Skip to main content
Log in

Multistage distributionally robust mixed-integer programming with decision-dependent moment-based ambiguity sets

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

Abstract

We study multistage distributionally robust mixed-integer programs under endogenous uncertainty, where the probability distribution of stage-wise uncertainty depends on the decisions made in previous stages. We first consider two ambiguity sets defined by decision-dependent bounds on the first and second moments of uncertain parameters and by mean and covariance matrix that exactly match decision-dependent empirical ones, respectively. For both sets, we show that the subproblem in each stage can be recast as a mixed-integer linear program (MILP). Moreover, we extend the general moment-based ambiguity set in Delage and Ye (Oper Res 58(3):595–612, 2010) to the multistage decision-dependent setting, and derive mixed-integer semidefinite programming (MISDP) reformulations of stage-wise subproblems. We develop methods for attaining lower and upper bounds of the optimal objective value of the multistage MISDPs, and approximate them using a series of MILPs. We deploy the Stochastic Dual Dynamic integer Programming (SDDiP) method for solving the problem under the three ambiguity sets with risk-neutral or risk-averse objective functions, and conduct numerical studies on multistage facility-location instances having diverse sizes under different parameter and uncertainty settings. Our results show that the SDDiP quickly finds optimal solutions for moderate-sized instances under the first two ambiguity sets, and also finds good approximate bounds for the multistage MISDPs derived under the third ambiguity set. We also demonstrate the efficacy of incorporating decision-dependent distributional ambiguity in multistage decision-making processes.

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.

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

Similar content being viewed by others

References

  1. Ahmadi, A.A., Hall, G.: Sum of squares basis pursuit with linear and second order cone programming. Algebraic Geom. Methods Discrete Math. 685, 27–53 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  2. Basciftci, B., Ahmed, S., Shen, S.: Distributionally robust facility location problem under decision-dependent stochastic demand (2019). arXiv preprint arXiv:1912.05577

  3. Ben-Tal, A., El Ghaoui, L., Nemirovski, A.: Robust Optimization, vol. 28. Princeton University Press, Princeton (2009)

    Book  MATH  Google Scholar 

  4. Bertsimas, D., Brown, D.B., Caramanis, C.: Theory and applications of robust optimization. SIAM Rev. 53(3), 464–501 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bertsimas, D., Sim, M., Zhang, M.: Adaptive distributionally robust optimization. Manag. Sci. 65(2), 604–618 (2018)

    Article  Google Scholar 

  6. Birge, J.R., Louveaux, F.: Introduction to Stochastic Programming, 2nd edn. Springer, Berlin (2011)

    Book  MATH  Google Scholar 

  7. Blanchet, J., Murthy, K.: Quantifying distributional model risk via optimal transport. Math. Oper. Res. 44(2), 565–600 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  8. Delage, E., Ye, Y.: Distributionally robust optimization under moment uncertainty with application to data-driven problems. Oper. Res. 58(3), 595–612 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  9. Esfahani, P.M., Kuhn, D.: Data-driven distributionally robust optimization using the Wasserstein metric: Performance guarantees and tractable reformulations. Math. Program. 171(1–2), 115–166 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  10. Gao, R., Kleywegt, A.J.: Distributionally robust stochastic optimization with Wasserstein distance (2016). arXiv preprint arXiv:1604.02199

  11. Girardeau, P., Leclere, V., Philpott, A.B.: On the convergence of decomposition methods for multistage stochastic convex programs. Math. Oper. Res. 40(1), 130–145 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  12. Goel, V., Grossmann, I.E.: A class of stochastic programs with decision dependent uncertainty. Math. Program. 108(2–3), 355–394 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  13. Goh, J., Sim, M.: Distributionally robust optimization and its tractable approximations. Oper. Res. 58(4–part–1), 902–917 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  14. Guigues, V.: Convergence analysis of sampling-based decomposition methods for risk-averse multistage stochastic convex programs. SIAM J. Optim. 26(4), 2468–2494 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  15. Hu, J., Li, J., Mehrotra, S.: A data-driven functionally robust approach for simultaneous pricing and order quantity decisions with unknown demand function. Oper. Res. 67(6), 1564–1585 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  16. Jiang, R., Guan, Y.: Data-driven chance constrained stochastic program. Math. Program. 158(1–2), 291–327 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  17. Jiang, R., Guan, Y.: Risk-averse two-stage stochastic program with distributional ambiguity. Oper. Res. 66(5), 1390–1405 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  18. Jonsbråten, T.W., Wets, R.J., Woodruff, D.L.: A class of stochastic programs with decision dependent random elements. Ann. Oper. Res. 82, 83–106 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  19. Kung, L.C., Liao, W.H.: An approximation algorithm for a competitive facility location problem with network effects. Eur. J. Oper. Res. 267(1), 176–186 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  20. Lappas, N.H., Gounaris, C.E.: The use of decision-dependent uncertainty sets in robust optimization. In: Proceedings of Foundations of Computer-Aided Process Operations/Chemical Process Control 2017 (2017)

  21. Lappas, N.H., Gounaris, C.E.: Robust optimization for decision-making under endogenous uncertainty. Comput. Chem. Eng. 111, 252–266 (2018)

    Article  Google Scholar 

  22. Lee, S., Homem-de Mello, T., Kleywegt, A.J.: Newsvendor-type models with decision-dependent uncertainty. Math. Methods Oper. Res. 76(2), 189–221 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  23. Lofberg, J.: YALMIP: a toolbox for modeling and optimization in MATLAB. In: 2004 IEEE International Conference on Robotics and Automation (IEEE Cat. No. 04CH37508). IEEE, pp. 284–289 (2004)

  24. Luo, F., Mehrotra, S.: Distributionally robust optimization with decision dependent ambiguity sets. Optim. Lett. (2020). https://doi.org/10.1007/s11590-020-01574-3

  25. McCormick, G.P.: Computability of global solutions to factorable nonconvex programs: part I—convex underestimating problems. Math. Program. 10(1), 147–175 (1976)

    Article  MATH  Google Scholar 

  26. Mehrotra, S., Papp, D.: A cutting surface algorithm for semi-infinite convex programming with an application to moment robust optimization. SIAM J. Optim. 24(4), 1670–1697 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  27. Nohadani, O., Sharma, K.: Optimization under decision-dependent uncertainty. SIAM J. Optim. 28(2), 1773–1795 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  28. Noyan, N., Rudolf, G., Lejeune, M.: Distributionally robust optimization with decision-dependent ambiguity set (2018). http://www.optimization-online.org/DB_FILE/2018/09/6821.pdf. Accessed 16 Oct 2020

  29. Pereira, M.V., Pinto, L.M.: Multi-stage stochastic optimization applied to energy planning. Math. Program. 52(1–3), 359–375 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  30. Philpott, A., de Matos, V., Kapelevich, L.: Distributionally robust SDDP. Comput. Manag. Sci. 15(3–4), 431–454 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  31. Philpott, A.B., Guan, Z.: On the convergence of stochastic dual dynamic programming and related methods. Oper. Res. Lett. 36(4), 450–455 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  32. Poss, M.: Robust combinatorial optimization with variable budgeted uncertainty. 4OR 11(1), 75–92 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  33. Rockafellar, R.T., Uryasev, S.: Conditional value-at-risk for general loss distributions. J. Bank. Finance 26(7), 1443–1471 (2002)

    Article  Google Scholar 

  34. Rockafellar, R.T., Uryasev, S., et al.: Optimization of conditional value-at-risk. J. Risk 2(3), 21–42 (2000)

    Article  Google Scholar 

  35. Shapiro, A.: On duality theory of conic linear problems. In: Goberna, M. Á., López, M.A. (eds.) Semi-infinite programming. Nonconvex Optimization and Its Applications, vol. 57. Springer, Boston, MA (2001). https://doi.org/10.1007/978-1-4757-3403-4_7

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

    Book  MATH  Google Scholar 

  37. Spacey, S.A., Wiesemann, W., Kuhn, D., Luk, W.: Robust software partitioning with multiple instantiation. INFORMS J. Comput. 24(3), 500–515 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  38. Vayanos, P., Kuhn, D., Rustem, B.: Decision rules for information discovery in multi-stage stochastic programming. In: 2011 50th IEEE Conference on Decision and Control and European Control Conference. IEEE, pp. 7368–7373 (2011)

  39. Vayanos, P., Georghiou, A., Yu, H.: Robust optimization with decision-dependent information discovery (2020). arXiv preprint arXiv:2004.08490

  40. Wagner, M.R.: Stochastic 0–1 linear programming under limited distributional information. Oper. Res. Lett. 36(2), 150–156 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  41. Webster, M., Santen, N., Parpas, P.: An approximate dynamic programming framework for modeling global climate policy under decision-dependent uncertainty. Comput. Manag. Sci. 9(3), 339–362 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  42. Yu, X., Ahmed, S., Shen, S.: On the value of multistage stochastic facility location with (or without) risk aversion. Technical report. University of Michigan, Department of Industrial and Operations Engineering (2019)

  43. Zhang, Y., Jiang, R., Shen, S.: Ambiguous chance-constrained binary programs under mean-covariance information. SIAM J. Optim. 28(4), 2922–2944 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  44. Zou, J., Ahmed, S., Sun, X.A.: Stochastic dual dynamic integer programming. Math. Program. 175(1–2), 461–502 (2019)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors sincerely thank the Associate Editor and two reviewers for their helpful review and constructive feedback. The authors are grateful for the support from the United States National Science Foundation Grants #1727618, #1709094, and Department of Engineering (DoE) Grant #DE-SC0018018 for this project.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Siqian Shen.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The United States National Science Foundation Grants #1727618, #1709094, and Department of Engineering (DoE) Grant #DE-SC0018018.

Appendices

Appendix

Reformulations of N-DDDR having continuous supports

A continuous version of the Type 1 ambiguity set \({\mathcal {P}_{t+1}^{D_1}}(\varvec{x}_{t})\) in Sect. 3.1 is given by

$$\begin{aligned} { \mathcal {P}_{t+1}^{C_1}} (\varvec{x}_{t})&:=\left\{ P\in \mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})\ |\ \underline{\nu }(\varvec{x}_t) \le P\le \bar{\nu }(\varvec{x}_t),\right. \nonumber \\&\quad \left. \int _{{\varXi }_{t+1}}f_s(\varvec{\xi }_{t+1})P(d\varvec{\xi }_{t+1})\in [l_s(\varvec{x}_{t}),u_s(\varvec{x}_{t})],\ \forall s\in [m]\right\} ,\ \end{aligned}$$
(A.1)

where \(\mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})\) represents the set of all positive measures defined on \((\varXi _{t+1},\mathcal {F}_{t+1})\), and \(\underline{\nu }(\varvec{x}_t),\ \bar{\nu }(\varvec{x}_t)\in \mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})\) are two given measures that are lower and upper bounds for the true probability measure, respectively. To ensure that P is a probability distribution, let \(l_1(\varvec{x}_t)=u_1(\varvec{x}_t)=f_1(\varvec{\xi }_{t+1})=1\) (see (5a) for more details).

Let \(\varXi _{t+1}\) be a closed and bounded set in the Euclidean space, and the probability measures \(P,\ \underline{\nu },\ \bar{\nu }\) be defined on the measurable space \((\varXi _{t+1},\mathcal {F}_{t+1})\), where the \(\sigma \)-algebra \(\mathcal {F}_{t+1}\) contains all singleton subsets, i.e., \(\{\xi \}\in \mathcal {F}_{t+1}\) for all \(\xi \in \varXi _{t+1}\). Then, based on the ambiguity set \({ \mathcal {P}_{t+1}^{C_1}} (\varvec{x}_{t})\) in (A.1), we describe a reformulation of the Bellman equation (2) as an analogy to Theorem 1 for the continuous support case.

Theorem A.1

If for any feasible \(\varvec{x}_t\in \hat{X}_t\), the ambiguity set (A.1) has a non-empty relative interior, then the Bellman equation (2) can be reformulated as \(Q_t(\varvec{x}_{t-1},\varvec{\xi }_t)=\)

$$\begin{aligned} \min _{\varvec{\alpha },\varvec{\beta },{\underline{\gamma },\bar{\gamma },}\varvec{x}_t,\varvec{y}_t}\quad&g_t(\varvec{x}_t,\varvec{y}_t)-{\varvec{\alpha }}^{\mathsf T} \varvec{l}(\varvec{x}_t)+{\varvec{\beta }}^{\mathsf T} \varvec{u}(\varvec{x}_t)\\&\quad -\int _{\varvec{\xi }_{t+1}\in \varXi _{t+1}}\underline{\gamma }(\varvec{\xi }_{t+1})\underline{\nu }(\varvec{x}_t,\varvec{\xi }_{t+1})d\varvec{\xi }_{t+1}\\&\quad + \int _{\varvec{\xi }_{t+1}\in \varXi _{t+1}}\bar{\gamma }(\varvec{\xi }_{t+1})\bar{\nu }(\varvec{x}_t,\varvec{\xi }_{t+1})d\varvec{\xi }_{t+1}\\ \text {s.t.}\quad&(-\varvec{\alpha }+\varvec{\beta })^{\mathsf T} \varvec{f}(\varvec{\xi }_{t+1}) {-\underline{\gamma }(\varvec{\xi }_{t+1})+\bar{\gamma }(\varvec{\xi }_{t+1})}\ge Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}),\ \forall \varvec{\xi }_{t+1}\in \varXi _{t+1},\\&(\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t),\\&\varvec{\alpha },\ \varvec{\beta },\ {\varvec{\underline{\gamma }}(\varvec{\xi }_{t+1}),\ \varvec{\bar{\gamma }}(\varvec{\xi }_{t+1})}\ge 0, \ \forall \varvec{\xi }_{t+1}\in \varXi _{t+1}. \end{aligned}$$

Next, a continuous version of the Type 2 ambiguity set \({ \mathcal {P}_{t+1}^{D_2}} (\varvec{x}_{t})\) in (10) is given by

$$\begin{aligned} { \mathcal {P}^{C_2}_{t+1}}(\varvec{x}_t)&:=\Biggl \{P\in \mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})\ | \mathbb {E}_P[\varvec{\xi }_{t+1}]=\varvec{\mu }(\varvec{x}_t), \end{aligned}$$
(A.2a)
$$\begin{aligned}&\qquad \mathbb {E}_P[(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}]=\varvec{\varSigma }(\varvec{x}_t)\Biggr \}, \end{aligned}$$
(A.2b)

Then, the following result is an analogy to Theorem 2 based on the continuous ambiguity set (A.2).

Theorem A.2

If for any feasible \(\varvec{x}_t\in \hat{X}_t\), the ambiguity set (A.2) has a non-empty relative interior, then the Bellman equation (2) can be reformulated as

$$\begin{aligned} Q_t(\varvec{x}_{t-1},\varvec{\xi }_t)=\min _{\varvec{x}_t,\varvec{y}_t,s,\varvec{u},\varvec{Y}}\quad&g_t(\varvec{x}_t,\varvec{y}_t)+s+\varvec{u}^{\mathsf T}\varvec{\mu }(\varvec{x}_t)+\varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{Y} \end{aligned}$$
(A.3a)
$$\begin{aligned} \text {s.t.}\quad&s+\varvec{u}^{\mathsf T}\varvec{\xi }_{t+1} +(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\cdot \varvec{Y}\nonumber \\&\quad \ge Q_{t+1}(\varvec{x}_t,\varvec{\xi _{t+1}}),\nonumber \\&\qquad \forall \varvec{\xi }_{t+1}\in \varXi _{t+1},\nonumber \\&(\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t). \end{aligned}$$
(A.3b)

Finally, a continuous version of the Type 3 ambiguity set \({ \mathcal {P}_{t+1}^{D_3}}(\varvec{x}_{t})\) in (14) is given by

$$\begin{aligned} { \mathcal {P}_{t+1}^{C_3}}(\varvec{x}_t)&:=\Biggl \{P\in \mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})\ | (\mathbb {E}_P[\varvec{\xi }_{t+1}]-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\varvec{\varSigma }(\varvec{x}_t)^{-1}(\mathbb {E}_P[\varvec{\xi }_{t+1}]-\varvec{\mu }(\varvec{x}_t))\le \gamma , \end{aligned}$$
(A.4a)
$$\begin{aligned}&\qquad \mathbb {E}_P[(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}]\preceq \eta \varvec{\varSigma }(\varvec{x}_t)\Biggr \}. \end{aligned}$$
(A.4b)

We present a reformulation of the Bellman equation (2) in the following theorem that is an analogy to Theorem 3, but given the continuous ambiguity set (A.4).

Theorem A.3

Suppose that the Slater’s constraint qualification conditions are satisfied, i.e., for any feasible \(\varvec{x}_t\in \hat{X}_t\), there exists a probability measure \(P\in \mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})\) such that \((\mathbb {E}_P[\varvec{\xi }_{t+1}]-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\varvec{\varSigma }(\varvec{x}_t)^{-1}(\mathbb {E}_P[\varvec{\xi }_{t+1}]-\varvec{\mu }(\varvec{x}_t))< \gamma \), and \(\mathbb {E}_P[(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}]\prec \eta \varvec{\varSigma }(\varvec{x}_t)\). Using the ambiguity set defined in (A.4), the Bellman equation (2) can be recast as

$$\begin{aligned} Q_t(\varvec{x}_{t-1},\varvec{\xi }_t)=\min _{\varvec{x}_t,\varvec{y}_t,s,\varvec{Z},\varvec{Y}}\quad&g_t(\varvec{x}_t,\varvec{y}_t)+s+\varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{z}_1 -2\varvec{\mu }(\varvec{x}_t)^{\mathsf T}\varvec{z}_2+\gamma z_3 +\eta \varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{Y} \end{aligned}$$
(A.5a)
$$\begin{aligned} \text {s.t.}\quad&s-2\varvec{z}_2^{\mathsf T}\varvec{\xi }_{t+1}+(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\cdot Y\ge Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}),\nonumber \\&\qquad \forall \varvec{\xi }_{t+1}\in \varXi _{t+1}, \end{aligned}$$
(A.5b)
$$\begin{aligned}&\varvec{Z}=\begin{pmatrix} \varvec{z}_1&{}\varvec{z}_2\\ \varvec{z}_2^{\mathsf T}&{}z_3\end{pmatrix}\succeq 0,\ \varvec{Y}\succeq 0, \end{aligned}$$
(A.5c)
$$\begin{aligned}&(\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t). \end{aligned}$$
(A.5d)

The proof of Theorem A.1 is similar to the proof in [24] for the two-stage continuous-support case, and we omit its details. We provide detailed proofs for Theorems A.2 and A.3 in “Appendix C”. Note that the above three reformulations for ambiguity sets with continuous support are semi-infinite programs and thus cannot be optimized directly.

Risk-averse multistage DRO with endogenous uncertainty

We can extend N-DDDR in (1) to a more general setting. Previously, the robust counterpart chooses the worst-case distribution P from a risk-neutral aspect using expectation to measure the uncertain cost over multiple stages. However, a decision maker may measure the worst-case distribution in a risk-averse fashion, and we accordingly replace the expectations by coherent risk measures \(\rho _t, \ \forall t=2,\ldots ,T\). The corresponding risk-averse multistage decision-dependent DRO model is:

$$\begin{aligned} {\mathbf{A}}\mathrm{-}{\mathbf{DDDR:}}&~\nonumber \\ \quad \min _{(\varvec{\varvec{x}_1},\varvec{y}_1)\in X_1}&\Big \{g_1(\varvec{x}_1,\varvec{y}_1)+\max _{P_2\in \mathcal {P}_2(\varvec{x}_1)}\rho _2\Big [\min _{(\varvec{x}_2,\varvec{y}_2)\in X_2(\varvec{x}_1,\varvec{\xi }_2)}g_2(\varvec{x}_2,\varvec{y}_2)+\cdots \nonumber \\&+\max _{P_{t}\in \mathcal {P}_{t}(\varvec{x}_{t-1})}\rho _{t}\Big [\min _{(\varvec{x}_{t},\varvec{y}_{t})\in X_{t}(\varvec{x}_{t-1},\varvec{\xi }_{t})}g_{t}(\varvec{x}_{t},\varvec{y}_{t})+\cdots \nonumber \\&+\max _{P_T\in \mathcal {P}_T(\varvec{x}_{T-1})}\rho _{T}\Big [\min _{(\varvec{x}_T,\varvec{y}_T)\in X_T(\varvec{x}_{T-1},\varvec{\xi }_T)}g_T(\varvec{x}_T,\varvec{y}_T)\Big ]\Big \}. \end{aligned}$$
(B.6)

We consider a special class of coherent risk measures, which is a convex combination of expectation and Conditional Value-at-Risk (CVaR) [34]:

$$\begin{aligned} \rho _t(Z)=(1-\lambda _t)\mathbb {E}[Z]+\lambda _t \text {CVaR}_{\alpha _t}[Z], \end{aligned}$$

where \(\lambda _t\in [0,1]\) is a parameter that balances the expectation and CVaR measure at \(\alpha _t\in (0,1)\) risk level. This risk measure is more general than expectation and it becomes the risk-neutral case when \(\lambda _t=0\).

The Bellman equations for A-DDDR (B.6) then become:

$$\begin{aligned} Q_1=\min _{(\varvec{x}_1,\varvec{y}_1)\in X_1} g_1(\varvec{x}_1,\varvec{y}_1)+\max _{P_2\in \mathcal {P}_2(\varvec{x}_1)}\rho _2[Q_2(\varvec{x}_1,\varvec{\xi }_2)], \end{aligned}$$

where for \(t=2,\ldots ,T-1\),

$$\begin{aligned} Q_t(\varvec{x}_{t-1},\varvec{\xi }_t)=\min _{(\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t)} g_t(\varvec{x}_t,\varvec{y}_t)+\max _{P_{t+1}\in \mathcal {P}_{t+1}(\varvec{x}_t)}\rho _{t+1}[Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1})], \end{aligned}$$
(B.7)

and

$$\begin{aligned} Q_T(\varvec{x}_{T-1},\varvec{\xi }_T)=\min _{(\varvec{x}_T,\varvec{y}_T)\in X_T(\varvec{x}_{T-1},\varvec{\xi }_T)} g_T(\varvec{x}_T,\varvec{y}_T). \end{aligned}$$

Following the results by [33], CVaR can be attained by solving the following optimization problem:

$$\begin{aligned} \text {CVaR}_{\alpha _t}[Z]:=\inf _{\eta \in \mathbb {R}}\left\{ \eta +\frac{1}{1-\alpha _t}\mathbb {E}[Z-\eta ]_{+}\right\} , \end{aligned}$$

where \([a]_{+}:=\max \{a,0\},\) and \(\eta \) is an auxiliary variable. To linearize \([Z-\eta ]_{+}\), we replace it by a variable m with two additional constraints: \(m\ge 0,\ m\ge Z-\eta \).

Recall that by assumption, every \(P_{t+1}\in \mathcal {P}_{t+1}(\varvec{x}_{t})\) has a decision-independent finite support \(\varXi _{t+1}:=\{\varvec{\xi }_{t+1}^k\}_{k=1}^K,\ \forall \varvec{x}_{t}\in X_{t}\) for a fixed K and all \(t\in [T-1]\). Each realization \(k\in [K]\) is associated with probability \(p_k\), and therefore the inner maximization problem \(\max _{P_{t+1}\in \mathcal {P}_{t+1}(\varvec{x}_t)}\rho _{t+1}[Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1})]\) in (B.7) can be reformulated as

$$\begin{aligned} \max _{P_{t+1}\in \mathcal {P}_{t+1}(\varvec{x}_t)}\min _{m,\eta }\quad&\lambda _{t+1}\eta +\sum _{k=1}^Kp_k(\frac{\lambda _{t+1}}{1-\alpha _{t+1}}m+(1-\lambda _{t+1})Q_{t+1}^k) \end{aligned}$$
(B.8a)
$$\begin{aligned} \text {s.t.}\quad&m+\eta \ge Q_{t+1}^k,\ \forall k\in [K], \end{aligned}$$
(B.8b)
$$\begin{aligned}&m\ge 0,\ \forall k\in [K]. \end{aligned}$$
(B.8c)

We further simplify the notation of the recursive function \(Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}^k)\) as \(Q_{t+1}^k\). Associating dual variables \(q_k\) with constraints (B.8b) and applying strong duality result, we have

$$\begin{aligned}&\max _{P_{t+1}\in \mathcal {P}_{t+1}(\varvec{x}_t)}\Biggl \{(1-\lambda _{t+1})\sum _{k=1}^Kp_kQ_{t+1}^k+\max _{\varvec{q}}\sum _{k=1}^Kq_kQ_{t+1}^k \end{aligned}$$
(B.9a)
$$\begin{aligned}&\quad \text {s.t.}\quad q_k\le p_k\frac{\lambda _{t+1}}{1-\alpha _{t+1}},\ \forall k\in [K], \end{aligned}$$
(B.9b)
$$\begin{aligned}&\quad \sum _{k=1}^Kq_k=\lambda _{t+1}, \end{aligned}$$
(B.9c)
$$\begin{aligned}&\quad q_k\ge 0,\ \forall k\in [K].\Biggr \} \end{aligned}$$
(B.9d)

Merging the two layers of maximization problems, for each \(t\in [T-1]\), we solve

$$\begin{aligned} \max _{\varvec{p,q}}\quad&(1-\lambda _{t+1})\sum _{k=1}^K p_kQ_{t+1}^k+q_kQ_{t+1}^k \end{aligned}$$
(B.10a)
$$\begin{aligned} \text {s.t.}\quad&\text{(B-10b), } \text{(B-10c) },\nonumber \\&\varvec{p}\in \mathcal {P}_{t+1}(\varvec{x}_t), \end{aligned}$$
(B.10b)
$$\begin{aligned}&q_k\ge 0,\ \forall k\in [K] . \end{aligned}$$
(B.10c)

In the following subsections, we present reformulations of A-DDDR in (B.6) under the three types of ambiguity sets mentioned in Sect. 3.

1.1 Solving A-DDDR under Type 1 ambiguity set

Using the ambiguity set defined in (3), the inner maximization problem (B.10) can be recast as

$$\begin{aligned} \max _{\varvec{p,q}}\quad&(1-\lambda _{t+1})\sum _{k=1}^Kp_kQ_{t+1}^k+q_kQ_{t+1}^k \end{aligned}$$
(B.11a)
$$\begin{aligned} \text {s.t.}\quad&\text{(B-10b), } \text{(B-10c), } \text{(C-17b)-(C-17e) }\nonumber \\&p_k,\ q_k\ge 0,\ \forall k\in [K] . \end{aligned}$$
(B.11b)

Theorem B.4

If for any feasible \(\varvec{x}_t\in \hat{X}_t\), problem (B.11) is feasible, then the Bellman equation (B.7) can be reformulated as \(Q_t(\varvec{x}_{t-1},\varvec{\xi }_t)=\)

$$\begin{aligned} \min _{\varvec{\alpha },\varvec{\beta },\varvec{x}_t,\varvec{y}_t} \quad&g_t(\varvec{x}_t,\varvec{y}_t)+\lambda _{t+1}\theta -{\varvec{\alpha }}^{\mathsf T} \varvec{l}(\varvec{x}_t)+{\varvec{\beta }}^{\mathsf T} \varvec{u}(\varvec{x}_t){-\underline{\varvec{\gamma }}^{\mathsf T}\underline{\varvec{p}}(\varvec{x}_t) + \bar{\varvec{\gamma }}^{\mathsf T}\bar{\varvec{p}}(\varvec{x}_t)} \end{aligned}$$
(B.12a)
$$\begin{aligned} \text {s.t.}\quad&\pi _k+\theta \ge Q_{t+1}^k,\ \forall k\in [K], \end{aligned}$$
(B.12b)
$$\begin{aligned}&-\frac{\lambda _{t+1}}{1-\alpha _{t+1}}\pi _k+(-\varvec{\alpha }+\varvec{\beta })^{\mathsf T} \varvec{f}(\varvec{\xi }_{t+1}^k){-\underline{\gamma }_k+\bar{\gamma }_k}\ge (1-\lambda _{t+1}) Q_{t+1}^k, \nonumber \\&\quad \forall k\in [K], \end{aligned}$$
(B.12c)
$$\begin{aligned}&(\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t), \end{aligned}$$
(B.12d)
$$\begin{aligned}&\pi _k,\ \varvec{\alpha },\ \varvec{\beta },\ {\varvec{\underline{\gamma }},\ \varvec{\bar{\gamma }}}\ge 0. \end{aligned}$$
(B.12e)

The proof of Theorem B.4 is similar to the one of Theorem 1 in “Appendix C”, where the only difference is that we introduce the two more dual variables \(\pi _k\) and \(\theta \), associated with constraints (B.9b) and (B.9c), respectively.

Notice here when \(\lambda _{t+1}=0\), model (B.12) reduces to the risk-neutral case (4). This reformulation also has similar computational complexity as model (4) in Theorem 1. Therefore, with the same specific ambiguity set considered in (6) in Sect. 3, we can apply McCormick envelopes to obtain a multistage stochastic MILP and deploy SDDiP to solve it.

1.2 Solving A-DDDR under Type 2 ambiguity set

Under Type 2 ambiguity set in (10), the inner maximization problem (B.10) can be recast as

$$\begin{aligned} \max _{\varvec{p,q}}\quad&(1-\lambda _{t+1})\sum _{k=1}^K p_kQ_{t+1}^k+q_kQ_{t+1}^k \end{aligned}$$
(B.13a)
$$\begin{aligned} \text {s.t.}\quad&\text{(B-10b), } \text{(B-10c), } \text{(C-18b)-(C-18d) }\nonumber \\&p_k,\ q_k\ge 0, \quad \forall k\in [K]. \end{aligned}$$
(B.13b)

Theorem B.5

If for any feasible \(\varvec{x}_t\in \hat{X}_t\), problem (B.13) is feasible, then the Bellman equation (B.7) can be reformulated as \(Q_t(\varvec{x}_{t-1},\varvec{\xi }_t)=\)

$$\begin{aligned} \min _{\varvec{x}_t,\varvec{y}_t,s,\varvec{u},\varvec{Y}} \quad&g_t(\varvec{x}_t,\varvec{y}_t)-\lambda _{t+1}\theta +s+\varvec{u}^{\mathsf T}\varvec{\mu }(\varvec{x}_t)+\varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{Y} \end{aligned}$$
(B.14a)
$$\begin{aligned} \text {s.t.}\quad&s+\varvec{u}^{\mathsf T}\varvec{\xi }_{t+1}^k +(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\cdot \varvec{Y}\nonumber \\&\quad -\pi _k\frac{\lambda _{t+1}}{1-\alpha _{t+1}}\ge (1-\lambda _{t+1})Q_{t+1}^k,\nonumber \\&\qquad \forall k\in [K], \end{aligned}$$
(B.14b)
$$\begin{aligned}&\pi _k-\theta \ge Q_{t+1}^k,\quad \forall k\in [K], \end{aligned}$$
(B.14c)
$$\begin{aligned}&\pi _k\ge 0 ,\quad \forall k\in [K], \end{aligned}$$
(B.14d)
$$\begin{aligned}&(\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t). \end{aligned}$$
(B.14e)

The proof of Theorem B.5 is similar to the one of Theorem 2 in “Appendix C”, where the only difference is that we introduce the two more dual variables \(\pi _k\) and \(\theta \), associated with constraints (B.9b) and (B.9c), respectively.

Notice here when \(\lambda _{t+1}=0\), the risk-averse model (B.14) reduces to the risk-neutral case (11). We can apply McCormick envelopes to get a multistage stochastic MILP and use SDDiP algorithm to attain optimal solutions as in Sect. 3.2.

1.3 Solving A-DDDR under Type 3 ambiguity set

Given Type 3 ambiguity set defined in (14), the inner maximization problem (B.10) can be recast as

$$\begin{aligned} \max _{\varvec{p,q}}\quad&(1-\lambda _{t+1})\sum _{k=1}^Kp_kQ_{t+1}^k+q_kQ_{t+1}^k\\ \text {s.t.}\quad&\text{(B-10b), } \text{(B-10c), } \text{(C-20b)(C-20e) }\nonumber \\&p_k\ge 0,\ q_k\ge 0, \quad \forall k\in [K]. \end{aligned}$$

Theorem B.6

Suppose that Slater’s constraint qualification conditions are satisfied, i.e., for any feasible \(\varvec{x}_t\in \hat{X}_t\), there exists a vector \(p = (p_1,p_2,\ldots ,p_K)^{\mathsf T}\) such that \(\sum _{k=1}^Kp_{k}=1\), \((\sum _{k=1}^K p_k\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\varvec{\varSigma }(\varvec{x}_t)^{-1}(\sum _{k=1}^K p_k\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))< \gamma \), and \(\sum _{k=1}^Kp_k(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\prec \eta \varvec{\varSigma }(\varvec{x}_t)\). Using the ambiguity set defined in (14), the Bellman equation (B.7) can be recast as \(Q_t(\varvec{x}_{t-1},\varvec{\xi }_t)=\)

$$\begin{aligned} \min _{\varvec{x}_t,\varvec{y}_t,s,\varvec{Z},\varvec{Y}}\quad&g_t(\varvec{x}_t,\varvec{y}_t)-\lambda \theta +s+\varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{z}_1 -2\varvec{\mu }(\varvec{x}_t)^{\mathsf T}\varvec{z}_2+\gamma z_3 +\eta \varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{Y} \end{aligned}$$
(B.15a)
$$\begin{aligned} \text {s.t.}\quad&s-2\varvec{z}_2^{\mathsf T}\varvec{\xi }_{t+1}^k+(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\cdot Y-\pi _k\frac{\lambda _{t+1}}{1-\alpha _{t+1}}\ge (1-\lambda _{t+1})Q_{t+1}^k,\nonumber \\&\qquad \forall k\in [K], \end{aligned}$$
(B.15b)
$$\begin{aligned}&\varvec{Z}=\begin{pmatrix} \varvec{z}_1&{}\varvec{z}_2\\ \varvec{z}_2^{\mathsf T}&{}z_3\end{pmatrix}\succeq 0,\ \varvec{Y}\succeq 0, \end{aligned}$$
(B.15c)
$$\begin{aligned}&\pi _k - \theta \ge Q_{t+1}^k,\ \pi _k\ge 0 ,\quad \forall k\in [K], \end{aligned}$$
(B.15d)
$$\begin{aligned}&(\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t). \end{aligned}$$
(B.15e)

The proof is similar to the one of Theorem 3. All the proofs in this section are omitted here due to similarity.

Notice at we obtain an MISDP in each stage and when \(\lambda _{t+1}=0\), the risk-averse model (B.15) reduces to the risk-neutral case (15). We can apply McCormick envelopes and approximation schemes to obtain valid upper and lower bounds similar to the procedures in Sects. 4.2.14.2.2.

Details of all needed proofs

Proof

(Theorem 1) The proof follows Theorem 3.1 in [24]. Using the ambiguity set defined in (3), the inner maximization problem of (2) can be expressed as

$$\begin{aligned} \max _{\varvec{p}\in \mathbb {R}^K}\quad&\sum _{k=1}^Kp_kQ_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}^k) \end{aligned}$$
(C.16a)
$$\begin{aligned} \text {s.t.}\quad&\sum _{k=1}^Kp_k\varvec{f}(\varvec{\xi }_{t+1}^k)\ge \varvec{l}(\varvec{x}_{t}), \end{aligned}$$
(C.16b)
$$\begin{aligned}&\sum _{k=1}^Kp_k\varvec{f}(\varvec{\xi }_{t+1}^k)\le \varvec{u}(\varvec{x}_t), \end{aligned}$$
(C.16c)
$$\begin{aligned}&{p_k\ge \underline{p}_k(\varvec{x}_t),\ \forall k\in [K],} \end{aligned}$$
(C.16d)
$$\begin{aligned}&{p_k\le \bar{p}_k(\varvec{x}_t),\ \forall k\in [K],} \end{aligned}$$
(C.16e)
$$\begin{aligned}&p_k\ge 0, \ \forall k\in [K]. \end{aligned}$$
(C.16f)

We associate dual variables \(\varvec{\alpha },\ \varvec{\beta }\in \mathbb {R}^m\) with Constraints (C.16b) and (C.16c), dual variables \(\varvec{\underline{\gamma }}\) and \(\varvec{\bar{\gamma }}\in \mathbb {R}^K\) with Constraints (C.16d) and (C.16e), respectively. When (C.16) is feasible, strong duality holds and the Bellman equation (2) can be reformulated as (4), which completes the proof. \(\square \)

The proof of Theorem A.1 is omitted due to its similarity to the proof of Theorem 3.3 in [24].

Proof

(Theorem 2) Following Type 2 ambiguity set in (10), the inner maximization problem in (2) can be recast as

$$\begin{aligned} \max _{\varvec{p}\in \mathbb {R}^K}\quad&\sum _{k=1}^Kp_kQ_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}^k) \end{aligned}$$
(C.17a)
$$\begin{aligned} \text {s.t.}\quad&\sum _{k=1}^K p_k=1, \end{aligned}$$
(C.17b)
$$\begin{aligned}&\sum _{k=1}^K p_k\varvec{\xi }_{t+1}^k=\varvec{\mu }(\varvec{x}_t), \end{aligned}$$
(C.17c)
$$\begin{aligned}&\sum _{k=1}^K p_k(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}=\varvec{\varSigma }(\varvec{x}_t), \end{aligned}$$
(C.17d)
$$\begin{aligned}&p_k\ge 0,\ \forall k\in [K]. \end{aligned}$$
(C.17e)

If the above linear program is feasible, then strong duality holds. Associate dual variables \(s\in \mathbb {R},\ \varvec{u}\in \mathbb {R}^J, \ \varvec{Y}\in \mathbb {R}^{J\times J}\) with the three sets of constraints, respectively, and recast the inner maximization problem as a minimization problem. After including constraints \((\varvec{x}_t,\varvec{y}_t)\in X_t(\varvec{x}_{t-1},\varvec{\xi }_t)\), the Bellman equation (2) is equivalent to (11), and we complete the proof. \(\square \)

Proof

(Theorem A.2) The proof follows the conic duality in functional spaces [35]. Using the ambiguity set defined in (A.2), the inner maximization problem of (2) can be formulated as a conic linear program in a functional space as follows:

$$\begin{aligned} \max _{P\in \mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})}\quad&\mathbb {E}_P[Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1})] \end{aligned}$$
(C.18a)
$$\begin{aligned} \text {s.t.}\quad&\mathbb {E}_P[1]= 1, \end{aligned}$$
(C.18b)
$$\begin{aligned}&\mathbb {E}_P[\varvec{\xi }_{t+1}]=\varvec{\mu }(\varvec{x}_t), \end{aligned}$$
(C.18c)
$$\begin{aligned}&\mathbb {E}_P[(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}]=\varvec{\varSigma }(\varvec{x}_t) \end{aligned}$$
(C.18d)

We associate dual variables \(s\in \mathbb {R},\ \varvec{u}\in \mathbb {R}^J, \ \varvec{Y}\in \mathbb {R}^{J\times J}\) with the three sets of constraints, respectively. Because the primal problem has a non-empty relative interior, strong duality holds and the dual problem can be formulated as (A.3), which completes the proof. \(\square \)

Proof

(Theorem 3) Given Type 3 ambiguity set (14), the inner maximization problem in (2) can be recast as

$$\begin{aligned} \max _{\varvec{p}\in \mathbb {R}^K, \varvec{\tau }\in \mathbb {R}^J}\quad&\sum _{k=1}^Kp_kQ_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}^k) \end{aligned}$$
(C.19a)
$$\begin{aligned} \text {s.t.}\quad&\sum _{k=1}^K p_k=1, \end{aligned}$$
(C.19b)
$$\begin{aligned}&\sum _{k=1}^K p_k\varvec{\xi }_{t+1}^k=\varvec{\tau }, \end{aligned}$$
(C.19c)
$$\begin{aligned}&(\varvec{\tau }-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\varvec{\varSigma }(\varvec{x}_t)^{-1}(\varvec{\tau }-\varvec{\mu }(\varvec{x}_t))\le \gamma , \end{aligned}$$
(C.19d)
$$\begin{aligned}&\sum _{k=1}^K p_k(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\preceq \eta \varvec{\varSigma }(\varvec{x}_t), \end{aligned}$$
(C.19e)
$$\begin{aligned}&p_k\ge 0,\ \forall k\in [K]. \end{aligned}$$
(C.19f)

We rewrite Constraint (C.19d) as

$$\begin{aligned} \begin{pmatrix} \varvec{\varSigma }(\varvec{x}_t) &{} \varvec{\tau }-\varvec{\mu }(\varvec{x}_t)\\ (\varvec{\tau }-\varvec{\mu }(\varvec{x}_t))^{\mathsf T} &{} \gamma \end{pmatrix} \succeq 0, \end{aligned}$$

and associate dual variables \(s\in \mathbb {R},\ \varvec{u}\in \mathbb {R}^d, \ \varvec{Z}=\begin{pmatrix} \varvec{z}_1&{}\varvec{z}_2\\ \varvec{z}_2^{\mathsf T}&{}z_3\end{pmatrix}\succeq 0,\ \varvec{Y}\succeq 0\) with Constraints (C.19b)–(C.19e), respectively. The Lagrangian function of (C.19) has the following form:

$$\begin{aligned}&L(\varvec{p},\varvec{\tau },s, \varvec{u}, \varvec{Z},\varvec{Y})=\sum _{k=1}^Kp_kQ_{t+1}\left( \varvec{x}_t,\varvec{\xi }_{t+1}^k\right) -s\left( \sum _{k=1}^Kp_k-1\right) +\varvec{u}^{\mathsf T}\left( \varvec{\tau }-\sum _{k=1}^K p_k\varvec{\xi }_{t+1}^k\right) \nonumber \\&\qquad +\begin{pmatrix} \varvec{\varSigma }(\varvec{x}_t) &{} \varvec{\tau }-\varvec{\mu }(\varvec{x}_t)\\ (\varvec{\tau }-\varvec{\mu }(\varvec{x}_t))^{\mathsf T} &{} \gamma \end{pmatrix}\cdot \varvec{Z}\nonumber \\&\qquad + \left( \eta \varvec{\varSigma }(\varvec{x}_t)-\sum _{k=1}^K p_k(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\right) \cdot \varvec{Y}\nonumber \\&\quad = \sum _{k=1}^K p_k\left( Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}^k)-s-\varvec{u}^{\mathsf T}\varvec{\xi }_{t+1}^k-(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\cdot \varvec{Y}\right) \nonumber \\&\qquad +\varvec{\tau }^{\mathsf T}(\varvec{u}+2\varvec{z}_2)+s+\varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{z}_1 -\varvec{\mu }(\varvec{x}_t)^{\mathsf T}(2\varvec{z}_2)+\gamma z_3 +\eta \varvec{\varSigma }(\varvec{x}_t)\cdot Y. \end{aligned}$$
(C.20)

Because problem (C.19) is convex and under the Slater’s conditions, strong duality holds. The maximization problem (C.19) can be recast as

$$\begin{aligned} \min _{s,\varvec{u},\varvec{Z},\varvec{Y}}\left\{ \max _{\varvec{p},\varvec{\tau }}\left\{ L(\varvec{p},\varvec{\tau },s, \varvec{u}, \varvec{Z},\varvec{Y}): \varvec{p\ge 0},\ \varvec{\tau }\in \mathbb {R}^J\right\} \right\} . \end{aligned}$$
(C.21)

Following the Lagrangian function (C.20), after solving the inner maximization problem in (C.21) over \(\varvec{p},\varvec{\tau }\), we have

$$\begin{aligned} \min _{s,\varvec{u},\varvec{Z},\varvec{Y}}\quad&s+\varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{z}_1 -\varvec{\mu }(\varvec{x}_t)^{\mathsf T}(2\varvec{z}_2)+\gamma z_3 +\eta \varvec{\varSigma }(\varvec{x}_t)\cdot \varvec{Y}\\ \text {s.t.}\quad&Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1}^k)-s-\varvec{u}^{\mathsf T}\varvec{\xi }_{t+1}^k-(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}^k-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\cdot \varvec{Y}\le 0,\\&\quad \forall k\in [K],\\&\varvec{u}+2\varvec{z}_2=0,\\&\varvec{Z}=\begin{pmatrix} \varvec{z}_1&{}\varvec{z}_2\\ \varvec{z}_2^{\mathsf T}&{}z_3\end{pmatrix}\succeq 0,\\&\varvec{Y}\succeq 0. \end{aligned}$$

Substituting \( \varvec{u}=-2\varvec{z}_2\) and combining with the outer minimization problem in (2), we complete the proof. \(\square \)

Proof

(Theorem A.3) The proof follows the conic duality in functional spaces [35]. Using the ambiguity set defined in (A.4), the inner maximization problem of (2) can be formulated as a conic linear program in a functional space as follows:

$$\begin{aligned} \max _{P\in \mathcal {M}(\varXi _{t+1},\mathcal {F}_{t+1})}\quad&\mathbb {E}_P[Q_{t+1}(\varvec{x}_t,\varvec{\xi }_{t+1})] \end{aligned}$$
(C.22a)
$$\begin{aligned} \text {s.t.}\quad&\mathbb {E}_P[1]= 1, \end{aligned}$$
(C.22b)
$$\begin{aligned}&(\mathbb {E}[\varvec{\xi }_{t+1}]-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}\varvec{\varSigma }(\varvec{x}_t)^{-1}(\mathbb {E}[\varvec{\xi }_{t+1}]-\varvec{\mu }(\varvec{x}_t))\le \gamma , \end{aligned}$$
(C.22c)
$$\begin{aligned}&\mathbb {E}[(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))(\varvec{\xi }_{t+1}-\varvec{\mu }(\varvec{x}_t))^{\mathsf T}]\preceq \eta \varvec{\varSigma }(\varvec{x}_t) \end{aligned}$$
(C.22d)

We associate dual variables \(s\in \mathbb {R}, \ \varvec{Z}=\begin{pmatrix} \varvec{z}_1&{}\varvec{z}_2\\ \varvec{z}_2^{\mathsf T}&{}z_3\end{pmatrix}\succeq 0,\ \varvec{Y}\succeq 0\) with Constraints (C.22b)–(C.22d), respectively. The Slater’s constraint qualification conditions ensure that the primal problem has a non-empty relative interior. Therefore, strong duality holds and the dual problem can be formulated as (A.5), which completes the proof.    \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yu, X., Shen, S. Multistage distributionally robust mixed-integer programming with decision-dependent moment-based ambiguity sets. Math. Program. 196, 1025–1064 (2022). https://doi.org/10.1007/s10107-020-01580-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-020-01580-4

Keywords

Mathematics Subject Classification

Navigation