Skip to main content
Log in

Mean occupancy time: linking mechanistic movement models, population dynamics and landscape ecology to population persistence

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Reaction–diffusion models for the dynamics of a biological population in a fragmented landscape can incorporate detailed descriptions of movement and behavior, but are difficult to analyze and hard to parameterize. Patch models, on the other hand, are fairly easy to analyze and can be parameterized reasonably well, but miss many details of the movement process within and between patches. We develop a framework to scale up from a reaction–diffusion process to a patch model and, in particular, to determine movement rates between patches based on behavioral rules for individuals. Our approach is based on the mean occupancy time, the mean time that an individuals spends in a certain area of the landscape before it exits that area or dies. We illustrate our approach using several different landscape configurations. We demonstrate that the resulting patch model most closely captures persistence conditions and steady state densities as compared with the reaction–diffusion model.

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

Similar content being viewed by others

References

  • Artiles W, Carvalho PGS, Kraenkel RA (2008) Patch-size and isolation effects in the Fisher-Kolmogorov equation. J Math Biol 57:521–535

    Article  MATH  MathSciNet  Google Scholar 

  • Cantrell R, Cosner C (1999) Diffusion models for population dynamics incorporating individual behavior at boundaries: applications to refuge design. Theor Popul Biol 55:189–207

    Article  MATH  Google Scholar 

  • Cantrell R, Cosner C (2003) Spatial ecology via reaction-diffusion equations. Mathematical and computational biology. Wiley

  • Cantrell R, Cosner C (2007) Density dependent behaviour at habitat boundaries and the allee effect. Bull Math Biol 69(7):2339–2360

    Article  MATH  MathSciNet  Google Scholar 

  • Cantrell R, Cosner C, Fagan W (2012) The implications of model formulation when transitioning from spatial to landscape ecology. Math Biosci Eng 9(1):27–60

    Article  MATH  MathSciNet  Google Scholar 

  • Cobbold C, Lewis M, Lutscher F (2005) How parasitism affects critical patch size in a host-parasitoid system: application to Forest Tent Caterpillar. Theor Popul Biol 67(2):109–125

    Article  MATH  Google Scholar 

  • Fagan W, Lutscher F (2006) The average dispersal success approximation: a bridge linking home range size, natal dispersal, and metapopulation dynamics to critical patch size and reserve design. Ecol Appl 16(2):820–828

    Article  Google Scholar 

  • Frair J, Merrill E, Visscher D, Fortin D, Beyer H, Morales J (2005) Scales of movement by elk (Cervus elaphus) in response to heterogeneity in forage resources and predation risk. Landsc Ecol 20:273–287

    Article  Google Scholar 

  • Hadeler K (2000) Reaction transport equations in biological modeling. Math Comput Model 31:75–81

    Article  Google Scholar 

  • Hanggi P, Talkner P, Borkovec M (1990) Reaction rate theory: fifty years after Kramers. Rev Modern Phys 62(2):251–342

    Article  MathSciNet  Google Scholar 

  • Hanski I, Ovaskainen O (2000) The metapopulation capacity of a fragmented landscape. Nature 404:755–758

    Article  Google Scholar 

  • Kierstead H, Slobodkin LB (1953) The size of water masses containing plankton blooms. J Mar Res 12:141–147

    Google Scholar 

  • Kolpas A, Nisbet R (2010) Effects of demographic stochasiticity on population persistence in advective media. Bull Math Biol 72(5):1254–1270

    Article  MATH  MathSciNet  Google Scholar 

  • Kot M (2001) Elements of mathematical ecology. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Ludwig D, Aronson DG, Weinberger HF (1979) Spatial patterning of the spruce budworm. J Math Biol 8:217–258

    Article  MATH  MathSciNet  Google Scholar 

  • Lutscher F, Lewis MA (2004) Spatially-explicit matrix models. A mathematical analysis of stage-structured integrodifference equations. J Math Biol 48:293–324

    Article  MATH  MathSciNet  Google Scholar 

  • Maciel G, Lutscher F (2013) How individual response to habitat edges affects population persistence and spatial spread (Submitted)

  • Matkowsky B, Schuss Z (1977) The exit problem for randomly perturbed dynamical systems. SIAM Appl Math 33(2):365–382

    Article  MATH  MathSciNet  Google Scholar 

  • McKenzie H (2006) Linear features impact predator-prey encounters: analysis with first passage time. Master’s thesis, University of Alberta

  • McKenzie H, Lewis M, Merrill E (2009) First passage time analysis of animal movement and insights into the functional response. Bull Math Biol 71(1):107–129

    Article  MATH  MathSciNet  Google Scholar 

  • Othmer H, Adler F, Lewis M, Dallon J (1997) Mathematical modeling in biology: case studies in ecology, physiology and cell biology. Prentice Hall

  • Ovaskainen O (2008) Analytical and numerical tools for diffusion-based movement models. Theor Popul Biol 73:198–211

    Article  MATH  Google Scholar 

  • Ovaskainen O, Cornell S (2003) Biased movement at a boundary and conditional occupancy times for diffusion processes. J Appl Prob 40(3):557–580

    Article  MATH  MathSciNet  Google Scholar 

  • Redner S (2001) A guide to first-passage processes. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Samia Y, Lutscher F (2012) Persistence probabilities for stream populations. Bull Math Biol 74(7):1629–1650

    Article  MATH  MathSciNet  Google Scholar 

  • Schultz C, Crone E (2001) Edge-mediated dispersal behavior in a prairie butterfly. Ecology 82(7):1879–1892

    Article  Google Scholar 

  • Shigesada N, Kawasaki K, Teramoto E (1986) Traveling periodic waves in heterogeneous environments. Theor Popul Biol 30:143–160

    Article  MATH  MathSciNet  Google Scholar 

  • Singer A, Schuss Z, Osipov A, Holcman D (2008) Partially reflected diffusion. SIAM J Appl Math 28(3):844–868

    Article  MathSciNet  Google Scholar 

  • Skellam JG (1951) Random dispersal in theoretical populations. Biometrika 38:196–218

    MATH  MathSciNet  Google Scholar 

  • Strohm S, Tyson R (2012) The effect of habitat fragmentation on cyclic population dynamics: a reduction to ordinary differential equations. Theor Ecol. doi:10.1007/s12080-011-0141-1

    Google Scholar 

  • Turchin P (1998) Quantitative analysis of movement: measuring and modeling population redistribution of plants and animals. Sinauer, Sunderland

    Google Scholar 

  • Van Kirk RW, Lewis MA (1997) Integrodifference models for persistence in fragmented habitats. Bull Math Biol 59(1):107–137

    Article  MATH  Google Scholar 

  • Van Kirk RW, Lewis MA (1999) Edge permeability and population persistence in isolated habitat patches. Nat Resour Model 12:37–64

    Article  MATH  Google Scholar 

  • Vasilyeva O, Lutscher F (2012) Competition of three species in an advective environment. Nonlinear Anal: Real World Appl 13(4):1730–1748

    MATH  MathSciNet  Google Scholar 

  • Wakano J, Ikeda K, Miki T, Mimura M (2011) Effective dispersal rate is a function of habitat size and corridor shape: mechanistic formulation of a two-patch compartment model for spatially continuous systems. Oikos 120(11):1712–1720

    Article  Google Scholar 

  • Weins J, Stenseth N, Van Horne B, Ims R (1993) Ecological mechanisms and landscape ecology. Oikos 66:369–380

    Article  Google Scholar 

Download references

Acknowledgments

Some of this work was carried out while the first author visited Ottawa, funded by the Carnegie Trust for the Universities of Scotland. Additional support came from a grant to FL by the Canadian research network MITACS. FL is partially funded by a Discovery Grant from the Natural Sciences and Engineering Research Council of Canada.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christina A. Cobbold.

Appendices

Appendix A: Core and buffer habitat

We consider a core habitat, say a reserve, of length \(2L_1,\) with a buffer zone of length \(L_2\) attached to the core on either side (Cantrell and Cosner 1999). Since the one-dimensional model is symmetric with respect to the middle of the core habitat, we use this symmetry to reduce the model to the core (\(\varOmega _1=[0,L_1]\)) and the buffer (\(\varOmega _2=[-L_2,0]\)). The linearized equations for the density of individuals in the core and buffer are

$$\begin{aligned} u_t=D_i u_{xx}-m_i u +b_i u,\quad x\in \varOmega _i \end{aligned}$$
(70)

with \(r_i=b_i-m_i\) and \(r_1>r_2.\) We have a hostile boundary at \(x=-L_2\) and, due to symmetry, no-flux conditions at \(x=L_1,\) i.e.

$$\begin{aligned} u_x(L_1,t)=0,\quad u(-L_2,t)=0. \end{aligned}$$
(71)

At the interface between core and buffer, we require the flux to be continuous, but the density can be discontinuous if either there is movement preference by the individuals or if the diffusion rates differ between core and buffer (Ovaskainen and Cornell 2003), i.e.

$$\begin{aligned} D_1 u_x(0^+,t)=D_2 u_x(0^-,t),\quad u(0^+,t)=k u(0^-,t), \end{aligned}$$
(72)

where superscripts \(\pm \) stand for right- and left-sided limits. Parameter \(k\) summarizes bias and changes in movement. If individuals prefer the core habitat with probability \(p,\) then either (I) \(k=\frac{p\sqrt{D_2}}{(1-p) \sqrt{D_1}}\) or (II) \(k=\frac{p D_2}{(1-p)D_1},\) see (38).

We denote \(\mathcal M \) as the operator defined by (7072) with \(b_1=b_2=0.\) To determine the adjoint operator, we use the definition

$$\begin{aligned} \int \limits _{-L_2}^{L_1}v(x)\mathcal M u(\cdot ) dx=\int \limits _{-L_2}^{L_1}u(x)\mathcal M ^* v(\cdot ) dx \end{aligned}$$
(73)

We find that \(\mathcal M ^*\) is given by \(D_iv_{xx}-m_iv\) for \(x\in \varOmega _i\) with interface and boundary conditions,

$$\begin{aligned} kD_1 v_x(0^+,t)&= D_2 v_x(0^-,t),\quad v(0^+,t)=v(0^-,t),\end{aligned}$$
(74)
$$\begin{aligned} v_x(L_1,t)&= 0,\quad v(-L_2,t)=0. \end{aligned}$$
(75)

Occupancy time in the core (\(T_{\varOmega _1}\)) and buffer (\(T_{\varOmega _2}\)) satisfy (20), which yields the following ODEs

$$\begin{aligned}&D_1\frac{d^2T_{\varOmega _1}}{dy^2}-m_1T_{\varOmega _1}=-1,\quad y\in \varOmega _1;\quad D_2\frac{d^2T_{\varOmega _1}}{dy^2}-m_2T_{\varOmega _1}=0,\quad y\in \varOmega _2;\qquad \quad \end{aligned}$$
(76)
$$\begin{aligned}&D_1\frac{d^2T_{\varOmega _2}}{dy^2}-m_1T_{\varOmega _2}=0,\quad y\in \varOmega _1;\quad D_2\frac{d^2T_{\varOmega _2}}{dy^2}-m_2T_{\varOmega _2}=-1,\quad y\in \varOmega _2.\qquad \quad \end{aligned}$$
(77)

Setting \(m_1=0,\) and using the boundary condition at \(x=L_1,\) we obtain

$$\begin{aligned} T_{\varOmega _1}(y)=\frac{1}{D_1}\left(-\frac{y^2}{2}+L_1 y+C\right),\quad y\in \varOmega _1. \end{aligned}$$
(78)

A convenient representation of the solution on \(\varOmega _2\) that takes the boundary at \(x=-L_2\) into account, is

$$\begin{aligned} T_{\varOmega _1}(y)=A\sinh \left(\sqrt{\frac{m_2}{D_2}}(y+L_2)\right),\quad y\in \varOmega _2. \end{aligned}$$
(79)

The interface condition for \(T_{\varOmega _1}\) gives

$$\begin{aligned} C=D_1 A \sinh \left(\sqrt{\frac{m_2}{D_2}}L_2\right), \end{aligned}$$
(80)

whereas the condition on \(T_{\varOmega _1}^{\prime }\) gives

$$\begin{aligned} A=\frac{kL_1}{\sqrt{D_2 m_2}\cosh \left(\sqrt{\frac{m_2}{D_2}}L_2\right)}. \end{aligned}$$
(81)

The calculations for \(T_{\varOmega _2}\) are similar.

The definitions in (23) now give the elements of the matrix \(\mathcal T .\) Finally, the average population growth rate is given by the eigenvalues of the matrix

$$\begin{aligned} \begin{bmatrix} b_1&\quad 0 \\ 0&\quad b_{2} \end{bmatrix}-\mathcal T ^{-1}. \end{aligned}$$
(82)

Appendix B: MFPT of a rectangle

We consider the rectangular domain \(\varOmega =[0,a]\times [0,b]\) and the diffusion equation

$$\begin{aligned} \frac{\partial u}{\partial t}=D\left(\frac{\partial ^2 u}{\partial x^2}+\frac{\partial ^2 u}{\partial y^2}\right), \quad (x,y)\in \varOmega \end{aligned}$$
(83)

with hostile boundary conditions \(u=0\) for \(x\in \{0,a\}\) and \(y\in \{0,b\}.\)

The dominant eigenvalue can be calculated by standard separation of variables as

$$\begin{aligned} \lambda =-\pi ^2 D\frac{a^2+b^2}{a^2 b^2}. \end{aligned}$$
(84)

When we keep the area of the rectangle constant, we set \(b=1/a.\) In that case, the dominant eigenvalue is maximal when \(a=1.\)

The MFPT is given by the equation

$$\begin{aligned} D\left(\frac{\partial ^2 T}{\partial x^2}+\frac{\partial ^2 T}{\partial y^2}\right)=-1\quad \mathrm{on}\quad \varOmega . \end{aligned}$$
(85)

The ansatz

$$\begin{aligned} T(x,y)=\sum _{m,n} B_{m,n}\sin \left(\frac{m\pi x}{a}\right)\sin \left(\frac{n\pi x}{b}\right) \end{aligned}$$
(86)

leads to the condition

$$\begin{aligned} \sum _{m,n} \hat{B}_{m,n}\sin \left(\frac{m\pi x}{a}\right)\sin \left(\frac{n\pi x}{b}\right)=1,\quad \hat{B}_{m,n}= D\pi ^2 B_{m,n}\left(\frac{m^2}{a^2}+\frac{n^2}{b^2}\right). \end{aligned}$$
(87)

The coefficients must satisfy \(\hat{B}_{m,n}=16/(nm\pi ^2)\) when \(n\) and \(m\) are odd and \(\hat{B}_{m,n}=0\) otherwise. Accordingly, we obtain

$$\begin{aligned} T(x,y)=\sum _{m,n \;\mathrm{odd}} \frac{16}{nmD\pi ^2}\left(\frac{m^2}{a^2}+\frac{n^2}{b^2}\right)^{-1}\sin \left(\frac{m\pi x}{a}\right)\sin \left(\frac{n\pi x}{b}\right). \end{aligned}$$
(88)

After integrating, we find

$$\begin{aligned} \overline{T}=\sum _{m,n \;\mathrm{odd}} \frac{16}{nmD\pi ^2}\frac{a^2 b^2}{a^2 n^2+ b^2 m^2}. \end{aligned}$$
(89)

We evaluate this expression numerically with \(b=1/a\) and plot the result as a function of \(a\in [0,1]\) in Fig. 8.

Fig. 8
figure 8

Plot of MOT equation (89) for a rectangular domain with hostile boundary conditions. The area is fixed to 1, while the dimensions of the domain are varied, so illustrate how shape can effect MOT. The parameter \(D=1\)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Cobbold, C.A., Lutscher, F. Mean occupancy time: linking mechanistic movement models, population dynamics and landscape ecology to population persistence. J. Math. Biol. 68, 549–579 (2014). https://doi.org/10.1007/s00285-013-0642-1

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-013-0642-1

Keywords

Mathematics Subject Classification (2000)

Navigation