
A double-time-delay Holling Ⅱ predator model with weak Allee effect and age structure was studied in this paper. First, the model was converted into an abstract Cauchy problem. We also discussed the well-posedness of the model and the existence of the equilibrium solution. We analyzed the global stability of boundary equilibrium points, the local stability of positive equilibrium points, and the conditions of the Hopf bifurcation for the system. The conclusion was verified by numerical simulation.
Citation: Yanhe Qiao, Hui Cao, Guoming Xu. A double time-delay Holling Ⅱ predation model with weak Allee effect and age-structure[J]. Electronic Research Archive, 2024, 32(3): 1749-1769. doi: 10.3934/era.2024080
[1] | Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109 |
[2] | Mengting Sui, Yanfei Du . Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay. Electronic Research Archive, 2023, 31(9): 5124-5150. doi: 10.3934/era.2023262 |
[3] | Yichao Shao, Hengguo Yu, Chenglei Jin, Jingzhe Fang, Min Zhao . Dynamics analysis of a predator-prey model with Allee effect and harvesting effort. Electronic Research Archive, 2024, 32(10): 5682-5716. doi: 10.3934/era.2024263 |
[4] | Yuan Tian, Yang Liu, Kaibiao Sun . Complex dynamics of a predator-prey fishery model: The impact of the Allee effect and bilateral intervention. Electronic Research Archive, 2024, 32(11): 6379-6404. doi: 10.3934/era.2024297 |
[5] | Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150 |
[6] | Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215 |
[7] | Hui Cao, Mengmeng Han, Yunxiao Bai, Suxia Zhang . Hopf bifurcation of the age-structured SIRS model with the varying population sizes. Electronic Research Archive, 2022, 30(10): 3811-3824. doi: 10.3934/era.2022194 |
[8] | San-Xing Wu, Xin-You Meng . Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation. Electronic Research Archive, 2025, 33(2): 995-1036. doi: 10.3934/era.2025045 |
[9] | Zhuo Ba, Xianyi Li . Period-doubling bifurcation and Neimark-Sacker bifurcation of a discrete predator-prey model with Allee effect and cannibalism. Electronic Research Archive, 2023, 31(3): 1405-1438. doi: 10.3934/era.2023072 |
[10] | Xiaowen Zhang, Wufei Huang, Jiaxin Ma, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior. Electronic Research Archive, 2022, 30(7): 2510-2523. doi: 10.3934/era.2022128 |
A double-time-delay Holling Ⅱ predator model with weak Allee effect and age structure was studied in this paper. First, the model was converted into an abstract Cauchy problem. We also discussed the well-posedness of the model and the existence of the equilibrium solution. We analyzed the global stability of boundary equilibrium points, the local stability of positive equilibrium points, and the conditions of the Hopf bifurcation for the system. The conclusion was verified by numerical simulation.
This paper considers the following double-time-delay Holling Ⅱ predation model with weak Allee effect and age-structure:
{dV(t)dt=γV(t)(1−V(t)K)(V(t)m+1)−αV(t)β+V(t)∫+∞0p(t,a)da,∂p(t,a)∂t+∂p(t,a)∂a=−σp(t,a),p(t,0)=ηαV(t−τ1)β+V(t−τ1)∫+∞0δ(a)p(t,a)da. | (1.1) |
The boundary conditions are
V0=χ∈C([−τ1,0],R),p(0,a)=p0(a)∈L1+(0,+∞). |
Here, V(t) and p(t,a) represent the number of predator densities at time t and age a, γ is the rate of prey' intrinsic growth, and Λ=γ+μ, where Λ is the birth and μ is the mortality rate of the prey. K indicates the maximum environmental load of prey, while σ and η indicate the predator mortality rate and the conversion coefficient of predator intake to each prey. In addition, τ1 is the time delay effect, and 0<m<K indicates the survival threshold of the food bait in the weak Allee effect. Meanwhile F(V(t))=αV(t)β+V(t) represents the Holling Ⅱ functional reaction function, where α represents the capture rate and β is a semi-saturated constant. In addition, the reproductive generation of species in the ecosystem generally takes some time to mature to have fertility, i.e., "age-dependent fertility". Fertility δ(a)∈L∞+((0,+∞),A) depends on age, for which it is often assumed that:
Assumption 1.1. Suppose that
δ(a)={δ∗,a≥τ2,0,a<τ2, |
where τ2>0,δ∗>0, and∫+∞0δ(a)e−σada=1.
It is well known that there is a long history of mathematical modeling of the interactions between predators and bait, and that different biological properties are considered in classical models, thus developing various types of predation models. Besides, there is growing evidence that functional responses play a crucial role in the interaction between predators and prey. The Gauss predation math model is given as
{dVdt=γ(1−VK)−F(V)P,dPdt=−dP+ηF(V)P, |
where, V(t) indicates the species density of the prey at time t, and P(t) indicates the species number of predators at time t. The normal constants K and γ denote the environmental capacity and inherent growth rate, respectively. d and η denote the predator mortality and the conversion coefficient of predator intake to each prey, respectively. The functional reaction function F(x) indicates the feeding rate of predators feeding on their prey.
For functional reaction functions, the earliest function form is the Lotka-Volterra type functional response F(x)=αx. This was followed by Holling Ⅱ: F(x)=αxβ+x, Holling Ⅲ: F(x)=αx2β+x2, Holling Ⅳ: F(x)=αx2β+x, and Ivlev-type: F(x)=1−e−αx. For example, a class of Leslie-Gower and Holling Ⅱ predator models are proposed in [1], which gives the global stability of the bounds of understanding, the existence of attracting sets, and the equilibrium points of coexistence:
{dxdt=x(r1−b1x−a1yx+k1),dydt=y(r2−a2yx+k2). |
However, in the real biological world, due to food digestion, reproduction, and other reasons, organisms need reaction time, so the time delay effect is a very important factor in many biological systems. In recent years, many researchers have studied the stability of predation models with time delay, and time delay systems, such as [2,3,4,5,6,7].
At the same time, recent literature has done extensive analysis of predator-prey models of age structure, see [8,9,10]. To study age structure models, a common method is to convert the original model to a time delay differential equation, such as [11]. Another critical method is to turn it into an abstract Cauchy problem, thus applying semi-group theory ([12,13]). In [8], Yang proposed a class of age-structure predation models containing the functional response of the Holling Ⅱ with a prey refuge:
{dv1(t)dt=rv1(t)(1−v1(t)K)−σ1v1(t)+σ2v2(t)−mv1(t)∫+∞0u(t,a)da1+cv1(t),dv2(t)dt=Λ+σ1v1(t)−σ2v2(t)−νv2(t),∂u(t,a)∂t+∂u(t,a)∂a=−μu(t,a),u(t,0)=ηmv1(t)∫+∞0β(a)u(t,a)da1+cv1(t), |
and obtained the Hopf bifurcation of the model at the internal equilibrium point, indicating that the model has a special periodic orbit that bifurcations from the internal equilibrium point when the parameter τ exceeds the bifurcation threshold τ0. The validity of the theoretical analysis is verified by numerical simulation.
In recent years, the Allee effect in predator-prey models has also been widely studied [14,15,16,17]. The Allee effect is defined as the relationship between population size and fitness. In a predator-prey model, the impact of the Allee effect on logistic growth is expressed by including an V−m form multiplier, where m is the Allee threshold. The Allee effect is broadly divided into two categories: the strong Allee effect and the weak Allee effect [18]. The strong Allee effect indicates negative population growth when the population size is below a certain threshold. In contrast, the weak Allee effect indicates a positive population growth trend below a certain threshold. In [16], the author considered a kind of generalized Holling Ⅲ-type functional reaction, predation model with weak Allee effect, and explored the existence conditions of model equilibrium and singularity, as well as some properties of equilibrium stability:
{dxdt=(r(1−xK)(x−m)−qxyx2+bx+a)x,dydt=s(1−ynx)y. |
Inspired by the above work, in this paper we will study the dynamical behavior of the predation model (1.1) of the double time-delay Holling Ⅱ functional response function with weak Allee effect and age structure.
The organization plan for this article is as follows: In Section 2, the transformation of Cauchy's problem is given, and the system well-posedness is obtained. In Section 3, the equilibrium solution of the system is studied, and the linearized system is obtained. In Section 4, the dynamical behavior of the system is studied. In Section 5, some numerical simulations and discussions are conducted.
First we normalize τ2 in the system (1.1), which gives
˜t=tτ2,˜a=aτ2, |
and we consider the distribution
˜V(˜t)=V(τ2˜t),˜p(˜t,˜a)=τ2p(τ2˜t,τ2˜a). |
After the wave is removed, system (1.1) becomes
{dV(t)dt=τ2[γmKV(t)(K−V(t))(V(t)+m)−αV(t)β+V(t)∫+∞0p(t,a)da],∂p(t,a)∂t+∂p(t,a)∂a=−τ2σp(t,a),p(t,0)=τ2ηαV(t−τ1τ2)β+V(t−τ1τ2)∫+∞0δ(a)p(t,a)da, | (2.1) |
where
V0=ˉχ∈C([−τ1τ2,0],R),p(0,a)=p0(a)∈L1((0,+∞),R). |
The new function δ(a) is
δ(a):=δ∗1[1,+∞](a)={δ∗,a≥1,0,otherwise, |
and
∫+∞τ2δ∗e−σada=1. |
i.e., δ∗=σe−στ2,(τ2>0).
Based on the integral semigroup theory, the suitability of the solution of system (2.1) is discussed below. For this purpose, (2.1) will be rewritten as the abstract cauchy problem(ACP). First, two lemmas about operator semigroups are introduced.
Lemma 2.1. [19,20] Let (G,T(G)) be the Hille-Yosida operator on the Banach space Y, A∈L(Y) is the set of all bounded linear operators on Y, and C=G+A are the Hille-Yosida operators.
Lemma 2.2. [19,20] Let G0 be part of the operator G on Y0:=¯T(G), defined as: G0x=Gx, where x∈T(G0)={x∈T(G):Gx∈Y0}. If (G,T(G)) is the Hille-Yosida operator on the Banach space, then (G0,T(G0)) generates a C0-semigroup on Y0.
First, let
V(t)=∫+∞0v(t,a)da, |
and model (2.1) is converted to:
{∂v(t,a)∂t+∂v(t,a)∂a=−τ2μv(t,a),v(0,a)=v0∈L1((0,+∞),R), |
and
v(t,0)=τ2R(v(t,a),p(t,a)), |
where
R(v(t,a),p(t,a))=Λ∫+∞0v(t,a)da+γ(K−m)mK(∫+∞0v(t,a)da)2−γmK(∫+∞0v(t,a)da)3−α∫+∞0v(t,a)da∫+∞0p(t,a)daβ+∫+∞0v(t,a)da. |
Further, let
w(t,a)=(v(t,a)p(t,a)), |
then system (2.1) becomes
{∂w(t,a)∂t+∂w(t,a)∂t=−Mw(t,a),w0(θ,a)=(v0(t,a)p0(t,a))∈C([−τ1τ2,0],L1((0,+∞),G2)),w(t,0)=A(wt(θ,a)), | (2.2) |
where
M=(τ2μ00τ2d),A(wt(θ,a))=(τ2R(v(t,a),p(t,a))τ2ηαV(t−τ1τ2)⋅∫+∞0δ(a)p(t,a)daβ+V(t−τ1τ2)). |
Here, we introduce the Banach space
Y:=R2×L1((0,+∞),R2). |
We have the following usual product norm
‖(φg)‖=‖φ‖R2+‖g‖L1,(φg)∈Y. |
Further, G:T(G)⊂Y→Y is defined as
G(0g)=(−g(0)−g′−Mg), |
with domain
T(G)={0R2}×W1,1((0,+∞),R2), |
and then
Y0:=¯T(G)={0R2}×L1((0,+∞),R2). |
Next, we introduce the space
IG:={(ζ(⋅)ρ(⋅))∈I([−τ1τ2,0],X):ζ(0)=0}, |
and define the map H:IG→Y as
H((ζ(⋅)ρ(⋅)))=(A(ρ(⋅))0L1), |
where
A(ρ(⋅)) = (τ2R(ρ1(0)(a),ρ2(0)(a))τ2ηα∫+∞0ρ1(−τ1τ2)(a)da⋅∫+∞0δ(a)ρ2(0)(a)daβ+∫+∞0ρ1(−τ1τ2)(a)da),ρ(⋅)=(ρ1(⋅)ρ2(⋅)). |
Let h(t)=(0R2w(t)), where w(t)=w(t)(a)=w(t,a), then system (2.2) becomes
{dh(t)dt=Gh(t)+H(ht),h0=χ∈IG, | (2.3) |
where ht∈IG,ht(θ)=h(t+θ), and h0(θ)=(0w0(θ,⋅)). Obviously this is an abstract time delay differential equation, which is further rewritten (2.3) as the ACP for applying the theory of the integral semigroup. Define y∈I([0,+∞]×[−τ1τ2,0]:Y), where y(t,θ)=h(t+θ),t≥0, and θ∈[−τ1τ2,0]. Thus, we can get the following equation:
{∂y(t,θ)∂t−∂y(t,θ)∂θ=0,θ∈[−τ1τ2,0],∂y(t,0)∂θ=Gy(t,0)+H(y(t,⋅)),t≥0,y(0,⋅)=h0∈IG. | (2.4) |
Below we take the product of space Y and space I as the state space Z, so
Z=Y×I,I:=I([−τ1τ2,0],Y) |
and the usual product norm
‖(yχ)‖=‖y‖Y+‖χ‖C,(yχ)∈Z. |
Therefore, system (2.4) can be rewritten as the Cauchy problem of an abstract non-dense definition, and the linear operator T:T(L)⊂Z→Z is defined as follows:
L(0Yχ)=(−χ′(0)χ′),(0Yχ)∈T(L),T(L)={0Y}×{χ∈I1([−τ1τ2,0],Y),χ(0)∈T(G)}. |
Since Z0:=¯T(L)={0Y}×IG≠Z, we obtain that L is apparently a non-dense linear operator defined in Z. Furthermore, we take the following operator H:Z0→Z:
H(0Yχ)=(H(χ)0IG). |
Finally, let z(t)=(0y(t)),y(t)=y(t)(θ)=y(t,θ). Then, (2.4) is the non-dense definition of the Cauchy problem
{dz(t)dt=Lz(t)+H(z(t)),z(0)=(0Yh0)∈Z0. | (2.5) |
[19,21] studied the global existence and uniqueness of solutions containing (2.5).
Let
Ω:={λ∈C:Re(λ)>−ς},ς:=min{τ2μ,τ2d}>0, |
and then the following conclusion holds:
Theorem 2.1. For G and L, we can get:
(i) If λ∈Ω, then λ∈w(G) and
(λ−G)−1(˜φ˜g)=(0g)⇔g(a)=e−∫a0(λI+M)dl˜φ+∫a0e−∫a0(λI+M)dl˜g(q)dq, |
where (˜φ˜g)∈Y,(0g)∈T(G);
(ii) w(L)=w(G). Also, for the each λ∈w(L), we can obtain the explicit formula of L' resolvent
(λ−L)−1(˜x˜χ)=(0χ)⇔χ(θ)=eλθ(λ−G)−1(˜χ(0))+∫0θeλ(θ−q)˜χ(q)dq; |
(iii) L and G are Hille−Yosida operators on Z and Y, respectively.
Proof. The first two proof methods are shown in [22], and here we only need to prove (iii). Let λ∈(−ς,+∞), then from (i) we can get
g(a)=e−(λI+M)a˜φ+∫a0e−(λI+M)(a−q)˜g(q)dq, |
the integral for a, then
‖g‖L1=∫+∞0|e−(λa+τ2μa)˜α1+∫a0e−(λ(a−q)+τ2μ(a−q))˜g1(q)dq|da+∫+∞0|e−(λa+τ2da)˜α1+∫a0e−(λ(a−q)+τ2d(a−q))˜g2(q)dq|da≤2∫+∞0e−(λ+ς)ada|˜α|+2∑i=1∫+∞0∫a0e−(λ+ς)(a−q)|˜gi(q)|dqda=2∫+∞0e−(λ+ς)ada|˜α|+2∑i=1∫+∞0∫aqe−(λ+ς)(a−q)da|˜gi(q)|dq≤2λ+ς(|˜α|+‖g‖L1), |
‖(λI−G)−1‖≤2λ+ς,(λ>−ς). |
This indicates that (G,T(G)) is the Hille-Yosida operator. By the proof method in [21], we get that (L,T(L)) is a Hille-Yosida operator.
Finally, let Z0+:=Z0∩Z+, and
Z+:=Y+×I([−τ1τ2,0],Y+),Y+:=R2+×L1((0,+∞),R2+). |
Since both G and L are Hille−Yosida operators, using the theory about the integral semigroup, we can get the well-posedness of model (2.5) as follows:
Theorem 2.2. There exists a unique continuous semigroup {U(t)}t≥0 on Z0+ such that for any z∈Z0+, we have that t→U(t)z is the unique integral solution for the following problem:
{dU(t)zdt=LU(t)z+H(U(t)z),U(0)z=z, |
or
U(t)=z+L∫t0U(q)zdq+∫t0H(T(q)z)dq,t≥0. |
Now we prove the existence and linearization of the equilibrium points of system (2.5). Suppose ˉz=(0Yˉχ)∈T(L) is the steady state solution of system (2.5), where
ˉχ=(ˉς(⋅)ˉρ(⋅))∈I1([−τ1τ2,0],Y),ˉχ(0)∈T(L),ˉρ(⋅)=(ˉρ1(⋅)ˉρ2(⋅)), |
then we have Lˉz+H(ˉz)=0, that is,
{H(ˉχ)−ˉχ′(0)+Gˉχ(0)=0,ˉχ′=0. | (3.1) |
The following conclusions can be obtained from (3.1):
Theorem 3.1. System (2.5) always has equilibrium:
ˉz0=(0Y(ˉς0(⋅)(ˉρ01(⋅)ˉρ02(⋅)))),ˉz1=(0Y(ˉς1(⋅)(ˉρ11(⋅)ˉρ12(⋅)))), |
where
ˉς0(θ)=ˉς0(θ)=0R2,(ˉρ01(θ)(a)ˉρ02(θ)(a))=(00),(ˉρ11(θ)(a)ˉρ12(θ)(a))=(τ2μKe−τ2μa0). |
In addition, (2.5) has only the positive equilibrium solution
ˉz∗=(0Y(ˉς(⋅)(ˉρ1(⋅)ˉρ2(⋅)))), |
if
αη>1,K(αη−1)>β, |
where
ˉς(θ)=0R2,(ˉρ1(θ)(a)ˉρ2(θ)(a))=(τ2μβαη−1e−τ2μaτ2σγβη[K(αη−1)−β][β+m(αη−1)]mK(αη−1)3e−τ2σa). |
Therefore, the following theorem holds for system (1.1):
Theorem 3.2. (i) System (1.1) always has equilibriums E0(0,0),E1(m,0),E2(K,0);
(ii) when αη>1,K(αη−1)>β, the positive equilibrium point E∗(V∗,p∗(a)) exists in the system, where
V∗=βαη−1,p∗(a)=σγβη[K(αη−1)−β][β+m(αη−1)]mK(αη−1)3e−σa. |
Next, we linearize system (2.5) at the equilibrium point, set ˉz as the steady state of the system (2.5), let ϖ(t)=z(t)−ˉz, and replace it with (2.5). Then,
{ϖ(t)dt=Lϖ(t)(t)+H(ϖ(t)(t)+ˉz)−H(ϖ(t)(t)),t≥0,ϖ(t)(0)=(0ω0−ˉχ):=ϖ(t)0∈T(L). |
Thus, the linearization system around ˉz is as follows:
{dϖ(t)(t)dt=Lϖ(t)(t)+TH(ϖ(t))(ϖ(t)(t)),t≥0,ϖ(t)(0)=ϖ(t)0∈T(L), | (3.2) |
with
TH(ˉz)(0Yχ)=(TH(ˉχ)(χ)0CG),(0Yχ)∈T(L),χ=(ς(⋅)ρ(⋅)), | (3.3) |
and
TH(ˉχ)(χ)=(TH(ˉρ)(ρ)0L1), |
where
TH(ˉρ)(ρ)=(τ2Λ+2τ2γ(K−m)mKM2−3τ2γmKM22−τ2αβN2(β+M2)2−τ2αM2β+M200)∫+∞0ρ(0)(a)da+(00τ2αηβN1(β+M1)20)∫+∞0ρ(−τ1τ2)(a)da+(000τ2αηM1β+M1)∫+∞0δ(a)ρ(0)(a)da, |
M1=∫+∞0ˉρ1(−τ1τ2)(a)da,M2=∫+∞0ˉρ1(0)(a)da,N1=∫+∞0δ(a)ˉρ2(0)(a)da,N2=∫+∞0ˉρ2(0)(a)da. |
According to Lemma 2.1 and Theorem 2.1, we get Theorems 3.3 and 3.4:
Theorem 3.3. L+TH(ˉz) is a Hille-Yosida operator.
Then, by Lemma 2.2, we can get the following:
Theorem 3.4. (L,T(L)),(L+TH(ˉz),T(L+TH(ˉz))) generate C0-semigroups (I(t))t≥0, (J(t))t≥0 on space Z0, respectively.
Based on the proof of Theorem 2.1, the Hille−Yosida estimate domain is ‖I(t)‖≤e−dt. Moreover, TH(ˉz)I(t):Z0→Z is clearly compact for any t>0. Then, we have
J(t)=eTH(ˉz)tI(t)=I(t)++∞∑k=1(TH(ˉz)t)kk!I(t). |
And then we can get (J(t))t≥0 is quasi-compact. According to [18], the quasi-compact related conclusion for strong continuous semigroups, when all eigenvalues of L+TH(ˉz) are negative, then for ˉd>0, when t→+∞, eˉdt‖J(t)‖→0.
Theorem 3.5. The solution semigroup T(t) of the system (2.5) satisfies the following: the solution of the steady state ˉz(t) is locally asymptotically stable (LAS), when all eigenvalues of L+TH(ˉz) have strictly negative real parts; the solution of the steady state ˉz(t) is unstable, when the presence of L+TH(ˉz) has a strictly positive real eigenvalue.
Obviously, E0 and E1 are unstable equilibrium points. Next we consider E2's stability.
Theorem 4.1. When αηKβ+K<1, the equilibrium state ˉz1 of system (2.5), i.e., the equilibrium point E1(K,0) of system (1.1), is LAS; when αηKβ+K>1, E1(K,0) is unstable.
Proof. Let
ˉz(t)=z(t)−ˉz1=(0Y0R2˜w(t)(⋅))T, |
where
˜w(t)(⋅)=(˜v(t)(⋅)˜p(t)(⋅))=(v(t,⋅)−τ2μKe−τ2μap(t,⋅)). |
From this, the linearized system ˉz1 can be written as
{∂˜w(t,a)∂t+∂˜w(t,a)∂a=−M˜w(t,a),˜w(t,0)=Q1∫+∞0˜w(t,a)da+Q2∫+∞0δ(a)˜w(t,a)da, |
where
Q1=(τ2Λ+2τ2γ(K−m)m−3τ2γKm−τ2mKβ+K00),Q2=(000τ2αηKβ+K), |
which is equivalent to
{˜V′(t)=−τ2γ(m+K)m˜V(t)−τ2mKβ+K∫+∞0˜p(t,a)da,∂˜p(t,a)∂t+∂˜p(t,a)∂a=−τ2σ˜p(t,a),˜p(t,0)=τ2αηKβ+K∫+∞0δ(a)˜p(t,a)da, | (4.1) |
where ˜V(t)=∫+∞0˜v(t,a)da.
Let ˜V(t)=˜V0eλt,˜p(t,a)=˜p0(a)eλt, and substituting this into (4.1), the characteristic equations of (4.1)
Δ0(λ)=(τ2αηKβ+K∫+∞0δ(a)e−(λ+τ2σ)ada−1)(λ+τ2γ(m+K)m)=f0(λ)g0(λ)=0. |
Let f0(λ)=τ2ηαβKβ+K∫+∞0δ(a)e−(λ+τ2σ)ada−1. Then
f0(0)=ηαβKβ+K−1,limλ→∞f0(λ)=−1. |
Obviously, the root of g0(λ)=0 is negative, and for f0(λ)=τ2αηKβ+K∫+∞0δ(a)e−(λ+τ2σ)ada−1, we have
f0(0)=αηKβ+K−1,limλ→∞f0(λ)=−1. |
Because f0(λ) is strictly decreasing and satisfies continuous real functions, we have:
When ηαKβ+K−1>0, f0(λ)=0 has at least one positive root, and E2 is unstable.
When ηαKβ+K−1<0, f0(λ)=0 has no complex solution with real root and no negative, suppose that λ0=θ+ωi is the solution, so
1=|f(λ0)+1|=|τ2αηKβ+K∫+∞0δ(a)e−(θ+τ2σ)a−ωaida−1|≤τ2αηKβ+K∫+∞0δ(a)e−(θ+τ2σ)ada=f0(θ)+1≤f0(0)+1=τ2αηKβ+K<1. |
Clearly, this is contradictory, so the solution of the characteristic equation must have negative real parts, that is, when ηαKβ+K−1<0, f0(λ)=0, and E2 is LAS.
This section demonstrates the global stability of E2 using asymptotic autonomous semigroup theory.
Theorem 4.2. When ηαKβ+K−1<0, E2 is globally asymptotically stable.
Proof. From dV(t)dt of system (2.1), we can obtain
dVdt≤τ2γV(1−VK)(Vm+1), |
and, by the comparison principle, we have:
limt→∞(supV(t))≤K. |
Therefore, for any κ>0, there exists t1 such that V(t−τ1τ2)≤K+κ, when t≥t1+τ1τ2, and then
p(t,0)≤τ2ηαK+κβ+(K+κ)∫+∞0δ(a)p(t,a)da≤τ2ηα∫+∞0δ(a)p(t,a)da,(t≥t1+τ1τ2). |
Now, we consider the following system:
{∂ˆp∂t+∂ˆp∂a=−τ2σˆp,ˆp(t,0)=τ2ηα∫+∞0δ(a)ˆp(t,a)da. | (4.2) |
Using the same method as in Theorem 4.1, the solution of (4.2) exists in the form of ˆp(t,a)=ˆp0(a)eλ0t, where ˆp0(a) is non-negative and λ0 is the root of the characteristic equation of (4.2), i.e.,
Δ0(λ0)=τ2ηα∫+∞0δ(a)e−(λ0+τ2σ)ada−1=0. |
From the second equation of (2.1), we can get
p(t,a)={p(t−a,0)e−τ2σa,a≤t,p0(a−t)e−τ2σt,a<t, |
namely p(t,a)≤ˆp(t,a) for t≥t1+τ1τ2. So, p(t,a)≤ˆp0(a)eλ0t.
From Theorem 4.1, when ηαKβ+K<1, limt→∞p(t,a)=0. Thus, when t→∞, dVdt of (2.1) converges to
dˆVdt=τ2γˆV(1−ˆVK)(ˆVm+1), |
which illustrates that limt→∞ˆV(t)=K.
Applying the related theories from [23], we can get limt→∞V(t)=K. Hence, when ηαKβ+K<1, E2 is globally asymptotically stable.
First, we need to obtain the characteristic equation of system (3.2). Let K=TF(ˉz), where ˉz represents the equilibrium state of system (2.5).
Now, we note ˉA=λI−(L+K), ˉB=I−K(λI−L)−1, ˉC=λI−L. Suppose that λ∈Ω. Because ˉC is reversible, then ˉA is equivalent to
ˉA=ˉBˉC. | (4.3) |
From this, we get that
ˉA is reversible and ˉB is reversible. |
If ˉB is reversible, then
ˉA−1=ˉB−1ˉC−1 |
Applying Theorem 2.1 and (3.3), we can get that, for
ξ=(ϑg),˜ξ=(˜ϑ˜g)∈Y,χ∈(ς(⋅)ρ(⋅)),˜χ∈(˜ς(⋅)˜ρ(⋅))∈C1([−τ1τ2,0],Y), |
we have
ˉB(ξχ)=(˜ξ˜χ), | (4.4) |
which is equivalent to
{ξ−TH(˜χ)(eλθ(λI−G)−1(χ(0)+ξ)+∫0θeλ(θ−s)χ(s)ds)=˜ξ,χ=˜χ, |
i.e.,
{(I−TH(˜χ)(eλθ(λI−G)−1))ξ=˜ξ+TH(˜χ)(eλθ(λI−G)−1χ(0)+∫0θeλ(θ−s)χ(s)ds),χ=˜χ, |
Let
(I−TH(˜χ)(eλθ(λI−G)−1))ξ=(χ1χ2), | (4.5) |
and
(χ1χ2)=(˜ϑ+TA(ˉρ)(eλθ(λI−G)−1˜ρ(0)+∫0θeλ(θ−s)˜ρ(s)ds)˜g), |
then we have
(ϑg)−(TA(ˉρ)[eλθ(e−∫a0(λI−G)−1dlϑ+∫a0e−∫a0(λI−G)−1dlg(s)ds)]0)=(χ1χ2). |
From this, we can derive that
{(I−TA(ˉρ)(eλθe−∫a0(λI+T)dl))ϑ=χ1+TA(ˉρ)(eλθe−∫as(λI+T)dlg(s)ds).g=χ2. | (4.6) |
Let
Δ(λ)=I−TA(ˉρ)(eλθe−∫a0(λI+T)dl),ˉΦ(λ,χ2)=TA(ˉρ)(eλθ∫a0e−∫as(λI+T)dlχ2(s)ds). |
From the first equation of (4.6), we can get Δ(λ)ϑ=χ1+ˉΦ(λ,χ2). That is, when Δ(λ) is reversible, we have
ϑ=(Δ(λ))−1(χ1+ˉΦ(λ,χ2)). |
Thus, ˉA is reversible, i.e.,
ˉB is reversible ⇔Δ(λ) is reversible. |
So, Theorem 4.3 can be obtained:
Theorem 4.3. The following conclusion holds: σ(L+K)∩Ω=σP(L+K)∩Ω={λ∈Ω:det(Δ(λ))=0}, and if λ∈w(L+K)∩Ω, then the resolvents formula is
(λI−(L+K))−1(˜x˜χ)=(0eλθ(λI−G)−1(˜χ(0)+x)+∫0θeλ(θ−s)˜χ(s)ds), | (4.7) |
where
x=((Δ(λ))−1[˜ϑ+TA(ˉρ)(eλθ(λI−G)−1˜ρ(0)+∫0θeλθ−λs˜ρ(s)ds)+ˉΦ(λ,˜g)]˜g). | (4.8) |
Proof. Because λ∈Ω with det(Δ(λ))≠0, we can get that (I−TH(ˉχ)(eλθ(λI−G)−1)) is reversible, and then from (4.5) we can get
(I−TH(ˉχ)(eλθ(λI−G)−1))−1(χ1χ2)=x, |
and
x=((Δ(λ))−1(χ1+ˉΘ(λ,χ2))χ2). |
Thus, ˉB is reversible for any (˜x˜χ)∈Z,
ˉB(˜x˜χ)=(xχ), |
where x as shown in formula (4.8), χ=˜χ. Therefore, we obtain (4.7) and
{λ∈Ω:det(Δ(λ))=0}⊂w(L+K)∩Ω,σ(L+K)∩Ω⊂{λ∈Ω:det(Δ(λ))=0}. |
Further, take λ∈Ω:det(Δ(λ))=0, and according to (4.3), there exists (0χ0)∈T(L),(0χ0)≠0, so
(λ−(L+K))(0χ0)=0, | (4.9) |
is established if and only if there exists ( ⌣ x0 ⌣ χ0)∈Z, and ( ⌣ x0 ⌣ χ0)≠0 satisfies
[I−K(λI−L)−1]( ⌣x0 ⌣χ0)=0. | (4.10) |
Let the (˜ξ˜χ) of the Eq (4.4) be equal to 0, and then we can obtain the existence of ( ⌣ x0 ⌣ χ0)∈Z∖{0} is equivalent to (4.10), where ⌣ x0=( ⌣ ϑ0 ⌣ f0) satisfies
{Δ(λ) ⌣ ϑ0=0, ⌣g0=0, ⌣ χ0=0. |
Thus, (4.9) has solutions if and only if there exists ⌣ ϑ0≠0, so Δ(λ) ⌣ ϑ0=0, and λ∈σP(L+K). Therefore, {λ∈Ω:det(Δ(λ))=0}⊂σP(L+K).
The above analysis shows that det(Δ(λ))=0 is the characteristic equation of (3.2) about ˉz∗.
Below, we analyze the stability at E∗ and the Hopf bifurcation's existence. Due to the complexity of τ1≠τ2, only τ1=τ2=τ is considered below, and then the characteristic equation is:
det(Δ(λ))=λ2+τp1λ+τ2p0+(τq1λ+τ2q0)e−λ(λ+στ)(λ+μτ)=f(λ)g(λ)=0, |
where
p1=σ−[γ+2βγ(K−m)mK(αη−1)−3γβ2mK(αη−1)2−γ[K(αη−1)−β][β+m(αη−1)]Kmηα(αη−1)],p0=−σ[γ2βγ(K−m)mK(αη−1)−3γβ2mK(αη−1)2−γ[K(αη−1)−β][β+m(αη−1)]Kmηα(αη−1)],q1=−σ,q0=σ[γ+2βγ(K−m)mK(αη−1)−3γβ2mK(αη−1)2]. |
Let λ=τϑ, then
f(λ)=f(τϑ)=τ2[ϑ2+p1ϑ+p0+(q1ϑ+q0)e−τϑ]=τ2h(ϑ). | (4.11) |
Because of g(λ)≠0, there is
{λ∈Ω:det(Δ(λ))=0}={τϑ∈Ω:h(ϑ)=0}. |
First, when τ=0, we have h(ϑ)=ϑ2+(p1+q1)ϑ+(p0+q0)=0, where p0+q0=σγ[K(αη−1)−β][β+m(αη−1)]Kmηα(αη−1)>0, hence we get the following theorem:
Theorem 4.4. When τ=0, if p1+q1>0, then E∗ is locally asymptotically stable; otherwise, it is unstable.
This section considers the Hopf bifurcation problem when τ>0. Since the age structure model studied is infinite and the central manifold theory needs to be applied to the abstract non-dense Cauchy problem, we can simplify the system considering the finite-dimensional equation on the central manifold. Thus, the Hopf bifurcation theorem of Hassard remains valid. Therefore, we will use Hassard's theorem directly below to explore the existence of the Hopf bifurcation.
Let τ>0, so the root of h(ϑ)=0 has continuous dependence on τ0. As τ increases, the root of h(ϑ)=0 can pass through the imaginary axis to the right. Let ϑ=iω(ω>0) be the purely imaginary roots of h(ϑ)=0 and substitute it into h(ϑ)=0, which gives
−ω2+ip1ω+p0+iq1ωe−iωτ+q0e−iωτ=0, |
disassociating the real part and imaginary part,
{−ω2+p0=−q1ωsinωτ−q0cosωτ,p1ω=q0sinωτ−q1ωcosωτ, |
i.e.,
{sinωτ=q1ω3+(p1q0−p0q1)ωq20+q21ω2,cosωτ=(q0−p1q1)ω2−p0q0q20+q21ω2, |
and
(ω2−p0)2+p21ω2=q21ω2+q20, |
which is
ω4+(p21−2p0−q21)ω2+(p20−q20)=0. | (4.12) |
Let Θ=ω2, then above equation becomes
Θ2+(p21−2p0−q21)Θ+(p20−q20)=0. |
Due to p0+q0>0,
p21−2p0−q21=[γ+2βγ(K−m)mK(αη−1)−3γβ2mK(αη−1)2−γ[K(αη−1)−β][β+m(αη−1)]Kmηα(αη−1)]2>0, |
and, when p0−q0<0, the above equation has the sole positive root, and denote it as Θ∗. That means (4.12) has the only positive root ω∗=√Θ∗, hence h(ϑ)=0,(τ=τk) has a pair of purely imaginary roots, with
τk={1ω∗(arccos(q0−p1q1)ω2∗−p0q0q20+q21ω2∗+2kπ),c≥0,1ω∗(2π−arccos(q0−p1q1)ω2∗−p0q0q20+q21ω2∗+2kπ),c<0, |
and
c=q1ω3∗+(p1q0−p0q1)ω∗q20+q21ω2∗. |
Lemma 4.5. In the case of Assumption 1.1 holding, when αη>1, p1+q1>0 and p0−q0<0, then
dh(ϑ)dϑ|ϑ=iω∗≠0, |
and, at this time, ϑ=iω∗ is the unique root of h(ϑ)=0.
Proof. From (4.11), we can get that
dh(ϑ)dϑ=2iω∗+p1+q1e−iω∗τk−q0τke−iω∗τk−iq1ω∗τke−iω∗τk. |
Because of h(ϑ)=0, we derive that
[2ϑ+p1+q1e−ϑτ−τ(q1ϑ+q0)e−ϑτ]dh(ϑ)dϑ=ϑ(q1ϑ+q0)e−ϑτ. |
If dh(ϑ)dϑ|ϑ=iω∗=0 is correct, then iω∗(iq1ω∗+q0)e−iω∗τ=0, i.e., iq1ω∗+q0=0, hence q1=q0=0. Because q1=−σ<0, which contradicts the conclusion, dh(ϑ)dϑ|ϑ=iω∗≠0.
Let ϑ(τ)=ˉα(τ)+iˉω(τ) become the root of h(ϑ)=0 where ˉα(τk)=0,ˉω(τk)=ω∗. Evaluating τ on both sides of h(ϑ)=0, we get that
(dϑdτ)−1|ϑ=iω∗=2ϑ+p1+q1e−ϑτ−τ(q1ϑ+q0)e−ϑτϑ(q1ϑ+q0)e−ϑτ|ϑ=iω∗=(−τϑ+q1ϑ(q1ϑ+q0)−2ϑ+p1ϑ(ϑ2+p1ϑ+p0))|ϑ=iω∗, |
so
Re((dϑdτ)−1|ϑ=iω∗)=−q21q21ω2∗+q20+2ω2∗+p21−2p0p21ω2∗+(p0−ω2∗)2=2ω2∗+p21−2p0−q21q21ω2∗+q20. |
Besides,
ω2∗=−(p21−2p0−q21)+√(p21−2p0−q21)2−4(p20−q20)2. |
Replace ω2∗ with Re((dϑdτ)−1|ϑ=iω∗), so
sign((dRe(ϑ)dτ)−1|τ=τk)=sign(Re((dϑdτ)−1|ϑ=iω∗))=sign(2ω2∗+p21−2p0−q21q21ω2∗+q20)>0. |
According to the correlation theorem of the Hopf bifurcation in [24], we get Theorem 4.6:
Theorem 4.6. In the case of Assumption 1.1 holding, when αη>1,K(αη−1)>β, p1+q1>0, and p0−q0<0, then
(i) when τ∈[0,τ0), E∗ is asymptotically stable, and when τ>τ0, it is unstable;
(ii) when τ=τk, system (1.1) undergoes a Hopf bifurcation at the equilibrium E∗.
This section uses the MATLAB software to simulate the model numerically. First, the parameters are:
γ=1,K=20,α=1.01,β=4,σ=0.02,η=1.1,m=5. |
By calculating, we can get ηαKb+K−1=−0.074<0, which satisfies the condition of Theorem 4.2, that E2(20,0) is globally asymptotically stable at this time.
Let τ1=1,τ2=2. The available time series diagrams and phase diagrams are shown in Figure 1, and E2 are globally stable at this time.
Next, let the parameters become
γ=1,K=20,α=1.082,β=1.09,σ=0.4,η=1.0045,m=5, |
and set V(0)=14,p(0,a)=16e−a. By calculating, we have
V∗=12.548,p∗(a)=6.593e−0.4a,∫+∞0p∗(a)da=16.48,ηα−1=0.086>0,K(ηα−1)−b=0.647>0,p1+q1=0.027>0,p0−q0=−0.005<0, |
satisfying the condition of Theorem 4.6, and we can get τ0=3.185. First, let τ1=τ2=2. E∗(12.548,6.593e−0.4a) is asymptotically stable at this time, and the available time series diagrams and phase diagrams are shown in Figure 2. Second, let τ1=τ2=4, E∗ pass through a Hopf bifurcation. For aesthetics, we modify the initial value to V(0)=25,p(0,a)=5.395e−a. The system has periodic solutions, and the available time series diagrams and phase diagrams are shown in Figure 3.
Finally, we observe the dynamics of the system under the current parameter conditions τ1≠τ2, and let τ1=1,τ2=0. E∗ is asymptotically stable at this time, and the available time series diagrams and phase diagrams are shown in Figure 4.
As can be seen from the above analysis, the time delay effects of predation processes, energy conversion, reproductive reproduction, etc., can cause changes in the dynamics behavior of the predation system over a later period of time. At a time when the delay is less than a certain threshold, the final size of the two species is in a state of coexistence and tends to stabilize. When this threshold is exceeded, the two species still coexist, but, because the system undergoes a Hopf bifurcation, the number of the two species is subject to periodic oscillations. We know that the weak Allee effect affects the population size of the predator system by influencing its time delay threshold, which is essential for the study of the predator system.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work is supported by National Natural Science Foundation of China Grants 12071268 and 11971281, Inner Mongolia Autonomous Region University Science and Technology Research Project NJZY22036 and by Innovation Capability Support Program of Shaanxi Province (Program No. 2023-CX-TD-61).
The authors declare there are no conflicts of interest.
[1] |
M. A. Aziz-Alaoui, M. D. Okiye, Boundedness and global stability for a predator-prey model with modified Leslie-Gower and Holling-type Ⅱ schemes, Appl. Math. Lett., 16 (2003), 1069–1075. https://doi.org/10.1016/S0893-9659(03)90096-6 doi: 10.1016/S0893-9659(03)90096-6
![]() |
[2] |
F. A. Rihan, H. J. Alsakaji, Stochastic delay differential equations of three-species prey-predator system with cooperation among prey species, Discrete Contin. Dyn. Syst. - Ser. S, 15 (2022), 245–263. https://doi.org/10.3934/dcdss.2020468 doi: 10.3934/dcdss.2020468
![]() |
[3] |
Y. H. Fan, W. T. Li, Permanence for a delayed discrete ratio-dependent predator–prey system with Holling type functional response, J. Math. Anal. Appl., 299 (2004), 357–374. https://doi.org/10.1016/j.jmaa.2004.02.061 doi: 10.1016/j.jmaa.2004.02.061
![]() |
[4] |
Y. Zhang, Q. L. Zhang, Dynamical analysis of a delayed singular prey–predator economic model with stochastic fluctuations, Complexity, 19 (2014), 23–29. https://doi.org/10.1002/cplx.21486 doi: 10.1002/cplx.21486
![]() |
[5] |
U. Das, T. K. Kar, Bifurcation analysis of a delayed predator-prey model with Holling type Ⅲ functional response and predator harvesting, J. Nonlinear Dyn., 2014 (2014), 543041. https://doi.org/10.1155/2014/543041 doi: 10.1155/2014/543041
![]() |
[6] |
Q. B. Gao, N. Olgac, Bounds of imaginary spectra of LTI systems in the domain of two of the multiple time delays, Automatica, 72 (2016), 235–241. https://doi.org/10.1016/j.automatica.2016.05.011 doi: 10.1016/j.automatica.2016.05.011
![]() |
[7] |
J. Z. Cai, Q. B. Gao, Y. F. Liu, A. G. Wu, Generalized dixon resultant for strong delay-independent stability of linear systems with multiple delays, IEEE Trans. Autom. Control, 1 (2023), 1–8. https://doi.org/10.1109/TAC.2023.3337691 doi: 10.1109/TAC.2023.3337691
![]() |
[8] |
P. Yang, Hopf bifurcation of an age-structured prey Cpredator model with Holling type Ⅱ functional response incorporating a prey refuge, Nonlinear Anal. Real World Appl., 49 (2019), 368–385. https://doi.org/10.1016/j.nonrwa.2019.03.014 doi: 10.1016/j.nonrwa.2019.03.014
![]() |
[9] |
D. X. Yan, Y. Cao, Y. Yuan, Stability and Hopf bifurcation analysis of a delayed predator–prey model with age-structure and Holling Ⅲ functional response, Z. Angew. Math. Phys., 74 (2023), 148–172. https://doi.org/10.1007/s00033-023-02036-3 doi: 10.1007/s00033-023-02036-3
![]() |
[10] |
D. X. Yan, Y. Yuan, X. L. Fu, Asymptotic analysis of an age-structured predator-prey model with ratio-dependent Holling Ⅲ functional response and delays, Evol. Equations Control Theory, 12 (2023), 391–414. https://doi.org/10.3934/eect.2022034 doi: 10.3934/eect.2022034
![]() |
[11] |
G. Zhu, J. J. Wei, Global stability and bifurcation analysis of a delayed predator-prey system with prey immigration, Electron. J. Qual. Theory Differ. Equations, 13 (2016), 1–20. https://doi.org/10.14232/ejqtde.2016.1.13 doi: 10.14232/ejqtde.2016.1.13
![]() |
[12] |
L. J. Wang, C. J. Dai, M. Zhao, Hopf bifurcation in an age-structured prey-predator model with Holling Ⅲ response function, Math. Biosci. Eng., 18 (2021), 3144–3159. https://doi.org/10.3934/mbe.2021156 doi: 10.3934/mbe.2021156
![]() |
[13] |
X. M. Zhang, Z. H. Liu, Periodic oscillations in age-structured ratio-dependent predator–prey model with Michaelis–Menten type functional response, Physica D, 389 (2019), 51–63. https://doi.org/10.1016/j.physd.2018.10.002 doi: 10.1016/j.physd.2018.10.002
![]() |
[14] |
N. N. Li, W. X. Sun, S. Q. Liu, A stage-structured predator-prey model with Crowley-Martin functional response, Discrete Contin. Dyn. Syst. - Ser. B, 28 (2023), 2463–2489. https://doi.org/10.3934/dcdsb.2022177 doi: 10.3934/dcdsb.2022177
![]() |
[15] |
B. T. Mulugeta, L. P. Yu, Q. G. Yuan, J. L. Ren. Bifurcation analysis of a predator-prey model with strong Allee effect and Beddington-DeAngelis functional response, Discrete Contin. Dyn. Syst. - Ser. B, 28 (2023), 1938–1963. https://doi.org/10.3934/dcdsb.2022153 doi: 10.3934/dcdsb.2022153
![]() |
[16] |
C. A. Ibarra, J. Flores, Dynamics of a Leslie-Gower predator-prey model with Holling type Ⅱ functional response, Allee effect and a generalist predator, Math. Comput. Simul., 188 (2021), 1–22. https://doi.org/10.1016/j.matcom.2021.03.035 doi: 10.1016/j.matcom.2021.03.035
![]() |
[17] |
H. Y. Wang, S. J. Guo, S. Z. Li, Stationary solutions of advective Lotka–Volterra models with a weak Allee effect and large diffusion, Nonlinear Anal. Real World Appl., 56 (2020), 103171. https://doi.org/10.1016/j.nonrwa.2020.103171 doi: 10.1016/j.nonrwa.2020.103171
![]() |
[18] |
M. H. Wang, M. Kot, Speeds of invasion in a model with strong or weak Allee effects, Math. Biosci., 171 (2001), 83–97. https://doi.org/10.1016/S0025-5564(01)00048-7 doi: 10.1016/S0025-5564(01)00048-7
![]() |
[19] | P. Magal, S. G. Ruan, Theory and Applications of Abstract Semilinear Cauchy Problems, Springer Cham, Switzerland, 2018. https://doi.org/10.1007/978-3-030-01506-0 |
[20] | A. Pazy, Semigroups of Linear Operators and Applications to Partial Differential Equations, Springer-Verlag, New York, 1983. https://doi.org/10.1007/978-1-4612-5561-1 |
[21] | P. Magal, Compact attractors for time-periodic age-structured population models, Electron. J. Differ. Equations, 65 (2001), 1–35. |
[22] | P. Magal, S. G. Ruan, Infinite Dimensional Dynamical Systems, Springer, New York, 2013. https://doi.org/10.1007/978-1-4614-4523-4 |
[23] |
H. R. Thieme, Convergence results and a Poincare-Bendixson trichotomy for asymptotically autonomous differential equations, J. Math. Biol., 30 (1992), 755–763. https://doi.org/10.1007/BF00173267 doi: 10.1007/BF00173267
![]() |
[24] |
Z. H. Liu, P. Magal, S. G. Ruan, Hopf bifurcation for non-densely defined Cauchy problems, Z. Angew. Math. Phys., 62 (2011), 191–222. https://doi.org/10.1007/s00033-010-0088-x doi: 10.1007/s00033-010-0088-x
![]() |