Skip to main content
Log in

Rapid prey evolution and the dynamics of two-predator food webs

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

Abstract

Traits affecting ecological interactions can evolve on the same time scale as population and community dynamics, creating the potential for feedbacks between evolutionary and ecological dynamics. Theory and experiments have shown in particular that rapid evolution of traits conferring defense against predation can radically change the qualitative dynamics of a predator–prey food chain. Here, we ask whether such dramatic effects are likely to be seen in more complex food webs having two predators rather than one, or whether the greater complexity of the ecological interactions will mask any potential impacts of rapid evolution. If one prey genotype can be well-defended against both predators, the dynamics are like those of a predator–prey food chain. But if defense traits are predator-specific and incompatible, so that each genotype is vulnerable to attack by at least one predator, then rapid evolution produces distinctive behaviors at the population level: population typically oscillate in ways very different from either the food chain or a two-predator food web without rapid prey evolution. When many prey genotypes coexist, chaotic dynamics become likely. The effects of rapid evolution can still be detected by analyzing relationships between prey abundance and predator population growth rates using methods from functional data analysis.

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
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Abrams PA (2000) The evolution of predator–prey interactions: theory and evidence. Ann Rev Ecolog Syst 31:79–105

    Article  Google Scholar 

  • Abrams PA, Matsuda H (1997) Prey adaptation as a cause of predator–prey cycles. Evolution 51:1742–1750

    Article  Google Scholar 

  • Becks L, Hilker FM, Malchow H, Jurgens K, Arndt H (2005) Experimental demonstration of chaos in a microbial food web. Nature 435:1226–1229

    Article  PubMed  CAS  Google Scholar 

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

    PubMed  Google Scholar 

  • Bellman R, Roth RS (1971) The use of splines with unknown end points in the identification of systems. J Math Anal Appl 34:26–33

    Article  Google Scholar 

  • Benincà E, Huisman J, Heerkloss R, Jöhnk KD, Branco P, Van Nes EH, Scheffer M, Ellner SP (2008) Chaos in a long-term experiment with a plankton community. Nature 451:822–825

    Article  PubMed  Google Scholar 

  • Benincà E, Jöhnk KD, Heerkloss R, Huisman J (2009) Coupled predator–prey oscillations in a chaotic food web. Ecol Lett 12:1367–1378

    Article  PubMed  Google Scholar 

  • Berlow EL, Neutel AM, Cohen JE, Ruiter PCd, Ebenman B, Emmerson M, Fox JW, Jansen VAA, Jones JI, Kokkoris GD, Logofet DO, McKane AJ, Montoya JM, Petchey O (2004) Interaction strengths in food webs: issues and opportunities. J Anim Ecol 73:585–598

    Article  Google Scholar 

  • Brunel J-B (2008) Parameter estimation of ODE’s via nonparametric estimators. Electron J Stat 2:1242–1267

    Article  Google Scholar 

  • Bulmer MG (1975) Phase relations in the ten-year cycle. J Anim Ecol 44:609–621

    Article  Google Scholar 

  • Cortez M, Ellner S (2010) Understanding rapid evolution in predator–prey interactions using the theory of fast-slow dynamical systems. Am Nat 176:E109–E127

    Article  PubMed  Google Scholar 

  • Cousyn C, De Meester L, Colbourne JK, Brendonck L, Verschuren D, Volckaert F (2001) Rapid, local adaptation of zooplankton behavior to changes in predation pressure in the absence of neutral genetic changes. Proc Natl Acad Sci USA 98:6256–6260

    Article  PubMed  CAS  Google Scholar 

  • Dercole F, Ferriere R, Gragnani A, Rinaldi S (2006) Coevolution of slow-fast populations: evolutionary sliding, evolutionary pseudo-equilibria and complex Red Queen dynamics. Proc R Soc B Biol Sci 273:983–990

    Article  CAS  Google Scholar 

  • Doebeli M, Koella JC (1995) Evolution of simple population-dynamics. Proc R Soc Lond B Biol Sci 260:119–125

    Article  Google Scholar 

  • Doebeli M, Koella J (1996) Chaos and evolution. Trends Ecol Evol 11:220–220

    Article  CAS  Google Scholar 

  • Dube D, Kim K, Alker AP, Harvell CD (2002) Size structure and geographic variation in chemical resistance of sea fan corals Gorgonia ventalina to a fungal pathogen. Mar Ecol Prog Ser 231:139–150

    Article  Google Scholar 

  • Duffy MA, Sivars-Becker L (2007) Rapid evolution and ecological host-parasite dynamics. Ecol Lett 10:44–53

    Article  PubMed  Google Scholar 

  • Duffy MA, Hall SR (2008) Selective predation and rapid evolution can jointly dampen effects of virulent parasites on daphnia populations. Am Nat 171:499–510

    Article  PubMed  Google Scholar 

  • Duffy MA, Brassil CE, Hall SR, Tessier AJ, Caceres CE, Conner JK (2008) Parasite-mediated disruptive selection in a natural Daphnia population. BMC Evol Biol 8:80

    Article  PubMed  Google Scholar 

  • Duffy MA, Hall SR, Caceres CE, Ives AR (2009) Rapid evolution, seasonality, and the termination of parasite epidemics. Ecology 90:1441–1448

    Article  PubMed  Google Scholar 

  • Duncan AB, Little TJ (2007) Parasite-driven genetic change in a natural population of Daphnia. Evolution 61:796–803

    Article  PubMed  Google Scholar 

  • Ellner S, Turchin P (1995) Chaos in a noisy world—new methods and evidence from time-series analysis. Am Nat 145:343–375

    Article  Google Scholar 

  • Ellner SP, Kendall BE, Wood SN, McCauley E, Briggs, CJ (1997) Inferring mechanism from time-series data: delay-differential equations. Physica D 110:182–194

    Article  Google Scholar 

  • Ellner SP, Hairston NG Jr, Kearns CM, Babai D (1999) The roles of fluctuating selection and long-term diapause in microevolution of diapause timing in a freshwater copepod. Evolution 53:111–122

    Article  Google Scholar 

  • Ellner SP, Seifu Y, Smith RH (2002) Fitting population dynamic models to time-series data by gradient matching. Ecology 83:2256–2270

    Article  Google Scholar 

  • Ezard THG, Cote SD, Pelletier F (2009) Eco-evolutionary dynamics: disentangling phenotypic, environmental and population fluctuations. Philos Trans R Soc B Biol Sci 364:1491–1498

    Article  Google Scholar 

  • Ferriere R, Fox GA (1995) Chaos and evolution. Trends Ecol Evol 10:480–485

    Article  CAS  Google Scholar 

  • Fussmann GF, Loreau M, Abrams PA (2007) Eco-evolutionary dynamics of communities and ecosystems. Funct Ecol 21:465–477

    Article  Google Scholar 

  • Gagneux S, Long CD, Small PM, Van T, Schoolnik GK, Bohannan BJM (2006) The competitive cost of antibiotic resistance in Mycobacterium tuberculosis. Science 312:1944–1946

    Article  PubMed  CAS  Google Scholar 

  • Grant PR, Grant BR (2002) Unpredictable evolution in a 30-year study of darwin’s finches. Science 296:707–711

    Article  PubMed  CAS  Google Scholar 

  • Hairston NG Jr, Walton WE (1986) Rapid evolution of a life-history trait. Proc Natl Acad Sci USA 83:4831–4833

    Article  PubMed  CAS  Google Scholar 

  • Hairston NG Jr, Lampert W, Caceres CE, Holtmeier CL, Weider LJ, Gaedke U, Fischer JM, Fox JA, Post DM (1999) Lake ecosystems—rapid evolution revealed by dormant eggs. Nature 401:446–446

    Article  Google Scholar 

  • Hampton SE, Scheuerell MD, Schindler DE (2006) Coalescence in the Lake Washington story: Interaction strengths in a planktonic food web. Limnol Oceanogr 51:2042–2051

    Article  Google Scholar 

  • Heath DD, Heath JW, Bryden CA, Johnson RM, Fox CW (2003) Rapid evolution of egg size in captive salmon. Science 299:1738–1740

    Article  PubMed  CAS  Google Scholar 

  • Hendry AP, Wenburg JK, Bentzen P, Volk EC, Quinn TP (2000) Rapid evolution of reproductive isolation in the wild: evidence from introduced salmon. Science 290:516–518

    Article  PubMed  CAS  Google Scholar 

  • Hooker G (2009) Forcing function diagnostics for nonlinear dynamics. Biometrics 65:928–936

    Article  PubMed  Google Scholar 

  • Ives AR, Carpenter SR, Dennis B (1999) Community interaction webs and zooplankton responses to planktivory manipulations. Ecology 80:1405–1421

    Article  Google Scholar 

  • Ives AR, Dennis B, Cottingham KL, Carpenter SR (2003) Estimating community stability and ecological interactions from time-series data. Ecol Monogr 73:301–330

    Article  Google Scholar 

  • Johst K, Doebeli M, Brandl R (1999) Evolution of complex dynamics in spatially structured populations. Proc R Soc Lond B Biol Sci 266:1147–1154

    Article  Google Scholar 

  • Jones LE, Ellner SP (2004) Evolutionary tradeoff and equilibrium in an aquatic predator–prey system. Bull Math Biol 66:1547–1573

    Article  PubMed  Google Scholar 

  • Jones LE, Ellner SP (2007) Effects of rapid prey evolution on predator–prey cycles. J Math Biol 55:541–573

    Article  PubMed  Google Scholar 

  • Kassen R (2009) Toward a general theory of adaptive radiation: insights from microbial experimental evolution. Year in evolutionary biology 2009. Ann NY Acad Sci 1168:3–22

    Article  PubMed  Google Scholar 

  • Kendall BE, Prendergast J, Bjornstad ON (1998) The macroecology of population dynamics: taxonomic and biogeographic patterns in population cycles. Ecol Lett 1:160–164

    Article  Google Scholar 

  • Khibnik AI, Kondrashov AS (1997) Three mechanisms of Red Queen dynamics. Proc R Soc Lond B Biol Sci 264:1049–1056

    Article  Google Scholar 

  • Kinnison M, Hairston NG Jr (2007) Eco-evolutionary conservation biology: contemporary evolution and the dynamics of persistence. Funct Ecol 21:444–454

    Article  Google Scholar 

  • Kinnison MT, Unwin MJ, Quinn TP (2008) Eco-evolutionary vs habitat contributions to invasion in salmon: experimental evaluation in the wild. Mol Ecol 17:405–414

    Article  PubMed  Google Scholar 

  • Kondoh M (2007) Anti-predator defence and the complexity-stability relationship of food webs. Proc R Soc B Biol Sci 274:1617–1624

    Article  Google Scholar 

  • Konishi S, Kitagawa G (2008) Information criteria and statistical modeling. Springer Series in Statistics, Springer, New York

  • Levin SA (1972) Mathematical analysis of genetic feedback mechanism. Am Nat 106:145–164

    Article  Google Scholar 

  • Levin SA, Udovic DJ (1977) Mathematical model of co-evolving populations. Am Nat 111:657–675

    Article  Google Scholar 

  • Meyer JR, Ellner SP, Hairston NG Jr, Jones LE, Yoshida T (2006) Prey evolution on the time scale of predator–prey dynamics revealed by allele-specific quantitative PCR. Proc Natl Acad Sci USA 103:10690–10695

    Article  PubMed  CAS  Google Scholar 

  • Olsen EM, Heino M, Lilly GR, Morgan MJ, Brattey J, Ernande B, Dieckmann U (2004) Maturation trends indicative of rapid evolution preceded the collapse of northern cod. Nature 428:932–935

    Article  PubMed  CAS  Google Scholar 

  • Palkovacs EP, Marshall MC, Lamphere BA, Lynch BR, Weese DJ, Fraser DF, Reznick DN, Pringle CM, Kinnison MT (2009) Experimental evaluation of evolution and coevolution as agents of ecosystem change in Trinidadian streams. Philos Trans R Soc B Biol Sci 364:1617–1628

    Article  Google Scholar 

  • Reznick DN, Ghalambor CK (2001) The population ecology of contemporary adaptations: what empirical studies reveal about the conditions that promote adaptive evolution. Genetica 112:183–198

    Article  PubMed  Google Scholar 

  • Reznick DN, Shaw FH, Rodd FH, Shaw RG (1997) Evaluation of the rate of evolution in natural populations of guppies (Poecilia reticulata). Science 275:1934–1937

    Article  PubMed  CAS  Google Scholar 

  • Shertzer KW, Ellner SP, Fussmann GF, Hairston NG Jr (2002) Predator–prey cycles in an aquatic microcosm: testing hypotheses of mechanism. J Anim Ecol 71:802–815

    Article  Google Scholar 

  • Sinervo B, Svensson E, Comendant T (2000) Density cycles and an offspring quantity and quality game driven by natural selection. Nature 406:985–988

    Article  PubMed  CAS  Google Scholar 

  • Swain DP, Sinclair AF, Mark Hanson J (2007) Evolutionary response to size-selective mortality in an exploited fish population. Proc R Soc Lond B Biol Sci 274:1015–1022

    Article  Google Scholar 

  • Swartz J, Bremermann H (1975) Discussion of parameter estimation in biological modelling: algorithms for estimation and evaluation of the estimates. J Math Biol 1:241–257

    Article  Google Scholar 

  • Thompson JN (1998) Rapid evolution as an ecological process. Trends Ecol Evol 13:329–332

    Article  CAS  Google Scholar 

  • Tien R (2010) Two theoretical studies investigating predator–prey interactions. PhD Thesis, Department of Ecology and Evolutionary Biology, Cornell University, Ithaca NY 14853

  • Tyerman JG, Bertrand M, Spencer CC, Doebeli M (2008) Experimental demonstration of ecological character displacement. BMC Evol Biol 8:34

    Article  PubMed  Google Scholar 

  • Varah J (1982) A spline least squares method for numerical parameter estimation in differential equations. SIAM J Sci Statist Comput 3:28–46

    Article  Google Scholar 

  • Weitz JS, Hartman H, Levin SA (2005) Co-evolutionary arms races between bacteria and bacteriophage. Proc Natl Acad Sci USA 102: 9535–9540

    Article  PubMed  CAS  Google Scholar 

  • Wootton JT, Emmerson M (2005) Measurement of interaction strength in nature. Ann Rev Ecolog Evol Syst 36:419–444

    Article  Google Scholar 

  • Yoshida T, Jones LE, Ellner SP, Fussmann GF, Hairston NJ Jr (2003) Rapid evolution drives ecological dynamics in a predator–prey system. Nature 424:303–306

    Article  PubMed  CAS  Google Scholar 

  • Yoshida T, Hairston NG Jr, Ellner SP (2004) Evolutionary trade-off between defence against grazing and competitive ability in a simple unicellular alga, Chlorella vulgaris. Proc R Soc Lond B Biol Sci 271:1947–1953

    Article  Google Scholar 

  • Yoshida T, Ellner SP, Jones LE, Bohannan BJM, Lenski RE, Hairston NG Jr (2007) Cryptic population dynamics: Rapid evolution masks trophic interactions. PLoS Biology 5:1868–1879

    Article  CAS  Google Scholar 

  • Zbinden M, Haag CR, Ebert D (2008) Experimental evolution of field populations of Daphnia magna in response to parasite treatment. J Evol Biol 21:1068–1078

    Article  PubMed  CAS  Google Scholar 

  • Zeineddine M, Jansen VAA (2005) The evolution of stability in a competitive system. J Theor Biol 236:208–215

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the Cornell EEB chemostat group (Aldo Barreiro Felpeto, Michael Cortez, Nelson G. Hairston Jr., Teppo Hiltunen, Giles Hooker, Laura E. Jones, Joseph Simonis, Yuefeng Wu) and two perceptive anonymous referees for their advice and comments on the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stephen P. Ellner.

Additional information

Research supported by a grant from the James S. McDonnell Foundation and US National Science Foundation grant DEB-0813743. We dedicate this paper to Simon A. Levin, a pioneer in the study of eco-evolutionary dynamics, on the occasion of his 70th birthday.

Electronic supplementary material

Below is the link to the electronic supplementary material.

(PDF 699 kb)

Appendix

Appendix

A Trait dynamics in clonal versus sexually reproducing prey

Here, we show that many of our results are also relevant to sexually reproducing prey because our clonal evolution model can be re-expressed as a gradient dynamics model for trait evolution in a sexual species.

Let x 1 and x 2 be the abundances of the two prey clones, X = x 1 + x 2, and let f = x 2/(x 1 + x 2). We need to re-write the model in terms of X, the total prey population, and its “trait mean” f, rather than x 1 and x 2. In the predator and substrate equations, we just substitute x 2 = fX, x 1 = (1 − f)X, so that for example Q = ((1 − f)p 1 + f p 2)X. The dynamic equation for f is (omitting some algebra)

$$\begin{array}{rll} \dot f &=& \frac{d}{dt}{\left( \frac{x_2}{x_1+x_2} \right)} = f(1-f)\left[\frac{\dot x_2}{x_2} - \frac{\dot x_1}{x_1} \right]\\ &=& f(1-f)(W_2 - W_1)\end{array}$$
(6)

where W i is the fitness of prey type i. Note that f is the population mean of the trait whose value is 1 for clone-2 individuals, and 0 for clone-1 individuals, and f(1 − f) is the population variance for this trait. The dynamic equation for X is

$$ \dot X = \dot x_1 + \dot x_2 = x_1 W_1 + x_2 W_2 = X\left[(1-f)W_1 + f W_2 \right]. $$
(7)

Comparing the last two equations, we see that the rate of change in f is the trait variance f(1 − f) multiplying the gradient with respect to f of prey fitness \(\dot X/X\) with the predator’s grazing rate g/(k y  + Q) held constant, which is the fitness gradient for a rare invader. Thus, our two-clone model is equivalent to a gradient-dynamics model for frequency-dependent evolution of the defense trait in a sexually reproducing species. Inherited from this model’s asexual origins are the form of the mean-variance relationship, and the relationship between mean trait value f and per-capita nutrient uptake rate in the quantitative trait model, namely \(\frac{(1-f)m_1 S}{k_1 + S} + \frac{fm_2 S}{k_2 + S}.\) But mean-variance relationships like the one in (6) have often been assumed to model traits limited to a finite interval (such as fractional allocations), and while the nutrient uptake rate is not a standard model, it’s not any less reasonable than specifying m and k directly as a function of the mean defense level.

With three or more clones the correspondence between clonal and sexual models is no longer exact. Let f i denote the population mean of the trait which has value 1 for clone i, for all other clones (i.e., the frequency of clone i). Then calculations similar to those above show that \(\dot f_i = f_i(1-f_i)\frac{\partial}{\partial f_i}\frac{\dot X}{X}\), if the derivative with respect to f i is taken by increasing f i and decreasing all other clone frequencies by a constant fraction of their value while maintaining \(\sum\limits_j{f_j} = 1\) and holding the predator grazing rate constant. So we get gradient dynamics, but not a standard model for correlated quantitative traits in a sexual species. It remains an open question whether the complex dynamics seen in our model with three clones would also be seen in a model for sexually reproducing prey.

B Calculating phase lags

Here we explain our method for computing approximate phase lags, which we implemented using computer algebra. The oscillations in our model appear to always arise through a supercritical Hopf bifurcation, as in the one-predator model. We compute approximate phase lags for parameters just past that bifurcation. Solutions x(t) are then small-amplitude oscillations about an equilibrium with common period T, and x j (t) is approximately proportional to sin(λt + ψ j ) where λ = 2 π/T and ψ j is the phase for component j.

To compute relative phases, we choose one solution component (say x 0) as the reference and set the origin of time so that ψ 0 = 0. The other components (which we still denote x(t)) then satisfy a forced nonlinear system \(\dot{{\bf{x}}} = \mathbf{F}({\bf{x}},x_0(t))\) where x 0(t) = εsin(λt) is regarded as an exogenous forcing term. The linearized system for deviations from equilibrium can be written as

$$ \dot{{\bf{x}}} = \mathbf{J} {\bf{x}}+{\bf{u}}(t) $$
(8)

where J is the Jacobian matrix of F at the equilibrium, x(t) is now the m-vector of deviations from the equilibrium, and \({\bf{u}}(t)=\varepsilon \sin(\lambda t) \frac{\partial \mathbf{F}}{\partial x_0}\) is the forcing. The asymptotic solutions to (8) are

$$\begin{array}{rll} x_i(t) & =& A_i \cos(\lambda t) + B_i \sin(\lambda t), \\ \mbox {with } u_i(t) & =& b_i \sin(\lambda t), \qquad b_i=\varepsilon \frac{\partial \mathbf{F_i}}{\partial x_0}. \end{array}$$
(9)

The unknown A’s and B’s, which determine the phase lags, can be found by writing (8) as a matrix equation. Differentiation of x is accomplished by transforming A i λB i , B i → − λA i . Defining

$$\begin{array}{rll}{\bf{X}} &=& (B_1,B_2,\ldots,B_m,A_1,A_2,\ldots,A_m)\\ {\bf{U}} &=& (b_1,b_2,\ldots,b_m,0,0,\ldots,0) \end{array}$$
(10)

differentation of x is represented by the matrix \( \mathbf{D} = \left[ {\begin{array}{*{20}c} 0 & { - \lambda \mathbf{I}} \\ {\lambda \mathbf{I}} & 0 \\ \end{array} } \right] \) and multiplication by J by the matrix \(\mathbf{J}_2 = \left[ {\begin{array}{*{20}c} \mathbf{J} & 0 \\ 0 & \mathbf{J} \\ \end{array} } \right]\). The coefficient representation of (8) is therefore D X = J 2 X + U with solution \({\bf{X}}=(\mathbf{D} - \mathbf{J}_2)^{-1} {\bf{U}}\). The blocks in D − J 2 commute, so defining R = (J 2 + λ 2 I) − 1, we have

$$ (\mathbf{D} - \mathbf{J}_2)^{-1} = \left[ {\begin{array}{*{20}c} { - \mathbf{R}\mathbf{J}} & {\lambda \mathbf{R}} \\ { - \lambda \mathbf{R}} & { - \mathbf{R}\mathbf{J}} \\ \end{array} } \right]. $$
(1)

The addition formula sin(λt + ψ) = sin(ψ) cos(λt) + cos(ψ) sin(λt) shows that the phase ψ i of x i (t) relative to sin(λt) has the properties

$$\begin{array}{rll} \sin(\psi_i) &=& A_i/\rho, \mbox{ where } \rho=\sqrt{A_i^2 + B_i^2} \\ \cos(\psi_i) &=& B_i/\rho \\ \tan(\psi_i) &=& A_i/B_i. \end{array}$$
(12)

The value of ψ i is therefore uniquely determined by the signs and ratio of A i and B i , so we can compute any convenient positive multiple of (A i , B i ). As a visual aid for readers, Fig. 18 in the Electronic supplementary material shows how the scaled coefficients (A i , B i ) relate to phase differences with the forcing variable.

The main limit to this method is that the expressions for A and B become very complicated as the number of state variables goes up, because they are symbolic solutions of a high-dimensional, non-sparse linear system. For scenarios with one or two prey genotypes we can make sense of the expressions for phase lags, at least in ecologically relevant limits of the parameter values, but with three genotypes it exceeds our abilities and we have to study the system numerically.

C LFC model without prey evolution

Here we consider LFC model (2) when there is only one genotype in the focal prey species x, assuming that all state variables have small oscillations around their mean values. The top predator only feeds on x, so it lags x by a quarter period (Bulmer 1975). The intermediate predator can behave differently because it is driven by both x and W. Writing \(x=\bar x (1+ \varepsilon \sin(\lambda t))\) where λ = 2 π/T and T is the cycle period, the relevant subsystem is

$$\begin{array}{rll} \dot z & = & z(hx(t) + \phi w -1) \\ \dot w & = & M-w-\phi z w \end{array}$$
(13)

Without forcing (i.e. ε = 0), (13) has steady state

$$ \bar w = \frac{1-h\bar x}{\phi}, \bar z = \frac{M-\bar w}{\phi \bar w}. $$
(14)

This is a stable node whenever parameters are such that \(\bar w\) and \(\bar z\) are positive (the Jacobian has trace \(-(1+\phi \bar z)<0\) and determinant \(\phi^2 \bar z \bar w >0\)), and numerical evidence suggests that it is globally stable (so far, no luck with the Bendixson–Dulac negative criterion). The forced system (ε > 0) is therefore “chasing” a moving equilibrium \((\bar z(t), \bar w(t))\) in which \(\bar z(t)\) is a monotonically increasing function of x(t), hence z(t) exhibits oscillations that track x(t) but lag behind it.

To be more precise, the phase lag for small oscillations can be calculated by the method in Appendix B for the system (13) linearized about the steady-state \((\bar z, \bar w, \bar x)\) for ε = 0. These calculations yield that \(z(t)-\bar z\) is proportional to sin(λt + ψ) where sin(ψ) < 0,cos(ψ) > 0 and

$$ \tan(\psi) = -\frac{\lambda^3+\left(\left(\phi \bar z+1\right)^2-\phi^2 \bar w \bar z \right)\lambda} {\phi^2 \bar w \bar z \left(\phi \bar z+1 \right)} $$
(15)

(note that because \(\phi \bar w = 1- h \bar x < 1\), the numerator on the right-hand side of (15) is positive, hence tan(ψ) < 0). These properties imply that ψ ∈ ( − π/2, 0), so z lags behind x by an amount that approaches the upper limit of one quarter cycle period when tan(ψ) is very large in magnitude, and approaches zero (relative to T) when tan(ψ) is small.

Although the expression for tan(ψ) is complicated, some situations can be identified in which the lag is predicted to be short vs. long relative to T.

  • Because λ → 0 as T →  ∞, the ratio between the lag and the cycle period goes to zero as the cycle period increases. That is, when the cycle period is very long, the dynamics of z and W become “slaved” to x(t), and z(t) is almost exactly at the value of \(\bar z\) determined by the current value of x(t).

  • The right-hand side of (15) can be written as

    $$ - \frac{\lambda \Omega + \lambda^3(1-h \bar x)^2}{\phi M (1-h\bar x)(\phi M + h\bar x -1)}. $$
    (16)

    where Ω > 0 and all terms in the denominator of (16) are positive so long as \(\bar w, \bar z >0\). The lag therefore approaches the quarter-period limit when the denominator goes to 0. One way that this happens is when \(h \bar x \to 1\), meaning that the intermediate predator is grazing heavily on the focal prey species and reaching high enough densities that its alternate food W is driven to very low levels (\(\bar w \to 0\)). The intermediate predator is then almost a specialist on the focal prey species, so it has the classical quarter-period lag behind the focal prey.

  • The other way for the denominator of (16) to go to zero is if ϕ decreases to \((1-h\bar x)/M\). This is the value of ϕ at which \(\bar z =0\) when \(x = \bar x\). The intermediate predator therefore increases only when the focal prey species is more abundant (above \(\bar x\)), and is heading for extinction whenever the focal prey species is less abundant. The intermediate predator therefore behaves as if the focal prey were its main food source, even if \(h \bar x\) is small.

Numerical evaluations of (15) indicate that the effects of h and ϕ hold generally: any increase in h or decrease in ϕ causes the linearized system’s phase lag to increase. The accuracy of (16) for the nonlinear system (13) is very good for small fluctuations (Fig. 19a in the Electronic supplementary material) and remains fairly good for substantial fluctuations (19b, in which the cycle peak is roughly three times the trough). The linearized analysis generally under-predicts the lag by a bit, and is most accurate near the extremes of zero or quarter-period lag.

D IC model without prey evolution

As the second baseline case, we calculate here the approximate phase relations between x,y and z in the Intermediate Consumer model with only one prey type. It is simplest to take x(t) as the forcing function and determine the phase lags of y and z relative to x. The equations for this scenario are

$$\begin{array}{rll} \dot y & =& y\left[\frac{gx(t)}{k_y + x(t)} +\eta z - 1\right] \\ \dot z & =& z\left[hx(t) - \eta y - 1 \right] \end{array}$$
(17)

As before we set \(x(t)=\bar x (1+ \varepsilon \sin(\lambda t))\), and assume small oscillations of y and z about \(\bar y\) and \(\bar z\), the equilibria of (17) when \(x(t) \equiv \bar x\),

$$\begin{array}{lll} &&{\kern-6pt}\bar z = (1- G \bar x)/\eta, \quad \bar y = (h \bar x-1)/\eta \\ &&{\kern-6pt}\mbox{where} \ G=g/(k_y + \bar x). \end{array}$$
(18)

We must have \(G \bar x <1, h\bar x >1\) for the equilibrium to exist. These inequalities correspond to z being the more effective predator, able to persist on x alone, while y depends on consuming z as well as x in order to persist.

The method of Appendix B can be applied to the linearization of (17). Note that these calculations are relevant to the full three-variable food web model at parameter values that generate a small-amplitude limit cycle, rather than to the dynamics of (17) on its own without any feedbacks from y and z to x. The results (up to a positive factor) are

$$ \begin{aligned} A_y &= -g k_y \lambda, & B_y &= -h(\bar x + k_y)^2 \eta \bar z \\ A_z &= -h(\bar x + k_y)^2 \lambda, & B_z &= g k_y \bar y. \end{aligned} $$
(19)

The signs of the A’s and B’s imply that ψ y  ∈ ( − π, − π/2), meaning a lag between quarter and half the cycle period, while ψ z  ∈ ( − π/2, 0), a phase lag of under a quarter period. These are exactly what the food web structure suggests: because y and z both feed on x and therefore should lag x by a quarter period, but at the same time y feeds on z so it should lag z by a quarter period. The outcome is a compromise between these two expectations determined by the relative importance of the direct and indirect paths between x and y. In particular, when either predator becomes rare, the other converges to a quarter period lag behind x (B = cos(ψ) → 0).

The “positive factor” in the previous paragraph requires some comment. Dividing out some strictly positive factors (such as \(\bar x + k_y\)) it is

$$ D = (\bar x+k_y)\lambda^2+ (g-1)h \bar x^2+ (1 -h k_y-g)\bar x+k_y. $$
(20)

The sign of D is not obvious, and it depends on the cycle period T. Setting λ = 0 in D, the result is a negative multiple of \(\bar z\). To determine the actual sign of D we therefore must estimate the cycle period of the full food chain. This was done by determining the period at the point where limit cycles arise through a Hopf bifurcation, for two cases. The first is the S,x,y system with z absent. This system reduces to two dimensions because asymptotically S + x + y ≡ 1. At the Hopf bifurcation point the eigenvalues of the Jacobian J are ±i λ, so \(\lambda^2 = \det(\mathbf{J})\). Substituting the resulting value of λ 2 into D, gives \(D = \lambda^2 g k_y/(g-1)\) which is positive because g > 1 must hold for the predator to persist. The second case is the complete x,y,z food web but with exponential growth of the prey (i.e., S is absent, and \(\dot x = rx -\) predation. In that case λ 2 at the Hopf bifurcation equals the product of all Jacobian eigenvalues divided by their sum, which is c 2/c 0 where c k is the k th order coefficient in the characteristic polynomial of J. The resulting value of (20) is the product of strictly positive terms and some of the c j , which by the Routh–Hurwitz criterion are all positive at the Hopf bifurcation point, hence D is positive. Code for these calculations in Maxima (maxima.sourceforge.net) is available on request from SPE. The full S, x, y, z model has not yet succumbed to this approach, but numerical evidence supports the belief that D is positive whenever the system has a limit cycle.

E Two prey genotypes

In this Appendix we consider scenarios in which the ESC consists of two prey genotypes. In order for eco-evolutionary cycles to occur, one genotype must be defended against top predator y while the other is vulnerable to predation by y. For simplicity we assume that defense is 100% effective, though numerical solutions of the model assume that defense is imperfect. Also, we assume that λ ≪ 1 because eco-evolutionary cycles typically have a long period (Jones and Ellner 2007), and usually discard terms of O(λ 2) or higher in calculations.

E.1 One predator

The starting point is a system with only the top predator y, the scenario leading to “evolutionary cycles” in our previous studies. In the limit of vanishingly small cost for defense, we have shown (Jones and Ellner 2007) that the two prey genotypes oscillate almost exactly out of phase, while y (lagging the edible prey type by a quarter period) is out of phase with total prey abundance. However numerical solutions of the model (e.g., Yoshida et al. 2003) show the same pattern even when the defense cost is substantial. Here we confirm the numerical results, as a general approximation when the cycle period is long and prey population growth rate is sensitive to changes in limiting substrate concentration.

The persisting prey genotypes are x 0 and x 2 (Fig. 2). Because y feeds only on x 2, it will lag x 2 by a quarter period. To determine the other lags we can therefore regard S and x 0 as forced by the oscillations in x 2. The equations for that subsystem are

$$\begin{array}{rll} \dot S & =& 1- S - \frac{mS (x_2(t)+x_0)}{k_c + S} \\ \dot x_0 & =& x_0\left[\frac{m S}{k_c + S}- 1 \right] \end{array}$$
(21)

with \(x_2(t)=\bar x_2 (1+ \varepsilon \sin(\lambda t))\).

The only nonlinear term is x 0’s functional response f(S) = mS/(k c  + S), which produces in entries containing \(f'(\bar S)\) in the first column of the Jacobian for (21). Defining \(b=f'(\bar S)\) the Jacobian is

$$\begin{array}{rll} \mathbf{J}&=& \begin{bmatrix} -1- b(\bar x_2 + \bar x_0) & - \dfrac{m\bar S}{k_c + \bar S} \\ b\bar x_0& 0 \end{bmatrix}\notag\\ &=& \begin{bmatrix} -1- b(\bar x_2 + \bar x_0) &-1 \\ b\bar x_0& 0 \end{bmatrix}. \end{array}$$

In the notation of Appendix B, the forcing coefficient vector U is proportional to ( − 1, 0). The procedure in Appendix B produces for x 0 (up to a positive rescaling that depends on the forcing amplitude)

$$ A = \lambda \left(b^{-1} + \bar x_2 + \bar x_0\right), \quad B= \lambda^2 b^{-1} - \bar x_0. $$
(22)

In this paper typically λ ≪ 1 because the cycles are long-period cycles driven by evolution, and b will be large because prey are potentially rapid-growing but limited by nutrient scarcity at equilibrium. Then to leading order (22) becomes \(A = \lambda(\bar x_2 + \bar x_0), B = -\bar x_0\). This implies that to leading order in λ, the phase of x 0 relative to x 2 is \(\psi = \arctan(A/B) = - \pi - \lambda(1 + \bar x_2/\bar x_0)\). Therefore

$$\begin{array}{rll} \sin(\lambda t + \psi) &\approx& \sin\left(\lambda t - \pi - \lambda(1+ \bar x_2/\bar x_0)\right)\\ &=& \sin\left(\lambda \left(t- 1 - \bar x_2/\bar x_0\right) - \pi\right). \end{array}$$
(23)

The defended prey type therefore lags the edible prey type by slightly more than half a cycle period, with the lag exceeding half the period by at least one unit of scaled time (which in dimensional time units is the inverse of the dilution rate). The defended prey type is only linked to S, so it lags S by a quarter period, hence S must lag the edible prey by a quarter period plus (to leading order) \(1+\bar x_2/\bar x_0\) time units.

Figure 2 shows a numerical example, and as predicted the defended prey lags the edible prey by slightly more than half the cycle period. This small asymmetry has the important consequence that the peak in total prey abundance occurs in between a peak in defended prey and the subsequent peak in edible prey, and roughly halfway in between. Peaks therefore occur in the sequence

$$ \mbox{defended type} \rightarrow \mbox{total prey} \rightarrow \mbox{edible type} \rightarrow \mbox{predator} $$

with roughly a quarter-period lag between each. The slight deviation of the prey types from being perfectly out of phase is thus the underlying cause for the half-period lag between predator and total prey abundance, which is the defining feature of “evolutionary cycles” in the one-predator food chain.

E.2 Two predators

Now we add back the second predator z, still assuming that there are two prey genotypes, one highly vulnerable to y and the other well defended against y.

Firstly, suppose the defense against y is also effective against z, so the persisting prey clones are x 2 and x 0. Then (as discussed in the main text) the (x 2, y, z) system is equivalent to the two-predator systems with uniform non-evolving prey x considered in Appendices C and D, and the results from those Appendices apply. x 0 is linked to the (x 2, y, z) system only through the substrate, so the results in Appendix E.1 apply, meaning that x 0 will lag x 2 by slightly more than half the cycle period. Figure 13 illustrates these for the IC model; for the LFC model see Fig. 7 in the main text.

Secondly, suppose that defenses against y and against z are incompatible, so the persisting prey clones are x y and x z . Then in the linked food chains model, the (y, x y ) and the (z, x z ) subsystems are both equivalent to a system of one predator feeding on a non-evolving prey type, so y lags x y by a quarter period, and z lags x z by a quarter period or less depending on how much it feeds on W. To determine the lag between the prey types we can take x y as the forcing variable and consider the subsystem

$$\begin{array}{rll} \dot S &=& 1- S - x_z f_z(S) - x_y(t) f_y(S) \\ \dot x_z &=& x_z\left[f_z(S)- hz -1 \right] \\ \dot z &=& z\left[ h x_z + \phi W - 1 \right] \\ \dot W &=& M - W - \phi zW \end{array}$$
(24)

where f y , f z are the algal types’ functional responses. Applying Appendix B gives that the prey types are roughly out of phase: as λ → 0 we have A → 0, B < 0 for the phase lag of x z relative to x y , which implies a half-period lag. The O(λ) term in A is a complicated expression whose sign cannot be determined in general, but it is negative when z’s second food source is relatively unimportant (\(\phi \bar w \ll 1\)) implying that the lag is slightly under a half period.

In the IC model, again considering the other variables as forced by x y , the equations are

$$\begin{array}{rll} \dot S &=& 1- S - x_z f_z(S) - x_y(t) f_y(S) \\ \dot x_z & =& x_z\left[f_z(S)- hz -1 \right] \\ \dot z & =& z\left[ hx_z - \eta y - 1 \right] \\ \dot y & =& y\left[\frac{gx_y(t)}{k_y + x_y(t)} + \eta z - 1\right] \end{array}$$
(25)

Applying Appendix B, we obtain the following results for λ ≪ 1. In the limit h = η = 0 the food web (25) is equivalent to the one-predator evolutionary cycles scenario, so x z lags x y by a half period plus an amount of O(λ). With h and η increasing from 0 in constant proportion (η = dh), the derivative of tan(ψ) is O(λ) > 0, indicating a decrease in lag, and tan(ψ) = O(λ/h) < 0 as h → ∞, so the lag converges to exactly a half period. The h → ∞ limit is not actually relevant, because the x y y link has to be dominant for cycles to occur, but the limit indicates the direction of change as the x z zy pathway increases in importance. y lags x y by a quarter period for h = η = 0 (A = 0(1) < 0,B = O(λ)), and the lag increases to a half period as h →  ∞ with η increasing in proportion (A = O(λ/h) > 0, B = O(1) < 0). z is only linked to x z when η = 0, so it then lags x z by a quarter period. As the x z zy path is strengthened, the linearized analysis predicts an increase in lag to a half period within O(λ), meaning that z’s lag behind x z shrinks.

Figure 14 illustrates these predictions, for η = 0.5,h = 4 meaning that the \(z \leftrightarrow y\) link is weaker than the \(z \leftrightarrow x\) link, but not vanishingly small. As predicted by the linear analysis, peaks in y lag peaks in x y by slightly more than a quarter period (30%), while peaks in z lag peaks in x z by slightly less (18%).

The final scenario with two prey types is coexistence of x z and x 2, the types with no defense against z (Fig. 10). For the LFC model, with x 2 as the forcing function the equations for the forced variables are

$$\begin{array}{rll} \dot S &=& 1- S - x_z f_z(S) - x_2(t) f_2(S) \\ \dot x_z &=& x_z\left[f_z(S)- hz -1 \right] \\ \dot y &=& y\left[\frac{gx_2(t)}{k_y + x_2(t)} - 1\right] \\ \dot z &=& z\left[ h (x_z + x_2(t)) + \phi W - 1 \right] \\ \dot W &=& M - W - \phi zW \end{array}$$
(26)

where f i (S) are the prey functional responses. The method in Appendix B yields that the lag of x z relative to x 2 when λ ≪ 1 is a half-period plus an amount of order λ (specifically, tan(ψ) = A/B = O(λ) with A = O(λ) > 0, B = O(1) < 0 for λ small). Everything else follows from the food web structure and previous results, as described in the main text.

For the IC model, the equations for the forced variables are the first two lines of (26) plus

$$\begin{array}{rll} \dot y &=& y\left[\frac{gx_2(t)}{k_y + x_2(t)} + \eta z - 1\right] \\ \dot z &=& z\left[ h (x_z + x_2(t)) -\eta y - 1 \right] \end{array}$$
(27)

For h, η → 0 the (y, x 2, x z ) subsystem reduces to one-predator evolutionary cycles, so x z and x 2 are nearly out of phase, y is out of phase with total prey and z lags total prey by a quarter period. When the intermediate predator becomes important (increasing h and η in constant proportion), x z and x 2 remain out of phase (A = O(λ/h) > 0, B = O(1) < 0), but the predator lags are completely different: y becomes synchronized with x 2 (A = O(λ) > 0, B = O(1) > 0), while z is out of phase with x 2 (A = O(λ/h) > 0, B = O(1) < 0), and therefore synchronized within O(λ) with x z . The result is that the two predators are out of phase with each other (Fig. 11). Because the two prey types become almost exactly synchronous, two things occur: the cycles of total prey abundance become “cryptic” (Yoshida et al. 2007), and peaks in total prey coincide with peaks in one of the two prey types (at some time when \(\dot{x}_z = \dot{x}_2 = 0\)). As a result, one predator is in phase with total prey abundance, and the other is out of phase with total prey abundance.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ellner, S.P., Becks, L. Rapid prey evolution and the dynamics of two-predator food webs. Theor Ecol 4, 133–152 (2011). https://doi.org/10.1007/s12080-010-0096-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12080-010-0096-7

Keywords

Navigation