Skip to main content
Log in

Solving joint chance constrained problems using regularization and Benders’ decomposition

  • S.I.: Stochastic Optimization:Theory&Applications in Memory of M.Bertocchi
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

We consider stochastic programs with joint chance constraints with discrete random distribution. We reformulate the problem by adding auxiliary variables. Since the resulting problem has a non-regular feasible set, we regularize it by increasing the feasible set. We solve the regularized problem by iteratively solving a master problem while adding Benders’ cuts from a slave problem. Since the number of variables of the slave problem equals to the number of scenarios, we express its solution in a closed form. We show convergence properties of the solutions. On a gas network design problem, we perform a numerical study by increasing the number of scenarios and compare our solution with a solution obtained by solving the same problem with the continuous distribution.

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

Similar content being viewed by others

Notes

  1. http://staff.utia.cas.cz/adam/research.html.

References

  • Adam, L., & Branda, M. (2016). Nonlinear chance constrained problems: Optimality conditions, regularization and solvers. Journal of Optimization Theory and Applications, 170(2), 419–436.

    Google Scholar 

  • Ahmed, S. (2014). Convex relaxations of chance constrained optimization problems. Optimization Letters, 8(1), 1–12.

    Google Scholar 

  • Allevi, E., Bertocchi, M. I., Vespucci, M. T., & Innorta, M. (2007). A mixed integer nonlinear optimization model for gas sale company. Optimization Letters, 1(1), 61–69.

    Google Scholar 

  • Barrera, J., Homem-de Mello, T., Moreno, E., Pagnoncelli, B. K., & Canessa, G. (2016). Chance-constrained problems and rare events: an importance sampling approach. Mathematical Programming, 157(1), 153–189.

    Google Scholar 

  • Ben-Tal, A., & Nemirovski, A. (1998). Robust convex optimization. Mathematics of Operations Research, 23(4), 769–805.

    Google Scholar 

  • Beraldi, P., & Bruni, M. (2010). An exact approach for solving integer problems under probabilistic constraints with random technology matrix. Annals of Operations Research, 177(1), 127–137.

    Google Scholar 

  • Birge, J. R., & Louveaux, F. (2011). Introduction to stochastic programming. Berlin: Springer.

    Google Scholar 

  • Branda, M., & Dupačová, J. (2012). Approximation and contamination bounds for probabilistic programs. Annals of Operations Research, 193(1), 3–19.

    Google Scholar 

  • Chen, W., Sim, M., Sun, J., & Teo, C. P. (2010). From cvar to uncertainty set: Implications in joint chance-constrained optimization. Operations Research, 58(2), 470–485.

    Google Scholar 

  • Cheng, J., & Lisser, A. (2012). A second-order cone programming approach for linear programs with joint probabilistic constraints. Operations Research Letters, 40(5), 325–328.

    Google Scholar 

  • Curtis, F., Wächter, A., & Zavala, V. (2018). A sequential algorithm for solving nonlinear optimization problems with chance constraints. SIAM Journal on Optimization, 28(1), 930–958.

    Google Scholar 

  • Déak, I. (2000). Subroutines for computing normal probabilities of sets-computer experiences. Annals of Operations Research, 100, 103–122.

    Google Scholar 

  • Dentcheva, D., & Martinez, G. (2012). Augmented Lagrangian method for probabilistic optimization. Annals of Operations Research, 200(1), 109–130.

    Google Scholar 

  • Dentcheva, D., & Martinez, G. (2013). Regularization methods for optimization problems with probabilistic constraints. Mathematical Programming, 138(1–2), 223–251.

    Google Scholar 

  • Dentcheva, D., Prékopa, A., & Ruszczyński, A. (2000). Concavity and efficient points of discrete distributions in probabilistic programming. Mathematical Programming, 89(1), 55–77.

    Google Scholar 

  • Dick, J., & Pillichshammer, F. (2010). Digital nets and sequences: Discrepancy theory and quasi-monte carlo integration. New York: Cambridge University Press.

    Google Scholar 

  • Geletu, A., Hoffmann, A., Kläppel, M., & Li, P. (2017). An inner-outer approximation approach to chance constrained optimization. SIAM Journal on Optimization, 27(3), 1834–1857.

    Google Scholar 

  • Genz, A., & Bretz, F. (2009). Computation of multivariate normal and t-probabilities, Lecture Notes in Statistics (Vol. 195). Heidelberg: Springer.

    Google Scholar 

  • Gonzalez Grandon, T., Heitsch, H., & Henrion, R. (2017). A joint model of probabilistic /robust constraints for gas transport management in stationary networks. Computational Management Science, 14, 443–460.

    Google Scholar 

  • Gotzes, C., Heitsch, H., Henrion, R., & Schultz, R. (2016). On the quantification of nomination feasibility in stationary gas networks with random load. Mathematical Methods of Operations Research, 84(2), 427–457.

    Google Scholar 

  • Haneveld, W., & van der Vlerk, M. (2006). Integrated chance constraints: Reduced forms and an algorithm. Computational Management Science, 3(4), 245–269.

    Google Scholar 

  • Hastie, T., Tibshirani, R., & Friedman, J. H. (2001). The elements of statistical learning. Berlin: Springer.

    Google Scholar 

  • Henrion, R. (2013). A critical note on empirical (Sample Average, Monte Carlo) approximation of solutions to chance constrained programs. In D. Hömberg & F. Tröltzsch (Eds.), System modeling and optimization, IFIP advances in information and communication Technology (Vol. 391, pp. 25–37). Berlin Heidelberg: Springer.

    Google Scholar 

  • Henrion, R., & Möller, A. (2012). A gradient formula for linear chance constraints under Gaussian distribution. Mathematics of Operations Research, 37(3), 475–488.

    Google Scholar 

  • Henrion, R., & Römisch, W. (2004). Hölder and Lipschitz stability of solution sets in programs with probabilistic constraints. Mathematical Programming, 100(3), 589–611.

    Google Scholar 

  • Henrion, R., & Strugarek, C. (2008). Convexity of chance constraints with independent random variables. Computational Optimization and Applications, 41(2), 263–276.

    Google Scholar 

  • Ioffe, A. D., & Outrata, J. (2008). On metric and calmness qualification conditions in subdifferential calculus. Set-Valued Analysis, 16(2–3), 199–227.

    Google Scholar 

  • Kaňková, V. (1990). On the convergence rate of empirical estimates in chance constrained stochastic programming. Kybernetika, 26(6), 449–461.

    Google Scholar 

  • Koch, T., Hiller, B., Pfetsch, M.E., & Schewe, L. (eds.) (2015). Evaluating gas network capacities, MOS-SIAM Series on Optimization, vol. 21. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA.

  • Küçükyavuz, S. (2012). On mixing sets arising in chance-constrained programming. Mathematical Programming, 132(1), 31–56.

    Google Scholar 

  • Lejeune, M. (2012). Pattern definition of the p-efficiency concept. Annals of Operations Research, 200(1), 23–36.

    Google Scholar 

  • Lejeune, M., & Noyan, N. (2010). Mathematical programming approaches for generating p-efficient points. European Journal of Operational Research, 207(2), 590–600.

    Google Scholar 

  • Luedtke, J. (2014). A branch-and-cut decomposition algorithm for solving chance-constrained mathematical programs with finite support. Mathematical Programming, 146(1), 219–244.

    Google Scholar 

  • Luedtke, J., & Ahmed, S. (2008). A sample approximation approach for optimization with probabilistic constraints. SIAM Journal on Optimization, 19(2), 674–699.

    Google Scholar 

  • Luedtke, J., Ahmed, S., & Nemhauser, G. (2010). An integer programming approach for linear programs with probabilistic constraints. Mathematical Programming, 122(2), 247–272.

    Google Scholar 

  • Maggioni, F., Bertocchi, M., Giacometti, R., Vespucci, M. T., Innorta, M., & Allevi, E. (2010). A stochastic optimization model for gas retail with temperature scenarios and oil price parameters. IMA Journal of Management Mathematics, 21(2), 149–163.

    Google Scholar 

  • Nemirovski, A., & Shapiro, A. (2007). Convex approximations of chance constrained programs. SIAM Journal on Optimization, 17(4), 969–996.

    Google Scholar 

  • Prékopa, A. (1990). Dual method for the solution of a one-stage stochastic programming problem with random RHS obeying a discrete probability distribution. Zeitschrift für Operations Research, 34(6), 441–461.

    Google Scholar 

  • Prékopa, A. (1995). Stochastic programming. Budapest: Kluwer, Dordrecht and Académiai Kiadó.

    Google Scholar 

  • Prékopa, A. (2003). Probabilistic programming. In A. Ruszczyński & A. Shapiro (Eds.), Stochastic programming, handbooks in operations research and management science (Vol. 10, pp. 267–351). Amsterdam: Elsevier.

    Google Scholar 

  • Rahmaniani, R., Crainic, T. G., Gendreau, M., & Rei, W. (2017). The benders decomposition algorithm: A literature review. European Journal of Operational Research, 259(3), 801–817.

    Google Scholar 

  • Rockafellar, R., & Uryasev, S. (2000). Optimization of conditional value-at-risk. Journal of Risk, 2, 21–41.

    Google Scholar 

  • Rockafellar, R., & Uryasev, S. (2002). Conditional value-at-risk for general loss distributions. Journal of Banking & Finance, 26(7), 1443–1471.

    Google Scholar 

  • Rockafellar, R. T., & Wets, R. J. B. (1998). Variational analysis. Berlin: Springer.

    Google Scholar 

  • Royset, J., & Polak, E. (2007). Extensions of stochastic optimization results to problems with system failure probability functions. Journal of Optimization Theory and Applications, 133, 1–18.

    Google Scholar 

  • Scholtes, S. (2001). Convergence properties of a regularization scheme for mathematical programs with complementarity constraints. SIAM Journal on Optimization, 11(4), 918–936.

    Google Scholar 

  • Shan, F., Zhang, L., & Xiao, X. (2014). A smoothing function approach to joint chance-constrained programs. Journal of Optimization Theory and Applications, 163(1), 181–199.

    Google Scholar 

  • Shapiro, A., Dentcheva, D., & Ruszczyński, A. (2009). Lectures on Stochastic Programming. MOS-SIAM Series on Optimization. SIAM-Society for Industrial and Applied Mathematics, Philadelphia.

  • Sun, H., Xu, H., & Wang, Y. (2014). Asymptotic analysis of sample average approximation for stochastic optimization problems with joint chance constraints via conditional value at risk and difference of convex functions. Journal of Optimization Theory and Applications, 161(1), 257–284.

    Google Scholar 

  • van Ackooij, W. (2015). Eventual convexity of chance constrained feasible sets. Optimization, 64(5), 1263–1284.

    Google Scholar 

  • van Ackooij, W., Berge, V., de Oliveira, W., & Sagastizábal, C. (2017). Probabilistic optimization via approximate p-efficient points and bundle methods. Computers and Operations Research, 77, 177–193.

    Google Scholar 

  • van Ackooij, W., Frangioni, A., & de Oliveira, W. (2016). Inexact stabilized benders’ decomposition approaches with application to chance-constrained problems with finite support. Computational Optimization and Applications, 65(3), 637–669.

    Google Scholar 

  • van Ackooij, W., & Henrion, R. (2014). Gradient formulae for nonlinear probabilistic constraints with gaussian and gaussian-like distributions. SIAM Journal on Optimization, 24, 1864–1889.

    Google Scholar 

  • van Ackooij, W., & Henrion, R. (2017). (Sub-) gradient formulae for probability functions of random inequality systems under gaussian distribution. SIAM/ASA Journal Uncertainty Quantification, 5, 63–87.

    Google Scholar 

  • van Ackooij, W., & Malick, J. (2018). Eventual convexity of probability constraints with elliptical distributions. Mathematical Programming.

  • Wiecek, M.M., & Dranichak, G.M. (2016). Robust multiobjective optimization for decision making under uncertainty and conflict, chap. 4, pp. 84–114.

  • Wozabal, D., Hochreiter, R., & Pflug, G. C. (2010). A difference of convex formulation of value-at-risk constrained optimization. Optimization, 59(3), 377–400.

    Google Scholar 

  • Xie, W., & Ahmed, S. (2017). On quantile cuts and their closure for chance constrained optimization problems. Mathematical Programming.

  • Zhang, Y., Wang, J., Zeng, B., & Hu, Z. (2017). Chance-constrained two-stage unit commitment under uncertain load and wind power output using bilinear benders decomposition. IEEE Transactions on Power Systems, 32(5), 3637–3647.

    Google Scholar 

  • Zhao, M., Huang, K., & Zeng, B. (2017). A polyhedral study on chance constrained program with random right-hand side. Mathematical Programming, 166(1), 19–64.

    Google Scholar 

  • Zymler, S., Kuhn, D., & Rustem, B. (2013). Distributionally robust joint chance constraints with second-order moment information. Mathematical Programming, 137(1–2), 167–198.

    Google Scholar 

Download references

Acknowledgements

We would like to thank two anonymous reviewers for carefully reading the manuscript and for suggesting many improvements.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lukáš Adam.

Additional information

This paper is dedicated to Marida Bertocchi.

This work was supported by the National Natural Science Foundation of China (Grant No. 61850410534), the Program for Guangdong Introducing Innovative and Enterpreneurial Teams (Grant No. 2017ZT07X386), Shenzhen Peacock Plan (Grant No. KQTD2016112514355531), the Grant Agency of the Czech Republic (18-04145S, 18-05631S), and by the Deutsche Forschungsgemeinschaft within project B04 in the Sonderforschungsbereich / Transregio 154 Mathematical Modelling, Simulation and Optimization.

Appendices

A New result for sufficient optimality conditions for hierarchical problems

In this short section, we present a new result which may play a crucial role in deriving sufficient optimality conditions for hierarchical problems such as bilevel problems, mathematical problems with equilibrium/complementarity/vanishing constraints and so on. In these problems, the feasible set is usually rather nasty but may be written as a finite union of nice sets. We show that if these nice sets are convex, then strong stationary points are immediately local minima.

Theorem 5

Consider a convex differentiable function \(f:{\mathbb {R}}^n\rightarrow {\mathbb {R}}\), a set \(X\subset {\mathbb {R}}^n\), a point \(\bar{x}\in X\) and an optimization problem

$$\begin{aligned} \begin{aligned} \hbox {minimize}&f(x)\\ \hbox {subject to}\,&x\in X. \end{aligned} \end{aligned}$$
(23)

Assume that \(\bar{x}\) is its S-stationary point, thus a point with

$$\begin{aligned} 0\in \nabla f(\bar{x}) + \hat{N}_X(\bar{x}), \end{aligned}$$

where \(\hat{N}_X(\bar{x})\) stands for the Fréchet normal cone of X at \(\bar{x}\). If X can be locally around \(\bar{x}\) written as a union of a finite number of (possibly overlapping) convex sets, then \(\bar{x}\) is a local minimum of problem (23).

Proof

From the theorem statement, there are convex sets \(X_i\), \(i=1,\ldots ,I\) such that locally around \(\bar{x}\) we have that X coincides with \(\cup _{i=1}^IX_i\). First, we realize that

$$\begin{aligned} \hat{N}_X(\bar{x}) = (T_X(\bar{x}))^* = (\cup _{i=1}^I T_{X_i}(\bar{x}))^* = \cap _{i=1}^I (T_{X_i}(\bar{x}))^* = \cap _{i=1}^I N_{X_i}(\bar{x}). \end{aligned}$$

Since \(\bar{x}\) is a S-stationary point of (23), we have

$$\begin{aligned} 0\in \nabla f(\bar{x}) + \hat{N}_X(\bar{x}) = \nabla f(\bar{x}) + \bigcap _{i=1}^I N_{X_i}(\bar{x}) = \bigcap _{i=1}^I\Big (\nabla f(\bar{x}) + N_{X_i}(\bar{x})\Big ). \end{aligned}$$

Fix now any i. From the equation above we obtain that \(\bar{x}\) is a stationary point of

$$\begin{aligned} \begin{aligned} {\hbox {minimize}}&f(x)\\ \hbox {subject to}\,&x\in X_i. \end{aligned} \end{aligned}$$

Due to the data convexity, it is a local minimum of the above problem and thus for all \(x\in X_i\) sufficiently close to \(\bar{x}\) we have \(f(x)\ge f(\bar{x})\). But since i was chosen arbitrarily, we obtain that \(\bar{x}\) is a local minimum of problem (23). \(\square \)

B Proofs

In this section, we collect the skipped proofs from Sect. 2.

Proof

(Proposition 1) It is sufficient to follow the proof of Lemma 3.1 in Adam and Branda (2016). \(\square \)

Proof

(Theorem 1) Denote the feasible set of problem (2) by Z and consider a point \(\bar{x}\in Z\). Note that Z can be written as

$$\begin{aligned} Z = \bigcup _{I\subset \{1,\ldots ,S\}, \sum _{i\in I}p_i\ge 1-\varepsilon } {\left\{ x:\begin{aligned} g_k(x,\xi _i)&\le 0,\ i\in I,\ k=1,\ldots ,K \\ h_j(x)&\le 0,\ j\in J \end{aligned} \right\} }. \end{aligned}$$

Then Z coincides locally around \(\bar{x}\) with

$$\begin{aligned} \bigcup _{I\in \mathscr {I}(\bar{x})} Z_I := \bigcup _{I\in \mathscr {I}(\bar{x})} {\left\{ x:\begin{aligned} g_k(x,\xi _i)&\le 0,\ i\in I_0(\bar{x})\cap I,\ k \in K_{0}^i(\bar{x})\\ h_j(x)&\le 0,\ j\in J_0 \end{aligned} \right\} }, \end{aligned}$$
(24)

which means that

$$\begin{aligned} \hat{N}_Z(\bar{x}) = (T_Z(\bar{x}))^* = \left( \bigcup _{I\in \mathscr {I}(\bar{x})}T_{Z_I}(\bar{x})\right) ^* = \bigcap _{I\in \mathscr {I}(\bar{x})}\hat{N}_{Z_I}(\bar{x}). \end{aligned}$$

By (Rockafellar and Wets 1998, Theorem 6.12) we obtain that \(0\in \nabla f(\bar{x})+\hat{N}_Z(\bar{x})\) is a necessary optimality condition for chance constrained problem (2). To obtain the first statement, it suffices to realize that \(Z_I\) can be due to (24) written as \(F(x)\le 0\) for some function F combining \(g_k\) and \(h_j\) and to use chain rule (Rockafellar and Wets 1998, Theorem 6.14).

The proof of the second part goes in a similar way. Due to (Rockafellar and Wets 1998, Theorem 6.12), the necessary optimality conditions for problem (4) read

$$\begin{aligned} 0\in \begin{pmatrix}\nabla f(\bar{x})\\ 0\end{pmatrix}+ \hat{N}_Z(\bar{x},\bar{y}) \end{aligned}$$
(25)

where Z is the feasible set of problem (4). For the computation of the normal cone, realize first that Z locally around \((\bar{x},\bar{y})\) coincides with the union of \(Z_I:=Z_I^x\times Z_I^y\) with respect to all \(I\subset I_{00}(\bar{x},\bar{y})\), where

$$\begin{aligned} \begin{aligned} Z_I^x&:= {\left\{ x:\begin{aligned}&g_k(x,\xi _i)\le 0,\ i\in I\cup I_{0+}(\bar{x},\bar{y}),\ k \in K_0^i(\bar{x})\\&h_j(x)\le 0,\ j\in J_0 \\ \end{aligned} \right\} }\\ Z_I^y&:= {\left\{ y:\begin{aligned}&p^\top y\ge 1-\varepsilon \\&y_i\in [0,1],\ i\in I\cup I_{0+}(\bar{x},\bar{y})\cup {\left\{ i:\textstyle {\max _k}\ g_k(\bar{x},\xi _i)<0\right\} } \\&y_i=0,\ i\in (I_{00}(\bar{x},\bar{y})\setminus I)\cup {\left\{ i:\textstyle {\max _k}\ g_k(\bar{x},\xi _i)>0\right\} } \end{aligned} \right\} }. \end{aligned} \end{aligned}$$

As before, we have

$$\begin{aligned} \hat{N}_Z(\bar{x},\bar{y})&= \bigcap _{I\subset I_{00}(\bar{x},\bar{y})}\hat{N}_{Z_I}(\bar{x},\bar{y}) = \bigcap _{I\subset I_{00}(\bar{x},\bar{y})}\hat{N}_{Z_I^x}(\bar{x})\times \bigcap _{I\subset I_{00}(\bar{x},\bar{y})}\hat{N}_{Z_I^y}(\bar{y}). \end{aligned}$$

Since zero always belongs to a normal cone, the optimality condition (25) is equivalent to

$$\begin{aligned} 0\in \nabla f(\bar{x}) + \bigcap _{I\subset I_{00}(\bar{x},\bar{y})}\hat{N}_{Z_I^x}(\bar{x}). \end{aligned}$$
(26)

To finish the proof, it suffices to realize that the intersection in (26) is attained for \(I=\emptyset \) and to use either (Ioffe and Outrata 2008, Proposition 3.4) (if the first part of Assumption 1 holds true) or (Rockafellar and Wets 1998, Theorem 6.14) (if the second part of Assumption 1 holds true). \(\square \)

The proof of Theorem 3 is more complicated. For notational simplicity, we consider only the case of \(J_0(\bar{x})=\emptyset \). First, we write down the stationarity conditions of (8), then show two preliminary lemmas and only then proof the theorem itself.

The necessary optimality conditions for problem (8) at a point \((\bar{x}^t,\bar{y}^t)\) read as follows: there exist multipliers \(\alpha ^t\in {\mathbb {R}}\), \(\beta ^t\in {\mathbb {R}}^S\) and \(\gamma ^t\in {\mathbb {R}}^{SK}\) such that the first-order optimality conditions

$$\begin{aligned} 0&= \nabla f(\bar{x}^t) - \sum _{i=1}^S\sum _{k=1}^K \gamma _{ik}^t \phi _t'(g_k(\bar{x}^t,\xi _i))\nabla _x g_k(\bar{x}^t,\xi _i), \end{aligned}$$
(27a)
$$\begin{aligned} 0&= -\alpha ^t p_i + \beta _i^t + \sum _{k=1}^K \gamma _{ik}^t,\ i=1,\ldots ,S \end{aligned}$$
(27b)

and the complementarity conditions

$$\begin{aligned} \alpha ^t(1-\varepsilon -p^\top \bar{y}^t)&= 0, \end{aligned}$$
(28a)
$$\begin{aligned} \beta _i^t&{\left\{ \begin{array}{ll} \ge 0 &{}\text {if }\bar{y}_i^t = 1,\\ = 0 &{}\text {if }0<\bar{y}_i^t<1,\\ \le 0 &{}\text {if }\bar{y}_i^t = 0,\end{array}\right. }\end{aligned}$$
(28b)
$$\begin{aligned} \gamma _{ik}^t(\bar{y}_i^t-\phi _t(g_k(\bar{x}^t,\xi _i)))&= 0 \end{aligned}$$
(28c)

are satisfied. Moreover, the sign restrictions \(\alpha ^t\ge 0\) and \(\gamma _{ik}^t\ge 0\) hold true.

Lemma 2

Assume that \((\bar{x}^t,\bar{y}^t)\) is a stationary point of problem (8). Then the following assertions hold true:

  1. 1.

    \(g_k(\bar{x}^t,\xi _i)<0 \implies \gamma _{ik}^t=0;\)

  2. 2.

    \(\alpha ^t = 0 \implies \beta _i^t=\gamma _{ik}^t=0\) for all i and k;

Proof

If \(g_k(\bar{x}^t,\xi _i)<0\), then from the definition of \(\phi _t\) we have \(\phi _t(g_k(\bar{x}^t,\xi _i))>1\). Since \(\bar{y}_i^t\in [0,1]\), this directly implies \(\gamma _{ik}^t=0\) due to (28c). For the second part, consider the case of \(\alpha ^t = 0\). Then due to (27b) we have \(\beta _i^t + \sum _{k=1}^K \gamma _{ik}^t=0\) for all i. Assume that \(\beta _i^t <0\). Then there exists some k such that \(\gamma _{ik}^t>0\). From (28b) we further get \(\bar{y}_i^t=0\), and thus \(\bar{y}_i^t-\phi _t(g_k(\bar{x}^t,\xi _i))<0\). But this is a contradiction with (28c) and thus we have \(\beta _i^t\ge 0\). But this together with \(\gamma _{ik}^t\ge 0\) and \(\beta _i+\gamma _{ik}^t=0\) implies the second statement. \(\square \)

In the following text, by \(\bar{y}_i^t\searrow 0\) we understand that the sequence \(\bar{y}_i^t\) is positive and converges monotonically to 0.

Lemma 3

If for all t we have \(p^\top \bar{y}^t = 1-\varepsilon \) and for some i and k we have \(\phi _t(g_k(\bar{x}^t,\xi _i)) = \bar{y}_i^t\searrow 0\), then there exists a subsequence in t such that \(\gamma _{ik}^t=0\) for all t or such that there exists index j such that

$$\begin{aligned} \frac{ -\gamma _{ik}^t \phi _t'(g_k(\bar{x}^t,\xi _i))}{ -\sum _{\tilde{k}=1}^K \gamma _{j\tilde{k}}^t \phi _t'(g_{\tilde{k}}(\bar{x}^t,\xi _j))} \rightarrow 0. \end{aligned}$$
(29)

Proof

Since \(\bar{y}_i^t\searrow 0\) and since \(p^\top \bar{y}=1-\varepsilon \), there exists index j, and possibly a subsequence in t, such that \(\bar{y}_j^t\) is strictly increasing. This implies that \(0<\bar{y}_j^t<1\) and \(\beta _i^t=\beta _j^t=0\) for all t by (28b). If \(\gamma _{ik}^t=0\), then the proof is finished. In the opposite case of \(\gamma _{ik}^t>0\), we realize that \(\alpha ^t>0\) due to Lemma 2. Since \(\gamma _{ik}^t\ge 0\) and \(\phi _t'(g_k(\bar{x}^t,\xi _i))<0\), we deduce

$$\begin{aligned} \begin{aligned} 0\le \frac{- \gamma _{ik}^t \phi _t'(g_k(\bar{x}^t,\xi _i))}{- \sum _{\tilde{k}=1}^K \gamma _{j\tilde{k}}^t \phi _t'(g_{\tilde{k}}(\bar{x}^t,\xi _j))}&\le \frac{- \left( \sum _{\tilde{k}=1}^K \gamma _{i\tilde{k}}^t \right) \phi _t'(g_k(\bar{x}^t,\xi _i)) }{- \sum _{\tilde{k}=1}^K \gamma _{j\tilde{k}}^t \phi _t'(g_{\tilde{k}}(\bar{x}^t,\xi _j))}\\&\le \frac{- \sum _{\tilde{k}=1}^K \gamma _{i\tilde{k}}^t }{- \sum _{\tilde{k}=1}^K \gamma _{j\tilde{k}}^t} \frac{ \phi _t'(g_k(\bar{x}^t,\xi _i))}{\phi _t'(g_{\hat{k}^t}(\bar{x}^t,\xi _j))} \\&= \frac{- \alpha ^t p_i}{- \alpha ^t p_j} \frac{ \phi _t'(g_k(\bar{x}^t,\xi _i))}{\phi _t'(g_{\hat{k}^t}(\bar{x}^t,\xi _j))} \rightarrow 0 \end{aligned} \end{aligned}$$
(30)

where \(\hat{k}^t := \arg \min _{\tilde{k}} -\phi _t'(g_{\tilde{k}}(\bar{x}^t,\xi _j)\), the last equality follows from (27b) as well as \(\beta _i^t=\beta _j^t=0\) and the convergence follows from assumption (9d), for which we realize that \(\phi _t(g_{\tilde{k}^t}(\bar{x}^t,\xi _j)) \ge \bar{y}_j^t\), the fact that \(\bar{y}_j^t\) is a strictly increasing sequence and the assumed convergence \(\phi _t(g_k(\bar{x}^t,\xi _i)) = \bar{y}_i^t\searrow 0\). \(\square \)

Proof

(Theorem 3) We will show first that \((\bar{x},\bar{y})\) is a feasible point of (4). Due to continuity, it is sufficient to show that \(g_k(\bar{x},\xi _i)\bar{y}_i\le 0\). Since this relation is obvious whenever \(g_k(\bar{x},\xi _i)\le 0\) for all k, we consider scenario i with \(g_k(\bar{x},\xi _i)>0\) for some k. But then \(g_k(\bar{x}^t,\xi _i)>0\) for sufficiently large t and thus \(0\le \bar{y}_i^t\le \phi _t(g_k(\bar{x}^t,\xi _i))\). But since \(g_k(\bar{x}^t,\xi _i)\rightarrow g_k(\bar{x},\xi _i)>0\), assumption (9c) implies that \(\bar{y}_i=0\), and thus \((\bar{x},\bar{y})\) is a feasible point of problem (4).

Define now

$$\begin{aligned} \lambda _{ik}^t := - \gamma _{ik}^t \phi _t'(g_k(\bar{x}^t,\xi _i)) \ge 0, \end{aligned}$$

where the nonnegativity follows from the property that \(\phi _t\) is decreasing. If \(g_k(\bar{x},\xi _i) <0\), then for all sufficiently large t we have \(g_k(\bar{x}^t,\xi _i) <0\) and due to Lemma 2 we deduce \(\gamma _{ik}^t=0\) and subsequently \(\lambda _{ik}^t=0\). Then for sufficiently large t, optimality condition (27a) reads

(31)

Here we can omit pairs of indices (ik) with \(g_k(\bar{x},\xi _i) <0\) due to the discussion above.

We claim now that \(\sum _{k=1}^{K}\lambda _{ik}^t\) is uniformly bounded in i and t. If this is not the case, then we have

$$\begin{aligned} \lambda _{max}^t:= \max _{i=1,\ldots ,S} \sum _{k=1}^{K} \lambda _{ik}^t \rightarrow \infty . \end{aligned}$$
(32)

Then dividing equation (31) by \(\lambda _{max}^t\) yields

(33)

When taking the limit with respect to \(t\rightarrow \infty \), the first term vanishes. Consider now the third term. If \(p^\top \bar{y}^t > 1-\varepsilon \), from Lemma 2 we have \(\lambda _{ik}^t=0\) for all i and k. Assume thus that \(p^\top \bar{y}^t = 1-\varepsilon \) for all t. If \(\phi _t(g_k(\bar{x}^t,\xi _i)) > \bar{y}_i^t\), then from (28c) we have \(\lambda _{ik}^t=0\). In the opposite case, we may use Lemma 3 to obtain again that \(\lambda _{ik}^t=0\) or there exists j such that \(\frac{\lambda _{ik}^t}{\sum _{\tilde{k}=1}^{K} \lambda _{j\tilde{k}}^t}\rightarrow 0\). But this implies that the last term in (33) vanishes as well. This means that

$$\begin{aligned} 0 = \sum _{{\left\{ (i,k):g_k(\bar{x},\xi _i)=0,\, \bar{y}_i>0\right\} }}\lim _{t\rightarrow \infty }\frac{\lambda _{ik}^t}{{\lambda _{max}^t}} \nabla _x g_k(\bar{x}^t,\xi _i). \end{aligned}$$

Since \(\frac{\lambda _{ik}^t}{\lambda _{max}^t}\in [0,1]\) and the numerators sum to \(\lambda _{max}^t\) due to (32) and the discussion above, at least one of these fractions converges to a positive number. However, the existence of such positive limit contradicts Assumption 1 and thus \(\sum _{k=1}^{K} \lambda _{ik}^t\) is indeed bounded. Since it is a sum of nonnegative elements, these elements \(\lambda _{ik}^t\) are uniformly bounded in i, k and t.

This means that we may pass to a converging subsequence, say \(\lambda _{ik}^t\rightarrow \lambda _{ik}\). Since \(\lambda _{ik}^t\ge 0\) for all t, the same property holds for \(\lambda _{ik}\). In the light of (31), to finish the proof it suffices to show that \(\lambda _{ik}=0\) for all pairs (ik) such that \(g_k(\bar{x},\xi _i)\ge 0\) and \(\bar{y}_i=0\). But this may be shown as in the previous paragraph via applying Lemmas 2 and 3. Thus \((\bar{x},\bar{y})\) is indeed a stationary point of problem (4). \(\square \)

C Cut reduction

To propose a method for cut reduction, we start with the following technical lemma.

Lemma 4

Consider two points \(\hat{x}^1,\hat{x}^2\) such that for all i we have

$$\begin{aligned} \max _k g_k(\hat{x}^2,\xi _i)> 0 \implies \max _k g_k(\hat{x}^1,\xi _i) > 0,\ \arg \max _k g_k(\hat{x}^1,\xi _i) = \arg \max _k g_k(\hat{x}^2,\xi _i). \end{aligned}$$
(34)

Define now for \(j=1,2\) mappings \(\mathscr {K}^j:\{1,\ldots ,S\}\rightarrow \{0,1,\ldots ,K\}\) and \(v^j:{\mathbb {R}}^n\rightarrow {\mathbb {R}}\) by

$$\begin{aligned} \begin{aligned} \mathscr {K}^j(i)&:= {\left\{ \begin{array}{ll} 0 &{} \text {if } \max _k g_k(\hat{x}^j,\xi _i)\le 0, \\ \arg \max _k g_k(\hat{x}^j,\xi _i) &{} \text {otherwise,}\end{array}\right. } \\ v^j(x)&:= \sum _{\{i|\ \mathscr {K}^j(i) > 0\}} p_i\phi _t(g_{\mathscr {K}^j(i)}(x,\xi _i)) - 1 + \varepsilon + \sum _{\{i|\ \mathscr {K}^j(i) = 0\}} p_i. \end{aligned} \end{aligned}$$

Then for any x we have

$$\begin{aligned} v^1(x) \ge 0 \implies v^2(x) \ge -(\sup \phi _t(\cdot ) - 1). \end{aligned}$$

Proof

Due to (34) we have \(\{i|\ \mathscr {K}^2(i)>0\}\subset \{i|\ \mathscr {K}^1(i)>0\}\) and subsequently

$$\begin{aligned} \begin{aligned} v^2(x)&= v^1(x) - \sum _{\{i|\ \mathscr {K}^1(i)>0,\ \mathscr {K}^2(i)=0\}} p_i\phi _t(g_{\mathscr {K}^1(i)}(x,\xi _i)) + \sum _{\{i|\ \mathscr {K}^1(i)>0,\ \mathscr {K}^2(i)=0\}} p_i \\&\ge v^1(x) - \sum _{\{i|\ \mathscr {K}^1(i)>0,\ \mathscr {K}^2(i)=0\}} p_i(\sup \phi _t(\cdot ) - 1) \ge v^1(x) - (\sup \phi _t(\cdot ) - 1), \end{aligned} \end{aligned}$$

which finishes the proof. \(\square \)

Note that the cuts generated by (17) equal to \(\{x|\ v^j(x)\ge 0\}\). This lemma states that if (34) holds true, then the cut generated by \(\hat{x}^1\) is tighter than the one generated by \(\hat{x}^2\) up to a margin \(\sup \phi _t(\cdot ) - 1\). Since \(\phi _t(0)=1\), this margin may be made arbitrarily small by a proper choice of \(\phi _t\), the best cut in (17) is generated whenever \(\max _k g_k(\hat{x},\xi _i)>0\) for all i.

This also gives rise to the following heuristic cut reduction technique. If we add a new cut generated by some \(\hat{x}^1\), we remove all the previously included cuts, generated by \(\hat{x}^2\), for which (34) is satisfied. Alternatively, we remove all previous cuts, for which (34) is violated only for a small number of scenarios.

Finally, the next lemma states that the added cuts in (17) are optimal as they cannot be linear combination of each other. Recall that a direction d is an extremal direction of a cone C if there do not exist directions \(d_1,d_2\in C\) different from d and a scalar \(\kappa \in (0,1)\) such that \(d=\kappa d_1+(1-\kappa )d_2\).

Lemma 5

The direction defined in (15) is an extremal direction of the feasible set of (13).

Proof

Denote the feasible set of problem (13) by Z. For contradiction assume that \((\hat{u},\hat{v},\hat{w})\) is not an extremal direction of Z. Then there are some \((u^1, v^1, w^1)\in Z\) and \((u^2, v^2, w^2)\in Z\) different from \((\hat{u}, \hat{v}, \hat{w})\) and some \(\kappa \in (0,1)\) such that

$$\begin{aligned} (\hat{u}, \hat{v}, \hat{w}) = \kappa (u^1, v^1, w^1) + (1-\kappa )(u^2, v^2, w^2). \end{aligned}$$
(35)

Note now that the role of \(\hat{u}_{i\hat{k}_i}\) and \(\hat{w}_i\) is symmetric in (15) and thus, it suffices to consider only the first equation in (15). Define

$$\begin{aligned} Z_1 :=\{(v,w_1)|\ vp_1+w_1\le 0,\ v\ge 0,\ w_1\le 0\} \end{aligned}$$

and observe that \((\hat{v},\hat{w}_1), (v^1, w_1^1), (v^2, w_1^2)\in Z_1\). But this due to (35) means that \((\hat{v},\hat{w}_1)\) is not an extremal direction of \(Z_1\). But since the extremal directions of \(Z_1\) amount to \((0,-p_1)\) and \((1,-p_1)\), this is a contradiction with (15). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adam, L., Branda, M., Heitsch, H. et al. Solving joint chance constrained problems using regularization and Benders’ decomposition. Ann Oper Res 292, 683–709 (2020). https://doi.org/10.1007/s10479-018-3091-9

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-018-3091-9

Keywords

Mathematics Subject Classification

Navigation