Skip to main content
Log in

A Comparison Study of Predator–Prey Model in Deterministic and Stochastic Environments with the Impacts of Fear and Habitat Complexity

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

In theoretical ecology, recent field experiments on terrestrial vertebrates observe that the predator–prey interaction can not only be curtailed by direct consumption but also governed by some indirect effects such as the fear of predator which may reduce the reproduction rate of prey individuals. Based on this fact, we have developed and explored the predator–prey interaction with the influence of both cost and benefit of fear effect (felt by prey). A Holling type III functional response with the effect of habitat complexity has been taken to consume the prey biomass. Positivity and boundedness of the studied system prove that the model is well-behaved. The uniform persistence of the studied system is derived analytically under some parametric restrictions. The feasibility conditions and stability criteria of each equilibrium points have been discussed. Next, we have exhibited the existence of Hopf-bifurcation around the interior equilibrium point. Our mathematical analyses show that habitat complexity and fear effect both have a great impact on the persistence of the predator biomass. Furthermore, we have investigated the effect of breeding delay parameter such that the system loses its stability behaviour and enters into a limit cycle oscillations through Hopf-bifurcation. Numerical simulations are illustrated to verify our analytical outcomes. Numerically, we have perturbed the death rates of prey and predator species with Gaussian white noise terms due to the effects of environmental fluctuations.

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
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27

Similar content being viewed by others

Data Availability

The data used to support the findings of the study are available within the article.

References

  • August PV (1983) The role of habitat complexity and heterogeneity in structuring tropical mammal communities. Ecology 64:1495–1507

    Article  Google Scholar 

  • Beretta E, Kuang Y (1996) Convergence results in a well known delayed predator-prey system. J Math Anal Appl 204:840–853

    Article  MathSciNet  Google Scholar 

  • Canion CR, Heck KL (2009) Effect of habitat complexity on predation success: re-evaluating the current paradigm in seagrass beds. Mar Ecol Prog Ser 393:37–46

    Article  Google Scholar 

  • Creel S, Christianson D (2008) Relationships between direct predation and risk effects. Trends Ecol Evol 23(4):194–201

    Article  Google Scholar 

  • Creel S, Christianson D, Liley S, Winnie JA (2007) Predation risk affects reproductive physiology and demography of elk. Science 315(5814):960–960

    Article  Google Scholar 

  • Cresswell W (2010) Predation in bird populations. J Ornithol 152(S1):251–263

    Google Scholar 

  • Dalziel BD, Thomann E, Medlock J, Leenheer PD (2020) Global analysis of a predator-prey model with variable predator search rate. J Math Biol 81:159–183

    Article  MathSciNet  Google Scholar 

  • Das A, Samanta GP (2018) Modelling the fear effect on a stochastic prey-predator system with additional food for predator. J Phys A: Math Theor 51:465601

    Article  Google Scholar 

  • Freedman HI, Rao VSH (1983) The trade-off between mutual interference and time lags in predator-prey systems. Bull Math Biol 45:991–1004

    Article  MathSciNet  Google Scholar 

  • Freedman HI, Ruan S (1995) Uniform persistence in functional differential equations. J Differ Equ 115:173–192

    Article  MathSciNet  Google Scholar 

  • Hale JK (1977) Theory of functional differential equations. Springer, New York

    Book  Google Scholar 

  • Holling CS (1959) The components of predation as revealed by a study of small-mammal predation of the European pine sawfly. Can Entomol 91(5):293–320

    Article  Google Scholar 

  • Holling CS (1959) Some characteristics of simple types of predation and parasitism. Can Entomol 91(7):385–398

    Article  Google Scholar 

  • Holling CS (1965) The functional response of predators to prey density and its role in mimicry and population regulation. Mem Entomol Soc Can 97(S45):5–60

    Article  Google Scholar 

  • Kuang Y (1993) Delay differential equation with application in population dynamics. Academic Press, NY

    MATH  Google Scholar 

  • Kuang Y, Freedman HI (1988) Uniqueness of limit cycles in Gauss-type models of predator-prey systems. Math Biosci 88:67–84

    Article  MathSciNet  Google Scholar 

  • Lima SL (1998) Nonlethal effects in the ecology of Predator-Prey interactions: what are the ecological effects of anti-predator decision-making? Bioscience 48(1):25–34

    Article  Google Scholar 

  • Lotka AJ (1920) Elements of physical biology. Williams and Wilkins, Baltimore

    MATH  Google Scholar 

  • Ma Z, Wang S (2018) A delay-induced predator-prey model with Holling type functional response and habitat complexity. Nonlinear Dyn 93:1519–1544

    Article  Google Scholar 

  • Malthus TR (1798) An essay on the principle of population as it affects the future improvement of society, with remarks on the speculations of Mr. Godwin, M. Condorcet, and other writers. The Lawbook Exchange, Ltd

  • Mao X (2007) Stochastic differential equations and applications. Horwood Publishing, Chichester

    MATH  Google Scholar 

  • Meiss JD (2007) Differential dynamical systems. Society for Industrial and Applied Mathematics, Philadelphia

    Book  Google Scholar 

  • Mondal S, Maiti A, Samanta GP (2018) Effects of fear and additional food in a delayed predator-prey model. Biophys Rev Lett 13(4):157–177

    Article  Google Scholar 

  • Mondal S, Samanta GP (2020) Dynamics of a delayed predator-prey interaction incorporating nonlinear prey refuge under the influence of fear effect and additional food. J Phys A: Math Theor 53:295601

    Article  MathSciNet  Google Scholar 

  • Mondal S, Samanta G (2021) Time-delayed predator-prey interaction with the benefit of antipredation response in presence of refuge. Zeitschrift fur Naturforschung A 76:23–42. https://doi.org/10.1515/zna-2020-0195

    Article  Google Scholar 

  • Mondal S, Samanta GP (2021) Impact of fear on a predator-prey system with prey-dependent search rate in deterministic and stochastic environment. Nonlinear Dyn 104:2931–2959. https://doi.org/10.1007/s11071-021-06435-x

    Article  Google Scholar 

  • Murray JD (1993) Mathematical biology. Springer, New york

    Book  Google Scholar 

  • Perko L (2001) Differential equations and dynamical systems. Springer, New York

    Book  Google Scholar 

  • Rosenzweig ML, MacArthur RH (1963) Graphical representation and stability conditions of predator-prey interactions. Am Nat 97:209–223

    Article  Google Scholar 

  • Verhulst PF (1838) Notice sur la loi que la population poursuit dans son accroissement. Correspondance mathématique et physique 10:113–121

    Google Scholar 

  • Volterra V (1926) Variazioni e fluttuazioni del numero d’individui in specie animali conviventi. Memorie della Reale Accademia Nazionale dei Lincei 6:31–113

    MATH  Google Scholar 

  • Wang Y, Zou X (2020) On a predator-prey system with digestion delay and anti-predation strategy. J Nonlinear Sci 30:1579–1605

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Wang S, Tang H, Ma Z (2021) Hopf bifurcation of a multiple-delayed predator-prey system with habitat complexity. Math Comput Simul 180:1–23

    Article  MathSciNet  Google Scholar 

  • Winfield IJ (1986) The influence of simulated aquatic macrophytes on the zooplankton consumption rate of juvenile roach, Rutilus rutilus, rudd, Scardinius erythrophthalmus, and perch, Perca fluviatilis. J Fish Biol 29:37–48

    Article  Google Scholar 

  • Zanette LY, Allen MC, White AF, Clinchy M (2011) Perceived predation risk reduces the number of offspring songbirds produce per year. Science 334(6061):1398–1401

    Article  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors are grateful to the learned editor, reviewers and Professor Matthew Simpson (Editor-in-Chief) for their careful reading, valuable comments and helpful suggestions, which have helped them to improve the presentation of this work significantly. They are also thankful to Mr. Nirapada Santra, JRF, IIEST, Shibpur for helping to draw Fig. 2. The third author (Manuel De la Sen) is grateful to the Spanish Government for its support through Grant RTI2018-094336-B-I00 (MCIU/AEI/FEDER, UE) and to the Basque Government for its support through Grant IT1555-22.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Manuel De la Sen.

Ethics declarations

Conflict of Interest

The authors declare that they have no conflict of interest regarding this work.

Additional information

Publisher's Note

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

Appendices

Appendix A

Here, we will deduce the conditions for existence of equilibrium points and perform their stability analysis corresponding to system (3) (\(\theta =0\)).

System (3) has three feasible equilibrium points:

(i) \(\widetilde{E_0}(0, 0)\) (unstable), (ii) \(\widetilde{E_1}(k, 0)\) (LAS if \(m > 1-\frac{d_{2}}{\beta k^2 (\gamma -d_{2}h)}\)) and (iii) \({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\) where \({\widetilde{x}}=\sqrt{\frac{d_{2}}{\beta (1-m)(\gamma -d_{2}h)}}\) and \({\widetilde{y}}=\frac{\gamma r {\widetilde{x}}\left( 1-\frac{{\widetilde{x}}}{k}\right) }{d_{2}}\). Now, \({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\) exists if \(1-\frac{{\widetilde{x}}}{k} > 0 \displaystyle \Rightarrow 0< m < 1-\frac{d_{2}}{\beta k^2 (\gamma -d_{2}h)}\) provided \(d_{2} < \frac{\gamma }{h}\). Otherwise, \({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\) goes towards predator-free equilibrium point \(\widetilde{E_1}(k, 0)\). In ecological point of view, interior equilibrium point exists if the strength of habitat complexity is less than its threshold value \({\widetilde{m}}=1-\frac{d_{2}}{\beta k^2 (\gamma -d_{2}h)}\).

\({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\) has two eigenvalues with negative real parts if

$$\begin{aligned} r\left( 1-\frac{2{\widetilde{x}}}{k}\right)&< \frac{2\beta (1-m){\widetilde{x}} {\widetilde{y}}}{\left( 1+\beta (1-m)h{\widetilde{x}}^2\right) ^2}\\ \Rightarrow \left( 1-\frac{2{\widetilde{x}}}{k}\right)&<\frac{2\gamma \beta (1-m){\widetilde{x}}^2\left( 1-\frac{{\widetilde{x}}}{k}\right) }{d_{2}\left( 1+\beta (1-m)h{\widetilde{x}}^2\right) ^2} \left[ \text {putting the value of } {\widetilde{y}}\right] \\ \Rightarrow \left( 1-\frac{2{\widetilde{x}}}{k}\right)&<\frac{2\left( 1-\frac{{\widetilde{x}}}{k}\right) (\gamma -d_{2}h)}{\gamma }\\&\quad \left[ \text {using the fact } \beta (1-m){\widetilde{x}}^2=\frac{d_2}{\gamma -d_{2}h}\right] \\ \Rightarrow {\widetilde{x}}&> \frac{k(2d_{2}h-\gamma )}{2d_{2}h}\\ \Rightarrow \sqrt{\frac{d_{2}}{\beta (1-m)(\gamma -d_{2}h)}}&> \frac{k(2d_{2}h-\gamma )}{2d_{2}h}\\ \Rightarrow m&> 1-\frac{4d_{2}^3h^2}{\beta k^2(\gamma -d_{2}h)(2d_{2}h-\gamma )^2 } \text {, provided } \frac{\gamma }{2h} \\&< d_{2} < \frac{\gamma }{h}. \end{aligned}$$

Theorem 8.1

\({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\) of system (3) exists and is LAS if the strength of habitat complexity satisfies the following:

$$\begin{aligned} 1-\frac{4d_{2}^3h^2}{\beta k^2(\gamma -d_{2}h)(2d_{2}h-\gamma )^2 }< m< 1-\frac{d_{2}}{\beta k^2(\gamma -d_{2}h)}\quad \text {with}\quad \frac{\gamma }{2h}< d_{2} < \frac{\gamma }{h}. \end{aligned}$$

Now, we rewrite system (3) in the form:

$$\begin{aligned} \frac{\mathrm{d}x}{\mathrm{d}t}= & {} xg(x)- yp(x),\quad x(0)> 0 \nonumber \\ \frac{\mathrm{d}y}{\mathrm{d}t}= & {} \left( \gamma p(x)-d_{2}\right) y, \quad y(0) > 0 \end{aligned}$$
(38)

where \(g(x)=r\left( 1-\frac{x}{k}\right) \), \(p(x)=\frac{\beta (1-m) x^2}{1+\beta (1-m) h x^2}\).

Following Theorem 4.2 of Kuang and Freedman (1988), we arrive at Lemma 8.2:

Lemma 8.2

Suppose in system (38):

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}x}\left\{ \frac{g(x)+xg'(x)-xg(x)\frac{p'(x)}{p(x)}}{-d_{2}+\gamma p(x)}\right\} \le 0 \end{aligned}$$

in \(0 \le x < {\widetilde{x}}\) and \({\widetilde{x}} < x \le k\). Then system (38) has unique limit cycle which is globally asymptotically stable in the set \(\left\{ (x, y)\big |x>0, y> 0\right\} \setminus \left\{ {\widetilde{E}}\right\} \), where \(g'(x)=\frac{\mathrm{d}}{\mathrm{d}x} g(x)\) and \(p'(x)=\frac{\mathrm{d}}{\mathrm{d}x}p(x)\).

Theorem 8.3

For system (3), if \( 0< m < 1-\frac{4d_{2}^3h^2}{\beta k^2(\gamma -d_{2}h)(2d_{2}h-\gamma )^2 }\), then \({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\) is unstable in int \({\mathbb {R}}_{+}^2\) and there exists a unique stable limit cycle which is globally asymptotically stable.

Proof

Proof is similar of Theorem 4.2.1 in Wang et al. (2021). \(\square \)

Remark 1

\({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\) of system (3) is globally asymptotically stable if

$$\begin{aligned} 1-\frac{4d_{2}^3h^2}{\beta k^2(\gamma -d_{2}h)(2d_{2}h-\gamma )^2 }< m< 1-\frac{d_{2}}{\beta k^2(\gamma -d_{2}h)} \quad \text {with}\quad \frac{\gamma }{2h}< d_2 < \frac{\gamma }{h}. \end{aligned}$$

Remark 2

When the prey and predator population coexist at \({\widetilde{E}}({\widetilde{x}}, {\widetilde{y}})\), the biomass of both the prey and predator population depend on the parameter m (strength of habitat complexity). We have: \(\frac{\mathrm{d}{\widetilde{x}}}{\mathrm{d}m}=\sqrt{\frac{d_{2}}{\beta \left\{ \gamma -d_{2}h\right\} }}\left( \frac{1}{2(1-m)^\frac{3}{2}}\right) >0 \text { since } m \in (0, 1)\), i.e. increasing the value of m can increase the biomass of prey population.

Similarly, we have: \(\frac{\mathrm{d}{\widetilde{y}}}{\mathrm{d}m}=\frac{\gamma r}{d_2}\left( 1-\frac{2{\widetilde{x}}}{k}\right) \frac{\mathrm{d}{\widetilde{x}}}{\mathrm{d}m}<0 \text { if } m > 1-\frac{4d_{2}}{\beta k^2\left\{ \gamma -d_{2}h\right\} }\), i.e. \({\widetilde{y}}\) is a decreasing function of m when \(m \in \left( 1-\frac{4d_{2}}{\beta k^2\left\{ \gamma -d_{2}h\right\} }, 1-\frac{d_{2}}{\beta k^2\left\{ \gamma -d_{2}h\right\} }\right) \).

Appendix B

Using the transformation \(x=X+{\widehat{x}}\) and \(y=Y+{\widehat{y}}\), system (20) can be rewritten as follows:

$$\begin{aligned} \frac{\mathrm{d}X}{\mathrm{d}t}&=\frac{b(X(t-\tau )+{\widehat{x}})}{1+\theta (Y(t-\tau )+{\widehat{y}})}-d_{1}(X+{\widehat{x}})-a(X+{\widehat{x}})^2\nonumber \\&\quad -\frac{\beta (1-m) (X+{\widehat{x}})^2 (Y+{\widehat{y}})}{\left( 1+\beta (1-m) h (X+{\widehat{x}})^2\right) (1+\theta )} \nonumber \\&=F(t, X(t), Y(t), X(t-\tau ), Y(t-\tau ))\nonumber \\ \frac{\mathrm{d}Y}{\mathrm{d}t}&= \frac{\gamma \beta (1-m) (X+{\widehat{x}})^2 (Y+{\widehat{y}})}{\left( 1+\beta (1-m) h (X+{\widehat{x}})^2\right) (1+\theta )}-d_{2}(Y+{\widehat{y}})\nonumber \\&=G(t, X(t), Y(t), X(t-\tau ), Y(t-\tau )). \end{aligned}$$
(39)

At the interior equilibrium point \(({\widehat{x}}, {\widehat{y}})\),

$$\begin{aligned} \frac{\partial F}{\partial X}\bigg |_{({\widehat{x}}, {\widehat{y}})}&=\frac{be^{-\lambda \tau }}{1+\theta {\widehat{y}}}-d_1-2a{\widehat{x}}^2-\frac{2\beta (1-m){\widehat{x}} {\widehat{y}}}{\left( 1+\beta (1-m)h{\widehat{x}}^2\right) (1+\theta )}\\&\quad +\frac{2\beta ^2(1-m)^2h{\widehat{x}}^3{\widehat{y}}}{\left( 1+\beta (1-m)h{\widehat{x}}^2\right) ^2(1+\theta )}, \\ \frac{\partial F}{\partial Y}\bigg |_{({\widehat{x}}, {\widehat{y}})}&=-\frac{b\theta {\widehat{x}}e^{-\lambda \tau }}{\left( 1+\theta {\widehat{y}}\right) ^2}-\frac{\beta (1-m){\widehat{x}}^2}{\left( 1+\beta (1-m)h{\widehat{x}}^2\right) (1+\theta )}, \\ \frac{\partial G}{\partial X}\bigg |_{({\widehat{x}}, {\widehat{y}})}&=\frac{2\gamma \beta (1-m){\widehat{x}}{\widehat{y}}}{\left( 1+\beta (1-m)h{\widehat{x}}^2\right) ^2(1+\theta )}, \\ \frac{\partial G}{\partial Y}\bigg |_{({\widehat{x}}, {\widehat{y}})}&=0. \end{aligned}$$

So, the linearization of system (20) at \(({\widehat{x}}, {\widehat{y}})\) has the form:

$$\begin{aligned} \frac{\mathrm{d}Z}{\mathrm{d}t}=A_{1}^{'}Z(t-\tau )+A_{2}^{'}Z(t) \end{aligned}$$

where \(Z=[X, Y]^T\), \(A_{1}^{'}=\begin{bmatrix} c_{2} &{} c_{3}\\ 0&{} 0 \end{bmatrix}\) and \(A_{2}^{'}=\begin{bmatrix} q_{11}+ c_{1} &{} c_{4} \\ q_{21} &{} q_{22} \end{bmatrix}\), \(c_{1}=-\frac{b}{1+\theta {\widehat{y}}}\), \(c_{2}=\frac{b}{1+\theta {\widehat{y}}}\), \(c_{3}=-\frac{b\theta {\widehat{x}}}{\left( 1+\theta {\widehat{y}}\right) ^2}\), \(c_{4}=q_{12}+\frac{b\theta {\widehat{x}}}{\left( 1+\theta {\widehat{y}}\right) ^2}\), \(q_{11}={\widehat{x}}\left\{ -a-\frac{\beta (1-m){\widehat{y}}\left[ 1-\beta (1-m)h {\widehat{x}}^2\right] }{\left( 1+\beta (1-m)h{\widehat{x}}^2\right) ^2(1+\theta )}\right\} \), \(q_{12}=-\frac{b\theta {\widehat{x}}}{(1+\theta {\widehat{y}})^2}-\frac{\beta (1-m){\widehat{x}}^2}{\left( 1+\beta (1-m)h{\widehat{x}}^2\right) (1+\theta )}\), \(q_{21}=\frac{2\gamma \beta (1-m){\widehat{x}}{\widehat{y}}}{\left( 1+\beta (1-m)h{\widehat{x}}^2\right) ^2(1+\theta )}\) and \(q_{22}=0\).

Rights and permissions

Springer Nature or its licensor 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

Mondal, S., Samanta, G. & De la Sen, M. A Comparison Study of Predator–Prey Model in Deterministic and Stochastic Environments with the Impacts of Fear and Habitat Complexity. Bull Math Biol 84, 115 (2022). https://doi.org/10.1007/s11538-022-01067-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-022-01067-7

Keywords

Mathematics Subject Classification

Navigation