Skip to main content
Log in

A delayed eco-epidemiological model with nonlinear incidence rate and Crowley–Martin functional response for infected prey and predator

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

Abstract

A prey–predator model with disease in prey population is considered here. The prey population is classified into two categories: susceptible prey and infected prey. The nonlinear incidence rate, at which the susceptible prey becomes infected, is incorporated. Here, predator consumes infected prey only according to Crowely–Martin functional response. Time lag is introduced to convert the susceptible prey to infected one, and the present model becomes delayed eco-epidemic model. The positivity, boundedness and permanence of the solutions of both models are guaranteed here. The existence of equilibrium points is assured. Local as well as global stability analyses of both the models around the equilibrium points are performed and finally summarized in a table, comparing the conditions between two models. The occurrence of bifurcation (saddle-node, transcritical, Hopf and Bogdanov–Takens) in the non-delay model is explored here. The nature and direction of Hopf bifurcation are searched using normal form and centre manifold theorems for delay model. Lastly, the numerical simulation is carried out to verify the analytical findings. An effort is made to search the parameters, by paying special attention to newly introduced parameters, which control the dynamics of both models. Accordingly, stable/unstable zones on \(x\tau _\mathrm{c}\)-planes, where x is any considered parameter and \(\tau _\mathrm{c}\) is the value of critical delay time, are proposed. The stranger attractor is searched in the delay model. Bifurcated diagrams of population with respect to newly introduced parameters are exemplified and discussed. When the population (S, I and P) starts oscillating, their time-averaged values are presented to show their variation due to aforesaid parameters. Moreover, pattern of all the curves \(S=S(x)\), \(I=I(x)\) and \(P=P(x)\), where x is any aforesaid parameters, are certified using their respective gradient.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

References

  1. Agiza, H.N., ELabbasy, E.M., EL-Metwally, H., Elsadany, A.A.: Chaotic dynamics of a discrete prey–predator model with Holling type II. Nonlinear Anal. Real World Appl. 10, 116–129 (2009)

    Article  MathSciNet  Google Scholar 

  2. Biswas, S., Sasmal, S.K., Samanta, S., Saifuddin, Md, Pal, N., Chattopadhyay, J.: Optimal harvesting and complex dynamics in a delayed eco-epidemiological model with weak Allee effects. Nonlinear Dyn. 87, 1553–1573 (2017)

    Article  Google Scholar 

  3. Celik, C., Merdan, H.: Hopf bifurcation analysis of a system of coupled delayed-differential equations. Appl. Math. Comput. 219, 6605–6617 (2013)

    MathSciNet  MATH  Google Scholar 

  4. Chakraborty, S., Pal, S., Bairagi, N.: Dynamics of a ratio-dependent eco-epidemiological system with prey harvesting. Nonlinear Anal. RWA. 11, 1862–1877 (2010)

    Article  MathSciNet  Google Scholar 

  5. Chakraborty, K., Das, K., Haldar, S., Kar, T.K.: A mathematical study of an eco-epidemiological system on disease persistence and extinction perspective. Appl. Math. Comput. 254, 99–112 (2015)

    MathSciNet  MATH  Google Scholar 

  6. Chattopadhyay, J., Bairagi, N.: Pelicans at risk in Salton sea an eco-epidemiological study. Ecol. Model. 135, 103–112 (2001)

    Article  Google Scholar 

  7. Gumel, A.B., Moghadas, S.M.: A qualitative study of a vaccination model with non-linear incidence. Appl. Math. Comput. 143, 409–419 (2003)

    MathSciNet  MATH  Google Scholar 

  8. Haldar, S., Chakraborty, K., Das, K., Kar, T.K.: Bifurcation and control of an eco-epidemiological system with environmental fluctuations: a stochastic approach. Nonlinear Dyn. 80, 1187–1207 (2015)

    Article  MathSciNet  Google Scholar 

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

    MATH  Google Scholar 

  10. Hu, G.P., Li, X.L.: Stability and Hopf bifurcation for a delayed predator-prey model with disease in the prey. Chaos, Solitons & Fractrals. 45, 229–237 (2012)

    Article  MathSciNet  Google Scholar 

  11. Hu, Z., Teng, Z., Zhang, T., Zhou, Q., Chen, Xi: Globally asymptotically stable analysis in a discrete time eco-epidemiological system. Chaos, Solitons Fractals 99, 20–31 (2017)

    Article  MathSciNet  Google Scholar 

  12. Jana, S., Kar, T.K.: A mathematical study of a prey-predator model in relevance to pest control. Nonlinear Dyn. 74, 667–683 (2013)

    Article  MathSciNet  Google Scholar 

  13. Jana, S., Kar, T.K.: Modeling and analysis of a prey-predator system with disease in the prey. Chaos, Solitons Fractrals 47, 42–53 (2013)

    Article  MathSciNet  Google Scholar 

  14. Jana, S., Guria, S., Das, U., Kar, T.K., Ghorai, A.: Effect of harvesting and infection on predator in a prey-predator system. Nonlinear Dyn. 81, 917–930 (2015)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  16. Li, G., Wang, W.: Bifurcation behaviour of an epidemic model with a nonlinear incidence. Appl. Math. Comput. 214, 411–423 (2009)

    MathSciNet  MATH  Google Scholar 

  17. Liu, X.: Bifurcation of an eco-epidemiological model with a non-linear incidence rate. Appl. Math. Comput. 218, 2300–2309 (2011)

    Article  MathSciNet  Google Scholar 

  18. Liu, W-m, Levin, S.A., Iwasa, Y.: Influence of nonlinear incidence rates upon the behaviour of SIRS epidemiological models. J. Math. Biol. 23(2), 187–204 (1986)

    Article  MathSciNet  Google Scholar 

  19. Liu, W-m, Hethcote, H.W., Levin, S.A.: Dynamical behaviour of epidemiological models with non-linear incidence rates. J. Math. Biol. 25, 359–380 (1987)

    Article  MathSciNet  Google Scholar 

  20. Perko, L.: Differential Equations and Dynamical Systems, 3rd edn. Springer, New York (2004)

    MATH  Google Scholar 

  21. Prasad, K.D., Prasad, B.S.R.V.: Qualitative analysis of additional food provided predator-prey system with anti-predator behaviour in prey. Nonlinear Dyn. 96, 1765–1793 (2019)

    Article  Google Scholar 

  22. Roy, P., Upadhyay, R.K.: Assessment of rabbit hemorrhagic disease in controlling the population of red fox: a measure to preserve endangered species in Australia. Ecol. Complex. 26, 6–20 (2016)

    Article  Google Scholar 

  23. Ruan, S., Wang, W.: Dynamical behaviour of an epidemic model with a non-linear incidence rate. J. Differ. Equ. 188, 135–163 (2003)

    Article  Google Scholar 

  24. Sahoo, B.: Role of additional food in eco-epidemiological system with disease in the prey. Appl. Math. Comput. 259, 61–79 (2015)

    MathSciNet  MATH  Google Scholar 

  25. Tripathi, J.P., Tyagi, S., Abbas, S.: Global analysis of a delayed density dependent predator-prey model with Crowley–Martin functional response. Common Nonlinear Sci. Numer. Simul. 30, 45–69 (2016)

    Article  MathSciNet  Google Scholar 

  26. Upadhyay, R.K., Roy, P.: Spread of a disease and its effects on population dynamics in an eco-epidemiological system. Commun. Nonlinear Sci. Numer. Simul. 19, 4170–4184 (2014)

    Article  MathSciNet  Google Scholar 

  27. Wang, N., Zhao, M., Yu, H., Dai, C., Wang, B., Wang, P.: Bifurcation behaviour analysis in a predator-prey model. Hindawi Publ. Corp. 2016, 1–11 (2016)

    Google Scholar 

  28. Zhang, J.-F., Li, W.-T., Yan, X.-P.: Hopf bifurcation and stability of periodic solutions in a delayed eco-epidemiological system. Appl. Math. Comput. 198, 865–876 (2008)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Authors A. P. Maiti and D. K. Maiti acknowledge the support received from WOS-A, DST(INDIA) (Grant No. SR/WOS-A/PM-1059/2015), and DST-SERB (INDIA) (San No. SR/S4/MS/820/13 dated 07.05.2015), respectively.

Funding

WOS-A, DST(INDIA) (Grant No. SR/WOS-A/PM-1059/2015), and DST-SERB (INDIA) (San No. SR/S4/MS/820/13 dated 07.05.2015).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dilip Kumar Maiti.

Ethics declarations

Ethical approval

This article does not contain any studies with human participants or animals performed by any of the authors.

Additional information

Publisher's Note

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

Appendices

Appendix A

Consider the following positive definite function about the equilibrium point \(E^*\)

$$\begin{aligned} V(S,I,P)&=\left[ S-S^*-S^*\ln \frac{S}{S^*}\right] \\&\quad +C_1\left[ I-I^*-\ln \frac{I}{I^*}\right] \\&\quad +C_2\left[ P-P^*-P^*\ln \frac{P}{P^*}\right] . \end{aligned}$$

where \(C_1\) and \(C_2\) are constants. Differentiating V w.r.t. time t along the solution of the system and while choosing \(C_1= \frac{1}{1+\alpha I^*} \) and \(C_2=\frac{m(1+aI^*)}{m_1(1+bP^*)(1+\alpha I^*)}\), then we have

$$\begin{aligned} \frac{dV}{\mathrm{d}t}= & {} -\left[ b_{11}(S-S^*)^2+b_{22}(I-I^*)^2\right. \\&\left. +\,b_{33}(P-P^*)^2+b_{12}(S-S^*)(I-I^*)\right] . \end{aligned}$$

where \(b_{11}=\frac{r}{k},\ b_{22}=C_1\left[ \frac{\alpha \beta S^*}{(1+\alpha I)(1+\alpha I^*)}- \frac{amP^*}{(1+aI)(1+aI^*)(1+bp^*)}\right] ,\ b_{33}= \left[ \frac{m_1bI^*}{(1+aI^*)(1+bP)(1+bP^*)}+e\right] C_2\) and \(\ b_{12}=\frac{r}{k}\). Clearly, \(\frac{dV}{\mathrm{d}t}=0\) if and only if \(S=S^*\), \(I=I^*\) and \(P=P^*\). The sufficient condition for \(\frac{dV}{\mathrm{d}t}\) to be negative definite around \(E_3(S^*,I^*,P^*)\) is that \( \ b_{22}>0\) and \( \ 4b_{11}b_{22}>b_{12}^2\). These conditions are satisfied under condition of Theorem 4.5. Therefore, \(E_3(S^*,I^*,P^*)\) is globally asymptotically stable under condition of Theorem 4.5.

Appendix B

$$\begin{aligned} {\bar{L}}_1&=\left[ \frac{2\alpha \beta ^2 K^*I^*}{r(1+\alpha I^*)^3}+2am{\bar{L}}_4\left\{ P^*+\frac{m_1I^*}{{\bar{L}}_3} \right\} \right. \\&\quad +\frac{2bmm_1^2I^*{\bar{L}}_4}{{\bar{L}}_3^2}\left\{ 1+\frac{m_1I^*}{P^*{{\bar{L}}_3}^3}\right\} \\&\quad \left. + \frac{2emm_1^2I^*(1+bP^*)^2}{P^*{\bar{L}}_3^3(1+aI^*)^2} \right] ,\\ {\bar{L}}_2&=\left[ \frac{2\beta K^*}{r(1+\alpha I^*)^2}\left\{ \frac{r}{K^*}+\frac{\beta }{(1+\alpha I^*)^2} \right\} \right. \\&\quad \left. +\frac{2\alpha \beta S^*}{(1+\alpha I^*)^3}+\frac{2mm_1{\bar{L}}_4}{{\bar{L}}_3}\left\{ 1+\frac{m_1 I^*}{P^*{\bar{L}}_3} \right\} \right] ,\\ {\bar{L}}_3&=m_1bI^*+e(1+aI^*)(1+bP^*)^2 \ \text {and} \\ {\bar{L}}_4&=\left\{ (1+aI^*)^3(1+bP^*) \right\} ^{-1}. \end{aligned}$$

Appendix C

The characteristic equation corresponding to \(E_2\) is

$$\begin{aligned}&\left[ \lambda -\left( \frac{m_1\hat{I}}{1+a\hat{I}}-d \right) \right] \left[ \left\{ {\lambda }^2 +\lambda \left( c+\frac{r\hat{s}}{k}\right) \right. \right. \\&\quad \left. \left. +\frac{rc}{k}\left( \hat{S}+\hat{I}\right) \right\} + e^{-\lambda \tau }\left\{ -\frac{c^2}{\beta \hat{S}}\lambda \right. \right. \\&\quad \left. \left. +\left( \frac{c^3\hat{I}}{\beta \hat{S}^2}-\frac{c^2r}{\beta k}\right) \right\} \right] =0. \end{aligned}$$

Clearly, one eigenvalue is \(\frac{m_1\hat{I}}{1+a\hat{I}}-d<0 \) if \(\hat{I}<\frac{d}{m_1-ad}\) (under this condition non-delay system was asymptotically stable) and other two eigenvalues are the root of the equation

$$\begin{aligned} \lambda ^2+l_1\lambda +l_2=e^{-\lambda \tau }(l_3\lambda +l_4) \end{aligned}$$
(6.1)

where \(l_1=c+\frac{r\hat{s}}{k}\), \(l_2=\frac{rc}{k}\left( \hat{S}+\hat{I}\right) \), \(l_3=-\frac{c^2}{\beta \hat{S}}\) and \(l_4=\frac{c^3\hat{I}}{\beta \hat{S}^2}-\frac{c^2r}{\beta k}\).

Now we shall look for the condition under which the system is locally asymptotically stable and undergoes a Hopf bifurcation near \(E_2\). Putting \(\lambda =\mu +i\omega \) with \(\mu =0\), we have

$$\begin{aligned} l_3\omega \sin {\omega \tau }+l_4\cos {\omega \tau }+\left( \omega ^2-l_2\right) =0 \end{aligned}$$
(6.2)

and

$$\begin{aligned} l_4\sin {\omega \tau }-l_3\omega \cos {\omega \tau }+l_1\omega =0. \end{aligned}$$
(6.3)

With the help of Eqs. (6.2) and (6.3), we have

$$\begin{aligned} \rho _1^2+d_4\rho _1+d_5=0. \end{aligned}$$
(6.4)

where \(\rho _1={\hat{\omega }}^2\), \(d_4=l_1^2-2l_2-l_3^2\) and \(d_5=l_2^2-l_4^2\). Now if \(d_4<0,d_5<0\) or \(d_4>0,d_5>0\), then this quadratic Eq. (6.4) will have at least one positive root and consequently a positive root for \(\omega \), say \({\hat{\omega }}\), and corresponding critical value of \(\tau \), say \(\tau _\mathrm{c}\), is given by

$$\begin{aligned} \tau _\mathrm{c}^{(n)}&=\frac{1}{{\hat{\omega }}}\arcsin \frac{1}{\Delta _1}\left( l_2l_3{\hat{\omega }}-l_1l_4{\hat{\omega }}-l_3{\hat{\omega }}^3\right) \nonumber \\&\quad +\frac{2n\pi }{{\hat{\omega }}} \qquad \qquad \ \ n=0,1,2,3.... \end{aligned}$$
(6.5)

where \(\Delta _1=l_3^2{\hat{\omega }}^2+l_4^2\).

If \(E_2\) is locally asymptotically for \(\tau =0\), then by G.J. Butler’s lemma, \(E_2\) remains stable for \(\tau <\tau _\mathrm{c}\). The transversality condition for the Hopf bifurcation at \(\tau _\mathrm{c}\) is \(Re\left( \frac{d\lambda }{d\tau }\right) _{\tau =\tau _\mathrm{c}}>0\). So differentiating (6.1) with respect to \(\tau \) and putting \(\tau =\tau _\mathrm{c}\), \(\omega ={\hat{\omega }}\), \(\mu =0\), we get

$$\begin{aligned} \text {Re}\left( \frac{d\lambda }{d\tau }\right) _{\tau =\tau _\mathrm{c}}=\frac{P_1R_1-Q_1S_1}{P_1^2+Q_1^2}>0 \end{aligned}$$
(6.6)

provided \(P_1R_1-Q_1S_1>0\) where \(P_1=l_1-l_3\cos {{\hat{\omega }}\tau _\mathrm{c}}+l_3{{\hat{\omega }}\tau _\mathrm{c}}\sin {{\hat{\omega }}\tau _\mathrm{c}}+l_4\tau _\mathrm{c}\cos {{\hat{\omega }}\tau _\mathrm{c}} \), \(Q_1=-2{\hat{\omega }}-l_3{\hat{\omega }}\tau _\mathrm{c}\cos {{\hat{\omega }}\tau _\mathrm{c}}-l_3\sin {{\hat{\omega }}\tau _\mathrm{c}}+l_4\tau _\mathrm{c}\sin {{\hat{\omega }}\tau _\mathrm{c}}\), \(R_1=l_3{\hat{\omega }}^2\cos {{\hat{\omega }}\tau _\mathrm{c}}-l_4{\hat{\omega }}\sin {{\hat{\omega }}\tau _\mathrm{c}}\), and \(S_1=-l_3{\hat{\omega }}^2\sin {{\hat{\omega }}\tau _\mathrm{c}}-l_4{\hat{\omega }}\cos {{\hat{\omega }}\tau _\mathrm{c}}\). From (6.6), it is clear that the condition for the occurrence of Hopf bifurcation at \(\tau =\tau _\mathrm{c}\) is established.

Appendix D

Let \(\tau =\tau _\mathrm{c}+\mu \) where \(\mu =0\) is a Hopf bifurcation value of the system. Define the space of continuous real valued function as \(C=C([-1,0],R^3)\). Using the transformation \(u_1(t)=S-S^*\), \(u_2(t)=I-I^*\) and \(u_3(t)=P-P^*\), the system (4.14.4) is transformed to a functional differential equation in C as

$$\begin{aligned} \frac{\mathrm{d}U}{\mathrm{d}t}=L_{\mu }(U_t)+H(\mu ,U_t) \end{aligned}$$
(6.7)

where \(U=(u_1(t),u_2(t),u_3(t))^T\in R^3\), \(U_t(\Theta )=U(t+\Theta ) \ \ \text {for} \ \Theta \in [-1,0]\).

With \(\phi (t)=(\phi _1(t),\phi _2(t),\phi _3(t))^T\in C([-\tau ,0],R^3), \)\(L_\mu :C\rightarrow R^3\) and \(H:R\times C\rightarrow R^3\) are given as follows :

$$\begin{aligned} L_\mu (\phi )= & {} (\tau _\mathrm{c}+\mu )[M\phi (0)+N\phi (-1)],\end{aligned}$$
(6.8)
$$\begin{aligned} H(\mu ,\phi )= & {} (\tau _\mathrm{c}+\mu )\nonumber \\&\times \left[ \begin{array}{c} -\frac{r}{k}\phi _1^2(0)-\frac{r}{k}\phi _1(0)\phi _2(0)-\beta F_1\\ \beta F_1-mF_2 \\ m_1F_2-e\phi _3^2(0) \end{array}\right] \nonumber \\ \end{aligned}$$
(6.9)

where

$$\begin{aligned} F_1= & {} \sum \limits _{i+j\ge 2}\frac{1}{i!j!}\left[ \frac{\partial ^{i+j}}{\partial S^i \partial I^j}\left( \frac{IS}{1+\alpha I} \right) \right] _{(S,I)=(S^*,I^*)}\nonumber \\&\phi _1^i(0)\phi _2^j(-1),\end{aligned}$$
(6.10)
$$\begin{aligned} F_2= & {} \sum \limits _{i+j\ge 2}\frac{1}{i!j!}\left[ \frac{\partial ^{i+j}}{\partial I^i \partial P^j}\left( \frac{IP}{(1+aI)(1+bP)} \right) \right] _{(I,P)=(I^*,P^*)}\nonumber \\&\phi _2^i(0)\phi _3^j(0). \end{aligned}$$
(6.11)

Using the Riesz representation theorem, there exists a matrix function \(\eta (\theta ,\mu )\) whose components are of bounded variation for \(\theta \in [-1,0]\) such that

$$\begin{aligned} L_\mu (\phi )=\int _{-1}^{0}d\eta (\theta ,\mu )\phi (\theta ). \end{aligned}$$
(6.12)

Now we choose

$$\begin{aligned} \eta (\theta ,\mu )=(\tau _\mathrm{c}+\mu )[M\delta (\theta )-N\delta (\theta +1)] \end{aligned}$$
(6.13)

where \(\delta \) is the Dirac delta function. For \(\phi \in C([-1,0],R^3)\), we define

$$\begin{aligned} A(\mu )\phi&=\left\{ \begin{array}{ll} \frac{d\phi (\theta )}{d\theta }, &{}\quad \theta \in [-1,0)\\ \int _{-1}^{0}d\eta (\mu ,s)\phi (s), &{} \quad \theta =0, \end{array}\right. \end{aligned}$$
(6.14)
$$\begin{aligned} R(\mu )\phi&=\left\{ \begin{array}{ll} 0&{}\qquad \theta \in [-1,0)\\ H(\mu ,\phi ), &{}\qquad \theta =0. \end{array}\right. \end{aligned}$$
(6.15)

Then Eq. (6.7) can be transformed into following operator differential equation

$$\begin{aligned} \frac{\mathrm{d}U}{\mathrm{d}t}=A(\mu )U_t+R(\mu )U_t. \end{aligned}$$
(6.16)

For \(\psi \in C^1([-1,0],(R^3)^*)\), the adjoint operator \(A^*\) of A is defined as

$$\begin{aligned} A^*\psi (s)=\left\{ \begin{array}{ll} -\frac{d\psi (s)}{ds} &{}\qquad for\ s\in (0,1]\\ \int _{-1}^{0}d\eta ^T(t,0)\psi (-t) &{}\qquad for \ s=0 \end{array}\right. \nonumber \\ \end{aligned}$$
(6.17)

and a bilinear inner product is defined as

$$\begin{aligned}&<\psi (s),\phi (\theta )>\nonumber \\&\quad ={\bar{\psi }}(0)\phi (0)\nonumber \\&\qquad -\int _{-1}^{0} \int _{\xi =0}^{\theta }{\bar{\psi }}(\xi -\theta )d\eta (\theta )\phi (\xi )d\xi \end{aligned}$$
(6.18)

where \(\eta (\theta )=\eta (\theta ,0)\). We know that \(\pm i\omega _0\tau _\mathrm{c} \) are eigenvalues of A(0) and consequently that are of \(A^*\) also. Now we need to find eigenvectors of A and \(A^*\) corresponding to \(+i\omega _0\tau _\mathrm{c}\) and \(-i\omega _0\tau _\mathrm{c}\), respectively. Let us assume \(q(\theta )=(1,\alpha _1,\beta _1)^Te^{i\omega _0\tau _\mathrm{c}\theta }\) be the eigenvector of A(0) corresponding to \(i\omega _0\tau _\mathrm{c}\) where \(\alpha _1=\frac{i\omega _0-m_{11}}{m_{12}+n_{12}e^{-i\omega _0\tau _\mathrm{c}}}\) and \(\beta _1=\frac{m_{32}}{i\omega _0-m_{33}}\alpha _1\). Similarly assuming \(q^*(s)=D(1,\alpha ^*_1,\beta ^*_1)^Te^{i\omega _0\tau _\mathrm{c}s}\) as the eigenvector of \(A^*\) corresponding to \(-i\omega _0\tau _\mathrm{c}\), we have \(\alpha ^*_1=-\frac{i\omega _0+m_{11}}{m_{21}} \) and \(\beta ^*_1=-\frac{1}{m_{32}}[(m_{12}+n_{12}e^{i\omega _0\tau _\mathrm{c}})+(m_{22}+n_{22}e^{i\omega _0\tau _\mathrm{c}}+i\omega _0 )\alpha ^*_1]\). Now we have to calculate the value of D from \(<q^*(s),q(\theta )>=1\). Using Eq. (6.18),

$$\begin{aligned}&<q^*(s),q(\theta )>\\&\quad ={\bar{D}}(1,\bar{\alpha ^*_1},\bar{\beta ^*_1}) (1,\alpha _1,\beta _1)^T\\&\qquad -\int _{-1}^{0}\int _{\xi =0}^{\theta }{\bar{D}} (1,\bar{\alpha ^*_1},\bar{\beta ^*_1})e^{-i\omega _0\tau _\mathrm{c} (\xi -\theta )}d\eta (\theta )\\&\qquad (1,\alpha _1,\beta _1)^Te^{i\omega _0\tau _\mathrm{c}\xi }d\xi \\&\quad ={\bar{D}}\left[ 1+\alpha _1\bar{\alpha ^*_1}+\beta _1\bar{\beta ^*_1}\right. \\&\qquad \left. +\frac{\beta S^*}{(1+\alpha I^*)^2}\alpha _1\tau _\mathrm{c}(\bar{\alpha ^*_1}-1) e^{-i\omega _0\tau _\mathrm{c}}\right] \end{aligned}$$

where \({\bar{D}}=\left[ 1+\alpha _1\bar{\alpha ^*_1}+\beta _1\bar{\beta ^*_1}+\frac{\beta S^*}{(1+\alpha I^*)^2}\alpha _1\tau _\mathrm{c}(\bar{\alpha ^*_1}-1)e^{-i\omega _0\tau _\mathrm{c}}\right] ^{-1}\).

Now we will calculate co-ordinates to describe the centre manifold \(C_0\) at \(\mu =0\). For each \(U\in D(A)\), we may choose the pair (yV) where

$$\begin{aligned} y(t)=\,<q^*,U_t>, \ V(t,\theta )= U_t(\theta )-2\text {Re}\left\{ y(t)q(\theta )\right\} .\nonumber \\ \end{aligned}$$
(6.19)

On the centre manifold \(C_0\), we have \(V(t,\theta )=V(y(t),{\bar{y}}(t),\theta )\) where

$$\begin{aligned} V(y,{\bar{y}},\theta )= & {} V_{20}(\theta )\frac{y^2}{2}+V_{11}(\theta )y{\bar{y}} +V_{02}(\theta )\frac{{\bar{y}}^2}{2}\nonumber \\&+V_{30}(\theta )\frac{y^3}{6}+\cdots \end{aligned}$$
(6.20)

where y and \({\bar{y}}\) are local co-ordinates for centre manifold \(C_0\) in the direction of \(q^*\) and \(\bar{q^*}\) . Clearly, V is real if \(U_t\) is real. We consider real solutions only. For solution \(U_t\in C_0\) of Eq. (6.16), since \(\mu =0\), we have

$$\begin{aligned} \dot{y}&=\,<q^*,\dot{U_t}>\, = i\omega _0\tau _\mathrm{c}y(t)+g(y,{\bar{y}}) \end{aligned}$$

where

$$\begin{aligned} g(y,{\bar{y}})= & {} \bar{q^*}(0)H_0(y,{\bar{y}})=g_{20}\frac{y^2}{2} +g_{11}y{\bar{y}}\nonumber \\&+g_{02}\frac{{\bar{y}}^2}{2}+g_{21}\frac{y^2{\bar{y}}}{2}+\cdots \end{aligned}$$
(6.21)

From Eq. (6.19), we have

$$\begin{aligned} U_t(\theta )&=V(t,\theta )+2\text {Re}\left\{ y(t)q(\theta )\right\} \\&= V(t,\theta )+yq(\theta )+{\bar{y}}{\bar{q}}(\theta ). \end{aligned}$$

For \(U_t(\theta )=(U_{1t}(\theta ),U_{2t}(\theta ),U_{3t}(\theta ))\), it follows from Eq. (6.9) that

$$\begin{aligned} g(y,{\bar{y}})&=\bar{q^*}(0)H_0(y,{\bar{y}})=\bar{q^*}(0)H(0,U_t)\nonumber \\&={\bar{D}}(1,\bar{\alpha ^*_1},\bar{\beta ^*_1)}\tau _\mathrm{c}\nonumber \\&\quad \left[ \begin{array}{c} -\frac{r}{k}U_{1t}^2(0)-\frac{r}{k}U_{1t}(0)U_{2t}(0)-\beta F_3\\ \beta F_3-mF_4 \\ m_1F_4-eU_{3t}^2(0) \end{array}\right] \nonumber \\&={\bar{D}}\tau _\mathrm{c}(r_1y^2+r_2y{\bar{y}}+r_3{\bar{y}}^2+r_4y^2{\bar{y}})+\cdots \end{aligned}$$
(6.22)

where

$$\begin{aligned} F_3= & {} \sum \limits _{i+j\ge 2}\frac{1}{i!j!}\left[ \frac{\partial ^{i+j}}{\partial S^i \partial I^j}\left( \frac{IS}{1+\alpha I} \right) \right] _{(S,I)=(S^*,I^*)}\nonumber \\&U_{1t}^i(0)U_{2t}^j(-1),\end{aligned}$$
(6.23)
$$\begin{aligned} F_4= & {} \sum \limits _{i+j\ge 2}\frac{1}{i!j!}\left[ \frac{\partial ^{i+j}}{\partial I^i \partial P^j}\left( \frac{IP}{(1+aI)(1+bP)} \right) \right] _{(I,P)=(I^*,P^*)}\nonumber \\&U_{2t}^i(0)U_{3t}^j(0). \end{aligned}$$
(6.24)

Comparing Eqs. (6.21) and (6.22), we have \( g_{20}=2{\bar{D}}\tau _\mathrm{c}r_1,\ g_{11}={\bar{D}}\tau _\mathrm{c}r_2,\ g_{02}=2{\bar{D}}\tau _\mathrm{c}r_3,\ g_{21}=2{\bar{D}}\tau _\mathrm{c}r_4\). Now we calculate \(V_{20}(\theta )\) and \(V_{11}(\theta )\) as they are present in \(g_{21}\). From Eqs. (6.16) and (6.19), we have

$$\begin{aligned} \dot{V}&=\dot{U_t}-\dot{y}q-\dot{{\bar{y}}}{\bar{q}}\nonumber \\&= \left\{ \begin{array}{ll} AV-2\text {Re}[\bar{q^*}(0)H_0q(\theta )] &{} \qquad -1\le \theta <0\\ AV-2\text {Re}[\bar{q^*}(0)H_0q(\theta )]+H_0 &{}\qquad \theta =0 \end{array}\right. \nonumber \\&=AV+G(y,{\bar{y}},\theta ) \end{aligned}$$
(6.25)

where

$$\begin{aligned} G(y,{\bar{y}},\theta )= & {} G_{20}(\theta )\frac{y^2}{2}+G_{11}(\theta )y{\bar{y}}\nonumber \\&+\,G_{02}(\theta )\frac{{\bar{y}}^2}{2}+\cdots \end{aligned}$$
(6.26)

Again near the origin

$$\begin{aligned} \dot{V}&=V_y\dot{y}+V_{{\bar{y}}}\dot{{\bar{y}}}\nonumber \\&=(V_{20}(\theta )y+V_{11}(\theta ){\bar{y}}+\cdots )(i\omega _0\tau _\mathrm{c}y(t))\nonumber \\&\quad +g(y,{\bar{y}}))+(V_{11}(\theta )y+V_{02}(\theta ){\bar{y}}+\cdots )\nonumber \\&\qquad (-i\omega _0\tau _\mathrm{c}{\bar{y}}(t))+{\bar{g}}(y,{\bar{y}})). \end{aligned}$$
(6.27)

Comparing Eqs. (6.26) and (6.27), we get

$$\begin{aligned}&(A(0)-2i\omega _0\tau _\mathrm{c}I_3)V_{20}(\theta )=-G_{20}(\theta ) \ \ \ \text {and} \nonumber \\&\quad A(0)V_{11}(\theta )=-G_{11}(\theta ).\nonumber \\ \end{aligned}$$
(6.28)

Case 1 :\(\theta \in [-1,0)\).

From Eqs. (6.19), (6.20), (6.25) and (6.26) for \(\theta \in [-1,0)\), we have

$$\begin{aligned} G(y,{\bar{y}},\theta )&=-g(y,{\bar{y}})q(\theta )-{\bar{g}}(y,{\bar{y}}) {\bar{q}}(\theta )\nonumber \\&=\left( g_{20}(\theta )\frac{y^2}{2}+g_{11}(\theta )y{\bar{y}}\right. \nonumber \\&\left. \quad +\,g_{02}(\theta ) \frac{{\bar{y}}^2}{2}+\cdots \right) \nonumber \\&\quad -\,\left( {\bar{g}}_{20}(\theta )\frac{{\bar{y}}^2}{2} +{\bar{g}}_{11}(\theta )y{\bar{y}}+{\bar{g}}_{02}(\theta )\frac{y^2}{2}+\cdots \right) . \end{aligned}$$
(6.29)

Comparing the coefficient of \(y^2\) and \(y{\bar{y}}\) between Eqs. (6.26) and (6.28),

$$\begin{aligned} G_{20}(\theta )&=-g_{20}q(\theta )-{\bar{g}}_{20}{\bar{q}}(\theta ) \ \ \ \text {and} \nonumber \\ G_{11}(\theta )&=-g_{11}q(\theta )-{\bar{g}}_{11}{\bar{q}}(\theta ). \end{aligned}$$
(6.30)

From Eqs. (6.28) and (6.8) using the definition of A,

$$\begin{aligned} V_{20}(\theta )&=\frac{ig_{20}}{\omega _0\tau _\mathrm{c}}q(0)e^{i\omega _0\tau _\mathrm{c}\theta }\nonumber \\&\quad +\frac{i{\bar{g}}_{02}}{3\omega _0\tau _\mathrm{c}}{\bar{q}}(0)e^{-i\omega _0\tau _\mathrm{c}\theta } +E_1e^{2i\omega _0\tau _\mathrm{c}\theta }. \end{aligned}$$
(6.31)

Similarly,

$$\begin{aligned} V_{11}(\theta )&=-\frac{ig_{11}}{\omega _0\tau _\mathrm{c}}q(0)e^{i\omega _0\tau _\mathrm{c}\theta }\nonumber \\&\quad +\frac{i{\bar{g}}_{11}}{\omega _0\tau _\mathrm{c}}{\bar{q}}(0)e^{-i\omega _0\tau _\mathrm{c}\theta }+E_2 \end{aligned}$$
(6.32)

where \(E_1=(E_1^{(1)},E_1^{(2)},E_1^{(3)})\) and \(E_2=(E_2^{(1)},E_2^{(2)},E_2^{(3)})\) are constant vector in \(R^{3}\).

Case 2 :\(\theta =0\).

From Eqs. (6.25) and (6.26),

$$\begin{aligned} G_{20}(0)&=-g_{20}q(0)-{\bar{g}}_{02}{\bar{q}}(0)+2\tau _\mathrm{c}(c_1,c_2,c_3)^T, \end{aligned}$$
(6.33)
$$\begin{aligned} G_{11}(0)&=-g_{11}q(0)-{\bar{g}}_{11}{\bar{q}}(0)+\tau _\mathrm{c}(c_4,c_5,c_6)^T \end{aligned}$$
(6.34)

where \(M_1=(c_1,c_2,c_3)^T\) and \(M_2=(c_4,c_5,c_6)^T\) are the coefficient of \(y^2\) and \(y{\bar{y}}\), respectively, in \(H_0(y,{\bar{y}})\) where

$$\begin{aligned} c_1&=-\frac{r}{k}(1+\alpha _1)-\frac{\beta \alpha _1}{(1+\alpha I^*)^{2}} e^{-i\omega _0\tau _\mathrm{c}}\\&\quad +\frac{\beta S^*}{(1+\alpha I^*)^{3}} \alpha _1^2e^{-2i\omega _0\tau _\mathrm{c}},\\ c_2&=\frac{\beta \alpha _1}{(1+\alpha I^*)^{2}}e^{-i\omega _0\tau _\mathrm{c}}\\&\quad -\frac{\beta S^*}{(1+\alpha I^*)^{3}}\alpha _1^2e^{-2i\omega _0\tau _\mathrm{c}} -\frac{m\alpha _1\beta _1}{(1+a I^*)^{2}(1+b P^*)^{2}}\\&\quad +\frac{mP^*}{(1+a I^*)^{3}(1+b P^*)}\alpha _1^2+\frac{mI^*}{(1+a I^*)(1+b P^*)^3} \beta _1^2,\\ c_3&=\frac{m_1\alpha _1\beta _1}{(1+a I^*)^{2}(1+b P^*)^{2}}-\frac{m_1P^*}{(1+a I^*)^{3} (1+b P^*)}\alpha _1^2\\&\quad -\left\{ \frac{m_1I^*}{(1+a I^*)(1+b P^*)^3}+e\right\} \beta _1^2,\\ c_4&=-\frac{r}{k}(2+\alpha _1+\bar{\alpha _1})\\&\quad -\frac{\beta }{(1+\alpha I^*)^{2}} \left\{ \bar{\alpha _1}e^{i\omega _0\tau _\mathrm{c}} +\alpha _1e^{-i\omega _0\tau _\mathrm{c}}\right\} \\&\quad +\frac{2\alpha _1\bar{\alpha _1}\beta S^*}{(1+\alpha I^*)^{3}}\\ c_5&=\frac{\beta }{(1+\alpha I^*)^{2}}\left\{ \bar{\alpha _1}e^{i\omega _0\tau _\mathrm{c}} +\alpha _1e^{-i\omega _0\tau _\mathrm{c}}\right\} \\&\quad -\frac{2\alpha _1\bar{\alpha _1}\beta S^*}{(1+\alpha I^*)^{3}}-\frac{m(\alpha _1\bar{\beta _1}+\bar{\alpha _1}\beta _1)}{(1+a I^*)^{2}(1+b P^*)^{2}}\\&\quad +\frac{2\alpha _1\bar{\alpha _1}mP^*}{(1+a I^*)^{3}(1+b P^*)}\\&\quad +\frac{2\beta _1\bar{\beta _1}mI^*}{(1+a I^*)(1+b P^*)^3},\\ c_6&=\frac{m_1(\alpha _1\bar{\beta _1}+\bar{\alpha _1}\beta _1)}{(1+a I^*)^{2} (1+b P^*)^{2}}-\frac{2\alpha _1\bar{\alpha _1}m_1P^*}{(1+a I^*)^{3}(1+b P^*)}\\&\quad -\left\{ \frac{m_1I^*}{(1+a I^*)(1+b P^*)^3}+e\right\} 2\beta _1\bar{\beta _1}. \end{aligned}$$

Now using Eqs. (6.31) and (6.33),

$$\begin{aligned} \left\{ 2i\omega _0\tau _\mathrm{c}I_3-\int _{-1}^{0}e^{2i\omega _0\tau _\mathrm{c}\theta }d\eta (\theta )\right\} E_1=2\tau _\mathrm{c}M_1. \end{aligned}$$

This gives \(M_3E_1=2M_1\) where

$$\begin{aligned} M_3= \left[ \begin{array}{ccc} 2i\omega _0-m_{11}&{} -m_{12}-n_{12}e^{-2i\omega _0\tau _\mathrm{c}} &{} 0 \\ -m_{21} &{} 2i\omega _0-m_{22}-n_{22}e^{-2i\omega _0\tau _\mathrm{c}} &{} -m_{23} \\ 0 &{} -m_{32} &{} 2i\omega _0-m_{33}\\ \end{array}\right] . \end{aligned}$$

By simple calculation, we have \(E_1^{(i)}=\frac{2\Delta _1^{(i)}}{\Delta _1}\) where \(\Delta _1=\det {(M_3)}\), and \(\Delta _1^{(i)}\) is the determinant value of \(M_3^{(i)}\) obtained by replacing \(i \ th \ (i=1,2,3)\) column vector of \(M_3\) by another column vector \((c_1,c_2,c_3)^T\). Similarly we have \(M_4E_2=2M_2\) where

$$\begin{aligned} M_4= \left[ \begin{array}{ccc} -m_{11}&{} -m_{12}-n_{12}&{} 0 \\ -m_{21} &{} -m_{22}-n_{22} &{} -m_{23} \\ 0 &{} -m_{32} &{}-m_{33}\\ \end{array}\right] . \end{aligned}$$

This gives \(E_2^{(j)}=\frac{2\Delta _2^{(j)}}{\Delta _2}\), where \(\Delta _2=\det {(M_4)}\) and \(\Delta _2^{(j)}\) is the determinant value of \(M_4^{(j)}\) obtained by replacing \(j \ th \ (j=1,2,3)\) column vector of \(M_4\) by another column vector \((c_4,c_5,c_6)^T\). Thus we can compute \(V_{20}(\theta )\) and \(V_{11}(\theta )\) from Eqs. (6.31) and (6.32) and consequently \(g_{21} \) also. Moreover, the following values can be computed as

$$\begin{aligned} C_1(0)&=\frac{i}{2\omega _0\tau _\mathrm{c}}\left( g_{20}g_{11}-2|g_{11}|^2 -\frac{|g_{02}|^2}{3} \right) +\frac{g_{21}}{2}\nonumber ,\\ \mu _2&=-\frac{\text {Re}\left\{ C_1(0)\right\} }{\text {Re} \left\{ \frac{d\lambda (\tau _\mathrm{c})}{d\tau } \right\} }\nonumber ,\\ \beta _2&=2\text {Re}\left\{ C_1(0)\right\} \nonumber ,\\ T_2&=-\frac{\text {Im}\left\{ C_1(0)\right\} +\mu _2 \text {Im} \left\{ \frac{d\lambda (\tau _\mathrm{c})}{d\tau } \right\} }{\omega _0\tau _\mathrm{c}}. \end{aligned}$$
(6.35)

These quantities determine the direction of Hopf bifurcation at \(\tau =\tau _\mathrm{c}\).

Appendix E

The derivatives of SI and P with respect to \(b, \alpha \) and \(\beta \) are obtained from following equations :

$$\begin{aligned}&rS\left( 1-\frac{S+I}{k}\right) -\frac{\beta IS}{1+\alpha I}-\gamma S=0,\\&\frac{\beta IS}{1+\alpha I}-\frac{mIP}{(1+aI)(1+bP)}-cI=0,\\&\text {and} \ \ \ \frac{m_1IP}{(1+aI)(1+bP)}-\mathrm{d}P-eP^2=0. \end{aligned}$$

b

\(\beta \)

\(\alpha \)

\(\frac{dS}{db}=-g_1 \frac{dI}{db}\)

\(\frac{dS}{d\beta }=-g_1\frac{dI}{d\beta }-\frac{kI}{r(1+\alpha I)}\)

\(\frac{dS}{d\alpha }=-g_1\frac{dI}{d\alpha }+\frac{\beta kI^2}{r(1+\alpha I)^2}\)

\(\frac{dI}{db}=\frac{(d+eP)(1+aI)^2P+m_1g_2P^2}{m_1(1-g_2g_3)}\)

\(\frac{dI}{d\beta }=g_2\frac{\mathrm{d}P}{d\beta }\)

\(\frac{dI}{d\alpha }=g_2\frac{\mathrm{d}P}{d\alpha }\)

\(\frac{\mathrm{d}P}{db}=g_3 \frac{dI}{db}+P^2 \)

\(\frac{\mathrm{d}P}{d\beta }=g_4\left\{ \frac{S}{1+\alpha I}-\frac{\beta kI}{r(1+\alpha I)^2} \right\} \)

\(\frac{\mathrm{d}P}{d\alpha }=\frac{(1+aI)(1+bP)^2}{m(1-g_2g_3)}\left\{ \frac{k\beta ^2 I^2}{r(1+\alpha I)^3}-\frac{\beta IS}{(1+\alpha I)^2}\right\} \)

where \(g_1=1+\frac{\beta k}{r(1+\alpha I)^2} \), \( g_2=\frac{(e+bd+2beP)(1+aI)^2}{m_1}\), \(g_3=\left[ \frac{amP}{(1+aI)^2(1+bP)}-\frac{\alpha \beta S}{(1+\alpha I)^2}-\frac{\beta g_1}{1+\alpha I}\right] \frac{(1+aI)(1+bP)^2}{m}\) and \(g_4=\left[ \frac{m}{(1+aI)(1+bP)^2}+\frac{\alpha \beta g_2S}{(1+\alpha I)^2}+\frac{\beta g_1g_2}{1+\alpha I}-\frac{amg_2P}{(1+aI)^2(1+bP)} \right] ^{-1}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Maiti, A.P., Jana, C. & Maiti, D.K. A delayed eco-epidemiological model with nonlinear incidence rate and Crowley–Martin functional response for infected prey and predator. Nonlinear Dyn 98, 1137–1167 (2019). https://doi.org/10.1007/s11071-019-05253-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-019-05253-6

Keywords

Navigation