1 Introduction

Nowadays, well-posedness has been drawing great attention in the field of optimization problems and related problems such as variational inequalities, hemivariational inequalities, fixed point problems, equilibrium problems, and inclusion problems (see [1, 5, 9, 11, 17, 19, 21, 23, 33]). The classical concept of well-posedness for a global minimization problem was first introduced by Tikhonov [35], which required the existence and uniqueness of a solution to the global minimization problem and the convergence of every minimizing sequence toward the unique solution. Thereafter, the concept of well-posedness has been generalized to variational inequalities. The initial notion of well-posedness for variational inequality is due to Lucchetti and Patrone [28]. Fang [13, 14] generalized two kinds of well-posedness for a mixed variational inequality problem in a Banach space. For further results on the well-posedness of variational inequalities, we refer to [2, 4, 1214, 16, 22, 27, 28] and the references therein.

As an important and useful generalization of variational inequality, hemivariational inequality, which was first studied by Panagiotopoulos [32], has a great development in recent years by several works [6, 29, 31]. Many authors are interested in generalizing the concept of well-posedness to hemivariational inequalities. In 1995, Goeleven and Mentagui [15] generalized the concept of the well-posedness to a hemivariational inequality and presented some basic results concerning the well-posed hemivariational inequality. Recently, using the concept of approximating sequence, Xiao et al. [37, 38] introduced a concept of well-posedness for a hemivariational inequality and a variational-hemivariational inequality. Ceng, Lur, and Wen [3] considered an extension of well-posedness for a minimization problem to a class of generalized variational-hemivariational inequalities with perturbations in reflexive Banach spaces. For more recent works on the well-posedness for variational-hemivariational inequalities, we refer to [3, 15, 18, 19, 26, 37, 38] and the references therein.

In the last years, many authors studied the existence results for some types of hemivariational inequalities involving set-valued operators [34, 36, 39]. In 2011, Zhang and He [39] studied a kind of hemivariational inequalities of Hartman–Stampacchia type by introducing the concept of stable quasimonotonicity. They supposed that the constraint set is a bounded (or unbounded), closed, and convex subset in a reflexive Banach space. The authors gave sufficient conditions for the existence and boundedness of solutions. In 2013, Tang and Huang [34] generalized the result of [39] by introducing the concept of stable ϕ-quasimonotonicity and obtained some existence theorems when the constrained set is nonempty, bounded (or unbounded), closed, and convex in a reflexive Banach space. Hereafter, Wangkeeree and Preechasilp [36] generalized the results of [34] and [39] by introducing the concept of stable f-quasimonotonicity. Very recently, Liu and Zeng obtained some existence results for a class of hemivariational inequalities involving the stable \((g,f,\alpha)\)-quasimonotonicity [25], a result on the well-posedness for mixed quasivariational hemivariational inequalities [26], and some existence results for a class of quasimixed equilibrium problems involving the \((f,g,h)\)-quasimonotonicity [24].

Let K be a nonempty, closed, and convex subset of a real Banach space X with its dual \(X^{*}\), and let \(F:K\rightarrow P(X^{*})\) be a set-valued operator, where \(P(X^{*})\) is the set of all nonempty subsets of \(X^{*}\). Let \(T:K\rightarrow X^{*}\) be a perturbation, and let \(f\in X^{*}\) be a given element. Let \(g:K\times K\rightarrow \overline{R}:=R\cup \{\pm \infty \}\) be a function such that \(\mathcal{D}(g)=\{u\in K:g(u,v)\neq -\infty, \forall v\in K\}\neq \emptyset \). Let \(J:X\to R\) be a locally Lipschitz function, and let \(J^{\circ }(u,v)\) denote the generalized directional derivative in the sense of Clarke of a locally Lipschitz functional \(J:X\rightarrow R\) at u in the direction v. In this paper, we discuss the following generalized variational-hemivariational inequality (GVHVI):

Find \(u\in K\) such that, for some \(u^{*}\in F(u)\),

$$\begin{aligned} \bigl\langle u^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq 0, \quad \forall v\in K. \end{aligned}$$

Now, let us consider some particular cases of GVHVI.

  1. (a)

    If \(T\equiv 0\), \(f\equiv 0\), and \(g\equiv 0\), then GVHVI is reduced to the following form:

    Find \(u\in K\) and \(u^{*}\in F(u)\) such that

    $$\begin{aligned} \bigl\langle u^{*},v-u\bigr\rangle +J^{\circ }(u;v-u)\geq 0, \quad \forall v\in K. \end{aligned}$$

    The existence of solutions to this inequality was recently studied by Zhang and He [39].

  2. (b)

    If \(T\equiv 0\) and \(f\equiv 0\), and \(g(u,v)=\phi (v)-\phi (u)\) for all \(u,v\in K\), then GVHVI is reduced to the following form:

    Find \(u\in K\) and \(u^{*}\in F(u)\) such that

    $$\begin{aligned} \bigl\langle u^{*},v-u\bigr\rangle +\phi (v)-\phi (u)+J^{\circ }(u;v-u) \geq 0, \quad \forall v\in K. \end{aligned}$$

    The existence of solutions to this inequality was studied by Tang and Huang [34].

  3. (c)

    If \(T\equiv 0\) and \(f\equiv 0\), then GVHVI is reduced to the following form:

    Find \(u\in K\) and \(u^{*}\in F(u)\) such that

    $$\begin{aligned} \bigl\langle u^{*},v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u)\geq 0, \quad \forall v\in K. \end{aligned}$$

The existence of solutions to this inequality was studied by Wangkeeree and Preechasilp [36].

Inspired by previous works, we study the well-posedness for GVHVI, which generalizes many known works. Under relatively weak conditions, we establish some equivalence results and some metric characterizations for the strong and weak α-well-posed GVHVI in the generalized sense. In particular, we present equivalence results on weak α-well-posedness for GVHVI, which were considered by few authors.

This paper is organized as follows. In Sect. 2, we recall some basic preliminaries of single-valued and set-valued mappings, metric concepts, Clarke’s generalized directional derivative, and some classes of well-posedness for GVHVI. In Sect. 3, we show some equivalence results for the well-posedness for GVHVI and some corresponding metric characterizations. Theorems 3.3, 3.5, and 3.6 are the main results in this section. In the last section, we also present the well-posedness for a class of generalized mixed equilibrium problems.

2 Preliminaries

Let R, \(R_{+}\), and N be the sets of real numbers, nonnegative real numbers, and natural numbers, respectively. Let X be a real Banach space with norm \(\|\cdot \|_{X}\). Denote by \(X^{*}\) its dual space and by \(\langle \cdot,\cdot \rangle_{X}\) the duality pairing between \(X^{*}\) and X. Let \(X_{w}\) be the Banach space X with weak topology.

Definition 2.1

Let K be a nonempty subset of X. A function \(f:K\rightarrow R\) is said to be

  1. (i)

    convex on K if for all finite subsets \(\{u_{1},\ldots,u_{n} \}\subset K\) and \(\{\lambda_{1},\ldots,\lambda_{n}\}\subset R_{+}\) such that \(\sum_{i=1}^{n}\lambda_{i}=1\) and \(\sum_{i=1}^{n}\lambda_{i}u _{i}\in K\), we have

    $$\begin{aligned} f\Biggl(\sum_{i=1}^{n}\lambda_{i}u_{i} \Biggr)\leq \sum_{i=1}^{n} \lambda_{i}f(u _{i}); \end{aligned}$$
  2. (ii)

    (weakly) upper semicontinuous (u.s.c. for short) at u if for any sequence \(\{u_{n}\}_{n\geq 1}\subset K\) with (\(u_{n}\rightharpoonup u\)) \(u_{n}\rightarrow u\), we have

    $$\begin{aligned} \limsup_{n\rightarrow \infty }f(u_{n})\leq f(u). \end{aligned}$$
  3. (iii)

    (weakly) lower semicontinuous (l.s.c. for short) at u, if for any sequence \(\{u_{n}\}_{n\geq 1}\subset K\) with (\(u_{n}\rightharpoonup u\)) \(u_{n}\rightarrow u\), we have

    $$\begin{aligned} \liminf_{n\rightarrow \infty }f(u_{n})\geq f(u). \end{aligned}$$

    The function f is said to be (weakly) u.s.c. (l.s.c.) on K if f is (weakly) u.s.c. (l.s.c.) at all \(u\in K\).

Definition 2.2

([20])

Let K be a nonempty subset of X. An operator \(\beta:K\rightarrow X\) is said to be affine if for any \(u_{i}\in K\) (\(i=1,2,\ldots,n\)) and \(\lambda_{i}\in [0,1]\) with \(\sum_{i=1} ^{n}\lambda_{i}=1\), we have

$$\begin{aligned} \beta \Biggl(\sum_{i=1}^{n} \lambda_{i}v_{i}\Biggr)=\sum_{i=1}^{n} \lambda_{i}\beta (u_{i}). \end{aligned}$$

Definition 2.3

A set-valued operator \(F:K\rightarrow P(X^{*})\) is said to be

  1. (i)

    lower semicontinuous (l.s.c.) at \(u_{0}\) if for any \(u_{0}^{*}\in F(u_{0})\) and sequence \(\{u_{n}\}_{n\geq 1}\subset K\) with \(u_{n}\rightarrow u_{0}\), there exists a sequence \(u_{n}^{*}\in F(u _{n})\) that converges to \(u_{0}^{*}\).

  2. (ii)

    lower hemicontinuous (l.h.c.) if the restriction of F to every line segment of K is lower semicontinuous with respect to the weak topology in \(X^{*}\).

Definition 2.4

A set-valued operator \(F:K\rightarrow P(X^{*})\) is said to be monotone if for all \(u,v\in K\),

$$\begin{aligned} \bigl\langle v^{*}-u^{*},v-u\bigr\rangle \geq 0, \quad \forall u^{*}\in F(u),\forall v^{*}\in F(v). \end{aligned}$$

Definition 2.5

Let S be a nonempty subset of X. The measure μ of noncompactness for the set S is defined by

$$\begin{aligned} \mu (S):=\inf \Biggl\{ \epsilon >0:S=\bigcup_{i=1} ^{n} S_{i}, \operatorname{diam} \vert S_{i} \vert < \epsilon,i=1,2,\ldots,n\Biggr\} , \end{aligned}$$

where diam\(|S_{i}|\) is the diameter of the set \(S_{i}\).

Now, let us recall the definitions of the Clarke generalized directional derivative and generalized gradient for a locally Lipschitz function \(\varphi:X\rightarrow R\) (see [6, 10]). The Clarke generalized directional derivative \(\varphi^{0}(u;v)\) of φ at the point \(u\in X\) in the direction \(v\in X\) is defined as

$$ \varphi^{0}(u;v):=\limsup_{\lambda \rightarrow 0^{+},\zeta \rightarrow u}\frac{\varphi (\zeta +\lambda v)-\varphi (\zeta)}{\lambda }. $$

The Clarke subdifferential or generalized gradient of φ at \(u\in X\), denoted by \(\partial \varphi (u)\), is the subset of \(X^{*}\) given by

$$ \partial \varphi (u):=\bigl\{ u^{*}\in X^{*}: \varphi^{0}(u;v)\geq \bigl\langle u^{*},v\bigr\rangle _{X}, \forall v\in X\bigr\} . $$

Lemma 2.6

([6], Proposition 2.1.1)

Let \(\varphi:X\rightarrow R\) be locally Lipschitz of rank \(L_{u}>0\) near u. Then

  1. (i)

    \(\varphi^{0}(u;v)\) is u.s.c. as a function of \((u,v)\) and, as a function of v alone, is Lipschitz of rank \(L_{u}\) near u on X and satisfies

    $$ \bigl\vert \varphi^{0}(u;v) \bigr\vert \leq L_{u} \Vert v \Vert _{X}; $$
  2. (ii)

    the gradient \(\partial \varphi (u)\) is a nonempty, convex, and weakly compact subset of \(X^{*}\) bounded by a Lipschitz constant \(L_{u}\) near x;

  3. (iii)

    for every \(v\in X\), we have

    $$ \varphi^{0}(u;v)= \max \bigl\{ \bigl\langle u^{*},v\bigr\rangle |u^{*}\in \partial \varphi (u)\bigr\} . $$

We end this section with the notions of several classes of α-approximating sequences and α-well-posedness for GVHVI. Let \(\alpha:X\to R_{+}\) be a functional.

Definition 2.7

A sequence \(\{u_{n}\}\) in K is an α-approximating sequence for GVHVI if there exist \(\{u^{*}_{n}\}\) in \(X^{*}\) with \(u^{*}_{n}\in F(u _{n})\) and a nonnegative sequence \(\{\epsilon_{n}\}\) with \(\epsilon _{n}\rightarrow 0\) as \(n\rightarrow \infty \) such that, for every \(n\in N\),

$$\begin{aligned} \bigl\langle u_{n}^{*}+ Tu_{n}-f,v-u_{n} \bigr\rangle +g(u_{n},v)+J^{\circ }(u _{n};v-u_{n}) \geq -\epsilon_{n}\alpha (v-u_{n}),\quad \forall v\in K. \end{aligned}$$

In particular, if \(\alpha (\cdot) = \|\cdot \|_{X}\), then \(\{u_{n}\}\) is said to be an approximating sequence for GVHVI.

Definition 2.8

GVHVI is said to be strongly (respectively, weakly) α-well-posed if it has a unique solution u and every α-approximating sequence \(\{u_{n}\}\) strongly (respectively, weakly) converges to u. In particular, if \(\alpha (\cdot) = \|\cdot \|_{X}\), then GVHVI is said to be strongly (respectively, weakly) well-posed.

Definition 2.9

GVHVI is said to be strongly (respectively, weakly) α-well-posed in the generalized sense if the solution set Γ of GVHVI is nonempty and every α-approximating sequence \(\{u_{n}\}\) has a subsequence that strongly (respectively, weakly) converges to some point of Γ. In particular, if \(\alpha (\cdot) = \|\cdot \|_{X}\), then GVHVI is said to be strongly (respectively, weakly) well-posed in the generalized sense.

Remark 2.10

Strong α-well-posedness (in the generalized sense) implies weak α-well-posedness (in the generalized sense), but the converse is not true in general.

3 The characterizations of well-posedness for GVHVI

In this section, we establish metric characterizations and derive some conditions under which GVHVI is strongly (weakly) α-well-posed.

For any \(\epsilon >0\), we define the following two sets:

$$\begin{aligned} \Omega_{\alpha }(\epsilon) =&\bigl\{ u\in K: \exists u^{*}\in F(u)\mbox{ such that }\bigl\langle u^{*}+ Tu-f,v-u\bigr\rangle +g(u,v) \\ & {} +J^{\circ }(u;v-u)\geq -\epsilon \alpha (v-u), \forall v \in K \bigr\} \end{aligned}$$

and

$$\begin{aligned} \Phi_{\alpha }(\epsilon) =&\bigl\{ u\in K:\bigl\langle v^{*}+Tu-f,v-u \bigr\rangle +g(u,v)+J ^{\circ }(u;v-u) \\ & {}\geq -\epsilon \alpha (v-u), \forall v\in K,\forall v^{*} \in F(v) \bigr\} . \end{aligned}$$

Denote by Γ the set of solutions to GVHVI. It is clear that \(\Gamma =\Omega_{0}(\epsilon)\).

Lemma 3.1

Assume that:

  1. (i)

    K is a nonempty closed subset of a real Banach space X;

  2. (ii)

    \(T:K \rightarrow X^{*}_{w}\) is continuous;

  3. (iii)

    \(g:K\times K\rightarrow R\) is u.s.c. with respect to the first variable;

  4. (iv)

    \(\alpha:X\to R_{+}\) is such that \(\liminf_{n\rightarrow \infty }\alpha (v_{n})\le \alpha (v)\) whenever \(v_{n}\rightarrow v\).

Then, for every \(\epsilon >0\), the set \(\Phi_{\alpha }(\epsilon)\) is closed in X.

Proof

Let \(\{u_{n}\}\subset \Phi_{\alpha }(\epsilon)\) be s sequence such that \(u_{n} \rightarrow u\) in X. Then \(u\in K\), and, for all \(v\in K\) and \(v^{*}\in F(v)\),

$$\begin{aligned} \bigl\langle v^{*}+ Tu_{n}-f,v-u_{n}\bigr\rangle +g(u_{n},v)+J^{\circ }(u_{n};v-u _{n})\geq - \epsilon \alpha (v-u_{n}). \end{aligned}$$

By the assumptions and the properties of \(J^{\circ }\) we have

$$\begin{aligned}& \bigl\langle v^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \\& \quad \geq \limsup_{n\rightarrow \infty }\bigl[\bigl\langle v^{*}+ Tu_{n}-f,v-u_{n} \bigr\rangle +g(u_{n},v)+J^{\circ }(u_{n};v-u_{n}) \bigr] \\& \quad \geq \limsup_{n\rightarrow \infty }-\epsilon \alpha (v-u_{n}) \\& \quad =-\epsilon \liminf_{n\rightarrow \infty }\alpha (v-u_{n}) \\& \quad \ge -\epsilon \alpha (v-u), \end{aligned}$$

and hence

$$\begin{aligned} \bigl\langle v^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u)- \epsilon \alpha (v-u), \quad \forall v\in K,\forall v^{*}\in F(v), \end{aligned}$$

which shows that \(u\in \Phi_{\alpha }(\epsilon)\). □

Lemma 3.2

Assume that:

  1. (i)

    K is a nonempty convex subset of a real Banach space X;

  2. (ii)

    \(F:K\rightarrow P(X^{*})\) is l.h.c. and monotone;

  3. (iii)

    \(g:K\times K\rightarrow R\) is convex with respect to the second variable;

  4. (iv)

    \(\alpha:X\to R_{+}\) is convex with \(\alpha (tv)=t\alpha (v)\) for all \(t\ge 0\) and \(v\in X\).

Then \(\Omega_{\alpha }(\epsilon)=\Phi_{\alpha }(\epsilon)\) for all \(\epsilon >0\).

Proof

We first show that \(\Omega_{\alpha }(\epsilon)\subset \Phi_{\alpha }( \epsilon)\). Indeed, take arbitrary \(u\in \Omega_{\alpha }(\epsilon)\). Then there exists \(u^{*}\in F(u)\) such that

$$\begin{aligned} \bigl\langle u^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq -\epsilon \alpha (v-u), \quad \forall v\in K. \end{aligned}$$

According to the monotonicity of F, we obtain

$$\begin{aligned} \bigl\langle v^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq -\epsilon \alpha (v-u), \quad \forall v\in K,\forall v^{*}\in F(v), \end{aligned}$$

which means that \(u\in \Phi_{\alpha }(\epsilon)\). Therefore \(\Omega_{\alpha }(\epsilon)\subset \Phi_{\alpha }(\epsilon)\).

Now we show that \(\Phi_{\alpha }(\epsilon)\subset \Omega_{\alpha }( \epsilon)\). Indeed, take arbitrary \(u\in \Phi_{\alpha }(\epsilon)\). Then

$$\begin{aligned} \bigl\langle v^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq -\epsilon \alpha (v-u), \quad \forall v\in K,\forall v^{*}\in F(v). \end{aligned}$$

Since the set K is convex, for any \(v\in K\) and \(\lambda \in [0,1]\), taking \(v_{\lambda }:=\lambda v+(1-\lambda)u\in K\) in this inequality, we have

$$\begin{aligned} \bigl\langle v_{\lambda }^{*}+ Tu-f,v_{\lambda }-u\bigr\rangle +g(u,v_{\lambda })+J^{\circ }(u;v_{\lambda }-u) \geq -\epsilon \alpha (v_{\lambda }-u), \quad \forall v_{\lambda }^{*}\in F(v_{\lambda }). \end{aligned}$$

Then by (iii), (iv), and the properties of \(J^{\circ }\) we obtain

$$\begin{aligned} \bigl\langle v_{\lambda }^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq -\epsilon \alpha (v-u), \quad \forall v_{\lambda }^{*}\in F(v_{\lambda }). \end{aligned}$$
(3.1)

Let \(u^{*}\in F(u)\) be fixed, and let \(v_{\lambda }^{*}\in F(v_{ \lambda })\) be such that \(v_{\lambda }^{*}\rightharpoonup u^{*}\) in \(X^{*}\) (the existence of such a sequence is ensured by the fact that F is l.h.c.). Taking the limit as \(\lambda \rightarrow 0\) in (3.1), we obtain

$$\begin{aligned}& \bigl\langle u^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \\& \quad =\lim_{\lambda \rightarrow 0}\bigl[\bigl\langle v_{\lambda }^{*}+ Tu-f,v-u \bigr\rangle +g(u,v)+J^{\circ }(u;v-u)\bigr] \\& \quad \geq -\epsilon \alpha (v-u), \end{aligned}$$

which implies that \(u\in \Omega_{\alpha }(\epsilon)\). The proof is complete. □

The following result is a consequence of Lemmas 3.1 and 3.2.

Theorem 3.3

Assume that:

  1. (i)

    K is a nonempty closed convex subset of a real Banach space X;

  2. (ii)

    \(F:K\rightarrow P(X^{*})\) is l.h.c. and monotone;

  3. (iii)

    \(T:K\rightarrow X^{*}_{w}\) is continuous;

  4. (iv)

    \(g:K\times K\rightarrow R\) is u.s.c. with respect to the first variable and convex with respect to the second variable;

  5. (v)

    \(\alpha:X\to R_{+}\) is continuous and convex with \(\alpha (tv)=t \alpha (v)\) for all \(t\ge 0\) and \(v\in X\).

Then \(\Omega_{\alpha }(\epsilon)=\Phi_{\alpha }(\epsilon)\) is closed in X for all \(\epsilon >0\). Moreover, \(\Gamma =\Omega_{0}(\epsilon)=\Phi_{0}(\epsilon)\), that is, GVHVI is equivalent to the following problem:

Find \(u\in K\) such that

$$\begin{aligned} \bigl\langle v^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq 0, \quad \forall v\in K,v^{*}\in F(v). \end{aligned}$$

Theorem 3.4

GVHVI is strongly α-well-posed if and only if Γ is nonempty and

$$\begin{aligned} \lim_{\epsilon \rightarrow 0}\operatorname{diam}\bigl( \Omega_{\alpha }( \epsilon)\bigr)=0. \end{aligned}$$

Proof

The proof is similar to that of Theorem 4.3 in [26] by the assumptions of g. □

Theorem 3.5

Assume that all the assumptions of Theorem 3.3 are satisfied. Then GVHVI is strongly α-well-posed if and only if

$$\begin{aligned} \Omega_{\alpha }(\epsilon)\neq \emptyset \quad \forall \epsilon \geq 0 \quad \textit{and}\quad \lim_{\epsilon \rightarrow 0} \operatorname{diam}\bigl( \Omega_{\alpha }(\epsilon)\bigr)=0. \end{aligned}$$
(3.2)

Proof

Suppose that GVHVI is strongly α-well-posed. Then GVHVI has a unique solution \(u\in K\), and thus \(\Gamma \neq \emptyset \). Now, we prove that (3.2) holds. Clearly, \(\Omega_{\alpha }(\epsilon) \supset \Gamma \neq \emptyset \). For the second part of (3.2), arguing by contradiction, let us assume that \(\operatorname{diam}( \Omega_{\alpha }(\epsilon))\) does not tend to 0 as \(\epsilon \rightarrow 0\). Thus for any nonnegative sequence \(\{\epsilon_{n}\}\) with \(\epsilon_{n}\rightarrow 0\) as \(n\rightarrow \infty \), there exists a constant \(\beta >0\) such that, for each \(n\in N\), there exist \(u_{n}^{(1)},u_{n}^{(2)}\in \Omega_{\alpha }(\epsilon_{n})\) satisfying

$$\begin{aligned} \bigl\Vert u_{n}^{(1)}-u_{n}^{(2)} \bigr\Vert >\beta >0. \end{aligned}$$
(3.3)

Since \(u_{n}^{(1)},u_{n}^{(2)}\in \Omega_{\alpha }(\epsilon_{n})\), we know that the sequences \(\{u_{n}^{(1)}\}\) and \(\{u_{n}^{(2)}\}\) are both α-approximating sequences of GVHVI, and thus

$$\begin{aligned} \lim_{n\rightarrow }u_{n}^{(1)}=\lim _{n\rightarrow }u _{n}^{(2)}=u. \end{aligned}$$
(3.4)

From (3.3) and (3.4) we have

$$\begin{aligned} 0< \beta < \bigl\Vert u_{n}^{(1)}-u_{n}^{(2)} \bigr\Vert \leq \bigl\Vert u_{n}^{(1)}-u \bigr\Vert + \bigl\Vert u_{n}^{(2)}-u \bigr\Vert \rightarrow 0, \end{aligned}$$

which is a contradiction.

Conversely, assume that condition (3.2) holds. Let \(\{u_{n}\}\) in K be an α-approximating sequence for GVHVI. Then, there exist \(\{u^{*}_{n}\}\) in \(X^{*}\) with \(u_{n}^{*}\in F(u_{n})\) and a nonnegative sequence \(\{\epsilon_{n}\}\) with \(\epsilon_{n}\rightarrow 0\) as \(n\rightarrow \infty \) such that, for every \(n\in N\),

$$\begin{aligned} \bigl\langle u^{*}_{n}+ Tu_{n}-f,v-u_{n} \bigr\rangle +g(u_{n},v)+J^{\circ }(u _{n};v-u_{n}) \geq -\epsilon_{n}\alpha (v-u_{n}), \quad \forall v\in K, \end{aligned}$$

that is, \(u_{n}\in \Omega_{\alpha }(\epsilon_{n})\) for all \(n\in N\). By condition (3.2) we deduce that the sequence \(\{u_{n}\}\) is a Cauchy sequence, and so \(\{u_{n}\}\) converges strongly to some point \(u\in K\). Let us show that \(u\in K\) is a solution for GVHVI. By the monotonicity of F we obtain that, for every \(n\in N\),

$$\begin{aligned}& \bigl\langle v^{*}+ Tu_{n}-f,v-u_{n}\bigr\rangle +g(u_{n},v)+J^{\circ }(u _{n};v-u_{n}) \\& \quad \geq \bigl\langle u^{*}_{n}+Tu_{n}-f,v-u_{n} \bigr\rangle +g(u_{n},v)+J^{ \circ }(u_{n};v-u_{n}) \\& \quad \geq -\epsilon_{n}\alpha (v-u_{n}), \quad \forall v\in K,v^{*}\in F(v). \end{aligned}$$

By the assumptions we obtain that

$$\begin{aligned}& \bigl\langle v^{*}+Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \\& \quad \geq \limsup_{n\rightarrow \infty }\bigl[\bigl\langle v^{*}+ Tu_{n}-f,v-u _{n}\bigr\rangle +g(u_{n},v)+J^{\circ }(u_{n};v-u_{n}) \bigr] \\& \quad \geq \limsup_{n\rightarrow \infty }-\epsilon_{n}\alpha (v-u _{n}) \\& \quad =\limsup_{n\rightarrow \infty }\alpha \bigl(-\epsilon_{n}(v-u_{n}) \bigr) \\& \quad =0, \end{aligned}$$

which implies that

$$\begin{aligned} \bigl\langle v^{*}+Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq 0, \quad \forall v\in K,\forall v^{*}\in F(v). \end{aligned}$$

It follows from Theorem 3.3 that there exists \(u^{*}\in F(u)\) such that

$$\begin{aligned} \bigl\langle u^{*}+ Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \geq 0, \quad \forall v\in K. \end{aligned}$$

Then \(u\in K\) is a solution of GVHVI.

Finally, we prove that the solution u is unique. If there exists another solution \(u'\in K\), then \(u,u_{1}\in \Omega_{\alpha }(\epsilon)\) for all \(\epsilon >0\), and

$$\begin{aligned} 0< \bigl\Vert u-u' \bigr\Vert \leq \operatorname{diam}\bigl( \Omega_{\alpha }(\epsilon)\bigr)\rightarrow 0\quad \mbox{as }\epsilon \rightarrow 0, \end{aligned}$$

which is a contradiction. This completes the proof. □

Theorem 3.6

Assume that:

  1. (i)

    K is a nonempty closed convex subset of a real reflexive Banach space X;

  2. (ii)

    \(F:K\rightarrow P(X^{*})\) is l.h.c. and monotone;

  3. (iii)

    \(T:K\rightarrow X^{*}\) is compact;

  4. (iv)

    \(g:K\times K\rightarrow R\) is weakly u.s.c. with respect to the first variable and convex with respect to the second variable;

  5. (v)

    \(\limsup_{n\rightarrow \infty }J^{\circ }(u_{n};v-u_{n})\le J^{ \circ }(u;v-u)\) for all \(v\in X\) whenever \(u_{n}\rightharpoonup u\) as \(n\rightarrow \infty \);

  6. (vi)

    \(\alpha:X\to R_{+}\) is a continuous and convex functional with \(\alpha (tv)=t\alpha (v)\) for all \(t\ge 0\) and \(v\in X\).

Then GVHVI is weakly α-well-posed if and only if GVHVI has a unique solution and there exists \(\epsilon_{0}>0\) such that \(\Omega_{\alpha }(\epsilon_{0})\) is nonempty and bounded.

Proof

The necessity is obvious. We now prove the sufficiency. Let \(\{u_{n}\}\) be an α-approximating sequence for GVHVI. Then, there exist \(\{u^{*}_{n}\}\) in \(X^{*}\) with \(u_{n}^{*}\in F(u_{n})\) and a nonnegative sequence \(\{\epsilon_{n}\}\) with \(\epsilon_{n}\rightarrow 0\) as \(n\rightarrow \infty \) such that, for every \(n\in N\),

$$\begin{aligned} \bigl\langle u^{*}_{n}+ Tu_{n}-f,v-u_{n} \bigr\rangle +g(u_{n},v)+J^{\circ }(u _{n};v-u_{n}) \geq -\epsilon_{n}\alpha (v-u_{n}) \end{aligned}$$

for all \(v\in K\). We claim that the sequence \(\{u_{n}\}\) is bounded in X. Indeed, since \(\Omega_{\alpha }(\epsilon_{0})\) is bounded and \(\Omega_{\alpha }(\epsilon)\subset \Omega_{\alpha }(\epsilon_{0})\) for all \(\epsilon \in (0,\varepsilon_{0})\), there exists \(n_{0}\in N\) such that \(\epsilon_{n_{0}}\in (0,\varepsilon_{0})\) and \(u_{n}\in \Omega_{\alpha }(\epsilon_{0})\) for all \(n\ge n_{0}\), which shows that \(\{u_{n}\}\) is bounded in X.

Since the Banach space X is reflexive, we can choose a subsequence of \(\{u_{n}\}\), denoted by \(\{u_{n}\}\) again, such that \(u_{n}\rightharpoonup \overline{u}\) as \(n\rightarrow \infty \) for some \(\overline{u}\in X\). Let us show that \(\overline{u}\in K\) is a solution for GVHVI. Obviously, \(\overline{u}\in K\). By the monotonicity of F we obtain that

$$\begin{aligned}& \bigl\langle v^{*}+ Tu_{n}-f,v-u_{n}\bigr\rangle +g(u_{n},v)+J^{\circ }(u _{n};v-u_{n}) \\& \quad \geq \bigl\langle u^{*}_{n}+Tu_{n}-f,v-u_{n} \bigr\rangle +g(u_{n},v)+J^{ \circ }(u_{n};v-u_{n}) \\& \quad \geq -\epsilon_{n}\alpha (v-u_{n}),\quad \forall v\in K,v^{*}\in F(v), \forall n\in N. \end{aligned}$$

By the assumptions, we obtain that

$$\begin{aligned}& \bigl\langle v^{*}+T\overline{u}-f,v-\overline{u}\bigr\rangle +g( \overline{u},v)+J^{\circ }(\overline{u};v-\overline{u}) \\& \quad \geq \limsup_{n\rightarrow \infty }\bigl[\bigl\langle v^{*}+ Tu_{n}-f,v-u _{n}\bigr\rangle +g(u_{n},v)+J^{\circ }(u_{n};v-u_{n}) \bigr] \\& \quad \geq \limsup_{n\rightarrow \infty }-\epsilon_{n}\alpha (v-u _{n}) \\& \quad =\limsup_{n\rightarrow \infty }\alpha \bigl(-\epsilon_{n}(v-u_{n}) \bigr) \\& \quad =0, \end{aligned}$$

which implies that

$$\begin{aligned} \bigl\langle v^{*}+T\overline{u}-f,v-\overline{u}\bigr\rangle +g( \overline{u},v)+J ^{\circ }(\overline{u};v-\overline{u})\geq 0, \quad \forall v\in K,\forall v^{*}\in F(v). \end{aligned}$$

It follows from Theorem 3.3 that there exists \(\overline{u}^{*}\in F(\overline{u})\) such that

$$\begin{aligned} \bigl\langle u^{*}+ T\,\overline{u}-f,v-\overline{u}\bigr\rangle +g( \overline{u},v)+J ^{\circ }(\overline{u};v-\overline{u})\geq 0, \quad \forall v\in K, \end{aligned}$$

Therefore \(\overline{u}\in K\) is a solution to problem GVHVI, and so we get that GVHVI is weakly α-well-posed by the uniqueness of the solution to problem GVHVI. This completes the proof. □

Remark 3.7

In the theorem, condition (v) can be found in [30], and the condition that there exists \(\epsilon_{0}>0\) such that \(\Omega_{ \alpha }(\epsilon_{0})\) is nonempty and bounded can be replaced by the conditions that K is bounded or that there exists \(n_{0}\in N\) such that, for every \(u\in K\setminus B_{n_{0}}\), there exists \(v\in K\) with \(\|v\|<\|u\|\) such that

$$\begin{aligned} \sup_{u^{*}\in F(u)}\bigl\langle u^{*}+Tu-f,v-u\bigr\rangle +g(u,v)+J^{\circ }(u;v-u) \leq -\frac{1}{n_{0}}. \end{aligned}$$

See [34, 36, 39] for more detail.

Next, we give some equivalence results for the strong α-posedness in the generalized sense.

Theorem 3.8

Assume that all the assumptions of Theorem 3.5 are satisfied. Then GVHVI is strongly α-well-posed in the generalized sense if and only if Γ is nonempty compact and

$$\begin{aligned} \lim_{\epsilon \rightarrow 0}e\bigl(\Omega_{\alpha }( \epsilon),\Gamma\bigr)=0, \end{aligned}$$

where \(e(A,B):=\sup_{a\in A}d(a,B)\) with \(d(a,B):=\inf_{b\in B}\|a-b \|\).

Proof

The proof is similar to that of Theorem 5.1 in [26] by the assumptions of g. □

Theorem 3.9

Assume that all the assumptions of Theorem 3.5 are satisfied. Then GVHVI is strongly α-well-posed in the generalized sense if and only if

$$\begin{aligned} \Omega_{\alpha }(\epsilon)\neq \emptyset, \quad \forall \epsilon >0,\quad \textit{and}\quad \lim_{\epsilon \rightarrow 0}\mu \bigl(\Omega_{\alpha }( \epsilon)\bigr)=0. \end{aligned}$$

Proof

The proof is similar to that of Theorem 3.2 in [3] by the assumptions of g. □

Theorem 3.10

Assume that all the assumptions of Theorem 3.6 are satisfied. Then GVHVI is weakly α-well-posed in the generalized sense if and only if there exists \(\epsilon_{0}>0\) such that \(\Omega_{\alpha }(\epsilon_{0})\) is nonempty and bounded.

Proof

The proof is similar to that of Theorem 3.6 by the assumptions of g. □

4 Well-posedness for GMEP

In this section, we consider the following generalized mixed equilibrium problem (GMEP):

Find \(u\in K\) such that, for some \(u^{*}\in F(u)\),

$$\begin{aligned} \bigl\langle u^{*},\eta (u,v)\bigr\rangle +\langle Tu-f,v-u\rangle +g(u,v)+h(u,v) \geq 0, \quad \forall v\in K, \end{aligned}$$

where \(\eta:K\times K\rightarrow X\) is an operator. The existence of solutions to this problem when \(T\equiv 0\) and \(f\equiv 0\) can be found in [25].

To study GMEP, we introduce the concept of η-monotonicity (see [7, 8]).

Definition 4.1

Let \(F:K\rightarrow P(X^{*})\) be a set-valued operator. F is said to be η-monotone if there exists a function \(\eta:K\times K\rightarrow X\) such that, for all \(u,v\in K\),

$$\begin{aligned} \bigl\langle v^{*}-u^{*},\eta (u,v)\bigr\rangle \geq 0, \quad \forall u^{*}\in F(u),\forall v^{*}\in F(v). \end{aligned}$$
(4.1)

Remark 4.2

If \(\eta (u,v)=v-u\) for all \(u,v\in X\), then (4.1) becomes

$$\begin{aligned} \bigl\langle v^{*}-u^{*},v-u\bigr\rangle \geq 0, \quad \forall u^{*}\in F(u),\forall v^{*}\in F(v), \end{aligned}$$

that is, F is monotone.

For any \(\epsilon >0\), we define the following two sets:

$$\begin{aligned} \Omega_{\eta,\alpha }(\epsilon) =&\bigl\{ u\in K: \exists u^{*}\in F(u)\mbox{ such that }\bigl\langle u^{*},\eta (u,v)\bigr\rangle + \langle Tu-f,v-u\rangle +g(u,v) \\ & {}+h(u,v))\geq -\epsilon \alpha (v-u),\forall v\in K \bigr\} \end{aligned}$$

and

$$\begin{aligned} \Phi_{\eta,\alpha }(\epsilon) =&\bigl\{ u\in K:\bigl\langle v^{*},\eta (u,v) \bigr\rangle +\langle Tu-f,v-u\rangle +g(u,v) \\ & {}+h(u,v)\geq -\epsilon \alpha (v-u), \forall v\in K,\forall v ^{*} \in F(v) \bigr\} . \end{aligned}$$

Denote by \(\Gamma_{\eta }\) the set of solutions to GMEP. It is clear that \(\Gamma =\Omega_{0}(\epsilon)\).

We can obtain similar results.

Theorem 4.3

Assume that all the assumptions of Theorem 3.3 are satisfied and, in addition, \(\eta:K\times K\rightarrow X\) is continuous on \(K\times K\) with \(\eta (u,u)=0\) for any \(u\in K\) and affine with respect to the first variable. Let \(h:K\times K\rightarrow R\) be such that:

  1. (i)

    \(h(u,u)=0\) for all \(u\in X\),

  2. (ii)

    for all \(v\in K\), \(h(\cdot,v)\) is u.s.c.,

  3. (iii)

    for all \(u\in K\), \(h(u,\cdot)\) is convex.

Then \(\Omega_{\eta,\alpha }(\epsilon)=\Phi_{\eta,\alpha }(\epsilon)\) is closed in X for all \(\epsilon >0\). Moreover, \(\Gamma_{\eta }= \Omega_{\eta,0}(\epsilon)=\Phi_{\eta,0}(\epsilon)\), that is, GMEP is equivalent to the following problem:

Find \(u\in K\) such that

$$\begin{aligned} \bigl\langle v^{*}+ Tu-f,\eta (u,v)\bigr\rangle +g(u,v)+h(u,v)\geq 0, \quad \forall v\in K,v^{*}\in F(v). \end{aligned}$$

Theorem 4.4

Assume that all the assumptions of Theorem 3.5 are satisfied and, in addition, \(\eta:K\times K\rightarrow X\) is continuous on \(K\times K\) with \(\eta (u,u)=0\) for any \(u\in K\) and affine with respect to the first variable. Let \(h:K\times K\rightarrow R\) be such that:

  1. (i)

    \(h(u,u)=0\) for all \(u\in X\),

  2. (ii)

    for all \(v\in K\), \(h(\cdot,v)\) is u.s.c.,

  3. (iii)

    for all \(u\in K\), \(h(u,\cdot)\) is convex.

Then GMVHVI is strongly α-well-posed if and only if

$$\begin{aligned} \Omega_{\eta,\alpha }(\epsilon)\neq \emptyset,\quad \forall \epsilon \geq 0, \quad \textit{and}\quad \lim_{\epsilon \rightarrow 0} \operatorname{diam}\bigl(\Omega_{\eta,\alpha }( \epsilon)\bigr)=0. \end{aligned}$$

Theorem 4.5

Assume that all the assumptions of Theorem 3.6 are satisfied and, in addition, \(\eta:K\times K\rightarrow X\) is continuous on \(K\times K\) with \(\eta (u,u)=0\) for any \(u\in K\) and affine with respect to the first variable. Let \(h:K\times K\rightarrow R\) be such that:

  1. (i)

    \(h(u,u)=0\) for all \(u\in X\),

  2. (ii)

    for all \(v\in K\), \(h(\cdot,v)\) is weakly u.s.c.,

  3. (iii)

    for all \(u\in K\), \(h(u,\cdot)\) is convex.

Then GMEP is weakly α-well-posed if and only if GMEP has a unique solution and there exists \(\epsilon_{0}>0\) such that \(\Omega_{\alpha }(\epsilon_{0})\) is nonempty and bounded.

Theorem 4.6

Assume that all the assumptions of Theorem 3.5 are satisfied and, in addition, \(\eta:K\times K\rightarrow X\) is continuous on \(K\times K\) with \(\eta (u,u)=0\) for any \(u\in K\) and affine with respect to the first variable. Let \(h:K\times K\rightarrow R\) is such that:

  1. (i)

    \(h(u,u)=0\) for all \(u\in X\),

  2. (ii)

    for all \(v\in K\), \(h(\cdot,v)\) is u.s.c.,

  3. (iii)

    for all \(u\in K\), \(h(u,\cdot)\) is convex.

Then GMEP is strongly α-well-posed in the generalized sense if and only if

$$\begin{aligned} \Omega_{\eta,\alpha }(\epsilon)\neq \emptyset, \quad \forall \epsilon >0,\quad \textit{and}\quad \lim_{\epsilon \rightarrow 0}\mu \bigl(\Omega_{\eta,\alpha }( \epsilon)\bigr)=0. \end{aligned}$$

Theorem 4.7

Assume that all the assumptions of Theorem 3.6 are satisfied and, in addition, \(\eta:K\times K\rightarrow X\) is continuous on \(K\times K\) with \(\eta (u,u)=0\) for any \(u\in K\) and affine with respect to the first variable. Let \(h:K\times K\rightarrow R\) be such that:

  1. (i)

    \(h(u,u)=0\) for all \(u\in X\),

  2. (ii)

    for all \(v\in K\), \(h(\cdot,v)\) is weakly u.s.c.,

  3. (iii)

    for all \(u\in K\), \(h(u,\cdot)\) is convex.

Then GMEP is weakly α-well-posed in the generalized sense if and only if there exists \(\epsilon_{0}>0\) such that \(\Omega_{\alpha }( \epsilon_{0})\) is nonempty and bounded.

5 Conclusion

In this paper, inspired by the previous works, we study the well-posedness for GVHVI. Under relatively weak conditions for the data F, T, g, J (see Theorems 3.3 and 3.6), we provide some equivalence results for the strong and weak α-well-posed GVHVI in the generalized sense. Our results generalize and improve many known results and can be applied to many other problems.