Skip to main content
Log in

Trait diversity promotes stability of community dynamics

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

Abstract

The theoretical exploration of how diversity influences stability has traditionally been approached by species-centric methods. Here we offer an alternative approach to the diversity–stability problem by examining the stability and dynamics of size and trait distributions of individuals. The analysis is performed by comparing the properties of two size spectrum models. The first model considers all individuals as belonging to the same “average” species, i.e., without a description of diversity. The second model introduces diversity by further considering individuals by a trait, here asymptotic body size. The dynamic properties of the models are described by a stability analysis of equilibrium solutions and by the non-equilibrium dynamics. We find that the introduction of trait diversity expands the set of parameters for which the equilibrium is stable and, if the community is unstable, makes the oscillations smaller, slower, and more regular. The stabilizing mechanism is the variation in growth rate between individuals with the same body size but different trait values.

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.

Institutional subscriptions

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

Similar content being viewed by others

References

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

    Article  Google Scholar 

  • Andersen KH, Beyer JE, Lundberg P (2009) Trophic and individual efficiencies of size-structured communities. Proc R Soc B 276:109–114

    Article  PubMed  CAS  Google Scholar 

  • Andersen KH, Pedersen M (2010) Damped trophic cascades driven by fishing in model marine ecosystems. Proc R Soc B 277:795–802

    Article  PubMed  CAS  Google Scholar 

  • Benoît E, Rochet MJ (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 JE, Jennings S, Law R, Castle MD, McCloghrie P, Rochet MJ, 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 

  • 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 

  • Brooks JL, Dodson SI (1965) Predation, body size, and composition of plankton. Science 150:28–35

    Article  PubMed  CAS  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 

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

    Article  Google Scholar 

  • Chambers RC (1997) Environmental influences on egg and propagule sizes in marine fishes. In: Chambers RC, Trippel EA (eds) Early life history and recruitment in fish population. Chapman and Hall, London, pp 63–102

    Chapter  Google Scholar 

  • Cohen JE, Pimm SL, Yodzis P, Saldana J (1993) Body sizes of animal predators and animal prey in food webs. J Anim Ecol 62:67–78

    Article  Google Scholar 

  • Cohen JE, Jonsson T, Carpenter SR (2003) Ecological community description using the food web, species abundance, and body size. Proc Natl Acad Sci U S A 100:1781–1786

    Article  PubMed  CAS  Google Scholar 

  • Cross MC, Hohenberg PC (1993) Pattern formation outside of equilibrium. Rev Mod Phys 65:851–1112

    Article  CAS  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Datta S, Delius GW, Law R, Plank MJ (2011) A stability analysis of the power-law steady state of marine size spectrum. J Math Biol. doi:101007/s00285-010-0387-z

    PubMed  Google Scholar 

  • Dunne JA, Brose U, Williams RJ, Martinez ND (2005) Modeling food web dynamics: complexity–stability implications. In: Belgrano A, Scharler U, Dunne JA, Ulanowicz RE (eds) Aquatic food webs. Oxford University Press, New York, pp 117–129

    Chapter  Google Scholar 

  • Gross T, Rudolf L, Levin SA, Dieckmann U (2009) Generalized models reveal stabilizing factors in food webs. Science 325:747–750

    Article  PubMed  CAS  Google Scholar 

  • Hofbauer J, Sigmund K (1988) The theory of evolution and dynamical systems. Cambridge University Press, Cambridge

    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 

  • Jennings S, Mackinson S (2003) Abundance–body mass relationships in size-structured food webs. Ecol Lett 6:971–974

    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 

  • Kondoh M (2003) Foraging adaptation and the relationship between food-web complexity and stability. Science 299:1388–1391

    Article  PubMed  CAS  Google Scholar 

  • Kuznetsov YA (1994) Elements of applied bifurcation theory. Springer, New York

    Google Scholar 

  • Law R, Morton RD (1993) Alternative permanent states of ecological communities. Ecology 74:1347–1361

    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 

  • May RM (1972) Will a large complex system be stable. Nature 18:413–414

    Article  Google Scholar 

  • McCann K, Hastings A, Huxel GR (1998) Weak trophic interactions and the balance of nature. Nature 395:794–798

    Article  CAS  Google Scholar 

  • McKendrick AG (1926) Applications of mathematics to medical problems. P Edinburgh Math Soc 44:98–130

    Article  Google Scholar 

  • Oksanen L, Fretwell SD, Arruda J, Niemela P (1981) Exploitation ecosystems in gradients of primary productivity. Am Nat 2:264–261

    Google Scholar 

  • Peters R (1983) The ecological implications of body size. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Plank MJ, Law R (2011) Ecological drivers of stability and instability in marine ecosystems. Theor Ecol. doi:10.1007/s12080-011-0137-x

    Google Scholar 

  • Rice J, Gislason H (1996) Patterns of change in the size-spectra of numbers and diversity of the North Sea fish assemblage, as reflected in surveys and models. ICES J Mar Sci 53:1214–1225

    Article  Google Scholar 

  • Rochet MJ, 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 

  • Sheldon RW, Parsons T (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 WHJ (1972) The size distribution of particles in the ocean. Limnol Oceanogr 17:327–340

    Article  Google Scholar 

  • Sheldon RW, Sutcliffe WH, 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 

  • Shin YJ, Rochet MJ, Jennings S, Field JG, Gislason H (2005) Using size-based indicators to evaluate the ecosystem effects of fishing. ICES J Mar Sci 62:384–396

    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 

  • Ursin E (1973) On the prey size preferences of cod and dab. Medd Dan Fisk Havunders, New Ser 7:84–98

    Google Scholar 

  • von Foerster H (1959) Some remarks on changing populations. In Stohlman F (editor) The kinetics of cellular prolifeation. Grune and Stratton, New York, pp 381–407

  • Yodzis P (1981) The stability of real ecosystems. Nature 289:674–676

    Article  Google Scholar 

  • Yodzis P, Innes S (1992) Body size and consumer-resource dynamics. Am Nat 6:1151–1175

    Article  Google Scholar 

Download references

Acknowledgements

Ken Haste Andersen acknowledges financial support from EU FP7 project FACTS. Martin Hartvig is kindly acknowledged for helpful discussions. We greatly thank the reviewers for constructive comments on this paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lai Zhang.

Appendices

Appendix 1: Boundary condition

We derive the lower boundary condition N(m b,M) in the trait-based model on the basis of equilibrium size spectrum (ESS) theory (Andersen and Beyer 2006). Moreover, we assume that the total offspring recruitment in the trait-based model equals the recruitment in the community model, that is,

$$ \int_{M_0}^{M_{\rm b}} N(m_{\rm b},M){\rm d}M = \kappa m_{\rm b}^{-\uplambda}. $$
(11)

The ESS theory involves two assumptions. On one hand, the community size spectrum is of a power law state, that is, \(\overline{N}_{\rm c}(m)=\kappa m^{-\uplambda}\). On the other hand, the feeding level, indicating the mount of energy routed to individual somatic growth, is constant. Under these two conditions, Andersen and Beyer (2006) found the analytical solution to Eq. 1

$$ \overline{N}(m,M) = k_0 M^{2n-q-3+a}m^{-n-a}\left(1-\left(\frac{m}{M}\right)^{r-n}\right)^{\frac{a}{r-n}-1}, $$
(12)

where a = β 2n − q − 1/α is the physiological predation rate and k 0 a constant satisfying \(\int_m^{\infty} \overline{N}(m,M){\rm d}M = \kappa m^{-\uplambda}\). The exponent of community size spectrum is found to be \(\uplambda = 2+q-n\). Due to Eq. 11, the lower boundary condition is set as

$$ N(m_{\rm b}, M) = f_0 \overline{N}(m_{\rm b}, M)/k_0 \simeq f_0 M^{2n-q-3+a}m_{\rm b}^{-n-a}. $$
(13)

Therefore, f 0 is calculated by evaluating

$$ f_0 = k_0 \kappa m_{\rm b}^{-\uplambda+n+a}/\int_{M_0}^{M_{\rm b}}\overline{N}(m_{\rm b}, M){\rm d}M .$$
(14)

Since resource spectrum is fixed, encountered food of offspring can be calculated analytically

$$\begin{array}{rll} E(m_{\rm b}) & = & \gamma m_{\rm b}^{q} \int_{10^{-6}}^{10^{-3}}\kappa w^{1-\uplambda}\varphi(m_{\rm b}/w){\rm d}w \\ &\simeq& \gamma m_{\rm b}^{q} \int_{0}^{\infty}\kappa w^{1-\uplambda}\varphi(m_{\rm b}/w){\rm d}w \\ &=& \gamma\kappa\beta^{\uplambda-2}e^{(\uplambda-2)^2\sigma^2/2}m_{\rm b}^{2+q-\uplambda}. \end{array}$$

Note that \(\uplambda =2\) when q = n. Thus, E(m b) is independent of the width of selection function σ and the preferred predator–prey mass ratio β.

Appendix 2: Averaging growth efficiency

The average of growth efficiency is calculated based on the ESS theory, and our calculation is in principle a replicate of Andersen and Beyer (2006) and Andersen et al. (2009). In order for completeness, we present it here briefly.

The ESS theory assumes an infinite size range for community size spectrum, i.e., \(\overline{N}_{\rm c}(m) = {km}^{-\uplambda}, \ 0< m < \infty\). The advantage is that we can get a closed form when doing integration to obtain individual encountered food (Eq. 15) and mortality (Eq. 18), whereas in our trait-based model and community model, size spectrum is, however, truncated for numerical calculations. The individual encountered food can be calculated as

$$ E(m) = \gamma m^q \int_{0}^{\infty} w\overline{N}_{\rm c}(w)\varphi(m/w){\rm d}w = \kappa \alpha_e m^n, $$
(15)

where \(\alpha_e =\gamma \beta^{\uplambda-2}e^{(\uplambda-2)^2\sigma^2/2}\). Thus, individual somatic growth rate is

$$ g(m,M) = \alpha \left(1-\left(\frac{m}{M}\right)^{r-n}\right)E(m). $$
(16)

According to Eqs. 12 and 16, the average growth rate is calculated as

$$ g_{\rm c}(m) = \frac{1}{\overline{N}_{\rm c}}\int_m^{\infty} g(m,M)\overline{N}(m,M){\rm d}M = \alpha_{\rm c} m^n , $$
(17)

where the coefficient α c is to be determined.

The mortality rate is

$$ \mu(m) = \int_0^{\infty} \gamma w^q \overline{N}_{\rm c}(m) \varphi(w/m){\rm d}w = \alpha_p m^{1+q-\uplambda}, $$
(18)

where \(\alpha_p = \gamma \kappa e^{(1+q-\uplambda)^2\sigma^2/2}\beta^{1+q-\uplambda}\). Inserting the average growth rate (Eqs. 16 and 18) into the stationary equation of Eq. 10, we obtain

$$ \frac{\partial }{\partial m}(N_{\rm c}(m) g_{\rm c}(m)) = \mu(m)N_{\rm c}(m). $$
(19)

By solving the equation (Eq. 19), the exponent of the community size spectrum is found to be \(\uplambda = n + \alpha_p/\alpha_{\rm c}\). Notice that \(\uplambda = 2+q-n\), and then α c = α p / (2 + q − 2n).

Dividing Eqs. 17 by 15, the average growth efficiency is

$$ \overline{\varepsilon}_I = \frac{\beta^{2n-q-1}}{2+q-n} e^{((n-1)^2-(q-n)^2)\sigma^2/2} . $$
(20)

Appendix 3: Tracing equilibrium and determining stability

To solve the non-equilibrium dynamics to the trait-based size-spectrum model, we refer readers to Hartvig et al. (2011) where detailed numerical program is included and the key point is how to discretize the McKendrick–von Foerster equation which is presented in Eq. 21. Analogously, the dynamics of the community size spectrum can also be solved. In the following, the focus is how to continuously trace equilibrium solution as a function of a free parameter (e.g., the diet width σ). To this end, we apply Newton’s parameter continuation (Kuznetsov 1994) combined with the semi-implicit upwind scheme (Hartvig et al. 2011). Since body size spans several orders of magnitude, logarithmical size is preferred. Details of the logarithmical transformation can be referred to Benoît and Rochet (2004). We describe our technical approach by taking the community model as an example.

Discretization

Set x = log(m) and discretize x as x i  = log(m b) + Δx, i = 0, ⋯ , k, where x k  < log(M b) < x k + 1. In addition, let \(u_i^t = u(x_i, t)\Delta x = N(m(x_i))m(x_i)\Delta x\) represent the total number of individuals in the range of [x i − 1, x i ]. Then the discretization of the McKendrick–von Foerster Eq. 10 can be described as follows:

$$ \frac{u_i^{t+\Delta t}-u_i^{t}}{\Delta t} = -\frac{ g_i^t u_i^{t+\Delta t} -g_{i-1}^t u_{i-1}^{t+\Delta t} }{\Delta x} - \mu_i^t u_i^{t+\Delta t}, $$
(21)

where

$$ g_i^t = \overline{\varepsilon}_I\gamma e^{(q-1)x_i}\sum\limits_j e^{x_j} u_j^t \varphi(x_i-x_j), $$
(22)

and

$$ \mu_i^t = \gamma \sum\limits_j e^{qx_j}u_j^t\varphi(x_j-x_i). $$
(23)

Set \(U^t = (u_1^t, \cdots, u_k^t)^{\prime}\), \(B = (\frac{\Delta t}{\Delta x}g_0 u_0, 0, \cdots, 0)^{\prime}\), where the subscript indicates the transpose of a matrix or a vector. g 0 is completely determined by the resource spectrum and thus constant.

Equation 21 has an equivalent form

$$ A(U^t)U^{t+\Delta t} - U^t -B = 0, $$
(24)

where A(U) is a do-bidiagonal matrix and along the main and lower diagonals are vectors a = (a i ) and b = (b i ), respectively. Moreover, the entries in these two vectors are

$$ a_i = 1+ \frac{\Delta t}{\Delta x}g_i^t + \Delta t \mu_i^t $$
(25)

and

$$ b_i = -\frac{\Delta t}{\Delta x}g_i^t. $$
(26)

SetF(U): = A(U t)U t + Δt − U t − B. Making an explicit dependence of F on the free parameter leads to

$$ F(U, \sigma)= 0. $$
(27)

Newton’s continuation

Now it is ready to perform the Newton’s continuation using Eq. 27. Detailed description is referred to Kuznetsov (1994). We here present how to calculate the Jacobian matrix of Eq. 27 for given equilibrium U and σ.

Define dFdU to be the Jacobian matrix, which is of k 2 dimension, and set c j to be a row vector of dimension k whose entries are all zero except for the j-th component which is exactly 1. The dFdU can be calculated as follows: Noticing the special structure of the matrix A(U), the first row of the Jacobian is

$$ {\rm d}F{\rm d}U(1,:) = \left(\frac{\Delta t}{\Delta x}\frac{{\rm d}g_1}{{\rm d}U} + \Delta t \frac{{\rm d}\mu_1}{{\rm d}U} \right)u_1 + (a_1-1)c_1, $$
(28)

and for j > 1

$$ \begin{array}{rll} {\rm d}F{\rm d}U(j,:) &=& -\frac{\Delta t}{\Delta x} \frac{{\rm d}g_{j-1}}{{\rm d}U}u_{j-1} + b_{j-1}c_{j-1} \\ & & + \left(\frac{\Delta t}{\Delta x}\frac{{\rm d}g_j}{{\rm d}U} + \Delta t \frac{{\rm d}\mu_j}{{\rm d}U} \right)u_j+ (a_j-1)c_j, \end{array}$$
(29)

where dg j /dU and dμ j /dU are row vectors with the following form:

$$ \frac{{\rm d}g_j}{{\rm d}U} = \overline{\varepsilon}_I \gamma e^{({(q-1)}x_j)} [e^{x_1}\varphi(x_j-x_1), \cdots, e^{x_k}\varphi(x_j-x_k) ], $$
(30)
$$ \frac{{\rm d}\mu_j}{{\rm d}U} = \gamma [e^{qx_1}\varphi(x_1-x_j), \cdots, e^{qx_k}\varphi(x_k-x_j) ]. $$
(31)

To calculate the derivative of F with respect to the free parameter, we use the forward finite difference as an approximation, i.e.,

$$ {\rm d}F{\rm d}\sigma = (F(U, \sigma+ \delta) -F(U, \sigma))/ \delta, $$
(32)

where δ =  max {σδ 1, δ 2}.

Implementation of the parameter continuation in the trait-based model is similar but more challenging. Apart from the discretization along the direction of individual body size, the trait is also discretized evenly on the logarithmical scale. Thus, the matrix in Eq. 24 is now a black don-bidiagonal of K 2 dimension with \(K=\sum_{i=1}^{L}k_i\), where k i is the number of discretized points for species i and L is the number of discretized species. Each block matrix has the same structure as the matrix in Eq. 24 with entries Eqs. 25 and 26.

After some initial experimentation, we found that setting \(\Delta x =0.1, \Delta t= 0.02, \delta_1 = 10^{-4}, \delta_1 = 10^{-7}, L = 20\) could produce reliable result without causing too much computational effort, even though the precise choice of discretization affects the resolution of the approximation.

Determining the stability of equilibrium

In analogy to Eq. 21, discretization of the McKendric–von Foerster equation at given equilibrium U gives rise to

$$ \frac{\partial }{\partial t} u_i(t) = - \frac{g_iu_i-g_{i-1}u_{i-1}}{\Delta x} -\mu_iu_i, $$
(33)

where g i and μ i are similar with Eqs. 22 and 23. Rewriting the right-hand side of Eq. 33 as a matrix leads to

$$ G(U) = A_{\rm s}(U) +B_{\rm s}, $$
(34)

where A s and B s are similar to A and B but the entries are \(-\frac{a_i-1}{\Delta t}\) and \(-\frac{b_i}{\Delta t}\) in A s and B/Δt in B s. The Jacobian matrix of G(U) with respect to U can be calculated in analogy to Eqs. 28 and 29. Stability of equilibrium U is determined through the maximum real part of the Jacobian matrix of G(U) with positive value meaning unstable and vice versa.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Zhang, L., Thygesen, U.H., Knudsen, K. et al. Trait diversity promotes stability of community dynamics. Theor Ecol 6, 57–69 (2013). https://doi.org/10.1007/s12080-012-0160-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12080-012-0160-6

Keywords

Navigation