On chaos, transient chaos and ghosts in single population models with Allee effects

15
Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 Contents lists available at SciVerse ScienceDirect Nonlinear Analysis: Real World Applications journal homepage: www.elsevier.com/locate/nonrwa On chaos, transient chaos and ghosts in single population models with Allee effects Jorge Duarte a,b,, Cristina Januário a , Nuno Martins b , Josep Sardanyés c,∗∗,1 a ISEL - Engineering Superior Institute of Lisbon, Department of Mathematics, Lisboa, Portugal b Centro de Análise Matemática, Geometria e Sistemas Dinâmicos, Departamento de Matemática, Instituto Superior Técnico, Lisboa, Portugal c Instituto de Biología Molecular y Celular de Plantas, Consejo Superior de Investigaciones Científicas-Universitat Politècnica de València, València, Spain article info Article history: Received 15 August 2011 Accepted 30 November 2011 Keywords: Allee effects Chaos Extinction transients Scaling laws Single species dynamics Theoretical ecology Topological entropy abstract Density-dependent effects, both positive or negative, can have an important impact on the population dynamics of species by modifying their population per-capita growth rates. An important type of such density-dependent factors is given by the so-called Allee effects, widely studied in theoretical and field population biology. In this study, we analyze two discrete single population models with overcompensating density-dependence and Allee effects due to predator saturation and mating limitation using symbolic dynamics theory. We focus on the scenarios of persistence and bistability, in which the species dynamics can be chaotic. For the chaotic regimes, we compute the topological entropy as well as the Lyapunov exponent under ecological key parameters and different initial conditions. We also provide co-dimension two bifurcation diagrams for both systems computing the periods of the orbits, also characterizing the period-ordering routes toward the boundary crisis responsible for species extinction via transient chaos. Our results show that the topological entropy increases as we approach to the parametric regions involving transient chaos, being maximum when the full shift R(L) occurs, and the system enters into the essential extinction regime. Finally, we characterize analytically, using a complex variable approach, and numerically the inverse square-root scaling law arising in the vicinity of a saddle–node bifurcation responsible for the extinction scenario in the two studied models. The results are discussed in the context of species fragility under differential Allee effects. © 2011 Elsevier Ltd. All rights reserved. 1. Introduction The presence of deterministic fluctuations in biological systems has been widely studied over the last decades, especially in the field of theoretical ecology. Time-continuous predator–prey dynamical systems have shown periodic or quasiperiodic dynamics [1,2], as well as chaos when three or more interacting species are considered [3–5]. Interestingly, some other theoretical studies also showed potential chaos in this type of models considering ecologically meaningful parameters based on empirical allometric scaling relationships [6,7]. Equivalent results showing chaotic dynamics were found in discrete theoretical models for single population dynamics [8], metapopulations [9] or high-dimensional ecosystems [10]. More recently, a real full trophic web was studied in laboratory experiments, with the dynamics of the populations suggesting the presence of deterministic chaos [11]. Laboratory experiments with insect populations also hinted chaotic fluctuations to be possible in population dynamics [12,13]. Corresponding author at: ISEL - Engineering Superior Institute of Lisbon, Department of Mathematics, Lisboa, Portugal. ∗∗ Corresponding author. E-mail addresses: [email protected] (J. Duarte), [email protected] (J. Sardanyés). 1 Current affiliation: Gladstone Institute of Virology and Immunology, University of California SF, San Francisco, CA, USA. 1468-1218/$ – see front matter © 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.nonrwa.2011.11.022

Transcript of On chaos, transient chaos and ghosts in single population models with Allee effects

Page 1: On chaos, transient chaos and ghosts in single population models with Allee effects

Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

Contents lists available at SciVerse ScienceDirect

Nonlinear Analysis: Real World Applications

journal homepage: www.elsevier.com/locate/nonrwa

On chaos, transient chaos and ghosts in single population models withAllee effectsJorge Duarte a,b,∗, Cristina Januário a, Nuno Martins b, Josep Sardanyés c,∗∗,1

a ISEL - Engineering Superior Institute of Lisbon, Department of Mathematics, Lisboa, Portugalb Centro de Análise Matemática, Geometria e Sistemas Dinâmicos, Departamento de Matemática, Instituto Superior Técnico, Lisboa, Portugalc Instituto de Biología Molecular y Celular de Plantas, Consejo Superior de Investigaciones Científicas-Universitat Politècnica de València, València, Spain

a r t i c l e i n f o

Article history:Received 15 August 2011Accepted 30 November 2011

Keywords:Allee effectsChaosExtinction transientsScaling lawsSingle species dynamicsTheoretical ecologyTopological entropy

a b s t r a c t

Density-dependent effects, both positive or negative, can have an important impact on thepopulation dynamics of species by modifying their population per-capita growth rates. Animportant type of such density-dependent factors is given by the so-called Allee effects,widely studied in theoretical and field population biology. In this study, we analyze twodiscrete single population models with overcompensating density-dependence and Alleeeffects due to predator saturation and mating limitation using symbolic dynamics theory.We focus on the scenarios of persistence and bistability, in which the species dynamicscan be chaotic. For the chaotic regimes, we compute the topological entropy as well asthe Lyapunov exponent under ecological key parameters and different initial conditions.We also provide co-dimension two bifurcation diagrams for both systems computing theperiods of the orbits, also characterizing the period-ordering routes toward the boundarycrisis responsible for species extinction via transient chaos. Our results show that thetopological entropy increases as we approach to the parametric regions involving transientchaos, being maximum when the full shift R(L)∞ occurs, and the system enters into theessential extinction regime. Finally, we characterize analytically, using a complex variableapproach, and numerically the inverse square-root scaling law arising in the vicinity of asaddle–node bifurcation responsible for the extinction scenario in the two studiedmodels.The results are discussed in the context of species fragility under differential Allee effects.

© 2011 Elsevier Ltd. All rights reserved.

1. Introduction

The presence of deterministic fluctuations in biological systems has beenwidely studied over the last decades, especiallyin the field of theoretical ecology. Time-continuous predator–prey dynamical systems have shown periodic or quasiperiodicdynamics [1,2], as well as chaos when three or more interacting species are considered [3–5]. Interestingly, some othertheoretical studies also showed potential chaos in this type ofmodels considering ecologicallymeaningful parameters basedon empirical allometric scaling relationships [6,7]. Equivalent results showing chaotic dynamics were found in discretetheoretical models for single population dynamics [8], metapopulations [9] or high-dimensional ecosystems [10]. Morerecently, a real full trophic webwas studied in laboratory experiments, with the dynamics of the populations suggesting thepresence of deterministic chaos [11]. Laboratory experiments with insect populations also hinted chaotic fluctuations to bepossible in population dynamics [12,13].

∗ Corresponding author at: ISEL - Engineering Superior Institute of Lisbon, Department of Mathematics, Lisboa, Portugal.∗∗ Corresponding author.

E-mail addresses: [email protected] (J. Duarte), [email protected] (J. Sardanyés).1 Current affiliation: Gladstone Institute of Virology and Immunology, University of California SF, San Francisco, CA, USA.

1468-1218/$ – see front matter© 2011 Elsevier Ltd. All rights reserved.doi:10.1016/j.nonrwa.2011.11.022

Page 2: On chaos, transient chaos and ghosts in single population models with Allee effects

1648 J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

Species fluctuations canhave a negative impact onpopulations survival under the so-calledAllee effect, as theremay exista critical population threshold which, if overcome, could drive species to extinction. Generically, the Allee effect describesa positive correlation between any measure of species fitness and population numbers [14,15]. The Allee effect occurswhen positive density-dependence dominates at low population densities. Positive density-dependent factors have beensuggested to occur in populations with cooperative hunting, conspecific enhancement of reproduction, predator saturationand increased availability of mates [16–18,15], among others. For the case of predator saturation, the Allee effect occursin species predated by a predator with a saturating functional response. Here, the risk of predation for a given individualdecreases as the population density increases. Examples of Allee effects due to predator saturation have been found in fieldstudies on Quercus robur, in which the per-capita rate of acorn loss due to invertebrate seed predators was greatest amongstlow acorn crops and lower on large acorn crops [19]. Moreover, Williams and collaborators [20] showed that some birdscould consume almost the whole population of adult cicadas in their summer decline, but their consumption was lower formuch larger population densities. In addition, some other studies about predation on bird eggs and nestlings [21] as well aspredation on fish species [22], showed that increasing prey population size minimizes vulnerability due to predation.

The Allee effect due tomating limitationmay also be important in sexually reproducing populations,where findingmatesbecomes difficult at low densities. This phenomenon has been shown to be relevant in the pollinization of plants by animalvectors [23] as well as in the fertilization of gametes in benthic invertebrates [24,25]. For this latter case, the fertilization byfree spawning gametes for the species Strongylocentrotus franciscanus can become insufficient at low densities. However, afertilization of 82.2% was found for larger colonies.

The dynamics associated to the previously described Allee effects can be studied for single populations with non-overlapping generations using discrete models, represented by the general difference equation

Nn+1 = Nnf (Nn). (1)Here, Nn is the population density at generation n, and f (Nn) corresponds to the per-capita growth rate of the population.For the sake of clarity, we will briefly explain the dynamics of such models, studied in detail by Schreiber [26]. In thesemodels, very large population densities at time n can give place to very small population numbers at time n + 1. Moreover,there is a population density at time n, namely C , that leads to the maximum population density at n + 1. Two differentdynamical regimes can be found for such systems: (i) persistence and (ii) extinction. Scenario (i) occurs if f (0) > 1 andf (N) > 0, ∀N > 0. Under such conditions, the population will always persist for all positive densities. Scenario (ii), inwhich the population becomes extinct independently of the initial condition, occurs if f (N) < 1, ∀N .

The previous models can undergo an Allee effect when the per-capita growth rate increases at low population numbers.There exists a critical equilibrium density, named A in [26], such that if f (N) < 1 for population densities below A, orf (N) > 1 for some population numbers greater than A, a so-called strong Allee effect takes place. If the population densityfalls at some time below this critical threshold extinction takes place. It is known that under the strong Allee effect, twoother dynamical scenarios are possible (see [26]): (iii) bistability and (iv) essential extinction. In the bistability scenario, thepopulationwill persist orwill disappear depending on the initial densities. In scenario (iv) extinction occurs for almost everyinitial population density (i.e., for a randomly chosen initial condition extinction occurs with probability one).

An interesting phenomenon associated to scenario (iv) is given by transient chaos, which asymptotically drives thepopulations to extinction via a transitory chaotic dynamics [27,26]. Generally talking, transient chaos is tied to the conceptof chaotic saddles [28–30]. Chaotic saddles are bounded sets with fractal structure in both stable and unstable directions.The fractality in the unstable direction implies the existence of an infinite number of gaps of all sizes along its unstablemanifold [31]. Such topological structures have been identified in predator–prey continuous dynamical systems [32,7,31].

The mechanisms generating transient chaos can be different. For instance, chaotic transients can arise due to thecollision of limit sets as occurs in the unstable–unstable pair bifurcations or in riddling bifurcations, responsible forsuperpersistent chaotic transients. In the first type of bifurcation, an unstable periodic orbit of a chaotic attractor collideswith another unstable periodic orbit on the separatrix, giving place to the so-called chaotic crisis [33,34]. Crises have beendescribed in discrete [35,36] and continuous-time predator–prey models [7,31]. More recently, such a phenomenon hasbeen investigated in the context of cooperative hunting [37]. Riddling and riddled basins [30], which have been describedin experiments with coupled electronic circuits [38,39], occur when a fixed point loses its transverse stability as a controlparameter crosses its critical value, colliding with two repellers, which are located symmetrically to an invariant subspace.This type of saddle–repeller pitchfork bifurcation causes the appearance of a tongue near the invariant subspace responsiblefor superpersistent chaotic transients [40]. The chaotic transients found in the models studied in this work are associated tothe collision of the basin of attraction of the attractor corresponding to population persistence with the basin of attractionof the origin [26]. After such a collision, named boundary crisis, the population becomes extinct via transient chaos.

In this study we will analyze with symbolic dynamics theory two discrete ecological models considering Allee effectsdue to predator saturation and mating limitation (see [26] for more details of the studied models). By doing so, we cancharacterize new measures such as the topological entropy for the chaotic dynamics and study their changes toward theboundary crisis found in such models. As we will explain later on, we also provide a general mathematical framework tocalculate the time-to-extinction for any mathematical model represented by a quadratic one-dimensional map undergoinga saddle–node bifurcation here illustrated with the two studied ecological models. Themodels investigated in our study canbe represented with the following general difference equation:

Nn+1 = Nner(1−Nn/K)β(Nn), (2)

Page 3: On chaos, transient chaos and ghosts in single population models with Allee effects

J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 1649

where er(1−Nn/K) is a negative density-dependent factor and the termβ(Nn) is used to introduce a positive density-dependentfactor. Here r is the intrinsic growth rate and K the carrying capacity.

The manuscript is organized as follows. In Section 2 we compute the Lyapunov exponent as well as the topologicalentropy for the model with predation saturation. These two measures are used to investigate the chaotic dynamicsassociated to persistent and transient chaos. In Section 3 we apply the samemeasures for the model with mating limitation.For both models we characterize co-dimension two bifurcation diagrams for key ecological parameters, also showing thedependence of the asymptotic dynamics on the initial condition, characterized in one-dimensional parameter space. InSection 4 we study, both analytically and numerically, the dynamical properties of extinction transients in the context ofdelayed transitions for both models. Using a complex variable approach, we derive the inverse square-root scaling law tiedto saddle–node bifurcations giving place to the extinction scenario.

2. Model with predator saturation

In this section we study the Allee effects due to predator saturation. To do so, we use β(N) = e−m/(1+sN) (see Eq. (2)).Here β(N) corresponds to the probability of escaping predation by a predator with a saturating functional response, beingmthe intensity of predation and s a parameter proportional to the handling time [41]. By non-dimensionalizing Eq. (2), takingxn = NnK−1, we obtain the model under study, given by

xn+1 = xn expr(1 − xn) −

m1 + sKxn

. (3)

Focusing on the dynamics associated to scenarios (i) persistence and (iii) bistability, we first study stability properties ofthe previous model by means of numerical evaluation of the Lyapunov exponent. The Lyapunov exponent, λ, for a point x0is obtained from

λ(x0) = limn→+∞

sup1n

n−1j=0

ln |f ′(xj)|, (4)

where xj = f j(x0). If the absolute value of f ′(xj) is greater than one, the Lyapunov exponent is positive, and then the systemis chaotic having sensitive dependence on the initial conditions. For our system we have

f ′(x) = er(1−x)− m1+sKx

1 + x

msK

(1 + sKx)2− r

.

In Fig. 1 we display several bifurcation diagrams showing the dynamics, together with the corresponding Lyapunovexponent for the chaotic semistability scenario. In all the studied parameters, the system undergoes the period-doublingroute to chaos, with a first flip bifurcation involving periodic dynamics that further evolves by the period-doubling cascadeleading to the Feigenbaum scenario. Under the selected parameter values, the increase in the handling time (i.e., increasein sK ) involves the emergence of chaos, but at some critical threshold (indicated with the dashed red line) the populationcollapses, entering into the essential extinction (iv) scenario. The positive Lyapunov exponent is shown to increase towardthe critical parameter value involving scenario (iv). Fig. 1(b) shows that an increase in the intensity of predation,m, stabilizesthe dynamics, although for large values ofm the population becomes extinct due to the entry into the scenario of extinction(ii). As shown in Fig. 2(d), where the unimodal map is below the diagonal, this is due to a saddle–node bifurcation. Thestabilization of the dynamics at increasing m is not a continuous process since, for some parameter values, the system fallsinto the scenario of essential extinction (iv), where the population has overcome the Allee effect threshold and becomesextinct via transient chaos (see Figs. 1(b) and 2(b)). Finally, the increase in the growth rate, r , displays a similar behavior asthe increase in the handling time, that is, after a saddle–node bifurcation the population reaches a nontrivial steady stateand after a flip bifurcation, the population achieve chaos via period-doubling cascade, and then finally collapses enteringinto scenario (iv).

The investigation of one-dimensional basins of attractionmay be useful to further understand the role of key parameterstogether with the initial population numbers. Fig. 3 shows the asymptotic dynamics for several initial conditions (we use0 < x0 ≤ 2) at increasing predator’s handling time, sK (Fig. 3(1.a)), and predation intensity, m (Fig. 3(1.b)). We heredifferentiate between chaos (blue), order (i.e., not chaotic, shown in red) and extinction (gray) in population dynamics.Chaotic dynamics were identified computing the Lyapunov exponent (see Fig. 3(2.a) and (2.b)). These analyses show thatfor small values of sK the population is in the extinction (ii) scenario, but increasing handling time involves survival ofthe populations which enter into the bistability scenario. Further increase in the handling time involves the appearance ofchaotic dynamics, and the initial population numbers driving the system to extinction is reduced. As previously mentioned,very large values of sK cause the extinction for all possible values of the initial population, now in the essential extinctionscenario. The basins of attraction for predator’s intensity,m, are different. Here, for small values ofm the populations persistchaotically, and after an increase of m, the populations enter into the essential extinction scenario where transient chaoscan occur. Higher predation intensity drives the population to the bistability scenario, which is initially chaotic and thenis stabilized at growing m (see Fig. 3(1.b)). The increase of m in the bistability scenario involves an increase in the initial

Page 4: On chaos, transient chaos and ghosts in single population models with Allee effects

1650 J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

a

b

c

Fig. 1. Bifurcation diagrams for Eq. (3). In (a) r = 4.5 and m = 8. In (b) r = 4.5 and sK = 16. In (c) m = 8 and sK = 16. The diagrams are obtained usingrandom initial conditions x0 ∈ [0, 1], plotting x100 for each parameter value. The dashed red lines indicate the parameter values causing the boundarycrisis. The black arrows indicate the parameter regions where the trajectories reach the stable equilibrium x∗

= 0, for any arbitrary value of x0 . We alsoshow the Lyapunov exponent, λ, computed (using x0 = 0.4) for the following parameter ranges: (a) 4.85 ≤ sK ≤ 12.25, (b) 10.3 ≤ m ≤ 18 and (c)3.25 ≤ r ≤ 4.25. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

a b c d

Fig. 2. Cobweb maps representing the dynamics for the four different scenarios shown in Fig. 1(b), using r = 4.5 and sK = 16. In (a) the system falls inthe (i) persistence scenario, with m = 3 and x0 = c = 0.3144. In (b) we show the dynamics for the (iv) essential extinctions where transient chaos canoccur usingm = 10.2913 and x0 = 0.925. In (c) we show the map for the (iii) chaotic semistability scenario withm = 12 and x0 = c = 0.4898. Finally, in(d) we show the (ii) extinction scenario which occurs after a saddle–node bifurcation, here with m = 20.5 and x0 = c = 0.6075. The arrow indicates thebottleneck region of the ghost causing the delayed transition.

population numbers with asymptotic extinctions. After a critical value of m ≈ 20, the saddle–node bifurcation occurs andthe system also enters into the extinction scenario (ii).

In the following lines we will characterize co-dimension two bifurcation diagrams computing the periods of the orbitsin the survival regimes using symbolic dynamics theory, either in the scenarios of (i) persistence and (iii) bistability. Forthe chaotic semistability behavior we will also compute the topological entropy using this theory. For the sake of clarity,we will explain in detail the steps followed to compute these measures using the theory of unimodal maps [42–44]. Let usconsider an iterated map f on the interval I = [a, b], which is a 2 -piecewise monotonic map with one turning point c . Thus

Page 5: On chaos, transient chaos and ghosts in single population models with Allee effects

J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 1651

x 0

x 0

0.693

0.00

0

0x

x

sK

m

.

(ii) extinction

(iv) essential extinction

order

chaos

order

m

sK

(ii) extinction

(iv) essential extinction

(iii) bistability

(iii)

bis

tabi

lity

–0.75

0.00

–0.75

0.670

(i) persistencechaos

(1.a)

(1.b)

(2.a)

(2.b)

Fig. 3. (1) Basins of attraction for different initial conditions (x0) computed in one-dimensional parameter space using sK (with m = 8) (a); and m (withsK = 16) (b), both with r = 4.5. Extinction domains are indicated in gray color. The initial conditions involving chaotic and ordered persistence areindicated in blue and red, respectively. The dotted line indicates the parametric boundaries after which sustained chaos switches to transient chaos. Thedashed line separates the scenarios of (iii) bistability from (ii) extinction. (2) Computation of the Lyapunov exponent, λ, for magnified regions of the basinsof attraction shown in (1). The transparent planes intersect at the value λ = 0, where bifurcations occur. (For interpretation of the references to colour inthis figure legend, the reader is referred to the web version of this article.)

I is subdivided into the following sets:

IL = [a, c[, IC∗ = {c}, IR = ]c, b],

in such way that the restriction of f to interval IL is strictly increasing and the restriction of f to interval IR is decreasing.Each of such maximal intervals on which the function f is monotone is called a lap of f , and the number ℓ = ℓ (f ) ofdistinct laps is called the lap number of f . Starting with the critical point of f , c (relative extremum), we obtain the orbitO(c) =

xi : xi = f i(c), i ∈ N

. With the purpose of studying the topological properties, we associate to the orbit O(c) a

sequence of symbols, which leads to the itinerary (i(x))j = S = S1S2 · · · Sj · · · , where Sj ∈ A = {L, C∗, R} and

Sj = L if f j(x) < c,

Sj = C∗ if f j(x) = c,

Sj = R if f j(x) > c.

The turning point c plays an important role, since the dynamics of the interval is characterized by the symbolic sequenceassociated to the critical point orbit. When O(c) is a k-periodic orbit, we obtain a sequence of symbols that can becharacterized by a block of length k, the kneading sequence S(k)

= S1S2 · · · Sk−1C∗. The concept of kneading incrementsand kneading-matrix was introduced by Milnor and Thurston [42]. These are power series that measure the discontinuityevaluated at the turning points ci, i = 1, 2, . . . ,m, ofm-modal maps. For the case of unimodal maps we have one kneading-increment defined by ν(t) = θc+(t) − θc−(t), where θx(t) is the invariant coordinate of the sequence S0S1S2 · · · Sj · · ·associated to the itinerary of point x. The invariant coordinate is defined by θx(t) =

j=0 τjt jSj, where τj =J−1

i=0 ε(Si)for j > 0, τ0 = 1 for j = 0,

ε(Si) =

−1 if Si = R0 if Si = C∗

1 if Si = L,

and θc±(t) = limx→c± θx(t).

Page 6: On chaos, transient chaos and ghosts in single population models with Allee effects

1652 J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

Taking as an example the kneading sequence RLLLLLLRRLLC obtained for m = 4 and sK = 9.365 (with r = 4.5) withperiod12,wewill illustrate the computation of the topological entropy. The symbolic sequences that correspond to the orbitsof the points c+ and c−, are c+

−→ R(RLLLLLLRRLLR)∞ and c−−→ L(RLLLLLLRRLLR)∞. Note that the block RLLLLLLRRLLR

corresponds to the sequence RLLLLLLRRLLC where the symbol C is replaced by R because the parity of the block RLLLLLLRRLLis odd. In fact, according to the kneading theory, the shared block of the orbits of the points c+ and c−, RLLLLLLRRLLC , mustbe even, that is, the symbol C is replaced by R in order to have an even number of R symbols. The invariant coordinates are

θc+(t) = R − Rt + Lt2 + Lt3 + Lt4 + Lt5 + Lt6 + Lt7 + Rt8 − Rt9 + Lt10 + Lt11 + Rt12 − Rt13 + Lt14 + · · ·

= R − tR − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9 − Lt10 − Rt11

− t13(R − Lt − · · · − Rt11) − · · ·

= R − tR − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9 − Lt10 − Rt11

(1 + t12 + t24 + · · ·)

= R − tR − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9 − Lt10 − Rt11

/1 − t12

and

θc−(t) = L + Rt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Lt7 − Rt8 + Rt9 − Lt10 − Lt11 − Rt12 + Rt13 − Lt14 − · · ·

= L + tR − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9 − Lt10 − Rt11

+ t13(R − Lt − · · · − Rt11) + · · ·

= L + tR − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9 − Lt10 − Rt11

(1 + t12 + t24 + · · ·)

= L + tR − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9 − Lt10 − Rt11

/1 − t12

.

Separating the terms associated to the symbols L and R, we obtain

ν(t) = N11(t)L + N12(t)R.

In our case, the kneading increment is given by

ν(t) = θc+(t) − θc−(t)= R − t

R − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9

/1 − t12

L + t

R − Lt − Lt2 − Lt3 − Lt4 − Lt5 − Lt6 − Rt7 + Rt8 − Lt9

/1 − t12

=

−1 +

2t2 + 2t3 + 2t4 + 2t5 + 2t6 + 2t7 + 2t10 + 2t11

1 − t12

L +

1 +

−2t + 2t8 − 2t9 + 2t12

1 − t12

R.

The general form of the kneading-matrix, N(t), can be expressed as follows

N(t) =N11(t) N12(t)

.

For the system under study we have

[N(t)]T =

−1 +2t2 + 2t3 + 2t4 + 2t5 + 2t6 + 2t7 + 2t10 + 2t11

1 − t12

1 +−2t + 2t8 − 2t9 + 2t12

1 − t12

T

.

The corresponding kneading determinant, D(t), is

D(t) =D1(t)

1 − ε(L)t= −

D2(t)1 − ε(R)t

=D1(t)1 − t

= −D2(t)1 + t

,

where D1(t) = N12(t) and D2(t) = N11(t). Again, for our model we obtain

D(t) =1 − 2t + 2t8 − 2t9 + t12

(1 − t)1 − t12

.

The topological entropy of f , denoted by htop(f ), is given by htop(f ) = log s, where s is the growth number of a unimodalinterval map f , and is given by s = 1/t∗, being t∗ the root of D(t), which has the lowest modulus. Therefore t∗ =

0.502140 . . . , and the topological entropy is given by

htop(f ) = log

1t∗

= 0.688874 . . . .

Some of the kneading sequences and the topological entropy for our system, obtained as described above, are shown in thefollowing table for small periods (n ≤ 5):

Page 7: On chaos, transient chaos and ghosts in single population models with Allee effects

J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 1653

a

0

5

10

15

20

0

5

10

15

20c

0

5

10

15

20b

0 20151050

5

10

15

20d

mm

m m

(ii) extinction

(iv) essential extinction

(ii) extinction

(iv) e

ssen

tial e

xtin

ctio

n

(iv) e

ssen

tial e

xtin

ction

(ii) extinction

(ii) extinction

(iv) essential extinction

sK

sK

sK

sK

5

4

5

5

3

1

2

4

4

5

1

2

4

5

3

5

1

2

4

5

5

1

2

4

4

35

35

5

5

4

Fig. 4. Periods (in solid black lines) of the dynamics in the parameter space (m, sK) using: (a) r = 3.8269, (b) r = 4, (c) r = 4.5 and (d) r = 4.8. Wealso indicate the regions with asymptotic extinctions given by scenarios (ii) extinction and (iv) essential extinctions. The periods are found in the parametricregions delimited by (i) persistence and (iii) bistability. The number of the periods are indicated on each period line inside a black circle. For all the diagrams,the kneading sequences (at decreasingm) are as follows: period one C , period two (RC)∞ , period four (RLRC)∞ , period five (RLRRC)∞ , period three (RLC)∞ ,period five (RLLRC)∞ , period four (RLLC)∞ and period five (RLLLC)∞ .

Kneadingsequences

Topologicalentropy

C 0RC 0RLRC 0RLRRC 0.414013. . .RLC 0.481212. . .RLLRC 0.543535. . .RLLC 0.609378. . .RLLLC 0.656256. . .

As we mentioned before, we also use the previous steps to first characterize co-dimension two bifurcation diagrams, whichallow us to distinguish the scenarios where species persist, showing the periods of the orbits, as well as to identify theparametric regions involving extinction of the populations. We specifically study the dynamics in the parameter space(m, sK), using four different and increasing values of the intrinsic growth rate of the population, r (see Fig. 4). We show thatthe region of extinction tied to scenario (ii), which is delimited by the period-1 line, is reduced at increasing r . Moreover,increasing r enlarges the region of essential extinctions (iv), where transient chaos can occur. The enlargement of the regionfor scenario (iv) involves that the region where the periods appear is stretched and thus the change of period is found inmore narrow regions of the parameter values.

Page 8: On chaos, transient chaos and ghosts in single population models with Allee effects

1654 J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

0 5 10 150.

0.2

0.4

0.6

htop

b

h top

m

0 5 10 15 200.

0.2

0.4

0.6

Sk

htop

a

sK

h top

Fig. 5. Topological entropy, htop , computed at increasing sK (a) and m (b) using x0 = c and r = 4.5. In (a) we use ( from left to right): m = 4,m = 8and m = 12. In (b) we show, with black points, the values of htop corresponding to sK = 4 (left) and sK = 8 (right). The arrows indicate the values ofsk where the full shift R(L)∞ takes place, and where transient chaos can appear. For this case, the topological entropy remains at the maximum value atincreasing sK values (results not shown). In red, we show the values of htop considering the phases of (i) persistence, (iv) essential extinctions and (iii)chaotic semistability, using sK = 12 (see Fig. 1(b)). Note that in scenario (iv), the topological entropy is maximum, also corresponding to the full shiftR(L)∞ , where the asymptotic dynamics is the extinction of the population. (For interpretation of the references to colour in this figure legend, the readeris referred to the web version of this article.)

The topological entropy, htop, is computed using as control parameters sK (Fig. 5(a)) andm (Fig. 5(b)). In agreement withFigs. 1(a) and 3(a), the topological entropy is positive once the dynamics falls into the chaotic semistability (scenario (iii)),and, as we show, the chaotic region starts at higher values of sK as m increases, that is, as predation intensity increases,chaos appears for longer handling times. This effect can be interpreted in another way: an increase in predation intensitystabilizes the dynamics. For the three values of m studied in Fig. 5(a), we show that the topological entropy increases atgrowing sK , being maximum once the population dynamics enters into the essential extinction regime, which takes placewhen the full shift (RL)∞ occurs and the topological entropy is maximum.

The values of the topological entropy at increasing predation intensity, m, are similar. For the example with sK = 12,indicated in red in Fig. 5(b), we show that for very small values of m (corresponding to the persistence scenario (i)), thetopological entropy is positive because the dynamics is chaotic (equivalently to the bifurcation diagram shown in Fig. 1(b)).Once parameter m reaches the critical value ( for this case m ≈ 4) the topological entropy achieves the full shift (RL)∞,being maximum, and then decreases again once the system enters again into the bistability region. A further increase of mdiminishes the topological entropy as the system undergoes the inverse period-doubling bifurcations.

3. Model with mating limitation

In this section we further analyze the model with Allee effects due to mating limitation studied in [26]. For this caseβ(Nn) = αN/(αN + 1) (see Eq. (2)). Here β(Nn) is the probability of finding a mate and α is the searching efficiency for anindividual. We also non-dimensionalize the system by setting xn = Nn/K , thus obtaining the model that we will explore,given by

xn+1 = xn exp (r(1 − xn))αKxn

1 + αKxn. (5)

We also computed the Lyapunov exponent using Eq. (4), now with

f ′(x) =xαKer(1−x)

1 + xαK

2 −

xαK1 + xαK

− rx

.

Similarly to the previous model studied in Section 2, the system undergoes the period-doubling cascade at increasingboth the individual searching efficiency, αK , and the intrinsic growth rate, r (see Fig. 6). For both cases, at low values of

Page 9: On chaos, transient chaos and ghosts in single population models with Allee effects

J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 1655

x

x

a

b

Fig. 6. Same as in Fig. 1 for Eq. (5) with (a) r = 4.5 and (b) αK = 2. For the Lyapunov exponent diagrams we explore the ranges: (a) 0.15 ≤ αK ≤ 1.42and (b) 2.5 ≤ r ≤ 4.25, using the initial condition x0 = 0.4.

a b c

Fig. 7. Same as in Fig. 2 now showing three cases of the dynamics of the model with mating limitation. In (a) we display the cobweb map for the (ii)extinction scenario, with r = 2, αK = 1.03 and x0 = c = 0.7776 (here the arrow also indicates the slow passage of the iterations through the bottleneckregion of the ghost). In (b) the system is in the (iii) chaotic semistability scenario with r = 4.5, αK = 1 and x0 = c = 0.3829. Finally, we show in (c) the(iv) essential extinction scenario with transient chaos, using r = 4.5 and αK = 1.2 and x0 = 1.4.

the parameters, the system is at the (ii) extinction scenario, and the increase of the parameters involves the saddle–nodebifurcation (see Fig. 7(a)), where the population can persist in the bistability scenario (iii), via fixed point, periodic orquasi-periodic and chaotic dynamics (see Fig. 7(b)). After the critical parameter values, the system enters into the essentialextinction scenario (here also indicated with the dashed red lines), where transient chaos occurs (see Fig. 7(c)).

As we already did for the previous model, we characterize one-dimensional basins of attraction at increasing thesearching efficiency (see Fig. 8). For this case, small values of αK involve extinction independently of the initial condition,thus corresponding to scenario (ii) essential extinction. As the searching efficiency increases, and because of the appearanceof the fixed point allowing persistence due to the saddle–node bifurcation, the dynamics falls into the bistability scenario(iii). The increase of the searching efficiency in this region has two effects: the basin of attraction involving extinction isreduced and the dynamics in the survival phase becomes chaotic. Once a critical threshold in αK is overcome (αK ≈ 14.1in the analysis shown in Fig. 8), the populations enter into the essential extinction regime, and populations become extinctvia transient chaos (as shown in Fig. 7(c)).

Co-dimension two bifurcation diagrams are also investigated in the parameter space (r, αK) (see Fig. 9). We showthat the periods of the orbits found for intermediate values of the searching efficiency, and as the intrinsic growth rateincreases, follow the next sequence: period-1, (C)∞; period-2, (RC)∞; period-4, (RLRC)∞; period-5, (RLRRC)∞; period-3,(RLC)∞; period-5, (RLLRC)∞; period-4, (RLLC)∞; and period-5, (RLLLC)∞. Note that the entrance to the essential extinctioncorresponds to the chaotic dynamics with period 5. In Fig. 9 we show the values of the topological entropy, htop, computed

Page 10: On chaos, transient chaos and ghosts in single population models with Allee effects

1656 J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

0.6635

0.00

x0

chaos

order (iii)

bis

tabi

lity

(iv) essential extinction

(ii) extinction

.

–0.750x

Kα0.0 0.5 1.0 1.5 2.0

0.0

0.0

0.5

0.5

1.01.0

1.5

1.5

2.0

2.0

–0.5

0.0

0.5

0.0

0.5

1.0

1.5

2.0

Fig. 8. Same as in Fig. 3 for αK using r = 4.5.

0 1 2 3

0.

0.2

0.4

0.6

k

hto ph top

b

1 2 3 4 5 60

2

3

4

5

1

(ii) extinction

extinction

(iv) essential

r

a

2

4

5

35 4

5

1

Fig. 9. (a) Period boundaries for the model with mating limitation given by Eq. (5) using as control parameters r and αK . The line in between period1 and period 2 corresponds to the particular dynamics f (c) = c. (b) Topological entropy, htop , at increasing αK , with x0 = c and ( from left to right):r = 4, r = 4.4 and r = 4.8. Note that here the three curves also achieve the (iv) essential extinction scenario (see Fig. 8) reaching the maximum value oftopological entropy, corresponding to the full shift R(L)∞ (indicated with the arrows), which is also found at increasing values of αK (results not shown).

using the searching efficiency as control parameters for three values of the growth rate. These results are in agreement withthe previous figures, which showed that the increase of the searching efficiency involves the period-doubling bifurcationscenario and the entrance into the chaotic semistability regime, which has positive topological entropy. As we found inthe previous model, the topological entropy increases as we approach to the critical value of the parameter involving thecollision of the basins of attraction of the persistence attractor and the origin. Once the full shift (RL)∞, where the topologicalentropy is maximum, occurs, the population goes to extinction via transient chaos.

4. Delayed extinctions: dynamics near a saddle–node bifurcation

In this section we study in detail the extinction scenario (ii) for the system with predator saturation and for thesystem with mating limitation. More precisely, we analytically derive the inverse square-root scaling law arising near thesaddle–node bifurcation by using complex variable techniques. The same approachwas recently applied to study the scalingproperties for saddle–node bifurcations in a general one-dimensional quadratic map [45]. For the sake of clarity, we willsummarize the calculations in the following lines.

For a discrete dynamical system, the evolution of an iterate xn can be represented by the formula

xn+1 = f (xn).

As pointed out in [45], the equation can also be viewed as a difference equation

xn+1 − xn = f (xn) − xn.

Hence, defining g(x) ≡ f (x) − x, gives

xn+1 − xn = g(xn) × 1,

Page 11: On chaos, transient chaos and ghosts in single population models with Allee effects

J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 1657

which can be read ‘‘as n changes by 1 unit, x changes by g(x)’’. In this section we are going to consider firstly the model withpredator saturation to study the scaling properties for the saddle–node bifurcation, analyzing

xn+1 = f (xn,m) = xn expr(1 − xn) −

m1 + sK xn

. (6)

Therefore, definingg(x,m) = f (x,m) − x, (7)

we obtain xn+1 − xn = g(xn,m), with

g(xn,m) = xn

exp

r(1 − xn) −

m1 + sKxn

− 1

.

The equilibrium points occur when g(xn,m) = 0. Besides the trivial solution x∗n = 0, we have two equilibrium points,

x∗

n =sK − 12sK

±

r2(1 + 2sK + (sK)2) − 4rsKm

2rsK,

which are central in the study of the saddle–node bifurcation. We have the following general scenarios:1. For r2(1 + 2sK + (sK)2) − 4rsKm > 0, that is, m < mc , with mc = r2

1 + 2sK + (sK)2

/(4rsK), there are two real

equilibrium points, a stable and an unstable one.2. Form = mc the saddle–node bifurcation occurs. We denote this bifurcation value of the parameter bymc .3. Form > mc , there are two complex equilibrium points.

When m is extremely close to mc (m > mc), many iterations of the map are required to pass through the bifurcationpoint. The reason for this local delay is the presence of a bottleneckingmanifold, which is also named ghost [46]. The scalinglaw measures the number of iterates required to pass through xc = (sK − 1)/(2sK), as a function of the difference of theparameterm to its bifurcation valuemc .

The saddle–node bifurcation for the map displayed in Fig. 2(d) takes place at x = xc when m = mc . The function(7) is analytic with respect to x in the neighborhood of xc . Sufficient conditions for such a bifurcation, when (xc,mc) =

sK−12sK ,

r2(1+2sK+(sK)2)4rsK

, are

g(xc,mc) = 0, Dmg(xc,mc) = 0, (8)and

Dxg(xc,mc) = 0, D2xg(xc,mc) = 0. (9)

Indeed,

g(xn,m) = xn

exp

r(1 − xn) −

m1 + sKxn

− 1

⇒ g(xc,mc) = 0;

Dmg(xn,m) = −xn

1 + sKxnexp

r(1 − xn) −

m1 + sKxn

⇒ Dmg(xc,mc) = −0.05514 . . . = 0;

Dxg(xn,m) = expr(1 − xn) −

m1 + sKxn

xn

−r +

msK

(1 + sKxn)2

+ 1

− 1

⇒ Dxg(xc,mc) = 0;D2xg(xc,mc) = −7.94117 . . . = 0.

In the following lines we consider a = Dmg(xc,mc) and b = D2xg(xc,mc)/2!. As previously mentioned, the equilibrium

points are the solutions of g(xn,m) = 0. Condition (8) allows us to apply the implicit function theorem and obtain that,locally, the zeros of g close to (xc,mc) are the points of the graph of an analytic function m = h(xn), such that h(xc) = mc .Differentiating implicitly g(xn, h(xn)) = 0, we obtain Dxg(xn,m) + Dmg(xn,m)Dh(xn) = 0 and Dh(xn) = −

Dxg(xn,m)

Dmg(xn,m).

For (xn,m) = (xc,mc), we have Dh(xc) = −Dxg(xc ,mc )Dmg(xc ,mc )

. Since Dxg(xc,mc) = 0, we obtain Dh(xc) = 0. Therefore, wemust proceed by computing the second derivative for the purposes of constructing the function h. From Dx[Dxg(xn,m) +

Dmg(xn,m)Dh(xn)] = 0, and since Dh(xc) = 0, we have D2xh(xc) = −

D2x g(xc ,mc )

Dmg(xc ,mc ). Using a Taylor expansion to construct

m = h(x), we obtain

m = h(x) = h(xc) +D2xh(xc)2!

(x − xc)2 + · · · , and thus,

m = mc −ba(x − xc)2 + O(x − xc).

Page 12: On chaos, transient chaos and ghosts in single population models with Allee effects

1658 J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

With the purpose of having the zeros in terms of m, we have to invert the previous expression. Except for m = mc , we getthe following solutions

x = xc ±

ab(m − mc) + O(m − mc).

Form < mc , we have two real solutions

xj(m) = xc +

ab(m − mc)

12eiπ j

+ O(m − mc),

with j = 0, 1 and i =√

−1. For m > mc , we obtain two complex solutions

xj(m) = xc +

ab(m − mc)

12ei

2j+12

π

+ O(m − mc),

also with j = 0, 1.As we already mentioned, our goal is to estimate the number of iterates needed to go from x0 to xn in terms of m(>mc).

Let us consider the equation presented at the beginning of this section, now expressed as

xn+1 − xn = g(xn,m) × 1, (10)

meaning that as n changes by 1 unit, x changes g(xn,m). Thus, the number of iterates required to go from x0 to xn can berepresented by

N =

ni=0

xi+1 − xig(xi,m)

. (11)

Let us consider again expression (10) andmake1x = xn+1 − xn. As a consequence of the application of the implicit theoremfunction, we notice that g(xn,m) = 0 ⇒ 1x → 0, thus,1x gives place to dx. Therefore, in order to compute an approximatevalue of the sum (11), we consider the path integral

γ

dxg(x,m)

= 2π ip−1j=0

Res1g, xj(m)

, (12)

where xj, with 0 ≤ j ≤ p − 1, are the poles in the upper-half plane inside the path γ . In our case, we only consider x0(m).

Therefore,γ

dxg(x,m)

= 2π i Res

1g , xj(m)

, with j = 0 (see [47]). The computation of the residue is given by

Res1g, xj(m)

= lim

x→xj(m)(x − xj(m))

1g(x,m)

.

Since g(xj(m),m) = 0, we can write

Res1g, xj(m)

= lim

x→xj(m)

x − xj(m)

g(x,m) − g(xj(m),m),

which leads to

Res1g, xj(m)

=

1Dxg(xj(m),m)

.

Taking j = 0, the integral (12) is given byγ

dxg(x,m)

=2π i

Dxg(x0(m),m)+ K + O(

√m − mc), m > mc,

where K is a constant independent of m. The number of iterates, N , required to pass through the bifurcation point for themodel with predator saturation, given by xn+1 = xn exp

r(1 − xn) −

m1+sK xn

, is approximated by

Nm =2π i

Dxg(x0(m),m), (13)

where

Dxg(x0(m),m) = expr(1 − x0(m)) −

m1 + sKx0(m)

x0(m)

−r +

msK

(1 + sKx0(m))2

+ 1

− 1;

and x0(m) = xc + ab (m − mc)

12 i.

Page 13: On chaos, transient chaos and ghosts in single population models with Allee effects

J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 1659

φm α K

extt

b

φ

–1/2

a

extt–1/2

Fig. 10. Delayed transitions after the saddle–node bifurcations found in the system with predator saturation (a) and mating limitation (b). In both plotswe show the time to extinction, text , at increasing the bifurcation parameters beyond the critical bifurcation values. In (a) we use φm = m − mc , wheremc = 20.3203125, r = 4.5, sK = 16 and x0 = c = 0.605293. In (b) we use φαK = αKc − αK where αKc = 1.06976, r = 2 and x0 = c = 0.77359.The blue dots show the extinction times computed numerically from (a) Eq. (3); and (b) Eq. (5). The solid lines correspond to the analytical solutions, (13)and (14) characterized in Section 4 for each model. We notice that delayed transitions occur in all discrete models given by unimodal maps undergoing asaddle–node bifurcation. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Now, we consider the system with mating limitation

xn+1 = f (xn, αK) = xn exp (r(1 − xn))αKxn

1 + αKxn,

and follow precisely the procedure described above. We start by defining

g(x, αK) = xn

exp (r(1 − xn))

αKxn1 + αKxn

− 1

.

The equilibrium points occur when g(xn, αK) = 0 and the saddle–node bifurcation of the one-dimensional map of Fig. 7(a) takes place at x = xc = 0.2157056 . . . , when αK = αKc = 0.1400549, for r = 4.5. When αK < αKc , there are twocomplex equilibrium points.

ForαK extremely close toαKc (αK < αKc), many iterations of themap are required to pass through the bifurcation point.In this case, the number of iterates, N , required to pass through the bifurcation point for the model with mating limitation,is approximated by

NαK =2π i

Dxg(x0(αK), αK), (14)

where

Dxg(x0(αK), αK) = −1 −exp(r(1 − x0(αK)))x20(αK)(αK)2

(1 + x0(αK)αK)2

+2 exp(r(1 − x0(αK)))x0(αK)αK

1 + x0(αK)αK−

exp(r(1 − x0(αK)))rx20(αK)αK1 + x0(αK)αK

,

and x0(αK) = xc + ab (αK − αKc)

12 i, with a = DαKg(xc, αKc) = 1.49499 . . . and b = D2

xg(xc, αKc)/2! = −4.63196 . . . .Fig. 10(a) displays the extinction time, text (number of iterates until the fixed point x∗

= 0 is reached), as the intensity ofpredation parameter grows above the limit point bifurcation.We display the extinction times obtained numerically from Eq.(3) (blue dots) and analytically (solid black line) as described above. Note that the analytical results are in perfect agreementwith the numerical ones. The equivalent results are displayed for the model with mating limitation given by Eq. (5) (seeFig. 10(b)).

5. Conclusions

The investigation of Allee effects in population ecology is a current subject of research. Field studies have shown theimportance of population densities in the overall population dynamics under different ecological interactions [19–25]. Thestudy of Allee effects in complex ecosystems has been also carried out by investigating ecological dynamical systems. Forinstance, two classic examples of Allee effects, given by predator saturation and mating limitation, have been thoroughly

Page 14: On chaos, transient chaos and ghosts in single population models with Allee effects

1660 J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661

studied by using one-dimensional discrete models [48,26]. Some works have also extended the previous one-dimensionalsystems to two-species competition models with predator–prey dynamics and Allee effects [49,50], and several conditionsfor species extinction and permanence have been mathematically determined [51]. Other studies analyzing Allee effects incontinuous-time dynamical systems have identified periodic solutions [52–54] and limit cycles governing the coexistencedynamics in predator–prey systems [49]. The study of the spatio-temporal dynamics of predator–prey systems with Alleeeffects have been also carried out in the context of species survival and diffusion-induced chaos [55,56].

In the present article we have studied one-dimensional maps describing single-species dynamics with predatorsaturation and mating limitation Allee effects. We have focused on the chaotic dynamics governing species survival in,also studying the parametric relations involving the essential extinction scenario where transient chaos can occur. Thetheory of symbolic dynamics allowed us to compute the topological entropy and co-dimension two bifurcation diagramsby identifying the periods of the orbits. Both models show common properties. In particular, we found the same period-ordering toward the critical parameter values responsible for transient chaos. We have also found that the appearance oftransient chaos corresponds to themaximum topological entropy, with dynamics (RL)∞, where the full shift takes place andthe population becomes asymptotically extinct.

From an ecological point of view, the first model reveals that longer predator handling times unstabilize the dynamics ofthe populations or even are responsible for extinctions via transient chaos, which are independent of the initial populationvalues. In the bistability scenario,where species can either survive or extinguish, the Allee effect at increasing handling timesmight not be so important because as the dynamics switches from the ordered to the chaotic regime, the basin of attraction tothe extinction dynamics is reduced. That is, only a very small initial populationwill become extinct. The inverse phenomenonis found at increasing predation intensity. For this case, the bistability scenario is mainly characterized by an initial chaoticregime, in which small initial populations will become extinct, but as the dynamics changes to the non-chaotic regime, thebasin of attraction for small population enlarges, and might experience a stronger Allee effect. Interestingly, the regionswith smaller basins of attraction involving extinction are found when the dynamics in the bistable scenario are chaotic. Thesame phenomenon is also found in the model with mating limitation, where the largest basin of attraction for populationextinctions in the bistability scenario is found in the non-chaotic regime. These results indicate that the extinction of thepopulation is less probable at low initial population densities for the chaotic regime. In this context, Allen and co-workersfound that chaos could decrease species extinction in metapopulation dynamical systems under random disturbances, forwhich the extinction probability increased under periodic or quasi-periodic regimes [9]. Our work suggests that, under thechaotic regime, the initial population values driving the population to extinctionmight be less than the ones for the orderedregime.

Finally, we have also investigated delayed extinction transients due to saddle–node bifurcations found in both studiedmodels. Delayed transitions tied to saddle–node bifurcations have been studied in theoretical models for catalytic networks(e.g., hypercycles) (see [57]) and in models of charge density waves [58]. This delaying phenomenon was also empiricallyidentified in an electronic circuit modeling Duffing’s oscillator [59]. Here, we have analytically derived the inverse square-root scaling law near the bifurcation for the time-to-extinction by using a complex variable approach. The expressionsderived analytically for such a time delay are in complete agreement with the numerical results carried out for the twostudied ecological models.

Acknowledgments

This work was partially funded by the Fundação para a Ciência e a Tecnologia through the Program POCI 2010/FEDERand by the Human Frontier Science Program Organization (Grant RGP12/2008). JS wants to thank the Kavli Institute forTheoretical Physics (University of California Santa Barbara) where part of this work was developed (National ScienceFoundation Grant No. NSF PHY05-51164).

References

[1] G.J. Butler, P.Waltman, Bifurcation from a limit cycle in a two predator-one prey ecosystemmodeled on a chemostat, J. Math. Biol. 12 (1981) 295–310.[2] J.D. Murray, in: S.A. Levin (Ed.), Mathematical Biology, Springer-Verlag, USA, 1989.[3] M.E. Gilpin, Spiral chaos in a predator–prey model, Am. Nat. 113 (1979) 306–308.[4] J. Guckenheimer, P. Holmes, Nonlinear Oscillations, Dynamical Systems and Bifurcations of Vector Fields, Springer-Verlag, New York, USA, 1983.[5] W.M. Schaffer, Order and chaos in ecological systems, Ecology 66 (1985) 93–106.[6] A. Hastings, T. Powell, Chaos in a three-species food chain, Ecology 72 (3) (1991) 896–903.[7] K. McCann, P. Yodzis, Nonlinear dynamics and population disappearances, Am. Nat. 144 (1994) 873–879.[8] R.M. May, Simple mathematical models with very complicated dynamics, Nature 261 (1976) 459–467.[9] J.C. Allen, W.M. Schaffer, D. Rosko, Chaos reduces species extinction by amplifying local population noise, Nature 364 (1993) 229–232.

[10] K. Kaneko, T. Ikegami, Homeochaos: dynamics stability of a symbiotic network with population dynamics and evolving mutation rates, Physica D 56(1992) 406–429.

[11] E. Benincà, J. Huisman, R. Weerkloss, K.D. Jönk, P. Branco, E.H. van Nes, M. Scheffer, S.P. Ellner, Chaos in a long term experiment with a planktoncommunity, Nature 451 (2008) 822–825.

[12] R.F. Constantino, R.A. Desharnais, J.M. Cushing, B. Dennis, Chaotic dynamics in an insect populatio, Science 275 (1997) 389–391.[13] J.M. Cushing, R.F. Constantino, B. Dennis, R.A. Desharnais, S.M. Henson, Nonlinear population dynamics: models, experiments and data, J. Theoret.

Biol. 194 (1998) 1–9.[14] W.C. Allee, Animal Aggregations: A Study in General Sociology, University of Chicago Press, Chicago, 1931.[15] F. Courchamp, L. Berec, J. Gascoigne, Allee Effects in Ecology and Conservation, Oxford University Press Inc., New York, 2008.

Page 15: On chaos, transient chaos and ghosts in single population models with Allee effects

J. Duarte et al. / Nonlinear Analysis: Real World Applications 13 (2012) 1647–1661 1661

[16] F. Courchamp, T. Clutton-Brock, B. Grenfell, Inverse density dependence and the Allee effect, Tree 14 (1999) 405–410.[17] P.A. Stephens, W.J. Sutherland, Consequences of the Allee effect for behavior, ecology and conservation, Tree 14 (1999) 401–405.[18] P.A. Stephens, W.J. Sutherland, R.P. Freckleton, What is the Allee effect? Oikos 87 (1999) 185–190.[19] M.J. Crawley, C.R. Long, Alternate bearing, predator satiation and seeding recruitment in Quercus robur , L. J. Ecol. 83 (1995) 683–696.[20] K.S. Williams, K.G. Smith, F.M. Stephen, Emergence of 13-yr periodical cicadas (Cicadidae: Magicicada): phenology, mortality, and predators satiation,

Ecology 74 (1993) 1143–1152.[21] C.G. Wiklund, M. Andersson, Natural selection of colony size in a passerine bird, J. Anim. Ecol. 63 (1994) 765–774.[22] P.R. Elrich, The population biology of coral reef fishes, Annu. Rev. Ecol. Syst. 6 (1975) 211–247.[23] M.J. Groom, Allee effects limit population viability of an annual plant, Am. Nat. 151 (1998) 487–496.[24] N. Knowlton, Threshold and multiple states in coral reef community dynamics, Am. Zool. 32 (1992) 674–682.[25] D.R. Levitan, M.A. Sewell, F. Chia, How distribution and abundance influence fertilization success in the sea urchin Strongylocentrotus franciscanus,

Ecology 73 (1992) 248–254.[26] S.J. Schreiber, Allee effects, extinctions and chaotic transients in simple population models, Theor. Popul. Biol. 64 (2003) 201–209.[27] M. Gyllenberg, A.V. Osipov, G. Söderbacka, Bifurcation analysis of a metapopulation model with sources and sinks, J. Nonlinear Sci. 6 (1996) 329–366.[28] Y.-C. Lai, T. Tél, Transient Chaos, Springer, 2010.[29] K.T. Alligood, T.D. Sauer, J.A. Yorke, Chaos: An Introduction to Dynamical Systems, Springer, 1997.[30] J. Aguirre, R.L. Viana, M.A.F. Sanjuán, Fractal structures in nonlinear dynamics, Rev. Modern Phys. 81 (2009) 333–386.[31] M. Dhamala, Y.-C. Lai, Controlling transient chaos in deterministic flows with applications to electrical power systems and ecology, Phys. Rev. E 59

(1999) 1646–1655.[32] P. Yodzis, S. Ines, Body size and consumer-resource dynamics, Am. Nat. 139 (1992) 1151–1175.[33] C. Grebogi, E. Ott, J.A. Yorke, Chaotic attractors in crisis, Phys. Rev. Lett. 48 (1982) 1507–1510.[34] C. Grebogi, E. Ott, J.A. Yorke, Crises, sudden changes in chaotic attractors, and transient chaos, Physica D 7 (1983) 181–200.[35] K.P. Hadeler, I. Gerstmann, The discrete Rosenzweig model, Math. Biosci. 98 (1989) 49–72.[36] M.G. Neubert, M. Kot, The subcritical collapse of predator populations in discrete-time predator–prey models, Math. Biosci. 110 (1992) 45–56.[37] J. Duarte, C. Januário, N. Martins, J. Sardanyés, Chaos and crisis in a model for cooperative hunting: a symbolic dynamics approach, Chaos 59 (2009)

043102.[38] J.F. Heagy, T.L. Carroll, L.M. Pecora, Experimental and numerical evidence for riddled basins in coupled chaotic systems, Phys. Rev. Lett. 73 (1994)

3528–3531.[39] P. Ashwin, J. Buescu, I.N. Stewart, Bubbling of attractors and synchronisation of chaotic oscillators, Phys. Lett. A 193 (1994) 126–139.[40] Y.-C. Lai, C. Grebogi, J.A. Yorke, S.C. Venkataramani, Riddling bifurcations in chaotic dynamical systems, Phys. Rev. Lett. 77 (1996) 55–58.[41] M.P. Hassell, J.H. Lawton, J.R. Beddington, The components of arthropod predation. I. The prey death rate, J. Anim. Ecol. 45 (1976) 135–164.[42] J. Milnor, W. Thurston, On iterated maps of the interval, Lecture Notes in Math. 1342 (1988) 465–563.[43] Bai-Lin Hao, Wei-Mou Zheng, Applied Symbolic Dynamics and Chaos. Directions in Chaos, vol. 7, World Scientific Publishing Co. Inc., River Edge, NJ,

1998.[44] M. Misiurewicz, W. Szlenk, Entropy of piecewise monotone mappings, Studia Math. 67 (1980) 45–63.[45] J. Duarte, C. Januário, N. Martins, J. Sardanyés, Scaling law in saddle-node bifurcations for one-dimensional maps: a complex variable approach,

Nonlinear Dynam. 67 (2012) 541–547.[46] S.H. Strogatz, Nonlinear Dynamics and Chaos with Applications to Physics, Biology, Chemistry, and Engineering, Westview Press, USA, 2000.[47] E. Fontich, J. Sardanyés, General scaling law in the saddle-node bifurcation: A complex phase space study, J. Phys. A 41 (2008) 015102-1.[48] S.J. Schreiber, Chaos and sudden extinction in simple ecological models, J. Math. Biol. 42 (2001) 239–260.[49] P. Aguirre, E.G. Olivares, E. Sáez, Two limit cycles in a Leslie–Gower predator–prey model with additive Allee effect, Nonlinear Anal. RWA 10 (2009)

1401–1416.[50] W.-X. Wang, Y.-B. Zhang, C.-Z. Liu, Analysis of a discrete-time predator–prey system with Allee effect, Ecol. Complex. 8 (2011) 81–85.[51] Y. Kang, A.-A. Yakubu, Weak Allee effects and species coexistence, Nonlinear Anal. RWA 12 (2011) 3329–3345.[52] S. Padhi, P.D.N. Srinivasu, G.K. Kumar, Periodic solutions for an equation governing dynamics of a renewable resource subjected to Allee effects,

Nonlinear Anal. RWA 11 (2010) 2610–2618.[53] J.R. Graef, S. Padhi, S. Pati, Periodic solutions of some models with strong Allee effects, Nonlinear Anal. RWA 13 (2012) 569–581.[54] F.B. Rizaner, S.P. Rogovchenko, Dynamics of a single species under periodic habitat fluctuations and Allee effect, Nonlinear Anal. RWA 13 (2012)

141–157.[55] L.A.D. Rodrigues, D.C. Mistro, S. Petrovskii, Pattern formation, long-term transients, and Turing–Hopf bifurcation in a space- and time-discrete

predator–prey system, Bull. Math. Biol. 73 (2011) 1812–1840.[56] L.A.D. Rodrigues, D.C. Mistro, S. Petrovskii, Pattern formation in a space- and time-discrete predator–prey system with a strong Allee effect, Theor.

Ecol. (2011) (in press).[57] J. Sardanyés, Dynamics, evolution and information in nonlinear dynamical systems of replicators, Ph.D. Thesis, Universitat Pompeu Fabra, UPF,

Barcelona, 2009.[58] S.H. Strogatz, R.M. Westervelt, Predicted power laws for delayed switching of charge-density waves, Phys. Rev. B 40 (1989) 10501–10508.[59] S.T. Trickey, L.N. Virgin, Bottlenecking phenomenon near a saddle-node remnant in a Duffing oscillator, Phys. Lett. A 248 (1998) 185–190.