Skip to main content
Log in

Successive Lagrangian relaxation algorithm for nonconvex quadratic optimization

  • Published:
Journal of Global Optimization Aims and scope Submit manuscript

Abstract

Optimization problems whose objective function and constraints are quadratic polynomials are called quadratically constrained quadratic programs (QCQPs). QCQPs are NP-hard in general and are important in optimization theory and practice. There have been many studies on solving QCQPs approximately. Among them, a semidefinite program (SDP) relaxation is a well-known convex relaxation method. In recent years, many researchers have tried to find better relaxed solutions by adding linear constraints as valid inequalities. On the other hand, the SDP relaxation requires a long computation time, and it has high space complexity for large-scale problems in practice; therefore, the SDP relaxation may not be useful for such problems. In this paper, we propose a new convex relaxation method that is weaker but faster than SDP relaxation methods. The proposed method transforms a QCQP into a Lagrangian dual optimization problem and successively solves subproblems while updating the Lagrange multipliers. The subproblem in our method is a QCQP with only one constraint for which we propose an efficient algorithm. Numerical experiments confirm that our method can quickly find a relaxed solution with an appropriate termination condition.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Notes

  1. Block-SDP could not solve one of ten instances with \(n=10^3\) when \(m= \text{ floor }(0.3 n)\) due to the numerical errors that would have been caused by large scale calculations.

References

  1. Adachi, S., Iwata, S., Nakatsukasa, Y., Takeda, A.: Solving the trust region subproblem by a generalized eigenvalue problem. SIAM J. Optim. 27, 269–291 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  2. Al-Khayyal, F.A., Larsen, C., Voorhis, T.V.: A relaxation method for nonconvex quadratically constrained quadratic programs. J. Global Optim. 6, 215–230 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  3. Anstreicher, K.M.: Semidefinite programming versus the reformulation-linearization technique for nonconvex quadratically constrained quadratic programming. J. Global Optim. 43, 471–484 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  4. Armijo, L.: Minimization of functions having Lipschitz continuous first partial derivatives. Pac. J. Math. 16, 1–3 (1966)

    Article  MathSciNet  MATH  Google Scholar 

  5. Audet, C., Hansen, P., Jaumard, B., Savard, G.: A branch and cut algorithm for nonconvex quadratically constrained quadratic programming. Math. Program. 87, 131–152 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bai, L., Mitchell, J.E., Pang, J.S.: Using quadratic convex reformulation to tighten the convex relaxation of a quadratic program with complementarity constraints. Optim. Lett. 8, 811–822 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  7. Boyd, S., Vandenberghe, L.: Convex Optimization. Cambridge University Press, Cambridge (2010)

    MATH  Google Scholar 

  8. Boyd, S., Xiao, L., Mutapcic, A.: Subgradient Methods. Lecture Notes of EE392o, Autumn Quarter. Stanford University, Stanford (2003)

    Google Scholar 

  9. Burer, S., Kim, S., Kojima, M.: Faster, but weaker, relaxations for quadratically constrained quadratic programs. Comput. Optim. Appl. 59, 27–45 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  10. Chen, Y., Ye, X.: Projection onto a simplex (2011), ArXiv:1101.6081v2. Accessed 17 May 2016

  11. Fujie, T., Kojima, M.: Semidefinite programming relaxation for nonconvex quadratic programs. J. Global Optim. 10, 367–380 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  12. Goemans, M.X.: Semidefinite programming in combinatorial optimization. Math. Program. 79, 143–161 (1997)

    MathSciNet  MATH  Google Scholar 

  13. Goemans, M.X., Williamson, D.P.: Improved approximation algorithms for maximum cut and satisfiability problems using semidefinite programming. J. Assoc. Comput. Mach. 42, 1115–1145 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  14. Hu, Y., Yang, X., Sim, C.: Inexact subgradient methods for quasi-convex optimization problems. Eur. J. Oper. Res. 240, 315–327 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  15. Hu, Y., Yu, C., Li, C.: Stochastic subgradient method for quasi-convex optimization problems. J. Nonlinear Convex Anal. 17, 711–724 (2016)

    MathSciNet  MATH  Google Scholar 

  16. Jiang, R., Li, D.: Convex relaxations with second order cone constraints for nonconvex quadratically constrained quadratic programming (2016). ArXiv:1608.02096v1. Accessed 8 Dec 2016

  17. Kim, S., Kojima, M.: Second order cone programming relaxation of nonconvex quadratic optimization problems. Optim. Methods Softw. 15, 201–224 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  18. Lu, C., Fang, S., Jin, Q., Wang, Z., Xing, W.: KKT solution and conic relaxation for solving quadratically constrained quadratic programming problems. SIAM J. Optim. 21, 1475–1490 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  19. Luo, Z.Q., Ma, W.K., So, A.M.C., Ye, Y., Zhang, S.: Semidefinite relaxation of quadratic optimization problems. IEEE Signal Process. Mag. 27, 20–34 (2010)

    Article  Google Scholar 

  20. Moré, J.J., Sorensen, D.C.: Computing a trust region step. SIAM J. Sci. Stat. Comput. 4, 553–572 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  21. Nocedal, J., Wright, S.J.: Numerical Optimization. Springer, Berlin (2006)

    MATH  Google Scholar 

  22. Novak, I.: Dual bounds and optimality cuts for all-quadratic programs with convex constraints. J. Global Optim. 18, 337–356 (2000)

    Article  MathSciNet  Google Scholar 

  23. Pardalos, P.M., Vavasis, S.A.: Quadratic programming with one negative eigenvalue is NP-hard. J. Global Optim. 1, 15–22 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  24. Rendl, F., Rinaldi, G., Wiegele, A.: Solving max-cut to optimality by intersecting semidefinite and polyhedral relaxations. Math. Program. Ser. A 121, 307–335 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  25. SeDuMi optimization over symmetric cones. http://sedumi.ie.lehigh.edu/

  26. Sherali, H.D., Fraticelli, B.M.P.: Enhancing RLT relaxation via a new class of semidefinite cuts. J. Global Optim. 22, 233–261 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  27. Sturm, J.F., Zhang, S.: On cones of nonnegative quadratic functions. Math. Oper. Res. 28, 246–267 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  28. Tuy, H.: On solving nonconvex optimization problems by reducing the duality gap. J. Global Optim. 32, 349–365 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  29. Vandenberghe, L., Boyd, S.: Semidefinite programming. SIAM Rev. 38, 49–95 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  30. Voorhis, T.V.: A global optimization algorithm using Lagrangian underestimates and the interval newton method. J. Global Optim. 24, 349–370 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  31. Zheng, X.J., Sun, X.L., Li, D.: Convex relaxations for nonconvex quadratically constrained quadratic programming: matrix cone decomposition and polyhedral approximation. Math. Program. 129, 301–329 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  32. Zheng, X.J., Sun, X.L., Li, D.: Nonconvex quadratically constrained quadratic programming: best D.C.decompositions and their SDP representations. J. Global Optim. 50, 695–712 (2011)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Akiko Takeda.

Additional information

This research was done at The University of Tokyo before joining the company, not a research conducted by the company.

Proofs of Theorems

Proofs of Theorems

Proof of Theorem 1

Proof

The vector \(\tilde{{\varvec{g}}}_{{{\varvec{\lambda }}}}\) of (21) can be found from

$$\begin{aligned} \psi ({\varvec{\lambda }})&=\phi (\bar{\mu }_{{{\varvec{\lambda }}}}{\varvec{\lambda }})\nonumber \\&=\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}^\top \left( Q_0+\bar{\mu }_{{{\varvec{\lambda }}}}\sum _{i=1}^m\lambda _i Q_i\right) \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+2\left( {\varvec{q}}_0+\bar{\mu }_{{{\varvec{\lambda }}}}\sum _{i=1}^m\lambda _i {\varvec{q}}_i\right) ^\top \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+\gamma _0+\bar{\mu }_{{{\varvec{\lambda }}}}\sum _{i=1}^m\lambda _i \gamma _i.\nonumber \end{aligned}$$

We prove that the vector \(\tilde{{\varvec{g}}}_{{{\varvec{\lambda }}}}\) is in the quasi-subdifferential \(\partial \psi \) defined by (22). Note that in [14, 15], the quasi-subdifferential is defined for a quasi-convex function, but \(\psi \) is quasi-concave. Therefore (22) is modified from the original definition of \(\partial \psi \) for a quasi-convex function. We further consider (22) as

$$\begin{aligned} \partial \psi ({\varvec{\lambda }})&=\{{\varvec{s}}\mid \psi ({\varvec{\nu }})\le \psi ({\varvec{\lambda }}),\;\;\;\forall {\varvec{\nu }};\;{\varvec{s}}^\top ({\varvec{\nu }}-{\varvec{\lambda }})< 0\}\nonumber \\&=\{{\varvec{s}}\mid \psi ({\varvec{\nu }})\le \psi ({\varvec{\lambda }}),\;\;\;\forall {\varvec{\nu }};\;{\varvec{s}}^\top {\varvec{\nu }}<{\varvec{s}}^\top {\varvec{\lambda }}\} \end{aligned}$$
(37)

Now we show that \(\tilde{{\varvec{g}}}_{{{\varvec{\lambda }}}}\) is in (37). When \(\bar{\mu }_{{{\varvec{\lambda }}}}=0\), \(\tilde{{\varvec{g}}}_{{{\varvec{\lambda }}}}={\varvec{0}}\) satisfies (22) and \(\tilde{{\varvec{g}}}_{{{\varvec{\lambda }}}}\in \partial \psi ({\varvec{\lambda }})\) holds. When \(\bar{\mu }_{{{\varvec{\lambda }}}}>0\), it is sufficient to consider the vector

$$\begin{aligned} {\varvec{g}}_{{{\varvec{\lambda }}}}:=\begin{pmatrix} \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}^\top Q_1\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+2{\varvec{q}}_1^\top \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+\gamma _1\\ \vdots \\ \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}^\top Q_m\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+2{\varvec{q}}_m^\top \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+\gamma _m \end{pmatrix} \end{aligned}$$

instead of \(\tilde{{\varvec{g}}}_{{{\varvec{\lambda }}}}\) because \(\partial \psi ({\varvec{\lambda }})\) forms a cone. Then, since \(\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}\) is feasible for \({\varvec{\lambda }}\), we have

$$\begin{aligned} \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}^\top \left( \sum _{i=1}^m\lambda _iQ_i\right) \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+2\left( \sum _{i=1}^m\lambda _i{\varvec{q}}_i\right) ^\top \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}+\sum _{i=1}^m\lambda _i\gamma _i\le 0. \end{aligned}$$
(38)

From the definition of \({\varvec{g}}_{{{\varvec{\lambda }}}}\), we can rewrite (38) as

$$\begin{aligned} {{\varvec{g}}_{{{\varvec{\lambda }}}}}^\top {\varvec{\lambda }}\le 0. \end{aligned}$$
(39)

Then, we have to prove that for an arbitrary vector \({\varvec{\nu }}\) which satisfies \({{\varvec{g}}_{{{\varvec{\lambda }}}}}^\top {\varvec{\nu }}<{{\varvec{g}}_{{{\varvec{\lambda }}}}}^\top {\varvec{\lambda }}\),

$$\begin{aligned} \psi ({\varvec{\nu }})\le \psi ({\varvec{\lambda }}) \end{aligned}$$

holds. From (39), we get \({{\varvec{g}}_{{{\varvec{\lambda }}}}}^\top {\varvec{\nu }}\le 0\) and it means that \(\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}\) is feasible for \({\varvec{\nu }}\). This implies that at an optimal solution \(\bar{{\varvec{x}}}_{{{\varvec{\nu }}}}\) for \({\varvec{\nu }}\), the optimal value of (17) is less than or equal to the one at \(\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}}\). Therefore, we have \(\psi ({\varvec{\nu }})\le \psi ({\varvec{\lambda }})\).

Next, we prove (ii)–(iv). First, \(\phi ({\varvec{\lambda }})\) defined by (3) is concave for \({\varvec{\lambda }}\). It is a general property of the objective function of the Lagrangian dual problem (e.g. [7]). Furthermore, note that \(\psi ({\varvec{\lambda }})\) is the maximum value of \(\phi (\mu {\varvec{\lambda }})\) with respect to \(\mu \ge 0\). Therefore, \(\psi ({\varvec{\lambda }})\ge \phi ({\varvec{0}})\) holds for all \({\varvec{\lambda }}\in \Lambda _{\mathrm {s}}\).

We show (ii) first. Let \({\varvec{\lambda }}_1,{\varvec{\lambda }}_2\;({\varvec{\lambda }}_1\ne {\varvec{\lambda }}_2)\) be arbitrary points in \(\Lambda _{\mathrm {s}}\), and \(\bar{\mu }_1\) and \(\bar{\mu }_2\) be optimal solutions of (16) with fixed \({\varvec{\lambda }}_1\) and \({\varvec{\lambda }}_2\), respectively. Without loss of generality, we assume that \(\psi ({\varvec{\lambda }}_1)\ge \psi ({\varvec{\lambda }}_2)\). Now, it is sufficient to prove that for any fixed \(\alpha \in [0,1]\),

$$\begin{aligned} \psi ({\varvec{\lambda }}_\alpha )\ge \psi ({\varvec{\lambda }}_2), \end{aligned}$$
(40)

where \({\varvec{\lambda }}_\alpha :=\alpha {\varvec{\lambda }}_1+(1-\alpha ){\varvec{\lambda }}_2\). If \({\varvec{\lambda }}_1\notin \Lambda _+\) or \({\varvec{\lambda }}_2\notin \Lambda _+\) holds, we get \(\psi ({\varvec{\lambda }}_2)=\phi ({\varvec{0}})\) and (40) holds. Therefore, we only have to consider the case when both \({\varvec{\lambda }}_1\) and \({\varvec{\lambda }}_2\) are in \(\Lambda _+\), implying that \(\bar{\mu }_1\) and \(\bar{\mu }_2\) are positive by (20). Since \(\phi ({\varvec{\lambda }})\) is concave for \({\varvec{\lambda }}\), we can see that for any \(\beta \in [0,1]\),

$$\begin{aligned} \phi ({\varvec{\xi }}(\beta ))&\ge \beta \phi (\bar{\mu }_1{\varvec{\lambda }}_1)+(1-\beta )\phi (\bar{\mu }_2{\varvec{\lambda }}_2)\\&=\beta \psi ({\varvec{\lambda }}_1)+(1-\beta )\psi ({\varvec{\lambda }}_2), \end{aligned}$$

where \({\varvec{\xi }}(\beta ):=\beta \bar{\mu }_1{\varvec{\lambda }}_1+(1-\beta )\bar{\mu }_2{\varvec{\lambda }}_2\). Accordingly, we can confirm that there exists

$$\begin{aligned} \bar{\beta }:=\frac{\bar{\mu }_2\alpha }{\bar{\mu }_1(1-\alpha )+\bar{\mu }_2\alpha }\in [0,1] \end{aligned}$$

which satisfies

$$\begin{aligned} {\varvec{\xi }}(\bar{\beta })=\frac{\bar{\mu }_1\bar{\mu }_2}{\bar{\mu }_1(1-\alpha )+\bar{\mu }_2\alpha }{\varvec{\lambda }}_\alpha \end{aligned}$$

For this \(\bar{\beta }\), we get

$$\begin{aligned} \psi ({\varvec{\lambda }}_\alpha )&=\phi (\bar{\mu }_{{{\varvec{\lambda }}}_\alpha }{\varvec{\lambda }}_\alpha )\nonumber \\&\ge \phi \left( \frac{\bar{\mu }_1\bar{\mu }_2}{\bar{\mu }_1(1-\alpha )+\bar{\mu }_2\alpha }{\varvec{\lambda }}_\alpha \right) \nonumber \\&=\phi ({\varvec{\xi }}(\bar{\beta }))\nonumber \\&\ge \bar{\beta }\psi ({\varvec{\lambda }}_1)+(1-\bar{\beta })\psi ({\varvec{\lambda }}_2)\nonumber \\&\ge \psi ({\varvec{\lambda }}_2), \end{aligned}$$
(41)

where \(\bar{\mu }_{{{\varvec{\lambda }}}_\alpha }\) is an optimal solution for \({\varvec{\lambda }}_\alpha \). Therefore, (ii) holds.

We can easily prove (iii). In the above proof of (ii), we assume \({\varvec{\lambda }}_1,{\varvec{\lambda }}_2\) is in \(\Lambda _+\). Then, (41) means that \(\psi ({\varvec{\lambda }}_\alpha )\ge \psi ({\varvec{\lambda }}_2)>\phi ({\varvec{0}})\), and we get \({\varvec{\lambda }}_\alpha \in \Lambda _+\) for any \(\alpha \in [0,1]\).

Lastly, we prove (iv). Let \({\varvec{\lambda }}^\dag \) be an arbitrary stationary point in \(\Lambda _+\), and let \(\bar{\mu }_{{{\varvec{\lambda }}}^\dag }\) be an optimal solution of (16) for \({\varvec{\lambda }}^\dag \). Moreover, \(\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }\) denotes an optimal solution of \(\phi (\bar{\mu }_{{{\varvec{\lambda }}}^\dag }{\varvec{\lambda }}^\dag )\). From (21) and (20), we have

$$\begin{aligned} \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }^\top Q_i\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }+2{\varvec{q}}_i^\top \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }+\gamma _i=0, \qquad i=1,\ldots ,m. \end{aligned}$$
(42)

On the other hand, it can be confirmed that

$$\begin{aligned} \begin{pmatrix} \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }^\top Q_1\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }+2{\varvec{q}}_1^\top \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }+\gamma _1\\ \vdots \\ \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }^\top Q_m\bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }+2{\varvec{q}}_m^\top \bar{{\varvec{x}}}_{{{\varvec{\lambda }}}^\dag }+\gamma _m \end{pmatrix} \end{aligned}$$

is a subgradient vector for \(\phi ({\varvec{\xi }}^\dag )\), where \({\varvec{\xi }}^\dag =\bar{\mu }_{{{\varvec{\lambda }}}^\dag }{\varvec{\lambda }}^\dag \). Hence, (42) implies that \(\bar{\mu }_{{{\varvec{\lambda }}}^\dag }{\varvec{\lambda }}^\dag \) is a stationary point of \(\phi \). From the properties of concave functions, all stationary points are global optimal solutions. Therefore, \(\bar{\mu }_{{{\varvec{\lambda }}}^\dag }{\varvec{\lambda }}^\dag \) is a global optimal solution of \(\phi \) and \({\varvec{\lambda }}^\dag \) is also a global optimal solution of \(\psi \). \(\square \)

Proof of Theorem 2

Proof

Let \(\Delta \) be the feasible region for (\({\varvec{x}},t\)) of (23) and \(\Delta _{\mathrm {rel}}\) be the feasible region of the relaxed problem (26). Let \(\mathrm {conv}(\Delta )\) be the convex hull of \(\Delta \). We will prove that \(\Delta _{\mathrm {rel}}=\mathrm {conv}(\Delta )\). In this proof, we write “\({\varvec{a}}\xleftarrow {conv}\{{\varvec{b}},{\varvec{c}}\}\)” if

$$\begin{aligned} \exists s\in [0,1]\;\;\mathrm {s.t.}\;\;{\varvec{a}}=s{\varvec{b}}+(1-s){\varvec{c}} \end{aligned}$$

holds. From the definition, it is obvious that \(\Delta _{\mathrm {rel}}\) is convex and \(\Delta \subseteq \Delta _{\mathrm {rel}}\) holds. Meanwhile, the definition of the convex hull is the minimum convex set that includes \(\Delta \), which implies \(\mathrm {conv}(\Delta )\subseteq \Delta _{\mathrm {rel}}\) is obvious. The convex hull consists of all the points obtained as convex combinations of any points in \(\Delta \). Therefore, if the proposition,

$$\begin{aligned} \forall {\varvec{z}}\in \Delta _{\mathrm {rel}},\;\;\exists {\varvec{z}}_1,{\varvec{z}}_2 \in \Delta ; \;\;{\varvec{z}}\xleftarrow {conv}\{{\varvec{z}}_1,{\varvec{z}}_2\} \end{aligned}$$
(43)

holds, it leads to \(\Delta _{\mathrm {rel}}\subseteq \mathrm {conv}(\Delta )\) and we get \(\Delta _{\mathrm {rel}}=\mathrm {conv}(\Delta )\). To show (43), let us choose an arbitrary point \(({\varvec{x}}^*,t)\in \Delta _{\mathrm {rel}}\) and let \(t^*\) be the lower bound of t for \({\varvec{x}}^*\) in \(\Delta _{\mathrm {rel}}\). Since \(({\varvec{x}}^*,t)\in \Delta _{\mathrm {rel}}\) holds for all \(t\ge t^*\), if

$$\begin{aligned} \forall ({\varvec{x}}^*,t^*)\in \Delta _{\mathrm {rel}},\;\;\exists ({\varvec{x}}_1,t_1),({\varvec{x}}_2,t_2) \in \Delta ;\;\;\;({\varvec{x}}^*,t^*)\xleftarrow {conv}\{({\varvec{x}}_1,t_1),({\varvec{x}}_2,t_2)\} \end{aligned}$$
(44)

holds, then for any \(\delta _t\ge 0\), \(({\varvec{x}}^*,t^*+\delta _t)\in \Delta _{\mathrm {rel}}\) \(\xleftarrow {conv} \{({\varvec{x}}_1,t_1+\delta _t), ({\varvec{x}}_2,t_2+\delta _t)\}\). These points are in \(\Delta \), and therefore, it is sufficient to focus on the case of \(t=t^*\).

To prove (44), we claim that if a point \(({\varvec{x}},t)\in \Delta _{\mathrm {rel}}\) satisfies both inequalities of (26) with equality, then the point is also in \(\Delta \). Since \(\underline{\sigma }<\bar{\sigma }\), we can see that \({\varvec{x}}^\top Q_{{{\varvec{\lambda }}}}{\varvec{x}}+2{\varvec{q}}_{{{\varvec{\lambda }}}}^\top {\varvec{x}}+\gamma _{{{\varvec{\lambda }}}}=0\) by setting the inequalities of (26) to equality and taking their difference. Then, we can easily get \({\varvec{x}}^\top Q_0{\varvec{x}}+2{\varvec{q}}_0^\top {\varvec{x}}+\gamma _0=t\). Therefore, \(({\varvec{x}},t)\) is feasible for (23) and in \(\Delta \). In what follows, we focus on when only one of the two inequalities is active.

Then, we have to prove (44) for when \(Q_{{{\varvec{\lambda }}}}\succeq O\) and \(Q_{{{\varvec{\lambda }}}}\nsucceq O\). However, due to space limitations, we will only show the harder case, i.e., when \(Q_{{{\varvec{\lambda }}}}\nsucceq O\), implying \(0<\bar{\sigma }<\infty \). The proof of the other case is almost the same. In the following explanation, we want to find two points in \(\Delta \) (i.e., points which satisfy both inequalities of (26) with equality). Figure 3 illustrates \(\Delta \) and \(\Delta _{\mathrm {rel}}\). In the figure, we want to find P and Q.

The optimal solution \(({\varvec{x}}^*,t^*)\) of the relaxation problem (26) satisfies at least one of the two inequalities with equality. Here, we claim that the matrix \(Q_0+\sigma Q_{{{\varvec{\lambda }}}}\) (\(\sigma \in \{\bar{\sigma },\underline{\sigma }\}\)) in the active inequality has at least one zero eigenvalue and the kernel is not empty if \(({\varvec{x}}^*,t^*)\notin \Delta \) (the claim is proved at the end of this proof). We denote the matrix in the inactive inequality as \(Q_0+\sigma ' Q_{{{\varvec{\lambda }}}}\) (\(\sigma '\in \{\bar{\sigma },\underline{\sigma }\}\)). By using \(\sigma \) and \(\sigma '\), \(({\varvec{x}}^*,t^*)\) satisfies (26) as follows:

$$\begin{aligned} {\left\{ \begin{array}{ll} &{}{{\varvec{x}}^*}^\top (Q_0+\sigma Q_{{{\varvec{\lambda }}}}){\varvec{x}}^*+2({\varvec{q}}_0+\sigma {\varvec{q}}_{{{\varvec{\lambda }}}})^\top {\varvec{x}}^*+\gamma _0+\sigma \gamma _{{{\varvec{\lambda }}}}=t^*, \\ &{}{{\varvec{x}}^*}^\top (Q_0+\sigma 'Q_{{{\varvec{\lambda }}}}){\varvec{x}}^*+2({\varvec{q}}_0+\sigma '{\varvec{q}}_{{{\varvec{\lambda }}}})^\top {\varvec{x}}^*+\gamma _0+\sigma '\gamma _{{{\varvec{\lambda }}}}\le t^*. \end{array}\right. } \end{aligned}$$
(45)

Since \(Q_0+\sigma Q_{{{\varvec{\lambda }}}}\;(\succeq O)\) has a zero eigenvalue, we can decompose \({\varvec{x}}^*\) into

$$\begin{aligned} {\varvec{x}}^*={\varvec{u}}^*+\tau ^*{\varvec{v}}^*\;\;\;\mathrm {s.t.}\;\;\;{\varvec{u}}^*&\in \mathrm {Ker}(Q_0+\sigma Q_{{{\varvec{\lambda }}}})^\bot ,\nonumber \\ {\varvec{v}}^*&\in \mathrm {Ker}(Q_0+\sigma Q_{{{\varvec{\lambda }}}}),\;\Vert {\varvec{v}}^*\Vert _2=1,\nonumber \\ \tau ^*&\in \mathbb {R}. \end{aligned}$$
(46)

Substituting these expressions into the constraints of (45), we get

$$\begin{aligned} {\left\{ \begin{array}{ll} ({\varvec{u}}^*+\tau ^*{\varvec{v}}^*)^\top (Q_0+\sigma Q_{{{\varvec{\lambda }}}})({\varvec{u}}^*+\tau ^*{\varvec{v}}^*)+2({\varvec{q}}_0+\sigma {\varvec{q}}_{{{\varvec{\lambda }}}})^\top ({\varvec{u}}^*+\tau ^*{\varvec{v}}^*)+\gamma _0+\sigma \gamma _{{{\varvec{\lambda }}}}=t^*,\\ ({\varvec{u}}^*+\tau ^*{\varvec{v}}^*)^\top (Q_0+\sigma ' Q_{{{\varvec{\lambda }}}})({\varvec{u}}^*+\tau ^*{\varvec{v}}^*)+2({\varvec{q}}_0+\sigma '{\varvec{q}}_{{{\varvec{\lambda }}}})^\top ({\varvec{u}}^*+\tau ^*{\varvec{v}}^*)+\gamma _0+\sigma '\gamma _{{{\varvec{\lambda }}}}\le t^*. \end{array}\right. } \end{aligned}$$
(47)

By fixing \({\varvec{u}}^*\) and \({\varvec{v}}^*\), we can see that (47) is of the form,

$$\begin{aligned} {\left\{ \begin{array}{ll}A+\alpha \tau ^*=t^*,\\ B+\beta \tau ^*+\gamma (\tau ^*)^2\le t^*, \end{array}\right. } \end{aligned}$$
(48)

where \((A,B,\alpha ,\beta ,\gamma )\) are appropriate constants. Here, we regard \(\tau ^*\) and \(t^*\) in (48) as variables \((\tau , t)\) and illustrate (48) in Fig. 9. The feasible region of (26) for fixed \({\varvec{u}}^*\) and \({\varvec{v}}^*\) is shown by the bold line. Note that the line and the parabola have at least one intersection point \(({\varvec{x}}^*, t^*)\). Here, both points P\((\tau _1,t_1)\) and Q\((\tau _2,t_2)\) in Fig. 9 satisfy both formulas of (48) with equality, so these points are in \(\Delta \). Furthermore, it is obvious from Fig. 9 that \((\tau ^*,t^*)\xleftarrow {conv}\{(\tau _1,t_1),(\tau _2,t_2)\}\). Therefore, (44) holds for any \(({\varvec{x}}^*,t^*)\).

Fig. 9
figure 9

The feasible region of \((\tau ,t)\)

Now let us check that \(\gamma >0\) and thereby show that the second formula of (48) actually forms a parabola. In (48), \(\gamma ={{\varvec{v}}^*}^\top (Q_0+\sigma 'Q_{{{\varvec{\lambda }}}}){\varvec{v}}^*\ge 0\). However, if \(\gamma =0\), then \({\varvec{v}}^*\in \mathrm {Ker}(Q_0+\sigma 'Q_{{{\varvec{\lambda }}}})\), so \((Q_0+\sigma 'Q_{{{\varvec{\lambda }}}}){\varvec{v}}^*=0\) holds. Meanwhile, from the definition of \({\varvec{v}}^*\), we have \((Q_0+\sigma Q_{{{\varvec{\lambda }}}}){\varvec{v}}^*=0\). Since \(\sigma \ne \sigma '\), we get \({\varvec{v}}^*\in \mathrm {Ker}(Q_0)\cap \mathrm {Ker}(Q_{{{\varvec{\lambda }}}})\), and this contradicts the Dual Slater condition.

Finally, we prove that \(Q_0+\sigma Q_{{{\varvec{\lambda }}}}\;(\succeq O)\) has a zero eigenvalue if \(({\varvec{x}}^*,t^*)\notin \Delta \). From the definition of \(\bar{\sigma }\) and \(\underline{\sigma }\) (see (24) and (25)), \(Q_0+\bar{\sigma }Q_{{{\varvec{\lambda }}}}\) or \(Q_0+\underline{\sigma }Q_{{{\varvec{\lambda }}}}\) has a zero eigenvalue if \(\bar{\sigma }\) or \(\underline{\sigma }\) is positive. Moreover, from the Dual Slater condition, \(\bar{\sigma }>0\) holds, so \(Q_0+\bar{\sigma }Q_{{{\varvec{\lambda }}}}\) always has a zero eigenvalue. Therefore, we only have to consider the case when \(\underline{\sigma }=0\), i.e., \(Q_0+\underline{\sigma } Q_{{{\varvec{\lambda }}}}=Q_0\), implying \(Q_0\succeq O\). If \(Q_0\) does not have a zero eigenvalue (i.e. \(Q_0\succ O\)), the claim does not hold. However, we can confirm that in this case \(({\varvec{x}}^*,t^*)\) is already feasible for (23) (i.e. \(({\varvec{x}}^*,t^*)\in \Delta \)) and we do not need to consider this case. We can check its feasibility for (23) by subtracting an equality with \(\underline{\sigma }\;(=0)\) from an inequality with \(\bar{\sigma }\) and dividing the resulting inequality by \(\bar{\sigma }\;(>0)\). \(\square \)

Remark 3

This proof suggests that if an optimal solution \({\varvec{x}}^*\) of (26) is found and \({\varvec{x}}^*\) is not feasible for (23), we can find an optimal feasible solution of (23). If \(({\varvec{x}}^*,t^*)\) is an optimal solution of (26), then in Fig. 9, the slope \(\alpha \) of the line \(A+\alpha \tau ^*=t^*\) is zero or \(({\varvec{x}}^*,t^*)\) is equal to the end point P or Q and is already feasible for (23) because \(t^*\) is the minimum value such that we can not obtain a t any smaller than \(t^*\) in the bold line part in Fig. 9. In the former case, we can find an optimal feasible solution of (23) by moving \({\varvec{x}}^*\) in the direction of \(\pm {\varvec{v}}^*\) defined in (46). We find an optimal feasible solution of (\(\mathrm {P}_k\)) in this way in SLR.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yamada, S., Takeda, A. Successive Lagrangian relaxation algorithm for nonconvex quadratic optimization. J Glob Optim 71, 313–339 (2018). https://doi.org/10.1007/s10898-018-0617-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10898-018-0617-2

Keywords

Navigation