Skip to main content
Log in

First order optimality conditions for mathematical programs with semidefinite cone complementarity constraints

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

Abstract

In this paper we consider a mathematical program with semidefinite cone complementarity constraints (SDCMPCC). Such a problem is a matrix analogue of the mathematical program with (vector) complementarity constraints (MPCC) and includes MPCC as a special case. We first derive explicit formulas for the proximal and limiting normal cone of the graph of the normal cone to the positive semidefinite cone. Using these formulas and classical nonsmooth first order necessary optimality conditions we derive explicit expressions for the strong-, Mordukhovich- and Clarke- (S-, M- and C-)stationary conditions. Moreover we give constraint qualifications under which a local solution of SDCMPCC is a S-, M- and C-stationary point. Moreover we show that applying these results to MPCC produces new and weaker necessary optimality conditions.

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. Aubin, J.-P.: Lipschitz behavior of solutions to convex minimization problems. Math. Oper. Res. 9, 87–111 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  2. Ben-Tal, A., Nemirovski, A.: Robust convex optimization. Math. Oper. Res. 23, 769–805 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  3. Ben-Tal, A., Nemirovski, A.: Robust convex optimization-methodology and applications. Math. Program. 92, 453–480 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bhatia, R.: Matrix Analysis. Springer, New York (1997)

    Book  Google Scholar 

  5. Bi, S., Han, L., Pan, S.: Approximation of rank function and its application to the nearest low-rank correlation matrix. J. Glob. Optim. 57, 1113–1137 (2013)

    Article  MathSciNet  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  7. Brigo, D., Mercurio, F.: Calibrating LIBOR. Risk Mag. 15, 117–122 (2002)

    Google Scholar 

  8. Burge, J.P., Luenberger, D.G., Wenger, D.L.: Estimation of structured covariance matrices. Proc. IEEE 70, 963–974 (1982)

    Article  Google Scholar 

  9. Clarke, F.H.: Optimization and Nonsmooth Analysis. Wiley-Interscience, New York (1983)

    MATH  Google Scholar 

  10. Clarke, F.H., Ledyaev, Yu.S, Stern, R.J., Wolenski, P.R.: Nonsmooth Analysis and Control Theory. Springer, New York (1998)

  11. de Gaston, R.R.E., Safonov, M.G.: Exact calculation of the multiloop stability margin. IEEE Trans. Autom. Control 33, 156–171 (1988)

    Article  MATH  Google Scholar 

  12. Dempe, S.: Foundations of Bilevel Programming. Kluwer, Berlin (2002)

    MATH  Google Scholar 

  13. Eaves, B.C.: On the basic theorem for complementarity. Math. Program. 1, 68–75 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  14. Faccchinei, F., Pang, J.S.: Finite-Dimensional Variational Inequalities and Complementarity Problem. Springer, New York (2003)

    Google Scholar 

  15. Fazel, M.: Matrix Rank Minimization with Applications. PhD thesis, Stanford University (2002)

  16. Fletcher, R.: Semi-definite matrix constraints in optimization. SIAM J. Control Optim. 23, 493–513 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  17. Flegel, M.L., Kanzow, C.: On the Guignard constraint qualification for mathematical programs with equilibrium constraints. Optimization 54, 517–534 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  18. Goh, K.C., Ly, J.C., Safonov, M.G., Papavassilopoulos, G., Turan, L.: Biaffine matrix inequality properties and computational methods. In: Proceeding of the American Control Conference, Baltimore, Maryland, pp. 850–855 (1994)

  19. Henrion, R., Outrata, J.: On the calmness of a class of multifunctions. SIAM J. Optim. 13, 603–618 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  20. Henrion, R., Outrata, J.: Calmness of constraint systems with applications. Math. Program. Ser. B 104, 437–464 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  21. Hobbs, B.F., Metzler, C.B., Pang, J.S.: Strategic gaming analysis for electric power systems: an MPEC approach. IEEE Trans. Power Syst. 15, 638–645 (2000)

    Article  Google Scholar 

  22. Lewis, A.S.: Nonsmooth analysis of eigenvalues. Math. Program. 84, 1–24 (1999)

    MathSciNet  MATH  Google Scholar 

  23. Li, Q.N., Qi, H.D.: A sequential semismooth Newton method for the nearest low-rank correlation matrix problem. SIAM J. Optim. 21, 1641–1666 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  24. Lillo, F., Mantegna, R.N.: Spectral density of the correlation matrix of factor models: a random matrix theory approach. Phys. Rev. E 72, 016219-1–016219-10 (2005)

    Article  MathSciNet  Google Scholar 

  25. Löwner, K.: Über monotone matrixfunktionen. Mathematische Zeitschrift 38, 177–216 (1934)

    Article  MathSciNet  Google Scholar 

  26. Luo, Z.Q., Pang, J.S., Ralph, D.: Mathematical Programs with Equilibrium Constraints. Cambridge University Press, Cambridge (1996)

    Book  Google Scholar 

  27. Hiriart-Urruty, J.-B., Ye, D.: Sensitivity analysis of all eigenvalues of a symmetric matrix. Numerische Mathematik 70, 45–72 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  28. Hoge, W.: A subspace identification extension to the phase correlation method. IEEE Trans. Med. Imaging 22, 277–280 (2003)

    Article  Google Scholar 

  29. Meng, F., Sun, D.F., Zhao, G.Y.: Semismoothness of solutions to generalized equations and Moreau-Yosida regularization. Mathe. Program. 104, 561–581 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  30. Mordukhovich, B.S.: Generalized differential calculus for nonsmooth and set-valued mappings. J. Math. Anal. Appl. 183, 250–288 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  31. Mordukhovich, B.S.: Variational Analysis and Generalized Differentiation, I: Basic Theory, Grundlehren Series (Fundamental Principles of Mathematical Sciences), vol. 330. Springer, Berlin (2006)

    Google Scholar 

  32. Mordukhovich, B.S.: Variational Analysis and Generalized Differentiation, II: Applications, Grundlehren Series (Fundamental Principles of Mathematical Sciences), vol. 331. Springer, Berlin (2006)

    Google Scholar 

  33. Mordukhovich, B.S., Shao, Y.: Nonsmooth sequential analysis in Asplund space. Trans. Am. Math. Soc. 348, 215–220 (1996)

    Article  MathSciNet  Google Scholar 

  34. Outrata, J.V., Koc̆vara, M., Zowe, J.: Nonsmooth Approach to Optimization Problem with Equilibrium Constraints: Theory, Application and Numerical Results. Kluwer, Dordrecht (1998)

    Book  Google Scholar 

  35. Overton, M., Womersley, R.S.: On the sum of the largest eigenvalues of a symmetric matrix. SIAM J. Matrix Anal. Appl. 13, 41–45 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  36. Overton, M., Womersley, R.S.: Optimality conditions and duality theory for minimizing sums of the largest eigenvalues of symmetric matrices. Math. Program. 62, 321–357 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  37. Psarris, P., Floudas, C.A.: Robust stability analysis of linear and nonlinear systems with real parameter uncertainty. AIChE Annual Meeting, p. 127e. Florida, Miami Beach (1992)

  38. Qi, H.D., Fusek, P.: Metric regularity and strong regularity in linear and nonlinear semidefinite programming. Technical Report, School of Mathematics, University of Southampton (2007)

  39. Robinson, S.M.: Stability theory for systems of inequalities, part I: linear systems. SIAM J. Numer. Anal. 12, 754–769 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  40. Robinson, S.M.: Stability theory for systems of inequalities, part II: nonlinear systems. SIAM J. Numer. Anal. 13, 473–513 (1976)

    Article  MathSciNet  Google Scholar 

  41. Robinson, S.M.: First order conditions for general nonlinear optimization. SIAM J. Appl. Math. 30, 597–607 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  42. Robinson, S.M.: Some continuity properties of polyhedral multifunctions. Math. Program. Stud. 14, 206–214 (1981)

    Article  MATH  Google Scholar 

  43. Rockafellar, R.T.: Convex Analysis. Princeton University Press, Princeton (1970)

    MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  45. Ryoo, H.S., Sahinidis, N.V.: Global optimization of nonconvex NLPs and MINLPs with applications in process design. Comput. Chem. Eng. 19, 551–566 (1995)

    Article  Google Scholar 

  46. Safonov, M.G., Goh, K.C., Ly, J.H.: Control system synthesis via bilinear matrix inequalities. In: Proceeding of the American Control Conference, pp. 45–49. Baltimore, Maryland (1994)

  47. Scheel, H., Scholtes, S.: Mathematical programs with complementarity constraints: stationarity, optimality and sensitivity. Math. Oper. Res. 25, 1–22 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  48. Simon, D.: Reduced order Kalman filtering without model reduction. Control Intell. Syst. 35, 169–174 (2007)

    MathSciNet  Google Scholar 

  49. Sun, D.F.: The strong second-order sufficient condition and constraint nondegeneracy in nonlinear semidefinite programming and their implications. Math. Oper. Res. 31, 761–776 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  50. Sun, D.F., Sun, J.: Semismooth matrix valued functions. Math. Oper. Res. 27, 150–169 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  51. Sun, D.F., Sun, J.: Strong semismoothness of eigenvalues of symmetric matrices and its applications in inverse eigenvalue problems. SIAM J. Numer. Anal. 40, 2352–2367 (2003)

    Article  MATH  Google Scholar 

  52. Tsing, N.K., Fan, M.K.H., Verriest, E.I.: On analyticity of functions involving eigenvalues. Linear Algebra Appl. 207, 159–180 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  53. VanAntwerp, J.G., Braatz, R.D., Sahinidis, N.V.: Globally optimal robust control for systems with nonlinear time-varying perturbations. Comput. Chem. Eng. 21, S125–S130 (1997)

    Article  Google Scholar 

  54. Visweswaran, V., Floudas, C.A.: A global optimization algorithm (GOP) for certain classes of nonconvex NLPs—I. Theory. Comput. Chem. Eng. 14, 1397–1417 (1990)

    Google Scholar 

  55. Visweswaran, V., Floudas, C.A.: A global optimization algorithm (GOP) for certain classes of nonconvex NLPs—II. Application of theory and test problems. Comput. Chem. Eng. 14, 1419–1434 (1990)

    Google Scholar 

  56. Wu, L.X.: Fast at-the-money calibration of the LIBOR market model using Lagrange multipliers. J Comput. Financ. 6, 39–77 (2003)

    Google Scholar 

  57. Wu, Z., Ye, J.J.: First and second order condition for error bounds. SIAM J. Optim. 14, 621–645 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  58. Yan, T., Fukushima, M.: Smoothing method for mathematical programs with symmetric cone complementarity constraints. Optimization 60, 113–128 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  59. Ye, J.J.: Optimality conditions for optimization problems with complementarity constraints. SIAM J. Optim. 9, 374–387 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  60. Ye, J.J.: Constraint qualifications and necessary optimality conditions for optimization problems with variational inequality constraints. SIAM J. Optim. 10, 943–962 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  61. Ye, J.J.: Necessary and sufficient optimality conditions for mathematical programs with equilibrium constraints. J. Math. Anal. Appl. 307, 305–369 (2005)

    Article  MathSciNet  Google Scholar 

  62. Ye, J.J., Ye, X.Y.: Necessary optimality conditions for optimization problems with variational inequality constraints. Math. Oper. Res. 22, 977–977 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  63. Ye, J.J., Zhu, D.L., Zhu, Q.J.: Exact penalization and necessary optimality conditions for generalized bilevel programming problems. SIAM J. Optim. 7, 481–507 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  64. Zhang, Z.Y., Wu, L.X.: Optimal low-rank approximation to a correlation matrix. Linear Algebra Appl. 364, 161–187 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  65. Zhao, Y.B.: An approximation theory of matrix rank minimization and its application to quadratic equations. Linear Algebra Appl. 437, 77–93 (2012)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The authors are grateful to the anonymous referees for their constructive suggestions and comments which helped to improve the presentation of the materials in this paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jane J. Ye.

Additional information

D. Sun’s research is supported in part by Academic Research Fund under grant R-146-000-149-112.

The research of J. J. Ye was partially supported by NSERC.

Part of this work was done while C. Ding was with Department of Mathematics, National University of Singapore. The research of this author is supported by the National Science Foundation for Distinguished Young Scholars of China (Grant No. 11301515).

Appendix

Appendix

Proof of Proposition 2.6

Firstly, we will show that (16) holds for the case that \(A=\Lambda (A)\). For any \(H\in \mathcal{S}^{n}\), denote \(Y:=A+H\). Let \(P\in \mathcal{O}^{n}\) (depending on \(H\)) be such that

$$\begin{aligned} \Lambda (A)+H=P\Lambda (Y)P^{T}. \end{aligned}$$
(57)

Let \(\delta >0\) be any fixed number such that \(0<\delta <\frac{\lambda _{|\alpha |}}{2}\) if \(\alpha \ne \emptyset \) and be any fixed positive number otherwise. Then, define the following continuous scalar function

$$\begin{aligned} f (t):=\left\{ \begin{array}{l@{\quad }l@{\quad }l} t &{} \mathrm{if} &{} t>\delta ,\\ 2t-\delta &{} \mathrm{if} &{} \frac{\delta }{2} < t< \delta , \\ 0 &{} \mathrm{if} &{}t< \frac{\delta }{2}. \end{array} \right. \end{aligned}$$

Therefore, we have

$$\begin{aligned} \{\lambda _{1}(A), \ldots , \lambda _{|\alpha |}(A)\}\in (\delta ,+\infty )\quad \mathrm{and} \quad \{\lambda _{|\alpha |+1}(A),\ldots , \lambda _{n}(A) \}\in \left( -\infty , \frac{\delta }{2}\right) . \end{aligned}$$

For the scalar function \(f\), let \(F:\mathcal{S}^{n}\rightarrow \mathcal{S}^{n}\) be the corresponding Löwner’s operator [25], i.e., for any \(Z\in \mathcal{S}^{n}\),

$$\begin{aligned} F(Z):= \sum _{i=1}^n f(\lambda _{i}(Z)) u_i u_i^T, \end{aligned}$$
(58)

where \(U\in \mathcal{O}^{n}\) satisfies that \(Z=U\Lambda (Z)U^{T}\). Since \(f \) is real analytic on the open set \((-\infty , \frac{\delta }{2})\cup (\delta ,+\infty )\), we know from [52, Theorem 3.1] that \(F\) is analytic at \(A\). Therefore, since \(A=\Lambda (A)\), it is well-known (see e.g., [4, Theorem V.3.3]) that for \(H\) sufficiently close to zero,

$$\begin{aligned} F(A+H)-F(A)-F'(A)H=O(\Vert H\Vert ^{2}) \end{aligned}$$
(59)

and

$$\begin{aligned} F'(A)H=\left[ \begin{array}{l@{\quad }l@{\quad }l} H_ {\alpha \alpha } &{} H_{\alpha \beta } &{} \Sigma _{\alpha \gamma } \circ H_{\alpha \gamma }\\ H_{\alpha \beta }^T &{} 0&{} 0 \\ {\Sigma }_{\alpha \gamma }^T \circ H_{\alpha \gamma }^T &{} 0 &{} 0 \end{array} \right] , \end{aligned}$$

where \(\Sigma \in \mathcal{S}^{n}\) is given by (15) . Let \(R(\cdot ):=\Pi _{\mathcal{S}_{+}^{n}}(\cdot )-F(\cdot )\). By the definition of \(f \), we know that \(F(A) =A_{+}:=\Pi _{\mathcal{S}_{+}^{n}}(A)\), which implies that \(R(A)=0\). Meanwhile, it is clear that the matrix valued function \(R\) is directionally differentiable at \(A\), and from (14), the directional derivative of \(R\) for any given direction \(H\in \mathcal{S}^{n}\), is given by

$$\begin{aligned} R' (A;H)=\Pi '_{\mathcal{S}_{+}^{n}}(A;H)-F '(A)H=\left[ \begin{array}{l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0\\ 0&{} \Pi _{\mathcal{S}_{+}^{|\beta |}}(H_{\beta \beta }) &{} 0 \\ 0 &{} 0 &{} 0 \end{array} \right] . \end{aligned}$$
(60)

By the Lipschitz continuity of \(\lambda (\cdot )\), we know that for \(H\) sufficiently close to zero,

$$\begin{aligned} \{\lambda _{1}(Y), \ldots , \lambda _{|\alpha |}(Y) \}\in (\delta ,+\infty ), \quad \{\lambda _{|\alpha |+1}(Y),\ldots , \lambda _{|\beta |}(Y) \}\in \left( -\infty , \frac{\delta }{2}\right) \end{aligned}$$

and

$$\begin{aligned} \{\lambda _{|\beta |+1}(Y),\ldots , \lambda _{n}(Y) \}\in (-\infty , 0). \end{aligned}$$

Therefore, by the definition of \(F \), we know that for \(H\) sufficiently close to zero,

$$\begin{aligned} R(A+H)=\Pi _{\mathcal{S}_{+}^{n}}(A+H)-F (A+H) = P\left[ \begin{array}{l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0 \\ 0 &{} (\Lambda (Y)_{\beta \beta })_{+} &{} 0 \\ 0 &{} 0 &{} 0 \end{array} \right] P^{T}. \end{aligned}$$
(61)

Since \(P\) satisfies (57), we know that for any \(\mathcal{S}^{n}\ni H\rightarrow 0\), there exists an orthogonal matrix \(Q\in \mathcal{O}^{|\beta |}\) such that

$$\begin{aligned} P_{\beta }=\left[ \begin{array}{l} O(\Vert H\Vert ) \\ P_{\beta \beta } \\ O(\Vert H\Vert ) \end{array} \right] \quad \mathrm{and}\quad P_{\beta \beta }=Q+O(\Vert H\Vert ^{2}), \end{aligned}$$
(62)

which was stated in [51] and was essentially proved in the derivation of Lemma 4.12 in [50]. Therefore, by noting that \((\Lambda (Y)_{\beta \beta })_{+}=O(\Vert H\Vert )\), we obtain from (60), (61) and (62) that

$$\begin{aligned}&R(A+H)-R(A)-R'(A;H)\\&\quad = \left[ \begin{array}{l@{\quad }l@{\quad }l} O(\Vert H\Vert ^{3}) &{} O(\Vert H\Vert ^{2}) &{} O(\Vert H\Vert ^{3})\\ O(\Vert H\Vert ^{2}) &{} P_{\beta \beta }( \Lambda (Y)_{\beta \beta })_{+}P_{\beta \beta }^{T}-\Pi _{\mathcal{S}_{+}^{|\beta |}}(H_{\beta \beta }) &{} O(\Vert H\Vert ^{2}) \\ O(\Vert H\Vert ^{3}) &{} O(\Vert H\Vert ^{2}) &{} O(\Vert H\Vert ^{3}) \end{array} \right] \\&\quad = \left[ \begin{array}{l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0\\ 0 &{} Q( \Lambda (Y)_{\beta \beta })_{+}Q^{T}-\Pi _{\mathcal{S}_{+}^{|\beta |}}(H_{\beta \beta })&{}0\\ 0 &{} 0 &{} 0 \end{array} \right] +O(\Vert H\Vert ^{2}). \end{aligned}$$

By (57) and (62), we know that

$$\begin{aligned} \Lambda (Y)_{\beta \beta }\!=\!P_{\beta }^{T}\Lambda (A)P_{\beta }\!+\!P_{\beta }^{T}HP_{\beta }\!=\! P_{\beta \beta }^{T}H_{\beta \beta }P_{\beta \beta }\!+\!O(\Vert H\Vert ^{2})\!=\! Q^{T}H_{\beta \beta }Q\!+\!O(\Vert H\Vert ^{2})\!. \end{aligned}$$

Since \(Q\in \mathcal{O}^{|\beta |}\), we have

$$\begin{aligned} H_{\beta \beta }=Q\Lambda (Y)_{\beta \beta }Q^{T}+O(\Vert H\Vert ^{2}). \end{aligned}$$

By noting that \(\Pi _{\mathcal{S}_{+}^{|\beta |}}(\cdot )\) is globally Lipschitz continuous and \(\Pi _{\mathcal{S}_{+}^{|\beta |}}(Q \Lambda (Y)_{\beta \beta }Q^{T})=Q( \Lambda (Y)_{\beta \beta })_{+}Q^{T}\), we obtain that

$$\begin{aligned}&Q(\Lambda (Y)_{\beta \beta })_{+}Q^{T}-\Pi _{\mathcal{S}_{+}^{|\beta |}}(H_{\beta \beta })\\&\quad =Q(\Lambda (Y)_{\beta \beta })_{+}Q^{T}-\Pi _{\mathcal{S}_{+}^{|\beta |}}(Q \Lambda (Y)_{\beta \beta }Q^{T})+O(\Vert H\Vert ^{2})\\&\quad =O(\Vert H\Vert ^{2}). \end{aligned}$$

Therefore,

$$\begin{aligned} R(A+H)-R(A)-R'(A;H)=O(\Vert H\Vert ^{2}). \end{aligned}$$
(63)

By combining (59) and (63), we know that for any \(\mathcal{S}^{n}\ni H\rightarrow 0\),

$$\begin{aligned} \Pi _{\mathcal{S}_{+}^{n}}(\Lambda (A)+H)-\Pi _{\mathcal{S}_{+}^{n}}(\Lambda (A))-\Pi '_{\mathcal{S}_{+}^{n}}(\Lambda (A);H)=O(\Vert H\Vert ^{2}). \end{aligned}$$
(64)

Next, consider the case that \(A=\overline{P}^{T}\Lambda (A)\overline{P}\). Re-write (57) as

$$\begin{aligned} \Lambda (A)+\overline{P}^{T}H\overline{P}=\overline{P}^{T}P \Lambda (Y)P^{T}\overline{P}. \end{aligned}$$

Let \(\widetilde{H}:=\overline{P}^{T}H\overline{P}\). Then, we have

$$\begin{aligned} \Pi _{\mathcal{S}^{n}_{+}}(A+H)=\overline{P}\,\Pi _{\mathcal{S}^{n}_{+}}(\Lambda (A)+\widetilde{H})\overline{P}^{T}. \end{aligned}$$

Therefore, since \(\overline{P}\in \mathcal{O}^{n}\), we know from (64) and (14) that for any \(\mathcal{S}^{n}\ni H\rightarrow 0\), (16) holds. \(\square \)

Proof of Proposition 3.3

Denote the set in the righthand side of (27) by \(\mathcal N\). We first show that \(N_{\mathrm{gph}\,N_{\mathcal{S}^{n}_+}} (0,0)\subseteq \mathcal{N}\). By the definition of the limiting normal cone in (8), we know that \((U^*,V^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^{|\beta |}_+}} (0,0)\) if and only if there exist two sequences \(\{ ({U^k}^*,{V^k}^*)\}\) converging to \((U^*,V^*)\) and \(\{(U^k,V^k)\}\) converging to \((0,0)\) with \(({U^k}^*,{V^k}^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^{n}_+}}^\pi (U^k,V^k)\) and \((U^k,V^k)\in \mathrm{gph}\,N_{\mathcal{S}^{n}_+}\) for each \(k\).

For each \(k\), denote \(A^{k}\!:=\!U^{k}\!+\!V^{k}\in \mathcal{S}^{n}\) and let \(A^{k}\!=\!P^{k}\Lambda (A^{k})(P^{k})^{T}\) with \(P^{k}\!\in \!\mathcal{O}^{n}\) be the eigenvalue decomposition of \(A^{k}\). Then for any \(i\!\in \!\{1,\ldots ,n\}\), we have

$$\begin{aligned} \lim _{k\rightarrow \infty } \lambda _{i}(A^{k})=0. \end{aligned}$$

Since \(\{P^k\}_{k=1}^{\infty }\) is uniformly bounded, by taking a subsequence if necessary, we may assume that \(\{P^k\}_{k=1}^{\infty }\) converges to an orthogonal matrix \(Q := \displaystyle {\lim \nolimits _{k\rightarrow \infty }} P^k\in \mathcal{O}^{n}\). For each \(k\), we know that the vector \(\lambda (A^{k})\) is an element of \({\mathfrak {R}}^{n}_{\gtrsim }\). By taking a subsequence if necessary, we may assume that for each \(k\), \(\Lambda (A^{k})\) has the same form, i.e.,

$$\begin{aligned} \Lambda (A^{k})=\left[ \begin{array}{l@{\quad }l@{\quad }l}\Lambda (A^{k})_{\beta _{+}\beta _{+}} &{} 0 &{} 0 \\ 0 &{} \Lambda (A^{k})_{\beta _{0}\beta _{0}} &{} 0 \\ 0 &{} 0 &{} \Lambda (A^{k})_{\beta _{-}\beta _{-}}\end{array}\right] , \end{aligned}$$

where \(\beta _{+}\), \(\beta _{0}\) and \(\beta _{-}\) are the three index sets defined by

$$\begin{aligned} \beta _+:=\{i:\, \lambda _{i}(A^{k})>0\}, \quad \beta _0:=\{i:\, \lambda _{i}(A^{k})=0\}\quad \mathrm{and} \quad \beta _-:= \{i:\, \lambda _{i}(A^{k})<0\}. \end{aligned}$$

Since \(({U^k}^*,{V^k}^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^{n}_+}}^\pi (U^k,V^k) \), we know from Proposition 3.2 that for each \(k\), there exist

$$\begin{aligned} \Theta ^{k}_{1} =\left[ \begin{array}{l@{\quad }l@{\quad }l} E_ {\beta _{+} \beta _{+}} &{} E_{\beta _{+} \beta _{0}} &{} \Sigma _{\beta _{+} \beta _{-}}^{k}\\ E_{\beta _{+} \beta _{0}}^T &{} 0 &{} 0 \\ ({\Sigma }_{\beta _{+} \beta _{-}}^{k})^T &{} 0 &{} 0 \end{array} \right] \end{aligned}$$

and

$$\begin{aligned} \Theta ^{k}_{2} =\left[ \begin{array}{l@{\quad }l@{\quad }l} 0 &{} 0 &{} E_{\beta _{+}\beta _{-}}-\Sigma _{\beta _{+} \beta _{-}}^{k}\\ 0 &{} 0 &{} E_{\beta _{0}\beta _{-}} \\ (E_{\beta _{+}\beta _{-}}-\Sigma _{\beta _{+} \beta _{-}}^{k}) ^T &{} (E_{\beta _{0}\beta _{-}})^{T} &{} E_{\beta _{-}\beta _{-}} \end{array} \right] \end{aligned}$$

such that

$$\begin{aligned} \Theta _{1}^{k}\circ {\widetilde{U^{*}}}^k+\Theta _{2}^{k}\circ \widetilde{{V^{k}}}^*=0, \quad \widetilde{{U^{k}}}^*_{\beta _{0}\beta _{0}}\preceq 0 \quad \mathrm{and}\quad \widetilde{{V^{k}}}^*_{\beta _{0}\beta _{0}}\succeq 0, \end{aligned}$$
(65)

where \(\widetilde{{U^{k}}}^*=(P^{k})^{T}{U^{k}}^*P^{k}\), \(\widetilde{{V^{k}}}^*=(P^{k})^{T}{V^{k}}^*P^{k}\) and

$$\begin{aligned} (\Sigma ^{k})_{i,j}= \frac{\max \{ \lambda _i(A^{k}) ,0\}-\max \{ \lambda _j(A^{k}) ,0\}}{\lambda _i(A^{k})-\lambda _j(A^{k})} \quad \forall \,(i,j) \in \beta _{+}\times \beta _{-}. \end{aligned}$$
(66)

Since for each \(k\), each element of \({\Sigma }^{k}_{\beta _{+}\beta _{-}}\) belongs to the interval \([0,1]\), by further taking a subsequence if necessary, we may assume that the limit of \(\{ {\Sigma }^{k}_{\beta _{+}\beta _{-}} \}_{k=1}^{\infty }\) exists. Therefore, by the definition of \(\mathcal{U}_{n}\) in (24), we know that

$$\begin{aligned} \lim _{k\rightarrow \infty }\Theta _{1}^{k}= \Xi _{1}\in \mathcal{U}_{n} \quad \mathrm{and} \quad \lim _{k\rightarrow \infty }\Theta _{2}^{k}=\Xi _{2}, \end{aligned}$$

where \(\Xi _{1}\) and \(\Xi _{2}\) are given by (26). Therefore, we obtain from (65) that \((U^{*},V^{*})\in \mathcal N.\)

The other direction, i.e., \(N_{\mathrm{gph}\,N_{\mathcal{S}^{n}_+}} (0,0)\supseteq \mathcal{N}\) can be proved in a similar but simpler way to that of the second part of Theorem 3.1. We omit it here. \(\square \)

Proof of Theorem 3.1

\(\Longrightarrow \)” Suppose that \((X^*,Y^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^n_+}} (X,Y)\). By the definition of the limiting normal cone in (8), we know that \((X^*,Y^*)=\displaystyle {\lim \nolimits _{k \rightarrow \infty }} ({X^k}^*,{Y^k}^*) \) with

$$\begin{aligned} ({X^k}^*,{Y^k}^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^n_+}}^\pi (X^k,Y^k)\quad k=1,2,\ldots , \end{aligned}$$

where \((X^k,Y^k) \rightarrow (X,Y)\) and \((X^k,Y^k)\in \mathrm{gph}\,N_{\mathcal{S}^n_+}\). For each \(k\), denote \(A^{k}:=X^{k}+Y^{k}\) and let \(A^{k}=P^{k}\Lambda (A^{k})(P^{k})^{T}\) be the eigenvalue decomposition of \(A^{k}\). Since \( \Lambda (A) = \displaystyle {\lim \nolimits _{k\rightarrow \infty }} \Lambda (A^{k}) \), we know that \(\Lambda (A^{k})_{\alpha \alpha }\succ 0\), \(\Lambda (A^{k})_{\gamma \gamma }\prec 0\) for k sufficiently large and \(\displaystyle {\lim \nolimits _{k \rightarrow \infty }} \displaystyle {\Lambda (A^{k})_{\beta \beta }} = 0\).

Since \(\{P^k\}_{k=1}^{\infty }\) is uniformly bounded, by taking a subsequence if necessary, we may assume that \(\{P^k\}_{k=1}^{\infty }\) converges to an orthogonal matrix \(\widehat{P}\in \mathcal{O}^{n}(A)\). We can write \(\widehat{P}=\left[ \overline{P}_{\alpha } \ \ \overline{P}_{\beta }Q \ \ \overline{P}_{\gamma }\right] \), where \(Q\in \mathcal {O}^{|\beta |}\) can be any \(|\beta |\times |\beta |\) orthogonal matrix. By further taking a subsequence if necessary, we may also assume that there exists a partition \(\pi (\beta )=(\beta _{+}, \beta _{0}, \beta _{-})\) of \(\beta \) such that for each \(k\),

$$\begin{aligned} \lambda _{i}(A^k) >0\quad \forall \, i\in \beta _{+}, \quad \lambda _{i}(A^k) =0\quad \forall \, i\in \beta _{0} \quad \text{ and } \quad \lambda _{i}(A^k) <0\quad \forall \, i\in \beta _{-}. \end{aligned}$$

This implies that for each \(k\),

$$\begin{aligned} \{i:\lambda _{i}(A^{k})>0\}\!=\!\alpha \cup \beta _{+}, \{i:\lambda _{i}(A^{k})=0\}\!=\!\beta _{0} \quad \text{ and } \quad \{i:\lambda _{i}(A^{k})<0\}\!=\!\beta _{-}\cup \gamma . \end{aligned}$$

Then, for each \(k\), since \(({X^k}^*,{Y^k}^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^n_+}}^\pi (X^k,Y^k)\), we know from Proposition 3.2 that there exist

$$\begin{aligned} \Theta ^{k}_{1}=\left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l@{\quad }l} E_ {\alpha \alpha } &{} E_{\alpha \beta _{+}} &{} E_{\alpha \beta _{0}} &{} \Sigma _{\alpha \beta _{-}}^{k} &{} \Sigma _{\alpha \gamma }^{k}\\ E_{\alpha \beta _{+}}^T &{} E_{\beta _{+}\beta _{+}} &{} E_{\beta _{+}\beta _{0}} &{} \Sigma _{\beta _{+}\beta _{-}}^{k} &{} \Sigma _{\beta _{+}\gamma }^{k} \\ E_{\alpha \beta _{0}}^T &{} E_{\beta _{+}\beta _{0}}^{T} &{} 0 &{} 0 &{} 0 \\ {\Sigma ^{k}_{\alpha \beta _{-}}}^{T} &{} {\Sigma ^{k}_{\beta _{+}\beta _{-}}}^{T} &{} 0 &{} 0 &{} 0 \\ {\Sigma ^{k}_{\alpha \gamma }}^{T} &{} {\Sigma ^{k}_{\beta _{+}\gamma }}^{T} &{} 0 &{} 0 &{} 0 \end{array} \right] \end{aligned}$$

and

$$\begin{aligned} \Theta ^{k}_{2}=\left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0 &{} E_{\alpha \beta _{-}}-\Sigma _{\alpha \beta _{-}}^{k} &{} E_{\alpha \gamma }-\Sigma _{\alpha \gamma }^{k}\\ 0 &{} 0 &{} 0 &{} E_{\beta _{+}\beta _{-}}-\Sigma _{\beta _{+}\beta _{-}}^{k} &{} E_{\beta _{+}\gamma }-\Sigma _{\beta _{+}\gamma }^{k} \\ 0 &{} 0 &{} 0 &{} E_{\beta _{0}\beta _{-}} &{} E_{\beta _{0}\gamma } \\ (E_{\alpha \beta _{-}}-\Sigma _{\alpha \beta _{-}}^{k})^{T} &{} (E_{\beta _{+}\beta _{-}}-\Sigma _{\beta _{+}\beta _{-}}^{k})^{T} &{} E_{\beta _{0}\beta _{-}}^{T} &{} E_{\beta _{-}\beta _{-}} &{} E_{\beta _{-}\gamma } \\ (E_{\alpha \gamma }-\Sigma _{\alpha \gamma }^{k})^{T} &{} (E_{\beta _{+}\gamma }-\Sigma _{\beta _{+}\gamma }^{k})^{T} &{} E_{\beta _{0}\gamma }^{T} &{} E_{\beta _{-}\gamma }^{T} &{} E_{\gamma \gamma } \end{array} \right] \end{aligned}$$

such that

$$\begin{aligned} \Theta _{1}^{k}\circ \widetilde{{X^{k}}}^*+\Theta _{2}^{k}\circ \widetilde{{Y^{k}}}^*=0, \quad \widetilde{{X^{k}}}^*_{\beta _{0}\beta _{0}}\preceq 0 \quad \mathrm{and} \quad \widetilde{{Y^{k}}}^*_{\beta _{0}\beta _{0}}\succeq 0, \end{aligned}$$
(67)

where \(\widetilde{{X^{k}}}^*=(P^{k})^{T}{X^{k}}^*P^{k}, \widetilde{{Y^{k}}}^*=(P^{k})^{T}{Y^{k}}^*P^{k}\) and

$$\begin{aligned} (\Sigma ^{k})_{i,j}= \frac{\max \{ \lambda _i(A^{k}) ,0\}-\max \{ \lambda _j(A^{k}) ,0\}}{\lambda _i(A^{k})-\lambda _j(A^{k})} \quad \forall \, (i,j) \in (\alpha \cup \beta _{+}) \times (\beta _{-}\cup \gamma ). \end{aligned}$$
(68)

By taking limits as \(k\rightarrow \infty \), we obtain that

$$\begin{aligned} \widetilde{{X^{k}}}^*\rightarrow \widehat{P}^{T}X^{*} \widehat{P}=\left[ \begin{array}{l@{\quad }l@{\quad }l} \widetilde{X}_{\alpha \alpha }^{*} &{} \widetilde{X}_{\alpha \beta }^{*}Q &{} \widetilde{X}_{\alpha \gamma }^{*}\\ (\widetilde{X}_{\alpha \beta }^{*}Q)^{T} &{} Q^{T}\widetilde{X}_{\beta \beta }^{*}Q &{} Q^{T}\widetilde{X}_{\beta \gamma }^{*}\\ (\widetilde{X}_{\alpha \gamma }^{*})^{T} &{} (Q^{T}\widetilde{X}_{\beta \gamma }^{*})^{T} &{} \widetilde{X}_{\gamma \gamma } \end{array} \right] \end{aligned}$$

and

$$\begin{aligned} \widetilde{{Y^{k}}}^*\rightarrow \widehat{P}^{T}Y^{*} \widehat{P}=\left[ \begin{array}{l@{\quad }l@{\quad }l} \widetilde{Y}_{\alpha \alpha }^{*} &{} \widetilde{Y}_{\alpha \beta }^{*}Q &{} \widetilde{Y}_{\alpha \gamma }^{*}\\ (\widetilde{Y}_{\alpha \beta }^{*}Q)^{T} &{} Q^{T}\widetilde{Y}_{\beta \beta }^{*}Q &{} Q^{T}\widetilde{Y}_{\beta \gamma }^{*}\\ (\widetilde{Y}_{\alpha \gamma }^{*})^{T} &{} (Q^{T}\widetilde{Y}_{\beta \gamma }^{*})^{T} &{} \widetilde{Y}_{\gamma \gamma } \end{array} \right] . \end{aligned}$$

By simple calculations, we obtain from (68) that

$$\begin{aligned} \lim _{k\rightarrow \infty }\Sigma _{\alpha \beta _{-}}^{k}=E_{\alpha \beta _{-}}, \quad \lim _{k\rightarrow \infty }\Sigma _{\beta _{+}\gamma }^{k}=0 \quad \text{ and } \quad \lim _{k\rightarrow \infty }\Sigma _{\alpha \gamma }^{k}=\Sigma _{\alpha \gamma }. \end{aligned}$$

This, together with the definition of \(\mathcal{U}_{|\beta |}\), shows that there exist \(\Xi _{1}\in \mathcal{U}_{|\beta |}\) and the corresponding \(\Xi _{2}\) such that

$$\begin{aligned} \lim _{k\rightarrow \infty }\Theta _{1}^{k}=\left[ \begin{array}{l@{\quad }l@{\quad }l} E_{\alpha \alpha } &{} E_{\alpha \beta } &{} \Sigma _{\alpha \gamma }\\ E_{\beta \alpha } &{} \Xi _{1} &{} 0\\ \Sigma _{\alpha \gamma }^{T} &{} 0 &{} 0 \end{array} \right] =\Theta _{1}+\left[ \begin{array}{l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0 \\ 0 &{} \Xi _{1} &{} 0\\ 0 &{} 0 &{} 0 \end{array} \right] \end{aligned}$$

and

$$\begin{aligned} \lim _{k\rightarrow \infty }\Theta _{2}^{k}=\left[ \begin{array}{l@{\quad }l@{\quad }l} 0 &{} 0 &{} E_{\alpha \gamma }-\Sigma _{\alpha \gamma }\\ 0 &{} \Xi _{2} &{} E_{\beta \gamma }\\ (E_{\alpha \gamma }-\Sigma _{\alpha \gamma })^{T} &{} E_{\gamma \beta } &{} E_{\gamma \gamma } \end{array} \right] =\Theta _{2}+\left[ \begin{array}{l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0 \\ 0 &{} \Xi _{2} &{} 0\\ 0 &{} 0 &{} 0 \end{array} \right] , \end{aligned}$$

where \(\Theta _{1}\) and \(\Theta _{2}\) are given by (22). Meanwhile, since \(Q\in \mathcal{O}^{|\beta |}\), by taking limits in (67) as \(k\rightarrow \infty \), we obtain that

$$\begin{aligned} \Theta _{1}\circ \widetilde{X}^{*}+\Theta _{2}\circ \widetilde{Y}^{*}=0, \quad \Xi _{1}\circ Q^{T}\widetilde{X}_{\beta \beta }^{*}Q+\Xi _{2}\circ Q^{T}\widetilde{Y}_{\beta \beta }^{*}Q=0 \end{aligned}$$
(69)

and

$$\begin{aligned} Q_{\beta _{0}}^{T}{\widetilde{X}}_{\beta \beta }^{*}{Q_{\beta _{0}}}\preceq 0 \quad \text{ and } \quad Q_{\beta _{0}}^{T}{\widetilde{Y}}_{\beta \beta }^{*}{Q_{\beta _{0}}} \succeq 0. \end{aligned}$$

Hence, by Proposition 3.3, we conclude that \((\widetilde{X}_{\beta \beta }^{*}, \widetilde{Y}_{\beta \beta }^{*})\in N_{\mathrm{gph}\,N_{\mathcal{S}^{|\beta |}_+}} (0,0)\). From (69), it is easy to check that \((X^{*},Y^{*})\) satisfies the conditions (28) and (29).

\(\Longleftarrow \)” Let \((X^{*},Y^{*})\) satisfies (28) and (29). We shall show that there exist two sequences \(\{(X^k,Y^k)\}\) converging to \((X,Y)\) and \(\{({X^k}^*,{Y^k}^*)\}\) converging to \((X^*,Y^*)\) with \( (X^k,Y^k)\in \mathrm{gph}\,N_{\mathcal{S}^n_+}\) and \(({X^k}^*,{Y^k}^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^n_+}}^\pi (X^k,Y^k)\) for each \(k\).

Since \((\widetilde{X}_{\beta \beta }^{*}, \widetilde{Y}_{\beta \beta }^{*})\in N_{\mathrm{gph}\,N_{\mathcal{S}^{|\beta |}_+}} (0,0)\), by Proposition 3.3, we know that there exist an orthogonal matrix \(Q\in \mathcal{O}^{|\beta |}\) and \(\Xi _1\in \mathcal{U}_{|\beta |}\) such that

$$\begin{aligned} \Xi _{1}\circ Q^{T}\widetilde{X}_{\beta \beta }^{*}Q+\Xi _{2}\circ Q^{T}\widetilde{Y}_{\beta \beta }^{*}Q=0, \quad Q_{\beta _{0}}^{T}\widetilde{X}_{\beta \beta }^{*}Q_{\beta _{0}}\preceq 0 \quad \text{ and } \quad Q_{\beta _{0}}^{T}\widetilde{Y}_{\beta \beta }^{*} Q_{\beta _{0}}\succeq 0. \end{aligned}$$
(70)

Since \(\Xi _1\in \mathcal{U}_{|\beta |}\), we know that there exists a sequence \(\{z^{k}\}\in {\mathfrak {R}}^{|\beta |}_{\gtrsim }\) converging to \(0\) such that \(\Xi _1=\displaystyle {\lim \nolimits _{k\rightarrow \infty }}D(z^k)\). Without loss of generality, we can assume that there exists a partition \(\pi (\beta )=(\beta _{+},\beta _{0},\beta _{-})\in \fancyscript{P}(\beta )\) such that for all \(k\),

$$\begin{aligned} z_{i}^{k}>0 \quad \forall \, i\in \beta _{+}, \quad z_{i}^{k}=0 \quad \forall \, i\in \beta _{0} \quad \mathrm{and} \quad \quad z_{i}^{k}<0 \quad \forall \, i\in \beta _{-}. \end{aligned}$$

For each \(k\), let

$$\begin{aligned} X^{k}=\widehat{P} \left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l@{\quad }l} \Lambda (A)_{\alpha \alpha } &{} 0 &{} 0 &{} 0 &{} 0 \\ 0 &{} (z^{k})_{+} &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0\\ \end{array} \right] \widehat{P}^{T} \quad \mathrm{and} \quad Y^{k}\!=\!\widehat{P} \left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} (z^{k})_{-} &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} \Lambda (A)_{\gamma \gamma } \end{array} \right] \widehat{P}^{T}\!, \end{aligned}$$

where \(\widehat{P}=\left[ \overline{P}_{\alpha } \, \, \overline{P}_{\beta }Q \, \, \overline{P}_{\gamma }\right] \in \mathcal{O}^{n}(A)\). Then, it is clear that \(\{(X^k,Y^k)\}\in \mathrm{gph}\,N_{\mathcal{S}^n_+}\) converging to \((X,Y)\). For each \(k\), denote

$$\begin{aligned} A^{k}=X^{k}+Y^{k}, \quad \Theta ^{k}_{1}=\left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l@{\quad }l} E_ {\alpha \alpha } &{} E_{\alpha \beta _{+}} &{} E_{\alpha \beta _{0}} &{} \Sigma _{\alpha \beta _{-}}^{k} &{} \Sigma _{\alpha \gamma }\\ E_{\alpha \beta _{+}}^T &{} E_{\beta _{+}\beta _{+}} &{} E_{\beta _{+}\beta _{0}} &{} \Sigma _{\beta _{+}\beta _{-}}^{k} &{} \Sigma _{\beta _{+}\gamma }^{k} \\ E_{\alpha \beta _{0}}^T &{} E_{\beta _{+}\beta _{0}}^{T} &{} 0 &{} 0 &{} 0 \\ (\Sigma ^{k}_{\alpha \beta _{-}})^{T} &{} (\Sigma ^{k}_{\beta _{+}\beta _{-}})^{T} &{} 0 &{} 0 &{} 0 \\ (\Sigma _{\alpha \gamma })^{T} &{} ({\Sigma ^{k}}_{\beta _{+}\gamma })^{T} &{} 0 &{} 0 &{} 0 \end{array} \right] \end{aligned}$$

and

$$\begin{aligned} \Theta ^{k}_{2}=\left[ \begin{array}{l@{\quad }l@{\quad }l@{\quad }l@{\quad }l} 0 &{} 0 &{} 0 &{} E_{\alpha \beta _{-}}-\Sigma _{\alpha \beta _{-}}^{k} &{} E_{\alpha \gamma }-\Sigma _{\alpha \gamma }\\ 0 &{} 0 &{} 0 &{} E_{\beta _{+}\beta _{-}}-\Sigma _{\beta _{+}\beta _{-}}^{k} &{} E_{\beta _{+}\gamma }-\Sigma _{\beta _{+}\gamma }^{k} \\ 0 &{} 0 &{} 0 &{} E_{\beta _{0}\beta _{-}} &{} E_{\beta _{0}\gamma } \\ (E_{\alpha \beta _{-}}-\Sigma _{\alpha \beta _{-}}^{k})^{T} &{} (E_{\beta _{+}\beta _{-}}-\Sigma _{\beta _{+}\beta _{-}}^{k})^{T} &{} E_{\beta _{0}\beta _{-}}^{T} &{} E_{\beta _{-}\beta _{-}} &{} E_{\beta _{-}\gamma } \\ (E_{\alpha \gamma }-\Sigma _{\alpha \gamma })^{T} &{} (E_{\beta _{+}\gamma }-\Sigma _{\beta _{+}\gamma }^{k})^{T} &{} E_{\beta _{0}\gamma }^{T} &{} E_{\beta _{-}\gamma }^{T} &{} E_{\gamma \gamma } \end{array} \right] , \end{aligned}$$

where

$$\begin{aligned} (\Sigma ^{k})_{i,j}= \frac{\max \{ \lambda _i(A^{k})) ,0\}-\max \{ \lambda _j(A^{k})) ,0\}}{\lambda _i(A^{k})-\lambda _j(A^{k})} \quad \forall \, (i,j) \in (\alpha \cup \beta _{+}) \times (\beta _{-}\cup \gamma ). \end{aligned}$$

Next, for each \(k\), we define two matrices \(\widehat{{X^{k}}}^*, \widehat{{Y^{k}}}^*\in \mathcal{S}^{n}\). Let \(i,j\in \{1,\ldots ,n\}\). If \((i,j)\) and \((j,i)\notin (\alpha \times \beta _{-}) \cup (\beta _{+}\times \gamma )\cup (\beta \times \beta )\). We define

$$\begin{aligned} {\widehat{{X^{k}}}^*}_{i,j}\equiv { \widetilde{X}^*}_{i,j}, \quad { \widehat{{Y^{k}}}^*}_{i,j}\equiv { \widetilde{Y}^*}_{i,j},\quad k=1,2,\ldots . \end{aligned}$$
(71)

Otherwise, denote \(c^{k}:=(\Sigma ^{k})_{i,j}\), \(k=1,2,\ldots \). We consider the following four cases.

Case 1 \((i,j)\) or \((j,i)\in \alpha \times \beta _{-}\). In this case, we know from (28) that \({\widetilde{X}^*}_{i,j}=0\). Since \(c_{k}\ne 0\) for all \(k\) and \(c^{k}\rightarrow 1\) as \(k\rightarrow \infty \), we define

$$\begin{aligned} {\widehat{{Y^{k}}}^*}_{i,j}\equiv {\widetilde{Y}^*}_{i,j}\quad \mathrm{and} \quad {\widehat{{X^{k}}}^*}_{i,j}=\frac{c^{k}-1}{c^{k}}{\widehat{{Y^{k}}}^*}_{i,j},\quad k=1,2,\ldots . \end{aligned}$$
(72)

Then, we have

$$\begin{aligned} c_{k}{\widehat{{X^{k}}}^*}_{i,j}+(1-c_{k}){\widehat{{Y^{k}}}^*}_{i,j}=0 \quad \forall \, k \quad \mathrm{and} \quad ({\widehat{{X^{k}}}^*}_{i,j},{\widehat{{Y^{k}}}^*}_{i,j})\rightarrow ({\widetilde{X}^*}_{i,j},{\widetilde{Y}^*}_{i,j})\quad \mathrm{as}\ k\rightarrow \infty . \end{aligned}$$

Case 2 \((i,j)\) or \((j,i)\in \beta _{+}\times \gamma \). In this case, we know from (28) that \({\widetilde{Y}^*}_{i,j}=0\). Since \(c_{k}\ne 1\) for all \(k\) and \(c^{k}\rightarrow 0\) as \(k\rightarrow \infty \), we define

$$\begin{aligned} {\widehat{{X^{k}}}^*}_{i,j}\equiv {\widetilde{X}^*}_{i,j}\quad \mathrm{and} \quad {\widehat{{Y^{k}}}^*}_{i,j}=\frac{c^{k}}{c^{k}-1}{\widehat{{X^{k}}}^*}_{i,j}, \quad k=1,2,\ldots . \end{aligned}$$
(73)

Then, we know that

$$\begin{aligned} c_{k}{\widehat{{X^{k}}}^*}_{i,j}+(1-c_{k}){\widehat{{Y^{k}}}^*}_{i,j}=0 \quad \forall \, k \quad \mathrm{and} \quad ({\widehat{{X^{k}}}^*}_{i,j},{\widehat{{Y^{k}}}^*}_{i,j})\rightarrow ({\widetilde{X}^*}_{i,j},{\widetilde{Y}^*}_{i,j})\quad \mathrm{as}\ k\rightarrow \infty . \end{aligned}$$

Case 3 \((i,j)\) or \((j,i)\in (\beta \times \beta ){\setminus } (\beta _{+}\times \beta _{-}) \). In this case, we define

$$\begin{aligned} {\widehat{{X^{k}}}^*}_{i,j} \equiv Q_{i}^{T}{\widetilde{X}^*}_{\beta \beta }Q_{j}, \quad {\widehat{{Y^{k}}}^*}_{i,j}\equiv Q_{i}^{T}{\widetilde{Y}^*}_{\beta \beta }Q_{j},\quad k=1,2,\ldots . \end{aligned}$$
(74)

Case 4 \((i,j)\) or \((j,i)\in \beta _{+}\times \beta _{-}\). Since \(c\in [0,1]\), we consider the following two sub-cases:

Case 4.1 \(c\ne 1\). Since \(c_{k}\ne 1\) for all \(k\) large enough, we define

$$\begin{aligned} {\widehat{{X^{k}}}^*}_{i,j}\equiv Q_{i}^{T}{\widetilde{X}^*}_{\beta \beta }Q_{j}\quad \mathrm{and} \quad {\widehat{{Y^{k}}}^*}_{i,j}=\frac{c^{k}}{c^{k}-1}{\widehat{{X^{k}}}^*}_{i,j}, \quad k=1,2,\ldots . \end{aligned}$$
(75)

Then, from (70), we know that

$$\begin{aligned} {\widehat{{Y^{k}}}^*}_{i,j}\rightarrow \frac{c}{c-1}Q_{i}^{T}{\widetilde{X}^*}_{\beta \beta }Q_{j}=Q_{i}^{T}{\widetilde{Y}^*}_{\beta \beta }Q_{j}\quad \mathrm{as}\ k\rightarrow \infty . \end{aligned}$$

Case 4.2 \(c=1\). Since \(c_{k}\ne 0\) for all \(k\) large enough, we define

$$\begin{aligned} {\widehat{{Y^{k}}}^*}_{i,j}\equiv Q_{i}^{T}{\widetilde{Y}^*}_{\beta \beta }Q_{j}\quad \mathrm{and} \quad {\widehat{{X^{k}}}^*}_{i,j}=\frac{c^{k}-1}{c^{k}}{\widehat{{Y^{k}}}^*}_{i,j}, \quad k=1,2,\ldots . \end{aligned}$$
(76)

Then, again from (70), we know that

$$\begin{aligned} {\widehat{{X^{k}}}^*}_{i,j}\rightarrow \frac{c-1}{c}Q_{i}^{T}{\widetilde{Y}^*}_{\beta \beta }Q_{j}=Q_{i}^{T}{\widetilde{X}^*}_{\beta \beta }Q_{j}\quad \mathrm{as}\ k\rightarrow \infty . \end{aligned}$$

For each \(k\), define \({X^k}^*=\widehat{P} \widehat{{X^{k}}}^*\widehat{P}^{T}\) and \({Y^k}^*=\widehat{P} \widehat{{Y^{k}}}^*\widehat{P}^{T}\). Then, from (71)–(76) we obtain that

$$\begin{aligned} \Theta _{1}^{k}\circ \widehat{P}^{T}{X^{k}}^{*}\widehat{P}+\Theta _{2}^{k}\circ \widehat{P}^{T}{Y^{k}}^{*}\widehat{P}=0,\quad k=1,2,\ldots . \end{aligned}$$

and

$$\begin{aligned} (\widehat{P}^{T}{X^{k}}^{*}\widehat{P}, \widehat{P}^{T}{Y^{k}}^{*}\widehat{P})\rightarrow (\widehat{P}^{T}{X}^{*}\widehat{P}, \widehat{P}^{T}{Y}^{*}\widehat{P})\quad \mathrm{as}\ k\rightarrow \infty . \end{aligned}$$
(77)

Moreover, from (74) and (70), we have

$$\begin{aligned}&Q_{\beta _{0}}^{T}{\widetilde{{X^{k}}}^*}_{\beta \beta }Q_{\beta _{0}}\equiv Q_{\beta _{0}}^{T}\widetilde{X}_{\beta \beta }^{*}Q_{\beta _{0}}\preceq 0 \quad \mathrm{and} \quad Q_{\beta _{0}}^{T}{\widetilde{{Y^{k}}}^*}_{\beta \beta }Q_{\beta _{0}} \equiv Q_{\beta _{0}}^{T}\widetilde{Y}_{\beta \beta }^{*} Q_{\beta _{0}}\succeq 0,\\&\quad k=1,2,\ldots . \end{aligned}$$

From Proposition 3.2 and (77), we know that

$$\begin{aligned} ({X^k}^*,{Y^k}^*)\in N_{\mathrm{gph}\,N_{\mathcal{S}^n_+}}^\pi (X^k,Y^k)\quad \mathrm{and} \quad (X^*,Y^*)=\displaystyle {\lim \limits _{k \rightarrow \infty }} ({X^k}^*,{Y^k}^*). \end{aligned}$$

Hence, the assertion of the theorem follows.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ding, C., Sun, D. & Ye, J.J. First order optimality conditions for mathematical programs with semidefinite cone complementarity constraints. Math. Program. 147, 539–579 (2014). https://doi.org/10.1007/s10107-013-0735-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-013-0735-z

Keywords

Mathematics Subject Classification

Navigation