , we define

\begin{align*} \alpha ^* & =\max \limits _{i\in \mathcal{I},j\in \mathcal{J}}\{|\alpha _{i,j}|\}, \,\,\gamma ^*=\max \limits _{i\in \mathcal{I},j\in \mathcal{J}}\{|\gamma _{i,j}|\},\,\,b^*=\max \limits _{i\in \mathcal{I},j\in \mathcal{J}}\{|b_{i,j}|\},\,\,r^*=\max \limits _{i,k\in \mathcal{I}}\{|r_{i,k}|\},\\ t^* & =\max \limits _{i\in \mathcal{I},j\in \mathcal{J}}\{|t_{i,j}|\},\,\,a^*=\max \limits _{i\in \mathcal{I}}\{|a_i|\}, \,\,\beta ^*=\max \limits _{j\in \mathcal{J}}\{|\beta _j|\},\,\,\beta _*=\min \limits _{j\in \mathcal{J}}\{\beta _j\}\,\,\text{if}\,\,\beta _j\gt 0. \end{align*}

2. Preliminaries

In this section, we list some lemmas which will be frequently used throughout this paper. Under some certain assumption on the initial data, we assert that Cauchy problem (1.5) admits a local classical solution. A number of fundamental properties, such as uniqueness, positivity, and regularity, have also been obtained in the following lemma.

Lemma 2.1. Suppose that $\boldsymbol{{u}}_0\in [C^0(\mathbb{R}^2)\cap L^1(\mathbb{R}^2)\cap L^{\infty }(\mathbb{R}^2)]^n$ and $ \boldsymbol{{v}}_0\in [W^{1,p}(\mathbb{R}^2)\cap W^{1,1}(\mathbb{R}^2)]^m$ for some $p\gt 2$ . Then, there exists a positive constant $T_{\max }\in (0,\infty ]$ such that the Cauchy problem (1.5) has a unique solution $(\boldsymbol{{u}},\boldsymbol{{v}})$ satisfying

(2.1) \begin{equation} \begin{cases} \boldsymbol{{u}}\in \left [C^0([0,T_{\max }];\ L^1(\mathbb{R}^2))\cap C^{2,1}(\mathbb{R}^2\times (0,T_{\max }))\right ]^n,\\[5pt]\boldsymbol{{v}}\in \left [C^0([0,T_{\max }];\ W^{1,p}(\mathbb{R}^2)\cap W^{1,1}(\mathbb{R}^2))\cap C^{2,1}(\mathbb{R}^2\times (0,T_{\max }))\right ]^m. \end{cases} \end{equation}

Moreover, it holds that

  1. (i) $(\boldsymbol{{u}},\boldsymbol{{v}})$ solves (1.5) classically in $\mathbb{R}^2\times (0,T_{\max })$ ;

  2. (ii) If $T_{\max }\lt \infty$ , then

    \begin{align*} \limsup _{t\rightarrow T_{\max }}\left (\sum \limits ^n_{i=1}\|u_i(\cdot,t)\|_{L^{\infty }(\mathbb{R}^2)}\right )=\infty ; \end{align*}
  3. (iii) $u_i\gt 0$ in $\mathbb{R}^2\times (0,T_{\max })$ , $i\in \mathcal{I}$ ;

  4. (iv) For $t\in (0,T_{\max })$ , $\|u_i(\cdot,t)\|_{L^1(\mathbb{R}^2)}=\|u_{i0}\|_{L^1(\mathbb{R}^2)}=m_i$ , $i\in \mathcal{I}$ ;

  5. (v) For $q\geq 1$ , $T\in (0,T_{\max })$ , then there exists a constant $A_q=A_q(q,\beta ^*,\gamma ^*,\|\boldsymbol{{u}}_0\|_{L^1(\mathbb{R}^2)},\|\boldsymbol{{v}}_0\|_{L^q(\mathbb{R}^2)},T)\gt 0$ such that

    \begin{align*} \sum \limits ^m_{j=1}\|v_j(\cdot,t)\|_{L^q(\mathbb{R}^2)}\leq A_q,\quad t\in (0,T). \end{align*}
    Moreover, $A_q$ is independent of $T$ if $\beta _j\gt 0$ for all $j\in \mathcal{J}$ .

Proof. For the case $|\mathcal{I}|=|\mathcal{J}|=1$ to system (1.5) in $\mathbb{R}^n$ ( $n\geq 3$ ), Winkler has proved these properties in [Reference Winkler41, Proposition 1.1]. One can apply similar arguments to obtain the desired results. Please see the proof in appendix.

The following inequalities are very important to derive a prior estimates for solutions.

Lemma 2.2. Let $\eta \in (0,1)$ . Then for any non-negative function $f\in L^1(\mathbb{R}^2)\cap H^1(\mathbb{R}^2)$ , one has

(2.2) \begin{align}\int _{\mathbb{R}^2}(f+1)\log\! (f+1)dx&\leq \eta \!\left (\int _{\mathbb{R}^2}fdx\right )\left (\int _{\mathbb{R}^2}\frac{|\nabla f|^2}{f+1}dx\right )+c\int _{\mathbb{R}^2}fdx, \end{align}
(2.3) \begin{align} \int _{\mathbb{R}^2}f^2dx&\leq \frac{1+\eta }{4\pi }\left (\int _{\mathbb{R}^2}fdx\right )\left (\int _{\mathbb{R}^2}\frac{|\nabla f|^2}{f+1}dx\right )+\frac{2}{\eta }\int _{\mathbb{R}^2}fdx, \end{align}
(2.4) \begin{align} \int _{\mathbb{R}^2}f^3dx&\leq \eta \!\left (\int _{\mathbb{R}^2}(f+1)\log\! (f+1)dx\right )\left (\int _{\mathbb{R}^2}|\nabla f|^2dx\right )+c\int _{\mathbb{R}^2}fdx, \end{align}

where $c=c(\eta )\rightarrow \infty$ as $\eta \rightarrow 0$ .

Proof. Inequality (2.2) has been shown in [Reference Yamada43, Lemma 2.3], and (2.3)–(2.4) have been proved in [Reference Nagai and Yamada37, Lemmas 2.3–2.4].

Now, the Gagliardo–Nirenberg inequality will be given as follows.

Lemma 2.3. Let $1\leq p\lt \infty$ , $1\leq q,r\leq \infty$ and $\theta \in [0,1]$ such that

\begin{align*} \frac{1}{p}=\theta \!\left (\frac{1}{r}-\frac{1}{2}\right )+(1-\theta )\frac{1}{q}. \end{align*}

Then for any $u(x)\in W^{1,r}(\mathbb{R}^2)\cap L^{q}(\mathbb{R}^2)$ , there exists a constant $c=c(p,q,r)\gt 0$ such that

\begin{align*} \|u\|_{L^{p}(\mathbb{R}^2)}\leq c\|u\|^{1-\theta }_{L^q(\mathbb{R}^2)}\|\nabla u\|^{\theta }_{L^r(\mathbb{R}^2)}. \end{align*}

Proof. The proof of this lemma has been given in [Reference Giga, Giga and Saal14, Theorem 6.1.1].

3. Proof of Theorem 1.1

In this section, the proof of Theorem 1.1 will be divided into several steps.

Now we would like to give an equality for the modified free energy functional $ \mathcal{F}$ . Notice that the equality for one-single variable can be found in [Reference Nagai and Ogawa34, Proposition 4.1].

Lemma 3.1. Let $\alpha _{i,j},\beta _{j},\gamma _{i,j}\in \mathbb{R}$ , $i\in \mathcal{I}$ , $j\in \mathcal{J}$ and $T\gt 0$ . Let $(\boldsymbol{{u}},\boldsymbol{{v}})$ be a local solution of (1.5) with initial data $(\boldsymbol{{u}}_0,\boldsymbol{{v}}_0)$ . Assume that there exist positive constants $a_1,\cdots,a_n$ and a positive definite matrix $\boldsymbol{{B}}=(b_{j,l})_{m\times m}$ such that

(3.1) \begin{equation} \boldsymbol{{B}}\boldsymbol{\gamma }_i=a_i\boldsymbol{\alpha }_i,\quad \forall \,\,i\in \mathcal{I}. \end{equation}

Then,

(3.2) \begin{align} \frac{d}{dt}&\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]+\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\left (\partial _t\boldsymbol{{v}}\right )dx +\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\Bigg \{u_i\!\left |\nabla \!\left (\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2\nonumber\\[5pt] &+\left |\nabla \!\left (\log\! (u_i+1)-\frac{1}{2}\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2\Bigg \}dx\\[5pt] =&\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j \right )dx+\frac{1}{4}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\left |\nabla \!\left (\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx \nonumber \end{align}

for $ t\in (0,T)$ , where $\mathcal{F}$ is given by

\begin{align*} \mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]=&\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx. \end{align*}

Moreover, there exists a constant $c\gt 0$ such that

(3.3) \begin{align} \frac{d}{dt}&\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]+\frac{1}{2}\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dx\nonumber\\[9pt] &+\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\Bigg \{u_i\left |\nabla \!\left (\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2\nonumber\\[9pt] &+\left |\nabla \!\left (\log\! (u_i+1)-\frac{1}{2}\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2\Bigg \}dx\\[9pt] \leq &\frac{1}{4}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\left |\nabla \!\left (\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx+c,\quad t\in (0,T).\nonumber \end{align}

Proof. Multiplying both sides of $i$ -th equation in (1.5) by $a_i\log\! (u_i+1)$ , integrating by parts and summing them with respect to $i$ , we have

(3.4) \begin{align} \frac{d}{dt}&\left [\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx\right ]\nonumber\\[9pt] =&-\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx +\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\frac{u_i}{u_i+1}\nabla u_i\cdot \nabla v_jdx. \end{align}

By the symmetry of $\boldsymbol{{B}}$ , it is clear that

\begin{align*} \frac{d}{dt}&\left [\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx\right ]\\[8pt] =&\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left [\nabla \!\left (\partial _tv_j\right )\cdot \nabla v_l+\nabla v_j\cdot \nabla \!\left (\partial _tv_l\right ) \right ]dx\\[8pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_l\partial _tv_j+v_j\partial _tv_l\right )dx\\[8pt] =&\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla \!\left (\partial _tv_j\right )\cdot \nabla v_ldx+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _l\int _{\mathbb{R}^2}\left (v_l\partial _tv_j+v_j\partial _tv_l\right )dx\\[8pt] =&-\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\partial _tv_j\right )\left (\Delta v_l-\beta _{l}v_l\right )dx\\[8pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j\right )dx, \end{align*}

and then using (3.1) and the $j$ -th equation in (1.5), it implies that

\begin{align*} \begin{split} \frac{d}{dt}&\left [\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx\right ]\\[5pt] =&-\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\partial _tv_j\partial _tv_ldx+\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}b_{j,l}\gamma _{i,l}\int _{\mathbb{R}^2}u_i\partial _tv_jdx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j \right )dx\\[5pt] =&-\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\left (\partial _t\boldsymbol{{v}}\right )dx+\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_i\partial _tv_jdx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j\right )dx. \end{split} \end{align*}

Since

\begin{align*} \begin{split} \frac{d}{dt}\left ({-}\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx\right )=&-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}v_j\partial _tu_idx\\[5pt] &-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_i\partial _tv_jdx \end{split} \end{align*}

and

\begin{align*} \begin{split} -\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}v_j\partial _tu_idx=&-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\left (\Delta u_i-\sum \limits ^m_{l=1}\alpha _{i,l}\nabla \cdot (u_i\nabla v_l)\right )v_jdx\\[5pt] =&\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\nabla u_i\cdot \nabla v_jdx\\[5pt] &-\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\left |\nabla \!\left (\sum \limits ^{m}_{j=1}\alpha _{i,j} v_j\right )\right |^2dx, \end{split} \end{align*}

one can obtain

\begin{align*} \begin{split} \frac{d}{dt}&\left [\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx\right ]\\[5pt] =&-\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dx+\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\nabla u_i\cdot \nabla v_jdx\\[5pt] &-\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\left |\nabla \!\left (\sum \limits ^{m}_{j=1}\alpha _{i,j} v_j\right )\right |^2dx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j\right )dx, \end{split} \end{align*}

which together with (3.4) yields that

\begin{align*} \begin{split} \frac{d}{dt}\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]=&-\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dx-\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx\\[5pt] &+\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{2u_i+1}{u_i+1}\nabla u_i\cdot \nabla \!\left (\sum \limits ^{m}_{j=1}\alpha _{i,j} v_j\right )dx-\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\left |\nabla \!\left (\sum \limits ^{m}_{j=1}\alpha _{i,j} v_j\right )\right |^2dx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j\right )dx\\[5pt] =&-\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dx-\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\left |\nabla \!\left (\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx\\[5pt] &-\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\left |\nabla \!\left (\log\! (u_i+1)-\frac{1}{2}\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx\\[5pt] &+\frac{1}{4}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\left |\nabla \!\left (\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j\right )dx. \end{split} \end{align*}

Hence, we have proved (3.2).

By means of the positivity of $\boldsymbol{{B}}$ , there exists a constant $c_1\gt 0$ such that

\begin{align*} c_1\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\partial _tv_j\partial _tv_ldx\geq \sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\partial _tv_j\right |^2dx, \end{align*}

then we have

\begin{align*} &\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}\left (v_j\partial _tv_l-v_l\partial _tv_j\right )dx\\[5pt]& \leq \beta ^*\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}|b_{j,l}|\int _{\mathbb{R}^2}\left |\partial _tv_j\right | |v_l|dx\\[5pt]& \leq \frac{1}{2c_1}\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\partial _tv_j\right |^2dx+\frac{c_1(b^*\beta ^*|\mathcal{J}|)^2}{2}\sum \limits ^m_{j=1}\int _{\mathbb{R}^2}v^2_jdx\\[5pt]& \leq \frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\partial _tv_j\partial _tv_ldx+\frac{c_1(b^*\beta ^*|\mathcal{J}|)^2}{2}\sum \limits ^m_{j=1}\int _{\mathbb{R}^2}v^2_jdx \end{align*}

due to Young’s inequality. Employing (3.2) and the boundedness of $v_j$ in $L^2$ space, we can obtain a constant $c\gt 0$ such that (3.3) holds for all $t\in (0,T)$ .

Remark 3. Let us define

\begin{align*} \begin{split} \mathcal{E}[\boldsymbol{{u}},\boldsymbol{{v}}]=&\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\log u_idx-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx. \end{split} \end{align*}

Then, $d\mathcal{E}[\boldsymbol{{u}},\boldsymbol{{v}}]/dt\leq 0$ for $t\in (0,T)$ if one has (3.1) and $\boldsymbol{{B}}D_{\boldsymbol{\beta }}$ is symmetric.

We list the Moser–Trudinger inequality for system on the two-dimensional unit sphere [Reference Shafrir and Wolansky38, Theorem 2 (ii)].

Lemma 3.2. Let $\mathbb{S}^2\subset \mathbb{R}^3$ be the unit sphere. Assume $\boldsymbol{{S}}=(s_{i,k})_{n\times n}$ is a positive definite matrix with $s_{i,k}\geq 0$ , $i,k\in \mathcal{I}$ , and $\boldsymbol{{M}}\in (\mathbb{R}_{+})^n$ . Then for $\rho _i\in H^1(\mathbb{S}^2)$ satisfying $\int _{\mathbb{S}^2}\rho _i=0,\,\,i\in \mathcal{I}$ ,

(3.5) \begin{equation} \left \{ \begin{array}{l@{\quad}l@{\quad}l@{\quad}l} \Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}}(\boldsymbol{{M}})\geq 0,\quad \forall \,\emptyset \neq \mathcal{K}\subset \mathcal{I},\\[5pt]\text{if}\,\,\Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}}(\boldsymbol{{M}})=0\,\,\textit{for some}\,\,\mathcal{K},\,\,\textit{then}\,\,s_{i,i}+\Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}\setminus \{i\}}(\boldsymbol{{M}})\gt 0,\,\,\forall \,{i\in \mathcal{K}}, \end{array} \right . \end{equation}

is the necessary and sufficient condition for the existence of a constant $B\gt 0$ such that

(3.6) \begin{equation} \begin{split} \Phi _{\mathbb{S}^2}(\boldsymbol{\rho })= \frac{1}{2}\sum _{i,k\in \mathcal{I}}s_{i,k}\int _{\mathbb{S}^2}\nabla \rho _i\cdot \nabla \rho _k-\sum _{i\in \mathcal{I}}M_i\log \!\left (\frac{1}{4\pi }\int _{\mathbb{S}^2}\exp \!\left(\sum \limits _{k\in \mathcal{I}}s_{i,k}\rho _k\right)\right )\geq -B. \end{split} \end{equation}

Now we use the stereographic projection $\mathcal{S}$ to transform the inequality for system in Lemma 3.2 to $\mathbb{R}^2$ . In fact, we associate with each $\rho _i\,:\,\mathbb{S}^2\rightarrow \mathbb{R}^2$ a function $\widetilde{\rho }_i\,:\,\mathbb{R}^2\rightarrow \mathbb{S}^2$ via the transformation

(3.7) \begin{equation} \begin{split} \begin{cases} \widetilde{\rho }_i\leftrightarrow &\rho _i=\widetilde{\rho }_i\circ \mathcal{S},\\[5pt] \rho _i\leftrightarrow &\widetilde{\rho }_i= \rho _i\circ \mathcal{S}^{-1},\quad i\in \mathcal{I}. \end{cases} \end{split} \end{equation}

By a simple calculation, we have

Lemma 3.3. Let $\boldsymbol{{S}}=(s_{i,k})_{n\times n}$ be a positive definite matrix with $s_{i,k}\geq 0$ , $i,k\in \mathcal{I}$ . Then for $\widetilde{\rho }_i\in H^1(\mathbb{R}^2)$ , $i\in \mathcal{I}$ , condition (3.5) is the necessary and sufficient condition for the existence of a constant $B\gt 0$ such that

(3.8) \begin{equation} \begin{split} \frac{1}{2}&\sum _{i,k\in \mathcal{I}}s_{i,k}\int _{\mathbb{R}^2}\nabla \widetilde{\rho }_i\cdot \nabla \widetilde{\rho }_kdx-\sum _{i\in \mathcal{I}}M_i\log \!\left (\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left(\sum \limits _{k\in \mathcal{I}}s_{i,k} \widetilde{\rho }_k\right)H(x)dx\right )\\[5pt] &+\frac{1}{4\pi }\sum \limits _{i,k\in \mathcal{I}}M_is_{i,k}\int _{\mathbb{R}^2}\widetilde{\rho }_kH(x)dx\geq -B, \end{split} \end{equation}

where $H(x)=4/\big(1+|x|^2\big)^2$ .

Proof. Let $\rho _i\in H^1(\mathbb{S}^2)$ , $i\in \mathcal{I}$ . Then, we take

\begin{align*} \rho _i-\frac{1}{4\pi }\int _{\mathbb{S}^2}\rho _i,\quad i\in \mathcal{I}, \end{align*}

for $\rho _i$ in (3.6), and obtain that

(3.9) \begin{equation} \begin{split} \frac{1}{2}&\sum _{i,k\in \mathcal{I}}s_{i,k}\int _{\mathbb{S}^2}\nabla \rho _i\cdot \nabla \rho _k-\sum _{i\in \mathcal{I}}M_i\log \!\left (\frac{1}{4\pi }\int _{\mathbb{S}^2}\exp \!\left(\sum \limits _{k\in \mathcal{I}}s_{i,k}\rho _k\right)\right )\\[5pt] &+\frac{1}{4\pi }\sum \limits _{i,k\in \mathcal{I}}M_is_{i,k}\int _{\mathbb{S}^2}\rho _k\geq -B \end{split} \end{equation}

with some constant $B\gt 0$ if and only if (3.5) is valid. Using the transformation (3.7), one derives that $\int _{\mathbb{S}^2}\rho _k=\int _{\mathbb{R}^2}\widetilde{\rho }_kH(x)dx$ , and

\begin{align*} \int _{\mathbb{S}^2}\nabla \rho _i\cdot \nabla \rho _k=\int _{\mathbb{R}^2}\nabla \widetilde{\rho }_i\cdot \nabla \widetilde{\rho }_kdx, \end{align*}

as well as

\begin{align*} \int _{\mathbb{S}^2}\exp \!\left(\sum \limits _{k\in \mathcal{I}}s_{i,k}\rho _k\right)=\int _{\mathbb{R}^2}\exp \!\left(\sum \limits _{k\in \mathcal{I}}s_{i,k}\widetilde{\rho }_k\right)H(x)dx. \end{align*}

Hence, we obtain (3.8) from (3.9).

Under the condition (1.10), utilising the above Moser–Trudinger inequality for system in the whole space, we can give an estimate on the interaction term consisting of $u_i$ and $v_j$ in $ \mathcal{F}$ at the first step. The idea is mainly combined with some of the work in [Reference Calvez and Corrias5] and [Reference Mizoguchi31] and has been applied to one-species or two-species chemotaxis system with two chemicals [Reference Hong, Wang, Yu and Zhang19, Reference Nagai, Seki and Yamada35].

Lemma 3.4. Suppose $\boldsymbol{\alpha }=(\alpha _{i,j})_{n\times m}$ with $R(\boldsymbol{\alpha })=m$ , $\beta _{j}\in \mathbb{R}$ , $j\in \mathcal{J}$ , $\boldsymbol{\gamma }=(\gamma _{i,j})_{n\times m}$ and $T\gt 0$ . Assume that $(\boldsymbol{{u}},\boldsymbol{{v}})$ is a local solution of Cauchy problem (1.5) with initial data $(\boldsymbol{{u}}_0,\boldsymbol{{v}}_0)$ . Suppose that there exist positive constants $a_1,\cdots,a_n$ and a positive definite matrix $\boldsymbol{{R}}=(r_{i,k})_{n\times n}$ with $r_{i,k}\geq 0$ , $i,k\in \mathcal{I}$ , such that

\begin{align*} \boldsymbol{\alpha }^T\boldsymbol{{R}}^{-1}\boldsymbol{\alpha }\boldsymbol{\gamma }_i=a_i\boldsymbol{\alpha }_i,\quad \forall \,\,i\in \mathcal{I}. \end{align*}

Then for any

(3.10) \begin{equation}{} \begin{split} 8\pi \sum \limits _{i\in \mathcal{K}}a_im_i\gt \sum \limits _{i,k\in \mathcal{K}}a_ia_kr_{i,k}m_im_k,\quad \forall \,\,\emptyset \neq \mathcal{K} \subset \mathcal{I}, \end{split} \end{equation}

there exist a small $\epsilon \gt 0$ and a constant $c\gt 0$ such that

\begin{align*} \sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx &\leq \frac{1}{2(1+\epsilon )}\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla v_j\cdot \nabla v_ldx\\[5pt] & \quad +\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1) dx\\[5pt] &\quad -\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+c, \end{align*}

where $\boldsymbol{{B}}=(b_{j,l})_{m\times m}=\boldsymbol{\alpha }^T\boldsymbol{{R}}^{-1}\boldsymbol{\alpha }$ and $H(x)=4/\big(1+|x|^2\big)^2$ .

Proof. From our assumption, $\boldsymbol{{B}}$ is a positive definite matrix satisfying $\boldsymbol{{B}}\boldsymbol{\gamma }_i=a_i\boldsymbol{\alpha }_i$ , $\forall \,i\in \mathcal{I}$ . Define $\boldsymbol{{T}}\,:\!=\,(t_{i,j})_{n\times m}=\boldsymbol{{R}}^{-1}\boldsymbol{\alpha }$ . Then it is easy to find that

\begin{align*} \alpha _{i,j}=\sum \limits ^n_{k=1}r_{i,k}t_{k,j} \end{align*}

and

\begin{align*} b_{j,l}=\sum \limits ^n_{i=1}\sum \limits ^n_{k=1}t_{i,j}r_{i,k}t_{k,l}. \end{align*}

Moreover, one can pick a positive definite matrix $\boldsymbol{{S}}=(s_{i,k})_{n\times n}$ with $s_{i,k}=(1+\epsilon )r_{i,k}\geq 0$ , $i,k\in \mathcal{I}$ , such that

\begin{align*} \sum \limits ^n_{k=1}s_{i,k}t_{k,j}=(1+\epsilon )\alpha _{i,j} \end{align*}

and

\begin{align*} \sum \limits ^n_{i=1}\sum \limits ^n_{k=1}t_{i,j}s_{i,k}t_{k,l}=(1+\epsilon )b_{j,l}. \end{align*}

Note that Lemma 2.1 $(v)$ will help us to find a constant $A_1\gt 0$ such that

\begin{align*} \sum \limits ^m_{j=1}\|v_j(\cdot,t)\|_{L^1(\mathbb{R}^2)}\leq A_{1}\quad \text{for}\,\,t\in (0,T). \end{align*}

Chosen $\epsilon \gt 0$ small enough, (3.10) implies

(3.11) \begin{equation}{} \begin{split} \sum \limits _{i\in \mathcal{K}}a_im_i\left [8\pi -(1+\epsilon )^2\sum \limits _{k\in \mathcal{K}}a_kr_{i,k}m_k\right ]\gt 0,\quad \forall \,\,\emptyset \neq \mathcal{K} \subset \mathcal{I}. \end{split} \end{equation}

Let

\begin{align*} \overline{\rho }_i(t)& \,:\!=\,\left (\sum \limits ^m_{j=1}\alpha _{i,j}v_j(t)-s\sum \limits ^n_{k=1}r_{i,k}\right )_{+},\quad \widetilde{\rho }_i(t)\,:\!=\,\left (\sum \limits ^m_{j=1}t_{i,j}v_j(t)-s\right )_{+},\quad i\in \mathcal{I},\\[5pt] \overline{\Omega }_i(t) & \,:\!=\,\left \{x\in \mathbb{R}^2\,:\,\sum \limits ^m_{j=1}\alpha _{i,j}v_j(t)\gt s\sum \limits ^n_{k=1}r_{i,k}\right \},\quad \widetilde{\Omega }_i(t)\,:\!=\,\left \{x\in \mathbb{R}^2:\sum \limits ^m_{j=1}t_{i,j}v_j(t)\gt s\right \},\quad i\in \mathcal{I}, \end{align*}

and

\begin{align*} \Omega (t)\,:\!=\,\bigcup \limits ^{n}_{i=1}\widetilde{\Omega }_i(t),\quad m_i(t)\,:\!=\,\int _{\overline{\Omega }_i(t)}u_idx\leq m_i,\quad i\in \mathcal{I}. \end{align*}

We claim some facts in the following. First, since $\widetilde{\rho }_i(t)\in H^1_0(\widetilde{\Omega }_i(t))$ and $\nabla \widetilde{\rho }_i(t)=\sum \limits ^m_{j=1}t_{i,j}\nabla v_j(t)$ in ${\Omega }(t)$ , then $\widetilde{\rho }_i(t)\in H^1_0({\Omega }(t))$ for all $i\in \mathcal{I}$ . Second, the Lebesgue measure of $\Omega (t)$ , denoted by $|\Omega (t)|$ , is finite. This is because

\begin{align*} s\cdot |\Omega (t)|\leq s\sum \limits ^n_{i=1}|\widetilde{\Omega }_i(t)| \leq \sum \limits ^n_{i=1}\left \|\sum \limits ^m_{j=1}t_{i,j}v_j(\cdot,t)\right \|_{L^1(\mathbb{R}^2)}\leq c_1 \end{align*}

implies that $|\Omega (t)|\leq c_{1}/s$ . Third, $|\overline{\Omega }_i(t)|\leq c_{1}/s$ holds out due to

\begin{align*} \overline{\Omega }_i(t)\subset \Omega (t), \quad \forall \,\,i\in \mathcal{I}. \end{align*}

Finally, without loss of generality, we assume $|\Omega (t)|\geq 1$ and $|\overline{\Omega }_i(t)|\gt 0$ for all $i\in \mathcal{I}$ . If $|\overline{\Omega }_i(t)|=0$ for some $i\in \mathcal{I}$ , classical techniques are sufficient to analyse this case.

Fixing $i\in \mathcal{I}$ , it is obvious that

\begin{align*} a_i\sum \limits ^{m}_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx & = a_i\int _{\overline{\Omega }_i(t)}u_i\!\left (\sum \limits ^{m}_{j=1}\alpha _{i,j}v_j\right )dx+a_i\int _{\mathbb{R}^2\setminus \overline{\Omega }_i(t)}u_i\!\left (\sum \limits ^{m}_{j=1}\alpha _{i,j}v_j\right )dx\\[5pt]& \leq a_i\int _{\overline{\Omega }_i(t)}u_i \overline{\rho }_idx+sa_im_i\left (\sum \limits ^n_{k=1}r_{i,k}\right )\\[5pt]& \leq a_i\int _{\mathbb{R}^2}u_i \overline{\rho }_idx+sa_im_i\left (\sum \limits ^n_{k=1}r_{i,k}\right ). \end{align*}

Denote $u^*_i=m_i\exp ({(1+\epsilon )\overline{\rho }_i(x,t)})H(x)\left (\int _{\mathbb{R}^2}\exp ({(1+\epsilon )\overline{\rho }_i(x,t)})H(x)dx\right )^{-1}$ . Then, $\|u^*_i\|_1=m_i$ and a classical entropy minimisation in [Reference Calvez and Corrias5, Lemma 2.1] implies that the function

\begin{align*} \mathcal{E}(u_i;\,\psi )=\int _{\mathbb{R}^2}(u_i(x)\log u_i(x)-u_i(x)\psi (x))dx\quad \text{with any}\,\,\exp (\psi )\in L^1(\mathbb{R}^2), \end{align*}

satisfies

\begin{align*} \mathcal{E}\!\left (u_i;\,(1+\epsilon )\overline{\rho }_i+\log H\right ) & \geq \mathcal{E}\!\left (u^*_i;\,(1+\epsilon )\overline{\rho }_i+\log H\right )\\[5pt] & =m_i\log m_i-m_i\log \!\left (\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp ({(1+\epsilon )\overline{\rho }_i(x,t)})H(x)dx\right )-m_i\log\! (4\pi ). \end{align*}

Combining the aforementioned findings, we arrive at the following

\begin{align*} &(1+\epsilon )a_i\sum \limits ^{m}_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx- a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1) dx\\[5pt]& \leq (1+\epsilon )a_i\int _{\mathbb{R}^2}u_i\overline{\rho }_idx- a_i\int _{\mathbb{R}^2}u_i\log u_i dx+(1+\epsilon )sa_im_i\left (\sum \limits ^n_{k=1}r_{i,k}\right )\\[5pt]& \leq a_im_i\log \!\left \{\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left [{(1+\epsilon )\overline{\rho }_i(x,t)}\right ]H(x)dx\right \}-a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx\\[5pt] & \quad +(1+\epsilon )sa_im_i\left (\sum \limits ^n_{k=1}r_{i,k}\right )+a_im_i\log \frac{4\pi }{m_i}\\[5pt]& \leq a_im_i\log \!\left \{\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left [(1+\epsilon )\left (\sum \limits ^m_{j=1}\alpha _{i,j}v_j(t)-s\sum \limits ^n_{k=1}r_{i,k}\right )_+\right ]H(x) dx\right \}\\[5pt] &\quad -a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+(1+\epsilon )sa_im_i\left (\sum \limits ^n_{k=1}r_{i,k}\right )+a_im_i\log \frac{4\pi }{m_i}, \end{align*}

where the choice of matrix $\boldsymbol{{S}}$ allows one to conclude that

\begin{align*} & (1+\epsilon )a_i\sum \limits ^{m}_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx- a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1) dx\\[5pt] &\leq a_im_i\log \!\left \{\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left [\sum \limits ^n_{k=1}s_{i,k}\left (\sum \limits ^m_{j=1}t_{k,j}v_j-s\right )_{+}\right ]H(x)dx\right \}\\[5pt] & \quad -a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+c_{2i}\\[5pt] &= a_im_i\log \!\left [\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left (\sum \limits ^n_{k=1}s_{i,k}\widetilde{\rho }_k\right )H(x)dx\right ]-a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+c_{2i}, \end{align*}

where $c_{2i}=(1+\epsilon )sa_im_i\left (\sum \limits ^n_{k=1}r_{i,k}\right )+a_im_i\log\! (4\pi/m_i)$ , $i\in \mathcal{I}$ . Then, summing it with respect to $i$ from $i=1$ to $i=n$ , we get

(3.12) \begin{align} \sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx\leq &\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_im_i\log \left [\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left (\sum \limits ^n_{k=1}s_{i,k}\widetilde{\rho }_k\right )H(x)dx\right ]\nonumber\\[5pt] &+\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1) dx\\[5pt] &-\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}c_{2i}.\nonumber \end{align}

Choose

\begin{align*} M_i=a_im_i(1+\epsilon ),\quad i\in \mathcal{I}. \end{align*}

Since (3.11) implies that

\begin{align*} \Lambda ^{\boldsymbol{{S}}}_{\mathcal{K}}(\boldsymbol{{M}}) & = 8\pi \sum \limits _{i\in \mathcal{K}}M_i-\sum \limits _{i,k\in \mathcal{K}}s_{i,k}M_iM_k\\[5pt] &=(1+\epsilon )\left [8\pi \sum \limits _{i\in \mathcal{K}}a_im_i-(1+\epsilon )^2\sum \limits _{i,k\in \mathcal{K}}a_ia_kr_{i,k}m_im_k\right ]\\[5pt]& \gt 0,\quad \forall \,\,\emptyset \neq \mathcal{K}\subset \mathcal{I}, \end{align*}

then the Moser–Trudinger inequality for system in Lemma 3.3 helps us to get that

\begin{align*} &\sum \limits ^n_{i=1} M_i\log \!\left [\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left (\sum \limits ^n_{k=1}s_{i,k}\widetilde{\rho }_k\right )H(x)dx\right ]\\[5pt] &=(1+\epsilon )\sum \limits ^n_{i=1}a_im_i\log \!\left [\frac{1}{4\pi }\int _{\mathbb{R}^2}\exp \!\left (\sum \limits ^n_{k=1}s_{i,k}\widetilde{\rho }_k\right )H(x)dx\right ]\\[5pt]& \leq \frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^n_{k=1}s_{i,k}\int _{\mathbb{R}^2}\nabla \widetilde{\rho }_i\cdot \nabla \widetilde{\rho }_k dx+\frac{1}{4\pi }\sum \limits _{i,k\in \mathcal{I}}M_is_{i,k}\int _{\mathbb{R}^2}\widetilde{\rho }_kH(x)dx+B\\[5pt]& \leq \frac{1}{2}(1+\epsilon )\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla v_j\cdot \nabla v_ldx+c_3, \end{align*}

where we have used the bound of $\|v_j\|_{L^1(\mathbb{R}^2)}$ , $j\in \mathcal{J}$ , and $\int _{\mathbb{R}^2}H(x)dx=4\pi$ . This together with (3.12) and the positivity of $\boldsymbol{{B}}$ implies that

\begin{align*} \sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx & \leq \frac{1}{2(1+\epsilon )}\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla v_j\cdot \nabla v_ldx\\[5pt] &+\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1) dx\\[5pt] &-\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}c_{2i}+\frac{c_3}{(1+\epsilon )^2}. \end{align*}

Therefore, we have finished the proof of this lemma.

As a consequence of Lemma 3.4, the bound on modified total entropy $\mathcal{S}$ could be obtained.

Lemma 3.5. Let $T\gt 0$ . Under the same assumptions in Lemma 3.4, there exists a constant $c=c(T)\gt 0$ such that

(3.13) \begin{equation} \sum \limits ^n_{i=1}\int _{\mathbb{R}^2}(u_i(x,t)+1)\log\! (u_i(x,t)+1)dx\leq c \end{equation}

and

(3.14) \begin{equation} \sum \limits ^m_{j=1}\int ^t_0\int _{\mathbb{R}^2}\left |\partial _tv_j\right |^2dxd\tau \leq c \end{equation}

hold out for $t\in (0,T)$ .

Proof. Notice that positive definite matrix $\boldsymbol{{B}}=(b_{j,l})_{m\times m}= \boldsymbol{\alpha }^{T}{\boldsymbol{{R}}}^{-1}\boldsymbol{\alpha }$ satisfies

\begin{align*} \boldsymbol{{B}}\boldsymbol{\gamma }_i=a_i\boldsymbol{\alpha }_i,\quad \forall \,\, i\in \mathcal{I}. \end{align*}

Then thanks to Lemma 3.1, there exists a modified free energy functional $\mathcal{F}$ given by

(3.15) \begin{equation} \begin{split} \mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]=&\sum \limits ^{n}_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx \end{split} \end{equation}

satisfying

(3.16) \begin{align} \frac{d}{dt}\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]&+\frac{1}{2}\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dx+\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\bigg \{u_i\left |\nabla \!\left (\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2\nonumber\\[5pt] &+\left |\nabla \!\left (\log\! (u_i+1)-\frac{1}{2}\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2\bigg \}dx\nonumber\\[5pt]& \leq \frac{1}{4}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\left |\nabla \!\left (\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx+c_1\\[5pt]& \leq \frac{a^*\alpha ^*|\mathcal{I}||\mathcal{J}|}{4}\sum \limits ^m_{j=1}\int _{\mathbb{R}^2}\left |\nabla v_j\right |^2dx+c_1\nonumber \end{align}

with some constant $c_1\gt 0$ . Moreover, Lemma 3.4 implies the existence of small $\epsilon \gt 0$ such that

(3.17) \begin{align} \sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx \leq &\frac{1}{2(1+\epsilon )}\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla v_j\cdot \nabla v_ldx\nonumber\\[5pt] &+\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1) dx\\[5pt] &-\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+c_2\nonumber \end{align}

is true with some $c_2\gt 0$ . On the one hand, one has

(3.18) \begin{equation} \begin{split} \mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]\geq &\frac{\epsilon }{1+\epsilon }\sum \limits ^{n}_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx+\frac{\epsilon }{2(1+\epsilon )}\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla v_j\cdot \nabla v_ldx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}v_jv_ldx+\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx-c_2 \end{split} \end{equation}

from (3.15) and (3.17). On the other hand, we have

\begin{align*} \begin{split} \frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla v_j\cdot \nabla v_ldx&=\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]-\sum \limits ^{n}_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx\\[5pt] &- \frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}v_jv_ldx+\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}u_iv_jdx, \end{split} \end{align*}

which together with (3.17) ensures that

\begin{align*} \begin{split} \frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\nabla v_j\cdot \nabla v_ldx\leq &\frac{1+\epsilon }{\epsilon }\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]-\sum \limits ^{n}_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx\\[5pt] &- \frac{1+\epsilon }{2\epsilon }\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _{l}\int _{\mathbb{R}^2}v_jv_ldx\\[5pt] &-\frac{1}{\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+\frac{(1+\epsilon )c_2}{\epsilon }. \end{split} \end{align*}

Further, using the positivity of $\boldsymbol{{B}}$ and the bound on the $\|v_j\|_{L^2(\mathbb{R}^2)}$ by Lemma 2.1 $(v)$ , there exists a constant $c_3\gt 0$ such that

\begin{align*} \sum \limits ^m_{j=1}\int _{\mathbb{R}^2}|\nabla v_j|^2dx\leq &\frac{(1+\epsilon )c_3}{\epsilon }\left (\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]-\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+1\right ). \end{align*}

Applying above inequalities and from (3.16), it follows that

(3.19) \begin{equation} \begin{split} \frac{d}{dt}&\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]+\frac{1}{2}\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dx+\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\left |\nabla \!\left (\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx\\[5pt] \leq & \frac{(1+\epsilon )c_3a^*\alpha ^*|\mathcal{I}||\mathcal{J}|}{4\epsilon }\left (\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]-\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx+1\right )+c_1 \end{split} \end{equation}

for $t\in (0,T)$ . To estimate the second term on the right side of (3.19), we first observe that

\begin{align*} \begin{split} -\frac{1}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log H(x)dx=&\frac{2}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log\! \big(1+|x|^2\big)dx\\[5pt] &-\frac{2\log 2}{1+\epsilon }\sum \limits ^n_{i=1}a_im_i\\[5pt] \leq &\frac{2}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log\! \big(1+|x|^2\big)dx, \end{split} \end{align*}

where we take the derivative of the right term to see that

\begin{align*} \begin{split} \frac{d}{dt}&\left (\frac{2}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log\! \big(1+|x|^2\big)dx\right )\\[5pt] =&-\frac{2}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\nabla \!\left [\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right ]\cdot \nabla \log\! \big(1+|x|^2\big)dx\\[5pt] &-\frac{2}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{\nabla u_i}{u_i+1}\cdot \nabla \log\! \big(1+|x|^2\big)dx \\ \leq &\frac{1}{2}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\left |\nabla \!\left (\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx\\[5pt] &+\frac{2}{(1+\epsilon )^2}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\big|\nabla \log\! \big(1+|x|^2\big)\big|^2dx\\[5pt] &+\frac{2}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\log\! (u_i+1)\Delta \log\! \big(1+|x|^2\big)dx\\[5pt] \leq &\frac{1}{2}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i\left |\nabla \!\left (\log\! (u_i+1)-\sum \limits ^m_{j=1}\alpha _{i,j}v_j\right )\right |^2dx+10\sum \limits ^n_{i=1}a_im_i \quad \text{for}\,\,t\in (0,T), \end{split} \end{align*}

by Young’s inequality, since

\begin{align*} \big|\nabla \log\! \big(1+|x|^2\big)\big|= \left |\frac{2x}{1+|x|^2}\right |\leq 1,\quad \big|\Delta \log\! \big(1+|x|^2\big)\big|= \left |\frac{4}{\big(1+|x|^2\big)^2}\right |\leq 4, \end{align*}

and $\log\! (s+1)\leq s$ for all $s\gt 0$ . Thereby, denoting

\begin{align*} y(t)\,:\!=\,\mathcal{F}[\boldsymbol{{u}},\boldsymbol{{v}}]+\frac{2}{1+\epsilon }\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}u_i(x)\log\! \big(1+|x|^2\big)dx\quad \text{for}\,\,t\in (0,T), \end{align*}

one derives that

\begin{align*} y^{\prime}(t)+\frac{1}{2}\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dx\leq \frac{(1+\epsilon )c_3a^*\alpha ^*|\mathcal{I}||\mathcal{J}|}{4\epsilon }\left (y(t)+1\right )+c_4\quad \text{for}\,\,t\in (0,T), \end{align*}

with $c_4=10\sum \limits ^n_{i=1}a_im_i+c_1$ , where the Gronwall argument means that

\begin{align*}y(t)+&\frac{1}{2}\int ^t_0\int _{\mathbb{R}^2}\left (\partial _t\boldsymbol{{v}}\right )^T\boldsymbol{{B}}\!\left (\partial _t\boldsymbol{{v}}\right )dxd\tau \leq \left (y(0)+1\right )e^{\frac{(1+\epsilon )c_3a^*\alpha ^*|\mathcal{I}||\mathcal{J}|}{4\epsilon }T}+c_4 \quad \text{for}\,\,t\in (0,T). \end{align*}

Hence, we have proved (3.13)–(3.14) due to (3.18), the choices of positive $a_i$ , $i\in \mathcal{I}$ , the positivity of $\boldsymbol{{B}}$ and the bound on the $\|v_j\|_{L^2(\mathbb{R}^2)}$ , $j\in \mathcal{J}$ .

A straightforward argument [Reference Nagai, Senba and Yoshida36, Lemma 3.6] could be indeed used to obtain $L^2$ estimates for the solutions by the bound on $\mathcal{S}$ .

Lemma 3.6. For $T\gt 0$ , there exists a constant $c=c(T)\gt 0$ such that

\begin{align*} \sum \limits ^n_{i=1}\|u_i(\cdot,t)\|_{L^2(\mathbb{R}^2)}\leq c\quad \text{for}\,\,\,\,t\in (0,T). \end{align*}

Proof. From Lemma 3.5, there exists a constant $c_1=c_1(T)\gt 0$ such that

\begin{align*} \sum \limits ^n_{i=1}\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx\leq c_1 \end{align*}

and

(3.20) \begin{equation} \begin{split} \sum \limits ^m_{j=1}\int ^T_0\int _{\mathbb{R}^2}\left |\partial _tv_j\right |^2dxd\tau \leq c_1. \end{split} \end{equation}

We multiply the $i$ -the equation in (1.5) by $u_i$ and integrate them over $\mathbb{R}^2$ to have

(3.21) \begin{equation} \begin{split} \frac{1}{2}\frac{d}{dt}\left (\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}u^2_idx\right )=&-\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}|\nabla u_i|^2dx-\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u^2_i\Delta v_jdx\\[5pt] =&-\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}|\nabla u_i|^2dx-\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u^2_i\left (\partial _t v_j+\beta _jv_j-\sum \limits ^n_{k=1}\gamma _{k,j}u_k\right )dx\\[5pt] =&-\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}|\nabla u_i|^2dx+\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^n_{k=1}\sum \limits ^m_{j=1}\alpha _{i,j}\gamma _{k,j}\int _{\mathbb{R}^2}u^2_iu_kdx\\[5pt] &-\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\alpha _{i,j}\beta _j\int _{\mathbb{R}^2}u^2_iv_jdx-\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u^2_i\partial _tv_jdx \end{split} \end{equation}

for $t\in (0,T)$ . It is clear that

\begin{align*} \begin{split} &\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^n_{k=1}\sum \limits ^m_{j=1}\alpha _{i,j}\gamma _{k,j}\int _{\mathbb{R}^2}u^2_iu_kdx-\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\alpha _{i,j}\beta _j\int _{\mathbb{R}^2}u^2_iv_jdx\\[5pt] &\leq \frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^n_{k=1}\sum \limits ^m_{j=1}|\alpha _{i,j}||\gamma _{k,j}|\int _{\mathbb{R}^2}\left(u^3_i+u^3_k\right)dx+\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}|\alpha _{i,j}||\beta _j|\int _{\mathbb{R}^2}\left(u^3_i+v^3_j\right)dx\\[5pt] &\leq \frac{\alpha ^*\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )|\mathcal{J}|}{2}\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}u^3_idx+\frac{\alpha ^*\beta ^*|\mathcal{I}|A^3_3}{2}\\[5pt] &\leq \frac{c_1\alpha ^*\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )|\mathcal{J}|}{2}\sum \limits ^n_{i=1}\eta _i\int _{\mathbb{R}^2}|\nabla u_i|^2dx\\[5pt] &\qquad +\frac{\alpha ^*\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )|\mathcal{J}|}{2}\sum \limits ^n_{i=1}c_{2i}m_i+\frac{\alpha ^*\beta ^*|\mathcal{I}|A^3_3}{2}, \end{split} \end{align*}

where we have used Young’s inequality and the following facts $\sum \limits ^m_{j=1}\|v_j\|_{L^3(\mathbb{R}^2)}\leq A_3$ hold due to Lemma 2.1 $(v)$ , and for any $\eta _i\in (0,1)$ , $i\in \mathcal{I}$ ,

\begin{align*} \begin{split} \int _{\mathbb{R}^2}u^3_idx&\leq \eta _i\left (\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx\right )\left (\int _{\mathbb{R}^2}|\nabla u_i|^2dx\right )+c_{2i}\int _{\mathbb{R}^2}u_idx\\[5pt] &\leq c_1\eta _i \int _{\mathbb{R}^2}|\nabla u_i|^2dx+c_{2i}m_i,\quad i\in \mathcal{I}, \end{split} \end{align*}

exists with $c_{2i}=c_{2i}(\eta _i)\gt 0$ from (2.4). As for the rightmost integral of (3.21), we first use Hölder’s inequality to find that

\begin{align*} -\frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u^2_i\partial _tv_jdx\leq \frac{1}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}|\alpha _{i,j}|\|u_i\|^2_{L^4(\mathbb{R}^2)}\|\partial _tv_j\|_{L^2(\mathbb{R}^2)}. \end{align*}

Applying the Gagliardo–Nirenberg inequality with $c_3\gt 0$ to have

\begin{align*} \|u_i\|^2_{L^4(\mathbb{R}^2)}\leq c_3\|\nabla u_i\|_{L^2(\mathbb{R}^2)}\| u_i\|_{L^2(\mathbb{R}^2)}, \end{align*}

and it infers that

\begin{align*} \begin{split} -\frac{1}{2}&\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}u^2_i\partial _tv_jdx\\[5pt] \leq &\frac{c_3}{2}\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}|\alpha _{i,j}|\|\nabla u_i\|_{L^2(\mathbb{R}^2)}\| u_i\|_{L^2(\mathbb{R}^2)}\|\partial _tv_j\|_{L^2(\mathbb{R}^2)}\\[5pt] \leq &\frac{c_3\alpha ^*|\mathcal{J}|}{2}\sum \limits ^n_{i=1}\eta _i\|\nabla u_i\|^2_{L^2(\mathbb{R}^2)}+\frac{c_3\alpha ^*}{8}\left (\sum \limits ^n_{i=1}\frac{1}{\eta _i}\|u_i\|^2_{L^2(\mathbb{R}^2)}\right ) \cdot \left (\sum \limits ^m_{j=1}\|\partial _tv_j\|^2_{L^2(\mathbb{R}^2)}\right ) \end{split} \end{align*}

by Young’s inequality. Hence, (3.21) gives us that

(3.22) \begin{equation} \begin{split} &\frac{d}{dt}\left (\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}u^2_idx\right )+\sum \limits ^n_{i=1}\left \{2-\alpha ^*\left [c_1\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )+c_3\right ]|\mathcal{J}|\eta _i\right \}\int _{\mathbb{R}^2}|\nabla u_i|^2dx\\[5pt] &\leq \frac{c_3\alpha ^*}{4}\left (\sum \limits ^n_{i=1}\frac{1}{\eta _i} \|u_i\|^2_{L^2(\mathbb{R}^2)}\right )\left (\sum \limits ^m_{j=1}\|\partial _tv_j\|^2_{L^2(\mathbb{R}^2)}\right )\\[5pt] &\quad +\alpha ^*\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )|\mathcal{J}|\sum \limits ^n_{i=1}c_{2i}m_i+\alpha ^*\beta ^*|\mathcal{I}|A^3_3\quad \text{for}\,\,t\in (0,T). \end{split} \end{equation}

Because

\begin{align*} \|u_i\|^2_{L^2(\mathbb{R}^2)}&\leq c_4\|\nabla u_i\|_{L^2(\mathbb{R}^2)}\| u_i\|_{L^1(\mathbb{R}^2)}=c_4m_i\|\nabla u_i\|_{L^2(\mathbb{R}^2)}\\[5pt]& \leq \|\nabla u_i\|^2_{L^2(\mathbb{R}^2)}+\frac{c^2_4m^2_i}{4} \end{align*}

is right for some $c_4\gt 0$ , we take $\eta _i=1/\{\alpha ^*\left [c_1\!\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )+c_3\right ]|\mathcal{J}|\}\gt 0$ , $i\in \mathcal{I}$ , small enough in (3.22) to arrive at

\begin{align*} \begin{split} y^{\prime}(t)&+\left [1-\frac{c_3(\alpha ^*)^2\left [c_1\!\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )+c_3\right ]|\mathcal{J}|}{4}\left (\sum \limits ^m_{j=1}\|\partial _tv_j\|^2_{L^2(\mathbb{R}^2)}\right )\right ]y(t)\\[5pt] \leq\ & \alpha ^*\left (\beta ^*+2\gamma ^*|\mathcal{I}|\right )|\mathcal{J}|\sum \limits ^n_{i=1}c_{2i}m_i+\frac{c^2_4}{4}\sum \limits ^n_{i=1}m^2_i+\alpha ^*\beta ^*|\mathcal{I}|A^3_3 \end{split} \end{align*}

for $t\in (0,T)$ , where $y(t)\,:\!=\,\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}u^2_idx$ . Together with (3.20), the $L^2$ estimates for the solutions can be obtained by solving this ODE.

Proof of Theorem 1.1. Let $0\lt T\leq \infty$ . Once one has $L^2$ estimates on $u_i$ , $i\in \mathcal{I}$ , then $L^p$ - $L^q$ estimates for the heat semigroup in Lemma 5.1 ensure that for $r\in (1,\infty )$ , we have

(3.23) \begin{equation} \begin{split} \sum \limits ^m_{j=1}\|\nabla v_j(\cdot,t)\|_{L^{r}(\mathbb{R}^2)}\leq c\quad \text{for}\quad t\in (0,T) \end{split} \end{equation}

with some $c\gt 0$ . Consequently, applying the Moser iteration technique [Reference Alikakos1] with (3.23), it means that

\begin{align*} \sum \limits ^n_{i=1}\|u_i(\cdot,t)\|_{L^{\infty }(\mathbb{R}^2)}\lt \infty \quad \text{for}\quad t\in (0,T). \end{align*}

Therefore, the global result in Theorem 1.1 is an immediate consequence of the extensibility criterion in Lemma 2.1.

4. Proof of Theorem 1.2

Under strong self-repelling effect, the global existence of the Cauchy problem (1.5) with arbitrary initial data will be established in this section. As with the above treatment of the proof of Theorem 1.1, the main approach is to give a bound on the modified total entropy $\mathcal{S}$ . For this purpose, we would like to create a differential inequality for $\mathcal{G}$ .

Lemma 4.1. Assume that there exist positive constants $a_1,\cdots,a_n$ and a positive definite matrix $\boldsymbol{{B}}=(b_{j,l})_{m\times m}$ such that

(4.1) \begin{equation} \boldsymbol{{B}}\boldsymbol{\gamma }_i=-a_i\boldsymbol{\alpha }_i,\quad \forall \,\,i\in \mathcal{I}. \end{equation}

Let

\begin{align*} D_1\boldsymbol{\xi }^T\boldsymbol{\xi }\leq \boldsymbol{\xi }^T\boldsymbol{{B}}\boldsymbol{\xi }\leq D_2\boldsymbol{\xi }^T\boldsymbol{\xi },\quad \forall \,\, \boldsymbol{\xi }=(\xi _1,\cdots,\xi _m)^T\in \mathbb{R}^m, \end{align*}

with some $D_1,D_2\gt 0$ . Then, there exist a constant $c_1=c_1(a^*,b^*,\alpha ^*,\beta ^*,A_2,D_1,D_2,\|\boldsymbol{{u}}_0\|_{L^1(\mathbb{R}^2)},|\mathcal{I}|$ , $|\mathcal{J}|)\gt 0$ such that

\begin{align*} \frac{d}{dt} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]+\frac{1}{2}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx\leq & c_1\quad \text{for}\quad t\in (0,T), \end{align*}

where $\mathcal{G}$ is given by

\begin{align*} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]=&\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _{l}v_jv_l\right )dx. \end{align*}

Moreover, if $\beta _j\gt 0$ for all $j\in \mathcal{J}$ , then there exist a constant $c_2\gt 0$ independent of $T$ such that

\begin{align*} \frac{d}{dt} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]+\frac{D_1}{D_2}\beta _* \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]\leq c_2\quad \text{for}\quad t\in (0,T). \end{align*}

Proof. Given $a_i\gt 0$ , testing the $i$ -th equation in (1.5) by $a_i\log\! (u_i+1)$ and summing the results with respect to $i$ , we get

(4.2) \begin{equation} \begin{split} &\frac{d}{dt}\left [\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx\right ] +\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx\\[5pt] &=\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\frac{u_i}{u_i+1}\nabla u_i\cdot \nabla v_jdx\\[5pt] &=\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\nabla u_i\cdot \nabla v_jdx+\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\Delta v_j\log\! (u_i+1)dx\quad \text{for}\quad t\in (0,T). \end{split} \end{equation}

Moreover, we observe that

(4.3) \begin{equation*} \begin{split} \frac{1}{2}&\frac{d}{dt}\left [\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _lv_jv_l\right )\right ]dx\\[5pt] =&-\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}(\Delta v_l-\beta _lv_l)\partial _tv_jdx+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _l\int _{\mathbb{R}^2}(v_j\partial _tv_l-v_l\partial _tv_j)dx \\ =&-\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\Delta v_j-\beta _jv_j\right )\left (\Delta v_l-\beta _lv_l\right )dx\\[5pt] &+\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\gamma _{i,j}\int _{\mathbb{R}^2}\nabla u_i\cdot \nabla v_ldx+\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\gamma _{i,j}\beta _l\int _{\mathbb{R}^2}u_iv_ldx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _l\int _{\mathbb{R}^2}(v_j\partial _tv_l-v_l\partial _tv_j)dx\quad \text{for}\quad t\in (0,T) \end{split} \end{equation*}

and that

(4.4) \begin{equation} \begin{split} D_1\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\Delta v_j-\beta _jv_j\right |^2dx\leq & \sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\Delta v_j-\beta _jv_j\right )\left (\Delta v_l-\beta _lv_l\right )dx\\[5pt] \leq &D_2\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\Delta v_j-\beta _jv_j\right |^2dx. \end{split} \end{equation}

Then employing $\sum \limits ^m_{j=1}b_{j,l}\gamma _{i,j}=-a_i\alpha _{i,l}$ by (4.1), it is obvious that

(4.5) \begin{equation} \begin{split} \frac{d}{dt}&\mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]+\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx+D_1\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\Delta v_j-\beta _jv_j\right |^2dx\\[5pt] \leq &\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\Delta v_j\log\! (u_i+1)dx-\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\beta _j\int _{\mathbb{R}^2}u_iv_jdx\\[5pt] &+\frac{1}{2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _l\int _{\mathbb{R}^2}(v_j\partial _tv_l-v_l\partial _tv_j)dx\quad \text{for}\quad t\in (0,T) \end{split} \end{equation}

due to (4.2) and (4.3). Since $\log\! (u_i+1)\leq \sqrt{u}_i$ , $\|u_i\|_{L^1(\mathbb{R}^2)}=m_i$ , $i\in \mathcal{I}$ , and

\begin{align*} \sum \limits ^m_{j=1}\|v_j\|_{L^2(\mathbb{R}^2)}\leq A_2 \end{align*}

hold out, an application of Young’s inequality gives that

(4.6) \begin{equation} \begin{split} &\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}\Delta v_j\log\! (u_i+1)dx\\[5pt] &=\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\int _{\mathbb{R}^2}(\Delta v_j-\beta _jv_j)\log\! (u_i+1)dx+\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\beta _j\int _{\mathbb{R}^2}v_j\log\! (u_i+1)dx\\[5pt] &\leq \sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i|\alpha _{i,j}|\int _{\mathbb{R}^2}|\Delta v_j-\beta _jv_j|\sqrt{u_i}dx+\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i|\alpha _{i,j}||\beta _j|\int _{\mathbb{R}^2}|v_j|\sqrt{u_i}dx\\[5pt] &\leq \frac{D_1}{4}\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\Delta v_j-\beta _jv_j\right |^2dx+\frac{(a^*\alpha ^*)^2|\mathcal{I}||\mathcal{J}|}{D_1}\sum \limits ^n_{i=1}m_i+a^*\alpha ^*\beta ^*A_2\sum \limits ^n_{i=1}m^{1/2}_i. \end{split} \end{equation}

Note that $A_2$ is uniformly bounded with respect to time variable if $\beta _j\gt 0$ for all $j\in \mathcal{J}$ . However, Lemma 2.2 tells us

\begin{align*} \int _{\mathbb{R}^2}u^2_idx&\leq \frac{1}{2\pi }\left (\int _{\mathbb{R}^2}u_idx\right )\left (\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx\right )+2\int _{\mathbb{R}^2}u_idx. \end{align*}

Then, one has

(4.7) \begin{equation} \begin{split} -\sum \limits ^n_{i=1}\sum \limits ^{m}_{j=1}a_i\alpha _{i,j}\beta _j\int _{\mathbb{R}^2}u_iv_jdx\leq & \sum \limits ^n_{i=1}\frac{\pi a_i}{2m_i}\int _{\mathbb{R}^2}u^2_idx+\frac{(\alpha ^*\beta ^*)^2|\mathcal{J}|}{2\pi }\left (\sum \limits ^n_{i=1}a_im_i\right )\left (\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}v^2_jdx\right )\\[5pt] \leq &\frac{1}{4}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx+\pi \sum \limits ^n_{i=1}a_i+\frac{(\alpha ^*\beta ^*A_2|\mathcal{J}|)^2}{2\pi }\left (\sum \limits ^n_{i=1}a_im_i\right ) \end{split} \end{equation}

and

(4.8) \begin{align} \frac{1}{2}&\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\beta _l\int _{\mathbb{R}^2}(v_j\partial _tv_l-v_l\partial _tv_j)dx\nonumber\\[5pt] \leq &\beta ^*\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}|b_{j,l}|\int _{\mathbb{R}^2}|v_j||\partial _tv_l|dx\nonumber\\[5pt] \leq & \beta ^*\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}|b_{j,l}|\int _{\mathbb{R}^2}|v_j||\Delta v_l-\beta _lv_l|dx+\beta ^*\sum \limits ^m_{j=1}\sum \limits ^m_{l=1}|b_{j,l}|\int _{\mathbb{R}^2}|v_j|\left |\sum \limits ^n_{i=1}\alpha _{i,l}u_i\right |dx\\[5pt] \leq &\frac{1}{4}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx+\frac{D_1}{4}\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\Delta v_j-\beta _jv_j\right |^2dx\nonumber\\[5pt] &+\frac{\left(\alpha ^*b^*\beta ^*A_2|\mathcal{J}|^2\right)^2}{2\pi }\sum \limits ^n_{i=1}\frac{m_i}{a_i}+\pi \sum \limits ^n_{i=1}a_i+\frac{(b^*\beta ^*A_2|\mathcal{J}|)^2}{D_1}\quad \text{for}\quad t\in (0,T).\nonumber \end{align}

Putting (4.5)–(4.8) together, then there exists a constant $c_1\gt 0$ such that $\mathcal{G}$ satisfies

\begin{align*} \frac{d}{dt} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]+\frac{1}{2}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx+\frac{D_1}{2}\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\Delta v_j-\beta _jv_j\right |^2dx\leq & c_1\quad \text{for}\quad t\in (0,T). \end{align*}

Now if $\beta _j\gt 0$ for all $j\in \mathcal{J}$ , combining (2.2) with (4.4) yields that

\begin{align*} \begin{split} \frac{D_1}{2}\sum \limits ^{m}_{j=1}\int _{\mathbb{R}^2}\left |\Delta v_j-\beta _jv_j\right |^2dx\,\geq\, &\frac{D_1}{2D_2}\sum \limits ^{m}_{j=1}\sum \limits ^{m}_{l=1}b_{j,l}\int _{\mathbb{R}^2}(\Delta v_j-\beta _jv_j)(\Delta v_l-\beta _lv_l)dx\\[5pt] \geq\, &\frac{D_1\beta _*}{2D_2}\sum \limits ^{m}_{j=1}\sum \limits ^m_{l=1}b_{j,l}\int _{\mathbb{R}^2}\left (\nabla v_j\cdot \nabla v_l+\beta _lv_jv_l\right )dx\\[5pt] =\, &\frac{D_1\beta _*}{D_2} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]- \frac{D_1\beta _*}{D_2}\sum \limits ^n_{i=1}a_i\int _{\mathbb{R}^2}(u_i+1)\log\! (u_i+1)dx\\[5pt] \geq\, &\frac{D_1\beta _*}{D_2} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]-\frac{D_1\beta _*}{D_2}\sum \limits ^n_{i=1}a_im_i\eta _i\int _{\mathbb{R}^2}\frac{|\nabla u_i|^2}{u_i+1}dx-c_2\beta _*\sum \limits ^n_{i=1}a_im_i, \end{split} \end{align*}

where $\eta _i\in (0,1)$ , $i\in \mathcal{I}$ and $c_2\gt 0$ are constants. Taking $\eta _i=\frac{D_2}{2D_1m_i\beta _*}$ , $i\in \mathcal{I}$ , we collect the above two inequalities to obtain

\begin{align*} \frac{d}{dt} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]+\frac{D_1\beta _*}{D_2} \mathcal{G}[\boldsymbol{{u}},\boldsymbol{{v}}]\leq c_1+c_2\beta _*\sum \limits ^n_{i=1}a_im_i\quad \text{for}\quad t\in (0,T). \end{align*}

Proof of Theorem 1.2. Thanks to Lemma 4.1, we conclude that $ \mathcal{G}$ is bounded. Therefore, there exists a constant $c\gt 0$ such that

\begin{align*} & \sum \limits ^n_{i=1}\|(u(\cdot,t)+1)\log\! (u(\cdot,t)+1)\|_{L^1(\mathbb{R}^2)}\leq c\quad \text{for}\quad t\in (0,T),\\& \sum \limits ^m_{j=1}\|\nabla v_j(\cdot,t)\|_{L^2(\mathbb{R}^2)}\leq c\quad \text{for}\quad t\in (0,T). \end{align*}

Similar to the proof of Theorem 1.1, the global result can be obtained through a classic and standard method.

Acknowledgements

The author thanks the anonymous referee very much for carefully reading our manuscript and giving positive and valuable comments.

Funding

K. Lin is supported by Natural Science Foundation of Sichuan Province (No. 2022NSFSC1837) and Guanghua Talent Project of Southwestern University of Finance and Economics.

Competing interest

None

Appendix

In this section, we will prove Lemma 2.1. The following lemma collects some basic facts on the asymptotics of the heat semigroup $(e^{t\Delta })_{t\geq 0}$ , given by

\begin{align*} (e^{t\Delta }\phi )(x)\,:\!=\,\int _{\mathbb{R}^2}G(x-y,t)\phi (y)dy,\quad x\in \mathbb{R}^2,\,\,t\gt 0, \end{align*}

where $\phi \in C^0(\mathbb{R}^2)\cap L^{\infty }(\mathbb{R}^2)$ and the Gaussian heat kernel is denoted by $G(z,t)\,:\!=\,(4\pi t)^{-1}e^{-\frac{|z|^2}{4t}}$ , $z\in \mathbb{R}^2$ , $t\gt 0$ .

Lemma 5.1. Let $(e^{t\Delta })_{t\geq 0}$ be the heat semigroup in $\mathbb{R}^2$ . Then, the following properties are true.

$(i)$ Let $\omega \in \mathbb{N}^n_0$ . Then

\begin{align*} D^\omega _xe^{t\Delta }\phi =e^{t\Delta }D^{\omega }_x\phi \quad \text{for all}\,\,t\gt 0 \end{align*}

is valid for all $\phi \in C^{|\omega |}(\mathbb{R}^2)\cap W^{|\omega |,\infty }(\mathbb{R}^2)$ .

$(ii)$ If $1\leq r_1\leq r_2\leq \infty$ and $\omega \in \mathbb{N}^n_0$ , then there exist a constant $c(r_1,r_2,|\omega |)\gt 0$ such that

\begin{align*} \|D^\omega _xe^{t\Delta }\phi \|_{L^{r_2}(\mathbb{R}^2)}\leq c(r_1,r_2,\omega )t^{-\frac{|\omega |}{2}-(\frac{1}{r_1}-\frac{1}{r_2})}\|\phi \|_{L^{r_1}(\mathbb{R}^2)}\quad \text{for all}\,\,t\gt 0 \end{align*}

holds for all $\phi \in L^{r_1}(\mathbb{R}^2)$ . In particular, $c(r_1,r_2,|\omega |)=1$ if $|\omega |=0$ and $r_1=r_2$ .

Proof. Please see [Reference Winkler41, Lemma 2.1] and [Reference Giga, Giga and Saal14] for details.

Proof of Lemma 2.1 . The proof of Lemma 2.1 will be divided into several steps.

First Step: local existence. The contraction mapping theorem will be used to prove the local existence of mild solutions. Let

\begin{align*} R\,:\!=\,\sum \limits ^n_{i=1}\left (\|u_{i0}\|_{L^{\infty }(\mathbb{R}^2)}+\|u_{i0}\|_{L^1(\mathbb{R}^2)}\right )+\sum \limits ^m_{j=1}\left (\|v_{j0}\|_{L^{1}(\mathbb{R}^2)}+\|\nabla v_{j0}\|_{L^{p}(\mathbb{R}^2)}+\|\nabla v_{j0}\|_{L^{1}(\mathbb{R}^2)}\right )+1, \end{align*}

and let $T$ be a fixed positive number below. Set

\begin{align*} X\,:\!=\,C^0\left ([0,T];\left (C^0(\mathbb{R}^2)\cap L^1(\mathbb{R}^2)\cap L^{\infty }(\mathbb{R}^2)\right )^n\times \left (W^{1,p}(\mathbb{R}^2)\cap W^{1,1}(\mathbb{R}^2) \right )^m\right ) \end{align*}

equipped with the norm

\begin{align*} \begin{split} \|(\boldsymbol{{u}},\boldsymbol{{v}})\|_{X}=\max \limits _{0\leq t\leq T}\Big \{&\sum \limits ^n_{i=1}\left [\|u_i(\cdot,t)\|_{L^{\infty }(\mathbb{R}^2)}+\|u_i(\cdot,t)\|_{L^1(\mathbb{R}^2)}\right ]\\[5pt] &+\sum \limits ^m_{j=1}\left [\|v_j(\cdot,t)\|_{L^{1}(\mathbb{R}^2)}+\|\nabla v_j(\cdot,t)\|_{L^{p}(\mathbb{R}^2)}+\|\nabla v_j(\cdot,t)\|_{L^{1}(\mathbb{R}^2)}\right ]\Big \} \end{split} \end{align*}

for $t\in [0,T]$ . Moreover, define

\begin{align*} E\,:\!=\,\left \{(\boldsymbol{{u}},\boldsymbol{{v}})\in X\Big |(\boldsymbol{{u}},\boldsymbol{{v}})(\cdot,0)=(\boldsymbol{{u}}_0,\boldsymbol{{v}}_0)\quad \text{and}\quad \|(\boldsymbol{{u}},\boldsymbol{{v}})\|_{X}\leq R\right \}. \end{align*}

Then, it is easy to see that $E$ is a closed convex subset of $X$ . Consider a nonlinear mapping $\Pi \,:\,E\mapsto X$ such that for any $(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\in E$ ,

\begin{align*} (\boldsymbol{{u}},\boldsymbol{{v}})=\Pi (\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}}), \end{align*}

where $\boldsymbol{{u}}=(u_1,\cdots,u_n)$ and $\boldsymbol{{v}}=(v_1,\cdots,v_m)$ satisfy

(A1) \begin{equation}{} \begin{split} u_i(\cdot,t)=e^{t\Delta }u_{i0}-\sum \limits ^m_{j=1}\alpha _{i,j}\int ^t_0\nabla \cdot e^{(t-s)\Delta }[\widetilde{u}_i(\cdot,s)\nabla \widetilde{v}_j(\cdot,s)]ds,\quad i\in \mathcal{I},\quad t\in [0,T], \end{split} \end{equation}

and

(A2) \begin{equation}{} \begin{split} v_j(\cdot,t)=e^{t(\Delta -\beta _j)}v_{j0}+\sum \limits ^n_{i=1}\gamma _{i,j}\int ^t_0e^{(t-s)(\Delta -\beta _j)}\widetilde{u}_i(\cdot,s)ds,\quad j\in \mathcal{J},\quad t\in [0,T], \end{split} \end{equation}

respectively. By the estimates for the heat semigroup in Lemma 5.1 $(ii)$ to (A1), there exists a constant $c_1=c_1(p)\gt 0$ such that

\begin{align*} \begin{split} \sum \limits ^n_{i=1}\|u_i(\cdot,t)\|_{L^{\infty }(\mathbb{R}^2)}\leq &\sum \limits ^n_{i=1}\|e^{t\Delta }u_{i0}\|_{L^{\infty }(\mathbb{R}^2)}+c_1\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}|\alpha _{i,j}|\int ^t_0(t-s)^{-\frac{1}{2}-\frac{1}{p}} \|\widetilde{u}_i(\cdot,s)\nabla \widetilde{v}_j(\cdot,s)\|_{L^p({\mathbb{R}^2})}ds,\\[5pt] \leq &\sum \limits ^n_{i=1}\|u_{i0}\|_{L^{\infty }(\mathbb{R}^2)} +c_1\alpha ^*\int ^t_0(t-s)^{-\frac{1}{2}-\frac{1}{p}} \left (\sum \limits ^n_{i=1}\|\widetilde{u}_i(\cdot,s)\|_{L^\infty ({\mathbb{R}^2})}\right )\\[5pt] &\cdot \left (\sum \limits ^m_{j=1}\|\nabla \widetilde{v}_j(\cdot,s)\|_{L^p({\mathbb{R}^2})}\right )ds\\[5pt] \leq &\sum \limits ^n_{i=1}\|u_{i0}\|_{L^{\infty }(\mathbb{R}^2)}+\frac{2c_1p\alpha ^*R^2}{p-2}T^{\frac{1}{2}-\frac{1}{p}},\quad t\in [0,T]. \end{split} \end{align*}

Similarly, one also has

\begin{align*} \begin{split} \sum \limits ^n_{i=1}\|u_i(\cdot,t)\|_{L^{1}(\mathbb{R}^2)} \leq &\sum \limits ^n_{i=1}\|u_{i0}\|_{L^{1}(\mathbb{R}^2)} +c_2\alpha ^*\int ^t_0(t-s)^{-\frac{1}{2}} \left (\sum \limits ^n_{i=1}\|\widetilde{u}_i(\cdot,s)\|_{L^\infty ({\mathbb{R}^2})}\right )\\[1pt] &\cdot \left (\sum \limits ^m_{j=1}\|\nabla \widetilde{v}_j(\cdot,s)\|_{L^1({\mathbb{R}^2})}\right )ds\\[1pt] \leq &\sum \limits ^n_{i=1}\|u_{i0}\|_{L^{1}(\mathbb{R}^2)}+2c_2\alpha ^*R^2T^{\frac{1}{2}},\quad t\in [0,T], \end{split} \end{align*}

with some $c_2\gt 0$ . Moreover, one can apply Lemma 5.1 $(ii)$ to find constants $c_3$ , $c_4\gt 0$ such that

\begin{align*} \begin{split} \sum \limits ^m_{j=1}\|v_j(\cdot,t)\|_{L^{1}(\mathbb{R}^2)} \leq &\sum \limits ^m_{j=1}e^{|\beta _j|t}\|v_{j0}\|_{L^{1}(\mathbb{R}^2)} +\gamma ^*\sum \limits ^m_{j=1}e^{|\beta _j|t}\int ^t_0 \left \|\sum \limits ^n_{i=1}\widetilde{u}_i(\cdot,s)\right \|_{L^1({\mathbb{R}^2})}ds\\[1pt] \leq\, &e^{\beta ^*T}\sum \limits ^m_{j=1}\|v_{j0}\|_{L^{1}(\mathbb{R}^2)}+\gamma ^*e^{\beta ^*T}R|\mathcal{J}|T,\quad t\in [0,T], \end{split} \end{align*}
(A3) \begin{equation}{} \begin{split} \sum \limits ^m_{j=1}\|\nabla v_j(\cdot,t)\|_{L^{p}(\mathbb{R}^2)}\leq & \sum \limits ^m_{j=1} e^{|\beta _j|t}\|\nabla v_{j0}\|_{L^{p}(\mathbb{R}^2)}+c_3\sum \limits ^n_{i=1}\sum \limits ^m_{j=1}|\gamma _{i,j}|e^{|\beta _j|t}\int ^t_0(t-s)^{-\frac{1}{2}}\|\widetilde{u}_i(\cdot,s)\|_{L^p(\mathbb{R}^2)}ds\\[5pt] \leq\, &e^{\beta ^*T}\sum \limits ^m_{j=1}\|\nabla v_{j0}\|_{L^{p}(\mathbb{R}^2)}+c_3\gamma ^*e^{\beta ^*t}|\mathcal{J}|\sum \limits ^n_{i=1}\int ^t_0(t-s)^{-\frac{1}{2}}\\[5pt] &\cdot \|\widetilde{u}_i(\cdot,s)\|^{\frac{p-1}{p}}_{L^{\infty }(\mathbb{R}^2)}\|\widetilde{u}_i(\cdot,s)\|^{\frac{1}{p}}_{L^{1}(\mathbb{R}^2)}ds\\[5pt] \leq\, &e^{\beta ^*T}\sum \limits ^m_{j=1}\|\nabla v_{j0}\|_{L^{p}(\mathbb{R}^2)}+2c_3\gamma ^*e^{\beta ^*T}R|\mathcal{J}|T^{\frac{1}{2}},\quad t\in [0,T], \end{split} \end{equation}
(A4) \begin{equation} \sum \limits ^m_{j=1}\|\nabla v_j(\cdot,t)\|_{L^{1}(\mathbb{R}^2)} \leq e^{\beta ^*T}\sum \limits ^m_{j=1}\|\nabla v_{j0}\|_{L^{1}(\mathbb{R}^2)}+2c_4\gamma ^*e^{\beta ^*T}R|\mathcal{J}|T^{\frac{1}{2}},\quad t\in [0,T]. \end{equation}

Hence, $\Pi$ maps $E$ into $E$ if we choose $T$ small enough.

We now show that the mapping is a contraction. Indeed, for $(\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})\in E$ , $(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\in E$ , we have

\begin{align*} \|\Pi (\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})-\Pi (\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\|_{X}=\max \limits _{0\leq t\leq T}\left \{I_1(t)+I_2(t)+I_3(t)+I_4(t)+I_5(t)\right \}, \end{align*}

where $I_i$ , $i=1,2,\cdots,5$ is introduced as follows,

\begin{align*} I_1(t)=&\sum \limits ^n_{i=1}\Bigg \|\int ^t_0\nabla \cdot e^{(t-s)\Delta }\Big [(\overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s))\Big (\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \overline{v}_j(\cdot,s)\Big )\\[5pt] &+\widetilde{u}_i(\cdot,s)\sum \limits ^m_{j=1}\alpha _{i,j}\Big ( \nabla \overline{v}_j(\cdot,s)-\widetilde{v}_j(\cdot,s)\Big )\Big ]ds\Bigg \|_{L^{\infty }(\mathbb{R}^2)}\\[5pt] \leq\, &c_1\sum \limits ^n_{i=1}\int ^t_0(t-s)^{-\frac{1}{2}-\frac{1}{p}}\left \|\overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s)\right \|_{L^{\infty }(\mathbb{R}^2)}\left \|\sum \limits ^m_{j=1}\alpha _{i,j} \nabla \overline{v}_j(\cdot,s)\right \|_{L^p(\mathbb{R}^2)}ds\\[5pt] &+c_1\sum \limits ^n_{i=1}\int ^t_0(t-s)^{-\frac{1}{2}-\frac{1}{p}}\left \|\widetilde{u}_i(\cdot,s)\right \|_{L^{\infty }(\mathbb{R}^2)}\left \|\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \Big ( \overline{v}_j(\cdot,s)-\widetilde{v}_j(\cdot,s)\Big )\right \|_{L^p(\mathbb{R}^2)}ds\\[5pt] \leq\, &\frac{4c_1p\alpha ^*R}{p-2}\|(\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})-(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\|_{X},\quad t\in [0,T], \end{align*}
\begin{align*} \begin{split} I_2(t)=&\sum \limits ^n_{i=1}\Bigg \|\int ^t_0\nabla \cdot e^{(t-s)\Delta }\Big [(\overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s))\Big (\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \overline{v}_j(\cdot,s)\Big )\\[5pt] &+\widetilde{u}_i(\cdot,s)\sum \limits ^m_{j=1}\alpha _{i,j}\Big ( \nabla \overline{v}_j(\cdot,s)-\widetilde{v}_j(\cdot,s)\Big )\Big ]ds\Bigg \|_{L^{1}(\mathbb{R}^2)}\\[5pt] \leq\, &c_2\sum \limits ^n_{i=1}\int ^t_0(t-s)^{-\frac{1}{2}}\left \|\overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s)\right \|_{L^{\infty }(\mathbb{R}^2)}\left \|\sum \limits ^m_{j=1}\alpha _{i,j} \nabla \overline{v}_j(\cdot,s)\right \|_{L^1(\mathbb{R}^2)}ds\\[5pt] &+c_2\sum \limits ^n_{i=1}\int ^t_0(t-s)^{-\frac{1}{2}}\left \|\widetilde{u}_i(\cdot,s)\right \|_{L^{\infty }(\mathbb{R}^2)}\left \|\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \Big ( \overline{v}_j(\cdot,s)-\widetilde{v}_j(\cdot,s)\Big )\right \|_{L^1(\mathbb{R}^2)}ds\\[5pt] \leq\, &4c_2\alpha ^*RT^{\frac{1}{2}}\|(\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})-(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\|_{X},\quad t\in [0,T], \end{split} \end{align*}
\begin{align*} \begin{split} I_3(t)=&\sum \limits ^m_{j=1}\left \|\int ^t_0 e^{(t-s)(\Delta -\beta _j)}\left (\sum \limits ^n_{i=1}\gamma _{i,j}\left (\overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s)\right )\right )ds\right \|_{L^{1}(\mathbb{R}^2)}\\[5pt] \leq\, &\gamma ^*e^{|\beta ^*|T}|\mathcal{J}|\sum \limits ^n_{i=1}\int ^t_0\left \| \overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s)\right \|_{L^{1}(\mathbb{R}^2)}ds\\[5pt] \leq\, &\gamma ^*e^{|\beta ^*|T}|\mathcal{J}|T\|(\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})-(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\|_{X},\quad t\in [0,T], \end{split} \end{align*}
\begin{align*} \begin{split} I_4(t)=&\sum \limits ^m_{j=1}\left \|\nabla \left [\int ^t_0 e^{(t-s)(\Delta -\beta _j)}\left (\sum \limits ^n_{i=1}\gamma _{i,j}\left (\overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s)\right )\right )\right ]\right \|_{L^{p}(\mathbb{R}^2)}\\[5pt] \leq\, &2c_3\gamma ^*e^{\beta ^*T}|\mathcal{J}|RT^{\frac{1}{2}}\|(\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})-(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\|_{X},\quad t\in [0,T], \end{split} \end{align*}
\begin{align*} \begin{split} I_5(t)=&\sum \limits ^m_{j=1}\left \|\nabla \left [\int ^t_0 e^{(t-s)(\Delta -\beta _j)}\left (\sum \limits ^n_{i=1}\gamma _{i,j}\left (\overline{u}_i(\cdot,s)-\widetilde{u}_i(\cdot,s)\right )\right )\right ]\right \|_{L^{1}(\mathbb{R}^2)}\\[5pt] \leq\, &2c_4\gamma ^*e^{\beta ^*T}|\mathcal{J}|RT^{\frac{1}{2}}\|(\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})-(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})\|_{X},\quad t\in [0,T]. \end{split} \end{align*}

So $\Pi$ is a contraction if $T$ is sufficiently small. Thus from Banach’s fixed point theorem, $\Pi$ has a fixed point in the sense that $(\boldsymbol{{u}},\boldsymbol{{v}})=\Pi (\boldsymbol{{u}},\boldsymbol{{v}})$ . Since the choice of above $T$ depends only on $R$ , $\alpha ^*$ , $\beta ^*$ , $\gamma ^*$ , $p$ and $|\mathcal{I}|$ , $|\mathcal{J}|$ , a standard argument implies that $(\boldsymbol{{u}},\boldsymbol{{v}})$ can be extended up to some $T_{\max }$ , and

(A5) \begin{equation}{} \begin{split} \limsup _{t\rightarrow T_{\max }}&\left\{\sum \limits ^n_{i=1}\left [\|u_i(\cdot,t)\|_{L^{\infty }(\mathbb{R}^2)}+\|u_i(\cdot,t)\|_{L^1(\mathbb{R}^2)}\right ]\right.\\[5pt] &\left. +\sum \limits ^m_{j=1}\left [\|v_j(\cdot,t)\|_{L^{1}(\mathbb{R}^2)}+\|\nabla v_j(\cdot,t)\|_{L^{p}(\mathbb{R}^2)}+\|\nabla v_j(\cdot,t)\|_{L^{1}(\mathbb{R}^2)}\right ]\right\}=\infty \end{split} \end{equation}

holds if $T_{\max }\lt \infty$ .

Second Step: Regularity. Since $e^{t\Delta }$ and $\nabla \cdot$ commute on $C^1(\mathbb{R}^2;\,\mathbb{R}^2)\cap L^1(\mathbb{R}^2;\,\mathbb{R}^2)$ , a straightforward regularity argument in [Reference Hillen, Painter and Winkler17, Lemma 3.3] which includes standard semigroup techniques and bootstrap procedure, and the parabolic Schauder estimates [Reference Ladyzenskaja, Solonnikov and Ural’ceva26] imply that $(\boldsymbol{{u}},\boldsymbol{{v}})\in \left [C^{2,1}(\mathbb{R}^2\times (0,T_{\max }))\right ]^{m+n}$ . In fact, abbreviating $F_i(x,t)=\sum \limits ^m_{j=1}\alpha _{i,j}u_i(x,t)\nabla v_j(x,t)$ for some $i\in \mathcal{I}$ and from the regularity for the mild solution, we rewritten (1.5) $_i$ as

(A6) \begin{equation}{} \begin{split} \partial _t u_i=\Delta u_i-\nabla \cdot F_i\quad \text{in}\,\,\mathbb{R}^2\times (0,T) \end{split} \end{equation}

with continuous and bounded $\mathcal{F}_i$ in $\mathbb{R}^2\times [0,T]$ . Then the Step 2 in [Reference Hillen, Painter and Winkler17, Lemma 3.3] tells that $u_i$ is a very weak solution to (A6), i.e.

\begin{align*} \begin{split} &-\int ^T_0\int _{\mathbb{R}^2}u_i\phi _t-\int _{\mathbb{R}^2}u_{i0}\phi (\cdot,0)=\int ^T_0\int _{\mathbb{R}^2}u_i\Delta \phi \\[5pt] &+\int ^T_0\int _{\mathbb{R}^2}F_i\cdot \nabla \phi \quad \text{for all}\,\,\phi \in C^{\infty }_0(\mathbb{R}^2\times [0,T)). \end{split} \end{align*}

Moreover, one can improve the regularity of very weak solution by introducing another solution $\overline{u}_i$ to the following initial boundary problem

\begin{align*} \begin{cases} \partial _t \overline{u}_i=\Delta \overline{u}_i-\nabla \cdot F_i,\quad x\in B_R,\,\,t\in (\tau,T), \\[5pt] \overline{u}_i|_{\partial B_R}=u_i, \quad t\in (\tau,T),\\[5pt] \overline{u}(x,\tau )=u(x,\tau ),\quad x\in B_R, \\[5pt] \end{cases} \end{align*}

where $0\lt \tau \lt T$ , $R\gt 0$ . Then a similar way in the Step 3 in [Reference Hillen, Painter and Winkler17, Lemma 3.3] makes sure that $\nabla u_i\in L^2_{loc}(\mathbb{R}^2\times (\tau,T))$ and

\begin{align*} \begin{split} &-\int ^{T}_\tau \int _{\mathbb{R}^2}u_i\phi _t-\int _{\mathbb{R}^2}u_{i}(\cdot,\tau )\phi (\cdot,\tau )=-\int ^T_\tau \int _{\mathbb{R}^2}\nabla u_i\cdot \nabla \phi \\[5pt] &+\int ^T_\tau \int _{\mathbb{R}^2}F_i\cdot \nabla \phi \quad \text{for all}\,\,\phi \in C^{\infty }_0(\mathbb{R}^2\times [\tau,T)). \end{split} \end{align*}

Hence, $\nabla \cdot F_i\in L^2_{loc}(\mathbb{R}^2\times [\tau,T])$ , which together with parabolic regularity theory [Reference Knútsdóttir, Pálsson and Edelstein-Keshet25] asserts that $u_i\in L^2((\tau,T);W^{2,2}_{loc}(\mathbb{R}^2))$ and $u_i\in L^p((\tau,T);W^{2,p}_{loc}(\mathbb{R}^2))$ for all $p\in (1,\infty )$ by the embedding theorem. Then invoking parabolic Schauder theory, we have $u_i\in C^{2+\gamma,1+\frac{\gamma }{2}}_{loc}(\mathbb{R}^2\times (\tau,T))$ with some $\gamma \in (0, 1)$ , and $u_i$ solves the $i$ -th equation in (1.5) classically in $\mathbb{R}^2\times (\tau,T)$ . The proof is complete due to the arbitrary choice of $\tau$ .

Third Step: Uniqueness, positivity and mass conservation. Construct a non-increasing cut-off function $h(x)\in C^{\infty }(\mathbb{R})$ to fulfil $h(x)\equiv 1$ in $({-}\infty, 0]$ and $h(x)\equiv 0$ in $[1,\infty )$ . And for $K\gt 0$ , set $\xi _K(x)\,:\!=\,h(|x|-K)$ , $x\in \mathbb{R}^2$ . Under the help of cut-off function $\xi _K$ , one can utilise localisation arguments to prove uniqueness, positivity and mass conservation of solutions to (1.5). Let us point out that such results were already obtained by Winkler in single-species case. We just describe the following main steps of the proof and refer to [Reference Winkler41, Lemmas 2.4–2.7] for more details.

Now we prove the uniqueness. Proceeding as in [Reference Winkler41, Lemma 2.4], given $T\gt 0$ and two solutions $(\overline{\boldsymbol{{u}}},\overline{\boldsymbol{{v}}})$ and $(\widetilde{\boldsymbol{{u}}},\widetilde{\boldsymbol{{v}}})$ in $\mathbb{R}^2\times (0,T)$ , we let $\boldsymbol{{w}}=\overline{\boldsymbol{{u}}}-\widetilde{\boldsymbol{{u}}}$ and $\boldsymbol{{z}}=\overline{\boldsymbol{{v}}}-\widetilde{\boldsymbol{{v}}}$ and obtain by applying straightforward procedure to (1.5) that

\begin{align*} \partial _tw_{i}=\Delta w_i-\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \cdot (w_i\nabla \overline{v}_j)-\sum \limits ^m_{j=1}\alpha _{i,j}\nabla \cdot (\widetilde{u}_i\nabla z_j),\quad i\in \mathcal{I},\quad t\in [0,T], \end{align*}
\begin{align*} \partial _tz_{j}=\Delta z_j-\beta _jz_j+\sum \limits ^n_{i=1}\gamma _{i,j}w_i,\quad j\in \mathcal{J},\quad t\in [0,T]. \end{align*}

With the help of cut-off function $\xi _K$ and Young’s inequality, we have

\begin{align*} \begin{split} \frac{1}{2}\frac{d}{dt}\int _{\mathbb{R}^2}\xi ^2_Kw^2_{i}dx\leq &-\frac{1}{4}\int _{\mathbb{R}^2}\xi ^2_K|\nabla w_i|^2dx+4\int _{\mathbb{R}^2}|\nabla \xi _K|^2w^2_idx\\[5pt] &+(\alpha ^*)^2|\mathcal{J}|\int _{\mathbb{R}^2}\xi ^2_Kw^2_i\left (\sum \limits ^m_{j=1}|\nabla \overline{v}_j|^2\right )dx\\[5pt] &+(\alpha ^*)^2|\mathcal{J}|\int _{\mathbb{R}^2}\xi ^2_K(\widetilde{u}_i)^2\left (\sum \limits ^m_{j=1}|\nabla z_j|^2\right )dx\\[5pt] &+2\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}\xi _kw^2_i\nabla \xi _K\cdot \nabla \overline{v}_jdx\\[5pt] &+2\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}\xi _K\widetilde{u}_iw_i\nabla \xi _K\cdot \nabla z_jdx \end{split} \end{align*}
\begin{align*} \begin{split} \frac{1}{2}\frac{d}{dt}\int _{\mathbb{R}^2}\xi ^2_K|\nabla z_{j}|^2dx\leq &\int _{\mathbb{R}^2}|\nabla \xi _K|^2\|\nabla z_j|^2dx+(\beta ^*+(\gamma ^*)^2)\int _{\mathbb{R}^2}\xi ^2_K|\nabla z_{j}|^2dx\\[5pt] &+\frac{|\mathcal{I}|}{4}\int _{\mathbb{R}^2}\xi ^2_K\left (\sum \limits ^n_{i=1}|\nabla w_i|^2\right )dx. \end{split} \end{align*}

By Hölder’s, Young’s and the Gagliardo–Nirenberg inequalities,

(A7) \begin{equation}{} \begin{split} (\alpha ^*)^2|\mathcal{J}|\int _{\mathbb{R}^2}\xi ^2_Kw^2_i\left (\sum \limits ^m_{j=1}|\nabla \overline{v}_j|^2\right )dx\leq & c_5(\alpha ^*)^2|\mathcal{J}|\|\xi _Kw_i\|^2_{L^{\frac{2q}{q-2}}(\mathbb{R}^2)}\\[5pt] \leq &c_5c_6(\alpha ^*)^2|\mathcal{J}|\|\nabla (\xi _Kw_i)\|^{\frac{4}{q}}_{L^2(\mathbb{R}^2)}\|\xi _Kw_i\|^{\frac{2(q-2)}{q}}_{L^2(\mathbb{R}^2)}\\[5pt] \leq &\frac{1}{8}\int _{\mathbb{R}^2}\xi ^2_K|\nabla w_i|^2dx+\frac{1}{8}\int _{\mathbb{R}^2}|\nabla \xi _K|^2w^2_idx\\[5pt] &+c_7\int _{\mathbb{R}^2}\xi ^2_Kw^2_idx, \end{split} \end{equation}
\begin{align*} (\alpha ^*)^2|\mathcal{J}|\int _{\mathbb{R}^2}\xi ^2_K(\widetilde{u}_i)^2\left (\sum \limits ^m_{j=1}|\nabla z_j|^2\right )dx \leq & c^2_8(\alpha ^*)^2|\mathcal{J}|\sum \limits ^m_{j=1}\int _{\mathbb{R}^2}\xi ^2_K|\nabla z_j|^2dx, \end{align*}

where we set $c_5=\sup _{t\in (0,T)}\sum \limits ^m_{j=1}\|\nabla v_j(\cdot,t)\|^2_{L^q(\mathbb{R}^2)}$ and $c_8=\sup _{t\in (0,T)}\sum \limits ^n_{i=1}\|\widetilde{u}_i\|_{L^{\infty }(\mathbb{R}^2)}$ . By the finiteness $\|(\nabla \overline{v}_j,\nabla z_j)\|_{L^q(\mathbb{R}^2)}$ for $j\in \mathcal{J}$ , $\|(\widetilde{u}_i, w_i)\|_{L^{\infty }(\mathbb{R}^2)}$ for $i\in \mathcal{I}$ and $supp\,\,\xi _K\subset ({-}K,K)$ , we find that $y_K(t)\,:\!=\,\frac{1}{2}\sum \limits ^n_{i=1}\int _{\mathbb{R}^2}\xi ^2_Kw^2_{i}dx+\frac{1}{4|\mathcal{I}|}\sum \limits ^m_{j=1}\int _{\mathbb{R}^2}\xi ^2_K|\nabla z_{j}|^2dx$ satisfies

\begin{align*} \begin{split} y^{\prime}_K(t)\leq & c_9y_K(t)+c_9\sum \limits ^n_{i=1}\|w_i\|^{\frac{q-1}{q}}_{L^1(B_{R+1}\backslash B_R)}+c_9\sum \limits ^n_{i=1}\|w_i\|_{L^1(B_{R+1}\backslash B_R)}\\[5pt] &+c_9\sum \limits ^m_{j=1}\|\nabla z_j\|^{\frac{q-2}{q-1}}_{L^1(B_{R+1}\backslash B_R)}\quad \text{for}\,\,t\in (0,T), \end{split} \end{align*}

where due to $y_K(0)=0$ and (2.1), an integration over $(0,T)$ shows that $y_K(t)\rightarrow 0$ as $K\rightarrow \infty$ . Hence $\overline{\boldsymbol{{u}}}=\widetilde{\boldsymbol{{u}}}$ and $\overline{\boldsymbol{{v}}}=\widetilde{\boldsymbol{{v}}}$ in $\mathbb{R}^2\times (0,T)$ .

To prove the positivity of $u_i$ , $i\in \mathcal{I}$ , it is sufficient to make sure that $u_i$ is non-negative in $\mathbb{R}^2\times (0,T)$ for each $T\in (0,T_{\max })$ by the strong maximum principle and (1.6). Denote $u^{-}_{i}=\max \{-u_i,0\}$ . A direct computation shows that

\begin{align*} \begin{split} \frac{1}{2}\frac{d}{dt}\int _{\mathbb{R}^2}\xi ^2_K(u^{-}_{i})^2dx\leq &-\frac{1}{2}\int _{\mathbb{R}^2}\xi ^2_K|\nabla u^{-}_i|^2dx+(\alpha ^*)^2|\mathcal{J}|\int _{\mathbb{R}^2}\xi ^2_K|u^{-}_i|^2\left (\sum \limits ^m_{j=1}|\nabla v_j|^2\right )dx\\[5pt] &+4\int _{\mathbb{R}^2}|u^{-}_i|^2|\nabla \xi _K|^2dx+2\sum \limits ^m_{j=1}\alpha _{i,j}\int _{\mathbb{R}^2}\xi _K(u^{-}_i)^2\nabla \xi _K\cdot \nabla v_jdx. \end{split} \end{align*}

Since $\nabla v_j\in L^{\infty }((0,T);\,L^{p}(\mathbb{R}^2))$ with $p\gt 2$ , using a similar approach in (A7) shows the existence of $c_{10}\gt 0$ such that

\begin{align*} \begin{split} (\alpha ^*)^2&|\mathcal{J}|\int _{\mathbb{R}^2}\xi ^2_K|u^{-}_i|^2\left (\sum \limits ^m_{j=1}|\nabla v_j|^2\right )dx\\[5pt] \leq & c_5(\alpha ^*)^2|\mathcal{J}|\|\xi _Ku^{-}_i\|^2_{L^{\frac{2q}{q-2}}(\mathbb{R}^2)}\\[5pt] \leq & \frac{1}{2}\int _{\mathbb{R}^2}\xi ^2_K|\nabla u^{-}_i|^2dx+\frac{1}{2}\int _{\mathbb{R}^2}|\nabla \xi _K|^2(u^{-}_i)^2dx+c_{10}\int _{\mathbb{R}^2}\xi ^2_K(u^{-}_i)^2dx. \end{split} \end{align*}

On the other hand, we follow a procedure in proving uniqueness and conclude that $g_K(t)\,:\!=\,\int _{\mathbb{R}^2}\xi ^2_K(u^{-}_{i})^2dx$ fulfils $g_K(t)\rightarrow 0$ as $K\rightarrow \infty$ . Hence non-negativity of $u_i$ , $i\in \mathcal{I}$ , has been proved.

Fourth Step: $L^q$ estimates for $v_j$ , $j\in \mathcal{J}$ . Integrating $j$ -th equation in (1.5) over $\mathbb{R}^2\times (0,T)$ directly, it results in

\begin{align*} \|v_j(\cdot,t)\|_{L^1(\mathbb{R}^2)}\leq \begin{cases} &\|v_{j0}\|_{L^1(\mathbb{R}^2)}+t\sum \limits ^n_{i=1}|\gamma _{i,j}|\|u_{i0}\|_{L^1(\mathbb{R}^2)},\quad \text{if}\quad \beta _j=0,\\[5pt] &e^{-\beta _jt}\|v_{j0}\|_{L^1(\mathbb{R}^2)}+\frac{1}{|\beta _j|}\left |1-e^{-\beta _jt}\right |\sum \limits ^n_{i=1}|\gamma _{i,j}|\|u_{i0}\|_{L^1(\mathbb{R}^2)},\quad \text{if}\quad \beta _j\neq 0. \end{cases} \end{align*}

For $q\gt 1$ , applying Lemma 5.1 to (A2) we infer that

\begin{align*} \begin{split} \|v_j(\cdot,t)\|_{L^{q}(\mathbb{R}^2)} \leq &e^{|\beta _j|t}\|v_{j0}\|_{L^{q}(\mathbb{R}^2)} +e^{|\beta _j|t}\sum \limits ^n_{i=1}|\gamma _{i,j}|\int ^t_0(t-s)^{-1+\frac{1}{q}} \left \|\widetilde{u}_i(\cdot,s)\right \|_{L^1({\mathbb{R}^2})}ds\\[5pt] \leq &e^{\beta ^*T}\|v_{j0}\|_{L^{q}(\mathbb{R}^2)}+q\gamma ^*e^{\beta ^*T}T^{\frac{1}{q}}\sum \limits ^n_{i=1}\|u_{i0}\|_{L^1(\mathbb{R}^2)},\quad j\in \mathcal{J},\quad t\in [0,T]. \end{split} \end{align*}

Hence, we have obtained $L^q$ estimates for $v_j$ , $j\in \mathcal{J}$ , and found that the upper bound is independent of time variable if $\beta _j\gt 0$ , $\forall \,\,j\in \mathcal{J}$ .

Fifth Step: Criterion. If $T_{\max }\lt \infty$ and there exists a constant $c_{11}\gt 0$ such that

\begin{align*} \sum \limits ^n_{i=1}\|u_i(\cdot,t)\|_{L^{\infty }(\mathbb{R}^2)}\leq c_{11}. \end{align*}

Then from the mass conservation $\|u_i(\cdot,t)\|_{L^1(\mathbb{R}^2)}=\|u_{i0}\|_{L^1(\mathbb{R}^2)}$ , $i\in \mathcal{I}$ , and the boundedness of $L^1$ estimate for $v_j(\cdot,t)$ , $j\in \mathcal{J}$ , for all $t\in (0,T_{\max })$ and the following fact

\begin{align*} \sum \limits ^m_{j=1}\left (\|\nabla v_j(\cdot,t)\|_{L^{p}(\mathbb{R}^2)}+\|\nabla v_j(\cdot,t)\|_{L^{1}(\mathbb{R}^2)}\right )\leq c_{12},\quad t\in (0,T_{\max }), \end{align*}

is right with some $c_{12}\gt 0$ because of (A3)–(A4), we claim that (A5) implies $T_{\max }$ cannot be finite.

References

Alikakos, N. D. (1979) $L^p$ bounds of solutions of reaction-diffusion equations. Commun. Partial Differ. Equations 4, 827868.Google Scholar
Beckner, W. (1993) Sharp Sobolev inequalities on the sphere and the Moser-Trudinger inequality. Ann. Math. 138(1), 213242.Google Scholar
Biler, P. & Nadzieja, T. (1994) Existence and nonexistence of solutions for a model of gravitational interaction of particles, I. Colloq. Math. 66(2), 319334.Google Scholar
Blanchet, A., Dolbeault, J. & Perthame, B. (2006) Two-dimensional Keller-Segel model: optimal critical mass and qualitative properties of the solutions. Electron. J. Differ. Equations 44, 32 pp. (electronic).Google Scholar
Calvez, V. & Corrias, L. (2008) The parabolic-parabolic Keller-Segel model in $\mathbb{R}^2$ . Commun. Math. Sci. 6, 417447.Google Scholar
Chang, S. Y. A. & Yang, P. C. (1988) Conformal deformation of metrics on $S^{2}$ . J. Differ. Geom. 27, 259296.Google Scholar
Conca, C., Espejo, E. & Vilches, K. (2011) Remarks on the blowup and global existence for a two species chemotactic Keller-Segel system in $\mathbb{R}^2$ . Eur. J. Appl. Math. 22, 553580.Google Scholar
Diaz, J. I., Nagai, T. & Rakotoson, J. M. (1998) Symmetrization techniques on unbounded domains: application to a chemotaxis system on $\mathbb{R}^n$ . J. Differ. Equations 145,156183.Google Scholar
Espejo, E., Stevens, A. & Suzuki, T. (2012) Simultaneous blowup and mass separation during collapse in an interacting system of chemotactic species. Differ. Integr. Equations 25, 251288.CrossRefGoogle Scholar
Espejo, E., Stevens, A. & Velázquez, J. J. L. (2009) Simultaneous finite time blow-up in a two-species model for chemotaxis. Analysis (Munich) 29(3), 317338.Google Scholar
Espejo, E., Stevens, A. & Velázquez, J. J. L. (2010) A note on nonsimultaneous blow-up for a drift-diffusion model. Differ. Integr. Equations 23, 451462.Google Scholar
Espejo, E., Vilches, K. & Conca, C. (2013) Sharp condition for blow-up and global existence in a two species chemotactic Keller-Segel system in $\mathbb{R}^2$ . Eur. J. Appl. Math. 24, 297313.Google Scholar
Espejo, E., Vilches, K. & Conca, C. (2019) A simultaneous blow-up problem arising in tumor modeling. J. Math. Biol. 79(4), 13571399.Google ScholarPubMed
Giga, M.-H., Giga, Y. & Saal, J. (2010) Nonlinear Partial Differential Equations: Asymptotic Behavior of Solutions and Self-Similar Solutions, Birkhäuser, Boston.Google Scholar
He, S. & Tadmor, E. (2021) Multi-species Patlak-Keller-Segel system. Indiana Univ. Math. J. 70, 15771624.Google Scholar
Herrero, M. A. & Velázquez, J. J. L. (1997) A blow-up mechanism for a chemotaxis model. Ann. Sc. Norm. Super 24, 633683.Google Scholar
Hillen, T., Painter, K. J. & Winkler, M. (2018) Global solvability and explicit bounds for a non-local adhesion model. Eur. J. Appl. Math. 29(4), 645684.Google Scholar
Hong, L., Wang, J. H. & Wang, W. (2022) Critical curve for a two-species chemotaxis model with two chemicals in $\mathbb{R}^2$ . Nonlinearity 35, 55305552.CrossRefGoogle Scholar
Hong, L., Wang, J. H., Yu, H. & Zhang, Y. (2019) Critical mass for a two-species chemotaxis model with two chemicals in $\mathbb{R}^2$ . Nonlinearity 32, 47624778.CrossRefGoogle Scholar
Horstmann, D. (2011) Generalizing the Keller-Segel model: Lyapunov functionals, steady state analysis, and blow-up results for multi-species chemotaxis models in the presence of attraction and repulsion between competitive interacting species. J. Nonlinear Sci. 21(2), 231270.Google Scholar
Horstmann, D. & Lucia, M. (2009) Nonlocal elliptic boundary value problems relate to chemotactic movement of mobile species. RIMS Kôkyûroku Bessatsu B 15, 3972.Google Scholar
Karmakar, D. & Wolansky, G. (2019) On Patlak-Keller-Segel system for several populations: a gradient flow approach. J. Differ. Equations 267, 74837520.Google Scholar
Karmakar, D. & Wolansky, G. (2022) On the critical mass Patlak-Keller-Segel for multi-species populations: global existence and infinite time aggregation. Indiana Univ. Math. J. 71(6), 24772527.CrossRefGoogle Scholar
Keller, E. F. & Segel, L. A. (1970) Initiation of slime mold aggregation viewed as an instability. J. Theor. Biol. 26(3), 399415.Google ScholarPubMed
Knútsdóttir, H., Pálsson, E. & Edelstein-Keshet, L. (2014) Mathematical model of macrophage-facilitated breast cancer cells invasion. J. Theor. Biol. 357, 184199.Google ScholarPubMed
Ladyzenskaja, O. A., Solonnikov, V. A. & Ural’ceva, N. N. (1968) Linear and Quasi-linear Equations of Parabolic Type, AMS, Providence, RI.Google Scholar
Lin, K. (2023) On the global existence of solutions to chemotaxis system for two populations in dimension two. Proc. R. Soc. Edinburgh Sect. A 153(6), 21062128.CrossRefGoogle Scholar
Lin, K. & Xiang, T. (2019) On global solutions and blow-up for a short-ranged chemical signaling loop. J. Nonlinear Sci. 29(2), 551591.Google Scholar
Lin, K. & Xiang, T. (2020) On boundedness, blow-up and convergence in a two-species and two-stimuli chemotaxis system with/without loop. Calc. Var. Partial Differ. 59, 135.Google Scholar
Lin, K. & Zhong, H. (2023) Global existence and blow-up for multi-species chemotaxis model in a two-dimensional bounded domain. submitted.CrossRefGoogle Scholar
Mizoguchi, N. (2013) Global existence for the Cauchy problem of the parabolic-parabolic Keller-Segel system on the plane. Calc. Var. Partial Differ. 48, 491505.Google Scholar
Moser, J. (1971) A sharp form of an inequality by N. Trudinger. Indiana Univ. Math. J. 20(11), 10771092.Google Scholar
Nagai, T. (2011) Global existence and decay estimates of solutions to a parabolic-elliptic system of drift-diffusion type in $\mathbb{R}^2$ . Differ. Integr. Equations 24, 2968.Google Scholar
Nagai, T. & Ogawa, T. (2011) Brezis-Merle inequalities and application to the global existence of the Cauchy problem of the Keller-Segel system,. Commun. Contemp. Math. 13(05), 795812.Google Scholar
Nagai, T., Seki, Y. & Yamada, T. (2021) Global existence of solutions to a parabolic attraction-repulsion chemotaxis system in $\mathbb{R}^2$ : the attractive dominant case. Nonlinear Anal. Real World Appl. 62, 103357.CrossRefGoogle Scholar
Nagai, T., Senba, T. & Yoshida, K. (1997) Application of the Trudinger-Moser inequality to a parabolic system of chemotaxis. Funkc. Ekvacioj 40, 411433.Google Scholar
Nagai, T. & Yamada, T. (2020) Boundedness of solutions to the Cauchy problem for an attraction-repulsion chemotaxis system in two-dimensional space. Rend. Istit. Mat. Univ. Trieste 52, 131149.Google Scholar
Shafrir, I. & Wolansky, G. (2005) Moser-Trudinger and logarithmic HLS inequalities for systems. J. Eur. Math. Soc. 7, 413448.CrossRefGoogle Scholar
Trudinger, N. S. (1967) On imbeddings into Orlicz spaces and some applications. J. Math. Mech. 17(5), 473483.Google Scholar
Weijer, C. (2004) Dictyostelium morphogenesis. Curr. Opin. Genet. Dev. 14(4), 392398.Google ScholarPubMed
Winkler, M. (2020) Single-point blow-up in the Cauchy problem for the higher-dimensional Keller-Segel system. Nonlinearity 33(10), 50075048.Google Scholar
Wolansky, G. (2002) Multi-components chemotactic system in the absence of conflicts. Eur. J. Appl. Math. 13(6), 641661.Google Scholar
Yamada, T. (2022) Global existence and boundedness of solutions to a parabolic attraction-repulsion chemotaxis system in $\mathbb{R}^2$ : the repulsive dominant case. J. Differ. Equations 315, 254269.Google Scholar