
Let Ω be a smooth bounded domain of
{−Δu−uΔ(u2)=a(x)1u in Ω,u>0 in Ω,u=0 on ∂Ω, | (1) |
where a(x)>0 a.e. in Ω and a∈Lp(Ω) with p>2. This type of equations is closely related to the standing wave solutions of the following quasilinear Schrodinger equation
i∂tψ=−Δψ+V(x)ψ−h(x,|ψ|2)ψ−κΔ[ρ(|ψ|2)]ρ′(|ψ|2)ψ, | (2) |
which has wide applications to physical models, suchas the superfluid film equation in plasma physics when ρ(s)=s[1], the self-channelling of a high-power ultra short laser in matter when
For the quasilinear singular equation involving a singular function f
{−Δu−uΔ(u2)=f(x,u) in Ω,u>0 in Ω,u=0 on ∂Ω, | (3) |
there are only a few results available about the existence of solutions by now. In Ref. [10], Marcos do Ó and Moameni considered the existence of radially symmetric solutions when Ω is a ball centered at the origin and the nonlinearity f(u)=λu3-u-u-γ with -γ∈(-1, 0). In Ref. [11], Santos, Yang and Zhou studied the case f(x, u)=λa(x)u-γ+b(x)uβ with -γ∈(-1, 0), β∈(1, 2·2*-1) and obtained the existence and multiplicity of solutions. For strongly singular quasilinear equations, in Ref. [12] Alves and Reis used the techniques developed in Ref. [13] by the second author for pure strongly singular equations and established the sufficient and necessary condition for the existence of solutions in the case when f(x, u)=h(x)u-γ+g(x, u) with -γ < -1. When -γ=-1, Alves[14] and Liu et al.[15] established the existence results for all -γ < 0 and f(x, u)=a(x)u-γ+λup with coefficient function a(x) satisfying: for each -γ < 0 there exists 0≤φ0∈C(Ω) and p>N such that a(x)φ0-γ∈Lp(Ω), which is used to construct the upper-lower solutions to overcome the singularity. In this paper we study the case -γ=-1 and face it in the quasilinear singular problem (1) for general a(x)∈Lp(Ω) with p>2.
1 The main resultTheorem 1.1 Let
By solutions of (1) we mean here solutions in H01(Ω), i.e. u∈H01(Ω) satisfying u(x)>0 in Ω and for all ψ∈H01(Ω),
∫Ω∇u∇ψ+2u2∇u∇ψ+2uψ|∇u|2−a(x)1uψ dx=0. |
We consider the following functional on H01(Ω):
J(u)=12∫Ω(1+2u2)|∇u|2 dx−∫Ωa(x)ln(u)dx. |
In this situation, one must find the integrability of both
{h′(t)=1√1+2h2(t),t⩾0,h(t)=−h(−t),t⩽0. |
(c.f. Ref. [16]) and the change, defined by ω(x)=h-1(u(x)), ∀x∈Ω, make singular terms more delicate. We give a way to deal with the singular functional after the function transform and establish the existence of solutions for more general coefficient functions of singular nonlinearities.
Notation. In the paper we make use of the following notation:
c denotes (possibly different) constants;
We denote by
First, we collect some properties of the function h(t) defined by the solution of the nonlinear equation:
\begin{cases}h^{\prime}(t)=\frac{1}{\sqrt{1+2 h^2(t)}}, & t \geqslant 0, \\ h(t)=-h(-t), & t \leqslant 0 .\end{cases} |
Lemma 2.1 (c.f. Refs. [5, 16-18]) Let h be defined as above. Then h has the following properties:
1) h″(t)=-2h(t)(h′(t))4, t>0;
2) h is unique, invertible and
3) 0≤h′(t)≤1,
4) |h(t)|≤|t| and
5)
6)
7)
8)
9) h(t)h′(t)/t, t>0 is decreasing,
10) h(t)-αh′(t)t is strictly decreasing on (0, +∞) with α≥2.
Proof We prove the point 10) of this lemma. The properties 1)-9) have been proved in Refs. [5, 16-18]. It is easily checked, thanks to point 5) of this lemma, that
t \geqslant h\left(1+2 h^2\right)^{\frac{1}{2}} \frac{1}{\alpha} \geqslant h \frac{\left(1+2 h^2\right)^{\frac{3}{2}}}{\left(1+\left(2+\frac{2}{\alpha}\right) h^2\right)} \frac{1}{\alpha}, |
since α≥2. Then it follows that
\begin{gathered} \frac{\mathrm{d}}{\mathrm{~d} t}\left[h(t)^{-\alpha} h^{\prime}(t) t\right] \\ =\frac{h\left(1+2 h^2\right)^{\frac{3}{2}}-\left(\alpha+(2 \alpha+2) h^2\right) t}{h^{\alpha+1}\left(1+2 h^2\right)^2}<0 .□ \end{gathered} |
By the change of function ω(x)=h-1(u(x)), we see that if ω∈H01(Ω) is a solution of
\left\{\begin{array}{l} -\Delta \omega=a(x) \frac{1}{h(\omega)} h^{\prime}(\omega) \text { in } \varOmega, \\ \omega>0 \text { in } \varOmega, \omega=0 \text { on } \partial \varOmega, \end{array}\right. | (4) |
namely,
\int_{\varOmega} \nabla \omega \nabla \psi-a(x) \frac{h^{\prime}(\omega)}{h(\omega)} \psi \mathrm{d} x=0, \quad \forall \psi \in H_0^1(\varOmega) |
then h(ω)∈H01(Ω) is the solution of (1). Now we write that
I(\omega)=\frac{1}{2} \int_{\varOmega}|\nabla \omega|^2 \mathrm{~d} x-\int_{\varOmega} a(x) \ln h(\omega(x)) \mathrm{d} x . |
Note that, after the transformation of u(x)∈H01(Ω) into h-1(u(x))=ω(x)∈H01(Ω), the difficulty for the integrability of
\begin{aligned} & X=\left\{\omega \in H_0^1(\varOmega): \omega>0 \text { a. e. in } \varOmega, \right. \\ & \left.a(x) \ln \omega(x) \in L^1(\varOmega)\right\}, \\ & \mathcal{N}=\left\{\omega \in X:\|\omega\|^2-\int_{\varOmega} a(x) \frac{h^{\prime}(\omega)}{h(\omega)} \omega \mathrm{d} x=0\right\} . \end{aligned} |
Indeed, since h(0)=0 and h(t) is strictly increasing on [0, +∞), thanks to properties 4) and 6) of Lemma 2.1, we get
\begin{aligned} \mid & \int_{A_{+}} a(x) \ln h(\omega(x)) \mathrm{d} x\left|\leqslant \int_{A_{+}} a(x)\right| \ln h(\omega(x)) \mid \mathrm{d} x \\ = & \int_{A_{+}} a(x) \ln h(\omega(x)) \mathrm{d} x \leqslant \int_{A_{+}} a(x) \ln \omega(x) \mathrm{d} x<+\infty \\ & \left|\int_{A_{-}} a(x) \ln h(\omega(x)) \mathrm{d} x\right| \leqslant \int_{A_{-}} a(x)|\ln h(\omega(x))| \mathrm{d} x \\ = & \int_{A_{-}} a(x) \ln \left(\frac{1}{h(\omega(x))}\right) \mathrm{d} x \leqslant \\ & \int_{A_{-} \cap[\omega<1]} a(x)\left[\ln \left(\frac{1}{h(1)}\right)+\ln \left(\frac{1}{\omega(x)}\right)\right] \mathrm{d} x+ \\ & \int_{A_{-} \cap[\omega \geqslant 1]} a(x)\left[\ln \left(\frac{1}{h(1)}\right)+\frac{1}{2} \ln \left(\frac{1}{\omega(x)}\right)\right] \mathrm{d} x \\ \leqslant & 2|\ln h(1)| \int_{\varOmega} a(x) \mathrm{d} x+\frac{3}{2} \int_{\varOmega} a(x)|\ln \omega(x)| \mathrm{d} x<+\infty \end{aligned} |
where
\begin{aligned} & A_{+}=\{x \in \varOmega: \ln h(\omega(x))>0\}, \\ & A_{-}=\{x \in \varOmega: \ln h(\omega(x)) \leqslant 0\}, \end{aligned} |
which clearly implies that the singular functional I is well defined on X. In addition, for any t>0 and ω∈X,
I(t \omega)=t^2 \frac{1}{2}\|\omega\|^2-\int_{\varOmega} a(x) \ln h(t \omega) \mathrm{d} x, |
so that
\frac{\mathrm{d}}{\mathrm{~d} t} I(t \omega)=t\|\omega\|^2-\int_{\varOmega} a(x) \frac{h^{\prime}(t \omega)}{h(t \omega)} \omega \mathrm{d} x . |
Note that, thanks to property 5) of Lemma 2.1, we get
\frac{h^{\prime}(t \omega(x))}{h(t \omega(x))} \omega(x) \leqslant(t \omega(x))^{-1} \omega(x)=t^{-1}, \forall x \in \varOmega . |
Thus
Claim 2.1 The set X is not empty.
Proof of Claim 2.1 We let φ1 be the first positive eigenfunction of-Δ in Ω with Dirichlet boundary condition, that is, -Δφ1=λ1φ1, φ1| ∂Ω=0, with λ1 the first Dirichlet eigenvalue of-Δ. It is well known that
\int_{\varOmega} \varphi_1^{-\gamma}(x) \mathrm{d} x<\infty | (5) |
if and only if -γ>-1 (c.f. Ref. [19]). By the result of Ref. [19](Theorem 1 and 2), we also know that: for any -α∈(-3, -1), there exist two positive constants d0, d1 and ωα(x)∈C2(Ω)∩C(Ω)∩H01(Ω), ωα(x)>0 in Ω such that
d_0 \varphi_1^{\frac{2}{1+\alpha}}(x) \leqslant \omega_\alpha(x) \leqslant d_1 \varphi_1^{\frac{2}{1+\alpha}}(x), \forall x \in \Omega. | (6) |
We choose
\begin{aligned} A_{\omega_{\alpha_0}} & =\left\{x \in \varOmega: \ln \omega_{\alpha_0}<-1\right\}, \\ B_{\omega_{\alpha_0}} & =\left\{x \in \varOmega:\left|\ln \omega_{\alpha_0}\right| \leqslant 1\right\}, \\ D_{\omega_{\alpha_0}} & =\left\{x \in \varOmega: \ln \omega_{\alpha_0}>1\right\}, \end{aligned} |
and then we have
\begin{aligned} & \left|\int_{B_{\omega_{\alpha_0}}} a(x) \ln \omega_{\alpha_0} \mathrm{~d} x\right| \leqslant \int_{B_{\omega_{\alpha_0}}} a(x)\left|\ln \omega_{\alpha_0}\right| \mathrm{d} x \\ & \leqslant \int_{B_{\omega_{\alpha_0}}} a(x) \mathrm{d} x \leqslant \int_{\varOmega} a(x) \mathrm{d} x<+\infty, \\ & \left|\int_{D_{\omega_{\alpha_0}}} a(x) \ln \omega_{\alpha_0} \mathrm{~d} x\right|=\int_{D_{\omega_{\alpha_0}}} a(x) \ln \omega_{\alpha_0} \mathrm{~d} x \\ & \leqslant \int_{\varOmega} a(x) \omega_{\alpha_0} \mathrm{~d} x \leqslant c\left(\int_{\varOmega} a(x)^p \mathrm{~d} x\right)^{\frac{1}{p}}\left\|\omega_{\alpha_0}\right\|<+\infty, \\ & \left|\int_{A_{\omega_{\alpha_0}}} a(x) \ln \omega_{\alpha_0} \mathrm{~d} x\right| \leqslant \int_{A_{\omega_{\alpha_0}}} a(x)\left|\ln \omega_{\alpha_0}\right| \mathrm{d} x \\ & \leqslant \int_{A_{\omega_{\alpha_0}}} a(x) \frac{1}{\omega_{\alpha_0}} \mathrm{~d} x \leqslant \int_{\varOmega} a(x) \frac{1}{\omega_{\alpha_0}} \mathrm{~d} x<+\infty, \end{aligned} |
where we have used the fact that
\begin{aligned} & \int_{\varOmega} a(x) \omega_{\alpha_0}^{-1} \mathrm{~d} x \leqslant d_0^{-1} \int_{\varOmega} a(x) \varphi_1^{\frac{2}{1+\alpha_0}}(x) \mathrm{d} x \\ & \leqslant d_0^{-1}\left(\int_{\varOmega} a^p(x) \mathrm{d} x\right)^{\frac{1}{p}}\left(\int_{\varOmega} \varphi_1^{-\frac{2 p^{\prime}}{1+\alpha_0}}(x) \mathrm{d} x\right)^{\frac{1}{p^{\prime}}}, \end{aligned} |
so that a(x)lnωα0(x)∈L1(Ω) and the set X is non-empty. This ends the proof of Claim 2.1.□
Claim 2.2 For every ω∈X there exists some t(ω)>0 (which may be not unique) such that
Proof of Claim 2.2 Fix ω∈X. We set
f(t)=t\|\omega\|^2-\int_{\varOmega} a(x) \frac{1}{h(t \omega)} h^{\prime}(t \omega) \omega \mathrm{d} x, \forall t>0 . |
Thanks to property 5) of Lemma 2.1, we have that
\frac{1}{2} \leqslant \frac{h^{\prime}(t \omega(x))}{h(t \omega(x))} t \omega(x) \leqslant 1, \quad \forall x \in \varOmega . |
We set
g_{i, \omega}(t)=t\|\omega\|^2-\frac{1}{i t} \int_{\varOmega} a(x) \mathrm{d} x, \forall t>0, i=1, 2, | (7) |
so that f(t) verifies
g_{1, \omega}(t) \leqslant f(t) \leqslant g_{2, \omega}(t), \forall t>0 . |
Clearly, gi, ω(t), i=1, 2, is strictly increasing for all t>0 and satisfies
g_{i, \omega}(t) \rightarrow+\infty \text { as } t \rightarrow+\infty, g_{i, \omega}(t) \rightarrow-\infty \text { as } t \rightarrow 0^{+}, |
then it follows that
f(t) \rightarrow+\infty \text { as } t \rightarrow+\infty, f(t) \rightarrow-\infty \text { as } t \rightarrow 0^{+} . |
Moreover, since
f(t(\omega))=\left.\frac{\mathrm{d}}{\mathrm{~d} t}\right|_{t=t(\omega)} I(t \omega)=0 . | (8) |
We thus obtain, thanks to ω∈X and (8), that
a(x) \ln (t(\omega) \omega(x)) \in L^1(\varOmega), t(\omega) \omega(x) \in \mathcal{N}.□ |
Proof of Theorem 1.1 We know from the property 4) of Lemma 2.1 that there exists
\begin{aligned} & I(\omega)=\frac{1}{2} \int_{\varOmega}|\nabla \omega|^2 \mathrm{~d} x-\int_{\varOmega} a(x) \ln h(\omega(x)) \mathrm{d} x \\ \geqslant & \frac{1}{2}\|\omega\|^2-\int_{\varOmega} a(x) h(\omega(x)) \mathrm{d} x\\ & \geqslant \frac{1}{2}\|\omega\|^2-\int_{\Omega} a(x) \omega(x) \mathrm{d} x \geqslant \frac{1}{2}\|\omega\|^2- \\ & \quad\left(\int_{\varOmega} a(x)^p \mathrm{~d} x\right)^{\frac{1}{p}}\left(\int_{\varOmega} \omega^{p^{\prime}}(x) \mathrm{d} x\right)^{\frac{1}{p^{\prime}}} \\ & \geqslant \frac{1}{2}\|\omega\|^2-c\|\omega\| \geqslant c_X, \quad \forall \omega \in X, \end{aligned} | (9) |
where we have used the fact that lns≤s, ∀s>0. This, with Claim 2.1 implies
\inf \limits_{\forall \omega \in X} I(\omega) \leqslant \inf \limits_{\forall \omega \in \mathcal{N}} I(\omega) \leqslant I(t(\omega) \omega) \leqslant I(\omega), \forall \omega \in X . |
Thus, we have obtained
\inf _X I=\inf _N I \in \mathbb{R} . |
We let {ωn} ⊂X be the sequence such that
\left\{\begin{array}{l} \omega \rightarrow \omega_0 \text { in } H_0^1(\varOmega), \\ \omega_n \rightarrow \omega_0 \text { a.e. in } \varOmega, \\ \omega_n \rightarrow \omega_0 \text { in } L^2(\varOmega), \\ \left|\omega_n(x)\right| \leqslant g(x), \left|\omega_0(x)\right| \leqslant g(x) \text { a. e. in } \varOmega . \end{array}\right. | (10) |
Since {ωn} ⊂X and ωn(x)>0 a.e. x∈Ω, we know ω0(x)≥0, a.e. x∈Ω. We prove now ω0(x)>0 a.e. x∈Ω and a(x)lnω0(x)∈L1(Ω) i.e. ω0∈X. Since {ωn} ⊂X, using (9), we write that
\begin{aligned} & \frac{1}{2}\left\|\omega_n\right\|^2-c_X \geqslant \int_{\varOmega} a(x) \omega_n(x) \mathrm{d} x \\ \geqslant & \int_{\varOmega} a(x) \ln \omega_n(x) \mathrm{d} x \geqslant \int_{\varOmega} a(x) \ln h\left(\omega_n(x)\right) \mathrm{d} x \\ = & \frac{1}{2}\left\|\omega_n\right\|^2-I\left(\omega_n\right), \end{aligned} |
where we have used property 4) of Lemma 2.1 for h. By the boundedness of {ωn} in H01(Ω) and
\begin{aligned} & \int_{\varOmega} a(x) \ln h\left(\omega_n(x)\right) \mathrm{d} x=O(1), \\ & \int_{\varOmega} a(x) \ln \omega_n(x) \mathrm{d} x=O(1), \\ & \int_{\varOmega} a(x) \omega_n(x) \mathrm{d} x=O(1) . \end{aligned} | (11) |
Taking advantage of a(x)(lnωn(x)-ωn(x))≤0 a.e. in Ω and using Fatou's Lemma, one also gets
\begin{aligned} & \limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x)\left(\ln \omega_n(x)-\omega_n(x)\right) \mathrm{d} x \\ \leqslant & \int_{\varOmega} a(x) \limsup _{n \rightarrow+\infty}\left(\ln \omega_n(x)-\omega_n(x)\right) \mathrm{d} x, \end{aligned} | (12) |
so that, since
\begin{aligned} & a(x) \limsup _{n \rightarrow+\infty}\left(\ln \omega_n(x)-\omega_n(x)\right) \\ = & \begin{cases}-\infty, & \omega_0(x)=0, \\ a(x)\left[\ln \omega_0(x)-\omega_0(x)\right], & \omega_0(x)>0, \end{cases} \end{aligned} |
we thus obtain, thanks to (11) and (12) and a(x)[lnω0(x)-ω0(x)]≤0, ∀x∈Ω, that
\omega_0(x)>0 \text { a. e. } x \in \varOmega . |
Since p>2 (so p′ < 2) and ωn→ω0 in L2(Ω), using Hölder's inequality, we get
\lim \limits_{n \rightarrow+\infty} \int_{\varOmega} a(x) \omega_n(x) \mathrm{d} x=\int_{\varOmega} a(x) \omega_0(x) \mathrm{d} x . | (13) |
We write now, with ω0(x)>0 a.e. in Ω, lnωn(x)-ωn(x)≤0 a.e. in Ω, and using Fatou's lemma, that
\begin{aligned} & \limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x) \ln \omega_n(x) \mathrm{d} x-\int_{\varOmega} a(x) \omega_0(x) \mathrm{d} x \\ & =\limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x) \ln \omega_n(x) \mathrm{d} x-\limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x) \omega_n(x) \mathrm{d} x \\ & \leqslant \limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x)\left(\ln \omega_n(x)-\omega_n(x)\right) \mathrm{d} x \\ & \leqslant \int_{\varOmega} a(x) \limsup _{n \rightarrow+\infty}\left(\ln \omega_n(x)-\omega_n(x)\right) \mathrm{d} x \\ & =\int_{\varOmega} a(x)\left(\ln \omega_0(x)-\omega_0(x)\right) \mathrm{d} x \\ & =\int_{\varOmega} a(x) \ln \omega_0(x) \mathrm{d} x-\int_{\varOmega} a(x) \omega_0(x) \mathrm{d} x . \end{aligned} |
Since
\begin{aligned} & -\infty<\limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x) \ln \omega_n(x) \mathrm{d} x \\ \leqslant & \int_{\varOmega} a(x) \ln \omega_0(x) \mathrm{d} x \leqslant \int_{\varOmega} a(x) \omega_0(x) \mathrm{d} x<+\infty, \end{aligned} |
so that we obtain a(x)lnω0(x)∈L1(Ω). It is worthy remarking here that the main difference between dealing with the singularities u-1 and u-ν(ν≠1) lies in the energy controlling:
Then we prove that
\begin{aligned} & a(x) \ln h\left(\omega_n(x)\right) \leqslant a(x) \ln \left(\omega_n(x)\right) \\ \leqslant & a(x) \omega_n(x) \leqslant a(x) g(x) \end{aligned} |
and a(x)g(x)∈L1(Ω), using Fatou's lemma and ωn(x)→ω0 a.e. in Ω, we get
\begin{gathered} \limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x) \ln h\left(\omega_n(x)\right) \mathrm{d} x \leqslant \\ \int_{\varOmega} a(x) \ln h\left(\omega_0(x)\right) \mathrm{d} x, \end{gathered} | (14) |
so that
\begin{aligned} & \inf _X I=\lim _{n \rightarrow+\infty} I\left(\omega_n\right)=\liminf _{n \rightarrow+\infty} \int_{\varOmega} \frac{1}{2}\left|\nabla \omega_n\right|^2- \\ & a(x) \ln h\left(\omega_n(x)\right) \mathrm{d} x \geqslant \liminf _{n \rightarrow+\infty} \frac{1}{2} \int_{\Omega}\left|\nabla \omega_n\right|^2 \mathrm{~d} x+ \\ & \liminf _{n \rightarrow+\infty} \int_{\varOmega}-a(x) \ln h\left(\omega_n(x)\right) \mathrm{d} x \\ & \geqslant \frac{1}{2}\left\|\omega_0\right\|^2+\int_{\varOmega}-a(x) \ln h\left(\omega_0(x)\right) \mathrm{d} x=I\left(\omega_0\right) \\ & \geqslant I\left(t\left(\omega_0\right) \omega_0\right) \geqslant \inf _N I=\inf _X I, \end{aligned} |
where we have used the weak lower semicontinuity of norm, (14), and
\begin{aligned} & \liminf _{n \rightarrow+\infty} \int_{\varOmega}-a(x) \ln h\left(\omega_n(x)\right) \mathrm{d} x \\ = & -\limsup _{n \rightarrow+\infty} \int_{\varOmega} a(x) \ln h\left(\omega_n(x)\right) \mathrm{d} x . \end{aligned} |
This proves that
I\left(\omega_0\right)=\inf _X I | (15) |
This leads to
\frac{{\rm{d}}}{{{\rm{d}}t}}\left| {_{t = 1}I\left( {t{\omega _0}} \right)} \right. = 0, |
that is,
We now show that h(ω0) is a solution of problem (1). Suppose φ∈H01(Ω), φ(x)≥0, ∀x∈Ω and ε>0. Set v=ω0+εφ, we divide the domain Ω into three parts:
\begin{gathered} A_v=\{x \in \varOmega: \ln v<-1\}, \\ B_v=\{x \in \varOmega:|\ln v| \leqslant 1\}, \\ D_v=\{x \in \varOmega: \ln v>1\}, \end{gathered} |
and then we have
\begin{aligned} & \left|\int_{R_v} a(x) \ln \left(\omega_0+\varepsilon \varphi\right) \mathrm{d} x\right| \\ & \leqslant \int_{B_v} a(x)\left|\ln \left(\omega_0+\varepsilon \varphi\right)\right| \mathrm{d} x \\ & \leqslant \int_{B_v} a(x) \mathrm{d} x \leqslant \int_{\Omega} a(x) \mathrm{d} x<+\infty, \\ & \left|\int_{D_v} a(x) \ln \left(\omega_0+\varepsilon \varphi\right) \mathrm{d} x\right|=\int_{D_v} a(x) \ln \left(\omega_0+\varepsilon \varphi\right) \mathrm{d} x \\ & \leqslant \int_{\varOmega} a(x)\left(\omega_0+\varepsilon \varphi\right) \mathrm{d} x \\ & \leqslant c\left(\int_{\varOmega} a(x)^p \mathrm{~d} x\right)^{\frac{1}{p}}\left\|\omega_0+\varepsilon \varphi\right\|<+\infty, \\ & \left|\int_{A_v} a(x) \ln \left(\omega_0+\varepsilon \varphi\right) \mathrm{d} x\right| \\ & \leqslant \int_{A_v} a(x)\left|\ln \left(\omega_0+\varepsilon \varphi\right)\right| \mathrm{d} x \\ & =\int_{A_v} a(x) \ln \frac{1}{\left(\omega_0+\varepsilon \varphi\right)} \mathrm{d} x \leqslant \int_{A_v} a(x) \ln \frac{1}{\omega_0} \mathrm{~d} x \\ & \leqslant \int_{A_v} a(x)\left|\ln \omega_0\right| \mathrm{d} x \\ & \leqslant \int_{\varOmega} a(x)\left|\ln \omega_0\right| \mathrm{d} x<+\infty \end{aligned} |
so ω0+εφ∈X. By Claim 2.2 and (15) we can conclude that
\begin{aligned} I\left(\omega_0+\varepsilon \varphi\right) & \geqslant I\left(t\left(\omega_0+\varepsilon \varphi\right)\left(\omega_0+\varepsilon \varphi\right)\right) \\ & \geqslant \inf _N I=I\left(\omega_0\right), \end{aligned} |
and then we have
\begin{aligned} & \frac{\left\|\omega_0+\varepsilon \varphi\right\|^2-\left\|\omega_0\right\|^2}{2} \\ \geqslant & \int_{\varOmega} a(x) \ln h\left(\omega_0+\varepsilon \varphi\right) \mathrm{d} x-\int_{\varOmega} a(x) \ln h\left(\omega_0\right) \mathrm{d} x \end{aligned} | (16) |
Dividing (16) by ε>0 and letting ε→0+ we conclude with Fatou's Lemma that
\begin{aligned} & \int_{\varOmega} \nabla \omega_0 \nabla \varphi \mathrm{~d} x \\ \geqslant & \liminf _{\varepsilon \rightarrow 0^{+}} \frac{\int_{\varOmega} a(x)\left[\ln h\left(\omega_0+\varepsilon \varphi\right)-\ln h\left(\omega_0\right)\right] \mathrm{d} x}{\varepsilon} \\ \geqslant & \int_{\varOmega} \liminf _{\varepsilon \rightarrow 0^{+}} \frac{a(x)\left[\ln h\left(\omega_0+\varepsilon \varphi\right)-\ln h\left(\omega_0\right)\right]}{\varepsilon} \mathrm{d} x \\ = & \int_{\varOmega} a(x) \frac{h^{\prime}\left(\omega_0\right)}{h\left(\omega_0\right)} \varphi \mathrm{d} x . \end{aligned} | (17) |
Suppose ψ∈H01(Ω) and t>0. Inserting φ=(ω0+tψ)+ into (17), we get
\begin{aligned} & 0 \leqslant \frac{1}{t} \int_{\varOmega} \nabla \omega_0 \nabla\left(\omega_0+t \psi\right)^{+}- \\ & a(x) \frac{h^{\prime}\left(\omega_0\right)}{h\left(\omega_0\right)}\left(\omega_0+t \psi\right)^{+} \mathrm{d} x=\frac{1}{t}\left(\int_{\varOmega}-\int_{\varOmega \cap\left[\omega_{0}+t \psi<0\right]}\right) \\ & \left(\nabla \omega_0 \nabla\left(\omega_0+t \psi\right)-a(x) \frac{h^{\prime}\left(\omega_0\right)}{h\left(\omega_0\right)}\left(\omega_0+t \psi\right)\right) \mathrm{d} x \\ & \leqslant \int_{\varOmega} \nabla \omega_0 \nabla \psi-a(x) \frac{h^{\prime}\left(\omega_0\right)}{h\left(\omega_0\right)} \psi \mathrm{d} x- \\ & \int_{\varOmega \cap\left[\omega_0+t \psi<0\right]} \nabla \omega_0 \nabla \psi \mathrm{~d} x . \end{aligned} | (18) |
Since meas[ω0+tψ]→0 as t→0+, by passing to the limit as t→0+ we can conclude that
0 \leqslant \int_{\varOmega} \nabla \omega_0 \nabla \psi-a(x) \frac{h^{\prime}\left(\omega_0\right)}{h\left(\omega_0\right)} \psi \mathrm{d} x . |
The same conclusion can be drawn for -ψ in place of ψ, thus it follows that h(ω0) is indeed a H01(Ω)-solution of (1).□
[1] |
Kurihara S. Large-amplitude quasi-solitons in superfluid films[J]. Journal of the Physical Society of Japan, 1981, 50(10): 3262-3267. DOI:10.1143/jpsj.50.3262 |
[2] |
Hartmann B, Zakrzewski W J. Electrons on hexagonal lattices and applications to nanotubes[J]. Physical Review B, 2003, 68(18): 184302. DOI:10.1103/physrevb.68.184302 |
[3] |
Kosevich A M, Ivanov B A, Kovalev A S. Magnetic solitons[J]. Physics Reports, 1990, 194(3/4): 117-238. DOI:10.1016/0370-1573(90)90130-T |
[4] |
Makhankov V G, Fedyanin V K. Non-linear effects in quasi-one-dimensional models of condensed matter theory[J]. Physics Reports, 1984, 104(1): 1-86. DOI:10.1016/0370-1573(84)90106-6 |
[5] |
Liu J Q, Wang Y Q, Wang Z Q. Soliton solutions for quasilinear Schrödinger equations, Ⅱ[J]. Journal of Differential Equations, 2003, 187(2): 473-493. DOI:10.1016/s0022-0396(02)00064-5 |
[6] |
Fang X D, Szulkin A. Multiple solutions for a quasilinear Schrödinger equation[J]. Physics Reports, 2013, 254(4): 2015-2032. DOI:10.1016/j.jde.2012.11.017 |
[7] |
Liu X Q, Liu J Q, Wang Z Q. Quasilinear elliptic equations via perturbation method[J]. Proceedings of the American Mathematical Society, 2013, 141(1): 253-263. DOI:10.1090/s0002-9939-2012-11293-6 |
[8] |
Agarwal R P, O'Regan D. Singular differential and integral equations with applications[M]. Dordretht, Netherlands: Springer Science & Business Media, 2003. DOI:10.1007/978-94-017-3004-4
|
[9] |
Hernández J, Mancebo F J. Singular elliptic and parabolic equations[M]//Handbook of differential equations: stationary partial differential equations, Amsterdam; Boston: Elsevier/North Holland. 2006, 3: 317-400. DOI: 10.1016/s1874-5733(06)80008-2.
|
[10] |
Marcos do Ó J, Moameni A. Solutions for singular quasilinear Schrödinger equations with one parameter[J]. Communications on Pure & Applied Analysis, 2010, 9(4): 1011-1023. DOI:10.3934/cpaa.2010.9.1011 |
[11] |
Santos C A, Yang M B, Zhou J Z. Global multiplicity of solutions for a modified elliptic problem with singular terms[J]. Nonlinearity, 2021, 34(11): 7842-7871. DOI:10.1088/1361-6544/ac2a50 |
[12] |
Alves R L, Reis M. About existence and regularity of positive solutions for a quasilinear Schr ödinger equation with singular nonlinearity[J]. Electronic Journal of Qualitative Theory of Differential Equations, 2020, 60: 1-23. DOI:10.14232/ejqtde.2020.1.60 |
[13] |
Sun Y J. Compatibility phenomena in singular problems[J]. Proceedings of the Royal Society of Edinburgh Section A: Mathematics, 2013, 143(6): 1321-1330. DOI:10.1017/s030821051100117x |
[14] |
Alves R L. Existence, nonexistence and multiplicity of positive solutions for singular quasilinear problems[J]. Electronic Journal of Qualitative Theory of Differential Equations, 2022, 13: 1-29. DOI:10.14232/ejqtde.2022.1.13 |
[15] |
Liu J Y, Liu D C, Zhao P H. Soliton solutions for a singular Schr ödinger equation with any growth exponents[J]. Acta Applicandae Mathematicae, 2017, 148(1): 179-199. DOI:10.1007/s10440-016-0084-z |
[16] |
do Ó J M B, Miyagaki O H, Soares S H M. Soliton solutions for quasilinear Schrödinger equations: the critical exponential case[J]. Nonlinear Analysis: Theory, Methods & Applications, 2007, 67(12): 3357-3372. DOI:10.1016/j.na.2006.10.018 |
[17] |
do Ó J M B, Miyagaki O H, Soares S H M. Soliton solutions for quasilinear Schrödinger equations with critical growth[J]. Journal of Differential Equations, 2010, 248(4): 722-744. DOI:10.1016/j.jde.2009.11.030 |
[18] |
Colin M, Jeanjean L. Solutions for a quasilinear Schr ödinger equation: a dual approach[J]. Nonlinear Analysis: Theory, Methods & Applications, 2004, 56(2): 213-226. DOI:10.1016/j.na.2003.09.008 |
[19] |
Lazer A C, McKenna P J. On a singular nonlinear elliptic boundary-value problem[J]. Proceedings of the American Mathematical Society, 1991, 111(3): 721. DOI:10.1090/s0002-9939-1991-1037213-9 |