Skip to main content

Theory and Modern Applications

Delay-induced Hopf bifurcation in a diffusive Holling–Tanner predator–prey model with ratio-dependent response and Smith growth

Abstract

In this paper, we study a diffusion Holling–Tanner predator–prey model with ratio-dependent functional response and Simth growth subject to a homogeneous Neumann boundary condition. Firstly, we use iteration technique and eigenvalue analysis to get the local stability and a Hopf bifurcation at the positive equilibrium. Secondly, by choosing the constant related to delay as bifurcation parameter we obtain periodic solutions near the positive equilibrium. Besides, by using center manifold theory and normal form theory we reflect the stability with Hopf bifurcating periodic solution and bifurcating direction.

1 Introduction

Mathematical ecology is a fast and active branch of biomathematics, and the dynamics of biological models are very rich. Thus it is very important for research. The predator–prey model is a very important population dynamics model. It is mainly reflected the factors influencing on the model in the functional response, based on ratio, age, or gender structure, reaction–diffusion, etc. The results of the study are mainly reflected in the stability of the equilibrium point, periodic solution, almost periodic solution, limit cycle, Hopf bifurcation, etc. The functional response is a very important factor to predator–prey model [15], It can be divided into several types, such as Hassell–Varley type, Holling types, Beddington–DeAngelies type, Crowley–Martin type, and so on.

In recent years, many authors investigate the diffusive predator–prey model with logistic prey growth rate [610] or strong Allee effect in prey [11]. It came out that this assumption is not realistic for a food-limited population under the effect of environmental toxicants. Thus the population dynamics growth restriction is based on the unused available resources [1216]. This model, also known as the Holling–Tanner model, has been studied for both its mathematical properties and its efficacy for describing real ecological systems such as mite/spider mite, lynx/hare, sparrow/sparrow hawk, and so one by Holling [2], Tanner [3] and Wollkind, Collings, and Logan [4].

In particular, Yue and Wang [16] studied the stability and Hopf bifurcation of a diffusive Holling–Tanner predator–prey model with smith growth subject to Neumann boundary condition:

$$ \textstyle\begin{cases} \frac{\partial u(x,t)}{\partial t}=d_{1}\Delta u(x,t)+\frac{u(x,t) ( 1-u(x,t) ) }{1+cu(x,t)}- \frac{\beta u(x,t)v(x,t)}{\alpha v(x,t)+u(x,t)},&x\in (0,\pi),t>0, \\ \frac{\partial v(x,t)}{\partial t}=d_{2}\Delta v(x,t)+\delta v(x,t) ( 1-\frac{\eta v(x,t)}{u(x,t)} ), & x\in (0,\pi),t>0, \\ u_{x}(0,t)=u_{x}(\pi,t)=v_{x}(0,t)=v_{x}(\pi,t)=0, & t\geq 0, \\ u(x,0)=u_{0}(x)\geq 0,\qquad v(x,0)=v_{0}(x)\geq 0, & x\in (0,\pi). \end{cases} $$
(1.1)

In realistic systems, delay exists everywhere. We have obtained many interesting conclusions since time delays may have very complex influence on the dynamical behaviors of systems. In the survey papers [12, 1719], the authors deduced that delay difference equations are much more complicated than ordinary differential equations, because time delay can lead to unstable equilibrium and induce bifurcation. Many results on the Hopf bifurcation of diffusion system can be found in [2027]. The Hopf bifurcation of the diffusive predator–prey system with delay subject to homogeneous Neumann boundary conditions was studied in [2023]. The Hopf bifurcation for a delayed predator–prey diffusion system with Dirichlet boundary condition wasonsidered in [24]. The Hopf bifurcation for a predator-prey model with age structure was studied in [27].

Chen et al. [20] have considered a delayed diffusive Leslie–Gower predator–prey system with homogeneous Neumann boundary conditions

$$\begin{aligned} \textstyle\begin{cases} \frac{\partial u(x,t)}{\partial t}=d_{1}\Delta u(x,t)+u(x,t) ( p- \alpha u(x,t)-v ( x,t-\tau_{1} ) ),& x\in (0,\pi),t>0, \\ \frac{\partial v(x,t)}{\partial t}=d_{2}\Delta v(x,t)+\mu v(x,t) ( 1-\frac{ v(x,t)}{u ( x,t-\tau_{2} ) } ), & x\in (0,\pi),t>0, \\ u_{x}(0,t)=u_{x}(\pi,t)=v_{x}(0,t)=v_{x}(\pi,t)=0, & t\geq 0, \\ u(x,t)=\phi (x,t)\geq 0, & (x,t)\in [0,\pi ]\times [-\tau_{2},0], \\ v(x,t)=\psi (x,t)\geq 0, & (x,t)\in [0,\pi ]\times [-\tau_{1},0]. \end{cases}\displaystyle \end{aligned}$$

Some related work considered the Hopf bifurcation in predator–prey models. According to the results of [20] and the analysis of the stability and bifurcation system (1.1) in [16], we mainly consider the dynamics of the following system:

$$ \textstyle\begin{cases} \frac{\partial u(x,t)}{\partial t}=d_{1}\Delta u(x,t)+\frac{u(x,t) ( 1-u(x,t) ) }{1+cu(x,t)}-\frac{\beta u(x,t)v(x,t-\tau_{1})}{ \alpha v(x,t-\tau_{1})+u(x,t)}, &x\in (0,\pi),t>0, \\ \frac{\partial v(x,t)}{\partial t}=d_{2}\Delta v(x,t)+\delta v(x,t) ( 1-\frac{\eta v(x,t)}{u(x,t-\tau_{2})} ), & x\in (0,\pi),t>0, \\ u_{x}(0,t)=u_{x}(\pi,t)=v_{x}(0,t)=v_{x}(\pi,t)=0, & t\geq 0, \\ u(x,t)=\phi (x,t)\geq 0, & (x,t)\in [0,\pi ]\times [-\tau_{2},0], \\ v(x,t)=\psi (x,t)\geq 0, & (x,t)\in [0,\pi ]\times [-\tau_{1},0]. \end{cases} $$
(1.2)

As far as we know, there are no results on the Hopf bifurcation of system (1.2). In this paper, we investigate the stability of the positive equilibrium, delay-induced Hopf bifurcation, and the properties of Hopf bifurcation such as the direction of the bifurcation and stability of the bifurcating periodic solutions.

2 The stability of equilibrium and the existence of Hopf bifurcation

In this paper, we only investigate the effect of the delay on the stability of the positive equilibrium \(E^{*}(u^{*},v^{*})\) of system (1.2), where

$$\begin{aligned} u^{*}=\frac{\alpha +\eta -\beta }{\alpha +\eta +\beta c},\qquad v^{*}=\frac{u ^{*}}{\eta }, \end{aligned}$$

under condition \((H)\): \(\beta <\alpha +\eta \).

For simplicity, in this paper, we only study the case \(\tau_{1}=\tau _{2}=\tau \). Let

$$\begin{aligned}& g^{(1)}(u,v)=\frac{u(x,t) ( 1-u(x,t) ) }{1+cu(x,t)}-\frac{ \beta u(x,t)v(x,t-\tau)}{\alpha v(x,t-\tau)+u(x,t)}, \\& g^{(2)}(u,v)=\delta v(x,t) \biggl( 1-\frac{\eta v(x,t)}{u(x,t-\tau)} \biggr). \end{aligned}$$

The linearization of (1.2) at the equilibrium \(E^{*}\) is

( u ( x , t ) t v ( x , t ) t ) =DΔ ( u ( x , t ) v ( x , t ) ) + A 0 ( u ( x , t ) v ( x , t ) ) + A 1 ( u ( x , t τ ) v ( x , t τ ) ) ,
(2.1)

where

D Δ = ( d 1 Δ 0 0 d 2 Δ ) , A 0 = ( b 11 0 0 b 22 ) , A 1 = ( 0 b 12 b 21 0 ) ,

with

$$\begin{aligned}& b_{11}=\frac{ ( 1-2u^{*} ) ( 1+cu^{*} ) -cu^{*} ( 1-u ^{*} ) }{ ( 1+cu^{*} ) ^{2}}-\frac{\alpha \beta v^{*2}}{ ( \alpha v^{*}+u^{*} ) ^{2}}, \\& b_{12}=\frac{-\beta u^{*2}}{ ( \alpha v^{*}+u^{*} ) ^{2}},\qquad b _{21}=\frac{\delta }{\eta }, \qquad b_{22}=-\delta. \end{aligned}$$

So, the characteristic equation of (2.1) is

$$ \operatorname{det}\bigl( \lambda I-M_{k}-A_{0}-A_{1}e^{-\lambda \tau } \bigr) =0, $$
(2.2)

where I is the \(2\times 2\) identity matrix, and \(M_{k}=-k^{2}\operatorname{diag}\{ d_{1},d_{2} \} \), \(k\in N_{0}=\{0,1,2,3,\ldots \}\). This is

$$\begin{aligned}& \lambda^{2}+ \bigl[ ( d_{1}+d_{2} ) k^{2}- ( b_{11}+b_{22} ) \bigr] \lambda \\& \quad {}+ \bigl[ d_{1}d_{2}k^{4}-(b_{11}d_{2}+b_{22}d_{1})k^{2}+ \bigl( b _{11}b_{22}-b_{12}b_{21}e^{-2\lambda \tau } \bigr) \bigr] =0. \end{aligned}$$
(2.3)

When \(\tau =0\), it becomes

$$ \lambda^{2}+T_{k}\lambda +D_{k}=0, $$

where

$$\begin{aligned}& T_{k}= ( d_{1}+d_{2} ) k^{2}- ( b_{11}+b_{22} ), \\& D_{k}=d_{1}d_{2}k^{4}- ( b_{11}d_{2}+b_{22}d_{1} ) k^{2}+ ( b_{11}b_{22}-b_{12}b_{21} ). \end{aligned}$$

When \(\tau \neq 0\), assume that (\(\omega >0\)) is the root of (2.3). Then we have

$$\begin{aligned}& -\omega^{2}+i \bigl[ (d_{1}+d_{2})k^{2}- ( b_{11}+b_{22} ) \bigr] \omega \\& \quad {}+ \bigl[ d_{1}d_{2}k^{4}- ( b_{11}d_{2}+b_{22}d_{1} ) k ^{2}+ \bigl( b_{11}b_{22}-b_{12}b_{21}e^{-2i\omega \tau } \bigr) \bigr] =0. \end{aligned}$$
(2.4)

Separating the real and imaginary parts of (2.4), we get

$$ \textstyle\begin{cases} -\omega^{2}+d_{1}d_{2}k^{4}- ( b_{11}d_{2}+b_{22}d_{1} ) k^{2}+b _{11}b_{22}-b_{12}b_{21}\cos 2\omega \tau =0, \\ [ ( d_{1}+d_{2} ) k^{2}- ( b_{11}+b_{22} ) ] \omega +b_{12}b_{21}\sin 2\omega \tau =0, \end{cases} $$
(2.5)

which implies that

$$ \omega^{4}+P_{k}\omega^{2}+Q_{k}=0, $$
(2.6)

where

$$\begin{aligned}& P_{k}= \bigl[ d_{1}k^{2}-b_{11} \bigr] ^{2}+ \bigl[ d_{2}k^{2}-b_{22} \bigr] ^{2}, \\& Q_{k}=D_{k} \bigl[ d_{1}d_{2}k^{4}- ( b_{11}d_{2}+b_{22}d_{1} ) k ^{2}+ ( b_{11}+b_{22} ) \bigr] . \end{aligned}$$

For \(0< k< N_{1}\), we can conclude that (2.6) has only one positive real root

$$ \omega_{k}=\sqrt{\frac{-P_{k}+\sqrt{P_{k}^{2}-4Q_{k}}}{2}},\quad k \in \{0,1,2, \ldots,N_{1}\}. $$
(2.7)

By (2.5) we obtain

$$ \tau_{k}^{j}=\tau_{k}^{0}+ \frac{\pi j}{\omega_{k}},\qquad \tau_{k}^{0}=\frac{1}{2 \omega_{k}} \arccos \frac{-\omega_{k}^{2}+d_{1}d_{2}k^{4}- ( b_{11}d _{2}+b_{22}d_{1} ) k^{2}+b_{11}b_{22}}{b_{12}b_{21}}, $$
(2.8)

for \(k\in \{ 0,1,2,\ldots,N_{1} \} \).

Lemma 2.1

If condition \((H)\) hold, then

$$\begin{aligned} \tau_{N_{1}}^{j}\geq \tau_{k+1}^{j}\geq \tau_{k}^{j}\geq \cdots \geq \tau_{1}^{j} \geq \tau_{0}^{j} \end{aligned}$$

for \(j\in N_{0}\).

Lemma 2.2

Let condition \((H)\) hold, \(T_{k}>0\) and \(D_{k}>0\) for \(k\in N_{0}\), and let \(\omega_{k}\) and \(\tau_{k}^{j}\) be defined by (2.7) and (2.8), respectively. Then (2.2) has a pair of purely imaginary roots \(i\omega_{k}\) for each \(k\in \{ 0,1,2, \ldots,N_{1} \} \), and (2.2) has no purely imaginary roots for \(k\geq N_{1}+1\).

Clearly, \(\tau_{k}^{0}=\min_{j\in N_{0}} \{ \tau_{k}^{j} \}\), \(k \in \{0,1,2,\ldots,N_{1}\}\), and from Lemmas 2.1 and 2.2 we know that \(\tau_{0}^{0}=\min \{ \tau_{k}^{j}: 0\leq k\leq N_{1}, j\in N_{0} \} \). Denote \(\tau^{*}=\tau_{0}^{0}\). Let \(\lambda ( \tau)=\alpha (\tau)+i\delta (\tau)\) be the pair of roots of (2.2) near \(\tau =\tau_{k}^{j}\) satisfying \(\alpha ( \tau_{k}^{j} ) =0\) and \(\delta ( \tau_{k}^{j} ) = \omega_{k}\). Then we have the following transversality condition.

Lemma 2.3

For \(k\in \{0,1,2,\ldots,N_{1}\}\) and \(j\in N_{0}\), \(\frac{d \operatorname{Re}(\lambda)}{d \tau }|_{\tau =\tau_{k}^{j}}>0\).

Proof

Differentiating two sides of (2.2), we get

$$\operatorname{Re}\biggl( \frac{d\lambda }{d\tau } \biggr) ^{-1}=\operatorname{Re}\biggl[ \frac{(2\lambda +T_{k})e^{2\lambda \tau }}{-2a_{12}a_{21}\lambda }-\frac{\tau }{2 \lambda } \biggr] . $$

Thus, by (2.5) and (2.7) we have

$$\begin{aligned} \operatorname{Re}\biggl( \biggl( \frac{d\lambda }{d\tau } \biggr) ^{-1} \biggr) \bigg|_{\tau =\tau _{k}^{j}} =&\operatorname{Re}\biggl[ \frac{(2\lambda +T_{k})e^{2\lambda \tau }}{-a _{12}a_{21}\lambda }-\frac{\tau }{\lambda } \biggr] \bigg|_{\tau =\tau_{k} ^{j}} \\ =&\operatorname{Re}\biggl[ \frac{(2i\omega_{k}+T_{k})e^{2i\omega_{k}\tau_{k}^{j}}}{-ia _{12}a_{21}\omega_{k}}-\frac{\tau_{k}^{j}}{i\omega_{k}} \biggr] \\ =&\frac{T_{k}\sin 2\omega_{k}\tau_{k}^{j}-2\omega_{k}\cos 2\omega _{k}\tau_{k}^{j}}{a_{12}a_{21}\omega_{k}}=\frac{\omega_{k}^{2}+P_{k}}{ ( a_{12}a_{21} ) ^{2}}>0. \end{aligned}$$

From Lemmas 2.12.3 and from the qualitative theory of partial differential equations we obtain the following results. □

Theorem 2.1

Assume that condition \((H)\) holds and \(\omega_{k}\) and \(\tau_{k}^{j}\) are defined by (2.7) and (2.8), respectively. Denote the minimum of the critical values of delay by \(\tau^{*}=\min_{k,j} \{ \tau_{k}^{j} \} \).

  1. (a)

    The positive equilibrium \(( u^{*},v^{*} ) \) of system (1.2) is asymptotically stable for \(\tau \in ( 0,\tau^{*} ) \) and unstable for \(( \tau^{*},+\infty ) \);

  2. (b)

    System (1.2) undergoes Hopf bifurcations near the positive equilibrium \(( u^{*},v^{*} ) \) at \(\tau_{k}^{j}\) for \(k\in \{ 0, 1, 2, \ldots, N_{1} \} \) and \(j\in N_{0}\).

3 Direction of Hopf bifurcation and stability of bifurcating periodic solution

From Theorem 2.1 we know that model (1.2) undergoes Hopf bifurcations near \(E^{*}\) at \(\tau =\tau_{k}^{j}\) and that a family homogeneous and inhomogeneous periodic solutions bifurcate from \(E^{*}\) of (1.2).

In this section, we investigate the stability of these Hopf bifurcations and bifurcating direction by using the normal formal theory of partial differential equation [2123]. Without loss of generality, denote any of these critical values by \(\tau_{0}\), at which the characteristic equation (1.2) has a pair of simple purely imaginary roots \(i\omega _{0}\).

Setting \(\tilde{u}(\cdot,t)=u(\cdot,\tau t)-u^{*}\), \(\tilde{v}(\cdot,t)=v( \cdot,\tau t)-v^{*}\), \(\tilde{U}(t)= ( \tilde{u}(\cdot,\tau t), \tilde{v}(\cdot,\tau t) ) \), and \(\mathcal{C}= C([-1,0],X)\), \(X= \{ (u,v)\in W^{2,2}(0,\pi)|\frac{\partial u}{\partial x}=\frac{ \partial v}{\partial x}=0 \mbox{ at } x=0,\pi \} \) and then dropping the tildes for simplification of notation, system (1.2) can be written as the equation in the space \(\mathcal{C}\):

$$ \frac{dU(t)}{dt}=\tau D_{0}\Delta U(t)+L(\tau) (U_{t})+ \tilde{G}(U _{t},\tau), $$
(3.1)

where for \(\varphi = ( \varphi_{1},\varphi_{2} ) ^{T}\in \mathcal{C}\), \(L(\mu)(\cdot):\mathcal{C}\rightarrow X \) and \(\tilde{G}:\mathcal{C}\times R\rightarrow X\) are given by

D 0 Δ U = ( d 1 Δ u d 2 Δ v ) , L ( τ ) φ = ( b 11 φ 1 ( 0 ) b 12 φ 2 ( 1 ) b 21 φ 1 ( 1 ) b 22 φ 2 ( 0 ) ) , G ˜ ( φ , ε ) = τ ( g ( 1 ) ( τ ) g ( 2 ) ( τ ) ) = ( i + j + k + l 2 1 i ! j ! k ! l ! g i j k l ( 1 ) φ 1 i ( 0 ) φ 2 j ( 0 ) φ 1 k ( 1 ) φ 2 l ( 1 ) i + j + k + l 2 1 i ! j ! k ! l ! g i j k l ( 2 ) φ 1 i ( 0 ) φ 2 j ( 0 ) φ 1 k ( 1 ) φ 2 l ( 1 ) )

with \(g_{ijkl}^{(n)}=\frac{\partial^{i+j+k+l}\tilde{g}^{(n)} ( u ^{*},v^{*},u^{*},v^{*} ) }{\partial u^{i} \partial v^{j} \partial w^{k}\partial s^{l}}\), \(n=1,2\), and

$$\begin{aligned}& \tilde{g}^{(1)} ( u,v,w,s ) =\frac{ ( u+u^{*} ) ( 1- ( u+u^{*} ) ) }{1+c ( u+u^{*} ) }-\frac{\beta ( u+u^{*} ) ( s+v^{*} ) }{\alpha ( s+v^{*} ) + ( u+u^{*} ) }, \\& \tilde{g}^{(2)} ( u,v,w,s ) =\delta \bigl( v+v^{*} \bigr) \biggl( 1-\frac{\eta ( v+v^{*} ) }{ ( w+u^{*} ) } \biggr). \end{aligned}$$

By direct computation we obtain \(g_{0200}=g_{0002}=g_{1100}=g_{1010}=g _{0011}=0\).

Setting \(\tau =\tau_{0}+\varepsilon \), (3.1) is written as

$$ \frac{dU(t)}{dt}=\tau_{0} D_{0}\Delta U(t)+L( \tau_{0}) (U_{t})+ \tilde{G}(U_{t}, \varepsilon), $$
(3.2)

where

$$\begin{aligned} \tilde{G}(U_{t},\varepsilon)=\varepsilon D_{0}\Delta \varphi (0)+L( \varepsilon) (\varphi)+g(\varphi,\tau_{0}+\varepsilon) \quad \text{for } \varphi \in \mathcal{C}. \end{aligned}$$

So, \(\varepsilon =0\) is the Hopf bifurcation value for (3.2), and \(\Lambda_{0}= \{ -i\omega_{0}\tau_{0},i\omega_{0}\tau_{0} \} \) is the set of eigenvalues on the imaginary axis of the infinitesimal generator associated with the flow of the following linearized system of (3.2) at the origin:

$$\begin{aligned} \frac{dU(t)}{dt}=\tau_{0} D_{0}\Delta U(t)+L( \tau_{0}) (U_{t}). \end{aligned}$$

The eigenvalues of \(\tau_{0}D_{0}\Delta \) on X are \(\mu^{i}_{k}=-d _{i}\tau_{0}k^{2}\), \(i=1,2\), \(k\in N_{0}\), with the corresponding normalized eigenfunctions \(\beta^{i}_{k}\), where

β k 1 = ( γ k ( x ) 0 ) , β k 2 = ( 0 γ k ( x ) ) , γ k (x)= cos k x cos k x 2 , 2 ,k N 0 .

Let \(\mathbf{B}_{k}=\operatorname{span}\{ \langle v(\cdot),\beta_{k}^{i}\rangle \beta_{k}^{i}|v \in \mathcal{C},i=1,2 \} \). Then it is easy to verify that \(L(\tau_{0})(\mathbf{B}_{k})\subset \operatorname{span}\{ \beta_{k}^{1},\beta _{k}^{2} \} \). Assume that \(z_{t}(\theta)\in \mathcal{C}=C ( [-1,0],R ^{2} ) \) and

z t T (θ) ( β k 1 β k 2 ) B k .

Then, the linear PDE restricted on \(\mathbf{B}_{k}\) is equivalent to the following FDE on \(\mathcal{C}=C ( [-1,0],R^{2} ) \):

z ˙ (t)= ( μ k 1 0 0 μ k 2 ) z(t)+L( τ 0 )( z t ).

When \(\tau =\tau_{0}\), define \(\eta (\theta)\in BV ( [-1,0],R ) \) such that

$$\begin{aligned} \mu_{k}\varphi (0)+L(\tau_{0})\varphi = \int_{-1}^{0}d\eta (\theta) \varphi (\theta) \end{aligned}$$

and the adjoint bilinear form on \(\mathcal{C}^{*}\times \mathcal{C}\), \(\mathcal{C}^{*}=C ( [0,1],R^{2*} ) \) by

$$\begin{aligned} \bigl\langle \psi (s),\phi (\theta)\bigr\rangle =\psi (0)\phi (0)- \int_{-1} ^{0} \int_{0}^{\theta }\psi (\xi -\theta)\,d\eta (\theta)\phi (\xi)\,d\xi. \end{aligned}$$

Then, for (3.2) with fixed k, the dual bases \(\Phi_{k}\) and \(\Psi_{k}\) for its the eigenspace P and its dual space \(P^{*}\) are given by

$$\begin{aligned} \Phi_{k}= \bigl( pe^{i\omega_{0}\tau_{0}\theta },\overline{p}e^{-i\omega _{0}\tau_{0}\theta } \bigr),\qquad \Psi_{k}=\operatorname{col}\bigl( q^{T}e^{-i\omega_{0} \tau_{0}s}, \overline{q}^{T}e^{i\omega_{0}\tau_{0}s} \bigr), \end{aligned}$$

where \(\langle \Phi_{k},\Psi_{k}\rangle =I_{2}\), the \(2\times 2\) identity matrix, and

p= ( p 1 p 2 ) = ( 1 i ω 0 + d 1 k 2 b 11 b 12 e i ω 0 τ 0 ) ,q= ( q 1 q 2 ) =D ( 1 i ω 0 + d 1 k 2 b 11 b 21 e i ω 0 τ 0 ) ,

with \(D= ( 1+2\tau_{0} ( i\omega_{0}+d_{1}k^{2}-b_{11} ) +\frac{ ( i\omega_{0}+d_{1}k^{2}-b_{11} ) e^{2i\omega_{0}\tau_{0}}}{b _{12}b_{21}} ) ^{-1}\).

Following the standard procedure in [21, 23], especially [22], we can obtain the following normal form on the center manifold:

z ˙ =Bz+ ( A k 1 z 1 ε A k 1 z 2 ε ) + ( A k 2 z 1 2 z 2 A k 2 z 1 z 2 2 ) +O ( | z | ε 2 + | z | 4 ) ,
(3.3)

where

$$\begin{aligned}& A_{k1}=-k^{2} ( d_{1}q_{1}p_{1}+d_{2}q_{2}p_{2} ) +i\omega _{0}q^{T}p, \\& A_{k2}=\frac{i}{2\omega_{0}\tau_{0}} \biggl( B_{k20}B_{k11}-2 \vert B_{k11} \vert ^{2}- \frac{1}{3} \vert B_{k02} \vert ^{2} \biggr) +\frac{1}{2} ( B_{k21}+D_{k21} ), \end{aligned}$$

with

$$\begin{aligned}& B_{k20}=\textstyle\begin{cases} \frac{\tau_{0}}{\sqrt{\pi }} ( c_{1}q_{1}+c_{2}q_{2} ), &k=0, \\ 0, & k\neq 0, \end{cases}\displaystyle \qquad B_{k11}=\textstyle\begin{cases} \frac{\tau_{0}}{\sqrt{\pi }} ( c_{3}q_{1}+c_{4}q_{2} ),& k=0, \\ 0, & k\neq 0. \end{cases}\displaystyle \\& B_{k02}= \textstyle\begin{cases} \frac{\tau_{0}}{\sqrt{\pi }} ( \overline{c_{1}}q_{1}+\overline{c _{2}}q_{2} ),& k=0, \\ 0, & k\neq 0, \end{cases}\displaystyle \qquad B_{k21}= \textstyle\begin{cases} \frac{\tau_{0}}{\pi }c_{5}, & k=0, \\ \frac{3\tau_{0}}{2\pi }c_{5},& k\neq 0, \end{cases}\displaystyle \end{aligned}$$

where

$$\begin{aligned}& c_{1} =g_{2000}^{(1)}p_{1}^{2}+2g_{1001}^{(1)}p_{1}p_{2}e^{-i\omega _{0}\tau_{0}}, \\& c_{2}=g_{0200}^{(2)}p_{2}^{2}+2g_{0110}^{(2)}p_{1}p _{2}e^{-i\omega_{0}\tau_{0}}+g_{0020}^{(2)}p_{1}^{2}e^{-2i\omega_{0} \tau_{0}}, \\& c_{3} =g_{0200}^{(1)} \vert p_{1} \vert ^{2}+2g_{1001}^{(1)}\operatorname{Re}\bigl\{ p_{1} \overline{p _{2}}e^{i\omega_{0}\tau_{0}} \bigr\} , \\& c_{4}=g_{0200}^{(2)} \vert p_{2} \vert ^{2}+2g _{0110}^{(2)}\operatorname{Re}\bigl\{ p_{1} \overline{p_{2}}e^{-i\omega_{0}\tau_{0}} \bigr\} +g _{0020}^{(2)} \vert p_{1} \vert ^{2}, \\& c_{5} =q_{1} \bigl( g_{3000}^{(1)}p_{1} \vert p_{1} \vert ^{2}+g_{2001}^{(1)} \bigl( p _{1}^{2}\overline{p_{2}}e^{i\omega_{0}\tau_{0}}+2 \vert p_{1} \vert ^{2}p_{2}e ^{-i\omega_{0}\tau_{0}} \bigr) \bigr) +q_{2} \bigl( g_{0210}^{(2)} \bigl( 2p _{1} \vert p_{2} \vert ^{2}e^{-i\omega_{0}\tau_{0}} \\& \hphantom{c_{5}=} {}+p_{2}^{2} \overline{p_{1}}e^{i\omega_{0}\tau_{0}} \bigr) +g _{0120}^{(2)} \bigl( p_{1}^{2}\overline{p_{2}}e^{-2i\omega_{0}\tau_{0}}+2 \vert p _{1} \vert ^{2}p_{2} \bigr) +f_{0030}^{(2)}p_{1} \vert p_{1} \vert ^{2}e^{-i\omega_{0} \tau_{0}} \bigr), \end{aligned}$$

and

$$\begin{aligned} D_{k21}= \textstyle\begin{cases} E_{0}, & k=0, \\ E_{0}+\frac{\sqrt{2}}{2}E_{2k}, &k\neq 0, \end{cases}\displaystyle \end{aligned}$$

where

E j = 2 τ 0 π ( S 1 h j 11 ( 1 ) ( 0 ) + S 1 h j 20 ( 1 ) ( 0 ) + S 2 h j 11 ( 2 ) ( 1 ) + S 2 h j 20 ( 2 ) ( 1 ) + S 3 h j 11 ( 2 ) ( 0 ) + S 3 h j 20 ( 2 ) ( 0 ) S 4 h j 11 ( 1 ) ( 1 ) + S 4 h j 20 ( 1 ) ( 1 ) + S 5 h j 11 ( 2 ) ( 0 ) + S 5 h j 20 ( 2 ) ( 0 ) )

with

$$\begin{aligned}& S_{1}=g_{2000}^{(1)}p_{1},\qquad S_{2}=g_{1001}^{(1)}p_{1},\qquad S_{3}=g_{1001} ^{(2)}p_{2}e^{-i\omega_{0}\tau_{0}}, \\& S_{4}=g_{0110}^{(2)}p_{2}+g_{0020}^{(2)}p_{1}e^{-i\omega_{0}\tau _{0}}, \qquad S_{5}=g_{0200}^{(2)}p_{2}+g_{0110}^{(2)}p_{1}e^{-i\omega_{0} \tau_{0}}, \end{aligned}$$

where \(h_{k20}(\theta)\) and \(h_{k11}(\theta)\) are determined by

$$\begin{aligned}& \textstyle\begin{cases} \dot{h}_{k20}(\theta)-2i\omega_{0}\tau_{0}h_{k20}(\theta)=\Phi_{k}( \theta) \big({\scriptsize\begin{matrix}{} B_{k20} \cr \overline{B_{k20}} \end{matrix}} \big) , \\ \dot{h}_{k20}(0)-L(\tau_{0})(h_{k20})=\tau_{0}c_{kj} \big({\scriptsize\begin{matrix}{} c_{1} \cr c_{2} \end{matrix}} \big) , \end{cases}\displaystyle \\& \textstyle\begin{cases} \dot{h}_{k11}(\theta)=2\Phi_{k}(\theta) \big({\scriptsize\begin{matrix}{} B_{k11} \cr \overline{B_{k11}} \end{matrix}} \big) , \\ \dot{h}_{k11}(0)-L(\tau_{0})(h_{k11})=2\tau_{0}c_{kj} \big({\scriptsize\begin{matrix}{} c_{1} \cr c_{2} \end{matrix}} \big) , \end{cases}\displaystyle \end{aligned}$$

with

$$\begin{aligned} c_{kj}=\textstyle\begin{cases} \frac{1}{\sqrt{\pi }}, & j=k=0, \\ \frac{1}{\sqrt{\pi }} , & j=0,k\neq 0, \\ \frac{1}{\sqrt{2\pi }}, & j=2k\neq 0, \\ 0 & \text{otherwise}. \end{cases}\displaystyle \end{aligned}$$

Through the change of variables \(z_{1}=\omega_{1}-i\omega_{2}\), \(z_{2}= \omega_{1}+i\omega_{2}\), and \(\omega_{1}=\rho_{1}\cos \phi\), \(\omega _{2}=\rho_{1}\sin \phi \), the normal form [20] becomes the following polar coordinate system:

$$\begin{aligned} \textstyle\begin{cases} \dot{\rho }=\kappa_{k1}\alpha \rho +\kappa_{k2}\rho^{3}+O ( \alpha ^{2}\rho + \vert (\rho,\alpha) \vert ^{4} ), \\ \dot{\phi }=-\omega_{0}\tau_{0}+O ( \vert (\rho,\alpha) \vert ), \end{cases}\displaystyle \end{aligned}$$

where \(\kappa_{k1}=\operatorname{Re}A_{k1}\) and \(\kappa_{k2}=\operatorname{Re}A_{k2}\). Thus, from [28] we see that the sign of \(\kappa_{k1}\kappa_{k2}\) determines the direction of the bifurcation and the sign of \(\kappa_{k2}\) determines the stability of the nontrivial periodic orbits and have following results.

  1. (a)

    When \(\kappa_{k1}\kappa_{k2}<0\), system (1.2) undergoes Hopf bifurcation at the critical value \(\tau =\tau^{*}\), which is a supercritical bifurcation. Moreover, if \(\kappa_{k2}<0\), then the bifurcating periodic solution is stable; if \(\kappa_{k2}>0\), then the bifurcating periodic solution is unstable.

  2. (b)

    When \(\kappa_{k1}\kappa_{k2}>0\), system (1.2) undergoes Hopf bifurcation at the critical value \(\tau =\tau^{*}\), which is a subcritical bifurcation. Moreover, if \(\kappa_{k2}<0\), then the bifurcating periodic solution is stable; if \(\kappa_{k2}>0\), then the bifurcating periodic solution is unstable.

4 Numerical simulations

In this section, we give some numerical simulations to support and extend our results by using the mathematical software Matlab.

For system (1.2), we choose \(d_{1}=0.2\), \(d_{2}=1\), \(\delta =2\), \(\alpha =0.05\), \(\beta =0.3\), \(h=0.8\), \(c=0.1\), and \((u(x,0),v(x,0))=(0.62+0.005\cos x,0.78+0.005 \cos x)\). Then, a series of calculations show that \((u^{*},v^{*})=(0.6250,0.7813)\) and \(\tau^{*}=6.8222\). Hence \((0.6250,0.7813)\) is locally stable when \(\tau \in [0,\tau^{*})\). When τ crosses through the critical \(\tau^{*}\), \((0.6250,0.7813)\) loses its stability, Hopf bifurcation occurs, and a family of periodic solutions are bifurcation from \((0.6250,0.7813)\). The direction and stability of Hopf bifurcation can be determined by the signs \(\kappa_{k1}\) and \(\kappa_{k2}\); by the procedure of Sect. 3, \(\kappa_{k1}=0.0150\) and \(\kappa_{k2}=1.1544\). So, the Hopf bifurcation occurring at \(\tau^{*}\) is subcritical, and the corresponding Hopf bifurcation periodic orbits are unstable. Taking \(\tau =6.5<6.8222\), the numerical simulation results of system (1.2) are depicted Fig. 1. And taking \(\tau =6.85>6.8222\), the numerical simulation results of system (1.2) are depicted Fig. 2, which is in agreement with the theoretical results.

Figure 1
figure 1

The positive equilibrium \(( u^{*},v^{*} ) =(0.6250,0.7813)\) of (1.2) is asymptotically stable when \(\tau =6.5< 6.8222\). Here we set parameter values \(d_{1}=0.2\), \(d_{2}=1\), \(\delta =2\), \(\alpha =0.05\), \(\beta =0.3\), \(h=0.8\), \(c=0.1\) and the initial value \((u(x,0),v(x,0))=(0.62+0.005\cos x,0.78+0.005\cos x)\)

Figure 2
figure 2

There exists unstable spatially homogenous periodic bifurcating from the positive equilibrium \(( u^{*},v^{*} ) =(0.6250,0.7813)\) of (1.2) when \(\tau =6.85>6.8222\). Here we set parameter values \(d_{1}=0.2\), \(d_{2}=1\), \(\delta =2\), \(\alpha =0.05\), \(\beta =0.3\), \(h=0.8\), \(c=0.1\). (A)–(B) The initial value \((u(x,0),v(x,0))=(0.62+0.05\cos x,0.78+0.05\cos x)\); (C)–(D) The initial value \((u(x,0),v(x,0))=(0.62+0.05\sin x,0.78+0.05\sin x)\)

5 Conclusions

This paper mainly consider about the effects of delay τ on dynamics of a diffusive predator-prey model with Simth growth rate under Neumann boundary conditions. Based on the analysis of the characteristic equations, we study the stability of the equilibrium and the existence of delay-induced Hopf bifurcations. Then, by the normal forms on the center manifold, we obtain the results determining the direction and stability of Hopf bifurcation. Finally, choosing the parameter values \(d_{1}=0.2\), \(d_{2}=1\), \(\delta =2\), \(\alpha =0.05\), \(\beta =0.3\), \(h=0.8\), \(c=0.1\), we get the critical value of delay \(\tau^{*}=6.8222\) with a direct computation. From Theorem 2.1 we know that the positive equilibrium \(( u^{*},v^{*} ) =(0.6250,0.7813)\) is asymptotically stable for \(\tau <\tau^{*}\) and unstable for \(\tau >\tau^{*}\). So, system (1.2) undergoes Hopf bifurcation near the positive equilibrium \(( u^{*},v ^{*} ) =(0.6250,0.7813)\) when the delay τ increasingly crosses through the critical value \(\tau^{*}\).

For the predator–prey systems, pattern dynamics is also an interesting topic. Many authors have discussed the pattern dynamics of the diffusion equations [2933]. In the next topics, we will discuss the pattern dynamics of a diffusive predator–prey model with Simth growth rate.

References

  1. Gause, G.F.: The Struggle for Existence. Hafner, New York (1969)

    Google Scholar 

  2. Holling, C.: The functional response of predator to prey density and its role in mimicry and population regulation. Mem. Entomol. Soc. Can. 45, 1–60 (1965)

    Google Scholar 

  3. Tanner, J.T.: The stability and the intrinsic growth rates of prey and predator populations. Ecology 56, 855–886 (1975)

    Article  Google Scholar 

  4. Wollkind, D.J., Collings, J.B., Logan, J.A.: Metastability in a temperature dependent model system for trees. Bull. Math. Biol. 50, 379–409 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  5. May, R.M.: Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton (1973)

    Google Scholar 

  6. Li, X., Jiang, W.: Hopf bifurcation and Turing instability in the reaction–diffusion Holling–Tanner predator–prey model. IMA J. Appl. Math. 78, 287–306 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  7. Murray, J.D.: Mathematical Biology II. Springer, Heidelberg (2002)

    Google Scholar 

  8. Shi, H., Ruan, S.: Spatial, temporal and spatiotemporal patterns of diffusive predator–prey models with mutual interference. IMA J. Appl. Math. 80, 1534–1568 (2015). https://doi.org/10.1093/imamat/hxv006

    Article  MathSciNet  MATH  Google Scholar 

  9. Yang, R., Wei, J.: Stability and bifurcation analysis of a diffusive prey–predator system in Holling type III with a prey refuge. Nonlinear Dyn. 79, 631–646 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  10. Yi, F., Wei, J., Shi, J.: Bifurcation and spatio-temporal patterns in a homogeneous diffusive predator–prey system. J. Differ. Equ. 246, 1944–1977 (2009)

    Article  MATH  Google Scholar 

  11. Wang, J., Shi, J., Wei, J.: Dynamics and pattern formation in a diffusive predator–prey system with strong Allee effect in prey. J. Differ. Equ. 251, 1276–1304 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  12. Fan, M., Wang, K.: Periodicity in a food-limited population model with toxicants and time delays. Acta Math. Appl. Sin. 18, 309–314 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  13. Gopalsamy, K., Kulenovic, M.R.S., Ladas, G.: Environmental periodicity and time delays in a food-limited population model. J. Math. Anal. Appl. 147, 545–555 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  14. Smith, F.E.: Population dynamics in Daphnia Magna and a new model for population growth. Ecology 44, 651–663 (1963)

    Article  Google Scholar 

  15. Sivakumar, M., Sambath, M., Balachandran, K.: Stability and Hopf bifurcation analysis of a diffusive predator–prey model with Smith growth. Int. J. Biomath. 8(1), 1550013 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  16. Yue, Z., Wang, W.: Qualitative analysis of a diffusive ratio-dependent Holling–Tanner predator–prey model with Smith growth. Discrete Dyn. Nat. Soc. 2013, Article ID 267173 (2013). https://doi.org/10.1155/2013/267173

    Article  MathSciNet  MATH  Google Scholar 

  17. Ruan, S.: On nonlinear dynamics of predator–prey models with discrete delay. Math. Model. Nat. Phenom. 4(2), 140–188 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  18. Cao, J., Yuan, R.: Bifurcation analysis in a modified Lesile–Gower model with Holling-type II functional response and delay. Nonlinear Dyn. 84(3), 1341–1352 (2016). https://doi.org/10.1007/s11071-015-2572-5

    Article  MathSciNet  MATH  Google Scholar 

  19. Lian, F., Xu, Y.: Hopf bifurcation analysis of a predator–prey system with Holling-type IV functional response and time delay. Appl. Math. Comput. 215, 1484–1495 (2009)

    MathSciNet  MATH  Google Scholar 

  20. Chen, S., Shi, J., Wei, J.: Global stability and Hopf bifurcation in a delayed diffusive Leslie–Gower predator–prey system. Int. J. Bifurc. Chaos 22, 1250061 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  21. Faria, T.: Stability and bifurcation for a delayed predator–prey model and the effect of diffusion. J. Math. Anal. Appl. 254, 433–463 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  22. Faria, T.: Normal forms and Hopf bifurcation for partial differential equations with delay. Trans. Am. Math. Soc. 352, 2217–2238 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  23. Song, Y., Peng, Y., Zou, X.: Persistence, stability and Hopf bifurcation in a diffusive ratio-dependent predator–prey model with delay. Int. J. Bifurc. Chaos 24, 1450093 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  24. Sun, G., Wang, S., Ren, Q.: Effects of time delay and space on herbivore dynamics: linking inducible defenses of plants to herbivore outbreak. Sci. Rep. 5, 11246 (2015)

    Article  Google Scholar 

  25. Li, L., Jin, Z., Li, J.: Periodic solutions in a herbivore–plant system with time delay and spatial diffusion. Appl. Math. Model. 40(7–8), 4765–4777 (2016)

    Article  MathSciNet  Google Scholar 

  26. Ma, Z., Huo, H., Xiang, H.: Hopf bifurcation for a delayed predator–prey diffusion system with Dirichlet boundary condition. Appl. Math. Comput. 311, 1–18 (2017)

    Article  MathSciNet  Google Scholar 

  27. Tang, H., Liu, Z.: Hopf bifurcation for a predator–prey model with age structure. Appl. Math. Model. 40(2), 726–737 (2016)

    Article  MathSciNet  Google Scholar 

  28. Chow, S.N., Hale, J.K.: Methods of Bifurcation Theory. Springer, New York (1982)

    Book  MATH  Google Scholar 

  29. Guin, L.N.: Existence of spatial patterns in a predator–prey model with self- and cross-diffusion. Appl. Math. Comput. 226, 320–335 (2014)

    MathSciNet  MATH  Google Scholar 

  30. Sun, G., Wu, Z., Wang, Z., Jin, Z.: Influence of isolation degree of spatial patterns on persistence of populations. Nonlinear Dyn. 83(1–2), 811–819 (2016)

    Article  MathSciNet  Google Scholar 

  31. Sun, G.: Mathematical modeling of population dynamics with Allee effect. Nonlinear Dyn. 85(1), 1–12 (2016)

    Article  MathSciNet  Google Scholar 

  32. Sun, G., Wang, C., Wu, Z.: Pattern dynamics of a Gierer–Meinhardt model with spatial effects. Nonlinear Dyn. 88(2), 1385–1396 (2017)

    Article  Google Scholar 

  33. Peng, Y., Ling, H.: Pattern formation in a ratio-dependent predator–prey model with cross-diffusion. Appl. Math. Comput. 331, 307–318 (2018)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors are grateful to the referee for carefully reading the manuscript and for providing some comments and suggestions which led to improvements in the paper.

Funding

The research of H. Fang was supported by the key project of Provincial Excellent Talents in University of Anhui Province (gxyqZD2016300, gxyqZD2018077) and the Natural Science Foundation of Huangshan University (2017xkjq001). The research of L. Hu was supported by the National Natural Science Foundation of China (11701208). The research of Y. Wu was supported by the Natural Science Foundation of Anhui Province (1708085MA04) and the Key NSF of Anhui Educational Committee (KJ2018A0428).

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed equally and read and approved the final manuscript.

Corresponding author

Correspondence to Yongfeng Wu.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fang, H., Hu, L. & Wu, Y. Delay-induced Hopf bifurcation in a diffusive Holling–Tanner predator–prey model with ratio-dependent response and Smith growth. Adv Differ Equ 2018, 285 (2018). https://doi.org/10.1186/s13662-018-1726-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13662-018-1726-4

MSC

Keywords