1,017
Views
1
CrossRef citations to date
0
Altmetric
Research Article

Combining impact of velocity, fear and refuge for the predator–prey dynamics

&

Abstract

We develop a deterministic predator–prey compartmental model to investigate the impact of their velocities on their interactions. Prey hides in a refuge area and comes out of this area when predation pressure declines. To avoid predation, prey can limit their velocity. For antipredator behaviour, we examined that prey mortality increases when either predator or prey velocity increases while raising antipredator behaviour increases prey density. We proved that predator free equilibrium is globally asymptotically stable and co-existing equilibrium will be globally stable under certain conditions. We find that transcritical bifurcations occur at predator-free equilibrium at the certain value of the death rate of the predator.

1. Introduction

In theory, ecology is the scientific study and aesthetic analysis of interactions among organisms and their environments. It consists of organisms, the communities they create and the nonliving components of their environment that are in dynamic interaction with each other. Historically, predators and their prey have been and will remain an important theme in ecology and mathematical ecology due to their universality. The interaction between them is a key feature of population dynamics.

A prey in the natural world protects itself from predators by hiding in an area where the predator cannot easily discover it. A refuge area is the name given to this location. By using the refuge, some members of the prey population are protected from predators, resulting in a lower predation rate. As a result, refuges are proportionate to the prey density while these are inversely proportional to the predator density. Prey refuge rate has been used as a crucial parameter in a prey–predator model in several studies [Citation7,Citation19,Citation35,Citation36]. Das and Samanta [Citation7] investigated the impact of predator fear on prey species and where both species become infected due to environmental toxicant. A diffusive predator–prey model with prey refuge was examined by [Citation1]. Mangroves, according to Glazner et al. [Citation11] can change predator–prey interactions by increasing the value of prey refuge in a mangrove-marsh ecotone. A stage-organized predator–prey model with prey refuge was examined for uniform persistence and global asymptotic stability by [Citation13]. The global stability of a stage structure predator–prey system with prey refuge was studied in [Citation33]. In [Citation33], the authors found that the model undergoes a Hopf bifurcation when the delay crosses specific critical values. Xie et al. [Citation34] investigated the persistence and stability of a modified Leslie–Gower predator–prey model with prey refuge, and they found that the prey refuge had a favourable effect on the persistence feature. Some mathematical models of biological systems have been developed to capture the effect of prey refuge, which protects some of the prey from predators [Citation2,Citation12].

There is a tendency for predator studies to concentrate only on direct killing by predators since such deaths can easily be observed in nature. However, prey species respond to predation risk in a variety of ways, including habitat change, foraging, and alertness, and different physiological changes. Based on numerous experiments, Zanette et al. [Citation37] have demonstrated that song sparrows reduce by 40% the offspring by predation fear apart from direct killing of birds. However, many field data show that the indirect impact of predator species on prey species has a major impact on population dynamics [Citation3,Citation4,Citation15,Citation25]. Das and Samanta [Citation8] investigated the impact of predator fear on prey species and where both species become infected due to environmental toxicant. In [Citation38], the authors explored a Holling-II predator–prey model with a prey refuge and the fear effect. They discovered that in a positive equilibrium, the fear effect not only reduces the predator's population density, but also stabilizes the system. The stability of a delayed predator–prey model with prey refuge and fear effect was investigated by [Citation14], and it was discovered that refuge below the threshold level is beneficial to the system. Impact of fear in a delay-induced predator–prey system with intraspecific competition within predator species was studied by Das et al. [Citation5]. Xiao and Li [Citation32] looked at a predator–prey model with mutual interference and a fear effect. They concluded that mutual interference can stabilize the predator–prey system when compared to the corresponding predator–prey model without mutual interference. Sasmal and Takeuchi [Citation26] investigated the multistability and Hopf bifurcation of a predator–prey system with fear effect. Das and Samanta [Citation6] examined a stochastic model where fear is in a prey population and an alternative food is available for predators. A comparison study of predator–prey system in deterministic and stochastic environments influenced by fear and its carry-over effects was examined by [Citation17]. In this regard, [Citation9,Citation16,Citation18,Citation20,Citation21,Citation29,Citation30] have more information.

Antipredator sensitivity refers to a shift in a consumer's behaviour in response to the presence of a predator. To avoid predators, prey can limit their speed, but this reduces their capacity to absorb resources, resulting in a trade-off between hiding and foraging. Sadowski and Grosholz [Citation24] created a basic tritrophic food chain that included predator and prey velocities. Antipredator behaviour, according to their hypothesis, allows prey to adopt to predators by reducing their speed in response to predator density. They studied the impact of both fast-moving (mobile) predators and slow-moving (sit and wait) predators on prey density and predator–prey cycle amplitude in equilibrium. Active search is advantageous when both predator and prey move randomly, but the relative advantage of active search reduces when the predator and prey adopt more directional movement [Citation27]. In [Citation10], the authors found that active searching predators are anticipated to prey on slow-moving species, while ambush predators are predicted to prey on fast-moving animals in a search mode model based on search in a three-dimensional space. Ross and Winterhalder [Citation23] recently created a two-dimensional predator–prey encounter model that includes random prey and predator movement.

In a grassland of small grasses with trees in small numbers, commonly found predators are lions, tigers, etc. They feed upon other animals like zebras, deer, etc. It is very easy for predators to locate their prey in small grassland. Prey animals run fast to reach a safe place and escape from predators. To protect from predators the velocity of the weaker animals helps them to escape from predators. In this way due to their velocity they protect themselves and their chances of survival increase. Some recent studies verified it experimentally and revealed that fast velocity is unrelated to survival. Slow prey has higher chances to escape from predators in a careful and skilful way. Studies found that escape success depends on both velocity and manoeuvrability [Citation24,Citation27]. Ross and Winterhalder [Citation23] examined that the encounter rate of prey and predator increases on increasing prey velocity. Slow prey has better chance to escape predation with high manoeuvrability. Large size prey needs to be faster and manoeuvrable to escape predation than the smaller size prey. It is verified that zebra and impala evade capture if they turn quickly which is possible at slow speed. If prey run very fast, then there are chance for prey to slip and fall and be easily captured by predators. Sadowski and Grosholz [Citation24] also examined and found that animals should run slower with manoeuvrability and fastest velocity is not always best as when animals run fast their control and manoeuvrability decline. Preys with lower velocity than their predators are always in an advantageous position because they can rapidly change directions [Citation27]. The fast running animals are likely to lose controls on their movements. High velocity reduces control and increases the chances of mistakes that goes in favour of predators. Preys of small size like Dick-dik are vulnerable by a greater range of predators. They are unable to defend themselves and to out run predators. They do not move far away their territory and make small feeding groups. They move slowly and feed cautiously. Predators detect prey only when they change their habitats. Prey prefers to hide in one place and remain still to detect ambush predators [Citation24].

We investigate our model by including predator fear, which is based on [Citation28] paradigm. Preys leave the refuge only when they are no longer afraid, otherwise, they return to the safe region. By enabling predator density to alter prey velocity, which specifies how prey behaviour changes in response to predator increases, we included predator and prey velocities with antipredator behaviour in our model as studied in [Citation24]. Our present study explore the combine effects of fear, refuge and velocities on the predator–prey dynamics.

2. Model formulation

We consider three-species model, a predator species (y) and two different habitats of prey species. One is called the prey in the refuge habitat (x1) and the others are called the prey outside the refuge habitat (x2). It is assumed that the prey species are available inside the refuge habitat and that their population grows logistically, where a is the growth rate of the prey species inside the refuge, k is the carrying capacity of the prey in the refuge habitat. It is also assumed that the prey species inside the refuge habitat is saved from predation. The prey species moves to a second habitat outside the refuge when the pressure of predation fear is decreased, and in this habitat the prey species can be killed by predators under the law of mass action. μ is the fear parameter, α is the migration rate from the refuge habitat. As predation fear increases in the prey species due to the presence of predators, the prey species migrate to the refuge habitat, where β represents the immigration rate into the refuge habitat and b is the feeding rate of the predator on the prey outside the refuge habitat. c is the conversion rate of prey to predator. In the absence of a prey species, predators die exponentially, where d is the death rate of the predator. We included predator and prey velocities in this model, with w representing the predator velocity. The prey velocity is a decreasing function of y, is γeθy. The maximum prey movement rate, defined as the velocity of the prey in the absence of a predator, is specified by γ. Preys are hampered by increasing their encounter rate with predators as the velocity rises. As a result, we include the antipredator behaviour θ in our model by allowing predator density to affect the prey velocity. Antipredator sensitivity describes how prey behaviour changes in response to predator increases. The following system of differential equations is used to simulate the predator–prey interaction: (1) dx1dt=ax1(1x1k)αx11+μy+βx2dx2dt=αx11+μyβx2bx2yw2+γ2e2θydydt=cx2yw2+γ2e2θydy(1)

3. Positivity of solutions

We will prove that all solutions of the model with positive initial data will remain positive for all time t>0. dx1dtx1=0=βx20dx2dtx2=0=αx11+μy0dydty=0=00 Hence, all solutions will remain positive for all time [Citation28].

4. Boundedness

Theorem 4.1

All solutions (x1(t),x2(t),y(t)) of the system Equation1 with initial conditions x1(0)0,x2(0)0, y(0)0 are bounded for all t>0.

Proof.

Let w=x1+x2+y dwdt=dx1dt+dx2dt+dydt=ax1(1x1k)αx11+μy+βx2+αx11+μyβx2bx2yw2+γ2e2θy+cx2yw2+γ2e2θydy=ax1(1x1k)(bc)x2yw2+γ2e2θydy From system (Equation1), the feeding rate of predator on the prey outside the refuge habitat will always be greater than the conversion rate of prey to predator. Therefore, b>c and so dwdtax1(1x1k)dydwdtax1dy For any positive q, we have dwdt+qwax1+qx1+qx2+(qd)y Since the carrying capacity k is the maximum population size of prey in the refuge, so x1<k and x2<k where x2 is a part of x1. Therefore, dwdt+qw(a+2q)k+(qd)y If q<d, then dwdt+qw(a+2q)k=L(say) The above differential inequality can be expressed by ddt(wLq)q(wLq) Using theory of differential inequality for w(t), we get 0w(x1,x2,y)Lq(1eqt)+w(0)eqt where w(0)=w(x1(0),x2(0),y(0)). Moreover, for t, 0<wLq. Hence all the solutions for the system (Equation1), (initiating in R+3) are confined in the region

Δ={(x1,x2,y)R+3:w(t)Lq+ϵ,foranyϵ>0}. This shows that the solution of the system represented by (Equation1) is bounded.

5. Equilibrium points of the system

We find three biologically meaningful equilibrium points:

  • The extinction of all populations E¯0=(0,0,0).

  • Predator Free Equilibrium E¯1=(x¯1,x¯2,0)=(k,αkβ,0), where the prey species only survive and the predator goes to extinction.

  • Co-existing equilibrium E¯2=(x1,x2,y) where

    x2=dcw2+γ2e2θy, x1=(1+μyα)x2(β+byw2+γ2e2θy) and y is the positive root of the following equation:

(2) (1+μyα)[β+byw2+γ2e2θy][aak(1+μyα)(dcw2+γ2e2θy)×(dcw2+γ2e2θy)(β+byw2+γ2e2θy)α1+μy]+β=0(2)

6. Stability analysis

6.1. Local stability analysis

Theorem 6.1

  

  1. The trivial equilibrium point E¯0=(0,0,0) is always unstable.

  2. Predator free equilibrium E¯1=(k,αkβ,0) is locally asymptotically stable if cx2¯w2+γ2<d

  3. The co-existing equilibrium E¯2=(x1,x2,y) will be locally asymptotically stable if (ax1k+βx2x1)(αx1μ(1+μy)2+bx2w2+γ2e2θy[w2+γ2e2θy(1θy)])>α2x1μ(1+μy)3

Proof.

The stability matrix for the system (Equation1) linearized about the equilibrium point is (3) (a2ax1kα1+μyβαx1μ(1+μy)2α1+μyβbyw2+γ2e2θyαx1μ(1+μy)2bx2w2+γ2e2θy×[w2+γ2e2θy(1θy)]0cyw2+γ2e2θyd+cx2w2+γ2e2θy×[w2+γ2e2θy(1θy)])(3)

  1. For E¯0=(0,0,0), the stability matrix is (aαλβ0αβλ000dλ) The corresponding characteristic equation is (d+λ)[λ2+λ(βa+α)aβ]=0. Therefore, E¯0=(0,0,0) is unstable

  2. Now the stability for E¯1=(k,αkβ,0) is ((a+α)λβαμkαβλαx¯1μbx¯2w2+γ200(cx2w2+γ2d)λ) The corresponding characteristic equation is ((cx2w2+γ2d)λ)[λ2+λ(a+α+β)+aβ]=0. Hence, the equilibrium will be locally asymptotically stable if cx2w2+γ2<d

  3. The stability for E¯2 is (p1λp2p3q1q2λq30γ2γ3λ) where p1=ax1kβx2x1<0

    p2=β>0

    p3=αx1μ(1+μy)2>0

    q1=α1+μy>0

    q2=αx1x2(1+μy)<0

    q3=αx1μ(1+μy)2bx2w2+γ2e2θy[w2+γ2e2θy(1θy)]<0 if θ<1y

    γ2=ydx2>0

    γ3=dγ2θye2θy<0

    The corresponding characteristic equation is (4) λ3+λ2(q2γ3p1)+λ((q2+γ3)p1+q2γ3γ2q3p2q1)+(p1q2γ3+p1γ2q3+p2q1γ3p3q1γ2)=0(4) Equation (Equation4) can be expressed as (5) λ3+a1λ2+a2λ+a3=0(5) where a1=(q2+γ3+p1)>0

    a2=(q2+γ3)p1+q2γ3γ2q3p2q1

    a3=(p1q2γ3+p1γ2q3+p2q1γ3p3q1γ2)=γ3(p1q2+p2q1)+γ2(p1q3p3q1)

    where γ3<0, p1q2+p2q1<0 if p2q1<p1q2

    or βα1+μy<aαx12kx2(1+μy)+βα1+μy, aαx12kx2(1+μy)>0 which is true that γ3(p1q2+p2q1)>0,

    For γ2(p1q3p3q1) where γ2>0, if p1q3>p3q1 then a3>0

    or (ax1k+βx2x1)(αx1μ(1+μy)2+bx2w2+γ2e2θy[w2+γ2e2θy(1θy)])>α2x1μ(1+μy)3

    Now, a1a2>a3(q2γ3p1)[(q2+γ3)p1+q2γ3γ2q3p2q1]>p1q2γ3+p1γ2q3+p2q1γ3p3q1γ2 or (6) p1q22+2p1q2γ3+q22γ3q2γ2q3p2q1q2+p1γ32+q2γ32γ2q3γ3+q2p12+γ3p12p2p1q1p3q1γ2<0(6) sign of various terms of Equation (Equation6) are mentioned

    Now, p1q22p2p1q1+q2p12p2q1q2=q2(p1q2p2q1)+p1(p1q2p2q1)=(p1q2p2q1)(q2+p1)

    we proved p1q2p2q1>0 and q2+p1<0. So, a1a2>a3. Hence non-zero equilibrium will be asymptotically stable if p1q3>p3q1, or (ax1k+βx2x1)(αx1μ(1+μy)2+bx2w2+γ2e2θy[w2+γ2e2θy(1θy)])>α2x1μ(1+μy)3

Theorem 6.2

If the equilibrium point y exists, be the positive real root of the Equation (Equation2), and the inequality p1q3>p3q1 is true then coexisting equilibrium E¯2=(x1,x2,y) will be asymptotically stable.

6.2. Global stability analysis

Theorem 6.3

  

  1. Predator free system (Equation1) equilibrium E=(x¯1,x¯2) is globally asymptotically stable.

  2. Co-existing equilibrium E¯2=(x1,x2,y) is globally asymptotically stable if ax1+a4k(k+x1)2+αx11+μi3βi2x1s1αi1x2s2(1+μs3)+βx2+bx2s3w2+γ2+dy0

Proof.

  1. We construct a Lyapunov function as follows:

    L=x1x¯1x¯1ln(x1x¯1)+x2x¯2x¯2ln(x2x¯2) (7) dLdt=(1x¯1x1)dx1dt+(1x¯2x2)dx2dt=(x1x¯1x1)[ax1(1x1k)αx1+βx2]+(x2x¯2x2)(αx1βx2)(7) at equilibrium, ax¯1(1x¯1k)αx¯1+βx¯2=0

    i.e. α=a(1x¯1k)+βx¯2x¯1 and αx¯1=βx¯2β=αx¯1x¯2

    using above values of α and β into Equation (Equation7) we get dLdt=(x1x¯1)[a(1x1k)a(1x¯1k)βx¯2x¯1+βx2x1]+(x2x¯2)[αx1x2αx¯1x¯2] or dLdt=ak(x1x¯1)2+β(x1x¯1)(x¯1x2x¯2x1x1x¯1)+α(x2x¯2)(x¯2x1x¯1x2x2x¯2). Since the carrying capacity k is the maximum population size of prey in the refuge, so, x¯1<k and x¯2<k. Therefore, we have dLdtak(x1x¯1)2(x1x¯2x2x¯1)[βk(x1x¯1x1)βk(x2x¯2x2)] or dLdtak(x1x¯1)2(x1x¯2x2x¯1)βk(x¯1x1+x¯2x2)ak(x1x¯1)2(x1x¯2x2x¯1)2x1x2βk0.

  2. From the second equation of (Equation1) and using the comparison theorem [Citation31], we can obtain dx2dtαx1βx2 At supremum of x2, dx2dt=0, then (8) αx1βx2(8) Therefore, (9) x2αx1β.(9) From the first equation of system (Equation1), we have dx1dtax1(1x1k)+βx2 By using Equation (Equation8), we get dx1dtax1(1x1k)+αx1. At supremum of x1, dx1dt=0, then 0aax1k+αx1ka(a+α). Therefore, the supremum of x1=ka(a+α)=s1 and from (Equation9) the supremum of x2=αkβa(a+α)=s2

    Now using system (Equation1)3, we have dydt=cx2yw2+γ2e2θydy At supremum of y, dydt=0, therefore cx2w2+γ2e2θy=dcs2w2+γ2e2θydw2+γ2e2θydcs2e2θyγ2[(dcs2)2w2]y12θln[γ2[(dcs2)2w2]] Therefore, the supremum of y = 12θln[γ2[(dcs2)2w2]]=s3.

    Similarly, from the first equation of system (Equation1) we have dx1dtax1(1x1k)αx1 At infimum of x1, dx1dt=0, therefore 0a(1x1k)αx1ka(aα) Therefore, infimum of x1=ka(aα)=i1.

    Now, from the third equation of system (Equation1)3 we have dydt=cx2yw2+γ2e2θydy At infimum of y, dydt=0, therefore cx2w2+γ2e2θy=dcx2w2+γ2dx2dcw2+γ2. Therefore, infimum of x2=dcw2+γ2=i2.

    Now, from the second equation of system (Equation1) we have dx2dt=αx11+μyβx2bx2yw2+γ2e2θydx2dtαi11+μyβx2bx2yw2+γ2e2θy At infimum of x2, dx2dt=0, therefore βx2+bx2yw2+γ2e2θyαi11+μyβs2+bs2yw2+γ2αi11+μyy2[bs2μw2+γ2]+y[[μβs2+bs2w2+γ2]]+(βs2αi1)0. Therefore, infimum of y=i3 where

    i3=[μβs2+bs2w2+γ2]+(μβs2+bs2w2+γ2)24(bs2μw2+γ2)(βs2αi1)2bs2μw2+γ2

    provided β<αi1s2.

    Now, we construct a Lyapunov function as follows: L=x1x1x1ln(x1x1)+x2x2x2ln(x2x2)+yyyln(yy)dLdt=(1x1x1)dx1dt+(1x2x2)dx2dt+(1yy)dydt=(x1x1x1)[ax1(1x1k)αx11+μy+βx2]+(x2x2x2)[αx11+μyβx2bx2yw2+γ2e2θy]+(yyy)[cx2yw2+γ2e2θydy]dLdt=(x1x1)[a(1x1k)α1+μy+βx2x1]+(x2x2)[αx1x2(1+μy)βbyw2+γ2e2θy]+(yy)[cx2w2+γ2e2θyd]dLdt=ak(x1k+x12)2ax1+a4k(k+x1)2+αx11+μyβx2x1x1αx1x2x2(1+μy)+βx2(bc)x2yw2+γ2e2θy+bx2yw2+γ2e2θycx2yw2+γ2e2θydy+dydLdtax1+a4k(k+x1)2+αx11+μyβx2x1x1αx1x2x2(1+μy)+βx2+bx2yw2+γ2+dydLdtax1+a4k(k+x1)2+αx11+μi3βi2x1s1αi1x2s2(1+μs3)+βx2+bx2s3w2+γ2+dy. Therefore, dLdt0ifax1+a4k(k+x1)2+αx11+μi3βi2x1s1αi1x2s2(1+μs3)+βx2+bx2s3w2+γ2+dy0.

7. Transcritical bifurcation

We investigate the transcritical bifurcation for the system (Equation1) taking d=d0 as the bifurcation parameter. According to the Sotomayor theorem [Citation22] for local bifurcation, the following conditions are satisfied:

  1. wTfd(E¯1,d0)=0

  2. wT(Dfd(E¯1,d0).υ)0

  3. wTD2f(E¯1,d0)(υ,υ)0.

Then the system (Equation1) experiences a transcritical bifurcation at the free predator equilibrium point E¯1=(k,αkβ,0) as the parameter d passes through the bifurcation value d=d0=cx¯2w2+γ2 where Df is denoted by the matrix of partial derivatives of the components of f with respect to the components of X=(x1,x2,y) and fd is denoted for the vector of partial derivatives of the components of f with respect to the scalar d. (υ) is an eigenvector of A=Df(E¯1,d0) corresponding to the eigenvalue λ3=0 and ω is an eigenvector of AT corresponding to the eigenvalue λ3=0,

Df(X)=(f1x1f1x2f1yf2x1f2x2f2yf3x1f3x2f3y)

D2f(X)=(2f1x1x12f1x1x22f1x1y2f1x2x12f1x2x22f1x2y2f1yx12f1yx22f1yy2f2x1x12f2x1x22f2x1y2f2x2x12f2x2x22f2x2y2f2yx12f2yx22f2yy2f3x1x12f3x1x22f3x1y2f3x2x12f3x2x22f3x2y2f3yx12f3yx22f3yy)

Theorem 7.1

The system (Equation1) undergoes a transcritical bifurcation at the free predator equilibrium point E¯1=(k,αkβ,0) when d=d0=cx¯2w2+γ2.

The linearized system around the equilibrium point E¯1 is ((a+α)βαμkαβαx¯1μbx¯2w2+γ200cx¯2w2+γ2d)

Now λ3=0 at d=cx¯2w2+γ2, So d0=cx¯2w2+γ2 J(E¯1,d0)=((a+α)βαμkαβαx¯1μbx¯2w2+γ2000) Let us define υ=(υ1,υ2,υ3)T and ω=(ω1,ω2,ω3)T are respectively the right and left eigenvectors of λ3=0.

Now solving J(E¯1,d0)υ=0 ((a+α)υ1+βυ2+μαkυ3αυ1βυ2(μαk+bx¯2w2+γ2)υ30)=(000) implies that

υ=((bax¯2w2+γ2)υ3,(αbaβx¯2w2+γ2(μαkβ+bβx¯2w2+γ2))υ3,υ3), where υ3 is any nonzero real number.

Then solving JT(E¯1,d0)ω=0 ((a+α)ω1+αω2βω1βω2μαkω1(μαk+bx¯2w2+γ2)ω2)=(000) we get ω1=ω2=0 and ω3 is any nonzero real number, so ω=(0,0,ω3).

Now, system (Equation1) can be rewritten as in the following vector form: (10) X˙=f(X),(10) where X=(x1,x2,y) and f(X)=(ax1(1x1k)αx11+μy+βx2αx11+μyβx2bx2yw2+γ2e2θycx2yw2+γ2e2θydy)

Taking the derivative of f(X) with respect to d, we get (11) fd(X)=(00y)(11) then fd(E¯1,d0)=(000), Hence ωTfd(E¯1,d0)=0, So, the first condition is satisfied.

Next, taking the derivative of fd(X) with respect to X=(x1,x2,y)T, we get Dfd(X)=(000000001) then Dfd(E¯1,d0)=(000000001) Therefore, we have ωTDfd(E¯1,d0).υ= (00ω3)(000000001) (υ1υ2υ3)=ω3υ30.

Therefore, the second condition is satisfied.

Furthermore, D2f(X)=(A1A2A3A4A5A6A7A8A9A10A11A12A13A14A15A16A17A18A19A20A21A22A23A24A25A26A27) where

A1=2ak, A2=0, A3=αμ(1+μy)2, A4=0, A5=0, A6=0, A7=αμ(1+μy)2, A8=0, A9=2αx1μ2(1+μy)3, A10=0, A11=αμ(1+μy)2, A12=0, A13=0, A14=0, A15=bw2+γ2, A16=αμ(1+μy)2, A17=bw2+γ2, A18=2αx1μ2(1+μy)3, A19=0, A20=0, A21=0, A22=0, A23=0, A24=cw2+γ2,A25=0, A26=cw2+γ2, A27=0.

The tensor product (υ,υ)=(υ12υ1υ2υ1υ3υ2υ1υ22υ2υ3υ3υ1υ3υ2υ32) so, ωTD2f(E1,d0)(υ,υ)=(00ω3)(A1A2A3A4A5A6A7A8A9A10A11A12A13A14A15A16A17A18A19A20A21A22A23A24A25A26A27)(υ12υ1υ2υ1υ3υ2υ1υ22υ2υ3υ3υ1υ3υ2υ32)=2cw2+γ2ω3(υ2υ3)0. Therefore, according to the Sotomayor theorem [Citation22] for local bifurcation, system (Equation1) has a transcritical bifurcation at steady state at the free predator equilibrium point E¯1=(k,αkβ,0) as the parameter d passes through the bifurcation value d=d0=cx¯2w2+γ2 (Figure ).

Figure 1. The transcritical bifurcation at d = 0.266204905858465, the parameters are a=0.7,k=0.8,α=0.035,μ=5,β=0.0119,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,w=2. (The red colour signifies the stable equilibrium and the black color shows the unstable equilibrium).

Figure 1. The transcritical bifurcation at d = 0.266204905858465, the parameters are a=0.7,k=0.8,α=0.035,μ=5,β=0.0119,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,w=2. (The red colour signifies the stable equilibrium and the black color shows the unstable equilibrium).

8. Numerical analysis

In this section, numerical simulation results are given to confirm the above theoretical results. Table  contains the parameter values that are used, some of them might be changed to study their effect.

Table 1. Parameter values used for simulations.

Figures  and  show the convergence of the free predator equilibrium point (k,αkβ,0) and the coexistence equilibrium point (x1,x2,y) respectively.

Figure 2. The convergence of the free predator equilibrium point (k,αkβ,0), the parameters are a=0.7,k=0.8,α=0.035,μ=5,β=0.0119,b=0.0112,c=0.04,d=0.266205,θ=0.69,γ=2,ω=2, E¯1=(0.8,2.352941176,0).

Figure 2. The convergence of the free predator equilibrium point (k,αkβ,0), the parameters are a=0.7,k=0.8,α=0.035,μ=5,β=0.0119,b=0.0112,c=0.04,d=0.266205,θ=0.69,γ=2,ω=2, E¯1=(0.8,2.352941176,0).

Figure 3. The convergence of the coexistence equilibrium point E¯2, the parameters are a = 0.7, k = 0.8, α=0.035, μ=5,β=0.0119,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,ω=2.

Figure 3. The convergence of the coexistence equilibrium point E¯2, the parameters are a = 0.7, k = 0.8, α=0.035, μ=5,β=0.0119,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,ω=2.

8.1. Impact of predator and prey velocities on their populations

From figures  we find that increasing the predator velocity has no effect on the prey population in the refuge but this will decrease the prey population outside the refuge habitat. On the other side, when the predator velocity rises, so does the predator population.

Figure 4. The effect of predator velocity on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,β=0.0119,c=0.04,d=0.07,θ=0.69,γ=2.

Figure 4. The effect of predator velocity on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,β=0.0119,c=0.04,d=0.07,θ=0.69,γ=2.

Figure  represents that as the prey velocity increases, the predator population grows while the prey population outside the refuge decreases. Furthermore, the prey population in the refuge habitat is unaffected by the increase in the prey velocity. Hence the high speed of prey becomes more vulnerable to predators. Therefore, any increase in predator or prey velocity raises the prey mortality and hence the prey population declines outside the refuge. Refuge area is safe from predation so, the prey population in the refuge reaches its equilibrium value. Numerical results confirm the claim that the slow speed of prey is useful for the survival of the prey population and at the same time the high speed of the predator is good for the predator population.

Figure 5. The effect of prey velocity on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,β=0.0119,c=0.04,d=0.07,θ=0.69,w=2.

Figure 5. The effect of prey velocity on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,β=0.0119,c=0.04,d=0.07,θ=0.69,w=2.

8.2. Effect of antipredator sensitivity (θ)

As demonstrated in figures , raising antipredator sensitivity increases prey density outside of the refuge habitat and as a result of this decreases the predator population density. The prey density inside the refuge is uneffected by antipredator sensitivity and will remain constant.

Figure 6. The effect of antipredator sensitivity on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,β=0.0119,c=0.04,d=0.07,w=2,γ=2.

Figure 6. The effect of antipredator sensitivity on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,β=0.0119,c=0.04,d=0.07,w=2,γ=2.

8.3. Effect of refuge parameter (β)

Figure  indicates that as the refuge parameter is increased, then more and more prey will be able to take shelter inside the refuge habitat and so prey population inside the refuge increases until it meet carrying capacity. Due to this process the prey population outside the refuge declines and so the predator population which depends on the prey population will also decline and eventually go extinct.

Figure 7. The effect of refuge parameter on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,w=2.

Figure 7. The effect of refuge parameter on the population densities, parameters are a=0.7,k=0.8,α=0.035,μ=5,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,w=2.

8.4. Effect of fear parameter (μ)

Figure  represents that as the fear of the predator increases the prey density inside the refuge habitat increases because of predator fear the prey will prefer to stay safe inside the refuge. The population of the prey outside the refuge and the predator population decreases. Since predators will get less food due to less number of prey available outside the refuge. So, by heavy predation on less number of prey the population of prey declines and ultimately the predator population will also decline.

Figure 8. The effect of fear parameter on the population densities, parameters are a=0.7,k=0.8,α=0.035,β=0.0119,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,w=2.

Figure 8. The effect of fear parameter on the population densities, parameters are a=0.7,k=0.8,α=0.035,β=0.0119,b=0.0112,c=0.04,d=0.07,θ=0.69,γ=2,w=2.

9. Conclusion

Predators provide a threat to all creatures, and they are all at risk of being devoured. Fear of predators causes prey populations to change their behaviour. We investigated the dynamics of predator–prey interactions in two settings, namely, refuge and out-of refuge, in this article. The prey is protected from predatory slaughter in the refuge, and it has enough supplies to survive, and the sanctuary's population rises logistically. Predators engage with prey outside of their safe haven and may kill them. Prey lives in fear of predators, but as that fear fades, prey emerges from their hiding places. Prey seeks refuge in the refuge as their fear of predation grows. By allowing predator density to influence prey velocity, we included predator and prey velocities with antipredator behaviour in our model.

Three biologically viable equilibria are obtained, and their stability is discussed. The equilibrium without prey and predator populations will always exist, but it will be unstable, implying that prey and predator populations will never go extinct. There will always be an equilibrium with zero predator population and a nonzero prey population, and it is globally stable. The nonzero equilibrium will be asymptotically stable if (ax1k+βx2x1)(αx1μ(1+μy)2+bx2w2+γ2e2θy[w2+γ2e2θy(1θy)])>α2x1μ(1+μy)3. At a specific value of the predator's mortality rate, a transcritical bifurcation occurs at predator free equilibrium.

We conducted a numerical analysis to see how the factors affect our system, and we observed that prey outside the refuge habitat and predator population decline and become extinct as the refuge parameter rises, whereas prey in the refuge increases until it meets the carrying capacity. Moreover, the prey density inside the refuge habitat grows as the fear parameter increases, but the prey density outside the refuge and predator density decreases. Furthermore, we discovered that when the antipredator sensitivity is low, predator and prey velocities have the same influence on the final equilibrium prey density. Any increase in predator or prey velocity causes prey mortality to rise and the equilibrium prey density to decline. However, boosting antipredator sensitivity increases the density of prey outside the refuge habitat while reducing the density of the predator population.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

We are very thankful to the reviewers for their valuable comments and suggestions which led to an improvement of our original manuscript. Research is funded by Sultan Qaboos University, Oman.

References

  • B. Chakraborty and N. Bairagi, Complexity in a prey-predator model with prey refuge and diffusion, Ecol. Complex. 37 (2019), pp. 11–23.
  • L. Chen, F. Chen, and L. Chen, Qualitative analysis of a predator–prey model with holling type II functional response incorporating a constant prey refuge, Nonlinear Anal. Real World Appl. 11 (2010), pp. 246–252.
  • S. Creel and D. Christianson, Relationships between direct predation and risk effects, Trends Ecol. Evol. 23(4) (2008), pp. 194–201.
  • W. Cresswell, Predation in bird populations, J. Ornitho. 152(S1) (2011), pp. 251–263.
  • B.K. Das, D. Sahoo, and G.P. Samanta, Impact of fear in a delay-induced predator-prey system with intraspecific competition within predator species, Math. Comput. Simul. 191 (2022), pp. 134–156.
  • A. Das and G.P. Samanta, Modeling the fear effect on a stochastic prey-predator system with additional food for the predator, J. Phys. A Math. Theor. 51(46) (2018), pp. 465601.
  • A. Das and G.P. Samanta, A prey-predator model with refuge for prey and additional food for predator in a fluctuating environment physica A, Stat. Mech. Appl. 538 (2020), pp. 122844.
  • A. Das and G.P. Samanta, Modelling the fear effect in a two-species predator–prey system under the influence of toxic substances, Rend. Del Circ. Mat. Di Palermo Ser. 2 70(3) (2021), pp. 1501–1526.
  • P. Dutta, D. Sahoo, S. Mondal, and G. Samanta, Dynamical complexity of a delay-induced eco-epidemic model with Beddington–DeAngelis incidence rate, Math. Comput. Simul. 197 (2022), pp. 45–90.
  • J. Gerritsen and J.R. Strickler, Encounter probabilities and community structure in zooplankton a mathematical model, J. Fish. Board Can. 34(1) (1977), pp. 73–82.
  • R. Glazner, J. Blennau, and A.R. Armitage, Mangroves alter predator-prey interactions by enhancing prey refuge value in a mangrove-marsh ecotone, J. Exp. Marine Biol. Ecol. 526 (2020), pp. 151336.
  • L. Ji and C. Wu, Qualitative analysis of a predator-prey model with constant-rate prey harvesting incorparating a constant prey refuge, Nonlinear Anal. Real World Appl. 11(4) (2010), pp. 2285–2295.
  • S. Khajanchi and S. Banerjee, Role of constant prey refuge on stage structure predator–prey model with ratio dependent functional response, Appl. Math. Comput. 314 (2017), pp. 193–198.
  • A. Kumar and B. Dubey, Modeling the effect of fear in a prey-predator system with prey refuge and gestation delay, Int. J. Bifurc. Chaos. 29(14) (2019), pp. 1950195.
  • S.L. Lima, Predators and the breeding bird behavioral and reproductive flexibility under the risk of predation, Bio. Rev. 84(3) (2009), pp. 485–513.
  • S. Mondal and G.P. Samanta, Impact of fear on a predator–prey system with prey-dependent search rate in deterministic and stochastic environment, Nonlinear. Dyn. 104(3) (2021), pp. 2931–2959.
  • S. Mondal and G. Samanta, A comparison study of predator–prey system in deterministic and stochastic environments influenced by fear and its carry-over effects, Eur. Phys. J. Plus 137(1) (2022), pp. 70.
  • S. Mondal, G.P. Samanta, and J.J. Nieto, Dynamics of a predator–prey population in the presence of resource subsidy under the influence of nonlinear prey refuge and fear effect, Complexity 2021 (2021), pp. 1–38.
  • D. Mukherjee and C. Maji, Bifurcation analysis of a holling type II predator–prey model with refuge, Chin. Phy. 65 (2020), pp. 153–162.
  • S. Pal, N. Pal, and S. Samanta, Effect of hunting cooperation and fear in a predator–prey model, Ecol. Complex. 39 (2019), pp. 100770.
  • S. Pal, N. Pal, and S. Samanta, Fear effect in prey and hunting cooperation among predators in a Leslie–Gower model, Math. Biosci. Eng. 16(5) (2019), pp. 5146–5179.
  • L. Perko, Differential Equations and Dynamical Systems, Springer, New York, 2001.
  • C.T. Ross and B. Winterhalder, Sit-and-wait versus active-search hunting: a behavioral ecological model of optimal search mode, J. Theor. Bio. 387 (2015), pp. 76–87.
  • J.S. Sadowski and E.D. Grosholz, Predator foraging mode controls the effect of antipredator behavior in a tritrophic model, Theor. Ecol. 12(4) (2019), pp. 531–544.
  • D. Sahoo and G.P. Samanta, Impact of fear effect in a two prey–one predator system with switching behaviour in predation, Differ. Equ. Dyn. Syst.,2021, pp. 1–23. doi:10.1007/s12591-021-00575-7.
  • S.K. Sasmal and Y. Takeuchi, Dynamics of a predator–prey system with fear and group defense, J. Math. Anal. Appl. 481(1) (2020), pp. 123471.
  • I. Scharf, E. Nulman, O. Ovadia, and A. Bouskila, Efficiency evaluation of two competing foraging modes under different conditions, Am. Nat. 168(3) (2006), pp. 350–357.
  • H. Wang, S. Thanarajah, and P. Gaudreau, Refuge-mediated predator-prey dynamics and biomass pyramids, Math. Biosci. 298 (2018), pp. 29–45.
  • X. Wang and X. Zou, Modeling the fear effect in predator–prey interactions with adaptive avoidance of predators, Bull. Math. Biol. 79(6) (2017), pp. 1325–1359.
  • X. Wang and X. Zou, Pattern formation of a predator–prey model with the cost of anti-predator behaviors, Math. Biosci. Eng. 15(3) (2017), pp. 775–805
  • Y. Xiao and L. Chen, Modelling and analysis of a predator–prey model with disease in the prey, Math. Biosci. 171 (1) (2001), pp. 59–82.
  • Z. Xiao and Z. Li, Stability analysis of a mutual interference predator–prey model with the fear effect, J. Appl. Sci. Eng. 22(2) (2019), pp. 205–211.
  • Z. Xiao, Z. Li, and Z. Zhu, Hopf bifurcation and stability in a beddington-DeAngelis predator–prey model with stage structure for predator and time delay incorporating prey refuge, Open Math. 17(1) (2019), pp. 141–159.
  • X. Xie, Y. Xue, and J. Chen, Permanence and global attractivity of a nonautonomous modified Leslie–Gower predator-prey model with holling-type II schemes and a prey refuge, Adv. Differ. Equ. 1 (2016), pp. 1–11.
  • P. Yang, Hopf bifurcation of an age-structured prey-predator model with Holling type II functional response incorporating a prey refuge, Nonlinear Anal. RWA 49 (2019), pp. 368–385.
  • Q. Yue, Dynamics of a modified Leslie–Gower predator-prey model with Holling-type II schemes and a prey refuge, SpringerPlus 5(1) (2016), pp. 1–12.
  • L.Y. Zanette, A.F. White, and M.C. Allen, Perceived predation risk reduces the number of offspring songbirds produce per year, Science 334(6061) (2011), pp. 1398–1401.
  • H. Zhang, Y. Cai, and S. Fu, Impact of the fear effect in a prey–predator model incorporating a prey refuge, Appl. Math. Comput. 356 (2019), pp. 46–66.