1 Introduction
Continuous -time population models derived by using differential equations are the earliest mathematical models that were used to understand population dynamics [Reference Lotka24, Reference Voltera38]. Such models are appropriate for studying the dynamical behaviour of species that have overlapping generations [Reference Hale19]. One of these types of models that has been used to investigate the relationship between prey and predator species is the prey–predator model. This model has been extensively studied in the realm of population dynamics, with numerous studies focusing on stability analysis of steady state, permanence, extinction and others [Reference Allen3, Reference Murray29, Reference Olsen and Degn30, Reference Seno36, Reference Wiggins39]. Furthermore, prey–predator models are often used in not only ecology and biology, but also other research areas, including engineering and economics, to analyse a variety of issues [Reference Brauer and Castillo-Chavez8, Reference Freedman17, Reference Mark26, Reference Strogaz37].
In the context of ecological systems, Hopf bifurcation is a particularly salient concept, as it provides a theoretical framework for understanding the transition from stable equilibria to periodic oscillations [Reference Hale and Koçak20, Reference Hassard, Kazarinoff and Wan21, Reference Perko33]. These oscillations are a common occurrence in ecological systems and frequently model real-world ecological phenomena, such as predator–prey oscillations and the cyclic dynamics of interacting species (see, for example, [Reference Çelik and Deǧerli11, Reference Dash and Khajanchi14, Reference Feng, Liu, Sun and Jiang16, Reference Merdan28, Reference Ou, Xu, Cui, Pang, Liu, Shen, Baber, Farman and Ahmad31] and references therein).
In prey–predator models, the population size of each species is generally influenced by a variety of factors such as species competition, parasitic infections, harvesting, etc. Recently, particular attention has been given to population models incorporating the Allee effect, which have been deemed more congruent with reality than those lacking this effect [Reference Abbasi and Samreen1, Reference Pal, Mandal, Guin and Saha32, Reference Pervenecki, Bewick, Otto, Fagan and Li34, Reference Xing, He and Li40, Reference Xue41]. So then, why should the Allee effect be included in mathematical models or population models? Let us consider in the prey–predator model where both species have exponential growth. In this situation, even if the population density is low at the beginning, if there is a favourable environment (adequate food, suitable environmental conditions, etc.) for the population to live in, it will begin to reproduce over time. However, Warder Clyde Allee has shown that such a situation is not always true. As a result of an experiment that has been conducted with goldfish, he has concluded that individuals in a population benefit from their own species, and as a natural consequence of this situation, individuals are negatively affected by a fall in population density [Reference Allee2]. Additionally, thanks to the observations of Allee, the positive correlation between population density and individual fitness is accepted as the definition of the Allee effect [Reference Courchamp, Berec and Gascoigne12].
There are several different mechanisms that prove positive density dependence, and mate-finding is one of them. The mate-finding Allee effect is a well-established ecological phenomenon that occurs in populations where individuals face difficulty in finding mates at low densities, leading to reduced reproductive success. To illustrate, in prey–predator dynamics, if the prey population has difficulty finding a suitable mate to reproduce at low population density, this may also adversely affect not only the prey population, but also the predator population, whose only food source is the prey population.
The Allee effect can have a significant impact on prey–predator models, because it can lead to unexpected stabilizing and destabilizing effects on population dynamics. In this article, to assess the potential significance of Allee effects in population dynamics, we consider the following continuous-time and Leslie-type prey–predator system where the prey population has a mate-finding Allee effect:

Here,
$N(t)$
and
$P(t)$
present prey density and predator density at time t, respectively, and
$\beta>0$
denotes the strength of the Allee effect on prey population. All other parameters, namely
$r_1,r_2,\epsilon $
and
$\theta $
are also positive constants. In system (1.1),
$r_1N(t)$
presents the growth of the prey population, where
$r_1$
is the intrinsic growth rate. It is hypothesized that the observed growth is constrained by the functional response term,
$\epsilon P(t)$
, in conjunction with the mate-finding Allee effect, whose functional form is
$\alpha (N(t))=N(t)/(N(t)+\beta ) $
. Here, decreasing
$\beta $
diminishes the strength of the Allee effect. To put it another way,
$\alpha (N(t)) \rightarrow 1$
, while
$\beta \rightarrow 0$
. If we look closely,
$\alpha (N(t))$
is a strictly increasing function of N
$(\alpha '(N(t))=\beta /(N(t)+\beta )^2>0)$
. It means that as the population density increases, the probability of finding a mate (and hence, the birth rate) increases, whereas the Allee effect decreases. If the number of individuals is very large, we cannot talk about the Allee effect since individuals can definitely find mates, that is,
$\lim _{N(t) \rightarrow \infty }N(t)/(N(t)+\beta )=1$
.
Unlike prey, predator populations grow logistically as a function of growth rate
$r_2$
. This growth depends on the number of prey per predator. In other words, the prey density directly affects the carrying capacity of the predator’s environment with the Leslie–Gower term,
$P(t)/N(t)$
, which measures the deprivation of a predator population due to its available food, that is, the number of available prey for the predator. In the second equation,
$\theta $
represents the strength of density-dependent interactions and the term
$-\theta P^2(t)/N(t)$
represents a form of intraspecific competition among predators. In circumstances where prey populations are limited, the repercussions of predation become pronounced, as predators find themselves contending more vigorously for scarce resources.
System (1.1) is inspired by the following continuous-time prey–predator model:

The nonlinear dynamics of system (1.2) has been analysed by several researchers: Zhou et al. determined the stability conditions of the positive equilibrium point of the system and analysed the impact of the Allee effect on its dynamics [Reference Zhou, Liu and Wang42]. A Hopf bifurcation analysis of this system involving a single discrete time delay is studied by Celik [Reference Çelik10]. In addition to these studies, Karaoglu and Merdan gave a detailed Hopf bifurcation analysis of system (1.2) involving two discrete time delays [Reference Karaoglu and Merdan22]. Furthermore, the complex dynamical behaviour of the discrete-time counterpart of system (1.2) is explored by Baydemir et al. [Reference Baydemir, Merdan, Karaoglu and Sucu7]. In addition to all these studies, although there are similar studies in the literature, the impact of the Allee effect on the dynamics of the prey and predator populations will be given in detail in this article to help researchers develop more accurate and effective models of ecological systems, which are essential for predicting and managing population dynamics in conservation biology and management. Additionally, the Hopf bifurcation analysis, which refers to a phenomenon that occurs in dynamical systems where a small variation in a parameter value leads to a qualitative change in the behaviour of the system and can provide insights into the sustainability and stability of prey–predator systems, will be subjected to a more accurate mathematical approach.
Within the relevant system, some input parameters might only marginally impact the uncertainty of the output values, whilst some input parameters, such as the Allee effect, may have a strong effect. In many areas of model examination, sensitivity analysis can be used to comprehend the impact of the various input values on the model behaviour [Reference Asghari, Topol, Markert and Merodio4]. Whereas local sensitivity analysis typically focuses on calculating the derivative of the model response with respect to the model input parameters, global sensitivity analysis, in particular, attempts to allocate variation in the output variable(s) to variation in the model input parameters [Reference Geng, Lou, Liu, Qi, Liu and Chen18].
The FAST (Fourier amplitude sensitivity test) procedure (proposed by Cukier et al. [Reference Cukier, Fortuin, Shuler, Petschek and Schaibly13]) is ideally suited to the task of determining the global sensitivity of nonlinear models subjected to variations of arbitrary size in the system parameters [Reference Asghari, Topol, Markert and Merodio5, Reference Büyükkahraman, Sabine, Kojouharov, Chen-Charpentier, McMahan and Liao9, Reference Maaboudalla and Atalla25, Reference Pianosi, Sarrazin and Wagener35]. In contrast to local sensitivity methods, which evaluate the effect of small changes in parameters around an equilibrium point, the FAST method provides a comprehensive view of the global influence of parameters over the entire parameter space. This property renders it especially well suited for the analysis of ecological models, in which parameter uncertainty is a prevalent occurrence.
In the present investigation, the FAST approach was selected, because it enables the systematic quantification of the individual contribution of each parameter. The motivation for this choice stems from the recognition that a comprehensive understanding is imperative for ecological systems, such as the Leslie-type prey–predator model with an Allee effect, where multiple factors interact in complex and often unpredictable ways. Furthermore, in the literature, no previous studies have used the FAST method in the context of this particular prey–predator model. This application thus constitutes a novel contribution within the existing body of research. Therefore, in this study, after the basic sensitivity analysis is demonstrated mathematically, the FAST method is performed to show the individual contribution of each parameter to the changes in density of prey and predator populations over time [Reference Baydemir and Merdan6].
This paper is organized as follows. In Section 2, we first determine the steady state of system (1.1) and analyse its local stability conditions through the Allee effect. Section 3 discusses the prerequisites for the periodic solutions to be seen while the Allee effect changes in the system. Further, the existence of the Hopf bifurcation is investigated and to obtain the formula of the Lyapunov coefficient for determining the direction of the Hopf bifurcation, the stability and period of the periodic solution are derived by normal form theory. The sensitivity analysis is presented in Section 4. In Section 5, we perform some numerical simulations using MATLAB to support and extend our theoretical results. The final section, that is Section 6, is devoted to the concluding remarks and discussions.
2 Stability analysis of system (1.1)
In this section, the future behaviour of prey and predator populations under the Allee effect will be investigated. For that, we first determine the equilibrium points of system (1.1). It is clear that
$(\bar {N},\bar {P})=(0,0)$
is an equilibrium point of this system, but there is nothing biologically interesting about that equilibrium point. The model has been designed with the assumption that the prey population is the only source of food for the predator population, and the survival of the predator is dependent on the number of prey in the environment. Then, system (1.1) has an equilibrium point of the form

Notice that this equilibrium point is positive so that the Allee coefficient stated in (2.1) can be taken as
$ \beta =k{\theta r_1}/{\epsilon r_2} $
for
$k \in \mathbb {R} $
and
$k\in (0,1) $
. As a result,

is the unique positive equilibrium point of system (1.1) for a constant value of k. If we look closely, as the value of k increases, the Allee constant
$\beta $
increases, so that the equilibrium point diminishes due to the Allee effect. This is in line with the biological interpretation of the decrease in the prey population that cannot find a mate owing to the Allee effect and, accordingly, the decrease in the predator population. The Jacobean matrix of system (1.1) at its steady state is

where
$ tr(J)=r_1 k (1-k)-r_2 $
and
$ \det (J)=r_1r_2(1-k)^2 $
. Let us define
$ a_1:= -tr (J) $
and
$ a_2:= \det (J) $
. So, we can write the characteristic polynomial associated with the Jacobean matrix as follows:

Then, the equilibrium point
$ (\bar {N},\bar {P}) $
is locally asymptotically stable when
$a_1>0$
(that is,
$tr({J})<0$
) and
$a_2>0$
(that is,
$\det ({J})>0$
). Obviously,
$a_2 =\det (J) =r_1r_2(1-k)^2$
is always positive. Therefore, let us determine the conditions required for the coefficient
$a_1$
to be positive.
The equation
$a_1=r_2-r_1k(1-k)=r_1k^2-r_1k+r_2$
is a second-order polynomial according to the parameter k. Let us express this equation by
$ a_1(k) $
. When
$r^2_1-4r_1r_2\geq 0$
, the equation
$a_1(k)$
has two real roots of the form:

We are dealing with two situations here. If
$r_1>4r_2$
, then
$r^2_1-4r_1r_2> 0$
and then
$k_{1,2}>0$
is satisfied since
$ \sqrt {r^2_1-4r_1r_2} < \lvert r_1 \rvert $
. Moreover, since

are written as above,
$k_1<k_2 $
is true and
$a_1(k)>0 $
under the condition that
$k \in (0,k_1) \cup (k_2, 1) $
. However, if
$r_1=4r_2 $
, then
$k_{1,2}={1}/{2}$
and
$a_1(k)=0$
. Much as the characteristic polynomial
$ F(\lambda ) $
has a pair of purely imaginary roots when
$r_1=4r_2$
, that is,
$a_1(k)=0 $
,
$a_2>0 $
and
$\lambda _{1,2}=\mp i \sqrt {a_2}$
, the stability condition of the equilibrium point does not change as k changes. Consequently, with the next theorem, we can express the required conditions for the stable point of system (1.1) to be a sink.
Theorem 2.1. The equilibrium point
$ (\bar {N},\bar {P}) $
of system (1.1) is a sink; in other words, it is locally asymptotically stable when the following condition is satisfied:
(C1)
$k \in (0,k_1) \cup (k_2, 1)$
when
$r_1>4r_2 $
.
Here, the coefficients
$k_1 $
and
$k_2 $
are formulated as given in (2.4).
3 Bifurcation analysis
We have shown in Section 2 that the stability structure of the equilibrium point
$ (\bar {N},\bar {P}) $
of system (1.1) changes as parameter k and, accordingly, the Allee constant,
$\beta $
, changes. It may therefore be said that there are qualitative changes, namely bifurcations, in the dynamics of system (1.1) under the influence of Allee. Here, a change in the stability condition of the equilibrium point means that the negative real part of the eigenvalues is zero at the same value, which is called the bifurcation parameter, and the real part of the same eigenvalues is positive immediately following the bifurcation parameter. This occurs if the system has zero or a pair of purely imaginary eigenvalues, and Hopf bifurcation occurs in the system when there is a pair of purely imaginary eigenvalues. So, let us investigate under which conditions system (1.1) has Hopf bifurcation under the Allee effect.
Lemma 3.1. Assume that the characteristic polynomial
$ F(\lambda )= \lambda ^2+a_1\lambda +a_2$
, associated to the Jacobian matrix at the equilibrium point
$ (\bar {N},\bar {P}) $
of system (1.1), is a second-order polynomial with real coefficients. So, for this polynomial to have a pair of purely imaginary eigenvalues, a necessary and sufficient condition is that
$a_1=0 $
and
$a_2>0 $
.
Proof. Let us consider that
$ \lambda = i w $
is a root of
$ F(\lambda ) $
for
$ w \in \mathbb {R} $
and
$ w>0 $
. Since
$ \lambda = i w $
satisfies the equation of
$F(\lambda )=0$
, we obtain the following:

Now, suppose that
$a_1=0 $
and
$a_2>0 $
. Then, from
$ F(\lambda ) = \lambda ^2+a_2 $
, we get the roots as
$ \lambda _{1,2}= \mp i \sqrt {a_2} $
. Thus, we have concluded the proof.
Theorem 3.2. There is a pair of purely imaginary eigenvalues of system (1.1) if the characteristic polynomial
$ F(\lambda ) $
satisfies the following conditions:
(C2)
$k=k_1$
or
$k=k_2$
, where
$k_1$
and
$k_2$
are stated in Theorem 2.1.
Proof. First, if
$k=k_1$
or
$k=k_2$
, then
$a_1=0$
. However, from Lemma 3.1, polynomial
$F(\lambda )$
has a pair of purely imaginary eigenvalues since
$a_2>0$
.
Now, suppose that the characteristic polynomial
$ F(\lambda ) $
has a pair of complex conjugate eigenvalues in the form of
$\lambda _{1,2}(k)=m(k)\mp in(k)$
, where
$m(k), n(k) \in \mathbb {R}$
and
$n(k) \neq 0$
. Here, parameter k is taken as a bifurcation parameter and to prove the existence of Hopf bifurcation arising from the stable point
$(\bar {N},\bar {P})$
of system (1.1), we need to show that
$m(k_c)=0$
and
$n(k_c)=n_0>0$
are satisfied by choosing
$k_c $
as a critical bifurcation parameter.
We next assume that
$ \lambda _1= m(k) + in(k) $
, and take
$ m(k)=m $
and
$ n(k)=n $
to make representation easier. Since
$ \lambda _1 $
is a root of
$ F(\lambda ) $
,

Hence, the equalities
$ m^2-n^2+a_1m+a_2=0 $
and
$ 2mn+a_1n =0 $
must be satisfied simultaneously. In a similar way, for
$ \lambda _2= m- in $
,

Again, we have the following equations to provide the last equality:


If we substitute (3.1) into (3.2), we obtain the above equation,

Obviously, we are dealing with three situations here.
-
• If
$a_2<n^2 $ , then
$ m^2<0 $ and there is a contradiction to the fact that
$ m \in \mathbb {R} $ .
-
• If
$a_2=n^2 $ , then
$ m=0 $ and then
$ \lambda _{1,2}=\mp i \sqrt {a_2} $ .
-
• If
$a_2>n^2 $ , then
$ m=\mp \sqrt {a_2-n^2} \neq 0$ . However, in this situation, one cannot find any critical value
$k_c $ that solves the equation
$ m(k_c)=0 $ .
Let us now bear in mind (3.2). This equation is a second-order polynomial with real coefficients depending on the parameter
$ m $
. Moreover, when
$a_1^2-4(a_2-n^2)\geq 0 $
, (3.2) has two real roots as stated below:

If we pay attention,
$a_1,a_2 $
and
$ n $
are functions of parameter k, and
$a_2 $
is always positive. So, in the next cases, we will look at the scenarios where the coefficient
$a_1 $
is zero, negative or positive.
Case 1: If
$a_1=0 $
, then from (3.1), we have
$ m=0 $
and putting this value into (3.2),
$ n^2=a_2 \Rightarrow n= \mp \sqrt {a_2}$
is found. Thus, the polynomial
$ F(\lambda ) $
has a pair of purely imaginary eigenvalues in the form of
$ \lambda _{1,2}=\mp i \sqrt {a_2} $
.
Case 2: Again consider (3.1); if
$a_1>0 $
, then
$ m=- {a_1}/{2} <0 $
must be satisfied. Moreover, let us consider the roots
$ m_1 $
and
$ m_2 $
separately for (3.2), and to do this, first assume that
$a_1^2-4(a_2-n^2)>0 $
holds.
-
• If
$a_2=n^2 $ , then
$ m_1=0 $ and this is inconsistent with
$ m<0 $ .
-
• If
$a_2<n^2 $ , then
$ m_1>0 $ and again it goes counter to
$ m<0 $ .
-
• If
$a_2>n^2 $ , then
$ m_1<0 $ is found.
As a result, if
$ m$
is taken as
$ m=m_1 $
when
$a_2>n^2 $
under the condition that
$a_1>0 $
and
$a_1^2-4(a_2-n^2)>0 $
, then the polynomial
$ F(\lambda ) $
has a complex conjugate root whose real part is negative. Therefore, the equilibrium point is locally asymptotically stable. Similarly:
-
•
$ m_2=-a_1<0 $ holds when
$a_2=n^2 $ ;
-
•
$ m_2<0 $ whether
$a_2<n^2 $ or
$a_2>n^2 $ .
Then, for the values
$a_1>0 $
and
$a_1^2-4(a_2-n^2)>0 $
, if
$ m=m_2 $
is chosen, the polynomial
$ F(\lambda ) $
has a complex conjugate root with negative real part and this means that the equilibrium point of system (1.1) is a sink.
Case 3: If
$a_1<0 $
, then
$ m=-{a_1}/{2}>0 $
from (3.1). Also,
$a_1^2-4(a_2-n^2)>0 $
must be held for the real value of
$ m $
. Under these assumptions:
-
• if
$a_2=n^2 $ , then
$ m_1=-a_1>0 $ ;
-
• if
$a_2<n^2 $ or
$a_2>n^2 $ , then
$ m_1>0 $ .
Therefore, if
$ m=m_1 $
is taken when
$a_1<0 $
or
$a_1^2-4(a_2-n^2)>0 $
, the equilibrium point becomes unstable since the real part of the polynomial
$ F(\lambda ) $
has a positive complex conjugate root. Now, let us investigate the sign of
$ m_2 $
under the same conditions:
-
•
$ m_2=0 $ if
$a_2=n^2 $ holds, but this contradicts with
$ m>0 $ ;
-
•
$ m_2<0 $ is found for
$a_2<n^2 $ and again there is a contradiction to the hypothesis
$ m>0 $ ;
-
•
$ m_2>0 $ if
$a_2>n^2 $ is taken.
As a result of the above findings,
$ F(\lambda ) $
has a complex conjugate root with the positive real part if we choose
$ m=m_2 $
when
$a_2>n^2 $
under the assumption that
$a_1<0 $
and
$a_1^2-4(a_2-n^2)>0 $
. This indicates that the equilibrium point is unstable.
Notice that
$a_1 $
, that is
$a_1(k) $
, determines the critical bifurcation value. Therefore, the solution of
$a_1(k)=0$
will give the critical bifurcation value. Let us denote this critical bifurcation value as
$k_c $
. When
$r_1>4r_2 $
, if
$k_c=k_1 $
or
$k_c=k_2 $
are taken (here,
$k_1$
and
$k_2 $
are the roots of the equation
$a_1(k)=0$
are given as (2.3)), then
$ m(k_c)=0 $
,
$ n(k_c)=\sqrt {a_2} $
, and consequently
$ \lambda _{1,2}= \mp i \sqrt {a_2} $
is found. This completes the proof.
The characteristic polynomial
$ F(\lambda ) $
has a pair of purely imaginary eigenvalues under the condition
$ (C2) $
given in Theorem 3.2. Also, when
$ 0<k<k_1 $
or
$k_2<k<1 $
(that is, the condition
$ (C1) $
of Theorem 2.1 holds), the equilibrium point of system (1.1) is locally asymptotically stable (see Case 2). However, if
$k_1<k<k_2 $
, then the equilibrium point of system (1.1) is unstable (see Case 3).
Now, let us give the following results for the complex eigenvalues we achieved under the above assumptions.
Lemma 3.3. Assume that the coefficients
$a_1 $
and
$ m $
are a function of the parameter k. In that case, if
$ {d(a_1)}/{dk} |_{k_c} \neq 0$
, then
${dm}/{dk}|_{k_c} \neq 0 $
.
Proof. The proof is completed by taking the derivative of both sides of the equation
$a_1 = -2m$
. Indeed, for

since
$k_c\neq {1}/{2}.$
If we take a close look,
$r_1=4r_2 $
must be held to satisfy the equation

However,
$r_1>4r_2 $
from the definition of
$k_c $
. So,
$k_c $
cannot be equal to
$ 1/2 $
.
Lemma 3.4.
$Re ({d(\lambda (k))}/{dk}) |_{k_c} \neq 0$
.
Proof.
$Re ({d(\lambda (k))}/{dk}) |_{k_c} = {dm}/{dk} |_{k_c} \neq 0$
since
$ {dm}/{dk} |_{k_c} \neq 0 $
from Lemma 3.3.
3.1 Hopf bifurcation
In the former section, we have concluded that the positive equilibrium point
$ (\bar {N},\bar {P})=((1-k){\theta r_1}/{\epsilon r_2},(1-k){r_1}/{\epsilon }) $
of system (1.1) is locally asymptotically stable by Theorem 2.1. We also know from Theorem 3.2 that the characteristic polynomial
$ F(\lambda ) $
has a pair of purely imaginary eigenvalues if
$k=k_1 $
or
$k=k_2 $
. Hence, Hopf bifurcation occurs when
$r_1>4r_2 $
in this case.
We now show the existence of the Hopf bifurcation by using the centre manifold theorem and bifurcation theory given in [Reference Kuznetsov23] (see the Appendix). First, let us shift the equilibrium point
$ (\bar {N},\bar {P}) $
of system (1.1) to the origin for simplicity and, for this purpose, denote the right-hand sides of the equations in system (1.1) as follows:

After that, using Taylor expansion of f and g defined by (3.3) at
$(\bar {N},\bar {P})$
and following transformations

we obtain the following system with equilibrium point
$(0, 0)$
:

where

in which H.O.T. represents the higher order terms in
$f_1$
and
$f_2$
. Notice that not only the equilibrium point but also the critical bifurcation value
$ \tilde {k} $
(when
$k=k_c $
) has shifted to the origin in system (3.4). Therefore, we can rewrite the functions
$f_1$
and
$f_2$
as follows:

Here, the coefficients
$ s_i $
for
$ i=1, \ldots , 9 $
are stated:

Now, assume that
$k_c=k_1 $
, then
$ \lambda _{1,2}(k_c)=m(k_c)\mp in(k_c), $
where
$ m(k_c)=0 $
and
$ n(k_c)=\sqrt {a_2} $
based on Theorem 3.2. In addition, Lemma 3.4 is also satisfied and the conditions of (H.1) and (H.2) of the Hopf bifurcation theorem are held. So, let us reduce system (3.4) to normal form so that we can apply the Hopf bifurcation theorem and calculate the Lyapunov coefficient
$ c_1(k_c) $
, which gives the directional analysis of this bifurcation.
The Jacobian matrix of system (3.4) at the critical bifurcation value is

We now construct a matrix T whose columns consist of real and imaginary parts of an eigenvector corresponding to
$ \lambda _1=-in(k_c) $
or
$ \lambda _1=-in $
in short as follows:

Using the following transformation:

system (3.4) is reduced to the following system of equations:

where

in which the coefficients are given as follows:

Again, considering system (3.5) and applying the complex transformation
$ Z(t)=X(t)+iY(t) $
, we obtain

where

Here, the coefficients are calculated as

Now, we reduce system (3.6) to its normal form in the following two steps.
Step 1. Since
$ \lambda (k_c)=in(k_c) $
in (3.6), both
$ \lambda =in $
and
$ \bar {\lambda }=-in $
are not equal to zero. Hence, for a sufficiently small value of k, using an invertible parameter-dependent complex coordinate transformation

and taking

(3.6) is obtained as the following equation without quadratic terms:

where the coefficients are given as

Step 2. Again, for sufficiently small values of k, using an invertible parameter- dependent change of complex coordinate transformation as follows:

and choosing

cubic terms disappear except
$ W^2(t)\bar {W}(t) $
and (3.8) can transform into the following equation:

Notice that if
$ h_{21}={\tau _{21}}/({\lambda +\bar {\lambda }}) $
is chosen, then the leading coefficient of the term
$ W^2(t)\bar {W}(t) $
can be zero. However, this option cannot be made because
$ \lambda (k_c)=\sqrt {a_2} $
and
$\bar {\lambda }(k_c)=-\sqrt {a_2} $
would be
$ \lambda +\bar {\lambda }=0 $
. Thus, since
$ \lambda +\bar {\lambda }=0$
in (3.9), the coefficient of the term
$ W^2(t)\bar {W}(t) $
is found as
${\tau _{21}}/{2}$
and taking this coefficient as
$ c_1 $
, one obtains the next equation:

where the coefficient
$ c_1=c_1(k_c) $
has the following form using (3.7) and (3.8) with
$ \lambda (k_c)=in(k_c)=in $
and
$ \bar {\lambda }(k_c)=-in(k_c)=-in $
:

As it is known, the mate-finding Allee effect is particularly important for species that are sensitive to low population densities. Heretofore, the stability structure of system (1.1) was examined depending on the k parameter and, consequently, the Allee constant
$\beta $
value, while the other parameters remained constant. Does the Allee effect, that is, the parameter
$\beta $
, play a more effective role in the dynamics of system (1.1) than the other parameters? To what extent is system (1.1) sensitive to the parameters? In the former section, we attempted to answer these questions.
4 Sensitivity analysis
The parameters of the equations that comprise the model determine how species interact with one another and vary over time in the prey–predator system. So, we concentrate on a sensitivity analysis to find out how resilient the model is to parameter values that are associated with the
$k_1 $
and
$k_2 $
, at which the dynamics of the system changes. Additionally, to do this, we calculate the sensitivity indices according to the sensitivity approach given in [Reference Martcheva27] and derive an analytical expression for the sensitivity index (elasticity) of
$k_1 $
and
$k_2 $
with respect to each parameter according to the explicit formula of these values as follows:

Since
$k_1,k_2 \in (0,1) $
and
$r_1,r_2>0 $
, we have the following corollary.
Corollary 4.1. The elasticities
$ E^{r_1}_{k_1} $
and
$ E^{r_2}_{k_2} $
are always negative, while the elasticities
$ E^{r_2}_{k_1} $
and
$ E^{r_1}_{k_2} $
are always positive.
According to the above result, since
$E^{r_1}_{k_1} < 0$
, the increase in
$r_1$
, which expresses the growth rate of the prey population, will reduce
$k_1$
, that is, the Allee effect. It is true that a higher
$r_1$
would result in a higher prey population birth rate and, consequently, a higher rate of mate finding, which would mitigate the Allee effect. However, since
$E^{r_2}_{k_1}> 0$
, an increase in
$r_2$
, which represents the growth rate of the predator population, will result in a decrease in the number of prey and a lower chance of finding a mate. For other sensitivity formulae, similar inferences can be made.
5 Numerical examples and simulations
In this section, the theoretical results found in Sections 2, 3 and 4 will be supported by numerical examples, and the impact of the Allee effect on the dynamics of the given system will be demonstrated. To perform numerical calculations, we used the MATLAB ODE package.
To begin with, we consider the next example to exhibit the local stability of system (1.1). We use the parameter values inspired by [Reference Baydemir, Merdan, Karaoglu and Sucu7], which are
$r_1=0.5$
,
$r_2=0.08$
,
$\epsilon =0.03$
and
$\theta =0.05$
.
Notice that the equilibrium point of system (1.1) with constant
$r_1,r_2,\epsilon $
and
$ \theta $
parameter values is

and it varies depending on the k parameter. In addition, as the value of k increases (correspondingly, as the Allee constant increases), the equilibrium point decreases. In Table 1, the Allee coefficient
$\beta $
and equilibrium point
$ (\bar {N},\bar {P}) $
of system (1.1) are given, corresponding to some k values.
Table 1 The Allee constant and equilibrium point of system (1.1) according to the chosen k values.

Since
$r_1>4r_2 $
,
$k_1=0.2 $
and
$k_2=0.8 $
, if we take
$k=0.02 $
, then condition
$ (C1) $
holds in Theorem 2.1. Therefore, based on the fixed parameter values above, the unique equilibrium point
$ (\bar {N}, \bar {P})=(10.21,16.33) $
is locally asymptotically stable. We have chosen the initial conditions as
$N(0)=10$
and
$P(0)=15$
for the numerical simulations of system (1.1). Figure 1(a) shows that if k is taken as less than
$k_c =k_1$
, then the equilibrium point is locally asymptotically stable. As seen in Figure 1(b), the equilibrium point loses its stability and periodic solutions occur when k is equal to the critical bifurcation value
$k_1 $
. However, Figure 1(c) illustrates that the equilibrium point is unstable if we select
$k=0.3 $
as larger than
$k_1 $
.

Figure 1 Density-time graphs of prey (left), density-time graphs of predator (middle) and phase portrait of system (1.1) (right) for the different values of k: (a)
$k=0.02 $
; (b)
$k=0.2 $
and (c)
$k=0.3 $
. The initial conditions are
$ N(0)=10 $
and
$ P(0)=15 $
for each simulation.
In Table 1, one can easily see that the equilibrium point is calculated as
$ (\bar {N},\bar {P})=(8.3333, 13.3333) $
and the Allee constant is calculated as
$ \beta =2.0833 $
for the crucial bifurcation value of
$k_1=0.2 $
. At this critical bifurcation value, system (1.1) has a pair of purely imaginary roots as
$ \lambda _{1,2}=m(k_1)\mp i n(k_1)=\mp 0.16 i $
. Figure 2(a) and (b) depict the decline in the equilibrium number of prey and predators as the Allee constant rises, and “H” is the point that designates the Allee value at which Hopf bifurcation appears. In addition to these, in accordance with the formulae in the Hopf bifurcation theorem expressed in the Appendix, the coefficients giving the stability, direction and period of the Hopf bifurcation are calculated as follows:

Here,
$ m'(k_1)>0$
and
$\alpha (k_1)>0 $
. So, the direction of the Hopf bifurcation is supercritical, and bifurcating periodic solutions are stable since
$ Re(c_1(k_1))<0 $
. Moreover, the period of the limit cycles increases since
$ \tilde {T}>0 $
. The aforementioned Hopf bifurcation diagram is shown in Figure 2(c). This diagram was created by using the MATCONT package, which is a graphical MATLAB software, for which further details can be found in [Reference Dhooge, Govaerts and Kuznetsov15].

Figure 2 Panels (a) and (b) represent prey and predator equilibria as the Allee coefficient
$\beta $
increases, and point H presents the Hopf bifurcation point. (c) Limit cycles that undergo Hopf bifurcation in
$(\beta , N, P) $
space.
Now, let us examine Figure 3, which includes the solution graphs of system (1.1) and the phase portrait of this system for the values of k that are greater than the first critical bifurcation value
$k_1 $
at the same parameter values and initial conditions. Figure 3 presents that the Allee effect changes the stability of the equilibrium point according to each specific k value from unstable to stable. One can also see from Figure 3(a) and (b) that equilibrium point
$ (\bar {N},\bar {P})=(3.1250, 5.0000) $
is unstable for
$k=0.7<k_2=0.8 $
, and periodic solutions arise in a neighbourhood of the equilibrium point
$ (\bar {N},\bar {P})=(2.0833, 3.3333) $
when k passes through another critical bifurcation value
$k_c=k_2=0.8 $
.

Figure 3 The trajectory of prey and predator density of system (1.1) versus time and the phase portrait of prey density versus predator density, respectively, with the initial conditions
$ N(0)=10 $
and
$ P(0)=15 $
when (a)
$k=0.7 $
, (b)
$k=0.8 $
and (c)
$k=0.9 $
.
Finally, the equilibrium point
$ (\bar {N},\bar {P})=(1.0417, 1.6667) $
is stable for the larger values of the bifurcation values and this case is simulated for
$k=0.9>k_2 $
in Figure 3(c). Therefore, Hopf bifurcation occurs at the critical bifurcation value of
$k_2 $
. Indeed, for
$k=k_2=0.8 $
, the eigenvalues of system (1.1) have the form of
$ \lambda _{1,2}=m(k_2)\mp i n(k_2)=\mp 0.04 i $
.
Again, similar calculations are done to determine the stability, direction and period of the Hopf bifurcation for the value of
$k_2 $
:

The fourth case in the Hopf bifurcation theorem exists because
$ m'(k_2)<0 $
and
$\alpha (k_2)>0$
. Since
$Re(c_1(k_2))>0$
, unstable periodic solutions are seen. Also,
$ \tilde {T} -35.4544<0 $
, and as can be seen from Figure 3, periodic solutions are unstable (that is, subcritical) with a decreasing period.
Note that the Allee effect has a significant impact on the dynamical behaviour of system (1.1). Since critical values
$k_1$
and
$k_2$
(accordingly Allee constant) are functions of the parameters
$r_1$
and
$r_2$
, it is expected that densities of populations are sensitive to the changing of these values. We provide numerical evidence for this notion in Figure 4, which also serves as a numerical example to bolster Corollary 4.1. It should be noted that in the provided sensitivity analysis example, the other parameter values are kept constant while the impact of a single parameter value change on the system solution is examined.

Figure 4 (a,c) Trajectories of prey and predator densities versus time when
$r_1=0.5 $
(square),
$r_1=0.55 $
(diamond) and
$r_1=0.65 $
(star), while
$r_2, \epsilon , \theta $
are constant values in the case of
$k=k_1 $
. (b,d) Trajectories of prey and predator densities versus time where
$r_2=0.08$
,
$0.09$
,
$0.12$
are marked with square, diamond and star, respectively, while other parameters are fixed in the same case. The initial conditions are
$N(0)=10$
and
$P(0)=15$
for these simulations.
The question is, which parameter, when all other parameter values fluctuate over time, has the greatest effect on the populations of prey and predators? The FAST (Fourier amplitude sensitivity test) approach of the MATLAB SAFE (Sensitivity analysis for everyone) program [Reference Pianosi, Sarrazin and Wagener35] was used to provide an answer to this query. Here, the sensitivity index of the model parameters was calculated based on the ranges given in Tables 2 and 3 for the variation of parameter values, and the results are displayed in Figures 5 and 6. As can be seen from Figure 5, when the parameter values change so that they remain in the stability region of system (1.1), the most effective parameter for the system is the
$r_1$
parameter, which shows the growth rate of the prey. Changes in the
$r_1$
and
$r_2$
parameters also impact the
$\beta $
parameter since
$\beta =k{\theta r_1}/{\epsilon r_2}$
. Consequently, it seems inevitable that the dynamics of system (1.1) would be more susceptible to variations in these three parameters.
Table 2 Upper and lower limits for parameter values in system (1.1).

Table 3 Upper and lower limits for parameter values in system (1.1) where prey population is low.

Another issue that occurs in real life is the potential for the prey population to not increase enough once the system is stable. Table 3 shows example values where the prey growth rate
$r_1$
is reduced and its range is restricted.
For these values, the sensitivity indices are shown in Figure 6. Here, the Allee effect
$\beta $
has a greater sensitivity index. From a biological perspective, the Allee effect drives a species towards extinction if the prey population grows slowly or does not have a sufficient growth rate, which results in difficulties for the population in finding a mate and reproducing. In this case, the dynamics of the system is more sensitive to changes in the Allee parameter compared with other parameters. However, we would anticipate a decrease in the Allee effect if the prey population is sufficiently large, the reproductive rate is high and finding a partner is not difficult. This is in line with Figure 5.
6 Conclusions
The study of bifurcations in dynamical systems has been a central topic in mathematics for many decades. One such bifurcation is the Hopf bifurcation, which occurs when a system undergoes a qualitative change in behaviour as a parameter is varied. In ecological systems, the Leslie-type prey–predator model is a popular choice for modelling the interactions between prey and predator populations. Research on population models includes changes in population dynamics caused by various mechanisms or interactions, and the Allee effect is one of these mechanisms. Incorporating the Allee effect into prey–predator models allows for a more accurate representation of the actual ecological dynamics involved, and understanding the Allee effect can be particularly important when considering species that may be particularly vulnerable to small population sizes.
In this paper, we have considered a continuous-time prey–predator model with the Leslie type in which the growth of prey is affected by the mate-finding Allee effect. First, we have examined the local stabilities of the equilibrium point of system (1.1) and we have put some reasonable restrictions on parameters so that the positive equilibrium point of the model is locally asymptotically stable. After that, Hopf bifurcation analysis has been carried out choosing the parameter k (accordingly Allee effect) as the bifurcation parameter, and we have found that there are two critical bifurcation values denoted as
$k_1 $
and
$k_2 $
at which Hopf bifurcation arises. More precisely, we have shown the existence of this bifurcation at these critical values by using normal form theory in [Reference Kuznetsov23] and we have obtained Lyapunov coefficient formulae that determine the stability of the bifurcation stated in the Hopf bifurcation theorem given in the Appendix.
The Hopf bifurcation analysis showed that the Allee effect significantly changes the stability of the system. The biological implication is that as the Allee effect strengthens, prey populations at low densities may have difficulty finding mates, leading to lower reproduction rates. This, in turn, reduces the amount of prey available to predators and has an impact on their population growth. What is more, the appearance of supercritical and subcritical Hopf bifurcations can be interpreted as population boom/bust cycles observed in real ecosystems. In more detail, the identification of critical bifurcation values (
$k_1 $
and
$k_2 $
) highlights thresholds that can be seen as critical population densities or environmental conditions at which species adapt to their unstable environments for survival. For instance, as the Allee effect increases, prey populations may exhibit behavioural or physiological alterations aimed at enhancing their capacity to locate mates, thus ensuring their reproduction and survival under adverse conditions. Correspondingly, predator populations could adapt their hunting strategies or modify their dietary habits in response to the diminished availability of prey, thereby contributing to the long-term stability of the ecosystem.
Sensitivity analysis subsequently provided insight into how sensitive it is to the growth rate of prey and predator populations, which are the parameters of the critical bifurcation value at which system dynamics change. Furthermore, we have provided a numerical example to support the consistency of analytical results and also to exhibit the Allee effect of system (1.1) in Section 5. This example system (1.1) and its numerical simulations (see Figures 1, 2 and 3) lay weight on the Allee effect on the dynamics of the continuous-time Leslie-type prey–predator model.
Sensitivity indices are numerical values that quantify the influence of input variables on the output of a system; these indices indicate how changes in each input variable individually contribute to the overall output variability. Values between
$ 0 $
and
$ 1 $
represent intermediate levels of sensitivity, where the corresponding input variables have partial contributions to the output variability. If the sensitivity index of an input parameter is close to zero, it can be taken as a constant value since the change in that parameter has a negligible effect on the system solution. However, since the parameter value with the highest sensitivity index can have the potential to drastically alter the dynamics of the system, it would be wise to take this parameter value into account as a control (or bifurcation) parameter. From a biological standpoint, this is advantageous for the management and conservation of ecosystems.
Our analysis has indicated that the Allee effect plays a crucial role in determining the stability and dynamics of the system, particularly in determining the critical threshold for the Hopf bifurcation since small changes in the Allee parameter have a significant effect on the occurrence and stability of the bifurcation with low population densities. The sensitivity indices come into play as well; Figures 5 and 6 provide further evidence of this.
As a result, the findings that we have concluded have a number of important implications for future practice in understanding population behaviours and the natural survival strategies of endangered populations and in providing valuable insight into the interplay between population dynamics, ecology, conservation and control theory. In addition, the numerical examples of sensitivity analysis may help guide data gathering and increase the potential to figure out important processes for ecosystem functioning.
Appendix
Hopf bifurcation theorem: Assume that the following two-dimensional one-parameter system:

has a pair of complex conjugate eigenvalues
$\lambda (k)=m (k) \mp i n (k)$
at the fixed point
$(0,0)$
, where
$k=k_c$
, and the following conditions be satisfied:
(H.1)
$m(k_c)=0$
and
$m' (k_c)\neq 0;$
(H.2)
$n (k_c):=n_0>0$
.
Then, there is a neighbourhood of the origin in which a unique closed curve bifurcates from origin as k passes through
$k_c $
if
$a(k_c)=-{Re(c_1(k_c))}/{\mu ' (k_c)} \neq 0$
. Here,

Remark A.1 The sign of
$m'(k_c)$
and
$a(k_c)$
determines the stability and direction of the appearance of the invariant curve in a generic system exhibiting the Hopf bifurcation.
-
• If
$m'(k_c)>0$ and
$a(k_c)>0$ , then prior to the critical bifurcation value
$k_c $ , the equilibrium point is stable; following this value, it is unstable. Moreover, periodic solutions appear after the critical bifurcation value. Since
$Re(c_1(k_c))<0$ , the bifurcating periodic solutions are stable. Hence, supercritical Hopf bifurcation takes place at the critical bifurcation value.
-
• If
$m'(k_c)>0$ and
$a(k_c)<0$ , then the equilibrium point is stable up to the crucial bifurcation value and unstable after this
$k_c $ . Before the critical bifurcation value, periodic solutions are seen. The bifurcating periodic solutions are unstable because
$Re(c_1(k_c))>0$ . Thereby, there is a subcritical Hopf bifurcation.
-
• If
$ m'(k_c)<0 $ and
$\alpha (k_c)<0$ , then the equilibrium point is unstable before
$k_c $ , but stable after this value, and before the bifurcation value, periodic solutions arise. Periodic solutions are stable due to the fact that
$Re(c_1(k_c))<0$ . As a result, supercritical Hopf bifurcation occurs at the crucial bifurcation value
$k_c $ .
-
• If
$ m'(k_c)<0 $ and
$\alpha (k_c)>0$ , then before the critical bifurcation value, the equilibrium point is unstable; however, it becomes stable after this value. After the critical bifurcation value, periodic solutions emerge and the bifurcating periodic solutions are unstable since
$Re(c_1(k_c))>0$ . So, there is a subcritical Hopf bifurcation.
In addition, the period of periodic solutions is calculated with the
$ \tilde {T} $
defined as

-
• If
$\tilde {T}>0$ , the period of periodic solutions increases as
$k_c $ bifurcation value increases.
-
• If
$\tilde {T} <0$ , the period of periodic solutions decreases as
$k_c $ bifurcation value increases.