+ All Categories
Home > Documents > Image space analysis for variational inequalities with cone constraints and applications to traffic...

Image space analysis for variational inequalities with cone constraints and applications to traffic...

Date post: 25-Aug-2016
Category:
Upload: jun-li
View: 214 times
Download: 2 times
Share this document with a friend
18
SCIENCE CHINA Mathematics . ARTICLES . April 2012 Vol. 55 No. 4: 851–868 doi: 10.1007/s11425-011-4287-5 c Science China Press and Springer-Verlag Berlin Heidelberg 2011 math.scichina.com www.springerlink.com Image space analysis for variational inequalities with cone constraints and applications to traffic equilibria LI Jun 1, & HUANG NanJing 2 1 College of Mathematics and Information, China West Normal University, Nanchong 637009, China; 2 Department of Mathematics, Sichuan University, Chengdu 610064, China Email: junli1026@163.com, [email protected] Received July 25, 2010; accepted May 2, 2011; published online September 6, 2011 Abstract In this paper, the image space analysis (for short, ISA) is employed to investigate variational in- equalities (for short, VI) with cone constraints. Linear separation for VI with cone constraints is characterized by using the normal cone to a regularization of the image, and saddle points of the generalized Lagrangian func- tion. Lagrangian-type necessary and sufficient optimality conditions for VI with cone constraints are presented by using a separation theorem. Gap functions and weak sharpness for VI with cone constraints are also investi- gated. Finally, the obtained results are applied to standard and time-dependent traffic equilibria introduced by Daniele, Maugeri and Oettli. Keywords variational inequality, image space analysis, linear separation, necessary and sufficient optimality condition, traffic equilibrium MSC(2010) 90C33, 49J40 Citation: Li J, Huang N J. Image space analysis for variational inequalities with cone constraints and applications to traffic equilibria. Sci China Math, 2012, 55(4): 851–868, doi: 10.1007/s11425-011-4287-5 1 Introduction The image of a constrained extremum problem was developed in [15] by exploiting previous results on theorems of the alternative [14]. The image space analysis (for short, ISA) is a powerful tool and a unifying scheme for studying both variational inequalities and optimization problems (for short, VI and OP, respectively). This approach can be applied to any kind of problem that can be expressed under the form of the impossibility of a parametric system. The impossibility of such a system is reduced to the disjunction of two suitable subsets of the image space. The disjunction between the two suitable subsets is proved by showing that they lie in two disjoint level sets of a separating functional. Recently, there has been an increasing interest in the ISA of vector variational inequalities and vector optimization problems (for short, VVI and VOP, respectively) [16, 17, 26, 31, 32]. The first step in this direction was made by Giannessi, Mastroeni and Pellegrini [19]. In [34, 35], Moldovan and Pellegrini employed the ISA to present necessary and sufficient optimality conditions for constrained extremum problems in finite-dimensional spaces. In [29, 30], Madani, Mastroeni and Moldovan dealt with ISA for constrained extremum problems having infinite-dimensional image and established sufficient and neces- sary optimality conditions for these problems. Some other related works on the ISA for VI, OP, VVI and VOP, can be found in [18,21]. Corresponding author
Transcript

SCIENCE CHINAMathematics

. ARTICLES . April 2012 Vol. 55 No. 4: 851–868

doi: 10.1007/s11425-011-4287-5

c© Science China Press and Springer-Verlag Berlin Heidelberg 2011 math.scichina.com www.springerlink.com

Image space analysis for variational inequalities withcone constraints and applications to traffic equilibria

LI Jun1,∗ & HUANG NanJing2

1College of Mathematics and Information, China West Normal University, Nanchong 637009, China;2Department of Mathematics, Sichuan University, Chengdu 610064, China

Email: [email protected], [email protected]

Received July 25, 2010; accepted May 2, 2011; published online September 6, 2011

Abstract In this paper, the image space analysis (for short, ISA) is employed to investigate variational in-

equalities (for short, VI) with cone constraints. Linear separation for VI with cone constraints is characterized

by using the normal cone to a regularization of the image, and saddle points of the generalized Lagrangian func-

tion. Lagrangian-type necessary and sufficient optimality conditions for VI with cone constraints are presented

by using a separation theorem. Gap functions and weak sharpness for VI with cone constraints are also investi-

gated. Finally, the obtained results are applied to standard and time-dependent traffic equilibria introduced by

Daniele, Maugeri and Oettli.

Keywords variational inequality, image space analysis, linear separation, necessary and sufficient optimality

condition, traffic equilibrium

MSC(2010) 90C33, 49J40

Citation: Li J, Huang N J. Image space analysis for variational inequalities with cone constraints and applications

to traffic equilibria. Sci China Math, 2012, 55(4): 851–868, doi: 10.1007/s11425-011-4287-5

1 Introduction

The image of a constrained extremum problem was developed in [15] by exploiting previous results on

theorems of the alternative [14]. The image space analysis (for short, ISA) is a powerful tool and a

unifying scheme for studying both variational inequalities and optimization problems (for short, VI and

OP, respectively). This approach can be applied to any kind of problem that can be expressed under the

form of the impossibility of a parametric system. The impossibility of such a system is reduced to the

disjunction of two suitable subsets of the image space. The disjunction between the two suitable subsets

is proved by showing that they lie in two disjoint level sets of a separating functional.

Recently, there has been an increasing interest in the ISA of vector variational inequalities and vector

optimization problems (for short, VVI and VOP, respectively) [16, 17, 26, 31, 32]. The first step in this

direction was made by Giannessi, Mastroeni and Pellegrini [19]. In [34, 35], Moldovan and Pellegrini

employed the ISA to present necessary and sufficient optimality conditions for constrained extremum

problems in finite-dimensional spaces. In [29, 30], Madani, Mastroeni and Moldovan dealt with ISA for

constrained extremum problems having infinite-dimensional image and established sufficient and neces-

sary optimality conditions for these problems. Some other related works on the ISA for VI, OP, VVI and

VOP, can be found in [18, 21].

∗Corresponding author

852 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

The gap function for VI was first introduced by Auslender [1]. The notion of weak sharp minima

is a generalization of the notion of sharp minima due to Polyak to include the possibility of a non-

unique solution set. The terminology weak sharp minima was introduced by Ferris [13]. The primary

motivations for this study are the impact this notion has on error bounds in convex programming (see,

for example, [12, 37, 39]), on the sensitivity analysis of optimization problems and on the convergence

analysis of a wide range of optimization algorithms.

Hitherto, almost all the models for predicting urban traffic equilibria are based on the well-known

Wardrop principle introduced in 1952. Later on, it was proved by Smith in 1979 and Dafermos in 1980 that

a traffic network can be modelled by means of a VI. Solutions to VI correspond to the equilibrium flows

on the traffic network (see, for example, [36]). In [9, 10], Daniele, Maugeri and Oettli introduced traffic

equilibria in the case where the numerical data of the traffic network depend on time and characterized it

by a suitably generalized form of Wardrop principle in terms of a VI. In order to derive the Lagrangian-

type necessary and sufficient optimality conditions for standard and time-dependent traffic equilibria

introduced in [9, 10], we shall carry on the ISA of VI with cone constraints by following the analysis

developed in [31]. Since the convex subset of the image space and the regularization of the image of

time-dependent traffic equilibria, for instance, H and E (see Section 6), have empty interior, the classic

separation theorem for convex sets cannot be applied. To this aim, we shall introduce the notion of

quasi-relative interior and apply separation theorems related to this notion.

An outline of the paper is as follows. We begin in Section 2 by recalling the preliminary results

and presenting the problem formulation of VI with cone constraints. In Section 3, we carry on the

ISA of VI with cone constraints. We characterize in Section 4 the linear separation for VI with cone

constraints in terms of the normal cone to a regularization of the image, the saddle points of generalized

Lagrangian function and derive the Lagrangian-type necessary and sufficient optimality conditions for

VI with cone constraints by utilizing a separation theorem. We also investigate the gap functions for

VI with cone constraints and prove that the solutions set of VI with cone constraints is weakly sharp

for certain functions in Section 5. As applications, we investigate the Lagrangian-type necessary and

sufficient optimality conditions for standard and time-dependent traffic equilibria introduced by Daniele,

Maugeri and Oettli.

2 Preliminaries

Let Rk be the k-dimensional Euclidean space, where k is a given positive integer. Denote by Rk+ := {x :=

(x1, . . . , xk)� : xi � 0, i = 1, . . . , k}, where the superscript � denotes the transposition. Let E ⊆ R

k.

For x ∈ Rk, denote by d(x,E) the distance from x to E, i.e., d(x,E) := infe∈E ‖x − e‖, where the ‖ · ‖

denotes the Euclidean norm. A nonempty subset P of Rk is said to be a cone iff λP ⊆ P for all λ � 0.

P is said to be a convex cone iff P is a cone and P + P = P . P is called a pointed cone iff P is a cone

and P ∩ (−P ) = {0}. The dual cone (or positive polar cone) of a nonempty subset P of Rk is given by

P ∗ := {z ∈ Rk : 〈z, x〉 � 0, ∀x ∈ P},

where 〈·, ·〉 denotes the inner product. Assume that P is a convex cone. Then, P ∗∗ = clP and so P ∗∗ = P

if P is closed, where the cl denotes the closure (see, for example, [22, 23, 38]).

Let P ⊆ Rk be a closed convex cone. We will use the following notations:

x �P y ⇔ x− y ∈ P, ∀x, y ∈ Rk,

and

x >P y ⇔ x− y ∈ intP, ∀x, y ∈ Rk,

where int denotes the interior.

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 853

Let P ⊆ Rk be a closed convex cone, Ω ⊆ R

l be convex and h : Ω → Rk. Recall that h is said to be

P -convex on Ω iff for any x, y ∈ Ω,

(1− t)h(x) + th(y) �P h((1− t)x+ ty), ∀ t ∈ (0, 1).

h is called convex-like on Ω with respect to the cone P iff the set h(Ω) + P is convex. Clearly, if h is

P -convex on Ω, then it is convex-like on Ω with respect to the cone P .

Let K be a nonempty subset of Rk. The normal cone NK(x0) to K at x0 ∈ K is defined by

NK(x0) := {z ∈ Rk : 〈z, y − x0〉 � 0, ∀ y ∈ K}.

It is clear that if x0 = 0 ∈ K, then NK(x0) = −K∗. If x0 ∈ intK, then NK(x0) = {0}. That is to say, if

NK(x0) �= {0}, then x0 ∈ bdK, where bd denotes the boundary (see, for example, [22, 23, 38]).

In this paper, without other specifications, let D ⊆ Rn be a closed convex cone with intD �= ∅ and let

X ⊆ Rm be convex. Let F := (F1, F2, . . . , Fm)� : Rm → R

m and g := (g1, g2, . . . , gn)� : X → R

n. We

consider the following VI with cone constraints:

find x ∈ K such that 〈F (x), y − x〉 � 0, ∀ y ∈ K,

where K := {x ∈ X : g(x) �D 0}. We always suppose that K is nonempty. It is obvious that x ∈ K

solves VI with cone constraints iff −F (x) ∈ NK(x).

Notice that for suitable choices of D, the inequality g(x) �D 0 collapses to some equalities and

inequalities. As a consequence, VI with cone constraints includes VI with equalities and inequalities

constraint as a special case.

We also need the following lemmas.

Lemma 2.1 (See [28]). Let h : X → Rn be D-convex on X. Then the function x �→ 〈θ, h(x)〉 is convex

on X, where θ ∈ D∗.

The following result is well-known (see, for example, [24]).

Lemma 2.2. Let intD �= ∅. Then the following hold:

(i) Let θ >D∗ 0. Then 〈θ, v〉 > 0 for all v ∈ D\{0};(ii) Let θ >D 0. Then 〈θ, v〉 > 0 for all v ∈ D∗\{0}.

Lemma 2.3. Let X ⊆ Rm be convex. Let f : Rm → R and h : X → R

n. If f is affine and −h is

D-convex on X, then the function (f,−h) : X → R× Rn is convex-like with respect to R+ ×D.

Proof. We need to prove that the set (f,−h)(X) + R+ ×D is convex. Let t ∈ (0, 1), xi ∈ X,αi ∈ R+

and di ∈ D(i = 1, 2) be given. Set x := (1 − t)x1 + tx2, α := (1 − t)α1 + tα2 and d := (1 − t)d1 + td2.

Since X,R+ and D are convex, x ∈ X,α ∈ R+ and d ∈ D. Again, since −h is D-convex on X , one has

−(1− t)h(x1)− th(x2) ∈ −h(x) +D. Consequently,

(1− t)[(f,−h)(x1) + (α1, d1)] + t[(f,−h)(x2) + (α2, d2)]

= ((1 − t)f(x1) + tf(x2) + (1− t)α1 + tα2,−(1− t)h(x1)− th(x2) + (1− t)d1 + td2)

= (f(x) + α,−(1− t)h(x1)− th(x2) + d) (from the affinity of f)

= (f(x),−(1 − t)h(x1)− th(x2)) + (α, d)

∈ (f(x),−h(x)) + R+ ×D (since D is a convex cone)

= (f,−h)(x) + R+ ×D

⊆ (f,−h)(X) + R+ ×D,

which establishes the convexity of (f,−h)(X) + R+ ×D.

3 ISA for VI with cone constraints

In this section, we shall carry on the ISA for VI with cone constraints.

854 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

Observe that, x ∈ K solves VI with cone constraints iff the system (in the unknown y):

⎧⎪⎪⎨

⎪⎪⎩

〈F (x), x − y〉 > 0,

g(y) �D 0,

y ∈ X

(3.1)

is impossible. We can associate VI with cone constraints with the following sets:

H := {(u, v) ∈ R× Rn : u > 0, v �D 0},

K(x) := {(u, v) ∈ R× Rn : u = 〈F (x), x− y〉, v = g(y), y ∈ X}.

The set K(x) is called the image of VI with cone constraints at x ∈ K, while the space R × Rn is the

image space associated with VI with cone constraints at x ∈ K.

It is easily seen that the following holds.

Proposition 3.1. Let x ∈ K. System (3.1) is impossible iff

H ∩K(x) = ∅. (3.2)

Consequently, x ∈ K is a solution to VI with cone constraints iff (3.2) is true.

Since D ⊆ Rn is a closed convex cone, we have 0 ∈ D. Set Hu := {(u, v) ∈ R× R

n : u > 0, v = 0}.As pointed out in [17–19, 32], to prove directly whether or not (3.2) holds is generally too difficult.

Therefore, in order to show such a disjunction, one must prove that the two sets, or the set H and

an extension of the image depending on H, admit a separating hyperplane. In general, the image of a

generalized system is not convex even when the functions involved enjoy some convexity properties. To

overcome this difficulty, similar to that in [18, 19, 32], we introduce a regularization of the image K(x),

namely, the extension with respect to the cone clH, denoted by E :

E := K(x)− clH = {(u, v) ∈ R× Rn : u � 〈F (x), x− y〉, v �D g(y), y ∈ X}.

Remark 3.1. As a direct consequence of the definition of E , we have E = −[(f,−g)(X) + R+ ×D],

where f(x) := 〈F (x), x− x〉. In [8,25,33], a constraint qualification involving the tangent cone to the set

E was employed to investigate infinite-dimensional duality.

The following proposition is useful in deriving the linear separation and the Lagrangian-type necessary

and sufficient optimality conditions for VI with cone constraints.

Proposition 3.2. Let x ∈ K. The following statements are true :

(i) If either (f,−g) is convex-like on X with respect to R+ × D or −g is D-convex on X, where

f(x) := 〈F (x), x − x〉, then the set E is convex ; if g is continuous and X is compact, then the set E is

closed ;

(ii) System (3.1) is impossible, or (3.2) holds iff

H ∩ E = ∅, (3.3)

iff

Hu ∩ E = ∅. (3.4)

Proof. (i) If (f,−g) is convex-like on X with respect to R+ ×D, where f(x) := 〈F (x), x − x〉, then it

follows from Remark 3.1 that the set E is convex.

Suppose that −g isD-convex onX . Since f is affine, from Lemma 2.3 one has that (f,−g) is convex-likeon X with respect to R+ ×D. Therefore, the set E is convex.

Notice that f(x) = 〈F (x), x − x〉 is continuous and R+ ×D is compact. If g is continuous and X is

compact, then from Remark 3.1 we know that the set E is closed.

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 855

(ii) Since K(x) ⊆ E , it suffices to prove that (3.2) implies (3.3) holds. Let System (3.2) hold. Suppose

to the contrary that H ∩ E �= ∅. Then there is (u, v) ∈ H ∩ E , i.e., there exists y ∈ X such that

u � 〈F (x), x− y〉, v �D g(y), and u > 0, v �D 0.

Therefore, one has

0 < 〈F (x), x− y〉, 0 �D g(y),

which contradicts the assumption H ∩K(x) = ∅.We now establish the equivalence between (3.3) and (3.4). We only need to prove that (3.4) implies

(3.3) since Hu ⊆ H. The proof is similar to that in [18]. For the completeness, we include it here. Assume

that (3.4) holds. Suppose to the contrary that there is (u, v) ∈ H ∩ E . Then (0, v) ∈ clH. Since

E − clH = K(x)− (clH+ clH) = K(x)− clH = E ,

one has (u, v)− (0, v) = (u, 0) ∈ E , which is a contradiction since (u, 0) ∈ Hu.

4 Linear separation and Lagrangian-type optimality conditions for VI with

cone constraints

As it is well-known, linear separation [7, 11, 18, 31] is closely related to the Lagrangian type optimality

conditions. In this section, by following the analysis developed in [31], we shall characterize the linear

separation in terms of the normal cone to E , a saddle point condition of generalized Lagrangian function

associated with the system (3.1) and present the Lagrangian-type necessary and sufficient optimality

conditions for VI with cone constraints.

Definition 4.1. Let x ∈ K. The sets H and K(x) admit a linear separation iff there exists (λ, θ) ∈R+ ×D∗, with (λ, θ) �= (0, 0), such that

λ〈F (x), x − y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X. (4.1)

Let

B := {(λ, θ) ∈ R+ ×D∗ : ‖(λ, θ)‖ = 1}.Then B is a compact subset of R+ ×D∗.

It is easy to see that the following conclusion holds.

Proposition 4.1. Let x ∈ K. Then the sets H and K(x) admit a linear separation iff there exists

(λ, θ) ∈ B, such that (4.1) holds.

We first characterize the linear separation of the sets H and E by using the normal cone to E at zero.

Theorem 4.1. Let x ∈ K. Then the sets H and K(x) admit a linear separation iff there exists

(λ, θ) ∈ NE(0, 0), or equivalently, (λ, θ) ∈ −E∗, with (λ, θ) �= (0, 0).

Proof. It is obvious that (0, 0) ∈ E and so (λ, θ) ∈ NE(0, 0) iff (λ, θ) ∈ −E∗.Necessity. Suppose that the sets H and K(x) admit a linear separation. Then there exists (λ, θ) ∈

R+ ×D∗, with (λ, θ) �= (0, 0), such that

λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X.

As a consequence,

−λf(y)− λα+ 〈θ, g(y)〉 − 〈θ, d〉 � 0, ∀ (α, d, y) ∈ R+ ×D ×X,

where f(y) := 〈F (x), y − x〉, or equivalently,

〈(λ, θ),−[(f,−g)(y) + (α, d)]〉 � 0, ∀ (α, d, y) ∈ R+ ×D ×X.

856 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

From Remark 3.1, the above inequality implies that (λ, θ) ∈ NE(0, 0).Sufficiency. Let (λ, θ) ∈ NE(0, 0), with (λ, θ) �= (0, 0). Then it follows from Remark 3.1 that

〈(λ, θ),−[(f,−g)(y) + (α, d)]〉 � 0, ∀ (α, d, y) ∈ R+ ×D ×X,

i.e.,

λ〈F (x), x− y〉 − λα+ 〈θ, g(y)〉 − 〈θ, d〉 � 0, ∀ (α, d, y) ∈ R+ ×D ×X. (4.2)

Setting (α, d) := (0, 0) in (4.2) leads to (4.1). Next, we show that (λ, θ) ∈ R+ ×D∗.If λ < 0, then letting (d, y) := (0, x) in (4.2) allows that

〈θ, g(x)〉 � λα, ∀α ∈ R+,

which is a contradiction.

Letting (α, y) := (0, x) in (4.2) allows that

〈θ, g(x)〉 � 〈θ, d〉, ∀ d ∈ D, (4.3)

and so 〈θ, g(x)〉 � 0. If θ �∈ D∗, then there is d0 ∈ D such that 〈θ, d0〉 < 0. Since D is a cone, one has

td0 ∈ D for all t � 0 and hence

〈θ, td0〉 = t〈θ, d0〉 → −∞, as t→ +∞,

a contradiction with (4.3). Consequently, θ ∈ D∗ and the sets H and K(x) admit a linear separation.

From Theorem 4.1, we have the following:

Remark 4.1. The sets H and K(x) admit a linear separation iff (0, 0) �∈ qiE , where qiE denotes the

quasi interior of E . For more details on the quasi interior, please see Subsection 6.2.

Let x ∈ K. Consider the generalized Lagrangian function associated with the system (3.1), defined by

L : R+ ×D∗ ×X → R,

L(x;λ, θ, y) := λ〈F (x), y − x〉 − 〈θ, g(y)〉, ∀ (λ, θ, y) ∈ R+ ×D∗ ×X.

If λ ≡ 1, then the function (θ, y) �→ L(x; 1, θ, y) given above is the Lagrangian function associated with

the system (3.1).

Definition 4.2. The point (λ, θ, x) ∈ R+ × D∗ × X is said to be a saddle point of the generalized

Lagrangian function L(x;λ, θ, y) on R+ ×D∗ ×X iff the following inequalities hold :

L(x;λ, θ, x) � L(x; λ, θ, x) � L(x; λ, θ, y), ∀ (λ, θ, y) ∈ R+ ×D∗ ×X.

If (λ, θ, x) ∈ R+ × D∗ × X is a saddle point of the generalized Lagrangian function L(x;λ, θ, y) on

R+ ×D∗ ×X , then

L(x; λ, θ, x) = infy∈K

L(x; λ, θ, y) = sup(λ,θ)∈R+×D∗

L(x;λ, θ, x).

Consequently, the saddle point (λ, θ, x) ∈ R+ × D∗ × X can be characterized by suitable minimax

problems [38],

L(x; λ, θ, x) = sup(λ,θ)∈R+×D∗

infy∈K

L(x;λ, θ, y) = infy∈K

sup(λ,θ)∈R+×D∗

L(x;λ, θ, y).

We next characterize the linear separation for VI with cone constraints by using the saddle point

conditions of the generalized Lagrangian function.

Theorem 4.2. Let x ∈ K. Then the sets H and K(x) admit a linear separation iff there exists

(λ, θ) ∈ B such that the point (λ, θ, x) is a saddle point for L(x;λ, θ, y) on R+ ×D∗ ×X.

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 857

Proof. Necessity. Suppose that H and K(x) admit a linear separation. Then from Proposition 4.1,

there is (λ, θ) ∈ B such that (4.1) holds. Letting y := x in (4.1) allows 〈θ, g(x)〉 � 0. Since x ∈ K, one

has g(x) �D 0 and so 〈θ, g(x)〉 � 0. As a consequence, 〈θ, g(x)〉 = 0 and

0 = L(x; λ, θ, x) � L(x; λ, θ, y), ∀ y ∈ X.

Moreover, we have

L(x;λ, θ, x) = −〈θ, g(x)〉 � 0, ∀ (λ, θ) ∈ R+ ×D∗.

This proves that (λ, θ, x) is a saddle point for L(x;λ, θ, y) on R+ ×D∗ ×X .

Sufficiency. Let (λ, θ) ∈ B such that (λ, θ, x) is a saddle point for L(x;λ, θ, y) on R+ ×D∗ ×X . Then

−〈θ, g(x)〉 � −〈θ, g(x)〉 � λ〈F (x), y − x〉 − 〈θ, g(y)〉, ∀ (λ, θ, y) ∈ R+ ×D∗ ×X.

Letting θ := 0 in the first inequality leads to 〈θ, g(x)〉 = 0 and therefore (4.1) holds. From Proposition 4.1,

the sets H and K(x) admit a linear separation.

Linear separation for VI with cone constraints is characterized by using the Lagrangian type optimality

conditions as follows:

Theorem 4.3. Let x ∈ K. Then the sets H and K(x) admit a linear separation iff there is a vector

(λ, θ) ∈ B, such that :

(i) λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X ;

(ii) 〈θ, g(x)〉 = 0.

Proof. Necessity. Suppose that H and K(x) admit a linear separation. Then there exists (λ, θ) ∈ B

such that (4.1) holds. From the proof of necessity in Theorem 4.2, one has 〈θ, g(x)〉 = 0. Again by (4.1),

we obtain

λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X.

Sufficiency. Suppose that there is a vector (λ, θ) ∈ B such that (i) and (ii) hold. Then

λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X,

It follows from Proposition 4.1 that the conclusion is true.

Moreover, if X is an open convex set in Rm, then we have the following:

Theorem 4.4. Let X be an open convex set in Rm, g be differentiable and −g be D-convex on X.

Then (λ, θ, x) with (λ, θ) �= (0, 0) is a saddle point for L(x;λ, θ, y) on R+ ×D∗ ×X iff it is a solution to

the following system,

⎧⎪⎪⎪⎪⎨

⎪⎪⎪⎪⎩

λF (y) =

n∑

i=1

(θi∇gi(y)),

〈θ, g(y)〉 = 0,

g(y) �D 0, y ∈ X, (λ, θ) ∈ R+ ×D∗, with (λ, θ) �= (0, 0).

(4.4)

Proof. Necessity. Assume that (λ, θ, x) is a saddle point for L(x;λ, θ, y) on R+ ×D∗ ×X . Then

−〈θ, g(x)〉 � −〈θ, g(x)〉 � λ〈F (x), y − x〉 − 〈θ, g(y)〉, ∀ (λ, θ, y) ∈ R+ ×D∗ ×X. (4.5)

Since D is closed and convex, D∗∗ = D. We declare that g(x) �D 0, i.e., g(x) ∈ D. In fact, if g(x) �∈ D,

then there is θ∗ ∈ D∗ such that 〈θ∗, g(x)〉 < 0. Since D∗ is a cone, we have tθ∗ ∈ D∗ for all t � 0 and

thus

−〈tθ∗, g(x)〉 = −t〈θ∗, g(x)〉 → +∞, as t→ +∞,

858 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

which is a contradiction with the first inequality in (4.5).

Since θ ∈ D∗ and g(x) �D 0, one has 〈θ, g(x)〉 � 0. Setting θ := 0 in the first inequality in (4.5) leads

to 〈θ, g(x)〉 � 0 and therefore 〈θ, g(x)〉 = 0.

From the second inequality in (4.5), one has that x ∈ X is a global minimum point of L(x; λ, θ, y) onX . Since g is differentiable, each component of g, gi (i = 1, 2, . . . , n), is differentiable. Since X is an open

convex set in Rm, it follows that

0 = ∇yL(x; λ, θ, x) = λF (x)−n∑

i=1

θi∇gi(x),

i.e.,

λF (x) =

n∑

i=1

θi∇gi(x).

Sufficiency. Let (λ, θ, x) be a solution to the system (4.4). Since −g isD-convex onX , from Lemma 2.1,

one has that y �→ 〈λ,−g(y)〉 is convex. Since g is differentiable, the function y �→ L(x; λ, θ, y) is convexand differentiable. Now, the equality λF (x) =

∑ni=1 θi∇gi(x) implies that ∇yL(x; λ, θ, x) = 0 and thus

L(x; λ, θ, x) � L(x; λ, θ, y), ∀ y ∈ X.

Then, the complementarity condition 〈θ, g(x)〉 = 0 allows to

L(x;λ, θ, x) = −〈θ, g(x)〉 � −〈θ, g(x)〉 = L(x; λ, θ, x), ∀ (λ, θ) ∈ R+ ×D∗.

The proof is complete.

From Theorems 4.2 and 4.4, we have

Remark 4.2. Let X be an open convex set in Rm, g be differentiable and −g be D-convex on X .

Suppose that the sets H and K(x) admit a linear separation, or equivalently, there exists (λ, θ) ∈ B such

that the point (λ, θ, x) is a saddle point for L(x;λ, θ, y) on R+ ×D∗ ×X .

(i) If λ = 0, then θ �= 0 and

n∑

i=1

(θi∇gi(x)) = 0,

which implies that ∇gi (x) (i = 1, 2, . . . , n) are linearly dependent.

(ii) If θ �= 0 and λ �= 0, then

F (x) =

n∑

i=1

(θi

λ∇gi(x)

)

.

As a consequence, Lagrangian duality for VI with cone constraints can be derived. For further details,

please see [2, 3].

Next, we shall present the Lagrangian-type necessary and sufficient optimality conditions for VI with

cone constraints.

First, we obtain the following necessary optimality condition for VI with cone constraints via ISA.

Theorem 4.5. Assume that either (f,−g) is convex-like on X with respect to R+ ×D or −g is D-

convex on X, where f(x) := 〈F (x), x − x〉. If x ∈ K is a solution to VI with cone constraints, then the

sets H and K(x) admit a linear separation, or equivalently, there is a vector (λ, θ) ∈ B such that :

(i) λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X ;

(ii) 〈θ, g(x)〉 = 0,

where B = {(λ, θ) ∈ R+ ×D∗ : ‖(λ, θ)‖ = 1}.

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 859

Proof. Let x ∈ K. Clearly, the set H is convex and from (i) of Proposition 3.2, E is convex. If x ∈ K

is a solution of VI with cone constraints, then (ii) of Proposition 3.2 implies that

H ∩ E = ∅.

Therefore, by a separation theorem [38], there exists a vector (λ, θ) ∈ R × Rn, with (λ, θ) �= (0, 0), such

that

λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X, (4.6)

and

λu+ 〈θ, v〉 � 0, ∀ (u, v) ∈ H. (4.7)

It follows immediately from (4.7) that (λ, θ) ∈ R+ ×D∗. Since (λ, θ) �= (0, 0), without loss of generality,

we can suppose that (λ, θ) ∈ B. From Proposition 4.1, Inequality (4.6) implies that the sets H and K(x)

admit a linear separation. The left proof follows directly from Theorem 4.3.

If X is an open convex set in Rm, then from Theorems 4.2, 4.4 and 4.5 one has the following classical

Lagrangian-type necessary optimality condition for VI with cone constraints.

Corollary 4.1. Let X be an open convex set in Rm and let x ∈ K be a solution of VI with cone

constraints. Suppose that g is differentiable and −g is D-convex on X. Then the sets H and K(x) admit

a linear separation, or equivalently, there exists (λ, θ) ∈ B such that the following hold :

(i) λF (x) =∑n

i=1(θi∇gi(x));(ii) 〈θ, g(x)〉 = 0.

Under certain assumptions, we obtain the following sufficient optimality condition for VI with cone

constraints.

Theorem 4.6. Let x ∈ K. Assume that there is a vector (λ, θ) ∈ B such that λ〈F (x), x−y〉+〈θ, g(y)〉 �0, ∀ y ∈ X. If either λ > 0 or θ >D∗ 0 and g(y) �= 0 for all y ∈ K, then x solves VI with cone constraints.

Proof. Suppose that x is not a solution of VI with cone constraints. Then from Proposition 3.1, there

is a y0 ∈ X such that

〈F (x), x− y0〉 > 0 and g(y0) �D 0.

Since (λ, θ) ∈ B, it follows from the assumption that λ〈F (x), x − y0〉+ 〈θ, g(y0)〉 � 0. If either λ > 0 or

θ >D∗ 0 and g(y) �= 0 for all y ∈ K, then from Lemma 2.2(i) the above inequality is strict, contradicting

the assumption.

5 Gap functions and weak sharpness for VI with cone constraints

In this section, we investigate the gap functions for VI with cone constraints by following similar idea

to that in [31], and prove that the solutions set of VI with cone constraints is weakly sharp for certain

functions under strong monotonicity.

Definition 5.1. Let K be the domain of VI with cone constraints. A function ψ : K → R ∪ {+∞} is

said to be a gap function for VI with cone constraints iff it satisfies the following properties :

(1) ψ(x) � 0, ∀x ∈ K;

(2) ψ(x) = 0, iff x solves VI with cone constraints.

Definition 5.2. Let K be the domain of VI with cone constraints and ψ : K → R∪{+∞} be a function

derived from VI with cone constraints. Denote by S the solutions set of VI with cone constraints. We

say that the set S is weakly sharp for the function ψ on K iff there exists μ > 0 such that

d(x, S) � μψ(x), ∀x ∈ K.

860 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

The notion of weak sharpness is closely related to that of error bounds. If S = {x ∈ K : ψ(x) = 0}and there exists μ > 0 such that

d(x, S) � μψ(x), ∀x ∈ K,

then we say that an error bound holds for the set S with ψ and with respect to the set K. For more

details, please see [12, 37].

Let

ϕ(λ, θ, x) := supy∈X

[λ〈F (x), x − y〉+ 〈θ, g(y)〉], ∀ (λ, θ, x) ∈ B ×K.

Define φ : K → R ∪ {+∞} by

φ(x) := min(λ,θ)∈B

ϕ(λ, θ, x) = min(λ,θ)∈B

supy∈X

[λ〈F (x), x − y〉+ 〈θ, g(y)〉], ∀x ∈ K,

where B = {(λ, θ) ∈ R+ ×D∗ : ‖(λ, θ)‖ = 1} is as that in Section 4.

It is clear that the following holds.

Remark 5.1. The function (λ, θ) �→ ϕ(λ, θ, x), being the supremum of a collection of linear functions,

is a convex function. As a consequence, φ(x) = min(λ,θ)∈B ϕ(λ, θ, x) is the optimal value of a parametric

problem on a compact set, with a convex objective function.

Lemma 5.1. Let x ∈ K. Assume that the sets H and K(x) admit a linear separation.

(i) If the following system is possible :⎧⎪⎪⎨

⎪⎪⎩

〈F (x), x − y〉 > 0,

g(y) >D 0,

y ∈ X,

then we can suppose that λ > 0 in (4.1).

(ii) If there exists y ∈ X such that g(y) >D 0 (Slater condition), then we can suppose that λ > 0 in

(4.1).

Proof. Clearly, Statement (i) follows from (ii). We only need to prove Statement (ii). Assume that

the sets H and K(x) admit a linear separation. From Proposition 4.1, there is (λ, θ) ∈ B such that (4.1)

holds and hence

λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0. (5.1)

Suppose to the contrary that λ = 0 in (4.1). Then θ �= 0. Since g(y) >D 0, again from Lemma 2.2(ii) it

holds 0 < 〈θ, g(y)〉 = λ〈F (x), x− y〉+ 〈θ, g(y)〉, a contradiction with (5.1). �

Definition 5.3. A mapping T : X → X is said to be strongly monotone on X with modulus δ > 0 iff

for any x, y ∈ X,

〈T (x)− T (y), x− y〉 � δ‖x− y‖2.Under suitable assumptions, we prove that φ is a gap function for VI with cone constraints, and show

that the solutions set of VI with cone constraints is weakly sharp for the function

x �→ (ϕ(λ, θ, x)− 〈θ, g(x)〉) 12 ,

where (λ, θ) is some vector in B.

Theorem 5.1. Let S := {x ∈ K : φ(x) = 0} �= ∅. Assume that either (f,−g) is convex-like on X

with respect to R+ ×D or −g is D-convex on X, where f(x) = 〈F (x), x − x〉, and the following system

is possible :⎧⎪⎪⎨

⎪⎪⎩

〈F (x), x − y〉 > 0,

g(y) >D 0,

y ∈ X.

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 861

Then φ is a gap function for VI with cone constraints. Furthermore, if F is strongly monotone on X

with modulus δ > 0, then there is (λ, θ) ∈ B, with λ > 0, such that S = {x}, the solutions set of VI with

cone constraints, is weakly sharp for the function x �→ (ϕ(λ, θ, x)− 〈θ, g(x)〉) 12 on K.

Proof. We first show that φ is a gap function for VI with cone constraints. Let x ∈ K be given. Then

x ∈ X , g(x) �D 0 and so

λ〈F (x), x − x〉+ 〈θ, g(x)〉 = 〈θ, g(x)〉 � 0, ∀ (λ, θ) ∈ B,

which yields φ(x) � 0, ∀x ∈ K.

Suppose that x solves VI with cone constraints. Then it follows from Theorem 4.5 that the sets H and

K(x) admit a linear separation. From Theorem 4.2, there exists (λ, θ) ∈ B such that the point (λ, θ, x)

is a saddle point for L(x;λ, θ, y) on R+ × D∗ × X and L(x; λ, θ, x) = 0 (see the proof of necessity in

Theorem 4.2). It follows that

min(λ,θ)∈R+×D∗

supy∈X

[λ〈F (x), x− y〉+ 〈θ, g(y)〉] = min(λ,θ)∈R+×D∗

supy∈X

[−L(x;λ, θ, y)]

= −L(x; λ, θ, x) = 0, (5.2)

which implies that φ(x) = 0 since (λ, θ) ∈ B.

Let φ(x) = 0. Then there exists (λ, θ) ∈ B such that

λ〈F (x), x− y〉+ 〈θ, g(y)〉 � 0, ∀ y ∈ X.

From Lemma 5.1, we have λ > 0. It thus follows from Theorem 4.6 that x solves VI with cone constraints.

Assume that F is strongly monotone on X with modulus δ > 0. Denote by S = {x}. Again from

Theorem 4.5, the sets H and K(x) admit a linear separation, or equivalently, there is a vector (λ, θ) ∈ B

such that

〈θ, g(x)〉 = 0 (5.3)

and

λ〈F (x), x− x〉 � 〈θ, g(x)〉, ∀x ∈ X. (5.4)

From Lemma 5.1, λ > 0. Since F is strongly monotone on X with modulus δ > 0, it follows that

ϕ(λ, θ, x) = supy∈X

[λ〈F (x), x − y〉+ 〈θ, g(y)〉]

� λ〈F (x), x − x〉+ 〈θ, g(x)〉� λ〈F (x), x− x〉+ λδ‖x− x‖2 + 〈θ, g(x)〉� λδ‖x− x‖2 + 〈θ, g(x)〉, ∀x ∈ K,

and consequently,

d(x, S) = ‖x− x‖ � (ϕ(λ, θ, x)− 〈θ, g(x)〉) 12

(λδ)12

, ∀x ∈ K. (5.5)

The proof is complete.

Remark 5.2. Let r(x) := (ϕ(λ, θ, x) − 〈θ, g(x)〉) 12 . From (5.3), we have 〈θ, g(x)〉 = 0 and from (5.2)

and (5.4), we have ϕ(λ, θ, x) = supy∈X [λ〈F (x), x − y〉 + 〈θ, g(y)〉] = 0. Then it follows that r(x) = 0.

From (5.5) one has

d(x, S) = ‖x− x‖ � r(x)

(λδ)12

, ∀x ∈ K,

which implies that x ∈ S = {x} if r(x) = 0. As a consequence, x ∈ S = {x} iff r(x) = 0 and so the above

inequality yields that an error bound holds for S with r(x) and with respect to K.

862 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

6 Applications to traffic equilibria

In [9, 10], Daniele, Maugeri and Oettli introduced traffic equilibria in the case where the numerical data

of the traffic network depend on time and characterized it by a suitably generalized form of Wardrop

principle in terms of a VI. In this section, we shall apply the results obtained in Sections 3 and 4 to

investigate the Lagrangian-type necessary and sufficient optimality conditions for traffic networks. The

notation that we employ is, for the most part, the same as that in [9, 10].

6.1 Applications to standard traffic networks

In the traffic network, one has a set W of origin-destination pairs and a set R of routes. Each route links

exactly one origin-destination pair w ∈ W . The set of all r ∈ R which links a given w ∈ W is denoted

by R(w).

We consider the flow vectors H = (Hr)�r∈R ∈ R

R, where Hr, r ∈ R, denotes the flow in route r. A

feasible flow has to satisfy the capacity restrictions,

λr � Hr � μr, ∀ r ∈ R,

and demand requirements,

r∈R(w)

Hr = ρw, ∀w ∈ W ,

where λ <R

R+μ are given in R

R and ρ �R

W+

0 is given in RW .

Thus, the set of all feasible flows is given by

K :=

{

H ∈ RR : λ �

RR+H �

RR+μ,

r∈R(w)

Hr = ρw, ∀w ∈ W

}

.

Furthermore, we are given a cost function C : K → RR. Then, to every feasible flow H ∈ K, there

corresponds a cost vector C(H) ∈ RR; Cr(H) gives the marginal cost of sending one additional unit of

flow through route r, when the flow H is already present.

Definition 6.1. A flow H ∈ RR is called an equilibrium flow iff

H ∈ K and 〈C(H), H − H〉 � 0, ∀H ∈ K.

Let

X := {H ∈ RR : λ �

RR+H �

RR+μ},

G1(H) :=

(∑

r∈R(w)

Hr − ρw

)�

w∈W

, ∀H ∈ RR,

G2(H) := −G1(H) =

(

ρw −∑

r∈R(w)

Hr

)�

w∈W

, ∀H ∈ RR,

G(H) :=

(G1(H)

G2(H)

)

, ∀H ∈ RR.

Clearly, X is compact and convex, the function G is affine. Then we can rewrite the set of all feasible

flows K as follows: K = {H ∈ X : G(H) �R

2W+

0}.Let H be an equilibrium flow. Set

H := {(u, v) ∈ R× R2W : u > 0, v �

R2W+

0},K(H) := {(u, v) ∈ R× R

2W : u = 〈C(H), H −H〉, v = G(H), H ∈ X},

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 863

Hu := {(u, v) ∈ R× R2W : u > 0, v = 0},

E := K(H)− clH= {(u, v) ∈ R× R

2W : u � 〈C(H), H −H〉, v �R

2W+

G(H), H ∈ X}= −[(f,−G)(X) + R+ × R

2W+ ],

where f(H) := 〈C(H), H − H〉. The set K(H) is called the image of the above traffic network at H ∈ K,

while the space R× R2W is the image space associated with the above traffic network at H ∈ K.

Based on the results presented in Sections 3 and 4, one can easily check the following results hold.

Proposition 6.1. That H ∈ K is an equilibrium flow iff

H ∩K(H) = ∅,iff

H ∩ E = ∅,iff

Hu ∩ E = ∅.Theorem 6.1. If H ∈ K is an equilibrium flow, then the sets H and K(H) admit a linear separation,

or equivalently, there is a vector (u, v) ∈ B such that :

(i) u〈C(H), H −H〉+ 〈v, G(H)〉 � 0, ∀H ∈ X ;

(ii) 〈v, G(H)〉 = 0,

where B := {(u, v) ∈ R+ × R2W+ : ‖(u, v)‖ = 1}.

Proof. Since the function G ia affine, −G is R2W+ -convex on X . Thus the desired conclusion follows

directly from Theorem 4.5.

Similarly, we have the following:

Theorem 6.2. Let H ∈ K. Assume that there is a vector (u, v) ∈ B such that u〈C(H), H − H〉 +〈v, G(H)〉 � 0, ∀H ∈ X, where B is as that in Theorem 6.1. If u > 0, then H is an equilibrium flow.

6.2 Applications to time-dependent traffic networks

In [9, 10], Daniele, Maugeri and Oettli considered the dynamic case of traffic networks. The traffic

network, whose geometry remains fixed, is considered at all times t ∈ T , where T := [0, T ].

For each time t ∈ T , we have a route-flow vector H(t) ∈ RR. H : T → R

R is the flow (trajectory)

over time. For technical reasons, let the functional setting for the flow trajectories be the reflexive Banach

space Lp(T ,RR) with p > 1. For simplicity, let L := Lp(T ,RR). The dual space of Lp(T ,RR), for

instance, Lq(T ,RR), will be denoted by L ∗, where 1q +

1p = 1. The canonical bilinear form on L ∗ ×L

is given by

� G,H �:=

T

〈G(t), H(t)〉dt, ∀ (G,H) ∈ L ∗ × L .

The feasible flows have to satisfy the time-dependent capacity constraints and demand requirements;

namely, almost everywhere (a.e.) on T ,

λ(t) �R

R+H(t) �

RR+μ(t) and

r∈R(w)

Hr(t) = ρw(t), ∀w ∈ W ,

where λ < μ (i.e., λ(t) <R

R+μ(t) for all t ∈ T ) are given in L and ρ � 0 (i.e., ρ(t) �

RW+

0 for all t ∈ T )

is given in Lp(T ,RW ). The set of feasible flows is given by

K :=

{

H ∈ L : λ(t) �R

R+H(t) �

RR+μ(t) a.e. onT ;

864 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

r∈R(w)

Hr(t) = ρw(t) a.e. onT , ∀w ∈ W

}

.

We assume that∑

r∈R(w)

λr(t) � ρw(t) �∑

r∈R(w)

μr(t), a.e. onT , ∀w ∈ W .

Then, K is nonempty. It is easy to see that K is closed, convex and bounded, hence weakly compact.

Furthermore, we are given a mapping C : K → L ∗, which assigns to each flow trajectory H ∈ K the

cost trajectory C(H) ∈ L ∗.

Definition 6.2. A flow H ∈ L is called an equilibrium flow iff

H ∈ K and � C(H), H − H �� 0, ∀H ∈ K.

In [9, 10], Daniele, Maugeri and Oettli presented Wardrop principle for the dynamic traffic networks.

We will derive the Lagrangian-type necessary and sufficient optimality conditions for the dynamic traffic

networks.

To begin with, let

X := {H ∈ L : λ(t) �R

R+H(t) �

RR+μ(t) a.e. onT },

G1(H) :=

(∑

r∈R(w)

Hr − ρw

)�

w∈W

, ∀H ∈ L ,

G2(H) := −G1(H) =

(

ρw −∑

r∈R(w)

Hr

)�

w∈W

, ∀H ∈ L ,

G(H) :=

(G1(H)

G2(H)

)

=

(∑

r∈R(w)

Hr,−∑

r∈R(w)

Hr

)�

w∈W

+ (−ρw, ρw)�w∈W , ∀H ∈ L .

Clearly, X is also closed, convex and bounded, hence weakly compact, the function G is affine. Denote

by Z := Lp(T ,R2W ) and by Z∗ := Lq(T ,R2W ) the dual space of Lp(T ,R2W ), where 1q + 1

p = 1.

Let

D := {I ∈ Z : I(t) �R

2W+

0, a.e. onT }

and

D+ := {I ∈ Z : I(t) >R

2W+

0, a.e. onT }.

Then D is a closed convex cone in Z. The dual cone (or positive polar cone) of D is given by

D∗ := {V ∗ ∈ Z∗ :� V ∗, I �� 0, ∀ I ∈ D}.Clearly, D∗ = {V ∗ ∈ Z∗ : V ∗(t) �

R2W+

0, a.e. onT }.We will use the following notations:

I �D J ⇔ I − J ∈ D, ∀ I, J ∈ Z ,and

I >D+ J ⇔ I − J ∈ D+, ∀ I, J ∈ Z .

Then, we can rewrite the set of all feasible flows K as follows: K = {H ∈ X : G(H) �D 0}.Let H be an equilibrium flow. Set

H := {(u, V ) ∈ R×Z : u > 0, V �D 0},

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 865

K(H) := {(u, V ) ∈ R×Z : u =� C(H), H −H �, V = G(H), H ∈ X},Hu := {(u, V ) ∈ R×Z : u > 0, V = 0},E := K(H)− clH

= {(u, V ) ∈ R×Z : u �� C(H), H −H �, V �D G(H), H ∈ X}= −[(f,−G)(X) + R+ × D],

where f(H) :=� C(H), H − H �. The set K(H) is called the image of the above time-dependent traffic

network at H ∈ K, while the space R× Z is the image space associated with the above time-dependent

traffic network at H ∈ K.

Similarly, one has the following:

Proposition 6.2. That H ∈ K is an equilibrium flow iff

H ∩K(H) = ∅,

iff

H ∩ E = ∅,

iff

Hu ∩ E = ∅.

Since the sets H and E have empty interior, the classic separation theorem for convex sets cannot be

applied. To this aim, we introduce the notion of quasi-relative interior and apply separation theorems

related to this notion.

Let M and N be two subsets of the topological vector space Y . Denote by coneN := {y ∈ Y : y =

λx, λ � 0, x ∈ N} the cone generated by N and by NM (x) := {x∗ ∈ Y ∗ : 〈x∗, y − x〉 � 0, ∀ y ∈ M}the normal cone to M at x, where Y ∗ is the topological dual space of Y . We say that x ∈ M is a quasi

interior point of M , denoted by x ∈ qiM , iff cl cone(M − x) = Y , or equivalently, NM (x) = {0}; we say

that x ∈ M is a quasi relative interior point of M , denoted by x ∈ qriM , iff cl cone(M − x) is a linear

subspace of Y , or equivalently, NM (x) is a linear subspace of Y ∗. For more details, see [4, 40].

It follows from the definitions above that qri{x} = {x}, ∀x ∈ Y . For any convex set M , we have that

qiM ⊆ qriM and, intM �= ∅ implies intM = qiM . Similarly, if qiM �= ∅, then qiM = qriM [5, 27].

Moreover, if Y is a finite-dimensional space and intM �= ∅, then qriM =intM and qriM = riM [4,5,27].

We need the following lemmas.

Lemma 6.1. Let us consider M and N to be two convex subsets of the topological vector space Y .

Then the following statements are true :

(i) qri(M ×N)=qriM×qriN ;

(ii) Let x ∈M and 0 ∈ qi(M −M). Then x ∈ qiM⇔ x ∈ qriM ;

(iii) If qiN �= ∅, then M−qiN ⊆qi(M −N).

Proof. Statement (i) can be found in [4–6, 27] and Statement (ii) can be found in [20]. We only need

to prove (iii). Let x ∈M and y ∈qiN . Then

cl cone(M −N − (x− y)) ⊇ −cl cone(N − y) = Y,

which implies that x− y ∈qi(M −N). �

Lemma 6.2 (See [5, 6]). Let M be a nonempty convex subset of the topological vector space Y and

x0 ∈ Y such that qriM �= ∅ and cl cone(M − x0) is not a linear subspace of Y . Then, there exists

x∗ ∈ Y ∗\{0} such that 〈x∗, x〉 � 〈x∗, x0〉 for all x ∈M , where Y ∗ is the topological dual space of Y .

The following proposition plays an important role in investigating the Lagrangian-type necessary and

sufficient optimality conditions for time-dependent traffic networks.

866 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

Proposition 6.3. Let H ∈ K be an equilibrium flow. Then the following statements are true :

(i) qiH �= ∅ and qiE �= ∅;(ii) (0, 0) �∈ qiE, i.e., cl coneE �= R×Z, or equivalently, NE(0, 0) �= {(0, 0)}.

Proof. (i) Clearly, clH = R+ × D, qriR+ = qri{u ∈ R : u > 0} = {u ∈ R : u > 0}. From [4, 20, 27], we

have qriD = D+. Since {u ∈ R : u > 0} − {u ∈ R : u > 0} = R+ − R+ = R and D − D = Z, from (ii) in

Lemma 6.1 one has qiR+ = qi{u ∈ R : u > 0} = {u ∈ R : u > 0} and qiD = D+. As a consequence, (i)

in Lemma 6.1 allows that qiH =qi (clH) = {(u, V ) ∈ R×Z : u > 0, V >D+ 0}. Since E = K(H)− clH,

it follows from (iii) in Lemma 6.1 that qiE �= ∅.(ii) Since H ∈ K is an equilibrium flow, (0, 0) ∈ E . Suppose to the contrary that (0, 0) ∈ qiE , i.e.,

cl coneE = R×Z , or equivalently, NE(0, 0) = {(0, 0)}. Let us consider (0, V ∗) ∈ R×Z∗ with

V ∗ =

(V ∗0

V ∗0

)

such that V ∗0 ∈ D∗

0 with V ∗0 �= 0, where D∗

0 = {I ∈ Lq(T ,RW ) : I(t) �R

W+

0, a.e. onT }. Then V ∗ ∈ D∗.Since

V ∗ =

(V ∗0

V ∗0

)

and G(H) =

(G1(H)

−G1(H)

)

,

for any H ∈ X, t ∈ R+, V ∈ D, we have

〈(0, V ∗), (−f(H)− t, G(H)− V )〉= 〈V ∗, G(H)− V 〉 = 〈V ∗, G(H)〉 − 〈V ∗, V 〉= 〈V ∗

0 , G1(H)〉 − 〈V ∗0 , G1(H)〉 − 〈V ∗, V 〉

� 0,

which implies that (0, V ∗) ∈ NE(0, 0), a contradiction with the assumption since (0, V ∗) �= (0, 0).

Now, we apply separation theorems related to quasi relative interior point to derive the Lagrangian-type

necessary and sufficient optimality conditions for time-dependent traffic networks.

Theorem 6.3. If H ∈ K is an equilibrium flow, then the sets H and K(H) admit a linear separation,

or equivalently, there is a vector (u, V ) ∈ B such that :

(i) u� C(H), H −H � + � V , G(H) �� 0, ∀H ∈ X ;

(ii) � V , G(H) �= 0,

where B := {(u, V ∗) ∈ R+ ×D∗ : ‖(u, V ∗)‖ = 1}.Proof. Let H ∈ K be an equilibrium flow. Since the function G is affine, −G is D-convex on X . Then

the sets H and E are convex. From Proposition 6.3, we have (0, 0) �∈ qiE . Since qiE =qriE , we obtain

(0, 0) �∈ qriE , equivalently, cl coneE is not a linear subspace of R × Z . Setting M := E and x0 := 0, it

follows from Lemma 6.2 that there exists a vector (u, V ) ∈ R×Z∗, with (u, V ) �= (0, 0), such that

u� C(H), H −H � + � V , G(H) �� uu+ � V , V �, ∀H ∈ X, ∀ (u, V ) ∈ H. (6.1)

Since H = {(u, V ) ∈ R×Z : u > 0, V �D 0}, letting (u, V ) → (0, 0) in (6.1) yields

u� C(H), H −H � + � V , G(H) �� 0, ∀H ∈ X, (6.2)

proving (i).

We declare that

0 � uu+ � V , V �, ∀ (u, V ) ∈ H. (6.3)

If there is (u0, V0) ∈ H such that uu0+ � V , V0 �< 0, then, as t→ +∞,

u(tu0)+ � V , tV0 �= t(uu0+ � V , V0 �) → −∞,

Li J et al. Sci China Math April 2012 Vol. 55 No. 4 867

which is a contradiction with (6.1) since (tu0, tV0) ∈ H.

It follows from (6.3) that (u, V ) ∈ R+ × D∗. Since (u, V ) �= (0, 0), without loss of generality, we can

suppose that (u, V ) ∈ B.Setting H := H in (6.2), we obtain

� V , G(H) �� 0,

and so

� V , G(H) �= 0,

since G(H) �D 0 and V ∈ D∗. This completes the proof. �

Theorem 6.4. Let H ∈ K. Assume that there is a vector (u, V ) ∈ B such that

u� C(H), H −H � + � V , G(H) �� 0, ∀H ∈ X,

where B is as that in Theorem 6.3. If u > 0, then H is an equilibrium flow.

Proof. The proof is similar to that in Theorem 4.6.

Acknowledgements This work was supported by National Natural Science Foundation of China (Grants Nos.

60804065, 70831005), the Key Project of Chinese Ministry of Education (Grant No. 211163), Sichuan Youth

Science and Technology Foundation and the Research Foundation of China West Normal University (Grant No.

08B075). The first author would like to gratefully thank Prof. Giannessi F and Dr. Mastroeni G for their

hospitality during the period he visited Department of Mathematics, University of Pisa, from March 19, 2011 to

March 27, 2011. The authors are in debt to Dr. Mastroeni G for his valuable suggestions on Sections 5 and 6,

which have essentially upgraded the quality of the paper. Moreover, the authors thank the editors and referees

for their insightful comments.

References

1 Auslender A. Optimization, Methodes Numeriques. Paris: Masson, 1976

2 Auslender A, Teboulle M. Lagrangian duality and related multiplier methods for variational inequality problems. SIAM

J Optim, 2000, 10: 1097–1115

3 Auslender A, Teboulle M. Asymptotic Cones and Functions in Optimization and Variational Inequalities. New York:

Springer-Verlag, 2003

4 Borwein J M, Lewis A S. Partially finite convex programming, part I: Quasi-relative interiors and duality theory. Math

Program Ser A, 1992, 57: 15–48

5 Bot R I, Csetnek E R, Moldovan A. Revisiting some duality theorems via the quasirelative interior in convex optimiza-

tion. J Optim Theory Appl, 2008, 139: 67–84

6 Cammaroto F, Bella B Di. Separation theorem based on the quasirelative interior and application to duality theory.

J Optim Theory Appl, 2005, 125: 223–229

7 Castellani M, Mastroeni G, Pappalardo M. On regularity for generalized systems and applications. In: Pillo G Di,

Giannessi F, eds. Nonlinear Optimization and Applications. New York: Plenum Press, 1996, 13–26

8 Daniele P, Giuffr S, Idone G, et al. Infinite dimensional duality and applications. Math Ann, 2007, 339: 221–239

9 Daniele P, Maugeri A, Oettli W. Variational inequalities and time-dependent traffic equilibria. C R Acad Sci Paris Ser

I Math, 1998, 326: 1059–1062

10 Daniele P, Maugeri A, Oettli W. Time-dependent traffic equilibria. J Optim Theory Appl, 1999, 103: 543–555

11 Dien P H, Mastroeni G, Pappalardo M, et al. Regularity conditions for constrained extremum problems via image

space: the linear case. In: Komlosi S, Rapcsak T, Schaible S, eds. Generalized Convexity. Lecture Notes in Economics

and Mathematical Systems. Berlin: Springer-Verlag, 1994, 145–152

12 Facchinei F, Pang J S. Finite-dimensional Variational Inequalities and Complementarity Problems. New York:

Springer-Verlag, 2003

13 Ferris M C. Weak Sharp Minima and Penalty Functions in Mathematical Programming. PhD Thesis. Cambridge:

University of Cambridge, 1988

14 Giannessi F. Theorems of the alternative and optimality conditions. J Optim Theory Appl, 1984, 60: 331–365

15 Giannessi F. Semidifferentiable functions and necessary optimality conditions. J Optim Theory Appl, 1989, 60: 191–241

868 Li J et al. Sci China Math April 2012 Vol. 55 No. 4

16 Giannessi F. Vector Variational Inequalities and Vector Equilibria. Dordrecht-Boston-London: Kluwer Academic

Publishers, 2000

17 Giannessi F. Constrained Optimization and Image Space Analysis. New York: Springer-Verlag, 2005

18 Giannessi F, Mastroeni G. Separation of sets and Wolfe duality. J Global Optim, 2008, 42: 401–412

19 Giannessi F, Mastroeni G, Pellegrini L. On the theory of vector optimization and variational inequalities: Image space

analysis and separation. In: Giannessi F, ed. Vector Variational Inequalities and Vector Equilibria. Dordrecht-Boston-

London: Kluwer Academic Publishers, 2000, 153–215

20 Grad A. Quasi-relative interior-type constraint qualifications ensuring strong Lagrange duality for optimization prob-

lems with cone and affine constraints. J Math Anal Appl, 2010, 361: 86–95

21 Guu S M, Huang N J, Li J. Scalarization approaches for set-valued vector optimization problems and vector variational

inequalities. J Math Anal Appl, 2009, 356: 564–576

22 Hiriart-Urruty J-B, LemarechalC. Convex Analysis and Minimization Algorithms I. Berlin: Springer-Verlag, 1993

23 Hiriart-Urruty J-B, Lemarechal C. Fundamentals of Convex Analysis. Berlin: Springer-Verlag, 2001

24 Jahn J. Vector Optimization, Theory, Applications, and Extensions. Berlin: Springer-Verlag, 2004

25 Jahn J. Introduction to the Theory of Nonlinear Optimization, 3rd ed. Berlin: Springer-Verlag, 2007

26 Li J, Huang N J. Image space analysis for vector variational inequalities with matrix inequality constraints and

applications. J Optim Theory Appl, 2010, 145: 459–477

27 Limber M A, Goodrich R K. Quasi interiors, Lagrange multipliers, and Lp spectral estimation with lattice bounds. J

Optim Theory Appl, 1993, 78: 143–161

28 Luc D T. Theory of Vector Optimization. Berlin: Springer-Verlag, 1989

29 Madani K, Mastroeni G, Moldovan A. Constrained extremum problems with infinite-dimensional image: Selection and

necessary conditions. J Optim Theory Appl, 2007: 135: 37–53

30 Madani K, Mastroeni G, Moldovan A. Constrained extremum problems with infinite dimensional image: Selection and

saddle point. J Global Optim, 2008, 40: 197–208

31 Mastroeni G. Separation methods for vector variational inequalities: Saddle point and gap function. In: Pillo G Di,

Giannessi F, eds. Nonlinear Optimization and Applications. Dordrecht: Kluwer Academic Publishers, 2000, 207–217

32 Mastroeni G, Pellegrini L. On the image space analysis for vector variational inequalities. J Ind Manag Optim, 2005,

1: 123–132

33 Maugeri A, Raciti F. Remarks on infinite dimensional duality. J Global Optim, 2010, 46: 581–588

34 Moldovan A, Pellegrini L. On regularity for constrained extremum problems, part 1: Sufficient optimality conditions.

J Optim Theory Appl, 2009, 142: 147–163

35 Moldovan A, Pellegrini L. On regularity for constrained extremum problems, part 2: Necessary optimality conditions.

J Optim Theory Appl, 2009, 142: 165–183

36 Nagurney A. Network Economics: A Variational Inequality Approach. Dordrecht: Kluwer Academic Publishers, 1993

37 Pang J S. Error bounds in mathematical programming. Math Program Ser A, 1997, 79: 299–332

38 Rockafellar R T. Convex Analysis. Princeton: Princeton Univ Press, 1970

39 Zheng X Y, Yang X Q. The structure of weak Pareto solution sets in piecewise linear multiobjective optimization in

normed spaces. Sci China Ser A, 2008, 51: 1243–1256

40 Zalinescu C. Convex Analysis in General Vector Spaces. River Edge, NJ: World Scientific, 2002


Recommended