• Nem Talált Eredményt

fixed boundary or free boundary

N/A
N/A
Protected

Academic year: 2022

Ossza meg "fixed boundary or free boundary"

Copied!
31
0
0

Teljes szövegt

(1)

On a reaction–diffusion–advection system:

fixed boundary or free boundary

Ying Xu, Dandan Zhu and Jingli Ren

B

School of Mathematics and Statistics, Zhengzhou University, Zhengzhou 450001, China Received 15 January 2018, appeared 14 May 2018

Communicated by Maria Alessandra Ragusa

Abstract. This paper is devoted to the asymptotic behaviors of the solution to a reaction–diffusion–advection system in a homogeneous environment with fixed bound- ary or free boundary. For the fixed boundary problem, the global asymptotic stability of nonconstant semi-trivial states is obtained. It is also shown that there exists a stable nonconstant co-existence state under some appropriate conditions. Numerical simu- lations are given not only to illustrate the theoretical results, but also to exhibit the advection-induced difference between the left and right boundaries as time proceeds.

For the free boundary problem, the spreading–vanishing dichotomy is proved, i.e., the solution either spreads or vanishes finally. Besides, the criteria for spreading and van- ishing are further established.

Keywords: reaction–diffusion–advection, fixed boundary, free boundary, nonconstant steady states, spreading–vanishing.

2010 Mathematics Subject Classification: 35Q92, 35K57, 35R35, 93D20.

1 Introduction

Consider









ut =d1uxxβ1ux+u(r1−a1u−b1v), (x,t)∈ (0,L)×(0,), vt =d2vxxβ2vx+v(r2−a2v−b2u), (x,t)∈ (0,L)×(0,∞), u(x, 0) =u0(x)>0, x ∈(0,L),

v(x, 0) =v0(x)>0, x ∈(0,L).

(1.1)

Here, di,βi,ri,ai,bi are given constants, which implies in a homogeneous environment. For convenience,i=1, 2 in the whole text whenever it is mentioned.

Particularly, fordi =0 andβi =0, (1.1) is a classical ordinary differential system, and mas- sive outstanding researches have been proposed, see [1–4,14,20–22] and references therein; for diR+ and βiR, (1.1) is a reaction–diffusion–advection (RDA) system. Such RDA prob- lems were extensively used to understand the spatial behavior of populations, the dynamics of information diffusion, and so on. Up to now, many remarkable results have been achieved, see [5–10,12,13,15–17,24–27,29], etc.

BCorresponding author. Email: renjl@zzu.edu.cn

(2)

For a population growth model,u(x,t),v(x,t)in (1.1) respectively represent the densities of two species at location x and time t. di > 0 denotes the random dispersal rate of the species; advection rateβiRis the moving speed of individuals towards their more favorable habitats. As noted in [12], βi > 0 means advection points towards larger x, while βi < 0 implies advection points towards smallerx. ri > 0 accounts for intrinsic growth rate; ai > 0 is intra-specific interaction rate;biRis interspecific interaction rate. What described above means that system (1.1) is a competition model withbi >0, see [16,17,29], and a predator-prey problem forb1> 0,b2 <0, see [24,26,27,30]. For a information diffusion model, the meaning ofu(x,t),v(x,t),di,βi,ri,ai,bi can refer to [22,25].

The way to formulate boundary conditions for reaction-diffusion models is based on how the flux of individuals crosses a boundary. As mentioned in [5], the flux −→

J = −d∇u+

−→

βu across the boundary at any given point is proportional to the density with constant of proportionality, that is

−d∇u+−→ βu

· −→

n =αu, (1.2)

where d > 0 is the diffusion rate, −→

β is the advection velocity, −→

n is the outward pointing normal vector,αis the proportionality coefficient. Ifα→∞, then the boundary condition (1.2) becomesu=0, which is a Dirichlet condition; if−→

β =0 andα=0, we have(−d∇u)· −→ n =0, i.e., ∂u

n = 0, which is known as a Neumann condition; if −→

β = 0 and α < 0, then (1.2) is in the form ofd ∂u

nαu=0, which is referred to a Robin condition; ifα=0, then (1.2) becomes d∂u

n −−→

β · −→

n u=0, which is called a no-flux or reflecting boundary condition, since it means that individuals encountering the boundary are always reflected back so they do not leave the domain, that is, no individual crosses the boundary.

Here we focus on the no-flux boundary conditions

(d1ux(0,t)−β1u(0,t) =d1ux(L,t)−β1u(L,t) =0, t ∈(0,∞),

d2vx(0,t)−β2v(0,t) =d2vx(L,t)−β2v(L,t) =0, t ∈(0,∞). (1.3) In fact, Lou et al. in [16] first qualitatively analyzed (1.1) with no-flux boundary (1.3) under d1 =d2 and

r1 =r2, a1 =a2=b1 =b2 =1. (1.4) It is shown that the movement with either smaller advection or no advection is eventually stable. Afterwards, based on the assumptions of (1.4), Zhou in [29] further investigated (1.1) with the addition of (1.3) to understand the joint effects of diffusion and advection on the outcome of competition. Overcoming the mathematical difficulties arising out of d1 6= d2, Zhou in [29] obtained much richer observations: the movement with smaller diffusion, smaller advection and smaller ratio of advection to diffusion, or with larger diffusion and smaller advection, wins the competition.

However, for a more general model withdi,ri,aiR+,βi,biRand without assumption (1.4), there have been no results so far. Due to this reason, in this paper, we study (1.1) with fixed boundary (1.3) and present a thorough understanding: forbi >0, problem (1.1) and (1.3) may finally stabilize to a nonconstant semi-trivial steady state ifβi >0, but admit a stable co- existence state ifβ1·β2< 0; forb1 >0,b2 <0, the two semi-trivial steady semi-trivial steady states of (1.1) with the addition of (1.3) are both unstable. Furthermore, we give numerical simulations to find out that advection can induce great difference between the left and right boundaries as time goes on; and the problem with b1 > 0,b2 < 0 may have a co-existence state.

(3)

On the other hand, influenced by human activity, the habitat of some species often changes with time, which can be described by a free boundary. In this case, we go one step further and discuss the corresponding free boundary problem

















ut =d1uxxβ1ux+u(r1−a1u−b1v), 0<x <h(t),t >0, vt=d2vxxβ2vx+v(r2−a2v−b2u), 0<x <h(t),t >0, u(0,t) =v(0,t) =u(h(t),t) =v(h(t),t) =0, t >0,

h0(t) =−µ[ux(h(t),t) +ρvx(h(t),t)], t >0, u(x, 0) =u0(x),v(x, 0) =v0(x),h(0) =h0, 0<x <h0.

(1.5)

Here µ,ρ and h0 are given positive constants. x = h(t) is the free boundary, and the initial functionu0(x),v0(x)∈Σ(h0)for someh0>0, where

Σ(h0) ={φ∈C2([0,h0]):φ(0) =φ(h0) =0, φ(x)>0 in(0,h0)}. (1.6) Wang et al. in [26] studied system (1.5) with βi =0 andbi >0 and obtained the long time behavior of two competing species spreading via a free boundary. Based on the assumptions of βi = 0, b1 > 0, and b2 < 0, Wang in [24] investigated system (1.5) to get a spreading- vanishing dichotomy and set the criteria for spreading and vanishing, moreover, Wang in [24]

gave the estimation of asymptotic spreading speed when spreading successfully.

Motivated by the works in [24,26], we study system (1.5) withdi > 0,βi ≥0,ai >0,b1>0 andb2R. We prove that the spreading-vanishing dichotomy still holds, i.e. the solution to problem (1.5) is vanishing ifh < +∞, on the other hand, it is spreading ifh = +under some proper conditions. Furthermore, we determine the criteria for spreading and vanishing.

The rest of this paper is organized as follows. In Section 2, the fixed boundary prob- lem is analyzed, including the global asymptotic stability of nonconstant semi-trivial steady states, and the existence of a stable nonconstant co-existence state. Numerical simulations are presented to illustrate the results. Section 3 is devoted to the free boundary problem. The spreading-vanishing dichotomy is obtained and the criteria for spreading and vanishing are determined.

2 The fixed boundary problem

2.1 Existence of semi-trivial steady states First, we consider the problem





d1uxxβ1ux+u(r1−a1u) =0, x∈(0,L), d1ux(0)−β1u(0) =0,

d1ux(L)−β1u(L) =0,

(2.1)

and the following statement is valid.

Lemma 2.1. For anyβ1R, and d1, r1, a1R+, problem(2.1)admits a unique positive solutionu.˜ Proof. For β1≥0, we rewrite problem (2.1) as





d1uxxβ1ux+u(r1−a1u) =0, x∈(0,L), d1ux(0)−β1u(0) =0,

d1ux(L) +β1u(L) =2β1u(L).

(2.2)

(4)

Let

u¯ = M1 a1 e

β1 d1x

, u= M2 a1 e

β1 d1x

, where M1>r1e

|β1| d1 L

, 0< M2<r1e

−|β1| d1 L

. After some simple computations, we have





d1xxβ1x+u¯(r1−a1u¯)<0, x∈ (0,L), d1x(0)−β1u¯(0) =0,

d1x(L) +β1u¯(L) =2β1u¯(L),

(2.3)

and 





d1uxxβ1ux+u(r1−a1u)>0, x∈ (0,L), d1ux(0)−β1u(0) =0,

d1ux(L) +β1u(L) =1u(L).

(2.4) According to the definition of upper and lower solutions in [19], one can see that ¯u andu are upper and lower solutions to problem (2.2).

Set

F(u):= u(r1−a1u), x ∈(0,L), and

G(u):=

(0, x =0, 2β1u, x =L.

Foru ≤u2 <u1 ≤u, one can see that there are constants¯ K1>0,K2>0 such that F(u1)−F(u2)≥ −K1(u1−u2), x ∈(0,L),

and

G(u1)−G(u2)≥ −K2(u1−u2), x=0,L.

Thanks to Theorem 4.4.1 and Corollary 4.4.1 in [19], problem (2.2) has a positive solution ˜u∈ C2([0,L]), satisfying

u≤u˜ ≤u.¯

The proof of uniqueness is similar to [16, Lemma 2.1]. For the readers’ convenience, we outline the main ideas. Suppose ˜u1 and ˜u2 are two different positive solutions to (2.2). We may assume ˜u1> u˜2>0. ˜u1and ˜u2 satisfy





d11xxβ11x+u˜1(r1−a11) =0, x∈(0,L), d11x(0)−β11(0) =0,

d11x(L)−β11(L) =0,

(2.5)

and 





d12xxβ12x+u˜2(r1−a12) =0, x∈(0,L), d12x(0)−β12(0) =0,

d12x(L)−β12(L) =0,

(2.6)

respectively. Multiplying the first equation of (2.5) by e

β1 d1x

˜

u2 and the first equation of (2.6) bye

β1 d1x

˜

u1, subtract the resulting equations and integrate over[0,L], and then we get Z L

0 a112e

β1 d1x

(u˜2−u˜1) =0,

(5)

which contradicts ˜u1 >u˜2 >0. Therefore, we obtain the positive solution to (2.1) is unique.

Forβ1<0, we can make some minor modifications to get the existence and uniqueness of the positive solution to problem (2.1), so we omit the details.

Similarly, the problem





d2vxxβ2vx+v(r2−a2v) =0, x∈ (0,L), d2vx(0)−β2v(0) =0,

d2vx(L)−β2v(L) =0,

(2.7)

also has a unique positive solution ˜v.

Thus, the following result follows from Lemma2.1directly.

Lemma 2.2. For any βi, biRand di, ri, aiR+, i=1, 2, system(1.1)with the addition of (1.3) has two semi-trivial steady states, denoted by(u, 0˜ )and(0, ˜v)respectively.

Furthermore, as for ˜u and ˜v, we have the following result, which is vital to our later analysis.

Lemma 2.3. Suppose0<d1 <d2, aiR+, then (i) 0< u˜u˜x < βd1

1βd2β1

2d1, if0< β1 <β2and βd1

1βd2

2, x ∈(0,L);

β1

d1 < u˜u˜x <0, ifβ1<0,x∈ (0,L); (ii) 0< v˜v˜x < βd2

2βd2β1

2d1, if0<β1< β2and βd1

1βd2

2, x∈(0,L);

β2

d2 < v˜v˜x <0, ifβ2 <0,x ∈(0,L). Proof. For part (i), note that(u, 0˜ )satisfies





d1xxβ1x+u˜(r1−a1u˜) =0, x∈(0,L), d1x(0)−β1u˜(0) =0,

d1x(L)−β1u˜(L) =0.

(2.8)

Set p:= u˜u˜x. Then some straightforward computations yield

(−d1pxx+ (β1−2d1p)px+a1up˜ =0, x∈(0,L), p(0) = p(L) = βd1

1. (2.9)

By using maximum principle [18, Theorem 3.6], it is clear that (0< p< βd1

1, if β1>0, x∈(0,L);

β1

d1 < p<0, if β1<0, x∈(0,L). (2.10) According to [29, Lemma 2.4], the conditions of 0< d1 < d2, 0< β1 < β2 and βd1

1βd2

2 imply that

β1

d1β2β1

d2−d1. (2.11)

Then part (i) of Lemma2.3follows from (2.10) and (2.11).

Similarly, we can prove part (ii).

(6)

2.2 Local stability of semi-trivial steady states

In this subsection, assumedi, ri, ai,∈R+andβi,biR, we focus on the local stability of the semi-trivial steady states of problem (1.1) with the addition of (1.3). Beginning with(u, 0˜ ), and its stability is governed by the equations









ut =d1uxxβ1ux+u(r1−2a1u˜)−vb1u,˜ (x,t)∈(0,L)×(0,∞), vt =d2vxxβ2vx+v(r2−b2u˜), (x,t)∈(0,L)×(0,∞), d1ux(0,t)−β1u(0,t) =d1ux(L,t)−β1u(L,t) =0, t>0,

d2vx(0,t)−β2v(0,t) =d2vx(L,t)−β2v(L,t) =0, t>0.

(2.12)

The corresponding eigenvalue problem is









d1Φxxβ1Φx+Φ(r1−2a1u˜)−ωb2u˜+λΦ=0, x∈ (0,L), d2ωxxβ2ωx+ω(r2−b2u˜) +λω=0, x∈ (0,L), d1Φx(0)−β1Φ(0) =d1Φx(L)−β1Φ(L) =0,

d2ωx(0)−β2ω(0) =d2ωx(L)−β2ω(L) =0.

(2.13)

One can find that the second equation in (2.13) is decoupled from the first. As a result, we only consider the eigenvalue problem





d2ωxxβ2ωx+ω(r2−b2u˜) +λω =0, x ∈(0,L), d2ωx(0)−β2ω(0) =0,

d2ωx(L)−β2ω(L) =0.

(2.14)

Similarly, in order to investigate the stability of(0, ˜v), we consider the eigenvalue problem





d1ϕxxβ1ϕx+ϕ(r1−b1v˜) +µϕ=0, x∈ (0,L), d1ϕx(0)−β1ϕ(0) =0,

d1ϕx(L)−β1ϕ(L) =0.

(2.15)

For convenience, the general formula of eigenvalue problems (2.14) and (2.15) is given as follows





xxγδx+δm(x) +σδ =0, x∈(0,L), dδx(0)−γδ(0) =0,

x(L)−γδ(L) =0.

(2.16)

Then we have the following statements.

Lemma 2.4. Suppose d ∈ R+, γR, m(x)∈ C1([0,L]), then the eigenvalue problem(2.16) has a simple principle eigenvalueσ0, and the corresponding eigenfunctionδ0(x)can be chosen asδ0(x)0.

Proof. Consider





Lu=duxxγux+um(x), x∈(0,L), dux(0)−γu(0) =0,

dux(L)−γu(L) =0,

(2.17)

(7)

whered ∈R+,γR,m(x)∈C1([0,L]). Letw=eγdxu, then (2.17) becomes (Lu=eγdx[dwxx+γwx+wm(x)], x ∈(0,L),

wx(0) =wx(L) =0. (2.18)

Denote

Lw=dwxx+γwx+wm(x), x∈(0,L), (2.19) and then

Z L

0

wLudx=

Z L

0

uLwdx. (2.20)

So L can be seen the adjoint operator of L.

By [23, Theorem 7.6.1], problem

(Lδ+σδ=0, x∈(0,L),

δx(0) =δx(L) =0, (2.21)

has a simple principle eigenvalueσ0, and the corresponding eigenfunctionδ0(x)can be chosen asδ0(x)0.

Thanks to [5, Corollary 2.13], we get thatσ0 is also the principle eigenvalue of





Lδ+σδ =0, x∈ (0,L), dδx(0)−γδ(0) =0,

x(L)−γδ(L) =0.

(2.22)

Accordingly, Lemma2.4is established.

Let λ0 (resp. µ0) be the principal eigenvalues of (2.14) (resp. (2.15)), and ω0(x) (resp.

ϕ0(x)) be the corresponding eigenfunction satisfying ω0(x)(resp. ϕ0(x)) 0. By Lemma 2.4,(λ0,ω0(x))and(µ0,ϕ0(x))must exist, moreover,λ0 andµ0 are simple.

For simplicity, in the following, we denoteω0(x), ϕ0(x),δ0(x), ∂ω∂x0(x), ∂ϕ∂x0(x) and ∂δ∂x0(x) by ω0,ϕ0,δ0,ω0x,ϕ0xandδ0x respectively.

Lemma 2.5. Suppose d∈(R+),γR, m(x)∈C1([0,L]), then we have (i) δδ0x

0 < γd in(0,L), if mx ≤,6≡0in[0,L]; (ii) δδ0x

0 > γd in(0,L), if mx ≥,6≡0in[0,L]. Proof. Leth:= δ0x

δ0 , thenh(0) =h(L) = γd. Taking derivative ofh, we derive hx = δ0xx

δ0 −h2. (2.23)

Note that(σ0,δ0)satisfies (2.16), thanks to (2.23), then some direct computations yield

−dhx−dh2+γh−m(x) =σ0. (2.24) Taking derivative of (2.24) in view ofx, we obtain

(−dhxx+ (γ−2dh)hx =mx, x∈ (0,L),

h(0) =h(L) = γd. (2.25)

It follows from the maximum principle that h < γ/d if mx ≤,6≡ 0; h > γ/d if mx ≥,6≡0 in [0,L].

(8)

As for the stability of semi-trivial steady states of system (1.1) with the addition of (1.3), the following result in [23] is of great concern in our subsequent analysis.

Lemma 2.6. Suppose di, ri, aiR+, andβi,biR, then the semi-trivial steady state(u, 0˜ )is linearly stable (resp. unstable) ifλ0 is positive (resp. negative); the semi-trivial steady state (0, ˜v)is linearly stable (resp. unstable) ifµ0is positive (resp. negative).

For clarity, we first state two propositions, which are vital to judge the stability of (u, 0˜ ) and(0, ˜v).

Proposition 2.7. λ0>0(resp.λ0<0) if and only if

λ(d1,d2,β1,β2,r1,r2,a1,b2)>0 (resp.λ(d1,d2,β1,β2,r1,r2,a1,b2)<0), where

λ(d1,d2,β1,β2,r1,r2,a1,b2) =

Z L

0 e

β2 d2x

˜

0[(r1−r2) + (b2−a1)u˜]dx +

Z L

0 e

β2 d2x

ω0xβ2 d2ω0

[(d2−d1)u˜x+ (β1β2)u˜]dx.

(2.26)

Proof. Rewrite (2.8) as





d2xxβ2x+u˜(r1−a1u˜) = (d2−d1)u˜xx+ (β1β2)u˜x, x ∈(0,L), d2x(0)−β2u˜(0) = (d2−d1)u˜x(0) + (β1β2)u˜(0),

d2x(L)−β2u˜(L) = (d2−d1)u˜x(L) + (β1β2)u˜(L).

(2.27)

Note that(λ0,w0)satisfies





d2ω0xxβ2ω0x+ω0(r2−b2u˜) =−λ0ω0, x∈(0,L), d2ω0x(0)−β2ω0(0) =0,

d2ω0x(L)−β2ω0(L) =0.

(2.28)

Multiplying the first equation of (2.27) bye

β2 d2x

ω0, and integrating the resulting equation over [0,L], we get

(d2xβ2u˜)e

β2 d2x

ω0

L 0

Z L

0

(d2xβ2u˜)e

β2 d2x

ω0xβ2 d2ω0

dx +

Z L

0

e

β2 d2x

ω0u˜(r1−a1u˜)dx

= [(d2−d1)u˜x+ (β1β2)u˜]e

β2 d2x

ω0

L 0

Z L

0

[(d2−d1)u˜x+ (β1β2)u˜]e

β2 d2x

ω0xβ2 d2ω0

dx.

(2.29)

By the boundary conditions of (2.27), (2.29) can be simplified to Z L

0 e

β2 d2x

ω0u˜(r1−a1u˜)dx+

Z L

0

[(d2−d1)u˜x+ (β1β2)u˜]e

β2 d2x

ω0xβ2 d2ω0

dx

=

Z L

0

(d2xβ2u˜)e

β2 d2x

ω0xβ2 d2ω0

dx.

(2.30)

(9)

Multiplying the first equation of (2.28) bye

β2 d2x

˜

u, and integrating the resulting equation over [0,L], according to boundary conditions, we have

λ0 Z L

0 uω˜ 0e

β2 d2x

dx

=

Z L

0

(d2ω0xβ2ω0)e

β2 d2x

˜ uxβ2

d2

dx−

Z L

0

(r2−b2u˜)ω0e

β2 d2x

˜ udx.

(2.31)

Combining (2.30) and (2.31), one can find λ0

Z L

0 uω˜ 0e

β2 d2x

dx=

Z L

0 e

β2 d2x

˜

0[(r1−r2) + (b2−a1)u˜]dx +

Z L

0 e

β2 d2x

ω0xβ2 d2ω0

[(d2−d1)u˜x+ (β1β2)u˜]dx.

(2.32)

By the definition ofλ(d1,d2,β1,β2,r1,r2,a1,b2), (2.32) can be rewritten as λ0

Z L

0

uω˜ 0e

β2 d2x

dx= λ(d1,d2,β1,β2,r1,r2,a1,b2). (2.33) Clearly, the sign ofλ0 is the same as that ofλ, which completes the proof of Proposition2.7.

Similarly, we have the following proposition concerningµ0. Proposition 2.8. µ0 >0(resp.µ0<0) if and only if

µ(d1,d2,β1,β2,r1,r2,b1,a2)>0 (resp. µ(d1,d2,β1,β2,r1,r2,b1,a2)<0), where

µ(d1,d2,β1,β2,r1,r2,b1,a2)

=

Z L

0

e

β1 d1x

˜

vϕ0[(r2−r1) + (b1−a2)v˜]dx +

Z L

0 e

β1 d1x

ϕ0xβ1 d1ϕ0

[(d1−d2)v˜x+ (β2β1)v˜]dx.

(2.34)

Proof. By a similar method noted in the proof of Proposition2.7, we can find µ0

Z L

0

ϕ0e

β1 d1x

dx=

Z L

0 e

β1 d1x

ϕ0[(r2−r1) + (b1−a2)v˜]dx +

Z L

0 e

β1 d1x

ϕ0xβ1 d1ϕ0

[(d1−d2)v˜x+ (β2β1)v˜]dx.

(2.35)

According to the definition ofµ(d1,d2,β1,β2,r1,r2,b1,a2), (2.35) can be simplified as µ0

Z L

0ϕ0e

β1 d1x

dx=µ(d1,d2,β1,β2,r1,r2,b1,a2), (2.36) which indicates thatµ0 andµ have the same sign.

Now, we can establish the local stability of(u, 0˜ )and(0, ˜v)respectively.

(10)

Lemma 2.9. Assume di,ri,a1R+,βi,b2R, then (i) if d1 <d2,0<β1< β2, βd1

1βd2

2, r2≤r1and0< a1 ≤b2,(u, 0˜ )is stable locally;

(ii) if d1 <d2,0<β2< β1, r1 ≤r2and0≤ b2 ≤a1,(u, 0˜ )is unstable;

(iii) if d1 ≤d2,β1·β2<0, r1≤r2and0≤b2≤ a1,(u, 0˜ )is unstable;

(iv) if d1 <d2,0<β1< β2, βd1

1βd2

2, r1≤r2and b2<0,(u, 0˜ )is unstable.

Proof. For part (i), according to part (i) of Lemma2.3, 0< u˜x

˜ u < β1

d1β2β1

d2−d1. (2.37)

Here, 0< u˜u˜x < βd1

1 implies ˜ux >0, which indicates that

mx = [r2−b2u˜]x =−b2x <0, and then

ω0x ω0

< β2

d2, (2.38)

from Lemma2.5.

Moreover, it directly follows that from 0<r2 ≤r1 and 0<a1≤ b2

(r1−r2) + (b2−a1)u˜ ≥0. (2.39) Therefore, thanks to (2.37), (2.38) and (2.39), we get λ(d1,d2,β1,β2,r1,r2,a1,b2) > 0 if 0 <

d1 < d2, 0< β1 < β2, βd1

1βd2

2, 0 < r2 ≤ r1, 0< a1 ≤ b2. By Lemma2.6 and Proposition2.7, the proof of part (i) is finished.

Actually,λ(d1,d2,β1,β2,r1,r2,a1,b2)<0 directly follows from the conditions of 0<d1<

d2, 0<β2< β1, 0<r1 ≤r2 and 0< b2 ≤a1. Thus, we can get(u, 0˜ )is unstable.

For part (iii), the case of d1 = d2 > 0 is easy to check, so we only verify the case of 0<d1 <d2.

β1·β2 < 0 implies β1 > 0 > β2 or β2 > 0 > β1. We first consider β1 > 0 > β2. By the similar argument to part (i), it is easy to find

((d2−d1)u˜x+ (β1β2)u˜ >0, ω0xβd2

2ω0 <0. (2.40)

Combining (2.40) with the conditions of 0<r1 <r2 and 0<b2< a1, we deduce λ(d1,d2,β1,β2,r1,r2,a1,b2)<0,

which shows that(u, 0˜ )is unstable by Proposition2.7and Lemma2.6.

For β2 > 0 > β1, combining this condition with part (i) of Lemma 2.3 and part (ii) of Lemma2.5, we have

((d2−d1)u˜x+ (β1β2)u˜ <0, ω0xβd2

2ω0 >0, (2.41)

which yields

λ(d1,d2,β1,β2,r1,r2,a1,b2)<0,

(11)

if 0<r1≤r2and 0<b2≤a1. Therefore,(u, 0˜ )is unstable.

For part (iv), in the case ofb2 <0, again by part (i) of Lemma2.3and part (ii) of Lemma2.5, we get

(0< u˜u˜x < βd2β1

2d1, x ∈(0,L), 0< βd2

2 < ww0x

0, x ∈(0,L). (2.42)

Since 0< r1 ≤r2andb2< 0<a1, we deriveλ(d1,d2,β1,β2,r1,r2,a1,b2)<0. Then the proof of part (iii) is completed by Proposition2.7and Lemma2.6.

Lemma 2.10. Suppose that di,ri,a2,b1R+,βiR, then (i) if d1< d2,0< β1 <β2, βd1

1βd2

2, r2 ≤r1and b1≤ a2,(0, ˜v)is unstable;

(ii) if d1< d2,0< β2 <β1, r1≤r2and a2 ≤b1,(0, ˜v)is stable locally;

(iii) if d1≤ d2,β1·β2<0, r2 ≤r1and b1≤ a2,(0, ˜v)is unstable.

Proof. By using the similar argument to the one applied in the proof of Lemma2.9, one can prove that µ(d1,d2,β1,β2,r1,r2,b1,a2)< 0 ford1 <d2, 0 <β1 <β2, βd1

1βd2

2,r2 ≤r1,b1 ≤a2; µ(d1,d2,β1,β2,r1,r2,b1,a2) > 0 for d1 < d2, 0< β2 < β1, r1 ≤ r2 and a2 ≤ b1. Thus, part (i) and part (ii) directly follows from Proposition2.7 and Lemma2.6.

Part (iii), we only consider d1 < d2, because the case of d1 = d2 > 0 can be verified in a similar argument.

For β1>0> β2, it follows from part (ii) of Lemma2.3that 0<−v˜x

˜

v <−β2

d2 < β1β2

d2−d1, that is,

(d1−d2)v˜x+ (β2β1)v˜<0. (2.43) Moreover, due to part (i) of Lemma2.5, we deduce

ϕ0x

ϕ0 > β1

d1, (2.44)

since mx = (r1−b1v˜)x = −b1x > 0 if β2 < 0. So (0, ˜v)is unstable under the conditions of r2≤r1andb1≤ a2.

For β2>0> β1, we can use a similar argument to show(0, ˜v)is unstable.

2.3 The non-existence of coexistence steady state

In this subsection, we show the nonexistence of coexistence steady states under some proper conditions.

Lemma 2.11. Suppose that0<d1 <d2, then (i) if0 < β1 < β2, βd1

1βd2

2, 0 < r2 ≤ r1,0 < a1 ≤ b2 and0 < b1 ≤ a2, system (1.1) with the addition of (1.3)has no coexistence steady state;

(ii) if 0 < β2 < β1,0 < r1 ≤ r2, 0 < b2 ≤ a1 and0 < a2 ≤ b1, system(1.1) with the addition of (1.3)has no coexistence steady state.

(12)

Proof. For part (i), arguing indirectly, we assume that system (1.1) with the addition of (1.3) has a coexistence steady state(U,V), then









d1Uxxβ1Ux+U(r1−a1U−b1V) =0, x ∈(0,L), d2Vxxβ2Vx+V(r2−a2V−b2U) =0, x ∈(0,L), d1Ux(0)−β1U(0) =d1Ux(L)−β1U(L) =0,

d2Vx(0)−β2V(0) =d2Vx(L)−β2V(L) =0.

(2.45)

Define

f(x):=d1Uxβ1U, x∈[0,L]; (2.46) and

g(x):=d2Vxβ2V, x ∈[0,L]. (2.47) By the boundary conditions of (2.45), we have

f(0) = f(L) = g(0) =g(L) =0. (2.48) In the following, we show 8 claims to finish the proof.

Claim 1.

(i) If f0(x)≥0 (resp.>0),g0(x)≥0 (resp.>0);

(ii) Ifg0(x)≤0 (resp. <0), f0(x)≤0 (resp.<0).

Combining (2.45) with the definition of f(x)andg(x), we find f0(x) =d1Uxxβ1Ux =U(a1U+b1V−r1), and

g0(x) =d2Vxxβ2Vx =V(a2V+b2U−r2). It follows that from f0(x)≥0 (resp.>0)

a1U+b1V−r1 ≥0 (resp. >0). Therefore, 0<r2 ≤r1, 0<a1≤b2and 0<b1≤ a2directly yield

a2V+b2U−r2≥a1U+b1V−r1≥0 (resp. >0). Hence,g0(x)≥0 (resp.>0).

Part (ii) can be proved similarly.

Claim 2.

(i) There is smallε >0 such that f(x)<0 in(0,ε]; (ii) There is smallδ >0 such thatg(x)<0 in[L−δ,L).

Part (i), if not, there exists some small ε0 such that f(x)> 0 or f(x) ≡ 0 in(0,ε0]. Next, we show contradictions respectively. Define

T:= Ux

U, x ∈[0,L]; (2.49)

(13)

and

S:= Vx

V, x∈ [0,L]. (2.50)

Due to (2.45), some straightforward calculations yield









−d1Txx+ (β1−2d1T)Tx+a1TU+b1SV =0, x∈ (0,L),

−d2Sxx+ (β2−2d2S)Sx+a2SV+b2TU=0, x∈ (0,L), T(0) =T(L) = βd1

1 >0, S(0) =S(L) = βd2

2 >0.

(2.51)

When f(x)>0 in (0,ε0], there must beε1 ∈(0,ε0]such that f0(x)>0, x∈ (0,ε1]. According to Claim 1, we have

g0(x)>0, x∈(0,ε1]. Due to g(0) =0, we get

g(x)>0, x∈(0,ε1].

Denote the first zero point of f(x)in(0,L]byy1, and the first zero ofg(x)byx1. For brevity, we let y1 ≤x1. Actually, the following expression is similar wheny1≥x1.

Thus, we have

f(0) = f(y1) =0, f(x)>0, x∈(0,y1); (2.52) and

g(0) =0, g(y1)≥0, g(x)>0, x∈(0,y1). (2.53) Thanks to (2.46), (2.47), (2.49), (2.50), (2.52) and (2.53), we find

T(0) =T(y1) = β1

d1, T(x)> β1

d1 >0, x∈(0,y1); (2.54) and

S(0) = β1

d2, S(y1)≥ β2

d2 >0, S(x)> β2

d2 >0, x∈(0,y1). (2.55) From (2.54), T must attain a positive local maximum at some point denoted by z1, z1 ∈ (0,y1). By the first equation of (2.51), we getS(z1)<0, which contradicts (2.55). Consequently, the statement f(x)>0 in (0,ε0]does not hold.

Actually, if f(x) ≡ 0 in (0,ε0], f0(x) ≡ 0. By Claim 1, g0(x) ≥ 0, together with g(0) = 0, there must exist ε2∈(0,ε0]such thatg(x)≥0,x∈ (0,ε2)], deducing that

S≥ β2 d2

, x ∈(0,ε2].

On the other hand, by (2.46) and (2.49), f(x)≡ 0 in(0,ε0]guarantees T ≡ βd1

1. According to the first equation of (2.51), we deriveS<0 in(0,ε0), a contradiction toS≥ βd2

2 in(0,ε2). Consequently, part (i) of Claim 2 is set up. Part (ii) can be verified by the similar method, so we omit the details.

Claim 2 implies that f andg can not be identically zero in[0,L]. Besides, f andgare real analytic. Therefore, all zero points of f andgare isolated.

Ábra

Figure 2.1: The asymptotic behaviors of u ( x, t ) in ( a ) and v ( x, t ) in ( b ) with β 1 = 0.2, β 2 = 1, r 1 = 1, r 2 = 0.5, a 1 = 1.1, a 2 = 2.1, b 1 = 1.2, b 2 = 2.
Example 2.19. Figure 2.1 (resp. Figure 2.2) describes part (i) (resp. (ii)) of Theorem 2.12, and Figure 2.3 displays Theorem 2.15
Figure 2.5: The illustration of u ( x, t ) in ( a ) and v ( x, t ) in ( b ) with β 1 = 0.2, β 2 = 1, r 1 = r 2 = 1, a 1 = 1.2, a 2 = 2.1, b 1 = 1.1, b 2 = − 2.

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

Consequently, we assume that the underlying floating exchange rate will be fixed at the same time as the target zone system will be replaced by the fixed system and the

We establish the existence of at least two positive solutions for the 3 n th order three-point boundary value problem on time scales by using Avery-Henderson fixed point theorem..

The main contributions are as follows: (a) we present problems with linear boundary value conditions, and on this basis we obtain the existence of the extremal solutions for

Williams fixed point theorem, commonly referred to as an Avery type fixed point theo- rem, to a second order boundary value problem with antiperiodic boundary conditions.. Under

Furthermore, we demonstrate by numerical simulations that typical Turing patterns such as spotted and spot- stripelike mixtures patterns, a stable limit cycle due to Hopf

Mart´ınez-Aparicio, Elliptic obstacle problems with natural growth on the gradiant and singular nonlinear terms, Adv.. Mart´ınez-Aparicio, Quasilinear equations with natural

Insisting on the question of the relationship between Egypt and Israel, Carter was intrested in the conditions of the diplomatic connection and the problem of the boundary between

We will deal with a first order func- tional boundary value problem on an infinite interval, seeking positive non-zero solutions.. In Section 1 we state the boundary value problem