Skip to main content
Log in

Hopf Bifurcation Analysis of a Predator–Prey Model with Prey Refuge and Fear Effect Under Non-diffusion and Diffusion

  • Published:
Qualitative Theory of Dynamical Systems Aims and scope Submit manuscript

Abstract

In this paper, we propose a predator–prey model with prey refuge and fear effect under non-diffusion and diffusion. For the non-diffusion ODE model, we first analyze the existence and stability of equilibria. Then, the existence of transcritical bifurcation, Hopf bifurcation and limit cycle is discussed, respectively. We find that when the cost of minimum fear \(\eta \) is taken as the bifurcation parameter, it not only influence the occurrence of Hopf bifurcation but also alters its direction. For diffusion predator–prey model under homogeneous Neumann boundary conditions, we observe that the Turing instability does not occur, but the Hopf bifurcation will manifest near the interior equilibrium. By considering \(\eta \) as the bifurcation parameter, the direction and stability of spatially homogeneous periodic orbits are established. At last, the validity of the theoretical analysis are verified by a series of numerical simulations. The results indicate that prey refuge and fear effect play an key role in the stability of populations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Data Availability

All data generated or analysed during this study are included in this published article.

References

  1. Ritchie, E.G., Johnson, C.N.: Predator interactions, mesopredator release and biodiversity conservation. Ecol. Lett. 12(9), 982–998 (2009)

    Google Scholar 

  2. Holling, C.S.: The functional response of invertebrate predators to prey density. Mem. Entomol. Soc. Can. 98, 5–86 (1966)

    Google Scholar 

  3. Beddington, J.R.: Mutual interference between parasites or predators and its effect on searching efficiency. J. Anim. Ecol. 44, 331–340 (1975)

    Google Scholar 

  4. DeAngelis, D.L., Goldstein, R.A., Oneill, R.V.: A model for tropic interaction. Ecology 56, 881–892 (1975)

    Google Scholar 

  5. Abrams, P.A., Ginzburg, L.R.: The nature of predation: prey dependent, ratio dependent or neither? Trends Ecol. Evol. 15(8), 337–341 (2000)

    Google Scholar 

  6. Ivlev, V.S.: Experimental ecology of the feeding of fishes, University Microfilms, (1961)

  7. Hassell, M.P., Varley, G.C.: New inductive population model for insect parasites and its bearing on biological control. Nature 223, 1133–1137 (1969)

    Google Scholar 

  8. Hu, D., Li, Y., Liu, M., Bai, Y.: Stability and Hopf bifurcation for a delayed predator-prey model with stage structure for prey and Ivlev-type functional response. Nonlinear Dynam. 99, 3323–3350 (2020)

    MATH  Google Scholar 

  9. Hu, D., Cao, H.: Stability and bifurcation analysis in a predator-prey system with Michaelis–Menten type predator harvesting. Nonlinear Anal-Real. 33, 58–82 (2017)

    MathSciNet  MATH  Google Scholar 

  10. Hu, D., Cao, H.: Bifurcation and chaos in a discrete-time predator-prey system of Holling and Leslie type. Commun. Nonlinear Sci. 22(1–3), 702–715 (2015)

    MathSciNet  MATH  Google Scholar 

  11. Hu, D., Yu, X., Zheng, Z., Zhang, C., Liu, M.: Multiple bifurcations in a discrete bazykin predator-prey model with predator intraspecific interactions and ratio-dependent functional response. Qual. Theor. Dyn. Syst. 22(3), 99 (2023)

    MathSciNet  MATH  Google Scholar 

  12. Zhang, W., Zhao, S., Meng, X., Zhang, T.: Evolutionary analysis of adaptive dynamics model under variation of noise environment. Appl. Math. Model. 84, 222–239 (2020)

    MathSciNet  MATH  Google Scholar 

  13. Xiang, C., Huang, J., Ruan, S., Xiao, D.: Bifurcation analysis in a host-generalist parasitoid model with Holling II functional response. J. Differ. Equations 268(8), 4618–4662 (2020)

    MathSciNet  MATH  Google Scholar 

  14. Qi, H., Meng, X.: Threshold behavior of a stochastic predatorCprey system with prey refuge and fear effect. Appl. Math. Lett. 113, 106846 (2021)

    MATH  Google Scholar 

  15. Qi, H., Meng, X., Hayat, T., Hobiny, A.: Stationary distribution of a stochastic predator-prey model with hunting cooperation. Appl. Math. Lett. 124, 107662 (2022)

    MathSciNet  MATH  Google Scholar 

  16. Liu, Q., Jiang, D.: Influence of the fear factor on the dynamics of a stochastic predator-prey model. Appl. Math. Lett. 112, 106756 (2021)

    MathSciNet  MATH  Google Scholar 

  17. Zhang, H., Cai, Y., Fu, S., Wang, W.: Impact of the fear effect in a prey-predator model incorporating a prey refuge. Appl. Math. Comput. 356, 328–337 (2019)

    MathSciNet  MATH  Google Scholar 

  18. McNair, J.N.: Stability effects of prey refuges with entry-exit dynamics. J. Theor. Biol. 125, 449–464 (1987)

    MathSciNet  Google Scholar 

  19. Kar, T.K.: Stability analysis of a prey-predator model incorporating a prey refuge. Commun. Nonlinear Sci. Numer. Simulat. 10, 681–691 (2005)

    MathSciNet  MATH  Google Scholar 

  20. Chen, L., Chen, F., Chen, L.: Qualitative analysis of a predator-prey model with Holling type II functional response incorporating a constant prey refuge. Nonlinear Anal-Real. 11(1), 246–252 (2010)

    MathSciNet  MATH  Google Scholar 

  21. Ma, Z., Wang, S., Li, W., Li, Z.: The effect of prey refuge in a patchy predator-prey system. Math. Biosci. 243(1), 126–130 (2013)

    MathSciNet  MATH  Google Scholar 

  22. Zhou, Y., Sun, W., Song, Y., Zheng, Z., Lu, J., Chen, S.: Hopf bifurcation analysis of a predator-prey model with Holling-II type functional response and a prey refuge. Nonlinear Dynam. 97(2), 1439–1450 (2019)

    MATH  Google Scholar 

  23. Chang, X., Wei, J.: Stability and Hopf bifurcation in a diffusive predator-prey system incorporating a prey refuge. Math. Biosci. Eng. 10(4), 979–996 (2013)

    MathSciNet  MATH  Google Scholar 

  24. Altendorf, K.B., Laundré, J.W., González, C.A.L., Brown, J.S.: Assessing effects of predation risk on foraging behavior of mule deer. J. Mammal. 82(2), 430–439 (2001)

    Google Scholar 

  25. Laundré, J.W., Hernández, L., Altendorf, K.B.: Wolves, elk, and bison: reestablishing the landscape of fear in yellowstone National Park, USA. Can. J. Zool. 79(8), 1401–1409 (2001)

    Google Scholar 

  26. Creel, S., Christianson, D., Liley, S., Winnie, J.A.: Predation risk affects reproductive physiology and demography of elk. Science 315, 960–960 (2007)

    Google Scholar 

  27. Zanette, L.Y., White, A.F., Allen, M.C., Clinchy, M.: Perceived predation risk reduces the number of offspring songbirds produce per year. Science 334(6061), 1398–1401 (2011)

    Google Scholar 

  28. Cresswell, W.: Predation in bird populations. J. Ornithol. 152(1), 251–263 (2011)

    Google Scholar 

  29. Wang, X., Zanette, L., Zou, X.: Modelling the fear effect in predator-prey interactions. J. Math. Biol. 73, 1179–1204 (2016)

    MathSciNet  MATH  Google Scholar 

  30. Wang, X., Zou, X.: Modeling the fear effect in predator-prey interactions with adaptive avoidance of predators. Bull. Math. Biol. 79(6), 1325–1359 (2017)

    MathSciNet  MATH  Google Scholar 

  31. Sarkara, K., Khajanchi, S.: Impact of fear effect on the growth of prey in a predator-prey interaction model. Ecol. Complex. 42, 100826 (2020)

    Google Scholar 

  32. Sasmal, S.K., Takeuchi, Y.: Dynamics of a predator-prey system with fear and group defense. J. Math. Anal. Appl. 481(11), 123471 (2020)

    MathSciNet  MATH  Google Scholar 

  33. Roy, J., Alam, S.: Fear factor in a prey-predator system in deterministic and stochastic environment. Physica A. 541, 123359 (2020)

    MathSciNet  MATH  Google Scholar 

  34. Ko, W., Ryu, K.: Qualitative analysis of a predator-prey model with Holling type II functional response incorporating a prey refuge. J. Differ. Equations 231(2), 534–550 (2006)

    MathSciNet  MATH  Google Scholar 

  35. Peng, R., Shi, J.: Non-existence of non-constant positive steady states of two Holling type-II predator-prey systems: strong interaction case. J. Differ. Equations 247, 866–886 (2009)

    MathSciNet  MATH  Google Scholar 

  36. Li, C., Wang, X., Shao, Y.: Steady states of a predator-prey model with prey-taxis. Nonlinear Anal. 97, 155–168 (2014)

    MathSciNet  MATH  Google Scholar 

  37. Qiao, T., Cai, Y., Fu, S., Wang, W.: Stability and hopf bifurcation in a predator-prey model with the cost of anti-predator behaviors. Int. J. Bifurcat. Chaos 29(13), 1950185 (2019)

    MathSciNet  MATH  Google Scholar 

  38. Song, Y., Jiang, H., Yuan, Y.: Turing-Hopf bifurcation in the reaction-diffusion system with delay and application to a diffusive predator-prey model. J. Appl. Anal. Comput. 9(3), 1132–1164 (2019)

    MathSciNet  MATH  Google Scholar 

  39. Upadhyay, R.K., Roy, P., Datta, J.: Complex dynamics of ecological systems under nonlinear harvesting: Hopf bifurcation and turing instability. Nonlinear Dynam. 79, 2251–2270 (2015)

    MathSciNet  MATH  Google Scholar 

  40. Tiwari, V., Tripathi, J.P., Mishra, S., Upadhyay, R.K.: Modeling the fear effect and stability of non-equilibrium patterns in mutually interfering predator-prey systems. Appl. Math. Comput. 371, 124948 (2020)

    MathSciNet  MATH  Google Scholar 

  41. Song, D., Li, C., Song, Y.: Stability and cross-diffusion-driven instability in a diffusive predator-prey system with hunting cooperation functional response. Nonlinear Anal-Real. 54, 103106 (2020)

    MathSciNet  MATH  Google Scholar 

  42. Li, Q., Liu, Z., Yuan, S.: Cross-diffusion induced Turing instability for a competition model with saturation effec. Appl. Math. Comput. 347, 64–77 (2019)

    MathSciNet  MATH  Google Scholar 

  43. Yan, S., Jia, D., Zhang, T., Yuan, S.: Pattern dynamics in a diffusive predator-prey model with hunting cooperations. Chaos Soliton. Fract. 130, 109428 (2020)

    MathSciNet  MATH  Google Scholar 

  44. Song, Y., Tang, X.: Stability, steady-state bifurcations, and turing patterns in a predator-prey model with herd behavior and prey-taxis. Stud. Appl. Math. 139(3), 371–404 (2017)

    MathSciNet  MATH  Google Scholar 

  45. Yang, R., Jin, D., Wang, W.: A diffusive predator-prey model with generalist predator and time delay. AIMS Math. 7(3), 4574–4591 (2022)

    MathSciNet  Google Scholar 

  46. Duan, D., Niu, B., Wei, J.: Hopf-Hopf bifurcation and chaotic attractors in a delayed diffusive predator-prey model with fear effect, Chaos. Soliton. Fract. 123, 206–216 (2019)

    MATH  Google Scholar 

  47. Qi, H., Meng, X., Hayat, T., Hobiny, A.: Bifurcation dynamics of a reaction-diffusion predator-prey model with fear effect in a predator-poisoned environment. Math. Method. Appl. Sci. 45(10), 6217–6254 (2022)

    MathSciNet  MATH  Google Scholar 

  48. Lou, Y., Ni, W.: Diffusion, self-diffusion and cross-diffusion. J. Differ. Equations 131, 79–131 (1996)

    MathSciNet  MATH  Google Scholar 

  49. Yang, R., Nie, C., Jin, D.: Spatiotemporal dynamics induced by nonlocal competition in a diffusive predator-prey system with habitat complexity. Nonlinear Dynam. 110, 879–900 (2022)

    Google Scholar 

  50. Yang, R., Wang, F., Jin, D.: Spatially inhomogeneous bifurcating periodic solutions induced by nonlocal competition in a predator-prey system with additional food. Math. Method. Appl. Sci. 45(16), 9967–9978 (2022)

    MathSciNet  Google Scholar 

  51. Sotomayor, J.: Generic bifurcation of dynamical system. Dynam Syst (1973). https://doi.org/10.1016/B978-0-12-550350-1.50047-3

    Article  MATH  Google Scholar 

  52. Meiss JD.: Differential Dynamical Systems. Society for Industrial and Applied Mathematics, Philadelphia (2007)

  53. Wiggins, S., Golubitsky, M.: Introduction to applied nonlinear dynamical systems and chaos, vol. 2. Springer, Berlin (1990)

    MATH  Google Scholar 

  54. Hassard, B.D., Kazarinoff, N.D., Wan, Y.: Theory and applications of Hopf Bifurcation. Cambridge University Press, Cambridge (1981)

    MATH  Google Scholar 

  55. Garvie, M.R.: Finite-difference schemes for reaction-diffusion equations modeling predator-prey interactions in MATLAB. Bull. Math. Biol. 69(3), 931–956 (2007)

    MathSciNet  MATH  Google Scholar 

  56. Burton, T.A.: Volterra integral and differential equations. Academic Press, Inc., Orlando (1983)

    MATH  Google Scholar 

Download references

Acknowledgements

This work was supported by the Doctor Research Project Foundation of Liaoning Province of China (No. 2023-BS-210).

Author information

Authors and Affiliations

Authors

Contributions

HZ and HQ wrote the main manuscript text. HQ prepared all figures. All authors reviewed the manuscript.

Corresponding author

Correspondence to Haokun Qi.

Ethics declarations

Conflicts of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

Appendices

Appendix A

Proof of Theorem 2.1

From model (2), denote

$$\begin{aligned} \left\{ \begin{aligned} H_1(u,v)=&ru\left( \eta +\frac{\alpha (1-\eta ) }{\alpha +v}\right) -\mu _1 u-\gamma u^2-\frac{\beta (1-\delta )uv}{1+a(1-\delta ) u},\\ H_2(u,v)=&\frac{\theta \beta (1-\delta ) uv}{1+a(1-\delta ) u}-\mu _2v. \end{aligned}\right. \end{aligned}$$
(A.1)

It is obvious that we have two boundary equilibrium points

$$\begin{aligned} E_0=(0,0), \quad E_1=(u_1,0)=\left( \frac{r-\mu _1}{\gamma },0\right) , \end{aligned}$$

where \(E_0\) is always existed in model (2), \(E_1\) is existed when \(R_0>1\).

For the positive equilibrium point \(E^*=(u^*,v^*)\): if conditions \(\mathbf {C.0}\) hold, \(u^*=\frac{\mu _2}{(\theta \beta -a\mu _2)(1-\delta )}\) exists and \(v^*\) is the positive root of following equation

$$\begin{aligned} v^2+B_1 v+B_2=0, \end{aligned}$$

where \(B_1= {\mu _2\theta \gamma }/{(\theta \beta -a\mu _2)^2(1-\delta )^2}+{\theta (\mu _1-r\eta )}/{(\theta \beta -a\mu _2)(1-\delta )}+\alpha \), \(B_2={\alpha \theta [\mu _2\gamma -(r-\mu _1)(\theta \beta -a\mu _2)(1-\delta )]}/{(\theta \beta -a\mu _2)^2(1-\delta )^2}\).

For \(B_2<0\), that is \(\delta ^{*}> \delta \ge 0\), the above equation has a unique positive root given by

$$\begin{aligned} v^*=\frac{-B_1+\sqrt{B_1^2-4 B_2}}{2}. \end{aligned}$$

\(\square \)

Proof of Theorem 2.2

(I) For equilibrium \(E_0\): The corresponding Jacobian matrix of the model (2) is

$$\begin{aligned} J_{E_0}:=\left( \begin{array}{cc} r-\mu _1&{}0\\ 0&{} -\mu _2 \end{array} \right) , \end{aligned}$$

which has two eigenvalues \(\lambda _1=r-\mu _1\), \(\lambda _2=-\mu _2<0\). If \(R_0<1\), then the equilibrium \(E_0\) of model (2) is locally asymptotically stable node point. But if \(R_0>1\), \(E_0\) is a saddle.

(II) For equilibrium \(E_1\): The corresponding Jacobian matrix of the model (2) is

$$\begin{aligned} J_{E_1}:=\left( \begin{array}{cc} -r+\mu _1&{} -\frac{r(1-\eta )u_1}{\alpha }-\frac{\beta (1-\delta ) u_1}{1+a(1-\delta )u_1} \\ 0&{}\frac{\theta \beta (1-\delta ) u_1}{1+a(1-\delta )u_1}-\mu _2 \end{array} \right) . \end{aligned}$$

Obviously, we have two eigenvalues \(\lambda _1=-r+\mu _1\), \(\lambda _2={\theta \beta (1-\delta ) u_1}/{1+a(1-\delta )u_1}-\mu _2\). Noting that \(\lambda _1<0\) is equivalent to \(R_0>1\) and \(\lambda _2<0\) is equivalent to

$$\begin{aligned} \begin{aligned} \mu _2>\frac{1}{\gamma }(\beta \theta -a\mu _2)(r-\mu _1)(1-\delta ). \end{aligned} \end{aligned}$$
(A.2)

When \(R_1<1\), that is \(\beta \theta <a\mu _2\), which means that the right side of the inequality (A.2) is less than zero. Then we obtain \(\lambda _2<0\). When \(R_1>1\), that is \(\beta \theta >a\mu _2\), we obtain \(1>\delta >\delta ^{*}\) by calculating the inequality (A.2). Thus, we obtain conditions for locally asymptotically stability of \(E_1\), i.e., \(R_1<1\) or \(R_1>1\) and \(1>\delta >\delta ^{*}\).

(III) For positive equilibrium \(E^*=(u^*,v^*)\): The corresponding Jacobian matrix of model (2) is

$$\begin{aligned} J_{E^*}=\left( \begin{array}{cc} a_{11}&{} a_{12}\\ a_{21}&{} a_{22} \end{array} \right) =\left( \begin{array}{cc} -\gamma u^*+\frac{a\beta (1-\delta )^2u^*v^*}{(1+a(1-\delta )u^*)^2}&{}-\left( \frac{r\alpha (1-\eta )u^*}{(\alpha +v^*)^2}+\frac{\beta (1-\delta ) u^*}{1+a(1-\delta )u^*}\right) \\ \frac{\theta \beta (1-\delta ) v^*}{(1+a(1-\delta )u^*)^2}&{} 0 \end{array} \right) .\nonumber \\ \end{aligned}$$
(A.3)

We can obtain the following characteristic equation

$$\begin{aligned} \begin{aligned} \lambda ^2-\textrm{tr}(J_{E^*})\lambda +\textrm{det}(J_{E^*})=0, \end{aligned} \end{aligned}$$
(A.4)

where

$$\begin{aligned} \left\{ \begin{aligned} \textrm{tr}(J_{E^*})=&-\gamma u^*+\frac{a\beta (1-\delta )^2u^*v^*}{(1+a(1-\delta )u^*)^2},\\ \textrm{det}(J_{E^*})=&\frac{\theta \beta (1-\delta ) v^*}{(1+a(1-\delta )u^*)^2} \left( \frac{r\alpha (1-\eta )u^*}{(\alpha +v^*)^2}+\frac{\beta (1-\delta ) u^*}{1+a(1-\delta )u^*}\right) >0. \end{aligned}\right. \nonumber \\ \end{aligned}$$
(A.5)

From (A.5), we have \(\textrm{det}(J_{E^*})>0\) always hold. And we know that \(\textrm{tr}(J_{E^*})<0\) is equivalent to

$$\begin{aligned} \begin{aligned}\alpha \left( \frac{r(1-\eta )}{\frac{\gamma (\theta \beta +a\mu _2)}{a(\theta \beta -a\mu _2)(1-\delta )}-r\eta +\mu _1}-1\right) <v^*=\frac{-B_1+\sqrt{B_1^2-4 B_2}}{2}. \end{aligned} \end{aligned}$$
(A.6)

When conditions \(\mathbf {C.0}\) and \(\delta ^{*}> \delta \ge \delta ^{**}\) are satisfied, the left side of the inequality (A.6) is less than zero, then we obtain \(\textrm{tr}(J_{E^*})<0\). If the condition \(\delta ^{**}> \delta \ge 0\) are satisfied, by calculating the inequality (A.6), we obtain \(\eta ^*>\eta \ge 0\) which can make \(\textrm{tr}(J_{E^*})<0\). Then we know that the positive equilibrium \(E^*\) is a locally asymptotically stable node for \(\textrm{tr}^2(J_{E^*})>4\textrm{det}(J_{E^*})\) and stable focus for \(\textrm{tr}^2(J_{E^*})<4\textrm{det}(J_{E^*})\).

Now, we consider the global asymptotic stability of the positive equilibrium point \(E^*\) under conditions \(\mathbf {C.0}\) and \(\delta ^{*}> \delta \ge \delta ^{**}\) of Theorem 2.2.

Denote a Dulac function \(B(u,v)=(\alpha +v)(1+a(1-\delta )u)u^{-1}v^b\), where b is to be specified later, then

$$\begin{aligned} \begin{aligned} D=&\frac{\partial (H_1(u,v)B(u,v))}{\partial u}+\frac{\partial (H_2(u,v)B(u,v))}{\partial v}\\ =&u^{-1}v^b\left( \frac{1}{\alpha }h_1(u)v+h_2(u)\right) , \end{aligned} \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} h_1(u,b)=&-2a\alpha \gamma (1-\delta )u^2+[(1-\delta )[(2+b)(\theta \beta -a\mu _2)\\&+a(r\eta -\mu _1)]-\gamma ]\alpha u-(2+b)\mu _2\alpha ,\\ h_2(u,b)=&-2a\alpha \gamma (1-\delta )u^2+[(1-\delta )[(1+b)(\theta \beta -a\mu _2)\\&+a(r-\mu _1)]-\gamma ]\alpha u-(1+b)\mu _2\alpha . \end{aligned} \end{aligned}$$

When \(1-\frac{\theta \beta -a\mu _2}{ar}\ge \eta \ge 0\) holds, we have

$$\begin{aligned} \begin{aligned} h_2(u,b)-h_1(u,b)=[ar(1-\eta )-(\theta \beta -a\mu _2)]\alpha (1-\delta )u+\mu _2\alpha \ge 0 \end{aligned} \end{aligned}$$

for \(u\in {\mathbb {R}}_+\).

Thus, we need to prove \(h_2(u,b)\le 0\) for all \(u\in {\mathbb {R}}_+\) such that \(D\le 0\) for all \((u,v)\in {\mathbb {R}}_+^2\).

Since \(a\alpha \gamma (1-\delta )>0\) always holds, then we can obtain that

$$\begin{aligned} \begin{aligned} g(b+1)=g_1(b+1)^2+g_2(b+1)+g_3\le 0 \end{aligned} \end{aligned}$$

such that \(D\le 0\) for all \((u,v)\in {\mathbb {R}}_+^2\), where

$$\begin{aligned} \begin{aligned} g_1=&\alpha ^2(\beta \theta -a\mu _2)^2(1-\delta )^2,\\ g_2=&2\alpha ^2(1-\delta )[(\beta \theta -a\mu _2)(a(r-\mu _1 )(1-\delta )-\gamma )-4a\gamma \mu _2],\\ g_3=&\alpha ^2(a(r-\mu _1 )(1-\delta )-\gamma )^2. \end{aligned} \end{aligned}$$

That means that there is a \(b+1\) such that

$$\begin{aligned} \begin{aligned} g_2^2-4g_1g_3\ge 0, \end{aligned} \end{aligned}$$

which is equivalent to

$$\begin{aligned} \begin{aligned} 2a\gamma \mu _2\ge (a(r-\mu _1 )(1-\delta )-\gamma )(\beta \theta -a\mu _2). \end{aligned} \end{aligned}$$

This simplifies to \(\delta >\delta ^{**}\).

Thus, there exists b such that \(D\le 0\) for all \((u,v)\in {\mathbb {R}}_+^2\). By the Bendixson-Dulac theorem [56], we obtain that \(E^*\) is globally asymptotically stable under both conditions \(\mathbf {C.0}\), \(\delta ^{*}> \delta \ge \delta ^{**}\) and \(1-{\theta \beta -a\mu _2}/{ar}>\eta \ge 0\). \(\square \)

Appendix B

Proof of Theorem 2.7

Let \(\widehat{u}=u-u^*\) and \(\widehat{v}=v-v^*\), then we can rewrite the model (2) by the Taylar expansions as follows (for simplicity, u and v represent \(\widehat{u}\) and \(\widehat{v}\) again, respectively)

$$\begin{aligned} \left\{ \begin{aligned} \frac{\text {d}{u}}{\text {d} t}=&r(u+u^*)\left( \eta +\frac{\alpha (1-\eta ) }{\alpha +(v+v^*)}\right) \\&-\mu _1 (u+u^*)-\gamma (u+u^*)^2-\frac{\beta (1-\delta )(u+u^*)(v+v^*)}{1+a(1-\delta )(u+u^*)},\\ \frac{\text {d}{v}}{\text {d} t}=&\frac{\theta \beta (1-\delta )(u+u^*)(v+v^*)}{1+a(1-\delta )(u+u^*)}-\mu _2(v+v^*). \end{aligned}\right. \end{aligned}$$
(B.1)

The model (B.1) can be rewritten as

$$\begin{aligned} \begin{aligned} \left( \begin{array}{cc} u_t\\ v_t \end{array} \right) =J_{E^*}\left( \begin{array}{cc} u\\ v \end{array} \right) +\left( \begin{array}{cc} g(u,v,\eta )\\ h(u,v,\eta ) \end{array} \right) , \end{aligned} \end{aligned}$$
(B.2)

where

$$\begin{aligned} \left\{ \begin{aligned} g(u,v,\eta )=&g_{20}u^2+g_{11}uv+g_{02}v^2+g_{30}u^3+g_{21}u^2v+g_{12}uv^2+g_{03}v^3+\cdots ,\\ h(u,v,\eta )=&h_{20}u^2+h_{11}uv+h_{02}v^2+h_{30}u^3+h_{21}u^2v+h_{12}uv^2+h_{03}v^3+\cdots ,\\ \end{aligned}\right. \end{aligned}$$
(B.3)

with

$$\begin{aligned} \begin{aligned} g_{20}=&- \gamma + \frac{a\beta (1-\delta )^2v^*}{(1+a(1-\delta )u^*)^3},\; g_{11}=-\frac{r\alpha (1-\eta )}{(\alpha + v^*)^2} - \frac{\beta (1-\delta )}{(1+a(1-\delta )u^*)^2},\\ g_{02}=&\frac{r\alpha (1-\eta )u^*}{(\alpha + v^*)^3},\\ g_{30}=&-\frac{a^2\beta (1-\delta )^3v^*}{(1+a(1-\delta )u^*)^4},\;g_{21}=\frac{a\beta (1-\delta )^2}{(1+a(1-\delta )u^*)^3},\;g_{12}=\frac{r\alpha (1-\eta )}{(\alpha + v^*)^3},\\ g_{03}=&-\frac{r\alpha (1-\eta )u^*}{(\alpha + v^*)^4}, \\ h_{20}=&-\frac{a\theta \beta (1-\delta )^2v^*}{(1+a(1-\delta )u^*)^3},\;h_{11}=\frac{\theta \beta (1-\delta )}{(1+a(1-\delta )u^*)^2},\;h_{02}=0, \\ h_{30}=&\frac{a^2\theta \beta (1-\delta )^3v^*}{(1+a(1-\delta )u^*)^4},\;h_{21}=-\frac{a\theta \beta (1-\delta )^2}{(1+a(1-\delta )u^*)^3},\;h_{12}=0,\;h_{03}=0. \end{aligned} \end{aligned}$$

We know that the characteristic roots of \(J_{E^*}\) are

$$\begin{aligned} \lambda _{1,2}=S\pm iT. \end{aligned}$$

When \({-a_{12}a_{21}>S^2}\), the eigenvalues \(\lambda _1\) and \(\lambda _2\) will be complex conjugate, especially, when \(\eta =\eta ^*\), \(\lambda _1\) and \(\lambda _2\) will be purely imaginary, that is, \(S(\eta ^*)=0\) and \(\lambda _{1,2}=\pm i T(\eta ^*)\).

The eigenvectors of \(J_{E^*}\) corresponding to the eigenvalues of \(\lambda _{1,2}=S + i T\) are given by

$$\begin{aligned} X=\left[ \begin{array}{cc} 1\\ Y-iZ \end{array} \right] , \end{aligned}$$

where Y and Z are defined in (5).

Let the matrix

$$\begin{aligned} G=\left[ \begin{array}{cc} 1&{}0\\ Y&{}Z \end{array} \right] . \end{aligned}$$

By using the transformation

$$\begin{aligned} \left[ \begin{array}{cc} u\\ v \end{array} \right] =G\left[ \begin{array}{cc} x\\ y \end{array} \right] , \end{aligned}$$

then, the model (B.2) becomes

$$\begin{aligned} \begin{aligned} \left[ \begin{array}{cc} \frac{dx}{dt}\\ \frac{dy}{dt} \end{array} \right] =J^*_{E^*}\left[ \begin{array}{cc} x\\ y \end{array} \right] +\left[ \begin{array}{cc} \Phi (x,y)\\ \Psi (x,y) \end{array} \right] , \end{aligned} \end{aligned}$$
(B.4)

where

$$\begin{aligned}{} & {} J^*_{E^*}=\left[ \begin{array}{cc} S&{}-T\\ T&{}S \end{array} \right] ,\\{} & {} \Phi (x,y) = \bigg [- \gamma + \frac{a\beta (1-\delta )^2v^*}{(1+a(1-\delta )u^*)^3}- \frac{\beta (1-\delta )Y}{(1+a(1-\delta )u^*)^2}\\{} & {} \quad +\frac{r\alpha (1-\eta )(u^*Y-\alpha -v^*)Y}{(\alpha + v^*)^3}\bigg ]x^2\\{} & {} \quad -\bigg [ \frac{\beta (1-\delta )Z}{(1+a(1-\delta )u^*)^2}+\frac{r\alpha (1-\eta )(2u^*Y+\alpha +v^*)}{(\alpha + v^*)^3}Z\bigg ]\\{} & {} xy+\frac{r\alpha (1-\eta )u^*}{(\alpha + v^*)^3}Z^2y^2\\{} & {} +\bigg [-\frac{a^2\beta (1-\delta )^3v^*}{(1+a(1-\delta )u^*)^4}+\frac{a\beta (1-\delta )^2 Y}{(1+a(1-\delta )u^*)^3}-\frac{r\alpha (1-\eta )(u^*Y-\alpha -v^*)}{(\alpha + v^*)^4}Y^2\bigg ]x^3\\{} & {} -\frac{r\alpha (1-\eta )(3u^*Y-\alpha -v^*)}{(\alpha + v^*)^4}Z^2 xy^2+\cdots , \Psi (x,y)\\{} & {} =\frac{1}{Z}\bigg [-\frac{a\beta (1-\delta )^2v^*(\theta +Y)}{(1+a(1-\delta )u^*)^3}+\gamma Y+ \frac{\beta (1-\delta )(\theta +Y)Y}{(1+a(1-\delta )u^*)^2}\\{} & {} -\frac{r\alpha (1-\eta )(u^*Y-\alpha -v^*)}{(\alpha + v^*)^3}Y^2\bigg ]x^2\\{} & {} +\bigg [\frac{\beta (1-\delta )(\theta +Y)}{(1+a(1-\delta )u^*)^2}-\frac{r\alpha (1-\eta )(2u^*Y+\alpha +v^*)}{(\alpha + v^*)^3}Y\bigg ]xy\\{} & {} -\frac{r\alpha (1-\eta )u^*}{(\alpha + v^*)^3}YZy^2\\{} & {} +\frac{r\alpha (1-\eta )u^*}{(\alpha + v^*)^4}YZ^2y^3+\bigg [-\frac{a\beta (1-\delta )^2(\theta +Y)}{(1+a(1-\delta )u^*)^3}\\{} & {} +\frac{r\alpha (1-\eta )(3u^*Y-2\alpha -2v^*)}{(\alpha + v^*)^4}Y^2\bigg ]x^2y+\cdots . \end{aligned}$$

The model (B.4) can be written in the polar form as

$$\begin{aligned} \begin{aligned}&\frac{dr}{dt}=S(\eta )r+m(\eta )r^3+\cdots ,\\&\frac{d\theta }{dt}=T(\eta )+h(\eta )r^2+\cdots . \end{aligned} \end{aligned}$$
(B.5)

The Taylors series expansion of (B.5) at \(\eta =\eta ^*\) gives

$$\begin{aligned} \begin{aligned}&\frac{dr}{dt}=S'(\eta ^*)(\eta -\eta ^*)r+m(\eta ^*)r^3+\cdots ,\\&\frac{d\theta }{dt}=T(\eta )+T'(\eta ^*)(\eta -\eta ^*)+h(\eta ^*)r^2+\cdots . \end{aligned} \end{aligned}$$

where \(m(\eta ^*)\) is defined in (4).

Owing to

$$\begin{aligned} \begin{aligned} \textrm{sgn}{\left[ \frac{\partial (\textrm{Re}S)}{\partial \eta }\right] }_{\eta =\eta ^*}=\frac{ra\mu _2}{4\theta \beta }\frac{\sqrt{B_1^2-4 B_2}-B_1}{\sqrt{B_1^2-4 B_2}}|_{\eta =\eta ^*}>0. \end{aligned} \end{aligned}$$

Therefore, we have \(\Lambda =-\frac{m(\eta ^*)}{S'(\eta ^*)}\), which determines the stability of Hopf bifurcating periodic solution. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, H., Qi, H. Hopf Bifurcation Analysis of a Predator–Prey Model with Prey Refuge and Fear Effect Under Non-diffusion and Diffusion. Qual. Theory Dyn. Syst. 22, 135 (2023). https://doi.org/10.1007/s12346-023-00837-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12346-023-00837-5

Keywords

Mathematics Subject Classification

Navigation