Hindawi Publishing Corporation Abstract and Applied Analysis Volume 2013, Article ID 602735, 21 pages http://dx.doi.org/10.1155/2013/602735 Research Article Nonsingularity Conditions for FB System of Reformulating Nonlinear Second-Order Cone Programming Shaohua Pan,1 Shujun Bi,1 and Jein-Shan Chen2 1 2 Department of Mathematics, South China University of Technology, Guangzhou 510641, China Mathematics Division, National Center for Theoretical Sciences, Taipei 11677, Taiwan Correspondence should be addressed to Jein-Shan Chen; jschen@math.ntnu.edu.tw Received 2 June 2012; Accepted 9 December 2012 Academic Editor: Narcisa C. Apreutesei Copyright © 2013 Shaohua Pan et al. 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. This paper is a counterpart of Bi et al., 2011. For a locally optimal solution to the nonlinear second-order cone programming (SOCP), specifically, under Robinson’s constraint qualification, we establish the equivalence among the following three conditions: the nonsingularity of Clarke’s Jacobian of Fischer-Burmeister (FB) nonsmooth system for the Karush-Kuhn-Tucker conditions, the strong second-order sufficient condition and constraint nondegeneracy, and the strong regularity of the Karush-Kuhn-Tucker point. 1. Introduction The nonlinear second-order cone programming (SOCP) problem can be stated as By introducing a slack variable to the second constraint, the SOCP (1) is equivalent to min đ,đĽ∈Rđ minđ đ (đ) â (đ) = 0, where đ : Rđ → R, â : Rđ → Rđ , and đ : Rđ → Rđ are given twice continuously differentiable functions, and K is the Cartesian product of some second-order cones, that is, đĽ ∈ K. In this paper, we will concentrate on this equivalent formulation of problem (1). Let đż : Rđ × Rđ × Rđ × Rđ × K → R be the Lagrangian function of problem (4) đż (đ, đĽ, đ, đ , đŚ) := đ (đ) + â¨đ, â (đ)⊠+ â¨đ (đ) − đĽ, đ ⊠− â¨đĽ, đŚâŠ, (5) (2) đđ with đ1 + ⋅ ⋅ ⋅ + đđ = đ and K being the second-order cone (SOC) in Rđđ defined by óľŠ óľŠ Kđđ := {(đĽđ1 , đĽđ2 ) ∈ R × Rđđ −1 | đĽđ1 ≥ óľŠóľŠóľŠóľŠđĽđ2 óľŠóľŠóľŠóľŠ} . đ (đ) − đĽ = 0, (1) đ (đ) ∈ K, K := Kđ1 × Kđ2 × ⋅ ⋅ ⋅ × Kđđ , (4) s.t. â (đ) = 0, đ∈R s.t. đ (đ) and denote by NK (đĽ) the normal cone of K at đĽ in the sense of convex analysis [1]: NK (đĽ) = { (3) {đ ∈ Rđ : â¨đ, đ§ − đĽâŠ ≤ 0 ∀đ§ ∈ K} , if đĽ ∈ K, 0, if đĽ ∉ K. (6) 2 Abstract and Applied Analysis Then the Karush-Kuhn-Tucker (KKT) conditions for (4) take the following form: Jđ,đĽ đż (đ, đĽ, đ, đ , đŚ) = 0, đ (đ) − đĽ = 0, â (đ) = 0, −đŚ ∈ NK (đĽ) , (7) where Jđ,đĽ đż(đ, đĽ, đ, đ , đŚ) is the derivative of đż at (đ, đĽ, đ, đ , đŚ) with respect to (đ, đĽ). Recall that đsoc is an SOC complementarity function associated with the cone K if đsoc (đĽ, đŚ) = 0 ⇐⇒ đĽ ∈ K, đŚ ∈ K, â¨đĽ, đŚâŠ = 0 ⇐⇒ −đŚ ∈ NK (đĽ) . (8) With an SOC complementarity function đsoc associated with K, we may reformulate the KKT optimality conditions in (7) as the following nonsmooth system: Jđ,đĽ đż (đ, đĽ, đ, đ , đŚ) ] â (đ) ] = 0. ] đ (đ) − đĽ soc đ (đĽ, đŚ) ] [ [ đ¸ (đ, đĽ, đ, đ , đŚ) := [ [ (9) The most popular SOC complementarity functions include the vector-valued natural residual (NR) function and Fischer-Burmeister (FB) function, respectively, defined as đNR (đĽ, đŚ) := đĽ − ΠK (đĽ − đŚ) , đFB (đĽ, đŚ) := (đĽ + đŚ) − √đĽ2 + đŚ2 , ∀đĽ, đŚ ∈ Rđ , ∀đĽ, đŚ ∈ Rđ , (10) where ΠK (⋅) is the projection operator onto the closed convex cone K, đĽ2 = đĽ â đĽ means the Jordan product of đĽ and itself, and √đĽ denotes the unique square root of đĽ ∈ K. It turns out that the FB SOC complementarity function đFB enjoys almost all favorable properties of the NR SOC complementarity function đNR (see [2]). Also, the squared norm of đFB induces a continuously differentiable merit function with globally Lipschitz continuous derivative [3, 4]. This greatly facilitates the globalization of the semismooth Newton method [5, 6] for solving the FB nonsmooth system of KKT conditions: Jđ,đĽ đż (đ, đĽ, đ, đ , đŚ) ] â (đ) ] = 0. ] đ (đ) − đĽ đ (đĽ, đŚ) FB [ ] [ đ¸FB (đ, đĽ, đ, đ , đŚ) := [ [ (11) Recently, with the help of [7, Theorem 30] and [8, Lemma 11], Wang and Zhang [9] gave a characterization for the strong regularity of the KKT point of the SOCP (1) via the nonsingularity study of Clarke’s Jacobian of the NR nonsmooth system Jđ,đĽ đż (đ, đĽ, đ, đ , đŚ) ] â (đ) ] = 0. ] đ (đ) − đĽ đNR (đĽ, đŚ) ] [ [ đ¸NR (đ, đĽ, đ, đ , đŚ) := [ [ (12) They showed that the strong regularity of the KKT point, the nonsingularity of Clarke’s Jacobian of đ¸NR at the KKT point, and the strong second-order sufficient condition and constraint nondegeneracy [7] are all equivalent. These nonsingularity conditions are better structured than those of [10] for the nonsingularity of the đľ-subdifferential of the NR system. Then, it is natural to ask the following: is it possible to obtain a characterization for the strong regularity of the KKT point by studying the nonsingularity of Clarke’s Jacobian of đ¸FB . Note that up till now one even does not know whether the đľ-subdifferential of the FB system is nonsingular or not without the strict complementarity assumption. In this work, for a locally optimal solution to the nonlinear SOCP (4), under Robinson’s constraint qualification, we show that the strong second-order sufficient condition and constraint nondegeneracy introduced in [7], the nonsingularity of Clarke’s Jacobian of đ¸FB at the KKT point, and the strong regularity of the KKT point are equivalent to each other. This, on the one hand, gives a new characterization for the strong regularity of the KKT point and, on the other hand, provides a mild condition to guarantee the quadratic convergence rate of the semismooth Newton method [5, 6] for the FB system. Note that parallel results are obtained recently for the FB system of the nonlinear semidefinite programming (see [11]); however, we do not duplicate them. As will be seen in Sections 3 and 4, the analysis techniques here are totally different from those in [11]. It seems hard to put them together in a unified framework under the Euclidean Jordan algebra. The main reason causing this is due to completely different analysis when dealing with the Clarke Jacobians associated with FB SOC complementarity function and FB semidefinite cone complementarity function. Throughout this paper, đź denotes an identity matrix of appropriate dimension, Rđ (đ > 1) denotes the space of đdimensional real column vectors, and Rđ1 × ⋅ ⋅ ⋅ × Rđđ is identified with Rđ1 +⋅⋅⋅+đđ . Thus, (đĽ1 , . . . , đĽđ ) ∈ Rđ1 × ⋅ ⋅ ⋅ × Rđđ is viewed as a column vector in Rđ1 +⋅⋅⋅+đđ . The notations int Kđ , bd Kđ , and bd+ Kđ denote the interior, the boundary, and the boundary excluding the origin of Kđ , respectively. For any đĽ ∈ Rđ , we write đĽ ⪰Kđ 0 (resp., đĽ âťKđ 0) if đĽ ∈ Kđ (resp., đĽ ∈ int Kđ ). For any given real symmetric matrix đ´, we write đ´ ⪰ 0 (resp., đ´ âť 0) if đ´ is positive semidefinite (resp., positive definite). In addition, Jđ đ(đ) and J2đđ đ(đ) denote the derivative and the second-order derivative, respectively, of a twice differentiable function đ with respect to the variable đ. 2. Preliminary Results First we recall from [12] the definition of Jordan product and spectral factorization. Definition 1. The Jordan product of đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) ∈ R × Rđ−1 is given by đĽ â đŚ := (â¨đĽ, đŚâŠ , đĽ1 đŚ2 + đŚ1 đĽ2 ) . (13) Unlike scalar or matrix multiplication, the Jordan product is not associative in general. The identity element under this product is đ := (1, 0, . . . , 0)đ ∈ Rđ , that is, đ â đĽ = đĽ for all Abstract and Applied Analysis 3 đĽ ∈ Rđ . For each đĽ = (đĽ1 , đĽ2 ) ∈ R × Rđ−1 , we define the associated arrow matrix by đż đĽ := [ đĽ1 đĽ2đ ]. đĽ2 đĽ1 đź (14) Then it is easy to verify that đż đĽ đŚ = đĽ â đŚ for any đĽ, đŚ ∈ Rđ . Recall that each đĽ = (đĽ1 , đĽ2 ) ∈ R × Rđ−1 admits a spectral factorization, associated with Kđ , of the form đĽ= đ 1 (đĽ) đ˘đĽ(1) + đ 2 (đĽ) đ˘đĽ(2) , đ˘đĽ(1) , đ˘đĽ(2) (15) đ ∈ R are the spectral where đ 1 (đĽ), đ 2 (đĽ) ∈ R and values and the associated spectral vectors of đĽ, respectively, with respect to the Jordan product, defined by óľŠ óľŠ đ đ (đĽ) := đĽ1 + (−1)đ óľŠóľŠóľŠđĽ2 óľŠóľŠóľŠ , đ˘đĽ(đ) 1 1 := ( đ ), 2 (−1) đĽĚ2 (16) for đ = 1, 2, with đĽĚ2 = đĽ2 /âđĽ2 â if đĽ2 ≠ 0 and otherwise being any vector in Rđ−1 satisfying âđĽĚ2 â = 1. đ Definition 2. The determinant of a vector đĽ ∈ R is defined as det(đĽ) := đ 1 (đĽ)đ 2 (đĽ), and a vector đĽ is said to be invertible if its determinant det(đĽ) is nonzero. By the formula of spectral factorization, it is easy to compute that the projection of đĽ ∈ Rđ onto the closed convex cone Kđ , denoted by ΠKđ (đĽ), has the expression The following lemma states a result for the arrow matrices associated with đĽ, đŚ ∈ Rđ and đ§ ⪰Kđ √đĽ2 + đŚ2 , which will be used in the next section to characterize an important property for the elements of Clarke’s Jacobian of đFB at a general point. Lemma 4. For any given đĽ, đŚ ∈ Rđ and đ§ âťKđ 0, if đ§2 ⪰Kđ đĽ2 + đŚ2 , then óľŠóľŠ óľŠóľŠ −1 óľŠ ≤ 1, óľŠóľŠ[đż đ§ đż đĽ đż−1 đ§ đż đŚ ]óľŠ óľŠ2 óľŠ (20) where âđ´â2 means the spectral norm of a real matrix đ´. Consequently, it holds that óľŠóľŠ −1 óľŠóľŠ √ óľŠóľŠđż đ§ đż đĽ Δđ˘ + đż−1 óľŠ ≤ âΔđ˘â2 + âΔđŁâ2 , đ§ đż đŚ ΔđŁóľŠ óľŠ óľŠ ∀Δđ˘, ΔđŁ ∈ Rđ . (21) −1 Proof. Let đ´ = [đż−1 đ§ đż đĽ đż đ§ đż đŚ ]. From [13, Proposition 3.4], it follows that 2 2 −1 −1 2 −1 đ´đ´đ = đż−1 đ§ (đż đĽ + đż đŚ ) đż đ§ ⪯ đż đ§ đż đ§ đż đ§ = đź. (22) This shows that âđ´â2 ≤ 1, and the first part follows. Note that, for any đ ∈ R2đ , óľŠóľŠ óľŠóľŠ2 óľŠ óľŠ2 óľŠ óľŠ 2 đ đ đ óľŠóľŠđ´đóľŠóľŠ = đ đ´ đ´đ ≤ đ max (đ´ đ´) óľŠóľŠóľŠđóľŠóľŠóľŠ ≤ óľŠóľŠóľŠđóľŠóľŠóľŠ . (23) By letting đ = (Δđ˘, ΔđŁ) ∈ Rđ × Rđ , we immediately obtain the second part. ΠKđ (đĽ) = max (0, đ 1 (đĽ)) đ˘đĽ(1) + max (0, đ 2 (đĽ)) đ˘đĽ(2) . (17) The following two lemmas state the properties of đĽ, đŚ with đĽ2 + đŚ2 ∈ bd Kđ which are often used in the subsequent sections. The proof of Lemma 5 is given in [3, Lemma 2]. Define |đĽ| := 2ΠKđ (đĽ) − đĽ. Then, using the expression of ΠKđ (đĽ), it follows that Lemma 5. For any đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) ∈ R × Rđ−1 with đĽ2 + đŚ2 ∈ bd Kđ , one has óľ¨ óľ¨ óľ¨ óľ¨ |đĽ| = óľ¨óľ¨óľ¨đ 1 (đĽ)óľ¨óľ¨óľ¨ đ˘đĽ(1) + óľ¨óľ¨óľ¨đ 2 (đĽ)óľ¨óľ¨óľ¨ đ˘đĽ(2) . (18) óľŠ óľŠ2 đĽ12 = óľŠóľŠóľŠđĽ2 óľŠóľŠóľŠ , The spectral factorization of the vectors đĽ, đĽ2 , √đĽ and the matrix đż đĽ have various interesting properties (see [13]). We list several properties that we will use later. đĽ1 đŚ1 = đĽ2đ đŚ2 , Property 3. For any đĽ = (đĽ1 , đĽ2 ) ∈ R × Rđ−1 with spectral factorization (15), we have the following. (a) đĽ2 = đ21 (đĽ)đ˘đĽ(1) + đ22 (đĽ)đ˘đĽ(2) ∈ Kđ . (b) If đĽ ∈ Kđ , then 0 ≤ đ 1 (đĽ) ≤ đ 2 (đĽ) and √đĽ = √đ 1 (đĽ)đ˘đĽ(1) + √đ 2 (đĽ)đ˘đĽ(2) . đ (c) If đĽ ∈ int K , then 0 < đ 1 (đĽ) ≤ đ 2 (đĽ) and đż đĽ is invertible with đż−1 đĽ = đĽ1 1 [ [ det (đĽ) −đĽ2 [ −đĽ2đ ] đĽ đĽđ ] . det (đĽ) đź+ 2 2 đĽ1 đĽ1 ] óľŠ óľŠ2 đŚ12 = óľŠóľŠóľŠđŚ2 óľŠóľŠóľŠ , đĽ1 đŚ2 = đŚ1 đĽ2 . (24) Lemma 6. For any đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) ∈ R × Rđ−1 , let đ¤ = (đ¤1 , đ¤2 ) := đĽ2 + đŚ2 . (a) If đ¤ ∈ bd Kđ , then for any đ = (đ1 , đ2 ), â = (â1 , â2 ) ∈ R × Rđ−1 , it holds that đ (đĽ1 đĽ2 + đŚ1 đŚ2 ) (đĽ1 đ2 + đ1 đĽ2 + đŚ1 â2 + â1 đŚ2 ) = (đĽ12 + đŚ12 ) (đĽđ đ + đŚđ â) . (25) (b) If đ¤ ∈ bd+ Kđ , then the following four equalities hold (19) đĽ1 đ¤2 óľŠóľŠ óľŠóľŠ = đĽ2 , óľŠóľŠđ¤2 óľŠóľŠ đĽ2đ đ¤2 óľŠóľŠ óľŠóľŠ = đĽ1 , óľŠóľŠđ¤2 óľŠóľŠ (d) đż đĽ ⪰ 0 (resp., đż đĽ âť 0) if and only if đĽ ∈ Kđ (resp., đĽ ∈ int Kđ ). đŚ1 đ¤2 óľŠóľŠ óľŠóľŠ = đŚ2 , óľŠóľŠđ¤2 óľŠóľŠ đŚ2đ đ¤2 óľŠóľŠ óľŠóľŠ = đŚ1 ; óľŠóľŠđ¤2 óľŠóľŠ (26) 4 Abstract and Applied Analysis and consequently the expression of đFB (đĽ, đŚ) can be simplified as đĽ1 + đŚ1 − √đĽ12 + đŚ12 đFB (đĽ, đŚ) = (đĽ + đŚ − đĽ1 đĽ2 + đŚ1 đŚ2 ) . 2 2 √đĽ12 + đŚ12 (27) Proof. (a) The result is direct by the equalities of Lemma 5 since đĽ2 + đŚ2 ∈ bd Kđ . (b) Since đ¤ ∈ bd+ Kđ , we must have đ¤2 = 2(đĽ1 đĽ2 + đŚ1 đŚ2 ) ≠ 0. Using Lemma 5, đ¤2 = 2(đĽ1 đĽ2 + đŚ1 đŚ2 ) and âđ¤2 â = đ¤1 = 2(đĽ12 + đŚ12 ), we easily obtain the first part. Note that đFB (đĽ, đŚ) = (đĽ + đŚ) − √đ¤. Using Property 3(b) and Lemma 5 yields (27). đ When đĽ, đŚ ∈ bd K satisfies the complementary condition, we have the following result. Lemma 7. For any given đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) ∈ R × Rđ−1 , if đĽ, đŚ ∈ bd Kđ and â¨đĽ, đŚâŠ = 0, then there exists a constant đź > 0 such that đĽ1 = đźđŚ1 and đĽ2 = −đźđŚ2 . Proof. Since đĽ, đŚ ∈ bd Kđ , we have that đĽ1 = âđĽ2 â and đŚ1 = âđŚ2 â, and consequently, óľŠ óľŠóľŠ óľŠ (28) 0 = â¨đĽ, đŚâŠ = đĽ1 đŚ1 + đĽ2đ đŚ2 = óľŠóľŠóľŠđĽ2 óľŠóľŠóľŠ óľŠóľŠóľŠđŚ2 óľŠóľŠóľŠ + đĽ2đ đŚ2 . This means that there exists đź > 0 such that đĽ2 = −đźđŚ2 , and then đĽ1 = đźđŚ1 . For the concept of (strong) semismoothness, please refer to the literature [5, 6]. Unless otherwise stated, in the rest of this paper, for any đĽ ∈ Rđ (đ > 1), we write đĽ = (đĽ1 , đĽ2 ), where đĽ1 is the first component of đĽ and đĽ2 is a column vector consisting of the remaining đ − 1 entries of đĽ. For any đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) ∈ R × Rđ−1 , let đ¤ = đ¤ (đĽ, đŚ) := đĽ2 + đŚ2 , if đ¤2 ≠ 0, Proposition 9. For any given đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) ∈ ó¸ R × Rđ−1 , the directional derivative đFB ((đĽ, đŚ); (đ, â)) of đFB at (đĽ, đŚ) with the direction (đ, â) has the following form. ó¸ ((đĽ, đŚ); (đ, â)) = đFB (đ, â). (a) If (đĽ, đŚ) = (0, 0), then đFB ó¸ (b) If đĽ2 + đŚ2 ∈ int Kđ , then đFB ((đĽ, đŚ); (đ, â)) = (đź − −1 −1 đż đ§ đż đĽ )đ + (đź − đż đ§ đż đŚ )â. (c) If đĽ2 + đŚ2 ∈ bd+ Kđ , then ó¸ đFB ((đĽ, đŚ); (đ, â)) = (đ + â) − Definition 8. We say that đ§ is a strongly regular solution of the generalized equation (29) if there exist neighborhood B of the origin 0 ∈ Z and V of đ§ such that for every đż ∈ B, the linearized generalized equation đż ∈ đ(đ§) + Jđ§ đ(đ§)(đ§ − đ§) + Nđˇ(đ§) has a unique solution in V, denoted by đ§V (đż), and the mapping đ§V : B → V is Lipschitz continuous. To close this section, we recall from [15] Clarke’s (generalized) Jacobian of a locally Lipschitz mapping. Let đ ⊂ Rđ be an open set and Ξ : đ → Rđ a locally Lipschitz continuous function on đ. By Rademacher’s theorem, Ξ is almost everywhere đš(reĚchet)-differentiable in đ. We denote by đΞ the set of points in đ where Ξ is đš-differentiable. Then Clarke’s Jacobian of Ξ at đŚ is defined by đΞ(đŚ) := conv{đđľ Ξ(đŚ)}, where “conv” means the convex hull, and đľsubdifferential đđľ Ξ(đŚ), a name coined in [16], has the form đ đ đ đđľ Ξ (đŚ) := {đ : đ = lim JđŚ Ξ (đŚ ) , đŚ ół¨→ đŚ, đŚ ∈ đΞ } . đ→∞ (30) đ§ = đ§ (đĽ, đŚ) = √đ¤ (đĽ, đŚ). The function đFB is directionally differentiable everywhere by [2, Corollary 3.3]. But, to the best of our knowledge, the expression of its directional derivative is not given in the literature. In this section, we derive its expression and then prove that the đľ-subdifferential of đFB at a general point coincides with that of its directional derivative function at the origin. Throughout this section, we assume that K = Kđ . (29) where đ is a continuously differentiable mapping from a finite dimensional real vector space Z to itself, đˇ is a closed convex set in Z, and Nđˇ(đ§) is the normal cone of đˇ at đ§. As will be shown in Section 4, the KKT condition (7) can be written in the form of (29). (31) 3. Directional Derivative and đľ-Subdifferential Next we recall from [14] the strong regularity for a solution of generalized equation 0 ∈ đ (đ§) + Nđˇ (đ§) , đ¤ Ě2 := óľŠ 2 óľŠ , đ¤ óľŠóľŠđ¤2 óľŠóľŠ óľŠ óľŠ − đ (đ, â) 1 đĽđđ + đŚ2đ â2 0 )+ 2 2 ( ) ( Ě2 đ¤ −Ě đ¤2 2 2√đĽ12 + đŚ12 1 2√đĽ12 + đŚ12 đĽđđ + đŚđ â ( ), 2đĽ1 đ2 + đ1 đĽ2 + 2đŚ1 â2 + â1 đŚ2 (32) where đ = (đ1 , đ2 ), â = (â1 , â2 ) ∈ R × Rđ−1 , and đ : Rđ × Rđ → R is defined by đ (đ, â) := √(đĽ1 đ1 −đĽ2đ đ2 +đŚ1 â1 −đŚ2đ â2)2 + óľŠóľŠóľŠóľŠđĽ1 â2 −â1 đĽ2 +đ1 đŚ2 −đŚ1 đ2óľŠóľŠóľŠóľŠ2 √đĽ12 + đŚ12 . (33) Proof. Part (a) is immediate by noting that đFB is a positively homogeneous function. Part (b) is due to [13, Proposition 5.2]. We next prove part (c) by two subcases as shown in the following. In the rest of proof, we let đ 1 , đ 2 with đ 1 ≤ đ 2 Abstract and Applied Analysis 5 denote the spectral values of đ¤. Since đ¤ = đĽ2 + đŚ2 ∈ bd+ Kđ , we have đ¤2 ≠ 0, and from Lemma 6(b) it follows that óľŠ óľŠ óľŠ óľŠ đ¤1 = óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ = 2 óľŠóľŠóľŠđĽ1 đĽ2 + đŚ1 đŚ2 óľŠóľŠóľŠ [ − lim [ đĄ↓0 óľŠ óľŠ Ě2 + đŚ12 đ¤ Ě2 óľŠóľŠóľŠóľŠ = 2 (đĽ12 + đŚ12 ) , = 2 óľŠóľŠóľŠóľŠđĽ12 đ¤ óľŠ óľŠ đ 1 = đ¤1 − óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ = 0, = (đ2 + â2 ) − óľŠ óľŠ đ 2 = đ¤1 + óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ = 4 (đĽ12 + đŚ12 ) . (34) 2 + đ (c.1): (đĽ + đĄđ) + (đŚ + đĄâ) ∈ bd K for sufficiently small đĄ > 0. In this case, from Lemma 6(b), we know that đFB (đĽ + đĄđ, đŚ + đĄâ) has the following expression: √đĽ12 + đŚ12 đĽ1 đĽ2 +đŚ1 đŚ2 2 2 đĄ√(đĽ1 +đĄđ1 ) + (đŚ1 +đĄâ1 ) = (đ2 + â2 ) − + 2 [ đ1 đĽ2 + đĽ1 đ2 + đŚ1 â2 + â1 đŚ2 − đĽ1 đĽ2 +đŚ1 đŚ2 ] ] đĄ√đĽ12 +đŚ12 ] đ1 đĽ2 + đĽ1 đ2 + đŚ1 â2 + â1 đŚ2 √đĽ12 + đŚ12 (đĽ1 đĽ2 + đŚ1 đŚ2 ) (đĽ1 đ1 + đŚ1 â1 ) (đĽ12 + đŚ12 ) √đĽ12 + đŚ12 = (đ2 + â2 ) − đĽ1 đ2 + đŚ1 â2 √đĽ12 + đŚ12 , (36) 2 2 (đĽ1 + đŚ1 ) + đĄ (đ1 + â1 ) − √(đĽ1 + đĄđ1 ) + (đŚ1 + đĄâ1 ) (đĽ2 + đŚ2 ) + đĄ (đ2 + â2 ) ). ( (đĽ1 + đĄđ1 ) (đĽ2 + đĄđ2 ) + (đŚ1 + đĄâ1 ) (đŚ2 + đĄâ2 ) − √(đĽ1 + đĄđ1 )2 + (đŚ1 + đĄâ1 )2 (35) Let [đFB (đĽ, đŚ)]1 be the first element of đFB (đĽ, đŚ) and [đFB (đĽ, đŚ)]2 the vector consisting of the rest đ−1 components of đFB (đĽ, đŚ). By the above expression of đFB (đĽ + đĄđ, đŚ + đĄâ), where the last equality is using đĽ1 đŚ2 = đŚ1 đĽ2 by Lemma 5. The above two limits imply ó¸ đFB ((đĽ, đŚ); (đ, â)) = (đ + â) − đĽ1 √đĽ12 + đŚ12 đ− đŚ1 â. √đĽ12 + đŚ12 (37) (c.2): (đĽ + đĄđ)2 + (đŚ + đĄâ)2 ∈ int Kđ for sufficiently small đĄ > 0. Let đ˘ = (đ˘1 , đ˘2 ) := (đĽ + đĄđ)2 + (đŚ + đĄâ)2 with the spectral values đ1 , đ2 . An elementary calculation gives óľŠ2 óľŠ óľŠ2 óľŠ đ˘1 = óľŠóľŠóľŠđĽ + đĄđóľŠóľŠóľŠ + óľŠóľŠóľŠđŚ + đĄâóľŠóľŠóľŠ = đ¤1 + 2đĄ (đĽđ đ + đŚđ â) óľŠ óľŠ2 + đĄ2 (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) , lim đĄ↓0 [đFB (đĽ + đĄđ, đŚ + đĄâ)]1 − [đFB (đĽ, đŚ)]1 đĄ = (đ1 + â1 ) − lim = (đ1 + â1 ) − lim đĄ↓0 đ˘2 = 2 (đĽ1 + đĄđ1 ) (đĽ2 + đĄđ2 ) + 2 (đŚ1 + đĄâ1 ) (đŚ2 + đĄâ2 ) = đ¤2 + 2đĄ (đĽ1 đ2 + đ1 đĽ2 + đŚ1 â2 + â1 đŚ2 ) √(đĽ1 +đĄđ1 )2 + (đŚ1 +đĄâ1 )2 − √đĽ12 +đŚ12 đĄ đĄ↓0 đĽ1 đ1 + đŚ1 â1 √đĽ12 + đŚ12 , [đFB (đĽ + đĄđ, đŚ + đĄâ)]2 − [đFB (đĽ, đŚ)]2 đĄ = (đ2 + â2 ) (39) + 2đĄ2 (đ1 đ2 + â1 â2 ) . Also, since đ¤2 ≠ 0, applying the Taylor formula of â ⋅ â at đ¤2 and Lemma 6(a) yields đ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ đ¤2 (đ˘2 − đ¤2 ) + đ (đĄ) óľŠóľŠđ˘2 óľŠóľŠ = óľŠóľŠđ¤2 óľŠóľŠ + óľŠóľŠ óľŠóľŠ óľŠóľŠđ¤2 óľŠóľŠ óľŠ óľŠ = óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ + 2đĄ (đĽđ đ + đŚđ â) + đ (đĄ) . (40) Now using the definition of đFB and noting that đ 1 = 0 and đ¤2 ≠ 0, we have that đFB (đĽ + đĄđ, đŚ + đĄâ) − đFB (đĽ, đŚ) [ (đĽ +đĄđ1 ) (đĽ2 +đĄđ2 ) + (đŚ1 +đĄâ1 ) (đŚ2 +đĄâ2 ) − lim [ 1 2 2 đĄ↓0 đĄ√(đĽ1 + đĄđ1 ) + (đŚ1 + đĄâ1 ) [ − (38) đĽ1 đĽ2 + đŚ1 đŚ2 ] ] đĄ√đĽ12 + đŚ12 ] = (đĽ + đĄđ + đŚ + đĄâ) − √đ˘ − (đĽ + đŚ) + √đ¤ √đ1 + √đ2 − √đ 2 2 = đĄ (đ + â) − ( đ − đ √đ 2 đ¤2 ) , √ 2 √ 1 đ˘2 − óľŠóľŠ óľŠóľŠ 2 2 óľŠóľŠóľŠóľŠđ¤2 óľŠóľŠóľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ (41) 6 Abstract and Applied Analysis Using âđ¤2 â = 2(đĽ12 + đŚ12 ) and (40), we simplify the sum of the first two terms in (45) as which in turn implies that ó¸ ((đĽ, đŚ); (đ, â)) đFB óľŠ óľŠ 4đĄ óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ (đĽđ đ + đŚđ â) 2đĄ (đĽ đ + đŚ â) − óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠđ˘ óľŠ − óľŠđ¤ óľŠ = 2đĄ (đĽđ đ + đŚđ â) óľŠóľŠóľŠ 2 óľŠóľŠóľŠ óľŠóľŠóľŠ 2 óľŠóľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ √đ1 + √đ2 − √đ 2 đĄ↓0 2đĄ ). = (đ + â) − ( √đ 2 đ¤2 √đ2 − √đ1 đ˘2 − ) lim ( óľŠóľŠ óľŠóľŠ đĄ↓0 2đĄ 2đĄ óľŠóľŠóľŠóľŠđ¤2 óľŠóľŠóľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ (42) lim đ (43) = 4đĄ (đĽđ đ + đŚđ â) + đ (đĄ) , 4đĄ2 (đĽđ đ + đŚđ â) 2 = óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ + đ (đĄ ) . óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ Then, from (45) and âđ¤2 â = 2(đĽ12 + đŚ12 ), we obtain that and consequently, lim đĄ↓0 lim đ − đ2 1 √đ2 − √đ 2 = lim 2 ⋅ đĄ↓0 đĄ đĄ √đ2 + √đ 2 = đĽđ đ + đŚđ â 2√đ 2 = đĽđ đ + đŚđ â √đĽ12 + đŚ12 đĄ↓0 (44) đ¤2đ (đĽ1 đ2 + đ1 đĽ2 + đŚ1 â2 + â1 đŚ2 ) óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ óľŠ óľŠ2 + đĄ2 (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) óľŠóľŠ − 4đĄ − 4đĄ2 − 4đĄ − 8đĄ2 . (47) óľŠ2 óľŠ óľŠ2 óľŠ (đĽ12 + đŚ12 ) (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) − óľŠóľŠóľŠđ1 đĽ2 + đĽ1 đ2 + đŚ1 â2 + â1 đŚ2 óľŠóľŠóľŠ óľŠ2 óľŠ óľŠ2 óľŠ = (đĽ12 + đŚ12 ) (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) − óľŠóľŠóľŠđ1 đĽ2 + đĽ1 đ2 óľŠóľŠóľŠ đ (đĽ12 + đŚ12 ) (đĽđ đ + đŚđ â) óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ 2 đ (đĽđ đ + đŚđ â) − 2(đĽ1 đĽ2 + đŚ1 đŚ2 ) (đ1 đ2 + â1 â2 ) 2 2 = (đĽ1 đ1 + đĽ2đ đ2 ) + (đŚ1 â1 + đŚ2đ â2 ) + 2đĽđ đđŚđ â óľŠ2 + đĽ1 đ2 + đŚ1 â2 + â1 đŚ2 óľŠóľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ đ − 2(đĽ1 đĽ2 + đŚ1 đŚ2 ) (đ1 đ2 + â1 â2 ) 2 2 2 2 = (đĽ1 đ1 ) + (đĽ2đ đ2 ) + (đŚ1 â1 ) + (đŚ2đ â2 ) đ (đĽ1 đĽ2 + đŚ1 đŚ2 ) (đ1 đ2 + â1 â2 ) . óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ đĽ12 + đŚ12 − 2(đ1 đĽ2 + đĽ1 đ2 ) (đŚ1 â2 + â1 đŚ2 ) , óľŠ óľŠ2 + đĄ2 (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) + đ (đĄ2 ) óľŠóľŠ đ óľŠ óľŠ2 = đĽ12 âââ2 + đŚ12 óľŠóľŠóľŠđóľŠóľŠóľŠ − 2đĽ1 đ1 đĽ2đ đ2 − 2đŚ1 â1 đŚ2đ â2 đ¤2đ (đ1 đ2 + â1 â2 ) 2 óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ + đ (đĄ ) óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ 2óľŠ óľŠđ1 đĽ2 + (đĽđ đ + đŚđ â) − 2(đĽ1 đĽ2 + đŚ1 đŚ2 ) (đ1 đ2 + â1 â2 ) đ óľŠ2 óľŠ − óľŠóľŠóľŠđŚ1 â2 + â1 đŚ2 óľŠóľŠóľŠ − 2(đ1 đĽ2 + đĽ1 đ2 ) (đŚ1 â2 + â1 đŚ2 ) óľŠ2 + đĽ1 đ2 + đŚ1 â2 + â1 đŚ2 óľŠóľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ = 2đĄ (đĽđ đ + đŚđ â) − 8đĄ đĽ12 + đŚ12 We next make simplification for the numerator of the right hand side of (47). Note that óľŠóľŠ óľŠóľŠ2 óľŠóľŠ óľŠóľŠ2 óľŠđ˘ óľŠ − óľŠđ¤ óľŠ = (đ˘1 − đ¤1 ) − óľŠóľŠóľŠ 2 óľŠóľŠóľŠ óľŠóľŠóľŠ 2 óľŠóľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđ¤2 óľŠóľŠ 2óľŠ óľŠđ1 đĽ2 óľŠ2 óľŠ óľŠ2 óľŠ (đĽ12 + đŚ12 ) (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) − óľŠóľŠóľŠđ1 đĽ2 + đĽ1 đ2 + đŚ1 â2 + â1 đŚ2 óľŠóľŠóľŠ 2 óľŠ óľŠ óľŠ óľŠ đ1 = (đ˘1 − đ¤1 ) − (óľŠóľŠóľŠđ˘2 óľŠóľŠóľŠ − óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ) = 2đĄ (đĽđ đ + đŚđ â) − 4đĄ = đ1 đĄ2 . We next calculate limđĄ↓0 (√đ1 /đĄ). Since đ¤1 − âđ¤2 â = 0, using (38)-(39) and Lemma 6(a), (46) 2 We first calculate limđĄ↓0 ((√đ2 − √đ 2 )/đĄ). Using (38) and (40), it is easy to see that óľŠ óľŠ óľŠ óľŠ đ2 − đ 2 = (đ˘1 − đ¤1 ) + (óľŠóľŠóľŠđ˘2 óľŠóľŠóľŠ − óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ) đ (45) + 2đĽđ đđŚđ â − 2đĽ1 â1 đĽ2đ â2 − 2đ1 đŚ1 đ2đ đŚ2 . (48) Abstract and Applied Analysis 7 Therefore, adding the last two equalities and using Lemma 5 yield that Together with (44) and (50), we have that lim [ đĄ↓0 óľŠ2 óľŠ óľŠ2 óľŠ (đĽ12 + đŚ12 ) (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) − óľŠóľŠóľŠđ1 đĽ2 + đĽ1 đ2 + đŚ1 â2 + â1 đŚ2 óľŠóľŠóľŠ 2 √đ 2 đ¤2 √đ2 − √đ1 đ˘2 − ] óľŠ óľŠ óľŠ óľŠ 2đĄ 2đĄ óľŠóľŠóľŠóľŠđ¤2 óľŠóľŠóľŠóľŠ óľŠóľŠđ˘2 óľŠóľŠ = −lim √đ 2 đ¤2 √đ1 đ˘2 √đ2 đ˘2 + lim [ − ] óľŠ óľŠ óľŠ óľŠ óľŠ óľŠ óľŠ óľŠ đĄ↓0 2đĄ óľŠóľŠđ˘2 óľŠóľŠ 2đĄ óľŠóľŠđ˘2 óľŠóľŠ 2đĄ óľŠóľŠóľŠóľŠđ¤2 óľŠóľŠóľŠóľŠ = −lim √đ1 đ˘2 √đ2 − √đ 2 đ˘2 + lim óľŠóľŠ óľŠóľŠ 2đĄ óľŠóľŠóľŠóľŠđ˘2 óľŠóľŠóľŠóľŠ đĄ↓0 2đĄ óľŠóľŠđ˘2 óľŠóľŠ đĄ↓0 đ + (đĽđ đ + đŚđ â) − 2(đĽ1 đĽ2 + đŚ1 đŚ2 ) (đ1 đ2 + â1 â2 ) óľŠ óľŠ2 = (đĽ12 âââ2 − 2đĽ1 â1 đĽ2đ â2 ) + (đŚ12 óľŠóľŠóľŠđóľŠóľŠóľŠ − 2đ1 đŚ1 đ2đ đŚ2 ) 2 2 2 2 đĄ↓0 √đ 2 (óľŠóľŠóľŠóľŠđ¤2 óľŠóľŠóľŠóľŠ đ˘2 − óľŠóľŠóľŠóľŠđ˘2 óľŠóľŠóľŠóľŠ đ¤2 ) + lim óľŠ óľŠóľŠ óľŠ đĄ↓0 2đĄ óľŠóľŠóľŠđ˘2 óľŠóľŠóľŠ óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ + ((đĽ1 đ1 ) + (đĽ2đ đ2 ) − 2đĽ1 đ1 đĽ2đ đ2 ) + ((đŚ1 â1 ) + (đŚ2đ â2 ) − 2đŚ1 â1 đŚ2đ â2 ) + 2đĽđ đđŚđ â =− đ − 2(đ1 đĽ2 + đĽ1 đ2 ) (đŚ1 â2 + â1 đŚ2 ) 1 óľŠ óľŠ2 óľŠ óľŠ2 = óľŠóľŠóľŠđĽ1 â2 − â1 đĽ2 óľŠóľŠóľŠ + óľŠóľŠóľŠđ1 đŚ2 − đŚ1 đ2 óľŠóľŠóľŠ + (đĽ1 đ1 − đĽ2đ đ2 ) 2 − 2 + (đŚ1 â1 − đŚ2đ â2 ) + 2 (đ1 đĽ1 + đ2đ đĽ2 ) (đŚ1 â1 + đŚ2đ â2 ) đ =− − 2(đ1 đĽ2 + đĽ1 đ2 ) (đŚ1 â2 + â1 đŚ2 ) 2 óľŠ óľŠ2 óľŠ óľŠ2 = óľŠóľŠóľŠđĽ1 â2 − â1 đĽ2 óľŠóľŠóľŠ + óľŠóľŠóľŠđ1 đŚ2 − đŚ1 đ2 óľŠóľŠóľŠ + (đĽ1 đ1 − đĽ2đ đ2 ) 2 đ (đ, â) đĽ đ + đ1 đĽ2 + đŚ1 â2 + â1 đŚ2 Ě2 + 1 2 đ¤ 2 √ đĽ 2 + đŚ2 đ + (đŚ1 â1 − đŚ2đ â2 ) + 2(đĽ1 â2 − â1 đĽ2 ) (đ1 đŚ2 − đ2 đŚ1 ) đĽ đ đ + đŚđ â 2√đĽ12 + đŚ12 1 Ě2 đ¤ đ (đ, â) 2đĽ đ + đ1 đĽ2 + 2đŚ1 â2 + â1 đŚ2 Ě2 + 1 2 đ¤ 2 2√đĽ2 + đŚ2 1 − đĽ2đ đ2 + đŚ2đ â2 2√đĽ12 + đŚ12 1 Ě2 , đ¤ (52) + 2 (đĽ1 đ1 − đĽ2đ đ2 ) (đŚ1 â1 − đŚ2đ â2 ) óľŠ2 óľŠ = óľŠóľŠóľŠđĽ1 â2 − â1 đĽ2 + đ1 đŚ2 − đŚ1 đ2 óľŠóľŠóľŠ 2 + (đĽ1 đ1 − đĽ2đ đ2 + đŚ1 â1 − đŚ2đ â2 ) . (49) Ě2 = đĽ2 and đŚ1 đ¤ Ě2 = đŚ2 . where the last equality is using đĽ1 đ¤ Combining with (42), (44), and (50), a suitable rearrangeó¸ ((đĽ, đŚ); (đ, â)) has the expression (32). ment shows that đFB Finally, we show that, when (đĽ + đĄđ)2 + (đŚ + đĄâ)2 ∈ bd+ Kđ for sufficiently small đĄ > 0, the formula in (32) reduces to the one in (37). Indeed, an elementary calculation yields 2 Combining this equality with (47) and using the definition of đ in (33), we readily get lim đĄ↓0 √đ1 = đ (đ, â) . đĄ 2 đ 1 ((đĽ + đĄđ) + (đŚ + đĄâ) ) óľŠ2 óľŠ óľŠ2 2 óľŠ = [óľŠóľŠóľŠđĽ + đĄđóľŠóľŠóľŠ + óľŠóľŠóľŠđŚ + đĄâóľŠóľŠóľŠ ] óľŠ óľŠ2 − 4óľŠóľŠóľŠ(đĽ1 + đĄđ1 ) (đĽ2 + đĄđ2 ) + (đŚ1 + đĄâ1 ) (đŚ2 + đĄâ2 )óľŠóľŠóľŠ (50) = 4đĄ2 đ (đ, â) √đĽ12 + đŚ12 óľŠ óľŠ2 + 4đĄ3 (đĽđ đ + đŚđ â) (óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) We next calculate limđĄ↓0 [((√đ2 − √đ1 )/2đĄ)(đ˘2 /âđ˘2 â) − ((√đ 2 /2đĄ)(đ¤2 /âđ¤2 â))]. To this end, we also need to take a look at âđ¤2 âđ˘2 − âđ˘2 âđ¤2 . From (38)-(39) and (40), it follows that óľŠ óľŠ óľŠóľŠ óľŠóľŠ óľŠóľŠđ¤2 óľŠóľŠ đ˘2 − óľŠóľŠóľŠđ˘2 óľŠóľŠóľŠ đ¤2 óľŠ óľŠ = 2đĄ óľŠóľŠóľŠđ¤2 óľŠóľŠóľŠ [ (đĽ1 đ2 + đ1 đĽ2 + đŚ1 â2 + â1 đŚ2 ) Ě2 ] + đ (đĄ) . − (đĽđ đ + đŚđ â) đ¤ đ − 8đĄ2 (đĽ1 đ2 + đ1 đĽ2 + đŚ1 â2 + â1 đŚ2 ) (đ1 đ2 + â1 â2 ) 2 óľŠ2 óľŠ óľŠ2 óľŠ + đĄ4 [(óľŠóľŠóľŠđóľŠóľŠóľŠ + âââ2 ) − 2óľŠóľŠóľŠđ1 đ2 + â1 â2 óľŠóľŠóľŠ ] = 4đĄ2 đ (đ, â) √đĽ12 + đŚ12 + đ (đĄ2 ) . (53) (51) This implies that if (đĽ+đĄđ)2 +(đŚ+đĄâ)2 ∈ bd+ Kđ for sufficiently small đĄ > 0, that is, đ 1 ((đĽ + đĄđ)2 + (đŚ + đĄâ)2 ) = 0 for sufficiently small đĄ > 0, then đ(đ, â) = 0, and hence đĽ1 đ1 + đŚ1 â1 − (đĽ2đ đ2 + 8 Abstract and Applied Analysis ó¸ đŚ2đ â2 ) = 0. Thus, đFB ((đĽ, đŚ); (đ, â)) in (32) can be simplified as (đ + â) + đĽ1 đ1 + đŚ1 â1 2√đĽ12 + đŚ12 0 ( )− Ě2 đ¤ 1 2√đĽ12 + đŚ12 2đĽ1 đ1 + 2đŚ1 â1 ×( ) 2đĽ1 đ2 + đ1 đĽ2 + 2đŚ1 â2 + â1 đŚ2 = (đ + â) − đĽ1 √đĽ12 + đŚ12 đ− đŚ1 √đĽ12 + đŚ12 (54) â, As a consequence of Proposition 9, we have the following sufficient and necessary characterizations for the (continuously) differentiable points of đFB and đ(đĽ) := |đĽ|. Corollary 10. (a) The function đFB is (continuously) differentiable at (đĽ, đŚ) ∈ Rđ ×Rđ if and only if đ§(đĽ, đŚ) ∈ int Kđ , where đ§(đĽ, đŚ) is defined in (31). Also, when đ§(đĽ, đŚ) ∈ int Kđ , one has JđFB (đĽ, đŚ) = [đź − đź− đż−1 đ§(đĽ,đŚ) đż đŚ ] . (55) (b) The function đ(đĽ) := |đĽ| is (continuously) differentiable at đĽ ∈ Rđ if and only if đĽ is invertible. Also, when đĽ is invertible, Jđ(đĽ) = đż−1 |đĽ| đż đĽ . Proof. (a) The “if ” part is direct by [13, Proposition 5.2]. We next prove the “only if ” part. If đFB is differentiable at (đĽ, đŚ), ó¸ ((đĽ, đŚ); (đ, â)) is a linear function of (đ, â), which by then đFB Proposition 9 holds only if đ§(đĽ, đŚ) ∈ int Kđ . The formula of JđFB (đĽ, đŚ) is given in [13]. (b) Since đ(đĽ) = √đĽ2 , by part (a) đ is (continuously) differentiable at đĽ ∈ Rđ if and only if |đĽ| ∈ int Kđ , which is equivalent to requiring that đĽ is invertible since |đĽ| ∈ Kđ always holds. When đĽ is invertible, the formula of Jđ(đĽ) follows from part (a). For any given đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) ∈ R × Rđ−1 with 2 đĽ + đŚ2 ∈ bd+ Kđ , define đ1 (đ, â) := đĽ1 đ1 − đĽ2đđ2 + đŚ1 â1 − đŚ2đ â2 , đ2 (đ, â) := đĽ1 â2 − â1 đĽ2 + đ1 đŚ2 − đŚ1 đ2 (56) for any đ = (đ1 , đ2 ), â = (â1 , â2 ) ∈ R×Rđ−1 , and let đ(đ, â) := (đ1 (đ, â), đ2 (đ, â)). Then, comparing with (33), we can rewrite the function đ as óľŠ óľŠóľŠ óľŠđ (đ, â)óľŠóľŠóľŠ đ (đ, â) = óľŠ , √đĽ12 + đŚ12 ∀đ, â ∈ Rđ . Lemma 11. For any given (đĽ, đŚ) with đĽ2 + đŚ2 ∈ bd+ Kđ , let đ be defined by (33). Then, the B-subdifferential of the function đ at (0, 0) takes the following form: { {[ đđ đŚ + đ1 đĽ1 −đŚ1 đ2đ − đ1 đĽ2đ , ), đđľ đ (0, 0) = {[( 2 2 { √đĽ12 + đŚ12 √đĽ12 + đŚ12 {[ Ě2 , â1 đŚ2 = â1 đŚ1 đ¤ Ě2 . where the equality is using đ1 đĽ2 = đ1 đĽ1 đ¤ The proof is complete. đż−1 đ§(đĽ,đŚ) đż đĽ differentiable at (đ, â) if and only if đ(đ, â) > 0. The following lemma characterizes the đľ-subdifferential of the function đ at the origin. (57) Note that the Euclidean norm â ⋅ â is globally Lipschitz continuous and strongly semismooth everywhere in Rđ , and đ(⋅, ⋅) is a linear function. Then, (57) implies that đ is globally Lipschitz continuous and strongly semismooth everywhere in Rđ × Rđ by [17, Theorem 19]. Also, the function đ is ( −đ2đ đĽ2 + đ1 đŚ1 đĽ1 đ2đ − đ1 đŚ2đ ] , )] : √đĽ12 + đŚ12 √đĽ12 + đŚ12 ] } óľŠ óľŠ2 } (đ1 , đ2 ) ∈ R × Rđ−1 đ đđĄđđ đđđđ đ12 + óľŠóľŠóľŠđ2 óľŠóľŠóľŠ = 1} . } } (58) Proof. Let (đ˘đ , đŁđ ) ∈ đđľ đ(0, 0). By the definition of the elements in đđľ đ(0, 0), there exists a sequence {(đđ , âđ )} in Rđ × Rđ converging to (0, 0) with đ(đđ , âđ ) > 0 such that (đ˘đ , đŁđ ) = lim Jđ (đđ , âđ ) . đ→∞ (59) By (57), a simple computation shows that such (đ˘đ, đŁđ ) belongs to the set on the right hand side of (58). Thus, đđľ đ(0, 0) is included in the set on the right hand side of (58). In fact, đ1 and đ2 in (58) are the limit points of {đ1 (đđ , âđ )/âđ(đđ , âđ )â} and {đ2 (đđ , âđ )/âđ(đđ , âđ )â}, respectively. Conversely, let (đ˘đ , đŁđ ) be an arbitrary element of the set on the right hand side of (58). Then, there exists a (đ1 , đ2 ) ∈ R × Rđ−1 with đ12 + âđ2 â2 = 1 such that đ˘=( đ2đ đŚ2 + đ1 đĽ1 −đŚ1 đ2 − đ1 đĽ2 , ), √đĽ12 + đŚ12 √đĽ12 + đŚ12 (60) đŁ=( −đ2đ đĽ2 + đ1 đŚ1 đĽ1 đ2 − đ1 đŚ2 , ). √đĽ12 + đŚ12 √đĽ12 + đŚ12 Take the sequence {(đđ , âđ )} ⊂ Rđ × Rđ with đđ = (đĽ1 đ1 / đ, −đŚ1 đ2 /đ) and âđ = (đŚ1 đ1 /đ, đĽ1 đ2 /đ). Clearly, (đđ , âđ ) → (0, 0) as đ → ∞. Also, by Lemma 5, it is easy to verify that đ1 (đđ , âđ ) = 1 2 (đĽ + đŚ12 ) đ1 , đ 1 đ2 (đđ , âđ ) = 1 2 (đĽ + đŚ12 ) đ2 . đ 1 (61) This shows that đ(đđ , âđ ) > 0 and Jđ(đđ , âđ ) = (đ˘đ , đŁđ ). Hence, (đ˘đ , đŁđ ) ∈ đđľ đ(0, 0). Thus, the set on the right hand side of (58) is included in đđľ đ(0, 0). Abstract and Applied Analysis 9 Now we may prove the equivalence between the đľsubdifferential of đFB at a general point (đĽ, đŚ) and that of its ó¸ ((đĽ, đŚ); (⋅, ⋅)) at (0, 0). This directional derivative function đFB result corresponds to that of [18, Lemma 14] established for the NR SOC function đNR . = − (62) − Proof. The result is direct by Proposition 9(a)-(b) and Lemma A.1 in the Appendix. Using Lemma 12, we may present an upper estimation for Clarke’s Jacobian of đFB at the point (đĽ, đŚ) with đĽ2 + đŚ2 ∈ bd+ Kđ , which will be used in the next section. Proposition 13. For any given đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) with đĽ2 + đŚ2 ∈ bd+ Kđ , one has đđFB (đĽ, đŚ) ⊆ {[đź − đ − 1 1 ) đ˘đ ( đ¤2 2 −Ě 2√đĽ12 + đŚ12 + Lemma 12. For any given (đĽ, đŚ) ∈ Rđ × Rđ , let đFB (⋅, ⋅) ≡ ó¸ ((đĽ, đŚ); (⋅, ⋅)). Then, đFB đđľ đFB (đĽ, đŚ) = đđľ đFB (0, 0) . 1 0 ( ) (0 đĽ2đ ) đ Ě2 đ¤ 1 2√đĽ12 + đŚ12 1 2√đĽ12 + đŚ12 1 2√đĽ12 + đŚ12 0 ( ) (0 đŚ2đ ) â Ě2 đ¤ đĽ đĽ2đ ( 1 )đ đĽ2 2đĽ1 đź đŚ đŚ2đ ( 1 )â đŚ2 2đŚ1 đź = −đđ − đâ, (65) where the last equality is using đ¤1 = 2(đĽ12 + đŚ12 ). Therefore, from (32), we have ó¸ ((đĽ, đŚ); (đ, â)) = (đ + â) − đFB óľ¨óľ¨ 1 1 óľ¨ đź−đ− ( ) đŁđ ] óľ¨óľ¨óľ¨ đ¤2 óľ¨óľ¨ 2 −Ě đ (đ, â) 1 ) − đđ − đâ, ( −Ě đ¤2 2 ∀đ, â ∈ Rđ . (66) đđđ đ˘ = (đ˘1 , đ˘2 ) , đŁ = (đŁ1 , đŁ2 ) đ đđĄđđ đđŚđđđ đĄâđ đđđđđ˘đđđđĄđđđ óľ¨ óľ¨ óľŠ óľŠ óľ¨óľ¨ óľ¨óľ¨ óľŠóľŠ óľŠóľŠ óľ¨óľ¨đ˘1 óľ¨óľ¨ ≤ óľŠóľŠđ˘2 óľŠóľŠ ≤ 1, óľ¨óľ¨óľ¨đŁ1 óľ¨óľ¨óľ¨ ≤ óľŠóľŠóľŠđŁ2 óľŠóľŠóľŠ ≤ 1, óľŠ óľŠ (63) óľŠ óľŠ (đ˘1 − đŁ1 ) ≤ óľŠóľŠóľŠđ˘2 − đŁ2 óľŠóľŠóľŠ , (đ˘1 + đŁ1 ) ≤ óľŠóľŠóľŠđ˘2 + đŁ2 óľŠóľŠóľŠ , 2 óľŠ2 óľŠ (đ˘1 − đŁ1 ) + óľŠóľŠóľŠđ˘2 + đŁ2 óľŠóľŠóľŠ ≤ 2, 2 óľŠ2 óľŠ (đ˘1 + đŁ1 ) + óľŠóľŠóľŠđ˘2 − đŁ2 óľŠóľŠóľŠ ≤ 2, Ě2đ ) đ˘ = 0, (1, đ¤ Ě2đ ) đŁ = 0, (1, đ¤ 1 đŚ2đ đŚ đ= ( 1 ). Ě2 đŚ2đ √2đ¤1 đŚ2 2đŚ1 đź − đ¤ 2√đĽ12 + đŚ12 0 ( )− Ě2 đ¤ 1 2√đĽ12 + đŚ12 đĽ đ đ + đŚđ â ×( ) 2đĽ1 đ2 + đ1 đĽ2 + 2đŚ1 â2 + â1 đŚ2 óľ¨óľ¨ 1 1 1 1 óľ¨ ) đ˘đ đź − đ − ( ) đŁđ ] óľ¨óľ¨óľ¨ ( đ¤2 đ¤2 óľ¨óľ¨ 2 −Ě 2 −Ě for đ˘ = (đ˘1 , đ˘2 ) , đŁ = (đŁ1 , đŁ2 ) (67) { {[ đđ đŚ + đ1 đĽ1 −đŚ1 đ2đ − đ1 đĽ2đ , ), đđ (0, 0) = {[( 2 2 { √đĽ12 + đŚ12 √đĽ12 + đŚ12 {[ (64) Proof. We first make simplifications for the last two terms in (32) by đ, đ. Note that đĽ2đ đ2 + đŚ2đ â2 = {[đź − đ − where, by Lemma 11 and the definition of Clarke’s Jacobian, where đ and đ are đ × đ real symmetric matrices defined as follows: 1 đĽ2đ đĽ ( 1 ), Ě2 đĽ2đ √2đ¤1 đĽ2 2đĽ1 đź − đ¤ đđFB (đĽ, đŚ) with (đ˘đ, đŁđ ) ∈ đđ (0, 0) } , (1, −Ě đ¤2đ ) đ˘ = 2đ˘1 , (1, −Ě đ¤2đ ) đŁ = 2đŁ1 } , đ= Now, applying Lemma 12, we immediately obtain that ( −đ2đ đĽ2 + đ1 đŚ1 đĽ1 đ2đ − đ1 đŚ2đ ] , )] : √đĽ12 + đŚ12 √đĽ12 + đŚ12 ] } } óľŠ óľŠ2 (đ1 , đ2 ) ∈ R × Rđ−1 satisfies đ12 + óľŠóľŠóľŠđ2 óľŠóľŠóľŠ ≤ 1} . } } (68) Let (đ˘đ , đŁđ ) ∈ đđ(0, 0) with đ˘ = (đ˘1 , đ˘2 ), đŁ = (đŁ1 , đŁ2 ) ∈ R × Rđ−1 . Then, it suffices to prove that such đ˘ and đŁ satisfy 10 Abstract and Applied Analysis all inequalities and equalities in (63). By (68), there exists a vector (đ1 , đ2 ) ∈ R × Rđ−1 with đ12 + âđ2 â2 ≤ 1 such that đ˘1 = đŁ1 = đ2đ đŚ2 + đ1 đĽ1 √đĽ12 + đŚ12 , −đ2đ đĽ2 + đ1 đŚ1 √đĽ12 + đŚ12 đ˘2 = , −đŚ1 đ2 − đ1 đĽ2 đŁ2 = , √đĽ12 + đŚ12 đĽ1 đ2 − đ1 đŚ2 √đĽ12 + đŚ12 (69) . đđľ đNR (đĽ, đŚ) ⊆ đđľ đFB (đĽ, đŚ) . óľŠ2 óľŠ (70) = óľŠóľŠóľŠ−đŚ1 đ2 − đ1 đĽ2 óľŠóľŠóľŠ óľŠ óľŠ2 óľŠ óľŠ2 óľŠ óľŠ2 ≤ óľŠóľŠóľŠđ2 óľŠóľŠóľŠ đŚ12 + đ12 đĽ12 + óľŠóľŠóľŠđ2 óľŠóľŠóľŠ đĽ12 + đ12 óľŠóľŠóľŠđŚ2 óľŠóľŠóľŠ óľŠ óľŠ2 = (óľŠóľŠóľŠđ2 óľŠóľŠóľŠ + đ12 ) (đĽ12 + đŚ12 ) ≤ đĽ12 + đŚ12 . This means that |đ˘1 | ≤ âđ˘2 â ≤ 1. Similarly, we also have |đŁ1 | ≤ âđŁ2 â ≤ 1. We next prove that (đ˘1 − đŁ1 )2 − âđ˘2 − đŁ2 â2 ≤ 0. By Lemma 5, it is easy to verify that (đĽ2 − đŚ2 ) (đĽ1 + đŚ1 ) = (đĽ2 + đŚ2 ) (đĽ1 − đŚ1 ) , 2 óľŠ óľŠ2 (đĽ1 − đŚ1 ) = óľŠóľŠóľŠđĽ2 − đŚ2 óľŠóľŠóľŠ . Using the two equalities, it is not hard to calculate that 2 óľŠ2 óľŠ (đ˘1 − đŁ1 ) − óľŠóľŠóľŠđ˘2 − đŁ2 óľŠóľŠóľŠ = (71) đ (đĽ, đŚ) := 2 đĽ12 + đŚ12 ≤ = 2 óľŠ2 óľŠ óľŠ2 óľŠ + óľŠóľŠóľŠđ2 óľŠóľŠóľŠ ) [(đŚ1 − đĽ1 ) + óľŠóľŠóľŠđĽ2 + đŚ2 óľŠóľŠóľŠ ] đĽ12 + đŚ12 2 (đ12 óľŠ óľŠ2 + óľŠóľŠóľŠđ2 óľŠóľŠóľŠ ) (đĽ12 + đŚ12 ) ≤ 2. đĽ12 + đŚ12 (75) óľ¨ óľ¨óľ¨ óľ¨đĽ − đŚóľ¨óľ¨óľ¨ − (đĽ − đŚ) đŚ= óľ¨ . 2 (76) 1 óľ¨ óľ¨ [(đĽ + đŚ) + óľ¨óľ¨óľ¨đĽ − đŚóľ¨óľ¨óľ¨] 2 ∀đĽ, đŚ ∈ Rđ . (77) Comparing with the definition of đFB , we only need to prove the following inclusion: (đ2đ đŚ2 + đ1 đĽ1 + đ2đ đĽ2 − đ1 đŚ1 ) (đ12 đŚ = đ1 đ˘(1) + đ2 đ˘(2) . Indeed, if đĽ = 0 or đŚ = 0, then the statement clearly holds. If đĽ ≠ 0 and đŚ ≠ 0, then the condition that đĽ ∈ Kđ , đŚ ∈ Kđ , and â¨đĽ, đŚâŠ = 0 implies đĽ, đŚ ∈ bd Kđ . From Lemma 7, there exists đź > 0 such that đĽ2 = −đźđŚ2 and đĽ1 = đźđŚ1 . Together with the spectral factorizations of đĽ and đŚ, the conclusion in (75) follows. Since â¨đ˘(1) , đ˘(2) ⊠= 0 and âđ˘(1) â = âđ˘(2) â = 1/√2, using (75) and â¨đĽ, đŚâŠ = 0 yields that đ 1 đ1 = đ 2 đ2 = 0. This, along with the nonnegativity of đ 1 , đ 2 and đ1 , đ2 , implies (72) Similarly, we also have (đ˘1 + đŁ1 )2 ≤ âđ˘2 + đŁ2 â2 . In addition, we have that 2 óľŠ2 óľŠ (đ˘1 − đŁ1 ) + óľŠóľŠóľŠđ˘2 + đŁ2 óľŠóľŠóľŠ óľŠ2 óľŠóľŠ óľŠ−đŚ đ − đ1 đĽ2 + đĽ1 đ2 − đ1 đŚ2 óľŠóľŠóľŠ +óľŠ 1 2 đĽ12 + đŚ12 đĽ = đ 1 đ˘(1) + đ 2 đ˘(2) , By the definition of đNR , we have đNR (đĽ, đŚ) = (đĽ+đŚ)−đ(đĽ, đŚ) for đĽ, đŚ ∈ Rđ , where đĽ12 + đŚ12 2 2 óľŠ óľŠ2 (đ2đ đŚ2 + đ2đ đĽ2 ) − óľŠóľŠóľŠđ2 óľŠóľŠóľŠ (đŚ1 + đĽ1 ) = ≤ 0. đĽ12 + đŚ12 = (74) Proof. Since (đĽ, đŚ) satisfies đĽ ∈ Kđ , đŚ ∈ Kđ and â¨đĽ, đŚâŠ = 0, there exist spectral vectors đ˘(1) , đ˘(2) ∈ Rđ and nonnegative real numbers đ 1 , đ 2 and đ1 , đ2 such that óľ¨ óľ¨óľ¨ óľ¨đĽ − đŚóľ¨óľ¨óľ¨ + (đĽ − đŚ) đĽ= óľ¨ , 2 2 óľŠ2 óľŠóľŠ óľŠóľŠ−đŚ1 đ2 − đ1 đĽ2 − đĽ1 đ2 + đ1 đŚ2 óľŠóľŠóľŠ − đĽ12 + đŚ12 To close this section, we establish a relation between the đľ-subdifferential of đFB at a complementarity point pair and that of đNR at the corresponding point pair. Lemma 14. Let (đĽ, đŚ) ∈ Rđ × Rđ satisfy đĽ ∈ Kđ , đŚ ∈ Kđ and â¨đĽ, đŚâŠ = 0. Then, Using Lemma 5, it is immediate to obtain that 2 óľ¨2 óľ¨óľ¨ đ óľ¨óľ¨đ2 đŚ2 + đ1 đĽ1 óľ¨óľ¨óľ¨ = (đ2đ đŚ2 ) + (đ1 đĽ1 )2 + 2đ1 đĽ1 đ2đ đŚ2 óľ¨ óľ¨ óľŠ óľŠ2 ≤ óľŠóľŠóľŠđ2 óľŠóľŠóľŠ đŚ12 + đ12 đĽ12 + 2đ1 đŚ1 đ2đ đĽ2 (đ2đ đŚ2 + đ1 đĽ1 + đ2đ đĽ2 − đ1 đŚ1 ) A similar argument also yields (đ˘1 + đŁ1 )2 + âđ˘2 − đŁ2 â2 ≤ 2. The last four equalities in (63) are direct by Lemma 5 and the expression of đ˘1 , đ˘2 , đŁ1 , and đŁ2 . đđľ đ (đĽ, đŚ) ⊆ đđľ đ§ (đĽ, đŚ) . (78) For this purpose, let [đ đ] ∈ đđľ đ(đĽ, đŚ). By the definition of the elements in đđľ đ(đĽ, đŚ) and Corollary 10(b), there exists a sequence {(đĽđ , đŚđ )} ⊂ Rđ × Rđ converging to (đĽ, đŚ) with đđ ≡ đĽđ − đŚđ invertible such that [đ đ] = lim Jđ (đĽđ , đŚđ ) (73) đ→∞ 1 1 = lim [ (đź + đż−1đđ đż đđ ) (đź − đż−1đđ đż đđ )] . | | | | đ→∞ 2 2 (79) For each đ, let đĽĚđ = (|đđ | + đđ )/2 and đŚĚđ = (|đđ | − đđ )/ 2. Then, using đđ = đĽđ − đŚđ and (76), we have that đĽĚđ → đĽ and đŚĚđ → đŚ as đ → ∞. Also, đ§(đĽĚđ , đŚĚđ ) = √(đđ )2 = Abstract and Applied Analysis 11 |đđ |âťKđ 0. By Corollary 10(a), the function đ§(⋅, ⋅) is continuously differentiable at (đĽĚđ , đŚĚđ ) with −1 −1 Jđ§ (đĽĚđ , đŚĚđ ) = [đż đ§(đĽĚđ ,đŚĚđ ) đż đĽĚđ đż đ§(đĽĚđ ,đŚĚđ ) đż đŚĚđ ] such that đđ = đđ đˇđ đđđ , đđ = đđ Λ đ đđđ , where đˇđ and Λ đ take one of the forms đˇđ = 0, đż đ đ đż−1 đż đ đ = [đż−1 |đđ | (|đ |+đ )/2 |đđ | (|đ |−đ )/2 ] . (80) Λ đ = đź, đˇđ = diag (0, 0, . . . , 0, 1) , Λđ = ( 4. Nonsingularity Conditions This section studies the nonsingularity of Clarke’s Jacobian of đ¸FB at a KKT point. Let (đ, đĽ, đ, đ , đŚ) ∈ Rđ × K × Rđ × Rđ × K be a KKT point of the SOCP (4), that is, đ (đ) − đĽ = 0, â (đ) = 0, (81) −đŚ ∈ NK (đĽ) . Taking into account that −đŚ ∈ NK (đĽ) if and only if đĽ and đŚ satisfy đĽđ ∈ Kđđ , đŚđ ∈ Kđđ , â¨đĽđ , đŚđ ⊠= 0, đ = 1, 2, . . . , đ, (82) we introduce the following index sets associated with đĽ and đŚ: đ˝đľ0 := {đ | đĽđ ∈ bd K , đŚđ = 0} , đđ = 1 1 ( ), √2 đŚĚđ2 From [19], we learn that the above six index sets form a partition of {1, 2, . . . , đ}. First of all, let us take a careful look at the properties of the elements in đđFB (đĽđ , đŚđ ) for đ ∈ đ˝đľ0 ∪ đ˝0đľ , as stated in the following. The proof of Lemma 15 is given in the Appendix. Lemma 15. For (đĽ, đŚ) satisfying (82), let [đđ đđ ] ∈ đđFB (đĽđ , đŚđ ) for each đ. Then, (a) when đ ∈ đ˝đľ0 , there exists an đđ × đđ orthogonal matrix Ěđ đó¸ ] with đđ = [đđ đ đ đđó¸ = 1 1 ( ), √2 −đĽĚđ2 đĽđ2 đĽĚđ2 = óľŠóľŠ óľŠóľŠ óľŠóľŠđĽđ2 óľŠóľŠ óľŠ óľŠ (84) (85) 1 1 ( ), √2 −đŚĚđ2 đŚđ2 đŚĚđ2 = óľŠóľŠ óľŠóľŠ óľŠóľŠđŚđ2 óľŠóľŠ óľŠ óľŠ (86) such that đđ = đđ đˇđ đđđ , đđ = đđ Λ đ đđđ , where đˇđ and Λ đ take one of the forms đˇđ = đź, Λ đ = 0, 0 − (83) đ˝00 := {đ | đĽđ = 0, đŚđ = 0} . 1 1 ( ), √2 đĽĚđ2 ) 1 đĚ đŁđ đđ 1 − đŁđ1 √2 đđó¸ = 1 0 đ˝0đľ := {đ | đĽđ = 0, đŚđ ∈ bd+ Kđđ } , đđ = 0 − 0 0 (b) when đ ∈ đ˝0đľ , there exists an đđ × đđ orthogonal matrix Ěđ đó¸ ] with đđ = [đđ đ đ đˇđ = ( đđ 0 đź with đ˘đ = (đ˘đ1 , đ˘đ2 ), đŁđ = (đŁđ1 , đŁđ2 ) ∈ R × Rđđ −1 satisfying đ˘đ1 , đŁđ1 < 1, đ˝0đź := {đ | đĽđ = 0, đŚđ ∈ int Kđđ } , + 1 0 đˇđ = diag (1, 1, . . . , 1, 0) , đ˝đź0 := { đ | đĽđ ∈ int Kđđ , đŚđ = 0} , đ˝đľđľ := {đ | đĽđ , đŚđ ∈ bd+ Kđđ } , Λ đ = diag (1, 1, . . . , 1, 0) ; đˇđ = diag (0, 0, . . . , 0, 1 − đ˘đ1 ) , Together with (79), we have that [đ đ] = limđ → ∞ Jđ§(đĽĚđ , đŚĚđ ). This shows that [đ đ] ∈ đđľ đ§(đĽ, đŚ), and the inclusion in (78) follows. The proof is complete. Jđ,đĽ đż (đ, đĽ, đ, đ , đŚ) = 0, đđ = đź; đđ = đź; Λ đ = diag (0, 0, . . . , 0, 1) ; (87) 0 đź 0 0 ), 1 đĚ đ˘đ đđ 1 − đ˘đ1 √2 (88) Λ đ = diag (0, 0, . . . , 0, 1 − đŁđ1 ) with đ˘đ = (đ˘đ1 , đ˘đ2 ), đŁđ = (đŁđ1 , đŁđ2 ) ∈ R × Rđđ −1 satisfying đ˘đ1 , đŁđ1 < 1. The following proposition plays a key role in achieving the main result of this section. Proposition 16. For (đĽ, đŚ) satisfying (82), let [đđ đđ ] ∈ đđFB (đĽđ , đŚđ ) for đ = 1, . . . , đ. Then, for any (Δđ˘)đ , (ΔđŁ)đ ∈ Rđđ , it holds that đđ (Δđ˘)đ + đđ (ΔđŁ)đ (ΔđŁ)đ = 0, { { { (Δđ˘) { đ = 0, { { â¨(Δđ˘) , đŚ ⊠= 0, â¨(ΔđŁ)đ , đĽđ ⊠= 0, = 0 ół¨⇒ { đ đ { { { (Δđ˘)đ ∈ R (đŚđ1 , −đŚđ2 ) , { { (ΔđŁ)đ ∈ R (đĽđ1 , −đĽđ2 ) , { đđ đ ∈ đ˝đź0 , đđ đ ∈ đ˝0đź , đđ đ ∈ đ˝đľđľ , đđ đ ∈ đ˝0đľ , đđ đ ∈ đ˝đľ0 . (89) 12 Abstract and Applied Analysis Particularly, for each đ ∈ đ˝đľđľ , the following implication also holds: đđ (Δđ˘)đ + đđ (ΔđŁ)đ { { â¨(Δđ˘)đ , (ΔđŁ)đ ⊠= { { { { = 0 ół¨⇒ { { { { { {â¨(Δđ˘)đ , (ΔđŁ)đ ⊠= { đŚđ1 đĽđ1 đĽđ1 đŚđ1 óľŠ2 óľŠ ((Δđ˘)2đ1 − óľŠóľŠóľŠóľŠ(Δđ˘)đ2 óľŠóľŠóľŠóľŠ ) , óľŠ2 óľŠ ((ΔđŁ)2đ1 − óľŠóľŠóľŠóľŠ(ΔđŁ)đ2 óľŠóľŠóľŠóľŠ ) . (90) đĽ2đ đŚ2đ Proof. Throughout the proof, let đ¤đ = + and đ§đ = √đ¤đ for đ = 1, 2, . . . , đ. We prove the conclusion by discussing five cases as shown in the following arguments. Case 1 (đ ∈ đ˝đź0 ). In this case đ§đ = đĽđ ∈ int Kđđ . From Corollary 10(a), it follows that the function đFB is continuously differentiable at (đĽđ , đŚđ ). Therefore, −1 −1 = {[đź − đż đ§đ đż đĽđ đź − đż đ§đ đż đŚđ ]} = {[0 đź]} . (91) Then, đđ = 0 and đđ = đź. Together with đđ (Δđ˘)đ + đđ (ΔđŁ)đ = 0, we get (ΔđŁ)đ = 0. Case 2 (đ ∈ đ˝0đź ). Using the same arguments as in Case 1 readily yields (Δđ˘)đ = 0. Case 3 (đ ∈ đ˝đľđľ ). Now đĽđ ∈ bd+ Kđđ and đŚđ ∈ bd+ Kđđ . By (82) and Lemma 7, there exists đđ > 0 such that đĽđ1 = đđ đŚđ1 and đĽđ2 = −đđ đŚđ2 , and consequently, đđ2 ) đŚ2đ1 (1 + đđ ) đŚđ1 đ§đ = ( ) = đĽ đ + đŚđ . (1 − đđ ) đŚđ2 (92) −1 −1 đđ = đź − đż−1 đ§đ đż đŚđ = đż đ§đ đż đ§đ −đŚđ = đż đ§đ đż đĽđ . Together with (94), we have â¨(Δđ˘)đ , đŚđ ⊠= 0 and â¨(ΔđŁ)đ , đĽđ ⊠= 0. We next prove the implication in (90). By the expressions of đ§đ , đđ , and đđ , đđ (Δđ˘)đ + đđ (ΔđŁ)đ = 0 ⇐⇒ [ đŚđ1 đŚđđ2 đĽ đĽđ ] (Δđ˘)đ + [ đ1 đ2 ] (ΔđŁ)đ = 0 đŚđ2 đŚđ1 đź đĽđ2 đĽđ1 đź đĽđ1 đŚđ1 −đŚđđ2 đŚ đŚđ ] (ΔđŁ)đ = 0 [ ⇐⇒ [ đ1 đ2 ] (Δđ˘)đ + đŚđ2 đŚđ1 đź đŚđ1 −đŚđ2 đŚđ1 đź ⇐⇒ [−đŚđ2 đŚđ1 đź] [ đĽđ1 (ΔđŁ)đ1 −(Δđ˘)đ1 ( )+( )] = 0, (ΔđŁ) (Δđ˘)đ2 đŚđ1 đ2 (97) for some đ ∈ R. (98) Making an inner product with (Δđ˘)đ for the equality and using â¨(Δđ˘)đ , đŚđ ⊠= 0, we get óľŠ2 óľŠ â¨(Δđ˘)đ , (ΔđŁ)đ ⊠= (Δđ˘)2đ1 − óľŠóľŠóľŠóľŠ(Δđ˘)đ2 óľŠóľŠóľŠóľŠ . (99) Using the similar arguments as above and noting that â¨(ΔđŁ)đ , đĽđ ⊠= 0, we may obtain đŚđ1 Thus, đđ (Δđ˘)đ + đđ (ΔđŁ)đ = 0 implies that đż đŚđ (Δđ˘)đ + đż đĽđ (ΔđŁ)đ = 0, and consequently, [−đŚđ2 đŚđ1 đź] (ΔđŁ)đ = 0 đĽđ1 (ΔđŁ)đ1 −(Δđ˘)đ1 ( )+( ) = đđŚđ (Δđ˘)đ2 đŚđ1 (ΔđŁ)đ2 đŚđ1 (93) đĽđ1 đŚđ1 where the second equivalence is due to đĽđ1 = đđ đŚđ1 and đĽđ2 = −đđ đŚđ2 . Since the rank of [−đŚđ2 đŚđ1 đź] is đđ −1, the dimension of the solution space for the system [−đŚđ2 đŚđ1 đź]đ = 0 equals 1. Note that đŚđ is a nonzero solution of this linear system. Therefore, đĽđ1 From Corollary 10(a), đFB is continuously differentiable at (đĽđ , đŚđ ), and hence −1 −1 đđ = đź − đż−1 đ§đ đż đĽđ = đż đ§đ đż đ§đ −đĽđ = đż đ§đ đż đŚđ , − (đĽđ1 (ΔđŁ)đ1 + đĽđđ2 (ΔđŁ)đ2 ) + (đŚđ1 (Δđ˘)đ1 + đŚđđ2 (Δđ˘)đ2 ) = 0. (96) ół¨⇒ [đŚđ2 đŚđ1 đź] (Δđ˘)đ + đđFB (đĽđ , đŚđ ) = {JđFB (đĽđ , đŚđ )} 2 (1 + đ¤đ = ( ) ∈ int Kđđ , 2 2 (1 − đđ ) đŚđ1 đŚđ2 Making an inner product with đŚđ2 /âđŚđ2 â, we have from đĽđ , đŚđ ∈ bd+ Kđđ and Lemma 5 that đĽđ1 óľŠ2 óľŠ â¨(Δđ˘)đ , (ΔđŁ)đ ⊠= (ΔđŁ)2đ1 − óľŠóľŠóľŠóľŠ(ΔđŁ)đ2 óľŠóľŠóľŠóľŠ . (100) The last two equalities show that the implication in (90) holds. (đĽđ1 (ΔđŁ)đ1 + đĽđđ2 (ΔđŁ)đ2 ) + (đŚđ1 (Δđ˘)đ1 + đŚđđ2 (Δđ˘)đ2 ) = 0, (94) Case 4 (đ ∈ đ˝0đľ ). By Lemma 15(b), there exists an đđ × đđ Ěđ đó¸ ] such that đđ = đđ đˇđ đđ orthogonal matrix đđ = [đđ đ đ đ đ and đđ = đđ Λ đ đđ , where đđ and đđó¸ are given by (86), and đđ and Λ đ take one of the form in (87)-(88). Thus, we have (đĽđ2 (ΔđŁ)đ1 + đĽđ1 (ΔđŁ)đ2 ) + (đŚđ2 (Δđ˘)đ1 + đŚđ1 (Δđ˘)đ2 ) = 0. (95) đđ (Δđ˘)đ + đđ (ΔđŁ)đ = 0 ⇐⇒ đˇđ đđđ (Δđ˘)đ + Λ đ đđđ (ΔđŁ)đ = 0. (101) Abstract and Applied Analysis 13 When đˇđ = đź and Λ đ = 0, we have đđđ (Δđ˘)đ = 0, and then (Δđ˘)đ = 0. When đˇđ and Λ đ take the form in (87), we have From the continuity and convexity of the Euclidean norm â ⋅ â and Lemma 4, we get đđ [ đĚđđ ] (Δđ˘)đ = 0. Consequently, â(đź − đ) Δđ˘ + (đź − đ) ΔđŁâ đ (Δđ˘)đ = đđ đđđ (Δđ˘)đ Ěđ đó¸ ] [ = [đđ đ đ [ [ đ = (đđó¸ ) (Δđ˘)đ √2đŚđ1 [ 0 0 đ óľŠ −1 óľŠóľŠóľŠ ≤ lim ∑đđ óľŠóľŠóľŠóľŠ(đż−1 đ§đđ đż đĽđđ Δđ˘ + đż đ§đđ đż đŚđđ ΔđŁ)óľŠ óľŠ đ→∞ đ=1 ] ] đ (đđó¸ ) (Δđ˘)đ ] (106) ≤ √âΔđ˘â2 + âΔđŁâ2 . (102) Now assume that đΔđ˘+đΔđŁ = 0. Then, by the last inequality, we get the result. đŚđ1 ], −đŚđ2 where the last equality is using the definition of đđó¸ . When đˇđ đđ and Λ đ take the form in (88), we also have [ đĚđđ ] (Δđ˘)đ = 0. Proposition 18. For (đĽ, đŚ) satisfying (82), let [đ đ] đđFB (đĽ, đŚ). Then it holds that ∈ đΔđ˘ + đΔđŁ = 0 ół¨⇒ â¨Δđ˘, ΔđŁâŠ đ đ Using the same arguments as above, we have that (Δđ˘)đ has the form of (102). Thus, we prove that (Δđ˘)đ ∈ R(đŚđ1 , −đŚđ2 ). ≤ ∑ ΥđĽđ (đŚđ , (Δđ˘)đ ) ∀Δđ˘, ΔđŁ ∈ Rđ , (107) đ=1 Case 5 (đ ∈ đ˝đľ0 ). Using Lemma 15(a) and following the same arguments as in Case 4, the result can be checked routinely. So, we omit the proof. where, for any given đ ∈ Rđ , Υđđ : Rđđ × Rđđ → R is the linear-quadratic function: The following lemma states an important property for the elements of Clarke’s Jacobian of đFB at a general point, which will be used to prove Proposition 18. đ { { đ1 đđ (1 0 ) đđ , đđ đđ ∈ bd+ Kđđ , Υđđ (đđ , đđ ) := { đđ1 đ 0 −đź { 0, đđĄâđđđ¤đđ đ. { (108) Lemma 17. For any given (đĽ, đŚ) ∈ Rđ × Rđ , let [đ đ] ∈ đđFB (đĽ, đŚ). Then, đΔđ˘ + đΔđŁ = 0 ół¨⇒ â¨Δđ˘, ΔđŁâŠ ≤ 0, ∀Δđ˘, ΔđŁ ∈ Rđ . (103) Proof. Since [đ đ] ∈ đđFB (đĽ, đŚ), by CaratheĚodory’s theorem, there exist a positive integer đ and [đđ đđ ] ∈ đđľ đFB (đĽ, đŚ) for đ = 1, . . . , đ such that [đ đ] = ∑đ đ=1 đđ [đđ đđ ] where đđ ≥ 0, đ = 1, . . . , đ , and ∑đ đ=1 đđ = 1. For each đ ∈ {1, . . . , đ }, by Corollary 10(a) and the definition of the elements in đđľ đFB (đĽ, đŚ), there exists a sequence {(đĽđđ , đŚđđ )} in Rđ × Rđ converging to (đĽ, đŚ) with đ§đđ = đ§(đĽđđ , đŚđđ ) âť 0 such that [đđ đđ ] = lim JđFB (đĽđđ , đŚđđ ) đ→∞ đż đ đź − đż−1 đż đ = lim [đź − đż−1 đ§đđ đĽ đ đ§đđ đŚ đ ] . (104) đ→∞ đ→∞ đ=1 đ = 1, 2, . . . , đ. (109) From Lemma 17 and (90) of Proposition 16, it then follows that â¨Δđ˘, ΔđŁâŠ ≤ ∑ â¨(Δđ˘)đ , (ΔđŁ)đ ⊠đ∈đ˝đľđľ = ∑ đŚđ1 đ∈đ˝đľđľ đĽđ1 [(Δđ˘)2đ1 óľŠ2 óľŠ − óľŠóľŠóľŠóľŠ(Δđ˘)đ2 óľŠóľŠóľŠóľŠ ] . (110) In addition, by the definition of Υđđ , ΥđĽđ (đŚđ , ⋅) = 0 for all đ ∈ đ˝đź0 ∪ đ˝0đź ∪ đ˝0đľ ∪ đ˝00 since đĽđ ∉ bd+ Kđđ , and ΥđĽđ (đŚđ , ⋅) = 0 for đ ∈ đ˝đľ0 since đŚđ1 = 0. This means that đ đ=1 đ∈đ˝đľđľ = ∑ đŚđ1 đ∈đ˝đľđľ đĽđ1 (đź − đ) Δđ˘ + (đź − đ) ΔđŁ −1 = lim ∑đđ (đż−1 đ§đđ đż đĽđđ Δđ˘ + đż đ§đđ đż đŚđđ ΔđŁ) . đđ (Δđ˘)đ + đđ (ΔđŁ)đ = 0, ∑ ΥđĽđ (đŚđ , (Δđ˘)đ ) = ∑ ΥđĽđ (đŚđ , (Δđ˘)đ ) Consequently, for any Δđ˘, ΔđŁ ∈ Rđ , we have that đ Proof. Fix any Δđ˘, ΔđŁ ∈ Rđ . Since đ = diag(đ1 , . . . , đđ ) and đ = diag(đ1 , . . . , đđ ) with [đđ đđ ] ∈ đđFB (đĽđ , đŚđ ) for đ = 1, 2, . . . , đ, we may rewrite đΔđ˘ + đΔđŁ = 0 as (105) [(Δđ˘)2đ1 óľŠ2 óľŠ − óľŠóľŠóľŠóľŠ(Δđ˘)đ2 óľŠóľŠóľŠóľŠ ] . (111) From the above two equations, we immediately obtain the desired result. 14 Abstract and Applied Analysis Before stating the main result of this section, we also need to recall several concepts, including constraint nondegeneracy, Robinson’s constraint qualification (CQ) (see [20]), and the strong second-order sufficient condition introduced in [7, Theorem 30]. To this end, let đ ≡ (đ, đĽ) ∈ R2đ and define đĚ : R2đ → R, Ěâ : R2đ → Rđ+đ , đĚ : R2đ → Rđ by đĚ (đ) := đ (đ) , Ěâ (đ) := ( â (đ) ) , đ (đ) − đĽ Then, we may rewrite the nonlinear SOCP (4) succinctly as follows: min đĚ (đ) Definition 19. A feasible vector đ = (đ, đĽ) of (4) is called constraint nondegenerate if {0} Rđ+đ J Ěâ (đ) ) R2đ + ( ( đ ) = ( đ ), lin (TK (đĽ)) R Jđ đĚ (đ) (114) where lin(TK (đĽ)) is the largest linear space of TK (đĽ), that is, lin(TK (đĽ)) = TK (đĽ) ∩ −TK (đĽ). Definition 20. Robinson’s CQ is said to hold at a feasible solution đ = (đ, đĽ) to (4) if Ěâ (đ) {0} J Ěâ (đ) )+( đ ) R2đ − ( )} , 0 ∈ int {( K đĚ (đ) Jđ đĚ (đ) (115) đ which, since K is a closed convex set in R , can be equivalently written as Rđ+đ {0} J Ěâ (đ) ) R2đ + ( ( đ ) = ( đ ). TK (đĽ) R Jđ đĚ (đ) (116) Clearly, the constraint nondegenerate condition (114) implies Robinson’s CQ (116). If đ = (đ, đĽ) is a locally optimal solution to (4) and Robinson’s CQ holds at đ, then there exists a Lagrange multiplier (đ, đ , đŚ), together with đ, satisfying the KKT conditions: â (đ) = 0, −đŚ ∈ NK (đĽ) . (117) In the sequel, we let Λ(đ) denote the set of Lagrange multipliers satisfying (117). Let (đ, đĽ, đ, đ , đŚ) ∈ Rđ × K × Rđ × Rđ × K be a KKT point of the SOCP (4). From [7, Lemma 25], it follows that the tangent cone of K at đĽ takes the form of TK (đĽ) = {đ ∈ Rđ | đđ ∈ Kđđ for đ ∈ đ˝0đź ∪ đ˝0đľ ∪ đ˝00 ; đđđ (đĽđ1 , −đĽđ2 ) ≥ 0 for đ ∈ đ˝đľđľ ∪ đ˝đľ0 } , (118) (119) đđđ (đĽđ1 , −đĽđ2 ) = 0 for đ ∈ đ˝đľđľ ∪ đ˝đľ0 } . We next recall the critical cone of problem (4) at a feasible đ0 which is defined as C (đ0 ) := {đ ∈ R2đ | Jđ Ěâ (đ0 ) đ = 0, (113) đĚ (đ) ∈ K. đ (đ) − đĽ = 0, = {đ ∈ Rđ | đđ = 0 for đ ∈ đ˝0đź ∪ đ˝0đľ ∪ đ˝00 ; Jđ đĚ (đ0 ) đ ∈ TK (đĚ (đ0 )) , s.t. Ěâ (đ) = 0, Jđ,đĽ đż (đ, đĽ, đ, đ , đŚ) = 0, lin (TK (đĽ)) đĚ (đ) := đĽ. (112) đ∈R2đ which implies that the largest linear space in TK (đĽ) has the following form: (120) Jđ đĚ (đ0 ) đ ≤ 0} . The critical cone C(đ0 ) represents those directions for which the linearization of (4) does not provide any information about optimality of đ0 and is very important in studying second-order optimality conditions. Particularly, if the set of Lagrange multipliers Λ(đ0 ) at the point đ0 is nonempty, then C(đ0 ) can be rewritten as C (đ0 ) = {đ ∈ R2đ | Jđ Ěâ (đ0 ) đ = 0, (121) ⊥ Jđ đĚ (đ0 ) đ ∈ TK (đĚ (đ0 )) ∩ (đŚ0 ) } , where đŚ0 ∈ Λ(đ0 ) and (đŚ0 )⊥ means the orthogonal complementarity space of đŚ0 . Now let (đ, đ , đŚ) ∈ Λ(đ). Then, using Ě = đĽ and the expression of TK (đĽ), we have that đ(đ) C (đ) = {đ ∈ R2đ | Jđ đĚđ (đ) đ = 0 for đ ∈ đ˝0đź , Jđ đĚđ (đ) đ ∈ R+ (đŚđ1 , −đŚđ2 ) for đ ∈ đ˝0đľ , â¨Jđ đĚđ (đ) đ, đŚđ ⊠= 0 for đ ∈ đ˝đľđľ , Jđ đĚđ (đ) đ ∈ Kđđ for đ ∈ đ˝00 , (122) (Jđ đĚđ (đ) đ) (đĽđ1 , −đĽđ2 ) ≥ 0 for đ ∈ đ˝đľ0 ; Jđ Ěâ (đ) đ = 0} . Definition 21. Let đ = (đ, đĽ) be a stationary point of the SOCP (4) such that Λ(đ) = {(đ, đ , đŚ)}. We say that the strong second-order sufficient condition holds at đ if â¨đ, J2đđ đż (đ, đĽ, đ, đ , đŚ) đ⊠+ â¨đ, đť (đ, đŚ) đ⊠> 0, ∀đ ∈ aff (C (đ)) \ {0} , (123) Abstract and Applied Analysis 15 where đť(đ, đŚ) = ∑đđ=1 đťđ (đ, đŚ) ∈ R2đ×2đ with đťđ (đ, đŚ) ∈ R2đ×2đ defined by From the expression of đ¸FB , we know that there exists a [đ đ] ∈ đđFB (đĽ, đŚ) such that đ đťđ (đ, đŚ) đŚ { { − đ1 Jđ đĚđ (đ)đ (1 0 ) Jđ đĚđ (đ) , if đ ∈ đ˝đľđľ ∪ đ˝đľ0 ; 0 −đź := { đĽđ1 { 0, otherwise, { (124) and aff(C(đ)) denotes the affine hull of C(đ) and is now equivalent to the span of C(đ): đ Jđđ đ (đ, đ, đ ) Δđ + Jđ â(đ) Δđ + Jđ đ(đ) Δđ ] [ −Δđ − ΔđŚ ] [ ] [ ] = 0, (129) [ Jđ â (đ) Δđ ] [ ] [ Jđ đ (đ) Δđ − ΔđĽ đΔđĽ + đΔđŚ ] [ where đ = diag(đ1 , . . . , đđ ) and đ = diag(đ1 , . . . , đđ ) with [đđ đđ ] ∈ đđFB (đĽđ , đŚđ ) for all đ = 1, 2, . . . , đ. The last system can be simplified as đ aff (C (đ)) = {đ ∈ R2đ | Jđ Ěâ (đ) đ = 0; Jđ đĚđ (đ) đ = 0 for đ ∈ đ˝0đź , â¨Jđ đĚđ (đ) đ, đŚđ ⊠= 0 for đ ∈ đ˝đľđľ , By the second and the third equations of (130) and (89), we get Jđ đĚđ (đ) đ ∈ R (đŚđ1 , −đŚđ2 ) for đ ∈ đ˝0đľ } . (125) Now we are in a position to prove the nonsingularity of Clarke’s Jacobian of đ¸FB under the strong second-order sufficient condition and constraint nondegeneracy. Proposition 22. Let (đ, đĽ, đ, đ , đŚ) be a KKT point of the nonlinear SOCP (4). Suppose that the strong second-order sufficient condition (123) holds at đ = (đ, đĽ) and đ is constraint nondegenerate, then any element in đđ¸FB (đ, đĽ, đ, đ , đŚ) is nonsingular. Proof. Since the nondegeneracy condition (114) is assumed to hold at (đ, đĽ), from [21, Proposition 4.75], we know that Ě Λ(đ) = {(đ, đ , đŚ)}. Then, by the definition of đż and đ, the strong second-order sufficient condition (123) takes the following form: Ě (đĽ, đŚ) đ⊠> 0, â¨đ, J2đđ đ (đ, đ, đ ) đ⊠− â¨đ, đť ∀ (đ, đ) ∈ aff (C (đ)) \ {(0, 0)} , (126) Ě đŚ) = where đ(đ, đ, đ ) := đ(đ) + â¨đ, â(đ)⊠+ â¨đ(đ), đ ⊠and đť(đĽ, đ đ×đ Ě ∑đ=1 đťđ (đĽ, đŚ) ∈ R with đŚđ1 { { Ěđ (đĽ, đŚ) := − đĽ đť { đ1 { { 1 0 ( ) , if đ ∈ đ˝đľđľ ∪ đ˝đľ0 ; 0 −đź 0, otherwise. đ 2 [Jđđ đ (đ, đ, đ ) Δđ + Jđ â(đ) Δđ − Jđ đ(đ) ΔđŚ] ] = 0. [ Jđ â (đ) Δđ ] [ đ (đ) Δđ + đΔđŚ đJ đ ] [ (130) Jđ â (đ) Δđ = 0, (Jđ đ (đ) Δđ)đ = R (đŚđ1 , −đŚđ2 ) , (Jđ đ (đ) Δđ)đ = 0, for đ ∈ đ˝0đľ , (131) for đ ∈ đ˝0đź , â¨(Jđ đ (đ) Δđ)đ , đŚđ ⊠= 0, for đ ∈ đ˝đľđľ . Comparing with the definition of aff(C(đ)) in Definition 21, it follows that (Δđ, Jđ đ (đ) Δđ) ∈ aff (C (đ)) . (132) From the first and the second equations of (130), it is not hard to verify đ đ 0 = â¨Δđ, J2đđ đ (đ, đ, đ ) Δđ + Jđ â(đ) Δđ − Jđ đ(đ) ΔđŚâŠ = â¨Δđ, J2đđ đ (đ, đ, đ ) Δđ⊠− â¨Jđ đ (đ) Δđ, ΔđŚâŠ , (133) which, together with the third equation of (130) and Proposition 18, implies that đ 0 ≥ â¨Δđ, J2đđ đ (đ, đ, đ ) Δđ⊠− ∑ΥđĽđ (đŚđ , Jđ đđ (đ) Δđ) . đ=1 (134) (127) This, together with (132) and (126), yields that Δđ = 0. Thus, (130) reduces to Let đ be an arbitrary element in đđ¸FB (đ, đĽ, đ, đ , đŚ). To prove that đ is nonsingular, let (Δđ, ΔđĽ, Δđ, Δđ , ΔđŚ) ∈ Rđ × Rđ × Rđ × Rđ × Rđ such that [Jđ â(đ) Δđ − Jđ đ(đ) ΔđŚ] = 0. đΔđŚ đ (Δđ, ΔđĽ, Δđ, Δđ , ΔđŚ) = 0. From the second equation of (135), we have đđ (ΔđŚ)đ = 0 for đ ∈ đ˝đľđľ . In addition, by the arguments for Case 3 of (128) đ đ (135) 16 Abstract and Applied Analysis Proposition 16, đđ = đż−1 đ§đ đż đĽđ for đ ∈ đ˝đľđľ , and so đż đĽđ (ΔđŚ)đ = 0 for đ ∈ đ˝đľđľ . Since đĽđ ∈ bd+ Kđđ for đ ∈ đ˝đľđľ , đż đĽđ has the two single eigenvalues 0 and 2đĽđ1 as well as the eigenvalues đĽđ1 with multiplicity đđ − 2, and (1, −đĽđ2 /đĽđ1 ) is the eigenvector corresponding to eigenvalue 0. Thus, from đż đĽđ (ΔđŚ)đ = 0, we deduce (ΔđŚ)đ ∈ R (đĽđ1 , −đĽđ2 ) , for đ ∈ đ˝đľđľ . (136) By the second equation of (135), we use Proposition 16 with Δđ˘ = 0, ΔđŁ = ΔđŚ to yield (ΔđŚ)đ = 0, for đ ∈ đ˝đź0 , (ΔđŚ)đ ∈ R (đĽđ1 , −đĽđ2 ) , for đ ∈ đ˝đľ0 . (137) Using the constraint nondegeneracy condition (114), we know that there exist a vector (đ1 , đ2 ) ∈ Rđ × Rđ and a vector đ ∈ lin(TK (đĽ)) such that Jđ â (đ) đ1 = Δđ, Jđ đ (đ) đ1 − đ2 = −ΔđŚ, đ2 + đ = −ΔđŚ. (138) Since đ ∈ lin(TK (đĽ)), from (119), it follows that đđ = 0, for đ ∈ đ˝0đź ∪ đ˝0đľ ∪ đ˝00 , đđđ (đĽđ1 , −đĽđ2 ) = 0, for đ ∈ đ˝đľđľ ∪ đ˝đľ0 . (139) Combining the last four equations with the first equation of (135), we obtain â¨Δđ, Δđ⊠+ 2 â¨ΔđŚ, ΔđŚâŠ = â¨Jđ â (đ) đ1 , Δđ⊠− â¨Jđ đ (đ) đ1 − đ2 , ΔđŚâŠ − â¨đ2 + đ, ΔđŚâŠ đ (140) đ = â¨đ1 , Jđ â(đ) Δđ − Jđ đ(đ) ΔđŚâŠ − â¨đ, ΔđŚâŠ = − â¨đ, ΔđŚâŠ = 0. Thus, Δđ = 0 and ΔđŚ = 0. Along with Δđ = 0, we show that đ is nonsingular. Note that −đŚ ∈ NK (đĽ) if and only if −đĽ ∈ NK (đŚ). The KKT conditions in (7) can be equivalently written as the following generalized equation Jđ,đĽ đż (đ, đĽ, đ, đ , đŚ) NRđ ×Rđ (đ, đĽ) ] [ NRđ (đ) ] [ â (đ) ] ]+[ 0∈[ ] [ NRđ (đ ) ] , [ đ (đ) − đĽ đĽ ] [ NK (đŚ) ] [ Theorem 23. Let (đ, đĽ) be a locally optimal solution to the nonlinear SOCP (4). Suppose that Robinson’s CQ holds at this point. Let (đ, đ , đŚ) ∈ Rđ ×Rđ ×K be such that (đ, đĽ, đ, đ , đŚ) is a KKT point of (4). Then the following statements are equivalent. (a) The strong second-order sufficient condition in Definition 21 holds at (đ, đĽ) and (đ, đĽ) is constraint nondegenerate. (b) Any element in đđ¸FB (đ, đĽ, đ, đ , đŚ) is nonsingular. (c) Any element in đđ¸NR (đ, đĽ, đ, đ , đŚ) is nonsingular. (d) (đ, đĽ, đ, đ , đŚ) is a strongly regular solution of the generalized equation (141). Proof. First, Lemma 14 and the definition of đ¸FB and đ¸NR imply the following inclusion: đđ¸NR (đ, đĽ, đ, đ , đŚ) ⊆ đđ¸FB (đ, đĽ, đ, đ , đŚ) . (142) Using this inclusion and Proposition 22, we have that (a) ⇒ (b) ⇒ (c). Since the SOCP (4) is obtained from (1) by introducing a slack variable, we know from [9, Theorem 3.1] that (a) ⇔ (c) ⇔ (d). Thus, we complete the proof of this theorem. 5. Conclusions In this paper, for a locally optimal solution of the nonlinear SOCP, we established the equivalence between the nonsingularity of Clarke’s Jacobian of the FB system and the strong regularity of the corresponding KKT point. This provides a new characterization for the strong regularity of the nonlinear SOCPs and extends the result of [22, Corollary 3.7] for the FB system of variational inequalities with the polyhedral cone Rđ+ constraints to the setting of SOCs. Also, this result implies that the semismooth Newton method [5, 6] applied to the FB system is locally quadratically convergent to a KKT point under the strong second-order sufficient condition and constraint nondegeneracy. We point it out that we have also established parallel (not exactly the same) results for SDP case in [11] recently. However, it seems hard to put them together in a unified framework under Euclidean Jordan algebra. The main reason causing this is due to that the analysis and techniques are totally different when dealing with the Clarke Jacobians associated with FB SOC complementarity function and FB SDP complementarity function. Appendix (141) which is clear in the form of the generalized equation given by (29). Now using Proposition 22 and [9, Theorem 3.1], we may establish the main result of this paper, which states that Clarke’s Jacobian of đ¸FB at a KKT point is nonsingular if and only if the KKT point is a strongly regular solution to the generalized equation (141). Lemma A.1. For any given đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) with đĽ2 + đŚ2 ∈ bd+ Kđ , it holds that đđľ đFB (đĽ, đŚ) = đđľ đFB (0, 0) ó¸ đ¤đđĄâ đFB (⋅, ⋅) ≡ đFB ((đĽ, đŚ) ; (⋅, ⋅)) . (A.1) Proof. We first prove that đđľ đFB (0, 0) ⊆ đđľ đFB (đĽ, đŚ). Let [đ đ] ∈ đđľ đFB (0, 0). By the formula (32) and Lemma 11, Abstract and Applied Analysis 17 there exists a vector (đ1 , đ2 ) ∈ R × Rđ−1 with đ12 + âđ2 â2 = 1 such that đ˘ and đŁ defined by (60) satisfy where the last two equalities are using Lemma 5. In addition, we compute that 1 1 ) đŁđ , ( đ¤2 2 −Ě (A.2) đż 1 (đ¤đ ) đż đĽđ = where đ and đ are defined by (64). Take the sequences {đĽđ } and {đŚđ } with đż 1 (đ¤đ ) đż đŚđ = đ=đź−đ− 1 1 ) đ˘đ , ( đ¤2 2 −Ě 1 đĽ1 + đĽ1 đ1 đ đĽ ≡( ), 1 đĽ2 − đŚ1 đ2 đ đ=đź−đ− 1 đŚ1 + đŚ1 đ1 đ đŚ ≡( ) . (A.3) 1 đŚ2 + đĽ1 đ2 đ đ đ Let đ¤đ = (đ¤1đ , đ¤2đ ) = (đĽđ )2 + (đŚđ )2 . By Lemma 5, a simple computation yields that đ¤2đ = (1 + đ1 ) đ¤2 , đ đ 1 (đ¤đ ) = đĽ12 + đŚ12 , đ2 1 đ 2 (đ¤ ) = đ 2 (đ¤) + đ ( ) . đ (A.4) Clearly, đ§đ = √đ¤đ ∈ int Kđ . From Corollary 10(a), it then follows that lim JđĽ đFB (đĽđ , đŚđ ) = đź − lim đż−1 đ§đ đż đĽđ , đ→∞ đ đ lim JđŚ đFB (đĽ , đŚ ) = đź − đ→∞ lim đż−1đ đż đŚđ . đ→∞ đ§ (A.5) 1 đż 1 (đ¤ ) := 2√đ 1 (đ¤đ ) 1 = đ −(Ě đ¤2đ ) ( đ đ đ đ) , Ě2 (Ě −Ě đ¤2 đ¤ đ¤2 ) 2√đ 2 (đ¤đ ) (Ě đ¤2đ ) 1 Ě2đ đ¤ = Ě2đ đĽ1đ − (đĽ2đ ) đ¤ √đ 1 (đ¤đ ) , Ě2đ đĽ2đ − đĽ1đ đ¤ đ˘2đ = √đ 1 (đ¤đ ) đŁ1đ = Ě2đ đŚ1đ − (đŚ2đ ) đ¤ √đ 1 (đ¤đ ) , (A.9) , Ě2đ đŚ2đ − đŚ1đ đ¤ đŁ2đ = √đ 1 (đ¤đ ) . Together with the definition of đĽđ and đŚđ and (A.4), we can verify that lim (đ˘1đ , đ˘2đ ) = đ˘, đ→∞ lim (đŁ1đ , đŁ2đ ) = đŁ. đ→∞ (A.10) lim JđĽ đFB (đĽđ , đŚđ ) = đź − đ − 1 1 ) đ˘đ ; ( đ¤2 2 −Ě lim JđŚ đFB (đĽđ , đŚđ ) = đź − đ − 1 1 ) đŁđ . ( đ¤2 2 −Ě đ→∞ (A.11) Comparing this with (A.2), we have [đ đ] ∈ đđľ đFB (đĽ, đŚ). So, đđľ đFB (0, 0) ⊆ đđľ đFB (đĽ, đŚ). In what follows, we show that đđľ đFB (đĽ, đŚ) ⊆ đđľ đFB (0, 0). Note that đĽ2 + đŚ2 ∈ bd+ Kđ is equivalent to (đĽ, đŚ) ≠ (0, 0) and âđĽâ2 + âđŚâ2 = 2âđĽ1 đĽ2 + đŚ1 đŚ2 â, which is equivalent to 1 ×( đ đ˘1đ đ→∞ đż 2 (đ¤đ ) := where đ˘đ = (đ˘1đ , đ˘2đ ), đŁđ = (đŁ1đ , đŁ2đ ) ∈ R × Rđ−1 are defined as follows: Thus, the above arguments show that Ě2đ = đ¤2đ /âđ¤2đ â. Using the formula (19), we have that đż−1 Let đ¤ đ§đ đż(đ¤1đ ) + đż(đ¤2đ ), where đ đŁ2đ đŁđ đ 1 1 ) (đŁđ ) , ( đ1 đ đ) = ( −Ě đ¤2đ 2 −đŁ1 đ¤ Ě2 −Ě đ¤2đ (đŁ2đ ) (A.8) đ đ đ→∞ đ˘2đ đ˘đ 1 1 đ đ ( đ1 đ đ ) (đ˘ ) , đ đ đ ) = (−Ě đ¤ 2 −đ˘1 đ¤ Ě2 −Ě đ¤2 (đ˘2 ) 2 4√đ 2 (đ¤đ ) (đź − đ đ Ě2đ (Ě đ¤ đ¤2đ ) ) √đ 2 (đ¤đ ) + √đ 1 (đ¤đ ) Ě2đ (Ě đ¤2đ ) +đ¤ đ) . óľ¨ óľ¨ óľŠ óľŠ2 óľ¨ óľ¨ óľŠ óľŠ2 (óľ¨óľ¨óľ¨đĽ1 óľ¨óľ¨óľ¨ − óľŠóľŠóľŠđĽ2 óľŠóľŠóľŠ) + (óľ¨óľ¨óľ¨đŚ1 óľ¨óľ¨óľ¨ − óľŠóľŠóľŠđŚ2 óľŠóľŠóľŠ) óľŠ óľ¨ óľ¨óľŠ óľŠ óľ¨ óľ¨óľŠ óľŠ óľŠ + 2 (óľ¨óľ¨óľ¨đĽ1 óľ¨óľ¨óľ¨ óľŠóľŠóľŠđĽ2 óľŠóľŠóľŠ + óľ¨óľ¨óľ¨đŚ1 óľ¨óľ¨óľ¨ óľŠóľŠóľŠđŚ2 óľŠóľŠóľŠ − óľŠóľŠóľŠđĽ1 đĽ2 + đŚ1 đŚ2 óľŠóľŠóľŠ) = 0. (A.6) (A.12) Ě2đ → đ¤ Ě2 as Since đ 1 (đ¤đ ) → 0, đ 2 (đ¤đ ) → 2đ¤1 > 0 and đ¤ đ → ∞, it follows that Hence, đĽ2 + đŚ2 ∈ bd+ Kđ is equivalent to saying that đĽ = (đĽ1 , đĽ2 ), đŚ = (đŚ1 , đŚ2 ) satisfy lim đż 2 (đ¤đ ) đż đĽđ = đ→∞ lim đż 2 (đ¤đ ) đż đŚđ = đ→∞ 1 Ě2đ 1 đ¤ ( ) đż = đ, Ě2 4đź − 3Ě Ě2đ đĽ đ¤2 đ¤ 2√2đ¤1 đ¤ Ě2đ đ¤ 1 1 ( ) đż = đ, Ě Ě2đ đŚ đ¤2 đ¤ đ¤ 2√2đ¤1 2 4đź − 3Ě (A.7) óľ¨óľ¨ óľ¨óľ¨ óľŠóľŠ óľŠóľŠ óľ¨óľ¨ óľ¨óľ¨ óľŠóľŠ óľŠóľŠ óľ¨óľ¨đŚ1 óľ¨óľ¨ = óľŠóľŠđŚ2 óľŠóľŠ , óľ¨óľ¨đĽ1 óľ¨óľ¨ = óľŠóľŠđĽ2 óľŠóľŠ , óľ¨ óľ¨óľ¨ óľ¨óľŠ óľŠóľŠ óľŠ đĽ1 đŚ1 đĽ2đ đŚ2 = óľ¨óľ¨óľ¨đĽ1 óľ¨óľ¨óľ¨ óľ¨óľ¨óľ¨đŚ1 óľ¨óľ¨óľ¨ óľŠóľŠóľŠđĽ2 óľŠóľŠóľŠ óľŠóľŠóľŠđŚ2 óľŠóľŠóľŠ , (đĽ, đŚ) ≠ (0, 0) . (A.13) This means that đĽ, đŚ must satisfy one of the following cases: (i) đĽ = đźđŚ for some đź ∈ R and đŚ2 ∈ bd+ Kđ ; (ii) đŚ = đ˝đĽ 18 Abstract and Applied Analysis for some đ˝ ∈ R and đĽ2 ∈ bd+ Kđ . Since đFB and đFB are symmetric with respect to two arguments, we only need to prove one of the two cases. In the following arguments, we assume that đĽ = đźđŚ for some đź ∈ R and đŚ2 ∈ bd+ Kđ . Noting that đŚ1 ≠ 0 since đŚ2 ∈ bd+ Kđ , we without loss of generality assume that đŚ1 > 0. From (32) and Lemma 11, it is not hard to see that đđľ đFB (0, 0) đŁ=( (A.14) đ đ = đź − đ − lim (A.15) (đŁ1đ , đŁ2đ ) defined by with đ˘ = = (A.9). Thus, in order to prove that [đ đ] ∈ đđľ đFB (0, 0), it suffices to argue that the following limits: đ lim đ→∞ đ đŚĚ + đźđ1 −đ2 − đźđ1 đŚĚ2 =( 2 2 , ), √1 + đź2 √1 + đź2 đ lim (đŁ1đ , đŁ2đ ) = ( đ→∞ −đźđ2 đŚĚ2 + đ1 đźđ2 − đ1 đŚĚ2 , ), √1 + đź2 √1 + đź2 Ě2đ = đ¤2đ /âđ¤2 â and đ¤2đ = 2âđĽ1đ đĽ2đ + đŚ1đ đŚ2đ â. An where đ¤ elementary computation yields that 2 óľŠ2 óľŠ = (1 + đź2 ) đ 1 (đ¤đ ) − óľŠóľŠóľŠóľŠđŚ2đ + đźđĽ2đ óľŠóľŠóľŠóľŠ Let [đ đ] ∈ đđľ đFB (đĽ, đŚ). By the definition of the elements in đđľ đFB (đĽ, đŚ), there exists a sequence {(đĽđ , đŚđ )} with đ¤đ = (đĽđ )2 +(đŚđ )2 ∈ int Kđ converging to (đĽ, đŚ) such that [đ đ] = limđ → ∞ JđFB (đĽđ , đŚđ ). From the arguments for the first part, we know that (đ˘1đ , đ˘2đ ) (A.17) (1 + đź2 ) đ 1 (đ¤đ ) óľŠ2 óľŠ Ě2đ óľŠóľŠóľŠóľŠ + óľŠóľŠóľŠóľŠ(đźđŚ2đ − đĽ2đ ) − (đźđŚ1đ − đĽ1đ ) đ¤ are defined by (64) } . (đ˘1đ , đ˘2đ ) and đŁđ 2 Ě2đ ] [(đźđĽ1đ + đŚ1đ ) − (đźđĽ2đ + đŚ2đ ) đ¤ đ the đ × đ symmetric matrices đ and đ đ 2 óľŠ2 óľŠ (đźđ˘1đ + đŁ1đ ) + óľŠóľŠóľŠóľŠđźđŁ2đ − đ˘2đ óľŠóľŠóľŠóľŠ 1 + đź2 Ě2đ ] [(đźđĽ1đ + đŚ1đ ) − (đźđĽ2đ + đŚ2đ ) đ¤ đŚ where đŚĚ2 = óľŠóľŠ 2 óľŠóľŠ and óľŠóľŠđŚ2 óľŠóľŠ đ 1 1 ) (đŁ1đ , (đŁ2đ ) ) , đ = đź − đ − lim ( đ¤2 đ → ∞ 2 −Ě 2 óľŠóľŠ óľŠ2 óľŠóľŠ(đźđŚ2đ − đĽ2đ ) − (đźđŚ1đ − đĽ1đ ) đ¤ Ě2đ óľŠóľŠóľŠóľŠ óľŠ , + (1 + đź2 ) đ 1 (đ¤đ ) óľŠ óľŠ2 with đ = (đ1 , đ2 ) satisfying đ12 + óľŠóľŠóľŠđ2 óľŠóľŠóľŠ = 1, đ 1 1 ) (đ˘1đ , (đ˘2đ ) ) , ( đ¤2 đ → ∞ 2 −Ě 2 Step 1. To prove limđ → ∞ (((đźđ˘1đ đŁ1đ ) +âđźđŁ2đ − đ˘2đ â )/(1+đź2 )) = 1 (taking a subsequence if necessary). For each đ, by the expressions of đ˘1đ , đ˘2đ , and đŁ1đ , đŁ2đ , it is easy to see that = đ2đ đŚĚ2 + đźđ1 −đ2 − đźđ1 đŚĚ2 , ), √1 + đź2 √1 + đź2 −đźđ2đ đŚĚ2 + đ1 đźđ2 − đ1 đŚĚ2 , ) √1 + đź2 √1 + đź2 2 đ óľ¨óľ¨ 1 1 1 1 óľ¨ = {[đź − đ − ( ) đ˘đ đź − đ − ( ) đŁđ ] óľ¨óľ¨óľ¨ đ¤2 đ¤2 óľ¨óľ¨ 2 −Ě 2 −Ě for some đ˘ = ( 2 hold for some đ = (đ1 , đ2 ) with đ1 + âđ2 â = 1. We proceed the arguments by two steps. (A.18) 2 [(đźđĽ2đ + đŚ2đ ) (đĽ1đ đĽ2đ + đŚ1đ đŚ2đ )] + . óľŠóľŠ đ đ óľŠ2 óľŠóľŠđĽ1 đĽ2 + đŚ1đ đŚ2đ óľŠóľŠóľŠ From the last two equations, we immediately obtain that 2 óľŠ2 óľŠ (đźđ˘1đ + đŁ1đ ) + óľŠóľŠóľŠóľŠđźđŁ2đ − đ˘2đ óľŠóľŠóľŠóľŠ 1 + đź2 óľŠóľŠ đ óľŠ2 óľŠóľŠđŚ2 + đźđĽ2đ óľŠóľŠóľŠ óľŠ óľŠ =1− (1 + đź2 ) đ 1 (đ¤đ ) đ (A.19) 2 [(đźđĽ2đ + đŚ2đ ) (đĽ1đ đĽ2đ + đŚ1đ đŚ2đ )] + óľŠ2 . óľŠ (1 + đź2 ) đ 1 (đ¤đ ) óľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠ This shows that, in order to achieve the result, it suffices to prove that 2 óľŠóľŠ đ đóľŠ óľŠóľŠ2 [(đźđĽđ + đŚđ )đ (đĽđ đĽđ + đŚđ đŚđ )] óľŠ đŚ + đźđĽ 2 2 1 2 1 2 óľŠ óľŠ 2 2 [ ] óľŠ − lim [ óľŠ ] = 0, óľŠ2 óľŠ đ→∞ đ 1 (đ¤đ ) đ 1 (đ¤đ ) óľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠ [ ] (A.20) which is equivalent to arguing that (A.16) đ (đĽđ , đŚđ ) óľŠ2 = 0 óľŠ đ đ đ → ∞ đ (đ¤đ ) óľŠ óľŠóľŠđĽ1 đĽ2 + đŚ1đ đŚ2đ óľŠóľŠóľŠ 1 lim (A.21) Abstract and Applied Analysis 19 Taking the limit đ → ∞ to the inequality, we obtain the limit in (A.21). with óľŠ2 óľŠ óľŠ2 óľŠ đ (đĽđ , đŚđ ) := óľŠóľŠóľŠóľŠđŚ2đ + đźđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠóľŠ 2 đ − [(đźđĽ2đ + đŚ2đ ) (đĽ1đ đĽ2đ + đŚ1đ đŚ2đ )] . (A.22) 2 2 2 óľŠ óľŠ2 óľŠ óľŠ2 2 óľŠ óľŠ2 2 (đ¤1đ ) − óľŠóľŠóľŠóľŠđ¤2đ óľŠóľŠóľŠóľŠ = ((đĽ1đ ) − óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ ) + ((đŚ1đ ) − óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ ) An elementary computation yields that óľŠ óľŠóľŠ óľŠ 2 + 2(đĽ1đ đŚ1đ + óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ) đ (đĽđ , đŚđ ) óľŠ2 óľŠ đ 1 (đ¤đ ) óľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠ = = (đĽ1đ − 2 đźđŚ1đ ) óľŠ óľŠ2 óľŠ óľŠ2 [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ óľŠ đ 1 (đ¤đ ) óľŠóľŠóľŠđĽ1đ đĽ2đ + (đĽ1đ − Substep 1.2 (đź < 0). Since đŚ1 > 0 and đĽ1 = đźđŚ1 < 0, we have đĽ1đ đŚ1đ < 0 for sufficiently large đ. Now, from (A.24), it is easy to obtain that − óľŠ2 óľŠ + 2óľŠóľŠóľŠóľŠđĽ1đ đŚ2đ + đŚ1đ đĽ2đ óľŠóľŠóľŠóľŠ 2 đ ((đĽ2đ ) đŚ2đ ) ] đ óľŠ óľŠóľŠ óľŠ − 8đĽ1đ đŚ1đ [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ + (đĽ2đ ) đŚ2đ ] óľŠ2 đŚ1đ đŚ2đ óľŠóľŠóľŠ 2 đ óľŠ óľŠ2 óľŠ óľŠ2 [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ − ((đĽ2đ ) đŚ2đ ) ] đ 2 (đ¤đ ) . 2 óľŠóľŠ đ đ óľŠ2 óľŠ óľŠ2 óľŠóľŠđĽ1 đĽ2 + đŚ1đ đŚ2đ óľŠóľŠóľŠ [(đ¤1đ ) − óľŠóľŠóľŠđ¤2đ óľŠóľŠóľŠ ] (A.23) 2 đźđŚ1đ ) By the expressions of đ¤1đ and đ¤2đ , we compute that 2 2 2 óľŠ óľŠ2 óľŠ óľŠ2 2 óľŠ óľŠ2 2 (đ¤1đ ) − óľŠóľŠóľŠóľŠđ¤2đ óľŠóľŠóľŠóľŠ = ((đĽ1đ ) − óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ ) + ((đŚ1đ ) − óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ ) 2 2 óľŠ óľŠ2 óľŠ óľŠ2 + 2 ((đĽ1đ ) + óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ ) ((đŚ1đ ) + óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ ) đ − 8đĽ1đ đŚ1đ (đĽ2đ ) đŚ2đ . (A.24) Substep 1.1 (đź > 0). Since đŚ1 > 0 and đĽ1 = đźđŚ1 > 0, we have đĽ1đ đŚ1đ > 0 for sufficiently large đ. In addition, from (A.24), it is not difficult to obtain that 2 2 2 óľŠ óľŠ2 óľŠ óľŠ2 2 óľŠ óľŠ2 2 (đ¤1đ ) − óľŠóľŠóľŠóľŠđ¤2đ óľŠóľŠóľŠóľŠ = ((đĽ1đ ) − óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ ) + ((đŚ1đ ) − óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ ) óľŠ óľŠóľŠ óľŠ 2 + 2(đĽ1đ đŚ1đ − óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ) óľŠ2 óľŠ + 2óľŠóľŠóľŠóľŠđĽ1đ đŚ2đ − đŚ1đ đĽ2đ óľŠóľŠóľŠóľŠ (A.27) Together with (A.23) and đĽ1đ đŚ1đ < 0, it follows that đ (đĽđ , đŚđ ) 0≤ óľŠ2 óľŠ đ 1 (đ¤đ ) óľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠ 2 óľŠ óľŠóľŠ óľŠ đ (đĽ1đ − đźđŚ1đ ) [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ − (đĽ2đ ) đŚ2đ ] đ 2 (đ¤đ ) ≤ . óľŠ óľŠ2 −8đĽ1đ đŚ1đ óľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠ đ óľŠ óľŠóľŠ óľŠ ≥ 8đĽ1đ đŚ1đ [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ − (đĽ2đ ) đŚ2đ ] . (A.25) Substep 1.3 (đź = 0). Now we must have đĽ1đ > 0 or đĽ1đ < 0 for sufficiently large đ. Then, using the same arguments as in Substepss 1.1 and 1.2, we get the limit in (A.21). Let đ˘ = (đ˘1 , đ˘2 ) = limđ → ∞ đ˘đ and đŁ = (đŁ1 , đŁ2 ) = limđ → ∞ đŁđ with đ˘đ and đŁđ defined by (A.9). Step 2. To prove that đźđ˘2 + đźđ˘1 đŚĚ2 + đŁ2 + đŁ1 đŚĚ2 = 0. By the Ě2đ = −đ˘1đ expression of đ˘đ and đŁđ , it is easy to verify that (đ˘2đ )đ đ¤ đ đ đ đ Ě2 = −đŁ1 , which implies that and (đŁ2 ) đ¤ đŁ2đ đŚĚ2 = −đŁ1 . (A.29) By (A.29), we can verify that âđźđ˘2 + đźđ˘1 đŚĚ2 + đŁ2 + đŁ1 đŚĚ2 â = 0 is equivalent to âđźđ˘2 + đŁ2 â2 = (đźđ˘1 + đŁ1 )2 . From Step 1, we have that (đźđ˘1 + đŁ1 )2 + âđźđŁ2 − đ˘2 â2 = 1 + đź2 . This implies that, to prove that đźđ˘2 + đźđ˘1 đŚĚ2 + đŁ2 + đŁ1 đŚĚ2 = 0, it suffices to argue that óľŠóľŠ óľŠóľŠ2 óľŠóľŠ óľŠóľŠ2 óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđŁ2 óľŠóľŠ = 1. Together with (A.23) and đĽ1đ đŚ1đ > 0, we have that (A.28) Taking the limit đ → ∞ to this inequality, we readily obtain the limit in (A.21). đ˘2đ đŚĚ2 = −đ˘1 , đ óľŠ óľŠóľŠ óľŠ + 8đĽ1đ đŚ1đ [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ − (đĽ2đ ) đŚ2đ ] đ (đĽđ , đŚđ ) 0≤ óľŠ2 óľŠ đ 1 (đ¤đ ) óľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠ 2 óľŠ óľŠóľŠ óľŠ đ (đĽ1đ − đźđŚ1đ ) [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ + (đĽ2đ ) đŚ2đ ] đ 2 (đ¤đ ) ≤ . óľŠ óľŠ2 8đĽ1đ đŚ1đ óľŠóľŠóľŠđĽ1đ đĽ2đ + đŚ1đ đŚ2đ óľŠóľŠóľŠ đ óľŠ óľŠóľŠ óľŠ ≥ − 8đĽ1đ đŚ1đ [óľŠóľŠóľŠóľŠđĽ2đ óľŠóľŠóľŠóľŠ óľŠóľŠóľŠóľŠđŚ2đ óľŠóľŠóľŠóľŠ + (đĽ2đ ) đŚ2đ ] . 2 (A.30) 2 Ě2 â + âđŚ2đ − đŚ1đ đ¤ Ě2đ â = đ 1 (đ¤đ ), Indeed, noting that âđĽ2đ − đĽ1đ đ¤ we readily have that (A.26) óľŠóľŠ đ óľŠ2 óľŠóľŠđĽ2 − đĽ1đ đ¤ Ě2 óľŠóľŠóľŠóľŠ óľŠóľŠ óľŠóľŠ2 óľŠóľŠ óľŠóľŠ2 óľŠ + óľŠóľŠđ˘2 óľŠóľŠ + óľŠóľŠđŁ2 óľŠóľŠ = lim đ→∞ đ (đ¤đ ) 1 óľŠ2 óľŠóľŠ đ óľŠóľŠđŚ2 − đŚ1đ đ¤ Ě2đ óľŠóľŠóľŠóľŠ óľŠ = 1. đ (đ¤đ ) (A.31) 20 Abstract and Applied Analysis Now let đ1 = (đźđ˘1 + đŁ1 )/√1 + đź2 and đ2 = (đźđŁ2 − đ˘2 )/ 2 đ1 2 √1 + đź2 . Clearly, + âđ2 â = 1. Also, using the results of Steps 1 and 2 and (A.29), we can verify that đ˘1 = đŁ1 = đ đ2 đŚĚ2 + đźđ1 , √1 + đź2 đ −đźđ2 đŚĚ2 + đ1 , √1 + đź2 đ˘2 = đˇđ = diag (0, 0, . . . , 0, 1 ) − −đ2 − đźđ1 đŚĚ2 , √1 + đź2 đŁ2 = Together with (A.33), we obtain đđ = đđ đˇđ đđđ and đđ = đđ Λ đ đđđ , where 1 đ ó¸ đ đ đđŁ đ. Λđ = đź − √2 đ đ đ đ (A.32) đźđ2 − đ1 đŚĚ2 . √1 + đź2 This, along with (A.15), implied that [đ đ] ∈ đđľ đFB (0, 0). The result then follows. đ đĽĚđ2 1 1 1 ó¸ đ đđ˘ , ( đ)− Ě Ě Ě 2đź − đĽ đĽ đĽ √2 đ đ 2 đ2 đ2 đ2 đđ = đź − (A.33) 1 ó¸ đ đđŁ , √2 đ đ where óľ¨óľ¨ óľ¨óľ¨ óľŠóľŠ óľŠóľŠ óľ¨óľ¨đ˘đ1 óľ¨óľ¨ ≤ óľŠóľŠđ˘đ2 óľŠóľŠ ≤ 1, óľ¨ óľ¨ óľŠ óľŠ óľ¨óľ¨ óľ¨óľ¨ óľŠóľŠ óľŠóľŠ óľ¨óľ¨đŁđ1 óľ¨óľ¨ ≤ óľŠóľŠđŁđ2 óľŠóľŠ ≤ 1, óľ¨ óľ¨ óľŠ óľŠ óľŠ óľŠóľŠ (đ˘đ1 − đŁđ1 ) ≤ óľŠóľŠóľŠđ˘đ2 − đŁđ2 óľŠóľŠóľŠóľŠ , óľŠ óľŠ (đ˘đ1 + đŁđ1 ) ≤ óľŠóľŠóľŠóľŠđ˘đ2 + đŁđ2 óľŠóľŠóľŠóľŠ , 2 óľŠ2 óľŠ (đ˘đ1 − đŁđ1 ) + óľŠóľŠóľŠóľŠđ˘đ2 + đŁđ2 óľŠóľŠóľŠóľŠ ≤ 2, 2 óľŠ2 óľŠ (đ˘đ1 + đŁđ1 ) + óľŠóľŠóľŠóľŠđ˘đ2 − đŁđ2 óľŠóľŠóľŠóľŠ ≤ 2, đđđđ˘đ = 0, đ (đđó¸ ) đ˘đ = √2đ˘đ1 , (A.34) Rđđ ×(đđ −2) . Along with đđđ đđó¸ = (0, 0, . . . , 0, 1)đ , we have that 0 0 đˇđ = ( 0 − Λđ = ( 1 0 0 − 0 0 0 0 0 đź 0 0 ), 1 đĚ đ˘đ đđ 1 − đ˘đ1 √2 (A.37) ). 1 đĚ đŁđ đđ 1 − đŁđ1 √2 Since |đ˘đ1 | ≤ âđ˘đ2 â ≤ 1, |đŁđ1 | ≤ âđŁđ2 â ≤ 1 and |đ˘đ1 + đŁđ1 | ≤ √2, there are exactly three cases for the vectors đ˘đ and đŁđ satisfying (A.34): (1) đ˘đ1 = 1, đŁđ1 < 1; (2) đ˘đ1 < 1, đŁđ1 = 1; (3) đ˘đ1 < 1, đŁđ1 < 1. We next proceed the arguments by the three cases. Case 1 (đ˘đ1 = 1, đŁđ1 < 1). Now we have âđ˘đ â = √2. From the equality (đđó¸ )đ đ˘đ = √2 in (A.34) and âđđó¸ â = 1, we deduce that Ěđ = 0. In addition, from the last đ˘đ = √2đđó¸ , and hence đ˘đđ đ two inequalities of (A.34), âđ˘đ â2 + âđŁđ â2 ≤ 2, which together Ěđ = 0, đ˘đ1 = with âđ˘đ â2 = 2 implies đŁđ = 0. Now plugging đ˘đđđ 1, đŁđ = 0 into (A.37) yields đˇđ = 0 and Λ đ = đź. Therefore, đđ can be taken as an identity matrix. Case 2 (đ˘đ1 < 1, đŁđ1 = 1). Under this case, since (đđó¸ )đ đŁđ = √2, âđŁđ â = √2 and âđđó¸ â = 1, using the same arguments as Ěđ = 0 and đ˘đ = 0. Now plugging in Case (1) then yields đŁđđ đ Ěđ = 0, đŁđ1 = 1, đ˘đ = 0 into (A.37), đˇđ and Λ đ become the đŁđđ đ one given by (85). đđđ đŁđ = 0, đ (đđó¸ ) đŁđ = √2đŁđ1 . 1 It is not hard to verify that the matrix đź − (1/2) ( đĽĚ đ2 đ đĽĚđ2 đ 2đź−đĽĚđ2 đĽĚđ2 ) has eigenvalue 0 of multiplicity đđ − 1 and a single eigenvalue 0 1, with the corresponding eigenvectors being đđ , đĚđ = ( đŁĚđ ) ó¸ for đ = 1, . . . , đđ − 2, and đđ , respectively, where đŁĚ1 , . . . , đŁĚđđ −2 are any unit vectors that span the linear subspace {ĚđŁ ∈ Rđđ −1 | đŁĚđ đĽĚđ2 = 0}. Let đđ = [ đđ đĚ1 ⋅⋅⋅ đĚđđ −2 đđó¸ ]. Then such đđ is an đđ × đđ orthogonal matrix satisfying đź− (A.36) Ěđ , √2đ˘đ1 ), Using the equalities in (A.34) yields đ˘đđ đđ = (0, đ˘đđđ đ đĚ √ Ě đ Ě and đŁđ đđ = (0, đŁđ đđ , 2đŁđ1 ) with đđ = [ 1 ⋅⋅⋅ đĚđđ −2 ] ∈ Ěđ = đ¤đ /âđ¤đ â with đ¤đ = (đĽ2đ + đŚ2đ ) for Proof of Lemma 15. Let đ¤ đ = 1, . . . , đ. Ěđ2 = đĽĚđ2 . By (a) Since đĽđ ∈ bd+ Kđđ and đŚđ = 0, we have đ¤ Proposition 13, there exist some đ˘đ = (đ˘đ1 , đ˘đ2 ) ∈ R × Rđđ −1 and đŁđ = (đŁđ1 , đŁđ2 ) ∈ R × Rđđ −1 such that đđ = đź − 1 đ ó¸ đ đ đđ˘ đ, √2 đ đ đ đ đ đĽĚđ2 1 1 đ ( đ ) = đđ diag (0, 0, . . . , 0, 1) đđ . 2 đĽĚđ2 2đź − đĽĚđ2 đĽĚđ2 (A.35) Case 3 (đ˘đ1 < 1, đŁđ1 < 1). By the expressions of đˇđ and Λ đ , we calculate that đˇđ Λđđ + Λ đ đˇđđ 0 0 0 1 Ěđ 0 0 − đđ đ˘đ =( ). √2 1 đĚ đ Ě Ěđ 0 − đ˘đ đđ 2 (1 − đ˘đ1 ) (1 − đŁđ1 ) + đ˘đ đđ đđ đŁđ √2 (A.38) Since [đđ đđ ] ∈ đđFB (đĽđ , đŚđ ), the definition of the elements in đđFB (đĽđ , đŚđ ) and the proof of [3, Lemma 6(b)] imply that Abstract and Applied Analysis đđ đđđ +đđ đđđ ⪰ 0, and hence đˇđ Λđđ +Λ đ đˇđđ ⪰ 0. Thus, the zero Ěđ = 0, and đˇđ and Λ đ have the expression diagonals imply đ˘đđ đ of (86). (b) In view of the symmetry of đĽ and đŚ in đđ and đđ , the results readily follow by using similar arguments as in part (a). Thus, we complete the proof. Acknowledgments This work was supported by National Young Natural Science Foundation (no. 10901058) and Guangdong Natural Science Foundation (no. 9251802902000001), the Fundamental Research Funds for the Central Universities, and Project of Liaoning Innovative Research Team in University WT2010004. The author’s work is supported by National Science Council of Taiwan, Department of Mathematics, National Taiwan Normal University, Taipei, Taiwan 11677. References [1] R. T. Rockafellar, Convex Analysis, Princeton University Press, Princeton, NJ, USA, 1970. [2] D. F. Sun and J. Sun, “Strong semismoothness of the FischerBurmeister SDC and SOC complementarity functions,” Mathematical Programming, vol. 103, no. 3, pp. 575–581, 2005. [3] J.-S. Chen and P. Tseng, “An unconstrained smooth minimization reformulation of the second-order cone complementarity problem,” Mathematical Programming, vol. 104, no. 2-3, pp. 293– 327, 2005. [4] J.-S. Chen, D. F. Sun, and J. Sun, “The đđś1 property of the squared norm of the SOC Fischer-Burmeister function,” Operations Research Letters, vol. 36, no. 3, pp. 385–392, 2008. [5] J.-S. Pang and L. Q. Qi, “Nonsmooth equations: motivation and algorithms,” SIAM Journal on Optimization, vol. 3, no. 3, pp. 443–465, 1993. [6] L. Q. Qi and J. Sun, “A nonsmooth version of Newton’s method,” Mathematical Programming, vol. 58, no. 3, pp. 353–367, 1993. [7] J.-F. Bonnans and H. C. RamÄąĚrez, “Perturbation analysis of second-order cone programming problems,” Mathematical Programming, vol. 104, no. 2-3, pp. 205–227, 2005. [8] Z. X. Chan and D. F. Sun, “Constraint nondegeneracy, strong regularity, and nonsingularity in semidefinite programming,” SIAM Journal on Optimization, vol. 19, no. 1, pp. 370–396, 2008. [9] Y. Wang and L. W. Zhang, “Properties of equation reformulation of the Karush-Kuhn-Tucker condition for nonlinear second order cone optimization problems,” Mathematical Methods of Operations Research, vol. 70, no. 2, pp. 195–218, 2009. [10] C. Kanzow, I. Ferenczi, and M. Fukushima, “On the local convergence of semismooth Newton methods for linear and nonlinear second-order cone programs without strict complementarity,” SIAM Journal on Optimization, vol. 20, no. 1, pp. 297–320, 2009. [11] S.-J. Bi, S.-H. Pan, and J.-S. Chen, “Nonsingularity conditions for the Fischer-Burmeister system of nonlinear SDPs,” SIAM Journal on Optimization, vol. 21, no. 4, pp. 1392–1417, 2011. [12] J. Faraut and A. KoraĚnyi, Analysis on Symmetric Cones, Oxford University Press, New York, NY, USA, 1994. [13] M. Fukushima, Z.-Q. Luo, and P. Tseng, “Smoothing functions for second-order-cone complementarity problems,” SIAM Journal on Optimization, vol. 12, no. 2, pp. 436–460, 2002. 21 [14] S. M. Robinson, “Strongly regular generalized equations,” Mathematics of Operations Research, vol. 5, no. 1, pp. 43–62, 1980. [15] F. H. Clarke, Optimization and Nonsmooth Analysis, John Wiley & Sons, New York, NY, USA, 1983, reprinted by Society for Industrial and Applied Mathematics, Philadelphia, Pa, USA, 1990. [16] L. Q. Qi, “Convergence analysis of some algorithms for solving nonsmooth equations,” Mathematics of Operations Research, vol. 18, no. 1, pp. 227–244, 1993. [17] A. Fischer, “Solution of monotone complementarity problems with locally Lipschitzian functions,” Mathematical Programming, vol. 76, no. 3, pp. 513–532, 1997. [18] J.-S. Pang, D. Sun, and J. Sun, “Semismooth homeomorphisms and strong stability of semidefinite and Lorentz complementarity problems,” Mathematics of Operations Research, vol. 28, no. 1, pp. 39–63, 2003. [19] F. Alizadeh and D. Goldfarb, “Second-order cone programming,” Mathematical Programmingy, vol. 95, no. 1, pp. 3–51, 2003. [20] S. M. Robinson, “First order conditions for general nonlinear optimization,” SIAM Journal on Applied Mathematics, vol. 30, no. 4, pp. 597–607, 1976. [21] J. F. Bonnans and A. Shapiro, Perturbation Analysis of Optimization Problems, Springer, New York, NY, USA, 2000. [22] F. Facchinei, A. Fischer, and C. Kanzow, “Regularity properties of a semismooth reformulation of variational inequalities,” SIAM Journal on Optimization, vol. 8, no. 3, pp. 850–869, 1998. 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