Skip to main content
Log in

Combined impact of predatory insects and bio-pesticide over pest population: impulsive model-based study

  • Original Article
  • Published:
Energy, Ecology and Environment Aims and scope Submit manuscript

Abstract

The production with increasing demands maintaining the balance of nature and natural diversity is the most challenging part of the agricultural system. However, pests and other insect populations are significant obstacles to the continuous food supply. This study proposes a crop pest management mathematical model using the predator (pests’ natural enemy) and viral infection through bio-pesticides. Impulsive differential equations have been implemented to study the dynamics between all populations, considering the repetitive release of virus micro-pesticides and predator insects in the crop field. The hypothesized model gives the outlook of complex natural dynamics. Two types of scenarios have been analyzed here using the model: One deals with the complete eradication of the field’s pest population, and another is sounder from a biodiversity conservation perspective, that defines the minimum pest population below the economic injury level, which is, nowadays, the major challenge in the agricultural field. Numerical examples show that pest management is successful when considering the minimum pest level that keeps the economic threshold by optimizing predator and virus levels cost-effectively.

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

Similar content being viewed by others

References

  • Abbas N, Shad SA, Razaq M (2012) Fitness cost, cross resistance and realized heritability of resistance to imidacloprid in Spodoptera litura (Lepidoptera: Noctuidae). Pestic Biochem Phys 103(3):181–188

    Article  Google Scholar 

  • Abraha T, Al Basir F, Obsu LL, Torres DF (2021) Pest control using farming awareness: Impact of time delays and optimal use of biopesticides. Chaos Soliton Fract 146:110869

    Article  MathSciNet  Google Scholar 

  • Al Basir F, Elaiw AM, Ray S (2019) Effect of time delay in controlling crop pest using farming awareness. Int J Appl Comput Math 5(4):1–9

    Article  MathSciNet  Google Scholar 

  • Al Basir F (2020) A multi-delay model for pest control with awareness induced interventions-Hopf bifurcation and optimal control analysis. Int J Biomath 13(06):2050047

    Article  MathSciNet  Google Scholar 

  • Barclay HJ (1982) Models for pest control using predator release, habitat management and pesticide release in combination. J Appl Ecol 19:337–348

    Article  Google Scholar 

  • Barclay HJ (1987) Models for pest control: complementary effects of periodic releases of sterile pests and parasitoids. Theor Popul Biol 32(1):76–89

    Article  Google Scholar 

  • Bhattacharyya S, Bhattacharya DK (2006) Pest control through viral disease: mathematical modeling and analysis. J Theor Biol 238(1):177–197

    Article  MathSciNet  Google Scholar 

  • Croft BA (1990) Arthropod biological control agents and pesticides. Wiley, London

    Google Scholar 

  • Chowdhury J, Al Basir F, Pal J, Roy PK (2016) Pest control for Jatropha curcas plant through viral disease: a mathematical approach. Nonlinear Stud 23(4):515–30

    MathSciNet  MATH  Google Scholar 

  • Chattopadhyay P, Banerjee G, Mukherjee S (2017) Recent trends of modern bacterial insecticides for pest control practice in integrated crop management system. 3 Biotech 7(1):60

    Article  Google Scholar 

  • Eizi Y (2005) Effects of intraguild predation and interspecific competition among biological control agents in augmentative biological control in greenhouses. USDA Forest Service Publication, Davos, pp 523–530

    Google Scholar 

  • Flexner J, Lighthart B, Croft BA (1986) The effects of microbial pesticides on non-target, beneficial arthropods. Agric Ecosyst Environ 16(3–4):203–254

    Article  Google Scholar 

  • Freedman HI (1976) Graphical stability, enrichment, and pest control by a natural enemy. Math Biosci 31(3–4):207–225

    Article  MathSciNet  Google Scholar 

  • Georgescu P, Zhang H (2010) An impulsively controlled predator-pest model with disease in the pest. Nonlinear Anal Real World Appl 11(1):270–287

    Article  MathSciNet  Google Scholar 

  • Ghosh S, Bhattacharyya S, Bhattacharya DK (2007) The role of viral infection in pest control: a mathematical study. Bull Math Biol 69(8):2649–2691

    Article  MathSciNet  Google Scholar 

  • Idris AL, Fan X, Muhammad MH, Guo Y, Guan X, Huang T (2020) Ecologically controlling insect and mite pests of tea plants with microbial pesticides: a review. Arch Microbiol 17:1

    Google Scholar 

  • Jana S, Kar TK (2013) A mathematical study of a prey–predator model in relevance to pest control. Nonlinear Dyn 74(3):667–83

    Article  MathSciNet  Google Scholar 

  • Kar TK, Ghorai A, Jana S (2012) Dynamics of pest and its predator model with disease in the pest and optimal use of pesticide. J Theor Biol 310:187–198

    Article  MathSciNet  Google Scholar 

  • Katti G (2013) Biopesticides for insect pest management in rice-present status and future scope. J Rice Res 6(1):1–15

    Article  Google Scholar 

  • Kropff MJ, Teng PS, Rabbinge R (1995) The challenge of linking pest and crop models. Agric Syst 49(4):413–434

    Article  Google Scholar 

  • Liu B, Teng Z, Chen L (2006) Analysis of a predator-prey model with Holling II functional response concerning impulsive control strategy. J Comput Appl Math 193(1):347–362

    Article  MathSciNet  Google Scholar 

  • Liu B, Zhang Y, Chen L (2004) Dynamic complexities of a Holling I predator–prey model concerning periodic biological and chemical control. Chaos Soliton Fract 22(1):123–134

    Article  MathSciNet  Google Scholar 

  • Liu B, Zhang Y, Chen L (2005) Dynamic complexities in a Lotka–Volterra predator–prey model concerning impulsive control strategy. Int J Bifur Chaos 15(02):517–531

    Article  MathSciNet  Google Scholar 

  • Lakshmikantham V, Simeonov PS (1989) Theory of impulsive differential equations, vol 6. World Scientific, Singapore

    Book  Google Scholar 

  • Tian Y, Tang S, Cheke RA (2019) Dynamic complexity of a predator–prey model for IPM with nonlinear impulsive control incorporating a regulatory factor for predator releases. Math Model Anal 24(1):134–54

    Article  MathSciNet  Google Scholar 

  • Perdikis D, Fantinou A, Lykouressis D (2011) Enhancing pest control in annual crops by conservation of predatory Heteroptera. Biol Control 59(1):13–21

    Article  Google Scholar 

  • Luff ML (1983) The potential of predators for pest control. Agric Ecosyst Environ 10(2):159–81

    Article  Google Scholar 

  • Ostad-Ali-Askari K, Shayannejad M (2021) Quantity and quality modelling of groundwater to manage water resources in Isfahan-Borkhar Aquifer. Environ Dev Sustain. https://doi.org/10.1007/s10668-021-01323-1

    Article  Google Scholar 

  • Ostad-Ali-Askari K, Shayannejad M, Ghorbanizadeh-Kharazi H (2017) Artificial neural network for modeling nitrate pollution of groundwater in marginal area of Zayandeh-rood River, Isfahan, Iran. KSCE J Civ Eng 21(1):134–140. https://doi.org/10.1007/s12205-016-0572-8

    Article  Google Scholar 

  • Páez CJ, Jungmann D, Siegmund S (2017) Modeling and analysis of integrated pest control strategies via impulsive differential equations. Int J Diff Eqns. https://doi.org/10.1155/2017/1820607

    Article  MATH  Google Scholar 

  • Páez CJ, Jungmann D, Siegmund S (2018) A comparative study of integrated pest management strategies based on impulsive control. J Biol Dyn 12(1):318–41

    Article  MathSciNet  Google Scholar 

  • Prokopy R, Kogan M (2009) Integrated pest management. In: Encyclopedia of insects. Academic Press, pp 523–528

  • Sajap AS, Kotulai JR, Kadir HA, Hussein MY (1999) Impact of prey infected by nuclear polyhedrosis virus on a predator, Sycanus leucomesus Walk (Hem., Reduviidae). J Appl Entomol 123(2):93–97

    Article  Google Scholar 

  • Samada LH, Tambunan US (2020) Biopesticides as promising alternatives to chemical pesticides: a review of their current and future status. OnLine J Biol Sci 20(2):66–76

    Article  Google Scholar 

  • Tang S, Chen L (2004) Modelling and analysis of integrated pest management strategy. Dis Cont Dyn Syst B 4(3):759

    MathSciNet  MATH  Google Scholar 

  • Tang S, Tang G, Cheke RA (2010) Optimum timing for integrated pest management: modelling rates of pesticide application and natural enemy releases. J Theor Biol 264(2):623–638

    Article  MathSciNet  Google Scholar 

  • van Lenteren JC et al (2018) Biological control using invertebrates and microorganisms: plenty of new opportunities. Biocontrol 63(1):39–59

    Article  Google Scholar 

  • Tuan S-J et al (2017) Economic injury level and demography-based control timing projection of Spodoptera litura (Lepidoptera: Noctuidae) at different growth stages of Arachis hypogaea. J Econom Entomol 110(2):755–762

    Article  Google Scholar 

  • Wei DAI et al (2019) Selectivity and sublethal effects of some frequently-used biopesticides on the predator Cyrtorhinus lividipennis Reuter (Hemiptera: Miridae). J Integrative Agric 18(1):124–133

    Article  Google Scholar 

  • Wilby A, Thomas MB (2002) Are the ecological concepts of assembly and function of biodiversity useful frameworks for understanding natural pest control? Agric For Entomol 4(4):237–43

    Article  Google Scholar 

  • Yu H, Zhong S, Agarwal RP (2011) Mathematics analysis and chaos in an ecological model with an impulsive control strategy. Commun Nonlinear Sci Numer Simul 16(2):776–786

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

Authors are grateful to the reviewer for their useful comments and suggestions that enrich the quality of the paper. Suvendu Das’s research is supported by the University Grants Commission, Government of India.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Fahad Al Basir.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Appendices

Appendix A

Coefficients of equation (11) The coefficients of (11) are given as follows:

$$\begin{aligned} A_1& {} = - F_{11}- F_{22}- F_{33}- F_{44}\\ A_2& {} = - F_{12} F_{21} + F_{11} F_{22} - F_{13} F_{31} - F_{23} F_{32} + F_{11} F_{33}\\&\quad+F_{22} F_{33} - F_{24} F_{42} + F_{11} F_{44} + F_{22} F_{44} + F_{33} F_{44}\\ A_3& {} = F_{13} F_{22} F_{31}- F_{12} F_{23} F_{31}- F_{13} F_{21} F_{32}\\&\quad+ F_{11} F_{23} F_{32}+ F_{12} F_{21} F_{33} - F_{11} F_{22} F_{33} - F_{14} F_{21} F_{42}\\&\quad + F_{11} F_{24} F_{42}+ F_{24} F_{33} F_{42}+ F_{12} F_{21} F_{44}\\&- F_{11} F_{22} F_{44}+ F_{13} F_{31} F_{44}+ F_{23} F_{32} F_{44}\\&\quad- F_{11} F_{33} F_{44}- F_{22} F_{33} F_{44}\\ A_4& {} = -F_{14} F_{23} F_{31} F_{42} + F_{13} F_{24} F_{31} F_{42}\\&\quad+ F_{14} {F_{21}} F_{33} F_{42} -F_{11} F_{24} F_{33} F_{42}\\&- F_{13} F_{22} F_{31} F_{44} + F_{12} F_{23} F_{31} F_{44} + F_{13} F_{21} F_{32} F_{44} \\&\quad- F_{11} F_{23} F_{32} F_{44}\\&\quad- F_{12} F_{21} F_{33} F_{44} + F_{11} F_{22} F_{33} F_{44}. \end{aligned}$$

Appendix B

Proof of Theorem 3:

Proof

Here, we want to check the stability of the system (4) for two cases.

(i) Release predator(P) and Virus(V) at same time interval \(\mathbf{t = n\tau }\).

Local Stability: Here, we discuss stability of the system through small amplitude perturbation method at the periodic solution \((0,0 ,P^*,V^*)\). Let,

$$\begin{aligned}&S(t)=\epsilon _1(t), I(t)=\epsilon _2(t),\nonumber \\&P(t)=P^*(t)+\epsilon _3(t),V(t)=V^*(t)+\epsilon _4(t). \end{aligned}$$
(25)

Here, \(\epsilon _1, \epsilon _2, \epsilon _3 \text{ and }\epsilon _4\) denote small amplitude perturbation.

First, we consider both predator and virus release at time \(t= n\tau\), then the corresponding system of (4) at \((0,0 ,P^*,V^*)\) is given as For \(t\ne n\tau\)

$$\begin{aligned} \frac{\mathrm{d}\epsilon _1}{\mathrm{d}t}& {} = r\epsilon _1(t)\left\{ 1-\frac{\epsilon _1(t)+\epsilon _2(t)}{K}\right\} -\lambda \epsilon _1(t) (V^*+\epsilon _4(t))-\alpha \epsilon _1(t) (P^*+\epsilon _3(t))\nonumber \\ \frac{\mathrm{d}\epsilon _2}{\mathrm{d}t}& {} = \lambda \epsilon _1(t) (V^*+\epsilon _4(t))-\delta \epsilon _2-\beta \epsilon _2(t) (P^*+\epsilon _3(t))\nonumber \\ \frac{\mathrm{d}\epsilon _3}{\mathrm{d}t}& {} = m \alpha \epsilon _1(t) (P^*+\epsilon _3(t))+m\beta \epsilon _2(t) (P^*+\epsilon _3(t))-d (P^*+\epsilon _3(t)) \nonumber \\ \frac{\mathrm{d}\epsilon _4}{\mathrm{d}t}& {} = n \delta \epsilon _2-\mu (V^*+\epsilon _4(t)). \end{aligned}$$
(26)

At \(t=n\tau\), i.e., for impulsive system,

$$\begin{aligned}&\triangle \epsilon _1(t) = 0,\nonumber \\&\triangle \epsilon _2(t)=0,\nonumber \\&\triangle (P^*(t)+\epsilon _3(t))=\rho ,\nonumber \\&\triangle (V^*(t)+\epsilon _4(t))=\nu . \end{aligned}$$
(27)

Since \(\epsilon _i(i=1,2,3,4)\) are very small, we can neglect second higher degree terms of \(\epsilon _i(i=1,2,3,4)\). Hence, the corresponding linear system of above system is given as, For \(t\ne n\tau\)

$$\begin{aligned} \frac{\mathrm{d}\epsilon _1}{\mathrm{d}t}& {} = r\epsilon _1(t)-\lambda \epsilon _1(t) V^*-\alpha \epsilon _1(t) P^*\nonumber \\ \frac{\mathrm{d}\epsilon _2}{\mathrm{d}t}& {} = \lambda \epsilon _1(t) V^*-\delta \epsilon _2-\beta \epsilon _2(t)P^*\nonumber \\ \frac{\mathrm{d}\epsilon _3}{\mathrm{d}t}& {} = m \alpha \epsilon _1(t) P^*+m\beta \epsilon _2(t) P^*-d (P^*+\epsilon _3(t)) \nonumber \\ \frac{\mathrm{d}\epsilon _4}{\mathrm{d}t}& {} = n \delta \epsilon _2-\mu (V^*+\epsilon _4(t)). \end{aligned}$$
(28)

At \(t=n\tau\), i.e., for impulsive system,

$$\begin{aligned}&\triangle \epsilon _1(t) = 0,\nonumber \\&\triangle \epsilon _2(t)=0,\nonumber \\&\triangle (P^*(t)+\epsilon _3(t))=\rho ,\nonumber \\&\triangle (V^*(t)+\epsilon _4(t))=\nu . \end{aligned}$$
(29)

Fundamental matrix M(t) of (28) will be \(\frac{\mathrm{d}M_f(t)}{\mathrm{d}t}=\left[ \begin{array}{ccccc} r(t)-\lambda (t) V^*-\alpha (t) P^* &{} 0 &{} 0 &{} 0 \\ \lambda V^* &{} -\delta -\beta P^*&{} 0 &{} 0 \\ m \alpha P^* &{} m\beta P^* &{} -d &{} 0 \\ 0 &{} n\delta &{} 0 &{} -\mu \\ \end{array}\right] .\)

with initial condition \(M(0)=I_4\)(the identity matrix). Now, fundamental solution matrix is given as \(M_f(t)=\left[ \begin{array}{ccccc} \exp \int _0^t{(r(t)-\lambda (t) V^*-\alpha (t) P^*)\mathrm{d}s}&{} 0 &{} 0 &{} 0 \\ \psi _1 &{} \exp\int _0^t{(-\delta -\beta P^*)\mathrm{d}s}&{} 0 &{} 0 \\ \psi _2 &{} \psi _3 &{} \exp(-dt) &{} 0 \\ 0 &{} \psi _4 &{} 0 &{} \exp(-\mu t) \\ \end{array}\right] .\) Here, \(\psi _i(t)(i=1,2,3,4)\) is not required to calculate for our further analysis. Now, the monodromy matrix of (28) and (29) is given as \(M(t)=I_4 M_f(t)\), where \(I_4\) is the identity matrix . According to Bionov et al. 1993[], the periodic solution \(E=(0,0,P^*,V^*)\) is local stable if and only if absolute value of all eigenvalue of \(M_(t)\) is less than one. The eigenvalue of \(M_(t)\) is

$$\begin{aligned}&\lambda _1(t)=\exp\int _0^t{(r(t)-\lambda (t) V^*-\alpha (t) P^*)\mathrm{d}s}\\&\lambda _2(t)=\exp\int _0^t{(-\delta -\beta P^*)\mathrm{d}s},\lambda _3(t)=\exp(-\mathrm{d}t), \\&\lambda _5(t)=\exp(-\mu t) \end{aligned}$$

Now, clearly, \(0<\lambda _2<1\), \(0<\lambda _3<1\) and \(0<\lambda _4<1\) thus the system (4) locally stable around the periodic solution if \(0<\lambda _1<1,\) i.e., if

$$\begin{aligned} \exp\int _0^t{(r(t)-\lambda (t) V^*-\alpha (t) P^*)\mathrm{d}s}<1 \end{aligned}$$
(30)

Global stability From Eq. (30), we can choose \(\delta _1>0\) such that,

$$\begin{aligned} \eta _1=\exp\int _{n\tau }^{(n+1)\tau }(r(t)-\lambda (t) (V^*-\delta _1)-\alpha (t) (P^*-\delta _1))\mathrm{d}s<1 \end{aligned}$$

Since all state variable are positive,

$$\begin{aligned} \frac{\mathrm{d}P}{\mathrm{d}t} \ge -\mathrm{d}P;~ \frac{\mathrm{d}V}{\mathrm{d}t} \ge -\mu V \end{aligned}$$

Thus, according to comparison theorem and Eqs. (21) and (22), for \(\delta _1>0\) there exist \(t_0>0\) such that \(P(t)\ge P^* - \delta _1\), \(V(t)\ge V^* - \delta _1\) for all \(t>t_0\).

Now, from the first equation of the system, (4) can be written as

$$\begin{aligned} \dot{S(t)} & {} \le S(t)(r(t)-\lambda (t) (V^*-\delta )-\alpha (t) (P^*-\delta _1)),~ t\ne n\tau \nonumber \\ S(t^+)& {} = S(t),~ t= n\tau \end{aligned}$$
(31)

Integrating (31) on \([n\tau , (n+1)\tau ]\), it can be shown that

$$\begin{aligned} S\{(n+1)\tau \} & {} \le S(n\tau )\exp\int _{n\tau }^{(n+1)\tau }(r(t)-\lambda (t) (V^*-\delta _1)\nonumber \\&-\alpha (t) (P^*-\delta _1)){\rm d}t,\nonumber \\& {} = S(n\tau )\eta _1 \end{aligned}$$
(32)

Similarly,

$$\begin{aligned} S\{n\tau \} & {} \le S\{(n-1)\tau \}\eta _1. \eta _1 \end{aligned}$$
(33)

Hence, from (32) and (33),

$$\begin{aligned} S\{(n+1)\tau \} & {} \le S\{(n-1)\tau \}\eta _1^2 \end{aligned}$$

Proceeding in this way, we get

$$\begin{aligned} S\{(n+1)\tau \} & {} \le S(\tau )\eta _1^n \end{aligned}$$
(34)

Since \(\eta _1<1\), hence\(\eta _1^n\rightarrow 0\), whenever \(n\rightarrow \infty\). Hence, \(S\{(n+1)\tau \}\rightarrow 0\) as \(n\rightarrow \infty\). Now, we take \(n\tau <t~\le ~(n+1)\tau\) then clearly \(0<S(t)~\le ~S(n\tau )~\exp~{( r~\tau )}\). Thus, \(S(t)\rightarrow 0\) as \(t\rightarrow \infty\). Again since

$$\begin{aligned} \eta _2=\exp\int _{n\tau }^{(n+1)\tau }(-\delta _1 -\beta (P^*-\delta _1))\mathrm{d}t<1, \end{aligned}$$

similarly we can prove that \(I(t)\rightarrow 0\) as \(t\rightarrow \infty\).

Now, we prove that \(P(t)\rightarrow P^*(t)\) as \(t\rightarrow \infty\). Since \(S(t)\rightarrow 0\) and \(I(t)\rightarrow 0\) as \(t\rightarrow \infty\), then for some \(0<\delta _2<\frac{d}{m(\alpha +\beta )}\) there exist \(t_1>0\) such that \(0<S(t)<\delta _2\) and \(0<I(t)<\delta _2\), for all \(t>t_1\). Thus, for \(t>t_1\), from third equation of the system (4), we can write

$$\begin{aligned} (m \alpha \delta _2+ m \beta \delta _2 -d)P(t)\ge {\dot{P}}(t)\ge (-m \alpha \delta _2- m \beta \delta _2 -d)P(t) \end{aligned}$$
(35)

Let \(\acute{P}(t)\) and \(\acute{\acute{P}}(t)\) be the solution of the following equation, respectively,

$$\begin{aligned} \dot{ \acute{P}}(t)& {} = (-m \alpha \delta _2- m \beta \delta _2 -d)\acute{\acute{P}}(t), ~t\ne n\tau \\ \acute{P}(t^+)& {} = \acute{P}(t)+\rho ~t=n\tau \end{aligned}$$

and

$$\begin{aligned} \dot{ \acute{P}}(t)& {} = (m \alpha \delta _2+ m \beta \delta _2 -d)\acute{\acute{P}}(t), ~t\ne n\tau \\ \acute{P}(t^+)& {} = \acute{P}(t)+\rho ~t=n\tau \end{aligned}$$

Then, the solution will be

$$\begin{aligned} \acute{P}^*(t)& {} = \frac{\rho e^{-(m \alpha \delta _2+ m \beta \delta _2 +d) (t-n\tau )}}{1-e^{-(m \alpha \delta _2+ m \beta \delta _2 +d) \tau }},\nonumber \\ \acute{\acute{P}}^*(t)& {} = \frac{\rho e^{(m \alpha \delta _2+ m \beta \delta _2 -d) (t-n\tau )}}{1-e^{(m \alpha \delta _2+ m \beta \delta _2 -d) \tau }}, \end{aligned}$$
(36)

From (36), it is clear that when \(\delta _2\rightarrow 0,\) then, \(\acute{P}^*(t)\rightarrow P^*(t)\) and \(\acute{\acute{P}}^*(t)\rightarrow P^*(t)\). Hence, it follows from (35) that \(P(t)\rightarrow P^*(t)\) as \(t\rightarrow \infty\).

Similarly, we can choose \(0<\delta _3<\frac{\mu }{ n \delta }\) and in the same way, we can prove \(V(t)\rightarrow V^*(t)\) as \(t\rightarrow \infty\).

From Eqs. (21), (22) and (30), we can say that the system (4) is locally, as well as globally, stable if

$$\begin{aligned} r t-\frac{\lambda \nu e^{n\mu \tau }( 1-e^{-\mu t})}{\mu (1-e^{-d \tau })}-\frac{\alpha \rho e^{d n\tau } (1-e^{-d t})}{d(1-e^{-\mu \tau })}<0. \end{aligned}$$
(37)

(ii) Release predator(P) and Virus(V) at different time interval: In this case, two subcases arise.

Subcase I: Release predator(P) at time \(t\mathbf{=n\tau _1}\)

In this subcase, \(\nu =0\), hence, the system (4) is locally stable around the periodic solution if

$$\begin{aligned} r t-\frac{\alpha \rho e^{d n\tau _1} (1-e^{-d t})}{d(1-e^{-\mu \tau _1})}<0. \end{aligned}$$
(38)

Subcase II: Release Virus(V) at time \(t\mathbf{=n\tau _2}\) In this subcase, \(\rho =0\), hence, the system (4) is locally stable around the periodic solution if

$$\begin{aligned} r t-\frac{\lambda \nu e^{n\mu \tau _2}( 1-e^{-\mu t})}{\mu (1-e^{-d \tau _2})}<0. \end{aligned}$$
(39)

\(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Al Basir, F., Chowdhury, J., Das, S. et al. Combined impact of predatory insects and bio-pesticide over pest population: impulsive model-based study. Energ. Ecol. Environ. 7, 173–185 (2022). https://doi.org/10.1007/s40974-021-00226-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s40974-021-00226-1

Keywords

Navigation