Skip to main content
Log in

Dynamic analysis of a biocontrol of sea lice by age-structured model

  • Original paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

Due to the responsibility for many illnesses on farmed salmon and inflicting huge financial losses, sea lice treatment has become one of the highest priorities in aquaculture studies. From the viewpoint of dynamics, we consider three stages in sea louse life cycle and study the predator–prey interaction between cleaner fish and sea lice. Through mathematical analysis, we provide two important indexes: the adult reproduction number \({\mathcal {R}}_s\) for sea lice and the net reproductive number of cleaner fish \({\mathcal {R}}_f\), which address the global dynamics relating to \({\mathcal {R}}_s\) and \({\mathcal {R}}_f\) theoretically, including the global/local stability of the equilibria, uniformly persistence and possible Hopf bifurcation. Numerical simulation is provided to show the oscillation behavior in the system.

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

Access this article

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

Instant access to the full article PDF.

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

Similar content being viewed by others

References

  1. Davenport, J.C., Black, K., Burnell, G., Cross, T., Culloty, S., Ekaratne, S., Furness, B., Mulcahy, M., Thetmeyer, H.: Aquaculture: The Ecological Issues. Wiley-Blackwell, New York (2003)

    Google Scholar 

  2. Treasurer, J.W.: A review of potential pathogens of sea lice and the application of cleaner fish in biological control. Pest Manag. Sci. 58, 546–558 (2002)

    Article  Google Scholar 

  3. Lawrence, E.: Henderson’s Dictionary of Biology. Pearson, London (2011)

    Google Scholar 

  4. Allen, E., Victory, H.: Modelling and simulation of a schistosomiasis infection with biological control. Acta Trop. 87, 251–267 (2003)

    Article  Google Scholar 

  5. Gourley, S., Lou, Y.: A mathematical model for the spatial spread and biocontrol of the asian longhorned beetle. SIAM J. Appl. Math. 74, 864–884 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  6. Lou, Y., Zhao, X.-Q.: Modelling malaria control by introduction of larvivorous fish. Bull. Math. Biol. 74, 2384–2407 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  7. Deady, S., Varian, S.J.A., Fives, J.M.: The use of cleaner-fish to control sea lice on two irish salmon (Salmo salar) farms with particular reference to wrasse behaviour in salmon cages. Aquaculture 131, 73–90 (1995)

    Article  Google Scholar 

  8. Skiftesvik, A., Bjelland, R., Durif, C.M., Johansen, I.S., Browman, H.I.: Delousing of atlantic salmon (Salmo salar) by cultured vs. wild ballan wrasse (Labrus bergylta). Aquaculture 402, 113–118 (2013)

    Article  Google Scholar 

  9. Liu, Q., Jiang, D., Hayat, T., Alsaedi, A.: Dynamics of a stochastic predator–prey model with stage structure for predator and Holling type II functional response. J. Nonlinear Sci. 28, 1151–1187 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  10. Al-Darabsah, I., Tang, X., Yuan, Y.: A prey–predator model with migrations and delays. Discrete Contin. Dyn. Syst. Ser. B 21, 737–761 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  11. Mohr, M., Barbarossa, M.V., Kuttler, C.: Predator–prey interactions, age structures and delay equations. Math. Model. Nat. Phenom. 9, 92–107 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  12. Song, Y., Xiao, W., Qi, X.: Stability and Hopf bifurcation of a predator–prey model with stage structure and time delay for the prey. Nonlinear Dyn. 83, 1409–1418 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  13. Dubey, B., Kumar, A.: Dynamics of prey–predator model with stage structure in prey including maturation and gestation delays. Nonlinear Dyn. 96, 2653–2679 (2019)

    Article  Google Scholar 

  14. Groner, M.L., Cox, R., Gettinby, G., Revie, C.W.: Use of agent-based modelling to predict benefits of cleaner fish in controlling sea lice, Lepeophtheirus salmonis, infestations on farmed atlantic salmon, Salmo salar L. J. Fish Dis. 36, 195–208 (2013)

    Article  Google Scholar 

  15. Aldrin, M., Huseby, R.B., Stien, A., Grøntvedt, R.N., Viljugrein, H., Jansen, P.A.: A stage-structured Bayesian hierarchical model for salmon lice populations at individual salmon farms-estimated from multiple farm data sets. Ecol. Model. 359, 333–348 (2017)

    Article  Google Scholar 

  16. Tian, Y., Al-Darabsah, I., Yuan, Y.: Global dynamics in sea lice model with stage structure. Nonlinear Anal. Real World Appl. 44, 283–304 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  17. Allee, W., Bowen, E.: Studies in animal aggregations: mass protection against colloidal silver among goldfishes. J. Exp. Zool. 61, 185–207 (1932)

    Article  Google Scholar 

  18. Kearn, G.C.: Leeches, Lice and Lampreys: A Natural History of Skin and Gill Parasites of Fishes. Springer, Berlin (2004)

    Google Scholar 

  19. Armstrong, R.: Stable model structures for representing biogeochemical diversity and size spectra in plankton communities. J. Plankton Res. 21, 445–464 (1999)

    Article  Google Scholar 

  20. Oaten, A., Murdoch, W.W.: Functional response and stability in predator–prey systems. Am. Nat. 109, 289–298 (1975)

    Article  Google Scholar 

  21. Cushing, J.M.: An Introduction to Structured Population Dynamics. SIAM, Philadelphia (1998)

    Book  MATH  Google Scholar 

  22. Cushing, J.M., Saleem, M.: A predator prey model with age structure. J. Math. Biol. 14, 231–250 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  23. Allen, L.J.: An Introduction to Mathematical Biology. Pearson, London (2006)

    Google Scholar 

  24. Smorynski, C.: History of Mathematics. Springer, New York (2008)

    MATH  Google Scholar 

  25. James, R.: Mathematics Dictionary. Springer, Berlin (1992)

    Book  MATH  Google Scholar 

  26. Beretta, E., Kuang, Y.: Geometric stability switch criteria in delay differential systems with delay dependent parameters. SIAM J. Math. Anal. 33, 1144–1165 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  27. Baker, J., Muller, R., Rollinson, D. (eds.): Advances in Parasitology, vol. 44. Academic Press, London (1999)

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

    Article  Google Scholar 

  29. Begon, M., Townsend, C.R., Harper, J.L.: Ecology: From Individuals to Ecosystems. Wiley-Blackwell, New York (2005)

    Google Scholar 

  30. Skiftesvik, A.B., Blom, G., Agnalt, A.L., Durif, C.M., Browman, H.I., Bjelland, R.M., Harkestad, L.S., Farestveit, E., Paulsen, O.I., Fauske, M., Havelin, T., Johnsen, K., Mortensen, S.: Wrasse (labridae) as cleaner fish in salmonid aquaculture the hardangerfjord as a case study. Mar. Biol. Res. 10, 289–300 (2014)

    Article  Google Scholar 

  31. Boxshall, G.A., Defays, D.: Pathogens of Wild and Farmed Fish: Sea Lice. CRC Press, Boca Raton (1993)

    Google Scholar 

  32. Sayer, M.D., Treasurer, J.W., Costello, M.J.: Wrasse: Biology and Use in Aquaculture. Wiley-Blackwell, New York (1996)

    Google Scholar 

  33. Hale, J., Lunel, S.V.: Introduction to Functional Differential Equations. Springer, New York (1993)

    Book  MATH  Google Scholar 

  34. Smith, H.L.: Monotone Dynamical Systems: An Introduction to the Theory of Competitive and Cooperative Systems. American Mathematical Society, Providence (1995)

    MATH  Google Scholar 

  35. Zhao, X.-Q., Jing, Z.: Global asymptotic behavior in some cooperative systems of functional differential equations. Can. Appl. Math. Q. 4, 421–444 (1996)

    MathSciNet  MATH  Google Scholar 

  36. Zhao, X.-Q.: Dynamical Systems in Population Biology. Springer, New York (2017)

    Book  MATH  Google Scholar 

  37. Hirsch, M.W., Smith, H.L., Zhao, X.-Q.: Chain transitivity, attractivity, and strong repellors for semidynamical systems. J. Dyn. Differ. Equ. 13, 107–131 (2001)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the editor and anonymous referee for their comments that greatly improved our original manuscript.

Funding

This study was funded in part by the NSERC of Canada (203786463102000) and National NSF of China (11571257).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yuan Yuan.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

Appendices

Appendix A

The proof of Theorem 1.

Proof

The set \(D_X\) is closed in X. Given \(\phi \in D_X\), we define

$$\begin{aligned} G(\phi )=(G_1(\phi ),G_2(\phi ),G_3(\phi )), \end{aligned}$$

with

$$\begin{aligned} G_1(\phi )=&\frac{{b\left( {\phi _2(0)} \right) }}{{P\left( { \phi _2(0)} \right) }}{\mathrm{e}^{ - {\mu _N}{\hat{\tau }}({ {\phi _2}})}}\\&- \left( {{\mu _{C}} + \beta H} \right) \frac{{ \phi _1(-m)}}{{P\left( { \phi _2(-m)} \right) }},\\ G_2(\phi )=&\frac{1}{{P\left( { \phi _2(-m)} \right) }}\left[ \beta H \phi _1(-m) \right. \\&\left. \!\!- {\mu _{A}} \phi _2(-m)\! -\phi _3(-m)\phi _2(-m)f(\phi _2(-m)) \right] ,\\ G_3(\phi )=&\frac{1}{{P\left( { \phi _2(-m)} \right) }}\left[ \gamma \phi _3(-m)\phi _2(-m)f(\phi _2(-m))\right. \\&\left. -{\mu _W}\phi _3(-m)\right] , \end{aligned}$$

where

$$\begin{aligned} \hat{\tau }({ {\phi _2}}) = \int \limits _{ - m}^0 {\frac{1}{{P\left( {\phi _2 (m+\theta )} \right) }}} \mathrm{d}\theta . \end{aligned}$$

Obviously, \(G(\phi )\) is Lipschitz continuous in \(\phi \) in any compact set in \(D_X\). Thus, there is a unique solution in (4) on its maximal existence interval \([0,\sigma _{\phi })\) through \(\phi \) for any \(\phi \in D_X\) [33, Theorem 2.2.3].

Furthermore, it is easy to check that \(G_i(\phi ) \ge 0\) or any \(\phi \in D_X\) with \(\phi _i(m)=0, i=1,2,3\). Hence, by [34, Theorem 5.2.1], \(\phi _i(T)\ge 0\) for all \(t\in [0,\sigma _{\phi }), i=1,2,3\). Therefore, any solution of (4) is nonnegative for any \(t \in [0,\sigma _{\phi })\). In particular, \({\mathcal {C}}(T)>0\) when \({\mathcal {C}}(m)>0\) because

$$\begin{aligned} {\mathcal {C}}(T) \ge {\mathcal {C}}(m)\exp \left\{ { - \int \limits _m^T {\frac{{{\mu _C} + \beta H}}{{P\left( { {\mathcal {A}}(\theta )} \right) }}\mathrm{d}\theta } } \right\} . \end{aligned}$$

for any when \(\phi \in D_X\). Similarly, the positivity of \({\mathcal {A}}(T)\) and \({\mathcal {W}}(T)\) is followed when \({\mathcal {A}}(m)>0\) and \({\mathcal {W}}(m)>0\).

To show the boundedness of the system, let \({\tilde{\mu }}=\min \left\{ \mu _{C},\mu _{A},\mu _{W}\right\} \). Since \(b(t)\le B(t)\) is a nondecreasing function and both \({\mathcal {C}}\) and \({\mathcal {W}}\) are nonnegative, it follows from (4) that

$$\begin{aligned}&\frac{{\mathrm{d}({{{\mathcal {C}}}} + {{{\mathcal {A}}}} + {{{\mathcal {W}}}})}}{{\mathrm{d}T}} \le \frac{1}{{{P_0}}}b\left( {{{{\mathcal {A}}}}(T - m) } \right) \\&\quad - \frac{{{{\tilde{\mu }}} }}{{{P_\infty }}}\left( {{{{\mathcal {C}}}}(T) + {{{\mathcal {A}}}}(T) + {{{\mathcal {W}}}}(T)} \right) , \\&\quad \le \frac{1}{{{P_0}}}B\left( {{{{\mathcal {C}}}}(T - m) + {{{\mathcal {A}}}}(T - m) + {{{\mathcal {W}}}}(T - m)} \right) \\&\quad - \frac{{{{\tilde{\mu }}} }}{{{P_\infty }}}\left( {{{{\mathcal {C}}}}(T) + {{{\mathcal {A}}}}(T) + {{{\mathcal {W}}}}(T)} \right) . \end{aligned}$$

Therefore, \({\mathcal {C}}+{\mathcal {A}}+{\mathcal {W}}\) is bounded ( [35, Theorem 3.2]). Thus, all solution of (4) are bounded because \({\mathcal {C}}, {\mathcal {A}}\) and \({\mathcal {W}}\) are nonnegative. Hence, by [33, Theorem 2.3.1], we have \(\sigma _{\phi }=+\infty \). Thus, all the solutions are ultimately bounded and exist globally. Thus, there exists \(T_1>0\) and \(K>0\) such that \(0 \le \{{\mathcal {C}}(T),{\mathcal {A}}(T),{\mathcal {W}}(T)\} \le {K}\) for \(T>T_1\). \(\square \)

Appendix B

The proof of Theorem 2.

Proof

From (4a) and (4b), we have

$$\begin{aligned} \frac{{\mathrm{d} {\mathcal {C}}}}{{\mathrm{d}T}}&= \frac{{b\left( {{\mathcal {A}}(T - m)} \right) }}{{P\left( { {\mathcal {A}}(T - m)} \right) }}{\mathrm{e}^{ - {\mu _N}{\hat{\tau }}({ {\mathcal {A}}_T})}} \nonumber \\&- \left( {{\mu _{C}} + \beta H} \right) \frac{{ {\mathcal {C}}(T)}}{{P\left( { {\mathcal {A}}(T)} \right) }},\nonumber \\ \frac{{\mathrm{d} {\mathcal {A}}}}{{\mathrm{d}T}}&\le \frac{1}{{P\left( { {\mathcal {A}}(T)} \right) }}\left[ {\beta H {\mathcal {C}}(T) - {\mu _{A}} {\mathcal {A}}(T)} \right] . \end{aligned}$$
(26)

When \({\mathcal {R}}_s<1\), we know that (0, 0) is a globally attractive equilibrium point in the system obtained from (26) by replacing \(\le \) with \(=\), see [16, Theorem 4.1]. By the comparison principle (see e.g., [34, Theorem 5.1.1]) and the nonnegativity of \({{{\mathcal {C}}}}(T)\) and \({{{\mathcal {A}}}}(T)\), we obtain

$$\begin{aligned} \mathop {\lim }\limits _{T \rightarrow \infty } {{{\mathcal {C}}}}(T) = 0 \qquad \text {and}\qquad \mathop {\lim }\limits _{T \rightarrow \infty } {{{\mathcal {A}}}}(T) =0. \end{aligned}$$

Denote \(\Phi (T)(\phi _{_{\mathcal {CW}}}):= {\hat{z}}_T(\phi _{_{\mathcal {CW}}})\) be the solution semiflow associated with (4). Let \(\omega =\omega (\phi _{_{\mathcal {CW}}})\) be the omega limit set of \(\Phi (T)\). By [36, Lemma 1.2.1], \(\omega \) is an internally chain transitive set for \(\Phi (T)\). Thus, \(\omega =\{(0,0)\}\times {\bar{\omega }}\) for some \({\bar{\omega }}\subset {\mathbb {R}}\). It is easy to see that

$$\begin{aligned} \Phi (0,0,{\mathcal {W}}_m)=(0,0,\hat{\Phi }({\mathcal {W}}_m)) \end{aligned}$$

where \(\hat{\Phi }(T)\) is the solution semiflow associated with the equation

$$\begin{aligned} \frac{{\mathrm{d}{\mathcal {W}}}}{{\mathrm{d}T}} =\frac{-{\mu _W}}{P_0}{\mathcal {W}}(T). \end{aligned}$$
(27)

Since \(\omega \) is an internally chain transitive set for \(\Phi (T)\), it easily follows that \({\bar{\omega }}\) is an internally chain transitive set for \(\hat{\Phi }\).

Obviously, \(\{0\}\) is globally asymptotically stable for (27), and we get \({{{\bar{\omega }}} } \cap W^s\left( 0 \right) \ne \emptyset \), where \(W^s\left( 0 \right) \) is the stable manifold of 0. Thus, \({\bar{\omega }}=\{0\}\) (see e.g., [37, Theorem 3.2 and Remark 4.6] or [36, Theorem 1.2.1]). Therefore, we have \(\omega =\{(0,0,0)\}\), and hence

$$\begin{aligned} \mathop {\lim }\limits _{T \rightarrow \infty } \left( {{{{\mathcal {C}}}}(T),{{{\mathcal {A}}}}(T),{{{\mathcal {W}}}}(T)} \right) = \left( {0,0,0} \right) . \end{aligned}$$

Together with the local stability of (0, 0, 0) (similar proof as in [16] [Theorem 3.1], we obtain the global asymptotic stability of (0, 0, 0) when \({\mathcal {R}}_s<1\) and the instability as \({\mathcal {R}}_s>1\). \(\square \)

Appendix C

The proof of Theorem 3.

Proof

From [16, Theorem 4.3], when \({\mathcal {R}}_s>1\), \(({\mathcal {C}}_1^*,{\mathcal {A}}_1^*)\) is a globally attractive equilibrium point in the system

$$\begin{aligned} \frac{{\mathrm{d} {\mathcal {C}}}}{{\mathrm{d}T}}&= \frac{{b\left( {{\mathcal {A}}(T - m)} \right) }}{{P\left( { {\mathcal {A}}(T - m)} \right) }}{\mathrm{e}^{ - {\mu _N}{\hat{\tau }}({ {\mathcal {A}}_T})}} \nonumber \\&- \left( {{\mu _{C}} + \beta H} \right) \frac{{ {\mathcal {C}}(T)}}{{P\left( { {\mathcal {A}}(T)} \right) }}, \end{aligned}$$
(28a)
$$\begin{aligned} \frac{{\mathrm{d} {\mathcal {A}}}}{{\mathrm{d}T}}&= \frac{1}{{P\left( { {\mathcal {A}}(T)} \right) }}\left[ {\beta H {\mathcal {C}}(T) - {\mu _{A}} {\mathcal {A}}(T)} \right] . \end{aligned}$$
(28b)

By the comparison principle (see e.g., [34, Theorem 5.1.1]), we obtain

$$\begin{aligned} \mathop {\lim \sup }\limits _{T \rightarrow \infty } {{{\mathcal {C}}}}(T) \le {{{\mathcal {C}}}}_1^*\qquad \text {and}\qquad \mathop {\lim \sup }\limits _{T \rightarrow \infty } {{{\mathcal {A}}}}(T) \le {{{\mathcal {A}}}}_1^*. \end{aligned}$$

Next, we prove \({\mathcal {W}}(T)\rightarrow 0\) as \(T\rightarrow \infty \) when \({\mathcal {R}}_f<1\). Let \({T}_2\) be sufficiently large, from (4c), we get

$$\begin{aligned} \frac{{d{{{\mathcal {W}}}}}}{{\mathrm{d}T}}\le & {} \frac{{\gamma {{{\mathcal {A}}}}_1^*f({{{\mathcal {A}}}}_1^*) - {\mu _W}}}{{P(A)}}{{{\mathcal {W}}}}(T)\\\le & {} \frac{{\gamma {{{\mathcal {A}}}}_1^*f({{{\mathcal {A}}}}_1^*) - {\mu _W}}}{{P_0}}{{{\mathcal {W}}}}(T) \\= & {} - \frac{1}{{{\mu _A}P_0}}\left( {1 - {{{{\mathcal {R}}}}_f}} \right) {{{\mathcal {W}}}}(T) \end{aligned}$$

for \(T\ge {T}_2\). Therefore,

$$\begin{aligned} \mathop {\lim \sup }\limits _{T \rightarrow \infty } {{{\mathcal {W}}}}(T) \le 0. \end{aligned}$$

Hence,

$$\begin{aligned} \mathop {\lim }\limits _{T \rightarrow \infty } {{{\mathcal {W}}}}(T) = 0 \end{aligned}$$

due to the nonnegativity of \({\mathcal {W}}\).

Hence, the limiting system of system (4) is same as (28). Thus, the omega limit set of the solution semiflow \(\Phi (T)\) of system (4) is \(\omega =\bar{{\bar{\omega }}} \times 0\) for some \( \bar{{\bar{\omega }}} \subset {\mathbb {R}}\times {\mathbb {R}}\).

Claim 1

\(\bar{{{\bar{\omega }}} }\not \subset \left\{ {\left( {0,0} \right) } \right\} \cup \left\{ {\left( {0,{\bar{y}}} \right) :{\bar{y}} \ne 0} \right\} \cup \left\{ {\left( {{\bar{x}},0} \right) :}\right. \left. {{\bar{x}} \ne 0} \right\} \).

By contrary, first, assume \(\bar{{{\bar{\omega }}} }\subset \left\{ {\left( {0,{\bar{y}}} \right) :{\bar{y}} \ne 0} \right\} \). Then,

$$\begin{aligned} \mathop {\lim }\limits _{T \rightarrow \infty }{\mathcal {C}}(T) = 0. \end{aligned}$$

When \({\mathcal {R}}_s>1\), the solutions of (28) satisfy \(\liminf \limits _{T\rightarrow \infty } ({\mathcal {C}}(T),{\mathcal {A}}(T))\ge (\eta _1,\eta _1)\) for some positive number \(\eta _1>0\) [16, Theorem 4.2]. Thus,

$$\begin{aligned} ~&\mathop {\lim }\limits _{T \rightarrow \infty } \left[ \frac{{b\left( {{{{\mathcal {A}}}}(T - m)} \right) }}{{P\left( {{{{\mathcal {A}}}}(T - m)} \right) }}{\mathrm{e}^{ - {\mu _N}{\hat{\tau }} ({{{{\mathcal {A}}}}_T})}} \right. \\&\left. - \left( {{\mu _C} + \beta H} \right) \frac{{{{{\mathcal {C}}}}(T)}}{{P\left( {{{{\mathcal {A}}}}(T)} \right) }}\right] \\ =&\quad \mathop {\lim }\limits _{T \rightarrow \infty } \frac{{b\left( {{{{\mathcal {A}}}}(T - m)} \right) }}{{P\left( {{{{\mathcal {A}}}}(T - m)} \right) }}{\mathrm{e}^{ - {\mu _N}{\hat{\tau }} ({{{{\mathcal {A}}}}_T})}}\\&\quad> \frac{1}{2}\frac{{b\left( {{\eta _1}} \right) }}{{{P_\infty }}}\exp \left\{ {\frac{{ - m{\mu _N}}}{{P\left( {{\eta _1}} \right) }}} \right\} > 0. \end{aligned}$$

Let \({T}_3>0\) be sufficiently large. Then, it follows from (28a) that

$$\begin{aligned} \frac{{d{{{\mathcal {C}}}}}}{{\mathrm{d}T}}> \frac{1}{2}\frac{{b\left( {{\eta _1}} \right) }}{{{P_\infty }}}\exp \left\{ {\frac{{ - m{\mu _N}}}{{P\left( {{\eta _1}} \right) }}} \right\} > 0,\qquad \forall T\ge {T}_3. \end{aligned}$$

Therefore, \({\mathcal {C}}(T)\rightarrow \infty \) as \(T\rightarrow \infty \), a contradiction.

Now, assume that \(\bar{{{\bar{\omega }}} }\subset \left\{ {\left( {{\bar{x}},0} \right) :{\bar{x}} \ne 0} \right\} \). Then,

$$\begin{aligned} \mathop {\lim }\limits _{T \rightarrow \infty }{\mathcal {A}}(T) = 0. \end{aligned}$$

Let \({T}_4>0\) be sufficiently large. Then, it follows from (28b) that

$$\begin{aligned} \frac{{d{{{\mathcal {A}}}}}}{{\mathrm{d}T}}> \frac{1}{2}\frac{\beta H\eta _1}{P_{\infty }} > 0,\qquad \forall T\ge {T}_4. \end{aligned}$$

Therefore, \({\mathcal {A}}(T)\rightarrow \infty \) as \(T\rightarrow \infty \), which contradicts the boundedness of solutions. This proves the claim.

For any \(\phi _{_{\mathcal {CW}}}\in {D_{Y}}\), we have

$$\begin{aligned} \Phi ({\mathcal {C}}_m,\hat{\phi },0)=( {{\tilde{\Phi }}}({\mathcal {C}}_m,\hat{\phi }),0) \end{aligned}$$

where \({\tilde{\Phi }}(T)\) is the solution semiflow associated with system (28). By [36, Lemma 1.2.1], \(\omega \) is an internally chain transitive set, and hence, \(\bar{{{\bar{\omega }}} }\) is an internally chain transitive set for \({\tilde{\Phi }}\). Since \(({\mathcal {C}}_1^*,{\mathcal {A}}_1^*)\) is a globally attractive in (28) and from Claim 1, we get \(\bar{{{\bar{\omega }}} } \cap W^s\left( {\mathcal {C}}_1^*,{\mathcal {A}}_1^* \right) \ne \emptyset \), where \(W^s\left( {\mathcal {C}}_1^*,{\mathcal {A}}_1^* \right) \) is the stable manifold of \(\left( {\mathcal {C}}_1^*,{\mathcal {A}}_1^* \right) \). Then, it follows from [36, Theorem 1.2.1] (or [37, Theorem 3.2 and Remark 4.6]) that \(\bar{{{\bar{\omega }}} }=\left( {\mathcal {C}}_1^*,{\mathcal {A}}_1^* \right) \). This proves that \(\omega =\left( {{{\mathcal {C}}}}_1^*, {{{{\mathcal {A}}}}_1^*,0} \right) \). Consequently,

$$\begin{aligned} \lim \limits _{T\rightarrow \infty } \left( {\mathcal {C}}(T),{\mathcal {A}}(T),{\mathcal {W}}(T)\right) = \left( {\mathcal {C}}_1^*,{\mathcal {A}}_1^*,0 \right) \end{aligned}$$

for any \(\phi _{_{\mathcal {CW}}}\in {D_{Y}}\) with \({\mathcal {C}}_m>0, \hat{\phi }(m)>0\), that is, \(\left( {\mathcal {C}}_1^*,{\mathcal {A}}_1^*,0 \right) \) is globally attractive. This completes the proof. \(\square \)

Appendix D

The proof of Theorem 4.

Proof

Let

$$\begin{aligned} Y_{0}&=\{\phi _{_{\mathcal {CW}}}\in {D_{Y}}\ : \ {\mathcal {C}}_m>0,\ \hat{\phi }(m)>0\ \text {and}\ {\mathcal {W}}_m>0\},\\ Y_{1}&=\{\phi _{_{\mathcal {CW}}}\in {D_{Y}}\ : \ {\mathcal {C}}_m=0\},\\ Y_{2}&=\{\phi _{_{\mathcal {CW}}}\in {D_{Y}}\ : \ \hat{\phi }(m)=0\},\\ Y_{3}&=\{\phi _{_{\mathcal {CW}}}\in {D_{Y}}\ : \ {\mathcal {W}}_m=0\}, \end{aligned}$$

and

$$\begin{aligned} M_{\partial }=\{\phi _{_{\mathcal {CW}}}\in {D_{Y}}\ :\ {\hat{z}}_T(\phi _{_{\mathcal {CW}}})\in \partial Y_{0},T\ge m\} \end{aligned}$$

where \(\partial Y_{0}=Y \setminus Y_{0}=Y_{1}\cup Y_{2}\cup Y_{3}\).

For \(i=1,2\), fix a small \(\epsilon _i>0\). Since \(\mathop {\lim }\limits _{{{{\mathcal {A}}}} \rightarrow 0} \frac{{b\left( {{{\mathcal {A}}}} \right) }}{{{{{\mathcal {A}}}}P({\mathcal {A}})}} = \frac{{b'(0)}}{{{P_0}}}\) and \(\mathop {\lim }\limits _{{{{\mathcal {A}}}} \rightarrow 0} {{{\mathcal {A}}}}f({{{\mathcal {A}}}}) = 0\), in a neighborhood of \({\mathcal {A}}=0\), we have

$$\begin{aligned} \left| {\frac{{b\left( {{{\mathcal {A}}}} \right) }}{{{{\mathcal {A}}}}P({\mathcal {A}})} - \frac{b'(0)}{P_0}} \right|<\epsilon _1\ \text {and}\ \left| {{{{\mathcal {A}}}}f({{{\mathcal {A}}}})} \right| <\epsilon _2, \end{aligned}$$
(29)

respectively. Set \(\epsilon _3=\max \{\epsilon _1,\epsilon _2\}\).

Claim 2

There exists a \(\delta _1(\epsilon _3)>0\), such that for any \(\phi _{_{\mathcal {CW}}}\in Y_{0}\),

$$\begin{aligned} \limsup \limits _{T\rightarrow \infty } \Vert {\hat{z}}_{T}(\phi _{_{\mathcal {CW}}})-E_0\Vert \ge \delta _1(\epsilon _3). \end{aligned}$$

By contradiction, suppose that \(\limsup \limits _{T\rightarrow \infty } \Vert {\hat{z}}_{T}(\psi _{_{\mathcal {CW}}})-E_0\Vert < \delta _1(\epsilon _3)\) for some \(\psi _{_{\mathcal {CW}}}\in Y_0\). Thus, there exists \(T_5 > m\) such that \( \left| {\mathcal {A}}(T)\right| <\delta _1(\epsilon _3)\) for \(T>T_5+m\). Hence, (29) is satisfied.

From (4), we obtain

$$\begin{aligned} \frac{{\mathrm{d} {\mathcal {C}}}}{{\mathrm{d}T}}> & {} \left( \frac{{b'(0)}}{{{P_0}}} - {\epsilon _3}\right) \mathrm{e}^{\frac{{ - m{\mu _N}}}{{{P_0}}}} A(T - m)\nonumber \\&- \frac{{{\mu _C} + \beta H}}{{{P_0}}}C(T),\nonumber \\ \frac{{\mathrm{d} {\mathcal {A}}}}{{\mathrm{d}T}}> & {} \frac{{\beta H}}{{{P(\delta _1(\epsilon _3))}}}{{{\mathcal {C}}}}(T) - \frac{{{\mu _A}}}{{{P_0 }}}{{{\mathcal {A}}}}(T)-\frac{\epsilon _3}{P_0}{\mathcal {W}}(T),\nonumber \\ \frac{{\mathrm{d} {\mathcal {W}}}}{{\mathrm{d}T}}> & {} -\frac{\mu _W}{P_0}{\mathcal {W}}(T). \end{aligned}$$
(30)

The characteristic equation of the system obtained from (30) by replacing > with \(=\) is

$$\begin{aligned} (\lambda +\frac{\mu _W}{P_0}){\Delta _2}(\lambda )=0 \end{aligned}$$

where

$$\begin{aligned} {\Delta _2}(\lambda )&={\lambda ^2} + \left( {\frac{{{\mu _A} + {\mu _C} + \beta H}}{{{P_0}}}} \right) \lambda + \frac{{{\mu _A}\left( {{\mu _C} + \beta H} \right) }}{{P_0^2}}\nonumber \\&\quad -\, \left( {\frac{{b'(0)}}{{{P_0}}} - {\epsilon _3}} \right) \frac{{{\mathrm{e}^{\frac{{ - {\mu _N}m}}{{{P_0}}}}}\beta H}}{{P\left( {{\delta _1(\epsilon _3)}} \right) }}{\mathrm{e}^{ - \lambda m}} = 0. \end{aligned}$$
(31)

Obviously, it is enough to study the roots in \({\Delta _2}(\lambda )=0\). Let \(\lambda _2(\epsilon _3)\) be the principal eigenvalue. When \(\epsilon _3=0\), the obtained system is irreducible and quasimonotone; hence, it is enough to study only the real roots in (31) (see [34, Theorem 5.5.1]). It is clear that \(\Delta _2(0)<0\) when \({\mathcal {R}}_s>1\). Therefore, due to the continuity of \({\Delta _2}(\lambda )\) and \(\mathop {\lim }\limits _{\lambda \rightarrow \infty } {\Delta _2}(\lambda ) = + \infty \), there exists \({\bar{\lambda }}>0\) such that \({\Delta _2}({\bar{\lambda }} )=0\). Thus, \(\lambda _2(0)>0\), and hence, \(\lambda _2(\epsilon _3)>0\) for sufficiently small \(\epsilon _3>0\) because of the continuity of \(\lambda _2\). Thus, there exists a solution \(\mathbf {V_1}(T)=\mathrm{e}^{\lambda _2(\epsilon _3)T}\hat{\xi }\), where \(\hat{\xi }\) is the positive eigenfunction corresponding to \(\lambda _2(\epsilon _3), \mathbf {V_1}\) and \(\hat{\xi }\) are vectors with three components. Since \({\mathcal {C}}(T), {\mathcal {A}}(T)\) and \({\mathcal {W}}(T)\) are nonnegative for all \(T > m\), the comparison arguments imply that there exists a small \(\ell _1 > 0\) such that \(({\mathcal {C}}(T),{\mathcal {A}}(T),{\mathcal {W}}(T))\ge \ell _1 \mathrm{e}^{\lambda _2(\epsilon _3)T}\hat{\xi }\) for all \(T\ge T_5+m\) [34, Theorem 5.1.1]. Since \(\lambda _2(\epsilon _3)>0\), we have \(\mathop {\lim }\limits _{T \rightarrow \infty } ({\mathcal {C}}(T),{\mathcal {A}}(T),{\mathcal {W}}(T)) = \infty \), which is a contradiction because all solutions in system (4) are bounded. This proves the claim.

Since \(\mathop {\lim }\limits _{{{{\mathcal {A}}}} \rightarrow {\mathcal {A}}_1^*} {{{\mathcal {A}}}}f({{{\mathcal {A}}}}) = {\mathcal {A}}_1^*f({\mathcal {A}}_1^*)\), in a neighborhood of \({\mathcal {A}}={\mathcal {A}}_1^*\), there exists a small \(\epsilon _4>0\) such that

$$\begin{aligned} {{{\mathcal {A}}}}_1^*f({{{\mathcal {A}}}}_1^*) - {\epsilon _4}< {{{\mathcal {A}}}}f({{{\mathcal {A}}}}) < {{{\mathcal {A}}}}_1^*f({{{\mathcal {A}}}}_1^*) + {\epsilon _4}. \end{aligned}$$

Claim 3

There exists a \(\delta _2(\epsilon _4)=\delta _2>0\), such that for any \(\phi _{_{\mathcal {CW}}}\in Y_{0}\),

$$\begin{aligned} \limsup \limits _{T\rightarrow \infty } \Vert {\hat{z}}_{T}(\phi _{_{\mathcal {CW}}})-E_1\Vert \ge \delta _2(\epsilon _4). \end{aligned}$$

By contradiction, suppose that \(\limsup \limits _{T\rightarrow \infty } \Vert {\hat{z}}_{T}({\bar{\psi }}_{_{\mathcal {CW}}})-E_1\Vert < \delta _2(\epsilon _4)\) for some \({\bar{\psi }}_{_{\mathcal {CW}}}\in Y_0\). Hence, there exists \(T_6 > m\) such that \( \left| {\mathcal {A}}(T)-{\mathcal {A}}_1^*\right| <\delta _2(\epsilon _4)\) for \(T>T_6+m\). From (4), we have

$$\begin{aligned} \frac{{d{{{\mathcal {C}}}}}}{{\mathrm{d}T}}> & {} \frac{{ - \left( {{\mu _C} + \beta H} \right) }}{{{P({\mathcal {A}}_1^*-\delta _2(\epsilon _4))}}}{{{\mathcal {C}}}} (T),\nonumber \\ \frac{{d{{{\mathcal {A}}}}}}{{\mathrm{d}T}}> & {} \frac{{\beta H}}{{{P({\mathcal {A}}_1^*+\delta _2(\epsilon _4))}}}{{{\mathcal {C}}}}(T)\nonumber \\&-\frac{{ {\mu _A}}}{{{P({\mathcal {A}}_1^*-\delta _2(\epsilon _4))}}}{{{\mathcal {A}}}}(T)\nonumber \\&- \frac{{{{{\mathcal {A}}}}_1^*f({{{\mathcal {A}}}}_1^*) +\epsilon _4 }}{{{P({\mathcal {A}}_1^*-\delta _2(\epsilon _4))}}}{{{\mathcal {W}}}}(T),\\ \frac{{d{{{\mathcal {W}}}}}}{{\mathrm{d}T}}> & {} \left( \gamma {\frac{{ {{{\mathcal {A}}}}_1^*f({{{\mathcal {A}}}}_1^*) -\epsilon _4 }}{{P({\mathcal {A}}_1^*+\delta _2(\epsilon _4))}} - \frac{{{\mu _W}}}{{{P({\mathcal {A}}_1^*-\delta _2(\epsilon _4))}}}} \right) {{{\mathcal {W}}}}(T).\nonumber \end{aligned}$$
(32)

The characteristic equation of the system obtained from (32) by replacing > with \(=\) is

$$\begin{aligned} {\Delta _3}(\lambda )= & {} \left( {\lambda + \frac{{{\mu _C} + \beta H}}{{{P({\mathcal {A}}_1^*-\delta _2(\epsilon _4))}}}} \right) \\&\left( {\lambda + \frac{{{\mu _A}}}{{{P({\mathcal {A}}_1^*-\delta _2(\epsilon _4))}}}} \right) \\&\left( \lambda -\gamma \frac{{ {{{\mathcal {A}}}}_1^*f({{{\mathcal {A}}}}_1^*) - \epsilon _4}}{{{P({\mathcal {A}}_1^*+\delta _2(\epsilon _4)) }}}\right. \\&\left. + \frac{{{\mu _W}}}{{{P({\mathcal {A}}_1^*-\delta _2(\epsilon _4))}}} \right) =0. \end{aligned}$$

When \(\epsilon _4=0\), \(\lambda =\frac{{\mathcal {R}}_f-1}{\mu _WP({\mathcal {A}}^*_1)}\) is a positive root in \({\Delta _3}(\lambda )=0\) if \({\mathcal {R}}_f >1\). Hence, the principal eigenvalue is positive. By similar arguments as those in the proof of Claim 2, we have \(({\mathcal {C}}(T),{\mathcal {A}}(T),{\mathcal {W}}(T))\rightarrow \infty \) as \(T\rightarrow \infty \), resulting a contradiction.

Let \(\hat{\omega }(\phi _{_{\mathcal {CW}}})\) be the omega limit set of the orbit \({\hat{z}}_{T}(\phi _{_{\mathcal {CW}}})\) through \(v\in {D_{Y}}\).

Claim 4

\(\bigcup \left\{ \hat{\omega }(\phi _{_{\mathcal {CW}}}):\phi _{_{\mathcal {CW}}} \in {M_\partial } \right\} = E_0\cup E_1\).

For any \(\phi _{_{\mathcal {CW}}}\in M_\partial \), we have \({\hat{z}}_T(\phi _{_{\mathcal {CW}}})\in Y_1\) or \({\hat{z}}_T(\phi _{_{\mathcal {CW}}})\in Y_2\) or \({\hat{z}}_T(\phi _{_{\mathcal {CW}}})\in Y_3\). If \({\hat{z}}_T(\phi _{_{\mathcal {CW}}})\in Y_1\), i.e., \({\mathcal {C}}(T)\equiv 0\), then from (4b) we have

$$\begin{aligned} \frac{{d{{{\mathcal {A}}}}}}{{\mathrm{d}T}} \le - {{{\mathcal {A}}}}(T)\left( {\frac{{{\mu _A} + {{{\mathcal {W}}}}(T)f({{{\mathcal {A}}}}(T))}}{{P_{\infty }}}} \right) . \end{aligned}$$

By the comparison arguments, we have \(\lim _{T\rightarrow \infty }{\mathcal {A}}(T)=0\). It then follows from (4c) that

$$\begin{aligned} \frac{{d{{{\mathcal {W}}}}}}{{\mathrm{d}T}} \le \frac{{ - {\mu _W}}}{{{P_\infty }}}{{{\mathcal {W}}}}(T). \end{aligned}$$

Hence, \(\lim _{T\rightarrow \infty }{\mathcal {W}}(T)=0\).

Similarly, when \({\hat{z}}_T(\phi _{_{\mathcal {CW}}})\in Y_2\), i.e., \({\mathcal {A}}(T)\equiv 0\), it follows from (4a) and (4c) that

$$\begin{aligned} \frac{{\mathrm{d} {\mathcal {C}}}}{{\mathrm{d}T}} = - \frac{{{\mu _{C}} + \beta H}}{P_0}{\mathcal {C}}(T)\qquad \text {and}\qquad \\ \frac{{d{{{\mathcal {W}}}}}}{{\mathrm{d}T}} = \frac{{ - {\mu _W}}}{{{P_0}}}{{{\mathcal {W}}}}(T), \end{aligned}$$

respectively. Thus, \({\mathcal {C}}(T)\rightarrow 0\) and \({\mathcal {W}}(T)\rightarrow 0\) as \(T\rightarrow \infty \).

If \({\hat{z}}_T(\phi _{_{\mathcal {CW}}})\in Y_3\), i.e., \({\mathcal {W}}(T)\equiv 0\), then (4) becomes (28). Since \({\mathcal {R}}_{s} > 1\), (\(\hbox {Q}_1\))–(\(\hbox {Q}_3\)), (\(\hbox {H}_2\)) and (\(\hbox {H}_3\)) hold, it follows from [16, Theorem 4.3] that \( \left( {\mathcal {C}}(T),{\mathcal {A}}(T)\right) \rightarrow \left( {\mathcal {C}}_1^*,{\mathcal {A}}_1^* \right) \) as \(T\rightarrow \infty \). Hence, \(\bigcup \left\{ \hat{\omega }(\phi _{_{\mathcal {CW}}}):\phi _{_{\mathcal {CW}}} \in {M_\partial } \right\} = E_0\cup E_1\). Thus, the claim holds.

Define a continuous function \(\rho _2:D_{Y}\rightarrow {\mathbb {R}}_+\) by

$$\begin{aligned} \rho (\phi _{_{\mathcal {CW}}})=\min \{{\mathcal {C}}_m,\hat{\phi }(m),{\mathcal {W}}_m\}, ~\forall \phi _{_{\mathcal {CW}}}\in {D_{Y}}.\end{aligned}$$

It is clear that \(\rho ^{-1}(0,\infty )\subset Y_0\) and if \(\rho (\phi _{_{\mathcal {CW}}})>0\) then \(\rho ({\hat{z}}_T(\phi _{_{\mathcal {CW}}}))>0\) for all \(T>m\). By Claim 4, we know that any forward orbit of \({\hat{z}}_T\) in \(M_{\partial }\) converges to either \(E_0\) or \(E_1\). We further conclude \(E_0\) and \(E_1\) are two isolated invariant in Y, and \(\left( W^s(E_0)\cup W^s(E_1)\right) \cap Y_0 = \emptyset \) and no subset of \(\{E_0,E_1\}\) form a cycle in \(\partial Y_0\), from Claims 2 and 3. Thus, there exists \(\eta _2>0\) such that \(\liminf \limits _{T\rightarrow \infty } ({\mathcal {C}}(T),{\mathcal {A}}(T),{\mathcal {W}}(T))\ge (\eta _2,\eta _2,\eta _2)\) for any \(({\mathcal {C}}_m,\hat{\phi },{\mathcal {W}}_m)\in Y_0\) implying the uniform persistence. The proof is complete. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Al-Darabsah, I., Yuan, Y. Dynamic analysis of a biocontrol of sea lice by age-structured model. Nonlinear Dyn 97, 1649–1666 (2019). https://doi.org/10.1007/s11071-019-05088-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-019-05088-1

Keywords

Navigation