In this paper, by using the KKM theory and the properties of Γ-convexity and RC-mapping, we investigate the existence of collectively fixed points for a family with a finite number of set-valued mappings on the product space of noncompact abstract convex spaces. Consequently, as applications, some existence theorems of generalized weighted Nash equilibria and generalized Pareto Nash equilibria for constrained multiobjective games, some nonempty intersection theorems with applications to the Fan analytic alternative formulation and the existence of Nash equilibria, and some existence theorems of solutions for generalized weak implicit inclusion problems in noncompact abstract convex spaces are given. The results obtained in this paper extend and generalize many corresponding results of the existing literature.
Citation: Haishu Lu, Kai Zhang, Rong Li. Collectively fixed point theorems in noncompact abstract convex spaces with applications[J]. AIMS Mathematics, 2021, 6(11): 12422-12459. doi: 10.3934/math.2021718
[1] | Shunyou Xia, Chongyi Zhong, Chunrong Mo . A common fixed point theorem and its applications in abstract convex spaces. AIMS Mathematics, 2025, 10(3): 5236-5245. doi: 10.3934/math.2025240 |
[2] | Chengqing Pan, Haishu Lu . On the existence of solutions for systems of generalized vector quasi-variational equilibrium problems in abstract convex spaces with applications. AIMS Mathematics, 2024, 9(11): 29942-29973. doi: 10.3934/math.20241447 |
[3] | Ateq Alsaadi, Manochehr Kazemi, Mohamed M. A. Metwali . On generalization of Petryshyn's fixed point theorem and its application to the product of $ n $-nonlinear integral equations. AIMS Mathematics, 2023, 8(12): 30562-30573. doi: 10.3934/math.20231562 |
[4] | Haishu Lu, Xiaoqiu Liu, Rong Li . Upper semicontinuous selections for fuzzy mappings in noncompact $ WPH $-spaces with applications. AIMS Mathematics, 2022, 7(8): 13994-14028. doi: 10.3934/math.2022773 |
[5] | Anam Arif, Muhammad Nazam, Aftab Hussain, Mujahid Abbas . The ordered implicit relations and related fixed point problems in the cone $ b $-metric spaces. AIMS Mathematics, 2022, 7(4): 5199-5219. doi: 10.3934/math.2022290 |
[6] | Meiying Wang, Luoyi Shi, Cuijuan Guo . An inertial iterative method for solving split equality problem in Banach spaces. AIMS Mathematics, 2022, 7(10): 17628-17646. doi: 10.3934/math.2022971 |
[7] | Hui Huang, Xue Qian . Common fixed point of nonlinear contractive mappings. AIMS Mathematics, 2023, 8(1): 607-621. doi: 10.3934/math.2023028 |
[8] | Noor Muhammad, Ali Asghar, Samina Irum, Ali Akgül, E. M. Khalil, Mustafa Inc . Approximation of fixed point of generalized non-expansive mapping via new faster iterative scheme in metric domain. AIMS Mathematics, 2023, 8(2): 2856-2870. doi: 10.3934/math.2023149 |
[9] | Sani Salisu, Poom Kumam, Songpon Sriwongsa . One step proximal point schemes for monotone vector field inclusion problems. AIMS Mathematics, 2022, 7(5): 7385-7402. doi: 10.3934/math.2022412 |
[10] | Siniša N. Ješić, Nataša A. Ćirović, Rale M. Nikolić, Branislav M. Ranƌelović . A fixed point theorem in strictly convex $ b $-fuzzy metric spaces. AIMS Mathematics, 2023, 8(9): 20989-21000. doi: 10.3934/math.20231068 |
In this paper, by using the KKM theory and the properties of Γ-convexity and RC-mapping, we investigate the existence of collectively fixed points for a family with a finite number of set-valued mappings on the product space of noncompact abstract convex spaces. Consequently, as applications, some existence theorems of generalized weighted Nash equilibria and generalized Pareto Nash equilibria for constrained multiobjective games, some nonempty intersection theorems with applications to the Fan analytic alternative formulation and the existence of Nash equilibria, and some existence theorems of solutions for generalized weak implicit inclusion problems in noncompact abstract convex spaces are given. The results obtained in this paper extend and generalize many corresponding results of the existing literature.
Collectively fixed point theorems for a family of set-valued mappings play a key role in studying pure and applied mathematical problems, which can be seen as natural generalizations of fixed point theorems. In 1991, Tarafdar [1] first established a collectively fixed point theorem in the framework of nonempty compact convex subsets of Hausdorff topological vector spaces and then provided its applications in the existence problem of equilibrium points for abstract economies. Since then, many authors have investigated and developed this topic under different assumptions in Hausdorff topological vector spaces. See, for example, [2,3,4,5,6,7] and the references therein.
On the other hand, to broaden the application of the collectively fixed point theory, many authors have studied the collectively fixed point problem in the framework of Hausdorff topological spaces without linear structure. In 1992, Tarafdar [8] extended the collectively fixed point theorem in [1] to compact H-spaces and then gave some applications to the nonempty intersection problem of sets with H-convex sections and existence problem of equilibrium points for an abstract economy. In 1999, Park [9] proved a collectively fixed point theorem which generalizes the collectively fixed point theorems in [1,8] to compact G-convex spaces. In 2003, Yu and Lin [10] generalized the collectively fixed point theorem in [9] to noncompact G-convex spaces. In 2007, Ding [11] and Zhang and Cheng [12] obtained some collectively fixed point theorems in noncompact FC-spaces. In 2010, Al-Homidan et al. [13] derived a collectively fixed point theorem and a maximal element theorem in noncompact topological semilattice spaces and presented applications to problems on generalized abstract economy, systems of vector quasi-equilibrium, and constrained Nash equilibrium. In 2011, Khanh et al. [14] proved some collectively fixed point theorems in noncompact GFC-spaces and gave applications to collectively coincidence point theorems and systems of variational relations. Recently, by means of the technique of partition of unity and Tikhonov fixed point theorem, Khanh and Quan [15] proved the existence of collectively fixed points for a family of set-valued mappings defined on the product set of nonempty sets which have topologically based structures and do not possess linear or convexity structures. Furthermore, they gave applications to coincidence points of a family of set-valued mappings and intersection points of a family of sets.
The abstract convex space is first introduced by Park [16], which includes the spaces mentioned above as special cases. So far, a small part of the literature discussed the problem of collectively fixed points in abstract convex spaces. In 2010, by using a Fan-Browder type fixed point theorem in [17], Park [18] obtained a collectively fixed point theorem for finite families of compact abstract convex spaces and then used this collectively fixed point theorem to obtain a Fan-type nonempty intersection theorem for sets with Γ-convex sections. Recently, Lu and Hu [19] proved a new collectively fixed point theorem for finite families of noncompact abstract convex spaces and gave its applications to equilibria for generalized abstract economies. It is needed to point out that the Hausdorffness of the spaces involved in the collectively fixed point theorems in [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15] for a family of set-valued mappings is necessary since these theorems are proved based on the partition of unity argument. Note that the proofs of the collectively fixed point theorems in [18,19] are based on the Fan-Browder-type fixed point theorem in abstract convex spaces whose Hausdorff separation property can be dropped. Thus, in this sense, the corresponding collectively fixed point theorems in these two cases cannot be deduced from each other.
Motivated and inspired by the work mentioned above, in this paper, the main goal of this paper is to prove the existence of collectively fixed points for a family with a finite number of set-valued mappings defined on the product space of noncompact abstract convex spaces. These obtained collectively fixed point theorems have two alternative coercivity conditions. Furthermore, as applications, in the framework of noncompact abstract convex spaces, some existence theorems of generalized weighted Nash equilibria and generalized Pareto Nash equilibria for constrained multiobjective games, some nonempty intersection theorems for sets with abstract convex sections, and some existence theorems of solutions for generalized weak implicit inclusion problems are established.
The rest of this paper is organized as follows. In Section 2, we introduce some notation, definitions, and lemmas for further investigations. Section 3 is devoted to theorems on collectively fixed points in noncompact abstract convex spaces. The following sections give applications of collectively fixed points in noncompact abstract convex spaces. Section 4 contains existence results for generalized weighted Nash equilibra and generalized Pareto Nash equilibria for constrained multiobjective games. In Section 5, we deal with some nonempty intersection theorems for sets with abstract convex sections and give applications to the Fan analytic alternative formulation and the existence of Nash equilibria for noncooperative games in noncompact abstract convex spaces. Finally, in Section 6, by using a maximal element theorem which is essentially equivalent to fixed point theorem, we obtain some existence results of solutions for generalized weak implicit inclusion problems in the setting of noncompact abstract convex spaces.
In this section, we give some notation, definitions, and lemmas for later use.
Let R and N denote the set of the real numbers and the set of the natural numbers, respectively. For a nonempty set X, let 2X and ⟨X⟩ denote by the family of all subsets of X and by the family of nonempty finite subsets of X, respectively. Let T:X→2Y be a set-valued mapping, where X and Y are two nonempty sets. Then the graph of T is defined by the set {(x,y)∈X×Y:y∈T(x)} and the set-valued mapping T−1:Y→2X is defined by T−1(y)={x∈X:y∈T(x)} for each y∈Y. For each y∈Y, we call T−1(y) the lower section of T. For every X0⊆X, T(X0):=∪x∈X0T(x). If A and B are subsets of a topological space X such that A⊆B, then we denote the closure (respectively, interior) of A in B by clBA (respectively, intBA). When B=X, clA (respectively, intA) denotes the closure (respectively, interior) of A. A topological space X is said to be first-countable if for each x∈X, there exists a sequence {N1,N2,…} of neighbourhoods of x such that for any neighbourhood N of x, there exists an integer k such that Nk⊆N. The product of countable first-countable topological spaces is first-countable, although uncountable product needs not be. Let A be a subset of a first countable topological space X. Then x∈clA if and only if there exists a sequence {xn} in A such that xn→x. We should point out that if A is a subset of a topological space X, then x∈clA if and only if there exists a net {xα} in A such that xα→x.
Definition 2.1 ([20]). Let X and Y be two topological spaces. A set-valued mapping T:X→2Y is called to be:
(i) upper semicontinuous (respectively, lower semicontinuous) at x∈X if for each open set U in Y with T(x)⊆U (respectively, T(x)⋂U≠∅), there is a neighborhood V(x) of x such that T(x′)⊆U (respectively, T(x′)⋂U≠∅) for every x′∈V(x);
(ii) upper semicontinuous (respectively, lower semicontinuous) on X if it is upper semicontinuous (respectively, lower semicontinuous) at every point x∈X;
(iii) continuous on X if it is both upper semicontinuous and lower semicontinuous on X;
(iv) closed if its graph Gr(T)={(x,y)∈X×Y:y∈T(x)} is closed in X×Y.
Lemma 2.1 ([20]). Let T:X→2Y be a set-valued mapping, where X is a topological space and Y is a compact topological space. If the graph of T is closed in X×Y, then T is upper semicontinuous.
Lemma 2.2 ([21]). Let X and Y be two topological spaces and T:X→2Y be a set-valued mapping. Then T is lower semicontinuous at x∈X if and only if for each y∈T(x) and each net {xα}⊆X such that xα→x, there is a net {yα}⊆Y such that yα∈T(xα) for every α and yα→y.
Lemma 2.3 ([21]). Let X and Y be two topological spaces and T:X→2Y be a set-valued mapping. If either T is upper semicontinuous on X with compact values and Y is Hausdorff, or T is upper semicontinuous on X with closed values and Y is regular, then T is closed, that is, the graph of T is closed in X×Y.
Lemma 2.4 ([22]). Let X and Y be two topological spaces and T:X→2Y be a set-valued mapping. If T has compact values, then T is upper semicontinuous at x∈X if and only if for each net {xα}⊆X such that xα→x and for each net {yα}⊆T(xα) for every α, there exist y∈T(x) and a subsbet {yβ} of {yα} such that yβ→y.
In what follows, we introduce some basic definitions and lemmas related to abstract convex spaces. For more details, the reader may refer to [16,17,18,23,27,28,29].
Definition 2.2 ([23]). If X is a topological space, Y is a nonempty set, and Γ:⟨Y⟩→2X is a set-valued mapping with nonempty values ΓA:=Γ(A) for every A∈⟨Y⟩, then the family (X,Y;Γ) is called to be an abstract convex space. When X=Y, we denote (X,X;Γ) by (X;Γ).
Remark 2.1. It is worthwhile noticing that abstract convex spaces contain convex spaces due to Lassonde [24], H-spaces introduced by Horvath [25], G-convex spaces due to Park and Kim [9], L-spaces due to Ben-El-Mechaiekh et al. [26], GFC-spaces due to Khanh et al. [14], FC-spaces due to Ding [11], and many other topological spaces with generalized convex structure (for example, see [18] and references therein).
Definition 2.3 ([23]). Given an abstract convex space (X,Y;Γ) and a nonempty subset Y′ of Y, we define the Γ-convex hull of Y′ by coΓ(Y′)=⋃{ΓA:A∈⟨Y′⟩}.
Definition 2.4 ([23]). Let (X,Y;Γ) be an abstract convex space. A nonempty subset X′ of X is said to be a Γ-convex subset of (X,Y;Γ) relative to a nonempty subset Y′ of Y if we have ΓN⊆X′ for every N∈⟨Y′⟩, that is, coΓ(Y′)⊆X′. In case X=Y, a nonempty subset X′ of X is said to be Γ-convex if coΓ(X′)⊆X′, that is, X′ is Γ-convex relative to itself.
Remark 2.2. Given an abstract convex space (X,Y;Γ), by Definition 2.3, we can see that if a nonempty subset X′ of X is a Γ-convex subset of (X,Y;Γ) relative to a nonempty subset Y′ of Y, then (X′,Y′;Γ|⟨Y′⟩) itself is an abstract convex space which is called to be a subspace of (X,Y;Γ).
Definition 2.5 ([23]). Let (X,Y;Γ) be an abstract convex space and Z be a set. For a set-valued mapping H:X→2Z with nonempty values, if a set-valued mapping G:Y→2Z satisfies H(ΓA)⊆G(A) for every A∈⟨Y⟩, then G is called to be a KKM mapping with respect to H. A KKM mapping G:Y→2X is a KKM mapping with respect to the identity mapping 1X.
Definition 2.6 ([23]). Let (X,Y;Γ) be an abstract convex space and Z be a topological space. A set-valued mapping H:X→2Z is called to be a RC-mapping, if for any closed-valued KKM mapping G:Y→2Z with respect to H, the family {G(y):y∈Y} has the finite intersection property. We denote RC(X,Z):={H:X→2Z| H is a RC-mapping}.
Definition 2.7 ([27]). Let (X,Y;Γ) be an abstract convex space. A function f:X→R is said to be quasi-convex (respectively, quasi-concave) relative to a nonempty subset Y′ of Y if the set {x∈X:f(x)<t} (respectively, {x∈X:f(x)>t}) is Γ-convex relative to Y′ for every r∈R. In case X=Y, a function f:X→R is said to be quasi-convex (respectively, quasi-concave) if the set {x∈X:f(x)<t} (respectively, {x∈X:f(x)>t}) is Γ-convex for every r∈R
Lemma 2.5 ([28]). Let {(Xi,Yi;Γi)}i∈I be a family of abstract convex spaces, where I is a finite (or infinite) index set. Let X:=∏i∈IXi be equipped with the product topology and Y:=∏i∈IYi. For each i∈I, let πi:Y→Yi be the projection. Define Γ=∏i∈IΓi:⟨Y⟩→2E by Γ(A):=∏i∈IΓi(πi(A)) for each A∈⟨Y⟩, where πi(A) is the projection of A onto Xi. Then (X,Y;Γ) is an abstract convex space.
Lemma 2.6 ([29]). Let (X,Y;Γ) be an abstract convex space, (X′,Y′;Γ|⟨Y′⟩) be a subspace of (X,Y;Γ), and Z be a topological space. If H∈KC(X,Z), then H|X′∈KC(X′,cl(H(X′))).
Let (X;Γ) be an abstract convex space and C be a nonempty subset of X. We define the Γ-convex combination of C, denoted by Γ-co(C) as follows.
Γ-co(C)=⋂{D⊆X:D is Γ-convex and C⊆D}. |
We can see that Γ-co(C) is the smallest Γ-convex subset containing C. In fact, for any Γ-convex subset D of X with C⊆D, it follows from the definition of Γ-co(C) that Γ-co(C)⊆D. Next, we show that Γ-co(C) is Γ-convex. Indeed, let A∈⟨Γ-co(C)⟩. Then for each Γ-convex subset D of X with C⊆D, we have A⊆Γ-co(C)⊆D. Since D is Γ-convex, it follows that ΓA⊆D and thus, ΓA⊆Γ-co(C) which implies that Γ-co(C) is Γ-convex. It is obvious that C is Γ-convex if and only if C=Γ-co(C).
Lemma 2.7. Let (X;Γ) be an abstract convex space and C be a nonempty subset of X. Then Γ-co(C)=⋃{Γ-co(A):A∈⟨C⟩}.
Proof. Let A∈⟨C⟩. Then by the fact that Γ-co(A) is the smallest Γ-convex subset containing A and Γ-co(C) is the smallest Γ-convex subset containing C, we have Γ-co(A)⊆Γ-co(C). Therefore, ⋃{Γ-co(A):A∈⟨C⟩}⊆Γ-co(C). Next, we prove that Γ-co(C)⊆⋃{Γ-co(A):A∈⟨C⟩}. Since ⋃{Γ-co(A):A∈⟨C⟩}⊇C, it suffices to show that ⋃{Γ-co(A):A∈⟨C⟩} is Γ-convex. Let B={x0,x1,…,xn}∈⟨⋃{Γ-co(A):A∈⟨C⟩}⟩. Then there exist finite subsets A0,A1,…,An of C such that xi∈Γ-co(Ai), i=0,1,…,n. Let ˆA=⋃ni=0Ai. Then we have ˆA∈⟨C⟩ and xi∈Γ-co(ˆA), i=0,1,…,n. Therefore, by the fact that Γ-co(ˆA) is Γ-convex, we get ΓB⊆Γ-co(ˆA)⊆⋃{Γ-co(A):A∈⟨C⟩}, which implies that ⋃{Γ-co(A):A∈⟨C⟩} is Γ-convex subset containing C. Hence, Γ-co(C)⊆⋃{Γ-co(A):A∈⟨C⟩}. This completes the proof.
Remark 2.3. Lemma 2.7 extends Lemma 1 obtained by Tarafdar [30] in H-spaces, Lemma 2.1 by Tan and Zhang [31] in G-convex spaces, and Lemma 2.1 by Ding [32] in FC-spaces to abstract convex spaces.
Lemma 2.8. Let (X;Γ) be an abstract convex space, Y be a topological space, and F:Y→2X be a set-valued mapping such that F−1(x) is open in Y for every x∈X. Then the set-valued mapping Γ-co(F):Y→2X defined by Γ-co(F)(y)=Γ-co(F(y)) for every y∈Y, has the property that (Γ-co(F))−1(x) is open in Y.
Proof. Let x∈X and y∈(Γ-co(F))−1(x) be any given. Then it suffices to find an open neighborhood O of y in Y such that O⊆(Γ-co(F))−1(x). Since x∈Γ-co(F(y)), it follows from Lemma 2.7 that there exists A={x0,…,xn}∈⟨F(y)⟩ such that x∈Γ-co(A). Let O=⋂ni=0F−1(xi). Since F−1(xi) is open in Y and y∈F−1(xi) for every i=0,…,n, it follows that O is an open neighborhood of y in Y. We show that O⊆(Γ-co(F))−1(x). In fact, let w∈O be any given. Then xi∈F(w) for all i=0,…,n. Hence, we have x∈Γ-co(A)⊆Γ-co(F(w)) and so, w∈(Γ-co(F))−1(x). This implies that (Γ-co(F))−1(x) is open in Y for every x∈X. This completes the proof.
Remark 2.4. Lemma 2.2 due to Ding [32] with underlying FC-spaces, Lemma 3.1 due to Ding [33] for a H-space setting, and Lemma 2.2 due to Tan and Zhang [31] for the framework of a G-convex space are special cases of Lemma 2.8.
In this section, by using the KKM method, we obtain the following theorem which characterizes the existence of collectively fixed points for finite families of set-valued mappings in noncompact abstract convex spaces.
Theorem 3.1. Let {(Xi;Γi)}i∈I be a family of abstract convex spaces such that (X;Γ):=(∏i∈IXi;Γ) is an abstract convex space defined as in Lemma 2.5, where I is a finite index set. Let K be a nonempty compact subset of X. For each i∈I, let Si, Ti:X→2Xi be two set-valued mappings satisfying
(i) for each x∈X, Si(x)⊆Ti(x) and Ti(x) is Γi-convex;
(ii) for each ui∈Xi, S−1i(ui) is open in X;
(iii) for each x∈K, Si(x)≠∅;
(iv) one of the following two conditions holds:
(iv)1 for each Ni∈⟨Xi⟩, there exists a compact Γi-convex subset LNi of (Xi;Γi) containing Ni, such that for L:=∏i∈ILNi, we have
L∖K⊆⋃u∈LintL(⋂i∈IT−1i(ui)⋂L); |
(iv)2 there exists u0∈X such that cl(X∖⋂i∈IT−1i(u0i))⊆K.
If (X;Γ) satisfies 1X∈RC(X,X), then there exists ¯x=(¯xi)i∈I∈X such that ¯xi∈Ti(¯x) for every i∈I.
Proof. Define two set-valued mappings S,T:X→2X by S(x)=∏i∈ISi(x) and T(x)=∏i∈ITi(x) for every x∈X, respectively. We distinguish the following two cases for proving the conclusion that there exists ¯x∈X such that ¯x∈T(¯x).
Case I. If (iv)1 holds, then we suppose contrary to the assertion that x∉T(x) for every x∈X. Define two set-valued mappings ˜S,˜T:X→2X by ˜S(u)=(X∖S−1(u))⋂K and ˜T(u)=cl(X∖T−1(u))⋂K for every u∈X, respectively. We show that the family {˜T(u):u∈X} has the finite intersection property. Indeed, let N∈⟨X⟩ be any given and let πi be the projection from X to Xi for every i∈I. Then for each i∈I, we have πi(N)=Ni∈⟨Xi⟩ and thus, it follows from (iv)1 that there is a compact Γi-convex subset LNi of (Xi;Γi) containing Ni such that L=∏i∈ILNi. Further, let us define two set-valued mappings S′,T′:L→2L by S′(u)=L∖S−1(u) and T′(u)=clL(L∖T−1(u)) for every u∈L, respectively. For each u∈X, by the definition of S, we have
S−1(u)={x∈X:u∈S(x)}={x∈X:u∈∏i∈ISi(x)}={x∈X:ui∈Si(x),∀i∈I}={x∈X:x∈S−1i(ui),∀i∈I}=⋂i∈IS−1i(ui). |
Similarly, we have T−1(u)=⋂i∈IT−1i(ui) for every u∈X. Since I is a finite index set, it follows from (ii) that S−1(u) is open in X for every u∈X. By (i), we can see that T′(u)⊆S′(u) for every u∈X. Now, we check that the set-valued mapping T″:L→2L defined by T″(u)=L∖T−1(u) for every u∈L, is a KKM mapping. In fact, if this were not, then there exist A∈⟨L⟩ and x0∈Γ(A)⊆L such that
x0∉⋃u∈AT″(u)=L∖(⋂u∈AT−1(u)), |
which implies that x0∈⋂u∈AT−1(u) and thus, A⊆T(x0). By (i) again, we can deduce that T(x0) is Γ-convex. Therefore, we have x0∈Γ(A)⊆T(x0), which contradicts our assumption that x∉T(x) for every x∈X. Hence, T″:L→2L is a KKM mapping and so is T′. Since L is Γ-convex, it follows from Remark 2.2 that (L;Γ|⟨L⟩) be a subspace of (X;Γ). So, by Lemma 2.6 and the fact that 1X∈RC(X,X), we have 1L∈RC(L,L). Since T′ is a KKM mapping with closed compact values and (iv)1 holds, it follows that ∅≠⋂u∈LT′(u)=⋂u∈LclL(L∖T−1(u))⊆L⋂K. Let x0∈⋂u∈LT′(u). Then we have
x0∈⋂u∈LT′(u)⊆⋂u∈N(T′(u)⋂K)⊆⋂u∈N˜T(u). |
This implies that the family {˜T(u):u∈X} has the finite intersection property. By the compactness of K, we obtain ⋂u∈X˜T(u)≠∅. Since ˜T(u)⊆˜S(u) for every u∈X, we have
∅≠⋂u∈X˜S(u)=⋂u∈X(X∖S−1(u))⋂K=K∖⋃u∈XS−1(u), |
which implies that there exists x∗∈K such that S(x∗)=∅. By the definition of S again, there exists i0∈I such that Si0(x∗)=∅, which contradicts (iii). Therefore, there exists ¯x∈X such that ¯x∈T(¯x). By the definition of T, we have ¯xi∈Ti(¯x) for every i∈I. This completes the proof.
Case II. Assume that (iv)2 hold. Suppose to the contrary that x∉T(x) for every x∈X. Define two set-valued mappings ˜S,˜T:X→2X by ˜S(u)=(X∖S−1(u)) and ˜T(u)=cl(X∖T−1(u)) for every u∈X, respectively. By (i), (ii), and the expressions of S−1(u) and T−1(u) in Case I, we have ˜T(u)⊆˜S(u) for every u∈X. We show that Γ(A)⊆⋃u∈A˜T(u) for every A∈⟨X⟩, that is, ˜T is a KKM mapping. Otherwise, there exist A∈⟨X⟩ and a point x0∈Γ(A) such that x0∉⋃u∈A˜T(u)=X∖⋂u∈AintT−1(u). It follows that x0∈⋂u∈AT−1(u). Therefore, we have A⊆T(x0). According to (i) and the definition of T, we can see that T(x0) is Γ-convex and thus, x0∈Γ(A)⊆T(x0). This creates a contradiction. Hence, ˜T is a KKM mapping. Since 1X∈RC(X,X) and ˜T(u) is closed in X for every u∈X, it follows that the family {˜T(u):u∈X} has the finite intersection property. By (iv)2, there exists u0∈X such that
˜T(u0)=cl(X∖T−1(u0))=cl(X∖⋂i∈IT−1i(u0i))⊆K, |
which implies that ˜T(u0) is compact. Consequently, the intersection of the family {˜T(u):u∈X} is nonempty. Let x0∈⋂u∈X˜T(u). Then we have x0∈K⋂(⋂u∈X˜S(u)). Thus, we get S(x0)=∅. It follows from the definition of S that there exists i0∈I such that Si0(x0)=∅, which contradicts (iii). Therefore, there exists ¯x∈X such that ¯x∈T(¯x). By the definition of T again, we have ¯xi∈Ti(¯x) for every i∈I. The proof is complete.
Remark 3.1. (1) Unlike Theorem 6.1 obtained by Park [18], the abstract convex spaces involved in Theorem 3.1 is not required to be compact.
(2) Theorem 3.1 cannot be regarded as a special case of Theorem 10 due to Lu and Hu [19]. Although (i)–(iii) of Theorem 3.1 are stronger than the corresponding conditions of Theorem 10 in Lu and Hu [19], Theorem 3.1 has two coercive conditions to be selected, and both the first coercive condition of Theorem 3.1 and the corresponding coercive condition of Theorem 10 in Lu and Hu [19] are independent of each other. Thus, Theorem 3.1 and Theorem 10 obtained by Lu and Hu [19] cannot be deduced from each other. In addition, the methods of proving these two theorems are also different. The proof of our theorem is based on KKM theory in abstract convex spaces, and the proof of Theorem 10 in Lu and Hu [19] is to use a fixed point theorem in abstract convex spaces.
Theorem 3.2. Let {(Xi;Γi)}i∈I be a family of abstract convex spaces such that (X;Γ):=(∏i∈IXi;Γ) is an abstract convex space defined as in Lemma 2.5, where I is a finite index set. Let K be a nonempty compact subset of X. For each i∈I, let Si, Ti:X→2Xi be two set-valued mappings satisfying
(i) for each x∈X, Si(x)⊆Γ-co(Ti(x));
(ii) for each ui∈Xi, S−1i(ui) is open in X;
(iii) for each x∈K, Si(x)≠∅;
(iv) one of the following two conditions holds:
(iv)1 for each Ni∈⟨Xi⟩, there exists a compact Γi-convex subset LNi of (Xi;Γi) containing Ni, such that for L:=∏i∈ILNi, we have
L∖K⊆⋃u∈LintL(⋂i∈IT−1i(ui)⋂L); |
(iv)2 there exists u0∈X such that cl(X∖⋂i∈IT−1i(u0i))⊆K.
If (X;Γ) satisfies 1X∈RC(X,X), then there exists ¯x=(¯xi)i∈I∈X such that ¯xi∈Γ-co(Ti(¯x)) for every i∈I.
Proof. For each i∈I, we define a set-valued mapping ~Ti:X→2Xi by ~Ti(x)=Γ-co(Ti(x)) for every x∈X. By (i) the definition of Γ-convex combination, we can see that Si(x)⊆~Ti(x) and ~Ti(x) is Γi-convex for every i∈I and every x∈X. From (iv) and the definition of ~Ti, one can see that one of the following two conditions holds:
∙ for each Ni∈⟨Xi⟩, there exists a compact Γi-convex subset LNi of (Xi;Γi) containing Ni, such that for L:=∏i∈ILNi, we have
L∖K⊆⋃u∈LintL(⋂i∈I˜T−1i(ui)⋂L); |
∙ there exists u0∈X such that cl(X∖⋂i∈I˜T−1i(u0i))⊆K.
So far, combined with (ii) and (iii), we can see that all the conditions of Theorem 3.1 are fulfilled. Thus, by Theorem 3.1, there exists ¯x=(¯xi)i∈I∈X such that ¯xi∈Γ-co(Ti(¯x)) for every i∈I. This completes the proof.
Remark 3.2. Theorem 3.1 is equivalent to Theorem 3.2. In fact, we only need to show that theorem 3.2 implies Theorem 3.1. By (i) of Theorem 3.1 and the definition of Γ-convex combination, we have Ti(x)=Γ-co(Ti(x)) for every i∈I and every x∈X. Therefore, it follows from Theorem 3.2 that there exists ¯x=(¯xi)i∈I∈X such that ¯xi∈Γ-co(Ti(¯x))=Ti(x) for every i∈I.
Let I in Theorem 3.1 be a singleton. Then we have the following fixed point theorem.
Theorem 3.3. Let (X;Γ) be an abstract convex space, K be a nonempty compact subset of X, and S, T:X→2X be two set-valued mappings such that
(i) for each x∈X, S(x)⊆T(x) and T(x) is Γ-convex;
(ii) for each u∈X, S−1(u) is open in X;
(iii) for each x∈K, S(x)≠∅;
(iv) one of the following two conditions holds:
(iv)1 for each N∈⟨X⟩, there exists a compact Γ-convex subset LN of (X;Γ) containing N such that
LN∖K⊆⋃u∈LNintLN(T−1(u)⋂LN); |
(iv)2 there exists u0∈X such that cl(X∖T−1(u0))⊆K.
If (X;Γ) satisfies 1X∈RC(X,X), then there exists ¯x∈X such that ¯x∈T(¯x).
Remark 3.3. Theorem 3.3 extends the famous Fan-Browder fixed point theorem due to Browder [34], Corollary 1 obtained by Horvath and Ciscar [35], Theorem 3.2 by Yannelis and Prabhakar [36], Corollary 1 by Ansari and Yao [3], Corollary 3.1 by Al-Homidan and Ansari [37], Theorem 2.4 by Luo [38], and several other fixed point theorems in the literature to noncompact abstract convex spaces (see Park [18] and the references therein).
When I is a singleton and S=T, it is obvious that the following maximal element theorem can be obtained from Theorem 3.1 (or Theorem 3.3). We omit the proof.
Theorem 3.4. Let {(X;Γ1)} and {(Y;Γ2)} be two abstract convex spaces such that (X×Y;Γ1×Γ2) is an abstract convex space defined as in Lemma 2.5. Let K be a nonempty compact subset of X×Y. Let T:X×Y→2X×Y be a set-valued mapping satisfying
(i) for each (x,y)∈X×Y, T(x,y) is Γ1×Γ2-convex;
(ii) for each (u,v)∈X×Y, T−1(u,v) is open in X×Y;
(iii) for each (x,y)∈X×Y, (x,y)∉T(x,y);
(iv) one of the following two conditions holds:
(iv)1 for each N0×N1∈⟨X×Y⟩, there exist a compact Γ1-convex subset LN0 of (X;Γ1) containing N0 and a compact Γ2-convex subset LN1 of (Y;Γ2) containing N1 such that for L:=LN0×LN1, one has L∖K⊆⋃(u,v)∈LT−1(u,v);
(iv)2 there exists (u0,v0)∈X×Y such that X×Y∖T−1(u0,v0)⊆K.
If (X×Y;Γ1×Γ2) satisfies 1X×Y∈RC(X×Y,X×Y), then there exists (¯x,¯y)∈K such that T(¯x,¯y)=∅.
Remark 3.4. (1) It is obvious that (iv)1 of Theorem 3.4 is equivalent to the following condition:
(iv)1′ for each N0×N1∈⟨X×Y⟩, there exist a compact Γ1-convex subset LN0 of (X;Γ1) containing N0 and a compact Γ2-convex subset LN1 of (Y;Γ2) containing N1 such that for L:=LN0×LN1, one has L∖K⊆⋃(u,v)∈L(T−1(u,v)⋂L).
(2) If we drop (i) of Theorem 3.4, then (iii) of Theorem 3.4 can be replaced by the following stronger condition:
(x,y)∉Γ1×Γ2-co(T(x,y)), ∀(x,y)∈X×Y. | (3.1) |
In fact, we can show that the conclusion of Theorem 3.4 still holds when (3.1) is satisfied. Define a set-valued mapping ˜T:X×Y→2X×Y by ˜T(x,y)=Γ1×Γ2-co(T(x,y)) for every (x,y)∈X×Y. It is obvious that ˜T(x,y) is Γ1×Γ2-convex for every (x,y)∈X×Y. By Lemma 2.8, ˜T−1(u,v) is open in X×Y for every (u,v)∈X×Y. It follows from (3.1) that (x,y)∉˜T(x,y) for every (x,y)∈X×Y. Finally, by (iv), we can see that one of the following two conditions holds:
∙ for each N0×N1∈⟨X×Y⟩, there exist a compact Γ1-convex subset LN0 of (X;Γ1) containing N0 and a compact Γ2-convex subset LN1 of (Y;Γ2) containing N1 such that for L:=LN0×LN1, one has
L∖K⊆⋃(u,v)∈LT−1(u,v)⊆⋃(u,v)∈L˜T−1(u,v); |
∙ there exists (u0,v0)∈X×Y such that
X×Y∖˜T−1(u0,v0)⊆X×Y∖T−1(u0,v0)⊆K. |
Thus, all the hypotheses of Theorem 3.4 are satisfied. Therefore, by Theorem 3.4, there exists (¯x,¯y)∈K such that ˜T(¯x,¯y)=∅ and so, T(¯x,¯y)=∅.
(3) Combining the above arguments in (2), we can see that Theorem 3.4 generalizes Lemma 2.1 of Balaj and Lin [39] in the following aspects: (a) from noncompact topological vector spaces to noncompact abstract convex spaces; (b) the Hausdorffness of the topological spaces in Theorem 3.4 is redundant, while the topological spaces in Lemma 2.1 of Balaj and Lin [39] are assumed to be Hausdorff; (c) from one coercivity condition to two alternative coercivity conditions; (d) the conclusion of our Theorem 3.4 is stronger than that of Lemma 2.1 of Balaj and Lin [39] since the maximal elements of T can be found in K instead of X.
In this section, we shall consider the constrained multiobjective game in its strategic form Θ:=((Xi;Γi),Ui,Ai,Bi)i∈I, where I={1,2,…,n} is a finite set of player. For each i∈I, Xi is the strategy set of player i such that (Xi;Γi) is an abstract convex space, Ai,Bi:X=∏i∈IXi→2Xi are two constraint set-valued mappings of the ith player, and Ui:X=Πi∈IXi→Rki is the payoff function of the ith player, where ki∈N. For each i∈I, we denote Xˆi:=∏j∈I∖iXj. If x=(x1,x2,…,xn)∈X, then we write xˆi:=(x1,…,xi−1,xi+1,…,xn) for every i∈I. If xi∈Xi, zi∈Xi and xˆi∈Xˆi, then we use the notation (xˆi,xi):=(x1,…,xi−1,xi,xi+1,…,xn)=x∈X and the natation (xˆi,zi):=(x1,…,xi−1,zi,xi+1,…,xn)∈X. If a choice x=(x1,…,xn) is played, each player i is trying to find his/her vector payoff function Ui(x):=(ui1(x),…,uiki(x)) consisting of non-commensurable outcomes. Each player i has a preference ⪰i over the outcome space Rki. For each i∈I, the ith player's preference ⪰i is defined by
z1⪰iz2 if and only if z1j≥z2j for each j=1,…,ki, |
where z1=(z11,…,z1ki)∈Rki and z2=(z21,…,z2ki)∈Rki. The players' preference relations induce the preferences on X which is defined by x⪰iy ⇔ Ui(x)⪰iUi(y) for each player i and their choices x=(x1,…,xn), y=(y1,…,yn)∈X.
If A(x)=Bi(x)≠Xi for every i∈I and every x∈X, then the model of constrained multiobjective games with two constrained set-valued mappings reduces to the model of constrained multiobjective games with one constrained set-valued mapping considered by Ding [40] and Kim and Ding [41]. If A(x)=Bi(x)=Xi for every i∈I and every x∈X, then the constrained multiobjective game model reduces to the multiobjective game model studied by Wang [42], Yuan and Tarafdar [43], and Yu and Yuan [44].
We need to point out that the constrained multiobjective game model in this paper is a non-cooperative game model, which implies that there is no communicating between players and so, players act as free agents, and each player is trying to minimize his/her own payoff function according to his/her preference.
For a multiobjective game, as it is well known, in general, there does not exist a strategy ˆx∈X to minimize all uijs for each player i∈I; see, for example, Yu [45] and the references therein. Hence, we need to give some solution concepts for the multicriteria games with constraint set-valued mappings.
Throughout this paper, for each m∈N, we shall denote by Rm+:={q:=(q1,…,qm)∈Rm:qj≥0,∀j=1,…,m} and intRm+:={q:=(q1,…,qm)∈Rm:qj>0,∀j=1,…,m} the nonnegative orthant of Rm and the nonempty interior with the topology induced by the Euclidean metric, respectively. For each u,v∈Rm, u⋅v denotes the standard Euclidian inner product.
Let ˆx=(ˆx1,…,ˆxn)∈X. Now, we have the following definitions.
Definition 4.1. A strategy ˆxi∈Xi of player i is said to be a generalized Pareto efficient strategy (respectively, a generalized weak Pareto efficient strategy) of the constrained multiobjective game Θ=((Xi;Γi),Ui,Ai,Bi)i∈I with respect to ˆx if ˆxi∈Bi(ˆx) and there is no strategy xi∈Ai(ˆx) such that
Ui(ˆx)−Ui(ˆxˆi,xi)∈Rki+∖{0} (respectively, Ui(ˆx)−Ui(ˆxˆi,xi)∈intRki+). |
Definition 4.2. A strategy ˆx∈X is said to be a generalized Pareto equilibrium (respectively, a generalized weak Pareto equilibrium) of the constrained multiobjective game Θ=((Xi;Γi),Ui,Ai,Bi)i∈I if for each player i, ˆxi∈Xi is a generalized Pareto efficient strategy (respectively, a generalized weak Pareto efficient strategy) of the constrained multiobjective game Θ:=((Xi;Γi),Ui,Ai,Bi)i∈I with respect to ˆx.
Remark 4.1. The above two definitions generalize the corresponding definitions in [42,43,44]. It is clear that every generalized Pareto equilibrium is a generalized weak Pareto equilibrium, but the converse is not always true.
Definition 4.3. A strategy ˆx∈X is said to be a generalized weighted Nash equilibrium with respect to the weight vector W=(Wi)i∈I with Wi=(Wi,1,Wi,2…,Wi,ki)∈Rki+ of the constrained multiobjective game Θ=((Xi;Γi),Ui,Ai,Bi)i∈I if for each player i, we have
(i) ˆxi∈Bi(ˆx);
(ii) Wi∈Rki+∖{0};
(iii) Wi⋅Ui(ˆx)≤Wi⋅Ui(ˆxˆi,xi) for every xi∈Ai(ˆx), where ⋅ denotes the inner product in Rki.
Remark 4.2. When Wi∈Rki+ with ∑kij=1Wij=1 for every i∈I, the strategy ˆx∈X is said to a normalized form of generalized weighted Nash equilibrium with respect to the weight vector W. In addition, it follows from the above definition that ˆx∈X is a generalized weighted Nash equilibrium with respect to the weight vector W=(Wi)i∈I of the constrained multiobjective game Θ=((Xi;Γi),Ui,Ai,Bi)i∈I if and only if ˆx∈X is a solution of the constrained optimization problem as follows: find ˆx∈X such that for each i∈I, ˆxi∈Bi(ˆx) and min.
The following lemma shows that the existence problem of generalized weak Pareto equilibrium (respectively, generalized Pareto equilibrium) for a constrained multiobjective game can be reduced to the existence problem of generalized weighted Nash equilibrium under certain conditions.
Lemma 4.1. Let \Theta = ((X_i; \Gamma_i), U^i, A_i, B_i)_{i\in I} be a constrained multiobjective game. Then a normalized form of generalized weighted Nash equilibrium \widehat{x}\in X with respect to a weight W = (W_1, \ldots, W_n) , W_i\in\mathbb{R}_{+}^{k_i}\setminus\{0\} (respectively, W_i\in\text{int}\mathbb{R}_{+}^{k_i} ) and \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I , is a generalized weak Pareto equilibrium (respectively, a generalized Pareto equilibrium) of the game \Theta .
Proof. Suppose to the contrary that \widehat{x} is not a generalized weak Pareto equilibrium. Then by Definitions 4.1 and 4.2, there exists some i_0\in I such that \widehat{x}_{i_0}\not\in B_{i_0}(\widehat{x}) or there exists an x_{i_0}\in A_{i_0}(\widehat{x}) such that
U^i(\widehat{x}_{\widehat{i_0}}, \widehat{x}_{i_0})-U^i(\widehat{x}_{\widehat{i_0}}, x_{i_0})\in \text{int}\mathbb{R}_{+}^{k_{i_0}}. |
It is obvious that \widehat{x}_{i_0}\not\in B_{i_0}(\widehat{x}) contradicts the the assumption that \widehat{x} is a normalized generalized weighted Nash equilibrium with respect to the weight W = (W_1, \ldots, W_n) . Thus, we only consider the second case that there exists an x_{i_0}\in A_{i_0}(\widehat{x}) such that U^i(\widehat{x}_{\widehat{i_0}}, \widehat{x}_{i_0})-U^i(\widehat{x}_{\widehat{i_0}}, x_{i_0})\in \text{int}\mathbb{R}_{+}^{k_{i_0}} . In fact, since W_{i_0}\in\mathbb{R}_{+}^{k_{i_0}}\setminus\{0\} with \sum_{j = 1}^{k_{i_0}}W_{i_0, j} = 1 , it follows that W_{i_0}\cdot U^i(\widehat{x}_{\widehat{i_0}}, \widehat{x}_{i_0}) > W_{i_0}\cdot U^i(\widehat{x}_{\widehat{i_0}}, x_{i_0}) , which also contradicts the fact that \widehat{x} is a normalized form of generalized weighted Nash equilibrium with respect to the weight W = (W_1, \ldots, W_n) . Therefore, \widehat{x} is a generalized weak Pareto equilibrium. Now, we suppose that W_i\in\text{int}\mathbb{R}_{+}^{k_i} and \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I . We show that \widehat{x} is a generalized Pareto equilibrium by contradiction. If this was not the case, then by Definitions 5.1 and 5.2, there exists i_0\in I such that \widehat{x}_{i_0}\not\in B_{i_0}(\widehat{x}) or there exists an x_{i_0}\in A_{i_0}(\widehat{x}) such that
U^i(\widehat{x}_{\widehat{i_0}}, \widehat{x}_{i_0})-U^i(\widehat{x}_{\widehat{i_0}}, x_{i_0})\in \mathbb{R}_{+}^{k_{i_0}}\setminus\{0\}. |
By using the same argument as in the above, we get contradictions. Therefore, \widehat{x} is a generalized Pareto equilibrium. This completes the proof.
Remark 4.3. It should be noted that the conclusion of Lemma 4.1 still holds if \widehat{x}\in X is a generalized weighted Nash equilibrium with respect to a weight W = (W_1, \ldots, W_n) satisfying W_i\in\mathbb{R}_{+}^{k_i}\setminus\{0\} (respectively, W_i\in\text{int}\mathbb{R}_{+}^{k_i} ) for every i\in I . Also, we point out that a generalized Pareto equilibrium is not necessarily a generalized weighted Nash equilibrium.
Lemma 4.2 ([41]). Let X and Y be two topological spaces. Let T:X\rightarrow 2^Y be a continuous set-valued mapping such that T(x) is nonempty compact subset of Y for every x\in X . Suppose that f:X\times Y\rightarrow \mathbb{R} is a continuous function. Then the function \xi:X\rightarrow \mathbb{R} defined by \xi(x): = \text{min}_{y\in T(x)}f(x, y) for every x\in X , is a continuous function on X .
Now, as applications of Theorems 3.1 and 3.3, we have the following existence theorems of generalized weighted Nash equilibria and generalized Pareto equilibria for constrained multiobjective games.
Theorem 4.1. Let \Theta = ((X_i; \Gamma_i), U^i, A_i, B_i)_{i\in I} be a constrained multiobjective game such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K is a nonempty compact subset of X , where I is a finite index set. For each i\in I and each u_i\in X_i , A_i^{-1}(u_i) is open in X . Assume that there exists a weight vector W = (W_1, \ldots, W_n) with W_i\in \mathbb{R}_{+}^{k_i}\setminus \{0\} such that for each i\in I , the following conditions are satisfied:
(i) for each x\in X , \emptyset\neq A_i(x)\subseteq B_i(x) , and B_i(x) is \Gamma_i -convex;
(ii) for each x\in X , the set \{u_i\in X_i:W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is \Gamma_i -convex;
(iii) for each u_i\in X_i , the set \{x\in X:W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is open in X ;
(iv) the set \mathfrak{F}_i = \{x\in X:\text{there exists}\ u_i\in A_i(x)\ \text{such that}\ W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is a closed subset of X ;
(v) one of the following conditions holds:
(v) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that L\setminus K\subseteq \bigcup_{u\in L}\text{int}_{L}(\bigcap_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap B_i^{-1}(u_i))\bigcap L) , where L: = \prod_{i\in I}L_{N_{i}} ;
(v) _2 there exists u_0\in X such that \text{cl}(X\setminus \bigcap_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap B_i^{-1}(u_{0i})))\subseteq K .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the game \Theta has a generalized weighted Nash equilibrium \widehat{x}\in X with respect to the weight vector W = (W_i)_{i\in I} and hence it has a generalized weak Pareto equilibrium. Further, if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I , then \Theta has a generalized Pareto equilibrium.
Proof. We shall prove this theorem by considering the following two cases:
Case I. Suppose that the set \mathfrak{F}_i = \{x\in X:\text{there exists}\ u_i\in A_i(x)\ \text{such that}\ W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is empty for every i\in I . Then we have W_i\cdot U^i(x_{\widehat{i}}, u_i)\geq W_i\cdot U^i(x_{\widehat{i}}, x_i) for every i\in I , x\in X , and every u_i\in A_i(x) . By (v), we know that the one of the following conditions holds:
\bullet for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that L\setminus K\subseteq \bigcup_{u\in L}\text{int}_{L}(\bigcap_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap B_i^{-1}(u_i))\bigcap L)\subseteq\bigcup_{u\in L}\text{int}_{L}(\bigcap_{i\in I}B_i^{-1}(u_i)\bigcap L) , where L: = \prod_{i\in I}L_{N_{i}} .
\bullet there exists u_0\in X such that \text{cl}(X\setminus \bigcap_{i\in I}(B_i^{-1}(u_{0i})))\subseteq \text{cl}(X\setminus \bigcap_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap B_i^{-1}(u_{0i})))\subseteq K .
By combining (i) and the fact that A_i^{-1}(u_i) is open in X for every u_i\in X_i , we can see that all the hypotheses of Theorem 3.1 are satisfied. Thus, by Theorem 3.1, there exists \widehat{x}\in X such that \widehat{x}_i\in B_i(\widehat{x}) for every i\in I . Therefore, for each i\in I , \widehat{x}_i\in B_i(\widehat{x}) and W_i\cdot U^i(\widehat{x})\leq W_i\cdot U^i(\widehat{x}_{\widehat{i}}, x_i) for every x_i\in A_i(\widehat{x}) , which implies that \widehat{x}\in X is a generalized weighted Nash equilibrium of the game \Theta with respect to the weight vector W = (W_i)_{i\in I} . It follows from Lemma 4.1 that \widehat{x}\in X is also a generalized weak Pareto equilibrium of \Theta , and a generalized Pareto equilibrium of \Theta if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I .
Case II. Suppose that the set \mathfrak{F}_i = \{x\in X:\text{there exists}\ u_i\in A_i(x) \text{such that}\ W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is nonempty for every i\in I . Define a set-valued mapping Q_i:X\rightarrow 2^{X_i} by
Q_i(x) = \{u_i\in X_i:W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\}, \ \forall i\in I\ \text{and}\ x\in X. | (4.1) |
By (4.1), we get
x_i\notin Q_i(x), \ \forall i\in I\ \text{and}\ x\in X. | (4.2) |
Further, for each i\in I , we define two set-valued mappings S_i, T_i:X\rightarrow 2^{X_i} by setting, for each x\in X ,
\begin{eqnarray*} S_i(x) = \left\{ \begin{array}{ll} Q_i(x)\bigcap A_i(x), \ & \text{if}\ x\in \mathfrak{F}_i, \\ A_i(x), \ & \text{if}\ x\in X\setminus \mathfrak{F}_i, \end{array} \right. \end{eqnarray*} |
\begin{eqnarray*} \ \ \ \ \ T_i(x) = \left\{ \begin{array}{ll} Q_i(x)\bigcap B_i(x), \ & \text{if}\ x\in \mathfrak{F}_i, \\ B_i(x), \ & \text{if}\ x\in X\setminus \mathfrak{F}_i. \end{array} \right. \end{eqnarray*} |
It follows from (i), (ii), and the definitions of \mathfrak{F}_i and Q_i that S_i(x)\subseteq T_i(x) , T_i(x) is \Gamma_i -convex, and S_i(x)\neq\emptyset for every i\in I and every x\in X . For each i\in I and each u_i\in X_i , we have
\begin{eqnarray*} S_i^{-1}(u_i)& = &\bigg{\{}x\in X:u_i\in S_i(x)\bigg{\}}\\ & = &\bigg{\{}x\in \mathfrak{F}_i:u_i\in Q_i(x)\bigcap A_i(x)\bigg{\}}\bigcup\bigg{\{}x\in X\setminus \mathfrak{F}_i:u_i\in A_i(x)\bigg{\}}\\ & = &\bigg{(}(X\setminus \mathfrak{F}_i)\bigcap A_i^{-1}(u_i)\bigg{)}\bigcup\bigg{(}\mathfrak{F}_i\bigcap Q_i^{-1}(u_i)\bigcap A_i^{-1}(u_i)\bigg{)}\\ & = &\bigg{(}(X\setminus \mathfrak{F}_i)\bigcap A_i^{-1}(u_i)\bigg{)}\bigcup\bigg{(}Q_i^{-1}(u_i)\bigcap A_i^{-1}(u_i)\bigg{)}. \end{eqnarray*} |
Then by (iii), (iv), and the definition of Q_i , we can see that S_i^{-1}(u_i) is open in X . Similarly, we get
\begin{eqnarray*} T_i^{-1}(u_i)& = &\bigg{(}(X\setminus \mathfrak{F}_i)\bigcap B_i^{-1}(u_i)\bigg{)}\bigcup \bigg{(}Q_i^{-1}(u_i)\bigcap B_i^{-1}(u_i)\bigg{)}. \end{eqnarray*} |
Next, we show that (iv) of Theorem 3.1 is fulfilled. Indeed, by (v) and the expression of T_i^{-1}(u_i) , we can see that one of the following conditions holds:
\bullet for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that for L: = \prod_{i\in I}L_{N_{i}} , we have
\begin{eqnarray*} L\setminus K&\subseteq&\bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}\bigcap\limits_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap B_i^{-1}(u_i))\bigcap L\bigg{)}\\ &\subseteq& \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}\bigcap\limits_{i\in I}T_i^{-1}(u_i)\bigcap L\bigg{)}. \end{eqnarray*} |
\bullet there exists u_0\in X such that
\begin{eqnarray*} \text{cl}\bigg{(}X\setminus \bigcap\limits_{i\in I}T_i^{-1}(u_{0i})\bigg{)}&\subseteq&\text{cl}\bigg{(}X\setminus \bigcap\limits_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap B_i^{-1}(u_{0i}))\bigg{)}\\ &\subseteq& K. \end{eqnarray*} |
Thus, we can see that all the conditions of Theorem 3.1 are satisfied. Therefore, it follows from Theorem 3.1 that there exists \widehat{x}\in X such that \widehat{x}_i\in T_i(\widehat{x}) for every i\in I . If \widehat{x}_i\in \mathfrak{F}_i for some i\in I , then it follows from the definition of T_i that \widehat{x}_i\in Q_i(\widehat{x})\bigcap B_i(\widehat{x}) . Hence, \widehat{x}_i\in Q_i(\widehat{x}) , which contradicts (4.2). Therefore, we have \widehat{x}_i\in X\setminus \mathfrak{F}_i for every i\in I . By the definitions of Q_i , \mathfrak{F}_i , and T_i , we can deduce that for each i\in I , \widehat{x}_i\in B_i(\widehat{x}) and Q_i(\widehat{x})\cap A_i(\widehat{x}) = \emptyset , that is, for each i\in I , \widehat{x}_i\in B_i(\widehat{x}) and W_i\cdot U^i(\widehat{x})\leq W_i\cdot U^i(\widehat{x}_{\widehat{i}}, x_i) for every x_i\in A_i(\widehat{x}) , which implies that \widehat{x}\in X is a generalized weighted Nash equilibrium of the game \Theta with respect to the weight vector W = (W_i)_{i\in I} . By Lemma 4.1, one can see that \widehat{x}\in X is also a generalized weak Pareto equilibrium of \Theta , and a generalized Pareto equilibrium of \Theta if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I . This completes the proof.
Theorem 4.2. Let \Theta = ((X_i; \Gamma_i), U^i, A_i, B_i)_{i\in I} be a constrained multiobjective game such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is a compact abstract convex space defined as in Lemma 2.5, where I is a finite index set. For each i\in I , the graph of B_i is closed in X \times X_i and A_i is a continuous set-valued mapping such that each A_i(x) is a \Gamma_i -convex subset of X_i . Assume that there exists a weight vector W = (W_1, \ldots, W_n) with W_i\in \mathbb{R}_{+}^{k_i}\setminus \{0\} such that for each i\in I , the following conditions are satisfied:
(i) for each x\in X , \emptyset\neq A_i(x)\subseteq B_i(x) , and B_i(x) is \Gamma_i -convex;
(ii) for each u_i\in X_i , B_i^{-1}(u_i) is open in X ;
(iii) the function (x, u)\mapsto W_i\cdot U^i(x_{\widehat{i}}, u_i) is jointly continuous on X\times X ;
(iv) for each x\in X , the function u\mapsto W_i\cdot U^i(x_{\widehat{i}}, u_i) is quasi-convex on X .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the game \Theta has a generalized weighted Nash equilibrium \widehat{x}\in X with respect to the weight vector W = (W_i)_{i\in I} and hence it has a generalized weak Pareto equilibrium. Further, if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I , then \Theta has a generalized Pareto equilibrium.
Proof. For each m\in \mathbb{N} , define a set-valued mapping T_m:X\rightarrow 2^X as follows:
T_m(x) = \prod\limits_{i\in I}B_i(x)\bigcap\prod\limits_{i\in I}\bigg{(}\{u_i\in X_i:W_i\cdot U^i(x_{\widehat{i}}, u_i) < \text{min}_{y_i\in A_i(x)}W_i\cdot U^i(x_{\widehat{i}}, y_i)+\frac{1}{m}\}\bigg{)}, \ \forall x\in X. |
Thus, we have T_m(x) = \prod_{i\in I}\{u_i\in B_i(x):W_i\cdot U^i(x_{\widehat{i}}, u_i) < \text{min}_{y_i\in A_i(x)}W_i\cdot U^i(x_{\widehat{i}}, y_i)+\frac{1}{m}\} for every x\in X . By (i) and (iv), we can see that T_m(x) is a nonempty \Gamma -convex subset of X for every x\in X . Note that for each u\in X , we have
\begin{eqnarray*} T_m^{-1}(u)& = &\bigg{\{}x\in X:u\in T_m(x)\bigg{\}}\\ & = &\bigg{\{}x\in X:u\in \prod\limits_{i\in I}\{u_i\in B_i(x):W_i\cdot U^i(x_{\widehat{i}}, u_i) < \text{min}_{y_i\in A_i(x)}W_i\cdot U^i(x_{\widehat{i}}, y_i)+\frac{1}{m}\}\bigg{\}}\\ & = &\bigg{\{}x\in X:u_i\in B_i(x)\ \text{and }\ W_i\cdot U^i(x_{\widehat{i}}, u_i) < \text{min}_{y_i\in A_i(x)}W_i\cdot U^i(x_{\widehat{i}}, y_i)+\frac{1}{m}, \ \forall i\in I\bigg{\}}\\ & = &\bigg{(}\bigcap\limits_{i\in I}B_i^{-1}(u_i)\bigg{)}\bigcap\bigg{(}\bigcap\limits_{i\in I}\{x\in X:W_i\cdot U^i(x_{\widehat{i}}, u_i) < \text{min}_{y_i\in A_i(x)}W_i\cdot U^i(x_{\widehat{i}}, y_i)+\frac{1}{m}\}\bigg{)}. \end{eqnarray*} |
By (ii), (iii), and Lemma 4.2, we have that T_m^{-1}(u) is open in X for every u\in X . Therefore, by Theorem 3.2 with K = X and S = T , T_m has a fixed point x(m)\in X . Then it follows from the definition of T_m that W_i\cdot U^i(x_{\widehat{i}}(m), x_i(m)) < \text{min}_{y_i\in A_i(x(m))}W_i\cdot U^i(x_{\widehat{i}}(m), y_i)+\frac{1}{m} for every i\in I . Since X is compact, we may assume that x(m)\rightarrow \widehat{x}\in X without loss of generality. Since x_i(m)\in B_i(x(m)) and the graph of B_i is closed in X \times X_i , we have \widehat{x}_i\in B_i(\widehat{x}) . By (iii) and Lemma 4.2 again, we have
\begin{eqnarray*} W_i\cdot U^i(\widehat{x}_{\widehat{i}}, \widehat{x}_i)& = &\lim\limits_{m\rightarrow \infty}W_i\cdot U^i(x_{\widehat{i}}(m), x_i(m))\\ &\leq&\lim\limits_{m\rightarrow \infty}\text{min}_{y_i\in A_i(x(m))}W_i\cdot U^i(x_{\widehat{i}}(m), y_i)\\ & = &\text{min}_{y_i\in A_i(\widehat{x})}W_i\cdot U^i(\widehat{x}_{\widehat{i}}, y_i)\\ &\leq&\text{min}_{y_i\in B_i(\widehat{x})}W_i\cdot U^i(\widehat{x}_{\widehat{i}}, y_i). \end{eqnarray*} |
Since \widehat{x}_i\in B_i(\widehat{x}) for every i\in I , we have W_i\cdot U^i(\widehat{x}_{\widehat{i}}, \widehat{x}_i) = \text{min}_{y_i\in A_i(\widehat{x})}W_i\cdot U^i(\widehat{x}_{\widehat{i}}, y_i) , which implies that \widehat{x}\in X is a generalized weighted Nash equilibrium of the game \Theta with respect to the weight vector W = (W_i)_{i\in I} . By Lemma 4.1, we can see that \widehat{x}\in X is also a generalized weak Pareto equilibrium of \Theta , and a generalized Pareto equilibrium of \Theta if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I . This completes the proof.
Remark 4.4. Theorem 4.2 generalizes Theorem 2 due to Kim and Ding [41] in the following aspects: (a) from topological vector spaces to abstract convex spaces without any linear and convex structure; (b) the topological spaces in Theorem 4.2 need not possess Hausdorff property; (c) from constrained multiobjective games with one constrained set-valued mapping to constrained multiobjective games with two constrained set-valued mappings. Theorem 4.2 also generalizes Theorem 3.1 due to Wang [42] and Theorem 1 due to Yu and Yuan [44] to abstract convex spaces under much weaker assumptions.
If A_i = B_i for every i\in I , then by Theorems 4.1 and 4.2, we have the following two theorems.
Theorem 4.3. Let \Theta = ((X_i; \Gamma_i), U^i, A_i)_{i\in I} be a constrained multiobjective game such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K is a nonempty compact subset of X , where I is a finite index set. For each i\in I and each u_i\in X_i , A_i^{-1}(u_i) is open in X . Assume that there exists a weight vector W = (W_1, \ldots, W_n) with W_i\in \mathbb{R}_{+}^{k_i}\setminus \{0\} such that for each i\in I , the following conditions are satisfied:
(i) for each x\in X , A_i(x) is nonempty \Gamma_i -convex;
(ii) for each x\in X , the set \{u_i\in X_i:W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is \Gamma_i -convex;
(iii) for each u_i\in X_i , the set \{x\in X:W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is open in X ;
(iv) the set \mathfrak{F}_i = \{x\in X:\text{there exists}\ u_i\in A_i(x) \text{such that}\ W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is a nonempty closed subset of X ;
(v) one of the following conditions holds:
(v) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that L\setminus K\subseteq \bigcup_{u\in L}\text{int}_{L}(\bigcap_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap A_i^{-1}(u_i))\bigcap L) , where L: = \prod_{i\in I}L_{N_{i}} ;
(v) _2 there exists u_0\in X such that \text{cl}(X\setminus \bigcap_{i\in I}((X\setminus \mathfrak{F}_i)\bigcap A_i^{-1}(u_{0i})))\subseteq K .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the game \Theta has a generalized weighted Nash equilibrium \widehat{x}\in X with respect to the weight vector W = (W_i)_{i\in I} and hence it has a generalized weak Pareto equilibrium. Further, if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I , then \Theta has a generalized Pareto equilibrium.
Theorem 4.4. Let \Theta = ((X_i; \Gamma_i), U^i, A_i)_{i\in I} be a constrained multiobjective game such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is a compact abstract convex space defined as in Lemma 2.5, where I is a finite index set. For each i\in I , the graph of A_i is closed in X \times X_i . Assume that there exists a weight vector W = (W_1, \ldots, W_n) with W_i\in \mathbb{R}_{+}^{k_i}\setminus \{0\} such that for each i\in I , the following conditions are satisfied:
(i) for each x\in X , A_i(x) is nonempty \Gamma_i -convex;
(ii) for each u_i\in X_i , A_i^{-1}(u_i) is open in X ;
(iii) the function (x, u)\mapsto W_i\cdot U^i(x_{\widehat{i}}, u_i) is jointly continuous on X\times X ;
(iv) for each x\in X , the function u\mapsto W_i\cdot U^i(x_{\widehat{i}}, u_i) is quasi-convex on X .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the game \Theta has a generalized weighted Nash equilibrium \widehat{x}\in X with respect to the weight vector W = (W_i)_{i\in I} and hence it has a generalized weak Pareto equilibrium. Further, if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I , then \Theta has a generalized Pareto equilibrium.
Proof. It suffices to prove that A_i is a continuous set-valued mapping for every i\in I . In fact, since the graph of A_i is closed in X \times X_i and X_i is compact topological space for every i\in I , it follows from Lemma 2.1 that A_i is an upper semicontinuous set-valued mapping. We note that each A_i has open lower sections and so, A_i is a lower semicontinuous set-valued mapping. Therefore, A_i is a continuous set-valued mapping. Let A_i = B_i for every i\in I . Then by Theorem 4.2, the conclusion of Theorem 4.4 holds. This completes the proof.
By setting A_i(x)\equiv X_i for every i\in I and every x\in X , we have the following corollaries from Theorems 4.3-4.4. These two corollaries characterize the existence of weighted Nash equilibria for the multiobjective games without constrained set-valued mappings.
Corollary 4.1. Let \Theta = ((X_i; \Gamma_i), U^i)_{i\in I} be a multiobjective game such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K is a nonempty compact subset of X , where I is a finite index set. Assume that there exists a weight vector W = (W_1, \ldots, W_n) with W_i\in \mathbb{R}_{+}^{k_i}\setminus \{0\} such that for each i\in I , the following conditions are satisfied:
(i) for each x\in X , the set \{u_i\in X_i:W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is \Gamma_i -convex;
(ii) for each u_i\in X_i , the set \{x\in X:W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is open in X ;
(iii) the set \mathfrak{F}_i = \{x\in X:\text{there exists}\ u_i\in X_i \text{such that}\ W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is closed in X ;
(iv) one of the following conditions holds:
(iv) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that L\setminus K\subseteq \bigcup_{u\in L}\text{int}_{L}(\bigcap_{i\in I}(X\setminus \mathfrak{F}_i)\bigcap L) , where L: = \prod_{i\in I}L_{N_{i}} ;
(iv) _2 there exists u_0\in X such that \text{cl}(X\setminus \bigcap_{i\in I}(X\setminus \mathfrak{F}_i))\subseteq K .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the game \Theta has a weighted Nash equilibrium \widehat{x}\in X with respect to the weight vector W = (W_i)_{i\in I} and hence it has a weak Pareto equilibrium. Further, if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I , then \Theta has a Pareto equilibrium.
Remark 4.5. If \{(X_i; \Gamma_i)\}_{i\in I} is a family of abstract convex spaces such that X_i is a first-countable topological space for every i\in I , then (iii) of Corollary 4.1 can be replaced with the following condition:
(iii) ' for each i\in I , the graph of the set-valued mapping Q_i:X\rightarrow 2^{X_i} defined by Q_i(x) = \{u_i\in X_i:W_i \cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} for each x\in X , is closed in X\times X_i and for each compact subset Z\subseteq X , the set Q_i(Z) is compact subset of X_i .
In fact, let i\in I be fixed. For each x\in \text{cl}(\{x\in X:Q_i(x)\neq\emptyset\}) , since each X_i is a first-countable topological space, it follows that X = \prod_{i\in I}X_i is a first-countable topological space. By Theorem 2.40 due to Aliprantis and Border [21], there exists a sequence \{x_n\}_{n\in\mathbb{N}}\subseteq \{x\in X:Q_i(x)\neq\emptyset\} such that x_n\rightarrow x\in X . Thus, we have Q_i(x_n)\neq\emptyset and thus, for every n\in\mathbb{N} , there exists u_{in}\in X_i such that u_{in}\in Q_i(x_n) . Let L = \{x_n\}_{n\in\mathbb{N}}\cup \{x\} . Then by Theorem 2.38 due to Aliprantis and Border [21], L is compact subset of X . By (iii) ' , the set Q_i(L) = \cup_{x\in L}Q_i(x) is compact subset of X_i . Since \{u_{in}\}_{n\in \mathbb{N}}\subseteq Q_i(L) , it follows that \{u_{in}\}_{n\in \mathbb{N}} has a convergent subnet with limit u_i^* . Without loss of generality, we may assume that u_{in}\rightarrow u_i . Since the graph of Q_i is closed, we have u_i\in Q_i(x) , which implies that
x\in \{x\in X:Q_i(x)\neq\emptyset\}. |
Therefore, the set \{x\in X:Q_i(x)\neq\emptyset\} = \{x\in X:\text{there exists}\ u_i\in X_i\ \text{such that} W_i\cdot U^i(x_{\widehat{i}}, u_i) < W_i\cdot U^i(x_{\widehat{i}}, x_i)\} is closed in X .
Corollary 4.2. Let \Theta = ((X_i; \Gamma_i), U^i)_{i\in I} be a multiobjective game such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is a compact abstract convex space defined as in Lemma 2.5, where I is a finite index set. Assume that there exists a weight vector W = (W_1, \ldots, W_n) with W_i\in \mathbb{R}_{+}^{k_i}\setminus \{0\} such that for each i\in I , the following conditions are satisfied:
(i) the function (x, u)\mapsto W_i\cdot U^i(x_{\widehat{i}}, u_i) is jointly continuous on X\times X ;
(ii) for each x\in X , the function u\mapsto W_i\cdot U^i(x_{\widehat{i}}, u_i) is quasi-convex on X .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the game \Theta has a weighted Nash equilibrium \widehat{x}\in X with respect to the weight vector W = (W_i)_{i\in I} and hence it has a weak Pareto equilibrium. Further, if W_i\in\text{int}\mathbb{R}_{+}^{k_i} with \sum_{j = 1}^{k_i}W_{i, j} = 1 for every i\in I , then \Theta has a Pareto equilibrium..
Remark 4.6. Corollary 4.2 is different from Corollary 4.1 in the following aspects: (a) the topological spaces in Corollary 4.1 may be noncompact, while the topological spaces in Corollary 4.2 need to be compact; (b) (i) and (ii) of Corollary 4.1 are respectively weaker than (i) and (ii) of Corollary 4.2; (c) in order to guarantee the conclusion of Corollary 4.1 holds, the closeness condition of the set \mathfrak{F}_i and the coercive condition, that is, (iii) and (iv) of Corollary 4.1 must be satisfied, but Corollary 4.2 does not need theses conditions.
In this section, by using Theorems 3.1 and 3.2, we establish some new nonempty intersection theorems for sets with abstract convex sections. Furthermore, as applications of nonempty intersection property for sets with abstract convex sections, we obtain an analytic alternative formulation and two existence results of Nash equilibria for noncooperative games in noncompact abstract convex spaces.
Theorem 5.1. Let \{(X_i; \Gamma_i)\}_{i\in I} be a family of abstract convex spaces such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K = \prod_{i\in I}K_i is a nonempty compact subset of X , where I is a finite index set. For each i\in I , let P_i and Q_i be two subsets of X satisfying the following conditions:
(i) for each x_{\widehat{i}}\in X_{\widehat{i}} , \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}\subseteq \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\} and \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\} is \Gamma_i -convex;
(ii) for each u_i\in X_i , \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\} is open in X_{\widehat{i}} ;
(iii) for each x_{\widehat{i}}\in K_{\widehat{i}} , \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}\neq\emptyset ;
(iv) one of the following two conditions holds:
(iv) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} , such that for L: = \prod_{i\in I}L_{N_{i}} , we have
L\setminus K\subseteq \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in Q_i\}\times X_i))\bigcap L\bigg{)}; |
(iv) _2 there exists u_0 = (u_{0i})_{i\in I}\in X such that \text{cl}(X\setminus \bigcap_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_{0i})\in Q_i\}\times X_i))\subseteq K .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then \bigcap_{i\in I}Q_i\neq\emptyset .
Proof. For each i\in I , let us define two set-valued mappings S_i, T_i:X\rightarrow 2^{X_i} by S_i(x) = \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\} and T_i(x) = \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\} for every x = (x_i)_{i\in I}\in X . Then by (i), we have S_i(x)\subseteq T_i(x) and T_i(x) is \Gamma_i -convex for every i\in I and every x\in X . For each i\in I and each u_i\in X_i , we have S_i^{-1}(u_i) = \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\}\times X_i which is an open subset of X by (ii) and the definition of S_i . For each i\in I , it follows from (iii) and the definition of S_i that S_i(x)\neq\emptyset for every x\in K . Finally, we show that (iv) of Theorem 3.1 is fulfilled. Indeed, by (iv) and the fact that T_i^{-1}(u_i) = \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in Q_i\}\times X_i , one can see that one of the following conditions holds:
\bullet for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that for L: = \prod_{i\in I}L_{N_{i}} , we have
\begin{eqnarray*} L\setminus K&\subseteq&\bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in Q_i\}\times X_i))\bigcap L\bigg{)}\\ &\subseteq& \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}\bigcap\limits_{i\in I}T_i^{-1}(u_i)\bigcap L\bigg{)}. \end{eqnarray*} |
\bullet there exists u_0 = (u_{0i})_{i\in I}\in X such that
\begin{eqnarray*} \text{cl}\bigg{(}X\setminus \bigcap\limits_{i\in I}T_i^{-1}(u_{0i})\bigg{)}& = &\text{cl}(X\setminus \bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_{0i})\in Q_i\}\times X_i))\\ &\subseteq& K. \end{eqnarray*} |
Thus, we can see that all the conditions of Theorem 3.1 are satisfied. Therefore, it follows from Theorem 3.1 that there exists \widehat{x}\in X such that \widehat{x}_i\in T_i(\widehat{x}) = \{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in Q_i\} for every i\in I , that is, \widehat{x} = (\widehat{x}_{\widehat{i}}, \widehat{x}_i)\in Q_i for every i\in I and thus, \bigcap_{i\in I}Q_i\neq\emptyset . Our proof is complete.
Remark 5.1. Theorem 5.1 extends Theorem 7.1 in Park [18], Theorem 22 in Park [23], Theorem 4.15 in Bielawski [46], and Theorem 5.2 in Kirk et al. [47] to noncompact abstract convex spaces.
Theorem 5.2. Let \{(X_i; \Gamma_i)\}_{i\in I} be a family of abstract convex spaces such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K = \prod_{i\in I}K_i is a nonempty compact subset of X , where I is a finite index set. For each i\in I , let P_i and Q_i be two subsets of X satisfying the following conditions:
(i) for each x_{\widehat{i}}\in X_{\widehat{i}} , \Gamma\text{-}\text{co}(\{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\})\subseteq \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\} ;
(ii) for each u_i\in X_i , \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\} is open in X_{\widehat{i}} ;
(iii) for each x_{\widehat{i}}\in K_{\widehat{i}} , \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}\neq\emptyset ;
(iv) one of the following two conditions holds:
(iv) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} , such that for L: = \prod_{i\in I}L_{N_{i}} , we have
L\setminus K\subseteq \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\}\times X_i))\bigcap L\bigg{)}; |
(iv) _2 there exists u_0 = (u_{0i})_{i\in I}\in X such that \text{cl}(X\setminus \bigcap_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_{0i})\in P_i\}\times X_i))\subseteq K .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then \bigcap_{i\in I}Q_i\neq\emptyset .
Proof. For each i\in I , we define two set-valued mappings S_i, \widetilde{S}_i:X\rightarrow 2^{X_i} by S_i(x) = \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\} and \widetilde{S}_i(x) = \Gamma\text{-}\text{co}(\{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}) = \Gamma\text{-}\text{co}(S_i(x)) for every x = (x_i)_{i\in I}\in X . It is obvious that \Gamma\text{-}\text{co}(S_i(x)) is \Gamma_i -convex for all i\in I and all x = (x_i)_{i\in I}\in X . From (ii) and the definition of S_i , it follows that S_i^{-1}(u_i) = \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\}\times X_i is an open subset of X for every i\in I and every u_i\in X_i . Thus, by Lemma 2.8, \widetilde{S}_i^{-1}(u_i) is also an open subset of X for every i\in I and every u_i\in X_i . By (iii), we have \widetilde{S}_i(x)\supseteq S_i(x)\neq\emptyset for every i\in I and every x\in K . Since S_i^{-1}(u_i) = \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\}\times X_i\subseteq \widetilde{S}_i^{-1}(u_i) for every i\in I and every u_i\in X_i , it follows from (iv) that that one of the following conditions holds:
\bullet for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that for L: = \prod_{i\in I}L_{N_{i}} , we have
\begin{eqnarray*} L\setminus K&\subseteq&\bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\}\times X_i))\bigcap L\bigg{)}\\ & = &\bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}\bigcap\limits_{i\in I}S_i^{-1}(u_i)\bigcap L\bigg{)}\\ &\subseteq& \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}\bigcap\limits_{i\in I}\widetilde{S}_i^{-1}(u_i)\bigcap L\bigg{)}. \end{eqnarray*} |
\bullet there exists u_0 = (u_{0i})_{i\in I}\in X such that
\begin{eqnarray*} \text{cl}\bigg{(}X\setminus \bigcap\limits_{i\in I}\widetilde{S}_i^{-1}(u_{0i})\bigg{)}&\subseteq&\text{cl}\bigg{(}X\setminus \bigcap\limits_{i\in I}S_i^{-1}(u_{0i})\bigg{)}\\ & = &\text{cl}(X\setminus \bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_{0i})\in P_i\}\times X_i))\\ &\subseteq& K. \end{eqnarray*} |
Thus, we can see that all the conditions of Theorem 3.1 with S_i = T_i are satisfied. Therefore, we know that there exists \widehat{x}\in X such that \widehat{x}_i\in \widetilde{S}_i(\widehat{x}) = \Gamma\text{-}\text{co}(S_i(\widehat{x})) = \Gamma\text{-}\text{co}(\{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in P_i\}) for every i\in I . For this \widehat{x} , by (i), we have \widehat{x}_i\in \Gamma\text{-}\text{co}(\{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in P_i\})\subseteq \{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in Q_i\} for every i\in I , which implies that \widehat{x} = (\widehat{x}_{\widehat{i}}, \widehat{x}_i)\in Q_i for every i\in I . Therefore, we get \bigcap_{i\in I}Q_i\neq\emptyset . This completes the proof.
Remark 5.2. Except that the condition that the index set of Theorem 5.2 is finite is stronger than the condition that the index set of Theorem 16 due to Fan [48] is arbitrary, Theorem 5.2 partially generalizes Theorem 16 of Fan [48] in the following aspects: (a) from compact topological vector spaces to noncompact abstract convex spaces without any linear and convex structure; (b) there is no Hausdorff separation requirement for the abstract convex spaces involved Theorem 5.3. The topological vector spaces in Theorem 16 of Fan [48] need to meet the Hausdorff separation requirement because the continuous unity partition theory is used in the proof of this theorem; (c) even if we strengthen the abstract convex spaces in Theorem 5.2 to be topological vector spaces, (iii) of Theorem 5.2 is still weaker than the first half of (b) of Theorem 16 due to Fan [48].
Theorem 5.3. Let \{(X_i; \Gamma_i)\}_{i\in I} be a family of abstract convex spaces such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K = \prod_{i\in I}K_i is a nonempty compact subset of X , where I is a finite index set. For each i\in I , let P_i and Q_i be two subsets of X satisfying the following conditions:
(i) for each x_{\widehat{i}}\in X_{\widehat{i}} , \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}\subseteq \Gamma\text{-}\text{co}(\{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\}) ;
(ii) for each u_i\in X_i , \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\} is open in X_{\widehat{i}} ;
(iii) for each x_{\widehat{i}}\in K_{\widehat{i}} , \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}\neq\emptyset ;
(iv) one of the following two conditions holds:
(iv) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} , such that for L: = \prod_{i\in I}L_{N_{i}} , we have
L\setminus K\subseteq \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in Q_i\}\times X_i))\bigcap L\bigg{)}; |
(iv) _2 there exists u_0 = (u_{0i})_{i\in I}\in X such that \text{cl}(X\setminus \bigcap_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_{0i})\in Q_i\}\times X_i))\subseteq K .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then there exists \widehat{x}\in X such that \widehat{x}_i\in \Gamma\text{-}\text{co}(\{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in Q_i\}) for every i\in I .
Proof. For each i\in I , define two set-valued mappings S_i, T_i:X\rightarrow 2^{X_i} by S_i(x) = \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\} and T_i(x) = \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\} for every x = (x_i)_{i\in I}\in X . Then it is easy to verify that S_i and T_i satisfy all the requirements of Theorem 3.2. Therefore, by Theorem 3.2, there exists \widehat{x}\in X such that \widehat{x}_i\in \Gamma\text{-}\text{co}(T_i(\widehat{x})) = \Gamma\text{-}\text{co}(\{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in Q_i\}) for every i\in I . This completes the proof.
Remark 5.3. We can compare Theorem 5.3 and Theorem 2.3 obtained by Lan and Webb [2] from the following aspects: (a) Theorem 5.3 is based on noncompact abstract convex spaces without any linear and convex structure. The Hausdorffness of the abstract convex spaces involved Theorem 5.3 is redundant. Theorem 2.3 due to by Lan and Webb [2] is established in the framework of Hausdorff topological vector spaces; (b) Theorem 5.3 has two coercive conditions to be available, and Theorem 2.3 obtained by Lan and Webb [2] has only one coercive condition; (c) there are two families of subsets of X in Theorem 5.3. In Theorem 2.3 obtained by Lan and Webb [2], there is only one family of subsets of X ; (d) even the abstract convex spaces in Theorem 5.3 are strengthened to be topological vector spaces, (iii) of Theorem 5.3 is weaker than ( S_1 ) of Theorem 2.3 due to Lan and Webb [2]; (e) Theorem 5.3 deals with nonempty intersection of finite number of sets with abstract convex sections, and Theorem 2.3 in Lan and Webb [2] concerns on nonempty intersection of arbitrary number of sets with convex sections.
Theorem 5.4. Suppose that all the requirements of Theorem 5.3 are satisfied. For each i\in I , let V_i be a subset of X such that for each x\in X , there is a subset I(x) of I such that \Gamma\text{-}\text{co}(\{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\})\subseteq \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in V_i\} for every i\in I(x) . Then there exists \widehat{x}\in X such that \bigcap_{i\in I(\widehat{x})}V_i\neq\emptyset .
Proof. By Theorem 5.3, there exists \widehat{x}\in X such that \widehat{x}_i\in \Gamma\text{-}\text{co}(\{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in Q_i\}) for every i\in I . Therefore, for this \widehat{x} , we have \widehat{x}_i\in \{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in V_i\} for every i\in I(\widehat{x}) , which implies that there exists a point \widehat{x}\in X such that \bigcap_{i\in I(\widehat{x})}V_i\neq\emptyset . This completes the proof.
Now, we present the following analytical formulation of Theorem 5.3.
Theorem 5.5. Let \{(X_i; \Gamma_i)\}_{i\in I} be a family of abstract convex spaces such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K = \prod_{i\in I}K_i is a nonempty compact subset of X , where I is a finite index set. For each i\in I , let \xi_i , \rho_i , \upsilon_i:X\rightarrow \mathbb{R} be three real-valued functions and let t_i be a real number satisfying the following conditions:
(i) for each x\in X , \xi_i(x)\leq\rho_i(x)\leq\upsilon_i(x) ;
(ii) for each u_i\in X_i , \xi_i(., u_i) is lower semicontinuous on X_{\widehat{i}} ;
(iii) for each x_{\widehat{i}}\in X_{\widehat{i}} , \upsilon_i(x_{\widehat{i}}, .) is quasiconcave on X_i ;
(iv) one of the following two conditions holds:
(iv) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} , such that for L: = \prod_{i\in I}L_{N_{i}} , we have
L\setminus K\subseteq \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\rho_i(x_{\widehat{i}}, u_i) > t_i\}\times X_i))\bigcap L\bigg{)}; |
(iv) _2 there exists u_0 = (u_{0i})_{i\in I}\in X such that \text{cl}(X\setminus \bigcap_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\rho_i(x_{\widehat{i}}, u_{0i}) > t_i\}\times X_i))\subseteq K .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then either there exist an i\in I and an x_{\widehat{i}}\in K_{\widehat{i}} such that \xi_i(x_{\widehat{i}}, y_i)\leq t_i for every y_i\in X_i or there exists \widehat{x}\in X such that \upsilon_i(\widehat{x}) > t_i for every i\in I .
Proof. Suppose that for each i\in I and each x_{\widehat{i}}\in K_{\widehat{i}} , there is y_i\in X_i satisfying \xi_i(x_{\widehat{i}}, y_i) > t_i . For each i\in I , we define P_i = \{x\in X:\xi_i(x) > t_i\} , Q_i = \{x\in X:\rho_i(x) > t_i\} , and V_i = \{x\in X:\upsilon_i(x) > t_i\} . Then by (i), for each i\in I and each x_{\widehat{i}}\in X_{\widehat{i}} , we have
\begin{eqnarray*} \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}&\subseteq&\{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\}\\ &\subseteq&\Gamma\text{-}\text{co}(\{y_i\in X_i:(x_{\widehat{i}}, y_i)\in Q_i\}). \end{eqnarray*} |
By (ii), it follows that the set \{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in P_i\} is an open subset of X_{\widehat{i}} for every u_i\in X_i . From the beginning of the proof, we can see that \{y_i\in X_i:(x_{\widehat{i}}, y_i)\in P_i\}\neq\emptyset for all i\in I and all x_{\widehat{i}}\in K_{\widehat{i}} . By (iv), one of the following two conditions holds:
\bullet for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} such that for L: = \prod_{i\in I}L_{N_{i}} , we have
\begin{eqnarray*} L\setminus K&\subseteq&\bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\rho_i(x_{\widehat{i}}, u_i) > t_i\}\times X_i))\bigcap L\bigg{)}\\ & = &\bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_i)\in Q_i\}\times X_i))\bigcap L\bigg{)}. \end{eqnarray*} |
\bullet there exists u_0 = (u_{0i})_{i\in I}\in X such that
\begin{eqnarray*} K&\supseteq&\text{cl}(X\setminus \bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\rho_i(x_{\widehat{i}}, u_{0i}) > t_i\}\times X_i))\\ & = &\text{cl}(X\setminus \bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:(x_{\widehat{i}}, u_{0i})\in Q_i\}\times X_i)). \end{eqnarray*} |
Therefore, it follows from Theorem 5.3 that there exists there exists \widehat{x}\in X such that \widehat{x}_i\in \Gamma\text{-}\text{co}(\{y_i\in X_i:(\widehat{x}_{\widehat{i}}, y_i)\in Q_i\}) for every i\in I . By (iii) and the fact that \rho_i(x)\leq\upsilon_i(x) for every x\in X , we have \widehat{x}\in\bigcap_{i\in I}V_i , which implies that there exists \widehat{x}\in X such that \upsilon_i(\widehat{x}) > t_i for every i\in I . The proof is finished.
Remark 5.4. Theorem 5.5 generalizes Theorem 8.1 of Park [18] in the following two aspects: (a) from compact abstract convex spaces to noncompact abstract convex spaces; (b) from two families of real-valued functions to three families of real-valued functions.
Theorem 5.6. Let \{(X_i; \Gamma_i)\}_{i\in I} be a family of abstract convex spaces such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is the abstract convex space defined as in Lemma 2.5 and K = \prod_{i\in I}K_i is a nonempty compact subset of X , where I is a finite index set. For each i\in I , let \xi_i , \rho_i , \upsilon_i:X\rightarrow \mathbb{R} be three real-valued functions satisfying the following conditions:
(i) for each x\in X , \xi_i(x)\leq\rho_i(x)\leq\upsilon_i(x) ;
(ii) for each u_i\in X_i , \xi_i(., u_i) is lower semicontinuous on X_{\widehat{i}} ;
(iii) for each x_{\widehat{i}}\in X_{\widehat{i}} , \upsilon_i(x_{\widehat{i}}, .) is quasiconcave on X_i ;
(iv) for each x_{\widehat{i}}\in X_{\widehat{i}} , \xi_i(x_{\widehat{i}}, .) is bounded on X_{i} and for any \varepsilon > 0 , suppose that one of the following two conditions holds:
(iv) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} , such that for L: = \prod_{i\in I}L_{N_{i}} , we have
L\setminus K\subseteq \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\rho_i(x_{\widehat{i}}, u_i) > \sup\limits_{y_i\in X_i}\xi_i(x_{\widehat{i}}, y_i)-\varepsilon\}\times X_i))\bigcap L\bigg{)}; |
(iv) _2 there exists u_0 = (u_{0i})_{i\in I}\in X such that
\text{cl}(X\setminus \bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\rho_i(x_{\widehat{i}}, u_{0i}) > \sup\limits_{y_i\in X_i}\xi_i(x_{\widehat{i}}, y_i)-\varepsilon\}\times X_i))\subseteq K. |
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then there exists \widehat{x}^{\varepsilon} = (\widehat{x}_{\widehat{i}}^{\varepsilon}, \widehat{x}_{i}^{\varepsilon})\in X such that \upsilon_i(\widehat{x}^{\varepsilon}) > \sup_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}^{\varepsilon}, y_i)-\varepsilon for every i\in I .
Proof. Set t_i: = \sup_{y_i\in X_i}\xi_i(x_{\widehat{i}}, y_i)-\varepsilon\in \mathbb{R} for all i\in I and all x_{\widehat{i}}\in X_{\widehat{i}} . Then it is easy to see that for each i\in I and each x_{\widehat{i}}\in X_{\widehat{i}} , there exists y_i\in X_i such that \xi_i(x_{\widehat{i}}, y_i) > t_i . Thus, it follows from Theorem 5.5 that there exists \widehat{x}^{\varepsilon} = (\widehat{x}_{\widehat{i}}^{\varepsilon}, \widehat{x}_{i}^{\varepsilon})\in X such that \upsilon_i(\widehat{x}^{\varepsilon}) > t_i = \sup_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}^{\varepsilon}, y_i)-\varepsilon for every i\in I . This completes the proof.
Remark 5.5. {Under the conditions of Theorem 9.1 due to Park [18], only the conclusion similar to that of Theorem 5.6 can be obtained. This is because \widehat{x}\in X varies with \varepsilon and the conditions of Theorem 9.1 in Park [18] are not sufficient to guarantee the continuity of the function x_{\widehat{i}}\mapsto\max_{y_i\in X_i}f_i(x_{\widehat{i}}, y_i) . Thus, from this perspective, Theorem 5.6 generalizes Theorem 9.1 of Park [18] in the following aspects:} (a) from compact abstract convex spaces to noncompact abstract convex spaces; (b) from two families of real-valued functions to three families of real-valued functions; (c) the condition that \xi_i(x_{\widehat{i}}, .) is bounded on X_{i} for every x_{\widehat{i}}\in X_{\widehat{i}} , is weaker than (9.2) of Theorem 9.1 due to Park [18].
From Theorem 5.6 for \xi_i = \rho_i = \upsilon_i , we can derive the following existence theorem of \varepsilon -Nash equilibria for noncooperative games in noncompact abstract convex spaces.
Theorem 5.7. Let \{(X_i; \Gamma_i)\}_{i\in I} be a family of abstract convex spaces such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5 and K = \prod_{i\in I}K_i is a nonempty compact subset of X , where I is a finite index set. For each i\in I , let \xi_i:X\rightarrow \mathbb{R} be a real-valued function satisfying the following conditions:
(i) for each u_i\in X_i , \xi_i(., u_i) is lower semicontinuous on X_{\widehat{i}} ;
(ii) for each x_{\widehat{i}}\in X_{\widehat{i}} , \xi_i(x_{\widehat{i}}, .) is quasiconcave on X_i ;
(iii) for each x_{\widehat{i}}\in X_{\widehat{i}} , \xi_i(x_{\widehat{i}}, .) is bounded on X_{i} and for any \varepsilon > 0 , suppose that one of the following two conditions holds:
(iii) _1 for each N_{i}\in \langle X_i\rangle , there exists a compact \Gamma_i -convex subset L_{N_{i}} of (X_i; \Gamma_i) containing N_{i} , such that for L: = \prod_{i\in I}L_{N_{i}} , we have
L\setminus K\subseteq \bigcup\limits_{u\in L}\text{int}_{L}\bigg{(}(\bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\xi_i(x_{\widehat{i}}, u_i) > \sup\limits_{y_i\in X_i}\xi_i(x_{\widehat{i}}, y_i)-\varepsilon\}\times X_i))\bigcap L\bigg{)}; |
(iii) _2 there exists u_0 = (u_{0i})_{i\in I}\in X such that
\text{cl}(X\setminus \bigcap\limits_{i\in I}(\{x_{\widehat{i}}\in X_{\widehat{i}}:\xi_i(x_{\widehat{i}}, u_{0i}) > \sup\limits_{y_i\in X_i}\xi_i(x_{\widehat{i}}, y_i)-\varepsilon\}\times X_i))\subseteq K. |
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then there exists \widehat{x}^{\varepsilon} = (\widehat{x}_{\widehat{i}}^{\varepsilon}, \widehat{x}_{i}^{\varepsilon})\in X such that \xi_i(\widehat{x}^{\varepsilon}) > \sup_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}^{\varepsilon}, y_i)-\varepsilon for every i\in I .
By using a special case of Theorem 5.7, we have the following existence theorem of Nash equilibria for noncooperative games in compact abstract convex spaces.
Corollary 5.1. Let \{(X_i; \Gamma_i)\}_{i\in I} be a family of compact abstract convex spaces such that (X; \Gamma): = (\prod_{i\in I}X_i; \Gamma) is an abstract convex space defined as in Lemma 2.5, where I is a finite index set. For each i\in I , let \xi_i:X\rightarrow \mathbb{R} be a real-valued function such that:
(i) \xi_i is upper semicontinuous on X ;
(ii) for each u_i\in X_i , \xi_i(., u_i) is lower semicontinuous on X_{\widehat{i}} ;
(iii) for each x_{\widehat{i}}\in X_{\widehat{i}} , \xi_i(x_{\widehat{i}}, .) is quasiconcave on X_i .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then there exists \widehat{x}\in X such that \xi_i(\widehat{x}) = \max_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}, y_i) for every i\in I .
Proof. Let \varepsilon > 0 . Then by Theorem 5.7 with each X_i being a compact abstract convex space, it follows there exists \widehat{x}^{\varepsilon} = (\widehat{x}_{\widehat{i}}^{\varepsilon}, \widehat{x}_{i}^{\varepsilon})\in X such that \xi_i(\widehat{x}^{\varepsilon}) > \max_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}^{\varepsilon}, y_i)-\varepsilon . Let \varepsilon\rightarrow 0 . By the compactness of X and \{\widehat{x}^{\varepsilon}\}\subseteq X , we assume that \widehat{x}^{\varepsilon}\rightarrow \widehat{x} without loss of generality. By (i) and (ii), it follows from Lemma 2 of Yu and Yuan [44] that the function x_{\widehat{i}}\mapsto\max_{y_i\in X_i}\xi_i(x_{\widehat{i}}, y_i) is continuous. Immediately using (i) again, we get \xi_i(\widehat{x}_{\widehat{i}}, \widehat{x}_{i})\geq\varlimsup_{\varepsilon\rightarrow 0}\xi_i(\widehat{x}_{\widehat{i}}^{\varepsilon}, \widehat{x}_{i}^{\varepsilon})\geq\varlimsup_{\varepsilon\rightarrow 0}\max_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}^{\varepsilon}, y_i) = \max_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}, y_i) . Thus, we have \xi_i(\widehat{x}) = \max_{y_i\in X_i}\xi_i(\widehat{x}_{\widehat{i}}, y_i) for every i\in I . This completes the proof.
In this section, we use Theorem 3.4 to establish some existence results of solutions for generalized weak implicit inclusion problems in noncompact abstract convex spaces. We first formulate the problems in the following.
Let (X; \Gamma^1) and (Y; \Gamma^2) be two abstract convex spaces and let Z be a nonempty set. Let A, B:X\rightarrow 2^X , F:X\rightarrow 2^Y , G:X\rightarrow 2^Z , and H:Y\times Z\rightarrow 2^X be five set-valued mappings. We consider the F-generalized weak implicit inclusion problem denoted by (FGWIIP): find (\widehat{x}, \widehat{y})\in X\times Y such that \widehat{x}\in A(\widehat{x}) , \widehat{y}\in F(\widehat{x}) , and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{y}, z) and the S-generalized weak implicit inclusion problem denoted by (SGWIIP): find (\widehat{x}, \widehat{y})\in X\times Y such that \widehat{x}\in A(\widehat{x}) , \widehat{y}\in F(\widehat{x}) , and u\in H(\widehat{y}, z) for every u\in B(\widehat{x}) and every z\in G(\widehat{x}) . If X = Z and G is the identity mapping on X , then (FGWIIP) coincides with (SGWIIP).
Note that if X = Y and F is the identity mapping on X , then (FGWIIP) reduces to the generalized weak implicit inclusion problem denoted by (GWIIP): find \widehat{x}\in X such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{x}, z) . If A(x) = B(x)\equiv X for every x\in X , then (GWIIP) reduces to the generalized implicit inclusion problem denoted by (GIIP): find \widehat{x}\in X such that for each u\in X , there exists z\in G(\widehat{x}) : u\in H(\widehat{x}, z) , which was discussed by Wang and Huang [49] under the condition that X and Z are two Hausdorff topological vector spaces. If X = Z and G is the identity mapping on X , then (GWIIP) reduces to the extended weak inclusion problem denoted by (EWIP): find \widehat{x}\in X such that \widehat{x}\in A(\widehat{x}) and B(\widehat{x})\subseteq H(\widehat{x}, \widehat{x}) . If A(x) = B(x)\equiv X for every x\in X , then (EWIP) reduces to the extended inclusion problem (for short, EIP): find \widehat{x}\in X such that X\subseteq H(\widehat{x}, \widehat{x}) , which was studied by Fang and Huang [50] under the condition that X is a real Banach space. If X = Z , G is the identity mapping on X , and H(x, z) = H(x) for every (x, z)\in X\times X , then (GIIP) reduces to the inclusion problem denoted by (IP): find \widehat{x}\in X such that X\subseteq H(\widehat{x}) , which was investigated by Di Bella [51] when X is a Hausdorff topological vector space.
From these special cases, we can see that (FGWIIP) extends and unifies the corresponding models in [49,50,51].
Definition 6.1. Let (X; \Gamma) be an abstract convex space and let Y and Z be two nonempty sets. Let F:X\rightarrow 2^Y and G:X\rightarrow 2^Z be two set-valued mappings. A set-valued mapping H:Y\times Z\rightarrow 2^X is said to be \Gamma -quasiconvex-like with respect to F and G if for each N = \{u_{0}, u_{1}, \ldots, u_{n}\}\in \langle X\rangle , each x\in \Gamma\text{-}\text{co}(N) , and for each y\in F(x) , there exist j\in \{0, 1, \ldots, n\} and z\in G(x) such that u_j\in H(y, z) .
Definition 6.2. Let (X; \Gamma) be an abstract convex space and let Y and Z be two nonempty sets. Let F:X\rightarrow 2^Y and G:X\rightarrow 2^Z be two set-valued mappings. A set-valued mapping H:Y\times Z\rightarrow 2^X is said to be strong \Gamma -quasiconvex-like with respect to F and G if for each N = \{u_{0}, u_{1}, \ldots, u_{n}\}\in \langle X\rangle and for each x\in \Gamma\text{-}\text{co}(N) and each y\in F(x) , there exists j\in \{0, 1, \ldots, n\} such that u_j\in H(y, z) for every z\in G(x) .
Definition 6.3. Let (X; \Gamma) be an abstract convex space, Y be a topological vector space, C\subseteq Y be a nonempty convex cone, and \eta:X\times X\rightarrow X be a single-valued mapping. A set-valued mapping F:X\rightarrow 2^Y is said to be C - \Gamma -quasiconvex in the second argument of \eta if for each x\in X , each A = \{y_0, y_1, \ldots, y_n\}\in \langle X\rangle and each z\in \Gamma(A) , there exists j\in \{0, 1, \ldots, n\} such that F(\eta(x, z))\subseteq F(\eta(x, y_j))-C .
Theorem 6.1. Let (X; \Gamma^1) and (Y; \Gamma^2) be two abstract convex spaces such that (X\times Y; \Gamma^1\times \Gamma^2) is an abstract convex space defined as in Lemma 2.5. Let K be a nonempty compact subset of X\times Y and Z be a nonempty set. Let A, B:X\rightarrow 2^X , F:X\rightarrow 2^Y , G:X\rightarrow 2^Z , and H:Y\times Z\rightarrow 2^X be five set-valued mappings satisfying
(i) for each x\in X , B(x)\subseteq A(x) ;
(ii) B and F have nonempty \Gamma^1 -convex and \Gamma^2 -convex values and open lower sections;
(iii) the set \mathfrak{F} = \{(x, y)\in X\times Y:x\in A(x)\ \text{and}\ y\in F(x)\} is closed in X\times Y ;
(iv) for each u\in X , the set \{(x, y)\in X\times Y:u\not\in H(y, z)\ \text{for every}\ z\in G(x)\} is open in X\times Y ;
(v) for each x\in X and each y\in F(x) , x\not\in \Gamma^1\text{-}\text{co}(\{u\in X:u\not\in H(y, z)\ \text{for every}\ z\in G(x)\}) ;
(vi) one of the following conditions holds:
(vi) _1 for each N_{0}\times N_{1}\in \langle X\times Y\rangle , there exist a compact \Gamma^1 -convex subset L_{N_{0}} of (X; \Gamma^1) containing N_{0} and a compact \Gamma^2 -convex subset L_{N_{1}} of (Y; \Gamma^2) containing N_{1} such that for L: = L_{N_{0}}\times L_{N_{1}} and for each (x, y)\in L\setminus K , there exists (u, v)\in L such that u\in B(x) , v\in F(x) , and u\not\in H(y, z) for every z\in G(x) ;
(vi) _2 there exists (u_0, v_0)\in X\times Y such that for each (x, y)\in X\times Y\setminus K , one has u_0\in B(x) , v_0\in F(x) , and u_0\not\in H(y, z) for every z\in G(x) .
If (X\times Y; \Gamma^1\times \Gamma^2) satisfies 1_{X\times Y}\in{\frak{RC}}(X\times Y, X\times Y) , then (FGWIIP) is solvable, that is, there exists (\widehat{x}, \widehat{y})\in K such that \widehat{x}\in A(\widehat{x}) , \widehat{y}\in F(\widehat{x}) , and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{y}, z) .
Proof. By (ii), B has nonempty \Gamma^1 -convex values and open lower sections. Let \pi(K) denotes the projection of K onto X . Then it is clear that \pi(K) is a compact subset of X . For each N_{0}\times N_{1}\in \langle X\times Y\rangle , it follows from (vi) _1 that there exist a compact \Gamma^1 -convex subset L_{N_{0}} of (X; \Gamma^1) containing N_{0} and a compact \Gamma^2 -convex subset L_{N_{1}} of (Y; \Gamma^2) containing N_{1} . Let x\in L_{N_0}\setminus \pi(K) and y\in L_{N_1} be given arbitrarily. Then we have (x, y)\in L\setminus K . By (vi) _1 again, for each x\in L_{N_0}\setminus \pi(K) , there exists u\in L_{N_0} such that u\in B(x) , which implies that L_{N_0}\setminus \pi(K)\subseteq \bigcup_{u\in L_{N_0}}(B^{-1}(u)\bigcap L_{N_0}) . Similarly, let x\in X\setminus \pi(K) and y\in Y be any given. Then we have (x, y)\in X\times Y\setminus K and thus, by (vi) _2 , there exists u_0\in X such that for each x\in X\setminus \pi(K) , we have u_0\in B(x) , which implies that X\setminus B^{-1}(u_{0})\subseteq \pi(K) . Therefore, all the conditions of Corollary 3.1 with S = T are fulfilled and thus, it follows that there exists x_0\in X such that x_0\in B(x_0)\subseteq A(x_0) . Then we have x_0\times F(x_0)\subseteq \mathfrak{F} and hence, the set \mathfrak{F} is nonempty.
Define a set-valued mapping T:X\times Y\rightarrow 2^{X\times Y} by setting, for each (x, y)\in X\times Y ,
\begin{eqnarray*} \ \ \ \ \ T(x, y) = \left\{ \begin{array}{ll} (B(x)\bigcap J(x, y))\times F(x), \ & \text{if}\ (x, y)\in \mathfrak{F}, \\ B(x)\times F(x), \ & \text{if}\ (x, y)\in X\times Y\setminus \mathfrak{F}, \end{array} \right. \end{eqnarray*} |
where J:X\times Y\rightarrow 2^X is defined by J(x, y) = \{u\in X:u\not\in H(y, z)\ \text{for every}\ z\in G(x)\} for every (x, y)\in X\times Y . For each (u, v)\in X\times Y , we have
\begin{eqnarray*} T^{-1}(u, v)& = &\bigg{(}(X\times Y\setminus \mathfrak{F})\bigcap (B^{-1}(u)\times Y)\bigcap (F^{-1}(v)\times Y)\bigg{)}\\ &\bigcup& \bigg{(}J^{-1}(u)\bigcap (B^{-1}(u)\times Y)\bigcap (F^{-1}(v)\times Y)\bigg{)}. \end{eqnarray*} |
Since J^{-1}(u) = \{(x, y)\in X\times Y:u\not\in H(y, z)\ \text{for every}\ z\in G(x)\} for every u\in X , it follows from (iv) that J^{-1}(u) is open in X\times Y . By (ii) and (iii), one can see that T^{-1}(u, v) is open in X\times Y for every (u, v)\in X\times Y . By (v) and the definition of J , we have
(x, y)\not\in \Gamma^1\text{-}\text{co}(B(x)\bigcap J(x, y))\times F(x) = \Gamma^1\text{-}\text{co}(B(x)\bigcap J(x, y))\times \Gamma^2\text{-}\text{co}(F(x)), \ \forall (x, y)\in \mathfrak{F}. |
Since \Gamma^1\text{-}\text{co}(B(x)\bigcap J(x, y))\times \Gamma^2\text{-}\text{co}(F(x)) is a \Gamma^1\times \Gamma^2 -convex subset, (B(x)\bigcap J(x, y))\times F(x)\subseteq \Gamma^1\text{-}\text{co}(B(x)\bigcap J(x, y))\times \Gamma^2\text{-}\text{co}(F(x)) , and \Gamma^1\times\Gamma^2\text{-}\text{co}((B(x)\bigcap J(x, y))\times F(x)) is the smallest \Gamma^1\times \Gamma^2 -convex subset containing (B(x)\bigcap J(x, y))\times F(x) , it follows that (x, y)\not\in \Gamma^1\times\Gamma^2\text{-}\text{co}((B(x)\bigcap J(x, y))\times F(x)) for every (x, y)\in \mathfrak{F} . It is easy to see that (x, y)\not\in B(x)\times F(x) = \Gamma^1\times\Gamma^2\text{-}\text{co}(B(x)\times F(x)) for every (x, y)\in X\times Y\setminus \mathfrak{F} . Therefore, in both two cases, we have (x, y)\not\in \Gamma^1\times\Gamma^2\text{-}\text{co}(T(x, y)) for every (x, y)\in X\times Y . By (vi), we know that one of the following two conditions holds:
\bullet for each N_{0}\times N_{1}\in \langle X\times Y\rangle , there exist a compact \Gamma^1 -convex subset L_{N_{0}} of (X; \Gamma^1) containing N_{0} and a compact \Gamma^2 -convex subset L_{N_{1}} of (Y; \Gamma^2) containing N_{1} such that for L: = L_{N_{0}}\times L_{N_{1}} , we have L\setminus K\subseteq \bigcup_{(u, v)\in L}T^{-1}(u, v) .
\bullet there exists (u_0, v_0)\in X\times Y such that} X\times Y\setminus T^{-1}(u_{0}, v_{0})\subseteq K .
Thus, by Theorem 3.4 and Remark 3.4, there exists (\widehat{x}, \widehat{y})\in K such that T(\widehat{x}, \widehat{y}) = \emptyset . Since B and F have nonempty values, we can conclude that (\widehat{x}, \widehat{y})\in \mathfrak{F} . Thus, \widehat{x}\in A(\widehat{x}) , \widehat{y}\in F(\widehat{x}) , and B(\widehat{x})\bigcap J(\widehat{x}, \widehat{y}) = \emptyset . Therefore, for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{y}, z) . This completes the proof.
Remark 6.1. (1) (v) of Theorem 6.1 can be replaced by the following stronger condition:
(v) ' H is \Gamma^1 -quasiconvex-like with respect to F and G .
In fact, suppose to the contrary that there exist x\in X and y\in F(x) such that x\in \Gamma^1\text{-}\text{co}(\{u\in X:u\not\in H(y, z) \text{for every}\ z\in G(x)\}) . Then by Lemma 2.7, there exists \{u_0, u_1, \ldots, u_n\}\in \langle\{u\in X:u\not\in H(y, z) \text{for every}\ z\in G(x)\}\rangle such that x\in\Gamma^1\text{-}\text{co}(\{u_0, u_1, \ldots, u_n\}) . By (v) ' , there exists j\in \{0, 1, \ldots, n\} and z\in G(x) such that u_j\in H(y, z) , which contradicts that u_j\not\in H(y, z) for every z\in G(x) . Therefore, (v) ' implies (v) of Theorem 6.1.
(2) the following two conditions imply that (v) ' holds.
(a) for each x\in X and each y\in F(x) , the set \{u\in X:u\not\in H(y, z)\ \text{for every}\ z\in G(x)\} is \Gamma^1 -convex.
(b) for each x\in X and each y\in F(x) , there exists z\in G(x) such that x\in H(y, z) .
Indeed, by way of contradiction, suppose that for some N = \{u_{0}, u_{1}, \ldots, u_{n}\}\in \langle X\rangle , some x\in \Gamma\text{-}\text{co}(N) , and for some y\in F(x) , u_j\not \in H(y, z) for every j\in \{0, 1, \ldots, n\} and every z\in G(x) . By (a), we have x\not\in H(y, z) , which contradicts (b).
(3) If we assume that X has Hausdorff property and Z is a topological space, then (iv) of Theorem 6.1 can be replaced by the following condition:
(iv) ' G and H are two upper semicontinuous set-valued mappings with compact values.
In fact, it suffices to prove that the set \{(x, y)\in X\times Y:\text{there exists}\ z\in G(x) \text{such that}\ u\in H(y, z)\} is closed in X\times Y for every u\in X . Let \{(x_\alpha, y_\alpha)\}\subseteq \{(x, y)\in X\times Y:\text{there exists}\ z\in G(x) \text{such that}\ u\in H(y, z)\} be an arbitrary net such that (x_\alpha, y_\alpha)\rightarrow (x_0, y_0) . Then for each \alpha , there exists z_\alpha\in G(x_\alpha) such that u\in H(y_\alpha, z_\alpha) . Since G is an upper semicontinuous set-valued mapping with compact values, it follows from Lemma 2.4 that there exist z_0\in G(x_0) and a subnet \{z_\beta\} of \{z_\alpha\} such that z_\beta\rightarrow z_0 . Since H is an upper semicontinuous set-valued mapping with compact values, it follows from Lemma 2.3 that H is closed. In addition, for each \beta , we have u\in H(y_\beta, z_\beta) and (y_\beta, z_\beta)\rightarrow (y_0, z_0) , so, u\in H(y_0, z_0) . Therefore, (x_0, y_0)\in \{(x, y)\in X\times Y:\text{there exists}\ z\in G(x)\ \text{such that}\ u\in H(y, z)\} , which implies that the set \{(x, y)\in X\times Y:\text{there exists}\ z\in G(x) \text{such that}\ u\in H(y, z)\} is closed in X\times Y for every u\in X and thus, the set \{(x, y)\in X\times Y:u\not\in H(y, z) \text{for every}\ z\in G(x)\} is open in X\times Y for every u\in X .
By using the similar arguments as in Theorem 6.1, we have the following existence result of solutions for (GWIIP). We omit the proof.
Theorem 6.2. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , and Z be a nonempty set. Let A, B:X\rightarrow 2^X , G:X\rightarrow 2^Z , and H:X\times Z\rightarrow 2^X be four set-valued mappings satisfying
(i) for each x\in X , B(x)\subseteq A(x) ;
(ii) B has nonempty \Gamma -convex values and open lower sections;
(iii) the set \mathfrak{F} = \{x\in X:x\in A(x)\} is closed in X ;
(iv) for each u\in X , the set \{x\in X:u\not\in H(x, z)\ \text{for every}\ z\in G(x)\} is open in X ;
(v) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:u\not\in H(x, z)\ \text{for every}\ z\in G(x)\}) ;
(vi) one of the following conditions holds:
(vi) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\in B(x) and u\not\in H(x, z) for every z\in G(x) ;
(vi) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\in B(x) and u_0\not\in H(x, z) for every z\in G(x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then (GWIIP) is solvable, that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{x}, z) .
Remark 6.2. Theorem 6.2 generalizes Theorem 3.1 of Wang and Huang [49] in the following aspects: (a) from noncompact Hausdorff topological vector spaces to noncompact abstract convex spaces without any linear and convex structure. In fact, for X in Theorem 3.1 of Wang and Huang [49], let \Gamma_A = \text{co}(A) for every A\in \langle X\rangle , where \text{co}(A) denotes the convex hull of A . Then (X; \Gamma) forms an abstract convex space. Further, we can prove that 1_{X}\in{\frak{RC}}(X, X) (for details, see the proof of the following Theorem 6.3); (b) from two set-valued mappings to four set-valued mappings; (c) from one coercivity condition to two alternative coercivity conditions. And K in Theorem 6.2 only needs to be compact, while D in Theorem 3.1 of Wang and Huang [49] needs to be compact convex; (d) (v) of Theorem 6.2 is weaker than (i) and (ii) of Theorem 3.1 due to Wang and Huang [49]. In such an abstract convex space perspective, it is easy to see that (i) and (ii) of Theorem 3.1 due to Wang and Huang [49] can deduce (v) of Theorem 6.2; (e) concerns on the more general set Z without any topological and linear structure instead of the nonempty set Y in Theorem 3.1 of Wang and Huang [49], which is a subset of a Hausdorff topological vector space. In addition, the proof of Theorem 6.2 originates from the existence of maximal elements in noncompact abstract convex spaces, while Theorem 3.1 of Wang and Huang [49] is proved based on the famous FKKM theorem. Therefore, the proof method of Theorem 6.2 is different from that of Theorem 3.1 of Wang and Huang [49].
In Theorem 6.2, if X is a Banach space, then the compactness of L_{N_{0}} can be weakened to weak compactness.
Theorem 6.3. Let X be a real Banach space, K be a nonempty weak compact subset of X , and Z be a nonempty set. Let A, B:X\rightarrow 2^X , G:X\rightarrow 2^Z , and H:X\times Z\rightarrow 2^X be four set-valued mappings satisfying
(i) for each x\in X , B(x)\subseteq A(x) ;
(ii) B has nonempty convex values and weakly open lower sections;
(iii) the set \mathfrak{F} = \{x\in X:x\in A(x)\} is weakly closed in X ;
(iv) for each u\in X , the set \{x\in X:u\not\in H(x, z)\ \text{for every}\ z\in G(x)\} is weakly open in X ;
(v) for each x\in X , x\not\in \text{co}(\{u\in X:u\not\in H(x, z)\ \text{for every}\ z\in G(x)\}) ;
(vi) one of the following conditions holds:
(vi) _1 for each N_{0}\in \langle X\rangle , there exists a weak compact convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\in B(x) and u\not\in H(x, z) for every z\in G(x) ;
(vi) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\in B(x) and u_0\not\in H(x, z) for every z\in G(x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then (GWIIP) is solvable, that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{x}, z) .
Proof. Let \Gamma:\langle X\rangle\rightarrow 2^X be a set-valued mapping defined by \Gamma_A = \text{co}(A) for every A\in \langle X\rangle , where \text{co}(A) denotes the convex hull of A . Endowing X with the weak topology, we can see that (X; \Gamma) forms an abstract convex space and (i)-(vi) of Theorem 6.2 are satisfied. Now, we show that 1_{X}\in{\frak{RC}}(X, X) . In fact, let G:X\rightarrow 2^X is a KKM mapping with respect to the identity mapping 1_X such that each G(x) is weakly closed in X . Then for each A = \{x_0, x_1, \ldots, x_n\}\in \langle X\rangle , we have \Gamma_A = \text{co}(\{x_0, x_1, \ldots, x_n\})\subseteq \bigcup_{i = 0}^nG(x_i) and further, we can define a mapping \sigma:\Delta_n\rightarrow \text{co}(\{x_0, x_1, \ldots, x_n\}) by \sigma(t) = \sum_{i = 0}^nt_ix_i for every t = (t_0, t_1, \ldots, t_n)\in \Delta_n with \sum_{i = 0}^nt_i = 1 and t_i\geq 0 , where \Delta_n denotes the standard n -dimensional simplex with vertices \{e_0, e_1, \ldots, e_n\} . Considering the norm topology on \text{co}(\{x_0, x_1, \ldots, x_n\}) , we can see that the continuity of \sigma can be guaranteed by the fact that \|\sigma(t_1)-\sigma(t_2)\|\leq\sum_{i = 0}^n|t_{i1}-t_{i2}|\|x_i\| for every t_1 = (t_{01}, t_{11}, \ldots, t_{n1})\in \Delta_n with \sum_{i = 0}^nt_{i1} = 1 , t_{i1}\geq 0 and every t_2 = (t_{02}, t_{12}, \ldots, t_{n2})\in \Delta_n with \sum_{i = 0}^nt_{i2} = 1 , t_{i2}\geq 0 . For every i\in \{0, 1, \ldots, n\} , let E_i = \sigma^{-1}(\text{co}(\{x_0, x_1, \ldots, x_n\})\bigcap G(x_i)) . Since each G(x_i) is weakly closed in X , it follows that G(x_i) is closed in X . Thus, we can see that \text{co}(\{x_0, x_1, \ldots, x_n\})\bigcap G(x_i) is a closed subset of \text{co}(\{x_0, x_1, \ldots, x_n\}) . By the continuity of \sigma , each E_i is closed in \Delta_n . Next, let us prove that \text{co}(\{e_i:i\in I\})\subseteq \bigcup_{i\in I}E_i for every I = \{i_1, i_2, \ldots, i_k\}\in \langle\{0, 1, \ldots, n\}\rangle . In fact, let t = \sum_{j = 1}^kt_{i_j}e_{i_j}\in \text{co}(\{e_i:i\in I\}) be any given such that \sum_{j = 1}^kt_{i_j} = 1 and t_{i_j}\geq 0 . By the definition of \sigma and the hypothesis that G is a KKM mapping with respect to the identity mapping 1_X , we have \sigma(t)\in \text{co}\{x_{i_1}, x_{i_2}, \ldots, x_{i_k}\}\subseteq \bigcup_{j = 1}^kG(x_{i_j}) . Thus, there exists j\in \{1, 2, \ldots, k\} such that \sigma(t)\in \text{co}(\{x_0, x_1, \ldots, x_n\})\bigcap G(x_{i_j}) and consequently, t\in E_{i_j} . By applying the classical KKM principle to the family \{E_i\}_{i = 0}^n , there exists t_0\in \text{co}(\{e_0, e_1, \ldots, e_n\}) such that t_0\in \bigcap_{i = 0}^nE_i and so, \sigma(t_0)\in \bigcap_{i = 0}^nG(x_i) , which implies that the family \{G(x):x\in X\} has the finite intersection property. Therefore, as a consequence of Theorem 6.2, (GWIIP) is solvable, that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{x}, z) . This completes the proof.
In Theorem 6.2, if A(x) = B(x)\equiv X for every x\in X , then we have the following existence result of solutions for (GIIP).
Theorem 6.4. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , and Z be a nonempty set. Let G:X\rightarrow 2^Z and H:X\times Z\rightarrow 2^X be two set-valued mappings satisfying
(i) for each u\in X , the set \{x\in X:u\not\in H(x, z)\ \text{for every}\ z\in G(x)\} is open in X ;
(ii) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:u\not\in H(x, z)\ \text{for every}\ z\in G(x)\}) ;
(iii) one of the following conditions holds:
(iii) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\not\in H(x, z) for every z\in G(x) ;
(iii) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\not\in H(x, z) for every z\in G(x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then (GIIP) is solvable, that is, there exists \widehat{x}\in K such that for each u\in X , there exists z\in G(\widehat{x}) for which u\in H(\widehat{x}, z) .
In Theorem 6.2, if X = Z and G is the identity mapping on X , then we obtain the following existence theorem of solutions for (EWIP).
Theorem 6.4. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , and let A, B:X\rightarrow 2^X , and H:X\times X\rightarrow 2^X be three set-valued mappings satisfying
(i) for each x\in X , B(x)\subseteq A(x) ;
(ii) B has nonempty \Gamma -convex values and open lower sections;
(iii) the set \mathfrak{F} = \{x\in X:x\in A(x)\} is closed in X ;
(iv) for each u\in X , the set \{x\in X:u\not\in H(x, x)\} is open in X ;
(v) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:u\not\in H(x, x)\}) ;
(vi) one of the following conditions holds:
(vi) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\in B(x) and u\not\in H(x, x) ;
(vi) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\in B(x) and u_0\not\in H(x, x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then (EWIP) is solvable, that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and B(\widehat{x})\subseteq H(\widehat{x}, \widehat{x}) .
In Theorem 6.4, by setting A(x) = B(x)\equiv X for every x\in X , we have the following existence result of solutions for (EIP).
Corollary 6.1. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , and let H:X\times X\rightarrow 2^X be a set-valued mapping satisfying
(i) for each u\in X , the set \{x\in X:u\not\in H(x, x)\} is open in X ;
(ii) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:u\not\in H(x, x)\}) ;
(iii) one of the following conditions holds:
(iii) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\not\in H(x, x) ;
(iii) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\not\in H(x, x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then (EIP) is solvable, that is, there exists \widehat{x}\in K such that X\subseteq H(\widehat{x}, \widehat{x}) .
Remark 6.3. (1) Corollary 6.1 generalizes Theorem 3.4 of Wang and Huang [49] in the following aspects: (a) from noncompact Hausdorff topological vector spaces to noncompact abstract convex spaces without any linear and convex structure; (b) from one coercivity condition to two alternative coercivity conditions; (c) (ii) of Corollary 6.1 is weaker than (i) and (ii) of Theorem 3.4 due to Wang and Huang [49]. In addition, the proof of Corollary 6.1 is different from that of Theorem 3.4 due to Wang and Huang [49]. In fact, Corollary 6.1 is proved based on the existence of maximal elements in noncompact abstract convex spaces, while Theorem 3.4 of Wang and Huang [49] is proved using the famous FKKM theorem.
(2) Corollary 6.1 is different from Theorem 2.3 of Fang and Huang [50] in the following two ways: (a) X needs not be a Banach space; (b) the proof technique is different. Corollary 6.1 is established based on the existence of maximal elements in noncompact abstract convex spaces, while the proof of Theorem 2.3 of Fang and Huang [50] is proved by using the Kakutani-Fan-Glicksberg fixed point theorem.
By Corollary 6.1, we have the following corollary which is an existence result of solutions for (IP).
Corollary 6.2. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , and let H:X\rightarrow 2^X be a set-valued mapping satisfying
(i) for each u\in X , the set \{x\in X:u\not\in H(x)\} is open in X ;
(ii) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:u\not\in H(x)\}) ;
(iii) one of the following conditions holds:
(iii) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\not\in H(x) ;
(iii) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\not\in H(x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then (IP) is solvable, that is, there exists \widehat{x}\in K such that X\subseteq H(\widehat{x}) .
Proof. Define a set-valued mapping \widetilde{H}:X\times X\rightarrow 2^{X} by \widetilde{H}(x, z) = H(x) for every (x, z)\in X\times X . Then we can see that all the conditions of Corollary 6.1 are fulfilled. Thus, it follows from Corollary 6.1 that there exists \widehat{x}\in K such that X\subseteq \widetilde{H}(\widehat{x}, \widehat{x}) = H(\widehat{x}) , that is, (IP) is solvable. This completes the proof.
Now, as applications of Theorem 6.2, we have the following existence theorems of solutions for generalized set-valued implicit Stampacchia-type vector equilibrium problems and generalized set-valued implicit weak vector equilibrium problems in the framework of noncompact abstract convex spaces.
Theorem 6.5. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , Y be a topological vector space, and Z be a nonempty set. Let A, B:X\rightarrow 2^X , C:X\rightarrow 2^Y , G:X\rightarrow 2^Z , and F:X\times Z\times X\rightarrow 2^Y be five set-valued mappings satisfying
(i) for each x\in X , C(x) is a convex cone;
(ii) for each x\in X , B(x)\subseteq A(x) ;
(iii) B has nonempty \Gamma -convex values and open lower sections;
(iv) the set \mathfrak{F} = \{x\in X:x\in A(x)\} is closed in X ;
(v) for each u\in X , the set \{x\in X:F(x, z, u)\subseteq -C(x)\setminus\{0\}\ \text{for every}\ z\in G(x)\} is open in X ;
(vi) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:F(x, z, u)\subseteq -C(x)\setminus\{0\}\ \text{for every}\ z\in G(x)\}) ;
(vii) one of the following conditions holds:
(vii) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\in B(x) and F(x, z, u)\subseteq -C(x)\setminus\{0\} for every z\in G(x) ;
(vii) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\in B(x) and F(x, z, u_0)\subseteq -C(x)\setminus\{0\} for every z\in G(x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the generalized set-valued implicit Stampacchia-type vector equilibrium problem is solvable, that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which F(\widehat{x}, z, u)\not\subseteq -C(\widehat{x})\setminus\{0\} .
Proof. Define a set-valued mapping H:X\times Z\rightarrow 2^{X} by H(x, z) = \{u\in X:F(x, z, u)\not\subseteq -C(x)\setminus\{0\}\} for every (x, z)\in X\times Z . Then it is easy to see that all the conditions of Theorem 6.2 are satisfied. Thus, by Theorem 6.2, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{x}, z) , that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which F(\widehat{x}, z, u)\not\subseteq -C(\widehat{x})\setminus\{0\} . This completes the proof.
Remark 6.4. Theorem 6.5 generalizes Theorem 4.6 of Wang and Huang [49] in the following aspects: (a) from three set-valued mappings to five set-valued mappings; (b) from one coercivity condition to two alternative coercivity conditions. And the K in Theorem 6.5 only needs to be compact, while the D in Theorem 4.6 of Wang and Huang [49] needs to be compact convex; (c) (v) of Theorem 6.5 is weaker than (ii) and (iii) of Theorem 4.6 due to Wang and Huang [49]; (d) concerns on the more general set Z without any topological and linear structure instead of the nonempty set Y in Theorem 4.6 of Wang and Huang [49], which is a subset of a Hausdorff topological vector space. In addition, the proof of Theorem 6.5 is based on the existence of maximal elements in noncompact abstract convex spaces, while Theorem 4.6 of Wang and Huang [49] is proved using the famous FKKM theorem. Therefore, the proof technique of Theorem 6.5 is different from that of Theorem 4.6 of Wang and Huang [49].
Theorem 6.6. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , and Y be a topological vector space. Let A, B:X\rightarrow 2^X , C:X\rightarrow 2^Y , and F:X\times X\rightarrow 2^Y be four set-valued mappings satisfying
(i) for each x\in X , C(x) is a convex cone;
(ii) for each x\in X , B(x)\subseteq A(x) ;
(iii) B has nonempty \Gamma -convex values and open lower sections;
(iv) the set \mathfrak{F} = \{x\in X:x\in A(x)\} is closed in X ;
(v) for each u\in X , the set \{x\in X:F(x, u)\subseteq -C(x)\setminus\{0\}\} is open in X ;
(vi) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:F(x, u)\subseteq -C(x)\setminus\{0\}\}) ;
(vii) one of the following conditions holds:
(vii) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\in B(x) and F(x, u)\subseteq -C(x)\setminus\{0\} ;
(vii) _2 there exists u_0\in X such that for each x\in X\setminus K , one has u_0\in B(x) and F(x, u_0)\subseteq -C(x)\setminus\{0\} .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the generalized set-valued Stampacchia-type vector equilibrium problem is solvable, that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and F(\widehat{x}, u)\not\subseteq -C(\widehat{x})\setminus\{0\} for every u\in B(\widehat{x}) .
Proof. Let Z = X . Then we define two set-valued mappings \widetilde{F}:X\times Z\times X\rightarrow 2^{Y} and G:X\rightarrow 2^Z by \widetilde{F}(x, z, u) = F(x, u) for every (x, z, u)\in X\times Z\times X and G(x) = \{x\} for every x\in X , respectively. It is easy to see that all the requirements of Theorem 6.5 are fulfilled. Therefore, it follows from Theorem 6.5 that the conclusion of Theorem 6.6 holds. This completes the proof.
Remark 6.5. Theorem 6.6 generalizes Theorem 2.1 of Kazmi and Khan [52] in the following aspects: (a) from real Bananch spaces to noncompact abstract convex spaces without any linear and convex structure; (b) from a single-valued mapping to four set-valued mappings; (c) concerns the more general generalized set-valued Stampacchia-type vector equilibrium problems with movable convex cones instead of the generalized system problems with a fixed solid, pointed, closed and convex cone with apex at the origin; (d) in Theorem 6.6, the topological spaces X and Y need not to be Hausdorff spaces, while the spaces X and Y in Theorem 2.1 of Kazmi and Khan [52] have Hausdorff property. In fact, It can be seen from the proof of Theorem 2.1 of Kazmi and Khan [52] that the Hausdorff property of X is indispensable. In addition, the proof of Theorem 6.6 is essentially based on the existence of maximal elements in noncompact abstract convex spaces, while Theorem 2.1 of Kazmi and Khan [52] is proved by using the famous Brouwer's fixed point theorem. Thus, the proof method of Theorem 6.6 is different from that of Theorem 2.1 of Kazmi and Khan [52].
Theorem 6.7. Let (X; \Gamma) be an abstract convex space, K be a nonempty compact subset of X , Y be a topological vector space, and Z be a topological space. Let A, B:X\rightarrow 2^X , C:X\rightarrow 2^Y , G:X\rightarrow 2^Z , and F:X\times X\rightarrow 2^Y be five set-valued mappings. Let \zeta:X\times Z\rightarrow X be a continuous mapping and \eta:X\times X\rightarrow X be a continuous mapping in the first argument. Suppose that:
(i) for each x\in X , C(x) is a convex cone with \text{int}C(x)\neq\emptyset and the set-valued mapping W:X\rightarrow 2^Y defined by W(x) = Y\setminus \{-\text{int}C(x)\} for every x\in X , is closed;
(ii) G and F are two upper semicontinuous set-valued mappings with compact values;
(iii) for each x\in X , B(x)\subseteq A(x) ;
(iv) B has nonempty \Gamma -convex values and open lower sections;
(v) the set \mathfrak{F} = \{x\in X:x\in A(x)\} is closed in X ;
(vi) for each x\in X , x\not\in \Gamma\text{-}\text{co}(\{u\in X:F(\zeta(x, z), \eta(x, u))\subseteq -\text{int}C(x)\ \text{for every}\ z\in G(x)\}) ;
(vii) one of the following conditions holds:
(vii) _1 for each N_{0}\in \langle X\rangle , there exists a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\in B(x) and F(\zeta(x, z), \eta(x, u))\subseteq -\text{int}C(x) for every z\in G(x) ;
(vii) _2 there is u_0\in X such that for each x\in X\setminus K , one has u_0\in B(x) and F(\zeta(x, z), \eta(x, u_0))\subseteq -\text{int}C(x) for every z\in G(x) .
If (X; \Gamma) satisfies 1_{X}\in{\frak{RC}}(X, X) , then the generalized set-valued implicit weak vector equilibrium problem is solvable, that is, there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which F(\zeta(\widehat{x}, z), \eta(\widehat{x}, u))\not\subseteq -\text{int}C(\widehat{x}) .
Proof. Define a set-valued mapping H:X\times Z\rightarrow 2^X by H(x, z) = \{u\in X:F(\zeta(x, z), \eta(x, u))\not\subseteq -\text{int}C(x)\} for every (x, z)\in X\times Z . By (vi), we get x\not\in \Gamma\text{-}\text{co}(\{u\in X:u\not\in H(x, z) \text{for every}\ z\in G(x)\}) for every x\in X . Now, we show that the set \{x\in X:\text{there exists}\ z\in G(x)\ \text{such that}\ u\in H(x, z)\} = \{x\in X:\text{there exists}\ z\in G(x)\ \text{such that}\ F(\zeta(x, z), \eta(x, u))\not\subseteq -\text{int}C(x)\} is closed in X for every u\in X . In fact, let \{x_\alpha\} be an arbitrary net of \{x\in X:\text{there exists}\ z\in G(x)\ \text{such that} F(\zeta(x, z), \eta(x, u))\not\subseteq -\text{int}C(x)\} such that x_\alpha\rightarrow x_0 . Then for each \alpha , there exists z_\alpha\in G(x_\alpha) such that F(\zeta(x_\alpha, z_\alpha), \eta(x_\alpha, u))\not\subseteq -\text{int}C(x_\alpha) and thus, for each \alpha , there exists \vartheta_\alpha\in F(\zeta(x_\alpha, z_\alpha), \eta(x_\alpha, u)) such that \vartheta_\alpha\not\in-\text{int}C(x_\alpha) , which implies that \vartheta_\alpha\in Y\setminus \{-\text{int}C(x_\alpha)\} = W(x_\alpha) . Since G is upper semicontinuous with compact vales by (ii), it follows from Lemma 2.4 that there exist z_0\in G(x_0) and a subnet \{z_\beta\} of \{z_\alpha\} such that z_\beta\rightarrow z_0 . Further, Since F is upper semicontinuous with compact vales by (ii) again, \zeta is continuous and \eta is continuous in the first argument, by Lemma 2.4 again, there exist \vartheta_0\in F(\zeta(x_0, z_0), \eta(x_0, u)) and a subnet \{\vartheta_\gamma\} of \{\vartheta_\beta\} such that \vartheta_\gamma\rightarrow \vartheta_0 . Therefore, we have (x_\gamma, \vartheta_\gamma)\rightarrow (x_0, \vartheta_0) and \vartheta_\gamma\in W(x_\vartheta) for every \gamma . Since the graph of W is closed in X\times Y by (i), it follows that \vartheta_0\in W(x_0) = Y\setminus \{-\text{int}C(x_0)\} . Combining the fact that \vartheta_0\in F(\zeta(x_0, z_0), \eta(x_0, u)) , we know that F(\zeta(x_0, z_0), \eta(x_0, u))\not\subseteq -\text{int}C(x_0) . Thus, we have
x_0\in \{x\in X:\text{there exists}\ z\in G(x)\ \text{such that}\ u\in H(x, z)\}, |
which implies that the set \{x\in X:\text{there exists}\ z\in G(x)\ \text{such that}\ u\in H(x, z)\} is closed in X for every u\in X . Thus, (iv) of Theorem 6.2 is satisfied. By (vii) and the definition of H , we know that one of the following two conditions holds:
\bullet for each N_{0}\in \langle X\rangle , there exist a compact \Gamma -convex subset L_{N_{0}} of (X; \Gamma) containing N_{0} such that for each x\in L_{N_{0}}\setminus K , there exists u\in L_{N_{0}} such that u\in B(x) and u\not\in H(x, z) for every z\in G(x) .
\bullet there exists u_0\in X such that for each x\in X\setminus K , one has u_0\in B(x) and u_0\not\in H(x, z) for every z\in G(x) .
Combining (iii)-(v), we can see that all the requirements of Theorem 6.2 are fulfilled. Thus, it follows from Theorem 6.2 that there exists \widehat{x}\in K such that \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which u\in H(\widehat{x}, z) , that is, \widehat{x}\in A(\widehat{x}) and for each u\in B(\widehat{x}) , there exists z\in G(\widehat{x}) for which F(\zeta(\widehat{x}, z), \eta(\widehat{x}, u))\not\subseteq -\text{int}C(\widehat{x}) . This completes the proof.
Remark 6.6. Wang and Huang [49] studied the implicit set-valued weak vector equilibrium problem in the setting of Hausdorff topological vector spaces. Under some linear and convex assumptions, Wang and Huang [49] obtained an existence theorem of solutions for the implicit set-valued weak vector equilibrium problem. However, in the setting of noncompact abstract convex spaces without any linear and convex structure, Theorem 6.7 characterizes the existence of solutions for the generalized set-valued implicit weak vector equilibrium problem which is more general than the implicit set-valued weak vector equilibrium problem studied by Wang and Huang [49].
Remark 6.7. (vi) of Theorem 6.7 can be replaced by the following two conditions:
(vi) ' for each x\in X and each z\in G(x) , F(\zeta(x, z), \cdot) is C(x) - \Gamma -quasiconvex in the second argument of \eta .
(vi) '' for each x\in X , there exists z\in G(x) such that F(\zeta(x, z), \eta(x, x))\not\subseteq -\text{int}C(x) .
Indeed, we first show that the set D = \{u\in X:F(\zeta(x, z), \eta(x, u))\subseteq -\text{int}C(x)\ \text{for every}\ z\in G(x)\} is \Gamma -convex for every x\in X . In fact, let A = \{u_{0}, u_{1}, \ldots, u_{n}\}\in \langle D\rangle and u\in \Gamma(A) be given arbitrarily. Then by (vi) ' , there exists j\in \{0, 1, \ldots, n\} such that for each x\in X and each z\in G(x) , we have
\begin{eqnarray*} F(\zeta(x, z), \eta(x, u))&\subseteq&F(\zeta(x, z), \eta(x, u_j))-C(x)\\ &\subseteq&-\text{int}C(x)-C(x)\\ &\subseteq&-\text{int}C(x), \end{eqnarray*} |
which implies that
\Gamma(A)\subseteq D. |
Then it follows that the set D = \{u\in X:F(\zeta(x, z), \eta(x, u))\subseteq -\text{int}C(x)\ \text{for every}\ z\in G(x)\} is \Gamma -convex for every x\in X . Secondly, by this fact and (vi) '' , we have x\not\in \{u\in X:F(\zeta(x, z), \eta(x, u)) \subseteq -\text{int}C(x)\ \text{for every}\ z\in G(x)\} = \Gamma\text{-}\text{co}(\{u\in X:F(\zeta(x, z), \eta(x, u))\subseteq -\text{int}C(x) \text{for every}\ z\in G(x)\}) for every x\in X .
Finally, by using Theorem 3.4 and the same arguments as in Theorem 6.1, we obtain the following existence theorem of solutions for (SGWIIP).
Theorem 6.8. Let (X; \Gamma^1) and (Y; \Gamma^2) be two abstract convex spaces such that (X\times Y; \Gamma^1\times \Gamma^2) is an abstract convex space defined as in Lemma 2.5. Let K be a nonempty compact subset of X\times Y and Z be a nonempty set. Let A, B:X\rightarrow 2^X , F:X\rightarrow 2^Y , G:X\rightarrow 2^Z , and H:Y\times Z\rightarrow 2^X be five set-valued mappings satisfying
(i) for each x\in X , B(x)\subseteq A(x) ;
(ii) B and F have nonempty \Gamma^1 -convex and \Gamma^2 -convex values and open lower sections;
(iii) the set \mathfrak{F} = \{(x, y)\in X\times Y:x\in A(x)\ \text{and}\ y\in F(x)\} is closed in X\times Y ;
(iv) for each u\in X , the set \{(x, y)\in X\times Y:u\not\in H(y, z)\ \text{for some}\ z\in G(x)\} is open in X\times Y ;
(v) for each x\in X and each y\in F(x) , x\not\in \Gamma^1\text{-}\text{co}(\{u\in X:u\not\in H(y, z)\ \text{for some}\ z\in G(x)\}) ;
(vi) one of the following conditions holds:
(vi) _1 for each N_{0}\times N_{1}\in \langle X\times Y\rangle , there exist a compact \Gamma^1 -convex subset L_{N_{0}} of (X; \Gamma^1) containing N_{0} and a compact \Gamma^2 -convex subset L_{N_{1}} of (Y; \Gamma^2) containing N_{1} such that for L: = L_{N_{0}}\times L_{N_{1}} and for each (x, y)\in L\setminus K , there exists (u, v)\in L such that u\in B(x) , v\in F(x) , and u\not\in H(y, z) for some z\in G(x) ;
(vi) _2 there exists (u_0, v_0)\in X\times Y such that for each (x, y)\in X\times Y\setminus K , one has u_0\in B(x) , v_0\in F(x) , and u_0\not\in H(y, z) for some z\in G(x) .
If (X\times Y; \Gamma^1\times \Gamma^2) satisfies 1_{X\times Y}\in{\frak{RC}}(X\times Y, X\times Y) , then (SGWIIP) is solvable, that is, there exists (\widehat{x}, \widehat{y})\in K such that \widehat{x}\in A(\widehat{x}) , \widehat{y}\in F(\widehat{x}) , and u\in H(\widehat{y}, z) for every u\in B(\widehat{x}) and every z\in G(\widehat{x}) .
Proof. By using the same arguments as in Theorem 6.1, we can show that the set \mathfrak{F} is nonempty. Define a set-valued mapping J:X\times Y\rightarrow 2^X is defined by J(x, y) = \{u\in X:u\not\in H(y, z)\ \text{for some}\ z\in G(x)\} for every (x, y)\in X\times Y . Further, let us define a set-valued mapping T:X\times Y\rightarrow 2^{X\times Y} by setting, for each (x, y)\in X\times Y ,
\begin{eqnarray*} \ \ \ \ \ T(x, y) = \left\{ \begin{array}{ll} (B(x)\bigcap J(x, y))\times F(x), \ & \text{if}\ (x, y)\in \mathfrak{F}, \\ B(x)\times F(x), \ & \text{if}\ (x, y)\in X\times Y\setminus \mathfrak{F}. \end{array} \right. \end{eqnarray*} |
For each (u, v)\in X\times Y , we have
\begin{eqnarray*} T^{-1}(u, v)& = &\bigg{(}(X\times Y\setminus \mathfrak{F})\bigcap (B^{-1}(u)\times Y)\bigcap (F^{-1}(v)\times Y)\bigg{)}\\ &\bigcup& \bigg{(}J^{-1}(u)\bigcap (B^{-1}(u)\times Y)\bigcap (F^{-1}(v)\times Y)\bigg{)}. \end{eqnarray*} |
By (iv), the set J^{-1}(u) is open in X\times Y for every u\in X . Thus, it follows from (ii) and (iii) that T^{-1}(u, v) is open in X\times Y for every (u, v)\in X\times Y . By (v) and using the same arguments as in Theorem 6.1, we can deduce that (x, y)\not\in \Gamma^1\times\Gamma^2\text{-}\text{co}(T(x, y)) for every (x, y)\in X\times Y . By (vi), it follows that one of the following two conditions holds:
\bullet for each N_{0}\times N_{1}\in \langle X\times Y\rangle , there exist a compact \Gamma^1 -convex subset L_{N_{0}} of (X; \Gamma^1) containing N_{0} and a compact \Gamma^2 -convex subset L_{N_{1}} of (Y; \Gamma^2) containing N_{1} such that for L: = L_{N_{0}}\times L_{N_{1}} , we have L\setminus K\subseteq \bigcup_{(u, v)\in L}T^{-1}(u, v) .
\bullet there exists (u_0, v_0)\in X\times Y such that} X\times Y\setminus T^{-1}(u_{0}, v_{0})\subseteq K .
Thus, by Theorem 3.4 and Remark 3.4, there exists (\widehat{x}, \widehat{y})\in K such that T(\widehat{x}, \widehat{y}) = \emptyset . Since B and F have nonempty values, we can conclude that (\widehat{x}, \widehat{y})\in \mathfrak{F} . Thus, \widehat{x}\in A(\widehat{x}) , \widehat{y}\in F(\widehat{x}) , and B(\widehat{x})\bigcap J(\widehat{x}, \widehat{y}) = \emptyset . Therefore, u\in H(\widehat{y}, z) for every u\in B(\widehat{x}) and every z\in G(\widehat{x}) . This completes the proof.
Remark 6.8. (1) (v) of Theorem 6.8 can be replaced by the following stronger condition:
(v) ' H is strong \Gamma^1 -quasiconvex-like with respect to F and G .
In fact, suppose to the contrary that there exist x\in X and y\in F(x) such that x\in \Gamma^1\text{-}\text{co}(\{u\in X:u\not\in H(y, z) \text{for some}\ z\in G(x)\}) . Then it follows from Lemma 2.7 that there exists \{u_0, u_1, \ldots, u_n\}\in \langle\{u\in X:u\not\in H(y, z) \text{for some}\ z\in G(x)\}\rangle such that x\in\Gamma^1\text{-}\text{co}(\{u_0, u_1, \ldots, u_n\}) . By (v) ' , there exists j\in \{0, 1, \ldots, n\} and u_j\in H(y, z) for every z\in G(x) , which contradicts that u_j\not\in H(y, z) for some z\in G(x) . Therefore, (v) ' implies (v) of Theorem 6.5.
(2) the following two conditions imply that (v) ' holds.
(a) for each x\in X and each y\in F(x) , the set \{u\in X:u\not\in H(y, z)\ \text{for some}\ z\in G(x)\} is \Gamma^1 -convex.
(b) for each x\in X and each y\in F(x) , x\in H(y, z) for every z\in G(x) .
Indeed, by way of contradiction, suppose that for some N = \{u_{0}, u_{1}, \ldots, u_{n}\}\in \langle X\rangle , some x\in \Gamma\text{-}\text{co}(N) , there exists a point y\in F(x) such that for each j\in \{0, 1, \ldots, n\} , u_j\not \in H(y, z) for some z\in G(x) . By (a), we have x\not\in H(y, z) , which contradicts (b).
(3) If we assume that Z is a topological space, then (iv) of Theorem 6.8 can be replaced by the following condition:
(iv) ' G is a lower semicontinuous set-valued mapping and H is closed.
In fact, it is sufficient to prove that the set \{(x, y)\in X\times Y:\ u\in H(y, z) \text{for every}\ z\in G(x)\} is closed in X\times Y for every u\in X . Let (x^*, y^*)\in \text{cl}(\{(x, y)\in X\times Y:\ u\in H(y, z) \text{for every}\ z\in G(x)\}) any given. Then there is a net \{(x_\alpha, y_\alpha)\}\subseteq \{(x, y)\in X\times Y:\ u\in H(y, z) \text{for every}\ z\in G(x)\} such that (x_\alpha, y_\alpha)\rightarrow (x^*, y^*) . Therefore, we have u\in H(y_\alpha, z_\alpha) for every z'\in G(x_\alpha) . Since G is a lower semicontinuous set-valued mapping, it follows from Lemma 2.2 that for each z\in G(x^*) , there exists z_\alpha\in G(x_\alpha) such that z_\alpha\rightarrow z . Since H is closed, we have u\in H(y^*, z) . This shows that (x^*, y^*)\in \{(x, y)\in X\times Y:\ u\in H(y, z) \text{for every}\ z\in G(x)\} and so, the set \{(x, y)\in X\times Y:\ u\in H(y, z) \text{for every}\ z\in G(x)\} is closed in X\times Y for every u\in X . Thus, the set \{(x, y)\in X\times Y:u\not\in H(y, z) \text{for some}\ z\in G(x)\} is open in X\times Y for every u\in X .
In this paper, based on the KKM theory and the properties of \Gamma -convexity and {\frak{RC}} -mapping, we have dealt with the existence of collectively fixed points in the framework of noncompact abstract convex spaces and provided applications to some existence theorems of generalized weighted Nash equilibria and generalized Pareto Nash equilibria for constrained multiobjective games, some nonempty intersection theorems for sets with abstract convex sections, and some existence theorems of solutions for generalized weak implicit inclusion problems. In our view, future research should focus on considering how to further generalize and improve the collectively fixed point theorems obtained in this paper in the framework of noncompact abstract convex spaces. Furthermore, on this basis, the existence of generalized weighted Nash equilibria and generalized Pareto Nash equilibria for constrained multiobjective games with infinite players and the existence of solutions for systems of generalized vector quasi-variational equilibrium problems should be investigated.
The authors would like to thank the referees for their valuable comments and helpful suggestions which improve the exposition of the paper. This work was supported by the Planning Foundation for Humanities and Social Sciences of Ministry of Education of China (No. 18YJA790058).
The authors declare that they have no competing interests.
[1] |
E. Tarafdar, A fixed point theorem and equilibrium point of an abstract economy, J. Math. Econ., 20 (1991), 211–218. doi: 10.1016/0304-4068(91)90010-Q
![]() |
[2] |
K. Lan, J. Webb, New fixed point theorems for a family of mappings and applications to problems on sets with convex sections, Proc. Amer. Math. Soc., 126 (1998), 1127–1132. doi: 10.1090/S0002-9939-98-04347-0
![]() |
[3] | Q. H. Ansari, J. C. Yao, A fixed point theorem and its applications to the system of variational inequalities, B. Aust. Math. Soc., 54 (1999), 433–442. |
[4] | Q. H. Ansari, A. Idzik, J. C. Yao, Coincidence and fixed point theorems with applications, Topol. Method. Nonl. An., 15 (2000), 191–202. |
[5] |
S. P. Singh, E. Tarafdar, B. Watson, A generalized fixed point theorem and equilibrium point of an abstract economy, J. Comput. Appl. Math., 113 (2000), 65–71. doi: 10.1016/S0377-0427(99)00244-7
![]() |
[6] |
L. J. Lin, Z. T. Yu, Q. H. Ansari, L. P. Lai, Fixed point and maximal element theorems with applications to abstract economies and minimax inequalities, J. Math. Anal. Appl., 284 (2003), 656–671. doi: 10.1016/S0022-247X(03)00385-8
![]() |
[7] |
M. Balaj, L. J. Lin, Selecting families and their applications, Comput. Math. Appl., 55 (2008), 1257–1261. doi: 10.1016/j.camwa.2007.06.011
![]() |
[8] |
E. Tarafdar, Fixed point theorems in H-spaces and equilibrium point of abstract economies, J. Aust. Math. Soc., 53 (1992), 252–260. doi: 10.1017/S1446788700035825
![]() |
[9] |
S. Park, Continuous selection theorems in generalized convex spaces, Numer. Funct. Anal. Opt., 20 (1999), 567–583. doi: 10.1080/01630569908816911
![]() |
[10] |
Z. T. Yu, L. J. Lin, Continuous selections and fixed point theorems, Nonlinear Anal. Theor., 52 (2003), 445–455. doi: 10.1016/S0362-546X(02)00107-4
![]() |
[11] |
X. P. Ding, Collectively fixed points and a system of coincidence theorems in product FC-spaces and applications, Nonlinear Anal. Theor., 66 (2007), 2604–2617. doi: 10.1016/j.na.2006.03.043
![]() |
[12] |
Q. B. Zhang, C. Z. Cheng, Some fixed-point theorems and minimax inequalities in FC-space, J. Math. Anal. Appl., 328 (2007), 1369–1377. doi: 10.1016/j.jmaa.2006.06.027
![]() |
[13] | S. Al-Homidan, Q. H. Ansari, J. C. Yao, Collectively fixed point and maximal element theorems in topological semilattice spaces, Appl. Anal., 90 (2010), 865–888. |
[14] |
P. Q. Khanh, V. S. T. Long, N. H. Quan, Continuous selections, collectively fixed points and weak Knaster-Kuratowski-Mazurkiewicz mappings in optimization, J. Optimiz. Theory Appl., 151 (2011), 552–572. doi: 10.1007/s10957-011-9889-0
![]() |
[15] |
P.Q. Khanh, N.H. Quan, A fixed-component point theorem and applications, B. Malays. Math. Sci. Soc., 42 (2019), 503–520. doi: 10.1007/s40840-017-0496-6
![]() |
[16] | S. Park, On generalizations of the KKM principle on abstract convex spaces, Nonlinear Anal. Forum, 11 (2006), 67–77. |
[17] | S. Park, New foundations of the KKM theory, J. Nonlinear Convex Anal., 9 (2008), 331–350. |
[18] | S. Park, Generalizations of the Nash equilibrium theorem in the KKM theory, Fixed Point Theory A., 2010 (2010), 1–23. |
[19] | H. S. Lu, Q. W. Hu, A collectively fixed point theorem in abstract convex spaces and its applications, J. Funct. Space. Appl., 2013 (2013), 1–10. |
[20] | J. P. Aubin, I. Ekeland, Applied nonlinear analysis, New York: John Wiley & Sons, 1984. |
[21] | C. D. Aliprantis, K. C. Border, Infinite dimensional analysis, New York: Springer-Verlag, 1994. |
[22] |
F. Ferro, Optimization and stability results through cone lower semicontinuity, Set-Valued Anal., 5 (1997), 365–375. doi: 10.1023/A:1008653120360
![]() |
[23] |
S. Park, Elements of the KKM theory on convex spaces, J. Korean Math. Soc., 45 (2008), 1–27. doi: 10.4134/JKMS.2008.45.1.001
![]() |
[24] | M. Lassonde, On the use of KKM multifunctions in fixed point theory and related topics, J. Math. Anal. Appl., 97 (1983), 151–201. |
[25] |
C. D. Horvath, Contractibility and generalized convexity, J. Math. Anal. Appl., 156 (1991), 341–357. doi: 10.1016/0022-247X(91)90402-L
![]() |
[26] |
H. Ben-E1-Mechaiekh, S. Chebbi, M. Flonzano, J. V. Llinares, Abstract convexity and fixed points, J. Math. Anal. Appl., 222 (1998), 138–150. doi: 10.1006/jmaa.1998.5918
![]() |
[27] |
S. Park, New generalizations of basic theorems in the KKM theory, Nonlinear Anal. Theor., 74 (2011), 3000–3010. doi: 10.1016/j.na.2011.01.020
![]() |
[28] |
S. Park, On the von Neumann-Sion minimax theorem in KKM spaces, Appl. Math. Lett., 23 (2010), 1269–1273. doi: 10.1016/j.aml.2010.06.011
![]() |
[29] | S. Park, A genesis of general KKM theorems for abstract convex spaces, J. Nonlinear Anal. Optimiz., 2 (2011), 141–154. |
[30] |
E. Tarafdar, A fixed point theorem in H-space and related results, B. Aust. Math. Soc., 42 (1990), 133–140. doi: 10.1017/S0004972700028239
![]() |
[31] | K. K. Tan, X. L. Zhang, Fixed point theorems on G-convex spaces and applications, Proc. Nonlinear Funct. Anal. Appl., 1 (1996), 1–19. |
[32] |
X. P. Ding, Maximal elements of GKKM-majorized mappings in product FC-spaces and applications (I), Nonlinear Anal. Theor., 67 (2007), 963–973. doi: 10.1016/j.na.2006.06.037
![]() |
[33] | X. Ding, Fixed points, minimax inequalities and equilibria of noncompact generalized games, Taiwan. J. Math., 2 (1998), 25–55. |
[34] |
F.E. Browder, The fixed point theory of multivalued mappings in topological vector spaces, Math. Ann., 177 (1968), 283–301. doi: 10.1007/BF01350721
![]() |
[35] |
C. D. Horvath, J. V. L. Ciscar, Maximal elements and fixed points for binary relations on topological ordered spaces, J. Math. Econ., 25 (1996), 291–306. doi: 10.1016/0304-4068(95)00732-6
![]() |
[36] |
N. C. Yannelis, N. D. Prabhakar, Existence of maximal elements and equilibria in linear topological spaces, J. Math. Econ., 12 (1983), 233–245. doi: 10.1016/0304-4068(83)90041-1
![]() |
[37] | S. Al-Homidan, Q. H. Ansari, Fixed point theorems on product topological semilattice spaces, generalized abstract economies and systems of generalized vector quasi-equilibrium problems, Taiwan. Math., 15 (2011), 307–330. |
[38] |
Q. Luo, Ky Fan's section theorem and its applications in topological ordered spaces, Appl. Math. Lett., 17 (2004), 1113–1119. doi: 10.1016/j.aml.2003.12.003
![]() |
[39] |
M. Balaj, L. J. Lin, Existence criteria for the solutions of two types of variational relation problems, J. Optimiz. Theory Appl., 156 (2013), 232–246. doi: 10.1007/s10957-012-0136-0
![]() |
[40] | X. P. Ding, Existence af Pareto equilibria for constrained multiobjective gmes in H-space, Comput. Math. Appl., 39 (2000), 125–134. |
[41] |
K. Kim, X. P. Ding, On generalized weight nash equilibria for generalized multiobjective games, J. Korean Math. Soc., 40 (2003), 883–899. doi: 10.4134/JKMS.2003.40.5.883
![]() |
[42] |
S. Y. Wang, Existence of a pareto equilibrium, J. Optimiz. Theory Appl., 79 (1993), 373–384. doi: 10.1007/BF00940586
![]() |
[43] |
X. Z. Yuan, E. Tarafdar, Non-compact pareto equilibria for multiobjective games, J. Math. Anal. Appl., 204 (1996), 156–163. doi: 10.1006/jmaa.1996.0429
![]() |
[44] | J. Yu, G. X. Z. Yuan, The study of Pareto equilibria for multiobjective games by fixed point and Ky Fan minimax inequality methods, Comput. Math. Appl., 35 (1998), 17–24. |
[45] |
P. L. Yu, Second-order game problems: Decision dynamics in gaming phenomena, J. Optimiz. Theory Appl., 27 (1979), 147–166. doi: 10.1007/BF00933332
![]() |
[46] |
R. Bielawski, Simplicial convexity and its applications, J. Math. Anal. Appl., 127 (1987), 155–171. doi: 10.1016/0022-247X(87)90148-X
![]() |
[47] |
W. A. Kirk, B. Sims, G. X. Z. Yuan, The Knaster-Kuratowski and Mazurkiewicz theory in hyperconvex metric spaces and some of its applications, Nonlinear Anal. Theor., 39 (2000), 611–627. doi: 10.1016/S0362-546X(98)00225-9
![]() |
[48] |
K. Fan, Some properties of convex sets related to fixed point theorems, Math. Ann., 266 (1984), 519–537. doi: 10.1007/BF01458545
![]() |
[49] |
S. H. Wang, N. J. Huang, Generalized implicit inclusion problems on noncompact sets with applications, B. Aust. Math. Soc., 84 (2011), 261–279. doi: 10.1017/S0004972710002091
![]() |
[50] | Y. P. Fang, N. J. Huang, An existence result for a class of extended inclusion problems with applications to equilibrium problems, J. Anal. Appl., 25 (2006), 257–264. |
[51] | B. Di Bella, An existence theorem for a class of inclusions, Appl. Math. Lett., 13 (2000), 15–19. |
[52] |
K. R. Kazmi, S. A. Khan, Existence of solutions to a generalized system, J. Optimiz. Theory Appl., 142 (2009), 355–361. doi: 10.1007/s10957-009-9530-7
![]() |