• 沒有找到結果。

Alternative proofs for some results of vector-valued functions associated with second-order cone

N/A
N/A
Protected

Academic year: 2022

Share "Alternative proofs for some results of vector-valued functions associated with second-order cone"

Copied!
31
0
0

加載中.... (立即查看全文)

全文

(1)

Journal of Nonlinear Analysis and Convex Analysis, vol. 6, pp. 297-325, 2005

Alternative proofs for some results of vector-valued functions associated with second-order cone

Jein-Shan Chen 1 Department of Mathematics National Taiwan Normal University

Taipei, Taiwan 11677

September 14, 2004 (Revised, February 22, 2005)

Abstract. Let Kn be the Lorentz/second-order cone in IRn. For any function f from IR to IR, one can define a corresponding vector-valued function fsoc(x) on IRn by applying f to the spectral values of the spectral decomposition of x ∈ IRn with respect to Kn. It was shown by J.-S. Chen, X. Chen and P. Tseng that this vector-valued function inher- its from f the properties of continuity, Lipschitz continuity, directional differentiability, Fr´echet differentiability, continuous differentiability, as well as semismoothness. It was also proved by D. Sun and J. Sun that the vector-valued Fischer-Burmeister function as- sociated with second-order cone is strongly semismooth. All proofs for the above results are based on a special relation between the vector-valued function and the matrix-valued function over symmetric matrices. In this paper, we provide a straightforward and in- tuitive way to prove all the above results by using the simple structure of second-order cone and spectral decomposition.

Key words. Second-order cone, vector-valued function, semismooth function, comple- mentarity, spectral decomposition.

AMS subject classifications. 26A27, 26B05, 26B35, 49J52, 90C33, 65K05

1 Introduction

The second-order cone (SOC) in IRn, also called the Lorentz cone, is defined to be Kn:= {(x1, x2) ∈ IR × IRn−1| kx2k ≤ x1},

1E-mail: jschen@math.ntnu.edu.tw, TEL: 886-2-29320206, FAX: 886-2-29332342.

(2)

where k · k denotes the Euclidean norm. If n = 1, K1 is the set of nonnegative reals IR+. Recently, there have been much study on second-order cone in optimization, partic- ularly in the context of applications and solution methods for second-order cone program (SOCP) [1, 2, 9, 12, 14, 18, 22]. For any x = (x1, x2) ∈ IR × IRn−1, we can decompose x as

x = λ1u(1)+ λ2u(2), (1)

where λ1, λ2 and u(1), u(2) are the spectral values and the associated spectral vectors of x, with respect to Kn, given by

λi = x1+ (−1)ikx2k, (2)

u(i) =

1 2

³1, (−1)i x2 kx2k

´, if x2 6= 0,

1 2

³1, (−1)iw´, if x2 = 0, (3)

for i = 1, 2, with w being any vector in IRn−1 satisfying kwk = 1. If x2 6= 0, the decomposition (1) is unique. With this spectral decomposition, for any function f : IR → IR, the following vector-valued function associated with Kn (n ≥ 1) was considered (see [9]):

fsoc(x) = f (λ1)u(1)+ f (λ2)u(2) ∀x = (x1, x2) ∈ IR × IRn−1. (4) If f is defined only on a subset of IR, then fsoc is defined on the corresponding subset of IRn. The definition (4) is unambiguous whether x2 6= 0 or x2 = 0. The above definition (4) is analogous to one associated with the semidefinite cone Sn, see [20, 21].

The study of this function is motivated by second-order cone complementarity prob- lem (SOCCP), see [3, 4] and references therein. In fact, in the paper [4], it studied the continuity and differentiability properties of the vector-valued function fsoc. In partic- ular, it showed that the properties of continuity, strict continuity, Lipschitz continuity, directional differentiability, differentiability, continuous differentiability, and (ρ-order) semismoothness are each inherited by fsoc from f . These results parallel those obtained recently in [5] for matrix-valued functions and are useful in the design and analysis of smoothing and nonsmooth methods for solving SOCP and SOCCP. The proofs are based on an elegant relation between the vector-valued function fsoc and its matrix-valued coun- terpart (see Lemma 4.1 of [4]). This relation enables applying the results from [5] for matrix-valued functions to the vector-valued function fsoc. In this paper, we study an intuitive way to prove all the aforementioned results without using the relation as will be seen in Sec. 3.

A popular approach to solving SOCCP is to reformulate it as an unconstrained min- imization problem. Specifically, it is to find a smooth (continuously differentiable) func- tion ψ : IRn× IRn→ IR+ such that

ψ(x, y) = 0 ⇐⇒ hx, yi = 0, x ∈ Kn, y ∈ Kn, (5)

(3)

which yields that the SOCCP can be expressed as an unconstrained smooth (continuously differentiable) minimization problem: min

ζ∈IRnf (ζ) := ψ(F (ζ), G(ζ)), for some F and G. For detailed reformulation, please refer to [3]. Such a ψ is usually called a merit function. A popular choice of ψ is

ψ(x, y) = 1

2kφ(x, y)k2 x, y ∈ IRn, (6)

where

φ(x, y) := (x2+ y2)1/2− x − y. (7) Here (·)2 and (·)1/2 are well-defined via the Jordan product as will be explained in Sec.

2. The function φ is called Fischer-Burmeister function. It is the natural extension of Fischer-Burmeister function over IRn to Kn. D. Sun and J. Sun proved that φ is strongly semismooth in [19], while ψ was proved smooth (continuously differentiable) everywhere by J.-S. Chen and P. Tseng in [3]. In this paper, we also provide an alternative proof for property of strong semismoothness of φ in Sec. 4.

In what follows, for any differentiable (in the Fr´echet sense) mapping F : IRn→ IRm, we denote its Jacobian(not transposed) at x ∈ IRn by ∇F (x) ∈ IRm×n, i.e., (F (x + u) − F (x) − ∇F (x)u)/kuk → 0 as u → 0. “ := ” means “define”. We write z = O(α) (respectively, z = o(α)), with α ∈ IR and z ∈ IRn, to mean kzk/|α| is uniformly bounded (respectively, tends to zero) as α → 0.

2 Basic Concepts and Known Results

In this section, we review some basic materials regarding vector-valued functions. These contain continuity, (local) Lipschitz continuity, directional differentiability, differentiabil- ity, continuous differentiability, as well as (ρ-order) semismoothness. We also recall some known results for vector-valued functions for which we will provide alternative proofs later.

Let the mapping F : IRn → IRm. Then F is continuous at x ∈ IRn if F (y) → F (x) as y → x; and F is continuous if F is continuous at every x ∈ IRn. We say F is strictly continuous (also called ‘locally Lipschitz continuous’) at x ∈ IRn if there exist scalars κ > 0 and δ > 0 such that

kF (y) − F (z)k ≤ κky − zk ∀y, z ∈ IRn with ky − xk ≤ δ, kz − xk ≤ δ;

and F is strictly continuous if F is strictly continuous at every x ∈ IRn. We say F is directionally differentiable at x ∈ IRn if

F0(x; h) := lim

t→0+

F (x + th) − F (x)

t exists ∀h ∈ IRn;

(4)

and F is directionally differentiable if F is directionally differentiable at every x ∈ IRn. F is differentiable (in the Fr´echet sense) at x ∈ IRn if there exists a linear mapping

∇F (x) : IRn → IRm such that

F (x + h) − F (x) − ∇F (x)h = o(khk).

If F is differentiable at every x ∈ IRn and ∇F is continuous, then F is continuously differentiable. We notice that, in the above expression about strict continuity of F , if δ can be taken to be ∞, then F is called Lipschitz continuous with Lipschitz constant κ.

It is well-known that if F is strictly continuous, then F is almost everywhere differ- entiable by Rademacher’s Theorem–see [6] and [17, Sec. 9J]. In this case, the generalized Jacobian ∂F (x) of F at x (in the Clarke sense) can be defined as the convex hull of the generalized Jacobian ∂BF (x), where

BF (x) :=

½

xlimj→x∇F (xj)|F is differentiable at xj ∈ IRn

¾

.

The notation ∂B is adopted from [15]. In [17, Chap. 9], the case of m = 1 is considered and the notations “ ¯∇” and “ ¯∂” are used instead of, respectively, “∂B” and “∂”. Assume F : IRn → IRm is strictly continuous, then F is said to be semismooth at x if F is directionally differentiable at x and, for any V ∈ ∂F (x + h), we have

F (x + h) − F (x) − V h = o(khk).

Moreover, F is called ρ-order semismooth at x (0 < ρ < ∞) if F is semismooth at x and, for any V ∈ ∂F (x + h), we have

F (x + h) − F (x) − V h = O(khk1+ρ).

The following lemma, proven by D. Sun and J. Sun [20, Thm. 3.6] using the defini- tion of generalized Jacobian, enables one to study the semismooth property of fsoc by examining only those points x ∈ IRnwhere fsoc is differentiable and thus work only with the Jacobian of fsoc, rather than the generalized Jacobian. It is a very useful working lemma for verifying semismoothness property.

Lemma 2.1 Suppose F : IRn→ IRn is strictly continuous and directionally differentiable in a neighborhood of x ∈ IRn. Then, for any 0 < ρ < ∞, the following two statements are equivalent:

(a) For any v ∈ ∂F (x + h) and h → 0,

F (x + h) − F (x) − vh = o(khk) (respectively, O(khk)1+ρ).

(5)

(b) For any h → 0 such that F is differentiable at x + h,

F (x + h) − F (x) − ∇F (x + h)h = o(khk) (respectively, O(khk)1+ρ).

We say F is semismooth (respectively, ρ-order semismooth) if F is semismooth (re- spectively, ρ-order semismooth) at every x ∈ IRn. We say F is strongly semismooth if it is 1-order semismooth. Convex functions and piecewise continuously differentiable functions are examples of semismooth functions. The composition of two (respectively, ρ-order) semismooth functions is also a (respectively, ρ-order) semismooth function. The prop- erty of semismoothness, as introduced by Mifflin [13] for functionals and scalar-valued functions and further extended by L. Qi and J. Sun [16] for vector-valued functions, is of particular interest due to the key role it plays in the superlinear convergence analysis of certain generalized Newton methods [10, 11, 15, 16, 23]. For extensive discussions of semismooth functions, see [8, 13, 16].

For any x = (x1, x2) ∈ IR × IRn−1 and y = (y1, y2) ∈ IR × IRn−1, we define their Jordan product as

x ◦ y =³xTy, y1x2+ x1y2´. (8) We write x2 to mean x ◦ x and write x + y to mean the usual componentwise addition of vectors. Then, ◦, +, together with e = (1, 0, . . . , 0) ∈ IRn, give rise to a Jordan algebra associated with Kn[7, Chap. II]. If x ∈ Kn, then there exists a unique vector in Kn, which we denote by x1/2, such that (x1/2)2 = x1/2◦ x1/2 = x. For any x = (x1, x2) ∈ IR × IRn−1, we also define the symmetric matrix

Lx =

"

x1 xT2 x2 x1I

#

, (9)

viewed as a linear mapping from IRn to IRn×n. The matrix Lx has various interesting properties that were studied in [9]. Especially, we have Lx· y = x ◦ y for any x, y ∈ IRn. Now, we summarize the results shown in [4] for which we will provide alternative proofs that are straightforward and intuitive in the subsequent sections.

Proposition 2.2 For any f : IR → IR, the following results hold:

(a) fsoc is continuous at an x ∈ IRn with eigenvalues λ1, λ2 if and only if f is continuous at λ1, λ2.

(b) fsoc is directionally differentiable at an x ∈ IRn with eigenvalues λ1, λ2 if and only if f is directionally differentiable at λ1, λ2

(c) fsoc is differentiable at an x ∈ IRn with eigenvalues λ1, λ2 if and only if f is differ- entiable at λ1, λ2.

(6)

(d) fsoc is continuously differentiable at an x ∈ IRn with eigenvalues λ1, λ2 if and only if f is continuously differentiable at λ1, λ2.

(e) fsoc is strictly continuous at an x ∈ IRn with eigenvalues λ1, λ2 if and only if f is strictly continuous at λ1, λ2.

(f) fsoc is Lipschitz continuous (with respect to k · k) with constant κ if and only if f is Lipschitz continuous with constant κ.

(g) fsoc is semismooth if and only if f is semismooth.

Proposition 2.3 The vector-valued Fischer-Burmeister function associated with second- order cone defined as (7) is strongly semismooth.

3 Alternative Proofs of Continuity and Differentia- bility

In this section, we present alternative proofs for Prop. 2.2 of Sec. 2 which is one of the main purposes of this paper. Unlike the existed proofs which employed an elegant lemma ([4, Lem. 4.1]), our arguments come from an intuitive way only using the simple structure of second-order cone and basic definitions. We need some technical lemmas before starting the alternative proofs.

Lemma 3.1 Let λ1 ≤ λ2 be the spectral values of x ∈ IRn and m1 ≤ m2 be the spectral values of y ∈ IRn. Then we have

1− m1|2+ |λ2− m2|2 ≤ 2 kx − yk2 , (10) and hence, |λi− mi| ≤√

2 kx − yk , ∀i = 1, 2.

Proof. The proof follows from a direct computation. 2

Lemma 3.2 Let x = (x1, x2) ∈ IR × IRn−1 and y = (y1, y2) ∈ IR × IRn−1. (a) If x2 6= 0, y2 6= 0, then we have

ku(i)− v(i)k ≤ 1

kx2kkx − yk , ∀i = 1, 2 , (11) where u(i), v(i) are the unique spectral vectors of x and y, respectively.

(7)

(b) If either x2 = 0 or y2 = 0, then we can choose u(i), v(i) such that the left hand side of inequality (11) is zero.

Proof. (a) From the spectral factorization (1)-(3), we know that u(i) = 1

2

µ

1 , (−1)i x2 kx2k

, v(i) = 1 2

µ

1 , (−1)i y2 ky2k

,

where ui), v(i) are unique. Thus, we have u(i)− v(i) = 12

µ

0 , (−1)i(kxx2

2k kyy2

2k)

. Then

ku(i)− v(i)k = 1 2

°°

°°

°

x2

kx2k y2

ky2k

°°

°°

°= 1 2

°°

°°

°

x2− y2

kx2k +(ky2k − kx2k)y2

kx2k · ky2k

°°

°°

°

1 2

µ 1

kx2kkx2− y2k + 1 kx2k

¯¯

¯¯ky2k − kx2k

¯¯

¯¯

1 2

µ 1

kx2kkx2− y2k + 1

kx2kkx2− y2k

1

kx2kkx − yk, where the first inequality follows from the triangle inequality.

(b) We can choose the same spectral vectors for x and y by the spectral factorization (1)-(3) since either x2 = 0 or y2 = 0. Then, it is obvious. 2

Lemma 3.3 For any w 6= 0 ∈ IRn, we have ∇w

µ w kwk

= 1

kwk

µ

I − wwT kwk2

.

Proof. The verification is routine, so we omit it. 2

Now, we are ready to present our alternative proofs for Prop. 2.2. As mentioned, all of our proofs are from intuitive definitions as well as the structure of second-order cone.

Some portion of the proofs are similar to the original ones, we omit them when there is the case.

Proof. (a) “⇐” Suppose f is continuous at λ1, λ2. For any fixed x ∈ IRnand y → x, let the spectral factorizations of x, y be x = λ1u(1)+ λ2u(2) and y = m1v(1)+ m2v(2). Then, we discuss two cases.

Case (i): If x2 6= 0, then we have

fsoc(y) − fsoc(x) (12)

= f (m1)

µ

v(1)− u(1)

+

µ

f (m1) − f (λ1)

u(1)+ f (m2)

µ

v(2)− u(2)

+

µ

f (m2) − f (λ2)

u(2). Since f is continuous at λ1, λ2, and from Lemma 3.1, |mi− λi| ≤√

2 ky − xk, we obtain f (mi) −→ f (λi) as y → x. Also by Lemma 3.2, we know that kv(i)− u(i)k −→ 0 as y →

(8)

x. Thus, equation (12) yields fsoc(y) −→ fsoc(x) as y → x, since both f (mi) and ku(i)k are bounded. Hence, fsoc is continuous at x ∈ IRn .

Case (ii): If x2 = 0, no matter y2 is zero or not, we can arrange that x, y have the same spectral vectors. Thus, fsoc(y) − fsoc(x) =

µ

f (m1) − f (λ1)

u(1)+

µ

f (m2) − f (λ2)

u(2) . Then, fsoc is continuous at x ∈ IRn by similar arguments.

“⇒” The proof for this direction is straightforward and similar to the arguments in [4, Prop. 5.2]. 2

Proof. (b) “⇐” Suppose f is directionally differentiable at λ1, λ2. Fix any x = (x1, x2) ∈ IR × IRn−1, then we discuss two cases as below.

Case (i): If x2 6= 0, then we have fsoc(x) = f (λ1)u(1)+f (λ2)u(2)where λi = x1+(−1)ikx2k and u(i) = 1

2

µ

1 , (−1)i x2 kx2k

for all i = 1, 2. From Lemma 3.3, we know that u(i) is Fr´echet-differentiable with respect to x, with

xu(i) = (−1)i 2kx2k

0 0

0 I − x2xT2 kx2k2

, ∀i = 1, 2. (13)

Also by the expression of λi, we know that λi is Fr´echet-differentiable with respect to x, with

xλi =

µ

1 , (−1)i x2 kx2k

= 2u(i) , ∀i = 1, 2. (14) Since f is directionally differentiable at λ1, λ2, then the chain rule and product rule for directional differentiation give

(fsoc)0(x; h)

= f (λ1)∇xu(1)h + u(1)f01; h)(∇xλ1)T + f (λ2)∇xu(2)h + u(2)f02; h)(∇xλ2)T

= f (λ2) − f (λ1) 2kx2k

0 0

0 I − x2xT2 kx2k2

h + 2f01; h)u(1)(u(1))T + 2f02; h)u(2)(u(2))T

= f (λ2) − f (λ1) λ2− λ1

0 0

0 I − x2xT2 kx2k2

h + 2f01; h)u(1)(u(1))T + 2f02; h)u(2)(u(2))T,

where the second equality uses equations (13) and (14), and the last equality employs the fact that λ2− λ1 = 2kx2k. Notice that we can obtain u(i)(u(i))T as follows by direct computation :

u(i)(u(i))T = 1 4

1 (−1)i xT2 kx2k (−1)i x2

kx2k

x2xT2 kx2k2

, ∀i = 1, 2.

(9)

Now let

˜a = f (λ2) − f (λ1)

λ2− λ1 h , ˜b = f02; h) + f01; h)

2 , ˜c = f02; h) − f01; h)

2 . (15)

Then, we can rewrite the previous expression of (fsoc)0(x; h) as

(fsoc)0(x; h) = ˜a

0 0

0 I − x2xT2 kx2k2

+

˜b ˜cxT2 kx2k

˜cx2

kx2k

˜bx2xT2 kx2k2

=

˜b ˜cxT2 kx2k

˜cx2

kx2k ˜aI + (˜b − ˜a)kxx2xT2

2k2

.

(16) This enables that fsoc is directionally differentiable at x when x2 6= 0 with (fsoc)0(x; h) being in form of (16).

Case (ii): If x2 = 0, we compute the directional derivative (fsoc)0(x; h) at x for any direction h by definition. Let h = (h1, h2) ∈ IR × IRn−1. We have two subcases. First, consider the subcase of h2 6= 0. From the spectral factorization, we can choose u(i) =

1 2

µ

1 , (−1)i hkh22k

for all i = 1, 2, such that

( fsoc(x + th) = f (λ + 4λ1)u(1)+ f (λ + 4λ2)u(2) fsoc(x) = f (λ)u(1)+ f (λ)u(2),

where λ = x1 and 4λi = t

µ

h1+ (−1)ikh2k

for all i = 1, 2. Thus, we obtain

fsoc(x + th) − fsoc(x) =

µ

f (λ + 4λ1) − f (λ)

u(1)+

µ

f (λ + 4λ2) − f (λ)

u(2). The fact that

t→0lim+

f (λ + 4λ1) − f (λ)

t = lim

t→0+

f (λ + t(h1− kh2k)) − f (λ)

t = f0(λ; h1− kh2k) , and

t→0lim+

f (λ + 4λ2) − f (λ)

t = lim

t→0+

f (λ + t(h1+ kh2k)) − f (λ)

t = f0(λ; h1+ kh2k), yields

t→0lim+

fsoc(x + th) − fsoc(x)

t = lim

t→0+

f (λ + 4λ1) − f (λ)

t u(1)+ lim

t→0+

f (λ + 4λ2) − f (λ)

t u(2)

= f0(λ; h1− kh2k)u(1)+ f0(λ; h1+ kh2k)u(2) , (17) where u(1) = 1

2

µ

1, − h2 kh2k

, u(2) = 1 2

µ

1, h2 kh2k

. Hence, (fsoc)0(x; h) exists with form of (17).

(10)

Secondly, for the subcase of h2 = 0, the same argument applies except h2/kh2k is replaced by any w ∈ IRn−1 with kwk = 1, i.e., choosing u(i) = 1

2(1, (−1)iw), for all i = 1, 2.

Analogously,

t→0lim+

fsoc(x + th) − fsoc(x)

t = f0(λ; h1)u(1)+ f0(λ; h1)u(2). (18) Hence, (fsoc)0(x; h) exists with form of (18). From the above, it shows that fsoc is di- rectionally differentiable at x when x2 = 0 and its directional derivative (fsoc)0(x; h) is either in form of (17) or (18).

“⇒” Suppose fsoc is directionally differentiable at x ∈ IRn with spectral values λ1, λ2, we will prove that f is directionally differentiable at λ1, λ2. For λ1 ∈ IR and any direction d1 ∈ IR, let h := d1u(1)+ 0u(2) where x = λ1u(1)+ λ2u(2). Then, x + th = (λ1+ td1)u(1)+ λ2u(2), and

fsoc(x + th) − fsoc(x)

t = f (λ1+ td1) − f (λ1) t u(1).

Since fsoc is directionally differentiable at x, the above equation yields that f01; d1) = lim

t→0+

f (λ1+ td1) − f (λ1)

t exists.

This means f is directionally differentiable at λ1. Similarly, f is also directionally differ- entiable at λ2. 2

Proof. (c) “⇐” The proof of this direction is identical to the proof shown as in (b), but with “directionally differentiable” replaced by “differentiable”. We omit the proof and only present the formula of fsoc(x) as below. For x2 6= 0, we have

∇fsoc(x) =

b cxT2

kx2k cx2

kx2k aI + (b − a)x2xT2 kx2k2

, (19)

where

a = f (λ2) − f (λ1)

λ2− λ1 , b = f02) + f01)

2 , c = f02) − f01)

2 . (20)

If x2 = 0, then

∇fsoc(x) = f0(λ)I. (21)

“⇒” Let fsoc be Fr´echet-differentiable at x ∈ IRn with spectral eigenvalues λ1, λ2, we will show that f is Fr´echet-differentiable at λ1, λ2. Suppose not, then f is not Fr´echet- differentiable at λi for some i ∈ {1, 2}. Thus, either f is not directionally differentiable at λi or, if it is, the right- and left-directional derivatives of f at λi are unequal. In

(11)

either case, this implies that there exist two sequences of non-zero scalars tν and τν, ν = 1, 2, . . . , converging to zero, such that the limits

ν→∞lim

f (λi+ tν) − f (λi)

tν , lim

ν→∞

f (λi+ τν) − f (λi) τν

exist (possible ∞ or −∞) and either are unequal or both equal to ∞ or are both equal to −∞. Now for any x = λ1u(1) + λ2u(2), let h := 1 · u(1) + 0 · u(2) = u(1). Then, x + th = (λ1+ t)u(1)+ λ2u(2) and fsoc(x + th) = f (λ1+ t)u(1)+ f (λ2)u(2). Thus,

ν→∞lim

fsoc(x + tνh) − fsoc(x)

tν = lim

ν→∞

fsoc1 + tν) − fsoc1)

tν u(1)

ν→∞lim

fsoc(x + τνh) − fsoc(x)

τν = lim

ν→∞

fsoc1 + τν) − fsoc1)

τν u(1) .

It follows that these two limits either are unequal or are both non-finite. This implies that fsoc is not Fr´echet-differentiable at x where is a contradiction. 2

Proof. (d) “⇐” Suppose f is continuously differentiable. From equation (19), it can been seen that ∇fsoc is continuous at every x with x2 6= 0. It remains to show that ∇fsoc is continuous at every x with x2 = 0. Fix any x = (x1, 0) ∈ IRn, so λ1 = λ2 = x1. Then, from equation (20), we have

y→xlima = lim

y→x

f (λ2) − f (λ1)

λ2− λ1 = f0(x1)

y→xlimb = lim

y→x

1

2(f02) + f01)) = f0(x1)

y→xlimc = lim

y→x

1

2(f02) − f01)) = 0 . Taking the limit in equation (19) as y → x yields lim

y→x∇fsoc(y) = f0(x1)I = ∇fsoc(x), which says ∇fsoc is continuous at every x ∈ IRn .

“⇒” The proof for this direction is similar to the original proof of [4, Prop. 5.4], so we omit it. 2

Proof. (e) and (f) The original proofs in [4] use the working Lemma 2.1 directly which is the same idea as the one used for the whole paper, so the proofs for part(e) and (f) are identical to theirs. We therefore omit them. 2

Proof. (g) “⇒” Suppose fsoc is semismooth, then fsoc is strictly continuous and di- rectionally differentiable. By part (b) and (e), f is strictly continuous and directionally differentiable. Now, for any α ∈ IR and any η ∈ IR such that f is differentiable at α + η, part (c) yields that fsoc is differentiable at x + h, where x := (α, 0) ∈ IR × IRn−1 and

(12)

h := (η, 0) ∈ IR × IRn−1. Hence, we can choose the same spectral vectors for x + h and x

such that (

fsoc(x + h) = f (α + η)u(1)+ f (α + η)u(2), fsoc(x) = f (α)u(1)+ f (α)u(2).

Since fsoc is semismooth, by Lemma 2.1, we have

fsoc(x + h) − fsoc(x) − ∇fsoc(x + h)h = o(khk). (22) On the other hand, (21) says ∇fsoc(x+h)h = f0(α+η)Ih =

µ

f0(α+η)η , 0

. Plugging this into equation (22), it yields that f (α + η) − f (α) − f0(α + η)η = o(|η|). Thus, by Lemma 2.1 again, it says that f is semismooth at α. Since α is arbitrary, f is semismooth.

“⇐” Suppose f is semismooth, then f is strictly continuous and directionally differen- tiable. By part (b) and (c), fsoc is strictly continuous and directionally differentiable.

For any x ∈ IRn and h ∈ IRn such that fsoc is differentiable at x + h, we will verify that fsoc(x + h) − fsoc(x) − ∇fsoc(x + h)h = o(khk).

Case (i): If x2 6= 0, let λi be the spectral eigenvalues of x and u(i) be the associated vectors. We denote x + h by z for convenience, i.e., z := x + h and let mi be the spectral values of z with the associated vectors v(i). Hence, we have

( fsoc(x) = f (λ1)u(1)+ f (λ2)u(2), fsoc(x + h) = f (m1)v(1)+ f (m2)v(2). In addition, from (19), we know

∇fsoc(x + h) =

ˆb ˆcz2T kz2k ˆcz2

kz2k ˆaI − (ˆb − ˆa)z2z2T kz2k2

,

where

ˆa = f (m2) − f (m1)

m2− m1 , ˆb = f0(m2) + f0(m1)

2 , ˆc = f0(m2) − f0(m1)

2 .

With this, we can write out fsoc(x+h)−fsoc(x)−∇fsoc(x+h)h := (Ξ1, Ξ2) where Ξ1 ∈ IR and Ξ2 ∈ IRn−1. Since the expansion is very long, for simplicity, we denote Ξ1 be the first component and Ξ2 be the second component of the expansion. We will show that Ξ1 and Ξ2 are both o(khk).

First, we compute the first component Ξ1: Ξ1 = 1

2

(

f (m1) − f (λ1) − f0(m1)(h1−zT2h2 kz2k)

)

+ 1 2

(

f (m2) − f (λ2) − f0(m2)(h1+ z2Th2 kz2k)

)

(13)

= 1 2

½

f (m1) − f (λ1) − f0(m1)

µ

h1− (kz2k − kx2k)

+ o(khk)

¾

+1 2

½

f (m2) − f (λ2) − f0(m2)

µ

h1+ (kz2k − kx2k)

+ o(khk)

¾

= o

µ

h1− (kz2k − kx2k)

+ o(khk) + o

µ

h1+ (kz2k − kx2k)

+ o(khk).

In the above expression of Ξ1, the third equality holds since the following:

zT2h2

kz2k = zT2(z2− x2)

kz2k = kz2k − kz2kkx2k kz2k cos θ

= kz2k − kx2k

µ

1 + O(θ2)

= kz2k − kx2k

µ

1 + O(khk2)

= kz2k − kx2k

µ

1 + o(khk)

,

where θ is the angle between x2 and z2 and note that z2−x2 = h2 gives O(θ2) = O(khk2).

Also the last equality in expression of Ξ1 holds since f is semismooth and mi− λi = h1+ (−1)i(kz2k − kx2k).

On the other hand, due to h1+(−1)i(kz2k−kx2k) ≤ h1+kz2−x2k = h1+kh2k , we observe that when khk → 0 then h1 + (−1)i(kz2k − kx2k) → 0 and h1+ (−1)i(kz2k − kx2k) = O(khk). Thus, o

µ

h1 + (−1)i(kz2k − kx2k)

= o(khk), which yields the first component Ξ1 is o(khk).

Now consider the second component Ξ2: Ξ2 = −1

2f (m1) z2

kz2k +1

2f (m2) z2

kz2k+ 1

2f (λ1) x2

kx2k 1

2f (λ2) x2

kx2k

1 2

µ

f0(m2) − f0(m1)

z2h1

kz2k f (m2) − f (m1) m2− m1 h2

(1 2

µ

f0(m2) − f0(m1)

f (m2) − f (m1) m2− m1

)z2z2Th2

kz2k2

= −1 2

(

f (m1) z2

kz2k − f (λ1) x2

kx2k − f0(m1)z2h1

kz2k 2f (m1) m2− m1h2 +

µ

f0(m1) + 2f (m1) m2− m1

z2z2Th2 kz2k2

)

+1 2

(

f (m2) z2

kz2k − f (λ2) x2

kx2k − f0(m2)z2h1

kz2k− 2f (m2) m2 − m1h2

µ

f0(m2) + 2f (m2) m2− m1

z2z2Th2 kz2k2

)

:= Ξ(1)2 + Ξ(2)2 ,

(14)

where Ξ(1)2 denotes the first half part while Ξ(2)2 denotes the second half part. We will show that both Ξ(1)2 and Ξ(2)2 are o(khk). For symmetry, it is enough to show that Ξ(1)2 is o(khk). From the observations that

z2 = x2+ h2, m2− m1 = 2kz2k,

mi− λi = h1 + (−1)i(kz2k − kx2k) = O(khk), we have the following:

Ξ(1)2 = −1 2

(

f (m1) z2

kz2k − f (λ1) x2

kx2k − f0(m1)z2h1

kz2k 2f (m1) m2− m1h2 +

µ

f0(m1) + 2f (m1) m2− m1

z2z2Th2

kz2k2

)

= −1 2

z2 kz2k

(

f (m1) − f (λ1) − f0(m1)

µ

h1 z2Th2 kz2k

¶)

+1 2

µ

f (m1) − f (λ1)

¶µ−h2

kz2k + z2z2Th2 kz2k3

1 2f (λ1)

µ z2

kz2k x2

kx2k− h2

kz2k + z2z2Th2

kz2k3

.

Following the same arguments as in the first component Ξ1, it can be seen that f (m1) − f (λ1) − f0(m1)

µ

h1 z2Th2 kz2k

= o(khk).

Since m1− λ1 = h1− (kz2k − kx2k) = O(khk) and f is strictly continuous, then f (m1) − f (λ1) = O(khk). In addition, −h2/kz2k + z2z2Th2/kz2k3 = O(khk). Hence,

µ

f (m1) − f (λ1)

¶µ−h2

kz2k +z2zT2h2 kz2k3

= O(khk2) = o(khk) .

Therefore, it remains to show that the last part of Ξ(1)2 is o(khk). Now, note that z2

kz2k x2

kx2k h2

kz2k+ z2z2Th2 kz2k3 = x2

µ 1

kz2k− 1

kx2k + zT2h2 kz2k3

+ O(khk2) .

Let θ(z2) := −1/kz2k, then ∇θ(z2) = − −1 kz2k2

z2

kz2k = z2

kz2k3. This implies that 1

kz2k 1

kx2k+ z2Th2

kz2k3 = θ(x2) − θ(z2) − ∇θ(z2)(x2− z2) = O(khk2), where the last equality is from first Taylor approximation. Thus, we obtain

f (λ1)

µ z2

kz2k x2

kx2k− h2

kz2k + z2z2Th2 kz2k3

= o(khk) .

參考文獻

相關文件

Abstract In this paper, we consider the smoothing Newton method for solving a type of absolute value equations associated with second order cone (SOCAVE for short), which.. 1

Abstract Like the matrix-valued functions used in solutions methods for semidefinite pro- gram (SDP) and semidefinite complementarity problem (SDCP), the vector-valued func-

Chen, Conditions for error bounds and bounded level sets of some merit func- tions for the second-order cone complementarity problem, Journal of Optimization Theory and

Chen, Properties of circular cone and spectral factorization associated with circular cone, to appear in Journal of Nonlinear and Convex Analysis, 2013.

Chen, “Alternative proofs for some results of vector- valued functions associated with second-order cone,” Journal of Nonlinear and Convex Analysis, vol.. Chen, “The convex and

In this paper, by using the special structure of circular cone, we mainly establish the B-subdifferential (the approach we considered here is more directly and depended on the

Abstract We investigate some properties related to the generalized Newton method for the Fischer-Burmeister (FB) function over second-order cones, which allows us to reformulate

Taking second-order cone optimization and complementarity problems for example, there have proposed many ef- fective solution methods, including the interior point methods [1, 2, 3,