Skip to main content

Advertisement

Log in

Bimodal trait distributions with large variances question the reliability of trait-based aggregate models

  • ORIGINAL PAPER
  • Published:
Theoretical Ecology Aims and scope Submit manuscript

Abstract

Functionally diverse communities can adjust their species composition to altered environmental conditions, which may influence food web dynamics. Trait-based aggregate models cope with this complexity by ignoring details about species identities and focusing on their functional characteristics (traits). They describe the temporal changes of the aggregate properties of entire communities, including their total biomasses, mean trait values, and trait variances. The applicability of aggregate models depends on the validity of their underlying assumptions that trait distributions are normal and exhibit small variances. We investigated to what extent this can be expected to work by comparing an innovative model that accounts for the full trait distributions of predator and prey communities to a corresponding aggregate model. We used a food web structure with well-established trade-offs among traits promoting mutual adjustments between prey edibility and predator selectivity in response to selection. We altered the shape of the trade-offs to compare the outcome of the two models under different selection regimes, leading to trait distributions increasingly deviating from normality. Their biomass and trait dynamics agreed very well for stabilizing selection and reasonably well for directional selection, under which different trait values are favored at different times. However, for disruptive selection, the results of the aggregate model strongly deviated from the full trait distribution model that showed bimodal trait distributions with large variances. Hence, the outcome of aggregate models is reliable under ideal conditions but has to be questioned when confronted with more complex selection regimes and trait distributions, which are commonly observed in nature.

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

  • Abrams PA (2006) The prerequisites for and likelihood of generalist-specialist coexistence. Am Nat 167:329–342

    Article  PubMed  Google Scholar 

  • Abrams PA, Harada Y, Matsuda H (1993) On the relationship between quantitative genetic and ESS models. Evolution 47:982–985

    Article  Google Scholar 

  • Acevedo-Trejos E, Brandt G, Bruggeman J, Merico A (2015) Mechanisms shaping size structure and functional diversity of phytoplankton communities in the ocean. Scientific reports 5

  • Akpalu W (2009) Economics of biodiversity and sustainable fisheries management. Ecol Econ 68(10):2729–2733

    Article  Google Scholar 

  • Allen CR, Garmestani A, Havlicek T, Marquet P, Peterson G (2006) Patterns in body mass distributions: sifting among alternative hypotheses. Ecol Lett 9:630–643

    Article  CAS  PubMed  Google Scholar 

  • Banas NS (2011) Adding complex trophic interactions to a size-spectral plankton model: emergent diversity patterns and limits on predictability. Ecol Model 222:2663–2675

    Article  CAS  Google Scholar 

  • Baptestini EM, de Aguiar MA, Bolnick DI, Araujo MS (2009) The shape of the competition and carrying capacity kernels affects the likelihood of disruptive selection. J Theor Biol 259:5–11

    Article  PubMed  Google Scholar 

  • Bauer B, Vos M, Klauschies T, Gaedke U (2014) Diversity, functional similarity, and top-down control drive synchronization and the reliability of ecosystem function. Am Nat 183(3):394–409

    Article  PubMed  Google Scholar 

  • Becks L, Ellner SP, Jones LE, Hairston Jr. NG (2010) Reduction of adaptive genetic diversity radically alters eco-evolutionary community dynamics. Ecol Lett 13(8):989–997

    PubMed  Google Scholar 

  • Blest DC (2003) A new measure of kurtosis adjusted for skewness. Australian & New Zealand J Stat 45 (2):175–179

    Article  Google Scholar 

  • Bruggeman J, Kooijman SA (2007) A biodiversity-inspired approach to aquatic ecosystem modeling. Limnol Oceanogr 52(4):1533–1544

    Article  Google Scholar 

  • Chisholm SW (1992) Phytoplankton size. In: Falkowski PG, Woodhead AD (eds) Primary productivity and biogeochemical cycles in the sea. Springer, pp 213–237

  • DeCarlo LT (1997) On the meaning and use of kurtosis. Psychol Methods 2(3):292

    Article  Google Scholar 

  • Dieckmann U, Doebeli M, Metz J, Tautz D (2004) Adaptive speciation. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Doebeli M (1996) A quantitative genetic competition model for sympatric speciation. J Evol Biol 9:893–909

    Article  Google Scholar 

  • Doebeli M, Blok HJ, Leimar O, Dieckmann U (2007) Multimodal pattern formation in phenotype distributions of sexual populations. Proc R Soc Lond B Biol Sci 274(1608):347–357

    Article  Google Scholar 

  • Downing A, Hajdu S, Hjerne O, Otto S, Blenckner T, Larsson U, Winder M (2014) Zooming in on size distribution patterns underlying species coexistence in baltic sea phytoplankton. Ecol Lett 17:1219–1227

    Article  CAS  PubMed  Google Scholar 

  • Fox JW, Nelson WA, McCauley E (2010) Coexistence mechanisms and the paradox of the plankton: quantifying selection from noisy data. Ecology 91:1774–1786

    Article  PubMed  Google Scholar 

  • Gaedke U (1992) The size distribution of plankton biomass in a large lake and its seasonal variability. Limnol Oceanogr 37(6):1202–1220

    Article  Google Scholar 

  • Gaedke U, Wickham S (2004) Ciliate dynamics in response to changing biotic and abiotic conditions in a large, deep lake (lake constance). Aquat Microb Ecol 34:247–261

    Article  Google Scholar 

  • Hansen PJ, Bjornsen PK, Hansen BW (1997) Zooplankton grazing and growth: scaling within the 2-2000- μm body size range. Limnol Oceanogr 42:687–704

    Article  Google Scholar 

  • Havlicek TD, Carpenter SR (2001) Pelagic species size distributions in lakes: are they discontinuous? Limnol Oceanogr 46(5):1021–1033

    Article  Google Scholar 

  • Hindmarsh AC (1983) Odepack, a systematized collection of ode solvers. In: et al RSS (ed) Scientific computing, IMACS transactions on scientific computation, vol 1. North-Holland, Amsterdam, pp 55–64

  • Holling C (1965) The functional response of predators to prey density and its role in mimicry and population regulation. Memoirs of the Entomological Society of Canada 45:1–60

    Google Scholar 

  • Holling C (1992) Cross-scale morphology, geometry and dynamics of ecosystems. Ecol Monogr 62(4):447–502

    Article  Google Scholar 

  • Horn H, Paul L, Horn W, Petzoldt T (2011) Long-term trends in the diatom composition of the spring bloom of a German reservoir: is aulacoseira subarctica favoured by warm winters? Freshw Biol 56:2483–2499

    Article  Google Scholar 

  • Jones E, Oliphant T, Peterson P, et al. (2001) SciPy: open source scientific tools for Python. http://www.scipy.org/, [Online; accessed 2015-07-25]

  • Jones M, Rosco J, Pewsey A (2011) Skewness-invariant measures of kurtosis. Am Stat 65(2):89–95

    Article  Google Scholar 

  • Kingsolver JG, Hoekstra HE, Hoekstra JM, Berrigan D, Vignieri SN, Hill C, Hoang A, Gibert P, Beerli P (2001) The strength of phenotypic selection in natural populations. Am Nat 157(3):245– 261

    Article  CAS  PubMed  Google Scholar 

  • Klaassen CA, Mokveld PJ, Van Es B (2000) Squared skewness minus kurtosis bounded by 186/125 for unimodal distributions. Statistics & Probability Letters 50(2):131–135

    Article  Google Scholar 

  • Lande R (1982) A quantitative genetic theory of life history evolution. Ecology 63:607–615

    Article  Google Scholar 

  • Leibold MA, Norberg J (2004) Biodiversity in metacommunities: plankton as complex adaptive systems? Limnol Oceanogr 49:1278–1289

    Article  Google Scholar 

  • Levins R (1962) Theory of fitness in a heterogeneous environment. i. the fitness set and adaptive function. Am Nat 96(891):361–373

    Article  Google Scholar 

  • Maharjan R, Nilsson S, Sung J, Haynes K, Beardmore RE, Hurst LD, Ferenci T, Gudelj I (2013) The form of a trade-off determines the response to competition. Ecol Lett 16:1267–1276

    Article  PubMed  Google Scholar 

  • Martin RA, Pfennig DW (2012) Widespread disruptive selection in the wild is associated with intense resource competition. BMC Evol Biol 12(1):136

    Article  PubMed  PubMed Central  Google Scholar 

  • Maurer BA, Brown JH, Dayan T, Enquist BJ, Ernest SM, Hadly EA, Haskell JP, Jablonski D, Jones KE, Kaufman DM, et al. (2004) Similarities in body size distributions of small-bodied flying vertebrates. Evol Ecol Res 6(6):783–797

    Google Scholar 

  • Merico A, Bruggeman J, Wirtz K (2009) A trait-based approach for downscaling complexity in plankton ecosystem models. Ecol Model 220(21):3001–3010

    Article  CAS  Google Scholar 

  • Merico A, Brandt G, Smith SL, Oliver M (2014) Sustaining diversity in trait-based models of phytoplankton communities. Population Dynamics 2:59

    Google Scholar 

  • Moors J (1986) The meaning of kurtosis: Darlington reexamined. Am Stat 40(4):283–284

    Google Scholar 

  • Norberg J (2004) Biodiversity and ecosystem functioning: a complex adaptive systems approach. Limnol Oceanogr 49:1269– 1277

    Article  Google Scholar 

  • Norberg J, Swaney DP, Dushoff J, Lin J, Casagrandi R, Levin SA (2001) Phenotypic diversity and ecosystem functioning in changing environments: a theoretical framework. Proc Natl Acad Sci 98(20):11,376–11,381

    Article  CAS  Google Scholar 

  • Norberg J, Urban MC, Vellend M, Klausmeier CA, Loeuille N (2012) Eco-evolutionary responses of biodiversity to climate change. Nat Clim Chang:1–5

  • Pearson K (1929) Editorial note on inequalities for moments of frequency functions and for various statistical constants. Biometrika:370–375

  • Pigolotti S, López C, Hernández-García E, Andersen KH (2010) How gaussian competition leads to lumpy or uniform species distributions. Theor Ecol 3:89–96

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

    Article  Google Scholar 

  • Ruel J, Ayres M (1999) Jensen’s inequality predicts effects of environmental variation. TREE 14:361–366

    PubMed  Google Scholar 

  • Saloniemi I (1993) A coevolutionary predator-prey model with quantitative characters. Am Nat 141(6):880–896

    Article  CAS  PubMed  Google Scholar 

  • Savage VM, Webb CT, Norberg J (2007) A general multi-trait-based framework for studying the effects of biodiversity on ecosystem functioning. J Theor Biol 247(2):213–229

    Article  PubMed  PubMed Central  Google Scholar 

  • Schartau M, Landry MR, Armstrong RA (2010) Density estimation of plankton size spectra: a reanalysis of ironex ii data. J Plankton Res 32(8):1167–1184

    Article  Google Scholar 

  • Scheffer M, van Nes EH (2006) Self-organized similarity, the evolutionary emergence of groups of similar species. Proc Natl Acad Sci 103(16):6230–6235

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Scheffer M, Rinaldi S, Gragnani A, Mur LR, van Nes EH (1997) On the dominance of filamentous cyanobacteria in shallow, turbid lakes. Ecology 78:272–282

    Article  Google Scholar 

  • Segura A, Kruk C, Calliari D, García-rodriguez F, Conde D, Widdicombe C, Fort H (2013) Competition drives clumpy species coexistence in estuarine phytoplankton. Scientific reports 3

  • Sommer U, Adrian R, De Senerpont Domis L, Elser JJ, Gaedke U, Ibelings B, Jeppesen E, Lürling M, Molinero JC, Mooij WM, et al. (2012) Beyond the plankton ecology group (peg) model: mechanisms driving plankton succession. Annu Rev Ecol Evol Syst 43:429–448

    Article  Google Scholar 

  • Straub C, Ives A, Gratton C (2011) Evidence for a trade-off between host-range breadth and host-use efficiency in aphid parasitoids. Am Nat 177:389–395

    Article  PubMed  Google Scholar 

  • Strauss S, Rudgers J, Lau J, Irwin R (2002) Direct and ecological costs of resistance to herbivory. Trends Ecol Evol 17:278–285

    Article  Google Scholar 

  • Tackx MLM, Herman PMJ, van Rijswijk P, Vink M, Bakker C (1994) Plankton size distributions and trophic relations before and after the construction of the storm-surge barrier in the oosterschelde estuary. Hydrobiologia 282(/283):145–152

    Article  Google Scholar 

  • Tang EP (1995) The allometry of algal growth rates. J Plankton Res 17:1325–1335

    Article  Google Scholar 

  • Terseleer N, Bruggeman J, Lancelot C, Gypens N (2014) Trait-based representation of diatom functional diversity in a plankton functional type model of the eutrophied southern north sea. Limnol Oceanogr 59(6):1958–1972

    Article  Google Scholar 

  • Thibault KM, White EP, Hurlbert AH, Ernest S (2011) Multimodality in the individual size distributions of bird communities. Glob Ecol Biogeogr 20(1):145–153

    Article  Google Scholar 

  • Tirok K, Gaedke U (2007) The effect of irradiance, vertical mixing and temperature on spring phytoplankton dynamics under climate change: long-term observations and model analysis. Oecologia 150(4):625–642

    Article  PubMed  Google Scholar 

  • Tirok K, Gaedke U (2010) Internally driven alternation of functional traits in a multispecies predator-prey system. Ecology 91:1748–1762

    Article  PubMed  Google Scholar 

  • Tirok K, Bauer B, Wirtz K, Gaedke U (2011) Predator-prey dynamics driven by feedback between functionally diverse trophic levels. PLoS ONE 6(11):e27,357. doi:10.1371/journal.pone.0027357. http://www.ncbi.nlm.nih.gov/pubmed/22096560

    Article  CAS  Google Scholar 

  • Vasseur DA, Gaedke U (2007) Spectral analysis unmasks synchronous and compensatory dynamics in plankton communities. Ecology 88(8):2058–2071

    Article  PubMed  Google Scholar 

  • Vergnon R, Dulvy NK, Freckleton RP (2009) Niches versus neutrality: uncovering the drivers of diversity in a species-rich community. Ecol lett 12(10):1079–1090

    Article  PubMed  Google Scholar 

  • Webb CT, Hoeting JA, Ames GM, Pyne MI, n LeRoy P (2010) A structured and dynamic framework to advance traits-based theory and prediction in ecology. Ecol Lett 13:267–283

    Article  PubMed  Google Scholar 

  • Wirtz KW, Eckhardt B (1996) Effective variables in ecosystem models with an application to phytoplankton succession. Ecol Model 92(1):33–53

    Article  Google Scholar 

Download references

Acknowledgments

We thank E. van Velzen and two anonymous referees for helpful comments and suggestions. RMC was supported through grant #2012/05949-6, São Paulo Research Foundation (FAPESP), Brazil, and also through a post-doctoral PNPD fellowship from CAPES (Brazil). TK was funded by the German Research Foundation (DFG, GA 401/19-1 and 26-1).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Toni Klauschies.

Additional information

Renato Mendes Coutinho and Toni Klauschies contributed equally to this work.

Electronic supplementary material

Below is the link to the electronic supplementary material.

(MP4 1.52 MB)

(MP4 1.44 MB)

(MP4 1.64 KB)

Appendices

Appendix A: Derivation of the aggregate model

The FTD model describes the temporal changes in the biomass density of each particular trait value and thus yields, at every moment in time, the biomass densities of the prey and predator communities (or functional groups) as a function of their respective traits ϕ and ω. As a consequence, the magnitude and shape of the biomass-trait distributions may change in time. Instead of resolving the rate of change of the biomass density for each particular trait value, one may directly describe the temporal changes in the aggregate properties of the prey and predator communities using moment approximation methods. In this case, the dynamics of the FTD model is approximated by a corresponding aggregate model which tracks only the temporal changes in the total biomasses of the prey and predator communities and the locations (mean trait values) and widths (trait variances) of the corresponding trait distributions. These aggregate quantities are related to the trait distribution by integrals over the ecologically feasible trait range, in accordance with Eq. 5. In this way, aggregate models explicitly represent the trait-dependent growth and loss terms that determine the dynamics of the aggregate properties of the trait distribution (cf. Tirok et al. 2011).

To derive the aggregate model (68) from the FTD model (13), one has to write down the rate of change of the aggregate properties using Eqs. 13 and expand all functions in Taylor series around the mean trait value, since the average over the distribution of the per-capita net growth rates determines the overall trait and biomass dynamics. This leads to expressions written in terms of (central) moments of the trait distributions. Afterwards, a moment approximation is performed by truncating these series up to the second order. This will be a good approximation when the growth and grazing functions have only weak non-linearities and thus all higher-order derivatives quickly vanish, or when the trait distribution is very narrow, that is, has a small variance. Hence, the error in the biomass and trait dynamics will be most pronounced if both the higher-order derivatives of the fitness landscape, i.e., the per-capita net-growth rate as a function of the trait, evaluated at the mean trait value and the higher-order moments of the trait distribution obtain large absolute values at the same time. For example, if trait distributions with a sufficiently large variance exhibit large values of skewness or kurtosis, the quality of the approximation made by the corresponding aggregate model will be rather poor when the fitness landscape is highly non-linear. This holds since the performance of species with trait values away from the community average \(\bar {\phi }\) will likely differ from the performance of a species with a trait value equal to \(\bar {\phi }\) (cf. Ruel and Ayres 1999; Tirok et al. 2011).

For sake of brevity, we will derive the prey equations in detail below, as the derivation of the equations for the predator are completely analogous. The Taylor expansion of a function f around the prey’s mean trait value \(\bar {\phi }\) is given by:

$$ f(\phi) = \sum\limits_{n=0}^{\infty} \frac{(\phi-\bar{\phi})^{n}}{n!} \frac{d^{n}f(\bar{\phi})}{d\phi^{n}} \,. $$
(13)

Since we are dealing with aggregate quantities such as prey biomass summed over the whole trait range, we will need the expansion of integrals over the prey’s trait distribution A(ϕ), such as:

figure b

where M n is the nth central moment of the distribution A(ϕ) (in particular, M 2 = ν ϕ ), defined by

$$ M_{n} = \frac{1}{A_{T}}\int (\phi-\bar{\phi})^{n}A(\phi)d\phi \,. $$
(14)

The rate of change of the total prey biomass A T is

$$\begin{array}{@{}rcl@{}} \frac{dA_{T}}{dt} &=& \frac{d}{dt} \int A(\phi)d\phi = \int \frac{dA(\phi)}{dt} d\phi\\ &=& \int \left[A(\phi) \tilde{R}(\phi) + I_{A} \right] d\phi, \end{array} $$
(15)

where \(\tilde {R}\) is the per capita net growth rate from Eq. 1:

$$ \tilde{R}(\phi) = r(\phi) + B(\phi) - \frac{1}{A(\phi)}\int g(\phi, \omega) C(\omega) d\omega . $$
(16)

Now, the grazing term already contains an integral over \(\phi ^{\prime }\), independent of ϕ, so we expand it first:

$$\begin{array}{@{}rcl@{}} &&\frac{1}{A(\phi)} \int g(\phi, \omega) C(\omega) d\omega\\ &=& \frac{g_{m}\, q(\phi, \omega)}{M(\omega) + \int q(\phi^{\prime}, \omega) A(\phi^{\prime})d\phi^{\prime}}\\ &=& \frac{g_{m}\, q(\phi, \omega)}{M(\omega) + A_{T}\left[q(\bar{\phi}, \omega) + \left.\frac{\nu_{\phi}}{2} \frac{\partial^{2}q(\phi,\omega)} {\partial\phi^{2}}\right|_{\phi=\bar{\phi}} + \dots\right]} , \end{array} $$
(17)

and here we drop terms of order higher than two (denoted by the dots), which gives us simply G(ϕ,ω) defined by Eq. 8. We now proceed expanding the integral in terms of ω:

$$\begin{array}{@{}rcl@{}} \tilde{R}(\phi)&=& r(\phi) + B(\phi) - \int G(\phi, \omega) C(\omega) d\omega\\ &=& r(\phi) + B(\phi) \\ &&+ C_{T} \left[ G(\phi, \bar{\omega}) +\left. \frac{\nu_{\omega}}{2}\frac{\partial^{2}G(\phi, \omega)}{\partial\omega^{2}}\right|_{\omega=\bar{\omega}} + \dots\right],\\ \end{array} $$
(18)

and, stopping once again at the second order, we arrive at the expression for R A (ϕ) (7). Substituting it back into Eq. 15 and expanding:

$$\begin{array}{@{}rcl@{}} \frac{dA_{T}}{dt} &=& \int \left[A(\phi) R_{A}(\phi) + I_{A} \right] d\phi\\ &=& A_{T} \left( R_{A}(\bar{\phi}) + \left.\frac{\nu_{\phi}}{2}\frac{\partial^{2}R_{A}}{\partial \phi^{2}}\right|_{\phi=\bar{\phi}} + {\cdots} \right) + I_{A} , \end{array} $$
(19)

which yields Eq. 6 when we truncate at the second order.

The rate of change of the mean trait value is derived in a similar fashion:

$$\begin{array}{@{}rcl@{}} \frac{d\bar{\phi}}{dt} &=& \frac{1}{A_{T}} \left[\frac{d}{dt}\int \phi A(\phi) d\phi - \frac{1}{A_{T}} \frac{dA_{T}}{dt} \int \phi A(\phi) d\phi\right]\\ &=& \frac{1}{A_{T}} \left[\int \phi \frac{dA(\phi)}{dt} d\phi - \bar{\phi} \frac{dA_{T}}{dt} \right]\\ &=& \frac{1}{A_{T}} \int (\phi-\bar{\phi}) \frac{dA(\phi)}{dt} d\phi\\ &=&\frac{1}{A_{T}} \int (\phi-\bar{\phi}) \left( A(\phi) R_{A}(\phi) + I_{A}\right) d\phi\\ &=& \frac{1}{A_{T}} \int (\phi-\bar{\phi}) A(\phi) R_{A}(\phi) d\phi \\&&+ \frac{1}{A_{T}} \int (\phi-\bar{\phi}) I_{A} d\phi\\ &=&\left( \left.\nu_{\phi} \frac{\partial R_{A}}{\partial\phi}\right|_{\phi=\bar{\phi}} + \left.\frac{M_{3}}{2}\frac{\partial^{2} R_{A}}{\partial\phi^{2}}\right|_{\phi=\bar{\phi}} + \dots\right)\\ &&+ \frac{I_{A}}{A_{T}} \left( \frac{1}{2} - \bar{\phi} \right) . \end{array} $$
(20)

Finally, the trait variance will be expressed by:

$$\begin{array}{@{}rcl@{}} \frac{d\nu_{\phi}}{dt} &=& \frac{1}{A_{T}} \left[\frac{d}{dt}\int (\phi-\bar{\phi})^{2} A(\phi) d\phi \right. \\ && \qquad\left.- \frac{1}{A_{T}} \frac{dA_{T}}{dt} \int (\phi-\bar{\phi})^{2} A(\phi) d\phi\right]\\ &=& \frac{1}{A_{T}} \left[\int (\phi-\bar{\phi})^{2}\, \frac{dA(\phi)}{dt} d\phi - \nu_{\phi} \frac{dA_{T}}{dt} \right]. \end{array} $$
(21)

The second term can be expanded directly using Eq. 19

$$ \frac{\nu_{\phi}}{A_{T}} \frac{dA_{T}}{dt} = \nu_{\phi} \left[\left( R_{A}(\bar{\phi}) + \left.\frac{\nu_{\phi}}{2}\frac{\partial^{2}R_{A}}{\partial\phi^{2}}\right|_{\phi=\bar{\phi}} \,+\, {\cdots} \right) \,+\,\frac{I_{A}}{A_{T}} \right] , $$
(22)

while the first term can be written as:

$$\begin{array}{@{}rcl@{}} && \frac{1}{A_{T}} \int (\phi-\bar{\phi})^{2} \left( A(\phi) R_{A}(\phi) + I_{A} \right) d\phi = \nu_{\phi} R_{A}(\bar{\phi})\\ &&+\left. M_{3} \frac{\partial R_{A}}{\partial\phi}\right|_{\phi=\bar{\phi}} + \left.\frac{M_{4}}{2} \frac{\partial^{2} R_{A}}{\partial\phi^{2}}\right|_{\phi=\bar{\phi}}\\ &&+ \frac{I_{A}}{A_{T}} \left( \frac{1}{3}+\bar{\phi}^{2}-\bar{\phi}\right). \end{array} $$
(23)

Given the fact that the terms proportional to ν ϕ in Eq. 23 cancel with the ones from Eq. 22 and that the terms proportional to \(\nu _{\phi }^{2}\) combine with those from Eq. 22, the rate of change of the trait variance can be approximated by:

$$\begin{array}{@{}rcl@{}} \frac{d\nu_{\phi}}{dt} &=& M_{3} \left.\frac{\partial R_{A}}{\partial\phi}\right|_{\phi=\bar{\phi}} + \frac{\left( M_{4} - {\nu_{\phi}}^{2}\right)}{2} \left.\frac{\partial^{2} R_{A}}{\partial\phi^{2}}\right|_{\phi=\bar{\phi}}\\ &&+ \frac{I_{A}}{A_{T}} \left[\frac{1}{12} - \nu_{\phi} + \left( \frac{1}{2}-\bar{\phi}\right)^{2}\right] . \end{array} $$
(24)

As seen in Eqs. 20 and 24, temporal changes in the aggregate properties (lower-order moments) generally depend on higher-order moments. To derive self-contained expressions for the evolution of the aggregate properties, and thus to close the system of differential equations, one has to describe the higher-order moments in the corresponding equations by lower-order moments. Different moment closure techniques were described in detail by Norberg et al. (2001) and Merico et al. (2009). For instance, Norberg et al. (2001) described the third and fourth central moments by power-functions of the first (mean trait) and second (variance) central moments and the optimal trait value. The parameters were estimated from the trait distributions obtained from simulations of the underlying multispecies model. Another approach is to assume the trait distributions to be well represented by a Gaussian (Wirtz and Eckhardt 1996; Merico et al. 2009; Tirok et al. 2011) for which the higher-order moments are either zero (odd moments) or fully determined by the variance alone (even moments).

Following Wirtz and Eckhardt (1996) and Merico et al. (2009), we assume traits to be normally distributed. Using the fact that the third central moment of a Gaussian distribution equals zero (M 3=0; the normal distribution is symmetric), and that its fourth central moment is proportional to the square of its variance (\(M_{4} = 3 {M_{2}^{2}}\)), that is, its excess kurtosis is 0, we can approximate the rate of change of the mean trait value Eq. 20 and of the trait variance Eq. 24 by Eq. 6.

Appendix: B: Model version with mutation

The immigration term in Eqs. 1 and 6 prevents the extinction of species by providing a very small constant influx of immigrants of all trait values. This is generally realistic and necessary to ensure that some variance is retained in the aggregate model, but may in principle strongly influence the dynamics of both types of models. Here, we build versions of the FTD model and the aggregate model without immigration, but with mutation instead. Mutation can be an important process when dealing with narrowly defined functional groups, where the trait variance may originate mostly from intra-populational processes rather than standing trait variance among different species.

Here, mutation is understood as a small probability of offspring to have trait values slightly different then their parents (we do not account for sexual mixing). It is modeled as a (Fokker-Planck) diffusive term, corresponding to the second derivative of the population density times the birth rate with respect to the trait value. In the case of prey, we assign all density-dependent terms to the mortality, and thus their birth rate, ρ A , is simply the intrinsic growth rate (\(r^{\prime }\)). The birth rate ρ C of predators is their grazing rate. This leads to the modified FTD model:

$$\begin{array}{@{}rcl@{}} \frac{\partial A(\phi, t)}{\partial t} &=& \left( r(\phi) + B(\phi)\right) A - \int g(\phi, \omega) C d\omega \\ &&+ \mu_{A} \frac{\partial^{2} \left[\rho_{A} A(\phi)\right]}{\partial \phi^{2}}\\ \frac{\partial C(\omega, t)}{\partial t} &=& \left( e \int g(\phi, \omega)d\phi - d + B(\omega)\right) C\\ &&+ \mu_{C} \frac{\partial^{2} \left[\rho_{C} C(\omega)\right]}{\partial \omega^{2}}\\ \rho_{A}(\phi) &=& r^{\prime}(\phi) \qquad \rho_{C}(\omega) = e \int g(\phi, \omega)d\phi , \end{array} $$
(25)

where μ A and μ C are the mutation rates for prey and predator, respectively, and all other terms and parameters are as before. We derive the aggregate model from the FTD model from this equation together with Eq. 5. Due to the boundary functions imposed, the population densities go to zero very quickly outside the range of traits from 0 to 1. Therefore, we assume that A(ϕ), as well as its derivatives, go fast enough to 0 as ϕ goes to (plus or minus) infinity, which guarantees that terms calculated at infinity appearing via integration by parts vanish.

The derivation proceeds exactly as in “AppendixA:Derivation of the aggregate model”. We begin by calculating the rate of change of total biomass, \(\frac {dA_{T}}{dt}\). The new term corresponding to mutation in Eq. 15 will be:

$$ \mu_{A} \int \frac{\partial^{2} \left[\rho_{A} A(\phi)\right]}{\partial \phi^{2}} d\phi= \left.\mu_{A}\frac{\partial \left[\rho_{A} A(\phi)\right]}{\partial \phi} \right|_{-\infty}^{\infty} = 0 , $$
(26)

which means that mutation does not contribute any extra term to the total biomass equation. Proceeding to the mean trait equation, \(\frac {d\bar {\phi }}{dt}\), we now get the following contribution of the mutation term:

$$\begin{array}{@{}rcl@{}} \frac{\mu_{A}}{A_{T}} \int \phi \frac{\partial^{2} \left[\rho_{A} A(\phi)\right]}{\partial \phi^{2}} d\phi &=&\frac{\mu_{A}}{A_{T}} \left.\left[\phi \frac{\partial \left[\rho_{A} A(\phi)\right]}{\partial \phi} \right] \right|_{-\infty}^{\infty} \\ &&- \frac{\mu_{A}}{A_{T}} \int \frac{\partial \left[\rho_{A} A(\phi)\right]}{\partial \phi} d\phi\\ &=& \left.0 - \frac{\mu_{A}}{A_{T}} \left[\rho_{A} A(\phi)\right] \right|_{-\infty}^{\infty}\\ &=& 0 , \end{array} $$
(27)

and hence mutation does not affect the mean trait value either. Finally, the equation for the rate of change of the trait variance, \(\frac {d\nu _{\phi }}{dt}\), has a term due to mutation given by:

$$\begin{array}{@{}rcl@{}} &&\frac{\mu_{A}}{A_{T}} \int \phi^{2} \frac{\partial^{2} \left[\rho_{A} A(\phi)\right]}{\partial \phi^{2}}d\phi = \frac{\mu_{A}}{A_{T}} \left.\left[\phi^{2} \frac{\partial \left[\rho_{A} A(\phi)\right]}{\partial \phi} \right] \right|_{-\infty}^{\infty}\\ &&- 2 \frac{\mu_{A}}{A_{T}} \int \phi \frac{\partial \left[\rho_{A} A(\phi)\right]}{\partial \phi} d\phi \\ &\,=\,&0 - 2 \frac{\mu_{A}}{A_{T}} \left\{\left.{\vphantom{\frac{\mu_{A}}{A_{T}}}}\left[\phi \rho_{A} A(\phi)\right] \right|_{-\infty}^{\infty}- \frac{\mu_{A}}{A_{T}} \int \rho_{A} A(\phi) d\phi \right\}\\ &\,=\,& 2 \frac{\mu_{A}}{A_{T}} \int \rho_{A} A(\phi) d\phi\\ &\,=\,& 2 \mu_{A} \left[\rho_{A}(\bar{\phi}) + \left.\frac{\nu_{\phi}}{2} \frac{\partial^{2} \rho_{A}(\phi)}{\partial \phi^{2}}\right|_{\phi=\bar{\phi}} \,+\, {\dots} \right] . \end{array} $$
(28)

Thus the equation for trait variance becomes:

$$ \frac{d\nu_{\phi}}{dt} = \nu_{\phi}^{2} \left.\frac{\partial^{2} R_{A}}{\partial\phi^{2}}\right|_{\phi=\bar{\phi}} + 2 \mu_{A} \rho_{A}(\bar{\phi}) + \nu_{\phi} \mu_{A} \left.\frac{\partial^{2} \rho_{A}(\phi)}{\partial \phi^{2}}\right|_{\phi=\bar{\phi}} . $$
(29)

We analyzed the resulting dynamics of the FTD model and the aggregate model as described in the methods of the main text. For all selection regimes, Fig. 8 shows very similar patterns for the FTD model with mutation compared to the FTD model with immigration (cf. Fig. 3). In contrast, the aggregate model no longer displayed large oscillations in its mean trait values for intermediate values of β, but high trait variances instead, indicating that even the type of dynamics it generates is not robust to such structural modification. Importantly, immigration influenced the mean trait values directly, pulling toward intermediate values and facilitating recurrent changes in trait values. Even for stabilizing selection, the agreement between the two types of models in the temporal averages of the total biomasses and mean trait values is worse than in the scenario with immigration, since variance levels of the FTD model are higher in the mutation variant (4×10−3 compared to 2×10−5, when β=0.2), leading to worse performance of the aggregate model due to Jensen’s inequality.

Fig. 8
figure 8

Temporal averages (lines) and variation (± 1 standard deviation; shades) of biomass, mean trait, and standing trait variance of the solutions of the FTD (red) and aggregate (green) models with mutation instead of immigration (2529), as the shape of the trade-offs (β) is varied. Mutation rates were set to μ A =10−4 and μ C =2×10−5, with all other parameters as in Table 1. The dotted lines correspond to alternative long-term solutions, obtained by exploring different initial conditions, leading to distinct basins of attraction

In the absence of immigration, the FTD model no longer exhibited bistability. Also, the transition between states of very different dominant trait values occurred more slowly, leading to longer cycles. Nonetheless, it still presented bimodal trait distributions for a wide range of parameters. For instance, P was above 1.2 (see “Numerical simulations and data analysis” section) during 12 and 68 % of the time for prey and predators, respectively, with β=3, and during 5 and 45 % for β=5. This confirms that bimodal trait distributions were stabilized but not caused by immigration, which is in accordance with findings from Pigolotti et al. (2010).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Coutinho, R.M., Klauschies, T. & Gaedke, U. Bimodal trait distributions with large variances question the reliability of trait-based aggregate models. Theor Ecol 9, 389–408 (2016). https://doi.org/10.1007/s12080-016-0297-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12080-016-0297-9

Keywords

Navigation