In this paper, we are concerned with the existence of subharmonic solutions for the degenerate periodic systems of Lotka-Volterra type with impulsive effects. In our degenerate model, the variation of the predator and prey populations may vanish on a time interval, which imitates the (real) possibility that the predation is seasonally absent. Our proof is based on the Poincaré-Birkhoff theorem. By using phase plane analysis, we can find the large gap in the rotation numbers between the "small" solutions and the "large" solutions, which guarantees a suitable twist property. By applying the Poincaré-Birkhoff theorem, we then obtain the existence of subharmonic solutions. Our main theorem extends the associated results by J. López-Gómez et al.
Citation: Yinyin Wu, Fanfan Chen, Qingchi Ma, Dingbian Qian. Subharmonic solutions for degenerate periodic systems of Lotka-Volterra type with impulsive effects[J]. AIMS Mathematics, 2023, 8(9): 20080-20096. doi: 10.3934/math.20231023
[1] | Li Wang, Hui Zhang, Suying Liu . On the existence of almost periodic solutions of impulsive non-autonomous Lotka-Volterra predator-prey system with harvesting terms. AIMS Mathematics, 2022, 7(1): 925-938. doi: 10.3934/math.2022055 |
[2] | Chunmei Song, Qihuai Liu, Guirong Jiang . Harmonic and subharmonic solutions of quadratic Liénard type systems with sublinearity. AIMS Mathematics, 2021, 6(11): 12913-12928. doi: 10.3934/math.2021747 |
[3] | Marimuthu Mohan Raja, Velusamy Vijayakumar, Anurag Shukla, Kottakkaran Sooppy Nisar, Wedad Albalawi, Abdel-Haleem Abdel-Aty . A new discussion concerning to exact controllability for fractional mixed Volterra-Fredholm integrodifferential equations of order $ {r} \in (1, 2) $ with impulses. AIMS Mathematics, 2023, 8(5): 10802-10821. doi: 10.3934/math.2023548 |
[4] | Lini Fang, N'gbo N'gbo, Yonghui Xia . Almost periodic solutions of a discrete Lotka-Volterra model via exponential dichotomy theory. AIMS Mathematics, 2022, 7(3): 3788-3801. doi: 10.3934/math.2022210 |
[5] | Haihua Lu, Tianjue Shi . Multiple periodic solutions of parameterized systems coupling asymmetric components and linear components. AIMS Mathematics, 2025, 10(4): 8988-9010. doi: 10.3934/math.2025412 |
[6] | Yuanlin Ding . Existence and stability analysis of solutions for periodic conformable differential systems with non-instantaneous impulses. AIMS Mathematics, 2025, 10(2): 4040-4066. doi: 10.3934/math.2025188 |
[7] | Yaw Chang, Wei Feng, Michael Freeze, Xin Lu, Charles Smith . Elimination, permanence, and exclusion in a competition model under Allee effects. AIMS Mathematics, 2023, 8(4): 7787-7805. doi: 10.3934/math.2023391 |
[8] | Yinyin Wu, Dingbian Qian, Shuang Wang . Periodic bouncing solutions for sublinear impact oscillator. AIMS Mathematics, 2021, 6(7): 7170-7186. doi: 10.3934/math.2021420 |
[9] | Shuting Chen, Ke Wang, Jiang Liu, Xiaojie Lin . Periodic solutions of Cohen-Grossberg-type Bi-directional associative memory neural networks with neutral delays and impulses. AIMS Mathematics, 2021, 6(3): 2539-2558. doi: 10.3934/math.2021154 |
[10] | M. Adel, M. M. Khader, Hijaz Ahmad, T. A. Assiri . Approximate analytical solutions for the blood ethanol concentration system and predator-prey equations by using variational iteration method. AIMS Mathematics, 2023, 8(8): 19083-19096. doi: 10.3934/math.2023974 |
In this paper, we are concerned with the existence of subharmonic solutions for the degenerate periodic systems of Lotka-Volterra type with impulsive effects. In our degenerate model, the variation of the predator and prey populations may vanish on a time interval, which imitates the (real) possibility that the predation is seasonally absent. Our proof is based on the Poincaré-Birkhoff theorem. By using phase plane analysis, we can find the large gap in the rotation numbers between the "small" solutions and the "large" solutions, which guarantees a suitable twist property. By applying the Poincaré-Birkhoff theorem, we then obtain the existence of subharmonic solutions. Our main theorem extends the associated results by J. López-Gómez et al.
In this paper, we consider the periodic systems of Lotka-Volterra type with impulsive effects
{x′=−α(t)f(y),y′=β(t)g(x),△x(tj)=Ij(x(t−j),y(t−j)),△y(tj)=Jj(x(t−j),y(t−j)),j=±1,±2,⋯, | (1.1) |
where α, β are nonnegative T-periodic continuous functions, △x(tj)=x(t+j)−x(t−j),△y(tj)=y(t+j)−y(t−j), the impulses Ij,Jj:R×R→R are continuous and area-preserving for j=±1,±2,⋯.
In addition, we assume that the impulsive time is T-periodic and there exists an integer l>0 such that
0≤t1<t2<⋯<tl<T,tj+l=tj+TandIj+l=Ij,Jj+l=Jj,j=±1,±2,⋯. | (1.2) |
Throughout this paper, we write
A:=∫T0α(t)dtandB:=∫T0β(t)dt. |
Extensive and interesting research has been conducted on the existence and multiplicity of periodic solutions for second order differential equations without impulsive effects, see [1,2,3,4] by using critical point theory, [5,6,7,8,9] by using Poincaré-Birkhoff theorem, and other references. At the same time, many researchers are also interested in periodic solutions for impulsive systems, see [10,11,12,13]. For the Lotka-Volterra type systems, one can refer to [14,15,16,17,18].
In above previous researches, there has often been some requirements for the preservation of the sign of equations. For example, in (1.1), one of the two coefficients α(t) and β(t) is required to be strictly positive. From the point of view of using the Poincaré-Birkhoff theorem and phase plane analysis, these strictly positive conditions are used to obtain suitable twist properties. Otherwise, if both coefficient functions change sign or vanish on a subinterval, the associated phase plane analysis is considerably more difficult.
Recently, many researchers have been paying attention to the system where coefficient functions are degenerate, that is, α(t) and β(t) could both be zero. These degenerate conditions imitate the (real) possibility that the predation is seasonally absent.
In [19], López-Gómez et al. considered a Hamiltonian system
{x′=−λα(t)f(y),y′=λβ(t)g(x), | (1.3) |
where parameter λ>0. By the Poincaré-Birkhoff theorem, the authors obtained the existence of subharmonic solutions for sufficiently large λ, under the following assumptions:
(A1) Let α⪈0 and β⪈0 be T-periodic continuous functions such that
α(t0)β(t0)>0 for some t0∈[0,T]. | (1.4) |
(A2) Let f,g∈C(R) be locally Lipschitz functions such that f,g∈C1 on a neighborhood of the origin and
{f(0)=0,f(y)y>0 for all y≠0,g(0)=0,g(x)x>0 for all x≠0,f′(0)>0,g′(0)>0. |
(A3) Either f, or g, satisfies, at least, one of the following conditions:
(f−) f is bounded in R−,(f+) f is bounded in R+,(g−) g is bounded in R−,(g+) g is bounded in R+. |
Meanwhile, a predator-prey model of Lotka-Volterra type has been considered in [20], in which the coefficients satisfy αβ=0, that is,
{u′=α(t)u(1−v),v′=β(t)v(−1+u), | (1.5) |
where α(t) and β(t) are real continuous T-periodic functions such that
α(t)>0,∀t∈(0,T2),α(t)=0,∀t∈[T2,T];β(t)=0,∀t∈[0,T2],β(t)>0,∀t∈(T2,T). |
By constructing the iterates of the monodromy operator of the system, it was shown in [20] that the system (1.5) possesses exactly two 2T-periodic solutions if AB>4.
Further, López-Gómez et al. [21] considered the degenerate model (1.3) with (A1) and (A2). Here "degenerate" means that the set
Z:=supp(α)∩supp(β) |
has Lebesgue measure zero, that is |Z|=0. According to some geometric configurations of α and β, for large λ>0, the existence of a large number of subharmonic solutions is obtained in [21].
The strategy used in [19,21] is to expand the parameter λ>0 such that the rotational motions of "large" solutions and "small" solutions produce enough angular gap in mT time in phase plane, which implies a twist property. Then the Poincaré-Birkhoff theorem can be applied. Thus, a natural and interesting problem is whether we can apply Poincaré-Birkhoff theorem to degenerate impulsive systems (1.1) without expanding parameter λ. This is one of the motivations for the research presented in this paper.
Due to the presence of pulses, even the simplest pulse function can potentially give rise to complex dynamical phenomena, posing challenges for subsequent research. There are only a few results on the existence of periodic solutions to impulsive equations of Lotka-Volterra type. In [22], Tang and Chen investigated a classical periodic Lotka-Volterra predator-prey system with impulsive effect. By using the method of coincidence degree, the authors proved the existence of strictly positive periodic solution.
In the following, we consider the impulsive periodic systems of Lotka-Volterra type (1.1). We preserve (A2)–(A3) and assume that
(H1) Let α⪈0, β⪈0 be T-periodic continuous functions such that A>0 and B>0.
By (A2), there exists a constant η>0 such that
min{f′(0),g′(0)}>η. | (1.6) |
Then by the limit definition, we find ε0>0 such that,
f(ζ)ζ≥ηζ2, g(ζ)ζ≥ηζ2,for |ζ|≤ε0. |
Moreover, we give the following assumptions about the impulse functions.
(H2) For some η satisfying (1.6), there exists a constant δ1, such that
|Ij(x,y)|<μr, |Jj(x,y)|<μr,for r<δ1, |
where r=√x2+y2 and the constant μ satisfies
larcsin(√2μ)<μ∗:=min{π6(l+1),π12,ηAsin2π12,ηBsin2π12}, |
where l as defined in (1.2).
(H3) There exists M0>0 such that
|Ij(x,y)|<M0,|Jj(x,y)|<M0,j=±1,±2,⋯. |
The main result of this paper is the following:
Theorem 1.1. Assume that (A2)–(A3), (H1)–(H3) hold. Then for every positive integer k≥1, there exists a positive integer m∗(k) such that, for every integer m≥m∗(k), the system (1.1) has at least two mT-periodic solutions having k as a rotation number.
Remark 1.1. Theorem 1.1 generalizes the relevant results [19,21] of López-Gómez et al.
(1) In contrast to the non-degenerate condition (A1), condition (H1) allows the coefficient functions α and β to vanish on a subinterval. This implies that the solutions near the origin may enter the origin in the phase plane which leads to a bad evaluation of the rotations. By careful phase plane analysis, we can overcome this problem.
(2) Our model does not have a parameter λ. In [19,21], as the parameter λ increases, the gap in the rotation number between the small and the large solutions increases. But this approach is not directly applicable to our case. For our system (1.1) without parameter, it is difficult to obtain enough gap between the rotation numbers of small and large solutions. In Section 3, we overcome all difficulties by phase plane analysis.
The rest of the paper is organized as follows. In Section 2, we consider the degenerate systems without impulses and give some preliminary results. In Section 3, we provide the proof of Theorem 1.1. Firstly, we prove that the solutions of (1.1) near the origin can complete many turns around the origin via phase plane analysis. Next, we show that the solutions of (1.1) far from the origin cannot complete one turn. Finally, we obtain the existence of subharmonic solutions by Poincaré-Birkhoff theorem.
For the sake of convenience, we introduce some basic lemmas and tools that will be used in the next section. First, we consider the Hamiltonian system without impulsive terms
{x′=−α(t)f(y),y′=β(t)g(x), | (2.1) |
where α, β, f and g satisfy (A2)–(A3) and (H1).
Apparently, (0,0) is a solution of (2.1). By the uniqueness of the solution, if the initial value (x(0),y(0))≠(0,0), then the solution (x(t),y(t))≠(0,0) of (2.1), for t∈R. Introducing polar coordinates
x(t)=r(t)cosθ(t),y(t)=r(t)sinθ(t). |
Let
D1={(x,y): x>0, y≥0},D2={(x,y): x≤0, y>0}, |
D3={(x,y): x<0, y≤0},D4={(x,y): x≥0, y<0}. |
Similar to the discussion in [19, Remark 1], by (A2) and (H1), all solutions of (2.1) exists globally. Further, we have the following spiral property.
Lemma 2.1. Assume that (A2)–(A3) and (H1) hold. For any fixed positive integer m, there are a sufficiently large R∗>0 and two strictly monotonically increasing functions ξ±m:[R∗,+∞)→R such that
ξ±m(s)→+∞⟺s→+∞. |
Moreover, let z(t)=(x(t),y(t)) be the solution of (2.1) with R0=|z(0)|≥R∗. Then we have either
ξ−m(R0)≤|z(t)|≤ξ+m(R0),t∈[0,mT], |
or there exists t∗∈(0,mT) such that
θ(t∗)−θ(t0)=2π, |
and
ξ−m(R0)≤|z(t)|≤ξ+m(R0),t∈[0,t∗]. |
Proof. Without loss of generality we assume that z(t) is the solution of (2.1) with |z(0)|=R0, where R0 is sufficiently large.
Step 1. The estimation of upper bound on the solution z(t).
As a first case, we assume that z(0)∈D2. If z(t)∈D2 for all t∈[0,mT]. Notice that x′(t)≤0 and y′(t)≤0, we have 0≤y(t)≤y(0). Taking
M=max{|f(y)|:0≤y≤y(0)}, |
then
x(t)=∫t0x′dt≥−∫mT0α(t)Mdt=−mAM:=−R1. | (2.2) |
Otherwise, there exists a t1∈[0,mT], such that y(t1)=0 and z(t)∈D2 for t∈[0,t1), then (2.2) still holds.
If z(t)∈D3 for all t∈[t1,mT]. Then, we have x′(t)≥0, y′(t)≤0. Therefore, −R1≤x(t)≤0. Taking
N=max{|g(x)|:−R1≤x≤0}, |
then we have
y(t)=∫tt1y′dt≥−∫mT0β(t)Ndt=−mBN:=−R2. | (2.3) |
Otherwise, there exists a t2∈(t1,mT], such that x(t2)=0 and z(t)∈D3 for t∈[t1,t2). Thus, for t∈[t1,t2], (2.3) still holds.
If z(t)∈D4 for all t∈[t2,mT], we have x′(t)≥0, y′(t)≥0, which implies that −R2≤y≤0. Taking
M1=max{|f(y)|:−R2≤y≤0},x(t)=∫tt2x′dt≤∫mT0α(t)M1dt=mAM1:=R3. | (2.4) |
Otherwise, there exists a t3∈(t2,mT], such that y(t3)=0 and z(t)∈D4 for t∈[t2,t3). Then for t∈[t2,t3], (2.4) still holds.
If z(t)∈D1 for all t∈[t3,mT], we have x′(t)≤0, y′(t)≥0, which implies that 0≤x(t)≤R3. Taking
N1=max{|g(x)|:0≤x≤R3}. |
It follows that
y(t)=∫tt1y′dt≤∫mT0β(t)N1dt=mBN1:=R4. | (2.5) |
Otherwise, there exists a t∗∈(t3,mT], such that x(t∗)=0 and z(t)∈D1 for t∈[t3,t∗). Then for t∈[t3,t∗], (2.5) still holds.
Let ξ+m(z(0))=√2max{R0,R1,R2,R3,R4}. Then
ξ+m(z(0))→+∞⟺|z(0)|→+∞. |
For the other case, that is z(0)∈D3, z(0)∈D4 and z(0)∈D1, similar to the discussion above, we can find the an upper bound on each quadrant. For briefness, we still denote by Ri, i=0,⋯,4 the associated upper bound on each quadrant. Note that Ri, i=0,⋯,4, depends on z(0)∈R2, and therefore, the choice of ξ+m only depends on z(0).
Step 2. The estimation of lower bound on the solution z(t).
To complete our analysis, we need to look at what happens in the every quadrants. At first, we consider the case of z(t)∈D2.
For z(t)∈D2, there exists t′1∈[0,t1] such that y(t′1)=−x(t′1). we claim that
|x(t′1)|→+∞⟺|z0|→+∞. | (2.6) |
For otherwise, we assume that there exists a constant J1 which is independent of |z0|, such that
0<−x(t)≤J1,for t∈(0,t′1]. |
On the other hand,
y(t′1)=y(0)+∫t′10y′dt≥|z0|−([t′1T]+1)BM′g, |
where M′g=max{|g(x)|:0<−x≤J1}. Then y(t′1)→+∞ as |z0|→+∞. But this is a contradiction with the fact that y(t′1)≤J.
Since x′≤0 and y′≤0, we have
|z(t)|≥y(t)≥y(t′1)for t∈[0,t′1] | (2.7) |
and
|z(t)|≥−x(t)≥−x(t′1)for t∈[t′1,t1]. | (2.8) |
Combine (2.6)–(2.8), there exists ξ−m,2(z0) such that
ξ−m,2(z(0))→+∞⟺|z(0)|→+∞and|z(t)|≥ξ−m,2(z(0)) for t∈[0,t1]. |
In particular, |z(t1)|→+∞⟺|z0|→+∞.
Similarly, we can discuss the cases of z(t)∈D3, D4, D1. In conclusion, we can find ξ−m,3(z(t1)), ξ−m,4(z(t2)) and ξ−m,1(z(t3)) such that
ξ−m,3(z(t1))→+∞⟺|z(t1)|→+∞,|z(t)|≥ξ−m,3(z(t1)) for t∈[t1,t2], |
ξ−m,4(z(t2))→+∞⟺|z(t2)|→+∞,|z(t)|≥ξ−m,4(z(t2)) for t∈[t2,t3], |
ξ−m,1(z(t3))→+∞⟺|z(t3)|→+∞,|z(t)|≥ξ−m,1(z(t3)) for t∈[t3,t∗]. |
Set
ξ−m(z(0))=ξ−m,1∘ξ−m,4∘ξ−m,3∘ξ−m,2(z(0)). |
Then,
|z(t)|≥ξ−m(z(0))for t∈[0,t∗]. |
Notice that the discussions in Steps 1 and 2 are true for not only for initial time t0=0 but also for all initial time t0∈[0,T]. Therefore, we can rewrite ξ±m(z(0)) as ξ±m(R0), which means that the upper and lower bounds of the solution depend only on R0=|z(t0)|. The proof of the lemma is complete.
Now let us return to the impulsive Eq (1.1). Notice that the motion of the solution of (1.1) is the same as that of (2.1) until it meets the next impulsive time. We consider the behavior of small solutions. Denote by DR the disc of radius R centered at zero.
According to [23], (A2) guarantee the existence and uniqueness of the solution. By boundedness of impulses, similar to the proof of [19, Proposition 1], we have the following result.
Proposition 1. Assume that (A2)–(A3), (H1)–(H3) hold. For every integer m≥1 and ε>0, there exists δ=δ(m,ε)> 0 such that if (x0,y0)∈Dδ, then the unique solution of (1.1), (x(t),y(t)), with (x(0),y(0))=(x0,y0), satisfies (x(t),y(t))∈Dε for all t∈[0,mT].
By Proposition 1, there exists δ0=δ0(m,δ1) such that if z(t)=(x(t),y(t)) is any solution of (1.1) with initial value z0=(x(0),y(0))∈Dδ0, then z(t)∈Dδ1. For the nontrivial solution z(t)∈Dδ1, by (H2), we have √2μ<12, then Δr(tj)=√(△x(tj))2+(△y(tj))2<12r(t−j). This implies that z(t) never passes the origin, that is, if z0≠(0,0) then z(t)≠(0,0) for t∈R. Then the rotation number of the solution z(t) associated with the initial value z0 can be defined as
rot(z0;[0,mT]):=θ(mT)−θ(0)2π. |
The rotation number is an algebraic counter of the counterclockwise turns of the solution z(t) around the origin during the time-interval [0,mT].
Now we introduce a generalized version of the Poincaré-Birkhoff theorem [11, Theorem 2.1] to prove the existence of subharmonic solutions.
Theorem 2.1. Let A be an annular region bounded by two strictly star-shaped curves around the origin, Γ−and Γ+,Γ−⊂int(Γ+), where int(Γ+)denotes the interior domain bounded by Γ+. Suppose that F:¯int(Γ+)→R2 is an area-preserving homeomorphism and F|A admits a lifting, with the standard covering projection Π:(θ,r)↦z=(rcosθ,rsinθ), of the form
˜F|A:(θ,r)↦(θ+h(θ,r),w(θ,r)), |
where h and w are continuous functions of period T in the first variable. Correspondingly, for ˜Γ−=Π−1(Γ−) and ˜Γ+=Π−1(Γ+), assume the twist condition
h(θ,r)>0on ˜Γ−,h(θ,r)<0on ˜Γ+. | (2.9) |
Then, F has two fixed points z1,z2 in the interior of A, such that
h(Π−1(z1))=h(Π−1(z2))=0. |
Remark 2.1. In (2.9), the function h(θ,r) means ΔθF(z), the difference between polar angles of z and its image through F(z). The continuity of this function is essential to verify the twist condition. In fact, Poincaré map of impulsive system is an iterative map by the section maps of flows and the jump maps. See [11] for more details.
In order to apply the Poincaré-Birkhoff theorem to the impulsive Eq (1.1) where α and β may vanish on a subinterval, we need to characterize a suitable twist condition. Therefore, we should evaluate the rotation number of small and large solutions (see Lemmas 3.1 and 3.3).
Let z(t;z0) be the solution of (1.1) and (θ(t;θ0,r0),r(t;θ0,r0)) the polar coordinates of z(t;z0). First of all, we find some small solutions of (1.1) with a given lower bound on the rotation number on the interval [0,mT]. More precisely, we have the following lemma.
Lemma 3.1. Assume that (A2), (H1)–(H3) hold. For any given positive integer k>1, there exists a integer m∗(k)≥1 such that, for every integer m≥m∗(k), there exists r−(m)>0 such that for
θ(mT;θ0,r0)−θ0>2kπ,for θ0∈R and r0=r−(m). |
Proof. For simplicity, let z(t)=z(t;z0) and (θ(t),r(t))=(θ(t;θ0,r0),r(t;θ0,r0)). We assume that (x(t),y(t))∈Dε. Let's introduce notations
Δijθ=:θ(t−j)−θ(t+j−1),j=(i−1)l+1,⋯,il, |
Δiil+1θ=:θ(iT)−θ(t+il)andΔiΘ=:il∑j=(i−1)l+1(θ(t+j)−θ(t−j)), |
where l as defined in (1.2).
We first consider the case where i=1. For (x(t),y(t))∈Dδ1 t∈[0,T], t≠tj, j=1,⋯,l, we have
θ′(t)=y′(t)x(t)−y(t)x′(t)x2(t)+y2(t)=α(t)f(y(t))y(t)+β(t)g(x(t))x(t)x2(t)+y2(t)≥η(α(t)sin2θ+β(t)cos2θ). |
By (H2), we get
sin(|θ(t+j)−θ(t−j)|)=√△2x(tj)+△2y(tj)x2(t−j)+y2(t−j)<√2μr(t−j)r(t−j)=√2μ, |
it follows that |θ(t+j)−θ(t−j)|<arcsin(√2μ), thus
|Δ1Θ|≤l∑j=1|θ(t+j)−θ(t−j)|<larcsin(√2μ). | (3.1) |
There are two cases to be considered.
Case 1.1. Let θ(0)∈[16π,23π]⋃[76π,53π]. For t∈[0,T], t≠tj, j=1,⋯,l, we have θ′(t)≥0. So either
Δ1jθ≥π6(l+1),for some j∈{1,⋯,l+1}, |
or
0≤Δ1jθ≤π6(l+1),for all j∈{1,⋯,l+1}. |
For the former, by (H2) and (3.1), we have
θ(T)−θ(0)=∑l+1j=1Δ1jθ+Δ1Θ≥π6(l+1)−larcsin(√2μ)>0. |
For the latter, by (H2) and (3.1), we obtain |ΔiΘ|<π12. Hence,
θ(t)∈[112π,1112π]⋃[1312π,2312π],for t∈[0,T], |
which implies that
θ′(t)≥ηα(t)sin2112π, for t≠tj, j=1,⋯,l. |
Integrating both sides of the above inequality from 0 to T, we have
θ(T)−θ(0)=l∑j=1∫tjtj−1θ′(t)dt+∫Ttlθ′(t)dt+Δ1Θ≥ηAsin2112π−larcsin(√2μ)>0. |
Case 1.2. Let θ(0)∈[−13π,16π]⋃[23π,76π]. Similarly, either
Δ1jθ≥π6(l+1),for some j∈{1,⋯,l+1}, |
or
0≤Δ1jθ≤π6(l+1),for all j∈{1,⋯,l+1}. |
For the former, we get
θ(T)−θ(0)=l+1∑j=1Δ1jθ+Δ1Θ≥π6(l+1)−larcsin(√2μ)>0. |
For the latter, we see
θ(t)∈[−512π,512π]⋃[712π,1712π],for t∈[0,T]. |
Therefore,
θ′(t)≥ηβ(t)cos2512π=ηβ(t)sin2112π, for t≠tj, j=1,⋯,l. |
Integrating both sides of the above inequality from 0 to T, we find
θ(T)−θ(0)=l∑j=1∫tjtj−1θ′(t)dt+∫Ttlθ′(t)dt+Δ1Θ≥ηBsin2112π−larcsin(√2μ)>0. |
We take Θ∗=μ∗−larcsin(√2μ)>0 and obtain
θ(T)−θ(0)≥Θ∗. | (3.2) |
Let m∗(k)=[2πkΘ∗]+1. Thus, for a given positive integer m≥m∗(k), there exists a small circle r=r−(m,δ1) such that the solution of (1.1) starting from r=r−(m) completes k counterclockwise rotations on [0,mT], that is,
θ(mT;θ0,r0)−θ0>2kπ. |
As is well-known, the global existence of solutions is a crucial requirement for applying the Poincaré-Birkhoff theorem. To this end, we will prove that the solution of (1.1) possesses a spiral property in the phase plane, which guarantees the global existence of the solution.
Lemma 3.2. Assume that (A2) and (A3), (H1)–(H3) hold. For fixed m≥m∗(k) and sufficiently large R∗>0, there exists R±m:[R∗,+∞)→R such that
R±m(s)→+∞⟺s→+∞. |
Moreover, let z(t) be the solution of (1.1) with r0=|z(t0)|≥R∗. Then we have either
R−m(r0)≤|z(t)|≤R+m(r0),∀t∈[t0,t0+mT], |
or there exists t∗∈(t0,t0+mT) such that
θ(t∗)−θ(t0)=2π, |
and
R−m(r0)≤|z(t)|≤R+m(r0),∀t∈[t0,t∗]. |
Proof. Let z(t) be the solution of (1.1) satisfying the initial value z(t0), where |z(t0)| is sufficiently large. Notice that for impulsive Eq (1.1), the motion of the solution is same as the motion of the corresponding equation without impulses until it meets the next impulse time. By Lemma 2.1, we can find curves ξ±j,m(r(t+j−1)):=ξ±m(r(t+j−1)), j=1,⋯,ml+1, such that
ξ±j,m(r(t+j−1))→+∞⟺r(t+j−1)→+∞, j=1,⋯,ml+1. |
Moreover,
ξ−j,m(r(t+j−1))≤r(t)≤ξ+j,m(r(t+j−1)),t∈(t+j−1,t−j), j=1,⋯,ml, |
and
ξ−ml+1,m(r(t+ml))≤r(t)≤ξ+ml+1,m(r(t+ml)),t∈(t+ml,mT]. |
In particular,
ξ−j,m(r(t+j−1))≤r(t−j)≤ξ+j,m(r(t+j−1)),j=1,⋯,ml. | (3.3) |
By (H3), we get
|r(t+j)−r(t−j)|=√I2j+J2j<√2M0,j=1,⋯,ml. |
Hence, for j=1,⋯,ml,
Q−j(r(t+j)):=ξ−j,m(r(t+j−1))−√2M0≤r(t+j)≤ξ+j,m(r(t+j−1))+√2M0=:Qj(r(t+j)). | (3.4) |
Let
R±m(r0):=ξ±ml+1,m∘Q±ml∘⋯ξ±2,m∘Q±1∘ξ±1,m(r0) |
Combining (3.3) with (3.4), we have
R±m(r0)→+∞⟺r0→+∞, |
and
R−m(r0)≤r(t)≤R+m(r0),∀t∈[0,mT]. |
Further, we prove that the solution z(t) with sufficiently large |z(t0)|, cannot complete one turn around the origin on [0,mT], where m is given by Lemma 3.1.
Lemma 3.3. Assume that (A3), (H1)–(H3) hold. For any given m∈N, there exists r+(m)>0 such that,
θ(mT;θ0,r0)−θ0<2π,for θ0∈R and r0=r+(m). |
Proof. Without loss of generality we assume that (H3) holds with g satisfying (g−). Then there exists Mg>0 such that
supx≤0|g(x)|≤Mg. |
Let z(t)=(x(t),y(t)) be a nontrivial solution of (1.1) with the initial value z0=(x0,y0). There are two cases to discuss.
Case 3.1. Let z0∈D1⋃D3⋃D4. We claim that z(t) cannot cross entirely D2 during [0,mT], if |z0| is sufficiently large.
Proof by contradiction. Assume that there exists [τ1,τ2]⊂[0,mT], τ1,τ2≠tj, j=1,⋯,ml, such that
0≥x(τ1)≥−M0, y(τ1)>0,0>x(τ2), M0≥y(τ2)≥0, |
where M0 is given by (H5). Then,
|y(τ1)|≤|y(τ2)|+|Mg∫τ2τ1β(s)ds|+∑τ1<tj<τ2|Jj(x(t−j),y(t−j))|≤M0+mMgB+mlM0:=R∗. |
By Lemma 3.2, we choose r+(m) such that for r0=r+(m), |z(t)|≥R−m(r0)>R∗+M0, t∈[0,mT]. This leads to a contradiction. Hence,
θ(mT;θ0,r0)−θ0<2π. |
Case 3.2. Let z0∈D2. We claim that z(t) cannot cross entirely D3 during [0,mT], if |z0| is sufficiently large.
Similarly, by contradiction, we assume that there exists [τ3,τ4]⊂[0,mT], τ3,τ4≠tj, j=1,⋯,ml, such that
0>x(τ3), 0≥y(τ3)≥−M0,0≥x(τ4)≥−M0, 0≥y(τ4), |
Thus,
|y(t)|≤|y(τ3)|+M∫τ4τ3β(s)ds+∑τ3<tj<τ4|Jj(x(t−j),y(t−j))|≤M0+mMgB+mlM0=R∗. |
Let
N∗:=max{|f(y)|:|y|≤R∗}, |
then
|x(τ3)|≤|x(τ4)|+N∗∫τ4τ3α(s)ds+∑τ3<tj<τ4|Ij(x(t−j),y(t−j))|≤M0+mN∗A+mlM0:=R∗∗. |
By Lemma 3.2, we choose r+(m) such that for r0=r+(m), |z(t)|≥R−m(r0)>R∗∗+M0, t∈[0,mT]. This leads to a contradiction. Hence,
θ(mT;θ0,r0)−θ0<2π. |
Now we are in a position to consider the Poincaré mapping associated with (1.1). According to Lemma 3.2, the solutions of (1.1) exists globally. Denote by (x(t;x0,y0),y(t;x0,y0)) the solution of (1.1) with the initial value (x0,y0)=(x(0;x0,y0),y(0;x0,y0)) and define
P0:(x0,y0)→(x(t1−;x0,y0),y(t1−;x0,y0)),P1:(x1,y1)→(x(t2−;x1,y1),y(t2−;x1,y1)), ⋮Pj:(xj,yj)→(x(tj+1−;xj,yj),y(tj+1−;xj,yj)), ⋮Pml−1:(xml−1,yml−1)→(x(t−ml;xml−1,yml−1),y(t−ml;xml−1,yml−1)),Pml:(xml,yml)→(x(mT;xml,yml),y(mT;xml,yml)), |
where (xj,yj)=(x(tj+;xj,yj),y(tj+;xj,yj)), j=1,2,⋯,ml. The jumping map is defined by
Φj:(x,y)⟼(x+Ij(x,y),y+Jj(x,y)),j=1,2,⋯,ml. |
Then the Poincaré mapping P associated with (1.1) can be written in the form
P=Pml∘Φml∘⋯P1∘Φ1∘P0. |
Since
{x′=−α(t)f(y),y′=β(t)g(x), |
is conservative, Pj, j=0,2,⋯,ml, are symplectic. And Φj, j=1,2,⋯,ml, are area-preserving homeomorphisms. Hence, the Poincaré mapping P is area-preserving. Moreover, by Lemmas 3.2 and 3.3, P satisfies the boundary twist condition.
Finally, we apply the Poincaré-Birkhoff theorem to prove the existence of subharmonic solutions.
Applying Theorem 2.1, P has at least two geometrically distinct fixed points
(xi,yi)=(ricosθi,risinθi),i=1,2, |
which correspond to two mT-periodic solutions of system (1.1) with
θ(mT;θi,ri)−θi=2kπ,i=1,2, |
that is, rot(zi,0;[0,mT])=k, i=1,2.
The proof of Theorem 1.1 is thus completed.
Remark 3.1. For any positive integer m′≥m∗(k), there exist a new r±(m′) such that P satisfies the boundary twist condition. Then, by Theorem 1.1, P has at least two geometrically distinct fixed points
(x(m′)i,y(m′)i)=(r(m′)icosθ(m′)i,r(m′)isinθ(m′)i),i=1,2, |
which correspond to two mT-periodic solutions of system (1.1) with rot(zm′i,0;[0,mT])=k, i=1,2. Therefore, system (1.1) has infinitely many geometrically distinct mT-periodic solutions.
Our system (1.1) covers many mathematical models with biological properties. We briefly illustrate our degenerate systems with examples.
A classical predator-prey model of Lotka-Volterra type [19,20,21] has the form
{x′=x(a(t)−b(t)y),y′=y(−c(t)+d(t)x), | (4.1) |
where a,b,c,d:R→R are T-periodic functions and b(t)⪈0,d(t)⪈0. If system (4.1) possesses a T-periodic coexistence state (˜x(t),˜y(t)), namely, a positive componentwise T-periodic solution of the system, then by change of variables
x(t)=u(t)˜x(t),y(t)=v(t)˜y(t), |
system (4.1) is changed into the equivalent system
{u′=b(t)˜y(t)u(1−v),v′=d(t)˜x(t)v(−1+u). | (4.2) |
Commonly, (1, 1) is regarded as a trivial coexistence state of system (4.2). From a geometric point of view, nontrivial componentwise positive mT-periodic solutions are the trajectories inside the first quadrant wound around the equilibrium point.
For the convenience of study, we can move the equilibrium point to the origin. System (4.2) is changed, via the change of variables x=lnu and y=lnv, into the system
{x′=−α(t)(−1+ey),y′=β(t)(−1+ex). | (4.3) |
with α(t)⪈0 and β(t)⪈0. Obviously, (4.3) satisfies (A2) and (A3) and is a particular case of (1.1).
Now, let's provide some biological explanations for (H1). In nature, a predator or prey can maintain a fairly constant density despite fluctuations in the density of its prey. At certain periodic times (certain seasons) of the year, the predators hunt preys and collect them. At other periodic times, they stop hunting the preys and simply consume what they have collected, during which the prey is unaffected by the predator.
Introducing a simple prototype model for system (1.1). In Mediterranean pine forests, there is a species of moth called the pine processionary. Typically, in late summer or early autumn, they lay their eggs on pine trees. After a period of incubation, the larvae emerge, and the larval stage can last for several months, during which the caterpillars actively feed on pine needles. the caterpillars actively feed on pine needles, causing defoliation. During this period, the density of pine trees remains constant, while the density of pine processionary caterpillars fluctuates. Depending on the impact of defoliation, the population of the next generation of caterpillars can increase or decrease before the start of a new cycle. Coefficient functions α and β vanish on some subintervals to simulate this property, which makes our model more realistic.
On the other hand, models (4.1) and (4.2) are affected by short-term perturbations, which are usually considered as impulse terms. It is well known that many biological systems are disturbed by human activities, such as biological control, feeding, harvesting, and planting in fisheries and forest management, chemotherapy to cancer cells, and regular release of toxins from environmental pollution. In order to describe this system more accurately, continuous human activity is often removed from the model and replaced with impulses [22,24,25,26].
A commonly employed method in biological pest control is the release of natural enemies to reduce the population of pests. For example, parasitic wasps are natural enemies of many pests and are commonly used in biological pest control. In greenhouse production, pests often experience outbreaks during the summer season. In this period of time, we can release parasitic wasps to control the growth of pests, and in some cases, even eliminate them.
In this paper, based on Poincaré-Birkhoff theorem, by using phase plane analysis, we show that for every integer m≥m∗(k), impulsive systems (1.1) has at least two mT-periodic solutions. From a biological perspective, Theorem 1.1 indicates that small bounded disturbances caused by human activities do not affect the existence of infinitely many equilibrium states.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
We would like to thank the referees, who made many useful comments and suggestions for improvements of this paper. This work is supported by National Natural Science Foundation of China (No. 12071327) and Fundamental Research Funds of Zhejiang Sci-Tech University (No. 11432832612202).
All authors declare no conflicts of interest in this paper.
[1] | P. Rabinowitz, Minimax methods in critical point theory with applications to differential equations, In: CBMS Regional Conference Series in Mathematics, 1986. https://dx.doi.org/10.1090/cbms/065 |
[2] | J. Mawhin, M. Willem, Critical point theory and Hamiltonian systems, New York: Springer, 1989. https://dx.doi.org/10.1007/978-1-4757-2061-7 |
[3] |
Y. Long, Multiple solutions of perturbed superquadratic second order Hamiltonian systems, Trans. Amer. Math. Soc., 311 (1989), 749–780. https://dx.doi.org/10.2307/2001151 doi: 10.2307/2001151
![]() |
[4] | A. Fonda, Playing around resonance: An invitation to the search of periodic solutions for second order ordinary differential equations, Cham: Springer, 2016. https://dx.doi.org/10.1007/978-3-319-47090-0 |
[5] |
H. Jacobowitz, Periodic solutions of x″+f(x,t)=0 via the Poincaré-Birkhoff theorem, J. Differ. Equations, 20 (1976), 37–52. https://dx.doi.org/10.1016/0022-0396(76)90094-2 doi: 10.1016/0022-0396(76)90094-2
![]() |
[6] |
A. Margheri, C. Rebelo, F. Zanolin, Maslov index, Poincaré-Birkhoff theorem and periodic solutions of asymptotically linear planar Hamiltonian systems, J. Differ. Equations, 183 (2002), 342–367. http://dx.doi.org/10.1006/jdeq.2001.4122 doi: 10.1006/jdeq.2001.4122
![]() |
[7] |
A. Fonda, Positively homogeneous Hamiltonian systems in the plane, J. Differ. Equations, 200 (2004), 162–184. http://dx.doi.org/10.1016/j.jde.2004.02.001 doi: 10.1016/j.jde.2004.02.001
![]() |
[8] |
D. Qian, P. J. Torres, Periodic motions of linear impact oscillators via the successor map, SIAM J. Math. Anal., 36 (2005), 1707–1725. http://dx.doi.org/10.1137/S003614100343771X doi: 10.1137/S003614100343771X
![]() |
[9] |
D. Qian, P. J. Torres, P. Wang, Periodic solutions of second order equations via rotation numbers, J. Differ. Equations, 266 (2019), 4746–4768. http://dx.doi.org/10.1016/j.jde.2018.10.010 doi: 10.1016/j.jde.2018.10.010
![]() |
[10] |
F. Jiang, J. Shen, Y. Zeng, Applications of the Poincaré-Birkhoff theorem to impulsive Duffing equations at resonance, Nonlinear Anal. Real, 13 (2012), 1292–1305. http://dx.doi.org/10.1016/j.nonrwa.2011.10.006 doi: 10.1016/j.nonrwa.2011.10.006
![]() |
[11] |
D. Qian, L. Chen, X. Sun, Periodic solutions of superlinear impulsive differential equations: A geometric approach, J. Differ. Equations, 258 (2015), 3088–3106. http://dx.doi.org/10.1016/j.jde.2015.01.003 doi: 10.1016/j.jde.2015.01.003
![]() |
[12] |
Y. Niu, X. Li, Periodic solutions of semilinear Duffing equations with impulsive effects, J. Math. Anal. Appl., 467 (2018), 349–370. http://dx.doi.org/10.1016/j.jmaa.2018.07.008 doi: 10.1016/j.jmaa.2018.07.008
![]() |
[13] |
J. Shen, L. Chen, X. Yuan, Lagrange stability for impulsive Duffing equations, J. Differ. Equations, 266 (2019), 6924–6962. http://dx.doi.org/10.1016/j.jde.2018.11.022 doi: 10.1016/j.jde.2018.11.022
![]() |
[14] |
A. Hausrath, R. Manásevich, Periodic solutions of a periodically perturbed Lotka-Volterra equation using the Poincaré-Birkhoff theorem, J. Math. Anal. Appl., 157 (1991), 1–9. http://dx.doi.org/10.1016/0022-247X(91)90132-J doi: 10.1016/0022-247X(91)90132-J
![]() |
[15] | T. Ding, F. Zanolin, Harmonic solutions and subharmonic solutions for periodic Lotka-Volterra systems, In: Dynamical systems: Proceedings Of the special program at Nankai institute of mathematics, World Scientific, 1993. |
[16] | T. Ding, F. Zanolin, Periodic solutions and subharmonic solutions for a class of planar systems of Lotka-Volterra type, In: World congress of nonlinear analysts'92, Boston: De Gruyter, 1996. http://dx.doi.org/10.1515/9783110883237.395 |
[17] |
A. Fonda, R. Toader, Subharmonic solutions of Hamiltonian systems displaying some kind of sublinear growth, Adv. Nonlinear Anal., 8 (2019), 583–602. http://dx.doi.org/10.1515/anona-2017-0040 doi: 10.1515/anona-2017-0040
![]() |
[18] |
T. Dondé, F. Zanolin, Multiple periodic solutions for one-sided sublinear systems: A refinement of the Poincaré-Birkhoff approach, Topol. Methods Nonlinear Anal., 55 (2020), 565–581. http://dx.doi.org/10.12775/TMNA.2019.104 doi: 10.12775/TMNA.2019.104
![]() |
[19] |
J. López-Gómez, E. Muñoz-Hernández, F. Zanolin, On the applicability of the poincaré-Birkhoff twist theorem to a class of planar periodic predator-prey models, Discrete Contin. Dyn. Syst., 40 (2020), 2393–2419. http://dx.doi.org/10.3934/dcds.2020119 doi: 10.3934/dcds.2020119
![]() |
[20] |
J. López-Gómez, E. Muñoz-Hernández, Global structure of subharmonics in a class of periodic predator-prey models, Nonlinearity, 33 (2020), 34. http://dx.doi.org/10.1088/1361-6544/ab49e1 doi: 10.1088/1361-6544/ab49e1
![]() |
[21] |
J. López-Gómez, E. Muñoz-Hernández, F. Zanolin, The Poincaré-Birkhoff theorem for a class of degenerate planar Hamiltonian systems, Adv. Nonlinear Stud., 21 (2021), 489–499. http://dx.doi.org/10.1515/ans-2021-2137 doi: 10.1515/ans-2021-2137
![]() |
[22] |
S. Tang, L. Chen, The periodic predator-prey Lotka-Volterra model with impulsive effect, J. Mech. Med. Biol., 2 (2002), 267–296. http://dx.doi.org/10.1142/S021951940200040X doi: 10.1142/S021951940200040X
![]() |
[23] | D. Bainov, P. Simenov, Impulsive differential equations: Periodic solutions and applications, CRC Press, 1993. |
[24] |
Q. Liu, Q. Chen, Dynamics of stochastic delay Lotka-Volterra systems with impulsive toxicant input and Lévy noise in polluted environments, Appl. Math. Comput., 256 (2015), 52–67. http://dx.doi.org/10.1016/j.amc.2015.01.009 doi: 10.1016/j.amc.2015.01.009
![]() |
[25] |
L. Chen, J. Sun, F. Chen, L. Zhao, Extinction in a Lotka-Volterra competitive system with impulse and the effect of toxic substances, Appl. Math. Model., 40 (2016), 2015–2024. http://dx.doi.org/10.1016/j.apm.2015.09.057 doi: 10.1016/j.apm.2015.09.057
![]() |
[26] |
R. Tuladhar, F. Santamaria, I. Stamova, Fractional Lotka-Volterra-type cooperation models: Impulsive control on their stability, Entropy, 22 (2020), 970. https://doi.org/10.3390/e22090970 doi: 10.3390/e22090970
![]() |
1. | Liping Wu, Zhongyi Xiang, Dynamic analysis of a predator-prey impulse model with action threshold depending on the density of the predator and its rate of change, 2024, 9, 2473-6988, 10659, 10.3934/math.2024520 |