1. Introduction
Differential equations with fractional orders are widely used in science, and this extensive use has recently sparked much research interest. The so-called fractional-order integral or derivative operators are the main producers of fractional-order differential equations. Due to the nonlocal nature of their integral and derivative operators, fractional-order operators are more effective than other classical deterministic operators at predicting the future state of the system, because they depend not only on the current state, but also on all of its past states. The most popular and widely used fractional-order operators in applications of fractional-order differential equations are the Riemann–Liouville, Caputo and Grünwald–Letnikov kinds [Reference Hoan, Akinlar, Inc, Gómez-Aguilar, Chu and Almohsen28]. Now, let us give the definition of fractional-order integration and differentiation [Reference Podlubny44, Reference Samko, Kilbas and Marichev46].
Definition 1.1. The fractional integral (or the Riemann Liouville integral) of order
$\beta \in \mathbb {R}^+$
of the function
$f(t), \hspace {0.5em} t>0$
, is defined by

The fractional derivative of order
$\alpha \in (n-1,n)$
is defined by the following two (nonequivalent) ways.
-
(1) Riemann–Liouville fractional derivative: take the fractional integral of order
$(n-\alpha )$ , and then take the nth derivative as follows:
$$ \begin{align*} D^\alpha f(t)= D^{n} I_{a}^{(n-\alpha)} f(t), \quad D =\frac{d}{dt}, \quad n=1,2,3,\ldots. \end{align*} $$
-
(2) Caputo-fractional derivative: take the nth derivative, and then take a fractional integral of order
$(n-\alpha )$
$$ \begin{align*} D^\alpha f(t)=I_{a}^{(n-\alpha)} D^{n} f(t). \end{align*} $$
Even though it is more restrictive than the Riemann–Liouville, we consider the fractional derivative provided by Caputo in this study [Reference Caputo11]. This is because this definition is more suitable for problems consisting of fractional differential equations with initial conditions. The following result contains the main properties from fractional calculus [Reference Rihan45].
Remark 1.2. Let
$L^1=L^1[a,b]$
be the class of Lebesgue integrable functions on
$[a,b], a<b<\infty $
,
$\beta , \gamma \in \mathbb {R^+}$
and
$\alpha \in (0,1)$
. Then:
-
(1) if
$I_{a}^{\beta }:L^1\rightarrow L^1$ and
$f(x) \in L^1$ , then
$I_{a}^{\gamma }I_{a}^{\beta }f(x)=I_{a}^{\gamma +\beta } f(x)$ ;
-
(2)
$\lim _{\beta \to n }I_{a}^{\beta }f(x){\kern-1pt}={\kern-2pt} I_{a}^{n}f(x)$ uniformly on
$[a,b], n{\kern-1pt}={\kern-1pt}1,2,3,\ldots ,$ where
$I_{a}^{1}f(t){\kern-1pt}={\kern-2.5pt}\int _0^t f(s) ds$ ;
-
(3)
$\lim _{\beta \to 0 }I_{a}^{\beta }f(t)=f(t)$ weakly;
-
(4) if
$f(t)$ is absolutely continuous on
$[a,b]$ , then
$\lim _{\alpha \to 1} D^\alpha f(t)={df(t)}/{dt}$ .
In applied mathematics, the process of converting continuous models defined by differential equations into their discrete equivalent is known as discretization. Analysing the dynamics of the discretization is crucial for understanding how well the discrete model approximates the original continuous system. Discretization is often applied to systems to perform numerical simulations or to design algorithms that run on digital platforms. As the step size approaches zero, the discrete system converges to the original continuous system. In this limit, the discrete-time model approaches the behaviour of the differential equations, ensuring that any discrepancies between the two systems vanish. In the literature, there are a large number of nonlinear fractional differential equations without an analytical solution, so they need to be solved with the help of some numerical and discretization techniques such as the Adomian decomposition method [Reference Afreen and Raheem1, Reference Guo22], differential transform method [Reference Nazari and Shahmorad40, Reference Wang and Wang51], Euler method [Reference Kim and Parovik34], extrapolation method [Reference Diethelm and Walz18], Grünwald–Letnikov method [Reference Ihtisham, Nigar and Nisar30] and variational iteration method [Reference Wu55, Reference Wu and Baleanu56].
In the last decade, we see a new type of discretization technique for fractional equations, called piecewise discretization, which is constructed with the help of the piecewise constant arguments [Reference Agarwal, El-Sayed and Salman2, Reference El Raheem and Salman19, Reference El-Sayed and Salman20, Reference Gürcan, Kaya and Kartal23, Reference Kartal and Gürcan32, Reference Kaya, Kartal and Gürcan33, Reference Matouk, Elsadany, Ahmed and Agiza39, Reference Panigoro, Rayungsari and Suryanto42, Reference Xin, Peng and Guerrini57, Reference Xin, Peng, Kwon and Liu58, Reference Yousef, Semmar and Al Nasr60, Reference Yousef, Semmar and Al Nasr61]. The systematic study of problems involving piecewise constant arguments began in the early 1980s with the work of Shah and Wiener [Reference Shah and Wiener50], who introduced the term “differential equations with piecewise constant argument” (DEPCA). A comprehensive source on this class of equations is provided in [Reference Wiener52]. Busenberg and Cooke [Reference Busenberg and Cooke10] were pioneers in applying such deviating arguments to mathematical models, specifically in the study of vertically transmitted diseases, by reducing their analysis to discrete equations. The main source for DEPCA theory are the papers [Reference Cooke, Wiener, Busenberg and Martelli17, Reference Wiener52]. In 1991, Györi was the first who used the piecewise constant argument for approximation [Reference Györi24]. After this work, there has been a significant interest in the usage of the piecewise constant argument for a numerical approach [Reference Cooke and Györi16, Reference Györi and Hartung25–Reference Györi, Hartung and Turi27]. In the late 2010s, Akhmet [Reference Akhmet4] introduced the equation

where
$\gamma (t)$
is a piecewise constant argument of generalized type, that is, given
$(t_{k})_{k \in \mathbb {Z}}$
and
$(\zeta _{k})_{k \in \mathbb {Z}}$
such that
$t_{k}<t_{k+1}$
for all
$ k \in \mathbb {Z}$
with
$\lim _{ k\rightarrow \pm \infty }t_{k}=\pm \infty $
and
$t_{k} \leq \zeta _{k}\leq t_{k+1}$
, if
$t \in I_{k}=[t_{k}, t_{k+1})$
, then
$\gamma (t)=\zeta _{k}$
. These type of equations are called “differential equations with piecewise constant argument of generalized” (DEPCAG) type. They have continuous solutions, even when
$\gamma (t)$
is not, producing a discrete equation. Several aspects of this equation have garnered much interest, see [Reference Akhmet3, Reference Akhmet, Aruğaslan and Yılmaz5, Reference Chiu13–Reference Chiu and Pinto15] and the references therein.
In this paper, we propose to use a piecewise discretization method. For this aim, let us first explain the usage of the method for any fractional-order differential equation. Agarwal et al. [Reference Agarwal, El-Sayed and Salman2] considered the equation

as a discretized version of the equation

where r is the discretization parameter. Let
$t \in [0,r)$
, then
${t}/{r} \in [0,1) $
. So from (1.1),

Thus,

Let
$t \in [r,2r)$
, so
${t}/{r} \in [1,2) $
. Thus,

whose solution is

Let
$t \in [2r,3r)$
, then
${t}/{r} \in [2,3) $
. Thus, we obtain

So,

Repeating the process, when
$t \in [nr,(n+1)r)$
,

which yields

Letting
$t \rightarrow (n+1)r$
in the above, the corresponding difference equation is obtained as

with
$x(nr)=y_n$
.
Prey and predator live in the same ecosystem, and both of them have their own, but mostly the same, biotic factors, such as speed, stealth and camouflage (to hide while pursuing the prey or to hide from the predator). They also possess good senses of smell, sight and hearing (to locate the prey or to detect the predator), and poison (to kill the prey or to spray when approached or bitten), among other traits. In mathematical biology, these factors are usually considered as the main subjects that affect the model. However, recently, there has been a significant increase in the number of studies exploring the impact of abiotic effects. These include the influence of predator-induced fear [Reference Barman, Roy and Alam6, Reference Barman, Roy, Alrabaiah, Panja, Mondal and Alam8, Reference Liu, Cai, Tan and Chen36, Reference Liu and Jiang37], climate change [Reference Sekerci48, Reference Sekerci49, Reference Wilmers, Post and Hastings54], seasonal variations [Reference Charles, Makinde and Kunǵaro12, Reference Oksanen, Oksanen, Vuorinen, Wolf, Mäkynen, Olofsson, Ripple, Virtanen and Utsi41, Reference Sauve, Taylor and Barraquand47] and wind [Reference Barman, Roy and Alam7, Reference Panja43]. These elements can of course be increased.
Now, let us turn the wheels to our problem. Barman et al. [Reference Barman, Roy and Alam7] investigated the effect of wind in a prey–predator model. They constructed two types of functional responses to describe the whole dynamics of the considered system under the fact that wind may either decrease or increase the predation rate of predators. They showed positivity and boundedness of the solutions of the systems and examined the stability of the equilibrium points. Also for particular cases, it is demonstrated that wind can change the stability of the equilibrium point through Hopf bifurcation. Furthermore, considering the fact that wind flow cannot be constant for a time period, they developed the discussed system and investigated this system with the help of numerical simulations. In their work, they considered

with the initial conditions

The variables and all positive parameters in this model are defined in Table 1.
Table 1 Description of the variables and parameters in (1.3).

Our aim is to investigate the fractional-order discrete version of this problem to see the effect of both the fractional order and the discretization parameter in the occurrence of flip and Neimark–Sacker bifurcation.
In recent years, there has been a considerable amount of work on flip and Neimark–Sacker bifurcations: in [Reference Kangalgil and Işık31], a discrete-time predator–prey system is considered, and the chaos control is investigated. The effect of prey refuge proportional to the predator in a discrete-time prey–predator model is studied in [Reference Mahapatra, Santra and Bonyah38]. The authors consider the dynamical behaviour of a discrete-time prey–predator system with Leslie type in [Reference Baydemir, Merdan, Karaoglu and Sucu9], and in [Reference Yíldíz, Bilazeroglu and Merdan59], a discrete Lotka–Volterra-type predator–prey system with refuge effect is analysed. The reader is also referred to the references therein.
Before stating our main problem, let us first apply nondimensionalization to (1.3) to use one of the most important benefits: reduction of the numbers of the parameters. For this aim, let us define the variables

Then, the derivatives become

Substituting (1.5) and (1.6) in the initial value problem (1.3)–(1.4) gives the nondimensionalized form

with

where

Now, let us consider this initial value problem with fractional derivatives and apply the piecewise discretization process. For
$\alpha \in (0,1)$
and
$h>0$
as the discretization parameter, it can be written as

with

The equations in (1.7) are both in the form of (1.1), so the corresponding difference equation, from (1.2), is

where
$\rho =h^{\alpha }$
,

with the initial conditions

The presentation in this paper is organized as follows. Section 1 is devoted to the introduction of the stated problem. Preliminaries needed for the paper are given in Section 2. Stability and bifurcation analysis are investigated in Sections 3 and 4, respectively. Numerical examples are stated in Section 5 and, finally, Section 6 presents the conclusion.
2. Preliminaries
In this section, we recall some definitions and lemmas that will be useful in the later sections.
Consider the difference system

The Jacobian of this system at any fixed point
$(\bar {M},\bar {N})$
is

and the corresponding characteristic equation is

Definition 2.1 [Reference Kangalgil and Işık31].
Let
$\lambda _1$
and
$\lambda _2$
be the characteristic roots of (2.1). A fixed point
$(\bar {M},\bar {N})$
is called:
-
(i) sink if
$\lvert \lambda _1 \rvert <1$ and
$\lvert \lambda _2 \rvert <1$ , and it is locally asymptotically stable;
-
(ii) source if
$\lvert \lambda _1 \rvert>1$ and
$\lvert \lambda _2 \rvert>1$ , and it is locally unstable;
-
(iii) saddle if
$\lvert \lambda _1 \rvert>1$ and
$\lvert \lambda _2 \rvert <1$ or
$\lvert \lambda _1 \rvert <1$ and
$\lvert \lambda _2 \rvert>1$ ;
-
(iv) nonhyperbolic if either
$\lvert \lambda _1 \rvert =1$ or
$\lvert \lambda _2 \rvert =1$ .
Lemma 2.2 [Reference Hu, Teng and Zhang29].
Let
$F(\lambda ) =\lambda ^2 + B\lambda + C,$
where B and C are constants. Suppose
$F(1)> 0$
, and
$\lambda _{1}$
and
$\lambda _{2}$
are two roots of
$F(\lambda ) = 0.$
Then:
-
(i)
$\lvert \lambda _{1} \rvert < 1$ and
$\lvert \lambda _{2} \rvert < 1$ if and only if
$F(-1)>0$ and
$C<1$ ;
-
(ii)
$\lambda _{1}=-1$ and
$\lvert \lambda _{2} \rvert \neq 1$ if and only if
$F(-1)=0$ , and
$B\neq 0,2$ ;
-
(iii)
$\lvert \lambda _{1} \rvert> 1$ and
$\lvert \lambda _{2} \rvert < 1$ if and only if
$F(-1)<0$ ;
-
(iv)
$\lvert \lambda _{1} \rvert> 1$ and
$\lvert \lambda _{2} \rvert> 1$ if and only if
$F(-1)>0$ and
$C>1$ ;
-
(v)
$\lambda _{1} $ and
$ \lambda _{2} $ are the conjugate complex roots and
$\lvert \lambda _{1} \rvert = \lvert \lambda _{2} \rvert =1$ if and only if
${B^2-4C<0}$ and
$C = 1.$
Definition 2.3 [Reference Kuznetsov, Kuznetsov and Kuznetsov35].
The bifurcation associated with the appearance of
$\mu _{1}=-1$
is called a flip (or period-doubling) bifurcation, where
$\mu _{1}$
is an eigenvalue of the nonhyperbolic fixed point of the the system

with smooth f.
Theorem 2.4 [Reference Kuznetsov, Kuznetsov and Kuznetsov35].
Suppose that a one-dimensional system

with smooth f, has at
$\alpha = 0$
the fixed point
$ x_0= 0$
and let
$\mu =f_x (0,0)=-1.$
Assume that the following nondegeneracy conditions are satisfied:
-
(i)
$\tfrac {1}{2}( f_{xx}(0,0))^{2} + \tfrac {1}{3} f_{xxx}(0,0)\neq 0$ ;
-
(ii)
$f_{x \alpha }(0,0)\neq 0$ .
Then, there is a flip bifurcation in the system.
Definition 2.5 [Reference Kuznetsov, Kuznetsov and Kuznetsov35].
The bifurcation corresponding to the presence of
$\mu _{1,2}=e^{\pm i \theta _{0}}$
,
$0<\theta _{0}<\pi $
, is called a Neimark–Sacker bifurcation, where
$\mu _{1,2}$
are the eigenvalues of the nonhyperbolic fixed point of the system (2.2).
Theorem 2.6 [Reference Kuznetsov, Kuznetsov and Kuznetsov35].
Suppose that the system

with a parameter
$\delta $
, has a pair of complex conjugate eigenvalues
$\lambda _{1,2}=r(\delta )e^{\pm i \theta (\delta )}$
at the fixed point
$(0,0)$
, where
$\delta _c$
is the critical parameter value,
$r(\delta _c)=1$
and
$\theta (\delta _c)=\theta _0$
. Assume that the following conditions are satisfied:
-
(i)
${d(r(\delta ))}/{d \delta } \rvert _{\delta =\delta _c} \neq 0$ ;
-
(ii)
$e^{i k \theta _0} \neq 1$ for
$k=1,2,3,4$ and
$a(\delta _c)= Re (e^{i\theta _0}c_{1}(\delta _c)) \neq 0$ .
Then, there is a neighbourhood of origin in which a unique closed curve bifurcates from the origin as
$\delta $
passes through the critical value
$\delta _c$
. It should be mentioned here that the sign of
$a(\delta _c)$
determines the direction of the appearance of the invariant curve in a generic system exhibiting the Neimark–Sacker bifurcation: if
$a(\delta _c)<0$
, then there is a supercritical Neimark–Sacker bifurcation which is stable; if
$a(\delta _c)>0$
, then there is a subcritical Neimark–Sacker bifurcation which is unstable.
3. Stability analysis
The difference system (1.8) has three equilibrium points given by

with

When simulating biological systems, positivity ensures that the model can make meaningful predictions about real-world behaviour. Negative values in simulations would lead to outcomes that are not interpretable or applicable to the biological phenomenon being studied. To ensure the positivity of the point
$E_{*}=( M_{*}, N_{*} )$
, it is assumed that

Now, let us investigate the stability of each point.
3.1. Stability of
$E_0$
Theorem 3.1.
$E_0=(0,0)$
is:
-
(i) saddle if
$0<\rho < ({2\Gamma (\alpha +1)})/{b_{2}}$ ;
-
(ii) source if
$\rho> ({2\Gamma (\alpha +1)})/{b_{2}}$ ,
where
$\rho =h^\alpha $
, h is the discretization parameter and
$\alpha $
is the order of the equation.
Proof. The Jacobian matrix at
$E_{0}$
is

The characteristic polynomial corresponding to (3.4) is calculated as

whose eigenvalues are

Hence, we have

3.2. Stability of
$E_1$
Theorem 3.2. Assume that
$b_1<(a_1+a_2) b_2.$
Then,
$E_{1}=(1,0)$
is:
-
(i) sink if
$0<\rho <\min \{\delta _{1}, \delta _{2}\}$ ;
-
(ii) saddle if
$\min \{\delta _{1}, \delta _{2}\}<\rho <\max \{\delta _{1}, \delta _{2}\}$ ;
-
(iii) source if
$\rho>\max \{\delta _{1}, \delta _{2}\},$
where

Proof. The Jacobian matrix at
$E_{1}$
is obtained as

which has the characteristic polynomial

Hence, the following eigenvalues are found:

Since
$b_1<(a_1+a_2) b_2$
,

and

3.3. Stability of
$E_*$
Theorem 3.3. Assume that
$ b_1(a_1 - a_2) + a_2 b_2(a_1 + a_2)> 0$
and (3.3) is true. Then, the positive equilibrium point
$E_{*}$
of system (1.8) is:
-
(i) sink fixed point if one of the following conditions holds:
$$ \begin{align*} 0<\rho<\rho_1\quad \textit{when}\ \xi_2> 4\xi_1;\\ 0<\rho<\frac{1}{\xi_1}\quad \textit{when} \ \xi_2\leq 4\xi_1, \end{align*} $$
-
(ii) source fixed point if one of the following conditions holds:
$$ \begin{align*} \rho>\rho_2\quad \textit{when} \ \xi_2> 4\xi_1;\\ \rho>\frac{1}{\xi_1}\quad \textit{when} \ \xi_2\leq 4\xi_1, \end{align*} $$
-
(iii) saddle fixed point if the following condition holds:
$$ \begin{align*} \rho_1<\rho<\rho_2\quad \textit{when} \ \xi_2> 4\xi_1, \end{align*} $$
where
$\rho =h^\alpha $
and

Proof. The Jacobian matrix at
$E_{*}$
, given in (3.1) and (3.2), is

whose characteristic equation is

where

To apply Lemma 2.2(i), consider the left-hand side of (3.7) as a function of
$\lambda $
, that is,

or

with
$h^\alpha =\rho $
, and
$\xi _1\: and \: \xi _2 $
are given in (3.5). Equation (3.8) is called the characteristic polynomial corresponding to the positive equilibrium point
$E_{*}$
.
Taking
$\lambda =1$
and
$\lambda =-1$
above, and considering
$\Phi $
and
$ \Psi $
,
$F(1)$
and
$F(-1)$
are calculated as

From (3.3), it is clear that
$F(1)>0$
. To ensure the positivity of
$F(-1)$
, it is written as

where
$\xi _1, \: \xi _2, \: \rho _1, \: \rho _2$
are given in (3.5). So, the roots of
$F(-1)$
are obtained as

First, we assume that
$\xi _2>4\xi _1$
, then
$\rho _1$
and
$\rho _2$
are real and distinct. Under conditions (3.3) and
$a_1>a_2$
, it is obtained that
$\rho _1$
and
$\rho _2$
are positive with
$0<\rho <\rho _1$
or
$\rho>\rho _2$
, and this implies that
$F(-1)>0$
. Moreover, when
$0<\rho <\rho _1,$
we get
$\rho <\rho _1<{1}/{\xi _1}<\rho _2$
and
$C=\xi _1\xi _2\rho ^2-\xi _2\rho +1<1.$
When
$\rho>\rho _2,$
it is found
$\rho>{1}/{\xi _1}$
and then
$C>1$
.
Second, if
$\:\xi _2=4\xi _1$
and
$\:\xi _2<4\xi _1$
, then
$F(-1)$
has real roots
$ \rho _{1}= \rho _{2}={1}/{\xi _1}$
and does not have real roots, respectively, and for each case,
$F(-1)$
becomes positive. Also, if
$0<\rho <{1}/{\xi _1},$
then
$C<1$
or if
$\rho>{1}/{\xi _1},$
then
$C>1.$
So, from Definition 2.1 and Lemma 2.2, the proof is complete.
4. Bifurcation analysis
This section is devoted to the investigation of the flip and Neimark–Sacker bifurcation analysis of the equilibrium point
$E_*$
given in (3.1)–(3.2). Here, we choose
$\rho $
as a bifurcation parameter.
4.1. Flip bifurcation
In this part, we study flip bifurcation for the system (1.8). We prove that under certain conditions, the flip bifurcation arises from the positive fixed point
$E_*$
.
Theorem 4.1. If the following conditions are satisfied by the characteristic polynomial (3.8), then one may conclude that there is a flip bifurcation at
$E_*$
. When
$\xi _2> 4\xi _1$
, then
$F(-1)$
has the roots
$ \rho =\rho _1$
and
$\rho =\rho _2$
such that

where
$\rho _1$
and
$\rho _2$
are given in (3.5).
Proof. From Lemma 2.2(ii), it is known that if

then there may be flip bifurcation at the equilibrium point
$E_*$
. From Theorem 3.3, it is known that when
$\xi _2> 4\xi _1$
, then
$F(-1)$
has two real roots, which means that if we take
$\rho =\rho _1$
or
$\rho _2$
, then we get
$F(-1)=0$
. Moreover, considering
$ \rho \neq {2}/{\xi _2}$
and
$ \rho \neq {4}/{\xi _2}$
leads us to
$\rho \xi _2-2 \neq \{0,2\},$
and thus the proof is complete.
However, let us take

then the characteristic roots of (3.9) are found as
$\lambda _1=-1$
and
$\lambda _2=3-\rho _1 \xi _2$
, with
$\lvert \lambda _{2} \rvert \neq 1$
.
We will now focus on the analytic construction of the flip bifurcation, or, to put it another way, use Theorem 2.4 to establish the following theorem (Theorem 4.2) at the point
$E_*$
. However, to apply Theorem 2.4, we must first apply to our system (1.8) the centre manifold theorem [Reference Baydemir, Merdan, Karaoglu and Sucu9, Reference Kuznetsov, Kuznetsov and Kuznetsov35, Reference Wiggins and Golubitsky53].
Theorem 4.2. If the following conditions are satisfied, then there is a flip bifurcation for the system (1.8):
-
(i)
$\xi _2>4 \xi _1$ ;
-
(ii)
$2 m_1+2( h_1 m_2 + m_5 ) \neq 0$ ;
-
(iii)
$m_3 \neq 0$ .
Furthermore, if
$2 m_1+2( h_1 m_2 + m_5 )$
is positive, then the flip bifurcation is supercritical, otherwise it becomes subcritical. Here,
$\xi _1$
and
$\xi _2$
are given in (3.5), and
$h_1, m_1, m_2, m_3, m_5$
are given in the following proof in (4.6) and (4.8).
Proof. For the application of the centre manifold theorem, let us define the right-hand side of the equations in system (1.8) as

Expanding the Taylor series of these functions around the fixed point

gives us the following equivalent system of (1.8):

where

and

with
$\rho =\rho _{1}$
as in (4.1), where
$\xi _1$
,
$\xi _2$
are given in (3.5), and H.O.T. represents the higher order terms.
In this expansion, we take
$\bar {\rho }=\rho _1$
and
$M_{n}-M_{*}=X_{n}, \: N_{n}-N_{*}=Y_{n}, \: \rho -\bar {\rho }=\tilde {\rho }$
, which transform the fixed point
$E_*$
to the origin and the bifurcation parameter’s critical value to zero, that is,
$\rho _c=0$
. Next, we construct the matrix

whose columns are the eigenvectors corresponding to the eigenvalues of
$J_*$
, given in (3.6), when
$\rho =\bar {\rho }=\rho _1$
, which are
$ \lambda _1=-1 \quad \text {and} \quad \lambda _2 = 3-\rho _1 \xi _2.$
Here,
$\rho _{1}$
is given in (4.1) with
$\xi _1$
and
$\xi _2$
as in (3.5), and
$B_{21}$
is given in (4.3). Using the transformation

in (4.2), the following system is obtained:


where


with

and
$B_{21}$
is given in (4.3). Applying centre manifold theorem [Reference Baydemir, Merdan, Karaoglu and Sucu9, Reference Kuznetsov, Kuznetsov and Kuznetsov35, Reference Wiggins and Golubitsky53] to the system (4.5), one may obtain the one-dimensional system

with

Now, let us apply Theorem 2.4 to (4.7). Considering the right-hand side of (4.7) as
$f( U_n,\tilde {\rho })$
, we calculate that

So, from Theorem 2.4, we may conclude that there exists a flip bifurcation for the system (1.8) under the conditions (i), (ii), (iii) in Theorem 4.2. So the proof of Theorem 4.2 is complete.
4.2. Neimark–Sacker bifurcation
This section focuses on the construction of the Neimark–Sacker bifurcation at the point

with the help of condition (v) in Lemma 2.2.
Theorem 4.3. If

then there exists a Neimark–Sacker bifurcation for the equilibrium point
$E_*$
.
Proof. It is known that if the characteristic roots of the Jacobian matrix at
$E_*$
are complex conjugate with modulus 1, then there may be a Neimark–Sacker bifurcation at this point. In our problem, the characteristic equation is given by

Under (4.9), conditions given in Lemma 2.2(v) are satisfied and the proof is complete.
To show the existence of the Neimark–Sacker bifurcation, we apply Theorem 2.6. For this aim, let us obtain the characteristic roots and the bifurcation parameter explicitly under the condition
$ b_1(a_1 - a_2) + a_2 b_2(a_1 + a_2)> 0.$
The complex characteristic roots are in the form

with

Note that when
$\rho ={1}/{\xi _1}$
,

and our critical bifurcation parameter is
$\rho _c={1}/{\xi _1}$
and

because of the condition
$a_1>a_2$
. Furthermore, since
${\xi _2}>0$
, it is seen that
$( \lambda _{1,2}(\rho _c))^k=e^{\pm i k \theta =0} \neq 1$
for
$k=1,2,3,4$
. So, we conclude that Neimark–Sacker bifurcation occurs as for the fixed point
$E_*$
as
$\rho $
passes
$\rho _c={1}/{\xi _1}$
from left to right.
To construct the normal form of this bifurcation, let us consider the system (4.2) again as follows:

where
$B_{11}, B_{12}, B_{21}$
and
$f_1(X_n,Y_n,\tilde {\rho },\bar {\rho })$
,
$g_1(X_n,Y_n,\tilde {\rho },\bar {\rho })$
are given in (4.3) and (4.4), respectively. With this system, we transform the fixed point
$E_*$
to the origin and the bifurcation parameter
$\rho _c=\bar {\rho }$
to zero.
Now, let us consider the following Jacobian matrix:


We now create a matrix

whose columns are made up of the real and imaginary parts of an eigenvector that correspond to
$\lambda _1=\mu -i \omega $
, where

Next, we apply the transformation

to the system (4.10) and procure the following system:

where

with

and

However, the direction of this bifurcation is calculated with the help of the formula [Reference Guckenheimer and Holmes21]

with the following functions, which are all calculated at the point
$(0,0,{1}/{\xi _1})$
:

and it is said that the bifurcation is supercritical if
$ a({1}/{\xi _1})<0$
and subcritical if
$a({1}/{\xi _1})>0$
. As a result of the above discussion, we state the following theorem.
Theorem 4.4. If
${\xi _2}/{4}<\xi _1$
and
$ a({1}/{\xi _1})\neq 0$
, then system (1.8) has a Neimark–Sacker bifurcation at the fixed point
$E_{*}=(M_{*},N_{*})$
and this bifurcation is supercritical if
$ a({1}/{\xi _1})<0$
and subcritical if
$a({1}/{\xi _1})>0$
.
5. Examples
This section provides two numerical examples to strengthen and broaden the conclusions drawn from the theory in the other sections.
Example 5.1. This example is constructed by choosing the parameters in [Reference Barman, Roy and Alam7]. Here, we obtain the system

with the initial conditions

The equilibrium point for this system is determined to be
$E_*=(0.947664,3.30649)$
. As the first condition in Theorem 3.3 holds, we conclude that this point is a sink as illustrated in Figure 1(a) and the phase portrait of the system is given in Figure 1(b).

Figure 1 (a) Stability of
$E_*=(0.947664,3.30649)$
of system (1.8) and (b) phase portrait for
$(M_{n},N_{n}).$
However, since the conditions of Theorem 4.2 are satisfied, we have established that there exists a flip bifurcation when
$\rho $
exceeds the critical value
$1.93915$
. This can also be observed in Figures 2(a) and 2(b), where the flip bifurcation takes place within the range
$[0, 3],$
which includes the critical value
$\rho _c=1.93915$
.

Figure 2 Bifurcation diagrams in
$\rho \in [0,3]$
with initial condition
$(M_{0},N_{0})=(0.1,0.1).$
Example 5.2. Let us consider the following system:

with the initial conditions

The equilibrium
$E_*=(0.276294,26.6608)$
of this system is a sink, as it satisfies the first condition of Theorem 3.3. This can be seen in Figure 3(a) and the phase portrait of the system is given in Figure 3(b). Meanwhile, since the conditions in Theorem 2.6 are true, there exists a Neimark–Sacker bifurcation at
$\rho _{c}=1.83218$
. This bifurcation is supercritical in view of (4.11), which is calculated as
$-0.000042416 < 0$
. This behaviour is presented in Figures 4(a) and 4(b), which occurs within the range
$[0, 3]$
.

Figure 3 (a) Stability of
$E_*=(0.276294,26.6608)$
of system (1.8) and (b) phase portrait for
$(M_{n},N_{n}).$

Figure 4 Bifurcation diagrams in
$\rho \in [0,3]$
with initial condition
$(M_{0},N_{0})=(0.1,0.1).$
6. Conclusion
This paper explores the bifurcation properties of a fractional-order prey–predator model influenced by the wind, which is one of the most important abiotic factors often overlooked in ecological studies. The model was discretized using a piecewise discretization approach, allowing a detailed analysis of its discrete-time dynamics. We choose
$\rho =h^\alpha $
as the bifurcation parameter that consists of both the discretization parameter h and the fractional order
$\alpha $
of the system. The local stability of fixed points was investigated, and it was shown using the centre-manifold theorem and bifurcation theory that the system experiences both flip and Neimark–Sacker bifurcations at a positive fixed point. Numerical simulations were provided to illustrate and validate the theoretical results.
This work underscores the importance of including the abiotic factors such as wind in ecological models and demonstrates the advantages of fractional-order systems for capturing complex dynamics. Future research could focus on extending the model to include additional ecological interactions or exploring control strategies for practical applications. Further, validating the theoretical findings with empirical data could enhance the model’s relevance and applicability to the real-world ecological scenarios.
Acknowledgement
We would like to express our sincere gratitude to the referees for their meticulous review and valuable suggestions. The constructive feedback provided has greatly contributed to the clarity and rigour of this work. The referee’s detailed comments have been instrumental in improving both the mathematical exposition and the overall quality of the manuscript.