Skip to main content
Log in

A Mathematical Biologist’s Guide to Absolute and Convective Instability

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

Abstract

Mathematical models have been highly successful at reproducing the complex spatiotemporal phenomena seen in many biological systems. However, the ability to numerically simulate such phenomena currently far outstrips detailed mathematical understanding. This paper reviews the theory of absolute and convective instability, which has the potential to redress this inbalance in some cases. In spatiotemporal systems, unstable steady states subdivide into two categories. Those that are absolutely unstable are not relevant in applications except as generators of spatial or spatiotemporal patterns, but convectively unstable steady states can occur as persistent features of solutions. The authors explain the concepts of absolute and convective instability, and also the related concepts of remnant and transient instability. They give examples of their use in explaining qualitative transitions in solution behaviour. They then describe how to distinguish different types of instability, focussing on the relatively new approach of the absolute spectrum. They also discuss the use of the theory for making quantitative predictions on how spatiotemporal solutions change with model parameters. The discussion is illustrated throughout by numerical simulations of a model for river-based predator–prey systems.

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

Access this article

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

Instant access to the full article PDF.

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

Similar content being viewed by others

Notes

  1. Instead, curves of absolute spectrum can emanate from “triple points”, defined by \(\operatorname{Im} k_{i}=\operatorname {Im} k_{i+1}=\operatorname{Im} k_{i+2}\) for some i (Rademacher et al. 2007; Smith et al. 2009). See Fig. 6c for an example of parts of an absolute spectrum emanating from triple points.

References

  • Anderson, K. E., Nisbet, R. M., Diehl, S., & Cooper, S. D. (2005). Scaling population responses to spatial environmental variability in advection-dominated systems. Ecol. Lett., 8, 933–943.

    Article  Google Scholar 

  • Anderson, K. E., Paul, A. J., McCauley, E., Jackson, L. J., Post, J. R., & Nisbet, R. M. (2006). Instream flow needs in streams and rivers: the importance of understanding ecological dynamics. Front. Ecol. Environ., 4, 309–318.

    Article  Google Scholar 

  • Anderson, K. E., Hilker, F. M., & Nisbet, R. M. (2012). Directional dispersal and emigration behavior drive a flow-induced instability in a stream consumer-resource model. Ecol. Lett., 15, 209–217.

    Article  Google Scholar 

  • Aranson, I. S., & Kramer, L. (2002). The world of the complex Ginzburg–Landau equation. Rev. Mod. Phys., 74, 99–143.

    Article  MathSciNet  MATH  Google Scholar 

  • Aranson, I. S., Aranson, L., Kramer, L., & Weber, A. (1992). Stability limits of spirals and traveling waves in nonequilibrium media. Phys. Rev. A, 46, R2992–R2995.

    Article  Google Scholar 

  • Armstrong, R. A., & McGehee, R. (1980). Competitive exclusion. Am. Nat., 115, 151–170.

    Article  MathSciNet  Google Scholar 

  • Beyn, W.-J., & Lorenz, J. (1999). Stability of travelling waves: dichotomies and eigenvalue conditions on finite intervals. Numer. Funct. Anal. Optim., 20, 201–244.

    Article  MathSciNet  MATH  Google Scholar 

  • Biancofiore, L., Gallaire, F., & Pasquetti, R. (2011). Influence of confinement on a two-dimensional wake. J. Fluid Mech., 688, 297–320.

    Article  MATH  Google Scholar 

  • Brandt, M. J., & Lambin, X. (2007). Movement patterns of a specialist predator, the weasel Mustela nivalis exploiting asynchronous cyclic field vole Microtus agrestis populations. Acta Theriol., 52, 13–25.

    Article  Google Scholar 

  • Brevdo, L. (1988). A study of absolute and convective instabilities with an application to the Eady model. Geophys. Astrophys. Fluid Dyn., 40, 1–92.

    Article  MathSciNet  MATH  Google Scholar 

  • Brevdo, L. (1995). Convectively unstable wave packets in the Blasius boundary layer. Z. Angew. Math. Mech., 75, 423–436.

    Article  MathSciNet  MATH  Google Scholar 

  • Brevdo, L., & Bridges, T. J. (1996). Absolute and convective instabilities of spatially periodic flows. Philos. Trans. R. Soc. Lond. A, 354, 1027–1064.

    Article  MathSciNet  MATH  Google Scholar 

  • Brevdo, L., & Bridges, T. J. (1997a). Absolute and convective instabilities of temporally oscillating flows. Z. Angew. Math. Phys., 48, 290–309.

    MathSciNet  MATH  Google Scholar 

  • Brevdo, L., & Bridges, T. J. (1997b). Local and global instabilities of spatially developing flows: cautionary examples. Proc. R. Soc. Lond. A, 453, 1345–1364.

    Article  MathSciNet  MATH  Google Scholar 

  • Brevdo, L., Laure, P., Dias, F., & Bridges, T. J. (1999). Linear pulse structure and signalling in a film flow on an inclined plane. J. Fluid Mech., 396, 37–71.

    Article  MathSciNet  MATH  Google Scholar 

  • Briggs, R. J. (1964). Electron-stream interaction with plasmas. Cambridge: MIT Press.

    Google Scholar 

  • Chomaz, J. M. (2004). Transition to turbulence in open flows: what linear and fully nonlinear local and global theories tell us. Eur. J. Mech. B, 23, 385–399.

    Article  MathSciNet  MATH  Google Scholar 

  • Chomaz, J. M. (2005). Global instabilities in spatially developing flows: non-normality and nonlinearity. Annu. Rev. Fluid Mech., 37, 357–392.

    Article  MathSciNet  MATH  Google Scholar 

  • Cooper, S. D., Diehl, S., Kratz, K., & Sarnelle, O. (1998). Implications of scale for patterns and processes in stream ecology. Aust. J. Ecol., 23, 27–40.

    Article  Google Scholar 

  • Dagbovie, A. S., & Sherratt, J. A. (2013, accepted). Absolute stability and dynamical stabilisation in predator–prey systems. J. Math. Biol. doi:10.1007/s00285-013-0672-8.

  • Dunbar, S. R. (1986). Traveling waves in diffusive predator–prey equations—periodic orbits and point-to-periodic heteroclinic orbits. SIAM J. Appl. Math., 46, 1057–1078.

    Article  MathSciNet  MATH  Google Scholar 

  • Fausch, K. D., Torgersen, C. E., Baxter, C. V., & Li, H. W. (2002). Landscapes to riverscapes: bridging the gap between research and conservation of stream fishes. Bioscience, 52, 483–498.

    Article  Google Scholar 

  • Fox, P. J., & Proctor, M. R. E. (1998). Effects of distant boundaries on pattern forming instabilities. Phys. Rev. E, 57, 491–494.

    Article  Google Scholar 

  • Fraile, J. M., & Sabina, J. C. (1989). General conditions for the existence of a critical point–periodic wave front connection for reaction–diffusion systems. Nonlinear Anal.-Theor., 13, 767–786.

    Article  MathSciNet  MATH  Google Scholar 

  • Gaylord, B., & Gaines, S. D. (2000). Temperature or transport? Range limits in marine species mediated solely by flow. Am. Nat., 155, 769–789.

    Article  Google Scholar 

  • Grimm, V., & Wissel, C. (1997). Babel, or the ecological stability discussions: an inventory and analysis of terminology and a guide for avoiding confusion. Oecologia, 109, 323–334.

    Article  Google Scholar 

  • Hauzy, C., Hulot, F. D., & Gins, A. (2007). Intra and interspecific density-dependent dispersal in an aquatic prey–predator system. J. Anim. Ecol., 76, 552–558.

    Article  Google Scholar 

  • Hilker, F. M., & Lewis, M. A. (2010). Predator-prey systems in streams and rivers. Theor. Ecol., 3, 175–193.

    Article  Google Scholar 

  • Huerre, P., & Monkewitz, P. A. (1990). Local and global instabilities in spatially developing flows. Annu. Rev. Fluid Mech., 22, 473–537.

    Article  MathSciNet  MATH  Google Scholar 

  • Huisman, J., & Weissing, F. J. (1999). Biodiversity of plankton by species oscillations and chaos. Nature, 402, 407–410.

    Article  Google Scholar 

  • Klausmeier, C. A. (1999). Regular and irregular patterns in semiarid vegetation. Science, 284, 1826–1828.

    Article  Google Scholar 

  • Kopell, N., & Howard, L. N. (1973). Plane wave solutions to reaction–diffusion equations. Stud. Appl. Math., 52, 291–328.

    Article  MathSciNet  MATH  Google Scholar 

  • Lee, S.-H. (2012). Effects of uniform rotational flow on predator–prey system. Physica A, 391, 6008–6015.

    Article  Google Scholar 

  • Levine, J. M. (2003). A patch modeling approach to the community-level consequences of directional dispersal. Ecology, 84, 1215–1224.

    Article  Google Scholar 

  • Lutscher, F., Lewis, M. A., & McCauley, E. (2006). Effects of heterogeneity on spread and persistence in rivers. Bull. Math. Biol., 68, 2129–2160.

    Article  MathSciNet  MATH  Google Scholar 

  • Malchow, H. (2000). Motional instabilities in predator–prey systems. J. Theor. Biol., 204, 639–647.

    Article  Google Scholar 

  • Malchow, H., & Petrovskii, S. V. (2002). Dynamical stabilization of an unstable equilibrium in chemical and biological systems. Math. Comput. Model., 36, 307–319.

    Article  MathSciNet  MATH  Google Scholar 

  • Merchant, S. M., & Nagata, W. (2010). Wave train selection behind invasion fronts in reaction–diffusion predator–prey models. Physica D, 239, 1670–1680.

    Article  MathSciNet  MATH  Google Scholar 

  • Merchant, S. M., & Nagata, W. (2011). Instabilities and spatiotemporal patterns behind predator invasions with nonlocal prey competition. Theor. Popul. Biol., 80, 289–297.

    Article  MATH  Google Scholar 

  • Morozov, A. Yu., Petrovskii, S. V, & Li, B.-L. (2006). Spatiotemporal complexity of the patchy invasion in a predator–prey system with the Allee effect. J. Theor. Biol., 238, 18–35.

    Article  MathSciNet  Google Scholar 

  • Nauman, E. B. (2008). Chemical reactor design, optimization, and scaleup (2nd ed.). Hoboken: Wiley.

    Book  Google Scholar 

  • Owen, M. R., & Lewis, M. A. (2001). How predation can slow, stop or reverse a prey invasion. Bull. Math. Biol., 63, 655–684.

    Article  MATH  Google Scholar 

  • Perumpanani, A. J., Sherratt, J. A., & Maini, P. K. (1995). Phase differences in reaction–diffusion–advection systems and applications to morphogenesis. IMA J. Appl. Math., 55, 19–33.

    Article  MathSciNet  MATH  Google Scholar 

  • Petrovskii, S. V., & Malchow, H. (2000). Critical phenomena in plankton communities: kiss model revisited. Nonlinear Anal., Real World Appl., 1, 37–51.

    Article  MathSciNet  MATH  Google Scholar 

  • Petrovskii, S., Li, B.-L., & Malchow, H. (2004). Transition to spatiotemporal chaos can resolve the paradox of enrichment. Ecol. Complex., 1, 37–47.

    Article  Google Scholar 

  • Rademacher, J. D. M. (2006). Geometric relations of absolute and essential spectra of wave trains. SIAM J. Appl. Dyn. Syst., 5, 634–649.

    Article  MathSciNet  MATH  Google Scholar 

  • Rademacher, J. D. M., Sandstede, B., & Scheel, A. (2007). Computing absolute and essential spectra using continuation. Physica D, 229, 166–183.

    Article  MathSciNet  MATH  Google Scholar 

  • Rietkerk, M., Boerlijst, M. C., van Langevelde, F., HilleRisLambers, R., van de Koppel, J., Prins, H. H. T., & de Roos, A. (2002). Self-organisation of vegetation in arid ecosystems. Am. Nat., 160, 524–530.

    Article  Google Scholar 

  • Rosenzweig, M. L., & MacArthur, R. H. (1963). Graphical representation and stability conditions of predator–prey interactions. Am. Nat., 97, 209–223.

    Article  Google Scholar 

  • Rovinsky, A. B., & Menzinger, M. (1992). Chemical instability induced by a differential flow. Phys. Rev. Lett., 69, 1193–1196.

    Article  Google Scholar 

  • Sandstede, B. (2002). Stability of travelling waves. In B. Fiedler (Ed.), Handbook of dynamical systems II (pp. 983–1055). Amsterdam: North-Holland.

    Chapter  Google Scholar 

  • Sandstede, B., & Scheel, A. (2000a). Absolute versus convective instability of spiral waves. Phys. Rev. E, 62, 7708–7714.

    Article  MathSciNet  Google Scholar 

  • Sandstede, B., & Scheel, A. (2000b). Absolute and convective instabilities of waves on unbounded and large bounded domains. Physica D, 145, 233–277.

    Article  MathSciNet  MATH  Google Scholar 

  • Scheuring, I., Károlyi, G., Péntek, A., Tel, T., & Toroczkai, Z. (2000). A model for resolving the plankton paradox: coexistence in open flows. Freshw. Biol., 45, 123–132.

    Article  MATH  Google Scholar 

  • Sherratt, J. A. (2005). An analysis of vegetation stripe formation in semi-arid landscapes. J. Math. Biol., 51, 183–197.

    Article  MathSciNet  MATH  Google Scholar 

  • Sherratt, J. A. (2010). Pattern solutions of the Klausmeier model for banded vegetation in semi-arid environments I. Nonlinearity, 23, 2657–2675.

    Article  MathSciNet  MATH  Google Scholar 

  • Sherratt, J. A., Lewis, M. A., & Fowler, A. C. (1995). Ecological chaos in the wake of invasion. Proc. Natl. Acad. Sci. USA, 92, 2524–2528.

    Article  MATH  Google Scholar 

  • Sherratt, J. A., Smith, M. J., & Rademacher, J. D. M. (2009). Locating the transition from periodic oscillations to spatiotemporal chaos in the wake of invasion. Proc. Natl. Acad. Sci. USA, 106, 10890–10895.

    Article  MATH  Google Scholar 

  • Smith, M. J., & Sherratt, J. A. (2009). Propagating fronts in the complex Ginzburg–Landau equation generate fixed-width bands of plane waves. Phys. Rev. E, 80, 046209.

    Article  Google Scholar 

  • Smith, M. J., Sherratt, J. A., & Lambin, X. (2008). The effects of density-dependent dispersal on the spatiotemporal dynamics of cyclic populations. J. Theor. Biol., 254, 264–274.

    Article  MathSciNet  Google Scholar 

  • Smith, M. J., Rademacher, J. D. M., & Sherratt, J. A. (2009). Absolute stability of wavetrains can explain spatiotemporal dynamics in reaction–diffusion systems of lambda–omega type. SIAM J. Appl. Dyn. Syst., 8, 1136–1159.

    Article  MathSciNet  MATH  Google Scholar 

  • Suslov, S. A. (2001). Searching convective/absolute instability boundary for flows with fully numerical dispersion relation. Comput. Phys. Commun., 142, 322–325.

    Article  MathSciNet  MATH  Google Scholar 

  • Suslov, S. A. (2006). Numerical aspects of searching convective/absolute instability transition. J. Comp. Physiol., 212, 188–217.

    Article  MathSciNet  MATH  Google Scholar 

  • Suslov, S. A. (2009). Analysis of instability patterns in non-Boussinesq mixed convection using a direct numerical evaluation of disturbance integrals. Comput. Fluids, 38, 590–601.

    Article  MathSciNet  Google Scholar 

  • Suslov, S. A., & Paolucci, S. (2004). Stability of non-Boussinesq convection via the complex Ginzburg–Landau model. Fluid Dyn. Res., 35, 159–203.

    Article  MathSciNet  MATH  Google Scholar 

  • Tobias, S. M., Proctor, M. R. E., & Knobloch, E. (1998). Convective and absolute instabilities of fluid flows in finite geometries. Physica D, 113, 43–72.

    Article  MathSciNet  MATH  Google Scholar 

  • Turchin, P. (2003). Complex population dynamics. A Theoretical/Empirical synthesis. Princeton: Princeton University Press.

    MATH  Google Scholar 

  • van Saarloos, W. (2003). Front propagation into unstable states. Phys. Rep., 386, 29–222.

    Article  MATH  Google Scholar 

  • Wheeler, P., & Barkley, D. (2006). Computation of spiral spectra. SIAM J. Appl. Dyn. Syst., 5, 157–177.

    Article  MathSciNet  MATH  Google Scholar 

  • Wieters, E. A., Gaines, S. D., Navarrete, S. A., Blanchette, C. A., & Menge, B. A. (2008). Scales of dispersal and the biogeography of marine predator–prey interactions. Am. Nat., 171, 405–417.

    Article  Google Scholar 

  • Worledge, D., Knobloch, E., Tobias, S., & Proctor, M. (1997). Dynamo waves in semi-infinite and finite domains. Proc. R. Soc. Lond. A, 453, 119–143.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

J.A.S. acknowledges discussions with Leonid Brevdo (Louis Pasteur University, Strasbourg), Jens Rademacher (CWI, Amsterdam), Björn Sandstede (Brown University) and Matthew Smith (Microsoft Research, Cambridge). A.S.D. was supported by the Centre for Analysis and Nonlinear PDEs funded by the UK EPSRC grant EP/E03635X and the Scottish Funding Council. F.M.H. acknowledges discussions with Mark Lewis (University of Alberta), Sergei Petrovskii (University of Leicester) and Frithjof Lutscher (University of Ottawa).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jonathan A. Sherratt.

Appendix: Examples of Calculating Branch Points

Appendix: Examples of Calculating Branch Points

In this Appendix, we show how to calculate the branch points for the Rosenzweig–MacArthur model (1a), (1b) and the Klausmeier model (2a), (2b). We present these calculations in some detail, with the aim of providing templates that readers can follow when performing corresponding calculations for their own models.

1.1 A.1 Branch Points for the Rosenzweig–MacArthur Model

Recall that the Rosenzweig–MacArthur model (1a), (1b) has a unique homogeneous co-existence steady state (h s ,p s ) where h s =1/(μ) and p s =(1−h s )(1+μh s )/μ. We begin by linearising (1a), (1b) about (h s ,p s ) giving

$$\begin{aligned} \tilde{p}_t &=\alpha\tilde{p}+\beta\tilde{h}+c\tilde{p}_x+d \tilde {p}_{xx}, \\ \tilde{h}_t &=\gamma\tilde{p}+\delta\tilde{h}+c \tilde{h}_x+\tilde{h}_{xx}, \end{aligned}$$

where \(\tilde{p}=p-p_{s}\), \(\tilde{h}=h-h_{s}\), and α, β, γ, δ are coefficients from linearisation and are given by

$$\begin{aligned} \alpha&=\frac{\mu h_s}{b(1+\mu h_s)}-\frac{1}{ab}, \\ \beta&=\frac{\mu p_s}{b(1+\mu h_s)^2}, \\ \gamma&=1-2h_s-\frac{\mu p_s}{(1+\mu h_s)^2}, \\ \delta&=-\frac{\mu h_s}{1+\mu h_s} . \end{aligned}$$

Substituting \((\tilde{p},\tilde{h})=(\bar{p},\bar{h})\exp (ikx+\lambda t)\) into (1a), (1b) and requiring \(\bar{p}\) and \(\bar{h}\) to be non-zero gives the dispersion relation

$$\begin{aligned} \mathcal{D}(\lambda,k) =& dk^4-cik^3(d+1) -k^2\bigl(\alpha-\lambda+d\gamma-d\lambda+c^2\bigr) \\ &{}+cik(\gamma+\alpha-2\lambda)+ (\alpha-\lambda) (\gamma-\lambda)-\delta\beta=0 . \end{aligned}$$
(9)

Branch points are double roots of the dispersion relation for k, and satisfy (9) and also

$$ \begin{aligned} 0&=\partial\mathcal{D}/\partial k \\ &= 4dk^3-3k^2(1+d)ci-2k \bigl(\gamma+d\alpha-(1+d)\lambda\bigr)+c^2+ ci(\alpha+\gamma-2 \lambda)\quad \Rightarrow \\ \lambda&= \frac{4dk^3-3cidk^2-2dk\gamma+\alpha ci-2c^2k-3cik^2-2\alpha k+ci\gamma}{-2(k+dk-ci)} . \end{aligned} $$
(10)

Substituting (10) into (9) gives the following hexic polynomial in k:

$$\begin{aligned} & \bigl[-4d(d-1)^2 \bigr]k^6+ \bigl[2ci(d+1) (d-1)^2 \bigr]k^5 + \bigl[(d-1) \bigl(c^2d+8d\alpha-8d\gamma-c^2\bigr) \bigr]k^4 \\ &\quad{} + \bigl[4ci(-\alpha+\gamma) (d-1) (d+1) \bigr]k^3 \\ &\quad{} + \bigl[2c^2(d-1) (\gamma-\alpha)-4\beta\delta(d+1)^2 -4d( \alpha-\gamma)^2 \bigr]k^2 \\ &\quad{} + \bigl[2i\bigl(4\beta\delta+\alpha^2-2\alpha\gamma + \gamma^2\bigr)c(d+1) \bigr]k+c^2\bigl(4\beta\delta + \alpha^2-2\alpha\gamma+\gamma^2\bigr)=0 . \end{aligned}$$
(11)

We must now proceed numerically and we fix a=1.3, b=4.0, c=−1, d=2, and μ=9. These parameter values satisfy μ>μ crit, so that the coexistence steady state is unstable. Substituting these values into (11), we obtain six roots for k, two real and two pairs of complex conjugates. We then substitute each into (10) to find the corresponding value of λ. To determine whether these branch points belong to the absolute spectrum, we substitute each λ value into (9) and solve for k, giving the repeated roots found from (11) and two others.

Branch point k=0.676i.:

Substituting this value of k into (10) gives λ=1.380. Substituting this value of λ back into (9) gives a quartic polynomial for k whose roots are −1.727i, −1.125i, 0.676i, 0.676i. Recall that a branch point is in the absolute spectrum if the repeated roots are k 2 and k 3, when the roots k 1, k 2, k 3 and k 4 of (9) are labelled in increasing order of their imaginary parts. In this case, the repeated roots are k 3 and k 4 so that the branch point is not in the absolute spectrum.

Branch point k=−0.473−0.013i.:

Substituting this value of k into (10) gives λ=−0.255+0.483i. Substituting this value of λ back into (9) gives a quartic polynomial for k whose roots are 0.570−0.943i, 0.377−0.530i, −0.473−0.013i, −0.473−0.013i Therefore, the repeated roots are k 3 and k 4 so that the branch point is not in the absolute spectrum.

Branch point k=0.473−0.013i.:

Substituting this value of k into (10) gives λ=−0.255−0.483i. Substituting this value of λ back into (9) gives a quartic polynomial for k whose roots are −0.570−0.943i, −0.377−0.530i, 0.473−0.013i, 0.473−0.013i. Therefore, the repeated roots are k 3 and k 4 so that the branch point is not in the absolute spectrum.

Branch point k=0.001−0.334i.:

Substituting this value of k into (10) gives λ=−0.110−0.167i. Substituting this value of λ back into (9) gives a quartic polynomial for k whose roots are −0.314−0.816i, 0.001−0.334i, 0.001−0.334i, 0.312−0.0165i. Therefore, the repeated roots are k 2 and k 3 so that the branch point is in the absolute spectrum.

Branch point k=−0.001−0.334i.:

Substituting this value of k into (10) gives λ=−0.110+0.167i. Substituting this value of λ back into (9) gives a quartic polynomial for k whose roots are 0.314−0.816i, −0.001−0.334i, −0.001−0.334i, −0.312−0.0165i. Therefore, the repeated roots are k 2 and k 3 so that the branch point is in the absolute spectrum.

Branch point k=−0.732i.:

Substituting this value of k into (10) gives λ=0.000. Substituting this value of λ back into (9) gives a quartic polynomial for k whose roots are −0.732i, −0.732i, 0.160−0.017i, −0.160−0.017i. Therefore, the repeated roots are k 1 and k 2 so that the branch point is not in the absolute spectrum.

Therefore, of the six branch points, two are in the absolute spectrum, with the corresponding eigenvalues being −0.110±0.167i. Since these eigenvalues have negative real parts, the steady state (h s ,p s ) is absolutely stable. To determine whether the convective instability is of transient or remnant type, it is necessary to calculate the absolute spectrum. This can be done via numerical continuation of the generalised absolute spectrum, using the six branch points listed above as starting points, as discussed in Sect. 5. This shows that the branch points are the most unstable points in the absolute spectrum, so that the steady state has a transient convective instability.

1.2 A.2 Branch Points for the Klausmeier Model

For all parameters, the Klausmeier model (2a), (2b) has a “desert” steady state m=0, w=A. When A≥2B, there are two further steady states (m ±,w ±) where

$$m_\pm=\frac{2B}{A\pm\sqrt{A^2-4B^2}}, \qquad w_\pm=\frac{A\pm\sqrt{A^2-4B^2}}{2} . $$

Ecologically realistic values of B are relatively small, and in particular satisfy B<2 (Klausmeier 1999; Rietkerk et al. 2002). Under this constraint, (m ,w ) is stable as a solution of the kinetics odes, although it can be destabilised by the diffusion and advection terms, leading to spatial patterns (Klausmeier 1999; Sherratt 2005, 2010). However, (m +,w +) is unstable as a solution of the kinetic odes, and we will consider the nature of its instability as a solution of the pdes (2a), (2b).

We begin by linearising (2a), (2b) about (m +,w +), giving

$$\begin{aligned} \tilde{m}_t =&\tilde{\alpha}\tilde{m}+\tilde{\beta}\tilde{w}+\tilde{m}_{xx}, \end{aligned}$$
(12a)
$$\begin{aligned} \tilde{w}_t =&\tilde{\gamma}\tilde{m}+\tilde{\delta}\tilde{w}+\nu\tilde{w}_x, \end{aligned}$$
(12b)

where \(\tilde{m}=m-m_{+}\), \(\tilde{w}=w-w_{+}\), and the linear coefficients α, β, γ, δ are given by

$$\begin{aligned} \tilde{\alpha} =&B, \end{aligned}$$
(13a)
$$\begin{aligned} \tilde{\beta} =&\frac{A-\sqrt{A^2-4B^2}}{A+\sqrt{A^2-4B^2}}, \end{aligned}$$
(13b)
$$\begin{aligned} \tilde{\gamma} =&-2B, \end{aligned}$$
(13c)
$$\begin{aligned} \tilde{\delta} =&\frac{-2A}{A+\sqrt{A^2-4B^2}} . \end{aligned}$$
(13d)

Substituting \((\tilde{m},\tilde{w} )= (\overline{m},\overline{w} ) \exp(ikx+\lambda t)\) into (12a), (12b) and requiring \(\overline{m}\) and \(\overline{w}\) to be non-zero gives the dispersion relation

$$ {\tilde{\mathcal{D}}}(\lambda,k)= \lambda^2+\lambda \bigl(k^2-\tilde{\alpha}-ik\nu -\tilde{ \delta} \bigr) +\bigl(\tilde{\alpha}-k^2\bigr) (ik\nu+\tilde{\delta})- \tilde{\beta}\tilde {\gamma}=0 . $$
(14)

Branch points are double roots (for k) of the dispersion relation, and satisfy (14) and also

$$\begin{aligned} \begin{aligned} 0&=\partial{\tilde{\mathcal{D}}}/\partial k= \lambda (2k-i\nu ) - \bigl(3ik^2\nu+2\tilde{\delta}k-i\nu\tilde{\alpha} \bigr)\quad \Rightarrow \\ \lambda&= \bigl(3ik^2\nu+2\tilde{\delta}k-i\nu \tilde{\alpha} \bigr) / (2k-i\nu ) . \end{aligned} \end{aligned}$$
(15)

Substituting (15) into (14) gives a quintic polynomial in k:

$$\begin{aligned} &\bigl(3ik^2\nu+2\tilde{\delta}k-i\nu\tilde{\alpha} \bigr)^2 + (2k-i\nu ) \bigl(3ik^2\nu+2\tilde{\delta}k-i\nu \tilde{\alpha} \bigr) \bigl(k^2-\tilde{\alpha}-ik\nu-\tilde{\delta} \bigr) \\ &\quad{} + (2k-i\nu )^2 \bigl[\bigl(\tilde{\alpha}-k^2\bigr) (ik \nu+\tilde {\delta}) -\tilde{\beta}\tilde{\gamma} \bigr]=0 . \end{aligned}$$
(16)

We must now proceed numerically, and we will fix the parameter values to be A=2, B=0.5 and ν=20. These parameters satisfy the condition A>2B, but otherwise they are chosen arbitrarily. The value ν=20 is too small for ecological realism: The formula for the dimensionless parameter ν involves the ratio of the advection rate of water and the (square root of the) plant diffusion coefficient (Klausmeier 1999; Sherratt 2005), so that ν is relatively large, with Klausmeier’s (1999) estimate being 182.5. However, we use the smaller value to improve the clarity of the numerical calculations. Substituting the parameter values into (16) gives five distinct roots for k, two complex and three pure imaginary. For each, we substitute into (15) to find the corresponding value of λ. We then substitute this value of λ into (14) to determine whether the branch point is in the absolute spectrum. We performed all of the various calculations using the software package maple with 20 decimal places, but for clarity we give results to 3 decimal places.

Branch point k=−0.099+i0.012.:

Substituting this value of k into (15) gives λ=0.473+i0.022. Substituting this value of λ back into (14) gives a cubic polynomial for k whose roots are −0.099+i0.012, −0.099+i0.012, 0.199−i0.101. Recall from Sect. 5 that the branch point is in the absolute spectrum if the repeated roots are k 2 and k 3, when the roots k 1, k 2, k 3 of (14) are labelled in increasing order of their imaginary parts. Therefore, in this case the branch point is in the absolute spectrum.

Branch point k=−i19.950.:

Substituting this value of k into (15) gives λ=398.112. Substituting this value of λ back into (14) gives a cubic polynomial for k whose roots are −i19.950, −i19.950 and i19.94. Therefore, the repeated roots are k 1 and k 2, so that this branch point is not in the absolute spectrum.

Branch point k=−i19.892.:

Substituting this value of k into (15) gives λ=396.587. Substituting this value of λ back into (14) gives a cubic polynomial for k whose roots are −i19.892, −i19.892 and i19.902. Therefore, the repeated roots are k 1 and k 2, so that this branch point is not in the absolute spectrum.

Branch point k=−i0.181.:

Substituting this value of k into (15) gives λ=0.569. Substituting this value of λ back into (14) gives a cubic polynomial for k whose roots are −i0.181, −i0.181, and i0.281. Therefore, the repeated roots are k 1 and k 2, so that this branch point is not in the absolute spectrum.

Branch point k=0.099+i0.012.:

Substituting this value of k into (15) gives λ=0.473−i0.022. Substituting this value of λ back into (14) gives a cubic polynomial for k whose roots are −0.199−i0.101, 0.099+i0.012, and 0.099+i0.012. Therefore, the repeated roots are k 2 and k 3, so that this branch point is in the absolute spectrum.

Therefore, of the five branch points, two are in the absolute spectrum, with the corresponding eigenvalues being 0.473±i0.022. Since these eigenvalues have positive real part, the steady state (m +,w +) is absolutely unstable.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Sherratt, J.A., Dagbovie, A.S. & Hilker, F.M. A Mathematical Biologist’s Guide to Absolute and Convective Instability. Bull Math Biol 76, 1–26 (2014). https://doi.org/10.1007/s11538-013-9911-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-013-9911-9

Keywords

Navigation