1. Introduction
In this paper, we consider the second-order cone constrained variational inequalities (SOCCVI) problem: finding
such that
where
is the Euclidean inner product,
,
are continuously differentiable,
is a closed convex set, and
is the Cartesian product of second-order cones (SOCs) as follows:
where
,
, and
is a second-order cone in
. We denote
and
for
. Then, we have the following equivalent relations:
Variational inequalities are widely used in physics, mechanics, economics, optimization, control, and equilibrium models in transportation, etc. The second-order cone constrained variational inequalities problem has also attracted the attention of many scholars. Recently, Sun et al. [
1] constructed an implementable augmented Lagrangian method for solving the second-order cone constrained variational inequalities (SOCCVI) problem (
1). Nazemi and Sabeghi [
2] reduced the variational inequalities to a convex second-order cone programming (CSOCP) problem and solved the convex second-order cone constrained variational inequalities problem by using a gradient neural network model. Nazemi and Sabeghi [
3] used a new neural network model as a simple solution to the convex second-order cone constrained variational inequalities problem.
The optimality condition is a necessary and sufficient condition that the objective function and the constraint function must satisfy at the optimal point of the optimization problem. The optimality conditions play an important role in the termination rule of the algorithm and the proof of optimization theory. In order to ensure the convergence of algorithms, we should analyze the stability of the solution to the optimization problems (or perturbation analysis). Stability analysis was used for the first time in the study of the sensitivity analysis of linear programming by Manne [
4]. Danskin [
5] computed the directional differentiability and directional derivative at the optimal solution. Bonnans and Ramirez [
6] proved that the strong regularity of the KKT point for the second-order cone programming (SOCP) problem is equivalent to the constrained nondegenerate condition and the strong second-order sufficient condition, which was the perturbation analysis of the second-order cone programming problems. Wang and Zhang [
7] studied the stability analysis for the second-order cone programming (SOCP) problem and proved that the nonsingularity of Clarke generalized Jacobian and the strong regularity of the KKT point for the non-smooth KKT system are equivalent to each other.
In this paper, we shall study the KKT condition, the first-order necessary condition, the second-order sufficient condition, and Clarke’s generalized differential nonsingularity of the non-smooth mapping corresponding to the KKT condition for the second-order cone constrained variational inequalities (SOCCVI) problem (
1). We will demonstrate the relationship between the strong second-order sufficient condition and the nonsingularity of Clarke’s generalized Jacobian for the KKT system of the SOCCVI problem. And the strong regularity of the KKT point, the strong stability of the KKT point, and the local Lipschitz homeomorphism near the KKT point will be equivalent to each other.
The content of this paper is outlined as follows: In
Section 2, we give some preliminaries that will be used in the following sections.
Section 3 studies the KKT conditions of the SOCCVI problem (
1) and proves the first-order necessary condition, the second-order sufficient condition, and the Clarke generalized differential nonsingularity of the non-smooth mapping corresponding to the KKT conditions. Finally, we establish and prove the relationship between the strong second-order sufficient condition, the nonsingularity of Clarke’s generalized differentiation, the strong regularity of the KKT point, the strong stability of the KKT point, and the local Lipschitz homeomorphism near the KKT point for the SOCCVI problems in
Section 4.
2. Preliminaries
The projection operator to a convex set is quite useful in reformulating the box-constrained variational inequality as a non-smooth equation. Let
C be a convex closed set, and for every
, there is a unique
in
C such that
The point is the projection of x onto C, denoted by . The projection operator is well defined over , and it is a nonexpensive mapping.
Lemma 1 ([
8]).
Let H be a real Hilbert space and be a closed convex set. For a given satisfies the inequalityif and only if . More information on the projection operator can be found in Section 3 of the book by Goebel and Reich [
9].
The second-order cone (SOC) in
(
), also called the Lorentz cone or the ice-cream cone, is defined as
If , is the set of nonnegative reals .
For any two vectors for
and
, the Jordan product of
x and
y is denoted by
As a result, the normal addition “+”, “∘”, and produce the Jordan algebra associated with the second-order cone.
Any
has the following spectral decomposition (see Faraut and Kor
nyi [
10]).
where
and
are the spectral values of
x with formulas
while
and
are the spectral vectors associated with
x given by
where
,
, and
.
The projection of
x onto
, denoted by
, can be represented as follows:
where
We can calculate it as follows:
Importantly,
is strongly semi-smooth over
(see Chen et al. [
11]).
In order to obtain the stability analysis for the SOCCVI problem (
1), we recall the differential properties in the following.
Lemma 2 ([
12]).
is directionally differentiable at x for any . Moreover, the directional derivative is described bywhere, and
For the convenience of further discussion, we need the definition of the tangent cone, regular normal cone, and the second-order tangent cone of a closed set at a point. All the concepts are found in Rockafellar and Wets [
13].
Definition 1 ([
13]).
For a closed set and a point , the following sets are defined.The tangent (Bouligand) cone is as follows:where represents that t decreases towards 0. The regular normal cone is as follows: The limiting (in the sense of Mordukhovich) normal cone is as follows: If
K is a closed convex set, then
and
where
denotes the polar of
K. According to Bonnans and Shapiro [
14] and the definition of the tangent, we obtain that for any
, there exist
and
such that
, and for
.
Lemma 3 ([
13]).
The definitions of the tangent cone and the second-order tangent cone of at areand Next, we recall the smoothing metric projection operator as follows:
for any
, where
and
. Notice that when
, then
, and
is continuously differentiable at any
if
. Moreover,
is globally Lipschitz continuous and strongly semi-smooth for any
.
Let
x have the spectral decomposition
. For simplicity, we denote
for
. It is easy to verify that
has the following spectral decomposition:
where
, for any
. Thus, the function
has the following expression:
If
, we can calculate the derivative of
as follows:
where
and
(i) if
, we have
where
and
(ii) if
, we obtain that
If , we can calculate the B-subdifferential of as follows:
Lemma 4 ([
15]).
If and , can be calculated as follows:(i) if , we have (ii) if , we obtain that (iii) if , we obtain thatwherewhere and are defined by Now, we give an expression of
in terms of
and
. Let
. If
,
,
,
,
,
, for
, we can obtain that
for some
. Therefore, we have the following:
The following conclusions will play an important role in the subsequent discussions on optimality and stability.
Lemma 5 ([
16]).
Suppose that and (i.e., and , for all i = 1,⋯,r). Then, =0, if and only if = Arw()Arw()e=0 for i = 1, ⋯, r, or equivalently,- (i)
,
- (ii)
,
Lemma 6 ([
17]).
For any , there exists with such that Proposition 1 ([
17]).
Let be a Karush–Kuhn–Tucker point. Then, for any and such that , it holds thatwhere , and is defined bywhere is a case of the index set, and the specific index set is as follows:where . 3. Optimality Condition
The Lagrange function of the SOCCVI problem (
1) is as follows:
where
.
The KKT condition of the SOCCVI problem (
1) is as follows:
where “∘” is the Jordan product.
Let us consider the second-order cone constrained optimization (SOCCOP) problem.
where
is defined by (
2),
is a twice continuously differentiable function, and
is a twice continuously differentiable mapping.
It is easy to know that the Lagrange function of the SOCCOP problem (
21) can be written as follows:
where
.
The KKT condition of the SOCCOP problem (
23) is as follows:
Clearly, if
, the KKT mappings of the SOCCOP problem (
23) and the SOCCVI problem (
1) are the same. Moreover, if
f is a twice continuously differentiable convex function, these two problems are equivalent. Due to the close relationship between the SOCCVI problem (
1) and SOCCOP (
23), we will study the optimality conditions and stability analysis of the SOCCVI problem (
1) in the framework of SOCCOP (
23). The following research is assumed to be true of the condition
.
If
satisfies the KKT condition (
21), then
is a stationary point of the SOCCVI problem (
1), and the set of multipliers is denoted by
. If the constraint nondegeneracy condition holds at
, that is,
then
is nondegenerate.
Suppose that
is a stationary point of the SOCCVI problem (
1); then, Robinson’s constraint qualification holds at
; that is,
and the Lagrange multiplier set
is nonempty compact.
It follows from Corollary 26 in Bonnans and Ramirez [
6] that we can deduce the critical cone at a stable point
of the SOCCVI problem (
1).
According to Definition 1 and Lemma 5, the above critical cone
can be expressed as follows:
Next, we will obtain the second-order sufficient condition for the stationary point of the SOCCVI problem (
1).
Theorem 1. Suppose that is a stationary point of the SOCCVI problem (1) and satisfies the Robinson’s Constraint Qualification (25); then, the second-order growth condition holds at if and only if the second-order sufficient condition holds at ; that is,
where
,
, and
Proof. If the second-order growth condition holds, and the set
is the second-order regular at
along
, then we obtain that
where
is the second-order tangent set at
along the direction
, and
is the support function of the set
. Therefore, we only prove that (
27) and (
29) are equivalent.
According to the definition of the second-order tangent set, we set
and
). Then, we have the following from Lemma 4:
Since
, we obtain that
. According to the definition of the second-order tangent set,
and
, we obtain that
If
, we obtain that
. From (
30), when the condition
,
, or
holds, we have
.
Next, considering
and
, we let
. From (
30), for
, we can obtain that
From the KKT condition
, we obtain that
. Then, we have
For
, we know that
Since
, we have
Thus, we show that (
27) and (
29) are equivalent; that is, the second-order sufficient condition at
is as follows:
This completes the proof. □
If
is a stationary point of the SOCCVI problem (
1), the Robinson’s constraint qualification is satisfied at
; that is,
From the definition of normal cone, we have the Karush–Kuhn–Tucker condition.
where
.
The set of multipliers is denoted by
, which is a nonempty closed convex and bounded set. Then, the second expression can be as follows:
The Constraint Nondegenerate Condition of the SOCCVI problem (
1) holds at
in the following:
In order to prove the stability theorem, the affine space of the critical cone and the strong second-order sufficient condition are defined below.
Let
be a locally optimal solution of the SOCCVI problem (
1); then, the affine space of the critical cone at
is as follows:
where the index set is defined by the following:
Definition 2. Let be a feasible point of the SOCCVI problem (1) such that is a singleton. We say that the strong second-order sufficient condition holds at ifwhere , and 4. Stability Analysis
Let
be a stationary point of the SOCCVI problem (
1). Then,
satisfies the Karush–Kuhn–Tucker condition (
32) if and only if
, where
The smoothing function
of the Karush–Kuhn–Tucker function
is defined as follows:
where
is defined in (
6). It is obvious that
Moreover, if
is a Karush–Kuhn–Tucker point of the SOCCVI problem (
1), then
.
Next, we introduce the concepts and theorems related to stability analysis.
Let
be the Karush–Kuhn–Tucker point of the SOCCVI problem (
1); then,
if and only if
is a solution of the following inclusion:
The Equation (
40) can be described equivalently as the following generalized equation:
where
,
, and
is a continuous differentiable map defined by
Next, we give the definition of the strongly regular solution of the KKT condition.
Definition 3. We say that is the strongly regular solution of the KKT condition. If there exist a neighborhood V of and a neighborhood of such that for any , then the following linearization systemhas a unique solution , and this solution is Lipschtiz continuous. The perturbation problem of the SOCCOP problem (
23) is as follows:
where
is a finite-dimensional space,
, and
,
are second continuously differentiable, satisfying
,
. We denote by
the corresponding perturbation problem of the SOCCOP problem (
23).
Similar to the discussion in Bonnans and Shapiro [
14], we give the uniform second-order growth condition for the stationary point of the SOCCOP problem (
23).
Definition 4. Let be a stationary point of the SOCCOP problem (23). Suppose is a -smooth parametrization. The uniform second-order growth condition holds at if there exist , a neighborhood V of , and a neighborhood of such that, for any and any stationary point of the problem (44), the following inequality holds: If (45) holds for any -smooth parametrization, then the uniform second-order growth condition holds at . Further, we can give the definition of a strong stable point as in Bonnans and Shapiro [
14].
Definition 5. Let be a stationary point of the SOCCOP problem (23). If, for any -smooth parametrization , there exist the neighborhood V of and the neighborhood of such that, for any , the problem (44) has a unique stationary point , and is continuously at , then is strongly stable. The next theorem demonstrates the relationship between the strong second-order sufficient condition under the constraint nondegeneracy condition and the nonsingularity of the generalized Jacobian of the smoothing mapping
E defined by (
39).
Theorem 2. Suppose that is a Karush–Kuhn–Tucker point of the SOCCVI problem (1), and the constraint nondegeneracy condition and the strong second-order sufficient condition (37) hold at ; then, any element in is nonsingular. Proof. Let
W be an arbitrary element in
, and
satisfy
then, there exists
satisfying
Obviously, we have
. Thus, the third equation in (
46) becomes
We will discuss the following three cases:
(a) If
, we know that
and
. Then, we obtain
. Thus, for any
, we obtain
(b) If
, we obtain
and
. Then, there exists
such that
, and we have
For
, it has the following form:
where
From (
47), we obtain that
Thus, for
, we have
(c) If
, we have
and
. Then,
can be expressed
, where
,
,
. Since there exists
such that
and thus,
we can obtain
which implies that, for any
,
From the second and third equations of (
46), we have
Combining with the last Equation (
46), Proposition 1, and (
37), the following is implied:
Then, from the strong second-order sufficient condition (
37) and (
46), we conclude that
. Therefore, (
46) reduces to
From the second equation of (
49), the index sets (
20), Lemma 5, and B, we could easily obtain
Then, it follows from the constraint nondegeneracy condition that there exists a vector
such that
Then, we have
which implies that
Together with (
51), the conditions
, and (
50) imply that
W is nonsingular. This completes the proof. □
Corollary 1. Let be a Karush–Kuhn–Tucker point of the SOCCVI problem (1). If the constraint nondegeneracy condition and the strong second-order sufficient condition (37) hold at , then any element in is nonsingular. Proof. For any element
, we compute
where
. From Lemma 6, there exists
such that
Then, we obtain
where
is defined by (
46).
Under the constraint nondegeneracy condition and the strong second-order sufficient condition (
37), it follows from Theorem 2 that
which implies that
; that is,
is nonsingular. This completes the proof. □
The next theorem demonstrates the relationship between the strongly regular solution and the strong second-order sufficient condition.
Theorem 3. Let be a locally optimal solution of the SOCCVI problem (1), and is the corresponding Lagrange multiplier. Then, is the strongly regular solution of the generalized Equation (41) if and only if is nondegenerate and the following strongly second-order sufficient condition holds at : Proof. From Theorem 5.25 in Bonnans and Ramirez [
6], we know that
is nondegenerate, and the uniform second-order growth condition holds at
if and only if
is a strongly regular solution of the generalized Equation (
41). Hence, under the nondegenerate condition, we only prove that the strong second-order sufficient condition is equivalent to the uniform second-order growth condition.
If
is nondegenerate, then the affine space of the critical cone at
is as follows:
The uniform second-order growth condition implies that the strong second-order sufficient condition holds. Consider that the vector space
E is defined by
Obviously, we obtain that aff .
Next, we consider the perturbation problem of the second-order cone constrained optimization (SOCCOP) problem (
23).
where
denotes the first element of the natural basis of
, and
is the perturbed parameter. In problem (
56),
is still a locally optimal solution with the same Lagrange multiplier
such that a bigger critical cone is equal to
E.
Now, we prove that the strong second-order sufficient condition of the perturbation problem of (
56) can be obtained when
,
.
Let denote the matrices in the expression of the strong second-order sufficient condition.
For all
, we obtain that
; however, for all
, we obtain that
where
. We set
If
, then we have
,
. Hence, we obtain that
is equivalent to
. Combining with (
59), we can obtain
for all
and
for all
aff
.
For sufficiently small
u, the uniform second-order growth of the perturbation problem (
56) implies that
Since
for
, we have
from (
61). Thus, the strongly second-order sufficient condition (
53) holds at
. For any
and
, there exist
and
such that
is a feasible point of the perturbation problem (
44); that is,
and
For sufficiently large
n, there exists a unique Lagrange multiplier
of the stationary point
of the perturbation problem (
44). Since
, we have
. We assume that
converge to some
. Let us check that
aff
. For
, we know that
It follows from
that
Dividing it by
implies that
Let
in the above inequality, and we obtain that
. Combining the first equation of the KKT condition and
in (
62), we obtain that
According to the characterization of aff, we consider the following three cases:
Case I: For
,
, we obtain that (
65) holds.
Case II: For
, it follows from
that
and
for sufficiently large
n. There exists
such that
. Then, for any unit vector
z,
for sufficiently large
n, we compute the scalar product of
and (
64). Moreover, let
, and we obtain that
The above inequality holds for any
z. Using (
65), we have
Case III: For and , we will prove that .
By (
60) and (
65), we obtain that
By Taylor expansion in (
62), we deduce that
Combining with the first-order optimality conditions of the perturbation problem (
44), we deduce that
Similarly to (
63) and (
64), we can obtain that
Since
is self-dual,
is bounded, and
, we deduce that
Since
, we have
for some sequence
and
. Substituting (
69) into the inequality
and using
, we obtain that
Using (
71) and
, we obtain that
From the above three cases, we prove that aff.
Let
and
; then,
means that
. We denote by
the Lagrange multiplier associated with
, and we know that
Substituting
in (
73) and using the complementarity conditions at
, we deduce that
Adding (
62) and (
74), we can the following:
Similarly to Lemma 27 in Bonnans and Ramirez [
6], let
in (
75), and we can obtain that
, which is a contradiction with the strongly second-order sufficiently condition, since
aff
. Hence, the uniform second-order growth condition holds. This completes the proof. □
From the definition of the function
Q in (
38), we know that the Karush–Kuhn–Tucker point
of the SOCCVI problem (
1) satisfies
. The directional derivative of
Q along
is as follows:
Since is Lipschitz continuous, is well defined.
Proposition 2. Let and be defined by (76); then, Proof. Let
be as follows:
where
and
.
Let
. From Lemma 14 in Pang et al. [
18],
holds. We have
and together with (
76) and Lemma 4, we obtain that (
76) holds. This completes the proof. □
According to Clarke’s inverse theorem in Clarke [
19,
20], we can obtain that any element in
is nonsingular if and only if the function
Q is a locally Lipschitz homeomorphism near the Karush–Kuhn–Tucker point
in the following theorem:
Theorem 4. If is nonsingular, then there exist the neighborhood U of and V of , as well as the Lipschtiz function such that
The next lemma demonstrates that the relationship between the strongly regular solution and the function Q is that Q is a locally Lipschitz homeomorphism near the Karush–Kuhn–Tucker point .
Lemma 7. Suppose that , is defined by (38), and is a KKT point of the SOCCVI problem (1). Then, Q is a locally Lipschitz homeomorphism at if and only if is a strongly regular solution of the generalized Equation (41). Proof. Assuming that Q is a locally Lipschitz homeomorphism at , there exists an open neighborhood of which satisfies the condition that is an open neighborhood of the origin . For any , the equation has a unique solution in , and is Lipschitz continuous.
For any
, let
be a solution of the generalized Equation (
41). Denote
, and then we have
which implies that
that is,
Then,
uniquely exists in
and
Thus, is Lipschitz continuous on .
Assuming that
is a strongly regular solution of the generalized Equation (
41), there exist neighborhoods
of the origin
and
of
, as well as a locally Lipschitz function
, which satisfy the condition that
is the unique solution in
for any
. According to the above proof, we can conclude that
has a unique solution for any
and
where
and
which implies that
is Lipschitz continuous on
. Thus,
Q is Lipschitz homeomorphism at
. This completes the proof. □
Now, we demonstrate the stability analysis theorem of the SOCCVI problem (
1).
Theorem 5. Let be a locally optimal solution of the SOCCVI problem (1), suppose that Robinson’s CQ holds at , and let be a Lagrangian multiplier associated with . Then, the following conclusions are equivalent to each other: - (a)
The strong second-order sufficient condition (37) holds at , and is constraint nondegenerate. - (b)
Any element in is nonsingular.
- (c)
Any element in is nonsingular.
- (d)
The KKT point is a strongly regular solution of the generalized Equation (40). - (e)
The point is nondegenerate, and the uniform second-order growth condition holds at .
- (f)
The point is strongly stable, and is constraint nondegenerate.
- (g)
The function Q is a locally Lipschitz homeomorphism at the KKT point .
- (h)
Any element in is nonsingular.
Proof. Obviously, it follows from Theorem 2 and Corollary 1 that (a)⇒(b)⇒(c). According to Theorem 5.24 and Theorem 5.35 in Bonnans and Shapiro [
14], we can obtain that (d)⇔(e)⇔(f). From Theorem 3, we have (a)⇔(d). Proposition 2 implies that (h)⇔(c). From Clarke’s inverse theorem, we have (c)⇔(g). From Lemma 7, we have (d)⇔(g). This completes the proof. □