• Aucun résultat trouvé

A linearly convergent derivative-free descent method for the second-order cone complementarity problem

N/A
N/A
Protected

Academic year: 2022

Partager "A linearly convergent derivative-free descent method for the second-order cone complementarity problem"

Copied!
25
0
0

Texte intégral

(1)

Vol. 59, No. 8, November 2010, 1173–1197

A linearly convergent derivative-free descent method for the second-order cone complementarity problem

Shaohua Panaand Jein-Shan Chenb*

aSchool of Mathematical Sciences, South China University of Technology, Guangzhou 510640, China;bDepartment of Mathematics, National Taiwan

Normal University, Taipei 11677, Taiwan

(Received 25 August 2008; final version received 1 May 2009) We consider a class of derivative-free descent methods for solving the second-order cone complementarity problem (SOCCP). The algorithm is based on the Fischer–Burmeister (FB) unconstrained minimization reformulation of the SOCCP, and utilizes a convex combination of the negative partial gradients of the FB merit function FB as the search direction. We establish the global convergence results of the algorithm under monotonicity and the uniform JordanP-property, and show that under strong monotonicity the merit function value sequence generated converges at a linear rate to zero. Particularly, the rate of convergence is dependent on the structure of second-order cones. Numerical comparisons are also made with the limited BFGS method used by Chen and Tseng (An unconstrained smooth minimization reformulation of the second-order cone complementarity problem, Math. Program. 104(2005), pp. 293–327), which confirm the theoretical results and the effectiveness of the algorithm.

Keywords: second-order cone complementarity problem; Fischer–

Burmeister function; descent algorithms; derivative-free methods; linear convergence

1. Introduction

We consider the conic complementarity problem of finding a vector2IRnsuch that

2 K, FðÞ 2 K, h,FðÞi ¼0, ð1Þ

where F:IRn!IRn is a mapping assumed to be continuously differentiable throughout this article, and K is the Cartesian product of second-order cones (SOCs). In other words,

K ¼ Kn1 Kn2 Knm, ð2Þ wherem,n1,. . .,nm1,n1þ þnm¼n, and

Kni:¼ ðx 1,x2Þ 2IRIRni1 jx1 kx2k

, ð3Þ

*Corresponding author. Email: jschen@math.ntnu.edu.tw.

ISSN 0233–1934 print/ISSN 1029–4945 online ß2010 Taylor & Francis

DOI: 10.1080/02331930903085359 http://www.informaworld.com

(2)

withkkdenoting the Euclidean norm andK1denoting the set of non-negative reals IRþ. We will refer to (1)–(2) as the second-order cone complementarity problem (SOCCP).

As a direct extension of the non-linear complementarity problem (NCP), the SOCCP includes as a special case the Karush-Kuhn-Tucker (KKT) system of SOC programming, which has a wide range of applications in engineering design, control, finance, robust optimization and combinatorial optimization; see [1,18] and the references therein. Now there have been various methods proposed for solving the SOCCP, which include the merit function method [5], the smoothing Newton methods [6,10,12], the semismooth Newton methods [16,20], and the interior-point method [25]. We observe that the last three kinds of methods in each iteration involve the solution of a linear system of equations, which makes them unsuitable for handling large-scale SOCCPs. On the contrary, the merit function method [5], based on the Fischer–Burmeister (FB) unconstrained minimization reformulation of the SOCCP, requires much less computation work in each iteration and consequently has a certain potential for solving large-scale SOCCPs.

The FB merit function associated with the coneKn is given by

FBðx,yÞ:¼12kFBðx,yÞk2, ð4Þ

whereFB :IRnIRn!IRn is the FB function associated withKn, defined by FBðx,yÞ:¼ ðx2þy2Þ1=2 ðxþyÞ ð5Þ withx2¼xxdenoting the Jordan product ofxand itself,x1/2being a vector such that (x1/2)2¼x, and xþy meaning the componentwise addition of vectors. The functions FBandFBwere studied in the papers [2,5,10,21], where FBwas shown in [10] to satisfy

FBðx,yÞ ¼0()x2 Kn, y2 Kn, hx,yi ¼0, ð6Þ

and its continuous differentiability was established by Chen and Tseng [5], andFB was proved to be strongly semismooth in [21] and [2] via different ways. By equivalence (6), clearly, the SOCCP can be reformulated as an unconstrained minimization problem

min2IRn FBðÞ:¼Xm

i¼1

FBði,FiðÞÞ, ð7Þ

where¼(1,. . .,m),F()¼(F1(),. . .,Fm()) withi2IRni andFi:IRn !IRni. The merit function method in [5] was developed by applying the limited BFGS method directly for the minimization reformulation (7). In this article, we propose another merit function method based on the same reformulation, which can be viewed as an extension of the method in [23] for the NCP. Different from the limited BFGS method adopted by Chen and Tseng [5], our method does not exploit the derivative of the mapping F, but utilizes some convex combination of the negative partial gradients of FB, i.e. the vector of the form rx FB ð1Þry FB with 2(0,1), as the search direction. Since the computation of the search direction and the step size does not involve the Jacobian of F, our derivative-free algorithm

(3)

requires less computation work and lower memory in each iteration than the existing methods mentioned above. We show that the algorithm is globally convergent under monotonicity and the uniform Jordan P-property of F, and particularly that the merit function value sequencef FBðkÞggenerated converges at a linear rate to zero if Fis strongly monotone. But, unlike the NCP case, the rate of convergence depends on the structure ofK (Remark 5.1 (a)).

The literature on derivative-free methods for solving the NCP is vast; see, for example, [11,15,17,19,24,23]. Nevertheless, to the best of our knowledge, there are no papers to study derivative-free methods for the SOCCP except [3] where a different unconstrained reformulation and a different descent direction were employed, and no rate of convergence result was established. The main difficulty is to extend the growth relation between the FB function and the natural residual function established in [22]

to the SOCCP case. In addition, numerical results were not reported for the above derivative-free methods, so the practical performance of these methods cannot be judged. In this article, we obtain the rate of convergence result for the proposed derivative-free descent algorithm by using the favourable properties of the gradients of the function FB(Propositions 3.1 and 3.2), as well as compare the performance of the algorithm with that of the limited BFGS method in [5], which indicates that our method is comparable to the limited BFGS method for some test problems.

Throughout this article, IRn denotes the space of n-dimensional real column vectors, and IRn1 IRnm is identified with IRn1þ þnm. Thus, ðx1,. . .,xmÞ 2 IRn1 IRnmis viewed as a column vector in IRn1þþnm. The notationImeans an identity matrix of suitable dimension, and intðKnÞdenotes the interior ofKn. For any x, y in IRn, we write xKn y if xy2 Kn; and write xKn y if xy2 Kn. For a differentiable mapping F:IRn!IRm,rFðxÞ 2IRnm denotes the transposed Jacobian of F at x. For a symmetric matrix A, we write AO (respectively, AO) to mean A is positive semidefinite (respectively, positive definite).

In addition, we use diag(1,. . .,n) to denote a diagonal matrix with 1,. . .,n as the diagonal elements.

2. Preliminaries

This section recalls some background materials that will be used in the subsequent sections. It is known that Kn is a closed convex self-dual cone with non-empty interior

intðKnÞ:¼x¼ ðx1,x2Þ 2IRIRn1j x14kx2k :

For anyx¼ ðx1,x2Þ,y¼ ðy1,y2Þ 2IRIRn1, we define their Jordan product [8] by xy:¼ ðhx,yi, y1x2þx1y2Þ: ð8Þ The Jordan product, unlike scalar or matrix multiplication, is not associative, which is a main source of complication in the analysis of SOCCP. The identity element under this product is e:¼ ð1, 0,. . ., 0ÞT2IRn. Given a vector x¼ ðx1,x2Þ 2 IRIRn1, let

Lx:¼ x1 xT2 x2 x1I

,

(4)

which can be viewed as a linear mapping from IRn to IRn withLxy¼xy for any y2IRn. It is easy to verify thatLxforx2intðKnÞis invertible with the inverseL1x given by

L1x ¼ 1 detðxÞ

x1 xT2

x2

detðxÞ x1 Iþ 1

x1x2xT2 2

4

3

5, ð9Þ

where detðxÞ:¼x21 kx2k2 denotes the determinant ofx.

We recall from [8,10] that each x¼ ðx1,x2Þ 2IRIRn1 admits a spectral factorization associated with Kn in the form of x¼1ðxÞ uð1Þx þ2ðxÞ uð2Þx , where i(x) anduðiÞx fori¼1, 2 are the spectral values ofxand the corresponding spectral vectors, defined by

iðxÞ:¼x1þ ð1Þikx2k, uðiÞx :¼1

21, ð1Þix2

, ð10Þ

with x2¼kxx2

2k if x26¼0, and otherwise x2 being any vector in IRn1 satisfying kx2k ¼1. Ifx26¼0, the factorization is unique. The spectral factorization ofxand the matrixLxhave various interesting properties; see [10]. We list several ones that will be used later.

LEMMA 2.1

(a) For any x2IRn,x2¼ ð1ðxÞÞ2uð1Þx þ ð2ðxÞÞ2uð2Þx 2 Kn: (b) For any x2 Kn,x1=2¼ ffiffiffiffiffiffiffiffiffiffiffi

1ðxÞ

p uð1Þx þ ffiffiffiffiffiffiffiffiffiffiffi 2ðxÞ

p uð2Þx 2 Kn.

(c) xKn0()1ðxÞ 0()LxO and xKn 0()1ðxÞ40()LxO.

The following lemma is a representation of Problem 7 in [13 p. 468] for the real symmetric matrix case. In view of its importance, we here include its proof.

LEMMA 2.2 Let B,C2IRnnbe symmetric matrices with BO.Then BþCO if and only if every eigenvalue of CB1is greater than1.

Proof By Corollary 7.6.5 of [13], there exists a non-singular matrixD2IRnnsuch thatDTCD¼diag(1,. . .,n) andDTBD¼I. Consequently,

CB1¼ ðDTÞ1diagð1,. . .,nÞD1 ðDTÞ1D11

¼ ðDTÞ1diagð1,. . .,nÞDT:

This implies that CB1 is similar to the diagonal matrix diag(1,. . .,n), and therefore1,. . .,nare the eigenvalues ofCB1including the multiplicities. On the other hand,

BþC¼ ðDTÞ1diagð1þ1,. . ., 1þnÞD1,

which means thatBþCOif and only ifi41 for alli¼1, 2,. . .,n. Combining

the two sides, we then obtain the desired result. g

Next, we review the definitions of the monotonicity and the P-property of a mapping.

Definition 2.1 The mappingF¼(F1,. . .,Fm) withFi:IRn!IRni is said to (a) be monotone if, for every,2IRn,h, FðÞ FðÞi 0;

(5)

(b) be strongly monotone if there exists a40 such that, for every,2IRn, h, FðÞ FðÞi kk2;

(c) have the uniform Jordan P-property if there exists a 40 such that, for every¼ ð1,. . .,mÞ,¼ ð1,. . .,mÞ 2IRn, there exists2{1,2,. . .,m} such that

2½ðÞ ðFðÞ FðÞÞ kk2;

(d) have the uniform CartesianP-property if there exists a 40 such that, for every¼ ð1,. . .,mÞ,¼ ð1,. . .,mÞ 2IRn, there exist2{1,2,. . .,m} such that

h,FðÞ FðÞi kk2:

From Definition 2.1, clearly, the uniform Cartesian P-property implies the uniform JordanP-property, and ifFis strongly monotone with modulus40, then F has the uniform Jordan P-property and the uniform Cartesian P-property with modulus/m. Also, when Fis continuously differentiable, Fis strongly monotone with modulus40 if and only ifrF() is uniformly positive definite with modulus 40, i.e.

dTrFðÞdkdk2 for all, d2IRn:

In addition, we see that the uniform Jordan P-property does not imply the monotonicity.

Unless otherwise stated, in the subsequent three sections, we assume K ¼ Kn, and all analysis can be carried over to the case whereKhas the Cartesian structure as in (2).

3. Some properties ofwFB and)FB

In this section, we present some important properties for the gradient of FB which play a crucial role in analysing the convergence results of the descent algorithm proposed in the next section. In addition, we establish the coerciveness ofFB under two mild conditions. Throughout this section, for any x¼ ðx1,x2Þ,y¼ ðy1,y2Þ 2IRIRn1, we write

w¼ ðw1,w2Þ:¼x2þy2 and z:¼ ðz1,z2Þ ¼ ðx2þy2Þ1=2: ð11Þ First, from Propositions 1 and 2 of [5], we know that the function FB is continuously differentiable everywhere and its gradient is given as in the following lemma.

LEMMA 3.1 The function FB in [4] is continuously differentiable everywhere.

Moreover,rx FBð0, 0Þ ¼ ry FBð0, 0Þ ¼0:If x2þy22intðKnÞ,then rx FBðx,yÞ ¼

LxL1z I

FBðx,yÞ, ry FBðx,yÞ ¼

LyL1z I

FBðx,yÞ:

(6)

If x2þy22=intðKnÞ and (x,y)6¼(0,0),then x21þy216¼0 and rx FBðx,yÞ ¼ x1

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x21þy21

q 1

!

FBðx,yÞ,

ry FBðx,yÞ ¼ y1

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x21þy21

q 1

!

FBðx,yÞ:

For the partial gradients rx FB and ry FB, from [5 Lemma 9] and [4, Theorem 3.1], we readily obtain the following favourable properties whose proofs will be omitted.

PROPOSITION 3.1 The gradientsrx FBandry FBof FBhave the following properties:

(a) hrx FBðx,yÞ,ry FBðx,yÞi 0for all x,y2IRn,and furthermore, the equality holds if and only if FBðx,yÞ ¼0.

(b) For all x,y2IRn,rx FBðx,yÞ ¼ ry FBðx,yÞ ¼0if and only if FBðx,yÞ ¼0.

(c) r FB is globally Lipschitz continuous, i.e. there exists a constant L40 such that

krx FBðx,yÞ rx FBðx, yÞk Lkðx,yÞ ðx, yÞk, kry FBðx,yÞ ry FBðx, yÞk Lkðx,yÞ ðx, yÞk:

for allðx,yÞ,ðx, yÞ 2 IRnIRn,where L is dependent on the dimension n.

Next we will establish another three important properties for the gradientsrx FB and ry FB (Proposition 3.2) which are crucial to analyse the convergent results in Sections 4 and 5. To the end, we need the following technical lemmas. The first one is an extension of [5, Lemma 3], which will be used to give a tighter upper bound for LxþyL1z .

LEMMA 3.2 For any x¼ ðx1,x2Þ,y¼ ðy1,y2Þ 2IRIRn1such that w26¼0,we have ðx1þy1Þ þ ð1Þiðx2þy2ÞTw2

2 ðx 2þy2Þ þ ð1Þiðx1þy1Þw222iðwÞ ð12Þ for i¼1,2,wherew2¼w2=kw2k.

Proof The first inequality can be easily obtained by expanding the square on both sides and using the Cauchy–Schwartz inequality. We next show that the second inequality holds wheni¼1, which is equivalent to proving the following inequality:

ðx2þy2Þkw2k ðx1þy1Þw2

221ðwÞkw2k2: ð13Þ

Let L and R denote the left-hand side and the right-hand side of (13), respectively. Then, by plugging inw2¼2(x1x2þy1y2), it is easy to compute that

L¼ kx2þy2k2kw2k2þ ðx1þy1Þ2kw2k2

4 x21kx2k2þx1y1xT2y2þx21xT2y2þx1y1kx2k2 kw2k 4 y21ky2k2þx1y1xT2y2þy21xT2y2þx1y1ky2k2

kw2k, R¼2ðx21þy21Þkw2k2þ2ðkx2k2þ ky2k2Þkw2k2

4 2x21kx2k2þ2y21ky2k2þ4x1y1xT2y2

kw2k:

(7)

Using the last two equalities, it then follows that RL¼ ðx1y1Þ2kw2k2þ kx2y2k2kw2k2

4 x21kx2k2þy21ky2k2þ2x1y1xT2y2

kw2k þ4x21xT2y2þx1y1ky2k2þy21xT2y2þx1y1kx2k2

kw2k

¼ ðx1y1Þ2kw2k2þ kx2y2k2kw2k22ðx1y1Þðx2y2ÞTw2kw2k

¼ ðx 1y1Þw2 ðx2y2Þkw2k20:

This implies (13), and consequently the inequality (12) holds fori¼1. Using similar arguments, we can prove that the inequality (12) holds fori¼2. g LEMMA 3.3 For any x¼ ðx1,x2Þ,y¼ ðy1,y2Þ 2IRIRn1 such that x2þy2 2 intðKnÞ,

LxþyL1z

2 2ð ffiffiffiffiffiffiffiffiffiffiffi n1 p

þ2 ffiffiffi p2

Þ,

wherekAk2denotes the Frobenius norm (Euclidean norm)of the matrix A2IRnn. Proof Let1,2be the spectral values ofw. Then, by the definition ofz, we have

z1¼ ffiffiffiffiffi 2 p þ ffiffiffiffiffi

1 p

2 , z2¼

ffiffiffiffiffi 2 p ffiffiffiffiffi

1 p

2 w2 ð14Þ

with w2¼kww2

2k if w26¼0, and otherwise w2 being any vector in IRn1 satisfying kw2k ¼1.

Ifw2¼0, then1¼2¼w1¼ kxk2þ kyk2. From formula (9), it follows that LxþyL1z ¼ 1

ffiffiffiffiffiffi w1

p Lxþy¼ 1 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi kxk2þ kyk2

p Lxþy:

Consequently,

LxþyL1z

2

2¼nðx1þy1Þ2þ2kx2þy2k2 kxk2þ kyk2 2n, which immediately implies the desired result.

Ifw26¼0, then by applying formula (9), it is not difficult to compute that

LxþyL1z ¼

ðx1þy1Þz1 ðx2þy2ÞTz2

ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p ðx1þy1ÞzT2 ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p þðx2þy2ÞT

z1 þðx2þy2ÞTz2zT2 z1 ffiffiffiffiffi

1

p ffiffiffiffiffi 2

p ðx2þy2Þz1 ðx1þy1Þz2

ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p ðx2þy2ÞzT2 ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p þðx1þy1ÞI

z1 þðx1þy1Þz2zT2 z1 ffiffiffiffiffi

1

p ffiffiffiffiffi 2

p 2

66 64

3 77 75

:¼ b1ðx,yÞ b2ðx,yÞT c2ðx,yÞ B2ðx,yÞ

" #

:

(8)

Substituting the expressions ofz1,z2in (14) into the entries of the above matrix, we get

b1ðx,yÞ ¼ðx1þy1Þ þ ðx2þy2ÞTw2

2 ffiffiffiffiffi 2

p þðx1þy1Þ ðx2þy2ÞTw2

2 ffiffiffiffiffi 1

p ,

c2ðx,yÞ ¼ðx2þy2Þ þ ðx1þy1Þw2

2 ffiffiffiffiffi 2

p þðx2þy2Þ ðx1þy1Þw2

2 ffiffiffiffiffi 1

p ,

b2ðx,yÞ ¼1½ðx1þy1Þ þ ðx2þy2ÞTw2w2

2 ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p ð ffiffiffiffiffi 1

p þ ffiffiffiffiffi 2

p Þ 2½ðx1þy1Þ ðx2þy2ÞTw2w2

2 ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p ð ffiffiffiffiffi 1

p þ ffiffiffiffiffi 2

p Þ þ2ðx2þy2Þ ðx2þy2ÞTw2w2

ffiffiffiffiffi 1 p þ ffiffiffiffiffi

2

p ,

B2ðx,yÞ ¼1½ðx2þy2Þ þ ðx1þy1Þw2wT2 2 ffiffiffiffiffi

1

p ffiffiffiffiffi 2

p ð ffiffiffiffiffi 1

p þ ffiffiffiffiffi 2

p Þ 2½ðx2þy2Þ ðx1þy1Þw2wT2 2 ffiffiffiffiffi

1

p ffiffiffiffiffi 2

p ð ffiffiffiffiffi 1

p þ ffiffiffiffiffi 2

p Þ þ ðx1þy1Þ

ffiffiffiffiffi 1 p þ ffiffiffiffiffi

2

p ð2Iw2wT2Þ:

Now, using Lemma 3.2, we can verify that the following inequalities hold:

ðx1þy1Þ þ ðx2þy2ÞTw2

2 ffiffiffiffiffi 2 p

ðx2þy2Þ þ ðx1þy1Þw2

2 ffiffiffiffiffi 2 p

1

ffiffiffi2 p , ðx1þy1Þ ðx2þy2ÞTw2

2 ffiffiffiffiffi 1 p

ðx2þy2Þ ðx1þy1Þw2 2 ffiffiffiffiffi

1 p

1

ffiffiffi2 p , and

1½ðx1þy1Þ þ ðx2þy2ÞTw2w2

2 ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p ð ffiffiffiffiffi 1

p þ ffiffiffiffiffi 2

p Þ 2½ðx1þy1Þ ðx2þy2ÞTw2w2

2 ffiffiffiffiffi 1

p ffiffiffiffiffi 2

p ð ffiffiffiffiffi 1

p þ ffiffiffiffiffi 2

p Þ

ffiffiffi

2 p

, 1½ðx2þy2Þ þ ðx1þy1Þw2wT2

2 ffiffiffiffiffi 1 p ffiffiffiffiffi

2 p ð ffiffiffiffiffi

1 p þ ffiffiffiffiffi

2

p Þ 2½ðx2þy2Þ ðx1þy1Þw2wT2 2 ffiffiffiffiffi

1 p ffiffiffiffiffi

2 p ð ffiffiffiffiffi

1 p þ ffiffiffiffiffi

2 p Þ

2

ffiffiffi 2 p

: This together withjx1þy1j ffiffiffiffiffi

1

p þ ffiffiffiffiffi 2

p andkx2þy2k ffiffiffiffiffi 1

p þ ffiffiffiffiffi 2

p implies that jb1ðx,yÞj kc2ðx,yÞk ffiffiffi

p2

, kb2ðx,yÞk ffiffiffi p2

þ3, kB2ðx,yÞk22 ffiffiffiffiffiffiffiffiffiffiffi n1 p

þ1þ ffiffiffi p2

: Consequently,LxþyL1z

22 ffiffiffiffiffiffiffiffiffiffiffi n1 p

þ4 ffiffiffi p2

. The proof is thus completed. g It should be pointed out that using Lemmas 3–4 of [5] we may also get a upper bound forkLxþyL1z k2, but such a upper bound is not tighter than the one given here.

By using Lemma 3.3, we can further obtain the following result. Its proof is simple, however, as will be shown below, this result is a key to establish Proposition 3.2 (b).

LEMMA 3.4 For any given x,y2IRn such that x2þy22intðKnÞ, let A:¼ L2zðxþyÞL1z and pAðtÞ ¼tnþa1ðx,yÞtn1þ þan1ðx,yÞtþanðx,yÞ be its charac- teristic polynomial. Then, there exists a constant c1(n)41 dependent on n such that

kAn1þa1ðx,yÞAn2þ þan1ðx,yÞAk2c1ðnÞ: ð15Þ

(9)

Proof For any givenx,y2IRnsuch thatx2þy22intðKnÞ, sinceA¼2ILxþyL1z , applying Lemma 3.3 yields

kAk22ð ffiffiffi pn

þ ffiffiffiffiffiffiffiffiffiffiffi n1 p

þ2 ffiffiffi p2

Þ: ð16Þ

Let c2ðnÞ:¼2ð ffiffiffi pn

þ ffiffiffiffiffiffiffiffiffiffiffi n1 p

þ2 ffiffiffi p2

Þ. Then, from the inequality (3.1.11) of [14], we have

jiðAÞj c2ðnÞ, i¼1, 2,. . .,n,

where1(A),. . .,n(A) are the eigenvalues ofAincluding multiplicities. Sinceak(x,y) is the sum of all kn k-fold products of distinct items from1(A),. . .,n(A), i.e.

akðx,yÞ ¼ X

1i155ikn

Yk

j¼1

ijðAÞ, k¼1, 2,. . .,n,

there exists a positive constantc3(n) only dependent on the dimensionnsuch that jakðx,yÞj c3ðnÞ, k¼1, 2,. . .,n: ð17Þ Combining Equations (16) and (17), we immediately obtain (15) with

c1ðnÞ:¼max 1, c2ðnÞn1þc3ðnÞc2ðnÞn2þ þc3ðnÞc2ðnÞ ,

and consequently the desired result follows. g

Now we are in a position to present the three crucial properties of rx FB and ry FB.

PROPOSITION 3.2 The gradientsrx FBandry FBof FBhave the following properties:

(a) krx FBðx,yÞ þ ry FBðx,yÞk 2ð ffiffiffi pn

þ ffiffiffiffiffiffiffiffiffiffiffi n1 p þ2 ffiffiffi

p2

ÞkFBðx,yÞk for all x,y2IRn;

(b) rx FBðx,yÞ þ ry FBðx,yÞð32 ffiffi2 p

Þn

2nc1ðnÞ kFBðx,yÞk for all x,y2IRn, where c1ðnÞis the constant from Lemma3.4.

(c) krx FBðx,yÞ þ ry FBðx,yÞk ¼0 if and only if x2 K, y2 K, hx,yi ¼0.

Proof (a) We prove the result by the following three cases:

Case1 (x,y)¼(0,0). In this case, the result is clear by Lemma 3.1 andFB(0,0)¼0.

Case2 x2þy22intðKnÞ. Using Lemmas 3.1 and 3.3, it follows that rx FBðx,yÞ þ ry FBðx,yÞ

¼ ð2I LxþyL1z ÞFBðx,yÞ k2ILxþyL1z k2kFBðx,yÞk 2ð ffiffiffi

pn

þ ffiffiffiffiffiffiffiffiffiffiffi n1 p

þ2 ffiffiffi 2 p

ÞkFBðx,yÞk: ð18Þ Case3 x2þy22=intðKnÞand (x,y)6¼(0, 0). From Lemma 3.1 we have that

rx FBðx,yÞ þ ry FBðx,yÞ

¼ x1þy1

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x21þy21

q 2

!

FBðx,yÞ

¼ 2x1þy1

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x21þy21 q

!

kFBðx,yÞk

kFBðx,yÞk, ð19Þ

(10)

where the second equality is due toffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðx1þy1Þ22ðx21þy21Þ, and the inequality is since

x21þy21

q x1þy1 by the non-negativity ofx1,y1.

(b) Similar to part (a), we also proceed the proof by the three cases.

Case1 (x,y)¼(0,0). The result is clear by Lemma 3.1 andFB(0,0)¼0.

Case2 x2þy22intðKnÞ. In this case, from Lemma 3.1 it follows that rx FBðx,yÞ þ ry FBðx,yÞ

¼L2zðxþyÞL1z FBðx,yÞ:

Notice that zKn 0 and 4z2 ðxþyÞ2¼2z2þ ðxyÞ2zKn 0. From [10, Proposition 3.4] we have 2z ðxþyÞ Kn 0, which by Lemma 2.1 (c) implies L2z(xþy)O. Consequently,

rx FBðx,yÞ þ ry FBðx,yÞ

¼ kFBðx,yÞk

L2zðxþyÞL1z

1

2

¼ kFBðx,yÞk LzL12zðxþyÞ

2

: ð20Þ

We next prove that all eigenvalues of LzL12zðxþyÞ are bounded. Since L2z(xþy)OandL2z(xþy)þLzO, setting B¼L2z(xþy),C¼Lzand apply- ing Lemma 2.2 then yields that every eigenvalue ofCB1is greater than –1, i.e.

i LzL12zðxþyÞ

4 1, i¼1, 2,. . .,n: ð21Þ On the other hand, since zKn 0 and 2z2 ðxþyÞ2¼ ðxyÞ2xKn0, we have from [10, Proposition 3.4] that ffiffiffi

p2

z ðxþyÞxKn y ffiffiffi p2

z jxþyjxKn0.

Consequently,

½2z ðxþyÞ 3=2 ffiffiffi p2

z¼ ð1=2Þzþ ffiffiffi p2

z ðxþyÞ Kn 0:

This in turn implies L2zðxþyÞLð3=2pffiffi2

ÞzO. Setting B¼L2zðxþyÞ,C¼ Lð3=2 ffiffi

2 p

Þzand applying Lemma 2.2 again, we have i Lð3=2 ffiffi

2 p

ÞzL12zðxþyÞ

41, i¼1, 2,. . .,n, and therefore,

iLzL12zðxþyÞ

5 2 32 ffiffiffi

p2, i¼1, 2,. . .,n: ð22Þ Combining (21) and (22) shows that all eigenvalues of LzL12zðxþyÞ are bounded and

iLzL12zðxþyÞ

5 2

32 ffiffiffi

p2, i¼1, 2,. . .,n: ð23Þ Now letA¼L2zðxþyÞL1z andpA(t) be the characteristic polynomial ofAdefined as in Lemma 3.4. Then, using the fact thatpA(A)¼0, we obtain

An1þa1ðx,yÞAn2þ þan1ðx,yÞ þanðx,yÞA1 ¼0,

(11)

which in turn implies that A1¼ 1

anðx,yÞ An1þa1ðx,yÞAn2þ þan1ðx,yÞ

¼ 1

1ðAÞ nðAÞ An1þa1ðx,yÞAn2þ þan1ðx,yÞ

¼ 1ðA1Þ nðA1Þ An1þa1ðx,yÞAn2þ þan1ðx,yÞ

: ð24Þ Note thatA1is preciselyLzL12zðxþyÞ. Hence, from (23) to (24) and Lemma 3.4, we have

kLzL12zðxþyÞk2¼ kA1k2 2 32 ffiffiffi

p2

n

c1ðnÞ:

This together with (20) yields the desired result.

Case3 ðx,yÞ2=intðKnÞand (x,y)6¼(0,0). Using (19) andjx1þy1j pffiffiffi2 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x21þy21

q ,

rx FBðx,yÞ þ ry FBðx,yÞ

ð2 ffiffiffi

2 p

ÞkFBðx,yÞk:

Noting that 2 ffiffiffi p2

ð32pffiffi2Þn

2nc1ðnÞ sincec1(n)41, the desired result follows.

(c) This is direct by using parts (a)–(b) and the equivalence (6). g In what follows, we establish the coerciveness of the functionFBunder some mild assumptions of F. For this purpose, we assume that K is given by (2), and corresponding to the Cartesian structure of K, write ¼(1,. . .,m) and F¼ (F1,. . .,Fm) withi2IRniandFi:IRn!IRni. The following lemma and assumptions will be needed.

LEMMA 3.5 [20 Lemma 5.2] Let FB be defined by (5). For any sequence fðxk,ykÞg IRnIRn, let 1k2k and 1k2k denote the spectral values of xk andyk,respectively.

(a) Iff1kg ! 1or f1kg ! 1,then {kFB(xk,yk)k}! 1.

(b) If f1kg and f1kg are bounded below, but f2kg,f2kg ! þ1 and

xk kxkkkyykkk

n o

6!0,thenfkFBðxk,ykÞkg ! 1.

ASSUMPTION 3.1 For any sequence fkg IRn satisfying limk!1kkk ¼ 1,if there exists2{1,. . .,m}such that the sequencesf1ðkÞg,f1ðFðkÞÞgare bounded below, butf2ðkÞg,f2ðFðkÞÞg ! 1,then there holds that

k

kkk FðkÞ

kFðkÞk6!0 as k! 1: ð25Þ ASSUMPTION 3.2 There exist40 and r2(0,1]such that the mapping F satisfies

kFðÞk kFð0Þk þkkr for any2IRn:

PROPOSITION 3.3 Let FBbe given by (7)and F¼(F1,. . .,Fm)with Fi:IRn!IRni. Then, the functionFBis coercive under one of the following conditions:

(a) F has the uniform Jordan P-property and Assumption3.1holds;

(b) F has the uniform Jordan P-property and Assumption3.2holds.

(12)

Proof The proof is by contradiction. Assume that a sequence {k} exists such that limk!1 kkk ¼ 1 and the sequence {FB(k)} is bounded. Corresponding to the structure ofK, for eachk we writek¼ ð1k,. . .,mkÞwith ik2IRni. Define the index set

J:¼i2 f1, 2,. . .,mg j fikgis unbounded :

Clearly, J6¼ ; since {k} is unbounded. Let {k} be a bounded sequence with k¼ ð1k,. . .,mkÞandik2IRni for i¼1, 2,. . .,m, where ik for eachkis defined as follows:

ik ¼ 0 ifi2J, ik otherwise:

(a) From the uniform JordanP-property ofF, there exists 40 such that kkkk2 max

i¼1,...,m2 ðikikÞ ðFiðkÞ FiðkÞÞ

¼max

i2J 2 ik ðFiðkÞ FiðkÞÞ

¼2 k ðFðkÞ FðkÞÞ kk ðFðkÞ FðkÞÞk ffiffiffi

2 p

kkkkFðkÞ FðkÞk, ð26Þ whereis one of the indices for which the maximum is attained and which we have, without loss of generality, assumed to be independent ofk, and the last inequality is easily shown by (8). Since 2J, we assume without loss of generality that fkkkg ! 1. Since kkkk2 kkkk2¼ kkk2, dividing the both sides of (26) bykkkthen yields

kkk ffiffiffi p2

kFðkÞ FðkÞk ffiffiffi p2

kFðkÞk þ kFðkÞk

:

This, together with the boundedness of {F(k)}, impliesfkFðkÞkg ! 1. Thus, fkkkg ! 1 and fkFðkÞkg ! 1: ð27Þ Now if f1ðkÞg ! 1 or f1ðFðkÞÞg ! 1, then using Lemma 3.5 (a) readily yieldsfkFBðk,FðkÞÞkg ! 1and hence {FB(k)}! 1, which gives a contradic- tion to the boundedness of {FB(k)}. Otherwise, from (27) we havef2ðkÞg ! 1 and {2(F(k))}! 1. By the given assumption, condition (25) holds. Then, {k} satisfies Lemma 3.5 (b), which in turn implies {FB(k)}! 1. This is clearly impossible.

(b) From the above discussions, Equations (26)–(27) still hold for this case.

If f1ðkÞg ! 1 or {2(F(k))}! 1, then from part (a) it is impossible.

Otherwise, from (27) we have f2ðkÞg ! 1 and {2(F(k))}! 1. We next show thatkkk

kkFFðkÞ

ðkÞk6!0 ask! 1. If not, by the continuity of2() and Equation (27),

k!1lim

2 k ðFðkÞ FðkÞÞ kkkkFðkÞk lim

k!12 k

kkk FðkÞ kFðkÞk

þ lim

k!12

kFðkÞ kkkkFðkÞk

¼0, ð28Þ

(13)

where the inequality is easily shown by (8) and the equality is due to the boundedness of {F(k)}. On the other hand, from Assumption 3.2, there exist 40 andr2(0,1]

such thatkFðkÞk kFðkÞk kFð0Þk þkkkrfor eachk, and hence,

k!1lim

kkkk2 kkkkFðkÞk lim

k!1

kkkk2 kkkðkFð0Þk þkkkrÞ

40:

This together with the first inequality of (26) yields a contradiction to (28). Thus, we verify that the sequencesfkgandfFðkÞgsatisfy the conditions of Lemma 3.5 (b).

Consequently, we have {FB(k)}! 1. This is clearly impossible. g Since the uniform CartesianP-property implies the uniform JordanP-property, the condition of Proposition 3.3 (a) is weaker than that of Proposition 5.2 in [20].

We also see that Assumption 3.2 is weaker than the Lipschitz continuity ofF. WhenK reduces to the non-negative orthant cone IRnþand the Jordan product ‘’ becomes the component wise product of the vectors, since Assumption 3.1 automatically holds and the uniform Jordan P-property of F is equivalent to saying that F is a uniform P-function, we readily recover the result of [7, Theorem 4.2] from Proposition 3.3 (a).

4. A descent method and global convergence

In this section, we propose a derivative-free descent algorithm based on the minimization reformulation (7). The algorithm will make use of the vector of the following form:

dð,Þ:¼ rx FBð,FðÞÞ ð1Þry FBð,FðÞÞ ð29Þ as the search direction, where 2[0,1) is a parameter. Note that d(,) for any 2[0,1) may not be a descent direction ofFBat. But, the following lemma states that, whenFis monotone, there always existsðÞ 2 ð0, 1 such thatd(,) for any 2 ½0,ðÞÞ is a descent direction. The idea for constructing such a direction is borrowed from [23].

LEMMA 4.1 Suppose that F is monotone. If is not a solution of the SOCCP, then there existsðÞ 2 ð0, 1 such thatrFB()Td(,)50 for all2 ½0,ðÞÞ.

Proof Since F is continuously differentiable, the function FB() is also continuously differentiable by Lemma 3.1. Using the chain rule, the gradient of FBat is

rFBðÞ ¼ rx FBð,FðÞÞ þ rFðÞry FBð,FðÞÞ: ð30Þ This together with the definition ofd(,) yields that

rFBðÞTdð,Þ ¼ krx FBð,FðÞÞk2rx FBð,FðÞÞ,rFðÞry FBð,FðÞÞ ð1Þ r x FBð,FðÞÞ,ry FBð,FðÞÞ

ð1Þ r y FBð,FðÞÞ,rFðÞry FBð,FðÞÞ

: ð31Þ Let

qðÞ:¼ krx FBð,FðÞÞk2 r x FBð,FðÞÞ,rFðÞry FBð,FðÞÞ

(14)

and

pðÞ:¼ r x FBð,FðÞÞ,ry FBð,FðÞÞ

r y FBð,FðÞÞ,rFðÞry FBð,FðÞÞ : Then, (31) can be rewritten as

rFBðÞTdð,Þ ¼ ð1ÞpðÞ þqðÞ:

Note that the first term ofp() is negative by Proposition 3.1 (a) since is not a solution of the SOCCP, whereas the second term is non-positive sinceFis monotone.

Therefore, we havep()50. LetðÞ be defined as follows:

ðÞ :¼

pðÞ

qðÞ pðÞ ifqðÞ4pðÞand pðÞ qðÞ pðÞ1;

1 otherwise:

8<

:

We see that for all2 ½0,ðÞÞ, the search direction d(,) defined by (29) satisfies the descent conditionrFB()Td(,)50. The proof is thus completed. g Lemma 4.1 motivates us to propose the following descent algorithm withd(,).

Algorithm 4.1

Step0. Choose02IRn, 0, 2 ð0, 1=2Þand,2(0,1) with4. Setk:¼0.

Step1. IfFB(k), then stop andkis an approximate solution of the SOCCP.

Step2. Let lkbe the smallest non-negative integerlsatisfying FBðkþlk,lÞÞ FBðkÞ

2lkrx FBðk,FðkÞÞ þ ry FBðk,FðkÞÞk2, ð32Þ whered(,) is defined as in (29), and set

dkðlkÞ:¼dðk,lkÞ and kþ1kþlkdkðlkÞ:

Step3. Let k:¼kþ1, and then go to Step 1.

Algorithm 4.1 is similar to the one proposed in [23] for the NCP with a regularized FB merit function. Since there is no need to compute the gradient ofFB

and the Jacobian ofF(), Algorithm 4.1 is suitable for large-scale problems, as well as applications where the Jacobians ofF() are not available or are costly to compute. In addition, the stepsize and the search direction are adjusted during the backtracking search of Armijo-type, which may be regarded as a kind of curvilinear search.

In what follows, we analyse the global convergence of Algorithm 4.1. Without loss of generality, we assume that¼0. We first show that under the monotonicity of Fevery accumulation point of the sequence {k} is a solution of the SOCCP.

THEOREM 4.1 Suppose that F is monotone. Then, Algorithm 4.1 is well-defined for any initial point 0.Furthermore, if* is an accumulation point of the sequence{k} generated by Algorithm4.1,then*is a solution of the SOCCP.

Proof The proofs are similar to those of [23, Theorem 4.1]. We first show that, whenever k is not a solution, there exists a non-negative integer lk in Step 3 of

(15)

Algorithm 4.1 such that (32) holds. Suppose not, then for any positive integerl, we have

FBðkþlk,lÞÞ FBðkÞ4 2lkrxFBðk,FðkÞÞ þ ryFBðk,FðkÞÞk2: Dividing the above inequality byland passing to the limit l! 1, we get

l!1lim

FBðkþlk,lÞÞ FBðkÞ

l 0: ð33Þ On the other hand, using the mean-value theorem, it follows that

FBðkþlk,lÞÞ FBðkþlk, 0ÞÞ

¼lrFBkþlk, 0Þ þtlðdðk,lÞ dðk, 0ÞÞT

k,lÞ dðk, 0Þ

¼llrFBkþlk, 0Þ þtllkÞT

kÞ,

where t is a constant such that t2(0, 1) and hðkÞ:¼ ry FBðk,FðkÞÞ rx FBðk,FðkÞÞ. From this and the continuity ofrFB, we immediately obtain

l!1lim

FBðkþlk,lÞÞ FBðkþlk, 0ÞÞ l ¼0:

Consequently,

l!1lim

FBðkþlk,lÞÞ FBðkÞ l

¼ lim

l!1

FBðkþlk,lÞÞ FBðkþlk, 0ÞÞ l

þlim

l!1

FBðkþlk, 0ÞÞ FBðkÞ l

¼ rFBðkÞTk, 0Þ: ð34Þ

Combining (34) with (33) then yieldsrFB (k)Td(k,0)0. This gives a contradiction, since, by Lemma 4.1,d(k,0) must be a descent direction of FBat kifkis not a solution of the SOCCP. Thus, Algorithm 4.1 is well defined.

Next, we prove that any accumulation point * of {k} is a solution of the SOCCP. Let {k}k2K} be a subsequence converging to *. From the definition of d(,), we see thatd(,) is continuous, which implies thatdkðlkÞ ¼dðk,lkÞ !d ask(2K)! 1. SinceFB(k) decreases at each iteration, the right-hand side of (32) tends to 0. We next proceed the discussions by two cases: {lk}k2Kis bounded and {lk}k2Kis unbounded.

Case1 {lk}k2Kis bounded. In this case,flkgk2Kdoes not approach 0. Consequently, krx FBð,FðÞÞ þ ry FBð,FðÞÞk2 ¼0:

From Proposition 3.2 (c), it then follows that * is a solution of the SOCCP.

Références

Documents relatifs

An efficient numerical method for solving a symmetric positive definite second-order cone linear complementarity problem (SOCLCP) is proposed.. The method is shown to be more

(i) spectral factorization associated with circular cones; (ii) smooth and nonsmooth analysis for f L θ given as in (7); (iii) the so-called L θ -convexity; and (iv) L θ

Although we have obtained the global and superlinear convergence properties of Algorithm 3.1 under mild conditions, this does not mean that Algorithm 3.1 is practi- cally efficient,

block multiconvex, block coordinate descent, Kurdyka–Lojasiewicz inequality, Nash equilibrium, nonnegative matrix and tensor factorization, matrix completion, tensor

More precisely, we estimate monotone non-increasing densities in a Bayesian setting and de- rive concentration rate for the posterior distribution for a Dirichlet process and

This would imply the Mumford–Shah conjecture from [19] that says that for reduced minimizers of the Mumford–Shah functional on simple planar domains (see the definition at the

It would be interesting to state the problem in which water can form such reservoirs and show monotonicity properties of the free boundary also in this case.. and

In this paper, our main aim is not to require milder conditions to the operator F , but stronger, which pursue a modification, not a restriction, of the valid starting points