1 Introduction

In recent years, a fractional Laplacian (more precisely, the fractional powers of the Laplacian) attracted attention of many scientists due to its applications in various areas. It appears including in probability (cf. [1, 6, 7, 13]), economics and finance (cf. [1, 14]), mechanics (cf. [5, 7]), optimal control theory (cf. [9, 20]), fluid mechanics and hydrodynamics (cf. [8, 10,11,12, 28,29,30]). There exist many definitions of such operators (e.g., Fourier transform [15, 17], hypersingular integral [15], Riesz potential operator [24], Bochner’s definition [27], spectral decomposition (cf. [4, 14, 18]).

In this paper, we are concerned with the study of solutions to the following differential inclusion

$$\begin{aligned} \sum \limits _{i=0}^k\alpha _i[(-{\varDelta })_\omega ]^{\beta _i}u(x)\in \partial _uF(x,u(x)),\quad x\in {\varOmega }\,\,\hbox {a.e.}, \end{aligned}$$
(1)

where \(\alpha _i>0\), \(i=0,\dots ,k\) (\(k\in {\mathbb {N}}\cup \{0\}\)), \(0\le \beta _0<\beta _1<\dots <\beta _k\), \({\varOmega }\subset {\mathbb {R}}^N\) is an open and bounded set, \(F:{\varOmega }\times {\mathbb {R}}\rightarrow [0,\infty )\), \(\partial _uF\) denotes a subgradient of F with respect to u, \([(-{\varDelta }_\omega )]^{\gamma }\) denotes a weak fractional Laplace operator of order \(\gamma >0\) with zero Dirichlet boundary values on \(\partial {\varOmega }\) (the term “weak” is explained in Sect. 2). This work is based on the spectral definition of the mentioned operator (we shall call it a weak fractional Dirichlet–Laplace operator). More precisely, the definition of the Dirichlet Laplacian comes from the functional calculus for unbounded self-adjoint operators in a Hilbert space [16, 21] and is based on the spectral integral representation theorem for such operators [23, 25].

Let us note that if \(F(t,\cdot )\) is differentiable on \({\mathbb {R}}\), then (1) reduces to the following boundary value problem

$$\begin{aligned} \sum \limits _{i=0}^k\alpha _i[(-{\varDelta })_\omega ]^{\beta _i}u(x)= D_uF(x,u(x)),\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$
(2)

Motivated by the paper [3], we characterize solutions of the problem (1). They are minimizers of a some integral functional J related to (1) due to the Legendre–Fenchel transform of the function F and the Fenchel–Young inequality (cf. Theorem 3). Using a standard method of the calculus of variations, we prove the existence and uniqueness of minimizers of J on an appropriate space of functions, being the space of solutions of (1) (cf. Theorem 4). Finally, obtained results are illustrated by some examples. In particular, we show that if F is adequately chosen, then the minimizer of J is a solution of (2). To the best of our knowledge, the problem of a type (1) was not investigated by other authors.

2 Preliminaries

In this section, we give some preliminary notions and results that will be used in the remaining part of this paper.

2.1 Weak Fractional Dirichlet–Laplace Operator

In a whole paper we assume that \({\varOmega }\subset {\mathbb {R}}^N\) is an open and bounded set.

Let \(T_0=-{\varDelta }:C^\infty _c({\varOmega },{\mathbb {R}})\subset L^2({\varOmega },{\mathbb {R}})\rightarrow L^2({\varOmega },{\mathbb {R}})\) be the classical Dirichlet–Laplace operator.

Definition 1

(cf. [21]) We say that \(u:{\varOmega }\rightarrow {\mathbb {R}}\) has a weak (minus) Dirichlet–Laplacian if \(u\in H^1_0({\varOmega },{\mathbb {R}})\) and there exists a function \(g\in L^2({\varOmega },{\mathbb {R}})\) such that

$$\begin{aligned} \int \limits _{{\varOmega }}\nabla u(x)\nabla v(x)\hbox {d}x=\int \limits _{{\varOmega }}g(x)v(x)\hbox {d}x \end{aligned}$$

for any \(v\in H^1_0({\varOmega },{\mathbb {R}})\). The function g will be called the weak Dirichlet–Laplacian and denoted by \((-{\varDelta })_\omega u\).

The operator \((-{\varDelta })_\omega \) is called the weak Dirichlet–Laplace operator and the set

$$\begin{aligned} \begin{aligned}&D((-{\varDelta })_\omega ) \\&\quad :=\left\{ u\in H^1_0({\varOmega },{\mathbb {R}});\quad \exists _{g\in L^2},\,\,\forall _{v\in H^1_{0}}\quad \int \limits _{{\varOmega }}\nabla u(x)\nabla v(x)\hbox {d}x=\int \limits _{{\varOmega }}g(x)v(x)\hbox {d}x\right\} \end{aligned} \end{aligned}$$

is named the domain of the operator \((-{\varDelta })_\omega \).

Using the Friedrich’s extension procedure, we obtain (cf. [21, Theorem 3.1], [23])

Theorem 1

The operator

$$\begin{aligned} (-{\varDelta })_\omega : D((-{\varDelta })_\omega )\subset L^2({\varOmega },{\mathbb {R}})\rightarrow L^2({\varOmega },{\mathbb {R}}) \end{aligned}$$

is the self-adjoint extension of the operator \(T_0\) and

$$\begin{aligned} T_0 \subset -{\varDelta }\subset (-{\varDelta })_\omega , \end{aligned}$$

where \(-{\varDelta }:H^1_0({\varOmega },{\mathbb {R}})\cap H^2({\varOmega },{\mathbb {R}})\subset L^2({\varOmega },{\mathbb {R}})\rightarrow L^2({\varOmega },{\mathbb {R}})\) is the (strong) Dirichlet–Laplace operator.

Remark 1

In [2, Section 8.2] \((-{\varDelta })_\omega \) is called the Laplace–Dirichlet operator (without “weak”) and denoted by \(-{\varDelta }\).

It is known (cf. [21]) that the spectrum \(\sigma ((-{\varDelta })_\omega )\) of \((-{\varDelta })_\omega \) contains only the eigenvalues of \((-{\varDelta })_\omega \) that can be written in a non-decreasing sequence, repeating each eigenvalue according to its multiplicity \(0<\lambda _1\le \lambda _2\le \dots \le \lambda _j\rightarrow \infty \). Moreover, a system \(\{e_j\}\) of eigenfunctions of the operator \((-{\varDelta })_\omega \), corresponding to \(\lambda _j\), is a hilbertian basis in \(L^2({\varOmega },{\mathbb {R}})\).

Remark 2

In paper [21], the author proved that if \({\varOmega }\) is an open and bounded set of class \(C^{1,1}\) or a convex polygon in \({\mathbb {R}}^2\), then the weak and strong Dirichlet–Laplace operators coincide.

Now, let \(\beta >0\). By the weak fractional Dirichlet–Laplace operator of order \(\beta \) we mean the operator \([(-{\varDelta })_{\omega }]^\beta :D([(-{\varDelta })_{\omega }]^\beta )\subset L^2({\varOmega },{\mathbb {R}})\rightarrow L^2({\varOmega },{\mathbb {R}})\) defined in the following way (cf. [21, Section 3]):

$$\begin{aligned} \left( [(-{\varDelta })_{\omega }]^\beta u\right) (t)=\left( \left( \int \limits _{\sigma ((-{\varDelta })_\omega )}\lambda ^\beta E(\hbox {d}\lambda )\right) u\right) (t)=\sum \limits _{j=1}^\infty (\lambda _j)^\beta a_j e_j(t), \end{aligned}$$

where

$$\begin{aligned} D\left( [(-{\varDelta })_{\omega }]^\beta \right)&=\left\{ u\in L^2({\varOmega },{\mathbb {R}});\quad \int \limits _{\sigma ((-{\varDelta })_\omega )}|\lambda ^\beta |^2\Vert E(\hbox {d}\lambda )u\Vert ^2=\sum \limits _{j=1}^\infty ((\lambda _j)^\beta )^2 a^2_j<\infty ,\right. \\&\quad \quad \hbox {where}\,\, a_j\text {-s}\,\, \hbox {is}\,\, \hbox {such}\,\, \hbox {that}\,\, u(t)=\left( \Bigg (\int \limits _{\sigma ((-{\varDelta })_\omega )}1 E(\hbox {d}\lambda )\Bigg )u\right) (t) \\&\left. =\sum \limits _{j=1}^\infty a_j e_j(t)\right\} \end{aligned}$$

(here E is the spectral measure given by \((-{\varDelta })_\omega \) and the convergence of the series is meant in \(L^2({\varOmega },{\mathbb {R}})\)).

It is well known that the operator \([(-{\varDelta })_{\omega }]^\beta \) is self-adjoint, its spectrum \(\sigma ([(-{\varDelta })_{\omega }]^\beta )\) contains only proper values \((\lambda _j)^\beta \), \(j\in {\mathbb {N}}\). Moreover, eigenspaces, corresponding to \((\lambda _j)^\beta \)-s and eigenspaces for \([(-{\varDelta })_{\omega }]\), corresponding to \(\lambda _j\)-s are the same.

In the space \(D([(-{\varDelta })_{\omega }]^\beta )\), we define two scalar products:

$$\begin{aligned} \langle u,v\rangle _{\beta }&:=\langle u,v\rangle _{L^2}+\left\langle [(-{\varDelta })_\omega ]^\beta u,[(-{\varDelta })_\omega ]^\beta v\right\rangle _{L^2}, \\ \langle u,v\rangle _{\sim \beta }&:=\left\langle [(-{\varDelta })_\omega ]^\beta u,[(-{\varDelta })_\omega ]^\beta v\right\rangle _{L^2} \end{aligned}$$

which generate norms:

$$\begin{aligned} \Vert u\Vert _{\beta }&=\left( \Vert u\Vert ^2_{L^2}+\Vert [(-{\varDelta })_\omega ]^\beta u\Vert ^2_{L^2}\right) ^{\frac{1}{2}}, \end{aligned}$$
(3)
$$\begin{aligned} \Vert u\Vert _{\sim \beta }&=\Vert [(-{\varDelta })_\omega ]^\beta u\Vert _{L^2}, \end{aligned}$$
(4)

respectively. Norms (3) and (4) are equivalent due to the following Poincaré inequality in \(D([(-{\varDelta })_{\omega }]^\beta )\) (cf. [21, inequality (3.2)]):

$$\begin{aligned} \Vert u\Vert ^2_{L^2}\le M_\beta \Vert u\Vert ^2_{\sim \beta }, \end{aligned}$$
(5)

where

$$\begin{aligned} M_\beta ={\left\{ \begin{array}{ll} 1&{}\hbox {if}\quad \lambda _1\ge 1\\ \tfrac{1}{((\lambda _1)^\beta )^2}&{}\hbox {if}\quad \lambda _1<1 \end{array}\right. } \end{aligned}$$

(here \(\lambda _1>0\) is the first (the smallest) eigenvalue of the operator \((-{\varDelta })_\omega \)).

Now, let us consider the function

$$\begin{aligned} w:{\mathbb {R}}\ni \lambda \rightarrow {\left\{ \begin{array}{ll} 0&{}\lambda <0\\ \alpha _k\lambda ^{\beta _k}+\dots +\alpha _1\lambda ^{\beta _1}+ \alpha _0\lambda ^{\beta _0}&{}\lambda \ge 0, \end{array}\right. } \end{aligned}$$
(6)

where \(0\le \beta _0<\beta _1<\dots <\beta _k\), \(k\in {\mathbb {N}}\cup \{0\}\), \(\alpha _i>0\), \(i=0,\dots ,k\). Then, we can define the operator \(w((-{\varDelta })_{\omega }):D(w((-{\varDelta })_{\omega }))\subset L^2({\varOmega },{\mathbb {R}})\rightarrow L^2({\varOmega },{\mathbb {R}})\) as follows:

$$\begin{aligned} w((-{\varDelta })_{\omega })u:=\sum \limits _{i=0}^k\alpha _i((-{\varDelta })_\omega )^{\beta _i}u. \end{aligned}$$

The operator \(w((-{\varDelta })_{\omega })\) is self-adjoint and \(D(w((-{\varDelta })_{\omega }))=D(((-{\varDelta })_{\omega })^{\beta _k})\) (cf. [21]). Moreover, in \(D([(-{\varDelta })_{\omega }]^{\beta _k})\) one can introduce a new scalar product of the form

$$\begin{aligned} \langle u,v\rangle _{w((-{\varDelta })_{\omega })}:=\langle w((-{\varDelta })_{\omega }) u,w((-{\varDelta })_{\omega }) v\rangle _{L^2} \end{aligned}$$
(7)

which generates a norm

$$\begin{aligned} \Vert u\Vert _{w((-{\varDelta })_{\omega })}=\Vert w((-{\varDelta })_{\omega }) u\Vert _{L^2}. \end{aligned}$$
(8)

Norms (4) and (8) (the norm (8) will be used in the main part of this article) are equivalent in \(D([(-{\varDelta })_{\omega }]^{\beta _k})\) (cf. [21, Lemma 3.6]) and

$$\begin{aligned} \Vert u\Vert ^2_{\sim \beta _k}\le \frac{1}{\alpha _k^2}\Vert u\Vert ^2_{w((-{\varDelta })_{\omega })} \le \frac{C_1}{\alpha _k^2}\left( 1+kC_2^2\right) \Vert u\Vert ^2_{\sim \beta _k}, \end{aligned}$$
(9)

where \(C_1, C_2>0\) are constants from [21, Lemma 3.6]. Hence and from (5) we obtain the following Poincaré inequality in \(D([(-{\varDelta })_{\omega }]^{\beta _k})\):

$$\begin{aligned} \Vert u\Vert ^2_{L^2}\le \frac{1}{ M_\beta \alpha _k^2}\Vert u\Vert ^2_{w((-{\varDelta })_{\omega })}. \end{aligned}$$
(10)

Moreover, the space \(D(w((-{\varDelta })_{\omega }))=D(((-{\varDelta })_{\omega })^{\beta _k})\) with the scalar product (7) is complete (cf. [21, Lemma 3.6]).

2.2 Basic Facts from Convex Analysis

In this part we recall some basic definitions and facts concerning the convex analysis. More details can be found in [19].

Let \(f : {\mathbb {R}}^n\rightarrow {\mathbb {R}}\). We shall say that a vector \(y\in {\mathbb {R}}^n\) is a subgradient of f at \(x\in {\mathbb {R}}^n\) if

$$\begin{aligned} f(z)-f(x)\geqslant \langle y,z-x\rangle _{{\mathbb {R}}^n},\quad \hbox {for}\,\, \hbox {every}\,\, z\in {\mathbb {R}}^n. \end{aligned}$$

The set of all subgradients of f at x is called a subdifferential and is denoted by \(\partial f(x)\). If \(\partial f(x)\ne \emptyset \) then the function f is called subdifferentiable at x. If f is a convex function, then the subdifferential is a nonempty, convex and compact set. Moreover, if f is Gateaux differentiable at x, then \(\partial f(x)=\{\nabla f(x)\}\).

The function \(f^*:{\mathbb {R}}^n\rightarrow {\mathbb {R}}\cup \{\infty \}\) defined by

$$\begin{aligned} f^*(y):=\underset{x\in {\mathbb {R}}^n}{\sup }\{\langle y,x\rangle _{{\mathbb {R}}^n}-f(x)\} \end{aligned}$$

is called the Legendre–Fenchel transform of the function f.

In the main part of this paper we use the following facts:

$$\begin{aligned} f(x)\le g(x), \,\,x\in {\mathbb {R}}^n\quad \Longrightarrow \quad f^*(y)\ge g^*(y),\,\,y\in {\mathbb {R}}^n. \end{aligned}$$
(11)

Theorem 2

(Fenchel–Young inequality) Let \(f : {\mathbb {R}}^n\rightarrow {\mathbb {R}}\). Any points \(x,y\in {\mathbb {R}}^n\) satisfy the inequality

$$\begin{aligned} f(x)+f^*(y)\geqslant \langle y,x\rangle . \end{aligned}$$

Equality holds if and only if \(y\in \partial f(x)\).

In conclusion of this section we give an another useful fact (cf. [22, Part II, Lemma 4.3.1]).

Lemma 1

Let \({\varOmega }\subset {\mathbb {R}}^d\) be open, \(f:{\varOmega }\times {\mathbb {R}}^d\rightarrow {\mathbb {R}}\), with \(f(\cdot ,v)\) measurable for all \(v\in {\mathbb {R}}^d\), \(f(x,\cdot )\) continuous for almost all \(x\in {\varOmega }\), and

$$\begin{aligned} f(x,v)\ge -a(x)+b|v|^p,\quad x\in {\varOmega }\,\,\mathrm{a.e.},\,\,v\in {\mathbb {R}}^d, \end{aligned}$$

where \(a\in L^1({\varOmega },{\mathbb {R}})\), \(b\in {\mathbb {R}}\), \(p\ge 1\). Then, the functional \({\varPhi }:L^p({\varOmega },{\mathbb {R}}^d)\rightarrow {\mathbb {R}}\cup \{\infty \}\),

$$\begin{aligned} {\varPhi }(v):=\int \limits _{\varOmega }f(x,v(x))\mathrm{d}x \end{aligned}$$

is a sequentially lower semicontinuous on \(L^p({\varOmega },{\mathbb {R}}^d)\).

Remark 3

The above lemma can be also proved in case of the function f defined on \({\varOmega }\times {\mathbb {R}}^n\), where \(n\ne d\).

3 Differential Inclusion with Fractional Dirichlet–Laplace Operators

In this section we prove the main result of this paper, namely a theorem on the existence of minimizers of a some functional associated with the problem (1).

By a solution of inclusion (1) we mean a function \(u\in D(w((-{\varDelta })_\omega ))\) satisfying (1) a.e. on \({\varOmega }\).

Let us assume that F satisfies the following conditions:

(F1)::

\(F(\cdot ,{u})\) is measurable on \({\varOmega }\) for all \(u\in {\mathbb {R}}\) and \(F(x,\cdot )\) is convex on \({\mathbb {R}}\) for a.e. \(x\in {\varOmega }\).

(F2)::

\(F(x,\cdot )\) is coercive on \({\mathbb {R}}\) for a.e. \(x\in {\varOmega }\), i.e.

$$\begin{aligned} \underset{|u|\rightarrow \infty }{\lim }\tfrac{F(x,u)}{|u|}=\infty ,\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$
(12)
(F3)::

there exist a function \(v\in L^1({\varOmega },{\mathbb {R}})\) and a constant \(c>0\) such that

$$\begin{aligned} F(x,u)\le v(x)+c|u|^2,\quad x\in {\varOmega }\,\,\hbox {a.e.}, u\in {\mathbb {R}}. \end{aligned}$$
(13)

Remark 4

Of course, since \(F(x,\cdot )\) is convex, therefore it is continuous on \({\mathbb {R}}\) for a.e. \(x\in {\varOmega }\) (cf. [26, Corollary 10.1.1]).

Let us consider the function \(f:{\mathbb {R}}\rightarrow {\mathbb {R}}\) given by

$$\begin{aligned} f(u):=v+c|u|^2, \end{aligned}$$
(14)

where \(c>0\) and \(v\in {\mathbb {R}}\) are fixed. Then, it easy to check that

$$\begin{aligned} f^*(z)= -v+\frac{1}{4c}|z|^2,\quad z\in {\mathbb {R}}. \end{aligned}$$
(15)

In the next result we shall use the following

Lemma 2

Let \(G:{\varOmega }\times {\mathbb {R}}\rightarrow {\mathbb {R}}\) be a Carathéodory function and the function \(H:{\varOmega }\rightarrow {\mathbb {R}}\cup \{\infty \}\) is defined by

$$\begin{aligned} H(x):=\underset{w\in {\mathbb {R}}}{\sup }\{G(x,w)\},\quad x\in {\varOmega }\,\,\mathrm{a.e.} \end{aligned}$$

Then H is measurable on \({\varOmega }\).

Proof

From the fact that \(G(\cdot ,w)\) is continuous on \({\mathbb {R}}\) it follows that

$$\begin{aligned} H(x)=\underset{w\in {\mathbb {Q}}}{\sup }\{G(x,w)\},\quad x\in {\varOmega }\,\,\hbox {a.e.}, \end{aligned}$$

where \({\mathbb {Q}}\) denotes the set of rational numbers. Since the supremum of the countable set of measurable functions is a measurable set, therefore H is measurable on \({\varOmega }\).\(\square \) \(\square \)

Now, we formulate and prove the following result.

Proposition 1

Let us assume that conditions (F1)–(F3) are satisfied. Then, the Legendre–Fenchel transform of the function F with respect to u \(F^*:{\varOmega }\times {\mathbb {R}}\rightarrow (-\infty ,\infty ]\) given by

$$\begin{aligned} F^*(x,z):=\underset{u\in {\mathbb {R}}}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-F(x,u)\right] ,\quad x\in {\varOmega }\,\,\mathrm{a.e.},\,\,z\in {\mathbb {R}} \end{aligned}$$

has the following properties:

  1. (a)

    \(F^*(x,z)\) is finite for a.e. \(x\in {\varOmega }\) and all \(z\in {\mathbb {R}}\),

  2. (b)

    \(F^*(\cdot ,z)\) is measurable on \({\varOmega }\) for all \(z\in {\mathbb {R}}\) and \(F^*(x,\cdot )\) is continuous on \({\mathbb {R}}\) for a.e. \(x\in {\varOmega }\),

  3. (c)

    for a.e. \(x\in {\varOmega }\) and all \(z\in {\mathbb {R}}\)

    $$\begin{aligned} F^*(x,z)\ge -v(x)+\frac{1}{4c}|z|^2, \end{aligned}$$
    (16)

    where the constant c and the function v are from assumption (F3).

Proof

proof of the property (a): Let us fix any \(z\in {\mathbb {R}}\) and \(x\in S\subset {\varOmega }\), where \(\mu ({\varOmega }{\setminus } S)=0\) (\(\mu \) denotes the Lebesque measure on \({\mathbb {R}}^N\)). From (F2) it follows that for any \(R>0\) there exists \(r>0\) such that for \(|u|>r\) we have \(F(x,u)\ge R|u|\). In particular, putting \(R=|z|\), we get

$$\begin{aligned} F(x,u)\ge |z||u|,\quad |u|>r. \end{aligned}$$

Consequently,

$$\begin{aligned} \langle u,z\rangle _{{\mathbb {R}}}-F(x,u)\le |u||z|-R|u|=0,\quad |u|>r. \end{aligned}$$

Hence and from continuity of the function \(F(\cdot ,u)\) we obtain

$$\begin{aligned} F^*(x,z)&=\underset{u\in {\mathbb {R}}}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-F(x,u)\right] \\&=\max \left\{ \underset{|u|\le r}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-F(x,u)\right] ,\underset{|u|>r}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-F(x,u)\right] \right\} \\&\le \max \{\hbox {const},0\}<\infty . \end{aligned}$$

This means that \(F^*\) is finite for a.e. \(x\in {\varOmega }\) and all \(z\in {\mathbb {R}}\).

proof of the property (b): Measurability of \(F^*(\cdot ,z)\) on \({\varOmega }\) for any fixed \(z\in {\mathbb {R}}\) follows from Lemma 2. From the definition of the Legendre–Fenchel transform it follows that \(F^*(x,\cdot )\) is convex on \({\mathbb {R}}\) for a.e. \(x\in {\varOmega }\). Thus and from the proved condition (a) of this proposition we obtain continuity of \(F^*\) with respect to z for a.e. \(x\in {\varOmega }\) (cf. [26, Corollary 10.1.1]).

proof of the property (c): The condition (16) follows from assumption (F3), property (11) and equality (15). \(\square \)

Now, let us consider the functional \(J:D(w((-{\varDelta })_\omega ))\longrightarrow {\mathbb {R}}\cup \{\infty \}\) given by:

$$\begin{aligned} J(u):=\int \limits _{\varOmega }\big (F(x,u(x))+F^*(x,w((-{\varDelta })_\omega )u(x))-\langle u(x),w((-{\varDelta })_\omega )u(x)\rangle _{\mathbb {R}}\big )\hbox {d}x. \end{aligned}$$
(17)

Let us note that since all terms in the above integral are measurable on \({\varOmega }\) (the first and third terms are even summable on \({\varOmega }\)), therefore J is well defined.

In the proof of the main result we use the following fact.

Lemma 3

If assumptions (F1)–(F3) are satisfied, then the functional

\(I:D(w((-{\varDelta })_{\omega }))\rightarrow {\mathbb {R}}\cup \{\infty \}\) given by

$$\begin{aligned} I(u)=\int \limits _{{\varOmega }}F^*(x,w((-{\varDelta })_{\omega })u(x)))\mathrm{d}x \end{aligned}$$
(18)

is sequentially weakly lower semicontinuous on \(D(w((-{\varDelta })_{\omega }))\).

Proof

Let us consider the functional \(I_1:L^2({\varOmega },{\mathbb {R}})\rightarrow {\mathbb {R}}\cup \{\infty \}\) given by

$$\begin{aligned} I_1(z)=\int \limits _{{\varOmega }}F^*(x,z(x))\hbox {d}x. \end{aligned}$$

From Proposition 1, Lemma 1 and Remark 3 it follows that \(I_1\) is a sequentially lower semicontinuous functional on \(L^2({\varOmega },{\mathbb {R}})\).

Let \((u_n)_{n\in {\mathbb {N}}}\subset D(w((-{\varDelta })_{\omega }))\) be a sequence convergent to a some element u in the space \(D(w((-{\varDelta })_{\omega }))\). Then, the sequence \((w((-{\varDelta })\omega )u_n)_{n\in {\mathbb {N}}}\subset L^2({\varOmega },{\mathbb {R}})\) is convergent to \(w((-{\varDelta })\omega )u\) in \(L^2({\varOmega },{\mathbb {R}})\). From the proved part of this Lemma it follows that the functional I is sequentially lower semicontinuous on \(D(w((-{\varDelta })_{\omega }))\). Since I is convex on \(D(w((-{\varDelta })_{\omega }))\) (\(F^*(x,\cdot )\) is convex on \({\mathbb {R}}\)), therefore it is sequential weak lower semicontinuous on \(D(w((-{\varDelta })_{\omega }))\).

The proof is completed. \(\square \)

The next theorem plays a key role in the study of the existence of solutions to problem (1).

Theorem 3

Assume that conditions (F1)–(F3) are satisfied. A function \(u_*\in D(w((-{\varDelta })_{\omega }))\) is a solution of the inclusion (1) if and only if \(u_*\) is a minimizer of J on \(D(w((-{\varDelta })_{\omega }))\) and \(J(u_*)=0\).

Proof

Let \(u_*\in D(w((-{\varDelta })_{\omega }))\) be a solution of the problem (1). From Theorem 2 it follows that

$$\begin{aligned}&F(x,u_*(x))+F^*(x,w((-{\varDelta })_{\omega })u_*(x))-\langle u_*(x),w((-{\varDelta })_\omega )u_*(x)\rangle _{\mathbb {R}}=0, \\&\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

Consequently, \(J(u_*)=0\). On the other hand, for every \(u\in D(w((-{\varDelta })_{\omega }))\), using once again Theorem 2, we have

$$\begin{aligned} F(x,u(x))+F^*(x,w((-{\varDelta })_{\omega })u(x))-\langle u(x),w((-{\varDelta })_\omega )u(x)\rangle _{\mathbb {R}}\ge 0,\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$
(19)

This means that \(J(u)\ge 0\) for every \(u\in D(w((-{\varDelta })_{\omega }))\), so \(u_*\) is a minimizer of J on \(D(w((-{\varDelta })_{\omega }))\).

Now, we assume that \(u_*\) is a minimizer of J on \(D(w((-{\varDelta })_{\omega }))\) and \(J(u_*)=0\). This means in particular that

$$\begin{aligned} 0= & {} J(u_*) \\= & {} \int \limits _{\varOmega }\big (F(x,u_*(x))+F^*(x,w((-{\varDelta })_\omega ) u_*(x))-\langle u_*(x),w((-{\varDelta })_\omega )u_*(x)\rangle _{\mathbb {R}}\big )\hbox {d}x. \end{aligned}$$

On the other hand, from Theorem 2 it follows that

$$\begin{aligned}&F(x,u_*(x))+F^*(x,w((-{\varDelta })_{\omega })u_*(x))-\langle u_*(x),w((-{\varDelta })_\omega )u_*(x)\rangle _{\mathbb {R}}\ge 0, \\&\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

Consequently,

$$\begin{aligned}&F(x,u_*(x))+F^*(x,w((-{\varDelta })_{\omega })u_*(x))-\langle u_*(x),w((-{\varDelta })_\omega )u_*(x)\rangle _{\mathbb {R}}=0, \\&\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

The second part of Theorem 2 guaranties that \(u_*\in D(w((-{\varDelta })_{\omega }))\) is a solution of the inclusion (1).

The proof is completed. \(\square \)

Remark 5

From the above theorem it follows that if \(u_*\in D(w((-{\varDelta })_{\omega }))\) is a solution of the problem (1) then it is a minimizer of J. Consequently, only minimizers of J can be solutions of (1) (they are in effect solutions of (1) if additionally \(J(u_*)=0\), where \(u_*\) denotes a minimizer of J).

Now, we formulate and prove the main result of this paper, namely a theorem on the existence of a minimizer of the functional J on \(D(w((-{\varDelta })_{\omega }))\). We have

Theorem 4

Assume that conditions (F1)–(F3) are satisfied and

$$\begin{aligned} c<\frac{\sqrt{M_{\beta _k}}\alpha _k}{4}, \end{aligned}$$
(20)

where \(\lambda _1\) is the first eigenvalue of \((-{\varDelta })_\omega \) and c is a constant from the assumption (F3). Then there exists a minimizer \(u_*\in D(w((-{\varDelta })_{\omega }))\) of the functional J. Moreover, if \(F(x,\cdot )\) is strictly convex on \({\mathbb {R}}\) for a.e. \(x\in {\varOmega }\), then the minimizer is unique.

Proof

Since the space \(D(w((-{\varDelta })_{\omega }))\) is reflexive (as the Hilbert space), therefore it is sufficient to show that the functional J is coercive and sequentially weakly lower semicontinuous on \(D(w((-{\varDelta })_{\omega }))\).

Coercivity: Let \((u_l)_{l\in {\mathbb {N}}}\subset D(w((-{\varDelta })_{\omega }))\) be any sequence such that

$$\begin{aligned} \Vert u_l\Vert _{w((-{\varDelta })_{\omega })} \underset{l\rightarrow \infty }{\longrightarrow }\infty . \end{aligned}$$

From the proof of Theorem 3 it follows that \(J(u_l)\ge 0\). Moreover, using nonnegativity of F, conditions (16), (20), the Hölder inequality and the Poincaré inequality (10) we obtain

$$\begin{aligned} J(u_l)=&\int \limits _{\varOmega }\big (F(x,u_l(x))+F^*(x,w((-{\varDelta })_\omega ) u_l(x))-\langle u_l(x),w((-{\varDelta })_\omega )u_l(x)\rangle _{\mathbb {R}}\big )\hbox {d}x\\ \ge&\int \limits _{\varOmega }F^*(x,w((-{\varDelta })_\omega )u_l(x))\hbox {d}x-\int \limits _{\varOmega }\langle u_l(x),w((-{\varDelta })_\omega )u_l(x)\rangle _{\mathbb {R}}\hbox {d}x\\ \ge&-\int \limits _{\varOmega }v(x)\hbox {d}x+\frac{1}{4c}\Vert w((-{\varDelta })_\omega )u_l \Vert ^2_{L^2}-\Vert u_l\Vert _{L^2}\Vert w((-{\varDelta })_\omega )u_l\Vert _{L^2}\\ \ge&\,C+\frac{1}{4c}\Vert u_l\Vert ^2_{w((-{\varDelta })_{\omega })}- \frac{1}{\sqrt{M_{\beta _k}}\alpha _k}\Vert u_l\Vert ^2_{w((-{\varDelta })_{\omega })}\\ =&\,C+\left( \frac{1}{4c}-\frac{1}{\sqrt{M_{\beta _k}}\alpha _k}\right) \Vert u_l\Vert ^2_{w((-{\varDelta })_{\omega })} \underset{l\rightarrow \infty }{\longrightarrow }\infty , \end{aligned}$$

where \(C=-\int \limits _{\varOmega }v(x)\hbox {d}x\). This means that J is coercive.

Sequential weak lower semicontinuity of J :  Assume that \((u_l)_{l\in {\mathbb {N}}}\subset D(w((-{\varDelta })_{\omega }))\) is a sequence such that

$$\begin{aligned} u_{l}\underset{l\rightarrow \infty }{\rightharpoonup }u_*\quad \hbox {weakly}\,\,\hbox {in}\,\, D(w((-{\varDelta })_{\omega })) \end{aligned}$$

and let

$$\begin{aligned} J(u)=J_1(u)+J_2(u)-J_3(u), \end{aligned}$$

where

$$\begin{aligned} J_1(u)= & {} \int \limits _{\varOmega }F(x,u(x))\hbox {d}x,\quad J_2(u)=\int \limits _{\varOmega }F^*(x,w((-{\varDelta })_\omega )u(x))\hbox {d}x, \\ J_3(u)= & {} \int \limits _{\varOmega }\langle u(x),w((-{\varDelta })_\omega )u(x)\rangle _{\mathbb {R}}\hbox {d}x. \end{aligned}$$

First, let us note that since F is a Carathéodory function, therefore \((F(\cdot ,u_{l}(\cdot )))_{l\in {\mathbb {N}}}\) is the sequence of measurable functions. Moreover all terms of this sequence are nonnegative. Consequently, using Fatou’s Lemma, we conclude

$$\begin{aligned} \int \limits _{{\varOmega }}\underset{l\rightarrow \infty }{\liminf }F(x,u_{l}(x)) \hbox {d}x\le \underset{l\rightarrow \infty }{\liminf }\int \limits _{{\varOmega }}F(x,u_{l}(x))\hbox {d}x. \end{aligned}$$
(21)

[21, Proposition 3.10] implies

$$\begin{aligned} u_{l}\underset{l\rightarrow \infty }{\longrightarrow }u_*\quad \hbox {strongly}\,\,\hbox {in}\,\,L^2({\varOmega },{\mathbb {R}}). \end{aligned}$$
(22)

So, there exists a subsequent \((u_{l_j})_{j\in {\mathbb {N}}}\) such that

$$\begin{aligned} u_{l_j}(x)\underset{j\rightarrow \infty }{\longrightarrow }u_*(x),\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

Hence and from the fact that F is continuous with respect to the second variable we get

$$\begin{aligned} \underset{j\rightarrow \infty }{\lim }F(x,u_{l_j}(x))=F(x,u_{*}(x)),\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

Supposing contrary and repeating the above argumentation we assert that

$$\begin{aligned} \underset{l\rightarrow \infty }{\lim }F(x,u_{l}(x))=F(x,u_{*}(x)),\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

Consequently, the inequality (21) can be written as

$$\begin{aligned} \int \limits _{{\varOmega }}F(x,u_{*}(x))\hbox {d}x\le \underset{l\rightarrow \infty }{\liminf }\int \limits _{{\varOmega }}F(x,u_{l}(x))\hbox {d}x. \end{aligned}$$

This means that \(J_1\) is sequentially weakly lower semicontinuous on \(D(w((-{\varDelta })_{\omega }))\). The functional \(J_2\) has also such a property due to Lemma 3.

Now, we show that \(J_3\) is sequentially weakly continuous on \(D(w((-{\varDelta })_{\omega }))\). Indeed, using once again [21, Proposition 3.10] we assert that

$$\begin{aligned} w((-{\varDelta })_\omega ) u_{l}\underset{l\rightarrow \infty }{\rightharpoonup }w ((-{\varDelta })_\omega )u_*\quad \hbox {weakly}\,\,\hbox {in}\,\, L^2({\varOmega },{\mathbb {R}}). \end{aligned}$$

This fact and convergence (22) imply

$$\begin{aligned} \underset{l\rightarrow \infty }{\lim }\int \limits _{\varOmega }\langle u_l(x),w((-{\varDelta })_\omega )u_l(x)\rangle _{\mathbb {R}}\hbox {d}x=\int \limits _{\varOmega }\langle u_*(x),w((-{\varDelta })_\omega )u_*(x)\rangle _{\mathbb {R}}\hbox {d}x. \end{aligned}$$

Finally, we conclude that J is sequentially weakly lower semicontinuous on \(D(w((-{\varDelta })_{\omega }))\).

Uniqueness of a solution: It is clear that since F and \(F^*\) are convex with respect to the second variable, therefore J is convex on \(D(w((-{\varDelta })_{\omega }))\). Let us suppose that \(u_*\) and \(v_*\) are two different minimizers of J on \(D(w((-{\varDelta })_{\omega }))\) and assume that \(F(x,\cdot )\) is strictly convex on \({\mathbb {R}}\) for a.e. \(x\in {\varOmega }\). This implies that \(J_1\), so also J, are strictly convex on \(D(w((-{\varDelta })_{\omega }))\). This means in particular that

$$\begin{aligned} J\left( \frac{u_*+v_*}{2}\right) <J(u_*)=J(v_*). \end{aligned}$$

This contradicts the assumption that \(u_*\) (\(v_*\)) is a minimizer of J on \(D(w((-{\varDelta })_{\omega }))\).

The proof is completed. \(\square \)

Remark 6

Let us note that only the order \(\beta _k\) and the value \(\alpha _k\) impact the existence of a minimizer of J. This is a consequence of the fact that domains \(D(((-{\varDelta })_\omega )^{\beta _k})\) and \(D(w((-{\varDelta })_{\omega }))\) coincide.

4 Illustrative Examples

In this section we present two simple theoretical examples which illustrate results obtained in Sect. 3.

Example 1

Let us consider the following problem:

$$\begin{aligned} ((-{\varDelta })_\omega )^{\frac{1}{3}}u(x)+ 5((-{\varDelta })_\omega )^{\frac{1}{2}}u(x)\in \partial _u F(x,u(x)),\quad x=(x_1,x_2)\in {\varOmega }\,\,\hbox {a.e.}, \end{aligned}$$
(23)

where \({\varOmega }=(0,\pi )\times (0,\pi )\subset {\mathbb {R}}^2\) and \(F:{\varOmega }\times {\mathbb {R}}\rightarrow [0,\infty )\) is given by

$$\begin{aligned} F(x,u):={\left\{ \begin{array}{ll} |u|&{}\hbox {if}\,\,|u|\le 1\\ u^2&{}\hbox {if}\,\,|u|>1. \end{array}\right. } \end{aligned}$$

It is easy to check that

$$\begin{aligned} \partial _u F(x,u)={\left\{ \begin{array}{ll} {[}-2,-1]&{}\hbox {if}\,\,u=-\,1\\ \{-1\}&{}\hbox {if}\,\,u\in (-\,1,0)\\ {[}-\,1,1]&{}\hbox {if}\,\,u=0\\ \{1\}&{}\hbox {if}\,\,u\in (0,1)\\ {[}1,2]&{}\hbox {if}\,\,u=1\\ \{2u\}&{}\hbox {if}\,\,|u|>1. \end{array}\right. } \end{aligned}$$

Moreover,

$$\begin{aligned} F^*(x,z)&= \underset{u\in {\mathbb {R}}}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-F(x,u)\right] \\&=\max \left\{ \underset{|u|\le 1}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-|u|\right] ,\underset{|u|>1}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-u^2\right] \right\} . \end{aligned}$$

We check that

$$\begin{aligned} \underset{|u|\le 1}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-|u|\right] = {\left\{ \begin{array}{ll} -z-1&{}\hbox {if}\,\,z< -1\\ z-1&{}\hbox {if}\,\,z>1\\ 0&{}\hbox {if}\,\,|z|\le 1 \end{array}\right. } \end{aligned}$$

and

$$\begin{aligned} \underset{|u|>1}{\sup }\left[ \langle u,z\rangle _{{\mathbb {R}}}-u^2\right] = {\left\{ \begin{array}{ll} -z-1&{}\hbox {if}\,\,z\in [-2,0]\\ z-1&{}\hbox {if}\,\,z\in [0,2]\\ \frac{1}{4}z^4&{}\hbox {if}\,\,|z|>2. \end{array}\right. } \end{aligned}$$

Consequently,

$$\begin{aligned} F^*(x,z) = {\left\{ \begin{array}{ll} 0&{}\hbox {if}\,\,|z|<1\\ -z-1&{}\hbox {if}\,\,z\in [-2,-1]\\ z-1&{}\hbox {if}\,\,z\in [1,2]\\ \frac{1}{4}z^4&{}\hbox {if}\,\,|z|>2. \end{array}\right. } \end{aligned}$$

Of course assumptions (F1)–(F3) are satisfied. In particular,

$$\begin{aligned} F(x,u)\le 1+|u|^2, \end{aligned}$$

so the condition (13) holds with \(c=1\). Moreover, it is well known (cf. [2, Proposition 8.5.3]) that the first eigenvalue of \((-{\varDelta })_\omega \) equals to 2, so the assumption (20) is also satisfied. Consequently, using Theorem 4, we assert that there exists a minimizer (not necessarily unique) of the functional J given by (17) which can be the only possible solution of the problem (23) (cf. Remark 5).

Example 2

Let us consider the following boundary value problem:

$$\begin{aligned} \sum \limits _{i=0}^k\alpha _i((-{\varDelta })_\omega )^{\beta _i}u(x)=D_uF(x,u(x)),\quad x\in {\varOmega }\,\,\hbox {a.e.}, \end{aligned}$$
(24)

where \(\alpha _i>0\), \(i=0,\dots ,k\) (\(k\in {\mathbb {N}}\cup \{0\}\)), \(0\le \beta _0<\beta _1<\dots <\beta _k\), \({\varOmega }\subset {\mathbb {R}}^N\), is any open and bounded set, \(F:{\varOmega }\times {\mathbb {R}}\rightarrow [0,\infty )\) is given by

$$\begin{aligned} F(x,u):=\frac{a(x)}{p}|u|^p, \end{aligned}$$

where \(1<p<2\) and \(a\in L^\infty ({\varOmega },{\mathbb {R}}_+)\).

We shall show that the above problem has the only trivial solution. Indeed, first let us note that

$$\begin{aligned} F^*(x,z)=\frac{1}{qa(x)}|z|^q,\quad \tfrac{1}{p}+\tfrac{1}{q}=1 \end{aligned}$$

and the partial derivative of F with respect to u is given by

$$\begin{aligned} D_u F(x,u)=a(x)|u|^{p-2}u. \end{aligned}$$

Moreover, assumptions (F1)–(F3) hold. In particular, for any \(c>0\) there exists a sufficiently large constant \(R>0\) (dependent on the constants cp and the function a) such that

$$\begin{aligned} \frac{a(x)}{p}|u|^p\le \frac{\Vert a\Vert _{L^\infty }}{p}|u|^p\le {\left\{ \begin{array}{ll} c|u|^2,&{}|u|>R\\ \frac{\Vert a\Vert _{L^\infty }}{p}R^p+c|u|^2,&{}|u|\le R, \end{array}\right. } \end{aligned}$$

so, the condition (13) is satisfied for any \(c>0\).

Consequently, since all assumptions of Theorem 4 are satisfied and F is strictly convex with respect to u for a.e. \(x\in {\varOmega }\), therefore the functional J given by

$$\begin{aligned} J(u)=\int \limits _{\varOmega }\left( \frac{a(x)}{p}|u(x)|^p+\frac{1}{qa(x)} \Big |[w((-{\varDelta })_\omega )u(x)]\Big |^q-u(x)w((-{\varDelta })_\omega )u(x)\right) \hbox {d}x \end{aligned}$$
(25)

has a unique minimizer \(u_*\in D(w((-{\varDelta })_{\omega }))\).

On the other hand it is clear that \(u_*=0\) is a solution of the problem (24), so it must minimize J on \(D(w((-{\varDelta })_{\omega }))\). This means that \(u_*=0\) is the only solution of (24).

Remark 7

If \(p=2\) in the above example then the condition (13) holds for \(c=\frac{\Vert a\Vert _{L^\infty }}{2}\). Consequently, the linear problem of the form

$$\begin{aligned} \sum \limits _{i=0}^k\alpha _i((-{\varDelta })_\omega )^{\beta _i}u(x)=a(x)u(x),\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

has the only trivial solution provided that

$$\begin{aligned} \Vert a\Vert _{L^\infty }<\frac{\sqrt{M_{\beta _k}}\alpha _k}{2}. \end{aligned}$$
(26)

Here it is worth to note that replacing the functional J given by (25) (for \(p=q=2\)) with the following one

$$\begin{aligned} J_1(u)&=\int \limits _{\varOmega }\left( \frac{a(x)}{2}|u(x)|^2+\frac{1}{2a(x)} \Big |[w((-{\varDelta })_\omega )u(x)-f(x)]\Big |^2 \right. \\&\qquad \left. -u(x)\big (w((-{\varDelta })_\omega ) u(x)-f(x)\big ) \right) \hbox {d}x \end{aligned}$$

we can prove (similarly as in Sect. 3) counterparts of Theorems 3 and 4 for the functional \(J_1\) (in particular, under assumption (26) one can obtain the existence of a unique minimizer of \(J_1\) on \(D(w((-{\varDelta })_{\omega }))\)) and the following, more general linear problem

$$\begin{aligned} \sum \limits _{i=0}^k\alpha _i((-{\varDelta })_\omega )^{\beta _i}u(x)=a(x)u(x)+f(x),\quad x\in {\varOmega }\,\,\hbox {a.e.}, \end{aligned}$$
(27)

where \(f\in L^2({\varOmega },{\mathbb {R}})\). Additionally, if a is a constant function, then the problem (27) has a unique solution. Indeed, since \(J_1\) has a unique minimizer \(u_*\) on \(D(w((-{\varDelta })_{\omega }))\), therefore

$$\begin{aligned} \delta J_1(u_*,h)=0.\quad h\in D(w((-{\varDelta })_{\omega })), \end{aligned}$$
(28)

where \(\delta J_1(u_*,h)\) denotes the first variation of \(J_1\) at \(u_*\) in the direction h.

So,

$$\begin{aligned} \delta J_1(u_*,h)&= \int \limits _{\varOmega }\frac{1}{a} \Big (w((-{\varDelta })_\omega ) u_*(x)-au_*(x)-f(x)\Big ) \nonumber \\&\quad \Big (w((-{\varDelta })_\omega )h(x)-ah(x)\Big )\hbox {d}x=0 \end{aligned}$$
(29)

for any \(h\in D(w((-{\varDelta })_{\omega }))\).

Since \(a<\frac{\sqrt{M_{\beta _k}}\alpha _k}{2}\), therefore the constant a is not the eigenvalue of the operator \(w((-{\varDelta })_{\omega })\). Consequently, the kernel of the self-adjoint operator \(L{:}D(w((-{\varDelta })_{\omega }))\rightarrow L^2({\varOmega },{\mathbb {R}})\), defined as

$$\begin{aligned} Lh:=w((-{\varDelta })_\omega )h-ah,\quad a>0, \end{aligned}$$

is trivial, so the image of L satisfies the equality \(R(L)=L^2({\varOmega },{\mathbb {R}})\). Hence and from (29) we conclude

$$\begin{aligned} w((-{\varDelta })_\omega )u_*(x)-au_*(x)-f(x)=0,\quad x\in {\varOmega }\,\,\hbox {a.e.} \end{aligned}$$

This means that \(u_*\) is a unique solution of (27).