In this paper, we studied the existence of normalized solutions to the following Kirchhoff equation with a perturbation:
{−(a+b∫RN|∇u|2dx)Δu+λu=|u|p−2u+h(x)|u|q−2u, in RN,∫RN|u|2dx=c,u∈H1(RN),
where 1≤N≤3,a,b,c>0,1≤q<2, λ∈R. We treated three cases:
(i) When 2<p<2+4N,h(x)≥0, we obtained the existence of a global constraint minimizer.
(ii) When 2+8N<p<2∗,h(x)≥0, we proved the existence of a mountain pass solution.
(iii) When 2+8N<p<2∗,h(x)≤0, we established the existence of a bound state solution.
Citation: Xin Qiu, Zeng Qi Ou, Ying Lv. Normalized solutions to nonautonomous Kirchhoff equation[J]. Communications in Analysis and Mechanics, 2024, 16(3): 457-486. doi: 10.3934/cam.2024022
[1] | Yangyu Ni, Jijiang Sun, Jianhua Chen . Multiplicity and concentration of normalized solutions for a Kirchhoff type problem with L2-subcritical nonlinearities. Communications in Analysis and Mechanics, 2024, 16(3): 633-654. doi: 10.3934/cam.2024029 |
[2] | Xueqi Sun, Yongqiang Fu, Sihua Liang . Normalized solutions for pseudo-relativistic Schrödinger equations. Communications in Analysis and Mechanics, 2024, 16(1): 217-236. doi: 10.3934/cam.2024010 |
[3] | 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 |
[4] | Shengbing Deng, Qiaoran Wu . Existence of normalized solutions for the Schrödinger equation. Communications in Analysis and Mechanics, 2023, 15(3): 575-585. doi: 10.3934/cam.2023028 |
[5] | 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 |
[6] | Yan-Fei Yang, Chun-Lei Tang . Positive and sign-changing solutions for Kirchhoff equations with indefinite potential. Communications in Analysis and Mechanics, 2025, 17(1): 159-187. doi: 10.3934/cam.2025008 |
[7] | Henryk Żołądek . Normal forms, invariant manifolds and Lyapunov theorems. Communications in Analysis and Mechanics, 2023, 15(2): 300-341. doi: 10.3934/cam.2023016 |
[8] | Panyu Deng, Jun Zheng, Guchuan Zhu . Well-posedness and stability for a nonlinear Euler-Bernoulli beam equation. Communications in Analysis and Mechanics, 2024, 16(1): 193-216. doi: 10.3934/cam.2024009 |
[9] | Chang-Jian Wang, Jia-Yue Zhu . Global existence and uniform boundedness to a bi-attraction chemotaxis system with nonlinear indirect signal mechanisms. Communications in Analysis and Mechanics, 2023, 15(4): 743-762. doi: 10.3934/cam.2023036 |
[10] | Maomao Wu, Haidong Liu . Multiple solutions for quasi-linear elliptic equations with Berestycki-Lions type nonlinearity. Communications in Analysis and Mechanics, 2024, 16(2): 334-344. doi: 10.3934/cam.2024016 |
In this paper, we studied the existence of normalized solutions to the following Kirchhoff equation with a perturbation:
{−(a+b∫RN|∇u|2dx)Δu+λu=|u|p−2u+h(x)|u|q−2u, in RN,∫RN|u|2dx=c,u∈H1(RN),
where 1≤N≤3,a,b,c>0,1≤q<2, λ∈R. We treated three cases:
(i) When 2<p<2+4N,h(x)≥0, we obtained the existence of a global constraint minimizer.
(ii) When 2+8N<p<2∗,h(x)≥0, we proved the existence of a mountain pass solution.
(iii) When 2+8N<p<2∗,h(x)≤0, we established the existence of a bound state solution.
In this paper, we consider the existence of solutions with prescribed L2-norm to the following Kirchhoff problem with a perturbation
{−(a+b∫RN|∇u|2dx)Δu+λu=|u|p−2u+h(x)|u|q−2u, in RN,∫RN|u|2dx=c,u∈H1(RN), | (1.1) |
where 1≤N≤3,a,b,c>0,p∈(2,2∗),q∈[1,2), h(x):RN→R is a potential, 2∗=6 if N=3, and 2∗=+∞ if N=1,2. Based on these observations, we establish the existence of normalized solutions under different assumptions on h(x).
The energy functional of Eq.(1.1) is defined by
I(u)=a2∫RN|∇u|2dx+b4(∫RN|∇u|2dx)2−1p∫RN|u|pdx−1q∫RNh(x)|u|qdx | (1.2) |
constrained on the L2-spheres in H1(RN):
Sc={u∈H1(RN):‖u‖22=c>0}. |
In 1883, Kirchhoff [1] first proposed the following nonlinear wave equation
ρ∂2u∂t2−(P0h+E2L∫L0|∂u∂x|2dx)∂2u∂x2=0, |
which extends the original wave equation by describing the transversal oscillations of a stretched string and, particularly, by considering the subsequent change in string length caused by oscillations. Thereafter, there was a boom in the study of the Kirchhoff-type equation. We can refer to [2,3,4] for the physical background about Kirchhoff problem.
Mathematically, Eq.(1.1) is not a pointwise identity as a result of the emergence of the term (b∫RN|∇u|2dx)Δu. This causes some mathematical difficulties. In the renowned paper [5], J.L. Lions raised an abstract framework that has received much attention. There are two ways to study the Kirchhoff-type equation. The first approach is to consider fixing the parameter λ∈R. In this case, there are a lot of results, which have been widely studied by using variational methods. We can refer to [6,7,8,9] and the references therein. Another way is to fix the L2-norm. In this case, the desired solutions have a priori prescribed L2-norm, which are usually referred to as normalized solutions in the literature; that is, for any fixed c>0, we take (uc,λc)∈H1(RN)×R as a normalized solution with ‖uc‖22=c, λc is a Lagrange multiplier. From a physical perspective, the L2-prescribed norm represents the number of particles of each component in Bose-Einstein condensates or the power supply in a nonlinear optics framework. In addition, the L2-prescribed norm can provide a better insight on the dynamical properties, like orbital stability or instability, and can describe attractive Bose-Einstein condensates.
For the local case, i.e., b=0, Eq.(1.1) reduces to the general Schrödinger type:
{−Δu+λu=f(x,u), in RN,∫RN|u|2dx=c,u∈H1(RN), | (1.3) |
which dates back to the groundbreaking work by Stuart. In [10], Stuart tackled the problem (1.3) for f(x,u)=|u|p−2u and p∈(2,2+4N) (L2-subcritical case); here, 2+4N is called the L2-critical exponent. For L2-subcritical case, the minimization method is the conventional method to find normalized solutions. When f is L2-supercritical growth, a groundbreaking work in the L2-supercritical case was accomplished by Jeanjean[11]. Jeanjean developed a novel argument related to the mountain pass geometry by the stretched functional. Bartsch and Soave [12,13] also proposed a new approach by using a minimax principle based on the homotopy stable family to prove the existence of normalized solutions for the problem (1.3). Moreover, Soave in [14] studied the combined nonlinearity case f(x,u)=|u|p−2u+μ|u|q−2u, 2<q≤2+4N≤p<2∗ and q<p, where 2∗=∞ if N≤2 and 2∗=2NN−2 if N≥3. Soave showed that nonlinear terms with different power strongly affects the geometry of the functional and the existence and properties of ground states.
When f(x,u)=a(x)f(u), the solutions to the nonautonomous problem were first studied by Chen and Tang [15]. Compared with the autonomous problems, the main challenge of the problem is constructing a (PS) sequence with an additional property to recover the compactness. Very recently, Chen and Zou [16] studied the following problem with a perturbation
{−Δu+λu=|u|p−2u+h(x), in RN,∫RN|u|2dx=c,u∈H1(RN), | (1.4) |
where h(x)≥0. For p∈(2,2+4N) and an arbitrarily positive perturbation, Chen and Zou proved that there exists a global minimizer with negative energy. The existence of a mountain pass solution with positive energy for p∈(2+4N,2∗) was studied. We can refer to [17,18,19] for more details.
For the nonlocal case, i.e., b>0, the more general form of Eq.(1.1) is the following equation
{−(a+b∫RN|∇u|2dx)Δu+λu=f(x,u), in RN,∫RN|u|2dx=c,u∈H1(RN), | (1.5) |
which has attracted considerable attention. When f(x,u)=|u|p−2u (i.e., the limited problem of Eq.(1.1)), the problem (1.5) turns to
{−(a+b∫RN|∇u|2dx)Δu+λu=|u|p−2u, in RN,∫RN|u|2dx=c,u∈H1(RN), | (1.6) |
where a,b,c>0 are constants, 1≤N≤3, and p∈(2,2∗). The energy functional of (1.6) is
I∞(u)=a2∫RN|∇u|2dx+b4(∫RN|∇u|2dx)2−1p∫RN|u|pdx. | (1.7) |
By the Gagliardo-Nirenberg inequality [20] for any p∈(2,2∗)
‖u‖p≤CN,p‖∇u‖γp2‖u‖1−γp2 | (1.8) |
where γp=N(p−2)2p, we can get L2-critical exponent ˉp=2+8N of the Kirchhoff problem. It is well known that Ye [21] obtained the sharp existence of global constraint minimizers for Eq.(1.6) in the case of p∈(2,ˉp). When p∈(2+4N,ˉp), Ye proved a local minimizer, which is a critical point of I∞|Sc. By considering a global minimization problem
l∞,c:=infScI∞(u), | (1.9) |
we have
{l∞,c∈(−∞,0], if p∈(2,ˉp),l∞,c=−∞, if p∈(ˉp,2∗), | (1.10) |
for any given c>0. We can see that the minimization method is not feasible for p∈(ˉp,2∗). Then, Ye proved the existence of normalized solutions by taking advantage of the Pohozaev constraint method in the case of p∈(ˉp,2∗). For the L2-critical case of ˉp=2+8N, Ye [22] showed the existence and mass concentration of critical points. Using some simple energy estimates instead of the concentration-compactness principles introduced in [21], Zeng studied the existence and uniqueness of normalized solutions for p∈(2,2∗) in [23].
Additionally, Li, Luo, and Yang [24] proved the existence and asymptotic properties of solutions to the following equation with combined nonlinearity
{−(a+b∫RN|∇u|2dx)Δu+λu=|u|p−2u+μ|u|q−2u, in R3,∫RN|u|2dx=c,u∈H1(RN), | (1.11) |
where a,b,c,μ>0, 2<q<143<p≤6 or 143<q<p≤6. They showed a multiplicity result for the case of 2<q<103 and 143<p<6 and obtained the existence of ground state normalized solutions for 2<q<103<p=6 or 143<q<p≤6. They also showed some asymptotic results on the obtained solutions. For the case μ≤0, in [25], Carri˜ao, Miyagaki, and Vicente studied the ground states existence of Eq.(1.11) for 2<q<2∗,p=2∗ or 2<q≤ˉp<p<2∗. For the nonautonomous problem, when f(x,u)=|u|p−2u+V(x)|u|q−2u, N=3, p=143, q=4 and V∈L∞loc(R3), Ye [26] considered the existence of minimizers to the nonautonomous problem. Moreover, V(x) satisfies
V(x)≥0,lim|x|→∞V(x)=0. |
By the concentration compactness principle, if b<b0, Ye showed that there exists a0,c0>0 such that the above problem has a minimizer for all a<a0 and c<c0. Additionally, when f(x,u)=K(x)f(u), Chen and Tang [27] considered the existence of ground state solutions, where K(x)∈C(R3,R+) and f(u) is L2-supercritical. When 2+4N<p<2+8N, the geometric structure of the energy functional is more complex, especially when h(x)>0, and there are very few works studying this range with potential. Other results about normalized solutions of Kirchhoff equation in a more general form can be found in [28,29,30,31].
Motivated by the results above, when μ of Eq.(1.11) is replaced by a potential function h(x) and 1≤q<2, there are no results in studying normalized solutions of such nonautonomous Kirchhoff equations with a small perturbation. In the present paper, we first obtain the normalized solution of this type of equation, which can be seen as an extension of some known results in the literature.
Let us now outline the main strategy to prove the three results of this paper under different assumptions on h(x). First, we treat the mass-subcritical case 2<p<2+4N: for any c>0, we set
lc:=infScI(u). | (1.12) |
It is standard that the minimizers of lc are critical points of I|Sc. We introduce the following assumptions on h(x).
(h1) h∈L22−q(RN) and h(x)>0 on a set with positive measure. |
Now we state the main results of this paper:
Theorem 1.1. Suppose 1≤N≤3, 2<p<2+4N and h(x)≥0 satisfies (h1). Then, for all c>0, lc has a minimizer, hence Eq.(1.1) has a normalized ground state solution.
Remark 1.1. Notice that the minimizer obtained in Theorem 1.1 is a global minimizer rather than a local minimizer. It is easy to find that the energy functional is coercive on Sc, which hints that each minimizing sequence {un} is bounded on Sc. The main difficulty of proof is to show that the minimizing sequence {un} converges strongly to u≠0 in H1(RN). The key step is to establish the inequality lc1+c2≤lc1+l∞,c2 for c1,c2>0 (see Lemma 2.2), which is crucial to recover the compactness.
Next, while addressing the L2-supercritical case, the functional is unbounded from below on Sc, thus the minimizing approach on Sc is not valid anymore. Ye [21] proved that l∞,c=−∞ for all c>0 if p∈(2+8N,2∗), and proved the existence of one normalized solution by a suitable submanifold of Sc. In this paper, after the appearance of a very small perturbation term, we want to show that the energy functional I has a mountain pass geometry and show the existence of a mountain pass solution with positive energy level for p∈(2+8N,2∗). We require the perturbation h(x) to have a higher regularity. We need to assume that:
(h2) h∈Lpp−q(RN)∩C1(RN), ⟨∇h,x⟩∈L22−q(RN) and h(x)≥0. |
We have the following result.
Theorem 1.2. Suppose 1≤N≤3, 2+8N<p<2∗ and h(x) satisfies (h2). Let c>0 be fixed. Moreover,
‖h‖pp−q<aq(pγp−2)2CqN,pγp(p−q)(ap(2−qγp)2γp(p−q)CpN,p)2−qγppγp−2c−(1−γp)(p−q)pγp−2, | (1.13) |
‖∇h⋅x‖22−q<q(2p−Np+2N)p−2mcc−q2. | (1.14) |
Then, Eq.(1.1) has a mountain pass solution u at a positive energy level.
Remark 1.2. We are going to use the minimax characterization to find a critical point. Although the mountain pass geometry of the functional on Sc can be obtained easily, unfortunately the boundedness of the obtained (PS) sequence is not yet clear. In this paper, we adopt a similar idea to [11] and construct an auxiliary map ˜I(t,u):=I(t⋆u), which on R×Sc has the same type of geometric structure as I on Sc. Besides, the (PS) sequence of I satisfies the additional condition (see Lemma 3.5), which is the key ingredient to obtain the boundedness of the (PS) sequence.
Finally, we will discuss h(x)≤0, and the problem becomes more delicate and difficult. Although the mountain pass structure by Jeanjean [11] is destroyed, Bartsch et al.[32] established a new variational principle exploiting the Pohozaev identity. For convenience, we define ˉh(x):=−h(x)≥0. Next, we state our basic assumptions on ˉh(x).
(h3) ˉh(x)∈L22−q(RN)∩C1(RN), ⟨∇ˉh(x),x⟩∈L22−q(RN) and ˉh(x)≥0. For some constants Υ>0, ˉh(x) satisfies
|x⋅∇ˉh(x)|≤Υˉh(x). |
Theorem 1.3. Assume 1≤N≤3, 2+8N<p<2∗. If (h3) holds and ˉh(x) satisfies
0<‖ˉh‖22−q<min{1,2p(1−γp)2(p−q)+(p−2)Υ}⋅qmccq2. | (1.15) |
Then Eq.(1.1) has a couple of solutions (u,λ)∈H1(RN)×R and λ>0.
Remark 1.3. Indeed, when h(x)≤0, the problem is made more difficult by the simultaneous appearance of a negative potential and nonlocal term. We refer to Bartsch et al. [32] constructing a suitable linking geometry method to obtain the existence of bound state solutions with high Morse index. The crucial step is to estimate the minimax level mc<Lh,c<2mc (see Lemma 4.3 and Lemma 4.5) to recover the compactness.
Notations: We introduce some notations that will clarify what follows:
∙ H1(RN) is the usual Sobolev space with the norm ‖u‖=(∫RN|∇u|2+|u|2dx)12.
∙ Lp(RN) with p∈[1,∞) is the Lebesgue space with the norm ‖u‖p=(∫RN|u|pdx)1p.
∙ The arrows ′⇀′ and ′→′ denote the weak convergence and strong convergence, respectively.
∙ C,Ci denote positive constants, which may vary from line to line.
∙ (t⋆u)(x):=tN2u(tx) for t∈R+ and u∈H1(RN).
In this section, for 2<p<2+4N and h(x)≥0 we prove Theorem 1.1. By the Gagliardo-Nirenberg inequality (1.8), the Hölder inequality, and the assumption (h1), we have
I(u)=a2‖∇u‖22+b4‖∇u‖42−1p‖u‖pp−1q∫RNh(x)|u|qdx≥a2‖∇u‖22+b4‖∇u‖42−1pCpN,p‖∇u‖pγp2‖u‖p(1−γp)2−1q‖h‖22−q‖u‖q2, | (2.1) |
thus I is bounded from below on Sc since 0<pγp<2.
For 1≤N≤3 and 2<p<2+4N, the existence and uniqueness of positive normalized solutions of the limited problem (1.6) have been studied in [21]. In order to find the minimizer of I on Sc, first we state some fundamental properties of l∞,c, which will be crucial to recover the compactness later on. The proof of the next lemma can be found in [28,Theorem 1.1 and Lemma 2.5].
Lemma 2.1. Suppose 1≤N≤3 and 2<p<2+4N. Then, for all c>0, we have
(i) the strict sub-additivity for l∞,c, i.e.,
l∞,c1+c2<l∞,c1+l∞,c2 for c1,c2>0; |
(ii) the limited problem (1.6) has a couple of ground state solutions (u∞,λc)∈H1(RN)×R, i.e.,
l∞,c=infScI∞(u)=I∞(u∞)<0. |
Next, we introduce the inequality lc1+c2≤lc1+l∞,c2, which plays a crucial role in proving the convergence of the minimizing sequence.
Lemma 2.2. Suppose 2<p<2+4N and h(x) satisfies (h1), then the following holds
(i)−∞<lc<l∞,c<0 for c>0;
(ii)lc1+c2≤lc1+l∞,c2 for c1,c2>0.
Proof. (i) It is obvious that lc>−∞ by (2.1). Moreover, by Lemma 2.1, we have
lc≤I(u∞)=a2∫RN|∇u∞|2dx+b4(∫RN|∇u∞|2dx)2−1p∫RN|u∞|pdx−1q∫RNh|u∞‖qdx<I∞(u∞)=l∞,c<0, |
since u∞>0 and h(x) satisfies (h1).
(ii) For any ε>0, c=c1+c2, we can find φε,ψε∈C∞0(RN) such that
φε∈Sc1,I(φε)<lc1+ε2,ψε∈Sc2,I∞(ψε)<l∞,c2+ε2. |
Let uε,n(x):=φε(x)+ψε(x−ne1), where e1 is the unit vector (1,0,⋯) in RN. Since φε and ψε have compact support, we see that uε,n∈Sc and
lc≤I(uε,n)=I(φε)+I(ψε(x−ne1)), |
for large n. Moreover, thanks to h∈L22−q(RN), we have that ∫RNh(x)ψqε(x−ne1)dx→0 as n→∞, hence I(ψε(⋅−ne1))→I∞(ψε) as n→∞. It follows that
lc≤lim supn→∞I(uε,n)=lim supn→∞(I(φε)+I(ψε(⋅−ne1)))=I(φε)+I∞(ψε)<lc1+l∞,c2+ε. |
Passing to the limit, thus lc≤lc1+l∞,c2 since ε>0 is arbitrary.
Let {un}⊂Sc be a minimizing sequence for lc. By (2.1), we know that I(u) is coercive on Sc and deduce that {un} is bounded in H1(RN). Thus, there exists a subsequence such that un⇀u0 and
I(u0)≤lim infn→∞I(un)=lc,c1:=‖u0‖22≤‖un‖22=c. |
We need to prove I(u0)=lc and ‖u0‖22=c. Now we argue by contradiction to prove this.
Lemma 2.3. Suppose 2<p<2+4N and h(x) satisfies (h1). Then, every minimizing sequence for lc has a strong convergent subsequence in L2(RN).
Proof. We argue by contradiction and assume that c1<c. We divide the proof into four steps.
Step 1: There exists {yn}⊂RN and μ0∈H1(RN)∖{0} such that
|yn|→∞, un(⋅+yn)⇀μ0 in H1(RN). | (2.2) |
First, we show by contradiction that
δ0:=lim infn→∞supy∈RN∫B1(y)|un−u0|2dx>0, | (2.3) |
where B1(y)={x∈RN:|x−y|≤1}. Suppose, on the contrary, that δ0=0. Then, un→u0 strongly in Lp(RN). Since un⇀u0 in H1(RN),h∈L22−q(RN), we see that ∫RNh|un|qdx→∫RNh|u0|qdx. Combined with Lemma 2.1 (ii), for c−c1>0, we have that
lc=I(un)+o(1)=I(u0)+I(un−u0)+o(1)=I(u0)+a2∫RN|∇(un−u0)|2dx+b4(∫RN|∇(un−u0)|2dx)2+o(1)>lc1+l∞,c−c1, |
which is a contradiction with Lemma 2.2 (ii). Therefore, (2.3) holds. From (2.3) and un→u0 in L2loc(RN), we can find {yn}⊂RN such that ∫B1(yn)|un−u0|2dx→c0>0 and |yn|→∞. Let un(⋅+yn)⇀μ0 weakly in H1(RN). Note that μ0≠0 since c0>0. Therefore, {yn} and μ0 satisfy (2.2). Thus, the proof of Step 1 is complete.
Step 2: We show that {yn} and (u0,μ0) satisfy
limn→∞‖un−u0−μ0(⋅−yn)‖22=0. | (2.4) |
Since |yn|→∞, we have that
‖un−u0−μ0(⋅−yn)‖22=‖un‖22+‖u0‖22+‖μ0‖22−2⟨un,u0⟩L2−2⟨un(⋅+yn),μ0⟩L2+o(1)=‖un‖22−‖u0‖22−‖μ0‖22+o(1). | (2.5) |
According to (2.5), we could let δ1:=limn→∞‖un−u0−μ0(⋅−yn)‖22. Then, we have δ1=c−c1−c2, where c2:=‖μ0‖22. We want to show that δ1=0. Suppose on the contrary that δ1>0, by direct calculations we have
‖∇un‖22−‖∇u0‖22−‖∇μ0(⋅−yn)‖22−‖∇(un−u0−μ0(⋅−yn))‖22=−2‖∇u0‖22−2‖∇μ0‖22+2⟨∇un,∇u0⟩L2+2⟨∇un(⋅+yn),∇μ0⟩L2=o(1). | (2.6) |
From the Brezis-Lieb Lemma, we have
∫RN|un|pdx=∫RN|u0|pdx+∫RN|μ0(⋅−yn)|pdx+∫RN|un−u0−μ0(⋅−yn)|pdx+o(1). | (2.7) |
Similarly,
∫RNh|un|qdx=∫RNh|u0|qdx+∫RNh|μ0(⋅−yn)|qdx+∫RNh|(un−u0−μ0(⋅−yn))|qdx+o(1). | (2.8) |
Combining (2.6)–(2.8), we have
I(un)−I(u0)−I(μ0(⋅−yn))−I(un−u0−μ0(⋅−yn))=o(1). | (2.9) |
Since un⇀u0 in H1(RN), |yn|→∞ and h∈L22−q(RN), we have
∫RNh|un−u0−μ0(⋅−yn)|qdx→0. | (2.10) |
Recalling that l∞,c is continuous with respect to c>0 (see [33], Theorem 2.1), we have that
lim infn→∞I(un−u0−μ0(⋅−yn))=lim infn→∞I∞(un−u0−μ0(⋅−yn))≥l∞,δ1, | (2.11) |
and
lim infn→∞I(μ0(⋅−yn))≥l∞,c2. | (2.12) |
Hence by (2.9)–(2.12), we have
lc≥lc1+l∞,c2+l∞,δ1. | (2.13) |
However, using Lemma 2.1 (i), for any c2,δ1>0, there exists l∞,c2+δ1<l∞,c2+l∞,δ1. Hence, we also have
lc≥lc1+l∞,c2+l∞,δ1>lc1+l∞,c2+δ1≥lc1+c2+δ1=lc. | (2.14) |
This gives a contradiction and thus we have that δ1=0.
Step 3: Moreover, the following holds
I(u0)=lc1, I∞(μ0)=l∞,c2, | (2.15) |
and
lc=lc1+l∞,c2. | (2.16) |
By (2.9)–(2.12) and δ1=0, we have that
lc=limn→∞I(un)=lim infn→∞(I(u0)+I(μ0(⋅+yn)))≥I(u0)+I∞(μ0)≥lc1+l∞,c2. | (2.17) |
Combined with Lemma 2.2 (ii), we see that lc=lc1+l∞,c2. I(u0)=lc1 and I∞(μ0)=l∞,c2. Thus, Step 3 is proved.
Step 4: Now, we prove the precompactness of minimizing sequence, i.e., un→u0 in L2(RN).
We can suppose that {un} are nonnegative. Using the strong maximum principle, we have u0,μ0>0 and h(x)>0 on a set with positive measure, we have that
∫RNh|√u20+μ20|qdx>∫RNh|u0|qdx. |
Combine with the two following inequalities:
∫RN|∇√u20+μ20|2dx≤∫RN(|∇u0|2+|∇μ0|2)dx, | (2.18) |
∫RN|√u20+μ20|pdx≥∫RN(|u0|p+|μ0|p)dx. | (2.19) |
So we have
lc≤I(√u20+μ20)=a2∫RN|∇√u20+μ20|2dx+b4(∫RN|∇√u20+μ20|2dx)2−1p∫RN|√u20+μ20|pdx−1q∫RNh|√u20+μ20|qdx<I(u0)+I∞(μ0)=lc1+l∞,c−c1=lc, | (2.20) |
which is a contradiction. Thus the proof of Lemma 2.3 is completed.
Proof of Theorem 1.1. From Lemma 2.3, the minimizing sequence {un} satisfies un→u0 in L2(RN) and lc=I(u0), c=c1. Since {un}⊂Sc is the minimizing sequence of lc, we have dI|Sc(un)→0 and there exists a sequence of real numbers {λn} such that
I′(un)[φ]+λn∫RNunφdx→0,as n→∞, | (2.21) |
for every φ∈H1(RN). Hence, by (2.21), we have that
{−(a+b∫RN|∇u0|2dx)Δu0+ˉλu0=|u0|p−2u0+h(x)|u0|q−2u0 in RN,∫RN|u0|2dx=c. | (2.22) |
Notice that h(x)≥0, then by the maximum principle, u0>0, and we finish the proof of Theorem 1.1.
In this section, we study the mass-supercritical and Sobolev-subcritical case: 2+8N<p<2∗, 1≤N≤3, and h(x) satisfies the assumption (h2). First, we show that the energy functional I possesses a mountain pass geometry, which implies the existence of the (PS) sequence. Next, we prove that the limit of the sequence of the Lagrange multipliers related to the (PS) sequence is positive. Then, by applying the splitting lemma, we recover the compactness for this sequence, which yields the existence of solutions for Eq.(1.1).
In order to study the behavior of (PS) sequence, we introduce the splitting lemma, which plays a crucial role in overcoming the lack of compactness. For λ>0, we set
Iλ(u)=a2∫RN|∇u|2dx+b4(∫RN|∇u|2dx)2+12∫RNλu2dx−1p∫RN|u|pdx−1q∫RNh|u|qdx |
and
I∞,λ(u)=a2∫RN|∇u|2dx+b4(∫RN|∇u|2dx)2+12∫RNλu2dx−1p∫RN|u|pdx. |
Lemma 3.1. Let {un}⊂H1(RN) be a (PS) sequence for Iλ such that un⇀u in H1(RN) and limn→∞‖∇un‖22=A2. Then, there exists an integer k≥0, k nontrivial solutions ω1,⋯,ωk∈H1(RN) to the following problem
−(a+bA2)Δω+λω=|ω|p−2ω, | (3.1) |
and k sequences {yjn}⊂RN,1≤j≤k, such that as n→∞,|yjn|→∞,|yj1n−yj2n|→∞ for each 1≤j1,j2≤k,j1≠j2, and
‖un−u−k∑j=1ωj(⋅−yjn)‖→0, | (3.2) |
A2=‖∇u‖22+k∑j=1‖∇ωj‖22, | (3.3) |
‖un‖22=‖u‖22+k∑j=1‖wj‖22+o(1), | (3.4) |
and
Iλ(un)→Jh,λ(u)+k∑j=1J∞,λ(ωj), | (3.5) |
as n→∞ where
Jh,λ(u):=(a2+bA24)∫RN|∇u|2dx+λ2∫RNu2dx−1p∫RN|u|pdx−1q∫RNh|u|qdx |
and
J∞,λ(u):=(a2+bA24)∫RN|∇u|2dx+λ2∫RNu2dx−1p∫RN|u|pdx. |
Proof. The proof is similar to [34,Proposition 2.1] and [28,Lemma 1.6]; therefore, we omit it.
Lemma 3.2. Let X be a Hilbert manifold and let F∈C1(X,R) be a given functional. Let K⊆X be compact and consider a subset.
E⊂{E⊂X:E is compact, K⊂E}, |
which is invariant with respect to deformations leaving K fixed. Assume that
maxu∈KF(u)<c:=infE∈Emaxu∈EF(u)∈R. |
Let σn∈R be such that σn→0 and En∈E be a sequence such that
c≤maxu∈EnF(u)<c+σn. |
Then, there exists a sequence vn∈X such that
1. c≤F(vn)<c+σn,
2. ‖∇XF(vn)‖<˜c√σn,
3. dist(vn,En)<˜c√σn,
for some constant ˜c>0.
We shall prove that I on Sc possesses a kind of mountain pass geometrical structure. To this aim, we establish two preliminary lemmas.
Lemma 3.3. Assume that h∈Lpp−q(RN) and let u∈Sc be arbitrary but fixed. Then, we have:
(i) I(t⋆u)→0 as t→0;
(ii) I(t⋆u)→−∞ as t→+∞.
Proof. (i) By the Gagliardo-Nirenberg inequality (1.8), the Hölder inequality, and the assumption (h2), then we have that
|I(t⋆u)|≤a2∫RN|∇(t⋆u)|2dx+b4(∫RN|∇(t⋆u)|2dx)2+1p∫RN|t⋆u|pdx+1q∫RNh|t⋆u|qdx≤at22‖∇u‖22+bt44‖∇u‖42+tpγppCpN,pcp−pγp2‖∇u‖pγp2+1qtqγpCqN,pcq(1−γp)2‖h‖pp−q‖∇u‖qγp2→0 |
as t→0+, since pγp,qγp>0.
(ii) Similarly, we have that
I(t⋆u)≤at22‖∇u‖22+bt44‖∇u‖42−1p∫RN|t⋆u|pdx+1q∫RNh|t⋆u|qdx≤at22‖∇u‖22+bt44‖∇u‖42−tpγpp∫RN|u|pdx+1qtqγpCqN,pcq(1−γp)2‖h‖pp−q‖∇u‖qγp2→−∞ |
as t→+∞, since pγp>4.
Again, using the Gagliardo-Nirenberg inequality and the Hölder inequality,
I(u)≥a2‖∇u‖22+b4‖∇u‖42−1pCpN,pcp−pγp2‖∇u‖pγp2−1qCqN,pcq(1−γp)2‖h‖pp−q‖∇u‖qγp2≥a2‖∇u‖22−1pCpN,pcp−pγp2‖∇u‖pγp2−1qCqN,pcq(1−γp)2‖h‖pp−q‖∇u‖qγp2. | (3.6) |
To understand the geometry of the functional I on Sc, it is useful to consider the function φ:R+→R defined by
φ(t):=a2t2−1pCpN,pcp−pγp2tpγp−1qCqN,pcq(1−γp)2‖h‖pp−qtqγp. | (3.7) |
Since 0<qγp<2<pγp, we have that φ(0+)=0− and φ(+∞)=−∞. The role of assumption (1.13) is clarified by the following lemma.
Lemma 3.4. Under the assumption (h2), if (1.13) holds, then the function φ has a local strict minimum at negative level and a global strict maximum at positive level. Moreover, there exists 0<R1<R2, both depending on c, such that φ(R1)=0=φ(R2) and φ(t)>0 if and only if t∈(R1,R2).
Proof. For t>0, we see that φ(t)>0 if and only if
ψ(t)>1qCqN,pcq(1−γp)2‖h‖pp−q, |
where
ψ(t):=a2t2−qγp−1pCpN,pcp−pγp2tpγp−qγp. |
Observe that pγp−qγp>2−qγp>0, then ψ has a unique critical point ˉt on (0,+∞), which is a global maximum point at positive level. In fact, the expression of ˉt is
ˉt=(ap(2−qγp)2γp(p−q)CpN,pcp−pγp2)1pγp−2, |
and the maximum value of ψ is
ψ(ˉt)=a(pγp−2)2γp(p−q)(ap(2−qγp)2γp(p−q)CpN,p)2−qγppγp−2c−p(1−γp)(2−qγp)2(pγp−2). | (3.8) |
Therefore, if (1.13) holds, then ψ(ˉt)>1qCqN,pcq(1−γp)2‖h‖pp−q, thus the equation φ=0 has two roots R1,R2 and φ is positive on (R1,R2). Moreover, φ has a global maximum point t2 at positive level. According to the expression of φ, we can deduce that φ also has a local minimum point t1 at negative level in (0,R1).
Set
Aι:={u∈Sc:‖∇u‖2<ι},Ik:={u∈Sc:I(u)<k}. |
By Lemmas 3.3 and 3.4, there exists a ι1>0 small enough, such that
I(u)<12φ(t2), for any u∈Aι1. |
Moreover, Iφ(t1)⊂{‖∇u‖2>R2} since I(u)≥φ(‖∇u‖2). Now we can get a mountain pass structure of I on manifold Sc.
Γ:={γ∈C([0,1],Sc):γ(0)∈Aι,γ(1)∈Iφ(t1)}, | (3.9) |
and the mountain pass value is
mh,c:=infγ∈Γmaxt∈[0,1]I(γ(t)). | (3.10) |
Remark 3.1.
I∞(vc)=mc=infγ∈Γmaxt∈[0,1]I∞(γ(t)) |
where vc satisfies
{−(a+b∫RN|∇vc|2dx)Δvc+λvc=|vc|p−2vc in RN,∫RN|vc|2dx=c,u∈H1(RN), |
i.e., the solution vc of the problem (1.6) is a mountain pass critical point of I∞ constrained on Sc. (see [35]). It is immediately seen that
mh,c<mc. | (3.11) |
Lemma 3.5. Under the assumption (h2), suppose that h satisfies (1.14), then there exists a (PS) sequence {un} of I|Sc, which satisfies
I(un)→mh,c, | (3.12) |
I′∣Sc(un)→0, | (3.13) |
P(un)→0, | (3.14) |
as n→∞, where
P(u)=a‖∇u‖22+b‖∇u‖42−γp∫RN|u|pdx−γq∫RNh|u|qdx+1q∫RN⟨∇h,x⟩|u|qdx, |
and
limn→∞‖(un)−‖=0. | (3.15) |
We remark that (3.13) means that there exists {λn}n≥1, such that for any φ∈C∞0(RN), there holds
I′(un)[φ]+λn∫RNunφdx→0,as n→∞. | (3.16) |
Moreover, {un} is bounded in H1(RN) and the related Lagrange multipliers {λn} in (3.16) are also bounded, up to a subsequence, λn→ˉλ, with ˉλ>0.
Proof. We divide the proof into three steps.
Step 1: Existence of the Palais-Smale sequence. The existence of the (PS) sequence that verifies (3.14) and (3.15) closely follows the arguments in [32], where the authors adapt some ideas from [11]. We recall the main strategy, referring to [32] for the details. A key tool is to set
˜I(t,u):=I(t⋆u) for all (t,u)∈R×H1(RN). |
The corresponding minimax structure of ˜I on R×Sc, as follows
˜Γ:={γ=(γ1,γ2)∈C([0,1],R×Sc):γ(0)∈(0,Aι1),γ(1)∈(0,Iφ(t1))}, | (3.17) |
and its minimax value is
˜mh,c:=infγ∈˜Γmaxt∈[0,1]˜I(γ(t)). | (3.18) |
It turns out that ˜mh,c=mh,c and that, if (tn,vn)n is a (PS)c sequence for ˜I with tn→0, then un=tn⋆vn is a (PS)c sequence for I. Now, let us consider a sequence ξn∈Γ such that
mh,c≤maxt∈[0,1]I(ξn(t))<mh,c+1n. |
We observe that, since I(u)=I(|u|) for every u∈H1(RN), we can take ξn(t)≥0 in RN, for every t∈[0,1] and n∈N. We are in a position to apply Lemma 3.2 to ˜I with
X:=R×Sc,K:={(0,Aι1),(0,Iφ(t1))},E=˜Γ,En:={(0,ξn(t)):t∈[0,1]}. |
As a consequence, there exists a sequence (tn,vn)∈R×Sc and ˜c>0 such that
mh,c−1n<˜I(tn,vn)<mh,c+1n,mint∈[0,1]‖(tn,vn)−(0,ξn(t))‖R×H1(RN)<˜c√n,‖∇R×Sc˜I(tn,vn)‖<˜c√n. | (3.19) |
Now, we can define
un=tn⋆vn. |
We observe that, by differentiating ˜I with respect to t, we get the "almost" Pohozaev identity (3.14), differentiating with respect to the second variable on the tangent space to Sc, and by (3.19) and ξn(t)≥0 we get (3.15).
Step 2: Boundedness of the (PS) sequence.
By (3.12), for the (PS) sequence {un}⊂Sc, there holds
mh,c=I(un)+o(1)=a2‖∇un‖22+b4‖∇un‖42−1p∫RN|un|pdx−1q∫RNh|un|qdx+o(1). | (3.20) |
Combining with (3.14),
mh,c=a(N(p−2)−4)2N(p−2)‖∇un‖22+b(N(p−2)−8)4N(p−2)‖∇un‖42−p−qq(p−2)∫RNh|un|qdx−2qN(p−2)∫RN⟨∇h,x⟩|un|qdx+o(1)≥a(N(p−2)−4)2N(p−2)‖∇un‖22−p−qq(p−2)∫RNh|un|qdx−2qN(p−2)∫RN⟨∇h,x⟩|un|qdx+o(1)≥a(N(p−2)−4)2N(p−2)‖∇un‖22−p−qq(p−2)CqN,pcq(1−γp)2‖h‖pp−q‖∇un‖qγp2−2qN(p−2)‖∇h⋅x‖22−qcq2+o(1). | (3.21) |
Thus {un} is bounded in H1(RN) since h∈Lpp−q(RN) and ‖∇h⋅x‖22−q<∞.
Step3: Positivity of the Lagrange multiplier.
By taking un as a test function for (3.16), we obtain that
o(1)‖un‖H1=a‖∇un‖22+b‖∇un‖42−‖un‖pp−∫RNh|un|q+λnc. |
So
|λn|=1c|o(1)‖un‖H1−a‖∇un‖22−b‖∇un‖42+‖un‖pp+∫RNh|un|q|<+∞. |
Thus the Lagrange multipliers {λn} are also bounded. Next, we show that {λn} has a positive lower bound. In fact, according to (3.14) and (3.16),
λnc=λn∫RN|un|2dx=−a‖∇un‖22−b‖∇un‖42+‖un‖pp+∫RNh|un|qdx+o(1)=(1−γp)‖un‖pp+(1−γq)∫RNh|un|qdx+1q∫RN⟨∇h,x⟩|un|qdx+o(1). | (3.22) |
We also have that
mh,c=a2‖∇un‖22+b4‖∇un‖42−1p‖un‖pp−1q∫RNh|un|qdx+o(1)=−b4‖∇un‖42+N(p−2)−44p‖un‖pp+N(q−2)−44q∫RNh|un|qdx−12q∫RN⟨∇h,x⟩|un|qdx+o(1). | (3.23) |
Then, combined with the assumption (1.14), we have that
(3.24) |
since
Now we prove the convergence of the sequence and hence we complete the proof of Theorem 1.2.
Proof of Theorem 1.2. Next, we prove the existence of solutions of (1.1) with a positive energy level when . We consider the bounded sequence given by Lemma 3.5. Then, there exists such that due to the boundedness of . We claim that strongly in
For any , satisfies
Using the boundedness of again, we obtain that
And hence
which implies that is a sequence for at level , so that we can apply the Splitting Lemma 3.1, getting
Assume by contradiction that , or, equivalently, that . In addition, if , then and (see [28]). Therefore,
(3.25) |
where , . By (3.4), we have
Thus, combined with (3.25), we obtain
(3.26) |
Since , we have , which is a contradiction of (3.11). Thus . That is strongly in and is a solution of Eq.(1.1).
In this section, we assume that , , , and . By using a min-max argument, we can find the existence of normalized solutions of Eq.(1.1). First, we show that the energy functional corresponding to Eq.(1.1) has a linking geometry. For and , we introduce the scaling
which preserves the -norm: for all . For and , which will be determined later, we set
where is the closed ball of radius around 0 in For , define
where satisfies
We define
To prove that the energy functional has a linking geometry, it is necessary to find the suitable , such that
at least for some suitable choice of . Now, we recall the notion of barycenter of a function , which has been introduced in [36] and in [37]. Setting
we observe that is bounded and continuous, so the function
is well defined, continuous, and has compact support. Therefore, we can define as
The map is well defined, because has compact support, and it is not difficult to verify that it enjoys the following properties:
(i) is continuous in ;
(ii) if is a radial function, then ;
(iii) for all and for all ;
(iv) setting for and there holds .
Now, we define
and
It has been proved in [28] that
where
Lemma 4.1. .
Proof. Clearly , so that . It remains to prove that and .
Arguing by contradiction, we assume that . Then, for some , hence for some , here is the -neighborhood of . Observe that is open and connected, so it is path-connected. Therefore, there exists a path such that , a contradiction.
The inequality follows from the fact that the set satisfies
Lemma 4.2. .
Proof. Using and Lemma 4.1, we have
(4.1) |
Now, we argue by contradiction and assume that there exists a sequence such that
In view of (4.1), we also have
Adapting an argument from [11,Lemma 2.4], we consider the functional
constrained to . We apply Lemma 3.2 with
and
Observe that
because , hence , and for any we have and
hence . Hence, Lemma 3.2 yields a sequence such that
(1) as ;
(2) as ;
(3) as
Then is a sequence for on at , and there exists Lagrange multipliers such that
as . So, combining those properties, we can infer that
and
Therefore, is bounded in and is bounded in . We may assume that in , , and . In fact, is a sequence for at level As a consequence of Lemma 3.1, can be rewritten as
in , where and are solutions to
and . Moreover, we get
(4.2) |
(4.3) |
and hence,
By (4.2), we have
If and , we get from (4.3), we have
Similarly, we have . Thus,
we get a contradiction. Therefore, and , or and . If and , then . On the other hand, due to point (3) that , we obtain
which contradicts the fact that is continuous and .
If and , then in . Using again point (3), we also have . Hence, by the uniqueness, in . This implies
which is a contradiction.
Lemma 4.3. For any , then holds.
Proof. Similar to [32,Proposition 3.5], so we omit it.
Lemma 4.4. For any and for any , there exists and such that for with the following holds:
Proof. We have
and
Moreover, there holds
because satisfies (1.15), thus for all , we have
As a consequence, we deduce
provided is small enough and is large enough. Moreover, for large enough and , we choose such that , so that we have
The first integral is bounded by
as and
as , which concludes the proof.
By Lemma 4.3 and 4.4, we may choose and such that
Therefore, has a linking geometry and there exists a sequence at the level . In order to estimate , we have the following Lemma.
Lemma 4.5. If are large enough, then
Proof. This follows from
provided are large enough.
By the Lemma 4.3 and Lemma 4.5, we can get
Next, we construct a bounded sequence of I at by adopting the approach from [11] and Lemma 3.2. We define a auxiliary functional
and
Lemma 4.6. (1)
(2) If is a sequence for at level and , then is a sequence for at level .
Proof. The proof is similar to that of [11] and is omitted.
Lemma 4.7. Let be a sequence such that
Then, there exists a sequence and such that
The last inequality means:
for all
Proof. Apply Lemma 3.2 to with
Lemma 4.8. Under the assumption , then there exists a bounded sequence of , which satisfies
(4.4) |
(4.5) |
(4.6) |
as , where
(4.7) |
Moreover, the sequence of Lagrange multipliers satisfies, up to subsequence .
Proof. First, the existence of the sequence that verifies (4.6) and (4.7) closely follows the arguments in Lemma 3.5. The proof is omitted.
Next, we prove is bounded in By (4.4), for the sequence , there holds
(4.8) |
Combining with (4.6),
(4.9) |
Thus is bounded in , since .
Then, we prove the positivity of the Lagrange multiplier in the same way as lemma 3.5. By (4.5), we obtain that
Thus, the Lagrange multipliers are also bounded. In fact, according to (4.5) and (4.6), we have that
(4.10) |
thus provided
So
which is given in (1.15).
Proof of Theorem 1.3. Since is bounded, after passing to a subsequence it converges weakly in to . By (4.7) and weak convergence, is a nonnegative weak solution of
(4.11) |
such that , where . We note that is a bounded sequence of at level , therefore, by Lemma 3.1, there exists an integer , non-trivial solutions to the equation
and sequences , such that as .
Moreover, we have
(4.12) |
and
(4.13) |
as . It remains to show , so that strongly in and we are done. Thus, by contradiction, we can assume that , or equivalently .
First, we exclude the case . In fact, if and , we have and and so that (4.13) would give , which is not possible due to Lemma 5.3. On the other hand, if , we get and , thus , which contradicts with Lemma 4.5.
Therefore, from now on, we will assume and . From (4.13) and , we deduce
Using (4.12), we have
Then, from and , we have
Similarly, we have . Thus,
we get a contradiction. Thus and converges strongly to in .
Xin Qiu: Writing-original draft, Writing-review & editing; Zeng Qi Ou: Supervision, Formal Analysis; Ying Lv: Writing-review & editing, Methodology, Supervision.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This research is supported by National Natural Science Foundation of China (No.12371120), Innovation Research 2035 Pilot Plan of Southwest University(SWU-XDPY22015) and Chongqing Postgraduate Research and Innovation Programme(CYS240130).
The authors declare there is no conflict of interest.
[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] |
M. Cavalcanti, V. Cavalcanti, J. Soriano, Global existence and uniform decay rates for the Kirchhoff-Carrier equation with nonlinear dissipation, Adv. Differential Equations, 6 (2001), 701–730. https://doi.org/10.57262/ade/1357140586 doi: 10.57262/ade/1357140586
![]() |
[4] |
P. D'Ancona, 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
![]() |
[5] |
J. L. Lions, On some questions in boundary value problems of mathmatical physics, North-Holland Math. Stud., 30 (1978), 284–346. https://doi.org/10.1016/S0304-0208(08)70870-3 doi: 10.1016/S0304-0208(08)70870-3
![]() |
[6] |
G. Figueiredo, J. R. Santos, Multiplicity and concentration behavior of positive solutions for a Schrödinger-Kirchhoff type problem via penalization method, ESAIM Control Optim. Calc. Var., 20 (2014), 389–415. https://doi.org/10.1051/cocv/2013068 doi: 10.1051/cocv/2013068
![]() |
[7] |
Z. J. Guo, Ground states for Kirchhoff equations without compact condition, J. Differential Equations, 259 (2015), 2884–2902. https://doi.org/10.1016/j.jde.2015.04.005 doi: 10.1016/j.jde.2015.04.005
![]() |
[8] |
X. M. He, W. M. Zou, Ground states for nonlinear kirchhoff equations with critical growth, Ann. Mat. Pura Appl., 193 (2014), 473–500. https://doi.org/10.1007/s10231-012-0286-6 doi: 10.1007/s10231-012-0286-6
![]() |
[9] |
A. M. Mao, Z. T. Zhang, Sign-changing and multiple solutions of Kirchhoff type problems without the P.S. condition, Nonlinear Anal., 70 (2009), 1275–1287. https://doi.org/10.1016/j.na.2008.02.011 doi: 10.1016/j.na.2008.02.011
![]() |
[10] |
C. A. Stuart, Bifurcation for Dirichlet problems without eigenvalues, Proc. London Math. Soc., 45 (1982), 169–192. https://doi.org/10.1112/plms/s3-45.1.169 doi: 10.1112/plms/s3-45.1.169
![]() |
[11] |
L. Jeanjean, Existence of solutions with prescribed norm for semilinear elliptic equations, Nonlinear Anal., 28 (1997), 1633–1659. https://doi.org/10.1016/S0362-546X(96)00021-1 doi: 10.1016/S0362-546X(96)00021-1
![]() |
[12] |
T. Bartsch, N. Soave, A natural constraint approach to normalized solutions of nonlinear Schrödinger equations and systems, J. Funct. Anal., 272 (2017), 4998–5037. https://doi.org/10.1016/j.jfa.2017.01.025 doi: 10.1016/j.jfa.2017.01.025
![]() |
[13] | T. Bartsch, N. Soave, Multiple normalized solutions for a competing system of Schrödinger equations, Calc. Var. Partial Differential Equations, 58 (2019). https://doi.org/10.1007/s00526-018-1476-x |
[14] |
N. Soave, Normalized ground states for the NLS equation with combined nonlinearities, J. Differential Equations, 269 (2020), 6941–6987. https://doi.org/10.1016/j.jde.2020.05.016 doi: 10.1016/j.jde.2020.05.016
![]() |
[15] |
S. T. Chen, X. H. Tang, Normalized solutions for nonautonomous Schrödinger equations on a suitable manifold, J. Geom. Anal., 30 (2020), 1637–1660. https://doi.org/10.1007/s12220-019-00274-4 doi: 10.1007/s12220-019-00274-4
![]() |
[16] | Z. Chen, W. M. Zou, Existence of Normalized Positive Solutions for a Class of Nonhomogeneous Elliptic Equations, J. Geom. Anal., 33 (2023). https://doi.org/10.1007/s12220-023-01199-9 |
[17] | C. O. Alves, On existence of multiple normalized solutions to a class of elliptic problems in whole , Z. Angew. Math. Phys., 73 (2022). https://doi.org/10.1007/s00033-022-01741-9 |
[18] | D. M. Cao, E. S. Noussair, Multiplicity of positive and nodal solutions for nonlinear elliptic problems in , Ann. Inst. H. Poincar C Anal. Non Linaire., 13 (1996), 567–588. https://doi.org/10.1016/S0294-1449(16)30115-9 |
[19] | P. H. Zhang, Z. Q. Han, Normalized ground states for Kirchhoff equations in with a critical nonlinearity, J. Math. Phys., 63 (2022). https://doi.org/10.1063/5.0067520 |
[20] |
M. I. Weinstein, Nonlinear Schrödinger equations and sharp interpolation estimates, Commun. Math. Phys., 87 (1983), 567–576. https://doi.org/10.1007/BF01208265 doi: 10.1007/BF01208265
![]() |
[21] |
H. Y. Ye, The sharp existence of constrained minimizers for a class of nonlinear Kirchhoff equations, Math. Methods Appl. Sci., 38 (2015), 2663–2679. https://doi.org/10.1002/mma.3247 doi: 10.1002/mma.3247
![]() |
[22] | H. Y. Ye, The mass concentration phenomenon for -critical constrained problems related to Kirchhoff equations, Z. Angew. Math. Phys., 67 (2016). https://doi.org/10.1007/s00033-016-0624-4 |
[23] |
X. Y. Zeng, Y. M. Zhang, Existence and uniqueness of normalized solutions for the Kirchhoff equation, Appl. Math. Lett., 74 (2017), 52–59. https://doi.org/10.1016/j.aml.2017.05.012 doi: 10.1016/j.aml.2017.05.012
![]() |
[24] |
G. B. Li, X. Luo, T. Yang, Normalized solutions to a class of Kirchhoff equations with Sobolev critical exponent, Ann. Fenn. Math., 47 (2022), 895–925. https://doi.org/10.54330/afm.120247 doi: 10.54330/afm.120247
![]() |
[25] | P. C. Carrio, O. H. Miyagaki, A. Vicente, Normalized solutions of Kirchhoff equations with critical and subcritical nonlinearities: the defocusing case, Partial Differ. Equ. Appl., 3 (2022). https://doi.org/10.1007/s42985-022-00201-3 |
[26] |
H. Y. Ye, The existence of normalized solutions for -critical constrained problems related to Kirchhoff equations, Z. Angew. Math. Phys., 66 (2015), 1483–1497. https://doi.org/10.1007/s00033-014-0474-x doi: 10.1007/s00033-014-0474-x
![]() |
[27] |
S. T. Chen, V. Rdulescu, X. H. Tang, Normalized Solutions of nonautonomous Kirchhoff equations: sub- and super-critical cases, Appl. Math. Optim., 84 (2021), 773–806. https://doi.org/10.1007/s00245-020-09661-8 doi: 10.1007/s00245-020-09661-8
![]() |
[28] | L. Cai, F. B. Zhang, Normalized Solutions of Mass Supercritical Kirchhoff Equation with Potential, J. Geom. Anal., 33 (2023). https://doi.org/10.1007/s12220-022-01148-y |
[29] | A. Fiscella, A. Pinamonti, Existence and multiplicity results for Kirchhoff-type problems on a double-phase setting, Mediterr. J. Math., 20 (2023). https://doi.org/10.1007/s00009-022-02245-6 |
[30] |
A. Fiscella, G. Marino, A. Pinamonti, S. Verzellesi, Multiple solutions for nonlinear boundary value problems of Kirchhoff type on a double phase setting, Rev. Mat. Complut., 37 (2024), 205–236. https://doi.org/10.1007/s13163-022-00453-y doi: 10.1007/s13163-022-00453-y
![]() |
[31] |
W. H. Xie, H. B. Chen, Existence and multiplicity of normalized solutions for the nonlinear Kirchhoff type problems, Comput. Math. Appl., 76 (2018), 579–591. https://doi.org/10.1016/j.camwa.2018.04.038 doi: 10.1016/j.camwa.2018.04.038
![]() |
[32] |
T. Bartsch, R. Molle, M. Rizzi, M. Verzini, Normalized solutions of mass supercritical Schrödinger equations with potential, Comm. Partial Differential Equations, 46 (2021), 1729–1756. https://doi.org/10.1080/03605302.2021.1893747 doi: 10.1080/03605302.2021.1893747
![]() |
[33] |
J. Bellazzini, G. Siciliano, Scaling properties of functionals and existence of constrained minimizers, J. Funct. Anal., 261 (2011), 2486–2507. https://doi.org/10.1016/j.jfa.2011.06.014 doi: 10.1016/j.jfa.2011.06.014
![]() |
[34] |
Q. L. Xie, S. W. Ma, X. Zhang, Bound state solutions of Kirchhoff type problems with critical exponent, J. Differential Equations, 261 (2016), 890–924. https://doi.org/10.1016/j.jde.2016.03.028 doi: 10.1016/j.jde.2016.03.028
![]() |
[35] | Q. Wang, A. Qian, Normalized Solutions to the Kirchhoff Equation with Potential Term: Mass Super-Critical Case, Bull. Malays. Math. Sci. Soc., 46 (2023). https://doi.org/10.1007/s40840-022-01444-4 |
[36] | T. Bartsch, T. Weth, Three nodal solutions of singularly perturbed elliptic equations on domains without topology, Ann. Inst. H. Poincar C Anal. Non Linaire., 22 (2005), 259–281. https://doi.org/10.1016/j.anihpc.2004.07.005 |
[37] |
G. Cerami, D. Passaseo, The effect of concentrating potentials in some singularly perturbed problems, Calc. Var. Partial Differential Equations, 17 (2003), 257–281. https://doi.org/10.1007/s00526-002-0169-6 doi: 10.1007/s00526-002-0169-6
![]() |