1 Introduction and preliminaries

In this article, we study the existence and uniqueness of mild and strong solutions of nonlocal nonlinear diffusion problems of p-Laplacian type with nonlinear boundary conditions. The problems are posed in a subset W of a metric random walk space [Xdm] with a reversible measure \(\nu \) for the random walk m (see Subsect. 1.1 for details). The nonlocal diffusion can hold either in W, in its nonlocal boundary \(\partial _mW\), or in both at the same time. We will assume that \(W\cup \partial _mW\) is m-connected and \(\nu \)-finite. The formulations of the diffusion problems that we study are the following:

$$\begin{aligned} \left\{ \begin{array}{ll} v_t(t,x) - \hbox {div}_m\textbf{a}_p u(t,x)=f(t,x), &{}x\in W,\ 0<t<T, \\ \\ \displaystyle v(t,x)\in \gamma \big (u(t,x)\big ), &{} x\in W,\ 0<t<T, \\ \\ -\mathcal {N}^{\textbf{a}_p}_\textbf{1} u(t,x) \in \beta \big (u(t,x)\big ), &{}x\in \partial _mW, \ 0<t<T, \\ \\ v(0,x) = v_0(x), &{}x\in W, \end{array} \right. \end{aligned}$$
(1.1)

and, for nonlinear dynamical boundary conditions,

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle v_t(t,x) - \hbox {div}_m\textbf{a}_p u(t,x)=f(t,x), &{}x\in W,\ 0<t<T, \\ \\ \displaystyle v(t,x)\in \gamma \big (u(t,x)\big ), &{} x\in W,\ 0<t<T, \\ \\ w_t(t,x)+\mathcal {N}^{\textbf{a}_p}_\textbf{1} u(t,x)=g(t,x), &{}x\in \partial _m W, \ 0<t<T, \\ \\ w(t,x) \in \beta \big (u(t,x)\big ), &{}x\in \partial _m W, \ 0<t<T, \\ \\ v(0,x) = v_0(x), &{}x\in W, \\ \\ w(0,x) = w_0(x), &{}x\in \partial _m W, \end{array} \right. \end{aligned}$$
(1.2)

where \(\gamma \) and \(\beta \) are maximal monotone (multivalued) graphs in \(\mathbb {R}\times \mathbb {R}\), \(\hbox {div}_m\textbf{a}_p\) is a nonlocal Leray–Lions-type operator whose model is the nonlocal p-Laplacian type diffusion operator, and \(\mathcal {N}^{\textbf{a}_p}_\textbf{1}\) is a nonlocal Neumann boundary operator (see Subsect. 2.1 for details). In fact, we solve these problems with greater generality, as we will not only consider them for a set W and its nonlocal boundary \(\partial _m W\), but rather for any two disjoint subsets \(\Omega _1\) and \(\Omega _2\) of X such that their union is m-connected.

These problems can be seen as the nonlocal counterpart of local diffusion problems governed by the p-Laplacian diffusion operator (or a Leray–Lions operator) where two further nonlinearities are induced by \(\gamma \) and \(\beta \) (see, for example, [4, 15] for local problems). In [8], and the references therein, one can find an interpretation of the nonlocal diffusion process involved in this kind of problems. On the nonlinearities (brought about by) \(\gamma \) and \(\beta \), we do not impose any further assumptions aside from the natural one (see Bénilan, Crandall and Sacks [15]):

$$\begin{aligned} 0\in \gamma (0)\cap \beta (0), \end{aligned}$$

and (in order for diffusion to take place)

$$\begin{aligned} \nu (W)\varGamma ^- + \nu (\partial _mW)\mathfrak {B}^-< \nu (W)\varGamma ^+ + \nu (\partial _mW)\mathfrak {B}^+, \end{aligned}$$

where

$$\begin{aligned} \varGamma ^-=\inf \text{ Ran }(\gamma ), \ \varGamma ^+=\sup \text{ Ran }(\gamma ), \ \mathfrak {B}^-=\inf \text{ Ran }(\beta ) \ \hbox { and } \ \mathfrak {B}^+=\sup \text{ Ran }(\beta ). \end{aligned}$$

Therefore, we work with a rather general class of nonlocal nonlinear diffusion problems with nonlinear boundary conditions. We are able to directly cover: obstacle problems, with unilateral or bilateral obstacles (either in W, in \(\partial _mW\), or in both at the same time); the nonlocal counterpart of Stefan-like problems that involve monotone graphs like the graph inverse of

$$\begin{aligned} \theta _S(r)=\left\{ \begin{array}{ll} r &{} \hbox {if } r<0,\\ \left[ 0,\lambda \right] &{}\hbox {if } r=0,\\ \lambda +r &{} \hbox {if } r>0, \end{array}\right. \end{aligned}$$

for \(\lambda >0\); diffusion problems in porous media, where monotone graphs like \(p_s(r)=|r|^{s-1}r\), \(s>0\), are involved; and Hele–Shaw type problems, which involve graphs like

$$\begin{aligned} H(r)=\left\{ \begin{array}{ll} 0 &{} \hbox {if } r<0,\\ \,[0,1]&{}\hbox {if } r=0,\\ 1 &{} \hbox {if } r>0. \end{array}\right. \end{aligned}$$

Moreover, if \(\gamma =0\) in problem (1.1), then the dynamics only appear in the nonlocal boundary and we obtain the evolution problem for a nonlocal Dirichlet-to-Neumann operator as a particular case. In addition, the homogeneous Dirichlet boundary condition (\(\beta =\{0\}\times \mathbb {R}\)) and the Neumann boundary condition (\(\beta =\mathbb {R}\times \{0\}\)) are also covered.

Nonlocal diffusion problems of p-Laplacian type involving nonlocal Neumann boundary operators have been recently studied in [43] inspired by the nonlocal Neumann boundary operators for the linear case studied in [29, 35]. Nevertheless, due to the generality of the hypotheses considered in this study, the results that we obtain lead to new existence and uniqueness results, which do not follow from previous works, for a great range of problems. This is true even when the problems are considered on weighted discrete graphs or \(\mathbb {R}^N\) with a random walk induced by a nonsingular kernel, spaces for which only some particular cases of these problems have been studied. Some references are given afterwards. For these ambient spaces and for the precise choice of the nonlocal p-Laplacian operator, Problem (1.1) has the following formulations (see Subsect. 1.1, in particular Examples 1.1 and 1.2, and Definition 1.4, for the necessary definitions and notations):

$$\begin{aligned} \left\{ \begin{array}{ll} v_t(t,x) = \displaystyle \frac{1}{d_x} \sum _{y \in V(G)} w_{x,y} \vert u(y)-u(x)\vert ^{p-2}(u(y) - u(x)), \quad &{}x\in W,\ 0<t<T, \\ \displaystyle v(t,x)\in \gamma \big (u(t,x)\big ), &{} x\in W,\ 0<t<T, \\ \displaystyle \frac{1}{d_x} \sum _{y \in W_{m^G}} w_{x,y} \vert u(y)-u(x)\vert ^{p-2}(u(y) - u(x)) \in \beta (u(t,x)), \quad &{}x\in \partial _{m^G}W, \ 0<t<T, \\ u(x,0) = u_0(x), &{}x\in W, \end{array} \right. \end{aligned}$$

for weighted discrete graphs, and

$$\begin{aligned} \left\{ \begin{array}{ll} v_t(t,x) = \displaystyle \int _{{\mathbb {R}}^N} J(y-x) \vert u(y)-u(x)\vert ^{p-2}(u(y) - u(x)) {\text {d}}y, \quad &{}x\in W,\ 0<t<T, \\ \displaystyle v(t,x)\in \gamma \big (u(t,x)\big ), &{} x\in W,\ 0<t<T, \\ \displaystyle \int _{W_{m^J}} J(y-x) \vert u(y) -u(x)\vert ^{p-2}(u(y) - u(x)) {\text {d}}y \in \beta (u(t,x)), \quad &{}x\in \partial _{m^J}W, \ 0<t<T, \\ v(x,0) = v_0(x), &{}x\in W. \end{array} \right. \end{aligned}$$

for the case of \(\mathbb {R}^N\) with the random walk induced by the nonsingular kernel J. We have detailed these problems with well-known formulations in order to show the extent to which Problems (1.1) and (1.2) cover specific nonlocal problems of great interest.

Nonlinear semigroup theory will be the basis for the study of the existence and uniqueness of solutions of the above problems. This study is developed in Sect. 3, where we prove, as a particular case of Theorem 3.4, the existence of mild solutions of Problem (1.2) for general data in \(L^1\), and of strong solutions assuming extra integrability conditions on the data. Moreover, a contraction and comparison principle is obtained. The same is done for Problem (1.1) in Theorem 3.10. See [9,10,11, 21, 30, 31] and [32], for details on such theory, which is completely covered in the well-known unpublished manuscript Evolution equations governed by accretive operators written by Ph. Bénilan, M. G. Crandall and A. Pazy. A summary of it can be found in [8, Appendix].

To apply the nonlinear semigroup theory, our first aim is to prove the existence and uniqueness of solutions of the problem

$$\begin{aligned} \left\{ \begin{array}{ll} \gamma \big (u(x))-\hbox {div}_m\textbf{a}_p u(x) \ni \varphi (x), \quad &{} x\in W, \\ \mathcal {N}^{\textbf{a}_p}_1 u(x)+\beta \big (u(x)\big )\ni \varphi (x), \quad &{} x\in \partial _m W, \end{array} \right. \end{aligned}$$
(1.3)

for general maximal monotone graphs \(\gamma \) and \(\beta \). This is the nonlocal counterpart of (local) quasilinear elliptic problems with nonlinear boundary conditions (see [5] and [15] for the general study of the local case) and is an interesting problem in itself due to the generality with which we address it. To this aim, we make use of a kind of nonlocal Poincaré-type inequalities (see Appendix A) which help us obtain boundedness arguments. These boundedness arguments together with some monotonicity arguments allow us to prove our results by adapting some of the ideas used in [5] and [15] (see also [7] for a very particular case). The same holds for the diffusion problems. The study of Problem (1.3) is developed in Sect. 2, where we prove, for a more general problem, the existence of solutions (Theorem 2.7) and a contraction and comparison principle (Theorem 2.6). At the end of that section, we deal with another nonlocal Neumann boundary operator.

For linear or quasilinear elliptic problems with boundary conditions, obstacles complicate the existence of solutions. The appearance of this difficulty is better understood when one takes into account the continuity of the solution between the inside of the domain and the boundary via the trace. In fact, for a bounded smooth domain \(\Omega \) in \(\mathbb {R}^N\), \(\gamma \) with bounded domain [0, 1] and \(\beta (r)=0\) for all r, it is not possible to find a weak solution of

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u+\gamma (u)\ni \varphi &{}\hbox {in }\Omega ,\\ \nabla u\cdot \eta ={\widetilde{\varphi }}&{}\hbox {in }\partial \Omega , \end{array} \right. \end{aligned}$$

for data satisfying \(\varphi \le 0\), \({\widetilde{\varphi }}\le 0\) and \( {\widetilde{\varphi }}\not \equiv 0\) (see [5]). However, in our nonlocal setting this sort of continuity is not present and the study of these nonlocal diffusion problems with obstacles hence differs from the study of the local ones (see [6] for a detailed study of these local problems). In particular, we do not need to impose any assumptions on the nonlinearities \(\gamma \) and \(\beta \) aside from the natural ones.

There is a very long list of references for the local elliptic and parabolic counterparts of the problems that we study; see, for example, [4, 5, 11,12,13, 15, 24, 44, 45], and the references therein. See also [38] for a Hele–Shaw problem with dynamical boundary conditions and the references therein. For some particular nonlocal problems we refer to [7, 8, 16, 19, 23, 36, 39, 43]. For fractional diffusion problems, we refer, for example, to [40], where Dirichlet and Neumann boundary conditions are considered; to [17, 18, 25, 26, 34], where fractional porous medium equations are studied, see also J. L. Vázquez’s survey [46] and the references therein; and to [27, 28] for fractional diffusion problems for the Stefan problem.

We now introduce the framework space considered and some other concepts that will be used later on.

1.1 Metric random walk spaces

Let (Xd) be a Polish metric space equipped with its Borel \(\sigma \)-algebra. In the following, whenever we consider a measure on X we assume that it is defined on this \(\sigma \)-algebra.

As introduced in [47], a random walk m on X is a family of Borel probability measures \(m_x\) on X, \(x \in X\), satisfying the two technical conditions: (1) the measures \(m_x\) depend measurably on the point \(x \in X\), i.e., for any Borel set A of X and any Borel set B of \({\mathbb {R}}\), the set \(\{ x \in X:m_x(A) \in B \}\) is Borel; (2) each measure \(m_x\) has finite first moment, i.e., for some (hence any) \(z \in X\), and for any \(x \in X\) one has \(\int _X d(z,y) dm_x(y) < +\infty \).

A metric random walk space [Xdm] is a Polish metric space (Xd) together with a random walk m.

A \(\sigma \)-finite measure \(\nu \) on X is invariant with respect to the random walk \(m=(m_x)\) if

$$\begin{aligned} \nu (A):=\int _X m_x(A){\text {d}}\nu (x)\ \ \hbox { for every Borel set } A. \end{aligned}$$

Moreover, the measure \(\nu \) is said to be reversible with respect to m if the following balance condition holds:

$$\begin{aligned} dm_x(y)d\nu (x)=dm_y(x){\text {d}}\nu (y), \end{aligned}$$

that is, for any Borel set \(C \subset X \times X\),

Under suitable assumptions on the metric random walk space [Xdm], such a reversible measure \(\nu \) exists and is unique. Note that the reversibility condition implies the invariance condition.

Assumption 1

From this point onwards, [Xdm] is a metric random walk space equipped with a \(\sigma \)-finite measure \(\nu \) which is reversible (thus invariant) with respect to m.

Let \(\mathcal {B}\) be the Borel \(\sigma \)-algebra of (Xd). Since \(\nu \) is a \(\sigma \)-finite measure on \((X,\mathcal {B})\) and m is a stochastic kernel on \((X,\mathcal {B})\), we may define the tensor product \(\nu \otimes m_x\) of \(\nu \) and m (see, for example, [33, Section 1.2.2], see also [1, Section 2.5]), which is a measure on \((X\times X, \mathcal {B}\otimes \mathcal {B})\), by

$$\begin{aligned} \nu \otimes m_x (A\times B):=\int _A m_x(B) {\text {d}}\nu (x) \quad \hbox {for every } A,\, B\in \mathcal {B}. \end{aligned}$$

Then, a \(\sigma \)-finite measure \(\nu \) invariant with respect to m is reversible if, and only if, the measure \(\nu \otimes m_x\) is symmetric. Note that, for every \(g\in L^1(X\times X,\nu \otimes m_x)\),

$$\begin{aligned} \int _{X \times X} g {\text {d}}(\nu \otimes m_x) = \int _X \int _X g(x,y) dm_x(y) {\text {d}}\nu (x). \end{aligned}$$

Example 1.1

An important class of examples of metric random walk spaces is composed by those which are obtained from weighted discrete graphs. Let \(G = (V(G), E(G),(w_{xy})_{x,y\in V(G)})\) be a weighted discrete graph, where V(G) is the set of vertices, E(G) is the set of and \(w_{xy} = w_{yx}\) is the nonnegative weight assigned to the edge \((x,y) \in E(G)\). We suppose that \(w_{xy} = 0\) if \((x,y) \not \in E(G)\) for \(x,y\in V(G)\). In this case, the following probability measures define a random walk on \((V(G),d_G)\) (here, \(d_G\) is the standard graph distance):

$$\begin{aligned} m_x^G:=\frac{1}{d_x}\sum _{y\in V(G)}w_{xy}, \end{aligned}$$

where \(d_x:= \sum _{y\sim x} w_{xy} = \sum _{y\in V(G)} w_{xy}\). Note that, if \(w_{x,y}=1\) for every \((x,y)\in E(G)\), then \(d_x\) coincides with the degree of the vertex x in the graph, that is, the number of edges containing the vertex x. Moreover, the measure \(\nu _G\) defined by

$$\begin{aligned} \nu _G(A):= \sum _{x \in A} d_x, \ \ \ A \subset V(G), \end{aligned}$$

is a reversible measure with respect to this random walk.

Example 1.2

Another important class of examples is given by those of the form \([{\mathbb {R}}^N, d,m^J]\) where d is the Euclidean distance and \(m^J\) is defined as follows: let \(J:{\mathbb {R}}^N\rightarrow [0,+\infty [\) be a measurable, nonnegative and radially symmetric function satisfying \(\int _{{\mathbb {R}}^N}J(z){\text {d}}\mathcal {L}^N(z)=1\) (\(\mathcal {L}^N\) is the Lebesgue measure) and set

$$\begin{aligned} m^J_x(A):= \int _A J(x - y) {\text {d}}\mathcal {L}^N(y) \quad \hbox { for every Borel set } A \subset {\mathbb {R}}^N \hbox { and }x\in {\mathbb {R}}^N. \end{aligned}$$

In this case, \(\mathcal {L}^N\) is a reversible measure with respect to this random walk.

See [41] (in particular [41, Example 1.2]) for a more detailed exposition of these and other examples.

Definition 1.3

Given two measurable subsets A, \(B \subset X\), we define the m-interaction between A and B as:

$$\begin{aligned} L_m(A,B):= \int _A \int _B dm_x(y) {\text {d}}\nu (x). \end{aligned}$$

Note that, whenever \(L_m(A,B) < +\infty \), if \(\nu \) is reversible with respect to m,

$$\begin{aligned} L_m(A,B)=L_m(B,A). \end{aligned}$$

Definition 1.4

Given a measurable set \(\Omega \subset X\), we define its m-boundary as:

$$\begin{aligned} \partial _m\Omega :=\{ x\in X\setminus \Omega : m_x(\Omega )>0 \} \end{aligned}$$

and its m-closure as:

$$\begin{aligned} \Omega _m:=\Omega \cup \partial _m\Omega . \end{aligned}$$

Moreover, we define the following ergodicity property.

Definition 1.5

Let [Xdm] be a metric random walk space with a reversible measure \(\nu \) with respect to m, and let \(\Omega \subset X\) be a measurable and non-\(\nu \)-null subset. We say that \(\Omega \) is m-connected if \(L_m(A,B)>0\) for every pair of measurable non-\(\nu \)-null sets A, \(B\subset \Omega \) such that \(A\cup B=\Omega \) (see [41]).

We recall the following nonlocal notions of gradient and divergence.

Definition 1.6

Given a function \(u: X \rightarrow {\mathbb {R}}\) we define its nonlocal gradient \(\nabla u: X \times X \rightarrow {\mathbb {R}}\) as:

$$\begin{aligned} \nabla u (x,y):= u(y) - u(x), \quad \, x,y \in X. \end{aligned}$$

For a function \(\textbf{z}: X \times X \rightarrow {\mathbb {R}}\), its m-divergence \(\textrm{div}_m \textbf{z}: X \rightarrow {\mathbb {R}}\) is defined as:

$$\begin{aligned} (\textrm{div}_m \textbf{z})(x):= \frac{1}{2} \int _{X} (\textbf{z}(x,y) - \textbf{z}(y,x)) dm_x(y), \quad x\in X. \end{aligned}$$

1.2 Yosida approximation and a Bénilan–Crandall relation

Given a maximal monotone graph \(\vartheta \) in \({\mathbb {R}}\times {\mathbb {R}}\) (see [21]) and \(\lambda >0\), let us denote by

$$\begin{aligned} \vartheta _\lambda :=\lambda \left( I-\left( I+\frac{1}{\lambda }\vartheta \right) ^{-1}\right) \end{aligned}$$

the Yosida approximation of \(\vartheta \) of parameter \(1/\lambda \).

The function \(\vartheta _\lambda \) is maximal monotone and Lipschitz continuous with Lipschitz constant \(\lambda \) (see [21, Proposition 2.6]. Moreover, \(\lim _{\lambda \rightarrow +\infty } \vartheta _\lambda (s) = \vartheta ^0 (s)\) where

$$\begin{aligned} \vartheta ^0(s):=\left\{ \begin{array}{ll} \hbox {the element of minimal absolute value of } \vartheta (s) &{} \hbox {if } s\in D(\vartheta ), \\ +\infty &{} \hbox {if } [s,+\infty )\cap D(\vartheta )=\emptyset , \\ -\infty &{} \hbox {if } (-\infty ,s]\cap D(\vartheta )=\emptyset , \end{array}\right. \end{aligned}$$

is an extension to \({\mathbb {R}}\) of the minimal section of \(\vartheta \). Furthermore, if \(s\in D(\vartheta )\), \(|\vartheta _\lambda (s)|\le |\vartheta ^0(s)|\) for every \(\lambda >0\), and \(|\vartheta _\lambda (s)|\) is nondecreasing in \(\lambda \).

Given a maximal monotone graph \(\vartheta \) in \({\mathbb {R}}\times {\mathbb {R}}\) with \(0\in \vartheta (0)\), we define, for \(s\in D(\vartheta )\),

$$\begin{aligned} \vartheta _+(s):=\left\{ \begin{array}{ll} \vartheta (s) &{} \hbox { if } s>0,\\ \vartheta (0)\cap [0,+\infty ) &{} \hbox { if } s=0,\\ \{0\} &{} \hbox { if } s<0, \end{array}\right. \end{aligned}$$

and

$$\begin{aligned} \vartheta _-(s):=\left\{ \begin{array}{ll} \{0\} &{} \hbox { if } s>0,\\ \vartheta (0)\cap (-\infty ,0] &{} \hbox { if } s=0,\\ \vartheta (s) &{} \hbox { if } s<0. \end{array}\right. \end{aligned}$$

Note that the Yosida approximation \((\vartheta _+)_\lambda \) of \(\vartheta _+\) is nondecreasing in \(\lambda >0\) and \((\vartheta _-)_\lambda \) is nonincreasing in \(\lambda >0\). Observe also that \((\vartheta _+)_\lambda (s)=0\) for \(s\le 0\) and \((\vartheta _-)_\lambda (s)=0\) for \(s\ge 0\), for every \(\lambda >0\), and \(\vartheta _++\vartheta _-=\vartheta \).

Given a maximal monotone graph \(\vartheta \) with \(0\in D(\vartheta )\), \(j_\vartheta (r):=\int _0^r\vartheta ^0(s){\text {d}}s\), \(r\in {\mathbb {R}}\), defines a convex and lower semicontinuous function such that \(\vartheta \) is equal to the subdifferential of \(j_\vartheta \):

$$\begin{aligned} \vartheta =\partial j_\vartheta . \end{aligned}$$

Moreover, if \(j_\vartheta {}^*\) is the Legendre transform of \(j_\vartheta \), then

$$\begin{aligned} \vartheta {}^{-1}=\partial j_\vartheta {}^*. \end{aligned}$$

We now recall a Bénilan–Crandall relation between functions \(u, v\in L^1(\Omega ,\nu )\). Denote by \(J_0\) and \(P_0\) the following sets of functions:

$$\begin{aligned} J_0:= & {} \{ j: {\mathbb {R}}\rightarrow [0, +\infty ] : \ j \text{ is } \text{ convex, } \text{ lower } \text{ semicontinuous } \text{ and } \ j(0) = 0 \}, \\ P_0:= & {} \left\{ \rho \in C^\infty ({\mathbb {R}}) : \ 0\le \rho '\le 1, \hbox { supp}(\rho ') \hbox { is compact and } 0\notin \hbox {supp}(\rho ) \right\} . \end{aligned}$$

Assume that \(\nu (\Omega ) < +\infty \) and let \(u,v\in L^1(\Omega ,\nu )\). The following relation between u and v is defined in [14]:

$$\begin{aligned} u\ll v \ \hbox { if} \ \int _{\Omega } j(u)\, {\text {d}}\nu \le \int _{\Omega } j(v) \, {\text {d}}\nu \ \ \hbox {for every} \ j \in J_0. \end{aligned}$$

Moreover, the following equivalences are proved in [14, Proposition 2.2] (we only give the particular cases that we use):

$$\begin{aligned}{} & {} \int _\Omega v\rho (u) {\text {d}}\nu \ge 0\quad \hbox {for every } \rho \in P_0\ \Longleftrightarrow \ u\ll u+\lambda v\quad \hbox {for every } \lambda>0,\\{} & {} \int _\Omega v\rho (u){\text {d}}\nu \ge 0\quad \hbox {for every } \rho \in P_0\ \Longleftrightarrow \ \int _{\{u<-h\}}v{\text {d}}\nu \le 0\le \int _{\{u>h\}}v{\text {d}}\nu \quad \hbox {for every } h>0. \end{aligned}$$

2 Nonlocal stationary problems

In this section, we give our main results concerning the existence and uniqueness of solutions of the nonlocal stationary Problem (1.3). We start by recalling the class of nonlocal Leray–Lions-type operators and the Neumann boundary operators that we will be working with, which were introduced in [43].

2.1 Nonlocal diffusion operators of Leray–Lions-type and nonlocal Neumann boundary operators

For \(1<p<+\infty \), let us consider a function \(\textbf{a}_p:X\times X\times \mathbb {R}\rightarrow \mathbb {R}\) such that

$$\begin{aligned}{} & {} (x,y)\mapsto \textbf{a}_p(x,y,r) \quad \hbox {is measurable for every } r\in {\mathbb {R}};\nonumber \\{} & {} \textbf{a}_p(x,y,.) \hbox { is continuous for } \nu \otimes m_x\hbox {-a.e } (x,y)\in X\times X; \end{aligned}$$
(2.1)
$$\begin{aligned}{} & {} \textbf{a}_p(x,y,r)=-\textbf{a}_p(y,x,-r) \quad \hbox {for } \nu \otimes m_x\hbox {-a.e } (x,y)\in X\times X \hbox { and for every } r\in {\mathbb {R}};\nonumber \\ \end{aligned}$$
(2.2)
$$\begin{aligned}{} & {} (\textbf{a}_p(x,y,r)-\textbf{a}_p(x,y,s))(r-s) > 0 \quad \hbox {for } \nu \otimes m_x\hbox {-a.e. } (x,y)\in X \times X \nonumber \\ {}{} & {} \hbox { and for every } r\ne s; \end{aligned}$$
(2.3)

there exist constants \(c_p,C_p>0\) such that

$$\begin{aligned} |\textbf{a}_p(x,y,r)|\le C_p\left( 1+|r|^{p-1}\right) \quad \hbox {for } \nu \otimes m_x\hbox {-a.e.} (x,y)\in X\times X\hbox { and for every } r\in {\mathbb {R}}, \nonumber \\ \end{aligned}$$
(2.4)

and

$$\begin{aligned} \textbf{a}_p(x,y,r)r\ge c_p\vert r \vert ^p \quad \hbox {for } \nu \otimes m_x\hbox {-a.e. }(x,y)\in X\times X\hbox { and for every } r\in {\mathbb {R}}. \nonumber \\ \end{aligned}$$
(2.5)

Condition (2.2) and the last condition imply that

$$\begin{aligned}{} & {} \textbf{a}_p(x,y,0)=0 \ \hbox { and } \ \hbox {sign}_0(\textbf{a}_p(x,y,r))=\hbox {sign}_0(r) \end{aligned}$$

for \(\nu \otimes m_x\)-a.e. \((x,y)\in X \times X\hbox { and for every } r\in {\mathbb {R}}\).

For \(u:X\rightarrow \mathbb {R}\), let us define \(\textbf{z}_{\textbf{a}_p,u}:X\times X\rightarrow {\mathbb {R}}\) by

$$\begin{aligned} \textbf{z}_{\textbf{a}_p,u}(x,y):=\textbf{a}_p(x,y, \nabla u(x,y)). \end{aligned}$$

Then, by Definition 1.6 and on account of (2.2),

$$\begin{aligned} \displaystyle \hbox {div}_m \textbf{z}_{\textbf{a}_p,u} (x)= & {} \frac{1}{2} \int _{X}\big (\textbf{a}_p(x,y,u(y)-u(x)) - \textbf{a}_p(y,x,u(x)-u(y))\big ) dm_x(y) \\= & {} \int _X \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y). \end{aligned}$$

For simplicity, we write

$$\begin{aligned} \hbox {div}_m \textbf{a}_p u(x)=\hbox {div}_m \textbf{z}_{\textbf{a}_p,u} (x). \end{aligned}$$

An example of a function \(\textbf{a}_p\) satisfying the above assumptions is

$$\begin{aligned} \textbf{a}_p(x,y,r):=\frac{\varphi (x)+\varphi (y)}{2}|r|^{p-2}r, \end{aligned}$$

where \(\varphi :X\rightarrow {\mathbb {R}}\) is a measurable function satisfying \(0<{c}\le \varphi \le {C}\), where c and C are constants. In particular, if \(\varphi (x)=2\) for every \( x\in X\),

$$\begin{aligned} \hbox {div}_m \textbf{a}_p u(x)= & {} \int _{X} |u(y)-u(x)|^{p-2}(u(y)-u(x)) dm_x(y) \nonumber \\= & {} \int _{X} |\nabla u(x,y)|^{p-2}\nabla u(x,y) dm_x(y) \end{aligned}$$

is the (nonlocal) p-Laplacian operator on the metric random walk space [Xdm].

Observe that \(\hbox {div}_m \textbf{a}_p u(x)\) defines a kind of Leray–Lions operator for the random walk m.

We now recall the nonlocal Neumann boundary operators introduced in [43]. Let us consider a measurable set \(W\subset X\) with \(\nu (W)>0\). The Gunzburger–Lehoucq-type Neumann boundary operator on \(\partial _mW\) is given by

$$\begin{aligned} \mathcal {N}^{\textbf{a}_p}_1 u(x):= -\int _{W_m} \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y), \quad x \in \partial _mW, \end{aligned}$$

where, taking into account the supports of the \(m_x\), we have that in fact, the integral is being calculated over the nonlocal tubular boundary \(\partial _mW\cup \partial _m(X\setminus W)\) of W. On the other hand, the Dipierro–Ros-Oton–Valdinoci-type Neumann boundary operator on \(\partial _mW\) is given by

$$\begin{aligned} \mathcal {N}^{\textbf{a}_p}_2 u(x):= -\int _{W} \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y) \quad x \in \partial _mW, \end{aligned}$$

for which, in this case, the integral is being calculated over the nonlocal boundary \(\partial _m(X\setminus W)\) of \(X\setminus W\).

For each of these Neumann boundary operators and for \(\varphi \) defined on \(W_m=W \cup \partial _m W\), we can look for solutions of the following problem:

$$\begin{aligned} \left\{ \begin{array}{ll} \gamma \big (u(x))-\hbox {div}_m\textbf{a}_p u(x) \ni \varphi (x), \quad &{} x\in W, \\ \mathcal {N}^{\textbf{a}_p}_\textbf{j} u(x)+\beta \big (u(x)\big )\ni \varphi (x), \quad &{} x\in \partial _m W, \end{array} \right. \end{aligned}$$

\(\textbf{j}\in \{1,2\}\). Observe that, by the reversibility of \(\nu \) with respect to m and recalling the definitions of \(\partial _m W\) and \(W_m\) (Definition 1.4), \(m_x(X{\setminus } W_m)=0\) for \(\nu \)-a.e. \(x\in W\). Indeed,

$$\begin{aligned} \displaystyle \int _{W}m_x(X\setminus W_m){\text {d}}\nu (x)=\int _{X\setminus W_m}m_x(W){\text {d}}\nu (x)=0. \end{aligned}$$

Consequently,

$$\begin{aligned} \hbox {div}_m\textbf{a}_p u (x) =\int _{W_m} \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y) \quad \hbox {for every}\quad x\in W. \end{aligned}$$
(2.6)

Lemma 2.1

Let \(\Omega \subset X\) be a \(\nu \)-finite set and let \(\{u_k\}_{k\in {\mathbb {N}}}\subset L^p(\Omega ,\nu )\) such that \(u_k{\mathop {\longrightarrow }\limits ^{k}} u\in L^p(\Omega ,\nu )\) in \(L^p(\Omega ,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega \). Suppose also that there exists \(h\in L^p(\Omega ,\nu )\) such that \(|u_k|\le h\) \(\nu \)-a.e. in \(\Omega \). Then,

$$\begin{aligned} \textbf{z}_{\textbf{a}_p,u_k}{\mathop {\longrightarrow }\limits ^{k}} \textbf{z}_{\textbf{a}_p,u} \ \hbox { in } L^{p'}(\Omega \times \Omega ,\nu \otimes m_x) \end{aligned}$$

and, in particular,

$$\begin{aligned} \int _\Omega \textbf{a}_p(\cdot ,y, \nabla u_{k}(\cdot ,y))\textrm{d}m_{(\cdot )}(y){\mathop {\longrightarrow }\limits ^{k}} \int _\Omega \textbf{a}_p(\cdot ,y, \nabla u(\cdot ,y))dm_{(\cdot )}(y) \ \hbox { in } L^{p'}(\Omega ,\nu ). \end{aligned}$$

Taking a subsequence if necessary, the \(\nu \)-a.e. pointwise convergence and the domination by the function h in the hypotheses are a consequence of the convergence in \(L^p(\Omega ,\nu )\).

Proof

Let \(A\subset \Omega \) be a \(\nu \)-null set such that \(|u_k(x)|\le h(x)<+\infty \) for every \(x\in \Omega {\setminus } A\) and every \(k\in {\mathbb {N}}\), and such that \(u_k(x){\mathop {\longrightarrow }\limits ^{k}} u(x)\) for every \(x\in \Omega \setminus A\). By (2.1), there exists a \(\nu \otimes m_x\)-null set \(N_1 \subset \Omega \times \Omega \) such that \(\textbf{a}_p(x,y,\cdot )\) is continuous for every \((x,y)\in (\Omega \times \Omega ){\setminus } N_1\). Therefore,

$$\begin{aligned} \textbf{a}_p(x,y, u_{k}(y)-u_{k}(x)){\mathop {\longrightarrow }\limits ^{k}} \textbf{a}_p(x,y, u(y)-u(x)) \end{aligned}$$

for every \((x,y)\in (\Omega \times \Omega ){\setminus } (N_1\cup (A\times \Omega )\cup (\Omega \times A))\), where, by the reversibility of \(\nu \) with respect to m, \(N_1\cup (A\times \Omega )\cup (\Omega \times A)\) is also \(\nu \otimes m_x\)-null. Moreover, by (2.4), there exists a \(\nu \otimes m_x\)-null set \(N_2 \subset \Omega \times \Omega \) such that

$$\begin{aligned} \begin{array}{rl} \displaystyle |\textbf{a}_p(x,y,u_{k}(x)-u_{k}(y))|&{} \displaystyle \le C_p\big (1+|u_{k}(x)-u_{k}(y)|^{p-1}\big )\\ &{} \displaystyle \le \widetilde{C}\big (1+|u_{k}(x)|^{p-1}+|u_{k}(y)|^{p-1}\big ) \\ &{} \displaystyle \le \widetilde{C} \big (1+|h(x)|^{p-1}+|h(y)|^{p-1}\big ) \end{array} \end{aligned}$$

for every \((x,y)\in (\Omega \times \Omega ){\setminus } (N_2\cup (A\times \Omega )\cup (\Omega \times A))\) and some constant \(\widetilde{C}\), where, again, \(N_2\cup (A\times \Omega )\cup (\Omega \times A)\) is \(\nu \otimes m_x\)-null. Then, taking \((x,y)\in (\Omega \times \Omega ){\setminus }(N_1\cup N_2\cup (A\times \Omega )\cup (\Omega \times A))\),

$$\begin{aligned} \textbf{a}_p(x,y, u_{k}(y)-u_{k}(x)){\mathop {\longrightarrow }\limits ^{k}} \textbf{a}_p(x,y, u(y)-u(x)) \end{aligned}$$

and

$$\begin{aligned} |\textbf{a}_p(x,y,u_{k}(x)-u_{k}(y))|\le \widetilde{C}\big (1+|h(x)|^{p-1}+|h(y)|^{p-1}\big ). \end{aligned}$$

Now, by the invariance of \(\nu \) with respect to m, since \(h\in L^{p}(\Omega ,m_x)\) and \(\nu (\Omega )<+\infty \), we have that for \({{\tilde{h}}}(x,y):= 1+|h(x)|^{p-1}+|h(y)|^{p-1}\), \({\tilde{h}}\in L^{p'}(\Omega \times \Omega ,\nu \otimes m_x)\), so we may apply the dominated convergence theorem to conclude. \(\square \)

2.2 Existence and uniqueness of solutions of doubly nonlinear stationary problems under nonlinear boundary conditions

As mentioned in the introduction, the aim here is to study the existence and uniqueness of solutions of the problem

$$\begin{aligned} \left\{ \begin{array}{ll} \gamma \big (u(x))-\hbox {div}_m\textbf{a}_p u(x) \ni \varphi (x), \quad &{} x\in W, \\ \mathcal {N}^{\textbf{a}_p}_1 u(x)+\beta \big (u(x)\big )\ni \varphi (x), \quad &{} x\in \partial _m W, \end{array} \right. \end{aligned}$$
(2.7)

where \(W\subset X\) is m-connected and \(\nu (W_m)<+\infty \). See [5, 15] for the reference local models. In Subsect. 2.3, we address this problem but with the nonlocal Neumann boundary operator \(\mathcal {N}^{\textbf{a}_p}_2\) instead.

Problem (2.7) is a particular case (recall (2.6)) of the following general, and interesting by itself, problem. Let \(\Omega _1,\Omega _2\subset X\) be disjoint measurable non-\(\nu \)-null sets and let

$$\begin{aligned} \Omega :=\Omega _1\cup \Omega _2. \end{aligned}$$

Given \(\varphi \in L^{1}(\Omega ,\nu )\), we consider the problem

$$\begin{aligned} (GP_\varphi ^{ \textbf{a}_p,\gamma ,\beta })\quad \left\{ \begin{array}{ll}\displaystyle \gamma \big (u(x))- \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y) \ni \varphi (x), \quad &{} x\in \Omega _1, \\ \displaystyle \beta \big (u(x)\big )- \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y)\ni \varphi (x), \quad &{} x\in \Omega _2. \end{array} \right. \nonumber \\ \end{aligned}$$
(2.8)

For simplicity, we generally use the notation \((GP_\varphi )\) in place of \((GP_\varphi ^{ \textbf{a}_p,\gamma ,\beta })\). However, we use the more detailed notation further on. Moreover, we make the following assumptions.

Assumption 2

We assume that \(\Omega =\Omega _1\cup \Omega _2\) is m-connected and \(\nu (\Omega )<+\infty \).

Remark 2.2

Observe that, given an m-connected set \(\Omega \subset X\) (recall Definition 1.5), \(m_x(\Omega )>0\) for \(\nu \)-a.e. \(x\in \Omega \). Indeed, if

$$\begin{aligned} N:=\{x\in \Omega \,: \, m_x(\Omega )=0\}, \end{aligned}$$

then

$$\begin{aligned} L_m(N,\Omega )=0, \end{aligned}$$

thus \(\nu (N)=0\).

Assumption 3

Let

where the notation means that and are mutually singular. We assume that

$$\begin{aligned} \nu \left( \mathcal {N}_\perp ^\Omega \right) =0. \end{aligned}$$

Remark 2.3

Note that, for \(x\in \Omega \) such that \(m_x(\Omega )>0\), if \(m_x\ll \nu \) (i.e., \(m_x\) is absolutely continuous with respect to \(\nu \), do not confuse the use of \(\ll \) in this context with its use in the notation in Subsect. 1.2) then . Therefore, by Remark 2.2, if \(m_x\ll \nu \) for \(\nu \)-a.e. \(x\in \Omega \) then \(\nu \left( \mathcal {N}_\perp ^\Omega \right) =0\). Hence, the above condition is weaker than assuming that \(m_x\ll \nu \) for \(\nu \)-a.e. \(x\in \Omega \).

Assumption 4

We assume, together with \(0\in \gamma (0)\cap \beta (0)\), that

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-< \mathcal {R}_{\gamma ,\beta }^+, \end{aligned}$$

where

$$\begin{aligned} \begin{array}{c} \mathcal {R}_{\gamma ,\beta }^-:=\nu (\Omega _1)\inf \text{ Ran }(\gamma ) + \nu (\Omega _2)\inf \text{ Ran }(\beta ), \\ \mathcal {R}_{\gamma ,\beta }^+:=\nu (\Omega _1)\sup \text{ Ran }(\gamma ) + \nu (\Omega _2)\sup \text{ Ran }(\beta ). \end{array} \end{aligned}$$

Assumption 5

We assume that the following generalised Poincaré type inequality holds: For every \(0<l\le \nu (\Omega )\), there exists a constant \(\Lambda >0\) such that, for every \(u \in L^p(\Omega ,\nu )\) and any measurable set \(Z\subset \Omega \) with \(\nu (Z)\ge l\),

$$\begin{aligned} \left\| u \right\| _{L^p(\Omega ,\nu )} \le \Lambda \left( \left( \int _{ \Omega \times \Omega } |u(y)-u(x)|^p dm_x(y) {\text {d}}\nu (x) \right) ^{\frac{1}{p}}+\left| \int _Z u\,{\text {d}}\nu \right| \right) . \end{aligned}$$

This assumption holds true in many important examples (see Appendix A).

From now on in this subsection, we work under Assumptions 1 to 5.

Definition 2.4

A solution of \((GP_\varphi )\) is a pair [uv] with \(u\in L^p(\Omega ,\nu )\) and \(v\in L^{p'}(\Omega ,\nu )\) such that

  1. 1.

    \(v(x)\in \gamma (u(x))\ \hbox { for}~\nu \hbox {-a.e. } x\in \Omega _1,\)

  2. 2.

    \(v(x)\in \beta (u(x))\ \hbox { for}~\nu \hbox {-a.e. } x\in \Omega _2,\)

  3. 3.

    \([(x,y)\mapsto a_p(x,y,u(y)-u(x))]\in L^{p'}(\Omega \times \Omega ,\nu \otimes m_x)\),

  4. 4.

    and

    $$\begin{aligned} v(x) - \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y)=\varphi (x), \quad x \in \Omega . \end{aligned}$$

A subsolution (supersolution) of \((GP_\varphi )\) is a pair [uv] with \(u\in L^p(\Omega ,\nu )\) and \(v\in L^1(\Omega ,\nu )\) satisfying 1., 2., 3. and

$$\begin{aligned} v(x) - \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y)\le \varphi (x), \quad x \in \Omega ,\\ \left( v(x) - \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y)\ge \varphi (x), \quad x \in \Omega \right) . \end{aligned}$$

Remark 2.5

(Integration by parts formula) The following integration by parts formula which results from the reversibility of \(\nu \) with respect to m, can be easily proved. Let u be a measurable function such that

$$\begin{aligned} {[}(x,y)\mapsto \textbf{a}_p(x,y,u(y)-u(x))]\in L^{q}( \Omega \times \Omega ,\nu \otimes m_x) \end{aligned}$$

and let \(w \in L^{q'}(\Omega ,\nu )\). Then,

$$\begin{aligned} \begin{array}{l} \displaystyle -\int _{\Omega }\int _{\Omega } \textbf{a}_p(x,y,u(y)- u (x))dm_x(y)w(x){\text {d}}\nu (x) \\ = \displaystyle \frac{1}{2} \int _{\Omega \times \Omega } \textbf{a}_p(x,y,u(y)-u(x)) (w(y) - w(x)) {\text {d}}(\nu \otimes m_x)(x,y). \end{array} \end{aligned}$$

Let us see, formally, the way in which we use the above integration by parts formula in what follows. Suppose that we are in the following situation:

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle -\int _{\Omega } \textbf{a}_p(x,y,u(y)- u (x))dm_x(y) = f(x), \quad &{} x\in \Omega _1, \\ \displaystyle -\int _{\Omega } \textbf{a}_p(x,y,u(y)- u (x))dm_x(y) = g(x), \quad &{} x \in \Omega _2. \end{array} \right. \end{aligned}$$

Then, multiplying both equations by a test function w, integrating them with respect to \(\nu \) over \(\Omega _1\) and \(\Omega _2\), respectively, adding them and using the integration by parts formula we get

$$\begin{aligned} \begin{array}{l} \displaystyle \frac{1}{2} \int _{\Omega \times \Omega } \textbf{a}_p(x,y,u(y)-u(x)) (w(y) - w(x)) {\text {d}}(\nu \otimes m_x)(x,y) \\ \displaystyle = \int _{\Omega _1} f(x)w(x){\text {d}}\nu (x) +\int _{\Omega _2} g(x)w(x) {\text {d}}\nu (x). \end{array} \end{aligned}$$

Moreover, as a consequence of (2.3), taking \(u=u_i\), \(f=f_i\) and \(g=g_i\), \(i=1,2\), in the above system and for every nondecreasing function \(T:\mathbb {R}\rightarrow \mathbb {R}\), we obtain

$$\begin{aligned} \displaystyle{} & {} \int _{\Omega _1} (f_1(x)-f_2(x))T(u_1(x)-u_2(x)){\text {d}}\nu (x) +\int _{\Omega _2} (g_1(x)-g_2(x))T(u_1(x)-u_2(x)) {\text {d}}\nu (x) \\{} & {} \qquad \displaystyle =\frac{1}{2} \int _{\Omega \times \Omega } \big (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x))\big ) \\{} & {} \displaystyle \qquad \qquad \qquad \qquad \qquad \times \big (T(u_1(y) - u_2(y)) -T(u_1(x)-u_2(x))\big ){\text {d}}(\nu \otimes m_x)(x,y)\,\ge 0. \end{aligned}$$

The next result gives a maximum principle for solutions of Problem \((GP_\varphi )\) given in (2.8) and, consequently, also for solutions of Problem (2.7).

Theorem 2.6

(Contraction and comparison principle) Let \(\varphi _1\), \(\varphi _2\in L^{1}(\Omega ,\nu )\). Let \([u_{1},v_1]\) be a subsolution of \((GP_{\varphi _1})\) and \([u_{2},v_2]\) be a supersolution of \((GP_{\varphi _2})\). Then,

$$\begin{aligned} \int _\Omega (v_1-v_2)^+{\text {d}}\nu \le \int _\Omega (\varphi _1-\varphi _2)^+{\text {d}}\nu . \end{aligned}$$
(2.9)

Moreover, if \(\varphi _1\le \varphi _2\) with \(\varphi _1\ne \varphi _2\), then \(v_1\le v_2\), \(v_1\ne v_2\), and \(u_1\le u_2\) \(\nu \)-a.e. in \(\Omega \).

Furthermore, if \(\varphi _1= \varphi _2\) and \([u_i,v_i]\) is a solution of \((GP_{\varphi _i})\), \(i=1,2\), then \(v_1=v_2\) \(\nu \)-a.e. in \(\Omega \) and \(u_1-u_2\) is \(\nu \)-a.e. equal to a constant.

Proof

By hypothesis,

$$\begin{aligned}{} & {} v_1(x)-v_2(x) - \int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x))) dm_x(y)\\{} & {} \le \varphi _1(x)-\varphi _2(x) \end{aligned}$$

for \(x \in \Omega \). Let \(k>0\) and \(T_k:{\mathbb {R}}\rightarrow [-k,k]\) be the truncation operator defined as:

$$\begin{aligned} T_k(r):= \left\{ \begin{array}{lll} -k &{}\quad \hbox {if} \ r< -k,\\ r &{}\quad \hbox {if} \ \vert r \vert \le k, \\ k &{}\quad \hbox {if} \ r >k, \end{array}\right. \end{aligned}$$
(2.10)

and denote \(T_k^+(s):=(T_k(s))^+\). Multiplying the above inequality by \(\frac{1}{k} T_k^+(u_{1}-u_{2}+ k \, \hbox {sign}_0^+(v_1-v_2))\) and integrating over \(\Omega \), we get

$$\begin{aligned}{} & {} \int _{\Omega }\left( v_1(x)-v_2(x)\right) \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)+ k \, \hbox {sign}_0^+(v_1(x)-v_2(x))) {\text {d}}\nu (x) \nonumber \\{} & {} \qquad - \int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x))) dm_x(y)\nonumber \\{} & {} \qquad \qquad \qquad \qquad \qquad \qquad \times \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)+ k \, \hbox {sign}_0^+(v_1(x)-v_2(x))){\text {d}}\nu (x) \nonumber \\{} & {} \quad \le \int _{\Omega }(\varphi _1(x)-\varphi _2(x)) \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)+ k \, \hbox {sign}_0^+(v_1(x)-v_2(x))){\text {d}}\nu (x)\nonumber \\{} & {} \quad \le \int _{\Omega }(\varphi _1(x)-\varphi _2(x))^+ {\text {d}}\nu (x). \end{aligned}$$
(2.11)

Moreover, by the integration by parts formula (Remark 2.5),

$$\begin{aligned}{} & {} \displaystyle - \int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x)))dm_x(y) \\{} & {} \displaystyle \quad \times \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)+ k \, \hbox {sign}_0^+(v_1(x)-v_2(x))) {\text {d}}\nu (x)\\ \\{} & {} \displaystyle =\frac{1}{2} \int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x)))\\{} & {} \displaystyle \quad \times \Big (\frac{1}{k} T_k^+(u_{1}(y)-u_{2}(y)+ k \, \hbox {sign}_0^+(v_1(y)-v_2(y)))\\{} & {} \displaystyle \quad -\frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)+ k \, \hbox {sign}_0^+(v_1(x)-v_2(x)))\Big ) dm_x(y){\text {d}}\nu (x). \end{aligned}$$

Now, since the integrand on the right-hand side is bounded from below by an integrable function, we can apply Fatou’s lemma to get (recall the last observation in Remark 2.5)

$$\begin{aligned} \displaystyle \liminf _{k\rightarrow 0^+}-&\int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x)))dm_x(y) \\&\qquad \quad \displaystyle \times \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)+ k \, \hbox {sign}_0^+(v_1(x)-v_2(x))) {\text {d}}\nu (x)\ge 0. \end{aligned}$$

Hence, taking limits in (2.11), we get

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega }\left( v_1(x)-v_2(x)\right) ^+ {\text {d}}\nu (x) \\{} & {} \displaystyle =\lim _{k\rightarrow 0^+}\int _{\Omega }\left( v_1(x)-v_2(x)\right) \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)+ k \, \hbox {sign}_0^+(v_1(x)-v_2(x))) {\text {d}}\nu (x) \\{} & {} \displaystyle \le \int _{\Omega }(\varphi _1(x)-\varphi _2(x))^+ {\text {d}}\nu (x), \end{aligned}$$

and (2.9) is proved.

Take now \(\varphi _1\le \varphi _2\) with \(\varphi _1\ne \varphi _2\), then, by (2.9), \(v_1\le v_2\) \(\nu \)-a.e. in \(\Omega \). Now, since \([u_{1},v_1]\) is a subsolution of \((GP_{\varphi _1})\)

$$\begin{aligned} v_1(x) - \int _{\Omega } \textbf{a}_p(x,y,u_1(y)-u_1(x)) dm_x(y)\le \varphi _1(x) \end{aligned}$$

thus

$$\begin{aligned} \int _\Omega v_1(x){\text {d}}\nu (x) - \underbrace{\int _\Omega \int _{\Omega } \textbf{a}_p(x,y,u_1(y)-u_1(x)) dm_x(y){\text {d}}\nu (x)}_{=0} \le \int _\Omega \varphi _1(x){\text {d}}\nu (x). \end{aligned}$$

Therefore, with the same calculation for \([u_{2},v_2]\),

$$\begin{aligned} \int _\Omega v_1(x){\text {d}}\nu (x) \le \int _\Omega \varphi _1(x){\text {d}}\nu (x)<\int _\Omega \varphi _2(x){\text {d}}\nu (x)\le \int _\Omega v_2(x){\text {d}}\nu (x) \end{aligned}$$

thus \(v_1\ne v_2\). Now, since \((\varphi _1-\varphi _2)^+=0\) and \((v_1-v_2)^+=0\), from (2.11) we get that

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega }\left( v_1(x)-v_2(x)\right) \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x))) {\text {d}}\nu (x) \\{} & {} \displaystyle - \int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x)))\\{} & {} \qquad \qquad \qquad \times \frac{1}{k} T_k^+(u_{1}(x)\displaystyle -u_{2}(x)) dm_x(y){\text {d}}\nu (x)\le 0. \end{aligned}$$

However, since \(v_i(x)\in \gamma (u_i(x))\) for \(\nu \)-a.e. \(x\in \Omega _1\) and \(v_i(x)\in \beta (u_i(x))\) for \(\nu \)-a.e. \(x\in \Omega _2\), \(i=1,2\), we get, by the monotonicity of the graphs, \(u_1(x)\le u_2(x)\) for \(\nu \)-a.e. \(x\in \Omega \) such that \(v_1(x)<v_2(x)\). Therefore, \(\left( v_1(x)-v_2(x)\right) \frac{1}{k} T_k^+(u_{1}(x)-u_{2}(x)))=0\) for \(\nu \)-a.e. \(x\in \Omega \) and thus

$$\begin{aligned} - \int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x)){} & {} -\textbf{a}_p(x,y,u_2(y)-u_2(x)))\\ \times \frac{1}{k} T_k^+(u_{1}(x){} & {} -u_{2}(x)) dm_x(y){\text {d}}\nu (x) \le 0. \end{aligned}$$

Now, recalling Remark 2.5 (that is, integration by parts), we obtain

$$\begin{aligned} \begin{array}{l} \displaystyle \int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x)))\\ \displaystyle \qquad \quad \times ((u_{1}(y)-u_{2}(y))^+-(u_{1}(x)-u_{2}(x))^+) dm_x(y){\text {d}}\nu (x) = 0, \end{array} \end{aligned}$$

and thus

$$\begin{aligned} (\textbf{a}_p(x,y,u_1(y)-u_1(x)){} & {} -\textbf{a}_p(x,y,u_2(y)-u_2(x)))\nonumber \\ {}{} & {} \times (u_{1}(y)-u_{1}(x) -(u_{2}(y)-u_{2}(x)))=0 \end{aligned}$$
(2.12)

for \((x,y)\in (\Omega \times \Omega )\setminus N\) where \(N\subset \Omega \times \Omega \) is a \(\nu \otimes m_x\)-null set. Let \(C\subset \Omega \) be a \(\nu \)-null set such that the section \(N_x:=\{y\in \Omega \,:\, (x,y)\in N\}\) of N is \(m_x\)-null for every \(x\in \Omega {\setminus } C\) and let us see that \(u_1(x)\le u_2(x)\) for every \(x\in \Omega \setminus (C\cup \mathcal {N}_\perp ^\Omega )\) (recall Assumption 3 for the definition of the \(\nu \)-null set \(\mathcal {N}_\perp ^\Omega \)). Suppose that there exists \(x_0\in \Omega \setminus (C\cup \mathcal {N}_\perp ^\Omega )\) such that \(u_1(x_0)-u_2(x_0)>0\). Then, from (2.12) (and (2.3)), we get that \(u_1(y)-u_2(y)=u_1(x_0)-u_2(x_0)>0\) for every \(y\in \Omega \setminus N_{x_0}\). Let

$$\begin{aligned} S:=\{y\in \Omega \,:\, u_1(y)-u_2(y)=u_1(x_0)-u_2(x_0)\}\supset \Omega \setminus N_{x_0}. \end{aligned}$$

Since \(x_0\not \in \mathcal {N}_\perp ^\Omega \) and \(m_{x_0}(N_{x_0})=0\), we must have \(\nu (S)\ge \nu (\Omega {\setminus } N_{x_0})>0\). Now, following the same argument as before, if \(x\in S\), then \(\Omega \setminus N_x\subset S\) thus \(m_x(\Omega {\setminus } S)\le m_x(N_x)=0\) and, therefore,

$$\begin{aligned} L_m(S,\Omega \setminus S)=0. \end{aligned}$$

However, since \(\Omega \) is m-connected and \(\nu (S)>0\), we must have \(\nu (\Omega \setminus S)=0\); thus, \(u_1(y)-u_2(y)=u_1(x_0)-u_2(x_0)>0\) for \(\nu \)-a.e. \(y\in \Omega \). This contradicts that \(v_1\le v_2\), \(v_1\ne v_2\), \(\nu \)-a.e. in \(\Omega \).

Finally, suppose that \([u_{1},v_1]\) and \([u_{2},v_2]\) are solutions of \((GP_{\varphi })\) for some \(\varphi \in L^1(\Omega ,\nu )\). Then,

$$\begin{aligned} v_1(x)-v_2(x) - \int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x))) dm_x(y)=0 \end{aligned}$$

thus, since \(v_1=v_2\) \(\nu \)-a.e. in \(\Omega \),

$$\begin{aligned} -\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x))-\textbf{a}_p(x,y,u_2(y)-u_2(x))) dm_x(y)=0. \end{aligned}$$

Multiplying this equation by \(u_1-u_2\), integrating over \(\Omega \) and using the integration by parts formula as in Remark 2.5 we get

$$\begin{aligned} \int _{\Omega }\int _{\Omega } (\textbf{a}_p(x,y,u_1(y)-u_1(x)){} & {} -\textbf{a}_p(x,y,u_2(y)-u_2(x)))\\{} & {} \times (u_1(y) -u_1(x)-(u_2(y)-u_2(x)))=0 \end{aligned}$$

thus, by (2.3) and positivity,

$$\begin{aligned} (\textbf{a}_p(x,y,u_1(y)-u_1(x)){} & {} -\textbf{a}_p(x,y,u_2(y)-u_2(x)))\nonumber \\{} & {} \times (u_1(y)-u_1(x)-(u_2(y)-u_2(x)))=0 \end{aligned}$$
(2.13)

for \((x,y)\in (\Omega \times \Omega )\setminus N'\) where \(N'\subset \Omega \times \Omega \) is a \(\nu \otimes m_x\)-null set. Let \(C'\subset \Omega \) be a \(\nu \)-null set such that the section \(N'_x:=\{y\in \Omega \,:\, (x,y)\in N'\}\) of \(N'\) is \(\nu \)-null for every \(x\in \Omega \setminus C'\), and let us see that there exists \(L\in {\mathbb {R}}\) such that \(u_1(x)- u_2(x)=L\) for \(\nu \)-a.e. \(x\in \Omega \). Let \(x_0\in \Omega \setminus C'\), \(L:=u_1(x_0)- u_2(x_0)\) and

$$\begin{aligned} S':=\{y\in \Omega \,:\, u_1(y)-u_2(y)=L \}\supset \Omega \setminus N'_{x_0}. \end{aligned}$$

By (2.13), \(\Omega \setminus C'_{x_0}\subset S'\). Proceeding as we did before to prove that \(\nu (\Omega \setminus S)=0\), we obtain that \(\nu (\Omega \setminus S')=0\). \(\square \)

In order to prove the existence of solutions of Problem (2.8) (Theorem 2.7), we first prove the existence of solutions of an approximate problem. Then, we obtain some monotonicity and boundedness properties of the solutions of these approximate problems that allow us to pass to the limit. This method lets us get around the loss of compactness results in our setting with respect to the local setting. Indeed, we follow ideas used in [5], but, as we have said, making the most of the monotonicity arguments since the Poincaré-type inequalities here only produce boundedness in \(L^{p}\) spaces (versus the boundedness in \(W^{1,p}\) spaces obtained in their local setting). This will be done in the following subsections.

2.2.1 Existence of solutions of an approximate problem

Take \(\varphi \in L^{\infty }(\Omega ,\nu )\). Let \(n, k\in \mathbb {N}\), \(K>0\) and

$$\begin{aligned} A:=A_{n,k}:L^p(\Omega ,\nu ) \rightarrow L^{p'}(\Omega ,\nu )\equiv L^{p'}(\Omega _1,\nu )\times L^{p'}(\Omega _2,\nu ) \end{aligned}$$

be defined by

$$\begin{aligned} A(u)= \big (A_1(u),A_2(u)\big ), \end{aligned}$$

where

$$\begin{aligned} \displaystyle A_1(u)(x):= & {} \displaystyle T_K((\gamma _+)_k(u(x)))+T_K((\gamma _-)_n(u(x)))-\int _{\Omega }\textbf{a}_p(x,y,u(y)-u(x))dm_x(y)\\{} & {} \displaystyle +\frac{1}{n} |u(x)|^{p-2}u^+(x) -\frac{1}{k} |u(x)|^{p-2}u^-(x), \end{aligned}$$

for \(x\in \Omega _1\), and

$$\begin{aligned} \displaystyle A_2(u)(x):= & {} \displaystyle T_K((\beta _+)_k(u(x)))+T_K((\beta _-)_n(u(x)))-\int _{\Omega }\textbf{a}_p(x,y,u(y)-u(x))dm_x(y)\\{} & {} \displaystyle +\frac{1}{n} |u(x)|^{p-2}u^+(x)-\frac{1}{k} |u(x)|^{p-2}u^-(x), \end{aligned}$$

for \(x\in \Omega _2\). Here, \(T_K\) is the truncation operator defined in (2.10) and \((\gamma _+)_k\), \((\gamma _-)_n\), \((\beta _+)_k\) and \((\beta _-)_n\) are Yosida approximations as defined in Subsect. 1.2.

It is easy to see that A is continuous and, moreover, it is monotone and coercive in \(L^p(\Omega , \nu )\). Indeed, the monotonicity results from the integration by parts formula (Remark 2.5) and the coercivity results from the following computation (where the term involving \(\textbf{a}_p\) has been neglected because it is nonnegative, as shown in Remark 2.5):

$$\begin{aligned} \int _{\Omega } A(u)u {\text {d}}\nu \ge \frac{1}{n} || u^+||_{L^p(\Omega ,\nu )}+\frac{1}{k}|| u^-||_{L^p(\Omega ,\nu )}. \end{aligned}$$

Therefore, since \(\varphi \in L^{\infty }(\Omega ,\nu )\subset L^{p'}(\Omega ,\nu )\), by [20, Corollary 30], there exist \(u_{n,k}\in L^{p}(\Omega , \nu )\), n, \(k\in {\mathbb {N}}\), such that

$$\begin{aligned} \big (A_1(u_{n,k}),A_2(u_{n,k})\big )=\varphi . \end{aligned}$$

That is,

$$\begin{aligned}{} & {} T_K((\gamma _+)_k(u_{n,k}(x)))+T_K((\gamma _-)_n(u_{n,k}(x))) -\displaystyle \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y) \nonumber \\{} & {} \quad \displaystyle +\frac{1}{n} |u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k} |u_{n,k}(x)|^{p-2}u_{n,k}^-(x) =\varphi (x) \ \hbox { for}~x\in \Omega _1, \end{aligned}$$
(2.14)

and

$$\begin{aligned}{} & {} T_K((\beta _+)_k(u_{n,k}(x))) +T_K((\beta _-)_n(u_{n,k}(x)))-\displaystyle \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y) \nonumber \\{} & {} \quad \displaystyle +\frac{1}{n} |u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k} |u_{n,k}(x)|^{p-2}u_{n,k}^-(x) =\varphi (x) \ \hbox { for } x\in \Omega _2. \end{aligned}$$
(2.15)

Let n, \(k\in {\mathbb {N}}\). We start by proving that \(u_{n,k}\in L^\infty (\Omega ,\nu )\). Set

$$\begin{aligned} M:=\left( (k+n)\Vert \varphi \Vert _{L^\infty (\Omega ,\nu )}\right) ^{\frac{1}{p-1}}. \end{aligned}$$

Then, multiplying (2.14) and (2.15) by \((u_{n,k}-M)^+\), integrating over \(\Omega _1\) and \(\Omega _2\), respectively, adding both equations and neglecting the terms which are zero, we get

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1} T_K((\gamma _+)_k(u_{n,k}(x))) (u_{n,k}(x)-M)^+ {\text {d}}\nu (x) \nonumber \\{} & {} \quad + \int _{\Omega _2} T_K((\beta _+)_k(u_{n,k}(x))) (u_{n,k}(x)-M)^+ {\text {d}}\nu (x) \nonumber \\{} & {} \quad - \displaystyle \int _{\Omega }\int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))(u_{n,k}(x)-M)^+dm_x(y) {\text {d}}\nu (x) \nonumber \\ {}{} & {} \displaystyle \quad + \frac{1}{n}\int _\Omega |u_{n,k}(x)|^{p-2}u_{n,k}^+(x)(u_{n,k}(x)-M)^+{\text {d}}\nu (x) \nonumber \\ {}{} & {} = \displaystyle \int _{\Omega } \varphi (x)(u_{n,k}(x)-M)^+ {\text {d}} \nu (x). \end{aligned}$$
(2.16)

Now, by the integration by parts formula (recall Remark 2.5),

$$\begin{aligned}{} & {} \displaystyle -\int _{\Omega }\int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))(u_{n,k}(x)-M)^+ dm_x(y){\text {d}}\nu (x) \\{} & {} \quad \displaystyle =\frac{1}{2}\int _{\Omega }\int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))\left( (u_{n,k}(y)-M)^+ \right. \\{} & {} \qquad \displaystyle \left. -(u_{n,k}(x)-M)^+\right) dm_x(y){\text {d}}\nu (x)\ge 0. \end{aligned}$$

Hence, neglecting nonnegative terms in (2.16), we get

$$\begin{aligned} \int _\Omega |u_{n,k}(x)|^{p-2}u_{n,k}^+(x)(u_{n,k}(x)-M)^+{\text {d}}\nu (x) \le n\displaystyle \int _{\Omega } \varphi (x)(u_{n,k}(x)-M)^+ {\text {d}} \nu (x), \end{aligned}$$

thus

$$\begin{aligned} \int _\Omega T_K(|u_{n,k}(x)|^{p-2}u_{n,k}^+(x))(u_{n,k}(x)-M)^+{\text {d}}\nu (x) \le n\displaystyle \int _{\Omega } \varphi (x)(u_{n,k}(x)-M)^+ {\text {d}} \nu (x). \end{aligned}$$

Now, subtracting \(\displaystyle \int _\Omega M^{p-1}(u_{n,k}(x)-M)^+{\text {d}}\nu (x)\) from both sides of the above inequality yields

$$\begin{aligned} \begin{array}{l} \displaystyle \int _\Omega \left( T_K(|u_{n,k}(x)|^{p-2}u_{n,k}^+(x))-M^{p-1}\right) (u_{n,k}(x)-M)^+{\text {d}}\nu (x) \\ \le \displaystyle n\int _{\Omega }\left( \varphi (x)-\frac{1}{n} M^{p-1}\right) (u_{n,k}(x)-M)^+ {\text {d}} \nu (x)\le 0 \end{array} \end{aligned}$$

and, consequently, taking \(K>M\), we get

$$\begin{aligned} u_{n,k}\le M\quad \nu \hbox {-a.e. in }\Omega . \end{aligned}$$

Similarly, taking \(w=(u_{n,k}+M)^-\), we get

$$\begin{aligned} \begin{array}{l} \displaystyle \int _\Omega \left( T_K(|u_{n,k}(x)|^{p-2}u_{n,k}^-(x))+M^{p-1}\right) (u_{n,k}(x)+M)^-{\text {d}}\nu (x) \\ \ge \displaystyle k\int _{\Omega }\left( \varphi (x)+\frac{1}{k} M^{p-1}\right) (u_{n,k}+M)^- {\text {d}} \nu (x)\ge 0 \end{array} \end{aligned}$$

which yields, taking also \(K>M\),

$$\begin{aligned} u_{n,k}\ge -M\quad \nu \hbox {-a.e. in } {\Omega }. \end{aligned}$$

Therefore,

$$\begin{aligned} \Vert u_{n,k}\Vert _{ L^\infty (\Omega ,\nu )}\le M \end{aligned}$$

as desired.

Now, taking

$$\begin{aligned} K>\max \left\{ M, (\gamma _+)_k(M), -(\gamma _-)_k(-M), (\beta _+)_n(M), -(\beta _-)_n(-M)\right\} , \end{aligned}$$

equations (2.14) and (2.15) yield

$$\begin{aligned}{} & {} (\gamma _+)_k(u_{n,k}(x))+(\gamma _-)_n (u_{n,k}(x))-\displaystyle \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y) \nonumber \\{} & {} \qquad \displaystyle +\frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)=\varphi (x), \ \ \ x\in \Omega _1, \end{aligned}$$
(2.17)

and

$$\begin{aligned}{} & {} (\beta _+)_k(u_{n,k}(x))+(\beta _-)_n (u_{n,k}(x))-\displaystyle \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y)\nonumber \\{} & {} \qquad \displaystyle +\frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)=\varphi (x), \ \ \ x\in \Omega _2. \end{aligned}$$
(2.18)

Take now \(\varphi \in L^{p'}(\Omega ,\nu )\) and, for \(n,k\in \mathbb {N}\), set

$$\begin{aligned} \varphi _{n,k}:=\sup \{\inf \{n,\varphi \},-k\}. \end{aligned}$$
(2.19)

Then, since \(\varphi _{n,k}\in L^\infty (\Omega ,\nu )\), by the previous computations leading to (2.17) and (2.18), there exists a solution \(u_{n,k}\in L^\infty (\Omega ,\nu )\) of the following approximate problem (2.20)–(2.21):

$$\begin{aligned}{} & {} (\gamma _+)_k(u_{n,k}(x))+(\gamma _-)_n(u_{n,k}(x))-\displaystyle \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y)\nonumber \\{} & {} \displaystyle \qquad +\frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)=\varphi _{n,k}(x) , \ \ \ x\in \Omega _1, \end{aligned}$$
(2.20)
$$\begin{aligned}{} & {} (\beta _+)_k(u_{n,k}(x))+(\beta _-)_n(u_{n,k}(x))-\displaystyle \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y)\nonumber \\{} & {} \displaystyle \qquad +\frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)=\varphi _{n,k}(x) , \ \ \ x\in \Omega _2. \end{aligned}$$
(2.21)

Moreover, we obtain the following estimates which will be used later on. Multiplying (2.20) and (2.21) by \(\frac{1}{s} T_s(u_{n,k}^+)\), integrating with respect to \(\nu \) over \(\Omega _1\) and \(\Omega _2\), respectively, adding both equations, applying the integration by parts formula (Remark 2.5), and letting \(s\downarrow 0\), we get, after neglecting some nonnegative terms, that

$$\begin{aligned} \frac{1}{n}{} & {} \int _{\Omega }|u_{n,k}|^{p-2}u_{n,k}^+{\text {d}}\nu +\int _{\Omega _1}(\gamma _+)_k(u_{n,k}){\text {d}}\nu +\int _{\Omega _2}(\beta _+)_k(u_{n,k}){\text {d}}\nu \le \int _{\Omega }\varphi _{n,k}^+{\text {d}}\nu \le \int _{\Omega }\varphi ^+{\text {d}}\nu .\nonumber \\ \end{aligned}$$
(2.22)

Similarly, multiplying by \(\frac{1}{s} T_s(u_{n,k}^-)\), we get

$$\begin{aligned} -\frac{1}{k} \int _{\Omega }|u_{n,k}|^{p-2}u_{n,k}^-{\text {d}}\nu{} & {} +\int _{\Omega _1}(\gamma _-)_n(u_{n,k}){\text {d}}\nu +\int _{\Omega _2}(\beta _-)_n(u_{n,k}){\text {d}}\nu \nonumber \\ {}{} & {} \ge -\int _{\Omega }\varphi _{n,k}^-{\text {d}}\nu \ge -\int _{\Omega }\varphi ^-{\text {d}}\nu . \end{aligned}$$
(2.23)

2.2.2 Monotonicity of the solutions of the approximate problems

Using that \(\varphi _{n,k}\) is nondecreasing in n and nonincreasing in k, and thanks to the way in which we have approximated the maximal monotone graphs \(\gamma \) and \(\beta \), we obtain monotonicity properties for the solutions of the approximate problems.

Fix \(k\in {\mathbb {N}}\). Let \(n_1<n_2\). Multiply equations (2.20) and (2.21) with \(n=n_1\) by \((u_{n_1,k}-u_{n_2,k})^+\), integrate with respect to \(\nu \) over \(\Omega _1\) and \(\Omega _2\), respectively, and add both equations. Then, doing the same with \(n=n_2\) and subtracting the resulting equation from the one that we have obtained for \(n=n_1\), we get

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1}\left( (\gamma _+)_k(u_{n_1,k}(x))-(\gamma _+)_k(u_{n_2,k}(x))\right) (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x)\\{} & {} \displaystyle \qquad +\int _{\Omega _1}\left( (\gamma _-)_{n_1}(u_{n_1,k}(x))-(\gamma _-)_{n_2}(u_{n_2,k}(x))\right) (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x)\\{} & {} \displaystyle \qquad +\int _{\Omega _2}\left( (\beta _+)_k(u_{n_1,k}(x))-(\beta _+)_k(u_{n_2,k}(x))\right) (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x)\\{} & {} \displaystyle \qquad +\int _{\Omega _2}\left( (\beta _-)_{n_1}(u_{n_1,k}(x))-(\beta _-)_{n_2}(u_{n_2,k}(x))\right) (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x)\\{} & {} \displaystyle \qquad -\int _{\Omega }\int _{\Omega }(\textbf{a}_p(x,y,u_{n_1,k}(y)-u_{n_1,k}(x))-\textbf{a}_p(x,y,u_{n_2,k}(y)-u_{n_2,k}(x)))\\{} & {} \qquad \qquad \quad \displaystyle \times (u_{n_1,k}(x)-u_{n_2,k}(x))^+ dm_x(y){\text {d}}\nu (x)\\{} & {} \displaystyle \qquad +\int _{\Omega }\left( \frac{1}{n_1}|u_{n_1,k}(x)|^{p-2}u_{n_1,k}^+(x) -\frac{1}{n_2}|u_{n_2,k}(x)|^{p-2}u_{n_2,k}^+(x) \right) \\{} & {} \quad \displaystyle \times (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x) \\{} & {} \displaystyle \qquad -\frac{1}{k}\int _{\Omega }\left( |u_{n_1,k}(x)|^{p-2}u_{n_1,k}^- (x)-|u_{n_2,k}(x)|^{p-2}u_{n_2,k}^-(x)\right) \\{} & {} \quad \displaystyle \times (u_{n_1,k}(x) -u_{n_2,k}(x))^+{\text {d}}\nu (x) \\{} & {} \displaystyle =\int _\Omega \left( \varphi _{n_1,k}(x)-\varphi _{n_2,k}(x)\right) (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x)\le 0. \end{aligned}$$

Since \((\gamma _+)_k\) and \((\beta _+)_k\) are maximal monotone, the first and third summands on the left-hand side are nonnegative, and the same is true for the second and fourth summands since \((\gamma _-)_{n_1}\ge (\gamma _-)_{n_2}\), \((\beta _-)_{n_1}\ge (\beta _-)_{n_2}\) and these are all maximal monotone. The fifth summand is also nonnegative as illustrated in Remark 2.5. Then, since the last two summands are obviously nonnegative, we get that, in fact,

$$\begin{aligned} \int _{\Omega }\left( \frac{1}{n_1}|u_{n_1,k}(x)|^{p-2}u_{n_1,k}^+(x) -\frac{1}{n_2}|u_{n_2,k}(x)|^{p-2}u_{n_2,k}^+(x) \right) (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x)=0 \end{aligned}$$

and

$$\begin{aligned} \frac{1}{k}\int _{\Omega }\left( |u_{n_1,k}(x)|^{p-2}u_{n_1,k}^-(x)-|u_{n_2,k}(x)|^{p-2}u_{n_2,k}^ -(x)\right) (u_{n_1,k}(x)-u_{n_2,k}(x))^+{\text {d}}\nu (x)=0 \end{aligned}$$

which together imply that

$$\begin{aligned} u_{n_1,k}(x)\le u_{n_2,k}(x)\quad \hbox {for } \nu \hbox {-a.e. } x\in \Omega . \end{aligned}$$

Similarly, we obtain that, for a fixed n, \(u_{n,k}\) is \(\nu \)-a.e. in \(\Omega \) nonincreasing in k.

2.2.3 An \(L^{p}\)-estimate for the solutions of the approximate problems

Multiplying (2.20) and (2.21) by

$$\begin{aligned} u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu , \end{aligned}$$

integrating with respect to \(\nu \) over \(\Omega _1\) and \(\Omega _2\), respectively, adding both equations and using the integration by parts formula (Remark 2.5) we get

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1}\left( (\gamma _+)_k(u_{n,k}(x))+(\gamma _-)_n (u_{n,k}(x))\right) \left( u_{n,k}(x)-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x) \nonumber \\ \qquad{} & {} \displaystyle \quad + \int _{\Omega _2}\left( (\beta _+)_k(u_{n,k}(x))+(\beta _-)_n(u_{n,k}(x))\right) \left( u_{n,k}(x)-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x) \nonumber \\ \qquad{} & {} \displaystyle \quad + \frac{1}{2} \int _{\Omega }\int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))(u_{n,k}(y)-u_{n,k}(x))dm_x(y){\text {d}}\nu (x) \nonumber \\ \qquad{} & {} \displaystyle \quad + \int _\Omega \left( \frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)\right) \nonumber \\ \qquad{} & {} \times \displaystyle \left( u_{n,k}(x)-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x) \nonumber \\{} & {} \displaystyle = \int _{\Omega }\varphi _{n,k}(x)\left( u_{n,k}(x)-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x). \end{aligned}$$
(2.24)

For the first summand on the left-hand side of (2.24), we have

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1}\left( (\gamma _+)_k(u_{n,k})+(\gamma _-)_n(u_{n,k})\right) \left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle =\int _{\Omega _1}\left( (\gamma _+)_k(u_{n,k})-(\gamma _+)_k\left( \frac{1}{\nu (\Omega _1)} \int _{\Omega _1}u_{n,k}\right) \right) \left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle + \int _{\Omega _1}\left( (\gamma _-)_n(u_{n,k})-(\gamma _-)_n\left( \frac{1}{\nu (\Omega _1)} \int _{\Omega _1}u_{n,k}\right) \right) \left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \ge 0, \end{aligned}$$

and for the second

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _2}\left( (\beta _+)_k(u_{n,k})+(\beta _-)_n (u_{n,k})\right) \left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle =\int _{\Omega _2}\left( (\beta _+)_k(u_{n,k})-(\beta _+)_k\left( \frac{1}{\nu (\Omega _2)} \int _{\Omega _2}u_{n,k}\right) \right) \left( u_{n,k}-\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle \qquad +\int _{\Omega _2}\left( (\beta _-)_n(u_{n,k})-(\beta _-)_n\left( \frac{1}{\nu (\Omega _2)} \int _{\Omega _2}u_{n,k}\right) \right) \left( u_{n,k}-\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle \qquad -\int _{\Omega _2}\left( (\beta _+)_k(u_{n,k})+(\beta _-)_n(u_{n,k})\right) \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle \ge -\int _{\Omega _2}\left( (\beta _+)_k(u_{n,k})+(\beta _-)_n(u_{n,k})\right) \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu . \end{aligned}$$

Since \( F_{n,k} (s):=\frac{1}{n} |s|^{p-2}s^+-\frac{1}{k} |s|^{p-2}s^-\) is nondecreasing, for the fourth summand on the left-hand side of (2.24) we have that

$$\begin{aligned} \begin{array}{l} \displaystyle \int _\Omega \left( \frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)\right) \left( u_{n,k}(x)-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x)\\ \displaystyle =\int _{\Omega _1} \left( F_{n,k} (u_{n,k}(x)) - F_{n,k} \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) \right) \left( u_{n,k}(x)-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x)\\ \displaystyle \qquad + \int _{\Omega _2} \left( F_{n,k} (u_{n,k}(x)) - F_{n,k} \left( \frac{1}{\nu (\Omega _2)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) \right) \left( u_{n,k}(x)-\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x)\\ \displaystyle \qquad - \int _{\Omega _2} F_{n,k} (u_{n,k}(x)) \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x)\\ \displaystyle \ge -\int _{\Omega _2} F_{n,k} (u_{n,k}(x)) \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu (x). \end{array} \end{aligned}$$

Finally, recalling (2.5) for the third summand in (2.24), we get

$$\begin{aligned}{} & {} \displaystyle \frac{c_p}{2}\int _{\Omega }\int _{\Omega }|u_{n,k}(y)-u_{n,k}(x)|^p dm_x(y){\text {d}}\nu (x) \\ {}{} & {} \displaystyle \le \int _{\Omega }\varphi _{n,k} \left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle \qquad +\int _{\Omega _2}\left( (\beta _+)_k(u_{n,k})+(\beta _-)_n(u_{n,k})\right) \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{r,n,k}{\text {d}}\nu \right) {\text {d}}\nu \\{} & {} \displaystyle \qquad + \int _{\Omega _2}\left( \frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x) -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)\right) \\{} & {} \displaystyle \qquad \times \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu . \end{aligned}$$

Now, by Hölder’s inequality and the generalised Poincaré-type inequality with \(l=\nu (\Omega _1)\) (let \(\Lambda _1\) denote the constant appearing in the generalised Poincaré-type inequality in Assumption 5),

$$\begin{aligned} \begin{array}{rl} &{} \displaystyle \int _{\Omega }\varphi _{n,k}\left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ &{}\displaystyle \le \Vert \varphi \Vert _{L^{p'}(\Omega ,\nu )}\left\| u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right\| _{L^p(\Omega ,\nu )} \\ &{}\displaystyle \le \Lambda _1\Vert \varphi \Vert _{L^{p'}(\Omega ,\nu )} \left( \int _{\Omega }\int _{\Omega }|u_{n,k}(y)-u_{n,k}(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^{\frac{1}{p}}, \end{array} \end{aligned}$$

and, by (2.22), (2.23) and the generalised Poincaré-type inequality with \(l=\nu (\Omega _1)\) and with \(l=\nu (\Omega _2)\) (let \(\Lambda _2\) denote the constant appearing in the Poincaré-type inequality for the latter case), we obtain

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _2}\left( \left( (\beta _+)_k(u_{n,k})+(\beta _-)_n(u_{n,k})\right) +\frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x)-\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)\right) \\{} & {} \displaystyle \qquad \times \left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right) {\text {d}}\nu \\ {}{} & {} \displaystyle \le \Vert \varphi \Vert _{L^{1}(\Omega ,\nu )} \left| \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu -\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right| \\{} & {} \displaystyle \le \Vert \varphi \Vert _{L^{1}(\Omega ,\nu )}\frac{1}{\nu (\Omega )^{\frac{1}{p}} } \left( \left\| u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right\| _{L^p(\Omega ,\nu )}\right. \\{} & {} \qquad \displaystyle \left. +\left\| u_{n,k}-\frac{1}{\nu (\Omega _2)}\int _{\Omega _2}u_{n,k}{\text {d}}\nu \right\| _{L^p(\Omega ,\nu )}\right) \\{} & {} \displaystyle \le \Vert \varphi \Vert _{L^{1}(\Omega ,\nu )} \frac{\Lambda _1+\Lambda _2}{\nu (\Omega )^{\frac{1}{p}} }\left( \int _{\Omega }\int _{\Omega }|u_{n,k}(y)-u_{n,k}(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^{\frac{1}{p}}. \end{aligned}$$

Therefore, by (2.24) and the subsequent equations,

$$\begin{aligned}{} & {} \frac{c_p}{2} \left( \int _{\Omega }\int _{\Omega }|u_{n,k}(y)-u_{n,k}(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^{\frac{1}{p'}}\nonumber \\{} & {} \displaystyle \le \Lambda _1\Vert \varphi \Vert _{L^{p'}(\Omega ,\nu )}+ \frac{\Lambda _1+\Lambda _2}{\nu (\Omega )^{\frac{1}{p}}}\Vert \varphi \Vert _{L^{1}(\Omega ,\nu )}. \end{aligned}$$
(2.25)

2.2.4 Existence of solutions of \((GP_\varphi )\)

Observe that a solution (uv) of \((GP_\varphi )\) satisfies

$$\begin{aligned} \int _{\Omega _1}v{\text {d}}\nu +\int _{\Omega _2}v{\text {d}}\nu =\int _\Omega \varphi , \end{aligned}$$

therefore, since \(v\in \gamma (u)\) in \(\Omega _1 \) and \(v\in \beta (u)\) in \(\Omega _2\), we need \(\varphi \) to satisfy

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-\le \int _\Omega \varphi {\text {d}}\nu \le \mathcal {R}_{\gamma ,\beta }^+. \end{aligned}$$

We prove the existence of solutions when the inequalities in the previous equation are strict. This suffices for what we need in the next section. Recall that we are working under the Assumptions 1 to 5.

Theorem 2.7

Given \(\varphi \in L^{p'}(\Omega ,\nu )\) such that

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\int _\Omega \varphi {\text {d}}\nu <\mathcal {R}_{\gamma ,\beta }^+, \end{aligned}$$

Problem \((GP_\varphi )\) stated in (2.8) has a solution.

Observe then that any solution (uv) of \((GP_\varphi )\) under such assumptions also satisfies

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\int _\Omega v{\text {d}}\nu <\mathcal {R}_{\gamma ,\beta }^+. \end{aligned}$$

This will be used later on.

We divide the proof into three cases.

Proof

(Proof of Theorem 2.7 when \(\mathcal {R}_{\gamma ,\beta }^\pm =\pm \infty \)) Suppose that

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-= -\infty \ \hbox { and } \mathcal {R}_{\gamma ,\beta }^+= +\infty . \end{aligned}$$

Let \(\varphi \in L^{p'}(\Omega ,\nu )\), \(\varphi _{n,k}\) defined as in (2.19) and let \(u_{n,k}\in L^\infty (\Omega ,\nu )\), \(n, k\in {\mathbb {N}}\), be solutions of the Approximate Problem (2.20)–(2.21).

Step A (Boundedness). Let us first see that \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\) is bounded.

Step 1. We start by proving that \(\{\Vert u_{n,k}^+\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\) is bounded. We see this case by case. Since \(\mathcal {R}_{\gamma ,\beta }^+=+\infty \), then \(\sup \text{ Ran }(\gamma )=+\infty \) or \(\sup \text{ Ran }(\beta )=+\infty \).

Case 1.1. Suppose that \(\sup \text{ Ran }(\gamma )=+\infty \). Then, by (2.22),

$$\begin{aligned} \int _{\Omega _1}(\gamma _+)_k(u_{n,k}){\text {d}}\nu \le M:=\int _\Omega \varphi {\text {d}}\nu \quad \hbox {for every } n, k\in {\mathbb {N}}. \end{aligned}$$

Let \(z^+_{n,k}:= (\gamma _+)_k(u_{n,k})\) and \(\widetilde{\Omega }_{n,k}:=\left\{ x\in \Omega _1: z_{n,k}^+(x)<\frac{2\,M}{\nu (\Omega _1)}\right\} \). Then,

$$\begin{aligned} 0\le \int _{\widetilde{\Omega }_{n,k}}z_{n,k}^+{\text {d}}\nu= & {} \int _{\Omega _1}z_{n,k}^+{\text {d}}\nu -\int _{\Omega _1\setminus \widetilde{\Omega }_{n,k}}z_{n,k}^+{\text {d}}\nu \le M-(\nu (\Omega _1)-\nu (\widetilde{\Omega }_{n,k}))\frac{2M}{\nu (\Omega _1)}\\= & {} \nu (\widetilde{\Omega }_{n,k})\frac{2M}{\nu (\Omega _1)}-M, \end{aligned}$$

from where

$$\begin{aligned} \nu (\widetilde{\Omega }_{n,k})\ge \frac{\nu (\Omega _1)}{2}. \end{aligned}$$

Case 1.1.1. Assume that \(\sup D(\gamma )=+\infty \). Let \(r_0\in {\mathbb {R}}\) be such that \(\gamma ^0(r_0)>2M/\nu (\Omega _1)\) and let \(k_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \frac{2M}{\nu (\Omega _1)}<(\gamma _+)_k(r_0)\le \gamma ^0(r_0)\quad \hbox {for } k\ge k_0. \end{aligned}$$
(2.26)

Then, since in \(\widetilde{\Omega }_{n,k}\), \((\gamma _+)_k(u_{n,k})=z^+_{n,k}<\frac{2\,M}{\nu (\Omega _1)}\), from (2.26) we get that

$$\begin{aligned} u_{n,k}^+ \le r_0\quad \hbox {in } \widetilde{\Omega }_{n,k}\quad \hbox {for every } k\ge k_0 \hbox { and every } n\in {\mathbb {N}}. \end{aligned}$$

Therefore, this bound, the generalised Poincaré-type inequality with \(l=\frac{\nu (\Omega _1)}{2}\) and (2.25) yield the boundedness of \(\{\Vert u_{n,k}^+\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\).

Case 1.1.2. Suppose that \(r_\gamma :=\sup D(\gamma )<+\infty \) and let \(h>0\). Since \(r_\gamma +h\not \in D(\gamma )\),

$$\begin{aligned} (\gamma _+)_k(r_\gamma +h)\uparrow +\infty \ \hbox { as }k\rightarrow +\infty . \end{aligned}$$

Take \(k_0\in {\mathbb {N}}\) such that \((\gamma _+)_k(r_\gamma +h)\ge \frac{2\,M}{\nu (\Omega _1)}\) for every \(k\ge k_0\). Then,

$$\begin{aligned} (\gamma _+)_k(u_{n,k}^+)<\frac{2M}{\nu (\Omega _1)}\le (\gamma _+)_k(r_\gamma +h) \quad \hbox {in } \widetilde{\Omega }_{n,k}\quad \hbox {for every } k\ge k_0 \hbox { and every } n\in {\mathbb {N}}, \end{aligned}$$

thus

$$\begin{aligned} u_{n,k}^+\le r_\gamma +h \quad \hbox {in } \widetilde{\Omega }_{n,k}\quad \hbox {for every } k\ge k_0 \hbox { and every } n\in {\mathbb {N}}. \end{aligned}$$

Therefore, again, this bound together with the generalised Poincaré-type inequality with \(l=\frac{\nu (\Omega _1)}{2}\) and (2.22) yield the boundedness of \(\{\Vert u_{n,k}^+\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\).

Case 1.2. If \( \sup \text{ Ran }(\beta )=+\infty \), we proceed similarly.

Step 2. Using that \(\mathcal {R}_{\gamma ,\beta }^-=-\infty \) we obtain that \(\{\Vert u_{n,k}^-\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\) is bounded with an analogous argument.

Consequently, we get that \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\) is bounded as desired.

Step B (Taking limits in n). The monotonicity properties obtained in Subsect. 2.2.2 together with the boundedness of \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\) allow us to apply the monotone convergence theorem to obtain \(u_k\in L^p(\Omega ,\nu )\), \(k\in {\mathbb {N}}\), and \(u\in L^p(\Omega ,\nu )\) such that, taking a subsequence if necessary, \(u_{n,k}{\mathop {\rightarrow }\limits ^{n}} u_k\) in \(L^p(\Omega ,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega \) for \(k\in {\mathbb {N}}\), and \(u_{k}{\mathop {\rightarrow }\limits ^{k}} u\) in \(L^p(\Omega ,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega \).

We now want to take limits, in n and then in k, in (2.20) and (2.21). Since \(u_{n,k}{\mathop {\rightarrow }\limits ^{n}} u_k\) in \(L^p(\Omega ,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega \),

$$\begin{aligned}&\displaystyle \int _{\Omega }\textbf{a}_p(\cdot ,y,u_{n,k}(y)\!-\!u_{n,k}(\cdot ))dm_{(\cdot )}(y) {\mathop {\longrightarrow }\limits ^{n}}\!\!\!\int _{\Omega }\textbf{a}_p(\cdot ,y,u_{k}(y)\!-\!u_{k}(\cdot ))dm_{(\cdot )}(y), \\&\displaystyle \frac{1}{n}|u_{n,k}|^{p-2}u_{n,k}^+{\mathop {\longrightarrow }\limits ^{n}} 0 \nonumber \end{aligned}$$
(2.27)

and

$$\begin{aligned} \frac{1}{k}|u_{n,k}|^{p-2}u_{n,k}^-{\mathop {\longrightarrow }\limits ^{n}} \frac{1}{k}|u_{k}|^{p-2}u_{k}^- \end{aligned}$$

in \(L^{p'}(\Omega ,\nu )\) and, up to a subsequence, for \(\nu \)-a.e. \(x\in \Omega \). Indeed, the second and third limits follow because \(|u_{n,k}|^{p-2}u_{n,k}^+{\mathop {\rightarrow }\limits ^{n}}|u_{k}|^{p-2}u_{k}^+\) in \(L^{p'}(\Omega ,\nu )\). Now, since \(\{u_{n,k}\}_n\) is nonincreasing in n, \(|u_{n,k}|\le \max \{|u_{1,k}|,|u_k|\}\) \(\nu \)-a.e. in \(\Omega \), for every n, \(k\in {\mathbb {N}}\), so Lemma 2.1 yields the convergence (2.27) in \(L^{p'}(\Omega ,\nu )\).

Now, isolating \((\gamma _+)_k(u_{n,k})+(\gamma _-)_n(u_{n,k})\) and \((\beta _+)_k(u_{n,k})+(\beta _-)_n(u_{n,k})\) in equations (2.20) and (2.21), respectively, and taking the positive parts, we get that

$$\begin{aligned} \displaystyle (\gamma _+)_k(u_{n,k}(x))= & {} \left( \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y)+\frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x)\right. \\{} & {} \displaystyle \left. -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x) +\varphi _{n,k}(x)\right) ^+ \end{aligned}$$

for \(x\in \Omega _1\), and

$$\begin{aligned} \displaystyle (\beta _+)_k(u_{n,k}(x))= & {} \left( \int _{\Omega }\textbf{a}_p(x,y,u_{n,k}(y)-u_{n,k}(x))dm_x(y)+\frac{1}{n}|u_{n,k}(x)|^{p-2}u_{n,k}^+(x)\right. \\{} & {} \displaystyle \left. -\frac{1}{k}|u_{n,k}(x)|^{p-2}u_{n,k}^-(x)+\varphi _{n,k}(x)\right) ^+ \end{aligned}$$

for \(x\in \Omega _2\). Therefore, since the right-hand sides of these equations converge in \(L^{p'}(\Omega _1,\nu )\) and \(L^{p'}(\Omega _2,\nu )\) (and also \(\nu \)-a.e. in \(\Omega _1\) and \(\Omega _2\)), respectively, there exist \(z^+_k\in L^{p'}(\Omega _1,\nu )\) and \(\omega ^+_k\in L^{p'}(\Omega _2,\nu )\) such that \((\gamma _+)_k(u_{n,k}){\mathop {\rightarrow }\limits ^{n}}z^+_k\) in \(L^{p'}(\Omega _1,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega _1\), and \((\beta _+)_k(u_{n,k}){\mathop {\rightarrow }\limits ^{n}}\omega ^+_k\) in \(L^{p'}(\Omega _2,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega _2\). Moreover, since \((\gamma _+)_k\) and \((\beta _+)_k\) are maximal monotone graphs, \(z^+_k= (\gamma _+)_k(u_{k})\) \(\nu \)-a.e. in \(\Omega _1\), and \(\omega ^+_k= (\beta _+)_k(u_{k})\) \(\nu \)-a.e. in \(\Omega _2\).

Similarly, taking the negative parts, there exist

$$\begin{aligned} \lim _{n\rightarrow +\infty }(\gamma _-)_n(u_{n,k}(x))=z_k^-(x)\ \hbox { in } L^{p'}(\Omega _1,\nu ) \hbox { and for } \nu \hbox {-a.e. } x\in \Omega _1, \end{aligned}$$

and

$$\begin{aligned} \lim _{n\rightarrow +\infty }(\beta _-)_n(u_{n,k}(x))=\omega _k^-(x) \ \hbox { in } L^{p'}(\Omega _2,\nu )\hbox { and for } \nu \hbox {-a.e. } x\in \Omega _2. \end{aligned}$$

Moreover, by [15, Lemma G], \(z^-_k\in \gamma _-(u_{k})\) and \(\omega ^-_k\in \beta _-(u_{k})\). Therefore, we have obtained that

$$\begin{aligned}{} & {} z_k^+(x)+z^-_k(x)-\int _{\Omega } \textbf{a}_p(x,y,u_{k}(y)-u_{k}(x))dm_x(y) \nonumber \\{} & {} -\frac{1}{k}|u_{k}(x)|^{p-2}u_{k}^-(x) = \varphi _{k}(x), \end{aligned}$$
(2.28)

for \(\nu \)-a.e. \(x\in \Omega _1\), and

$$\begin{aligned}{} & {} \omega _k^+(x)+\omega _k^-(x)- \int _{\Omega }\textbf{a}_p(x,y,u_{k}(y)-u_{k}(x))dm_x(y)\nonumber \\{} & {} -\frac{1}{k} |u_{k}(x)|^{p-2} u_{k}^-(x) =\varphi _{k}(x) \end{aligned}$$
(2.29)

for \(\nu \)-a.e. \(x\in \Omega _2\).

Step C (Taking limits in k). Now again, isolating \(z_k^++z^-_k\) and \(\omega _k^++\omega _k^-\) in equations (2.28) and (2.29), respectively, and taking the positive and negative parts as above, we get that there exist \(z^+\in L^{p'}(\Omega _1,\nu )\), \(z^-\in L^{p'}(\Omega _1,\nu )\), \(\omega ^+\in L^{p'}(\Omega _2,\nu )\) and \(\omega ^-\in L^{p'}(\Omega _2,\nu )\) such that \(z^+_k {\mathop {\rightarrow }\limits ^{k}}z^+\) and \(z^-_k {\mathop {\rightarrow }\limits ^{k}}z^-\) in \(L^{p'}(\Omega _1,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega _1\), and \(\omega ^+_k {\mathop {\rightarrow }\limits ^{k}}\omega ^+\) and \(\omega ^-_k {\mathop {\rightarrow }\limits ^{k}}\omega ^-\) in \(L^{p'}(\Omega _2,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega _2\). In addition, by the maximal monotonicity of \(\gamma _-\) and \(\beta _-\), \(z^-\in \gamma _-(u)\) and \(\omega ^-\in \beta _-(u)\) \(\nu \)-a.e. in \(\Omega _1\) and \(\Omega _2\), respectively. Moreover, by [15, Lemma G], \(z^+\in \gamma _+(u)\) and \(\omega ^+\in \beta _+(u)\) \(\nu \)-a.e. in \(\Omega _1\) and \(\Omega _2\), respectively.

Consequently,

$$\begin{aligned} z(x)-\int _{\Omega }\textbf{a}_p(x,y,u(y)-u(x))dm_x(y)=\varphi (x) \ \hbox { for } \nu \hbox {-a.e. } x\in \Omega _1, \end{aligned}$$

and

$$\begin{aligned} \omega (x)-\int _{\Omega }\textbf{a}_p(x,y,u(y)-u(x))dm_x(y)=\varphi (x) \ \hbox { for } \nu \hbox {-a.e. }x\in \Omega _2, \end{aligned}$$

where \(z=z^++z^- \in \gamma (u)\) \(\nu \)-a.e. in \(\Omega _1\) and \(\omega =\omega ^++\omega ^-\in \beta (u)\) \(\nu \)-a.e. in \(\Omega _2\). The proof of existence in this case is done. \(\square \)

Proof

(Proof of Theorem 2.7 when \(\mathcal {R}_{\gamma ,\beta }^\pm \) are finite) Suppose that

$$\begin{aligned} -\infty<\mathcal {R}_{\gamma ,\beta }^-<\mathcal {R}_{\gamma ,\beta }^+< +\infty . \end{aligned}$$

Let \(\varphi \in L^{p'}(\Omega ,\nu )\), and assume that it satisfies

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\int _{\Omega }\varphi {\text {d}}\nu < \mathcal {R}_{\gamma ,\beta }^+. \end{aligned}$$

Then, for \(\varphi _{n,k}\) defined as in (2.19), there exist \(M_1, M_2\in {\mathbb {R}}\) and \(n_0, k_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<M_2<\int _{\Omega }\varphi _{n,k}{\text {d}}\nu< M_1<\mathcal {R}_{\gamma ,\beta }^+ \end{aligned}$$
(2.30)

for every \(n\ge n_0\) and \(k\ge k_0\). For \(n, k\in {\mathbb {N}}\) let \(u_{n,k}\in L^\infty (\Omega ,\nu )\) be the solution of the Approximate Problem (2.20)–(2.21), and let

$$\begin{aligned} M_3:=\sup _{n,k\in {\mathbb {N}}} \left\| u_{n,k} - \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right\| _{L^p(\Omega ,\nu )}<+\infty . \end{aligned}$$
(2.31)

Observe that \(M_3\) is finite by the generalised Poincaré-type inequality together with (2.25). Let \(k_1\in {\mathbb {N}}\) such that \(k_1\ge k_0\) and \(M_1+\frac{1}{k}M_3\nu (\Omega )^{\frac{1}{p(p-1)}}<\mathcal {R}_{\gamma ,\beta }^+\) for every \(k\ge k_1\).

Step D (Boundedness of \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_n\) and passing to the limit in n) Let us see that, for each \(k\in \mathbb {N}\), \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_n\) is bounded. Fix \(k\ge k_1\) and suppose that \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_{n}\) is not bounded. Then, by (2.31), since \(u_{n,k}\) is nondecreasing in n,

$$\begin{aligned} \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu {\mathop {\longrightarrow }\limits ^{n\rightarrow +\infty }} +\infty . \end{aligned}$$

Thus, using again that \(u_{n,k}\) is nondecreasing in n, there exists \(n_1\ge n_0\) such that

$$\begin{aligned} \begin{array}{rl} \displaystyle u_{n,k}^-&{}\displaystyle \le \left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) ^-+\left( \frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) ^-\\ &{}\displaystyle =\left( u_{n,k}-\frac{1}{\nu (\Omega _1)}\int _{\Omega _1}u_{n,k}{\text {d}}\nu \right) ^- \end{array} \end{aligned}$$

for every \(n\ge n_1\), and thus

$$\begin{aligned} \Vert u_{n,k}^-\Vert _{L^p(\Omega ,\nu )}\le M_3 \ \hbox { for every } n\ge n_1. \end{aligned}$$

Consequently, \(\Vert u_{n,k}^-\Vert _{L^{p-1}(\Omega ,\nu )}\le M_3\nu (\Omega )^{\frac{1}{p(p-1)}}\) for \(n\ge n_1\). Then, with this bound and (2.30) at hand, integrating (2.20) and (2.21) with respect to \(\nu \) over \(\Omega _1\) and \(\Omega _2\), respectively, adding both equations and neglecting some nonnegative terms we get

$$\begin{aligned} \begin{array}{l} \displaystyle \int _{\Omega _1}\underbrace{(\gamma _+)_k(u_{n,k}(x))+(\gamma _-)_n(u_{n,k}(x))}_{z_{n,k}(x)} {\text {d}}\nu (x)+\int _{\Omega _2}\underbrace{(\beta _+)_k(u_{n,k}(x))+(\beta _-)_n(u_{n,k}(x))}_{\omega _{n,k}(x)}{\text {d}}\nu (x) \\ \displaystyle \ \le \underbrace{M_1 +\frac{1}{k}M_3\nu (\Omega )^{\frac{1}{p(p-1)}}}_{M_4}< \mathcal {R}_{\gamma ,\beta }^+. \end{array} \end{aligned}$$

Therefore, for each \(n\in {\mathbb {N}}\), either

$$\begin{aligned} \int _{\Omega _1}z_{n,k}{\text {d}}\nu <\nu (\Omega _1)\sup \text{ Ran }(\gamma )-\frac{\delta }{2} \end{aligned}$$
(2.32)

or

$$\begin{aligned} \int _{\Omega _2}\omega _{n,k} {\text {d}}\nu <\nu (\Omega _2)\sup \text{ Ran }(\beta )-\frac{\delta }{2}, \end{aligned}$$
(2.33)

where \(\delta :=\mathcal {R}_{\gamma ,\beta }^+- M_4>0\).

For \(n\in {\mathbb {N}}\) such that (2.32) holds let

$$\begin{aligned} K_{n,k}:=\left\{ x\in \Omega _1 \,: \, z_{n,k}(x)<\sup \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)}\right\} . \end{aligned}$$

Then

$$\begin{aligned} \int _{K_{n,k}}z_{n,k}{\text {d}}\nu =\int _{\Omega _1}z_{n,k}{\text {d}}\nu -\int _{\Omega _1\setminus K_{n,k}}z_{n,k}{\text {d}}\nu <-\frac{\delta }{4}+\nu (K_{n,k})\left( \sup \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)}\right) , \end{aligned}$$

and

$$\begin{aligned} \int _{K_{n,k}}z_{n,k}{\text {d}}\nu \ge \nu (K_{n,k})\inf \text{ Ran }(\gamma ). \end{aligned}$$

Therefore,

$$\begin{aligned} \nu (K_{n,k})\left( \sup \text{ Ran }(\gamma )-\inf \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)}\right) \ge \frac{\delta }{4}, \end{aligned}$$

thus \(\nu (K_{n,k})>0\), \(\displaystyle \sup \text{ Ran }(\gamma )-\inf \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)}>0\) and

$$\begin{aligned} \nu (K_{n,k})\ge \frac{\delta /4}{\sup \text{ Ran }(\gamma )-\inf \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)}}. \end{aligned}$$

Note that, if \(\sup \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)}\le 0\), then \(z_{n,k}\le 0\) in \(K_{n,k}\), thus \(u_{n,k}^+=0\) in \(K_{n,k}\) and, consequently, \(\Vert u_{n,k}^+ \Vert _{L^p(K_{n,k},\nu )}=0\). Therefore, by the generalised Poincaré-type inequality and (2.25) we get that \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_{n}\) is bounded, which is a contradiction. We may therefore suppose that \(\sup \text{ Ran }(\gamma ) -\frac{\delta }{4\nu (\Omega _1)} > 0\). Then, for \(k_2\ge k_1\) large enough so that \(\sup \text{ Ran }((\gamma _+)_{k})>\sup \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)}\) for \(k\ge k_2\),

$$\begin{aligned} \Vert u_{n,k}^+\Vert _{L^p(K_{n,k},\nu )}\le \nu (K_{n,k})^{\frac{1}{p}}(\gamma _+)_{k}^{-1}\left( \sup \text{ Ran }(\gamma )-\frac{\delta }{4\nu (\Omega _1)} \right) \end{aligned}$$

and by the generalised Poincaré-type inequality and (2.25) we get that \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_{n}\) is bounded, which is a contradiction. Similarly, for \(n\in {\mathbb {N}}\) such that (2.33) holds.

We have obtained that \(\{\Vert u_{n,k}\Vert _{L^p(\Omega ,\nu )}\}_{n}\) is bounded for each \(k\in {\mathbb {N}}\). Therefore, since \(\{u_{n,k}\}_n\) is nondecreasing in n, we may apply the monotone convergence theorem to obtain \(u_k\in L^p(\Omega ,\nu )\), \(k\in {\mathbb {N}}\), such that \(u_{n,k}{\mathop {\rightarrow }\limits ^{n}} u_k\) in \(L^p(\Omega ,\nu )\) and pointwise \(\nu \)-a.e. in \(\Omega \) for \(k\in {\mathbb {N}}\). Proceeding now like in Step B of the previous proof we get: \(z_k^+\in L^{p'}(\Omega _1,\nu )\) and \(\omega _k^+\in L^{p'}(\Omega _2,\nu )\) such that \(z^+_k\in \gamma _+(u_{k})\) and \(\omega ^+_k\in \beta _+(u_{k})\) \(\nu \)-a.e. in \(\Omega _1\) and \(\Omega _2\), respectively; and \(z_k^-\in L^{p'}(\Omega _1,\nu )\) and \(\omega _k^-\in L^{p'}(\Omega _2,\nu )\) with \(z^-_k\in \gamma _-(u_{k})\) and \(\omega ^-_k\in \beta _-(u_{k})\), \(\nu \)-a.e. \(\Omega _1\) and \(\Omega _2\), respectively, and such that

$$\begin{aligned}{} & {} z_k^+(x)+z^-_k(x)-\int _{\Omega }\textbf{a}_p (x,y,u_{k}(y)-u_{k}(x))dm_x(y) \nonumber \\ {}{} & {} -\frac{1}{k}|u_{k}(x)|^{p-2}u_{k}^-(x) = \varphi _{k}(x), \end{aligned}$$
(2.34)

for \(\nu \)-a.e. every \(x\in \Omega _1\), and

$$\begin{aligned}{} & {} \omega _k^+(x)+\omega _k^-(x)- \int _{\Omega }\textbf{a}_p(x,y,u_{k}(y)-u_{k}(x))dm_x(y) \nonumber \\ {}{} & {} -\frac{1}{k}|u_{k}(x)|^{p-2}u_{k}^-(x) =\varphi _{k}(x) \end{aligned}$$
(2.35)

for \(\nu \)-a.e. every \(x\in \Omega _2\).

Step E (Boundedness of \(\{\Vert u_{k}\Vert _{L^p(\Omega ,\nu )}\}_k\) and passing to the limit in k) We now see that \(\{\Vert u_{k}\Vert _{L^p(\Omega ,\nu )}\}_{k}\) is bounded. Since \(u_k^+\le u_1^+\), it is enough to see that \(\{\Vert u_{k}^-\Vert _{L^p(\Omega ,\nu )}\}_{k}\) is bounded.

Now, (2.34) and (2.35) yield

$$\begin{aligned} \begin{array}{l} \displaystyle \int _{\Omega _1}\underbrace{z_k^+(x)+z^-_k(x)}_{z_{k}(x)} {\text {d}}\nu (x)+\int _{\Omega _2}\underbrace{\omega _k^+(x)+\omega _k^-(x)}_{\omega _{k}(x)}{\text {d}}\nu (x) \ge M_2 > \mathcal {R}_{\gamma ,\beta }^-. \end{array} \end{aligned}$$

Therefore, for each \(k\in {\mathbb {N}}\), either

$$\begin{aligned} \int _{\Omega _1}z_{k}{\text {d}}\nu >\nu (\Omega _1)\inf \text{ Ran }(\gamma )+\frac{\delta '}{2} \end{aligned}$$
(2.36)

or

$$\begin{aligned} \int _{\Omega _2}\omega _{k} {\text {d}}\nu >\nu (\Omega _2)\inf \text{ Ran }(\beta )+\frac{\delta '}{2}, \end{aligned}$$
(2.37)

where \(\delta ':=M_2-\mathcal {R}_{\gamma ,\beta }^->0\).

For \(k\in {\mathbb {N}}\) such that (2.36) holds let \(K_{k}:=\{x\in \Omega _1 \,: \, z_{k}(x)>\inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)}\}\). Then,

$$\begin{aligned} \begin{array}{rl} \displaystyle \int _{K_{k}}z_{k}{\text {d}}\nu &{}\displaystyle =\int _{\Omega _1}z_{k}{\text {d}}\nu -\int _{\Omega _1\setminus K_{k}}z_{k}{\text {d}}\nu \\ &{}\displaystyle > \left( \nu (\Omega _1)\inf \text{ Ran }(\gamma ) +\frac{\delta '}{2}\right) -(\nu (\Omega _1)-\nu ( K_k))\left( \inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)}\right) \\ &{}\displaystyle =\frac{\delta '}{4}+\nu (K_{k})\left( \inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)}\right) , \end{array} \end{aligned}$$

and

$$\begin{aligned} \int _{K_{k}}z_{k}{\text {d}}\nu \le \nu (K_{k})\sup \text{ Ran }(\gamma ). \end{aligned}$$

Therefore,

$$\begin{aligned} \nu (K_{k})\left( \sup \text{ Ran }(\gamma ) -\inf \text{ Ran }(\gamma ) -\frac{\delta '}{4\nu (\Omega _1)}\right) \ge \frac{\delta '}{4}, \end{aligned}$$

thus \(\nu (K_{k})>0\), \(\displaystyle \sup \text{ Ran }(\gamma ) -\inf \text{ Ran }(\gamma ) -\frac{\delta '}{4\nu (\Omega _1)}>0\) and

$$\begin{aligned} \nu (K_{k})\ge \frac{\delta '/4}{\sup \text{ Ran }(\gamma ) -\inf \text{ Ran }(\gamma ) -\frac{\delta '}{4\nu (\Omega _1)}}. \end{aligned}$$

Now, if \(\inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)}\ge 0\) then \(z_{k}\ge 0\) in \(K_{k}\), thus \(u_{n,k}^-=0\) in \(K_{k}\) and \(\Vert u_{k}^-\Vert _{L^p(K_{k},\nu )}=0\); so by the generalised Poincaré-type inequality and (2.25) we get that \(\{\Vert u_{k}\Vert _{L^p(\Omega ,\nu )}\}_{n}\) is bounded. If \(\inf \text{ Ran }(\gamma )+\frac{\delta '}{4\nu (\Omega _1)} < 0\), then

$$\begin{aligned} \Vert u_{k}^-\Vert _{L^p(K_{n,k},\nu )}\le - \nu (K_{k})^{\frac{1}{p}}\gamma _-^{-1}\left( \inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)} \right) \end{aligned}$$

and by the generalised Poincaré inequality and (2.25) we get that \(\{\Vert u_{k}\Vert _{L^p(\Omega ,\nu )}\}_{k}\) is bounded. Similarly, for \(k\in {\mathbb {N}}\) such that (2.37) holds.

Now, proceeding as in Step C of the previous proof, we finish this proof. \(\square \)

Finally, we give the proof of the remaining case.

Proof

(Proof of Theorem 2.7 in the mixed case) Let us see the existence for

$$\begin{aligned} -\infty<\mathcal {R}_{\gamma ,\beta }^-<\mathcal {R}_{\gamma ,\beta }^+= +\infty , \end{aligned}$$
(2.38)

or

$$\begin{aligned} -\infty =\mathcal {R}_{\gamma ,\beta }^-<\mathcal {R}_{\gamma ,\beta }^+< +\infty . \end{aligned}$$
(2.39)

Suppose that (2.38) holds and let \(\varphi \in L^{p'}(\Omega ,\nu )\) satisfying

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\int _{\Omega }\varphi {\text {d}}\nu . \end{aligned}$$

If (2.39) holds and we have \(\varphi \in L^{p'}(\Omega ,\nu )\) satisfying \(\displaystyle \int _{\Omega }\varphi {\text {d}}\nu < \mathcal {R}_{\gamma ,\beta }^+\), the argument is analogous.

Let \(\varphi _{n,k}\) be defined as in (2.19) and let \(u_{n,k}\in L^\infty (\Omega ,\nu )\), \(n, k\in {\mathbb {N}}\), be the solution of the Approximate Problem (2.20)–(2.21). Then, by Lemma A.7 together with (2.22), \(\{\Vert u_{n,k}^+\Vert _{L^p(\Omega ,\nu )}\}_{n,k}\) is bounded. However, for a fixed \(k\in {\mathbb {N}}\), since \(u_{n,k}\) is nondecreasing in n, \(\{\Vert u_{n,k}^-\Vert _{L^p(\Omega ,\nu )}\}_{n}\) is also bounded. Therefore, proceeding as in Step B of the first case, we obtain \(u_k\in L^p(\Omega ,\nu )\), \(z_k^+\), \(z_k^-\in L^{p'}(\Omega _1,\nu )\) and \(\omega _k^+\), \(\omega _k^-\in L^{p'}(\Omega _2,\nu )\), \(k\in {\mathbb {N}}\), such that

$$\begin{aligned}{} & {} z_k^+(x)+z^-_k(x)-\int _{\Omega }\textbf{a}_p (x,y,u_{k}(y)-u_{k}(x))dm_x(y) \nonumber \\{} & {} -\frac{1}{k}|u_{k}(x)|^{p-2}u_{k}^-(x) = \varphi _{k}(x), \end{aligned}$$
(2.40)

for \(\nu \)-a.e. \(x\in \Omega _1\), and

$$\begin{aligned} \begin{array}{l}\displaystyle \omega _k^+(x)+\omega _k^-(x)-\int _{\Omega }\textbf{a}_p(x,y,u_{k}(y)-u_{k}(x))dm_x(y) -\frac{1}{k}|u_{k}(x)|^{p-2}u_{k}^-(x) =\varphi _{k}(x)\end{array} \end{aligned}$$

for \(\nu \)-a.e. \(x\in \Omega _2\); where, for \(k\in {\mathbb {N}}\),

$$\begin{aligned} z^+_k= (\gamma _+)_k(u_{k}),\ z^-_k\in \gamma _-(u_{k}) \ \nu \hbox {-a.e. in } \Omega _1, \end{aligned}$$

and

$$\begin{aligned} \omega ^+_k= (\beta _+)_k(u_{k}),\ \omega ^-_k\in \beta _-(u_{k}) \ \nu \hbox {-a.e. in } \Omega _2. \end{aligned}$$

We now prove that \(\{\Vert u_{k}\Vert _{L^p(\Omega ,\nu )}\}_{k}\) is bounded. Proceeding as in Step E of the previous proof and using the same notation, we get that for each \(k\in {\mathbb {N}}\), either

$$\begin{aligned} \int _{\Omega _1}z_{k}{\text {d}}\nu >\nu (\Omega _1)\inf \text{ Ran }(\gamma ) +\frac{\delta '}{2} \end{aligned}$$
(2.41)

or

$$\begin{aligned} \int _{\Omega _2}\omega _{k} {\text {d}}\nu >\nu (\Omega _2)\inf \text{ Ran }(\beta ) +\frac{\delta '}{2}. \end{aligned}$$
(2.42)

Case 1. For \(k\in {\mathbb {N}}\) such that (2.41) holds, let

$$\begin{aligned} K_{k}:=\left\{ x\in \Omega _1 \,: \, z_{k}(x)>\inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)}\right\} . \end{aligned}$$

Then,

$$\begin{aligned} \int _{K_{k}}z_{k}{\text {d}}\nu =\int _{\Omega _1}z_{k}{\text {d}}\nu -\int _{\Omega _1\setminus K_{k}}z_{k}{\text {d}}\nu >\frac{\delta '}{4}+\nu (K_{k})\left( \inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)}\right) . \nonumber \\ \end{aligned}$$
(2.43)

Now, by (2.40),

$$\begin{aligned}{} & {} \int _{\{x\in \Omega _1 \,: \,z_k(x)>h\}} z_k {\text {d}}\nu \le \int _{\{x\in \Omega _1 \,: \,z_k(x)>h\}}\frac{|z_k|^{p'}}{h^{p'-1}}{\text {d}}\nu \\ \\{} & {} \quad = \frac{1}{h^{p'-1}}\int _{\{x\in \Omega _1 \,: \,z_k(x)>h\}}\left| \int _{\Omega }a_p(x,y,u_k(y)-u_k(x)dm_x(y)+\varphi _k(x)\right| ^{p'}{\text {d}}\nu (x). \end{aligned}$$

Thus, for a constant \(D_1\) independent of k and h,

$$\begin{aligned}{} & {} \int _{\{x\in \Omega _1 \,: \,z_k(x)>h\}} z_k {\text {d}}\nu \\{} & {} \le \frac{D_1}{h^{p'-1}}\left( \int _{\Omega _1}\int _{\Omega }|a_p (x,y,u_k(y)-u_k(x)|^{p'}dm_x(y){\text {d}}\nu (x)+\int _{\Omega _1}|\varphi _k|^{p'}{\text {d}}\nu \right) . \end{aligned}$$

Hence, by (2.4) and (2.25), there exist constants \(D_2\) and \(D_3\), independent of k and h, such that

$$\begin{aligned}{} & {} \int _{\{x\in \Omega _1 \,: \,z_k(x)>h\}} z_k {\text {d}}\nu \\{} & {} \le \frac{D_2}{h^{p'-1}}\left( \int _{\Omega _1}\int _{\Omega }| u_k(y)-u_k(x)|^{p}dm_x(y){\text {d}}\nu (x)+\int _{\Omega _1}|\varphi _k|^{p'}{\text {d}}\nu +1\right) \le \frac{D_3}{h^{p'-1}}. \end{aligned}$$

Consequently, we may find \(h>0\) such that

$$\begin{aligned} \sup _{k\in {\mathbb {N}}}\int _{\{x\in \Omega _1 \,: \,z_k(x)>h\}} z_k {\text {d}}\nu < \frac{\delta '}{8}. \end{aligned}$$

Therefore,

$$\begin{aligned} \int _{K_{k}}z_{k}{\text {d}}\nu =\int _{K_{k}\cap \{z_k>h\}}z_{k}{\text {d}}\nu + \int _{K_{k}\cap \{z_k\le h\}}z_{k}{\text {d}}\nu \le \frac{\delta '}{8}+\nu (K_k)h. \end{aligned}$$

Recalling (2.43), we get

$$\begin{aligned} \frac{\delta '}{4}+\nu (K_{k})\left( \inf \text{ Ran }(\gamma ) +\frac{\delta '}{4\nu (\Omega _1)}\right) < \frac{\delta '}{8}+\nu (K_k)h, \end{aligned}$$

thus

$$\begin{aligned} \frac{\delta '}{8}< \nu (K_{k})\left( h-\inf \text{ Ran }(\gamma ) -\frac{\delta '}{4\nu (\Omega _1)}\right) . \end{aligned}$$

Consequently, \(\displaystyle h-\inf \text{ Ran }(\gamma ) -\frac{\delta '}{4\nu (\Omega _1)}>0\) and

$$\begin{aligned} \nu (K_{k})\ge \frac{\delta '/4}{h-\inf \text{ Ran }(\gamma ) -\frac{\delta '}{4\nu (\Omega _1)}}>0. \end{aligned}$$

From here we conclude as in the previous proof.

Case 2. For \(k\in {\mathbb {N}}\) such that (2.42) holds, let

$$\begin{aligned} {\tilde{K}}_{k}:=\{x\in \Omega _2 \,: \, w_{k}(x)>\inf \text{ Ran }(\beta ) +\frac{\delta '}{4\nu (\Omega _2)}\} \end{aligned}$$

and proceed similarly. \(\square \)

Remark 2.8

  1. (i)

    Taking limits in (2.25) we obtain that, if [uv] is a solution of \((GP_\varphi ^{\textbf{a}_p,\gamma ,\beta })\), then

    $$\begin{aligned} \displaystyle{} & {} \frac{c_p}{2}\left( \int _{\Omega }\int _{\Omega }|u(y)-u(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^{\frac{1}{p'}} \displaystyle \\{} & {} \le \Lambda _1\Vert \varphi \Vert _{L^{p'}(\Omega ,\nu )}+ \frac{\Lambda _1+\Lambda _2}{\nu (\Omega )^{\frac{1}{p}} }\Vert \varphi \Vert _{L^{1}(\Omega ,\nu )}, \end{aligned}$$

    where \(c_p\) is the constant in (2.5), and \(\Lambda _1\) and \(\Lambda _2\) come from the generalised Poincaré-type inequality and depend only on p, \(\Omega _1\) and \(\Omega _2\).

  2. (ii)

    Observe that, on account of (2.4) and the above estimate, we have

    $$\begin{aligned}{} & {} \left( \int _{\Omega }\left| \int _{\Omega }\textbf{a}_p(x,y,u(y)-u(x))dm_x(y)\right| ^{p'}{\text {d}}\nu (x)\right) ^{\frac{1}{p'}} \\{} & {} \le C_p\nu (\Omega )+ \frac{2C_p}{c_p}(2\Lambda _1+\Lambda _2)\Vert \varphi \Vert _{L^{p'}(\Omega ,\nu )}. \end{aligned}$$

    Therefore, since [uv] is a solution of \((GP_\varphi ^{\textbf{a}_p,\gamma ,\beta })\),

    $$\begin{aligned} \begin{array}{l} \displaystyle \Vert v\Vert _{L^{p'}(\Omega ,\nu )} \le C_p\nu (\Omega )+\left( \frac{2C_p}{c_p}(2\Lambda _1+\Lambda _2)+1\right) \Vert \varphi \Vert _{L^{p'}(\Omega ,\nu )}. \end{array} \end{aligned}$$
  3. (iii)

    When \(\varphi =0\) in \(\Omega _2\), we can easily get that \(v\ll \varphi \) in \(\Omega _1.\)

2.3 Other boundary conditions

We can now ask for existence and uniqueness of solutions of the following problem (which is introduced in Sect. 2.1)

$$\begin{aligned} \left\{ \begin{array}{ll} \gamma \big (u(x))-\hbox {div}_m\textbf{a}_p u(x) \ni \varphi (x), &{} x\in W, \\ \mathcal {N}^{\textbf{a}_p}_2 u(x)+\beta \big (u(x)\big )\ni \varphi (x), &{} x\in \partial _mW, \end{array} \right. \end{aligned}$$
(2.44)

or, of the more general problem,

$$\begin{aligned} \left\{ \begin{array}{ll}\displaystyle \gamma \big (u(x))- \int _{W\cup \Omega _2} \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y) \ni \varphi (x), &{} x\in \Omega _1=W, \\ \displaystyle \mathcal {N}^{\textbf{a}_p}_2 u(x)+\beta \big (u(x)\big )\ni \varphi (x), &{} x\in \Omega _2\subseteq \partial _mW. \end{array} \right. \end{aligned}$$

Recall that \(\mathcal {N}^{\textbf{a}_p}_2\) is defined as follows:

$$\begin{aligned} \mathcal {N}^{\textbf{a}_p}_2 u(x):= -\int _{W} \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y), \ \ x\in \partial _mW, \end{aligned}$$

which involves integration with respect to \(\nu \) only over W, or more specifically over \(\partial _m(X\setminus W)\).

For Problem (2.44), we know that, in general, we do not have an appropriate Poincaré-type inequality to work with (see Remark A.5). Therefore, other techniques must be used to obtain the existence of solutions. In the particular case of \(\gamma (r)=\beta (r)=r\), this was done in [43] by exploiting further monotonicity techniques.

However, if a generalised Poincaré-type inequality (as defined in Definition A.1) is satisfied on \((A,B)=(\Omega _1, \Omega _2)\), we could solve the above problem by using the same techniques that we have used to solve Problem (2.7). Indeed, we can work analogously but with the integration by parts formula given in Remark 2.9. Note that this kind of Poincaré-type inequality holds, for example, for finite graphs; even if \(\Omega _2=\partial _m W\).

Remark 2.9

Let \(\Omega :=\Omega _1\cup \Omega _2\). The following integration by parts formula holds: Let u be a measurable function such that

$$\begin{aligned}{}[(x,y)\mapsto \textbf{a}_p(x,y,u(y)-u(x))]\in L^{q}( \left( \Omega \times \Omega \right) \setminus \left( \Omega _2\times \Omega _2\right) ,\nu \otimes m_x) \end{aligned}$$

and let \(w \in L^{q'}(\Omega ,\nu )\). Then,

$$\begin{aligned}{} & {} \displaystyle -\int _{\Omega _1}\int _{\Omega } \textbf{a}_p(x,y,u(y)- u (x))dm_x(y)w(x){\text {d}}\nu (x)\\{} & {} \displaystyle \qquad -\int _{\Omega _2}\int _{\Omega _1} \textbf{a}_p(x,y,u(y)- u (x))dm_x(y)w(x){\text {d}}\nu (x) \\ ={} & {} \displaystyle \frac{1}{2} \int _{\left( \Omega \times \Omega \right) \setminus \left( \Omega _2\times \Omega _2\right) } \textbf{a}_p(x,y,u(y)-u(x)) (w(y) - w(x)) {\text {d}}(\nu \otimes m_x)(x,y). \end{aligned}$$

Remark 2.10

It is possible to consider this type of problems but with the random walk and the nonlocal Leray–Lions operator having a different behaviour on each subset \(\Omega _i\), \(i=1, 2\). For example, one could consider a problem, posed in \(\Omega _1\cup \Omega _2\subset \mathbb {R}^N\), such as the following:

$$\begin{aligned} \left\{ \begin{array}{ll}\displaystyle \gamma \big (u(x))- \int _{\Omega _1} \textbf{a}_p^1(x,y,u(y)-u(x)) J_1(x-y){\text {d}}y \\ \displaystyle - \int _{\Omega _2} \textbf{a}_p^3(x,y,u(y)-u(x)) J_3(x-y){\text {d}}x\ni \varphi (x), &{} x\in \Omega _1, \\ \displaystyle \beta \big (u(x)\big )- \int _{\Omega _1} \textbf{a}_p^3(x,y,u(y)-u(x)) J_3(x-y){\text {d}}y\\ \displaystyle -\int _{\Omega _2} \textbf{a}_p^2(x,y,u(y)-u(x)) J_2(x-y){\text {d}}x\ni \varphi (x), &{} x\in \Omega _2, \end{array} \right. \end{aligned}$$

where \(J_i\) are kernels like the one in Example 1.2, and \(\textbf{a}_p^i\) are functions like the one in Subsect. 2.1, \(i=1,2,3\). This could be done by obtaining a Poincaré-type inequality involving \(\frac{1}{\alpha _0}J_0\), where \(J_0\) is the minimum of the previous three kernels and \(\alpha _0=\int _{\mathbb {R}^N}J_0(z)dz\). This idea has been used in [22] to study a homogenization problem.

3 Doubly nonlinear diffusion problems

We study two kinds of nonlocal p-Laplacian-type diffusions problems. In one of them, we cover nonlocal nonlinear diffusion problems with nonlinear dynamical boundary condition; and on the other, we tackle nonlinear boundary conditions. We work under Assumptions 15 used in Subsect. 2.2.

3.1 Nonlinear dynamical boundary conditions

Our aim in this subsection is to study the following diffusion problem:

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle v_t(t,x) - \int _{\Omega } \textbf{a}_p(x,y,u(t,y)-u(t,x)) dm_x(y)=f(t,x), &{}x\in \Omega _1,\ 0<t<T, \\ \displaystyle v(t,x)\in \gamma \big (u(t,x)\big ), &{} x\in \Omega _1,\ 0<t<T, \\ \displaystyle w_t(t,x) - \int _{\Omega } \textbf{a}_p(x,y,u(t,y)-u(t,x)) dm_x(y)=g(t,x), &{}x\in \Omega _2, \ 0<t<T, \\ w(t,x) \in \beta \big (u(t,x)\big ), &{}x\in \Omega _2, \ 0<t<T, \\ v(0,x) = v_0(x), &{}x\in \Omega _1, \\ w(0,x) = w_0(x), &{}x\in \Omega _2, \end{array} \right. \nonumber \\ \end{aligned}$$
(3.1)

of which Problem (1.2) is a particular case and which covers the case of dynamic evolution on the boundary \(\partial _mW\) when \(\beta \ne \mathbb {R}\times \{0\}\). This includes, in particular, for \(\gamma =\mathbb {R}\times \{0\}\), the problem where the dynamic evolution occurs only on the boundary:

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle - \hbox {div}_m\textbf{a}_p u(t,x)=f(t,x), &{}x\in W,\ 0<t<T, \\ \displaystyle w_t(t,x)+\mathcal {N}^{\textbf{a}_p}_\textbf{1} u(t,x)=g(t,x), &{}x\in \partial _m W, \ 0<t<T, \\ w(t,x) \in \beta \big (u(t,x)\big ), &{}x\in \partial _m W, \ 0<t<T, \\ w(0,x) = w_0(x), &{}x\in \partial _m W. \end{array} \right. \end{aligned}$$

See [4] for the reference local model.

Note that we may abbreviate Problem (3.1) by using v instead of (vw) and f instead of (fg) as

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle v_t(t,x) - \int _{\Omega } \textbf{a}_p(x,y,u(t,y)-u(t,x)) dm_x(y)=f(t,x), &{}x\in \Omega ,\ 0<t<T, \\ \displaystyle v(t,x)\in \gamma \big (u(t,x)\big ), &{} x\in \Omega _1,\ 0<t<T, \\ v(t,x) \in \beta \big (u(t,x)\big ), &{}x\in \Omega _2, \ 0<t<T, \\ v(0,x) = v_0(x), &{}x\in \Omega .\end{array} \right. \nonumber \\ \end{aligned}$$
(3.2)

To solve this problem, we use nonlinear semigroup theory. To this end, we introduce a multivalued operator associated with Problem (3.2) that allows us to rewrite it as an abstract Cauchy problem. Observe that this operator is defined on \(L^1(\Omega ,\nu ) \equiv \left( L^1(\Omega _1,\nu )\times L^1(\Omega _2,\nu )\right) .\)

Definition 3.1

We say that \( (v, {\hat{v}}) \in \mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p}\) if \(v,{\hat{v}} \in L^1(\Omega ,\nu )\), and there exists \( u\in L^p(\Omega ,\nu )\) with

$$\begin{aligned} u\in \hbox {Dom}(\gamma ) \hbox { and } v\in \gamma (u) \quad \nu \hbox {-a.e. in } \Omega _1, \end{aligned}$$

and

$$\begin{aligned} u\in \hbox {Dom}(\beta ) \hbox { and } v\in \beta (u) \quad \nu \hbox {-a.e. in } \Omega _2, \end{aligned}$$

such that

$$\begin{aligned} (x,y)\mapsto a_p(x,y,u(y)-u(x))\in L^{p'}(\Omega \times \Omega ,\nu \otimes m_x) \end{aligned}$$

and

$$\begin{aligned} - \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y) = {\hat{v}} \quad \hbox {in} \ \ \Omega ; \end{aligned}$$

that is, [uv] is a solution of \((GP_{v+{{\hat{v}}}})\) (see (2.8) and Definition 2.4).

On account of the results given in Subsect. 2.2 (Theorems 2.6 and 2.7), we have the following result. Recall that an operator A in \(L^1(\Omega ,\nu )\) is T-accretive if

$$\begin{aligned} \Vert (u - {\hat{u}})^+ \Vert _{L^1(\Omega ,\nu )} \le \Vert (u - {\hat{u}} + \lambda (v - {\hat{v}}))^+ \Vert _{L^1(\Omega ,\nu )} \quad \hbox {for every } (u, v), ({\hat{u}}, {\hat{v}}) \in A \ \ \hbox {and} \ \lambda > 0. \end{aligned}$$

In fact, A is T-accretive if, and only if, its resolvents are contractions and order-preserving (see, for example, [8, Appendix] for further details).

Theorem 3.2

The operator \(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p}\) is T-accretive in \(L^1(\Omega ,\nu )\) and satisfies the range condition

With respect to the domain of such operator, we can prove the following result.

Theorem 3.3

It holds that

$$\begin{aligned}{} & {} \overline{D(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega ,\nu )}\\{} & {} =\left\{ v\in L^{p'}(\Omega ,\nu ) \,: \, \varGamma ^-\le v\le \varGamma ^+\ \nu \hbox {-a.e. in }\Omega _1, \, \mathfrak {B}^-\le v\le \mathfrak {B}^+\ \nu \hbox {-a.e. in }\Omega _2 \right\} . \end{aligned}$$

Therefore, we also have that

$$\begin{aligned}{} & {} \overline{D(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{1}(\Omega ,\nu )}\\{} & {} =\left\{ v\in L^{1}(\Omega ,\nu ) \,: \, \varGamma ^-\le v\le \varGamma ^+\ \nu \hbox {-a.e. in }\Omega _1, \, \mathfrak {B}^-\le v\le \mathfrak {B}^+\ \nu \hbox {-a.e. in }\Omega _2 \right\} . \end{aligned}$$

Proof

It is obvious that

$$\begin{aligned}{} & {} \overline{D(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega ,\nu )}\\{} & {} \subset \left\{ v\in L^{p'}(\Omega ,\nu ) \,: \, \varGamma ^-\le v\le \varGamma ^+\ \nu \hbox {-a.e. in }\Omega _1, \, \mathfrak {B}^-\le v\le \mathfrak {B}^+\ \nu \hbox {-a.e. in }\Omega _2 \right\} . \end{aligned}$$

For the other inclusion, it is enough to see that

$$\begin{aligned}{} & {} \left\{ v\in L^{\infty }(\Omega ,\nu ) \,: \, \varGamma ^-\le v\le \varGamma ^+\ \nu \hbox {-a.e. in }\Omega _1, \, \mathfrak {B}^-\le v\le \mathfrak {B}^+\ \nu \hbox {-a.e. in }\Omega _2 \right\} \\ {}{} & {} \subset \overline{D(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega ,\nu )}. \end{aligned}$$

Suppose first that \(\gamma \) and \(\beta \) satisfy

$$\begin{aligned} \begin{array}{ll}\varGamma ^-<0, &{}\varGamma ^+>0, \\ \mathfrak {B}^-=0, &{} \mathfrak {B}^+>0. \end{array} \end{aligned}$$

It is enough to see that for any \(v\in L^\infty (\Omega ,\nu )\) such that there exist \(m_1<0\), \(\widetilde{m_i}\in {\mathbb {R}}\), \(\widetilde{M_i}\in {\mathbb {R}}\), \(M_i>0\), \(i=1,2\), satisfying

$$\begin{aligned}{} & {} \varGamma ^-<m_1< \widetilde{m_1}\le v\le \widetilde{M_1}< M_1< \varGamma ^+\ \ \nu \hbox {-a.e. in }\Omega _1, \\{} & {} 0<\widetilde{m_2}\le v\le \widetilde{M_2}< M_2<\mathfrak {B}^+\ \ \nu \hbox {-a.e. in }\Omega _2, \end{aligned}$$

it holds that \(v\in \overline{D(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega ,\nu )}\).

By the results in Subsect. 2.2.4, we know that for \(n\in {\mathbb {N}}\), there exists \(u_n\in L^p(\Omega ,\nu )\) and \(v_n\in L^{p'}(\Omega ,\nu )\) such that \([u_n,v_n]\) is a solution of \(\displaystyle \left( GP_v^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \), i.e., \(v_n\in \gamma (u_n)\) \(\nu \)-a.e. in \(\Omega _1\), \(v_n\in \beta (u_n)\) \(\nu \)-a.e. in \(\Omega _2\) and

$$\begin{aligned} v_n(x) -\frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,u_n(y)-u_n(x)) dm_x(y) = v(x) \quad \hbox { for } \nu \hbox {-a.e. } x\in \Omega . \end{aligned}$$

In other words, \((v_n,n(v-v_n))\in \mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p}\) or, equivalently,

$$\begin{aligned} v_n:=\left( I+\frac{1}{n}\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p}\right) ^{-1}(v) \in D(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p}). \end{aligned}$$

Let us see that \(v_n{\mathop {\longrightarrow }\limits ^{n}} v\) in \(L^{p'}(\Omega ,\nu )\).

Let \(a_{m_1}\le 0\) and \(a_{M_1}\ge 0\) such that

$$\begin{aligned} m_1\in \gamma (a_{m_1})\,\hbox { and } \, M_1\in \gamma (a_{M_1}), \end{aligned}$$

and let \(b_{M_2}\ge 0\) such that

$$\begin{aligned} M_2\in \beta (b_{M_2}). \end{aligned}$$

Set

$$\begin{aligned} \widehat{v}(x):= & {} \left\{ \begin{array}{cc} M_1,\, &{} x\in \Omega _1, \\ M_2,\, &{} x\in \Omega _2, \end{array}\right. \\ \widehat{u}(x):= & {} \left\{ \begin{array}{cc} a_{M_1},\, &{} x\in \Omega _1, \\ b_{M_2},\, &{} x\in \Omega _2, \end{array}\right. \end{aligned}$$

and

$$\begin{aligned} \widehat{\varphi }_n(x):=\left\{ \begin{array}{cc} \displaystyle M_1-\frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,\widehat{u}(y)-\widehat{u}(x))dm_x(y),\, &{} x\in \Omega _1, \\ \displaystyle M_2-\frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,\widehat{u}(y)-\widehat{u}(x))dm_x(y),\, &{} x\in \Omega _2. \end{array}\right. \end{aligned}$$

Then, \([\widehat{u},\widehat{v}]\) is a solution of \(\displaystyle \left( GP_{\widehat{\varphi }_n}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \).

Similarly, for

$$\begin{aligned} \widetilde{v}(x):= & {} \left\{ \begin{array}{cc} m_1, \, &{} x\in \Omega _1, \\ 0, \, &{} x\in \Omega _2, \end{array}\right. \\ \widetilde{u}(x):= & {} \left\{ \begin{array}{cc} a_{m_1},\, &{} x\in \Omega _1, \\ 0,\, &{} x\in \Omega _2, \end{array}\right. \end{aligned}$$

and

$$\begin{aligned} \widetilde{\varphi }_n(x):=\left\{ \begin{array}{cc} \displaystyle m_1-\frac{1}{n}\int _{\Omega _2} \textbf{a}_p(x,y,-a_{m_1})dm_x(y), &{} x\in \Omega _1,\, \\ \displaystyle \frac{1}{n}\int _{\Omega _1} \textbf{a}_p(x,y,- a_{m_1})dm_x(y), &{} x\in \Omega _2, \end{array}\right. \end{aligned}$$

we have that \([\widetilde{u},\widetilde{v}]\) is a solution of \(\displaystyle \left( GP_{\widetilde{\varphi }_n}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \).

Now, recalling (2.4), we have that there exists \(n_0\in {\mathbb {N}}\) such that

and

for \(n\ge n_0\). Consequently, by the maximum principle (Theorem 2.6), we obtain that

$$\begin{aligned} \widetilde{u}\le u_n\le \widehat{u}, \end{aligned}$$

thus

$$\begin{aligned} \left\{ \Vert u_n\Vert _{L^{\infty }(\Omega ,\nu )}\right\} _{n} \hbox { is bounded.} \end{aligned}$$

Finally, since

$$\begin{aligned} v_n(x) - v(x) = \frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,u_n(y)-u_n(x)) dm_x(y)\quad \nu \hbox {-a.e. in} \ \Omega , \end{aligned}$$

we conclude that, on account of (2.4),

$$\begin{aligned} v_n{\mathop {\longrightarrow }\limits ^{n}} v \hbox { in } L^{p'}(\Omega ,\nu ). \end{aligned}$$

The other cases follow similarly, we see two of them. Note that, since \(\mathcal {R}_{\gamma ,\beta }^-<\mathcal {R}_{\gamma ,\beta }^+\), it is not possible to have \(\gamma =\mathbb {R}\times \{0\}\) and \(\beta =\mathbb {R}\times \{0\}\) simultaneously. For example, suppose that we have

$$\begin{aligned} \begin{array}{ll}\varGamma ^-=0,\, &{}\varGamma ^+>0,\\ \mathfrak {B}^-=0,\, &{} \mathfrak {B}^+>0. \end{array} \end{aligned}$$

We use the same notation. Let \(v\in L^\infty (\Omega ,\nu )\) such that there exist \(\widetilde{m_i}\in {\mathbb {R}}\), \(\widetilde{M_i}\in {\mathbb {R}}\), \(M_i>0\), \(i=1,2\), satisfying

$$\begin{aligned} \begin{array}{l}0< \widetilde{m_1}\le v\le \widetilde{M_1}< M_1< \varGamma ^+\ \nu \hbox {-a.e. in }\Omega _1, \\ 0<\widetilde{m_2}\le v\le \widetilde{M_2}< M_2<\mathfrak {B}^+\ \nu \hbox {-a.e. in }\Omega _2. \end{array} \end{aligned}$$

As before, the results in Subsect. 2.2.4 ensure that there exist \(u_n\in L^p(\Omega ,\nu )\) and \(v_n\in L^{p'}(\Omega ,\nu )\), \(n\in {\mathbb {N}}\), such that \([u_n,v_n]\) is a solution of \(\displaystyle \left( GP_v^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \). Let \(\ a_{M_1}\ge 0\) and \(b_{M_2}\ge 0\) such that

$$\begin{aligned} M_1\in \gamma (a_{M_1})\, \hbox {and} \, M_2\in \beta (b_{M_2}). \end{aligned}$$

Now again, let

$$\begin{aligned} \widehat{v}(x):= & {} \left\{ \begin{array}{cc} M_1,\, &{} x\in \Omega _1, \\ M_2,\, &{} x\in \Omega _2, \end{array}\right. \\ \widehat{u}(x):= & {} \left\{ \begin{array}{cc} a_{M_1},\, &{} x\in \Omega _1, \\ b_{M_2},\, &{} x\in \Omega _2, \end{array}\right. \end{aligned}$$

and

$$\begin{aligned} \widehat{\varphi }_n(x):=\left\{ \begin{array}{cc} \displaystyle M_1-\frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,\widehat{u}(y)-\widehat{u}(x))dm_x(y),\, &{} x\in \Omega _1, \\ \displaystyle M_2-\frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,\widehat{u}(y)-\widehat{u}(x))dm_x(y),\, &{} x\in \Omega _2. \end{array}\right. \end{aligned}$$

Then, as before, \([\widehat{u},\widehat{v}]\) is a solution of \(\displaystyle \left( GP_{\widehat{\varphi }_n}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \).

Now, taking \(\widetilde{v}\), \(\widetilde{u}\) and \({\widetilde{\varphi }}\) all equal to the null function in \(\Omega \) and recalling that \(\textbf{a}_p(x,y,0)=0\) for every \(x, y\in X\), we obviously have that \([\widetilde{u},\widetilde{v}]\) is a solution of \(\displaystyle \left( GP_{0}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \). Consequently, again by the second part of the maximum principle, we obtain, as desired, that \(0\le u_n\le \widehat{v}\) for n large enough.

Finally, as a further example of a case which does not follow exactly with the same argument, suppose that \(\gamma :={\mathbb {R}}\times \{0\}\) and, for example,

$$\begin{aligned} \mathfrak {B}^-=0, \ \mathfrak {B}^+>0. \end{aligned}$$

In this case, we take \(0\not \equiv v\in L^\infty (\Omega ,\nu )\) such that \(v=0\) in \(\Omega _1\) and

$$\begin{aligned} 0\le v <M_2 \ \nu \hbox {-a.e. in } \Omega _2\, {\text {for some constant}}\, M_2>0. \end{aligned}$$

As in the previous cases, there exist \(u_n\in L^p(\Omega ,\nu )\) and \(v_n\in L^{p'}(\Omega ,\nu )\), \(n\in {\mathbb {N}}\), such that \([u_n,v_n]\) is a solution of \(\displaystyle \left( GP_v^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \). Let \(b_{M_2}\ge 0 \) such that \(M_2\in \beta (b_{M_2})\),

$$\begin{aligned} \widehat{v}(x):= & {} \left\{ \begin{array}{cc} 0, &{} x\in \Omega _1, \\ M_2,\, &{} x\in \Omega _2, \end{array}\right. \\ \widehat{u}(x):= & {} b_{M_2}, \ x\in \Omega , \end{aligned}$$

and

$$\begin{aligned} \varphi _n(x):=\left\{ \begin{array}{lc} \displaystyle 0, &{} x\in \Omega _1, \\ \displaystyle M_2,\, &{} x\in \Omega _2. \end{array}\right. \end{aligned}$$

Then, \([\widehat{u},\widehat{v}]\) is a solution of \(\displaystyle \left( GP_{\varphi _n}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \). Finally, take \(\widetilde{v}\) and \(\widetilde{u}\) again equal to the null function in \(\Omega \) so that \([\widetilde{u},\widetilde{v}]\) is a solution of \(\displaystyle \left( GP_{0}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \). Consequently, for n large enough, we get that \(0\le u_n\le \widehat{v}\). \(\square \)

In the next result, we state the existence and uniqueness of solutions of Problem (3.2).

Theorem 3.4

Let \(T>0\). For any \(v_0\in L^{1}(\Omega ,\nu )\) and \(f\in L^1(0,T;L^{1}(\Omega ,\nu ))\) such that

$$\begin{aligned} \varGamma ^-\le v_0\le \varGamma ^+\quad \nu \hbox {-a.e. \,in}\, \Omega _1,\\ \mathfrak {B}^-\le v_0\le \mathfrak {B}^+\quad \nu \hbox {-a.e. \,in}\, \Omega _2, \end{aligned}$$

and

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\int _{\Omega }v_0{\text {d}}\nu +\int _0^t\int _\Omega fd\nu {\text {d}}s <\mathcal {R}_{\gamma ,\beta }^+ \quad \hbox {for every } 0 \le t\le T, \end{aligned}$$
(3.3)

there exists a unique mild-solution \(v\in C([0,T];L^1(\Omega ,\nu ))\) of Problem (3.2).

Let v and \({\widetilde{v}}\) be the mild solutions of Problem (3.2) with respective data \(v_0,\ {\widetilde{v}}_0\in L^{1}(\Omega ,\nu )\) and \(f,\ {\widetilde{f}}\in L^1(0,T;L^{1}(\Omega ,\nu ))\). Then,

$$\begin{aligned}{} & {} \displaystyle \int _\Omega \left( v(t,x)-{\widetilde{v}}(t,x)\right) ^+{\text {d}}\nu (x) \displaystyle \le \int _\Omega \left( v_0(x)-{\widetilde{v}}_0(x)\right) ^+{\text {d}}\nu (x)\\{} & {} \displaystyle +\int _0^t\int _{\Omega }\left( f(s,x)-{\widetilde{f}}(s,x)\right) ^+{\text {d}}\nu (x){\text {d}}s \quad \hbox {for every } 0\le t\le T. \end{aligned}$$

If, in addition to the previous assumptions on the data, we impose that

$$\begin{aligned} v_0\in L^{p'}(\Omega ,\nu ), \ f\in L^{p'}(0,T;L^{p'}(\Omega ,\nu )) \ \hbox { and } \ \int _{\Omega _1}j_\gamma ^*(v_0){\text {d}}\nu +\int _{\Omega _2}j_\beta ^*(v_0){\text {d}}\nu <+\infty , \end{aligned}$$
(3.4)

then the mild solution v belongs to \(W^{1,1}(0,T;L^{1}(\Omega ,\nu ))\) and satisfies

$$\begin{aligned} \left\{ \begin{array}{l}\partial _tv(t)+\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p}v(t)\ni f(t)\quad \hbox {for a.e. }t\in (0,T),\\ v(0)=v_0,\end{array}\right. \end{aligned}$$

that is, v is a strong solution.

Proof

We start by proving the existence of mild solutions. For \(n\in {\mathbb {N}}\), consider the partition

$$\begin{aligned} t_0^n=0<t_1^n<\cdots<t_{n-1}^n<t_n^n=T \end{aligned}$$

where \(t_i^n:=iT/n\), \(i=1,\ldots ,n\). Given \(\epsilon >0\), there exists \(n\in \mathbb {N}\), \(f_i^n\in L^{p'}(\Omega ,\nu )\), \(i=1,\ldots n\), and \(v_0^n \in \overline{D(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega ,\nu )}\) (i.e., \(v_0^n\in L^{p'}(\Omega ,\nu )\) satisfying \(\varGamma ^-\le v_0^n\le \varGamma ^+\) \(\nu \)-a.e. in \(\Omega _1\), and \(\mathfrak {B}^-\le v_0^n\le \mathfrak {B}^+\) \(\nu \)-a.e. in \(\Omega _2\)) such that \(T/n\le \epsilon \),

$$\begin{aligned} \sum _{i=1}^n\int _{t_{i-1}^n}^{t_i^n}\Vert f(t)- f_i^n\Vert _{L^{1}(\Omega ,\nu )}{\text {d}}t \le \epsilon \end{aligned}$$
(3.5)

and

$$\begin{aligned} \Vert v_0-v_0^n\Vert _{L^1(\Omega ,\nu )} \le \epsilon . \end{aligned}$$
(3.6)

Then, setting

$$\begin{aligned} f_n(t):=f_i^n \ \hbox { for } t\in ]t_{i-1}^n,t_i^n], i=1,\ldots , n, \end{aligned}$$

we have that

$$\begin{aligned} \int _0^T \Vert f(t)- f_n(t) \Vert _{L^1(\Omega ,\nu )}{\text {d}}t \le \epsilon . \end{aligned}$$

By the results in Subsect. 2.2.4 we see that, for n large enough, we may recursively find a solution \([u_i^n,v_i^n]\) of \(\displaystyle \left( GP^{\frac{T}{n}\textbf{a}_p,\gamma ,\beta }_{\frac{T}{n} f_i^n+v_{i-1}^n}\right) \), \(i=1,\ldots ,n\), in other words,

$$\begin{aligned} v_i^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)=\frac{T}{n} f_i^n(x)+v_{i-1}^n(x), \ \, x\in \Omega , \end{aligned}$$

or, equivalently,

$$\begin{aligned} \frac{v_i^n(x)-v_{i-1}^n(x)}{T/n}-\int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)= f_i^n(x), \ \, x\in \Omega , \end{aligned}$$
(3.7)

with \(v_i^n(x)\in \gamma (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _1\) and \(v_i^n(x)\in \beta (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _2\), \(i=1,\ldots ,n\). That is, we may find the unique solution \(v_i^n\) of the time discretization scheme associated with (3.2):

$$\begin{aligned} v_i^n+\frac{T}{n} \mathcal {B}_{\textbf{a}_p}^{m,\gamma ,\beta }(v_i^n)\ni \frac{T}{n} f_i^n + v_{i-1}^n \ \hbox { for} ~i=1,\ldots ,n. \end{aligned}$$

However, to apply the results in Subsect. 2.2.4, we must ensure that

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\int _\Omega \left( \frac{T}{n} f_i^n + v_{i-1}^n\right) {\text {d}}\nu <\mathcal {R}_{\gamma ,\beta }^+ \end{aligned}$$
(3.8)

holds for each step. For the first step we need that

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\int _{\Omega }v_0^n {\text {d}}\nu +\frac{T}{n}\int _\Omega f_1^n {\text {d}}\nu <\mathcal {R}_{\gamma ,\beta }^+ \end{aligned}$$

holds so that condition (3.8) is satisfied. Integrating (3.7) with respect to \(\nu \) over \(\Omega \), we get

$$\begin{aligned} \int _\Omega v_1^n {\text {d}}\nu =\int _{\Omega }v_0^n {\text {d}}\nu +\frac{T}{n}\int _\Omega f_1^n {\text {d}}\nu \end{aligned}$$

thus

$$\begin{aligned} \frac{T}{n} \int _\Omega f_2^n {\text {d}}\nu +\int _\Omega v_{1}^n {\text {d}}\nu =\frac{T}{n} \sum _{j=1}^2 \int _\Omega f_j^n {\text {d}}\nu +\int _\Omega v_{0}^n {\text {d}}\nu , \end{aligned}$$

so that, for the second step, we need

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-<\frac{T}{n} \sum _{j=1}^2 \int _\Omega f_j^n {\text {d}}\nu +\int _\Omega v_{0}^n {\text {d}}\nu <\mathcal {R}_{\gamma ,\beta }^+. \end{aligned}$$

Therefore, we recursively obtain that for each n and each step \(i=1,\ldots , n\), the following must be satisfied:

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-< \frac{T}{n} \sum _{j=1}^i \int _\Omega f_j^n {\text {d}}\nu +\int _\Omega v_{0}^{n} {\text {d}}\nu <\mathcal {R}_{\gamma ,\beta }^+. \end{aligned}$$

However, taking n large enough, this holds thanks to (3.3), (3.5) and (3.6).

Therefore,

$$\begin{aligned} v_n (t):=\left\{ \begin{array}{ll} v_0^n, &{} \hbox {if } t=0,\\ v_i^n, &{} \hbox {if } t\in ]t_{i-1}^n,t_i^n], i=1,\ldots ,n, \end{array} \right. \end{aligned}$$

is an \(\epsilon \)-approximate solution of Problem (3.2) as defined in nonlinear semigroup theory. Consequently, by nonlinear semigroup theory (see [11, 10, Theorem 4.1], or [8, Theorem A.27]) and on account of Theorem 3.2 and Theorem 3.3 we have that Problem (3.2) has a unique mild solution \(v\in C([0,T];L^1(\Omega ,\nu ))\) with

$$\begin{aligned} v_n(t){\mathop {\longrightarrow }\limits ^{n}} v(t) \ \hbox { in } L^1(\Omega ,\nu )\hbox { uniformly for } t\in [0,T]. \end{aligned}$$
(3.9)

Uniqueness and the maximum principle for mild solutions are guaranteed by the T-accretivity of the operator.

Let us now see that v is a strong solution of Problem (3.2) when (3.4) holds. Note that, since \(v_0\in L^{p'}(\Omega ,\nu )\), we may take \(v_0^n=v_0\) for every \(n\in {\mathbb {N}}\) in the previous computations and \(f_i^n\in L^{p'}(\Omega ,\nu )\), \(i=1,\ldots n\), additionally satisfying

$$\begin{aligned} \sum _{i=1}^n\int _{t_{i-1}^n}^{t_i^n}\Vert f(t)- f_i^n\Vert ^{p'}_{L^{p'}(\Omega ,\nu )}{\text {d}}t \le \epsilon . \end{aligned}$$

Let us define

$$\begin{aligned} u_n (t)= u_i^n \ \hbox { for } t\in ]t_{i-1}^n,t_i^n], \, i=1,\ldots ,n. \end{aligned}$$

Multiplying equation (3.7) by \(u_i^n\) and integrating over \(\Omega \) with respect to \(\nu \), we obtain

$$\begin{aligned}{} & {} \displaystyle \int _\Omega \frac{v_i^n(x)-v_{i-1}^n(x)}{T/n}u_i^n(x){\text {d}}\nu (x)\nonumber \\{} & {} \qquad -\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)u_i^n(x){\text {d}}\nu (x) \nonumber \\{} & {} \quad \displaystyle = \int _\Omega f_i^n(x)u_i^n(x) {\text {d}}\nu (x). \end{aligned}$$
(3.10)

Now, since \(v_i^n(x)\in \gamma (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _1\) and \(v_i^n(x)\in \beta (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _2\),

$$\begin{aligned} \left\{ \begin{array}{cc} u_i^n(x)\in \gamma ^{-1}(v_i^n(x))=\partial j_\gamma ^*(v_i^n(x))\, &{} \hbox {for } \nu \hbox {-a.e. } x\in \Omega _1, \\ u_i^n(x)\in \beta ^{-1}(v_i^n(x))=\partial j_\beta ^*(v_i^n(x))\, &{} \hbox {for } \nu \hbox {-a.e. } x\in \Omega _2. \end{array}\right. \end{aligned}$$

Consequently,

$$\begin{aligned} \left\{ \begin{array}{cc} j_\gamma ^*(v_{i-1}^n(x))- j_\gamma ^*(v_{i}^n(x))\ge (v_{i-1}^n(x)-v_i^n(x))u_i^n(x)\, &{} \hbox {for } \nu \hbox {-a.e. } x\in \Omega _1, \\ j_\beta ^*(v_{i-1}^n(x))- j_\beta ^*(v_{i}^n(x))\ge (v_{i-1}^n(x)-v_i^n(x))u_i^n(x)\, &{} \hbox {for } \nu \hbox {-a.e. } x\in \Omega _2. \end{array}\right. \end{aligned}$$

Therefore, from (3.10), it follows that

$$\begin{aligned}{} & {} \displaystyle \frac{1}{T/n}\int _{\Omega _1}(j_\gamma ^*(v_{i}^n(x))- j_\gamma ^*(v_{i-1}^n(x))){\text {d}}\nu (x)+ \frac{1}{T/n}\int _{\Omega _2}(j_\beta ^*(v_{i}^n(x))- j_\beta ^*(v_{i-1}^n(x))){\text {d}}\nu (x) \\ {}{} & {} \displaystyle \qquad -\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))u_i^n(x)dm_x(y){\text {d}}\nu (x)\\{} & {} \displaystyle \le \int _\Omega f_i^n(x)u_i^n(x) {\text {d}}\nu (x), \end{aligned}$$

\(i=1,\ldots ,n\). Then, integrating this equation over \(]t_{i-1}^n,t_i^n]\) and adding for \(1\le i \le n\), we get

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1}(j_\gamma ^*(v_{n}^n(x))- j_\gamma ^*(v_{0}(x))){\text {d}}\nu (x)+ \int _{\Omega _2}(j_\beta ^*(v_{n}^n(x))- j_\beta ^*(v_{0}(x))){\text {d}}\nu (x) \\{} & {} \displaystyle \qquad -\sum _{i=1}^n\int _{t_{i-1}^n}^{t_i^n} \int _\Omega \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)u_i^n(x){\text {d}}\nu (x){\text {d}}t\\{} & {} \displaystyle \le \sum _{i=1}^n\int _{t_{i-1}^n}^{t_i^n}\int _\Omega f_i^n(x)u_i^n(x) {\text {d}}\nu (x){\text {d}}t, \end{aligned}$$

which, recalling the definitions of \(f_n\), \(u_n\) and \(v_n\), and integrating by parts, can be rewritten as:

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1}(j_\gamma ^*(v_{n}^n(x))- j_\gamma ^*(v_{0}(x))){\text {d}}\nu (x)+ \int _{\Omega _2}(j_\beta ^*(v_{n}^n(x))- j_\beta ^*(v_{0}(x))){\text {d}}\nu (x) \nonumber \\{} & {} \displaystyle \qquad +\frac{1}{2}\int _0^T\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_n(t)(y)-u_n(t)(x))(u_n(t)(y)-u_n(t)(x))dm_x(y){\text {d}}\nu (x){\text {d}}t \nonumber \\{} & {} \displaystyle \le \int _0^T\int _\Omega f_n(t)(x)u_n(t)(x) {\text {d}}\nu (x){\text {d}}t. \end{aligned}$$
(3.11)

This, together with (2.5) and the fact that \(j^*_\gamma \) and \(j^*_\beta \) are nonnegative, yields

$$\begin{aligned} \begin{array}{l} \displaystyle \frac{c_p}{2}\int _0^T\int _\Omega \int _\Omega |u_n(t)(y)-u_n(t)(x)|^p dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \frac{1}{2}\int _0^T\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_n(t)(y)-u_n(t)(x))(u_n(t)(y)-u_n(t)(x))dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \int _{\Omega _1}( j_\gamma ^*(v_{0}(x))){\text {d}}\nu (x)+ \int _{\Omega _2}( j_\beta ^*(v_{0}(x))){\text {d}}\nu (x) + \int _0^T\int _\Omega f_n(t)(x)u_n(t)(x) {\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \int _{\Omega _1}( j_\gamma ^*(v_{0}(x))){\text {d}}\nu (x)+ \int _{\Omega _2}( j_\beta ^*(v_{0}(x))){\text {d}}\nu (x) + \int _0^T \Vert f_n(t)\Vert _{L^{p'}(\Omega ,\nu )} \Vert u_n(t) \Vert _{L^{p}(\Omega ,\nu )} {\text {d}}t. \end{array} \end{aligned}$$

Therefore, for any \(\delta >0\), by (3.4) and Young’s inequality, there exists \(C(\delta )>0\) such that

$$\begin{aligned}{} & {} \displaystyle \int _0^T\int _\Omega \int _\Omega |u_n(t)(y)-u_n(t)(x)|^p dm_x(y){\text {d}}\nu (x){\text {d}}t\nonumber \\{} & {} \quad \le C(\delta )+ \delta \int _0^T \Vert u_n(t) \Vert _{L^{p}(\Omega ,\nu )}^{p} {\text {d}}t. \end{aligned}$$
(3.12)

Now, by (3.9), if \(\mathcal {R}_{\gamma ,\beta }^+=+\infty \), there exists \(M>0\) and \(n_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \sup _{t\in [0,T]}\int _\Omega v_{n}^+(t)(x){\text {d}}\nu (x)<M \quad \hbox {for every } n\ge n_0, \end{aligned}$$

and, if \(\mathcal {R}_{\gamma ,\beta }^+<+\infty \), there exist \(M\in {\mathbb {R}}\), \(h>0\) and \(n_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \sup _{t\in [0,T]}\int _\Omega v_{n}(t)(x){\text {d}}\nu (x)<M<\mathcal {R}_{\gamma ,\beta }^+, \end{aligned}$$

and

$$\begin{aligned} \sup _{t\in [0,T]} \int _{\{x\in \Omega \,: \, v_{n}(t)(x)<-h\}}|v_{n}(t)(x)|{\text {d}}\nu (x)<\frac{\mathcal {R}_{\gamma ,\beta }^+-M}{8} \quad \hbox {for every } n\ge n_0. \end{aligned}$$

Consequently, Lemma A.7 and Lemma A.8 yield

$$\begin{aligned} \Vert u_{n}^+(t)\Vert _{L^{p}(\Omega ,\nu )}\le C_2\left( \left( \int _\Omega \int _\Omega |u_{n}^+(t)(y)-u_{n}^+(t)(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^\frac{1}{p}+1 \right) \end{aligned}$$

and for some constant \(C_2>0\). Similarly, we may find \(C_3>0\) such that

$$\begin{aligned} \Vert u_{n}^-(t)\Vert _{L^{p}(\Omega ,\nu )}\le C_3\left( \left( \int _\Omega \int _\Omega |u_{n}^-(t)(y)-u_{n}^-(t)(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^\frac{1}{p}+1 \right) . \end{aligned}$$

Consequently, by (3.12), choosing \(\delta \) small enough, we deduce that \(\{u_{n}\}_n\) is bounded in \(L^p(0,T;L^p(\Omega ,\nu ))\). Therefore, there exists a subsequence, which we continue to denote by \(\{u_{n}\}_n\), and \(u\in L^p(0,T;L^p(\Omega ,\nu ))\) such that

$$\begin{aligned} u_{n}{\mathop {\rightharpoonup }\limits ^{n}} u \ \hbox { weakly in } L^p(0,T;L^p(\Omega ,\nu )). \end{aligned}$$

Then, since \(\gamma \) and \(\beta \) are maximal monotone graphs, we conclude that \(v(t)(x)\in \gamma (u(t)(x))\) for \(\mathcal {L}^1\otimes \nu \)-a.e. \((t,x)\in (0,T)\times \Omega _1\) and \(v(t)(x)\in \beta (u(t)(x))\) for \(\mathcal {L}^1\otimes \nu \)-a.e. \((t,x)\in (0,T)\times \Omega _2\).

Note that, since, by (3.12),

$$\begin{aligned} \displaystyle \left\{ \int _0^T\int _\Omega \int _\Omega |u_{n}(t)(y)-u_{n}(t)(x)|^p dm_x(y){\text {d}}\nu (x){\text {d}}t\right\} _n \quad \hbox {is bounded}, \end{aligned}$$

then, by (2.4), \(\{[(t,x,y)\mapsto \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x))]\}_n\) is bounded in \(L^{p'}(0,T; L^{p'}(\Omega \times \Omega ,\nu \otimes m_x))\) so we may take a further subsequence, which we still denote in the same way, such that

$$\begin{aligned}{} & {} [(t,x,y)\mapsto \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x))]{\mathop {\rightharpoonup }\limits ^{n}} \Phi , \\{} & {} \hbox { weakly in } L^{p'}(0,T; L^{p'}(\Omega \times \Omega ,\nu \otimes m_x)). \end{aligned}$$

Note that, for any \(\xi \in L^p(\Omega ,\nu )\), by the integrations by parts formula we know that

$$\begin{aligned} \begin{array}{l} \displaystyle -\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x))\xi (x) dm_x(y){\text {d}}\nu (x)\\ \displaystyle =\frac{1}{2}\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x))(\xi (y)-\xi (x))dm_x(y){\text {d}}\nu (x) \end{array} \end{aligned}$$

for \(t\in [0,T]\), thus taking limits as \(n\rightarrow \infty \) we have

$$\begin{aligned}{} & {} -\int _\Omega \int _\Omega \Phi (t,x,y) \xi (x) dm_x(y){\text {d}}\nu (x)\nonumber \\{} & {} =\frac{1}{2}\int _\Omega \int _\Omega \Phi (t,x,y)(\xi (y)-\xi (x))dm_x(y){\text {d}}\nu (x). \end{aligned}$$
(3.13)

Now, from (3.7), we have that

$$\begin{aligned}{} & {} \frac{v_{n}(t)(x)-v_{n}(t-{T/n})(x)}{T/n}-\int _\Omega \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x))dm_x(y)\nonumber \\{} & {} \qquad = f_{n}(t)(x) \end{aligned}$$
(3.14)

for \(t\in [0,T]\) and \(x\in \Omega \). Let \(\Psi \in W^{1,1}_0(0,T;L^p(\Omega ,\nu ))\), \(\hbox {supp}(\Psi )\subset \subset [0,T]\), then

$$\begin{aligned}{} & {} \displaystyle \int _0^T \frac{v_{n}(t)(x)-v_{n}(t-{T/n})(x)}{T/n}\Psi (t)(x){\text {d}}t\\{} & {} \displaystyle =-\int _0^{T-T/n} v_{n}(t)(x)\frac{\Psi (t+T/n)(x)-\Psi (t)(x)}{T/n}{\text {d}}t+\int _{T-T/n}^T \frac{v_{n}\Psi (t)(x)}{T/n}{\text {d}}t\\{} & {} \quad -\int _0^{T/n} \frac{v_0\Psi (t)(x)}{T/n}{\text {d}}t \end{aligned}$$

for \(x\in \Omega \). Therefore, multiplying (3.14) (for the previously chosen subsequence) by \(\Psi \), integrating over \((0,T)\times \Omega \) with respect to \(\mathcal {L}^1\otimes \nu \) and taking limits, we get

$$\begin{aligned}{} & {} \displaystyle -\int _0^T \int _\Omega v(t)(x)\frac{d}{{\text {d}}t}\Psi (t)(x){\text {d}}\nu (x) {\text {d}}t- \int _0^T\int _\Omega \int _\Omega \Phi (t,x,y)dm_x(y)\Psi (t)(x){\text {d}}\nu (x){\text {d}}t \nonumber \\{} & {} \quad \displaystyle = \int _0^T\int _\Omega f(t)(x)\Psi (t)(x){\text {d}}\nu (x){\text {d}}t. \end{aligned}$$
(3.15)

Therefore, taking \(\Psi (t)(x)=\psi (t)\xi (x)\), where \(\psi \in C_c^\infty (0,T)\) and \(\xi \in L^p(\Omega ,\nu )\), we obtain that

$$\begin{aligned}{} & {} \displaystyle \int _0^T v(t)(x)\psi '(t) {\text {d}}t =-\int _0^T \int _\Omega \Phi (t,x,y)\psi (t)dm_x(y) {\text {d}}t\\{} & {} \qquad \qquad \qquad \qquad \qquad \quad \,\,- \int _0^T f(t)(x)\psi (t) {\text {d}}t, \, \hbox { for} ~\nu \hbox {-a.e.} ~x\in \Omega . \end{aligned}$$

It follows that

$$\begin{aligned} \displaystyle v'(t)(x) = \int _\Omega \Phi (t,x,y)dm_x(y) + f(t)(x) \quad \hbox {for a.e. } t\in (0,T) \hbox { and } \nu \hbox {-a.e. } x\in \Omega . \end{aligned}$$

Therefore, since \(v\in C([0,T];L^1(\Omega ,\nu ))\), \(\Phi \in L^{p'}(0,T; L^{p'}(\Omega \times \Omega ,\nu \otimes m_x))\) and \(f\in L^{p'}(0,T;L^{p'}(\Omega ,\nu ))\), we have \(v'\in L^{p'}(0,T;L^{p'}(\Omega ,\nu ))\) and \(v\in W^{1,1}(0,T; L^{1}(\Omega ,\nu ))\).

Hence, to conclude it remains to prove that

$$\begin{aligned} \int _\Omega \Phi (t,x,y)dm_x(y)=\int _\Omega \textbf{a}_p(x,y,u(t)(y)-u(t)(x))dm_x(y) \end{aligned}$$

for \(\mathcal {L}^1\otimes \nu \)-a.e. \((t,x)\in [0,T]\times \Omega \). To this aim, we make use of the following claim that will be proved later on:

$$\begin{aligned}{} & {} \displaystyle \limsup _n \int _0^T \int _{\Omega }\int _{\Omega } \textbf{a}_p(x,y,u_{n}(t) (y)-u_{n}(t)(x))(u_{n}(t)(y)-u_{n}(t)(x)))dm_x(y){\text {d}}\nu (x){\text {d}}t \nonumber \\{} & {} \quad \displaystyle \le \int _0^T \int _{\Omega }\int _{\Omega } \Phi (t,x,y)(u(t)(y)-u(t)(x)))dm_x(y)d{\text {d}}\nu (x){\text {d}}t. \end{aligned}$$
(3.16)

Now, let \(\rho \in L^p(0,T;L^p(\Omega ,\nu ))\). By (2.3), we have

$$\begin{aligned}{} & {} \displaystyle \int _0^T \int _\Omega \int _\Omega \textbf{a}_p(x,y,\rho (t)(y)-\rho (t)(x)) \\{} & {} \qquad \qquad \qquad \displaystyle \times (u_{n}(t)(y)-\rho (t)(y)-(u_{n}(t)(x)-\rho (t)(x)))dm_x(y){\text {d}}\nu (x){\text {d}}t\\{} & {} \quad \displaystyle \le \int _0^T \int _\Omega \int _\Omega \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x)) \\{} & {} \qquad \qquad \qquad \displaystyle \times (u_{n}(t)(y)-\rho (t)(y)-(u_{n}(t)(x)-\rho (t)(x)))dm_x(y){\text {d}}\nu (x){\text {d}}t. \end{aligned}$$

Thus, taking limits as \(n\rightarrow \infty \) and using (3.16), we obtain

$$\begin{aligned}{} & {} \displaystyle \int _0^T \int _\Omega \int _\Omega \textbf{a}_p(x,y,\rho (t)(y)-\rho (t)(x)) \\{} & {} \qquad \displaystyle \times (u(t)(y)-\rho (t)(y)-(u(t)(x)-\rho (t)(x)))dm_x(y){\text {d}}\nu (x){\text {d}}t \\{} & {} \quad \displaystyle \le \int _0^T \int _\Omega \int _\Omega \Phi (t,x,y)(u(t)(y)-\rho (t)(y)-(u(t)(x)-\rho (t)(x)))dm_x(y){\text {d}}\nu (x){\text {d}}t, \end{aligned}$$

which, integrating by parts and recalling (3.13), becomes

$$\begin{aligned}{} & {} \displaystyle \int _0^T \int _\Omega \int _\Omega \textbf{a}_p(x,y,\rho (t)(y)-\rho (t)(x))dm_x(y)(u(t)(x)-\rho (t)(x)){\text {d}}\nu (x){\text {d}}t \\{} & {} \displaystyle \ge \int _0^T \int _\Omega \int _\Omega \Phi (t,x,y)dm_x(y)(u(t)(x)-\rho (t)(x)){\text {d}}\nu (x){\text {d}}t. \end{aligned}$$

To conclude, take \(\rho =u\pm \lambda \xi \) for \(\lambda >0\) and \(\xi \in L^p(0,T;L^p(\Omega ,\nu ))\) to get

$$\begin{aligned} \begin{array}{l} \displaystyle \int _0^T \int _\Omega \int _\Omega \textbf{a}_p(x,y,(u\pm \lambda \xi )(t)(y)-(u\pm \lambda \xi )(t)(x))dm_x(y)\xi (t)(x){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \ge \int _0^T \int _\Omega \int _\Omega \Phi (t,x,y)dm_x(y)\xi (t)(x){\text {d}}\nu (x){\text {d}}t \end{array} \end{aligned}$$

which, letting \(\lambda \rightarrow 0 \) yields

$$\begin{aligned} \begin{array}{l} \displaystyle \int _0^T \int _\Omega \int _\Omega \textbf{a}_p(x,y,u(t)(y)-u(t)(x))dm_x(y)\xi (t)(x){\text {d}}\nu (x){\text {d}}t\\ \displaystyle =\int _0^T \int _\Omega \int _\Omega \Phi (t,x,y)dm_x(y)\xi (t)(x){\text {d}}\nu (x){\text {d}}t \end{array} \end{aligned}$$

for any \(\xi \in L^p(0,T;L^p(\Omega ,\nu ))\). Therefore,

$$\begin{aligned} \int _\Omega \textbf{a}_p(x,y,u(t)(y)-u(t)(x))dm_x(y)=\int _\Omega \Phi (t,x,y)dm_x(y) \end{aligned}$$

for \(\mathcal {L}^1\otimes \nu \)-a.e. \((t,x)\in [0,T]\times \Omega \).

Let us prove claim (3.16). By (3.11) and Fatou’s lemma, we have

$$\begin{aligned}{} & {} \displaystyle \limsup _n\frac{1}{2}\int _0^T\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x))(u_{n}(t)(y)-u_{n}(t)(x))dm_x(y){\text {d}}\nu (x){\text {d}}t \nonumber \\{} & {} \quad \displaystyle \le -\int _{\Omega _1}(j_\gamma ^*(v(T)(x))- j_\gamma ^*(v(0)(x))){\text {d}}\nu (x) -\int _{\Omega _2}(j_\beta ^*(v(T)(x))- j_\beta ^*(v(0)(x))){\text {d}}\nu (x) \nonumber \\{} & {} \qquad \displaystyle +\int _0^T\int _\Omega f(t)(x)u(t)(x) {\text {d}}\nu (x){\text {d}}t. \end{aligned}$$
(3.17)

Moreover, by (3.15),

$$\begin{aligned} \int _0^T v(t)(x)\frac{d}{{\text {d}}t}\Psi (t)(x){\text {d}}t=\int _0^T F(t)(x)\Psi (t)(x){\text {d}}t, \, \hbox { for } \nu \hbox {-a.e. } x\in \Omega , \end{aligned}$$
(3.18)

where F is given by

$$\begin{aligned} F(t)(x)=-\int _\Omega \Phi (t,x,y)dm_x(y)-f(t)(x), \ x\in \Omega . \end{aligned}$$
(3.19)

Let \(\psi \in C_c^\infty (0,T)\), \(\psi \ge 0\), \(\tau >0\) and

$$\begin{aligned} \eta _\tau (t)(x)=\frac{1}{\tau }\int _t^{t+\tau } u(s)(x) \psi (s){\text {d}}s, \ \ t\in [0,T], \, x\in \Omega . \end{aligned}$$

Then, for \(\tau \) small enough, \(\eta _\tau \in W^{1,1}_0(0,T;L^p(\Omega ,\nu ))\) and we may use it as a test function in (3.18) to obtain

$$\begin{aligned} \begin{array}{rl} \displaystyle \int _0^T F(t)(x)\eta _\tau (t)(x) {\text {d}}t&{}\displaystyle =\int _0^T v(t)(x)\frac{d}{{\text {d}}t}\eta _\tau (t)(x)\\ &{}\displaystyle = \int _0^T v(t)(x)\frac{u(t+\tau )(x)\psi (t+\tau )-u(t)(x)\psi (t)}{\tau } {\text {d}}t \\ &{}\displaystyle =\int _0^T \frac{v(t-\tau )(x)-v(t)(x)}{\tau } u(t)(x)\psi (t) {\text {d}}t. \end{array} \end{aligned}$$

Now,

$$\begin{aligned} \gamma ^{-1}(r)= \partial j_{\gamma ^{-1}}(r)=\partial \left( \int _0^r (\gamma ^{-1})^0(s) {\text {d}}s\right) , \end{aligned}$$

thus, for \(v,{\hat{v}} \in \gamma (u)\),

$$\begin{aligned} ({\hat{v}} -v)u\le \int _{v}^{{\hat{v}}}(\gamma ^{-1})^0(s){\text {d}}s. \end{aligned}$$

A similar fact holds for \(\beta \). Then, for \(\tau >0\) fixed, since \(v(t)(x)\in \gamma (u(t)(x))\) for \(\mathcal {L}^1\otimes \nu \)-a.e. \((t,x)\in (0,T)\times \Omega _1\) and \(v(t)(x)\in \beta (u(t)(x))\) for \(\mathcal {L}^1\otimes \nu \)-a.e. \((t,x)\in (0,T)\times \Omega _2\),

$$\begin{aligned} \begin{array}{l} \displaystyle \int _0^T \int _\Omega F(t)(x)\eta _\tau (t)(x){\text {d}}\nu (x) {\text {d}}t \\ \displaystyle \le \frac{1}{\tau }\int _0^T\int _{\Omega _1} \int _{v(t)(x)}^{v(t-\tau )(x)}(\gamma ^{-1})^0(s){\text {d}}s {\text {d}}\nu (x)\psi (t){\text {d}}t\\ \displaystyle \quad +\frac{1}{\tau }\int _0^T\int _{\Omega _2} \int _{v(t)(x)}^{v(t-\tau )(x)}(\beta ^{-1})^0(s){\text {d}}s {\text {d}}\nu (x)\psi (t){\text {d}}t\\ \displaystyle =\int _0^T\int _{\Omega _1}\int _0^{v(t)(x)}(\gamma ^{-1})^0(s){\text {d}}s{\text {d}}\nu (x) \frac{\psi (t+\tau )-\psi (t)}{\tau }{\text {d}}t\\ \displaystyle \qquad +\int _0^T\int _{\Omega _2}\int _0^{v(t)(x)}(\beta ^{-1})^0(s){\text {d}}s{\text {d}}\nu (x) \frac{\psi (t+\tau )-\psi (t)}{\tau }{\text {d}}t. \end{array} \end{aligned}$$

Letting \(\tau \rightarrow 0^+\) in the above expression, by the dominated convergence theorem,

$$\begin{aligned} \begin{array}{l} \displaystyle \int _0^T \int _\Omega F(t)(x)u(t)(x)\psi (t){\text {d}}\nu (x) {\text {d}}t\displaystyle \le \int _0^T\int _{\Omega _1}\int _0^{v(t)(x)}(\gamma ^{-1})^0(s){\text {d}}s {\text {d}}\nu (x)\psi '(t) {\text {d}}t\\ \displaystyle +\int _0^T\int _{\Omega _2}\int _0^{v(t)(x)}(\beta ^{-1})^0(s){\text {d}}s {\text {d}}\nu (x)\psi '(t) {\text {d}}t\\ \displaystyle =\int _0^T\int _{\Omega _1} j_{\gamma ^{-1}}(v(t)(x)){\text {d}}\nu (x) \psi '(t) {\text {d}}t+\int _0^T\int _{\Omega _2} j_{\beta ^{-1}}(v(t)(x)){\text {d}}\nu (x) \psi '(t) {\text {d}}t\\ \displaystyle = \int _0^T \int _{\Omega _1}j_{\gamma }^*(v(t)(x)){\text {d}}\nu (x) \psi '(t) {\text {d}}t+\int _0^T \int _{\Omega _2}j_{\beta }^*(v(t)(x)){\text {d}}\nu (x) \psi '(t) {\text {d}}t. \end{array} \end{aligned}$$

Taking

$$\begin{aligned} \tilde{\eta }_\tau (t)(x)=\frac{1}{\tau }\int _t^{t+\tau } u(s-\tau )(x) \psi (s){\text {d}}s \end{aligned}$$

yields the opposite inequality so that, in fact,

$$\begin{aligned} \begin{array}{l} \displaystyle \int _0^T\int _\Omega F(t)(x)u(t)(x){\text {d}}\nu (x)\psi (t) {\text {d}}t \displaystyle \\ \displaystyle = \int _0^T \int _{\Omega _1}j_{\gamma }^*(v(t)(x)){\text {d}}\nu (x) \psi '(t) {\text {d}}t+\int _0^T \int _{\Omega _2}j_{\beta }^*(v(t)(x)){\text {d}}\nu (x) \psi '(t) {\text {d}}t. \end{array} \end{aligned}$$

Then,

$$\begin{aligned} -\frac{d}{{\text {d}}t}\left( \int _{\Omega _1}j_{\gamma }^*(v(t)(x)){\text {d}}\nu (x)+\int _{\Omega _2}j_{\beta }^*(v(t)(x)){\text {d}}\nu (x)\right) =\int _\Omega F(t)(x)u(t)(x){\text {d}}\nu (x) \nonumber \\ \end{aligned}$$
(3.20)

in \(\mathcal {D}'(]0,T[)\), thus, in particular,

$$\begin{aligned} \int _{\Omega _1}j_{\gamma }^*(v(t)(x)){\text {d}}\nu (x)+\int _{\Omega _2}j_{\beta }^*(v(t)(x)){\text {d}}\nu (x)\in W^{1,1}(0,T). \end{aligned}$$

Therefore, integrating from 0 to T in (3.20) and recalling (3.19), we get

$$\begin{aligned} \begin{array}{l} \displaystyle \int _0^T\int _\Omega \int _\Omega \Phi (t,x,y)u(t)(x)dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle = -\int _{\Omega _1}(j_\gamma ^*(v(T)(x))- j_\gamma ^*(v(0)(x))){\text {d}}\nu (x)\\ \ \ \displaystyle -\int _{\Omega _2}(j_\beta ^*(v(T)(x))- j_\beta ^*(v(0)(x))){\text {d}}\nu (x)\\ \ \ \displaystyle +\int _0^T\int _\Omega f(t)(x)u(t)(x) {\text {d}}\nu (x){\text {d}}t \end{array} \end{aligned}$$

which, together with (3.17), yields the claim (3.16). \(\square \)

Observe that we have imposed the compatibility condition (3.3) because, for a strong solution,

$$\begin{aligned} \int _\Omega v_0{\text {d}}\nu + \int _0^t\int _\Omega f{\text {d}}\nu {\text {d}}s=\int _\Omega v(t){\text {d}}\nu , \hbox { for } t\in [0,T]. \end{aligned}$$

Example 3.5

Let \(W\subset X\) be a measurable set such that \(W_m\) is m-connected. Given \(f\in L^1(\partial _mW,\nu )\), we say that a function \(u\in L^1(W_m,\nu )\) is an \(\textbf{a}_p\)-lifting of f to \(W_m=W\cup \partial _mW\) if

$$\begin{aligned} \left\{ \begin{array}{ll} -\hbox {div}_m\textbf{a}_p u(x) = 0, &{} x\in W, \\ u(x)=f(x), &{} x\in \partial _m W. \end{array} \right. \end{aligned}$$

We define the Dirichlet-to-Neumann operator \(\mathfrak {D}_{\textbf{a}_p}\subset L^1(\partial _mW,\nu )\times L^1(\partial _mW,\nu )\) as follows: \((f,\psi )\in \mathfrak {D}_{\textbf{a}_p}\) if

$$\begin{aligned} \mathcal {N}^{\textbf{a}_p}_1 u(x)=\psi (x),\quad x\in \partial _mW, \end{aligned}$$

where u is an \(\textbf{a}_p\)-lifting of f to \(W_m\).

Then, rewriting the operator \(\mathfrak {D}_{\textbf{a}_p}\) as \(\mathcal {B}^{m,\gamma ,\beta }_{\textbf{a}_p}\) for \(\gamma (r)=0\) and \(\beta (r)=r\), \(r\in {\mathbb {R}}\), (\(\Omega _1=W\) and \(\Omega _2=\partial _mW\)), by the results in this subsection we have that \(\mathfrak {D}_{\textbf{a}_p}\) is T-accretive in \(L^1(\partial _mW,\nu )\) (it is easy to see that, in fact, in this situation, it is completely accretive), it satisfies the range condition

$$\begin{aligned} L^{p'}(\partial _mW,\nu )\subset R(I+\mathfrak {D}_{\textbf{a}_p}), \end{aligned}$$

and it has dense domain. The non-homogeneous Cauchy evolution problem for this nonlocal Dirichlet-to-Neumann operator is a particular case of Problem (3.2):

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle - {\text {div}}_m\textbf{a}_p(u)(x)=0, &{}x\in W,\ 0<t<T, \\ \displaystyle u_t(t,x) + \mathcal {N}^{\textbf{a}_p}_1 u(t,x)=g(t,x), &{}x\in \partial _mW, \ 0<t<T, \\ w(0,x) = w_0(x), &{}x\in \partial _mW. \end{array} \right. \end{aligned}$$

See, for example, [2, 3, 24, 37, 44] and the references therein, for other evolution problems with p-Dirichlet-to-Neumann operators, see [16] for the problem with convolution kernels.

3.2 Nonlinear boundary conditions

In this subsection, our aim is to study the following diffusion problem:

$$\begin{aligned} \left( {\text {DP}}_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \quad \left\{ \begin{array}{ll} \displaystyle v_t(t,x) - \int _{\Omega } \textbf{a}_p(x,y,u(t,y)-u(t,x)) dm_x(y)=f(t,x), &{}x\in \Omega _1,\ 0<t<T, \\ \displaystyle v(t,x)\in \gamma \big (u(t,x)\big ), &{} x\in \Omega _1,\ 0<t<T, \\ \displaystyle \int _{\Omega } \textbf{a}_p(x,y,u(t,y)-u(t,x)) dm_x(y) \in \beta \big (u(t,x)\big ), &{}x\in \Omega _2, \ 0<t<T, \\ v(0,x) = v_0(x), &{}x\in \Omega _1, \end{array} \right. \end{aligned}$$

that in particular covers Problem (1.1). See [15] for the reference local model.

We assume that

$$\begin{aligned} \varGamma ^-<\varGamma ^+ \end{aligned}$$

since, otherwise, we do not have an evolution problem. Hence, \(\mathcal {R}_{\gamma ,\beta }^-<\mathcal {R}_{\gamma ,\beta }^+.\) Moreover, we also assume that

$$\begin{aligned} \mathfrak {B}^-<\mathfrak {B}^+, \end{aligned}$$

since the case \( \mathfrak {B}^-=\mathfrak {B}^+\) (\(\beta ={\mathbb {R}}\times \{0\}\)) is treated with more generality in Subsect. 3.1.

We will again make use of nonlinear semigroup theory. To this end, we introduce the corresponding operator associated with \(\left( DP_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \), which is now defined in \(L^1(\Omega _1,\nu )\).

Definition 3.6

We say that \((v,{\hat{v}}) \in B^{m,\gamma ,\beta }_{\textbf{a}_p}\) if \( v,{\hat{v}} \in L^1(\Omega _1,\nu )\) and there exist \( u\in L^p(\Omega ,\nu )\) and \(w\in L^1(\Omega _2,\nu )\) with

$$\begin{aligned} u\in \hbox {Dom}(\gamma ) \hbox { and } v\in \gamma (u) \ \nu \hbox {-a.e. in }\Omega _1, \end{aligned}$$

and

$$\begin{aligned} u\in \hbox {Dom}(\beta ) \hbox { and } w\in \beta (u) \ \nu \hbox {-a.e. in }\Omega _2, \end{aligned}$$

such that

$$\begin{aligned} (x,y)\mapsto a_p(x,y,u(y)-u(x))\in L^{p'}(\Omega \times \Omega ,\nu \otimes m_x) \end{aligned}$$

and

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle - \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y) = {\hat{v}} \quad &{}\hbox {in} \ \ \Omega _1, \\ \displaystyle w- \int _{\Omega } \textbf{a}_p(x,y,u(y)-u(x)) dm_x(y)= 0 \quad &{}\hbox {in} \ \ \Omega _2; \end{array} \right. \end{aligned}$$

that is, [u, (vw)] is a solution of \((GP_{(v+{\hat{v}},\textbf{0})})\), where \(\textbf{0}\) is the null function in \(\Omega _2\) (see (2.8) and Definition 2.4).

Set

$$\begin{aligned} \begin{array}{c} \mathcal {R}_{\gamma ,\lambda \beta }^-:=\nu (\Omega _1)\varGamma ^- + \lambda \nu (\Omega _2) \mathfrak {B}^-, \\ \mathcal {R}_{\gamma ,\lambda \beta }^+:=\nu (\Omega _1)\varGamma ^+ + \lambda \nu (\Omega _2)\mathfrak {B}^+. \end{array} \end{aligned}$$

On account of the results given in Subsect. 2.2 (Theorems 2.6 and 2.7), we have:

Theorem 3.7

The operator \(B^{m,\gamma ,\beta }_{\textbf{a}_p}\) is T-accretive in \(L^1(\Omega ,\nu )\) and satisfies the range condition

$$\begin{aligned} \left\{ \varphi \in L^{p'}(\Omega _1,\nu ):\mathcal {R}_{\gamma ,\lambda \beta }^-<\int _{\Omega _1}\varphi {\text {d}}\nu <\mathcal {R}_{\gamma ,\lambda \beta }^+\right\} \subset R(I+ \lambda B^{m,\gamma ,\beta }_{\textbf{a}_p})\quad \hbox {for every } \lambda >0. \end{aligned}$$

Remark 3.8

Observe that if \(\mathcal {R}_{\gamma ,\beta }^-=-\infty \) and \(\mathcal {R}_{\gamma ,\beta }^+=+\infty \), then the closure of \(B^{m,\gamma ,\beta }_{\textbf{a}_p}\) is m-T-accretive in \(L^1(\Omega _1,\nu )\).

With respect to the domain of this operator, we prove the following result.

Theorem 3.9

$$\begin{aligned} \overline{D(B^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega _1,\nu )}=\big \{v\in L^{p'}(\Omega _1,\nu )\,:\, \varGamma ^-\le v\le \varGamma ^+ \big \}. \end{aligned}$$

Therefore, we also have

$$\begin{aligned} \overline{D(B^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{1}(\Omega _1,\nu )}=\big \{v\in L^{1}(\Omega _1,\nu )\,:\, \varGamma ^-\le v\le \varGamma ^+ \big \}. \end{aligned}$$

Proof

It is obvious that

$$\begin{aligned} \overline{D(B^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega _1,\nu )} \subset \left\{ v\in L^{p'}(\Omega _1,\nu ):\, \varGamma ^-\le v\le \varGamma ^+\ \nu \hbox {-a.e. in }\Omega _1 \right\} . \end{aligned}$$

For the other inclusion, it is enough to see that

$$\begin{aligned} \left\{ v\in L^{\infty }(\Omega _1,\nu ):\, \varGamma ^-\le v\le \varGamma ^+\ \nu \hbox {-a.e. in }\Omega _1 \right\} \subset \overline{D(B^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega _1,\nu )}. \end{aligned}$$

We work on a case-by-case basis.

(A) Suppose that \(\varGamma ^-<0<\varGamma ^+ \). It is enough to see that for any \(v\in L^\infty (\Omega _1,\nu )\) such that there exist \(m\in {\mathbb {R}}\), \(\widetilde{m}<0\), \(\widetilde{M}>0\), \(M\in {\mathbb {R}}\) satisfying

$$\begin{aligned} \varGamma ^-<m<\widetilde{m}< v<\widetilde{M}<M<\varGamma ^+ \ \nu \hbox {-a.e. in } \Omega _1 \end{aligned}$$

it holds that \(v\in \overline{D(B^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega _1,\nu )}\).

By the results in Subsect. 2.2.4 we know that, for \(n\in {\mathbb {N}}\), there exist \(u_n\in L^p(\Omega ,\nu )\), \(v_n\in L^{p'}(\Omega _1,\nu )\) and \(w_n\in L^{p'}(\Omega _2,\nu )\), such that \([u_n,(v_n,\frac{1}{n} w_n)]\) is a solution of \(\left( GP_{(v,\textbf{0})}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \), i.e., \(v_n\in \gamma (u_n)\) \(\nu \)-a.e. in \(\Omega _1\), \(w_n\in \beta (u_n)\) \(\nu \)-a.e. in \(\Omega _2\) and

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle v_n(x) -\frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,u_n(y)-u_n(x)) dm_x(y) = v(x), &{} \hbox {for } x\in \Omega _1,\\ \displaystyle w_n(x) - \int _{\Omega } \textbf{a}_p(x,y,u_n(y)-u_n(x)) dm_x(y) = 0, &{} \hbox {for } x\in \Omega _2. \end{array}\right. \end{aligned}$$

In other words, \((v_n,n(v-v_n))\in B^{m,\gamma ,\beta }_{\textbf{a}_p}\) or, equivalently,

$$\begin{aligned} v_n:=\left( I+\frac{1}{n}B^{m,\gamma ,\beta }_{\textbf{a}_p}\right) ^{-1}(v) \in D(B^{m,\gamma ,\beta }_{\textbf{a}_p}). \end{aligned}$$

Let us see that \(v_n{\mathop {\longrightarrow }\limits ^{n}} v\) in \(L^{p'}(\Omega _1,\nu )\).

(A1) Suppose first that \(\hbox {sup} D(\beta )=+\infty \). Take \(a_M>0\) such that \(M\in \gamma (a_M)\) and let \(N\in \beta (a_M)\). Let

$$\begin{aligned} \widehat{v}(x):= & {} \left\{ \begin{array}{cc} M, &{} x\in \Omega _1, \\ N, &{} x\in \Omega _2, \end{array}\right. \\ \widehat{u}(x):= & {} a_M, \ x\in \Omega , \end{aligned}$$

and

$$\begin{aligned} \varphi (x):=\left\{ \begin{array}{cc} \displaystyle M, &{} x\in \Omega _1, \\ \displaystyle 0, &{} x\in \Omega _2. \end{array}\right. \end{aligned}$$

Then, \([\widehat{u}, \widehat{v}]\) is a supersolution of \(\big (GP^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }_{\varphi }\big )\) and \((v,\textbf{0})\le \varphi \). Thus, by the maximum principle (Theorem 2.6),

$$\begin{aligned} u_n\le \widehat{u}=a_M \, \nu \hbox {-a.e. in }\Omega \quad \hbox {for every } n\in {\mathbb {N}}. \end{aligned}$$

(A2) Suppose now that \(\hbox {sup}D(\beta )=r_\beta <+\infty \). Again, by the results in Subsect. 2.2.4 we know that, for \(n\in {\mathbb {N}}\), there exist \(\widetilde{u}_n\in L^p(\Omega ,\nu )\), \(\widetilde{v}_n\in L^{p'}(\Omega _1,\nu )\) and \(\widetilde{w}_n\in L^{p'}(\Omega _2,\nu )\), such that \([\widetilde{u}_n,(\widetilde{v}_n,\frac{1}{n}\widetilde{w}_n)]\) is a solution of \(\big (GP_{(M,\textbf{0})}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\big )\). Therefore, by the maximum principle (Theorem 2.6),

$$\begin{aligned} v_n\le \widetilde{v}_n\quad \nu \hbox {-a.e. in }\Omega _1. \end{aligned}$$

Now, since \(\widetilde{v}_n\ll M\) in \(\Omega _1\) (recall Remark 2.8(iii)), we have that \(\widetilde{v}_n\le M\) and, consequently, also \(v_n\le M\). Hence, since \(M\le \widetilde{M}<\varGamma ^+ \),

$$\begin{aligned} u_n\le \inf \big (\gamma ^{-1}(\widetilde{M})\big ) \, \nu \hbox {-a.e. in }\Omega _1, \end{aligned}$$

but we also have

$$\begin{aligned} u_n\le r_\beta \, \nu \hbox {-a.e. in } \Omega _2\quad \hbox {for every } n\in {\mathbb {N}}. \end{aligned}$$

(B) For \(\varGamma ^-<0=\varGamma ^+\): let \(\varGamma ^-<m<\widetilde{m}< 0\), and \(v\in L^\infty (\Omega _1,\nu )\) be such that

$$\begin{aligned} \widetilde{m}\le v< 0. \end{aligned}$$

As in the previous case, by the results in Subsect. 2.2.4, we know that, for \(n\in {\mathbb {N}}\), there exist \(u_n\in L^p(\Omega ,\nu )\), \(v_n\in L^{p'}(\Omega _1,\nu )\) and \(w_n\in L^{p'}(\Omega _2,\nu )\), such that \([u_n,(v_n,\frac{1}{n} w_n)]\) is a solution of \(\left( GP_{(v,\textbf{0})}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \). Then, since for the null function \(\textbf{0}\) in \(\Omega \), \([\textbf{0},\textbf{0}]\) is a solution of \(\left( GP_\textbf{0}^{\frac{1}{n}\textbf{a}_p,\gamma ,\beta }\right) \) and \(v<0\), the maximum principle yields

$$\begin{aligned} u_n\le 0 \ \nu \text{-a.e. } \text{ in } \Omega \quad \hbox {for every } n\in {\mathbb {N}}. \end{aligned}$$

Therefore, in all the cases, \(\{u_n\}_n\) is \(L^\infty (\Omega ,\nu )\)-bounded from above. With a similar reasoning we obtain that, in any of these cases, \(\{u_n\}_n\) is also \(L^\infty (\Omega ,\nu )\)-bounded from below. Then, since

$$\begin{aligned} v_n(x)-v(x)=\frac{1}{n}\int _{\Omega } \textbf{a}_p(x,y,u_n(y)-u_n(x)) dm_x(y) \ \ \hbox { in} \ \Omega _1, \end{aligned}$$

we obtain that

$$\begin{aligned} v_n{\mathop {\longrightarrow }\limits ^{n}} v \ \hbox { in } L^{p'}(\Omega _1,\nu ) \end{aligned}$$

as desired. \(\square \)

The following theorem gives the existence and uniqueness of solutions of Problem \(\left( {\text {DP}}_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \). Recall that \(\varGamma ^-<\varGamma ^+\) and \(\mathfrak {B}^-<\mathfrak {B}^+\).

Theorem 3.10

Let \(T>0\). Let \(v_0\in L^{1}(\Omega _1,\nu )\) and \(f\in L^1(0,T;L^{1}(\Omega _1,\nu ))\). Assume

$$\begin{aligned} \varGamma ^-\le v_0\le \varGamma ^+ \nu \hbox {-a.e. in }\Omega _1, \end{aligned}$$

and

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^+=+\infty \ \hbox { or } \ \int _{\Omega _1}f(x,t){\text {d}}\nu (x)\le \nu (\Omega _2)\mathfrak {B}^+\quad \hbox {for every } 0<t<T, \end{aligned}$$

and

$$\begin{aligned} \mathcal {R}_{\gamma ,\beta }^-=-\infty \ \hbox { or } \ \displaystyle \int _{\Omega _1}f(x,t)d\nu (x)\ge \nu (\Omega _2)\mathfrak {B}^-\quad \hbox {for every } 0<t<T. \end{aligned}$$

Then, there exists a unique mild-solution \(v\in C([0,T];L^1(\Omega _1,\nu ))\) of \(\displaystyle \left( DP_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \).

Let v and \({\widetilde{v}}\) be the mild solutions of the problem with respective data \(v_0,\ {\widetilde{v}}_0\in L^{1}(\Omega _1,\nu )\) and \(f,\ {\widetilde{f}}\in L^1(0,T;L^{1}(\Omega _1,\nu ))\). Then,

$$\begin{aligned} \begin{array}{rl} &{} \displaystyle \int _{\Omega _1} \left( v(t,x)-{\widetilde{v}}(t,x)\right) ^+{\text {d}}\nu (x)\\ &{} \displaystyle \qquad \le \int _{\Omega _1} \left( v_0(x)-{\widetilde{v}}_0(x)\right) ^+{\text {d}}\nu (x)\\ &{} \displaystyle \qquad \quad +\int _0^t\int _{\Omega _1}\left( f(s,x)-{\widetilde{f}}(s,x)\right) ^+{\text {d}}\nu (x){\text {d}}s \quad \hbox {for every } 0\le t\le T. \end{array} \end{aligned}$$

Under the additional assumptions

$$\begin{aligned}{} & {} \displaystyle v_0\in L^{p'}(\Omega _1,\nu ) \hbox { and } f\in L^{p'}(0,T;L^{p'}(\Omega _1,\nu )) \hbox { with }\nonumber \\{} & {} \displaystyle \int _{\Omega _1}j_\gamma ^*(v_0){\text {d}}\nu<+\infty \hbox { and } \nonumber \\{} & {} \displaystyle \int _{\Omega _1}v_0^+{\text {d}}\nu +\int _0^T\int _{\Omega _1}f(s)^+{\text {d}}\nu {\text {d}}t< \nu (\Omega _1)\Gamma ^+,\nonumber \\{} & {} \displaystyle \int _{\Omega _1}v_0^-{\text {d}}\nu +\int _0^T\int _{\Omega _1}f(s)^-{\text {d}}\nu {\text {d}}t<-\nu (\Omega _1)\Gamma ^-, \end{aligned}$$
(3.21)

the mild solution v belongs to \(W^{1,1}(0,T;L^{1}(\Omega _1,\nu ))\) and satisfies the equation

$$\begin{aligned} \left\{ \begin{array}{l}\partial _tv(t)+B^{m,\gamma ,\beta }_{\textbf{a}_p}v(t)\ni f(t)\quad \hbox {for a.e. }t\in (0,T),\\ v(0)=v_0,\end{array}\right. \end{aligned}$$

that is, v is a strong solution.

The proof of this result differs, strongly at some points, from the proof of Theorem 3.4.

Proof

We start by proving the existence of mild solutions. For \(n\in {\mathbb {N}}\), consider the partition

$$\begin{aligned} t_0^n=0<t_1^n<\cdots<t_{n-1}^n<t_n^n=T \end{aligned}$$

where \(t_i^n:=iT/n\), \(i=1,\ldots ,n\). Given \(\epsilon >0\), since \(\mathfrak {B}^-<\mathfrak {B}^+\), there exist \(n\in \mathbb {N}\), \(v_0^n \in \overline{D(B^{m,\gamma ,\beta }_{\textbf{a}_p})}^{L^{p'}(\Omega _1,\nu )}\) (i.e., \(v_0^n\in L^{p'}(\Omega _1,\nu )\) satisfying \(\varGamma ^-\le v_0^n\le \varGamma ^+\)) and \(f_i^n \in L^{p'}(\Omega _1,\nu )\), \(i=1,\ldots n\), such that \(T/n\le \epsilon \),

$$\begin{aligned} \Vert v_0-v_0^n\Vert _{L^1(\Omega _1,\nu )} \le \epsilon , \end{aligned}$$
$$\begin{aligned} \sum _{i=1}^n\int _{t_{i-1}^n}^{t_i^n}\Vert f(t)- f_i^n\Vert _{L^1(\Omega _1,\nu )}{\text {d}}t \le \epsilon \end{aligned}$$
(3.22)

and

$$\begin{aligned} \nu (\Omega _2)\mathfrak {B}^-<\int _{\Omega _1}f_i^n{\text {d}}\nu <\nu (\Omega _2)\mathfrak {B}^+. \end{aligned}$$

Then, setting

$$\begin{aligned} f_n(t):=f_i^n, \ \hbox { for } t\in ]t_{i-1}^n,t_i^n], i=1,\ldots , n, \end{aligned}$$

we have that

$$\begin{aligned} \int _0^T \Vert f(t)- f_n(t) \Vert _{L^{1}(\Omega _1,\nu )}{\text {d}}t \le \epsilon . \end{aligned}$$

Using the results in Subsect. 2.2.4, we see that, for n large enough, we may recursively find a solution \([u_i^n,(v_i^n,\frac{T}{n} w_i^n)]\) of \(\displaystyle \left( GP^{\frac{T}{n}\textbf{a}_p,\gamma ,\frac{T}{n}\beta }_{\left( \frac{T}{n} f_i^n+v_{i-1}^n,\textbf{0}\right) }\right) \), \(i=1,\ldots ,n\), so that

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle v_i^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)=\frac{T}{n} f_i^n(x)+v_{i-1}^n(x), &{} \displaystyle x\in \Omega _1\\ \displaystyle w_i^n(x)- \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)=0, &{} \displaystyle x\in \Omega _2, \end{array}\right. \nonumber \\ \end{aligned}$$
(3.23)

or, equivalently,

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{v_i^n(x)-v_{i-1}^n(x)}{T/n}-\int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)= f_i^n(x), &{} \displaystyle x\in \Omega _1\\ \displaystyle w_i^n(x)- \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))dm_x(y)=0, &{} \displaystyle x\in \Omega _2, \end{array}\right. \nonumber \\ \end{aligned}$$
(3.24)

with \(v_i^n(x)\in \gamma (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _1\) and \(w_i^n(x)\in \beta (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _2\), \(i=1,\ldots ,n\). That is, we may find the unique solution \(v_i^n\) of the time discretization scheme associated with \(\left( DP_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \).

To apply these results, we must ensure that

$$\begin{aligned} \mathcal {R}_{\gamma ,\frac{T}{n}\beta }^-<\int _{\Omega _1} \left( \frac{T}{n} f_i^n + v_{i-1}^n\right) {\text {d}}\nu <\mathcal {R}_{\gamma ,\frac{T}{n}\beta }^+ \end{aligned}$$

holds for each step, but this holds true thanks to the choice of the \(f_i^n\), \(i=1,\ldots ,n\).

Therefore,

$$\begin{aligned} v_n (t):=\left\{ \begin{array}{ll} v_0^n, &{} \hbox {if } t=0,\\ v_i^n, &{} \hbox {if } t\in ]t_{i-1}^n,t_i^n],\, i=1,\ldots ,n, \end{array} \right. \end{aligned}$$

is an \(\epsilon \)-approximate solution of Problem \(\left( DP_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \). Consequently, by nonlinear semigroup theory ((see [11, 10, Theorem 4.1], or [8, Theorem A.27])) and on account of Theorem 3.7 and Theorem 3.9, we have that \(\left( DP_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \) has a unique mild solution \(v\in C([0,T];L^1(\Omega _1,\nu ))\) with

$$\begin{aligned} v_n(t){\mathop {\longrightarrow }\limits ^{n}} v(t) \ \hbox {in}\quad L^1(\Omega _1,\nu ) \,\hbox {uniformly for}\, t\in [0,T]. \end{aligned}$$
(3.25)

Uniqueness and the maximum principle for mild solutions is guaranteed by the T-accretivity of the operator.

We now prove, step by step, that these mild solutions are strong solutions of Problem \(\left( DP_{f,v_0}^{ \textbf{a}_p,\gamma ,\beta }\right) \) under the set of assumptions given in (3.21).

Let us define

$$\begin{aligned} u_n (t):= u_i^n \quad \hbox {for } t\in ]t_{i-1}^n,t_i^n],\, i=1,\ldots ,n, \end{aligned}$$

and

$$\begin{aligned} w_n (t):= w_i^n \quad \hbox {for } t\in ]t_{i-1}^n,t_i^n],\, i=1,\ldots ,n. \end{aligned}$$

Step 1. Suppose first that \(\mathcal {R}_{\gamma ,\beta }^-=-\infty \) and \(\mathcal {R}_{\gamma ,\beta }^+=+\infty \).

In the construction of the mild solution, we now take \(v_0^n=v_0\) (since \(v_0\in L^{p'}(\Omega _1,\nu )\)) and the functions \(f_i^n \in L^{p'}(\Omega _1,\nu )\), \(i=1,\ldots n\), additionally satisfying

$$\begin{aligned} \sum _{i=1}^n\int _{t_{i-1}^n}^{t_i^n}\Vert f(t)- f_i^n\Vert ^{p'}_{L^{p'}(\Omega _1,\nu )}{\text {d}}t\le \epsilon \end{aligned}$$

and

$$\begin{aligned} \nu (\Omega _2)\mathfrak {B}^-<\int _{\Omega _1}f_i^n{\text {d}}\nu <\nu (\Omega _2)\mathfrak {B}^+. \end{aligned}$$

Multiplying both equations in (3.24) by \(u_i^n\), integrating with respect to \(\nu \) over \(\Omega _1\) and \(\Omega _2\), respectively, and adding them, we obtain

$$\begin{aligned} \begin{array}{l} \displaystyle \int _{\Omega _1}\frac{v_i^n(x)-v_{i-1}^n(x)}{T/n}u_i^n(x){\text {d}}\nu (x)+\int _{\Omega _2}w_i^n(x) u_i^n(x){\text {d}}\nu (x) \\ \displaystyle \qquad -\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))u_i^n(x)dm_x(y){\text {d}}\nu (x)\\ \displaystyle = \int _{\Omega _1} f_i^n(x)u_i^n(x) {\text {d}}\nu (x). \end{array} \end{aligned}$$

Then, since \(w_i^n(x)\in \beta (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _2\) the second term on the left-hand side is nonnegative and integrating by parts the third term, we get

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1}\frac{v_i^n(x)-v_{i-1}^n(x)}{T/n}u_i^n(x){\text {d}}\nu (x) \nonumber \\{} & {} \quad +\frac{1}{2} \int _\Omega \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))(u_i^n(y)-u_i^n(x))dm_x(y){\text {d}}\nu (x) \nonumber \\{} & {} \displaystyle \le \int _{\Omega _1} f_i^n(x)u_i^n(x) {\text {d}}\nu (x). \end{aligned}$$
(3.26)

Now, since \(v_i^n(x)\in \gamma (u_i^n(x))\) for \(\nu \)-a.e. \(x\in \Omega _1\),

$$\begin{aligned} u_i^n(x)\in \gamma ^{-1}(v_i^n(x))=\partial j_\gamma ^*(v_i^n(x)) \ \hbox { for}\, \nu \hbox {-a.e.}\, x\in \Omega _1. \end{aligned}$$

Consequently,

$$\begin{aligned} j_\gamma ^*(v_{i-1}^n(x))- j_\gamma ^*(v_{i}^n(x))\ge (v_{i-1}^n(x)-v_i^n(x))u_i^n(x) \ \hbox { for}\, \nu \hbox {-a.e.}\, x\in \Omega _1. \end{aligned}$$

Therefore, from (3.26), it follows that

$$\begin{aligned}{} & {} \displaystyle \frac{n}{T}\int _{\Omega _1}(j_\gamma ^*(v_{i}^n(x))- j_\gamma ^*(v_{i-1}^n(x))){\text {d}}\nu (x)\\{} & {} \quad +\frac{1}{2} \int _\Omega \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))(u_i^n(y)-u_i^n(x))dm_x(y){\text {d}}\nu (x)\\{} & {} \displaystyle \le \int _{\Omega _1} f_i^n(x)u_i^n(x) {\text {d}}\nu (x), \end{aligned}$$

\(i=1,\ldots ,n\). Then, integrating this equation over \(]t_{i-1},t_i]\) and adding for \(1\le i \le n\), we get

$$\begin{aligned} \begin{array}{l} \displaystyle \int _{\Omega _1}(j_\gamma ^*(v_{n}^n(x))- j_\gamma ^*(v_{0}(x))){\text {d}}\nu (x)\\ \displaystyle \qquad +\frac{1}{2} \sum _{i=1}^n\int _{t_{i-1}}^{t_i}\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_i^n(y)-u_i^n(x))(u_i^n(y)-u_i^n(x))dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \sum _{i=1}^n\int _{t_{i-1}}^{t_i}\int _{\Omega _1} f_i^n(x)u_i^n(x) {\text {d}}\nu (x){\text {d}}t, \end{array} \end{aligned}$$

which, recalling the definitions of \(f_n\), \(u_n\), \(v_n\) and \(w_n\), can be rewritten as:

$$\begin{aligned}{} & {} \displaystyle \int _{\Omega _1}(j_\gamma ^*(v_{n}^n(x))- j_\gamma ^*(v_{0}(x))){\text {d}}\nu (x) \nonumber \\{} & {} \displaystyle \qquad +\frac{1}{2} \int _0^T\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_n(t)(y)-u_n(t)(x))(u_n(t)(y)-u_n(t)(x))dm_x(y){\text {d}}\nu (x){\text {d}}t \nonumber \\{} & {} \displaystyle \le \int _0^T\int _{\Omega _1} f_n(t)(x)u_n(t)(x) {\text {d}}\nu (x){\text {d}}t. \end{aligned}$$
(3.27)

This, together with (2.5) and the fact that \(j^*_\gamma \) is nonnegative, yields

$$\begin{aligned} \begin{array}{l} \displaystyle \frac{c_p}{2}\int _0^T\int _\Omega \int _\Omega |u_n(t)(y)-u_n(t)(x)|^p dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \frac{1}{2}\int _0^T\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_n(t)(y)-u_n(t)(x))(u_n(t)(y)-u_n(t)(x))dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \int _{\Omega _1} j_\gamma ^*(v_{0}(x)){\text {d}}\nu (x)+ \int _0^T\int _{\Omega _1} f_n(t)(x)u_n(t)(x) {\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \int _{\Omega _1} j_\gamma ^*(v_{0}(x)){\text {d}}\nu (x)+ \int _0^T \Vert f_n(t)\Vert _{L^{p'}(\Omega _1,\nu )} \Vert u_n(t) \Vert _{L^{p}(\Omega _1,\nu )} {\text {d}}t. \end{array} \end{aligned}$$

Therefore, for any \(\delta >0\), by (3.21) and Young’s inequality, there exists \(C(\delta )>0\) such that, in particular,

$$\begin{aligned} \displaystyle \int _0^T\int _{\Omega }\int _{\Omega } |u_n(t)(y)-u_n(t)(x)|^p dm_x(y){\text {d}}\nu (x){\text {d}}t \le C(\delta )+ \delta \int _0^T \Vert u_n(t) \Vert _{L^{p}(\Omega _1,\nu )}^{p} {\text {d}}t. \nonumber \\ \end{aligned}$$
(3.28)

Observe also that, for any \(n\in {\mathbb {N}}\) and \(i\in \{1,\ldots ,n\}\), and for \(t\in ]t_{i-1}^n,t_i^n]\),

$$\begin{aligned} \int _{\Omega _1}v_n^+(t) {\text {d}}\nu +\int _0^{t_i^n}\int _{\Omega _2}w_n^+(s) {\text {d}}\nu {\text {d}}s\le \int _{\Omega _1}v_{0}^+ {\text {d}}\nu + \int _0^{t_i^n}\int _{\Omega _1}f_n^+(s) {\text {d}}\nu {\text {d}}s. \nonumber \\ \end{aligned}$$
(3.29)

Indeed, multiplying the first equation in (3.23) by \(\frac{1}{r}T_r^+(u_i^n)\) and integrating with respect to \(\nu \) over \(\Omega _1\), then multiplying the second by \(\frac{T}{n}\frac{1}{r}T_r^+(u_i^n)\) and integrating with respect to \(\nu \) over \(\Omega _2\), adding both equations, neglecting the nonnegative term involving \(\textbf{a}_p\) (recall Remark 2.5) and letting \(r\downarrow 0\), we get that

$$\begin{aligned} \int _{\Omega _1}(v_i^n)^+ {\text {d}}\nu +\frac{T}{n}\int _{\Omega _2}(w_i^n)^+ {\text {d}}\nu \le \int _{\Omega _1}(v_{i-1}^n)^+ {\text {d}}\nu + \frac{T}{n}\int _{\Omega _1}(f_i^n)^+ {\text {d}}\nu , \end{aligned}$$

i.e.,

$$\begin{aligned} \int _{\Omega _1}(v_i^n)^+ {\text {d}}\nu \le \int _{\Omega _1}(v_{i-1}^n)^+ {\text {d}}\nu + \frac{T}{n}\int _{\Omega _1}(f_i^n)^+ {\text {d}}\nu -\frac{T}{n}\int _{\Omega _2}(w_i^n)^+ {\text {d}}\nu . \end{aligned}$$

Therefore,

$$\begin{aligned} \int _{\Omega _1}(v_i^n)^+ {\text {d}}\nu \le \int _{\Omega _1}(v_{0}^n)^+ {\text {d}}\nu + \sum _{j=1}^i\frac{T}{n}\int _{\Omega _1}(f_j^n)^+ {\text {d}}\nu -\sum _{j=1}^i\frac{T}{n}\int _{\Omega _2}(w_j^n)^+ {\text {d}}\nu \end{aligned}$$

which is equivalent to (3.29).

Now, by (3.25), if \(\varGamma ^+=+\infty \), there exists \(M>0\) such that

$$\begin{aligned} \sup _{t\in [0,T]}\int _{\Omega _1} v_n^+(t)(x){\text {d}}\nu (x)<M \ \hbox { for every } n\in {\mathbb {N}}. \end{aligned}$$

Consequently, Lemma A.7 applied for \(A=\Omega _1\), \(B=\emptyset \) and \(\alpha =\gamma \), yields

$$\begin{aligned} \Vert u_n^+(t)\Vert _{L^{p}(\Omega _1,\nu )}\le K_2\left( \left( \int _{\Omega _1} \int _{\Omega _1} |u_n^+(t)(y)-u_n^+(t)(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^\frac{1}{p}+1 \right) \end{aligned}$$

for every \(n\in {\mathbb {N}}\), every \(0\le t \le T\) and some constant \( K_2>0\).

Suppose now that \(\varGamma ^+<+\infty \). Then, by (3.29) we have that, for any \(n\in {\mathbb {N}}\) and \(i\in \{1,\ldots ,n\}\), and for \(t\in ]t_{i-1}^n,t_i^n]\) if \(i\ge 2\), or \(t\in [t_{0}^n,t_1^n]\) if \(i=1\),

$$\begin{aligned} \int _{\Omega _1}v_n^+(t) {\text {d}}\nu \le \int _{\Omega _1}v_{0}^+ {\text {d}}\nu + \int _0^{t_i^n}\int _{\Omega _1}f_n^+(s) {\text {d}}\nu {\text {d}}s \end{aligned}$$

thus, by the assumptions in (3.21) and by (3.22), there exists \(M\in {\mathbb {R}}\) such that

$$\begin{aligned} \sup _{t\in [0,T]}\int _{\Omega _1} v_n(t){\text {d}}\nu \le M<\nu (\Omega _1)\Gamma ^+ \end{aligned}$$

for n sufficiently large and, by (3.25), such that

$$\begin{aligned} \sup _{t\in [0,T]}\int _{\{x\in \Omega _1 \,:\, v_n(t)<-h\}}|v_n(t)|{\text {d}}\nu <\frac{\nu (\Omega _1)\Gamma ^+-M}{8} \end{aligned}$$

for n sufficiently large. Therefore, we may apply Lemma A.8 for \(A=\Omega _1\), \(B=\emptyset \) and \(\alpha =\gamma \) to conclude that there exists a constant \(K_2'>0\) such that

$$\begin{aligned}{} & {} \Vert u_n^+(t)\Vert _{L^{p}(\Omega _1,\nu )}\le K_2'\left( \left( \int _{\Omega _1} \int _{\Omega _1} |u_n^+(t)(y)-u_n^+(t)(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^\frac{1}{p}+1 \right) \\ {}{} & {} \quad \hbox {for every } 0\le t \le T, \end{aligned}$$

for n sufficiently large.

Similarly, we may find \( K_3>0\) such that

$$\begin{aligned}{} & {} \Vert u_n^-(t)\Vert _{L^{p}(\Omega _1,\nu )}\le K_3\left( \left( \int _{\Omega _1} \int _{\Omega _1} |u_n^-(t)(y)-u_n^-(t)(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^\frac{1}{p}+1 \right) \\ {}{} & {} \quad \hbox {for every } 0\le t \le T, \end{aligned}$$

for n sufficiently large.

Consequently, by the generalised Poincaré-type inequality together with (3.28) for \(\delta \) small enough, we get

$$\begin{aligned} \int _0^T \left\| u_n(t) \right\| _{L^p(\Omega ,\nu )}^{p}{\text {d}}t\le K_4 \quad \hbox {for every } n\in {\mathbb {N}}, \end{aligned}$$

for some constant \(K_4>0\), that is, \(\{u_n\}_n\) is bounded in \(L^p(0,T;L^p(\Omega ,\nu ))\). Therefore, there exists a subsequence, which we continue to denote by \(\{u_{n}\}_{n}\), and \(u\in L^p(0,T;L^p(\Omega ,\nu ))\) such that

$$\begin{aligned} u_{n}{\mathop {\rightharpoonup }\limits ^{n}} u \ \hbox { weakly in } L^p(0,T;L^p(\Omega ,\nu )). \end{aligned}$$

Note that, since \( \displaystyle \Big \{\int _0^T\int _\Omega \int _\Omega |u_n(t)(y)-u_n(t)(x)|^p dm_x(y){\text {d}}\nu (x){\text {d}}t\Big \}_n \) is bounded, then, by (2.4), we have that \(\{[(t,x,y)\mapsto \textbf{a}_p(x,y,u_n(t)(y)-u_n(t)(x))]\}_n\) is bounded in \(L^{p'}(0,T; L^{p'}(\Omega \times \Omega ,\nu \otimes m_x))\) so we may take a further subsequence, which we continue to denote in the same way, such that

$$\begin{aligned}{}[(t,x,y)\mapsto \textbf{a}_p(x,y,u_{n}(t)(y)-u_{n}(t)(x))]{\mathop {\rightharpoonup }\limits ^{n}} \Phi , \ \hbox { weakly in } L^{p'}(0,T; L^{p'}(\Omega \times \Omega ,\nu \otimes m_x)). \end{aligned}$$

Now, let \(\Psi \in W^{1,1}_0(0,T;L^p(\Omega ,\nu ))\), \(\hbox {supp}(\Psi )\subset \subset [0,T]\), then

$$\begin{aligned}{} & {} \int _0^T \frac{v_{n}(t)(x)-v_{n}(t-T/n)(x)}{T/n}\Psi (t)(x){\text {d}}t \nonumber \\{} & {} =-\int _0^{T-T/n} v_{n}(t)(x)\frac{\Psi (t+T/n)(x)-\Psi (t)(x)}{T/n}{\text {d}}t+\int _{T-T/n}^T \frac{v_{n}\Psi (t)(x)}{T/n}{\text {d}}t\nonumber \\{} & {} \quad -\int _0^{T/n} \frac{z_0\Psi (t)(x)}{T/n} \end{aligned}$$

for \(x\in \Omega _1\). Therefore, multiplying both equations in (3.24) by \(\Psi \), integrating the first one over \(\Omega _1\) and the second one over \(\Omega _2\) with respect to \(\nu \), adding them, and taking limits as \(n\rightarrow +\infty \) we get that

$$\begin{aligned}{} & {} \displaystyle -\int _0^T \int _{\Omega _1} v(t)(x)\frac{d}{{\text {d}}t}\Psi (t)(x){\text {d}}\nu (x) {\text {d}}t+\int _0^T \int _{\Omega _2}w(t)(x)\Psi (t)(x){\text {d}}\nu (x){\text {d}}t \\{} & {} \quad \displaystyle - \int _0^T\int _\Omega \int _\Omega \Phi (t,x,y)dm_x(y)\Psi (t)(x){\text {d}}\nu (x){\text {d}}t \\ {}{} & {} \displaystyle = \int _0^T\int _{\Omega _1} f(t)(x)\Psi (t)(x){\text {d}}\nu (x){\text {d}}t. \end{aligned}$$

Therefore, taking \(\Psi (t)(x)=\psi (t)\xi (x)\), where \(\psi \in C_c^\infty (0,T)\) and \(\xi \in L^p(\Omega ,\nu )\), we obtain that

$$\begin{aligned} \displaystyle \int _0^T v(t)(x)\psi '(t) {\text {d}}t=-\int _0^T \int _\Omega \Phi (t,x,y)\psi (t)dm_x(y) {\text {d}}t - \int _0^T f(t)(x)\psi (t) {\text {d}}t \end{aligned}$$

for \(\nu \)-a.e. \(x\in \Omega _1\).

It follows that

$$\begin{aligned} \displaystyle v'(t)(x) = \int _\Omega \Phi (t,x,y)dm_x(y) + f(t)(x) \quad \hbox {for a.e. } t\in (0,T) \hbox { and } \nu \hbox {-a.e. } x\in \Omega _1. \end{aligned}$$

Therefore, since \(v\in C([0,T];L^1(\Omega _1,\nu ))\), \(\Phi \in L^{p'}(0,T; L^{p'}(\Omega \times \Omega ,\nu \otimes m_x))\) and \(f\in L^{p'}(0,T;L^{p'} (\Omega _1,\nu ))\), we get that \(v'\in L^{p'}(0,T;L^{p'}(\Omega _1,\nu ))\) and \(v\in W^{1,1}(0,T;L^{1}(\Omega _1,\nu ))\).

Then, by Remark 3.8, we conclude that the mild solution v is, in fact, a strong solution (see [11] or [8, Corollary A.34]). Hence,

$$\begin{aligned}{} & {} v'(t)(x) - \int _\Omega \textbf{a}_p(x,y,u(t)(y)-u(t)(x))dm_x(y) = f(t)(x) \nonumber \\{} & {} \quad \hbox {for a.e. } t\in (0,T) \hbox { and } \nu \hbox {-a.e. } x\in \Omega _1. \end{aligned}$$
(3.30)

Let us see, for further use, that \(\displaystyle \int _{\Omega _1}j_{\gamma }^*(v(t)){\text {d}}\nu \in W^{1,1}(0,T)\). By (3.27) and Fatou’s lemma, we have

$$\begin{aligned} \begin{array}{l} \displaystyle \limsup _n\frac{1}{2}\int _0^T\int _\Omega \int _\Omega \textbf{a}_p(x,y,u_n(t)(y)-u_n(t)(x))(u_n(t)(y)-u_n(t)(x))dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le -\int _{\Omega _1}(j_\gamma ^*(v(T)(x))- j_\gamma ^*(v(0)(x))){\text {d}}\nu (x) +\int _0^T\int _{\Omega _1}f(t)(x)u(t)(x) {\text {d}}\nu (x){\text {d}}t. \end{array} \end{aligned}$$

Moreover, by (3.30),

$$\begin{aligned} \int _0^T v(t)(x)\frac{d}{{\text {d}}t}\Psi (t)(x){\text {d}}t=\int _0^T F(t)(x)\Psi (t)(x){\text {d}}t, \end{aligned}$$
(3.31)

where F is given by

$$\begin{aligned} F(t)(x)=- \int _\Omega \textbf{a}_p(x,y,u(t)(y)-u(t)(x))dm_x(y)-f(t)(x), \ \ \ x\in \Omega _1. \end{aligned}$$

Let \(\psi \in C_c^\infty (0,T)\), \(\psi \ge 0\), \(\tau >0\) and

$$\begin{aligned} \eta _\tau (t)(x):=\frac{1}{\tau }\int _t^{t+\tau } u(s)(x) \psi (s){\text {d}}s, \ \ \ t\in [0,T], \, x\in \Omega _1. \end{aligned}$$

Then, for \(\tau \) small enough, \(\eta _\tau \in W^{1,1}_0(0,T;L^p(\Omega _1,\nu ))\) and we may use it as a test function in (3.31) to obtain

$$\begin{aligned} \begin{array}{rl} &{} \displaystyle \int _0^T \int _{\Omega _1} F(t)(x)\eta _\tau (t)(x) {\text {d}}\nu (x) {\text {d}}t\\ &{}\displaystyle =\int _0^T \int _{\Omega _1} v(t)(x) \frac{d}{{\text {d}}t}\eta _\tau (t)(x){\text {d}}\nu (x){\text {d}}t\\ &{}\displaystyle = \int _0^T \int _{\Omega _1}v(t)(x)\frac{u(t+\tau )(x) \psi (t+\tau )-u(t)(x)\psi (t)}{\tau }{\text {d}}\nu (x) {\text {d}}t \\ &{}\displaystyle =\int _0^T \int _{\Omega _1}\frac{v(t-\tau )(x)-v(t)(x)}{\tau } u(t)(x)\psi (t) {\text {d}}\nu (x){\text {d}}t. \end{array} \end{aligned}$$

Now, since

$$\begin{aligned}{} & {} \gamma ^{-1}(r)= \partial j_{\gamma ^{-1}}(r)=\partial \left( \int _0^r (\gamma ^{-1})^0(s) {\text {d}}s\right) ,\\{} & {} \displaystyle \int _0^T \int _{\Omega _1}F(t)(x)\eta _\tau (t)(x) {\text {d}}\nu (x){\text {d}}t \displaystyle \le \frac{1}{\tau }\int _0^T\int _{\Omega _1}\int _{v(t) (x)}^{v(t-\tau )(x)}(\gamma ^{-1})^0(s){\text {d}}s \psi (t){\text {d}}\nu (x){\text {d}}t\\{} & {} \quad \displaystyle =\int _0^T\int _{\Omega _1}\int _0^{v(t)(x)}(\gamma ^{-1})^0(s){\text {d}}s \frac{\psi (t+\tau )-\psi (t)}{\tau }{\text {d}}\nu (x){\text {d}}t, \end{aligned}$$

which, letting \(\tau \rightarrow 0^+\) yields

$$\begin{aligned} \begin{array}{rl} \displaystyle \int _0^T \int _{\Omega _1} F(t)u(t)(x)\psi (t){\text {d}}\nu (x) {\text {d}}t &{}\displaystyle \le \int _0^T \int _{\Omega _1}\int _0^{v(t)(x)}(\gamma ^{-1})^0(s){\text {d}}s \Psi '(t) {\text {d}}\nu (x) {\text {d}}t\\ &{}\displaystyle =\int _0^T\int _{\Omega _1} j_{\gamma ^{-1}}(v(t)(x)) \psi '(t) {\text {d}}\nu (x) {\text {d}}t\\ &{}\displaystyle = \int _0^T \int _{\Omega _1} j_{\gamma }^*(v(t)(x)) \psi '(t) {\text {d}}\nu (x) {\text {d}}t. \end{array} \end{aligned}$$

Taking

$$\begin{aligned} \widetilde{\eta }_\tau (t)(x)=\frac{1}{\tau }\int _t^{t+\tau } u(s-\tau ) \Psi (s){\text {d}}s, \ \ t\in [0,T], \, x\in \Omega _1, \end{aligned}$$

yields the opposite inequalities so that, in fact,

$$\begin{aligned} \int _0^T \int _{\Omega _1}F(t)(x)u(t)(x){\text {d}}\nu (x)\psi (t) {\text {d}}t \displaystyle = \int _0^T \int _{\Omega _1} j_{\gamma }^*(v(t)(x)){\text {d}}\nu (x)\psi '(t) {\text {d}}t, \end{aligned}$$

i.e.,

$$\begin{aligned} -\frac{d}{{\text {d}}t}\int _{\Omega _1} j_{\gamma }^*(v(t)(x)){\text {d}}\nu (x) =\int _{\Omega _1}F(t)(x)u(t)(x){\text {d}}\nu (x) \ \hbox { in} \mathcal {D}'(]0,T[), \end{aligned}$$

thus, in particular,

$$\begin{aligned} \int _{\Omega _1}j_{\gamma }^*(v){\text {d}}\nu \in W^{1,1}(0,T). \end{aligned}$$
(3.32)

Step 2. Suppose now that, either \(\mathcal {R}_{\gamma ,\beta }^-=-\infty \) and \(\mathcal {R}_{\gamma ,\beta }^+<+\infty \), or \(\mathcal {R}_{\gamma ,\beta }^->-\infty \) and \(\mathcal {R}_{\gamma ,\beta }^+=+\infty \). Recall that we are assuming the hypotheses in (3.21) and that \(v_0^n=v_0\) for every \(n\in {\mathbb {N}}\). Suppose first that \(\mathcal {R}_{\gamma ,\beta }^-=-\infty \) and \(\mathcal {R}_{\gamma ,\beta }^+<+\infty \). Then, for \(k\in \mathbb {N}\), let \(\beta ^k:{\mathbb {R}}\rightarrow {\mathbb {R}}\) be the following maximal monotone graph:

$$\begin{aligned} \beta ^k(r):=\left\{ \begin{array}{l} \beta (r)\quad \hbox {if }r<k,\\ \,[\beta ^0(k),\mathfrak {B}^+]\quad \hbox {if }r=k,\\ \mathfrak {B}^++r-k\quad \hbox {if }r>k. \end{array}\right. \end{aligned}$$

We have that \(\beta ^k\rightarrow \beta \) in the sense of maximal monotone graphs. Indeed, given \(\lambda >0\) and \(s\in {\mathbb {R}}\), there exists \(r\in {\mathbb {R}}\) such that \(s\in r+\lambda \beta (r)\) thus, for \(k>r\), \(s\in r+\lambda \beta (r)= r+\lambda \beta ^k(r)\), i.e., \(r=(I+\lambda \beta )^{-1}(s)=(I+\lambda \beta ^k)^{-1}(s)\).

By Step 1, we know that, since \(\mathcal {R}_{\gamma ,\beta ^k}^-=-\infty \) and \(\mathcal {R}_{\gamma ,\beta ^k}^+=+\infty \), there exists a strong solution \(v_k\in W^{1,1}(0,T;L^{1}(\Omega _1,\nu ))\) of Problem \(\left( DP_{f-\frac{1}{k},v_0}^{ \textbf{a}_p,\gamma ,\beta ^k}\right) \), therefore, there exist \(u_k\in L^p(0,T;L^p(\Omega ,\nu ))\) and \(w_k\in L^{p'}(0,T;L^{p'}(\Omega _2,\nu ))\) such that

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle (v_k)_t(t)(x) - \int _{\Omega } \textbf{a}_p(x,y,u_k(t)(y)-u_k(t)(x)) dm_x(y)=f(t)(x)-\frac{1}{k}, &{}x\in \Omega _1,\ 0<t<T, \\ \displaystyle w_k(t)(x)- \int _{\Omega } \textbf{a}_p(x,y,u_k(t)(y)-u_k(t)(x)) dm_x(y) =0, &{}x\in \Omega _2, \ 0<t<T,\end{array} \right. \nonumber \\ \end{aligned}$$
(3.33)

with \(v_k\in \gamma (u_k)\) \(\nu \)-a.e. in \(\Omega _1\) and \(w_k\in \beta ^k(u_k)\) \(\nu \)-a.e. in \(\Omega _2\). Let us see that

$$\begin{aligned} u_k\le u_{k+1}, \ \nu \hbox {-a.e. in } \Omega , \, k\in {\mathbb {N}}, \end{aligned}$$
(3.34)

and

$$\begin{aligned} v_k\le v_{k+1}, \ \nu \hbox {-a.e. in } \Omega _1, \, k\in {\mathbb {N}}. \end{aligned}$$
(3.35)

Going back to the construction of the mild solution, in this case of \(\left( DP_{f-\frac{1}{k},v_0}^{ \textbf{a}_p,\gamma ,\beta ^k}\right) \), for each step \(n\in {\mathbb {N}}\) and for each \(i\in \{1,\ldots , n\}\), we have that there exists \(u_{k,i}^n\in L^p(\Omega ,\nu )\), \(v_{k,i}^n\in L^{p'}(\Omega _1,\nu )\) and \(w_{k,i}^n\in L^{p'}(\Omega _2,\nu )\) such that

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle v_{k,i}^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_{k,i}^n(y)-u_{k,i}^n(x))dm_x(y)=\frac{T}{n} \left( f_i^n(x)-\frac{1}{k}\right) +v_{k,i-1}^n(x), \ x\in \Omega _1\\ \displaystyle w_{k,i}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k,i}^n(y)-u_{k,i}^n(x))dm_x(y)=0, \ x\in \Omega _2, \end{array}\right. \end{aligned}$$

with \(v_{k,i}^n\in \gamma (u_{k,i}^n)\) \(\nu \)-a.e. in \(\Omega _1\) and \(w_{k,i}^n\in \beta ^k(u_{k,i}^n)\) \(\nu \)-a.e. in \(\Omega _2\). Let

$$\begin{aligned} z_{k,i}^n:=\left\{ \begin{array}{l} w_{k+1,i}^n\quad \hbox {if }u_{k+1,i}^n< k,\\ \mathfrak {B}^+\quad \hbox {if }u_{k+1,i}^n=k,\\ \beta ^k(u_{k+1,i}^n)\quad \hbox {if } u_{k+1,i}^n> k, \end{array} \right. \end{aligned}$$

for \(n\in {\mathbb {N}}\) and \(i\in \{1,\ldots ,n\}\) (observe that \(\beta ^k(r)\) is single-valued for \(r>k\) and coincides with \(\beta ^{k+1}(r)=\beta (r)\) for \(r<k\)). It is clear that \(z_{k,i}^n\in \beta ^k(u_{k+1,i}^n)\) and, since \(\beta ^k\ge \beta ^{k+1}\), \(z_{k,i}^n\ge w_{k+1,i}^n\). Then,

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle v_{k+1,1}^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_{k+1,1}^n(y)-u_{k+1,1}^n(x))dm_x(y)=\frac{T}{n} \left( f_1^n(x)-\frac{1}{k+1}\right) +v_{0}(x)\\ \displaystyle > \frac{T}{n} \left( f_1^n(x)-\frac{1}{k}\right) +v_{0}(x)=v_{k,1}^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_{k,1}^n(y)-u_{k,1}^n(x))dm_x(y), \ \ x\in \Omega _1\\ \displaystyle z_{k,1}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k+1,1}^n(y)-u_{k+1,i}^n(x))dm_x(y) \\ \displaystyle \ge w_{k+1,1}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k+1,1}^n(y)-u_{k+1,1}^n(x))dm_x(y) \\ \displaystyle =0= w_{k,1}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k,1}^n(y)-u_{k,1}^n(x))dm_x(y), \ \ x\in \Omega _2, \end{array}\right. \end{aligned}$$

for \(n\in {\mathbb {N}}\). Hence, by the maximum principle (Theorem 2.6),

$$\begin{aligned} v_{k,1}^n\le v_{k+1,1}^n \ \ \hbox { and } \ \ u_{k,1}^n\le u_{k+1,1}^n \ \ \nu \hbox {-a.e.} \end{aligned}$$

Proceeding in the same way, we get that

$$\begin{aligned} v_{k,i}^n\le v_{k+1,i}^n\ \ \hbox { and } \ \ u_{k,i}^n\le u_{k+1,i}^n \ \ \nu \hbox {-a.e.} \end{aligned}$$

for each \(n\in {\mathbb {N}}\) and \(i\in \{1,\ldots , n\}\). From here we get (3.34) and (3.35).

Since \(\gamma ^{-1}(r)= \partial j_{\gamma }^*(r)\) and \(u_k(t)\in \gamma ^{-1}(v_k(t))\) \(\nu \)-a.e. in \(\Omega _1\), we have

$$\begin{aligned}{} & {} \int _{\Omega _1}(v_k(t-\tau )(x)-v_k(t)(x))u_k(t)(x){\text {d}}\nu (x) \\{} & {} \quad \le \int _{\Omega _1}j_\gamma ^*(v_k(t-\tau )(x))-j_\gamma ^*(v_k(t)(x)){\text {d}}\nu (x). \end{aligned}$$

Integrating this equation over [0, T], dividing by \(\tau \), letting \(\tau \rightarrow 0^+\) and recalling that, by (3.32), \(\displaystyle \int _{\Omega _1}j_\gamma ^*(v_k){\text {d}} \nu \in W^{1,1}(0,T)\), we get

$$\begin{aligned} \begin{array}{rl} \displaystyle -\int _0^T\int _{\Omega _1}(v_k)_t(t)(x)u_k(t)(x){\text {d}}\nu (x){\text {d}}t&{} \displaystyle \le \int _{\Omega _1}j_\gamma ^*(v(0)(x))-j_\gamma ^*(v_k(T)(x)){\text {d}}\nu (x)\\ &{}\displaystyle \le \int _{\Omega _1}j_\gamma ^*(v(0)(x)){\text {d}}\nu (x). \end{array} \end{aligned}$$

Therefore, multiplying (3.33) by \(u_k\) and integrating with respect to \(\nu \), we get

$$\begin{aligned} \begin{array}{l} \displaystyle \frac{1}{2}\int _0^T\int _\Omega \int _{\Omega } \textbf{a}_p(x,y,u_k(t,y)-u_k(t)(x))(u_k(t)(y)-u_k(t)(x)) dm_x(y){\text {d}}\nu (x){\text {d}}t\\ \displaystyle \le \int _0^T\int _{\Omega _1}\left( f(t)(x)-\frac{1}{k}\right) u_k(t)(x)d \nu (x){\text {d}}t+\int _{\Omega _1}j_\gamma ^*(v(0)(x)){\text {d}}\nu (x). \end{array} \end{aligned}$$

Now, working as in the previous step, since \(\Gamma ^{+}<\infty \), we get that \( \displaystyle \left\{ \Vert u_k \Vert _{L^{p}(0,T;L^p(\Omega ,\nu ))}^{p}\right\} _k\) is bounded. Then, by the monotone convergence theorem, we get that there exists \(u\in L^p(0,T;L^p(\Omega ,\nu ))\) such that \(u_k{\mathop {\longrightarrow }\limits ^{k}}u\) in \(L^p(0,T;L^p(\Omega ,\nu ))\). From this we get, by [15, Lemma G], that \(v(t)(x)\in \gamma (u(t)(x))\) for a.e. \(t\in [0,T]\) and \(\nu \)-a.e. \(x\in \Omega _1\).

Therefore, from (3.33) and Lemma 2.1 (note that, by the monotonicity of \(\{u_k\}\), \(|u_k|\le \max \{|u_1|,|u|\}\in L^p(\Omega ,\nu )\)), we get that \((v_k)_t\) converges strongly in \(L^{p'}(0,T;L^{p'}(\Omega _1,\nu ))\) and \(w_k\) converges strongly in \(W^{1,1}(0,T;L^{1}(\Omega _1,\nu ))\). In particular, \(v\in W^{1,1} (0,T;L^1(\Omega _1,\nu ))\), \(w(t)(x)\in \beta (u(t)(x))\) for a.e. \(t\in [0,T]\) and \(\nu \)-a.e. \(x\in \Omega _2\), and

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle v_t(t)(x) - \int _{\Omega } \textbf{a}_p(x,y,u(t)(y)-u(t)(x)) dm_x(y)=f(t)(x), &{}x\in \Omega _1,\ 0<t<T, \\ \displaystyle w(t)(x)- \int _{\Omega } \textbf{a}_p(x,y,u(t)(y)-u(t)(x)) dm_x(y) =0, &{}x\in \Omega _2, \ 0<t<T.\end{array} \right. \end{aligned}$$

The case \(\mathcal {R}_{\gamma ,\beta }^->-\infty \) and \(\mathcal {R}_{\gamma ,\beta }^+=+\infty \) follows similarly by taking

$$\begin{aligned} {\widetilde{\beta }}^k:=\left\{ \begin{array}{l} \mathfrak {B}^-+r+k\quad \hbox {if }r<-k,\\ \,[\mathfrak {B}^-,\beta ^0(-k)]\quad \hbox {if }r=-k,\\ \beta (r)\quad \hbox {if }r>-k. \end{array}\right. \end{aligned}$$

instead of \(\beta ^k\), \(k\in {\mathbb {N}}\).

Step 3. Finally, assume that both \(\mathcal {R}_{\gamma ,\beta }^-\) and \(\mathcal {R}_{\gamma ,\beta }^+\) are finite. We define, for \(k\in \mathbb {N}\),

$$\begin{aligned} {\widetilde{\beta }}^k:=\left\{ \begin{array}{l} \mathfrak {B}^-+r+k\quad \hbox {if }r<-k,\\ \,[\mathfrak {B}^-,\beta ^0(-k)]\quad \hbox {if }r=-k,\\ \beta (r)\quad \hbox {if }r>-k. \end{array}\right. \end{aligned}$$

By the previous step, we have that for k large enough such that \(f+\frac{1}{k}\) satisfies

$$\begin{aligned} \int _{\Omega _1}v_0^+{\text {d}}\nu +\int _0^T\int _{\Omega _1}\left( f(s)^++\frac{1}{k}\right) {\text {d}}\nu {\text {d}}s< \nu (\Omega _1)\Gamma ^+, \end{aligned}$$

there exists a strong solution \(v_k\in W^{1,1}(0,T;L^{1}(\Omega _1,\nu ))\) of Problem \(\left( DP_{f+\frac{1}{k},v_0}^{ \textbf{a}_p,\gamma ,{\widetilde{\beta }}^k}\right) \), i.e., there exist \(u_k\in L^p(0,T;L^p(\Omega ,\nu ))\) and \(w_k\in L^{p'}(0,T;L^{p'}(\Omega _2,\nu ))\) such that

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle (v_k)_t(t)(x) - \int _{\Omega } \textbf{a}_p(x,y,u_k(t)(y)-u_k(t)(x)) dm_x(y)=f(t)(x)+\frac{1}{k}, &{}x\in \Omega _1,\ 0<t<T, \\ \displaystyle w_k(t)(x)- \int _{\Omega } \textbf{a}_p(x,y,u_k(t)(y)-u_k(t)(x)) dm_x(y) =0, &{}x\in \Omega _2, \ 0<t<T,\end{array} \right. \end{aligned}$$

with \(v_k\in \gamma (u_k)\) \(\nu \)-a.e. in \(\Omega _1\) and \(w_k\in {\tilde{\beta }}^k(u_k)\) \(\nu \)-a.e. in \(\Omega _2\).

Going back to the construction of the mild solution, in this case of \(\left( DP_{f+\frac{1}{k},v_0}^{ \textbf{a}_p,\gamma ,{\widetilde{\beta }}^k}\right) \), for each step \(n\in {\mathbb {N}}\) and for each \(i\in \{1,\ldots , n\}\), we have that there exists \(u_{k,i}^n\in L^p(\Omega ,\nu )\), \(v_{k,i}^n\in L^{p'}(\Omega _1,\nu )\) and \(w_{k,i}^n\in L^{p'}(\Omega _2,\nu )\) such that

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle v_{k,i}^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_{k,i}^n(y)-u_{k,i}^n(x))dm_x(y)=\frac{T}{n} \left( f_i^n(x)+\frac{1}{k}\right) +v_{k,i-1}^n(x), &{}x\in \Omega _1\\ \displaystyle w_{k,i}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k,i}^n(y)-u_{k,i}^n(x))dm_x(y)=0, &{}x\in \Omega _2, \end{array}\right. \end{aligned}$$

where \(v_{k,i}^n\in \gamma (u_{k,i}^n)\) \(\nu \)-a.e. in \(\Omega _1\) and \(w_{k,i}^n\in {\widetilde{\beta }}^k(u_{k,i}^n)\) \(\nu \)-a.e. in \(\Omega _2\). Let

$$\begin{aligned} z_{k,i}^n:=\left\{ \begin{array}{l} w_{k+1,i}^n\quad \hbox {if }u_{k+1,i}^n>- k,\\ \mathfrak {B}^-\quad \hbox {if }u_{k+1,i}^n=-k,\\ {\widetilde{\beta }}^k(u_{k+1,i}^n)\quad \hbox {if } u_{k+1,i}^n<- k, \end{array} \right. \end{aligned}$$

for \(n\in {\mathbb {N}}\) and \(i\in \{1,\ldots ,n\}\) (observe that \({\widetilde{\beta }}^k(r)\) is single-valued for \(r<-k\) and coincides with \({\widetilde{\beta }}^{k+1}(r)=\beta (r)\) for \(r>-k\)). It is clear that \(z_{k,i}^n\in {\widetilde{\beta }}^k(u_{k+1,i}^n)\) and, since \({\widetilde{\beta }}^k\le {\widetilde{\beta }}^{k+1}\), we have that \(z_{k,i}^n\le w_{k+1,i}^n\), \(i\in \{1,\ldots ,n\}\). Then,

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle v_{k+1,1}^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_{k+1,1}^n(y)-u_{k+1,1}^n(x))dm_x(y)=\frac{T}{n} \left( f_1^n(x)+\frac{1}{k+1}\right) +v_{0}^n(x)\\ \displaystyle < \frac{T}{n} \left( f_1^n(x)+\frac{1}{k}\right) +v_{0}^n(x)=v_{k,1}^n(x)-\frac{T}{n}\int _\Omega \textbf{a}_p(x,y,u_{k,1}^n(y)-u_{k,1}^n(x))dm_x(y), \ \ x\in \Omega _1\\ \displaystyle z_{k,1}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k+1,1}^n(y)-u_{k+1,i}^n(x))dm_x(y) \\ \displaystyle \le w_{k+1,1}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k+1,1}^n(y)-u_{k+1,1}^n(x))dm_x(y) \\ \displaystyle =0= w_{k,1}^n(x)- \int _\Omega \textbf{a}_p(x,y,u_{k,1}^n(y)-u_{k,1}^n(x))dm_x(y), \ \ x\in \Omega _2, \end{array}\right. \end{aligned}$$

for \(n\in {\mathbb {N}}\). Hence, by the maximum principle (Theorem 2.6),

$$\begin{aligned} v_{k,1}^n\ge v_{k+1,1}^n\ \ \hbox { and }\ \ u_{k,1}^n\ge u_{k+1,1}^n \ \ \nu \hbox {-a.e.} \end{aligned}$$

Proceeding in the same way we get that, for \(n\in {\mathbb {N}}\) and \(i\in \{1,\ldots ,n\}\),

$$\begin{aligned} v_{k,i}^n\ge v_{k+1,i}^n\ \ \hbox { and }\ \ u_{k,i}^n\ge u_{k+1,i}^n \ \ \nu \hbox {-a.e.} \end{aligned}$$

Therefore,

$$\begin{aligned} u_k\ge u_{k+1}, \ \nu \hbox {-a.e. in } \Omega , \, k\in {\mathbb {N}}, \end{aligned}$$

and

$$\begin{aligned} v_k\ge v_{k+1}, \ \nu \hbox {-a.e. in } \Omega _1, \, k\in {\mathbb {N}}. \end{aligned}$$

We can now conclude, as in the previous step, that

$$\begin{aligned} \int _0^T\Vert u_k^-(t)\Vert _{L^{p}(\Omega _1,\nu )}{\text {d}}t\le K_5\left( \int _0^T\left( \int _{\Omega _1} \int _{\Omega _1} |u_k^-(t)(y)-u_k^-(t)(x)|^p dm_x(y){\text {d}}\nu (x)\right) ^\frac{1}{p} {\text {d}}t+1 \right) \end{aligned}$$

for some constant \( K_5>0\). Moreover, by the monotonicity of \(\{u_k\}\), we get that \(\displaystyle \Big \{\int _0^T\Vert u_k^+(t)\Vert _{L^{p}(\Omega _1,\nu )}{\text {d}}t\Big \}_k\) is bounded. From this point, we can finish the proof as in the previous step. \(\square \)