Plan
Comptes Rendus

Ecology/Écologie
A focus on long-run sustainability of a harvested prey predator system in the presence of alternative prey
Comptes Rendus. Biologies, Volume 333 (2010) no. 11-12, pp. 841-849.

Résumé

Within the framework of a general equilibrium model we study the long-run dynamics of a prey-predator model in the presence of an alternative prey. Our results show that sustainability, i.e. a positive value of the population in the long run, essentially depends on individual harvesting efforts and digesting factors relative to alternative prey. A detailed bifurcation analysis evidences the richness of possible long-run dynamics. Our model clearly shows that the role of an alternative prey must be taken into consideration when studying prey-predator dynamics.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crvi.2010.09.001
Mots clés : Prey-predator, Stability, Bifurcation, Limit cycle, Sustainability
T.K. Kar 1 ; S.K. Chattopadhyay 2

1 Department of Mathematics, Bengal Engineering and Science University, Shibpur, Howrah 711103, India
2 Department of Mathematics, Sree Chaitanya College, Habra, 24-Pgs (North), West Bengal, India
@article{CRBIOL_2010__333_11-12_841_0,
     author = {T.K. Kar and S.K. Chattopadhyay},
     title = {A focus on long-run sustainability of a harvested prey predator system in the presence of alternative prey},
     journal = {Comptes Rendus. Biologies},
     pages = {841--849},
     publisher = {Elsevier},
     volume = {333},
     number = {11-12},
     year = {2010},
     doi = {10.1016/j.crvi.2010.09.001},
     language = {en},
}
TY  - JOUR
AU  - T.K. Kar
AU  - S.K. Chattopadhyay
TI  - A focus on long-run sustainability of a harvested prey predator system in the presence of alternative prey
JO  - Comptes Rendus. Biologies
PY  - 2010
SP  - 841
EP  - 849
VL  - 333
IS  - 11-12
PB  - Elsevier
DO  - 10.1016/j.crvi.2010.09.001
LA  - en
ID  - CRBIOL_2010__333_11-12_841_0
ER  - 
%0 Journal Article
%A T.K. Kar
%A S.K. Chattopadhyay
%T A focus on long-run sustainability of a harvested prey predator system in the presence of alternative prey
%J Comptes Rendus. Biologies
%D 2010
%P 841-849
%V 333
%N 11-12
%I Elsevier
%R 10.1016/j.crvi.2010.09.001
%G en
%F CRBIOL_2010__333_11-12_841_0
T.K. Kar; S.K. Chattopadhyay. A focus on long-run sustainability of a harvested prey predator system in the presence of alternative prey. Comptes Rendus. Biologies, Volume 333 (2010) no. 11-12, pp. 841-849. doi : 10.1016/j.crvi.2010.09.001. https://comptes-rendus.academie-sciences.fr/biologies/articles/10.1016/j.crvi.2010.09.001/

Version originale du texte intégral

1 Introduction

The modelling of commercial exploitations of renewable biological resources is a challenging task, as it includes the non-linear interaction of biological, economic and social components. In recent past years, many researchers have paid attention to the management of renewable resources, considering different parameters and externalities. In the management of common property renewable resources (such as fisheries, forestry and wildlife) harvested by competing individuals, societies or countries, the problem known as “The tragedy of commons” after Hardin [1]. Later on Clark [2], Masterton-Gibbons [3], Conrad [4] discussed many issues and techniques on this subject. Clark [2] discussed the problem of non-selective harvesting of two ecologically independent populations obeying logistic growth. Kar and Chaudhuri [5] discussed non-selective harvesting of a multispecies fishery. There are also some papers on competitive fish-species and on prey-predator models (see Gordon [6]; Datta and Mirman [7]; Bischi and Kopel [8]; Dubey et al. [9]; Bischi et al. [10]). Most of the prey-predator models consider a single prey called the focal prey; however, many authors have emphasized that the presence of alternative foods can effect biological control through a variety of mechanisms. For example, the presence of one prey species can have negative effects on the population of another prey species, by allowing the population of a shared predator to increase, thus leading to higher predation rates upon both prey items. In contrast, alternative prey can also lower predation on focal prey because of predator preference for alternative prey resources (Abram and Matsuda [11]). In such instances, the alternative prey can have a positive effect on population densities of the focal prey. Many massive piscivores, including spiny dogfish in the North-West Atlantic, and Pacific hake, undertake extensive seasonal migrations on a spatial scale much larger than the habitat occupied by some of their prey, and the development of realistic models for these prey-predator systems will require consideration of alternative prey. Vence [12], Spencer and Collie [13] have considered alternative prey in their models.

Methods used for managing predators, and associated public perception, are a crucial consideration in developing effective systems of management for predators and prey. Controlling predator populations to reduce predation on a threatened or endangered species may be difficult to achieve. Keeping wolf and cougar (wild American cat) populations in check might require reducing alternate prey. Control of mule deer populations in Sierra mountains of California might be required in addition to reductions in cougar numbers to prevent extirpation (complete destruction) of Sierra Nevada big horned sheep populations. The Alberta Government has issued additional hunting permits for moose and deer in an attempt to reduce alternate prey for wolves, hoping to reduce wolf numbers and thereby predation pressure on caribou.

In this study, we develop a simple, two species predator-prey model that incorporates the effect of alternative prey. We consider logistic production function of prey and a Holling type-II predator functional response (Holling [14]) so that there may present multiple equilibrium population levels. In our model, the growth rate of prey is formulated as

dNdT=rN(1N/K)α1NPa+N(1)
where N = N(T) = size of prey at time T, P = P(T) = size of predator at time T, K = environmental carrying capacity of prey, r = intrinsic growth rate of prey, α1 = predation co-efficient, a = half-saturation constant.

The growth rate of predator population is taken as

dPdT=β1α1NPa+N+d1P(1N/K)γ1P(2)
where β1 is a conversion factor (we assume β1 < 1, since the whole biomass of the prey is not transferred to the biomass of the predator), d1 is the digesting factor relative to alternative prey and γ1 is mortality rate of predator population.

Here, as the focal prey population (N) increases, the predator uses less alternative prey and when N → K the mass of alternative prey consumed tends to zero, and conversely, as the focal prey decreases, the predators increase their feeding on alternative prey. If N → 0 then dP/dT → (d1 − γ1)P. Thus even in case of the extinction of the focal prey the predator population maintains its growth rate, varying linearly with its density.

Now we introduce dimensionless variables by the following substitutionN = ax, P = ray/α1, T = t/rthen our Eqs. (1) and (2) become

dxdt=x(1x/Ka)xy1+x(3)
dydt=β1α1rxy1+x+d1ry(1x/Ka)γ1ry(4)

LettingKa=η,   β1α1r=β,   d1r=d,   γ1r=γwe can write the governing equations as

dxdt=x(1x/η)xy1+x(5)
dydt=βxy1+x+dy(1x/η)γy(6)

For considering the exploited prey-predator system, we introduce scaled harvesting efforts E1 and E2 for prey and predator, respectively and then the equations governing our model become

dxdt=x(1x/η)xy1+xE1x(7)
dydt=βxy1+x+dy(1x/η)γyE2y(8)

The Note is organized as follows. The existence of different equilibrium points are considered in section 2. Stability analysis and bifurcations are given in section 3. In section 4, detailed numerical simulations are given. The Note ends with a conclusion in section 5.

2 Equilibrium of the model

Lettingϕ1(x)=xx+1,   ϕ2(x)=(1x/ηE1)(x+1)andϕ3(x) = βϕ1(x) + d(1 − x/η) − (E2 + γ)the governing equations (7) and (8) become

dxdt=ϕ1(x)[ϕ2(x)y](9)
dydt=yϕ3(x)(10)

The prey zero growth lines are obtained by setting dx/dt = 0, which gives ϕ1(x) = 0, y = ϕ2(x) that is x = 0 (y-axis) and the curve y = ϕ2(x). The curve y = ϕ2(x) passes through the points (0, 1 – E1) and (η(1 − E1), 0).

We see that

ϕ2(x)=(1x/ηE1)+(1+x)(1/η)=1E11/η2x/η.

Thus, ϕ2(0)=1E11/η, andϕ2′(η(1 − E1)) = − (1 − E1 + 1/η)

Thus if η(1 − E1) > 1, then ϕ2(0) and ϕ2(η(1E1)) are of opposite signs and so ϕ2(x) has a local maximum between x = 0 and x = η(1 − E1). The predator zero growth lines are obtained by setting dy/dt = 0 i.e. y = 0 (x-axis) and ϕ3(x) = 0. Thus isoclines of predator other than x-axis are the zeros of ϕ3(x).

These are given by

βϕ1(x)+d(1x/η)(E2+γ)=0,
or,
dx2/η+(β+dd/ηE2γ)x+dE2γ=0(11)
The zeros of ϕ3(x) are given by
x2,   x1=η2d[(d+βd/ηE2γ)±(d+βd/ηE2γ)24dη(E2+γd)](12)

Here we investigate the different cases for positive zeros of ϕ3(x) to have interior equilibrium. From the above discussion we already have two boundary equilibria P0(0, 0), P1((1 − E1)η, 0) while P1 exists if E1 < 1.

Case 1.

If (d+βdηE2γ)>0, and E2 + γ − d < 0, that is, if E2<Min{(βdη)+dγ,   dγ}, then ϕ3(x) has only one positive zero given by

x2=η2d[(d+βd/ηE2γ)+(d+βd/ηE2γ)24dη(E2+γd)](13)
and the corresponding value of y is y2=ϕ2(x2)=(1+x2)(1x2ηE1), which lies in R2+ if E1 < 1 − x2/η.

Thus, in this case, the trivial equilibrium P0(0, 0) always exists, the boundary equilibrium P1((1 − E1)η, 0) exists if E1 < 1 and the only interior equilibrium P2(x2, ϕ2(x2)) exists if E1 < 1 − x2/η and E2<Min{(βdη)+dγ,dγ}.

Here it can be observed that if P2 exists then P1 also exists.

Case 2.

Let (d+βdηE2γ)<0, E2 + γ − d < 0. In this case only one positive interior equilibrium may be obtained if (d − γ) > 0, (β − d/η) < 0 i.e. if d > max{γ, βη} and (d+βdηγ)<E2<(dγ).

In this case x = x2 is a positive zero of ϕ3(x) and P2(x2, ϕ2(x2)) is a positive interior equilibrium while E1 < 1 − x2/η along with the above conditions. Therefore, in this case, the trivial equilibrium P0(0, 0) always exists, the boundary equilibrium P1((1 − E1)η, 0) exists if E1 < 1 and the interior equilibrium P2(x2, ϕ2(x2)) exists if E1 < 1 − x2/η, d > max{γ, βη} and (d+βdηγ)<E2<(dγ). Here also we see that the conditions of existence of P2 suffice the existence of P1.

Case 3.

Let

(d+βdηE2γ)>0,   E2+γd>0
and
(d+βdηE2γ)2>4dη(E2+γd)

More explicitly,

dγ<E2<dγ+βdη
and
(d+βdηE2γ)2>4dη(E2+γd)
for γ < d < ηβ, or, d<γ<(βdη+d).

Then there are two positive zeros of ϕ3(x) which are x1, x2 given in (12) and

ϕ3(x)=>0<0   for   x1<x<x2   for   0<x<x1,   x>x2,
i.e. ϕ3(x1)>0 and ϕ3(x2)>0. In this case all four equilibria exist. The boundary equilibrium P1((1 − E1)η, 0) exists if the condition:(H1): E1 < 1 is satisfied. One interior equilibrium P2(x1, ϕ2(x1)) exists if the following conditions are satisfied:(H2): (d+βdηE2γ)2>4dη(E2+γd),   E1<1x1/ηwith γ < d < ηβ, or d<γ<(βdη+d), and the other interior equilibrium P3(x2, ϕ2(x2)) exists if the conditions below are satisfied:(H3): dγ<E2<dγ+βdη,(d+βdηE2γ)2>4dη(E2+γd),   E1<1x2/ηwith γ < d < ηβ or d<γ<(βdη+d).

It is observed that at P2 & P3 both the species co-exist under some conditions stated above.

3 Stability analysis

The Jacobian matrix for our model is

J(x,y)=ϕ1(x)(ϕ2(x)y)+ϕ1(x)ϕ2(x)ϕ1(x)yϕ3(x)ϕ3(x).(14)

Since ϕ1(x)=x1+x, therefore ϕ1(x)=1(1+x)2 and from the expressions of ϕ2(x), ϕ3(x) we get

ϕ2(x)=1E11η2xη,ϕ3(x)=βϕ1(x)dη=β(1+x)2dη.

We first investigate the stability at the trivial equilibrium P0(0, 0), the boundary equilibrium P1((1 − E1)η, 0), and then the discussion of stability of interior equilibrium will be considered in different cases as they appear in terms of their existence.

At P0(0, 0) the corresponding community matrix is simplified to be

J(P0)=1E100dE2γ,
whose eigenvalues are (1 − E1) and (d − E2 − γ). Therefore the equilibrium at P0 is stable if
E1>1   and   E2>dγ(15)

At P1((1 − E1)η, 0) the matrix J is evaluated as

J(P1)=(1E1)ϕ1((1E1)η)0ϕ3((1E1)η)
whose eigen values are λ1 = − (1 − E1) < 0, since E1 < 1 for existence of P1 and λ2 = ϕ3((1 − E1)η).Now λ2 < 0 if
β(1E1)η1+(1E1)η+d(1(1E1)ηη)(E2+γ)   >0
i.e. if
E2>β(1E1)η1+(1E1)η+dE1γ.
Thus, P1 is a stable equilibrium if
E2>β(1E1)η1+(1E1)η+dE1γ(16)
and E1 < 1

Now we go through the following cases for discussing the stability at interior equilibrium, taking the references of the cases of existence of equilibrium points discussed earlier.

Case 1.

LetE2<Min{(βdη)+dγ,dγ}   and   E1<1x2/η,then only the positive interior equilibrium is P2(x2, ϕ2(x2)). At this equilibrium

J=ϕ1(x2)ϕ2(x2)ϕ1(x2)ϕ2(x2)ϕ3(x2)0

The characteristic equation is

λ2λϕ1(x2)ϕ2(x2)+ϕ1(x2)ϕ2(x2)ϕ3(x2)=0(17)

Since x1, x2 are the roots of ϕ3(x) = 0, therefore from Eq. (11) we have

ϕ3(x)=dη[(xx2)(xx1)]
where x2 > x1 and hence
ϕ3(x)=dη[(xx2)+(xx1)]
so that
ϕ3(x2)=dη(x2x1)<0

Thus ϕ1(x2)ϕ2(x2)ϕ3′(x2) < 0 as ϕ1(x2), ϕ2(x2) are positive and hence (17) has one positive and one negative root. Therefore in this case P2(x2, ϕ2(x2)) is an unstable equilibrium.

Case 2.

Let

(d+βdηE2γ)<0   and   E2+γd<0.

In this case P2(x2, ϕ2(x2)) is the only positive interior equilibrium. As in Case 1, ϕ3(x2)   <0 and the equilibrium is unstable at P2(x2, ϕ2(x2)).

Case 3.

Let

(d+βdηE2γ)>0,   E2+γd>0
and
(d+βdηE2γ)2>4dη(E2+γd).

In this case two interior equilibrium points P2(x1, ϕ2(x1)) and P3(x2, ϕ2(x2)) exist.

Now we investigate the stability at P2 and P3.

At P2(x1, ϕ2(x1)) the community matrix is reduced to

J(P2)=ϕ1(x1)ϕ2(x1)ϕ1(x1)ϕ2(x1)ϕ3(x1)0(18)
whose characteristic equation is
λ2λϕ1(x1)ϕ2(x1)+ϕ1(x1)ϕ2(x1)ϕ3(x1)=0(19)

Now ϕ1(x1)   ϕ2(x1)   ϕ3(x1) are all positive under the conditions of existence of P2(x1, ϕ2(x1)). Thus by Descartes’ rule of signs, P2 is a node or a focus or a centre. By Routh-Hurwitz condition, the equilibrium at P2(x1, ϕ2(x1)) is stable if ϕ2(x1)<0, and is unstable if ϕ2(x1)>0. A limit cycle is expected closed to the curve ϕ2(x1)=0. Now ϕ2(x1)=0 implies that 1E11η2x1η=0, or 1E1=1η(1+2x1).

Or,

E1=11η(1+2x1)=11η1d[(d+βdηE2γ)(d+βdηE2γ)24dη(E2+γd)]
i.e.
E1=ϕ(E2)(20)
where
ϕ(E2)=11η1d[(d+βdηE2γ)(d+βdηE2γ)24dη(E2+γd)](21)

Thus if ϕ2(x1)<0 i.e. if E1 > ϕ(E2) then P2(x1, ϕ2(x1)) is locally asymptotically stable and if E1 < ϕ(E2) then P2(x1, ϕ2(x1)) is unstable in the positive quadrant of x1x2 − plane. For E1 = ϕ(E2) i.e., for ϕ2(x1)=0 the roots of (19) ur discussion on stability at P3(x2, ϕ2(x2)) we see that the characteristic equation of the community matrix at P3 is

λ2λϕ1(x2)ϕ2(x2)+ϕ1(x2)ϕ2(x2)ϕ3(x2)=0(22)

Since we have earlier discussed that ϕ3(x2)<0,   ϕ1(x2)>0,   ϕ2(x2)>0 for the existence of P3; therefore by Descartes’ rule of signs, the roots of (22) are both real and of opposite sign. Thus P3(x2, ϕ2(x2)) is an unstable saddle point whatever the sign of ϕ2(x2) may be.

We now state the following theorem:

Theorem 1

If the conditions (H1), (H2) and (H3) are satisfied then:

(i) P0(0, 0) is stable if E1 > 1, E2 > d − γ.

(ii) P1((1 − E1)η, 0) is stable if E2>β(1E1)η1+(1E1)η+dE1γ

and is unstable when E2<β(1E1)η1+(1E1)η+dE1γ.

(iii) P2(x1, ϕ2(x1)) is stable if E1 > ϕ(E2) and is unstable when E1 < ϕ(E2).

(iv) P3(x2, ϕ2(x2)) is always unstable.

We notice that

ddE1(Trace   J)E1=ϕ(E2)=ddE1[ϕ1(x1)ϕ2(x1)]E1=ϕ(E2)=x11+x10.

Hence by the Hopf bifurcation theorem (Hassard et al. [15]) the system (9) and (10) enters into a Hopf type small amplitude periodic oscillation at the parametric values E1 = ϕ(E2) near the positive interior equilibrium P2(x1, ϕ2(x1)). Hence we may state the following theorem:

Theorem 2

For E1 = ϕ(E2) a super critical Hopf bifurcation takes place for the equilibrium P2.

Proof

Let us consider the Jacobian matrix (18). By direct calculation the eigen values are given by:

λ1,2=ϕ1(x1)ϕ2(x1)±{ϕ1(x1)ϕ2(x1)}24ϕ1(x1)ϕ2(x1)ϕ3(x1)2

Or, λ1,2=T(E1,   E2)±T24D(E1,E2)2, where T = trace of J and D = det J.

Now for T = 0, D > 0 and in corresponding to the values of E1 = ϕ(E2) the matrix has two purely imaginary eigenvalues λ1,2=±iD. Moreover d/dE1(Re(λ1,2))E1=ϕ(E2). Thus, all conditions of Hopf theorem are satisfied and a stable limit cycle for E1 < ϕ(E2) may be found. All these conditions together prove the existence of a super critical Hopf bifurcation for the equilibrium P2.

We also have the following result on global stability in the region of local stability:

Theorem 3

If P2(x1*, ϕ2(x1*)) is locally asymptotically stable, then it is globally stable in the interior of R2+.

Proof

If possible, let there be a periodic orbit Γ = (x(t), y(t)), 0 ≤ t ≤ T with the enclosed region Ω and consider the variational matrix J about the periodic orbit,

J(x,y)=ϕ1(x)(ϕ2(x)y)+ϕ1(x)ϕ2(x)ϕ1(x)yϕ3(x)ϕ3(x).

We compute

Δ=0Tϕ1(x)[ϕ2(x)y)+ϕ1(x)ϕ2(x)+ϕ3(x)]dt

For biological equilibrium in the region of local stability it can easily be seen that

Δ=0Tϕ1(x)ϕ2(x)dt<0

Therefore, Δ < 0 and the periodic orbit Γ is orbitally asymptotically stable (Cheng et al. [16] and Hale [17]). Since every periodic orbit is orbitally stable then there is a unique limit cycle. From the Poincare–Bendixson Theorem, it is impossible to have unique stable limit cycle enclose a stable equilibrium. This is the desired contradiction. Hence there is no limit cycle and P2 is globally stable.

As the analytical results already indicate, the existence and stability of the interior equilibrium, as well as the stability of the trivial and boundary equilibrium will depend on the specific parameters that determine the dynamic evolution of the resource stock and the population. In the next section we present a more in-depth analysis of the complexity of the long run dynamics of population and resource stocks as they depend on the parameters of the system. The results are obtained through a detailed numerical and global bifurcation analysis.

4 Numerical simulations

We consider numerical simulations of different cases to understand the theoretical findings.

For η = 60, β = 4, d = 0.6, γ = 0.9 we have the bifurcation diagram shown in Fig. 1, and the point (E1, E2) = (0.9, 2.69) marked by * is in stable region and the point (E1, E2) = (0.75, 2.69) marked by • is in the region of unstability.

Fig. 1

Bifurcation diagram in the (E1, E2) plane.

Fig. 1 illustrates a detailed picture of the possible long run dynamics as they depend on the values of the harvesting efforts E1 and E2 for the benchmark case. The various dynamic behaviors are described in separate figures that show the corresponding stable and unstable equilibria in phase space together with selected time paths.

In brief, the bifurcation diagram as presented in Fig. 1 allows us to investigate the specific combination of efforts E1 and E2 that may cause the model to shift from sustainable long run equilibrium towards a limit cycle and finally the collapse of the system. Local bifurcation theory is the numerical tool that determines the boundaries between these regions in parameter space (Fig. 2).

Fig. 2

Bifurcation diagram for d = 0.9.

We see that as d increases from 0.6 to 0.9 the area of region of stability decreases as shown in Fig. 2. If we increase d again to 1.0 the stable region of stability shrinks more, as depicted in the Fig. 3.

Fig. 3

Bifurcation diagram for d = 1.0.

Figs. 1, 2 and 3 show bifurcation diagrams in the (E1, E2)-plane for d = 0.6, 0.9 and 1, respectively. We see that the stability region becomes smaller as d becomes larger. We may say that d destabilizes the coexisting equilibrium P2. Thus for the long run sustainability of a prey-predator system, this might require reducing alternative prey (Fig. 4).

Fig. 4

The phase diagram for (E1, E2) = (0.9, 2.69) with d = 0.6.

The phase diagram for d = 0.06 is shown in Fig. 4. When d increases to 0.9 or 1.0 then the point (E1, E2) = (0.9,2.69) marked by * lies outside the stable region and the interior equilibrium becomes unstable. The corresponding phase diagrams are shown in the Figs. 5 and 6.

Fig. 5

The phase diagram for (E1, E2) = (0.9, 2.69) with d = 0.9.

Fig. 6

The phase diagram for (E1, E2) = (0.9, 2.69) with d = 1.0.

Figs. 7 and 8 shows the phase diagram of prey and predator with a fixed value of E2 = 2.69 and different values of E1. Fig. 9 has E1 = 0.39, and E2 = 3.60

Fig. 7

The phase diagram for (E1, E2) = (0.75, 2.69) from region-II of the Fig. 1.

Fig. 8

The phase diagram for (E1, E2) = (2, 2.69) in region-III of Fig. 1.

Fig. 9

The phase diagram for (E1, E2) = (0.39, 3.60) in the Region IV of Fig. 1.

Figs. 10 and 11 shows the isoclines of prey and predator with a fixed value of E2 = 2.69 and different values of E1. It is observed that the equilibrium point for the prey population gradually decreases when the respective fishing effort used to harvest prey population is simultaneously increased. In Fig 10, d = 0.6 and in Fig. 11 d = 0 (Figs. 12 and 13).

Fig. 10

For d = 0.6.

Fig. 11

For d = 0.

Fig. 12

For d = 0.6.

Fig. 13

For d = 0.

Figs. 12 and 13 show the isoclines of prey and predator population with a fixed value of E1 = 0.6 and different values of E2. It is observed that the equilibrium point for the prey population gradually increases as the effort E2 increases. It is natural as an increase in E2 decreases the predator population and hence enhancing the survival rate of the prey.

5 Conclusion

Managing predator-prey systems involves complex challenges for resource managers. Many studies have demonstrated populations consequences of predators on prey population, but how managers should use this information are not easy to decide. Predator control can be effective at enhancing survival and recruitment in populations of prey, but the potential role of predators in structuring ecological communities has been recognized for sometime. For example, overexploitation of sea otter populations resulted in increase in sea urchins that subsequently destroyed kelp beds that provided habitats for fish and other marine organisms. Because of such complexity of food webs in ecological communities it is difficult to anticipate the full ramifications of eliminating or restoring predators. For example, reducing predator numbers to increase abundance of prey can have counter-productive results such as increasing disease and parasite infection in prey.

We believe that prey-predator management requires ecosystem management and this must include careful consideration of habitats as well as the particular predator prey populations. Management actions include adjusting season length and bag limits for hunter and trappers, but must also include other activities such as management of habitats that provide secure areas for prey. Ecosystem management acknowledges the value of predators in the environment.

Thus unregulated exploitation and extinction of many natural and biological resources is a major problem of present day. This work pays attention to the exploitation or harvesting of such resources. It describes some strategies on harvesting efforts of a prey-predator model incorporating an alternative prey for the predators. It gives a study of a prey-predator model with an alternative prey, which shows some methods to control the system and how the state can be driven into equilibrium. We have considered all possible cases for the variation of parameters for the equilibrium and stability analysis of the model. It has also been discussed, how the system bifurcates from equilibrium. We have seen that the interior equilibrium in Case 3 is stable for E1 > ϕ(E2) and is unstable for E1 < ϕ(E2) and another interior equilibrium is always unstable. Some numerical examples also show how the system becomes stable at an interior equilibrium and how it bifurcates from the equilibrium and also some other features of the model.

Managing prey-predator populations for hunter harvest becomes more complex when predators are competing with humans for the same prey. Adjustments to harvest regimes may be necessary, but certainly we can have sustainable harvest of populations under predation. Elk on the northern range of Yellowstone National Park are harvested by hunters when they move into Montana during winter. The sustainability of this harvest is ensured because the Montana Department of Fish, Wildlife and Parks has density dependent harvest guidelines so that the number of tags issued for the late-Gardiner elk hunt increases with the number of elk censused on the northern range. This helps to balance the hunter harvest with wolf predation ensuring that the elk population is not driven to low levels by excessive hunter harvest.

Model that has been developed permit harvest guidelines while accommodating predators and these can be used to achieve sustainable yields. However application of this model will require data to manage a prey-predator systems.


Bibliographie

[1] G. Hardin The tragedy of Commons, Science, Volume 162 (1968), pp. 1234-1247

[2] C.W. Clark, Mathematical bioeconomics: the optimal management of renewable resources, Wiley, New York, 1990.

[3] M. Masterton-Gibbons Game-theoretic resource modeling, Natural Resource Modeling, Volume 7 (1993), pp. 93-147

[4] J.M. Conrad. Resources Economics, Cambridge University Press, Cambridge, 1999.

[5] T.K. Kar; K.S. Chaudhuri On non-selective harvesting of a multispecies fisheries, Int. J. Math. Edu. Sci. Technol., Volume 33 (2002) no. 4, pp. 543-556

[6] H.S. Gordon The economic theory of a common property resource: the fishery, Journal of Political Economy, Volume 62 (1954), pp. 124-142

[7] M. Datta; L.J. Mirman Externalities, market power, and resources extraction, Journal of Environmental Economics and Management, Volume 37 (1999), pp. 233-255

[8] G.I. Bischi; M. Kopel The Role of competition, expectations and harvesting costs in commercial fishing (T. Puu; I. Sushko, eds.), Oligopoly dynamics: models and tools, Springer, Berl, 2002, pp. 85-109

[9] B. Dubey; C. Peeyush; P. Sinha A model for fishery resources with reserve area, Nonlinear Analysis, Volume 4 (2003), pp. 625-637

[10] G.I. Bischi; F. Lamantia; L. Sbragia Compettition and co-operation in natural resources exploitation. An evolutionary game approach (C. Carraro; V. Fragnelli, eds.), Game practice and the environment, Edward Elgar Publishing, Northampton, 2004, pp. 187-211

[11] P.A. Abrams; H. Matsuda Positive indirect effects between prey species that share predators, Ecology, Volume 77 (1996), pp. 610-616

[12] P.R. Vence Predation and resource partitioning in one predator-two prey model communications, American Naturalist, Volume 112 (1978), pp. 797-813

[13] P.D. Spencer; J.S. Collie A simple predator-prey model of exploited marine fish populations incorporating alternative prey, ICES Journal of Marine science, Volume 53 (1995), pp. 615-628

[14] C.S. Holling The functional response of predators to prey density and its role in mimicry and population regulation, Memoirs of Entomological Society of Canada, Volume 45 (1965), pp. 1-60

[15] B.D. Hassard; N.D. Kazarinoff; Y.H. Wan Theory of application of Hopf bifurcation, Cambridge University Press, Cambridge, 1981

[16] K.S. Cheng; S.B. Hsu; S.S. Lin Some results on global stability of a predator-prey system, J. Math. Biol., Volume 12 (1981), pp. 115-126

[17] J.K. Hale Ordinary differential equations, Wiley, New York, 1969


Research of T.K. Kar was supported by the Council of Scientific and Industrial Research (CSIR), India (Grant No. 25 (0160)/08/EMR-II dated: 17.01.08) and research of S.K. Chattopadhyay was supported by the University Grants Commissions (UGC), India (Grant No. PSW-179/09-10(ERO) dated: 08.10.2009).

Commentaires - Politique


Ces articles pourraient vous intéresser

Combined harvesting of a stage structured prey–predator model incorporating cannibalism in competitive environment

Kunal Chakraborty; Kunal Das; Tapan Kumar Kar

C. R. Biol (2013)


Does mutual interference always stabilize predator–prey dynamics? A comparison of models

Roger Arditi; Jean-Marc Callois; Yuri Tyutyunov; ...

C. R. Biol (2004)


Chaotic dynamics of a three species prey–predator competition model with bionomic harvesting due to delayed environmental noise as external driving force

Kalyan Das; M.N. Srinivas; M.A.S. Srinivas; ...

C. R. Biol (2012)