17.1 Introduction
In this chapter we consider bilevel optimization models with uncertain parameters. Such models can be classified based on the chronology of decision and observation as well as the nature of the uncertainty involved. A bilevel stochastic program arises, if the uncertain parameter is realization of some random vector with known distribution, that can only be observed once the leader has submitted their decision. In contrast, the follower decides under complete information.
If upper and lower level objectives coincide, the bilevel stochastic program collapses to a classical stochastic optimization problem with recourse (cf. [1, Chap. 2]). Relations to other mathematical programming problems are explored in the seminal work [2] that also established the existence of solutions, Lipschitzian properties and directional differentiability of a risk-neutral formulation of a bilevel stochastic nonlinear model. Moreover, gradient descent and penalization methods were investigated to tackle discretely distributed stochastic mathematical programs with equilibrium constraints (SMPECs).
Reference [3] studies an application to topology optimization problems in structural mechanics. Many other applications are motivated by network related problems that inherit a natural order of successive decision making under uncertainty. Notable examples arise in telecommunications (cf. [4]), grid-based (energy) markets (cf. [5–8]) or transportation science (cf. [9, 10]). An extensive survey on bilevel stochastic programming literature is provided in [11, Chap. 1.4].
In two-stage stochastic bilevel programming leader and follower take two decisions: The decision on the respective first-stage variables is made in a here-and-now fashion, i.e. without knowledge of the realization of the random parameter. In contrast, the respective second-stage decisions are made in a wait-and-see manner, i.e. after observing the parameter (cf. [12]).
This chapter is organized as follows: In Sects. 17.2.1–17.2.5, we outline structural properties, existence and optimality conditions as well as stability results for bilevel stochastic linear problems while paying special attention to the modelling of risk-aversion via coherent or convex risk measures or stochastic dominance constraints. Sections 17.2.6 and 17.2.7 are devoted to the algorithmic treatment of bilevel stochastic linear problems, where the underlying distribution is finite discrete . An application of two-stage stochastic bilevel programming in the context of network pricing is discussed in Sect. 17.3. The chapter concludes with an overview of potential challenges for future research.
17.2 Bilevel Stochastic Linear Optimization
While the analysis in this section is confined to the bilevel stochastic linear problems with random right-hand side , the concepts and underlying principles can be easily transferred to stochastic extensions of more complex bilevel programming models.
17.2.1 Preliminaries
![$$\displaystyle \begin{aligned} \min_x \left\{c^\top x + \min_y \{q^\top y \; | \; y \in \Psi(x,z)\} \; | \; x \in X \right\}, \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ1.png)
![$$X \subseteq \mathbb {R}^n$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq1.png)
![$$c \in \mathbb {R}^n$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq2.png)
![$$q \in \mathbb {R}^m$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq3.png)
![$$z \in \mathbb {R}^s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq4.png)
![$$\Psi : \mathbb {R}^n \times \mathbb {R}^s \rightrightarrows \mathbb {R}^m$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq5.png)
![$$\displaystyle \begin{aligned}\Psi(x,z) := \underset{y}{\mbox{Argmin}} \; \{d^\top y \; | \; Ay \leq Tx + z\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equa.png)
![$$A \in \mathbb {R}^{s \times m},\ T \in \mathbb {R}^{s \times n}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq6.png)
![$$d \in \mathbb {R}^m$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq7.png)
![$$f: \mathbb {R}^n \times \mathbb {R}^s \to \mathbb {R} \cup \lbrace \pm \infty \rbrace $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq8.png)
![$$\displaystyle \begin{aligned}f(x,z) := c^\top x + \min_y \{q^\top y \; | \; y \in \Psi(x,z)\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equb.png)
Assume dom f ≠ ∅, then f is real-valued
and Lipschitz continuous
on the polyhedron
. △
![$$\displaystyle \begin{aligned} \begin{array}{rcl} |f(x,z)-f(x',z')| \; &\displaystyle = \; f(x,z)- c^\top x' - q^\top y' \; \leq \; c^\top x + q^\top y - c^\top x' - q^\top y' \\ &\displaystyle \leq \; \|c\| \|x-x'\| + \|q\| \|y-y'\| \end{array} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ2.png)
![$$\mathbb {B}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq10.png)
![$$\displaystyle \begin{aligned}\Psi(x',z')\subseteq \Psi(x,z)+ \Lambda \|(x,z)-(x',z')\|\mathbb{B} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equc.png)
An alternate proof for Lemma 17.2.1 is given in [17, Theorem 1]. However, the arguments above can be easily extended to lower level problems with convex quadratic objective function and linear constraints. △
Linear programming theory (cf. [18]) provides verifiable necessary and sufficient condition for dom f ≠ ∅:
![$$(x,z) \in \mathbb {R}^n \times \mathbb {R}^s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq11.png)
- a.
{y | Ay ≤ Tx + z} is nonempty,
- b.
there is some
satisfying A ⊤u = d and u ≤ 0, and
- c.
the function y↦q ⊤y is bounded from below on Ψ(x, z).
![$$\displaystyle \begin{aligned}\min_{y} \lbrace q^\top y \; | \; y \in \Psi(x',z') \rbrace \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equd.png)
is attained for any (x′, z′) ∈ P. △
17.2.2 Bilevel Stochastic Linear Programming Models
![$$(\Omega , \mathcal {F}, \mathbb {P})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq13.png)
![../images/480569_1_En_17_Chapter/480569_1_En_17_Eque_HTML.png](../images/480569_1_En_17_Chapter/480569_1_En_17_Eque_HTML.png)
![$$\mu _Z := \mathbb {P} \circ Z^{-1} \in \mathcal {P}(\mathbb {R}^s)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq14.png)
![$$\displaystyle \begin{aligned}X \subseteq P_Z := \lbrace x \in \mathbb{R}^n \; | \; (x,z) \in P \; \forall z \in \mbox{supp} \; \mu_Z \rbrace. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equf.png)
![$$\mathcal {R}: \mathcal {X} \to \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq15.png)
![$$\mathcal {X}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq16.png)
![$$L^0(\Omega , \mathcal {F}, \mathbb {P})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq17.png)
![$$\displaystyle \begin{aligned}\lbrace f(x,Z(\cdot)) \; | \; x \in X \rbrace \subseteq \mathcal{X}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equg.png)
![$$\displaystyle \begin{aligned} \min_x \left\{\mathcal{R}[f(x,Z(\cdot))] \; | \; x \in X \right\}. \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ3.png)
![$$L^p(\Omega , \mathcal {F}, \mathbb {P})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq18.png)
![$$\mathcal {X}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq19.png)
![$$\mathcal {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq20.png)
![$$\displaystyle \begin{aligned}\mathcal{M}^p_s := \left\{ \mu \in \mathcal{P}(\mathbb{R}^s) \; | \; \int_{\mathbb{R}^s} \|z\|{}^p~\mu(dz) < \infty \right\}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equh.png)
![$$\mathbb {R}^s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq21.png)
![$$\displaystyle \begin{aligned}\mathcal{M}^\infty_s := \left\{ \mu \in \mathcal{P}(\mathbb{R}^s) \; | \; \mbox{supp} \; \mu_Z \; \text{is bounded} \right\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equi.png)
Assume dom f ≠ ∅ and
for some p ∈ [1, ∞]. Then the mapping
given by F(x) := f(x, Z(⋅)) takes values in
and is Lipschitz continuous
with respect to the L
p-norm. △
![$$\displaystyle \begin{aligned} \begin{array}{rcl} \| F(x) \|{}_{L^p}^p &\displaystyle \leq 2^p |f(0,0)|{}^p + 2^p \int_{\mathbb{R}^s} |f(x,z) - f(0,0)|{}^p~\mu_Z(dz) \\ &\displaystyle \leq 2^p |f(0,0)|{}^p + 2^p L_f^p \|x\|{}^p + 2^p L_f^p \int_{\mathbb{R}^s} \|z\|{}^p~\mu_Z(dz) < \infty \end{array} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ4.png)
![$$\displaystyle \begin{aligned}\| F(x) - F(x') \|{}_{L^p} = \left( \int_{\mathbb{R}^s} |f(x,z) - f(x',z)|{}^p~\mu_Z(dz) \right)^{1/p} \leq L_f \|x - x' \|. \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equj.png)
![$$\mu _Z \in \mathcal {M}^\infty _s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq25.png)
![$$\displaystyle \begin{aligned}\|F(x)\|{}_{L^\infty} \leq \sup_{z \in \mbox{supp} \; \mu_Z} |f(x,z)| < \infty.\end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equk.png)
![$$\displaystyle \begin{aligned}\| F(x) - F(x') \|{}_{L^\infty} \leq \sup_{z \in \mbox{supp} \; \mu_Z} |f(x,z) - f(x',z)| \leq L_f \|x-x'\|.\end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equl.png)
The mapping in (17.2.1) can be used to measure the risk associated with the random variable
F(x).
![$$\mathcal {R}: \mathcal {X} \to \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq27.png)
![$$\mathcal {X}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq28.png)
![$$L^0(\Omega , \mathcal {F}, \mathbb {P})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq29.png)
- a.(Convexity) For any
and λ ∈ [0, 1] we have
- b.
(Monotonicity)
for all
satisfying Y 1 ≤ Y 2 with respect to the
-almost sure partial order.
- c.
(Translation equivariance)
for all
and
.
![$$\mathcal {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq37.png)
- d.
(Positive homogeneity)
for all
and t ∈ [0, ∞). △
A mapping is called law-invariant
if for all
with
we have
. △
Coherent risk measures have been introduced in [19], while the analysis of convex risk measures dates back to [20]. A thorough discussion of their analytical traits is provided in [21]. Below we list some risk measures that are commonly used in stochastic programming (cf. [1, Sect. 6.3.2]).
- (a)The expectation
,
is a law-invariant and coherent risk measure that turns (17.2.1) into the risk neutral bilevel stochastic program - (b)The expected excess of orderp ∈ [1, ∞) over a predefined level
is the mapping
given by
is law-invariant, convex and nondecreasing, but neither translation-equivariant nor positively homogeneous (cf. [1, Example 6.22]).
- (c)The mean upper semideviation of orderp ∈ [1, ∞) is the mapping
defined by
where ρ ∈ (0, 1] is a parameter.
is a law-invariant coherent risk measure (cf. [1, Example 6.20]).
- (d)The excess probability
over a prescribed target level
given by
is nondecreasing and law-invariant. However, it lacks convexity, translation-equivariance and positive homogeneity (cf. [22, Example 2.29]).
- (e)The Value-at-Risk
at level α ∈ (0, 1) defined by
is law-invariant, nondecreasing, translation-equivariant and positively homogeneous, but in general not convex (cf. [23]).
- (f)The Conditional Value-at-Risk
at level α ∈ (0, 1) given by
is a law-invariant coherent risk measure (cf. [23, Proposition 2]). The variational representation above was established in [24, Theorem 10].
- (g)The entropic risk measure
defined by
where α > 0 is a parameter, is a law-invariant convex (but not coherent) risk measure (cf. [21, Example 4.13, Example 4.34]).
- (h)The worst-case risk measure
given by
is law-invariant and coherent (cf. [21, Example 4.8]). This choice ofin (17.2.1) leads to the bilevel robust problem
Note that
only depends on the so called uncertainty set
. Thus, a bilevel robust problem can be formulated without knowledge of the distribution of the uncertain parameter. In robust optimization, the uncertainty set is often assumed to be finite, polyhedral or ellipsoidal (cf. [25]).
![$$L^p(\Omega , \mathcal {F}, \mathbb {P})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq59.png)
![$$\mathcal {R}: L^p(\Omega , \mathcal {F}, \mathbb {P}) \to \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq60.png)
![$$\mathbb {E} + \rho \cdot \mathcal {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq61.png)
![$$\displaystyle \begin{aligned}\min_x \left\{\mathbb{E}[F(x)] + \rho \cdot \mathcal{R}[F(x)] \; | \; x \in X \right\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equx.png)
![$$\displaystyle \begin{aligned}\min \left\{ \mathcal{R}[\min \Psi(x,Z)] \; | \; 1 \leq x \leq 6 \right\}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equy.png)
![$$\displaystyle \begin{aligned}\Psi(x,z) = \begin{cases} \lbrace x + 2 +z_1 \rbrace &\text{if} \; x \leq 3.25 + 0.5z_2 - 0.5z_1 \\ \lbrace -x + 8.5 + z_2 \rbrace &\text{else} \end{cases} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equz.png)
![$$\displaystyle \begin{aligned} \begin{array}{rcl} \mathbb{E}[\min \Psi(x,Z)] &\displaystyle = \int_{-0.5}^{0.5} \int_{-0.5}^{0.5} x + 2 + z_1~dz_1~dz_2 \\ &\displaystyle = x+2 \end{array} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ5.png)
![$$\displaystyle \begin{aligned} \begin{array}{rcl} \mathbb{E}[\min \Psi(x,Z)] &\displaystyle = \int_{-0.5}^{2x-6} \int_{-0.5}^{-2x + 6.5 + z_2} x + 2 + z_1~dz_1~dz_2 \\ &\displaystyle + \int_{2x-6}^{0.5} \int_{-0.5}^{0.5} x + 2 + z_1~dz_1~dz_2 \\ &\displaystyle +\int_{6-2x}^{0.5} \int_{-0.5}^{2x - 6.5 + z_1} -x + 8.5 + z_2~dz_2~dz_1 \\ &\displaystyle = -\frac{4}{3}x^3 + 11x^2 - \frac{117}{4} x + \frac{1427}{48} \end{array} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ6.png)
![$$\displaystyle \begin{aligned} \begin{array}{rcl} \mathbb{E}[\min \Psi(x,Z)] &\displaystyle = \int_{2x-7}^{0.5} \int_{-0.5}^{-2x + 6.5 + z_2} x + 2 + z_1~dz_1~dz_2 \\ &\displaystyle +\int_{-0.5}^{7-2x} \int_{-0.5}^{2x - 6.5 + z_1} -x + 8.5 + z_2~dz_2~dz_1 \\ &\displaystyle +\int_{7-2x}^{0.5} \int_{-0.5}^{0.5} -x + 8.5 + z_2~dz_2~dz_1 \\ &\displaystyle = \frac{4}{3} x^3 - 15 x^2 + \frac{221}{4} x - \frac{989}{16} \end{array} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ7.png)
![$$\displaystyle \begin{aligned} \begin{array}{rcl} \mathbb{E}[\min \Psi(x,Z)] &\displaystyle = \int_{-0.5}^{0.5} \int_{-0.5}^{0.5} -x + 8.5 + z_2~dz_2~dz_1 \\ &\displaystyle = -x + 8.5. \end{array} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ8.png)
![$$\mathbb {E}[\min \Psi (\cdot \,,Z)]$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq62.png)
![$$\displaystyle \begin{aligned}\min \left\{ \mathbb{E}[\min \Psi(x,Z)] \; | \; 1 \leq x \leq 6 \right\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equaa.png)
![$$\displaystyle \begin{aligned}\min \left\{ \mathcal{R}_{max}[\min \Psi(x,Z)] \; | \; 1 \leq x \leq 6 \right\}. \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equab.png)
![../images/480569_1_En_17_Chapter/480569_1_En_17_Fig1_HTML.png](../images/480569_1_En_17_Chapter/480569_1_En_17_Fig1_HTML.png)
The bold line depicts the graph of Ψ(⋅ , (0, 0)), while the dotted lines correspond to the graphs of Ψ(⋅ , (±0.5, ±0.5)) and Ψ(⋅ , (∓0.5, ±0.5))
17.2.3 Continuity and Differentiability
Continuity properties of carry over to Lipschitzian
properties of
,
.
![$$\mu _Z \in \mathcal {M}^p_s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq66.png)
![$$\mathcal {R}: L^p(\Omega , \mathcal {F}, \mathbb {P}) \to \mathbb {R}:$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq67.png)
- a.
is locally Lipschitz continuous if
is convex and continuous.
- b.
is locally Lipschitz continuous if
is convex and nondecreasing.
- c.
is locally Lipschitz continuous if
is a convex risk measure .
- d.
is Lipschitz continuous if
is Lipschitz continuous.
- e.
is Lipschitz continuous if
is a coherent risk measure.△
- a.
It is well-known that any real-valued convex and continuous mapping on a normed space is locally Lipschitz continuous (cf. [26]). The result is thus an immediate consequence of Lemma 17.2.4.
- b.
Any real-valued, convex and nondecreasing functional on the Banach lattice
is continuous (see e.g. [27, Theorem 4.1]).
- c.
By definition, any convex risk measure is convex and nondecreasing.
- d.
This is a straightforward conclusion from Lemma 17.2.4.
- e.
Any coherent risk measure on
is Lipschitz continuous by Inoue [28, Lemma 2.1].
Any coherent risk measure
is Lipschitz continuous
with constant 1 by Föllmer and Schied [21, Lemma 4.3]. Concrete Lipschitz constants for continuous coherent law-invariant risk measures
with p ∈ [1, ∞) may be obtained from representation results (see e.g. [29]). △
Proposition 17.2.8 allows to formulate sufficient conditions for the existence of optimal solutions to the bilevel stochastic linear program (17.2.1):
Assume dom f ≠ ∅,
for some p ∈ [1, ∞] and let X ⊆ P
Zbe nonempty and compact. Then (17.2.1) is solvable for any convex and nondecreasing mapping
. △
Due to the lack of convexity, Proposition 17.2.8 and the subsequent Corollary do not apply to the excess probability and the Value-at-Risk. However, invoking Lemma 17.2.1, the arguments used in the proof of [30, Proposition 3.3] can adapted to the setting of bilevel stochastic linear programming:
![$$\eta \in \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq84.png)
![$$\mathcal {Q}_{\mathit{\mbox{EP}}_\eta }$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq85.png)
![$$\displaystyle \begin{aligned}\mu_Z[\lbrace z \in \mathbb{R} \; | \; f(x,z) = \eta \rbrace] = 0. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equac.png)
![$$\displaystyle \begin{aligned}\min_x \left\{ \mathit{\mbox{EP}}_\eta [F(x)] \; | \; x \in X \right\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equad.png)
is solvable.△
has been analyzed in [17, Theorem 2]:
![$$\mathcal {Q}_{\mathit{\mbox{VaR}}_\alpha }$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq87.png)
![$$\displaystyle \begin{aligned}\min_x \left\{\mathit{\mbox{VaR}}_\alpha[F(x)] \; | \; x \in X \right\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equae.png)
is solvable.△
For specific risk measures, sufficient conditions for differentiability
of have been investigated in [31].
Assume dom f ≠ ∅ and that
is absolutely continuous with respect to the Lebesgue measure. Fix any
, then
and
are continuously differentiable at any x
0 ∈int P
Z. Furthermore, for any ρ ∈ [0, 1),
is continuously differentiable at any x
0 ∈int P
Zsatisfying
. △
Theorems 3.7, 3.8 and 3.9 in [31] provide more involved sufficient conditions for continuous differentiability of ,
and
that do not require μ
Z to be absolutely continuous.△
![$$\displaystyle \begin{aligned}y \leq x + 2 + z_1^{\prime}, \; y \leq -x + 8.5 + z_2^{\prime}, \; y \geq 1 + z_3^{\prime}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equaf.png)
![$$\mathbb {P} \circ Z^{\prime -1}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq98.png)
In the presence of differentiability, necessary optimality conditions for (17.2.1) can be formulated in terms of directional derivatives (cf. [31, Corollary 3.10]).
![$$\mu _Z \in \mathcal {M}^p_s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq99.png)
![$$\mathcal {Q}_{\mathcal {R}}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq100.png)
![$$\displaystyle \begin{aligned}\mathcal{Q}^{\prime}_{\mathcal{R}}(x_0)v \geq 0 \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equag.png)
![$$\displaystyle \begin{aligned}v \in \lbrace v \in \mathbb{R}^n \; | \; \exists \epsilon_0 > 0: \; x_0 + \epsilon v \in X \; \forall \epsilon \in [0,\epsilon_0] \rbrace. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equah.png)
△
17.2.4 Stability
While we have only considered as a functions of the leader’s decision x so far, it also depends on the underlying probability measure μ
Z. In stochastic programming, incomplete information about the true underlying distribution or the need for computational efficiency may lead to optimization models that employ an approximation of μ
Z. This section analysis deals with the behaviour of optimal values and (local) optimal solution sets of (17.2.1) under perturbations of the underlying distribution.
![$$\displaystyle \begin{aligned}P = \mathbb{R}^n \times \mathbb{R}^s \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equai.png)
![$$P = \mathbb {R}^n \times \mathbb {R}^s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq102.png)
holds if and only if u = 0 is the only non-negative solution to A
⊤u = 0. △
Throughout this section, we shall consider the situation that with p ∈ [1, ∞) is law-invariant, convex and nondecreasing. Furthermore, for the sake of notational simplicity (cf. [31, Remark 4.1]), we assume that the probability space
is atomless, i.e. for any
with
there exists some
with
and
.
![$$\mu \in \mathcal {M}^p_s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq111.png)
![$$(\delta _x \otimes \mu ) \circ f^{-1} \in \mathcal {M}^p_1$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq112.png)
![$$\delta _x \in \mathcal {P}(\mathbb {R}^n)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq113.png)
![$$(\Omega , \mathcal {F}, \mathbb {P})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq114.png)
![$$Y_{(x,\mu )} \in L^p(\Omega , \mathcal {F}, \mathbb {P})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq115.png)
![$$\mathbb {P} \circ Y_{(x,\mu )}^{-1} = (\delta _x \otimes \mu ) \circ f^{-1}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq116.png)
![$$\mathcal {Q}_{\mathcal {R}}: X \times \mathcal {M}^p_s \to \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq117.png)
![$$\displaystyle \begin{aligned}\mathcal{Q}_{\mathcal{R}}(x,\mu) := \mathcal{R}[Y_{(x,\mu)}]. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equaj.png)
![$$\mathcal {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq118.png)
![$$\displaystyle \begin{aligned} \min_x \lbrace \mathcal{Q}_{\mathcal{R}}(x,\mu) \; | \; x \in X \rbrace. \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ9.png)
![$$V \subseteq \mathbb {R}^n$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq119.png)
![$$\varphi _V: \mathcal {M}^p_s \to \overline {\mathbb {R}}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq120.png)
![$$\displaystyle \begin{aligned}\varphi_V(\mu) := \min_x \lbrace \mathcal{Q}_{\mathcal{R}}(x,\mu) \; | \; x \in X \cap \mbox{cl} \; V \rbrace, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equak.png)
![$$\phi _V: \mathcal {M}^p_s \rightrightarrows \mathbb {R}^n$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq121.png)
![$$\displaystyle \begin{aligned}\phi_V(\mu) := \underset{x}{\mbox{Argmin}} \lbrace \mathcal{Q}_{\mathcal{R}}(x,\mu) \; | \; x \in X \cap \mbox{cl} \; V \rbrace. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equal.png)
Given and an open set
, ϕ
V(μ) is called a complete local minimizing (CLM) set
of (Pμ
) w.r.t. V if ∅ ≠ ϕ
V(μ) ⊆ V . △
The set of global optimal solutions and any set of isolated minimizers are CLM sets. However, sets of strict local minimizers may fail to be CLM sets (cf. [33]). △
![$$\mathcal {P}(\mathbb {R}^s)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq125.png)
![$$\{\mu _l\}_{l \in \mathbb {N}} \subset \mathcal {P}(\mathbb {R}^s)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq126.png)
![$$\mu \in \mathcal {P}(\mathbb {R}^s)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq127.png)
![../images/480569_1_En_17_Chapter/480569_1_En_17_IEq128_HTML.gif](../images/480569_1_En_17_Chapter/480569_1_En_17_IEq128_HTML.gif)
![$$\displaystyle \begin{aligned} \lim_{l \to \infty} \int_{\mathbb{R}^s} h(t)~\mu_l(dt) = \int_{\mathbb{R}^s} h(t)~\mu(dt) \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equam.png)
![$$h:\mathbb {R}^s \to \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq129.png)
![$$\varphi _{\mathbb {R}^n}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq130.png)
![$$\mathcal {P}(\mathbb {R}^s)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq131.png)
![$$\displaystyle \begin{aligned}\min_x \left\{x + \int_{\mathbb{R}} z~\mu(dz) \; | \; 0 \leq x \leq 1 \right\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equan.png)
![$$\mathcal {R} = \mathbb {E}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq132.png)
![$$\Psi (x,z) = \{z\} \subsetneq \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq133.png)
![$$\displaystyle \begin{aligned}\min_x \{x \; | \; 0 \leq x \leq 1\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equao.png)
![$$\mu _l := (1 - \frac {1}{l}) \delta _{0} + \frac {1}{l} \delta _l$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq134.png)
![$$\displaystyle \begin{aligned}\min_x \left\{x + 1 \; | \; 0 \leq x \leq 1 \right\}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equap.png)
![$$l \in \mathbb {N}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq135.png)
We shall follow the approach of [22, 31] and [35] and confine the stability analysis to locally uniformly ∥⋅∥p-integrating sets.
![$$\mathcal {M} \subseteq \mathcal {M}^p_s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq136.png)
![$$\mathcal {N}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq137.png)
![$$\displaystyle \begin{aligned}\lim_{a \to \infty} \sup_{\nu \in \mathcal{M} \cap \mathcal{N}} \int_{\mathbb{R}^s \setminus a \mathbb{B}} \|z\|{}^p~\nu(dz) \leq \epsilon. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equaq.png)
△
A detailed discussion of locally uniformly ∥⋅∥p-integrating sets and their generalizations is provided in [21, 36, 37], and [38]. The following examples demonstrate the relevance of the concept.
- (a)Fix κ, 𝜖 > 0. Then by Föllmer and Schied [21, Corollary A.47 (c)], the set
of Borel probability measures with uniformly bounded moments of order p + 𝜖 is locally uniformly ∥⋅∥p-integrating.
- (b)
of Borel probability measures whose support is contained in Ξ is locally uniformly ∥⋅∥p-integrating.
The following result has been established in [31, Theorem 4.7]:
Assume dom f ≠ ∅ and
. Let
be locally uniformly ∥⋅∥p-integrating, then
- a.
is real-valued and weakly continuous.
- b.
is weakly upper semicontinuous.
In addition, assume that
is such that ϕ
V(μ
0) is a CLM set of
w.r.t. some open bounded set
. Then the following statements hold true:
- c.
is weakly continuous at μ 0.
- d.
is weakly upper semicontinuous at μ 0in the sense of Berge (cf. [ 39]), i.e. for any open set
with
there exists a weakly open neighborhood
of μ 0such that
for all
.
- e.
There exists some weakly open neighborhood
of μ 0such that ϕ V(μ) is a CLM set for (Pμ ) w.r.t. V for any
. △
Under the assumptions of Theorem 17.2.21d., any accumulation point x of a sequence local optimal solutions x
l ∈ ϕ
V(μ
l) as ,
, is a local optimal solution of (Pμ
). A detailed discussion of Berge’s notion of upper semicontinuity and related concepts is provided in [40, Chap. 5].△
As any Borel probability measure is the weak limit of a sequence of measures having finite support, Theorem 17.2.21 justifies an approach where the true underlying measure is approximated by a sequence of finite discrete ones. It is well known that approximation schemes based on discretization via empirical estimation [41, 42] or conditional expectations [43, 44] produce weakly converging sequences of discrete probability measures under mild assumptions.
17.2.5 Stochastic Dominance Constraints
![$$\displaystyle \begin{aligned}\min \lbrace f(x,Z(\cdot)) \; | \; x \in X \rbrace \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equav.png)
![$$g: \mathbb {R}^n \to \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq159.png)
![$$\displaystyle \begin{aligned} \min_x \left\{g(x) \; | \; x \in X,\; f(x, Z(\cdot)) \in \mathcal{A}\right\}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equaw.png)
![$$\mathcal {A} \subseteq f(X,Z) := \{f(x, Z(\cdot )) \; | \; x \in X\}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq160.png)
- (a)
is given by probabilistic constraints, i.e.
for bounds
and safety levels p 1, …, p l ∈ (0, 1).
- (b)
is given by first-order stochastic dominance constraints, i.e.
where
is a given benchmark variable. If b is discrete with a finite number of realizations, it is sufficient to impose the relation
for any β in a finite subset of
. In this case,
admits a description by a finite system of probabilistic constraints.
- (c)
is given by second-order stochastic dominance constraints, i.e.
where
is a given benchmark variable.
A discussion of general models involving probabilistic or stochastic dominance constraints can be found in [1, Chap. 8] and [45, Chap. 8.3].
![$$\nu := \mathbb {P} \circ b^{-1} \in \mathcal {P}(\mathbb {R})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq170.png)
![$$\displaystyle \begin{aligned}\left\{ x \in X \; | \; \mu_Z\big[ \lbrace z \in \mathbb{R}^s \; | \; f(x,z) \leq \beta \rbrace \big] \geq \nu\big[ \lbrace b \in \mathbb{R} \; | \; b \leq \beta \rbrace \big] \; \forall \beta \in \mathbb{R} \right\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equba.png)
![$$\mu \in \mathcal {M}^1_s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq171.png)
![$$\nu \in \mathcal {M}^1_1$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq172.png)
![$$\displaystyle \begin{aligned}\left\{ x \in X \, | \int_{\mathbb{R}^s} \max \lbrace f(x,z) - \eta, 0 \rbrace~\mu_Z(dz) \leq \int_{\mathbb{R}} \max \lbrace b - \eta, 0 \rbrace~\nu(db) \; \forall \eta \in \mathbb{N} \right\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbb.png)
![$$\mathcal {C}_1: \mathbb {P}(\mathbb {R}^s) \rightrightarrows \mathbb {R}^n$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq173.png)
![$$\displaystyle \begin{aligned}\mathcal{C}_1(\mu) = \left\{ x \in X \; | \; \mu \big[ \lbrace z \in \mathbb{R}^s \; | \; f(x,z) \leq \beta \rbrace \big] \geq \nu\big[ \lbrace b \in \mathbb{R} \; | \; b \leq \beta \rbrace \big] \; \forall \beta \in \mathbb{R} \right\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbc.png)
![$$\mathcal {C}_2: \mathcal {M}^1_s \rightrightarrows \mathbb {R}^n$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq174.png)
![$$\displaystyle \begin{aligned}C_2(\mu) := \left\{ x \in X \, | \int_{\mathbb{R}^s} \max \lbrace f(x,z) - \eta, 0 \rbrace~\mu(dz) \leq \int_{\mathbb{R}} \max \lbrace b - \eta, 0 \rbrace~\nu(db) \, \forall \eta \in \mathbb{N} \right\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbd.png)
Invoking Lemma 17.2.1, the following result can be obtained by adapting the proofs of [46, Proposition 2.1] and [47, Proposition 2.2] :
![$$P = \mathbb {R}^n \times \mathbb {R}^s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq175.png)
- a.
The multifunction C 1is closed w.r.t. the topology of weak convergence , i.e. for any sequences
and
with
,
for l →∞ and x l ∈ C 1(μ l) for all
it holds true that x ∈ C 1(μ).
- b.
Additionally assume that
, then the multifunction C 2is closed w.r.t. the topology of weak convergence.△
By considering the constant sequence μ
l = μ for all we obtain the closedness of the sets C
1(μ) and C
2(μ) under the conditions of Proposition 17.2.24. The closedness of the multifunctions C
1 and C
2 is also the key to proving the following stability result (cf. [46, Proposition 2.5]):
![$$P = \mathbb {R}^n \times \mathbb {R}^s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq183.png)
- a.The optimal value function
given by
is weakly lower semicontinuous on dom C 1.
- b.Additionally assume
, then the function
given by
is weakly lower semicontinuous on dom C 2. △
17.2.6 Finite Discrete Distributions
![$$Z_1, \ldots , Z_K \in \mathbb {R}^s$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq187.png)
![$$\displaystyle \begin{aligned}P_Z = \lbrace x \in \mathbb{R}^n \; | \; \forall k \in I \; \exists y \in \mathbb{R}^m: \; Ay \leq Tx + Z_k \rbrace. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbg.png)
![$$\lbrace y \in \mathbb {R}^m \; | \; Ay \leq Tx_0 + Z_k \rbrace = \emptyset $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq188.png)
In this setting, the bilevel stochastic linear problem can be reduced to a standard bilevel program, which allows to adapt optimality conditions and algorithms designed for the deterministic case (cf. [48]).
![../images/480569_1_En_17_Chapter/480569_1_En_17_IEq189_HTML.gif](../images/480569_1_En_17_Chapter/480569_1_En_17_IEq189_HTML.gif)
![../images/480569_1_En_17_Chapter/480569_1_En_17_IEq190_HTML.gif](../images/480569_1_En_17_Chapter/480569_1_En_17_IEq190_HTML.gif)
![$$ \mathcal {R}_{\max } \rbrace $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq191.png)
![$$\mathcal {R} \in \lbrace \mathit{\mbox{EP}}_{\eta }, \mathit{\mbox{VaR}}_{\alpha } \rbrace $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq192.png)
![$$\displaystyle \begin{aligned}\min_x \left\{ \mathcal{R}[F(x)] \; | \; x \in X \right\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbh.png)
![$$\displaystyle \begin{aligned}\min_x \Big\{\underbrace{\ \;\inf_{\eta \in \mathbb{R}}}_{\mathit{\text{if}}\ \mathcal{R} = \mathit{\mbox{CVaR}}_{\alpha}}\ \min_w \big\{a(x, w) \; | \; w \in \Psi_{\mathcal{R}}(x)\big\}| \; x \in X\Big\},\ \mathit{\text{or}} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbi.png)
![$$\displaystyle \begin{aligned}\min_x \Big\{c^\top x + \inf_{\eta \in \mathbb{R}}\big\{\eta \; | \; \min_w \{a(x, w) \; | \; w \in \Psi_{\mathcal{R}}(x)\} \geq \alpha\big\}| \; x \in X\Big\}\quad \mathit{\text{if}}\ \mathcal{R}= \mathit{\mbox{VaR}}_{\alpha}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbj.png)
![$$\Psi _{\mathcal {R}}: \mathbb {R}^n \rightrightarrows \mathbb {R}^{\mathit{\mbox{dim}}\;w}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq193.png)
![$$\displaystyle \begin{aligned}\Psi_{\mathcal{R}}(x) := \underset{w}{\mathit{\mbox{Argmin}}} \; \left\{\sum_{k \in I}d^\top y_k \; | \; Ay_k \leq Tx + Z_k,\;b_k(x, w) \geq 0\ \forall k \in I\right\}. \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbk.png)
Equivalent bilevel linear programs
| β | w | a(x, w) | b(x, w) |
---|---|---|---|---|
|
| c ⊤x+∑k ∈ Iπ kq ⊤y k | ||
|
|
| ∑k ∈ Iπ kv k |
|
| ρ ∈ (0, 1] |
|
|
|
EPη |
|
| ∑k ∈ Iπ kθ k | Mθ k−c ⊤x−q ⊤y k+η |
VaRα | α ∈ (0, 1) |
| ∑k ∈ Iπ kθ k |
|
CVaRα | α ∈ (0, 1) |
|
| cf. |
|
| maxk ∈ Ic ⊤x+q ⊤y k |
For , we refer to [31, Section 5].
![$$\mbox{EP}_{\eta }[F(x)] = \mathbb {P}\left [c^\top x + \inf _y \{q^\top y \; | \; y \in \Psi (x, Z(\cdot ))\} > \eta \right ]$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq215.png)
![$$M \in \mathbb {R}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq216.png)
![$$\displaystyle \begin{aligned} M > \sup\big\{c^\top x + \inf_{y_k} \big\{q^\top y_k \; | \; y_k \in \Psi(x, Z_k)\big\} \; | \; x \in X,\; k \in I\big\} - \eta\end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbl.png)
![$$\displaystyle \begin{aligned} \theta_k := \begin{cases} 0 & \text{if}\ c^\top x + q^\top y_k - \eta \leq 0,\\ 1 & \text{otherwise.} \end{cases} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbm.png)
![$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle \sum_{k \in I}\pi_k \inf_{y_k, \theta_k} \left\{\theta_k \; | \; M\theta_k \geq c^\top x + q^\top y_k - \eta,\; y_k \in \Psi(x, Z_k),\; \theta_k \in \{0, 1\}\right\}\\ &\displaystyle \left. \begin{aligned} = \inf_{\genfrac{}{}{0pt}{}{y_1, \ldots, y_K,}{\theta_1, \ldots, \theta_K}} \left\{\sum_{k \in I}\pi_k \theta_k \; | \; \right.&\displaystyle M\theta_k \geq c^\top x + q^\top y_k - \eta,\; y_k \in \Psi(x, Z_k),\; \theta_k \in \{0, 1\}\\ &\displaystyle \forall k \in I \end{aligned} \right\}. \end{array} \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ10.png)
![$$\mathcal {R} = \mbox{EP}_\eta $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq217.png)
![$$\mathbb {P}[f(x,Z(\cdot )) \leq \eta ]$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq218.png)
![$$\displaystyle \begin{aligned} &\sum_{k \in I}\pi_k \inf_{y_k, \theta_k} \left\{\theta_k \; | \; M(1 - \theta_k) \geq c^\top x + q^\top y_k - \eta,\; y_k \in \Psi(x, Z_k),\; \theta_k \in \{0, 1\} \right\}\\ &\left. \begin{aligned} = \left.\inf_{\genfrac{}{}{0pt}{}{y_1, \ldots, y_K,}{\theta_1, \ldots, \theta_K}} \right\{\sum_{k \in I}\pi_k \theta_k \; | \; &M(1 - \theta_i) \geq c^\top x + q^\top y_k - \eta,\; y_k \in \Psi(x, Z_k),\; \theta_k \in \{0, 1\} \\ &\forall k \in I \end{aligned} \right\}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbn.png)
![$$\displaystyle \begin{aligned}\theta_k := \begin{cases} 1 & \text{if}\ c^\top x + q^\top y_k - \eta \leq 0,\\ 0 & \text{otherwise.} \end{cases} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbo.png)
![$$\displaystyle \begin{aligned} \inf \left\{\eta \in \mathbb{R} \; | \; \inf_{\genfrac{}{}{0pt}{}{y_1, \ldots, y_K,}{\theta_1, \ldots, \theta_K}} \left\{\sum_{k \in I}\pi_k \theta_k \; | \; (y_1, \ldots, y_K, \theta_1, \ldots, \theta_K) \in \Psi_{\mbox{VaR}_{\eta}}(x) \right\} \geq \alpha\right\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbp.png)
![$$\sup _{k \in I} \left \{c^\top x + \min _{y_k} \{q^\top y \; | \; y {\in } \Psi (x, Z_k)\}\right \}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq219.png)
![$$\Psi _{\mathcal {R}_{\max }} = \Psi (x, Z_1) \times \ldots \times \Psi (x, Z_K)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq220.png)
- a.
The equivalent standard bilevel problem is linear provided that
.
- b.
Analogous to [31, Remarks 5.2, 5.4], the inner minimization problems of the standard bilevel linear programs for
can be decomposed into K scenario problems that only differ w.r.t. the right-hand side of the constraint system. For the other models, a similar decomposition is possible after Lagrangian relaxation of the coupling constraints involving different scenarios.
- c.
For
, every evaluation of the objective function in the standard bilevel linear program corresponds to solving a bilevel linear problem with scalar upper level variable η.
- d.
Alternate models for
are given in [17] and [49], where the considered bilevel stochastic linear problem is reduced to a mixed-integer nonlinear program and a mathematical programming problem with equilibrium constraints, respectively. A mean-risk model with
is used in [5, Sect. III].△
Similar reformulations can be obtained for the models discussed in Sect. 17.2.5 if we assume that the disutility function is linear.
![$$\displaystyle \begin{aligned} \min_x \left\{g^\top x \; | \; F(x) \in \mathcal{A},\; x \in X \right\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbq.png)
![$$\displaystyle \begin{aligned} \min_x \Big\{g^\top x \; | \; \inf_{w_j} \left\{a(w_j) \; |\; w_j \in \Psi_{\mathcal{R}}(x)\right\} \geq \delta_j \ \forall j = 1, \ldots, l,\; x \in X\Big\}. \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbr.png)
![$$\bar {a}_j := 1 - \mathbb {P}[b \leq a_j]$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq226.png)
![$$\tilde {a}_j := \int _{\mathbb {R}^s}\max \{b(z) - a_j, 0\}\;\mu _Z(dz)$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq227.png)
Equivalent programs, notation as in the Example in Sect. 17.2.5
| γ | w j | a(w j) |
| δ j |
---|---|---|---|---|---|
|
|
| ∑k ∈ Iπ kθ kj |
| p j |
|
|
| ∑k ∈ Iπ kθ kj |
|
|
|
|
| ∑k ∈ Iπ kv kj |
|
|
△
17.2.7 Solution Approaches
To solve bilevel problems, it is very common to use a single level reformulation. Often the lower level minimality condition is replaced by its Karush-Kuhn-Tucker or Fritz John conditions and the bilevel problem is reduced to a mathematical programming problem with equilibrium constraints (cf. [5, 17], [48, Chap. 3.5.1]).
![$$\mathcal {R} \in \lbrace \mathbb {E}, \mbox{EE}_\eta ^1, \mbox{SD}^1_\rho , \mbox{EP}_\eta , \mathcal {R}_{\max } \rbrace $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq244.png)
![$$\displaystyle \begin{aligned} \min_u \big\{g^\top u + \min_w \lbrace h^\top w \; | \; w \in \Psi(u) \rbrace \; | \; u \in U\big\}, \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ11.png)
![$$\Psi : \mathbb {R}^k \rightrightarrows \mathbb {R}^l$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq245.png)
![$$g \in \mathbb {R}^k$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq246.png)
![$$h, t \in \mathbb {R}^l$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq247.png)
![$$b \in \mathbb {R}^r$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq248.png)
![$$W \in \mathbb {R}^{r \times l}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq249.png)
![$$B \in \mathbb {R}^{r \times k}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq250.png)
![$$U \subseteq \mathbb {R}^k$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq251.png)
![$$\displaystyle \begin{aligned} \min_{u, w, v} \left\{ g^\top u + h^\top w \; \Bigg| \; \begin{aligned} &Ww \leq Bu + b, \; W^\top v = t, \; v \leq 0, \\ &v^\top (Ww - Bu - b) = 0, \; u \in U \end{aligned} \right\}. \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ12.png)
In [31, Sect. 6] it is also shown that is a local minimizer of the optimistic formulation, if
is an accumulation point of a sequence
of local minimizers of problem P(ε
n) for ε
n ↓ 0.
In the risk-neutral setting, problem (17.2.2) exhibits a block-structure (cf. Remark 17.2.27 b.). Adapting the solution method for general linear complementarity problems proposed in [50], this special structure has been used in [11, Chap. 6] to construct an efficient algorithm for the global resolution of bilevel stochastic linear problems based on dual decomposition.
Utilizing the lower level value function, problem (17.2.2) can be reformulated as a single level quasiconcave optimization problem (cf. [48, Chap. 3.6.5]). Solution methods based on a branch-and-bound scheme have been proposed in [51] and [52]. However, without modifications, these algorithms fail to exploit the block structure arising in risk-neutral bilevel stochastic linear optimization models (cf. [11, Chap. 4.2]). △
17.3 Two-Stage Stochastic Bilevel Programs
![../images/480569_1_En_17_Chapter/480569_1_En_17_Equbs_HTML.png](../images/480569_1_En_17_Chapter/480569_1_En_17_Equbs_HTML.png)
The bilevel stochastic linear problems considered in Sect. 17.2 can be understood as special two-stage bilevel programs, where the follower’s first-stage and the leader’s second stage decision do not influence the outcome. △
![$$\bar {\theta }$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq255.png)
![$$\displaystyle \begin{aligned}\text{``}\max_{x}\text{''} \left\{ \sum_{k \in K} x^\top y^k \; | \; (y,\bar{y}) \in \Psi(x) \right\}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbt.png)
![$$\displaystyle \begin{aligned}\Psi(x, c, d, b) := \underset{y,\bar{y}}{\mbox{Argmin}} \left\{ \sum_{k \in K} \left[ (c + x)^\top y^k + \bar{c}^\top \bar{y}^k \right] \; \Big| \; \begin{array}{l} y, \bar{y} \geq 0, \\ Ay^k + \bar{A} \bar{y}^k = b^k \; \forall k \in K \end{array} \right\}, \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbu.png)
![$$\bar {y}^k$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq256.png)
![$$\bar {c}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq257.png)
![$$\bar {\theta }$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq258.png)
![$$(A,\bar {A})$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_IEq259.png)
![$$\displaystyle \begin{aligned}b^k_i := \begin{cases} n^k, &\text{if} \; i = O(k) \\ -n^k, &\text{if} \; i = D(k) \\ 0, &\text{else} \end{cases} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbv.png)
![$$\displaystyle \begin{aligned} \text{``}\max_{x_1}\text{''} \left\{ \sum_{k \in K} x_1^\top y^k_1 + \mathbb{E}[\Phi(x_1, Z(\cdot))] \; | \; (y_1,\bar{y}_1) \in \Psi(x_1,c_1,d_1,b_1) \right\}, \end{aligned} $$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equ13.png)
![$$\displaystyle \begin{aligned}\Phi(x_1, Z(\omega)) := \text{``}\max_{x_2}\text{''} \left\{ \sum_{k \in K} x_2^\top y^k_2 \; | \; (x_1,x_2) \in \Gamma(\delta), \; (y_2,\bar{y}_2) \in \Psi(x_2,Z(\omega)) \right\} \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbw.png)
![$$\displaystyle \begin{aligned}\Gamma(\delta) := \Gamma_A(\delta) := \lbrace (x_1,x_2) \; | \; |x_{1, \theta} - x_{2, \theta}| \leq \delta_\theta \; \forall \theta \in \Theta \rbrace \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equbx.png)
![$$\displaystyle \begin{aligned}\Gamma(\delta) := \Gamma_R(\delta) := \lbrace (x_1,x_2) \; | \; |x_{1, \theta} - x_{2, \theta}| \leq \delta_\theta |x_{1, \theta}| \; \forall \theta \in \Theta \rbrace \end{aligned}$$](../images/480569_1_En_17_Chapter/480569_1_En_17_Chapter_TeX_Equby.png)
17.4 Challenges
We shall highlight some aspects of bilevel stochastic programming that are highly deserving of future research:
The first paper on bilevel stochastic programming has already outlined the basic principles as well as existence and sensitivity results for risk neutral models (cf. [2]). Nevertheless, so far, most of the research on bilevel stochastic nonlinear programming is still concerned with the risk-neutral case. Notable exceptions are [5] and [8], where models involving the Conditional Value-at-Risk are considered. In the first paper the problem of maximizing the medium-term revenue of an electricity retailer under uncertain pool prices, demand, and competitor prices is modeled as a bilevel stochastic quadratic problem, while the latter explores links between electricity swing option pricing and bilevel stochastic optimization. However, there exists no systematic analysis of bilevel stochastic nonlinear problems in the broader framework of coherent risk measures or higher stochastic dominance constraints. Future research may also consider distributionally robust models (cf. [54]).
Under finite discrete distributions many bilevel stochastic problems can be reformulated as standard bilevel programs. While this reformulation entails a blow-up of the dimension which is usually linear in the number of scenarios, the resulting problems often exhibit (quasi) block structures (cf. Remark 17.2.27b., [2]). For risk-neutral bilevel stochastic linear problems, [11, Chap. 6] utilizes these structures to enhance the mixed integer programming based solution algorithm of [50] resulting in a significant speed-up. Based on the structural similarities an analogous approach should be possible for risk-averse models after Lagrangian relaxation of coupling constraints.
While the analysis in the vast majority of papers on stochastic programming is confined to the case of purely exogenous stochasticity, this assumption is known to be unrealistic in economic models, where the decision maker holds market power. Therefore, models with decision dependent distributions are of particular interest in view of stochastic Stackelberg games (cf. [55]).
The second author thanks the Deutsche Forschungsgemeinschaft for its support via the Collaborative Research Center TRR 154.