We considered predator-prey models which incorporated both an Allee effect and a new fear factor effect together, and where the predator predated the prey with a Holling type I functional response. We started off with a two-dimensional model where we found possible equilibria and examined their stabilities. By using the predator mortality rate as the bifurcation parameter, the model exhibited Hopf-bifurcation for the coexistence equilibrium. Furthermore, our numerical illustrations demonstrated the effect of fear and the Allee effect on the population densities, and we found that the level of fear had little impact on the long-term prey population level. The population of predators, however, declined as the fear intensity rose, indicating that the fear effect might result in a decline in the predator population. The dynamics of the delayed system were examined and Hopf-bifurcation was discussed. Finally, we looked at an eco-epidemiological model that took into account the same cost of fear and the Allee effect. In this model, the prey was afflicted with a disease. The prey was either susceptible or infected. Numerical simulations were carried out to show that as the Allee threshold rose, the uninfected prey and predator decreased, while the population of infected prey increased. When the Allee threshold hit a certain value, all populations became extinct. As fear intensity increased, the population of uninfected prey decreased, and beyond a certain level of fear, habituation prevented the uninfected prey from changing. After a certain level of fear, the predator population went extinct and, as a result, the only interaction left was between uninfected and infected prey which increased disease transmission, and so the infected prey increased. Hopf-bifurcation was studied by taking the time delay as the bifurcation parameter. We estimated the delay length to preserve stability.
Citation: Kawkab Al Amri, Qamar J. A Khan, David Greenhalgh. Combined impact of fear and Allee effect in predator-prey interaction models on their growth[J]. Mathematical Biosciences and Engineering, 2024, 21(10): 7211-7252. doi: 10.3934/mbe.2024319
[1] | Zhongzi Zhao, Meng Yan . Positive radial solutions for the problem with Minkowski-curvature operator on an exterior domain. AIMS Mathematics, 2023, 8(9): 20654-20664. doi: 10.3934/math.20231052 |
[2] | Wenjia Li, Guanglan Wang, Guoliang Li . The local boundary estimate of weak solutions to fractional p-Laplace equations. AIMS Mathematics, 2025, 10(4): 8002-8021. doi: 10.3934/math.2025367 |
[3] | Sobajima Motohiro, Wakasugi Yuta . Remarks on an elliptic problem arising in weighted energy estimates for wave equations with space-dependent damping term in an exterior domain. AIMS Mathematics, 2017, 2(1): 1-15. doi: 10.3934/Math.2017.1.1 |
[4] | Limin Guo, Jiafa Xu, Donal O'Regan . Positive radial solutions for a boundary value problem associated to a system of elliptic equations with semipositone nonlinearities. AIMS Mathematics, 2023, 8(1): 1072-1089. doi: 10.3934/math.2023053 |
[5] | Keqiang Li, Shangjiu Wang . Symmetry of positive solutions of a p-Laplace equation with convex nonlinearites. AIMS Mathematics, 2023, 8(6): 13425-13431. doi: 10.3934/math.2023680 |
[6] | Zhanbing Bai, Wen Lian, Yongfang Wei, Sujing Sun . Solvability for some fourth order two-point boundary value problems. AIMS Mathematics, 2020, 5(5): 4983-4994. doi: 10.3934/math.2020319 |
[7] | Lin Zhao . Monotonicity and symmetry of positive solution for 1-Laplace equation. AIMS Mathematics, 2021, 6(6): 6255-6277. doi: 10.3934/math.2021367 |
[8] | Manal Alfulaij, Mohamed Jleli, Bessem Samet . A hyperbolic polyharmonic system in an exterior domain. AIMS Mathematics, 2025, 10(2): 2634-2651. doi: 10.3934/math.2025123 |
[9] | Maria Alessandra Ragusa, Abdolrahman Razani, Farzaneh Safari . Existence of positive radial solutions for a problem involving the weighted Heisenberg p(⋅)-Laplacian operator. AIMS Mathematics, 2023, 8(1): 404-422. doi: 10.3934/math.2023019 |
[10] | Zhiqian He, Liangying Miao . Multiplicity of positive radial solutions for systems with mean curvature operator in Minkowski space. AIMS Mathematics, 2021, 6(6): 6171-6179. doi: 10.3934/math.2021362 |
We considered predator-prey models which incorporated both an Allee effect and a new fear factor effect together, and where the predator predated the prey with a Holling type I functional response. We started off with a two-dimensional model where we found possible equilibria and examined their stabilities. By using the predator mortality rate as the bifurcation parameter, the model exhibited Hopf-bifurcation for the coexistence equilibrium. Furthermore, our numerical illustrations demonstrated the effect of fear and the Allee effect on the population densities, and we found that the level of fear had little impact on the long-term prey population level. The population of predators, however, declined as the fear intensity rose, indicating that the fear effect might result in a decline in the predator population. The dynamics of the delayed system were examined and Hopf-bifurcation was discussed. Finally, we looked at an eco-epidemiological model that took into account the same cost of fear and the Allee effect. In this model, the prey was afflicted with a disease. The prey was either susceptible or infected. Numerical simulations were carried out to show that as the Allee threshold rose, the uninfected prey and predator decreased, while the population of infected prey increased. When the Allee threshold hit a certain value, all populations became extinct. As fear intensity increased, the population of uninfected prey decreased, and beyond a certain level of fear, habituation prevented the uninfected prey from changing. After a certain level of fear, the predator population went extinct and, as a result, the only interaction left was between uninfected and infected prey which increased disease transmission, and so the infected prey increased. Hopf-bifurcation was studied by taking the time delay as the bifurcation parameter. We estimated the delay length to preserve stability.
Boundary value problems with p-Laplace operator Δpu=div(|∇u|p−2∇u) arise in many different areas of applied mathematics and physics, such as non-Newtonian fluids, reaction-diffusion problems, non-linear elasticity, etc. But little is known about the p-Laplace operator cases (p≠2) compared to the vast amount of knowledge for the Laplace operator (p=2). In this paper, we discuss the existence of positive radial solution for the p-Laplace boundary value problem (BVP)
{−Δpu=K(|x|)f(u),x∈Ω,∂u∂n=0,x∈∂Ω,lim|x|→∞u(x)=0, | (1.1) |
in the exterior domain Ω={x∈RN:|x|>r0}, where N≥2, r0>0, 1<p<N, ∂u∂n is the outward normal derivative of u on ∂Ω, K:[r0,∞)→R+ is a coefficient function, f:R+→R is a nonlinear function. Throughout this paper, we assume that the following conditions hold:
(A1) K∈C([r0,∞),R+) and 0<∫∞r0rN−1K(r)dr<∞;
(A2) f∈C(R+,R+);
For the special case of p=2, namely the Laplace boundary value problem
{−Δu=K(|x|)f(u),x∈Ω,∂u∂n=0,x∈∂Ω,lim|x|→∞u(x)=0, | (1.2) |
the existence of positive radial solutions has been discussed by many authors, see [1,2,3,4,5,6,7]. The authors of references[1,2,3,4,5,6] obtained some existence results by using upper and lower solutions method, priori estimates technique and fixed point index theory. In [7], the present author built an eigenvalue criteria of existing positive radial solutions. The eigenvalue criterion is related to the principle eigenvalue λ1 of the corresponding radially symmetric Laplace eigenvalue problem (EVP)
{−Δu=λK(|x|)u,x∈Ω,∂u∂n=0,x∈∂Ω,u=u(|x|),lim|x|→∞u(|x|)=0. | (1.3) |
Specifically, if f satisfies one of the following eigenvalue conditions:
(H1) f0<λ1, f∞>λ1;
(H2) f∞<λ1, f0>λ1,
the BVP(1.2) has a classical positive radial solution, where
f0=lim infu→0+f(u)u,f0=lim supu→0+f(u)u,f∞=lim infu→∞f(u)u,f∞=lim supu→∞f(u)u. |
See [7,Theorem 1.1]. This criterion first appeared in a boundary value problem of second-order ordinary differential equations, and built by Zhaoli Liu and Fuyi Li in [8]. Then it was extended to general boundary value problems of ordinary differential equations, See [9,10]. In [11,12], the radially symmetric solutions of the more general Hessian equations are discussed.
The purpose of this paper is to establish a similar existence result of positive radial solution of BVP (1.1). Our results are related to the principle eigenvalue λp,1 of the radially symmetric p-Laplce eigenvalue problem (EVP)
{−Δpu=λK(|x|)|u|p−2u,x∈Ω,∂u∂n=0,x∈∂Ω,u=u(|x|),lim|x|→∞u(|x|)=0. | (1.4) |
Different from EVP (1.3), EVP (1.4) is a nonlinear eigenvalue problem, and the spectral theory of linear operators is not applicable to it. In Section 2 we will prove that EVP (1.4) has a minimum positive real eigenvalue λp,1, see Lemma 2.3. For BVP (1.1), we conjecture that eigenvalue criteria is valid if f0, f0, f∞ and f∞ is replaced respectively by
fp0=lim infu→0+f(u)up−1,fp0=lim supu→0+f(u)up−1,fp∞=lim infu→∞f(u)up−1,fp∞=lim supu→∞f(u)up−1. | (1.5) |
But now we can only prove a weaker version of it: In second inequality of (H1) and (H2), λp,1 needs to be replaced by the larger number
B=[∫10Ψ(∫1stp−1a(t)dt)ds]−(p−1), | (1.6) |
where a∈C+(0,1] is given by (2.4) and Ψ∈C(R) is given by (2.7). Our result is as follows:
Theorem 1.1. Suppose that Assumptions (A1) and (A2) hold. If the nonlinear function f satisfies one of the the following conditions:
(H1)∗ fp0<λp,1, fp∞>B;
(H2)∗ fp∞<λp,1, fp0>B,
then BVP (1.1) has at least one classical positive radial solution.
As an example of the application of Theorem 1.1, we consider the following p-Laplace boundary value problem
{−Δpu=K(|x|)|u|γ,x∈Ω,∂u∂n=0,x∈∂Ω,lim|x|→∞u(x)=0. | (1.7) |
Corresponding to BVP (1.1), f(u)=|u|γ. If γ>p−1, by (1.5) fp0=0, fp∞=+∞, and (H1) holds. If 0<γ<p−1, then fp∞=0, fp0=+∞, and (H2) holds. Hence, by Theorem 1.1 we have
Corollary 1. Let K:[r0,∞)→R+ satisfy Assumption (A1), γ>0 and γ≠p−1. Then BVP (1.7) has a positive radial solution.
The proof of Theorem 1.1 is based on the fixed point index theory in cones, which will be given in Section 3. Some preliminaries to discuss BVP (1.1) are presented in Section 2.
For the radially symmetric solution u=u(|x|) of BVP (1.1), setting r=|x|, since
−Δpu=div(|∇u|p−2∇u)=−(|u′(r)|p−2u′(r))′−N−1r|u′(r)|p−2u′(r), |
BVP (1.1) becomes the ordinary differential equation BVP in [r0,∞)
{−(|u′(r)|p−2u′(r))′−N−1r|u′(r)|p−2u′(r)=K(r)f(u(r)),r∈[r0,∞),u′(r0)=0,u(∞)=0, | (2.1) |
where u(∞)=limr→∞u(r).
Let q>1 be the constant satisfying 1p+1q=1. To solve BVP (2.1), make the variable transformations
t=(r0r)(q−1)(N−p),r=r0t−1/(q−1)(N−p),v(t)=u(r(t)), | (2.2) |
Then BVP (2.1) is converted to the ordinary differential equation BVP in (0,1]
{−(|v′(t)|p−2v′(t))′=a(t)f(v(t)),t∈(0,1],v(0)=0,v′(1)=0, | (2.3) |
where
a(t)=rq(N−1)(t)(q−1)p(N−p)pr0q(N−p)K(r(t)),t∈(0,1]. | (2.4) |
BVP (2.3) is a quasilinear ordinary differential equation boundary value problem with singularity at t=0. A solution v of BVP (2.3) means that v∈C1[0,1] such that |v′|p−2v′∈C1(0,1] and it satisfies the Eq (2.3). Clearly, if v is a solution of BVP (2.3), then u(r)=v(t(r)) is a solution of BVP (2.1) and u(|x|) is a classical radial solution of BVP (1.1). We discuss BVP (2.3) to obtain positive radial solutions of BVP (1.1).
Let I=[0,1] and R+=[0,+∞). Let C(I) denote the Banach space of all continuous function v(t) on I with norm ‖v‖C=maxt∈I|v(t)|, C1(I) denote the Banach space of all continuous differentiable function on I. Let C+(I) be the cone of all nonnegative functions in C(I).
To discuss BVP (2.3), we first consider the corresponding simple boundary value problem
{−(|v′(t)|p−2v′(t))′=a(t)h(t),t∈(0,1],v(0)=0,v′(1)=0, | (2.5) |
where h∈C+(I) is a given function. Let
Φ(v)=|v|p−2v=|v|p−1sgnv,v∈R, | (2.6) |
then w=Φ(v) is a strictly monotone increasing continuous function on R and its inverse function
Φ−1(w):=Ψ(w)=|w|q−1sgnw,w∈R, | (2.7) |
is also a strictly monotone increasing continuous function.
Lemma 2.1. For every h∈C(I), BVP (2.5) has a unique solution v:=Sh∈C1(I). Moreover, the solution operator S:C(I)→C(I) is completely continuous and has the homogeneity
S(νh)=νq−1Sh,h∈C(I),ν≥0. | (2.8) |
Proof. By (2.4) and Assumption (A1), the coefficient a(t)∈C+(0,1] and satisfies
∫10a(t)dt=1[(q−1)(N−p)]p−1r0N−p∫∞r0rN−1K(r)dr<∞. | (2.9) |
Hence a∈L(I).
For every h∈C(I), we verify that
v(t)=∫t0Ψ(∫1sa(τ)h(τ)dτ)ds:=Sh(t),t∈I | (2.10) |
is a unique solution of BVP (2.5). Since the function G(s):=∫1sa(τ)h(τ)dτ∈C(I), from (2.10) it follows that v∈C1(I) and
v′(t)=Ψ(∫1ta(τ)h(τ)dτ),t∈I. | (2.11) |
Hence,
|v′(t)|p−2v′(t)=Φ(v′(t))=∫1ta(τ)h(τ)dτ,t∈I. |
This means that (|v′(t)|p−2v′(t)∈C1(0,1] and
(|v′(t)|p−2v′(t))′=−a(t)h(t),t∈(0,1], |
that is, v is a solution of BVP (2.5).
Conversely, if v is a solution of BVP (2.5), by the definition of the solution of BVP (2.5), it is easy to show that v can be expressed by (2.10). Hence, BVP (2.5) has a unique solution v=Sh.
By (2.10) and the continuity of Ψ, the solution operator S:C(I)→C(I) is continuous. Let D⊂C(I) be bounded. By (2.10) and (2.11) we can show that S(D) and its derivative set {v′|v∈S(D)} are bounded sets in C(I). By the Ascoli-Arzéla theorem, S(D) is a precompact subset of C(I). Thus, S:C(I)→C(I) is completely continuous.
By the uniqueness of solution of BVP (2.5), we easily verify that the solution operator S satisfies (2.8).
Lemma 2.2. If h∈C+(I), then the solution v=Sh of LBVP (2.5) satisfies: ‖v‖c=v(1), v(t)≥t‖v‖C for every t∈I.
Proof. Let h∈C+(I) and v=Sh. By (2.10) and (2.11), for every t∈I v(t)≥0 and v′(t)≥0. Hence, v(t) is a nonnegative monotone increasing function and ‖v‖C=maxt∈Iv(t)=v(1). From (2.11) and the monotonicity of Ψ, we notice that v′(t) is a monotone decreasing function on I. For every t∈(0,1), by Lagrange's mean value theorem, there exist ξ1∈(0,t) and ξ2∈(t,1), such that
(1−t)v(t)=(1−t)(v(t)−v(0))=v′(ξ1)t(1−t)≥v′(t)t(1−t),tv(t)=tv(1)−t(v(1)−v(t))=tv(1)−tv′(ξ2)(1−t)≥tv(1)−v′(t)t(1−t). |
Hence
v(t)=tv(t)+(1−t)v(t)≥tv(1)=t‖v‖C. |
Obviously, when t=0 or 1, this inequality also holds. The proof is completed.
Consider the radially symmetric p-Laplace eigenvalue problem EVP (1.3). We have
Lemma 2.3. EVP (1.4) has a minimum positive real eigenvalue λp,1, and λp,1 has a radially symmetric positive eigenfunction.
Proof. For the radially symmetric eigenvalue problem EVP (1.4), writing r=|x| and making the variable transformations of (2.2), it is converted to the one-dimensional weighted p-Laplace eigenvalue problem (EVP)
{−(|v′(t)|p−2v′(t))′=λa(t)|v(t)|p−2v(t),t∈(0,1],v(0)=0,v′(1)=0, | (2.12) |
where v(t)=u(r(t)). Clearly, λ∈R is an eigenvalue of EVP (1.4) if and only if it is an eigenvalue of EVP (2.12). By (2.4) and (2.9), a∈C+(0,1]∩L(I) and ∫10a(s)ds>0. This guarantees that EVP (2.12) has a minimum positive real eigenvalue λp,1, which given by
λp,1=inf{∫10|w′(t)|pdt∫10a(t)wp(t)dt|w∈C1(I),w(0)=0,w′(1)=0,∫10a(t)wp(t)dt≠0}. | (2.13) |
Moreover, λp,1 is simple and has a positive eigenfunction ϕ∈C+(I)∩C1(I). See [13, Theorem 5], [14, Theorem 1.1] or [15, Theorem 1.2]. Hence, λp,1 is also the minimum positive real eigenvalue of EVP (1.4), and ϕ((r0/|x|)(q−1)(N−p)) is corresponding positive eigenfunction.
Now we consider BVP (2.3). Define a closed convex cone K of C(I) by
K={v∈C(I)|v(t)≥t‖v‖C,t∈I}. | (2.14) |
By Lemma 2.2, S(C+(I))⊂K. Let f∈C(R+,R+), and define a mapping F:K→C+(I) by
F(v)(t):=f(v(t)),t∈I. | (2.15) |
Then F:K→C+(I) is continuous and it maps every bounded subset of K into a bounded subset of C+(I). Define the composite mapping by
A=S∘F. | (2.16) |
Then A:K→K is completely continuous by the complete continuity of the operator S:C+(I)→K. By the definitions of S and K, the positive solution of BVP (2.3) is equivalent to the nonzero fixed point of A.
Let E be a Banach space and K⊂E be a closed convex cone in E. Assume D is a bounded open subset of E with boundary ∂D, and K∩D≠∅. Let A:K∩¯D→K be a completely continuous mapping. If Av≠v for every v∈K∩∂D, then the fixed point index i(A,K∩D,K) is well defined. One important fact is that if i(A,K∩D,K)≠0, then A has a fixed point in K∩D. In next section, we will use the following two lemmas in [16,17] to find the nonzero fixed point of the mapping A defined by (2.16).
Lemma 2.4. Let D be a bounded open subset of E with 0∈D, and A:K∩¯D→K a completely continuous mapping. If μAv≠v for every v∈K∩∂D and 0<μ≤1, then i(A,K∩D,K)=1.
Lemma 2.5. Let D be a bounded open subset of E with 0∈D, and A:K∩¯D→K a completely continuous mapping. If ‖Av‖≥‖v‖ and Av≠v for every v∈K∩∂D, then i(A,K∩D,K)=0.
Proof of Theorem 1.1. We only consider the case that (H1)* holds, and the case that (H2)* holds can be proved by a similar way.
Let K⊂C(I) be the closed convex cone defined by (2.14) and A:K→K be the completely continuous mapping defined by (2.16). If v∈K is a nontrivial fixed point of A, then by the definitions of S and A, v(t) is a positive solution of BVP (2.3) and u=v(r0N−2/|x|N−2) is a classical positive radial solution of BVP (1.1). Let 0<R1<R2<+∞ and set
D1={v∈C(I):‖v‖C<R1},D2={v∈C(I):‖v‖C<R2}. | (3.1) |
We prove that A has a fixed point in K∩(¯D2∖D1) when R1 is small enough and R2 large enough.
Since fp0<λp,1, by the definition of fp0, there exist ε∈(0,λp,1) and δ>0, such that
f(u)≤(λp,1−ε)up−1,0≤u≤δ. | (3.2) |
Choosing R1∈(0,δ), we prove that A satisfies the condition of Lemma 2.4 in K∩∂D1, namely
μAv≠v,∀v∈K∩∂D1,0<μ≤1. | (3.3) |
In fact, if (3.3) does not hold, there exist v0∈K∩∂D1 and 0<μ0≤1 such that μ0Av0=v0. By the homogeneity of S, v0=μ0S(F(v0))=S(μ0p−1F(v0)). By the definition of S, v0 is the unique solution of BVP (2.5) for h=μ0p−1F(v0)∈C+(I). Hence, v0∈C1(I) satisfies the differential equation
{−(|v′0(t)|p−2v0′(t))′=μ0p−1a(t)f(v0(t)),t∈(0,1],v0(0)=0,v0′(1)=0. | (3.4) |
Since v0∈K∩∂D1, by the definitions of K and D1,
0≤v0(t)≤‖v0‖C=R1<δ,t∈I. |
Hence by (3.2),
f(v0(t))≤(λp,1−ε)v0p−1(t),t∈I. |
By this inequality and Eq (3.4), we have
−(|v′0(t)|p−2v0′(t))′≤μ0p−1(λp,1−ε)a(t)v0p−1(t),t∈(0,1]. |
Multiplying this inequality by v0(t) and integrating on (0,1], then using integration by parts for the left side, we have
∫10|v′0(t)|pdt≤μ0p−1(λp,1−ε)∫10a(t)v0p(t)dt≤(λp,1−ε)∫10a(t)v0p(t)dt. | (3.5) |
Since v0∈K∩∂D, by the definition of K,
∫10a(t)v0p(t)dt≥‖v0‖Cp∫10tpa(t)dt=R1p∫10tpa(t)dt>0. |
Hence, by (2.13) and (3.5) we obtain that
λp,1≤∫10|v′0(t)|pdt∫10a(t)v0p(t)dt≤λp,1−ε, |
which is a contradiction. This means that (3.3) holds, namely A satisfies the condition of Lemma 2.4 in K∩∂D1. By Lemma 2.4, we have
i(A,K∩D1,K)=1. | (3.6) |
On the other hand, by the definition (1.6) of B, we have
B<[∫1σΨ(∫1stp−1a(t)dt)ds]−(p−1)→B(σ→0+),σ∈(0,1). | (3.7) |
Since fp∞>B, by (3.7) there exists σ0∈(0,1), such that
B0:=[∫1σ0Ψ(∫1stp−1a(t)dt)ds]−(p−1)<fp∞. | (3.8) |
By this inequality and the definition of fp∞, there exists H>0 such that
f(u)>B0up−1,u>H. | (3.9) |
Choosing R2>max{δ,H/σ0}, we show that
‖Av‖C≥‖v‖C,v∈K∩∂D2. | (3.10) |
For ∀v∈K∩∂D2 and t∈[σ0,1], by the definitions of K and D2
v(t)≥t‖v‖C≥σ0R2>H. |
By this inequality and (3.9),
f(v(t))>B0vp−1(t)≥B0‖v‖p−1Ctp−1,t∈[σ0,1]. | (3.11) |
Since Av=S(F(v)), by the expression (2.10) of the solution operator S and (3.11), noticing (p−1)(q−1)=1, we have
‖Av‖C≥Av(1)=∫10Ψ(∫1sa(t)f(v(t))dt)ds≥∫1σ0Ψ(∫1sa(t)f(v(t))dt)ds≥∫1σ0Ψ(∫1sa(t)B0‖v‖p−1Ctp−1dt)ds=Bq−10‖v‖C∫1σ0Ψ(∫1stp−1a(t)dt)ds=‖v‖C. |
Namely, (3.10) holds. Suppose A has no fixed point on ∂D2. Then by (3.10), A satisfies the condition of Lemma 2.5 in K∩∂D2. By Lemma 2.5, we have
i(A,K∩D2,K)=0. | (3.12) |
By the additivity of fixed point index, (3.6) and (3.11), we have
i(A,K∩(D2∖¯D1),K)=i(A,K∩D2,K)−i(A,K∩D1,K)=−1. |
Hence A has a fixed point in K∩(D2∖¯D1).
The proof of Theorem 1.1 is complete.
The authors would like to express sincere thanks to the reviewers for their helpful comments and suggestions. This research was supported by National Natural Science Foundations of China (No.12061062, 11661071).
The authors declare that they have no competing interests.
[1] |
H. Ylönen, Weasels Mustela nivalis suppress reproduction in cyclic bank voles Clethrionomys glareolus, Oikos, 55 (1989), 138–140. https://doi.org/10.2307/3565886 doi: 10.2307/3565886
![]() |
[2] |
S. L. Lima, Stress and decision-making under the risk of predation: recent developments from behavioral, reproductive, and ecological perspectives, Adv. Stud. Behav., 27 (1998), 215–290. https://doi.org/10.1016/S0065-3454(08)60366-6 doi: 10.1016/S0065-3454(08)60366-6
![]() |
[3] |
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
![]() |
[4] | H. Ylönen, B. Jedzrejewska, W. Jedrzejewski, J. Heikilla, Antipredatory behaviour of Clethrionomys voles–'David and Goliath' arms race, Ann. Zool. Fenn., 29 (1992), 207–216. |
[5] |
W. Cresswell, Predation in bird populations, J. Ornithol., 152 (2011), 251–263. https://doi.org/10.1007/s10336-010-0638-1 doi: 10.1007/s10336-010-0638-1
![]() |
[6] |
E. L. Preisser, D. I. Bolnick, The many faces of fear: comparing the pathways and impacts of nonconsumptive predators effects on prey populations, PLoS One, 3 (2008), e2465. https://doi.org/10.1371/journal.pone.0002465 doi: 10.1371/journal.pone.0002465
![]() |
[7] |
H. Ronkainen, H. Ylönen, Behaviour of cyclic bank voles under risk of mustelid predation: do females avoid copulations?, Oecologia, 97 (1994), 377–381. https://doi.org/10.1007/BF00317328 doi: 10.1007/BF00317328
![]() |
[8] |
H. Ylönen, Vole cycles and anti predatory behaviour, Trends Ecol. Evol., 9 (1994), 426–430. https://doi.org/10.1016/0169-5347(94)90125-2 doi: 10.1016/0169-5347(94)90125-2
![]() |
[9] |
H. Ylönen, H. Ronkainen, Breeding suppression in the bank vole as anti predatory adaptation in a predictable environment, Evol. Ecol., 8 (1994), 658–666. https://doi.org/10.1007/BF01237848 doi: 10.1007/BF01237848
![]() |
[10] |
A. Kumar, B. Dubey, Modeling the effect of fear in a prey-predator system with prey refuge and gestation delay, Int. J. Bifurcat. Chaos, 29 (2019), 1950195. https://doi.org/10.1142/S0218127419501955 doi: 10.1142/S0218127419501955
![]() |
[11] |
M. Das, G. P. Samanta, Prey-predator fractional order model with fear effect and group defense, Int. J. Dynam. Control, 9 (2021), 334–349. https://doi.org/10.1007/s40435-020-00626-x doi: 10.1007/s40435-020-00626-x
![]() |
[12] |
X. Wang, L. Zanette, X. Zou, Modelling the fear effect in predator-prey interactions, J. Math. Biol., 73 (2016), 1179–1204. https://doi.org/10.1007/s00285-016-0989-1 doi: 10.1007/s00285-016-0989-1
![]() |
[13] |
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/s11538-017-0287-0 doi: 10.1007/s11538-017-0287-0
![]() |
[14] |
S. Mondal, A. Maiti, G. P. Samanta, Effects of fear and additional food in a delayed predator-prey model, Biophys. Rev. Lett., 13 (2018), 157–177. https://doi.org/10.1142/S1793048018500091 doi: 10.1142/S1793048018500091
![]() |
[15] |
H. Zhang, Y. Cai, S. Fu, W. Wang, Impact of the fear effect in a prey-predator model incorporating a prey refuge, Appl. Math. Comput., 356 (2019), 328–337. https://doi.org/10.1016/j.amc.2019.03.034 doi: 10.1016/j.amc.2019.03.034
![]() |
[16] |
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
![]() |
[17] |
D. Duan, B. Niu, J. Wei, Hopf-hopf bifurcation and chaotic attractors in a delayed diffusive predator-prey model with fear effect, Chaos Soliton. Fract., 123 (2019), 206–216. https://doi.org/10.1016/j.chaos.2019.04.012 doi: 10.1016/j.chaos.2019.04.012
![]() |
[18] |
N. Pal, S. Samanta, J. Chattopadhyay, Revisited Hastings and Powell model with omnivory and predator switching, Chaos Soliton. Fract., 66 (2014), 58–73. https://doi.org/10.1016/j.chaos.2014.05.003 doi: 10.1016/j.chaos.2014.05.003
![]() |
[19] |
S. Pal, N. Pal, S. Samanta, J. Chattopadhyay, Fear effect in prey and hunting cooperation among predators in a Leslie-Gower model, Math. Biosci. Eng., 16 (2019), 5146–5179. https://doi.org/10.3934/mbe.2019258 doi: 10.3934/mbe.2019258
![]() |
[20] |
A. Das, G. P. Samanta, Modeling the fear effect on a stochastic prey-predator system with additional food for the predator, J. Phys. A: Math. Theor., 51 (2018), 465601. https://doi.org/10.1088/1751-8121/aae4c6 doi: 10.1088/1751-8121/aae4c6
![]() |
[21] |
P. Panday, N. Pal, S. Samanta, J. Chattopadhyay, Stability and bifurcation analysis of a three-species food chain model with fear, Int. J. Bifurcat. Chaos, 28 (2018), 1850009. https://doi.org/10.1142/S0218127418500098 doi: 10.1142/S0218127418500098
![]() |
[22] |
A. Sha, S. Samanta, M. Martcheva, J. Chattopadhyay, Backward bifurcation, oscillations and chaos in an eco-epidemiological model with fear effect, J. Biol. Dynam., 13 (2019), 301–327. https://doi.org/10.1080/17513758.2019.1593525 doi: 10.1080/17513758.2019.1593525
![]() |
[23] |
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
![]() |
[24] |
V. Kumar, N. Kumari, Stability and bifurcation analysis of Hassell-Varley prey-predator system with fear effect, Int. J. Appl. Comput. Math., 6 (2020), 150. https://doi.org/10.1007/s40819-020-00899-y doi: 10.1007/s40819-020-00899-y
![]() |
[25] |
Y. Huang, Z. Zhu, Z. Li, Modeling the Allee effect and fear effect in predator-prey system incorporating a prey refuge, Adv. Differ. Equations, 2020 (2020), 321. https://doi.org/10.1186/s13662-020-02727-5 doi: 10.1186/s13662-020-02727-5
![]() |
[26] |
K. Sarkar, S. Khajanchi, Impact of fear effect on the growth of prey in a predator-prey interaction model, Ecol. Compl., 42 (2020), 100826. https://doi.org/10.1016/j.ecocom.2020.100826 doi: 10.1016/j.ecocom.2020.100826
![]() |
[27] |
W. M. Liu, H. W. Hethcote, S. A. Levin, Dynamical behavior of epidemiological models with nonlinear incidence rates, J. Math. Biology, 25 (1987), 359–380. https://doi.org/10.1007/BF00277162 doi: 10.1007/BF00277162
![]() |
[28] | R. O. Peterson, R. E. Page, Wolf Density as a Predictor of Predation Rate, Swedish Wildlife Research, Sweden, 1987. |
[29] |
Q. J. A. Khan, E. Balakrishnan, G. C. Wake, Analysis of a predator-prey system with predator switching, Bull. Math. Biol., 66 (2004), 109–123. https://doi.org/10.1016/j.bulm.2003.08.005 doi: 10.1016/j.bulm.2003.08.005
![]() |
[30] |
P. K. Tiwari, K. A. N. A. Amri, S. Samanta, Q. J. A. Khan, J. Chattopadhyay, A systematic study of autonomous and nonautonomous predator-prey models with combined effects of fear, migration and switching, Nonlinear Dyn., 103 (2021), 2125–2162. https://doi.org/10.1007/s11071-021-06210-y doi: 10.1007/s11071-021-06210-y
![]() |
[31] | L. Perko, Differential Equations and Dynamical Systems, Springer, New York, 2001. |
[32] | S. N. Chow, J. K. Hale, Methods of bifurcation theory, Springer-Verlag, New York, 1982. |
[33] |
L. H. Erbe, H. I. Freedman, V. S. H. Rao, Three-species food-chain models with mutual interference and time delays, Math. Biosci., 80 (1986), 57–80. https://doi.org/10.1016/0025-5564(86)90067-2 doi: 10.1016/0025-5564(86)90067-2
![]() |
[34] | V. Geetha, S. Balamuralitharan, Hopf bifurcation analysis of nonlinear HIV infection model and the effect of delayed immune response with drug therapies, Bound. Value Probl., 2020 (2020). https://doi.org/10.1186/s13661-020-01410-8 |
1. | Bo Yang, Radially Symmetric Positive Solutions of the Dirichlet Problem for the p-Laplace Equation, 2024, 12, 2227-7390, 2351, 10.3390/math12152351 | |
2. | Yongxiang Li, Pengbo Li, Radial solutions of p-Laplace equations with nonlinear gradient terms on exterior domains, 2023, 2023, 1029-242X, 10.1186/s13660-023-03069-y | |
3. | 旭莹 唐, The Existence of Positive Solutions to Quasilinear Differential Equation on Infinite Intervals, 2023, 13, 2160-7583, 2103, 10.12677/PM.2023.137217 |