Mathematical Biology

1 downloads 0 Views 2MB Size Report
Jul 8, 2014 - 1975; Oaten and Murdoch 1975; Briggs and Hoopes 2004; Gross et al. ..... Begon M, Harper JL, Townsend CR (1990) Ecology: individuals, ...
J. Math. Biol. DOI 10.1007/s00285-014-0820-9

Mathematical Biology

Predator interference and stability of predator–prey dynamics Lenka Pˇribylová · Ludˇek Berec

Received: 5 September 2013 / Revised: 8 July 2014 © Springer-Verlag Berlin Heidelberg 2014

Abstract Predator interference, that is, a decline in the per predator consumption rate as predator density increases, is generally thought to promote predator–prey stability. Indeed, this has been demonstrated in many theoretical studies on predator–prey dynamics. In virtually all of these studies, the stabilization role is demonstrated as a weakening of the paradox of enrichment. With predator interference, stable limit cycles that appear as a result of environmental enrichment occur for higher values of the environmental carrying capacity of prey, and even a complete absence of the limit cycles can happen. Here we study predator–prey dynamics using the Rosenzweig– MacArthur-like model in which the Holling type II functional response has been replaced by a predator-dependent family which generalizes many of the commonly used descriptions of predator interference. By means of a bifurcation analysis we show that sufficiently strong predator interference may bring about another stabilizing mechanism. In particular, hysteresis combined with (dis)appearance of stable limit cycles imply abrupt increases in both the prey and predator densities and enhanced

Electronic supplementary material The online version of this article (doi:10.1007/s00285-014-0820-9) contains supplementary material, which is available to authorized users. L. Pˇribylová (B) Section of Applied Mathematics, Department of Mathematics and Statistics, Faculty of Science, Masaryk University, Kotláˇrská 2, 611 37 Brno, Czech Republic e-mail: [email protected] L. Berec Department of Biosystematics and Ecology, Institute of Entomology, Biology Centre ASCR, ˇ Branišovská 31, 370 05 Ceské Budˇejovice, Czech Republic L. Berec Faculty of Science, Institute of Mathematics and Biomathematics, University of South Bohemia, ˇ Branišovská 31, 370 05 Ceské Budˇejovice, Czech Republic

123

L. Pˇribylová, L. Berec

persistence and resilience of the predator–prey system. We encourage refitting the previously collected data on predator consumption rates as well as for conducting further predation experiments to see what functional response from the explored family is the most appropriate. Keywords Population stability · Functional response · Intraspecific competition · Predation · Predator–prey model Mathematics Subject Classification

92B05 · 92D25 · 37G10 · 37G15

1 Introduction Predator–prey theory is still to a large extent marked by an effort to cope with the legacy of Lotka and Volterra (Lotka 1920; Volterra 1926; Kot 2001) and the paradox of enrichment (Rosenzweig 1971; Vos et al. 2004; Boukal et al. 2007). That is, it still largely dangles round the issue of mechanisms stabilizing and destabilizing predator–prey dynamics, which mostly translates to the issue of functional response (e.g. Murdoch and Oaten 1975; Magnússon 1999; Fryxell et al. 2007; Boukal et al. 2008; Berec 2010). A recently suggested family of functional responses and its effect on predator–prey dynamics is the topic of this study. The most common functional response that occurs in nature appears to be the type II functional response (Hassell et al. 1976; Begon et al. 1990). In its classical formulation, it assumes that predators need some time to handle their prey upon locating it, that searching for prey and handling prey are mutually exclusive activities, and that predators do not interfere in their searching and handling activities (Holling 1959). However, there are other plausible mechanisms that lead to the same functional response form, such as a combination of linear predation rate and density-dependent anti-predator behaviour of prey (Pavlova et al. 2010). Regardless of the underlying mechanism, the per capita predator consumption rate increases, at a decelerating rate, with increasing prey density (Gascoigne and Lipcius 2004). The type II functional response of predators is also a part of the prototype predator–prey model demonstrating the paradox of enrichment where stable oscillations bifurcate out of a stable equilibrium once the environmental carrying capacity of prey exceeds a critical value (Rosenzweig 1971; Kot 2001). The assumption of independence of functional response of predator density, dominating predator–prey theory, is not always satisfactory (Skalski and Gilliam 2001). Individual predators do not always act independently of conspecifics, effects of two or more predators do not always sum up, and competition among predators does not always occur only through prey depletion. Predator interference, occurring when the (per capita) predator consumption rate decreases as predator density increases, is a well-documented and wide-ranging phenomenon. Actually, it is a collective term embracing a number of specific mechanisms. One can think of (i) behavior typical of territorial animals where individuals “waste time” in direct contests thus decreasing time each could otherwise devote to searching for or handling prey (Spradbery 1970; Beddington 1975; Ruxton et al. 1992), (ii) predators that steal already subdued

123

Predator interference and stability

prey from one another (Arnqvist et al. 2006), (iii) prey that can escape predation even if they are already “locked in” by predators (Huisman and DeBoer 1997), (iv) prey which react to an increasing predator density by changing behavior which makes them more difficult to locate and/or capture (Ruxton 1995), (v) prey heterogeneous in their vulnerability to predation (Abrams 1994), or (vi) spatially aggregated predators and prey (Cosner et al. 1999). Predator interference has already found its way into predator–prey models; despite the variety of mechanisms, a common pattern that appears to emerge from modeling predator interference is its stabilizing effect on predator–prey dynamics (Rogers and Hassell 1974; Ruxton et al. 1992; Ruxton 1995; Huisman and DeBoer 1997; Kˇrivan and Vrkoˇc 2004). An intuitive explanation for this is that a decrease in individual consumption rate with increasing predator density can negatively affect reproduction, growth and/or survival of predators and hence the predator population is controlled by its own density in a form of implicit negative density dependence (Begon et al. 1990). A more quantitative, model-based view on predator interference as a stabilizing mechanism largely comes from the Beddington–DeAngelis functional response replacing the Holling type II functional response in the Rosenzweig–MacArthur predator– prey model with the logistic growth of prey. Incorporation of the predator-dependent Beddington–DeAngelis functional response causes the predator isocline to change from vertical to slanted to the right, which in turn has an effect of delaying occurrence of the paradox of enrichment—the Hopf bifurcation marking the onset of predator– prey oscillations occurs for higher values of the environmental carrying capacity of prey (Rosenzweig and MacArthur 1963; Cantrell and Cosner 2001). This stabilizing role of predator interference has even become sort of conventional wisdom in population ecology (Holt 2011). Berec (2010) introduced a novel family of functional responses dependent on predator density, of which the type II functional response as well as common types of functional responses modeling predator interference were special cases. In particular, all of the common types (Hassell–Varley, ratio-dependent, Beddington–DeAngelis) can be viewed as arising from encounter rates of predators with their prey that decrease with increasing predator density (Berec 2010). With this novel family, predator interference was shown to stabilize predator–prey dynamics once its strength was not too high. For other members of this family which modeled higher predator interference strengths three coexistence equilibria co-occurred (Berec 2010). Berec (2010) also showed why this previously unreported multi-equilibrial regime could not occur if only the Beddington–DeAngelis functional response was considered. The study by Berec (2010) considered only the number and character (i.e. stability) of coexistence equilibria. Here we extend this study by looking more thoroughly at the system dynamics when some of the coexistence equilibria are unstable, and characterize system bifurcations that may occur, for an even more general predatordependent functional response. In particular, we are interested in the occurrence of multi-stability regimes, aiming to seek for general conditions on predator-dependent functional responses that would produce multiple coexistence equilibria in a predator– prey model.

123

L. Pˇribylová, L. Berec

2 Model We consider the following predator–prey model with a predator-dependent functional response,   N dN = rN 1 − − P f (N , P), dt K (1) dP = e P f (N , P) − m P . dt In this model, N and P are prey and predator densities, respectively, r is the intrinsic per capita growth rate of prey, K is the environmental carrying capacity of prey, m is the per capita predator mortality rate, and e is the efficiency with which consumed prey are transformed into new predators. The prey population grows logistically in the absence of predators and the predator population dies out exponentially in the absence of prey. For the predator functional response f (N , P) that links prey and predator dynamics we consider a generalized Holling type II functional response, in which the encounter rate λ of predators with their prey is a smooth and declining function of the predator density P: f (N , P) =

λ(P)N 1 + hλ(P)N

(2)

where λ(0) = λ0 and λ (P) < 0; h is the predator handling time of one prey. We thus model predator interference, assuming that the higher is the predator density, the longer it takes any individual predator to encounter a prey. Combining the generic model (1) with the functional response (2), our primary model to be analyzed is   N λ(P)N dN = rN 1 − − P, dt K 1 + hλ(P)N (3) dP λ(P)N =e P − mP . dt 1 + hλ(P)N As a specific example of the function λ(P) we consider the function λ(P) =

λ0 , (b + P)w

(4)

with b > 0, λ0 > 0, and w > 0 (Berec 2010). Notice that the case w = 0 simplifies to a constant λ(P) = λ0 and f (N , P) becomes the Holling type II functional response; the resulting model (1), the Rosenzweig–MacArthur predator–prey model, is the prototype for occurrence of the paradox of enrichment (Rosenzweig 1971). For w = 1, we have f (N , P) =

(λ0 /b)N 1 + P/b + h(λ0 /b)N

(5)

which is actually the Beddington–DeAngelis functional response (Beddington 1975; DeAngelis et al. 1975). Hence, varying w in the predator encounter rate (4) does not produce a single functional response but rather a family of functional responses,

123

Predator interference and stability

whereas varying b or λ0 would give only one. Moreover, it is easy to show that for any fixed predator density P and any parameters λ0 , b and w, the functional response (2) combined with the predator encounter rate (4) is an increasing, concave function of N . On the other hand, for any fixed prey density N , it is a decreasing function of P, as expected for a model of predator interference. The model (1) has invariant isoclines N = 0 and P = 0 that correspond to exponential extinction of the predator without prey and logistic dynamics of the prey without predator, respectively. Stability properties of two boundary equilibria [0, 0] and [K , 0], related to the logistic dynamics of prey with predator extinction, were studied in Appendix A of Berec (2010): whereas the extinction equilibrium [0, 0] is always a saddle point, the prey-only equilibrium [K , 0] is locally stable for any K if e < mh and for K < m/[λ(0)(e − mh)] if e > mh; if K > m/[λ(0)(e − mh)] for e > mh then [K , 0] is a saddle point. In this paper, we focus on coexistence equilibria of the model (3). Berec (2010) showed that for 0 < w < 1, the model (3) with the predator encounter rate (4) could have at most one coexistence equilibrium. For w > 1, however, the non-trivial isoclines may intersect in zero, one or three points with positive N and P and we thus may have zero, one or three coexistence equilibria in this case (Berec 2010). In addition, in the case of three co-occurring coexistence equilibria a saddle point was shown to separate either two stable coexistence equilibria or one stable coexistence equilibrium and one unstable coexistence equilibrium; where just one coexistence equilibrium occurred, this could either be stable or unstable (Berec 2010). In what follows, we look more thoroughly at dynamics of the model (3), and characterize bifurcations that may occur. The non-linear phenomena that we are going to observe include bistability and hysteresis caused by a cusp bifurcation, appearance of a limit cycle due to a Hopf bifurcation, and disappearance of a stable limit cycle due to a homoclinic bifurcation. All of these phenomena have important consequences for predator–prey dynamics. 3 Results 3.1 Cusp bifurcation, hysteresis, and multiple coexistence equilibria A coexistence equilibrium [N ∗ , P ∗ ] of the model (3) satisfies C2 , λ(P ∗ ) λ(P ∗ ) − C2 P ∗ = C1 , λ2 (P ∗ )

N∗ = K

(6) (7)

er m where C1 = e−hm and C2 = K (e−hm) . Hence, if e − hm ≤ 0 then no coexistence equilibrium exists. Therefore, we assume e − hm > 0 from here on, and have C1 > 0 and C2 > 0. Since we assume λ (P) < 0 the Eq. (6) implies that as the predator equilibrium P ∗ > 0 increases so does the prey equilibrium N ∗ > 0. Moreover, the Eq. (7) gives a necessary condition

123

L. Pˇribylová, L. Berec

λ(P ∗ ) > C2

(8)

for [N ∗ , P ∗ ] to be a coexistence equilibrium of the model (3). The Eq. (7) defines an implicit function F(P ∗ ) := P ∗ λ2 (P ∗ ) − C1 (λ(P ∗ ) − C2 ) = 0

(9)

of the non-zero predator equilibrium P ∗ [(having one-to-one correspondence to the non-zero prey equilibrium N ∗ via the Eq. (6)]. In a limit point related to merging and disappearance of two equilibria on a fold of the equilibrium manifold the value of P ∗ satisfies F  (P ∗ ) = 0, which can be equivalently expressed as λ3 (P ∗ ) − C1 (2C2 − λ(P ∗ ))λ (P ∗ ) = 0.

(10)

(recall that we assume λ(P) is smooth). In Appendix A.1 we show that this condition is both necessary and sufficient for vanishing of the Jacobian of the model (3) at a coexistence equilibrium. Since λ (P ∗ ) < 0, the Eq. (10) also yields that for P ∗ to be a limit point we must have λ(P ∗ ) > 2C2 . In particular, λ(0) > 2C2 is a necessary condition for a fold bifurcation to occur. Hence, if λ(P ∗ ) ≤ 2C2 then no coexistence equilibrium of the model (3) with predator interference (i.e.} with λ (P) < 0) can undergo the fold bifurcation. For our specific family of functional responses (2)–(4) we are most interested in how dynamics of the model (3) vary with the strength of predator interference w, i.e. with individual functional responses forming this family. Substituting the predator encounter rate λ(P) (4) into the condition (10) and denoting Λ := λ(P ∗ ) =

λ0 , (b + P ∗ )w

(11)

we get the following necessary condition for the fold bifurcation to occur: bΛ2 + C1 (1 − w)Λ + C1 C2 (2w − 1) = 0 .

(12)

This implies that we can have one limit point for 0 < w < 21 , no limit point for 1 2 ≤ w ≤ 1 and up to two limit points for w > 1. At a critical value wc > 1 that satisfies (1 − wc )2 =

4bC2 (2wc − 1) C1

(13)

the two limit points merge at Λc =

C1 (wc − 1) . 2b

(14)

Note that the critical value wc does not depend on the predator handling time h. Merging of two branches of limit points in a two-parameter space is a cusp bifurcation (Kuznetsov 1998; Seydel 2010). Two-parametric bifurcation diagrams display this situation as a V-shaped curve with a cusp (Fig. 1a). Around the cusp point, this

123

Predator interference and stability

a

b

c

d

Fig. 1 Dynamics of the model (3)–(4) with the emphasis on fold and cusp bifurcations. a Bifurcation diagram in the (w, b) parameter space with the cusp point at wc = 2 and bc = 2 and the V-shaped region. Parameter values: m = 1, e = 1, r = 3, h = 0.25, K = 8 and λ0 = (16/3)2 . b Bifurcation diagram of the equilibrium manifold with w crossing  the cusp point wc = 2 and bc = 2. In this case, Λc = 1 is a double root of (12) and [N ∗ , P ∗ ] = 43 , 10 3 . c Bifurcation diagram of the equilibrium manifold with w crossing the V-shaped multiequilibria region for b = 1.4. d The phase plain portrait for w = 2.2 and b = 1.4, with three coexistence equilibria E 1 , E 2 and E 3 . LP denotes a limit point

V-shaped curve represents the non-linear phenomenon of hysteresis, that is, an Sshaped, single parameter-dependent equilibrium manifold (Fig. 1b, c). Let a single model parameter (in our case the parameter w) be within the V-shaped region (close to the cusp point); whether it changes to low or high values it leaves this region at a fold bifurcation (Fig. 1c). In our particular example the upper branch of the equilibrium manifold is a stable node branch, the middle branch is a branch of saddle points, and the lower branch is initially stable and then loses its stability due to another (Hopf) bifurcation (Fig. 1c), giving rise to a stable limit cycle (Fig. 1d). Leaving the V-shaped region on a branch undergoing the fold bifurcation can be viewed as a catastrophe if it concerns a sudden drop in density of a threatened or exploited population, or an asset if it concerns a sudden drop in density of a pest population, and vice-versa if it concerns a sudden population increase. 3.2 Hopf bifurcation and appearance of limit cycles While the middle branch of an S-shaped, single parameter-dependent equilibrium manifold is always a branch of saddle points, the outer branches need not consist only

123

L. Pˇribylová, L. Berec

of stable equilibria. Hopf bifurcations can occur on these branches in which case limit cycles do appear. Stable coexistence equilibria may lose their stability in a Hopf bifurcation (Fig. 3). A necessary condition for the Hopf bifurcation and concomitant appearance of a limit cycle to occur in a coexistence equilibrium [N ∗ , P ∗ ] is Λ = where we define

(e + hm − ΛK h(e − hm))Λ2 , e(ΛK (e − hm) − m) Λ := λ (P)| P=P ∗

(15)

(16)

We prove this statement in Appendix A.2, showing that the condition (15) is both necessary and sufficient for vanishing of the trace of the Jacobian of the model (3) at a coexistence equilibrium. Again, we make this result more specific by considering our specific family of functional responses (2)–(4). Substituting the specific predator encounter rate λ(P) (4) into the condition (15), we get the third-order polynomial AΛ3 + BΛ2 + CΛ + D = 0 , where

(17)

A = K 2 hb(e − hm)3 > 0, B = K (e − hm)2 (h K er − b(e + hm) − wK e(e − hm)), C = K e(e − hm)(wm(e − hm) − r (2hm + e)), D = emr (e + hm) > 0.

(18)

Numerical simulations show that for any w > 0 this polynomial has two distinct positive roots, the larger of which satisfies the necessary condition (8) for existence of the coexistence equilibrium of the model (3). For general properties of the polynomial (17) see Appendix A.3. As an example, for m = 4, e = 2, r = 1, h = 0.25, and K = 4 the polynomial (17) has the form . Λ3 − 5Λ2 − 4Λ + 6 = 0, with two positive roots Λ1 = 0.811 < C2 = 1 and . Λ2 = 5.527 > C2 . Using the formulas (6) and (7), the corresponding coexistence . equilibrium at which the Hopf bifurcation may occur is [N ∗ , P ∗ ] = [0.724, 0.296] for Λ = Λ2 . For given values of w and b, e.g. w = 0.5 and b = 1, the value . of λ0 can be calculated from the equality λ0 = Λ2 (b + P ∗ )w = 6.293. Figure 2a shows the parameter w crossing the Hopf bifurcation point w H = 0.5 for these selected parameters. This example also shows how we can apply our general results to predator–prey systems with specific forms of the predator encounter rate λ(P). We will not study uniqueness and stability of limit cycles here, but numerical simulations show that commonly they are unique and stable. However, sometimes two limit cycles may coexist (Fig. 2b): the unstable inner cycle comes from a subcritical Hopf bifurcation, while the stable outer cycle arises in a cycle limit point (Fig. 2c). Hence, for an interval of w we have a bistability of an equilibrium and a limit cycle

123

Predator interference and stability

a

b

d

c

Fig. 2 Dynamics of the model (3)–(4) with the emphasis on Hopf bifurcations. a Bifurcation diagram of the equilibrium manifold with the parameter w crossing the supercritical Hopf bifurcation point (HB) w H = 0.5 for K = 4. b Phase plane portrait with two limit cycles for w = 0.8 and K = 33. c Bifurcation diagram of the equilibrium manifold with the parameter w undergoing a subcritical Hopf bifurcation for K = 33. d Bifurcation diagram in the (w, K ) parameter space with supercritical and subcritical Hopf . bifurcation branches. Common parameter values: m = 4, e = 2, r = 1, h = 0.25, b = 1 and λ0 = 6.3. LP denotes a limit point, GH denotes the generalized Hopf point at which the character of the Hopf bifurcation changes from subcritical to supercritical, or vice versa

that is however of a different character than in the hysteresis case. Still, also here a sudden switch in population density may occur. Numerical bifurcations suggest that the interval of w for which the two limit cycles coexist is the one with w < 1. We already know that no fold bifurcations can occur in this interval. If no fold bifurcation exists or a hysteresis is present yet a (supercritical) Hopf bifurcation occurs below the bistability region (from the perspective of parameter w) then the stable limit cycle amplitude increases with decreasing w and relatively quickly approaches separatrices of the extinction saddle point [0, 0] and the prey-only saddle point [K , 0] (Fig. 3a, b). Practically this is fatal for one or both populations, since both can be arbitrarily close to zero for some time for sufficiently low values of w. Chance events can then easily make one or both populations extinct. If prey go extinct first, both populations eventually vanish. If predators go extinct first, prey may either vanish, too, or have good luck, recover and attain its environmental carrying capacity K . An analogous situation occurs if a stable coexistence equilibrium on the lower branch of a hysteresis-caused bistability region loses stability via a (supercritical) Hopf bifurcation and the emerging stable limit cycle succeeds to leave the bistability region (Fig. 3c).

123

L. Pˇribylová, L. Berec

a

c

b

d

Fig. 3 Dynamics of the model (3)–(4) and various locations of the Hopf bifurcation points. a Bifurcation diagram of the equilibrium manifold with the parameter w crossing the cusp point wc = 3, bc = 4 and other parameters set to m = 4, e = 2, r = 5, h = 0.25, λ0 = 933.12 and K = 8. Here Λc = 2.5 and [N ∗ , P ∗ ] = [ 58 , 16 5 ]. b Bifurcation diagram of the equilibrium manifold near the cusp point [wc , bc ] = [3, 4] with b = 3.9 and w crossing the narrow bistable region, distinct to the limit cycle existence region. c Bifurcation diagram of the equilibrium manifold with the parameter w undergoing a supercritical Hopf bifurcation within the bistability region; model parameters: b = 1.3, m = 1, e = 1, r = 3.5, h = 0.25, λ0 = (16/3)2 and K = 5. d Phase plane portrait corresponding to panel b for w = 3

3.3 Bogdanov–Takens bifurcation and disappearance of stable limit cycles The stable limit cycles that appear on the lower branch of the bistability region via a (supercritical) Hopf bifurcation, need not succeed to leave the bistability region. Instead, they may disappear in a homoclinic bifurcation for parameter values still within the bistability region. In particular, they split on the separatrix loop of a saddle point on the unstable middle branch of the hysteresis S-curve (Fig. 4). This causes global bifurcation, yet not species extinction—populations tend to the upper stable coexistence equilibrium that still exists (Fig. 4). Interestingly, the global homoclinic bifurcation is not a bad thing here, since the populations escape the proximity of zero density and prosper at a stable and relatively large abundance (Fig. 5). Nevertheless, as the predator interference strength w further decreases and crosses the left limit point, another homoclinic bifurcation occurs and a stable limit cycle appears that instantly has a large amplitude (Fig. 4). Since no stable equilibrium is left in this part of the parameter space, the anticlockwise trajectories tend to the stable limit cycle that soon

123

Predator interference and stability

a

b

c

d

e

f

g

h

Fig. 4 Dynamics of the model (3)–(4) with the emphasis on homoclinic bifurcations. a Bifurcation diagram of the equilibrium manifold near the cusp point [wc , bc ] = [2, 2] with b = 1.3 and w crossing the bistable region. Other parameters: m = 1, e = 1, r = 3, h = 0.25, λ0 = (16/3)2 and K = 8. b–h Phase plain portraits for various values of w, indicated in the lower left corner of the respective panels. In panel d, a homoclinic bifurcation occurs, at which the stable limit cycle ceases to exist. In panel g, disappearance of the stable node gives rise to another homoclinic bifurcation, at which another stable limit cycle arises (h)

123

L. Pˇribylová, L. Berec 6

Prey density, N

5 4 3 2 1 0

0

20

40

60

80

100

Time Fig. 5 Trajectories of the model (3)–(4) corresponding to different values of w and other parameters as in Fig. 4. Black: w = 2.35, for which two stable attractors (equilibrium and limit cycle) exist. Dark gray: w = 2.2, in which case the system underwent a homoclinic bifurcation and the limit cycle around the lower branch of the S-curve ceased to exist. Light gray: w = 2.1, for which the system underwent a fold bifurcation and no stable equilibrium exists, only a large stable limit cycle to the left of the S-curve

becomes very close to the extinction equilibrium [0, 0] (Fig. 4). Hence, as emphasized earlier, chance events can here easily make one or both populations extinct. It is a Bogdanov–Takens bifurcation that plays a key role in this global bifurcation. This bifurcation is related to the existence of two zero eigenvalues of the respective Jacobian and is the point where the homoclinic curve arises and where the fold, Hopf and homoclinic bifurcations co-occur (Fig. 6a). The bifurcation diagram near the Bogdanov–Takens bifurcation point is topologically equivalent to that shown in Fig. 7; see also Kuznetsov (1998). Since the fold bifurcation is related to merging of two equilibria, the domains 2, 3 and 4 in Fig. 7 correspond to a state space with two nearby equilibria, while the domain 1 corresponds to a state space with no corresponding equilibrium. Crossing the Hopf bifurcation curve H gives rise to a stable limit cycle that generically splits on the homoclinic trajectory P of the nearby saddle (the cycle split curve; Fig. 6a). This transition from fold through Hopf to homoclinic bifurcation is visualized in Fig. 4 for our specific form of the predator encounter rate λ(P) (4). Any Bogdanov–Takens bifurcation point simultaneously satisfies the conditions (10) and (15). For our specific form of λ(P) (4) this leads to the condition AΛ2 + BΛ + C = 0,

(19)

A = K 2 (e − hm)2 (r h − (e − hm)), B = K (e − hm)(em − er − hm 2 − 3mhr ), C = 2mr (e + hm) > 0,

(20)

where the coefficients are

and do not depend on w. Moreover, the expression (13) implies that wc =

123

bK (e−hm)2 Λc 2 +er K (e−hm)Λc −er m . er (K (e−hm)Λc −2m)

(21)

Predator interference and stability

a

b

c

Fig. 6 a Bifurcation diagram for the model (3)–(4) and the parameters w and K , with b = 1.3, m = 1, e = 0.7, r = 3, h = 0.25, and λ0 = 33.734. The Hopf bifurcation curve denotes appearance of a stable limit cycle in the gray region, and in the edge triangle of the V-shaped region near the cusp point we have the bistability region. In the gray region a stable limit cycle exists and disappears on the separatrix loop of the middle saddle point (or a saddle-node) of the S-shaped hysteresis curve; a stable node on the upper branch still exists. b Bifurcation diagram of the equilibrium manifold corresponding to panel a with w = 2.5. c Bifurcation diagram of the equilibrium manifold corresponding to panel a with w = 2.49

Hence, the critical Bogdanov–Takens value is given by w = wc and Λc that solves the Eq. (19). Since B 2 − 4 AC = K 2 (e − hm)3 (m 2 (e − hm) + r 2 (e − hm) + 6emr + 2m 2 hr ) > 0 for e − hm > 0, the Eq. (19) as a quadratic equation of Λ has always two real roots. If A > 0 and B > 0 both are negative and consequently not feasible biologically. If A > 0 and B < 0 two positive roots exist. In the other cases the Eq. (19) has one positive root. In Appendix A.4 we show that the Bogdanov–Takens bifurcation may exist for example for small enough handling times h. For a range of values of the predator interference strength w, two Hopf bifurcation points exist for two different values of prey carrying capacity K (Fig. 6a). Two specific situations, for w = 2.5 (Fig. 6b) and w = 2.49 (Fig. 6c), exemplify an ‘advanced’ stabilization role of predator interference for predator–prey dynamics. In both cases, as K increases the paradox of enrichment initially occurs through the first Hopf bifurcation. Then, for w = 2.5 and further increasing K , the limit cycle amplitude first increases, then starts to decrease, and eventually the limit cycles cease to exist in the

123

L. Pˇribylová, L. Berec Fig. 7 Bifurcation diagram near the Bogdanov–Takens bifurcation point of its normal form y˙1 = y2 , y˙2 = ε1 + ε2 y1 + y12 − y1 y2 . T + is the fold bifurcation curve with one zero and one positive eigenvalue, T − is the fold bifurcation curve with one zero and one negative eigenvalue, H is the Hopf bifurcation curve with two purely imaginary eigenvalues, and P is the homoclinic bifurcation curve. See also Kuznetsov (1998)

second Hopf bifurcation (Fig. 6b). For w = 2.49, however, an analogous sequence of events is interrupted by two homoclinic bifurcations at which the limit cycles first disappear and then reappear again (Fig. 6c). In this latter case, at the first (i.e. for lower K ) homoclinic bifurcation the system abruptly jumps to the upper branch of the multi-equilibrium domain and its persistence is thus enhanced. In both cases the bifurcation diagram is initially analogous to the one corresponding to w = 0 (i.e. to the Rosenzweig–MacArthur model), but the limit cycle eventually disappears due to the existence of hysteresis. The occurrence of the paradox of enrichment is therefore limited. 3.4 Sigmoidally decreasing predator encounter rate λ(P) Here we suggest one more functional form for the declining, predator-dependent encounter rate λ(P), and show how to apply our general theory to another specific example. The function we consider is λ(P) = λ0

bw , w > 1. bw + P w

(22)

It is a sigmoidally decreasing function of the predator density P; the parameter b is the value of P at which λ(P) = λ0 /2 and the parameter w describes the slope of λ(P) at this inflection point P = b. For this function, the formula (10) implies the following necessary condition for the fold bifurcation to occur:

123

Predator interference and stability

a

b

c

d

Fig. 8 Bifurcation diagrams for the model (3) with the sigmoidally decreasing predator encounter rate (22). a Bifurcation diagram for parameters w and λ0 with K = 8, showing the cusp bifurcation (CP). The solid black line is the line of fold bifurcation points (or limit points, LP), the gray line is the line of Hopf bifurcation points (H), and the dashed black line is the line of homoclinic bifurcation points (HL). b Bifurcation diagram of the prey equilibrium manifold near the cusp point. c Bifurcation diagram of the predator equilibrium manifold near the cusp point. In panels b and c, w = 3, BP denotes branching points, and the gray lines mark maximum and minimum of the respective stable limit cycles; going from left to right, the cycles first appear in the Hopf bifurcation, disappear in a homoclinic bifurcation, and appear again in another homoclinic bifurcation. d Bifurcation diagram for parameters λ0 and K with w = 3, showing the cusp bifurcation. Legend as in panel a. Common model parameters: m = 1, e = 1, r = 3, h = 0.25, b = 2.725 and K = 8

   w 2 2C2 Λ + C2 (2w − 1) = 0. Λ + 1−w 1+ λ0 λ0

(23)

We show in the Electronic Supplementary Material that a cusp bifurcation and hence hysteresis can occur for the sigmoidally decreasing predator encounter rate (22), and also how it can be found using our above-given derivations. Figure 8 corroborates this finding. It also demonstrates the existence of a (supercritical) Hopf bifurcation and of two homoclinic bifurcations (and hence a Bogdanov–Takens bifurcation), just as for the monotonically decreasing predator encounter rate (4). Everything said with respect to the latter function thus holds for the former, too.

123

L. Pˇribylová, L. Berec

4 Discussion Since the publication of the Lotka-Volterra predator–prey model in 1920’s (Lotka 1920; Volterra 1926), the stability of predator–prey interactions has been virtually a continuous enterprise in the theoretical ecological literature. Today, a fair amount of stabilizing and destabilizing mechanisms have been identified (Murdoch and Oaten 1975; Oaten and Murdoch 1975; Briggs and Hoopes 2004; Gross et al. 2004; Vos et al. 2004, and references therein). One of the more commonly studied mechanisms has been the intraspecific predator interactions, including predator cannibalism (Kohlmeier and Ebenhöh 1995; Magnússon 1999) and, more often, predator interference (Rogers and Hassell 1974; Ruxton et al. 1992; Ruxton 1995; Huisman and DeBoer 1997). In particular, interference between predators has been found to be a major stabilizing factor in predator–prey systems (DeAngelis et al. 1975; van Voorn et al. 2008). In this paper, we study the effect of a family of functional responses that generalizes many classical descriptions of predator interference on predator–prey dynamics. Most importantly, we show that changes in predator interference strength may give rise to hysteresis and (dis)appearance of stable limit cycles, phenomena that can make both species more vulnerable to extinction. Our major assumption is that the per predator encounter rate of prey λ(P) is a decreasing function of predator density; otherwise, we use the classical Rosenzweig– MacArthur predator–prey model formulation with logistic growth of prey and with predators limited in their consumption by a non-negligible prey handling time (e.g. Kot 2001). The Rosenzweig–MacArthur model of predator–prey dynamics assumes the Holling type II functional response. As a consequence, the predator isocline is vertical, intersects the prey isocline (at most) once, and a supercritical Hopf bifurcation with appearance of stable limit cycles occurs once the predator isocline crosses the vertex of the prey isocline from right to left (Rosenzweig 1971). When the Holling type II functional response is replaced by the Beddington–DeAngelis one, the predator isocline stays straight but leans to the right (Beddington 1975; DeAngelis et al. 1975). As a consequence, the paradox of enrichment occurs for larger carrying capacities when compared to the Holling type II functional response, if at all, which has given rise to the notion that predator interference is a major stabilizing factor of predator– prey dynamics (DeAngelis et al. 1975; van Voorn et al. 2008). Finally, for sufficiently strong predator interference functional responses from the family we consider cause the predator isocline to bend to the right, and if sufficiently bent, it may intersect the prey isocline at three points, thus giving rise to hysteresis and fold and cusp bifurcations. Fold and cusp bifurcations have not been observed in the previous two cases. Moreover, stable limit cycles that may appear on the lower branch of the hysteresis curve (due to the Hopf bifurcation) may break up on its middle branch (due to a homoclinic bifurcation). We have already emphasized in the introduction that varying the parameter w representing the strength of predator interference in the predator encounter rate (4) does not actually produce a single functional response but rather a family of functional responses. As this parameter varies from low to high values, we observe a number of dynamical transitions and may thus characterize predator interference strengths for which different dynamical regimes emerge. We thus reveal that

123

Predator interference and stability

whereas for low values of w just a single, low-density equilibrium exists which can be either stable or unstable, for very high values of w a unique, high-density equilibrium occurs that is always stable. Perhaps most interestingly, for intermediate values of w three coexistence equilibria emerge and the system may exist in two alternative stable states. These alternative states can either be both stable or the lower one, due to a supercritical Hopf bifurcation, may actually be a limit cycle. In the latter case, if the predator interference strength w decreases, the limit cycle amplitude increases which threatens one or both interacting populations. Such a potential threat can be mitigated by an occurrence of a homoclinic bifurcation at an intermediate value of w. This bifurcation causes an interruption of the stable limit cycle and makes the high-density coexistence equilibrium globally attracting. All that said, most persistent systems of predators and prey with respect to the predator interference strength are likely those with intermediate and very high values of w. In any case, all else being equal, functional responses with higher values of w have a stronger potential for stabilizing predator– prey dynamics with predator interference. In our opinion, the most important contribution of our study is that there is one more aspect of stabilization that the predator interference brings, and that this aspect is related to the occurrence of hysteresis. An interesting question is what would happen if we let the predator interference strength w to evolve in the predator by natural selection. Would the eventual evolutionary endpoints correspond to low or high persistence regimes of predator–prey dynamics, i.e. to low or high values of w? The way predator interference should go to evolutionarily is not a priori clear. Predators could decrease the degree of interference through, e.g. evolution of more effective protection of acquired food and/or more effective searching and handling strategies, but they could also invest in better ways of how to steal otherwise subdued prey to others. Does even a potential for polymorphism exist within the predator population? We are convinced that for understanding all of the facets of predator interference, such an evolutionary study would be more than interesting to conduct. Fold and cusp bifurcations mark abrupt changes in state variables and have been subjects of an extensive research. Recently, they have been observed also in both ecological models (Lade et al. 2013) and eco-epidemiological models (Hilker et al. 2009), but their history in ecology goes at least to the work of Ludwig et al. (1978) on the spruce budworm population dynamics. Actually, the whole currently very popular issue of regime shifts is based on the concepts of fold and cusp bifurcations (Scheffer et al. 2001; Scheffer and Carpenter 2003; Lade et al. 2013). And in all of these works, hysteresis has been shown to have immense practical significance with respect to population management and risk of population extinction. In particular, not all instances of hysteresis have a positive effect on population viability as we demonstrate here. For example, in the eco-epidemiological model of Hilker et al. (2009), bistability occurs between two levels of infection prevalence. When a stable limit cycle appears through a supercritical Hopf bifurcation on the lower branch of the hysteresis curve and then splits on the middle branch through a homoclinic bifurcation, the proportion of infectious individuals all of a sudden increases to much higher values. Similarly rich bifurcation structure as revealed in this paper has also been found in a stage-structured

123

L. Pˇribylová, L. Berec

population model (Baer et al. 2006) or recently in a predator–prey model with a strong Allee effect in prey (Aguirre et al. 2014). We give a general protocol for exploring Rosenzweig–MacArthur-like systems in which the predator encounter rate λ(P) of the Holling-type-II-like functional response is a decreasing function of the predator density P (Electronic Supplementary Material). Our analysis is made for a general λ(P) of this type, which depends on a number of model parameters. If we claim that a bifurcation occurs for a critical value of the parameter Λ, we can claim that it occurs also for all involved model parameters, the critical value of which is determined implicitly—and it exists. We use the general analysis to find important starting points for the numerical continuation and plotting of bifurcation diagrams in specific cases, such as finding a cusp point in Electronic Supplementary Material. For a general λ(P) we just can claim that if the predator density exceeds such and such value, the fold (cusp, Hopf, Bogdanov–Takens) bifurcation will not occur. So, for any specific λ(P), we know what to look for and where. Then, we need to use standard tools to draw a complete picture. Another question relevant to our study is whether experiments exist that collated data on predator functional response and that give some support for relatively strong predator interference w > 1 in our convex or sigmoidal form of λ(P). This is a difficult question to answer, since many empirical studies on predator functional response varied just the predator density, in the effort to distinguish among Holling type I–III functional responses. And this is also the reason why the Holling type II functional response appears to dominate predator–prey theory (Skalski and Gilliam 2001). However, Skalski and Gilliam (2001) collected data on 19 predator–prey systems and showed that predator-dependent functional responses they considered (Beddington– DeAngelis, Crowley-Martin, and Hassell–Varley) could provide better descriptions of predator feeding over a range of predator–prey abundances. In particular, treating the Holling type II model as the null hypothesis, comparison with the Beddington– DeAngelis model resulted in rejection of the Holling type II model in favor of the Beddington–DeAngelis model in 18 of the 19 data sets. Still, no single functional response was shown to best describe all of the 19 data sets (Skalski and Gilliam 2001). Given that the family of functional responses proposed in Berec (2010) and explored in this paper is a generalization of both the Beddington–DeAngelis and Hassell–Varley models, it would be more than interesting to refit these 19 data sets by our model and see what values of w best describe the data and whether this family describes all of the 19 data sets well. In general, when predator interference has been tested as an alternative to the Holling type II model, it has almost exclusively been modeled by the Beddington–DeAngelis model. Our family of functional responses provides an additional flexibility and moreover has non-standard dynamical implications for both the predator and prey. Further experiments allowing for a straightforward fitting of prey- and predator-dependent functional responses are also more than welcome. Acknowledgments LP acknowledges support from the project CZ.1.07/2.2.00/15.0203 “Univerzitní výuka matematiky v mˇenícím se svˇetˇe”. LB acknowledges institutional support RVO:60077344.

123

Predator interference and stability

Appendix A A.1 Fold bifurcation Here we show that if a coexistence equilibrium of the model (1) exists then the condition (10) is equivalent to the existence of one zero eigenvalue of the Jacobian of this model evaluated at this equilibrium. Let us define a function g(N , P) such that f (N , P) = N g(N , P) where f (N , P) is the predator functional response (2). In addition, let A denote the Jacobian of the model (1) and [N ∗ , P ∗ ] any coexistence equilibrium of this model. By direct computation we have  A=

 ∗ ∗ ∗ ∗ ∗ ∗ ∗ r − r 2N K − P g − P N g N −N g − P N g P , e P ∗ g + e P ∗ N ∗ gN eN ∗ g + e P ∗ N ∗ g P − m

(24)

where g = g(N ∗ , P ∗ ), g N = ∂g(N ∗ , P ∗ )/∂ N and g P = ∂g(N ∗ , P ∗ )/∂ P. Since for the coexistence equilibrium eN ∗ g = m, we get ∗

∗ ∗ ∗ ∗ ∗ tr A = r − r 2N K − P g − P N gN + e P N g P

and

 det A = r 1 −

2N ∗ K



e P∗ N ∗gP + m P∗g +

m2 eg

P ∗ gN .

(25)

(26)

Since the functional response (2) implies g(N , P) = λ(P)/[1 + hλ(P)N ], we have 2

g N = −g 2 h and g P = λ λg2 .

(27)

Substituting (27) and (6) into (26), we get   det A = P ∗ gm r 1 −

2C2 Λ



Λ Λ2

+

e−hm e



.

(28)

Consequently, the condition (10) is equivalent to the condition det A = 0 at a coexistence equilibrium [N ∗ , P ∗ ] and is a necessary condition for this equilibrium to be a limit point. Generic change in a single model parameter will change sign of the determinant det A. Consequently, the local phase space changes qualitatively. A.2 Hopf bifurcation For a Hopf bifurcation to occur at a parameter value in a two-dimensional system, real parts of the complex eigenvalues of the Jacobian evaluated at the respective system equilibrium have to transversally cross zero (i.e. change sign) when the parameter crosses that value. Denoting by a the bifurcation parameter and by μ(a) ± iω(a) the respective pair of complex eigenvalues, we require μ(ac ) = 0, ω(ac ) > 0, and μ (ac ) = 0 at a bifurcation point a = ac as a necessary conditions for a Hopf bifurcation to occur. we note that

123

L. Pˇribylová, L. Berec

μ(a) =

tr A(a) and ω(a) = det A(a) . 2

(29)

where A(a) denotes the Jacobian evaluated at the system equilibrium corresponding to the parameter a. Now we will prove that if a coexistence equilibrium of the model (1) exists then the condition (15) is necessary and sufficient for vanishing of the trace of the Jacobian of the model (1) at that equilibrium. Denoting g(N , P) = f (N , P)/N then at the coexistence equilibrium   ∗ (30) P ∗ g = r 1 − NK . Consequently, at this equilibrium the trace (25) of the Jacobian of the model (1) simplifies to     ∗ ∗ Λ (31) tr A = −r NK + r 1 − NK N ∗ g h + e Λ 2 . ∗

Since r > 0 and according to (6) NK = CΛ2 > 0 at the coexistence equilibrium, we have      ∗ Λ , (32) tr A = −r NK 1 + K CΛ2 − 1 g h + e Λ 2 m where C2 = K (e−hm) . Since the model (1) implies g = r λ/C1 , a necessary and sufficient condition for vanishing of the trace of the Jacobian of the model (1) at the coexistence equilibrium is

 1= K 1−

m K (e−hm)Λ

 Λ(e − hm)  e



Λ h + eΛ 2

 (33)

or

(K (e − hm)Λ − m)(hΛ2 + eΛ ) (34) eΛ2 which is equivalent to the condition (15). Similar manipulation with the expression (28) and replacement of Λ with the right-hand side of (15) gives the second necessary condition 1=

e − hm r (K Λ(e − hm) − 2m)(ΛK h(e − hm) − (e + hm))2 + > 0. K (e − hm)Λ2 e(ΛK (e − hm) − m) e

(35)

This second condition det A > 0 excludes the neutral saddle case (with two real opposite eigenvalues). A.3 Hopf bifurcation in the special case of λ(P) satisfying (4) For the special case of λ(P) satisfying (4), the condition (15) simplifies to (17). We give here some properties of the polynomial ϕ(Λ) = AΛ3 + BΛ2 + CΛ + D on the left-hand side of (17). For a coexistence equilibrium to exist, we must have e − hm > 0. Consequently, A = K 2 hb(e − hm)3 > 0 and D = er m(e + hm) > 0. The cubic polynomial ϕ(Λ)

123

Predator interference and stability

then satisfies ϕ(0) > 0 and limΛ→∞ ϕ(Λ) = ∞. Let w B =

h K er −b(e+hm) K e(e−hm)

value of w at which B = 0. Analogously, let wC = which C = 0. Since

be the value of w at

wC − w B =

r (2hm+e) m(e−hm)

(e + hm)(bm + K er ) >0 em K (e − hm)

be the

(36)

at the coexistence equilibrium, B decreases linearly with w and C increases linearly with w, B or C need to be negative for any w at the coexistence equilibrium. (The proof is evident: let us suppose that B ≥ 0 and C ≥ 0 concurrently, consequently w ≤ w B and w ≥ wC which is the contrary to w B < wC .) √ 2 −3AC B Let Λmin = − 3A + B 3A denote the local minimum of the cubic polynomial ϕ(Λ). Now we will show that this local minimum indeed exists, i.e.} that the discriminant d(w) = B 2 − 3AC is positive for any w at the coexistence equilibrium. The function d(w) is a quadratic convex function of w with minimum in K er −2be wmin = bhm+2h 2K e(e−hm) . From this, d(wmin ) =

3 (bhm 2 + 4mh K er + 4bem + 4e2 K r )hb(e − hm)4 K 2 > 0. 4

(37)

Consequently Λmin √ exists and for B < 0 it has to be positive. If B ≥ 0, then C < 0 and consequently B 2 − 3AC > B also gives positivity of Λmin . Simulations show that ϕ(Λmin ) ≤ 0 and hence that the polynomial (17) has a positive root for any w, but we don’t have any analytical proof for ϕ(Λmin ) ≤ 0 yet.

A.4 Bogdanov–Takens bifurcation can occur for small handling times h Let us have the handling time h small enough so that r < e−hm h . This implies A < 0 in (20) and since C > 0, there exists one positive real root Λc of the quadratic equation (19). We will show that this Λc satisfies the condition (8) necessary for the m . corresponding P ∗ to be a coexistence equilibrium. Let 0 < Λc ≤ C2 = K (e−hm) Consequently, 0 = AΛc + B +

C Λc

≥ AC2 + B +

C C2

= K er (e − hm) > 0,

(38)

which cannot happen and thus the contrary Λc > C2 is true. Consequently, the critical Bogdanov–Takens bifurcation point occurs at the coexistence equilibrium [N ∗ , P ∗ ] m . Generic change of any two of the that satisfies P ∗ = λ−1 (Λc ) and N ∗ = (e−hm)Λ c parameters λ0 , K , e, h, m, and r will change sign of the trace and determinant of the Jacobian at the corresponding coexistence equilibrium and violate the existence of two zero eigenvalues. This change then necessarily causes a saddle separatrix loop split near the critical parameters.

123

L. Pˇribylová, L. Berec

References Abrams PA (1994) The fallacies of “ratio-dependent” predation. Ecology 75:1842–1850 Aguirre P, Flores JD, González-Olivares E (2014) Bifurcations and global dynamics in a predator–prey model with a strong allee effect on the prey, and a ratio-dependent functional response. Nonlinear Anal Real World Appl 16:235–249 Arnqvist G, Jones TM, Elgar MA (2006) Sex-role reversed nuptial feeding reduces male kleptoparasitism of females in Zeus bugs (Heteroptera: Veliidae). Biol Lett 2:491–493 Baer SM, Kooi BW, Kuznetsov YA, Thieme HR (2006) Multiparametric bifurcation analysis of a basic two-stage population model. SIAM J Appl Math 66:1339–1365 Beddington JR (1975) Mutual interference between parasites or predators and its effect on searching efficiency. J Animal Ecol 44:331–340 Begon M, Harper JL, Townsend CR (1990) Ecology: individuals, populations and communities, 2nd edn. Blackwell Scientific Publications, Oxford Berec L (2010) Impacts of foraging facilitation among predators on predator–prey dynamics. Bull Math Biol 72:94–121 Boukal DS, Sabelis MW, Berec L (2007) How predator functional responses and Allee effects in prey affect the paradox of enrichment and population collapses. Theor Popul Biol 72:136–147 Boukal DS, Berec L, Krivan V (2008) Does sex-selective predation stabilize or destabilize predator–prey dynamics? PLoS One 3(7):e2687 Briggs CJ, Hoopes MF (2004) Stabilizing effects in spatial parasitoid–host and predator–prey models: a review. Theor Popul Biol 65:299–315 Cantrell RS, Cosner C (2001) On the dynamics of predator-prey models with the Beddington–DeAngelis functional response. J Math Anal Appl 257:206–222 Cosner C, DeAngelis D, Ault JS, Olson DB (1999) Effects of spatial grouping on the functional response of predators. Theor Popul Biol 56:65–75 DeAngelis DL, Goldstein RA, O’Neill RV (1975) A model for tropic interaction. Ecology 56:881–892 Fryxell JM, Mosser A, Sinclair ARE, Packer C (2007) Group formation stabilizes predator–prey dynamics. Nature 449:1041–1044 Gascoigne JC, Lipcius RN (2004) Allee effects driven by predation. J Appl Ecol 41:801–810 Gross T, Ebenhöh W, Feudel U (2004) Enrichment and foodchain stability: the impact of different forms of predator–prey interaction. J Theor Biol 227:349–358 Hassell MP, Lawton JH, Beddington JR (1976) The components of arthropod predation. 1. The prey death rate. J Animal Ecol 45:135–164 Hilker FM, Langlais M, Malchow H (2009) The Allee effect and infectious diseases: extinction, multistability, and the (dis-)appearance of oscillations. Am Nat 173:72–88 Holling CS (1959) Some characteristics of simple types of predation and parasitism. Can Entomol 91:385– 398 Holt RD (2011) Natural enemy–victim interactions: do we have a unified theory yet? In: Scheiner SM, Willig MR (eds) The theory of ecology. University of Chicago Press, Chicago, pp 125–162 Huisman G, DeBoer RJ (1997) A formal derivation of the “Beddington” functional response. J Theor Biol 185:389–400 Kohlmeier C, Ebenhöh W (1995) The stabilizing role of cannibalism in a predator–prey system. Bull Math Biol 57:401–411 Kot M (2001) Elements of mathematical ecology. Cambridge University Press, Cambridge Kuznetsov YA (1998) Elements of applied bifucation theory, 2nd edn. Applied Mathematical Sciences, vol 112. Springer, Berlin Kˇrivan V, Vrkoˇc I (2004) Should “handled” prey be considered? Some consequences for functional response, predator–prey dynamics and optimal foraging theory. J Theor Biol 227:167–174 Lade SJ, Tavoni A, Levin SA, Schlüter M (2013) Regime shifts in a social-ecological system. Theor Ecol 6:359–372 Lotka AJ (1920) Undamped oscillations derived from the law of mass action. J Am Chem Soc 42:1595–1599 Ludwig D, Jones DD, Holling CS (1978) Qualitative analysis of insect outbreak systems: the spruce budworm and forest. J Animal Ecol 47:315–332 Magnússon KG (1999) Destabilizing effect of cannibalism on a structured predator–prey system. Math Biosci 155:61–75

123

Predator interference and stability Murdoch WW, Oaten A (1975) Predation and population stability. In: Macfayden A (ed) Advances in ecological research, vol 9. Academic Press, London, pp 1–131 Oaten A, Murdoch WM (1975) Functional response and stability in predator–prey systems. Am Nat 109:289–298 Pavlova V, Berec L, Boukal DS (2010) Caught between two Allee effects: trade-off between reproduction and predation risk. J Theor Biol 264:787–798 Rogers DJ, Hassell MP (1974) General models for insect parasite and predator searching behaviour: interference. J Animal Ecol 43:239–253 Rosenzweig ML (1971) Paradox of enrichment: destabilization of exploitation ecosystems in ecological time. Science 171:385–387 Rosenzweig ML, MacArthur RH (1963) Graphical representation and stability conditions of predator–prey interactions. Am Nat 97:209–223 Ruxton GD (1995) Short term refuge use and stability of predator–prey models. Theor Popul Biol 47:1–17 Ruxton GD, Gurney WSC, Roos AMD (1992) Interference and generation cycles. Theor Popul Biol 42:235– 253 Scheffer M, Carpenter SR (2003) Catastrophic regime shifts in ecosystems: linking theory to observation. Trends Ecol Evol 18:648–656 Scheffer M, Carpenter SR, Foley JA, Folke C, Walker B (2001) Catastrophic shifts in ecosystems. Nature 413:591–596 Seydel R (2010) Practical bifurcation and stability analysis, 3r edn. Springer, Berlin Skalski GT, Gilliam JF (2001) Functional responses with predator interference: viable alternatives to the Holling type II model. Ecology 82:3083–3092 Spradbery JP (1970) Host finding by Rhyssa persuasoria (L.), an ichneumonid parasite of siricid woodwasps. Animal Behav 18:103–114 Volterra V (1926) Fluctuation in the abundance of a species considered mathematically. Nature 118:558–560 van Voorn GAK, Stiefs D, Gross T, Kooi BW, Feudel U, Kooijman SALM (2008) Stabilization due to predator interference: comparison of different analysis approaches. Math Biosci Eng 5:567–583 Vos M, Kooi BW, DeAngelis DL, Mooij WM (2004) Inducible defences and the paradox of enrichment. Oikos 105:471–480

123