In this paper we are concerned with the Lane-Emden-Fowler equation
{−Δu=un+2n−2−εinΩ,u>0inΩ,u=0on∂Ω,
where Ω⊂Rn (n≥3) is a nonconvex polygonal domain and ε>0. We study the asymptotic behavior of minimal energy solutions as ε>0 goes to zero. A main part is to show that the solution is uniformly bounded near the boundary with respect to ε>0. The moving plane method is difficult to apply for the nonconvex polygonal domain. To get around this difficulty, we derive a contradiction after assuming that the solution blows up near the boundary by using the Pohozaev identity and the Green's function.
Citation: Woocheol Choi. Energy minimizing solutions to slightly subcritical elliptic problems on nonconvex polygonal domains[J]. AIMS Mathematics, 2023, 8(11): 26134-26152. doi: 10.3934/math.20231332
[1] | Yasir Nadeem Anjam . Singularities and regularity of stationary Stokes and Navier-Stokes equations on polygonal domains and their treatments. AIMS Mathematics, 2020, 5(1): 440-466. doi: 10.3934/math.2020030 |
[2] | S. S. Santra, S. Priyadharshini, V. Sadhasivam, J. Kavitha, U. Fernandez-Gamiz, S. Noeiaghdam, K. M. Khedher . On the oscillation of certain class of conformable Emden-Fowler type elliptic partial differential equations. AIMS Mathematics, 2023, 8(6): 12622-12636. doi: 10.3934/math.2023634 |
[3] | Huafei Di, Yadong Shang, Jiali Yu . Blow-up analysis of a nonlinear pseudo-parabolic equation with memory term. AIMS Mathematics, 2020, 5(4): 3408-3422. doi: 10.3934/math.2020220 |
[4] | Cagnur Corekli . The SIPG method of Dirichlet boundary optimal control problems with weakly imposed boundary conditions. AIMS Mathematics, 2022, 7(4): 6711-6742. doi: 10.3934/math.2022375 |
[5] | Yaxin Zhao, Xiulan Wu . Asymptotic behavior and blow-up of solutions for a nonlocal parabolic equation with a special diffusion process. AIMS Mathematics, 2024, 9(8): 22883-22909. doi: 10.3934/math.20241113 |
[6] | Xinyue Zhang, Haibo Chen, Jie Yang . Blow up behavior of minimizers for a fractional p-Laplace problem with external potentials and mass critical nonlinearity. AIMS Mathematics, 2025, 10(2): 3597-3622. doi: 10.3934/math.2025166 |
[7] | Keqiang Li, Shangjiu Wang, Shaoyong Li . Symmetry of large solutions for semilinear elliptic equations in a symmetric convex domain. AIMS Mathematics, 2022, 7(6): 10860-10866. doi: 10.3934/math.2022607 |
[8] | Meixia Cai, Hui Jian, Min Gong . Global existence, blow-up and stability of standing waves for the Schrödinger-Choquard equation with harmonic potential. AIMS Mathematics, 2024, 9(1): 495-520. doi: 10.3934/math.2024027 |
[9] | Li-Ming Yeh . Lipschitz estimate for elliptic equations with oscillatory coefficients. AIMS Mathematics, 2024, 9(10): 29135-29166. doi: 10.3934/math.20241413 |
[10] | Sobajima Motohiro, Wakasugi Yuta . Remarks on an elliptic problem arising in weighted energy estimates for wave equations with space-dependent damping term in an exterior domain. AIMS Mathematics, 2017, 2(1): 1-15. doi: 10.3934/Math.2017.1.1 |
In this paper we are concerned with the Lane-Emden-Fowler equation
{−Δu=un+2n−2−εinΩ,u>0inΩ,u=0on∂Ω,
where Ω⊂Rn (n≥3) is a nonconvex polygonal domain and ε>0. We study the asymptotic behavior of minimal energy solutions as ε>0 goes to zero. A main part is to show that the solution is uniformly bounded near the boundary with respect to ε>0. The moving plane method is difficult to apply for the nonconvex polygonal domain. To get around this difficulty, we derive a contradiction after assuming that the solution blows up near the boundary by using the Pohozaev identity and the Green's function.
In this paper we study asymptotic profile of energy minimizing solutions to the Lane-Emden-Fowler equation
{−Δuε=up−εεinΩ,uε>0inΩ,uε=0on∂Ω, | (1.1) |
as ε>0 goes to zero. Here Ω⊂Rn (n≥3) is a bounded polygonal domain and p=n+2n−2 is the critical exponent. In the seminar papers Han [1] and Rey [2], the asymptotic behavior of energy minimizing solutions to (1.1) was obtained for smooth bounded domains Ω.
The asymptotic behavior was first studied by Atkinson and Peletier [3] when Ω is the unit ball in R3 using an ODE argument. The result was revisited by Brezis and Peletier [4] by applying PDE methods. Extensions to smooth bounded domains were obtained by Han [1] and Rey [2]. The asymptotic behavior have been studied by a lot of researchers for nonlinear elliptic equations with various settings (see e.g., [5,6,7,8,9,10,11,12,13,14,15,16]) and we note that most of the results have been obtained for elliptic problems on bounded smooth domains.
Pistoia and Rey [17] showed that as for problem (1.1) posed on a specific nonsmooth bounded domain constructed by Flucher-Garroni-Müller [18], the maximum point of uε may approach to the boundary point as ε→0. By the way, we mention that the arguments of Han [1] and Rey [2] work straight-forwardly for convex bounded domains, which may not be non-smooth. In fact a key part in the analysis of Han [1] and Rey [2] is that the maximum point of uε(x) is uniformly away from the boundary ∂Ω by showing that the solutions uε(x) are uniformly bounded for ε>0 and x near the boundary ∂Ω by the moving plane argument. If Ω is a smooth nonconvex domain, Han [1] obtained the unfirom boundedness by using the Kelvin transform to (1.1) on balls which touch the domain Ω by the boundary ∂Ω. However, the argument is difficult to apply when Ω is not smooth.
Given this result, a natural question is that can we extend the result of Han [2] and Rey [2] to certain class of nonsmooth convex domains? In this paper, we show that the results of Han [1] and Rey [2] to nonconvex polygonal domains. The following is the main result of this paper.
Theorem 1.1. For n≥3 we let Ω⊂Rn be a bounded polygonal domain. Assume that {uε}ε>0 is a set of solutions to (1.1) such that
limε→0(∫Ω|uε|p+1−εdx)1p+1−ε(∫Ω|∇uε|2dx)1/2=Sn, | (1.2) |
where Sn=[πn(n−2)Γ(n/2)/Γ(n)]−1 is the best Sobolev constant in Rn. Then the family of solutions {uε}ε>0 are uniformly bounded near the boundary, i.e., there are costants δ>0 and C>0 independent of ε>0 such that
supε>0sup{x∈Ω:dist(x,∂Ω)<δ}|uε(x)|≤C. |
Given the boundary estimates of Theorem 1.1, one may apply standard argument to deduce the following result [1].
Theorem 1.2. For n≥3 we let Ω⊂Rn be a bounded polygonal domain. Assume that {uε}ε>0 is a set of solutions to (1.1) such that (1.2) holds. Then, there exists a point x0∈Ω such that, up to a subsequence,
● The solution uε converges to 0 in C1(Ω∖{x0}).
● ∇R(x0)=0, where R(x)=H(x,x).
● We have
limε→0‖uε‖L∞(Ω)uε(x)=[n(n−2)](n−2)/2|Sn−1|G(x,x0). |
● We have
limε→0ε‖uε‖2L∞(Ω)=(n−2)|Sn−1|2[n(n−2)]n−2H(x0,x0). |
Here G denotes the Green's function and H is the regular part of G (see Section 2 for the detail).
In order to prove Theorem 1.1 we assume that contrary that the maximum point xε approaches to the boundary. Under this assumption, we shall deduce a contradiction from the following Pohozaev type identity on an annulus centered at the blow up point; 1≤j≤n,
∫∂B(xε,2dε)|∇uε|2νj−2(∂uε∂ν∂uε∂xj)dSx=2p+1−ε∫∂B(xε,2dε)up−ε+1ενjdSx, | (1.3) |
where xε∈Ω is the maximum point of uε and dε=dist(xε,∂Ωε)/4.
In fact we shall prove Theorem 1.1 for more general domain Ω satisfying the following assumption.
Assumption D. Consider a sequence of points {xk}k∈N in the domain Ω such that dk:=dist(xk,∂Ω) goes to zero as k→∞. Take zk∈∂Ω such that |xk−zk|=dk. Let Ωk:=1dk(Ω−zk). Note that we have 0∈Ωk, and also 1dk(xk−zk)∈Sn−1. Thus we can find a rotation Rk:Rn→Rn such that
Rk(1dk(xk−zk))=en=(0,⋯,0,1). |
Then, the domain Dk:=RkΩk converges to an infinite star-shaped domain P⊊Rn.
It is not difficult to see that any bounded polygonal domain Ω satisfies the above assumption. Under the above assumption we will obtain the following result on the regular part H of the Green's function.
Theorem 1.3. For n≥3 we let Ω⊂Rn be a bounded open domain satisfying Assumption D. Then, for any sequence of points {yk}k≥1 in Ω such that limk→∞dk=0, where dk:=dist(yk,∂Ω), there exists a constant c>0 and N∈N such that, for k≥N we have
sup1≤j≤n|∂H∂xj(yk,yk)|≥cdn−1k. | (1.4) |
If Ω is smooth, then the result of Theorem 1.3 was proved in Rey [19] by applying the Maximum principle. To obtain the above inequality for the nonsmooth domains, we shall rescale the function H in a suitable way and investigate its limit.
This paper is organized as follows. In Section 2, we are concerned about the properties of Green's function. Also we show that a sequence of the minimal energy solutions blows up as ε→0 and that the blow up point does not approach to the boundary too fast in some sense (see Lemma 2.2). In Section 3, we will obtain a sharp estimate of the function uε on an annulus centered at the blow up point. In Section 4, we prove Theorem 1.1. In Section 5, we give a proof of Theorem 1.2. Section 6 is devoted to prove Theorem 1.3.
Notations.
Here we list some notations which will be used throughout the paper.
- C>0 is a generic constant that may vary from line to line.
- For k∈N we denote by Bk(x0,r) the ball {x∈Rk:|x−x0|<r} for each x0∈Rk and r>0.
- For x∈Ω we denote by dist(x,∂Ω) the distance from x to ∂Ω and we denote d(x):=dist(x,∂Ω).
- For a domain D⊂Rn, the map ν=(ν1,⋯,νn):∂D→Rn denotes the outward pointing unit normal vector on ∂D.
- dS stands for the surface measure.
- |Sn−1|=2πn/2/Γ(n/2) denotes the Lebesgue measure of (n−1)-dimensional unit sphere Sn−1.
In this section we obtain preliminary results for a sequence of the solutions {uε}ε>0 satifsying (1.2). For this purpose, we first recall Green's function G of the Laplacian −Δ on Ω with the Dirichlet boundary condition. It is divided into a singular part and a regular part as
G(x,y)=cn|x−y|n−2−H(x,y), | (2.1) |
where cn=1/(n−2)|Sn−1| and the regular part H:Ω×Ω→R is the function such that
{−ΔxH(x,y)=0x∈Ω,H(x,y)=cn|x−y|n−2x∈∂Ω. | (2.2) |
Let d(x)=dist(x,∂Ω) for x∈Ω. Take a small constant δ>0.
We take a value λϵ>0 and a point xϵ∈Ω such that
λ2p−ε−1ϵ:=uε(xε)=maxx∈Ω{uϵ(x)}. | (2.3) |
Now we recall the sharp Sobolev embedding
(∫Rn|f(x)|2nn−2dx)n−22n≤Sn(∫Rn|∇f(x)|2dx)1/2∀f∈H1(Rn). | (2.4) |
If we replace the function f by (−Δ)−1/2f in the above inequality, we find the Hardy-Littlewood-Sobolev inequality:
‖(−Δ)−1/2f‖Lp+1(Rn)≤Sn‖f‖L2(Rn)∀f∈L2(Rn). | (2.5) |
We let K denote Green's function of the Laplacain on Rn, i.e.,
K(x,y)=cn|x−y|n−1. |
The estimate (2.5) is then written as
‖∫RnK(x,y)f(y)dy‖Lp+1(Rn)≤Sn‖f‖L2(Rn)∀f∈L2(Rn). |
For given a domain Q⊂Rn we denote by KQ:Q×Q→R Green's function of the Laplacian (−Δ)1/2 on domain Q with the Dirichlet zero boundary condition, i.e., for the solution u∈H1(Ω) to the problem
{(−Δ)1/2u=finΩu=0on∂Ω, |
with f∈L2(Ω) admits the representation
u(x)=∫ΩK(x,y)f(y)dy. |
Then, it is a classical fact that for any open subset Q⊂Rn with Q≠Rn, we have
KQ(x,y)<K(x,y)for all(x,y)∈Q×Q. | (2.6) |
Here we remark that (−Δ)1/2 is defined by the spectral decomposition of (−Δ) on domain Ω.
Lemma 2.1. The value λε>0 defined in (2.3) satisfies limε→0λε=∞.
Proof. In order to prove the lemma, we assume the contrary. Then there is a subsequence {εk}k∈N such that limk→∞εk=0 and supk∈Nλϵk<∞. This implies that the solutions {uϵk}k∈N are uniformly bounded in C1,α(Ω) for some α∈(0,1) by the standard regularity theory applied to (1.1). Up to a subsequence, the solution uϵk converges in C1(Ω) to a function u0∈C1(Ω), and taking k→∞ in the formula
uεk(x)=∫ΩG(x,y)up−εkεk(y)dy, |
we find
u0(x)=∫ΩG(x,y)up0(y)dy, |
and so
{−Δu0=up0inΩ,u0=0on∂Ω. | (2.7) |
On the other hand, by taking the limit k→∞ in (1.2) we get
‖u0‖Lp+1(Ω)=Sn‖∇u0‖L2(Ω). |
Let us set w0:Ω→¯R+ by w0(x)=(−ΔΩ)1/2u0(x) for x∈Ω. Then u0(x)=(−ΔΩ)−1/2w0(x) for x∈Ω and so we have
‖(−ΔΩ)−1/2w0‖Lp+1(Ω)=Sn‖w0‖L2(Ω). | (2.8) |
We extend the function w0 to set W0:Rn→¯R+ by
W0(x)={w0(x)forx∈Ω,0forx∉Ω. |
Then, using the inequality (2.6) and (2.8) we obtain the following estimate
Sn‖W0‖L2(Rn)=Sn‖w0‖L2(Ω)=‖(−ΔΩ)−1/2w0‖Lp+1(Ω)<‖(−ΔΩ)−1/2W0‖Lp+1(Ω)<‖(−Δ)−1/2W0‖Lp+1(Rn). |
However, this contradicts to the optimality of the constant Sn of the inequality (2.5). Therefore it should hold that limϵ→0λϵ=∞. The lemma is proved.
For each ε>0 we set Ωϵ:=λϵ(Ω−xϵ) and normalize the solution uε as follows
Uϵ(x):=λ−2p−ε−1ϵuϵ(λ−1ϵx+xϵ), | (2.9) |
so that
{−ΔUϵ=Up−εϵinΩϵ,Uϵ=0on∂Ωϵ, | (2.10) |
and maxx∈Ωϵ{Uϵ(x)}=1=Uϵ(0). In the next lemma, we obtain an estimate for the distance between the maximum point of the solutions and the boundary ∂Ω.
Lemma 2.2. We have limϵ→0λϵdist(xϵ,∂Ω)=∞.
Proof. We assume the contrary. Then, up to a subsequence, we have limϵ→0λϵdist(xϵ,∂Ω)=l for some l∈(0,∞). This implies that the extended domain Ωϵ converges to a infinite star-shaped domain P⊊Rn as ε→0. Also, the normalized functions Uϵ converge to a nontrivial solution ¯U in C2loc(P) of the problem
{−Δ¯U=¯UpinP,¯U=0on∂P, |
and we know that KP(x,y)<K(x,y) from (2.6). Then we can obtain a contradiction as in the proof of Lemma 2.1. Thus the result of the lemma is true.
We set dϵ:=14dist(xϵ,∂Ω) and Nε=dελε. Then we see from Lemma 2.2 that
dϵ=Nϵλϵandlimε→0Nϵ=∞. | (2.11) |
We remark that the fact Nε→∞ as ε→0 will be important in the proofs of Theorem 1.1. By Lemma 2.2 the domain Ωϵ converges to Rn as ϵ goes to zero, and so the rescaled solution Uϵ converges in C2loc(Rn) to a solution U of the problem
{−ΔU=UpinRn,U(y)>0y∈Rn,U(0)=1=maxx∈RnU(x),U→0as|y|→∞. | (2.12) |
Then it is well-known that the function U is equal to
U(x)=[n(n−2)](n−2)/4(ηη2+|x|2)(n−2)/2, |
where η=√n(n−2). Next we recall the following result from Corollary 1 and Lemma 3 in [1].
Lemma 2.3 ([1]). The value λε>0 defined in (1.2) and the rescaled solution Uε defined (2.9) satisfy the following.
(1) There is a constant C>0 independent of ϵ>0 such that
λϵϵ≤C. | (2.13) |
(2) There exists a constant C>0 such that
Uϵ(x)≤CU(x)∀ϵ>0. | (2.14) |
We end this section with a local version of the Pohozaev type identity for the problem (1.1).
Lemma 2.4. Let 1≤j≤n. Suppose that u∈C2(Ω)×C2(Ω) is a solution of (1.1). Then, for any open smooth subset D⊂Ω, we have the following identity.
−2∫∂D∂u∂ν(x)∂u∂xj(x)dSx+∫∂D|∇u(x)|2νjdSx=2p+1∫∂Dup+1(x)νjdSx, | (2.15) |
where D is an open subset of Ω.
Proof. Multiplying (1.1) by ∂u∂xj we get −Δu∂u∂xj=up∂u∂xj. Integrating this over the domain D and using an integration by part, we get
−∫∂D∂u∂ν∂u∂xjdSx+∫D∇u⋅∂∇u∂xjdSx=1p+1∫∂Dup+1νjdSx. | (2.16) |
We use an integration by parts to get
12∫D∂∂xj|∇u|2dx=12∫∂D|∇u|2νjdSx. |
The lemma is proved.
This section is devoted to prove the following lemma regarding a sharp estimate for uε and its derivatives on the annulus ∂B(xε,2dε).
Lemma 3.1. Assume that {uε}ε>0 is a sequence of solutions to (1.1) of type (ME) and that limε→0dε=0. Then, for x∈∂B(xε,2dε) we have the estimates
uϵ(x)=AUλ−[2−(n−2)ε]p−ε−1ϵG(x,xϵ)+o(d−(n−2)ϵλ−np+1ϵ) | (3.1) |
and
∇uϵ(x)=AUλ−[2−(n−2)ε]p−ε−1ϵ∇G(x,xϵ)+o(d−(n−1)ϵλ−np+1ϵ). | (3.2) |
Here the value AU is defined as
AU=∫RnUp(y)dy=[n(n−2)]n2cnn=[n(n−2)]n2−1|Sn−1|. | (3.3) |
In addition, the o-notation is uniform with respect to x∈∂B(xε,2dε), i.e., it holds that
limε→0supx∈∂B(xε,2dε)|o(d−kελ−np+1ε)|(d−kελ−np+1ε)=0fork=n−1orn−2. |
Proof. Since uϵ is a solution to (1.1), we have
uϵ(x)=∫ΩG(x,y)upϵ(y)dy=G(x,xϵ)(∫Ωuqϵ(y)dy)+∫Ω[G(x,y)−G(x,xϵ)]upϵ(y)dy. | (3.4) |
Given the estimate (2.14) we apply the dominated convergence theorem to find
limϵ→0λ[2−(n−2)ε]p−ε−1ϵ∫Ωup−εϵ(y)dy=limϵ→0∫ΩϵUp−εϵ(y)dy=∫RnUp(y)dy=AU. |
Using this and noting that G(x,xε)=O(|x−xε|−(n−2))=O(d−(n−2)ε) for x∈∂B(xε,2dε), we find
G(x,xε)(∫Ωup−εε(y)dy)=λ−[2−(n−2)ε]p−ε−1εAUG(x,xε)+o(λ−np+1εd−(n−2)ε), |
where we also used that
λ−[2−(n−2)ε]p−1−εε=O(λ−np+1ε |
due to the fact that 2p−1=np+1 and (2.13). Similarly, we may deduce
∇G(x,xε)(∫Ωup−εε(y)dy)=λ−[2−(n−2)ε]p−ε−1εAU∇G(x,xε)+o(λ−np+1εd−(n−1)ε). |
Therefore, in order to prove (3.1), we only need to estimate the last term of (3.4) as o(d−(n−2)ελ−np+1ε) and its derivatives as o(d−(n−1)ελ−np+1ε). For this aim, we decompose it into three parts as follows:
∫Ω[G(x,y)−G(x,xϵ)]up−εϵ(y)dy=I1(x)+I2(x)+I3(x), | (3.5) |
where
I1(x):=∫B(xϵ,dϵ)[G(x,y)−G(x,xϵ)]up−εϵ(y)dy,I2(x):=∫B(xϵ,4dϵ)∖B(xϵ,dϵ)[G(x,y)−G(x,xϵ)]up−εϵ(y)dy,I3(x):=∫Ω∖B(xϵ,4dϵ)[G(x,y)−G(x,xϵ)]up−εϵ(y)dy. | (3.6) |
We shall show that I1(x), I2(x), and I3(x) are estimated as o(d−(n−2)ελ−np+1ε) and their derivatives ∇I1(x), ∇I2(x), and ∇I3(x) are estimated as o(d−(n−1)ελ−np+1ε).
Estimate of I1. Since x∈∂B(xε,2dε), we have |x−y|≥dϵ for y∈B(xϵ,dϵ), and so
|∇yG(x,y)|≤Cd−(n−1)ϵand|∇x∇yG(x,y)|≤Cd−nε∀y∈B(xϵ,dϵ). |
Combining this with the mean value formula yields
|G(x,y)−G(x,xϵ)|≤C|y−xϵ|d−(n−1)ϵand|∇xG(x,y)−∇xG(x,xϵ)|≤C|y−xϵ|d−nϵ | (3.7) |
for all y∈B(xε,dε). Applying this and (2.14) we may estimate I1 as follows:
I1(x)≤Cd−(n−1)ϵ∫B(xϵ,dϵ/2)|y−xϵ|λ2(p−ε)p−ε−1ϵUp(λϵ(y−xϵ))dy≤Cd−(n−1)ϵλ2(p−ε)p−ε−1ϵλ−(n+1)ϵ∫B(0,Nϵ/2)|y|Uq(y)dy. | (3.8) |
Using (2.13) and that 2pp−1−(n+1)<−np+1 we find that I1(x)=o(d−(n−2)ελ−np+1ε). By the same way along with the second inequality of (3.7), we can obtain the estimate
∇I1(x)=o(d−(n−1)ελ−np+1ε). |
Estimate of I2. For y∈B(xϵ,4dϵ)∖B(xϵ,dϵ) we use the estimate (2.14) and (2.13) to find
uϵ(y)≤Cλnp+1εU(λϵ(y−xϵ))≤Cλnp+1−(n−2)εd−(n−2)ε. | (3.9) |
Noting that
|x−y|≤8dεfory∈B(xε,4dε)andx∈∂B(xε,2dε), | (3.10) |
we have
{|G(x,y)|+|G(x,xϵ)|≤cn|x−y|n−2+cnd(n−2)ϵ≤C|x−y|n−2,|∇xG(x,y)|+|∇xG(x,xϵ)|≤cn|x−y|n−1+cnd(n−1)ϵ≤C|x−y|n−1. | (3.11) |
Combining the first estimate of (3.11), (3.10) and (3.9) in (3.6) yields
I2(x)≤Cλpnp+1εd−(n−2)pελ−(n−2)pε∫B(xϵ,4dϵ)∖B(xϵ,dϵ)1|x−y|n−2dy≤Cλpnp+1εd2−(n−2)pελ−(n−2)pε=Cλ−np+1εd−(n−2)εNn−(n−2)pε. |
Due to the fact that p=n+2n−2 the above estimate gives the estimate I2(x)=o(λ−np+1εd−(n−2)ε). Similarly, using the second estimate of (3.11), we obtain
∇I2(x)=O(λ−np+1εd−(n−1)εNn−(n−2)pε)=o(λ−np+1εd−(n−1)ε). |
Estimate of I3. Since |x−xϵ|=2dϵ, we have the following estimates
{|G(x,y)−G(x,xϵ)|≤Cd−(n−2)ϵfory∈Ω∖B(xϵ,4dϵ),|∇xG(x,y)−∇xG(x,xϵ)|≤Cd−(n−1)ϵfory∈Ω∖B(xϵ,4dϵ). | (3.12) |
Applying the first inequality of (3.12), we have
I3(x)≤Cd−(n−2)ϵ∫Ω∖B(xϵ,4dϵ)up−εϵ(y)dy. |
Using (2.14) we deduce
∫Ω∖B(xϵ,4dϵ)up−εϵ(y)dy=λ−np−ε+1ϵ∫Ωϵ∖B(0,4Nϵ)Up−εϵ(y)dy≤Cλ−np−ε+1ϵ∫Rn∖B(0,4Nϵ)Up(y)dy≤Cλ−np+1ϵN−(n−2)p+nϵ. |
Combining the above two estimates, we arrive at the following estimate
I3(x)≤Cd−(n−2)ϵλ−np+1ϵN−(n−2)p+nϵ=o(d−(n−2)ϵλ−np+1ϵ), | (3.13) |
where the fact that (n−2)p>n was also used. Similarly, using the second estimate of (3.12), then we get
∇I3(x)=O(d−(n−1)ϵλ−np+1ϵN−(n−2)p+nϵ)=o(d−(n−1)ϵλ−np+1ϵ). |
Finally, gathering the above estimates on I1, I2, and I3, we finally get
I1(x)+I2(x)+I3(x)=o(d−(n−2)ϵλ−np+1ϵ) |
and
|∇xI1(x)|+|∇xI2(x)|+|∇xI3(x)|=o(d−(n−1)ϵλ−np+1ϵ). |
The lemma is proved.
This section is devoted to prove Theorem 1.1.
Proof of Theorem 1.1. Let dϵ=dist(xϵ,∂Ω). In view of (2.9), it is enough to show that infε>0dε>0. For this purpose, with a view to a contradiction, we assume the contrary that dϵ→0 as ϵ→0 in a subsequence.
We use the notation λε and Nε=dελε defined in (2.3) and (2.11). Then we recall from Lemma 2.2 that we have Nϵ→∞. Let us set Dϵ=B(xϵ,2dϵ) for each 1≤j≤n and we define the values Ljϵ and Rjϵ by
Ljϵ:=−2∫∂Dϵ∂uϵ∂ν∂uϵ∂xj(x)dSx+∫∂Dϵ|∇uϵ|2νjdSx,Rjϵ:=n−2n∫∂Dϵup+1ϵνjdSx. |
Applying Lemma 2.4 to uε with D=Dϵ, we find that
Ljϵ=Rjϵ. |
In what follows, we proceed to obtain sharp estimates of the values of Lϵj and Rϵj, which will lead to a contradiction.
First, we compute Ljϵ using the expression (3.2) as follows.
Lϵj=−2λ−2p−1−εϵA2U∫∂Dϵ(∂∂νG(x,xϵ)∂∂xjG(x,xϵ))dSx+λ−2p−1−εϵA2U∫∂Dϵ|∇G(x,xϵ)|2νjdSx+o(|∂Dϵ|λ−2np+1ϵd−2(n−1)ϵ)=−λ−2p−1−εϵA2UI(2dϵ)+o(d−(n−1)ϵλ−(n−2)ϵ), | (4.1) |
where we have set
I(r):=[∫∂B(xϵ,r)2∂G∂ν(x,xϵ)∂∂xjG(x,xϵ)−|∇G(x,xϵ)|2νjdSx]forr>0. |
In order to compute the value of I(2dϵ), we first notice that I(r) is independent of r>0. Indeed, it follows from that −ΔxG(x,xϵ)=0 for x∈Ar:=B(xε,2dε)∖B(xε,r) for each r∈(0,2dε), and an integration by parts performed as follows:
0=∫Ar(−ΔxG)(x,xϵ)∂G∂xj(x,xϵ)dx=−∫∂Ar∂G∂ν(x,xϵ)∂G∂xj(x,xϵ)dSx+∫Ar∇xG(x,xϵ)∂∇xG∂xj(x,xϵ)dx=−∫∂Ar∂G∂ν(x,xϵ)∂G∂xj(x,xϵ)dSx+12∫∂Ar|∇xG(x,xϵ)|2νjdSx, | (4.2) |
which means that I(r) is constant on (0,2dϵ]. Therefore we can evaluate I(2dε) by computing the following limit;
I(2dϵ)=limr→0I(r)=limr→0∫∂B(xϵ,r)2(−cn(n−2)|x−xϵ|n−∂H∂ν(x,xϵ))(−cn(n−2)(x−xϵ)j|x−xϵ|n−∂H∂xj(x,xϵ))−(−cn(n−2)(x−xϵ)|x−xϵ|n−∇H(x,xϵ))2νjdSx. |
Thanks to the oddness of the integrand, we have
∫∂B(xϵ,r)2(cn(n−2)|x−xϵ|n)(cn(n−2)(x−xϵ)j|x−xϵ|n)−[(cn(n−2)(x−xϵ)|x−xϵ|n)2νj]dSx=0. |
Also, since −ΔxH(x,xε)=0 holds for x∈B(xε,2dε), we may proceed as in (4.2) to get
∫∂B(xϵ,r)2(∂H∂ν(x,xϵ))(∂H∂xj(x,xϵ))−[(∇H(x,xϵ))2νj]dSx=0. |
Using the above estimates, we complete the estimation as follows.
I(2dϵ)=limr→0∫∂B(x,r)2c2n(n−2)∂H∂ν(x,xϵ)(x−xϵ)j|x−xϵ|n+2cn(n−2)|x−xϵ|n−1∂H∂xj(x,xϵ)dSx−2cn(n−2)(x−xϵ)|x−xϵ|n∇H(x,xϵ)νjdSx=[2cn(n−2)n∂H∂xj(xϵ,xϵ)+2cn(n−2)∂H∂xj(xϵ,xϵ)−2cn(n−2)n∂H∂xj(xϵ,xϵ)]|Sn−1|=2cn(n−2)|Sn−1|∂H∂xj(xϵ,xϵ). |
Plugging this into (4.1) shows that
\begin{equation} L_j^{\epsilon} = - \lambda_{\epsilon}^{-\frac{2}{p-1- \varepsilon}} c_n A_U^2 |S^{n-1}| 2(n-2) \frac{\partial H}{\partial x_j} (x_{\epsilon}, \; x_{\epsilon}) + o (d_{\epsilon}^{-(n-1)} \lambda_{\epsilon}^{-(n-2)}). \end{equation} | (4.3) |
Now, we take j \in \{1, \cdots, n\} such that \left| \frac{\partial H}{\partial x_j}(x_{ \varepsilon}, x_{ \varepsilon}) \right| \geq \frac{C}{d_{ \varepsilon}^{n-1}} , which is guaranteed by Theorem 1.3. Injecting this into (4.3) we have
\begin{equation} \begin{split} L_j^{\epsilon} \geq C \lambda_{\epsilon}^{-(n-2)} d_{\epsilon}^{-(n-1)} = C\lambda_{\epsilon} N_{\epsilon}^{-(n-1)}. \end{split} \end{equation} | (4.4) |
Next we shall find an upper bound of R_j^{\epsilon} . Applying (2.14) we have
\begin{equation*} u_{\epsilon} (x) \leq C\lambda_{\epsilon}^{\frac{n}{p+1}} U(\lambda_{\epsilon}(x-x_{\epsilon})) \leq C \lambda_{\epsilon}^{\frac{n}{p+1}} N_{\epsilon}^{-(n-2)}\quad \forall\; x \in \partial B(x_{ \varepsilon}, 2d_{ \varepsilon}). \end{equation*} |
Using this we estimate
\begin{equation} \begin{split} \left| \int_{\partial D_{\epsilon}} u_{\epsilon}^{p+1} \nu_j dS_x \right|& \leq C|\partial D_{\epsilon}| \lambda_{\epsilon}^n N_{\epsilon}^{-(n-2)(p+1)} \\ & \leq C d_{\epsilon}^{(n-1)} \lambda_{\epsilon}^{n} N_{\epsilon}^{-(n-2)(p+1)} \\ & = C \left( \frac{N_{\epsilon}}{\lambda_{\epsilon}}\right)^{(n-1)} \lambda_{\epsilon}^n N_{\epsilon}^{-(n-2)(p+1)} = C\lambda_{\epsilon} N_{\epsilon}^{(n-1)-(n-2)(p+1)}, \end{split} \end{equation} | (4.5) |
which yields
\begin{equation} |R_j^{\epsilon}| \leq C \lambda_{\epsilon} N_{\epsilon}^{(n-1) - (n-2)(p+1)}. \end{equation} | (4.6) |
Now we combine (4.4) and (4.6) to get
\begin{equation*} \begin{split} \lambda_{\epsilon} N_{\epsilon}^{-(n-1)} \leq L_j^{ \varepsilon} = R_j^{ \varepsilon} \leq C \lambda_{\epsilon} N_{\epsilon}^{(n-1) - (n-2)(p+1)}. \end{split} \end{equation*} |
Since N_{\epsilon} goes to infinity as \epsilon \rightarrow 0 , the above inequality yields that
\begin{equation*} -(n-1) \leq (n-1) - (n-2) (p+1), \end{equation*} |
which is equivalent to p \leq \frac{n}{n-2} . However this contradicts to the fact that p = \frac{n+2}{n-2} . Thus the assumption d_{\epsilon} \rightarrow 0 cannot hold, and so \inf_{ \varepsilon > 0} d_{ \varepsilon} > 0 . The proof is completed.
In this section we provide a proof of Theorem 1.2.
Proof of Theorem 1.1. From the result of Theorem 1.1, we know that the maximum point x_{ \varepsilon} of the solution u_{ \varepsilon} are uniformly away from the boundary \partial \Omega . Therefore, up to a subsequence, the point x_{ \varepsilon} converges to an interior point x_0 \in \Omega . By Lemma 2.3 we know the first statement of the theorem holds.
We may easily deduce the version of Lemma 3.1 under the assumption that x_{ \varepsilon} converges to an interior point x_0 . Indeed, it is direct to deduce from (3.4) that
\begin{equation*} u_{ \varepsilon}(x) = A_U \lambda_{ \varepsilon}^{-\frac{[2-(n-2) \varepsilon]}{p-1- \varepsilon}} G(x, x_0) + o (\lambda_{ \varepsilon}^{-\frac{n}{p+1- \varepsilon}}), \end{equation*} |
for x \in \Omega \setminus \{x_0\} . Thus we have
\begin{equation} \lim\limits_{ \varepsilon \rightarrow 0} \lambda_{ \varepsilon}^{\frac{[2-(n-2) \varepsilon]}{p-1- \varepsilon}} u_{ \varepsilon}(x) = A_U G(x, x_0) \quad {\rm{in}}\; C^1 (\Omega \setminus \{x_0\}). \end{equation} | (5.1) |
Proposition 5.1. We have \lim_{ \varepsilon \rightarrow 0} \lambda_{ \varepsilon}^{ \varepsilon} = 1 .
Proof. Let v_{ \varepsilon} = (x-x_0) \cdot \nabla u_{ \varepsilon} + ({2 \over p-1- \varepsilon_n}) u_{ \varepsilon} . Then it satisfies
- \Delta v_{ \varepsilon} = (p- \varepsilon)\, u_{ \varepsilon}^{p-1- \varepsilon} v_{ \varepsilon} \quad \text{in } \Omega. |
Therefore we have
\begin{equation} \lambda_{ \varepsilon}^{\frac{4-2(n-2) \varepsilon}{p-1- \varepsilon}}\int_{ \partial B^n(y, r)} ({ \partial u_{ \varepsilon} \over \partial \nu}v_{ \varepsilon} - { \partial v_{ \varepsilon} \over \partial \nu}u_{ \varepsilon}) dS_x = \lambda_{ \varepsilon}^{\frac{2n}{p+1- \varepsilon}}(p-1- \varepsilon) \int_{B^n(y, r)} u_{ \varepsilon}^{p- \varepsilon} v_{ \varepsilon}\, dx. \end{equation} | (5.2) |
By (5.1) we have
\lim\limits_{ \varepsilon \rightarrow 0} \lambda_{ \varepsilon}^{\frac{2-(n-2) \varepsilon}{p-1- \varepsilon}} v_{ \varepsilon}(x) = A_U \left[(x-x_0) \cdot \nabla G(x, x_0) + \frac{2}{p-1} G(x, x_0)\right]. |
Taking \varepsilon \rightarrow 0 we have
\begin{split} &\lim\limits_{ \varepsilon \rightarrow 0} \lambda_{ \varepsilon}^{\frac{4-2(n-2) \varepsilon}{p-1- \varepsilon}}\int_{ \partial B^n(y, r)} ({ \partial u_{ \varepsilon} \over \partial \nu}v_{ \varepsilon} - { \partial v_{ \varepsilon} \over \partial \nu}u_{ \varepsilon}) dS_x \\ & = A_U^2 \int_{ \partial B^n(y, r)} ({ \partial G(x, x_{0}) \over \partial \nu}\left[(x-x_0) \cdot \nabla G(x, x_0) + \frac{2}{p-1} G(x, x_0)\right] \\ &\quad\quad\quad\quad\quad\qquad - \frac{\partial}{\partial \nu} \left[(x-x_0) \cdot \nabla G(x, x_0) + \frac{2}{p-1} G(x, x_0)\right]G(x, x_0)) dS_x \\ & = A_U^2 (n-2) H(x_0, x_0), \end{split} | (5.3) |
where the last equality is derived in [1, pg. 170]. The right hand side of (5.2) is equal to
\begin{equation*} \begin{split} & \int_{B^n(y, r)} u_{ \varepsilon}^{p- \varepsilon} v_{ \varepsilon}\, dx \\ & = \int_{B^n (x_0, r)} u_{ \varepsilon}^{p- \varepsilon} \left[ (x-x_0) \cdot \nabla u_{ \varepsilon} + \frac{2}{p-1- \varepsilon} u_{ \varepsilon}\right] dx \\ & = \left(\frac{2}{p-1- \varepsilon} - \frac{n}{p+1- \varepsilon}\right) \int_{B (x_0, r)} u_{ \varepsilon}^{p+1- \varepsilon} (x) dx + \int_{\partial B (x_0, r)} u_{ \varepsilon}^{p+1- \varepsilon} (x-x_0) \cdot \nu dS_x \\ & = \left(\frac{2}{p-1- \varepsilon} - \frac{n}{p+1- \varepsilon}\right) \int_{B (x_0, r)} u_{ \varepsilon}^{p+1- \varepsilon} (x) dx + O(\lambda_{ \varepsilon}^{-n}). \end{split} \end{equation*} |
Using this we have
\begin{equation} \begin{split} &\lambda_{ \varepsilon}^{\frac{4-2(n-2) \varepsilon}{p-1- \varepsilon}}(p-1- \varepsilon) \int_{B^n(y, r)} u_{ \varepsilon}^{p- \varepsilon} v_{ \varepsilon}\, dx \\ &\qquad = \left(\lambda_{ \varepsilon}^{\frac{4-2(n-2) \varepsilon}{p-1- \varepsilon}} \varepsilon\right) \frac{(n-2)^2}{2n} \left(\int_{\mathbb{R}^n} U^{p+1} (x) dx + o(1)\right) + O (\lambda_{ \varepsilon}^{-2}). \end{split} \end{equation} | (5.4) |
Injecting (5.3) and (5.4) into (5.2) we get
\begin{equation} A_U^2 q_n H(x_0, x_0) = \lim\limits_{ \varepsilon \rightarrow 0} \left( \lambda_{ \varepsilon}^{\frac{2n}{p+1- \varepsilon}} \varepsilon\right). \end{equation} | (5.5) |
This implies that \lambda_{ \varepsilon} \leq C \varepsilon^{-\frac{2n}{p-1- \varepsilon}} \leq C \varepsilon^{-\frac{4n}{p-1}} for any small \varepsilon > 0 . Therefore we have
\begin{equation*} 1 \leq \lim\limits_{ \varepsilon \rightarrow 0} \lambda_{ \varepsilon}^{ \varepsilon} \leq \lim\limits_{ \varepsilon \rightarrow 0} C^{ \varepsilon} \varepsilon^{-\left(\frac{4n}{p-1}\right) \varepsilon} = 1. \end{equation*} |
The lemma is proved.
Given the result of Proposition 5.1, we deduce from (5.1) that
\begin{equation} \begin{split} \lim\limits_{ \varepsilon \rightarrow 0} \lambda_{ \varepsilon}^{\frac{n}{p+1}} u_{ \varepsilon} (x)& = \lim\limits_{ \varepsilon \rightarrow 0} \lambda_{ \varepsilon}^{\frac{2(n-2)^2 \varepsilon}{2(4- \varepsilon(n-2))}} \lambda_{ \varepsilon}^{\frac{2-(n-2) \varepsilon}{p-1- \varepsilon}} u_{ \varepsilon}(x) \\ & = A_U G(x, x_0)\quad {\rm{in}}\; C^1 (\Omega \setminus \{x_0\}), \end{split} \end{equation} | (5.6) |
where we used p = \frac{n+2}{n-2} in the first equality. This proves the third statement of Theorem 1.2. Next, taking D = B(x_0, r) and u = u_{ \varepsilon} in (2.15), we have
\begin{equation} \lambda_{ \varepsilon}^{n-2} \frac{(n-2)}{n} \int_{\partial B(x_0, r)} u_{ \varepsilon}^{p+1} \nu_j dS_x = \lambda_{ \varepsilon}^{n-2} \int_{\partial B(x_0, r)} |\nabla u_{ \varepsilon}(x)|^2 \nu_j - 2 \frac{\partial u_{ \varepsilon}}{\partial \nu} \frac{\partial u_{ \varepsilon}}{\partial x_j}(x) dS_x. \end{equation} | (5.7) |
By (2.9) we have u_{ \varepsilon} (x) \leq \lambda_{ \varepsilon}^{-\frac{(n-2)}{2}} for x \in \partial B(x_0, r) we have
\begin{equation*} \left|\lambda_{ \varepsilon}^{(n-2)} \int_{\partial B(x_0, r)} u_{ \varepsilon}^{p+1} \nu_j dS_x \right|\leq C\lambda_{ \varepsilon}^{(n-2)} \lambda_{ \varepsilon}^{-n}. \end{equation*} |
Using this and (5.1) we take limit \varepsilon \rightarrow 0 in (5.7) to get
\begin{equation*} 0 = A_U^2 \int_{\partial B(x_0, r)} |\nabla G(x, x_0)|^2 \nu_j - 2 \frac{\partial G(x, x_0)}{\partial \nu} \frac{\partial G(x, x_0)}{\partial x_j} dS_x = -A_U^2 \frac{(2n-1)}{n}\frac{\partial H}{\partial x_j} (x_0, x_0), \end{equation*} |
which yields the second statement of the theorem. Finally, given the result of Proposition 5.1, we get from (5.5) that
\begin{equation*} \lim\limits_{ \varepsilon\rightarrow 0} \left( \varepsilon \cdot\lambda_{ \varepsilon}^{\frac{2n}{p+1}} \right) = (n-2)A_U^2 H(x_0, x_0). \end{equation*} |
This proves the last statement of the theorem. The proof is finished.
We prove the second main theorem of this paper.
Proof of Theorem 1.3. Consider a sequence of points \{x^k\}_{k \in \mathbb{N}} in the domain \Omega such that {\bf d}_k: = {\rm{dist}}(x^k, \partial \Omega) goes to zero as k \rightarrow \infty . Take z^k \in \partial \Omega such that |x^k -z^k| = {\bf d}_k . Let \Omega_k: = \frac{1}{{\bf d}_k} (\Omega -z^k) . Note that we have 0 \in \Omega_k , and also \frac{1}{{\bf d}_k} (x^k - z^k) \in S^{n-1} . Thus we can find a rotation R_k : \mathbb{R}^n \rightarrow \mathbb{R}^n such that
\begin{equation} R_k \left( \frac{1}{{\bf d}_k}(x^k -z^k)\right) = e_n = (0, \cdots, 0, 1). \end{equation} | (6.1) |
Then, by Assumption D. the domain D_k : = R_k \Omega_k converges to an infinite star-shaped domain \mathbb{P} \subsetneq \mathbb{R}^n . To prove the estimate (1.4) we set the function W_{k} : D_{k} \rightarrow \mathbb{R} for each k \in \mathbb{N} by
\begin{equation} W_{k} (z) = {H}(R_k^{-1}{{\bf d}_k}z + z^k, x^k ) {{\bf d}_k^{n-2}}. \end{equation} | (6.2) |
Let G_{k} be Green's function of -\Delta on D_k with the Dirichlet boundary condition. For each y \in \mathbb{R}^{n}_{+} we denote y^* = (y_1, \cdots, y_{n-1}, -y_n) for y = (y_1, \cdots, y_n) \in \mathbb{R}^{n}_{+} . We consider the function {H}_0:\overline{\mathbb{P}} \times \overline{\mathbb{P}} \rightarrow \mathbb{R} satisfying
\begin{equation} \left\{\begin{aligned} -\Delta_z {H}_0 (z, y)& = 0 & \quad {\rm{for}}\; (z, y) \in \mathbb{P} \times \mathbb{P}, \\ H_0 (z, y)& = \frac{c_n}{|(z-y)|^{n-2}}& {\rm{for}}\; z \in \partial \mathbb{P}. \end{aligned}\right. \end{equation} | (6.3) |
Here c_n is the value defined in (2.1). Now we set W_0 : \overline{\mathbb{P}} \rightarrow \mathbb{R} by W_0 (z) : = {H}_0 (z, e_n) . Then we have the following result.
Lemma 6.1. As {{\bf d}_k} \rightarrow 0 , the function W_{k} converges to W_0 in C^1 (B(e_n, 1/4)) .
Proof. By definition (6.2) and (2.2), the function W_{k} satisfies
\begin{equation} -\Delta_w W_{k} (w) = 0 \quad {\rm{in}}\; D_k\quad {\rm{and}}\quad W_k (w) = \frac{c_n}{|R_k^{-1} {\bf d}_k w+ z^k -x^k|^{n-2}} \; {\rm{for}}\; w \in \partial D_k. \end{equation} | (6.4) |
Set the difference R_{k} : \Omega_{k} \rightarrow \mathbb{R} by R_{k} (x) = W_0 (x)- W_{k} (x) for x \in \Omega_{k} . Then, it suffices to show that R_{k} \rightarrow 0 in C_{loc}^1 (\mathbb{P}) . By (6.4) and (6.3) we have
\begin{equation} (-\Delta_w) R_{k} (w) = 0 \quad {\rm{in}}\; D_k. \end{equation} | (6.5) |
Let us prove the C^0 convergence of \mathcal{R}_{k} . Since \mathcal{R}_{k} is harmonic in \Omega_{k} , we only need to show that
\begin{equation*} \lim\limits_{k\rightarrow \infty} \sup\limits_{x \in \partial \Omega_{k}} |\mathcal{R}_{k} (x)| = 0. \end{equation*} |
Take a large number R > 0 . Then we have
\begin{equation*} \sup\limits_{x \in \partial D_k \cap B(0, R)^{c}} |W_k (x)| + |W_0 (x)|\leq \frac{C}{R^{n-2}}. \end{equation*} |
We note that for z\in \partial D_k , using (6.1) we have
\begin{equation} \begin{split} W_k (z)& = \frac{c_n}{\|z-e_n\|^{n-2}}, \end{split} \end{equation} | (6.6) |
and for z \in \partial \mathbb{P} ,
\begin{equation} W_0 (z) = \frac{c_n}{\|z-e_n\|^{n-2}}. \end{equation} | (6.7) |
For fixed R > 0 , we have
\begin{equation*} \lim\limits_{k \rightarrow \infty} (\partial D_k \cap B_R) = \partial \mathbb{P} \cap B_R, \end{equation*} |
and we note that \partial D_k \cap B_R is compact. Combining this fact with (6.6) and (6.7), we obtain
\begin{equation*} \lim\limits_{k \rightarrow \infty} \sup\limits_{x \in \partial D_k \cap B_R} |W_k (x) - W_0 (x)| = 0. \end{equation*} |
Thus,
\begin{equation*} \begin{split} &\lim\limits_{k \rightarrow \infty} \sup\limits_{x \in \partial D_k}|W_k (x) - W_0 (x)| \\ &\qquad\leq \lim\limits_{k \rightarrow \infty} \sup\limits_{x \in \partial D_k \cap B_R}|W_k (x) - W_0 (x)| + \lim\limits_{k \rightarrow \infty} \sup\limits_{x \in \partial D_k \cap B_R^{c}}|W_k (x) - W_0 (x)| \\ &\qquad \leq \frac{C}{R^{n-2}}. \end{split} \end{equation*} |
Since R > 0 is arbitrary, we have
\lim\limits_{k \rightarrow \infty} \sup\limits_{x \in \partial D_k}|W_k (x) - W_0 (x)| = 0. |
Combining the above two convergence results, we can deduce that R_{k} (x) \rightarrow 0 uniformly for x \in B(e_n, 1/4) . From (6.4) we know that R_{k} is contained in C^{1, \beta} (B(e_n, 1/4)) uniformly in k \in \mathbb{N} for some \beta > 0 . Thus R_{k} converges to a function f in C^1 (B(e_n, 1/4)) . In this paper we are concerned with the Lane-Emden-Fowler equation
\begin{equation} \left\{\begin{array}{rll}-\Delta u & = u^{\frac{n+2}{n-2}- \varepsilon}& {\rm{in}}\; \Omega, \\ u& > 0& {\rm{in}}\; \Omega, \\ u& = 0& {\rm{on}}\; \partial \Omega, \end{array} \right. \end{equation} | (6.8) |
where \Omega \subset \mathbb{R}^n ( n \geq 3 ) is a polygonal domain and \varepsilon > 0 . We study the asymptotic behavior of minimal energy solutions as \varepsilon > 0 goes to zero. we have f\equiv 0 since R_{k} converges to 0 in C^0 (B(e_n, 1/2)) . The lemma is proved.
Lemma 6.2. We have \frac{\partial}{\partial x_n} W_0 (e_n) \neq 0 .
Proof. Notice that H_0 (x, y) satisfies
\begin{equation*} \left\{ \begin{array}{ll}-\Delta_x H_0 (x, y) = 0&\quad x \in \mathbb{P}, \\ H_0 (x, y) = \frac{c_n}{|x-y|^{n-2}}&\quad x \in \partial \mathbb{P}. \end{array} \right. \end{equation*} |
Since H_0 is the regular part of Green's function on \mathbb{P} , we have
\begin{equation} H_0 (x, y) = H_0 (y, x). \end{equation} | (6.9) |
For given t > 0 consider the function f (x): = t^{n-2} H_0 (tx, te_n) defined on \frac{1}{t} \mathbb{P} = \mathbb{P} which satisfies
\begin{equation*} \left\{ \begin{array}{ll}-\Delta_x f(x) = 0&\quad x \in \mathbb{P}, \\ f(x) = \frac{c_n t^{n-2}}{|tx-te_n|^{n-2}} = \frac{c_n}{|x-e_n|^{n-2}}&\quad x \in \partial \mathbb{P}. \end{array} \right. \end{equation*} |
This exaclty means that f (x) = H_0 (x, e_n) , and so H_0 (x, e_n) = t^{n-2} H_0 (tx, t e_n) . Combining this with the symmetric property (6.9), we have
\begin{equation} \begin{split} \left.\frac{\partial}{\partial x_n}W_0 (x) \right|_{x = e_n}& = \left(\frac{\partial}{\partial x_n} H_0 (x, e_n) \right)_{x = e_n} \\ & = \frac{1}{2}\left( \frac{\partial}{\partial x_n} H_0 (x, x)\right)_{x = e_n} \\ & = \frac{1}{2}\left.\frac{\partial}{\partial t} H_0 (te_n, te_n)\right|_{t = 1} = \frac{(2-n)}{2} H(e_n, e_n). \end{split} \end{equation} | (6.10) |
Also we note that H_0 (e_n, e_n) \neq 0 by the maximum principle since (-\Delta)H_0 = 0 in \mathbb{P} and H_0 > 0 on \partial \mathbb{P} . Combining this fact with (6.10) we deduce that \frac{\partial}{\partial x_n} W_{0}(e_n) < 0 . The proof is finished.
Now we are ready to finish the proof of Theorem 1.3. By Lemma 6.1, we know that W_k (x) converges to W_0 (x) in C^1 (B(e_n, 1/4)) . Since \left|\frac{\partial}{\partial x_n} W_0 (e_n)\right| > c > 0 , we conclude that for large k \in \mathbb{N} , we have \left|\frac{\partial}{\partial x_n}W_k (e_n)\right| > c/2 . By definition of W_k given in (6.2), we have
\begin{equation*} \frac{\partial}{\partial x_n} W_k (z) = {\bf d}_k^{n-1}(R_k^{-1})_n \cdot \nabla H ({\bf d}_k R_k^{-1}(z), x^k). \end{equation*} |
Therefore we may conclude that for large k \in \mathbb{N} ,
\begin{equation*} |{\bf d}_k^{n-1} (R_k^{-1})_n \cdot \nabla H (x^k, x^k)| > c/2, \end{equation*} |
which implies that
\begin{equation*} \left|\nabla H (x^k, x^k)\right| > \frac{c}{2{\bf d}_k^{n-1}} \end{equation*} |
for k \in \mathbb{N} large enough. The proof is finished.
In this paper, we study the energy minimizing solutions to slightly subcritical elliptic problems on nonconvex polygonal domains. The main part for the analysis is to exclude the possibility that the peak of the solution approaches the boundary of the domain as the moving plane method is difficult to apply directly for the nonconvex polygonal domain. To address this challenge, we make use of the Pohozaev identity and the Green's function to show that a contradiction aries when we assume that the solution blows up near the boundary.
The authors declare they have not used artificial intelligence (AI) tools in the creation of this article.
This work was supported by the National Research Foundation of Korea (grant NRF-2021R1F1A1059671).
The authors declare no conflict of interest.
[1] |
Z. C. Han, Asymptotic approach to singular solutions for nonlinear elliptic equations involving critical Sobolev exponent, Ann. Inst. H. Poincaré Anal. Non Linéaire, 8 (1991), 159–174. https://doi.org/10.1016/S0294-1449(16)30270-0 doi: 10.1016/S0294-1449(16)30270-0
![]() |
[2] |
O. Rey, The role of the Green's function in a nonlinear elliptic equation involving the critical Sobolev exponent, J. Funct. Anal., 89 (1990), 1–52. https://doi.org/10.1016/0022-1236(90)90002-3 doi: 10.1016/0022-1236(90)90002-3
![]() |
[3] |
F. V. Atkinson, L. A. Peletier, Elliptic equations with nearly critical growth, J. Differ. Equ., 70 (1987), 349–365. https://doi.org/10.1016/0022-0396(87)90156-2 doi: 10.1016/0022-0396(87)90156-2
![]() |
[4] | H. Brezis, L. A. Peletier, Asymptotics for elliptic equations involving critical growth, In: Partial differential equations and the calculus of variations, Boston: Birkhäuser Boston, 1989. https://doi.org/10.1007/978-1-4615-9828-2_7 |
[5] |
B. Aharrouch, A. Aberqi, J. Bennouna, Existence and regularity of solutions to unilateral nonlinear elliptic equation in Marcinkiewicz space with variable exponent, Filomat, 37 (2023), 5785–5797. https://doi.org/10.2298/FIL2317785A doi: 10.2298/FIL2317785A
![]() |
[6] |
T. Bartsch, Q. Guo, Nodal blow-up solutions to slightly subcritical elliptic problems with Hardy-critical term, Adv. Nonlinear Stud., 17 (2017), 55–85. https://doi.org/10.1515/ans-2016-6008 doi: 10.1515/ans-2016-6008
![]() |
[7] |
G. Cora, A. Iacopetti, On the structure of the nodal set and asymptotics of least energy sign-changing radial solutions of the fractional Brezis-Nirenberg problem, Nonlinear Anal., 176 (2018), 226–271. https://doi.org/10.1016/j.na.2018.07.001 doi: 10.1016/j.na.2018.07.001
![]() |
[8] |
Y. Dammak, R. Ghoudi, Sign-changing tower of bubbles to an elliptic subcritical equation, Commun. Contemp. Math., 21 (2019), 1850052. https://doi.org/10.1142/S0219199718500529 doi: 10.1142/S0219199718500529
![]() |
[9] |
Q. Guo, Blowup analysis for integral equations on bounded domains, J. Differ. Equ., 266 (2019), 8258–8280. https://doi.org/10.1016/j.jde.2018.12.028 doi: 10.1016/j.jde.2018.12.028
![]() |
[10] |
W. Ma, Z. Zhao, B. Yan, Global existence and blow-up of solutions to a parabolic nonlocal equation arising in a theory of thermal explosion, J. Funct. Spaces, 2022 (2022), 4629799. https://doi.org/10.1155/2022/4629799 doi: 10.1155/2022/4629799
![]() |
[11] | M. Musso, A. Pistoia, Multispike solutions for a nonlinear elliptic problem involving the critical Sobolev exponent, Indiana Univ. Math. J., 51 (2002), 541–579. |
[12] |
M. Musso, A. Pistoia, Tower of bubbles for almost critical problems in general domains, J. Math. Pures Appl., 93 (2010), 1–40. https://doi.org/10.1016/j.matpur.2009.08.001 doi: 10.1016/j.matpur.2009.08.001
![]() |
[13] |
M. Ragusa, Local Hölder regularity for solutions of elliptic systems, Duke Math. J., 113 (2002), 385–397. https://doi.org/10.1215/S0012-7094-02-11327-1 doi: 10.1215/S0012-7094-02-11327-1
![]() |
[14] |
F. Takahashi, Asymptotic behavior of least energy solutions for a biharmonic problem with nearly critical growth, Asymptot. Anal., 60 (2008), 213–226. https://doi.org/10.3233/ASY-2008-0904 doi: 10.3233/ASY-2008-0904
![]() |
[15] |
D. Salazar, Sign changing bubbling solutions for a critical Neumann problem in 3D, Nonlinear Anal., 188 (2019), 500–539. https://doi.org/10.1016/j.na.2019.06.018 doi: 10.1016/j.na.2019.06.018
![]() |
[16] |
S. Santra, Existence and shape of the least energy solution of a fractional Laplacian, Calc. Var. Partial Differential Equations, 58 (2019), 48. https://doi.org/10.1007/s00526-019-1494-3 doi: 10.1007/s00526-019-1494-3
![]() |
[17] |
A. Pistoia, O. Rey, Boundary blow-up for a Brezis-Peletier problem on a singular domain, Calc. Var. Partial Differential Equations, 18 (2003), 243–251. https://doi.org/10.1007/s00526-003-0197-x doi: 10.1007/s00526-003-0197-x
![]() |
[18] |
M. Flucher, A. Garroni, S. Müller, Concentration of low energy extremals: Identification of concentration points, Calc. Var. Partial Differential Equations, 14 (2002), 483–516. https://doi.org/10.1016/S0294-1449(99)80015-8 doi: 10.1016/S0294-1449(99)80015-8
![]() |
[19] |
O. Rey, Blow-up points of solutions to elliptic equations with limiting nonlinearity, Differ. Integral Equ., 4 (1991), 1155–1167. https://doi.org/10.57262/die/1371154279 doi: 10.57262/die/1371154279
![]() |