We deal with the nonlinear Kirchhoff problem
−(a+b∫R3|∇u|2dx)Δu+V(x)u=f(u), x∈R3,
where a is a positive constant, b>0 is a parameter, the potential function V is allowed to change its sign, and the nonlinearity f∈C(R,R) exhibits subcritical growth. Under some suitable conditions on V, we first prove that the problem has a positive ground state solution for all b>0. Then, by using a more general global compactness lemma and a sign-changing Nehari manifold, combined with the method of constructing a sign-changing (PS)c sequence, we show the existence of a least energy sign-changing solution for b>0 that is sufficiently small. Moreover, the asymptotic behavior b↘0 is established.
Citation: Yan-Fei Yang, Chun-Lei Tang. Positive and sign-changing solutions for Kirchhoff equations with indefinite potential[J]. Communications in Analysis and Mechanics, 2025, 17(1): 159-187. doi: 10.3934/cam.2025008
[1] | Xiao Qing Huang, Jia Feng Liao . Existence and asymptotic behavior for ground state sign-changing solutions of fractional Schrödinger-Poisson system with steep potential well. Communications in Analysis and Mechanics, 2024, 16(2): 307-333. doi: 10.3934/cam.2024015 |
[2] | Fangyuan Dong . Multiple positive solutions for the logarithmic Schrödinger equation with a Coulomb potential. Communications in Analysis and Mechanics, 2024, 16(3): 487-508. doi: 10.3934/cam.2024023 |
[3] | Wenqian Lv . Ground states of a Kirchhoff equation with the potential on the lattice graphs. Communications in Analysis and Mechanics, 2023, 15(4): 792-810. doi: 10.3934/cam.2023038 |
[4] | Xin Qiu, Zeng Qi Ou, Ying Lv . Normalized solutions to nonautonomous Kirchhoff equation. Communications in Analysis and Mechanics, 2024, 16(3): 457-486. doi: 10.3934/cam.2024022 |
[5] | Yangyu Ni, Jijiang Sun, Jianhua Chen . Multiplicity and concentration of normalized solutions for a Kirchhoff type problem with $ L^2 $-subcritical nonlinearities. Communications in Analysis and Mechanics, 2024, 16(3): 633-654. doi: 10.3934/cam.2024029 |
[6] | Rui Sun, Weihua Deng . A generalized time fractional Schrödinger equation with signed potential. Communications in Analysis and Mechanics, 2024, 16(2): 262-277. doi: 10.3934/cam.2024012 |
[7] | Chen Yang, Chun-Lei Tang . Sign-changing solutions for the Schrödinger-Poisson system with concave-convex nonlinearities in $ \mathbb{R}^{3} $. Communications in Analysis and Mechanics, 2023, 15(4): 638-657. doi: 10.3934/cam.2023032 |
[8] | Xiao Han, Hui Wei . Multiplicity of the large periodic solutions to a super-linear wave equation with general variable coefficient. Communications in Analysis and Mechanics, 2024, 16(2): 278-292. doi: 10.3934/cam.2024013 |
[9] | Henryk Żoła̧dek . An example in Hamiltonian dynamics. Communications in Analysis and Mechanics, 2024, 16(2): 431-447. doi: 10.3934/cam.2024020 |
[10] | Qi Li, Yuzhu Han, Bin Guo . A critical Kirchhoff problem with a logarithmic type perturbation in high dimension. Communications in Analysis and Mechanics, 2024, 16(3): 578-598. doi: 10.3934/cam.2024027 |
We deal with the nonlinear Kirchhoff problem
−(a+b∫R3|∇u|2dx)Δu+V(x)u=f(u), x∈R3,
where a is a positive constant, b>0 is a parameter, the potential function V is allowed to change its sign, and the nonlinearity f∈C(R,R) exhibits subcritical growth. Under some suitable conditions on V, we first prove that the problem has a positive ground state solution for all b>0. Then, by using a more general global compactness lemma and a sign-changing Nehari manifold, combined with the method of constructing a sign-changing (PS)c sequence, we show the existence of a least energy sign-changing solution for b>0 that is sufficiently small. Moreover, the asymptotic behavior b↘0 is established.
In the past decades, the following Kirchhoff problem
−(a+b∫R3|∇u|2dx)Δu+V(x)u=f(x,u), x∈R3 | (1.1) |
has attracted considerable attention. As we know, the following Dirichlet problem is one of the important deformations of Equation (1.1), which can be degenerated from (1.1). That is, if V(x)=0 and Ω is a bounded subset of R3, then Equation (1.1) will become
{−(a+b∫Ω|∇u|2dx)Δu=f(x,u)inΩ,u=0on∂Ω. | (1.2) |
Problem (1.2) corresponds to
utt−(a+b∫Ω|∇u|2dx)Δu=f(x,u), |
which was advanced by Kirchhoff in [1]. As a generalization of the classical D'Alembert wave equation of the free vibration of elastic strings, the Kirchhoff model considers the changes of string length caused by lateral vibrations, which has important practical significance. For more mathematical and physical background about the Kirchhoff equations, we direct readers to [2,3] and the references quoted within them.
When b=0, Equation (1.1) degenerates to the Schrödinger problem
−aΔu+V(x)u=f(x,u), x∈RN, | (1.3) |
which has been studied in the past few decades; see, for instance, [4,5,6,7,8,9] and the references therein. One interesting characteristic is the potential V change sign on RN. In [6], Bahrouni, Ounaies, and Radulescu showed the existence of infinitely many solutions for Equation (1.3) with f(x,u)=a(x)|u|q−1u and 0<q<1. Further, Furtado, Maia, and Medeiros [10] investigated the Schrödinger equation (1.3) with a=1 and f(x,u)=f(u). Concretely, the nonlinearity f∈C1(R,R) is superlinear with subcritical growth. In addition, it verifies the Ambrosetti-Rabinowitz condition: for some θ>2, there is
(ˆf)0<θF(t)≤tf(t) for all t≠0, where F(t)=∫t0f(s)ds. |
Besides, V satisfies the following assumptions:
(V0) V∈Ltloc(R3) for some t>32;
(V1) 0<V∞:=lim|x|→∞V(x)<+∞;
(V2) ∫R3|V−(x)|32dx<S32, where V−:=max{−V,0} and S is the best constant for the Sobolev embedding, given by
S:=infu∈D1,2(R3)∖{0}∫R3|∇u|2dx(∫R3|u|6dx)13; |
(V3) V(x)≤V∞ for all x∈R3 and V≢V∞;
(V4) there exist γ>0 and CV>0 such that
V(x)≤V∞−CVe−γ|x|,for all x∈R3. |
By using variational methods and the concentration-compactness principle, they obtained a positive ground state solution, and also a nodal solution. Therefore, we know that Equation (1.3) has a least energy sign-changing solution v0∈M0 such that I0(v0)=c0,2:=infu∈M0I0(u), where M0:={u∈X:u±≠0, ⟨I′0(u),u+⟩=⟨I′0(u),u−⟩=0} with I0(u)=12∫R3(a|∇u|2+V(x)u2)dx−∫R3F(u)dx. For more results about problem (1.3) with indefinite potential, see [10,11,12,13] and the references therein.
When b≠0, due to the presence of ∫R3|∇u|2dxΔu, problem (1.1) becomes a nonlocal problem, which also brings some essential difficulties for our study. In the past, problems similar to (1.1) have attracted a lot of interest, and so there are many results. For instance, the existence of positive, ground state, and sign-changing solutions for problem (1.1) with various potential V and nonlinearity f has been extensively studied; see [14,15,16,17,18,19,20,21]. In a recent paper [22], Ni, Sun, and Chen also obtained the existence and multiplicity of normalized solutions for a Kirchhoff type problem by using minimization techniques and Lusternik-Schnirelmann theory.
In what follows, we are particularly interested in the case when the potential V involved in problem (1.1) is indefinite. When V can change its sign and satisfies conditions (V0)−(V2) and (V4), Batista and Furtado [23] studied problem (1.1) with f(x,u)=a(x)|u|p−2u, that is,
−(a+b∫R3|∇u|2dx)Δu+V(x)u=a(x)|u|p−2u, x∈R3, | (1.4) |
where 4<p<6 and a satisfies the following assumptions:
(a0) a∈L∞(R3);
(a1) there exist Ca, θ0>0 such that
a(x)⩽a∞−Cae−θ0|x|,for a.e. x∈R3, |
where
a∞:=lim|x|→+∞a(x)>0. |
Via the constraint variational methods and the quantitative deformation lemma, the authors not only obtained a non-negative ground state solution, but also a sign-changing solution of Equation (1.4). Here, we point out that the proof of the existence of sign-changing solutions depends on the radial symmetry of V in their paper. Indeed, once the potential V is radial, one can overcome the lack of compactness by considering in the radial subspace. However, if V is not radial, it makes no sense to restrict the problem to spaces of radial functions as the authors did in [23]. Besides, we know the embedding H1(R3)↪Lp(R3) is not compact for 2<p<6. Therefore, one may ask if there will be a sign-changing solution for problem (1.4) or problem (1.1) when the indefinite potential V is not radially symmetric and f is a general nonlinearity.
Next, we will provide an answer to the questions raised above. Specifically, we are going to investigate the problem
−(a+b∫R3|∇u|2dx)Δu+V(x)u=f(u), x∈R3, | (1.5) |
where a,b>0 and V is a sign-changing potential.
Moreover, we shall impose that f∈C(R,R) is odd. In addition, f also satisfies the following assumptions:
(f0) limt→0f(t)t=0;
(f1) lim|t|→∞f(t)|t|5=0;
(f2) lim|t|→∞F(t)|t|4=+∞, where F(t):=∫t0f(s)ds;
(f3) f(t)|t|3 is a non-decreasing function for t∈R∖{0};
(f4) F(t)≥0, for all t∈R.
Through (f0) and (f1), it can be known that for any ε>0, there is a positive constant Cε which makes
|f(t)|≤ε(|t|+|t|5)+Cε|t|q−1,for any t∈R | (1.6) |
where q∈(2,6). Then, let G(t):=14f(t)t−F(t). By using (f3), we can easily obtain
0≤G(t1)≤G(t2),for any 0≤t1≤t2. | (1.7) |
In fact, for any t2≥t1≥0, from (f3), one has f(t1)t31≤f(t)t3≤f(t2)t32 for all t1≤t≤t2. Then,
G(t1)=14f(t1)t1+∫t2t1f(t)dt−F(t2)=14f(t1)t1+∫t2t1f(t)t3t3dt−F(t2)≤14f(t1)t1+∫t2t1f(t2)t32t3dt−F(t2)≤14f(t1)t1+14f(t2)t32(t42−t41)−F(t2)≤14{f(t1)t1+f(t2)t2−f(t1)t1}−F(t2)=14f(t2)t2−F(t2)=G(t2). |
Before presenting our main results, we first discussed the basic framework in the space
X:={u∈H1(R3):∫R3V(x)u2dx<+∞} |
and the corresponding norm is given by
‖u‖X:=(∫R3(a|∇u|2+V(x)u2)dx)12,for any u∈X. |
Next, we present the main results.
Theorem 1.1. Assume that (V0)−(V3) and (f0)−(f3) are satisfied. Then, problem (1.5) has a positive ground state solution.
In the proof, we need to overcome the problems stemming from the lack of compactness of Sobolev embedding in the whole space R3. To solve this problem, we first analyze the relationship between the energy functional's minimax level and that of the limit problem. Subsequently, we obtain a positive ground state solution by applying a more general global compactness lemma (see Lemma 3.1).
In the second result, we mainly focused on researching the existence of least-energy sign-changing solutions. In order to obtain the result, we will use the method in [24].
Theorem 1.2. Assume that f satisfies (f0)−(f4). Assume also that V satisfies (V0)−(V2), and there exist positive constants M,C, and γ such that, for |x|≥M,
(V′4)V(x)≤V∞−C1+|x|γ. |
Then, there exists b∗>0 small enough such that Equation (1.5) possesses a least energy sign-changing solution for every b∈(0,b∗).
Remark 1.3. Here, we would like to provide an example of nonlinearity f, which is odd and satisfies the conditions (f0)−(f4), as shown below:
f(t)=l∑i=1ai|t|bit,for any t∈R, |
where ai>0 and 2<bi<4 for every i∈{1,2,…,l}.
In addition, we can find that a more obvious example that satisfies conditions (V0)−(V2),(V′4) is to take V(x)=V∞−C1+|x|γ, where C and γ are positive given by (V′4). Alternatively, we can give another example below that satisfies our hypothesis. That is, the potential function V is given by
V(x)={−α|x|β,if |x|<1;|x|21+|x|2,if |x|≥1, |
where β∈(0,2) and α>0 is a sufficiently small.
Remark 1.4. Compared with [10], we apply weaker conditions (f2) and (f4) instead of the Ambrosetti-Rabinowitz condition to investigate the existence of sign-changing solutions. Moreover, condition (f4) can be used to construct a sign-changing (PS)cb,2; please refer to Section 5 for the detailed process. We note that condition (f4) is only used here. Besides, it is not necessary for f to be differentiable. By using the quantitative deformation lemma, we can prove that the minimizer on the Nehari manifold is a critical point of the energy functional Ib given in Section 2, please refer to Theorem 1.1 below.
Remark 1.5. In this result, we apply the weaker condition (V′4) instead of the condition (V4) given in [23], which makes our results applicable to a wider range of potential functions. Furthermore, the potential function V does not need to be radial in our paper, which is different from [23]. Moreover, for the proof of Theorem 1.2, compared to [10], the essential problem we face is that the existence of the nonlocal term poses some difficulties to energy estimates. Here, the condition (V′4) and the restriction on the range of the parameter b are crucial for estimating energy (see Lemma 5.1). After that, we can restore the compactness of a bounded Palais–Smale sequence at a certain level with the help of a global compactness lemma, thereby obtaining the conclusion of Theorem 1.2.
Theorem 1.6. Assume that ubn are the least energy sign-changing solutions of Equation (1.5) obtained in Theorem 1.2. Then, for any sequence {bn} with bn↘0 as n→∞, there exists a subsequence, still denoted by {bn}, such that ubn→u0 in X as n→∞. Moreover, u0 is a least energy sign-changing solution of Equation (1.3) with f(x,u)=f(u).
Remark 1.7. In this paper, we not only weaken the potential condition, but also increase asymptotic behavior, which is different from [23]. In particular, our result regarding asymptotic behavior is new. The asymptotic research enriches the results of our paper, and at the same time, the sign-changing solution is closely related to the results of the Schrödinger equation in [10]. In practical terms, we study their asymptotic behavior as b↘0 under assumptions (V0)−(V2) and (V4), and we show that they converge to a least energy sign-changing solution of the Schrödinger equation (1.3) with f(x,u)=f(u), b↘0.
Now, we introduce the organizational structure of this paper. We first provide the notations and some necessary lemmas in Section 2. Then, in Section 3, our main job is to establish a more general global compactness result, which will be well applied in the proof of our main theorems. In Section 4, we first give energy estimates, then Theorem 1.1 is verified. Finally, we prove Theorem 1.2. After that, Theorem 1.6 is verifed in Section 5.
Next, we first give some notations and necessary lemmas, which are very helpful for us in proving the main theorems.
● "⇀" and "→" depict the weak and strong convergence, sequentially.
● |u|p=(∫R3|u|pdx)1p denotes the norm in Lp(R3) for p∈[1,+∞).
● Let H1(R3) be the Hilbert space with respect to the norm ‖u‖2H1:=∫R3(|∇u|2+u2)dx.
● We use |Ω| to represent the Lebesgue measure of the set Ω.
● o(1) denotes a quantity which goes to zero as n→∞.
● C, Ci (i=1,2,…) represent different positive constants.
● We denote V=V+−V− with V±:=max{±V,0}.
Lemma 2.1. Under the conditions of (V1) and (V2), the quadratic form
u↦∫R3(a|∇u|2+V(x)u2)dx | (2.1) |
defines a norm in H1(R3), which is equivalent to the usual one.
Proof. Due to (V1), one has that there is R>0, which satisfies
V∞2≤V+≤3V∞2,for any x∈R3∖BR(0), | (2.2) |
where BR(0):={x∈R3:|x|<R}. In general, we take a=1. Then, using the Hölder inequality, the definition of S, and (2.2), one can ascertain
∫R3(|∇u|2+Vu2)dx=∫R3(|∇u|2+V+u2)dx−∫R3V−u2dx≤∫R3|∇u|2dx+∫BR(0)V+u2dx+∫|x|≥RV+u2dx≤∫R3|∇u|2dx+|V+|L3/2(BR(0))|u|26+3V∞2∫|x|≥Ru2dx≤(1+S−1|V+|L3/2(BR(0)))∫R3|∇u|2dx+3V∞2∫R3u2dx≤max{1+S−1|V+|L3/2(BR(0)),3V∞2}∫R3(|∇u|2+u2)dx. |
Furthermore, from [10, Lemma 2.1], one deduces that there is C1>0, which holds the inequality ∫R3(|∇u|2+V+u2)dx≥C1∫R3(|∇u|2+u2)dx. Moreover, from (V2), one has
|∫R3V−u2dx|≤∫R3|V−|u2dx≤|V−|32|u|26≤S−1|V−|32∫R3|∇u|2dx. |
Hence, combining with V=V+−V−, it can be obtained that
∫R3(|∇u|2+Vu2)dx≥∫R3|∇u|2dx+∫R3V+u2dx−S−1|V−|32∫R3|∇u|2dx≥min{1−S−1|V−|32,1}∫R3(|∇u|2+V+u2)dx≥min{1−S−1|V−|32,1}C1∫R3(|∇u|2+u2)dx |
and the lemma is completed.
Remark 2.2. From Lemma 2.1, it is obvious to see that the norm given by (2.1) is equivalent to the usual norm of H1(R3). Moreover, we know that the embeddings H1(R3)↪Lp(R3) (see [9, Theorem 1.8]) and H1(R3)↪Lqloc(R3) (see [9, Theorem 1.9]) are continuous and compact, respectively, where p∈[2,6] and q∈[1,6). Therefore, the embedding X↪Lp(R3) is also continuous for all p∈[2,6]. Besides, one can see that the continuity of the embedding mentioned above can be represented by the following inequalities:
|u|pp≤Spp‖u‖pX,for any p∈[2,6], |
in which Spp>0 is a constant.
Define the energy functional Ib:X→R by
Ib(u)=12‖u‖2X+b4(∫R3|∇u|2dx)2−∫R3F(u)dx. |
One can see that u↦∫F(u)dx is well defined on X, so Ib(u) is also well defined. Through discussions, one can deduce that Ib∈C1(X,R). Moreover, for any v∈H1(R3),
⟨I′b(u),v⟩=∫R3(a∇u⋅∇v+V(x)uv)dx+b∫R3|∇u|2dx∫R3∇u⋅∇vdx−∫R3f(u)vdx. |
The limit problem associated with (1.5) is the autonomous problem below:
−(a+b∫R3|∇u|2dx)Δu+V∞u=f(u), x∈R3. | (P∞) |
The functional corresponding to Equation (P∞) is
Ib,∞(u)=12∫R3(a|∇u|2+V∞u2)dx+b4(∫R3|∇u|2dx)2−∫R3F(u)dx. |
Let
cb,1:=infu∈NbIb(u)andcb,∞:=infu∈Nb,∞Ib,∞(u), | (2.3) |
where
Nb:={u∈X∖{0}:⟨I′b(u),u⟩=0}andNb,∞:={u∈X∖{0}:⟨I′b,∞(u),u⟩=0}. |
In addition, we define
cb,2:=infu∈MbIb(u), |
where
Mb:={u∈X:u±≠0, ⟨I′b(u),u+⟩=⟨I′b(u),u−⟩=0}. |
Next, we will provide some important lemmas to prove Theorem 1.1 and Theorem 1.2.
Lemma 2.3. If {un}⊂Mb is a minimizing sequence for Ib, then C1≤‖u±n‖X≤C2 for some C1, C2>0.
Proof. Let {un}⊂Mb and Ib(un)→m as n→∞. On one hand, due to (f0), (f1), and the Sobolev inequality, one has
‖u±n‖2X≤‖u±n‖2X+b∫R3|∇un|2dx∫R3|∇u±n|2dx=∫R3f(un)u±ndx≤∫R3(ε|un|+Cε|un|5)u±ndx=ε∫R3|u±n|2dx+Cε∫R3|u±n|6dx≤εS22‖u±n‖2X+CεS66‖u±n‖6X. |
When ε is sufficiently small to make (1−εS22)>0, we can obtain that ‖u±n‖X≥C1 for some C1>0. On the other hand, in light of {un}⊂Mb⊂Nb, we have ⟨I′b(un),un⟩=0. In virtue of (1.7), one can arrive at
m+o(1)=Ib(un)=Ib(un)−14⟨I′b(un),un⟩=14‖un‖2X+∫R3(14f(un)un−F(un))dx≥14‖un‖2X, |
which means that m>0, so one can deduce that {un} is bounded in X. Namely, ‖u±n‖X≤C2 for some C2>0. Therefore, C1≤‖u±n‖X≤C2.
Remark 2.4. If {un}⊂X be a (PS)c sequence, then by using (1.7) and similar arguments to Lemma 2.3, one can get ‖un‖X≤C2 for some C2>0.
Similar to the discussion in [25], one can derive Lemma 2.5 and Lemma 2.6. Here, we omit the proof process.
Lemma 2.5. Assume that (V0)−(V2) and (f0)−(f3) hold. If u∈X with u±≠0, then there exists a unique pair (su,tu)∈(0,+∞)×(0,+∞) such that suu++tuu−∈Mb. Moreover,
Ib(suu++tuu−)=maxs,t≥0Ib(su++tu−). |
Lemma 2.6. Assume that (V0)−(V2) and (f0)−(f3) hold. If u∈X with u≠0, then there exists a unique su>0 such that suu∈Nb. Moreover,
Ib(suu)=maxs≥0Ib(su). |
Lemma 2.7. cb,∞ can be obtained by some positive and radially symmetric function ˉu∈Nb,∞, that corresponds to the ground state solution of (P∞) (see [26]). Furthermore, if f is odd, then for every 0<δ<√V∞, there is C=C(δ)>0 that satisfies
0<ˉu(x)≤Ce−δα|x|, for any x∈R3, | (2.4) |
where α=(a+b∫R3|∇ˉu|2dx)12.
Proof. The existence of the ground state solutions of (P∞) was proved in [26, Proposition 2.4]. For the properties of solutions of (P∞), see [27, Proposition 1.1].
Now, we will give a more general global compactness lemma, which is very useful.
Lemma 3.1. Let {un}⊂X be a sequence such that
Ib(un)→c and I′b(un)→0,as n→∞. |
Then, there exist u0∈H1(R3) and A∈R+, such that J′A(u0)=0, where
JA(u)=a+bA2∫R3|∇u|2dx+12∫R3V(x)u2dx−∫R3F(u)dx, |
and either
(i) un→u0 in X, or
(ii) there exists a number k∈N+, k sequences of points {yjn}⊂R3 with |yjn|→+∞, 1≤j≤k, and k functions {u1,u2,…,uk}⊂H1(R3), which are nontrivial weak solutions to
−(a+bA)Δu+V∞u=f(u) | (3.1) |
and
c+bA24=JA(u0)+k∑j=1J∞A(uj), | (3.2) |
‖un−u0−k∑j=1uj(⋅−yjn)‖X→0, |
A=|∇u0|22+k∑j=1|∇uj|22, | (3.3) |
where
J∞A(u)=a+bA2∫R3|∇u|2dx+12∫R3V∞u2dx−∫R3F(u)dx. |
Proof. In view of Remark 2.4, we know {un} is bounded in X. Then, there are u0∈X and A∈R+ that satisfy
un⇀u0 in X and ∫R3|∇un|2dx→A. | (3.4) |
Due to (3.4) and I′b(un)→0 as n→∞, we arrive at
∫R3(a∇u0⋅∇φ+V(x)u0φ)dx+bA∫R3∇u0⋅∇φdx−∫R3f(u0)φdx=0,∀φ∈X. |
That is, J′A(u0)=0. On the other hand, it is clear to see that
JA(un)=a+bA2∫R3|∇un|2dx+12∫R3V(x)u2ndx−∫R3F(un)dx=a2∫R3|∇un|2dx+12∫R3V(x)u2ndx+b4(∫R3|∇un|2dx)2−∫R3F(un)dx+bA24+o(1)=Ib(un)+bA24+o(1). | (3.5) |
Besides, for all φ∈C∞0(R3), one has
⟨J′A(un),φ⟩=(a+bA)∫R3∇un⋅∇φdx+∫R3V(x)unφdx−∫R3f(un)φdx=∫R3(a∇un⋅∇φ+V(x)unφ)dx+b∫R3|∇un|2dx∫R3∇un⋅∇φdx−∫R3f(un)φdx+o(1)=⟨I′b(un),φ⟩+o(1). | (3.6) |
It follows from (3.5) and (3.6) that
JA(un)→c+bA24 and J′A(un)→0,as n→∞. |
Next, we will demonstrate in three steps and provide detailed proof process.
Step 1. Letting u1n:=un−u0, we can obtain
(a1) J∞A(u1n)→c+bA24−JA(u0),
(b1) ⟨(J∞A)′(u1n),u1n⟩=⟨J′A(un),un⟩−⟨J′A(u0),u0⟩+o(1)=o(1).
To prove (a1), we can use the weak convergence of {un} and [28, Lemma 3] to conclude that
|∇u1n|22=|∇un|22−|∇u0|22+o(1), | (3.7) |
|u1n|22=|un|22−|u0|22+o(1), | (3.8) |
∫R3F(u1n)dx=∫R3F(un)dx−∫R3F(u0)dx+o(1), | (3.9) |
∫R3f(u1n)u1ndx=∫R3f(un)undx−∫R3f(u0)u0dx+o(1). | (3.10) |
Moreover, by virtue of (V1), we know ∀ε>0, ∃R>0 such that |V∞−V(x)|<ε on R3∖BR(0). Hence, it holds that
|∫R3(V∞−V)(u2n−u20)dx|≤∫BR(0)|V∞−V||u2n−u20|dx+ε∫R3∖BBR(0)|u2n−u20|dx≤C1|un−u0|L2(BBR(0))+C2ε=C2ε+o(1). |
From the arbitrariness of ε, it can be concluded that
limn→∞∫R3(V∞−V)(u2n−u20)dx=0. | (3.11) |
Using (3.7)-(3.9) and (3.11), we can show that
J∞A(u1n)−JA(un)+JA(u0)=a+bA2(|∇u1n|22−|∇un|22+|∇u0|22)+12∫R3V(x)(u20−u2n)dx+12∫R3V∞(u1n)2dx+∫R3(F(un)−F(u0)−F(u1n))dx=o(1). |
That is, (a1) is correct. As for (b1), by using a similar argument as before and (3.10), it is sufficient to get (b1). We omit the details here. Furthermore, by (a1), we can obtain that J∞A(u1n)≥0.
Step 2. Define
δ(1):=lim supn→+∞supy∈R3∫B1(y)|u1n|2dx. |
Case 1 (Vanishing): δ(1)=0. That is, as n→∞,
supy∈R3∫B1(y)|u1n|2dx→0. |
From the P.L. Lions lemma in [9], one has u1n→0 in Lt(R3) for any t∈(2,6). Thus, we deduce that limn→+∞∫R3f(u1n)u1ndx=0 by (1.6). Besides, it is easy to get ∫R3(V(x)−V∞)(u1n)2dx=o(1). Hence, by (b1), we can conclude ‖u1n‖X→0 as n→∞.
Case 2 (Non-vanishing): δ(1)>0. Assume that there exists {y1n}⊂R3 such that
∫B1(y1n)|u1n|2dx>δ(1)2>0. |
We now define a new sequence w1n:=u1n(⋅+y1n). It is easy to get that {w1n} is bounded in X. Moreover, we suppose that w1n⇀u1 in X. Since
∫B1(0)|w1n|2dx=∫B1(y1n)|u1n|2dx>δ(1)2>0, |
it follows from the Sobolev embedding that u1≠0. Moreover, u1n⇀0 in X implies that {y1n} is unbounded. That is, |y1n|→∞ as n→∞. Furthermore, we can show (J∞A)′(u1)=0.
Step 3. Setting u2n:=u1n−u1(⋅−y1n), we can check that
(a2) J∞A(u2n)→c+bA24−JA(u0)−J∞A(u1),
(b2) ⟨(J∞A)′(u2n),u2n⟩=⟨J′A(un),un⟩−⟨J′A(u0),u0⟩−⟨(J∞A)′(u1),u1⟩+o(1)=o(1).
Similar to Step 2, define
δ(2):=lim supn→+∞supy∈R3∫B1(y)|u2n|2dx. |
If δ(2)=0, we have ‖u2n‖X→0 as n→∞. That is, ‖un−u0−u1(⋅−y1n)‖X→0 as n→∞. By (3.7) and (a2), we have A=|∇u0|22+|∇u1|22 and c+bA24=JA(u0)+J∞A(u1). Furthermore, we know that J∞A(u2n)=o(1). If δ(2)>0, as the arguments as above, we know that there exists {y2n}⊂R3 unbounded, and a sequence w2n:=u2n(⋅+y2n) that satisfies w2n⇀u2 in X and u2≠0. Besides, we can obtain (J∞A)′(u2)=0. Iterating the above process, we can show that
ujn=uj−1n−uj−1(⋅−yj−1n) |
with |yjn|→∞ and
wj−1n=uj−1n(⋅+yj−1n)⇀uj−1 in X, |
where uj is the nontrivial weak solution of Equation (3.1). Moreover, we can conclude that
J∞A(ujn)=c+bA24−JA(u0)−j−1∑i=1J∞A(ui)+o(1). |
Noticing that ui is the nontrivial weak solution of Equation (3.1), in view of (1.7), we can obtain J∞A(ui)>0. Besides, similar to the above discussion, we know that when δ(j)=0, J∞A(ujn)=o(1). Hence, there is some finite constant k∈N. Moreover, the above process will stop after k iterations. Namely, the proof is completed.
Corollary 3.2. The functional Ib satisfies (PS)c condition for c∈(0,cb,∞).
Proof. Let {un}⊂X be a (PS)c sequence for c∈(0,cb,∞). Then,
Ib(un)→c∈(0,cb,∞) and I′b(un)→0,as n→∞. |
We only need to prove that {un} has a convergent subsequence in X next. From Remak 2.4, we first conclude that {un} is bounded in X. Hence, there is a subsequence of {un}, still denoted as {un}. Besides, there is also u0∈H1(R3) that satisfies un⇀u0 in X. If it is strongly convergent, then the proof is completed. If un↛u0 in X, from Lemma 3.1, there exist A=|∇u0|22+∑kj=1|∇uj|22, k∈N and {yjn}⊂R3 with |yjn|→+∞ for j=1,2,…,k and uj∈H1(R3) such that
J′A(u0)=0,‖un−u0−k∑j=1uj(⋅−yjn)‖X→0,c+bA24=JA(u0)+k∑j=1J∞A(uj), |
where uj are nontrivial critical points of J∞A for j=1,2,…,k.
We first give the following two claims.
Claim 1: If u0≠0 and there exists t0>0 such that t0u0∈Nb, then we claim that t0≤1.
Since t0u0∈Nb and J′A(u0)=0, we can obtain that
‖t0u0‖2X+b(∫R3|∇(t0u0)|2dx)2=∫R3f(t0u0)(t0u0)dx | (3.12) |
and
(a+bA)∫R3|∇u0|2dx+∫R3V(x)u20dx=∫R3f(u0)u0dx. | (3.13) |
It follows from (3.3), (3.12), and (3.13) that
(1−1t20)‖u0‖2X≤∫R3(f(u0)(u0)3−f(t0u0)(t0u0)3)|u0|4dx. | (3.14) |
From (f3) and (3.14), it is easy to obtain t0≤1. Therefore, the claim is proved.
Claim 2: If u0≠0, we claim that
JA(u0)≥cb,1+bA4∫R3|∇u0|2dx. | (3.15) |
Combining t0≤1 and (1.7), we can arrive at
∫R3(14f(t0u0)(t0u0)−F(t0u0))dx≤∫R3(14f(u0)u0−F(u0))dx. |
Hence, we can obtain
JA(u0)=JA(u0)−14⟨J′A(u0),u0⟩=a+bA4∫R3|∇u0|2dx+14∫R3V(x)u20dx+∫R3(14f(u0)u0−F(u0))dx≥t204∫R3(a|∇u0|2+V(x)u20)dx+bA4∫R3|∇u0|2dx+∫R3(14f(t0u0)(t0u0)−F(t0u0))dx=Ib(t0u0)−14⟨I′b(t0u0),t0u0⟩+bA4∫R3|∇u0|2dx=Ib(t0u0)+bA4∫R3|∇u0|2dx≥cb,1+bA4∫R3|∇u0|2dx. |
Therefore, the claim is proved. Similarly, since uj≠0 and (J∞A)′(uj)=0, it is easy to see that
J∞A(uj)≥cb,∞+bA4∫R3|∇uj|2dx. | (3.16) |
Now, we return to our proof. If u0≡0,k≥1, then A=∑kj=1|∇uj|22 and c+bA24=∑kj=1J∞A(uj)≥kcb,∞+bA24. Noticing that c<cb,∞, this is absurd. If u0≠0,k≥1, then from (3.2), (3.15), and (3.16), we can obtain c+bA24≥cb,1+kcb,∞+bA24. By using (f0)−(f3) and similar discussions to those in [9, Theorem 4.2], one ascertains cb,1>0. Combining c<cb,∞, we know this case cannot occur. Therefore, k=0 and the proof is completed.
Now, we prove Theorem 1.1. First, due to conditions (f0)−(f2), one can easily deduce that Ib satisfies the mountain pass geometry. Then, by conditions (f0)−(f3) and similar discussions to those in [9, Theorem 4.2], we have
cb,1=infγ∈Γmaxt∈[0,1]Ib(γ(t))=infu∈X∖{0}maxt≥0Ib(tu)>0, |
where Γ:={γ∈C([0,1],X):γ(0)=0, Ib(γ(1))<0}. Then, we show the relationship between cb,1 and cb,∞. Note that ˉu is a positive ground state solution of (P∞), so combined with Lemma 2.6, one can ascertain that there is tˉu>0, which makes tˉuˉu∈Nb.
Lemma 4.1. Assume that (V0)−(V3) and (f0)−(f3) hold. Then,
0<cb,1<cb,∞. |
Proof. We first claim that tˉu<1.
Since ˉu∈Nb,∞, one has that
∫R3(a|∇ˉu|2+V∞ˉu2)dx+b(∫R3|∇ˉu|2dx)2=∫R3f(ˉu)ˉudx. | (4.1) |
Due to tˉuˉu∈Nb, we ascertain that
t2ˉu∫R3(a|∇ˉu|2+V(x)ˉu2)dx+t4ˉub(∫R3|∇ˉu|2dx)2=∫R3f(tˉuˉu)(tˉuˉu)dx. | (4.2) |
Combining (4.1), (4.2), and (V3), we deduce that
(1t2ˉu−1)∫R3(a|∇ˉu|2+V∞ˉu2)dx>∫R3(f(tˉuˉu)(tˉuˉu)3−f(ˉu)ˉu3)|ˉu|4dx. |
If tˉu≥1, the above equation does not hold by (f3). Thus, we can obtain that tˉu<1.
It follows from (1.7), (V3), Lemma 2.6, and the above claim that
cb,1≤maxt≥0Ib(tˉu)=Ib(tˉuˉu)<t2ˉu4∫R3(a|∇ˉu|2+V∞ˉu2)dx+∫R3(14f(tˉuˉu)(tˉuˉu)−F(tˉuˉu))dx<14∫R3(a|∇ˉu|2+V∞ˉu2)dx+∫R3(14f(ˉu)ˉu−F(ˉu))dx=Ib,∞(ˉu)=cb,∞. |
The proof is completed.
Proof of Theorem 1.1. Through using [9, Theorem 1.15], one knows that there exists a sequence {un}⊂X such that
Ib(un)→cb,1andI′b(un)→0,as n→∞. |
In view of Corollary 3.2 and Lemma 4.1, the sequence {un} has a subsequence which strongly converges to u∈X. Besides, the function u satisfies Ib(u)=cb,1>0 and I′b(u)=0. We can easily get that u≠0. This indicates that u is a ground solution of Equation (1.5). Next, we prove that Equation (1.5) has a positive ground solution. First, based on f being an odd function, we can see Ib(|u|)=Ib(u)=cb,1 and |u|∈Nb. We can claim that I′b(|u|)=0 by using the deformation lemma, where f does not require differentiability. For convenience, let us note w=|u|. Then, we only need to prove I′b(w)=0.
By contradiction, we assume I′b(w)≠0. Then, there are ϱ>0 and δ>0 that satisfy
‖I′b(u)‖X≥ϱ,∀u∈X with ‖u−w‖X≤3δ. |
Let D:=(1−σ,1+σ), where σ∈(0,min{12,δ√2‖w‖X}). By using the fact that w∈Nb and the condition (f3), we have
⟨I′b(tw),tw⟩>0,if t<1 |
and
⟨I′b(tw),tw⟩<0,if t>1. |
Hence, we can take t1,t2∈D∖{1}, such that
⟨I′b(t1w),t1w⟩>0, ⟨I′b(t2w),t2w⟩<0. | (4.3) |
It follows from Lemma 2.6 that
ˉc:=max{Ib(t1w),Ib(t2w)}<Ib(w)=cb,1. | (4.4) |
For ε:=min{(cb,1−ˉc)/3,δϱ/8} and Sδ:={u∈X:‖u−w‖X≤δ}, due to the deformation lemma in [9, Lemma 2.3], one deduces that there exists η∈C([0,1]×X,X) that satisfies
(a) η(1,u)=u if u∉I−1b([cb,1−2ε,cb,1+2ε])∩S2δ;
(b) η(1,Icb,1+εb∩Sδ)⊂Icb,1−εb;
(c) Ib(η(1,u))≤Ib(u), for any u∈X.
From this deformation, with the help of Lemma 2.6, we can claim that
maxt∈DIb(η(1,tw))<cb,1. | (4.5) |
Indeed, on the one hand, from Lemma 2.6, we know ∀t∈D,
Ib(tw)≤Ib(w)=cb,1≤cb,1+ε. |
In addition, based on the definition of σ, it can be concluded that
‖tw−w‖2X≤σ2‖w‖2X≤12δ2<δ2, |
which means tw∈Sδ. Hence, tw∈Icb,1+εb∩Sδ. According to (b), it is easy to obtain (4.5).
In what follows, we can first claim that η(1,tw)∩Nb≠∅ for some t∈D. We define
Φ(t):=⟨I′b(η(1,tw)),η(1,tw)⟩,for t>0. |
From (4.4), the definition of ε and (a), we have η(1,t1w)=t1w and η(1,t2w)=t2w. From (4.3), one has
Φ(t1)=⟨I′b(t1w),t1w⟩>0, Φ(t2)=⟨I′b(t2w),t2w⟩<0. | (4.6) |
In view of (4.6) and the continuity of Φ, there exists t0∈[t1,t2]⊂D such that Φ(t0)=0. From the definition of Φ, one has η(1,t0w)∈Nb. Namely, η(1,tw)∩Nb≠∅ for some t∈[t1,t2]. Then, one gets cb,1≤Ib(η(1,t0w)), which is clearly contradictory to (4.5). Hence, the assumption is not valid and we ascertain that I′b(w)=0. Therefore, w is a non-negative solution of Equation (1.5). Finally, according to the maximum principle [29, Theorem 3.5], one can obtain w>0 in R3. Then, we say Equation (1.5) has a positive ground solution w. So far, the proof is completed.
Remark 4.2. Let u1∈X be the solution obtained from Theorem 1.1. From the theory of classical Schrödinger equations (see [30, Theorem 3.1]), it is easy to conclude that u1 decays exponentially as |x|→∞. Namely, for every δ>0, there exist C=C(δ)>0 and R>0 such that
u1(x)≤Ce−δ|x|,for any |x|≥R. | (4.7) |
In this section, we will prove the existence and asymptotic behavior of least-energy sign-changing solutions of Eq. (1.5). We first give the relationship among cb,1, cb,∞ and cb,2. Then, inspired by [24], we can construct a sign-changing (PS)cb,2 sequence for Ib. After that, we can use Lemma 3.1 to prove Theorem 1.2. Finally, we prove Theorem 1.6.
Lemma 5.1. Assume that (V0)−(V2), (V′4), and (f0)−(f3) hold. Then, there exists b∗>0 small enough such that for 0<b<b∗ we have
0<cb,2<cb,1+cb,∞<2cb,∞. |
Proof. Define ˉun(x):=ˉu(x−ne1) and e1:=(1,0,0), where ˉu is given in Lemma 2.7. In what follows, u1 represents the positive ground state solution of Equation (1.5), which is obtained from Theorem 1.1.
Claim 1: There exist s0,t0>0 such that s0u1−t0ˉun0∈Mb for some n0∈N large enough.
In fact, denote χ(a)=1au1−ˉun with a>0, and define a1,a2 by
a1=sup{a∈R+:χ+(a)≠0}anda2=inf{a∈R+:χ−(a)≠0}. |
By Lemma 2.5, there exists (s(χ(a)),t(χ(a))) such that s(χ(a))χ+(a)+t(χ(a))χ−(a)∈Mb. Because u1 is positive and ˉu is radial, we can prove that a1=+∞. Indeed, we first notice that ˉu is radially symmetric, as the same arguments presented in [31, Lemma 3.1.2] (see also [32, Radial Lemma 1]), one has that there exists a constant C>0 such that
|ˉu(x)|≤C‖ˉu‖X|x|,for every |x|≥1. | (5.1) |
We can obtain that for every x∈BR(0), there is |x−ne1|≥n−|x|≥n−R. By using (5.1), one can ascertain
ˉun(x)=ˉu(x−ne1)≤C1‖ˉu‖Xn−R≤C2n−R. |
Then, for fixed x∈BR(0), we will have
χ(a)=1au1−ˉun≥1au1−C21n−R. | (5.2) |
Therefore, we can take ε=u12aC2>0 and n0∈N, such that 1n−R<ε for every n≥n0. Combining with (5.2), we know that for all n≥n0 and a∈(0,+∞), there is
χ(a)=1au1−ˉun>12au1>0. |
Finally, we obtain that a1=+∞ by the definition of a1.
If a→a1=+∞, then 1au1→0 in X and χ+(a)→0. Similar to [33, Lemma 2.2 (ii)], one concludes that s(χ(a))→+∞ and {t(χ(a))} is bounded in R+. Hence, as a→a1, one has
s(χ(a))−t(χ(a))→+∞. | (5.3) |
Similarly, if a→a+2, χ−(a)→0, one has that t(χ(a))→+∞ and {s(χ(a))} is bounded in R+. So, as a→a+2, we have
s(χ(a))−t(χ(a))→−∞. | (5.4) |
From [33, Lemma 2.2 (i)], we can ascertain the continuity of s and t. Combining (5.3) and (5.4), we obtain that there exists a0∈(a2,a1) such that s(χ(a0))=t(χ(a0)) for n0. Hence, let s0=1a0s(χ(a0)) and t0=t(χ(a0)), it is easy to show that
s(χ(a0))χ(a0)=s0u1−t0ˉun0∈Mb. |
Claim 2: There exist b∗>0 small enough and n∈N+ large enough such that for 0<b<b∗ we have
maxs,t>0Ib(s0u1−t0ˉun)<cb,1+cb,∞. | (5.5) |
Obviously, Ib(su1−tˉun)<0 for s or t large enough. Next, we only need to consider this problem in a bounded interval. That is, we consider the case that s,t∈(0,C), where C>0. Moreover, one can check that there is tn>0 that satisfies tn→1 as n→∞ and tnˉun∈Nb. Then, by a direct calculation, we can obtain
Ib(su1−tˉun)=12∫R3(a|∇(su1−tˉun)|2+V(x)(su1−tˉun)2)dx+b4(∫R3|∇(su1−tˉun)|2dx)2−∫R3F(su1−tˉun)dx=12∫R3(a|∇(su1)|2+V(x)(su1)2)dx+b4(∫R3|∇(su1)|2dx)2−∫R3F(su1)dx+12∫R3(a|∇(tˉun)|2+V(x)(tˉun)2)dx+b4(∫R3|∇(tˉun)|2dx)2−∫R3F(tˉun)dx−∫R3a∇su1⋅∇tˉundx−∫R3V(x)(su1)(tˉun)dx+b4{4s2t2(∫R3∇u1⋅∇ˉundx)2−4s3t∫R3|∇u1|2dx∫R3∇u1⋅∇ˉundx+2s2t2∫R3|∇u1|2dx∫R3|∇ˉun|2dx−4st3∫R3|∇ˉun|2dx∫R3∇u1⋅∇ˉundx}−∫R3(F(su1−tˉun)−F(su1)−F(tˉun))dx≤Ib(su1)+Ib(tnˉun)+Bn+Cn+Dn=Ib(su1)+Ib,∞(tnˉun)+An+Bn+Cn+Dn, | (5.6) |
where
An=12∫R3(V(x)−V∞)(tnˉun)2dx,Bn=−st∫R3(a∇u1⋅∇ˉun+V(x)u1ˉun)dx,Cn=b4{4s2t2(∫R3∇u1⋅∇ˉundx)2−4s3t∫R3|∇u1|2dx∫R3∇u1⋅∇ˉundx+2s2t2∫R3|∇u1|2dx∫R3|∇ˉun|2dx−4st3∫R3|∇ˉun|2dx∫R3∇u1⋅∇ˉundx},Dn=∫R3(F(su1)+F(tˉun)−F(su1−tˉun))dx. |
First, similar to the conclusion presented in [10], we can show that
∫R3u1|ˉun|qdx≤C1e−δα(q+1)n+C2e−δq+1n, | (5.7) |
∫R3|u1|qˉundx≤C3e−δqα(q+1)n+C4e−δqq+1n≤C5e−δα(q+1)n+C6e−δq+1n, | (5.8) |
for any q∈[1,5].
For An, in view of (V′4), we can arrive at
An≤−14∫B1(0)(V∞−V(x+ne1))ˉu2dx≤−14C∫B1(0)ˉu21+|x+ne1|γdx≤−C7nγ. | (5.9) |
As for Dn, due to (1.6), one has
Dn≤2∫R3(f(su1)tˉun+f(tˉun)su1)dx≤2∫R3(ε(|su1|+|su1|5)tˉun+Cε|su1|q−1tˉun+ε(|tˉun|+|tˉun|5)su1+Cε|tˉun|q−1su1)dx≤C8e−δ6αn+C9e−δ6n. | (5.10) |
In what follows, we estimate Bn. Since ⟨I′b,∞(ˉun),u1⟩=0, we deduce that
∫R3(a∇ˉun⋅∇u1+V∞ˉunu1)dx+b∫R3|∇ˉun|2dx∫R3∇ˉun⋅∇u1dx=∫R3f(ˉun)u1dx. | (5.11) |
From (1.6), (5.7), and (5.11), we derive that
∫R3∇ˉun⋅∇u1dx=∫R3f(ˉun)u1dx−∫R3V∞ˉunu1dxa+b∫R3|∇ˉun|2dx≤C10(∫R3f(ˉun)u1dx+∫R3V∞ˉunu1dx)≤C11e−δ6αn+C12e−δ6n. | (5.12) |
Since ⟨I′b(u1),ˉun⟩=0, we can get that
∫R3(a∇u1⋅∇ˉun+V(x)u1ˉun)dx+b∫R3|∇u1|2dx∫R3∇u1⋅∇ˉundx=∫R3f(u1)ˉundx. |
Let us go back to the term Bn now. Combining (1.6), (5.8), (5.12), and the above equation, we can obtain that
Bn=st(b∫R3|∇u1|2dx∫R3∇u1⋅∇ˉundx−∫R3f(u1)ˉundx)≤st(b∫R3|∇u1|2dx∫R3∇u1⋅∇ˉundx+∫R3|f(u1)|ˉundx)≤C13e−δ6αn+C14e−δ6n. | (5.13) |
By (5.12), it is easy to conclude that
Cn≤C15e−δ6αn+C16e−δ6n+C42b∫R3|∇u1|2dx∫R3|∇ˉun|2dx≤C15e−δ6αn+C16e−δ6n+C17b. | (5.14) |
Then, due to (5.9), (5.10), (5.13), and (5.14), we can ascertain that
An+Bn+Cn+Dn≤−C7nγ+C18e−δ6αn+C19e−δ6n+C17b. | (5.15) |
Therefore, choosing n0∈N+ large enough, we can obtain −C7nγ0+C18e−δ6αn0+C19e−δ6n0<0. Hence, we can take b∗=C71nγ0−C18e−δ6αn0−C19e−δ6n02C17>0, and then, ∀b∈(0,b∗), we have
An+Bn+Cn+Dn<0, |
for all n≥n0. Noticing that (5.6), we deduce that
maxs,t∈(0,C)Ib(su1−tˉun)<maxs∈(0,C)Ib(su1)+Ib,∞(ˉun)≤Ib(u1)+Ib,∞(ˉun)=cb,1+cb,∞, |
for all n≥n0. Therefore, the claim is proved. Finally, combining Claim 1 and Claim 2, for any b∈(0,b∗), we can show that
cb,2≤Ib(s0u1−t0ˉun0)≤maxs,t>0Ib(su1−tˉun0)<cb,1+cb,∞. |
The proof is completed.
In order to construct a sign-changing (PS)cb,2 sequence of the functional Ib, we follow the method in [24]. Define
g(u,v):={∫R3f(u)udx‖u‖2X+b(∫R3|∇u|2dx)2+b∫R3|∇u|2dx∫R3|∇v|2dx,ifu≠0;0,ifu=0. | (5.16) |
First, since f being an odd function, we can obtain g(u,v)>0 if u≠0. Besides, we can see that u∈Mb if and only if g(u+,u−)=g(u−,u+)=1. Moreover, we can construct the following set U, which is larger than the set Mb. Namely, we define
U:={u∈X:|g(u+,u−)−1|<12,|g(u−,u+)−1|<12}. |
We know Mb⊂U. Furthermore, from Lemma 2.5, it can be concluded that Mb≠∅. That is, Mb⊂U≠∅. Now, we use P to represent the cone of the non-negative functions in X, D:=[0,1], E:=D×D, and Σ to represent the set of continuous maps σ, so that for all s,t∈D,
(i) σ∈C(E,X);
(ii) σ(s,0)=0,σ(0,t)∈P and σ(1,t)∈−P;
(iii) (Ib∘σ)(s,1)≤0 and ∫R3f(σ(s,1))σ(s,1)dx‖σ(s,1)‖2X+b(∫R3|∇(σ(s,1))|2dx)2≥2.
We can claim that Σ≠∅. In fact, for every u∈X and u±≠0, we can set σ(s,t)=μ(1−s)tu++μstu−, where μ>0 and s,t∈D. Then, through simple calculations, one can conclude σ(s,t)∈Σ for some μ>0. Moreover, we can also obtain the following lemmas.
Lemma 5.2. Assume that (V0)−(V2) and (f0)−(f4) hold. Then,
infσ∈Σsupu∈σ(E)Ib(u)=infu∈MbIb(u)=cb,2. |
Proof. On one hand, for all and , there is such that when is sufficiently large, . In light of Lemma 2.5, one concludes that . Hence, one obtains that
which implies that
(5.17) |
On the other hand, we can see that for every , there is . Therefore,
Then, one can arrive at
(5.18) |
Hence, combining (5.17) and (5.18), we can conclude that
Now it remains to prove the claim. Indeed, due to the definition of , we can know and , which holds for every and . Moreover, by using conditions and (1.7), it is easy to ascertain that for every , there is . Next, for convenience, one can define . Then, one can deduce
(5.19) |
and
(5.20) |
Then, we can use property and the inequality , where , to obtain that
Thus,
(5.21) |
In addition, we can verify . Hence,
(5.22) |
Therefore, using Miranda's theorem [34] and (5.19)-(5.22), there is that satisfies
As a result, one has
Namely, there exists for any . The proof is completed.
Lemma 5.3. Assume that and hold, then there exists a sequence satisfying and as .
Proof. Let be a minimizing sequence and , so
(5.23) |
We can claim that there is a sequence , which satisfies, as ,
(5.24) |
Suppose there is a contradiction, then, there exists such that for , which is sufficiently large, where
Through using [35], due to Hofer [36], there is , which satisfies the following properties for some and every .
, ;
for any , where ;
;
.
By (5.23), select to be sufficiently large so that
(5.25) |
Define for all . Then, it is clear to see that . From (5.25) and property , it can be inferred that . Therefore,
which is absurd, so the claim is valid.
Now, we begin to prove when is large enough. Indeed, in light of as , one can ascertain . After that, it is sufficient to prove , which means and . Therefore, one gets for large enough. Using (5.24), there is a sequence that satisfies
(5.26) |
To get , we only need to show that and for large enough. It derives from Lemma 2.3 and the fact of that for some . Next, we only need to prove and as . By contradiction, if as , in light of (5.26) and being continuous, one can get
Choosing such that
leads to a contradiction. Then, the above assumption is not valid. Therefore, for large enough.
Based on the previous lemma, we will now focus on the proof of Theorem 1.2.
Proof of Theorem 1.2. First, there is a sequence by Lemma 2.3 and Lemma 5.3, which is bounded. Then, we assume that there exists a subsequence, which satisfies in . We can claim in , thus we deduce that
From Lemma 2.3, by in , we get , namely . Hence, is a least energy sign-changing solution of Equation (1.5).
It remains to verify the above claim. In fact, due to Lemma 3.1, if case occurs, the proof is completed. If case occurs, since , it follows from (3.2) that . Hence, or .
If , in light of , one can deduce for and ,
(5.27) |
Since and , (5.27) and Lemma 2.3 imply that . Besides, due to , we get . Hence, we have by (3.16). Therefore,
which contradicts with . Thus, , combining , (3.2), (3.15), and (3.16), we can get that
Therefore, we can show that and in . The proof is completed.
From now on, we will study the asymptotic behavior of the above sign-changing solutions with respect to . In order to facilitate research, we set , and with , , and , where and respectively represent and with . Besides, from Theorem 1.2, we have obtained that Equation (1.5) has a least energy sign-changing solution for all under hypothesis and . Next, we denote it as and consider the problem with .
Proof of Theorem 1.6. First of all, in virtue of Theorem 1.2 and Lemma 5.1, we immediately know that is achievable by some for all and satisfies . Hence, there is a sequence that satisfies and . Similar to the proof in [37], we can get that , where is a positive constant. Due to , we have that
and we can easily infer that is bounded. Thus, there is a subsequence of and that satisfies in . Besides that, we can easily check that the sequence is bounded for all . Therefore, combining as , we can arrive at
Through simple calculations, one has
for all , which implies that .
In the following, we deduce that . Because of , we know that there is large enough such that for all . Due to Lemma 2.5, we get that there is that satisfies , where satisfies . Hence, we have by , and then . Therefore, we can conclude that
Hence,
(5.28) |
In light of and Fatou's Lemma, one gets
(5.29) |
Next, we show that and . First, combining the boundedness of and (5.28), one deduces that
(5.30) |
Besides, it is easy to check that
for all , which means that
(5.31) |
Hence, owing to (5.30) and (5.31), we know that there is a subsequence of , still denoted by , which satisfies
(5.32) |
In virtue of [10, Proposition 4.1], we get that . According to [10], we know that satisfies the condition, where . Hence, by (5.32), we get that the sequence has a convergent subsequence, still denoted by . Then, one has in . Moreover, in light of and Lemma 2.3, we can arrive at , and then . Noting that , we can show that . Therefore, we can obtain that . Combining with (5.29), we conclude that and . Namely, is a least energy sign-changing solution of Equation (1.3). The proof is completed.
Yan-Fei Yang: Writing-original draft, Writing-review & editing; Chun-Lei Tang: Formal analysis, Methodology, Supervision, Writing-review & editing.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This study was funded by the National Natural Science Foundation of China (No. 12371120).
All authors declare no conflicts of interest in this paper.
[1] | G. Kirchhoff, Mechanik, Teubner, Leipzig, 1883. |
[2] |
A. Arosio, S. Panizzi, On the well-posedness of the Kirchhoff string, Trans. Amer. Math. Soc., 348 (1996), 305–330. https://doi.org/10.1090/S0002-9947-96-01532-2 doi: 10.1090/S0002-9947-96-01532-2
![]() |
[3] |
P. Dancona, S. Spagnolo, Global solvability for the degenerate Kirchhoff equation with real analytic data, Invent. Math., 108 (1992), 247–262. https://doi.org/10.1007/BF02100605 doi: 10.1007/BF02100605
![]() |
[4] |
C. Alves, M. Souto, Existence of solutions for a class of nonlinear Schrödinger equations with potential vanishing at infinity, J. Differential Equations, 254 (2013), 1977–1991. https://doi.org/10.1016/j.jde.2012.11.013 doi: 10.1016/j.jde.2012.11.013
![]() |
[5] |
H. Berestycki, P. Lions, Nonlinear scalar field equations. I. Existence of a ground state, Arch. Ration. Mech. Anal., 82 (1983), 313–345. https://doi.org/10.1007/BF00250555 doi: 10.1007/BF00250555
![]() |
[6] |
A. Bahrouni, H. Ounaies, V. Radulescu, Infinitely many solutions for a class of sublinear Schrödinger equations with indefinite potentials, Proc. R. Soc. Edinb. Sect. A, 145 (2015), 445–465. https://doi.org/10.1017/S0308210513001169 doi: 10.1017/S0308210513001169
![]() |
[7] |
L. Jeanjean, K. Tanaka, Singularly perturbed elliptic problems with superlinear or asymptotically linear nonlinearities, Calc. Var. Partial Differential Equations, 21 (2004), 287–318. https://doi.org/10.1007/s00526-003-0261-6 doi: 10.1007/s00526-003-0261-6
![]() |
[8] |
R. Lehrer, L. Maia, Positive solutions of asymptotically linear equations via Pohozaev manifold, J. Funct. Anal., 266 (2014), 213–246. https://doi.org/10.1016/j.jfa.2013.09.002 doi: 10.1016/j.jfa.2013.09.002
![]() |
[9] | M. Willem, Minimax Theorems, Birkhäuser Boston, 1996. https://doi.org/10.1007/978-1-4612-4146-1 |
[10] |
M. Furtado, L. Maia, E. Medeiros, Positive and nodal solutions for a nonlinear Schrödinger equation with indefinite potential, Adv. Nonlinear Stud., 8 (2008), 353–373. https://doi.org/10.1515/ans-2008-0207 doi: 10.1515/ans-2008-0207
![]() |
[11] |
Y. Ding, A. Szulkin, Bound states for semilinear Schrödinger equations with sign-changing potential, Calc. Var. Partial Differential Equations, 29 (2007), 397–419. https://doi.org/10.1007/s00526-006-0071-8 doi: 10.1007/s00526-006-0071-8
![]() |
[12] |
X. Tang, Infinitely many solutions for semilinear Schrödinger equations with sign-changing potential and nonlinearity, J. Math. Anal. Appl., 401 (2013), 407–415. https://doi.org/10.1016/j.jmaa.2012.12.035 doi: 10.1016/j.jmaa.2012.12.035
![]() |
[13] |
Q. Zhang, B. Xu, Multiplicity of solutions for a class of semilinear Schrödinger equations with sign-changing potential, J. Math. Anal. Appl., 377 (2011), 834–840. https://doi.org/10.1016/j.jmaa.2010.11.059 doi: 10.1016/j.jmaa.2010.11.059
![]() |
[14] |
Y. Deng, S. Peng, W. Shuai, Existence and asymptotic behavior of nodal solutions for the Kirchhoff-type problems in , J. Funct. Anal., 269 (2015), 3500–3527. https://doi.org/10.1016/j.jfa.2015.09.012 doi: 10.1016/j.jfa.2015.09.012
![]() |
[15] |
W. Lv, Ground states of a Kirchhoff equation with the potential on the lattice graphs, Commun. Anal. Mech., 15 (2023), 792–810. https://doi.org/10.3934/cam.2023038 doi: 10.3934/cam.2023038
![]() |
[16] |
Y. Li, F. Li, J. Shi, Existence of a positive solution to Kirchhoff type problems without compactness conditions, J. Differential Equations, 253 (2012), 2285–2294. https://doi.org/10.1016/j.jde.2012.05.017 doi: 10.1016/j.jde.2012.05.017
![]() |
[17] |
G. Li, H. Ye, Existence of positive ground state solutions for the nonlinear Kirchhoff type equations in , J. Differential Equations, 257 (2014), 566–600. https://doi.org/10.1016/j.jde.2014.04.011 doi: 10.1016/j.jde.2014.04.011
![]() |
[18] |
J. Sun, L. Li, M. Cencelj, B. Gabrovšek, Infinitely many sign-changing solutions for Kirchhoff type problems in , Nonlinear Anal., 186 (2019), 33–54. https://doi.org/10.1016/j.na.2018.10.007 doi: 10.1016/j.na.2018.10.007
![]() |
[19] |
X. Tang, S. Chen, Ground state solutions of Nehari-Pohozaev type for Kirchhoff-type problems with general potentials, Calc. Var. Partial Differential Equations, 56 (2017), 110. https://doi.org/10.1007/s00526-017-1214-9 doi: 10.1007/s00526-017-1214-9
![]() |
[20] |
L. Wang, B. Zhang, K. Cheng, Ground state sign-changing solutions for the Schrödinger-Kirchhoff equation in , J. Math. Anal. Appl., 466 (2018), 1545–1569. https://doi.org/10.1016/j.jmaa.2018.06.071 doi: 10.1016/j.jmaa.2018.06.071
![]() |
[21] |
F. Zhang, M. Du, Existence and asymptotic behavior of positive solutions for Kirchhoff type problems with steep potential well, J. Differential Equations, 269 (2020), 10085–10106. https://doi.org/10.1016/j.jde.2020.07.013 doi: 10.1016/j.jde.2020.07.013
![]() |
[22] |
Y. Ni, J. Sun, J. Chen, Multiplicity and concentration of normalized solutions for a Kirchhoff type problem with -subcritical nonlinearities, Commun. Anal. Mech., 16 (2024), 633–654. https://doi.org/10.3934/cam.2024029 doi: 10.3934/cam.2024029
![]() |
[23] |
A. Batista, M. Furtado, Solutions for a Schrödinger-Kirchhoff equation with indefinite potentials, Milan J. Math., 86 (2018), 1–14. https://doi.org/10.1007/s00032-018-0276-2 doi: 10.1007/s00032-018-0276-2
![]() |
[24] |
G. Cerami, S. Solimini, M. struwe, Some existence results for superlinear elliptic boundary value problem involving critical exponents, J. Funct. Anal., 69 (1986), 289–306. https://doi.org/10.1016/0022-1236(86)90094-7 doi: 10.1016/0022-1236(86)90094-7
![]() |
[25] |
M. Wu, C. Tang, The existence and concentration of ground state sign-changing solutions for Kirchhoff-type equations with a steep potential well, Acta Math. Sci. Ser. B (Engl. Ed.), 43 (2023), 1781–1799. https://doi.org/10.1007/s10473-023-0419-6 doi: 10.1007/s10473-023-0419-6
![]() |
[26] |
X. He, W. Zou, Existence and concentration behavior of positive solutions for a Kirchhoff equation in , J. Differential Equations, 252 (2012), 1813–1834. https://doi.org/10.1016/j.jde.2011.08.035 doi: 10.1016/j.jde.2011.08.035
![]() |
[27] |
G. Li, P. Luo, S. Peng, C. Wang, C. Xiang, A singularly perturbed Kirchhoff problem revisited, J. Differential Equations, 268 (2020), 541–589. https://doi.org/10.1016/j.jde.2019.08.016 doi: 10.1016/j.jde.2019.08.016
![]() |
[28] |
C. Alves, Existence of positive solutions for a problem with lack of compactness involving the p-Laplacian, Nonlinear Anal., 51 (2002), 1187–1206. https://doi.org/10.1016/S0362-546X(01)00887-2 doi: 10.1016/S0362-546X(01)00887-2
![]() |
[29] | D. Gilbarg, N. Trudinger, Elliptic partial differential equations of second order, Springer-Verlag, Berlin, 2001. https://doi.org/10.1007/978-3-642-61798-0_3 |
[30] |
G. Li, S. Yan, Eigenvalue problems for quasilinear elliptic equations on , Comm. Partial Differential Equations, 14 (1989), 1291–1314. https://doi.org/10.1080/03605308908820654 doi: 10.1080/03605308908820654
![]() |
[31] | M. Badiale, E. Serra, Semilinear Elliptic Equations for Beginners. Existence Results via the Varia tional Approach, Springer-Verlag, London, 2011. https://doi.org/10.1007/978-0-85729-227-8 |
[32] |
W. Strauss, Existence of solitary waves in higher dimensions, Comm. Math. Phys., 55 (1977), 149–162. https://doi.org/10.1007/BF01626517 doi: 10.1007/BF01626517
![]() |
[33] |
Y. Zhao, X. Wu, C. Tang, Ground state sign-changing solutions for Schrödinger-Kirchhoff-type problem with critical growth, J. Math. Phys., 63 (2022), 101503. https://doi.org/10.1063/5.0092120 doi: 10.1063/5.0092120
![]() |
[34] | C. Miranda, Un'osservazione su un teorema di Brouwer, Boll. Unione Mat. Ital., 3 (1940), 5–7. |
[35] | P. Rabinowitz, Variational methods for nonlinear eigenvalue problems, in: G. Prodi (Ed.), Eigenvalues of Nonlinear Problems, CIME, (1974), 141–195. https://doi.org/10.1007/978-3-642-10940-9_4 |
[36] |
H. Hofer, Variational and topological methods in partially ordered Hilbert spaces, Math. Ann., 261 (1982), 493–514. https://doi.org/10.1007/BF01457453 doi: 10.1007/BF01457453
![]() |
[37] |
J. Chen, X. Tang, B. Cheng, Existence and concentration of ground state sign-changing solutions for Kirchhoff type equations with steep potential well and nonlinearity, Topol. Methods Nonlinear Anal., 51 (2018), 111–133. https://doi.org/10.12775/TMNA.2017.062 doi: 10.12775/TMNA.2017.062
![]() |