• 沒有找到結果。

3 Differentiability of the function ψ

N/A
N/A
Protected

Academic year: 2022

Share "3 Differentiability of the function ψ"

Copied!
34
0
0

加載中.... (立即查看全文)

全文

(1)

Journal of Mathematical Analysis and Applications, vol. 355, pp. 195-215, 2009

A one-parametric class of merit functions for the symmetric cone complementarity problem

Shaohua Pan1

School of Mathematical Sciences South China University of Technology

Guangzhou 510640, China E-mail: shhpan@scut.edu.cn

Jein-Shan Chen 2 Department of Mathematics National Taiwan Normal University

Taipei, Taiwan 11677 E-mail: jschen@math.ntnu.edu.tw

June 23, 2008

Abstract. In this paper, we extend the one-parametric class of merit functions proposed by Kanzow and Kleinmichel [14] for the nonnegative orthant complementarity problem to the general symmetric cone complementarity problem (SCCP). We show that the class of merit functions is continuously differentiable everywhere and has a globally Lipschitz continuous gradient mapping. From this, we particularly obtain the smoothness of the Fischer-Burmeister merit function associated with symmetric cones and the Lipschitz continuity of its gradient. In addition, we also consider a regularized formulation for the class of merit functions which is actually an extension of one of the NCP function classes studied by [18] to the SCCP. By exploiting the Cartesian P -properties for a nonlinear transformation, we show that the class of regularized merit functions provides a global error bound for the solution of the SCCP, and moreover, has bounded level sets under a rather weak condition which can be satisfied by the monotone SCCP with a strictly feasible point or the SCCP with the joint Cartesian R02-property. All of these results generalize some recent important works in [4, 25, 28] under a unified framework.

1The author’s work is partially supported by the Doctoral Starting-up Foundation (B13B6050640) of GuangDong Province.

2Member of Mathematics Division, National Center for Theoretical Sciences, Taipei Office. The author’s work is partially supported by National Science Council of Taiwan.

(2)

Key words. Symmetric cone complementarity problem, merit function, Jordan algebra, smoothness, Lipschitz continuity, Cartesian P -properties.

1 Introduction

Given a Euclidean Jordan algebra A = (V, ◦, h·, ·i) where V is a finite-dimensional vector space over the real field R endowed with the inner product h·, ·i and “◦” denotes the Jordan product. Let K be a symmetric cone in V and G, F : V → V be nonlinear trans- formations assumed to be continuously differentiable throughout this paper. Consider the symmetric cone complementarity problem (SCCP) of finding ζ ∈ V such that

G(ζ) ∈ K, F (ζ) ∈ K, hG(ζ), F (ζ)i = 0. (1) The model provides a simple, natural and unified framework for various existing comple- mentarity problems such as the nonnegative orthant nonlinear complementarity problem (NCP), the second-order cone complementarity problem (SOCCP), and the semidefinite complementarity problem (SDCP). In addition, the model itself is closely related to the KKT optimality conditions for the convex symmetric cone program (CSCP):

minimize g(x)

subject to hai, xi = bi, i = 1, 2, . . . , m, x ∈ K,

(2)

where ai ∈ V, bi ∈ R for i = 1, 2, . . . , m, and g : V → R is a convex twice continu- ously differentiable function. Therefore, the SCCP has wide applications in engineering, economics, management science and other fields; see [1, 11, 20, 29] and references therein.

During the past several years, interior-point methods have been well used for solving the symmetric cone linear programming problem (SCLP), i.e., the CSCP with g being a linear function (see [7, 8, 23, 24]). However, in view of the wide applications of the SCCP, it is worthwhile to explore other solution methods for the more general CSCP and SCCP. Recently, motivated by the successful applications of the merit function approach in the solution of NCPs, SOCCPs and SDCPs (see, e.g., [4, 10, 22, 28]), some researchers started with the investigation of merit functions or complementarity functions associated with symmetric cones. For example, Liu, Zhang and Wang [21] extended a class of merit functions proposed in [18] to the following special SCCP:

ζ ∈ K, F (ζ) ∈ K, hζ, F (ζ)i = 0; (3)

Kong, Tuncel and Xiu [17] studied the extension of the implicit Lagrangian function proposed by Mangasarian and Solodov [22] to symmetric cones; and Kong, Sun and Xiu [16] proposed a regularized smoothing method for the SCCP (3) based on the natural

(3)

residual complementarity function associated with symmetric cones. Following this line, in this paper we will consider the extension of the one-parametric class of merit functions proposed by Kanzow and Kleinmichel [14] and a class of regularized functions based on it.

We define the one-parametric class of vector-valued functions φτ : V × V → V by φτ(x, y) :=¡

x2 + y2+ (τ − 2)(x ◦ y)¢1/2

− (x + y), (4)

where τ ∈ (0, 4) is an arbitrary but fixed parameter, x2 = x ◦ x, x1/2 is a vector such that (x1/2)2 = x, and x + y means the usual componentwise addition of vectors. When τ = 2, φτ reduces to the vector-valued Fischer-Burmeister function given by

φFB(x, y) := (x2+ y2)1/2− (x + y); (5) whereas as τ → 0 it will become a multiple of the vector-valued residual function

ψNR(x, y) := x − (x − y)+

where (·)+ denotes the metric projection on K. In this sense, the one-parametric class of vector-valued functions covers the two popular complementarity functions associated with the symmetric cone K. In fact, from Lemma 3.1 later, it follows that the function φτ with any τ ∈ (0, 4) is a complementarity function associated with K, that is,

φτ(x, y) = 0 ⇐⇒ x ∈ K, y ∈ K, hx, yi = 0.

Consequently, its squared norm yields a merit function associated with K ψτ(x, y) := 1

2τ(x, y)k2, (6)

where k · k is the norm induced by h·, ·i, and the SCCP can be reformulated as

minζ∈V fτ(ζ) := ψτ(G(ζ), F (ζ)). (7)

To apply the effective unconstrained optimization methods, such as the quasi-Newton method, the trust-region method and the conjugate gradient method, for solving the un- constrained minimization reformulation (7) of the SCCP, the smoothness of the merit function ψτ and the Lipschitz continuity of its gradient will play an important role. In Section 3 and Section 4, we show that the function ψτ defined by (6) is continuously differentiable everywhere and has a globally Lipschitz continuous gradient with the Lip- schitz constant being a positive multiple of 1 + τ−1. These results generalize some recent important works in [4, 25, 28] under a unified framework, as well as improve the work [21] greatly in which only the differentiability of the merit function ψFB was given.

(4)

In addition, we also consider a class of regularized functions for fτ defined as

fbτ(ζ) := ψ0(G(ζ) ◦ F (ζ)) + ψτ(G(ζ), F (ζ)), (8) where ψ0 : V → R+ is continuously differentiable and satisfies

ψ0(u) = 0 ∀u ∈ −K and ψ0(u) ≥ βk(u)+k ∀u ∈ V (9) for some constant β > 0. Using the properties of ψ0 in (9), it is not hard to verify that fbτ is a merit function for the SCCP. The class of functions will reduce to the one studied in [21] if τ = 2 and G degenerates into an identity transformation. In Section 5, we show that the class of merit functions can provide a global error bound for the solution of the SCCP under the condition that G and F have the joint uniform Cartesian P -property.

In Section 6, we establish the boundedness of the level sets of bfτ under a weaker con- dition than the one used by [21], which can be satisfied by the monotone SCCP with a strictly feasible point or the SCCP with G and F having the joint Cartesian R02-property.

Throughout this paper, I denotes an identity operator, k · k represents the norm induced by the inner product h·, ·i, and int(K) denotes the interior of the symmetric cone K. All vectors are column ones and write the column vector (xT1, . . . , xTm)T as (x1, . . . , xm), where xi is a column vector from the subspace Vi. For any x ∈ V, we denote (x)+ and (x) by the metric projection of x onto K and −K, respectively, i.e., (x)+ := argminy∈K{kx − yk}. For any symmetric matrix A, the notation A º O means that A is positive semidefinite. For a differentiable mapping F : V → V, the notation

∇F (x) denotes the transposed Jacobian operator of F at a point x. We write x = o(α) (respectively, x = O(α)) if kxk/|α| → 0 (respectively, uniformly bounded) as α → 0.

2 Preliminaries

In this section, we recall some concepts and materials of Euclidean Jordan algebras that will be used in the subsequent sections. More detailed expositions of Euclidean Jordan algebras can be found in the monograph by Faraut and Kor´anyi [9]. Besides, one can find excellent summaries in the articles [2, 12, 24, 26].

A Euclidean Jordan algebra is a triple (V, ◦, h·, ·iV), where V is a finite-dimensional inner product space over the real field R and (x, y) 7→ x ◦ y : V × V → V is a bilinear mapping satisfying the following conditions:

(i) x ◦ y = y ◦ x for all x, y ∈ V,

(ii) x ◦ (x2◦ y) = x2◦ (x ◦ y) for all x, y ∈ V, where x2 := x ◦ x, and (iii) hx ◦ y, ziV = hx, y ◦ ziV for all x, y, z ∈ V.

(5)

We call x ◦ y the Jordan product of x and y. We also assume that there is an element e ∈ V, called the unit element, such that x ◦ e = x for all x ∈ V. For x ∈ V, let ζ(x) be the degree of the minimal polynomial of x, which can be equivalently defined as

ζ(x) := min©

k : {e, x, x2, . . . , xk} are linearly dependentª .

Since ζ(x) ≤ dim(V) where dim(V) denotes the dimension of V, the rank of V is well defined by r := max{ζ(x) : x ∈ V}. In a Euclidean Jordan algebra A = (V, ◦, h·, ·iV), we define the set of squares as K := {x2 : x ∈ V}. Then, by Theorem III. 2.1 of [9], K is a symmetric cone. This means that K is a self-dual closed convex cone with nonempty interior int(K) and for any two elements x, y ∈ int(K), there exists an invertible linear transformation T : V → V such that T (K) = K and T (x) = y.

A Euclidean Jordan algebra is said to be simple if it is not the direct sum of two Euclidean Jordan algebras. By Proposition III. 4.4 of [9], each Euclidean Jordan algebra is, in a unique way, a direct sum of simple Euclidean Jordan algebras. A common simple Euclidean Jordan algebra is (Sn, ◦, h·, ·iSn), where Sn is the space of n × n real symmetric matrices with the inner product hX, Y iSn := Tr(XY ), and the Jordan product is defined by X ◦ Y := (XY + Y X)/2. Here, XY is the usual matrix multiplication of X and Y and Tr(X) is the trace of X. The associate cone K is the set of all positive semidefinite matrices. Another one is the Lorentz algebra (Rn, ◦, h·, ·iRn), where Rn is the Euclidean space of dimension n with the standard inner product hx, yiRn = xTy, and the Jordan product is defined by x ◦ y := (hx, yiRn, x1y2+ y1x2) for any x = (x1, x2), y = (y1, y2) ∈ R × Rn−1. The associate cone, called the Lorentz cone or the second-order cone, is

K :=©

x = (x1, x2) ∈ R × Rn−1: kx2k ≤ x1ª .

Recall that an element c ∈ V is said to be idempotent if c2 = c. Two idempotents c and d are said to be orthogonal if c ◦ d = 0. One says that {c1, c2, . . . , ck} is a complete system of orthogonal idempotents if

c2j = cj, cj◦ ci = 0 if j 6= i, j, i = 1, 2, . . . , k, and Pk

j=1cj = e.

A nonzero idempotent is said to be primitive if it cannot be written as the sum of two other nonzero idempotents. We call a complete system of orthogonal primitive idempotents a Jordan frame. Then, we have the following spectral decomposition theorem.

Theorem 2.1 [9, Theorem III. 1.2] Suppose that A = (V, ◦, h·, ·iV) is a Euclidean Jordan algebra and the rank of A is r. Then for any x ∈ V, there exist a Jordan frame {c1, c2, . . . , cr} and real numbers λ1(x), λ2(x), . . . , λr(x), arranged in the decreasing order λ1(x) ≥ λ2(x) ≥ · · · ≥ λr(x), such that x =Pr

j=1λj(x)cj.

(6)

The numbers λj(x) (counting multiplicities), which are uniquely determined by x, are called the eigenvalue, and we write the maximum eigenvalue and the minimum eigenvalue of x as λmax(x) and λmin(x), respectively. The trace of x, denoted as tr(x), is defined by tr(x) :=Pr

j=1λj(x); whereas the determinant of x is defined by det(x) :=Qr

j=1λj(x).

By Proposition III. 1.5 of [9], a Jordan algebra over R with a unit element e ∈ V is Euclidean if and only if the symmetric bilinear form tr(x ◦ y) is positive definite. Hence, we may define an inner product h·, ·i on V by

hx, yi := tr(x ◦ y), ∀ x, y ∈ V. (10)

Unless otherwise states, the inner product h·, ·i appearing in this paper always means the one defined by (10). By the associativity of tr(·) (see [9, Proposition II. 4.3]), the inner product h·, ·i is associative, i.e., hx, y ◦ zi = hy, x ◦ zi for all x, y, z ∈ V. Let

L(x)y := x ◦ y for every y ∈ V.

Then, the linear operator L(x) for each x ∈ V is symmetric with respect to the inner product h·, ·i in the sense that hL(x)y, zi = hy, L(x)zi for any y, z ∈ V. Let k · k be the norm on V induced by the inner product h·, ·i, namely,

kxk := p

hx, xi = ³Pr

j=1λ2j(x)

´1/2

, ∀ x ∈ V.

It is not difficult to verify that for any x, y ∈ V, there always holds that hx, yi ≤ 1

2(kxk2+ kyk2) and kx + yk2 ≤ 2(kxk2+ kyk2). (11) Let ϕ : R → R be a scalar valued function. Then, it is natural to define a vector- valued function associated with the Euclidean Jordan algebra A = (V, ◦, h·, ·i) by

ϕV(x) := ϕ(λ1(x))c1+ ϕ(λ2(x))c2+ · · · + ϕ(λr(x))cr, (12) where x ∈ V has the spectral decomposition x = Pr

j=1λj(x)cj. The function ϕV is also called the L¨owner operator [26]. When ϕ(t) is chosen as max{0, t} and min{0, t} for t ∈ R, respectively, ϕV becomes the metric projection operator onto K and −K:

(x)+ :=

Xr j=1

max{0, λj(x)}cj and (x) :=

Xr j=1

min{0, λj(x)}cj. (13)

Lemma 2.1 [26, Theorem 13] For any x = Pr

j=1λj(x)cj, let ϕV be given as in (12).

Then ϕV is (continuously) differentiable at x if and only if ϕ is (continuously) differen- tiable at each λj(x), j = 1, 2, . . . , r. The derivative of ϕV at x, for any h ∈ V, is

ϕ0V(x)h = Xr

j=1

[1](λ(x))]jjhcj, hicj+ X

1≤j<l≤r

4[ϕ[1](λ(x))]jlcj◦ (cl◦ h)

(7)

where

[1](λ(x))]ij :=



ϕ(λi(x)) − ϕ(λj(x))

λi(x) − λj(x) if λi(x) 6= λj(x) ϕ0i(x)) if λi(x) = λj(x)

, i, j = 1, 2, . . . , r.

In fact, the Jacobian ϕ0V(·) is a linear and symmetric operator, which can be written as ϕ0V(x) =

Xr j=1

ϕ0j(x))Q(cj) + 2 Xr i,j=1,i6=j

[1](λ(x))]ijL(cj)L(ci) (14)

where Q(x) := 2L2(x) −L(x2) for any x ∈ V is called the quadratic representation of V.

In the sequel, unless otherwise stated, we assume that A = (V, ◦, h·, ·i) is a simple Euclidean Jordan algebra of rank r and dim(V) = n.

An important part in the theory of Euclidean Jordan algebras is the Peirce decompo- sition. Let c be a nonzero idempotent in A. Then, by [9, Proposition III. 1.3], c satisfies 2L3(c) − 3L2(c) + L(c) = 0 and the distinct eigenvalues of the symmetric operator L(c) are 0,1

2 and 1. Let V(c, 1), V(c,12) and V(c, 0) be the three corresponding eigenspaces, i.e.,

V(c, α) :=

n

x ∈ V : L(c)x = αx o

, α = 1, 1 2, 0.

Then V is the orthogonal direct sum of V(c, 1), V(c,12) and V(c, 0). The decomposition V = V(c, 1) ⊕ V(c,1

2) ⊕ V(c, 0)

is called the Peirce decomposition of V with respect to the nonzero idempotent c.

Let {c1, c2, . . . , cr} be a Jordan frame of A. For i, j ∈ {1, . . . , r}, define the eigenspaces Vii := V(ci, 1) = Rci,

Vij := V(ci,1

2) ∩ V(cj,1

2), i 6= j.

Then, from [9, Theorem IV. 2.1], it follows that the following conclusion holds.

Theorem 2.2 The space V is the orthogonal direct sum of subspaces Vij (1 ≤ i ≤ j ≤ r), i.e., V = ⊕i≤jVij. Furthermore,

Vij ◦ Vij ⊂ Vii+ Vjj, Vij ◦ Vjk ⊂ Vik, if i 6= k,

Vij ◦ Vkl = {0}, if {i, j} ∩ {k, l} = ∅.

(8)

Let x ∈ V have the spectral decomposition x =Pr

j=1λj(x)cj. For i, j ∈ {1, 2, . . . , r}, let Cij(x) be the orthogonal projection operator onto Vij. Then,

Cij(x) = Cij(x), Cij2(x) = Cij(x), Cij(x)Ckl(x) = 0 if {i, j} 6= {k, l}, i, j, k, l = 1, . . . , r (15)

and P

1≤i≤j≤rCij(x) = I, (16)

where Cij is the adjoint (operator) of Cij. In addition, by [9, Theorem IV. 2.1], Cjj(x) = Q(cj) and Cij(x) = 4L(ci)L(cj) = 4L(cj)L(ci) = Cji(x), i, j = 1, 2, . . . , r.

Note that the original notation in [9] for orthogonal projection operator is Pij. However, to avoid confusion with another orthogonal projector Pi(cj) onto V(c, α) and orthogonal matrix P which will be used later (Sections 3–4), we adopt Cij instead.

With the orthogonal projection operators {Cij(x) : i, j = 1, 2, . . . , r}, we have the following spectral decomposition theorem for L(x) and L(x2); see [15, Chapters VI–V].

Lemma 2.2 Let x ∈ V have the spectral decomposition x = Pr

j=1λj(x)cj. Then the symmetric operator L(x) has the spectral decomposition

L(x) = Xr j=1

λj(x)Cjj(x) + X

1≤j<l≤r

1

2j(x) + λl(x)) Cjl(x)

with the spectrum σ(L(x)) consisting of all distinct numbers in {12j(x) + λl(x)) : j, l = 1, 2, . . . , r}, and L(x2) has the spectral decomposition

L(x2) = Xr

j=1

λ2j(x)Cjj(x) + X

1≤j<l≤r

1 2

¡λ2j(x) + λ2l(x)¢

Cjl(x) (17)

with the spectrum σ(L(x2)) consisting of all distinct numbers in {12¡

λ2j(x) + λ2l(x)¢ : j, l = 1, 2, . . . , r}.

Proposition 2.1 For any x ∈ V, the operator L(x2) − L2(x) is positive semidefinite.

Proof. By Lemma 2.2 and (15), we can verify that L2(x) has the spectral decomposition L2(x) =

Xr j=1

λ2j(x)Cjj(x) + X

1≤j<l≤r

1

4j(x) + λl(x))2Cjl(x). (18) This means that the operator L(x2) − L2(x) has the spectral decomposition

L(x2) − L2(x) = X

1≤j<l≤r

·1 2

¡λ2j(x) + λ2l(x)¢

1

4j(x) + λl(x))2

¸

Cjl(x).

(9)

Noting that the orthogonal projection operator is positive semidefinite on V and λ2j(x) + λ2l(x)

2 j(x) + λl(x))2

4 for all j, l = 1, 2, . . . , r,

we readily obtain the conclusion from the spectral decomposition of L(x2) − L2(x). 2

3 Differentiability of the function ψ

τ

In this section, we show that ψτ is a merit function associated with K, and moreover, it is differentiable everywhere on V × V. By the definition of Jordan product,

x2+ y2 + (τ − 2)(x ◦ y) = µ

x +τ − 2 2 y

2

+τ (4 − τ ) 4 y2

= µ

y +τ − 2 2 x

2

+τ (4 − τ )

4 x2 ∈ K (19)

for any x, y ∈ V, and consequently the function φτ in (4) is well defined. The following lemma states that φτ and ψτ is respectively a complementarity function and a merit function associated with K.

Lemma 3.1 For any x, y ∈ V, let φτ and ψτ be given by (4) and (6), respectively. Then, ψτ(x, y) = 0 ⇐⇒ φτ(x, y) = 0 ⇐⇒ x ∈ K, y ∈ K, hx, yi = 0,

Proof. The first equivalence is clear by the definition of ψτ, and we only need to prove the second equivalence. Suppose that φτ(x, y) = 0. Then,

£x2+ y2+ (τ − 2)(x ◦ y)¤1/2

= (x + y). (20)

Squaring the two sides of (20) yields that

x2 + y2+ (τ − 2)(x ◦ y) = x2 + y2+ 2(x ◦ y),

which implies x ◦ y = 0 since τ ∈ (0, 4). Substituting x ◦ y = 0 into (20), we have that x =¡

x2+ y2¢1/2

− y and y = ¡

x2+ y2¢1/2

− x.

Since x2+ y2 ∈ K, x2 ∈ K and y2 ∈ K, from [12, Proposition 8] or [19, Corollary 9] it follows that x, y ∈ K. Consequently, the necessity holds. For the other direction, suppose x, y ∈ K and x ◦ y = 0. Then, (x + y)2 = x2+ y2. This, together with x ◦ y = 0, implies

that £

x2+ y2 + (τ − 2)(x ◦ y)¤1/2

− (x + y) = 0.

(10)

Consequently, the sufficiency follows. The proof is thus completed. 2

In what follows, we concentrate on the differentiability of the merit function ψτ. For this purpose, we need the following two crucial technical lemmas.

Lemma 3.2 For any x, y ∈ V, let u(x, y) := (x2+ y2)1/2. Then, the function u(x, y) is continuously differentiable at any point (x, y) such that x2+ y2 ∈ int(K). Furthermore,

xu(x, y) = L(x)L−1(u(x, y)) and ∇yu(x, y) = L(y)L−1(u(x, y)). (21) Proof. The first part is due to Lemma 2.1. It remains to derive the formulas in (21).

From the definition of u(x, y), it follows that

u2(x, y) = x2+ y2, ∀ x, y ∈ V. (22) By the formula (14), it is easy to verify that ∇x(x2) = 2L(x). Differentiating on both sides of (22) with respect to x then yields that

2∇xu(x, y)L(u(x, y)) = 2L(x).

This implies that ∇xu(x, y) = L(x)L−1(u(x, y)) since, by u(x, y) ∈ int(K), L(u(x, y)) is positive definite on V. Similarly, we have that ∇yu(x, y) = L(y)L−1(u(x, y)). 2

To present another lemma, we first introduce some related notations. For any 0 6=

z ∈ K and z /∈ int(K), suppose that z has the spectral decomposition z =Pr

j=1λj(z)cj, where {c1, c2, . . . , cr} is a Jordan frame and λ1(z), . . . , λr(z) are the eigenvalues arranged in the decreasing order λ1(z) ≥ λ2(z) ≥ · · · ≥ λr(z) = 0. Define the index

j := min n

j | λj(z) = 0, j = 1, 2, . . . , r o

(23) and let

cJ := Pj−1

l=1 cl. (24)

Clearly, j and cJ are well-defined since 0 6= z ∈ K and z /∈ int(K). Since cJ is an idem- potent and cJ 6= 0 (otherwise z = 0), V can be decomposed as the orthogonal direct sum of the subspaces V(cJ, 1), V(cJ,12) and V(cJ, 0). In the sequel, we write P1(cJ), P1

2(cJ) and P0(cJ) as the orthogonal projection onto V(cJ, 1), V(cJ,12) and V(cJ, 0), respectively.

From [21], we know that L(z) is positive definite on V(cJ, 1) and is a one-to-one mapping from V(cJ, 1) to V(cJ, 1). This means that L(z) an inverse L−1(z) on V(cJ, 1), i.e., for any u ∈ V(cJ, 1), L−1(z)u is the unique v ∈ V(cJ, 1) such that z ◦ v = u.

Lemma 3.3 For any x, y ∈ V, let z : V × V → V be the mapping defined as z = z(x, y) := £

x2+ y2+ (τ − 2)(x ◦ y)¤1/2

. (25)

If (x, y) 6= (0, 0) such that z(x, y) /∈ int(K), then the following results hold:

(11)

(a) The vectors x, y, x + y, x + τ −22 y and y + τ −22 x belong to the subspace V(cJ, 1).

(b) For any h ∈ V such that z2(x, y)+h ∈ K, let w = w(x, y) := [z2(x, y)+h]1/2−z(x, y).

Then, P1(cJ)w = 12L−1(z(x, y))[P1(cJ)h] + o(khk).

Proof. From (19) and the definition of z, it is clear that z(x, y) ∈ K for all x, y ∈ V.

Hence, using the similar arguments as Lemma 11 of [21] yields the desired result. 2

Now by Lemmas 3.2–3.3, we prove the differentiability of the merit function ψτ. Proposition 3.1 The function ψτ defined by (6) is differentiable everywhere on V × V.

Furthermore, ∇xψτ(0, 0) = ∇yψτ(0, 0) = 0, and if (x, y) 6= (0, 0), then

xψτ(x, y) =

· L

³

x +τ − 2 2 y

´

L−1(z(x, y)) − I

¸

φτ(x, y),

yψτ(x, y) =

· L

³

y +τ − 2 2 x

´

L−1(z(x, y)) − I

¸

φτ(x, y) (26)

where z(x, y) is given by (25).

Proof. We prove the conclusion by the following three cases.

Case (1): (x, y) = (0, 0). For any u, v ∈ V, suppose that u2+ v2+ (τ − 2)(u ◦ v) has the spectrum decomposition u2+ v2 + (τ − 2)(u ◦ v) = Pr

j=1µjdj, where {d1, d2, . . . , dr} is the corresponding Jordan frame. Then, for j = 1, 2, . . . , r, we have

µj = 1 kdjk2

DPr

j=1µjdj, djE

= ­

u2+ v2+ (τ − 2)(u ◦ v), dj®

=

u + τ − 2 2 v

2

+ τ (4 − τ ) 4 v2, dj

+

u + τ − 2 2 v

2

+ τ (4 − τ ) 4 v2, e

+

= kuk2+ (τ − 2)hu, vi + kvk2

≤ (τ /2)(kuk2+ kvk2), (27)

(12)

where the second equality is by kdjk = 1, the first inequality is due to e =Pr

j=1dj and dj ∈ K for j = 1, 2, . . . , r, and the last inequality is due to (11). Therefore,

ψτ(u, v) − ψτ(0, 0) = 1 2

°°

°[u2+ v2+ (τ − 2)(u ◦ v)]1/2− (u + v)

°°

°2

= 1 2

°°

°Pr

j=1

√µj dj− (u + v)

°°

°2

°°

°Pr

j=1

õj dj

°°

°2+ ku + vk2

Xr

j=1

µjkdjk2+ 2(kuk2+ kvk2)

µ1

2τ r + 2

¶¡

kuk2+ kvk2¢ ,

where the first two inequalities are due to (11), and the last one is from (27). This shows that ψτ is differentiable at (0, 0) with ∇xψτ(0, 0) = ∇yψτ(0, 0) = 0.

Case (2): z(x, y) ∈ int(K). Since φτ(x, y) = z(x, y) − (x + y), we have from Lemma 2.1 that φτ is continuously differentiable under this case. Notice that

ψτ(x, y) = 1 2

­e, φ2τ(x, y)® ,

and hence the function ψτ is continuously differentiable. Applying the chain rule yields

xψτ(x, y) = ∇xφτ(x, y)L(φτ(x, y))e = ∇xφτ(x, y)φτ(x, y). (28) On the other hand, from (19) it follows that

φτ(x, y) =

·³

x +τ − 2 2 y

´2

+ τ (4 − τ ) 4 y2

¸1/2

− (x + y), and therefore using the formulas in (21) gives that

xφτ(x, y) = L

³

x +τ − 2 2 y

´

L−1(z(x, y)) − I.

This, together with (28), immediately yields that

xψτ(x, y) =

· L

µ

x + τ − 2 2 y

L−1(z(x, y)) − I

¸

φτ(x, y).

For symmetry of x and y in ψτ(x, y), we also have that

yψτ(x, y) =

· L

µ

y + τ − 2 2 x

L−1(z(x, y)) − I

¸

φτ(x, y).

(13)

Case (3): (x, y) 6= (0, 0) and z(x, y) /∈ int(K). For any u, v ∈ V, define ˆ

z := 2ˆx ◦ u + 2ˆy ◦ v + u2+ v2+ (τ − 2)u ◦ v with ˆx = x + τ −22 y and ˆy = y + τ −22 x. It is not difficult to verify that

z2(x, y) + ˆz = µ

(x + u) + τ − 2

2 (y + v)

2

+ τ (4 − τ )

4 (y + v)2

= z2(x + u, y + v) ∈ K.

Let

w(x, y) := ¡

z2(x, y) + ˆz¢1/2

− z(x, y).

From the definitions of ψτ and z(x, y), it then follows that ψτ(x + u, y + v) − ψτ(x, y)

= 1 2

°[z2(x, y) + ˆz]1/2− (x + u + y + v)°°2− kz(x, y) − (x + y)k2 i

= 1 2

£hˆz, ei + ku + vk2¤

− hw(x, y), x + u + y + vi + hx + y − z(x, y), u + vi

= −hw(x, y), x + yi + hx + y − z(x, y), u + vi + hˆx, ui + hˆy, vi + o(k(u, v)k). (29) By Lemma 3.3 (a), x + y ∈ V(cJ, 1). Thus, using Lemma 3.3 (b), we have that

hw(x, y), x + yi = hP1(cJ)w(x, y), x + yi

=

¿1

2L−1(z(x, y))[P1(cJz] + o(kˆzk), x + y À

= 1

2

­P1(cJz, L−1(z(x, y))[x + y]®

+ o(kˆzk)

= D

P1(cJ) [ˆx ◦ u + ˆy ◦ v] , L−1(z(x, y))[x + y]

E

+ o(k(u, v)k)

= D

ˆ

x ◦ u + ˆy ◦ v, P1(cJ)[L−1(z(x, y))(x + y)]

E

+ o(k(u, v)k)

= D

ˆ

x ◦ u + ˆy ◦ v, L−1(z(x, y))(x + y) E

+ o(k(u, v)k)

= ­£

L−1(z(x, y))(x + y)¤

◦ ˆx, u® +­£

L−1(z(x, y))(x + y)¤

◦ ˆy, v®

+ o(k(u, v)k) (30)

where the first equality is since V = V(cJ, 1) ⊕ V(cJ,12) ⊕ V(cJ, 0), the fifth one is due to P1(cJ) = P1(cJ), and the sixth is from the fact that L−1(z(x, y))(x + y) ∈ V(cJ, 1).

Combining (29) with (30), we obtain that ψτ(x + u, y + v) − ψτ(x, y)

= D

ˆ

x + x + y − z(x, y) −£

L−1(z(x, y))(x + y)¤

◦ ˆx, u E

+ D

ˆ

y + x + y − z(x, y) −£

L−1(z(x, y))(x + y)¤

◦ ˆy, v E

+ o(k(u, v)k).

(14)

This implies that the function ψτ is differentiable at (x, y), and furthermore,

xψτ(x, y) = ˆx + x + y − z(x, y) −£

L−1(z(x, y))(x + y)¤

◦ ˆx,

yψτ(x, y) = ˆy + x + y − z(x, y) −£

L−1(z(x, y))(x + y)¤

◦ ˆy.

Notice that ˆ

x + x + y − z(x, y) −£

L−1(z(x, y))(x + y)¤

◦ ˆx

= ˆx − φτ(x, y) −£

L−1(z(x, y))(x + y)¤

µ

x +τ − 2 2 y

= x + τ − 2

2 y − φτ(x, y) − L µ

x +τ − 2 2 y

¶ h

L−1(z(x, y))(x + y) i

= L µ

x + τ − 2 2 y

L−1(z(x, y))[z(x, y) − x − y] − φτ(x, y)

=

· L

³

x + τ − 2 2 y

´

L−1(z(x, y)) − I

¸

φτ(x, y),

where the third equality is due to L−1(z(x, y))z(x, y) = e and the fact that x + τ − 2

2 y = L µ

x + τ − 2 2 y

e = L

µ

x +τ − 2 2 y

L−1(z(x, y))z(x, y).

Therefore,

xψτ(x, y) =

· L

³

x + τ − 2 2 y

´

L−1(z(x, y)) − I

¸

φτ(x, y).

Similarly, we also have that

yψτ(x, y) =

· L

µ

y + τ − 2 2 x

L−1(z(x, y)) − I

¸

φτ(x, y).

This shows that the conclusion holds under this case. The proof is thus completed. 2 It should be pointed out that the formula (26) is well-defined even if z(x, y) /∈ int(K) since in this case φτ(x, y) ∈ V(cJ, 1) by Lemma 3.3 (a). When A is specified as the Lorentz algebra (Rn, ◦, h·, ·iRn), the formula reduces to the one of [3, Proposition 3.2];

whereas when A is specified as (Sn, ◦, h·, ·iSn) and τ = 2, the formula is same as the one in [28, Lemma 6.3 (b)] by noting that z(x, y) = (x2+ y2)1/2 and

xψτ(x, y) = L(x)L−1(z(x, y))φFB(x, y) − φFB(x, y)

= x ◦ [L−1(z(x, y))φFB(x, y)] − L(z(x, y))L−1(z(x, y))φFB(x, y)

= x ◦ [L−1(z(x, y))φFB(x, y)] − z(x, y) ◦ [L−1(z(x, y))φFB(x, y)]

= [L−1(z(x, y))φFB(x, y)] ◦ (x − z(x, y)).

(15)

Thus, the formula (26) provides a unified framework for the SOCCP and the SDCP cases.

From Proposition 3.1, we readily obtain the following properties of ∇ψτ, which have been given in the setting of NCP [14] and the SOCCP [3], respectively.

Proposition 3.2 Let ψτ be given as in (6). Then, for any (x, y) ∈ V × V, we have (a) hx, ∇xψτ(x, y)i + hy, ∇yψτ(x, y)i = kφτ(x, y)k2.

(b) ∇ψτ(x, y) = 0 if and only if x ∈ K, y ∈ K, hx, yi = 0.

Proof. (a) If (x, y) = (0, 0), the result is clear. Otherwise, from (26) it follows that hx, ∇xψτ(x, y)i + hy, ∇yψτ(x, y)i

=

¿ x,

µ

x +τ − 2 2 y

◦ [L−1(z(x, y))φτ(x, y)]

À

− hx, φτ(x, y)i +

¿ y,

µ

y +τ − 2 2 x

◦ [L−1(z(x, y))φτ(x, y)]

À

− hy, φτ(x, y)i

=

¿ x ◦

µ

x +τ − 2 2 y

, L−1(z(x, y))φτ(x, y) À

− hx, φτ(x, y)i +

¿ y ◦

µ

y +τ − 2 2 x

, L−1(z(x, y))φτ(x, y) À

− hy, φτ(x, y)i

= hz2(x, y), L−1(z(x, y))φτ(x, y)i − hx + y, φτ(x, y)i

= hz(x, y), φτ(x, y)i − hx + y, φτ(x, y)i = kφτ(x, y)k2,

where the next to last equality is by z2 = L(z)z and the symmetry of L(z).

(b) The proof is direct by part (a), Lemma 3.1 and Proposition 3.1. 2

4 Lipschitz continuity of ∇ψ

τ

In this section, we investigate the continuity of the gradients ∇xψτ(x, y) and ∇yψτ(x, y).

To this end, for any ² > 0, we define the mapping z²: V × V → V by z² = z²(x, y) := ¡

x2+ y2+ (τ − 2)(x ◦ y) + ²e¢1/2

. (31)

From (19), clearly, z²(x, y) ∈ int(K) for any x, y ∈ V, and hence the operator L(z²(x, y)) is positive definite on V. Since the spectral function induced by ϕ(t) =√

t (t ≥ 0) is con- tinuous by Lemma 2.1, it follows that z²(x, y) → z(x, y) as ² → 0+ for any (x, y) ∈ V × V, where z(x, y) is given by (25). This means that L(z²(x, y)) → L(z(x, y)) as ² → 0+.

(16)

In what follows, we prove that the gradients ∇xψτ(x, y) and ∇yψτ(x, y) are Lipschitz continuous by arguing the Lipschitz continuity of z²(x, y) and the mapping

H²(x, y) := L

³

x +τ − 2 2 y

´

L−1(z²(x, y))(x + y). (32) To establish the Lipschitz continuity of z²(x, y), we need the following crucial lemma.

Lemma 4.1 For any (x, y) ∈ V × V and ² > 0, let z²(x, y) be defined as in (31). Then the function z²(x, y) is continuously differentiable everywhere with

xz²(x, y) = L

³

x +τ − 2 2 y

´

L−1(z²(x, y)),

yz²(x, y) = L

³

y + τ − 2 2 x

´

L−1(z²(x, y)). (33) Furthermore, there exists a constant C > 0, independent of x, y and ², τ , such that

k∇xz²(x, y)k ≤ C and k∇yz²(x, y)k ≤ C.

Proof. The first part follows from Lemma 3.2 and the following fact that z²(x, y) =

·³

x +τ − 2 2 y

´2

+τ (4 − τ )

4 y2+ ²e

¸1/2

=

·³

y +τ − 2 2 x´2

+τ (4 − τ )

4 x2+ ²e

¸1/2

. (34)

We next prove that the operator ∇xz²(x, y) is bounded for any x, y ∈ V and ² > 0. Let {u1, u2, . . . , un} be an orthonormal basis of V. For any x, y ∈ V, let L(z2), L(x + τ −22 y), L(z²) and L((x + τ −22 y)2) be the corresponding matrix representation of the operators L(z2), L(x + τ −22 y), L(z²) and L((x + τ −22 y)2) with respect to the basis {u1, u2, . . . , un}.

Then, by the formula (33), it suffices to prove that the matrix L(x + τ −22 y)L−1(z²) is bounded for any x, y ∈ V and ² > 0. The verifications are given as below.

Suppose that z = z(x, y) has the spectral decomposition z = Pr

j=1λj(z)cj, where λ1(z) ≥ λ2(z) ≥ · · · ≥ λr(z) ≥ 0 are the eigenvalue of z and {c1, c2, . . . , cr} is the corresponding Jordan frame. From Lemma 2.2, L(z) has the spectral decomposition

L(z) = Xr

j=1

λj(z)Cjj(z) + X

1≤j<l≤r

1

2j(z) + λl(z)) Cjl(z) (35) with the spectrum σ(L(z)) consisting of all distinct numbers in {12j(z) + λl(z)) : j, l = 1, 2, . . . , r}, and L(z2) has the spectral decomposition

L(z2) = Xr

j=1

λ2j(z)Cjj(z) + X

1≤j<l≤r

1 2

¡λ2j(z) + λ2l(z)¢

Cjl(z) (36)

(17)

with σ(L(z2)) consisting of all distinct numbers in {12¡

λ2j(z) + λ2l(z)¢

: j, l = 1, 2, . . . , r}.

By the definition of z²(x, y), it is easy to verify that z² = Pr

j=1

q

λ2j(z) + ² cj, and consequently the symmetric operator L(z²) has the spectral decomposition

L(z²) = Xr

j=1

q

λ2j(z) + ² Cjj(z) + X

1≤j<l≤r

1 2

³q

λ2j(z) + ² + q

λ2l(z) + ²

´

Cjl(z) (37)

with the spectrum σ(L(z²)) consisting of all distinct numbers in

½1 2

µq

λ2j(z) + ² + q

λ2l(z) + ²

: j, l = 1, 2, . . . , r

¾ .

We first prove that the matrix L(x + τ −22 y) (L(z2) + ²I)−1/2 is bounded for any x, y ∈ V and ² > 0. For this purpose, let P be an n × n orthogonal matrix such that

P L(z2)PT = diag¡

λ1(L(z2)), λ2(L(z2)), · · · , λn(L(z2))¢

(38) where λ1(L(z2)) ≥ λ2(L(z2)) · · · ≥ λn(L(z2)) ≥ 0 are the eigenvalues of L(z2). Then, it is not hard to verify that for any ² > 0,

P ¡

L(z2) + ²I¢−1/2

PT = diag Ã

p 1

λ1(L(z2)) + ², · · · , 1

pλn(L(z2)) + ²

! .

Write eU := P L¡

x +τ −22 y¢

PT. We can compute that L

³

x +τ − 2 2 y

´³

L(z2) + ²I

´−1/2

= PTUdiage Ã

p 1

λ1(L(z2)) + ², · · · , 1

pλn(L(z2)) + ²

! P

= PT

"

Ueik

pλk(L(z2)) + ²

#

1≤i≤n 1≤k≤n

P. (39)

Since L(z2) = L¡

(x + τ −22 y)2¢ + L³

τ (4−τ ) 4 y2´

and L(y2) is positive semidefinite, we get

L(z2) − L µ

(x + τ − 2 2 y)2

º O,

In addition, by Proposition 2.1 L[(x + τ −22 y)2] − L(x + τ −22 y)L(x + τ −22 y) is positive semidefinite, and hence we have that

L õ

x + τ − 2 2 y

2!

− L µ

x + τ − 2 2 y

L

µ

x +τ − 2 2 y

º O.

參考文獻

相關文件

Tseng, Growth behavior of a class of merit functions for the nonlinear comple- mentarity problem, Journal of Optimization Theory and Applications, vol. Fukushima, A new

Then, we tested the influence of θ for the rate of convergence of Algorithm 4.1, by using this algorithm with α = 15 and four different θ to solve a test ex- ample generated as

Numerical results are reported for some convex second-order cone programs (SOCPs) by solving the unconstrained minimization reformulation of the KKT optimality conditions,

Chen, The semismooth-related properties of a merit function and a descent method for the nonlinear complementarity problem, Journal of Global Optimization, vol.. Soares, A new

Particularly, combining the numerical results of the two papers, we may obtain such a conclusion that the merit function method based on ϕ p has a better a global convergence and

Then, it is easy to see that there are 9 problems for which the iterative numbers of the algorithm using ψ α,θ,p in the case of θ = 1 and p = 3 are less than the one of the

Numerical results are reported for some convex second-order cone programs (SOCPs) by solving the unconstrained minimization reformulation of the KKT optimality conditions,

11 (1998) 227–251] for the nonnegative orthant complementarity problem to the general symmet- ric cone complementarity problem (SCCP). We show that the class of merit functions