In this paper, we considered the Neumann boundary value problem for the strongly-coupled subelliptic system and the predator-prey subelliptic system on the Heisenberg group. We provide a priori estimates and the non-existence result for non-constant positive solutions for the strongly-coupled and predator-prey systems.
Citation: Xinjing Wang, Guangwei Du. Strongly-coupled and predator-prey subelliptic system on the Heisenberg group[J]. AIMS Mathematics, 2024, 9(10): 29529-29555. doi: 10.3934/math.20241430
[1] | Jin Liao, André Zegeling, Wentao Huang . The uniqueness of limit cycles in a predator-prey system with Ivlev-type group defense. AIMS Mathematics, 2024, 9(12): 33610-33631. doi: 10.3934/math.20241604 |
[2] | Shanshan Yu, Jiang Liu, Xiaojie Lin . Multiple positive periodic solutions of a Gause-type predator-prey model with Allee effect and functional responses. AIMS Mathematics, 2020, 5(6): 6135-6148. doi: 10.3934/math.2020394 |
[3] | Weili Kong, Yuanfu Shao . The effects of fear and delay on a predator-prey model with Crowley-Martin functional response and stage structure for predator. AIMS Mathematics, 2023, 8(12): 29260-29289. doi: 10.3934/math.20231498 |
[4] | Xuyang Cao, Qinglong Wang, Jie Liu . Hopf bifurcation in a predator-prey model under fuzzy parameters involving prey refuge and fear effects. AIMS Mathematics, 2024, 9(9): 23945-23970. doi: 10.3934/math.20241164 |
[5] | Binfeng Xie, Na Zhang . Influence of fear effect on a Holling type III prey-predator system with the prey refuge. AIMS Mathematics, 2022, 7(2): 1811-1830. doi: 10.3934/math.2022104 |
[6] | Wei Li, Qingkai Xu, Xingjian Wang, Chunrui Zhang . Dynamics analysis of spatiotemporal discrete predator-prey model based on coupled map lattices. AIMS Mathematics, 2025, 10(1): 1248-1299. doi: 10.3934/math.2025059 |
[7] | Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra . Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045 |
[8] | Yadigar Sekerci . Adaptation of species as response to climate change: Predator-prey mathematical model. AIMS Mathematics, 2020, 5(4): 3875-3898. doi: 10.3934/math.2020251 |
[9] | Yalong Xue . Analysis of a prey-predator system incorporating the additive Allee effect and intraspecific cooperation. AIMS Mathematics, 2024, 9(1): 1273-1290. doi: 10.3934/math.2024063 |
[10] | Lingling Li, Xuechen Li . The spatiotemporal dynamics of a diffusive predator-prey model with double Allee effect. AIMS Mathematics, 2024, 9(10): 26902-26915. doi: 10.3934/math.20241309 |
In this paper, we considered the Neumann boundary value problem for the strongly-coupled subelliptic system and the predator-prey subelliptic system on the Heisenberg group. We provide a priori estimates and the non-existence result for non-constant positive solutions for the strongly-coupled and predator-prey systems.
The spatial distribution pattern of an animal population in its natural environment may be the result of several biological effects. In a patchy environment, linear diffusional flows have a stabilizing effect on the coexistence of competitive species. Shigesada, Kawasaki, and Teramoto [22] studied the spatial segregation of interacting species and proposed the model
{∂u1∂t=Δ[(d1+α11u1+α12u2)u1]+u1(a1−b1u1−c1u2),in ΩT,∂u2∂t=Δ[(d2+α21u1+α22u2)u2]+u2(a2−b2u1−c2u2),in ΩT,∂u1∂ν=∂u2∂ν=0,on ∂ΩT,u1(x,0)=u1,0(x),u2(x,0)=u2,0(x),in Ω, |
where u1 and u2 represent the densities of two competing species, d1 and d2 are their diffusion rates, a1 and a2 denote the intrinsic growth rates, b1 and b2 account for intra-specific competitions, c1 and c2 are the coefficients of inter-specific competitions, α11 and α22 are usually referred as self-diffusion pressures, and α12 and α21 are cross-diffusion pressures. Here, Δ is the Laplace operator, Ω is a bounded smooth domain of RN with N≥1, ∂Ω and ¯Ω are the boundary and the closure of Ω, respectively, ΩT=Ω×[0,T) and ∂ΩT=∂Ω×[0,T) for some T∈(0,∞], ν is the outward unit normal vector on ∂Ω, di,ai,bi,ci(i=1,2) are all positive constants, and αij(i,j=1,2) denote non-negative constants. The initial values u1,0 and u2,0 are non-negative smooth functions that are not identically zero. For more details on the backgrounds of this model, we refer to [21,22]; for reaction diffusion, see [7,15].
Lou and Ni [16] considered positive steady-state solutions to the above strongly-coupled parabolic system and derived properties of these solutions, including a priori estimates, as well as conditions for existence and non-existence. To prove those results, they first considered the strongly-coupled elliptic system
{Δ[(d1+α11u1+α12u2)u1]+u1(a1−b1u1−c1u2)=0,in Ω,Δ[(d2+α21u1+α22u2)u2]+u2(a2−b2u1−c2u2)=0,in Ω,∂u1∂ν=∂u2∂ν=0,on ∂Ω,u1>0,u2>0,in Ω. |
For N=1,α11=α21=α22=0, Mimura and Kawasaki [19] demonstrated the existence of small amplitude solutions bifurcating from the trivial solution. Mimura [18] established that large amplitude solutions exist when α12 is suitably large. Mimura, Nishiura, Tesei, and Tsujikawa [20] proved the existence of non-constant solutions of this problem. Jia and Xue [14] investigated the non-existence of non-constant positive steady states in a generalized predator-prey system. Xue, Jia, Ren, and Li [28] proved both the existence and non-existence of non-constant positive stationary solutions for the general Gause-type predator-prey system with constant self-diffusion and cross-diffusion. For more information on the parabolic system, we refer to [21,27,29].
In this paper, we study the strongly-coupled subelliptic system on the Heisenberg group
{ΔH[(d1+α11u+α12v)u]+u(a1−b1u−c1v)=0,in Ω,ΔH[(d2+α21u+α22v)v]+v(a2−b2u−c2v)=0,in Ω,∂u∂ν=∂v∂ν=0,on ∂Ω,u>0,v>0,in Ω, | (1.1) |
where ΔH is the degenerate subelliptic (also called hypoelliptic in [12]) operator. Here, di,ai,bi,ci(i=1,2) are positive constants, and αij(i,j=1,2) are non-negative constants. For the degenerate of the ΔH, there are some different forms [14,16,28]; see Section 2 for further details.
Only one of the diffusion rates or one of the self-diffusion pressures needs to be large to prevent the formation of a non-constant solution to (1.1).
Theorem 1.1. Suppose that a1a2≠b1b2 and a1a2≠c1c2.
(i) There exists a positive constant C1=C1(di,ai,bi,ci,α12,α21) such that problem (1.1) has no non-constant solution if max{α11,α22}≥C1.
(ii) There exists a positive constant C2=C2(ai,bi,ci,αij) such that if max{d1,d2}≥C2, then problem (1.1) has no non-constant solution provided that both α11 and α22 are positive.
In the case of weak competition, if self-diffusion is weaker than diffusion, then (1.1) still has no non-constant solution.
To obtain some non-existence results from Theorem 1.1, we mainly study the effects of diffusion and self-diffusion in the strongly-coupled subelliptic system
{ΔH[(d1+α11u+α12v)u]+uf(u,v)=0,in Ω,ΔH[(d2+α21u+α22v)v]+vg(u,v)=0,in Ω,∂u∂ν=∂v∂ν=0,on ∂Ω,u>0,v>0,in Ω. | (1.2) |
For the sake of convenience, we collect here all the assumptions on f,g, some of which will be made at different times in this paper. Throughout this paper, we always follow the following hypotheses:
(H1) f(0,0)=a1,g(0,0)=a2,∂f∂u≤−b1,∂g∂u≤−b2,∂f∂v≤−c1,∂g∂v≤−c2, for all u≥0,v≥0, where ai,bi and ci are all positive constants for i=1,2.
(H1') f(0,0)=a1,g(0,0)=a2,∂f∂u≤−b1,∂g∂u≤0,∂f∂v≤0,∂g∂v≤−c2, for all u≥0,v≥0, where a1,a2,b1 and c2 are all positive constants.
(H2) Both {u>0|f(u,0)=g(u,0)=0} and {v>0|f(0,v)=g(0,v)=0} are empty.
(H3) f(u,v)=g(u,v)=0 has a unique positive root (u∗,v∗).
It is easy to see that (H1) is more restrictive than (H1′). From (H1′), it follows that if (u,v) is a positive root of f(u,v)=g(u,v)=0, then u≤a1b1 and v≤a2c2. For the special case f=u(a1−b1u−c1v) and g=v(a2−b2u−c2v), it is trivial to check that (H1) and (H1′) hold, and that (H2) is equivalent to a1a2≠b1b2 and a1a2≠c1c2, while (H3) is satisfied only in b1b2>a1a2>c1c2 and b1b2<a1a2<c1c2.
For the generalized predator-prey subelliptic system with cross-diffusion and homogeneous Neumann boundary conditions, we investigate the existence and non-existence of non-constant positive solutions to the following subelliptic system
{d1ΔH[(1+ˉα12v)u]+uq(u)−p(u)v=0,in Ω,d2ΔH[(1+ˉα21u)v]+v(−a2+c2p(u))=0,in Ω,∂u∂ν=∂v∂ν=0,on ∂Ω,u≥0,v≥0,in Ω. | (1.3) |
The functions q(u)∈C1([0,+∞)) and p(u)∈C1([0,+∞))⋂C2((0,+∞)) are assumed to satisfy the following two hypotheses throughout this paper:
(H4) q(0)>0,q′(u)<0, for all u≥0. And there exists a unique positive constant S, such that q(S)=0.
(H5) p(0)=0,limu→0+p′(u)<∞,cp(S)>a2 and p′(u)>0 for all 0<u≤S.
For the case ˉα21=0 of the subelliptic system (1.3), we have the following theorem.
Theorem 1.2. There are positive constants ˜d1,˜d2,˜α12 such that if d1≥˜d1,d2≥˜d2 and ˉα12≤˜α12, then problem (1.3) has no non-constant solution.
In general, for the subelliptic system (1.3), we obtain the following theorem.
Theorem 1.3. There are positive constants ˜d1,˜d2,˜α12,˜α21 such that if d1>˜d1,d2>˜d2 and ˉα12<˜α12,ˉα21<˜α21, then problem (1.3) has no non-constant solution.
The paper is organized as follows. In Section 2, we collect some well-known facts about Hn and the subelliptic operator ΔH. Section 3 gives an overview of competition-diffusion in the strongly-coupled model. Section 4 is devoted to studying diffusion and self-diffusion in the strongly-coupled model. In Section 5, we derive the predator-prey model.
In this section, we list some facts related to the Heisenberg group and sub-Laplacian ΔH. For proofs and more information, we refer, for example, to [3,4,8,9,12].
The Heisenberg group Hn is the Euclidean space R2n+1(n≥1) endowed with the group action ∘ defined by
ξ0∘ξ=(x+x0,y+y0,t+t0+2n∑i=1(xiyi0−yixi0)), | (2.1) |
where ξ=(x1,⋯,xn,y1,⋯,yn,t):=(x,y,t)∈Rn×Rn×R, ξ0=(x0,y0,t0). Let us denote by δλ the dilations on R2n+1, i.e.,
δλ(ξ)=(λx,λy,λ2t) | (2.2) |
which satisfies δλ(ξ0∘ξ)=δλ(ξ0)∘δλ(ξ).
The left invariant vector fields corresponding to Hn are of the form
Xi=∂∂xi+2yi∂∂t,i=1,⋯,n,Yi=∂∂yi−2xi∂∂t,i=1,⋯,n,T=∂∂t. |
The Heisenberg gradient of a function u is defined as
∇Hu=(X1u,⋯,Xnu,Y1u,⋯,Ynu). | (2.3) |
The sub-Laplacian ΔH on Hn is
ΔH=n∑i=1Xi2+Yi2, | (2.4) |
with the expansion
ΔH=n∑i=1∂2∂xi2+∂2∂yi2+4yi∂2∂xi∂t−4xi∂2∂yi∂t+4(xi2+yi2)∂2∂t2. |
It is easy to check that
[Xi,Yj]=−4Tδij,[Xi,Xj]=[Yi,Yj]=0,i,j=1,⋯,n |
and {X1,⋯,Xn,Y1,⋯Yn} satisfies the Hörmander's rank condition (see [12]). In particular, this implies that ΔH is hypoelliptic (see [12]), and the solution of equation including ΔH satisfies the maximum principle (see [2,4]).
Denote by Q=2n+2 the homogeneous dimension of Hn. The norm |ξ|H is the distance of ξ to the origin (see [8]),
ρ=|ξ|H=(n∑i=1(xi2+yi2)2+t2)14. | (2.5) |
Using this norm, one can define the distance between two points in Hn in the natural way
dH(ξ,η)=|η−1∘ξ|H, |
where η−1 denotes the inverse of η with respect to the group action ∘, i.e., η−1=−η.
The open ball of radius R>0 centered at ξ0 is the set
BH(ξ0,R)={η∈Hn|dH(η,ξ0)<R}. |
By the dilation of the group, ξ→|ξ|H is homogeneous of degree one with respect to δλ and
|BH(ξ0,R)|=|BH(0,R)|=|BH(0,1)|RQ, |
where |⋅| denotes the Lebesgue measure. Noting that Xi and Yi are homogeneous of degree minus one with respect to δλ, i.e.,
Xi(δλ)=λδλ(Xi),Yi(δλ)=λδλ(Yi), |
then ΔH is homogeneous of degree minus two and left invariant.
We denote the Sobolev space by
‖u‖Lp(Ω)=(∫Ω|u(ξ)|pdξ)1p,1≤p<∞,‖u‖L∞(Ω)=esssupξ∈Ω|u(ξ)|. |
and
W1,2(Ω)={u|u,∇Hu∈L2(Ω)}, |
which is a Banach space about the norm
‖u‖W1,2(Ω)=‖u‖L2(Ω)+‖∇Hu‖L2(Ω). |
Denote by W1,20(Ω) the closure of C∞0(Ω) in W1,2(Ω).
Let us state Sobolev's and Poincaré's inequalities in Hn, see [10,13,17].
Lemma 2.1. Let U be a bounded domain in Hn and Ω⊂⊂U. If 1<p<Q and u∈W1,p0(Ω), then there exists C>0 depending on n,p and Ω, such that for any 1≤q≤pQQ−p,
(∫Ω|u|q)1q≤C(∫Ω|∇Hu|p)1p. | (2.6) |
If 1≤p<∞ and u∈W1,p0(Ω), then
∫Ω|u|p≤C∫Ω|∇Hu|p. | (2.7) |
For the maximum principle, we refer to [2,4].
Lemma 2.2. Let Ω be a bounded domain and K(ξ)>0, u satisfies
−ΔHu+K(ξ)u≥0onΩ,u=0in∂Ω, |
then u≥0 on Ω. Furthermore, u>0 on Ω, unless u≡0.
The following Hopf-type lemma is from [4,6,26].
Lemma 2.3. For a domain V in ˆHn:=Hn×R⊂Cn+1, let the point P0∈∂V satisfy the interior Heisenberg ball condition (see [6]). Assume that U∈C2(V)∩C1(¯V) is a solution to
−LαU≥c1(z)U, |
for c1(z)∈L∞(V), where
Lα=∂2∂λ2+1−αλ∂∂λ+n∑i=1(∂2∂xi2+∂2∂yi2+4yi∂2∂xi∂t−4xi∂2∂yi∂t)+4(λ2+n∑i=1(xi2+yi2))∂2∂t2. |
If U(z)>U(P0)=0, z∈V, then
∂U∂ν(P0)>0, |
where ν is the outer unit normal to ∂V at P0.
If c1(z)=0, then the above conclusion is also valid when we drop the assumption U(P0)=0.
To handle the equations in this paper, we also give a maximum principle as follows.
Lemma 2.4. Suppose that h∈C(¯Ω×R).
(i) If w∈C2(Ω)⋂C1(¯Ω) satisfies
ΔHw+h(ξ,w(ξ))≥0onΩ,∂w∂ν≤0in∂Ω, | (2.8) |
and w(ξ0)=max¯Ωw, then h(ξ0,w(ξ0))≥0.
(ii) If w∈C2(Ω)⋂C1(¯Ω) satisfies
ΔHw+h(ξ,w(ξ))≤0onΩ,∂w∂ν≥0in∂Ω, | (2.9) |
and w(ξ0)=min¯Ωw, then h(ξ0,w(ξ0))≤0.
Proof. We prove (ⅰ) only since (ⅱ) can be established in a similar way.
If ξ0∈Ω. Since w(ξ0)=max¯Ωw, we have ΔHw(ξ0)≤0. Thus, the conclusion holds from (2.8).
If ξ0∈∂Ω. We argue by contradiction. Suppose that h(ξ0,w(ξ0))<0. Then, by the continuity of h and w, there exists a small ball BH in ¯Ω with ∂BH⋂∂Ω={ξ0} such that h(ξ,w(ξ))<0 for ξ∈BH. Therefore, by (2.8), we have ΔHw(ξ)>0 for all ξ∈BH.
Since w(ξ0)=max¯BHw, it follows from the Hopf boundary Lemma 2.3 that ∂w∂ν(ξ0)>0, which contradicts the boundary condition in (2.8).
For the following Harnack inequality, we refer to [3,23,25].
Lemma 2.5. Let Ω be a bounded domain and K(ξ)∈C(¯Ω), u satisfies
−ΔHu+K(ξ)u=0onΩ,u=0in∂Ω, |
then there exists a positive constant C=C(‖K(ξ)‖L∞(Ω),Ω), such that max¯Ωu≤Cmin¯Ωu.
In this section, we consider the non-existence of non-constant solutions to the following semilinear subelliptic system
{d1ΔHu+uf(u,v)=0,in Ω,d2ΔHv+vg(u,v)=0,in Ω,∂u∂ν=∂v∂ν=0,on ∂Ω,u>0,v>0,in Ω. | (3.1) |
Throughout this section, C and Ci will always denote generic positive constants depending only on f,g and/or Ω. Let (u∗,v∗) be a positive root to f(u,v)=g(u,v)=0,
M=(∂f∂u∂f∂v∂g∂u∂g∂v)(u,v)=(u∗,v∗) | (3.2) |
and |M| denotes the determinant of the matrix M.
Theorem 3.1. Suppose that (H1′) and (H3) hold. Then (u,v)=(u∗,v∗) is the only solution of problem (3.1) if either
(i) |M|>0 or
(ii) |M|≤0 and max{d1,d2}≥C1 for some constant C1.
To prove Theorem 3.1, we need some preliminary results. In this section, set
Γ1={(u,v)∈R+×R+|f(u,v)=0},Γ2={(u,v)∈R+×R+|g(u,v)=0},I1={(u,v)∈R+×R+|f(u,v)≥0≥g(u,v)},I2={(u,v)∈R+×R+|f(u,v)≤0≤g(u,v)}. |
Lemma 3.2. Suppose that (H1′) and (H3) hold.
(i) If |M|>0, then
I1⊂{(u,v)∈R+×R+|u≤u∗,v≥v∗}andI2⊂{(u,v)∈R+×R+|u≥u∗,v≤v∗}. | (3.3) |
(ii) |M|<0, then
I1⊂{(u,v)∈R+×R+|u≥u∗,v≤v∗}andI2⊂{(u,v)∈R+×R+|u≤u∗,v≥v∗}. | (3.4) |
(iii) |M|=0, then there are three possibilities: the two sets I1 and I2 satisfy (3.3), or they satisfy (3.4), or one of them is equal to the set {(u∗,v∗)}.
Proof. We shall show (ⅰ) only, since parts (ⅱ) and (ⅲ) can be shown in similar ways. By (H1′), the curves Γ1,Γ2 can be represented as
Γ1={u=F(v),0<v<∞},Γ2={v=G(u),0<u<∞}. |
It is easy to show that F,G are non-increasing functions with F(v∗)=u∗ and G(u∗)=v∗. Then, our conclusion follows from (H3) and the observation that if |M|>0, Γ1 lies above Γ2 for 0<u<u∗ in uv plane, and Γ1 is below Γ2 for u≥u∗.
Lemma 3.3. Suppose that (H1′) and (H3) hold.
(i) If |M|>0, then (u,v)=(u∗,v∗) is the only solution of problem (3.1).
(ii) If |M|≤0, then any solution (u,v) of problem (3.1) satisfies the following estimate:
max¯Ωu≥u∗≥min¯Ωu,max¯Ωv≥v∗≥min¯Ωv. | (3.5) |
Proof. Let u(ξ0)=max¯Ωu, by Lemma 2.4 and (H1′), we have
0≤f(u(ξ0),v(ξ0))≤f(max¯Ωu,min¯Ωv), | (3.6) |
and in a similar way, we can obtain that
f(min¯Ωu,max¯Ωv)≤0,g(min¯Ωu,max¯Ωv)≥0,g(max¯Ωu,min¯Ωv)≤0. | (3.7) |
The (3.6) and (3.7) show that
(max¯Ωu,min¯Ωv)∈I1and(min¯Ωu,max¯Ωv)∈I2. |
By Lemma 3.2, we have, if |M|>0,
max¯Ωu≤u∗≤min¯Ωu,max¯Ωv≤v∗≤min¯Ωv. |
This implies that (u,v)=(u∗,v∗), hence (i) is established. Part (ii) follows similarly from Lemma 3.2.
We shall present a priori estimates on solutions of the strongly-coupled subelliptic system (1.2).
Lemma 3.4. Suppose that (H1) holds. Then, there exists a positive constant ˜C=˜C(ai,bi,ci) such that for any solution (u,v) of problem (1.2) satisfying the following estimates:
max¯Ωu≤˜C(1+α12d1),max¯Ωv≤˜C(1+α21d2). | (3.8) |
Proof. Let Ψ=u(d1+α11u+α12v), then Ψ satisfies
{ΔHΨ+uf(u,v)=0,in Ω,∂Ψ∂ν=0,on ∂Ω. |
Let Ψ(ξ0)=max¯ΩΨ, then by Lemma 2.2 and the positivity of u, we have f(u(ξ0),v(ξ0))≥0. Therefore,
f(0,0)≥f(0,0)−f(u(ξ0),v(ξ0))=(f(0,0)−f(u(ξ0),0))+(f(u(ξ0),0)−f(u(ξ0),v(ξ0)))=(−∂f∂u(η1,0))u(ξ0)+(−∂f∂v(u(ξ0),η2))v(ξ0)≥b1u(ξ0)+c1v(ξ0), |
where the last inequality follows from the assumption (H1) and η1≥0,η2≥0. Hence, we have u(ξ0)≤a1b1 and v(ξ0)≤a1c1. Then,
max¯ΩΨ=Ψ(ξ0)≤a1b1(d1+α11a1b1+α12a1c1), |
which in turn implies that
(d1+α11max¯Ωu)max¯Ωu≤max¯ΩΨ≤a1b1(d1+α11a1b1+α12a1c1). | (3.9) |
If α11≤d1, it follows directly from (3.9) that
max¯Ωu≤a1b1(1+a1b1+α12a1d1c1). | (3.10) |
If α11≥d1, by (3.9) we obtain
α11(max¯Ωu)2≤a1b1(d1+α11a1b1+α12a1c1), |
then
(max¯Ωu)2≤a1b1(d1α11+a1b1+α12a1α11c1)≤a1b1(1+a1b1+α12a1d1c1). | (3.11) |
Combining (3.10) and (3.11), we obtain the first half of (3.8). The estimate of max¯Ωv can be obtained in a similar way.
Proof of Theorem 3.1. By Lemma 3.3, it suffices to consider the case |M|≤0. Let (u,v) be an arbitrary solution of (3.1). We claim that there exists a positive constant C, independent of (u,v), such that
‖u−ˉu‖L∞(Ω)≤Cd1, | (3.12) |
where ˉu is the average of u in Ω, i.e., ˉu=1|Ω|∫Ωu.
Following the proof of Lemma 3.4, by (H1′), we obtain
max{‖u‖L∞(Ω),‖v‖L∞(Ω)}≤C1=max{a1b1,a2c2}. | (3.13) |
Substituting u−ˉu into the problem (3.1), we have
{ΔH(u−ˉu)+˜fd1=0,in Ω,∂(u−ˉu)∂ν=0,on ∂Ω, | (3.14) |
where ˜f=uf(u,v) can be estimated by
‖˜f‖L∞(Ω)=‖uf(u,v)‖L∞(Ω)≤C=max0≤u,v≤C1|uf(u,v)|. | (3.15) |
Multiplying (3.14) by u−ˉu, by Green's identity, Hölder's inequality, and Poincaré's inequality, we derive
∫Ω|∇Hu|2≤‖˜f‖L∞d1∫Ω|u−ˉu|≤Cd1‖u−ˉu‖L2(Ω)≤Cd1‖∇Hu‖L2(Ω), |
which implies that
‖u−ˉu‖L2(Ω)≤Cd1. | (3.16) |
By Lemma 2.1 and (3.14), (3.15), we get
‖u−ˉu‖W2,2(Ω)≤C(‖u−ˉu‖L2(Ω)+‖˜f‖L∞(Ω)d1)≤Cd1, |
and hence, by Sobolev embedding theorem [3,11],
{‖u−ˉu‖L∞(Ω)≤Cd1ifQ≤4,‖u−ˉu‖L2QQ−4(Ω)≤Cd1ifQ≥5. |
Since 2QQ−4>2, this proves (3.16). Iterating this argument finitely many times, we establish (3.12). Furthermore, it follows from (3.12) that
|max¯Ωu−min¯Ωu|≤2‖u−ˉu‖L∞(Ω)≤Cd1. | (3.17) |
Then, we will show that there exists a positive constant C (independent of u and v), such that
‖u−u∗‖L∞(Ω)≤Cd1. | (3.18) |
It follows from the above process and Lemma 3.3 that
ˉu−Cd1≤min¯Ωu≤u∗≤max¯Ωu≤ˉu+Cd1, |
that is
|ˉu−u∗|≤Cd1, |
which in turn implies that
‖u−u∗‖L∞(Ω)≤‖u−ˉu‖L∞(Ω)+|ˉu−u∗|≤Cd1. |
Simultaneously, there exists a positive constant C (independent of u and v) such that the following inequality holds
‖v−v∗‖L∞(Ω)≤Cd1. | (3.19) |
From (3.7), it follows that for some ζ1>0,ζ2>0,
−∂g∂v(max¯Ωu,ζ2)(max¯Ωv−min¯Ωv)=g(max¯Ωu−min¯Ωv)−g(max¯Ωu−max¯Ωv)≤g(min¯Ωu−max¯Ωv)−g(max¯Ωu−max¯Ωv)=−∂g∂u(ζ1,max¯Ωv)(max¯Ωu−min¯Ωu). |
Hence, by (H1′) and (3.17), we deduce that
max¯Ωv−min¯Ωv≤‖∂g∂u‖L∞(Ω)c2(max¯Ωu−min¯Ωu)≤Cd1, |
which, together with Lemma 3.3, shows that (3.19) holds.
At last, we prove that there exists a constant C1 (independent of u and v), such that if max{d1,d2}≥C1, then the only solution of (3.1) is (u,v)=(u∗,v∗).
Multiplying the first equation of (3.1) with u−ˉu, by Green's identity, we obtain
d1∫Ω|∇Hu|2=∫Ω(u−ˉu)(uf(u,v)−ˉuf(ˉu,ˉv))=∫Ω(u−ˉu)((u−ˉu)f(u,v)+ˉu(f(u,v)−f(ˉu,ˉv)))≤C∫Ω|u−ˉu|2+C∫Ω|u−ˉu||v−ˉv|=Cε∫Ω|u−ˉu|2+ε∫Ω|v−ˉv|2. | (3.20) |
For the second equation of (3.1), we proceed slightly differently, as follows.
d2∫Ω|∇Hv|2=∫Ω(v−ˉv)(vg(u,v)−ˉvg(ˉu,ˉv))=∫Ω[g(u,v)|v−ˉv|2+ˉv(v−ˉv)(g(ˉu,v)−g(ˉu,ˉv))+ˉv(v−ˉv)(g(u,v)−g(ˉu,v))]=∫Ω{[g(u,v)+ˉv∂g∂v(ˉu,ς2)]|v−ˉv|2+ˉv∂g∂u(ς1,v)(u−ˉu)ˉv(v−ˉv)}≤C∫Ω(g(u,v)−c2ˉv)|v−ˉv|2+C∫Ω|u−ˉu||v−ˉv|=Cε∫Ω|u−ˉu|2+∫Ω(g(u,v)−c2ˉv+ε)|v−ˉv|2, | (3.21) |
where ς1(ξ) lies between ˉu and u(ξ), and ς2(ξ) lies between ˉv and v(ξ) for each ξ∈Ω. From the above conclusions, it follows that (3.18) and (3.19) hold. And then, by (3.18) and (3.19), if d1≥C,
g(u,v)−c2ˉv=g(u,v)−g(u∗,v∗)−c2ˉv=g(u,v)−g(u∗,v)+g(u∗,v)−g(u∗,v∗)−c2ˉv≤−b2‖u−u∗‖L∞(Ω)−c2‖v−v∗‖L∞(Ω)−c2ˉv≤−b2Cd1−c2Cd1−c2ˉv≤−c2v∗2. |
Choosing ε=c2v∗4 in (3.21), we have
d2∫Ω|∇Hv|2≤C∫Ω|u−ˉu|2−c2v∗4∫Ω|v−ˉv|2. | (3.22) |
Combing (3.20) and (3.22), we arrive at
d1∫Ω|∇Hu|2≤C∫Ω|u−ˉu|2≤C2∫Ω|∇Hu|2, |
which implies that if d1>C2, then ∇Hu≡0, i.e., u is constant.
Then, (3.22) guarantees that v≡ˉv, a non-negative constant.
In view of part (ii) of Lemma 3.3, we see that these constants must be positive. Hence, from the assumption (H3), we conclude that (u,v)≡(u∗,v∗).
A similar argument applies when d2 is large, leading to the same conclusion. This completes the proof of Theorem 3.1.
As a consequence of Theorem 3.1, we have the following corollary:
Corollary 3.5. If f=u(a1−b1u−c1v) and g=v(a2−b2u−c2v), then (u,v)=(u∗,v∗) is the only solution of problem (3.1) if either
(i) b1b2>a1a2>c1c2 or
(ii) b1b2<a1a2<c1c2 and max{d1,d2}≥C1 for some constant C1.
Remark 1. The equations in Theorem 3.1 and Corollary 3.5 involve subelliptic operators, which are more general than elliptic operator as described in [16], and the proof mainly relies on Lemma 2.4, which is the subelliptic case.
In this section, we mainly study the effects of diffusion and self-diffusion in the strongly-coupled subelliptic system (1.2). Throughout this section, C will always denote generic positive constants depending only on d1,d2,α12,α21 and the nonlinearity f,g, but independent of α11,α22.
Theorem 4.1. Suppose that the conditions (H1) and (H2) hold. Then, there exists a constant C such that if max{α11,α22}≥C, the problem (1.2) has no non-constant solution.
Lemma 4.2. Suppose that (H1) and (H2) hold.
(i) If f(u,v)=g(u,v)=0 has no positive root, then there exists a constant C such that (1.2) has no solution provided that max{α11,α22}≥C.
(ii) If f(u,v)=g(u,v)=0 has at least a positive root, then there every small ε>0, there exists a constant C(ε) such that if max{α11,α22}≥C(ε), for any solution (u,v) of (1.2), there are two positive constants ˆu,ˆv that f(ˆu,ˆv)=g(ˆu,ˆv)=0 and ‖u−ˆu‖L∞(Ω)+‖v−ˆv‖L∞(Ω)≤ε.
Proof. We prove (ⅱ) at first; suppose that the conclusion is false. Without loss of generality, we assume that there exists a constant ε0>0, and a sequence {α11,k,α22,k}∞k=1 with α11,k→∞, such that
‖uk−ˆu‖L∞(Ω)+‖vk−ˆv‖L∞(Ω)≥ε0 | (4.1) |
for any positive root (ˆu,ˆv) of f(u,v)=g(u,v)=0, where (uk,vk) is a solution to
{ΔH[(d1+α11,kuk+α12vk)uk]+ukf(uk,vk)=0,in Ω,ΔH[(d2+α21uk+α22,kvk)vk]+vkg(uk,vk)=0,in Ω,∂uk∂ν=∂vk∂ν=0,on ∂Ω,uk>0,vk>0,in Ω. | (4.2) |
We use the same notation of the subsequence of {uk}∞k=1 as for the original sequence {uk}∞k=1, such that uk converges uniformly to a constant as k→∞. Set
Φk=uk(uk+d1α11,k+α12α11,kvk), |
then Φk satisfies
{α11,kΔHΦk+ukf(uk,vk)=0,in Ω,∂Φk∂ν=0,on ∂Ω. |
By Lemma 3.4 and the fact α11,k→∞, we know that ‖Φk‖L∞(Ω)≤C. Hence by standard Lp estimates and the Sobolev embedding theorem [5,11,24], we obtain ‖Φk‖C1,α(¯Ω)≤C for some α∈(0,1). Therefore, a subsequence of {Φk}∞k=1 converges to some nonnegative function Φ in C1(¯Ω), and Φ must satisfy the following problem weakly
{ΔHΦ=0,in Ω,∂Φ∂ν=0,on ∂Ω. |
By standard subelliptic regularity theory, Φ∈C2(¯Ω) and therefore Φ=ˆΦ, where ˆΦ is a nonnegative constant. Letting ˆu=√ˆΦ, we get that
u2k−ˆu2=Φk−ˆΦ−d1α11,kuk−α12α11,kukvk→0 |
as k→∞. Hence uk→ˆu uniformly.
Next, we claim the subsequence of {vk}∞k=1 and also denote {vk}∞k=1, such that vk→ˆv uniformly as k→∞, where ˆv is some nonnegative constant.
Before establishing the above assertion, we show how to derive a contradiction via the fact that (uk,vk)→(ˆu,ˆv) uniformly as k→∞.
Integrating the equations of (4.2) in Ω, we have
∫Ωukf(uk,vk)=∫Ωvkg(uk,vk)=0. | (4.3) |
From this, we conclude that f(ˆu,ˆv)=g(ˆu,ˆv)=0 for (ˆu,ˆv). Suppose that f(ˆu,ˆv)≠0. Without loss of generality, we may assume that f(ˆu,ˆv)>0. Since (uk,vk)→(ˆu,ˆv) uniformly, f(uk,vk)→f(ˆu,ˆv) as k→∞. Hence, f(ˆuk,ˆvk)>0 for k large, and therefore
∫Ωukf(uk,vk)>0 |
for large k since uk is always positive, which contradicts (4.3). A similar contradiction can be deduced if g(ˆu,ˆv)≠0.
By (H2) and the assumption that f(0,0)=a1>0, g(0,0)=a2>0, we must have
ˆu>0,ˆv>0. |
That is, (uk,vk)→(ˆu,ˆv) uniformly with
ˆu>0,ˆv>0andf(ˆu,ˆv)=g(ˆu,ˆv)=0, |
which contradicts (4.1) and thus establishes (ⅱ) of Lemma 4.2.
To finish the proof of part (ⅱ) of Lemma 4.2, it remains to show the above assertion.
If {α22,k}∞k=1 is unbounded. We choose a subsequence of {α22,k}∞k=1, still denoted as {α22,k}∞k=1, such that α22,k→∞ as k→∞. We can then argue in very much the same way as before to conclude that vk→ˆv for some non-negative constant ˆv.
If {α22,k}∞k=1 is bounded. Without loss of generality, we may assume that α22,k→α22∈[0,∞). Set
Υk=(d2+α21uk+α22,kvk)vk. |
Since {α22,k}∞k=1 is bounded, by Lemma 3.4 it is easy to know that ‖Υk‖L∞(Ω)≤C. Hence, Υk satisfies
{ΔHΥk+vkg(uk,vk)=0,in Ω,∂Υk∂ν=0,on ∂Ω. | (4.4) |
By standard Lp estimate and the Sobolev embedding theorem, we obtain ‖Υk‖C1,α(¯Ω)≤C for some α∈(0,1). Then, by passing to a subsequence if necessary, we may assume that {Υk}∞k=1 converges to some nonnegative function Υ in C1(¯Ω). By the definition of Υk and the fact u→ˆu, we see
Υ−(d2+α21ˆuk+α22,kvk)vk→0 |
in C1(¯Ω). If α22>0, it is easy to get vk→˜v in C1(¯Ω), where
˜v=−(d2+α21ˆu)+√(d2+α21uk)2+4α22Υ2α22≥0. |
Letting k→∞ in (4.4), we can know that Υ satisfies the following problem weakly
{ΔHΥ+˜vg(ˆu,˜v)=0,in Ω,∂Υ∂ν=0,on ∂Ω. | (4.5) |
The standard subelliptic regularity theory ensures that Υ∈C2(¯Ω), and hence is a classical solution of (4.5). Note that Υ≥0. If Υ≡0, then we claim that vk→0 in C1(¯Ω). Since uk→ˆu, by (4.2) we can argue similarly as before to show that f(ˆu,0)=0 and ˆu>0, which contradicts (H2). Therefore, Υ≥0 and is not identically zero in Ω. The problem (4.5) can be rewritten as
{ΔHΥ+g(ˆu,˜v)d2+α21u+α22vΥ=0,in Ω,∂Υ∂ν=0,on ∂Ω. |
By Lemma 2.2, Υ>0, and thus ˜v>0 in ¯Ω. Since ˜v is a solution of
{ΔH[(d2+α21ˆu+α22˜v)˜v]+˜vg(ˆu,˜v)=0,in Ω,∂˜v∂ν=0,on ∂Ω, | (4.6) |
by Lemma 2.4 and the positivity of ˜v, we obtain g(ˆu,max¯Ω˜v)≥0. Thus, from assumption (H1), it follows that
g(ˆu,˜v(ξ))≥g(ˆu,max¯Ω˜v)≥0,∀ξ∈Ω. |
Integrating the equation of (4.6) in Ω shows
0=∫Ω˜vg(ˆu,˜v))≥∫Ω˜vg(ˆu,max¯Ω˜v)=g(ˆu,max¯Ω˜v)∫Ω˜v≥0, |
which implies that ˜v≡max¯Ω˜v>0. That is, if α22,k→α22>0, then there exists a subsequence of {α22,k}∞k=1 which converges uniformly to some positive constant.
If α22=0, we have already established that
vk→˜v=Υd2+α21ˆu |
in C1(¯Ω) as k→∞. Then, our conclusion that a subsequence of {vk}∞k=1 converges to some positive constant follows from the same arguments as in the case α22>0 with obvious modifications. This proves our assertion, and the proof of part (ii) is now complete.
Finally, we return to the proof of part (ⅰ). Suppose that the conclusion in (ⅰ) fails. Then, we can assume that there exists a sequence of solutions {(uk,vk)}∞k=1 to (4.2) with α11,k→∞.
Similarly to the processes in part (ⅱ), we show that there exists a subsequence of {(uk,vk)}∞k=1 that converges uniformly to some non-negative (ˆu,ˆv). Again, (4.3) and the arguments following it guarantee that f(ˆu,ˆv)=g(ˆu,ˆv)=0. By (H1) and (H2), we conclude that ˆu>0 and ˆv>0. However, this contradicts our assumption of (ⅰ).
Lemma 4.3. Suppose that (H1) and (H2) hold and min{d1,d2}≥ϵ.
(i) If f(u,v)=g(u,v)=0 have no positive root, then there exists some positive constant C1=C1(ϵ,α11, α12,α21,α22) such that (1.2) has no solution provided that max{d1,d2}≥C1.
(ii) If f(u,v)=g(u,v)=0 have a positive root, then for any small ε>0, there exists a positive constant C2=C2(ε,ϵ,α11,α12,α21,α22) such that if max{d1,d2}≥C2, for any solution (u,v) of (1.2), there are two positive constants ˆu,ˆv that f(ˆu,ˆv)=g(ˆu,ˆv)=0 and ‖u−ˆu‖L∞(Ω)+‖v−ˆv‖L∞(Ω)≤ε.
Proof. We shall only prove part (ⅱ), as (ⅰ) can be shown in a similar way. For the proof of (ⅱ), we still argue by contradiction. We assume that there exist two positive constants ϵ0 and ε0, and a sequence {d1,k,d2,k}∞k=1 with d1,k→∞ and d2,k≥ϵ0, such that
‖uk−ˆu‖L∞(Ω)+‖vk−ˆv‖L∞(Ω)≥ε0 | (4.7) |
for any positive root (ˆu,ˆv) of f(u,v)=g(u,v)=0, where (uk,vk) is a solution to
{ΔH[(d1,k+α11uk+α12vk)uk]+ukf(uk,vk)=0,in Ω,ΔH[(d2,k+α21uk+α22vk)vk]+vkg(uk,vk)=0,in Ω,∂uk∂ν=∂vk∂ν=0,on ∂Ω,uk>0,vk>0,in Ω. | (4.8) |
For the problem (4.8), Lemma 3.4 implies that
max¯Ω{uk,vk}≤C1=C1(ϵ,α11,α12,α21,α22). |
To show that uk converges to some constant, let
Φk=uk(1+α11d1,kuk+α12d1,kvk). | (4.9) |
Then by similar arguments as in the proof of Lemma 4.2, we see that Φk converges uniformly to some non-negative constant Φ. By (4.9) and the fact d1,k→∞, uk converges uniformly to Φ. If {d2,k}∞k=1 is unbounded, then it is easy to show that a subsequence of {d2,k}∞k=1 also converges to a non-negative constant. If {d2,k}∞k=1 bounded, setting
Υk=(d2,k+α21uk+α22vk)vk, |
we know that a subsequence of {Υk}∞k=1 converges to some non-negative function Υ, and hence a subsequence of {vk}∞k=1 converges to a nonnegative function ˜v. Then, we can proceed further as in the proof of Lemma 4.2 to show that ˜v is a constant that derives a contradiction.
Lemma 4.4. Suppose that (H1) and (H2) hold and α22>0.
(i) If f(u,v)=g(u,v)=0 has no positive root, then there exists a positive constant C3=C3(α11,α12, α21,α22) such that (1.2) has no solution provided that d1≥C3.
(ii) If f(u,v)=g(u,v)=0 has a positive root, then for any small ε>0, there exists a constant C4=C4(ε,α11,α12,α21,α22) such that if d1≥C4, for any solution (u,v) of (1.2), there exist two positive constants ˆu,ˆv that f(ˆu,ˆv)=g(ˆu,ˆv)=0 and ‖u−ˆu‖L∞(Ω)+‖v−ˆv‖L∞(Ω)≤ε.
Similar results hold if α11>0 and d2 is large enough.
Proof. In view of Lemma 4.3, it suffices to consider the case d1,k→∞ and d2,k→0. For this case, by following the proof of Lemma 4.3, we obtain
max¯Ωuk≤a1b1(1+a1b1+α12d1,ka1c1)≤a1b1(1+a1b1+a1c1), |
and
max¯Ωvk≤[a2b2(1+α21α22a2b2+a2c2)]2, |
for large k. Then, we can prove Lemma 4.4 in the same way as Lemma 4.3.
Proof of Theorem 4.1. In view of part (ⅰ) of Lemma 4.2, we may assume that f(u,v)=g(u,v)=0 has at least a positive root. Setting
S={(u,v)∈R+×R+|f(u,v)=g(u,v)=0}. |
By (H1) and (H2) we know
δ=inf(u,v)∈S{u,v}>0. |
Choosing ε=δ2 in Lemma 4.2, there is a positive constant C(δ) and C such that if max{α11,α22}≥C(δ), then for any solution (u,v) of (1.2),
δ2≤u(ξ),v(ξ)≤C,∀ξ∈Ω. | (4.10) |
Without loss of generality, we may assume that α11 is sufficiently large. Let (ˉu,ˉv) be the average of (u,v) in Ω. Multiplying the first equation of problem (1.2) by u−ˉu and integrating in Ω, by the same arguments as in (3.20), we get
∫Ω[(d1+2α11u+α12v)|∇Hu|2+α12u∇Hu⋅∇Hv]=∫Ω(u−ˉu)uf(u,v)≤Cε∫Ω|u−ˉu|2+ε∫Ω|v−ˉv|2. | (4.11) |
By Lemma 3.4, (4.10), and Poincaré's inequality, we have
|∫Ωα12u∇Hu⋅∇Hv|≤Cε∫Ω|∇Hu|2+ε∫Ω|∇Hv|2. |
Using (4.10) and Poincaré's inequality, we obtain
(α11δ−Cε)∫Ω|∇Hu|2≤ε(1+1λ1)∫Ω|∇Hv|2, | (4.12) |
where λ1 is the smallest positive eigenvalues of the sub-Laplace operator subject to the homogeneous Neumann boundary condition (see [1]). For the second equation of problem (2.1), we proceed as in (3.21) to obtain
∫Ω[(d2+α21u+2α22v)|∇Hv|2+α21v∇Hu⋅∇Hv]=∫Ω(v−ˉv)vg(u,v)≤Cε∫Ω|u−ˉu|2+∫Ω(g(u,v)−c2ˉv+ε)|v−ˉv|2. | (4.13) |
By Lemma 4.2, for any small ε, there exists C(ε) such that if α11≥C(ε), then
‖u−ˆu‖L∞(Ω)+‖v−ˆv‖L∞(Ω)≤ε |
for some (ˆu,ˆv)∈S. And then
‖g(u,v)‖L∞(Ω)=‖g(u,v)−g(ˆu,ˆv)‖L∞(Ω)≤Cε. |
As ˉv≥δ2, we know that for α11≥C(ε) and ε small enough,
g(u,v)−c2ˉv+ε≤(C+1)ε−c2δ2≤0. |
Therefore
d2∫Ω|∇Hv|2=∫Ωα21v|∇Hu||∇Hv|+Cε∫Ω|u−ˉu|2≤Cε∫Ω|∇Hu|2+ε∫Ω|∇Hv|2. | (4.14) |
Combining (4.12) and (4.14), we have
(α11δ−Cε)∫Ω|∇Hu|2+(d2−ε(2+1λ1))∫Ω|∇Hv|2≤0. | (4.15) |
Choosing ε small enough, for α11 sufficiently large, ∇Hu=∇Hv≡0, then (u,v) is constant.
Theorem 4.5. Suppose that the conditions (H1) and (H2) hold. For any ϵ>0, there exists some positive constant C5=C5(ϵ,α11,α12,α21,α22) such that if min{d1,d2}≥ϵ and max{d1,d2}≥C5, then problem (1.2) has no non-constant solution.
Proof. Replacing α11δ by d1 in both (4.12) and (4.15), and following the proof of Theorem 4.1 with the help of Lemma 4.3 instead, we see immediately that this theorem holds.
Theorem 4.6. Suppose that the conditions (H1) and (H2) hold.
(i) There exists a positive constant C6=C6(d2,α11,α12,α21,α22) such that if d1≥C6, problem (1.2) has no non-constant solution. Furthermore, if α22>0, then C6 can be chosen independent of d2.
(ii) There exists a positive constant C7=C7(d1,α11,α12,α21,α22) such that if d2≥C7, problem (1.2) has no non-constant solution. Furthermore, if α11>0, then C7 can be chosen independent of d1.
Proof. We shall establish part (ⅰ) only, since (ⅱ) can be shown in a similar way. By letting ϵ=d2 and C6=max{C5,d2} in Theorem 4.5, we know that the first assertion of (ⅰ) follows immediately from Theorem 4.5. To prove the second assertion, we first note that by choosing ε=δ2, from Lemma 4.4 it follows that min¯Ωv≥δ2. Then, we modify the proof of Theorem 4.1 by replacing the constant d2 in (4.14) and (4.15) by 2α22min¯Ωv, and the term α11δ by d1 in both (4.12) and (4.15). The remaining arguments are rather similar as before and are thus omitted.
It follows immediately from Theorem 4.6 that
Corollary 4.7. Suppose that the conditions (H1) and (H2) hold, α11>0 and α22>0. Then, there exists a positive constant C8=C8(α11,α12,α21,α22) such that if max{d1,d2}≥C8, problem (1.2) has no non-constant solution.
Remark 2. We note that Theorem 1.1 follows from Theorem 4.1 and Corollary 4.7. Moreover, from Theorem 4.1 and Corollary 4.7 we see that large self-diffusion seems to have a very similar effect to large diffusion, as observed in [16].
In this section, we mainly study the predator-prey system (1.3). Throughout this section, C will always denote generic positive constants.
At first, we study the case ˉα21=0, and give the proof of Theorem 1.2. As a by-product a priori estimate is established by using the maximum principle and the Harnack inequality.
Theorem 5.1. Suppose that ˜d1,˜d2, and ˜α12 are given positive constants. Then, there exists a positive constant C=C(a2,c2,˜d1,˜d2,˜α12) such that if d1≥˜d1,d2≥˜d2 and ˉα12≤˜α12, then every positive solution (u,v) of (1.3) satisfies C−1<u,v<C.
Proof. Assume that (u,v) is a positive solution of problem (1.3) and denote Π=(1+α12v)u. Then problem (1.3) becomes
{d1ΔHΠ+uq(u)−p(u)v=0,in Ω,d2ΔHv+v(−a2+c2p(u))=0,in Ω,∂Π∂ν=∂v∂ν=0,on ∂Ω. | (5.1) |
Let ξ1∈¯Ω be a point where Π(ξ1)=max¯ΩΠ(ξ). By Lemma 2.4, for the first equation of problem (5.1), we obtain that
u(ξ1)q(u(ξ1))−p(u(ξ1))v(ξ1)≥0. |
Therefore, u(ξ1)q(u(ξ1))≥0. By (H4), we have
0<u(ξ1)≤S |
and
0<v(ξ1)≤u(ξ1)q(u(ξ1))p(u(ξ1))≤Sq(u(ξ1))p(u(ξ1)):=M, |
here, the condition limu→0+p′(u)<∞ in (H5) shows that supu∈(0,S)uq(u)p(u)<∞. Thus,
max¯Ωu(ξ)≤max¯ΩΠ(ξ)=(1+ˉα12v(ξ1))u(ξ1)≤(1+ˉα12M)S:=C1. |
Multiplying c2 to the first equation of (1.3) and adding it to the second equation of (1.3) and then integrating over Ω, we obtain
∫Ω{c2d1ΔH[(1+ˉα12v)u]+d2ΔHv}=∫Ω[a2v−c2uq(u)]. |
By Green's identity, we know that
∫Ω{c2d1ΔH[(1+ˉα12v)u]+d2ΔHv}=0. |
So
a2∫Ωv=c2∫Ωuq(u)≤c2∫Ωq(0)C1=c2q(0)C1|Ω|, |
that is,
∫Ωv≤c2q(0)C1|Ω|a2. |
The problem (1.3) can also be written as
{−ΔHΠ=q(u)−p(u)uvd1(1+α12v)Π,in Ω,−ΔHv=v(−a2+c2p(u))d2,in Ω,∂Π∂ν=∂v∂ν=0,on ∂Ω. |
For u<S and d2≥˜d2, we see −a2+c2p(u)d2<c2p(S)˜d2<∞, so the Lemma 2.5 holds for v,
max¯Ωv≤C0min¯Ωv | (5.2) |
for some positive constant C0. Hence, we have
max¯Ωv≤C0min¯Ωv≤C0∫Ωv|Ω|≤c2q(0)C1C0a2:=C2. | (5.3) |
By integrating the first equation of problem (1.3) over Ω, we have
∫Ω(uq(u)−p(u)v)=0. | (5.4) |
Equation (5.4) implies that there exists a point ξ2∈Ω, such that
(u(ξ2)q(u(ξ2))−p(u(ξ2))v(ξ2))=0. |
By assumptions (H4) and (H5), it follows that 0<u(ξ2)<S. Then,
v(ξ2)=u(ξ2)q(u(ξ2)p(u(ξ2))>0. |
If min¯Ωv=0, by (5.2) it follows that max¯Ωv=0. That means that v≡0 uniformly in Ω, which is a contradiction. Thus v has a positive lower bound for d2≥˜d2.
In the following, we show that u has a positive lower bound.
By (H5) and p(u)∈C2((0,+∞)), it follows that
limu→0+p(u)u=limu→0+p′(u)<∞, |
there exists a positive constant ˉp such that p(u)u≤ˉp for small 0<u≤S. For d1≥˜d1, we have
q(u)−p(u)uvd1(1+ˉα12v)≤q(0)+ˉpC2˜d1<∞. |
Thus Lemma 2.5 holds for Π,
max¯ΩΠ≤˜C0min¯ΩΠ | (5.5) |
for some positive constant ˜C0. By (5.3) and (5.5), we get
max¯Ωumin¯Ωu≤max¯ΩΠmin¯ΩΠ⋅1+ˉα12max¯Ωv1+ˉα12min¯Ωv≤˜C0C1(1+ˉα12max¯Ωv)≤˜C0C1(1+ˉα12C2):=C3. | (5.6) |
To obtain a contradiction, assume that there exists a sequence {(d1,k,d2,k,ˉα12,k)}∞k=1, satisfying d1,k≥˜d1,d2,k≥˜d2 and ˉα12,k≤˜α12 for some ˜α12>0, such that the corresponding positive solutions (uk,vk) of problem (1.3) with (d1,d2,ˉα12)=(d1,k,d2,k,ˉα12,k) such that min¯Ωuk→0 as k→∞. Using (5.6), we have max¯Ωuk→0 as k→∞. By the regularity theory for subelliptic equations, there exists a subsequence of {(uk,vk)}, which will also be denoted by {(uk,vk)}, such that uk→0 uniformly as k→∞. Integrating the second equation of problem (1.3) with (u,v)=(uk,vk), we obtain
∫Ωvk(−a2+c2p(uk))=0. | (5.7) |
Since uk→0 as k→∞, we have −a2+c2p(uk)<0 in ¯Ω for any large k. This contradicts the integrating identity (5.7) as well as the fact that vk>0.
Theorem 5.2. Suppose that p(S)≤a2c2, then problem (1.3) has no non-constant solution.
Proof. Since q(u)<0 for u≥S, we only need to consider the case u<S. Suppose, on the contrary, that (1.3) has a non-constant positive solution (u,v) for p(S)≤a2c2. Then, v must be non-constant; otherwise, it is easily seen that u must be constant from the second equation of problem (1.3).
Using the fact that p(u) is increasing in u, and integrating the second equation of problem (1.3) over Ω, we have
0=−d2∫ΩΔHv=∫Ωv(−a2+c2p(u))<∫Ωv(−a2+c2p(S)). |
Since v>0, we have p(S)≥a2c2. This contradiction completes the proof. Thus, if (1.3) has a positive solution, it must be that p(S)≥a2c2, which is the condition that p(u) should satisfy according to (H5).
Remark 3. Theorem 5.2 is directly characterized by the function p(u). When ˉα21=0, we will prove the non-existence result. Theorem 1.2, which considers the self-diffusion and cross-diffusion rates d1, and d2, is given in [14].
Theorem 5.3. Suppose that ˜d1=λ1+λ−11(q(0)+c2λ1K) and ˜d2=λ−11(−a2+c2p(˜C)+c2K+(d1ˉα12˜C)24) where K=sup¯Ωˉvp′(u). If d1≥˜d1 and d2≥˜d2, then problem (1.3) has no non-constant solution.
Proof. Let (u,v) be a positive solution of problem (1.3). Multiplying the equations of problem (1.3) by u−ˉu,v−ˉv, and then integrating over Ω, using the mean value theorem, we get
∫Ω[d1(1+ˉα12v)|∇Hu|2+d2|∇Hv|2+d1ˉα12u∇Hu⋅∇Hv]=∫Ω[(u−ˉu)(uq(u)−p(u)v)+(v−ˉv)v(−a2+c2p(u))]=∫Ω(u−ˉu)[q(u)(u−ˉu)+ˉu(q(u)−q(ˉu))−p(u)(v−ˉv)−ˉv(p(u)−p(ˉu))]+∫Ω(v−ˉv)[−a2(v−ˉv)+c2p(u)(v−ˉv)+c2ˉv(p(u)−p(ˉu))]=∫Ω(u−ˉu)[q(u)(u−ˉu)+ˉuq′(η)(u−ˉu)−p(u)(v−ˉv)−ˉvp′(ζ)(u−ˉu)]+∫Ω(v−ˉv)[−a2(v−ˉv)+c2p(u)(v−ˉv)+c2ˉvp′(ς)(u−ˉu)]=∫Ω[(u−ˉu)2(q(u)+ˉuq′(η)−ˉvp′(ζ))+(v−ˉv)2(−a2+c2p(u))+(u−ˉu)(v−ˉv)(−p(u)+c2ˉvp′(ς))]<∫Ω[q(0)|u−ˉu|2+(−a2+c2p(˜C))|v−ˉv|2+c2K|u−ˉu||v−ˉv|], |
where 0<η,ζ,ς≤˜C and K=sup¯Ωˉvp′(u), we note here that p′(u) is bounded in any finite interval due to the assumptions p(u)∈C2((0,+∞)) and (H5).
By Theorem 5.1, Cauchy's inequality, and Poincaré's inequality, we see
∫Ω[d1(1+ˉα12v)|∇Hu|2+d2|∇Hv|2]<∫Ω[(q(0)+c2KT)|u−ˉu|2+(−a2+c2p(˜C)+c2K4T)|v−ˉv|2+T|∇Hu|2+(d1ˉα12u)24T|∇Hv|2]<∫Ω[(λ1+λ−11(q(0)+c2λ1K))|∇Hu|2+(λ−11(−a2+c2p(˜C)+c2K+(d1ˉα12˜C)24))|∇Hv|2], |
where T is taken as any positive constant, specifically λ1. Hence, by the assumptions d1≥˜d1,d2≥˜d2, we know that problem (1.3) has no non-constant positive solution.
The Theorem 1.2 can be obtained from Theorem 5.3.
Remark 4. If ˉα21=0 and ˉα12 is small enough as [14], then Theorem 1.2 shows that problem (1.3) does not admit a non-constant positive solution for some large enough d1,d2, which is consistent with the result Theorem of 1.1.
Next, we prove Theorem 1.3.
Theorem 5.4. Suppose that ˜d1,˜d2,˜α12,˜α21 are given positive constants. Then, there exists a positive constant C=C(a2,c2,˜d1,˜d2,˜α12,˜α21) such that if d1≥˜d1,d2≥˜d2,ˉα12≤˜α12 and ˉα21≤˜α21, then every positive solution (u,v) of (1.3) satisfies C−1<u,v<C.
The proof of Theorem 5.4 is similar to Theorem 5.1.
Theorem 5.5. Suppose that ˜d1=λ−11q(0)+c2K and ˜d2=λ−11(−a2+c2K+˜C24λ1) with K=sup¯Ωˉvp′(u). Then, there exists positive constants ˜d1,˜d2,˜α12,˜α21 such that if d1>˜d1 and d2>˜d2, then problem (1.3) has no non-constant solution when ˉα12<˜α12 and ˉα21<˜α21.
Proof. Let (u,v) be a positive solution of problem (1.3). Multiplying the equations of problem (1.3) by u−ˉu,v−ˉv, and then integrating over Ω, using the mean value theorem, we have
∫Ω{(u−ˉu)d1ΔH[(1+ˉα12v)u]+(v−ˉv)d2ΔH[(1+ˉα21u)v]}=∫Ω[d1(1+ˉα12v)|∇Hu|2+d2(1+ˉα21u)|∇Hv|2+(d1ˉα12u+d2ˉα21v)∇Hu⋅∇Hv]=∫Ω[(u−ˉu)(uq(u)−p(u)v)+(v−ˉv)v(−a2+c2p(u))]=∫Ω(u−ˉu)[q(u)(u−ˉu)+ˉu(q(u)−q(ˉu))−p(u)(v−ˉv)−ˉv(p(u)−p(ˉu))]+∫Ω(v−ˉv)[−a2(v−ˉv)+c2p(u)(v−ˉv)+c2ˉv(p(u)−p(ˉu))]=∫Ω(u−ˉu)[q(u)(u−ˉu)+ˉuq′(η)(u−ˉu)−p(u)(v−ˉv)−ˉvp′(ζ)(u−ˉu)]+∫Ω(v−ˉv)[−a2(v−ˉv)+c2p(u)(v−ˉv)+c2ˉvp′(ς)(u−ˉu)]=∫Ω[(u−ˉu)2(q(u)+ˉuq′(η)−ˉvp′(ζ))+(v−ˉv)2(−a2+c2p(u))+(u−ˉu)(v−ˉv)(−p(u)+c2ˉvp′(ς))]<∫Ω[q(0)|u−ˉu|2+(−a2+c2p(˜C))|v−ˉv|2+c2K|u−ˉu||v−ˉv|], |
where 0<η,ζ,ς≤˜C and K=sup¯Ωˉvp′(u); here, we note that p′(u) is bounded in any finite interval in view of the assumptions p(u)∈C2((0,+∞)) and (H5).
By Theorem 5.1, the Cauchy's inequality, and Poincaré's inequality, we have
∫Ω[d1|∇Hu|2+d2|∇Hv|2]≤∫Ω[d1(1+ˉα12v)|∇Hu|2+d2(1+ˉα21u)|∇Hv|2]<∫Ω[(q(0)+c2KT)|u−ˉu|2+(−a2+c2p(˜C)+c2K4T)|v−ˉv|2+(T(d1ˉα12)2+˜C24T)|∇Hu|2+(T(d2ˉα21)2+˜C24M)|∇Hv|2]<∫Ω[(λ−11(q(0)+c2λ1K+(λ1d1ˉα12)2+˜C24)|∇Hu|2+(λ−11(−a2+c2p(˜C)+c2K4λ1+(λ1d2ˉα21)2+˜C24)|∇Hv|2], |
where T is taken as any positive constant, specifically λ1. Recall that C1=(1+ˉα12M)S in the proof of Theorem 5.1. Hence,
d1>λ−11(q(0)+c2λ1K+(λ1d1ˉα12)2+˜C24)andd2>λ−11(−a2+c2p(˜C)+c2K4λ1+(λ1d2ˉα21)2+˜C24) |
i.e.,
ˉα12<˜α12=2√λ1d1−q(0)−c2λ1K2(λ1d2ˉα21)2+(KS)2andˉα21<˜α21=√λ1d2+a2−c2K+(1+ˉα12M)2S244λ1, |
we know that, under the given assumptions, Theorem 1.3 implies that problem (1.3) has no non-constant positive solution.
The Theorem 1.3 can be obtained from Theorem 5.5.
Remark 5. If ˉα12 and ˉα21 are small enough as [28], then Theorem 1.3 shows that problem (1.3) does not admit a non-constant positive solution for some large enough d1,d2, which is consistent with the result of Theorem 1.1.
We consider the Neumann boundary value problem for the strongly-coupled subelliptic system and the predator-prey subelliptic system on the Heisenberg group. We provide a priori estimates and non-existence results for non-constant positive solutions of the strongly-coupled and predator-prey systems with coefficients under different conditions. Only one of the diffusion rates or one of the self-diffusion pressures needs to be large to prevent the formation of non-constant solutions in the strongly-coupled subelliptic systems. For the predator-prey subelliptic system with cross-diffusion and homogeneous Neumann boundary conditions, we investigate the existence and non-existence of non-constant positive solutions.
Xinjing Wang: Investigation, Methodology, Validation, Writing-review and editing, Formal analysis; Guangwei Du: Methodology, Writing-original draft preparation, Visualization, Validation. All authors have read and agreed to the published version of the manuscript. Both authors contributed equally and significantly to this manuscript.
This work is supported by the Youth Backbone Teacher Funding Program in Huanghuai University, the Key Specialized Research and Development Breakthrough Program in Henan Province (No.222102310265), the Key Scientific Research Project of the Colleges and Universities in Henan Province (No.23B110009), and the China Scholarship Council (No.202408410284).
The authors declare no conflict of interest.
[1] |
A. Aribi, S. Dragomir, Dirichlet and Neumann eigenvalue problems on CR manifolds, Ricerche Mat., 67 (2018), 285–320. https://doi.org/10.1007/s11587-018-0420-x doi: 10.1007/s11587-018-0420-x
![]() |
[2] | I. Birindelli, A. Cutrì, A semi-linear problem for the Heisenberg Laplacian, Rend. Sem. Mat. Della Univ. Padova, 94 (1995), 137–153. |
[3] | A. Bonfiglioli, E. Lanconelli, F. Uguzzoni, Stratified Lie groups and potential theory for their sub-Laplacians, New York: Springer, 2007. https://doi.org/10.1007/978-3-540-71897-0 |
[4] | J. M. Bony, Principe du Maximum, Inegalite de Harnack et unicite du probleme de Cauchy pour les operateurs ellipitiques degeneres, Ann. Inst. Fourier Grenobles, 19 (1969), 277–304. |
[5] |
L. Capogna, D. Danielli, N. Garofalo, An embedding theorem and Harnack inequality for nonlinear subelliptic equations, Commun. Part. Diff. Equ., 18 (1993), 1765–1794. https://doi.org/10.1080/03605309308820992 doi: 10.1080/03605309308820992
![]() |
[6] |
E. Cinti, J. Tan, A nonlinear Liouville theorem for fractional equations in the Heisenberg group, J. Math. Anal. Appl., 433 (2016), 434–454. https://doi.org/10.1016/j.jmaa.2015.07.050 doi: 10.1016/j.jmaa.2015.07.050
![]() |
[7] |
K. Ding, Q. Zhu, T. Huang, Prefixed-time local intermittent sampling synchronization of stochastic multicoupling delay reaction-diffusion dynamic networks, IEEE Transact. Neur. Networks Learn. Syst., 35 (2024), 718–732. https://doi.org/10.1109/TNNLS.2022.3176648 doi: 10.1109/TNNLS.2022.3176648
![]() |
[8] |
G. B. Folland, A Fundamental solution for a subelliptic operators, Bull. Amer. Math. Soc., 79 (1979), 373–376. https://doi.org/10.1090/S0002-9904-1973-13171-4 doi: 10.1090/S0002-9904-1973-13171-4
![]() |
[9] |
G. B. Folland, E. M. Stein, Estimates for the complex and analysis on the Heisenberg group, CPAM, 27 (2006), 429–522. https://doi.org/10.1002/cpa.3160270403 doi: 10.1002/cpa.3160270403
![]() |
[10] |
N. Garofalo, D. M. Nhieu, Isoperimetric and Sobolev inequalities for Carnot-Carathéodory spaces and the existence of minimal surfaces, CPAM, 49 (1996), 11081–1144. https://doi.org/10.1002/(SICI)1097-0312(199610)49:10<1081::AID-CPA3>3.0.CO;2-A doi: 10.1002/(SICI)1097-0312(199610)49:10<1081::AID-CPA3>3.0.CO;2-A
![]() |
[11] | D. Gilbarg, N. S. Trudinger, Elliptic partial differential equations of second order, 2 Eds., Berlin: Springer, 1998. https://doi.org/10.1007/978-3-642-61798-0 |
[12] |
L. Hörmander, Hypoelliptic second order differential equations, Acta Math., 119 (1967), 147–171. https://doi.org/10.1007/BF02392081 doi: 10.1007/BF02392081
![]() |
[13] |
D. Jerison, The Poincare inequality for vector fields satisfying Hörmander's condition, Duke Math. J., 53 (1986), 503–523. https://doi.org/10.1215/s0012-7094-86-05329-9 doi: 10.1215/s0012-7094-86-05329-9
![]() |
[14] |
Y. Jia, P. Xue, Effects of the self- and cross-diffusion on positive steady states for a generalized predator-prey system, Nonlinear Anal. RWA., 32 (2016), 229–241. https://doi.org/10.1016/j.nonrwa.2016.04.012 doi: 10.1016/j.nonrwa.2016.04.012
![]() |
[15] |
F. Kong, Q. Zhu, K. H. Reza, Fixed-time periodic stabilization of discontinuous reaction-diffusion Cohen-Grossberg neural networks, Neur. Networks, 166 (2023), 354–365. https://doi.org/10.1016/j.neunet.2023.07.017 doi: 10.1016/j.neunet.2023.07.017
![]() |
[16] |
Y. Lou, W. M. Ni, Diffusion, self-diffusion and cross-diffusion, J. Diff. Equ., 131 (1996), 79–131. https://doi.org/10.1006/jdeq.1996.0157 doi: 10.1006/jdeq.1996.0157
![]() |
[17] |
G. Z. Lu, Weighted Poincaré and Sobolev inequalities for vector fields satisfying Hörmander's condition and applications, Rev. Mat. Iberoam., 8 (1992), 367–439. https://doi.org/10.4171/rmi/129 doi: 10.4171/rmi/129
![]() |
[18] |
M. Mimura, Stationary pattern of some density-dependent diffusion system with competitive dynamics, Hiroshima Math. J., 11 (1981), 621–635. https://doi.org/10.32917/hmj/1206133994 doi: 10.32917/hmj/1206133994
![]() |
[19] |
M. Mimura, K. Kawasaki, Spatial segregation in competitive interaction-diffusion equations, J. Math. Biol., 9 (1980), 49–64. https://doi.org/10.1007/BF00276035 doi: 10.1007/BF00276035
![]() |
[20] |
M. Mimura, Y. Nishiura, A. Tesei, T. Tsujikawa, Coexistence problem for two competing species models with density-dependent diffusion, Hiroshima Math. J., 14 (1984), 425–449. https://doi.org/10.32917/hmj/1206133048 doi: 10.32917/hmj/1206133048
![]() |
[21] | A. Okubo, Diffusion and ecological problems: mathematical models, Berlin: Springer, 1980. https://doi.org/10.1007/978-1-4757-4978-6 |
[22] |
N. Shigesada, K. Kawasaki, E. Teramoto, Spatial segregation of interacting species, J. Theoret. Biol., 79 (1979), 83–99. https://doi.org/10.1016/0022-5193(79)90258-3 doi: 10.1016/0022-5193(79)90258-3
![]() |
[23] | G. Tralli, Harnack inequalities in sub-Riemannian settings, Alma Mater Studiorum Università di Bologna, Dottorato di ricerca in Matematica, 2013. https://doi.org/10.6092/unibo/amsdottorato/5702 |
[24] | K. D. Tran, Regularity of solutions for quasilinear subelliptic equations on Heisenberg group, California: Irvine ProQuest Dissertations & Theses, 2011. |
[25] |
Y. Wang, X. Wang, Subelliptic equations with singular nonlinearities in the heisenberg group, Bound. Value Probl., 2018 (2018), 7. https://doi.org/10.1186/s13661-018-0925-y doi: 10.1186/s13661-018-0925-y
![]() |
[26] |
X. Wang, P. Niu, X. Cui, A Liouville type theorem to an extension problem relating to the Heisenberg group, Commun. Pure Appl. Anal., 17 (2018), 2379–2394. https://doi.org/10.3934/cpaa.2018113 doi: 10.3934/cpaa.2018113
![]() |
[27] |
H. Wang, Q. Zhu, Global stabilization of a class of stochastic nonlinear time-delay systems with SISS inverse dynamics, IEEE Trans. Automat. Control, 65 (2020), 4448–4455. https://doi.org/10.1109/TAC.2020.3005149 doi: 10.1109/TAC.2020.3005149
![]() |
[28] |
P. Xue, Y. Jia, C. Ren, X. Li, Non-constant positive solutions of a general Gause-type predator-prey system with self- and cross-diffusions, Math. Model. Nat. Phenom., 16 (2021), 25. https://doi.org/10.1051/mmnp/2021017 doi: 10.1051/mmnp/2021017
![]() |
[29] | Q. Zhu, Event-triggered sampling problem for exponential stability of stochastic nonlinear delay systems driven by Levy processes, IEEE Trans. Automat. Control, 2024. https://doi.org/10.1109/TAC.2024.3448128 |