Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2022, Volume 213, Issue 7, Pages 1004–1019
DOI: https://doi.org/10.4213/sm9655e
(Mi sm9655)
 

This article is cited in 1 scientific paper (total in 1 paper)

Semiregular solutions of elliptic boundary-value problems with discontinuous nonlinearities of exponential growth

V. N. Pavlenkoa, D. K. Potapovb

a Chelyabinsk State University, Chelyabinsk, Russia
b Saint Petersburg State University, St. Petersburg, Russia
References:
Abstract: An elliptic boundary-value problem with discontinuous nonlinearity of exponential growth at infinity is investigated. The existence theorem for a weak semiregular solution of this problem is deduced by the variational method. The semiregularity of a solution means that its values are points of continuity of the nonlinearity with respect to the phase variable almost everywhere in the domain where the boundary-value problem is considered. The variational approach used is based on the concept of a quasipotential operator, unlike the traditional approach, which uses Clarke's generalized derivative.
Bibliography: 29 titles.
Keywords: elliptic boundary-value problem, discontinuous nonlinearity, exponential growth, semiregular solution, variational method.
Funding agency Grant number
Russian Foundation for Basic Research 20-41-740003
The research was funded by RFBR and Chelyabinsk Region, project number 20-41-740003.
Received: 16.08.2021 and 17.03.2022
Russian version:
Matematicheskii Sbornik, 2022, Volume 213, Number 7, Pages 121–138
DOI: https://doi.org/10.4213/sm9655
Bibliographic databases:
Document Type: Article
MSC: Primary 35J20; Secondary 35R05
Language: English
Original paper language: Russian

§ 1. Introduction

In a bounded domain $\Omega\subset{\mathbb R}^2$ of class $C^{1,1}$ we consider an elliptic boundary-value problem of the form

$$ \begin{equation} Lu(x)=g(x,u(x)), \qquad x\in\Omega, \end{equation} \tag{1.1} $$
$$ \begin{equation} u(x)=0, \qquad x\in\partial\Omega. \end{equation} \tag{1.2} $$
The differential operator
$$ \begin{equation} Lu(x)\equiv-\sum_{i,j=1}^2(a_{ij}(x)u_{x_i})_{x_j}+c(x)u(x) \end{equation} \tag{1.3} $$
is uniformly elliptic in $\Omega$ with ellipticity constant $\chi$, $a_{ij}\in C^1(\overline\Omega)$, $a_{ij}(x)=a_{ji}(x)$, $c\in C(\overline\Omega)$ and $c(x)\geqslant 0$ in $\Omega$. The function $g(x,u)$ is superpositionally measurable on $\Omega\times\mathbb R$, that is, for any measurable function $u(x)$ on $\Omega$ the composition $g(x,u(x))$ is measurable. The function $g(x,\cdot\,)$ is assumed to have finite one-sided limits at any point $u\in\mathbb R$ for almost all $x\in\Omega$; we denote them by $g(x,u-)$ and $g(x,u+)$ (the limits from the left and right, respectively). In addition, the nonlinearity $g(x,u)$ admits a growth at infinity of order $|u|^\alpha \exp(A|u|^\tau)$ with respect to the phase variable $u$, where the constants $\alpha$, $A$, and $\tau$ are positive and $\tau<2$.

A weak solution of (1.1), (1.2) is a function $u(x)$ in the space $\mathring{W}^1_2(\Omega)$ such that for any $v\in\mathring{W}^1_2(\Omega)$

$$ \begin{equation*} \sum_{i,j=1}^2\int_\Omega a_{ij}(x)u_{x_i}v_{x_j}\,dx+\int_\Omega c(x)u(x)v(x)\,dx =\int_\Omega g(x,u(x))v(x)\,dx. \end{equation*} \notag $$

A weak solution of problem (1.1), (1.2) is called semiregular if the value $u(x)$ is a point of continuity of the function $g(x,\cdot\,)$ for almost all $x\in\Omega$.

We give a statement of the main result of our investigations.

Theorem 1. Assume that

1) for almost all $x\in\Omega$ the inequality $g(x,u-)\leqslant g(x,u+)$ holds and

$$ \begin{equation*} g(x,u)\in [g(x,u-),g(x,u+)] \quad \forall\, u\in\mathbb R; \end{equation*} \notag $$

2) there are positive constants $b$, $\alpha$, $A$, and $\tau$, $\tau<2$, such that for almost all $x\in\Omega$

$$ \begin{equation} |g(x,u)|\leqslant a(x)+b|u|^\alpha \exp(A|u|^\tau) \quad \forall\, u\in\mathbb R, \end{equation} \tag{1.4} $$
where $a\in L_N(\Omega)$, and $L_N(\Omega)$ is the Orlicz class of measurable functions on $\Omega$ that is associated with the $N$-function $N(u)$, the complementary function of an $N$-function $M(u)=|u|^{\alpha+2}\exp(|u|^s)$, $s\in (\tau,2)$ (see § 2.1 below);

3) for almost all $x\in\Omega$

$$ \begin{equation*} \int_0^ug(x,s)\,ds\leqslant \frac{ku^2+d(x)|u|^\theta+d_1(x)}2 \quad \forall\, u\in\mathbb R, \end{equation*} \notag $$
where $k<\chi/\|P_1\|^2$, $\chi$ is the ellipticity constant of the differential operator $L$ defined by (1.3), $P_1$ is the embedding operator of $\mathring{W}^1_2(\Omega)$ in $L_2(\Omega)$, $d\in L_{2/(2-\theta)}(\Omega)$, ${0<\theta<2}$ and $d_1$ is an integrable function on $\Omega$.

Then problem (1.1), (1.2) has a weak semiregular solution $u(x)$ in $\mathring W_2^1(\Omega)$.

Remark 1. In the proof of Theorem 1, we show that the integral $\displaystyle\int_\Omega \!g(x,u(x))v(x)\,dx$ in the definition of a weak solution of (1.1), (1.2) exists under the conditions of this theorem for any $u,v\in \mathring{W}^1_2(\Omega)$.

Remark 2. Condition 3) in Theorem 1 is a hard constraint on a nonlinearity $g(x,u)$ that is of exponential growth at infinity. It is not fulfilled if $g(x,u)\leqslant 0$ on $\Omega\times (-\infty,-u_0]$ or $g(x,u)\geqslant 0$ on $\Omega\times [u_0,+\infty)$ for some $u_0>0$ and this function has exponential growth as $u\to -\infty$ or $u\to +\infty$, respectively. However, if for some ${u_0>0}$ the function $g(x,u)$ is bounded on $\Omega\times [-u_0,u_0]$, positive on ${\Omega\times (-\infty,-u_0]}$ and negative on ${\Omega\times [u_0,+\infty)}$, then condition 3) in Theorem 1 holds for it. For example, the following function is of this type:

$$ \begin{equation*} g(x, u)= \begin{cases} \operatorname{sgn}u+u^2e^{|u|}, & u\leqslant 0, \\ \operatorname{sgn}u-u^2e^{|u|}, & u\geqslant 0. \end{cases} \end{equation*} \notag $$
Note that $g(x,u)$ of this type also satisfies conditions 1) and 2) in Theorem 1.

The existence problem for weak solutions of elliptic boundary-value problems with discontinuous nonlinearities of exponential growth was considered in a number of works. We note the papers [1]–[6], which are typical for this area.

In [1] a quasilinear elliptic-type equation with homogeneous Dirichlet boundary condition of the form

$$ \begin{equation} -\operatorname{div}(a(x,\nabla u))=\lambda h(x)\exp(\alpha_0|u|^{n/(n-1)})+f(x,u), \qquad x\in\Omega, \end{equation} \tag{1.5} $$
$$ \begin{equation} u(x)=0, \qquad x\in\partial\Omega, \end{equation} \tag{1.6} $$
was considered in a bounded domain $\Omega\subset{\mathbb R}^n$ ($n\geqslant 2$) with smooth boundary $\partial\Omega$. It was assumed that $a(x,t)$ satisfies the Leray-Lions conditions of growth, coercivity and monotonicity and there exists a Carathéodory function $A(x,t)$ such that $a(x,t)=\nabla_t A(x,t)$. The function $f(x,s)$ is superpositionally measurable on $\Omega\times\mathbb R$ (a function of this type can have discontinuities with respect to $s$) and nondecreasing with respect to $s$ on $\mathbb R$; $f(x,0)=0$ on $\Omega$; $f(x,s)$ can have exponential critical growth in $s$, that is, there is $\alpha_0>0$ such that $\lim_{|s|\to\infty}(|f(x,s)|/\exp(\alpha|s|^{n/(n-1)}))=0$ if $\alpha>\alpha_0$ and this limit is $+\infty$ if $\alpha<\alpha_0$. The parameter $\lambda$ is positive; $h\in L_q(\Omega)$, $q>1$, $h(x)\geqslant 0$ almost everywhere on $\Omega$ and $h(x)>0$ on a set of nonzero measure in $\Omega$.

A weak solution of problem (1.5), (1.6) is a function $u\in\mathring{W}^1_n(\Omega)$ such that

$$ \begin{equation*} \int_\Omega a(x,\nabla u)\cdot\nabla v\,dx\,{=}\,\lambda\int_\Omega h(x)\exp(\alpha_0|u|^{n/(n-1)})v\,dx+\int_\Omega f(x,u)v\,dx \quad\! \forall\, v\,{\in}\,\mathring{W}^1_n(\Omega). \end{equation*} \notag $$

It was established in [1] that for sufficiently small $\lambda$ a nonnegative weak solution exists under the above assumptions. The proof reduces to verifying the conditions of the general Carl-Heikkilä theorem on the existence of a fixed point for a monotone operator in a semiordered Banach space. We present its statement.

Theorem 2 (see [7], Corollary 2.2). Let $E$ be a Banach semilattice and let the space $E$ be reflexive. Then any monotone mapping $S\colon B\to B$, where $B\subset E$ is an arbitrary ball, has a fixed point.

A Banach space $E$ equipped with a partial order relation $\leqslant$ is called a Banach semilattice if $\sup\{x,y\}$ and $\inf\{x,y\}$ exist for any $x,y\in E$ and $\|x^\pm\|\leqslant \|x\|$ for any $x\in E$. Monotonicity in Theorem 2 is understood in the sense of semiordering: ${Sx\leqslant Sy}$ for any $x$, $y\in B$ such that $x\leqslant y$. Here $x^+=\sup\{0,x\}$ and ${x^-=-\inf\{0,x\}}$.

Note that the space $\mathring{W}^1_n(\Omega)$ with the natural partial order relation satisfies the assumptions of Theorem 2. In the case of the operator statement of problem (1.5), (1.6), the Trudinger-Moser inequality (see [8]) and theorems on the embedding of $\mathring{W}^1_n(\Omega)$ in Orlicz spaces (see [9]) are actually used.

In [2] the existence of a weak solution of the equation

$$ \begin{equation} -\operatorname{div} (|\nabla u|^{n-2}\nabla u)+V(x)|u|^{n-2}u=g(x,u)+\lambda h(x) \end{equation} \tag{1.7} $$
in ${\mathbb R}^n$ ($n\geqslant 2$) was established for sufficiently small values of the positive parameter $\lambda$. The potential $V(x)$ can change sign; the nonlinearity $g(x,u)$, which is nondecreasing in $u$, can have critical exponential growth and be discontinuous with respect to $u$. Solutions are considered in the space
$$ \begin{equation*} X=\biggl\{u\in W^1_n({\mathbb R}^n)\colon \int_{{\mathbb R}^n}V^+(x)|u|^n\,dx<\infty\biggr\} \end{equation*} \notag $$
with the norm
$$ \begin{equation*} \|u\|=\biggl(\int_{{\mathbb R}^n}\bigl (|\nabla u|^n+V^+(x)|u|^n\bigr)\,dx\biggr)^{1/n}, \end{equation*} \notag $$
where $V^+=\max\{V,0\}$.

A weak solution of equation (1.7) is a function $u\in X$ such that

$$ \begin{equation*} \int_{{\mathbb R}^n}|\nabla u|^{n-2}\nabla u\nabla v\,dx+\int_{{\mathbb R}^n}V(x)|u|^{n-2}uv\,dx= \int_{{\mathbb R}^n}g(x,u)v\,dx+\lambda\langle h,v\rangle \quad \forall\, v\in X, \end{equation*} \notag $$
where $\langle h,v\rangle$ is the value of $h\in X^*$ for $v\in X$ ($X^*$ is the dual space of $X$).

The existence theorem for weak solutions was proved in [2], just as in [1], by verifying that the conditions of Theorem 2 hold in the space $X$ with the use of a version of the Trudinger-Moser inequality for the whole space.

In [3] a generalization of (1.7) in ${\mathbb R}^n$ was considered. It was obtained by replacing the differential part in (1.7) by $-\operatorname{div}(a(x,\nabla u))$ and the nonlinearity $g(x,u)$ by $Q(x)f(u)$, where $Q(x)$ can be unbounded with respect to $x$ and the nondecreasing function $f(u)$ can be discontinuous and have critical exponential growth. Here $a(x,t)$ satisfies the same conditions as in [1], for $\Omega$ replaced by ${\mathbb R}^n$. A sufficient condition for a weak solution to exist in the space $X$ (for the definition of $X$, see above) for sufficiently small values of the parameter $\lambda$ was derived. Like in [1] and [2], the proof reduced to verifying the assumptions of Theorem 2 for the operator statement of problem (1.7) in $X$.

An analogue of the equation in [3] with monotone nonlinearity of critical exponential growth was considered in [4] on a compact Riemannian manifold without boundary and in [5] on a noncompact Riemannian manifold of dimension $n\geqslant 2$. By Theorem 2, existence theorems for weak solutions were obtained in a space defined similarly to $X$.

In [6] the variational approach was used. That paper considers the inclusion

$$ \begin{equation} -\Delta u+V(x)u-\varepsilon h(x)\in\partial_t F(x,u), \qquad x\in {\mathbb R}^2, \end{equation} \tag{1.8} $$
where $\varepsilon > 0$; $V$ is a continuous function such that $V(x) \geqslant V_0$ on ${\mathbb R}^2$, where $V_0$ is a positive constant, and $V^{-1} \in L_1({\mathbb R}^2)$; $h \in (W^1_2({\mathbb R}^2))^*$ and $\displaystyle 0 < \int_{{\mathbb R}^2}h(x)\,dx < + \infty$; $\displaystyle F(x,t)=\int_0^t f(x,s)\,ds$, $f(x,s)$ is a discontinuous function of exponential critical growth with respect to $s$; $\partial_t F(x,t)$ is the Clarke generalized derivative of $F(x,t)$ with respect to $t$. It is assumed that $\underline f(x,t):=\lim_{\eta\to t}\inf f(x,\eta)$ and $\overline f(x,t):=\lim_{\eta\to t}\sup f(x,\eta)$ are superpositionally measurable on ${\mathbb R}^2\times\mathbb R$ and there exists $t_0\geqslant 0$ such that $f(x,t)=0$ if $t<t_0$ and $f(x,t)>0$ if $t>t_0$ for any $x\in\mathbb R^2$. Problem (1.8) was considered in the Hilbert space
$$ \begin{equation*} E=\biggl\{u\in W^1_2({\mathbb R}^2)\colon \int_{{\mathbb R}^2}V(x)u^2(x)\,dx<+\infty\biggr\} \end{equation*} \notag $$
with inner product
$$ \begin{equation*} (u,v)=\int_{{\mathbb R}^2}\bigl(\nabla u\nabla v+V(x)u(x)v(x)\bigr)\,dx. \end{equation*} \notag $$

A generalized weak solution of problem (1.8) is a function $u\in E$ such that there exists $z(x)$ lying in the intervals $ [\underline f(x,u(x)),\overline f(x,u(x))]$ almost everywhere on ${\mathbb R}^2$ such that

$$ \begin{equation*} (u,v)-\int_{{\mathbb R}^2}z(x)v(x)\,dx-\varepsilon\int_{{\mathbb R}^2}h(x)v(x)\,dx=0 \end{equation*} \notag $$
for any $v\in E$ and $\operatorname{mes}\{x\in {\mathbb R}^2\colon u(x)>t_0\}\neq 0$.

The following restrictions were additionally imposed in [6]:

Note that the above conditions on the nonlinearity are satisfied if

$$ \begin{equation*} f(x,t)\equiv f(t)=2H(t-a)t^p \exp(t^2) \end{equation*} \notag $$
for all $t\in\mathbb R$. Here $a\geqslant 0$, $H(s)=0$ for $s\leqslant 0$ and $H(s)=1$ for $s>0$ (the Heaviside step function).

Under the above conditions the existence of two solutions of inclusion (1.8) was proved in [6] for sufficiently small $\varepsilon$ and $t_0$ and sufficiently large $\mu$ in condition 3). Ekeland’s variational principle and the mountain pass theorem for nondifferentiable functionals were used in the proof.

Semiregular solutions of elliptic-type equations with discontinuous nonlinearities were introduced by Krasnosel’skii and Pokrovskii [10]. In [11] they deduced an existence theorem for a semiregular solution of an elliptic boundary-value problem with discontinuous nonlinearity using the method of upper and lower solutions. The present authors obtained more general results in [12]. Abstract theorems in this area were proved in [13]. We mention the most recent papers [14]–[20] concerning semiregular solutions. Note that equations with nonlinearities of exponential growth were not considered in [10]–[20].

This paper presents a sufficient existence condition for a semiregular solution of problem (1.1), (1.2) with a discontinuous nonlinearity admitting an exponential growth with respect to the phase variable (Theorem 1) to exist. We do not assume that the nonlinearity $g(x,s)$ is nondecreasing with $s$ (a key requirement in [1]–[5]) and that $g(x,s)=0$ for $s<t_0$ and $g(x,s)>0$ for $s>t_0$ for some $t_0$ (one of the assumptions in [6]). In addition, in the proof of Theorem 1 we use the variational approach based on the concept of a quasipotential operator (see, for example, [21]–[23]), unlike the traditional approach, which uses the Clarke generalized derivative. One of the most recent papers where it was used was [20].

§ 2. Preliminary information

2.1. Orlicz spaces

The necessary notions and facts concerning Orlicz spaces have been borrowed from [24].

A continuous convex function $M(u)$ on $\mathbb R$ is called an $N$-function if it is even and satisfies the conditions $\lim_{u\to 0}M(u)/u\,{=}\,0$ and $\lim_{u\to\infty}M(u)/u\,{=}\,{+}\infty$.

Examples of $N$-functions are: $\varphi(u)=\exp(u^2)-1$ and $\psi(u)=|u|^{\alpha+2}\exp(|u|^s)$, where $\alpha$ and $s$ are positive constants.

The complementary function of an $N$-function $M(u)$ is its Legendre transform $N(u)=\max\{|u|t-M(t)\colon t\geqslant 0\}$. Note that $N(u)$ is also an $N$-function. The functions $M$ and $N$ are called mutually complementary or conjugate.

An $N$-function $M(u)$ is said to satisfy the $\Delta_2$-condition if $M(lu)\leqslant m(l)M(u)$ for sufficiently large $u$, where $l$ can be any number greater than 1. Note that functions satisfying the $\Delta_2$-condition grow not faster than power functions. In particular, the functions $\varphi(u)$ and $\psi(u)$ defined above do not satisfy the $\Delta_2$-condition. The natural question arises as to how we can verify whether the complementary function $N(u)$ of $M(u)$ satisfies the $\Delta_2$-condition. It turns out that the existence of $l>1$ and $v_0\geqslant 0$ such that $M(v)\leqslant M(lv)/(2l)$ for all $v\geqslant v_0$ is necessary and sufficient for this (see [24], Ch. I, § 4, Theorem 4.2). It immediately follows that the complementary functions of $\varphi(u)$ and $\psi(u)$ satisfy the $\Delta_2$-condition.

To prove Theorem 1 we need the inequality

$$ \begin{equation} N^{-1}(v)M^{-1}(v)\geqslant v \quad \forall\, v>0, \end{equation} \tag{2.1} $$
where $M$ and $N$ are mutually complementary functions, while $M^{-1}$ and $N^{-1}$ are the inverse functions.

Let $\Omega$ be a bounded domain in ${\mathbb R}^n$ and $M(u)$ be an $N$-function. The Orlicz class associated with $M(u)$ is the set

$$ \begin{equation*} \begin{aligned} \, L_M(\Omega)&=\biggl\{\text{measurable functions } u\colon \Omega\to {\mathbb R} \\ &\qquad\qquad\text{such that } \rho(u,M)=\int_\Omega M(u(x))\,dx<\infty\biggr\}. \end{aligned} \end{equation*} \notag $$
Functions that differ on a zero measure set are considered to be indistinguishable in this case. The notation $L_M$ is used instead of $L_M(\Omega)$ when this cannot lead to misunderstanding.

An Orlicz class is a convex set containing all bounded measurable functions. Its elements are integrable functions; however, not any integrable function belongs to $L_M$.

The class $L_M$ is a linear space if and only if $M(u)$ satisfies the $\Delta_2$-condition.

Let $M(u)$ and $N(v)$ be mutually complementary $N$-functions. We set

$$ \begin{equation*} \begin{aligned} \, L_M^*(\Omega)&=\biggl\{\text{measurable functions } u\colon\Omega\to{\mathbb R} \\ &\qquad\qquad\text{such that } (u,v)=\int_\Omega u(x)v(x)\,dx<\infty\ \forall\, v\in L_N(\Omega)\biggr\}. \end{aligned} \end{equation*} \notag $$
In this case, like in the definition of Orlicz classes, functions that differ on a zero measure set are considered to be indistinguishable. The notation $L_M^*$ is used instead of $L_M^*(\Omega)$ when this cannot lead to misunderstanding.

The set $L_M^*(\Omega)$ is a linear space and coincides with the span of $L_M(\Omega)$. It is endowed with the Orlicz norm

$$ \begin{equation*} \|u\|_M=\sup\biggl\{\biggl|\int_\Omega u(x)v(x)\,dx\biggr|\colon \rho(v,N)\leqslant 1\biggr\}. \end{equation*} \notag $$
This normed space is complete; it is denoted by $L_M^*(\Omega)$ and called the Orlicz space associated with the $N$-function $M(u)$.

The following two inequalities are important for further considerations:

$$ \begin{equation} \|u\|_M\leqslant\rho(u,M)+1\quad \forall\, u\in L_M; \end{equation} \tag{2.2} $$
if $\|u\|_M\leqslant 1$, then $u\in L_M$ and
$$ \begin{equation} \rho(u,M)\leqslant \|u\|_M. \end{equation} \tag{2.3} $$

The Luxemburg norm, which is equivalent to the Orlicz norm, can be introduced in the space $L_M^*$ by the formula

$$ \begin{equation*} \|u\|_{(M)}=\inf\biggl\{m\colon \rho\biggl(\frac um,M\biggr)\leqslant 1\biggr\}. \end{equation*} \notag $$
It is related to the Orlicz norm by
$$ \begin{equation*} \|u\|_{(M)}\leqslant \|u\|_M\leqslant 2\|u\|_{(M)}. \end{equation*} \notag $$
The closure of the set of bounded functions in $L_M^*$ is denoted by $E_M$. If $M(u)$ satisfies the $\Delta_2$-condition, then $E_M=L_M^*=L_M$. In the general case the dual space $E_M^*$ of $E_M$ coincides with $L_N^*$ endowed with the Luxemburg norm, where $N(u)$ is the complementary function of $M(u)$.

We switch to Trudinger’s embedding theorems for Orlicz spaces.

An $N$-function $M_1(u)$ is said to grow significantly faster than an $N$-function $M(u)$ if $\lim_{u\to\infty}M(\lambda u)/M_1(u)=0$ for any positive $\lambda$. In this case, following [9], we write $M\prec M_1$.

It is straightforward to verify that for the functions $\varphi(u)$ and $\psi(u)$ defined above we have $\psi\prec\varphi$.

The following theorem is true.

Theorem 3 (see [9], Theorem 2). Let $\Omega$ be a bounded domain in ${\mathbb R}^n$ satisfying the cone condition. Then the Sobolev space $W^k_p(\Omega)$, where $n=kp$, is continuously embedded in the Orlicz space $L_\Phi^*(\Omega)$, where $\Phi(t)=\exp(|t|^{n/(n-1)})-1$. In addition, for any $N$-function $\gamma(t)$ such that $\gamma\prec\Phi$ this embedding in $L_\gamma^*(\Omega)$ is compact.

We need this theorem in the two-dimensional case, for $k=1$ and $p=2$.

2.2. Quasipotential operators

Let $E$ be a real Banach space and let $E^*$ be the dual space of $E$. We let $\langle z,x \rangle$ denote the value of the functional $z\in E^*$ at an element $x\in E$. In what follows, an operator $T$ acts from $E$ to $E^*$.

A functional $f\colon E\to\mathbb R$ is said to be Gateaux differentiable at a point $x\in E$ if there is $y\in E^*$ such that the limit $\lim_{t\to 0}(f(x+th)-f(x))/{t}=\langle y,h \rangle$ exists for each $h\in E$. In this case $y$ is called the Gateaux derivative of $f$ at $x$ and is denoted by $f'(x)$.

An operator $T$ is said to be potential if there exists a Gateaux differentiable functional $f$ on $E$ such that $f'(x)=Tx$ for any $x\in E$. A functional of this type is called a potential of the operator $T$.

An operator $T$ is said to be radially integrable if for any $x$, $h\in E$ the function $\langle T(x+th),h \rangle$ is integrable on $[0,1]$.

An operator $T$ is said to be radially continuous at a point $x\in E$ if for any $h\in E$ we have $\lim_{t\to 0}\langle T(x+th),h \rangle=\langle Tx, h\rangle$.

An element $x\in E$ is called a point of discontinuity of an operator $T$ if there is $h\in E$ such that either $\lim_{t\to 0}\langle T(x+th), h\rangle$ does not exist or ${\lim_{t\to 0}\langle T(x+th), h\rangle\neq\langle Tx, h\rangle}$.

A radially integrable operator $T$ is said to be quasipotential (see [25], Ch. 5, § 17, Definition 17.15) if there exists a functional $f\colon E\,{\to}\,{\mathbb R}$ such that

$$ \begin{equation} f(x+h)-f(x)=\int^1_0\langle T(x+th), h\rangle\, dt \quad \forall\, x, h\in E. \end{equation} \tag{2.4} $$
In this case $f$ is called a quasipotential of $T$.

Note that if an operator $T$ is potential and radially continuous on $E$ and $f$ is a potential of it, then for all $x , h\in E$ we have

$$ \begin{equation*} f(x+h)-f(x)=\int_0^1\frac{d}{dt}f(x+th)\,dt=\int_0^1\langle T(x+th),h\rangle \,dt. \end{equation*} \notag $$
However, a quasipotential operator can be discontinuous. For example, if $E=\mathbb R$, $Tx=\operatorname{sgn}x$, and $f(x)=|x|$, then (2.4) holds, but $T$ is discontinuous at $x=0$.

To state the variational principle to be used in the proof of Theorem 1 we need a further important notion.

An element $x\in E$ is called a regular point of an operator $T$ if for some $h\in E$ we have $\lim_{t\to +0}\sup\langle T(x+th), h\rangle<0$.

The following theorem is valid.

Theorem 4 (see [23], Theorem 1). Let $x\in E$ be a minimum point of a quasipotential $f$ of a locally bounded operator $T\colon E\to E^*$ whose points of discontinuity are regular. Then $x$ is a radial continuity point of $T$ and $Tx=0$.

We present a sufficient condition from [22] for the regularity of points of discontinuity of an operator $T$: if for any $x$, $h\in E$ the nonpositive limit ${\lim_{t\to +0}\langle T(x+th)-Tx,h\rangle}$ exists, then all points of discontinuity of $T$ are regular. Now we prove this. Let $x$ be a point of discontinuity of $T$. By assumption $\lim_{t\to +0}\langle T(x+th),h\rangle$ exists for any $h\in E$ and does not exceed $\langle Tx,h\rangle$. Since $x$ is a point of discontinuity, there exists $z\in E$ such that $\lim_{t\to +0}\langle T(x+tz),z\rangle<\langle Tx,z\rangle$. Otherwise $x$ is a point of radial continuity of $T$. Furthermore, if $\langle Tx,z\rangle\leqslant 0$, then $\lim_{t\to +0}\langle T(x+tz),z\rangle<0$ because of the last inequality; hence $z$ is a regular point of the operator $T$. If $\langle Tx,z\rangle>0$, then $\lim_{t\to +0}\langle T(x+t(-z)),-z\rangle$ exists and satisfies the inequalities $\lim_{t\to +0}\langle T(x+t(-z)),-z\rangle\leqslant \langle Tx,-z\rangle =-\langle Tx,z\rangle<0$. Therefore, $x$ is a regular point of $T$.

We give another result, which is useful for further consideration.

Theorem 5 (see [26]). Let $T=T_1+T_2$, where $T_1$ is a monotone operator (that is, $\langle T_1u-T_1v,u-v\rangle\geqslant 0$ for any $u,v\in E$), $T_2$ is a compact operator (that is, it maps bounded sets in $E$ to precompact sets in $E^*$), and both $T_i$, $i=1,2$, are quasipotential. Then a quasipotential $f$ of $T$ is weakly semicontinuous from below on $E$ (that is, for any $x\in E$ and any sequence $(x_n)\subset E$ weakly converging to $x$ the inequality $\lim_{n\to\infty}\inf f(x_n)\geqslant f(x))$ holds.

§ 3. Proof of Theorem 1

Let $E=\mathring{W}^1_2(\Omega)$ and $E^*=W^{-1}_2(\Omega)$. The space $E$ is a reflexive Hilbert space with inner product

$$ \begin{equation*} (u,v)=\int_\Omega \nabla u\cdot\nabla v\,dx \quad \forall\, u,v\in E, \end{equation*} \notag $$
which induces the norm $\|u\|=\sqrt{(u,u)}$.

We define operators $T_i$, $i=1,2$, on $E$ by

$$ \begin{equation*} \langle T_1u,v\rangle=\sum_{i,j=1}^2\int_\Omega a_{ij}(x)u_{x_i}v_{x_j}\,dx+\int_\Omega c(x)u(x)v(x)\,dx, \end{equation*} \notag $$
$$ \begin{equation} \langle T_2u,v\rangle=\int_\Omega g(x,u(x))v(x)\,dx \end{equation} \tag{3.1} $$
for any $u$, $v\in E$.

In the proof of Theorem 1 we adhere to the following five-part plan.

1) We establish that the operator $T_1$ is bounded, quasipotential, and monotone.

2) We prove that the operator $T_2$ is quasipotential and compact.

By Theorem 5 it follows from parts 1) and 2) that the quasipotential $J(u)$ of the operator $T=T_1-T_2$ is weakly semicontinuous from below on $E$.

3) We prove that the functional $J(u)$ is coercive, that is, $\lim_{\|u\|\to\infty}\!J(u)\,{=}\,{+}\infty$.

Since a functional in a reflexive Banach space that is weakly semicontinuous from below and coercive attains its global minimum (see [25], Ch. III, § 9, Remark 9.1), there exists $u_0\in E$ such that $J(u_0)=\inf\{J(u)\colon u\in E\}$.

4) We prove that points of discontinuity of $T$ are regular. To do this we verify that the conditions in the regularity test for points of discontinuity of the operator $T$ are fulfilled. The statement and proof of this test are presented in § 2.2.

Theorem 4 implies that $u_0$ is a point of radial continuity of the operator $T$ and $Tu_0=0$. By the definition of $T$, a function $z(x)\in E$ is a weak solution of problem (1.1), (1.2) if and only if $Tz=0$. Therefore, $u_0(x)$ is a weak solution of (1.1), (1.2).

5) We prove that, as the operator $T$ is radially continuous at the point $u_0$, the set $U=\{x\in\Omega\colon u_0(x)$ is a point of discontinuity of the function $g(x,\cdot\,)\}$ has measure zero. Thus, $u_0(x)$ is a weak semiregular solution of problem (1.1), (1.2).

This completes the proof of Theorem 1.

We switch to implementing the above plan of the proof.

The operator $T_1$ is linear, bounded and self-adjoint (that is, $\langle T_1 u,v\rangle=\langle T_1v,u\rangle$). Since $c(x)\geqslant 0$ on $\Omega$, we have

$$ \begin{equation} \langle T_1u,u\rangle\geqslant\chi \|u\|^2 \quad \forall\, u\in E, \end{equation} \tag{3.2} $$
where $\chi$ is the ellipticity constant of the differential operator $L$. It follows from (3.2), since $T_1$ is linear, that $T_1$ is a monotone operator. This operator is potential, and has a potential $f_1(u)=(1/2)\langle T_1u,u\rangle$ ([25], Ch. II, § 5, Example 5.6).

Now we realize part 2) of the plan. We need to establish that the operator $T_2$ acts from $E$ to $E^*$, is compact and quasipotential. First we prove that the Nemytskii operator $G(u)=g(x,u(x))$ maps bounded sets in $E_M(\Omega)$ (where $M(u)$ is the $N$-function in condition 2) of Theorem 1) to bounded sets in $L_N^*(\Omega)$, $N(u)$ is the complementary function of $M(u)$. As noted above, $N(u)$ satisfies the $\Delta_2$-condition. Therefore, $L_N^*=L_N=E_N$. Since $E_M^*=L_N^*$, we have $E_M^*=E_N$ in this case. We fix $r>0$ and show that there exist $a_r(x)\in L_N$ and $b_r>0$ such that for almost all $x\in\Omega$ the right-hand side of (1.4) in condition 2) of Theorem 1 does not exceed $a_r(x)+b_rN^{-1}(M(u/r))$ for any $u\in\mathbb R$. In fact,

$$ \begin{equation*} \begin{aligned} \, &\lim_{u\to\infty}\frac{M(u/r)}{(|u|/r)|u|^\alpha \exp(A|u|^\tau)}= \lim_{u\to\infty}\frac{|u|^{\alpha+2}\exp(|u/r|^s)}{r^{\alpha+2}|u|^\alpha \exp(A|u|^\tau)}\,\frac{r}{|u|} \\ &\qquad =\lim_{u\to\infty}\frac{|u|\exp(|u/r|^s)}{r^{\alpha+1}\exp(A|u|^\tau)}=+\infty, \end{aligned} \end{equation*} \notag $$
since $s>\tau$ by virtue of condition 2) in Theorem 1. It follows that there exist $a_r(x)\in L_N(\Omega)$ and $b_r>1$ such that
$$ \begin{equation} a(x)+b|u|^\alpha \exp(A|u|^\tau)\leqslant a_r(x)+\frac{b_rM(u/r)}{|u|/r} \quad \forall\, u\in\mathbb R \end{equation} \tag{3.3} $$
for almost all $x\in\Omega$. Due to (2.1), we have $N^{-1}(M(u/r))\geqslant M(u/r)/(|u|/r)$ for each $ u\in\mathbb R$. This fact and (3.3) yield
$$ \begin{equation*} a(x)+b|u|^\alpha \exp(A|u|^\tau)\leqslant a_r(x)+b_rN^{-1}\biggl(M\biggl(\frac ur\biggr)\biggr) \quad \forall\, u\in\mathbb R \end{equation*} \notag $$
for almost all $x\in\Omega$. Thus, we have shown that for any $r>0$ there exist $a_r(x)\in L_N(\Omega)$ and $b_r>1$ such that
$$ \begin{equation} |g(x,u)|\leqslant a_r(x)+b_rN^{-1}\biggl(M\biggl(\frac ur\biggr)\biggr) \quad \forall\, u\in\mathbb R \end{equation} \tag{3.4} $$
for almost all $x\in\Omega$. It follows from this inequality that the Nemytskii operator $G$ is bounded on the ball $B_r=\{u\in E_M\colon \|u\|_M\leqslant r\}$, that is, there is a constant $C_r>0$ such that $\|Gu\|_N\leqslant C_r$ for any $u\in B_r$. In fact, since $N(u)$ is increasing and convex on $\mathbb R$, for almost all $x\in\Omega$ we have
$$ \begin{equation*} N(g(x,u))\leqslant\frac12N(a_r(x))+\frac12N\biggl(b_rN^{-1}\biggl(M\biggl(\frac ur\biggr)\biggr)\biggr) \quad \forall\, u\in\mathbb R. \end{equation*} \notag $$
Since $N(u)$ satisfies the $\Delta_2$-condition, there exist a number $l(b_r)>0$ and $u_0> 0$ such that $N(b_rN^{-1}(M(u/r)))\leqslant l(b_r)N(N^{-1}(M(u/r)))=l(b_r)M(u/r)$ for ${|u|\,{\geqslant}\, u_0}$. From this we deduce that there exists an integrable function $d_r(x)$ on $\Omega$ such that
$$ \begin{equation} N(g(x,u))\leqslant d_r(x)+\frac12l(b_r)M\biggl(\frac ur\biggr) \quad \forall\, u\in\mathbb R \end{equation} \tag{3.5} $$
for almost all $x\in\Omega$. It follows that if $u(x)\in E_M$ and $\|u\|_M\leqslant r$, then
$$ \begin{equation*} \begin{aligned} \, \int_\Omega N(g(x,u(x))\,dx &\leqslant\int_\Omega d_r(x)\,dx+\frac12l(b_r)\int_\Omega M\biggl(\frac{u(x)}r\biggr)\,dx \\ &\leqslant\int_\Omega d_r(x)\,dx+\frac12l(b_r)\biggl\|\frac ur\biggr\|_M \leqslant\int_\Omega d_r(x)\,dx+\frac12l(b_r)=A_r \end{aligned} \end{equation*} \notag $$
by (3.5) (we have used (2.3) since $\|u/r\|_M\leqslant 1$). By virtue of (2.2), we have
$$ \begin{equation} \|Gu\|_N\leqslant\int_\Omega N(g(x,u(x))\,dx+1\leqslant A_r+1 \end{equation} \tag{3.6} $$
if $u\in B_r$. The boundedness of the operator $G$ on the ball $B_r$ is proved.

We show that equality (3.1) defines the operator $T_2\colon E\to E^*$ consistently and establish that it is compact. As noted above, $M\prec\varphi$, where $M(u)$ is the $N$-function in condition 2) of Theorem 1 and $\varphi(u)=\exp(u^2)-1$. Then, according to Theorem 3 the embedding of $W^1_2(\Omega)$ in $L_M^*(\Omega)$ is compact. We show that $W_2^1(\Omega)\subset E_M$. Since the bounded domain $\Omega\subset {\mathbb R}^2$ is of class $C^{1,1}$, the set $C^\infty(\overline\Omega)$ is dense in $W_2^1(\Omega)$. The space $E_M$ is the closure of the set of measurable bounded functions on $\Omega$ in the space $L_M^*$. Consequently, $E_M\supset C^\infty(\overline\Omega)$. Since $E_M$ is a closed subset in $L_M^*$, this inclusion implies that $E_M\supset \overline{C^\infty(\overline\Omega)}$, where the right-hand side is the closure of $C^\infty(\overline\Omega)$ in $L_M^*$. Since the embedding of $W_2^1(\Omega)$ in $L_M^*$ is compact, we have ${\overline{C^\infty(\overline\Omega)}\supset W_2^1(\Omega)}$. It follows that $E_M\supset W_2^1(\Omega)$. Since $E$ is a closed subspace in $W_2^1(\Omega)$, the embedding of $E$ in $E_M$ is compact. We denote the embedding operator of $E$ in $E_M$ by $P$. The dual operator $P^*$ embeds $E_M^*=E_N$ in ${E^*=W_2^{-1}(\Omega)}$. It is compact since $P$ is ([27], Ch. 4, § 6, Theorem 3). Owing to the theorem on the general form of a linear bounded functional on $E_M$ ([24], Ch. II, § 14, Theorem 14.2), for arbitrary $u$, $v\in E$ we have

$$ \begin{equation*} \int_\Omega g(x,u(x))v(x)\,dx=\langle G(Pu),Pv\rangle=\langle P^*GP(u),v\rangle. \end{equation*} \notag $$
We conclude that $T_2u=P^*GP(u)$ for all $u\in E$. This representation of $T_2$ implies its compactness, since $G$ maps bounded sets in $E_M$ to bounded sets in $E_M^*=E_N$ and the embedding operators $P$ and $P^*$ are compact.

It remains to prove that the operator $T_2$ is quasipotential. For this purpose it suffices to prove that $G$ is quasipotential. In fact, if $G$ is a quasipotential operator and $f$ is a quasipotential of it, then we have

$$ \begin{equation*} f(u+h)-f(u)=\int_0^1\langle G(u+th),h\rangle \,dt \quad \forall\, u,h\in E_M. \end{equation*} \notag $$
Then it is true for arbitrary $u$, $h\in E$ that
$$ \begin{equation*} \begin{aligned} \, &f(Pu+Ph)-f(Pu)=\int_0^1\langle G(Pu+tPh),Ph\rangle\, dt \\ &\qquad =\int_0^1\langle P^*GP(u+th),h \rangle \,dt=\int_0^1\langle T_2(u+th),h\rangle \,dt, \end{aligned} \end{equation*} \notag $$
which means that the operator $T_2$ is quasipotential and $f_2(u)=f(Pu)$ is a quasipotential of it. Now we prove that $G$ is a quasipotential operator. In calculations below we use the following three classical results:

We define a functional $f$ on $E_M$ by the equality

$$ \begin{equation} f(u)=\int_\Omega dx\int_0^{u(x)}g(x,s)\,ds\quad \forall\, u\in E_M. \end{equation} \tag{3.7} $$
For arbitrary $u$, $h\in E_M$ we have
$$ \begin{equation} \begin{aligned} \, \notag f(u+h)-f(u) &=\int_\Omega dx\int_{u(x)}^{u(x)+th(x)}g(x,s)\,ds \\ \notag &=\int_\Omega dx\int_0^1\frac{d}{dt}\int_0^{u(x)+th(x)}g(x,s)\,ds\,dt \\ &=\int_0^1dt\int_\Omega g(x,u(x)+th(x))h(x)\,dx=\int_0^1\langle G(u+th),h\rangle \,dt. \end{aligned} \end{equation} \tag{3.8} $$
The integrability of $\psi(x,t)=g(x,u(x)+th(x))h(x)$ on $\Omega\times [0,1]$ follows from Hölder’s inequality (see [24], Ch. II, § 9, Theorem 9.3)
$$ \begin{equation*} \biggl|\int_\Omega z(x)y(x)\,dx\biggr|\leqslant \|z\|_M\, \|y\|_N \end{equation*} \notag $$
for any $z\in L_M$ and $y\in L_N$ and from estimate (3.6). Note that we can verify that $f$ is finite on $E_M$ by making this calculation in the reverse order for $u=0$. It follows from (3.8) that the operator $G$ is quasipotential; a quasipotential of it is specified by (3.7). As mentioned above, this implies that $T_2$ is a quasipotential operator, and $f_2(u)=f(Pu)$ is a quasipotential of it. The second part of the plan is realized.

So, $T=T_1-T_2$ is a quasipotential operator and $J(u)=(1/2)\langle T_1u,u\rangle-f(Pu)$ is a quasipotential of it.

By the definition of the operators $T_1$ and $T_2$, a function $z(x)\in E$ is a weak solution of problem (1.1), (1.2) if and only if $Tz=0$. By Theorem 5, since the operator $T_1$ is monotone and $T_2$ is compact, the functional $J(u)$ is weakly semicontinuous from below on $E$.

We switch to part 3) of the plan. We prove that the functional $J(u)$ is coercive, that is, $\lim_{\|u\|\to\infty}J(u)=+\infty$. In fact, by condition 3) in Theorem 1, for an arbitrary $u\in E$ we have

$$ \begin{equation} \begin{aligned} \, \notag J(u) &=\frac12\langle T_1u,u\rangle-f(Pu)\geqslant\frac{\chi}{2}\|u\|^2-\int_\Omega dx\int_0^{u(x)}g(x,s)\,ds \\ \notag &\geqslant\frac{\chi}{2}\|u\|^2-\frac12\int_\Omega \biggl(ku^2(x)+d(x)|u|^\theta+d_1(x)\biggr)\,dx \\ &\geqslant \frac{\chi-k\|P_1\|^2}{2}\|u\|^2-\frac{\|d\|_{2/(2-\theta)}}{2}\|u\|_2^\theta-\frac{\|d_1\|_1}{2}, \end{aligned} \end{equation} \tag{3.9} $$
where $\chi$ is the ellipticity constant of the operator $L$ defined by (1.3), $P_1$ is the embedding operator of $E$ in $L_2(\Omega)$ and $\|\cdot\|_s$ is the norm in $L_s(\Omega)$.

By condition 3) in Theorem 1, the constant $\chi-k\|P_1\|^2$ is positive; in addition, $\|u\|_2\leqslant \|P_1\|\,\|u\|$ and $0<\theta<2$. By virtue of (3.9) we derive from this that $J(u)$ is coercive.

A weakly lower semicontinuous coercive functional in a reflexive Banach space is bounded below and attains its global minimum (see [25], Ch. III, § 9, Remark 9.1). Therefore, there is $u_0\in E$ such that $J(u_0)=\inf\{J(u)\colon u\in E\}$.

We switch to implementing part 4) of the plan. We prove that all points of discontinuity of the operator $T$ are regular. Then, according to Theorem 4, $Tu_0=0$ and $u_0$ is a point of radial continuity of $T$. A sufficient condition for the regularity of points of discontinuity of $T$ is the inequality $\lim_{t\to +0}\langle T(u+th)-Tu,h\rangle\leqslant 0$ $\forall\, u, h\in E$. For any $u$, $h\in E$ and $t\in (0,1)$ we have

$$ \begin{equation*} \langle T_2(u+th)-T_2u,h\rangle=\int_\Omega g(x,u(x)+th(x))h(x)\,dx-\int_\Omega g(x,u(x))h(x)\,dx. \end{equation*} \notag $$
By condition 1) in Theorem 1, $\lim_{t\to +0}g(x,u(x)+th(x))h(x)$ exists for almost all $x \in \Omega$ and is no smaller than $g(x,u(x))h(x)$. This implies that the following limit exists:
$$ \begin{equation} \lim_{t\to +0}\langle T_2(u+th)-T_2u,h\rangle\geqslant 0, \end{equation} \tag{3.10} $$
provided that we justify taking the limit under the integral sign. To use Lebesgue’s theorem we need to prove that, given $u(x)$ and $h(x)$ in $E$, there exists an integrable function $\psi(x)$ on $\Omega$ such that
$$ \begin{equation} |g(x,u(x)+th(x))h(x)|\leqslant\psi(x) \end{equation} \tag{3.11} $$
almost everywhere on $\Omega$ for any $t\in (0,1)$. We take $r=\bigl\||u|+|h|\bigr\|_M$. As shown above, there exist $a_r(x)\in L_N$ and $b_r>1$ such that (3.4) holds for almost all $x\in\Omega$ and any $u\in\mathbb R$. We set $\psi_1(x)=a_r(x)+b_rN^{-1}(M((|u(x)|+|h(x)|)/r))$, $x\in\Omega$. By virtue of (3.4), taking account of the fact that $N^{-1}$ and $M$ are increasing functions, we see that (3.11) is valid for $\psi(x)=\psi_1(x)|h(x)|$, for any $t\in (0,1)$. Note that $\psi_1(x)\in L_N(\Omega)$ because
$$ \begin{equation*} N(\psi_1(x))\leqslant d_r(x)+\frac12l(b_r)M\biggl(\frac{|u(x)|+|h(x)|}{r}\biggr) \end{equation*} \notag $$
and $u(x)$, $h(x)\in E_M$. Here $d_r(x)$ and $l(b_r)$ are the same as in (3.5). Since $h\in L_M$, it follows that $\psi(x)$ is integrable on $\Omega$ by Hölder’s inequality. Inequality (3.10) is proved.

From (3.10) and the radial continuity of $T_1$ we derive that $\lim_{t\to +0}\langle T(u+th)-Tu,h\rangle\leqslant 0$ for any $u$, $h\in E$. The regularity of points of discontinuity of the operator $T$ is established. According to Theorem 4, $Tu_0=0$ and $u_0$ is a point of radial continuity of $T$.

To complete the proof of Theorem 1 it suffices to prove that if $T$ is radially continuous at the point $u_0$, then the set $U=\{x\in \Omega\colon u_0(x) \text{ is a point of discontinuity of} \text{the function } g(x,\cdot\,)\}$ has measure zero (part 5) of the plan). In view of condition 1) in Theorem 1 the set $U$ coincides with $\{x\in\Omega\colon g(x,u_0(x)-)<g(x,u_0(x)+)\}$ up to a nullset. Assume that $\operatorname{mes} U\neq 0$. Then there exist $\varepsilon>0$ and $\delta>0$ such that $U(\varepsilon)=\{x\in \Omega\colon g(x,u_0(x)+)-g(x,u_0(x)-)>\varepsilon\}$ has measure $\delta$. We set $r=\|u_0\|_M+\|l\|_M$ ($l\equiv 1$ on $\Omega$) and $\widehat{\psi}(x)=a_r(x)+b_rN^{-1}(M((|u_0(x)|+1)/r))$, $x\in\Omega$, where $a_r(x)$ and $b_r$ are from (3.4). Like in the proof of the regularity of points of discontinuity of $T$, we can prove that $\widehat{\psi}(x)\in L_N$ and thus $\widehat{\psi}(x)$ is integrable on $\Omega$.

Since the Lebesgue integral is absolutely continuous (see [27], Ch. V, § 5, Theorem 5), there exists $\nu>0$ such that if $\omega$ is a measurable subset of $\Omega$ and $\operatorname{mes}\omega\leqslant\nu$, then

$$ \begin{equation*} \int_\omega\widehat{\psi}(x)\,dx<\frac{\varepsilon\delta}{8}. \end{equation*} \notag $$
The set $U(\varepsilon)\subset\Omega$ is measurable. Therefore, there exist a closed set $F\subset U(\varepsilon)$ and an open set $H \supset F$ with closure $\overline H\subset\Omega$ such that $\operatorname{mes}F>\operatorname{mes} U(\varepsilon)/2=\delta/2$ and $\operatorname{mes}(H\setminus F)<\nu$ (see [28]). Let $h \in C^\infty(\overline{\Omega})$ be equal to 1 on $F$, to 0 outside $H$, and ${0\leqslant h(x)\leqslant 1}$ for $x\in H\setminus F$ (such a function exists by virtue of the lemma in [29], Ch. 14, § 2). Note that $h\in E$. In view of (3.4) and the fact that the functions $N^{-1}$ and $M$ are increasing we see that
$$ \begin{equation} |g(x,u_0(x)+th(x))|\leqslant\widehat{\psi}(x) \end{equation} \tag{3.12} $$
for almost all $x\in\Omega$, for $t\in (-1,1)$. Owing to Lebesgue’s dominated convergence theorem, from condition 1) in Theorem 1 and (3.12) we obtain
$$ \begin{equation*} \begin{aligned} \, &\lim_{t\to +0}\int_\Omega g\bigl(x,u_0(x)+th(x)\bigr)h(x)\,dx \\ &\qquad=\int_F g(x,u_0(x)+)\,dx+\int_{H\setminus F}\lim_{t\to +0}g\bigl(x,u_0(x)+th(x)\bigr)h(x)\,dx, \\ &\lim_{t\to -0}\int_\Omega g\bigl(x,u_0(x)+th(x)\bigr)h(x)\,dx \\ &\qquad=\int_F g(x,u_0(x)-)\,dx+\int_{H\setminus F}\lim_{t\to -0}g\bigl(x,u_0(x)+th(x)\bigr)h(x)\,dx. \end{aligned} \end{equation*} \notag $$
Consequently, the difference between the left-hand sides of the last two equalities is greater than
$$ \begin{equation*} \frac{\varepsilon\delta}{2}-2\frac{\varepsilon\delta}{8}=\frac{\varepsilon\delta}{4}>0, \end{equation*} \notag $$
which contradicts the radial continuity of $T$ at $u_0$, since $T_1$ is a radially continuous operator. Theorem 1 is proved.


Bibliography

1. M. de Souza, E. de Medeiros and U. Severo, “On a class of quasilinear elliptic problems involving Trudinger-Moser nonlinearities”, J. Math. Anal. Appl., 403:2 (2013), 357–364  crossref  mathscinet  zmath
2. M. de Souza, E. de Medeiros and U. Severo, “On a class of nonhomogeneous elliptic problems involving exponential critical growth”, Topol. Methods Nonlinear Anal., 44:2 (2014), 399–412  crossref  mathscinet  zmath
3. M. de Souza, “Existence of solutions to equations of $N$-Laplacian type with Trudinger-Moser nonlinearities”, Appl. Anal., 93:10 (2014), 2111–2125  crossref  mathscinet  zmath
4. M. de Souza, “On a class of nonhomogeneous elliptic equation on compact Riemannian manifold without boundary”, Mediterr. J. Math., 15:3 (2018), 101, 11 pp.  crossref  mathscinet  zmath
5. M. de Souza, “On a class of nonhomogeneous elliptic equations on noncompact Riemannian manifolds”, Complex Var. Elliptic Equ., 64:3 (2019), 386–397  crossref  mathscinet  zmath
6. C. O. Alves and J. A. Santos, “Multivalued elliptic equation with exponential critical growth in $\mathbb R^2$”, J. Differential Equations, 261:9 (2016), 4758–4788  crossref  mathscinet  zmath  adsnasa
7. S. Carl and S. Heikkilä, “Elliptic problems with lack of compactness via a new fixed point theorem”, J. Differential Equations, 186:1 (2002), 122–140  crossref  mathscinet  adsnasa
8. J. Moser, “A sharp form of an inequality by N. Trudinger”, Indiana Univ. Math. J., 20:11 (1971), 1077–1092  crossref  mathscinet  zmath
9. N. S. Trudinger, “On imbeddings into Orlicz spaces and some applications”, J. Math. Mech., 17:5 (1967), 473–483  mathscinet  zmath
10. M. A. Krasnosel'skii and A. V. Pokrovskii, “Regular solutions of equations with discontinuous nonlinearities”, Dokl. Akad. Nauk SSSR, 226:3 (1976), 506–509  mathnet  mathscinet  zmath; English transl. in Soviet Math. Dokl., 17:1 (1976), 128–132
11. M. A. Krasnosel'skii and A. V. Pokrovskii, “Elliptic equations with discontinuous nonlinearities”, Dokl. Ross. Akad. Nauk, 342:6 (1995), 731–734  mathnet  mathscinet  zmath; English transl. in Dokl. Math., 51:3 (1995), 415–418
12. V. N. Pavlenko and D. K. Potapov, “The existence of semiregular solutions to elliptic spectral problems with discontinuous nonlinearities”, Mat. Sb., 206:9 (2015), 121–138  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 206:9 (2015), 1281–1298  crossref  adsnasa
13. M. A. Krasnosel'skii and A. V. Lusnikov, “Regular fixed points and stable invariant subsets of monotone operators”, Funktsional. Anal. Prilozhen., 30:3 (1996), 34–46  mathnet  crossref  mathscinet  zmath; English transl. in Funct. Anal. Appl., 30:3 (1996), 174–183  crossref
14. V. N. Pavlenko and D. K. Potapov, “Existence of two nontrivial solutions for sufficiently large values of the spectral parameter in eigenvalue problems for equations with discontinuous right-hand sides”, Mat. Sb., 208:1 (2017), 165–182  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 208:1 (2017), 157–172  crossref  adsnasa
15. V. N. Pavlenko and D. K. Potapov, “Existence of three nontrivial solutions of an elliptic boundary-value problem with discontinuous nonlinearity in the case of strong resonance”, Mat. Zametki, 101:2 (2017), 247–261  mathnet  crossref  mathscinet  zmath; English transl. in Math. Notes, 101:2 (2017), 284–296  crossref
16. V. N. Pavlenko and D. K. Potapov, “Properties of the spectrum of an elliptic boundary value problem with a parameter and a discontinuous nonlinearity”, Mat. Sb., 210:7 (2019), 145–170  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 210:7 (2019), 1043–1066  crossref  adsnasa
17. V. N. Pavlenko and D. K. Potapov, “On a class of elliptic boundary-value problems with parameter and discontinuous non-linearity”, Izv. Ross. Akad. Nauk Ser. Mat., 84:3 (2020), 168–184  mathnet  crossref  mathscinet  zmath; English transl. in Izv. Math., 84:3 (2020), 592–607  crossref  adsnasa
18. V. N. Pavlenko and D. K. Potapov, “On the existence of three nontrivial solutions of a resonance elliptic boundary value problem with a discontinuous nonlinearity”, Differ. Uravn., 56:7 (2020), 861–871  crossref  mathscinet  zmath; English transl. in Differ. Equ., 56:7 (2020), 831–841  crossref
19. V. N. Pavlenko and D. K. Potapov, “Existence of semiregular solutions of elliptic systems with discontinuous nonlinearities”, Mat. Zametki, 110:2 (2021), 239–257  mathnet  crossref  mathscinet  zmath; English transl. in Math. Notes, 110:2 (2021), 226–241  crossref
20. V. N. Pavlenko and D. K. Potapov, “Variational method for elliptic systems with discontinuous nonlinearities”, Mat. Sb., 212:5 (2021), 133–152  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 212:5 (2021), 726–744  crossref  adsnasa
21. V. N. Pavlenko, “On the solvability of some nonlinear equations with discontinuous operators”, Dokl. Akad. Nauk SSSR, 204:6 (1972), 1320–1323  mathnet  mathscinet  zmath; English transl. in Soviet Math. Dokl., 13 (1972), 846–850
22. V. N. Pavlenko, “Variational method for equations with discontinuous operators”, Vestnik Chelyabinsk. Gos. Univ., 1994, no. 2, 87–95 (Russian)  mathnet  mathscinet  zmath
23. V. N. Pavlenko and D. K. Potapov, “Existence of a ray of eigenvalues for equations with discontinuous operators”, Sibirsk. Mat. Zh., 42:4 (2001), 911–919  mathnet  mathscinet  zmath; English transl. in Siberian Math. J., 42:4 (2001), 766–773  crossref
24. M. A. Krasnosel'skii and Ya. B. Rutickii, Convex functions and Orlicz spaces, Fizmatgiz, Moscow, 1958, 271 pp.  mathscinet  zmath; English transl., P. Noordhoff Ltd., Groningen, 1961, xi+249 pp.  mathscinet  zmath
25. M. M. Vainberg, Variational method and method of monotone operators in the theory of nonlinear equations, Nauka, Moscow, 1972, 416 pp.  mathscinet  zmath; English transl., Halsted Press (A division of John Wiley & Sons), New York–Toronto, ON; Israel Program for Scientific Translations, Jerusalem–London, 1973, xi+356 pp.  mathscinet  zmath
26. V. N. Pavlenko, “Existence theorems for elliptic variational inequalities with quasipotential operators”, Differ. Uravn., 24:8 (1988), 1397–1402  mathnet  mathscinet  zmath; English transl. in Differ. Equ., 24:8 (1988), 913–916
27. A. N. Kolmogorov and S. V. Fomin, Introductory real analysis, 3rd ed., Nauka, Moscow, 1972, 496 pp.  zmath; English transl. of 2nd ed., Rev. ed., Prentice-Hall, Inc., Englewood Cliffs, NY, 1970, xii+403 pp.  mathscinet  zmath
28. V. N. Pavlenko, “The existence of solutions for nonlinear equations with discontinuous monotone operators”, Vestnik Moskov. Univ. Ser. 1 Mat. Mekh., 1973, no. 6, 21–29  mathscinet  zmath; English transl. in Moscow Univ. Math. Bull., 28:6 (1974), 70–77
29. N. Dunford and J. T. Schwartz, Linear operators, v. II, Spectral theory. Self adjoint operators in Hilbert space, Intersci. Publ. John Wiley & Sons, New York–London, 1963, ix+859–1923+7 pp.  mathscinet  zmath

Citation: V. N. Pavlenko, D. K. Potapov, “Semiregular solutions of elliptic boundary-value problems with discontinuous nonlinearities of exponential growth”, Mat. Sb., 213:7 (2022), 121–138; Sb. Math., 213:7 (2022), 1004–1019
Citation in format AMSBIB
\Bibitem{PavPot22}
\by V.~N.~Pavlenko, D.~K.~Potapov
\paper Semiregular solutions of elliptic boundary-value problems with discontinuous nonlinearities of exponential growth
\jour Mat. Sb.
\yr 2022
\vol 213
\issue 7
\pages 121--138
\mathnet{http://mi.mathnet.ru/sm9655}
\crossref{https://doi.org/10.4213/sm9655}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4461460}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2022SbMat.213.1004P}
\transl
\jour Sb. Math.
\yr 2022
\vol 213
\issue 7
\pages 1004--1019
\crossref{https://doi.org/10.4213/sm9655e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=000992267100004}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85165933499}
Linking options:
  • https://www.mathnet.ru/eng/sm9655
  • https://doi.org/10.4213/sm9655e
  • https://www.mathnet.ru/eng/sm/v213/i7/p121
  • This publication is cited in the following 1 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:305
    Russian version PDF:16
    English version PDF:51
    Russian version HTML:133
    English version HTML:67
    References:68
    First page:10
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024