Hindawi Publishing Corporation Abstract and Applied Analysis Volume 2013, Article ID 678154, 6 pages http://dx.doi.org/10.1155/2013/678154 Research Article Positive Definiteness of High-Order Subdifferential and High-Order Optimality Conditions in Vector Optimization Problems He Qinghai1 and Zhang Binbin2 1 2 Department of Mathematics, Yunnan University, Kunming, Yunnan 650091, China School of Science, Kunming University of Science and Technology, Kunming, Yunnan 650500, China Correspondence should be addressed to He Qinghai; heqh@ynu.edu.cn Received 15 October 2012; Accepted 26 December 2012 Academic Editor: Gue Lee Copyright © 2013 H. Qinghai and Z. Binbin. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. We obtain a new Taylor’s formula in terms of the π + 1 order subdifferential of a πΆπ,1 function from π π to π π . As its applications in optimization problems, we build π + 1 order sufficient optimality conditions of this kind of functions and π + 1 order necessary conditions for strongly πΆ-quasiconvex functions. 1. Introduction For a function from π π to π , Luc [1] studied the π + 1 order subdifferential of it, established a Taylor-type formula in terms of such π + 1 order subdifferential, and applied such Taylor-type formula to consider two-order optimality conditions in vector optimization and characterizations of quasiconvex functions. In vector optimization, notions of Pareto solution, weak Pareto solution, sharp minima and weak sharp minima are very important; see [2–14] and the references therein. Some authors have attained many necessary or sufficient optimality conditions in optimization problems. In particular, Zheng and Yang provided some results on sharp minima, and weak sharp minima for high-order smooth vector optimization problems in Banach spaces. By the tools of nonsmooth analysis, many optimality conditions were obtained; for examples, one can see [6, 7, 15, 16] and the references therein. Such optimality conditions play a key role in many issues of mathematical programming such as sensitivity analysis and error bounds. Motivated by Luc [1] and Zheng and Yang [17], in this paper, we consider the π + 1 order subdifferential and optimality conditions of a πΆπ,1 vector-valued function from π π to π π . We will first prove a new Taylor’s formula in the terms of π + 1 order subdifferential for πΆπ,1 functions from π π to π π , which is analogous to that for real-valued functions in [1]. Then, under the positive definiteness assumption of π + 1 order subdifferential, we will use this formula to derive π + 1 order optimality conditions of weak Pareto and Pareto solutions in the terms of π + 1 order subdifferential for a πΆπ,1 function from π π to π π . Finally, we will define a kind of strongly πΆ-quasiconvex functions and prove a necessary condition in the terms of (π + 1)th order subdifferential for such kind of functions. Our results extend the corresponding results in [1] for πΆπ,1 functions from π π to π to that for πΆπ,1 vector-valued functions from π π to π π and in [17] for functions in smooth setting to that in nonsmooth setting, respectively. The outline of the paper is as follows. In the next section, we give some notions and preliminary results in vector optimization problems. In Section 3, we build our Taylor’s formula in the terms of π + 1 order subdifferential for a πΆπ,1 function from π π to π π . In Section 4, as applications in optimization problems, we establish some optimality conditions in terms of (π + 1)th order subdifferential. In Section 5, we give a necessary condition in the terms of 2 Abstract and Applied Analysis (π + 1)th order subdifferential for a strongly πΆ-quasiconvex vector-valued function. 2. Preliminaries Let π, π be Banach spaces, π∗ the dual space of π, πΆ ⊂ π a closed convex cone with int(πΆ) =ΜΈ 0, and πΆ+ the dual cone of πΆ; that is, πΆ+ = {π¦∗ ∈ π∗ : 0 ≤ β¨π¦∗ , πβ© ∀π ∈ πΆ} . π ∈ WE (π΄, πΆ) ⇐⇒ (π − int (πΆ)) ∩ π΄ = 0, (2) πΌ (π΄, πΆ) ⊂ πΈ (π΄, πΆ) ⊂ WE (π΄, πΆ) . π Let π := {(π₯1 , . . . , π₯π ) : π₯π ∈ π, π = 1, . . . , π} be equipped with the norm β(π₯1 , . . . , π₯π )β = ∑ππ=1 βπ₯π β. Let Φ : ππ → π be π-linear and symmetric mapping [17]; that is, for any π , π‘ ∈ R and π₯1 , π§1 , π₯2 , . . . , π₯π ∈ π, Φ (π π₯1 + π‘π§1 , π₯2 , . . . , π₯π ) = π Φ (π₯1 , π₯2 , . . . , π₯π ) + π‘Φ (π§1 , π₯2 , . . . , π₯π ) , (3) Φ (π₯1 , . . . , π₯π ) = Φ (π₯π1 , . . . , π₯ππ ) , where (π1 , . . . , ππ ) is an arbitrary permutation of (1, . . . , π). Let π : π → π be a mapping. It is known that its derivative π(π) (π₯) is π-linear, symmetric, and continuous mapping if π is π-time smooth. Let π be a function from π π to π π and πΆ ⊂ π π be a closed convex cone. Consider the following vector optimization problem πΆ − minπ π (π₯) . (4) π₯∈π A vector π₯ ∈ π is said to be a local weak Pareto (resp., Pareto and ideal) solution of (4) if there exists πΏ > 0 such that π(π₯) is a weak Pareto (resp., Pareto and ideal) point of π(π΅(π₯, πΏ)), where π΅(π₯, πΏ) denotes the open ball with center π₯ and radius πΏ. We say that π₯ is a sharp Pareto solution of (4) of order π if there exist π, πΏ ∈ (0, +∞) such that π π βπ₯ − π₯β ≤ [π (π₯) − π (π₯)]+ , ∀π₯ ∈ π΅ (π₯, πΏ) , ππ+1 π (π₯) := co {lim π·π+1 π (π₯π ) : π₯π σ³¨→ π₯, π·π+1 π (π₯π ) exists at π₯π } . (1) For π¦1 , π¦2 ∈ π, we define π¦1 <πΆ π¦2 and π¦1 ≤πΆ π¦2 if π¦2 − π¦1 ∈ int(πΆ) and π¦2 − π¦1 ∈ πΆ, respectively. Let π΄ be a subset of π and π ∈ π΄. Recall that (i) π is a weak Pareto point of π΄ if there exists no point π¦ ∈ π΄ such that π¦ <πΆ π; (ii) π is a Pareto point of π΄ if there exists no point π¦ ∈ π΄ \ {π} such that π¦ ≤πΆ π; (iii) π is an ideal point of π΄ if π ≤πΆ π¦ for all π¦ ∈ π΄. Let WE(π΄, πΆ), πΈ(π΄, πΆ), and πΌ(π΄, πΆ) denote the sets of all weak Pareto, Pareto, and ideal points of π΄, respectively. It is easy to verify that π ∈ πΈ (π΄, πΆ) ⇐⇒ (π − πΆ) ∩ π΄ = {π} , Lipschitz functions from π π to π π . By Rademacher’s theorem (see [18]), for any π ∈ πΆπ,1 , π(π₯) = (π1 (π₯), . . . , ππ (π₯)), its πth order derivative π·π π(π₯) is a function differentiable almost everywhere. The (π + 1)th order subdifferential of π at π₯ ∈ π π is defined as “generalized Jacobian” of π·π+1 π at π₯ in Clarke’s sense [18] as follows: (5) where [π(π₯) − π(π₯)]+ := π(π(π₯) − π(π₯), −πΆ). We denote by πΆπ,1 , π > 0, the class of π-time differentiable mappings from π π to π π whose πth order derivatives are locally Lipschitz mappings and by πΆ0,1 the class of locally (6) It is worth mentioning that each element in ππ+1 π(π₯) is a π+1 linear and symmetric mapping from (π π )π+1 to π π . For more details about ππ+1 π(π₯), we refer the reader to [18]. It is similar to the proof of Lemma 2.1 in [1], and one can verify the following chain rule. Lemma 1. Let π₯, π’ in π π , π be a function from π to π π defined by π(π‘) = π₯ + π‘π’ for every π‘ ∈ π , and let π be a πΆπ,1 function from π π to π π . Then, ππ+1 (π β π) (π‘) ⊆ ππ+1 π (π₯ + π‘π’) (π’, . . . , π’) . (7) 3. A New Taylor’s Formula in Form of High-Order Subdifferential By Lemma 1, we have the following Taylor-type formula for a πΆπ,1 vector-valued function from π π to π π which will be useful in the sequel. Theorem 2. Let π₯, π’, and π be as in Lemma 1. Then, there exists π΄ ∈ clco ππ+1 π(π₯, π₯ + π’) such that π 1 1 π (π₯ + π’) − π (π₯) = ∑ π·π π (π₯) (π’π ) + π΄ (π’π+1 ) , + 1)! π! (π π=1 (8) where π’π denotes (π’, . . . , π’) ∈ (π π )π and clco ππ+1 π (π₯, π₯ + π’) := clco ( β ππ+1 π (π₯ + π‘π’)) . (9) π‘∈(0,1) Proof. Let πΌ ∈ π π be a vector satisfying π 1 1 π (π₯ + π’) − π (π₯) = ∑ π·π π (π₯) (π’π ) + πΌ. π! + 1)! (π π=1 (10) We only need to show that there exists π΄ ∈ clco ππ+1 π(π₯, π₯ + π’) such that πΌ = π΄ (π’π+1 ) . (11) Let π be as Lemma 1. Set π(π‘) := (π β π)(π‘) and β (π‘) := π (1) − π (π‘) π 1 1 − ∑ π·π π (π‘) (1 − π‘)π − (1 − π‘)π+1 πΌ. + π! (π 1)! π=1 (12) Abstract and Applied Analysis 3 Let π¦ ∈ π π be arbitrarily given. Since the function β¨π¦, β(⋅)β© is locally Lipschitz and β(0) = β(1), applying Lebourg mean value theorem [18, Theorem 2.3.7 and Theorem 2.3.9], there exists π‘0 ∈ (0, 1) such that (13) Since π¦ is arbitrary in π π and clco ππ+1 π(π₯, π₯ + π’)(π’π+1 ) is convex and compact, by the separation theorem, we can easily show that πΌ ∈ clco ππ+1 π(π₯, π₯ + π’)(π’π+1 ). Hence, we can take π΄ ∈ clco ππ+1 π(π₯, π₯ + π’) such that πΌ = π΄(π’π+1 ). The proof is completed. Noting that π(⋅) and each π·π π(⋅)(1 − ⋅)π (1 ≤ π ≤ π − 1) have derivatives which are continuous, it follows that they are strictly π»-differentiable. We have Corollary 3. Let π be as in Theorem 2 and π ∈ π π . Then, for every π₯ ∈ π π , there exist π΄ π₯ ∈ ππ+1 π(π) and a (π + 1)-linear mapping π(π₯) from (π π )π+1 to π π such that 0 ∈ π β¨π¦, β (π‘0 )β© ⊂ β¨π¦, πβ (π‘0 )β© . π−1 1 πβ (π‘0 ) = − ππ (π‘0 ) − ∑ π (π·π π (⋅) (1 − ⋅)π ) (π‘0 ) π! π=1 lim βπ (π₯)β = 0, π₯→π π 1 π (π₯) = π (π) + ∑ π·π π (π) (π₯ − π)π π! π=1 1 − π (π·π π (⋅) (1 − ⋅)π ) (π‘0 ) π! − 1 π ((1 − ⋅)π+1 πΌ) (π‘0 ) (π + 1)! + π−1 1 π π−1 (π·π+1 π (π‘0 ) (1 − π‘0 ) − ππ·π π (π‘0 ) (1 − π‘0 ) ) π! π=1 −∑ = − − 1 1 π π (π·π π (⋅) (1 − ⋅)π ) (π‘0 ) − (1 − π‘0 ) πΌ π! π! 1 1 π π (π·π π (⋅) (1 − ⋅)π ) (π‘0 ) − (1 − π‘0 ) πΌ. π! π! π (π₯) := (14) Here, the first equation holds by Propositions 7.4.3(b), and 7.3.5 in [19] and the second holds by Proposition 7.3.9 in [19]. By the chain rule [19, Theorem 7.4.5(a)], we also have π π (π·π π (⋅) (1 − ⋅)π ) (π‘0 ) ⊂ ππ+1 π (π‘0 ) (1 − π‘0 ) π − ππ·π π (π‘0 ) (1 − π‘0 ) . (15) Hence, we have 1 1 π π πβ (π‘0 ) ⊂ − ππ+1 π (π‘0 ) (1 − π‘0 ) + (1 − π‘0 ) πΌ. π! π! (16) From (13) and (16), we have 1 π+1 1 π π π π (π‘0 ) (1 − π‘0 ) + (1 − π‘0 ) πΌβ© . π! π! (17) 1 1 π π (1 − π‘0 ) ππ+1 π (π₯ + π‘0 π’) (π’π+1 ) + (1−π‘0 ) πΌβ© ; π! π! (18) that is, β¨π¦, πΌβ© ∈ β¨π¦, ππ+1 π (π₯ + π‘0 π’) (π’π+1 )β© ⊂ β¨π¦, clco ππ+1 π (π₯, π₯ + π’) (π’π+1 )β© . π΅π₯ − π΄ π₯ . (π + 1)! (22) Then, from (21), we obtain the formula of the corollary. Moreover, since the mapping ππ+1 π is upper continuous, nonempty, convex, and compact valued (see [18]), for any π > 0, there exists πΏ > 0 such that, for all π¦ ∈ π + πΏπ΅π π (where π΅π π denotes the closed unit ball of π π ), ππ+1 π (π¦) ⊂ ππ+1 π (π) + (π + 1)!ππ΅πΏ((π π )π+1 ,π π ) , (23) where π΅πΏ((π π )π+1 , π π ) denotes the closed unit ball of the space πΏ((π π )π+1 , π π ) of all bounded linear operators from (π π )π+1 to π π . If π₯ ∈ π + πΏπ΅π π , then clco ππ+1 π (π, π + π₯) ⊂ ππ+1 π (π) + (π + 1)!ππ΅πΏ((π π )π+1 ,π π ) . (24) With this we obtain βπ(π₯)β ≤ π. The proof is completed. Together with Lemma 1, it follows that 0 ∈ β¨π¦, − π 1 1 π (π₯) = π (π) + ∑ π·π π (π) (π₯ − π)π + π΅π₯ (π₯ − π)π+1 . π! + 1)! (π π=1 (21) Let π΄ π₯ ∈ ππ+1 π(π) be an element minimizing the distance from π΅π₯ to the convex and compact set ππ+1 π(π). Set 1 π−1 π·π π (π‘0 ) (1 − π‘0 ) (π − 1)! 0 ∈ β¨π¦, − 1 π΄ (π₯ − π)π+1 + π (π₯) (π₯ − π)π+1 . (π + 1)! π₯ Proof. By Theorem 2, for a given π₯ ∈ π π , there exists π΅π₯ ∈ clco ππ+1 π(π, π + π₯) such that = − πσΈ (π‘0 ) − (20) (19) 4. The Positive Definiteness of High-Order Subdifferential and Optimality Conditions Recall [17] that π-linear symmetric mapping Φ : ππ → π is said to be positively definite (resp., positively semidefinite) with respect to the ordering cone πΆ if 0 <πΆ Φ (π₯π ) (resp., 0 ≤πΆ Φ (π₯π )) , ∀π₯ ∈ π \ {0} , (25) 4 Abstract and Applied Analysis where π₯π denotes (π₯, . . . , π₯). If π is odd and the ordering cone πΆ is pointed (i.e., πΆ ∩ (−πΆ) = {0}), then Φ is positively semidefinite if and only if Φ = 0; see [17]. By the separation theorem, it is easy to verify that a πlinear symmetric mapping Φ is positively semidefinite with respect to the ordering cone πΆ if and only if the composite π∗ β Φ is positively semidefinite for any π∗ ∈ πΆ+ . Recall that a mapping π : π → π is πΆ-convex if π (π‘π₯1 + (1 − π‘) π₯2 ) ≤πΆ π‘π (π₯1 ) + (1 − π‘) π (π₯2 ) , ∀π₯1 , π₯2 ∈ π, ∀π‘ ∈ [0, 1] . (26) Noting that π is πΆ-convex if and only if π∗ β π is convex for all π∗ ∈ πΆ+ , one can see that a twice differentiable function π is πΆ-convex if and only if πσΈ σΈ (π₯) is positively semidefinite for all π₯ ∈ π. Inspired by the notion of positive definiteness, we introduce positive definiteness of the (π+1)th order subdifferential for πΆπ,1 functions. Definition 4. Let π be a πΆπ,1 function from π π to π π and πΆ a closed convex cone of π π . We say that the (π + 1)th order subdifferential mapping ππ+1 π is positively definite at π₯ ∈ π π with respect to the ordering cone πΆ if each π΄ ∈ ππ+1 π(π₯) is positively definite with respect to πΆ. Proposition 5. Let π be a πΆπ,1 function from π π to π π , and let πΆ be a closed convex cone of π π . Suppose that the subdifferential mapping ππ+1 π is positively definite at π₯ ∈ π π with respect to πΆ. Then, there exists π > 0 such that π΄ (π₯π+1 ) + ππ΅π π ⊂ πΆ, ∀π΄ ∈ ππ+1 π (π₯) , π₯ ∈ ππ π , (27) where ππ π := {π₯ ∈ π π : βπ₯β = 1}. Proof. From [17, Proposition 3.4], for any π΄ ∈ ππ+1 π(π₯), there exists ππ΄ > 0 such that π΄ (π₯π+1 ) + ππ΄π΅π π ⊂ πΆ, ∀π₯ ∈ ππ π . (28) If the conclusion is not true, then, for every natural number π, there exist π΄ π ∈ ππ+1 π(π₯), π₯π ∈ ππ π and ππ ∈ π΅π π such that 1 π΄ π (π₯ππ+1 ) + ππ ∉ πΆ. π (29) π+1 Since π π(π₯) and ππ π are compact, we can assume that π΄ π → π΄ 0 ∈ ππ+1 π(π₯), π₯π → π₯0 ∈ ππ π (passing to a subsequence if necessary). Then, 1 π΄ 0 (π₯0π+1 ) + (π΄ π (π₯ππ+1 ) − π΄ 0 (π₯0π+1 )) + ππ ∉ πΆ π (30) for all π. But from (28), for large enough π, we have π΄ 0 (π₯0π+1 ) + (π΄ π (π₯ππ+1 ) − Theorem 6. Let π be a πΆπ,1 function from π π to π π , πΆ a closed convex cone of π π , and π₯ ∈ π π . Suppose that there exists π∗ ∈ πΆ+ with βπ∗ β = 1 such that ∑ππ=1 (1/π!)π∗ β π·π π(π₯) = 0, and that π(π+1) π is positively definite at π₯ with respect to the ordering cone πΆ. Then, π₯ is a local Pareto solution of (4), and there exist π, πΏ ∈ (0, +∞) such that 1/(π+1) π βπ₯ − π₯β ≤ [π (π₯) − π (π₯)]+ , ∀π₯ ∈ π΅ (π₯, πΏ) . (32) Proof. Since π(π+1) π(π₯) is positively definite with respect to πΆ, by Proposition 5, there exists π > 0 such that 1 π(π+1) π (π₯) (βπ+1 ) + 2ππ+1 π΅π π ⊂ πΆ, (π + 1)! ∀β ∈ ππ π . (33) Noting that π∗ ∈ πΆ+ and βπ∗ β = 1, we have that 1 β¨π∗ , π΄βπ+1 β© ≥ 2ππ+1 βββπ+1 , (π + 1)! ∀π΄ ∈ π (π+1) π (π₯) , (34) π β∈π . Let π(π₯) := β¨π∗ , π(π₯)β© for all π₯ ∈ π. Since π is a πΆπ,1 function, so is π. Noting that π·(π) π(π₯) = π∗ β π·(π) π(π₯) with Corollary 3, there exist π΄ π₯ ∈ ππ+1 π(π₯) and (π + 1)-linear mapping π(π₯) with limπ₯ → π₯ βπ(π₯)β = 0 such that π 1 1 π (π₯) = π (π₯) + ∑ π·(π) π (π₯) ((π₯ − π₯)π ) + π! + 1)! (π π=1 × β¨π∗ , π΄ π₯ (π₯−π₯)π+1 β©+β¨π∗ , π (π₯) (π₯−π₯)π+1 β©. (35) It follows that there exists πΏ > 0 such that π 1 π (π₯) − π (π₯) − ∑ π·(π) π (π₯) ((π₯ − π₯)π ) π! π=1 − 1 β¨π∗ , π΄ π₯ (π₯ − π₯)π+1 β© (π + 1)! (36) ≥ − ππ+1 βπ₯ − π₯βπ+1 , for all π₯ ∈ π΅(π₯, πΏ). Since ∑ππ=1 (1/π!)π∗ β π·π π(π₯) = 0, it follows from (34) and (36) that ππ+1 βπ₯ − π₯βπ+1 ≤ π (π₯) − π (π₯) , π΄ 0 (π₯0π+1 )) 1 + ππ ∈ π΄ 0 (π₯0π+1 ) + ππ΄ 0 π΅π π ⊂ πΆ, π Under the positive definiteness assumption, we will provide a (π + 1)th order sufficient condition for π₯ to be a sharp local Pareto solution of (4) for a πΆπ,1 function π. ∀π₯ ∈ π΅ (π₯, πΏ) . (37) On the other hand, for any π ∈ πΆ, one has (31) which is a contradiction with (29). The proof is completed. π (π₯) − π (π₯) = β¨π∗ , π (π₯) − π (π₯)β© σ΅© σ΅© ≤ β¨π∗ , π (π₯) − π (π₯) + πβ© ≤ σ΅©σ΅©σ΅©π (π₯) − π (π₯) + πσ΅©σ΅©σ΅© . (38) Abstract and Applied Analysis 5 This implies that (32) holds. It remains to show that π₯ is a local Pareto solution of (4). Let π₯ ∈ π΅(π₯, πΏ) such that π(π₯) ≤πΆ π(π₯). Then, βπ(π₯) − π(π₯)β+ = 0. It follows from (32) that π₯ = π₯, and hence π(π₯) = π(π₯). This shows that π₯ is a local Pareto solution of (4). It follows that In Theorem 6, if π is a πΆ-convex πΆπ,1 function, then π₯ is a global Pareto solution of (4). On the other hand, since ∑ππ=1 (1/π!)π·π π(π₯) = 0, with Corollary 3, we can assume that for any π₯ ∈ π π close to π₯, there exists π-linear symmetric and continuous mapping π(π₯) from (π π )π+1 to π π such that limπ₯ → π₯ π(π₯) = 0 and Theorem 7. Let π be a πΆ-convex πΆπ,1 function from π π to π π , πΆ a closed convex cone of π π , and π₯ ∈ π π . Suppose that there exists π∗ ∈ πΆ+ with βπ∗ β = 1 such that ∑ππ=1 (1/π!)π∗ β π·π π(π₯) = 0 and that ππ+1 π(π₯) is positively definite. Then, π₯ is a global Pareto solution of (4), and there exists π0 ∈ (0, +∞) such that π0 βπ₯ − π₯β ≤ max {[π (π₯) − 1/(π+1) π (π₯)]+ , [π (π₯) − π (π₯)]+ } , π πΏ ∀π₯ ∈ π π . π₯−π₯ ≤ [π (π₯ + πΏ ) − π (π₯)] βπ₯ − π₯β + πΏ ≤ (1 − ) [π (π₯) − π (π₯)]+ βπ₯ − π₯β πΏ + [π (π₯) − π (π₯)]+ βπ₯ − π₯β = (40) πΏ [π (π₯) − π (π₯)]+ . βπ₯ − π₯β With ∑ππ=1 (1/π!)π∗ β π·π π(π₯) = 0 in Theorem 7 replaced by a stronger assumption, we have the following sufficient condition for sharp ideal solutions of (4). Theorem 8. Let π be a πΆπ,1 function from π π to π π , πΆ a closed convex cone of π π , and π₯ ∈ π π . Suppose that ∑ππ=1 (1/π!)π·π π(π₯) = 0 and that ππ+1 π is positively definite at π₯ with respect to the ordering cone πΆ. Then, there exist π, πΏ ∈ (0, +∞) such that ∀π₯ ∈ π΅ (π₯, πΏ) , 1/(π+1) π βπ₯ − π₯β ≤ [π (π₯) − π (π₯)]+ , ∀π₯ ∈ π΅ (π₯, πΏ) . (41) (42) Proof. By Theorem 6, we need only to show that there exists πΏ > 0 such that (41) holds. Since ππ+1 π(π₯) is positively definite, there exists π > 0 such that 1 ππ+1 π (π₯) (βπ+1 ) + 2ππ΅π π ⊂ πΆ, (π + 1)! 1 π(π+1) π (π₯) ((π₯ − π₯)π+1 ) (π + 1)! π+1 + π (π₯) (π₯ − π₯) (44) (45) . ∀β ∈ ππ π . π (π₯) − π (π₯) ∈ 1 ππ+1 π (π₯) ((π₯ − π₯)π+1 ) (π + 1)! + πβπ₯ − π₯β π+1 π΅π π , (43) (46) ∀π₯ ∈ π΅ (π₯, πΏ) . This and (44) imply that (41) holds. The proof is completed. 5. (π+1)th Order Necessary Conditions for Strongly πΆ-Quasiconvex Functions We recall that a function π from π π to π is quasiconvex if, for every π₯, π¦ ∈ π π and for every π ∈ (0, 1), one has π(ππ₯ + (1 − π)π¦) ≤ max{π(π₯), π(π¦)}. Inspired by this, we introduce the notion of strong πΆ-quasiconvexity for functions from π π to π π . A function π from π π to π π is said to be strongly πΆquasiconvex if, for every π₯, π¦ ∈ π π and for every π ∈ (0, 1), one has π (ππ₯ + (1 − π) π¦) ∈ {π (π₯) , π (π¦)} − πΆ. Hence, ππ+1 πΏπ βπ₯ − π₯β ≤ [π(π₯) − π(π₯)]+ . Letting π0 := min{π, ππ+1 πΏπ }, it follows from (32) that (39) holds. The proof is completed. π (π₯) ≤πΆ π (π₯) , π (π₯) − π (π₯) ∈ ∀π₯ ∈ π π . Hence, there exists πΏ > 0 such that (39) Proof. Similar to the proof of Theorem 6, one can show that (39) implies that π₯ is a global Pareto solution of (4). It remains to show that (39) holds. By Theorem 6, there exist π, πΏ ∈ (0, +∞) such that (32) holds. Since π is πΆ-convex, it is easy to verify that π₯ σ³¨→ [π(π₯) − π(π₯)]+ is a convex function. Let π₯ ∈ π π \ π΅(π₯, πΏ). Then, π+1 π+1 1 ππ+1 π (π₯) (π₯π+1 ) + πβπ₯βπ+1 π΅π π ⊂ πΆ, (π + 1)! (47) Using the generalized Hessian (see [20]), Luc [1] gave a second-order criterion for quasiconvex functions. We will give a (π + 1)th order necessary codition for a function to be strongly πΆ-quasiconvex. Theorem 9. Let π be a strongly πΆ-quasiconvex function from π π to π π , π an odd number and πΆ ⊂ π π the closed pointed ordering cone. Then, for any π₯, π’ ∈ π π with π·π π(π₯)(π’π ) = 0 (π = 1, . . . , π), there exists π΄ 0 ∈ ππ+1 π(π₯) such that π΄ 0 (π’π+1 ) ∈ πΆ. Proof. Suppose that the conclusion is not true. Then, there exist some π₯, π’ ∈ π π , with π·π π(π₯)(π’π ) = 0 (π = 1, . . . , π) such that ππ+1 π(π₯)(π’π+1 ) ⊂ π π \ πΆ. Since π π \ πΆ is open and ππ+1 π(π₯)(π’π+1 ) is compact, there exists π > 0 such that ππ+1 π (π₯) (π’π+1 ) + ππ΅πΏ((π π )π+1 ,π π ) (π’π+1 ) ⊂ π π \ πΆ. (48) Since ππ+1 π(⋅) is upper continuous, for the previous π, there exists π‘0 > 0 such that ππ+1 π (π₯ + π‘π’) ⊂ ππ+1 π (π₯) + ππ΅πΏ((π π )π+1 ,π π ) , (49) 6 Abstract and Applied Analysis for any π‘ ∈ [−π‘0 , π‘0 ]. Noting that ππ+1 π(π₯) is closed convex, from (48) and (49), we have clco ππ+1 π (π₯ − π‘π’, π₯ + π‘π’) (π’π+1 ) ⊂ ππ+1 π (π₯) (π’π+1 ) + ππ΅πΏ((π π )π+1 ,π π ) (π’π+1 ) ⊂ π π \ πΆ. (50) From Theorem 2, for any π‘ ∈ [−π‘0 , π‘0 ], we can take π΄ π‘ ∈ clco ππ+1 π(π₯, π₯ + π‘π’) such that π 1 1 π (π₯+π‘π’)−π (π₯) = ∑ π‘π π·π π (π₯) (π’π )+ π‘π+1 π΄ π‘ (π’π+1 ) + 1)! π! (π π=1 = 1 π‘π+1 π΄ π‘ (π’π+1 ) ∈ π π \ πΆ. (π + 1)! (51) Noting that π + 1 is even, we have π(π₯ + π‘π’) − π(π₯) ∈ π π \ πΆ and π(π₯ − π‘π’) − π(π₯) ∈ π π \ πΆ, for all π‘ ∈ [−π‘0 , π‘0 ]. On the other hand, since π is πΆ-quasiconvex and π(π₯) = π((1/2)(π₯ + π‘π’) + (1/2)(π₯ − π‘π’)), one has π(π₯ + π‘π’) − π(π₯) ∈ πΆ or π(π₯ − π‘π’) − π(π₯) ∈ πΆ. This is a contradiction. If π = 1 and πΆ = π + , then πΆ+ = π + . We have the following. Corollary 10 (see [1]). Let π be a quasiconvex function from π π to π and π an odd number. Then, for any π₯, π’ ∈ π π with π·π π(π₯)(π’π ) = 0 (π = 1, . . . , π), one has π·+π+1 π(π₯; π’) ≥ 0, where π·+π+1 π(π₯; π’) := max{π΄(π’, . . . , π’) : π΄ ∈ ππ+1 π(π₯)}. Acknowledgments This research was supported by the National Natural Science Foundations, China (Grant no. 11061039, 11061038, and 11261067), an internal Grant of Hong Kong Polytechnic University (G-YF17), and IRTSTYN. References [1] D. T. Luc, “Taylor’s formula for πΆπ,1 functions,” SIAM Journal on Optimization, vol. 5, no. 3, pp. 659–669, 1995. [2] A. GoΜpfert, H. Riahi, C. Tammer, and C. ZaΜlinescu, Variational Methods in Partially Ordered Spaces, vol. 17 of CMS Books in Mathematics, Springer, New York, NY, USA, 2003. [3] V. V. Gorokhovik, Vypuklye i Negladkie Zadachi Vektornoi Optimizatsii, Nauka i Tekhnika, Minsk, Russia, 1990. [4] J. Jahn, Vector Optimization: Theory, Applications and Extensions, Operations Research and Decision Theory, Springer, New York, NY, USA, 2004. [5] J. J. Ye and Q. J. Zhu, “Multiobjective optimization problem with variational inequality constraints,” Mathematical Programming, vol. 96, no. 1, pp. 139–160, 2003. [6] X. Y. Zheng and X. Q. Yang, “Weak sharp minima for semiinfinite optimization problems with applications,” SIAM Journal on Optimization, vol. 18, no. 2, pp. 573–588, 2007. [7] M. Studniarski and D. E. Ward, “Weak sharp minima: characterizations and sufficient conditions,” SIAM Journal on Control and Optimization, vol. 38, no. 1, pp. 219–236, 1999. [8] M. I. Henig, “Existence and characterization of efficient decisions with respect to cones,” Mathematical Programming, vol. 23, no. 1, pp. 111–116, 1982. [9] G. Isac, “Pareto optimization in infinite-dimensional spaces: the importance of nuclear cones,” Journal of Mathematical Analysis and Applications, vol. 182, no. 2, pp. 393–404, 1994. [10] M. Minami, “Weak Pareto-optimal necessary conditions in a nondifferentiable multiobjective program on a Banach space,” Journal of Optimization Theory and Applications, vol. 41, no. 3, pp. 451–461, 1983. [11] A. Zaffaroni, “Degrees of efficiency and degrees of minimality,” SIAM Journal on Control and Optimization, vol. 42, no. 3, pp. 1071–1086, 2003. [12] X. Y. Zheng, X. M. Yang, and K. L. Teo, “Sharp minima for multiobjective optimization in Banach spaces,” Set-Valued Analysis, vol. 14, no. 4, pp. 327–345, 2006. [13] X. Y. Zheng and K. F. Ng, “The Fermat rule for multifunctions on Banach spaces,” Mathematical Programming, vol. 104, no. 1, pp. 69–90, 2005. [14] G. Y. Chen, X. X. Huang, and X. Q. Yang, Vector Optimization: Set-Valued and Variational Analysis, Springer, New York, NY, USA, 2005. [15] J. V. Burke and S. Deng, “Weak sharp minima revisited. II. Application to linear regularity and error bounds,” Mathematical Programming, vol. 104, no. 2-3, pp. 235–261, 2005. [16] R. Ellaia and A. Hassouni, “Characterization of nonsmooth functions through their generalized gradients,” Optimization, vol. 22, no. 3, pp. 401–416, 1991. [17] X. Y. Zheng and X. Q. Yang, “Conic positive definiteness and sharp minima of fractional orders in vector optimization problems,” Journal of Mathematical Analysis and Applications, vol. 391, no. 2, pp. 619–629, 2012. [18] F. H. Clarke, Optimization and Nonsmooth Analysis, John Wiley & Sons, New York, NY, USA, 1983. [19] S. Winfried, Nonsmooth Analysis, Springer, Berlin, Germany, 2007. [20] J. B. Hiriart-Urruty, J. J. Strodiot, and V. H. Nguyen, “Generalized Hessian matrix and second-order optimality conditions for problems with πΆ1,1 data,” Applied Mathematics and Optimization, vol. 11, no. 1, pp. 43–56, 1984. Advances in Operations Research Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Advances in Decision Sciences Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Mathematical Problems in Engineering Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Journal of Algebra Hindawi Publishing Corporation http://www.hindawi.com Probability and Statistics Volume 2014 The Scientific World Journal Hindawi Publishing Corporation http://www.hindawi.com Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 International Journal of Differential Equations Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Volume 2014 Submit your manuscripts at http://www.hindawi.com International Journal of Advances in Combinatorics Hindawi Publishing Corporation http://www.hindawi.com Mathematical Physics Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Journal of Complex Analysis Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 International Journal of Mathematics and Mathematical Sciences Journal of Hindawi Publishing Corporation http://www.hindawi.com Stochastic Analysis Abstract and Applied Analysis Hindawi Publishing Corporation http://www.hindawi.com Hindawi Publishing Corporation http://www.hindawi.com International Journal of Mathematics Volume 2014 Volume 2014 Discrete Dynamics in Nature and Society Volume 2014 Volume 2014 Journal of Journal of Discrete Mathematics Journal of Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Applied Mathematics Journal of Function Spaces Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Optimization Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014