Next Article in Journal
A New Uncertain Interest Rate Model with Application to Hibor
Next Article in Special Issue
A Class of Sparse Direct Broyden Method for Solving Sparse Nonlinear Equations
Previous Article in Journal
Symmetric Bernstein Polynomial Approach for the System of Volterra Integral Equations on Arbitrary Interval and Its Convergence Analysis
Previous Article in Special Issue
Kantorovich Type Generalization of Bernstein Type Rational Functions Based on (p,q)-Integers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterizations of Well-Posedness for Generalized Hemivariational Inequalities Systems with Derived Inclusion Problems Systems in Banach Spaces

Department of Mathematics, Shanghai Normal University, Shanghai 200234, China
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(7), 1341; https://doi.org/10.3390/sym14071341
Submission received: 6 June 2022 / Revised: 12 June 2022 / Accepted: 17 June 2022 / Published: 29 June 2022
(This article belongs to the Special Issue Symmetry in Nonlinear Functional Analysis and Optimization Theory II)

Abstract

:
In real Banach spaces, the concept of α -well-posedness is extended to the class of generalized hemivariational inequalities systems consisting of two parts which are of symmetric structure mutually. First, certain concepts of α -well-posedness for generalized hemivariational inequalities systems are put forward. Second, certain metric characterizations of α -well-posedness for generalized hemivariational inequalities systems are presented. Lastly, certain equivalence results between strong α -well-posedness of both the system of generalized hemivariational inequalities and its system of derived inclusion problems are established.

1. Introduction

Tykhonov’s well-posedness put forward in [1] has been playing an important role in the study of optimization problems and their related problems such as variational inequalities, inclusion problems, Nash equilibrium problems, etc. For more than the last 50 years, a large number of results regarding well-posedness for optimization problems have been established in the literature; these can be seen, e.g., in [2,3,4,5,6,7,8,9,10,11] and the references therein. In particular, Lucchetti and Patrone [12] extended the concept of well-posedness for optimization problems to the variational inequalities in 1981. Using Ekeland’s variational principle, they presented the characterization of Tykhonov’s well-posedness for minimization problems involving convex and lower semicontinuous (l.s.c.) functions on nonempty, convex and closed sets.
In 1995, Goeleven and Mentagui [13] first put forward the notion of well posedness for hemivariational inequalities (HVIs) and established certain elementary results for well-posed HVIs. Very recently, Wang et al. [14] built the equivalence between the well-posedness of both the hemivariational inequalities system (SHVI) and its derived inclusion problems system (SDIP), i.e., an inclusion problems system which is equivalent to the SHVI. Meanwhile, Ceng, Liou and Wen [15] extended the concept of α -well-posedness to the class of generalized hemivariational inequalities (GHVIs), gave certain metric characterizations of α -well-posedness for GHVIs, and established the equivalence between α -well-posedness of both the GHVI and its derived inclusion problem (DIP), i.e., an inclusion problem which is equivalent to the GHVI. Additionally, Ceng and Lin [16] introduced and considered the α -well-posedness for systems of mixed quasivariational-like inequalities (SMQVLIs) in Banach spaces, and furnished certain metric characterizations of α -well-posedness for SMQVLIs.
Suppose that V k is a real Banach space with its dual V k * for k = 1 , 2 . For k = 1 , 2 , we denote by · , · V k * × V k the duality pairing between V k and V k * and by | | · | | V k and | | · | | V k * the norms of spaces V k and V k * , respectively. It is well known that the product space V = V 1 × V 2 is still a real Banach space endowed with the norm below:
| | u | | V = | | u 1 | | V 1 + | | u 2 | | V 2 u = ( u 1 , u 2 ) V .
For k = 1 , 2 , let A k : V 1 × V 2 2 V k * be a nonempty set-valued mapping, J : V 1 × V 2 R be a locally Lipschitz functional on V and f k be a given point in V k * .
In this paper, we consider the system of generalized hemivariational inequalities (SGHVI), which consists of finding u = ( u 1 , u 2 ) V s.t. for certain ( ω 1 , ω 2 ) A 1 ( u 1 , u 2 ) × A 2 ( u 1 , u 2 ) ,
( SGHVI ) ω 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) 0 v 1 V 1 , ω 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) 0 v 2 V 2 ,
where, for k j = 1 , 2 , J k ( u k , u j ; v k u k ) indicates Clarke’s generalized directional derivative of functional J ( · , u j ) at u k in the direction v k u k , with J ( · , u j ) being a functional on V k for any fixed u j V j , that is,
J k ( u k , u j ; v k u k ) = lim sup w u k , λ 0 J ( w + λ ( v k u k ) , u j ) J ( w , u j ) λ .
It is worth pointing out that the above SGHVI consists of two parts, which are of symmetric structure mutually.
In particular, if A k is a single-valued mapping for k = 1 , 2 , then the above SGHVI reduces to the following system of hemivariational inequalities (SHVI) investigated in [14]:
Find u = ( u 1 , u 2 ) V s.t.
( SHVI ) A 1 ( u 1 , u 2 ) f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) 0 v 1 V 1 , A 2 ( u 1 , u 2 ) f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) 0 v 2 V 2 .
Inspired by the above research works on well posedness, we shall extend the concept of α -well-posedness to the class of SGHVIs in Banach spaces, present certain metric characterizations of α -well-posedness for SGHVIs, and establish the equivalence between the α -well-posedness of both the SGHVI and its SDIP. The architecture of this article is organized below: in Section 2, we present some concepts and basic tools for further use. In Section 3, we define certain notions of α -well-posedness for SGHVIs and, under two assumptions imposed on the operators involved, provide certain metric characterizations of α -well-posedness for SGHVIs. In Section 4, we establish two equivalence results between the α -well-posedness of both the SGHVI and its SDIP.

2. Preliminaries

First of all, we recall certain vital concepts and helpful results on nonlinear analysis, optimization theory and nonsmooth analysis, which can be found in [17,18,19,20,21]. Let E be a real Banach space with its dual E * . Let υ and { υ n } be a point and a sequence in E, and let υ * and { υ n * } be a point and a sequence in E * , respectively. We use the notations υ n υ , υ n υ and υ n * * υ * to represent the strong convergence of { υ n } to υ , the weak convergence of { υ n } to υ and the weak * convergence of { υ n * } to υ * , respectively. Recall that, if E is not reflexive, then the weak * topology of E * is weaker than its weak topology and that if E is reflexive, then the weak * topology of V * coincides with its weak topology. It is readily known that if { υ n } E , { υ n * } E * , υ n υ in E and υ n * * υ * in E * , then υ n * , υ n E * × E υ * , υ E * × E as n .
Definition 1.
Let φ : E R be a functional on E. φ is referred to as being
(i) Lipschitz continuous on E iff L > 0 s.t.
| φ ( υ 1 ) φ ( υ 2 ) | L | | υ 1 υ 2 | | E υ 1 , υ 2 E ;
(ii) Locally Lipschitz continuous on E iff υ E , (neighborhood) N ( υ ) and L υ > 0 s.t.
| φ ( υ 1 ) φ ( υ 2 ) | L υ | | υ 1 υ 2 | | E υ 1 , υ 2 N ( υ ) .
Definition 2.
Let V 1 , V 2 be two real Banach spaces and J : V 1 × V 2 R be a functional on V 1 × V 2 . The functional J is referred to as being:
(i) Lipschitz continuous in the first variable iff the functional J ( · , υ 2 ) : V 1 R is Lipschitz continuous on V 1 for any fixed υ 2 V 2 ;
(ii) Locally Lipschitz continuous in the first variable, iff the functional J ( · , u 2 ) : V 1 R is locally Lipschitz continuous on V 1 for any fixed υ 2 V 2 .
In a similar way, the Lipschitz continuity and locally Lipschitz continuity of the functional J : V 1 × V 2 R in the second variable can be formulated, respectively.
Suppose that φ : E R be a locally Lipschitz functional on E, u is a given point and υ is a directional vector in E. The Clarke’s generalized directional derivative (CGDD) of φ at the point u in the direction υ , denoted by φ ( u ; υ ) , is formulated below
φ ( u ; υ ) = lim sup w u , λ 0 φ ( w + λ υ ) φ ( w ) λ .
According to the CGDD, Clarke’s generalized subdifferential (CGS) of φ at u, denoted by φ ( u ) , is the set in the dual space E * , formulated below
φ ( u ) = { ξ E * : φ ( u ; υ ) ξ , υ E * × E , υ E } .
The following proposition provides some basic properties for the CGDD and the CGS; as can be seen in, e.g., [18,20,22,23,24] and the references therein.
Proposition 1.
Let φ : E R be a locally Lipschitz functional on E and let u , υ E be two given elements. Then:
(i) The function υ φ ( u ; υ ) is finite, positively homogeneous, subadditive and thus convex on E;
(ii) φ ( u ; υ ) is upper semicontinuous (u.s.c.) on E × E as a function of ( u , υ ) , as a function of υ alone, is Lipschitz continuous on E;
(iii) φ ( u ; υ ) = ( φ ) ( u ; υ ) ;
(iv) For all u E , φ ( u ) is a nonempty, convex, bounded and weak * -compact set in E * ;
(v) For all v E , one has
φ ( u ; υ ) = max { ξ , υ E * × E : ξ φ ( u ) } ;
(vi) The graph of the Clarke’s generalized subdifferential φ ( u ) is closed in E × ( w * E * ) topology, with ( w * E * ) being the space E * endowed with the weak * topology, i.e., if { u n } E and { u n * } E * are sequences s.t. u n * φ ( u n ) , u n u in E and u n * u * weakly * in E * , then u * φ ( u ) .
Definition 3.
(i) A single-valued operator T : E E * is referred to as being monotone, iff
T u T υ , u υ E * × E 0 u , υ E ;
(ii) A set-valued operator F : E 2 E * is referred to as being monotone, iff
u * υ * , u υ 0 u , υ E , u * F ( u ) , υ * F ( υ ) .
Definition 4
(see [19]). Let S be a nonempty set in E. The measure of noncompactness (MNC) μ of the set S is formulated below
μ ( S ) = inf { ϵ > 0 : S k = 1 n S k and diam ( S k ) < ϵ k { 1 , 2 , . . . , n } } ,
where diam ( S k ) indicates the diameter of set S k .
Let A 1 , A 2 be the nonempty subsets of E. The Hausdorff metric H ( · , · ) between A 1 and A 2 is formulated by
H ( A 1 , A 2 ) = max { e ( A 1 , A 2 ) , e ( A 2 , A 1 ) } ,
where e ( A 1 , A 2 ) = sup a A 1 d ( a , A 2 ) with d ( a , A 2 ) = inf b A 2 | | a b | | E . It is worth pointing out that certain additional properties of the Hausdorff metric between two sets can be found in [19]. In addition, we note that [25], if A 1 and A 2 are compact subsets in E, we know that a A 1 , b A 2 s.t.
| | a b | | E H ( A 1 , A 2 ) .
Definition 5
(see [26]). Let H ( · , · ) be the Hausdorff metric on the collection C B ( E * ) of all nonempty, closed and bounded subsets of E * , formulated below
H ( A , B ) = max { e ( A , B ) , e ( B , A ) } ,
for A and B in C B ( E * ) . A set-valued operator F : E C B ( E * ) is referred to as being
(i) H -hemicontinuous, if for any u , υ E , the function t H ( F ( u + t ( υ u ) ) , F ( u ) ) from [ 0 , 1 ] into [ 0 , + ) is continuous at 0 + ;
(ii) H -continuous, if ϵ > 0 and (fixed) u 0 E , δ > 0 s.t. υ E with | | υ u 0 | | E < δ , one has H ( F ( v ) , F ( u 0 ) ) < ϵ .
It is remarkable that the H -continuity ensures the H -hemicontinuity, but the converse is generally not true. In the end, we recall a theorem in [27], which is very vital for deducing our main results.
Theorem 1
(see [27]). Suppose that C is nonempty, closed and convex in E and C * is nonempty, closed, convex and bounded in E * . Let φ : E R be a proper convex l.s.c. functional and υ C be arbitrary. Assume that u C , u * ( u ) C * s.t.
u * ( u ) , u υ E * × E φ ( υ ) φ ( u ) .
Then, υ * C * s.t.
υ * , u υ E * × E φ ( υ ) φ ( u ) u C .

3. Metric Characterizations of Well-Posedness for SGHVIs

In this section, we introduce certain notions of α -well-posedness for SGHVIs and establish certain metric characterizations of α -well-posedness for SGHVIs under certain appropriate conditions.
On the basis of certain notions of well-posedness in [2,15,16,26,28,29,30,31,32,33,34], we first introduce certain definitions of α -well-posedness for SGHVIs. For k = 1 , 2 , let α k : V k [ 0 , + ) be convex, continuous, and positively homogeneous, i.e., α k ( s υ k ) = s α k ( υ k ) for all υ k V k and s 0 .
Definition 6.
A sequence { u n } V 1 × V 2 with u n = ( u 1 n , u 2 n ) is referred to as being an α-approximating sequence with α = ( α 1 , α 2 ) for the SGHVI iff ( ω 1 n , ω 2 n ) A 1 ( u 1 n , u 2 n ) × A 2 ( u 1 n , u 2 n ) , n N and { ϵ n } [ 0 , + ) with ϵ n 0 ( n ) s.t.
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 .
In particular, if for k = 1 , 2 , A k is single-valued and α k ( x k y k ) = | | x k y k | | V k x k , y k V k , then { u n } is referred to as being an approximating sequence for SHVI (see [14]).
Definition 7.
The SGHVI is referred to as being strongly (and weakly, respectively) α-well-posed with α = ( α 1 , α 2 ) iff it has a unique solution and every α-approximating sequence for the SGHVI converges strongly (and weakly, respectively) to the unique solution. In particular, if for k = 1 , 2 , A k is single-valued and α k ( x k y k ) = | | x k y k | | V k x k , y k V k , then the SHVI is referred to as being strongly (and weakly, respectively) well-posed (see [14]).
It is evident that the strong α -well-posedness of the SGHVI ensures the weak α -well-posedness of the SGHVI, but the converse is generally not valid.
Definition 8.
The SGHVI is referred to as being strongly (and weakly, respectively) α-well-posed in the generalized sense if the solution set of the SGHVI is nonempty and, for every α-approximating sequence, there always exists a subsequence converging strongly (and weakly, respectively) to some point of the solution set. In particular, if for k = 1 , 2 , A k is single-valued and α k ( x k y k ) = | | x k y k | | V k x k , y k V k , then the SHVI is referred to as being strongly (and weakly, respectively) well-posed in the generalized sense (see [14]).
In a similar way, the strong α -well-posedness in the generalized sense for the SGHVI ensures the weak α -well-posedness in the generalized sense for the SGHVI, but the converse is not valid in general. Obviously, the notions of strong and weak α -well-posedness of the SGHVI put forward in this paper are quite different from those of Definitions 3.1–3.2 and 3.4 in Wang et al. [14]. In order to establish the metric characterizations of α -well-posedness for SGHVI, for any ϵ > 0 , we first formulate two sets in V = V 1 × V 2 below:
Ω α ( ϵ ) = { ( u 1 , u 2 ) V 1 × V 2 : for some ( ω 1 , ω 2 ) A 1 ( u 1 , u 2 ) × A 2 ( u 1 , u 2 ) , ω 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) ϵ α 1 ( v 1 u 1 ) v 1 V 1 , ω 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) ϵ α 2 ( v 2 u 2 ) v 2 V 2 } ,
and
Δ α ( ϵ ) = { ( u 1 , u 2 ) V 1 × V 2 : for all ( v 1 , v 2 ) V 1 × V 2 , ν 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) ϵ α 1 ( v 1 u 1 ) ν 1 A 1 ( v 1 , u 2 ) , ν 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) ϵ α 2 ( v 2 u 2 ) ν 2 A 2 ( u 1 , v 2 ) } .
In order to show certain properties of sets Ω α ( ϵ ) and Δ α ( ϵ ) , we first impose certain hypotheses on the operators A 1 , A 2 and J in the SGHVI.
(HA): (a) A 1 : V 1 × V 2 2 V 1 * is monotone in the first variable, i.e., u 1 , v 1 V 1 and u 2 V 2 ,
v 1 * u 1 * , v 1 u 1 V 1 * × V 1 0 v 1 * A 1 ( v 1 , u 2 ) , u 1 * A 1 ( u 1 , u 2 ) ;
(b) A 2 : V 1 × V 2 2 V 2 * is monotone in the second variable, i.e., u 1 V 1 and u 2 , v 2 V 2 ,
v 2 * u 2 * , v 2 u 2 V 2 * × V 2 0 v 2 * A 2 ( u 1 , v 2 ) , u 2 * A 2 ( u 1 , u 2 ) ;
(c) A 1 : V 1 × V 2 2 V 1 * is a nonempty compact-valued mapping which is H -hemicontinuous;
(d) A 2 : V 1 × V 2 2 V 2 * is a nonempty compact-valued mapping which is H -hemicontinuous;
(e) A 1 : V 1 × V 2 2 V 1 * is a nonempty compact-valued mapping which is H -continuous;
(f) A 2 : V 1 × V 2 2 V 2 * is a nonempty compact-valued mapping which is H -continuous.
(HJ): (a) J : V 1 × V 2 R is locally Lipschitz with respect to the first variable and second variable on V 1 × V 2 ;
(b) J ( u 1 , u 2 ) + J ( v 1 , v 2 ) = J ( u 1 , v 2 ) + J ( v 1 , u 2 ) u = ( u 1 , u 2 ) and v = ( v 1 , v 2 ) in V = V 1 × V 2 .
Lemma 1
(see ([14], Lemma 3.6)). Suppose that the functional J : V 1 × V 2 R satisfies the hypotheses (a), (b) in (HJ). Then, for any sequence u n = ( u 1 n , u 2 n ) V strongly converging towards u = ( u 1 , u 2 ) V and v k n V k strongly converging towards v k V k , one has
lim sup n J k ( u 1 n , u 2 n ; v k n ) J k ( u 1 , u 2 ; v k ) ,
where k = 1 , 2 .
Proposition 2.
Suppose that A 1 : V 1 × V 2 2 V 1 * and A 2 : V 1 × V 2 2 V 2 * satisfy the hypotheses (a), (b), (c), (d) in(HA)and J : V 1 × V 2 R satisfies the hypothesis(HJ). Then, Ω α ( ϵ ) = Δ α ( ϵ ) ϵ > 0 .
Proof. 
From the monotonicity of operators A 1 in the first variable and A 2 in the second variable, it follows that ν 1 ω 1 , v 1 u 1 V 1 * × V 1 0 ν 1 A 1 ( v 1 , u 2 ) , ω 1 A 1 ( u 1 , u 2 ) , and ν 2 ω 2 , v 2 u 2 V 2 * × V 2 0 ν 2 A 2 ( u 1 , v 2 ) , ω 2 A 2 ( u 1 , u 2 ) . Hence, it is easy to see that Ω α ( ϵ ) Δ α ( ϵ ) for any ϵ > 0 . Thus, it is sufficient to show that Δ α ( ϵ ) Ω α ( ϵ ) . In fact, arbitrarily pick a fixed u = ( u 1 , u 2 ) Δ α ( ϵ ) . Then, ( v 1 , v 2 ) V 1 × V 2 , one has
ν 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) ϵ α 1 ( v 1 u 1 ) ν 1 A 1 ( v 1 , u 2 ) , ν 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) ϵ α 2 ( v 2 u 2 ) ν 2 A 2 ( u 1 , v 2 ) .
For any w = ( w 1 , w 2 ) V 1 × V 2 and t ( 0 , 1 ) , letting v 1 : = w 1 , t = u 1 + t ( w 1 u 1 ) and v 2 : = w 2 , t = u 2 + t ( w 2 u 2 ) in (2), we deduce from the positive homogeneousness of α 1 and α 2 that
ω 1 , t f 1 , t ( w 1 u 1 ) V 1 * × V 1 + J 1 ( u 1 , u 2 ; t ( w 1 u 1 ) ) ϵ t α 1 ( w 1 u 1 ) ω 1 , t A 1 ( w 1 , t , u 2 ) , ω 2 , t f 2 , t ( w 2 u 2 ) V 2 * × V 2 + J 2 ( u 1 , u 2 ; t ( w 2 u 2 ) ) ϵ t α 2 ( w 2 u 2 ) ω 2 , t A 2 ( u 1 , w 2 , t ) .
Using Proposition 1 (i), we know that the CGDD is of positive homogeneousness with respect to its direction. So it follows that
ω 1 , t f 1 , w 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; w 1 u 1 ) ϵ α 1 ( w 1 u 1 ) , ω 1 , t A 1 ( w 1 , t , u 2 ) , ω 2 , t f 2 , w 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; w 2 u 2 ) ϵ α 2 ( w 2 u 2 ) , ω 2 , t A 2 ( u 1 , w 2 , t ) .
Since A 1 : V 1 × V 2 2 V 1 * and A 2 : V 1 × V 2 2 V 2 * are nonempty compact-valued mappings, A 1 ( w 1 , t , u 2 ) , A 1 ( u 1 , u 2 ) , A 2 ( u 1 , w 2 , t ) and A 2 ( u 1 , u 2 ) are nonempty compact sets. Hence, by Nadler’s result [25], we deduce that t ( 0 , 1 ) , ω 1 , t A 1 ( w 1 , t , u 2 ) and ω 2 , t A 2 ( u 1 , w 2 , t ) , ν 1 , t A 1 ( u 1 , u 2 ) and ν 2 , t A 2 ( u 1 , u 2 ) s.t.
| | ω 1 , t ν 1 , t | | V 1 * H ( A 1 ( w 1 , t , u 2 ) , A 1 ( u 1 , u 2 ) ) , | | ω 2 , t ν 2 , t | | V 2 * H ( A 2 ( u 1 , w 2 , t ) , A 2 ( u 1 , u 2 ) ) .
Since for k = 1 , 2 , A k ( u 1 , u 2 ) is compact, without loss of generality, we may assume that ν k , t ω k A k ( u 1 , u 2 ) as t 0 + . It is obvious that ( w 1 , t , u 2 ) = ( u 1 , u 2 ) + t [ ( w 1 , u 2 ) ( u 1 , u 2 ) ] and ( u 1 , w 2 , t ) = ( u 1 , u 2 ) + t [ ( u 1 , w 2 ) ( u 1 , u 2 ) ] . Since A k is H -hemicontinuous for k = 1 , 2 , we obtain that
| | ω 1 , t ν 1 , t | | V 1 * H ( A 1 ( w 1 , t , u 2 ) , A 1 ( u 1 , u 2 ) ) 0 as t 0 + , | | ω 2 , t ν 2 , t | | V 2 * H ( A 2 ( u 1 , w 2 , t ) , A 2 ( u 1 , u 2 ) ) 0 as t 0 + ,
which immediately implies that for k = 1 , 2 ,
| | ω k , t ω k | | V k * | | ω k , t ν k , t | | V k * + | | ν k , t ω k | | V k * 0 as t 0 + .
Thus, taking the limit as t 0 + at both sides of the inequalities in (3), we infer from (4) that
ω 1 f 1 , w 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; w 1 u 1 ) ϵ α 1 ( w 1 u 1 ) , ω 2 f 2 , w 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; w 2 u 2 ) ϵ α 2 ( w 2 u 2 ) ,
which, together with the arbitrariness of w = ( w 1 , w 2 ) V 1 × V 2 , implies that Δ α ( ϵ ) Ω α ( ϵ ) . This completes the proof. □
Lemma 2.
Suppose that A 1 : V 1 × V 2 2 V 1 * and A 2 : V 1 × V 2 2 V 2 * satisfy the hypotheses (a), (b), (e), (f) in(HA), and J : V 1 × V 2 R satisfies the hypothesis(HJ). Then, for any ϵ > 0 , Ω α ( ϵ ) = Δ α ( ϵ ) is closed in V = V 1 × V 2 .
Proof. 
Since the H -continuity guarantees the H -hemicontinuity, using Proposition 2, one has Ω α ( ϵ ) = Δ α ( ϵ ) ϵ > 0 . Let u n = ( u 1 n , u 2 n ) Δ α ( ϵ ) be a sequence strongly converging towards u = ( u 1 , u 2 ) in V = V 1 × V 2 . Then, n 1 , ( ω 1 n , ω 2 n ) A 1 ( u 1 n , u 2 n ) × A 2 ( u 1 n , u 2 n ) s.t.
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) ϵ α 1 ( v 1 u 1 n ) , v 1 V 1 , ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) ϵ α 2 ( v 2 u 2 n ) , v 2 V 2 .
Since A 1 : V 1 × V 2 2 V 1 * and A 2 : V 1 × V 2 2 V 2 * are nonempty compact-valued mappings, A 1 ( u 1 n , u 2 n ) , A 1 ( u 1 , u 2 ) , A 2 ( u 1 n , u 2 n ) and A 2 ( u 1 , u 2 ) are nonempty compact sets. Hence, by Nadler’s result [25], one knows that for ω 1 n A 1 ( u 1 n , u 2 n ) and ω 2 n A 2 ( u 1 n , u 2 n ) , ν 1 n A 1 ( u 1 , u 2 ) and ν 2 n A 2 ( u 1 , u 2 ) s.t.
| | ω 1 n ν 1 n | | V 1 * H ( A 1 ( u 1 n , u 2 n ) , A 1 ( u 1 , u 2 ) ) , | | ω 2 n ν 2 n | | V 2 * H ( A 2 ( u 1 n , u 2 n ) , A 2 ( u 1 , u 2 ) ) .
Furthermore, since for k = 1 , 2 , A k ( u 1 , u 2 ) is compact, without loss of generality, we may assume that ν k n ω k A k ( u 1 , u 2 ) as n . For k = 1 , 2 , we note that A k is H -continuous. Thus, we obtain that
| | ω 1 n ν 1 n | | V 1 * H ( A 1 ( u 1 n , u 2 n ) , A 1 ( u 1 , u 2 ) ) 0 as n , | | ω 2 n ν 2 n | | V 2 * H ( A 2 ( u 1 n , u 2 n ) , A 2 ( u 1 , u 2 ) ) 0 as n ,
which immediately implies that, for k = 1 , 2 ,
| | ω k n ω k | | V k * | | ω k n ν k n | | V k * + | | ν k n ω k | | V k * 0 as n .
It therefore follows from (6) that
lim n ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 = ω 1 f 1 , v 1 u 1 V 1 * × V 1 , lim n ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 = ω 2 f 2 , v 2 u 2 V 2 * × V 2 .
Moreover, by the hypothesis (HJ) on the functional J, Lemma 1 ensures that
lim sup n J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) J 1 ( u 1 , u 2 ; v 1 u 1 ) , lim sup n J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) J 2 ( u 1 , u 2 ; v 2 u 2 ) .
Furthermore, using the continuity of α 1 and α 2 , we obtain that, for k = 1 , 2 ,
lim n α k ( v k u k n ) = α k ( v k u k ) .
Therefore, taking the limsup as n at both sides of the inequalities in (5), we conclude from (7)–(9) that
ω 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) ϵ α 1 ( v 1 u 1 ) v 1 V 1 , ω 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) ϵ α 2 ( v 2 u 2 ) v 2 V 2 ,
which implies that u = ( u 1 , u 2 ) Ω α ( ϵ ) = Δ α ( ϵ ) . Thus, Ω α ( ϵ ) = Δ α ( ϵ ) is closed in V = V 1 × V 2 . This completes the proof. □
Theorem 2.
Suppose that A 1 : V 1 × V 2 2 V 1 * satisfy the hypothesis (d) in(HA), A 2 : V 1 × V 2 2 V 2 * satisfy the hypothesis (e) in(HA), and J : V 1 × V 2 R satisfy the hypothesis(HJ). Then, the SGHVI is strongly α-well-posed if and only if
Ω α ( ϵ ) ϵ > 0 and diam ( Ω α ( ϵ ) ) 0 as ϵ 0 .
Proof. 
Necessity. Assume that the SGHVI is strongly α -well-posed. Then, the SGHVI admits a unique solution u = ( u 1 , u 2 ) V 1 × V 2 , i.e., for certain ( ω 1 , ω 2 ) A 1 ( u 1 , u 2 ) × A 2 ( u 1 , u 2 ) ,
SGHVI ω 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) 0 v 1 V 1 , ω 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) 0 v 2 V 2 .
This ensures that u Ω α ( ϵ ) ϵ > 0 , i.e., Ω α ( ϵ ) ϵ > 0 . If diam ( Ω α ( ϵ ) ) 0 as ϵ 0 , then there exists u n = ( u 1 n , u 2 n ) , p n = ( p 1 n , p 2 n ) Ω α ( ϵ n ) , d > 0 and 0 < ϵ n 0 such that
| | u n p n | | V 1 × V 2 = | | u 1 n p 1 n | | V 1 + | | u 2 n p 2 n | | V 2 > d .
By the definition of the α -approximating sequence for the SGHVI, { u n } and { p n } are two α -approximating sequences. Thus, it follows from the strong α -well-posedness of SGHVI that { u n } and { p n } both strongly converge towards the unique solution u , which contradicts (10).
Sufficiency. Suppose that Ω α ( ϵ ) and diam ( Ω α ( ϵ ) ) 0 as ϵ 0 . We claim that the SGHVI is strongly α -well-posed. In fact, let { u n } with u n = ( u 1 n , u 2 n ) be an α -approximating sequence for the SGHVI. Then, there exist ( ω 1 n , ω 2 n ) A 1 ( u 1 n , u 2 n ) × A 2 ( u 1 n , u 2 n ) , n N and a nonnegative sequence { ϵ n } with ϵ n 0 ( n ) such that
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 ,
which implies u n Ω α ( ϵ n ) n 1 . Since diam ( Ω α ( ϵ n ) ) 0 as n , { u n } is a Cauchy sequence in V = V 1 × V 2 . Without loss of generality, we may assume that { u n } strongly converges towards u = ( u 1 , u 2 ) in V = V 1 × V 2 .
Now, we claim that u is a unique solution to the SGHVI. Indeed, since operators A 1 and A 2 are H -continuous on V = V 1 × V 2 , the functional J satisfies the hypothesis (HJ), and α 1 and α 2 are continuous, so we can obtain by similar arguments to those in (7)–(9) that
ω 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) lim n ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + lim sup n J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) = lim sup n { ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) } lim n ϵ n α 1 ( v 1 u 1 n ) = lim n α 1 ( ϵ n ( v 1 u 1 n ) ) = 0 .
By a similar way, one has
ω 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) 0 .
Therefore, u is a solution to the SGHVI.
Finally, we claim the uniqueness of solutions of the SGHVI. Suppose that u is another solution to the SGHVI. Since, for any ϵ > 0 , u , u Ω α ( ϵ ) , | | u u | | V 1 × V 2 diam ( Ω α ( ϵ ) ) , which together with the condition diam ( Ω α ( ϵ ) ) 0 as ϵ 0 , guarantees that u = u . This completes the proof. □
Theorem 3.
Suppose that A 1 : V 1 × V 2 2 V 1 * and A 2 : V 1 × V 2 2 V 2 * satisfy the hypotheses (a), (b), (e) and (f) in(HA)and J : V 1 × V 2 R satisfy the hypothesis(HJ). Then, the SGHVI is strongly α-well-posed in the generalized sense if and only if
Ω α ( ϵ ) ϵ > 0 and μ ( Ω α ( ϵ ) ) 0 ( ϵ 0 ) .
Proof. 
Necessity. Suppose that the SGHVI is strongly α -well-posed in the generalized sense. Then, the solution set S of the SGHVI is nonempty, i.e., S . This ensures that Ω α ( ϵ ) ϵ > 0 because S Ω α ( ϵ ) . Moreover, we claim here that the solution set S of the SGHVI is compact. In fact, for any sequence { u n } S with u n = ( u 1 n , u 2 n ) , { u n } is an α -approximating sequence for the SGHVI and thus there exists a subsequence of { u n } strongly converging towards a certain element of S, which implies that S is compact. To complete the proof of the necessity, we claim that μ ( Ω α ( ϵ ) ) 0 as ϵ 0 . From S Ω α ( ϵ ) , it follows that
H ( Ω α ( ϵ ) , S ) = max { e ( Ω α ( ϵ ) , S ) , e ( S , Ω α ( ϵ ) ) } = e ( Ω α ( ϵ ) , S ) .
Since the solution set S is compact, one has
μ ( Ω α ( ϵ ) ) 2 H ( Ω α ( ϵ ) , S ) = 2 e ( Ω α ( ϵ ) , S ) .
Now, to prove μ ( Ω α ( ϵ ) ) 0 as ϵ 0 , it is sufficient to show that e ( Ω α ( ϵ ) , S ) 0 as ϵ 0 . On the contrary, assume that e ( Ω α ( ϵ ) , S ) 0 as ϵ 0 . Then, there exists a constant l > 0 , a sequence { ϵ n } [ 0 , ) with ϵ n 0 and u n Ω α ( ϵ n ) such that
u n S + B ( 0 , l ) ,
where B ( 0 , l ) is the closed ball centered at 0 with radius l. Since u n Ω α ( ϵ n ) with ϵ n 0 , { u n } is an α -approximating sequence for SGHVI. Thus, there exists a subsequence converging strongly towards a certain element u S due to the strong α -well-posedness in the generalized sense for SGHVI. This contradicts (11). Consequently, μ ( Ω α ( ϵ ) ) 0 as ϵ 0 .
Sufficiency. Assume that Ω α ( ϵ ) ϵ > 0 and μ ( Ω α ( ϵ ) ) 0 ( ϵ 0 ) . We claim that the SGHVI is strongly α -well-posed in the generalized sense. In fact, we observe that
S = ϵ > 0 Ω α ( ϵ ) .
Furthermore, since μ ( Ω α ( ϵ ) ) 0 ( ϵ 0 ) and Ω α ( ϵ ) is nonempty and closed for any ϵ > 0 (due to Lemma 2), it follows from the theorem in ([19], p. 412) that S is nonempty compact and
e ( Ω α ( ϵ ) , S ) = H ( Ω α ( ϵ ) , S ) 0 ϵ 0 .
Now, to show the strong α -well-posedness in the generalized sense for the SGHVI, let { u n } V 1 × V 2 with u n = ( u 1 n , u 2 n ) be an α -approximating sequence for the SGHVI. Then, there exists ( ω 1 n , ω 2 n ) A 1 ( u 1 n , u 2 n ) × A 2 ( u 1 n , u 2 n ) , n N and { ϵ n } [ 0 , + ) with ϵ n 0 ( n ) such that
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 ,
which yields u n Ω α ( ϵ n ) . This, together with (12), leads to
d ( u n , S ) e ( Ω α ( ϵ n ) , S ) 0 .
Since S is compact, there exists u ¯ n S such that
| | u n u ¯ n | | V 1 × V 2 = d ( u n , S ) 0 .
Again from the compactness of the solution set S, one knows that { u ¯ n } has a subsequence { u ¯ n k } strongly converging towards a certain element u ¯ S . Thus, it follows that
| | u n k u ¯ | | V 1 × V 2 | | u n k u ¯ n k | | V 1 × V 2 + | | u ¯ n k u ¯ | | V 1 × V 2 0 ,
which immediately implies that the subsequence { u n k } of { u n } strongly converges towards u ¯ . Therefore, the SGHVI is strongly α -well-posed in the generalized sense. This completes the proof. □
It is remarkable that Proposition 2, Lemma 2 and Theorems 2–3 improve, extend and develop Lemmas 3.7–3.8 and Theorems 3.10–3.11 in [14] to a great extent because the SGHVI is more general than the SHVI considered in Lemmas 3.7–3.8 and Theorems 3.10–3.11 of [14].

4. Equivalence for Well-Posedness of the SGHVI and SDIP

In this section, we first introduce the systems of inclusion problems (SIPs) in the product space V = V 1 × V 2 and then define the concept of α -well-posedness for SIPs. Moreover, we show the equivalence results between the α -well-posedness of the SGHVI and α -well-posedness of its SDIP.
Let V 1 and V 2 be two real Banach spaces with V 1 * and V 2 * being their dual spaces, respectively. Suppose that, for k = 1 , 2 , Γ k is a nonempty set-valued mapping from V 1 × V 2 to V k * . A system of inclusion problems (SIP) associated with mappings Γ 1 and Γ 2 is formulated below:
Find u 1 V 1 and u 2 V 2 such that
( SIP ) 0 1 Γ 1 ( u 1 , u 2 ) , 0 2 Γ 2 ( u 1 , u 2 ) ,
where for k = 1 , 2 , 0 k V k * represents the zero element in V k * . For simplicity, we use the symbols below:
u = ( u 1 , u 2 ) V 1 × V 2 , 0 = ( 0 1 , 0 2 ) V 1 * × V 2 * and Γ ( u ) = ( Γ 1 ( u ) , Γ 2 ( u ) ) V 1 * × V 2 * .
This allows us to simplify the SIP as follows:
Find u V = V 1 × V 2 such that
0 Γ ( u ) .
Definition 9.
A sequence { u n } V 1 × V 2 with u n = ( u 1 n , u 2 n ) is called an α-approximating sequence for the SIP if p n = ( p 1 n , p 2 n ) Γ ( u n ) , n N and { ϵ n } [ 0 , + ) with | | p n | | V 1 * × V 2 * + ϵ n 0 as n , s.t.
p 1 n , v 1 u 1 n V 1 * × V 1 ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , n N , p 2 n , v 2 u 2 n V 2 * × V 2 ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 , n N .
Definition 10.
The SIP is referred to as being strongly (and weakly, respectively) α-well-posed if it has a unique solution and every α-approximating sequence converges strongly (and weakly, respectively) to the unique solution of the SIP.
Definition 11.
The SIP is referred to as being strongly (and weakly, respectively) α-well-posed in the generalized sense if the solution set S of the SIP is nonempty and every α-approximating sequence has a subsequence strongly converging (and weakly, respectively) towards a certain element of the solution set S.
In order to show that the α -well-posedness for the SGHVI is equivalent to the α -well-posedness for its SDIP, we first furnish a lemma which establishes the equivalence between the SGHVI and SDIP.
Lemma 3.
u = ( u 1 , u 2 ) V 1 × V 2 is a solution to the SGHVI if and only if it solves the following SDIP:
Find u = ( u 1 , u 2 ) V 1 × V 2 such that
( SDIP ) f 1 A 1 ( u 1 , u 2 ) + 1 J ( u 1 , u 2 ) , f 2 A 2 ( u 1 , u 2 ) + 2 J ( u 1 , u 2 ) ,
where, for k j = 1 , 2 , k J ( u 1 , u 2 ) denotes the CGS of J ( · , u j ) at u k .
Proof. 
First of all, we claim the necessity. In fact, assume that u = ( u 1 , u 2 ) V = V 1 × V 2 is a solution of the SGHVI, i.e., for certain ( ω 1 , ω 2 ) A 1 ( u 1 , u 2 ) × A 2 ( u 1 , u 2 ) ,
ω 1 f 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) 0 v 1 V 1 , ω 2 f 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) 0 v 2 V 2 .
For any w = ( w 1 , w 2 ) V 1 × V 2 , letting v 1 = u 1 + w 1 V 1 and v 2 = u 2 + w 2 V 2 in (14), we obtain that
J 1 ( u 1 , u 2 ; w 1 ) f 1 ω 1 , w 1 V 1 * × V 1 , J 2 ( u 1 , u 2 ; w 2 ) f 2 ω 2 , w 2 V 2 * × V 2 .
It follows from the definition of the CGS and the arbitrariness of w k V k , k = 1 , 2 that
f 1 ω 1 + 1 J ( u 1 , u 2 ) A 1 ( u 1 , u 2 ) + 1 J ( u 1 , u 2 ) , f 2 ω 2 + 2 J ( u 1 , u 2 ) A 2 ( u 1 , u 2 ) + 2 J ( u 1 , u 2 ) ,
which implies that u = ( u 1 , u 2 ) V 1 × V 2 is a solution to the SDIP.
Sufficiency. Suppose that u = ( u 1 , u 2 ) V = V 1 × V 2 is a solution to the SDIP, i.e.,
f 1 A 1 ( u 1 , u 2 ) + 1 J ( u 1 , u 2 ) , f 2 A 2 ( u 1 , u 2 ) + 2 J ( u 1 , u 2 ) .
It follows that, for k = 1 , 2 , there exist ω k A k ( u 1 , u 2 ) and η k k J ( u 1 , u 2 ) such that
f 1 = ω 1 + η 1 , f 2 = ω 2 + η 2 .
For any v = ( v 1 , v 2 ) V = V 1 × V 2 , by multiplying both sides of the equalities in (15) with v 1 u 1 V 1 and v 2 u 2 V 2 , respectively, we deduce, by the definition of the CGS, that
f 1 , v 1 u 1 V 1 * × V 1 = ω 1 + η 1 , v 1 u 1 V 1 * × V 1 = ω 1 , v 1 u 1 V 1 * × V 1 + η 1 , v 1 u 1 V 1 * × V 1 ω 1 , v 1 u 1 V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 u 1 ) ,
and
f 2 , v 2 u 2 V 2 * × V 2 = ω 2 + η 2 , v 2 u 2 V 2 * × V 2 = ω 2 , v 2 u 2 V 2 * × V 2 + η 2 , v 2 u 2 V 2 * × V 2 ω 2 , v 2 u 2 V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 u 2 ) .
Therefore, u is a solution of the SGHVI. This completes the proof. □
Let E be a real reflexive Banach space with its dual E * . We denote by J the normalized duality mapping from E * to its dual E * * ( = E ) formulated by
J ( ν ) = { x E : ν , x E * × E = | | ν | | E * 2 = | | x | | E 2 } ν E * .
Theorem 4.
Let V 1 and V 2 be real reflexive Banach spaces. Then, the SGHVI is strongly α-well-posed if and only if its SDIP is strongly α-well-posed.
Proof. 
Necessity. Suppose that the SGHVI is strongly α -well-posed. Then there exists a unique u = ( u 1 , u 2 ) V = V 1 × V 2 settling the SGHVI. It follows from Lemma 3 that u is the unique solution of the SDIP. To show the strong α -well-posedness for the SDIP, we let u n = ( u 1 n , u 2 n ) be an α -approximating sequence for the SDIP. We claim that u n u as n . In fact, one knows that there exists a sequence p n = ( p 1 n , p 2 n ) V 1 * × V 2 * , n N and a sequence { ϵ n } [ 0 , + ) , such that for each k = 1 , 2 , p k n A k ( u 1 n , u 2 n ) f k + k J ( u 1 n , u 2 n ) , | | p n | | V 1 * × V 2 * + ϵ n 0 as n and
p 1 n , v 1 u 1 n V 1 * × V 1 ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , n N , p 2 n , v 2 u 2 n V 2 * × V 2 ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 , n N .
It is obvious that for k = 1 , 2 , there exists ω k n A k ( u 1 n , u 2 n ) and η k n k J ( u 1 n , u 2 n ) , such that
p 1 n = ω 1 n f 1 + η 1 n , p 2 n = ω 2 n f 2 + η 2 n .
For k j = 1 , 2 , using the definition of the CGS k J ( u 1 n , u 2 n ) of J ( · , u j n ) at u k n and multiplying both sides of the equalities in (17) with v k u k n V k , we obtain from (16) that
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + η 1 n , v 1 u 1 n V 1 * × V 1 = p 1 n , v 1 u 1 n V 1 * × V 1 ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 ,
and
ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + η 2 n , v 2 u 2 n V 2 * × V 2 = p 2 n , v 2 u 2 n V 2 * × V 2 ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 ,
Therefore, we deduce that { u n } is an α -approximating sequence for the SGHVI. Thus, it follows from the strong α -well-posedness for the SGHVI that { u n } strongly converges towards the unique solution u . This ensures that the SDIP is strongly α -well-posed.
Sufficiency. Suppose that the SDIP is strongly α -well-posed. Then, there exists a unique solution u of the SDIP, which, together with Lemma 3, implies that u is also the unique solution of the SGHVI. Let { u n } be an α -approximating sequence for the SGHVI. Then, there exist ( ω 1 n , ω 2 n ) A 1 ( u 1 n , u 2 n ) × A 2 ( u 1 n , u 2 n ) , n N and { ϵ n } [ 0 , + ) with ϵ n 0 ( n ) such that
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 .
Using Proposition 1 (v), one observes that
J 1 ( u 1 n , u 2 n ; v 1 u 1 n ) = max { h 1 , v 1 u 1 n V 1 * × V 1 : h 1 1 J ( u 1 n , u 2 n ) } , J 2 ( u 1 n , u 2 n ; v 2 u 2 n ) = max { h 2 , v 2 u 2 n V 2 * × V 2 : h 2 2 J ( u 1 n , u 2 n ) } .
Thus, for any ( v 1 , v 2 ) V 1 × V 2 , there exist h 1 ( u 1 n , u 2 n , v 1 ) 1 J ( u 1 n , u 2 n ) and h 2 ( u 1 n , u 2 n , v 2 ) 2 J ( u 1 n , u 2 n ) such that
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + h 1 ( u 1 n , u 2 n , v 1 ) , v 1 u 1 n V 1 * × V 1 ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + h 2 ( u 1 n , u 2 n , v 2 ) , v 2 u 2 n V 2 * × V 2 ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 .
By Proposition 1 (iv), we know that 1 J ( u 1 n , u 2 n ) and 2 J ( u 1 n , u 2 n ) are nonempty, convex, bounded and closed subsets in V 1 * and V 2 * , respectively, which imply that, for each k = 1 , 2 , the set { ω k n + h k f k : h k k J ( u 1 n , u 2 n ) } is also nonempty, convex, bounded and closed in V k * . Therefore, for each k = 1 , 2 , it follows from (19) and Theorem 1 with φ k ( · ) = ϵ n α k ( · u k n ) , which is proper, convex and continuous, that there exists a h k n k J ( u 1 n , u 2 n ) , which is independent on v k , such that
ω 1 n f 1 , v 1 u 1 n V 1 * × V 1 + h 1 n , v 1 u 1 n V 1 * × V 1 ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , ω 2 n f 2 , v 2 u 2 n V 2 * × V 2 + h 2 n , v 2 u 2 n V 2 * × V 2 ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 .
Therefore, it follows that
p 1 n , v 1 u 1 n V 1 * × V 1 ϵ n α 1 ( v 1 u 1 n ) v 1 V 1 , p 2 n , v 2 u 2 n V 2 * × V 2 ϵ n α 2 ( v 2 u 2 n ) v 2 V 2 ,
where p k n = ω k n f k + h k n for k = 1 , 2 . It is readily known that for k = 1 , 2 ,
p k n = ω k n f k + h k n A k ( u 1 n , u 2 n ) f k + k J ( u 1 n , u 2 n ) .
Then, to show that | | p n | | V 1 * × V 2 * 0 as n , it is sufficient to show that | | p k n | | V k * 0 as n for k = 1 , 2 , that is, for any ε > 0 , there exists an integer N 1 such that | | p k n | | V k * < ε for all n N . In fact, note that V k is reflexive, i.e., V k = V k * * . According to the normalized duality mapping J k from V k * to its dual V k * * ( = V k ) formulated below
J k ( ν k ) = { x k V k : ν k , x k V k * × V k = | | ν k | | V k * 2 = | | x k | | V k 2 } ν k V k * ,
we know that for each n N , there exists j ˜ k ( p k n ) J k ( p k n ) such that
p k n , j ˜ k ( p k n ) V k * × V k = | | p k n | | V k * 2 = | | j ˜ k ( p k n ) | | V k 2 .
For k = 1 , 2 , putting v k = u k n j ˜ k ( p k n ) in (21), we obtain
p k n , j ˜ k ( p k n ) V k * × V k ϵ n α k ( j ˜ k ( p k n ) ) ,
that is,
| | p k n | | V k * 2 ϵ n α k ( j ˜ k ( p k n ) ) .
If | | p k n | | V k * 0 as n , then there exists ε k > 0 and for each j 1 , there exists p k n j such that
| | p k n j | | V k * ε k .
Taking into account | | ϵ n j j ˜ k ( p k n j ) | | p k n j | | V k * | | V k 0 as j , and using the positive homogeneousness and continuity of α k , we conclude from (23) that
ε k | | p k n j | | V k * ϵ n j | | p k n j | | V k * α k ( j ˜ k ( p k n j ) ) = α k ( ϵ n j j ˜ k ( p k n j ) | | p k n j | | V k * ) 0 as j ,
which reaches a contradiction. This means that | | p k n | | V k * 0 as n for k = 1 , 2 . Hence, the sequence { u n } with u n = ( u 1 n , u 2 n ) is an α -approximating sequence for SDIP. Thus, it follows from the strong α -well-posedness for the SDIP that { u n } strongly converges towards the unique solution u in V 1 × V 2 . Therefore, the SGHVI is strongly α -well-posed. This completes the proof. □
Using arguments similar to those in the proof of Theorem 4, one can easily prove the following equivalence between the strong α -well-posedness in the generalized sense for the SGHVI and the strong α -well-posedness in the generalized sense for the SDIP. In fact, we first denote by ℧ the solution set of the SGHVI. Note that the SGHVI is strongly α -well-posed ⇔ = { u } and ∀ ( α -approximating sequence) { u n } for the SGHVI it holds u n u , and that the SGHVI is strongly α -well-posed in the generalized sense ⇔ and ∀ ( α -approximating sequence) { u n } , { u n j } { u n } s.t. u n j u for some u . After substituting the strong α -well-posedness in the generalized sense for the SGHVI (and SDIP, respectively) into the strong α -well-posedness for the SGHVI (and SDIP, respectively) in the proof of Theorem 4, we can deduce the conclusion of the following Theorem 5.
Theorem 5.
Let V 1 and V 2 be real reflexive Banach spaces. Then, the SGHVI is strongly α-well-posed in the generalized sense if and only if its SDIP is strongly α-well-posed in the generalized sense.
It is remarkable that, not only in [14] (Theorem 4.5), Wang et al. proved that the SHVI is strongly well-posed if and only if its SDIP is strongly well-posed, but also in [14] (Theorem 4.6), they proved that the SHVI is strongly well-posed in the generalized sense if and only if its SDIP is strongly well-posed in the generalized sense. Compared with Theorems 4.5 and 4.6 of [14], our Theorems 4 and 5 improve and extend them in the following aspects:
(i) The strong well-posedness for the SHVI and its SDIP in [14] (Theorem 4.5) is extended to develop the strong α -well-posedness for the SGHVI and its SDIP in our Theorem 4.
(ii) The strong well-posedness in the generalized sense for the SHVI and its SDIP in [14] (Theorem 4.6) is extended to develop the strong α -well-posedness in the generalized sense for the SGHVI and its SDIP in our Theorem 5.

5. Conclusions

In this article, we extended the concept of α -well-posedness to the class of generalized hemivariational inequalities systems (SGHVIs) consisting of the two parts which are of symmetric structure mutually. In real Banach spaces, we first put forward certain concepts of α -well-posedness for SGHVIs, and then provide certain metric characterizations of α -well-posedness for SGHVIs. Additionally, we establish certain equivalence results of strong α -well-posedness for both the SGHVI and its system of derived inclusion problems (SDIP). In particular, these equivalence results of strong α -well-posedness (i.e., Theorems 4 and 5) improve and extend Theorems 4.5 and 4.6 of [14] in the following aspects:
(i) The strong well-posedness for the SHVI and its SDIP in [14] (Theorem 4.5) is extended to develop the strong α -well-posedness for the SGHVI and its SDIP in our Theorem 4.
(ii) The strong well-posedness in the generalized sense for the SHVI and its SDIP in [14] (Theorem 4.6) is extended to develop the strong α -well-posedness in the generalized sense for the SGHVI and its SDIP in our Theorem 5.
On the other hand, for k = 1 , 2 , let G k : V k R { + } be a proper convex and lower semicontinuous functional, and g ¯ k : V k V k be a continuous mapping. Denote by dom G k the efficient domain of functional G k , that is, dom G k : = { u k V k : G k ( u k ) < + } . Consider the system of generalized strongly variational–hemivariational inequalities (SGSVHVI), which consists of finding u = ( u 1 , u 2 ) V = V 1 × V 2 such that for some ( ω 1 , ω 2 ) A 1 ( g ¯ 1 ( u 1 ) , u 2 ) × A 2 ( u 1 , g ¯ 2 ( u 2 ) ) ,
ω 1 f 1 , v 1 g ¯ 1 ( u 1 ) V 1 * × V 1 + J 1 ( u 1 , u 2 ; v 1 g ¯ 1 ( u 1 ) ) + G 1 ( v 1 ) G 1 ( g ¯ 1 ( u 1 ) ) 0 v 1 V 1 , ω 2 f 2 , v 2 g ¯ 2 ( u 2 ) V 2 * × V 2 + J 2 ( u 1 , u 2 ; v 2 g ¯ 2 ( u 2 ) ) + G 2 ( v 2 ) G 2 ( g ¯ 2 ( u 2 ) ) 0 v 2 V 2 .
It is worth mentioning that the above SGSVHVI also consists of two parts which are of symmetric structure mutually.
In particular, if G k ( v k ) = 0 v k V k and g ¯ k is the identity mapping on V k , then the above SGSVHVI reduces to the SGHVI considered in this article. Additionally, if A k is a single-valued mapping for k = 1 , 2 , then the above SGSVHVI reduces to the SHVI considered in [14].
Finally, it is worth mentioning that part of our future research is aiming to generalize and extend the well-posedness results for SGHVIs in this article to the above class of SGSVHVIs in real Banach spaces.

Author Contributions

Conceptualization, J.-Y.L., H.-Y.H. and Y.-L.C.; Data curation, F.-F.Z.; Formal analysis, J.-Y.L., C.-S.W., F.-F.Z. and L.H.; Funding acquisition, L.-C.C.; Investigation, L.-C.C., J.-Y.L., C.-S.W., F.-F.Z., H.-Y.H., Y.-L.C. and L.H.; Methodology, H.-Y.H. and Y.-L.C.; Project administration, L.-C.C.; Resources, J.-Y.L.; Software, C.-S.W. and L.H.; Supervision, L.-C.C.; Validation, L.H.; Writing-original draft, L.-C.C. and J.-Y.L.; Writing-review and editing, L.-C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially supported by the 2020 Shanghai Leading Talents Program of the Shanghai Municipal Human Resources and Social Security Bureau (20LJ2006100), the Innovation Program of Shanghai Municipal Education Commission (15ZZ068) and the Program for Outstanding Academic Leaders in Shanghai City (15XD1503100).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tykhonov, A.N. On the stability of the functional optimization problem. USSR Comput. Math. Math. Phys. 1966, 6, 28–33. [Google Scholar] [CrossRef]
  2. Ceng, L.C.; Hadjisavvas, N.; Schaible, S.; Yao, J.C. Well-posedness for mixed quasivariational-like inequalities. J. Optim. Theory Appl. 2008, 139, 109–125. [Google Scholar] [CrossRef]
  3. Chen, J.W.; Cho, Y.J.; Ou, X.Q. Levitin-Polyak well-posedness for set-valued optimization problems with constraints. Filomat 2014, 28, 1345–1352. [Google Scholar] [CrossRef]
  4. Fang, Y.P.; Hu, R.; Huang, N.J. Well-posedness for equilibrium problems and for optimization problems with equilibrium constraints. Comput. Math. Appl. 2008, 55, 89–100. [Google Scholar] [CrossRef] [Green Version]
  5. Hu, R.; Fang, Y.P.; Huang, N.J.; Wong, M.M. Well-posedness of systems of equilibrium problems. Taiwanese J. Math. 2010, 14, 2435–2446. [Google Scholar] [CrossRef]
  6. Huang, X.X.; Yang, X.Q. Generalized Levitin-Polyak well-posedness in constrained optimization. SIAM J. Optim. 2006, 17, 243–258. [Google Scholar] [CrossRef] [Green Version]
  7. Lemaire, B. Well-posedness, conditioning, and regularization of minimization, inclusion, and fixed-point problems. Pliska Stud. Math. Bulgar. 1998, 12, 71–84. [Google Scholar]
  8. Levitin, E.S.; Polyak, B.T. Convergence of minimizing sequences in conditional extremum problems. Soviet Math. Dokl. 1966, 7, 764–767. [Google Scholar]
  9. Lignola, M.B. Well-posedness and L-well-posedness for quasivariational inequalities. J. Optim. Theory Appl. 2006, 128, 119–138. [Google Scholar] [CrossRef] [Green Version]
  10. Lin, L.J.; Chuang, C.S. Well-posedness in the generalized sense for variational inclusion and disclusion problems and well-posedness for optimization problems with constraint. Nonlinear Anal. 2009, 70, 3609–3617. [Google Scholar] [CrossRef]
  11. Zolezzi, T. Extended well-posedness of optimization problems. J. Optim. Theory Appl. 1996, 91, 257–266. [Google Scholar] [CrossRef]
  12. Lucchetti, R.; Patrone, F. A characterization of Tykhonov well-posedness for minimum problems with applications to variational inequalities. Numer. Funct. Anal. Optim. 1981, 3, 461–476. [Google Scholar] [CrossRef]
  13. Goeleven, D.; Mentagui, D. Well-posed hemivariational inequalities. Numer. Funct. Anal. Optim. 1995, 16, 909–921. [Google Scholar] [CrossRef]
  14. Wang, Y.M.; Xiao, Y.B.; Wang, X.; Cho, Y.J. Equivalence of well-posedness between systems of hemivariational inequalities and inclusion problems. J. Nonlinear Sci. Appl. 2016, 9, 1178–1192. [Google Scholar] [CrossRef] [Green Version]
  15. Ceng, L.C.; Liou, Y.C.; Wen, C.F. On the well-posedness of generalized hemivariational inequalities and inclusion problems in Banach spaces. J. Nonlinear Sci. Appl. 2016, 9, 3879–3891. [Google Scholar] [CrossRef] [Green Version]
  16. Ceng, L.C.; Lin, Y.C. Metric characterizations of α-well-posedness for a system of mixed quasivariational-like inequalities in Banach spaces. J. Appl. Math. 2012, 2012, 264721. [Google Scholar] [CrossRef] [Green Version]
  17. Carl, S.; Le, V.K.; Motreanu, D. Nonsmooth Variational Problems and Their Inequalities: Comparison Principles and Applications; Springer: New York, NY, USA, 2007. [Google Scholar]
  18. Clarke, F.H. Optimization and Nonsmooth Analysis; SIAM: Philadelphia, PA, USA, 1990. [Google Scholar]
  19. Kuratowski, K. Topology; Academic Press: New York, NY, USA, 1968; Volumes 1–2. [Google Scholar]
  20. Migorski, S.; Ochal, A.; Sofonea, M. Nonlinear Inclusions and Hemivariational Inequalities: Models and Analysis of Contact Problems; Springer: New York, NY, USA, 2013. [Google Scholar]
  21. Zeidler, E. Nonlinear Functional Analysis and Its Applications; Springer: Berlin, Germany, 1990; Volume II. [Google Scholar]
  22. Mahalik, K.; Nahak, C. Existence results for a class of variational quasi-mixed hemivariational-like inequalities. Bull. Malays. Math. Sci. Soc. 2022, 45, 1877–1901. [Google Scholar] [CrossRef]
  23. Migorski, S.; Long, F.Z. Constrained variational-hemivariational inequalities on nonconvex star-shaped sets. Mathematics 2020, 8, 1824. [Google Scholar] [CrossRef]
  24. Ragusa, M.A. Parabolic Herz spaces and their applications. Appl. Math. Lett. 2012, 25, 1270–1273. [Google Scholar] [CrossRef] [Green Version]
  25. Nadler, S.B., Jr. Multi-valued contraction mappings. Pacific J. Math. 1969, 30, 475–488. [Google Scholar] [CrossRef]
  26. Ceng, L.C.; Yao, J.C. Well-posedness of generalized mixed variational inequalities, inclusion problems and fixed-point problems. Nonlinear Anal. TMA 2008, 69, 4585–4603. [Google Scholar] [CrossRef]
  27. Giannessi, F.; Khan, A. Regularization of non-coercive quasivariational inequalities. Control Cybern. 2000, 29, 91–110. [Google Scholar]
  28. Xiao, Y.B.; Huang, N.J.; Wong, M.M. Well-posedness of hemivariational inequalities and inclusion problems. Taiwanese J. Math. 2011, 15, 1261–1276. [Google Scholar] [CrossRef]
  29. Xiao, Y.B.; Yang, X.M.; Huang, N.J. Some equivalence results for well-posedness of hemivariational inequalities. J. Glob. Optim. 2015, 61, 789–802. [Google Scholar] [CrossRef]
  30. Li, X.B.; Xia, F.Q. Levitin-Polyak well-posedness of a generalized mixed variational inequality in Banach spaces. Nonlinear Anal. TMA 2012, 75, 2139–2153. [Google Scholar] [CrossRef]
  31. Ceng, L.C.; Gupta, H.; Wen, C.F. Well-posedness by perturbations of variational-hemivariational inequalities with perturbations. Filomat 2012, 26, 881–895. [Google Scholar] [CrossRef] [Green Version]
  32. Ceng, L.C.; Wen, C.F. Well-posedness by perturbations of generalized mixed variational inequalities in Banach spaces. J. Appl. Math. 2012, 2012, 194509. [Google Scholar] [CrossRef]
  33. Ceng, L.C.; Liou, Y.C.; Wen, C.F. Some equivalence results for well-posedness of generalized hemivariational inequalities with Clarke’s generalized directional derivative. J. Nonlinear Sci. Appl. 2016, 9, 2798–2812. [Google Scholar] [CrossRef] [Green Version]
  34. Ceng, L.C.; Wong, N.C.; Yao, J.C. Well-posedness for a class of strongly mixed variational-hemivariational inequalities with perturbations. J. Appl. Math. 2012, 2012, 712306. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ceng, L.-C.; Li, J.-Y.; Wang, C.-S.; Zhang, F.-F.; Hu, H.-Y.; Cui, Y.-L.; He, L. Characterizations of Well-Posedness for Generalized Hemivariational Inequalities Systems with Derived Inclusion Problems Systems in Banach Spaces. Symmetry 2022, 14, 1341. https://doi.org/10.3390/sym14071341

AMA Style

Ceng L-C, Li J-Y, Wang C-S, Zhang F-F, Hu H-Y, Cui Y-L, He L. Characterizations of Well-Posedness for Generalized Hemivariational Inequalities Systems with Derived Inclusion Problems Systems in Banach Spaces. Symmetry. 2022; 14(7):1341. https://doi.org/10.3390/sym14071341

Chicago/Turabian Style

Ceng, Lu-Chuan, Jian-Ye Li, Cong-Shan Wang, Fang-Fei Zhang, Hui-Ying Hu, Yun-Ling Cui, and Long He. 2022. "Characterizations of Well-Posedness for Generalized Hemivariational Inequalities Systems with Derived Inclusion Problems Systems in Banach Spaces" Symmetry 14, no. 7: 1341. https://doi.org/10.3390/sym14071341

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop