You are on page 1of 34

Accepted Manuscript

Dynamics of a Leslie–Gower predator–prey model with additive Allee effect

Yongli Cai, Caidi Zhao, Weiming Wang, Jinfeng Wang

PII: S0307-904X(14)00487-9
DOI: http://dx.doi.org/10.1016/j.apm.2014.09.038
Reference: APM 10167

To appear in: Appl. Math. Modelling

Please cite this article as: Y. Cai, C. Zhao, W. Wang, J. Wang, Dynamics of a Leslie–Gower predator–prey model
with additive Allee effect, Appl. Math. Modelling (2014), doi: http://dx.doi.org/10.1016/j.apm.2014.09.038

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Dynamics of a Leslie–Gower predator–prey model
with additive Allee effect
Yongli Caia,b , Caidi Zhaoa , Weiming Wang∗a , Jinfeng Wangc
a
College of Mathematics and Information Science, Wenzhou University, Wenzhou, 325035 P.R.China
b
School of Mathematics and Computational Science, Sun Yat-sen University, Guangzhou, 510275
P.R.China
c
Department of Mathematics, Harbin Normal University, Harbin, 150001 P.R. China

Abstract

In this paper, we investigate the complex dynamics of a Leslie–Gower predation model


with additive Allee effect on prey. Without Allee effect, the model has a unique globally
asymptotically stable equilibrium point under some conditions. With Allee effect on prey,
the model has none, one or two positive equilibria. Moreover, we prove the existence of
parameter subsets for which the model can have Hopf bifurcation. And we find that when
the model exhibits two positive equilibria there is a separatrix curve that separates the
behavior of trajectories of the system, implying that the model is highly sensitive to the
initial conditions. One of the most interesting findings is that Allee effect can increase
the risk of ecological extinction.
Keywords: Predator–prey model; Additive Allee effect; Leslie–Gower; Bistability.

1. Introduction

Predator–prey interaction is one of basic interspecies relations for ecological and so-
cial models, and it is also a basic block of more complicated food chain, food web and
biochemical network structure [1]. In recent years, one of the important predator–prey
models is Leslie–Gower model [2], which has been extensively studied. The Leslie–Gower
model is characterized by the predator growth equation being of logistic type, and the


Corresponding author. Tel.: +86 577 86689222.
Email address: weimingwang2003@163.com (Weiming Wang∗ )

Preprint submitted to Elsevier October 14, 2014


carrying capacity of predator’s environment is proportional to the number of prey [2].
In [3, 4, 5], the authors considered a modified version of Leslie–Gower model as follows
 ( )
 dN N c1 N P

 dt = rN 1 − K − N + a ,
( ) 1
(1)

 dP P
 = s1 P 1 − ,
dt N + a2
where N (t), P (t) represent population density of prey and predator at time t, respectively.
The parameters r, a1 , a2 , c1 , s1 are positive constants, r and s1 the intrinsic growth rate
of prey and predator, respectively, c1 the maximal predator per capita consumption rate,
i.e., the maximum number of prey that can be eaten by a predator in each time unit, a1
and a2 the extent to which the environment provides protection for prey and predator,
respectively.
( )

Let (N, P, t) = K Ñ , K P̃ , , for simplicity, and we still use variables N , P instead
r
of Ñ , P̃ , thus model (1) becomes:

 dN cN P

 dt = N (1 − N ) − N + k ,
( ) 1 (2)

 dP P
 = sP 1 − ,
dt N + k2
a1 a2 c1 s1
where k1 = , k2 = , c = , s = .
K K r r
In population dynamics, any mechanism that can lead to a positive relationship be-
tween a component of individual fitness and either the number or density of conspecific
can be termed an Allee effect [6, 7, 8, 9], starting with the pioneer work of Allee [10].
Allee effect can be caused by difficulties in finding mating partners at small density, ge-
netic inbreeding, demographic stochasticity or a reduction in cooperative interactions,
see [11, 12, 13].
Nowadays, it is widely accepted that Allee effect greatly increase the likelihood of
local and global extinction, and can lead to a rich variety of dynamical effects [13]. And
it is interesting and important to study the impact of Allee effect on the predator–prey
models [1, 14].

2
Allee effect has been modeled in different ways [15, 16, 17, 18]. In a classic predation
model, if N = N (t) indicates the population size, we assume that the growth function
G(N ) satisfies G(0) = 0, which means that there is a trivial equilibrium at the origin.
From an ecological point of view, Allee effect has been modeled into strong and weak
ones [9, 12, 15, 19, 20, 21, 22, 23, 24].
In particular, the growth rate function

m
G(N ) = 1 − N − , (3)
N +b

is called the additive Allee effect, which was first deduced in [11] and applied in [16, 17]. In
m
Eq. (3), the term can induce weak or strong Allee effects in the absence of predator.
N +b
And the positive parameters m and b are Allee effect constants, b is the population size
at which fitness is half its maximum value. The constant m will allow the severity of the
Allee effect to be modelled. In model (3), the carrying capacity of the prey equals 1, then
b < 1.
In a recent analytic approach by Wang and Kot [19], the Allee effect is “strong” if
the sign of the growth function N G(N ) in the limit of law density is negative, i.e.,

d(N G(N ))
< 0.
dt N =0

It is “weak” if the sign of the growth function G(N ) in the limit of law density is positive,
i.e.,
d(N G(N ))
> 0.
dt N =0

In the sense of [19], it then follows


(i) If 0 < m < b, the Allee effect in (3) is the weak one;
(ii) If m > b, the Allee effect in (3) is the strong one.
In this paper, we focus on a Leslie–Gower model with the additive Allee effect on

3
prey:
 ( )
 dN m cN P

 =N 1−N − − , f (N, P ),
 dt N +b N + k1
 ( ) (4)

 dP P
 = sP 1 − , g(N, P ),
dt N + k2

with the conditions R2+ = {(N, P ) ∈ R2 : N ≥ 0, P ≥ 0}.


There are a lot of excellent works on Leslie–Gower predator–prey models with Allee
effect, for example, see Refs. [16, 17, 18, 20] and the references therein. In Ref. [16, 17],
the authors showed that under certain conditions over the parameters, the Leslie–Gower
model with the additive Allee effect on prey allows the existence of two or three limit
cycles. González–Olivares and co–workers [18] studied the dynamical complexity in the
Leslie–Gower predator–prey model as consequences of Allee effect on prey, and proved the
existence of parameter subsets for which the system can have: a cusp point, homoclinic
bifurcation, Hopf bifurcation and the existence of two limit cycles, the innermost stable
and the outermost unstable. Mena–Lorca and co–workers [20] proved that by considering
Allee effect, the Leslie–Gower model modifies strongly the dynamics of original system
and shows the existence of different homoclinic and heteroclinic curves that divide the
behavior of the trajectories.
The main purpose of this paper is to study dynamical behavior of a Leslie–Gower
predator–prey model with the additive Allee effect. We will determine how Allee ef-
fect affects the existence and the stability of the equilibria of the model and focus on
bifurcation mechanism.
The rest of the paper is organized as follows. In the next section, we give some
properties to the solutions of the model. In Section 3, we investigate the existence of the
positive equilibria. In Section 4, we discuss the stability of the equilibria and bifurcation
of the model in detail, which is followed by concluding remarks in Section 5.

4
2. Dissipation and boundedness of the solutions

In this section, we are mainly concerned with some simple properties of the solution
to model (4). These properties will be used to prove the global stability.
dN
From the first equation of model (4) we have: < N (1 − N ), and a standard
dt
comparison theorem shows that

lim sup N (t) ≤ 1.


t→∞

As a result, for any ε > 0, there exists a T > 0, such that N (t) ≤ 1 + ε for t > T .
( )
dP P
Then from the second equation of model (4), we get ≤ sP 1 − .
dt 1 + ε + k2
The arbitrariness of ε then implies that

lim sup P (t) ≤ 1 + k2 .


t→∞

Thus, we obtain the following conclusion.

Lemma 2.1. Model (4) is dissipative.

Theorem 2.2. All the solutions of model (4) which are initiated in R2+ are uniformly
bounded.

Proof. Define the function


W (t) = N (t) + P (t),

the time derivative of W along the solution of (4), we get

dW dN dP
= +
dt dt( dt ) ( )
m cN P P
=N 1−N − − + sP 1−
N+
(b N + k)1 N + k2
P
≤ N (1 − N ) + sP 1 −
( )N + k2
1 P
≤ + sP 1 − .
4 N + k2

5
Then, ( )
dW 1 P
+ W ≤ + N + P + sP 1 −
dt 4 ( N + k2
)
5 sP
≤ +P 1+s−
4 1 + k2 (5)
5 (1 + k2 )(1 + s)2
≤ +
4 4s
, B.
Using differential inequality, for all t ≥ T ≥ 0, we have,

0 ≤ W (t) ≤ B − (B − W (T ))e−(t−T ) .

Hence, we have
lim sup(N (t) + P (t)) ≤ B.
t→∞

This completes the proof.

Let Γ be the set defined by

Γ = {(N, P ) ∈ R2+ : 0 ≤ N ≤ 1, 0 ≤ P ≤ 1 + k2 , 0 ≤ N + P ≤ B}. (6)

Then Γ is positive invariant.


Remark: It needs to note that in the case of m = 0, i.e., in model (4) without Allee
effect, or model (2), the dissipation and boundedness of the solution are similar to the
results in Theorem 2.2. That’s to say, the solutions of the model without Allee effect are
always dissipative and uniformly bounded.

3. Existence of the positive equilibria

In this section, we mainly focus on the existence of positive equilibria of model (4).
Any positive equilibrium (N, P ) of model (4) satisfies:

 m cP
 1−N − − = 0,
N + b N + k1

 1−
P
= 0,
N + k2

6
which yields P = N + k2 and

ϕ1 (N ) , N 3 +(b+c+k1 −1)N 2 +(m+bk1 +bc+ck2 −b−k1 )N +bck2 +k1 (m−b) = 0. (7)

Let N be a real positive root of the cubic

x3 + 3η1 x2 + 3η2 x + η3 = 0, (8)

where

3η1 = b + c + k1 − 1, 3η2 = bc + bk1 + ck2 + m − b − k1 , η3 = bck2 + k1 (m − b). (9)

By the transformation z = x + η1 , Eq. (8) is reduced to

h(z) = z 3 + 3pz + q = 0, (10)

where
p = η2 − η12 ,
(11)
q = η3 − 3η1 η2 + 2η13 .
In the following, we discuss the existence of positive roots of Eq. (10).

Lemma 3.1. (a) If q < 0, Eq. (10) has a single positive root;
(b) Suppose q > 0 and p < 0, then:
(b1) if q 2 + 4p3 = 0, Eq. (10) has a positive root of multiplicity two;
(b2) if q 2 + 4p3 < 0, Eq. (10) has two positive roots;
(c) If q = 0 and p < 0, Eq. (10) has a unique positive root.

The proof is postponed to Appendix.


Moreover, simple algebraic
√( computations show that, when Eq. (10) has two pos-
√ )2
3
−4 q + 4 4 p3 + q 2 − 4 p
z1
itive roots, one is z1 = √ √ and the other z2 = − +
2
2 3 −4 q + 4 4 p3 + q 2

7

z13 + 4q
√ . It is noted that if the cubic Eq. (10) has one positive root, it must be
2 √ z1
( √ )2
3
−4 q + 4 4 p3 + q 2 − 4 p
z1 = √ √ .
2 3 −4 q + 4 4 p3 + q 2

Considering (11), from p = 0 and q = 0, one can determine m = m∗1 and m = m∗2 ,
where
1
m∗1 , (b + c + k1 − 1)2 + b − bc − (b − 1)k1 − ck2 ,
3
2b − c − k1 + 1 ( 2 )
m∗2 , b + b − bc − bk1 − 2c2 + 4c − 4ck1 − 2k1 2 − 5k1 + 9ck2 − 2 .
9(b + c − 2k1 − 1)
Hence, we have the following Lemmas regarding the existence of the positive equilibria
of model (4) with weak (0 < m < b) and strong (m > b) Allee effect, respectively.

Lemma 3.2. (Existence of the positive equilibria of model (4) with weak Allee effect)
(a) If either of the following inequalities holds:

2k1 + 1 − c − b > 0, 0 < m < min{b, m∗2 };

2k1 + 1 − c − b < 0, 0 < m∗2 < m < b,

model (4) has a unique positive equilibrium Ew∗ = (Nw∗ , Pw∗ ) = (z1 − η1 , z1 − η1 + k2 ).
(b) If either of the following inequalities holds:

2k1 + 1 − c − b > 0, 0 < m∗2 < m < min{b, m∗1 };

2k1 + 1 − c − b < 0, 0 < m < min{b, m∗1 , m∗2 }.

(b1) Model (4) has two positive equilibrium, namely, Ew3 = (Nw3 , Pw3 ) = (z1 −
η1 , z1 − η1 + k2 ) and Ew4 = (Nw4 , Pw4 ) = (z2 − η1 , z2 − η1 + k2 ), with q 2 + 4p3 < 0;
√ √
(b2) Model (4) has a unique positive equilibrium Ew = (Nw , Pw ) = ( −p, −p +
k2 ), with q 2 + 4p3 = 0;
(c) If 0 < m < min{b, m∗1 } model (4) has a unique positive equilibrium, namely,
√ √
E˜w = (N˜w , P˜w ) = ( −3p, −3p + k2 ), with m∗2 = 0.

8
Lemma 3.3. (Existence of the positive equilibria of model (4) with strong Allee effect)
(a) If either of the following inequalities holds:

2k1 + 1 − c − b > 0, 0 < b < m < m∗2 ;

2k1 + 1 − c − b < 0, 0 < max{b, m∗2 } < m,

model (4) has a unique positive equilibrium Es∗ = (Ns∗ , Ps∗ ) = (z1 − η1 , z1 − η1 + k2 ).
(b) If either of the following inequalities holds:

2k1 + 1 − c − b > 0, 0 < max{b, m∗2 } < m < m∗1 ;

2k1 + 1 − c − b < 0, 0 < b < m < min{m∗1 , m∗2 }.

(b1) Model (4) has two positive equilibrium, namely, Es4 = (Ns4 , Ps4 ) = (z1 −
η1 , z1 − η1 + k2 ) and Es5 = (Ns5 , Ps5 ) = (z2 − η1 , z2 − η1 + k2 ), with q 2 + 4p3 < 0;
√ √
(b2) Model (4) has a unique positive equilibrium Es = (Ns , Ps ) = ( −p, −p +
k2 ), with q 2 + 4p3 = 0.
(c) If b < m < m∗1 model (4) has a unique positive equilibrium, namely, Ẽs =
√ √
(Ñs , P̃s ) = ( −3p, −3p + k2 ), with m∗2 = 0.

Remark: It’s worthy to note that the conditions of the existence of the positive
equilibria in model (4) are very complex, due to Allee effect. In fact, in the case without
Allee effect (i.e., m = 0 in model (4) ), or model (2), there is a unique positive equilibrium
E ∗ = (N ∗ , P ∗ ) with ck2 < k1 , where

1( √ )
N∗ = 1 − c − k1 + (1 − c − k1 )2 − 4(ck2 − k1 ) , P ∗ = N ∗ + k2 . (12)
2

4. Stability of the equilibria

In this section, we consider the stability of the equilibria of model (4). For the sake of
learning Allee effect on the stability of the model, we first present the stability analysis
in the case without Allee effect.

9
4.1. The case without Allee effect

For model (2), one can easily know there are three boundary equilibria E0 = (0, 0),
E1 = (0, k2 ), E2 = (1, 0) and a unique positive equilibrium E ∗ = (N ∗ , P ∗ ) with ck2 < k1 .
And the Jacabian matrix around E0 , E1 and E2 is given as follows respectively:
 
1 0
 
JE0 =  ,
0 s
 
ck2
 1 − k1 0 
JE1 = 

,

s −s
 
c
 −1 − 1 + k1 
JE2 = 

,

0 s
Then, one can easily obtain the following results.

Theorem 4.1. For model (2),


(a) E0 is an unstable node point;
(b) E1 is a saddle point;
(c) if ck2 < k1 , E2 is a saddle point, while ck2 > k1 , a stable node (nodal sink).

For the stability of unique positive equilibrium E ∗ , following Aziz-Alaoui and Okiye [3],
one can obtain the following theorem.

Theorem 4.2. The positive equilibrium E ∗ = (N ∗ , P ∗ ) is globally asymptotically stable


if
k1 (1 + k1 )
B< , k1 < 2k2 , 4(N ∗ + k1 ) < c
2c
hold, where B has the same definition as in (5).

For more details, see Ref. [3].

10
4.2. The case with weak Allee effect

In this section, we consider the stability of the equilibria of model (4) with weak Allee
effect, i.e, 0 < m < b.
In this situation,
( model (4) has three boundary
) equilibria Ew0 = (0, 0), Ew1 = (0, k2 )

1 − b + (1 − b) − 4(m − b)
2
and Ew2 = ,0 .
2
The Jacobian matrix J of model (4) is given as J = (ξij )2×2 , where
bm −ck1 P
ξ11 = 1 − 2N − − ,
(N + b)2 (N + k1 )2
cN
ξ12 = − ,
N + k1
sP 2
ξ21 = ,
(N + k2 )2
2sP
ξ22 =s− .
N + k2
The stability of model (4) around Ew0 , Ew1 and Ew2 is found as follows:

Theorem 4.3. (a) Ew0 is an unstable node;


( ) { ( )}
ck2 ck2
(b) if ck2 < k1 and m < b 1 − , Ew1 is a saddle point; while max 0, b 1 − <
k1 k1
m < b, it is a stable node (nodal sink);
(c) Ew2 is a saddle point.

m
Proof. (a) The eigenvalues of the Jacobian matrix at Ew0 = (0, 0) are 1 − > 0 and s.
b
Therefore, model (4) is always unstable around Ew0 which is, in fact, an unstable node.
m ck2
(b) The eigenvalues of the Jacobian matrix at Ew1 = (0, k2 ) are 1 − −
( ) b k1
ck2
and −s. Hence, if ck2 < k1 and m < b 1 − hold, Ew1 is a saddle point. If
{ ( )} k1
ck2
max 0, b 1 − < m < b holds, Ew1 is a stable node point (nodal sink).
k1 ( )

1−b+ (1−b)2 −4(m−b)
(c) The eigenvalues of the Jacobian matrix at Ew2 = 2
, 0 are s
and √ (√ )
2 (1 + b)2 − 4m (1 + b)2 − 4m + 1 + b − 2m
− √ .
(b + 1 + (1 + b)2 − 4m)2

11
It is noted that b < 1, we have 1 + b − 2m > 1 − b > 0. Hence, model (4) is unstable
around Ew2 which is a saddle point.

In the following, we shall discuss the stability of the positive equilibria Ew∗ = (Nw∗ , Pw∗ ),
Ew3 = (Nw3 , Pw3 ) and Ew4 = (Nw4 , Pw4 ), respectively.
Let (N, P ) = (N, N + k2 ) be any positive solution of model (4), then the Jacobian
matrix at (N, P ) is given as:
 
cN
 s
[N ]

N + k1 
J(N,N +k2 ) =

,
 (13)
s −s
( )
c (N + k2 ) m
where s [N ]
=N + −1 .
(N + k1 )2 (N + b)2
The determinant of J(N,N +k2 ) is
( )
sN (N + k1 ) ((N + b) − m) + c(N + b) (k1 − k2 )
2 2 2

det(J(N,N +k2 ) ) = ,
(N + k1 )2 (N + b)2

and the sign of det(J(N,N +k2 ) ) depends on the factor ϕ(N ),

ϕ(N ) = (N + k1 )2 ((N + b)2 − m) + c(N + b)2 (k1 − k2 )


= (N + b + k1 + 1 − c)ϕ1 (N ) + µ1 N 2 + µ2 N + µ2
= µ1 N 2 + µ2 N + µ3 ,

where ϕ1 (N ) is defined in Eq.(7), and

µ1 = 1 + b + k1 + bk1 + c2 + ck1 − 2 m − bc − 2 ck2 − 2 c,


µ2 = −ck2 + b − m + 2 bk1 − 2 bc + b2 − ck1 + c2 k2 + c2 b
−bm + cm − b2 c − 4 mk1 + k1 2 − 4 bck2 − ck2 k1 + b2 k1 + k1 + bk1 2 + 2 bk1 c, (14)

µ3 = b2 k1 2 − bck1 − bmk1 + mck1 + b2 ck1 + bc2 k2 − bck1 k2


−bck2 + bk1 − mk1 + b2 k1 + bk1 2 − 2 mk1 2 − 2 b2 ck2 .

Then, we can obtain the following results of the stability of the positive equilibria.

12
( )
∗] c (Nw∗ + k2 ) m
Theorem 4.4. Define s [Nw
= Nw∗ 2 +
∗2
2 − 1 and assume that µ1 Nw +

(Nw + k1 ) ∗
(Nw + b)
µ2 Nw∗ + µ3 > 0, then:
c (Nw∗ + k2 ) m c (Nw∗ + k2 ) m ∗
(a) If 2 + 2 < 1 or 2 + 2 > 1 and s[Nw ] < s,
(Nw∗ + k2 ) (Nw∗ + b) (Nw∗ + k2 ) (Nw∗ + b)
∗ ∗ ∗
Ew = (Nw , Pw ) is locally asymptotically stable.
(b) Ew∗ is globally asymptotically stable, if the following conditions hold:
m (Nw∗ + k1 ) c c
(b1) Nw∗ + k1 > + (1 + k 2 ) + ;
b (Nw∗ + b) k1 4k22
(b2) k1 < 2k2 .
c (Nw∗ + k2 ) m ∗]
(c) If 2 + 2 > 1 and s
[Nw
> s, Ew∗ is unstable.

(Nw + k2 ) ∗
(Nw + b)
c (Nw∗ + k2 ) m
(d) If 2 + > 1, the model undergoes a Hopf-bifurcation around

(Nw + k2 ) (Nw + b)2


Ew∗ at s = s[Nw ] .

Proof. (a) The trace of the Jacobian matrix J(N,N +k2 ) around Ew∗ is tr(J(Ew∗ )) = s[Nw ] −s.

Therefore, if s[Nw ] < s, then tr(J(Ew∗ )) < 0. Thus, Ew∗ is locally asymptotically stable.
(b) We adopt the Lyapunov function
∫ N ∫
∗ ξ − Nw∗ cPw∗ P η − Pw∗
V (N, P ) = (Nw + k1 ) dξ + dη, (15)

Nw ξ s Pw∗ η

which is defined and continuous on R+ 2 . It can be easily verified that V (N, P ) is zero
at the equilibrium Ew∗ and is positive for all other positive values of N and P , so Ew∗ is
the global minimum of V (N, P ).
From Lemma 2.1, we know that the solutions of the model are bounded and ultimately
enter the set Γ, and we restrict the study to this set. The time derivative of V along the
solutions of (4) is:
dV (Nw∗ + k1 )(N − Nw∗ ) dN cP ∗ (P − Pw∗ ) dP
= + w
dt N dt sP dt
( ) ( )
∗ ∗ m cP ∗ ∗ P
= (Nw + k1 )(N − Nw ) 1 − N − − + cPw (P − Pw ) 1 − .
N + b N + k1 N + k2
For model (4), Ew∗ satisfies

m cPw∗
1 − Nw∗ − − = 0, Pw∗ = Nw∗ + k2 ,
Nw∗ + b Nw∗ + k1

13
then,
( )
dV m(Nw∗ + k1 ) cP
= −(N − Nw∗ )2

Nw + k1 − − − c(P − Pw∗ )2
dt (Nw∗ + b)(N + b) N + k1
( ) (16)
P
−c 1 − (N − Nw∗ )(P − P ∗ ),
N + k2
which can be rewritten as
  
dV ( ) φ1 (N, P ) φ2 (N, P ) N − Nw∗
= − N − Nw∗ , P − Pw∗    ,
dt φ2 (N, P ) c P− Pw∗

where
m c(Nw∗ + k2 )
φ1 (N, P ) = 1 − − ,
(Nw∗ + b)(N + b) (Nw∗ + k1 )(N + k1 )
( )
c P
φ2 (N, P ) = 1− ,
2 N + k2
and set  
φ1 (N, P ) φ2 (N, P )
M= .
φ2 (N, P ) c
dV
From (16), it is obvious that < 0 if matrix M is positive definite. And M is
dt
positive definite iff all upper–left sub-matrices are positive (Sylvester’s criterion). That
is, since c > 0, iff
(i) φ1 (N, P ) > 0 and (ii) Φ(N, P ) = cφ1 (N, P ) − φ22 (N, P ) > 0.
Since Γ is an attracting positive invariant set, and in Γ, all solutions satisfy 0 ≤ N ≤
1, 0 ≤ P ≤ 1 + k2 and 0 ≤ N + P ≤ B, then
m(Nw∗ + k1 ) cP
φ1 (N, P ) = Nw∗ + k1 − ∗
− ,
(Nw + b)(N + b) N + k1
m(Nw∗ + k1 ) c(1 + k2 )
≥ Nw∗ + k1 − − .
b(Nw∗ + b) k1
Therefore, if (b1) holds, ∀(N, P ) ∈ Γ, φ1 (N, P ) < 0, for all t ≥ 0. On the other hand,
( ) ( )2
∗ m(Nw∗ + k1 ) cP c2 P
Φ(N, P ) = c Nw + k1 − − − 1− .
(Nw∗ + b)(N + b) N + k1 4 N + k2
Since
∂Φ(N, P ) c2 (N + k2 − P ) c2
= − ,
∂P 2 (N + k2 )2 N + k1

14
then
∂ 2 Φ(N, P ) c2
= − < 0.
∂P 2 2 (N + k2 )2
∂Φ(N, P )
Hence, is strictly decreasing in R2+ , with respect to P , and
∂P
∂Φ(N, P ) c2 (k1 − 2k2 − N )
= .
∂P P =0 2 (N + k1 ) (N + k2 )

∂Φ(N, P )
Consequently, if (b2) holds, < 0 in R2+ and so Φ(N, P ) is strictly decreasing
∂P
in R2+ . This yields

Φ(N, P ) > Φ(N, 1 + k2 )


( ) ( )2
∗ m (Nw∗ + k1 ) c(1 + k2 ) 1 1 + k2
= c Nw + k1 − − − c 1−
(Nw∗ + b) (N + b) N + k1 4 N + k2
( ∗
)
m (Nw + k1 ) c c
> c Nw∗ + k1 − − (1 + k 2 ) − .
b (Nw∗ + b) k1 4k22

Due to (b1), ∀(N, P ) ∈ Γ2 , Φ(N, P ) > 0.


dV
It follows that if the conditions (b1)–(b2) are satisfied, < 0 along all trajectories
dt
in the first quadrant except Ew∗ , hence, the equilibrium Ew∗ is globally asymptotically
stable.

(c) If s[Nw ] > s, Ew∗ is unstable, by Poincaré–Bendixson theorem, at least one limit
cycle surrounding the equilibrium Ew∗ exists.
(d) It is easy to see that:
( )
(d1) tr JEw∗ [N ∗ ] = 0;
( )
s=s w

(d2) det JEw∗ ) [N ∗ ] > s[Nw ] Nw∗ (µ1 N 2 + µ2 N + µ3 ) > 0;
s=s
( )
w

(d3) When Ew∗ exists, the characteristic equation is λ + det 2


J(Ew∗ ) ∗
= 0,
s=s[Nw ]
whose roots are purely imaginary;
d[ ( )]
(d4) tr JEw∗ ) ∗
= −1 ̸= 0.
ds s=s[Nw ]
From the Poincaré–Andronov–Hopf Bifurcation Theorem [25], we know that model (4)

undergoes a Hopf–bifurcation at Ew∗ as s passes through value s[Nw ] .

15
From Theorem 4.2 and Theorem 4.4, one can know, there exists a unique positive
equilibrium in both model (2) and model (4); however, the dynamics of Ew∗ in model (4)
with Allee effect is more complex than that of E ∗ in model (2), which is induced by
Allee effect. Figs. 1– 3 demonstrates the complex dynamics of model (4) with weak Allee
effect around the unique positive equilibrium Ew∗ = (Nw∗ , Pw∗ ). In Fig. 1, Ew∗ is locally
asymptotically stable; Fig. 2 shows a stable limit cycle around Ew∗ , which is an unstable
focus. Fig. 3 illustrates a Hopf–bifurcation situation of the model around Ew∗ .

Fig.1

Fig.2

Fig.3

Next, we will focus on the stability of the two positive equilibria Ew3 = (Nw3 , Pw3 )
and Ew4 = (Nw4 , Pw4 ).
( )
c (Ns3 + k2 ) m
Theorem 4.5. (a) Define s [Nw3 ]
= Nw3 + − 1 and assume
(Nw3 + k1 )2 (Nw3 + b)2
2
that µ1 Nw3 + µ2 Nw3 + µ3 > 0 and Ew3 = (Nw3 , Pw3 + k2 ) is locally asymptotically stable
c (Ns4 + k2 ) m c (Ns4 + k2 ) m
if 2 + 2 < 1 or 2 + > 1 and s[Nw3 ] < s.
(Ns4 + k1 ) (Ns4 + b) (Ns4 + k1 ) (Ns4 + b)2
(b) if θ < −(µ1 Nw3
2
+ µ2 Nw3 + µ3 ) < 0, Ew4 = (Nw4 , Pw4 ) is a saddle point, where

θ = (6 µ1 η1 − 3 µ2 ) Nw3 + 3 µ3 + 9 µ1 η1 2 − 6 µ2 η1

2(µ2 − µ1 Nw3 − 3µ1 η1 ) (Nw3 + η1 )3 + 4q
+ √
(Nw3 + η1 )
3 2
µ1 Nw3 + 3 µ1 η1 Nw3 + 3 µ1 η1 2 Nw3 + µ1 η1 3 + 4 µ1 q
+ ,
(Nw3 + η1 )

η1 , η2 , η3 , q and µ1 , µ2 , µ3 have the same definitions as in (9), (11) and (14), respectively.

Proof. (a) It is clear the proof is similar to that of part (a) in Theorem 4.4, and we omit
the proof.

16
(b) Since Nw3 = z1 − η1 , we have:

Nw3 + η1 (Nw3 + η1 )3 + 4q
Nw4 = z2 − η 1 = − + √ − η1 .
2 2 Nw3 + η1

The determinant of the Jacobian matrix around Ew4 is

sNw4 ϕ(Nw4 )
det(JEw4 ) = ,
(Nw4 + k1 )2 (Nw4 + b)2

and the sign of det(JEw4 ) depends on the factor ϕ(Nw4 ),

ϕ(Nw4 ) = µ1 Nw4 2 + µ2 Nw4 + µ3


( √ )2
Nw3 + η1 (Nw3 + η1 )3 + 4q
= µ1 − + √ − η1
2 2 Nw3 + η1
( √ )
Nw3 + η1 (Nw3 + η1 )3 + 4q
+µ2 − + √ − η1 + µ3
2 2 Nw3 + η1

(µ2 − µ1 Nw3 − 3µ1 η1 ) (Nw3 + η1 )3 + 4q
= √
2 (Nw3 + η1 )
3 2
µ1 Nw3 + 3 µ1 η1 Nw3 + 3 µ1 η1 2 ηw3 + µ1 η1 3 + 4 µ1 q
+
4(Nw3 + η1 )
1
+ (µ1 Nw32
+ (−2 µ2 + 6 µ1 η1 ) Ne1 − 6 µ2 η1 + 4 µ3 + 9 µ1 η1 2 ).
4
It follows that ϕ(Nw4 ) < 0 is valid if θ < −(µ1 Nw3
2
+ µ2 Nw3 + µ3 ) < 0 holds, i.e.,
det(JEw4 ) < 0, which completes the proof.

Considering Theorem 4.3 and Theorem 4.5 together, we can find that under certain
conditions, model (4) with weak Allee effect can be a bistable system. In Fig.4, we show
the bistable dynamics that there is a separatrix curve generated by the stable manifold of
the positive equilibrium Ew4 . The orbits initiating the right of the separatrix curve tend
to Ew3 that represents the coexistence of the prey N and predator P , while the orbits
initiating the left of the separatrix curve tend to Ew1 that represents the extinction of
the prey N . This shows that the model is highly sensitive to the initial conditions.

Fig.4

17
4.3. The case with strong Allee effect

In this subsection, we consider the stability of the equilibria of model (4) with strong
Allee effect, i.e, m > b.
In this case, model (4) always has two boundary equilibria Es0 = (0, 0) and Es1 =
(1 + b)2
(0, k2 ), when b < 1 and b < m < , there are another two boundary equilibria
( √ 4) ( √ )
1 − b + (1 − b)2 − 4(m − b) 1 − b − (1 − b)2 − 4(m − b)
Es2 = , 0 and Es3 = ,0 .
2 2
Obviously, with strong Allee effect, the number of the boundary equilibria is more than
the cases without Allee effect or with weak Allee effect.
The behavior of model (4) around these four boundary equilibria is found as follows:

Theorem 4.6. (a) Es0 is a saddle point;


(1 + b)2
(b) Es1 is always locally asymptotically stable. Moreover, if m ≥ and k1 ≤ k2
4
hold, E1 is globally asymptotically stable in R2+ ;
(c) Es2 is a saddle point;
(d) Es3 is an unstable node point.

m
Proof. (a) The eigenvalues of the Jacobian matrix at Es0 = (0, 0) are 1 − < 0 and
b
s. Therefore, model (4) is always unstable around Es0 which is, in fact, a saddle point
whose stable manifold is the P –axis.
m ck2
(b) The eigenvalues of the Jacobian matrix at Es1 = (0, k2 ) are 1 − − < 0 and
b k1
−s. Therefore, Es1 is locally asymptotically stable.
Furthermore, consider the Lyapunov function

c P
η − k2
V1 = N + dη. (17)
s k2 η

18
The derivation of (17) along the solution of model (4) is
( ) ( )
dV1 m cP P
=N 1−N − − + c(P − k2 ) 1 −
dt N + b N + k1 N + k2
N (−N 2 + (1 − b)N + b − m) cN P (k2 − k1 ) c(P − k2 )2 ck2 N
= − − −
N +b (N + k1 )(N + k2 ) N + k2 N + k2
N ((1 + b)2 − 4m) cN P (k2 − k1 ) c(P − k2 )2 ck2 N
≤ − − −
4(N + b) (N + k1 )(N + k2 ) N + k2 N + k2

≤ 0,
(1 + b)2 dV1
if m ≥ and k1 ≤ k2 , and = 0 at Es1 = (0, k2 ). This indicates that Es1 is
4 dt
globally asymptotically stable.
( √ )
1−b+ (1−b)2 −4(m−b)
(c) The eigenvalues of the Jacobian matrix at Es2 = 2
, 0 are s
and √ (√ )
2 (1 + b)2 − 4m (1 + b)2 − 4m + 1 + b − 2m
− √ .
(b + 1 + (1 + b)2 − 4m)2
(1 + b)2
It is noted that 1 + b − 2m > 1 + b − > 0. Hence, model (4) is always unstable
2
around Es2 which is a saddle point whose stable manifold is the P –axis.
( √ )
1−b− (1−b)2 −4(m−b)
(d) The eigenvalues of the Jacobian matrix at Es3 = 2
, 0 are s
and √ (√ )
2 (1 + b) − 4m
2 (1 + b) − 4m − (1 + b − 2m)
2
− √ .
(b + 1 + (1 + b)2 − 4m)2

It is noted that (1 + b)2 − 4m − (1 + b − 2m) < 0. Hence, model (4) is always unstable
around Es3 which is an unstable node point.

Comparing Theorem 4.1, Theorem 4.3 with Theorem 4.6, one can see that, the sta-
bility of the boundary equilibrium (0, k2 ) is different. In the case without Allee effect,
E1 = (0, k2 ) is a saddle point, which is unstable. In the case with weak Allee effect,
Ew1 = (0, k2 ) is a saddle point or stable node, which is locally stable. And in the case
with strong Allee effect, Es1 = (0, k2 ) is always locally stable.
For the stability of the positive equilibrium Es∗ = (Ns∗ , Ps∗ ), in the similar way to
Theorem 4.4, we can obtain the following results.

19
( )
[Ns∗ ] ∗ c (Ns∗ + k2 ) m ∗2
Theorem 4.7. Define s =N
∗ 2 + ∗ 2 − 1 and suppose that µ1 Ns +
(Ns + k1 ) (Ns + b)
µ2 Ns∗ + µ3 > 0, and we have
c (Ns∗ + k2 ) m c (Ns∗ + k2 ) m ∗
(a) If 2 + 2 < 1 or 2 + 2 > 1 and s[Ns ] < s, Es∗
(Ns∗ + k1 ) (Ns∗ + b) (Ns∗ + k1 ) (Ns∗ + b)
is locally asymptotically stable. Moreover,

(a1) if (s[Ns ] − s)2 < 4 det(JEs∗ ), Es∗ is a stable focus;

(a2) if (s[Ns ] − s)2 > 4 det(JEs∗ ), Es∗ is a stable node point.
c (Ns∗ + k2 ) m ∗
(b) If 2 + 2 > 1 and s[Ns ] > s, Es∗ is unstable.
(Ns∗ + k2 ) (Ns∗ + b)

c (Ns + k2 ) m
(c) If 2 + > 1, the model enters into a Hopf-bifurcation around

(Ns + k2 ) (Ns + b)2


Es∗ at s = s[Ns ] .

Furthermore, similar to Theorem 4.5, we can obtain the stability of two positive
equilibria Es4 = (Ns4 , Ps4 ) and Es5 = (Ns5 , Ps5 ). We only show the following theorem
and omit the proof.
( )
c (Ns4 + k2 ) m
Theorem 4.8. Define s [Ns4 ]
= Ns4 + − 1 and assume that
(Ns4 + k1 )2 (Ns4 + b)2
2
µ1 Ns4 + µ2 Ns4 + µ3 > 0, then:
c (Ns4 + k2 ) m c (Ns4 + k2 )
(a) Es4 is locally asymptotically stable if 2+ 2 < 1 or +
(Ns4 + k1 ) (Ns4 + b) (Ns4 + k1 )2
m [Ns4 ]
2 > 1 and s < s;
(Ns4 + b)
c (Ns4 + k2 ) m
(b) If 2 + > 1 and s[Ns4 ] > s, Es4 is unstable.
(Ns4 + k2 ) (Ns4 + b)2
(c) Es5 is a saddle point, if θ < −(µ1 Ns5 2
+ µ2 Ns5 + µ3 ) < 0, where

θ = (−3 µ2 + 6 µ1 η1 ) Ne1 + 3 µ3 + 9 µ1 η1 2 − 6 µ2 η1

2(µ2 − µ1 Ns4 − 3µ1 η1 ) (Ns4 + A1 )3 + 4q
+ √
(Ns4 + η1 )
3 2
µ1 Ns4 + 3 µ1 η1 Ns4 + 3 µ1 η1 2 Ns4 + µ1 η1 3 + 4 µ1 q
+ ,
(Ns4 + η1 )
η1 , η2 , η3 , q and µ1 , µ2 , µ3 are defined as in (9), (11) and (14), respectively.

Comparing Theorem 4.7 with Theorem 4.4, Theorem 4.8 with Theorem 4.5, it is clear
that Theorem 4.7 is similar to Theorem 4.4, and Theorem 4.8 similar to Theorem 4.5,

20
which means that the stabilities of positive equilibria of model (4) with strong Allee effect
is similar to that in the case with weak Allee effect.
From Theorem 4.8, one can know that if model (4) with strong Allee effect exhibits two
positive equilibria, then Es4 = (Ns4 , Ps4 ) can be a locally stable equilibrium. Considering
Theorem 4.6 again, we know that Es1 = (0, k2) is always stable, too. That’s to say,
if Es4 = (Ns4 , Ps4 ) is locally stable, then model (4) with strong Allee effect can be a
bistable system, a similar result to the case with weak Allee effect. However, there is
a slight difference between the two cases. In fact, in the case with weak Allee effect,
the stability of the boundary equilibrium Ew1 = (0, k2 ) is located in two cases: one
is saddle, and the other stable node (local stable). And in the case with strong Allee
effect, the stability of the boundary equilibrium Es1 = (0, k2 ) is always stable. In other
words, if the model exhibits two positive equilibria, and one of them is locally stable,
then model (4) with strong Allee effect must be a bistable system. But in the case with
weak Allee effect, if the model exhibits two positive equilibria, and one of them is locally
stable, model (4) with weak Allee effect is not necessarily a bistable system, because the
boundary equilibrium Es1 = (0, k2 ) can be a stable node under certain conditions.
In Fig.5, we show the bistable dynamics of model (4) with strong Allee effect. One
can see that there is a separatrix curve generated by the stable manifold of the positive
equilibrium Es5 . The orbits initiating the right of the separatrix curve tend to Es4 that
represents the coexistence of the prey N and predator P , while the orbits initiating the
left of the separatrix curve tend to Es1 that represents the extinction of the prey N .

Fig.5

5. Conclusions and Remarks

In this paper, we are concerned with the complex dynamics in a Lesile–Gower predator–
prey model with the additive Allee effect on prey. The value of this study lies in three

21
aspects. First, it presents the conditions for the dissipation and boundedness of the so-
lutions to the model. Second, it illustrates the existence of equilibria of the model with
Allee effect, which indicates that Allee effect can drive new steady state of the model.
Third, it discusses the stability of the equilibria in detail, which shows that there is a
separatrix curve that separates the behavior of trajectories of the system, implying that
the model is highly sensitive to the initial conditions.
It is well-known that model (2), the case without Allee effect, has a unique globally
asymptotically stable equilibrium point. However, it is shown that Allee effect signifi-
cantly modifies the original system dynamics, as the studied model (4) involves much
non-topological equivalent behavior.
In the case with (
Allee effect, the equilibrium (0, 0),
) which represents the extinction

1 − b + (1 − b) − 4(m − b)
2
of both species, and , 0 , which represents the extinction
2
of predator, are all unstable node point. Hence, the extinction of both the species or
predator only is impossible in model (4), which is similar to the case without Allee effect.
And for another boundary equilibrium (0, k2 ), which implies predator species tends to
change its food habits as predator has sufficient resources for alternative foods, in the
case without Allee effect, it is an unstable saddle; in the case with weak Allee effect, it
is an unstable saddle or a stable node; while in the case with strong Allee effect, it is
always stable, which means that the extinction of prey can occur in the case with strong
Allee effect, while the extinction of prey may occur in the case with weak Allee effect
under certain conditions, but the extinction of prey is impossible in the case without
Allee effect.
When model (4) has a unique positive equilibrium point, which means that predator
and prey can coexist in stable conditions and are independent of initial data, the model
undergoes complex dynamics. Especially, for certain parameter values, the model under-
goes a Hopf–bifurcation around the unique positive equilibrium, and may exhibit a stable
limit cycle, which can disappear when the parameters vary. That’s to say, the predator

22
and prey can coexist oscillating around the positive equilibrium. When the equilibrium is
an unstable node and the limit cycle disappears, the prey will ultimately go to extinction.
Furthermore, with Allee effect on prey, model (4) can have two positive equilibria.
One is stable, and the other is unstable. The model demonstrates the existence of a
bistability phenomenon. This is different from other Leslie–Gower models with Allee
effect analyzed above [18, 20]. We find that there exists a separatrix curve determined
by the stable manifold of the unstable positive equilibrium that separates the behavior of
trajectories of the system. The orbits initiating the right of the separatrix curve tend to
the positive equilibrium that represents the coexistence of the prey and predator, while
the orbits initiating the left of the separatrix curve tend to the boundary equilibrium that
represents the extinction of the prey. This means that, in this situation, the trajectories
of the model can have different behavior strongly depending on the initial conditions. It’s
worthy to note that there is a slight difference between the emergency of the bistability
with weak and strong Allee effect. If the model exhibits two positive equilibria, and one of
them is locally stable, then model (4) with strong Allee effect must be a bistable system.
But in the case with weak Allee effect, the model is not necessarily a bistable system,
because the boundary equilibrium can be a stable node under certain conditions. In this
viewpoint, to keep the prey and predator species coexist, we must regulate and control
the parameters in the right of the separatrix curve; otherwise, the predator species may
tend to change its food habits as predator has sufficient resources for alternative foods.
Hence, it is important for ecologists to be aware of the kind of bistability shown here.
In short, the consideration of Allee effect on prey population has a high impact on
the dynamics of the classic Leslie–Gower model, resulting in a new system with various
dynamical behavior, which has adequate biological meanings.

23
Appendix: Proof of Lamma 3.1

Proof. The maximum and minimum values of the function h(z) are respectively M1 =
√ √
q + 2 |p|3 , m1 = q − 2 |p|3 .
(a) It can be easily shown that if q < 0, then we have
(a1) If p ≥ 0, since h′ (z) = 3(z 2 + p) ≥ 0, we have h(z) is strictly increasing and
continuous in [0, +∞), this yields h(z) ≥ h(0) = q. Consequently, h(z) has a positive
root.
(a2) If p ≤ 0, Eq. (10) has a positive root.
Hence, if q < 0, Eq. (10) has a positive root.
(b) When q > 0, we have that p < 0, otherwise, h(z) = z 3 + 3pz + q may not be equal
to zero.
(b1) If m1 = 0, i.e., q 2 + 4p3 = 0, then Eq. (10) has a positive root of multiplicity
two.
(b2) If m1 < 0, i.e., q 2 + 4p3 < 0, then Eq. (10) has two positive roots.
(c) When q = 0, we have that p < 0, otherwise, h(z) = z 3 + 3pz + q may not be equal
to zero. And Eq. (10) has a unique positive root.

Acknowledgements

The authors would like to thank the anonymous referee for very helpful suggestions
and comments which led to improvements of our original manuscript. This research
was supported by the National Science Foundation of China (61373005 & 11271290) and
Zhejiang Provincial Natural Science Foundation (LY12A01014).

References

[1] J. Wang, J. Shi, and J. Wei, Predator–prey system with strong Allee effect in prey.
J. Math. Biol., 62(3)(2011)291–331.

24
[2] P.H. Leslie, Some further notes on the use of matrices in population mathematics.
Biometrika, 35(3-4)(1948)213–245, .

[3] M.A. Aziz-Alaoui and M.D. Okiye, Boundedness and global stability for a predator–
prey model with modified Leslie-Gower and Holling–type II schemes. Appl. Math.
Lett., 16(7)(2003)1069–1075.

[4] A.F. Nindjin, M.A. Aziz-Alaoui, M. Cadivel, Analysis of a predator–prey model


with modified Leslie–Gower and Holling–type II schemes with time delay. Nonl.
Anal. RWA, 7(5)(2006)1104–1118.

[5] B.I. Camara, M.A. Aziz-Alaoui, Dynamics of predator–prey model with diffusion.
Dynam. Cont. Dis. Sys. A, 15(2008)897–906.

[6] F. Courchamp, T. Clutton-Brock, B. Grenfell, Inverse density dependence and the


Allee effect, Trends Ecol. Evolut. 14(10)(1999)405–410.

[7] P.A. Stephens, W.J. Sutherland, R.P. Freckleton, What is the Allee effect? OIKOS,
87(1999)185–190.

[8] P.A. Stephens, W.J. Sutherland, Consequences of the Allee effect for behaviour,
ecology and conservation, Trends Ecol. Evolut. 14(1999)401–405.

[9] F. Courchamp, J. Berec and J. Gascoigne. Allee effects in ecology and conservation.
New York: Oxford University Press, 2008.

[10] W.C. Allee, Animal aggregations: a study in general sociology. University of


Chicago Press, Chicago, 1931.

[11] B. Dennis, Allee effects: population growth, critical density, and the chance of
extinction. Nat. Resour. Model., 3(4)(1989)481–538.

25
[12] G. Wang, X.G. Liang, F.Z. Wang, The competitive dynamics of populations subject
to an Allee effect. Ecol. Model., 124(1999)183–192.

[13] S.R. Zhou, Y.F. Liu, G. Wang, The stability of predator–prey systems subject to
the Allee effects. Theor. Popul. Biol., 67(1)(2005)23–31.

[14] J. Wang, J. Shi, J. Wei, Dynamics and pattern formation in a diffusive predator–prey
system with strong Allee effect in prey, J. Differential Equations, 251(2011)1276–
1304.

[15] L. Berec, E. Angulo, F. Courchamp, Multiple Allee effects and population manage-
ment. Trends Ecol. Evol., 22(4)(2007)185–191.

[16] P. Aguirre, E. González-Olivares, E. Sáez, Two limit cycles in a Leslie–Gower


predator–prey model with additive Allee effect. Nonl. Anal. RWA, 10(3)(2009)1401–
1416.

[17] P. Aguirre, E. González-Olivares, E. Sáez, Three limit cycles in a Leslie–


Gower predator–prey model with additive Allee effect. SIAM J. Appl. Math.,
69(5)(2009)1244–1269, 2009.

[18] E. González-Olivares, J. Mena-Lorca, A. Rojas-Palma, J. D. Flores, Dynamical com-


plexities in the Leslie–Gower predator–prey model as consequences of the Allee effect
on prey, Appl. Math. Model. 35(2011)366–381.

[19] M.H. Wang and M Kot, Speeds of invasion in a model with strong or weak Allee
effects, Math. Biosci., 171(2001)83–97.

[20] J. Mena-Lorca, E. González-Olivares, B. González-Yañez, The Leslie–Gower


predator–prey model with Allee effect on prey: A simple model with a rich and
interesting dynamics. Proceedings of the 2006 International Symposium on Mathe-
matical and Computational Biology BIOMAT, pp. 105–132, 2006.

26
[21] S.V. Petrovskii, A.Y. Morozov, E. Venturino, Allee effect makes possible patchy
invasion in a predator–prey system. Ecol. Lett., 5(3)(2002)345–352.

[22] C.M. Taylor, A. Hastings, Allee effects in biological invasions. Ecol. Lett.,
8(8)(2005)895–908.

[23] J. Shi, R. Shivaji, Persistence in reaction diffusion models with weak Allee effect.
J. Math. Biol., 52(6)(2006)807–829.

[24] G.A.K. van Voorn, L. Hemerik, M.P. Boer, B.W. Kooi, Heteroclinic orbits indicate
overexploitation in predator-prey systems with a strong Allee effect. Math. Biosci.,
209(2)(2007)451–469.

[25] S.Wiggins, Introduction to applied nonlinear dynamical systems and chaos. Springer,
New York, 1990.

27
Fig.1 The phase portrait of model (4) with weak Allee effect. The parameters are
taken as b = 0.4, c = 0.5, k1 = 0.3, k2 = 0.2, m = 0.24 and s = 0.125. Ew0 = (0, 0)
is an unstable node point, Ew1 = (0, 0.2) and Ew2 = (0.8, 0) are saddle points.
Ew∗ = (0.2, 0.4) is locally asymptotically stable. The dashed curve is the N –nullcline
f (N, P ) = 0, and the dotted curve is the P –nullcline g(N, P ) = 0.

Fig.2 The phase portrait of model (4) with weak Allee effect. s = 0.08. Other
parameters are taken as Fig. 1. Ew∗ = (0.2, 0.4) is an unstable focus surrounded by
a stable limit cycle.


Fig.3 The phase portrait of model (4) with weak Allee effect. s = s[Nw ] = 0.093333.
Other parameters are taken as Fig. 1. The model enters into a Hopf–bifurcation

around Ew∗ = (0.2, 0.4) at s = s[Nw ] .

Fig.4 The bistability of model (4) with weak Allee effect. The parameters are taken
as b = 0.4, c = 0.5, k1 = 1.2, k2 = 0.2, m = 0.39 and s = 0.1. In this case, Ew0 =
(0, 0), Ew2 = (0.61623, 0) and Ew4 = (0.085053, 0.285053) are saddle points, Ew1 =
(0, 0.2) is a stable node point, and Ew3 = (0.23227, 0.43227) is a stable focus. There
exists a separatrix curve determined by the stable manifold of the equilibrium point
Ew4 . The dashed curve is the N –nullcline f (N, P ) = 0, and the dotted curve is the
P –nullcline g(N, P ) = 0.

Fig.5 The bistability of model (4) with strong Allee effect. The parameters are
taken as b = 0.25, c = 0.1, k1 = 0.035, k2 = 0.2, m = 0.3 and s = 0.1. In this
case, Es0 = (0, 0), Es3 = (0.67604, 0) and Es5 = (0.27369, 0.47369) are saddle points,
Es1 = (0, 0.2) is a stable node point, Es2 = (0.07396, 0) is an unstable node point, and
Es4 = (0.40275, 0.60275) is locally stable. There exists a separatrix curve determined
by the stable manifold of the equilibrium point Es5 . The dashed curve is the N –
nullcline f (N, P ) = 0, and the dotted curve is the P –nullcline g(N, P ) = 0.

28
0.8
P

0.7

0.6

0.5

0.4 *
Ew

0.3

0.2 Ew1

0.1

Ew0 E
w2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
N

Figure 1: The phase portrait of model (4) with weak Allee effect. The parameters are taken as b = 0.4, c =
0.5, k1 = 0.3, k2 = 0.2, m = 0.24 and s = 0.125. Ew0 = (0, 0) is an unstable node point, Ew1 = (0, 0.2)

and Ew2 = (0.8, 0) are saddle points. Ew = (0.2, 0.4) is locally asymptotically stable. The dashed curve
is the N –nullcline f (N, P ) = 0, and the dotted curve is the P –nullcline g(N, P ) = 0.

29
0.8
P

0.7

0.6

0.5

0.4 *
Ew

0.3

0.2 E
w1

0.1

Ew0 Ew2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
N

Figure 2: The phase portrait of model (4) with weak Allee effect. s = 0.08. Other parameters are taken

as Fig. 1. Ew = (0.2, 0.4) is an unstable focus surrounded by a stable limit cycle.

30
0.8
P

0.7

0.6

0.5

0.4 *
Ew

0.3

0.2 Ew1

0.1

Ew0 Ew2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 N 0.8


Figure 3: The phase portrait of model (4) with weak Allee effect. s = s[Nw ] = 0.093333. Other parameters


are taken as Fig. 1. The model enters into a Hopf–bifurcation around Ew = (0.2, 0.4) at s = s[Nw ] .

31
0.8
P

0.7
Separatrix

0.6

0.5

E
w3
0.4

0.3
Ew4

0.2
Ew1

0.1

Ew0 Ew2
0
0 0.1 0.2 0.3 0.4 0.5 N 0.6

Figure 4: The bistability of model (4) with weak Allee effect. The parameters are taken as b = 0.4, c =
0.5, k1 = 1.2, k2 = 0.2, m = 0.39 and s = 0.1. In this case, Ew0 = (0, 0), Ew2 = (0.61623, 0)
and Ew4 = (0.085053, 0.285053) are saddle points, Ew1 = (0, 0.2) is a stable node point, and
Ew3 = (0.23227, 0.43227) is a stable focus. There exists a separatrix curve determined by the stable
manifold of the equilibrium point Ew4 . The dashed curve is the N –nullcline f (N, P ) = 0, and the dotted
curve is the P –nullcline g(N, P ) = 0.

32
0.7
P

Separatrix Es4
0.6

0.5
Es5

0.4

0.3

0.2
E
s1

0.1

E Es2 Es3
s0
0
0 0.1 0.2 0.3 0.4 0.5 0.6 N 0.7

Figure 5: The bistability of model (4) with strong Allee effect. The parameters are taken as b = 0.25, c =
0.1, k1 = 0.035, k2 = 0.2, m = 0.3 and s = 0.1. In this case, Es0 = (0, 0), Es3 = (0.67604, 0) and
Es5 = (0.27369, 0.47369) are saddle points, Es1 = (0, 0.2) is a stable node point, Es2 = (0.07396, 0) is
an unstable node point, and Es4 = (0.40275, 0.60275) is locally stable. There exists a separatrix curve
determined by the stable manifold of the equilibrium point Es5 . The dashed curve is the N –nullcline
f (N, P ) = 0, and the dotted curve is the P –nullcline g(N, P ) = 0.

33

You might also like