Skip to main content

Advertisement

Log in

Hopf bifurcation analysis in a delayed Leslie–Gower predator–prey model incorporating additional food for predators, refuge and threshold harvesting of preys

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

Abstract

In this paper, we formulate and analyze a modified Leslie–Gower predator–prey model. Our model incorporates refuge of preys, additional fixed food for predators, harvesting of preys through a continuous threshold policy and a time delay as to account for predators maturity time. We first carry out a qualitative analysis of the model without time delay, showing existence of extinction, prey-free, predator-free and coexistence equilibria. We further study their stability conditions. Relying only on theoretical results of the model, we construct bifurcation diagrams involving refuge and harvest limit parameters. This led to summarize different scenarios for the model including elimination of one species or competition of both species that are proved possible. Furthermore, considering the time delay as bifurcation parameter, we analyze the stability of the coexistence equilibria and prove the system can undergoes a Hopf bifurcation. The direction of that Hopf bifurcation and the stability of the bifurcated periodic solution are determined by applying the normal form theory and the center manifold theorem. Numerical simulations are presented to illustrate our theoretical results.

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

Similar content being viewed by others

References

  1. Alves, M.T.: Des interactions indirectes entre les proies : modélisation et influence du comportement du prédateur commun. Ph.D. thesis, Université de Nice Sophia Antipolis (2013)

  2. Arino, O., Hbid, F.M.L., Ait Dads, E.: Delay Differential Equations and Applications. Springer, New York (2006)

    MATH  Google Scholar 

  3. Auger, P., Lett, C., Poggiale, J.C.: Modélisation mathématique en écologie, cours et exercices corrigés. Dunod, Paris (2010)

    Google Scholar 

  4. Azbelev, N.V., Maksimov, V.P., Rakhmatullina, L.F.: Introduction to the Theory of Functional Differential Equations: Methods and Applications, Contemporary Mathematics and Its Applications, vol. 3. Springer, New York (2006)

    Google Scholar 

  5. Balachandran, B., Kalmr-Nagy, T., Gilsinn, D.E.: Delay Differential Equations. Springer, New York (2009)

    Google Scholar 

  6. Beretta, E., Kuang, Y.: Global analysis in some delayed ratio-dependent predator-prey systems. Nonlinear Anal. 32, 381–408 (1998)

    MathSciNet  MATH  Google Scholar 

  7. Bilde, T., Toft, S.: Quantifying food limitation of arthropod predators in the field. Oecologia 115, 54–58 (1998)

    Google Scholar 

  8. Bohn, J., Rebaza, J., Speer, K.: Continuous threshold prey harvesting in predator–prey models. Int. J. Math. Comput. Phys. Electr. Comput. Eng. 5(7), 996–1003 (2011)

    MathSciNet  Google Scholar 

  9. Boualeu, H., Brouzet, R., Elsner, B., Kaczmarek, L., Pennequin, D.: Mathématiques L1, cours complet avec 1000 tests et exercices corrigés. Pearson Education France, Paris (2007)

    Google Scholar 

  10. Brauer, F.: Stability of some population models with delay. Math. Biosci. 33, 345–358 (1977)

    MathSciNet  MATH  Google Scholar 

  11. Chen, F., Chen, L., Xie, X.: On a Leslie–Gower predator–prey model incorporating a prey refuge. Nonlinear Anal. RWA 10(5), 2905–2908 (2009)

    MathSciNet  MATH  Google Scholar 

  12. Chow, S.N., Li, C.Z., Wang, D.: Normal Forms and Bifurcation of Planar Vector Fields. Cambridge University Press, Cambridge (1994)

    MATH  Google Scholar 

  13. Clark, C.W.: Mathematical Bioeconomics: The Optimal Management of Renewable Resource, 2nd edn. Wiley, New York (1990)

    MATH  Google Scholar 

  14. Coll, M., Guershon, M.: Omnivoryin terrestrial arthropods: mixing plant and prey diets. Annu. Rev. Entomol. 47, 267–297 (2002)

    Google Scholar 

  15. Cooke, K.L., Grossman, Z.: Discrete delay, distributed delay and stability switches. J. Math. Anal. Appl. 86, 592–627 (1982)

    MathSciNet  MATH  Google Scholar 

  16. Culshaw, R.V., Ruan, S.: A delay differential equation model of HIV infection of CD4 T-cells. Math. Biosci. 169, 27–39 (2000)

    MATH  Google Scholar 

  17. Datta, J., Jana, D., Upadhyay, R.K.: Bifurcation and bio-economic analysis of a prey-generalist predator model with Holling type IV functional response and nonlinear age-selective prey harvesting. Chaos Solitons Fractals 122, 229–235 (2019)

    MathSciNet  Google Scholar 

  18. Diekmann, O., Stephan, A., Sjoerd, M.V.O., Verduyn, L., Hans-Otto, W.: Delay Equations Functional, Complex, and Nonlinear Analysis. Springer, New York (1995)

    MATH  Google Scholar 

  19. Dieudonne, J.: Foundations of Modern Analysis. Academic Press, New York (1960)

    MATH  Google Scholar 

  20. Feng, P.: Analysis of a delayed predator-prey model with ratio-dependent functional response and quadratic harvesting. J. Appl. Math. Comput. Sci. 44, 251–262 (2006)

    MathSciNet  MATH  Google Scholar 

  21. Gazi, N.H., Bandyopadhyay, M.: Effect of time delay on a harvested predator-prey model. J. Appl. Math. Comput. 26, 263–280 (2008)

    MathSciNet  MATH  Google Scholar 

  22. Gopalsamy, K.: Stability and Oscillations in Delay Differential Equations of Population Dynamics. Kluwer Academic, Dordrecht (1992)

    MATH  Google Scholar 

  23. Haque, M., Greenbach, D.: When a predator avoids infected prey: a model-based theoretical study. Math. Med. Biol. 27(1), 75–94 (2010)

    MathSciNet  MATH  Google Scholar 

  24. Harmon, J.P.: Indirect interactions among a generalist predator and its multiple foods, Ph.D. thesis, St Paul MN University of Minnesota (2003)

  25. Harwood, J.D., Obrycki, J.J.: The role of alternative prey in sustaining predator populations, In: Hoddle, M.S. (ed.) Proceedings of the second international symposium biology control of arthropods II, pp. 453–462 (2005)

  26. Hassard, B., Kazarinoff, D., Wan, Y.: Theory and Applications of Hopf Bifurcation. Cambridge University Press, Cambridge (1981)

    MATH  Google Scholar 

  27. Hirsch, M.W., Smale, B.S., Devaney, R.L.: Differential Equations, Dynamical Systems and an Introduction to Chaos. Elsevier Academic Press, Cambridge (2004)

    MATH  Google Scholar 

  28. 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(1), 672–683 (2006)

    MathSciNet  MATH  Google Scholar 

  29. Jordan, D.W., Smith, P.: Nonlinear Ordinary Differential Equations, An introduction for Scientists and Engineers Fourth Edition. Oxford University Press, Oxford (2007)

    MATH  Google Scholar 

  30. Kar, T.Kumar: Modelling and analysis of a harvested prey-predator system incorporating a prey refuge. J. Comput. Appl. Math. 185(1), 19–33 (2006)

    MathSciNet  MATH  Google Scholar 

  31. Kar, T.K., Ghorai, A.: Dynamic behaviour of a delayed predator-prey model with harvesting. Appl. Math. Comput. 217(22), 9085–9104 (2011)

    MathSciNet  MATH  Google Scholar 

  32. Kuang, Y.: Delay Differential Equations with Applications in Population Dynamics. Academic Press, New York (1993)

    MATH  Google Scholar 

  33. Lashkarian, R.S., Sharifabad, D.B.: Threshold harvesting policy and delayed ratio-dependent functional response predator-prey model. Comput. Methods Differ. Equ. 4(1), 1–18 (2016)

    MathSciNet  MATH  Google Scholar 

  34. Leard, B., Lewis, C., Rebaza, J.: Dynamics of ratio-dependent of predator-prey models with nonconstant harvesting. Dict. Cont. Dyn. Syst. S 1(7), 303–315 (2008)

    MathSciNet  MATH  Google Scholar 

  35. Lenzini, P., Rebaza, J.: Nonconstant predator harvesting on ratio-dependent predator-prey models. Appl. Math. Sci. 4(16), 791–803 (2010)

    MathSciNet  MATH  Google Scholar 

  36. Leslie, P.H.: Some further notes on the use of matrices in population mathematics. Biometrika 35, 213–245 (1948)

    MathSciNet  MATH  Google Scholar 

  37. Leslie, P.H.: A stochastic model for studying the properties of certain biological systems by numerical methods. Biometrika 45, 16–31 (1958)

    MathSciNet  MATH  Google Scholar 

  38. Liu, J., Zhang, Z., Fu, M.: Stability and bifurcation in a delayed Holling-Tanner predator-prey system with ratio-dependant functional response. J. Appl. Math. 5, 384293 (2012)

  39. Liu, J.: Dynamical analysis of a delayed predator-prey system with modified Leslie–Gower and Beddington–De angelis functional response. Adv. Differ. Equ. 314, 1–19 (2014)

    Google Scholar 

  40. Liu, Q., Lin, Y., Cao, J.: Global hopf bifurcation on two delays Leslie–Gower predator-prey system with a prey refuge. Comput. Math. Methods Med. 2014, 1–12 (2014)

    MathSciNet  MATH  Google Scholar 

  41. Liu, C., Lu, N., Zhang, Q., Li, J., Liu, P.: Modeling and analysis in a prey-predator system with commercial harvesting and double time delays. Appl. Math. Comput. 281, 77–101 (2016)

    MathSciNet  MATH  Google Scholar 

  42. Lotka, A.J.: Analytical note on certain rhythmic relations in organics systems. In: Proceeding of the National Academy of Sciences of the United States of America, vol. 6(7), pp. 410–415 (1920)

    Google Scholar 

  43. Ma, Y.F.: Global Hopf bifurcation in the Leslie–Gower predator–prey ststem model with two delays. Nonlinear Anal. RWA 13, 370–375 (2012)

    MATH  Google Scholar 

  44. Martin, A., Ruan, S.: predator-prey models with delay and prey harvesting. J. Math. Biol. 43, 247–267 (2001)

    MathSciNet  MATH  Google Scholar 

  45. Meza, M.E.M., Bhaya, A., Kaskurewiczk, M.I.S.C.: Threshold policies control for predator-prey systems using a control Liapunov function approach. Theor. Pop. Biol. 67, 273–284 (2005)

    MATH  Google Scholar 

  46. Moulin, F., Ruaud, J.F., Miquel, A., Sifre, J.C.: Mathematiques TOUT-EN-UN, premiere année, cours et exercices corrigés. Dunod, deuxième édition, MPSI-PCSI (2003)

  47. Murdoch, W.W., Chesson, J., Chesson, P.L.: Biological control in theory and practice. Am. Nat. 125(3), 344–366 (1985)

    Google Scholar 

  48. Nindjin, A.F., Aziz-Alaoui, M.A., Cadivel, M.: Analysis of a predator-prey model with modified Leslie–Gower and Holling-type II schemes with time delay. Nonlinear Anal. RWA 7(5), 1104–1118 (2006)

    MathSciNet  MATH  Google Scholar 

  49. Prasad, B.S.R.V., Banerjee, M., Srinivasu, P.D.N.: Dynamics of additional food provided predator-prey system with mutually interfering predators. Math. Bio Sci. 246, 176–190 (2013)

    MathSciNet  MATH  Google Scholar 

  50. Rayungsari, M., Kusumawinahyu, W.M.: Marsudi, dynamical analysis of predator–prey model with ratio-dependent functional response and predator harvesting. Appl. Math. Sci. 8(29), 1401–1410 (2014)

    Google Scholar 

  51. Sabelis, M.W., van Rijn, P.C.J.: When does alternative food promote biological pest control? In: Hoddle, M.S. (ed.) Proceedings of the second international symposium on biological control of arthropods II, pp. 428–437 (2005)

  52. Sabelis, M.W., Van Rijn, P.C.J.: When does alternative food promote biological pest control? IOBCWPRS Bull. 29, 195 (2006)

    Google Scholar 

  53. Sahoo, B.: Disease control through provision of alternative food to predator: a model based study. Int. J. Dyn. Control (2014). https://doi.org/10.1007/s40435-014-0099-0

    Article  Google Scholar 

  54. Sahoo, B., Poria, S.: Effects of additional food in a delayed predator-prey model. Math. Biosci. 270, 62–73 (2015)

    MathSciNet  MATH  Google Scholar 

  55. Sen, M., Srinivasu, P.D.N., Banerjee, M.: Global dynamics of an additional food provided predator-prey system with constant harvest in predators. Appl. Math. Comput. 250, 193–211 (2015)

    MathSciNet  MATH  Google Scholar 

  56. Srinivasu, P.D.N., Prasad, B.S.R.V.: Role of quantity of additional food to predators as a control in predatorprey systems with relevance to pest management and biological conservation. Bull. Math. Biol. 73(10), 2249–2276 (2011)

    MathSciNet  MATH  Google Scholar 

  57. Srinivasu, P.D.N., Prasad, B.S.R.V., Venkatesulu, M.: Biological control through provision of additional food to predators: a theoretical study. Theor. Popul. Biol. 72(1), 111–120 (2007)

    MATH  Google Scholar 

  58. Stephens, D.W., Krebs, J.R.: Foraging Theory. Monographs in Behavior and Ecology. Princeton University Press, Princeton (1986)

    Google Scholar 

  59. Tankam, I., Mouofo, P. T., Mendy, A., Lam, M., Tewa, J.J., Bowong, S.: Local bifurcations and optimal theory in delayed predator-prey model with threshold prey harvesting. Int. J. Bifurc. Chaos 25(7), 1540015 (2015)

    MathSciNet  MATH  Google Scholar 

  60. Ulfa, H.M., Suryanto, A., Darti, I.: Dynamics of Leslie-Gower predator-prey model with additional food for predators. Int. J. Pure Appl. Math. 115(2), 199–209 (2017)

    Google Scholar 

  61. Upadhyay, R.K., Roy, P., Datta, J.: Complex dynamics of ecological systems under nonlinear harvesting: Hopf bifurcation and Turing instability. Nonlinear Dyn. 79, 2251–2270 (2015)

    MathSciNet  MATH  Google Scholar 

  62. Van Baalen, M., Kivan, V., Van Rijn, P.C.J., Sabelis, M.W.: Alternative food, switching predators, and the persistence of predator–prey systems. Am. Nat. 157(5), 512–524 (2001)

    Google Scholar 

  63. Van Rijn, P.C.J., Van Houten, Y.M., Sabelis, M.W.: How plants benefit from providing food to predators even when it is also edible to herbivores. Ecology 83, 2664–2679 (2002)

    Google Scholar 

  64. Volterra, V.: Variations and fluctuations of the number of individuals in animal species living together. In: Chapman, R.N. (ed.) Animal Ecology. McGraw-Hill, New York (1931)

    Google Scholar 

  65. Xiao, D., Li, W.: Stability and bifurcation in a delayed ratio-dependent predator-prey system. Proc. Edinb. Math. Soc. 46, 205–220 (2003)

    MathSciNet  MATH  Google Scholar 

  66. Xu, C., Liao, M., He, X.: Stability and Hopf bifurcation analysis for a Lotka-Volterra predator-prey model with two delays. Int. J. Appl. Math. Comput. Sci. 21(1), 97–107 (2011)

    MathSciNet  MATH  Google Scholar 

  67. Yafia, R., El Adnani, F., Alaoui, H.T.: Stability of limit cycle in a predator-prey model with modified Leslie-Gower and Holling-type II schemes with time delay. Appl. Math. Sciences 1(1—-4), 119–131 (2007)

    MathSciNet  MATH  Google Scholar 

  68. Yan, X.P., Chu, Y.D.: Stability and bifurcation analysis for a delayed Lotka–Volterra predator-prey system. J. Comput. Appl. Math. 196, 198–210 (2006)

    MathSciNet  MATH  Google Scholar 

  69. Yan, X., Li, W.: Hopf bifurcation and global periodic solutions in a delayed predator-prey system. Appl. Math. Comput. 177, 427–445 (2006)

    MathSciNet  MATH  Google Scholar 

  70. Yuan, S., Song, Y.: Stability and Hopf bifurcations in a delayed Leslie-Gower predator-prey system. J. Math. Anal. Appl. 355, 82–100 (2009)

    MathSciNet  MATH  Google Scholar 

  71. Yuan, S.L., Song, Y.L.: Bifurcation and stability analysis for a delayed Leslie-Gower predator-prey system. IMA J. Appl. Math. 74, 574–603 (2009)

    MathSciNet  MATH  Google Scholar 

  72. Yuan, S.L., Song, Y.L., Zhang, J.M.: Bifurcation analysis in the delayed Leslie-Gower predator-prey system. Appl. Math. Model. 33, 4049–4061 (2009)

    MathSciNet  MATH  Google Scholar 

  73. Yuan, R., Jiang, W., Wang, Y.: Saddle-node-Hopf bifurcation in a modified Leslie-Gower predator-prey model with time-delay and prey harvesting. J. Math. Anal. Appli. 422, 1072–1090 (2015)

    MathSciNet  MATH  Google Scholar 

  74. Zhang, Z., Upadhyay, R.K., Datta, J.: Bifurcation analysis of a modified Leslie-Gower model with Holling type-IV functional response and nonlinear prey harvesting. Adv. Differ. Equ. 2018, 127 (2018)

    MathSciNet  MATH  Google Scholar 

  75. Zi-Zhen, Z., Hui-Zhong, Y.: Hopf bifurcation in a delayed predator-prey system with modified Leslie-Gower and Holling type III schemes. Acta Autom. Sin. 39(5), 610–616 (2013)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

Thanks to the reviewers for insightful comments that helped us to improve the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean Jules Tewa.

Ethics declarations

Conflict of interest

The authors Maximilien Onana, Boulchard Mewoli and Jean Jules Tewa 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

Proof of Theorem 1

We must show that each solution (xy) of system (2)–(3), defined and continuous on \([-\tau ,A[\) where \(A \in ]0,+\infty ]\), \(x(t)>0\) and \(y(t)>0\) for all \(t\in [0,A[\). Suppose that it is not true. Then there exists a value of T in ]0, A[ such that for all \(t\in [0,T[\), \(x(t)>0\) and \(y(t)>0\), and either \(x(T)=0\) or \(y(T)=0\).

For all \(t\in [0,T[\) and from equations of (2), we have,

$$\begin{aligned}&x(t)>x(0)\exp \left( \displaystyle \int _{0}^{t}\left( \displaystyle r_{1}-b_{1}x(s)-q\lambda _{1}(1-m)y(s)\right. \right. \\&\quad \left. \left. -\frac{h}{h+x(s)-T_{1}}\right) \mathrm{d}s\right) \end{aligned}$$

and

$$\begin{aligned}&y(t)=y(0)exp\left( \displaystyle \int _{0}^{t}\left( \displaystyle r_{2}\right. \right. \\&\left. \left. \quad \left( 1 -\displaystyle \frac{y(s - \tau )}{q\alpha _{1}(1-m)x(s - \tau ) + (1 - q)\alpha _{A}K_{A}}\right) \right) \mathrm{d}s\right) . \end{aligned}$$

As x and y are defined and continuous on the compact \([-\tau ,T]\), there exists \(M \ge 0\) such that:

$$\begin{aligned}&x(t)>x(0)exp\left( \displaystyle \int _{0}^{t}\left( \displaystyle r_{1}-b_{1}x(s)-a_{1}(1-m)y(s)\right. \right. \\&\quad \left. \left. -\frac{h}{h+x(s)-T_{1}}\right) \mathrm{d}s\right) \ge x(0)exp(-TM) \end{aligned}$$

and

$$\begin{aligned}&y(t)=y(0)exp\left( \displaystyle \int _{0}^{t}\left( \displaystyle r_{2}\right. \right. \\&\left. \left. \quad \left( 1 -\displaystyle \frac{y(s - \tau )}{q\alpha _{1}(1-m)x(s - \tau ) + (1 - q)\alpha _{A}K_{A}}\right) \right) \mathrm{d}s\right) \\&\quad \ge y(0)exp(-TM). \end{aligned}$$

Taking the limit as \(t \rightarrow T\) gives

$$\begin{aligned} x(T)\ge x(0)exp(-TM)> 0 \end{aligned}$$

and

$$\begin{aligned} y(T)\ge y(0)exp(-TM)> 0, \end{aligned}$$

which contradicts the fact that either \(x(T)= 0\) or \(y(T)= 0\). Then, a solution of System (2)–(3) which starts in the positive quadrant \({\mathbb {R}}_{+}^2\) remains there.

Proof of Theorem 2

Using the first equation of System (2), we always have

$$\begin{aligned} \dot{x}(t)\le r_{1}x(t)\left( \displaystyle 1 -\displaystyle \frac{x(t)}{K_{1}}\right) . \end{aligned}$$

Applying a differential inequality [Hale, 1980] gives

$$\begin{aligned} {x}(t)\le \frac{1}{\displaystyle {\frac{1}{K_{1}} + \left( \displaystyle \frac{1}{x(0)}-\frac{1}{K_{1}}\right) e^{-r_{1}t}}}. \end{aligned}$$

Using assumption (5) gives \(x(t)\le K_{1}\), for all positive value of t. It means that x(t) is bounded.

Let us use the second equation of System (3) to prove that y(t) is bounded. For all \(t\ge 0\), we have

$$\begin{aligned} \dot{y}(t)\le r_{2}y(t). \end{aligned}$$

Integrating that differential inequality from \(t-\tau \) to t gives

$$\begin{aligned} y(t-\tau )\ge y(t)e^{-r_{2}\tau }. \end{aligned}$$

Using the fact that \(x(t)\le K_{1}\) gives the following differential inequality

$$\begin{aligned} \dot{y}(t)\le r_{2}y(t)\left( \displaystyle 1 -\displaystyle \frac{y(t)}{K}\right) \end{aligned}$$

where \(K = \displaystyle \frac{(1-m)q\alpha _{1}K_{1}+(1-q)\alpha _{A}K_{A}}{e^{-r_{2}\tau }}\). Once more, applying a differential inequality [Hale, 1980] gives

$$\begin{aligned} {y}(t)\le \frac{1}{\displaystyle {\frac{1}{K} + \left( \displaystyle \frac{1}{y(0)}-\frac{1}{K}\right) e^{-r_{2}t}}}. \end{aligned}$$

Using assumption (6) gives \(y(t)\le K\), for all positive value of t. Thus y(t) is bounded. One can easily verify that, using assumption (6), the inequalities \(x(t)\le K_{1}\) and \(e^{r_{2}\tau }\ge 1\) lead to \(y(0)\le K\).

Proof of Theorem 3

A couple of variables (xy) is an equilibrium of system (2)–(3) if it is a solution of the following systems on \([0,T_1]\times {\mathbb {R}}_{+}\) and on \([T_1,K_1]\times {\mathbb {R}}_{+}\), respectively.

$$\begin{aligned}&\left\{ \begin{array}{lll} 0 &{}=&{} r_{1}\left( 1 -\displaystyle \frac{x(t)}{K_{1}}\right) x(t) - q\lambda _{1}(1-m)x(t)y(t), \\ 0 &{}=&{} r_{2}\left( 1 -\displaystyle \frac{y(t)}{q\alpha _{1}(1-m)x(t) + (1 - q)\alpha _{A}K_{A}}\right) y(t), \end{array} \right. \nonumber \\ \end{aligned}$$
(20)
$$\begin{aligned}&\left\{ \begin{array}{lll} 0 &{}=&{} r_{1}\left( 1 -\displaystyle \frac{x(t)}{K_{1}}\right) x(t) - q\lambda _{1}(1-m)x(t)y(t) \\ &{}&{}\quad - \displaystyle \frac{h(x-T_{1})}{h+x-T_{1}}, \\ 0 &{}=&{} r_{2}\left( 1 -\displaystyle \frac{y(t)}{q\alpha _{1}(1-m)x(t) + (1 - q)\alpha _{A}K_{A}}\right) y(t). \end{array} \right. \nonumber \\ \end{aligned}$$
(21)

Firstly, we solve system (20) on \([0,T_1]\times {\mathbb {R}}_{+}\). Using the first equation of the system (20), we have \(x=0\) or \(x = \displaystyle \frac{r_{1}-q\lambda _{1}(1-m)y}{b_{1}}\).

Replacing x by 0 in the second equation of system (20) gives \(y=0\) or \(y=(1-q)\alpha _{A}K_A\). Then we have the equilibria \(E_{1}^{0}\) and \(E_{2}^{0}\).

Replacing x by \(\displaystyle \frac{r_{1}-q\lambda _{1}(1-m)y}{b_{1}}\) in the second equation of system (20) gives \(y=0\) or \(y=(1-m)q\alpha _{1}x + (1-q)\alpha _{A}K_A\). If \(y=0\), we have \(x=\displaystyle \frac{r_{1}}{b_{1}}\). If \(y=(1-m)q\alpha _{1}x + (1-q)\alpha _{A}K_A\), we have \(x=\displaystyle \frac{r_{1}-\lambda _{1}(1-m)q(1-q)\alpha _{A}K_{A}}{b_{1}+q^{2}\lambda _{1}\alpha _{1}(1-m)^{2}}\) and \(y=\displaystyle \frac{r_{1}\alpha _{1}(1-m)q+(1-m)b_{1}\alpha _{A}K_{A}}{b_{1}+q^{2}\lambda _{1}\alpha _{1}(1-m)^{2}}\). Then we have the equilibria \(E_{3}^{0}\) and \(E_{4}^{0}\).

Secondly, we solve system (21) on \([T_1,K_1]\times {\mathbb {R}}_{+}\).

Using the second equation of system (21) gives \(y=0\) or \(y=(1-m)q\alpha _{1}x + (1-q)\alpha _{A}K_A\). If \(y=0\), we find that x is a solution of equation (10). Moreover, if we consider the function f defined by \(f(x) = -b_{1}x^{3}+(r_{1}-b_{1}(h-T_{1}))x^{2}+(r_{1}(h-T_{1})-h)x+hT_{1}\), then we have \(f(T_1) = hT_{1}(r_{1} - b_{1}T_{1})> 0\) using assumption (4). We also have \(f(K_1) = -h(K_{1} - T_{1}) < 0\) because \(T_{1}<K_{1}\). Thus, by the intermediate value theorem [9, 46], there exists at least one solution of equation (10). So we have the equilibria \(E_{1}^{\varphi }\). If \(y=(1-m)q\alpha _{1}x + (1-q)\alpha _{A}K_A\), x is a solution of equation (11). Moreover, if we consider the function g defined by \(g(x) = A_{1}^{\varphi }x^{3}+ A_{2}^{\varphi }x^{2}+A_{3}^{\varphi }x+hT_{1}\), we have \(g(T_{1}) = hT_{1}(r_{1} - b_{1}T_{1} - q\lambda _{1}(1 - m)(1 - q)\alpha _{A}K_A + q\alpha _{1}(1 - m)T_{1})\). We have \(g(K_1) = -(K_{1}(K_{1} - T_{1}) + hT_1)(q\lambda _{1}(1 - m)(1 - q)\alpha _{A}K_A + q^{2}\lambda _{1}\alpha _{1}K_{1}(1 - m)^{2})< 0\) because \(T_{1}<K_{1}\). Thus, if \(r_{1} - b_{1}T_{1} - q\lambda _{1}(1 - m)(1 - q)\alpha _{A}K_{A} + q(1 - m)\alpha _{1}T_{1} > 0\) then \(g(T_1)>0\). By the intermediate value theorem [9, 46], (11) has at least one solution in \([K_{1}, T_{1}]\). Thus we have the equilibrium \(E_{2}^{\varphi }\).

Proof of Theorem 4

  1. 1.

    The Jacobian matrix of system (2)–(3) at the equilibrium \(E_{1}^{0}\) is:

    $$\begin{aligned} J_{E_{1}^{0}}= \left( \begin{array}{cc} r_{1} &{}\quad 0\\ 0 &{}\quad r_{2} \end{array}\right) . \end{aligned}$$

    \(J_{E_{1}^{0}}\) has two positive eigenvalues \((r_{1} \quad and \quad r_{2})\). So the equilibrium \(E_{1}^{0}\) is an unstable node.

  2. 2.

    The Jacobian matrix of system (2)–(3) at the equilibrium \(E_{2}^{0}\) is:

    $$\begin{aligned} J_{E_{2}^{0}}= \left( \begin{array}{cc} r_{1} - q\lambda _{1}(1-m)(1-q)\alpha _{A}K_{A} &{}\quad 0\\ r_{2}(1-m)\alpha _{1}q &{}\quad -r_{2} \end{array}\right) . \end{aligned}$$

    \(J_{E_{2}^{0}}\) has one negative eigenvalue \((-r_{2})\). The stability of the equilibrium \(E_{2}^{0}\) depends on the sign of the second eigenvalue \(r_{1} - q\lambda _{1}(1-m)(1-q)\alpha _{A}K_{A}\). Moreover, the discriminant of the characteristic equation is \(\varDelta _{E_{2}^{0}} = (r_ + r_{1} - q\lambda _{1}(1-m)(1-q)\alpha _{A}K_{A})^2 > 0\). Thus if \(r_{1} < q\lambda _{1}(1-m)(1-q)\alpha _{A}K_{A}\), then the equilibrium \(E_{2}^{0}\) is stable node and the equilibrium \(E_4^0\) does not exist. If \(r_{1} > q\lambda _{1}(1-m)(1-q)\alpha _{A}K_{A}\), the equilibrium \(E_{2}^{0}\) is a saddle (unstable).

  3. 3.

    The Jacobian matrix of system (2)–(3) at the equilibrium \(E_{3}^{0}\) is:

    $$\begin{aligned} J_{E_{3}^{0}}= \left( \begin{array}{cc} -r_{1} &{}\quad -q\lambda _{1}(1-m)K_{1}\\ 0 &{}\quad r_{2} \end{array}\right) . \end{aligned}$$

    \(J_{E_{3}^{0}}\) has one positive eigenvalue \((r_{2})\). The equilibrium \(E_{3}^{0}\) is a saddle (unstable).

  4. 4.

    The Jacobian matrix of system (2)–(3) at the equilibrium \(E_{4}^{0}\) is:

    $$\begin{aligned}&J_{E_{4}^{0}}= \left( \begin{array}{cc} -b_{1}x_{4}^{0} &{}\quad -q\lambda _{1}(1-m)x_{4}^{0}\\ (1-m)\alpha _{1}r_{2}q &{}\quad -r_{2} \end{array}\right) .\\&Tr(J_{E_{4}^{0}})=-(b_{1}x_{4}^{0}+r_{2})< 0,\\&|(J_{E_{4}^{0}})|= r_{2}b_{1}x_{4}^{0}+(1-m)^{2}\alpha _{1}\lambda _{1}r_{2}q^{2}x_{4}^{0}> 0. \end{aligned}$$

    Moreover, the discriminant of the characteristic equation is \(\varDelta _{E_{4}^{0}} = (b_{1}x_{4}^{0} - r_2)^{2} - 4(1 - m)^{2}\alpha _{1}\lambda _{1}r_{2}q^{2}x_{4}^{0}.\) Therefore, if \(\varDelta _{E_{4}^{0}}> 0\), then the equilibrium \(E_{4}^{0}\) is a stable node. If \(\varDelta _{E_{4}^{0}}< 0\), then the equilibrium \(E_{4}^{0}\) is a stable spiral. If \(\varDelta _{E_{4}^{0}}= 0\), then the equilibrium \(E_{4}^{0}\) is a stable degenerate node.

  5. 5.

    The Jacobian matrix of system (2)–(3) at the equilibrium \(E_{1}^{\varphi }\) is:

    $$\begin{aligned} J_{E_{1}^{\varphi }}= \left( \begin{array}{cc} \displaystyle -b_{1}x_{1}^{\varphi }+ \frac{\varphi (x_{1}^{\varphi })}{x_{1}^{\varphi }}- \varphi '(x_{1}^{\varphi }) &{}\quad -q\lambda _{1}(1-m)x_{2}^{\varphi }\\ 0 &{}\quad r_{2} \end{array}\right) . \end{aligned}$$

    \(J_{E_{1}^{\varphi }}\) has one positive eigenvalue \((r_{2})\). Thus, the equilibrium \(E_{1}^{\varphi }\) is unstable. Moreover, if \(-b_{1}x_{1\varphi } + \displaystyle \frac{h(x_{1}^{\varphi } - T_1)^2 - h^{2}T_1}{x_{1}^{\varphi }(x_{1}^{\varphi } + h - T_1)^2} < 0\), then the equilibrium \(E_{1}^{\varphi }\) is a saddle. If \(-b_{1}x_{1}^{\varphi } + \displaystyle \frac{h(x_{1}^{\varphi } - T_1)^2 - h^{2}T_1}{x_{1}^{\varphi }(x_{1}^{\varphi } + h - T_1)^2} > 0\), then the equilibrium \(E_{1}^{\varphi }\) is an unstable node. If \(-b_{1}x_{1}^{\varphi } + \displaystyle \frac{h(x_{1}^{\varphi } - T_1)^2 - h^{2}T_1}{x_{1}^{\varphi }(x_{1}^{\varphi } + h - T_1)^2} = 0\), then the equilibrium \(E_{1}^{\varphi }\) is an unstable nonhyperbolic point.

  6. 6.

    The Jacobian matrix of system (2)–(3) at the equilibrium \(E_{2}^{\varphi }\) is:

    $$\begin{aligned}&J_{E_{2}^{\varphi }}= \left( \begin{array}{cc} \displaystyle -b_{1}x_{2}^{\varphi }+ \frac{\varphi (x_{2}^{\varphi })}{x_{2}^{\varphi }}- \varphi '(x_{2}^{\varphi }) &{}\quad -q\lambda _{1}(1-m)x_{2}^{\varphi }\\ (1-m)\alpha _{1}r_{2}q &{}\quad -r_{2} \\ \end{array}\right) .\\&Tr(J_{E_{2}^{\varphi }})= \displaystyle -b_{1}x_{2}^{\varphi } - r_{2}+ \displaystyle \frac{h(x_{2}^{\varphi } - T_1)^2 - h^{2}T_1}{x_{2}^{\varphi }(x_{2}^{\varphi } + h - T_1)^2},\\&|J_{E_{2}^{\varphi }}|= -r_{2}\left( \displaystyle -b_{1}x_{2}^{\varphi }+ \displaystyle \frac{h(x_{2}^{\varphi } - T_1)^2 - h^{2}T_1}{x_{2}^{\varphi }(x_{2}^{\varphi } + h - T_1)^2}\right) \\&\quad + (1-m)^{2}q^{2}\alpha _{1}\lambda _{1}r_{2} x_{2}^{\varphi }. \end{aligned}$$

    The discriminant of the characteristic equation is:

    $$\begin{aligned} \varDelta _{E_{2}^{\varphi }}= & {} \left( \displaystyle r_{2} + b_{1}x_{2}^{\varphi } - \displaystyle \frac{h(x_{2}^{\varphi } - T_1)^2 - h^{2}T_1}{x_{2}^{\varphi }(x_{2}^{\varphi } + h - T_1)^2} \right) ^{2} \\&- 4(1-m)^{2}q^{2}\alpha _{1}\lambda _{1}r_{2} x_{2}^{\varphi }. \end{aligned}$$

    Then using the signs of \(Tr(J_{E_{2}^{\varphi }})\), \(|J_{E_{2}^{\varphi }}|\) and \(\varDelta _{E_{2}^{\varphi }}\), and the table given in Jordan et al. [29], we have the type and the stability of the equilibrium \(E_{2}^{\varphi }\) as in theorem (4). (For stability of equilibria in the plane, one can also refer to Auger et al. [3] and Hirsch et al. [27].)

Proof of Theorem 5

We proved that it is possible to have purely imaginary roots for the characteristic equation (13) and the transversality condition is satisfied. Thus, we have the third item of theorem 5. In what follows, we prove the first and the second items of theorem 5. By Rouche’s theorem [19] and the continuity in \(\tau \), the characteristic equation (13) has roots with positive real parts if and only if it has purely imaginary roots. Let \(\lambda (\tau )= \mu (\tau )+i\omega (\tau )\) where \(\mu \) and \(\omega \) are real depending on \(\tau \). For \(\tau =0\), the equilibrium E is stable. Thus, we have \(\mu (0)<0\). By continuity, if \(\tau \) is sufficiently small, we still have \(\mu (\tau )<0\) and E is still stable. The change of stability will occur for some values of \(\tau \) for which \(\mu (\tau )=0\) and \(\omega (\tau )\ne 0\); it means that \(\lambda \) will be purely imaginary. Let \(\tau _{e}\) be such that \(\mu (\tau _{e})=0\) and \(\omega (\tau _{e})=\omega _{e}\ne 0\) with \(\lambda =i\omega (\tau _{e})\). In this case, the steady state loses stability and eventually becomes unstable when \(\mu (\tau )\) becomes positive. In other words, if such a value \(\omega _{e}\) does not exist, then the steady state E will remain stable for all \(\tau \).

Proof of Theorem 6

For convenience, let \(t=s\tau \), \(x(s\tau )=x_{1}(s)\), \(y(s\tau )=x_{2}(s)\) and \(\tau =\tau _{0}+\mu \), \(\mu \in {\mathbb {R}}\) so that \(\mu =0\) is the Hopf bifurcating value for system (2)–(3). Then system (2)–(3) becomes equivalent to the system:

$$\begin{aligned} \dot{u}(t)=L_{\mu }(u_{t})+f(\mu ,u_{t}), \end{aligned}$$
(22)

where \(u_{t}=(x_{1}(t),x_{2}(t))^{T} \in \mathcal {C}\) and \(u_{t}(\theta )=u(t+\theta )=(x_{1}(t+\theta ),x_{1}(t+\theta ))^{T}\in \mathcal {C}\).

\(L_{\mu }: \mathcal {C}\rightarrow {\mathbb {R}}^{2}\) is defined as follows:

$$\begin{aligned} L_{\mu }\phi= & {} (\tau _{0}+\mu )\left( \begin{array}{cc} A_{A\varphi } &{}\quad B_{A\varphi } \\ 0 &{}\quad 0 \end{array} \right) \left( \begin{array}{c} \phi _{1}(0) \\ \phi _{2}(0) \end{array} \right) \nonumber \\&+ (\tau _{0}+\mu )\left( \begin{array}{cc} 0 &{}\quad 0 \\ C_{A\varphi } &{}\quad D_{A\varphi } \end{array} \right) \left( \begin{array}{c} \phi _{1}(-1) \\ \phi _{2}(-1) \\ \end{array} \right) , \end{aligned}$$
(23)

where \(A_{A\varphi }=-b_{1}x_{e}+ \displaystyle \frac{\varphi (x_{e})}{x_{e}} - \varphi '(x_{e})\), \(B_{A\varphi }=-q\lambda _{1}(1-m)x_{e}\), \(C_{A\varphi }= qr_{2}\alpha _{1}(1-m)\), \(D_{A\varphi }=-r_{2}\).

\(f:\mathcal {R}\times \mathcal {C}\rightarrow {\mathbb {R}}^{2}\) is defined as follows:

$$\begin{aligned} f(\mu ,\phi )= (\tau _{0}+\mu )(f_{1},f_{2})^{T}, \end{aligned}$$
(24)

where \(\phi (\theta )=(\phi _{1}(\theta ),\phi _{2}(\theta ))^{T} \in \mathcal {C}\)

$$\begin{aligned} f_{1}=a_{11}\phi _{1}^{2}(0)+a_{12}\phi _{1}(0)\phi _{2}(0)+a_{13}\phi _{1}^{3}(0), \end{aligned}$$

and

$$\begin{aligned} f_{2}= & {} a_{21}\phi _{1}^{2}(-1)+a_{22}\phi _{1}(-1)\phi _{2}(0)\\&+a_{23}\phi _{1}(-1)\phi _{2}(-1)+a_{24}\phi _{2}(-1)\phi _{2}(0)\\&+a_{25}\phi _{1}^{3}(-1) +a_{26}\phi _{1}^{2}(-1)\phi _{2}(0)\\&+a_{27}\phi _{1}^{2}(-1)\phi _{2}(-1)\\&+a_{28}\phi _{1}(-1)\phi _{2}(0)\phi _{2}(-1), \end{aligned}$$

with

$$\begin{aligned} a_{11}= & {} -b_{1}+ \displaystyle \frac{h^{2}}{(x_{e}+h-T_1)^{2}} , a_{12}=-q\lambda _{1}(1-m), \\ a_{13}= & {} \displaystyle \frac{h^{2}}{(x_{e}+h-T_1)^{4}}\\ a_{21}= & {} - \displaystyle \frac{r_{2}(1-m)^{2}q^{2}\alpha _{1}^{2}}{y_{e}}, a_{22}= \displaystyle \frac{r_{2}(1-m)q\alpha _{1}}{y_{e}}, \\ a_{23}= & {} \displaystyle \frac{r_{2}(1-m)q\alpha _{1}}{y_{e}}, a_{24}=- \displaystyle \frac{r_{2}}{y_{e}}, \\ a_{25}= & {} -\displaystyle \frac{r_{2}(1-m)^{3}q^{3}\alpha _{1}^{3}}{y_{e}^{2}}, a_{26}=- \displaystyle \frac{r_{2}(1-m)^{2}q^{2}\alpha _{1}^{2}}{y_{e}^{2}},\\ a_{27}= & {} - \displaystyle \frac{r_{2}(1-m)^{2}q^{2}\alpha _{1}^{2}}{y_{e}^{2}}, a_{28}= \displaystyle \frac{r_{2}(1-m)q\alpha _{1}}{y_{e}^{2}}. \end{aligned}$$

\(L_{\mu }\) is a one parameter family of bounded linear operators in \(\mathcal {C}[-1,0]\rightarrow {\mathbb {R}}^{2}\). Then by the Riesz representation theorem, there exists a matrix whose components are bounded variation functions \(\eta (\theta ,\mu )\) in \([-1,0]\rightarrow {\mathbb {R}}^{2}\) such that

$$\begin{aligned} L_{\mu }=\int _{-1}^{0}\mathrm{d}\eta (\theta ,\mu )\phi (\theta ). \end{aligned}$$
(25)

In fact, we can choose

$$\begin{aligned}&\eta (\theta ,\mu )= (\tau _{0}+\mu )\left( \begin{array}{cc} A_{A\varphi } &{}\quad B_{A\varphi } \\ 0 &{}\quad 0 \\ \end{array} \right) \left( \begin{array}{c} \phi _{1}(0) \\ \phi _{2}(0) \\ \end{array} \right) \delta (\theta )\nonumber \\&\quad + (\tau _{0}+\mu )\left( \begin{array}{cc} 0 &{}\quad 0 \\ C_{A\varphi } &{}\quad D_{A\varphi } \end{array} \right) \left( \begin{array}{c} \phi _{1}(-1) \\ \phi _{2}(-1) \end{array} \right) \delta (\theta +1),\nonumber \\ \end{aligned}$$
(26)

where \(\delta \) is the Dirac function. Then equation (22) is satisfied.

For \(\phi \in \mathcal {C}^{1}[-1,0]\), let us define

$$\begin{aligned} A(\mu )\phi =\left\{ \begin{array}{ll} \displaystyle \frac{\mathrm{d}\phi (\theta )}{\mathrm{d}\theta } , &{}\quad \text{ if }\quad -1\le \theta <0 \\ \int _{-1}^{0}\mathrm{d}\eta (\theta ,\mu )\phi (\theta ), &{}\quad \text{ if }\quad \theta =0 \end{array}\right. \nonumber \\ \end{aligned}$$
(27)

and

$$\begin{aligned} R(\mu )\phi =\left\{ \begin{array}{ll} \left( \begin{array}{c} 0\ \\ 0\ \\ \end{array} \right) , &{}\quad \text{ if }\quad -1\le \theta <0, \\ f(\mu , \phi ), &{}\quad \text{ if }\quad \theta =0. \end{array}\right. \end{aligned}$$
(28)

System (2)–(3) is then transformed into the operator equation of the form (29) as follows, in order to study the Hopf bifurcation problem

$$\begin{aligned} \dot{u}(t)=A(\mu )u_{t}+Ru_{t}. \end{aligned}$$
(29)

Define the adjoint operator for \(\psi \in \mathcal {C}^{1}([0,1], ({\mathbb {R}}^{2})^{\star })\),

$$\begin{aligned} A^{\star }(\mu )\psi (s)=\left\{ \begin{array}{ll} \displaystyle -\frac{\mathrm{d}\phi (s)}{\mathrm{d}s} &{}\quad \text{ if }\quad 0 < s \le 1, \\ \int _{-1}^{0}\mathrm{d}\eta ^{T}(s,\mu )\psi (-s) &{}\quad \text{ if }\quad s = 0. \end{array}\right. \nonumber \\ \end{aligned}$$
(30)

In order to normalize the eigenvectors of the operator A and the adjoint operator \(A^{\star }\), we need to introduce the following bilinear form:

$$\begin{aligned}&\langle \psi , \phi \rangle = \overline{\psi }(0).\phi (0)\nonumber \\&\quad - \int _{\theta =-1}^{0}\int _{\xi =0}^{\theta }\overline{\psi }^{T}(\xi -\theta )\mathrm{d}\eta (\theta )\phi (\xi )\mathrm{d}\xi , \end{aligned}$$
(31)

where \(\eta (\theta )=\eta (\theta ,0)\).

By the discussion and the transformation \(t=s\tau \), we know that \(i\tau _{0}\omega _{0}\) and \(-i\tau _{0}\omega _{0}\) are the eigenvalues of A(0) and other eigenvalues have strictly negative real parts. Hence, they are also eigenvalues of \(A^{\star }\). Now we are going to compute the eigenvectors of A(0) and \(A^{\star }\) corresponding to their respective eigenvalues \(i\tau _{0}\omega _{0}\) and \(-i\tau _{0}\omega _{0}\). If we suppose that \(q(\theta )= (q^{(1)}(\theta ), q^{(2)}(\theta ))^T = (1,q_{1})^{T}e^{i\tau _{0}\omega _{0}\theta }\) is the eigenvector of A(0) corresponding to the eigenvalue \(i\tau _{0}\omega _{0}\), then by the definition we have \(A(0)q(0)=i\tau _{0}\omega _{0}q(0)\). Then using the definition of A(0) and the expressions given by (22), (24) and (25) gives:

$$\begin{aligned}&\tau _{0}\left( \begin{array}{cc} A_{A\varphi } &{}\quad B_{A\varphi } \\ 0 &{}\quad 0 \\ \end{array}\right) q(0)+ \tau _{0}\left( \begin{array}{cc} 0 &{}\quad 0 \\ C_{A\varphi } &{}\quad \quad D_{A\varphi } \\ \end{array}\right) q(-1)\\&\quad = i\tau _{0}\omega _{0}q(0), \end{aligned}$$

or equivalently \( \tau _{0}\left( \begin{array}{cc} A_{A\varphi } &{}\quad B_{A\varphi } \\ 0 &{}\quad 0 \\ \end{array} \right) \left( \begin{array}{c} 1\ \\ q_{1} \end{array} \right) + \tau _{0}\left( \begin{array}{cc} 0 &{} \quad 0 \\ C_{A\varphi } &{}\quad D_{A\varphi } \\ \end{array} \right) \left( \begin{array}{c} 1\ \\ q_{1} \end{array} \right) e^{-i\tau _{0}\omega _{0}}= i\tau _{0}\omega _{0}\left( \begin{array}{c} 1\ \\ q_{1} \end{array} \right) \)

This implies \( \left\{ \begin{array}{ll} A_{A\varphi }+B_{A\varphi }q_{1} &{} = i\omega _{0},\\ (C_{A\varphi }+D_{A\varphi }q_{1})e^{-i\tau _{0}\omega _{0}} &{} = q_{1}i\omega _{0}, \end{array}\right. \)

and    \(q_{1}= \displaystyle \frac{C_{A\varphi }}{-D_{A\varphi } + i\omega _{0}e^{i\tau _{0}\omega _{0}}}.\)

Thus, \(q^{(1)}(\theta )= e^{i\tau _{0}\omega _{0}\theta }\) and \(q^{(2)}(\theta )= \displaystyle \frac{C_{A\varphi }e^{i\tau _{0}\omega _{0}}}{-D_{A\varphi } + i\omega _{0}e^{i\tau _{0}\omega _{0}}}.\)

Now let us compute the eigenvector \(q^{\star }\) of \(A^{\star }\). Suppose that we have \(q^{\star }(s) = G_{A\varphi }(1,q_{1}^{\star })^{T}e^{i\tau _{0}\omega _{0}s} , 0\le s \le 1\). Then we have the following relation

$$\begin{aligned}&\tau _{0}\left( \begin{array}{cc} A_{A\varphi } &{} \quad 0 \\ B_{A\varphi } &{} \quad 0 \\ \end{array}\right) q^{\star }(0)+ \tau _{0}\left( \begin{array}{cc} 0 &{} \quad C_{A\varphi } \\ 0 &{} \quad D_{A\varphi } \end{array}\right) q^{\star }(-1)\\&\quad = -i\tau _{0}\omega _{0}q^{\star }(0), \end{aligned}$$

which is equivalent to \(\tau _{0}\left( \begin{array}{cc} A_{A\varphi } &{} \quad 0 \\ B_{A\varphi } &{} \quad 0 \end{array} \right) \left( \begin{array}{c} 1\ \\ q^{\star }_{1} \end{array} \right) + \tau _{0}\left( \begin{array}{cc} 0 &{} \quad C_{A\varphi } \\ 0 &{} \quad D_{A\varphi } \end{array} \right) \left( \begin{array}{c} 1\ \\ q^{\star }_{1} \end{array} \right) e^{-i\tau _{0}\omega _{0}}= -i\tau _{0}\omega _{0}\left( \begin{array}{c} 1\ \\ q^{\star }_{1} \end{array} \right) .\)

This implies \( \left\{ \begin{array}{ll} A_{A\varphi }+C_{A\varphi }q^{\star }_{1}e^{-i\tau _{0}\omega _{0}} &{} = -i\omega _{0},\\ B_{A\varphi }+D_{A\varphi }q^{\star }_{1})e^{-i\tau _{0}\omega _{0}} &{} = -q_{1}^{\star }i\omega _{0}, \end{array}\right. \) and    \(q^{\star }_{1}= \displaystyle -\frac{B_{A\varphi }}{D_{A\varphi } + i\omega _{0}e^{i\tau _{0}\omega _{0}}}.\)

Now let’s compute \(G_{A\varphi }\) by using the orthogonality condition \(\langle q^{\star }(s), q(\theta ) \rangle = 1\). By using (30), we have:

$$\begin{aligned}&\langle q^{\star }(s), q(\theta )\rangle \\&\quad = \overline{q^{\star }}(0).q(0)- \displaystyle \int _{\theta =-1}^{0}\displaystyle \int _{\xi =0}^{\theta }\overline{q^{\star }}^{T}\\&\qquad (\xi -\theta )\mathrm{d}\eta (\theta )q(\xi )\mathrm{d}\xi \\&\quad = \overline{q^{\star }}(0).q(0)- \displaystyle \int _{\theta =-1}^{0}\displaystyle \int _{\xi =0}^{\theta }\overline{G_{A\varphi }}(1,\overline{q_{1}^{\star }}) e^{-i\omega _{0}\tau _{0}(\xi -\theta )}\\&\qquad \mathrm{d}\eta (\theta )(1,q_{1})^{T} e^{i\omega _{0}\tau _{0}\xi } \\&\quad = \overline{G_{A\varphi }}\left\{ \displaystyle 1+q_{1}\overline{q_{1}^{\star }}-(1,\overline{q_{1}^{\star }})\right. \\&\qquad \left. \int _{-1}^{0}\theta e^{i\omega _{0}\tau _{0}\theta } \mathrm{d}\eta (\theta )(1,q_{1})^{T}\right\} \\&\quad = \overline{G_{A\varphi }}\left\{ \displaystyle 1+q_{1}\overline{q_{1}^{\star }}+ \tau _{0}e^{i\omega _{0}\tau _{0}}(C_{A\varphi }+D_{A\varphi }q_{1})\overline{q_{1}}^{\star } \right\} . \end{aligned}$$

Then,

$$\begin{aligned} G_{A\varphi } = [1+\overline{q_{1}}q_{1}^{\star }+ \tau _{0}e^{-i\omega _{0}\tau _{0}}(C_{A\varphi } + D_{A\varphi } \overline{q_{1}})q_{1}^{\star }]^{-1}. \end{aligned}$$

Now we are going to compute the coordinates to describe the center manifold \(\mathcal {C}_0\) at \(\mu =0\). Let us define

$$\begin{aligned} z(t) = \langle q^{*},u_{t} \rangle \end{aligned}$$
(32)

and

$$\begin{aligned} W(t, \theta )= u_{t}(\theta )-2Re\{z(t)q(\theta )\}, \end{aligned}$$
(33)

where \(u_{t}\) is a solution of (28) when \(\mu =0\).

On the center manifold \(\mathcal {C}_{0}\), we have:

$$\begin{aligned} W(t, \theta )= W(z(t), \overline{z}(t), \theta ), \end{aligned}$$

where

$$\begin{aligned} W(t, \theta )= & {} W(z(t), \overline{z}(t), \theta )= W_{20}(\theta )\frac{z^{2}}{2}+W_{11}(\theta )z\overline{z}\nonumber \\&+W_{02}(\theta )\frac{\overline{z}^{2}}{2}+\cdots \end{aligned}$$
(34)

In (33), z and \(\overline{z}\) represent the local coordinates of the center manifold \(\mathcal {C}_{0}\) in the direction of q and \(q^{*}\), respectively.

Now let us reduce equation (28) to an ordinary differential equation using a single complex variable on the center manifold. Since \(\mu =0\) and for \(u_{t}\) a solution of (28) belonging to the center manifold \(\mathcal {C}_{0}\), we have:

$$\begin{aligned} \dot{z}(t)= & {} \langle q^{*},\dot{u}(t) \rangle =\langle q^{*},Au_{t}+Ru_{t} \rangle \\= & {} \langle q^{*},Au_{t}\rangle +\langle q^{*},Ru_{t}\rangle \\= & {} \langle A^{*} q^{*},u_{t}\rangle +\langle q^{*},Ru_{t}\rangle \\= & {} i\tau _{0}\omega _{0}z+\overline{q}^{*}f_{0}(z,\overline{z}). \end{aligned}$$

The following equation:

$$\begin{aligned} \dot{z}(t)= i\tau _{0}\omega _{0}z+\overline{q}^{*}f_{0}(z,\overline{z}) \end{aligned}$$
(35)

can be rewritten as:

$$\begin{aligned} \dot{z}(t)=i\tau _{0}\omega _{0}z+g(z,\overline{z}), \end{aligned}$$
(36)

where

$$\begin{aligned} g(z,\overline{z})= g_{20}\frac{z^{2}}{2}+ g_{11}z\overline{z}+g_{02}\frac{\overline{z}^{2}}{2}+g_{21}\frac{z^{2}\overline{z}}{2}+\cdots \end{aligned}$$
(37)

In what follows, we are going to expand g in powers of z and \(\overline{z}\) in order to obtain from the first three coefficients of this expansion, the value of \(\mu _2\) which indicates the direction of the Hopf bifurcation (that is to say if the Hopf bifurcation is supercritical or subcritical) and the value of \(\beta _2\) which determines the stability. To do so, we use the algorithm presented by Hassard et al. [26].

A substitution of (28) in (36) leads to:

$$\begin{aligned} \dot{W}(t)= & {} \dot{u_{t}}-\dot{z}q-\dot{\overline{z}}\overline{q}\\= & {} A(\mu )u_{t}+Ru_{t}-[i\tau _{0}\omega _{0}z+\overline{q}^{*}(0)f_{0}(z,\overline{z})]q\\&- [-i\tau _{0}\omega _{0}\overline{z}+{q}^{*}(0)\overline{f}_{0}(z,\overline{z})]\overline{q}\\= & {} AW+2ARe(zq)+Ru_{t}-2Re[\overline{q}^{*}(0)f_{0}(z,\overline{z})q(\theta )]\\&-2Re[i\tau _{0}\omega _{0}zq(\theta )]\\= & {} AW-2Re[\overline{q}^{*}(0)f_{0}(z,\overline{z})q(\theta )]+Ru_{t}. \end{aligned}$$

Then we obtain the following equations

$$\begin{aligned} \dot{W}=\left\{ \begin{array}{ll} AW-2Re[\overline{q}^{*}(0)f_{0}(z,\overline{z})q(\theta )]&{}\quad \text{ if }\quad -1\le \theta <0, \\ AW-2Re[\overline{q}^{*}(0)f_{0}(z,\overline{z})q(\theta )] + f &{}\quad \text{ if }\quad \theta =0. \end{array}\right. \nonumber \\ \end{aligned}$$
(38)

Equations (37) can be rewritten as:

$$\begin{aligned} \dot{W}=AW+H(z,\overline{z},\theta ), \end{aligned}$$
(39)

where

$$\begin{aligned} H(z,\overline{z},\theta )= & {} H_{20}(\theta )\frac{z^{2}}{2}+H_{11}(\theta )z\overline{z}\nonumber \\&+H_{02}(\theta )\frac{\overline{z}^{2}}{2}+\cdots \end{aligned}$$
(40)

Derivating W in (33) with respect to t, we have:

$$\begin{aligned} \dot{W}=W_{z}\dot{z}+W_{\overline{z}}\dot{\overline{z}}. \end{aligned}$$
(41)

Then using (33), (38) and (39) gives:

$$\begin{aligned} (A-2i\tau _{0}\omega _{0})W_{20}(\theta )=-H_{20}(\theta ), \end{aligned}$$
(42)

and

$$\begin{aligned} AW_{11}(\theta )=-H_{11}(\theta ). \end{aligned}$$
(43)

Using equations (34) and (35) gives:

$$\begin{aligned} g(z,\overline{z})= & {} \overline{q}^{*}(0)f_{0}(z,\overline{z})= \tau _{0}\overline{G_{A\varphi }}\left( \begin{array}{c} 1\ \\ q^{\star }_{1}\ \\ \end{array} \right) \left( \begin{array}{c} f_{1}\ \\ f_{2} \end{array} \right) \nonumber \\= & {} \tau _{0}\overline{G_{A\varphi }}(f_{1}+q^{\star }_{1}f_{2}), \end{aligned}$$
(44)

where

$$\begin{aligned} f_{1}= & {} a_{11}x_{1}^{2}(t)+a_{12}x_{1}(t)x_{2}(t)+a_{13}x_{1}^{3}(t), \\ f_{2}= & {} a_{21}x_{1}^{2}(t-1)+a_{22}x_{1}(t-1)x_{2}(t)\\&+ a_{23}x_{1}(t-1)x_{2}(t-1)+a_{24}x_{2}(t-1)x_{2}(t)\\&+ a_{25}x_{1}^{3}(t-1) \\&+ a_{26}x_{1}^{2}(t-1)x_{2}(t)+a_{27}x_{1}^{2}(t-1)x_{2}(t-1)\\&+ a_{28}x_{1}(t-1)x_{2}(t)x_{2}(t-1), \end{aligned}$$

and

$$\begin{aligned} x_{1t}(0)= & {} q^{(1)}(0)z+\overline{q}^{(1)}(0)\overline{z}+W_{20}^{(1)}(0)\frac{z^{2}}{2}\\&+W_{11}^{(1)}(0)z\overline{z}+ W_{02}^{(1)}(0)\frac{\overline{z}^{2}}{2}+ \cdots \\ x_{2t}(0)= & {} q^{(2)}(0)z+\overline{q}^{(2)}(0)\overline{z}+W_{20}^{(2)}(0)\frac{z^{2}}{2}\\&+W_{11}^{(2)}(0)z\overline{z}+ W_{02}^{(2)}(0)\frac{\overline{z}^{2}}{2}+\cdots \\ x_{1t}(-1)= & {} q^{(1)}(-1)z+\overline{q}^{(1)}(-1)\overline{z}+W_{20}^{(1)}(-1)\frac{z^{2}}{2}\\&+W_{11}^{(1)}(-1)z\overline{z}+ W_{02}^{(1)}(-1)\frac{\overline{z}^{2}}{2}+\cdots \\ x_{2t}(-1)= & {} q^{(2)}(-1)z+\overline{q}^{(2)}(-1)\overline{z}+W_{20}^{(2)}(-1)\frac{z^{2}}{2}\\&+W_{11}^{(2)}(-1)z\overline{z}+ W_{02}^{(2)}(-1)\frac{\overline{z}^{2}}{2}+ \cdots \end{aligned}$$

An identification by using (34) and (43) gives us the following coefficients of g:

$$\begin{aligned} g_{20}= & {} 2\tau _{0}\overline{G_{A\varphi }}\{ a_{11}(q^{(1)}(0))^{2}+a_{12}q^{1}(0)q^{(2)}(0)\\&+ q_{1}^{*}(a_{21}(q^{(1)}(-1))^{2} +a_{22}q^{(1)}(-1)q^{(2)}(0)\\&+ a_{23}q^{(1)}(-1)q^{(2)}(-1)+a_{24}q^{(2)}(0)q^{(2)}(-1)) \}, \\ g_{11}= & {} \tau _{0}\overline{G_{A\varphi }}\{2a_{11}q^{(1)}(0)\overline{q}^{(1)}(0)\\&+ a_{12}(q^{(1)}(0)q^{(2)}(0)+\overline{q}^{(1)}(0)q^{(2)}(0))\\&+ q_{1}^{*}(2a_{21}q^{(1)}(-1)\overline{q}^{(1)}(-1)\\&+ a_{22}(q^{(1)}(-1)\overline{q}^{(2)}(0)+q^{(2)}(0)\overline{q}^{(1)}(-1))\\&+ a_{23}(q^{(1)}(-1)\overline{q}^{(2)}(-1)+q^{(2)}(-1)\overline{q}^{(1)}(-1))\\&+ a_{24}(q^{(2)}(-1)\overline{q}^{(2)}(0)+q^{(2)}(0)\overline{q}^{(2)}(-1)))\},\\ g_{02}= & {} 2\tau _{0}\overline{G_{A\varphi }}\{a_{11}(\overline{q}^{(1)}(0))^{2}\\&+ a_{21}\overline{q}^{1}(0)\overline{q}^{(2)}(0)+ q_{1}^{*}(a_{21}(\overline{q}^{(2)}(-1))^{2}\\&+ a_{22}\overline{q}^{(1)}(-1)\overline{q}^{(2)}(0)\\&+ a_{23}\overline{q}^{(1)}(-1)\overline{q}^{(2)}(-1)\\&+ a_{24}\overline{q}^{(2)}(0)\overline{q}^{(2)}(-1))\}, \\ g_{21}= & {} 2\tau _{0}\overline{G_{A\varphi }}\{a_{11}(\overline{q}^{(1)}(0)W_{20}^{(1)}(0)\\&+ 2q^{(1)}(0)W_{11}^{(1)}(0))+ a_{12}(q^{(1)}(0)W_{11}^{(2)}(0)\\&+ \frac{1}{2}\overline{q}^{(1)}(0)W_{20}^{(2)}(0)\\&+ \frac{1}{2}\overline{q}^{(2)}(0)W_{20}^{(1)}(0)\\&+ \overline{q}^{(2)}(0)W_{11}^{(1)}(0))+3a_{13}(q^{(1)}(0))^{2}\overline{q}^{(1)}(0)\\&+ q_{1}^{*}(a_{21}(\overline{q}^{(1)}(-1)W_{20}^{(1)}(-1)\\&+ 2q^{(1)}(-1)W_{11}^{(1)}(-1))+a_{22}(q^{(1)}(-1)W_{11}^{(2)}(0)\\&+ \frac{1}{2}\overline{q}^{(1)}(-1)W_{20}^{(2)}(0)\\&+ \frac{1}{2}\overline{q}^{(2)}(0)W_{20}^{(1)}(-1)\\&+ q^{(2)}(0)W_{11}^{(1)}(-1))+a_{23}(q^{(1)}(-1)W_{11}^{(2)}(-1)\\&+ \frac{1}{2} \overline{q}^{(1)}(-1)W_{20}^{(2)}(-1)+q^{(2)}(-1)W_{11}^{(1)}(-1)\\&+ \frac{1}{2}\overline{q}^{(2)}(-1)W_{20}^{(1)}(-1))+a_{24}(q^{(2)}(0)W_{11}^{(2)}(-1)\\&+ \frac{1}{2}\overline{q}^{(2)}(0)W_{20}^{(2)}(-1)+q^{(2)}(-1)W_{11}^{(2)}(0)\\&+\frac{1}{2}\overline{q}^{(2)}(-1)W_{20}^{(2)}(0))\\&+ 3a_{25}(q^{(1)}(-1))^{2}\overline{q}^{(1)}(-1)\\&+ a_{26}((q^{(1)}(-1))^{2}\overline{q}^{(2)}(0)\\&+ 2q^{(1)}(-1)\overline{q}^{(1)}(-1)q^{(2)}(0))\\&+ a_{27}((q^{(1)}(-1))^{2}\overline{q}^{(2)}(-1)\\&+ 2q^{(1)}(-1)\overline{q}^{(1)}(-1)q^{(2)}(-1))\\&+ a_{28}(q^{(1)}(-1)\overline{q}^{(2)}(0)q^{(2)}(-1))\\&+ q^{(1)}(-1)\overline{q}^{(2)}(-1)q^{(1)}(0)\\&+ q^{(2)}(-1)\overline{q}^{(1)}(-1)q^{(2)}(0))\}. \end{aligned}$$

Now we reach to the step of computation of \(W_{20}(\theta )\). Using (27) and (38) gives:

$$\begin{aligned} H(z,\overline{z}, \theta )= & {} -2Re[\overline{q}^{*}(0)f_{0}(z,\overline{z})q(\theta )]+Ru_{t}\\= & {} -gq(\theta )-\overline{g}\overline{q}(\theta )+Ru_{t} \end{aligned}$$
$$\begin{aligned}= & {} -\left( \displaystyle g_{20}\frac{z^{2}}{2}+g_{11}z\overline{z}+g_{02}\frac{\overline{z}^{2}}{2}\right) q(\theta )\nonumber \\&-\left( \displaystyle \overline{g}_{20}\frac{\overline{z}^{2}}{2}+\overline{g}_{11}z\overline{z}+\overline{g}_{02}\frac{{z}^{2}}{2}\right) \overline{q}(\theta )+Ru_{t}, \end{aligned}$$
(45)

where

$$\begin{aligned} Ru_{t}= \tau _{0} \left( \begin{array}{c} f_{1}\ \\ f_{2} \end{array} \right) =2\tau _{0}A^{A\varphi }\frac{z^{2}}{2}+\tau _{0}B^{A\varphi }z\overline{z}, \end{aligned}$$

with

$$\begin{aligned} A^{A\varphi }= & {} \left( \begin{array}{c} A_{11}^{A\varphi }\ \\ A_{21}^{A\varphi } \end{array} \right) , \\ B^{A\varphi }= & {} \left( \begin{array}{c} B_{11}^{A\varphi }\ \\ B_{21}^{A\varphi } \end{array} \right) , \end{aligned}$$

where

$$\begin{aligned} A_{11}^{A\varphi }= & {} a_{11}(q^{(1)}(0))^{2}+a_{12}q^{(1)}(0)q^{(2)}(0), \\ A_{21}^{A\varphi }= & {} a_{21}(q^{(1)}(-1))^{2}+a_{22}q^{(1)}(-1)q^{(2)}(0)\\&+ a_{23}q^{(1)}(-1)q^{(2)}(-1)\\&+ a_{24}q^{(2)}(-1)q^{(2)}(0), \\ B_{11}^{A\varphi }= & {} 2a_{11}q^{(1)}(0)\overline{q}^{(1)}(0)\\&+ a_{12}(q^{(1)}(0)\overline{q}^{(2)}(0)+\overline{q}^{(1)}(0){q}^{(2)}(0)), \\ B_{21}^{A\varphi }= & {} 2a_{21}q^{(1)}(-1)\overline{q}^{(1)}(-1)\\&+ a_{22}(q^{(1)}(-1)\overline{q}^{(2)}(0)+\overline{q}^{(1)}(-1){q}^{(2)}(0)) \\&+ a_{23}(q^{(1)}(-1)\overline{q}^{(2)}(-1)+\overline{q}^{(1)}(-1){q}^{(2)}(-1))\\&+ a_{24}(q^{(2)}(-1)\overline{q}^{(2)}(0)+\overline{q}^{(2)}(-1){q}^{(2)}(0)). \end{aligned}$$

A comparison of the coefficients of equations (39) and (44) gives the following equalities:

$$\begin{aligned} H_{20}(\theta )= & {} \left\{ \begin{array}{ll} -gq(\theta )-\overline{g}\overline{q}(\theta )&{}\quad \text{ if }\quad -1\le \theta <0, \\ -gq(\theta )-\overline{g}\overline{q}(\theta )+2\tau _{0}A &{}\quad \text{ if }\quad \theta =0. \end{array}\right. \nonumber \\ \end{aligned}$$
(46)
$$\begin{aligned} H_{11}(\theta )= & {} \left\{ \begin{array}{ll} -gq(\theta )-\overline{g}\overline{q}(\theta )&{}\quad \text{ if }\quad -1\le \theta <0, \\ -gq(\theta )-\overline{g}\overline{q}(\theta )+\tau _{0}B &{}\quad \text{ if }\quad \theta =0. \end{array}\right. \nonumber \\ \end{aligned}$$
(47)

When we substitute (45) in (41) and (46) in (42), respectively, we obtain the following differential equations:

$$\begin{aligned} \left\{ \begin{array}{l} \dot{W}_{20}(\theta )= 2i\tau _{0}\omega _{0}W_{20}(\theta )+g_{20}q(\theta )+\overline{g}_{20}\overline{q}(\theta ), \\ \dot{W}_{11}(\theta )=g_{11}q(\theta )+\overline{g}_{11}\overline{q}(\theta ), \end{array}\right. \nonumber \\ \end{aligned}$$
(48)

which have the following solutions:

$$\begin{aligned} \left\{ \begin{array}{l} {W}_{20}(\theta )= \displaystyle \frac{ig_{20}}{\tau _{0}\omega _{0}}q(0)e^{i\tau _{0}\omega _{0}\theta }+ \frac{\overline{g}_{02}}{3i\tau _{0}\omega _{0}}\overline{q}(0)e^{-i\tau _{0}\omega _{0}\theta }\\ \qquad \quad +E_{1A\varphi }e^{2i\tau _{0} \omega _{0}\theta } , \\ {W}_{11}(\theta )= \displaystyle \frac{g_{11}}{i\tau _{0}\omega _{0}}q(0)e^{i\tau _{0}\omega _{0}\theta }+ \frac{\overline{g}_{11}}{i\tau _{0}\omega _{0}}\overline{q}(0)e^{-i\tau _{0}\omega _{0}\theta }\\ \qquad +E_{2A\varphi }, \end{array}\right. \nonumber \\ \end{aligned}$$
(49)

where \(E_{1A\varphi }=(E_{1A\varphi }^{(1)},E_{1A\varphi }^{(2)})^{T}\) and \(E_{2A\varphi }=(E_{2A\varphi }^{(1)},E_{2A\varphi }^{(2)})^{T}\) are constant vectors belonging to \({\mathbb {R}}^{2}\).

Now, let us compute the constant vectors \(E_{1A\varphi }\) and \(E_{2A\varphi }\) by using (41) and (42). We have:

$$\begin{aligned} \int _{-1}^{0}\mathrm{d}\eta (\theta )W_{20}(\theta )=2i\tau _{0}\omega _{0}W_{20}(0)-H_{20}(0), \end{aligned}$$
(50)

and

$$\begin{aligned} \int _{-1}^{0}\mathrm{d}\eta (\theta )W_{11}(\theta )=-H_{11}(0). \end{aligned}$$
(51)

Then, using the fact that,

$$\begin{aligned}&\left( \displaystyle i\tau _{0}\omega _{0}I-\int _{-1}^{0}e^{i\tau _{0}\omega _{0}\theta }\mathrm{d}\eta (\theta )\right) q(0)=0, \\&\left( \displaystyle -i\tau _{0}\omega _{0}I-\int _{-1}^{0}e^{-i\tau _{0}\omega _{0}\theta }\mathrm{d}\eta (\theta )\right) \overline{q}(0)=0, \end{aligned}$$

and using the first equation of (48), equations (44) and (49), we have:

$$\begin{aligned} \left( \displaystyle 2i\tau _{0}\omega _{0}I-\int _{-1}^{0}e^{2i\tau _{0}\omega _{0}\theta }\mathrm{d}\eta (\theta )\right) E_{1}=2\tau _{0}\left( \begin{array}{c} A_{11}^{A\varphi }\ \\ A_{21}^{A\varphi } \end{array} \right) \end{aligned}$$

which implies \(\left( \begin{array}{cc} 2i\omega _{0}-A_{A\varphi } &{}\quad -B_{A\varphi }\\ -C_{A\varphi }e^{-2i\tau _{0}\omega _{0}} &{}\quad 2i\omega _{0}-D_{A\varphi }e^{-2i\tau _{0}\omega _{0}} \\ \end{array}\right) \left( \begin{array}{c} E_{1A\varphi }^{(1)}\ \\ E_{1A\varphi }^{(2)} \end{array} \right) =2\left( \begin{array}{c} A_{11}^{A\varphi }\ \\ A_{21}^{A\varphi } \end{array} \right) \) and finally

$$\begin{aligned} \left\{ \begin{array}{l} E_{1A\varphi }^{(1)}= \displaystyle \frac{2B_{A\varphi }A_{21}^{A\varphi }+2A_{11}^{A\varphi }(2i\omega _{0}-D_{A\varphi }e^{-2i\tau _{0}\omega _{0}})}{(A_{A\varphi }D_{A\varphi }-B_{A\varphi }C_{A\varphi })e^{-2i\tau _{0}\omega _{0}}-2i\omega _{0}D_{A\varphi } e^{-2i\tau _{0}\omega _{0}}-2i\omega _{0}A_{A\varphi }-4\omega _{0}^{2}} , \\ E_{1A\varphi }^{(2)}= \displaystyle \frac{2A_{11}^{A\varphi }C_{A\varphi }e^{-2i\tau _{0}\omega _{0}}+2A_{21}^{A\varphi }(2i\omega _{0}-A_{A\varphi })}{(A_{A\varphi }D_{A\varphi }-B_{A\varphi }C_{A\varphi })e^{-2i\tau _{0}\omega _{0}}-2i\omega _{0}D_{A\varphi } e^{-2i\tau _{0}\omega _{0}}-2i\omega _{0}A_{A\varphi }-4\omega _{0}^{2}}. \end{array}\right. \end{aligned}$$
(52)

Similarly, using the second equation of (48), (46) and (50) gives:

$$\begin{aligned}&\left( \displaystyle \int _{-1}^{0}e^{-i\tau _{0}\omega _{0}}\mathrm{d}\eta (\theta )\right) E_{2A\varphi }=\tau _{0}\left( \begin{array}{c} B_{11}^{A\varphi }\ \\ B_{21}^{^{A\varphi }} \end{array} \right) \\&\quad \Rightarrow \left( \begin{array}{cc} A_{A\varphi } &{} B_{A\varphi } \\ C_{A\varphi } &{} D_{A\varphi } \end{array}\right) \left( \begin{array}{c} E_{2A\varphi }^{(1)}\ \\ E_{2A\varphi }^{(2)} \end{array} \right) =-\left( \begin{array}{c} B_{11}^{A\varphi }\ \\ B_{21}^{A\varphi } \end{array} \right) , \end{aligned}$$

and finally

$$\begin{aligned} \left\{ \begin{array}{l} E_{2A\varphi }^{(1)}= \displaystyle \frac{B_{A\varphi }B_{21}^{A\varphi }-D_{A\varphi }B_{11}^{A\varphi }}{A_{A\varphi }D_{A\varphi } - B_{A\varphi }C_{A\varphi }}, \\ E_{2A\varphi }^{(2)}= \displaystyle \frac{C_{A\varphi }B_{11}^{A\varphi }-A_{A\varphi }B_{21}^{A\varphi }}{A_{A\varphi }D_{A\varphi } - B_{A\varphi }C_{A\varphi }}. \end{array}\right. \end{aligned}$$
(53)

Finally, from (48), (50) and (52) we can now calculate values which are useful for the determination of the period of the bifurcating solution and its stability. We have the following values:

$$\begin{aligned} C_{1}(0)= & {} \frac{i}{2\tau _{0}\omega _{0}}\left( \displaystyle g_{20}g_{11}-2|g_{11}|^{2}-\frac{1}{3}|g_{02}|^{2}\right) +\frac{1}{2}g_{21}, \\ \mu _{2}= & {} -\frac{ReC_{1}(0)}{Re(\frac{\mathrm{d}\lambda }{\mathrm{d}\tau }(\tau _{0}))}, \\ \beta _{2}= & {} 2ReC_{1}(0), \\ T_{2}= & {} -\frac{Im(C_{1}(0))+\mu _{2}Im(\frac{\mathrm{d}\lambda }{\mathrm{d}\tau }(\tau _{0}))}{\tau _{0}\omega _{0}}. \end{aligned}$$

This ends the proof.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Onana, M., Mewoli, B. & Tewa, J.J. Hopf bifurcation analysis in a delayed Leslie–Gower predator–prey model incorporating additional food for predators, refuge and threshold harvesting of preys. Nonlinear Dyn 100, 3007–3028 (2020). https://doi.org/10.1007/s11071-020-05659-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-020-05659-7

Keywords

Navigation