Skip to main content

Advertisement

Log in

Adaptive Release of Natural Enemies in a Pest-Natural Enemy System with Pesticide Resistance

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

Abstract

Integrated pest management options such as combining chemical and biological control are optimal for combating pesticide resistance, but pose questions if a pest is to be controlled to extinction. These questions include (i) what is the relationship between the evolution of pesticide resistance and the number of natural enemies released? (ii) How does the cumulative number of natural enemies dying affect the number of natural enemies to be released? To address these questions, we developed two novel pest-natural enemy interaction models incorporating the evolution of pesticide resistance. We investigated the number of natural enemies to be released when threshold conditions for the extinction of the pest population in two different control tactics are reached. Our results show that the number of natural enemies to be released to ensure pest eradication in the presence of increasing pesticide resistance can be determined analytically and depends on the cumulative number of dead natural enemies before the next scheduled release time.

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

Similar content being viewed by others

References

  • Artstein, Z. (1980). Discrete and continuous bang-bang and facial spaces, or: Look for the extreme points. SIAM Rev., 22, 172–185.

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  • Beddington, J. R., Free, C. A., & Lawton, J. R. (1978). Characteristics of successful natural enemies in models of biological control of insect pests. Nature, 273, 513–519.

    Article  Google Scholar 

  • Bonhoeffer, S., & Nowak, M. A. (1997). Pre-existence and emergence of drug resistance in HIV-1 infection. Proc. R. Soc. Lond. B, 264, 631–637.

    Article  Google Scholar 

  • Debach, P. (1974). Biological control by natural enemies. London: Cambridge University Press.

    Google Scholar 

  • Flint, M. L. (1987). Integrated pest management for walnuts. Statewide Integrated Pest Management Project, Division of Agriculture and Natural Resources (2nd ed., p. 3270). Oakland: University of California.

    Google Scholar 

  • Georghiou, G. P. (1990). Overview of insecticide resistance. In M. B. Green, H. M. LeBaron, & W. K. Moberg (Eds.), Managing resistance to agrochemicals: from fundamental research to practical strategies (pp. 18–41). Washington: American Chemical Society.

    Chapter  Google Scholar 

  • Gourley, S. A., Liu, R. S., & Wu, J. H. (2007). Eradicating vector-borne diseases via age-structured culling. J. Math. Biol., 54, 309–335.

    Article  MathSciNet  MATH  Google Scholar 

  • Greathead, D. J. (1992). Natural enemies of tropical locusts and grasshoppers: their impact and potential as biological control agents. In C. J. Lomer & C. Prior (Eds.), Biological control of locusts and grasshoppers (pp. 105–121). Wallingford: C.A.B. International.

    Google Scholar 

  • Gubbins, S., & Gilligan, C. A. (1999). Invasion thresholds for fungicide resistance: deterministic and stochastic analyses. Proc. R. Soc. Lond. B, 266, 2539–2549.

    Article  Google Scholar 

  • Hall, R. J., Gubbins, S., & Gilligan, C. A. (2004). Invasion of drug and pesticide resistance is determined by a trade-off between treatment efficacy and relative fitness. Bull. Math. Biol., 66, 825–840.

    Article  Google Scholar 

  • Hoffmann, M. P., & Frodsham, A. C. (1993). Natural enemies of vegetable insect pests. Cooperative Extension (p. 63). Ithaca: Cornell University.

    Google Scholar 

  • Kotchen, M. J. (1999). Incorporating resistance in pesticide masnagement: a dynamic regional approach (pp. 126–135). New York: Springer.

    Google Scholar 

  • Laxminarayan, R., & Simpson, R. D. (2000). Biological limits on agricultural intensification: an example from resistance management. Resources for the Future, 47–60.

  • Liang, J. H., & Tang, S. Y. (2010). Optimal dosage and economic threshold of multiple pesticide applications for pest control. Math. Comput. Model., 51, 487–503.

    Article  MathSciNet  MATH  Google Scholar 

  • Lilley, R., & Campbell, C. A. M. (1999). Biological, chemical and integrated control of two-spotted spider mite Tetranychus urticae on dwarf hops. Biocontrol Sci. Technol., 9, 467–473.

    Article  Google Scholar 

  • Milgroom, M. G., Levin, S. A., & Fry, W. E. (1989). Population genetics theory and fungicide resistance. In K. J. Leonard & W. E. Fry (Eds.), Plant Disease Epidemiology 2: Genetics, Resistance and Management (pp. 340–367). New York: McGraw-Hill.

    Google Scholar 

  • Milgroom, M. G. (1990). A stochastic model for the initial occurrence and development of fungicide resistance in plant pathogen populations. Phytopathology, 80, 410–416.

    Article  Google Scholar 

  • Neuenschwander, P., & Herren, H. R. (1988). Biological control of the Cassava Mealybug, Phenacoccus manihoti, by the exotic parasitoid Epidinocarsis lopezi in Africa. Philos. Trans. R. Soc. Lond. B, Biol. Sci., 318, 319–333.

    Article  Google Scholar 

  • Oi, D. H., Williams, D. F., Pereira, R. M., Horton, P. M., Davis, T. S., Hyder, A. H., Bolton, H. T., Zeichner, B. C., Porter, S. D., Hoch, A. L., Boswell, M. L., & Williams, G. (2008). Combining biological and chemical controls for the management of red imported fire ants (Hymenoptera: Formicidae). Amat. Entomol., B07-011, 46–55.

    Article  Google Scholar 

  • Onstad, D. W. (2008). Insect Resistance Management. Amsterdam: Elsevier.

    Google Scholar 

  • Parker, F. D. (1971). Management of pest populations by manipulating densities of both host and parasites through periodic releases. In C. B. Huffaker (Ed.), Biological control. New York: Plenum.

    Google Scholar 

  • Peck, S. L., & Ellner, S. P. (1997). The effect of economic thresholds and life-history parameters on the evolution of pesticide resistance in a regional setting. Am. Nat., 149, 43–63.

    Article  Google Scholar 

  • Ruberson, J. R., Nemoto, H., & Hirose, Y. (1998). Pesticides and conservation of natural enemies in pest management. In P. Barbosa (Ed.), Conservation Biological Control (pp. 207–220). New York: Academic Press.

    Chapter  Google Scholar 

  • Karaagac, S. U. (2012). Insecticide Resistance. In Farzana Perveen, Insecticides and Advances in Integrated Pest Management (pp. 469–478). London: InTech Press.

    Google Scholar 

  • Tang, S. Y., Xiao, Y. N., Chen, L. S., & Cheke, R. A. (2005). Integrated pest management models and their dynamical behaviour. Bull. Math. Biol., 67, 115–135.

    Article  MathSciNet  Google Scholar 

  • Tang, S. Y., Xiao, Y. N., & Cheke, R. A. (2008). Multiple attractors of host-parasitoid models with integrated pest management strategies: Eradication, persistence and outbreak. Theor. Popul. Biol., 73, 181–197.

    Article  MATH  Google Scholar 

  • Tang, S. Y., & Cheke, R. A. (2008). Models for integrated pest control and their biological implications. Math. Biosci., 215, 115–125.

    Article  MathSciNet  MATH  Google Scholar 

  • Tang, S. Y., Xiao, Y. N., & Cheke, R. A. (2009). Effects of predator and prey dispersal on success or failure of biological control. Bull. Math. Biol., 71, 2025–2047.

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Tang, S. Y., Liang, J. H., Tan, Y. S., & Cheke, R. A. (2013). Threshold conditions for integrated pest management models with pesticides that have residual effects. J. Math. Biol., 66, 1–35.

    Article  MathSciNet  MATH  Google Scholar 

  • Tang, S. Y., Liang, J. H., Xiao, Y. N., & Cheke, R. A. (2012). Sliding bifurcations of Filippov two stage pest control models with economic thresholds. SIAM J. Appl. Math., 72, 1061–1080.

    Article  MathSciNet  MATH  Google Scholar 

  • Tang, S. Y., & Liang, J. H. (2013). Global qualitative analysis of a non-smooth Gause predator-prey model with a refuge. Nonlinear Anal., 76, 165–180.

    Article  MathSciNet  MATH  Google Scholar 

  • Terry, A. J., & Gourley, S. A. (2010). Perverse Consequences of Infrequently Culling a Pest. Bull. Math. Biol., 72, 1666–1695.

    Article  MathSciNet  MATH  Google Scholar 

  • Thomas, M. B. (1999). Ecological approaches and the development of “truly integrated” pest management. Proc. Natl. Acad. Sci., 96, 5944–5951.

    Article  Google Scholar 

  • Van Lenteren, J. C., & Woets, J. (1988). Biological and integrated pest control in greenhouses. Annu. Rev. Entomol., 33, 239–250.

    Article  Google Scholar 

  • Van Lenteren, J. C. (1995). Integrated pest management in protected crops. In D. Dent (Ed.), Integrated pest management (pp. 311–320). London: Chapman & Hall.

    Google Scholar 

  • Van Lenteren, J. C. (2000). Measures of success in biological control of arthropods by augmentation of natural enemies. In S. Wratten & G. Gurr (Eds.), Measures of Success in Biological Control (pp. 77–89). Dordrecht: Kluwer Academic.

    Chapter  Google Scholar 

  • Vance, R. R., & Coddington, E. A. (1989). A nonautonomous model of population growth. Math. Biosci., 27, 491–506.

    MathSciNet  MATH  Google Scholar 

  • Zchori-Fein, E., Roush, R. T., & Sanderson, J. P. (1994). Potential for integration of biological and chemical control of greenhouse whitefly(Homoptera: Aleyrodidae) using Encarsia formosa (Hymenoptera: Aphelinidae) and abamectin. Environ. Entomol., 23, 1277–1282.

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by the Fundamental Research Funds for the Central Universities (GK201104009), and by the National Natural Science Foundation of China (NSFC, 11171199), and by the International Development Research Centre.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Juhua Liang or Sanyi Tang.

Appendices

Appendix A: The Deduction of Eq. (2) and Its Solution

Because P s =ωP and P r =(1−ω)P, and d 1 is the mortality rate of susceptible pests, d 2 is the mortality rate of resistant pests when the pesticide is sprayed, so we have

$$\left \{ \begin{aligned} &\frac{dP_s}{dt}=\omega rP(1-\eta P)-d_1P_s, \\ &\frac{dP_r}{dt}=(1-\omega) rP(1-\eta P)-d_2P_r. \end{aligned} \right . $$

However, for simplification we assume that the resistant pests display near-complete resistance to the pesticide, which means that d 2≈0. Therefore,

$$\frac{dP}{dt}=\frac{dP_s}{dt}+\frac{dP_r}{dt}=rP(1-\eta P)-\omega d_1P. $$

Then we get

$$\frac{d\omega}{dt} =\frac{d}{dt} \biggl(\frac{P_s}{P} \biggr)=d_1\omega(\omega-1). $$

This resistance evolution equation has been widely used recently in different fields (Hall et al. 2004; Bonhoeffer and Nowak 1997; Gubbins and Gilligan 1999; Laxminarayan and Simpson 2000; Milgroom et al. 1989).

In practice, the evolution of pesticide resistance is dependent on the dosage of the pesticide applications, the frequency of applications or the pesticide application period. Although linking the evolution of pesticide resistance to the pest growth model is a great challenge, it is well known that the less frequent are the pesticide applications (i.e. the longer the period between them), the slower the development of the pest resistance. One possible way is to consider the effect of each pulse spraying of pesticides on the evolution of pesticide resistance as a perturbation constant, i.e. we have

$$\frac{d\omega(t)}{dt} =d_1\omega \bigl(\omega^{q_i}-1 \bigr), \quad \tau_{i-1}\leq t\leq \tau_i,\ i\in \mathcal{N}, $$

with initial value ω(τ i−1) at each time interval t∈[τ i−1,τ i ] and ω(τ 0)=ω(0)=ω 0 is given. Where q i denotes the perturbation constant of the ith pulse spraying of pesticides, which depends the number of pesticide applications, dosage X i of the ith pesticide application and time interval Δτ i . This equation resembles the classical Richards equation. As such, we anticipate the approach in resistant equation can be used to describe analytically the functional format of q i .

The analytical solution of ω(t) can be determined as follows:

$$ \omega(t)= \bigl(1+e^{q_id_1(t-\tau_{i-1})} \bigl(\bigl(\omega( \tau_{i-1})\bigr)^{-q_i} -1 \bigr) \bigr)^{-\frac{1}{q_i}},\quad \tau_{i-1}\leq t\leq\tau_i, $$
(22)

which indicates that

$$ \omega(\tau_{i})= \bigl(1+e^{id_1} \bigl( \omega(\tau_{i-1})^{-q_i} -1 \bigr) \bigr)^{-1/q_i}. $$
(23)

For simplification, we assume that the same dosage of pesticides is applied each time, and without loss of generality we let X i =1 for all \(i\in \mathcal{N}\). It follows from the main text that one possible definition of q i is as follows: q i =iτ i , and then Δτ i =T for all \(i\in \mathcal{N}\). For this special case, the evolution of ω at each time point nT can be expressed as

$$ \omega(nT)= \bigl(1+e^{nd_1} \bigl(\omega\bigl((n-1)T \bigr)^{-n/T} -1 \bigr) \bigr)^{-T/n},\quad n\in \mathcal{N}. $$
(24)

Appendix B: The Proof of Theorem 2.1

For any given ε>0 (ε small enough), we will first show that there exists a τ(ε)>0 such that

$$ P(\tau)\leq \varepsilon e^{-\beta T}. $$
(25)

Otherwise, P(t)≥εexp(−βT) for all t≥0. It follows from (b) that ∂F/∂P≤0 and then

$$F\bigl(s, P(s)\bigr)\leq F \bigl(s, \varepsilon e^{-\beta T} \bigr),\quad \mbox{for } s\geq 0. $$

Since

$$F \bigl(s,\varepsilon e^{-\beta T} \bigr)-F \biggl(s,\frac{\varepsilon}{2} e^{-\beta T} \biggr)=\frac{\varepsilon}{2} e^{-\beta T}\frac{\partial F}{\partial P} \bigl(s,\eta (s)\bigr), $$

where \(\frac{\varepsilon}{2} \exp{(-\beta T)}<\eta (s)<\varepsilon \exp{(-\beta T)}\). Again assumption (b) implies that

$$\frac{\partial F}{\partial P}\bigl(s,\eta (s)\bigr)\leq -\varphi_0 \lambda(s), $$

where φ 0=min{φ(P): exp(−βT)ε/2≤Pεexp(−βT)}. Thus, we get

$$F\bigl(s, P(s)\bigr)\leq F\biggl(s,\frac{\varepsilon}{2} e^{-\beta T}\biggr)- \frac{\varepsilon}{2} e^{-\beta T}\varphi_0 \lambda(s)\leq F(s,0)- \frac{\varepsilon}{2} e^{-\beta T}\varphi_0 \lambda(s). $$

For any t≥0, there is a \(h\in \mathcal{N}\), such that (h−1)T<thT. Therefore,

$$\begin{aligned} \varepsilon e^{-\beta T} \leq&P(t) \\ =&\prod^{h-1}_{i=1}q(iT)P_0 \exp \biggl(\!\biggl(\int^{T}_{0}+\int ^{2T}_{T}+\cdots +\int^{(h-1)T}_{(h-2)T}+ \int^{t}_{(h-1)T}\biggr)F\bigl(s,P(s)\bigr)ds \biggr) \\ \leq& \prod^{h-1}_{i=1}q(iT)P_0 \exp \biggl(\biggl(\int^{T}_{0}+\cdots +\int ^{(h-1)T}_{(h-2)T}+\int^{t}_{(h-1)T} \biggr)\\ &{}\times\biggl(F(s,0)-\frac{\varepsilon}{2} e^{-\beta T}\varphi_0 \lambda(s) \biggr)ds \biggr) \\ =&\prod^{h-1}_{i=1}q(iT)P_0 \exp \biggl(\biggl(\int^{T}_{0}+\cdots +\int ^{t}_{(h-1)T}\biggr)F(s,0)ds \biggr)\\ &{}\times\exp \biggl(\int ^{t}_{0}\biggl(-\frac{\varepsilon}{2} e^{-\beta T}\varphi_0 \lambda(s)\biggr)ds \biggr). \end{aligned}$$

For ε>0 small enough and assumption (c) yields

$$q(iT)\exp{\int^{iT}_{(i-1)T}F \biggl(s, \frac{\varepsilon}{2} e^{-\beta T} \biggr)ds}\leq 1,\quad i=1,2,\ldots,h, $$

and F(s,0)≤β. Hence,

$$\varepsilon e^{-\beta T}\leq P(t)\leq P_0e^{\beta T}\exp \biggl(\int^{t}_{0}\biggl(-\frac{\varepsilon}{2} e^{-\beta T}\varphi_0 \lambda(s)\biggr)ds \biggr)\to 0 $$

as t→∞, in particular, P(t)<εexp(−βT) for t large enough. This contradiction establishes that (25) is true for some τ.

Now we will prove that

$$ P(t)\leq\varepsilon\quad \mbox{for } t\geq \tau. $$
(26)

Suppose that P(τ 1)>εεexp(−βT) for some τ 1>τ and there is a \(n\in \mathcal{N}\), such that (n−1)Tτ 1<nT. Then there exists a t 1∈(τ,τ 1), such that

$$P(t_1)=\varepsilon e^{-\beta T}, $$

and \(P(t)>\varepsilon \exp{(-\beta T)},\ \rm{for}\ t\in (t_{1},\tau_{1}] \), which means that for any \(k\in \mathcal{N}\), t 1kT, due to 0≤q(kT)<1. Therefore, either t 1>(n−1)T or t 1<(n−1)T. If t 1>(n−1)T, (this indicates τ 1>(n−1)T), then τ 1t 1<T and

$$\begin{aligned} \varepsilon<P(\tau_1) =&P(t_1)\exp \biggl(\int ^{\tau_1}_{t_1}F\bigl(s,P(s)\bigr)ds \biggr) \\ \leq &\varepsilon e^{-\beta T}\exp \biggl(\int^{\tau_1}_{t_1}F \bigl(s,\varepsilon e^{-\beta T}\bigr)ds \biggr) \\ \leq &\varepsilon e^{-\beta T}e^{\beta T}=\varepsilon, \end{aligned}$$

this is a contradiction. So t 1<(n−1)T, then there is a \(m\in \mathcal{N}\), m<n such that (m−1)T<t 1<mT (let t 1=(m−1)T+l, 0<l<T) and τ 1t 1<T, otherwise, τ 1t 1T, i.e. τ 1T+t 1, and P(t)>εexp(−βT), t∈(t 1,t 1+T]. Therefore, (m−1)T<t 1<mT<t 1+T<(m+1)T, and solving (3) from t 1 to t 1+T, we get

$$\begin{aligned} \varepsilon e^{-\beta T} <&P(t_1+T) \\ =&q(mT)P(t_1)\exp \biggl(\int^{mT}_{t_1}F \bigl(s,P(s)\bigr)ds \biggr) \exp \biggl(\int_{mT}^{t_1+T}F \bigl(s,P(s)\bigr)ds \biggr) \\ \leq&q(mT)\varepsilon e^{-\beta T}\exp \biggl(\int^{t_1+T}_{t_1}F \bigl(s,\varepsilon e^{-\beta T}\bigr)ds \biggr) \\ =&q(mT)\varepsilon e^{-\beta T}\exp \biggl(\int^{mT+l}_{(m-1)T+l}F \bigl(s,\varepsilon e^{-\beta T}\bigr)ds \biggr) \\ \leq&\varepsilon e^{-\beta T}, \end{aligned}$$

this is a contradiction, thus τ 1t 1<T. Therefore, if τ 1>(n−1)T, then

$$\begin{aligned} \varepsilon <&P(\tau_1) \\ =&q\bigl((n-1)T\bigr)P(t_1)\exp \biggl(\int^{(n-1)T}_{t_1}F \bigl(s,P(s)\bigr)ds \biggr) \exp \biggl(\int_{(n-1)T}^{\tau_1}F \bigl(s,P(s)\bigr)ds \biggr) \\ =&q\bigl((n-1)T\bigr)P(t_1)\exp \biggl(\int_{t_1}^{\tau_1}F \bigl(s,P(s)\bigr)ds \biggr) \\ \leq&q\bigl((n-1)T\bigr)P(t_1)\exp \biggl(\int _{t_1}^{\tau_1}F\bigl(s,\varepsilon e^{-\beta T} \bigr)ds \biggr) \\ \leq&\varepsilon e^{-\beta T}e^{\beta T}=\varepsilon, \end{aligned}$$

it is a contradiction, if τ 1=(n−1)T, then

$$\begin{aligned} \varepsilon <&P(\tau_1) \\ =&P(t_1)\exp \biggl(\int^{(n-1)T}_{t_1}F \bigl(s,P(s)\bigr)ds \biggr) \\ \leq&P(t_1)\exp \biggl(\int_{t_1}^{(n-1)T}F \bigl(s,\varepsilon e^{-\beta T}\bigr)ds \biggr) \\ \leq&\varepsilon e^{-\beta T}e^{\beta T}=\varepsilon. \end{aligned}$$

This is a contradiction. Hence, such τ 1 does not exist. Therefore, P(t)≤ε for tτ, that is lim t→∞ P(t)=0. The proof is complete.

Appendix C: The Proof of Theorem 3.1

It is seen from the second equation of system (4) that dN(t)/dt>−dN(t). Considering the following impulsive differential equation:

$$ \left \{ \begin{aligned} &\frac{dy(t)}{dt}=-dy(t), \quad t\neq nT, \\ &y\bigl(t^+\bigr)=y(t)+\delta_n,\quad t=nT, \\ &y\bigl(0^+\bigr)=N_0+\delta_0. \end{aligned} \right . $$
(27)

According to the comparison theorem on impulsive differential equations, B1 yields N(t)≥y(t)=N (t). It follows from the first equation of system (4) that

$$\frac{dP(t)}{dt}\leq rP(t) \bigl(1-\eta P(t)\bigr)-\beta P(t)N^*(t). $$

Now we consider the following impulsive differential equation

$$ \left \{ \begin{aligned} &\frac{dx(t)}{dt}= rx(t) \bigl(1-\eta x(t) \bigr)-\beta x(t)N^*(t),\quad t\neq nT, \\ &x\bigl(nT^{+}\bigr)=\bigl(1-\omega (nT) d_1\bigr)x(nT), \quad t= nT, \\ &\frac{d\omega(t)}{dt}=d_1\omega(t) \bigl(\omega(t)^{q_n}-1 \bigr), \\ &x\bigl(0^+\bigr)=P(0)\doteq P_0. \end{aligned} \right . $$
(28)

Again according to the comparison theorem on impulsive differential equations we have P(t)≤x(t).

By using the formula of (24), we can easily have

$$\begin{aligned} q(nT) \doteq &1-\omega (nT)d_1 \\ =&1-\frac{d_1}{ (1+e^{nd_1} ((\omega((n-1)T))^{-\frac{n}{T}} -1 ) )^{\frac{T}{n}}} \end{aligned}$$

and

$$F(s,x)\doteq r-r\eta x(s)-\beta N^*(s). $$

Now we test and verify the conditions of Theorem 2.1. It is easy to see that condition (a) holds true naturally, and

$$F(s,0)=r-\beta N^*(s)\leq r $$

and

$$\frac{\partial F(s,x)}{\partial x}=-r\eta,\qquad \int^{\infty}_{0}\eta ds=\infty. $$

Therefore,

$$\begin{aligned} \exp \biggl(\int^{l+nT}_{l+(n-1)T}F(s,0)ds \biggr) =& \exp \biggl(\int^{l+nT}_{l+(n-1)T}r-\beta N^*(s)ds \biggr) \\ =&e^{rT}\exp \biggl(\int^{l+nT}_{l+(n-1)T} \bigl(-\beta N^*(s)\bigr)ds \biggr) \end{aligned}$$

with

$$\begin{aligned} &\exp \biggl(\int^{l+nT}_{l+(n-1)T}\bigl(-\beta N^*(s) \bigr)ds \biggr) \\ &\quad{}=\exp \Biggl( \Biggl(\frac{\beta N_0}{d}e^{-d(l+(n-1)T)}+\sum ^{n-1}_{i=0}\frac{\beta \delta_i}{d}e^{-d(l+(n-1-i)T)} \Biggr) \bigl(e^{-dT}-1 \bigr)\\ &\qquad{}+\frac{\beta \delta_n}{d}\bigl(e^{-dl}-1\bigr) \Biggr) \\ &\quad{}=\exp{ \bigl(M(n,l) \bigr)}. \end{aligned}$$

Thus,

$$\begin{aligned} &q(nT)\exp{\biggl(\int^{l+nT}_{l+(n-1)T}F(s,0)ds\biggr)} \\ &\quad{}=R_0(n,T)\exp{ \bigl(M(n,l) \bigr)} \\ &\quad{}\doteq R_0^{N}(n,T,l). \end{aligned}$$

According to Theorem 2.1, we can see that if \(R_{0}^{N}(n,T,l)\leq 1\), then x(t)→0 as t→∞. Consequently, we have P(t)→0 as t→∞ provided \(R_{0}^{N}(n,T,l)\leq1\).

Next, we prove that N(t)→N (t) as t→∞. For any 0<ε<d/(λβ), there exists a t 1>0 such that 0<P(t)<ε for all tt 1. Without loss of generality, we may assume that 0<P(t)<ε holds true for all t>0, then we have

$$-dN(t)\leq\frac{dN(t)}{dt}\leq(\lambda\beta\varepsilon-d)N(t). $$

For the left-hand inequality, it follows from impulsive differential equation (27) that N(t)≥y(t)=N (t). For the right-hand inequality, considering the following impulsive differential equation:

$$ \left \{ \begin{aligned} &\frac{dz(t)}{dt}=(\lambda\beta \varepsilon-d)z(t), \quad t\neq nT, \\ &z\bigl(t^+\bigr)=z(t)+\delta_n,\quad t=nT, \\ &z\bigl(0^+\bigr)=N_0+\delta_0. \end{aligned} \right . $$
(29)

The analytical solution of the above system at any impulsive interval ((n−1)T,nT] gives

$$ z^*(t)=N_0e^{(\lambda\beta\varepsilon-d)t}+\sum ^{n-1}_{i=0}\delta_ie^{(\lambda\beta\varepsilon-d)(t-iT)}, \quad (n-1)T<t\leq nT. $$
(30)

Therefore, for any ε 1>0, there exists a t 2>0 such that

$$N^*(t)-\varepsilon_1<N(t)<z^*(t)+\varepsilon_1 $$

for t>t 2. Let ε→0, then we have

$$N^*(t)-\varepsilon_1<N(t)<N^*(t)+\varepsilon_1 $$

for t>t 2, which indicates that N(t)→N (t) as t→∞. Therefore, the pest-free solution (7) is globally attractive if \(R_{0}^{N}(n,T,l)\leq1\). The proof is complete.

Appendix D: Determining the New Number of Natural Enemies to be Released for Case 3.2

In this case \(R^{N}_{0_{\max}}(n,T,l)=R_{0}^{N}(n,T,T)= R_{C}\). It follows from (8) that we have M(n,T)=ln(R c /R 0(n,T)), that is

$$\Biggl(\frac{\beta N_0}{d}e^{-dnT}+\sum^{n}_{i=0} \frac{\beta \delta_i}{d}e^{-d(n-i)T} \Biggr) \bigl(e^{-dT}-1 \bigr)=\ln \biggl(\frac{R_C}{R_0(n,T)} \biggr), $$

this indicates that

$$ G_{n^{'}}+\sum^{n}_{i=n^\prime+1} \delta_i e^{diT}= -\frac{de^{dnT}}{\beta (1-e^{-dT} )}\ln \biggl( \frac{R_C}{R_0(n,T)} \biggr). $$
(31)

That is

$$\sum^{n}_{i=n^\prime+1} \delta_i e^{diT}= -\frac{de^{dnT}}{\beta (1-e^{-dT} )}\ln \biggl(\frac{R_C}{R_0(n,T)} \biggr)-G_{n^{'}}\doteq A_n. $$

Therefore, when n=n′+1 we get

$$\delta_{n^\prime+1}=A_{n^\prime+1}e^{-d(n^\prime+1)T}. $$

When n=n′+2, we have

$$\delta_{n^\prime+1}e^{d(n^\prime+1)T} +\delta_{n^\prime+2}e^{d(n^\prime+2)T}=A_{n^\prime+2}, $$

that is

$$\delta_{n^\prime+2}=(A_{n^\prime+2}-A_{n^\prime+1})e^{-d(n^\prime+2)T}. $$

Similarly, when n=n′+3 we get

$$\delta_{n^\prime+1}e^{d(n^\prime+1)T} +\delta_{n^\prime+2}e^{d(n^\prime+2)T}+ \delta_{n^\prime+3}e^{d(n^\prime+3)T}=A_{n^\prime+3}, $$

that is

$$\delta_{n^\prime+3}=(A_{n^\prime+3}-A_{n^\prime+2})e^{d(n^\prime+3)T}. $$

By induction, the new number of natural enemies to be released δ n can be determined as follows:

$$\delta_n=\left \{\begin{array}{l} \delta_c, \quad \mbox{if } n\leq n^{\prime},\\ A_{n^{\prime}+1}e^{-d(n^\prime+1)T},\quad n=n^{\prime}+1,\\ (A_{n}-A_{n-1})e^{-dnT},\quad \mbox{if } n>n^{\prime}+1. \end{array} \right . $$

Appendix E: The Proof of Theorem 4.1

It is seen from the second equation of system (15) that dN(t)/dt>−dN(t). Considering the following impulsive differential equation:

$$ \left \{ \begin{aligned} &\frac{dy(t)}{dt}=-dy(t), \quad t \neq hT_N, \\ &y\bigl(t^+\bigr)=y(t)+\delta_h,\quad t=hT_N, \\ &y\bigl(0^+\bigr)=N_0.\end{aligned} \right . $$
(32)

According to the comparison theorem on impulsive differential equations, (E1) yields N(t)≥y(t)=N (t). It follows from the first equation of system (15) that

$$\frac{dP(t)}{dt}\leq rP(t) \bigl(1-\eta P(t)\bigr)-\beta P(t)N^*(t). $$

Now we consider the following impulsive differential equation:

$$ \left \{ \begin{aligned} &\frac{dx(t)}{dt}= rx(t) \bigl(1-\eta x(t) \bigr)-\beta x(t)N^*(t),\quad t\neq \tau_n, \\ &x\bigl(hT_p^{+}\bigr)=\bigl(1-\omega (hT_p) d_1\bigr)x(hT_p),\quad t= \tau_n, \\ &\frac{d\omega(t)}{dt}=d_1\omega(t) \bigl(\omega(t)^{q_n}-1 \bigr), \\ &x\bigl(0^+\bigr)=P(0)\doteq P_0.\end{aligned} \right . $$
(33)

Again according to the comparison theorem on impulsive differential equations, we have P(t)≤x(t).

By using the formula of (24), we can easily have

$$\begin{aligned} q(nT) \doteq &1-\omega (nT)d_1 \\ =&1-\frac{d_1}{ (1+e^{nd_1} ((\omega((n-1)T))^{-\frac{n}{T}} -1 ) )^{\frac{T}{n}}} \end{aligned}$$

and

$$F(s,x)\doteq r-r\eta x(s)-\beta N^*(s). $$

Now we test and verify the conditions of Theorem 2.1. It is easy to see that condition (a) holds true naturally, and

$$F(s,0)=r-\beta N^*(s)\leq r $$

and

$$\frac{\partial F(s,x)}{\partial x}=-r\eta,\qquad \int^{\infty}_{0} \eta ds=\infty. $$

Therefore,

$$\begin{aligned} \exp \biggl(\int^{l+(h-1)T_N+\tau_k}_{l+(h-1)T_N+\tau_{k-1}}F(s,0)ds \biggr) =& \exp \biggl(\int^{l+(h-1)T_N+\tau_k}_{l+(h-1)T_N+\tau_{k-1}}r-\beta N^*(s)ds \biggr) \\ =&e^{r\tau}\exp \biggl(\int^{l+(h-1)T_N+\tau_k}_{l+(h-1)T_N+\tau_{k-1}} \bigl(-\beta N^*(s)\bigr)ds \biggr), \end{aligned}$$

where k≥1, τ k =, and

$$\begin{aligned} &\int^{l+(h-1)T_N+\tau_k}_{l+(h-1)T_N+\tau_{k-1}}\bigl(-\beta N^*(s)\bigr)ds \\ &\quad{}=-\beta\int^{l+(h-1)T_N+\tau_k}_{l+(h-1)T_N+\tau_{k-1}} \Biggl(N_0+\sum^{h-1}_{i=1} \delta_ie^{idT_N} \Biggr)e^{-ds}ds \\ &\quad{}=\frac{\beta}{d}e^{-d(l+(h-1)T_N+\tau_{k-1})} \Biggl(N_0+\sum ^{h-1}_{i=1}\delta_ie^{idT_N} \Biggr) \bigl(e^{-d\tau}-1 \bigr) \end{aligned}$$

for l∈(0,τ] and 1≤kk p , thus

$$\begin{aligned} &\hat{R}_{0}^{T_N}(h,k-1,T_N,l) \\ &\quad{}\doteq q \bigl((h-1)T_N+\tau_k \bigr)\exp \biggl(\int^{l+(h-1)T_N+\tau_k}_{l+(h-1)T_N+\tau_{k-1}}F(s,0)ds \biggr) \\ &\quad{}=R_0 \bigl((h-1)T_N+\tau_k \bigr) \\ &\qquad{}\times\exp{ \Biggl(\frac{\beta}{d}e^{-d(l+(h-1)T_N+\tau_{k-1})} \Biggl(N_0+\sum^{h-1}_{i=1} \delta_ie^{idT_N}\Biggr) \bigl(e^{-d\tau}-1\bigr) \Biggr)}, \end{aligned}$$

where R 0((h−1)T N +τ k ,τ)=(1−d 1 ω((h−1)T N +τ k ))exp().

Moreover,

$$\begin{aligned} &\int^{l+hT_N}_{l+(h-1)T_N+\tau_{k_p}}\bigl(-\beta N^*(s)\bigr)ds \\ &\quad{}=\int^{hT_N}_{l+(h-1)T_N+\tau_{k_p}}\bigl(-\beta N^*(s) \bigr)ds+\int^{l+hT_N}_{hT_N}\bigl(-\beta N^*(s)\bigr)ds \\ &\quad{}=-\beta\int^{hT_N}_{l+(h-1)T_N+\tau_{k_p}} \Biggl(N_0+\sum^{h-1}_{i=1} \delta_ie^{idT_N}\Biggr)e^{-ds}ds \\ &\qquad{}-\beta\int^{l+hT_N}_{hT_N} \Biggl(N_0+\sum^{h}_{i=1} \delta_ie^{idT_N}\Biggr)e^{-ds}ds \\ &\quad{}=\frac{\beta}{d}e^{-d(l+(h-1)T_N+\tau_{k_p})} \Biggl(N_0+\sum ^{h-1}_{i=1}\delta_ie^{idT_N} \Biggr) \bigl(e^{-d\tau}-1\bigr)+\frac{\beta\delta_{h}}{d}\bigl(e^{-dl}-1 \bigr) \\ &\quad{}\doteq \Psi(l), \end{aligned}$$

therefore,

$$\begin{aligned} \hat{R}_{0}^{T_N}(h,k_p,T_N,l) \doteq & q(hT)\exp{ \biggl(r-\int^{l+hT_N}_{l+(h-1)T_N+\tau_{k_p}}\beta N^*(s)ds \biggr)} \\ =& R_0(hT_N,\tau)e^{\Psi(l)}, \end{aligned}$$

for l∈(0,τ].

According to Theorem 2.1, we can see that if \(\hat{R}_{0}^{T_{N}}(h,k-1,T_{N},l)\leq 1\), then x(t)→0 as t→∞. Consequently, we have P(t)→0 as t→∞ provided \(\hat{R}_{0}^{T_{N}}(h,k-1,T_{N},l)\leq 1\), for k=1,2,…,k p +1.

Because \(\hat{R}_{0}^{T_{N}}(h,k-1,T_{N},l)\) increases with respect to k and l, for k=2,3,…,k p , we have

$$\begin{aligned} &\hat{R}_{0_{\max}}^{T_N}(h,k-1,T_N,l) \\ &\quad{}=\hat{R}_{0}^{T_N}(h,k_p-1,T_N, \tau) \\ &\quad{}=R_0\bigl((h-1)T_N+\tau_{k_p},\tau \bigr) \\ &\qquad{}\times\exp \Biggl(\frac{\beta}{d}e^{-d((h-1)T_N+\tau_{k_p})} \Biggl(N_0+ \sum^{h-1}_{i=1} \delta_ie^{idT_N}\Biggr) \bigl(e^{-d\tau}-1\bigr) \Biggr). \end{aligned}$$

From

$$\begin{aligned} \frac{\partial \hat{R}_{0}^{T_N}(h,k_p,T_N,l)}{\partial l} =&\hat{R}_{0}^{T_N}(h,k_p,T_N,l) \frac{\partial \Psi(l)}{\partial l} \\ =&\hat{R}_{0}^{T_N}(h,k_p,T_N,l) \beta e^{-dl}(D_{h-1}-\delta_{h}), \end{aligned}$$

we conclude that if D h−1>δ h then \(\hat{R}_{0}^{T_{N}}(h,k_{p},T_{N},l)\) increases with respect to l, so

$$\begin{aligned} &\hat{R}_{0_{\max}}^{T_N}(h,k_p,T_N,l) \\ &\quad{}=\hat{R}_{0}^{T_N}(h,k_p,T_N, \tau) \\ &\quad{}=R_0 (hT_N,\tau )\exp{ \Biggl( \frac{\beta}{d}e^{-dhT_N} \Biggl(N_0+\sum ^{h}_{i=1}\delta_ie^{idT_N} \Biggr) \bigl(e^{-d\tau}-1 \bigr) \Biggr)}. \end{aligned}$$

We can easily see that \(\hat{R}_{0_{\max}}^{T_{N}}(h,k-1,T_{N},l)<\hat{R}_{0_{\max}}^{T_{N}}(h,k_{p},T_{N},l)\). Therefore,

$$\begin{aligned} R^{T_N}_{0}(h,T_N) =&\max \bigl\{ \hat{R}_{0_{\max}}^{T_N}(h,k-1,T_N,l), \hat{R}_{0_{\max}}^{T_N}(h,k_p,T_N,l) \bigr\} \\ =&\max \bigl\{\hat{R}_{0}^{T_N}(h,k_p-1,T_N, \tau),\hat{R}_{0}^{T_N}(h,k_p,T_N, \tau) \bigr\} \\ =&R_0 (hT_N,\tau )\exp{ \Biggl( \frac{\beta}{d}e^{-dhT_N} \Biggl(N_0+\sum^{h}_{i=1} \delta_ie^{idT_N} \Biggr) \bigl(e^{-d\tau}-1 \bigr) \Biggr)}. \end{aligned}$$

If D h−1δ h then \(\hat{R}_{0}^{T_{N}}(h,k_{p},T_{N},l)\) decreases with respect to l, so

$$\begin{aligned} &\hat{R}_{0_{\max}}^{T_N}(h,k_p,T_N,l) \\ &\quad{}=\hat{R}_{0}^{T_N}(h,k_p,T_N,0) \\ &\quad{}=R_0 (hT_N,\tau )\exp{ \Biggl( \frac{\beta}{d}e^{-d((h-1)T_N+\tau_{k_p})}\Biggl(N_0+\sum ^{h-1}_{i=1}\delta_ie^{idT_N} \Biggr) \bigl(e^{-d\tau}-1\bigr) \Biggr)}. \end{aligned}$$

Due to the development of pest resistance, we can see that \(R_{0}(hT_{N},\tau)> R_{0}((h-1)T_{N}+\tau_{k_{p}},\tau)\), thus \(\hat{R}_{0}^{T_{N}}(h,k_{p},T_{N},0)>\hat{R}_{0}^{T_{N}}(h,k_{p}-1,T_{N},\tau)\), that is \(\hat{R}_{0_{\max}}^{T_{N}}(h,k-1,T_{N},l)<\hat{R}_{0_{\max}}^{T_{N}}(h,k_{p},T_{N},l)\), therefore,

$$\begin{aligned} &R^{T_N}_{0}(h,T_N)\\ &\quad{}=\max \bigl\{ \hat{R}_{0_{\max}}^{T_N}(h,k-1,T_N,l), \hat{R}_{0_{\max}}^{T_N}(h,k_p,T_N,l) \bigr\} \\ &\quad{}=\max \bigl\{\hat{R}_{0}^{T_N}(h,k_p-1,T_N, \tau),\hat{R}_{0}^{T_N}(h,k_p,T_N,0) \bigr\} \\ &\quad{}=R_0 (hT_N,\tau )\exp{ \Biggl( \frac{\beta}{d}e^{-d((h-1)T_N+\tau_{k_p})} \Biggl(N_0+\sum^{h-1}_{i=1} \delta_ie^{idT_N}\Biggr) \bigl(e^{-d\tau}-1\bigr) \Biggr)}. \end{aligned}$$

thus, if \(R^{T_{N}}_{0}(h,T_{N})<1\), then P(t)→0 as t→∞.

The following proof is the same as Theorem 3.1. The proof is complete.

Appendix F: Determining the New Number of Natural Enemies to Be Released for Case 4.2

In this case, solving equation

$$R^{T_N}_{0}(h,T_N)=R_C $$

with respect to δ i , yields

$$ \sum^{h}_{i=1} \delta_ie^{idT_N}=-\frac{de^{dhT_N}}{\beta(1-e^{-d\tau})} \ln \biggl( \frac{R_C}{R_0(hT_N,\tau)} \biggr)-N_0,\quad h\in \mathcal{N}. $$
(34)

Due to δ i =δ c , when ih′, we have

$$N_0+\delta_c\sum^{h^\prime}_{i=1}e^{idT_N}+ \sum^{h}_{i=h^\prime+1}\delta_ie^{idT_N} =-\frac{de^{dhT_N}}{\beta(1-e^{-d\tau})} \ln \biggl(\frac{R_C}{R_0(hT_N,\tau)} \biggr), $$

that is

$$\sum^{h}_{i=h^\prime+1}\delta_ie^{idT_N} =-\frac{de^{dhT_N}}{\beta(1-e^{-d\tau})} \ln \biggl(\frac{R_C}{R_0(hT_N,\tau)} \biggr) -N_0- \delta_c\sum^{h^\prime}_{i=1}e^{idT_N} \doteq A^\prime_{h}. $$

Therefore, when h=h′+1, we have

$$\delta_{h^\prime+1}=A^\prime_{h^\prime+1}e^{-d(h^\prime+1)T_N}. $$

When h=h′+2, we get

$$\delta_{h^\prime+1}e^{d(h^\prime+1)T_N}+\delta_{h^\prime+2}e^{d(h^\prime+2)T_N}=A^\prime_{h^\prime+2}, $$

that is

$$\delta_{h^\prime+2}=\bigl(A^\prime_{h^\prime+2}-A^\prime_{h^\prime+1} \bigr)e^{-d(h^\prime+2)T_N}. $$

Similarly, when h=h′+3, we get

$$\delta_{h^\prime+3}=\bigl(A^\prime_{h^\prime+3}-A^\prime_{h^\prime+2} \bigr)e^{-d(h^\prime+3)T_N}. $$

By induction, the new number of natural enemies to be released δ h can be determined as follows:

$$\delta_h=\left \{\begin{array}{l} \delta_c, \quad \mbox{if } h\leq h^{\prime},\\ A^\prime_{h^\prime+1}e^{-d(h^\prime+1)T_N},\quad h=h^{\prime}+1,\\ (A^\prime_{h}-A^\prime_{h-1})e^{-dhT_N},\quad \mbox{if } h>h^{\prime}+1. \end{array} \right . $$

Rights and permissions

Reprints and permissions

About this article

Cite this article

Liang, J., Tang, S., Cheke, R.A. et al. Adaptive Release of Natural Enemies in a Pest-Natural Enemy System with Pesticide Resistance. Bull Math Biol 75, 2167–2195 (2013). https://doi.org/10.1007/s11538-013-9886-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-013-9886-6

Keywords

Navigation