Citation: Huyuan Chen, Laurent Véron. Weak solutions of semilinear elliptic equations with Leray-Hardy potentials and measure data[J]. Mathematics in Engineering, 2019, 1(3): 391-418. doi: 10.3934/mine.2019.3.391
[1] | Luigi Montoro, Berardino Sciunzi . Qualitative properties of solutions to the Dirichlet problem for a Laplace equation involving the Hardy potential with possibly boundary singularity. Mathematics in Engineering, 2023, 5(1): 1-16. doi: 10.3934/mine.2023017 |
[2] | Lars Eric Hientzsch . On the low Mach number limit for 2D Navier–Stokes–Korteweg systems. Mathematics in Engineering, 2023, 5(2): 1-26. doi: 10.3934/mine.2023023 |
[3] | María Ángeles García-Ferrero, Angkana Rüland . Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 2019, 1(4): 715-774. doi: 10.3934/mine.2019.4.715 |
[4] | Konstantinos T. Gkikas . Nonlinear nonlocal equations involving subcritical or power nonlinearities and measure data. Mathematics in Engineering, 2024, 6(1): 45-80. doi: 10.3934/mine.2024003 |
[5] | Youchan Kim, Seungjin Ryu, Pilsoo Shin . Approximation of elliptic and parabolic equations with Dirichlet boundary conditions. Mathematics in Engineering, 2023, 5(4): 1-43. doi: 10.3934/mine.2023079 |
[6] | Feida Jiang . Weak solutions of generated Jacobian equations. Mathematics in Engineering, 2023, 5(3): 1-20. doi: 10.3934/mine.2023064 |
[7] | Giovanni Cupini, Paolo Marcellini, Elvira Mascolo . Local boundedness of weak solutions to elliptic equations with $ p, q- $growth. Mathematics in Engineering, 2023, 5(3): 1-28. doi: 10.3934/mine.2023065 |
[8] | François Murat, Alessio Porretta . The ergodic limit for weak solutions of elliptic equations with Neumann boundary condition. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021031 |
[9] | Edgard A. Pimentel, Miguel Walker . Potential estimates for fully nonlinear elliptic equations with bounded ingredients. Mathematics in Engineering, 2023, 5(3): 1-16. doi: 10.3934/mine.2023063 |
[10] | Boumediene Abdellaoui, Ireneo Peral, Ana Primo . A note on the Fujita exponent in fractional heat equation involving the Hardy potential. Mathematics in Engineering, 2020, 2(4): 639-656. doi: 10.3934/mine.2020029 |
Schrödinger operators with singular potentials under the form
u↦H(u):=−Δu+V(x)u,x∈R3 | (1.1) |
are at the core of the description of many aspects of nuclear physics. The associated energy, the sum of the momentum energy and the potential energy, endows the form
H(u)=12∫R3(|∇u|2+V(x)u2)dx. | (1.2) |
In classical physics V(x)=−κ|x|−1 (κ>0) is the Coulombian potential and H is not bounded from below and there is no ground state. In quantum physics there are reasons arising from its mathematical formulation which leads, at least in the case of the hydrogen atom, to V(x)=−κ|x|−2 (κ>0) and H is bounded from below provided κ≥−14. Furthermore, a form of the uncertainty principle is Hardy's inequality
∫R3|∇u|2dx≥14∫R3u2|x|2dxfor all u∈C∞0(R3). | (1.3) |
The meaning of this inequality is that if u is localized close to a point 0 (i.e., the right side term is large), then its momentum has to be large (i.e., the left side term is large), and the power |x|−2 is the consequence of a dimensional analysis (see [19,20]). Such potential is often called a Leray-Hardy potential. The study of the mathematical properties of generalisations of the operator H in particular in N-dimensional domains generated hundred of publications in the last thirty years. In this article we define the Schrödinger operator L in RN by
Lμ:=−Δ+μ|x|2, | (1.4) |
where μ is a real number satisfying
μ≥μ0:=−(N−2)24. | (1.5) |
Note that (N−2)24 achieves the value 14 when N=3. Let Ω⊂RN (N≥2) be a bounded, smooth domain containing the origin and g:R→R be a continuous nondecreasing function such that g(0)≥0, we are interested in the nonlinear Poisson equation
{Lμu+g(u)=νin Ω,Lμ+g(u)u=0on ∂Ω, | (1.6) |
where ν is a Radon measure in Ω. The reason for a measure framework is that the problem is essentially trivial if ν∈L2(Ω), more complicated if ν∈L1(Ω) and very rich if ν is a measure.
When μ=0, problem (1.6) reduces to
{−Δu+g(u)=νin Ω,−Δ+g(u)u=0on ∂Ω, | (1.7) |
which has been extensively studied by numerous authors in the last 30 years. A fundamental contribution is due to Brezis [6], Benilan and Brezis [2], where ν is bounded and the function g:R→R is nondecreasing, positive on (0,+∞) and satisfies the subcritical assumption in dimension N≥3:
∫+∞1(g(s)−g(−s))s−1−NN−2ds<+∞. | (1.8) |
They obtained the existence, uniqueness and stability of weak solutions for the problem. When N=2, Vàzquez [26] introduced the exponential orders of growth of g defined by
β+(g)=inf{b>0:∫∞1g(t)e−btdt<∞},β−(g)=sup{b<0:∫−1−∞g(t)ebtdt>−∞}, | (1.9) |
and proved that if ν is any bounded measure in Ω with Lebesgue decomposition
ν=νr+∑j∈Nαjδaj, |
where νr is part of ν with no atom, aj∈Ω and αj∈R satisfies
4πβ−(g)≤αj≤4πβ+(g), | (1.10) |
then (1.7) admits a (unique) weak solution. Later on, Baras and Pierre [1] studied (1.7) when g(u)=|u|p−1u for p>1 and they discovered that if p≥NN−2 the problem is well posed if and only if ν is absolutely continuous with respect to the Bessel capacity c2,p′ with p′=pp−1.
It is a well established fact that, by the improved Hardy inequality in [9] and Lax-Milgram Theorem, the non-homogeneous problem
Lμu=finΩ,u=0on∂Ω, | (1.11) |
with f∈L2(Ω), has a unique solution in H10(Ω) if μ>μ0, or in a weaker space H(Ω) if μ=μ0, see [18]. When f∉L2(Ω), a natural question is to find sharp conditions on f for the existence or nonexistence of solutions of (1.11) and the difficulty comes from the fact that the Hardy term |x|−2u may not be locally integrable in Ω. An attempt done by Dupaigne in [18] is to consider problem (1.11) when μ∈[μ0,0) and N≥3 in the sense of distributions
∫ΩuLμξdx=∫Ωfξdx,∀ξ∈C∞c(Ω). | (1.12) |
The corresponding semi-linear problem is studied in [5] with this approach.
We adopt here a different point of view in using a different notion of weak solutions. It is known that the equation Lμu=0 in RN∖{0} has two distinct radial solutions:
Φμ(x)={|x|τ−(μ)ifμ>μ0,|x|−N−22ln(1|x|)ifμ=μ0,andΓμ(x)=|x|τ+(μ), |
with
τ−(μ)=−N−22−√(N−2)24+μandτ+(μ)=−N−22+√(N−2)24+μ. |
In the remaining of the paper and when there is no ambiguity, we put τ+=τ+(μ), τ0+=τ+(μ0), τ−=τ−(μ) and τ0−=τ−(μ0). It is noticeable that identity (1.12) cannot be used to express that Φμ is a fundamental solution, i.e., f=δ0, since Φμ is not locally integrable if μ≥2N. Recently, Chen, Quaas and Zhou found in [12] that the function Φμ is the fundamental solution in the sense that
∫RNΦμL∗μξdγμ(x)=cμξ(0)for all ξ∈C1,10(RN), | (1.13) |
where
dγμ(x)=Γμ(x)dx, L∗μξ=−Δξ−2τ+|x|2⟨x,∇ξ⟩, | (1.14) |
and
cμ={2√μ−μ0∣SN−1∣if μ>μ0,|SN−1|if μ=μ0. | (1.15) |
With the power-absorption nonlinearity in Ω∗=Ω∖{0}, the precise behaviour near 0 of any positive solution of
Lμu+up=0in D′(Ω∗) | (1.16) |
is given in [22] when p>1. In this paper it appears a critical exponent
p∗μ=1−2τ− | (1.17) |
with the following properties: if p≥p∗μ any solution of (1.16) can be extended by continuity as a solution in D′(Ω). If 1<p<p∗μ any positive solution of (1.16) either satisfies
limx→0|x|2p−1u(x)=ℓ, | (1.18) |
where ℓ=ℓN,p,μ>0, or there exists k≥0 such that
limx→0u(x)Φμ(x)=k, | (1.19) |
and in that case u∈Lploc(Ω;dγμ). In view of [12], it implies that u satisfies
∫RN(uL∗μξ+upξ)dγμ(x)=cμkξ(0),∀ξ∈C1,10(RN). | (1.20) |
Note the threshold p∗μ and its role is put into light by the existence or non-existence of explicit solutions of (1.16) under the form x↦a|x|b, where necessarily b=−2p−1 and a=ℓ. It is also proved in [22] that when μ>μ0 and g:R→R+ is a continuous nondecreasing function satisfying
∫∞1(g(s)−g(−s))s−1−p∗μds<∞, | (1.21) |
then for any k>0 there exists a radial solution of
Lμu+g(u)=0in D′(B∗1) | (1.22) |
satisfying (1.19), where B∗1:=B1(0)∖{0}. When μ=μ0 and N≥3 it is proved in [22] that if there exists b>0 such that
∫10g(−bs−N−2N+2lns)ds<∞, | (1.23) |
then there exists a radial solution of (1.22) satisfying (1.19) with γ=(N+2)b2. In fact this condition is independent of b>0, by contrast to the case N=2 and μ=0 where the introduction of the exponential order of growth of g is a necessity. Moreover, it is easy to see that u satisfies
∫RN(uL∗μξ+g(u)ξ)dγμ(x)=cμγξ(0),∀ξ∈C1,10(RN). | (1.24) |
In view of these results and identity (1.13), we introduce a definition of weak solutions adapted to the operator Lμ in a measure framework. Since Γμ is singular at 0 if μ<0, there is need of defining specific set of measures and we denote by M(Ω∗;Γμ), the set of Radon measures ν in Ω∗ such that
∫Ω∗Γμd|ν|:=sup{∫Ω∗ζd|ν|:ζ∈C0(Ω∗),0≤ζ≤Γμ}<∞. | (1.25) |
If ν∈M+(Ω∗), we define its natural extension, with the same notation since there is no ambiguity, as a measure in Ω by
∫Ωζdν=sup{∫Ω∗ηdν:η∈C0(Ω∗),0≤η≤ζ}for all ζ∈C0(Ω),ζ≥0, | (1.26) |
a definition which is easily extended if ν=ν+−ν−∈M(Ω∗). Since the idea is to use the weight Γμ in the expression of the weak solution, the expression Γμν has to be defined properly if τ+<0. We denote by M(Ω;Γμ) the set of measures ν on Ω which coincide with the above natural extension of ν⌊Ω∗∈M+(Ω∗;Γμ). If ν∈M+(Ω;Γμ) we define the measure Γμν in the following way
∫Ωζd(Γμν)=sup{∫Ω∗ηΓμdν:η∈C0(Ω∗),0≤η≤ζ}for all ζ∈C0(Ω),ζ≥0. | (1.27) |
If ν=ν+−ν−, Γμν is defined accordingly. Notice that the Dirac mass at 0 does not belong to M(Ω;Γμ) although it is a limit of {νn}⊂M(Ω;Γμ). We denote by ¯M(Ω;Γμ) the set of measures which can be written under the form
ν=ν⌊Ω∗+kδ0, | (1.28) |
where ν⌊Ω∗∈M(Ω;Γμ) and k∈R. Before stating our main theorem we make precise the notion of weak solution used in this article. We denote ¯Ω∗:=¯Ω∖{0}, ρ(x)=dist(x,∂Ω) and
Xμ(Ω)={ξ∈C0(¯Ω)∩C1(¯Ω∗):|x|L∗μξ∈L∞(Ω)}. | (1.29) |
Clearly C1,10(¯Ω)⊂Xμ(Ω).
Definition 1.1. We say that u is a weak solution of (1.6) with ν∈¯M(Ω;Γμ) such that ν=ν⌊Ω∗+kδ0 if u∈L1(Ω,|x|−1dγμ), g(u)∈L1(Ω,ρdγμ) and
∫Ω[uL∗μξ+g(u)ξ]dγμ(x)=∫Ωξd(Γμν)+kξ(0)for all ξ∈Xμ(Ω), | (1.30) |
where L∗μ is given by (1.13) and cμ is defined in (1.15).
A measure for which problem (1.6) admits a solution is a g-good measure. In the regular case we prove the following
Theorem A. Let μ≥0 if N=2, μ≥μ0 if N≥3 and g:R→R be a Hölder continuous nondecreasing function such that g(r)r≥0 for any r∈R. Then for any ν∈L1(Ω,dγμ), problem (1.6) has a unique weak solution uν such that for some c1>0,
‖uν‖L1(Ω,|x|−1dγμ)≤c1‖ν‖L1(Ω,dγμ). |
Furthermore, if uν′ is the solution of (1.6) with right-hand side ν′∈L1(Ω,dγμ), there holds
∫Ω[|uν−uν′|L∗μξ+|g(uν)−g(uν′)|ξ]dγμ(x)≤∫Ω(ν−ν′)sgn(u−u′)ξdγμ(x), | (1.31) |
and
∫Ω[(uν−uν′)+L∗μξ+(g(uν)−g(uν′))+ξ]dγμ(x)≤∫Ω(ν−ν′)sgn+(u−u′)ξdγμ(x), | (1.32) |
for all ξ∈Xμ(Ω), ξ≥0.
Definition 1.2. A continuous function g:R→R such that rg(r)≥0 for all r∈R satisfies the weak Δ2-condition if there exists a positive nondecreasing function t∈R↦K(t) such that
|g(s+t)|≤K(t)(|g(s)|+|g(t)|)forall(s,t)∈R×Rs.t.st≥0. | (1.33) |
It satisfies the Δ2-condition if the above function K is constant.
The Δ2-condition has been intruduced in the study of Birnbaum-Orlicz spaces [4,23] and it is satisfied by power function r↦|r|p−1r, p>0, but not by exponential functions r↦ear. It plays a key role in the study of semilinear equation with a power type reaction term (see eg., [29,30]). The new weak Δ2-condition is more general and it is also satisfied by exponential functions.
Theorem B. Let μ>0 if N=2 or μ>μ0 if N≥3 and g:R→R be a nondecreasing continuous function such that g(r)r≥0 for any r∈R. If g satisfies the weak Δ2-condition and
∫∞1(g(s)−g(−s))s−1−min{p∗μ,p∗0}ds<∞, | (1.34) |
where p∗μ is given by (1.17), then for any ν∈¯M+(Ω;Γμ) problem (1.6) admits a unique weak solution uν.
Note that min{p∗μ,p∗0}=p∗μ for μ>0 and min{p∗μ,p∗0}=p∗0 if μ<0. Furthermore, the mapping: ν↦uν is increasing. In the case N≥3 and μ=μ0 we have a more precise result.
Theorem C. Assume that N≥3 and g:R→R is a continuous nondecreasing function such that g(r)r≥0 for any r∈R satisfying the weak Δ2-condition and (1.8). Then for any ν=ν⌊Ω∗+cμkδ0∈¯M+(Ω;Γμ) problem (1.6) admits a unique weak solution uν.
Furthermore, if ν⌊Ω∗=0, condition (1.8) can be replaced by the following weaker one
∫∞1(g(t)−g(−t))(lnt)N+2N−2t−2NN−2dt<∞. | (1.35) |
The optimality of these conditions depends whether the measure is concentrated at 0 or not. When the measure is of the form kδ0 the condition proved to be optimal in [22] and when it is of the type kδa with a≠0 optimality is shown in [28]. Normally, the estimates on the Green kernel plays an essential role for approximating the solution of elliptic problems with absorption and Radon measure data. However, we have avoided to use the estimates on the Green kernel for Hardy operators which are not easily tractable when 0>μ≥μ0, and our main idea is to separate the measure ν∗ in M(Ω;Γμ) and the Dirac mass at the origin, and then to glue the solutions with above measures respectively. This technique requires this new weak Δ2-condition.
In the previous result, it is noticeable that if k=0 (resp. ν⌊Ω∗=0) only condition (1.8) (resp. condition (1.35)) is needed. In the two cases the weak Δ2-condition is unnecessary. In the power case where g(u)=|u|p−1u:=gp(u),
{Lμu+gp(u)=ν in Ω,−−−− u=0 on ∂Ω, | (1.36) |
the following result follows from Theorem B and C.
Corollary D. Let μ≥μ0 if N≥3 and μ>0 if N=2. Any nonzero measure ν=ν⌊Ω∗+cμkδ0∈¯M+(Ω;Γμ) is gp-good if one of the following holds:
(i) 1<p<p∗μ in the case ν⌊Ω∗=0;
(ii) 1<p<p∗0 in the case k=0;
(iii) 1<p<min{p∗μ,p∗0} in the case k≠0 and ν⌊Ω∗≠0.
We remark that p∗μ is the sharp exponent for the existence of (1.36) when ν⌊Ω∗=0, while the critical exponent becomes p∗0 when k=0 and ν has atom in Ω∖{0}.
The supercritical case of equation (1.36) corresponds to the fact that not all measures are gp-good and the case where k≠0 is already treated.
Theorem E. Assume that N≥3. Then ν=ν⌊Ω∗∈M(Ω;Γμ) is gp-good if and only if for any ϵ>0, νϵ=νχBcϵ is absolutely continuous with respect to the c2,p′-Bessel capacity.
Finally we characterize the compact removable sets in Ω.
Theorem F. Assume that N≥3, p>1 and K is a compact set of Ω. Then any weak solution of
Lμu+gp(u)=0in Ω∖K | (1.37) |
can be extended a weak solution of the same equation in whole Ω if and only if
(i) c2,p′(K)=0 if 0∉K;
(ii) p≥pμ∗ if K={0};
(iii) c2,p′(K)=0 if μ≥0, 0∈K and K∖{0}≠∅;
(iv) c2,p′(K)=0 and p≥p∗μ if μ<0, 0∈K and K∖{0}≠∅.
The case (ⅰ) is already proved in [22,Theorem 1.2]. Notice also that if A≠∅ necessarily c2,p′(A)=0 holds only if p≥p0. Therefore, if μ≥0 there holds p≥p∗0≥p∗μ, while if μ<0, then p0<p∗μ.
The rest of this paper is organized as follows. In Section 2, we build the framework for weak solutions of (1.6) involving L1 data. Section 3 is devoted to solve existence and uniqueness of weak solution of (1.6), where the absorption is subcritical and ν is a related Radon measure. Finally, we deal with the super critical case in Section 4 by characterized by Bessel Capacity.
Throughout this section we assume N≥2 and μ≥μ0 and in what follows, we denote by ci with i∈N a generic positive constant. We first recall some classical comparison results for Hardy operator Lμ. The next lemma is proved in [12,Lemma 2.1], and in [15,Lemma 2.1] if h(s)=sp.
Lemma 2.1. Let G be a bounded domain in RN such that 0∉ˉG, L:G×[0,+∞)↦[0,+∞) be a continuous function satisfying for any x∈G,
h(x,s1)≥h(x,s2)ifs1≥s2, |
and functions u,v∈C1,1(G)∩C(¯G) satisfy
{Lμu+h(x,u)≥Lμv+h(x,v)in G,Lμ+h(x,u)u≥von ∂G, |
then
u≥vinG. |
As an immediate consequence we have
Lemma 2.2. Assume that Ω is a bounded C2 domain containing 0. If L is a continuous function as in Lemma 2.1 verifying that L(x,0)=0 for all x∈Ω, and u∈C1,1(Ω∗)∩C(¯Ω∗) satisfies
{LLμu+L(x,u)=0 in Ω∗,LLμ+L(x,u)u=0 on ∂Ω,limx→0u(x)Φ−1μ(x)=0. | (2.1) |
Then u=0.
We recall that if u∈L1(Ω,|x|−1dγμ) is a weak solution of
{Lμu=f in Ω,u=0 on ∂Ω, | (2.2) |
in the sense of Definition 1.1, then it satisfies that
∫ΩuL∗μ(ξ)dγμ(x)=∫Ωfξdγμ(x)forall ξ∈Xμ(Ω). | (2.3) |
If u is a weak solution of (2.2), there holds
Lμu=f in D′(Ω∗), | (2.4) |
and v=Γ−1μu verifies
L∗μv=Γ−1μf in D′(Ω∗), | (2.5) |
a fact which is expressed by the commutating formula
ΓμL∗μv=Lμ(Γμv). | (2.6) |
The following form of Kato's inequality, proved in [12,Proposition 2.1], plays an essential role in the obtention a priori estimates and uniqueness of weak solution of (1.6).
Proposition 2.1. If f∈L1(Ω,ρdγμ), then there exists a unique weak solution u∈L1(Ω,|x|−1dγμ) of (2.2). Furthermore, for any ξ∈Xμ(Ω), ξ≥0, we have
∫Ω|u|L∗μ(ξ)dγμ(x)≤∫Ωsign(u)fξdγμ(x) | (2.7) |
and
∫Ωu+L∗μ(ξ)dγμ(x)≤∫Ωsign+(u)fξdγμ(x). | (2.8) |
The proof is done if ξ∈C1,10(Ω), but it is valid if ξ∈Xμ(Ω). The next result is proved in [13,Lemma 2.3].
Lemma 2.3. Assume that μ>μ0 and f∈C1(Ω∗) verifies
0≤f(x)≤c2|x|τ−2, | (2.9) |
for some τ>τ−. Let uf be the solution of
{L–Lμu=f in Ω∗,LLμ–u=0 on ∂Ω,limx→0u(x)Φμ(x)=0. | (2.10) |
Then there holds:
(i) if τ−<τ<τ+,
0≤uf(x)≤c3|x|τ inΩ∗; | (2.11) |
(ii) if τ=τ+,
0≤uf(x)≤c4|x|τ(1+(−ln|x|)+) inΩ∗; | (2.12) |
(iii) if τ>τ+,
0≤uf(x)≤c5|x|τ+ inΩ∗. | (2.13) |
Proof of Theorem A. Let H1μ,0(Ω) be the closure of C∞0(Ω) under the norm of
‖u‖H1μ,0(Ω)=√∫Ω|∇u|2dx+μ∫Ωu2|x|2dx. | (2.14) |
Then H1μ,0(Ω) is a Hilbert space with inner product
⟨u,v⟩H1μ,0(Ω)=∫Ω⟨∇u,∇v⟩dx+μ∫Ωuv|x|2dx | (2.15) |
and the embedding H1μ,0(Ω)↪Lp(Ω) is continuous and compact for p∈[2,2∗) with 2∗=2NN−2 when N≥3 and any p∈[2,∞) if N=2. Furthermore, if η∈C10(¯Ω) has the value 1 in a neighborhood of 0, then ηΓμ∈H1μ,0(Ω). We put
G(v)=∫v0g(s)ds, |
then G is a convex nonnegative function. If ρν∈L2(Ω) we define the functional Jν in the space H1μ,0(Ω) by
Jν(v)={12‖v‖2H1μ,0(Ω)+∫ΩG(v)dx−∫Ωνvdxif G(v)∈L1(Ω,dγμ),∞if G(v)∉L1(Ω,dγμ). | (2.16) |
The functional J is strictly convex, lower semicontinuous and coercive in H1μ,0(Ω), hence it admits a unique minimum u which satisfies
⟨u,v⟩H1μ,0(Ω)+∫Ωg(u)vdx=∫Ωνvdxforallv∈H1μ,0(Ω). |
If ξ∈C1,10(Ω) then v=ξΓμ∈H1μ,0(Ω), then
⟨u,ξΓμ⟩H1μ,0(Ω)=∫Ω⟨∇u,∇ξ⟩dγμ(x)+∫Ω(⟨∇u,∇Γμ⟩+μΓμ|x|2)ξdx, | (2.17) |
and
∫Ω⟨∇u,∇Γμ⟩ξdx=−∫Ω⟨∇ξ,∇Γμ⟩udx−∫ΩuξΔΓμdx, |
since C∞0(Ω) is dense in H1μ,0(Ω). Furthermore, since u∈Lp(Ω) for any p<2∗, |x|−1u∈L1(Ω,dγμ), hence uL∗μξ∈L1(Ω,dγμ). Therefore
∫Ω(uL∗μξ+g(u)ξ)dγμ=∫Ωνξdγμ. | (2.18) |
Next, if ν∈L1(Ω,ρdγμ) we consider a sequence {νn}⊂C∞0(Ω) converging to ν in L1(Ω,ρdγμ) and denote by {un} the sequence of the corresponding minimizing problem in H1μ,0(Ω). By Proposition 2.1 we have that, for any ξ∈Xμ(Ω),
∫Ω(|un−um|L∗μξ+(g(un)−g(um))sgn(un−um)ξ)dγμ≤∫Ω(νn−νm)sgn(un−um)ξdγμ. | (2.19) |
We denote by η0 the solution of
L∗μη=1in Ω, η=0on ∂Ω. | (2.20) |
Its existence is proved in [12,Lemma 2.2], as well as the estimate 0≤η0≤c6ρ for some c6>0. Since g is monotone, we obtain from (2.19)
∫Ω(|un−um|+|g(un)−g(um)|η0)dγμ≤∫Ω|νn−νm|η0dγμ. | (2.21) |
Hence {un} is a Cauchy sequence in L1(Ω,dγμ). Next we construct a solution η1 to
L∗μη=|x|−1 in Ω∗,η=0on ∂Ω. | (2.22) |
For this aim, we consider for 0<θ<1, the function yθ(x)=1−|x|2−θN−θ+2τ+(μ) which verifies
L∗μyθ=|x|−θ in B1,yθ=0on ∂B1 |
(we can always assume that Ω⊂B1). As in the proof of [12,Lemma 2.2], for any x0∈Ω there exists r0>0 such that Br0(x0)⊂Ω and for t>0 small enough wt,x0(x)=t(r20−|x−x0|2) is a subsolution of (2.20), hence of (2.22). Therefore there exists ηθ such that
L∗μηθ=|x|−θ in Ω∗,ηθ=0on ∂Ω. | (2.23) |
Furthermore θ↦ηθ is increasing and bounded from above by y1, hence it converges to a function η1 which satisfies (2.23). Then
∫Ω(|un−um||x|−θ+|g(un)−g(um)|ηθ)dγμ≤∫Ω|νn−νm|ηθdγμ. | (2.24) |
Letting θ→1, we obtain as a complement of (2.21) that
∫Ω(|un−um||x|+|g(un)−g(um)|η1)dγμ≤∫Ω|νn−νm|η1dγμ. | (2.25) |
Hence {un} is a Cauchy sequence in L1(Ω,|x|−1dγμ) with limit u and {g(un)} is a Cauchy sequence in L1(Ω,ρdγμ) with limit g(u). Then (2.18) holds. As for (1.31) it is a consequence of (2.19) and (1.32) is proved similarly.
In this section as well as in the next one we always assume that N≥3 and μ≥μ0, or N=2 and μ>0, since the case N=2, μ=0, which necessitates specific tools, has already been completely treated in [26].
We recall that the set M(Ω∗;Γμ) of Radon measures is defined in the introduction as the set of measures in Ω∗ satisfying (1.25), and any positive measure ν∈M(Ω∗;Γμ) is naturaly extended by formula (1.26) as a positive measure in Ω. The space ¯M(Ω;Γμ) is the space of measures ν on C0(Ω) such that
ν=ν⌊Ω∗+kδ0, | (3.1) |
where ν⌊Ω∗∈M(Ω∗;Γμ).
Lemma 3.1. If ν∈¯M(Ω;Γμ), then there exists a unique weak solution u∈L1(Ω,|x|−1dγμ) to
{Lμu=νinΩ,Lμu=0on∂Ω. | (3.2) |
This solution is denoted by Gμ[ν], and this defines the Green operator of Lμ in Ω with homogeneous Dirichlet conditions.
Proof.By linearity and using the result of [12] on fundamental solution, we can assume that k=0 and ν≥0. Let {νn}⊂L1(Ω,ρdγμ) be a sequence such that νn≥0 and
∫ΩξΓμνndx→∫Ωξd(Γμν)forallξ∈Xμ(Ω), |
and by Proposition 2.1, we may let un be the unique, nonnegative weak solution of
{Lμun=νninΩ,Lμun=0on∂Ω, | (3.3) |
with n∈N. There holds
∫ΩunL∗μξdγμ(x)=∫ΩξνnΓμdx forallξ∈Xμ(Ω). | (3.4) |
Then un≥0 and using the function η1 defined in the proof of Theorem A for test function, we have
c6∫Ωun|x|dγμ=∫Ωη1Γμνndx≤c7‖ν‖M(Ω,Γμ), | (3.5) |
which implies that {un} is bounded in L1(Ω,1|x|dγμ(x)).
For any ϵ>0 sufficiently small, set the test function ξ in {ζ∈Xμ(Ω):ζ=0 in Bϵ}, then we have that
∫Ω∖Bϵ(0)unL∗μξdγμ(x)=∫Ω∖Bϵ(0)ξνnΓμdx forallξ∈Xμ(Ω). | (3.6) |
Therefore, for any open sets O and O′ verifying ˉO⊂O′⊂ˉO′⊂Ω∖Bϵ(0), there exists c8>0 independent of n such that
‖un‖L1(O′)≤c8‖ν‖M(Ω,Γμ). |
Note that in Ω∖Bϵ, the operator L∗μ is uniformly elliptic and the measure dγμ is equivalent to the N-dimensional Lebesgue measure dx, then [30,Corollary 2.8] could be applied to obtain that for some c9,c10>0 independent of n but dependent of O′,
‖un‖W1,q(O)≤c9‖un‖L1(O′)+‖˜νn‖L1(Ω,dγμ)≤c10‖ν‖M(Ω,Γμ). |
That is, {un} is uniformly bounded in W1,qloc(Ω∖{0}).
As a consequence, since ϵ is arbitrary, there exist a subsequence, still denoted by {un}n and a function u such that
un→ua.e. in Ω. |
Meanwhile, we deduce from Fatou's lemma,
∫Ωu|x|dγμ≤c11∫Ωη1Γμdν. | (3.7) |
Next we claim that un→u in L1(Ω,|x|−1dγμ). Let ω⊂Ω be a Borel set and ψω be the solution of
{L∗μψω=|x|−1χωin Ω,L∗μψω=0on ∂Ω. | (3.8) |
Then ψω≤η1, thus it is uniformly bounded. Assuming that Ω⊂B1, clearly ψω is bounded from above by the solution Ψω of
{L∗μΨω=|x|−1χωinB1,L∗μΨω=0on∂B1, | (3.9) |
and by standard rearrangement, supB1Ψω≤supB1Ψrω, where Ψrω solves
{L∗μΨrω=|x|−1Bϵ(|ω|)inB1,L∗μΨrω=0on ∂B1, | (3.10) |
where ϵ(|ω|)=(|ω||B1)1N. Then Ψrω is radially decreasing and lim|ω|→0Ψrω=0, uniformly on B1. This implies
lim|ω|→0ψω(x)=0uniformly in B1. | (3.11) |
Using (3.4) with ξ=ψω,
∫ωun|x|dγμ(x)=∫ωνnΓμψωdx≤supΩψω∫ωνnΓμdx→0 as|ω|→0. |
Therefore {un} is uniformly integrable for the measure |x|−1dγμ. Letting n→∞ in (3.4) implies the claim.
We assume that g:R→R is a continuous nondecreasing function such that rg(r)≥0 for all r∈R. The next lemma dealing with problem
{Lμu+g(u)=kδ0in Ω,Lμ+g(u)u=0on ∂Ω, | (3.12) |
is an extension of [22,Theorem 3.1,Theorem 3.2]. Actually it was quoted without demonstration in this article as Remark 3.1 and Remark 3.2 and we give here their proof. Notice also that when N≥3 and μ=μ0 we give a more complete result than [22,Theorem 3.2].
Lemma 3.2. Let k∈R and g:R→R be a continuous nondecreasing function such that rg(r)≥0 for all r∈R. Then problem (3.12) admits a unique solution u:=ukδ0 if one of the following conditions is satisfied:
(i) N≥2, μ>μ0 and g satisfies (1.21);
(ii) N≥3, μ=μ0 and g satisfies (1.35).
Proof.Without loss of generality we assume BR⊂Ω⊂B1 for some R∈(0,1).
(ⅰ) The case μ>μ0. It follows from [22,Theorem 3.1] that for any k∈R there exists a radial function vk,1 (resp. vk,R) defined in B∗1 (resp. B∗R) satisfying
Lμv+g(v)=0inB∗1(resp.inB∗R), | (3.13) |
vanishing respectively on ∂B1 and ∂BR and satisfying
limx→0vk,1(x)Φμ(x)=limx→0vk,R(x)Φμ(x)=kcμ. | (3.14) |
Furthermore g(vk,1)∈L1(B1,dγμ) (resp. g(vk,R)∈L1(BR,dγμ)). Assume that k>0, then 0≤vk,R≤vk,1 in B∗R and the extension of ˜vk,R by 0 in Ω∗ is a subsolution of (3.13) in Ω∗ and it is still smaller than vk,1 in Ω∗. By the well known method on super and subsolutions (see e.g., [32,Theorem 1.4.6]), there exists a function u in Ω∗ satisfying ˜vk,R≤u≤vk,1 in Ω∗ and
{Lμu+g(u)=0in Ω∗,Lμ+g(u)u=0on ∂Ω,,limx→0u(x)Φμ(x)=kcμ. | (3.15) |
By standard methods in the study of isolated singularities (see e.g., [22,29,16,17] for various extensions)
limx→0|x|1−τ−∇u(x)=τ−kcμx|x|. | (3.16) |
For any ϵ>0 and ξ∈Xμ(Ω),
0=∫Ω∖Bϵ(Lμu+g(u))Γμξdx0=∫Ω∖BϵuL∗μξdγμ(x)+(τ–τ+)kcμ|SN−1|ξ(0)(1+o(1)). |
Using (1.15), we obtain
∫ΩuL∗μξdγμ(x)=kξ(0). | (3.17) |
(ⅱ) The case μ=μ0. In [22,Theorem 3.2] it is proved that if for some b>0 there holds
I:=∫∞1g(btN−2N+2lnt)t−2dt<∞, | (3.18) |
then there exists a solution of (1.22) satisfying (1.19) with γ=(N+2)b2. Actually we claim that the finiteness of this integral is independent of the value of b. To see that, set s=tN−2N+2, then
I=N+2N−2∫∞1g(βslns)s−2NN−2ds, |
with β=N+2N−2b. Set τ=βslns, then
lns(1+lnlnslns+lnβlns)⟹lns=lnτ(1+o(1))ass→∞. |
We infer that for ϵ>0 there exists sϵ>2 and τϵ=sϵlnsϵ such that
(1−ϵ)βN+2N−2≤∫∞sϵg(βslns)s−2NN−2ds∫∞τϵg(τ)(lnτ)N+2N−2τ−2NN−2dτ≤(1+ϵ)βN+2N−2, | (3.19) |
which implies the claim. Next we prove as in case (i) the existence of vk,1 (resp. vk,R) defined in B∗1 (resp. B∗R) satisfying
Lμ0v+g(v)=0inB∗1(resp.inB∗R), | (3.20) |
vanishing respectively on ∂B1 and ∂BR and satisfying
limx→0vk,1(x)Φμ(x)=limx→0vk,R(x)Φμ(x)=kcμ0. | (3.21) |
We end the proof as above.
Remark.It is important to notice that conditions (1.21) and (1.35) (or equivalently (1.23)) are also necessary for the existence of radial solutions in a ball, hence their are also necessary for the existence of non radial solutions of the Dirichlet problem (3.12).
We consider now the problem
{Lμu+g(u)=νin Ω,Lμ+g(u)u=0on ∂Ω, | (3.22) |
where ν∈M(Ω∗;Γμ).
Lemma 3.1. Let μ≥μ0. Assume that g satisfies (1.8) if N≥3 or the β±(g) defined by (1.9) satisfy β−(g)<0<β+(g) if N=2, and let ν∈M(Ω∗;Γμ). If N=2, we assume that ν can be decomposed as ν=νr+∑jαjδaj where νr has no atom, the αj satisfy (1.10) and {aj}⊂Ω∗. Then problem (3.22) admits a unique weak solution.
Proof.We assume first that ν≥0 and let r0=dist(x,∂Ω). For 0<σ<r0, we set Ωσ=Ω∖{¯Bσ} and νσ=νχΩσ and for 0<ϵ<σ we consider the following problem in Ωϵ
{Lμu+g(u)=νσin Ωϵ,Lμ+g(u)u=0on ∂Ω,Lμ+g(u)u=0on ∂Bϵ. | (3.23) |
Since 0∉Ωϵ problem (3.23) admits a unique solution uνσ,ϵ which is smaller than Gμ[ν] and satisfies
0≤uνσ,ϵ≤uνσ′,ϵ′inΩϵforall0<ϵ′≤ϵand0<σ′≤σ. |
For any ξ∈C1,1c(Ω∗) and ϵ small enough so that supp(ξ)⊂Ωϵ, there holds
∫Ω(uνσ,ϵL∗μξ+g(uνσ,ϵ)ξ)dγμ=∫ΩξΓμdνσ. |
There exists uνσ=limϵ→0uνσ,ϵ and it satisfies the identity
∫Ω(uνσL∗μξ+g(uνσ)ξ)dγμ=∫ΩξΓμdνσforallξ∈C1,1c(Ω∗). | (3.24) |
As a consequence of the maximum principle and Lemma 3.1, there holds
0≤uνσ≤Gμ[νσ]≤Gμ[ν]. | (3.25) |
Since νσ vanishes in Bσ, Gμ[νσ](x)≤c12Φμ(x) in a neighborhood of 0, and uνσ is also bounded by c12Φμ in this neighborhood. This implies that Φ−1μ(x)uνσ(x)→c′ as x→0 for some c′≥0. Next let ξ∈C1,1c(Ω),
ℓn(r)={2−1(1+cos(2π|x|σ))if |x|≤σ2,0if |x|>σ2, |
and ξn=ξℓn. Then
∫Ω(uνσL∗μξn+g(uνσ)ξn)dγμ=∫ΩξnΓμdνσ. | (3.26) |
When n→∞,
∫ΩξnΓμdνσ→∫ΩξΓμdνσ |
and
∫Ωg(uσ)ξndγμ→∫Ωg(uσ)ξdγμ. |
Now for the first inegral term in (3.26), we have
∫ΩuνσL∗μξndγμ=∫ΩℓnuσL∗μξdγμ+In+IIn+IIIn, |
where
In=−∫Bσ2uσξΔℓndγμ, |
IIn=−2∫Bσ2uσ⟨∇ξ,∇ℓn⟩dγμ |
and
IIIn=−τ+∫Bσ2uσ⟨x|x|2,∇ℓn⟩dγμ. |
Using the fact that ξ(x)→ξ(0) and ∇ξ(x)→∇ξ(0) we easily infer that In, IIn and IIIn converge to 0 when n→∞, the most complicated case being the case when μ=μ0, which is the justification of introducing the explicit cut-off function ℓn. Therefore (3.24) is still valid if it is assumed that ξ∈C1,1c(Ω). This means that uνσ is a weak solution of
{Lμu+g(u)=νσin Ω,Lμ+g(u)u=0on ∂Ω. | (3.27) |
Furthermore uνσ is unique and uνσ is a decreasing function of σ with limit u when σ→0. Taking η1 as test function, we have
∫Ω(c|x|−1uνσ+η1g(uνσ))dγμ=∫Ωη1d(γμνσ)≤∫Ωη1d(γμν). |
By using the monotone convergence theorem we infer that uνσ→u in L1(Ω,|x|−1dγμ) and g(uνσ)→g(uν) in L1(Ω,dγμ). Hence u=uν is the weak solution of (3.22).
Next we consider a signed measure ν=ν+−ν−. We denote by uνσ+,ϵ, u−νσ−,ϵ and uνσ,ϵ the solutions of (3.23) in Ωϵ corresponding to νσ+, −νσ− and νσ,ϵ respectively. Then
u−νσ−,ϵ≤uνσ,ϵ≤uνσ+,ϵ. | (3.28) |
The correspondence ϵ↦uνσ+,ϵ and ϵ↦u−νσ−,ϵ are respectively increasing and decreasing. Furthermore uνσ,ϵ is locally bounded, hence by local compactness and up to a subsequence uνσ,ϵ converges a.e. in Bϵ to some function uνσ. Since u−νσ−,ϵ→u−νσ− and uνσ+,ϵ→uνσ+ in L1(Ω,|x|−1dγμ), it follows by Vitali's theorem that uνσ,ϵ→uνσ in L1(Ω,|x|−1dγμ). Similarly, using the monotonicity of g, g(uνσ,ϵ)→g(uνσ) in L1(Ω,dγμ). By local compactness, uνσ→u a.e. in Ω. Using the same argument of uniform integrability, we have that uνσ→u in L1(Ω,|x|−1dγμ) and g(uνσ)→g(u) in L1(Ω,dγμ) when σ→0 and u satisfies
∫Ω(uL∗μξ+g(u)ξ)dγμ=∫Ωξd(dγμν)for any ξ∈C1,1c(Ω∗). | (3.29) |
Finally the singularity at 0 is removable by the same argument as above which implies that u solves (3.29) and thus u=uν is the weak solution of (3.22).
The idea is to glue altogether two solutions one with the Dirac mass and the other with the measure in Ω∗, this is the reason why the weak Δ2 condition is introduced.
Lemma 3.3. Let ν=ν⌊Ω∗+kδ0∈¯M+(Ω;Γμ) and σ>0. We assume that ν⌊Ω∗(¯Bσ)=0. Then there exists a unique weak solution to (1.6).
Proof.Set νσ=ν⌊Ω∗. It has support in Ωσ=Ω∖¯Bσ. For 0<ϵ<σ we consider the approximate problem in Ωϵ=Ω∖¯Bϵ,
{Lμu+g(u)=νσ in Ωϵ,Lμ+g(u)u=0 on ∂Ω,Lμ+g(u)u=ukδ0 on ∂Bϵ, | (3.30) |
where ukδ0 is the solution of problem (3.12) obtained in Lemma 3.2. It follows from [30,Theorem 3.7] that problem (3.30) admits a unique weak solution denoted by Uνσ,ϵ, thanks to the fact that the operator is not singular in Ωϵ. We recall that uνσ,ϵ is the solution of (3.23) and Gμ[δ0] the fundamental solution in Ω. Then
max{ukδ0,uνσ,ϵ}≤Uνσ,ϵ≤uνσ+kGμ[δ0]inΩϵ. | (3.31) |
Furthermore one has Uνσ,ϵ≤Uνσ,ϵ′ in Ωϵ, for 0<ϵ′<ϵ. Since uνσ≤uν and both kGμ[δ0] and uν belong to L1(Ω,|x|−1dγμ), then it follows by the monotone convergence theorem that Uνσ,ϵ converges in L1(Ω,|x|−1dγμ) and almost everywhere to some function Uνσ∈L1(Ω,|x|−1dγμ). Since Γμ is a supersolution for equation Lμu+g(u)=0 in Bσ, for 0<ϵ0<σ there exists c13:=c13(ϵ0,σ)>0 such that
uνσ(x)≤c13|x|τ+forallx∈Bϵ0. |
For any δ>0, there exists ϵ0 such that uνσ(x)≤δGμ[δ0](x) in Bϵ0. Hence uνσ+kGμ[δ0]≤(k+δ)Gμ[δ0] in Bϵ0, which implies
g(Uνσ,ϵ)≤g((k+δ)Gμ[δ0])inBϵ0∖¯Bϵ, | (3.32) |
and
∫Ωg((k+δ)Gμ[δ0])dγμ(x)≤∫B1g(k+δcμ|x|τ−)|x|τ+dx=|SN−1|∫10g(k+δcμrτ−)rτ++N−1dr∫Ωg((k+δ)Gμ[δ0])dγμ(x)=c14∫∞k+δcμg(t)t−2+2τ−=c14∫∞k+δcμg(t)t−1−p∗μdt−−−−−−−−−−−<∞. |
Now, using the local Δ2-condition, with a′=kcμϵτ−0, we see that
g(Uνσ,ϵ)≤g(uνσ+kcμϵτ−0)≤K(a′)(g(uνσ)+g(a′))inΩϵ0. | (3.33) |
From (3.32) and (3.33) we infer that g(Uνσ,ϵ) is bounded in L1(Ωϵ,dγμ) independently of ϵ. If ξ∈C1,10(Ω∗), we have for ϵ>0 small enough so that supp(ξ)⊂Ωϵ
∫Ω(Uνσ,ϵL∗μξ+g(Uνσ,ϵ)ξ)dγμ=∫ΩξΓμdνσ. |
Letting ϵ→0 we obtain that
∫Ω(UνσL∗μξ+g(Uνσ)ξ)dγμ=∫ΩξΓμdνσ. | (3.34) |
Let ξ∈C1,10(¯Ω) and ηn∈C1,1(RN) be a nonnegative cut-off function such that 0≤ηn≤1, ηn≡1 in Bc2n, ηn≡0 in B1n, and choose ξηn for test function. Then
∫Ω(ηnUνσL∗μξ+g(Uνσ)ηnξ)dγμ−∫ΩUνσAndγμ=∫ΩξηnΓμdνσ, | (3.35) |
with
An=ξΔηn+2⟨∇ηn,∇ξ⟩+2τ+ξ⟨∇ηn,x|x|2⟩. | (3.36) |
Clearly
limn→∞∫Ω(ηnUνσL∗μξ+g(Uνσ)ηnξ)dγμ=∫Ω(UνσL∗μξ+g(Uνσ)ξ)dγμ, |
and
limn→∞∫ΩξηnΓμdνσ=∫ΩξΓμdνσ. |
We take
ηn(r)={12−12cos(nπ(r−1n)) if1n≤r≤2n,0 ifr<1n,1 ifr>2n. |
Then
An=n2π22cos(nπ(r−1n))+nπ2N−1+2τ+rsin(nπ(r−1n)). |
Letting ϵ→0 in (3.31), we have
Uνσ(x)=kGμ[δ0](x)(1+o(1))=kcμ|x|τ−(1+o(1))asx→0. |
Hence
limn→∞∫ΩUνσAndγμ=2k|SN−1|√μ−μ0cμ=k. | (3.37) |
This implies that Uνσ is the solution of (1.6) with ν replaced by νσ+kδ0.
Lemma 3.4. Let ν=ν⌊Ω∗+kδ0∈¯M+(Ω;Γμ). Then there exists a unique weak solution to (1.6).
Proof. Following the notations of Lemma 3.3, we set νσ=χBσν⌊Ω∗ and denote by Uνσ the solution of
{Lμu+g(u)=νσ+kδ0in Ω,Lμ+g(u)u=0on ∂Ω. | (3.38) |
It is a positive function and there holds
max{ukδ0,uνσ}≤Uνσ≤uνσ+kGμ[δ0]inΩ. | (3.39) |
Since the mapping σ↦Uνσ is decreasing, then there exists U=limσ→0Uνσ and U satisfies (3.39). As a consequence Uνσ→U in L1(Ω,|x|−1dγμ) as σ→0. We take η1 for test function in the weak formulation of (3.39), then
∫Ω(|x|−1Uνσ+η1g(Uνσ))dγμ=∫Ωη1Γμdνσ+kη1(0). |
By the monotone convergence theorem we obtain the identity
∫Ω(|x|−1U+η1g(U))dγμ=∫Ωη1d(γμν⌊Ω∗)+kη1(0)=∫Ωη1d(γμν), |
and the fact that g(Uνσ)→g(U) in L1(Ω,ρdγμ). Going to the limit as σ→0 in the weak formulation of (3.38), we infer that U=uν is the solution of (1.6).
Proof of Theorem B. Assume ν=ν⌊Ω∗+kδ0∈¯M(Ω;Γμ) satisfies k>0 and let ν+=ν+⌊Ω∗+kδ0 and ν−=ν−⌊Ω∗ the positive and the negative part of ν. We denote by uν+ and u−ν− the weak solutions of (1.6) with respective data ν+ and −ν−. For 0<ϵ<σ such that ¯Bσ⊂Ω, we set νσ=χBσν⌊Ω∗, with positive and negative part νσ+ and νσ− and denote by Uνσ+,ϵ, U−νσ−,ϵ and Uνσ,ϵ the respective solutions of
{Lμu+g(u)=νσ+in Ωϵ,Lμ+g(u)u=0on ∂Ω,Lμ+g(u)u=ukδ0on ∂Bϵ, | (3.40) |
{Lμu+g(u)=−νσ− in Ωϵ,Lμ+g(u)u=0 on ∂Ω∪∂Bϵ, | (3.41) |
and
{Lμu+g(u)=νσin Ωϵ,Lμ+g(u)u=0on ∂Ω,Lμ+g(u)u=ukδ0on ∂Bϵ, | (3.42) |
then
U−νσ−,ϵ≤Uνσ,ϵ≤Uνσ+,ϵ. | (3.43) |
Furthermore Uνσ+,ϵ satisfies (3.31) and, in coherence with the notations of Lemma 3.1 with νσ replaced by −νσ−,
u−νσ−≤U−νσ−,ϵ=u−νσ−,ϵ. | (3.44) |
By compactness, {Uνσ,ϵj}ϵj converges almost everywhere in Ω to some function U for some sequence {ϵj} converging to 0. Moreover Uνσ,ϵj converges to Uνσ in L1(Ω,|x|−1dγμ) because Uνσ+,ϵ→uνσ++kδ0 and u−νσ−,ϵ→u−νσ− in L1(Ω,|x|−1dγμ) by Lemma 3.1 and (3.43) holds. Similarly g(Uνσ,ϵj) converges to g(U) in L1(Ω,ρdγμ). This implies that U satisfies
∫Ω(UL∗μξ+g(U)ξ)dγμ=∫ΩξΓμdνσfor all ξ∈C1,10(Ω∗). |
In order to use test functions in C1,10(¯Ω), we proceed as in the proof of Lemma 3.3, using the inequality (derived from (3.43)) and the
u−νσ−≤Uνσ≤uνσ++kδ0. | (3.45) |
By (3.33), uνσ++kδ0(x)=kGμ[δ0](x)(1+o(1)) when x→0 and u−νσ−=o(Gμ[δ0]) near 0. This implies Uνσ(x)=kGμ[δ0](x)(1+o(1)) as x→0 and we conclude as in the proof of Lemma 3.3 that u=uνσ+kδ0.
At end we let σ→0. Up to a sequence {σj} converging to 0 such that uνσj+kδ0→U almost everywhere and
u−ν−≤U≤uν++kδ0. | (3.46) |
Since by Lemma 3.4, uνσ++kδ0→uν++kδ0 in L1(Ω,|x|−1dγμ) and g(uνσ++kδ0)→g(uν++kδ0) in L1(Ω,ρdγμ), we infer that the convergences of uνσj+kδ0→U and g(uνσj+kδ0)→g(U) occur respectively in the same space, therefore U=uν+kδ0, it is the weak solution of (1.6).
Remark.In the course of the proof we have used the following result which is independent of any assumption on g except for the monotonicity: If {νn}⊂¯M+(Ω;Γμ) is an increasing sequence of g-good measures converging to a measure ν∈¯M+(Ω;Γμ), then ν is a g-good measure, {uνn} converges to uν in L1(Ω,|x|−1dγμ) and {g(uνn)} converges to g(uν) in L1(Ω,ρdγμ).
The construction of a solution is essentially similar to the one of Theorem B, the only modifications lies in Lemma 3.3. Estimate (3.31) remains valid with
ukδ0(x)=k|SN−1||x|2−N2ln|x|−1(1+o(1))=kGμ[δ0](x)(1+o(1))asx→0. | (3.47) |
Since uνσ(x)≤c15|x|2−N2, (3.32) holds with δ>0 arbitrarily small. Next
∫Ωg((k+δ)Gμ[δ0])dγμ(x)≤∫B1g(k+δ|SN−1||x|2−N2ln|x|−1)|x|2−N2dx∫Ωg((k+δ)Gμ[δ0])dγμ(x)=|SN−1|∫10g(k+δ|SN−1|r2−N2lnr−1)rN2dr∫Ωg((k+δ)Gμ[δ0])dγμ(x)=c16∫∞c′g(tlnt)t−2NN−2<∞, |
by (3.19) and (1.35). The end of the proof for Theorem C is similar to the one of Theorem B.
Proof of Corollary D. If g(r)=gp(r)=|r|p−1r, p>1, the existence of a solution with ν=kδ0 is a direct consequence of conditions (1.34) and (1.35). If k=0 and ν⌊Ω∗≠0, the existence is ensured if (1.8) holds, hence p<NN−2. Assertion (ⅲ) follows.
The notion of reduced measures introduced by Brezis, Marcus and Ponce [8] turned out to be a useful tool in the construction of solutions in a measure framework. We will develop only the aspect needed for the proof of Theorem E. If k∈N∗, we set
gk(r)={min{g(r),g(k)}ifr≥0,max{g(r),g(−k)}ifr>0. | (4.1) |
Since gk satisfies (1.34) and (1.35), for any ν∈¯M+(Ω;Γμ) there exists a unique weak solution u=uν,k of
{Lμu+gk(u)=νin Ω,Lμ+gk(u)u=0on ∂Ω. | (4.2) |
Furthermore, from the proof of Lemma 3.4 and Kato's type estimates Proposition 2.1 we have that
0≤uν+,k′≤uν+,k forall k′≥k>0. | (4.3) |
Proposition 4.1. Let ν∈¯M+(Ω;Γμ). Then the sequence of weak solutions {uν,k} of
{Lμu+gk(u)=νin Ω,Lμ+gk(u)u=0on ∂Ω, | (4.4) |
decreases and converges, when k→∞, to some nonnegative function u, and there exists a measure ν∗∈¯M+(Ω;Γμ) such that 0≤ν∗≤ν and u=uν∗.
Proof.The proof is similar to the one of [8,Prop. 4.1]. Observe that uν,k↓u∗ and the sequence {uν,k} is uniformly integrable in L1(Ω,|x|−1dγμ). By Fatou's lemma u satisfies
∫Ω(u∗L∗μξ+g(u∗)ξ)dγμ(x)≤∫Ωξd(Γμν)forallξ∈Xμ(Ω),ξ≥0. | (4.5) |
Hence u∗ is a subsolution of (1.6) and by construction it is the largest of all nonnegative subsolutions. The mapping
ξ↦∫Ω(u∗L∗μξ+g(u∗)ξ)dγμ(x)forallξ∈C∞c(Ω), |
is a positive distribution, hence a measure ν∗, called the reduced measure of ν. It satisfies 0≤ν∗≤ν and u∗=uν∗.
Lemma 4.2. Let ν,ν′∈¯M+(Ω;Γμ). If ν′≤ν and ν=ν∗, then ν′=ν′∗.
Proof. Let uν′,k be the weak solution of the truncated equation
{Lμu+gk(u)=ν′in Ω,Lμ+gk(u)u=0on ∂Ω. | (4.6) |
Then 0≤uν′,k≤uν,k. By Proposition 4.1, we know that uν,k↓uν∗=uν and uν′,k↓u′∗ a.e. in L1(Ω,|x|−1dγμ) and then
Lμ(uν,k−uν)+gk(uν,k)−gk(uν)=g(uν)−gk(uν), |
from what follows, by Proposition 2.1,
∫Ω(uν,k−uν))|x|−1dγμ+∫Ω|gk(uν,k)−gk(uν)|η1dγμ≤∫Ω|g(uν)−gk(uν)|η1dγμ. |
By the increasing monotonicity of mapping k↦gk(uν), we have gk(uν)→g(uν) in L1(Ω,ρdγμ) as k→+∞, hence
∫Ω|gk(uν,k)−g(uν)|η1dγμ≤2∫Ω|g(uν)−gk(uν)|η1dγμ→0asn→∞. |
Because gk(uν′,k)≤gk(uν,k) it follows by Vitali's convergence theorem that gk(uν′,k)→g(u′∗) in L1(Ω,ρdγμ). Using the weak formulation of (4.6), we infer that u′∗ verifies
∫Ω(u′∗L∗μξ+g(u′∗)ξ)dγμ=∫Ωξd(γμν′) for all ξ∈Xμ(Ω). |
This yields u′∗=uν′.
The next result follows from Lemma 4.2.
Lemma 4.3. Assume that ν=ν⌊Ω∗+kδ0∈¯M+(Ω;Γμ), then ν∗=ν∗⌊Ω∗+k∗δ0∈¯M+(Ω;Γμ) with ν∗⌊Ω∗≤ν⌊Ω∗ and k∗≤k. More precisely,
(i) If μ>μ0 and g satisfies (1.34), then k=k∗.
(ii) If μ=μ0 and g satisfies (1.35), then k=k∗.
(ii) If μ>μ0 (resp. μ=μ0) and g does not satisfy (1.21) (resp. (1.35)), then k∗=0.
The next result is useful in applications.
Corollary 4.1. If ν∈¯M+(Ω;Γμ), then ν∗ is the largest g-good measure smaller or equal to ν.
Proof. Let λ∈¯M+(Ω;Γμ) be a g-good measure, λ≤ν. Then λ∗=λ≤ν∗. Since ν∗ is a g-good measure the result follows.
Proof of Theorem E. Assume that ν≥0. By Lemma 4.2 and Remark at the end of Section 3.5 the following assertions are equivalent:
(ⅰ) ν is gp-good.
(ⅱ) For any σ>0, νσ=χBcσν is gp-good.
If νσ is good, then uνσ satisfies
−Δuνσ+upνσ=νσ−μ|x|2uνσin D′(Ω∗) | (4.7) |
and since uνσ(x)≤c|x|τ+ if |x|≤σ2 (4.7) holds in D′(Ω). This implies that u∈Lp(Ω) and |x|−2uνσ∈Lα(Bσ2) for any α<N(2−τ+)+. Using [1] the measure νσ is absolutely continuous with respect to the c2,p′-Bessel capacity. If E⊂Ω is a Borel set such that c2,p′(E)=0, then c2,p′(E∩Bcσ)=0, hence ν(E∩Bcσ)=νσ(E∩Bcσ)=0. By the monotone convergence theorem ν(E)=0.
Conversely, if ν is nonnegative and absolutely continuous with respect to the c2,p′-Bessel capacity, then so is νσ=χBcσν. For 0≤ϵ≤σ2 we consider the problem
{−Δu+μ|x|2u+up=νσ in Ωϵ:=Ω∖Bϵ,−−−−−−−−u=0 on ∂Bϵ,−−−−−−−−u=0 on ∂Ω. | (4.8) |
Since μ|x|2 is bounded in Ωϵ and νσ is absolutely continuous with respect to the c2,p′ capacity there exists a solution uνσ,ϵ thanks to [1], unique by monotonicity. Now the mapping ϵ↦uνσ,ϵ is decreasing. We use the method developed in Lemma 3.1, when ϵ→0, we know that uνσ,ϵ increase to some uσ which is dominated by G[νσ] and satisfies
{−Δu+μ|x|2u+up=νσ in Ω∗,−−−−−−−−u=0 on ∂Ω. | (4.9) |
Because uσ≤G[νσ] and νσ=0 in Bσ, there holds u(x)≤c′11Γμ(x) in Bσ2, and then uσ is a solution in Ω and u=uνσ. Letting σ→0, we conclude as in Lemma 3.1 that uνσ converges to uν which is the weak solution of
{−Δu+μ|x|2u+up=ν in Ω,−−−−−−−−u=0 on ∂Ω. | (4.10) |
If ν is a signed measure absolutely continuous with respect to the c2,p′-capacity, so are ν+ and ν−. Hence there exists solutions uν+ and uν−. For 0<ϵ<σ2 we construct uνσ,ϵ with the property that −u−ν−σ,ϵ≤uνσ,ϵ≤uν+σ,ϵ, we let ϵ→0 and deduce the existence of uνσ which is eventually the weak solution of
{−Δu+μ|x|2u+|u|p−1u=νσ in Ω∗,−−−−−−−−−−u=0 on ∂Ω, | (4.11) |
and satisfies −u−ν−σ≤uνσ≤uν+σ. Letting σ→0 we obtain that u=limσ→0uνσ satisfies
{−Δu+μ|x|2u+|u|p−1u=ν in Ω∗,−−−−−−−−−−u=0 on ∂Ω. | (4.12) |
Hence u=uν and ν is a good solution.
Proof of Theorem F. Part 1. Without loss of generality we can assume that Ω is a bounded smooth domain. Let K⊂Ω be compact. If 0∈K and p<p∗μ there exists a solution ukδ0, hence K is not removable. If 0∉K and c2,p′(K)>0, there exists a capacitary measure νK∈W−2,p(Ω)∩M+(Ω) with support in K. This measure is gp-good by Theorem E, hence K is not removable.
Part 2. Conversely we first assume that 0∉K. Then there exists a subdomain D⊂Ω such that 0∉ˉD and K⊂D. Hence a solution u of (1.37) is also a solution of
−Δu+μ|x|2u+|u|p−1u=0 in D∖K, |
and the coefficient μ|x|2 is uniformly bounded in ˉD. By [1,Theorem 3.1] it can be extended as a C2 solution of the same equation in Ω′. Hence, if c2,p′(K)=0 the set K is removable.
If 0∈K we have to assume at least p≥p∗μ in order that 0 is removable and p≥p0 in order there exists non-empty set with zero c2,p′-capacity. Let ζ∈C1,10(Ω) with 0≤ζ≤1, vanishing in a compact neighborhood D of K. Since 0∉Ω∖D, we first consider the case where u is nonnegative and satisfies in the usual sense
−Δu+μ|x|2u+up=0inΩ∖D. |
Taking ζ2p′ for test function, we get
−2p′∫Ωuζ2p′−1Δζdx−2p′(2p′−1)∫Ωuζ2p′−2|∇ζ|2dx+μ∫Ωuζ2p′|x|2dx+∫Ωζ2p′updx=0. |
There holds
|∫Ωuζ2p′−1Δζdx|≤(∫Ωζ2p′updx)1p(∫Ω|Δζ|p′ζp′dx)1p′, |
0≤∫Ωuζ2p′−2|∇ζ|2dx≤(∫Ωζ2p′updx)1p(∫Ω|∇ζ|2p′dx)1p′, |
and
0≤∫Ωuζ2p′|x|2dx≤(∫Ωζ2p′updx)1p(∫Ωζ2p′|x|2p′dx)1p′. |
By standard elliptic equations regularity estimates and Gagliardo-Nirenberg inequality [21] (and since 0≤ζ≤1),
(∫Ω|Δζ|p′ζp′)1p′≤c17‖ζ‖W2,p′ |
and
(∫Ω|∇ζ|2p′dx)1p′≤c18‖ζ‖W2,p′. |
Finally, if p>p0:=NN−2, then 2p′<N which implies that there exists c19 independent of ζ (with value in [0,1]) such that
(∫Ωζ2p′|x|2p′dx)1p′≤(∫B1dx|x|2p′)1p′:=c19. |
Next we set
X=(∫Ωζ2p′updx)1p, |
and if μ≥0, p≥p0, we have
Xp−(2p′(2p′−1)c18−p′c18)‖ζ‖W2,p′X≤0; | (4.13) |
and if μ<0, p>p0, we have
Xp−((2p′(2p′−1)c18−p′c18)‖ζ‖W2,p′−c19μ)X≤0. | (4.14) |
However, the condition p>p0 is ensured when μ<0 since p≥p∗μ>p0. We consider a sequence {ηn}⊂S(RN) such that 0≤ηn≤1, ηn=0 on a neighborhood of K and such that ‖ηn‖W2,p′→0 when n→∞. Such a sequence exists by the result in [24] since c2,p′(K)=0. Let ξ∈C∞0(Ω) such that 0≤ξ≤1 and with value 1 in a neighborhood of K. We take ζ:=ζn=(1−ηn)ξ in the above estimates. Letting n→∞, then ζn→ξ in W2,p′ and finally
Xp−1=(∫Ωξ2p′updx)p−1p≤(2p′(2p′−1)c18−p′c18)‖ξ‖W2,p′+c19μ−, | (4.15) |
under the condition that p>p0 if μ<0, in which case there also holds
∫Ωuζ2p′|x|2dx≤c19X. | (4.16) |
However the condition p>p0 is not necessary in order the left-hand side of (4.16) is bounded, since we have
μ∫Ωuζ2p′|x|2dx+Xp≤(2p′(2p′−1)c18−p′c18)‖ζ‖W2,p′X, | (4.17) |
and X is bounded.
Next we take ζ:=ζn=(1−ηn)ξ for test function in (1.37) and get
−∫Ω((1−ηn)Δξ−ξΔηn−2⟨∇ηn,∇ξ⟩)udx+μ∫Ωuζn|x|2dx+∫Ωζnupdx=0. |
Since
∫ΩuξΔηndx≤(∫Ωupξdx)1p‖ηn‖W2,p′→0asn→∞, |
and
|∫Ωu⟨∇ηn,∇ξ⟩dx|≤(∫Ωup|∇ξ|dx)1p‖∇ξ‖L∞‖ηn‖W1,p′asn→∞, |
then we conclude that u satisfies
−∫ΩuΔξdx+μ∫Ωuξ|x|2dx+∫Ωξupdx=0, | (4.18) |
which proves that u satisfies the equation in the sense of distributions. By standard regularity u is C2 in Ω∗, and by the maximum principle u(x)≤c20Γμ(x) in Br0⊂Ω. Integrating by part as in the proof of Lemma 3.2 we obtain that u satisfies
∫Ω(uL∗μξ+ξup)dγμ(x)=0for every ξ∈Xμ(Ω). | (4.19) |
Finally, if u is a signed solution, then |u| is a subsolution. For ϵ>0 we set Kϵ={x∈RN:dist(x,K)≤ϵ}. If ϵ is small enough Kϵ⊂Ω. Let v:=vϵ be the solution of
{−Δv+μ|x|2v+vp=0in Ω∖Kϵ,−−−−−−− v=|u|⌊∂Kϵon ∂Kϵ,−−−−−−− v=|u|⌊∂Ωon ∂Ω. | (4.20) |
Then |u|≤vϵ. Furthermore, by Keller-Osserman estimate as in [22,Lemma 1.1], there holds
vϵ(x)≤c21dist(x,Kϵ)−2p−1forallx∈Ω∖Kϵ, | (4.21) |
where c21>0 depends on N, p and μ. Using local regularity theory and the Arzela-Ascoli Theorem, there exists a sequence {ϵn} converging to 0 and a function v∈C2(Ω∖K)∩C(ˉΩ∖K) such that {vϵn} converges to v locally uniformly in ˉΩ∖K and in the C2loc(Ω∖K)-topology. This implies that v is a positive solution of (1.37) in Ω∖K. Hence it is a solution in Ω. This implies that u∈Lp(Ω) and |u(x)|≤v(x)≤c20Γμ(x) in Ω∗. We conclude as in the nonnegative case that u is a weak solution in Ω.
H. Chen is supported by NSF of China, No: 11726614, 11661045, by the Jiangxi Provincial Natural Science Foundation, No: 20161ACB20007, by Doctoral Research Foundation of Jiangxi Normal University, and by the Alexander von Humboldt Foundation.
The authors declare no conflict of interest.
[1] |
Baras P, Pierre M (1984) Singularité séliminables pour des équations semi linéaires. Ann Inst Fourier Grenoble 34: 185–206. doi: 10.5802/aif.956
![]() |
[2] | Bénilan P, Brezis H (2003) Nonlinear problems related to the Thomas-Fermi equation. J Evol Equ 3: 673–770. |
[3] |
Bidaut-Véron M, Véron L (1991) Nonlinear elliptic equations on complete Riemannian manifolds and asymptotics of Emden equations. Invent Math 106: 489–539. doi: 10.1007/BF01243922
![]() |
[4] |
Birnbaum Z, Orlicz W (1931) über die Verallgemeinerung des Begriffes der zueinander Konjugierten Potenzen. Stud Math 3: 1–67. doi: 10.4064/sm-3-1-1-67
![]() |
[5] |
Boccardo L, Orsina L, Peral I (2006) A remark on existence and optimal summability of solutions of elliptic problems involving Hardy potential. Discrete Cont Dyn-A 16: 513–523. doi: 10.3934/dcds.2006.16.513
![]() |
[6] | Brezis H (1980) Some variational problems of the Thomas-Fermi type. Variational inequalities and complementarity problems. Proc Internat School, Erice, Wiley, Chichester 53–73. |
[7] |
Brezis H, Dupaigne L, Tesei A (2005) On a semilinear elliptic equation with inverse-square potential. Sel Math 11: 1–7. doi: 10.1007/s00029-005-0003-z
![]() |
[8] | Brezis H, Marcus M, Ponce A (2007) Nonlinear elliptic equations with measures revisited. Ann Math Stud 163: 55–109. |
[9] | Brezis H, Vázquez J (1997) Blow-up solutions of some nonlinear elliptic problems. Rev Mat Complut 10: 443–469. |
[10] |
Brezis H, Véron L (1980) Removable singularities of some nonlinear elliptic equations. Arch Ration Mech Anal 75: 1–6. doi: 10.1007/BF00284616
![]() |
[11] |
Chen H, Véron L (2014) Semilinear fractional elliptic equations with gradient nonlinearity involving measures. J Funct Anal 266: 5467–5492. doi: 10.1016/j.jfa.2013.11.009
![]() |
[12] | Chen H, Quaas A, Zhou F (2017) On nonhomogeneous elliptic equations with the Hardy-Leray potentials. arXiv:1705.08047. |
[13] | Chen H, Zhou F (2018) Isolated singularities for elliptic equations with inverse square potential and source nonlinearity. Discrete Cont Dyn-A 38: 2983–3002. |
[14] | Chen H, Alhomedan S, Hajaiej H, et al. (2017) Fundamental solutions for Schrödinger operators with general inverse square potentials. Appl Anal 1–24. |
[15] | Cignoli R, Cottlar M (1974) An Introduction to Functional Analysis. Amsterdam: North-Holland. |
[16] | Cîrstea F (2014) A complete classification of the isolated singularities for nonlinear elliptic equations with inverse square potentials.Mem Am Math Soc 227: No. 1068. |
[17] |
Cîrstea F, Du Y (2007) Asymptotic behavior of solutions of semilinear elliptic equations near an isolated singularity. J Funct Anal 250: 317–346. doi: 10.1016/j.jfa.2007.05.005
![]() |
[18] |
Dupaigne L (2002) A nonlinear elliptic PDE with the inverse square potential. J Anal Math 86: 359-398. doi: 10.1007/BF02786656
![]() |
[19] |
Folland G, Sitaram A (1997) The uncertainty principle: A mathematical survey. J Fourier Anal Appl 3: 207–238. doi: 10.1007/BF02649110
![]() |
[20] | Frank R (2011) Sobolev inequalities and uncertainty principles in mathematical physics. part I, unpublished notes of a course given at the LMU, Munich. Available From: http://www.math.caltech.edu/ rlfrank/sobweb1.pdf. |
[21] | Gilbarg D, Trudinger N (1983) Elliptic Partial Differential Equations of Second Order. Springer-Verlag, 224. |
[22] | Guerch B, Véron L (1991) Local properties of stationary solutions of some nonlinear singular Schrödinger equations. Rev Mat Iberoamericana 7: 65–114. |
[23] | Krasnosel'skii M, Rutickii Y (1961) Convex Functions and Orlicz Spaces. P. Noordhoff, Groningen. |
[24] |
Meyers N (1970) A theory of capacities for potentials of functions in Lebesgue classes. Math Scand 26: 255–292. doi: 10.7146/math.scand.a-10981
![]() |
[25] | Triebel H (1978) Interpolation Theory, Function Spaces, Differential Operators. North-Holland Pub Co. |
[26] |
Vàzquez J (1983) On a semilinear equation in RN involving bounded measures. Proc R Soc Edinburgh Sect A: Math 95: 181–202. doi: 10.1017/S0308210500012907
![]() |
[27] |
Véron L (1981) Singular solutions of some nonlinear elliptic equations. Nonlinear Anal Theory Methods Appl 5: 225–242. doi: 10.1016/0362-546X(81)90028-6
![]() |
[28] | Véron L (1986) Weak and strong singularities of nonlinear ellptic equations. Proc Symp Pure Math 45: 477–495. |
[29] | Véron L (1996) Singularities of Solutions of Second Order Quasilinear Equations. Pitman Research Notes in Mathematics. Series, 353. |
[30] | Véron L (2004) Elliptic equations involving measures, In: Chipot, M., Quittner, P., Editors,Handbook of Differential Equations: Stationary Partial Differential equations. Amsterdam: North-Holland, Vol I, 593–712. |
[31] | Véron L (2013) Existence and stability of solutions of general semilinear elliptic equations with measure data. Adv Nonlinear Stud 13: 447–460. |
[32] | Véron L (2017) Local and Global Aspects of Quasilinear Degenerate Elliptic Equations. Quasilinear Elliptic Singular Problems. World Scientific Publishing Co Pte Ltd, Hackensack, NJ, 457. |
1. | Huyuan Chen, Laurent Véron, Schrödinger operators with Leray-Hardy potential singular on the boundary, 2020, 269, 00220396, 2091, 10.1016/j.jde.2020.01.029 | |
2. | Ying Wang, Qingping Yin, On global estimates for Poisson problems with critical singular potentials, 2021, 210, 0362546X, 112372, 10.1016/j.na.2021.112372 | |
3. | Konstantinos T. Gkikas, Phuoc-Tai Nguyen, Martin kernel of Schrödinger operators with singular potentials and applications to B.V.P. for linear elliptic equations, 2022, 61, 0944-2669, 10.1007/s00526-021-02102-6 | |
4. | Huyuan Chen, Laurent Véron, Boundary singularities of semilinear elliptic equations with Leray-Hardy potential, 2022, 24, 0219-1997, 10.1142/S0219199721500516 | |
5. | Huyuan Chen, Hichem Hajaiej, Global $$W^{1,p}$$ regularity for elliptic problem with measure source and Leray–Hardy potential, 2022, 0035-5038, 10.1007/s11587-021-00615-y | |
6. | Huyuan Chen, Tobias Weth, The Poisson problem for the fractional Hardy operator: Distributional identities and singular solutions, 2021, 0002-9947, 10.1090/tran/8443 | |
7. | Huyuan Chen, Konstantinos T Gkikas, Phuoc-Tai Nguyen, Poisson problems involving fractional Hardy operators and measures, 2023, 36, 0951-7715, 7191, 10.1088/1361-6544/ad073e | |
8. | Huyuan Chen, Xiaowei Chen, Dirichlet problems involving the Hardy-Leray operators with multiple polars, 2023, 12, 2191-950X, 10.1515/anona-2022-0320 | |
9. | G. Barbatis, K. T. Gkikas, A. Tertikas, Heat and Martin Kernel estimates for Schrödinger operators with critical Hardy potentials, 2024, 389, 0025-5831, 2123, 10.1007/s00208-023-02693-9 | |
10. | Konstantinos T. Gkikas, Phuoc‐Tai Nguyen, Semilinear elliptic Schrödinger equations with singular potentials and absorption terms, 2024, 109, 0024-6107, 10.1112/jlms.12844 | |
11. | Mousomi Bhakta, Moshe Marcus, Phuoc-Tai Nguyen, Boundary value problems for semilinear Schrödinger equations with singular potentials and measure data, 2024, 390, 0025-5831, 351, 10.1007/s00208-023-02764-x | |
12. | Mohamed Jleli, Bessem Samet, Calogero Vetro, Higher order evolution inequalities involving Leray–Hardy potential singular on the boundary, 2024, 136, 18758576, 181, 10.3233/ASY-231873 | |
13. | Areej Bin Sultan, Mohamed Jleli, Bessem Samet, A system of wave inequalities with inverse-square potentials in an exterior domain, 2024, 103, 0003-6811, 843, 10.1080/00036811.2023.2210154 | |
14. | Mohamed Jleli, Bessem Samet, On a one-dimensional time-fractionally damped wave equation with a singular potential, 2024, 99, 0031-8949, 075244, 10.1088/1402-4896/ad5656 |