Processing math: 41%
Research article

Allee effect-driven complexity in a spatiotemporal predator-prey system with fear factor


  • In this paper, we propose a spatiotemporal prey-predator model with fear and Allee effects. We first establish the global existence of solution in time and provide some sufficient conditions for the existence of non-negative spatially homogeneous equilibria. Then, we study the stability and bifurcation for the non-negative equilibria and explore the bifurcation diagram, which revealed that the Allee effect and fear factor can induce complex bifurcation scenario. We discuss that large Allee effect-driven Turing instability and pattern transition for the considered system with the Holling-Ⅰ type functional response, and how small Allee effect stabilizes the system in nature. Finally, numerical simulations illustrate the effectiveness of theoretical results. The main contribution of this work is to discover that the Allee effect can induce both codimension-one bifurcations (transcritical, saddle-node, Hopf, Turing) and codimension-two bifurcations (cusp, Bogdanov-Takens and Turing-Hopf) in a spatiotemporal predator-prey model with a fear factor. In addition, we observe that the circular rings pattern loses its stability, and transitions to the coldspot and stripe pattern in Hopf region or the Turing-Hopf region for a special choice of initial condition.

    Citation: Yuhong Huo, Gourav Mandal, Lakshmi Narayan Guin, Santabrata Chakravarty, Renji Han. Allee effect-driven complexity in a spatiotemporal predator-prey system with fear factor[J]. Mathematical Biosciences and Engineering, 2023, 20(10): 18820-18860. doi: 10.3934/mbe.2023834

    Related Papers:

    [1] Fang Liu, Yanfei Du . Spatiotemporal dynamics of a diffusive predator-prey model with delay and Allee effect in predator. Mathematical Biosciences and Engineering, 2023, 20(11): 19372-19400. doi: 10.3934/mbe.2023857
    [2] Kalyan Manna, Malay Banerjee . Stability of Hopf-bifurcating limit cycles in a diffusion-driven prey-predator system with Allee effect and time delay. Mathematical Biosciences and Engineering, 2019, 16(4): 2411-2446. doi: 10.3934/mbe.2019121
    [3] Kawkab Al Amri, Qamar J. A Khan, David Greenhalgh . Combined impact of fear and Allee effect in predator-prey interaction models on their growth. Mathematical Biosciences and Engineering, 2024, 21(10): 7211-7252. doi: 10.3934/mbe.2024319
    [4] Honghua Bin, Daifeng Duan, Junjie Wei . Bifurcation analysis of a reaction-diffusion-advection predator-prey system with delay. Mathematical Biosciences and Engineering, 2023, 20(7): 12194-12210. doi: 10.3934/mbe.2023543
    [5] Shuo Yao, Jingen Yang, Sanling Yuan . Bifurcation analysis in a modified Leslie-Gower predator-prey model with fear effect and multiple delays. Mathematical Biosciences and Engineering, 2024, 21(4): 5658-5685. doi: 10.3934/mbe.2024249
    [6] Jin Zhong, Yue Xia, Lijuan Chen, Fengde Chen . Dynamical analysis of a predator-prey system with fear-induced dispersal between patches. Mathematical Biosciences and Engineering, 2025, 22(5): 1159-1184. doi: 10.3934/mbe.2025042
    [7] Tingting Ma, Xinzhu Meng . Global analysis and Hopf-bifurcation in a cross-diffusion prey-predator system with fear effect and predator cannibalism. Mathematical Biosciences and Engineering, 2022, 19(6): 6040-6071. doi: 10.3934/mbe.2022282
    [8] Ranjit Kumar Upadhyay, Swati Mishra . Population dynamic consequences of fearful prey in a spatiotemporal predator-prey system. Mathematical Biosciences and Engineering, 2019, 16(1): 338-372. doi: 10.3934/mbe.2019017
    [9] Moitri Sen, Malay Banerjee, Yasuhiro Takeuchi . Influence of Allee effect in prey populations on the dynamics of two-prey-one-predator model. Mathematical Biosciences and Engineering, 2018, 15(4): 883-904. doi: 10.3934/mbe.2018040
    [10] Zuolin Shen, Junjie Wei . Hopf bifurcation analysis in a diffusive predator-prey system with delay and surplus killing effect. Mathematical Biosciences and Engineering, 2018, 15(3): 693-715. doi: 10.3934/mbe.2018031
  • In this paper, we propose a spatiotemporal prey-predator model with fear and Allee effects. We first establish the global existence of solution in time and provide some sufficient conditions for the existence of non-negative spatially homogeneous equilibria. Then, we study the stability and bifurcation for the non-negative equilibria and explore the bifurcation diagram, which revealed that the Allee effect and fear factor can induce complex bifurcation scenario. We discuss that large Allee effect-driven Turing instability and pattern transition for the considered system with the Holling-Ⅰ type functional response, and how small Allee effect stabilizes the system in nature. Finally, numerical simulations illustrate the effectiveness of theoretical results. The main contribution of this work is to discover that the Allee effect can induce both codimension-one bifurcations (transcritical, saddle-node, Hopf, Turing) and codimension-two bifurcations (cusp, Bogdanov-Takens and Turing-Hopf) in a spatiotemporal predator-prey model with a fear factor. In addition, we observe that the circular rings pattern loses its stability, and transitions to the coldspot and stripe pattern in Hopf region or the Turing-Hopf region for a special choice of initial condition.



    The Allee effect is primarily associated with the relationship between per capita growth rate and population density [1,2,3,4]. The prevailing situation of the species being threatened due to the Allee effect has drawn major attention to researchers who are studying both the theoretical and practical ecological conservation in exploited environments. Biologists now perceive a diverse range of factors which include the benefit of reproduction, cooperative breeding [5,6,7], anti-predator character among the prey [8,9], foraging ability and environmental conditioning as well as the purpose of the influence of the Allee effect on the growth of prey [10,11,12]. A few reasons behind the emergence of the Allee effect may be stated as difficulties in mate finding, reproductive facilitation, predation, environmental conditioning and inbreeding depression [4,13]. In recent years, many eminent researchers have paid considerable attention to investigate the impact of the Allee effect in the growth of prey and the dynamical complexity of the predator-prey model [14,15,16,17]. The Allee effect is usually categorized by the manifestation of density dependence in the event of its relatively lower range. The population experiences a strong Allee effect resulting in the existence of a critical density where the population growth becomes negative. However, a weak Allee effect causes the population to have a reduced per capita growth rate at low population density, therefore exhibiting a positive per capita growth rate.

    A single species logistic model with adjusting for mating encounters was considered, which takes the following form [13]:

    dudt=ru(1uK)μhuu+h, (1.1)

    where r and K represent the intrinsic growth rate and the carrying capacity of the environment, respectively. h is a non-negative constant, which depicts the population density at which the probability of mating is half. The term hu+h is the probability of not mating, and the term μhuu+h represents the reduction of reproduction due to a mating shortage. Moreover, if 0<μ<r, it is called a weak Allee effect, and if μ>r then it has the strong Allee effect. Keeping the special aspects in mind, we rewrote the classical single species model (1.1) with additive Allee effect by using the following notation:

    dudt=aubucu2muu+h, (1.2)

    where a,b and c are positive constants that represent the birth rate, natural death rate and the death rate induced by intra-species competition, respectively. To obtain (1.1), we only need to set r=ab, K=abc and m=μh.

    Recently, Zanette et al. [8] found that the fear of predators could reduce about 40 of offspring that song sparrows could produce, the predation risk had an important effect on both the birth and survival rate of offspring. Based on the field observations in [8], Wang et al. [9] proposed a predator-prey model incorporating the fear effect in prey species as follows:

    {dudt=af0(k,v)ubucu2f(u)v,dvdt=v(d+ef(u)). (1.3)

    where the parameter k reflects the level of fear, f:R+R+ is the functional response of predators where R+ represents nonnegative real number set, d is the natural death rate of the predators, e is the conversion rate of prey's biomass to predator's biomass and the function f0 accounts for the cost of anti-predator defense due to fear and satisfies the following properties:

    {f0(0,v)=1,f0(k,0)=1,limkf0(k,v)=0,limvf0(k,v)=0,f0(k,v)k<0,f0(k,v)v<0.

    Moreover, their analyses showed that high levels of fear can stabilize the predator-prey system. However, relatively low levels of fear can induce multiple limit cycles through subcritical Hopf bifurcations, which is different from the classical predator-prey models, ignoring the fear factor where Hopf bifurcations are typically supercritical.

    Also, field observations show that fear can create an Allee effect [18]. Combining (1.2) and (1.3), we obtained a temporal predator-prey model with Allee and fear effects in prey species and density-dependent death factor of the predator species as follows:

    {dudt=f0(k,v)aubucu2muu+hpvg(u,v),dvdt=epvg(u,v)dvsv2, (1.4)

    where the parameters a,b,c,d,e,h,m and k have the same meanings as in (1.2) and (1.3). p represents predation factor of the predator species, s is the density-dependent death rate of the predators and g(u,v) denotes the functional response between predator and prey, which usually satisfies the following hypothesis:

    (H0) gC1(R2,R) satisfies g and is non-negative, and g(0,v)=0 for all v0. Furthermore, there exists a non-negative continuous function ϕ(u) and a polynomial function ψ(v) (of arbitrary degree) such that g(u,v)ϕ(u)ψ(v) for any (u,v)R2+, where R+=[0,).

    About the hypothesis (H0), the typical forms of the functional response g(u,v) are as follows:

    (i) Holling type Ⅰ g(u,v)=u, type Ⅱ g(u,v)=uu+κ, type Ⅲ g(u,v)=u2u2+κ;

    (ii) Beddington-DeAngelis type g(u,v)=u1+κ1u+κ2v;

    (iii) ratio-dependent type g(u,v)=uκ1u+κ2v;

    (iv) Crowley-Martin type g(u,v)=u(1+κ1u)(1+κ2v).

    Throughout the manuscript, attention is focused on the particular form of fear factor defined by f0(k,v)=11+kv [9]. By incorporating this function into the system of equations of (1.4), one obtains

    {dudt=au1+kvbucu2muu+hpvg(u,v),dvdt=epvg(u,v)dvsv2. (1.5)

    Considering the effect of spatial diffusion, one gets the corresponding spatiotemporal model which is given as follows:

    {u(x,t)t=f1(u,v)+du2u:=au1+kvbucu2muu+npvg(u,v)+du2u,xΩ,t>0,v(x,t)t=f2(u,v)+dv2v:=epvg(u,v)dvsv2+dv2v,xΩ,t>0,u(x,0)=u0(x)0,v(x,0)=v0(x)0,xΩ,uν=vν=0,xΩ,t>0. (1.6)

    Here, u(x,t) and v(x,t) are densities of the species u and v on spatial location x and time t, respectively. ΩR2 is a bounded domain with a smooth boundary having no-flux boundary conditions, where R represents the set of real number. du and dv are diffusion coefficients that represent the respective diffusive velocity of the two species.

    In fact, the effect of the fear factor and multiple Allee effects in a temporal or spatiotemporal predator-prey system have been investigated in some literatures (cf. [19,20,21]). The research investigations conducted by Li et al. [20] and Shi et al. [21] are primarily based on the original model established in [19], where it was assumed that the fear factor impacts the entire growth rate of the prey species. In [20], a one-parameter saddle-node bifurcation and degenerate Hopf bifurcation, as well as two-parameter cusp bifurcation and Bogdanov-Takens bifurcation, were studied in a predator-prey system with the Leslie-Gower functional response. In [21], the existence and non-existence of non-constant positive steady states and codimension one Hopf bifurcation were investigated in a diffusive predator-prey model with multiple Allee effects and the fear factor. However, they didn't consider the combination effect of Allee and fear effects on Turing instability and pattern selection.

    Recently, one-parameter bifurcation phenomena, including those within a predator-prey model with the Allee effect and the fear effect, were investigated by Lai et al. [22]. In this model, the Allee effect was suggested to be an additive, and the fear factor was assumed to influence the entire growth rate of the prey species. Their findings revealed that the density of predator species decreased as the fear factor increased. Notably, the manner in which the fear factor affected prey differed from previously documented approaches in the existing literature [19,22]. In actuality, field experiments have demonstrated that the fear factor leads to a reduction in prey species production [8]. Considering the long-term shift in prey populations, several experimental observations [9,23,24] indicated that the fear factor significantly impacted interactions between prey and predator species. Frightened song sparrow nestlings, for instance, were unable to survive due to a scarcity of food supplies caused by fear. Long-term evolution resulted in substantial changes in both behaviour and physiology driven by fearful predation rather than actual predation. Prey species, in an effort to minimize predation rates, would migrate from high-risk to low-risk zones. Examples of significantly reduced fear-driven reproduction rates include mule deer versus mountain lions, elk versus wolves, snowshoe hares versus dogs [26] and dugongs versus sharks [25], among others. Wang et al. [9] developed a mathematical model to investigate the reduction of the reproduction rate in the presence of fear and to explore the stability mechanism of the system. Subsequently, the influence of fear in predator-prey models, along with the inclusion of prey refuge, was studied by Wang et al. [27] and Samaddar et al. [28]. In [29], authors explored the combined effects of the Allee effect, fear factor and prey refuge on the dynamics of a fractional order prey-predator system. Liu, in the context of a spatiotemporal prey-predator model that incorporated fear and Allee effects, examined the existence and non-existence of non-constant positive solutions and bifurcations, including Hopf and steady-state bifurcations [30]. Most recently, Pal et al. investigated pattern formation in a cross-diffusive Leslie-Gower predator-prey model with fear and Allee effects [31], although they did not delve into pattern selection and codimension two bifurcations.

    The novelty of this study lies in the comprehensive exploration of the bifurcation phenomena, encompassing transcritical bifurcation, saddle-node bifurcation, and Hopf bifurcation, each of codimension one, extending to cusp and Bogdanov-Takens bifurcations of codimension two. These bifurcations are induced by both the Allee effect and the fear factor (refer to Theorems 3.3, 3.5, 3.6 and 3.9 below). Additionally, we investigated the Turing instability mechanism for the spatiotemporal system and interpreted the transition of patterns among three classical types: Hexagonal patterns, stripe patterns, and their combinations. It has been observed that the Allee effect plays a crucial role in inducing Turing instability. Furthermore, one may find that a small Allee effect factor stabilizes the considered system, whereas a large Allee effect factor destabilizes the unique positive coexistence steady state by inducing Turing instability (see Theorem 4.1 below), steady-state bifurcation (see Theorem 4.2 below), and spatiotemporal Hopf bifurcation (see Theorem 4.3 below). Given the significance of these factors, the primary objectives of this investigation revolve around ecological concerns:

    ● How does the Allee effect and fear factor influence the dynamics of the temporal/spatiotemporal system?

    ● What are the parametric criteria for stability and instability for both the spatiotemporal interacting species?

    ● Does intra-specific competition among predators encourage reduced biodiversity?

    ● What effects do the codimension two cusp and Bogdanov-Takens bifurcations around the coexistence equilibrium have on the environment?

    ● Is the Allee effect a key mechanism for inducing Turing instability, and if so, how does it impact the spatially heterogeneous distribution of species?

    A thorough analysis of stability and bifurcation is successfully carried out in the proposed model system in order to address all the issues mentioned above. The present article is organized as follows: Section 2 deals with the fundamental properties regarding global classical solvability, positivity, and uniformly boundedness of the spatiotemporal model entities for a very generic situation. The qualitative behavior of the temporal model with linear functional response is highlighted in Section 3 in terms of the occurrence of various bifurcations. Subsequently, the spatiotemporal model is analyzed and the corresponding numerical simulations are performed to illustrate the effectiveness of the theoretical findings in Section 4. Finally, the concluding remarks of the entire outcomes of the present model system are included in the concluding Section 5.

    This section consists of some basic properties such as positivity, uniqueness, and boundedness of solutions of the spatiotemporal model in order to verify that the considered model is well-posedness.

    Let ρ(n,). It is known that W1,ρ(Ω,R2) is continuously embedded in C(ˉΩ,R2). One may then consider the solution of system (1.6) in S={ωW1,ρ(Ω,R2)|wν=0Ω}. In view of Amann's theory (see Theorems 1, 14.4, 14.6, 14.7 and 17.1) [32,33,34]) and comparison principle of the parabolic equation, one can easily have the following:

    Theorem 2.1. Assume that all system parameters are positive and (H0) is satisfied. Furthermore, suppose that the initial data (u0(x),v0(x))S for some ρ>n and is nonnegative. Then there exists a maximal time Tmax such that the considered system (1.6) has a unique bounded nonnegative solution defined on Ω×(0,Tmax) fulfilling (u,v)C([0,Tmax),S)C2,1(ˉΩ×(0,Tmax),R2). Furthermore, it exists globally in time.

    Proof. Since the diffusion coefficient matrix D=diag{du,dv} is a positive definite and diagonal matrix, (1.6) is normally elliptic [32]. Since gC1, then the local existence, uniqueness, and regularity of the solution follows directly according to Theorem 14.4 and Theorem 14.6 [34]. Since u0(x)0 and v0(x)0, by the comparison principle of parabolic equation, u(x,t)0 and v(x,t)0 for all (x,t)Ω×[0,Tmax). To show the global existence, one needs to prove the boundedness of u(,t)L(Ω) and v(,t)L(Ω) for t[0,Tmax) according to Amann's theory (see Theorems 5.2 and 17.1) [33,34]. Since the non-negativity of u, v and g(u,v), one must conclude from the first equation of (1.6) that

    utau1+kvbucu2+du2u(ab)ucu2+du2u, (2.1)

    which implies that u(x,t)max{supΩ(u0(x),abc)}Mu for all (x,t)Ω×[0,), according to the comparison principle of the parabolic equation. On the other hand, since g is non-negative continuous and u is non-negative and bounded, we conclude that there always exists a constant C0>0 such that

    ef1(u,v)+f2(u,v)C0(1uv)

    for u,v0. Since u is a priori bounded, it follows from Theorem 1 [35] that there exists a constant Mv>0 such that v(x,t)Mv for all (x,t)Ω×[0,Tmax). Therefore, the solution of (1.6) exists globally in time. This completes the proof of the theorem.

    In this section, for convenience of analysis, it is chosen g(u,v)u, then temporal model (1.5) turns into

    {dudt=au1+kvbucu2muu+hpuv,t>0,dvdt=epuvdvsv2,t>0,u(0)=u00,v(0)=v00. (3.1)

    By using the following non-dimensional transformations

    ˜u=cua,˜v=cvea,˜t=at,

    one gets the corresponding non-dimensionalized form (after dropping tildes) given as follows:

    {dudt=u1+αvβuu2γuu+ρηuv,t>0,dvdt=ηuvδvσv2,t>0,u(0)=U00,v(0)=V00, (3.2)

    where U0=cu0a, V0=cv0ea, and

    α=kaec,β=ba,γ=mca2,ρ=cha,η=epc,δ=da,σ=esc.

    Equilibrium points are the of intersection of the zero growth prey curve and zero growth predator curve in the non-negative quadrant of the uv-plane. The biologically significant equilibrium points of (3.2) must satisfy the following equations

    {F1(u,v):=u1+αvβuu2γuu+ρηuv=0,F2(u,v):=ηuvδvσv2=0. (3.3)

    By solving the second equation of (3.3), we obtain the non-negative roots v=0, v=1σ(δ+ηu), provided u>δη. Clearly, (3.2) has the trivial equilibrium point E0=(0,0).

    To obtain the existence of arbitrary semi-trivial equilibrium point (u,0), substituting v=0 into the first equation of (3.3) yields the following equation with respect to u

    u2+A1u+A2=0, (3.4)

    where A1=β+ρ1 and A2=γ(1β)ρ. By analyzing the distribution of positive root of (3.4), one may obtain that the existence of the semi-trivial equilibrium point (u,0) has the following cases:

    Case-ⅰ: If A2<0, that is, γ<(1β)ρ, then (3.4) has a unique positive real root, u1=A1+A214A22, which means that (3.2) has only one semi-trivial equilibrium point E1=(u1,0);

    Case-ⅱ: If A2=0, and A1<0, that is, γ=(1β)ρ, and β+ρ<1, then (3.4) has a unique positive root u2=A1. Therefore, (3.2) has only one semi-trivial equilibrium point E2=(u2,0);

    Case-ⅲ: If A2>0, A1<0 and A214A2>0, that is, γ>(1β)ρ, β+ρ<1 and (βρ)2+1>2(βρ+2γ), then (3.4) has two positive roots as u3,4=A1±A214A22, and thus (3.2) has two semi-trivial equilibrium points E3=(u3,0) and E4=(u4,0);

    Case-ⅳ: If A2>0, A1<0 and A214A2=0, that is, γ>(1β)ρ, β+ρ<1 and (βρ)2=2(βρ+2γ)1, then (3.4) has only one positive real root, u5=A12. Therefore, (3.2) has a unique semi-trivial equilibrium point E5=(u5,0).

    To show the existence of the coexistence equilibrium point, substituting v=1σ(ηuδ) into the first equation of (3.3) and simplifying it, we obtain the following cubic equation of one variable:

    u3+3a1u2+3a2u+a3=0, (3.5)

    where,

    a1=σ2+η3αρ+σαρη2η2αδ+σαβη+ση2σαδ3(σαη+αη3),a2=2η2αρδαβσδησδ+ρη2σ+βσ2+αγσησ2+αβρση+αηδ2+ρσ2αρσδ3(σαη+αη3),a3=αηρδ2αγσδρησδ+βρσ2αβρσδρσ2+γσ2σαη+αη3.

    Let us choose z=u+a1, then (3.5) takes the form as follows:

    f(z):z3+b1z+b2=0, (3.6)

    with, b1=3(a2a21),b2=2a313a1a2+a3. Now, by analyzing the distribution of the positive roots of f(z), we obtain the distribution of coexistence equilibrium points of (3.2), which can be described as the following lemma.

    Lemma 3.1. For the proposed model system (3.2) , the number of coexistence equilibrium points can be described as follows:

    (i) If Δ=b3127+b224>0 and a3<0, then (3.2) has a unique coexistence equilibrium point given by E=(u,v) provided u<δη.

    (ii) If Δ=b3127+b224<0 and b1<0, then (3.2) has at most three coexistence equilibrium points denoted by, E1=(u1,v1), E2=(u2,v2) and E3=(u3,v3), more precisely,

    (b1:1:) E1,E2 and E3 coexist if a3<0, a1+a21a2<0 holds.

    (b2:2:) Only two of E1,E2 and E3 coexist if one of the following conditions attached here follows:

    (b2121) a3>0, a1a21a2<0 and (b2222) a3=0, a1+a21a2<0.

    (b3:3:) Only one of E1,E2 and E3 exists if one of the following conditions holds: (b3131) a3<0, a1+a21a2<0 and (b3232) a3=0, a1>0, a1a21a2<0.

    (iii) If Δ=b3127+b224=0 and b1<0, then the model system (3.2) has at most two positive equilibrium points, namely, E4=(u4,v4),E5=(u5,v5) or ˆE4=(ˆu4,ˆv4),ˆE5=(ˆu5,ˆv5). More specifically,

    (c1:1:) Let us suppose that b22(a21a2)32=0, then both E4:=E1=E2 and E5:=E3 coexists if a2>0 and 23a2<a1<a2 hold, where E4 has the multiplicity two, and only E4 exists if one of the following conditions holds: (c1111) a2>0, a123a2, (c1212) a2<0.

    (c2:2:) Let us consider b2+2(a21a2)23=0, then both ˆE4 and ˆE5 coexists if a2>0 and a1+a2<0 hold, where ˆE4 has the multiplicity two, and only ˆE5 exists if one of the following conditions holds: (c2121) a20, a21+a220, (c2222) a1>23a2.

    Proof. Denote by Δ=(b13)3+(b22)2=b3127+b224 the discriminant of cubic equation (3.6). By employing Cardano's formula, and Descartes' rule of signs [36], we have three cases:

    (i) If Δ>0, i.e., b3127+b224>0, then the cubic equation (3.6) has a unique real root denoted by z. For this case (3.5) has a unique real root u, which is positive if a3<0 follows. Furthermore, according to the Cardano's formula, the unique positive real root u is given by

    u=3b22+b224+b3127+3b22b224+b3127a1.

    (ii) If Δ<0 and b1<0, i.e., b3127+b224<0 and a21a2>0, according to the Cardano's formula, then (3.5) has three distinct real roots given by, u1,u2 and u3, respectively, where

    u1=2a21a2cosθ2π3a1,u2=2a21a2cosθ3a1,u3=2a21a2cosθ+2π3a1,θ=arccos(b22(a21a2)32). (3.7)

    More precisely, since θ[0,π], one obtains (u1)min=a1a21a2, (u2)min=a1+a21a2, and (u3)min=a12a21a2. (u1)max=a1+a21a2, (u2)max=a1+2a21a2, and (u3)max=a1a21a2. Thus, one can conclude the following facts.

    If a1+2a21a2<0 then (u1)min>0, (u2)min>0, and (u3)min>0, which implies that (b1) is satisfied;

    If a1a21a2<0, then (u2)min>0, thus u2>0. Since a3>0, it follows from Descartes' rule of signs that (3.5) has two positive real roots, which implies (b21) holds. On the other hand, if a3=0, then (3.5) has at most two positive real roots. Since a1+a21a2<0, it follows that (u1)min>0 and (u2)min>0, which manifests that (b22) holds.

    If a3<0, then (3.5) has at least one positive real roots. Since a1+a21a2<0, one obtains a1>0 and a2>0. Thus, (3.5) has exactly one positive real root. If a3=0, a1>0 and a1a21a2<0, one obtains a2<0, and thus, (3.5) has only one positive real root, which implies that the conclusion (b3) holds.

    (iii) If Δ=0 and b1<0, i.e., b3127+b224=0 and b1<0, then (3.5) has three real roots, one of them a double roots denoted by E4 and the other denoted by E5. More specifically, according to Cardano's formula, we have u4:=u1=u2=a21a2a1, u5:=u3=2a21a2a1 when we set b22(a21a2)32=0 in formula (3.7). Also, it is to be noted that u4>u5, so u4 and u5 are positive if a1+2a21a2<0 and a21a2>0 hold. These two conditions are identical to a2>0 and 23a2<a1<a2, whereas only u4 is positive when a2>0, a1+23a20, or a2<0, which implies that (c1) holds.

    When b2+2(a21a2)32=0 in (3.7), one has ˆu4:=u1=u3=a21a2a1, ˆu5:=u2=2a21a2a1. Since, ˆu5>ˆu4, it is to be noted that if a2>0, a1+a2<0, then one has that both ˆu5 and ˆu4 are positive. Moreover, there are two sets of conditions, a20,a21+a220 and a123a2>0, that can result in ˆu5>0, and thus, the conclusion (c2) holds.

    Based on the analysis above, one may easily prove Lemma 3.1, which ends the proof of the theorem.

    Remark 3.2. Lemma 3.1 provides an upper bound estimation on the number of coexistence equilibrium points E(u,v). One has to impose the condition uδη to ensure the existence of the exact number of coexistence equilibrium points. Additionally, one may obtain the range of original model parameters, which ensures the existence of coexistence equilibrium points by substituting the coefficients a1,a2 and a3 of (3.5) into the expressions b1 and b2.

    In Figure 1, graphical representation of the system equilibria are provided for various system parameter values through Lemma 3.1. Especially, this implies that the Allee effect and fear factor can affect the number of interior equilibrium points, as well as the results in a complex distribution of the population in the habitat.

    Figure 1.  Nullclines representation of (3.2) for the system parameter values (i) δ=0.04-no interior equilibrium points; δ=0.06-unique interior equilibrium point; δ=0.05-two interior equilibrium points. (ii) β=0.075-no interior equilibrium points; β=0.01-unique interior equilibrium point; β=0.055-two interior equilibrium points. (iii) γ=0.235-no interior equilibrium points; γ=0.21-unique interior equilibrium point; γ=0.225-two interior equilibrium points. (iv) ρ=0.045-no interior equilibrium points; ρ=0.08-unique interior equilibrium point; ρ=0.055-two interior equilibrium points. The remaining system parameter values for each case are chosen from the following set of parameter values α=0.38,β=0.01,γ=0.21,η=0.3,σ=0.2,ρ=0.08,δ=0.06. Here, the solid blue line is the prey nullcline and the red dashed line is the predator nullcline.

    This subsection consists of various dynamical behavior of the temporal model system (3.2) around the trivial equilibrium point E0 as well as the semi-trivial equilibrium points Ei for i=1,2,3,4,5, respectively. The Jacobian matrix of (3.2) at any points E(u,v) in the uv-plane is given by

    JE=[11+αvβ2uγu+ρ+γu(u+ρ)2ηvuα(1+αv)2ηuηvδ2σv+ηu][J11J12J21J22].

    Theorem 3.3. The trivial equilibrium point E0=(0,0) is

    (i) a hyperbolic attractor if γ>ρ(1β),

    (ii) a hyperbolic saddle point if γ<ρ(1β).

    Proof. The proof is trivial, so it is omitted here.

    By direct calculation, we have the following:

    Theorem 3.4. The model system (3.2) undergoes a transcritical bifurcation around the trivial equilibrium point E0=(0,0) as the system parameter γ crosses the threshold value γ=γ[TC]=ρ(1β), provided β+ρ<1.

    Proof. Since the occurrence of transcritical bifurcation around the trivial equilibrium point E0=(0,0), one of the eigenvalues of the corresponding Jacobian matrix must be zero, which implies that detJE0=0. One may easily verify that at the critical system parameter value γ[TC]=ρ(1β), we get detJE0=0.

    Let V and W be the eigenvectors of the Jacobian matrix JE0 and its transpose matrix JTE0, respectively, corresponding to the simple zero eigenvalue. Then, at the critical parameter value γ[TC]=ρ(1β), we obtain the eigenvectors as V=(v1,v2)=(1,0)T and W=(w1,w2)=(1,0)T. Now, we compute the transversality conditions Δ1,Δ2 and Δ3 as defined below:

    Δ1:WT[Fγ(0,0;γ[TC])]=(w1,w2)T(F1γF2γ)E0=(1,0)(00)=0,
    Δ2:WT[DFγ(0,0;γ[TC])(V)]=(w1,w2)T(ρ(u+ρ)2000)E0(v10)=ρ(u+ρ)20,
    Δ3:WT[D2F(0,0;γ[TC])(V,V)]=(w1,w2)T(2F1u2v21+22F1uvv1v2+2F1v2v222F2u2v21+22F2uvv1v2+2F2v2v22)E0=(1,0)(2F1u2v212F2u2v21)E0=(1,0)((2+2γρ2)v210)=2ρ(ρ+β1)v21.

    Thus, if β+ρ<1 then one obtains Δ3:WT[D2F(0,0;γ[TC])(V,V)]0. Therefore, (3.2) undergoes a transcritical bifurcation around the trivial equilibrium point E0=(0,0) whenever the parameter γ attains the threshold value γ[TC]=ρ(1β), according to Sotomayor's theorem [37].

    Direct applications of Routh-Hurwitz criteria give the following:

    Theorem 3.5. The semi-trivial equilibrium points Ei=(ui,0) with i=1,2,3,4,5 are locally asymptotically stable if γ<(ui+ρ)2 and ui<δη. Additionally, each Ei=(ui,0) for i=1,2,3,4,5 is saddle point if γ>(ui+ρ)2 and ui<δη or γ<(ui+ρ)2 and ui>δη.

    Theorem 3.6. The model system (3.2) undergoes a transcritical bifurcation around the semi-trivial equilibrium point ˉE=(ˉu,0) for the critical system parameter value γ=γ[TC] when γ satisfies δηˉu=0 under the two parametric restrictions

    γ(ˉu+ρ)2andγ(η(η+α)+σ)(ˉu+ρ)2σ, (3.8)

    where ˉu is determined by (3.4).

    Proof. To obtain the transcritical bifurcation requires that one of the eigenvalues of the corresponding Jacobian matrix at the semi-trivial equilibrium point ˉE must be zero, which implies that det(JˉE)=0. We accordingly obtained the critical system parameter value δ=δ[TC]=ηˉu.

    Next we checked Sotomayor's condition [37]. Let A and B be the eigenvectors of the Jacobian matrix JˉE and its transpose matrix JTˉE, respectively, corresponding to the simple zero eigenvalue. At the critical system parameter value γ=γ[TC], we obtain the eigenvectors A and B as A=(A1,A2)T(ˉJ12,ˉJ11)T and B=(B1,B2)T(0,ˉJ12)T. Now, we compute the three numerical values Δ1,Δ2 and Δ3 as follows:

    Δ1:BT[Fδ(ˉu,0;δ[TC])]=(B1,B2)(F1δF2δ)ˉE=(B1,B2)(00)=0,
    Δ2:BT[DFδ(ˉu,0;δ[TC])(A)]=(0,B2)(0001)(A1A2)=A2B2,
    Δ3:BT[D2F(ˉu,0;δ[TC])(A,A)]=(B1,B2)(2F1u2A21+22F1uvA1A2+2F1v2A222F2u2A21+22F2uvA1A2+2F2v2A22)ˉE=2A2B2(ηA1σA2).

    So, if γ(ˉu+ρ)2, then we have Δ2:BT[DFδ(ˉu,0;δ[TC])(A)]0; if σηA1A2, that is, γ(η(η+α)+σ)(ˉu+ρ)2σ, then we have Δ3:BT[D2F(ˉu,0;δ[TC])(A,A)]0. Hence, a transcritical bifurcation takes place into the system surrounding the semi-trivial equilibrium point ˉE=(ˉu,0) whenever the system parameter δ crosses the threshold value δ[TC]=ηˉu under the pre-assumptions as defined in (3.8).

    Remark 3.7. Biological interpretation: From Theorem 3.3 to Theorem 3.6, we obtained that large Allee effects are prone to causing both the prey and predator population extinction, and small Allee effects can only cause the predator species' extinction. However, the prey species u persists even when the prey species density is low.

    This section serves various types of bifurcation phenomena and takes place in (3.2) around the coexistence equilibrium point E(u,v) for some specific system parameter values. We showed that the temporal system undergoes codimension one bifurcations including saddle-node bifurcation, Hopf-bifurcation and codimension two cusp and Bogdanov-Takens bifurcations. As we know, at the coexistence equilibrium point E=(u,v), the characteristic equation of linearized system is given by

    λ2Tλ+D=0, (3.9)

    where T=tr(JE)=J11+J22 and D=det(JE)=J11J22J12J21 with

    J11=u+γu(u+ρ)2,J12=αu(1+αv)2ηu,J21=ηv,J22=σv. (3.10)

    Remark 3.8. System (3.2) may exhibit bi-stability between the trivial equilibrium E0 and the coexistence equilibrium E. For example, if one takes a set of system parameters as α=0.03,β=0.01,γ=0.21,ρ=0.08,η=0.3,δ=0.06 and σ=0.2, then it is easy to check that γρ(1β)=0.1292>0, tr(JE(0.3453,0.4579))=0.036<0 and det(JE(0.3453,0.4579))=0.0105>0 are satisfied. So, the trivial equilibrium E0 and the coexistence equilibrium E are all stable in this case.

    The following theorem follows directly from the Hopf bifurcation theorem [38].

    Theorem 3.9. The model system (3.2) exhibits a Hopf-bifurcation and gives birth of a limit cycle around the coexistence equilibrium point E=(u,v) at the critical system parameter value γ=γ[HB] provided by T(γ[HB])=0, D(γ[HB])>0 and ˙T(γ[HB])=dTdγ|γ=γ[HB]0.

    This section deals with the determination of the stability and direction of Hopf-bifurcation, which takes place in (3.2) around the coexistence equilibrium point E=(u,v), through the computation of the first Lyapunov number. For this purpose, here we will follow the normal form technique [36]. Let us choose the following transformation Q1=uu and Q2=vv, which shift the coexistence equilibrium point E=(u,v) to the origin. Using this transformation, (3.2) may be written as:

    dQ1dt=Q1+u1+α(Q2+v)β(Q1+u)(Q1+u)2γ(Q1+u)Q1+u+ρη(Q1+u)(Q2+v),dQ2dt=δ(Q2+v)σ(Q2+v)2+η(Q1+u)(Q2+v). (3.11)

    By Taylor's series expansion of the right hand side of (3.11) about the origin, and we obtain:

    ˙Q1=c10Q1+c01Q2+c20Q21+c11Q1Q2+c02Q22+c30Q31+c21Q21Q2+c12Q1Q22+c03Q32+o(||Q||4),˙Q2=d10Q1+d01Q2+d20Q21+d11Q1Q2+d02Q22+d30Q31+d21Q21Q2+d12Q1Q22+d03Q32+o(||Q||4), (3.12)

    where Q=(Q1,Q2)T and the coefficients cij,dij; i,j=0,1,2,3, are given by

    c10=11+αvβ2uγu+ρ+γu(u+ρ)2ηv,c01=uα(1+αv)2ηu,d01=δ2σv+ηu,d10=ηv,c20=2+2γρ(u+ρ)3,c11=α(1+αv)2η,c02=2α2u(1+αv)3,c30=6γρ(u+ρ)4,c21=0,c12=2α2(1+αv)3,d12=0,c03=6α3u(1+αv)4,d20=0,d11=η,d02=2σ,d30=0,d21=0,d03=0.

    System (3.12) may be rewritten as

    ˙Q=JEQ+Φ(Q), (3.13)

    where Φ=(c20Q21+c11Q1Q2+c02Q22+c30Q31+c12Q1Q22+c03Q32d11Q1Q2+d02Q22)=(Φ1Φ2). According to the argument in [7], after transformation, we get that (3.13) is equivalent to the following

    (˙y1˙y2)=(0ωω0)(y1y2)+(N1(y1,y2;σ=σ[HB])N2(y1,y2;σ=σ[HB])),

    where, the terms N1 and N2 are given by;

    N1(y1,y2;σ=σ[HB])=1c01Φ1,N2(y1,y2;σ=σ[HB])=1c01ω(c10Φ1+c01Φ2),

    with,

    Φ1=(c20c201c11c01c10+c02c210)y21+(2c02c10ωc11c01ω)y1y2+c02ω2y22+(c30c301+c12c01c210c03c310)y31+(2c12c01c10ω3c03c210ω)y21y23c03c10ω2y21y2c03ω3y32,Φ2=(d02c210d11c01c10)y21+(2d02c10ωd11c01ω)y1y2+d02ω2y22.

    Now, to observe the stability behavior of the bifurcating limit cycle, one may calculate the first Lyapunov number L1 by computing the following formula:

    L1=116[N1111+N1122+N2112+N2222]+116ω[N112(N111+N122N212(N211+N222)N111N211+N122N222)],

    where, Nkij=2Nkyiyj|(y1,y2;σ)=(0,0;σ[HB]), Nkijl=3Nkyiyjyl|(y1,y2;σ)=(0,0;σ[HB]). Therefore, depending upon the positivity of L1, one may consider the following cases in order to determine the stability-instability behavior of the limit cycle:

    Case-ⅰ: If L1>0(<0), then a sub-critical (super-critical) Hopf-bifurcation takes place into (3.2) or an unstable (stable) limit cycle will be created surrounding the coexistence equilibrium point E=(u,v).

    Case-ⅱ: Whenever the first Lyapunov number L1 vanishes, then a local bifurcation of co-dimension two, namely, generalized Hopf bifurcation or Bautin bifurcation appears into (3.2) about the coexistence equilibrium point E=(u,v).

    In this subsection, we studied the Hopf bifurcation numerically. The parameter values provided for numerical simulations of (3.2) are: α=0.38,β=0.01,γ=0.21,ρ=0.08,η=0.3,δ=0.06 and σ=0.2. To solve the nonlinear system (3.2) numerically we have used MATLAB R2016a by constructing a suitable code. In Figure 2, each system parameter is chosen as a bifurcation parameter to illustrate the occurrence of Hopf bifurcation, respectively. All numerical simulations suggest that Hopf bifurcations are subcritical in nature, which can be interpreted by the positive first Lyapunov numbers (Figure 2(ⅰ) with L1=2.901765; Figure 2(ⅱ) with L1=2.918016; Figure 2(ⅲ) with L1=2.848071; Figure 2(ⅳ) with L1=2.905451; Figure 2(ⅴ) with L1=2.940986; Figure 2(ⅵ) with L1=2.803728; Figure 2(ⅶ) with L1=2.863622).

    Figure 2.  Existence of limit cycle for (3.2) at the critical parameter value (i) γ=γ[HB]=0.211421 around the coexistence equilibrium point E=(0.352582,0.228873), (ii) δ=δ[HB]=0.058389 around the coexistence equilibrium point E=(0.350447,0.233725), (iii) α=α[HB]=0.402793 around E=(0.351249,0.226873), (iv) β=β[HB]=0.014360 around E=(0.351251,0.226876), (v) ρ=ρ[HB]=0.077660 around E=(0.353390,0.230085), (vi) σ=σ[HB]=0.193995 around E=(0.351248,0.233894), (vii) η=η[HB]=0.303690 around E=(0.350605,0.232377). The remaining parameter values are provided in α=0.38,β=0.01,γ=0.21,ρ=0.08,η=0.3,δ=0.06,σ=0.2 for each case.

    In this section, one may discuss the existence of saddle-node bifurcation in (3.2) around the coexistence equilibrium point E=(u,v) for the critical system parameter value γ=γ[SN]. For this purpose, here one may use Sotomayor's theorem [37].

    Theorem 3.10. The model system (3.2) may exhibit a saddle-node bifurcation around the coexistence equilibrium point E=(u,v) whenever the system parameter γ attains the critical value γ=γ[SN] and T0, where

    γ[SN]=(σ+2σαv+σα2v2+ηα+η2+2η2αv+η2α2v2)(u+ρ)2(1+αv)2σ,

    and T is specified in the proof of the theorem.

    Proof. It is easy to check that at the critical system parameter value γ=γ[SN] the determinant of the corresponding Jacobian matrix JE vanishes, that is, det(JE)=0, which implies that one of the eigenvalues of JE is equal to zero. Next, we checked the transversality condition. Let ˆV and ˆW be the eigenvectors of the Jacobian matrix JE and its transpose matrix JTE respectively corresponding to the simple zero eigenvalue. Then, at the critical system parameter value γ[SN], one may calculate the eigenvectors as

    ˆV=(J12J11)(ˆv1ˆv2),ˆW=(J21J11)(ˆw1ˆw2).

    Next, we calculated the transversality conditions Γ1 and Γ2 as follows:

    Γ1:ˆWT[Fγ(E;γ[SN])]=(ˆw1,ˆw2)(F1γF2γ)[E,γ[SN]]=(ˆw1,ˆw2)(uu+ρ0)=ηuvu+ρ0.

    Now, to compute the second transversality condition, one needs to compute the second order derivatives at the coexistence equilibrium point E=(u,v) as follows:

    2F1u2=2+2γρ(u+ρ)3,2F1uv=α(1+αv)2η,2F1v2=2α2u(1+αv)3,2F2u2=0,2F2uv=η,2F2v2=2σ.
    Γ2:ˆWT[D2F(E;γ[SN])(ˆV,ˆV)]=(ˆw1,ˆw2)(2F1u2ˆv21+22F1uvˆv1ˆv2+2F1v2ˆv222F2u2ˆv21+22F2uvˆv1ˆv2+2F2v2ˆv22),=ˆw1(2F1u2ˆv21+22F1uvˆv1ˆv2+2F1v2ˆv22)+ˆw2(2F2u2ˆv21+22F2uvˆv1ˆv2+2F2v2ˆv22)T.

    If T0 by Sotomayor's theorem, one may conclude that (3.2) undergoes a saddle-node bifurcation around the coexistence equilibrium point E=(u,v) whenever the system parameter γ attains the critical value γ[SN].

    Example 3.11. Now, to verify the analytical findings of the occurence of saddle-node bifurcation around the coexistence equilibrium point, here an attempt is made to preform a numerical example. For this purpose, the following set of system parameter values are taken into account; α=0.38,β=0.01,ρ=0.08,η=0.3,σ=0.2 and δ=0.06. Then, at the critical system parameter value γ[SN]=0.227779, (3.2) has a coexistence equilibrium point E=(0.258868,0.088301). Also, at that critical value the eigenvectors ˆV and ˆW are given by ˆV=(0.169746,0.254621)T,ˆW=(0.024690,0.254621)T, where T represents the transpose of matrix. Next, we calculated the transversality conditions Γ1 and Γ3 as follows:

    ˆWT[Fγ(E;γ[SN])]=0.0202360,T=ˆWT[D2F(E;γ[SN])(ˆV,ˆV)]=0.0021210.

    Hence, both of the transversality conditions for the occurence of saddle-node bifurcation holds good. Thus, one may conclude that (3.2) undergoes a saddle-node bifurcation around the coexistence equilibrium point E=(0.258868,0.088301) whenever the system parameter γ crosses the threshold value γ[SN]=0.227779 (cf. Figure 3).

    Figure 3.  Identification of saddle-node bifurcation in (3.2) for the specific system parameter value γ[SN]=0.227779 with the coexistence equilibrium point E=(0.258868,0.088301). The remaining system parameter values are α=0.38,β=0.01,ρ=0.08,η=0.3,σ=0.2 and δ=0.06.

    When in a system the saddle-node bifurcation curve and the Hopf bifurcation curve meet tangentially, at that point a local bifurcation of codimension two, namely a Bogdanov-Takens bifurcation takes place into the system [38]. Here, it is observed that, (3.2) exhibits Hopf-bifurcation for the critical system parameter value σ=σ[HB] surrounding the coexistence equilibrium point E=(u,v). Also, (3.2) undergoes a saddle-node bifurcation around the coexistence equilibrium point E=(u,v) whenever the system parameter γ attains the threshold value γ[SN]. Subsequently, a Bogdanov-Takens bifurcation takes place into (3.2) around the coexistence equilibrium point E=(u,v).

    Theorem 3.12. The temporal system (3.2) experiences a Bogdanov-Takens bifurcation surrounding the coexistence equilibrium E=(u,v) for the critical parameter value (σ,γ)=(σ[BT],γ[BT]) provided the following conditions must be followed:

    BT.11tr(JE;σ=σ[BT],γ=γ[BT])=0,det(JE;σ=σ[BT],γ=γ[BT])=0,BT.22d11(ϵ)+2c20(ϵ)0,d20(ϵ)0,BT.33det((ϑ1,ϑ2)(ϵ1,ϵ2)|(ϵ1,ϵ2)=(0,0))0,

    where c20(ϵ),d11(ϵ)d20(ϵ), ϑ1(ϵ), and ϑ2(ϵ) are given in the proof of the theorem. Then, there exists a parameter dependent nonlinear smooth invertible variable transformation, and a direction preserving time reparametrization, which together reduce (3.2) to the following equivalent norm form

    {dξ1dt=ξ2,dξ2dt=ϑ1+ϑ2ξ1+ξ12+sξ1ξ2, (3.14)

    where ϑ1 and ϑ2 are defined in (A.17), and s=±1.

    Proof. For the detailed proof please see the Appendix.

    Without loss of generality, let us start with the negative value of s, that is, let s=1, then for (3.2) there exists a neighbourhood of (ϑ1,ϑ2)=(0,0)R2, which divides the bifurcation plane into four regions through the following curves according to [38]:

    (i)SN+={(ϑ1,ϑ2):ϑ22=4ϑ1;ϑ2<0}, (3.15)
    (ii)SN={(ϑ1,ϑ2):ϑ22=4ϑ1;ϑ2>0}, (3.16)
    (iii)H={(ϑ1,ϑ2):ϑ1=0,ϑ2>0}, (3.17)
    (iv)HL={(ϑ1,ϑ2):ϑ1=625ϑ22+o(ϑ22),ϑ2<0}. (3.18)

    Here, SN represents the saddle-node bifurcation curve with two branches SN+ and SN, corresponding to the sign of ϑ2 such as ϑ2<0 and ϑ2>0, respectively, H represents the Hopf-bifurcation curve and HL represents the Homoclinic bifurcation curve. In addition, the argument against the case for the positive value s, that is, for s=1 is very much identical to the argument in favor of s=1. For this case the local representation of bifurcation curves in a very small neighborhood of (ϑ1,ϑ2)=(0,0) were generated using the following linear transformation given by (ξ1,ξ2,t,ϑ1,ϑ2)=(ξ1,ξ2,t,ϑ1,ϑ2).

    In this subsection, we devote to perform some numerical simulations to illustrate the effectiveness of Theorem 3.12. Without loss of generality, we choose γ and σ as the bifurcation parameters, and take a set of parameter values as follows:

    α=0.38,β=0.01,ρ=0.08,η=0.3,δ=0.06,σ=1.324208,γ=0.268561.

    Under this set of parameters, (3.2) has the interior equilibrium point E(0.4030,0.0459) and the Jacobian matrix associated with E has two zero eigenvalues. We now perturb the parameters γ and σ in the small neighbourhood of 1.324208 and 0.268561, respectively, that is, let σ=0.268561+ϵ1 and γ=1.324208+ϵ2, where ϵ1 and ϵ2 are small perturbation parameters. Then, we get the system corresponding to the form (A.9) as follows

    {dˉu1dt=G10+m10ˉu1+m01ˉv1+m20ˉu21+m11ˉu1ˉv1+m02ˉv21+ˉH1(ˉu1,ˉv1),dˉv1dt=H10+n10ˉu1+n01ˉv1+n20ˉu21+n11ˉu1ˉv1+n02ˉv21+ˉH2(ˉu1,ˉv1), (3.19)

    where G10=0.8345ϵ2, H10=0.0021ϵ1, m10=0.06090.3429ϵ2, n_{10} = 0.0138 , m_{01} = -0.2688 , n_{01} = -0.0609-0.092\epsilon_1 , m_{20} = -1.6187+1.4198*\epsilon_2 , m_{11} = -0.6671 , m_{02} = 0.1105 , n_{20} = 0 , n_{11} = 0.3 , and n_{02} = -0.537122-2\epsilon_1 .

    Since d_{11}(\epsilon^*)+2c_{20}^{*} = -5.0018\neq 0 and d_{20}(\epsilon^\ast) = -00892\neq 0 , we used the formulae presented in Section 3.3.5 to calculate \gamma_1(\epsilon), \gamma_2(\epsilon), \mathcal{D}(\epsilon) , and \mathcal{E}(\epsilon) . Then, we obtained the following norm form of the Bogdanov-Takens bifurcation corresponding to (A.16) by using sign calculation of MATLAB R2016a software:

    \begin{equation} \left\{\begin{array}{llll} \frac{d P}{d \tau} & = & Q, \\ \frac{d Q}{d \tau} & = & \gamma_{1}(\epsilon)+\gamma_{2}(\epsilon)P+\mathcal{D}(\epsilon)P^{2}+\mathcal{E}(\epsilon)PQ+H(P, Q), \end{array}\right. \end{equation} (3.20)

    where

    \begin{equation*} \begin{split} \gamma_1(\epsilon) = &\epsilon_1^2(0.1213\epsilon_2^3-0.084159\epsilon_2^2+0.014259\epsilon_2-0.000013488)-\epsilon_1(1.382\epsilon_2^4-1.5727\epsilon_2^3+0.42065\epsilon_2^2\\ &+0.053947\epsilon_2-0.00056868)+5.2189\times 10^{-15}-0.050817\epsilon_2-0.40822\epsilon_2^2+2.6904\epsilon_2^3-\\ &0.26569\epsilon_2^4+3.229\epsilon_2^5+O(||\epsilon||^6), \\ \gamma_2(\epsilon) = &\epsilon_1(9.617\epsilon_2^4-8.3014\epsilon_2^3+6.3419\epsilon_2^2-0.57442\epsilon_2+0.0011016)+\epsilon_1^2(-1.3453\epsilon_2^3+0.51871\epsilon_2^2-\\ &0.049249\epsilon_2+0.0042726)-\epsilon_1^3(-0.037202\epsilon_2^2+0.0048059\epsilon_2-4.5382\times 10^{-6})-1.1869\times 10^{-8}\\ &-1.0527\epsilon_2+10.3\epsilon_2^2-12.975\epsilon_2^3+1.5915\epsilon_2^4-21.546\epsilon_2^5+O(||\epsilon||^6), \end{split} \end{equation*}
    \begin{equation*} \begin{split} \mathcal{D}(\epsilon) = &\epsilon_1^4(0.00040496\epsilon_2-3.817\times 10^{-7})-\epsilon_1^3(0.10135\epsilon_2^2-0.011885\epsilon_2+0.0014745)-\epsilon_1^2(-6.6109\epsilon_2^3+\\ &2.6944\epsilon_2^2-0.31387\epsilon_2+0.023075)-\epsilon_1(38.14\epsilon_2^4-18.077\epsilon_2^3+26.184\epsilon_2^2- 4.2265\epsilon_2+0.13818)-\\ &0.089171+6.0452\epsilon_2-42.631\epsilon_2^2+15.871\epsilon_2^3-21.281\epsilon_2^4+30.583\epsilon_2^5+O(||\epsilon||^6), \\ \mathcal{E}(\epsilon) = &15.186\epsilon_2 + \epsilon_1(5.102\epsilon_2 - 0.67797) + 0.35954\epsilon_2^2 - 5.0018. \end{split} \end{equation*}

    Then, the generic norm form of B-T bifurcation for (3.2) near the origin for small \|\epsilon\| is locally topologically equivalent to the system:

    \begin{equation} \left\{\begin{array}{llll} \frac{d \xi_{1}}{d t} & = & \xi_{2}, \\ \frac{d \xi_{2}}{d t} & = & \vartheta_{1}+\vartheta_{2}\xi_{1}+{\xi_{1}}^{2}+s\xi_{1}\xi_{2}, \end{array}\right. \end{equation} (3.21)

    where \vartheta_1(\epsilon) and \vartheta_2(\epsilon) can be obtained with the expression of \gamma_1(\epsilon), \gamma_2(\epsilon), \mathcal{D}(\epsilon) , and \mathcal{E}(\epsilon) given in (3.20), s = sign\left(\frac{\theta}{d_{20}(\epsilon^\ast)}\right) = 1 , since \det\left(\frac{\partial(\vartheta_1, \vartheta_2)}{\partial(\epsilon_1, \epsilon_2)}\big{|}_{(\epsilon_1, \epsilon_2) = (0, 0)}\right) = 1507084.8078\neq 0, which means that the transversality condition BT. 3 is satisfied. In view of Theorem 3.12, (3.2) undergoes B-T bifurcation at the coexistence equilibrium E^{*} , and the local representation of bifurcation curves in \epsilon_1, \epsilon_2 in the small neighborhood of the origin are given in (3.15)–(3.18). This can be seen from the phase portrait diagrams depicted in Figure 4. Similarly, we obtained different B-T bifurcation phase portraits that occurred in the vicinity (\epsilon_{1}, \epsilon_{2}) of the chosen bifurcating parameters (see Figure 5). The detailed explanation in each figure corresponding to Bogdanov-Takens bifurcation are listed in Tables 1 and 2.

    Figure 4.  Phase portraits of Bogdanov-Takens bifurcation of (3.2) in the neighborhood of the unique coexistence equilibrium point E_{*} = (0.403015, 0.045993) in the uv -plane. The critical bifurcating parameter values are (\gamma^{[BT]}, \sigma^{[BT]}) = (0.268561, 1.324208) . For a small perturbation, \mathrm{(i)} \epsilon_{1} = 0, \epsilon_{2} = 0 , \mathrm{(ii)} \epsilon_{1} = -0.009472, \epsilon_{2} = -0.523889 , \mathrm{(iii)}\, \, \epsilon_{1} = -0.009472, \epsilon_{2} = -0.520889 , \mathrm{(iv)} \epsilon_{1} = -0.01172, \epsilon_{2} = -0.523889 , \mathrm{(v)}\, \, \epsilon_{1} = -0.007825, \epsilon_{2} = -0.523889 , \mathrm{(vi)} \epsilon_{1} = -0.009472, \epsilon_{2} = -0.53489 . The remaining parameter values are \alpha = 0.38, \beta = 0.01, \rho = 0.08, \eta = 0.3, \delta = 0.06 .
    Figure 5.  Phase portraits of Bogdanov-Takens bifurcation of (3.2) in the vicinity of the unique coexistence equilibrium point E_{*} = (0.281675, 0.107900) in the uv -plane. The critical bifurcating parameter values are (\eta^{[BT]}, \sigma^{[BT]}) = (0.818409, 1.580402) . In a small neighborhood (\eta^{[BT]}+\epsilon_{1}, \sigma^{[BT]}+\epsilon_{2}) , where \mathrm{(i)} \epsilon_{1} = 0, \epsilon_{2} = 0 , \mathrm{(ii)} \epsilon_{1} = -0.124587, \epsilon_{2} = -0.4438945 , \mathrm{(iii)}\, \, \epsilon_{1} = -0.126367, \epsilon_{2} = -0.4438945 , \mathrm{(iv)}\, \, \epsilon_{1} = -0.146887, \epsilon_{2} = -0.4438945 , \mathrm{(v)} \epsilon_{1} = -0.10887, \epsilon_{2} = -0.4438945 , \mathrm{(vi)} \epsilon_{1} = -0.124587, \epsilon_{2} = -0.4538945 . The remaining parameter values are \alpha = 0.38, \beta = 0.01, \gamma = 0.21, \rho = 0.08, \delta = 0.06 .
    Table 1.  Schematic summary of the phase portraits in the vicinity of the critical bifurcating parameter values (\sigma^{[BT]}, \gamma^{[BT]}) = (0.268561, 1.324208) illustrated through Figure 4.
    Value of Value of Equilibrium point Types of Phase
    \epsilon_{1} \epsilon_{2} Coexistence/interior Equilibrium point portrait
    0 0 E_{*}=(0.4030, 0.0459) Cusp of codimension two; Figure 4(ⅰ)
    Bogdanov-Takens bifurcation occurs
    -0.523889 -0.009472 E_{*}^{'}=(0.4003, 0.0751) Limit cycle creats surrounding E_{*}^{'} , Figure 4(ⅱ)
    E_{*}^{''}=(0.3522, 0.0571) while E_{*}^{''} is a saddle point
    -0.520889 -0.009472 E_{*}^{'}=(0.4018, 0.0754) Homoclinic orbit joining the saddle Figure 4(ⅲ)
    E_{*}^{''}=(0.3512, 0.0565) point E_{*}^{''} to itself surrounds E_{*}^{'}
    -0.523889 -0.01172 E_{*}^{'}=(0.4252, 0.0844) E_{*}^{'} is a spiral sink, while Figure 4(ⅳ)
    E_{*}^{''}=(0.3274, 0.0477) E_{*}^{''} is a saddle point
    -0.523889 -0.007825 Does not exist \ldots Figure 4(ⅴ)
    -0.53489 -0.009472 E_{*}^{'}=(0.3942, 0.0738) E_{*}^{'} is a spiral source, while Figure 4(ⅵ)
    E_{*}^{''}=(0.3566, 0.0595) E_{*}^{''} is a saddle point

     | Show Table
    DownLoad: CSV
    Table 2.  Schematic summary of the phase portraits in the neighborhood of (\eta^{[BT]}, \sigma^{[BT]}) = (0.818409, 1.580402) portrayed through Figure 5.
    Value of \epsilon_{1} Value of \epsilon_{2} Equilibrium point Coexistence/interior Types of Equilibrium point Phase portrait
    0 0 E_{*}=(0.2816, 0.1079) Cusp of codimension two; Figure 5(ⅰ)
    Bogdanov-Takens bifurcation occurs
    -0.124587 -0.4438945 E_{*}^{'}=(0.2978, 0.1291) Limit cycle creats surrounding E_{*}^{'} , Figure 5(ⅱ)
    E_{*}^{''}=(0.2597, 0.1058) while E_{*}^{''} is a saddle point
    -0.126367 -0.4438945 E_{*}^{'}=(0.3024, 0.1314) Homoclinic orbit joining the saddle Figure 5(ⅲ)
    E_{*}^{''}=(0.2562, 0.1032) point E_{*}^{''} to itself surrounds E_{*}^{'}
    -0.146887 -0.4438945 E_{*}^{'}=(0.3326, 0.1438) E_{*}^{'} is a spiral sink, while Figure 5(ⅳ)
    E_{*}^{''}=(0.2375, 0.0876) E_{*}^{''} is a saddle point
    -0.10887 0.4438945 Did not exists \ldots Figure 5(ⅴ)
    -0.124587 -0.4538945 E_{*}^{'}=(0.2899, 0.1253) E_{*}^{'} is a spiral source, while Figure 5(ⅵ)
    E_{*}^{''}=(0.2658, 0.1104) E_{*}^{''} is a saddle point

     | Show Table
    DownLoad: CSV

    In this present section we have explored the occurrence of various codimension one bifurcations, namely the Hopf-bifurcation, transcritical bifurcation and saddle-node bifurcation for some specific system parameter value surrounding the ecologically meaningful equilibrium point of (3.2). In Figure 6, we have illustrated the one-parameter bifurcation diagrams with respect to the system parameter as specified in the x -axis of each figure [39].

    Figure 6.  One parameter bifurcation diagrams corresponding to the system parameter as specified in the x -axis of each figures. The other system parameters used here are provided as (ⅰ) \alpha = 0.38, \beta = 0.01, \rho = 0.08, \eta = 0.3, \delta = 0.06, \sigma = 0.2 ; (ⅱ) \alpha = 0.38, \beta = 0.01, \gamma = 0.21, \rho = 0.08, \eta = 0.3, \sigma = 0.2 ; (ⅲ) \alpha = 0.38, \gamma = 0.21, \rho = 0.08, \eta = 0.3, \delta = 0.06, \sigma = 0.2 . Here, SN represents saddle-node bifurcation point, BP represents transcritical bifurcation point, H represents Hopf bifurcation point.

    In Figure 6(ⅰ) we have observed that a Hopf-bifurcation took place in (3.2) around the coexistence equilibrium point E_{*} = (0.352582, 0.228873) at the system parameter value \gamma^{[HB]} = 0.211421 , and we obtained the corresponding first Lyapunov number as \mathcal{L}_{1} = 2.901765 . Also, saddle-node bifurcation appeared in (3.2) at \gamma^{[SN]} = 0.227779 around E_{*} = (0.258868, 0.088301) . Additionally, (3.2) exhibited a transcritical bifurcation at the critical system parameter value \gamma^{[TC]} = 0.2212 about the axial equilibrium point \bar{E} = (0.2, 0) .

    The bifurcation diagram, with respect to the system parameter \delta as depicted through the Figure 6(ⅱ), illustrated that (3.2) exhibits a Hopf-bifurcation around the coexistence equilibrium point E_{*} = (0.350447, 0.233725) for the critical system parameter value \delta^{[HB]} = 0.058389 . At this critical system parameter value we obtained the first Lyapunov number as \mathcal{L}_{1} = 2.918016 , which implied that the appeared Hopf-bifurcation is subcritical. Additionally, we have observed that a saddle-node bifurcation takes place at (3.2) at the critical system parameter value \delta^{[SN]} = 0.043514 surrounding the coexistence equilibrium point E_{*} = (0.247418, 0.153555) . Moreover, we have observed that the system experiences a transcritical bifurcation about the predator free equilibrium point \bar{E} = (0.178911, 0) .

    From the bifurcation diagram of (3.2), with respect to the system parameter \beta as depicted in Figure 6(ⅲ), we have found that the model system exhibits a Hopf-bifurcation at the critical system parameter value \beta^{[HB]} = 0.014360 about the coexistence equilibrium point E_{*} = (0.351251, 0.226876) . Moreover, a saddle-node bifurcation took place in the system at \beta^{[SN]} = 0.063588 surrounding the equilibrium point E_{*} = (0.244684, 0.067026) . Additionally, the system experienced a transcritical bifurcation about the predator free equilibrium point \bar{E} = (0.2, 0) at \beta^{[TC]} = 0.05 .

    In this present section we have explored the occurrence of codimension two bifurcations, namely Bogdanov-Takens bifurcation, and cusp bifurcation for some specific system parameter value surrounding the ecologically meaningful equilibrium point by depicting the two-parameter bifurcation diagram [39]. It should be noted that we have dealt with a local bifurcation of codimension two, namely the cusp bifurcation in (3.2) due to the variation of two system parameters. It is a well-known fact that when two saddle-node bifurcation curves meet at a point, that is where a cusp bifurcation may take place. Also, one more fact is that when the saddle-node bifurcation curve vanishes on the transcritical bifurcation curve, a cusp bifurcation may arise. This suggests that three equilibrium points may be merged at the cusp bifurcation point due to the variation of two significant system parameters. For more detailed derivation about cusp bifurcation, please refer to Step-1 and Step-2 in the Appendix.

    In Figure 7(ⅰ) we have detected two cusp bifurcation points appearing on the saddle-node bifurcation curve, which indicates that (3.2) exhibits a cusp bifurcation at the critical system parameter values (\eta^{[CP]}, \gamma^{[CP]}) = (0.170821, 0.275520) about the coexistence equilibrium point E_{*} = (0.352761, 0.000693) . The normal form coefficient is given by -0.6960690 [39]. Similarly, at (\eta^{[CP]}, \gamma^{[CP]}) = (0.829643, 0.139782) about the coexistence equilibrium E_{*} = (0.072383, 0.000254) , the normal form coefficient is given by -0.3359687 .

    Figure 7.  Two parameter bifurcation diagrams corresponding to the system parameter as specified in the axis of each figure. The other system parameter values used here are provided as (ⅰ) \alpha = 0.38, \beta = 0.01, \rho = 0.08, \delta = 0.06, \sigma = 0.2 ; (ⅱ) \alpha = 0.38, \beta = 0.01, \gamma = 0.21, \rho = 0.08, \delta = 0.06 ; (ⅲ) \alpha = 0.38, \gamma = 0.21, \rho = 0.08, \eta = 0.3, \delta = 0.06 . Here, BT represents Bogdanov-Takens bifurcation point, CP represents cusp bifurcation point.

    It is evident from Figure 7(ⅱ) that, for the critical system parameter value (\eta^{[BT]}, \sigma^{[BT]}) = (0.818409, 1.580402) , (3.2) exhibits a Bogdanov-Takens bifurcation about the coexistence equilibrium point E_{*} = (0.281675, 0.1079) , which is an intersection point of Hopf bifurcation and saddle-node bifurcation curves. The normal form coefficients are given by (-0.1878629, -2.409671) [39]. Furthermore, a cusp bifurcation took place in the system at the system parameter value (\eta^{[CP]}, \sigma^{[CP]}) = (0.335815, 0.114164) about the equilibrium point E_{*} = (0.179915, 0.002268) , and the normal form coefficient is given by -0.9066523 [39].

    Similarly, a Bogdanov-Takens bifurcation point has been detected in Figure 7(ⅲ) at the system parameter value (\beta^{[BT]}, \sigma^{[BT]}) = (0.142135, 1.559992) . That is, a Bogdanov-Takens bifurcation takes place in the system for this system parameter value about the coexistence equilibrium point E_{*} = (0.351249, 0.029086) , and the normal form coefficient is given by (-0.0408996, -1.974066) . Also, a cusp bifurcation point has been detected in the Figure at the critical system parameter value (\beta^{[CP]}, \sigma^{[CP]}) = (0.05, 0.121539) about the predator free equilibrium point \bar{E} = (0.2, 0) , and the normal form coefficient is given by -0.218866 .

    In this subsection, stability and Turing instability mechanisms are studied. The corresponding non-dimensionalized form of the spatiotemporal model (1.6) with Holling type-Ⅰ functional response (i.e., g(u, v)\equiv 1 ) is given as follows:

    \begin{equation} \left\{\begin{array}{llll} \frac{\partial u({\bf x}, t)}{\partial t} = \frac{u}{1+\alpha v}-\beta u-u^2-\frac{\gamma u}{u+\rho}-\eta uv+\nabla^2 u, & {\bf x}\in\Omega, t > 0, \\ \frac{\partial v({\bf x}, t)}{\partial t} = -\delta v-\sigma v^2+\eta uv+d\nabla^2 v, & {\bf x}\in\Omega, t > 0, \\ u(x, y, 0) = u_0(x, y)\geq 0, \; \; v(x, y, 0) = v_0(x, y)\geq 0, & {\bf x}\in\Omega, \\ \frac{\partial u}{\partial \nu} = \frac{\partial v}{\partial \nu} = 0, & {\bf x}\in\partial\Omega, t > 0. \end{array}\right. \end{equation} (4.1)

    The linearized system of (4.1) can be written as

    \begin{equation} \frac{\partial}{\partial t}\left(\begin{array}{c} u\\ v \end{array}\right) = \left(\begin{array}{cc} \Delta +J_{11}^\ast & J_{12}^\ast\\ J_{21}^\ast & d\Delta+J_{22}^\ast \end{array}\right) \left(\begin{array}{c} u\\ v \end{array}\right)\triangleq L(d)\left(\begin{array}{c} u\\ v \end{array}\right). \end{equation} (4.2)

    Here, J_{ij}^\ast for i, j = 1, 2 are the same as those in (3.10). Let us suppose \Omega = [0, L]\times[0, L] . Then, k^2 = (\frac{m\pi}{L})^2+(\frac{n\pi}{L})^2 with (m, n)\in\mathbb{N}^2\setminus\{(0, 0)\} as the eigenvalues of -\nabla^2 on \Omega under homogeneous Neumann boundary conditions. The corresponding eigenfunctions are given by {\bf{w}}_{m, n}({\bf{x}}) = {\bf{C}}_{m, n}\cos\frac{m\pi x}{L}\cos\frac{n\pi y}{L} , where {\bf{x}} = (x, y) and {\bf{C}}_{m, n} are Fourier coefficients with m and n as integers. Denote

    \begin{equation} L_k(d) = \left(\begin{array}{cc} J_{11}^\ast-k^2 & J_{12}^\ast\\ J_{21}^\ast &J_{22}^\ast-dk^2 \end{array}\right). \end{equation} (4.3)

    It follows that the eigenvalues of L(d) are given by the eigenvalues of L_k(d) . So, the characteristic equation of (4.2) is

    \begin{equation} \lambda^2+A(k^2)\lambda+B(k^2) = 0 \end{equation} (4.4)

    with A(k^2) = (1+d)k^2-(J_{11}^\ast+J_{22}^\ast), \, \, B(k^2) = dk^4-(d J_{11}^\ast +J_{22}^\ast)k^2+J_{11}^\ast J_{22}^\ast-J_{12}^\ast J_{21}^\ast.

    Clearly, if \gamma\leq (u_\ast+\rho)^2 , E_\ast is locally asymptotically stable. In this subsection, one may always assume that \gamma > (u_\ast+\rho)^2 (i.e., J_{11}^{*} > 0 ) is satisfied and investigate the Turing instability mechanism for the considered spatial system. To investigate Turing instability, it is assumed that

    \begin{equation} \mathrm{tr}(J_{E_\ast}): = J_{11}^\ast+J_{22}^\ast < 0\, \, \mathrm{and}\, \, \mathrm{det}(J_{E_\ast}): = J_{11}^\ast J_{22}^\ast-J_{12}^\ast J_{21}^\ast > 0. \end{equation} (4.5)

    are satisfied.

    It is noted that A(k^2) > 0 for all non-negative wavenumber k under assumption (4.5). Hence, Turing instability can only be attained in the case of B(k^2) < 0 for some nonzero wavenumber k , which implies that the coefficient of k^2 must satisfy

    \begin{equation} d J_{11}^\ast+J_{22}^\ast > 0. \end{equation} (4.6)

    It should be noted that B(k^2) > 0 for all non-negative wavenumber k when \gamma = 0 , which implies that there is no Turing instability occurrence in this case although we can require d > 1 , which is a classical Turing instability mechanism. So, here Turing instability may occur only when the Allee effect and spatial diffusion are present, thus, calling this mechanism Allee-effect-driven Turing instability.

    The critical Turing instability condition can be found by imposing the minimum of B(k^2) satisfying \min\limits_{k^2 > 0}B(k^2) < 0 . Since (4.6), one should \min_{k^2 > 0}B(k^2) = B(k_m^2) with k_m^2 = \frac{d J_{11}^\ast+J_{22}^\ast}{2d}. Then \min\limits_{k^2 > 0}B(k^2) < 0 turns into

    \begin{equation} \frac{(dJ_{11}^\ast +J_{22}^\ast)^2}{4d} > J_{11}^\ast J_{22}^\ast-J_{12}^\ast J_{21}^\ast. \end{equation} (4.7)

    In fact, the parameter space defined by conditions (4.5)–(4.7) is called Turing instability space, where one can choose an arbitrary system parameter, including the parameter \gamma , as a bifurcation parameter to obtain Turing instability. To obtain a explicit condition revealing Turing instability, in what follows, we chose the diffusion coefficient d as a bifurcation parameter to derive it. Since

    \begin{equation*} \mathrm{sign}(J(E_\ast)) = \bigg{[}\begin{array}{cc} +& - \\ + & -\end{array}\bigg{]}, \end{equation*}

    one may find the following two threshold values

    \begin{equation} \chi_+ = \frac{J_{11}^\ast J_{22}^\ast-2J_{12}^\ast J_{21}^\ast +\sqrt{\Delta}}{J_{11}^{* 2}}, \quad\chi_- = \frac{J_{11}^\ast J_{22}^\ast-2J_{12}^\ast J_{21}^\ast -\sqrt{\Delta}}{J_{11}^{* 2}} \end{equation} (4.8)

    with \Delta = (2J_{12}^\ast J_{21}^\ast-J_{11}^\ast J_{22}^\ast)^2-J_{11}^{\ast 2} J_{22}^{\ast 2} > 0 . Moreover, one must d J_{11}^\ast+J_{22}^\ast = 0 when d = -\frac{J_{22}^\ast}{J_{11}^\ast}\triangleq d_\ast . Furthermore, one gets 0 < \chi_- < d_\ast < \chi_+ . It is easy to check that \min\limits_{k^2 > 0}B(k^2) < 0 and d J_{11}^\ast+J_{22}^\ast > 0 hold simultaneously when d > \chi_+\triangleq d_c , which means E_\ast is spatially unstable and Turing instability occurs in this case. Moreover, we have \min\limits_{k^2 > 0}B(k^2) > 0 when d < \chi_+ , which means that E_\ast is stable in this case. With these facts in hand, one may immediately have the following:

    Theorem 4.1 (Turing instability induced by Allee effect). Assume that the condition (4.5) and \gamma > (u_\ast+\rho)^2 are satisfied. Then, system (4.1) undergoes Turing instability and spatiotemporal forms at E_\ast when d > d_c: = \chi_+ . Furthermore, Turing bifurcation occurs at d = d_c .

    In Figure 8, we have plotted the critical d_c values against the fear effect \alpha with different Allee effect threshold values. We have also plotted the temporal Hopf bifurcation curves for the two different Allee effect threshold values, and these two kinds of curves divide the plane into four regions in each case: Stable region (SR), Hopf region (HR), Turing region (TR), and Turing-Hopf instability region (THR). Interestingly, from this figure, we find the d_c and \alpha^{[HB]} values decrease as the Allee effect factor increases, which means the Allee effect can affect Turing instability region or THR region. Next, we shall investigate pattern selection in the TR region and pattern formation in HR and THR regions, respectively.

    Figure 8.  Turing-Hopf bifurcation diagram in d - \alpha plane for different Allee effect threshold values. Here the other system parameters are \beta = 0.048, \gamma = 0.0826, \rho = 0.135, \eta = 0.8, \delta = 0.012, \sigma = 0.128 . The Hopf bifurcation threshold values \alpha^{[HB]} are 0.3254 and 0.316 for \gamma = 0.0826 and \gamma = 0.0829 , respectively.

    Another way to obtain Turing bifurcation is through the steady-state solution bifurcation theory [40,41]. In what follows, we shall choose the diffusion ratio d as a bifurcation parameter to investigate steady-state bifurcation near the positive steady state E_\ast . For this purpose, we rewrite A(k^2) and B(k^2) as A(k^2, d) and B(k^2, d) , respectively. In view of the theory frame in Section three [41], the sufficient condition for the occurrence of the steady state bifurcation is that there exists a non-zero wave number k > 0 and d_S > 0 such that

    \begin{equation} B(k^2, d^S) = 0, \quad A(k^2, d^S)\neq 0, \quad \mathrm{and}\quad A(j^2, d^S)\neq 0, \quad B(j^2, d^S)\neq 0\quad \mathrm{for}\, j\neq k \end{equation} (4.9)

    and

    \begin{equation} \frac{\partial B(k^2, d)}{\partial d}\Big{|}_{d = d^S}\neq 0. \end{equation} (4.10)

    In view of the discussion above, we immediately have:

    Theorem 4.2. Suppose that (4.5) is satisfied. Furthermore, suppose that

    \begin{equation*} d^S\triangleq d_k = \frac{\mathrm{det}(J_{E_\ast})-J_{22}^{*}k^2}{k^2(J_{11}^{*}-k^2)} > 0\quad \mathrm{and}\, \, d_k\neq d_j\quad \mathrm{for}\, \, \mathrm{any}\, \, j\neq k. \end{equation*}

    Then, (d_k, u_\ast, v_\ast) is a bifurcation point where a smooth curve \Gamma of the non-constant positive steady-state solutions bifurcates from the line of positive constant steady-state solution (u_\ast, v_\ast) , and \Gamma is contained in a connected component \mathcal{C} of the set of non-zero steady-state solutions of (4.1) in \mathbb{R}_+\times S_+ . Moreover, either \mathcal{C} is unbounded in \mathbb{R}_+\times S_+ , or \mathcal{C}\bigcap\, (\partial\mathbb{R}_+\times S_+)\neq \emptyset , or \mathcal{C} contains another bifurcation point (d_j, u_\ast, v_\ast) with d_j\neq d_k . More precisely, \Gamma is locally a curve near (d_k, u_{*}, v_{*}) in the form of \Gamma = \{(d(s), u(s), v(s)):s\in(-\epsilon, \epsilon)\} , where u(s) = u_{*}+sa_k\cos\frac{m\pi x}{L}\cos\frac{n\pi y}{L}+s\psi_1(s), v(s) = v_\ast+sb_k\cos\frac{m\pi x}{L}\cos\frac{n\pi y}{L}+s\psi_2(s) for s\in(-\epsilon, \epsilon) , and d:\, (-\epsilon, \epsilon)\rightarrow \mathbb{R} , (\psi_1, \psi_2):\, (-\epsilon, \epsilon)\rightarrow \mathcal{Z} are C^1 functions such that d(0) = d_k, \psi_1(0) = 0, \psi_2(0) = 0 . Here \mathcal{Z} = \Big{\{}(u, v)\in S_+\Big{|}\int_\Omega uu_{mn}+vv_{mn}dxdy = 0\Big{\}} with u_{mn} = a_k\cos\frac{m\pi x}{L}\cos\frac{n\pi y}{L}, v_{mn} = b_k\cos\frac{m\pi x}{L}\cos\frac{n\pi y}{L} , and a_k, b_k satisfy L_k(d_k)(a_k, b_k)^T = (0, 0)^T .

    Proof. From d^S = d_k = \frac{\mathrm{det}(J_{E_\ast})-J_{22}^{*}k^2}{k^2(J_{11}^{*}-k^2)} , we immediately have B(k^2, d^S) = B(k^2, d_k) = 0 , but B(j^2, d_k)\neq 0 for j\neq k . Clearly, A(k^2, d_k) = (1+d_k)k^2-(J_{11}^{*}+J_{22}^{*}) > 0 under the assumption (4.5) for all wavenumbers k . Furthermore, \frac{\partial B(k^2, d)}{\partial d}\Big{|}_{d = d^S} = k^2(k^2-J_{11}^{*})\neq 0 . Therefore, the conditions (4.9) and (4.10) are satisfied, which ends the proof of the theorem.

    In view of [41], the sufficient condition for the occurrence of Hopf bifurcation is that there exists a non-zero wavenumber k > 0 and d_H > 0 such that

    \begin{equation} A(k^2, d^H) = 0, \quad B(k^2, {d^H}) > 0, \quad \mathrm{and}\quad A(j^2, {d^H})\neq 0, \quad B(j^2, {d^H})\neq 0\quad \mathrm{for}\, j\neq k \end{equation} (4.11)

    and

    \begin{equation} \frac{\partial Re(\lambda(k^2, d))}{\partial d}\Big{|}_{d = {d^H}}\neq 0. \end{equation} (4.12)

    For any Hopf bifurcation point d^H , \alpha(\lambda)\pm \omega(\lambda) are the eigenvalues of L_k(d) , thus, (4.12) becomes

    \begin{equation} \alpha^\prime(d^H) = \frac{A^\prime(k^2, d)}{2}\Big{|}_{d = d^H} = \frac{k^2}{2}\neq 0. \end{equation} (4.13)

    On the other hand, from A(k^2, d^H) = 0 , one may have

    \begin{equation} d^H = \frac{J_{11}^{*}+J_{22}^{*}-k^2}{k^2} \end{equation} (4.14)

    when k\in(0, \sqrt{J_{11}^\ast+J_{22}^\ast}) with J_{11}^\ast+J_{22}^\ast > 0, which requires that J_{11}^{*} > 0. Clearly, A(k^2, d^H) = 0 and A(j^2, d^H)\neq 0 for j\neq k . Now, we only need to verify whether B(j^2, d^H)\neq 0 for k\in(0, \sqrt{J_{11}^\ast+J_{22}^\ast}) , and in particular, B(k^2, d^H) > 0 . Since

    \begin{equation*} \begin{split} \label{expression_of_B(k, d)} B(k^2, d^H) = &d^Hk^4-(d^HJ_{11}^\ast+J_{22}^\ast)k^2+J_{11}^{*}J_{22}^{*}-J_{12}^{*}J_{21}^{*}\\ = &k^2(J_{11}^{*}+J_{22}^{*}-k^2)-J_{11}^{*}(J_{11}^{*}+J_{22}^{*}-k^2)-J_{22}^{*}k^2+J_{11}^{*}J_{22}^{*}-J_{12}^{*}J_{21}^{*}\\ = & -k^4+2k^2J_{11}^{*}-J_{11}^{*^2}-J_{12}^{*}J_{21}^{*}, \end{split} \end{equation*}

    and \Delta_k = 4J_{11}^{*^2}-4(J_{11}^{*^2}+J_{12}^{*}J_{21}^{*}) = -4J_{12}^{*}J_{21}^{*} > 0 , it follows that B(k^2, d^H) > 0 for

    \begin{equation} k\in(k^{-}, k^{+}): = (J_{11}^{*}-\sqrt{-J_{12}^{*}J_{21}^{*}}, \, \, J_{11}^{*}+\sqrt{-J_{12}^{*}J_{21}^{*}}). \end{equation} (4.15)

    Suppose that

    \begin{equation} k^+ < \sqrt{J_{11}^{*}+J_{22}^{*}}, \end{equation} (4.16)

    then

    \begin{equation*} \begin{split} \label{expression_of_B(j, d)} B(j^2, d^H) = &d^Hj^4-(d^HJ_{11}^{*}+J_{22}^{*})j^2+J_{11}^{*}J_{22}^{*}-J_{12}^{*}J_{21}^{*}\\ \geq & d^Hj^4-\Big{(}\Big{(}\frac{L^2(J_{11}^\ast+J_{22}^\ast)}{\pi^2}-1\Big{)}J_{11}^{*}+J_{22}^{*}\Big{)}j^2+J_{11}^{*}J_{22}^{*}-J_{12}^{*}J_{21}^{*}\\ > & J_{11}^{*}J_{22}^{*}-J_{12}^{*}J_{21}^{*}-\frac{L^2(J_{11}^\ast+J_{22}^\ast)}{\pi^2}J_{11}^{*}j^2+\Big{(}\frac{J_{11}^{*}+J_{22}^{*}}{k_+^2}-1\Big{)}j^4. \end{split} \end{equation*}

    Set the quadratic function g(z) = J_{11}^{*}J_{22}^{*}-J_{12}^{*}J_{21}^{*}-\frac{L^2(J_{11}^\ast+J_{22}^\ast)}{\pi^2}J_{11}^{*}z+\Big{(}\frac{J_{11}^{*}+J_{22}^{*}}{k_+^2}-1\Big{)}z^2 , then it is positive for j\in\mathbb{R} if

    \begin{equation} \frac{L^4(J_{11}^{*}+J_{22}^{*})^2J_{11}^{*^2}}{\pi^4}-4(J_{11}^{*}J_{22}^{*}-J_{12}^{*}J_{21}^{*})\Big{(}\frac{J_{11}^{*}+J_{22}^{*}}{k_+^2}-1\Big{)} < 0. \end{equation} (4.17)

    We summarize the discussion above as following.

    Theorem 4.3. Suppose that J_{11}^{*}+J_{22}^{*} > 0 , k^+ < \sqrt{J_{11}^{*}+J_{22}^{*}} and (4.17) holds. Then, for any k\in(0, \sqrt{J_{11}^{*}+J_{22}^{*}})\cap(k^-, k^+) and k^{-} , k^+ are defined as in (4.15), the R-D system (4.1) undergoes a Hopf bifurcation at d = d^H = \frac{J_{11}^{*}+J_{22}^{*}-k^2}{k^2} and the bifurcating periodic solutions from d = d^H are spatially inhomogeneous.

    To interpret pattern selection among spot pattern, stripe pattern, and the mixture of them, we need to derive amplitude equations near the onset d = d_c . Following the method employed in [42], we obtain the amplitude equations as follows:

    \begin{equation} \left\{\begin{array}{llll} \tau_0\frac{\partial A_1}{\partial t} = \mu A_1+h_0\overline A_2\overline A_3-[g_1|A_1|^2+g_2(|A_2|^2+|A_3|^2)]A_1, \\ \tau_0\frac{\partial A_2}{\partial t} = \mu A_2+h_0\overline A_1\overline A_3-[g_1|A_2|^2+g_2(|A_1|^2+|A_3|^2)]A_2, \\ \tau_0\frac{\partial A_3}{\partial t} = \mu A_3+h_0\overline A_1\overline A_2-[g_1|A_3|^2+g_2(|A_1|^2+|A_2|^2)]A_3, \\ \end{array}\right. \end{equation} (4.18)

    where

    \begin{equation} \begin{split} &\tau_0 = -\frac{\varphi+\psi}{d_ck_c^2\psi}, \; \; \mu = \frac{d-d_c}{d_c}, \; g_1 = -\frac{G_1}{d_ck_c^2\psi\varphi^2}, \ g_2 = -\frac{G_2}{d_ck_c^2\psi\varphi^2}, \\ &h_0 = -\frac{c_{20}\varphi^2+2c_{11}\varphi+c_{02}+\psi(2d_{11}\varphi+d_{02})}{d_ck_c^2\psi\varphi}, \;G_1 = I_1+\psi J_1, \; G_2 = I_2+\psi J_2 \end{split} \end{equation} (4.19)

    with I_1 = -(u_{11}+u_{00})(\varphi c_{20}+c_{11})-(\varphi c_{11}+c_{02})(v_{11}+v_{00})-\frac{3}{2}c_{12}\varphi-\frac{1}{2}c_{30}\varphi^3-\frac{1}{2}c_{03}, \; I_2 = -(u_\star+u_{00})(\varphi c_{20}+c_{11})-(\varphi c_{11}+c_{02})(v_\star+v_{00})-3c_{12}\varphi-c_{30}\varphi^3-c_{03}, \; J_1 = -(u_{11}+u_{00})d_{11}-(\varphi d_{11}+d_{02})(v_{11}+v_{00})-\frac{3}{2}d_{12}\varphi, \; J_2 = -(u_{\star}+u_{00})d_{11}-(\varphi d_{11}+d_{02})(v_{\star}+v_{00})-3d_{12}\varphi . Here, \varphi, \psi, u_{00}, v_{00}, u_{11}, v_{11}, u_\star and v_\star possess the similar expressions as those in Section 4.3.1 [42]. Denote

    \begin{equation} \mu_1 = \frac{-h_0^2}{4(g_1+2g_2)}, \; \mu_2 = 0, \;\mu_3 = \frac{h_0^2g_1}{(g_2-g_1)^2}, \;\mu_4 = \frac{h_0^2(2g_1+g_2)}{(g_2-g_1)^2}. \end{equation} (4.20)

    Clearly, \mu_1 < \mu_2 < \mu_3 < \mu_4 under the assumption of g_1 > 0 and g_2 > 0 . Then, we have some results about the stability of the hexagonal pattern, the stripe pattern, and the mixture of them according to the range of \mu in Theorem 3 [42].

    In this subsection, we performed some numerical simulations to show the pattern formation over two dimensional spatial domain according to the obtained theoretical result (see Theorem 4.1), and indicated some interesting pattern solutions in either the Hopf instability region or Turing-Hopf region besides Turing region (see Figure 8). We solved the reaction-diffusion system (4.1) numerically by using a five-point finite difference scheme for the Laplacian operator and forward Euler scheme for the temporal part with appropriate care at the boundary to accommodate Neumann boundary conditions. All the programs for solving the spatiotemporal system (4.1) were performed by MATLAB R2016a. In addition, to ensure the stability of numerical method, the choice of time step size \Delta t , the space step sizes \Delta x and \Delta y and the diffusion coefficient d should satisfy the Courant-Friedrichs-Lewy stability criterion, i.e., \max\{d, 1\}\Delta t\bigg{(}\frac{1}{(\Delta x)^2}+\frac{1}{(\Delta y)^2}\bigg{)}\leq \frac{1}{2} . In addition, we have checked that the numerical simulation results for some other smaller values of the time and space step sizes and they remain the same qualitatively.

    In Figure 9, we observed pattern selections among the hexagonal pattern, stripe pattern and a mixture of hexagons and stripes for different diffusion coefficients with a set of fixed system parameters, where \alpha = 0.32, \beta = 0.048, \gamma = 0.0826, \rho = 0.135, \eta = 0.8, \delta = 0.012 and \sigma = 0.128 . Here, the initial condition is chosen as a random perturbation around the spatially homogeneous steady state E_\ast . Under this set of parameters, we obtained that the one of the Hopf bifurcation threshold values is \alpha = \alpha^{[HB]} = 0.3254 and Turing bifurcation threshold value is d = d_c = 37.9638 . In view of (4.19), we get the coefficients of the amplitude equations (4.18) are h_0 = -25.0863 < 0 and g_2 = 262.3196 > g_1 = 7.7015 > 0 . Moreover, we have \mu_1 = -0.2955 , \mu_2 = 0 , \mu_3 = 0.0748 and \mu_4 = 2.6959 , which will not change with the variation of diffusion coefficient d . So, when d = 40.7 , we obtained \mu = 0.0721\in(\mu_2, \mu_3) , which means the hexagonal pattern prevails over the whole domain as time goes forward (see Figure 9(ⅰ)). When d = 150 , we obtained \mu = 2.9511\in(\mu_4, \infty) , which implies that the stripe pattern will be formed as time evolves (see Figure 9(ⅲ)). However, when d = 60 , we obtained \mu = 0.5805\in(\mu_3, \mu_4) . This manifests that the mixture of hexagons and stripes forms over the two dimensional spatial domain (see Figure 9(ⅱ)).

    Figure 9.  Pattern selection of the prey species u for different diffusion coefficients in the super-critical bifurcation case 0 < g_1 < g_2 with random initial condition. The other system parameters used here are provided as \alpha = 0.32, \beta = 0.048, \gamma = 0.0826, \rho = 0.135, \eta = 0.8, \delta = 0.012, \sigma = 0.128 for each case. Here the space stepsizes are chosen as \Delta x = \Delta y = 1 and the time step size is chosen as \Delta t = 0.001 and d_c = 37.9638 . The parameter set is located in the TR (see, e.g., Figure 8 as a reference).

    In Figure 10, we took another set of system parameters as \alpha = 0.4, \beta = 0.042, \gamma = 0.1, \rho = 0.12, \eta = 0.86, \delta = 0.06 and \sigma = 0.18 with variational diffusion coefficients d = 55, 58, 62 in the TR. The initial condition was chosen as a random perturbation around the coexistence equilibrium. One of the critical Hopf bifurcation values for the temporal system and the critical Turing bifurcation value for the spatial system are \alpha^{[HB]} = 0.4334 and d_c = 53.6568 , respectively. We have plotted the spatiotemporal bifurcation diagram for Turing bifurcation and Hopf bifurcation (see Figure 10(ⅰ)). A direct calculation gives g_1 = -425.7471 < 0 and g_2 = -853.6242 < 0 , which suggests the amplitude equations (4.18) undergoes sub-critical bifurcation. Interestingly, we observed only stable hexagonal pattern solution prevails over the spatial domain as time evolved, although the diffusion coefficients are different (see Figure 10(ⅱ)(ⅳ)).

    Figure 10.  Turing-Hopf bifurcation diagram and coldspot pattern solutions of the prey species u for different diffusion coefficients in the sub-critical bifurcation case g_1 < 0 and g_2 < 0 with random initial condition. The other system parameters used here are provided as \alpha = 0.4, \beta = 0.042, \gamma = 0.1, \rho = 0.12, \eta = 0.86, \delta = 0.06, \sigma = 0.18 for each case. Here the space stepsizes are chosen as \Delta x = \Delta y = 1 and the time step size is chosen as \Delta t = 0.004 and d_c = 53.6568 . The parameter set is located in the TR.

    Apart from the stationary pattern solution, we are now interested in pattern dynamics in HR and THR regions, which are non-Turing regions. In Figure 11, the system parameters remain consistent with those depicted in Figure 10, with the exception of alterations to the fear factor \alpha and the diffusion coefficient d . Under this specific parameter configuration, it is observed that one of the Hopf-bifurcation threshold values is \alpha = \alpha{^{[HB]}} = 0.4334 . The initial choice is made to set \alpha = 0.5 , a value situated within the temporal Hopf instability region. Subsequently, the diffusion coefficient d is selected to be 40 , which falls below the Turing bifurcation threshold value ( d_c = 42.6939 ). This choice therefore positions the system parameters within the pure Hopf instability region, as indicated in Figure 10(ⅰ). Within Figure 11, the spatiotemporal distribution of the prey species u at various time points is demonstrated. It is noteworthy that the initial condition employed in [43] is utilized, elucidating that only a small number of populations from both species are confined to an elliptic domain.

    \begin{equation} \begin{split} u(0, x, y) = \left\{\begin{array}{cc} u_0, & \frac{(x-x_1)^2}{\Delta_{11}^2}+\frac{(y-y_1)^2}{\Delta_{12}^2}\leq 1\\ 0, & \mathrm{otherwise}\\ \end{array}\right.\, v(0, x, y) = \left\{\begin{array}{cc} v_0, & \frac{(x-x_2)^2}{\Delta_{21}^2}+\frac{(y-y_2)^2}{\Delta_{22}^2}\leq 1\\ 0, & \mathrm{otherwise}\\ \end{array}\right. \end{split} \end{equation} (4.21)

    Here, u_0 = 0.8 and v_0 = 0.2 , x_1 = 153.5, y_1 = 145, x_2 = y_2 = 150, \Delta_{11}^2 = \Delta_{12}^2 = 12.5, \Delta_{21}^2 = 5, \Delta_{22}^2 = 10 . We first observe some circular rings with different densities of the species on it (see Figure 11(ⅰ)) at t = 200 . As time goes forward, the expanding circular rings hit the domain boundary and break into the stripes (see Figure 11(ⅱ), (ⅲ)). With the further advancement of time, the circular rings gradually lose stability and break into some spots and stripes (see Figure 11(ⅳ)(ⅵ)), and eventually settle down to coldspot pattern (see Figure 11(ⅶ)).

    Figure 11.  Spatiotemporal patterns of the prey species u at different moments with initial condition (4.21). The other system parameters used here are provided as \alpha = 0.5, \beta = 0.042, \gamma = 0.1, \rho = 0.12, \eta = 0.86, \delta = 0.06, \sigma = 0.18 and d = 40 for each case. Here, the space step sizes are chosen as \Delta x = \Delta y = 1 and the time step size is chosen as \Delta t = 0.004 and d_c = 42.6939 . The parameter set is located in the HR.

    In Figure 12, we observe a similar spatiotemporal distribution of prey species except the difference of instability mechanism of the circular rings pattern, where the system parameters are chosen the same as those in Figure 11 besides the diffusion coefficient d = 45 , which above the Turing bifurcation threshold value d = d_c = 42.6939 . It is easy to check that here the system parameters are located in THR (see Figure 10(ⅰ)) and the initial conditions are in (4.21). After experiencing the rupture of the circular rings, the final pattern transition state is the mixture pattern of spots and stripes (see Figure 12(ⅶ)). It should be noted that we don't observe spatiotemporal chaotic pattern in THR for the certain system parameters and the special choice of initial condition.

    Figure 12.  Spatiotemporal patterns of the prey species u at different moments with initial condition (4.21). The other system parameters used here are provided as \alpha = 0.5, \beta = 0.042, \gamma = 0.1, \rho = 0.12, \eta = 0.86, \delta = 0.06, \sigma = 0.18 and d = 45 for each case. Here the space step sizes are chosen as \Delta x = \Delta y = 1 and the time stepsize is chosen as \Delta t = 0.004 and d_c = 42.6939 . The parameter set is located in the THR.

    The present article is dedicated to the derivation of comprehensive theoretical outcomes for various types of bifurcations, followed by their substantiation through numerical simulations. Each numerical simulation is underpinned by the selection of distinct system parameters, treating them as bifurcation parameters, thereby showcasing diverse dynamical responses. A particular focus is placed on the intricate influence of the Allee effect on both temporal and spatiotemporal systems.

    For the temporal system, a pivotal finding emerges, indicating that the Allee effect can instigate both codimension-one bifurcations (transcritical, saddle-node and Hopf) and codimension-two bifurcations (cusp and Bogdanov-Takens). This revelation imbues the system with a more intricate dynamical behavior, signifying an enrichment of the dynamics inherent to the model system. Notably, the induced Hopf bifurcations tend to be subcritical, aligning qualitatively with previous observations [9]. Additionally, it established that the temporal system can exhibit bistability between the trivial equilibrium E_0 and the coexistence equilibrium E^* (see Remark 3.1).

    For the spatiotemporal model, we investigated the system's dynamics and found that the Allee effect is a key mechanism for inducing Turing instability, and the small Allee effect usually plays a stabilization role (see Theorem 4.1). Furthermore, we have investigated its pattern dynamics and derived amplitude equation near the onset d = d_c . We have studied pattern transition among a hexagonal pattern, stripe pattern and a mixture of them and established an analytical formula to predict it, and we have preformed numerical simulations to illustrate the effectiveness of the theoretical results (see Figure 9). In addition, we found some interesting circular ring patterns in the HR and THR for a special choice of initial condition at the initial moment, which eventually transitioned to coldpattern or the mixture pattern consisting of coldspots and stripes (see Figures 11 and 12). This was different from the observed spatiotemporal chaotic pattern in a spatial slow-fast prey-predator system with the same initial condition [44].

    It was imperative to emphasize that the outcomes of this theoretical investigation hold relevance as long as the fear factor \alpha and the Allee parameter \gamma remained within their practical range of values. Beyond this range, the system experienced permanent collapse due to negative growth in the prey species. Future enhancements of the model could involve the introduction of multiplicative or double Allee effects, offering potential for exciting developments in this challenging field of research. These future investigations are anticipated to introduce a fresh dimension to the domain of mathematical ecology.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors are grateful to the anonymous reviewers for their valuable comments and suggestions. The first author gratefully acknowledges the financial support from Anhui Province University Excellent Talents Funding Project (Grant No. gxbjZD2021075) and Anhui Province University Research Project (Grant No. 2023AH051549). The last author is supported by Peak Discipline Construct Project of Zhejiang University of Science and Technology.

    The authors declare there is no conflict of interest.

    Step-1: (Derivation of non-degeneracy condition)

    Now, first of all it will be shown that the coexistence equilibrium point E_{*} = (u_{*}, v_{*}) is a cusp of codimension two. For this purpose, one may choose the following transformation A_{1} = u-u_{*}, A_{2} = v-v_{*} , which shifts the coexistence equilibrium point E_{*} into the origin. By power series expansion of the modified system into second order terms, one obtains

    \begin{equation} \left\{\begin{array}{llll} \frac{d A_{1}}{d t} & = & p_{10}A_{1}+p_{01}A_{2}+p_{20}A_{1}^{2}+p_{11}A_{1}A_{2}+p_{02}A_{2}^{2}+o(||A||^{3}), \\ \frac{d A_{2}}{d t} & = & q_{10}A_{1}+q_{01}A_{2}+q_{20}A_{1}^{2}+q_{11}A_{1}A_{2}+q_{02}A_{2}^{2}+o(||A||^{3}), \end{array}\right. \end{equation} (A.1)

    where A = (A_{1}, A_{2}) and the coefficients p_{ij} = c_{ij}, q_{ij} = d_{ij} for i, j = 0, 1, 2, 3, \ldots are all provided in the Subsection 3.3.2 . Therefore, one must have

    \begin{equation} \left\{\begin{array}{llll} p_{10}+q_{01} & = & 0, \\ p_{10}q_{01}-p_{01}q_{10} & = & 0. \end{array}\right. \end{equation} (A.2)

    Now, let us consider the transformation u_{1}^{'} = A_{1}, \; \; u_{2}^{'} = p_{10}A_{1}+p_{01}A_{2}. Using this transformation, (3.2) may be represented as

    \begin{equation} \left\{\begin{array}{llll} \frac{d u_{1}^{'}}{d t} & = & u_{2}^{'}+r_{20}{u_{1}^{'}}^{2}+r_{11}u_{1}^{'}u_{2}^{'}+r_{02}{u_{2}^{'}}^{2}+o(||u^{'}||^{3}), \\ \frac{d u_{2}^{'}}{d t} & = & s_{20}{u_{1}^{'}}^{2}+s_{11}u_{1}^{'}u_{2}^{'}+s_{02}{u_{2}^{'}}^{2}+o(||u^{'}||^{3}), \end{array}\right. \end{equation} (A.3)

    where u^{'} = (u_{1}^{'}, u_{2}^{'}) and the coefficients r_{ij}, s_{ij} with i, j = 0, 1, 2, 3, \ldots are given by

    \begin{eqnarray*} r_{20} & = & \frac{p_{02}p_{10}^{2}}{p_{01}^{2}}-\frac{p_{11}p_{10}}{p_{01}}+p_{20}, \; r_{11} = \frac{p_{11}}{p_{01}}-\frac{2p_{02}p_{10}}{p_{01}^{2}}, \; r_{02} = \frac{p_{02}}{p_{01}^{2}}, s_{02} = \frac{p_{02}p_{10}}{p_{01}^{2}}+\frac{q_{02}}{p_{01}}, \\ s_{20} & = & p_{01}q_{20}+p_{10}(p_{20}-q_{11})-\frac{p_{10}^{2}(p_{11}-q_{02})}{p_{01}}+\frac{p_{02}p_{10}^{3}}{p_{01}^{2}}, \; s_{11} = q_{11}-\frac{2p_{02}p_{10}^{2}}{p_{01}^{2}}+\frac{p_{10}(p_{11}-2q_{02})}{p_{01}}. \end{eqnarray*}

    Step-2: (Introducing a c^{\infty} transformation)

    Now, there must be a c^{\infty} invertible transformation given by

    \begin{equation} \left\{\begin{array}{llll} v_{1}^{'} & = & u_{1}^{'}+\frac{1}{2}\bigg{(}\frac{p_{10}p_{02}}{p_{01}^{2}}-\frac{p_{11}+q_{02}}{p_{01}}\bigg{)}{u_{1}^{'}}^{2}-\frac{p_{02}}{p_{01}^{2}}u_{1}^{'}u_{2}^{'}, \\ v_{2}^{'} & = & u_{2}^{'}+\bigg{(}\frac{p_{10}^{2}p_{02}}{p_{01}^{2}}-\frac{p_{10}p_{11}}{p_{01}}+p_{20}\bigg{)}{u_{1}^{'}}^{2}-\bigg{(}\frac{p_{10}p_{02}}{p_{01}^{2}}+\frac{q_{02}}{p_{01}}\bigg{)}u_{1}^{'}u_{2}^{'}. \end{array}\right. \end{equation} (A.4)

    This transformation reduced (A.3) to the following one:

    \begin{equation} \left\{\begin{array}{llll} \frac{d v_{1}^{'}}{d t} & = & v_{2}^{'}+o(||v^{'}||^{3}), \\ \frac{d v_{2}^{'}}{d t} & = & \mu_{1}{v_{1}^{'}}^{2}+\mu_{2}v_{1}^{'}v_{2}^{'}+o(||v^{'}||^{3}), \end{array}\right. \end{equation} (A.5)

    where v^{'} = (v_{1}^{'}, v_{2}^{'}) and the coefficients \mu_{1}, \mu_{2} are given by

    \begin{eqnarray*} \mu_{1} = s_{20}, \mu_{2} = -\frac{p_{10}}{p_{01}}(p_{11}+2q_{02})+2p_{20}+q_{11}. \end{eqnarray*}

    Step-3: (Introduction of a new transformation)

    Now, let us introduce a new transformation in a very small neighborhood of v^{'} = 0 with the form

    \begin{equation} \left\{\begin{array}{llll} \varkappa_{1} & = & v_{1}^{'}, \\ \varkappa_{2} & = & v_{2}^{'}+o(||v^{'}||^{3}). \end{array}\right. \end{equation} (A.6)

    By use of this transformation, (A.5) may be represented as

    \begin{equation*} \left\{\begin{array}{llll} \frac{d \varkappa_{1}}{d t} & = & \varkappa_{2}, \\ \frac{d \varkappa_{2}}{d t} & = & \mu_{1}\varkappa_{1}^{2}+\mu_{2}\varkappa_{1}\varkappa_{2}+o(||\varkappa||^{3}), \end{array}\right. \end{equation*}

    where \varkappa = (\varkappa_{1}, \varkappa_{2}) . Therefore, if \mu_{1}\mu_{2} \neq 0 (non-degeneracy condition) then a cusp bifurcation of codimension two occurs around the unique coexistence equilibrium point E_{*} = (u_{*}, v_{*}) [45].

    Step-4: (Deduction of generic normal form)

    Now, assume that the non-degeneracy condition has been well satisfied. So, one may try to calculate the generic normal form of the Bogdanov-Takens bifurcation according to the frame in [38]. First, we utilize the transformation \bar{u}_1 = u-u_\ast and \bar{v}_1 = v-v_\ast to shift the equilibrium point E_\ast of (3.2) to the origin, and then the system (3.2) becomes

    \begin{equation} \left\{\begin{array}{llll} \frac{d\bar{u}_1}{d t} & = & \frac{\bar{u}_1+u_\ast}{1+\alpha(\bar{v}_1+v_\ast)}-\beta (\bar u_1+u_\ast)-(\bar u_1+u_\ast)^{2}-\frac{\gamma (\bar u_1+u_\ast)}{\bar u_1+u_\ast+\rho}-\eta (\bar u_1+u_\ast)(\bar v_1+v_\ast), \\ \frac{d\bar{v}_1}{d t} & = & -\delta (\bar v_1+v_\ast)-\sigma (\bar v_1+v_\ast)^{2}+\eta (\bar u_1+u_\ast)(\bar v_1+v_\ast). \end{array}\right. \end{equation} (A.7)

    Let us perturb the bifurcating parameters (\sigma, \gamma) in a small neighborhood of \epsilon_{i} in such a way that (\sigma, \gamma) = (\sigma^{[BT]}+\epsilon_{1}, \gamma^{[BT]}+\epsilon_{2}) . Then, we obtain

    \begin{equation} \left\{\begin{array}{llll} \frac{d\bar{u}_1}{d t} & = & \frac{\bar{u}_1+u_\ast}{1+\alpha(\bar{v}_1+v_\ast)}-\beta (\bar u_1+u_\ast)-(\bar u_1+u_\ast)^{2}-\frac{(\gamma^{[BT]+\epsilon_2})(\bar u_1+u_\ast)}{\bar u_1+u_\ast+\rho}-\eta (\bar u_1+u_\ast)(\bar v_1+v_\ast), \\ \frac{d\bar{v}_1}{d t} & = & -\delta (\bar v_1+v_\ast)-(\sigma^{[BT]}+\epsilon_1)(\bar v_1+v_\ast)^{2}+\eta (\bar u_1+u_\ast)(\bar v_1+v_\ast). \end{array}\right. \end{equation} (A.8)

    By expanding the Taylor's series of (A.8) at (\bar u_1, \bar v_1) = (0, 0) up to the terms of second order, we get the following system

    \begin{equation} \left\{\begin{array}{llll} \frac{d \bar{u}_{1}}{d t} & = & G_{10}+m_{10}\bar{u}_{1}+m_{01}\bar{v}_{1}+m_{20}\bar{u}_{1}^{2}+m_{11}\bar{u}_{1}\bar{v}_{1}+m_{02}\bar{v}_{1}^{2}+\bar{H}_{1}(\bar{u}_{1}, \bar{v}_{1}), \\ \frac{d \bar{v}_{1}}{d t} & = & H_{10}+n_{10}\bar{u}_{1}+n_{01}\bar{v}_{1}+n_{20}\bar{u}_{1}^{2}+n_{11}\bar{u}_{1}\bar{v}_{1}+n_{02}\bar{v}_{1}^{2}+\bar{H}_{2}(\bar{u}_{1}, \bar{v}_{1}), \end{array}\right. \end{equation} (A.9)

    where the expression for m_{ij}, n_{ij} ; i, j = 0, 1, 2, 3, \ldots and G_{10}, H_{10} are defined by

    \begin{eqnarray*} G_{10} & = & -\frac{u_\ast}{u_\ast+\rho}\epsilon_2, \, \, H_{10} = -v_{*}^2\epsilon_1, m_{10} = \frac{1}{1+\alpha v_{*}}-\beta-2u_{*}-\frac{(\gamma^{[BT]}+\epsilon_{2})}{u_{*}+\rho}+\frac{(\gamma^{[BT]}+\epsilon_{2})u_{*}}{(u_{*}+\rho)^{2}}-\eta v_{*}, \\ m_{01} & = & -\frac{u_{*}\alpha}{(1+\alpha v_{*})^{2}}-\eta u_{*}, \; m_{20} = -2+\frac{2(\gamma^{[BT]}+\epsilon_{2})\rho}{(u_{*}+\rho)^{3}}, m_{11} = -\frac{\alpha}{(1+\alpha v_{*})^{2}}-\eta, m_{02} = \frac{2\alpha^{2}u_{*}}{(1+\alpha v_{*})^{3}}, \\ n_{10}& = &\eta v_{*}, n_{01} = -\delta-2(\sigma^{[BT]}+\epsilon_{1})v_{*}+\eta u_{*}, n_{20} = 0, n_{11} = \eta, n_{02} = -2(\sigma^{[BT]}+\epsilon_{1}), \end{eqnarray*}

    where the terms \bar{H}_{1}(\bar{u}_{1}, \bar{v}_{1}) and \bar{H}_{2}(\bar{u}_{1}, \bar{v}_{1}) are the power series of \bar{u}_{1} and \bar{v}_{1} and is the form for each of them \bar{u}_{1}^{i}\bar{v}_{1}^{j} ; i+j \geq 3 .

    Step-5: (Introduction of a affine transformation)

    Now, let us introduce a new affine transformation given by

    \begin{eqnarray*} p_{1} & = & \bar{u}_{1}, \; p_{2} = m_{10}\bar{u}_{1}+m_{01}\bar{v}_{1}. \end{eqnarray*}

    By using this affine transformation, (A.9) reduced to the following one:

    \begin{equation} \left\{\begin{array}{llll} \frac{d p_{1}}{d t} & = & G_{10}(\epsilon)+p_{2}+c_{20}(\epsilon)p_{1}^{2}+c_{11}(\epsilon)p_{1}p_{2}+c_{02}(\epsilon)p_{2}^{2}+H_{1}^{'}(p_{1}, p_{2}), \\ \frac{d p_{2}}{d t} & = & H_{10}^{'}(\epsilon)+a_{1}^{'}(\epsilon)p_{1}+b_{1}^{'}(\epsilon)p_{2}+d_{20}(\epsilon)p_{1}^{2}+d_{11}(\epsilon)p_{1}p_{2}+d_{02}(\epsilon)p_{2}^{2}+H_{2}^{'}(p_{1}, p_{2}), \end{array}\right. \end{equation} (A.10)

    where, the coefficients c_{ij}, d_{ij} ; i, j = 0, 1, 2, 3, \ldots , a_{1}^{'}, b_{1}^{'} and H_{10}^{'} are given by

    \begin{eqnarray*} H_{10}^{'}(\epsilon) & = & G_{10}m_{10}+H_{10}m_{01}, \; a_{1}^{'}(\epsilon) = m_{01}n_{10}-m_{10}n_{01}, \; b_{1}^{'}(\epsilon) = m_{10}+n_{01}, \\ c_{20}(\epsilon) & = & m_{20}-\frac{m_{11}m_{10}}{m_{01}}+\frac{m_{02}m_{10}^{2}}{m_{01}^{2}}, \; c_{11}(\epsilon) = \frac{m_{11}}{m_{01}}-\frac{2m_{02}m_{10}}{m_{01}^{2}}, \\ c_{02}(\epsilon) & = & \frac{m_{02}}{m_{01}^{2}}, \; d_{20} = m_{10}m_{20}-\frac{m_{11}m_{10}^{2}}{m_{01}}+\frac{m_{02}m_{10}^{3}}{m_{01}^{2}}+m_{01}n_{20}-m_{10}n_{11}+\frac{n_{02}m_{10}^{2}}{m_{01}}, \\ d_{11} & = & \frac{m_{10}m_{11}}{m_{01}^{2}}-\frac{2m_{02}m_{10}^{2}}{m_{01}^{2}}+n_{11}-\frac{2n_{02}m_{10}}{m_{01}}, \; d_{02} = \frac{m_{10}m_{02}}{m_{01}^{2}}+\frac{n_{02}m_{10}^{2}}{m_{01}}, \end{eqnarray*}

    while the remaining terms H_{1}^{'}(p_{1}, p_{2}) and H_{2}^{'}(p_{1}, p_{2}) are the power series expansion terms of p_{1} and p_{2} of the form p_{1}^{i}p_{2}^{j} ; i+j \geq 3 . Furthermore, for \epsilon^{*} = (0, 0) , we have H_{10}^{'}(\epsilon^{*}) = 0 , a_{1}^{'}(\epsilon^{*}) = 0 , b_{1}^{'}(\epsilon^{*}) = 0 .

    Step-6: (Setting of new variables)

    Now, let us choose a transformation as described by

    \begin{equation*} \left\{\begin{array}{llll} p_{1}^{'} & = & p_{1}, \\ p_{2}^{'} & = & G_{10}(\epsilon)+p_{2}+c_{20}(\epsilon)p_{1}^{2}+c_{11}(\epsilon)p_{1}p_{2}+c_{02}(\epsilon)p_{2}^{2}+H(p_{1}, p_{2}). \end{array}\right. \end{equation*}

    Employing this transformation into (A.10), one obtains

    \begin{equation} \left\{\begin{array}{llll} \frac{d p_{1}^{'}}{d t} = & p_{2}^{'}, \\ \frac{d p_{2}^{'}}{d t} = & L_{00}(\epsilon)+L_{10}(\epsilon)p_{1}^{'}+L_{01}(\epsilon)p_{2}^{'}+L_{20}(\epsilon){p_{1}^{'}}^{2}+L_{11}(\epsilon)p_{1}^{'}p_{2}^{'}+L_{02}(\epsilon){p_{2}^{'}}^{2}+I(p_{1}^{'}, p_{2}^{'}), \end{array}\right. \end{equation} (A.11)

    where I(p_{1}^{'}, p_{2}^{'}) is the power series of p_{1}^{'} and p_{2}^{'} of the form {p_{1}^{'i}}{p_{2}^{'j}} with i+j \geq 3 ; whereas the remaining coefficients L_{ij} ; i, j = 0, 1, 2, \ldots are given by

    \begin{eqnarray*} L_{00}(\epsilon) & = & H_{10}^{'}(\epsilon)-G_{10}(\epsilon)b_{1}^{'}(\epsilon)+\ldots, \\ L_{10}(\epsilon) & = & a_{1}^{'}(\epsilon)+c_{11}(\epsilon)H_{10}(\epsilon)-d_{11}(\epsilon)G_{10}(\epsilon)+\ldots, \\ L_{01}(\epsilon) & = & b_{1}^{'}(\epsilon)+2c_{02}(\epsilon)H_{10}^{'}(\epsilon)-c_{11}(\epsilon)G_{10}(\epsilon)-2d_{02}(\epsilon)G_{10}(\epsilon)+\ldots, \\ L_{20}(\epsilon) & = & d_{20}(\epsilon)-c_{20}(\epsilon)b_{1}^{'}(\epsilon)+a_{1}^{'}(\epsilon)c_{11}(\epsilon)+\ldots, \\ L_{11}(\epsilon) & = & d_{11}(\epsilon)+2c_{20}(\epsilon)+2c_{02}(\epsilon)a_{1}^{'}(\epsilon)-c_{11}(\epsilon)b_{1}^{'}(\epsilon)+\ldots, \\ L_{02}(\epsilon) & = & c_{11}(\epsilon)+d_{02}(\epsilon)-c_{02}(\epsilon)n_{01}(\epsilon)+\ldots. \end{eqnarray*}

    Also, one may easily verify that

    \begin{equation*} \begin{split} & L_{00}(\epsilon^{*}) = L_{10}(\epsilon^{*}) = L_{01}(\epsilon^{*}) = 0, \, \, L_{20}(\epsilon^{*}) = d_{20}(\epsilon^{*}), \\ & L_{02}(\epsilon^{*}) = c_{11}(\epsilon^{*})+d_{02}(\epsilon^{*}), \, \, L_{11}(\epsilon^{*}) = d_{11}(\epsilon^{*})+2c_{20}(\epsilon^{*}). \end{split} \end{equation*}

    Therefore, the system of equations of (A.11) may be written as

    \begin{equation} \left\{\begin{array}{llll} \frac{d p_{1}^{'}}{d t} & = & p_{2}^{'}, \\ \frac{d p_{2}^{'}}{d t} & = & (L_{00}+L_{10}p_{1}^{'}+L_{20}{p_{1}^{'}}^{2})+(L_{01}+L_{11}p_{1}^{'}+o(||p^{'}||^{2}))p_{2}^{'}+(L_{02}+o(||p^{'}||)){p_{2}^{'}}^{2}, \\ & = & \rho_{1}(p_{1}^{'}, \epsilon)+\rho_{2}(p_{1}^{'}, \epsilon)p_{2}^{'}+\Phi_{1}^{'}(p^{'}, \epsilon){p_{2}^{'}}^{2}, \end{array}\right. \end{equation} (A.12)

    where, the terms \rho_{1}, \rho_{2}, \Phi_{1}^{'} are all smooth functions and \rho_{1}(0, \epsilon^{*}) = L_{00}(\epsilon^{*}) = 0 , \frac{\partial \rho_{1}}{\partial p_{1}^{'}}\bigg{\rvert}_{(0, \; \epsilon^{*})} = L_{10}(\epsilon^{*}) = 0 , \frac{1}{2}\frac{\partial^{2} \rho_{1}}{\partial {p_{1}^{'}}^{2}}\bigg{\rvert}_{(0, \; \epsilon^{*})} = L_{20}(\epsilon^{*}) = d_{20}(\epsilon^{*})\neq 0 , \rho_{2}(0, \epsilon^{*}) = L_{01}(\epsilon^{*}) = 0 . Also, one obtains

    \begin{eqnarray*} \frac{\partial \rho_{2}}{\partial p_{1}^{'}}\bigg{\rvert}_{(0, \; \epsilon^{*})} & = & L_{11}(\epsilon^{*}) = d_{11}(\epsilon^{*})+2c_{20}(\epsilon^{*}), \\ & = & \frac{m_{10}m_{11}}{m_{01}}-\frac{2m_{02}m_{10}^{2}}{m_{01}^{2}}+n_{11}-\frac{2n_{02}m_{10}}{m_{01}}+2\bigg{(}m_{20}-\frac{m_{11}m_{10}}{m_{01}}+\frac{m_{02}m_{10}^{2}}{m_{01}^{2}}\bigg{)}, \\ & = & 2m_{20}+n_{11}-\frac{m_{10}}{m_{01}}(m_{11}+2n_{02}) \neq 0. \end{eqnarray*}

    Since we have \rho_{2}(0, \epsilon^{*}) = 0 and \frac{\partial \rho_{2}}{\partial p_{1}^{'}}\bigg{\rvert}_{(0, \; \epsilon^{*})} \neq 0 , by employing implicit function theorem, there exists a c^{\infty} function p_{1}^{'} = \phi_{1}^{'}(\epsilon) defined in a very small neighborhood N(\epsilon^{*}) of \epsilon = \epsilon^{*} in such a way that \phi_{1}^{'}(\epsilon^{*}) = 0 , \rho_{2}(\phi_1^{'}(\epsilon), \epsilon) = 0 , for any choices of \epsilon \in N(\epsilon^{*}) .

    Step-7: (Parameter dependent shift)

    Now, to annihilate the p_{2}^{'} -term from the second equation of (A.12), the following parameter dependent shift transformation has been taken into account

    \begin{eqnarray*} p_{1}^{'} = E_{1}+\phi_{1}^{'}(\epsilon), \;\;p_{2}^{'} = E_{2}. \end{eqnarray*}

    By using this shift transformation, (A.12) may be represented as

    \begin{equation} \left\{\begin{array}{llll} \frac{d E_{1}}{d t} & = & E_{2}, \\ \frac{d E_{2}}{d t} & = & L_{00}+L_{10}(E_{1}+\phi_{1}^{'})+L_{01}E_{2}+L_{20}(E_{1}+\phi_{1}^{'})^{2}+L_{11}(E_{1}+\phi_{1}^{'})E_{2}+L_{02}E_{2}^{2}, \\ & = & (e_{00}+e_{10}E_{1}+e_{20}E_{1}^{2}+o(||E_{1}||^{3}))+(e_{01}+e_{11}E_{1}+o(||E_{1}||^{2}))E_{2}+(e_{02}+o(||E_{1}||))E_{2}^{2}, \\ & = & \delta_{1}(E_{1}, \epsilon)+\delta_{2}(E_{1}, \epsilon)E_{2}+\Psi_{1}^{'}(E, \epsilon)E_{2}^{2}, \end{array}\right. \end{equation} (A.13)

    where E = (E_{1}, E_{2}) , and

    \begin{equation*} e_{00} = L_{00}+L_{10}\phi_{1}^{'}, e_{10} = L_{10}+2L_{20}\phi_{1}^{'}, e_{20} = L_{20}, e_{01} = L_{01}+L_{11}\phi_{1}^{'}, e_{11} = L_{11}, e_{02} = L_{02}. \end{equation*}

    The coefficient of E_{2} of the system of equations of (A.13) may be written as

    \begin{eqnarray*} e_{01} & = & \delta_{2}(0, \epsilon) = L_{01}+L_{11}\phi_{1}^{'}+o(||\phi_{1}^{'}||^{2}), \\ & = & (b_{1}^{'}+2c_{02}H_{10}^{'}-c_{11}G_{10}-2d_{02}G_{10}+\ldots)+(d_{11}+2c_{20}+2c_{02}a_{1}^{'}-c_{11}b_{1}^{'}+\ldots)\phi_{1}^{'}. \end{eqnarray*}

    From the first condition of \mathrm{BT. 2} , we conclude that e_{01}(0, \epsilon^{*}) = L_{01}(\epsilon^{*}) = 0 , \frac{\partial e_{01}}{\partial \phi_{1}^{'}}\bigg{|}_{(0, \epsilon^{*})} = L_{11}(\epsilon^{*}) = d_{11}(\epsilon^{*})+2c_{20}(\epsilon^{*}) = \theta \neq 0 , which indicates that there exists a smooth function \phi_{1}^{'} = \phi_{1}^{'}(\epsilon) in such a way that \phi_{1}^{'}(\epsilon^{*}) = 0 , e_{01}(\phi_{1}^{'}(\epsilon), \epsilon) = \delta_{2}(0, \epsilon) = 0 for any choices of \epsilon \in N(\epsilon^{*}) . So, to annihilate the E_{2} -term from the second equation of (A.13), we consider e_{01} = 0 and one obtains

    \begin{equation*} \phi_{1}^{'}(\epsilon) \approx -\frac{{L}_{01}(\epsilon)}{{L}_{11}(\epsilon)} \approx -\frac{{L}_{01}(\epsilon)}{\theta} = \frac{1}{\theta}\bigg{[}-b_{1}^{'}-2c_{02}H_{10}^{'}+c_{11}G_{10}+2d_{02}G_{10}+\ldots\bigg{]}. \end{equation*}

    Therefore, the system of equations of (A.13) may be written as

    \begin{equation*} \left\{\begin{array}{llll} \frac{d E_{1}}{d t} & = & E_{2}, \\ \frac{d E_{2}}{d t} & = & (e_{00}+e_{10}E_{1}+e_{20}E_{1}^{2}+o(||E_{1}||^{3}))+(e_{11}+o(||E_{1}||))E_{1}E_{2}+(e_{02}+o(||E_{1}||))E_{2}^{2}, \\ & = & \bar{\delta}_{1}(E_{1}, \epsilon)+\bar{\delta}_{2}(E_{1}, \epsilon)E_{1}E_{2}+\bar{\kappa}_{1}(E_{1}, \epsilon)E_{1}^{2}, \end{array}\right. \end{equation*}

    which may be represented as

    \begin{equation} \left\{\begin{array}{llll} \frac{d E_{1}}{d t} & = & E_{2}, \\ \frac{d E_{2}}{d t} & = & e_{00}(\epsilon)+e_{10}(\epsilon)E_{1}+e_{20}(\epsilon)E_{1}^{2}+e_{11}(\epsilon)E_{1}E_{2}+e_{02}(\epsilon)E_{2}^{2}+o(||E||^{3}), \end{array}\right. \end{equation} (A.14)

    where E = (E_{1}, E_{2}) .

    Step-8: (Time reparametrization)

    Furthermore, here e_{00}(\epsilon^{*}) = e_{10}(\epsilon^{*}) = 0 . Introducing a new time scaling defined by dt = (1+\omega E_{1})d\tau , and by using it (A.14) becomes

    \begin{equation} \left\{\begin{array}{llll} \frac{d E_{1}}{d \tau} & = & E_{2}(1+\omega E_{1}), \\ \frac{d E_{2}}{d \tau} & = & e_{00}+(e_{10}+\omega e_{00})E_{1}+(e_{20}+\omega e_{10})E_{1}^{2}+e_{11}E_{1}E_{2}+e_{02}E_{2}^{2}+o(||E||^{3}). \end{array}\right. \end{equation} (A.15)

    One may consider a new transformation of the form

    \begin{eqnarray*} P = E_{1}, \;\; Q = E_{2}(1+\omega E_{1}). \end{eqnarray*}

    By use of this transformation, (A.15) may be written as

    \begin{equation*} \left\{\begin{array}{llll} \frac{d P}{d \tau} & = & Q, \\ \frac{d Q}{d \tau} & = & f_{00}(\epsilon)+f_{10}(\epsilon)P+f_{20}(\epsilon)P^{2}+f_{11}(\epsilon)PQ+f_{02}(\epsilon)Q^{2}+H(P, Q), \end{array}\right. \end{equation*}

    where, H(P, Q) are the power series expansion of P and Q of the form P^{i}Q^{j} ; i+j \geq 3 . Also, the coefficients f_{ij} ; i, j = 0, 1, 2, 3, \ldots are given by

    \begin{eqnarray*} f_{00}(\epsilon) & = & e_{00}(\epsilon), \; f_{10}(\epsilon) = e_{10}(\epsilon)+2\omega(\epsilon)e_{00}(\epsilon), \; f_{20}(\epsilon) = e_{20}(\epsilon)+2\omega(\epsilon)e_{10}(\epsilon)+\omega^{2}(\epsilon)e_{00}(\epsilon), \\ f_{11}(\epsilon) & = & e_{11}(\epsilon), f_{02}(\epsilon) = e_{02}(\epsilon)+\omega(\epsilon). \end{eqnarray*}

    Now, to annihilate the Q^{2} -term, let us choose \omega(\epsilon) = -e_{02}(\epsilon) and we obtain

    \begin{equation} \left\{\begin{array}{llll} \frac{d P}{d \tau} & = & Q, \\ \frac{d Q}{d \tau} & = & \gamma_{1}(\epsilon)+\gamma_{2}(\epsilon)P+\mathcal{D}(\epsilon)P^{2}+\mathcal{E}(\epsilon)PQ+H(P, Q), \end{array}\right. \end{equation} (A.16)

    where \gamma_{1}(\epsilon) = e_{00}(\epsilon) , \gamma_{2}(\epsilon) = e_{10}(\epsilon)-2e_{00}(\epsilon)e_{02}(\epsilon) , \mathcal{D}(\epsilon) = e_{20}(\epsilon)-2e_{10}(\epsilon)e_{02}(\epsilon)+e_{00}(\epsilon)e_{02}^{2}(\epsilon) , \mathcal{E}(\epsilon) = e_{11}(\epsilon) . Also, one may verify that, \gamma_{1}(\epsilon^{*}) = 0 , \gamma_{2}(\epsilon^{*}) = 0 , and \mathcal{E}(\epsilon^{*}) = L_{11}(\epsilon^{*}) = \theta \neq 0 because of {{\bf{BT}}. \; {\bf{2}}} . is satisfied.

    Step-9: (Time rescaling)

    Since the second condition of {{\bf{BT}}. \; {\bf{2}}} is satisfied, \mathcal{D}(\epsilon^\ast) = e_{20}(\epsilon^\ast) = L_{20}(\epsilon^\ast) = d_{20}(\epsilon^\ast)\neq 0 . Thus we can choose a \delta^\ast- neighborhood such that \mathcal{D}(\epsilon)\neq 0 when \epsilon\in O(\epsilon_\ast, \delta^\ast) . Let us introduce a new time scaling that is of the form t = \bigg{|}\frac{\mathcal{E}(\epsilon)}{\mathcal{D}(\epsilon)}\bigg{|}\tau. Again, choose \xi_{1} = \frac{\mathcal{D}(\epsilon)}{\mathcal{E}^{2}(\epsilon)} , \xi_{2} = s\bigg{(}\frac{\mathcal{D}^{2}(\epsilon)}{\mathcal{E}^{3}(\epsilon)}\bigg{)}Q , where s = sign\bigg{(}\frac{\mathcal{E}(\epsilon)}{\mathcal{D}(\epsilon)}\bigg{)} = sign \bigg{(}\frac{\mathcal{E}(\epsilon^{*})}{\mathcal{D}(\epsilon^{*})}\bigg{)} = sign\bigg{(}\frac{\theta}{d_{20}(\epsilon^{*})}\bigg{)} = \pm 1 .

    Therefore, the reduced model system (A.16) takes the form

    \begin{equation} \left\{\begin{array}{llll} \frac{d \xi_{1}}{d t} & = & \xi_{2}, \\ \frac{d \xi_{2}}{d t} & = & \vartheta_{1}+\vartheta_{2}\xi_{1}+{\xi_{1}}^{2}+s\xi_{1}\xi_{2}+o(||\xi||), \end{array}\right. \end{equation} (A.17)

    where \xi = (\xi_{1}, \xi_{2}) ,

    \begin{equation*} \vartheta_{1}(\epsilon) = \bigg{(}\frac{\mathcal{E}^{4}(\epsilon)}{\mathcal{D}^{3}(\epsilon)}\bigg{)}\gamma_{1}(\epsilon), \quad \vartheta_{2}(\epsilon) = \bigg{(}\frac{\mathcal{E}^{2}(\epsilon)}{\mathcal{D}^{2}(\epsilon)}\bigg{)}\gamma_{2}(\epsilon). \end{equation*}

    In a very small neighborhood of the origin, (A.17) is topologically equivalent to the following one

    \begin{equation} \left\{\begin{array}{llll} \frac{d \xi_{1}}{d t} & = & \xi_{2}, \\ \frac{d \xi_{2}}{d t} & = & \vartheta_{1}+\vartheta_{2}\xi_{1}+{\xi_{1}}^{2}+s\xi_{1}\xi_{2}, \end{array}\right. \end{equation} (A.18)

    which is the generic normal form of Bogdanov-Takens bifurcation. Combining BT. 3, we can define an invertible change of parameters near the origin, which is also equivalent to the regularity of the map \epsilon\rightarrow (\vartheta_1(\epsilon), \vartheta_2(\epsilon)) at \epsilon = \epsilon^\ast = (0, 0) . Therefore, one may conclude that (3.2) exhibits a Bogdanov-Takens bifurcation around the coexistence equilibrium point E_{*} = (u_{*}, v_{*}) whenever the system parameters (\sigma, \gamma) attains the critical value (\sigma^{[BT]}, \gamma^{[BT]}) , which implies the proof of the theorem as complete.



    [1] F. Courchamp, T. Clutton-Brock, B. Grenfell, Inverse density dependence and the Allee effect, Trend. Ecol. Evol., 14 (1999), 405–410. https://doi.org/10.1016/S0169-5347(99)01683-3 doi: 10.1016/S0169-5347(99)01683-3
    [2] P. A. Stephens, W. J. Sutherland, R. P. Freckleton, What is the Allee effect, Oikos, 87 (1999), 185–190. https://doi.org/10.2307/3547011 doi: 10.2307/3547011
    [3] P. A. Stephens, W. J. Sutherland, Consequences of the Allee effect for behaviour, ecology and conservation, Trends Ecol. Evol., 14 (1999), 401–405. https://doi.org/10.1016/s0169-5347(99)01684-5 doi: 10.1016/s0169-5347(99)01684-5
    [4] F. Courchamp, L. Berec, J. Gascoigne, Allee Effects in Ecology and Conservation, Oxford University Press, Oxford, 2008. https://doi.org/10.1093/acprof:oso/9780198570301.001.0001
    [5] D. Scheel, C. Packer, Group hunting behavior of lions: a search for cooperation, Anim. Behav., 41 (1991), 697–709. https://doi.org/10.1016/S0003-3472(05)80907-8 doi: 10.1016/S0003-3472(05)80907-8
    [6] M. T. Alves, F. M. Hilker, Hunting cooperation and Allee effects in predators, J. Theor. Biol., 419 (2017), 13–22. https://doi.org/10.1016/j.jtbi.2017.02.002 doi: 10.1016/j.jtbi.2017.02.002
    [7] R. Han, G. Mandal, L. N. Guin, S. Chakravarty, Dynamical response of a reaction-diffusion predator-prey system with cooperative hunting and prey refuge, J. Stat. Mech., 2022 (2022), 103502. https://doi.org/10.1088/1742-5468/ac946d doi: 10.1088/1742-5468/ac946d
    [8] L. Y. Zanette, A. F. White, M. C. Allen, M. Clinchy, Perceived predation risk reduces the number of offspring songbirds produce per year, Science, 334 (2011), 1398–1401. https://doi.org/10.1126/science.1210908 doi: 10.1126/science.1210908
    [9] X. Y. Wang, L. Y. Zanette, X. F. Zou, Modelling the fear effect in predator-prey interaction, J. Math. Biol., 73 (2016), 1179–1204. https://doi.org/10.1007/s00285-016-0989-1 doi: 10.1007/s00285-016-0989-1
    [10] P. A. Schmidt, L. D. Mech, Wolf pack size and food acquisition, Am. Nat., 150 (1997), 513–517. https://doi.org/10.1086/286079 doi: 10.1086/286079
    [11] F. Courchamp, D. W. Macdonald, Crucial importance of pack size in the African wild dog Lycaon pictus, Anim. Conserv., 4 (2001), 169–174. https://doi.org/10.1017/S1367943001001196 doi: 10.1017/S1367943001001196
    [12] S. Vinoth, R. Sivasamy, K. Sathiyanathan, B. Unyong, G. Rajchakit, R. Vadivel, et al., The dynamics of a Leslie type predator-prey model with fear and Allee effect, Adv. Differ. Equ., 2021 (2021), 1–22. https://doi.org/10.1186/s13662-021-03490-x doi: 10.1186/s13662-021-03490-x
    [13] B. Dennis, Allee effects: population growth, critical density, and the chance of extinction, Nat. Resour. Model., 3 (1989), 481–538. https://doi.org/10.1111/j.1939-7445.1989.tb00119.x doi: 10.1111/j.1939-7445.1989.tb00119.x
    [14] A. D. Bazykin, Nonlinear Dynamics of Interacting Populations, World Scientific Publishing Company Incorporated, Singapore, 1998.
    [15] F. Courchamp, B. T. Grenfell, T. H. Clutton-Brock, Impact of natural enemies on obligately cooperative breeders, Oikos, 91 (2000), 311–322. https://doi.org/10.1034/j.1600-0706.2000.910212.x doi: 10.1034/j.1600-0706.2000.910212.x
    [16] A. Kent, C. P. Doncaster, T. Sluckin, Consequences for predators of rescue and Allee effects on prey, Ecol. Modell., 162 (2003), 233–245. https://doi.org/10.1016/S0304-3800(02)00343-5 doi: 10.1016/S0304-3800(02)00343-5
    [17] S. Zhou, Y. Zhou, G. Wang, The stability of predator-prey systems subject to the Allee effects, Theor. Popul. Biol., 67 (2005), 23–31. https://doi.org/10.1016/j.tpb.2004.06.007 doi: 10.1016/j.tpb.2004.06.007
    [18] K. H. Elliott, G. S. Betini, D. R. Norris, Fear creates an Allee effect: experimental evidence from seasonal populations, Proc. R. Soc. B Biol. Sci., 284 (2017), 20170878. https://doi.org/10.1098/rspb.2017.0878 doi: 10.1098/rspb.2017.0878
    [19] S. K. Sasmal, Population dynamics with multiple Allee effects induced by fear factors-A mathematical on prey-predator interactions, Appl. Math. Modell., 64 (2018), 1–14. https://doi.org/10.1016/j.apm.2018.07.021 doi: 10.1016/j.apm.2018.07.021
    [20] Y. Li, M. He, Z. Li, Dynamics of a ratio-dependent Leslie-Gower predator-prey model with Allee effect and fear effect, Math. Comput. Simul., 201 (2022), 417–439. https://doi.org/10.1016/j.matcom.2022.05.017 doi: 10.1016/j.matcom.2022.05.017
    [21] Y. Shi, J. H. Wu, Q. Cao, Analysis on a diffusive multiple Allee effects predator-prey model induced by fear factors, Nonlinear Anal. Real World Appl., 59 (2021), 103249. https://doi.org/10.1016/j.nonrwa.2020.103249 doi: 10.1016/j.nonrwa.2020.103249
    [22] L. Y. Lai, Z. L. Zhu, F. D. Chen, Stability and bifurcation analysis in a predator-prey model with the additive Allee effect and the fear effect, Mathematics, 8 (2020), 1280. https://doi.org/10.3390/math8081280 doi: 10.3390/math8081280
    [23] J. P. Suraci, M. Clinchy, L. M. Dill, D. Roberts, L. Y. Zanette, Fear of large carnivores causes a trophic cascade, Nat. Commun., 7 (2016), 10698. https://doi.org/10.1038/ncomms10698 doi: 10.1038/ncomms10698
    [24] X. Wang, X. Zou, Modeling the fear effect in predator-prey interactions with adaptive avoidance of predators, Bull. Math. Biol., 79 (2017), 1325–1359. https://doi.org/10.1007/s00285-016-0989-1 doi: 10.1007/s00285-016-0989-1
    [25] A. J. Wirsing, W. J. Ripple, A comparison of shark and wolf research reveals similar behavioral responses by prey, Front. Ecol. Environ., 9 (2011), 335–341. https://doi.org/10.1890/090226 doi: 10.1890/090226
    [26] M. J. Sheriff, C. J. Krebs, R. Boonstra, The sensitive hare: sublethal effects of predator stress on reproduction in snowshoe hares, J. Anim. Ecol., 78 (2009), 1249–1258. https://doi.org/10.1111/j.1365-2656.2009.01552. doi: 10.1111/j.1365-2656.2009.01552
    [27] J. Wang, Y. Cai, S. Fu, W. Wang, The effect of the fear factor on the dynamics of a predator-prey model incorporating the prey refuge, Chaos: Interdiscip. J. Nonlinear Sci., 29 (2019), 083109. https://doi.org/10.1063/1.5111121 doi: 10.1063/1.5111121
    [28] S. Samaddar, M. Dhar, P. Bhattacharya, Effect of fear on prey-predator dynamics: Exploring the role of prey refuge and additional food, Chaos: Interdiscip. J. Nonlinear Sci., 30 (2020), 063129. https://doi.org/10.1063/5.0006968 doi: 10.1063/5.0006968
    [29] B. F. Xie, Z. C. Zhang, Impact of Allee and fear effects in a fractional order prey-predator system incorporating prey refuge, Chaos: Interdiscip. J. Nonlinear Sci., 33 (2023), 013131. https://doi.org/10.1063/5.0130809 doi: 10.1063/5.0130809
    [30] J. Liu, Qualitative analysis of a diffusive predator-prey model with Allee and fear effects, Int. J. Biomath., 14 (2021), 2150037. https://doi.org/10.1142/S1793524521500376 doi: 10.1142/S1793524521500376
    [31] D. Pal, D. Kesh, D. Mukherjee, Qualitative study of cross-diffusion and pattern formation in Leslie-Gower predator-prey model with fear and Allee effects, Chaos, Solitons Fractals, 167 (2023), 113033. https://doi.org/10.1016/j.chaos.2022.113033 doi: 10.1016/j.chaos.2022.113033
    [32] H. Amann, Dynamic theory of quasilinear parabolic equations Ⅱ. reaction-diffusion systems, Differ. Integr. Equations, 3 (1990), 13–75. https://doi.org/10.57262/die/1371586185 doi: 10.57262/die/1371586185
    [33] H. Amann, Dynamics theory of quasilinear parabolic systems Ⅲ. Global existence, Math. Z., 202 (1989), 219–250. https://doi.org/10.1007/BF01215256 doi: 10.1007/BF01215256
    [34] H. Amann, Nonhomogeneous linear and quasiliear elliptic and parabolic boundary value problems, in Function Spaces, Differential Operators and Nonlinear Analysis (eds. H.J. Schmeisser and H. Triebel), Teubner-Texte zur Mathematik, 133 (1993), 9–126. https://doi.org/10.1007/978-3-663-11336-2_1
    [35] S. L. Hollis, R. H. Martin, M. Pierre, Global existence and boundedness in reaction-diffusion systems, SIAM J. Math. Anal., 18 (1987), 744–761. https://doi.org/10.1137/0518057 doi: 10.1137/0518057
    [36] S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, Springer, New York, 2003.
    [37] L. Perko, Differential Equations and Dynamical Systems, Springer-Verlag, New York, 2001.
    [38] Y. A. Kuznetsov, Elements of Applied Bifurcation Theory, Springer-Verlag, New York, 1998.
    [39] W. Govaerts, Y. A. Kuznetsov, H. G. E. Meijer, B. Al-Hdaibat, V. De Witte, A. Dhooge, et al., MATCONT: Continuation toolbox for ODEs in Matlab, Universiteit Gent, Belgium, 2019.
    [40] M. G. Crandall, P. H. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal., 8 (1971), 321–340. https://doi.org/10.1016/0022-1236(71)90015-2 doi: 10.1016/0022-1236(71)90015-2
    [41] F. Q. Yi, J. J. Wei, J. P. Shi, Bifurcation and spatiotemporal patters in a homogeneous diffusive predator-prey system, J. Differ. Equations, 246 (2009), 1944–1977. https://doi.org/10.1016/j.jde.2008.10.024 doi: 10.1016/j.jde.2008.10.024
    [42] R. Han, S. Dey, M. Banerjee, Spatio-temporal pattern selection in a prey-predator model with hunting cooperation and Allee effect in prey, Chaos, Solitons Fractals, 171 (2023), 113441. https://doi.org/10.1016/j.chaos.2023.113441 doi: 10.1016/j.chaos.2023.113441
    [43] A. Morozov, S. Petrovskii, B. L. Li, Spatiotemporal complexity of patchy invasion in a predator-prey system with the Allee effect, J. Theor. Biol., 238 (2006), 18–35. https://doi.org/10.1016/j.jtbi.2005.05.021 doi: 10.1016/j.jtbi.2005.05.021
    [44] P. R. Chowdhury, S. Petrovskii, M. Banerjee, Oscillation and pattern formation in a slow-fast prey-predator system, Bull. Math. Biol., 83 (2021), 1–41. https://doi.org/10.1007/s11538-021-00941-0 doi: 10.1007/s11538-021-00941-0
    [45] S. Chow, C. Li, D. Wang, Normal Forms and Bifurcation of Plannar Vector Fields, Cambridge University Press, 1994. https://doi.org/10.1017/CBO9780511665639
  • This article has been cited by:

    1. Gourav Mandal, Sukanya Das, Swagata Dutta, Lakshmi Narayan Guin, Santabrata Chakravarty, A Comparative Study of Allee Effects and Fear-Induced Responses: Exploring Hyperbolic and Ratio-Dependent Models, 2024, 10, 2349-5103, 10.1007/s40819-024-01773-x
    2. Gourav Mandal, Lakshmi Narayan Guin, Santabrata Chakravarty, Salih Djilali, Anwar Zeb, Depensation of supplementary food in a system of interacting species with refuge, 2024, 139, 2190-5444, 10.1140/epjp/s13360-023-04793-6
    3. Debjit Pal, Santu Ghorai, Dipak Kesh, Debasis Mukherjee, Hopf bifurcation and patterns formation in a diffusive two prey-one predator system with fear in preys and help, 2024, 481, 00963003, 128927, 10.1016/j.amc.2024.128927
    4. Liyang Wang, Yantao Luo, Zhidong Teng, Tingting Zheng, Stability and Hopf bifurcation for a multi-delay PSIS eco-epidemic model with saturation incidence and Beddington–DeAngelis functional response, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-05889-3
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1620) PDF downloads(110) Cited by(4)

Article outline

Figures and Tables

Figures(12)  /  Tables(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog