Skip to main content
Log in

SDDP for some interstage dependent risk-averse problems and application to hydro-thermal planning

  • Published:
Computational Optimization and Applications Aims and scope Submit manuscript

Abstract

We consider interstage dependent stochastic linear programs where both the random right-hand side and the model of the underlying stochastic process have a special structure. Namely, for equality constraints (resp. inequality constraints) the right-hand side is an affine function (resp. a given function b t ) of the process value for the current time step t. As for m-th component of the process at time step t, it depends on previous values of the process through a function h tm .

For this type of problem, to obtain an approximate policy under some assumptions for functions b t and h tm , we detail a stochastic dual dynamic programming algorithm. Our analysis includes some enhancements of this algorithm such as the definition of a state vector of minimal size, the computation of feasibility cuts without the assumption of relatively complete recourse, as well as efficient formulas for sharing optimality and feasibility cuts between nodes of the same stage. The algorithm is given for both a non-risk-averse and a risk-averse model. We finally provide preliminary results comparing the performances of the recourse functions corresponding to these two models for a real-life application.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

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

Similar content being viewed by others

Notes

  1. If all lags p t (m) are null, we recover the case when process (ξ t ) is interstage independent.

  2. The generalized autoregressive model is written using normalized random variables. For numerical reasons, it is recommended to use such formulation for the calibration of the model.

  3. Exchanges between subsystems can also be considered. Demand satisfaction constraints can be written as INEQ in this case too.

  4. We use notation from [2] and [15].

  5. Risk measures are defined on random variables representing costs, contrary to [4] where they are defined on random variables representing incomes. We easily switch from one setting to another since an income is the opposite of a cost.

  6. We adopt the convention 1 \(\mathrm{MWMonth}=\frac{365.25\times24}{12}\) MWh=730.5 MWh.

  7. The runs were done on a Dell PowerEdge 2900 server with 2 CPUs Intel Xeon E5345 (2.33 GHz, 8M of cache memory, 1333 MHz FSB), running under CentOS release 5, with 48 GB of RAM.

  8. The computational time was approximately 4 weeks.

References

  1. Birge, J.R.: Decomposition and partitioning methods for multistage stochastic linear programs. Oper. Res. 33, 989–1007 (1985)

    Article  MATH  MathSciNet  Google Scholar 

  2. Birge, J.R., Donohue, C.J.: The abridged nested decomposition method for multistage stochastic linear programs with relatively complete recourse. Algorithmic Oper. Res. 1, 20–30 (2001)

    MathSciNet  Google Scholar 

  3. Birge, J., Louveaux, F.: Introduction to Stochastic Programming. Springer, New York (1997)

    MATH  Google Scholar 

  4. Guigues, V., Römisch, W.: Sampling-based decomposition methods for multistage stochastic programs based on extended polyhedral risk measures. SIAM J. Optim. 22, 286–312 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  5. Guigues, V., Römisch, W.: SDDP for multistage stochastic linear programs based on spectral risk measures. Oper. Res. Lett. 40, 313–318 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  6. Guigues, V., Sagastizábal, C.: Exploiting the structure of autoregressive processes in chance-constrained multistage stochastic linear programs. Oper. Res. Lett. 40, 478–483 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  7. Guigues, V., Sagastizábal, C.: The value of rolling horizon policies for risk-averse hydro-thermal planning. Eur. J. Oper. Res. 217, 129–140 (2012)

    Article  MATH  Google Scholar 

  8. Guigues, V., Sagastizábal, C.: Risk-averse feasible policies for large-scale multistage stochastic linear programs. Math. Program. 138, 167–198 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  9. Homem de Mello, T., de Matos, V.L., Finardi, E.C.: Sampling strategies and stopping criteria for stochastic dual dynamic programming: a case study in long-term hydrothermal scheduling. Energy Syst. 2, 1–31 (2011)

    Article  Google Scholar 

  10. Infanger, G., Morton, D.: Cut sharing for multistage stochastic linear programs with interstage dependency. Math. Program. 75, 241–256 (1996)

    MATH  MathSciNet  Google Scholar 

  11. Maceira, M.E., Damázio, J.: The use of PAR(p) model in the stochastic dual dynamic programming optimization scheme used in the operation planning of the Brazilian hydropower system. In: VIII International Conference on Probabilistic Methods Applied to Power Systems, pp. 397–402 (2004)

    Google Scholar 

  12. Morton, D.P.: An enhanced decomposition algorithm for multistage stochastic hydroelectric scheduling. Ann. Oper. Res. 64, 211–235 (1996)

    Article  MATH  MathSciNet  Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

  14. Philpott, A., de Matos, V.: Dynamic sampling algorithms for multi-stage stochastic programs with risk aversion. Eur. J. Oper. Res. 218, 470–483 (2012)

    Article  MATH  Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  17. Ruszczyński, A., Shapiro, A.: Conditional risk mappings. Math. Oper. Res. 31, 544–561 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  18. Ruszczyński, A., Shapiro, A.: Optimization of convex risk functions. Math. Oper. Res. 31, 433–452 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  19. Shapiro, A.: Analysis of stochastic dual dynamic programming method. Eur. J. Oper. Res. 209, 63–72 (2011)

    Article  MATH  Google Scholar 

  20. Shapiro, A., Tekaya, W., da Costa, J.P., Soares, M.P.: Risk neutral and risk averse stochastic dual dynamic programming method. Eur. J. Oper. Res. 224, 375–391 (2013)

    Article  Google Scholar 

  21. Wittrock, R.J.: Advances in a nested decomposition algorithm for solving staircase linear programs. Technical Report SOL 83-2, Systems Optimization Laboratory, Department of Operations Research, Stanford University, Stanford, CA (1983)

Download references

Acknowledgements

The author’s research was partially supported by an FGV grant, PRONEX-Optimization, and CNPq grant No. 382.851/07-4. We would like to thank the reviewers and the Associate Editor for beneficial comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vincent Guigues.

Appendix

Appendix

Proof of Theorem 2.4

By duality, \(Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t j} )\) may be expressed as the optimal value of the following linear program (due to Assumption (A3) the dual and the primal have the same finite optimal value):

$$ \begin{aligned} &\max_{\lambda_t, \pi_{t 1}, \pi_{t 2}, \rho_t} \; g_1( \lambda_t, \pi_{t 1}, \pi_{t 2}, \rho_t, \eta_{t j} ) \\ &C_t^{\scriptscriptstyle \top }\pi_{t 1}^{\scriptscriptstyle \top }+ A_t^{\scriptscriptstyle \top }\pi_{t 2}^{\scriptscriptstyle \top }+ { \overrightarrow{E}_{t}^i}^{\scriptscriptstyle \top }\rho_t^{\scriptscriptstyle \top }\leq\beta _t^{\scriptscriptstyle \top }\lambda_t^{\scriptscriptstyle \top }\\ &\lambda_t e =1,\qquad \rho_t e =1, \qquad \rho_t \geq0,\qquad \lambda_t \geq0,\qquad \pi _{t 2} \geq0 \end{aligned} $$
(40)

where the objective function is given by

$$ \begin{aligned} &g_1( \lambda_t, \pi_{t 1}, \pi_{t 2}, \rho_t, \eta_{t j}) \\ &\quad = \lambda_t \alpha_t + \rho_t \bigl(\overrightarrow{e}_{t}^i + \overrightarrow {{\tilde{E}}}_{t}^i( {\tilde{\varPhi}}_{t} { \xi}_{[t-1]}+{\tilde{\varPsi}}_{t} \eta_{t j} + { \tilde{\varTheta}}_{t} ) \bigr) \\ &\qquad{} + \pi_{t 2} \bigl(b_t( \varPhi_{t} { \xi}_{[t-1]}+\varPsi_{t} \eta_{t j} + \varTheta_{t} )-B_t x_{t-1} \bigr) \\ &\qquad{} + \pi_{t 1} \bigl(D_t( \varPhi_{t} { \xi }_{[t-1]}+ \varPsi_{t} \eta_{t j} + \varTheta_{t} )-E_t x_{t-1} \bigr). \end{aligned} $$
(41)

For problem (40), optimal solutions are extremal points of the feasible set. Further, the feasible set neither depends on x t−1 nor on ξ [t−1] and for any (x t−1,ξ [t−1]), row vectors \(\lambda_{t}^{k j}, \pi_{t 1}^{k j}, \pi_{t 2}^{k j}, \rho_{t}^{k j}\) are extremal points of the feasible set of problem \(Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t j} )\) expressed as (40). It follows that \(Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t j})\) is bounded from below by

$$ \begin{aligned} &g_1\bigl( \lambda_t^{k j}, \pi_{t 1}^{k j}, \pi_{t 2}^{k j}, \rho_t^{k j}, \eta_{t j}\bigr) \\ &\quad = \lambda_t^{k j} \alpha_t + \rho_t^{k j} \bigl(\overrightarrow{e}_{t}^i + \overrightarrow{{\tilde{E}}}_{t}^i( {\tilde { \varPhi}}_{t} {\xi}_{[t-1]}+{\tilde{\varPsi}}_{t} \eta_{t j} + {\tilde {\varTheta }}_{t} ) \bigr) \\ &\qquad{} + \pi_{t 2}^{k j} \bigl(b_t( \varPhi_{t} {\xi}_{[t-1]}+ \varPsi_{t} \eta_{t j} + \varTheta_{t} )-B_t x_{t-1} \bigr) \\ &\qquad{} + \pi_{t 1}^{k j} \bigl(D_t( \varPhi _{t} {\xi}_{[t-1]}+ \varPsi_{t} \eta_{t j} + \varTheta_{t} )-E_t x_{t-1} \bigr) \end{aligned} $$
(42)

for j=1,…,q t . Next, from the convexity of b ti , we obtain

$$ b_{t i}( \varPhi_{t} {\xi}_{[t-1]}+ \varPsi_{t} \eta_{t j} + \varTheta_{t} ) \geq b_{t i} \bigl(\xi_{t j}^{k} \bigr) + s_{t i}^b \bigl( \xi_{t j}^{k} \bigr)^{\scriptscriptstyle \top }\varPhi _t \bigl(\xi_{[t-1]}- \xi_{[t-1]}^k \bigr), \quad i=1,\ldots,\ell_t, $$
(43)

and since \(\pi_{t 2}^{k j} \geq0\), we have

$$ \pi_{t 2}^{k j} b_t( \varPhi_{t} {\xi}_{[t-1]}+ \varPsi_{t} \eta_{t j} + \varTheta_{t} ) \geq\pi_{t 2}^{k j} \bigl[ b_t \bigl( \xi_{t j}^{k} \bigr) + s_t^b \bigl( \xi_{t j}^{k} \bigr) \varPhi_t \bigl(\xi_{[t-1]}-\xi_{[t-1]}^k \bigr) \bigr]. $$
(44)

Plugging (44) into lower bound (42) for \(Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t j})\) and since \(\mathcal{Q}_{t} ( x_{t-1}, \xi_{[t-1]})\) is bounded from below by \(\mathbb{E}_{\eta_{t}}[Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t})] = \sum_{j=1}^{q_{t}} \, p(t,j) Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t j})\), we obtain a cut of the form \(\theta_{t-1}^{k} + E_{t-1}^{k} x_{t-1} \geq {\tilde{E}}_{t-1}^{k} \xi_{[t-1]}+e_{t-1}^{k}\) and the result follows. □

Proof of Theorem 2.8

We show by induction, from t=T+1 down to t=2, that the announced cuts are valid and that \({\tilde{E}}_{t}^{k} \geq0\) for t=1,…,T, and k=0,1,…,iH. For t=T, we have \(\mathcal{Q}_{T+1}^{i}=\mathcal{Q}_{T+1} \equiv0\). As a result, all components of \(E_{T}^{k}, {\tilde{E}}_{T}^{k}\), and \(e_{T}^{k}\) are null for k=0,1,…,iH. In particular, we have that \({\tilde{E}}_{t}^{k} \geq0\). This achieves the first step of the induction.

Let us now assume that for some t∈{2,…,T}, valid cuts have been built for \(\mathcal{Q}_{\ell+1}, \ell=t,\ldots,T\), according to the formulas given in the theorem for \(E_{\ell}^{k}, {\tilde{E}}_{\ell}^{k}\), and \(e_{\ell}^{k}\), =t,…,T, k=0,1,…,iH, with all \({\tilde{E}}_{\ell}^{k} \geq0\).

We have \(\mathcal{Q}_{t}(x_{t-1}, \xi_{[t-1]})\) \(\geq\sum_{j=1}^{q_{t}} p(t,j) \mathcal{Q}_{t}^{i}(x_{t-1}, \xi_{[t-1]}, \eta_{t j})\) with

$$\begin{aligned} \mathcal{Q}_t^i(x_{t-1}, \xi_{[t-1]}, \eta_{t j}) \geq& g_2 \bigl( \lambda_t^{k j}, \pi_{t 2}^{k j}, \rho_t^{k j}, \eta_{t j} \bigr) \\ =& \lambda_t^{k j} \alpha_t + \rho_t^{k j} \overrightarrow{e}_{t}^i -\pi_{t 2}^{k j}B_t x_{t-1} +U_{t j}^{k} + V_{t j}^{k}+W_{t j}^{k} \end{aligned}$$
(45)

where

$$\begin{aligned} U_{t j}^{k} = & \sum_{m=1}^M \sum_{\ell=1}^{s_{t+1, m}-1} \Biggl[ \sum _{w=0}^{i H} \rho_{t}^{k j}(w) { \tilde{E}}_{t, m}^w(\ell+1) \Biggr] \xi _{t-\ell}(m), \end{aligned}$$
(46)
$$\begin{aligned} V_{t j}^{k} = & \sum_{m=1}^M \sum_{w=0}^{i H} \rho_{t}^{k j}(w) {\tilde{E}}_{t, m}^w(1) \xi_{t j}(m) , \end{aligned}$$
(47)
$$\begin{aligned} W_{t j}^{k} = & \sum_{\ell=1}^{\ell_t} \pi_{t 2}^{k j}(\ell)b_{t \ell } ( \xi_{t j} ). \end{aligned}$$
(48)

Let us first bound from below \(W_{t j}^{k}\). Using the convexity of b tℓ , we obtain

$$b_{t \ell}( \xi_{t j} ) \geq b_{t \ell} \bigl( \xi_{t j}^{k} \bigr) + s_{t \ell}^b \bigl( \xi_{t j}^{k} \bigr)^{\scriptscriptstyle \top }\bigl( \xi_{t j} - \xi_{t j}^{k} \bigr) $$

for every =1,…, t . Using these inequalities and the fact that \(\pi_{t 2}^{k j} \geq0\), we have

$$ W_{t j}^{k} = \pi_{t 2}^{k j} b_{t}( \xi_{t j} ) \geq\pi_{t 2}^{k j} b_{t} \bigl( \xi_{t j}^{k} \bigr) + \pi_{t 2}^{k j} s_{t}^b \bigl( \xi_{t j}^{k} \bigr) \bigl( \xi _{t j} - \xi_{t j}^{k} \bigr). $$
(49)

Similarly, using the convexity of h tm , we have

$$\begin{aligned} \xi_{t j}(m)-\xi_{t j}^{k}(m) =& h_{t m}\bigl(\xi_{t-1:t-p_{t}(m)}(m),\eta_{t j}(m) \bigr)-h_{t m}\bigl(\xi_{t-1:t-p_{t}(m)}^k(m), \eta_{t j}(m)\bigr) \\ \geq&\sum_{w=1}^{p_t(m)} s_{t m}^h \bigl(\xi _{t-1:t-p_{t}(m)}^k(m),\eta_{t j}(m) \bigr) (w) \bigl( \xi_{t-w}(m) \\ &{}- \xi _{t-w}^k(m) \bigr) \end{aligned}$$
(50)

for every m=1,…,M. Using Assumption (A4), each component of each subgradient of b tℓ is nonnegative. As a result, all elements in matrix \(s_{t}^{b}( \xi_{t j}^{k})\) are nonnegative. Using this observation and relations (49) and (50), we obtain for \(W_{t j}^{k}\) the lower bound \(\pi_{t 2}^{k j} b_{t}( \xi _{t j}^{k})\) plus

$$\begin{aligned} &\sum_{m=1}^M \sum _{w=1}^{p_t(m)} \Biggl[ \sum _{\ell=1}^{\ell_t} \pi_{t 2}^{k j}(\ell) s_{t \ell}^b \bigl( \xi_{t j}^{k} \bigr) (m) s_{t m}^h \bigl(\xi _{t-1:t-p_{t}(m)}^k(m), \eta_{t j}(m) \bigr) (w) \Biggr] \\ &\quad {} \times \bigl(\xi _{t-w}(m)-\xi _{t-w}^k(m) \bigr). \end{aligned}$$
(51)

Let us now bound from below \(V_{t j}^{k}\). Using relation (50) and the nonnegativeness of \({\tilde{E}}_{t, m}^{w}(1)\) (induction hypothesis) and of \(\rho_{t}^{k j}(w)\), we obtain for \(V_{t j}^{k}\) the lower bound

$$\begin{aligned} &\sum_{m=1}^M \sum_{w=0}^{i H} \rho_{t}^{k j}(w) {\tilde{E}}_{t, m}^w(1) \displaystyle{\sum _{u=1}^{p_t(m)}} s_{t m}^h \bigl( \xi _{t-1:t-p_{t}(m)}^k(m),\eta_{t j}(m) \bigr) (u) \bigl(\xi _{t-u}(m)-\xi _{t-u}^k(m) \bigr) \\ &\quad{} + \sum_{m=1}^M \sum _{w=0}^{i H} \rho_{t}^{k j}(w) { \tilde{E}}_{t, m}^w(1) \xi_{t j}^{k}(m). \end{aligned}$$
(52)

Plugging into (45) relation (46) as well as lower bounds (51) and (52) for respectively \(W_{t j}^{k}-\pi_{t 2}^{k j} b_{t}( \xi_{t j}^{k})\) and \(V_{t j}^{k}\), we obtain for \(\mathcal{Q}_{t}\) a cut of form \(-E_{t-1}^{k} x_{t-1}+{\tilde{E}}_{t-1}^{k} \xi_{[t]}+e_{t-1}^{k}\) with the desired values of \(E_{t-1}^{k}, {\tilde{E}}_{t-1}^{k}\), and \(e_{t-1}^{k}\). If remains to check that all components of \({\tilde{E}}_{t-1}^{k}\) are nonnegative. We had already observed that for all functions b ti , all components of all subgradients are nonnegative, due to Assumption (A4). The same remark holds for functions h tm , due to Assumption (A6). By induction hypothesis, all coefficients \(({\tilde{E}}_{t m}^{j}(\ell))_{j, \ell, m}\) are nonnegative. Using the nonnegativity of these coefficients, as well as the nonnegativity of row vectors \(\rho _{t}^{k j}\) and \(\pi_{t 2}^{k j}\), together with the formula for \({\tilde{E}}_{t-1}^{k}\), we obtain that \({\tilde{E}}_{t-1}^{k} \geq0,\;k=0,1,\ldots,iH\). □

Proof of Proposition 2.9

We show the cuts are valid when there are both equality and inequality constraints. A similar proof can be done when there are not equality constraints. Let k∈{(i−1)H+1,…,iH} and j∈{1,…,q t }. If \(\eta_{t j} \in\varOmega_{t}^{k}\) then \(Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta _{t j} )\) is bounded from below by \(g_{1}( \lambda_{t}^{k j}, \pi_{t 1}^{k j}, \pi_{t 2}^{k j}, \rho _{t}^{k j}, \eta_{t j} )\) where the expression of g 1 is given by (40). Next, for every j such that \(\eta_{t j} \notin\varOmega_{t}^{k}\), since all \((\lambda_{t}, \pi_{t 1}, \pi_{t 2}, \rho_{t}) \in\mathcal{M}_{t}^{i}\) belong to the feasible set of problem \(Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t j} )\), we have \(Q_{t}^{i} ( x_{t-1}, \xi_{[t-1]}, \eta_{t j} ) \geq g_{1}( \lambda_{t}, \pi _{t 1}, \pi_{t 2}, \rho_{t}, \eta_{t j} )\) for every \((\lambda_{t}, \pi_{t 1}, \pi_{t 2}, \rho_{t}) \in\mathcal {M}_{t}^{i}\). As a result,

$$\begin{aligned} Q_t^i ( x_{t-1}, \xi_{[t-1]}, \eta_{t j} ) \geq& \max_{(\lambda_t, \pi_{t 1}, \pi_{t 2}, \rho_t) \in\mathcal {M}_t^i} \; g_1( \lambda_t, \pi_{t 1}, \pi_{t 2}, \rho_t, \eta_{t j} ) \\ = & g_1\bigl( \lambda_t^{k j}, \pi_{t 1}^{k j}, \pi_{t 2}^{k j}, \rho_t^{k j}, \eta_{t j} \bigr) \quad \mbox{ using } \mbox{(25)}. \end{aligned}$$

We then conclude as in the proof of Theorem 2.4. □

Proof of Theorem 2.10

Let x t−1 be a feasible state at the end of time step t−1 at a given node of this time step with history ξ [t−1]. Since for one of the son nodes, the realization of η t is η tj , the optimal value of (32) is 0. As a result, the optimal value of the dual of (32) is 0. This dual problem can be written

$$ \begin{cases} \max_{\pi_{t 1}, \pi_{t 2}, \sigma_t } \;f( \pi_{t 1}, \pi_{t 2}, \sigma_t )\\ C_t^{\scriptscriptstyle \top }\pi_{t 1}^{\scriptscriptstyle \top }+ A_t^{\scriptscriptstyle \top }\pi_{t 2}^{\scriptscriptstyle \top }+ {\overrightarrow{F}}_{t}^{\scriptscriptstyle \top }\sigma_t^{\scriptscriptstyle \top }\leq0\\ -e \leq\pi_{t 1}^{\scriptscriptstyle \top }\leq e, \qquad 0 \leq\pi_{t 2}^{\scriptscriptstyle \top }\leq e, \qquad 0 \leq\sigma_t^{\scriptscriptstyle \top }\leq e, \end{cases} $$
(53)

where the objective function f is given by

$$\begin{aligned} &\pi_{t 1} \bigl[ D_t( \varPhi_t \xi_{[t-1]}+\varPsi_t \eta_{t j} + \varTheta_t ) -E_t x_{t-1} \bigr]+ \pi_{t 2} \bigl[b_t( \varPhi_t \xi_{[t-1]} \\ &\quad {}+\varPsi _t \eta _{t j} + \varTheta_t )-B_t x_{t-1} \bigr] \\ &\quad {}+ \sigma_t \bigl[ \overrightarrow{{\tilde{F}}}_{t}( {\tilde{\varPhi}}_t \xi _{[t-1]}+ {\tilde{\varPsi}}_t \eta_{t j} + {\tilde{\varTheta}}_t ) + \overrightarrow{f}_{t} \bigr]. \end{aligned}$$

For this dual problem, since the optimal value is 0 and since \((\pi_{t 1}^{k j}, \pi_{t 2}^{k j}, \sigma_{t}^{k j})\) is feasible, we obtain

$$0 \geq f \bigl( \pi_{t 1}^{k j}, \pi_{t 2}^{k j}, \sigma_t^{k j} \bigr). $$

We conclude using (43) and (44). □

Proof of Proposition 2.11

We follow the proofs of Theorems 2.8 and 2.10. □

Proof of Theorem 3.1

The proof is similar to the proof of Theorem 2.4. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Guigues, V. SDDP for some interstage dependent risk-averse problems and application to hydro-thermal planning. Comput Optim Appl 57, 167–203 (2014). https://doi.org/10.1007/s10589-013-9584-1

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10589-013-9584-1

Keywords

Navigation