
Citation: Chang-Yuan Cheng, Shyan-Shiou Chen, Xingfu Zou. On an age structured population model with density-dependent dispersals between two patches[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 4976-4998. doi: 10.3934/mbe.2019251
[1] | Marco Tosato, Xue Zhang, Jianhong Wu . A patchy model for tick population dynamics with patch-specific developmental delays. Mathematical Biosciences and Engineering, 2022, 19(5): 5329-5360. doi: 10.3934/mbe.2022250 |
[2] | James T. Cronin, Jerome Goddard II, Amila Muthunayake, Ratnasingham Shivaji . Modeling the effects of trait-mediated dispersal on coexistence of mutualists. Mathematical Biosciences and Engineering, 2020, 17(6): 7838-7861. doi: 10.3934/mbe.2020399 |
[3] | Jacques A. L. Silva, Flávia T. Giordani . Density-dependent dispersal in multiple species metapopulations. Mathematical Biosciences and Engineering, 2008, 5(4): 843-857. doi: 10.3934/mbe.2008.5.843 |
[4] | Nancy Azer, P. van den Driessche . Competition and Dispersal Delays in Patchy Environments. Mathematical Biosciences and Engineering, 2006, 3(2): 283-296. doi: 10.3934/mbe.2006.3.283 |
[5] | Yue Xia, Lijuan Chen, Vaibhava Srivastava, Rana D. Parshad . Stability and bifurcation analysis of a two-patch model with the Allee effect and dispersal. Mathematical Biosciences and Engineering, 2023, 20(11): 19781-19807. doi: 10.3934/mbe.2023876 |
[6] | Ali Mai, Guowei Sun, Lin Wang . The impacts of dispersal on the competition outcome of multi-patch competition models. Mathematical Biosciences and Engineering, 2019, 16(4): 2697-2716. doi: 10.3934/mbe.2019134 |
[7] | Tyler Cassidy, Morgan Craig, Antony R. Humphries . Equivalences between age structured models and state dependent distributed delay differential equations. Mathematical Biosciences and Engineering, 2019, 16(5): 5419-5450. doi: 10.3934/mbe.2019270 |
[8] | Yun Kang, Sourav Kumar Sasmal, Komi Messan . A two-patch prey-predator model with predator dispersal driven by the predation strength. Mathematical Biosciences and Engineering, 2017, 14(4): 843-880. doi: 10.3934/mbe.2017046 |
[9] | James T. Cronin, Nalin Fonseka, Jerome Goddard II, Jackson Leonard, Ratnasingham Shivaji . Modeling the effects of density dependent emigration, weak Allee effects, and matrix hostility on patch-level population persistence. Mathematical Biosciences and Engineering, 2020, 17(2): 1718-1742. doi: 10.3934/mbe.2020090 |
[10] | Joan Ponce, Horst R. Thieme . Can infectious diseases eradicate host species? The effect of infection-age structure. Mathematical Biosciences and Engineering, 2023, 20(10): 18717-18760. doi: 10.3934/mbe.2023830 |
Among the various features for population dynamics of a single species are the heterogeneity of different habitats for the species and the age structure of species. Due to some natural barriers such as rivers and mountains, man-made barriers such as buildings and highways, and the local habitation of insects and small animals, a patchy environment is often considered in population models. Random migrations betweens patches are often assumed which adopts constant dispersal rates between patches. However, for many species, their intention of dispersal is heavily influenced by the population distribution among the patches and the variance of resources in all patches. This is because individuals typically seek habitats with abundant sources and mates and flee from the threat of predators [1]. On the other hand, most species have a clear and well-defined age-structure consisting of immature (juvenile) and mature (adult) stages, and behave differently at different stages. Therefore, it is vital to explore the effects of density dependent dispersal rates on population dynamics that involve both heterogeneity among habitats and the age structure.
In [2], Levin proposed a two-patch model to explore the spatial heterogeneity of a single species population. In [3], the authors systematically developed age-structured population models to describe the evolution of populations over multiple life stages. Combining the two aspects in [2] and [3] and assuming constant dispersal rates between patches, So et. al. [4] derived a system of delayed differential equations (DDEs) to describe the adult populations of a single species living on two patches; and by analyzing the derived model, they were able to illustrate some effect of the immature death rate on global dynamics. In subsequent studies in [5,6,7] on the model assuming identical patches, rich dynamics including Hopf bifurcation, synchronized periodic oscillations and unstable phase-locked oscillations were obtained by applying local bifurcation theory to the DDE model; and global convergence to the trivial state or a positive homogeneous state was established under certain conditions in [6]. A recent study [8] examined the model with two general patches and three typical birth functions (linear, Ricker and Allee functions). For these types of birth function and in two patches, the author investigated the extinction and persistence of the species under consideration, the boundedness of the solutions to the model, and certain attractors of the model system.
In reality, the mechanisms of dispersion/migration of many species can be very complicated, and constant dispersal rate would be too simple and too ideal. In [1] the authors introduced an adaption mechanism of dispersion in a predator-prey model by assuming that according dispersal rates are dependent of the population densities; more specifically, they assumed that a higher density of prey population in a patch results in a lower migration rate of predators in the same patch, and a prey tends to flee from the regions with a high density of predators. They incorporated adaptive migration with multiple timescales, and numerically observed co-dimension two Bautin bifurcation. Recently, the authors in [9,10,11] constructed some mathematical models to examine the adaptive dispersal rates of predator-prey populations and compared the adaptive dispersal rate of competitive species between their habitats. The authors in [9] proposed a patch payoff concept to describe adaptive dispersal; this concept entails assuming that the net movement between patches is always toward the patch with the higher per capita growth rate. In [10,11], the authors described the fitness of individuals according to the local growth rate and assumed that each individual tends to move to a patch with higher fitness. By analyzing the local stability of equilibria, they obtained some results that can hardly occur in patch models with constant dispersal rates, such as an adaption-induced change in persistence patterns [10,11], oscillations and Neimark-Sacker bifurcation [9]. The terms of adaptive dispersal typically result in complex nonlinearity, and hence, the global dynamics remains an open and challenging problem in the above mentioned three studies.
Because of the limitation in resources, biological species tend to leave a crowded habitat. In this study, we incorporate a density-dependent dispersal strategy into the prototypical age-structured model proposed in [3]. Motivated by and following the derivation of the model in [4], we derive a DDE model for the adult population in the case of density dependent dispersal. In contrast to the model in [4], there are two additional difficulties in theoretically studying this new DDE model: the nonlinear instantaneous dispersal terms and the nonlinear delayed birth terms mediated by the nonlinear dispersals. Allowing a non-monotone birth function such as the Ricker or Allee functions is another common challenge in population models.
Global dynamics of a DDE system is generally very difficult to investigate, even in a scalar DDE (see, e.g., [12,13,14]). The authors in [12,13] used a domain decomposition method to study mono-stability and bi-stability in scaler DDEs respectively, and obtained some nice results, including a Poincaré-Bendixson type result, heteroclinic orbits, and basins of attraction. This method generally fails for systems of more than one DDE, and hence, the convergent dynamics in a two-variable DDE system is still a hard problem, especially when it involves a non-monotone feedback. Among the known attempts, Xu [6] investigated a population model in two identical patches and attacked this problem by seeking an attractor for the system and applying the monotone dynamical system theory; and for a species living in a general patchy environment with heterogeneity between two patches, Terry [15] examined the population dynamics of the model with impulsive culling of the adults by the comparison with certain linear DDE system.
The main goal of this study is to explore the impact of density-dependent migration on the dynamics of an age-structured population model in two heterogeneous patches. In Section 2, we derive our model which turns out to be a system of DDEs with nonlinearity in the dispersion terms and spatial non-locality in the delayed nonlinear birth terms. In Section 3, we show well-posedness of the proposed model, and focus on the case with only adult dispersal. In Section 4, by analyzing this model, we obtain some criteria for the existence of positive equilibrium and determine the uniform persistence of populations in the two patches. Making use of these results, we discuss how the density-dependent dispersal affects persistence or extinction of populations. In addition, we investigate the global convergence of the model by dividing the phase space, identifying an attractor for the model system, and applying the theory of monotone dynamical systems. We also identify a positive invariant set within which possible sustained oscillations may be induced by the maturation delay. We conclude the paper by Section 5, in which we present some numerical simulations and offer some discussion of our results.
We follow the procedure in [4] to derive our model with density dependent dispersals. Let ui(t,a) denote the density of a single species at time t (t≥0) of age a (a≥0) on patch i (i=1,2). In the case of density-independent dispersal rates between the patches as considered in [4], by the basic equation for age structured populations in [3], there hold
∂u1(t,a)∂t+∂u1(t,a)∂a=−d1(a)u1(t,a)+D2(a)u2(t,a)−D1(a)u1(t,a)∂u2(t,a)∂t+∂u2(t,a)∂a=−d2(a)u2(t,a)+D1(a)u1(t,a)−D2(a)u2(t,a). | (2.1) |
Here di(a) is the death rate of the individuals of age a in patch i, and Dj(a) denotes the dispersal rate of the species of age a from patch j to patch i, for 1≤j≠i≤2. So et. al. [4] considered two age stages: immatures and matures, and assumed that
di(a)={di,I(a)=dI(a),for0≤a≤r,di,M(a)=di,m,fora>r, | (2.2) |
and
Di(a)={Di,I(a)=DI(a),for0≤a≤r,Di,M(a)=Di,m,fora>r, | (2.3) |
where r≥0 denotes the maturation age, I and M stand for immature and mature respectively. Let wi(t) be the total number of adults at time t on patch i, that is,
wi(t)=∫∞rui(t,a)da. |
Noting that ui(t,∞)=0 and that only adults produce and hence
ui(t,0)=bi(wi(t)), | (2.4) |
where bi is a birth function of the population on the i-th patch, So et. al. derived the following system in [4] for the adult populations
dw1(t)dt=−d1,mw1(t)+D2,mw2(t)−D1,mw1(t)+σ[1−∫r0e−∫rθˆD(a)daD1(θ)dθ]b1(w1(t−r))+σ[∫r0e−∫rθˆD(a)daD2(θ)dθ]b2(w2(t−r)),dw2(t)dt=−d2,mw1(t)+D1,mw2(t)−D2,mw1(t)+σ[1−∫r0e−∫rθˆD(a)daD2(θ)dθ]b2(w2(t−r))+σ[∫r0e−∫rθˆD(a)daD1(θ)dθ]b1(w1(t−r)), | (2.5) |
where σ=e−∫r0dI(a)da and ˆD(a)=D1(a)+D2(a).
Now we modify (2.1) by incorporating density dependent dispersion between patches. In general, the dispersal rate of individuals may depend on the population of all ages. For simplicity of mathematics, in this paper we only consider a scenario that the dispersal rate of individuals depend only on the mature population and this leads to the following revision of (2.1):
∂u1(t,a)∂t+∂u1(t,a)∂a=−d1(a)u1(t,a)+˜D2(a,w2(t))u2(t,a)−˜D1(a,w1(t))u1(t,a)∂u2(t,a)∂t+∂u2(t,a)∂a=−d2(a)u2(t,a)+˜D1(a,w1(t))u1(t,a)−˜D2(a,w2(t))u2(t,a), | (2.6) |
where
˜Di(a,wi)=Di(a)fi(wi),i=1,2, | (2.7) |
with the functions fi satisfying
fi∈C1andf′i(w)≥0forw≥0,fi(0)=1,limw→∞fi(w)=1+ηi,i=1,2. | (2.8) |
Such a scenario can be justified as below. Firstly, for many species, each single mature individual reproduces a fixed numbers of offsprings. For examples, a couple of mature shorebirds lay fixed number of eggs in a clutch and then hatch and breed them [16,17]; some particular species of geckos and lizards also lay fixed number of eggs to maximize their reproductive success [18]. For such a species, it is reasonable to assume that the population at age a is proportional to the mature population. Secondly, the density-dependent dispersal is proposed to investigate the strategy of dispersing toward a relatively abundant resource, and many species compete for resources mainly for their mature individuals because the immature individuals consume relatively little or none resource directly but obtain all nutrition from their raisers. With the above two considerations, the dispersal rate given by (2.7) is reasonable for such species.
The value of ηi≥0 in (2.8) represents the strength of adaptive dispersal in patch i. We call Di(a) and ˜Di(a,wi) the intrinsic and adaptive dispersal rates respectively. When η1=η2=0, (2.6) reduces to (2.1) reflecting density-independent dispersal rates. In this work, we consider (2.6) with the assumptions (2.2), (2.3) and the typical function
fi(w)=1+ηiw1+w,forw≥0. |
By integrating (2.6) with respect to a from r to ∞ and making use of the fact ui(t,∞)=0, we obtain
dw1(t)dt=u1(t,r)−d1,mw1(t)+D2,mf2(w2(t))w2(t)−D1,mf1(w1(t))w1(t),dw2(t)dt=u2(t,r)−d2,mw1(t)+D1,mwf1(w1(t))2(t)−D2,mf2(w2(t))w1(t). | (2.9) |
To describe ui(t,r) in terms of the adult populations, we fix s and consider the function
Vsi(t):=ui(t,t−s),fors≤t≤s+randi=1,2. |
From (2.6),
dVsi(t)dt=−di(t−s)Vsi(t)+Dj(t−s)fj(wj(t))Vsj(t)−Di(t−s)fi(wi(t))Vsi(t) | (2.10) |
for s≤t≤s+r and 1≤i≠j≤2. Since di(t−s)=dI(t−s) for i=1,2 and s≤t≤s+r, we have
ddt[Vs1(t)+Vs2(t)]=−dI(t−s)[Vs1(t)+Vs2(t)], |
and then, by (2.4),
Vs1(t)+Vs2(t)=e−∫t−s0dI(a)da[b1(w1(s))+b2(w2(s))]. |
For s≤t≤s+r, we rewrite (2.10) as
dVsi(t)dt=−dI(t−s)Vsi(t)+Dj(t−s)fj(wj(t))[Vsi(t)+Vsj(t)]−[Di(t−s)fi(wi(t))+Dj(t−s)fj(wj(t))]Vsi(t)=−Ks(t−s)Vsi(t)+Dj(t−s)fj(wj(t))e−∫t−s0dI(a)da[bi(wi(s))+bj(wj(s))], | (2.11) |
where Ks(a)=dI(a)+Di(a)fi(wi(a+s))+Dj(a)fj(wj(a+s)). Solving (2.11), we obtain
Vsi(t)=e−∫tsKs(θ−s)dθVsi(s)+∫tse−∫tξKs(θ−s)dθDj(ξ−s)fj(wj(ξ))e−∫ξ−s0dI(a)dadξ[bi(wi(s))+bj(wj(s))]. |
Thus,
ui(t,r)=Vt−ri(t)=e−∫tt−rKt−r(θ−t+r)dθbi(wi(t−r))+∫tt−re−∫tξKt−r(θ−t+r)dθDj(ξ−t+r)fj(wj(ξ))e−∫ξ−t+r0dI(a)dadξ×[bi(wi(t−r))+bj(wj(t−r))]=e−∫r0[dI(a)+Γt−r(a)]dabi(wi(t−r))+e−∫r0dI(a)da×∫r0{e−∫rθΓt−r(a)daDj(θ)fj(wj(θ+t−r))}dθ[bi(wi(t−r))+bj(wj(t−r))], | (2.12) |
where
Γs(a):=D1(a)f1(w1(a+s))+D2(a)f2(w2(a+s)). |
In addition,
∫r0{e−∫rθΓt−r(a)daDj(θ)fj(wj(θ+t−r))}dθ=[e−∫rθΓt−r(a)da]θ=rθ=0−∫r0{e−∫rθΓt−r(a)daDi(θ)fi(wi(θ+t−r))}dθ=1−e−∫r0Γt−r(a)da−∫r0{e−∫rθΓt−r(a)daDi(θ)fi(wi(θ+t−r))}dθ. |
Combining this with (2.12), we have
ui(t,r)=σ{1−∫r0e−∫rθΓt−r(a)daDi(θ)fi(wi(θ+t−r))dθ}bi(wi(t−r))+σ∫r0e−∫rθΓt−r(a)daDj(θ)fj(wj(θ+t−r))dθbj(wi(t−r)), |
with σ=exp(−∫r0dI(a)da) as before. Plugging this into (2.9), we finally obtain the following system
dw1(t)dt=−d1,mw1(t)+D2,mf2(w2(t))w2(t)−D1,mf1(w1(t))w1(t)+Λ1(w1t,w2t),dw2(t)dt=−d2,mw2(t)+D1,mf1(w1(t))w2(t)−D2,mf2(w2(t))w2(t)+Λ2(w1t,w2t), | (2.13) |
where wit∈C([−r,0],R)→R is defined by wit(θ)=wi(t+θ) for θ∈[−r,0], i=1,2, and
Λi(w1t,w2t)=σ[1−∫r0e−∫rθΓt−r(a)daDi(θ)fi(wi(θ+t−r))dθ]bi(wi(t−r))+σ[∫r0e−∫rθΓt−r(a)daDj(θ)fj(wj(θ+t−r))dθ]bj(wj(t−r)),1≤i≠j≤2. | (2.14) |
Here bi(⋅):R+→R+,i=1,2, are still the birth functions which are assumed to be continuously differentiable. Depending on the particular species, they may have different forms. Comparing with (2.5), we note that both instantaneous and delayed dispersal terms have been modified in (2.13) by nonlinear functions fi(wi),i=1,2, representing the density dependent adaptive dispersal strategy.
The model system (2.13) is a system of delay differential equations, which is associated with the phase space C:=C([−r,0],R2). By the fundamental theory of functional differential equations (see, e.g., [19]) or directly by an argument of steps, one knows that with the initial condition
w0=(w10,w20)=ϕ=(ϕ1,ϕ2)∈C, |
(2.13) has a unique solution w(t)=(w1(t),w2(t)) for t>0. Next, we will show that if the initial functions ϕ1 and ϕ2 are nonnegative, that is,
w0=ϕ=(ϕ1,ϕ2)∈C+:={ϕ∈C;ϕ1(θ)≥0,ϕ2(θ)≥0,θ∈[−r,0]}, |
then the unique solution also remains nonnegative; moreover, if the birth function is bounded, then the solution is bounded.
Lemma 3.1. If ϕ∈C+, then the solution of (2.13) remains non-negative for all t≥0.
Proof. By the non-negativity of the functions fi(⋅) and bi(⋅),i=1,2, [Theorem 2.1, P81] in [20] applies, leading to the conclusion of the lemma.
Lemma 3.2. Assume that the birth functions bi(⋅),i=1,2, are bounded and let ϕ∈C+. Then the corresponding solution of (2.13) is bounded.
Proof. Let B=sup{bi(u);u∈R+,i=1,2.} and d_=min{d1,m,d2,m}. Then the total adult population W(t)=w1(t)+w2(t), satisfies
dW(t)dt=−d1,mw1(t)−d2,mw2(t)+σ[b1(w1(t−r))+b2(w1(t−r))]≤−d_W(t)+σB. |
This leads to
lim supt→∞W(t)≤σBd_, |
implying that W(t) is bounded. This together with the non-negativity of w1(t) and w2(t) then further implies the boundedness of w1(t) and w2(t), completing the proof.
In the model system (2.13)- (2.14), the general forms of the dispersion rate functions Di(a) seem to prevent us from going further in analyzing the mode, due to the occurrence of multiple iterated integrals. In the sequel, we only consider the following simple case for the dispersion rate functions:
Di(a)={Di,I(a)=0,a∈[0,r],Di,m(a)=Di,m=a positive constant,a∈(r,∞). | (3.1) |
With these simple dispersion rate functions, the delayed term (2.14) reduces to
Λi(w1t,w2t)=σbi(wi(t−r)). |
Also, the birth functions bi(⋅) have many choices depending on the particular species. In the rest of this paper, we will chose the form of the Ricker's birth function for both bi(⋅),i=1,2:
bi(w)=βiwe−γiw. | (3.2) |
For the above choices for the birth and dispersion rate functions, the model (2.13) becomes
dw1(t)dt=−d1,mw1(t)+D2,mf2(w2(t))w2(t)−D1,mf1(w1(t))w1(t)+σβ1w1(t−r)e−γ1w1(t−r)dw2(t)dt=−d2,mw2(t)+D1,mf1(w1(t))w1(t)−D2,mf2(w2(t))w2(t)+σβ2w2(t−r)e−γ2w2(t−r) | (3.3) |
First, we study the equilibria of the system (3.3) which are governed by the following system of algebraic equations:
L1: D2,mf2(w2)w2−d1,mw1−D1,mf1(w1)w1+σb1(w1)=0,L2: D1,mf1(w1)w1−d2,mw2−D2,mf2(w2)w2+σb2(w2)=0. | (4.1) |
For w≥0 and i=1,2, denote
Gi(w)=Di,mfi(w)w,Fi(w)=di,mw+Di,mfi(w)w−σbi(w). |
Then (4.1) is equivalent to Gj(wj)=Fi(wi) for 1≤i≠j≤2. Note that Gi(0)=0 and limw→∞Gi(w)=∞ for i=1,2. Obviously, each function Gi has an increasing inverse function G−1i on [0,∞) satisfying G−1(0)=0 and limw→∞G−1i(w)=∞. Hence, the curve Li in the first quadrant is described by
wj=G−1j(Fi(wi)):=Hi(wi), |
for wi≥0 and Fi(wi)>0. According to the feature of functions Fi(w), we obtain the following properties of the curves Li.
Lemma 4.1. The following hold:
(i) if
σβi≤di,m+Di,m, |
then the function Hi is increasing on [0,∞) and only has the trivial root w=0;
(ii) if
di,m+Di,m<σβi, |
then in addition to the trivial root, Hi also has a positive root wri>0 and is strictly increasing on [wri,∞).
Proof. First, we explore features of the functions Fi. Write it as Fi(w)=w[hi(w)−ki(w)] where hi(w):=di,m+Di,mfi(w) is a nondecreasing and ki(w):=σbi(w)/w=σβie−γiw is a nonincreasing. Thus, if σβi≤di,m+Di,m, then hi(0)≥ki(0) and Fi(w)=0 has no positive root and Fi(w) is increasing on [0,∞); and if di,m+Di,m<σβi, then in addition to the trivial root, Fi(w)=0 also has an unique positive root wri>0 and Fi(w) is increasing and positive on [wri,∞) and negative on [0,wri). Now combining these features of Fi(w) with the properties of G−1j(⋅) immediately leads to the assertions on Hi, completing the proof.
Lemma 4.2. The graph of Li in the first quadrant admits an asymptote wj=piwi+qi, 1≤i≠j≤2, for some constant qi and
pi:=di,m+Di,m(ηi+1)Dj,m(ηj+1) |
Proof. In the first quadrant, a point (wi,wj) on the graph of Li obeys
wjwi=di,m+Di,mfi(wi)−σbi(wi)/wiDj,mfj(wj)=di,m+Di,mfi(wi)−σβe−γiwiDj,mfj(wj). | (4.2) |
Note that limwi→∞Fi(w)=∞ and limw→∞Gj(w)=∞. In addition, a point (wi,wj) on the curve Li satisfies wj→∞ as wi→∞. The assertion follows from the above and the form of fi(w), completing the proof.
Based on previous results, we discuss existence of positive equilibria in the following.
Theorem 4.1. The system (3.3) always has the trivial equilibrium E0=(0,0). In addition, it admits an unique positive equilibrium when one of the following conditions holds:
(A) σβi≤di,m+Di,m for i=1,2, and Πk=1,2(dk,m+Dk,m−σβk)<D1,mD2,m;
(B) σβi>di,m+Di,m for i=1 or 2.
Proof. E0=(0,0) is obviously always an equilibrium of (3.3). An non-trivial equilibrium is governed by
wj=Hi(wi),for1≤i≠j≤2, | (4.3) |
for wi>0 with Hi(wi)>0. Note that each G−1i is increasing on [0,∞). As explored in Lemma 4.1, Fi is also increasing on [0,∞) under (A) (respectively on [wri,∞) under (B)), when Fi(w)=0 has one (respectively, two) nonnegative root. Therefore, the function Hi is increasing on [0,∞) under (A) (respectively on [wri,∞) under (B)). Next, we discuss the existence of roots to (4.3).
First, we consider the system (3.3) under the criterion(A). From Lemma 4.2, it always holds that
H′1(∞)=d1,m+D1,m(η1+1)D2,m(η2+1)>D1,m(η1+1)d2,m+D2,m(η2+1)=1H′2(∞)=(H−12)′(∞), | (4.4) |
where H′i(∞):=limw→∞H′i(w). The condition σβi<di,m+Di,m for i=1,2 implies that each Fi(w)=0 only has the trivial root and Fi(w)>0 for all w>0, and then the function G−1i is defined and increasing on [0,∞). Therefore, each Hi is defined, continuous and increasing on [0,∞). The further condition in (A),
Πk=1,2(dk,m+Dk,m−σβk)<D1,mD2,m |
is equivalent to
H′1(0)<(H−12)′(0). |
Since H1(0)=H−12(0) and H′1(0)<(H−12)′(0), there is a small ˇw>0 such that H1(ˇw)<H−12(ˇw). From H′1(∞)>(H−12)′(∞), there is a large enough ˆw such that H1(ˆw)>H−12(ˆw). Since functions H1 and H−12 are continuous on [0,∞), there exists at least one point (w∗1,w∗2) in the first quadrant such that H1(w∗1)=H−12(w∗1)=w∗2, giving a positive root to (4.3). From (4.3) and the monotonicity of each function Hi, i=1,2, the positive equilibrium is unique when it exists. See Figure 1(a) for a demonstration of this case.
Assume that the criterion(B) holds only for i=1 then F1(w)=0 has two nonnegative roots, 0, wr1, and F1(w)>0 for all w>wr1. Hence, the function H1 is defined, continuous and increasing on [wr1,∞). Note that H1(wr1)=0≤(H2)−1(0)<(H2)−1(wr1) and the fact (4.4) implies a large enough ˆw such that H1(ˆw)>H−12(ˆw). Similar to the case with (A), but considering [wr1,∞) as the domain of H1, there exists a unique positive root to (4.3). See Figure 1(b) for an illustration of this case. Similarly, the assertion holds when (B) holds only for i=2 or for both i=1,2. The proof is completed.
Theorem 4.2. Consider the system (3.3) and assume that neither(A) nor (B) holds. Then the population in (3.3) goes to extinction in both patches, that is, every solution (w1(t),w2(t)) of (3.3) with ϕ∈C+ satisfies limt→∞wi(t)=0,fori=1,2.
Proof. From (3.3), it holds that
dw1(t)dt≤−d1,mw1(t)+D2,mf2(w2(t))w2(t)−D1,mf1(w1(t))w1(t)+σ˜b1(w1(t−r)),dw2(t)dt≤−d2,mw2(t)+D1,mf1(w1(t))w1(t)−D2,mf2(w2(t))w2(t)+σ˜b2(w2(t−r)), |
where
˜bi(w)={bi(w),ifw∈[0,1/γi],bi(1/γi),ifw∈(1/γi,+∞). |
The auxiliary system
dx1(t)dt=−d1,mx1(t)+D2,mf2(x2(t))x2(t)−D1,mf1(x1(t))x1(t)+σ˜b1(x1(t−r)),dx2(t)dt=−d2,mx2(t)+D1,mf1(x1(t))x1(t)−D2,mf2(x2(t))x2(t)+σ˜b2(x2(t−r)), | (4.5) |
is cooperative and irreducible [20] and admits the unique trivial equilibrium when neither (A) nor (B) holds. Similar to the proof of Lemma 3.2, the solutions of (4.5) with initial conditions in C+ are uniformly bounded. From [20,Theorem 5.4.1], each solution of (4.5) converges to the unique trivial equilibrium, i.e. limt→∞xi(t)=0,fori=1,2. From the comparison principle [20,Theorem 5.1.1], wi(t)≤xi(t) for t≥0, i=1,2, and then it completes the proof.
Next, we consider permanence of the populations described by (3.3).
Theorem 4.3. Consider the system (3.3) and assume that either (A) or (B) holds. Then (3.3) is uniformly persistent in the sense that there is a positive constant ρ∗ such that every solution (w1(t),w2(t)) of (3.3) with ϕ∈C+∖{ˆ0}, where ˆk denotes the constant function on [−τ,0] taking value k, satisfies
lim inft→∞wi(t)≥ρ∗,fori=1,2. |
Proof. First, we suppose that (A) holds. Define
X=C([−r,0],R2+),X0={ϕ=(ϕ1,ϕ2)∈Xwithϕi≠ˆ0fori=1,2},∂X0=X∖X0={ϕ∈X:eitherϕ1=ˆ0orϕ2=ˆ0} |
Obviously, both X and X0 are positive invariant under the semiflow of (3.3) and X∖X0 is relatively close in X. From Lemma 3.2, system (3.3) is point dissipative. Set
M∂={ϕ∈X|Φ(t)ϕ∈∂X0,∀t≥0}, |
where Φ(t) is the semiflow generated by (3.3). Then obviously M∂⊂∂X0. We claim that M∂={ˆ0}. It is clear that {ˆ0}⊂M∂, so it suffice to show M∂⊂{ˆ0}. Assume the opposite, there is ϕ=(ϕ1,ϕ2)∈M∂ with ϕ≠ˆ0. Without loss of generality, we suppose ϕ1≠ˆ0, ϕ2=ˆ0. (ⅰ) If ϕ1(0)>0, we have dw2(0)dt>0 and then there is a small t0>0 such that w2(t)>0 for t∈(0,t0). Since ϕ1(0)>0, there is a t1≤t0 such that w1(t)>0 for t∈(0,t1). Thus, Φ(t)ϕ∈X0 for t∈(0,t1), a contradiction to the fact that ϕ∈M∂. (ⅱ) If ϕ1(0)=0 and ϕ1(−r)>0, we see that
dw1(0)dt=σb1(ϕ1(−r))>0. |
From continuity of solutions to (3.3) and ϕ∈C, there is a t2>0 such that
dw1(t)dt>σb1(ϕ1(−r))/2=:K1>0, |
for t∈[0,t2]. Then w1(t)>K1t, for t∈[0,t2]. In addition,
dw2(t)dt≥−d2,mw2(t)−D2,mf2(w2(t))w2(t)+D1,mK1t≥−K2w2(t)+D1,mK1t, |
where
K2:=maxt∈[0,t2]{d2,m+D2,mf2(w2(t))}, |
is finite. By a comparison theorem,
w2(t)≥e−K2tw2(0)+∫t0e−K2tD1,mK1sds>0, |
for t∈(0,t2], also a contradiction to ϕ∈M∂. (ⅲ) If ϕ1(0)=0 and ϕ1(−r)=0, we set r∗=sup{−θ|θ∈[−r,0],ϕ1(θ)≠0}<r. Then dw1(t)dt=0 and w1(t)=0 for t∈[0,r−r∗]. From the assumption ϕ2=ˆ0, it also holds that w2(t)=0 for t∈[0,r−r∗]. Since ϕ∈C, there is a small ϵ1>0 such that ϕ1(−r∗+ϵ1)>0. Define
ψ(θ)={ϕ(θ+r−r∗+ϵ1),ifθ∈[−r,r∗−ϵ1−r),w(θ+r−r∗+ϵ1),ifθ∈[r∗−ϵ1−r,0], |
then ψ2(θ)≥ˆ0 for θ∈[−r,0] and
ψ1(0)=w1(r−r∗+ϵ1)≥0,ψ1(−r)=ϕ1(−r∗+ϵ1)>0. |
By the comparison principle (for ψ2(θ)≥˜ψ2(θ):=ˆ0) and previous result (for ϕ1=ψ1 and ϕ2=˜ψ2), it yields that ψ∉M∂. Note that Φ(t)ψ=Φ(t+r−r∗+ϵ1)ϕ. By the positive invariance of the set M∂, it leads to ϕ∉M∂, a contradiction again. From the contradictions in all three cases (ⅰ)-(ⅲ), we conclude that M∂⊂{ˆ0}, and hence M∂={ˆ0} indeed, proving the claim.
From (A), there exist constants 0<ϵ2<1 and small ρ1>0 such that for 0<ρ<ρ1
ϵ2σβi<di,m+Di,mfi(ρ),fori=1,2,Πk=1,2(dk,m+Dk,mfk(ρ)−ϵ2σβk)<D1,mD2,m. | (4.6) |
For this ϵ2, there is a small 0<ρ∗<ρ1 such that
bi(ξ)≥ϵ2βiξ,forξ∈[0,ρ∗]. | (4.7) |
Now, we claim that
lim supt→∞maxi{wi(t)}>ρ∗,forallϕ∈X0. | (4.8) |
Suppose, for the sake of contradiction, that there exist an initial condition ϕ∈X0 and a t3>0 such that |wi(t)|≤ρ∗, i=1,2, for t≥t3−r. From (4.7), for t≥t3,
dwi(t)dt≥−di,mwi(t)+Dj,mwj(t)−Di,mfi(ρ∗)wi(t)+ϵ2σβiwi(t−r). |
We consider an auxiliary equation
dxi(t)dt=−di,mxi(t)+Dj,mxj(t)−Di,mfi(ρ∗)xi(t)+ϵ2σβixi(t−r), | (4.9) |
and its associated ordinary differential equation
dyi(t)dt=−di,myi(t)+Dj,myj(t)−Di,mfi(ρ∗)yi(t)+ϵ2σβiyi(t). | (4.10) |
From the comparison theory in [20,Theorem 5.5.1],
˜Φ(t)ϕ≤Φ(t)ϕ,fort≥0, |
where ˜Φ(t) is the semiflow of (4.9). In the ordinary differential equation (4.10), the characteristic equation of the related Jacobian matrix at the trivial equilibrium is
λ2+[d1,m+d2,m+D1,mf1(ρ∗)+D2,mf2(ρ∗)−ϵ2σβ1−ϵ2σβ2]λ+Πk=1,2(dk,m+Dk,mfk(ρ∗)−ϵ2σβk)−D1,mD2,m=0. |
From (4.6), it admits two roots, λ1<0<λ2, i.e. the stability modulus of (4.10) is positive. Also note that (4.9) is a cooperative irreducible system. From Theorem 5.5.1 and Corollary 5.5.2 in [20], the system (4.9) also admits a positive stability modulus associated with a positive eigenvector u. Note that the semiflow of (3.3) is eventually strong monotone in [ˆ0,ˆρ∗], see [20,Corollary 5.3.5], and ˆ0 is an equilibrium therein. There exist a t4>t3 and a small α>0 such that
ˆ0≪αˆu≪Φ(t4)ϕ. |
Hence, we have
˜Φ(t)αˆu≤Φ(t)αˆu≪Φ(t+t4)ϕ, |
for t≥0, which is a contradiction to boundedness of the semiflow Φ(t), and this contradiction proves (4.8).
Obviously, {ˆ0} is an isolated invariant set in X, and then the set M∂ consists with an acyclic equilibrium point. From (4.8), Ws(ˆ0)∩X0=∅, where Ws(ˆ0) denotes the stable manifold of ˆ0. By the persistence theory in [21,Theorem 4.6], the system (3.3) is uniformly persistent with respect to (X0,∂X0) and the assertion is proved.
Next, we suppose that (B) holds, i.e. σβi>di,m+Di,m for some i. Then there is a positive constant ρ2 such that
σβi>di,m+Di,mfi(ρ2) | (4.11) |
and
dwi(t)dt≥−di,mwi(t)−Di,mfi(ρ2)wi(t)+σbi(wi(t−r)). |
for 0≤wi(t)<ρ2. For this i, consider an auxiliary equation
dz(t)dt=−di,mz(t)−Di,mfi(ρ2)z(t)+σbi(z(t−r)). |
From (4.11), it obviously admits an unique positive equilibrium z∗. According to z∗≤1/γi or z∗>1/γi, we refer to Proposition 3.2 or Theorem 3.5 in [13] respectively to derive the existence of positive constants ρ3<ρ2 and t5 such that z(t)>ρ3 whenever t>t5. The comparison principle in [20,Theorem 5.1.1] implies that wi(t)≥z(t)≥ρ3 whenever 0≤wi(t)≤ρ2. Thus lim inft→∞wi(t)≥ρ3=:ρ∗i. If, in addition, σβj>dj,m+Dj,m for j≠i, then we have also lim inft→∞wj(t)≥ρ∗j and the assertion holds for ρ∗=min{ρ∗1,ρ∗2}. If σβj≤dj,m+Dj,m for j≠i, then from the equation of wj in (3.3),
dwj(t)dt≥−dj,mwj(t)+Di,mfi(ρ∗i)ρ∗i−Dj,mfj(wj(t))wj(t), |
for t>t5. Let
˜Fj(w)=dj,mw+Dj,mfj(w)w, | (4.12) |
which is obviously strictly increasing in w∈[0,∞) and satisfies ˜Fj(w)=0 and ˜Fj(∞)=∞. Thus, there is a unique ˜ρ∗j>0 such that ˜Fj(˜ρ∗j)=Di,mfi(ρ∗i)ρ∗i. Hence, for t>t5,
dwj(t)dt>−˜Fj(˜ρ∗j)+Di,mfi(ρ∗i)ρ∗i=0, |
whenever 0≤wj(t)<˜ρ∗j. Therefore, lim inft→∞wj(t)≥˜ρ∗j, and the assertion holds for ρ∗:=min{ρ∗i,˜ρ∗j}.
To demonstrate the convergent dynamics in system (3.3) that involves the non-monotone feedback, we identify an compact set that is invariant and attracting for (3.3) and in which, (3.3) is monotone dynamics and hence the monotone dynamical system theory is applicable therein. For convenience of formulation, we denote ˜wi=1γi for i=1,2, ˜w=(˜w1,˜w2) and an order interval
I:=[ˆ0,ˆ˜w]={ϕ=(ϕ1,ϕ2)|ϕi∈C([−r,0],[0,˜wi])}. |
Theorem 4.4. Consider the system (3.3). Suppose that either (A) or (B) holds. Suppose that (w∗1,w∗2)≤(˜w1,˜w2) and
σb1(˜w1)≤˜F1(˜w1)−G2(˜w2),σb2(˜w2)≤˜F2(˜w2)−G1(˜w1), | (4.13) |
where ˜Fi is defined in (4.12). Then the positive equilibrium, (w∗1,w∗2) attracts all solutions in X, i.e. limt→∞wi(t)=w∗i for i=1,2.
Proof. Denote, for l≥˜w1+˜w2,
Ωl:={ϕ=(ϕ1,ϕ2)|ϕi∈C([−r,0],[0,l−˜wj]),j≠i,ϕ1(θ)+ϕ2(θ)≤lforθ∈[−r,0]}, |
see Figure 2. We claim that each Ωl is positive invariant under (3.3). Except for those on the axes, there are three segments of the boundary of Ωl,
Bl1={(w1,w2)|w1=l−˜w2,0≤w2≤˜w2},Bl2={(w1,w2)|0≤w1≤˜w1,w2=l−˜w1},Bl3={(w1,w2)|w1+w2=l,wi≥˜wifori=1,2}. |
Denote the right side of (3.3) by J=(J1,J2)T, where T indicates the transpose. If ϕ∈Ωl and ϕ(0)∈Bl1, then we have
J1(ϕ)≤−d1,m(l−˜w2)+D2,mf2(˜w2)˜w2−D1,mf1(l−˜w2)(l−˜w2)+σb1(˜w1)≤−d1,m˜w1+D2,mf2(˜w2)˜w2−D1,mf1(˜w1)˜w1+σb1(˜w1)=−˜F1(˜w1)+G2(˜w2)+σb1(˜w1),≤0, |
where the last inequality follows from (4.13). Similarly, the case that ϕ∈Ωl and ϕ(0)∈Bl2 leads to J2(ϕ)≤0. If ϕ∈Ωl and ϕ(0)∈Bl3, then we have
J1(ϕ)+J2(ϕ)=−d1,mϕ1(0)−d2,mϕ2(0)+σ[b1(ϕ1(−r))+b2(ϕ2(−r))]≤−d1,m˜w1−d2,m˜w2+σ[b1(˜w1)+b2(˜w2)]=−˜F1(˜w1)−˜F2(˜w2)+G(˜w1)+G2(˜w2)+σ[b1(˜w1)+b2(˜w2)]≤0. |
Combining the result in Lemma 3.1 with [20,Remark 5.2.1], Ωl is positive invariant under the semiflow of (3.3).
We write ˜l=˜w1+˜w2 and note that liml→˜lΩl=I. Next, we prove that I attracts each solution of (3.3). It suffices to show that the omega limit set of each solution is contained in I. From boundedness in Lemma 3.2, the solution to (3.3) with initial value ϕ has a nonempty, compact and positively invariant omega limit set ω(ϕ). Define
l0=inf{l|l≥˜w1+˜w2,suchthatω(ϕ)⊂Ωl}. |
Then, there exists a ψ=(ψ1,ψ2)∈ω(ϕ) such that ψ(0)∈Bl01,Bl02 or Bl03. Also note that φi(θ)≤l0−˜wj for 1≤i≠j≤j, and φ1(θ)+φ2(θ)≤l0 for each φ=(φ1,φ2)∈ω(ϕ). From the invariance of ω(ϕ), there exists a ξ=(ξ1,ξ2)∈ω(ϕ) such that wr(ξ)=ψ. Now, suppose that l0>˜l. If ψ(0)∈Bl01, it holds that
dw1(r,ξ)dt=J1(ψ)≤−d1,m(l0−˜w2)+D2,mf2(˜w2)˜w2−D1,mf1(l0−˜w2)(l0−˜w2)+σb1(˜w1)<−d1,m˜w1+D2,mf2(˜w2)˜w2−D1,mf1(˜w1)˜w1+σb1(˜w1)≤0, |
where the strict inequality follows from l0>˜l. It implies a t1<r such that
w1(t1,ξ)>w1(r,ξ)=w1(0,ψ)=l0−˜w2, |
which contradicts to wt1(ξ)∈ω(ϕ)⊂Ωl0. Similarly, the case ψ(0)∈Bl02 also leads to a contradiction. If ψ(0)∈Bl03, it holds that
dw1(r,ξ)dt+dw1(r,ξ)dt=J1(ψ)+J2(ψ)=−d1,mψ1(0)−d2,mψ2(0)+σ[b1(ψ1(−r))+b2(ψ2(−r))]<−d1,m˜w1−d2,m˜w2+σ[b1(˜w1)+b2(˜w2)]≤0. |
It implies a t2<r such that
w1(t2,ξ)+w2(t2,ξ)>w1(r,ξ)+w2(r,ξ)=w1(0,ψ)+w2(0,ψ)=l0, |
which contradicts to wt2(ξ)∈ω(ϕ)⊂Ωl0. Therefore, the omega limit set ω(ϕ)∈Ω˜l=I.
From Theorem 4.3, each omega limit set ω(ϕ) is in fact contained in I∗:=[ˆρ∗,ˆ˜w] for some ρ∗>0. In addition, we see that (w∗1,w∗2) is the unique equilibrium of (3.3) in I∗ and this system is cooperative and irreducible over the space I∗. By [20,Proposition 5.4.2], the unique equilibrium (w∗1,w∗2) attracts all solutions in I∗, and therefore, it attracts all solutions in X.
When (w∗1,w∗2)≤(˜w1,˜w2) does not hold, numerical simulations (see Section 5) show that it is possible for (3.3) to have oscillatory solutions. To conclude this section, we establish some estimates for the dynamics of (3.3) for the case (w∗1,w∗2)>(˜w1,˜w2), and identify an invariant set for (3.3) for this case. To this end, we first have the following lemma.
Lemma 4.3. When (w∗1,w∗2)>(˜w1,˜w2), the system
˜F1(q1)−G2(q2)=σb1(˜w1),˜F1(p1)−G2(p2)=σb1(q1),˜F2(q2)−G1(q1)=σb2(˜w2),˜F2(p2)−G1(p1)=σb2(q2), | (4.14) |
has a unique positive solution for (p1,q1,p2,q2). Moreover, there holds the dichotomy: either pi<w∗i<qi for i=1,2, or pi=qi=w∗i=˜wi for i=1,2.
Proof. Consider the equation
˜F1(x)−G2(y)=σb1(˜w1),˜F2(y)−G1(x)=σb2(˜w2), |
which is equivalent to
˜F1(x)−G2(y)=σb1(˜w1),d1,mx+d2,my=σb1(˜w1)+σb2(˜w2):=M. | (4.15) |
Substituting y=(M−d1,mx)/d2,m into (4.15) leads to the equation
P(x):=˜F1(x)−G2(M−d1,mxd2,m)−σb1(˜w1)=0. |
From the facts that P(0)=−G2(Md2,m)−σb1(˜w1)<0, P(Md1,m)=˜F1(Md1,m)−σb1(˜w1)>0 and the continuous function P is strictly increasing on [0,Md1,m], there is a unique positive root satisfying P(q1)=0, and thus there are unique q1 and q2=(M−d1,mqi)/d2,m satisfying the two left equations in (4.14). A similar argument shows that the two right equations in (4.14) has a unique positive solution (p1,p2).
If w∗i>qi for i=1,2, we derive
σ[b1(w∗1)+b2(w∗2)]=d1,mw∗1+d2,mw∗2>d1,mq1+d2,mq2=σ[b1(˜w1)+b2(˜w2)], |
which is a contradiction. If w∗i>qi and w∗j≤qj for some i and j≠i, then strict monotonicity of functions ˜Fi and Gj implies
[˜Fi(w∗i)−Gj(w∗j)]−[˜Fi(qi)−Gj(qj)]>0. |
It contradicts to the fact from (4.14),
[˜Fi(w∗i)−Gj(w∗j)]−[˜Fi(qi)−Gj(qj)]=σ[bi(w∗i)−bi(˜wi)]≤0. |
Therefore, w∗i≤qi for i=1,2.
To show pi≤w∗i for i=1,2, define the strictly increasing functions
Qi(w):=˜Fi(w)−Gj(σ[b1(q1)+b2(q2)]−di,mwdj,m). |
Then
Qi(w∗i)=˜Fi(w∗i)−Gj(w∗j+σdj,m[b1(q1)+b2(q2)−b1(w∗1)−b2(w∗2)])≥˜Fi(w∗i)−Gj(w∗j)=σbi(w∗i)≥σbi(qi)=Qi(pi). |
The strict monotonicity of Qi implies pi≤w∗i. From previous argument, we see that either pi<w∗i<qi for i=1,2; or pi=qi=w∗i=˜wi for i=1,2.
We now show that the constants pi,qi in (4.14) can help us estimate all omega limit sets and identify a positive invariant set for (3.3).
Theorem 4.5. Assume that either (A) or (B) holds and (w∗1,w∗2)>(˜w1,˜w2). Let p=(p1,p2) and q=(q1,q2) be as in Lemma 4.3. Then for each ϕ∈C+, ω(ϕ)≤ˆq; moreover, if bi(qi)≤bi(pi) for i=1,2, then the set [ˆp,ˆq] is positively invariant under the semiflow of (3.3).
Proof. When wi(t)>qi and wj(t)≤qj for 1≤i≠j≤2,
dwi(t)dt=−˜Fi(wi(t))+Gj(wj(t))+σbi(wi(t−r))<−˜Fi(qi)+Gj(qj)+σbi(˜wi)=0. |
When wi(t)>qi for i=1,2,
dw1(t)dt+dw2(t)dt=−d1,mw1(t)−d2,mw2(t)+σb1(w1(t−r))+σb2(w2(t−r))<−d1,mq1−d2,mq2+σb1(˜w1)+σb2(˜w2)=0. |
Therefore, it holds that ω(ϕ)≤ˆq for all ϕ∈C.
Suppose that bi(qi)≤bi(pi) for i=1,2, let ϕ∈[ˆp,ˆq] and w(t)=(w1(t),w2(t)) be the corresponding solution with initial function ϕ. If there is a t1>0 such that wi(t1)=pi, then
dwi(t1)dt=−˜Fi(wi(t1))+Gj(wj(t1))+σbi(wi(t1−r))≥−˜Fi(pi)+Gj(pj)+σbi(qi)=0. |
If there is a t2>0 such that wi(t2)=qi, then
dwi(t2)dt=−˜Fi(wi(t2))+Gj(wj(t2))+σbi(wi(t2−r))≤−˜Fi(qi)+Gj(qj)+σbi(˜wi)=0. |
By [20,Theorem 5.2.1 and Remark 5.2.1], [ˆp,ˆq] is positively invariant under the semiflow of (3.3).
In this study, we incorporate adaptive dispersal, in the form of density dependent dispersal, into a two-patch population model with a maturation delay derived in [4] which assumes constant dispersion rates between patches, leading to a more realistic model. The improved model turns out to be a system of delay differential equation with spatial non-local birth terms resulted from the dispersals of the immature individual. The density dependent dispersals not only affect the instantaneous migrations of the mature population, they also have an impact on the nonlocal birth terms. For mathematical tractability, we have analyzed the case when the immatures only have constant dispersal rates, adopting the Ricker's birth function and a particular form for the density dependent dispersal rate functions. We have addressed the well-posedness of the model, structure of equilibria, threshold dynamics (in the sense of extinction and persistence) in terms of the conditions (A) and (B) (see Theorems 4.1-4.3). When the population is persistent, under a condition accounting for a monotone scenario (i.e., (w∗1,w∗2)<(˜w1,˜w2)), we have shown that the positive equilibrium is globally attractive; while when this condition is not satisfied (i.e., (w∗1,w∗2)>(˜w1,˜w2)), we have established an upper bound for omega limit sets of solutions and identified an invariant set within which, periodic oscillations may occur.
We remark that surprisingly the threshold conditions A and B for extinction/persistence do not depend on the ηi's. However, ηi's that reflect the level of density dependent dispersals do affect the value of the positive equilibrium (w∗1,w∗2), and accordingly, affect whether or not the condition (w∗1,w∗2)≤(˜w1,˜w2) holds. Thus, ηi's do have an impact on the long time dynamics of the solutions in the case of persistence. Below we numerically explore such an impact. To this end, we fix d1,m=0.4,D1,m=0.6,d2,m=0.2,D2,m=0.6,β1=7,γ1=2,β2=4,γ2=2,dI(a)≡0.05 and r=6. Then, by choosing various combinations for η1 and η2, we may observe different persistent dynamics. Figure 3 shows the results for three sets of (η1,η2) values, revealing that they may cause stability switch of the positive equilibrium.
Since the model is a DDE system, it is not surprising that the large delay r may destroy the stability of an equilibrium leading to stable periodic solutions with the period depending on the value of r. This is demonstrated in Figure 4, in which d1,m=0.45,D1,m=0.65,d2,m=0.35,D2,m=0.55,β1=8,γ1=2,β2=7,γ2=2,dI(a)≡0.05,η1=η2=1 and r=4,10 respectively. We point out that although the amplitudes of the oscillations change as r varies, these periodic solutions are contained in the invariant set [p,q]=[p1,q1]×[p2,q2] (see Section 3) which depends on the value of r since the value of σ in (4.14) depends on r.
It was discovered both in a two-patch discrete time competitive model [22] and a diffusion model for multiple competitive species [23], with random dispersal in spatially varying but temporally constant environments, that lower dispersal rates are beneficial to the population persistence. Our adoption of a density-dependent dispersal assumes a relatively higher migration rate than that of the random dispersal. This naturally raises a question as to whether or not density-dependent dispersal is advantageous for population's persistence in a single species model. Model (3.3) contains a non-monotone birth function in (3.2), potentially yielding a result different from that of competition models. As remarked above, according to Theorems 4.2 and 4.3, the levels of density dependence of the dispersal (ηi's) do not affect the threshold conditions for extinction/persistence. However, when it persists, the values of ηi's do play a significant role in affecting the final size of population, as the value of positive equilibrium (w∗1,w∗2) depends on ηi's. In Figure 5, we present some numeric results on the impact of ηi's on (w∗1,w∗2), with choices of balanced adaptive dispersals (η1=η2) and imbalanced adaption (η1≠η2), revealing diverse consequences. In Figure 5-(a)-(b), we consider two identical patches by assuming D1,m=D2,m=0.2,d1,m=d2,m=0.1,β1=β2=1,γ1=γ2=1. Take dI(a)≡0.1 and r=10ln5 leading to σ=0.2. Then, we can see in this case that balanced adaption does not influence the final sizes of population; and for the distribution among two patches while imbalanced adaption (with adaption in patch-1 and without adaption in patch-2) not only changes the distribution of population but also decreases the final total population (w∗1+w∗2) as adaption (η1) is increased. In Figure 5-(c)-(d), we set the same parameter values except d2,m=0.3 and β1=2 which means that patch-1 has lower mortality and higher reproduction and is thus a better habitat than patch-2. In Figure 5-(c), even if two groups adopt the same dispersion strategy, the final population decreases in (better) patch-1, so does the total population with respect to η1. The imbalanced adaption leads to a similar but worse consequence. In contrast, when patch-1 is a relative poorer habitat (with the same parameters as in (a) except d1,m=0.3 and β2=2), although Figure 5-(e) shows that the balanced adaption is also disadvantageous to the final population, the imbalanced strategy with only adaption in patch-1, on the other hand, does help boost the species's prosperity (total population) shown in Figure 5-(f).
The intrinsic age-structured dispersal rate, Di(a), varies among organisms. The general case involving a non-constant dispersal rate at each life stage is required for a population to exhibit its sensitive age structure. Our original model incorporating both nonlinear dispersal terms and highly complex delayed birth terms is a biologically meaningful and a mathematically intractable problem. In addition, the derived model with an Allee birth function, for example bi(w)=δiw2e−εiw where δi and εi are constants, can involve the dynamics of bi-stability, and the convergent dynamics in such a non-monotone feedback DDE is also interesting but very challenging. We will address these questions in future research.
C. Y. Cheng and S. S. Chen were partially supported by the Ministry of Science and Technology, Taiwan, R.O.C. (Grant No. MOST 106-2115-M-153-003 and 105-2115-M-003-008); X. Zou was partially supported by NSERC of Canada (RGPIN-2016-04665).
All authors declare no conflicts of interest in this paper.
[1] | A. E. Abdllaoui, P. Auger, B. W. Kooi, et al., Effects of density-dependent migrations on stability of a two-patch predator-prey model, Math. Biosci., 210 (2007), 335–354. |
[2] | S. A. Levin, Dispersion and population interactions, Amer. Natur., 108 (197), 207–228. |
[3] | J. A. J. Metz and O. Diekmann, The Dynamics of Physiologically Structured Populations, Springer-Verlag, New York, 1986. |
[4] | J. W. H. So, J. Wu and X. Zou, Structured population on two patches: modeling dispersal and delay, J. Math. Biol., 43 (2001), 37–51. |
[5] | P. Weng, C. Xiao and X. Zou, Rich dynamics in a non-local population model over three patches, Nonlinear Dynam., 59 (2010), 161–172. |
[6] | D. Xu, Global dynamics and Hopf bifurcation of a structured population model, Nonl. Anal. Real World Appl., 6 (2005), 461–476. |
[7] | C. Yu, J. Wei and X. Zou, Bifurcation analysis in an age-structured model of a single species living in two identical patches, Appl. Math. Model., 34 (2010), 1068–1077. |
[8] | A. J. Terry, Dynamics of a structured population on two patches, J. Math. Anal. Appl., 378 (2011), 1–15. |
[9] | R. Cressman and V. Křivan, Two-patch population models with adaptive dispersal: the effects of varying dispersal speed, J. Math. Biol., 67 (2013), 329–358. |
[10] | W. Wang and Y. Tacheuchi, Adaption of prey and predators between patches, J. Theo. Biol., 258 (2011), 603–613. |
[11] | X. Zhang and W. Wang, Importance of dispersal adaption of two competitive populations between patches, Ecol. Model., 222 (2011), 11–20. |
[12] | C.Huang, Z.Yang, T.Yi, et al., On the basins of attraction for a class of delay differential equations with non-monotone bistable nonlinearities, J. Diff. Eq., 256 (2014), 2101–2114. |
[13] | G. Rost and J. Wu, Domain-decomposition method for the global dynamics of delay differential equations with unimodal feedback, Proc. R. Soc. A, 463 (2007), 2655–2665. |
[14] | Y. Yuan and X. Q. Zhao, Global stability for non-monotone delay equations (with application to a model of blood cell production), J. Diff. Eq., 252 (2012), 2189–2209. |
[15] | A. J. Terry, Impulsive culling of a structured population on two patches, J. Math. Biol., 61 (2010), 843–875. |
[16] | H. C. J. Godfray, L. Partridge and P. H. Harvey, Clutch size, Annu. Rev. Ecol., 22 (1991), 409–429. |
[17] | B. K. Sandercock, Incubation capacity and clutch size determination in two calidrine sandpipers: a test of the four-egg threshold, Oecologia, 110 (1997), 50–59. |
[18] | B. A. Shanbhag, Reproductive strategies in the lizard, Calotes versicolor, Curr. Sci., 84 (2003), 646–652. |
[19] | J. K. Hale and S. M. Verduyn Lunel, Introduction to functional differential equations, Springer, New York, 1993. |
[20] | H. L. Smith, Monotone Dynamical Systems: An Introduction to the theory of Competitive and Cooperative Systems, Mathematical Surveys and Monographs, vol. 41, AMS, Providence, RI, 1995. |
[21] | H. R. Thieme, Persistence under relaxed point-dissipativity (with application to an endemic model), SIAM J. Math. Anal., 24 (1993), 407–435. |
[22] | M. A. McPeek and R. D. Holt, The evolution of dispersal in spatially and temporally varying environments, Amer. Natur., 140 (1992), 1010–1027. |
[23] | J. Dockery, V. Hutson, K. Mischaikow, et al., The evolution of slow dispersal rates: a reaction-diffusion model, J. Math. Biol., 37 (1998), 61–83. |
1. | Isabelle Bueno Silva de Godoy, Blake McGrane-Corrigan, Oliver Mason, Rafael de Andrade Moral, Wesley Augusto Conde Godoy, Plant-host shift, spatial persistence, and the viability of an invasive insect population, 2023, 475, 03043800, 110172, 10.1016/j.ecolmodel.2022.110172 | |
2. | Feng‐Bin Wang, Chang‐Yuan Cheng, Intraguild Predation and Competitions of Two Stage‐Structured Species in a Seasonal Patchy Model, 2025, 0170-4214, 10.1002/mma.10813 |