846
Views
2
CrossRef citations to date
0
Altmetric
Articles

On the Aubin property of solution maps to parameterized variational systems with implicit constraints

&
Pages 1681-1701 | Received 30 Oct 2018, Accepted 06 Aug 2019, Published online: 28 Aug 2019

Abstract

In the paper, a new sufficient condition for the Aubin property to a class of parameterized variational systems is derived. In these systems, the constraints depend both on the parameter as well as on the decision variable itself and they include, e.g. parameter-dependent quasi-variational inequalities and implicit complementarity problems. The result is based on a general condition ensuring the Aubin property of implicitly defined multifunctions which employs the recently introduced notion of the directional limiting coderivative. Our final condition can be verified, however, without an explicit computation of these coderivatives. The procedure is illustrated by an example.

AMS CLASSIFICATIONS:

1. Introduction

The Aubin (Lipschitz-like) property is probably the most important extension of the Lipschitz continuity to multifunctions. It has been introduced by J.-P. Aubin in [Citation1] (under a different name) and since that time it is widely used in variational analysis and its applications. In [Citation2], a new condition has been derived ensuring the Aubin property of implicitly defined multifunctions around a given reference point. To be precise, it concerns the multifunction S:RmRn defined by (1) S(p):={xRn|0M(p,x)},(1) where p is the parameter, x is the decision variable and multifunction M:Rm×RnRl is given. In the form (Equation1), we can write down a large class of parameterized optimization and equilibrium problems and so this condition can well be used, e.g. in post-optimal analysis (where p corresponds to uncertain problem data) or in problems with the so-called equilibrium constraints (where p represents the control variable).

The application of this condition requires, however, the computation of the graphical derivative and the directional limiting coderivative of M, which may be quite demanding, e.g. in case of solution maps to variational systems. In [Citation3], the authors investigated from this point of view a class of variational systems, in which the (rather general) constraints did not depend on the parameter p. In this paper, we intend to make a further step and consider a variational system, where the constraint set depends both on the parameter as well as on the decision variable itself. This generality permits to analyse the Aubin property of, among other things, rather complicated parameterized quasivariational inequalities (QVIs). The used model comes from [Citation4] where, to its analysis, the authors employed some advanced tools of the generalized differential calculus of B. Mordukhovich [Citation5,Citation6]. Among the results of [Citation4], one finds also a sufficient condition for the Aubin property of the associated solution map and our main aim here is the sharpening of that condition on the basis of the results from [Citation2]. At the same time, despite of the increased complexity of the considered model, our final condition (Theorem 5.3) seems to be more workable than its counterpart in [Citation3, Theorem 5].

The paper is organized as follows. In Section 2, we recall the used notions from variational analysis and formulate properly the considered problem. Sections 3 and 4 deal with the computation of the graphical derivative and the directional limiting coderivative of the investigated multifunction M, respectively. The results presented in Section 4 depend heavily on selected results from [Citation7], where a rich calculus for directional limiting notions (normal cones, subdifferentials and coderivatives) has been developed. The main results of the paper are then collected in Section 5. They include both the new criterion (sufficient condition) for the Aubin property of the solution map to the investigated variational system as well as a formula for the graphical derivative of this solution map which may be used, e.g. in some sensitivity issues. The usage of the suggested technique is illustrated by an academic example.

The following notation is employed. Given a set ARn, spA stands for the linear hull of A and A is the (negative) polar of A. For a convex cone K, lin K denotes the lineality space of K, i.e. the set KK. Further, B,S is the unit ball and the unit sphere, respectively. Given a vector aRn, [a] is the linear space generated by a and [a] stands for the orthogonal complement to [a]. Finally, A means the convergence within a set A and Limsup stands for the Painlevé–Kuratowski set limit.

2. Problem formulation and preliminaries

In the first part of this section, we introduce some notions from variational analysis which will be extensively used throughout the whole paper. Consider first a general closed-graph multifunction F:RnRz and its inverse F1:RzRn and assume that (u¯,v¯)gphF.

Definition 2.1

We say that F has the Aubin property around (u¯,v¯), provided there are neighbourhoods U of u¯, V of v¯ and a constant κ>0 such that F(u1)VF(u2)+κu1u2Bfor all u1,u2U.F is said to be calm at (u¯,v¯), provided there is a neighbourhood V of v¯ and a constant κ>0 such that F(u)VF(u¯)+κuu¯Bfor all uRn.

It is clear that the calmness is substantially weaker (less restrictive) than the Aubin property. Furthermore, it is known that F is calm at (u¯,v¯) if and only if F1 is metrically subregular at (v¯,u¯), i.e. there is a neighbourhood V of v¯ and a constant κ>0 such that (2) dist(v,F(u¯))κdist(u¯,F1(v)) for all vV,(2) cf. [Citation8, Exercise 3H.4].

To obtain directional versions of the above properties, consider a direction dRz, positive numbers ϱ,δ and define the set Vϱ,δ(d):=vϱB dvvdδvd. We say that a set V is a directional neighbourhood of d if there exist ϱ,δ>0 such that Vϱ,δ(d)V. Now, when the neighbourhood V in (Equation2) is replaced by the set v¯+V, we say that F1 is metrically subregular at (u¯,v¯) in direction d.

To conduct a thorough analysis of the above stability notions, one typically makes use of some basic notions of generalized differentiation, whose definitions are presented below.

Definition 2.2

Let A be a closed set in Rn and x¯A. Then

  1. TA(x¯):=Limsupt0(Ax¯)/t is the tangent (contingent, Bouligand) cone to A at x¯ and NˆA(x¯):=(TA(x¯)) is the regular (Fréchet) normal cone to A at x¯.

  2. NA(x¯):=Limsupxx¯ANˆA(x) is the limiting (Mordukhovich) normal cone to A at x¯ and, given a direction dRn, NA(x¯;d):=Limsupddt0NˆA(x¯+td) is the directional limiting normal cone to A at x¯ in direction d.

If A is convex, then NˆA(x¯)=NA(x¯) amounts to the classical normal cone in the sense of convex analysis and we will write NA(x¯). By the definition, the limiting normal cone coincides with the directional limiting normal cone in direction 0, i.e. NA(x¯)=NA(x¯;0), and NA(x¯;d)= whenever dTA(x¯).

In the sequel, we will also make an extensive use of the so-called critical cone. In the setting of Definition 2.2 with an additionally given vector dRn, the cone KA(x¯,d):=TA(x¯)[d] is called the critical cone to A at x¯ with respect to d.

The above listed cones enable us to describe the local behaviour of set-valued maps via various generalized derivatives. Consider again the multifunction F and the point (u¯,v¯)gphF.

Definition 2.3

  1. The multifunction DF(u¯,v¯):RnRz, defined by DF(u¯,v¯)(d):={hRz|(d,h)TgphF(u¯,v¯)},dRn is called the graphical derivative of F at (u¯,v¯).

  2. The multifunction DˆF(u¯,v¯):RzRn, defined by DˆF(u¯,v¯)(v):={uRn|(u,v)NˆgphF(u¯,v¯)},vRz is called the regular (Fréchet) coderivative of F at (u¯,v¯).

  3. The multifunction DF(u¯,v¯):RzRn, defined by DF(u¯,v¯)(v):={uRn|(u,v)NgphF(u¯,v¯)},vRz is called the limiting (Mordukhovich) coderivative of F at (u¯,v¯).

  4. Given a pair of directions (d,h)Rn×Rz, the multifunctionDF((u¯,v¯);(d,h)):RnRz, defined by DF((u¯,v¯);(d,h))(v):={uRn|(u,v)NgphF((u¯,v¯);(d,h))},vRz is called the directional limiting coderivative of F at (u¯,v¯) in direction (d,h).

For the properties of the cones TA(x¯), NˆA(x¯) and NA(x¯) from Definition 2.2 and generalized derivatives (i), (ii) and (iii) from Definition 2.3, we refer the interested reader to the monographs [Citation9] and [Citation6]. The directional limiting normal cone and coderivative were introduced by the first author in [Citation10] and various properties of these objects can be found also in [Citation2] and the references therein. Note that DF(u¯,v¯)=DF((u¯,v¯);(0,0)) and that domDF((u¯,v¯);(d,h))= whenever hDF(u¯,v¯)(d).

The above notions enable us to come back to the solution map (Equation1) and state the (already announced) sufficient condition for the Aubin property of S around (p¯,x¯) from [Citation2].

Theorem 2.1

[Citation2, Theorem 4.4]

Let M have a closed graph and assume that

  1. {kRn|0DM(p¯,x¯,0)(h,k)}for all hRm;

  2. M is metrically subregular at (p¯,x¯,0);

  3. For every nonzero (h,k)Rm×Rn verifying 0DM(p¯,x¯,0)(h,k) one has the implication (q,0)DM((p¯,x¯,0);(h,k,0))(v)q=0.Then S has the Aubin property around (p¯,x¯) and for any hRm DS(p¯,x¯)(h)={k|0DM(p¯,x¯,0)(h,k)}. The above assertions remain true provided assumptions (ii), (iii) are replaced by

  4. For every nonzero (h,k)Rm×Rn verifying 0DM(p¯,x¯,0)(h,k) one has the implication (q,0)DM((p¯,x¯,0);(h,k,0))(v)q=0,v=0.

We are now ready to proceed to the proper problem formulation. As announced in the Introduction, this paper is devoted to solution maps of a class of variational systems in which (3) M(p,x):=f(p,x)+NˆΓ(p,x)(x),(3) with f:Rm×RnRn being continuously differentiable and Γ:Rm×RnRn given via (4) Γ(p,x)={yRn|q(p,x,y)D}.(4) In (Equation4), q:Rm×Rn×RnRs is twice continuously differentiable and DRs is convex and polyhedral.

Consider the reference point (p¯,x¯) from the graph of the solution map S and, to unburden the notation, let us introduce the functions q~:Rm×RnRs and b:Rm×RnRs×n by q~(p,x)=q(p,x,x)and b(p,x)=3q(p,x,x). Throughout the whole paper, we will impose the following assumption:

  1. The implication (5) b(p¯,x¯)Tλ=0λspND(q~(p¯,x¯))λ=0(5) is fulfilled.

(A) entails in particular that the generalized equation (GE) (6) 0f(p,x)+NˆΓ(p,x)(x)(6) is locally, around (p¯,x¯), equivalent with the (possibly simpler) GE (7) 0f(p,x)+b(p,x)TND(q~(p,x))(7) which will be used as our basic model in the whole development. Indeed, as argued in [Citation4], this follows from a slight modification of amenability results in [Citation9, Chapter 10.F] when applied to the set Γ(p¯,x¯) at x¯Γ(p¯,x¯). In fact, this equivalence holds true even under a relaxation of (A), where the second line on the left-hand side of (Equation5) is replaced by λND(q~(p¯,x¯)). Note that this relaxed condition is imposed in [Citation4] instead of (A). Further note that both under assumption (A) and this relaxation we have NˆΓ(p,x)(x)=NΓ(p,x)(x) locally around (p¯,x¯), i.e. Clarke regularity holds.

Since D is polyhedral, (A) is equivalent with the nondegeneracy of Γ(p¯,x¯) at x¯, i.e. with the condition b(p¯,x¯)Rn+linTD(q~(p¯,x¯))=Rs. This follows from [Citation11, formula (4.172) and Example 3.139]. The polyhedrality of D implies further that we can employ the efficient representation of TgphND and its polar provided in [Citation12, Section 2].

Finally note that, given a yNˆΓ(p¯,x¯)(x¯), under (A) the relations (8) y=b(p¯,x¯)Tλ,λND(q~(p¯,x¯))(8) have a unique solution λ. Thanks to this fact, most formulas in the sequel are substantially simplified.

To derive the announced new criterion for the Aubin property of solution maps given by (Equation1) and (Equation3), we will in the first step apply Theorem 2.1 to GE (Equation7). The needed graphical derivative and directional limiting coderivative of the respective mapping M are computed in the next two sections.

3. Computation of the graphical derivative

The right-hand side of (Equation7) amounts to the sum of a smooth single-valued function f and the multifunction Q:Rm×RnRn defined via Q(p,x):=b(p,x)TND(q~(p,x)). The graphical derivative of Q is related with the one of the mapping Ψ:Rm×Rn×RnRn given by Ψ(p,x,y):=NˆΓ(p,x)(y). Note that Q(p¯,x¯)=Ψ(p¯,x¯,x¯). In what follows, we denote z¯:=(p¯,x¯,x¯) and for any z=(p,x,y)Rm×Rn×Rn we denote by π3 the canonical projection of z on its third component, i.e. π3(z)=y.

Proposition 3.1

Under assumption (A) for all yΨ(z¯) and all w=(h,k,l) we have (9) DΨ(z¯,y)(w)=(3q()Tλ)(z¯)w+π3(NKgphΓ(z¯,q(z¯)Tλ)(w))(9) (10) =(3q()Tλ)(z¯)w+3q(z¯)TNKD(q(z¯),λ)(q(z¯)w),(10) where λ is the unique solution of the system (11) 3q(z¯)Tλ=y, λND(q(z¯)).(11)

Proof.

Assumption (A) implies the weaker condition 3q(p¯,x¯,x¯)Tμ=0,μND(q(p¯,x¯,x¯))μ=0 which in turn is equivalent with the metric regularity of the mapping yq(p¯,x¯,y)D around (x¯,0), see [Citation9, Example 9.44]. Hence, by [Citation13, Corollary 3.7] we deduce that the system q(p,x,y)D enjoys the so-called Robinson stability property at (p¯,x¯,x¯), i.e. there is a constant κ>0 together with neighbourhoods V of x¯ and W of (p¯,x¯) such that dist(y,Γ(p,x))κdist(q(p,x,y),D)yV,(p,x)W. Because D is convex and polyhedral, we can apply [Citation14, Theorem 5.3] to compute the graphical derivative DΨ(z¯,y)(w) resulting in (Equation9). Since NKgphΓ(z,q(z)Tλ)(w)=(q(z)T(ND(q(z))+[λ]))[w], we have π3(NKgphΓ(z¯,q(z¯)Tλ)(w))={3q(z¯)TηηND(q(z¯))+[λ],ηTq(z¯)w=0}. Next, by using the identity (ND(q(z¯))+[λ])[q(z¯)w]=(TD(q(z¯))[λ])[q(z¯)w]=KD(q(z¯),λ)[q(z¯)w]=NKD(q(z¯),λ)(q(z¯)w), we obtain (Equation10) and the proof is complete.

Remark 3.1

Since NKD(q(z¯),λ)(q(z¯)w)ND(q(z¯))+[λ]spND(q(z¯)), for every yΨ(z¯), every direction w=(h,k,l)Rm×Rn×Rn and every vDΨ(z¯,y)(w) there is a unique element η satisfying (12) 3q(z¯)Tη=v(3q()Tλ)(z¯)w, ηNKD(q(z¯),λ)(q(z¯)w).(12)

In (Equation10), we dispose with a workable formula for DΨ(z¯,y) in terms of q, D and the multiplier λ associated with y. This enables us in the next statement to find the relationship between DΨ(z¯,y) and DQ(z¯,y) which is essential to capture the implicit nature of the considered constraint system.

Theorem 3.1

For all yQ(p¯,x¯) and all (h,k)Rm×Rn we have (13) DQ((p¯,x¯),y)(h,k)DΨ(z¯,y)(h,k,k).(13) Conversely, for all yΨ(z¯), all (h,k)Rm×Rn and all vDΨ(z¯,y)(h,k,k) such that the mapping (14) F(p,x,μ):=(q~(p,x),μ)gphND(14) is metrically subregular at ((p¯,x¯,λ),0) in direction (h,k,η) with λ and η given by (Equation11) and (Equation12) with w:=(h,k,k), respectively, we have vDQ((p¯,x¯),y)(h,k).

Proof.

The inclusion (Equation13) follows immediately from the definition of the graphical derivative and there remains to show the second statement. Consider vDΨ(z¯,y)(h,k,k) such that F is metrically subregular at ((p¯,x¯,λ),0) in direction (h,k,η). Then there are sequences tν0, wν:=(hν,kν,lν)w and vνv such that y+tνvν3q(z¯+tνwν)TND(q(z¯+tνwν)) for all ν. Due to (A) there is for all ν sufficiently large a unique multiplier λνND(q(z¯+tνwν)) satisfying y+tνvν3q(z¯+tνwν)Tλν. The sequence λν is uniformly bounded yielding, together with y=3q(z¯)Tλ, that tνvν=3q(z+tνwν)Tλνy=(3q(z¯)+tν(3q(z¯))wν)Tλν3q(z¯)Tλ+o(tν)=3q(z¯)T(λνλ)+tν(3q()Tλν)(z¯)wν+o(tν) and, consequently, limν3q(z¯)Tλνλtν=limν(vν(3q()Tλν)(z¯)wν+o(tν)tν)=v(3q()Tλ)(z¯)w. Since D is a convex polyhedral set, we have λνND(q(z¯+tνwν))ND(q(z¯)) for all ν sufficiently large and therefore (λνλ/tν)ND(q(z¯))+[λ]spND(q(z¯)). By virtue of (A), we conclude that (λνλ/tν) is convergent to η. Since dist((q~(p¯+tνhν,x¯+tνkν),λν),gphND)q~(p¯+tνhν,x¯+tνkν)q(z¯+tνwν)+dist((q(z¯+tνwν),λν),gphND)=q(p¯+tνhν,x¯+tνkν,x+tνkν)q(p¯+tνhν,x¯+tνkν,x¯+tνlν)=O(tνkνlν)=o(tν), by the assumed directional metric subregularity we can find for every ν sufficiently large some pν,xν,λ~ν with 0F(pν,xν,λ~ν) and pν(p¯+tνhν)+xν(x¯+tνkν)+λ~νλν=o(tν). Thus y+tνvν=3q(p¯+tνhν,x¯ν+tνkν,x¯+tνlν)Tλν=3q(pν,xν,xν)Tλ~ν+o(tν)=b(pν,xν)Tλ~ν+o(tν). This equality, together with λ~νND(q~(pν,xν)), implies the inclusion vDQ((p¯,x¯),y)(h,k) and we are done.

To ensure the directional metric subregularity of (Equation14), we may use the sufficient condition presented in Proposition 3.2. Recall that F is a face of a polyhedral convex cone K provided for some vector zK one has F=Kz.

Proposition 3.2

Let λND(q~(p¯,z¯)), let (h,k)Rm×Rn be a pair of directions satisfying q~(p¯,x¯)(h,k)KD(q~(p¯,x¯),λ) and let ηNKD(q~(p¯,x¯),λ)(q~(p¯,x¯)(h,k)). Further assume that for every pair of faces F1,F2 of the critical cone KD(q~(p¯,x¯),λ) with q~(p¯,x¯)(h,k)F2F1[η] there holds q~(p¯,x¯)Tμ=0,μ(F1F2)  μ=0. Then the mapping F given by (Equation14) is metrically subregular at ((p¯,x¯,λ),0) in direction (h,k,η).

Proof.

We claim that F is even metrically regular at ((p¯,x¯,λ),0) in direction ((h,k,η),0). In order to show this claim we invoke the characterization of directional metric regularity from [Citation15, Theorem 1], which reads in our case as q~(p¯,x¯)Tμ=0, ξ=0, (μ,ξ)NgphND((q~(p¯,x¯),λ);(q~(p¯,x¯)(h,k),η)) μ=0, ξ=0. By [Citation2, Theorem 2.12], NgphND((q~(p¯,x¯),λ);(q~(p¯,x¯)(h,k),η)) amounts to the union of all product sets K×K associated with cones K of the form F1F2, where F1,F2 are faces of the critical cone KD(q~(p¯,x¯),λ) with q~(p¯,x¯)(h,k)F2F1[η]. Thus our claim about the directional metric regularity of F holds true and the statement is proved.

Of course, for the verification of the directional metric subregularity of (Equation14) one could employ also some non-directional less fine criteria mentioned, e.g. in [Citation4,Citation16].

To write down the final formula for the graphical derivative of M, we associate now with the considered variational system for fixed λRs the Lagrangian mapping Lλ:Rm×RnRn via Lλ(p,x):=f(p,x)+b(p,x)Tλ. Under the assumptions of Theorem 3.1, we then obtain the formula DM(p¯,x¯,0)(h,k)=Lλ¯(p¯,x¯)(h,k)+b(p¯,x¯)TNKD(q~(p¯,x¯),λ¯)(q~(p¯,x¯)(h,k)), where λ¯ is the unique solution of the system (15) f(p¯,x¯)+b(p¯,x¯)Tλ=0, λND(q~(p¯,x¯)).(15)

4. Computation of the directional limiting coderivative

Given a pair of directions (h,k)Rm×Rn, the aim of this section is to provide possibly sharp estimates of the sets DM((p¯,x¯,0);(h,k,0))(z). Due to [Citation2, formula (2.4)] and the local equivalence of GEs (Equation6) and (Equation7), we have for any vRn the equality (16) DM((p¯,x¯,0);(h,k,0))(v)=f(p¯,x¯)Tv+DQ(p¯,x¯,f(p¯,x¯));×(h,k,f(p¯,x¯)(h,k))(v).(16) It suffices thus to compute just the directional limiting coderivative of Q. To this purpose, we observe that Q(p,x)=S2S1(p,x), where S1:Rm×RnRm×Rn×Rs is given by S1(p,x):=pxND(q~(p,x)) and S2:Rm×Rn×RsRn is given by S2(u1,u2,u3):=b(u1,u2)Tu3. Consider the intermediate mapping Ξ:Rm×Rn×RnRm×Rn×Rs defined by Ξ(p,x,y):={(u1,u2,u3)S1(p,x)|y=S2(u1,u2,u3)}={(u1,u2,u3)|u1=p,u2=x,u3ND(q~(p,x)),b(p,x)Tu3=y}.

Lemma 4.1

Let yQ(p¯,x¯). Then Ξ(p¯,x¯,y)={(p¯,x¯,λ)} with λ (uniquely) given by (Equation8). Moreover, the values Ξ(p,x,v) are bounded for all (p,x,v)domΞ close to (p¯,x¯,y).

Proof.

The first statement is directly implied by (A); see the mention at the end of Section 2. The boundedness follows by a standard argumentation even from a relaxed condition b(p¯,x¯)Tλ=0λND(q~(p¯,x¯))λ=0, see [Citation4, p.396].

Lemma 4.2

Let d¯=(p¯,x¯,λ)=Ξ(p¯,x¯,y). Then the set {ξS|ξDS1(p¯,x¯,d¯)(0), 0=S2(d¯,y)(ξ)} is empty.

Proof.

Clearly, (17) DS1(p¯,x¯,d¯)(0)={(0,0,η)Rm×Rn×Rs|ηD(NDq~)(p¯,x¯,λ)(0,0)}(17) and the condition 0=S2(d¯,y)(ξ) amounts to (18) 0=(b(p¯,x¯)Tλ)ξ1ξ2+b(p¯,x¯)Tξ3.(18) By comparing (Equation17) and (Equation18), it follows directly that ξ1=0,ξ2=0 and it remains to show that the conditions ηD(NDq~)(p¯,x¯,λ)(0,0),0=b(p¯,x¯)Tη imply η=0. Clearly, Tgph(NDq~)(p¯,x¯,λ¯)(h,k,η)q~(p¯,x¯)(h,k)ηTgphND(q~(p¯,x¯),λ). It follows that ηD(NDq~)(p¯,x¯,λ)(0,0) implies that (0,η)TgphND(q~(p¯,x¯),λ). Due to the polyhedrality of D, one has (cf. [Citation12, page 1093]) TgphND(q~(p¯,x¯),λ)={(a,b)Rs×Rs|aKD(q~(p¯,x¯),λ),b KD(q~(p¯,x¯),λ),a,b=0}, from which we infer that ηND(q~(p¯,x¯))+[λ]sp ND(q~(p¯,x¯)). Consequently, ξ3=η=0 by virtue of (A) and we are done.

As the last auxiliary result, we will now estimate the directional limiting coderivative of S1. Clearly, S1=ΣΩ, where Ω:Rm+nRm+n×Rm+n is defined by Ω(p,x)=(p,x)(p,x)(two copies), and Σ:Rm+n×Rm+nRm+n×Rs is defined by Σ(a1,a2)=a1(NDq~)(a2).

Lemma 4.3

Consider a direction (h,k,η)Rm×Rn×Rs and a point λ such that (p¯,x¯,λ)S1(p¯,x¯). Then one has for any d=(d1,d2,d3)Rm×Rn×Rs the inclusion (19) DS1((p¯,x¯,λ);(h,k,(h,k,η)))(d)d1d2+D(NDq~)((p¯,x¯,λ);(h,k,η))(d3).(19)

Proof.

The statement follows from [Citation7, Corollary 5.1], provided we verify the respective subregularity condition. To this aim, we observe that the implication (20) 0Ω(p¯,x¯)T(a1,a2)(a1,a2)DΣ((p¯,x¯),(p¯,x¯),(p¯,x¯,λ))(0,0)a1=0,a2=0(20) is fulfilled. Indeed, the relations on the left-hand side of (Equation20) imply that a1+a2=0 and a1=0, whence a2=0 as well. On the other hand, implication (Equation20) is a strengthened (non-directional) variant of condition (Equation32) in [Citation7], which ensures the subregularity condition in [Citation7, Corollary 5.1]. We obtain thus that DS1((p¯,x¯,λ);(h,k,(h,k,η)))(d)Ω(p¯,x¯)TDΣ(((p¯,x¯),(p¯,x¯),(p¯,x¯,λ));((h,k),(h,k),(h,k,η)))(d), which directly leads to inclusion (Equation19).

We are now in position to compute an estimate of the directional limiting coderivative of Q at the point (p¯,x¯,y¯) in the direction (h,k,l).

Theorem 4.1

Let yRn be given and let λRs be (uniquely) given by the relations (Equation8) and ηRs be (uniquely) given by (21) l=((b(p¯,x¯)Tλ)(h,k)+b(p¯,x¯)Tη, η NK(q~(p¯,x¯),λ)(q~(p¯,x¯)(h,k)).(21) Assume that the mapping (Equation14) is metrically subregular at ((p¯,x¯,λ),0) in direction (h,k,η).

Then for any vRn one has the estimate (22) DQ((p¯,x¯,y);(h,k,l))(v)(b(p¯,x¯)Tλ)Tv+q~(p¯,x¯)TDND((q~(p¯,x¯),λ);(q~(p¯,x¯)(h,k),η))(b(p¯,x¯)v).(22)

Proof.

We observe first that by virtue of (A) and Lemma 4.1 all assumptions of [Citation7, Corollary 5.2] are fulfilled and, thanks to Lemma 4.1 and Lemma 4.2, the inclusion in [Citation7, formula (26)] simplifies to (23) DQ((p¯,x¯,y);(h,k,l))ξDS1(p¯,x¯,u¯)(h,k)l=S2(u¯)ξDS1((p¯,x¯,u¯);(h,k,ξ))S2(u¯)T,(23) where u¯=(p¯,x¯,λ). The directional limiting coderivative of S1 has been estimated in Lemma 4.3 and so we compute now the graphical derivative of S1 and the Jacobian of S2. Since (24) (p¯,x¯,λ)gph(NDq~)  (q~(p¯,x¯),λ)gphND,(24) we have (h,k,η)Tgph(NDq~)(p¯,x¯,λ) if and only if there are sequences tν0, (hν,kν,ην)(h,k,η) such that (q~(p¯+tνhν,x¯+tνkν),λ+tνην)gphNDfor all ν and it follows that Tgph(NDq~)(p¯,x¯,λ)R:=(h,k,η)q~(p¯,x¯)(h,k)ηTgphND(q~(p¯,x¯),λ). On the other hand, given (h,k,η)R, there is a sequence tν0 such that dist((q~(p¯,x¯)+tνq~(p¯,x¯)(h,k),λ+tνη),gphND)=o(tν). Hence, dist((q~(p¯+tνh,x¯+tνk),λ+tνη),gphND)=o(tν) and we can employ the metric subregularity of (Equation14) at ((p¯,x¯,λ),0) to obtain (pν,xν,λν) satisfying (pν,xν,λν)(p¯+tνh,x¯+tνk,λ¯+tνη)=o(tν), (q~(pν,xν),λν)gphND. We conclude (h,k,η)Tgph(NDq~)(p¯,x¯,λ) and RTgph(NDq~)(p¯,x¯,λ) follows. Hence Tgph(NDq~)(p¯,x¯,λ)=R and this relation, together with [Citation8, Example 4A.4], implies that DS1(p¯,x¯,u¯)(h,k)=(h,k,η)ηNKD(q~(p¯,x¯),λ)(q~(p¯,x¯)(h,k)). Further, by a simple calculation we obtain that S2(u¯)=(b(u¯1,u¯2)Tu¯3, b(u¯1,u¯2)T), and thus the union in (Equation23) is taken over all ξ=(h,k,η) satisfying (Equation21). The uniqueness of η follows from the comparison of (Equation21) with (Equation12) and Remark 3.1. From (Equation23) we get now the inclusion (25) DQ((p¯,x¯,y);(h,k,l))(v)((b(p¯,x¯)T)λ)Tv+D(NDq~)(p¯,x¯,λ);(h,k,η))(b(p¯,x¯)v),(25) and it remains to rewrite the second term on the right-hand side of (Equation25) in a more tractable form. To this aim, we invoke [Citation7, Corollary 3.2]. Indeed, applying the equivalence (Equation24) as in the computation of DS1(p¯,x¯,u¯)(h,k), under the assumed directional metric subregularity of mapping (Equation14), we obtain the inclusion D(NDq~)((p¯,x¯,λ);(h,k,η))(b(p¯,x¯)v)q~(p¯,x¯)TDND((q~(p¯,x¯),λ);(q~(p¯,x¯)(h,k),η))(b(p¯,x¯)v), and the proof is complete.

We can now combine inclusion (Equation22) with relation (Equation16) to obtain a formula for DM((p¯,x¯,0);(h,k,0)) in terms of problem data. To this purpose, we introduce y¯=f(p¯,x¯) and denote by λ¯ the (unique) solution of (Equation15). Under the assumptions of Theorem 4.1, we obtain the estimate DM(p¯,x¯,0);(h,k,0))(v)Lλ¯(p¯,x¯)Tv+q~(p¯,x¯)TDND((q~(p¯,x¯)λ¯);(q~(p¯,x¯)(h,k),η))(b(p¯,x¯)v), where η is (uniquely) given by 0=Lλ¯(p¯,x¯)(h,k)+b(p¯,x¯)Tη,ηNKD(q~(p¯,x¯),λ¯)(q~(p¯,x¯)(h,k)). This will be utilized in the next section.

5. On the Aubin property of the solution map

Combining Theorem 2.1 with the formulas for DM(p¯,x¯,z¯)(h,k) and DM((p¯,x¯,z¯);(h,k,0)) derived in Sections 3 and 4, respectively, we arrive at the following result.

Theorem 5.1

Assume that λ¯ is the (unique) solution of system (Equation15) and for every hRm there is some kRn and some ηRs such that (26) 0=Lλ¯(p¯,x¯)(h,k)+b(p¯,x¯)Tη,ηNKD(q~(p¯,x¯),λ¯)(q~(p¯,x¯)(h,k)).(26) Further assume that for every non-zero pair (h,k)Rm×Rn and the corresponding (unique) ηRs satisfying (Equation26) the mapping F given by (Equation14) is metrically subregular at ((p¯,x¯,λ¯),0) in direction (h,k,η) and the implication (27) (p,0)=Lλ¯(p¯,x¯)Tv+q~(p¯,x¯)TwwDND((q~(p¯,x¯),λ¯);(q~(p¯,x¯)(h,k),η))(b(p¯,x¯)v)p=0,v=0(27) is fulfilled. Then the solution map defined via (Equation1) and (Equation3) has the Aubin property around (p¯,x¯). Moreover, one has (28) DS(p¯,x¯)(h)=k0Lλ¯(p¯,x¯)(h,k)+b(p¯,x¯)TNKD(q~(p¯,x¯),λ¯)(q~(p¯,x¯)(h,k)).(28)

Following Theorem 2.1, the implication (Equation27) could be weakened by omitting the requirement v = 0 on its right-hand side. Then, however, we have to impose an additional requirement that the mapping (Equation3) is metrically subregular at (p¯,x¯,0).

If the constraint mapping q (and hence also q~) does not depend on p, then (Equation26) attains the form 0=Lλ¯(p¯,x¯)(h,k)+b(x¯)Tη,ηNKD(q~(x¯),λ¯)(q~(x¯)k) and (Equation27) reduces to a substantially more tractable form 0=2Lλ¯(p¯,x¯)Tv+q~(x¯)TwwDND((q~(x¯),λ¯);(q~(x¯)k,η))(b(x¯)v)v=0. The polyhedrality of D enables us to avoid the computation of directional limiting coderivatives and to replace the verification of (Equation27) by a simpler procedure. The key argument comes from the already mentioned [Citation2, Theorem 2.12].

Theorem 5.2

In the setting of Theorem 5.1 replace the implication (Equation27) by the assumption that for every pair of faces F1,F2 of the critical cone KD(q~(p¯,x¯),λ¯) with q~(p¯,x¯)(h,k)F2F1[η] there holds (29) 2q~(p¯,x¯)Tμ=0,μ(F1F2)1q~(p¯,x¯)Tμ=0.(29) and for every w0 with b(p¯,x¯)wF1F2 there is some w~ with 2q~(p¯,x¯)w~F1F2 and (30) wT2Lλ¯(p¯,x¯)w~>0.(30) Then all assertions of Theorem 5.1 remain valid.

Proof.

We shall show that the conditions (Equation29), (Equation30) imply (Equation27). Assume on the contrary that there is some direction (0,0)(h,k) verifying (Equation26) together with some ηRs and some pair (p,v)(0,0) satisfying (p,0)Lλ¯(p¯,x¯)Tv+q~(p¯,x¯)TDND((q~(p¯,x¯),λ¯);(q~(p¯,x¯)(h,k),η))(b(p¯,x¯)v). Next we utilize [Citation2, Theorem 2.12] to find two faces F1,F2 of KD(q~(p¯,x¯),λ¯) and μ(F1F2) satisfying q~(p¯,x¯)(h,k)F2F1[η],b(p¯,x¯)vF1F2, and (p,0)=Lλ¯(p¯,x¯)Tv+q~(p¯,x¯)Tμ. In particular, we have 2Lλ¯(p¯,x¯)Tv=2q~(p¯,x¯)Tμ and p=1Lλ¯(p¯,x¯)Tv+1q~(p¯,x¯)Tμ. If v0, then, by taking w = −v, the imposed assumptions imply the existence of some w~ with 2q~(p¯,x¯)w~F1F2 fulfilling condition (Equation30). This results in the contradiction 0<wT2Lλ¯(p¯,x¯)w~=μT2q~(p¯,x¯)w~0, where the last inequality follows from μ(F1F2) and 2q~(p¯,x¯)w~F1F2. Thus one has v = 0 implying 2q~(p¯,x¯)Tμ=0 and 0p=1q~(p¯,x¯)Tμ. But from (Equation29), we obtain 1q~(p¯,x¯)Tμ=0 and consequently p=0, a contradiction. Hence (Equation27) holds true.

The metric subregularity assumption arising in Theorem 5.1 can be ensured together with condition (Equation27) in an elegant way shown in the next statement.

Corollary 5.1

Assume that for every hRm there is some kRn and some ηRs satisfying (Equation26) and assume that for every nonzero (h,k)Rm×Rn,ηRs verifying (Equation26) and for every pair of faces F1,F2 of the critical cone KD(q~(p¯,x¯),λ¯) with q~(p¯,x¯)(h,k)F2F1[η] there holds (31) 2q~(p¯,x¯)Tμ=0,μ(F1F2)μ=0,(31) and for every w0 with b(p¯,x¯)wF1F2 there is some w~ with 2q~(p¯,x¯)w~F1F2 and (32) wT2Lλ¯(p¯,x¯)w~>0.(32) Then all assertions of Theorem 5.1 remain valid.

Proof.

The proof easily follows from the observations that (Equation31) implies both (Equation27) and the metric subregularity of F at ((p¯,x¯,λ¯),0) in direction (h,k,η) by virtue of Proposition 3.2.

We now give a simpler criterion for verifying condition (Equation31). Consider the following lemma.

Lemma 5.1

Let KRs be a convex polyhedral cone and let vK. Then for every pair F1,F2 of faces of K with vF2F1 there holds (F1F2)spNK(v).

Proof.

Since K is assumed to be convex polyhedral, there are finitely many vectors a1,,atRs such that K={uRsaiTu0, i=1,,t}. Consider two faces F1,F2 of K satisfying vF2F1. Then we can find index sets Ij{1,,t}, j = 1, 2, such that Fj={uaiTu=0, iIj, aiTu0,i{1,,t}Ij},j=1,2 and, by a possible enlargement of I2, there exist some u¯F2 with aiTu¯=0, iI2, aiTu¯<0, i{1,,t}I2. Then I1I2. Indeed, assuming on the contrary that there is some i¯I1I2, we have ai¯Tu¯=0 because of u¯F2F1 and i¯I1. On the other hand we have ai¯Tu¯<0 because of i¯I2 and this is clearly impossible. Hence I1I2. Further we claim that F1F2=R:=uaiTu=0, iI1, aiTu0, iI2I1. The inclusion F1F2R immediately follows. To prove the opposite inclusion consider uR. Then we can choose λ0 large enough such that u1:=u+λu¯ fulfils aiTu1<0, i{1,,t}I2. Together with aiTu1=aiTu, iI2 and uR it follows that u1F1. Since u2:=λu¯F2, we have u=u1u2F1F2 showing RF1F2. Thus our claim holds true and we obtain (F1F2)={iI2μiaiμi0, iI2I1}. On the other hand, we have NK(v)={iI(v)μiaiμi0,iI(v)}, where I(v):={i{1,,t}aiTv=0} and thus spNK(v)={iI(v)μiaiμiR,iI(v)}. Because of vF2, we have I2I(v) and the asserted inclusion (F1F2)spNK(v) follows.

On the basis of Corollary 5.1 and Lemma 5.1, we can now state an efficient variant of Theorem 5.2 in which the manipulation with faces of KD(q~(p¯,x¯),λ¯) is reduced only to the verification of (Equation30).

Theorem 5.3

Assume that λ¯ is the (unique) solution of system (Equation15) and for every hRm there is some kRn and some ηRs fulfilling (Equation26).

Further assume that for every nonzero pair (h,k)Rm×Rn and the corresponding (unique) ηRs satisfying (Equation26) one has

  1. (33) 2q~(p¯,x¯)Tμ=0, μsp(NKD(q~(p¯,x¯),λ¯)(q~(p¯,x¯)(h,k)))  μ=0;(33)

  2. for every pair of faces F1,F2 of the critical cone KD(q~(p¯,x¯),λ¯) with q~(p¯,x¯)(h,k)F2F1[η] and for every w0 with b(p¯,x¯)wF1F2 there is some w~ with 2q~(p¯,x¯)w~F1F2 and (34) wT2Lλ¯(p¯,x¯)w~>0.(34)

Then all assertions of Theorem 5.1 remain valid.

Proof.

It follows immediately from Lemma 5.1 with K=KD(q~(p¯,x¯),λ) and v=q~(p¯,x¯)(h,k).

The next example illustrates the application of the preceding result.

Example 1

Consider the solution map S of the variational system defined via (Equation3) with f(p,x)=x1px2+x22,q(p,x,y)=px1+2y14y2x1+2y1+4y2,D=R2 at the reference point p¯=0, x¯=(0,0). Then b(p,x)=2424,q~(p,x)=p+x14x2x1+4x2 and (A) is fulfilled since b(p¯,x¯) has full rank. Further, λ¯=(0,0) is the unique solution of (Equation15), KD(q~(p¯,x¯),λ¯)=D=R2 and the system (Equation26) reads as (35) 00=h+k1k2+2244η,ηNR2h+k14k2k1+4k2.(35) Straightforward calculations yield that for every hR the set T(h):={(k,η)R2×R2(h,k,η) fulfils }(Equation35) is not empty and (36) T(h)={((h,0),(0,0)),((87h,27h),(0,114h)),((97h,47h),(17h,0))}if h<0,{((12h,18h),(2364h,2564h))}if h0.(36) Thus for every hR there is some pair (k,η)R2×R2 fulfilling (Equation26) and we shall now show that the other assumptions of Theorem 5.3 are fulfilled as well. Note that (Equation33) always holds because the matrix 2q~(p¯,x¯)=1414 has full rank. According to (Equation36) we have to consider the following four cases.

Case (i): h>0, k=(12h,18h), η=(2364h,2564h). Evidently, F2=F1={0} is the only face of KD(q~(p¯,x¯),λ¯)=R2 contained in [η]. Thus w = 0 is the solely element satisfying b(p¯,x¯)wF1F2={0} and we are done.

Case (ii): h<0, k=(h,0), η=(0,0). In this case, we have the requirement q~(p¯,x¯)(h,k)=2hhF2F1R2 resulting in F2=F1=R2 and F1F2=R2. Consider any wR2{0}. Then, by taking w~=(w1,w2) we have 2q~(p¯,x¯)w~F1F2 and wT2Lλ¯(p¯,x¯)w~=w12+w22>0 showing the validity of (Equation34).

Case (iii): h<0, k=(87h,27h), η=(0,114h). In this case, we conclude from the condition q~(p¯,x¯)(h,k)=237h0F2F1[η] that F2=F1=R×{0} and thus F1F2=R×{0}. Any w0 with b(p¯,x¯)wF1F2 satisfies w2=w1/20 and by choosing w~=(w1,w1/4) we have 2q~(p¯,x¯)w~F1F2 and wT2Lλ¯(p¯,x¯)w~=w12w2w~2=78w12>0 verifying again (Equation34).

Case (iv): h<0, k=(97h,47h), η=(17h,0). In this case, the faces F1,F2 satisfying q~(p¯,x¯)(h,k)=0257hF2F1[η] are F1=F2={0}×R and thus F1F2={0}×R. Any w0 with b(p¯,x¯)wF1F2 satisfies w2=w1/20 and w~=(w1,w1/4) fulfils 2q~(p¯,x¯)w~F1F2 and wT2Lλ¯(p¯,x¯)w~=w12w2w~2=78w12>0. Hence, (Equation34) holds in this case as well.

Thus all assumptions of Theorem 5.3 are fulfilled and the solution map S has the Aubin property around (p¯,x¯).

Note that this result cannot be obtained by condition (5.2) from [Citation4] which attains the form

p=v1+w10=v1+w1+w20=v2+4w1+4w2wDNR2(0,0)2v14v22v1+4v2p=v1=v2=0.

Indeed, the relations on the left-hand side have, e.g. the nontrivial solution v1=1, v2=0.5, p=2516.

Conclusion

This paper contains a thorough analysis of a parameterized variational system with implicit constraints. One can say that Boris Mordukhovich stands behind most important ingredients used in this development. Indeed, as pointed out in the Introduction, the model came from [Citation4] and the results in Section 4 are in fact directional variants of their counterparts in [Citation4, Section 3]. Furthermore, the development of the directional limiting calculus has been initiated in [Citation17] and also Theorem 2.1 [Citation2, Theorem 4.4] relies essentially on the so-called Mordukhovich criterion [Citation9, Chapter 9F]. Thus via this research the authors would like to give credit to their friend Boris on the occasion of his 70th birthday.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

The research of the first author was supported by the Austrian Science Fund (FWF) under grant P29190-N32. The research of the second author was supported by the Grant Agency of the Czech Republic, Project 17-04301S and the Australian Research Council, Project 10.13039/501100000923DP160100854.

References

  • Aubin JP. Lipschitz behavior of solutions to convex minimization problems. Math Oper Res. 1984;9:87–111.
  • Gfrerer H, Outrata JV. On Lipschitzian properties of implicit multifunctions. SIAM J Optimization. 2016;26:2160–2189.
  • Gfrerer H, Outrata JV. On the Aubin property of a class of parameterized variational systems. Math Meth Oper Res. 2017;86:443–467.
  • Mordukhovich BS, Outrata JV. Coderivative analysis of quasi-variational inequalities with applications to stability and optimization. SIAM J Optimization. 2007;18:389–412.
  • Mordukhovich BS. Generalized differential calculus for nonsmooth and set-valued mappings. J Math Anal Appl. 1994;183:250–288.
  • Mordukhovich BS. Variational analysis and generalized differentiation I. Heidelberg: Springer; 2006.
  • Benko M, Gfrerer H, Outrata JV. Calculus for directional limiting normal cones and subdifferentials. Set-Valued Var. Anal, doi:10.1007/s11228-018-04+2-5.
  • Dontchev AL, Rockafellar RT. Implicit functions and solution mappings. Heidelberg: Springer; 2014.
  • Rockafellar RT, Wets RJ-B. Variational analysis. Berlin: Springer; 1998.
  • Gfrerer H. On directional metric regularity, subregularity and optimality conditions for nonsmooth mathematical programs. Set-Valued Var Anal. 2013;21:151–176.
  • Bonnans JF, Shapiro A. Perturbation analysis of optimization problems. New York: Springer; 2000.
  • Dontchev AL, Rockafellar RT. Characterizations of strong regularity for variational inequalities over polyhedral convex sets. SIAM J Optim. 1996;7:1087–1105.
  • Gfrerer H, Mordukhovich BS. Robinson stability of parametric constraint systems via variational analysis. SIAM J Optim. 2017;27:438–465.
  • Gfrerer H, Mordukhovich BS. Second-order variational analysis of parametric constraint and variational systems. SIAM J Optim. 2019;29:423–453.
  • Gfrerer H, Klatte D. Lipschitz and Hölder stability of optimization problems and generalized equations. MathProg Ser A. 2016;158:35–75. doi:10.1007/s10107-015-0914-1.
  • Henrion R, Outrata JV, Surowiec T. On the co-derivative of normal cone mappings to inequality systems. Non Anal TMA. 2009;7:1213–1226.
  • Ginchev I, Mordukhovich BS. On directionally dependent subdifferentials. CR Bulg Acad Sci. 2011;64:497–508.