1. Introduction
Optimization problems have attracted significant attention due to their wide range of practical applications. In real-world scenarios, many problems often exhibit special structures, for example,
where
and
is regarded as a regularization term that imposes certain structures on
to yield a desired solution; see [
1,
2,
3,
4,
5] for more information. Particularly, the functions
and
often exhibit distinct properties, thus requiring different numerical treatment strategies. For example, in the LASSO problem,
is typically a smooth function, whereas
possesses non-smooth properties. Given a constrained optimization problem,
where
and
is a closed subset in
, the classical Lagrangian function provides a relaxation of the primal problem (
2). However if either
g or
is nonconvex, due to the presence of nonconvex constraints, this approach typically leads to a nonzero duality gap. Therefore, it is necessary to consider alternative strategies. In this case, we can employ the augmented Lagrangian function [
6,
7,
8]. However this approach requires the penalty parameter
to be sufficiently large to achieve the zero duality gap property between primal and dual problems, which inevitably introduces computational difficulties such as numerical instability.
Note that separable structures appear not only in the objective function as shown in (
1) but also in the constraint system. In fact, in some applications, the constraint system can also be categorized into different types [
9,
10,
11,
12,
13,
14]. For example, the author in [
14] considers the following optimization problem:
where
,
,
. The first two constraints are of the classical nonlinear programming form. The third constraint, however, requires at least one of the corresponding component functions of
G and
H to be zero, and hence, the problem is referred to as a mathematical program with switching constraints. If the third constraint is replaced by the complementarity constraint
,
,
, or equivalently
, then the corresponding problem is termed as a mathematical program with complementarity constraints. Due to the special structure of the third constraint, we need to apply distinct analytical approaches to the first two constraints and the third constraint separately. In this paper, we consider the following structured constrained optimization problem:
which corresponds to problem (
2) by setting
and
. The main reason for partitioning the constraint system into two components stems from their fundamentally different characteristics: in (3a),
is continuously differentiable with Lipschitz-continuous derivatives (i.e.,
), and
is a convex set; meanwhile, in (3b),
is twice continuously differentiable (i.e.,
), but
is a closed (but not necessarily convex) set. The objective function
is continuously differentiable with Lipschitz gradient, i.e.,
.
We adopt different strategies to handle these two distinct types of constraints in (3). For the convex constraint (3a), since
K is convex, we utilize the Moreau envelope for regularization. Recall that for a convex function
, the Moreau envelope with parameter
is defined as follows:
A key advantage of the Moreau envelope lies in its smoothing property: even if
g is nonsmooth,
remains smooth, and hence,
serves as a smooth approximation of
g. For the nonconvex constraint (3b), as mentioned above, since
D is nonconvex, the augmented Lagrangian function requires sufficiently large penalty parameters, which in turn leads to poor numerical performance. To overcome this drawback, we choose to remain this constraint explicitly rather than relaxing it into the objective function. Based on these considerations, in this paper we mainly study the following partial augmented Lagrangian function:
where
is the multiplier and
is a penalty parameter. The key difference between the proximal Lagrangian and the partial augmented Lagrangian lies in their treatment of nonconvex constraints: the former generally relaxes or regularizes nonconvex constraints by introducing a proximal term, whereas the latter retains nonconvex constraints directly within the constraint set and applies relaxation only to the convex part. The modified Lagrangian of this form has been studied in [
8] for sparse optimization problems, where
D denotes the sparse constraint, i.e.,
. In this case,
D is a union of polyhedra, but in this paper, we require
D to be merely a closed set, without any additional structure. In addition, when analyzing the second-order variational geometry of a set, it typically requires this set to be second-order regular or its second-order tangent set being nonempty [
15,
16,
17]. Our approach eliminates this requirement.
The main work of this paper is summarized in the following three aspects:
- (i)
Saddle points and Karush–Kuhn–Tucker (KKT) conditions. We first analyze the local/global saddle points of the partial augmented Lagrangian function (
5). The relationships between saddle points, minimizers, and KKT points of problem (3) are established. In particular, if
is a local/global saddle point of
, then
is a local/global minimizer of problem (3). Furthermore,
with
becomes a KKT point, provided that the metric subregularity constraint qualification (MSCQ) holds. Conversely, if
is a KKT point, then
is a global saddle point of
for all
, as problem (3) is convex, where we only require the set
K to be closed under addition, not necessarily to be a cone. The relationship between saddle points and the dual problem associated with the partial augmented Lagrangian is discussed.
- (ii)
Second-order analysis for
functions. According to the definition of saddle points (
15) below, we know that
is a minimizer of the following problem:
where
. Note that since
f and
G are
functions, then
also belongs to the
class, rather than being twice continuous differentiable. To establish second-order optimality conditions for problem (
6), we need to study the second-order approximation of
. Toward this end, we employ the second-order subdifferential
, defined as the coderivative of the gradient
. It enables us to obtain upper and lower bounds for the first-order Taylor expansion of
with respect to
x. In the process of theoretical analysis, the outer semicontinuity and local boundedness properties of
for
functions play an important role. Many works on second-order optimality conditions traditionally require the functions to be twice continuously differentiable; see [
7,
17,
18,
19,
20,
21]. Our work further relaxes this requirement from twice continuous differentiable to first-order continuous differentiability with Lipschitzian gradients, i.e., from
to
.
- (iii)
Second-order variational geometry of the constraint set
. Since the set
is explicitly retained in the constraint system, it is necessary to study its variational geometric properties. The traditional approach for describing second-order geometric information of sets utilizes the concept of second-order tangent sets. However, this set may be empty even for convex sets. To overcome this limitation, we further study the asymptotic second-order tangent cone. A key theoretical result is that the second-order tangent set and asymptotic second-order tangent cone cannot be empty simultaneously (Proposition 2.1 in [
22]). Therefore, the asymptotic second-order tangent cone serves as a supplement tool to second-order tangent sets, and their combination provides a complete characterization of the set’s second-order geometric information. By using the geometric analysis on constraint system and the aforementioned second-order analysis of objective function, we establish second-order optimality conditions for problem (
6).
The structure of this paper is organized as follows.
Section 2 presents fundamental notations and related results in the field of variational analysis.
Section 3 introduces the partial augmented Lagrangian function and investigates its saddle point properties.
Section 4 develops second-order optimality conditions for the partial augmented Lagrangian. Conclusion is drawn in
Section 5.
2. Basic Notations and Tools in Variational Analysis
In this section, we first recall some notations and fundamental results in variational analysis which are used throughout the paper.
Let
denote the closed unit ball in
. For a nonempty set
S, the support function is defined by
The indicator function is defined as follows:
if
, and
otherwise. The metric projection of a point
x onto the set
S is denoted by
. For a set-valued mapping
, its graph and inverse mapping are
and
The Painlevé-Kuratowski upper/outer limit of a set-valued mapping
F at a point
x is defined as
The Bouligand-Severi tangent/contingent cone to a closed set
S at a point
is
The Fréchet normal cone and the limiting/Mordukhovich/basic normal cone of
S at
are given by
where
represents the convergence of
x to
with
. If
S is convex, the Fréchet and limiting normal cones coincide. For a given direction
, the limiting normal cone to
S in direction
d at
is defined as
If, in particular,
, then
coincides with
.
Now we are ready to review two kinds of second-order tangent sets, both of which play fundamental roles in the second-order analysis later.
Definition 1 ([
16,
22])
. Let , and .- (i)
The outer second-order tangent set to S at in direction d is defined by - (ii)
The asymptotic second-order tangent cone to S at in direction d is defined by
The asymptotic second-order tangent cone was first introduced by Penot [
22] in the study of optimality conditions for scalar optimization. Note that the asymptotic second-order tangent cone is indeed a cone, while the second-order tangent set may not be a cone and even it may be empty; see, e.g., Bonnans and Shapiro (Example 3.29 in [
16]). An important fact is that both sets cannot be empty simultaneously (Proposition 2.1 in [
22]), i.e.,
. From this point, these two sets describe second-order information of the involved set sufficiently. In addition, according to definitions, it is easy to see
The following is the concept of directional neighborhood.
Definition 2 ([
23])
. Given a direction and positive numbers , the directional neighborhood of direction d is defined as follows: The concept of directional metric subregularity, as a directional version of constraint qualifications, plays an important role in developing variational geometric properties of constraint systems.
Definition 3 (Directional Metric Subregularity, [
23])
. Given a multifunction and , the mapping M is said to be metrically subregular at in direction , if there are positive numbers such thatThe infimum of κ over all the combinations , and κ satisfying the above relations is called the modulus of directional metric subregularity. In the case of , we simply say that M is metric subregular at . For the constraint system , we say that the metric subregularity constraint qualification (MSCQ) holds at in direction d if the set-valued mapping is metric subregularity at in direction d.
Lemma 1 (Lemma 2.2, [
21])
. Let and assume that MSCQ holds in direction for the constraint system with modulus . Then, The relations between second-order tangent set and asymptotic second-order tangent cone of the sets C and D, where , under directional MSCQ are given below.
Lemma 2 (Proposition 2.2, [
21])
. Let and . ThenandIf, in addition, assume that MSCQ holds in directional for the constraint system with modulus , then (9) and (10) hold as equality and the following estimationsand Let
be a single-valued map, and suppose that
satisfies
. The limiting (Mordukhovich) subdifferential of
at
is defined as
where
stands for the epigraph of
.
Definition 4 ([
24])
. Let be a multifunction and . The limiting (Mordukhovich) coderivative of F at is a multifunction with the valuesIf , one puts for any . We simple write when F is single-valued at and . By employing the notion of coderivative, we can establish the second-order generalized differential theory for extended-real-valued functions. This theoretical framework has become increasingly significant in areas such as second-order conditions, stability theory, and algorithmic analysis [
25,
26,
27].
Definition 5 ([
24])
. Let be a function with a finite value at . For any , the map with the valuesis said to be the limiting (Mordukhovich) second-order subdifferential of ϕ at relative to . If is a singleton, we simple write for convenience. If
is twice continuously differentiable in a neighborhood of
, by Proposition 1.119 in [
28],
where
denotes the Hessian matrix of
at
. Denote by
the class of real-valued functions
, which are Fréchet differentiable, and the gradient mapping
is locally Lipschitz. According to Proposition 1.120 in [
28], if
, one has
Lemma 3 (Proposition 2.6, [
29])
. Let . The following assertions hold:- (i)
For any , one has ,
- (ii)
For any the mapping is locally bounded. Moreover, if for all , then .
This following result establishes upper and lower bounds for the first-order Taylor approximation of functions by employing the limiting second-order subdifferential.
Lemma 4 (Theorem 3.1, [
30])
. Let and . Then, there exist where , and where , such that 3. Saddle Points and KKT Conditions
The set of minimum points corresponding to the Moreau envelope
(
4) is defined as
In particular, if
g is convex, then
is Fréchet differentiable and its gradient
is
Lipschitz-continuous. Because the set
K considered in this paper is convex, then
is continuously differentiable with respect to
x and the derivative takes the form
Since the projection operator
is Lipschitz with constant 1, then
is Lipschitz as well. Hence the partial augmented Lagrangian function
belongs to
.
For simplicity, we say that
is a Karush–Kuhn–Tucker (KKT) point of problem (3) if there exists
satisfying the following condition:
The definition of local and global saddle points is given below.
Definition 6 (Local/Global saddle point)
. A pair with is called a local saddle point of the partial augmented Lagrangian if there exists and a neighborhood U of such thatIf the restriction of U is omitted, then the pair is called a global saddle point and the infimum of all such is denoted by . Notice that there are two inequalities in the definition of saddle points. We begin by analyzing the first inequality.
Lemma 5. For a given , suppose that there exists such thatThen the following results hold. - (i)
is a feasible solution of problem (3);
- (ii)
;
- (iii)
.
Proof. (i) Suppose on the contrary that is not a feasible point of problem (3). Since by assumption, i.e., the constraint condition holds, then it remains to consider the case of .
Let
Then
since
. It follows from Example 6.16 in [
31] that
further implying
for all
, since
is cone. Note that
by (
17). This together with (
18) implies
where the second step above comes from the fact that
. Thus, we obtain from (
17) and (
19) that
Furthermore, as
, it follows from (
16) and (
20) that
Taking the limit as
in (
21) and using the fact that
leads to a contradiction with the finiteness of
by (
16). Thus,
.
(ii) If
x satisfies
and
, then
Hence,
In addition,
Since
is a feasible point by (i), according to (
22) and (
23), we have
Combining (
16) and (
24) yields
Thus,
for all
.
(iii) According to (
25), we have
which implies
because both
and
belong to
K and the projection onto a convex set is unique. Hence,
This completes the proof. □
In the above process of proof, it can be seen that
i.e., (
22) holds as an equality whenever
. In fact, since
, then
, and hence,
Lemma 6. If is a local/global saddle point of , then is a locally/globally optimal solution of problem (3).
Proof. The whole proof is divided into two parts. The first part proves that is a feasible point of problem (3) and the second part proves that is a local/global minimizer. Here we only consider the case of local saddle points, because the case of global saddle points can be proved by replacing the neighborhood U appeared in the following analysis by the whole space .
(i). Since
is a local saddle point of the partial augmented Lagrangian function
, then there exists
and a neighborhood
U of
such that
According to Lemma 5 (i), we know that
is a feasible solution of problem (3).
(ii). Since
is a local saddle point of the partial augmented Lagrangian function
, then for a feasible point
x of problem (3) satisfying
and
, one has
where the first equality comes from Lemma 5 (ii), the first inequality follows from (
27), and last step is due to (
22). Thus
is a locally optimal solution to the problem (3). □
The dual problem associated with the partial augmented Lagrangian
for problem (3) is defined as
where
and
. From (
22), the weak duality property between the primal problem (3) and its dual problem holds, i.e.,
where
denotes the feasible set of problem (3).
We say that the zero duality gap property holds for the partial augmented Lagrangian
, if
The relationship between saddle points and the zero duality gap property is given below.
Theorem 1. (i) If is a saddle point of the partial augmented Lagrangian , then for any , the pair is an optimal solution of the dual problem (28), and the zero duality gap property holds. (ii) If is an optimal solution of the dual problem (28), is an optimal solution of problem (3), and the zero duality gap property holds, then the pair is a global saddle point of and .
Proof. (i) Let
be a saddle point of
. For
, applying Lemma 5 yields
By the definition of
, it follows that for all
and
Taking the supremum over all
, we obtain from (
30) that for all
,
which implies
Therefore,
is an optimal solution of the dual problem (28), and the zero duality gap property holds, since
is an optimal solution of problem (3) by Lemma 6.
(ii) Let
be an optimal solution of the dual problem (28),
be an optimal solution of problem (3), and suppose that zero duality gap property holds, i.e.,
Note that
is nondecreasing in
for all
. It follows that
for all
. Since
is feasible, then
for any
. Therefore, by the definition of
, we have
which implies
. Since
is feasible for problem (3), it follows from (
29) that for all
Combining (
31) and (
32) yields
This means that
is a global saddle point of
and
. □
Under the metric subregularity constraint qualification, saddle points necessarily satisfy the KKT conditions.
Theorem 2. If is a local/global saddle point of , and the metric subregularity constraint qualification (MSCQ) holds at for the system , then there exists such that satisfies the KKT conditions for problem (3).
Proof. Since
is a local/global saddle point of
, then
by Lemma 5 (iii), which further implies
. Thus, according to (
13), we have
Under the MSCQ condition, it follows from (
8) that
Since
is continuously differentiable, then
where the second equality follows from (
33) and
by Exercise 8.14 in [
31].
Since
is a local/global saddle point of
, it follows from the definition that
is a local/global minimizer of
over
, i.e.,
is a local/global minimizer of
. By applying Fermat’s rule generalized (see Theorem 10.1 in [
31]), we have
where the second step follows from the fact that
since
, and the third step comes from (
35), and the last step is due to (
34).
The formula (
36) ensures the existence of
such that
This together with the fact
as shown in Lemma 5 (iii) yields that
satisfies the KKT condition (
14) for problem (3). □
In the remainder of this section, we show that the converse of Theorem 2 is valid when problem (3) is convex.
Definition 7. We say that problem (3) is convex if f is convex, the sets K and D are closed convex sets satisfying the additive closure properties:and the mappings G and H are convex with respect to K and D, respectively, i.e., for all and , we havewhere means . Here are some examples of convex sets that satisfy the additivity property.
Example 1. (i). The set K is a convex cone. It is well-known that for any convex cone K, the property holds, meaning that additivity is satisfied.
- (ii).
The set , where for , is a closed convex set that satisfies , but it is not a convex cone.
- (iii).
Let and . Define the set Then K is a closed convex set such that , but K is not a convex cone.
Let . Clearly f is continuous on the nonnegative orthant , which implies that K is closed. Note first that for any , we must have for all i, due to . Thus, K is a subset of the positive orthant . The function f is concave on , as it is the geometric mean function (Examples 3.1.5 in [32]). Hence, K is convex, since it is a superlevel set of a concave function. Now, take and consider their sum . Clearly, for all i. It remains to show that . By the inequality of arithmetic and geometric means, we have for each i. Therefore, Taking the th power on both sides yields Thus, , which implies . Note that K is not a cone. In fact, take any such that (for example, ). For , we have , so . Thus, K is not closed under scalar multiplication and hence is not a cone.
The following results show that the Moreau envelope preserves convexity properties under certain conditions.
Lemma 7. Let and K be a closed convex set. Assume that G is a convex mapping with respect to K with . Then andis convex, where . Proof. Since
G is a convex mapping with respect to
K, it follows from (
37) that for any
and
, we have
This ensures that
which is equivalent to the statement that if
, then
as well. Suppose on the contrary that
, then
where the last step is due to the fact that
. This leads to a contradiction with
. Note that
is convex, since
K is convex. This together with (
39) yields
which means that
is convex.
Since
K is convex, it follows from (
12) that
is differentiable with the gradient
Hence, for
, we have
where the fifth step follows from the fact that the metric projection is a non-expansive mapping, i.e.,
Thus, we show that
is monotone by (
40). Taking into account Theorem 12.17 in [
31], the function
is convex.
Note that
where the inequality comes from that fact that
by (
38) and
by assumption.
Since
is convex as shown above, we have
This together with (
41) yields
i.e.,
. This completes the proof. □
Theorem 3. Suppose that problem (3) is convex. If satisfies the KKT conditions, then is a global saddle point of for all .
Proof. Taking into account the convexity of problem (3) and Lemma 7, we know that the function
is convex. Hence,
where the first step follows from the fact
since
D is convex, and the second step comes from
by Exercise 8.14 and Theorem 10.6 in [
31].
Define
. Note that the function
is convex, since
is composed of
,
, and
, and these functions are convex with respect to
x by Lemma 7. Since
satisfies the KKT conditions (
14), then
and
. It then follows from (
33) that
where the forth step is ensured by (
42).
Since
is convex and
by (
43), we know that
is a global optimal solution of
. Therefore,
On the other hand, for any
and
, since
is feasible and
, it follows from (
22) and (
26) that
Therefore, putting (
44) and (
45) together yields
which means that
is a global saddle point of
. □
4. Optimality Conditions for Partial Augmented Lagrangian
In the previous section, we have discussed the first inequality in the definition of saddle points (
15). Now, we turn our attention to the second inequality that appears in (
15). This inequality indicates that
is a local optimal solution of the following optimization problem
The concept of directional local minimizer is given below.
Definition 8. A point is said to be a local optimal solution of (46) in direction , if there exist positive numbers such that The following result establishes the first-order optimality conditions for problem (
46).
Theorem 4. Let be a local optimal solution of problem (46) in direction . For the constraint system , suppose that MSCQ holds at in direction with modulus . Then
- (i)
- (ii)
if , then there exists such that
Proof. Since
is continuously differentiable with respect to
x, applying Proposition 3.1 in [
21] yields
- (a)
;
- (b)
if , then .
By (
13), we know
which together with (a) yields the desired conclusion (i).
According to (b) and Lemma 1, there exists
such that
and
. The desired results hold by further utilizing the formula of
given in (
47). □
The second-order necessary condition is obtained using second-order tangent sets and asymptotic second-order tangent cones.
Theorem 5. Let be a local optimal solution of problem (46) in direction with . Then,
- (i)
- (ii)
for any , there exists such that
Proof. (i) Pick
. Then, there exist
and
such that
and
for all
. Since
then
whenever
k is sufficiently large, and hence
. Note that
. Then
Since
, it follows from Lemma 4 that there exists
, where
, such that
According to Lemma 3, we have
which implies
for some
This together with (
48) and (
49) yields
By dividing both sides of the above inequality by
, we get
We claim that
is bounded. Let
. According to (
11) and (
50), we have
Note that
where
L is the Lipschitz constant of
. This implies that the subdifferential
is included in
. Since
by (
52), then
is bounded. We can assume without loss of generality that
. Since
, then
by Lemma 3. Using
and taking the limit in both sides of (
51) yields
(ii) Pick
. Then, there exist
and
such that
for all
. By an argument similar to that used for (
51) in case (i), we can obtain
where
and
. Taking limits yields
This completes the proof. □
Corollary 1. Let be a local optimal solution of problem (46) in direction with . For the constraint system , suppose that MSCQ holds at in direction d. Then,
- (i)
for every satisfying , we have .
- (ii)
for every satisfying , there exists such that .
Proof. The results follows immediately by applying Lemma 2 and Theorem 5. □
The following result develops second-order necessary conditions in terms of support functions.
Theorem 6. Let be a local optimal solution of problem (46) in direction with . If satisfies , then
- (i)
- (ii)
for each , there exists such that
Proof. (i) Pick
, i.e., there exists
such that
. Hence,
where the third step comes from the fact
by assumption and the last step follows from Theorem 5 (i).
(ii) If
, then the corresponding support function takes the value
, and in this case, (
53) holds trivially. If
, then there exists
such that
According to Theorem 5, for the above
, there exists
such that
Therefore,
where the third equality comes from the fact
by assumption, the fifth equality follows from (
54), and the last step is due to (
55). □
Lemma 8 (Lemma 3.4, [
33])
. Let . If the sequence converges to such that converges to some nonzero vector , where , then either converges to some vector , or there exists a sequence such that and converges to some vector , where denotes the orthogonal subspace to d. The second-order sufficient condition is derived below.
Theorem 7. Let be a feasible solution of problem (46), i.e., . Let with . Suppose that there exists such that and
- (i)
- (ii)
Then, the second-order growth conditions holds at in direction d, i.e., there exist such that Proof. Suppose on the contrary that the second-order growth condition in direction
d does not hold at
. This means that there exists sequence
such that
and
Pick
with
satisfying (i) and (ii). Let
and
. Then
and we can assume without loss of generality that
. Lemma 8 ensures that one of the following conditions hold:
- (a)
converges to some vector ;
- (b)
there exists a sequence such that and converges to some vector .
Case (1). If condition (a) holds, then
. Since
, it follows from Lemma 4 that there exists
, where
, such that
Note that
where the last step is due to
by assumption. Hence, it follows from (
57) and (
58) that
According to Lemma 3, we have
Hence,
with
. Following a similar argument as given for (
52), we can assume without loss of generality that
converge to some
by Lemma 3 (i). Hence
for some
.
It follows from (
56) and (
59) that
Since
, then
Putting (
60) and (
61) together gives
Dividing both sides of the above equation by
and letting
yields
Since
, then
and hence,
Note that
where the first step is due to
by assumption, the third step follows from (
64), and the last step comes from (
63). Hence, it follows from (
65) that
Recall that
and
. Hence applying (
7), the above formula can be rewritten as
which is a contradiction to assumption (ii).
Case (2). If condition (b) holds, then
. By following a similar argument on the formula (
62) as given in case (1), we can obtain
where
. Dividing both sides of the above equation by
yields
Taking the limit as
and noting that
yields
This together with the fact
by assumption implies
which is a contradiction to condition (i). □
Corollary 2. Let be a feasible solution of problem (46), i.e., , and let such that . Suppose that there exists satisfying and
- (i)
- (ii)
Then, the second-order growth conditions holds at in direction d.
Proof. The results follows immediately by applying Lemma 2 and Theorem 7. □
Example 2. Consider the following optimization problem:where For , the partial augmented Lagrangian isand its gradient is Let . Note that and . Hence, if and lies in some neighborhood of , then , which implies . So, For any nonzero direction , we show that either or and the second-order condition in Corollary 2 holds. In fact, since , then . Consider the following two cases. If , then . If , then by direct calculation, we obtain , , and
Let . Then, and for all . Hence, the condition (i) in Corollary 2 holds.
Define . Then,For , the gradient of ϕ is , and for , it is . Therefore, the subdifferential of ϕ at is given byIt is straightforward to compute thatFor any , we have with if , or if . So,The support function of the second-order tangent set at isCombining (66)–(68)
yieldsSo, the condition (ii) in Corollary 2 holds. According to the above analysis, we can conclude that is a local minimizer of . This establishes the second inequality in (15). The first inequality in (15) follows from the fact that by (26) since , and that for all by (22) due to being feasible. This shows that is a local saddle point of the problem.