Skip to main content
Log in

Quantitative stability analysis of stochastic quasi-variational inequality problems and applications

  • Full Length Paper
  • Series B
  • Published:
Mathematical Programming Submit manuscript

Abstract

We consider a parametric stochastic quasi-variational inequality problem (SQVIP for short) where the underlying normal cone is defined over the solution set of a parametric stochastic cone system. We investigate the impact of variation of the probability measure and the parameter on the solution of the SQVIP. By reformulating the SQVIP as a natural equation and treating the orthogonal projection over the solution set of the parametric stochastic cone system as an optimization problem, we effectively convert stability of the SQVIP into that of a one stage stochastic program with stochastic cone constraints. Under some moderate conditions, we derive Hölder outer semicontinuity and continuity of the solution set against the variation of the probability measure and the parameter. The stability results are applied to a mathematical program with stochastic semidefinite constraints and a mathematical program with SQVIP constraints.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Arutyunov, A.V., Izmailov, A.F.: Sensitivity analysis for cone-constrained optimization problems under the relaxed constraint qualifications. Math. Oper. Res. 30, 333–353 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  2. Aumann, R.J.: Integrals of set-valued functions. J. Math. Anal. Appl. 12, 1–12 (1965)

    Article  MathSciNet  MATH  Google Scholar 

  3. Aubin, J.P., Frankowska, H.: Set-Valued Analysis. Birkhäuser, Basel (1990)

    MATH  Google Scholar 

  4. Birbil, S., Gürkan, G., Listes, O.: Solving stochastic mathematical programs with complementarity constraints using simulation. Math. Oper. Res. 31, 739–760 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bonnans, J.F., Shapiro, A.: Perturbation Analysis of Optimization Problems. Springer, New York (2000)

    Book  MATH  Google Scholar 

  6. Chan, D., Pang, J.S.: The generalized quasi-variational inequality problem. Math. Oper. Res. 7, 211–222 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  7. Chen, X., Fukushima, M.: Expected residual minimization method for stochastic linear complementarity problems. Math. Oper. Res. 30, 1022–1038 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  8. Chen, X., Wets, R.B.-J., Zhang, Y.: Stochastic variational inequalities: residual minimization smoothing/sample average approximations. SIAM J. Optim. 22, 649–673 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  9. Ding, C., Sun, D.F., Zhang, L.W.: Characterization of the robust isolated calmness for a class of conic programming problems. http://arxiv.org/abs/1601.07418v1 (2016)

  10. Fiala, J., Kocvara, M., Stingl, M.: PENLAB: A MATLAB solver for nonlinear semidefinite optimization. http://arxiv.org/abs/1311.5240 (2013)

  11. Facchinei, F., Pang, J.S.: Finite-Dimensional Variational Inequalities and Complementarity Problems I–II. Springer, New York (2003)

    MATH  Google Scholar 

  12. Izmailov, A.F., Kurennoy, A.S., Solodov, M.V.: A note on upper Liptschitz stability, error bounds and critical multipliers for Liptschitz-continuous KKT systems. Math. Program. 142, 591–604 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  13. Jiang, H., Xu, H.: Stochastic approximation approaches to the stochastic variational inequality problem. IEEE Trans. Autom. Control 53, 1462–1475 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  14. King, A.J., Rockafellar, R.T.: Sensitivity analysis for nonsmooth generalized equations. Math. Program. 55, 193–212 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  15. King, A.J., Rockafellar, R.T.: Asymptotic theory for solutions in statistical estimation and stochastic programming. Math. Oper. Res. 18, 148–162 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  16. Klatte, D.: On the stability of local and global optimal solutions in parametric problems of nonlinear programming, part I: basic results. In: Humboldt-Universit\(\ddot{a}\)t Sektion Mathematik, vol. 75, pp. 1–21 (1985)

  17. Klatte, D., Kummer, B.: Nonsmooth Equations in Optimization: Regularity, Calculus, Methods and Applications. Kluwer, Dordrecht (2002)

    MATH  Google Scholar 

  18. Liu, Y., Römisch, W., Xu, H.: Quantitative stability analysis of stochastic generalized equations. SIAM J. Optim. 24, 467–497 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  19. Liu, Y., Xu, H., Lin, G.H.: Stability analysis of two stage stochastic mathematical programs with complementarity constraints via NLP-regularization. SIAM J. Optim. 21, 609–705 (2011)

    MathSciNet  MATH  Google Scholar 

  20. Lu, S., Budhiraja, A.: Confidence regions for stochastic variational inequalities. Math. Oper. Res. 38, 545–568 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  21. Mordukhovich, B.S., Outrata, J.V.: Coderivative analysis of quasi-variational inequalities with applications to stability and optimization. SIAM J. Optim. 18, 389–412 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  22. Pang, J.S., Fukushima, M.: Quasi-variational inequalities, generalized Nash equilibria, and multi-leader-follower games. CMS 2, 21–56 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  23. Pflug, G.Ch.: Stochastic optimization and statistical inference. In: Ruszczynski, A., Shapiro, A. (eds.) Stochastic Programming (Handbooks in Operations Research and Management Science), vol. 10, pp. 427–480. Elsevier, Amsterdam (2003)

  24. Pflug, G.Ch., Pichler, A.: Approximations for probability distributions and stochastic optimization problems. In: Bertocchi, M., Consigli, G., Dempster, M.A. (eds.) Stochastic Optimization Methods in Finance and Energy, vol. 163, pp. 343–387. Springer, New York (2011)

  25. Ravat, U., Shanbhag, U.V.: On the characterization of solution sets of smooth and nonsmooth convex stochastic Nash games. SIAM J. Optim. 21, 1168–1199 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  26. Rockafellar, R.T., Wets, R.J.B.: Variational Analysis. Springer, Berlin (1998)

    Book  MATH  Google Scholar 

  27. Robinson, S.M.: An application of error bounds for convex programming in linear space. SIAM J. Control Optim. 13, 271–273 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  28. Robinson, S.M.: Stability theory for systems of inequalities, part II: differentiable nonlinear systems. SIAM J. Numer. Anal. 13, 497–513 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  29. Robinson, S.M.: Generalized equations and their solutions, part I: basic theory. Math. Progr. Study 10, 128–141 (1979)

    Article  MATH  Google Scholar 

  30. Robinson, S.M.: Strongly regular generalized equations. Math. Oper. Res. 5, 43–62 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  31. Robinson, S.M.: Generalized equations and their solutions, part II: applications to nonlinear programming. Math. Progr. Study 19, 200–221 (1982)

    Article  MATH  Google Scholar 

  32. Rockafellar, R.T., Wets, R.J.B.: Stochastic variational inequalities: single-stage to multistage. Math. Program. (2016). doi:10.1007/s10107-016-0995-5

  33. Römisch, W.: Stability of stochastic programming problems. In: Ruszczynski, A., Shapiro, A. (eds.) Stochastic Programming. Handbooks in Operations Research and Management Science, vol. 10, pp. 483–554. Elsevier, Amsterdam (2003)

    Google Scholar 

  34. Rusczynski, A., Shapiro, A.: Stochastic Programming, Handbooks in OR and MS. Elsevier, Amsterdam (2003)

    Google Scholar 

  35. Shapiro, A.: Sensitivity analysis of parameterized variational inequalities. Math. Oper. Res. 30, 109–126 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  36. Shapiro, A.: Sensitivity analysis of generalized equations. J. Math. Sci. 115, 2554–2565 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  37. Shapiro, A., Dentcheva, D., Ruszczynski, A.: Lectures on Stochastic Programming: Modeling and Theory. SIAM, Philadelphia (2009)

    Book  MATH  Google Scholar 

  38. Wolfowitz, J.: Generalization of the theorem of Glivenko–Cantelli. Ann. Math. Stat. 25, 131–138 (1954)

    Article  MathSciNet  MATH  Google Scholar 

  39. Xu, H.: Sample average approximation methods for a class of stochastic variational inequality problems. Asian Pac. J. Oper. Res. 27, 103–119 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  40. Xu, H., Ye, J.J.: Approximating stationary points of stochastic mathematical programs with variational inequality constraints via sample averaging. Set-Valued Var. Anal. 19, 283–309 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  41. Yousefian, F., Nedić, A., Shanbhag, U. V.: A regularized smoothing stochastic approximation (RSSA) algorithm for stochastic variational inequality problems. In: Proceeding WSC ’13 Proceedings of the 2013 Winter Simulation Conference: Simulation: Making Decisions in a Complex World, pp. 933–944 (2013)

  42. Yousefian, F., Nedich, A., Shanbhag, U.: Self-tuned stochastic approximation schemes for non-Lipschitzian stochastic multi-user optimization and Nash games. IEEE Trans. Autom. Control 61, 1753–1766 (2016)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank the editor-in-chief and two anonymous referees for valuable comments which help us significantly consolidate the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Huifu Xu.

Additional information

J. Zhang: The research of this author is supported by the NSFC under Project Nos. 11671183 and 11201210, CPSF under Project No. 2014M560200 and Program for Liaoning Excellent Talents in University under Project No. LJQ2015059.

L.-W. Zhang: The research of this author was supported by the NSFC under Project Nos. 91330206 and 11571059.

Appendix

Appendix

Proof of Lemma 2.2

Part (i). Observe that (2.4) implies that \(\Psi (\bar{u}) \ne \emptyset \), and under uniform continuity w.r.t. u further that \(\Psi (u) \ne \emptyset \) when u is sufficiently close to \(\bar{u}\).

It is well known (see [3, Definition 1.4.6]) that

$$\begin{aligned} \liminf _{u\rightarrow \bar{u}}\Psi (u)\subseteq {\mathrm{cl}} \; \Psi (\bar{u})\subseteq \limsup _{u\rightarrow \bar{u}}\Psi (u). \end{aligned}$$

Since \(\Psi (\bar{u})\) is closed, to show continuity of \(\Psi (\cdot )\) at \(\bar{u}\), it suffices to show that the set-valued mapping is both upper and lower semicontinuous at \(\bar{u}\), that is,

$$\begin{aligned} \limsup _{u\rightarrow \bar{u}}\Psi (u)\subseteq \Psi (\bar{u})\subseteq \liminf _{u\rightarrow \bar{u}}\Psi (u). \end{aligned}$$
(6.1)

The upper semicontinuity can be easily verified under the continuity of \(\psi \) and closeness of \(\mathcal{K}\). Hence it suffices to show the lower semicontinuity.

Observe that condition (2.4) and continuity of \(\psi (\cdot ,\bar{u})\) imply that \({\mathrm{int}} \; \Psi (\bar{u})\ne \emptyset \). Moreover, since \(\psi (z,u)\) is uniformly continuous w.r.t. u under condition (b), it is easy to show that for any \(z \in {\mathrm{int}} \; \Psi (\bar{u})\), we can set u sufficiently close to \(\bar{u}\) such that \(z\in \Psi (u)\). This allows us to claim that

$$\begin{aligned} {\mathrm{int}} \; \Psi (\bar{u}) \subseteq \liminf _{u\rightarrow \bar{u}} \Psi (u). \end{aligned}$$
(6.2)

Moreover, since \(\Psi (\bar{u})\) is convex and closed and by Rockafellar and Wets [26, Proposition 4.4], \(\liminf _{u\rightarrow \bar{u}} \Psi (u)\) is closed, we have

$$\begin{aligned} \Psi (\bar{u}) = {\mathrm{cl}}\; \left( {\mathrm{int}} \; \Psi (\bar{u})\right) \subseteq {\mathrm{cl}}\;(\liminf _{u\rightarrow \bar{u}} \Psi (u)) = \liminf _{u\rightarrow \bar{u}} \Psi (u), \end{aligned}$$

which yields the second inclusion of (6.1).

Part (ii). Since \(\Psi (\bar{u})\) is compact and \(\Psi (\cdot )\) is closed at \(\bar{u}\), the continuity of \(\Psi (\cdot )\) at \(\bar{u}\) means that for any number \(\delta >0\), there exists \(\epsilon >0\) such that when \(\Vert u-\bar{u}\Vert _U\le \epsilon \),

$$\begin{aligned} \mathbb {H} (\Psi (u),\Psi (\bar{u})) \le \delta \end{aligned}$$

or equivalently

$$\begin{aligned} \Psi (u)\subseteq \Psi (\bar{u}) + \delta \mathcal{B}_m \; {\mathrm{and}} \; \Psi (\bar{u}) \subseteq \Psi (u) + \delta \mathcal{B}_m. \end{aligned}$$
(6.3)

In what follows, we use Lemma 2.1 (Robinson–Ursescu’s theorem) to derive an error bound for set-valued mapping \(\Psi (\cdot )\) at \(\bar{u}\) under condition (d) and Slater constraint qualification (2.4).

Let

$$\begin{aligned} \mathcal {F}_{u}(z):=\left\{ \begin{array}{cl} \psi (z,u)-\mathcal{K}, &{}\quad \text{ for }\,\, z\in \mathcal{Z}, \\ \emptyset , &{} \quad \text{ for }\,\, z\notin \mathcal{Z}. \end{array} \right. \end{aligned}$$

Condition (a) ensures that \(\mathcal {F}_{u}(z)\) is closed and convex set-valued over its domain \(\mathcal{Z}\). Moreover,

$$\begin{aligned} \Psi (u) = \mathcal {F}_{u}^{-1}(0), \end{aligned}$$
(6.4)

and \(z\in \Psi (u)\) if and only if \((z,0)\in \text{ gph }\; \mathcal {F}_{u}\). Furthermore, it follows from (2.4)

$$\begin{aligned} 0\in \text{ int }(\text{ range } \; \mathcal {F}_{\bar{u}}). \end{aligned}$$
(6.5)

Let \(z \in \Psi (\bar{u})\) be fixed, that is, \((z,0)\in \text{ gph }\; \mathcal {F}_{\bar{u}}\). By Lemma 2.1, there exist positive numbers \(\delta _{z}\), \(\eta _z\) and \(c_{z}\) (depending on z) such that

$$\begin{aligned} d(z', \mathcal{F}_{\bar{u}}^{-1}(q)) \le c_{z} d(q,\mathcal{F}_{\bar{u}}(z')). \end{aligned}$$
(6.6)

for each \(z'\in B_m(z,\delta _{z})\cap \mathcal{Z}\) and \(q\in B_W(0,\eta _z)\). In particular,

$$\begin{aligned} d(z', \Psi (\bar{u})) = d(z',\mathcal {F}_{\bar{u}}^{-1}(0)) \le c_{z} d(\psi (z',\bar{u}), \mathcal{K}). \end{aligned}$$
(6.7)

Under condition (d), we have

$$\begin{aligned} d(\psi (z',\bar{u}), \mathcal{K}) \le \Vert \psi (z', \bar{u})-\psi (z', u)\Vert _Z +d(\psi (z',u), \mathcal{K}) \le \sigma \Vert u-\bar{u}\Vert _U^\nu . \end{aligned}$$
(6.8)

for \(z'\in \Psi (u)\).

On the other hand, since z is arbitrarily chosen from \(\Psi (\bar{u})\), the set \(\Psi (\bar{u})\) may be covered by the union of a collection of \(\delta \)-balls, i.e.,

$$\begin{aligned} \Psi (\bar{u})\subseteq \bigcup _{z\in \Psi (\bar{u})}\text{ int }\,\, B_m(z,\delta _{z}). \end{aligned}$$

Since \(\Psi (\bar{u})\) is a compact set, by the finite covering theorem, there exist a finite number of points \(z_1, z_2,\dots , z_k \in \Psi (\bar{u})\) and positive constants \(\delta _{z_i}, c_{z_i}, i=1,\dots ,k\) such that

$$\begin{aligned} \Psi (\bar{u})\subseteq \bigcup _{i=1}^k\text{ int }\,\,B_m(z_i,\delta _{z_i}) \end{aligned}$$
(6.9)

and

$$\begin{aligned} d(z', \Psi (\bar{u}))\le c_{z_i} d(\psi (z',\bar{u}), \mathcal{K}) \end{aligned}$$
(6.10)

for any \(z'\in B_m(z_i,\delta _{z_i})\cap \mathcal{Z}\). Since \(\Psi (\bar{u})\) is a compact set and \(\bigcup _{i=1}^k\text{ int }\,\,B_m(z_i,\delta _{z_i})\) is an open set, there exists a positive constant \(\delta \) such that

$$\begin{aligned} \Psi (\bar{u})+\delta \mathcal{B}_m\subseteq \bigcup _{i=1}^k\text{ int }\,\,B_m(z_i,\delta _{z_i}). \end{aligned}$$
(6.11)

Let \(C:=\max \{c_{z_1},\dots ,c_{z_k}\}\). Combining (6.11) and (6.10), we obtain

$$\begin{aligned} d(z', \Psi (\bar{u}))\le Cd(\psi (z',\bar{u}), \mathcal{K}) \end{aligned}$$
(6.12)

for all \(z'\in \left( \Psi (\bar{u})+\delta \mathcal{B}_m\right) \cap \mathcal{Z}\). Through (6.3) and (6.8), we arrive at

$$\begin{aligned} \mathbb {D}(\Psi (u), \Psi (\bar{u}))\le Cd(\psi (z,\bar{u}), \mathcal{K}) \le C\sigma \Vert u-\bar{u}\Vert _U^\nu \end{aligned}$$

for all \(z'\in \Psi (u)\).

To complete the proof of (2.5), it suffices to show

$$\begin{aligned} \mathbb {D}(\Psi (\bar{u}),\Psi (u)) \le C\sigma \Vert u-\bar{u}\Vert _U^\nu . \end{aligned}$$
(6.13)

For any \(z\in \mathcal{Z}\), let \(q=\psi (z,\bar{u})-\psi (z,u)\). Under condition (d), \( \Vert q\Vert _W\le \sigma \Vert u-\bar{u}\Vert _U^\nu . \) Moreover, it is easy to see that \( \Psi (u) =\mathcal{F}^{-1}_{\bar{u}}(q). \) Let \(z\in \Psi (\bar{u})\), \(\delta _z\) be defined as in (6.6) and \(\epsilon \) be defined at the beginning of the proof of this part with \( \sigma \epsilon ^\nu \le \delta _z. \) Under condition (d), we have \(\Vert q\Vert \le \delta _z\). It follows from (6.6) (by setting \(z'=z\)) that

$$\begin{aligned} d(z, \Psi (u))= d(z,\mathcal {F}_{\bar{u}}^{-1}(q)) \le c_{z} d(q,\mathcal{F}_{\bar{u}}(z)). \end{aligned}$$
(6.14)

Since \(z\in \Psi (\bar{u})\), then \(0\in \mathcal{F}_{\bar{u}}(z)\) and hence

$$\begin{aligned} d(q,\mathcal{F}_{\bar{u}}(z)) \le \Vert q\Vert _W\le \sigma \Vert u-\bar{u}\Vert _U^\nu . \end{aligned}$$

Combining the two inequalities above, we obtain

$$\begin{aligned} d(z,\Psi (u)) \le c_z\sigma \Vert u-\bar{u}\Vert _U^\nu \end{aligned}$$

for any \(z\in \Psi (\bar{u})\) and \(\Vert u-\bar{u}\Vert _U\le \epsilon \). Utilizing the second inclusion in (6.3), we can apply the finite covering theorem to set \(\Psi (\bar{u})\) (similar to previous discussion) and find a positive constant C such that (6.13) holds. The proof is complete. \(\square \)

In the case when \(\psi \) is continuously differentiable, Lemma 2.2 coincides with [28, Corollary 1]. Thus we may regard our result as an extension of the Robinson’s result.

Proof of Lemma 2.3

Let \(\hat{\phi }(z,u):=\phi (z,u)+\delta _{\Psi (u)}(z)\), where \(\delta _C(\cdot )\) denotes the indicator function of a set \(C\subseteq \mathbb {R}^m\) with \(\delta _C(x)=0\) for \(x\in C\) and \(\delta _C(x)=+\infty \) for \(x\notin C\). Problem (2.3) can be equivalently written as

$$\begin{aligned} \min _{z\in \mathbb {R}^m}\hat{\phi }(z,u). \end{aligned}$$

The objective function \(\hat{\phi }\) has the following properties:

(i):

Since \(\phi \) and \(\psi \) are continuous, \(\mathcal{K}\) and \(\mathcal{Z}\) are closed and convex, we have graph\(\Psi \) is closed, which means that \(\mathrm{epi}\,\delta _{\Psi (\cdot )}(\cdot )=\{(u,z, \alpha ): z\in \Psi (u), \alpha \ge 0\}=\mathrm{graph} \Psi \times \mathbb {R}_+\) is closed. Hence \(\hat{\phi }\) is proper lower semicontinuous;

(ii):

Since \(\mathcal{Z}\) is compact, \(\hat{\phi }\) is level-bounded in z locally uniformly in u;

(iii):

Under the Slater condition (2.4), one has from Lemma 2.2 that

$$\begin{aligned} \lim _{u\rightarrow \bar{u}}\Psi (u)=\Psi (\bar{u}), \end{aligned}$$

which means that

$$\begin{aligned}\limsup _{u\rightarrow \bar{u}}\mathrm {epi}\, [\delta _{\Psi (u)}(\cdot )]=\mathrm {epi}\, [\delta _{\Psi (\bar{u})}(\cdot )],\end{aligned}$$

where

$$\begin{aligned}\mathrm {epi}\, [\delta _{\Psi (u)}(\cdot )]=\{(z,\alpha ): z\in \Psi (u), \alpha \ge 0\}=\Psi (u)\times \mathbb {R}_+.\end{aligned}$$

Then we obtain from Definition 2.1 that

$$\begin{aligned}e\!-\!\!\!\lim _{u\rightarrow \bar{u}}\delta _{\Psi (u)}(\cdot )=\delta _{\Psi (\bar{u})}(\cdot ),\end{aligned}$$

which, by Proposition 2.1, means that for any sequence \(u^{\nu }\rightarrow \bar{u}\) and point \(\bar{z}\in \mathbb {R}^m\),

$$\begin{aligned}\liminf _{\nu \rightarrow \infty }\delta _{\Psi (u^{\nu })}(z^{\nu })\ge \delta _{\Psi (\bar{u})}(\bar{z})\end{aligned}$$

for every sequence \(z^{\nu }\rightarrow \bar{z}\) and there exists some \(z^{\nu }\rightarrow \bar{z}\) such that

$$\begin{aligned}\limsup _{\nu \rightarrow \infty }\delta _{\Psi (u^{\nu })}(z^{\nu })\le \delta _{\Psi (\bar{u})}(\bar{z}).\end{aligned}$$

We know from the continuity of \(\phi \) that

$$\begin{aligned}\lim _{\nu \rightarrow \infty } \phi (z^{\nu }, u^{\nu })=\phi (\bar{z},\bar{u})\,\,\text{ as }\,\, \nu \rightarrow \infty ,\end{aligned}$$

which leads to

$$\begin{aligned}\liminf _{\nu \rightarrow \infty }\,[\phi (z^{\nu }, u^{\nu })+\delta _{\Psi (u^{\nu })}(z^{\nu })]\ge \phi (\bar{z}, \bar{u})+\delta _{\Psi (\bar{u})}(\bar{z})\end{aligned}$$

and there exists \(z^{\nu }\rightarrow \bar{z}\) such that

$$\begin{aligned} \limsup _{\nu \rightarrow \infty }\,[\phi (z^{\nu }, u^{\nu })+\delta _{\Psi (u^{\nu })}(z^{\nu })]\le \phi (\bar{z}, \bar{u})+\delta _{\Psi (\bar{u})}(\bar{z}).\end{aligned}$$

Then also by Proposition 2.1, we have that \(u\mapsto \hat{\phi }(\cdot ,u)\) is epi-continuous at \(\bar{u}\).

Hence all conditions in [26, Theorem 7.41] are satisfied, and from this theorem we obtain that \({\vartheta }(\cdot )\) is continuous at \(\bar{u}\) and \(\mathcal{Z}(\cdot )\) is outer semicontinuous at \(\bar{u}\). Noting that \(\mathcal{Z}(u)\subset \mathcal{Z}\) and it actually is locally bounded around \(\bar{u}\), we have from [26, Theorem 5.19] that \(\mathcal{Z}(\cdot )\) is upper semicontinuous at \(\bar{u}\). \(\square \)

Proof of Theorem 2.1

Part (i). First we show that

$$\begin{aligned} \mathbb {D}\left( \mathcal{Z}(u), \mathcal{Z}(\bar{u})\right) \le \max _{z \in \mathcal{Z}(u)}\Vert z-\Pi _{\Psi (\bar{u})}(z)\Vert + R(u), \end{aligned}$$
(6.15)

where \(\Pi _{\Psi (\bar{u})}(z)\) denotes orthogonal projection of z on \(\Psi (\bar{u})\),

$$\begin{aligned} R(u) := \left( \frac{1}{\alpha }\left[ 2\varrho \Vert u-\bar{u}\Vert _U^\gamma + L\left( \max _{z \in \mathcal{Z}(u)}\Vert z-\Pi _{\Psi (\bar{u})}(z)\Vert +\min _{z \in \mathcal{Z}(\bar{u})}\Vert z-\Pi _{\Psi (u)}(z)\Vert \right) \right] \right) ^{\frac{1}{2}}.\nonumber \\ \end{aligned}$$
(6.16)

Let \(z(u) \in \mathcal{Z}(u)\) and \(z(\bar{u}) \in \mathcal{Z}(\bar{u})\). Under conditions (b) and (c)

$$\begin{aligned} {\vartheta }(u)-\phi (\Pi _{\Psi (\bar{u})}(z(u)),\bar{u})= & {} \phi (z(u),u)-\phi (\Pi _{\Psi (\bar{u})}(z(u)),\bar{u}) \nonumber \\\ge & {} -L\Vert z(u) - \Pi _{\Psi (\bar{u})}(z(u))\Vert -\varrho \Vert u-\bar{u}\Vert _U^{\gamma } \quad \end{aligned}$$
(6.17)

for \(u\in B(\bar{u},\bar{\delta }_u)\). On the other hand, since \(\Pi _{\Psi (u)}(z(\bar{u}))\in \Psi (u)\) and z(u) is an optimal solution to \(\mathcal{P}_{u}\), we have

$$\begin{aligned} {\vartheta }(u) \le \phi \left( \Pi _{\Psi (u)}(z(\bar{u})),u\right) . \end{aligned}$$

Using this inequality, (2.8) and the growth condition (c), we have

$$\begin{aligned} {\vartheta }(u) - \phi (\Pi _{\Psi (\bar{u})}(z(u)),\bar{u})= & {} \phi (z(u), u)-\phi (z(\bar{u}),u)+(\phi (z(\bar{u}),u)-\phi (z(\bar{u}),\bar{u}))\nonumber \\&-\,(\phi (\Pi _{\Psi (\bar{u})}(z(u)),\bar{u})-\phi (z(\bar{u}),\bar{u}))\nonumber \\\le & {} \phi (\Pi _{\Psi (u)}(z(\bar{u})),u) -\phi (z(\bar{u}),u) + \varrho \Vert u-\bar{u}\Vert _U^{\gamma }\nonumber \\&-\,\alpha d\left( \Pi _{\Psi (\bar{u})}(z(u)), \mathcal{Z}(\bar{u})\right) ^2\nonumber \\\le & {} L\Vert \Pi _{\Psi (u)}(z(\bar{u}))-z(\bar{u})\Vert +\varrho \Vert u-\bar{u}\Vert _U^{\gamma }\nonumber \\&-\,\alpha d\left( \Pi _{\Psi (\bar{u})}(z(u)), \mathcal{Z}(\bar{u})\right) ^2 \end{aligned}$$
(6.18)

for \(u\in B(\bar{u},\bar{\delta }_u)\). Combining (6.17) and (6.18), we obtain

$$\begin{aligned}&d\left( \Pi _{\Psi (\bar{u})}(z(u)), \mathcal{Z}(\bar{u})\right) \nonumber \\&\quad \le \left( [L\Vert z(u)-\Pi _{\Psi (\bar{u})}(z(u))\Vert +2\varrho \Vert u-\bar{u}\Vert _U^{\gamma } +L\Vert z(\bar{u})-\Pi _{\Psi (u)}(z(\bar{u}))\Vert ]/\alpha \right) ^{\frac{1}{2}}, \nonumber \\ \end{aligned}$$
(6.19)

which yields (6.15) through the triangle inequality below

$$\begin{aligned} d\left( z(u), \mathcal{Z}(\bar{u})\right) \le \Vert z(u)-\Pi _{\Psi (\bar{u})}(z(u))\Vert +d\left( \Pi _{\Psi (\bar{u})}(z(u)),\mathcal{Z}(\bar{u})\right) \end{aligned}$$
(6.20)

because z(u) and \(z(\bar{u})\) are arbitrarily taken from the set of optimal solutions. Since \(\mathcal {Z}(u)\subseteq \Psi (u)\), it follows by Lemma 2.2 that there exists a constant \(\epsilon >0\) such that

$$\begin{aligned} \max \left\{ \max _{z\in \mathcal {Z}(u)}\Vert z-\Pi _{\Psi (\bar{u})}(z)\Vert , \min _{z\in \mathcal {Z}(\bar{u}) }\Vert z-\Pi _{\Psi (u)}(z)\Vert \right\} =O(\Vert u-\bar{u}\Vert _U^\nu ) \end{aligned}$$
(6.21)

when \(\Vert u-\bar{u}\Vert _U\le \epsilon \). Combining (6.15) and (6.21), we obtain (2.9).

Part (ii). We only need to show that

$$\begin{aligned} \mathbb {D}\left( \mathcal{Z}(\bar{u}), \mathcal{Z}(u)\right) \le c'\Vert u-\bar{u}\Vert _U^{\beta }. \end{aligned}$$

From the proof of Part (i), we can see that we can swap u with \(\bar{u}\) except that the first inequality in formulae (6.18) requires second order growth condition of \(\phi (\cdot ,u)\) over the optimal solution set \(\mathcal{Z}(u)\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, J., Xu, H. & Zhang, L. Quantitative stability analysis of stochastic quasi-variational inequality problems and applications. Math. Program. 165, 433–470 (2017). https://doi.org/10.1007/s10107-017-1116-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-017-1116-9

Keywords

Mathematics Subject Classification

Navigation