Skip to main content
Log in

Ecological drivers of stability and instability in marine ecosystems

  • Original Paper
  • Published:
Theoretical Ecology Aims and scope Submit manuscript

Abstract

A stability analysis of the steady state of marine ecosystems is described. The study was motivated by the approximate invariance of biomass in logarithmic size intervals, which is widely observed in marine ecosystems. This invariance is recovered as the steady state of dynamic models of size spectra, which, unlike traditional species-based models of food webs, explicitly account for the mass gained by an individual organism when it eats a prey item. Little is known about the ecological conditions affecting the stability of the steady state, and a new method is developed to examine this. The results show that stability is enhanced by: (a) decreasing the mean predator-to-prey mass ratio (PPMR), (b) increasing the diet breadth of predators, (c) increasing the strength of intrinsic mortality relative to predation mortality, (d) increasing the biomass conversion efficiency. When perturbed from steady state, size spectra develop a wave-like shape, with an average wavelength especially sensitive to the mean PPMR. These waves move from small to large body size at an average speed which depends on the rate of growth of organisms. In contrast to traditional food web models, stability is enhanced as connectance (diet breadth) increases and as food chain length is increased by reducing the PPMR.

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

Similar content being viewed by others

References

  • Allesina S, Pascual M (2008) Network structure, predator-prey modules, and stability in large food webs. Theor Ecol 1:55–64

    Article  Google Scholar 

  • Andersen KH, Beyer JE (2006) Asymptotic size determines species abundance in the marine size spectrum. Am Nat 168:54–61

    Article  Google Scholar 

  • Andersen KH, Beyer JE, Pedersen M, Andersen NG, Gislason H (2008) Life-history constraints on the success of the many small eggs reproductive strategy. Theor Popul Biol 73:490–497

    Article  PubMed  CAS  Google Scholar 

  • Anderson CNK, Hsieh C-h, Sandin SA, Hewitt R, Hollowed A, Beddington J, May RM, Sugihara G (2008) Why fishing magnifies fluctuations in fish abundance. Nature 452:835–839

    Article  PubMed  CAS  Google Scholar 

  • Arino O, Shin Y-J, Mullon C (2004) A mathematical derivation of size spectra in fish populations. Comptes Rendus Biol 327:245–254

    Article  Google Scholar 

  • Barnes C, Bethea DM, Brodeur RD, Spitz J, Ridoux V, Pusineri C, Chase BC, Hunsicker ME, Juanes F, Kellermann A, Lancaster J, Ménard F, Bard F-X, Munk P, Pinnegar JK, Scharf FS, Rountree RA, Stergiou KI, Sassa C, Sabates A, Jennings S (2008) Predator and prey body sizes in marine food webs. Ecology 89:881

    Article  Google Scholar 

  • Benoît E, Rochet M-J (2004) A continuous model of biomass size spectra governed by predation and the effects of fishing on them. J Theor Biol 226:9–21

    Article  PubMed  Google Scholar 

  • Blanchard JL (2008) The dynamics of size-structured ecosystems. Ph.D. thesis, University of York

  • Blanchard JL, Jennings S, Law R, Castle MD, McCloghrie P, Rochet M-J, Benoît E (2009) How does abundance scale with body size in coupled size-structured food webs? J Anim Ecol 78:270–280

    Article  PubMed  Google Scholar 

  • Blanchard JL, Law R, Castle MD, Jennings S (2010) Coupled energy pathways and the resilience of size-structured food webs. Theor Ecol 4:289–300

    Article  Google Scholar 

  • Blueweiss L, Fox H, Kudzma V, Nakashima D, Peters R, Sams S (1978) Relationships between body size and some life history parameters. Oecologia 37:257–272

    Article  Google Scholar 

  • Boudreau PR, Dickie LM (1992) Biomass spectra of aquatic ecosystems in relation to fisheries yield. Can J Fish Aquat Sci 49:1528–1538

    Article  Google Scholar 

  • Brose U, Williams RJ, Martinez ND (2006) Allometric scaling enhances stability in complex food webs. Ecol Lett 9:1228–1236

    Article  PubMed  Google Scholar 

  • Camacho J, Solé RV (2001) Scaling in ecological size spectra. Europhys Lett 55:774–780

    Article  CAS  Google Scholar 

  • Capitán JA, Delius GW (2010) Scale-invariant model of marine population dynamics. Phys Rev E 81:061901

    Article  Google Scholar 

  • Cushing DH (1969) The regularity of the spawning season of some fishes. J Cons Int Explor Mer 33:81–92

    Google Scholar 

  • Datta S, Delius GW, Law R (2010) A jump-growth model for predator–prey dynamics: derivation and application to marine ecosystems. Bull Math Biol 72:1361–1382

    Article  PubMed  Google Scholar 

  • Datta S, Delius GW, Law R, Plank MJ (2011) Stability analysis of marine ecosystems using the deterministic jump-growth equation. J Math Biol doi:10.1007/s00285-010-0387-z

    PubMed  Google Scholar 

  • Dunne JA, Williams RJ, Martinez ND (2002) Network structure and biodiversity loss in food webs: robustness increases with connectance. Ecol Lett 5:558–567

    Article  Google Scholar 

  • Emmerson MC, Raffaelli D (2004) Predator–prey body size, interaction strength and the stability of a real food web. J Anim Ecol 73:399–409

    Article  Google Scholar 

  • Fenchel T (1974) Intrinsic rate of natural increase: the relationship with body size. Oecologia 14:317–326

    Article  Google Scholar 

  • Frank KT, Petrie B, Choi JS, Leggett WC (2005) Trophic cascades in a formerly cod-dominated ecosystem. Science 308:1621–1623

    Article  PubMed  CAS  Google Scholar 

  • Gasol JM, Guerro R, Pedrós-Alió C (1991) Seasonal variations in size structure and procaryotic dominance in sulfurous Lake Cisó. Limnol Oceanog 36:860–872

    Article  Google Scholar 

  • Gaedke U (1992) The size distribution of plankton biomass in a large lake and its seasonal variability. Limnol Oceanog 37:1202–1220

    Article  Google Scholar 

  • Guckenheimer J, and Holmes P (1983) Nonlinear oscillations, dynamical systems and bifurcations of vector fields. Springer, New York

    Google Scholar 

  • Hartvig M, Andersen KH, Beyer JE (2011) Food web framework for size-structured populations. J Theor Biol 272:113–122

    Article  PubMed  Google Scholar 

  • Heath MR (1995) Size spectrum dynamics and the planktonic ecosystem of Loch Linnhe. ICES J Mar Sci 52:627–642

    Article  Google Scholar 

  • Hsieh C-h, Reiss CS, Hunter JR, Beddington JR, May RM, Sugihara G (2006) Fishing elevates variability in the abundance of exploited species. Nature 443:859–862

    Article  PubMed  CAS  Google Scholar 

  • Huete–Ortega M, Marañón, E, Varela M, Bode A (2010) General patterns in the size scaling of phytoplankton abundance in coastal waters during a 10-year time series. J Plankton Res 32:1–14

    Article  Google Scholar 

  • Jennings S, Pinnegar JK, Polunin NVC, Boon TW (2001) Weak cross-species relationships between body size and trophic level belie powerful size-based trophic structuring in fish communities. J Anim Ecol 70:934–944

    Article  Google Scholar 

  • Jennings S, Warr KJ (2003) Smaller predator–prey body size ratios in longer food chains. Proc Roy Soc B 270:1413–1417

    Article  Google Scholar 

  • Kerr SR, Dickie LM (2001) The biomass spectrum: a predator–prey theory of aquatic production. Columbia University Press, New York

    Google Scholar 

  • Knowlton N (2004) Multiple “stable” states and the conservation of marine ecosystems. Prog Oceanogr 60:387–396

    Article  Google Scholar 

  • Law R, Plank MJ, James A, Blanchard JL (2009) Size-spectra dynamics from stochastic predation and growth of individuals. Ecology 90:802–811

    Article  PubMed  Google Scholar 

  • Maury O, Faugeras B, Shin Y-J, Poggiale J-C, Ari TB, Marsac F (2007) Modelling environmental effects on the size-structured energy flow through marine ecosystems. Part 1: the model. Prog Oceanogr 74:479–499

    Article  Google Scholar 

  • May RM (1972) Will a large complex system be stable. Nature 238:413–414

    Article  PubMed  CAS  Google Scholar 

  • McKane AJ, Newman TJ (2005) Predator–prey cycles from resonant amplification of demographic stochasticity. Phys Rev Lett 94:218102

    Article  PubMed  CAS  Google Scholar 

  • McKendrick AG (1926) Applications of mathematics to medical problems. Proc Edinb Math Soc 40:98–130

    Google Scholar 

  • Neutel A-M, Heesterbeek JAP, de Ruiter PC (2002) Stability in real food webs: weak links in long loops. Science 296:1120–1123

    Article  PubMed  CAS  Google Scholar 

  • Pimm SL, Lawton JH (1977) Number of trophic levels in ecological communities. Nature 268:329–331

    Article  Google Scholar 

  • Platt T and Denman K (1978) The structure of pelagic marine ecosystems. Rapp P-v Réun Cons Int Explor Mer 173:60–65

    Google Scholar 

  • Platt T, Fuentes-Yaco C, Frank KT (2003) Spring algal bloom and larval fish survival. Nature 423:398–399

    Article  PubMed  CAS  Google Scholar 

  • Pope JG, Shepherd JG and Webb J (1994) Successful surf-riding on size spectra: the secret of survival in the sea. Phil Trans R Soc B 343:41–49

    Article  Google Scholar 

  • Post DM, Pace ML, Hairston NG Jr (2000) Ecosystem size determines food-chain length in lakes. Nature 405:1047–1049

    Article  PubMed  CAS  Google Scholar 

  • Post DM (2007) Testing the productive-space hypothesis: rational and power. Oecologia 153:973–984

    Article  PubMed  Google Scholar 

  • Quinones RA, Platt T, Rodríguez J (2003) Patterns of biomass-size spectra from oligotrophic waters of the Northwest Atlantic. Prog Oceanogr 57:405–427

    Article  Google Scholar 

  • Rochet M-J and Benoît E (2011) Fishing destabilizes the biomass flow in the marine size spectrum. Proc R Soc B. doi:10.1098/rspb.2011.0893

    PubMed  Google Scholar 

  • San Martin E, Irigoien X, Harris RP, López-Urrutia Á, Zubkov MZ, Heywood JL (2006) Variation in the transfer of energy in marine plankton along a productivity gradient in the Atlantic Ocean. Limnol Oceanogr 51:2084–2091

    Article  Google Scholar 

  • Sheldon RW, Parsons TR (1967) A continuous size spectrum for particulate matter in the sea. J Fish Res Board Can 24:909–915

    Article  Google Scholar 

  • Sheldon RW, Prakash A, Sutcliffe WH Jr (1972) The size distribution of particles in the ocean. Limnol Oceanogr 17:327–340

    Article  Google Scholar 

  • Sheldon RW, Sutcliffe WH Jr, Paranjape MA (1977) Structure of pelagic food chain and relationship between plankton and fish production. J Fish Res Board Can 34:2344–2353

    Article  Google Scholar 

  • Silvert W, Platt T (1978) Energy flux in the pelagic ecosystem: a time-dependent equation. Limnol Oceanogr 23:813–816

    Article  Google Scholar 

  • Silvert W, Platt T (1980) Dynamic energy-flow model of the particle size distribution in pelagic ecosystems. In: Kerfoot WC (ed) Evolution and ecology of zooplankton communities. University Press of New England, Hanover, pp 754–763

    Google Scholar 

  • Sterner RW, Bajpai A, Adams T (1997) The enigma of food chain length: absence of theoretical evidence for dynamic constraints. Ecol 78:2258–2262

    Article  Google Scholar 

  • Ursin E (1973) On the prey size preferences of cod and dab. Medd Dan Fisk Havunders 7:85–98

    Google Scholar 

  • von Foerster H (1959) Some remarks on changing populations. In: Stohlman JF (ed) The kinetics of cellular proliferation. Grune and Stratton, New York, pp 382–407

    Google Scholar 

  • Ware DM (1978) Bioenergetics of pelagic fish: theoretical change in swimming speed and ration with body size. J Fish Res Board Can 35:220–228

    Article  Google Scholar 

  • Zhou M, Huntley ME (1997) Population dynamics theory of plankton based on biomass spectra. Mar Ecol Prog Ser 159:61–73

    Article  Google Scholar 

  • Zhou M, Tande KS, Zhu Y, Basedow S (2009) Productivity, trophic levels and size spectra of zooplankton in northern Norwegian shelf regions. Deep Sea Res II 56:1934–1944

    Article  Google Scholar 

  • Zhou M, Carlotti F, Zhu Y (2010) A size-spectrum zooplankton closure model for ecosystem modelling J Plankton Res 32:1147–1165

    Article  Google Scholar 

Download references

Acknowledgements

This research was supported by the RSNZ Marsden fund, grant number 08-UOC-034. We thank Gustav Delius for his contributions to the analytical calculations and derivation of the convolution kernel. We also thank David Wall for illuminating discussions and José Capitán and Julia Blanchard for comments on an earlier version of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael John Plank.

Appendix

Appendix

Linear evolution equation for perturbations

We define the following dimensionless variables:

$$ x=\ln\left(\frac{w}{w_0}\right), \qquad \bar t=\frac{t}{t_0}, \qquad v(x) = e^{(\alpha+\rho+1) x} \frac{\phi(w)}{\phi(w_0)}. $$

If t 0 is chosen to be \(1/(Aw_0^{\alpha+1}\phi(w_0))\), then the dimensionless search rate constant \(\bar A\) is equal to 1. In these new variables, and with the constraint that ξ must equal ρ if intrinsic mortality is present (i.e. if μ > 0), the deterministic jump-growth equation (see Eqs. 1 and 2 in the main text) is

$$\begin{array}{rll} \frac{\partial }{\partial \bar t}v(x) &=& e^{-\rho x} \Bigg(-\bar\mu v(x) + \int s(r) \Big[-e^{(\alpha+\rho) r}v(x)v(x-r) \\ &&{\kern50pt}-e^{-\rho r} v(x)v(x+r) \\ &&{\kern50pt}+e^{(\alpha+\rho)r + (\alpha+2\rho)\psi(r)}v(x-\psi(r)) \\ &&{\kern55pt}\times v(x-r-\psi(r))\Big] {\rm d}r \Bigg), \end{array}$$
(11)

where ψ(r) = ln (1 + Ke  − r) and \(\bar\mu = t_0w_0^{-\rho}\mu\) (note that \(\bar\mu\) is the same as the dimensionless mortality parameter \(\hat\mu/(\hat Au_0)\) that is varied in Fig. 4 in the main text). Henceforth, all variables and parameters will be dimensionless and the overbars will be omitted. Note that v s(x) = 1 is a steady state of Eq. 11, provided that

$$ 0 \mu\int s(e^r) \left[ -e^{(\alpha+\rho)r}e^{-\rho r} + e^{(\alpha+\rho)r+(\alpha+2\rho)\psi(r)}\right] {\rm d}r. $$
(12)

This equation determines the intrinsic mortality exponent ρ as a function of α, K, μ and the feeding kernel s.

In order to investigate the behaviour of small perturbations to this steady state, we seek solutions of the form v(x, t) = 1 + ϵ(x, t), where |ϵ(x, t)| ≪ 1. Substituting this into Eq. 11, using Eq. 12 to eliminate terms of order ϵ 0 and neglecting terms of order ϵ 2 and higher, we obtain a linearised equation for ϵ:

$$\begin{array}{rll} \frac{\partial }{\partial t}\epsilon (x) &=& e^{-\rho x} \Bigg(-\mu\epsilon(x) + \int s(e^r)\Big[ -e^{(\alpha+\rho)r}(\epsilon(x)\notag\\ &&{\kern50pt}+\epsilon(x-r)) - e^{-\rho r}(\epsilon(x)+\epsilon(x+r)) \\ && {\kern50pt}+ e^{(\alpha+\rho)r+(\alpha+2\rho)\psi(r)}(\epsilon(x-\psi(r)) \\ &&{\kern50pt}+\epsilon(x-r-\psi(r))) \Big] {\rm d}r. \Bigg) \end{array}$$
(13)

In order to rewrite the right-hand side of this equation in the form of a linear integral operator acting on ϵ, the integral may be written as a sum of three individual integrals and change of integration variable z = f(r) made in each integral so that the integral contains ϵ(z). Recombining the resulting expression into a single integral then leads to

$$\begin{array}{rll} \frac{\partial }{\partial t}\epsilon (x) &=& e^{-\rho x}\left( b_0\epsilon(x) + \int G(x-r)\epsilon(r) {\rm d}r\right) \\ &&-\; M_{\rm s}(x)\epsilon(x), \end{array}$$
(14)

where

$$\begin{array}{rll} b_0 &=& -\mu - \int s(e^r) \left(e^{(\alpha+\rho)r}+e^{-\rho r}\right){\rm d}r, \\ G(r) &=& -s(e^r)e^{(\alpha+\rho)r} - s(e^{-r})e^{\rho r} \notag\\&&+\; s\left(\frac{K}{e^r-1}\right)\frac{K^{\alpha+\rho}e^{(\alpha+2\rho+1)r}}{(e^r-1)^{\alpha+\rho+1}} \nonumber\\ && +\; s\left(e^r-K\right) \frac{ e^{(\alpha+2\rho+1)r}}{(e^r-K)^{\rho+1}} ,\\ M_{\rm s}(x) &=& \left\{ \begin{array}{ll} \mu_{\rm s} e^{p(x-x_{\rm s})} & x\ge x_{\rm s}, \\ 0 & x<x_{\rm s}, \end{array}\right. . \end{array}$$

Here the perturbation damping term M s(x), introduced to reduce gradually the amplitude of the perturbations for x > x s, has also been included. Note that the integral term in the definition of b 0 corresponds to the constant C in Eq. 7 in the main text.

The McKendrick–von Foerster approximation can be obtained by carrying out a Taylor expansion in Eq. 13 up to terms of order K 2:

$$\begin{array}{rll} \frac{\partial }{\partial t}\epsilon (x) &=& -M_{\rm s}(x)\epsilon(x) + e^{-\rho x} \notag\\ &&\times\Bigg(\! -\mu \epsilon(x) \int s(e^r)\Big[\!-e^{-\rho r}\left(\epsilon(x)\epsilon(xr)\right) \\ &&{\kern14pt}\;\; + K e^{(\alpha+\rho-1) r}\left((\alpha+2\rho)(\epsilon(x)+\epsilon(x-r))\right.\notag\\ &&{\kern14pt}\;\; -\left.\epsilon^\prime(x)-\epsilon'(x-r)\right) + \frac{K^2}{2} e^{(\alpha+\rho-2)r}\nonumber \\ &&{\kern14pt}\;\; \times\Big((\alpha+2\rho)(\alpha+2\rho-1)(\epsilon(x)+\epsilon(x-r)) \\ &&{\kern14pt}\;\; + (1-2\alpha-4\rho)(\epsilon^\prime(x)+\epsilon^\prime(x-r)) \\ &&{\kern14pt}\;\; + \epsilon^{\prime\prime}(x)+\epsilon^{\prime\prime}(x-r)\Big) \Big] {\rm d}r \Bigg), \end{array}$$
(15)

where \(\epsilon^\prime(x) = (\partial/\partial x)\epsilon(x)\). As before, suitable changes of integration variables can be used to rewrite this equation in the form

$$\begin{array}{rll} \frac{\partial }{\partial t}\epsilon (x) &=& e^{-\rho x}\Bigg( b_0\epsilon(x) + b_1\epsilon^\prime(x) + b_2\epsilon^{\prime}(x) \nonumber \\ &&{\kern24pt}\quad + \int G_0(x-r)\epsilon(r) {\rm d}r\notag\\ &&{\kern24pt}\quad + \int G_1(x-r)\epsilon^{\prime}(r) {\rm d}r \notag\\ &&{\kern24pt}\quad +\int G_2(x-r)\epsilon^{\prime}(r) {\rm d}r \Bigg). \end{array}$$

As discussed below, numerical studies require the assumption that ϵ(x) is zero for x outside some finite interval, say 0 ≤ x ≤ a. In practice, a is taken to be sufficiently large that the damping term reduces ϵ(a) to a negligibly small value. Under the assumption that ϵ(a) = 0, integration by parts of the above equation gives

$$\begin{array}{rll} \frac{\partial }{\partial t}\epsilon (x) &=& e^{-\rho x}\Big( b_0\epsilon(x) + b_1\epsilon^\prime(x) + b_2\epsilon^{\prime\prime}(x) \notag\\ &&{\kern24pt}\quad -(G_1(x)+G_2^\prime(x))\epsilon(0) -G_2(x)\epsilon^\prime(0) \nonumber \\ &&{\kern24pt}\quad +\int G(x-r)\epsilon(r) {\rm d}r \Big) \\ &&-\, M_{\rm s}(x)\epsilon(x), \end{array}$$
(16)

where

$$\begin{array}{rll} b_0 &=& -\mu + \int s(e^r) \left( -e^{-\rho r} + (\alpha+2\rho)e^{(\alpha+\rho-1)r}\vphantom{\frac{K^2}{2}}\right.\notag\\&& {\kern23mm}\quad\quad\; \times \left.\left( K \frac{K^2}{2}(\alpha\rho)e^{-r}\right)\right) {\rm d}r,\\ b_1 &=& \int s(e^r)e^{(\alpha+\rho-1)r} \left(K \frac{K^2}{2}(1\alpha\rho)e^{-r}\right) {\rm d}r, \\ b_2 &=& \int s(e^r) \frac{K^2}{2} e^{(\alpha+\rho-2)r} {\rm d}r, \\ G(r) & = & \Big( -s(e^{-r})e^{\rho r} + K e^{(\alpha+\rho-1)r} \nonumber \\ &&{\kern8pt} \times\left((\rho+1)s(e^r)-e^r s^\prime(e^r)\right)+ \frac{K^2}{2} e^{(\alpha+\rho-2)r}\notag\\ &&{\kern8pt} \times \left( (\rho^2+3\rho+2) s(e^r) -2(\rho+1) \right.\\&& {\kern8pt} \times\left.e^r s'(e^r) + e^{2r} s''(e^r) \right) \Big). \end{array}$$

This is the second-order (K 2) approximation to the jump-growth equation. If all terms of order K 2 are also neglected, then the McKendrick–von Foerster (first-order) approximation to the jump-growth equation is obtained.

Both Eq. 14 and its approximation, Eq. 16, are in the form of a linear evolution equation:

$$ \frac{\partial \epsilon }{\partial t} = L\epsilon, $$
(17)

where L is a linear operator.

Numerical solution of the linear evolution equation

To study Eq. 17 numerically, the variable x is restricted to a finite range 0 ≤ x ≤ a. Biologically, this means that perturbations to the steady state are only allowed within this specified size range and the system is assumed to be at steady state outside this range. Furthermore, the perturbations in the range 0 ≤ x ≤ a are assumed to have no effect on the system outside this size range, which therefore remains at steady state. The variable x is also discretised using a fixed step size h. Thus, ϵ(x) is approximated by a finite vector \({\boldsymbol \epsilon}\) and Eq. 17 becomes a system of linear ordinary differential equations, \(\dot{\boldsymbol \epsilon} = M{\boldsymbol \epsilon}\), where M is a matrix defined by:

$$ M_{j,k} = \frac{\partial}{\partial {\boldsymbol \epsilon}_k} (L {\boldsymbol \epsilon})_j . $$

The solution to this system of differential equations is

$$ {\boldsymbol \epsilon}(t) = \sum\limits_{i=1}^N a_i {\boldsymbol v}_i e^{\lambda_i t}, $$
(18)

where λ i and \({\boldsymbol v}_i\) are the eigenvalues and eigenvectors of M and the a i are the coefficients of expanding the initial perturbation in terms of eigenvectors: \(\boldsymbol \epsilon(0) =\sum a_i {\boldsymbol v}_i\). If Re(λ i ) < 0 for all i, then small perturbations will always decay to zero in the long term. Conversely, if at least one eigenvalue has a positive real part, perturbations will grow.

The eigenvalues and eigenvectors of M can be computed by standard techniques. In the case of the full jump-growth equation with a realistic predator-to-prey mass ratio, taking a step size that is sufficiently small to resolve accurately the kernel G makes the problem computationally infeasible. However, the McKendrick–von Foerster approximations, with or without the order K 2 terms, are more tractable.

Calculating approximate wavelength and wave speed of eigenfunctions

The eigenfunctions of the linear operator L are not, in general, plane waves e ikx (unless ρ = 0) but are oscillatory complex functions of x. An approximate wavelength may be obtained from the average angular velocity of the eigenfunction in the complex plane. For a plane wave ϵ(x) = e ikx, the angular velocity dθ/dx (where \(\theta=\arg(\epsilon(x))\)) is constant and equal to k. The wavelength is equal to 2π/k. The wavelength of a general eigenfunction, at a particular value of x, may therefore be approximated by

$$ \textrm{wavelength} = 2\pi\left(\frac{{\rm d}\theta}{{\rm d}x}\right)^{-1}, $$

The average wavelength of the eigenfunction over the range 0 ≤ x ≤ a can be approximated by

$$ \textrm{average wavelength} = \frac{2\pi\Delta x}{\Delta \theta}, $$

where Δθ and Δx = a are the net change in \(\arg(\epsilon(x))\) and x, respectively, between x = 0 and x = a.

An initial perturbation that is equal to the eigenfunction ϵ(x) leads, after a period of time t, to a perturbation given by ϵ(x, t) = ϵ(x) e λt, where λ is the associated eigenvalue. In general, this is a complex-valued solution (both ϵ(x) and λ are complex-valued). A real-valued solution may be obtained by taking the sum of this solution with the corresponding solution for the complex conjugate eigenfunction and eigenvalue (this real-valued solution is the function that is plotted in Fig. 2.) This gives

$$ \epsilon(x,t) = \left( \epsilon_1(x)\cos(\omega t)+ \epsilon_2(x)\sin(\omega t)\right) e^{qt}, $$

where ϵ(x) = ϵ 1(x) +  2(x) and λ = q + . From this, it may be seen that the perturbations returns periodically to its original shape (with amplitude multiplied by e qt), with period T = 2π/ω. Therefore, the average wave speed is given by

$$ \textrm{average wave speed} = \frac{\textrm{average wavelength}}{\textrm{period}} = \frac{\omega \Delta x}{\Delta\theta}. $$

Rights and permissions

Reprints and permissions

About this article

Cite this article

Plank, M.J., Law, R. Ecological drivers of stability and instability in marine ecosystems. Theor Ecol 5, 465–480 (2012). https://doi.org/10.1007/s12080-011-0137-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12080-011-0137-x

Keywords

Navigation