For any R>0, infinitely many nonradial singular solutions can be constructed for the following equation:
−Δu=euinBR∖{0},(0.1)
where BR={x∈RN(N≥3):|x|<R}. To construct nonradial singular solutions, we need to consider asymptotic expansion at the isolated singular point x=0 of a prescribed solution of (0.1). Then, nonradial singular solutions of (0.1) can be constructed by using the asymptotic expansion and introducing suitable weighted Hölder spaces.
Citation: Jingyue Cao, Yunkang Shao, Fangshu Wan, Jiaqi Wang, Yifei Zhu. Nonradial singular solutions for elliptic equations with exponential nonlinearity[J]. Electronic Research Archive, 2024, 32(5): 3171-3201. doi: 10.3934/era.2024146
[1] | Zongming Guo, Fangshu Wan . Weighted fourth order elliptic problems in the unit ball. Electronic Research Archive, 2021, 29(6): 3775-3803. doi: 10.3934/era.2021061 |
[2] | Qihong Shi, Yaqian Jia, Xunyang Wang . Global solution in a weak energy class for Klein-Gordon-Schrödinger system. Electronic Research Archive, 2022, 30(2): 633-643. doi: 10.3934/era.2022033 |
[3] | Shuguan Ji, Yanshuo Li . Quasi-periodic solutions for the incompressible Navier-Stokes equations with nonlocal diffusion. Electronic Research Archive, 2023, 31(12): 7182-7194. doi: 10.3934/era.2023363 |
[4] | Abdeljabbar Ghanmi, Hadeel Z. Alzumi, Noureddine Zeddini . A sub-super solution method to continuous weak solutions for a semilinear elliptic boundary value problems on bounded and unbounded domains. Electronic Research Archive, 2024, 32(6): 3742-3757. doi: 10.3934/era.2024170 |
[5] | Xiao Su, Hongwei Zhang . On the global existence and blow-up for the double dispersion equation with exponential term. Electronic Research Archive, 2023, 31(1): 467-491. doi: 10.3934/era.2023023 |
[6] | Mufit San, Seyma Ramazan . A study for a higher order Riemann-Liouville fractional differential equation with weakly singularity. Electronic Research Archive, 2024, 32(5): 3092-3112. doi: 10.3934/era.2024141 |
[7] | Chang Hou, Hu Chen . Stability and pointwise-in-time convergence analysis of a finite difference scheme for a 2D nonlinear multi-term subdiffusion equation. Electronic Research Archive, 2025, 33(3): 1476-1489. doi: 10.3934/era.2025069 |
[8] | Shuting Chang, Yaojun Ye . Upper and lower bounds for the blow-up time of a fourth-order parabolic equation with exponential nonlinearity. Electronic Research Archive, 2024, 32(11): 6225-6234. doi: 10.3934/era.2024289 |
[9] | Gezi Chong, Jianxia He . On exponential decay properties of solutions of the (3 + 1)-dimensional modified Zakharov-Kuznetsov equation. Electronic Research Archive, 2025, 33(1): 447-470. doi: 10.3934/era.2025022 |
[10] | Hsueh-Chen Lee, Hyesuk Lee . An a posteriori error estimator based on least-squares finite element solutions for viscoelastic fluid flows. Electronic Research Archive, 2021, 29(4): 2755-2770. doi: 10.3934/era.2021012 |
For any R>0, infinitely many nonradial singular solutions can be constructed for the following equation:
−Δu=euinBR∖{0},(0.1)
where BR={x∈RN(N≥3):|x|<R}. To construct nonradial singular solutions, we need to consider asymptotic expansion at the isolated singular point x=0 of a prescribed solution of (0.1). Then, nonradial singular solutions of (0.1) can be constructed by using the asymptotic expansion and introducing suitable weighted Hölder spaces.
We are interested in singular solutions of the following equation with exponential nonlinearity:
Δu+eu=0inBR∖{0}, | (1.1) |
where R>0 and BR={x∈RN(N≥3):|x|<R} is a ball.
By a singular solution of (1.1) we mean that u∈C2(BR∖{0}) and 0 is a nonremovable singular point of u.
It is easily known that (1.1) admits a (trivial) radial singular solution:
Us(x)=Us(|x|):=−2ln|x|+ln[2(N−2)]. | (1.2) |
We are mainly concerned with nonradial singular solutions of (1.1) in this paper.
When N=2, by using the moving plane method, the authors of [1] proved that every solution of
{Δu+eu=0inR2,∫R2eudx<∞ | (1.3) |
has the form
u(x)=ln32λ2(4+λ2|x−x0|2)2,λ>0,x0∈R2. |
For n>0, symmetry and uniqueness results were obtained in [2] for the solutions of the following problem:
{Δu+|x|2(n−1)eu=0inR2,∫R2|x|2(n−1)eudx<∞. | (1.4) |
If n=1, problem (1.4) reduces to (1.3); also, classification of solutions of (1.4) can be found in [1]. Under the condition that n≥2 is an integer, the authors of [2] showed that problem (1.4) admits radial and nonradial solutions, but, when n>0 is not an integer, problem (1.4) only has radial solutions. Note that, for each n>0, if u(x) is a solution of (1.4), we can perform the following transformation:
v(x)=2(n−1)ln|x|+u(x), |
we see that v(x) satisfies the following equation
{Δv+ev=0inR2∖{0},∫R2∖{0}evdx<∞. | (1.5) |
The results in [2] imply that (1.5) admits a family of radial and nonradial singular solutions.
The asymptotic behavior of singular solutions of the problem given by
{Δu+eu=0inD1∖{0},∫D1∖{0}eudx<∞ | (1.6) |
where D1⊂R2 is the unit disc, was studied in [3]. The authors of [3] obtained that if u∈C2(D1∖{0}) is a singular solution of (1.6), then there is α>−2 such that
u(x)=αln|x|+O(1)as|x|→0. |
In a recent paper [4], the authors continued the study in [3] and obtained asymptotic expansions up to arbitrary orders for u(x) as |x|→0.
Under the condition that N≥2, the structure of finite Morse index solutions of the equation
Δu+eu=0inRN | (1.7) |
was studied in [5,6,7]. In particular, under the condition that N=3, the asymptotic behavior at x=0 of solutions u with |x|2eu∈L∞(R3) of the equation
Δu+eu=0inR3∖{0}, | (1.8) |
was classified in [8]. For the case that N=3, if we write
u(x)=−2ln|x|+Θ(θ) |
where (|x|,θ)∈(0,∞)×S2 denotes the spherical coordinates in R3∖{0}, we find that Θ(θ) must satisfy
ΔS2Θ−2+eΘ=0 | (1.9) |
on S2 where ΔS2 is the Laplace-Beltrami operator on (S2,g0) and g0 is the standard round metric. It means that the Gaussian curvature of the metric g=eΘg0 on S2 is 12. This and related equations have been studied for more than three decades. Chang and Yang[9] and Onofri[10] described all regular solutions of (1.9). Specifically, axially symmetric solutions of (1.9) can be written explicitly as Θ(θ)=log2−2log(√c2+1−ccosθ), where c∈R is constant and θ∈[0,π] is the geodesic distance from the north pole of S2. Hence,
u(x)=−2ln|x|+log2−2log(√c2+1−ccosθ) |
is a one-parameter family of non-radial singular solutions of (1.8).
Recently, singular solutions in different settings have also been studied in [11] and [12]. The authors of [12] obtained the existence and asymptotic behavior of singular solutions to quasilinear elliptic inequalities with nonlocal terms. Moreover, by using mini-max and asymptotic approximation methods, the existence of positive singular solutions to the planar logarithmic Choquard equation with exponential nonlinearity was established in [11].
In this paper, we study singular solutions of (1.1) in BR⊂RN(N≥3). We are interested in not only the asymptotic behavior of singular solutions of (1.1) at x=0, but also the existence of nonradial singular solutions of (1.1). The structure of nonradial singular solutions of the equation
Δu+eu=0inRN∖{0}withN≥3 | (1.10) |
remains largely open. Motivated by the main ideas in [13], the authors of [14,15] obtained infinitely many nonradial singular solutions of (1.10) of the following form given 4≤N≤10:
u(x)=−2ln|x|+Θ(θ), | (1.11) |
where Θ(θ) is a non-constant solution of the equation
ΔSN−1Θ−2(N−2)+eΘ=0 | (1.12) |
on SN−1, where ΔSN−1 is the Laplace-Beltrami operator on (SN−1,g0). They constructed infinitely many axially symmetric non-constant classical solutions of (1.12). The only singular solutions to (1.10) known so far are the (trivial) radial singular solution Us(x) and the solutions given in (1.11). It is clear that they are also the singular solutions to (1.1).
We will construct a new type of singular solutions of (1.1) in the following form:
u(x)−Us(x)=O(|x|ϵ)as |x|→0, | (1.13) |
for some ϵ>0.
Our main result is as follows.
Theorem 1.1. For any R>0, Eq (1.1) admits infinitely many nonradial singular solutions u(x) of the following form:
u(x)=Us(x)+O(|x|σ(2)+)as |x|→0, | (1.14) |
where
σ(2)+=−12(N−2)+12√N2−4N+20>0. |
It is known from Theorem 1.1 that the parameter ϵ in (1.13) is σ(2)+.
To prove Theorem 1.1, we firstly study the detailed asymptotic behavior at x=0 of a prescribed singular solution u of
Δu+eu=0in B∖{0} | (1.15) |
with the form (1.13), where B:=B1={x∈RN(N≥3):|x|<1} is the unit ball. Then, the infinitely many nonradial singular solutions of the form (1.14) can be constructed by using the asymptotic expansion and introducing suitable Hölder spaces.
This paper is organized as follows. In Section 2, we obtain asymptotic expansions near the isolated singular point x=0 of solutions of (1.15). In Section 3, we establish weighted Hölder spaces and invertible operators related to our Equation (1.15). In Section 4, we construct infinitely many singular solutions of (1.1) and show that the singular solutions that we have constructed are non-radial singular solutions.
We will establish asymptotic expansions of the singular solution u∈C2(B∖{0}) of (1.15) such that (1.13) is satisfied.
Let v(x)=u(x)−Us(x). Then v(x) satisfies
−Δv=2(N−2)|x|−2(ev−1)in B∖{0}. | (2.1) |
Making the following transformations:
t=lnr,w(t,θ)=v(r,θ), |
we see from (2.1) that w(t,θ) satisfies
wtt+(N−2)wt+ΔSN−1w+2(N−2)(ew−1)=0in (−∞,0)×SN−1. | (2.2) |
We write (2.2) in the following forms
wtt+(N−2)wt+ΔSN−1w+2(N−2)w+2(N−2)(ew−1−w)=0in (−∞,0)×SN−1 | (2.3) |
and
Lw+F(w)=0in (−∞,0)×SN−1, | (2.4) |
where
Lw=wtt+(N−2)wt+ΔSN−1w+2(N−2)w,F(w)=2(N−2)(ew−1−w). |
Moreover, (1.13) implies that
w(t,θ)=O(eϵt)uniformly for θ∈SN−1ast→−∞. | (2.5) |
Define a linearized operator
L=∂2∂t2+(N−2)∂∂t+ΔSN−1+2(N−2). | (2.6) |
Obviously, L can decouple into infinitely many ordinary differential operators, i.e.,
Lk=d2dt2+(N−2)ddt−λk+2(N−2) | (2.7) |
for k=0,1,2,…, where λk is the k-th eigenvalue of
−ΔSN−1Q=λQ | (2.8) |
and λk=k(N−2+k) with the following multiplicity:
mk=(N−2+2k)(N−3+k)!k!(N−2)!. |
The {Qk1(θ),…,Qkmk(θ)} with ‖Qkj‖L2(SN−1)=1 denotes the basis of the eigenspace Hk(SN−1)⊂L2(SN−1) corresponding to λk. Then two roots of characteristic polynomial of (2.7) are as follows:
σ(k)±=−12(N−2)±12√(N−2)(N−10)+4k(N−2+k). | (2.9) |
For k=0, we have from (2.9) that
σ(0)±={−12(N−2)±i2√(N−2)(10−N),for 3≤N≤9,−12(N−2)<0,for N=10,−12(N−2)±12√(N−2)(N−10)<0,for N≥11. | (2.10) |
For k=1,
σ(1)±=−(N−2)2±|N−4|2. | (2.11) |
Then,
σ(1)+={0,forN=3,−(N−2)2+|N−4|2<0,for N≥4 | (2.12) |
and
σ(1)−=−(N−2)2−|N−4|2<0,for N≥3. | (2.13) |
For k≥2, the fact that k(N−2+k)>2(N−2) implies that
(N−2)(N−10)+4k(N−2+k)>(N−2)2, |
we see from (2.9) that
σ(k)+>0,σ(k)−<0. | (2.14) |
It is clear that
σ(k+1)+>σ(k)+>0,σ(k+1)−<σ(k)−<0for any k≥2. |
Proposition 2.1. Assume that N≥3 and u=u(r) is a radial solution of (1.15) that satisfies
u(r)−Us(r)=O(rϵ)for r near 0 and some ϵ>0. |
Then
u(r)≡Us(r)for r∈(0,1]. |
Proof. Bt applying the following transformations:
v(r)=u(r)−Us(r),w(t)=v(r),t=logr, |
by (2.3), w(t) satisfies the following ordinary diifferential equation (ODE):
wtt+(N−2)wt+2(N−2)w+f(w)=0in (−∞,0), | (2.15) |
where
f(w)=2(N−2)(ew−1−w)=O(w2)=O(e2ϵt). |
Note that w(t)=O(eϵt) for t near −∞. Therefore, for 3≤N≤9,
w(t)=A1eτtcosγt+A2eτtsinγt−B1eτtcosγt∫t−∞e−τs[−f(w(s))]sinγsds−B2eτtsinγt∫t−∞e−τs[−f(w(s))]cosγsds, | (2.16) |
where |B1|=|B2|=1/γ, |f(w(t))|=O(e2ϵt),
τ=−12(N−2),γ=12√(N−2)(10−N). |
Since w(t)→0 as t→−∞, we obtain from (2.16) that A1=A2=0 and
w(t)=B1eτtcosγt∫t−∞e−τsO(w2(s))sinγsds+B2eτtsinγt∫t−∞e−τsO(w2(s))cosγsds, | (2.17) |
and
|w(t)|≤Ce2ϵt:=eβ+2ϵtfor some fixed β>0withC=eβandtnear−∞. | (2.18) |
Substituting (2.18) into (2.17), we see that
|w(t)|≤e3β+4ϵtfor tnear−∞. | (2.19) |
We can do the same process to obtain that w(t)≡0 for t≤−4βϵ. Since w(t) satisfies the ODE in (2.15), then w(t)≡0 for t∈(−∞,0).
For N=10, we see that
w(t)=A1eτt+A2teτt−B1eτt∫t−∞se−τs[−f(w(s))]ds−B2teτt∫t−∞e−τs[−f(w(s))]ds, | (2.20) |
where |B1|=|B2|=1. Note that τ=−(N−2)2<0. Since w(t)→0 as t→−∞, we see that A1=A2=0. Arguments similar to those in the proof for the case of 3≤N≤9 imply that w(t)≡0 for t∈(−∞,0).
For N≥11, we see that
w(t)=A1eσ(0)+t+A2eσ(0)−t−B1eσ(0)+t∫t−∞e−σ(0)+s[−f(w(s))]ds−B2eσ(0)−t∫t−∞e−σ(0)−s[−f(w(s))]ds, | (2.21) |
where |B1|=|B2|=|1σ(0)−−σ(0)+|. Note that σ(0)+<0 and σ(0)−<0. Since w(t)→0 as t→−∞, we see that A1=A2=0. Arguments similar to those in the proof for the case of 3≤N≤9 imply that w(t)≡0 for t∈(−∞,0). This completes the proof of this proposition.
Lemma 2.2. Assume that N≥3 and u∈C2(B∖{0}) is a singular solution of (1.15) that satisfies (1.13). Defining w(t,θ)=u(x)−Us(x) and t=lnr, it follows that w(t,θ)=O(eϵt) for t∈(−∞,−1], and
maxSN−1|w(t,θ)|≤Ceσ(2)+tfor t∈(−∞,−1]. | (2.22) |
Proof. Let
w(t,θ)=∞∑k=0mk∑j=1wkj(t)Qkj(θ). |
Then wkj(t) satisfies the following equation:
(wkj)″(t)+(N−2)(wkj)′(t)+[2(N−2)−λk]wkj(t)=−gkj(t), | (2.23) |
where
gkj(t)=∫SN−1F(w(t,θ))Qkj(θ)dθ. |
Note that
‖w‖2L2(SN−1)=∞∑k=0mk∑j=1(wkj(t))2,‖F(w)‖2L2(SN−1)=∞∑k=0mk∑j=1(gkj(t))2. |
Since F(w)=O(w2) and w(t,θ)=O(eϵt), we see that
‖F(w)‖L2(SN−1)=O(eϵt‖w‖L2(SN−1)). | (2.24) |
On the other hand, it follows from (2.23) that for k≥2, T≪−1 and t<T,
wkj(t)=Akj,1eσ(k)+t+Akj,2eσ(k)−t+Bkj,1∫Tteσ(k)+(t−s)[−gkj(s)]ds−Bkj,2∫t−∞eσ(k)−(t−s)[−gkj(s)]ds, |
where
|Bkj,1|=|Bkj,2|=|1σ(k)−−σ(k)+|. |
Since wkj(t)→0 as t→−∞, we have that Akj,2=0. Moreover,
wkj(T)=Akj,1eσ(k)+T−Bkj,2∫T−∞eσ(k)−(T−s)[−gkj(s)]ds |
and
Akj,1=O(e−σ(k)+T). |
Then,
wkj(t)=O(eσ(k)+(t−T))+Bkj,1∫Tteσ(k)+(t−s)[−gkj(s)]ds−Bkj,2∫t−∞eσ(k)−(t−s)[−gkj(s)]ds | (2.25) |
and for small enough δ>0,
[wkj(t)]2≤O(e2σ(k)+(t−T))+4(Bkj,1)2(∫Tteδ(t−s)ds)(∫Tte(2σ(k)+−δ)(t−s)(gkj(s))2ds)+4(Bkj,2)2(∫t−∞e−δ(t−s)ds)(∫t−∞e(2σ(k)−+δ)(t−s)(gkj(s))2ds)≤Ce2σ(k)+(t−T)+Cδ∫Tte(2σ(2)+−δ)(t−s)(gkj(s))2ds+Cδ∫t−∞e(2σ(2)−+δ)(t−s)(gkj(s))2ds, |
with constants C>0 and Cδ>0 being dependent on δ and independent of (j,k).
For k=1, T≪−1 and t<T,
w1j(t)=A1j,1eσ(1)+t+A1j,2eσ(1)−t−B1j,1∫t−∞eσ(1)+(t−s)[−g1j(s)]ds−B1j,2∫t−∞eσ(1)−(t−s)[−g1j(s)]ds, |
where
|B1j,1|=|B1j,2|=|1σ(1)−−σ(1)+|. |
Note that σ(1)+≤0 and σ(1)−<0. Since w1j(t)→0 as t→−∞, we have that A1j,1=A1j,2=0,
w1j(t)=−B1j,1∫t−∞eσ(1)+(t−s)[−g1j(s)]ds−B1j,2∫t−∞eσ(1)−(t−s)[−g1j(s)]ds | (2.26) |
and
(w1j(t))2=O(e4ϵt). | (2.27) |
Note that
(g1j(t))2≤C‖F(w)‖2L2(SN−1)≤Ce4ϵt. |
Similarly, we have
(g01(t))2=O(e4ϵt),(w01(t))2=O(e4ϵt). | (2.28) |
Then,
∞∑k=2mk∑j=1(wkj(t))2≤C∞∑k=2mk∑j=1e2σ(k)+(t−T)+Cδ∫Tte(2σ(2)+−δ)(t−s)∞∑k=2mk∑j=1(gkj(s))2ds+Cδ∫t−∞e(2σ(2)−+δ)(t−s)∞∑k=2mk∑j=1(gkj(s))2ds≤C∞∑k=2mk∑j=1e2σ(k)+(t−T)+C∫Tte(2σ(2)+−δ)(t−s)e4ϵsds+C∫Tte(2σ(2)+−δ)(t−s)e2ϵs∞∑k=2mk∑j=1(wkj(s))2ds+C∫t−∞e(2σ(2)−+δ)(t−s)e4ϵsds+C∫t−∞e(2σ(2)−+δ)(t−s)e2ϵs∞∑k=2mk∑j=1(wkj(s))2ds. |
Notice that
∞∑k=2mk∑j=1(gkj(t))2=‖F(w)‖2L2(SN−1)−[(g01(t))2+N∑j=1(g1j(t))2] |
and
‖F(w)‖2L2(SN−1)=O(e2ϵt)‖w‖2L2(SN−1)=O(e2ϵt)[∞∑k=2mk∑j=1(wkj(t))2+((w01(t))2+N∑j=1(w1j(t))2)], |
since F(w)=O(w2). We also know that
∞∑k=2mk∑j=1e2σ(k)+(t−T)=∞∑k=2mke2σ(k)+(t−T)=O(e2σ(2)+(t−T)), |
since
limk→∞mk+1e2(σ(k+1)+−σ(2)+)(t−T)mke2(σ(k)+−σ(2)+)(t−T)=limk→∞[mk+1mke2(σ(k+1)+−σ(k)+)(t−T)]=e2(t−T)<12. |
Let [W(t)]2=∑∞k=2∑mkj=1(wkj(t))2. We have that, if 4ϵ≠2σ(2)+−δ,
[W(t)]2≤Ce2σ(2)+(t−T)+Ce4ϵt+C∫Tte(2σ(2)+−δ)(t−s)e4ϵsds+C∫Tte(2σ(2)+−δ)(t−s)e2ϵs[W(s)]2ds+C∫t−∞e(2σ(2)−+δ)(t−s)e2ϵs[W(s)]2ds. |
Now we show that, for t<T≪−1,
‖w‖L2(SN−1)=[∞∑k=0mk∑j=1(wkj(t))2]12=O(eσ(2)+t). | (2.29) |
In fact, two cases can occur: (ⅰ) 4ϵ≥[2σ(2)+−δ] and (ⅱ) 4ϵ<[2σ(2)+−δ].
It suffices to consider that 4ϵ>[2σ(2)+−δ] in the case (ⅰ). So
[W(t)]2≤Ce(2σ(2)+−δ)(t−T)+C∫Tte(2σ(2)+−δ)(t−s)e2ϵs[W(s)]2ds+C∫t−∞e(2σ(2)−+δ)(t−s)e2ϵs[W(s)]2ds. | (2.30) |
Set
K1(t)=∫Tte(2σ(2)+−δ)(t−s)[W(s)]2ds,K2(t)=∫t−∞e(2σ(2)−+δ)(t−s)[W(s)]2ds. |
Then, if |T| is large,
(K2−K1)′(t)=(2σ(2)−+δ)K2(t)−(2σ(2)+−δ)K1(t)+2[W(t)]2≤(2σ(2)−+δ)K2(t)−(2σ(2)+−δ)K1(t)+Ce2ϵT(K1(t)+K2(t))+Ce(2σ(2)+−δ)(t−T)≤Ce(2σ(2)+−δ)(t−T) |
where σ(2)−<0, σ(2)+>0. Since K1(t)→0 and K2(t)→0 as t→−∞, then for t<T,
K2(t)≤K1(t)+Ce(2σ(2)+−δ)t. | (2.31) |
Substituting (2.31) into (2.30), we have
[W(t)]2≤Ce(2σ(2)+−δ)t+Ce2ϵT∫Tte(2σ(2)+−δ)(t−s)[W(s)]2ds. | (2.32) |
It follows from arguments similar to those in [16,17] that, for t<T (enlarge |T| if necessary),
[W(t)]2≤CϵTe(2σ(2)+−δ−ϵT)t, | (2.33) |
where ϵT=Ce2ϵT (i.e., C is independent of ϵ). On the other hand, (2.27) and (2.28) imply that
(w1j(t))2=O(e4ϵt)=O(e(2σ(2)+−δ)t),j=1,2,…,m1, |
(w01(t))2=O(e4ϵt)=O(e(2σ(2)+−δ)t) |
where 4ϵ>2σ(2)+−δ. Therefore, for t<T (i.e., T is sufficiently negative),
∞∑k=0mk∑j=1(wkj(t))2=O(e(2σ(2)+−δ−ϵT)t) | (2.34) |
and
‖w‖L2(SN−1)=O(e(σ(2)+−δ2−ϵT2)t). | (2.35) |
Using (2.35), we obtain
‖F(w)‖2L2(SN−1)=O(e2ϵt)‖w‖2L2(SN−1)=O(e(2σ(2)+−δ−ϵT+2ϵ)t). | (2.36) |
Choosing δ sufficiently small such that δ<2ϵ−ϵT, then
m2∑j=1|w2j(t)|≤Ceσ(2)+(t−T)+C∫Tteσ(2)+(t−s)m2∑j=1|g2j(s)|ds+C∫t−∞eσ(2)−(t−s)m2∑j=1|g2j(s)|ds≤Ceσ(2)+t+C∫Tteσ(2)+(t−s)e(σ(2)+−δ2−ϵT2+ϵ)sds+C∫t−∞eσ(2)−(t−s)e(σ(2)+−δ2−ϵT2+ϵ)sds≤Ceσ(2)+t. |
So for t<T,
m2∑j=1|w2j(t)|2≤[m2∑j=1|w2j(t)|]2≤Ce2σ(2)+t. | (2.37) |
Moreover, we choose 0<δ<min{2ϵ−ϵT,2σ(3)+−2σ(2)+}; then,
∞∑k=3mk∑j=1(wkj(t))2≤C∞∑k=3mk∑j=1e2σ(k)+(t−T)+Cδ∫Tte(2σ(3)+−δ)(t−s)∞∑k=3mk∑j=1(gkj(s))2ds+Cδ∫t−∞e(2σ(3)−+δ)(t−s)∞∑k=3mk∑j=1(gkj(s))2ds≤Ce2σ(3)+t+C∫Tte(2σ(3)+−δ)(t−s)e(2σ(2)+−δ−ϵT+2ϵ)sds+C∫t−∞e(2σ(3)−+δ)(t−s)e(2σ(2)+−δ−ϵT+2ϵ)sds≤Ce2σ(3)+t+Cmax{e(2σ(3)+−δ)t,e(2σ(2)+−δ−ϵT+2ϵ)t}+Ce(2σ(2)+−δ−ϵT+2ϵ)t≤Ce2σ(2)+t. | (2.38) |
It is known from (2.36) and δ<2ϵ−ϵT that
‖F(w)‖2L2(SN−1)=O(e2σ(2)+t),m1∑j=1(g1j)2=O(e2σ(2)+t),(g01)2=O(e2σ(2)+t). |
By (2.26) and g1j(t)=O(eσ(2)+t), we obtain
m1∑j=1(w1j(t))2=O(e2σ(2)+t). | (2.39) |
Similarly,
(w01(t))2=O(e2σ(2)+t). | (2.40) |
We obtain from (2.37)–(2.40) that
‖w‖L2(SN−1)=O(eσ(2)+t). | (2.41) |
For any fixed (t,θ)∈(−∞,T−1)×SN−1, by applying the interior L∞-estimate to (2.4) with (t−1,t+1)×SN−1, we obtain from (2.41) and (2.24) that
|w(t,θ)|≤C{‖w‖L2((t−1,t+1)×SN−1)+‖F(w)‖L2((t−1,t+1)×SN−1)}≤Ceσ(2)+t, | (2.42) |
where C>0 is independent of t. Note that we can also use arguments similar to those in the proof of [18] to obtain
maxθ∈SN−1|w(t,θ)|≤Meσ(2)+tfor t∈(−∞,T−1). | (2.43) |
Defining
v(r,θ)=w(t,θ),r=et, |
it follows that v(r,θ) satisfies
Δv+2(N−2)vr2+F(v)r2=0in BR∖{0}, | (2.44) |
where R=eT−1. For any x0∈BR∖{0}, denote r0=|x0|>0 and Ω=Br0/2(x0). Consider (2.44) to be a linear equation in Ω as in Lemma 5.1 and Theorem 5.1 of [18] with
k=k1=1,h(x)≡0,|c|=Qr20,k2=Qr20 |
where Q=Q(v)>0. Then, (2.41) implies that there is a positive constant (independent of r0)
M=M(k1/k,k2r20)=M(Q)=M(v) |
such that
supx∈Br0/4(x0)|v(x)|≤Mrσ(2)+0. |
In particular, we have
|v(x0)|≤Mrσ(2)+0, |
max|x|=r|v(x)|≤Mrσ(2)+. |
Hence, (2.43) follows for t∈(−∞,T−1).
For the case of 4ϵ=2σ(2)+−δ, we may choose δ′ a little larger than δ such that 0<δ<δ′ and 4ϵ>2σ(2)+−δ′. By similar arguments, we can prove (2.41) and (2.43).
For the case (ⅱ), by F(w)=O(e2ϵt), ∑∞k=2∑mkj=1(gkj(s))2=O(e4ϵt) and 4ϵ<2σ(2)+−δ<2σ(2)+, we can obtain
∞∑k=2mk∑j=1(wkj(t))2≤C∞∑k=2mk∑j=1e2σ(k)+(t−T)+C∫Tte(2σ(2)+−δ)(t−s)∞∑k=2mk∑j=1(gkj(s))2ds+C∫t−∞e(2σ(2)−+δ)(t−s)∞∑k=2mk∑j=1(gkj(s))2ds≤Ce4ϵt. |
Then
[W(t)]2≤Ce4ϵtfor t<T. |
Together with (2.27) and (2.28), we know that
‖w‖L2(SN−1)=O(e2ϵt). |
Arguments similar to those in the proof of (2.43) imply that
maxθ∈SN−1|w(t,θ)|≤Me2ϵtfor t∈(−∞,−1], | (2.45) |
where M:=M(w)>0. As a consequence,
maxSN−1|F(w)|≤Ce4ϵtfor t∈(−∞,−1]. | (2.46) |
Then (2.26) implies that
w1j(t)=O(e4ϵt),j=1,2,…m1. | (2.47) |
Similarly,
w01(t)=O(e4ϵt). | (2.48) |
Therefore,
[W(t)]2≤Ce2σ(2)+(t−T)+C∫Tte(2σ(2)+−δ)(t−s)e8ϵsds+C∫Tte(2σ(2)+−δ)(t−s)e2ϵs[W(s)]2ds+C∫t−∞e(2σ(2)−+δ)(t−s)e8ϵsds+C∫t−∞e(2σ(2)−+δ)(t−s)e2ϵs[W(s)]2ds. | (2.49) |
Note that
∞∑k=2mk∑j=1(gkj(t))2≤Ce2ϵt([W(t)]2+e8ϵt). |
In what follows, there are also two cases: (a) 8ϵ≥[2σ(2)+−δ] and (b) 8ϵ<[2σ(2)+−δ].
For the case (a), using (2.49) and arguments similar to those in the proof of (i), we can obtain (2.43).
The case (b) implies that F(w)=O(e4ϵt). Then
[W(t)]2≤Ce8ϵtfor t<T. |
This, (2.47) and (2.48) imply that
‖w‖L2(SN−1)=O(e4ϵt) | (2.50) |
and
‖F(w)‖L2(SN−1)=O(e5ϵt). | (2.51) |
Therefore,
g1j(t)=O(e5ϵt),w1j(t)=O(e5ϵt),w01(t)=O(e5ϵt). | (2.52) |
Then we have
∞∑k=2mk∑j=1(gkj(t))2≤Ce2ϵt([W(t)]2+e10ϵt) |
and
[W(t)]2≤Ce2σ(2)+(t−T)+C∫Tte(2σ(2)+−δ)(t−s)e10ϵsds+C∫Tte(2σ(2)+−δ)(t−s)e2ϵs[W(s)]2ds+C∫t−∞e(2σ(2)−+δ)(t−s)e10ϵsds+C∫t−∞e(2σ(2)−+δ)(t−s)e2ϵs[W(s)]2ds. | (2.53) |
Similarly, we still consider two cases: 10ϵ≥[2σ(2)+−δ] and 10ϵ<[2σ(2)+−δ]; then, we obtain (2.43).
The proof of this lemma is complete.
Theorem 2.3. Assume that N≥3 and u∈C2(B∖{0}) is a singular solution of (1.15) that satisfies (1.13). Defining w(t,θ)=u(x)−Us(x) and t=lnr, there is a positive number sequence {μk}k≥1, strictly increasing and converging to ∞ with
μ1=σ(2)+ | (2.54) |
such that for any positive integer n≫1 and any (t,θ)∈(−∞,−1)×SN−1,
w(t,θ)=n∑k=1k−1∑ℓ=0ckℓ(θ)tℓeμkt+O(|t|neμn+1t), | (2.55) |
where
ckℓ(θ)=Mkℓ∑i=0akℓiQi(θ) | (2.56) |
and Mkℓ is a nonnegative integer depending on N,k,ℓ; akℓi is constant and Qi(θ) is a linear combination of {Qi1(θ),Qi2(θ),…,Qimi(θ)}. Especially, for k=1,
c10(θ)=a102Q2(θ), |
where a102 is a constant.
Proof. By using the starting estimate (2.22), constructing the index set I, and examining the equation of w, the expansion of w(t,θ) can be established via similar arguments to those in Theorem 1.1 of [19].
Let {ρk}k≥1 be positive strictly increasing and converging to ∞:
ρ1=σ(2)+,ρ2=σ(3)+,…,ρk=σ(k+1)+,…. |
Also, let Z+ be the collection of nonnegative integers. Define the index set I by
I={∑k≥1nkρk:nk∈Z+with finitely many nk>0}. | (2.57) |
Set
Iρ={ρk:k≥1} | (2.58) |
and
I˜ρ={i∑k=1nkρk:nk∈Z+,i∑k=1nk≥2}. | (2.59) |
Assume that I˜ρ is given by a strictly increasing sequence {˜ρk}k≥1 with ˜ρ1=2ρ1. There may be identical elements in Iρ and I˜ρ.
For ˜ρk∈I˜ρ, there are nonnegative integers n1,…,ni1 such that
n1+…+ni1≥2,n1ρ1+…+ni1ρi1=˜ρk. | (2.60) |
The collections of nonnegative integers n1,…,ni1 that satisfy (2.60) are finite. Set
˜Mk=max{2n1+3n2+…+(i1+1)ni1:n1,…,ni1are nonnegative integers satisfying (2.60)}. | (2.61) |
Arrange I as follows:
ρ1<…<ρi1≤˜ρ1<…<˜ρl1≤ρi1+1<…<ρi2≤˜ρl1+1<…<˜ρl2≤ρi2+1<…. | (2.62) |
Note that if ρ1<˜ρ1<ρ2, we choose i1=1 and l1=1 and the arrangement of (2.62) becomes ρ1<˜ρ1<ρ2<…. Similarly, if ρik+1≤˜ρlk+1<ρik+2 for some k≥1, define ik+1=ik+1 and lk+1=lk+1. We do not consider the multiplicity of ρk here, since all terms containing eρkt in the expansions of w(t,θ) can be combined as one term. We know
L(w)=−F(w), |
where F(w)=∑∞k=2bkwk for |w|<ˆϵ with some sufficiently small ˆϵ>0; also, the expansion of F(w) consists of terms including ∑I(k)ℓ=0(∑˜Mki=0ckℓiQi(θ))tℓe˜ρkt for ˜ρk∈I˜ρ in (2.59).
Define
μ1=ρ1,μ2=ρ2,μ3=ρ3,…,μi1=ρi1,μi1+1=˜ρ1,… | (2.63) |
according to the arrangement in (2.62). To ensure that {μk}k≥1 is a strictly increasing sequence of positive constants, when ρi1=˜ρ1, define μi1=ρi1 and μi1+1=˜ρ2. In this case, an extra power of t term corresponding to μi1 may appear in the expansion of w(t,θ). Similarly, make the same choices of μk for the cases ˜ρl1=ρi1+1, ρi2=˜ρl1+1, ˜ρl2=ρi2+1, etc. As a consequence, for any positive integer n≫1 and any (t,θ)∈(−∞,−1)×SN−1,
w(t,θ)=n∑k=1k−1∑ℓ=0ckℓ(θ)tℓeμkt+O(|t|neμn+1t), | (2.64) |
where
ckℓ(θ)=Mkℓ∑i=0akℓiQi(θ) | (2.65) |
and Mkℓ is a nonnegative integer that is dependent on N,k,ℓ, akℓi is constant and Qi(θ) is in the span of Qi1(θ),Qi2(θ),…,Qimi(θ). Especially, for k=1,
c10(θ)=a102Q2(θ). |
The proof of this theorem is complete.
We will introduce the appropriate weighted Hölder spaces and obtain the inverse of the operator L on those spaces, where L is given in (2.4). We use some ideas from [20], where the authors derived singular solutions to the following equation:
Δu+N(N−2)4uN+2N−2=0in B∖{0}. |
Fix a t0<0. For a nonnegative integer i, α∈(0,1), and μ∈R, define
‖v‖Ciμ((−∞,t0]×SN−1)=i∑j=0sup(t,θ)∈(−∞,t0]×SN−1e−μt|∇jv(t,θ)|, |
and
‖v‖Ci,αμ((−∞,t0]×SN−1)=‖v‖Ciμ((−∞,t0]×SN−1)+supt≤t0−1e−μt[∇iv]Cα([t−1,t+1]×SN−1), |
where [⋅]Cα is the usual Hölder semi-norm.
Definition 3.1. The collection of functions v in Ci((−∞,t0]×SN−1) with a finite norm ‖v‖Ci,αμ((−∞,t0]×SN−1) is the weighted Hölder space Ci,αμ((−∞,t0]×SN−1).
For μ>0 and some g∈C0,αμ((−∞,t0]×SN−1), to consider the linear equation given by
Lv=gin(−∞,t0)×SN−1, | (3.1) |
we introduce a boundary condition on t=t0 such that
L:C2,αμ((−∞,t0]×SN−1)→C0,αμ((−∞,t0]×SN−1) |
has a bounded inverse. However, since signs of coefficients of zero order terms are inappropriate, we cannot directly apply the maximum principle to the following Dirichlet boundary-value problem
{Lv=gin (−∞,t0)×SN−1,v=φon {t0}×SN−1. | (3.2) |
Lemma 3.1. Let μ>0, g∈C0μ((−∞,t0]×SN−1), and φ∈C0(SN−1). Then, there is at most one solution v∈C2μ((−∞,t0]×SN−1) of (3.2).
Proof. Assume that g=0, φ=0 and v∈C2μ((−∞,t0]×SN−1) is a solution of (3.2). For each k≥0, define
vk(t)=∫SN−1v(t,θ)Qk(θ)dθ. |
So Lk(vk)=0 on (−∞,t0) and vk(t0)=0. This implies that vk is a linear combinations of the basis of Ker(Lk). In particular, for k=0,
v0(t)={c10eℜ(σ(0)+)tcosγt+c20eℜ(σ(0)+)tsinγt,for 3≤N≤9,c10eσ(0)+t+c20teσ(0)+t,forN=10,c10eσ(0)+t+c20eσ(0)−t,for N≥11, |
and for k≥1,
vk(t)=c1keσ(k)+t+c2keσ(k)−t, |
where c1k and c2k are constants for k=0,1,2,…. By the assumption, we have the following for any t∈(−∞,t0):
|e−μtvk(t)|≤C. | (3.3) |
Hence, vk=0 for k=0 and k=1. Note that
{ℜ(σ(0)+)<0,ℜ(σ(0)−)<0,for 3≤N≤9,σ(0)+<0,σ(0)−<0,for N≥10, |
σ(1)+≤0 and σ(1)−<0 for N≥3. Moreover, vk(t)=c1keσ(k)+t for k≥2, which decays exponentially as t→−∞ (note that c2k=0 since σ(k)−<0 for k≥2). Since vk(t0)=0, we can directly obtain c1k=0 and vk(t)≡0 for k≥2. In conclusion, vk=0 for all k≥0, i.e., v≡0.
Lemma 3.2. Let α∈(0,1), μ>0, g∈C0,αμ((−∞,t0]×SN−1), and φ∈C2,α(SN−1). Suppose that v∈C2,αμ((−∞,t0]×SN−1) is a solution of (3.2). Then
‖v‖C2,αμ((−∞,t0]×SN−1)≤C[‖v‖C0μ((−∞,t0]×SN−1)+‖g‖C0,αμ((−∞,t0]×SN−1)+e−μt0‖φ‖C2,α(SN−1)], | (3.4) |
where C is a positive constant that is only dependent on N,α,μ and is independent of t0.
Proof. Using similar arguments to that of Lemma 2.5 of [20], consider two cases:
(ⅰ) t<t0−2. We have
2∑j=0supSN−1|∇jv(t,⋅)|+[∇2v]Cα([t−1,t+1]×SN−1)≤C[‖v‖L∞([t−2,t+2]×SN−1)+‖g‖L∞([t−2,t+2]×SN−1)+[g]Cα([t−2,t+2]×SN−1)], |
where C is a positive constant that is independent of t. We estimate [g]Cα([t−2,t+2]×SN−1), by setting (t1,θ1), (t2,θ2)∈[t−2,t+2]×SN−1 with (t1,θ1)≠(t2,θ2). There are two cases: |t1−t2|≤2 and |t1−t2|>2.
When |t1−t2|≤2, choose t′∈[t−1,t+1] such that t1,t2∈[t′−1,t′+1] is satisfied. Then,
[g]Cα([t−2,t+2]×SN−1)≤max{supt′∈[t−1,t+1][g]Cα([t′−1,t′+1]×SN−1),‖g‖L∞([t−2,t+2]×SN−1)}. |
So,
2∑j=0supSN−1|∇jv(t,⋅)|+[∇2v]Cα([t−1,t+1]×SN−1)≤C[‖v‖L∞([t−2,t+2]×SN−1)+‖g‖L∞([t−2,t+2]×SN−1)+supt′∈[t−1,t+1][g]Cα([t′−1,t′+1]×SN−1)]. |
We multiply both sides by e−μt and take the supremum over t∈(−∞,t0−2). The following holds
2∑j=0supt∈(−∞,t0−2)supSN−1e−μt|∇jv(t,⋅)|+supt∈(−∞,t0−2)e−μt[∇2v]Cα([t−1,t+1]×SN−1)≤C[‖v‖C0μ((−∞,t0]×SN−1)+‖g‖C0,αμ((−∞,t0]×SN−1)], | (3.5) |
where C is a positive constant that is independent of t0.
(ⅱ) t0−2≤t≤t0. From the boundary Schauder estimate, we see that
2∑j=0supSN−1|∇jv(t,⋅)|+[∇2v]Cα([t0−3,t0]×SN−1)≤C[‖v‖L∞([t0−4,t0]×SN−1)+‖g‖L∞([t0−4,t0]×SN−1)+[g]Cα([t0−4,t0]×SN−1)+‖φ‖C2,α(SN−1)]. |
Similarly,
2∑j=0supt∈[t0−2,t0]supSN−1e−μt|∇jv(t,⋅)|+supt∈[t0−2,t0−1]e−μt[∇2v]Cα([t−1,t+1]×SN−1)≤C[‖v‖C0μ((−∞,t0]×SN−1)+‖g‖C0,αμ((−∞,t0]×SN−1)+e−μt0‖φ‖C2,α(SN−1)]. | (3.6) |
Combining (3.5) and (3.6), (3.4) holds.
Then, by arguments similar to those in [20,21,22], we obtain the L∞ estimates of solutions on finite cylinders to (3.2) with a 0 boundary value.
Lemma 3.3. Let μ>ρ1 and μ≠ρk for k≥1, T and t0 be constants with t0≤0 and T−t0≤−4, and g∈C0([T,t0]×SN−1). Suppose that v∈C2([T,t0]×SN−1) satisfies the following:
{Lv=gin (T,t0)×SN−1,v=0on ({T}∪{t0})×SN−1, |
and ∫SN−1v(t,θ)Qk(θ)dθ=0 for k=0,1,⋯,K, where K is the largest integer satisfying that ρK−1<μ. Then,
sup(t,θ)∈[T,t0]×SN−1e−μt|v(t,θ)|≤Csup(t,θ)∈[T,t0]×SN−1e−μt|g(t,θ)|, | (3.7) |
where C is a positive constant dependent only on N,μ and independent of T and t0.
Proof. Note that ρ1=σ(2)+.
Let the sequences {Ti}, {ti}, {vi} and {gi} with ti≤0 and Ti−ti≤−4, satisfy the following:
{Lvi=giin (Ti,ti)×SN−1,vi=0on ({Ti}∪{ti})×SN−1, |
and
sup(t,θ)∈[Ti,ti]×SN−1e−μt|gi(t,θ)|=1, |
sup(t,θ)∈[Ti,ti]×SN−1e−μt|vi(t,θ)|→∞as i→∞. |
There exists t∗i∈(Ti,ti) that satisfies
Mi=supSN−1e−μt∗i|vi(t∗i,⋅)|=sup(t,θ)∈[Ti,ti]×SN−1e−μt|vi(t,θ)|→∞as i→∞. |
Set
˜vi(t,θ)=M−1ie−μt∗ivi(t+t∗i,θ), | (3.8) |
˜gi(t,θ)=M−1ie−μt∗igi(t+t∗i,θ). | (3.9) |
The following holds:
supSN−1|˜vi(0,⋅)|=1, |
for any (t,θ)∈[Ti−t∗i,ti−t∗i]×SN−1,
|e−μt˜vi(t,θ)|≤1 | (3.10) |
and
L˜vi=˜gifor t∈(Ti−t∗i,ti−t∗i)×SN−1. |
Assume the following for some τ−∈R−∪{−∞} and τ+∈R+∪{∞}:
Ti−t∗i→τ−,ti−t∗i→τ+. | (3.11) |
From (3.10) we have
|˜vi|≤Ceμ(t∗i−Ti)on (Ti−t∗i,Ti−t∗i+2)×SN−1, |
and hence
|d2˜vidt2+(N−2)d˜vidt+Δθ˜vi|≤Ceμ(t∗i−Ti)on (Ti−t∗i,Ti−t∗i+2)×SN−1. |
Since ˜vi=0 on {Ti−t∗i}×SN−1, we have
|∇˜vi|≤Ceμ(t∗i−Ti)on (Ti−t∗i,Ti−t∗i+1)×SN−1. |
This implies that Ti−t∗i remains bounded away from zero. Similar arguments imply that ti−t∗i is bounded away from zero. As a consequence, 0∈(τ−,τ+). Let
˜vi→ˆvin the compact set of (τ−,τ+). | (3.12) |
Moreover, ˜gi→0 in every compact set of (τ−,τ+). So the following holds:
ˆv≠0,|e−μtˆv(t,θ)|≤1for any (t,θ)∈(τ−,τ+)×SN−1,Lˆv=0on (τ−,τ+)×SN−1, | (3.13) |
and
limt→τ∗ˆv(t,θ)=0 | (3.14) |
where τ∗=τ− or τ+ if it is finite.
Let
ˆvk(t)=∫SN−1ˆv(t,θ)Qk(θ)dθ. | (3.15) |
Then Lk(ˆvk)=0; hence, ˆvk is the linear combination of the basis of Ker(Lk). We now take k≥2 with ρk>μ. Then
ˆvk+1(t)=c1k+1eρkt+c2k+1eσ(k+1)−t, |
where c1k+1 and c2k+1 are constants. From (3.13), we obtain the following for any t∈(τ−,τ+):
|e−μtˆvk+1(t)|≤C. |
When τ+=∞, c1k+1=0 and hence ˆvk+1(t)=c2k+1eσ(k+1)−t. When τ+ is finite, limt→τ+ˆvk+1(t)=0 by (3.14). Similarly, when τ−=−∞, ˆvk+1(t)=c1k+1eρkt=c1k+1eσ(k+1)+t. When τ− is finite, limt→τ−ˆvk+1(t)=0 by (3.14). Thus,
∫τ+τ−[(∂tˆvk+1)2+[λk+1−2(N−2)]ˆv2k+1]dt=0. |
Since ρk>μ>0, it follows that λk+1>2N for each k. This implies that λk+1−2(N−2)>0 for each k. Therefore, ˆvk+1=0 for each k. By the assumption, we have
ˆv0=ˆv1=…=ˆvk=0 |
provided that ρk−1<μ. In conclusion, ˆvk=0 for any k≥0; hence, ˆv≡0. This is a contradiction.
Lemma 3.4. Let α∈(0,1), μ>ρK for some K≥1, and g∈C0,αμ((−∞,t0]×SN−1) with g(t,⋅)∈span{Q0,Q1,…,QK+1} for t≤t0. Then, there is a unique solution v∈C2,αμ((−∞,t0]×SN−1) of (3.1) with v(t,⋅)∈span{Q0,Q1,…,QK+1} for t≤t0. Furthermore, g↦v is linear, and
‖v‖C2,αμ((−∞,t0]×SN−1)≤C‖g‖C0,αμ((−∞,t0]×SN−1), |
where C is a positive constant that is only dependent on N,α,μ and independent of t0.
Proof. For k=0,1,…,K+1, define
gk(t)=∫SN−1g(t,θ)Qk(θ)dθ. |
Then,
‖gk‖C0,αμ((−∞,t0])≤C‖g‖C0,αμ((−∞,t0]×SN−1), |
and
g(t,θ)=K+1∑k=0gk(t)Qk(θ). | (3.16) |
Consider the following ODE:
Lkvk=gk. | (3.17) |
Suppose that there is a solution vk∈C2,αμ((−∞,t0]) of (3.17) and
‖vk‖C2,αμ((−∞,t0])≤C‖gk‖C0,αμ((−∞,t0]), | (3.18) |
where C is a constant that depends only on N,α,μ and independent of t0.
If k=0, it is known from Section 2 that Ker(L0) encompasses eτtcosγt and eτtsinγt with
τ=−12(N−2)<0,γ=12√(N−2)(10−N) |
provided that 3≤N≤9. Ker(L0) encompasses eσ(0)+t and teσ(0)+t with σ(0)+=σ(0)−<0 provided that N=10. Ker(L0) encompasses eσ(0)+t and eσ(0)−t with σ(0)+<0, σ(0)−<0 provided that N≥11. If k=1, Ker(L1) encompasses eσ(1)+t and eσ(1)−t with σ(1)+≤0, σ(1)−<0 provided that N≥3.
We now consider k=0 and k=1. For k=0, we see that
v0(t)={B10∫t−∞eτ(t−s)sinγ(t−s)g0(s)ds,for 3≤N≤9,∫t−∞seσ(0)+(t−s)g0(s)ds−t∫t−∞eσ(0)+(t−s)g0(s)ds,for N=10,B10∫t−∞eσ(0)+(t−s)g0(s)ds−B10∫t−∞eσ(0)−(t−s)g0(s)ds,for N≥11, | (3.19) |
where
|B10|={1γ,for 3≤N≤9,|1σ(0)+−σ(0)−|,for N≥11. |
We only consider v0(t) for N≥11 and v1(t). The other cases for v0(t) can be demonstrated similarly since τ<0 and σ(0)+<0 in these cases. Let
vk(t)=B1k∫t−∞eσ(k)+(t−s)gk(s)ds−B1k∫t−∞eσ(k)−(t−s)gk(s)ds, | (3.20) |
where
|B1k|=|1σ(k)−−σ(k)+|. |
Direct calculation implies the following for t≤t0:
e−μt|vk(t)|≤Csupt≤t0e−μt|gk(t)|=C‖gk‖C0μ((−∞,t0]), | (3.21) |
e−μt(|v′k(t)|+|v″k(t)|)≤C‖gk‖C0μ((−∞,t0]). | (3.22) |
Set
v″k(t)=R1(t)+R2(t), |
where
R1(t)=B1k(eσ(k)+t)″∫t−∞e−σ(k)+sgk(s)ds−B1k(eσ(k)−t)″∫t−∞e−σ(k)−sgk(s)ds, |
and
R2(t)=B1k(eσ(k)+t)′e−σ(k)+tgk(t)−B1k(eσ(k)−t)′e−σ(k)−tgk(t). |
Then,
R′1(t)=B1k(eσ(k)+t)‴∫t−∞e−σ(k)+sgk(s)ds−B1k(eσ(k)−t)‴∫t−∞e−σ(k)−sgk(s)ds+B1k(eσ(k)+t)″e−σ(k)+tgk(t)−B1k(eσ(k)−t)″e−σ(k)−tgk(t). |
Then we have
e−μt|R′1(t)|≤C‖gk‖C0μ((−∞,t0]), |
and hence, for t≤t0−1,
e−μt[R1]Cα([t−1,t+1])≤C‖gk‖C0μ((−∞,t0]), |
e−μt[R2]Cα([t−1,t+1])≤C‖gk‖C0,αμ((−∞,t0]). |
Therefore, for t≤t0−1,
e−μt[v″k]Cα([t−1,t+1])≤C‖gk‖C0,αμ((−∞,t0]). | (3.23) |
Combining (3.21), (3.22) and (3.23), (3.18) holds for k=0 with N≥11 and k=1.
For 2≤k≤K+1, since e−μt|gk(t)|≤C and μ>ρK, we can also define
vk(t)=B1k∫t−∞eσ(k)+(t−s)gk(s)ds−B1k∫t−∞eσ(k)−(t−s)gk(s)ds. | (3.24) |
Note that, at this time, σ(k)+>0 and σ(k)−<0 for k=2,…,K+1 and ρk=σ(k+1)+. By arguments similar to those in the proof of (3.18) for k=0 with N≥11 and k=1, we obtain (3.18) for 2≤k≤K+1.
With the solution vk of (3.17) for k=0,1,…,K+1, we set
v(t,θ)=K+1∑k=0vk(t)Qk(θ). |
Then, Lv=g and, by (3.18), we have
‖v‖C2,αμ((−∞,t0]×SN−1)≤CK+1∑k=0‖vk‖C2,αμ((−∞,t0])≤CK+1∑k=0‖gk‖C0,αμ((−∞,t0])≤C‖g‖C0,αμ((−∞,t0]×SN−1). |
Then the extra requirement v(t,⋅)∈ span{Q0,Q1,…,QK+1} implies the uniqueness of v.
Lemma 3.5. Let α∈(0,1), μ>ρ1 and μ≠ρk for any k≥1; also, g∈C0,αμ((−∞,t0]×SN−1), with ∫SN−1g(t,⋅)Qk(θ)dθ=0 for any k=0,1,…,K,K+1, where K is the largest integer such that ρK<μ, and t≤t0. Then, there exists a unique solution v∈C2,αμ((−∞,t0]×SN−1) of (3.1) with v=0 on {t0}×SN−1. Moreover,
‖v‖C2,αμ((−∞,t0]×SN−1)≤C‖g‖C0,αμ((−∞,t0]×SN−1), | (3.25) |
where C is a positive constant that is dependent on N,α and μ and independent of t0.
Proof. Assume that T≤t0−4. We claim that there is a solution vT∈C2,α([T,t0]×SN−1) to the following problem:
{LvT=gin (T,t0)×SN−1,vT=0on ({T}∪{t0})×SN−1. | (3.26) |
The problem described by (3.26) can be written as follows:
{∂∂t(eτt∂vT∂t)+eτtΔθvT+2(N−2)eτtvT=eτtgin (T,t0)×SN−1,vT=0on ({T}∪{t0})×SN−1. | (3.27) |
where τ=N−2. Consider the energy function
GT(v)=∫t0T∫SN−1[eτt(∂tv)2+eτt|∇θv|2−2(N−2)eτtv2+2eτtgv]dtdθ. |
Set
Γ={ϕ∈H1(SN−1):∫SN−1ϕ(θ)Qk(θ)dθ=0for k=0,1,…,K,K+1withρK<μ}. |
Then, for any ϕ∈Γ,
∫SN−1|∇θϕ|2dθ≥2N∫SN−1ϕ2dθ, |
since λk+1≥λ2=2N with μ>ρk≥ρ1. Hence, for any v∈H10((T,t0)×SN−1) with v(t,⋅)∈Γ for any t∈(T,t0), we have
GT(v)≥∫t0T∫SN−1[eτt(∂tv)2+[2N−2(N−2)]eτtv2+2eτtgv]dθdt. |
The inequality 2N−2(N−2)>0 implies that GT is coercive and weakly lower semi-continuous. Then there is a minimizer vT of GT in the following statement:
{v∈H10((T,t0)×SN−1):v(t,⋅)∈Γfor any t∈(T,t0)}. |
So vT is a solution of (3.26) that satisfies vT(t,⋅)∈Γ for any t∈(T,t0).
We obtain that by Lemma 3.3,
sup(t,θ)∈[T,t0]×SN−1e−μt|vT(t,θ)|≤Csup(t,θ)∈[T,t0]×SN−1e−μt|g(t,θ)|, |
where C is a positive constant that is dependent on N,μ and independent of T and t0. Fix T0<t0. By the interior and boundary Schauder estimates in [t0+T0,t0]×SN−1⊂[t0+T0−1,t0]×SN−1 and vT(t0,θ)=0, there exists a subsequence vT that converges to a C2,α-solution v of (3.1) in [t0+T0,t0]×SN−1 satisfying v=0 on {t0}×SN−1, as T→−∞. Then, we can obtain that vT converges to a C2,α-solution v of (3.1) in (−∞,t0]×SN−1 such that the following is satisfied: v=0 on {t0}×SN−1,
sup(t,θ)∈[T,t0]×SN−1e−μt|v(t,θ)|≤Csup(t,θ)∈[T,t0]×SN−1e−μt|g(t,θ)|, |
or
‖v‖C0μ((−∞,t0]×SN−1)≤C‖g‖C0μ((−∞,t0]×SN−1), | (3.28) |
Combining (3.28) and (3.4) with φ=0, (3.25) holds.
Theorem 3.6. Let α∈(0,1), μ>ρ1, μ≠ρk for any k≥1 and g∈C0,αμ((−∞,t0]×SN−1). Then (3.1) admits a solution v∈C2,αμ((−∞,t0]×SN−1) and
‖v‖C2,αμ((−∞,t0]×SN−1)≤C‖g‖C0,αμ((−∞,t0]×SN−1), | (3.29) |
where C is a positive constant that is dependent on N,α,μ, and independent of t0. Also, g↦v is linear.
Proof. Assume that K≥1 is the largest integer with ρK<μ. Define
gk(t)=∫SN−1g(t,θ)Qk(θ)dθfor k=0,1,…,K+1. |
Then v1∈C2,αμ((−∞,t0]×SN−1) is a solution of
L(v1)=K+1∑k=0gk(t)Qk(θ)in (−∞,t0]×SN−1 |
as in Lemma 3.4. By Lemma 3.5, let v2∈C2,αμ((−∞,t0]×SN−1) be the unique solution of the following problem:
{Lv=g−K+1∑k=0gkQkin (−∞,t0]×SN−1,v=0on {t0}×SN−1. |
Then v=v1+v2 is a solution of (3.1) satisfying that v(t0,θ)=v1(t0,θ)=∑K+1k=0vk(t0)Qk(θ), where vk(t) for k=0,1,…,K,K+1 is given in Lemma 3.4.
Remark 3.7. Theorem 3.6 implies that the bound of
L−1:C0,αμ((−∞,t0]×SN−1)→C2,αμ((−∞,t0]×SN−1) | (3.30) |
is independent of t0.
In what follows, singular solutions of (1.1) will be constructed.
We set
N(w)=wtt+(N−2)wt+ΔSN−1w+2(N−2)(ew−1). | (4.1) |
Then w satisfies
wtt+(N−2)wt+ΔSN−1w+2(N−2)(ew−1)=0in (−∞,0)×SN−1 | (4.2) |
if N(w)=0 in (−∞,0)×SN−1. This also implies that u(x)=Us(x)+w(ln|x|,θ) is a solution of (1.15) in B∖{0}.
Theorem 4.1. Let Us(x) be given as in (1.2), the index set Iρ be given as in (2.58), and μ>ρ1 with μ∉Iρ. Suppose that ˆw∈C2,α((−∞,0]×SN−1) satisfies
|ˆw(t,θ)|+|∇ˆw(t,θ)|→0as t→−∞uniformlyinθ∈SN−1, | (4.3) |
and for (t,θ)∈(−∞,0]×SN−1,
|N(ˆw)(t,θ)|+|∇(N(ˆw))(t,θ)|≤Ceμt, | (4.4) |
where C is a positive constant. Then, there exist t0<0 and a solution w∈C2,α((−∞,t0]×SN−1) of the equation in (4.2) such that the following is satisfied for (t,θ)∈(−∞,t0)×SN−1:
|w(t,θ)−ˆw(t,θ)|≤Ceμt, | (4.5) |
where C is a positive constant.
Proof. The proof consists of 4 steps.
Step 1. We claim that there is z∈C2,αμ((−∞,t0]×SN−1) such that
N(ˆw+z)=0. | (4.6) |
Rewrite this equation as
Lz+N(ˆw)+P(z)=0, | (4.7) |
and
z=L−1[−N(ˆw)−P(z)], | (4.8) |
where
P(z)=2(N−2)[eˆw+z−eˆw−z]. | (4.9) |
Defining T by
T(z)=L−1[−N(ˆw)−P(z)], | (4.10) |
we prove that T is a contraction on some ball in C2,αμ((−∞,t0]×SN−1), for some t0<0 with |t0| large. We set
ΓB,t0={z∈C2,αμ((−∞,t0]×SN−1):‖z‖C2,αμ((−∞,t0]×SN−1)≤B}. |
Step 2. We claim that T maps ΓB,t0 to itself, for some fixed B and any t0 with |t0| sufficiently large, namely, for any z∈C2,αμ((−∞,t0]×SN−1) with ‖z‖C2,αμ((−∞,t0]×SN−1)≤B, we have that T(z)∈C2,αμ((−∞,t0]×SN−1) and ‖T(z)‖C2,αμ((−∞,t0]×SN−1)≤B.
First, it follows from (4.4) that
‖N(ˆw)‖C1μ((−∞,t0]×SN−1)≤C1. |
Next, set
E(z)=2(N−2)∫10(eˆw+sz−1)ds. | (4.11) |
Then, P(z)=zE(z). Take any z∈C2,αμ((−∞,t0]×SN−1) with ‖z‖C2,αμ((−∞,t0]×SN−1)≤B where B will be determined later. Because
|ˆw|+|∇ˆw|≤ϵ(t), |
where ϵ is increasing such that ϵ(t)→0 as t→−∞ and
|z|+|∇z|≤Beμt. |
Then, for t≤t0,
|E(z)|+|∇E(z)|≤C2(ϵ(t)+Beμt), | (4.12) |
and hence,
‖P(z)‖C1μ((−∞,t0]×SN−1)≤C2(ϵ(t0)+Beμt0)‖z‖C1μ((−∞,t0]×SN−1)≤C2(ϵ(t0)+Beμt0)B. |
By Theorem 3.6, we have
‖T(z)‖C2,αμ((−∞,t0]×SN−1)≤C‖−N(ˆw)−P(z)‖C0,αμ((−∞,t0]×SN−1)≤C[C1+C2(ϵ(t0)+Beμt0)B], |
where C, C1 and C2 are constants that are independent of t0. Choose B≥2CC1 and t0 with |t0| sufficiently large such that CC2(ϵ(t0)+Beμt0)≤1/2. Then,
‖T(z)‖C2,αμ((−∞,t0]×SN−1)≤B. |
Step 3. We claim that T:ΓB,t0→ΓB,t0 is a contraction, i.e., for any z1,z2∈ΓB,t0,
‖T(z1)−T(z2)‖C2,αμ((−∞,t0]×SN−1)≤κ‖z1−z2‖C2,αμ((−∞,t0]×SN−1), | (4.13) |
for some constant κ∈(0,1).
Note that
T(z1)−T(z2)=L−1[P(z2)−P(z1)], |
and
P(z1)−P(z2)=z1E(z1)−z2E(z2)=(z1−z2)E(z1)+z2(E(z1)−E(z2)). |
By (4.11), we have
E(z1)−E(z2)=2(N−2)∫10[eˆw+sz1−eˆw+sz2]ds. |
Then,
|E(z1)−E(z2)|+|∇(E(z1)−E(z2))|≤C(|z1−z2|+|∇(z1−z2)|). |
By (4.12), we have the following for any t≤t0,
|P(z1)−P(z2)|+|∇(P(z1)−P(z2))|≤C(ϵ(t)+Beμt)(|z1−z2|+|∇(z1−z2)|), |
and hence
‖P(z1)−P(z2)‖C1μ((−∞,t0]×SN−1)≤C(ϵ(t0)+Beμt0)‖z1−z2‖C1μ((−∞,t0]×SN−1). |
By Theorem 3.6, we obtain
‖T(z1)−T(z2)‖C2,αμ((−∞,t0]×SN−1)≤C‖P(z1)−P(z2)‖C0,αμ((−∞,t0]×SN−1)≤C(ϵ(t0)+Beμt0)‖z1−z2‖C2,αμ((−∞,t0]×SN−1). |
We derive (4.13) by choosing t0 with t0 sufficiently negative.
Step 4. The contraction mapping principle implies that there exists z∈C2,αμ((−∞,t0]×SN−1) such that T(z)=z. Then z∈C2,αμ((−∞,t0]×SN−1) is a solution of (4.6) and hence w=ˆw+z is a solution of (4.2).
We call a function ˆw that satisfies (4.3) and (4.4) an approximate solution to (4.2) with order μ.
Lemma 4.2. Assume that Qk(θ) is the combination of Qk1(θ),…,Qkmk(θ). Then,
QkQl=k+l∑i=0Qi. | (4.14) |
Proof. The proof is similar to that of Lemma 2.4 in [19]. We omit the details here.
Proposition 4.3. Let the index sets Iρ and I˜ρ be respectively given as in (2.58) and (2.59) and μ>ρ1 with μ∉Iρ∪I˜ρ. Let η be a solution of L(η)=0 on R×SN−1 such that η(t,⋅)→0 as t→−∞ uniformly on SN−1. Therefore for some t0<0, there exists a smooth function ˜η on (−∞,t0]×SN−1 such that ˆw=η+˜η satisfies (4.3) and (4.4).
Proof. Choose ϕ with |ϕ|<ˆϵ on (−∞,0)×SN−1, where ˆϵ>0 is a sufficiently small number. Then, a simple computation yields
N(ϕ)=Lϕ+2(N−2)(eϕ−1−ϕ). |
Therefore,
N(ϕ)=L(ϕ)+∞∑i=2aiϕi. | (4.15) |
Assume K≥1 and ˜K represent the largest corresponding integer with ρK<μ and ˜ρ˜K<μ, respectively. There is no function that converges to 0 as t→−∞ in KerL0 and KerL1; for k≥2, ψ+k(t)=eσ(k)+ and ψ−k(t)=eσ(k)−t in KerLkt. Any term eρkt with k>K in η will produce the term e˜ρlt with ˜ρl>μ in N(η); set
η(t,θ)=K∑k=1ckQk+1(θ)eρkt, | (4.16) |
where ck denotes constants.
For the following case:
Iρ∩I˜ρ=∅, | (4.17) |
we will show that there ˜η0,˜η1,…,˜η˜K exists in succession such that, for any i=0,1,…,˜K,
N(η+˜η0+…+˜ηi)=O(e˜ρi+1t). | (4.18) |
Set ϕ=η. From (4.15) and L(η)=0, we obtain
N(η)=∑n1+…+ni1≥2an1…ni1e(n1ρ1+…+ni1ρi1)tQn12…Qni1i1+1, |
where n1,…,ni1 are nonnegative integers and an1…ni1 is a constant. By the definition of I˜ρ, n1ρ1+…+ni1ρi1 is some ˜ρk. Hence, by Lemma 4.2,
N(η)=˜K∑k=1{˜Mk∑m=0akmQm(θ)}e˜ρkt+O(e˜ρ˜K+1t), | (4.19) |
where ˜Mk is defined as in (2.61) and akm is constant. We see that
N(η)=O(e˜ρ1t), |
where ˜ρ1=2ρ1. Then ˜η0=0.
Assume that (4.18) holds for 0,1,…,i−1. Set
˜ηi(t,θ)=(˜Mi∑m=0cimQm(θ))e˜ρit, | (4.20) |
where cim is constant. This implies that
N(η+˜η0+…+˜ηi)=L(˜η1)+…+L(˜ηi)+˜K∑k=1{˜Mk∑m=0akmQm(θ)}e˜ρkt+O(e˜ρ˜K+1t), |
where akm is different from that in (4.19). The induction hypothesis implies that
N(η+˜η0+…+˜ηi)=L(˜ηi)+˜K∑k=i{˜Mk∑m=0akmQm(θ)}e˜ρkt+O(e˜ρ˜K+1t). |
Choose ˜ηi such that
L(˜ηi)=−{˜Mi∑m=0aimQm(θ)}e˜ρit. |
So (4.18) holds for i. For m=0,1,…,˜Mi, solve
Lm(cime˜ρit)=−aime˜ρit. | (4.21) |
Similar to that in Lemma 3.4, there is a formula for cime˜ρit in terms of aime˜ρit. For 0<ρm<˜ρi, the expression is in the form of (3.24). For ρm>˜ρi, the expression is similar. If m=0 or 1, the expression is (3.19) or (3.20). Therefore, define
˜η(t,θ)=˜K∑k=1{˜Mk∑m=0ckmQm(θ)}e˜ρkt, | (4.22) |
where ckm is a constant. We obtain
N(η+˜η)=O(e˜ρ˜K+1t)=O(eμt). |
The estimate of ∇N(η+˜η) is similar.
Then for the general case, ρk can be some ˜ρi. We only need to modify (4.21). When ρm=˜ρi, there exist constants ci0m and ci1m such that
Lm((ci0m+tci1m)e˜ρit)=−aime˜ρit. |
By iteration, there are more powers of t. Therefore, there exist constants denoted by cijm with j=0,1,…,J+1 such that
Lm(J+1∑j=0cijmtje˜ρit)=J∑j=0aijmtje˜ρit. |
In conclusion, instead of (4.22), apply
˜η(t,θ)=˜K∑k=1k∑j=0{˜Mk∑m=0ckjmQm(θ)}tje˜ρkt, | (4.23) |
where ckjm is a constant.
Proof of Theorem 1.1
Theorem 4.1 and Proposition 4.3 imply that we can obtain t0 and a solution w∈C2,α((−∞,t0]×SN−1) of the equation in (4.2) such that, for μ>ρ1 with μ∉Iρ∪I˜ρ and any (t,θ)∈(−∞,t0)×SN−1,
w(t,θ)=ˆη(t,θ)+O(eμt),ˆη(t,θ)=η(t,θ)+˜η(t,θ). | (4.24) |
Let r0=et0. Then 0<r0<1. Since u(x)=Us(x)+w(ln|x|,θ), we add easily see that
u(x)=Us(x)+O(|x|σ(2)+)for x∈Br0∖{0}. | (4.25) |
On the other hand, for any R>0, we see that ˜u(y):=u(x)+2ln(R−1r0), y=Rr−10x satisfies the following equations:
−Δy˜u=e˜uinBR∖{0} | (4.26) |
and
˜u(y)=Us(y)+O(|y|σ(2)+)as |y|→0+. | (4.27) |
This implies that ˜u is a singular solution of (1.1).
Now, suppose that ˜u is non-radial. It suffices to prove that u is non-radial in Br0∖{0}. Suppose that u(x)=u(|x|). Then from Proposition 2.1 we have
u(x)≡Us(x)for x∈Br0∖{0}. | (4.28) |
Moreover, we can easily see from (4.24) and (4.25) that
w(t,θ)≢ |
This implies that and are non-radial singular solutions of (1.15) in and (1.1) in , respectively.
Since depends on the parameter , for each , different coefficients of in (4.16) can determine infinitely many and may change. When restricting all coefficients of in (4.16) in a bounded interval, there is a minimal . Since of in (4.16) depends on , we can obtain infinitely many by choosing a sequence of parameters with . Hence, a family of nonradial singular solutions of (1.1) can be constructed.
In this manuscript, infinitely many nonradial singular solutions have been constructed for the equation
where .
The authors declare that they have not used artificial intelligence tools in the creation of this article.
The research was supported by the 2024 Training Program for College Students Innovation and Entrepreneurship in Anhui University (No. X20240024). The third author was supported by the National Key RD Program of China (No. 2020YFA0713100), the National Natural Science Foundation of China (No. 12201001), and the Natural Science Foundation of Universities of Anhui Province (No. KJ2020A0009).
The authors declare that there is no conflicts of interest.
[1] |
W. X. Chen, C. M. Li, Classification of solutions for some nonlinear elliptic equations, Duke Math. J., 63 (1991), 615–622. https://doi.org/10.1215/S0012-7094-91-06325-8 doi: 10.1215/S0012-7094-91-06325-8
![]() |
[2] |
J. Prajapat, G. Tarantello, On a class of elliptic problems in : symmetry and uniqueness results, Proc. R. Soc. Edinburgh Sect. A: Math., 131 (2001), 967–985. https://doi.org/10.1017/S0308210500001219 doi: 10.1017/S0308210500001219
![]() |
[3] |
K. S. Chou, Y. H. Wan, Asymptotic radial symmetry for solutions of in a punctured disc, Pacific J. Math., 163 (1994), 269–276. https://doi.org/10.2140/PJM.1994.163.269 doi: 10.2140/PJM.1994.163.269
![]() |
[4] |
Z. M. Guo, F. S. Wan, Y. Y. Yang, Asymptotic expansions for singular solutions of in a punctured disc, Calc. Var. Partial Differ. Equations, 60 (2021), 35. https://doi.org/10.1007/s00526-021-01926-6 doi: 10.1007/s00526-021-01926-6
![]() |
[5] |
E. N. Dancer, A. Farina, On the classification of solutions of on : stability outside a compact set and applications, Proc. Am. Math. Soc., 137 (2009), 1333–1338. https://doi.org/10.1090/S0002-9939-08-09772-4 doi: 10.1090/S0002-9939-08-09772-4
![]() |
[6] |
A. Farina, Stable solutions of on , C. R. Math., 345 (2007), 63–66. https://doi.org/10.1016/j.crma.2007.05.021 doi: 10.1016/j.crma.2007.05.021
![]() |
[7] |
C. Wang, D. Ye, Some Liouville theorems for Hénon type elliptic equations, J. Funct. Anal., 262 (2012), 1705–1727. https://doi.org/10.1016/j.jfa.2011.11.017 doi: 10.1016/j.jfa.2011.11.017
![]() |
[8] |
M. F. B. Véron, L. Véron, Nonlinear elliptic equations on compact Riemannian manifolds and asymptotics of Emden equations, Invent. Math., 106 (1991), 489–-539. https://doi.org/10.1007/BF01243922 doi: 10.1007/BF01243922
![]() |
[9] |
S. Y. Chang, P. C. Yang, Prescribing Gaussian curvature on , Acta Math., 159 (1987), 215–259. https://doi.org/10.1007/BF02392560 doi: 10.1007/BF02392560
![]() |
[10] |
E. Onofri, On the positivity of the effective action in a theory of random surfaces, Commun. Math. Phys., 86 (1982), 321–326. https://doi.org/10.1007/BF01212171 doi: 10.1007/BF01212171
![]() |
[11] |
D. Cassani, L. L. Du, Z. S. Liu, Positive solutions to the planar logarithmic Choquard equation with exponential nonlinearity, Nonlinear Anal., 241 (2024), 113479. https://doi.org/10.1016/j.na.2023.113479 doi: 10.1016/j.na.2023.113479
![]() |
[12] |
R. Filippucci, M. Ghergu, Singular solutions for coercive quasilinear elliptic inequalities with nonlocal terms, Nonlinear Anal., 197 (2020), 111857. https://doi.org/10.1016/j.na.2020.111857 doi: 10.1016/j.na.2020.111857
![]() |
[13] | E. N. Dancer, Z. M. Guo, J. Wei, Non-radial singular solutions of the Lane-Emden equation in , Indiana Univ. Math. J., 61 (2012), 1971–1996. https://www.jstor.org/stable/24904111 |
[14] | Z. M. Guo, J. C. Wei, Non-radial singular solutions of in , In press. |
[15] |
Y. Miyamoto, Infinitely many nonradial singular solutions of in , , Proc. R. Soc. Edinburgh Sect. A: Math., 148 (2018), 133–147. https://doi.org/10.1017/S0308210517000051 doi: 10.1017/S0308210517000051
![]() |
[16] |
Z. M. Guo, X. Huang, F. Zhou, Radial symmetry of entire solutions of a bi-harmonic equation with exponential nonlinearity, J. Funct. Anal., 268 (2015), 1972–2004. https://doi.org/10.1016/j.jfa.2014.12.010 doi: 10.1016/j.jfa.2014.12.010
![]() |
[17] |
Z. M. Guo, J. Y. Li, F. S. Wan, Asymptotic behavior at the isolated singularities of solutions of some equations on singular manifolds with conical metrics, Commun. Partial Differ. Equations, 45 (2020), 1647–1681. https://doi.org/10.1080/03605302.2020.1784210 doi: 10.1080/03605302.2020.1784210
![]() |
[18] |
H. H. Zou, Symmetry of positive solutions of in , J. Differ. Equations 120 (1995), 46–88. https://doi.org/10.1006/jdeq.1995.1105 doi: 10.1006/jdeq.1995.1105
![]() |
[19] |
Q. Han, X. X. Li, Y. C. Li, Asymptotic expansions of solutions of the Yamabe equation and the -Yamabe equation near isolated singular points, Commun. Pure Appl. Math., 74 (2021), 1915–1970. https://doi.org/10.1002/cpa.21943 doi: 10.1002/cpa.21943
![]() |
[20] |
Q. Han, Y. C. Li, Singular solutions to the Yamabe equation with prescribed asymptotics, J. Differ. Equations, 274 (2021), 127–150. https://doi.org/10.1016/j.jde.2020.12.006 doi: 10.1016/j.jde.2020.12.006
![]() |
[21] |
R. Mazzeo, F. Pacard, Constant scalar curvature metrics with isolated singulatities, Duke Math. J., 99 (1999), 353–418. https://doi.org/10.1215/S0012-7094-99-09913-1 doi: 10.1215/S0012-7094-99-09913-1
![]() |
[22] |
Z. M. Guo, F. S. Wan, F. Zhou, Positive singular solutions of a nonlinear Maxwell equation arising in mesoscopic electromagnetism, J. Differ. Equations, 366 (2023), 249–291. https://doi.org/10.1016/j.jde.2023.03.056 doi: 10.1016/j.jde.2023.03.056
![]() |