Skip to main content
Log in

The Effect of Time Delay in a Two-Patch Model with Random Dispersal

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

Abstract

We consider a two-patch model for a single species with dispersal and time delay. For some explicit range of dispersal rates, we show that there exists a critical value τ c for the time delay τ such that the unique positive equilibrium of the system is locally asymptotically stable for τ∈[0,τ c ) and unstable for τ>τ c .

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

Similar content being viewed by others

References

  • Busenberg, S., & Huang, W. (1996). Stability and Hopf bifurcation for a population delay model with diffusion effects. J. Differ. Equ., 124, 80–107.

    Article  MathSciNet  MATH  Google Scholar 

  • Cantrell, R. S., & Cosner, C. (2003). Spatial ecology via reaction-diffusion equations. Series in mathematical and computational biology. Chichester: Wiley.

    MATH  Google Scholar 

  • Chen, S., & Shi, J. (2012). Stability and Hopf bifurcation in a diffusive logistic population model with nonlocal delay effect. J. Differ. Equ., 253, 3440–3470.

    Article  MATH  MathSciNet  Google Scholar 

  • Chen, S., Shi, J., & Wei, J. (2013). Time delay-induced instabilities and Hopf bifurcations in general reaction-diffusion systems. J. Nonlinear Sci., 23, 1–38.

    Article  MathSciNet  MATH  Google Scholar 

  • Cooke, K., van den Driessche, P., & Zou, X. (1999). Interaction of maturation delay and nonlinear birth in population and epidemic models. J. Math. Biol., 39, 332–352.

    Article  MathSciNet  MATH  Google Scholar 

  • Davidson, F. S., & Gourley, S. A. (2001). The effects of temporal delays in a model for a food-limited, diffusing population. J. Math. Anal. Appl., 261, 633–648.

    Article  MathSciNet  MATH  Google Scholar 

  • Feng, W., & Lu, X. (1999). On diffusive population models with toxicants and time delays. J. Math. Anal. Appl., 233, 373–386.

    Article  MathSciNet  MATH  Google Scholar 

  • Gourley, S. A., & Ruan, S. (2000). Dynamics of the diffusive Nicholson’s blowflies equation with distributed delay. Proc. R. Soc. Edinb., Sect. A, 130, 1275–1291.

    Article  MathSciNet  MATH  Google Scholar 

  • Gourley, S. A., & So, J. W.-H. (2002). Dynamics of a food-limited population model incorporating nonlocal delays on a finite domain. J. Math. Biol., 44, 49–78.

    Article  MathSciNet  MATH  Google Scholar 

  • Hassard, B. D., Kazarinoff, N. D., & Wan, Y. H. (1981). Theory and applications of Hopf bifurcation. Cambridge: Cambridge University Press.

    MATH  Google Scholar 

  • He, X., Wu, J., & Zou, X. (1998). Dynamics of single species populations over a patchy environment. In L. Chen et al. (Eds.), Advanced topics in biomathematics, Singapore: World Scientific.

    Google Scholar 

  • Kuang, Y. (1993). Delay differential equations with applications in population dynamics. San Diego: Academic Press

    MATH  Google Scholar 

  • Liao, K.-L., Shih, C.-W., & Tseng, J.-P. (2012). Synchronized oscillations in mathematical model of segmentation in zebrafish. Nonlinearity, 25, 869–904.

    Article  MathSciNet  MATH  Google Scholar 

  • Madras, N., Wu, J., & Zou, X. (1996). Local-nonlocal interaction and spatio-temporal patterns in single-species population over a patch environment. Can. Appl. Math. Q., 4, 109–134.

    MathSciNet  MATH  Google Scholar 

  • Memory, M. C. (1989). Bifurcation and asymptotic behaviour of solutions of a delay-differential equation with diffusion. SIAM J. Math. Anal., 20, 533–546.

    Article  MathSciNet  MATH  Google Scholar 

  • Morita, Y. (1984). Destabilization of periodic solutions arising in delay-diffusion systems in several space dimensions. Jpn. J. Appl. Math., 1, 39–65.

    Article  MATH  MathSciNet  Google Scholar 

  • Shi, X., Cui, J., & Zhou, X. (2011). Stability and Hopf bifurcation analysis on an eco-epidemic model with a stage structure. Nonlinear Anal., 74, 1088–1106.

    Article  MathSciNet  MATH  Google Scholar 

  • Smith, H. L. (1996). Monotone dynamical system: an introduction to the theory of competitive and cooperative systems. Rhode Island: Am. Math. Soc.

    Google Scholar 

  • So, J. W.-H., Wu, J., & Yang, Y. (2000). Numerical steady state and Hopf bifurcation analysis on the diffusive Nicholson’s blowflies equation. Appl. Math. Comput., 111, 33–51.

    Article  MathSciNet  MATH  Google Scholar 

  • So, J. W.-H., & Yang, Y. (1998). Dirichlet problem for the diffusive Nicholson’s blowflies equation. J. Differ. Equ., 150, 317–348.

    Article  MathSciNet  MATH  Google Scholar 

  • Song, Y., Zhang, T., & Tadé, M. O. (2009). Stability switches, Hopf bifurcations, and spatio-temporal patterns in a delayed neural model with bidirectional coupling. J. Nonlinear Sci., 19, 597–632.

    Article  MathSciNet  MATH  Google Scholar 

  • Su, Y., Wei, J., & Shi, J. (2009). Hopf bifurcations in a reaction-diffusion population model with delay effect. J. Differ. Equ., 247, 1156–1184.

    Article  MathSciNet  MATH  Google Scholar 

  • Su, Y., Wei, J., & Shi, J. (2010). Bifurcation analysis in a delayed diffusive Nicholson’s blowflies equation. Nonlinear Anal., Real World Appl., 11, 1692–1703.

    Article  MathSciNet  MATH  Google Scholar 

  • Sun, C., Lin, Y., & Han, M. (2006). Stability and Hopf bifurcation for an epidemic disease model with delay. Chaos Solitons Fractals, 30, 204–216.

    Article  MathSciNet  MATH  Google Scholar 

  • Wan, A., & Zou, X. (2009). Hopf bifurcation analysis for a model of genetic regulatory system with delay. J. Math. Anal. Appl., 356, 464–476.

    Article  MathSciNet  MATH  Google Scholar 

  • Wang, L., & Zou, X. (2004). Hopf bifurcation in bidirectional associative memory neural networks with delays: analysis and computation. J. Comput. Appl. Math., 167, 73–90.

    Article  MathSciNet  MATH  Google Scholar 

  • Wang, L., & Zou, X. (2005). Stability and bifurcation of bidirectional associative memory neural networks with delayed self-feedback. Int. J. Bifurc. Chaos, 15, 2145–2159.

    Article  MathSciNet  MATH  Google Scholar 

  • Wei, J., & Zou, X. (2006). Bifurcation analysis of a population model and the resulting SIS epidemic model with delay. J. Comput. Appl. Math., 197, 169–187.

    Article  MathSciNet  MATH  Google Scholar 

  • Xu, C., & Li, P. (2012). Dynamical analysis in a delayed predator–prey model with two delays. Discrete Dyn. Nat. Soc.. doi:10.1155/2012/652947.

    MATH  MathSciNet  Google Scholar 

  • Xu, C., Tang, X., & Liao, M. (2010). Stability and bifurcation analysis of a delayed predator–prey model of prey dispersal in two-patch environments. Appl. Math. Comput., 216, 2920–2936.

    Article  MathSciNet  MATH  Google Scholar 

  • Xu, X., & Wu, Y. (2012). The effect of time delay on dynamical behavior in an ecoepidemiological model. J. Appl. Math. doi:10.1155/2012/286961.

    MATH  MathSciNet  Google Scholar 

  • Yi, T., & Zou, X. (2010). Map dynamics versus dynamics of associated delay reaction-diffusion equations with a Neumann condition. Proc. R. Soc. A, 466, 2955–2973.

    Article  MathSciNet  MATH  Google Scholar 

  • Yoshida, K. (1982). The Hopf bifurcation and its stability for semilinear diffusion equations with time delay arising in ecology. Hiroshima Math. J., 12, 321–348.

    MathSciNet  MATH  Google Scholar 

  • Yu, W., & Cao, J. (2006). Stability and Hopf bifurcation analysis on a four-neuron BAM neural network with time delays. Phys. Lett. A, 351, 64–78.

    Article  MATH  Google Scholar 

Download references

Acknowledgements

The authors are grateful to two anonymous reviewers for their helpful comments which substantially improved the manuscript. They thank Profs. Stephan van Gils, Chih-Wen Shih and Chang-Yuan Cheng and Kuang-Hui Lin for helpful discussions. This research was partially supported by the National Science Council of Taiwan, R.O.C., under No. NSC 101-2917-I-564-062 (K.-L. Liao), the NSF under Agreement No. 0635561 (K.-L. Liao), and the NSF Grant DMS-1021179 (Y. Lou), and has been supported in part by the Mathematical Biosciences Institute and the National Science Foundation under Grant DMS-0931642.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kang-Ling Liao.

Appendices

Appendix A

In this part, we display the main steps of using the center manifold theorem and the normal form method to analyze the stability and the bifurcating direction of the bifurcating periodic orbits for Neumann boundary problem (5). See the details in Hassard et al. (1981), Liao et al. (2012), Yu and Cao (2006). Herein, we use the supremum norm ∥ϕ∥=supτθ≤0|ϕ(θ)|. Hence, the phase space \(\mathcal{C}:=\mathcal {C}([-\tau,0],\mathbb{R}^{2 })\) is a Banach space of continuous functions.

Denote the solution of (5) as \({\mathbb{U}}(t)= (u_{1}(t), u_{2}(t))^{T}\), and set \({\mathbb{U}}_{t}(\theta)={\mathbb{U}}(t+\theta)\), θ∈[−τ,0], ν:=ττ c , where T denotes the transpose. Then, we can rewrite (11) in the form

$$ \dot{{\mathbb{U}}}(t)=L_{\nu}({\mathbb{U}}_{t})+G_{\nu}({\mathbb{U}}_{t}), $$
(100)

with the linear operator \(L_{\nu}:\mathcal{C}\rightarrow \mathbb{R}^{2}\) and the nonlinear operator \(G_{\nu}:\mathcal {C}\rightarrow\mathbb{R}^{2}\). More precisely, L ν  is defined by

$$L_{\nu}(\phi)=M_{0}\phi(0)+M_{1}\phi(- \tau), $$

with

$$ M_{0}=\left ( \begin{array}{c@{\quad}c} -d+\mu(r_1-a) & d\\ d & -d+\mu(r_2-b) \end{array} \right ),\qquad M_{1}=\left ( \begin{array}{c@{\quad}c} -\mu a & 0\\ 0 & -\mu b \end{array} \right ). $$
(101)

In addition, by using the steady-state system (14), M 0 can be rewritten as

$$\begin{aligned} M_{0}=\left ( \begin{array}{c@{\quad}c} -bd/a & d\\ d & -ad/b \end{array} \right ). \end{aligned}$$
(102)

We can choose

$$\eta(\theta,\nu)= M_{0}\delta(\theta)+M_{1} \delta(\theta+\tau), $$

where δ(⋅) is the Dirac function and η(⋅,μ) is obtained from the Riesz representation theorem, that its entries are functions of bounded variation on [−τ,0] such that

$$ L_{\nu}(\phi)=\int_{-\tau}^{0}d \eta(\theta,\nu)\phi(\theta). $$
(103)

On the other hand, the operator \(G_{\nu}:\mathcal{C}\rightarrow\mathbb{R}^{2}\) is defined by

$$ G_{\nu}(\phi)=\left ( \begin{array}{c} -\mu\phi_{1}(0)\phi_{1}(-\tau)+\mbox{h.o.t.}\\ -\mu\phi_{2}(0) \phi_{2}(-\tau)+\mbox{h.o.t.} \end{array} \right ). $$
(104)

Next, we define two operators on \(\mathcal{C}^{1} := \mathcal{C}^{1}([-\tau ,0],\mathbb{R}^{2})\):

$$\begin{aligned} (A_{\nu}\phi) (\theta) =&\left \{ \begin{array}{l@{\quad}l} d\phi(\theta)/d\theta, & \theta\in[-\tau,0), \\ \int_{-\tau}^{0}d\eta(\zeta,\nu) \phi(\zeta), & \theta=0; \end{array} \right . \end{aligned}$$
(105)
$$\begin{aligned} (R_{\nu}\phi) (\theta) =&\left \{ \begin{array}{l@{\quad}l} 0, & \theta\in[-\tau,0), \\ G_{\nu}(\phi), & \theta=0, \end{array} \right . \end{aligned}$$
(106)

and transform (100) into

$$ \dot{{\mathbb{U}}}_{t}=A_{\nu}{\mathbb{U}}_{t}+R_{\nu}{\mathbb{U}}_{t}. $$
(107)

Note that the adjoint operator \(A^{\ast}_{\nu}\) of A ν is defined as

$$\bigl(A^{\ast}_{\nu}\psi\bigr) \bigl( \theta^{\ast}\bigr)= \left \{ \begin{array}{l@{\quad}l} -d\psi(\theta^{\ast})/d\theta^{\ast}, & \theta^{\ast}\in (0,\tau], \\ \int_{-\tau}^{0}d\eta^{T}(\zeta,\nu)\psi(-\zeta), & \theta^{\ast}=0, \end{array} \right . $$

where \(\psi\in\mathcal{C}^{1}([0,\tau],\mathbb{R}^{2})\). For convenience, we allow functions to have range in \(\mathbb{C}^{2}\), instead of merely \(\mathbb{R}^{2}\), in the following computation.

We define the following bilinear form:

$$\langle\psi,\phi\rangle:=\overline{\psi}^T(0) \phi(0)- \int_{\theta=-\tau}^{0}\int_{\xi=0}^{\theta} \overline{\psi }^{T}(\xi-\theta)\,d\eta(\theta) \phi(\xi)\,d\xi, $$

for \(\phi\in\mathcal{C}([-\tau,0],\mathbb{C}^{2})\), \(\psi\in\mathcal {C}([0,\tau],\mathbb{C}^{2})\), where η(θ):=η(θ,0), to determine the coordinates of the center manifold near the origin of (100). Note that the overline stands for complex conjugate.

Next, we aim at finding the eigenvectors q(θ) and q (θ ) of A:=A 0 and \(A^{\ast}:=A^{\ast}_{0}\) with purely imaginary eigenvalues iw c and −iw c , respectively; namely,

$$ Aq(\theta)=iw_{c}q(\theta) \quad\mbox{and}\quad A^{\ast}q^{\ast}\bigl(\theta^{\ast}\bigr)=-iw_{c}q^{\ast} \bigl(\theta^{\ast}\bigr), $$
(108)

which satisfy the normalized conditions 〈q ,q〉=1 and \(\langle q^{\ast},\bar{q} \rangle=0\). Hence, we set

$$ q(\theta)=q(0)e^{iw_{c}\theta},\qquad q^{\ast}\bigl( \theta^{\ast}\bigr)=q^{\ast }(0)e^{iw_{c}\theta^{\ast}}, $$
(109)

for θ∈[−τ,0), θ ∈(0,τ], and \(q(0)=(q_{1}, q_{2})^{T}, q^{\ast}(0)=\frac{1}{\rho}(q_{1}^{\ast}, q^{\ast}_{2})^{T}\), and substitute (109) into (108) at θ=0 to yield

$$\begin{aligned} & q_{1}=1,\qquad q_{2}=\frac{b}{a}+ \frac{1}{d}\bigl(iw_c+\mu a e^{-iw_c\tau}\bigr), \\ & q_{1}^{\ast}=1,\qquad q_{2}^{\ast}= \frac{b}{a}+\frac{1}{d}\bigl(-iw_c+\mu a e^{iw_c \tau}\bigr). \end{aligned}$$

In the sequel, 〈q ,q〉=1 and \(\langle q^{\ast},\bar {q}\rangle=0\), if we set

$$ \rho=\overline{q_{1}}q_{1}^{\ast}+ \overline{q_{2}}q_{2}^{\ast}+\bar {J}_1 \tau e^{iw_c \tau}, $$
(110)

where \(J_{1}:=-\mu(a q_{1} \overline{q_{1}^{\ast}}+b q_{2} \overline{q_{2}^{\ast}})\).

Next, by using q and q , we can then construct a coordinate for the center manifold Ω 0 at ν=0. Denote \({\mathbb{U}}_{t}={\mathbb{U}}_{t}(\theta)=(u_{1,t}(\theta), u_{2,t}(\theta))^{T}\) as a solution of (107), and define

$$\begin{aligned} z(t) :=&\bigl\langle q^{\ast},{\mathbb{U}}_{t}\bigr\rangle , \end{aligned}$$
(111)
$$\begin{aligned} W(t,\theta) :=& {\mathbb{U}}_{t}(\theta)-2 \operatorname{Re} \bigl(z(t)q(\theta )\bigr). \end{aligned}$$
(112)

According to the center manifold reduction, we have \(W(t,\theta)= W(z(t),\bar{z}(t),\theta)\) on Ω 0. By the tangency to the center eigenspace at the equilibrium, we further have

$$ W(t,\theta)= W\bigl(z(t),\bar{z}(t),\theta\bigr)=W_{20}( \theta)\frac {z^{2}(t)}{2}+W_{11}(\theta)z(t)\bar{z}(t)+W_{02}( \theta)\frac{\bar {z}^{2}(t)}{2}+\cdots , $$
(113)

where z and \(\bar{z}\) are the local coordinates of the center manifold Ω 0 in directions q and \(\overline{{q}^{\ast}}\), respectively.

Accordingly, the solution \({\mathbb{U}}_{t}\in\varOmega_{0}\) of (107), at ν=0, satisfies

$$ \dot{z}(t)=iw_{c}z(t)+g\bigl(z(t),\bar{z}(t)\bigr), $$
(114)

with

$$ g(z,\bar{z})= \overline{q^{\ast}}^{T}(0) G_{0}({\mathbb{U}}_{t})=g_{20} \frac{z^{2}}{2}+g_{11}z\bar{z}+g_{02}\frac {\bar{z}^{2}}{2}+g_{21} \frac{z^{2}\bar{z}}{2}+\cdots, $$
(115)

and

$$ \begin{aligned} g_{20}&=-\frac{2}{\bar{\rho}}e^{-iw_c \tau}\mu\bigl(q_1^2 \overline {q_1^{\ast}}+q_2^2 \overline{q_2^{\ast}}\bigr), \\ g_{11}&=\frac{-\mu}{\bar{\rho}}\bigl(q_1 \overline{q_1}\overline{q_1^{\ast }}+q_2 \overline{q_2}\overline{q_2^{\ast}}\bigr) \bigl(e^{-iw_c \tau}+e^{iw_c \tau }\bigr), \\ g_{02}&=-\frac{2}{\bar{\rho}}e^{iw_c \tau}\mu\bigl(\overline {q_1^{2}}\overline{q_1^{\ast}}+ \overline{q_2^2}\overline{q_2^{\ast}} \bigr), \\ g_{21}&=\frac{-1}{\bar{\rho}}\bigl[\mu\bigl(2 q_1 \overline{q_1^{\ast }}W_{11}^{(1)}(-\tau)\\ &\quad{}+2 e^{-iw_c \tau}\bigl(q_1 \overline{q_1^{\ast }}W_{11}^{(1)}(0)+q_2 \overline{q_2^{\ast}}W_{11}^{(2)}(0)+2 q_2 \overline{q_2^{\ast}}W_{11}^{(2)}(- \tau) \\ &\quad{}+\overline{q_1}\overline{q_1^{\ast}}W_{20}^{(1)}(- \tau)+e^{iw_c \tau }\bigl(\overline{q_1} \overline{q_1^{\ast}}W_{20}^{(1)}(0)+ \overline {q_2}\overline{q_2^{\ast}}W_{20}^{(2)}(0) +\overline{q_2}\overline{q_2^{\ast}}W_{20}^{(2)}(- \tau)\bigr)\bigr)\bigr)\bigr], \end{aligned} $$
(116)

where \(W_{20}^{(k)}(\theta)\) and \(W_{11}^{(k)}(\theta)\) are the kth components of W 20(θ) and W 11(θ), respectively, for −τθ<0. Note that W 20(θ) and W 11(θ) are defined by

$$ \left \{ \begin{array}{l} W_{20}(\theta):=\frac{ig_{20}}{w_{c}}q(0)e^{iw_{c}\theta}-\frac{\bar {g}_{02}}{3iw_{c}}\bar{q}(0)e^{-iw_{c}\theta}+E_{1}e^{2iw_{c}\theta},\\ W_{11}(\theta):=\frac{g_{11}}{iw_{c}}q(0)e^{iw_{c}\theta}- \frac{\bar{g}_{11}}{iw_{c}}\bar{q}(0)e^{-iw_{c}\theta}+E_{2}, \end{array} \right . $$
(117)

where

$$\begin{aligned} E_{1} =&\biggl(2iw_{c}I-\int_{-\tau}^{0}e^{2iw_{c}\theta}d \eta(\theta,0)\biggr)^{-1} \left ( \begin{array}{c} -2\mu q_1^{2} e^{-iw_c \tau}\\ -2\mu q_2^{2} e^{-iw_c \tau} \end{array} \right ) \\ =&\left ( \begin{array}{c} \frac{2e^{-iw_c \tau}\mu(-d q_2^2-q_1^2(d\frac{a}{b}+2iw_c+b\mu e^{-2iw_c \tau}))}{-d^2+(d\frac{b}{a}+2iw_c+a\mu e^{-2iw_c \tau})(d\frac{a}{b}+2iw_c+b \mu e^{-2iw_c \tau})}\\ \frac{2 e^{-iw_c \tau}\mu(-d q_1^2-q_2^2(d \frac{b}{a}+2iw_c+a \mu e^{-2iw_c \tau}))}{-d^2+(d\frac{b}{a}+2iw_c+a \mu e^{-2iw_c \tau})(d\frac{a}{b}+2iw_c+b \mu e^{-2iw_c \tau})} \end{array} \right ), \end{aligned}$$
(118)
$$\begin{aligned} E_{2} =&\biggl(-\int_{-\tau}^{0}d\eta( \theta,0)\biggr)^{-1} \left ( \begin{array}{c} -\mu q_1 \overline{q_1}(e^{-iw_c \tau}+e^{iw_c \tau})\\ -\mu q_2 \overline{q_2}(e^{-iw_c \tau}+e^{iw_c \tau}) \end{array} \right ) \\ =& \left ( \begin{array}{c} -\frac{a e^{-iw_c \tau}(1+e^{2iw_c \tau})(q_1(a d+b^2 \mu)\overline {q_1}+b d q_2 \overline{q_2})}{a^3d+b^3d+a^2 b^2\mu} \\ -\frac{b e^{-iw_c \tau}(1+e^{2iw_c \tau})(a d q_1 \overline{q_1}+q_2(b d+a^2 \mu)\overline{q_2})}{a^3d+b^3d+a^2b^2\mu} \end{array} \right ). \end{aligned}$$
(119)

Accordingly, we already derive all formulas to find the values of g 20,g 11,g 02, and g 21. We can then determine the magnitudes of C 1(0),ν 2,ζ 2, and T 2.

Appendix B

Comparing with Appendix A, herein we consider the center manifold theorem and the normal form method for Dirichlet boundary problem (6) to analyze the stability and the bifurcating direction of the bifurcating orbits. In space \(\mathcal{C}:=\mathcal{C}([-\tau,0],\mathbb{R}^{2 })\) with the supremum norm ∥ϕ∥=supτθ≤0|ϕ(θ)|, we denote the solution of (6) as \(\tilde{{\mathbb{U}}}(t)= (u_{1}(t), u_{2}(t))^{T}\) and set \(\tilde{{\mathbb{U}}}_{t}(\theta)=\tilde{{\mathbb{U}}}(t+~\theta), \theta\in[-\tau, 0]\), and \(\nu:=\tau-\tilde{\tau}_{c}\). Then, system (40) can be rewritten as

$$ \dot{\tilde{{\mathbb{U}}}}(t)=\tilde{L}_{\nu}(\tilde{{ \mathbb{U}}}_{t})+\tilde{G}_{\nu}(\tilde{{\mathbb{U}}}_{t}), $$
(120)

with the linear operator \(\tilde{L}_{\nu}:\mathcal{C}\rightarrow \mathbb{R}^{2}\) and the nonlinear operator \(\tilde{G}_{\nu}:\mathcal {C}\rightarrow\mathbb{R}^{2}\). More precisely, \(\tilde{L}_{\nu}\) is defined by

$$\tilde{L}_{\nu}(\phi)=\tilde{M}_{0}\phi(0)+ \tilde{M}_{1}\phi(-\tau), $$

with

$$ \tilde{M}_{0}=\left ( \begin{array}{c@{\quad}c} -2d+\mu(r_1-\tilde{a}) & d\\ d & -2d+\mu(r_2-\tilde{b}) \end{array} \right ),\qquad \tilde{M}_{1}=\left ( \begin{array}{c@{\quad}c} -\mu\tilde{a} & 0\\ 0 & -\mu\tilde{b} \end{array} \right ). $$
(121)

By using the steady-state system (43), we rewrite \(\tilde{M}_{0}\) as

$$\begin{aligned} \tilde{M}_{0}=\left ( \begin{array}{c@{\quad}c} -\tilde{b}d/\tilde{a} & d\\ d & -\tilde{a}d/\tilde{b} \end{array} \right ). \end{aligned}$$
(122)

We then choose

$$\tilde{\eta}(\theta,\nu)= \tilde{M}_{0}\tilde{\delta}( \theta)+\tilde {M}_{1}\tilde{\delta}(\theta+\tau), $$

where \(\tilde{\delta}(\cdot)\) is the Dirac function and \(\tilde{\eta }(\cdot,\mu)\) is obtained from the Riesz representation theorem, that its entries are functions of bounded variation on [−τ,0] such that

$$ \tilde{L}_{\nu}(\phi)=\int_{-\tau}^{0}d \tilde{\eta}(\theta,\nu)\phi (\theta). $$
(123)

On the other hand, the operator \(\tilde{G}_{\nu}:\mathcal{C}\rightarrow \mathbb{R}^{2}\) is defined by

$$ \tilde{G}_{\nu}(\phi)=\left ( \begin{array}{c} -\mu\phi_{1}(0)\phi_{1}(-\tau)+\mbox{h.o.t.}\\ -\mu\phi_{2}(0) \phi_{2}(-\tau)+\mbox{h.o.t.} \end{array} \right ). $$
(124)

By using the steady-state systems (14) and (43), the structures of \(\tilde{M}_{0}\) in (122) and M 0 in (102) are totally the same. In addition, the structures of \(\tilde{M}_{1}\) and M 1 are also the same with each other and the formulas of the linear and nonlinear operators \(\tilde {L}_{\nu}\) and \(\tilde{G}_{\mu}\) of system (120) are very similar to L μ and G μ for system (100) in Appendix A. Accordingly, the analysis and formulas in center manifold theorem and the normal form method of these two systems (5) and (6) are almost the same. More precisely, all formulas for system (6) are obtain from replacing the values a,b by \(\tilde{a},\tilde{b}\) in Appendix A. Therefore, we omit the details of center manifold theorem and the normal form method for system (6).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Liao, KL., Lou, Y. The Effect of Time Delay in a Two-Patch Model with Random Dispersal. Bull Math Biol 76, 335–376 (2014). https://doi.org/10.1007/s11538-013-9921-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-013-9921-7

Keywords

Navigation