Skip to main content
Log in

Robust multicriteria risk-averse stochastic programming models

  • Original Paper
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

In this paper, we study risk-averse models for multicriteria optimization problems under uncertainty. We use a weighted sum-based scalarization and take a robust approach by considering a set of scalarization vectors to address the ambiguity and inconsistency in the relative weights of each criterion. We model the risk aversion of the decision makers via the concept of multivariate conditional value-at-risk (CVaR). First, we introduce a model that optimizes the worst-case multivariate CVaR and show that its optimal solution lies on a particular type of stochastic efficient frontier. To solve this model, we develop a finitely convergent delayed cut generation algorithm for finite probability spaces. We also show that the proposed model can be reformulated as a compact linear program under certain assumptions. In addition, for the cut generation problem, which is in general a mixed-integer program, we give a stronger formulation than the existing ones for the equiprobable case. Next, we observe that similar polyhedral enhancements are also useful for a related class of multivariate CVaR-constrained optimization problems that has attracted attention recently. In our computational study, we use a budget allocation application to benchmark our proposed maximin type risk-averse model against its risk-neutral counterpart and a related multivariate CVaR-constrained model. Finally, we illustrate the effectiveness of the proposed solution methods for both classes of models.

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

  • Artzner, P., Delbaen, F., Eber, J., & Heath, D. (1999). Coherent measures of risk. Mathematical Finance, 9(3), 203–228.

    Article  Google Scholar 

  • Balibek, E., & Köksalan, M. (2010). A multi-objective multi-period stochastic programming model for public debt management. European Journal of Operational Research, 205(1), 205–217.

    Article  Google Scholar 

  • Ben Abdelaziz, F. (2012). Solution approaches for the multiobjective stochastic programming. European Journal of Operational Research, 216(1), 1–16.

    Article  Google Scholar 

  • Ben Abdelaziz, F., Lang, P., & Nadeau, R. (1995). Distributional efficiency in multiobjective stochastic linear programming. European Journal of Operational Research, 85(2), 399–415.

    Article  Google Scholar 

  • Ben-Tal, A., Ghaoui, L., & Nemirovski, A. (2009). Robust optimization. Princeton, NJ: Princeton University Press.

    Book  Google Scholar 

  • Bertsimas, D., Brown, D. B., & Caramanis, C. (2011). Theory and applications of robust optimization. SIAM Review, 53(3), 464–501.

    Article  Google Scholar 

  • Borcherding, K., Eppel, T., & von Winterfeldt, D. (1991). Comparison of weighting judgments in multiattribute utility measurement. Management Science, 37(12), 1603–1619.

    Article  Google Scholar 

  • Burgert, C., & Rüschendorf, L. (2006). Consistent risk measures for portfolio vectors. Insurance: Mathematics and Economics, 38(2), 289–297.

    Google Scholar 

  • Dentcheva, D., & Ruszczyński, A. (2006). Portfolio optimization with stochastic dominance constraints. Journal of Banking and Finance, 30(2), 433–451.

    Article  Google Scholar 

  • Dentcheva, D., & Ruszczyński, A. (2009). Optimization with multivariate stochastic dominance constraints. Mathematical Programming, 117(1), 111–127.

    Article  Google Scholar 

  • Dentcheva, D., & Wolfhagen, E. (2013). Optimization with multivariate dominance constraints. In G. Deodatis, B. Ellingwood, & D. Frangopol (Eds.), Safety, reliability, risk and life-cycle performance of structures and infrastructures. Boca Raton: CRC Press LLC.

    Google Scholar 

  • Ehrgott, M. (2005). Multicriteria optimization. Berlin: Springer.

    Google Scholar 

  • Ehrgott, M., Ide, J., & Schöbel, A. (2014). Minmax robustness for multi-objective optimization problems. European Journal of Operational Research, 239(1), 17–31.

    Article  Google Scholar 

  • Ekeland, I., & Schachermayer, W. (2011). Law invariant risk measures on \(L^\infty (\mathbb{R}^d)\). Statistics and Risk Modeling with Applications in Finance and Insurance, 28(3), 195–225.

    Google Scholar 

  • Gupte, A., Ahmed, S., Dey, S. S., & Cheon, M. S. (2017). Relaxations and discretizations for the pooling problem. Journal of Global Optimization, 67(3), 631–669.

    Article  Google Scholar 

  • Gutjahr, W. J., & Pichler, A. (2016). Stochastic multi-objective optimization: A survey on non-scalarizing methods. Annals of Operations Research, 236(2), 475–499.

    Article  Google Scholar 

  • Hamel, A. H., Rudloff, B., & Yankova, M. (2013). Set-valued average value at risk and its computation. Mathematics and Financial Economics, 7(2), 229–246.

    Article  Google Scholar 

  • Homem-de-Mello, T., & Mehrotra, S. (2009). A cutting surface method for uncertain linear programs with linear stochastic dominance constraints. SIAM Journal on Optimization, 20(3), 1250–1273.

    Article  Google Scholar 

  • Hu, J., Homem-de-Mello, T., & Mehrotra, S. (2011). Risk-adjusted budget allocation models with application in homeland security. IIE Transactions, 43(12), 819–839.

    Article  Google Scholar 

  • Hu, J., Homem-de Mello, T., & Mehrotra, S. (2012). Sample average approximation of stochastic dominance constrained programs. Mathematical Programming, 133(1–2), 171–201.

    Article  Google Scholar 

  • Hu, J., & Mehrotra, S. (2012). Robust and stochastically weighted multiobjective optimization models and reformulations. Operations Research, 60(4), 936–953.

    Article  Google Scholar 

  • Jouini, E., Meddeb, M., & Touzi, N. (2004). Vector-valued coherent risk measures. Finance and Stochastics, 8(4), 531–552.

    Article  Google Scholar 

  • Köksalan, M., & Şakar, C. T. (2016). An interactive approach to stochastic programming-based portfolio optimization. Annals of Operations Research, 245(1), 47–66.

    Article  Google Scholar 

  • Küçükyavuz, S., & Noyan, N. (2016). Cut generation for optimization problems with multivariate risk constraints. Mathematical Programming, 159(1), 165–199.

    Article  Google Scholar 

  • Lehmann, E. (1955). Ordered families of distributions. Annals of Mathematical Statistics, 26(3), 399–419.

    Article  Google Scholar 

  • Levy, H. (1992). Stochastic dominance and expected utility: Survey and analysis. Management Science, 38(4), 555–593.

    Article  Google Scholar 

  • Mann, H., & Whitney, D. (1947). On a test of whether one of two random variables is stochastically larger than the other. Annals of Mathematical Statistics, 18(1), 50–60.

    Article  Google Scholar 

  • McCormick, G. (1976). Computability of global solutions to factorable nonconvex programs: Part I—Convex underestimating problems. Mathematical Programming, 10(1), 147–175.

    Article  Google Scholar 

  • Müller, A., & Stoyan, D. (2002). Comparison methods for stochastic models and risks. Chichester: Wiley.

    Google Scholar 

  • Noyan, N. (2012). Risk-averse two-stage stochastic programming with an application to disaster management. Computers and Operations Research, 39(3), 541–559.

    Article  Google Scholar 

  • Noyan, N., Balcik, B., & Atakan, S. (2016). A stochastic optimization model for designing last mile relief networks. Transportation Science, 50(3), 1092–1113.

    Article  Google Scholar 

  • Noyan, N., & Rudolf, G. (2013). Optimization with multivariate conditional value-at-risk-constraints. Operations Research, 61(4), 990–1013.

    Article  Google Scholar 

  • Ogryczak, W. (2010). On robust solutions to multi-objective linear programs. In T. Trzaskalik & T. Wachowicz (Eds.), Multiple Criteria Decision Making ’09 (pp. 197–212).

  • Ogryczak, W., & Ruszczyński, A. (2001). On consistency of stochastic dominance and mean-semideviation models. Mathematical Programming, 89(2), 217–232.

    Article  Google Scholar 

  • Pflug, G. C., & Römisch, W. (2007). Modelling, managing and measuring risk. Singapore: World Scientific Publishing.

    Book  Google Scholar 

  • Rockafellar, R., & Uryasev, S. (2000). Optimization of conditional value-at-risk. The Journal of Risk, 2(3), 21–41.

    Article  Google Scholar 

  • Rüschendorf, L. (2013). Risk measures for portfolio vectors. Mathematical risk analysis, springer series in operations research and financial engineering (pp. 167–188). Berlin: Springer.

    Google Scholar 

  • Saaty, T. (2000). Decision making for leaders; the analytical hierarchy process for decisions in a complex world. Pittsburgh: RWS Publications.

    Google Scholar 

  • Schoemaker, P. J. H., & Waid, C. C. (1982). An experimental comparison of different approaches to determining weights in additive utility models. Management Science, 28(2), 182–196.

    Article  Google Scholar 

  • Shaked, M., & Shanthikumar, J. G. (1994). Stochastic orders and their applications. Boston: Associated Press.

    Google Scholar 

  • Sherali, H. D., & Adams, W. P. (1994). A hierarchy of relaxations and convex hull representations for mixed-integer zero-one programming problems. Discrete Applied Mathematics, 52(1), 83–106.

    Article  Google Scholar 

  • Sherali, H. D., Adams, W. P., & Driscoll, P. J. (1998). Exploiting special structures in constructing a hierarchy of relaxations for 0–1 mixed integer problems. Operations Research, 46(3), 396–405.

    Article  Google Scholar 

  • Steuer, R. E. (1986). Multiple criteria optimization: Theory, computation, and application. New York: Wiley.

    Google Scholar 

  • von Winterfeldt, D., & Edwards, W. (1986). Decision analysis and behavioral research. Cambridge: Cambridge University Press.

    Google Scholar 

  • Willis, H. H., Morral, A. R., Kelly, T. K., & Medby, J. J. (2005). Estimating terrorism risk. Technical report, The RAND Corporation, Santa Monica, CA.

  • Wozabal, D. (2014). Robustifying convex risk measures for linear portfolios: A nonparametric approach. Operations Research, 62(6), 1302–1315.

    Article  Google Scholar 

  • Zhu, S., & Fukushima, M. (2009). Worst-case conditional value-at-risk with application to robust portfolio management. Operations Research, 57(5), 1155–1168.

    Article  Google Scholar 

Download references

Acknowledgements

We thank the two referees and the associate editor for their valuable comments that improved the presentation. Simge Küçükyavuz and Xiao Liu are supported, in part, by National Science Foundation Grants 1732364 and 1733001. Nilay Noyan acknowledges the support from Bilim Akademisi—The Science Academy, Turkey, under the BAGEP Program.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nilay Noyan.

Appendices

Appendix A: Stochastic dominance

In this section, we review the well-known stochastic dominance relations, which are essential for the stochastic Pareto optimality definitions presented in Sect. 2.

The stochastic dominance relations are fundamental concepts in comparing random variables (Mann and Whitney 1947; Lehmann 1955) and have been widely used in economics and finance (see, e.g., Levy 1992). Different from the approaches based on risk measures, in a stochastic dominance based approach, the random variables are compared by a point-wise comparison of some performance functions (constructed from their distribution functions when the order is greater than zero). We note that the lower order dominance relations (\(k=0,1,\) and 2) are the most common ones (referred to as ZSD, FSD, and SSD, respectively). We provide the formal definitions below and refer the reader to Müller and Stoyan (2002) and Shaked and Shanthikumar (1994) for further details.

  • We say that a random variable X dominates another random variable Y in the zeroth order if \(X\ge Y\) everywhere, i.e., \(X(\omega )\ge Y(\omega )\) for all \(\omega \in \varOmega \).

  • An integrable random variable X dominates another integrable Y in the first order (or X is stochastically larger than Y) if \(F_1(X,\eta ):=\mathbbm {P}(X \le \eta ) \le F_1(Y,\eta ):=\mathbbm {P}(Y \le \eta )\) for all \(\eta \in \mathbbm {R}\).

  • For \(k\ge 2\) we say that a k-integrable random variable X (i.e., \(\in \mathcal {L}^k\)) dominates another k-integrable random variable Y in the kth order if \(F_k(X,\eta ) \le F_k(Y,\eta )\) for all \(\eta \in \mathbbm {R}\), where \(F_k(X,\eta )=\int _{-\infty }^\eta F_{k-1}(X,t)~\mathrm {d}t\) for all \(\eta \in \mathbbm {R}\).

  • For \(k=0\), if \(X(\omega )> Y(\omega )\) for all \(\omega \in \varOmega \), we will refer to the relation as “strong ZSD” and denote it by \(X\succ _{(0)}Y\). For \(k \ge 1\), if all the inequalities \(F_k(X,\eta ) \le F_k(Y,\eta )\) are strict, then we refer to the relation as “strong kSD” and denote it by \(X\succ _{(k)}Y\). We remark that the notion of “strong kSD” is not analogous to the notion of strict kSD, which requires that at least one of the inequalities defining the dominance relation is strict.

Appendix B: A class of facets of \(\textit{conv}(\mathcal S)\)

Recall that

$$\begin{aligned} \mathcal S :=&\bigg \{(\varvec{\gamma }, \mathbf {c}, \varvec{\beta }, \eta , \mathbf {w}) \in \mathbbm {R}_+^{n \times d}\times \mathbbm {R}_+^d \times \{0,1\}^n \times \mathbbm {R}\times \mathbbm {R}^m_+ \; |&\;\varvec{\gamma }= \varvec{\beta }\mathbf {c}^\top , ~\sum _{j \in [d]} c_j = 1,~\\&\sum _{i \in [n]} \beta _i = k,\mathbf {c}^\top \mathbf {y}_l \ge \eta - w_l,~\forall ~l \in [m]\bigg \}. \end{aligned}$$

Before we study the facets of \(\textit{conv}(\mathcal S)\), we first need to establish its dimension.

Proposition B.1

\(\textit{Conv}(\mathcal S)\) is a polyhedron with dimension \(n + d + m - 1\).

Proof

Note that \(\textit{conv}(S)\) is a polyhedron, because \(\varvec{\beta }\in \{0,1\}^n\). Next, we show that the dimension of \(\textit{conv}(\mathcal S) \) is \(n + d + m - 1\). Clearly, in the original constraints defining \(\mathcal S\), there are two linearly independent equalities: \(\sum _{j \in [d]} c_j = 1\), \(\sum _{i \in [n]} \beta _i = k.\) In addition, there are nd equalities: \(\gamma _{ij} = c_j \beta _i\), for all \(i \in [n]\) and \(j \in [d]\). Hence, \(\textit{dim} (\textit{conv}(\mathcal S)) \le n + m + d - 1.\)

Consider the following set of points:

$$\begin{aligned}&\left( \mathbf {u}_v\mathbf {e}_1^\top , \mathbf {e}_1, \mathbf {u}_v, 0, 0\right)&\forall ~v \in [n], \\&\left( \mathbf {u}_1\mathbf {e}_j^\top , \mathbf {e}_j, \mathbf {u}_1, 0, 0\right)&\forall ~j \in [d] {\setminus } \{1\}, \\&\left( \mathbf {u}_1\mathbf {e}_1^\top , \mathbf {e}_1, \mathbf {u}_1, 0, e_l\right)&\forall ~l \in [m], \\&\left( \mathbf {u}_1\mathbf {e}_1^\top , \mathbf {e}_1, \mathbf {u}_1, -1, 0\right) , \end{aligned}$$

where \(\mathbf {u}_v\), for all \(v \in [n]\) are any affinely independent vectors with k elements equal to 1 and the remaining elements equal to 0. These vectors exist because the dimension of the following system:

$$\begin{aligned}&\varvec{\beta }\in \{0,1\}^n, \quad \sum _{i \in [n] } \beta _i = k, \end{aligned}$$
(38)

is \(n-1\). Clearly, this set of points is feasible and affinely independent. In addition, the cardinality of this set is \(n + m + d \). Hence, \(dim (conv( \mathcal S) ) \ge n + m + d - 1\), which completes the proof.

Proposition B.2

For any \(i \in [n]\), \(s \in [m]\), and \(t \in [m]{\setminus } \{s\}\), inequality (25) is facet-defining for \(conv(\mathcal S)\) if and only if \(s \in [m], t\in [m]{\setminus } \{s\}\) are such that \(y_{sj}< y_{tj}\) and \(y_{si} > y_{ti}\) for some \(i,j\in [d]\).

Proof

To show the necessity, we first note that if there exists a pair \(s \in [m], t\in [m]{\setminus } \{s\}\) such that \(y_{sj}\ge y_{tj}\) or \(y_{sj}\le y_{tj}\) for all \(j\in [d]\), in other words, when the realizations under a scenario are dominated by the realizations under another scenario, then the corresponding inequality (25) is dominated. To see this, suppose that \(y_{sj}\le y_{tj}\) for all \(j\in [d]\) for some pair \(\forall ~s \in [m], \forall ~t\in [m]{\setminus } \{s\}\). Then the corresponding inequality (25) is dominated by the original inequality \(\mathbf {c}^\top \mathbf {y}_s\ge \eta -w_s\), because the coefficients of \(\gamma _{ij}\) are \(y_{tj}-y_{sj}\ge 0\), and \(\gamma _{ij},w_t\ge 0\). Now consider the case that \(y_{sj}\ge y_{tj}\) for all \(j\in [d]\) for some pair \(\forall ~s \in [m], \forall ~t\in [m]{\setminus } \{s\}\). Then the corresponding inequality (25) is dominated by the original inequality \(\mathbf {c}^\top \mathbf {y}_t\ge \eta -w_t\). To see this, observe that we can rewrite inequality (25) for this choice of s and t as, \(\mathbf {c}^\top \mathbf {y}_t + \sum _{j\in [d]}(y_{sj}-y_{tj})(c_j-\gamma _{ij}) \ge \eta -w_t-w_s\). It is now easy to see that the inequality is dominated, because \(y_{sj}-y_{tj}\ge 0\), \(c_j\ge \gamma _{ij}\) and \(w_s\ge 0\).

To show sufficiency, we need to show that for any given \(i \in [n]\), \(s \in [m]\), and \(t \in [m]{\setminus } \{s\}\), there are \(n + m + d - 1\) affinely independent points that satisfy (25) at equality. From the necessity condition, we only need to consider the cases for which there exists an index \(j_1 \in [d]\), such that \(y_{sj_1} < y_{tj_1} \), and there exists an index \(j_2 \in [d]\), such that \(y_{sj_2} > y_{tj_2} \). In order to simplify the notation, and without loss of generality, throughout the rest of the proof, we let \(j_1 = 1\), and \(j_2 = 2\), or equivalently, \(y_{s1} < y_{t1}\), and \(y_{s2} > y_{t2}\).

First, we construct a set of points:

$$\begin{aligned}&\mathbf PT ^1_v= \left( \mathbf {u}_v \tilde{\mathbf {e}}_v^\top , {\tilde{\mathbf {e}}}_v, \mathbf {u}_v, \rho ^1_v ,\varvec{\xi }^1_v\right) ,&\forall ~v \in [n], \end{aligned}$$
(39)

where if \(u_{vi} = 0\), then \({\tilde{\mathbf {e}}}_v = \mathbf {e}_1\) and \(\rho ^1_v = y_{s1}\), else if \(u_{vi} = 1\), then \({\tilde{\mathbf {e}}}_v = \mathbf {e}_2\) and \(\rho ^1_v = y_{t2}\). In addition, \(\xi ^1_{vs} = \xi ^1_{vt}= 0\), and \(\xi ^1_{vl} = \max \{ {\tilde{M}}_s, {\tilde{M}}_t \}\) for all \(v \in [n] \) and \(l \in [m] {\setminus } \{s, t\}\). Clearly, the set of points defined in (39) are affinely independent feasible points, and satisfy (25) at equality. Next, we construct a set of points:

$$\begin{aligned}&\mathbf PT ^2_j= \left( \tilde{\mathbf {u}}_j\mathbf {e}_j^\top , \mathbf {e}_j, {\tilde{\mathbf {u}}}_j, \rho ^2_j, \varvec{\xi }^2_j \right) ,&\forall ~j \in [d] {\setminus } \{1,2\}, \end{aligned}$$
(40)

where \({\tilde{\mathbf {u}}}_j\) is any feasible point of (38) with \({\tilde{u}}_{ji} = 0\) if \(y_{sj} \le y_{tj} \), and \({\tilde{u}}_{ji} = 1\) otherwise (i.e., if \(y_{sj} \ge y_{tj} \)), for all \(j \in [d] {\setminus } \{1, 2\}\). In addition, \(\rho ^2_j = \min \{ y_{sj}, y_{tj}\}\), for all \(j \in [d] {\setminus } \{1, 2\}\). Furthermore, \(\xi ^2_{js} = \xi ^2_{jt}= 0\), and \(\xi ^2_{jl} = \max \{ {\tilde{M}}_s, {\tilde{M}}_t \}\) for all \(j \in [d] {\setminus } \{1, 2\} \) and \(l \in [m] {\setminus } \{s, t\}\). It is easy to see that the set of points defined in (40) are feasible, affinely independent from (39), and satisfy (25) at equality.

Furthermore, we construct the following set of points:

$$\begin{aligned}&\mathbf PT ^3_s = \left( {\bar{\mathbf {u}}}_1\mathbf {e}_1^\top , \mathbf {e}_1, {\bar{\mathbf {u}}}_1, y_{t1}, \varvec{\xi }^3_s\right) \end{aligned}$$
(41a)
$$\begin{aligned}&\mathbf PT ^3_t= \left( {\bar{\mathbf {u}}}_2\mathbf {e}_2^\top , \mathbf {e}_2, {\bar{\mathbf {u}}}_2, y_{s2}, \varvec{\xi }^3_t\right) \end{aligned}$$
(41b)
$$\begin{aligned}&\mathbf PT ^3_l= \mathbf PT ^3_s + \left( 0, 0, 0, 0, \mathbf {e}_l\right) ,&\forall ~l \in [m] {\setminus } \{s, t\}, \end{aligned}$$
(41c)

where \({\bar{\mathbf {u}}}_1\) is any feasible point of (38) with \({\bar{u}}_{1i} = 0\), and \({\bar{\mathbf {u}}}_2\) is any feasible point of (38) with \({\bar{u}}_{2i} = 1\). In addition, \(\xi ^3_{ss} = y_{t1} - y_{s1}\), \(\xi ^3_{st} = 0\), and \(\xi ^3_{sl} = \max \{ {\tilde{M}}_s, {\tilde{M}}_t \}\) for all \(l \in [m] {\setminus } \{s,t\}\). Similarly, \(\xi ^3_{ts} = 0\), \(\xi ^3_{tt} = y_{s2} - y_{t2}\), and \(\xi ^3_{tl} = \max \{ {\tilde{M}}_s, {\tilde{M}}_t \}\) for all \(l \in [m] {\setminus } \{s,t\}\). Clearly, the set of points defined by (41) are affinely independent feasible points which satisfy (25) at equality.

Finally, we construct the single point:

$$\begin{aligned} \mathbf PT ^4= \left( \mathbf {u}_1 {\mathbf {c}^*}^\top , \mathbf {c}^*, \mathbf {u}_1, \eta ^*, \varvec{\xi }^4\right) , \end{aligned}$$
(42)

where \(\mathbf {c}^* = (c^*_1, c^*_2, 0, \ldots , 0)\), and the parameters \((c^*_1, c^*_2, \eta ^*)\) are uniquely defined by the following linear system:

$$\begin{aligned}&c^*_1 + c^*_2 = 1 \\&y_{s1}c^*_1 + y_{s2}c^*_2 = \eta ^* \\&y_{t1}c^*_1 + y_{t2}c^*_2 = \eta ^*, \end{aligned}$$

or equivalently, \(c^*_1 = \frac{y_{s2} - y_{t2}}{y_{s2} - y_{t2} + y_{t1} - y_{s1} }\), \(c^*_2 = 1 - c^*_1\), and \(0<c^*_1 ,c^*_2<1 \). In addition, \(\xi ^4_s = \xi ^4_t = 0\), and \(\xi ^4_l =\max \{ {\tilde{M}}_s, {\tilde{M}}_t \}\), for all \(l \in [m] {\setminus }\{s,t\}\).

Clearly, PT \(^4\) is affinely independent from the points defined by (39), since the following matrix:

$$\begin{aligned} \begin{bmatrix} 1&0&y_{s1} \\ 0&1&y_{t2} \\ c^*_1&c^*_2&\eta ^*=y_{s1}c^*_1 + y_{s2}c^*_2 \end{bmatrix}, \end{aligned}$$
(43)

has full rank (due to \( y_{t2} < y_{s2})\). In addition, it is easy to check that (42) is affinely independent from the points defined by (40) and (41). Furthermore, it is also a feasible point which satisfies (25) at equality.

From (39)–(42), we obtain \(n + m +d - 1\) affinely independent feasible points which satisfy (25) at equality. Hence, inequalities (25) are facet defining. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, X., Küçükyavuz, S. & Noyan, N. Robust multicriteria risk-averse stochastic programming models. Ann Oper Res 259, 259–294 (2017). https://doi.org/10.1007/s10479-017-2526-z

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-017-2526-z

Keywords

Navigation