• Aucun résultat trouvé

Nonlinear Analysis

N/A
N/A
Protected

Academic year: 2022

Partager "Nonlinear Analysis"

Copied!
18
0
0

Texte intégral

(1)

Contents lists available atScienceDirect

Nonlinear Analysis

journal homepage:www.elsevier.com/locate/na

Analysis of nonsmooth vector-valued functions associated with infinite-dimensional second-order cones

Ching-Yu Yang, Yu-Lin Chang, Jein-Shan Chen

,1

Department of Mathematics, National Taiwan Normal University, Taipei 11677, Taiwan

a r t i c l e i n f o

Article history:

Received 28 July 2010 Accepted 21 May 2011 Communicated by Ravi Agarwal

Keywords:

Hilbert space

Infinite-dimensional second-order cone Strong semismoothness

a b s t r a c t

Given a Hilbert space H, the infinite-dimensional Lorentz/second-order cone K is introduced. For anyx ∈H, a spectral decomposition is introduced, and for any function f :R →R, we define a corresponding vector-valued functionfH(x)on Hilbert spaceH by applyingfto the spectral values of the spectral decomposition ofx∈Hwith respect to K. We show that this vector-valued function inherits fromf the properties of continuity, Lipschitz continuity, differentiability, smoothness, as well as s-semismoothness. These results can be helpful for designing and analyzing solution methods for solving infinite- dimensional second-order cone programs and complementarity problems.

©2011 Elsevier Ltd. All rights reserved.

1. Introduction

LetHbe a real Hilbert space endowed with an inner product

⟨· , ·⟩

, and we write the norm induced by

⟨· , ·⟩

as

‖ · ‖

. For any given closed convex coneK

H,

K

:= {

x

H

| ⟨

x

,

y

⟩ ≥

0

, ∀

y

K

}

is the dual cone ofK. A closed convex coneKinHis calledself-dualifKcoincides with its dual coneK; for example, the non-negative orthant coneRn+and the second-order cone (also called Lorentz cone)Kn

:= { (

r

,

x

) ∈

R

×

Rn1

|

r

≥ ‖

x

‖}

. As discussed in [1], this Lorentz coneKncan be rewritten as

Kn

:=

x

Rn

| ⟨

x

,

e

⟩ ≥ √

1 2

x

withe

= (

1

,

0

) ∈

R

×

Rn1

.

This motivates us to consider the following closed convex cone in the Hilbert spaceH: K

(

e

,

r

) := {

x

H

| ⟨

x

,

e

⟩ ≥

r

x

‖}

wheree

Hwith

e

‖ =

1 andr

Rwith 0

<

r

<

1. It can be seen thatK

(

e

,

r

)

is pointed, i.e.,K

(

e

,

r

) ∩ ( −

K

(

e

,

r

)) = {

0

}

. Moreover, by denoting

e

:= {

x

H

| ⟨

x

,

e

⟩ =

0

} ,

we may express the closed convex coneK

(

e

,

r

)

as K

(

e

,

r

) =

x

+ λ

e

H

|

x

∈ ⟨

e

and

λ ≥ √

r 1

r2

x

.

Corresponding author. Tel.: +886 2 29325417; fax: +886 2 29332342.

E-mail addresses:yangcy@math.ntnu.edu.tw(C.-Y. Yang),ylchang@math.ntnu.edu.tw(Y.-L. Chang),jschen@math.ntnu.edu.tw,jschen@ntnu.edu.tw (J.-S. Chen).

1 Member of Mathematics Division, National Center for Theoretical Sciences, Taipei Office.

0362-546X/$ – see front matter©2011 Elsevier Ltd. All rights reserved.

doi:10.1016/j.na.2011.05.068

(2)

C.-Y. Yang et al. / Nonlinear Analysis 74 (2011) 5766–5783 5767

WhenH

=

Rnande

= (

1

,

0

) ∈

R

×

Rn1,K

(

e

,

12

)

coincides withKn. In view of this, we shall callK

(

e

,

12

)

the infinite- dimensional second-order cone (or infinite-dimensional Lorentz cone) inHdetermined bye. In the rest of this paper, we shall only consider any fixed unit vectore

H, and denote

K

=

K

e

, √

1

2

since two infinite-dimensional second-order conesK

(

e1

)

andK

(

e2

)

associated with different unit elementse1ande2inH are isometric. This means there exists a bijective isometryPwhich mapsK

(

e1

)

ontoK

(

e2

)

such that

Px

‖ = ‖

x

for any x

K

(

e1

)

. For example, lete1

= (

1

,

0

,

0

)

ande2

= (

0

,

0

,

1

)

. Then, for anyx

K

(

e1

)

andy

K

(

e2

)

, we have the following relation:

y

=

Px

=

0 0 1

0 1 0

1 0 0

x

.

Moreover, this mapping preserves the Jordan algebra structure, i.e.,P

(

x

y

) =

Px

Py. In the infinite-dimensional Hilbert space,P is indeed a unitary operator. In light of this fact, we can consider the infinite-dimensional second-order cone associated with a fixed arbitrary unit element inH.

Unless specifically stated otherwise, we shall alternatively write any pointx

Hasx

=

x

+ λ

ewithx

∈ ⟨

e

and

λ = ⟨

x

,

e

. In addition, for anyx

,

y

H, we shall writex

Ky(respectively,x

Ky) ifx

y

intK(respectively,x

y

K).

Now, we introduce the spectral decomposition for any elementx

H. For anyx

=

x

+ λ

e

H, we can decomposexas

x

= α

1

(

x

) · v

x(1)

+ α

2

(

x

) · v

(x2)

,

(1)

where

α

1

(

x

), α

2

(

x

)

and

v

x(1)

, v

x(2)are the spectral values and the associated spectral vectors ofx, with respect toK, given by

α

i

(

x

) = λ + ( −

1

)

i

x

‖ ,

(2)

v

(xi)

=



 1 2

e

+ ( −

1

)

i

xx

,

x

̸=

0 1

2

(

e

+ ( −

1

)

i

w),

x

=

0

(3)

fori

=

1

,

2 with

w

being any vector inHsatisfying

‖ w ‖ =

1. With this spectral decomposition, for any functionf

:

R

R, the following vector-valued function associated withKis defined:

fH

(

x

) =

f

1

(

x

))v

(x1)

+

f

2

(

x

))v

(x2)

x

H

.

(4)

The above definition is analogous to the one in finite-dimensional second-order cone case [2,3].

The motivation of studyingfHdefined as in(4)is from concerning with the complementarity problem associated with infinite-dimensional second-order coneK, i.e., to find anx

Hsuch that

x

K

,

T

(

x

) ∈

K and

x

,

T

(

x

) ⟩ =

0

,

(5)

where T is a mapping from H to H. We denote this problem (5) as CP

(

K

,

T

)

. More specifically, when dealing with such complementarity problem by nonsmooth function approach, i.e., recasting it as a nonsmooth system of equations, we need to check what kind of properties of f can be inherited by fH so that we can know to what extent the convergence analysis of solutions methods based on such nonsmooth system can be obtained. Indeed, the format of the aforementioned complementarity problem CP

(

K

,

T

)

indeed follows the direction of complementarity problems associated with symmetric cones in Euclidean Jordan algebra. Recently, nonlinear symmetric cone optimization and complementarity problems in finite-dimensional spaces such as semidefinite cone optimization and complementarity problems, second- order cone optimization and complementarity problems, and general symmetric cone optimization and complementarity problems, become an active research field of mathematical programming. Taking second-order cone optimization and complementarity problems for example, there have proposed many effective solution methods, including the interior-point methods [4–7], the smoothing Newton methods [8,3,9], the semismooth Newton methods [10,11], and the merit function method [12,13]. However, there are very limited works about nonlinear symmetric cone optimization and complementarity problems in infinite-dimensional spaces, for instance [14], in which with the JB algebras of finite rank primal–dual interior- point methods are presented for some special type of infinite-dimensional cone optimization problems.

It is our intention to extend the above methods for infinite-dimensional complementarity problem CP

(

K

,

T

)

, in which the vector-valued functionfH will play a key role. In this paper, we study the continuity and differential properties of the vector-valued functionfH in general. In particular, we show that the properties of continuity, strict continuity (locally Lipschitz continuity), Lipschitz continuity, directional differentiability, differentiability, continuous differentiability, and s-semismoothness are each inherited byfH fromf. These results can give some concept in designing solutions methods for solving infinite-dimensional second-order cone programs and infinite-dimensional second-order cone complementarity problems.

(3)

2. Preliminaries

For anyx

=

x

+ λ

e

Handy

=

y

+ µ

e

H, we define the Jordan product ofxandyby

x

y

:= (µ

x

+ λ

y

) + ⟨

x

,

y

e

,

(6)

and writex2

=

x

x. Clearly, whenH

=

Rnande

= (

1

,

0

) ∈

R

×

Rn1, this definition coincides with the one given in [15, Chapter II] which is the case of finite-dimensional second-order cone associated with Euclidean Jordan algebra. The following technical lemmas will be frequently used in the subsequent analysis.

Lemma 2.1. Let

α

1

(

x

), α

2

(

x

)

be the spectral values of x

Hand

α

1

(

y

), α

2

(

y

)

be the spectral values of y

H. Then we have

| α

1

(

x

) − α

1

(

y

) |

2

+ | α

2

(

x

) − α

2

(

y

) |

2

2

x

y

2

,

(7) and hence,

| α

i

(

x

) − α

i

(

y

) | ≤

2

x

y

‖ , ∀

i

=

1

,

2.

Proof. The proof can be obtained by direct computation like in [2, Lemma 2].

Lemma 2.2. Let x

=

x

+ λ

e

Hand y

=

y

+ µ

e

H. (a) If x

̸=

0and y

̸=

0, then we have

‖ v

(xi)

− v

(yi)

‖ ≤

1

x

‖ ‖

x

y

‖ ∀

i

=

1

,

2

,

(8)

where

v

(xi)

, v

y(i)are the spectral vectors of x and y, respectively.

(b) If either x

=

0or y

=

0, then we can choose

v

x(i),

v

y(i)such that the left hand side of inequality(8)is zero.

Proof. The proof is similar to [16, Lemma 3.2], so we omit it here.

Lemma 2.3. For any x

̸=

0

H, the following hold.

(a) If g

(

x

) = ‖

x

, we have g

(

x

)

h

=

x,h

x. (b) If g

(

x

) =

x

x, we have g

(

x

)

h

=

h

x

x,h

x3x.

Proof. (a) See Example 3.1(V) of [1].

(b) First, we compute that

g

(

x

+

h

) −

g

(

x

) =

x

+

h

x

+

h

‖ −

x

x

=

h

x

+

h

‖ −

 1

x

‖ −

1

x

+

h

·

x

=

h

x

+

h

‖ −

√ ⟨

x

+

h

,

x

+

h

⟩ − √

x

,

x

x

,

x

⟩ · √

x

+

h

,

x

+

h

⟩ ·

x

=

h

x

+

h

‖ −

2

x

,

h

⟩ + ⟨

h

,

h

√ ⟨

x

,

x

⟩ · √

x

+

h

,

x

+

h

⟩ ( √

x

+

h

,

x

+

h

⟩ + √

x

,

x

⟩ ) ·

x

=

h

x

‖ − ⟨

x

,

h

x

3x

+

o

( ‖

h

‖ ).

From the above, it is clear thatg

(

x

)

h

=

h

x

x,h

x3x.

Semismooth function, as introduced by Mifflin [17] for functionals and further extended by Qi and Sun [18] for vector- valued functions, is of particular interest due to the central role it plays in the superlinear convergence analysis of certain generalized Newton methods, see [18,19] and references therein. Given a mappingF

:

Rn

Rm, it is well known that ifF is strictly continuous (locally Lipschitz continuous), thenFis almost everywhere differentiable by Rademacher’s Theorem

— see [20] and [21, Sec. 9J]. In this case, the generalized Jacobian

F

(

x

)

ofFatx(in the Clarke sense) can be defined as the convex hull of theB-subdifferential

BF

(

x

)

, where

BF

(

x

) :=

 lim

xjx

F

(

xj

) |

Fis differentiable atxj

Rn

.

The notation

Bis adopted from [19]. In [21, Chap. 9], the case ofm

=

1 is considered and the notations ‘‘

∇ ¯

’’ and ‘‘

∂ ¯

’’ are used instead of, respectively, ‘‘

B’’ and ‘‘

’’. AssumeF

:

Rn

Rmis strictly continuous, thenFis said to be semismooth atx ifFis directionally differentiable atxand, for anyV

∈ ∂

F

(

x

+

h

)

andh

0, we have

F

(

x

+

h

) −

F

(

x

) −

Vh

=

o

( ‖

h

‖ ).

(9)

(4)

C.-Y. Yang et al. / Nonlinear Analysis 74 (2011) 5766–5783 5769

Moreover,Fis called

ρ

-order semismooth atx(0

< ρ < ∞

) ifFis semismooth atxand, for anyV

∈ ∂

F

(

x

+

h

)

andh

0, we have

F

(

x

+

h

) −

F

(

x

) −

Vh

=

O

( ‖

h

1+ρ

).

The Rademacher theorem does not hold in function spaces, see [22]. Hence, the aforementioned definitions of generalized Jacobian and semismoothness cannot be used in infinite-dimensional spaces. To overcome this difficulty, in the paper [22], so-called slanting functions and slant differentiability of operators in Banach spaces are proposed and used to formulate a concept of semismoothness in infinite-dimensional spaces. We shall introduce them as below. LetX

,

Y

H. A function F

:

X

Yis said to be directionally differentiable atxif the limit

δ

+F

(

x

;

h

) :=

lim

t0+

F

(

x

+

th

) −

F

(

x

)

t (10)

exists, where

δ

+F

(

x

;

h

)

is called the directional derivative ofFatxwith respect to the directionh. A functionF

:

X

Yis said to beB-differentiable atxif it is directionally differentiable atxand

lim

h0

F

(

x

+

h

) −

F

(

x

) − δ

+F

(

x

;

h

)

h

‖ =

0 (11)

in which we call

δ

+F

(

x

; · )

theB-derivative ofFatx. In finite-dimensional Euclidean spaces, Shapiro [23] shows that a locally Lipschitz continuous functionFisB-differentiable atxif and only if it is directionally differentiable atx. From(9)and(11) (also see [18]), it can be seen thatFis semismooth atxif and only ifFisB-differentiable (hence directionally differentiable) atxand, for eachV

∈ ∂

F

(

x

+

h

)

, there has

δ

+F

(

x

;

h

) −

Vh

=

o

( ‖

h

‖ ).

As mentioned earlier, these results do not hold in infinite-dimensional spaces. Therefore, the slant differentiability is introduced to circumvent this hurdle. In what follows, we state its definition.

Definition 2.1. LetDbe an open domain inXandL

(

X

,

Y

)

denote the set of all bounded linear operators fromXontoY.

(a) A functionF

:

D

X

Y is said to be slantly differentiable atx

Dif there exists a mappingf

:

D

L

(

X

,

Y

)

such that the family

{

f

(

x

+

h

) }

of bounded linear operators is uniformly bounded in the operator norm forhsufficiently small and

lim

h0

F

(

x

+

h

) −

F

(

x

) −

f

(

x

+

h

)

h

h

‖ =

0

.

(12)

The functionfis called a slanting function forFatx.

(b) A functionF

:

D

X

Yis said to be slantly differentiable in an open domainD0

Dif there exists a mapping f

:

D

L

(

X

,

Y

)

such thatfis a slanting function forFat everyx

D0. In this case,fis called a slanting function for FinD0.

Definition 2.2. Suppose thatf

:

D

L

(

X

,

Y

)

is a slanting function forFatx

DWe denote the set

SF

(

x

) :=

xlimkxf

(

xk

)

(13) and call it the slant derivative ofF associated withfatx

D. Note thatf

(

x

) ∈ ∂

SF

(

x

)

which says

SF

(

x

)

is always nonempty.

A functionFmay be slantly differentiable at all points ofD, but there is no common slanting function ofFat all points ofD. Moreover, a slantly differentiable functionFatxcan have infinitely many slanting functions atx. A slanting function fforFatxis a single-valued function, but not continuous in general. In addition, a continuous function is not necessarily slantly differentiable. For more details about slanting functions and slantly differentiability, please refer to [22].

Definition 2.3. A mappingF

:

X

Yis said to bes-semismooth atxif there is a slanting functionfforFin a neighborhood Nxofxsuch thatfand the associated slant derivative satisfy the following two conditions.

(a) limt0+f

(

x

+

th

)

hexists for everyh

Xand lim

h‖→0

lim

t0+f

(

x

+

th

)

h

f

(

x

+

h

)

h

h

‖ =

0

.

(b)f

(

x

+

h

)

h

Vh

=

o

( ‖

h

‖ )

for allV

∈ ∂

SF

(

x

+

h

)

.

We point it out that the functionFdefined inDefinition 2.3was called semismooth in [22]. However, we here rename it as ‘‘s-semismooth’’ because whenX

,

Y are both finite-dimensional spaces it does not reduce to the original definition introduced by Qi and Sun [18] in finite-dimensional spaces. The main key causing this is the limits in

SF

(

x

)

and

BF

(

x

)

are approached by different ways. In order to distinguish such difference, we hence use the term ‘‘s-semismooth’’ to convey concept of semismoothness in infinite-dimensional spaces.

(5)

3. Continuous properties offH

In this section, we show properties of continuity and (local) Lipschitz continuity offH. The arguments are straightforward by checking their definitions.

Proposition 3.1. Suppose x

=

x

+ λ

e

Hwith spectral values

α

1

(

x

)

,

α

2

(

x

)

and spectral vectors

v

(x1),

v

(x2). Let fHbe defined as in(4). Then, fHis continuous at x

Hif and only if f is continuous at

α

1

(

x

)

,

α

2

(

x

)

.

Proof.

( ⇒ )

This part of proof is similar to the argument of [2, Proposition 2(a)].

( ⇐ )

This direction of proof is also similar to [16, Proposition 2.2(a)], we omit it.

Proposition 3.2. Suppose x

=

x

+ λ

e

Hwith spectral values

α

1

(

x

)

,

α

2

(

x

)

and spectral vectors

v

(x1),

v

(x2). Let fHbe defined as in(4). Then, the following hold.

(a) fHis strictly continuous at x

Hif and only if f is strictly continuous at

α

1

(

x

)

,

α

2

(

x

)

. (b) fHis Lipschitz continuous (with respect to

‖ · ‖

) if and only if f is Lipschitz continuous.

Proof. (a)

( ⇐ )

Supposef is strictly continuous at

α

1

(

x

)

,

α

2

(

x

)

. Then, there exist

κ

i

>

0 and

δ

i

>

0 fori

=

1

,

2, such that

|

f

(ξ) −

f

(ζ) | ≤ κ

i

| ξ − ζ | ∀ ξ, ζ ∈ [ α

i

(

x

) − δ

i

, α

i

(

x

) + δ

i

]

i

=

1

,

2

.

Let

δ =

1

2min

{ δ

1

, δ

2

}

and for anyy

,

z

B

(

x

, δ)

, we have

fH

(

y

) −

fH

(

z

) = (

f

1

(

y

))v

(y1)

+

f

2

(

y

))v

y(2)

) − (

f

1

(

z

))v

z(1)

+

f

2

(

z

))v

(z2)

)

=

f

1

(

y

))(v

(y1)

− v

z(1)

) + (

f

1

(

y

)) −

f

1

(

z

)))v

z(1)

+

f

2

(

y

))(v

y(2)

− v

(z2)

) + (

f

2

(

y

)) −

f

2

(

z

)))v

(z2) (14) wherey

= α

1

(

y

)v

y(1)

+ α

2

(

y

)v

(y2)andz

= α

1

(

z

)v

z(1)

+ α

2

(

z

)v

z(2). ByLemmas 2.1and2.2and the similar argument in [2, Proposition 6(a)], the proof can be obtained.

( ⇒ )

This part of proof is quite simple and similar to [2, Proposition 6(a)], we omit it here.

(b) The argument of proof is similar to [2, Proposition 6(c)].

4. Differential properties offH

In this section, we show properties of directional differentiability, differentiability, continuous differentiability and B-differentiability offH. For simplicity, in the arguments we sometimes abbreviate

α

i

(

x

)

as

α

iwhen there is no ambiguity in the context. Note that, unlike in finite-dimensional second-order cone case [2],Propositions 4.1and4.2are proved by different approaches since the chain rule for directional differentiability in infinite-dimensional space does not hold in general, see [23].

Proposition 4.1. Suppose x

=

x

+ λ

e

Hwith spectral values

α

1

(

x

)

,

α

2

(

x

)

and spectral vectors

v

(x1),

v

(x2). Let fHbe defined as in(4). Then, fHis directionally differentiable at x

Hif and only if f is directionally differentiable at

α

1

(

x

)

,

α

2

(

x

)

.

Proof.

( ⇐ )

Supposefis directionally differentiable at

α

1

(

x

)

,

α

2

(

x

)

. Fixx

=

x

+ λ

e

Hand any directionh

=

h

+

le

H, we discuss two cases as below.

Case(i). Ifx

̸=

0, then we havefH

(

x

) =

f

1

(

x

))v

x(1)

+

f

2

(

x

))v

x(2)where

α

i

(

x

) = λ + ( −

1

)

i

x

and

v

x(i)

=

12

(

e

+ ( −

1

)

i xx

)

fori

=

1

,

2. Nowx

+

th

= (

x

+

th

) + (λ +

tl

)

ewith spectral values

α

i

(

x

+

th

) = λ +

tl

+ ( −

1

)

i

x

+

th

and spectral vectors

v

(xi+)th

=

1

2

(

e

+ ( −

1

)

i xx++thth

)

fori

=

1

,

2. We consider Eq.(14)again in which replacingywithx

+

th, then we have fH

(

x

+

th

) −

fH

(

x

) =

f

1

(

x

+

th

))(v

(x1+)th

− v

x(1)

) + (

f

1

(

x

+

th

)) −

f

1

(

x

)))v

x(1)

+

f

2

(

x

+

th

))(v

x(2+)th

− v

(x2)

) + (

f

2

(

x

+

th

)) −

f

2

(

x

)))v

x(2)

.

(15) Because the process of checking argument is similar to [2, Proposition 3], we only present the result here.

By denoting

a

˜ =

f

2

(

x

)) −

f

1

(

x

)) α

2

(

x

) − α

1

(

x

) ,

b

˜ = δ

+f

2

(

x

) ;

k2

) + δ

+f

1

(

x

) ;

k1

)

2

,

(16)

c

˜ = δ

+f

2

(

x

) ;

k2

) − δ

+f

1

(

x

) ;

k1

)

2

,

(6)

C.-Y. Yang et al. / Nonlinear Analysis 74 (2011) 5766–5783 5771

whereki

= ⟨

h

,

e

⟩ + ( −

1

)

ixx,hfori

=

1

,

2, we can write the expression of

δ

+fH

(

x

;

h

)

as

δ

+fH

(

x

;

h

) = ˜

a

h

− ⟨

h

,

e

e

− ⟨

x

,

h

x

2x

+ ˜

be

+ ˜

c x

x

‖ .

(17)

Case(ii). If x

=

0, we compute the directional derivative

δ

+fH

(

x

;

h

)

atx

H for any directionhby definition. Let h

=

h

+

le

Hwithh

∈ ⟨

e

andl

R. We discuss two subcases.

Subcase(a). Ifh

̸=

0, from the spectral decomposition, we choose

v

x(i)

=

1

2

(

e

+ ( −

1

)

i hh

)

fori

=

1

,

2 such that fH

(

x

+

th

) =

f

(λ +

th1

)v

x(1)

+

f

(λ +

th2

)v

x(2)

fH

(

x

) =

f

(λ)v

(x1)

+

f

(λ)v

(x2)

wherehi

=

l

+ ( −

1

)

i

h

fori

=

1

,

2. Now, we compute lim

t0+

fH

(

x

+

th

) −

fH

(

x

)

t

=

lim

t0+

f

(λ +

th1

) −

f

(λ)

t

v

(x1)

+

lim

t0+

f

(λ +

th2

) −

f

(λ)

t

v

(x2)

= δ

+f

(λ ;

l

− ‖

h

‖ )v

x(1)

+ δ

+f

(λ ;

l

+ ‖

h

‖ )v

(x2)

.

(18) This shows that

δ

+fH

(

x

;

h

)

exists under this subcase.

Subcase(b). Ifh

=

0, we choose

v

(xi)

=

1

2

(

e

+ ( −

1

)

i

w)

for any

w ∈

Hwith

‖ w ‖ =

1. Analogous to(18), we have lim

t0+

fH

(

x

+

th

) −

fH

(

x

)

t

=

lim

t0+

f

(λ +

tl

) −

f

(λ)

t

v

(x1)

+

lim

t0+

f

(λ +

tl

) −

f

(λ)

t

v

(x2)

= δ

+f

(λ ;

l

)v

(x1)

+ δ

+f

(λ ;

l

)v

x(2)

.

(19) Hence,

δ

+fH

(

x

;

h

)

exists under this subcase.

From all the above, we have proved thatfHis directionally differentiable atx

Hwhenx

=

0 and its directional derivative

δ

+fH

(

x

;

h

)

is either in form of(18)or(19).

( ⇒ )

SupposefH is directionally differentiable atx

H, we will prove thatf is directionally differentiable at

α

1,

α

2. For

α

1

Rand any directiond1

R, leth

=

d1

v

x(1)

+

0

v

(x2)wherex

= α

1

v

(x1)

+ α

2

v

x(2). Then,x

+

th

= (α

1

+

td1

)v

(x1)

+ α

2

v

x(2)

and

fH

(

x

+

th

) −

fH

(

x

)

t

=

f

1

+

td1

) −

f

1

)

t

v

(x1)

.

SincefHis directionally differentiable atx, the above equation implies that

δ

+f

1

;

d1

) =

lim

t0+

f

1

+

td1

) −

f

1

)

t exists

.

This meansf is directionally differentiable at

α

1. Similarly, it can be verified thatf is also directionally differentiable at

α

2.

Proposition 4.2. Suppose x

=

x

+ λ

e

Hwith spectral values

α

1

(

x

)

,

α

2

(

x

)

and spectral vectors

v

x(1),

v

x(2). Let fH be defined as in(4). Then, fH is differentiable at x

Hif and only if f is differentiable at

α

1

(

x

)

,

α

2

(

x

)

.

Proof.

( ⇐ )

Supposefis differentiable at

α

1

, α

2. Fixx

=

x

+ λ

e

Handh

=

h

+

le

H, we discuss two cases as below.

Case(i). Ifx

̸=

0, then we havefH

(

x

) =

f

1

)v

(x1)

+

f

2

)v

(x2)where

α

i

= λ + ( −

1

)

i

x

and

v

(xi)

=

1

2

(

e

+ ( −

1

)

i xx

)

for i

=

1

,

2. By usingLemma 2.3and the chain rule and product rule for differentiation, the argument is similar to [2, Proposition 4] so we omit the process and present the result as following. Denoting

a

=

f

2

) −

f

1

)

α

2

− α

1

,

b

=

f

2

) +

f

1

)

2

,

c

=

f

2

) −

f

1

)

2

.

(20)

We can write the expression of

(

fH

)

(

x

)

has

(

fH

)

(

x

)

h

=

ah

+ (

b

a

)

h

,

e

e

+ ⟨

x

,

h

x

2x

+

c

x

‖ ( ⟨

x

,

h

e

+ ⟨

h

,

e

x

).

(21) Case(ii). The proof is identical to that of Case (ii) inProposition 4.1, but withthreplaced byh. We omit it and only present the formula of

(

fH

)

(

x

)

has below. Ifx

=

0, then

(

fH

)

(

x

)

h

=

f

(λ)

h

.

(22)

( ⇒ )

This part of proof is similar to [16, Proposition 2.2(c)].

Références

Documents relatifs

If a metric σ lies in GF (M ) but has some rank one cusps, its bending measured geodesic lamination λ has some compact leaves with a weight equal to π.. Thus we get that b GF is

From one side, we prove a Lipschitz regularity result for vector-valued functions which are almost-minimizers of a two-phase functional with variable coefficients (Theorem 1.2).. , u

The available theoretical estimates of the discrete LBB conditions may be very pessimistic compared to what is really observed in the numerical algorithms, but the fact that the

Therefore, also from this point of view, it is interesting to consider convolution processes obtained by taking f to be a sample path of a continuous Gaussian process with

Orlicz-Sobolev and Marcinkiewicz spaces and about the behaviour of norms in these spaces under rearrangements of functions.. Here, A is an N-function and A is its

Our main result is that if the obstacle function ?p has distributional derivatives of order one which are of class EP for some p &amp;#x3E; n and is Holder

L’accès aux archives de la revue « Annali della Scuola Normale Superiore di Pisa, Classe di Scienze » ( http://www.sns.it/it/edizioni/riviste/annaliscienze/ ) implique l’accord

In order to obtain this local minorization, we used Klingen- berg’s minorization for ig0(M) in terms of an upper bound for the sectional curvature of (M, go) and the