Hostname: page-component-f554764f5-sl7kg Total loading time: 0 Render date: 2025-04-11T23:39:23.899Z Has data issue: false hasContentIssue false

Pattern formation with jump discontinuity in a predator–prey model with Holling-II functional response

Published online by Cambridge University Press:  11 April 2025

Gaihui Guo*
Affiliation:
School of Mathematics and Data Science, Shaanxi University of Science and Technology, Xi’an, China
Xiaoyi Yang
Affiliation:
School of Mathematics and Data Science, Shaanxi University of Science and Technology, Xi’an, China
Conghui Zhang
Affiliation:
School of Science, Beijing University of Civil Engineering and Architecture, Beijing, China
Shanbing Li
Affiliation:
School of Mathematics and Statistics, Xidian University, Xian, China
*
Corresponding author: Gaihui Guo; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

This paper is focused on the existence and uniqueness of nonconstant steady states in a reaction–diffusion–ODE system, which models the predator–prey interaction with Holling-II functional response. Firstly, we aim to study the occurrence of regular stationary solutions through the application of bifurcation theory. Subsequently, by a generalized mountain pass lemma, we successfully demonstrate the existence of steady states with jump discontinuity. Furthermore, the structure of stationary solutions within a one-dimensional domain is investigated and a variety of steady-state solutions are built, which may exhibit monotonicity or symmetry. In the end, we create heterogeneous equilibrium states close to a constant equilibrium state using bifurcation theory and examine their stability.

Type
Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

In the seminal paper [Reference Turing26], Turing proposed the concept of diffusion-driven instability (DDI), which may explain the spontaneous formation of the pattern in developmental biology. Here, DDI is the spatial homogeneous instability caused by the interaction of two chemical substances with different diffusion rates. Since then, the Turing notion has become a paradigm for the pattern generation and inspired the emergence of various theoretical models, but its biological verification has remained elusive [Reference Akam1, Reference Gierer and Meinhardt10].

However, not all patterns are formed as a result of DDI. Some models incorporate a combination of a reaction-diffusion equation and an ordinary differential equation. A common example of migration involves macroalgae and herbivores, particularly since macroalgae are stationary and exist solely within the environment inhabited by herbivore species. Furthermore, the pattern occurs not just in classical reaction–diffusion systems in which all species diffuse [Reference Guo and Wang11, Reference Guo, You and Ahmed Abbakar12, Reference You and Guo14, Reference Song, Jiang, Liu and Yuan24] but also in degenerate systems in which certain species do not diffuse. The latter systems were modelled by reaction–diffusion–ODE systems, see [Reference Aronson, Tesei and Weinberger2, Reference Le, Tsujikawa and Yagi17, Reference Sherrat, Maini, Jager and Muller25]. A model consisting of free receptors, bound receptors and ligands was proposed by Sherrat et al. [Reference Sherrat, Maini, Jager and Muller25], which described the coupling of cell-localized processes with cell to cell communication via diffusion in a cell assembly. Free and bound receptors are located on the surface of the cell and therefore do not diffuse. Ligand diffuses and acts by binding itself to receptors, thereby triggering an intracellular response that leads to cell differentiation. Their model has a built-in spatial heterogeneity that triggers patterning. Marciniak–Czochra [Reference Marciniak–Czochra19, Reference Marciniak–Czochra20] later extended their model and demonstrated that nonlinear interactions of hysteresis type can result in the spontaneous emergence of the pattern, without the need for spatial heterogeneity. For a detailed mathematical analysis of their work, please refer to [Reference Harting, Marciniak–Czochra and Takagi15, Reference Kothe, Marciniak–Czochra and Takagi16, Reference Li, Marciniak–Czochra, Takagi and Wu18, Reference Marciniak–Czochra, Takagi and Nakayama21].

In order to analyse the contribution of non-diffusive components in the pattern development procedure, we concentrate on the following system

(1.1) \begin{equation} \left \{\begin{array}{lll} u_t=r\left (1-\dfrac {u}{K}\right )u-\dfrac {cuv}{m+bu}, & t\gt 0,& x\in \overline \Omega, \\[8pt] v_t=d_2\Delta v-av+\dfrac {\beta cuv}{m+bu},& t\gt 0,& x\in \Omega, \\ \partial _\tau v =0, & t\gt 0,& x\in \partial \Omega, \\ u(x,0)=u_0(x)\ge 0, \not \equiv 0 \quad v(x,0)=v_0(x)\ge 0,\not \equiv 0 & x\in {{\Omega }}, \end{array} \right . \end{equation}

where $u$ and $v$ represent the population density of the prey and predator, respectively; $d_2$ represents the predator diffusion rates; $\Omega$ is a bounded domain in the Euclidean space $R^N$ with smooth boundary, denoted as $\partial \Omega$ ; $\Delta$ is the Laplace operator in $R^N$ ; $\tau$ is the unit outer normal vector on $\partial \Omega$ . The parameters $r$ , $m$ , $c$ , $b$ , $a$ , $\beta$ , $K$ are positive constants.

The case is interesting because a scalar reaction–diffusion equation typically cannot produce stable spatially heterogeneous patterns [Reference Casten and Holland5]. While it is true that problem (1.1) does not exhibit stable Turing-type patterns, it is worth noting that DDI (Diffusion-Driven Instability) can still occur by selecting suitable parameters. Interestingly, under certain conditions, the stationary problem associated with equation (1.1) can be simplified into a boundary value problem for a single reaction–diffusion equation featuring a discontinuous nonlinearity, which leads to the emergence of positive solutions with jump discontinuity. There is a lot of work on such issues, such as [Reference Cygan, Marciniak–Czochra, Karch and Suzuki6, Reference Takagi and Zhang27, Reference Zhang and Yang32]. Hence, the focus of this paper is studying the stationary problem associated with equation (1.1).

(1.2) \begin{equation} \left \{\begin{array}{lll} r\left (1-\dfrac {u}{K}\right )u-\dfrac {cuv}{m+bu}=0,& x\in \overline \Omega, \\[8pt] d_2\Delta v-av+\dfrac {\beta cuv}{m+bu}=0,& x\in \Omega, \\ \partial _\tau v =0,& x\in \partial \Omega .\\ \end{array} \right . \end{equation}

For convenience, we let

\begin{equation*}f_1(u,v)=r\left (1-\dfrac {u}{K}\right )u-\dfrac {cuv}{m+bu},\qquad f_2(u,v)=-av+\dfrac {\beta cuv}{m+bu}.\end{equation*}

The main findings of our current work can be summarized as follows. To begin, we carefully select suitable coefficients $a$ , $c$ , $m$ , $b$ and $\beta$ in order to ensure that the kinetic system (without considering diffusion) of problem (1.1) possesses only one positive equilibrium $(u^*_2,v^*_2)$ located on the right branch, as depicted in Figure 1(b). Then, by a variational approach to bifurcation methods, we show the existence of regular stationary solutions of problem (1.2). Next, by transforming the problem into a boundary value problem for a single equation involving $v(x)$ , we establish the existence of a discontinuous solution $(u(x),v(x))$ for problem (1.2) using the generalized mountain pass lemma (Theorem4.1). This solution is characterized by a jump discontinuity in $u(x)$ and $\Delta v(x)$ . The innovation of our current research is from the presence of a discontinuous nonlinearity in the reduced problem for $v(x)$ , which results in invalidating the general mountain pass lemma introduced by Ambrosetti and Rabinowitz [Reference Ambrosetti and Rabinowitz3]. Fortunately, Chang [Reference Chang7] expanded the existing theory to handle problems involving partial differential equations that contain discontinuous nonlinearities. It appears that this approach is suitable for our specific issue, allowing us to solve challenges we faced.

Figure 1. Nullclines for $f_1(u,v)=f_2(u,v)=0$ . The blue curve represents the solution of $f_1(u,v)=0$ , while the red curve represents the solution of $f_2(u,v)=0$ . In $(a)$ , we select $a=0.4$ , $b=1$ , $m=0.3$ , $K=0.8$ , $c=1$ , $\beta =1.4$ , $r=0.3$ . In $(b)$ , we select $a=0.4$ , $b=1$ , $m=0.3$ , $K=0.8$ , $c=1$ , $\beta =0.6$ , $r=0.3$ .

We analyse problem (1.2) within the one-dimensional domain $[0,1]$ to know the structure of pattern formation. Under certain conditions on the coefficients, the equation $f_1(u,v)=0$ , where $u\geq 0$ , has three distinct branches. These branches can be represented as $u = h_0(v)\, :\!\equiv 0$ , $u = h_1(v)$ , and $u = h_2(v)$ , with the feature that $h_0(v) \lt h_1(v) \lt h_2(v)$ (see Figure 1). To begin, we select a non-negative constant $\gamma \in (0, v^*_2)$ and utilize the functions $u = h_0(v)$ and $u = h_2(v)$ in the following manner: $u = h_0(v)$ for $v\lt \gamma$ and $u = h_2(v)$ for $v\gt \gamma$ . Subsequently, the equation (1.2) is transformed into a boundary value problem for $v(x)$ , which has discontinuous nonlinearity.

Next, by considering all values of the diffusion coefficient $d_2$ , we are able to construct monotone solutions for this particular equation, and they are then used to construct symmetric solutions through the process of reflecting the monotone solutions, as described in Theorem5.1. In order to demonstrate Theorem5.1, we employ the shooting way, which was used in the research of Takagi and Zhang [Reference Takagi and Zhang28]. Furthermore, by selecting a smaller interval for $\beta$ within the range of $(0,v^*_2)$ , we can establish the uniqueness of solutions for any given $d_2$ . This is accomplished by employing another form of shooting method [Reference Mimura, Tabata and Hosono22], as demonstrated in Theorems5.2 and 5.3. Moreover, the mode of $(u(x),v(x))$ refers to the number of points at which $v''(x)$ is discontinuous. Specifically, an $n$ -mode solution $v_n(x)$ ( $n \geq 2$ ) implies that there are exactly $n$ points of discontinuity in $v''_n(x)$ . Notably, a one-mode solution $v_1(x)$ shows that $v_1(x)$ is either monotone increasing or monotone decreasing on $[0,1]$ .

Finally, with the aid of bifurcation theory [Reference Crandall and Rabinowitz8], we create nonconstant continuous stable states close to $(u^*_3, v^*_3)$ within the one-dimensional space domain of [0,1] and investigate their instability.

The paper is divided into five sections: In Section 2, we present preliminary results on nonlinear functions $f_1$ and $f_2$ that will be utilized in the subsequent sections of this paper. In Section 3, we construct regular stationary solutions of problem (1.2) utilizing the bifurcation theory. In Section 4, we prove the existence of discontinuous stationary solutions of problem (1.2). In Section 5, we not only construct steady states with jump discontinuities but also explore various types of these states, and these steady states can exhibit monotonic or symmetric behaviour. Additionally, we verify the uniqueness of these steady states under certain additional conditions. In Section 6, the investigation focuses on the stability of stationary solutions.

2. Preliminaries

We shall discuss certain properties of the functions $f_1$ and $f_2$ that will be applied in this paper.

Proposition 2.1. If $Kb\gt m$ and $u\geq 0$ hold, then $f_1(u,v)=0$ has three distinct branches: $u=h_0(v)=0$ for $v\in (\!-\infty, +\infty )$ , $u=h_1(v)$ for $v\in (rm/c,v_M)$ and $u=h_2(v)$ for $v\in (\!-\infty, v_M)$ , where

\begin{equation*}u_M=\dfrac {Kb-m}{2b},\quad v_M=\dfrac {r\left (1-\dfrac {u_M}{K}\right )(m+bu_M)}{c}\gt 0.\end{equation*}

Proof. If $f_1(u,v)=0$ , then

\begin{equation*}u=0 \quad \mbox {or} \quad r\left (1-\dfrac {u}{K}\right )-\dfrac {cv}{m+bu}=0.\end{equation*}

It is easy to obtain that $v=r\left (1-\dfrac {u}{K}\right )(m+bu)/c$ has a maximum point $(u_M,v_M)$ , where

(2.1) \begin{equation} u_M=\dfrac {Kb-m}{2b},\quad v_M=\dfrac {r\left (1-\dfrac {u_M}{K}\right )(m+bu_M)}{c}. \end{equation}

When $Kb\gt m$ , then $u_M\gt 0$ , which shows that

(2.2) \begin{equation} v=p(u)=r\left (1-\dfrac {u}{K}\right )(m+bu)/c \end{equation}

is monotone increasing in $(\!-\infty, u_M)$ , while monotone decreasing in $(u_M,+\infty )$ . As a result, for $v\in (p(0),v_M)=(rm/c,v_M)$ , $u = h_{1}(v)$ is monotone increasing with respect to $v$ , and for $v\in (\!-\infty, v_M)$ , $u=h_2(v)$ is monotone decreasing with respect to $v$ . Direct calculations provide $u_M\lt K$ . Based on the expression of $v_M$ in (2.1), we easily deduce that $v_M\gt 0$ .

Proposition 2.2. Assume that $Kb\gt m$ and $\beta c\gt ab$ hold. Then,

(i) $(u^*_2,v^*_2)$ is a positive solution of $f_1(u,v)=f_2(u,v)=0$ for $u_M\lt am/(\beta c-ab)\lt K$ and it is on the branch $u=h_2(v)$ , where $u^*_2=am/(\beta c-ab)$ and $v^*_2=r\left (1-\dfrac {u^*_2}{K}\right )(m+bu^*_2)/c$ ;

(ii) $(u^*_3,v^*_3)$ is a positive solution of $f_1(u,v)=f_2(u,v)=0$ for $0\lt am/(\beta c-ab)\lt u_M$ and it is on the branch $u=h_1(v)$ , where $u^*_3=am/(\beta c-ab)$ and $v^*_3=r\left (1-\dfrac {u^*_3}{K}\right )(m+bu^*_3)/c$ .

Proof. We omit the details because the proof is elementary.

Proposition 2.3. Assume that the Proposition 2.2 (i) holds, we have the following results.

(i) $f_2(h_0(v),v)\lt 0$ for $v\in (0,+\infty )$ and $f_2(h_2(v),v)\gt 0$ for $v\in (0,v^*_2)$ .

(ii) $\frac {d}{dv}f_2(h_0(v),v)\lt 0$ for $v\in (\!-\infty, +\infty )$ and there is a constant $\widetilde {d}\in (0,v^*_2)$ that ensures $\frac {d}{dv}f_2(h_2(v),v)\lt 0$ for $(\widetilde {d},v^*_2]$ .

Proof. (i) Obviously,

\begin{equation*}f_2(h_0(v),v)=-av\lt 0 \quad \mbox {for}\quad v\in (0,+\infty ).\end{equation*}

Then, we find that $u=h_2(v)\gt u^*_2$ for $v\in (\!-\infty, v^*_2)$ and $\beta cu/(m+bu)$ is monotone increasing with respect to $u$ . So

\begin{equation*}\dfrac {\beta ch_2(v)}{m+bh_2(v)}-a\gt \dfrac {\beta cu^*_2}{m+bu^*_2}-a=0.\end{equation*}

This shows that $f_2(h_2(v),v)\gt 0$ for $v\in (0,v^*_2)$ .

(ii) By direct calculations, we get

\begin{equation*}\frac {d}{dv}f_2(h_0(v),v)=-a\lt 0 \quad \mbox {for}\quad v\in (\!-\infty, +\infty ),\end{equation*}
\begin{equation*}\frac {d}{dv}f_2(h_2(v),v)=\left (\dfrac {\beta ch_2(v)}{m+bh_2(v)}-a\right )+v\dfrac {\beta cmh^{'}_R(v)}{(m+bh_2(v))^2}.\end{equation*}

Recall that $h^{'}_2(v)\lt 0$ for $v\in (\!-\infty, v_M)$ from the Proposition2.1 proof. Combining this with $-a+\beta ch_2(v^*_2)/(m+bh_2(v^*_2))=0,$ then we get $\frac {d}{dv}f_2(h_2(v),v)|_{v=v^*_2}\lt 0.$ By continuity, there exists a constant $\widetilde {d}\in (0,v^*_2)$ such that $\frac {d}{dv}f_2(h_2(v),v)\lt 0$ for $(\widetilde {d},v^*_2]$ .

3. Existence of regular stationary solutions

In this section, we mainly show the existence of regular stationary solutions of problem (1.2). Firstly, we review the results of [Reference Cygan, Marciniak–Czochra, Karch and Suzuki4] and express them in a form that [Reference Cygan, Marciniak–Czochra, Karch and Suzuki4] is already suitable to deal with a system consisting of PDEs and ODEs. Thus, we deal with a solution $(u,v)=(u(x),v(x))$ to the boundary value problem

(3.1) \begin{equation} \left \{\begin{array}{lll} \qquad \quad \;\;\; f(u,v)=0,& x\in \overline \Omega, \\ d_2\Delta _\tau v+g(u,v)=0,& x\in \Omega, \\ \end{array} \right . \end{equation}

with arbitrary $C^2-$ functions $f$ and $g$ , with constant $d_2 \gt 0$ , and in an open bounded domain $\Omega \subseteq R^N$ with a $C^2-$ boundary. $\Delta _\tau$ represents the Laplacian operator and Neumann boundary conditions.

Definition 1. ([Reference Cygan, Marciniak–Czochra, Karch and Suzuki4]) A solution $(u,v)=(u(x),v(x))$ of problem ( 3.1 ) is called weak if

(i) $u$ is measurable,

(ii) $v\in W^{1,2}(\Omega ),$

(iii) $g(u,v)\in (W^{1,2}(\Omega ))^*$ $($ the dual of the space $W^{1,2}(\Omega )$ $)$ ,

(iv) the equation $f(u(x),v(x))=0$ is satisfied for almost all $x\in \Omega, $

(v) the equality

\begin{equation*}-d_2\int _\Omega \nabla v(x)\cdot \nabla \zeta (x) dx+\int _\Omega g(u(x),v(x))\zeta (x)dx=0\end{equation*}

holds for all test functions $\zeta \in W^{1,2}(\Omega ).$

Definition 2. ([Reference Cygan, Marciniak–Czochra, Karch and Suzuki4]) The weak solution of problem ( 3.1 ) in the sense of Definition 1 is called a regular solution, if there is a $C^2$ -function $\theta\, :\, {R}\rightarrow {R}$ such that $u(x)=\theta (v(x))$ for all $x\in \Omega .$

Remark 1. It is easy to find that every regular solution of problem (3.1) satisfies

\begin{equation*}f(u(x),v(x))=f(\theta (v(x)),v(x))=0\quad \mbox {for all}\quad x\in \Omega, \end{equation*}

where $v=v(x)$ is a solution of the elliptic Neumann problem

(3.2) \begin{equation} d_2\Delta _\tau v+P(v)=0 \quad \mbox {for}\quad x\in \Omega \end{equation}

with $P(v)=g(\theta (v),v).$

Proposition 3.1. Assume that $N\le 6$ . Let $y\in C^2_b(R)$ satisfy $y(0)=y'(0)=0$ . There is a sequence of numbers $d_s\rightarrow d_2$ and a sequence of non-constant functions $v_s\in W^{1,2}(\Omega )$ such that $||v_s||_{W^{1,2}}\rightarrow 0$ and which satisfies the boundary value problem

(3.3) \begin{equation} d_s\Delta _\tau v_s+(\lambda _k+d_2-d_s)v_s+y(v_s)=0\quad \mbox {for}\quad x\in \Omega . \end{equation}

Proof. We prove this lemma using the Rabinowitz bifurcation theorem of the variational equation [Reference Rabinowitz23]. Then, assume that

(i) $M$ is a real Hilbert space,

(ii) $X\in C^2(M,R)$ with $X'(u)=Lu+Z(u)$ ,

(iii) $L$ is linear and $Z(u)=o(||u||)$ at $u=0,$

(iv) $\lambda$ is an isolated eigenvalue of $L$ of a finite multiplicity. If these assumptions hold, by [Reference Rabinowitz23], we know that $(\lambda, 0)\in R\times M$ is a bifurcation point of

(3.4) \begin{equation} A(\mu, v)\equiv Lv+Z(v)-\mu v=0. \end{equation}

Thus, for $||v||\neq 0$ , the solution $(\mu, v)$ of the equation (3.4) is present in each neighbourhood of $(\lambda, 0)$ . And we apply the usual Sobolev space $M=W^{1,2}(\Omega )$ with the equivalent scalar product

\begin{equation*}\left \langle u,v\right \rangle _{W^{1,2}(\Omega )}=\int _\Omega \nabla u\cdot \nabla v dx+\int _\Omega uv dx\end{equation*}

and the functional

\begin{equation*}X(v)=\dfrac {\lambda _k+d_2}{2}\int _\Omega v^2dx+\int _\Omega S(v)dx\end{equation*}

with $S(v)=\int _{0}^{v}y(s)ds$ . It is easy to find that $X\in C(W^{1,2}(\Omega ),R)$ and $X(v)$ is differentiable in the $Fr\acute {e}chet$ sense for each $v\in W^{1,2}(\Omega )$ . By simple calculation, we have

\begin{equation*}DX(v)\zeta =(\lambda _k+d_2)\int _\Omega v\zeta dx+\int _\Omega y(v)\zeta dx\end{equation*}

with $DX\in C(W^{1,2}(\Omega )$ , $\mbox {Lin}(W^{1,2}(\Omega ),R)).$ The second $Fr\acute {e}chet$ derivative at the point $v\in W^{1,2}(\Omega )$ is represented by the bilinear form

\begin{equation*}\left \langle D^2X(v)\zeta, \kappa \right \rangle =(\lambda _k+d_2)\int _\Omega \zeta \kappa dx+\int _\Omega y'(v)\zeta \kappa dx.\end{equation*}

Next, we prove that $D^2X(v)\in C\left (W^{1,2}(\Omega ),\mbox {Lin}(W^{1,2}(\Omega ),\mbox {Lin}(W^{1,2}(\Omega ),R))\right )$ . For $v_n\rightarrow v$ in $W^{1,2}(\Omega )$ and $\zeta, \kappa \in W^{1,2}(\Omega )$ , we estimate

\begin{align*}|\left \langle (D^2X(v_n)-D^2X(v))\zeta, \kappa \right \rangle |&\le \int _\Omega |y'(v_n)-y'(v)||\zeta ||\kappa |dx \\&\le ||y''||_\infty \int _\Omega |v_n-v||\zeta ||\kappa |dx\\&\le ||y''||_\infty ||v_n-v||_3||\zeta ||_3||\kappa ||_3\\&\le ||y''||_\infty ||v_n-v||_{W^{1,2}}||\zeta ||_{W^{1,2}}||\kappa ||_{W^{1,2}}.\end{align*}

The last inequality comes from the Sobolev embedding assuming $N\le 6$ .

Particularly, we have for each test function $\zeta \in W^{1,2}(\Omega )$

\begin{equation*}X'(v)(\zeta )=(\lambda _k+d_2)\int _\Omega v\zeta dx+\int _\Omega y(v)\zeta dx\equiv Lv(\zeta )+H(v)(\zeta ).\end{equation*}

Therefore, we can find that $H(v)=o(||v||_{W^{1,2}})$ as $||v||_{W^{1,2}}\rightarrow 0$ by assuming $y=y(v)$ . It is easy to find that $\mu =1$ is an isolated eigenvalue of the operator $L$ with finite multiplicity. So, we obtain

\begin{equation*}Lv(\zeta )=\left \langle v,\zeta \right \rangle _{W^{1,2}(\Omega )}\quad \mbox {for all}\quad \zeta \in W^{1,2}(\Omega ),\end{equation*}

that is

\begin{equation*}(\lambda _k+d_2)\int _\Omega v\zeta dx=\int _\Omega \nabla v\cdot \nabla \zeta dx+\int _\Omega v\zeta dx\quad \mbox {for all}\quad \zeta \in W^{1,2}(\Omega ).\end{equation*}

Obviously, we can reduce to the eigenvalue problem for $\Delta _\tau$ . Now, the property that $\lambda _k$ is an isolated eigenvalue with finite multiplicity is applied. So, by the Rabinowitz Theorem [Reference Rabinowitz23], we can find that $(1,0)$ is a bifurcation point of (3.4) which means that there is a sequence of numbers $d_s\rightarrow d_2$ and nonzero $\left \{ {v_s}\right \}\subset W^{1,2}(\Omega )$ such that $||v_s||_{W^{1,2}}\rightarrow 0$ , satisfying

\begin{equation*}Lv_s(\zeta )+Z(v_s)(\zeta )-d_s\left \langle v_s,\zeta \right \rangle _{1,2}=0\quad \mbox {for all}\quad \zeta \in W^{1,2}(\Omega ),\end{equation*}

which is equivalent to the equation satisfied by the weak solutions $v_s\in W^{1,2}(\Omega )$ to problem (3.3)

\begin{equation*}-d_s\int _\Omega \nabla v_s\cdot \nabla \zeta dx+(\lambda _k+d_2-d_s)\int _\Omega v_s\zeta dx+\int _\Omega y(v_s)\zeta dx=0\end{equation*}

for all $\zeta \in W^{1,2}(\Omega )$ .

Proposition 3.2. Assume that $N\le 6$ . Suppose that $(\overline u,\overline v)\in R^2$ is a constant solution of problem ( 1.2 ) such that $f(\overline u,\overline v)=0$ and $g(\overline u,\overline v)=0$ . We use the following notation

(3.5) \begin{equation} a_0=f_{u}(\overline u,\overline v),\quad b_0=f_{v}(\overline u,\overline v),\quad c_0=g_{u}(\overline u,\overline v),\quad d_0=g_{v}(\overline u,\overline v) \end{equation}

and assume that

(3.6) \begin{equation} a_0\neq 0 \quad \mbox {and}\quad \dfrac {1}{a_0}det{\left (\begin{array}{cc} a_0 & b_0\\ c_0 & d_0 \end{array}\right )}=d_2\lambda _k\gt 0, \end{equation}

for some $\lambda _k$ eigenvalues of $-\Delta _\tau$ . Then, there is a sequence of real numbers $d_s \rightarrow d_2$ such that the following perturbed problem

(3.7) \begin{equation} \left \{\begin{array}{ll} f(u,v)=0,& x\in \overline \Omega, \\ d_s\Delta _\tau v+(d_2-d_s)(v-\overline v)+g(u,v)=0,& x\in \Omega, \\ \end{array} \right . \end{equation}

has a non-constant regular solution.

Proof. We construct non-constant solutions to the reaction–diffusion–ODE system by Proposition3.1. In the following, we define an open ball with a radius of $\rho \gt 0$ that is centred at $\overline v$ as $B_\rho (\overline v)$ . First, we show that only a finite number of $v_s$ can be constant in Proposition3.1. If there is a constant subsequence $\left \{ {v_{l_n}}\right \}$ that satisfies the equation (3.3) such that $v_{l_n}\rightarrow 0$ , then we know that $y'(0)=-\lambda _k$ , which is obviously a contradiction.

Next, since det $f_{u}(\overline u,\overline v)\neq 0$ , for all $V\in (B_\rho (\overline v))$ , we obtain that there is a $\rho \gt 0$ and a function $\theta \in C^2(B_\rho (\overline v))$ such that $\theta (\overline v)=\overline u$ and $f(\theta (V),V)=0$ . Then, for all $V\in (B_\rho (\overline v))$ , we prove that $P(V)\equiv g(\theta (V),V)$ satisfies $P(\overline v)=0$ and $P'(\overline v)=d_2\lambda _k\gt 0.$ It is easy to find that $P(\overline v)=g(\theta (\overline v),\overline v)=g(\overline u,\overline v)=0$ . In addition, differentiating the function $P(V)=g(\theta (V),V)$ gets

(3.8) \begin{equation} P'(V)=g_{u}(\theta (V),V)\theta '(V)+g_{v}(\theta (V),V). \end{equation}

On the other hand, we differentiate the equation $f(\theta (V),V)=0$ to obtain $f_{u}(\theta (V),V)\theta '(V)+f_{v}(\theta (V),V)=0$ , or, equivalently,

(3.9) \begin{equation} \theta '(V)=-f_{u}^{-1}(\theta (V),V)f_{v}(\theta (V),V). \end{equation}

In the end, choosing $V=\overline v$ , substituting equation (3.9) into equation (3.8) and by (3.5) we have

\begin{equation*}P'(\overline v)=-g_{u}(\theta (\overline v),\overline v)f_{u}^{-1}(\theta (\overline v),\overline v)f_{v}(\theta (\overline v),\overline v)+g_{u}(\theta (\overline v),\overline v) =-c_0a_0^{-1}b_0+d_0.\end{equation*}

Notice that

(3.10) \begin{equation} \begin{array}{l} -c_0a_0^{-1}b_0+d_0=\mbox {det}{\left (\begin{array}{cc} 1 & 0\\ 0 & -c_0a_0^{-1}b_0+d_0 \end{array}\right )}=\dfrac {1}{a_0}\mbox {det}{\left (\begin{array}{cc} a_0 & 0\\ c_0 & -c_0a_0^{-1}b_0+d_0 \end{array}\right )}\\\\ \qquad \qquad \qquad \;\;=\dfrac {1}{a_0}\mbox {det}\left ({\left (\begin{array}{cc} a_0 & b_0\\ c_0 & d_0 \end{array}\right )}{\left (\begin{array}{cc} 1 & -a^{-1}_0b_0\\ 0 & 1 \end{array}\right )}\right )=\dfrac {1}{a_0}\mbox {det}{\left (\begin{array}{cc} a_0 & b_0\\ c_0 & d_0 \end{array}\right )}.\end{array} \end{equation}

So, $P'(\overline v)=d_2\lambda _k\gt 0.$

Next, $\overline P$ represents an arbitrary extension of the function $P$ to the whole line $R$ that satisfies

(3.11) \begin{equation} \overline P\in C^2_b(R)\quad \mbox {and}\quad \overline P(V)=P(V)\quad \mbox {for all}\quad V\in (B_\rho (\overline v)). \end{equation}

By Proposition3.1, we have a sequence $d_s\rightarrow d_2$ such that

(3.12) \begin{equation} d_s\Delta _\tau v_s+(d_2-d_s)(v_s-\overline v)+\overline P(v_s)=0\quad \mbox {for}\quad x\in \Omega \end{equation}

has a non-constant solution $v_s\in W^{1,2}(\Omega )$ . Indeed, It is sufficient to search for these solutions in the form $v_s=\overline v+m_s$ , where $m_s$ satisfies

(3.13) \begin{equation} d_s\Delta _\tau m_s+(\overline P'(\overline v)+d_2-d_s)m_s+ y(m_s)=0\quad \mbox {in}\quad \Omega \end{equation}

with $\overline h'(\overline v)=d_2\lambda _k$ and $y(m_s)=\overline h(\overline v+m_s)-\overline h'(\overline v)m_s$ satisfies $y\in C^2_b(R)$ , $y(0)=0$ , and $y'(0)=0$ .

Next, we can find the solutions $m_s$ of problem (3.13) by Proposition3.1. Therefore, according to the standard elliptic theory, we have $||m_s||_{W^{2,2}(\Omega )}\rightarrow 0$ . By the bootstrap arguments using the elliptic $L_p$ estimates and the Sobolev embedding theorem, we know that $||m_s||_{W^{2,q}(\Omega )}\rightarrow 0$ for $q\gt \frac {N}{2}$ and hence $||m_s||_{L^\infty (\Omega )}\rightarrow 0$ . Particularly, by (3.11), if $||m_s||_\infty \le \rho$ , we get $\overline h(v_s)=\overline h(\overline v+m_s)=h(\overline v+m_s)=h(v_s)$ . So, the nontrivial solution of problem (3.12) is $v_s=\overline v+m_s$ , where $\overline h$ is changed to $\rho$ . In the end, we define $u_s=\theta (v_s)$ to get a nontrivial regular solution of problem (3.7).

Now we apply the previous results to the specific reaction–diffusion–ODE model (1.2). So, (1.2) may be rewritten as

(3.14) \begin{equation} \left \{\begin{array}{lll} r\left (1-\dfrac {u}{K}\right )u-\dfrac {cuv}{m+bu}=0,& x\in \overline \Omega, \\ d_2\Delta _\tau v-av+\dfrac {\beta cuv}{m+bu}=0,& x\in \Omega .\\ \end{array} \right . \end{equation}

By simple calculation, it is easy to see that problem (3.14) has trivial equilibrium $(\overline u_1,\overline v_1)=(0,0)$ , semi-trivial equilibrium $(\overline u_2,\overline v_2)=(K,0)$ , and if $0\lt \dfrac {am}{\beta c-ab}$ , then $(\overline u_3,\overline v_3)=(u^*,v^*)$ , where

\begin{equation*}u^*=\dfrac {am}{\beta c-ab},\qquad v^*=\dfrac {\beta mr[K(\beta c-ab)-am]}{K(\beta c-ab)^2}=\dfrac {\beta ru^*(K-u^*)}{aK}.\end{equation*}

We always assume that $a\lt \dfrac {\beta cK}{m+bK}$ and $u^*\lt K$ .

Theorem 3.3. Assume that $2\overline u_3\gt K$ . For $a$ , $b$ , $c$ , $r$ , $m$ , $\beta$ , $K$ are all positive constants and for a discrete sequence of the diffusion coefficients $d_2\gt 0$ problem ( 3.14 ) has a regular stationary solution.

Proof. We consider a solution $(\overline u_3,\overline v_3)$ of problem (3.14) and use Proposition3.2 with the constant stationary solution $(\overline u,\overline v)=(\overline u_3,\overline v_3)$ . Since $2\overline u_3\gt K$ , $1-\dfrac {2\overline u_3}{K}\lt 0$ , that is, $a_0\lt 0$ . By simple calculation, we can find that

\begin{equation*}\mbox {det}{\left (\begin{array}{cc} a_0 & b_0\\ c_0 & d_0 \end{array}\right )}= \left |\begin{array}{cc} r-\dfrac {2r\overline u_3}{K}-\dfrac {cm\overline v_3}{(m+b\overline u_3)^2} & -\dfrac {c\overline u_3}{m+b\overline u_3}\\[8pt] \dfrac {\beta cm\overline v_3}{(m+b\overline u_3)^2} & 0 \end{array}\right |=\dfrac {\beta cm\overline v_3}{(m+b\overline u_3)^2}\cdot \dfrac {c\overline u_3}{m+b\overline u_3}\gt 0. \end{equation*}

As a result, for some eigenvalue $\lambda _k\gt 0$ , we may select $d_2\gt 0$ to satisfy equation (3.6).

4. Existence of steady states with jump discontinuous

In this section, we prove the existence of non-constant solutions of problem (1.2) by applying a generalized mountain pass lemma due to Chang [Reference Chang7, Reference Zhang31]. According to the first equation of (1.2), we obtain $u=h_0(v)$ , $u=h_1(v)$ and $u=h_2(v)$ . Applying the functions $u=h_0(v)$ and $u=h_2(v)$ , we get the following single boundary value problem for $v$ alone

(4.1) \begin{equation} \left \{\begin{array}{l} d_2\Delta v+f_2^\gamma (v)=0, \ x\in \Omega, \\ \partial _\tau v =0, \ x\in \partial \Omega, \\ \end{array} \right . \end{equation}

where

(4.2) \begin{equation} f_2^\gamma (v)= \left \{\begin{array}{l} f_2(h_0(v),v)=-av,\ v\lt \gamma, \\ f_2(h_2(v),v)=-av+\dfrac {\beta ch_2(v)v}{m+bh_2(v)},\ \gamma \lt v\leq v_M, \end{array} \right . \end{equation}

and $\gamma \in (\xi, v^*_2)$ . Note that $\xi =p(0)$ , where $p(0)$ has been defined in (2.2).

Theorem 4.1. Assume that the hypotheses of Proposition 2.2 (i) hold, then problem 4.1 has at least one classical nontrivial solution $v(x)$ so that $0\leq v(x)\leq v^*_2.$ Particularly, $v(x)$ must cross $\gamma$ .

Remark 2. By classical nontrivial solution, we mean a solution $v(x)$ of (4.1) such that $v(x)\not \equiv 0$ , $v(x)\not \equiv v^*_2$ , $v(x)\in C^1({\bar {\Omega }})$ and $\Delta v(x)$ on the $\bar {\Omega }$ has jump discontinuity.

To prove Theorem4.1, we first generalize $f_2^\gamma (v)$ to $\widetilde f_2^\gamma (v)$ , as follows

(4.3) \begin{equation} \widetilde f_2^\gamma (v)= \left \{\begin{array}{lll} f_2^\gamma (v) & \text{for}\ v\leq v_M,\\ -a(v-v_M)+v_M\left (\!-a+\dfrac {\beta ch_2(v_M)}{m+bh_2(v_M)}\right ) & \text{for}\ v\gt v_M,\\ \end{array} \right . \end{equation}

and consider

(4.4) \begin{equation} \left \{\begin{array}{lll} d_2\Delta v+\widetilde f_2^\gamma (v)=0,& x\in \Omega, \\ \partial _\tau v =0,& x\in \partial \Omega .\\ \end{array} \right . \end{equation}

Since $\widetilde f_2^\gamma (v)$ discontinues at $v=\gamma$ , we search for the solution of (4.4) in $W^{1,2}(\Omega )$ . The energy functional $J_{d_2}(v)$ connected to (4.4) is described as follows

(4.5) \begin{equation} J_{d_2}(v)=\dfrac {d_2}{2}\int _\Omega |\nabla v|^2dx-\int _\Omega G^\gamma (v)dx, \end{equation}

where $G^\gamma (v)=\int _{0}^{v} \widetilde f_2^\gamma (s) ds.$ Moreover, we endow $W^{1,2}(\Omega )$ with the norm

\begin{equation*}||v||=\left (\int _\Omega |\nabla v|^2dx+\int _\Omega v^2dx\right )^{\frac {1}{2}}.\end{equation*}

Note that $v=0$ and $v=v^*_2$ are two constant solutions of (4.4). As a result, the proof will be divided into two cases: Case 1 $J_{d_2}(0)\leq J_{d_2}(v^*_2)$ and Case 2 $J_{d_2}(0)\gt J_{d_2}(v^*_2)$ .

Firstly, we consider Case 1. Let $s=v-v^*_2$ and $Q_{d_2}(s)=J_{d_2}(s+v^*_2)-J_{d_2}(v^*_2).$ Then

(4.6) \begin{equation} Q_{d_2}(s)=\dfrac {d_2}{2}\int _\Omega |\nabla s|^2dx-\int _\Omega (G^\gamma (s+v^*_2)-G^\gamma (v^*_2))dx \end{equation}

Here, we rephrase the definitions of the generalized gradient, Palais-Smale condition (henceforth denoted by (PS)), and the generalized mountain pass lemma in the context of our problem.

Definition 3. If $Q\,:\, W^{1,2}(\Omega )\rightarrow R$ is a locally Lipschitz continuous function, then for each $\phi \in W^{1,2}(\Omega )$ , we can define the generalized directional derivative $Q^o(s;\,\phi )$ in the direction $\phi$ by

\begin{equation*}Q^o(s;\,\phi )={\mathop {\overline \lim }\limits _{u\to 0,r\downarrow 0}} \frac {Q(s+u+r\phi )-{Q_{{}}}(s+u)}{r},\end{equation*}

and the generalized gradient of $Q(s)$ at $s$ , denoted by $\partial Q(s)$ , is defined to be the subdifferential of the function $Q^o(s;\phi )$ at $0$ . That is, $\psi \in \partial Q(s)\subset (W^{1,2}(\Omega ))^*$ if and only if $\left \langle {\psi, \phi } \right \rangle \leq Q^o(s;\phi )$ for all $\phi \in W^{1,2}(\Omega )$ , where $(W^{1,2}(\Omega ))^*$ is the dual space of $W^{1,2}(\Omega )$ .

Definition 4. We say that a locally Lipschitz continuous function $Q$ satisfies the Palais–Smale condition (P.S.) if any sequence $\left \{ {s_n}\right \}\subset W^{1,2}(\Omega )$ for which $\left \{{Q(s_n)}\right \}$ is bounded and $\lambda (s_n)=\mathop {\min }\limits _{\psi \in \partial Q(s_n)}||\psi ||_ {(W^{1,2}(\Omega ))^*}\rightarrow 0$ possesses a convergent subsequence.

Theorem 4.2. ([Reference Chang7]) Let $Q(s)$ be a locally Lipschitz continuous function on $W^{1,2}(\Omega )$ which satisfies (P.S.) and assume that

(i) $Q(0)=0$ and there exist positive constants $\rho$ , $\alpha$ such that $Q\gt 0$ in $B_{\rho }\backslash \{0\}$ and $Q\gt \alpha$ on $\partial B_{\rho }$ ;

(ii) there is an $e\in W^{1,2}(\Omega )$ , $e\ne 0$ such that $Q(e)\le 0$ .

Then $Q(s)$ has a critical point. Here, $B_{\rho }=\left \{ {s\in W^{1,2}(\Omega )|\ ||s||_{W^{1,2}(\Omega )}\le \rho }\right \}$ .

Next, we prove that $Q_{d_2}(s)$ satisfies all the assumptions of Theorem4.2.

Remark 3. If $0\in \partial Q_{d_2}(s)$ , then the critical point of $Q_{d_2}$ is $s\in W^{1,2}(\Omega )$ , as stated in [7]. Once we obtain a critical point $s$ of $Q_{d_2}(s)$ , then $v=s+v^*_2$ is a critical point of $J_{d_2}(v)$ .

Proposition 4.3. Assume that hypotheses of Proposition 2.2 (i) hold. Then $Q_{d_2}(s)$ is a locally Lipschitz continuous function on $W^{1,2}(\Omega )$ .

Proof. According to the definition of $\widetilde f_2^\gamma (v)$ in (4.3), we rewrite equation (4.6) as

(4.7) \begin{equation} Q_{d_2}(s)=R^*-\int _\Omega \int _{0}^{s} (h^\gamma (w+v^*_2)-h^\gamma (v^*_2))dwdx, \end{equation}

where $R^*=\dfrac {d_2}{2}\int _\Omega |\nabla s|^2dx+\dfrac {a}{2}\int _\Omega s^2dx$ . Clearly, $R^*$ is $C^1$ on $W^{1,2}(\Omega )$ and hence locally Lipschitz continuous. On the other hand, we know that $h^\gamma (v)=0$ if $v\lt \gamma$ , $h^\gamma (v)={\beta ch_2(v)}v/({m+bh_2(v)})$ if $\gamma \lt v\leq v_M$ and $h^\gamma (v)={\beta ch_2(v_M)}v_M/({m+bh_2(v_M)})$ if $v\gt v_M$ . Thus, we find that there exist a constant $a_1$ such that $|h^\gamma (v)-h^\gamma (v^*_2)|\lt a_1$ for all $v\in R$ , which means that $|h^\gamma (s+v^*_2)-h^\gamma (v^*_2)|\lt a_1$ for all $s\in R$ . Let $H(s)=\int _{0}^{s} (h^\gamma (w+v^*_2)-h^\gamma (v^*_2))dw$ and $B(s)=\int _\Omega H(s)dx$ . Then

\begin{equation*}|H(s_1)-H(s_2)|\le |\int _{s_2}^{s_1}\left |h^\gamma (w+v^*_2)-h^\gamma (v^*_2)|dw\right |\lt a_1|s_1-s_2|;\end{equation*}

so that

(4.8) \begin{align} |B({s_1})-B({s_2})|&\lt {a_1}{\int }_\Omega |{s_1}-{s_2}|dx\le {a_1}mes{(\Omega )^{1/2}}||{s_1} - {s_2}|{|_{{L^2}(\Omega )}} \nonumber\\ &\quad \le {a_1}mes{(\Omega )^{1/2}}||{s_1}-{s_2}|{|_{W^{1,2}(\Omega )}}. \end{align}

Therefore, $B(s)$ is a locally Lipschitz continuous function on $L^2(\Omega )$ and $W^{1,2}(\Omega )$ . From this, it can be concluded that $Q_{d_2}(s)$ is locally Lipschitz continuous on $W^{1,2}(\Omega )$ .

Proposition 4.4. Assume Proposition 2.2 (i) holds. Let $\left \{ {s_n}\right \}\subset W^{1,2}(\Omega )$ be a sequence such that $\left \{{Q_{d_2}(s_n)}\right \}$ is bounded and $\lambda (s_n)=\mathop {\min }\limits _{\psi \in \partial Q_{d_2}(s_n)}||\psi ||_ {(W^{1,2}(\Omega ))^*}\rightarrow 0$ as $n\rightarrow \infty .$ Then $\left \{ {s_n}\right \}$ possesses a convergent subsequence.

Proof. By Proposition4.3, we know

\begin{equation*} Q_{d_2}(s_n)=\dfrac {d_2}{2}\int _\Omega |\nabla s_n|^2dx+\dfrac {a}{2}\int _\Omega s_n^2dx-\int _\Omega H(s_n)dx \end{equation*}
\begin{equation*}\qquad \qquad \qquad \quad \;\;\;\geq \dfrac {1}{2}\mbox {min}\left \{ d_2,a\right \}||s_n||_{W^{1,2}(\Omega )}^2-a_1\mbox {mes}{(\Omega )^{1/2}}||s_n||_{W^{1,2}(\Omega )}.\end{equation*}

Thus, $\left \{ {s_n}\right \}$ is bounded in $W^{1,2}(\Omega )$ and there is a weakly convergent subsequence $\left \{ {s_{n_i}}\right \}$ with limit $s_0$ in $W^{1,2}(\Omega ).$ Since $W^{1,2}(\Omega )\rightarrow L^2(\Omega )$ is compact, $\left \{ {s_{n_i}}\right \}$ is strongly convergent in $L^2(\Omega )$ . Recalling Proposition4.3, both $B(s)$ and $Q_{d_2}(s)$ are locally Lipschitz continuous. So, according to Definition3, the generalized gradients of $B(s)$ and $Q_{d_2}(s)$ with respect to $s$ do exist and they are denoted by $\partial B(s)$ and $\partial Q_{d_2}(s)$ , respectively. Note that

(4.9) \begin{equation} \partial Q_{d_2}(s_n)\subset \left \{ Ls_n\right \}-\partial B(s_n), \end{equation}

where $L$ is an elliptic differential operator such that $Ls=-d_2\Delta s+as.$ We have applied Propositions (4) and (3) of [Reference Chang7] to prove (4.9). Since $\lambda (s_n)\rightarrow 0$ as $n\rightarrow \infty, $ there is a sequence $\rho _{n_i}\in \partial B(s_n)$ such that

\begin{equation*}Ls_{n_i}-\rho _{n_i}\rightarrow 0 \quad \mbox {in}\quad {(W^{1,2}(\Omega ))^*}.\end{equation*}

Because of $\rho _{n_i}\in \partial B(s_n)$ , $\left \{ {\rho _{n_i}}\right \}$ is bounded in $L^2(\Omega )$ . This can be demonstrated by noting that $B(s_{n_i})$ is locally Lipschitz continuous on $L^2(\Omega )$ according to (4.8). Therefore, there is a subsequence $\left \{ {n'_i}\right \}$ of $\left \{ {{n_i}}\right \}$ , which satisfies that $\left \{ {{\rho _{n'_i}}}\right \}$ is weakly convergent to $\rho _0$ in $L^2(\Omega )$ . Therefore, it is strongly convergent in $(W^{1,2}(\Omega ))^*$ . So,

\begin{equation*}Ls_{n'_i}\rightarrow \rho _0 \quad \mbox {in}\quad {(W^{1,2}(\Omega ))^*},\end{equation*}

which shows that $s_{n'_i}\rightarrow (L)^{-1}\rho _0$ in $W^{1,2}(\Omega ).$

Proposition 4.5. Assume that Proposition 2.2 (i) holds. We get

(i) $Q_{d_2}(0)=0$ and there are constants $\rho \gt 0$ , $\alpha \gt 0$ , which satisfy that $Q_{d_2}\geq \alpha$ if $||s||_{W^{1,2}(\Omega )}=\rho$ ;

(ii) there is an $e\in W^{1,2}(\Omega )$ , $||e||_{W^{1,2}(\Omega )}\gt \rho$ , which satisfies that $Q_{d_2}(e)\le 0.$

Proof. (i) By (4.7), we know $Q_{d_2}(0)=0.$ Moreover, we can rewrite (4.7) as follows

(4.10) \begin{align} Q_{d_2}(s)&=\dfrac {d_2}{2}{\int }_\Omega |\nabla s|^2dx+\dfrac {1}{2}{\int }_\Omega (a-(h^\gamma )_v(v^*_2))s^2dx\nonumber \\ &\quad-{\int }_\Omega {\int }_{0}^{s} (h^\gamma (w+v^*_2)-h^\gamma (v^*_2)-(h^\gamma )_v(v^*_2)w)dwdx \end{align}

From the definitions of $f_2(h_2(v),v)$ in (4.2) and $h^\gamma (v)$ after (4.7), we know $f_2(h_2(v),v)=-av+h^\gamma (v),$ which implies $\partial _v f_{2}(h_2(v),v)=-a+(h^\gamma )_v(v).$ By the proof of Proposition2.3 (ii), we see $\partial _v f_{2}(h_2(v^*_2),v^*_2)\lt 0,$ so there is a constant $a_2\gt 0$ such that $a-(h^\gamma )_v(v^*_2)\gt a_2$ for all $x\in \overline {\Omega }$ .

Let $\Phi (s)=h^\gamma (s+v^*_2)-h^\gamma (v^*_2)-(h^\gamma )_v(v^*_2)s$ and $\Psi (s)=\int _{0}^{s}\Phi (w)dw.$ It is easy to see that $\Phi (s)=o(|s|)$ at $s=0$ uniformly in $x\in \overline {\Omega }$ . Thus, for any $\iota \gt 0$ , there exists a $\delta \gt 0$ such that $|\Psi (s)|\le \iota s^2$ if $|s|\le \delta$ . In addition, in the proof of Proposition4.3, we can recall the following that $|h^\gamma (w+v^*_2)-h^\gamma (v^*_2)|\lt a_1$ , which leads to the conclusion that for every $\varepsilon \in (1,(N+2)/(N-2))$ , there exists a constant $a_3\gt 0$ such that $|h^\gamma (w+v^*_2)-h^\gamma (v^*_2)|\lt a_1+a_3|w|^\varepsilon$ for all $w\in R$ . This means that there exists a constant $a_4\gt 0$ such that $|H(s)|=|\int _{0}^{s} (h^\gamma (w+v^*_2)-h^\gamma (v^*_2))dw|\le a_4|s|^{\varepsilon +1}$ for $|s|\gt \delta$ . Thanks to the Sobolev embedding theorem, we get

\begin{equation*}Q_{d_2}(s)=\dfrac {d_2}{2}\int _{|s|\gt \delta } |\nabla s|^2dx+\dfrac {a}{2}\int _{|s|\gt \delta } s^2dx-\int _{|s|\gt \delta } H(s)dx\end{equation*}
\begin{equation*}\qquad \qquad \qquad \qquad \;\;\;\;+\dfrac {d_2}{2}\int _{|s|\le \delta } |\nabla s|^2dx+\dfrac {1}{2}\int _{|s|\le \delta } (a-(h^\gamma )_v(v^*_2))s^2dx-\int _{|s|\le \delta }\Psi (s)dx\end{equation*}
\begin{equation*}\qquad \quad \geq \dfrac {d_2}{2}\int _{|s|\gt \delta } |\nabla s|^2dx+\dfrac {a}{2}\int _{|s|\gt \delta } s^2dx-c_5||s||^{\varepsilon +1}_{W^{1,2}(\Omega )}\end{equation*}
\begin{equation*}+\dfrac {d_2}{2}\int _{|s|\le \delta } |\nabla s|^2dx+(\dfrac {a_2}{2}-\iota )\int _{|s|\le \delta } s^2dx\end{equation*}
\begin{equation*}\qquad \qquad \qquad \qquad \;\;\;\geq a_6||s||^2_{W^{1,2}(\Omega )}-a_5||s||^{\varepsilon +1}_{W^{1,2}(\Omega )}=(a_6-a_5||s||^{\varepsilon -1}_{W^{1,2}(\Omega )})||s||^2_{W^{1,2} (\Omega )}\end{equation*}

with some positive constants $a_5$ and $a_6$ . Therefore, let $\rho =(a_6/2a_5)^{1/(\varepsilon -1)}$ , we can see that $Q_{d_2}(s)=0$ for $||s||_{W^{1,2}(\Omega )}\le \rho$ if and only if $s=0$ and that $Q_{d_2}(s)\geq (a_6/2)\rho ^2=\alpha$ for $||s||_{W^{1,2}(\Omega )}=\rho$ .

(ii) By selecting $e=-v^*_2$ , we derive $Q_{d_2}(e)=J_{d_2}(0)-J_{d_2}(v^*_2)\le 0$ from Case 1.

Using Theorem4.2, we find that $Q_{d_2}$ has a critical point $s(x)$ . Then $v(x)=s(x)+v^*_2$ is a critical point of $Q_{d_2}$ . Now, we examine Case 2. Similarly to Propositions4.3 and 4.4, we can show that $Q_{d_2}(v)$ is also locally Lipschitz continuous on $W^{1,2}(\Omega )$ and satisfies (PS). Hence, it is also crucial to establish the following Proposition.

Proposition 4.6. Assume that Proposition 2.2 (i) holds. We conclude that

(i) $Q_{d_2}(0)=0$ and there are constants $\rho _1\gt 0$ , $\alpha _1\gt 0$ , which satisfy that $Q_{d_2}\geq \alpha _1$ if $||v||_{W^{1,2}(\Omega )}=\rho _1$ ;

(ii) there is an $e_1\in W^{1,2}(\Omega )$ , $||e_1||_{W^{1,2}(\Omega )}\gt \rho _1$ , which satisfies that $Q_{d_2}(e_1)\le 0.$

Proof. (i) Obviously, we know $Q_{d_2}(0)=0$ by (4.7). Similarly to (4.10), (4.5) can be rewritten as follows

(4.11) \begin{equation} Q_{d_2}(v)=\dfrac {d_2}{2}\int _\Omega |\nabla v|^2dx+\dfrac {a}{2}\int _\Omega v^2dx-\int _\Omega \int _{0}^{v} h^\gamma (w)dw dx. \end{equation}

Since $h^\gamma (v)=0$ if $v\lt \gamma$ , we know that there exists a $\delta _1\in (0,\gamma )$ such that $\int _{0}^{v} h^\gamma (w)dw=0$ for $|v|\le \delta _1$ . Then, by the definition of $h^\gamma (v)$ , we obtain $| h^\gamma (w)|\lt a_7$ with a constant $a_7\gt 0$ , which implies that for every $\varepsilon \in (1,(N+2)/(N-2))$ , there exists a constant $a_8\gt 0$ such that $|h^\gamma (w)|\lt a_7+a_8|w|^\varepsilon$ for all $w\in R$ . This shows that there is a constant $a_9\gt 0$ such that $|\int _{0}^{v} h^\gamma (w)dw|\le a_9|v|^{\varepsilon +1}$ for $|v|\gt \delta _1$ . Then repeating the proof of Proposition4.5, we can show that there exist positive constants $\rho _1$ , $\alpha _1$ such that $Q_{d_2}\geq \alpha _1$ if $||v||_{W^{1,2}(\Omega )}=\rho _1$ . (ii) Let $e_1=v^*_2$ . Then it is easy to know that $Q_{d_2}(e_1)=J_{d_2}(v^*_2)-J_{d_2}(0)\lt 0$ according to Case 2.

Using Theorem4.2 again, we can show that $Q_{d_2}$ has a critical point $v(x)$ .

Proof of Theorem 4.1 . A critical point $v(x)$ of $Q_{d_2}$ has been found. In fact, $v(x)$ is a weak solution of (4.4). We can obtain that $v(x)$ is a classical solution of (4.4) by the elliptic regularity theorem [Reference Uhlenbeck29].

Now, we demonstrate that $0\le v(x)\le v^*_2$ . Because the proof for $0\le v(x)$ and $v(x)\le v^*_2$ are similar, we only show that $v(x)\le v^*_2$ . Let $v(x_1)=\mathop {\max }\limits _{x\in \overline {\Omega }}v(x)\gt v^*_2$ . If $x_1\in \Omega$ , then $d_2\Delta v|_{x=x_1}\le 0$ , which means that $\widetilde f_2^\gamma (v(x_1))\geq 0$ . But by the definition of $\widetilde f_2^\gamma (v)$ , we find $\widetilde f_2^\gamma (v(x_1))\lt 0$ , which is a contradiction. If $x_1\in {\Omega }$ , there is a ball $B_R(r_0)\subset \Omega$ centred at $r_0\in \Omega$ of radius $R$ , which satisfies $\partial \Omega \cap \overline B_R(r_0)=\left \{ {x_1}\right \}$ and $v(x)\lt v(x_1)$ in $B_R(r_0)$ . Since $v(x_1)\gt v^*_2$ , it follows from continuity that $v(x)\gt v^*_2$ in $B_R(r_0)$ , provided $r_0$ is sufficiently close to $x_1$ and $R$ is small enough. So, $\widetilde f_2^\gamma (v(x))\lt 0$ for $x\in B_R(r_0)$ , which shows that $d_2\Delta v\gt 0$ in $ B_R(r_0)$ . In addition, $v(x)\lt v(x_1)$ in $B_R(r_0)$ . By employing the Hopf boundary point Proposition (see, e.g., Chapters 8 and 9 of [Reference Gilbarg and Trudinger13]), we observe that $\partial _\tau v\gt 0$ . This is in contradiction with the boundary condition $\partial _\tau v=0$ . Thus, $v(x)\le v^*_2$ on $\overline {\Omega }$ . Since $f_2^\gamma (v)=\widetilde f_2^\gamma (v)$ for all $v\leq v_M$ , any classical solution $v(x)$ of (4.4) is also a classical solution of (4.1).

Finally, we demonstrate that $v(x)$ crosses $\gamma$ . Otherwise, we can suppose that $\gamma \lt v(x)\le v^*_2$ for all $x\in \overline {\Omega }$ . According to Proposition2.3 (i), we find that $f_2^\gamma (v)=f_2(h_2(v),v)\geq 0$ . So, $\int _\Omega f_2(h_2(v),v)dx\geq 0$ . In addition, integrating the first equation of (4.1) over $\Omega$ , we get $\int _\Omega f_2^\gamma (v)=\int _\Omega f_2(h_2(v),v)dx=0$ . This only holds if $v(x)\equiv v^*_2$ . So we have a contradiction. Similarly, we can demonstrate that $0\le v(x)\lt \gamma$ for all $x\in \overline {\Omega }$ is also invalid.

Remark 4. Suppose $v(x)$ is a solution of (4.1) and define

\begin{equation*}u(x)= \left \{\begin{array}{lll} 0 & v(x)\lt \gamma, \\ h_2(v(x)) & v(x)\gt \gamma .\\ \end{array} \right .\end{equation*}

Then, $(u(x),v(x))$ forms a stationary solution of problem (1.2).

5. Monotone and symmetric solutions

In this section, we focus on the construction of monotonic and symmetric solutions of (4.1) in the one-dimensional space domain $[0,1]$ through the method in [Reference Zhang31]. Therefore, (4.1) can be expressed as

(5.1) \begin{equation} \left \{\begin{array}{lll} d_2 v''+f_2^\gamma (v)=0,& x\in (0,1),\\ v'(0)=v'(1)=0.\\ \end{array} \right . \end{equation}

Firstly, to recall that in the introduction, a solution $V_{n,+}(x) (n\geq 2)$ for an $n$ -mode means that the number of points of discontinuity for $V''_{n,+}(x)$ is $n$ .

First of all, we fix $\gamma \in (\xi, v^*_2)$ and follow the method in [Reference Takagi and Zhang28] to construct monotonically increasing and symmetric solutions to (5.1) for each $d_2\gt 0$ ,where $\xi =p(0)$ has been defined in (2.2).

Theorem 5.1. Assume that Proposition 2.2 (i) and $d_2\gt 0$ hold. For every $\gamma \in (\xi, v^*_2)$ , problem ( 5.1 ) has a monotonic increasing solution $V_{1,+}(x;\,d_2)$ . Furthermore, the equation ( 5.1 ) has an $n$ -mode symmetric solution $V_{n,+}(x;\,\overline d_2)$ with $\overline d_2=n^2d_2$ for each value of $n \gt 2$ .

Proof. We divide the proof into two steps.

Step 1. We consider the following initial value problems

(5.2) \begin{equation} \left \{\begin{array}{lll} d_2 v''-av=0,& x\in (0,n_0),\\ v'(0)=0,\qquad v(n_0)=\gamma, \\ \end{array} \right . \end{equation}
(5.3) \begin{align} \left \{\begin{array}{lll} d_2 v''+f_2(h_2(v),v)=0,& x\in (n_0,1),\\ v(n_0)=\gamma, \qquad v'(1)=0,\\ \end{array} \right . \\[8pt]\nonumber\end{align}

where $n_0\in (0,1)$ . It is easy to find that $W_0(x;\,n_0,d_2)$ is a unique monotone increasing solution to problem (5.2) for every $d_2\gt 0$ , where

\begin{equation*}W_0(x;\,n_0,d_2)=\dfrac {\gamma }{\mbox {cosh}\sqrt {a/d_2}n_0}\mbox {cosh}\sqrt {a/d_2}x.\end{equation*}

Next, we prove that problem (5.3) has a unique monotone increasing solution $W_1(x;\,n_0,d_2)$ for every $d_2\gt 0$ . Since $f_2(h_2(v^*_2),v^*_2)=0$ , $f_2(h_2(v),v)\gt 0$ for all $v\in (0,v^*_2)$ and $\partial _v f_{2}(h_2(v^*_2),v^*_2)\lt 0$ by Proposition2.3, there exist $0\lt n_1\lt n_2$ such that $-n_1(v-v^*_2)\lt f_2(h_2(v),v)\lt -n_2(v-v^*_2)$ for $v\lt v^*_2$ . In order to find a solution of (5.3), we study the following initial value problems for $i=1,2$

(5.4) \begin{equation} \left \{\begin{array}{lll} d_2 v''-n_i(v-v^*_2)=0 & x\in (n_0,1),\\ v(n_0)=\gamma, \quad v'(1)=0.\\ \end{array} \right . \end{equation}

Let $Y_i(x;\,n_0,d_2)$ be respective solutions of problems (5.4) for $i=1,2.$ Then it is easy to find that $Y_1(x;\,n_0,d_2)$ is a lower solution of (5.3) and $Y_2(x;\,n_0,d_2)$ is an upper solution of (5.3). For simplicity, we let $G_{n_0}(x;\,a)=\mbox {cosh}(a(1-x))/\mbox {cosh}(a(1-n_0)).$ Simple calculations yield that for $i=1,2$ , $Y_i(x;\,n_0,d_2)=(\gamma -v^*_2)G_{n_0}(x;\,\sqrt {n_i/d_2})+v^*_2.$ We claim that $Y_1(x;\,n_0,d_2)\lt Y_2(x;\,n_0,d_2)$ , let $Y(x;\,n_0,d_2)=Y_1(x;\,n_0,d_2)-Y_2(x;\,n_0,d_2)$ . Then $Y(x;\,n_0,d_2)$ satisfies

\begin{equation*}\left \{\begin{array}{lll} d_2 Y''=n_1(Y_1-v^*_2)-n_2(Y_2-v^*_2)\gt n_1Y& x\in (n_0,1),\\ Y(k;\,n_0,d_2)=0,\quad Y'(1;\,n_0,d_2)=0.\\ \end{array} \right .\end{equation*}

If $Y(x_M;\,n_0,d_2)=\mathop {\max }\limits _{x\in [n_0,1]}Y(x;\,n_0,d_2)\gt 0$ at some $x_M\in (n_0,1),$ then

\begin{equation*}0\geq d_2 Y''(x_M;\,n_0,d_2)\gt n_1Y(x_M;\,n_0,d_2)\gt 0.\end{equation*}

This is a contradiction. So, $Y_1(x;\,n_0,d_2)\leq Y_2(x;\,n_0,d_2).$ We verify that (5.3) has a solution $W_1(x;\,n_0,d_2)$ by using the upper and lower solution approach.

Next, we show the uniqueness of a solution for (5.3). By the comparison method, we can guarantee the existence of a maximal solution $W_M(x)$ and a minimal solution $W_m(x)$ such that

\begin{equation*}Y_1(x)\lt W_m(x)\lt W_M(x)\lt Y_2(x).\end{equation*}

Let $\widetilde f_2(v)=f_2(h_2(v),v)/v$ , then $\widetilde f_2(v)$ is strictly decreasing in $v$ . Due to $W_M(x)$ and $W_m(x)$ satisfying (5.3), we see

(5.5) \begin{equation} d_2 W_M''+\widetilde f_2(W_M)W_M=0, \qquad x\in (n_0,1), \end{equation}

and

(5.6) \begin{equation} d_2 W_m''+\widetilde f_2(W_m)W_m=0, \qquad x\in (n_0,1). \end{equation}

Multiply (5.5) by $W_m$ and multiply (5.6) by $W_M$ . Then we obtain

(5.7) \begin{align} 0&={\int }_{n_0}^{1}((d_2 W_M''+\widetilde f_2(W_M)W_M)W_m-(d_2 W_m''+\widetilde f_2(W_m)W_m)W_M)dx\nonumber\\&=d_2( W_M'(x)W_m(x)-W_m'(x)W_M(x))\Bigm |^1_{n_0}+{\int }_{n_0}^{1}(\widetilde f_2(W_M)-\widetilde f_2(W_m))W_MW_m dx\nonumber\\&=d_2\gamma (W_m'(n_0)-W_M'(n_0))+{\int }_{n_0}^{1}(\widetilde f_2(W_M)-\widetilde f_2(W_m))W_MW_m dx. \end{align}

Since $W_M(x)\gt W_m(x)$ in $(n_0,1]$ and $\widetilde f_2(v)$ is strictly decreasing in $v$ , then

\begin{equation*}W_m'(n_0)-W_M'(n_0)\le 0,\quad \widetilde f_2(W_M)-\widetilde f_2(W_m) \le 0, \quad n_0 \le x \le 1.\end{equation*}

Due to $W_MW_m\gt 0$ , $W_M\equiv W_m$ can be seen from equation (5.7). This shows that we have completed the proof of the uniqueness of a solution for (5.3). Moreover, combining $d_2 W_1''+f_2(h_2(W_1),W_1)=0$ with $f_2(h_2(W_1),W_1)\gt 0$ and $W_1'(1;\,n_0,d_2)=0$ , we get that $W_1(x;\,n_0,d_2)$ is monotone increasing in $x$ .

Step 2. By simple calculation, we get

(5.8) \begin{equation} W_0'(n_0;\,n_0,d_2)=\gamma \sqrt {a/d_2}\mbox {tanh}(\sqrt {a/d_2}n_0),\, W_1'(n_0;\,n_0,d_2)=\dfrac {1}{d_2}{\int }_{n_0}^{1}f_2(h_2(W_1),W_1)dx. \end{equation}

Define $\rho _0$ to be a sufficiently small positive number. If $n_0=1-\rho _0$ , then we find $W_0'(1-\rho _0;\,1-\rho _0,d_2)\gt W_1'(1-\rho _0;\,1-\rho _0,d_2)$ by (5.8), since $f_2(h_2(v),v)$ is bounded for all $x\in [0,v^*_2]$ . In addition, if $n_0=\rho _0$ , then $W_0'(\rho _0;\,\rho _0,d_2)\lt W_1'(\rho _0;\,\rho _0,d_2)$ for sufficiently small $\rho _0\gt 0$ . Assume $\Theta (n_0,d_2)=W_0'(n_0;\,n_0,d_2)-W_1'(n_0;\,n_0,d_2)$ . Thus, we obtain

(5.9) \begin{equation} \Theta (1-\rho _0,d_2)\gt 0,\quad \Theta (\rho _0,d_2)\lt 0. \end{equation}

It is easy to see that $\Theta (n_0,d_2)$ is continuous with respect to $n_0$ for all $d_2\gt 0$ . The combination of this and (5.9), there exists a $n_0^*$ such that $\Theta (n_0^*,d_2)=0$ and

\begin{equation*}V_{1,+}(x,d_2)= \left \{\begin{array}{lll} W_0(x;\,n_0^*,d_2) & \mbox {for} \quad x\in [0,n_0^*],\\ W_1(x;\,n_0^*,d_2) & \mbox {for} \quad x\in [n_0^*,1],\\ \end{array} \right .\end{equation*}

is a monotone increasing solution of (5.1) for all $d_2\gt 0$ .

Following, using $V_{1,+}(x,d_2)$ and its reflection, we create symmetric solutions to (5.1) starting from the monotone increasing solution $V_{1,+}(x,d_2)$ . For all $n\geq 2$ , define a function $V_{n,+}(x,\overline d_2)$ on $0\le x\le 1$ by

\begin{equation*}V_{n,+}(x,\overline d_2)= \left \{\begin{array}{lll} V_{1,+}(nx-2j;\,\overline d_2) & \mbox {for} \quad x\in [2j/n,(2j+1)/n],\\ V_{1,+}(2(j+1)-nx;\,\overline d_2) & \mbox {for} \quad x\in [(2j+1)/n,2(j+1)/n],\\ \end{array} \right .\end{equation*}

where $j=0,1,2,\cdots, [n/2]$ and $\overline d_2=n^2d_2$ . Then $V_{n,+}(x,\overline d_2)$ is a symmetric solution of (5.1). This completes the proof.

Next, by using the shooting approach developed in the work of Mimura, Tabata and Hosono [Reference Mimura, Tabata and Hosono22], we further fix $\gamma \in (d^*,v^*_2)$ and demonstrate the existence and uniqueness of monotone increasing and symmetric solutions to problem (5.1). We consider the following two initial value problems

(5.10) \begin{equation} \left \{\begin{array}{l} d_2 v''-av=0,\ x\gt 0,\\ v'(0)=0,\ v(0)=b^*_0,\\ \end{array} \right . \end{equation}

and

(5.11) \begin{equation} \left \{\begin{array}{l} d_2 v''+f_2(h_2(v),v)=0,\ x\gt 0,\\ v'(0)=0,\ v(0)=b^*_1,\\ \end{array} \right . \end{equation}

where $d^*\lt b^*_0\lt \gamma \lt b^*_1\lt v^*_2.$

Let $V_0(x;\,b^*_0)$ and $V_1(x;\,b^*_1)$ be unique solutions of (5.10) and (5.11). We can see that $V_0(x;\,b^*_0)$ is monotonically increasing and $V_1(x;\,b^*_1)$ is monotonically decreasing (see Proposition5.4). For $j=0,1$ , let $x=l_j$ be the unique solution of $V_j(x;\,b^*_j)=\gamma$ and $\psi _j(l_j)$ satisfies $\psi _j(l_j)=\dfrac {\partial V_j}{\partial x}(l_j;b^*_j(l_j)).$ Next we set

(5.12) \begin{equation} \begin{array}{l} \overline l_0=\mathop {\lim \sup }\limits _{b^*_0\rightarrow d^*}l_0(b^*_0),\overline l_1=\mathop {\lim \sup }\limits _{b^*_1\rightarrow v^*_2}l_1(b^*_1),\overline \psi _0=\mathop {\lim }\limits _{l_0\rightarrow \overline l_0}\psi _0(l_0),\overline \psi _1=\mathop {\lim }\limits _{l_1\rightarrow \overline l_1}\psi _1(l_1),\\\\ z_0(\alpha _0)=\psi _0^{-1}(\alpha _0),z_1(\alpha _1)=\psi _1^{-1}(\alpha _1),\overline z_0=\mathop {\lim }\limits _{\alpha _0\rightarrow \overline \alpha }z_0(\alpha ),\overline z_1=\mathop {\lim }\limits _{\alpha _1\rightarrow -\overline \alpha }z_1(\alpha ), \end{array} \end{equation}

where $\alpha _j=\psi _j(l_j)$ and $\overline \alpha =\mbox {min}\left \{ {{\overline \psi _0,-\overline \psi _1}}\right \}$ . Here, we have utilized the facts that the inverses $b^*_j(l_j)$ of $l_j(b^*_j)$ and the inverses $\psi ^{-1}_j(\alpha _j)$ of $\alpha _j\ (j= 0,1)$ indeed exist. We will establish these findings in Proposition5.5. Then the following two theorems are main results of this section.

Theorem 5.2. Assume that Proposition 2.2 (i) holds. For each $\gamma \in (d^*,v^*_2)$ , let $1\le \overline z_0+\overline z_1.$ If $d_2\gt 0,$ $T_{1,+}(x)$ is a unique increasing solution of problem ( 5.1 ) and for every integer $n\geq 2$ , $T_{n,+}(x)$ is a unique $\mbox {n}$ -mode symmetric solution of problem ( 5.1 ).

Theorem 5.3. Assume that Proposition 2.2 (i) holds. For each $\gamma \in (d^*,v^*_2)$ , let $1\gt \overline z_0+\overline z_1.$ If $d_2\gt 0,$ $\widetilde B_{n,+}(x)$ is a unique $\mbox {n}$ -mode symmetric solution of problem ( 5.1 ) for every integer $n\geq N_0$ , where $N_0$ is the smallest positive integer greater than $1/(\overline z_0+\overline z_1)$ .

We start our discussion with the following propositions.

Proposition 5.4. Assume that Proposition 2.2 (i) holds. For all $b^*_0$ such that (5.10) has a unique positive and strictly increasing solution $V_0(x;\,b^*_0)$ defined for $x\gt 0$ , and ( 5.11 ) has a unique positive and strictly decreasing solution $V_1(x;\,b^*_1)$ for $0\le x\lt x_{b^*_1}$ , where $x_{b^*_1}$ is the solution of $V_1(x;\,b^*_1)=0.$

Proof. By $f_2(h_L(V_0),V_0)=f_2(0,V_0)=-aV_0,$ we know $d_2 V_0''=aV_0.$ By simple calculation

(5.13) \begin{equation} V_0(x;\,b^*_0)=b^*_0\mbox {cosh}\left (\sqrt {a/d_2}x\right )\quad \mbox {and}\quad V_0'(x;\,b^*_0)=b^*_0\sqrt {a/d_2}\mbox {sinh}\left (\sqrt {a/d_2}x\right ). \end{equation}

So, for all $x\gt 0,$ we can get $V_0(x;\,b^*_0)$ is positive and strictly increasing. By Proposition2.3 (i), we have $d_2 V_1''=-f_2(h_2(V_1),V_1)\lt 0.$ Thus, $V_1(x;\,b^*_1)$ is decreasing with respect to $x$ . Since $V_1'(0;\,b^*_1)=0,$ so $V_1'(x;\,b^*_1)\lt 0$ for $0\le x\lt x_{b^*_1}.$ These results show that $V_1(x;\,b^*_1)\gt 0$ is strictly monotone decreasing on $[0,x_{b^*_1})$ and $V_1(x_{b^*_1};b^*_1)=0.$

Proposition 5.5. Assume that Proposition 2.2 (i) holds. For $\gamma \in (d^*,v^*_2)$ , we have the following conclusions

(i) $\dfrac {\partial {l_0}}{\partial b^*_0}\lt 0$ , $\mathop {\lim }\limits _{b^*_0\rightarrow \gamma }l_0(b^*_0)=0$ and $\dfrac {\partial {l_1}}{\partial b^*_1}\gt 0$ , $\mathop {\lim }\limits _{b^*_1\rightarrow \gamma }l_1(b^*_1)=0$ ;

(ii) $\dfrac {\partial {\psi _0(l_0)}}{\partial l_0}\gt 0$ , $\mathop {\lim }\limits _{l_0\rightarrow 0}\psi _0(l_0)=0$ and $\dfrac {\partial {\psi _1(l_1)}}{\partial l_1}\lt 0$ , $\mathop {\lim }\limits _{l_1\rightarrow 0}\psi _1(l_1)=0$ ,

where for $j=0,1$ , $l_j$ and $\psi _j(l_j)$ are already defined in the previous section.

Proof. Since the assertions for $l_0$ and $\psi _0(l_0)$ may be treated similarly, it is sufficient to verify the statements for $l_1$ and $\psi _1(l_1)$ .

(i) Recall that $V_1(l_1(b^*_1);\,b^*_1)=\gamma$ . We differentiate it with respect to both sides with respect to $b^*_1$ and get

\begin{equation*}\dfrac {\partial {V_1}}{\partial x}\dfrac {dl_1}{db^*_1}+\dfrac {\partial {V_1}}{\partial b^*_1}=0,\end{equation*}

so

\begin{equation*}\dfrac {dl_1}{db^*_1}=-{\dfrac {\partial {V_1}}{\partial b^*_1}}/{\dfrac {\partial {V_1}}{\partial x}}.\end{equation*}

Moreover, let $\xi _1(x;\,b^*_1)=\dfrac {\partial {V_1}}{\partial x}(x;\,b^*_1)$ and $\eta _1(x;\,b^*_1)=\dfrac {\partial {V_1}}{\partial b^*_1}(x;\,b^*_1)$ . We get $\xi _1(x;\,b^*_1)\lt 0$ for all $x\in (0,l_1)$ by the proof of Proposition5.4. It is easy to find that $\eta _1(x;\,b^*_1)$ is a solution of

\begin{equation*}\left \{\begin{array}{lll} d_2\eta _1''+\dfrac {d}{dv}f_2(h_2(V_1),V_1)\eta _1=0 & \mbox {for} \quad x\in [0,1],\\ \eta _1'(0;\,b^*_1)=0, \quad \eta _1(0;\,b^*_1)=1.\\ \end{array} \right .\end{equation*}

We find that if $0\le x\le l_1$ , then $V_1(x;\,b^*_1)\in [\gamma, b^*_1]\subset (d^*,v^*_2].$ By Proposition2.3 (ii), we have $\dfrac {d}{dv}f_2(h_2(V_1),V_1)\lt 0$ for $V_1(x;\,b^*_1)\in [\gamma, b^*_1]\subset (d^*,v^*_2].$ So,

\begin{equation*}\eta _1''(0;\,b^*_1)=-\dfrac {1}{d_2}\dfrac {d}{dv}f_2(h_2(b^*_1),b^*_1)\eta _1(0;\,b^*_1)\gt 0.\end{equation*}

Expanding $\eta _1(x;\,b^*_1)$ near $x=0$ , we get

\begin{equation*}\eta _1(x;\,b^*_1)=\eta _1(0;\,b^*_1)+\eta _1'(0;\,b^*_1)x+\dfrac {1}{2}\eta _1''(0;\,b^*_1)x^2+\cdots \gt 0.\end{equation*}

Thus, we have $d_2\eta _1''=-\dfrac {d}{dV_1}f_2(h_2(V_1),V_1)\eta _1\gt 0$ for $\eta _1(x;\,b^*_1)\gt 0$ . Then, $\eta _1'(x;\,b^*_1)$ is strictly increasing with respect to $x$ . Since $\eta _1'(0;\,b^*_1)=0,$ we know $\eta _1'(x;\,b^*_1)\gt 0$ for $x\in (0,l_1]$ . Combining this with $\eta _1(0;\,b^*_1)=1$ , we get $\eta _1(x;\,b^*_1)\geq 1$ for all $x\in [0,l_1]$ . So,

\begin{equation*}\dfrac {dl_1}{db^*_1}=-{\dfrac {\partial {V_1}}{\partial b^*_1}}/{\dfrac {\partial {V_1}}{\partial x}}=-\dfrac {\eta _1(l_1(b^*_1);\,b^*_1)}{\xi _1(l_1(b^*_1);\,b^*_1)}\gt 0,\end{equation*}

and $l_1$ is strictly increasing in $b^*_1$ . Next, we show that $\mathop {\lim }\limits _{b^*_1\rightarrow \gamma }l_1(b^*_1)=0$ . Since $e^*\,:\!=f_2(h_2(\gamma ),\gamma )/d_2\gt 0$ and $V_1(x;\,b^*_1)$ satisfies

(5.14) \begin{equation} V_1(x;\,b^*_1)=b^*_1-\dfrac {e^*}{2}x^2+O(x^3) \quad \mbox {as}\quad x\rightarrow 0. \end{equation}

Therefore, by simple calculation, we have $l_1(b^*_1)=\sqrt {2(b^*_1-\gamma )/e^*}(1+o(1))$ as $b^*_1\rightarrow \gamma$ , which shows that $\mathop {\lim }\limits _{b^*_1\rightarrow \gamma }l_1(b^*_1)=0$ . As a result, the inverse of $l_1(b^*_1)$ exists and is represented as $b^*_1(l_1)$ .

(ii) Since $\psi _1(l_1)=\dfrac {\partial V_1}{\partial x}(l_1;\,b^*_1(l_1))=\xi _1(l_1;\,b^*_1(l_1)),$ from Proposition 4.2, we get

\begin{equation*}\dfrac {d\psi _1(l_1)}{dl_1}=\dfrac {\partial \xi _1}{\partial x}+\dfrac {\partial \xi _1}{\partial b^*_1}\dfrac {db^*_1}{dl_1}=\dfrac {\partial \xi _1}{\partial x}-\dfrac {\partial \xi _1}{\partial b^*_1}\dfrac {\xi _1(l_1;\,b^*_1(l_1))}{\eta _1(l_1;\,b^*_1(l_1))}\end{equation*}
\begin{equation*}\quad \quad \qquad =\dfrac {1}{\eta _1(l_1;\,b^*_1(l_1))}\left (\dfrac {\partial \xi _1}{\partial x}\eta _1-\dfrac {\partial \eta _1}{\partial x}\xi _1\right )(l_1;\,b^*_1(l_1))\end{equation*}
\begin{equation*}\quad \quad \qquad =\dfrac {1}{\eta _1(l_1;\,b^*_1(l_1))}\left (\dfrac {\partial \xi _1}{\partial x}\eta _1-\dfrac {\partial \eta _1}{\partial x}\xi _1\right )(0;\,b^*_1(l_1))\end{equation*}
\begin{equation*}=-\dfrac {f_2(h_2(b^*_1),b^*_1)}{d_2\eta _1(l_1;\,b^*_1(l_1))}\lt 0.\end{equation*}

Since $\left (\dfrac {\partial \xi _1}{\partial x}\eta _1-\dfrac {\partial \eta _1}{\partial x}\xi _1\right )'=\xi _1''\eta _1-\eta _1''\xi _1=0$ , we obtain $\dfrac {\partial \xi _1}{\partial x}\eta _1-\dfrac {\partial \eta _1}{\partial x}\xi _1$ is a constant. And by (5.14), we have $V_1^{'}(x;\,b^*_1)=-e^*x+O(x^2)$ as $x\rightarrow 0$ , so

\begin{equation*}\psi _1(l_1)=-\sqrt {2e^*(b^*_1-\gamma )}(1+o(1))\quad \mbox {as}\quad l_1\rightarrow 0.\end{equation*}

Hence, $\mathop {\lim }\limits _{l_1\rightarrow 0}\psi _1(l_1)=-\mathop {\lim }\limits _{b^*_1\rightarrow \gamma }\sqrt {2e^*(b^*_1-\gamma )}=0.$ As a result, $\psi _1(l_1)$ is strictly decreasing with respect to $l_1$ and the inverse $\psi ^{-1}_1(\alpha _1)$ of $\alpha _1$ does exist.

Proposition 5.6. For $\alpha \in [0,\overline \alpha ]$ , $z_0(\alpha )+z_1(\!-\alpha )$ are a strictly increasing functions of class $C^1$ such that

\begin{equation*}z_0(0)+z_1(0)=0\quad \mbox {and}\quad z_0(\overline \alpha )+z_1(\!-\overline \alpha )=\overline z_0+\overline z_1.\end{equation*}

Proof. By the implicit function theorem and Proposition5.5, we know that $z_0(\alpha )$ and $z_1(\!-\alpha )$ are a strictly increasing functions of class $C^1$ . Then, it follows from Proposition5.5 (ii) that $\mathop {\lim }\limits _{l_0\rightarrow 0}\psi _0(l_0)=0$ and $\mathop {\lim }\limits _{l_1\rightarrow 0}\psi _1(l_1)=0$ . Thus, $\mathop {\lim }\limits _{\alpha \rightarrow 0}z_0(\alpha )=0$ and $\mathop {\lim }\limits _{\alpha \rightarrow 0}z_1(\!-\alpha )=0$ , which show $z_0(0)+z_1(0)=0$ . Since $\overline z_0=\mathop {\lim }\limits _{\alpha _0\rightarrow \overline \alpha }z_0(\alpha ),\overline z_1=\mathop {\lim }\limits _{\alpha _1\rightarrow -\overline \alpha }z_1(\alpha )$ , we have $z_0(\overline \alpha )+z_1(\!-\overline \alpha )=\overline z_0+\overline z_1.$

Proof of Theorem 5.2. Since $1\le \overline z_0+\overline z_1$ , $z_0(\alpha )+z_1(\!-\alpha )$ increases in $\alpha$ according to Proposition5.6, so there is a unique $\alpha ^*\in (0,\overline \alpha ]$ that satisfies $z_0(\alpha ^*)+z_1(\!-\alpha ^*)=1.$ Let $z_0(\alpha ^*)=\Upsilon ^*$ for such $\alpha ^*$ . Then the definitions of $z_0$ and $z_1$ yield $\psi _0(\Upsilon ^*)=\alpha ^*$ and $\psi _1(1-\Upsilon ^*)=-\alpha ^*$ . So, $\psi _0(\Upsilon ^*)+\psi _1(1-\Upsilon ^*)=0$ . Define

\begin{equation*}T_{1,+}(x)=\left \{\begin{array}{lll} V_0(x;\,a^*_0(\Upsilon ^*)) & \mbox {for} \ x\in [0,\Upsilon ^*],\\ V_1(1-x;\,a^*_1(1-\Upsilon ^*)) & \mbox {for} \ x\in [\Upsilon ^*,1],\\ \end{array} \right .\end{equation*}

where $a^*_0(\Upsilon ^*)=a^*_0(z_0(\alpha ^*))$ and $a^*_0(1-\Upsilon ^*)=a^*_1(z_1(\!-\alpha ^*))$ .

Then it becomes a unique increasing solution of (5.1). Then, for each integer $n\geq 2,$ we demonstrate the existence of $T_{n,+}(x)$ . Let $\overline \alpha _n=\alpha _0=-\alpha _1$ such that $z_0(\overline \alpha _n)+z_1(\!-\overline \alpha _n)=\dfrac {1}{n}$ . Since $ n\geq 2$ , we know $\dfrac {1}{n}\lt 1\le \overline z_0+\overline z_1$ . As a result, we can construct a unique monotone increasing solution $\overline Z_{1,+}(x)$ on $[0,{1}/{n}]$ , where $\overline T_{1,+}(x)=V_0(x;\,\overline b^*_0)$ with $\overline b^*_0=b^*_0(z_0(\overline \alpha _n))$ for $x\in [0,z_0(\overline \alpha _n)]$ and $\overline T_{1,+}(x)=V_1(({1}/{n})-x;\,\overline b^*_1)$ with $\overline b^*_1=b^*_1(z_1(\!-\overline \alpha _n))$ for $x\in [z_0(\overline \alpha _n),{1}/{n}]$ . Now define

\begin{equation*}T_{n,+}(x)=\left \{\begin{array}{lll} \overline T_{1,+}((2j/n)+x) & \mbox {for} \ x\in [2j/n,(2j+1)/n],\\ \overline T_{1,+}([2(j+1)/n]-x) & \mbox {for} \ x\in [(2j+1)/n,2(j+1)/n],\\ \end{array} \right .\end{equation*}

where $j=0,1,2,\cdots, [n/2]$ . Then $T_{n,+}(x)$ is an $n$ -mode symmetric solution of (5.1). This completes the proof.

Proof of Theorem 5.3 Since $z_0(\alpha )+z_1(\!-\alpha )$ is a strictly increasing function for $\alpha \in [0,\overline \alpha ]$ from Proposition5.6, whose range is the same as $[0,\overline z_0+\overline z_1]\subset [0,1]$ . If $N_0$ is the smallest positive integer greater than $1/(\overline z_0+\overline z_1)$ , it follows that there is a unique $\alpha ^*_n$ such that $z_0(\alpha ^*_n)+z_1(\!-\alpha ^*_n)={1}/{n}(n\geq N_0)$ . Similar to the proof of Theorem5.5, we can create a unique monotone increasing solution $\widetilde B_{1,+}(x)$ on $[0,z_0(\alpha ^*_n)+z_1(\!-\alpha ^*_n)]$ , where

\begin{equation*}\widetilde B_{1,+}(x)=\left \{\begin{array}{lll} V_0(x;\,b^*_0) & \mbox {for} \ x\in [0,z_0(\alpha ^*_n)],\\ V_1(z_0(\alpha )+z_1(\!-\alpha )-x;\,b^*_1) & \mbox {for} \ x\in [z_0(\alpha ^*_n),E]\\ \end{array} \right .\end{equation*}

with $b^*_0=b^*_0(z_0(\alpha ^*_n))$ , $b^*_1=b^*_1(z_1(\!-\alpha ^*_n))$ and $E=z_0(\alpha ^*_n)+z_1(\!-\alpha ^*_n)$ . Then we extend $\widetilde B_{1,+}(x)$ to the interval $[0,2E]$ by

\begin{equation*}\widetilde B_{2,+}(x)=\left \{\begin{array}{lll} \widetilde B_{1,+}(x) & \mbox {for} \ x\in [0,E],\\ \widetilde B_{1,+}(2E-x) & \mbox {for} \ x\in [E,2E].\\ \end{array} \right .\end{equation*}

We continue this process until $x$ reaches $x=1$ . Then $\widetilde B_{n,+}(x)$ is an $n$ -mode symmetric solution of (5.1).

6. Existence and stability of bifurcation solutions

Firstly, in order to study the stability of this equilibrium solution for system (1.1) on one-dimensional domain [0, 1], we analyse the spectrum of the linearized operator through the method in [Reference Zhang31]. Let $(\widetilde {u},\widetilde {v})$ be any constant solution of system (1.2) and

\begin{equation*}f(u,v)=(f_1(u,v),d_2v{''}+f_2(u,v)),\end{equation*}

then the Fre’chet derivative with respect to $(u,v)$ of $F$ at $(\widetilde {u},\widetilde {v})$ is expressed as follows

\begin{equation*}L={\left (\begin{array}{cc} f_{11} & f_{12} \\ f_{21} & d_2\frac {d^2}{dx^2}+f_{22} \end{array}\right )},\end{equation*}

where $f_{11}=f_{1u}(\widetilde {u},\widetilde {v})$ , $f_{12}=f_{1v}(\widetilde {u},\widetilde {v})$ , $f_{21}=f_{2u}(\widetilde {u},\widetilde {v})$ , $f_{22}=f_{2v}(\widetilde {u},\widetilde {v})$ . Suppose that $\lambda$ is an eigenvalue of $L$ . Then we find that $\lambda$ satisfies the characteristic equation

(6.1) \begin{equation} \lambda ^2-(f_{11}+f_{22}-d_2L_j)\lambda +f_{11}f_{22}-f_{12}f_{21}-d_2f_{11}L_j=0 \end{equation}

for some $j\ge 0$ , where $L_j=(\pi j)^2$ , $j=0,1,2,\cdots, $ are the eigenvalues for $\frac {d^2}{dx^2}$ subject to Neumann boundary conditions. In addition, $\mbox {cos}(\pi jx)$ is an eigenfunction corresponding to $L_j$ and $\left \{ {\mbox {cos}(\pi jx)}\right \}_{j=0}^\infty$ forms a basis of $L^2(0,1)$ .

Theorem 6.1. For $d_2\gt 0$ , the following assertions are true.

(i) The trivial solution $(\widetilde {u},\widetilde {v})=(0, 0)$ and the semi-trivial solution $(\widetilde {u},\widetilde {v})=(K, 0)$ are unstable.

(ii) The positive solution $(\widetilde {u},\widetilde {v})=(u^*_2,v^*_2)$ is locally asymptotically stable under the assumption of Proposition 2.2 (i).

(iii) The positive solution $(\widetilde {u},\widetilde {v})=(u^*_3,v^*_3)$ is locally asymptotically stable under the assumption of Proposition 2.2 (ii).

Proof. The proof is simple. As a result, we omit the detail.

Then, we consider $d_2$ as a bifurcation parameter and study the bifurcation problem near the constant steady state $(u^*_3,v^*_3)$ in the boundary value problem

(6.2) \begin{equation} \left \{\begin{array}{lll} r\left (1-\dfrac {u}{K}\right )u-\dfrac {cuv}{m+bu}=0,& x\in [0,1],\\ d_2 v''-av+\dfrac {\beta cuv}{m+bu}=0,& x\in (0,1),\\ v'(0) =0,\quad v'(1)=0.\\ \end{array} \right . \end{equation}

Let

\begin{equation*}\widetilde X=\left \{ {(u,v)|u\in C^0([0,1]),v\in C^2([0,1]),v'(0)=v'(1)=0}\right \},\end{equation*}
\begin{equation*}\widetilde Y=C^0([0,1]) \times C^0([0,1]).\end{equation*}

Then the $Fr\acute{e}chet$ derivative $\widetilde {L}$ with respect to $(u,v)$ of $F$ at $(u^*_3,v^*_3)$ can be written as follows

\begin{equation*}\widetilde {L}={\left (\begin{array}{cc} \widetilde {f_{11}} & \widetilde {f_{12}} \\ \widetilde {f_{21}} & d_2\frac {d^2}{dx^2}+\widetilde {f_{22}} \end{array}\right )},\end{equation*}

where

\begin{equation*}\widetilde {f_{11}}=\dfrac {r(K-2u^*_3)}{K}-\dfrac {mr(K-u^*_3)}{K(m+bu^*_3)}\gt 0,\quad \widetilde {f_{12}}=-\dfrac {cu^*_3}{m+bu^*_3}\lt 0,\end{equation*}
\begin{equation*}\widetilde {f_{21}}=\dfrac {\beta cmv^*_3}{(m+bu^*_3)^2}\gt 0,\quad \widetilde {f_{22}}=0.\end{equation*}

Therefore, the characteristic equation (6.1) is transformed into

(6.3) \begin{equation} \lambda ^2-(\widetilde {f_{11}}-d_2L_j)\lambda -\widetilde {f_{12}}\widetilde {f_{21}}-d_2\widetilde {f_{11}}L_j=0. \end{equation}

Let $\widetilde {d_2}=-{\widetilde {f_{12}}\widetilde {f_{21}}}/{\widetilde {f_{11}}L_j}.$ Then $\widetilde {d_2}\gt 0$ for every $j\ge 1.$ If we assume that $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\neq 0,$ then at $d_2=\widetilde {d_2}$ for every $j\ge 1,$ $\widetilde {f_{11}}-d_2L_j=(\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}})/\widetilde {f_{11}}\neq 0$ . So, we see that the zero is a simple eigenvalue of (6.3). Therefore, the following conclusions can be drawn.

Theorem 6.2. Assume that Proposition 2.2 (ii) holds. If $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\neq 0,$ then $(\widetilde {d_2},(u^*_3,v^*_3))$ is a bifurcation point of $F=0$ . Furthermore, there is a $\delta _0$ , which satisfies that (6.2) admits a one-parameter family of non-constant solutions $\left \{ {(\widetilde {d_2}+d_2(l),(u(l),v(l))),|l|\lt \delta _0}\right \}$ of the form $u(l)=u^*_3+l\phi _j+o(l)$ , $v(l)=v^*_3+l\psi _j+o(l)$ , where $\phi _j=\mbox {cos} (\pi jx)$ , $\psi =-(\widetilde {f_{11}}\phi _j)/\widetilde {f_{12}}$ and $d_2(0)=0$ . Particularly, there exist no solutions other than $\left \{ {(\widetilde {d_2}+d_2(l),(u(l),v(l))),|l|\lt \delta _0}\right \}\cup \left \{ {(d_2,(u^*_3,v^*_3)),|d_2-\widetilde {d_2}|\lt \delta _0}\right \}$ in a small neighbourhood of $(\widetilde {d_2},(u^*_3,v^*_3))$ in ${R}\times X$ .

Proof. Assume that $\widetilde {\Phi }=(\widetilde {\phi },\widetilde {\psi })\in \mbox {ker}\widetilde {L}$ and let $\widetilde {\phi }=\sum _{ j}\widetilde {c_j}{\phi _j}$ , $\widetilde {\psi }=\sum _{ j}\widetilde {d_j}{\phi _j}$ , then $\sum _{ j}^\infty D_j{\left (\begin{array}{cc} \widetilde {c_j} \\ \widetilde {d_j} \end{array}\right )}\phi _j=0$ , where

\begin{equation*}D_j={\left (\begin{array}{cc} \widetilde {f_{11}} & \widetilde {f_{12}} \\ \widetilde {f_{21}} & -d_2L_j \end{array}\right )}.\end{equation*}

Obviously, $\mbox {det}D_j=0\Leftrightarrow d_2=\widetilde {d_2}$ . Let $d_2=\widetilde {d_2}$ , we get

\begin{equation*}\mbox {ker}\widetilde {L}=\mbox {span}\{ {\Phi _0}\},\quad \Phi _0= {\left (\begin{array}{cc} \phi _0 \\ \psi _0 \end{array}\right )}={\left (\begin{array}{cc} \mbox {cos} (\pi jx) \\ -\frac {\widetilde {f_{11}}}{\widetilde {f_{12}}}\mbox {cos} (\pi jx) \end{array}\right )}.\end{equation*}

Similar to this, it is simple to calculate an eigenvector $\Phi _0^*$ of $\widetilde {L}^*$ associated with $0$ having the following form

\begin{equation*}\mbox {ker}\widetilde {L}^*=\mbox {span}\{ {\Phi _0^*}\},\quad \Phi _0^*= {\left (\begin{array}{cc} \phi _0 \\ \psi _0 \end{array}\right )}={\left (\begin{array}{cc} \mbox {cos} (\pi jx) \\ -\frac {\widetilde {f_{11}}}{\widetilde {f_{21}}}\mbox {cos} (\pi jx) \end{array}\right )},\end{equation*}

where $\widetilde {L}^*$ is the adjoint operator of $\widetilde {L}$ , which is obtained by

\begin{equation*}\widetilde {L}^*={\left (\begin{array}{cc} \widetilde {f_{11}} & \widetilde {f_{21}} \\ \widetilde {f_{12}} & d_2\frac {d^2}{dx^2} \end{array}\right )}.\end{equation*}

Because $\mbox {rang}\widetilde {L}=(\mbox {ker}\widetilde {L}^*)^\bot$ , we have $\mbox {dim}\mbox {ker}\widetilde {L}^*=\mbox {codim}\,\mbox {rang}\,\widetilde {L}=1.$ Finally, since

\begin{equation*}\hat {L}=\dfrac {\partial \hat {L}}{\partial d_2}{\left (\begin{array}{cc} 0 & 0 \\ 0 & \frac {d^2}{dx^2} \end{array}\right )}\end{equation*}

and $\hat {L}\Phi _0\notin \mbox {rang}\,\widetilde {L}$ , we get that the conditions required for the standard bifurcation theorem to apply, based on the presence of a simple eigenvalue [Reference Crandall and Rabinowitz8], are satisfied.

Next, we study the stability of bifurcation solutions. Suppose that $L(l)$ represent the linearized operator $\partial _UF(\widetilde {d_2}+d_2(l), \overline {U}+l\Phi _0+o(l))$ , where $U=(u,v)$ , $\overline {U}=(u^*_3,v^*_3)$ and $(\widetilde {d_2}+d_2(l), \overline {U}+l\Phi _0+o(l))$ is a bifurcation solution obtained by Theorem6.2.

Definition 5. Let $B(X,Y)$ denote the set of bounded linear maps of $X$ into $Y$ . Let $T,K\in B(X,Y)$ . Then $\mu \in R$ is a $K$ -simple eigenvalue of $T$ if

(i) $dim N(T-\mu K)=codim R(T-\mu K)=1$ and, if $N(T-\mu K)=span\left \{x_0\right \}$ ,

(ii) $Kx_0\in R(T-\mu K)$ .

Proposition 6.3. For $d_2=\widetilde {d_2}$ , 0 is an $i$ -simple eigenvalue of $\widetilde {L}$ and $i$ is the inclusion mapping $\widetilde X\rightarrow \widetilde Y$ .

Proof. According to the proof procedure of theorem6.2, we can get $\mbox {dim}\mbox {ker}\widetilde {L}=\mbox {codim}\,\mbox {rang}\,\widetilde {L}=1$ . Then $i\Phi \notin \mbox {rang}\,\widetilde {L}$ , where $\Phi _0$ satisfies $\mbox {ker}\,\widetilde {L}=\{ {\Phi _0}\}$ . Thus, it is clear that $\widetilde {L}$ possesses $0$ as an $i$ -simple eigenvalue according to the definition of a $K$ -simple eigenvalue presented in [Reference Wu30].

We have identified an $i$ -simple eigenvalue $\lambda _j(d_2)$ for $\widetilde {L}$ near $d_2=\widetilde {d_2}$ , as well as an $i$ -simple eigenvalue $\lambda (l)$ for $L(l)$ when $|l|$ is small enough. By making use of the well-known theorem by Crandall and Rabinowitz [Reference Crandall and Rabinowitz9], we obtain

\begin{equation*}\lim \limits _{s\to 0,\lambda (l)\neq 0}-\dfrac {ld_2'(l)\lambda _j'(\widetilde {d_2})}{\lambda (l)}=1.\end{equation*}

Proposition 6.4. If $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\lt 0$ , then both $\lambda (l)$ and $-ld_2'(l)$ possess the same sign. But if $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\gt 0$ , then both $\lambda (l)$ and $ld_2'(l)$ possess the same sign.

Proof. From (6.3) we get $(\lambda _j(d_2))^2-(\widetilde {f_{11}}-d_2L_j)\lambda _j(d_2)-\widetilde {f_{12}}\widetilde {f_{21}}-d_2\widetilde {f_{11}}L_j=0$ . Taking the derivative of both sides with respect to $d_2$ , we get $2\lambda _j(d_2)\lambda _j'(d_2)+L_j\lambda _j(d_2)-(\widetilde {f_{11}}-d_2L_j)\lambda _j'(d_2)-\widetilde {f_{11}}L_j=0.$ So, $\lambda _j(\widetilde {d_2})=0$ shows

\begin{equation*}\lambda _j'(d_2)=\dfrac {-L_j\widetilde {f_{11}}}{\widetilde {f_{11}}-L_j\widetilde {d_2}}=\dfrac {-\widetilde {f_{11}}^2L_j}{\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}}.\end{equation*}

Thus, if $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\lt 0$ , then $\lambda _j'(d_2)\gt 0$ , which shows that both $\lambda (l)$ and $-ld_2'(l)$ possess the same sign. But if $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\gt 0$ , then $\lambda _j'(d_2)\lt 0$ , which implies that both $\lambda (l)$ and $ld_2'(l)$ possess the same sign.

Then we analyse the sign of $\lambda (l)$ . Because $(u^*_3,v^*_3)$ is on the branch $u=h_1(v)$ , we study the following boundary value problem

(6.4) \begin{equation} \left \{\begin{array}{lll} d_2 v''+d(v)=0,& x\in (0,1),\\ v'(0)=0,\qquad v'(1)=0,\\ \end{array} \right . \end{equation}

where $d(v)=f_2(h_1(v),v)$ .

Theorem 6.5. Let $C=d'(v^*_3)$ , $D=d''(v^*_3)$ , $E=d'''(v^*_3)$ and define $N=3CE-5D^2$ , then we have the following conclusions.

(i) If $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\lt 0$ and $N\gt 0$ , then $\lambda (l)\lt 0$ for $0\lt l\ll 1.$

(ii) If $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\lt 0$ and $N\lt 0$ , then $\lambda (l)\gt 0$ for $0\lt l\ll 1.$

(iii) If $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\gt 0$ and $N\gt 0$ , then $\lambda (l)\gt 0$ for $0\lt l\ll 1.$

(iv) If $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\gt 0$ and $N\lt 0$ , then $\lambda (l)\lt 0$ for $0\lt l\ll 1.$

Proof. We only demonstrate the assertion (i) here, as we can approach the proof of the other assertions in a similar manner. By Proposition6.4, it is sufficient to compute $d_2'(l)$ to reveal the sign of $\lambda (l)$ . We expand $v(x,l)$ and $d_2(l)$ in $l$ to obtain

\begin{align*}v(x,l)&=v^*_3+l\widetilde n_1(x)+l^2\widetilde n_2(x)+l^3\widetilde n_3(x)+\cdots, \\ d_2(l)&=\widetilde {d_2}+l\widetilde k_1+l^2\widetilde k_2+l^3\widetilde k_3+\cdots .\end{align*}

Due to $d(v^*_3)=f_2(h_1(v^*_3),v^*_3)=0$ , we obtain

\begin{equation*}d(v)=C(v-v^*_3)+\dfrac {1}{2}D(v-v^*_3)^2+\dfrac {1}{6}E(v-v^*_3)^3+\cdots, \end{equation*}

where

\begin{equation*}C=\dfrac {\beta cmv^*_3}{(m+bh_1(v^*_3))^2}h_1'(v^*_3),\end{equation*}
\begin{equation*}D=\dfrac {-2\beta cmbv^*_3}{(m+bh_1(v^*_3))^3}(h'_1(v^*_3))^2+\dfrac {2\beta cm}{(m+bh_1(v^*_3))^2}h'_1(v^*_3)+\dfrac {\beta cmv^*_3}{(m+bh_1(v^*_3))^2}h''_1(v^*_3),\end{equation*}
\begin{equation*}E=\dfrac {6\beta cmb^2v^*_3}{(m+bh_1(v^*_3))^4}(h'_1(v^*_3))^3- \dfrac {6\beta cmb}{(m+bh_1(v^*_3))^3}(h'_1(v^*_3))^2-\end{equation*}
\begin{equation*}\dfrac {6\beta cmbv^*_3}{(m+bh_1(v^*_3))^3}h'_1(v^*_3)h''_1(v^*_3)+\dfrac {3\beta cm}{(m+bh_1(v^*_3))^2}h''_1(v^*_3)+\dfrac {\beta cmv^*_3}{(m+bh_1(v^*_3))^2}h'''_1(v^*_3).\end{equation*}

In addition, it is easy to find that $f_1(h_1(v^*_3),v^*_3)=0.$ So, we get $h_1'(v^*_3)=-\widetilde {f_{12}}/\widetilde {f_{11}}.$ Consequently, based on the previous definition of $\widetilde {f_{21}}$ and $\widetilde {d_2}$ , we know

\begin{equation*}C=\dfrac {\beta cmv^*_3}{(m+bh_1(v^*_3))^2}h_1'(v^*_3)=\dfrac {\beta cmv^*_3}{(m+bh_1(v^*_3))^2}(\dfrac {-\widetilde {f_{12}}}{\widetilde {f_{11}}})=-\dfrac {\widetilde {d_2}\widetilde {f_{11}}L_j}{\widetilde {f_{12}}}(\dfrac {-\widetilde {f_{12}}}{\widetilde {f_{11}}})=\widetilde {d_2}\pi ^2j^2.\end{equation*}

By substituting these expressions into (6.4), we can obtain a sequence of equations by assigning a value of zero to the coefficient of each power of $l$ .

(6.5) \begin{align} \widetilde {d_2}\widetilde n''_1&+C\widetilde n_1=0, \end{align}
(6.6) \begin{align} \widetilde {d_2}\widetilde n''_2+\widetilde k_1\widetilde n''_1&+\dfrac {1}{2}D\widetilde n_1^2+C\widetilde n_2=0, \end{align}
(6.7) \begin{align} \widetilde {d_2}\widetilde n''_3+\widetilde k_1\widetilde n''_2+\widetilde k_2\widetilde n''_1&+C\widetilde n_3+D\widetilde n_1\widetilde n_2+\dfrac {1}{6}E\widetilde n_1^3=0. \\[8pt]\nonumber\end{align}

By (6.5), we know that $\widetilde {d_2}\frac {d^2}{dx^2}+C$ has an eigenvalue of 0. So, equation (6.6) can be solved only if

(6.8) \begin{equation} \widetilde k_1{\int }_{0}^{1}\widetilde n''_1\widetilde n_1\mbox {d}x+\dfrac {1}{2}D{\int }_{0}^{1}\widetilde n^3_1\mbox {d}x=0. \end{equation}

When $\widetilde n_1(x)=\mbox {cos}(\pi jx)$ , equation (6.5) is satisfied. By simple calculations, it can be deduced that $\widetilde k_1=0$ when it is substituted into (6.8). Therefore, (6.6) is rewritten as

\begin{equation*}\widetilde {d_2}\widetilde n''_2+\dfrac {1}{2}D\widetilde n_1^2+C\widetilde n_2=0.\end{equation*}

Since $\widetilde n_1^2=(1+\mbox {cos}(2\pi jx))/2$ , then we know

\begin{equation*}\widetilde {d_2}\widetilde n''_2+C\widetilde n_2=-\dfrac {1}{2}D\dfrac {1+\mbox {cos}(2\pi jx)}{2}.\end{equation*}

By simple calculations, it can be deduced that

\begin{equation*}\widetilde n_2(x)=-\dfrac {D}{4C}-\dfrac {D}{4(C-4\widetilde {d_2}\pi ^2j^2)}\mbox {cos}(2\pi jx)=-\dfrac {D}{4C}+\dfrac {D}{12C}\mbox {cos}(2\pi jx).\end{equation*}

Now we consider equation (6.7). Since $\widetilde k_1=0$ , it follows that (6.7) has a solution if and only if

(6.9) \begin{equation} \widetilde k_2{\int }_{0}^{1}\widetilde n''_1\widetilde n_1\mbox {d}x+D{\int }_{0}^{1}\widetilde n_1^2\widetilde n_2\mbox {d}x+\dfrac {1}{6}E{\int }_{0}^{1}\widetilde n^4_1\mbox {d}x=0. \end{equation}

It can be directly calculated that

\begin{equation*}{\int }_{0}^{1}\widetilde n''_1\widetilde n_1\mbox {d}x=-\dfrac {1}{2}L_j,\quad {\int }_{0}^{1}\widetilde n^2_1\widetilde n_2\mbox {d}x=-\dfrac {5D}{48C},\quad {\int }_{0}^{1}\widetilde n^4_1\mbox {d}x=\dfrac {3}{8}.\end{equation*}

So, (6.9) becomes

\begin{equation*}-\dfrac {1}{2}\widetilde k_2L_j-\dfrac {5D^2}{48C}+\dfrac {E}{16}=0;\quad \mbox {so that}\quad \widetilde k_2=\dfrac {1}{24CL_j}(3CE-5D^2).\end{equation*}

Since $\widetilde k_1=0$ and $l{d_2}'(l)=l(\widetilde k_1+2l\widetilde k_2+O(l^2)),$ we have $l{d_2}'(l)=2l^2\widetilde k_2+O(l^3))$ . Thus, for $|l|$ sufficiently small, the sign of $l{d_2}'(l)$ is the same as that of $\widetilde k_2$ . This implies that the sign of $-\lambda (l)$ is the same as that of $\widetilde k_2$ if $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\lt 0$ , according to Proposition6.4. So, if $\widetilde {f_{11}}^2+\widetilde {f_{12}}\widetilde {f_{21}}\lt 0$ and $N\gt 0$ , then $\widetilde k_2\gt 0$ ; so that $\lambda (l)\lt 0.$

7. Conclusions

In this paper, we examine a mechanism of pattern formation that occurs in a predator–prey model with Holling-II functional response. The model consists of a single reaction–diffusion equation coupled with an ordinary differential equation. The value of this paper reflects in three aspects.

7.1. Existence of non-constant regular solution

We prove the existence of regular stationary solutions for system (3.14), using the method in [Reference Cygan, Marciniak–Czochra, Karch and Suzuki4]. If the internal equilibrium $2\overline u_3$ system (3.14) is greater than the carrying capacity $K$ , and the other parameters are non-negative. Furthermore, if the diffusion coefficient $d_2\gt 0$ of the predator, then the system (3.14) produces a non-constant regular solution (Theorem3.3).

7.2. Existence and uniqueness of steady states with jump discontinuity

We apply various approaches to demonstrate the existence of steady states with jump discontinuities and investigate their characteristics on a one-dimensional spatial domain (refer to Theorem4.1 for domains of higher dimensions and Theorems5.15.3 for one-dimensional domains). These results show the existence of discontinuous steady-state solutions $(u(x), v(x))$ for system (1.1), where $u(x)$ displays a jump discontinuity while $v(x)$ is either monotonic or symmetric, depending on a fixed parameter $\gamma$ . Furthermore, it is observed that by selecting a smaller range for $\gamma$ , the solution becomes unique. This uniqueness stems from the fact that $f_2(h_2(v),v)$ is a strictly decreasing function in relation to $v$ within this interval. It should be emphasized that these phenomena differ significantly from those observed in systems where both species exhibit diffusion or non-diffusion.

7.3. Existence and stability of bifurcation solutions

In Section 6, we focus on the bifurcating solutions of the system (1.1). It has been observed that stable patterns emerge near the constant equilibrium state in a partial differential equation $(PDE)$ system with diffusion-driven instability $(DDI)$ property. However, the system (1.1) analysed in this paper exhibits the characteristic of $DDI$ (Theorem6.2), but all Turing-type patterns are unstable (Theorem6.5). This is significantly different from the classical diffusive model, exhibiting a notable difference.

Funding statement

The work was partially supported by the National Natural Science Foundation of China (61872227, 12126420), Cultivation Project Funds for Beijing University of Civil Engineering and Architecture (No. X24007).

Authors contributions

All authors contributed equally and significantly in writing this article. All authors read and approved the final manuscript.

Competing interest

None.

References

Akam, M. (1989) Making stripes inelegantly. Nature 341(6240), 282283.CrossRefGoogle ScholarPubMed
Aronson, D. G., Tesei, A. & Weinberger, H. (1988) A density–dependent diffusion system with stable discontinuous stationary solutions. Annali di Matematica Pura ed Appliecata 152(1), 259280.CrossRefGoogle Scholar
Ambrosetti, A. & Rabinowitz, P. H. (1973) Dual variational methods in critical point theory and applications. J. Funct. Anal. 14(4), 349381.CrossRefGoogle Scholar
Cygan, S., Marciniak–Czochra, A., Karch, G. & Suzuki, K. (2022) Instability of all regular stationary solutions to reaction–diffusion–ODE systems. J. Differ. Equations 337, 460482.CrossRefGoogle Scholar
Casten, R. & Holland, C. (1978) Instability results for reaction diffusion equations with Neumann boundary conditions. J. Differ. Equations 27(2), 266273.CrossRefGoogle Scholar
Cygan, S., Marciniak–Czochra, A., Karch, G. & Suzuki, K. (2023) Stable discontinuous stationary solutions to reaction–diffusion–ODE systems. Commun. Part. Diff. Eq. 48(3), 478510.CrossRefGoogle Scholar
Chang, K. C. (1981) Variational methods for non–differentiable functionals and their applications to partial differential equations. J. Math. Anal. Appl. 80(1), 102129.CrossRefGoogle Scholar
Crandall, M. & Rabinowitz, P. (1971) Bifurcation from simple eigenvalues. J. Funct. Anal. 8(2), 321340.CrossRefGoogle Scholar
Crandall, M. G. & Rabinowitz, P. H. (1973) Bifurcation, perturbation of simple eigenvalues and linearized stability. Arch. Ration. Mech. An. 52(2), 161180.CrossRefGoogle Scholar
Gierer, A. & Meinhardt, H. (1972) A theory of biological pattern formation. Kybernetik (Berlin) 12(1), 3039.CrossRefGoogle ScholarPubMed
Guo, G. H. & Wang, J. J. (2024) Pattern formation and qualitative analysis for a vegetation–water model with diffusion. Nonlinear Analysis: Real World Applications 76, 104008.Google Scholar
Guo, G. H., You, J. & Ahmed Abbakar, K. (2025) Pattern dynamics in a water-vegetation model with cross-diffusion and nonlocal delay. Math. Method. Appl. Sci. 48(3), 31903213.CrossRefGoogle Scholar
Gilbarg, D. & Trudinger, N. S. (2001) Elliptic Partial Differential Equations of Second Order, 1998 ed. Springer.CrossRefGoogle Scholar
You, J. & Guo, G. H. (2025) Pattern formation for a reversible biochemical reaction model with cross-diffusion and Michalis saturation. J. Math. Chem. 63(3), 888910.CrossRefGoogle Scholar
Harting, S., Marciniak–Czochra, A. & Takagi, I. (2017) Stable patterns with jump discontinuity in systems with Turing instability and hysteresis. Discret. Contin. Dyn. S. 37(2), 757800.CrossRefGoogle Scholar
Kothe, A., Marciniak–Czochra, A. & Takagi, I. (2020) Hysteresis–driven pattern formation in reaction–diffusion–ODE systems. Discret. Contin. Dyn. S. 40(6), 757800.Google Scholar
Le, H. C., Tsujikawa, T. & Yagi, A. (2009) Stationary solutions to forest kinematic model. Glasgow Math. J. 51(1), 117.Google Scholar
Li, Y., Marciniak–Czochra, A., Takagi, I. & Wu, B. Y. (2017) Bifurcation analysis of a diffusion–ODE model with Turing instability and hysteresis. Hiroshima Math. J. 47(2), 217247.CrossRefGoogle Scholar
Marciniak–Czochra, A. (2006) Receptor–based models with hysteresis for pattern formation in hydra. Math. Biosci. 199(1), 97119.CrossRefGoogle ScholarPubMed
Marciniak–Czochra, A. (2003) Receptor–based models with diffusion–driven instability for pattern formation in hydra. J. Biol. Syst. 11(03), 293324.CrossRefGoogle Scholar
Marciniak–Czochra, A., Takagi, I. & Nakayama, M. (2015) Pattern formation in a diffusion–ODE model with hysteresis. Differ. Integral Equ. 28(7/8), 655694.Google Scholar
Mimura, M., Tabata, M. & Hosono, Y. (1980) Multiple solutions of two–point boundary value problems of Neumann type with asmall parameter. SIAM J. Math. Anal. 11(4), 613631.CrossRefGoogle Scholar
Rabinowitz, P. H. (1977) A bifurcation theorem for potential operators. J. Funct. Anal. 25(4), 412424.CrossRefGoogle Scholar
Song, Y. L., Jiang, H. P., Liu, Q. X. & Yuan, Y. (2017) Spatiotemporal dynamics of the diffusive Mussel–Algae model near Turing–Hopf bifurcation. SIAM J. Appl. Dyn. Syst. 16(4), 20302062.CrossRefGoogle Scholar
Sherrat, J. A., Maini, P. K., Jager, W. & Muller, W. (1995) A receptor–based model for pattern formation in hydra. Forma 10, 7795.Google Scholar
Turing, A. M. (1952) The chemical basis of morphogenesis. Philo. Trans. R. Soc. B 237, 3772.Google Scholar
Takagi, I. & Zhang, C. H. (2021) Pattern formation in a reaction–diffusion–ODE model with hysteresis in spatially heterogeneous environments. J. Differ. Equations 280, 928966.CrossRefGoogle Scholar
Takagi, I. & Zhang, C. H. (2020) Existence and stability of patterns in a reaction–diffusion–ODE system with hysteresis in nonuniform media. Discret. Contin. Dyn. S. 41(7), 31093140.CrossRefGoogle Scholar
Uhlenbeck, K. (1977) Regularity for a class of non-linear elliptic systems. Acta Math. 138, 219240.CrossRefGoogle Scholar
Wu, J. H. (1995) Stability of steady–state solutions of the competition model in the chemostat. Chin. Sci. Ser. A 2(14), 14.Google Scholar
Zhang, C. H. (2021) Pattern formation with jump discontinuity in a macroalgae-herbivore model with strong Allee effect in macroalgae. J. Math. Anal. Appl. 504(1), 125371.CrossRefGoogle Scholar
Zhang, C. H. & Yang, W. B. (2020) Dynamic behaviors of a predator–prey model with weak additive Allee effect on prey. Nonlinear Analysis: Real World Applications 55, 103137.Google Scholar
Zhang, N. & Kao, Y. G. (2022) A fractional–order food chain system incorporating Holling–II type functional response and prey refuge. Int. J. Bifurcat. Chaos 32(10), 2250143.CrossRefGoogle Scholar
Zhou, Y., Sun, W., Song, Y. F., Zheng, Z. G., Lu, J. H. & Chen, S. H. (2019) Hopf bifurcation analysis of a predator–prey model with Holling–II type functional response and a prey refuge. Nonlinear Dynam. 97(2), 14391450.CrossRefGoogle Scholar
Figure 0

Figure 1. Nullclines for$f_1(u,v)=f_2(u,v)=0$. The blue curve represents the solution of$f_1(u,v)=0$, while the red curve represents the solution of$f_2(u,v)=0$. In$(a)$, we select$a=0.4$, $b=1$, $m=0.3$, $K=0.8$, $c=1$, $\beta =1.4$, $r=0.3$. In$(b)$, we select $a=0.4$, $b=1$, $m=0.3$, $K=0.8$, $c=1$, $\beta =0.6$, $r=0.3$.