Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Available online at www.sciencedirect.

com
ScienceDirect

Mathematics and Computers in Simulation 202 (2022) 526–539


www.elsevier.com/locate/matcom

Original articles

Dynamics and stability of two predators–one prey mathematical


model with fading memory in one predator
Zeynep Yılmaz, Selahattin Maden ∗, Aytül Gökçe ∗
Department of Mathematics, Faculty of Arts and Sciences, Ordu University, 52200 Ordu, Turkey
Received 12 March 2022; received in revised form 3 June 2022; accepted 24 July 2022
Available online 30 July 2022

Abstract
This paper concentrates on dynamics and stability analysis of two predators–one prey mathematical model with competition
between predators and fading memory in one predator. The investigation of the constructed model shows that there exist
five equilibria, e.g. trivial extinction state of all populations, extinction of both predators state, extinction of first or second
predator state and coexisting state. Investigating the eigenvalues of characteristic polynomial, conditions for the local stability
around each equilibrium are also determined depending on the parameter space. Analytical formulations are complemented with
numerical simulations, where time simulations and single parameter numerical continuation of each variable are performed with
respect to model parameters and multiple sub-and super-critical Hopf bifurcations, period doubling bifurcation and transcritical
bifurcation are detected for different values of memory related parameter. Our results show that fading memory and competition
between predators have substantial impact on the existence and dynamics of all three populations and may shed lights on further
understanding of interacting species in ecology.
© 2022 International Association for Mathematics and Computers in Simulation (IMACS). Published by Elsevier B.V. All rights
reserved.
Keywords: Two predators–one prey model; Stability analysis; Recent past; Bifurcation; Dynamical systems

1. Introduction
The complex interactions have a significant place in structuring ecological communities. The intermingling in
one place of ecological and biological systems is not limited to only two types of populations. In fact, interactions
of prey–predator species with more than two species have received a considerable attention by many researchers.
Recently several mathematical models with one prey and two predators have been thoroughly investigated. In
general, both predators are usually considered to have equal properties and to fed on the same prey [23,33,34]. On
the other hand, in some models, one of the predators can have different properties. For example, one predator can be
super-predator feeding on both prey and the other predator [27]. A general opinion is that one predator determines
its self territory for hunting and the presence of other predators in this territory is not tolerable. This leads to a
competition between predators and also has received a significant notice in mathematical ecology [3,5,30]. In an
early study, Hsu et al. investigated a three species model with one prey and two predators, and their results showed
that one of the predators tended to undergo extinction in time [20]. Dubey and Upadhyay analysed the conditions
∗ Corresponding authors.
E-mail addresses: zeynepyilmaz@odu.edu.tr (Z. Yılmaz), smaden@odu.edu.tr (S. Maden), aytulgokce@odu.edu.tr (A. Gökçe).

https://doi.org/10.1016/j.matcom.2022.07.023
0378-4754/© 2022 International Association for Mathematics and Computers in Simulation (IMACS). Published by Elsevier B.V. All rights
reserved.
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

for extinction and persistence of one prey and two predators model [10]. The role of fear in two predators–one
prey model is also studied by Mukherjee in [28]. More research focusing on two predators–one prey systems can
be found in [1,22,24]. In this paper, we are interested in a two predators–one prey model, where density of one
predator is dependent on fading memory.
One of the crucial mathematical components to predict various dynamics in ecology and biology is the interaction
function for species. Although many types of functional forms with specific manner of interactions have been
developed in the literature, non-linear Holling type functional responses saturating at large prey concentration
(I I, I I I and I V ) have been popular choice in modelling the real world phenomena [9,19,21]. Holling type II
response is particularly preferred as it presumes that the proportion of consumed prey (or rate of predation) is
proportional to prey density when prey species is scarce in the system. Therefore, although other type of functional
responses are possible, we chose the most common choice of Holling type II functional response [2,21,29,31].
It is usually considered that the growth rates of prey and predator species are respectively influenced by the
densities of current predator and current prey species. Although it is plausible to consider that the growth rate of
prey may depend on the predator density at present, a more realistic model can be achieved when the growth rate of
predator is associated with the prey density in the present and also that in the recent past [2,6,16,31]. Therefore there
has been a current interest in understanding the role of fading memory in the dynamics of prey–predator interactions.
In this context, a continuous density function, weighting the moments of the past, can be described [16,18,31].
Another important factor that critically affects the species interactions is the competition between predators.
Available resources for species are usually finite and thus it is more likely that predators living in the same territory
and searching for similar type of food will compete. Therefore, considering competitive predators may be another
determinant in the system and can be frequently observed in nature [4,7,8,33]. From the modelling perspective,
several models taking two predators competing on the same prey population have been proposed, for example
see [1,32,33].
A generic population model with one prey and two predators, that includes the role of recent past in one predator
and the effect of competition between predators, has not been investigated, elsewhere. A basic prey–predator model
with weak Allee effect including memory with a hump was proposed by Ferreira et al. using Holling type I functional
response [15]. Another model with general Holling interactions was provided by Sahoo et al., where harvesting
effort and the moments of the past are investigated for a prey–predator model [31]. Then, the influence of past in
a population system incorporating intraspecific competition and Allee effect was analysed by Gökçe [18]. These
models were proposed without considering the effect of the moments of the past and competition between multiple
predators. Compared to classical prey–predator models with Holling type II functional response, the aim of this study
is to explore the dynamics if another predator, density of which depends of the past quantities of prey, appears in
the system and competes with other predator.
This paper is organised in a couple of sections. The main model is described in Section 2. The preliminary results
of the model including non-dimensionalisation and the existence of various equilibria are presented in Sections 2.1
and 2.2. The conditions for the local stability corresponding to each equilibria are discussed in Section 3. Besides,
the extensive numerical simulation results are demonstrated in Section 4, where the effect of fading memory on
the system dynamics is emphasised using time evolution of all three populations and also using single parameter
numerical bifurcation analysis. Here, as an example, the results are presented for the numerical continuation along
half saturation constant and conversion ratios of prey biomass into both predators. Lastly, in Section 5, the outcomes
of this research are summarised with potential future directions.

2. The model description


In this section two predators–one prey model is considered with the following assumptions: (i) both predators
feed on the same prey, (ii) one of the predators is memory dependent, namely its density changes with both current
and recent past density of prey, (iii) both predators live in the same territory and compete, (iv) the intake rate of
predators as a function of prey is determined according to Holling type II functional response, (v) prey has a logistic
growth with a environmental carrying capacity in the absence of both predators.
A continuous density function is used to incorporate memory, by which predator density is associated with prey
density in present and also that in recent past. Following the ideas presented by∫Farkas [13,14], MacDonald [25,26]
t
and Sahoo and Poria [31], the moments of the past is weighed using R(t) = −∞ G(t˜) R̂(t − t˜)dt˜, where function
527
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Table 1
Description of the variables and parameters of the model (2)-(5).
Variable Description
G Prey population
Q1 First predator population
Q2 Second predator population
R Memory variable
Parameter Description
ρ Prey growth rate
K Environmental carrying capacity
β1 , β2 Capturing rates of the predators Q 1 and Q 2 on the prey G
α1 , α2 Half saturation constants
ϵ1 , ϵ2 Conversion rates of the prey to predators Q 1 and Q 2 , respectively
d1 , d2 Natural death rates of Q 1 and Q 2 , respectively
c1 , c2 Competition of Q 1 and Q 2 , respectively
ω Memory constant

∫∞
R̂ satisfies the conditions R̂ : [0, ∞) → R and 0 R̂(s)ds = 1. Differentiating this, one obtains the following
differential equation
dR(t)
= ω(G(t) − R(t)), (1)
dt
for which a basic probability density function is taken as R̂(t) = ωe−ωt . Here ω can be denoted as a memory
constant, namely smaller ω implies the existence of past influence for larger time interval [31]. In other words, 1/ω
implies the measure of the past effect. It is worth noting the exponentially decaying probability density function
is taken into account to weight moments of the time, assuming that the influence of past is exponentially fading
away. This idea is biologically plausible, as using a biologically meaningful choice of weight function, the largest
moments are given to moments in the neighbourhood of present [31]. However, when we go far into past, this
weight decreases exponentially and thus the dynamics of the predator depends on the recent past of prey quantity.
The papers published by Cavani and Farcas [6] and Sahoo and Poria [31] similarly present Holling type II functional
response, where present prey value in the time evolution of predator is replaced by time average of prey density
over the recent past. For details of this approach, we refer the reader to [6,16,31].
We now describe the model with four components for prey predator interactions as
β1 G Q 1 β2 G Q 2
( )
dG G
= ρG 1 − − − , (2)
dt K α1 + G α2 + G
dQ 1 ϵ 1 β1 G Q 1
= − d1 Q 1 − c1 Q 1 Q 2 , (3)
dt α1 + G
dQ 2 ϵ 2 β2 R Q 2
= − d2 Q 2 − c2 Q 1 Q 2 , (4)
dt α2 + R
dR
= ω(G − R), (5)
dt
where the predator Q 2 has a memory effect in its dynamics. The parameters and variables are described in Table 1.

2.1. Dimensionless model

Considering new variables as


Kρ ′ Kρ ′ 1 ′
G = K G ′, Q1 = Q , Q2 = Q , R = K R′, t= t,
β1 1 β2 2 ρ
and new parameters as
αi ϵi βi di
αi = , βi = , di = , i = 1, 2,
K ρ ρ
528
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

c1 K c2 K ω
c1′ = , c2′ = , ω′ = ,
β2 β2 ρ
the new dimensionless model can be written as
dG G Q1 G Q2
= G(1 − G) − − , (6)
dt α1 + G α2 + G
dQ 1 β1 G Q 1
= − d1 Q 1 − c1 Q 1 Q 2 , (7)
dt α1 + G
dQ 2 β2 R Q 2
= − d2 Q 2 − c2 Q 1 Q 2 , (8)
dt α2 + R
dR
= ω(G − R). (9)
dt
Here primes are omitted for simplicity. In the rest of the paper, the non-dimensional system of (6)–(9) will be
considered.

2.2. Equilibrium analysis

The system (6)–(9) holds five equilibrium points. Here it can be obviously seen that the extinction state always
exists for all parameter values. Eqs. (6)–(9) have one coexisting state which exists in a certain parameter space.
The other three equilibria are determined whether the system is first predator free, second predator free or both.
We determine all possible cases in the following.
1. Trivial State: E 0 = (0, 0, 0, 0).
2. Both predators free state: E 1 = (1, 0, 0, 1). Using Eqs. (6) and (9) it is obvious that G = 1 and G = R.
3. Predator Q 1 free state: E 2 = (G (1) (1)
∗ , 0, Q 2 ∗ , R∗ ). Substituting Q 1 = 0 in Eqs. (6) and (8) we obtain
(1)

G (1) (1)
∗ Q2∗
G (1) (1)
∗ (1 − G ∗ ) − = 0, (10)
α2 + G (1)

β2 G (1) (1)
∗ Q2∗
− d2 Q 2 (1)
∗ = 0, (11)
α2 + G (1)

and these lead to G (1) (1) (1)


∗ = d2 α2 /(β2 − d2 ) and Q 2 ∗ = (1 − G ∗ )(α2 + G ∗ ). Thus the predator Q 1 free
(1)

equilibrium is found as
d2 α2 β2 α2 (β2 − d2 − d2 α2 ) d2 α2
( )
E2 = , 0, , , (12)
β2 − d2 (β2 − d2 )2 β2 − d2
with β2 > d2 (1 + α2 ). This condition biologically implies that capturing rate of predator Q 2 should be larger
than its death rate for the existence of the first predator free equilibrium.
(2)
4. Predator Q 2 free state: E 3 = (G ∗ (2) , Q 1 (2) ∗ , 0, R∗ ). Similarly, using Eqs. (6) and (7), the second predator
free state is found as
d1 α1 α1 β1 (β1 − d1 − d1 α1 ) d1 α1
( )
E3 = , , 0, , (13)
β1 − d1 (β1 − d1 )2 β1 − d1
with β1 > d1 (1 + α1 ).
5. Nonzero state: E 4 = (G ∗ , Q ∗1 , Q ∗2 , R ∗ ). Using Eqs. (7) and (8), we obtain
β1 G ∗ β2 G ∗
( ) ( )
1 1
Q ∗2 = 2 .

− d1 and Q = − d (14)
c1 α1 + G ∗ 1
c 2 α2 + G ∗
Substituting Eq. (14) in Eq. (6) we obtain a third order polynomial of prey:
G ∗ 3 + f 0 G ∗ 2 + f 1 G ∗ + f 2 = 0, (15)
where
f 0 = −(1 − α1 − α2 ),
f 1 = −(α1 + α2 − α1 α2 ) + (β2 − d2 )/c2 + (β1 − d1 )/c1 , (16)
f 2 = −α1 α2 − (c1 α2 d2 + c2 α1 d1 )/(c1 c2 ).
529
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Thus the positive coexisting state is described by


G ∗ (β2 − d2 ) − d2 α2 G ∗ (β1 − d1 ) − d1 α1 ∗
( )
E4 = G ∗, , , G , (17)
c2 (α2 + G ∗ ) c1 (α1 + G ∗ )
for which according to Cardano’s theorem, the solution of Eq. (15) is obtained as
1( p)
G∗ = − f0 + C + , (18)
3 C
with
( √ )1/3
q + q 2 − 4 p3
C= ,
2
p = f 02 − 3 f 1 , q = 2 f 03 − 9 f 0 f 1 + 27 f 2 .
Here, the conditions G ∗ (β2 − d2 ) > d2 α2 and G ∗ (β1 − d1 ) > d1 α1 with a positive G ∗ should be satisfied for
the coexisting state.

3. Local stability analysis


The local stability of the model described by (6)–(9) can be investigated through linearisation argument around the
equilibrium point, i.e. at the coexisting equilibrium (G ∗ , Q ∗1 , Q ∗2 , R ∗ ). We use the linearisation by finding Jacobian
matrix of the system considering G = G ∗ + G ′ , Q 1 = Q ∗1 + Q ′1 , Q 2 = Q ∗2 + Q ′2 and R = R ∗ + R ′ . This leads
to a linearised model in the form of
dF
= A| E j F, j = 0, 1, 2, 3, 4. (19)
dt
T
Here F = G , Q 1 , Q ′2 , R ′ , where the variables with primes represent the perturbed variables. Besides, the
[ ′ ′ ]
Jacobian matrix is described by
⎡ ⎤⏐
f G f Q 1 f Q 2 f R ⏐⏐
⎢ gG g Q gQ 2 gR ⎥⎥⏐ , j = 0, 1, 2, 3, 4,

A| E j = ⎢ 1
⎣h G h Q h Q2 h R ⏐
⎦ ⏐
1
pG p Q 1 pQ pR ⏐
2 Ej

for which
α1 Q 1 α2 Q 2 G G
f G = 1 − 2G − − , fQ1 = , fQ2 = , f R = 0,
(α1 + G)2 (α2 + G)2 α1 + G α2 + G
β1 Q 1 α 1 β1 G
gG = , gQ 1 = − d1 − c1 Q 2 , g Q 2 = −c1 Q 1 g R = 0,
(α1 + G) 2 α1 + G
β2 R β2 Q 2 α 2
h G = 0, h Q 1 = −c2 Q 2 , h Q 2 = − d2 − c2 Q 1 , h R = ,
α2 + R (α2 + R)2
pG = ω, p Q 1 = 0, p Q 2 = 0 p R = −ω.
We now describe several theorems for the local stability of each equilibrium.

Theorem 3.1. The trivial equilibrium given by E 0 = (0, 0, 0, 0) always exists and it is unstable.

Proof. The characteristic matrix around the trivial state E 0 is given by


1−µ
⎡ ⎤
0 0 0
⎢ 0 −d 1 − µ 0 0 ⎥ ⎥,
A| E0 = ⎢
⎣ 0 0 −d 2 − µ 0 ⎦
ω 0 0 −ω − µ
the eigenvalues of which is found using Det A| E0 − µI4 = 0 leading to
[ ]

µ1 = 1, µ2 = −d1 , µ = −d2 and µ = −ω.


Since the first eigenvalue is positive, the system around the extinction state is an unstable node. □
530
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Theorem 3.2. The predators free state E 1 = (1, 0, 0, 1) is unstable if β2 > d2 (α2 + 1) or β1 > d1 (α1 + 1).

Proof. The stability of the system around the equilibrium E 1 = (1, 0, 0, 1) is computed using Det A| E1 − µI4 = 0
[ ]

leading to the characteristic polynomial:


(h Q 2 − µ)( p R − µ)( f G − µ)(g Q 1 − µ)⏐ E = 0.

(20)
1

Then the eigenvalues are given by


β2 β1
µ1 = − d2 , µ2 = −ω, µ3 = −1, µ4 = − d1 . □
α2 + 1 α1 + 1
Thus the system may be destabilised through eigenvalues µ1 and µ4 , respectively satisfying the conditions
β2 > d2 (α2 + 1) or β1 > d1 (α1 + 1).

Theorem 3.3. The stability of the predator Q 1 free state E 2 = (G ∗ , 0, Q ∗2 , R ∗ ) is determined depending on the
values of s and t given in Eqs. (22) and (23) or depending whether the condition β1 G (1) (1)
∗ < (α1 + G ∗ )(d1 + c1 Q 2 ∗ )
(1)

is fulfilled (stable).

Proof. The characteristic polynomial around the predator Q 1 free state E 2 is given by
µ − g Q 1 (µ − p R ) (µ − f G )(µ − h Q 2 ) + f Q 2 h R pG ⏐ E = 0.
( )[ ]⏐
2

Using Cardano’s technique to find the solutions of a cubic polynomial leads to eigenvalues
β1 G (1)

µ1 = g Q 1 = ∗ ,
− d1 − c1 Q 2 (1) (21)
α1 + G (1)

µ2 = s + t − a1 /3, (22)

s+t a1 i 3
µ3,4 =− − ± (s − t), (23)
2 3 2
where
)1/3 )1/3
s = r + (q 3 + r 2 )1/2 , t = r − (q 3 + r 2 )1/2 ,
( (

3a2 − a12 9a1 a2 − 27a3 − 2a13


q= , r= ,
(9 54
a1 = − f G + h Q 2 + p R , a2 = f G h Q 2 + f G p R + h Q 2 p R ,
)

a3 = f Q 2 h R p G − f G h Q 2 p R .
Thus, the system becomes unstable if the condition β1 G (1) (1)
∗ > (α1 + G ∗ )(d1 + c1 Q 2 ∗ ) is satisfied. Besides, the
(1)

stability around the predator Q 1 free state depends on the values of s and t (real or complex). □

Theorem 3.4.⏐ The predator Q 2 free state E 3 = (G ∗ , Q ∗1 , 0, R ∗ ) is unstable if β2 R∗(2) > (d2 + cQ 1 (2) (2)
∗ )(α2 + R∗ ) or
Tr( f G + g Q 1 ) E > 0.

3

Proof. The stability of the system around the equilibrium E 3 = (G ∗ , Q ∗1 , 0, R ∗ ) is computed using Det A| E3 − µI4
[ ]

= 0 leading to the characteristic polynomial:


[ ]
f f Q1
h Q 2 − µ ( p R − µ) µ2 − Tr(D)µ + Det(D) ⏐ E = 0, D = G ,
( ) [ ]⏐
3 gG g Q 1
and eigenvalues:

β2 R∗(2) Tr(D) Tr(D)2 − 4Det(D)
µ1 = −ω, µ2 = − d2 − ∗ ,
cQ 1 (2) µ3,4 = ± .
α2 + R∗(2) 2 4
Thus the system given by (6)–(9) around the equilibrium E 3 may be unstable when β2 R∗(2) > (d2 +cQ 1 (2) (2)
∗ )(α2 + R∗ ).
Besides, if 0 < Det(D) < Tr(D)2 /4, eigenvalues µ3,4 are found real and of the same sign (node); if 0 < Tr(D)2 /4 <
531
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Det(D), then eigenvalues µ3,4 become neither real nor purely imaginary (spiral). In both cases, the predator Q 2 free
state is found to be unstable with Tr(D) > 0. □
Similarly, the stability of the coexistence state E 4 = (G ∗ , Q ∗1 , Q ∗2 , R ∗ ) is determined using Det A| E4 − µI4 = 0
[ ]
which becomes
( p R − µ) h Q 2 − µ µ2 − Tr(D) + Det(D) − h Q 1 f G g Q 2 − f Q 2 gG
[( )( ) (
(24)
−g Q 2 µ + pG f Q 2 h R µ + f Q 1 g Q 2 h R − f Q 2 g Q 1 h R = 0,
)] ( )

which can be rewritten as µ4 + b1 µ3 + b2 µ2 + b3 µ + b4 = 0 where


b1 = − p R − h Q 2 − Tr(D),
b2 = h Q 2 + p R Tr(D) + Det(D) + p R h Q 2 − h Q 1 g Q 2 ,
( )
( )
b3 = − p R h Q 2 Tr(D) − p R + h Q 2 Det(D) + h Q 1 g Q 2 ( p R + f G ) + f Q 2 ( pG h R − h Q 1 gG ),
b4 = p R h Q 2 Det(D) − p R h Q 1 f G g Q 2 − f Q 2 gG + pG h R f Q 1 g Q 2 − f Q 2 g Q 1 .
( ) ( )

The eigenvalues corresponding to Eq. (24) are found using classical Ferrari’s method by changing the variables and
exploiting the formulas for the quadratic and cubic equations:

b1 1 y2
µ1,2 = − − u ± −4u 2 − 2y1 + , (25)
4 2 u

b1 1 y2
µ3,4 = − + u ± −4u 2 − 2y1 − , (26)
4 2 u
where
8b2 − 3b12 b3 − 4b1 b2 + 8b3
y1 = , y2 = 1 ,
8 8
√ ⎛ √ ⎞1/3
2 3
1 2 1
(
∆0
) ∆ 1 + ∆ 1 − 4∆0
u= − y1 + v+ , v=⎝ ⎠ ,
2 3 3 v 2

∆0 = b22 − 3b1 b3 + 12b4 , ∆1 = 2b23 − 9b1 b2 b3 + 27b12 b4 + 27b32 − 72b2 b4 .


Thus, eigenvalues calculated in Eqs. (25) and (26) are accounted as critical descriptors of the model stability
around the coexisting state (E 4 ) and depends on the system parameters. Here, the particular stability of the system
is associated with the existence of real and imaginary parts of the eigenvalues µ1,2 and µ3,4 , and also on the signs
of the real components of their values.

4. Numerical outcomes
In this section, firstly time simulations of prey and predators and corresponding phase portraits are shown for
low and high values of ω. Here small ω implies that the past effect exists for a larger time interval. In order to
clearly see the influence of past incorporated in one predator, e.g. Q 2 , we chose equal strength of half saturation (α1
and α2 ) and conversion efficacy (β1 and β2 ) constants associated with both predators Q 1 and Q 2 . Thus we assume
that α = α1 = α2 and β = β1 = β2 . Unless otherwise stated, the parameters for time simulations are fixed to
d1 = d2 = 0.1, c1 = c2 = 0.2 for small (ω = 0.1) and high (ω = 1) values of memory constant. By choosing equal
strength of parameters for both predators, our objective is to measure the effect of recent past in the dynamics.
In Fig. 1, time evolutions of prey and two predators are shown in the absence of competition for small and
large values of memory parameter ω. It is clear that long time memory (small ω) without competition leads to
unstable behaviour in the dynamics of prey G and predator Q 2 . Density of predator Q 1 starts with an oscillatory
nature and decrease to zero after t = 1000, see Fig. 1(a). As seen in Fig. 1(b), large values of ω in the absence
of competition remove the oscillatory behaviour and extinction tendency of solutions, leading to stable equilibria
for all populations. This case is actually close to the behaviour for which there is no memory effect in the system.
Given that predator density depends on the past density of prey for a larger time interval, e.g. Fig. 1(a), stability of
all species may switch from stable to unstable or even from stable to extinction. Thus, parameters associated with
weakening memory has an important impact on the stability and existence of the population densities.
532
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Fig. 1. Densities of prey and predator species when ω = 0.1 (a) and ω = 1 (b), and their corresponding phase portraits (c)–(d) in the absence
of competition. The phase portraits with respect to Q 1 (extinct and stable) are not shown. The initial and end points of the trajectories are
presented with red and magenta dots. Other parameters are α = 0.4, β = 0.2, d1 = d2 = 0.1, c1 = c2 = 0. The initial condition is chosen as
(0.5,0.4,0.4,0.5). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 2 concentrates on the species dynamics for small (ω = 0.1) and large (ω = 1) values of memory parameter
along with an equal strength of competition that affects both predators Q 1 and Q 2 . For both case, predator Q 1
undergoes extinction, yet the stability of predator Q 2 switches from unstable to stable with ω = 1. In fact, our
numerical results shows that usually either Q 1 or Q 2 tends to die out in time. This complements the results
previously given in the literature [20] that one of the predators tends to extinct, and here parameter ω, associated
with the effect of recent past, may become a key determinant in which predator will survive.
In Figs. 3, 4 and 5, the model given by (6)–(9) with various rates of memory term is investigated using single
parameter numerical continuation diagrams. As an example, the stability of the steady states E i i = 0, 1, 2, 3, 4 is
determined as α (half saturation constant) and β (strength of conversion efficacy) parameters are varied for low and
high memory value ω. It is worth reminding that smaller values of ω parameter implies that the past influence occurs
for a larger time interval. Line specifications and corresponding number of unstable eigenvalues are described in
Table 2.
In Fig. 3, stability analysis of prey population is shown with respect to half saturation constant (α) and conversion
efficacy of the prey G into new predators (β). Since our focus is to understand the role of memory, we choose
α1 = α2 and β1 = β2 and call these as α and β respectively. For all cases, the extinction state is always unstable
with Uµ = 1; and Q 1 and Q 2 free (E 1 ) states are unstable with Uµ = 2. For the coexisting state, a Hopf bifurcation
H1 is encountered at α = 0.5354, where the number of unstable eigenvalues changes from Uµ = 3 to Uµ = 1. The
solution branches of Q 1 free (E 2 ) or Q 2 free (E 3 ) are on top of each other in all figures. In Fig. 3(a), Q 1 free state
533
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Fig. 2. Time evolution of prey and predator populations when ω = 0.1 (a) and ω = 1 (b), and their corresponding phase portraits (c)–(d).
The phase portraits with respect to Q 1 (extinct) are not shown. The initial and end points of the trajectories are presented with red and
magenta dots. Other parameters are α = 0.64, β = 0.33, d1 = d2 = 0.1, c1 = c2 = 0.2. The initial condition is chosen as (0.5,0.4,0.4,0.5).
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Table 2
Descriptions of line colours in Figs. 3–5. (For interpretation of the references to
colour in this table, the reader is referred to the web version of this article.)
Line Colour Number of unstable eigenvalues Stability
Uµ = 0 for E i , i = {0, 1, 2, 3} stable
Uµ = 1 for E i unstable
Uµ = 2 for E i unstable
Uµ = 3 for E i unstable
Uµ = 2 for E 2 unstable
Uµ = 0 for E 3 stable
Uµ = 1 for E 0
Uµ = 2 for E 1 unstable
Uµ = 2 for E 2 or E 3
Uµ = 1 for E 0 unstable
Uµ = 2 for E 1 unstable
Uµ = 0 for E 2 or E 3 stable

is always unstable with Uµ = 2 for α ∈ [0, 1]. Stability of Q 2 free state switches from unstable (Uµ = 2) to stable
(Uµ = 0), where another Hopf bifurcation H2 occurs with a stable limit cycle arising from the Hopf bifurcation. The
case with ω = 1 is demonstrated in Fig. 3(b), where one Hopf bifurcation H2 occurs at α = 0.666 for Q 2 free state
534
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Fig. 3. Single parameter numerical continuation of prey population with respect to α and β parameters with ω = 0.1 (a)–(c) and ω = 1
(b)–(d). The number of eigenvalues that are positive or have positive real parts (Uµ ) are represented with different colours along the steady
states of the system, see Table 2. Other parameters are d1 = d2 = 0.1 and c1 = c2 = 0.2. The branches arising from the Hopf points H1 ,
H2 and H3 are the maximums of the periodic orbits. The other bifurcations Pi (i = 1, 2) and T1 are respectively for periodic doubling and
transcritical bifurcations. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)

and another Hopf bifurcation H3 occurs at α = 0.864 for Q 1 free state. In addition, a period doubling bifurcation
occurs at α = 0.4 along the periodic orbit arising from H3 . Similarly, bifurcation diagrams of G with respect to β
(= β1 = β2 ) for ω = 0.1 and ω = 1 are demonstrated in Fig. 3(c)–(d) respectively. Along the branch emanating
from the Hopf bifurcation H3 , two period doubling bifurcations occur where the number of unstable eigenvalues
changes from Uµ = 1 to Uµ = 0. In contrast to other Hopf bifurcations (all found to be supercritical) given in these
figures, the Hopf bifurcation H3 in Fig. 3(c) is found to be subcritical with Liapunov constant L(H3 ) = 0.1609.
Besides, the amplitude of the periodic orbit arising from Hopf point H1 for coexisting state is larger for small
values of ω. Apart from these findings, there is no substantial difference between bifurcation diagrams for ω = 0.1
in Fig. 3(c) and ω = 1 in Fig. 3(d).
535
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Fig. 4. Bifurcation diagram of predator Q 1 and predator Q 2 populations with respect to α and β parameters with ω = 0.1. The number of
eigenvalues that are positive or have positive real parts (Uµ ) are represented with different colours along the steady states of the system,
see Table 2. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

The main focus of this paper is to investigate the effect of memory on the dynamics of each predator, where
only Q 2 is directly exposed to the recent influence of past. In Fig. 4, we fix ω = 0.1, the past influence occurs for a
longer time interval. This can be observed in Eqs. (7)–(8), where the past influence is incorporated in the dynamics
of predator Q 2 . As seen in Fig. 4(a) and Table 2, considering ω = 0.1, extinction state E 0 and both predators free
state E 1 are found to be unstable with Uµ = 1 and Uµ = 2 respectively. Q 1 free state is also unstable with Uµ = 2.
Here, two Hopf bifurcations occur (one for coexisting state and the other is for Q 2 free state). The unstable (for H1 )
and stable (for H2 ) periodic orbits arising from these Hopf points intersect in two periodic doubling bifurcations
(P1 and P2 ). Dynamics of predator Q 2 which has been directly influenced by past (with ω = 0.1) is shown in
Fig. 3(b), where Q 1 free state becomes unstable for α ∈ [0, 1]. The branches emanating from H1 are restricted
in a small region. Note that, similar to Fig. 4(a), extinction state and the both predators free state are unstable.
Note that although it appears that the Hopf point for Q 2 free state occurs at α = 0.6685, the periodic oscillations
arising from this point has almost a zero amplitude, not shown. This may also be happening in time simulation
given in Fig. 2. Similar behaviour can also be obtained for numerical continuation of Q 1 and Q 2 with respect to
536
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

Fig. 5. Bifurcation diagram of predator Q 1 and predator Q 2 populations with respect to α parameter for ω = 1. The number of eigenvalues
that are positive or have positive real parts (Uµ ) are represented with different colours along the steady states of the system, see Table 2.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

β conversion efficacy. As seen a supercritical Hopf bifurcation is obtained for Q 2 free state for ω = 0.1 (c), and
this become a subcritical Hopf bifurcation for Q 1 free state with ω = 1 (d), for which stability of the periodic orbit
switches from unstable to stable via periodic doubling bifurcation. Similar to Fig. 4(b), periodic orbits associated
with the Hopf point along the coexisting state evolve in the same direction and approach towards zero.
In Fig. 5, we consider that the past influence occurs for a smaller time interval, e.g. ω = 1. Comparing Fig. 4(a)
with ω = 0.1 and Fig. 5(a) with ω = 1, there is no substantial change in the dynamics of predator Q 1 with respect
to α, with the exception of two periodic doubling bifurcation connecting stable orbits to a small unstable region.
However, there has been a considerable change in the dynamics of predator Q 2 from ω = 0.1 in Fig. 4(b) to ω = 1
in Fig. 5(b), particularly for Q 1 free and coexisting equilibria. The bifurcation diagrams can be performed for other
parameters, yet we only focus on the dynamics with respect to α and β. Therefore, considering a memory effect in
the dynamics of predator Q 2 has lead to a substantial change in its dynamical stability and existence.

5. Summary
This paper focuses on a memory dependent two predators–one prey mathematical model with competition
between predators. The contribution of this work is theoretical in nature and modelling-oriented, having a descriptive
purpose. As in many papers in theoretical ecology, we here concentrate on a generic prey–predator model and try to
understand how fading memory in a system comprising competing predators would affect the dynamics. The system
parameters are chosen as intuitively meaningful hypothetical population parameters. Considering fading memory in
both predators is also a natural choice (leading to same dynamics for both predators), yet the aim of our analysis is
to determine the impact of fading memory in competing predators and to justify our analytical results for a generic
mode to observe the role of memory. Thus, exponentially fading memory is considered for one of the predators,
implying that the density of one predator, e.g. Q 1 , only depends on the current density of prey. However, the density
of memory dependent predator, e.g. Q 2 , depends not only on the density of prey at present but also on the density
of prey at recent past. In this way, we investigate the direct influence of memory term in the system dynamics. Here,
the variable R denotes a continuous probability density function, the role of which is to weight the moments of the
past quantities of prey. Thus, this can be achieved by replacing prey variable in the evolution of second predator
with R.
The control parameter is chosen as ω, whose small values imply the existence of past influence for larger time
interval and whose large values imply the existence of past influence for a smaller time interval. Thus ω−1 can
be considered as the measure for past influence. We have performed our simulations for ω = 0.1 and ω = 1 and
537
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

obtained that, in the absence of competition, small ω induces instability of prey G and predator Q 2 . Here dynamics
of predator Q 1 tends to extinction. On the other hand, large ω induces stability of all three species, implying that
stability and persistence of all species are possible when past influence exists for smaller time interval. Our findings
show that one of the predators tends to undergo extinction in general. This can also be seen from Fig. 2, where
predator Q 1 dies out without depending on the memory parameter. This also confirms the ideas presented in the
introduction that one predator determines its territory and other predators are not tolerable in this territory. Besides,
although it is not shown here for brevity, predator Q 1 and Q 2 tends to go extinction with high and low rates of
ω respectively. Furthermore, bifurcation diagrams of all variables with respect to half saturation constant α and
conversion strength β are compared for ω = 0.1 and ω = 1, and significant differences are obtained, particularly
for the stability of states E 2 and E 3 where one predator (Q 1 or Q 2 ) is absent. Here it is obtained that the value
of memory parameter may affect the stability, may change the nature of Hopf bifurcations from subcritical to
supercritical, or may induce the existence or non-existence of periodic doubling bifurcation.
Thus, the weakening memory substantially influences the dynamics and stability of the model studied in this
paper. Since fading memory term incorporated in the model has a non-local nature, compared with classical local
prey–predator models, more complex and realistic results are possible to obtain. One potential future direction is
the inclusion of local and non-local delay terms for the predation terms as an additional degree of realism [17].
The time evolutions in this paper are performed using MATLAB’s ode45 package with a default tolerance.
The single parameter continuation diagrams are plotted using DDE-BIFTOOL package which is associated with
a collection of MATLAB routines and allows an effective performance of the local stability analysis around each
equilibria [11,12]. Here continuation of the Hopf bifurcations and their periodic orbits can also be determined under
two parameter variation, e.g. α and β. This way, it is also possible to compute all eigenvalues of corresponding
characteristic polynomial, where various branches are marked with different colours. For details of the numerical
scheme we refer the reader to [11,12].

CRediT authorship contribution statement


Zeynep Yılmaz: Conceptualization, Formal analysis, Investigation, Validation, Methodology, Software, Visual-
ization, Writing – review & editing. Selahattin Maden: Project management, Formal analysis, Validation, Concep-
tualization, Methodology, Project administration, Visualization, Writing – review & editing. Aytül Gökçe: Project
management, Conceptualization, Formal analysis, Investigation, Methodology, Project administration, Software,
Validation, Visualization, Writing – review & editing.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could
have appeared to influence the work reported in this paper.

References
[1] J. Alebraheem, Y.A. Hasan, Dynamics of a two predator–one prey system, Comput. Appl. Math. 33 (3) (2014) 767–780.
[2] M.R. Ali, S. Raut, S. Sarkar, U. Ghosh, Unraveling the combined actions of a holling type III predator–prey model incorporating allee
response and memory effects, Comput. Math. Methods 3 (2) (2021) e1130.
[3] R. Arditi, J. Abillon, J.V. Da Silva, A predator-prey model with satiation and intraspecific competition, Ecol. Model. 5 (3) (1978)
173–191.
[4] R. Arribas, J.C. Touchon, I. Gomez-Mestre, Predation and competition differentially affect the interactions and trophic niches of a
neotropical amphibian guild, Front. Ecol. Evol. 6 (2018) 28.
[5] E.N. Bodine, A.E. Yust, Predator-prey dynamics with intraspecific competition and an allee effect in the predator population, Lett.
Biomath. 4 (1) (2017) 23–38.
[6] M. Cavani, M. Farkas, Bifurcations in a predator-prey model with memory and diffusion. I: Andronov-Hopf bifurcation, Acta Math.
Hungar. 63 (3) (1994) 213–229.
[7] P. Chesson, Species competition and predation, in: R.A. Meyers (Ed.), Encyclopedia of Sustainability Science and Technology, Springer
New York, New York, NY, ISBN: 978-1-4939-2493-6, 2015, pp. 1–28, http://dx.doi.org/10.1007/978-1-4939-2493-6_579-3.
[8] P. Chesson, J.J. Kuang, The interaction between predation and competition, Nature 456 (7219) (2008) 235–238.
[9] S. Djilali, B. Ghanbari, Dynamical behavior of two predators–one prey model with generalized functional response and time-fractional
derivative, Adv. Difference Equ. 2021 (1) (2021) 1–19.
[10] B. Dubey, R. Upadhyay, Persistence and extinction of one-prey and two-predators system, Nonlinear Anal. Model. Control 9 (4) (2004)
307–329.
538
Z. Yılmaz, S. Maden and A. Gökçe Mathematics and Computers in Simulation 202 (2022) 526–539

[11] K. Engelborghs, T. Luzyanina, D. Roose, Numerical bifurcation analysis of delay differential equations using DDE-BIFTOOL, ACM
Trans. Math. Softw. 28 (1) (2002) 1–21.
[12] K. Engelborghs, T. Luzyanina, G. Samaey, et al., DDE-BIFTOOL: a matlab package for bifurcation analysis of delay differential
equations, TW Report 305 (2000) 1–36.
[13] M. Farkas, Stable oscillations in a predator-prey model with time lag, J. Math. Anal. Appl. 102 (1) (1984) 175–188.
[14] A. Farkas, M. Farkas, G. Szabó, Multiparameter bifurcation diagrams in predator-prey models with time lag, J. Math. Biol. 26 (1)
(1988) 93–103.
[15] J.D. Ferreira, C.A.T. Salazar, P.C. Tabares, Weak allee effect in a predator–prey model involving memory with a hump, Nonlinear
Anal. RWA 14 (1) (2013) 536–548.
[16] U. Ghosh, S. Pal, M. Banerjee, Memory effect on Bazykin’s prey-predator model: Stability and bifurcation analysis, Chaos Solitons
Fractals 143 (2021) 110531.
[17] A. Gökçe, A mathematical study for chaotic dynamics of dissolved oxygen-phytoplankton interactions under environmental driving
factors and time lag, Chaos Solitons Fractals 151 (2021) 111268.
[18] A. Gökçe, The influence of past in a population system involving intraspecific competition and allee effect, Eur. Phys. J. Plus 137 (2)
(2022) 200.
[19] C.S. Holling, Some characteristics of simple types of predation and parasitism1, Can. Entomol. 91 (7) (1959) 385–398.
[20] S. Hsu, S. Hubbell, P. Waltman, A contribution to the theory of competing predators, Ecol. Monograph 48 (3) (1978) 337–349.
[21] J.M. Jeschke, C. Laforsch, P. Diel, J.G. Diller, M. Horstmann, R. Tollrian, Predation, in: Reference Module in Earth Systems and
Environmental Sciences, Elsevier, ISBN: 978-0-12-409548-9, 2021, http://dx.doi.org/10.1016/B978-0-12-819166-8.00016-5.
[22] Q. Liu, D. Jiang, Stationary distribution and extinction of a stochastic one-prey two-predator model with Holling type II functional
response, Stoch. Anal. Appl. 37 (3) (2019) 321–345.
[23] M. Liu, P.S. Mandal, Dynamical behavior of a one-prey two-predator model with random perturbations, Commun. Nonlinear Sci.
Numer. Simul. 28 (1–3) (2015) 123–137.
[24] J. Llibre, D. Xiao, Global dynamics of a Lotka–Volterra model with two predators competing for one prey, SIAM J. Appl. Math. 74
(2) (2014) 434–453.
[25] N. MacDonald, Time delay in prey-predator models—II. Bifurcation theory, Math. Biosci. 33 (3–4) (1977) 227–234.
[26] N. MacDonald, Time delay in prey-predator models—II. Bifurcation theory, Math. Biosci. 33 (3–4) (1977) 227–234.
[27] W. Mbava, J. Mugisha, J.W. Gonsalves, Prey, predator and super-predator model with disease in the super-predator, Appl. Math.
Comput. 297 (2017) 92–114.
[28] D. Mukherjee, Effect of fear on two predator-one prey model in deterministic and fluctuating environment, Math. Appl. Sci. Eng. 2
(1) (2021) 55–71.
[29] G.M. Palamara, J.A. Capitán, D. Alonso, The stochastic nature of functional responses, Entropy 23 (5) (2021) 575.
[30] L. Persson, Behavioral response to predators reverses the outcome of competition between prey species, Behav. Ecol. Sociobiol. 28
(2) (1991) 101–105.
[31] B. Sahoo, S. Poria, Dynamics of predator–prey system with fading memory, Appl. Math. Comput. 347 (2019) 319–333.
[32] S. Sarwardi, P.K. Mandal, S. Ray, Dynamical behaviour of a two-predator model with prey refuge, J. Biol. Phys. 39 (4) (2013)
701–722.
[33] D. Savitri, A. Suryanto, W.M. Kusumawinahyu, et al., A dynamics behaviour of two predators and one prey interaction with competition
between predators, in: IOP Conference Series: Materials Science and Engineering, Vol. 546, no. 5, IOP Publishing, 2019, 052069.
[34] I. Sayekti, M. Malik, D. Aldila, One-prey two-predator model with prey harvesting in a food chain interaction, in: AIP Conference
Proceedings, Vol. 1862, no. 1, AIP Publishing LLC, 2017, 030124.

539

You might also like