Skip to main content
Log in

Comprehensive Analysis of Deterministic and Stochastic Eco-Epidemic Models Incorporating Fear, Refuge, Supplementary Resources, and Selective Predation Effects

  • Published:
Acta Applicandae Mathematicae Aims and scope Submit manuscript

Abstract

In this investigation, we delve into the dynamics of an ecoepidemic model, considering the intertwined influences of fear, refuge-seeking behavior, and alternative food sources for predators with selective predation. We extend our model to incorporate the impact of fluctuating environmental noise on system dynamics. The deterministic model undergoes thorough scrutiny to ensure the positivity and boundedness of solutions, with equilibria derived and their stability properties meticulously examined. Furthermore, we explore the potential for Hopf bifurcation within the system dynamics. In the stochastic counterpart, we prioritize discussions on the existence of a globally positive solution. Through simulations, we unveil the stabilizing effect of the fear factor on susceptible prey reproduction, juxtaposed against the destabilizing roles of prey refuge behavior and disease prevalence intensity. Notably, when disease prevalence intensity is too low, the infection can be eradicated from the ecosystem. Our deterministic analysis reveals a complex interplay of factors: the system destabilizes initially but then stabilizes as the fear factor suppressing disease prevalence intensifies, or as predators exhibit a stronger preference for infected prey over susceptible ones, or as predators are provided with more alternative food sources. Moreover, for the stochastic system, the oscillations tend to cluster around the coexistence equilibrium of the corresponding deterministic model when white noise intensity is low. However, with increasing white noise intensity, oscillation amplitudes escalate. Critically, very high levels of white noise can lead to the eradication of infection from the ecosystem.

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

Similar content being viewed by others

References

  1. Mondal, B., Roy, S., Ghosh, U., Tiwari, P.K.: A systematic study of autonomous and nonautonomous predator-prey models for the combined effects of fear, refuge, cooperation and harvesting. Eur. Phys. J. Plus 137(6), 724 (2022)

    Article  Google Scholar 

  2. Sk, N., Tiwari, P.K., Kang, Y., Pal, S.: A nonautonomous model for the interactive effects of fear, refuge and additional food in a prey-predator system. J. Biol. Syst. 29(1), 107–145 (2021)

    Article  MathSciNet  Google Scholar 

  3. Din, A., Li, Y., Yusuf, A.: Delayed hepatitis B epidemic model with stochastic analysis. Chaos Solitons Fractals 146, 110839 (2021)

    Article  MathSciNet  Google Scholar 

  4. Din, A., Li, Y.: Stationary distribution extinction and optimal control for the stochastic hepatitis B epidemic model with partial immunity. Phys. Scr. 96(7), 074005 (2021)

    Article  Google Scholar 

  5. Khan, F.M., Khan, Z.U.: Numerical analysis of fractional order drinking mathematical model. J. Math. Techn. Model. 1(1), 11–24 (2024)

    Google Scholar 

  6. Khan, W.A., Zarin, R., Zeb, A., Khan, Y., Khan, A.: Navigating food allergy dynamics via a novel fractional mathematical model for antacid-induced allergies. J. Math. Techn. Model. 1(1), 25–51 (2024)

    Google Scholar 

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

    Article  Google Scholar 

  8. Arditi, R., Ginzburg, L.R.: Coupling in predator-prey dynamics: ratio-dependence. J. Theor. Biol. 139, 311–326 (1989)

    Article  Google Scholar 

  9. Crowley, P.H., Martin, E.K.: Functional responses and interference within and between year classes of a dragonfly population. J. North Am. Benthol. Soc. 8, 211–221 (1989)

    Article  Google Scholar 

  10. Boonstra, R., Hik, D., Singleton, G.R., Tinnikov, A.: The impact of predator-induced stress on the snowshoe hare cycle. Ecol. Monogr. 68, 371–394 (1998)

    Article  Google Scholar 

  11. Buskirk, J.V., Yurewicz, K.L.: Effects of predators on prey growth rate: relative contributions of thinning and reduced activity. Oikos 82(1), 20–28 (1998)

    Article  Google Scholar 

  12. Nelson, E.H., Matthews, C.E., Rosenheim, J.A.: Predators reduce prey population growth by inducing change in prey behavior. Ecology 85(7), 1853–1858 (2004)

    Article  Google Scholar 

  13. Wang, X., Zanette, L., Zou, X.: Modelling the fear effect in predator-prey interactions. J. Math. Biol. 73(5), 1179–1204 (2016)

    Article  MathSciNet  Google Scholar 

  14. Panday, P., Pal, N., Samanta, S., Chattopadhyay, J.: Stability and bifurcation analysis of a three-species food chain model with fear. Int. J. Bifurc. Chaos 28(01), 185000 (2018)

    Article  MathSciNet  Google Scholar 

  15. Wang, J., Cai, Y., Fu, S., Wang, W.: The effect of the fear factor on the dynamics of a predator-prey model incorporating the prey refuge. Chaos 29(8), 083109 (2019)

    Article  MathSciNet  Google Scholar 

  16. Mondal, S., Samanta, G.P., Nieto, J.J.: Dynamics of a predator-prey population in the presence of resource subsidy under the influence of nonlinear prey refuge and fear effect. Complexity 2021, 9963031 (2021)

    Article  Google Scholar 

  17. 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. Methods Appl. Sci. 45, 6217–6254 (2022)

    Article  MathSciNet  Google Scholar 

  18. Hossain, M., Pal, S., Tiwari, P.K., Pal, N.: Bifurcations, chaos, and multistability in a nonautonomous predator-prey model with fear. Chaos 31(12), 123134 (2021)

    Article  MathSciNet  Google Scholar 

  19. Tiwari, P.K., Amri, K.A., Samanta, S., Khan, Q.J., Chattopadhyay, J.: A systematic study of autonomous and nonautonomous predator-prey models with combined effects of fear, migration and switching. Nonlinear Dyn. 103(2), 2125–2162 (2021)

    Article  Google Scholar 

  20. Sarkar, A., Tiwari, P.K., Pal, S.: A delay nonautonomous model for the effects of fear and refuge on predator-prey interactions with water-level fluctuations. Int. J. Model. Simul. Sci. Comput. 13(04), 2250033 (2022)

    Article  Google Scholar 

  21. Biswas, S., Tiwari, P.K., Pal, S.: Delay-induced chaos and its possible control in a seasonally forced eco-epidemiological model with fear effect and predator switching. Nonlinear Dyn. 104(3), 2901–2930 (2021)

    Article  Google Scholar 

  22. Hossain, M., Pal, N., Samanta, S.: Impact of fear on an eco-epidemiological model. Chaos Solitons Fractals 134, 109718 (2020)

    Article  MathSciNet  Google Scholar 

  23. Sasmal, S.K.: Population dynamics with multiple Allee effects induced by fear factors – a mathematical study on prey-predator interactions. Appl. Math. Model. 64, 1–14 (2018)

    Article  MathSciNet  Google Scholar 

  24. Sha, A., Samanta, S., Martcheva, M., Chattopadhyay, J.: Backward bifurcation, oscillation and chaos in an eco-epidemilogical model with fear effect. J. Biol. Dyn. 13(8), 301–327 (2019)

    Article  MathSciNet  Google Scholar 

  25. Huang, Y., Chen, F., Zhong, L.: Stability analysis of a prey-predator model with Holling type III response function incorporating a prey refuge. Appl. Math. Comput. 182, 672–683 (2006)

    MathSciNet  Google Scholar 

  26. Gause, G.F.: The struggle for existence. In: The Williams and Wilkins Comapny, Baltimore (1934)

    Google Scholar 

  27. Dolbeer, R.A., Clark, W.R.: Population ecology of snowshoe hares in the central Rocky Mountains. J. Wildl. Manag. 39(3), 535–549 (1975)

    Article  Google Scholar 

  28. 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 World Appl. 11, 246–252 (2010)

    Article  MathSciNet  Google Scholar 

  29. Chakroborty, S., Tiwari, P.K., Sasmal, S.K., et al.: Interactive effects of prey refuge and additional food for predator in diffusive predator-prey system. Appl. Math. Model. 47, 128–140 (2017)

    Article  MathSciNet  Google Scholar 

  30. Collings, J.B.: Bifurcation and stability analysis of a temperature-dependent mite predator-prey interaction model incorporating a prey refuge. Bull. Math. Biol. 57, 63–76 (1995)

    Article  Google Scholar 

  31. Gonzalez-Olivares, E., Ramos-Jiliberto, R.: Dynamic consequences of prey refuges in a simple model system: more prey, fewer predators and enhanced stability. Ecol. Model. 166, 135–146 (2003)

    Article  Google Scholar 

  32. Sk, N., Tiwari, P.K., Pal, S.: A delay nonautonomous model for the impacts of fear and refuge in a three species food chain model with hunting cooperation. Math. Comput. Simul. 192, 136–166 (2022)

    Article  MathSciNet  Google Scholar 

  33. Tiwari, P.K., Verma, M., Pal, S., Kang, Y., Misra, A.K.: A delay nonautonomous predator-prey model for the effects of fear, refuge and hunting cooperation. J. Biol. Syst. 21(4), 1–43 (2021)

    MathSciNet  Google Scholar 

  34. Ghosh, J., Sahoo, B., Poria, S.: Prey-predator dynamics with prey refuge providing additional food to predator. Chaos Solitons Fractals 96, 110–119 (2017)

    Article  MathSciNet  Google Scholar 

  35. Bai, Y., Li, Y.: Stability and Hopf-bifurcation for a stage-structured predator-prey model incorporating refuge for prey and additional food for predator. Adv. Differ. Equ. 42 (2019)

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

    Article  Google Scholar 

  37. Sk, N., Tiwari, P.K., Pal, S., Martcheva, M.: A delay non-autonomous model for the combined effects of fear, prey refuge and additional food for predator. J. Biol. Dyn. 15(1), 580–622 (2021)

    Article  MathSciNet  Google Scholar 

  38. Bairagi, N., Adak, D.: Complex dynamics of a predator-prey-parasite system: an interplay among infection rate, predator’s reproductive gain and preference. Ecol. Complex. 22, 1–12 (2015)

    Article  Google Scholar 

  39. Biswas, S., Tiwari, P.K., Kang, Y., Pal, S.: Effects of zooplankton selectivity on phytoplankton in an ecosystem affected by free-viruses and environmental toxins. Math. Biosci. Eng. 17(2), 1272–1317 (2020)

    Article  MathSciNet  Google Scholar 

  40. Bairagi, N., Saha, S., Chaudhuri, S., Dana, S.K.: Zooplankton selectivity and nutritional value of phytoplankton influences a rich variety of dynamics in a plankton population model. Phys. Rev. E 99, 012406 (2019)

    Article  Google Scholar 

  41. Maity, S.S., Tiwari, P.K., Pal, S.: An ecoepidemic seasonally forced model for the combined effects of fear, additional foods and selective predation. J. Biol. Syst. 30(2), 285–321 (2022)

    Article  MathSciNet  Google Scholar 

  42. Roy, S., Tiwari, P.K., Nayak, H., Martcheva, M.: Effects of fear, refuge and hunting cooperation in a seasonally forced eco-epidemic model with selective predation. Eur. Phys. J. Plus 137(5), 528 (2022)

    Article  Google Scholar 

  43. Upadhyay, R.K., Mukhopadhyay, A., Iyengar, S.R.K.: Influence of environmental noise on the dynamics of a realistic ecological model. Fluct. Noise Lett. 7(01), 61–77 (2007)

    Article  Google Scholar 

  44. Rudnicki, R.: Long-time behaviour of a stochastic prey-predator model. Stoch. Process. Appl. 108(1), 93–107 (2003)

    Article  MathSciNet  Google Scholar 

  45. Liu, M., Wang, K.: Persistence and extinction of a stochastic single-specie model under regime switching in a polluted environment. J. Theor. Biol. 264(3), 934–944 (2010)

    Article  MathSciNet  Google Scholar 

  46. Liu, M., Wang, K.: Population dynamical behavior of Lotka-Volterra cooperative systems with random perturbations. Discrete Contin. Dyn. Syst. 33(6), 2495 (2013)

    Article  MathSciNet  Google Scholar 

  47. Liu, M., Wang, K.: Dynamics of a two-prey one-predator system in random environments. J. Nonlinear Sci. 23(5), 751–775 (2013)

    Article  MathSciNet  Google Scholar 

  48. Ji, C., Jiang, D., Shi, N.: Analysis of a predator-prey model with modified Leslie-Gower and Holling-type II schemes with stochastic perturbation. J. Math. Anal. Appl. 359(2), 482–498 (2009)

    Article  MathSciNet  Google Scholar 

  49. Zhang, X., Li, Y., Jiang, D.: Dynamics of a stochastic Holling type II predator-prey model with hyperbolic mortality. Nonlinear Dyn. 87(3), 2011–2020 (2017)

    Article  Google Scholar 

  50. Roy, J., Alam, S.: Fear factor in a prey-predator system in deterministic and stochastic environment. Phys. A, Stat. Mech. Appl. 541, 123359 (2020)

    Article  MathSciNet  Google Scholar 

  51. Sk, N., Pal, S.: Dynamics of an infected prey-generalist predator system with the effects of fear, refuge and harvesting: deterministic and stochastic approach. Eur. Phys. J. Plus 137(1), 138 (2022)

    Article  Google Scholar 

  52. Lafferty, K.D., Morris, A.K.: Altered behaviour of parasitized killfish increases susceptibility to predation by bird final hosts. Ecology 77, 1390–1397 (1996)

    Article  Google Scholar 

  53. Lafferty, K.D.: Foraging on prey that are modified by parasites. Am. Nat. 140, 854–867 (1992)

    Article  Google Scholar 

  54. Mandal, A., Tiwari, P.K., Samanta, S., Venturino, E., Pal, S.: A nonautonomous model for the effect of environmental toxins on plankton dynamics. Nonlinear Dyn. 99, 3373–3405 (2020)

    Article  Google Scholar 

  55. Fiasconaro, A., Valenti, D., Spagnolo, B.: Nonmonotonic behavior of spatiotemporal pattern formation in a noisy Lotka-Volterra system. Acta Phys. Pol. B 35(4), 1491–1500 (2004)

    Google Scholar 

  56. Chichigina, O., Dubkov, A., Valenti, D., Spagnolo, B.: Stability in a system subject to noise with regulated periodicity. Phys. Rev. E 84, 021134 (2011)

    Article  Google Scholar 

  57. Valenti, D., Denaro, G., La Cognata, A., et al.: Picophytoplankton dynamics in noisy marine environment. Acta Phys. Pol. B 43(5), 1227–1240 (2012)

    Article  Google Scholar 

  58. Ciuchi, S., de Pasquale, F., Spagnolo, B.: Nonlinear relaxation in the presence of an absorbing barrier. Phys. Rev. E 47(6), 3915 (1993)

    Article  Google Scholar 

  59. Ciuchi, S., de Pasquale, F., Spagnolo, B.: Self-regulation mechanism of an ecosystem in a non-Gaussian fluctuation regime. Phys. Rev. E 54(1), 706 (1996)

    Article  Google Scholar 

  60. May, R.M.: Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton (2001)

    Book  Google Scholar 

  61. Slatkin, M.: The dynamics of a population in a Markovian environment. Ecology 59(2), 249–256 (1978)

    Article  Google Scholar 

  62. Du, N.H., Kon, R., Sato, K., Takeuchi, Y.: Dynamical behavior of Lokta-Volterra competition systems: non-autonomous bistable case and the effect of telegraph noise. J. Comput. Appl. Math. 170(2), 399–422 (2004)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors extend their appreciation to the reviewers whose feedback and recommendations contributed to the enhancements of this paper.

Funding

The research of Samares Pal is partially supported by Science and Engineering Research Board, Government of India (Grant No. MTR/2020/000542).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Samares Pal.

Ethics declarations

Ethical standard

The authors state that this research complies with ethical standards. This research does not involve either human participants or animals.

Competing Interests

The authors declare that there is no conflict of interests regarding the publication of this article.

Additional information

Publisher’s Note

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

Appendices

Appendix A

For any solution \((S(t),I(t),P(t))\) of system (1) with the positive initial values \((S_{0},I_{0},P_{0})\), we have

$$\begin{aligned} S(t) =&S_{0}\exp \biggl[\displaystyle \int ^{t}_{0}\biggl\{ \frac{r_{0}}{1+ k_{1}P(u)}-r_{1}-r_{2}S(u)- \frac{\lambda I(u)}{1+k_{2}P(u)}\\ &{}- \frac{n(1-mP(u))P(u)}{b+S(u)+\theta I(u)}\biggr\} du \biggr], \\ I(t) =&I_{0}\exp \left [\displaystyle \int ^{t}_{0}\left \{ \frac{\lambda S(u)}{1+k_{2}P(u)}- \frac{n\theta (1-mP(u))P(u)}{b+S(u)+\theta I(u)}-d\right \}du\right ], \\ P(t) =&P_{0}\exp \biggl[\displaystyle \int ^{t}_{0} \biggl\{ r_{3} \biggl(1-\frac{P(u)}{K} \biggr)+n(1-mP(u)) \frac{c_{1}S(u)+c_{2}\theta I(u)}{b+S(u)+\theta I(u)}\\ &{}-\delta _{1}- \delta _{2} P(u)\biggr\} du \biggr]. \end{aligned}$$

Positivity of the solutions of system (1) is guaranteed from the above expressions. Let \(\Phi =S+I+P\), then the time derivative of \(\Phi \) will be

$$\begin{aligned} \dot{\Phi} =&\frac{r_{0}S}{1+k_{1}P}+r_{3}P\left (1-\frac{P}{K} \right )- \frac{(1-c_{1})(1-mP)nSP+(1-c_{2})n\theta (1-mP)IP}{b+S+\theta I} \\ &-d I-\delta _{1} P-\delta _{2}P^{2}-r_{1}S-r_{2}S^{2} \\ \leq & r_{0}S+r_{3}P-r_{1}S-d I-\delta _{1} P-r_{2}S^{2}. \end{aligned}$$

For \(\delta >0\), we have \(\dot{\Phi}+\delta \Phi \leq S(-r_{2}S+r_{0}-r_{1}+\delta )-I(d- \delta )-P(\delta _{1}-r_{3}-\delta )\). For \(0<\delta \leq \min \{d,\delta _{1}-r_{3}\}\), we get \(\dot{\Phi}+\delta \Phi \leq S(-r_{2}S+r_{0}-r_{1}+\delta )\). Note that the maximum value of \(\displaystyle S(-r_{2}S+r_{0}-r_{1}+\delta )\) is \(\displaystyle \frac{(r_{0}-r_{1}+\delta )^{2}}{4r_{2}}\), which is always positive; denote this number by \(\kappa \). Thus, we get \(\displaystyle \dot{\Phi}+\delta \Phi \leq \kappa \). By the theorem of differential inequality, we have

$$\begin{aligned} 0< \Phi (S,I,P)\leq \frac{\kappa (1-e^{-\delta t})}{\delta}+\Phi (S_{0},I_{0},P_{0})e^{- \delta t}. \end{aligned}$$

Thus, for sufficiently large values of t, we have \(0<\Phi (t)\leq \kappa /\delta \).

Appendix B

Jacobian matrix \(V=[V_{ij}]_{3\times 3}\) of system (1) has the following entries:

$$\begin{aligned} &V_{11}=\frac{r_{0}}{1+k_{1} P}-r_{1}-2r_{2}S- \frac{\lambda I}{1+k_{2}P}- \frac{n(b+\theta I)(1-mP)P}{(b+S+\theta I)^{2}}, \\ & V_{12}=-\frac{\lambda S}{1+k_{2}P}+ \frac{n\theta (1-mP)SP}{(b+S+\theta I)^{2}}, \\ & V_{13}=- \frac{r_{0}k_{1} S}{(1+k_{1}P)^{2}}+ \frac{\lambda k_{2}IS}{(1+k_{2}P)^{2}}-\frac{nS(1-2mP)}{b+S+\theta I}, \\ & V_{21}=\frac{\lambda I}{1+k_{2}P}+ \frac{n\theta (1-mP)IP}{(b+S+\theta I)^{2}}, \ V_{22}= \frac{\lambda S}{1+k_{2}P}- \frac{n\theta (b+S)(1-mP)P}{(b+S+\theta I)^{2}}-d, \\ & V_{23}=-\frac{\lambda k_{2}IS}{(1+k_{2}P)^{2}}- \frac{n\theta (1-mP)I}{b+S+\theta I},\ V_{31}= \frac{n(1-mP)P\{bc_{1}+(c_{1}-c_{2})\theta I\}}{(b+S+\theta I)^{2}}, \\ &V_{32}= \frac{n\theta (1-mP)P\{bc_{2}+(c_{2}-c_{1}S)\}}{(b+S+\theta I)^{2}}, \\ &V_{33}=r_{3}\left (1-\frac{P}{K}\right )-\frac{r_{3}P}{K}+ \frac{n(1-2mP)(c_{1}S+c_{2}\theta I)}{b+S+\theta I}-\delta _{1}-2 \delta _{2} P. \end{aligned}$$
  1. 1.

    The three eigenvalues, after evaluating the matrix \(V\) at the equilibrium \(E_{0}\), are obtained as \(r_{0}-r_{1}\), \(-d\) and \(r_{3}-\delta _{1}\). Clearly, one eigenvalue is always negative while the other two will be negative if the conditions in (4) hold. Thus, the equilibrium \(E_{0}\) is stable in view of conditions in (4).

  2. 2.

    At the equilibrium point \(E_{1}\), eigenvalues of the Jacobian matrix \(V\) are obtained as \(-(r_{0}-r_{1})\), \(\displaystyle \frac{\lambda (r_{0}-r_{1})}{r_{2}}-d\) and \(\displaystyle r_{3}-\delta _{1}+ \frac{nc_{1}(r_{0}-r_{1})}{br_{2}+r_{0}-r_{1}}\). Note that in view of the feasibility condition of the equilibrium \(E_{1}\), one eigenvalue is negative while the other two will be negative if the conditions in (5) hold. Thus, the conditions stated in (5) lead to stability of the equilibrium \(E_{1}\).

  3. 3.

    The Jacobian matrix of the system (1) at the equilibrium \(E_{2}\) has one eigenvalue as \(\displaystyle r_{3}\left (1-\frac{2P_{2}}{K}\right )-\delta _{1}-2 \delta _{2} P_{2}\) whereas the others two are the roots of the following quadratic equation:

    $$\begin{aligned} &\Lambda ^{2}-\left \{\frac{r_{0}}{1+k_{1}P_{2}}-r_{1}-n\left ( \frac{1}{b}+\frac{\theta}{d}\right )(1-mP_{2})P_{2}\right \}\Lambda \\ & \hspace{2cm}+\frac{n^{2}\theta (1-mP_{2})^{2}P_{2}^{2}}{bd}-\left ( \frac{r_{0}}{1+k_{1}P_{2}}-r_{1}\right ) \frac{n\theta (1-mP_{2})P_{2}}{d}=0. \end{aligned}$$

    In view of the conditions in (6), the equilibrium \(E_{2}\) is stable as under those conditions, eigenvalues are either negative or have negative real parts. Therefore, if the conditions in (6) are satisfied, the equilibrium \(E_{3}\) is locally stable.

  4. 4.

    At the equilibrium \(E_{3}\), the Jacobian matrix \(V\) has one eigenvalue as \(\displaystyle r_{3}-\delta _{1}+ \frac{n(C_{1}S_{3}+C_{2}\theta I_{3})}{b+S_{3}+\theta I_{3}}\) whereas the others two are the roots of the following quadratic equation:

    $$\begin{aligned} \Lambda ^{2}-\left (r_{0}-2r_{2}S_{3}-r_{1}-\lambda I_{3}\right ) \Lambda +\lambda ^{2}S_{3}I_{3}=0. \end{aligned}$$

    If the conditions in (7) hold, all the eigenvalues are either negative or negative real parts, and hence the equilibrium \(E_{3}\) is stable locally.

  5. 5.

    The Jacobian matrix of the system (1) at the equilibrium \(E_{4}\) immediately gives one eigenvalue as \(\displaystyle \frac{\lambda S_{4}}{1+k_{2}P_{4}}- \frac{n\theta (1-mP_{4})P_{4}}{b+S_{4}}-d\). The other two eigenvalues are the roots of the following quadratic equation: \(\Lambda ^{2}-(F_{1}+F_{2})\Lambda +(F_{1}F_{2}-F_{3}F_{4})=0\),

    where

    $$\begin{aligned} &F_{1}=-r_{2}S_{4}+\frac{n(1-mP_{4})S_{4}P_{4}}{(b+S_{4})^{2}}, \ F_{2}=- \left [\frac{r_{3}}{K}+\delta _{2}+\frac{c_{1}mnS_{4}}{b+S_{4}} \right ]P_{4}, \\ &F_{3}=-\frac{r_{0}k_{1}S_{4}}{(1+k_{1}S_{4})^{2}}- \frac{n(1-2mP_{4})S_{4}}{b+S_{4}},\ F_{4}= \frac{bc_{1}n(1-mP_{4})P_{4}}{(b+S_{4})^{2}}. \end{aligned}$$

    If the conditions in (8) hold, all the eigenvalues are either negative or have negative real parts. Thus, the equilibrium \(E_{4}\) is stable under the conditions stated in (8).

  6. 6.

    The matrix obtained after evaluating the Jacobian matrix of system (1) at the equilibrium \(E_{*}\) is \(V_{E^{*}}=[m_{ij}]_{3\times 3}\), whose entries are as follows:

    $$\begin{aligned} &m_{11}=-r_{2}S^{*}+ \frac{n(1-mP^{*})S^{*}P^{*}}{(b+S^{*}+\theta I^{*})^{2}}, \ m_{12}=- \frac{\lambda S^{*}}{1+k_{2}P^{*}}+ \frac{n\theta (1-mP^{*})S^{*}P^{*}}{(b+S^{*}+\theta I^{*})^{2}}, \\ &m_{13}=-\frac{r_{0}k_{1} S^{*}}{(1+k_{1} P^{*})^{2}}+ \frac{\lambda k_{2}S^{*}I^{*}}{(1+k_{2}P^{*})^{2}}- \frac{n(1-2mP^{*})S^{*}}{b+S^{*}+\theta I^{*}}, \\ &m_{21}=\frac{\lambda I^{*}}{1+k_{2}P^{*}}+ \frac{n\theta (1-mP^{*}) I^{*}P^{*}}{(b+S^{*}+\theta I^{*})^{2}}, \ m_{22}= \frac{n\theta ^{2}(1-mP^{*}) I^{*}P^{*}}{(b+S^{*}+\theta I^{*})^{2}}, \\ & m_{23}=-\frac{\lambda k_{2}S^{*}I^{*}}{(1+k_{2}P^{*})^{2}}- \frac{n\theta (1-2mP^{*})I^{*}}{b+S^{*}+\theta I^{*}},\\ & m_{31}= \frac{n\{bc_{1}+(c_{1}-c_{2})\theta I^{*}\}(1-mP^{*})P^{*}}{(b+S^{*}+\theta I^{*})^{2}}, \\ &m_{32}= \frac{n\theta \{bc_{2}+(c_{2}-c_{1})S^{*}\}(1-mP^{*}) P^{*}}{(b+S^{*}+\theta I^{*})^{2}}, \\ & m_{33}=-\left [\frac{r_{3}}{K}+\delta _{2}+ \frac{mn(c_{1}S^{*}+c_{2}\theta I^{*})}{b+S^{*}+\theta I^{*}}\right ]P^{*}. \end{aligned}$$

    The characteristic equation corresponding to the matrix \(J_{E^{*}}\) is given by

    $$\begin{aligned} \Lambda ^{3}+A_{1}\Lambda ^{2}+A_{2}\Lambda +A_{3}=0, \end{aligned}$$
    (11)

    where

    $$\begin{aligned} & A_{1}=-(m_{11}+m_{22}+m_{33}), \\ & A_{2}=m_{22}m_{33}-m_{23}m_{32}+m_{11}m_{33}-m_{13}m_{31}+m_{11}m_{22}-m_{12}m_{21}, \\ & A_{3}=m_{12}(m_{21}m_{33}-m_{11}(m_{22}m_{33}-m_{23}m_{32})-m_{23}m_{31})-m_{13}(m_{21}m_{32}-m_{22}m_{31}). \end{aligned}$$

    In view of Routh-Hurwitz criterion, roots of the equation (11) will be either negative or have negative real parts if and if only if the three conditions in (9) hold. Therefore, the equilibrium \(E_{*}\) is locally asymptotically stable if and only if the conditions mentioned in (9) are satisfied.

Appendix C

Let the critical value of \(\theta \) (say \(\widehat{\theta}\)) be such that

$$\begin{aligned} A_{1}(\widehat{\theta})A_{2}(\widehat{\theta})-A_{3}(\widehat{\theta})=0. \end{aligned}$$
(12)

Thus, at \(\theta =\widehat{\theta}\), the characteristic equation (11) can be rewritten as \((\Lambda +A_{1})(\Lambda ^{2}+A_{2})=0\). This equation has three roots, a pair of purely imaginary roots \(\Lambda _{1,2}=\pm i\sqrt{A_{2}}\) and a negative root \(\Lambda _{3}=-A_{1}\). Thus, Hopf bifurcation occurs in the system (1).

Now, we consider a point \(\theta \) in an \(\epsilon -\) neighborhood of \(\widehat{\theta}\), then the above roots become function of \(\theta \), namely \(\Lambda _{1,2}=\eta (\theta )\pm i\rho (\theta )\). Thus, from equation (12), after separating real and imaginary parts, we get

$$\begin{aligned} \eta ^{3}-3\eta \rho ^{2}+A_{1}(\eta ^{2}-\rho ^{2})+A_{2}\eta +A_{3} =&0, \end{aligned}$$
(13)
$$\begin{aligned} 3\eta ^{2}\rho -\rho ^{3}+2A_{1}\eta \rho +A_{2}\rho =&0. \end{aligned}$$
(14)

Since \(\rho (\theta )\neq 0\), from equation (14), it follows that \(\rho ^{2}=3\eta ^{2}+2A_{1}\eta +A_{2}\). Substituting this in equation (13), we get

$$\begin{aligned} 8\eta ^{3}+8A_{1}\eta ^{2}+2\eta (A^{2}_{1}+A_{2})+A_{1}A_{2}-A_{3}=0. \end{aligned}$$
(15)

Since \(\eta (\widehat{\theta})=0\), from the above equation, we get

$$\begin{aligned} \left [\frac{d\eta}{d\theta}\right ]_{\theta =\widehat{\theta}}= - \left [\frac{1}{2(A^{2}_{1}+A_{2})}\frac{d}{d\theta}(A_{1}A_{2}-A_{3}) \right ]_{\theta =\widehat{\theta}}. \end{aligned}$$

The right side of the above equation does not vanish provided \(\displaystyle \left [\frac{d}{d\theta}(A_{1}A_{2}-A_{3})\right ]_{ \theta =\widehat{\theta}}\neq 0\).

Appendix D

Letting \(S(t)=e^{x}\), \(I(t)=e^{y}\) and \(P(t)=e^{z}\), the system (10) transforms into the following form:

$$\begin{aligned} &dx=\left [\frac{r_{0}}{1+k_{1}e^{z}}-r_{2}e^{x}-r_{1}- \frac{\sigma ^{2}_{1}}{2}-\frac{\lambda e^{y}}{1+k_{2}e^{z}}- \frac{n(1-me^{z})e^{z}}{b+e^{x}+\theta e^{y}}\right ]dt+\sigma _{1}d \beta _{1}(t), \\ &dy=\left [\frac{\lambda e^{x}}{1+k_{2}e^{z}}- \frac{n\theta (1-me^{z})e^{z}}{b+e^{x}+\theta e^{y}}-d- \frac{\sigma ^{2}_{2}}{2}\right ]dt+\sigma _{2}d\beta _{2}(t), \\ &dz=\left [r_{3}\left (1-\frac{e^{z}}{K}\right )+ \frac{n(1-me^{z})(e^{x}+\theta e^{y})}{b+e^{x}+\theta e^{y}}-\delta _{1}- \delta _{2}e^{z}-\frac{\sigma ^{2}_{3}}{2}\right ]dt+\sigma _{3}d \beta _{3}(t) \end{aligned}$$
(16)

with initial value \((x_{0},y_{0},z_{0})=(\ln S_{0},\ln I_{0},\ln P_{0} )\). Obviously, the coefficients of system (16) are locally Lipschitz continuous. Thus, there is a unique maximal local solution \((x(t),y(t),z(t))\) of system (16) for \(t\in [0,\tau _{e})\), where \(\tau _{e}\) denotes the explosion time. Therefore, \((S(t),I(t),P(t))=(e^{x(t)},e^{y(t)},e^{z(t)})\) is the unique positive local solution of the system (16) with positive initial value \((S_{0},I_{0},P_{0})\) on the interval \([0,\tau _{e})\).

Now, for the global positive solution, we need to verify \(\tau _{e}=\infty \) a.s. Since \((S(t),I(t), P(t))\) is always positive, from the first equation of the system (10), we can write

$$\begin{aligned} dS(t)\leq S(t)(r_{0}-r_{2}S(t))dt+\sigma _{1}S(t)d\beta _{1}(t). \end{aligned}$$

Now, we consider the following stochastic differential equation:

$$\begin{aligned} d\Phi _{1}(t)=\Phi _{1}(t)\left [r_{0}-r_{2}\Phi _{1}(t)\right ]dt+ \sigma _{1}\Phi _{1}(t)d\beta _{1}(t), \ \Phi _{1}(0)=S_{0}. \end{aligned}$$
(17)

Then, \(\displaystyle \Phi _{1}(t)= \frac{e^{\left (r_{0}-\frac{\sigma _{1}^{2}}{2}\right )t+\sigma _{1}\beta _{1}(t)}}{\frac{1}{S_{0}}+r_{2}\int _{0}^{t}e^{\left (r_{0}-\frac{\sigma _{1}^{2}}{2}\right )s+\sigma _{1}\beta _{1}(s)}ds}\) is the unique solution of system (17). Thus, from the comparison theorem of stochastic differential equation, we have \(S(t)\leq \Phi _{1}(t)\), \(t\in [0,\tau _{e})\) a.s.

From the second equation of system (10), we can write

$$\begin{aligned} dI(t)\leq I(t)[(\lambda S(t)-d)]dt+\sigma _{2}I(t)d\beta _{2}(t). \end{aligned}$$

Consider the following auxiliary system:

$$\begin{aligned} d\Phi _{2}(t)=\Phi _{2}(t)[\lambda \Phi _{1}(t)-d]dt+\sigma _{2}\Phi _{2}(t)d \beta _{2}(t), \ \Phi _{2}(0)=I_{0}. \end{aligned}$$
(18)

Then, \(\displaystyle \Phi _{2}(t)=I_{0}e^{\int _{0}^{t}\left (\lambda \Phi _{1}(s)-d- \frac{\sigma _{2}^{2}}{2}\right )ds+\sigma _{2}\beta _{2}(t)}\) is the unique solution of system (18). Thus, from the comparison theorem of stochastic differential equation, we have \(I(t)\leq \Phi _{2}(t)\), \(t\in [0,\tau _{e})\) a.s.

Similarly, for the following auxiliary system:

$$\begin{aligned} &d\Phi _{3}(t)=\Phi _{3}(t)\left [\left \{r_{3}-\delta _{1}+ \frac{n\Phi _{1}(t)+\theta \Phi _{2}(t)}{b+\Phi _{1}(t)+\theta \Phi _{2}(t)} \right \}-\delta _{2}\Phi _{3}(t)\right ]dt+\sigma _{3}\Phi _{3}(t)d \beta _{3}(t), \\ & \Phi _{3}(0)=P_{0}.\\ &\Phi _{3}(t)= \frac{\exp \left [\int _{0}^{t}{\left (r_{3}-\delta _{1}-\frac{\sigma _{3}^{2}}{2}+\frac{n\Phi _{1}(s)+\theta \Phi _{2}(s)}{b+\Phi _{1}(s)+\theta \Phi _{2}(s)}\right )ds+\sigma _{3}\beta _{3}(t)}\right ]}{\frac{1}{P_{0}}+\delta _{2}\int _{0}^{t}\exp \left [\int _{0}^{s}{\left (r_{3}-\delta _{1}-\frac{\sigma _{3}^{2}}{2}+\frac{n\Phi _{1}(\tau )+\theta \Phi _{2}(\tau )}{b+\Phi _{1}(\tau )+\theta \Phi _{2}(\tau )}\right )d\tau +\sigma _{3}\beta _{3}(s)}\right ]ds} \end{aligned}$$
(19)

is the unique solution. Thus, from the comparison theorem of stochastic differential equation, we have \(P(t)\leq \Phi _{3}(t)\), \(t\in [0,\tau _{e})\) a.s. Using the positivity of \(S(t)\), we get the following inequality:

$$\begin{aligned} dS(t)\geq S(t)\left [(r_{0}-r_{1}-n)-r_{2}S(t)\right ]dt+\sigma _{1}S(t)d \beta _{1}(t). \end{aligned}$$

Clearly,

$$\begin{aligned} \phi _{1}(t)= \frac{\exp \left \{(r_{0}-r_{1}-n-\frac{\sigma _{1}^{2}}{2})t+\sigma _{1}\beta _{1}(t)\right \}}{\frac{1}{a_{10}}+r\int _{0}^{t}\exp \left [(r_{0}-r_{1}-n-\frac{\sigma _{1}^{2}}{2})s+\sigma _{1}\beta _{1}(s)\right ]ds} \end{aligned}$$

is the unique solution to the initial value problem

$$\begin{aligned} &d\phi _{1}(t)=\phi _{1}(t)\left [(r_{0}-r_{1}-n)-r_{2}\phi _{1}(t) \right ]dt+\sigma _{1}\phi _{1}(t)d\beta _{1}(t),\ \phi _{1}(0)=a_{10}. \end{aligned}$$

Hence, by comparison theorem of stochastic differential equation, we get \(S(t)\geq \phi _{1}(t)\) for \(t\in [0,\tau _{e})\) a.s. and \(I(t)\geq \phi _{2}(t)\) for \(t\in [0,\tau _{e})\) a.s., where

$$\begin{aligned} \phi _{2}(t)=a_{20}\exp \left [\int _{0}^{t}\left (\lambda \phi _{1}(s)-n \theta -d-\frac{\sigma _{2}^{2}}{2}\right )ds+\sigma _{2}\beta _{2}(t) \right ] \end{aligned}$$

is the unique solution of the following auxiliary equation:

$$\begin{aligned} &d\phi _{2}(t)=\phi _{2}(t)\left [(\lambda \phi _{1}(t)-n\theta -d) \phi _{2}(t)\right ]dt+\sigma _{2}\phi _{2}(t)d\beta _{2}(t),\ \phi _{2}(0)=a_{20}. \end{aligned}$$

Again, \(P(t)\geq \phi _{3}(t)\ for \ t\in [0,\tau _{e})\) a.s., where

$$\begin{aligned} \phi _{3}(t)= \frac{\exp \left \{(r_{3}+n-\delta _{1}-\frac{\sigma _{3}^{2}}{2})t+\sigma _{3}\beta _{3}(t)\right \}}{\frac{1}{a_{30}}+(\frac{r_{3}}{K}+\delta _{2})\int _{0}^{t}\exp \left \{(r_{3}+n-\delta _{1}-\frac{\sigma _{3}^{2}}{2})s+\sigma _{3}\beta _{3}(s)\right \}ds} \end{aligned}$$

is the unique solution of the auxiliary equation

$$\begin{aligned} &d\phi _{3}(t)=\phi _{3}(t)\left [(r_{3}+n-\delta _{1})-\left ( \frac{r_{3}}{K}+\delta _{2}\right )\phi _{3}(t)\right ]dt+\sigma _{3} \phi _{3}(t)d\beta _{3}(t),\ \phi _{3}(0)=a_{30}. \end{aligned}$$

Therefore, we get

$$\begin{aligned} \phi _{1}(t)\leq S(t)\leq \Phi _{1}(t), \ \phi _{2}(t)\leq I(t)\leq \Phi _{2}(t),\ \phi _{3}(t)\leq P(t)\leq \Phi _{3}(t),\ \forall \ t \in [0,\tau _{e})\ \textit{a.s.} \end{aligned}$$

From the above, it is clear that the values of \(\Phi _{1}(t)\), \(\Phi _{2}(t)\), \(\Phi _{3}(t)\), \(\phi _{1}(t)\), \(\phi _{2}(t)\) and \(\phi _{3}(t)\) will not be explored at any finite time, and are well defined for any large values of \(\tau _{e}\). Hence, \((S(t),I(t),P(t))\) is a global solution of system (10).

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

Maity, S.S., Tiwari, P.K. & Pal, S. Comprehensive Analysis of Deterministic and Stochastic Eco-Epidemic Models Incorporating Fear, Refuge, Supplementary Resources, and Selective Predation Effects. Acta Appl Math 191, 5 (2024). https://doi.org/10.1007/s10440-024-00654-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10440-024-00654-1

Keywords

Navigation