1 Introduction

The main objective of this paper is to provide theoretical foundations of the theory of dynamic risk evaluation for Markov processes with only partial observation of the state and to derive the corresponding dynamic programming equations.

The theory of dynamic risk measures in discrete time has been intensively developed in the last 10 years (see Scandolo 2003; Riedel 2004; Roorda et al. 2005; Föllmer and Penner 2006; Cheridito et al. 2006; Ruszczyński and Shapiro 2006b; Artzner et al. 2007; Pflug and Römisch 2007; Klöppel and Schweizer 2007; Jobert and Rogers 2008; Cheridito and Kupper 2011 and the references therein). The basic setting is the following: we have a probability space \((\varOmega ,\mathscr {F},P)\), a filtration \(\{\mathscr {F}_t\}_{t=1,\ldots ,T}\) with a trivial \(\mathscr {F}_1\), and we define appropriate spaces \(\mathscr {Z}_t\) of \(\mathscr {F}_t\)-measurable random variables, \(t=1,\ldots ,T\). For each \(t=1,\ldots ,T\), a mapping \(\rho _{t,T}:\mathscr {Z}_T\rightarrow \mathscr {Z}_t\) is called a conditional risk measure. The central role in the theory is played by the concept of time consistency, which regulates relations between the mappings \(\rho _{t,T}\) and \(\rho _{s,T}\) for different s and t. One definition employed in the literature (Cheridito et al. 2006; Ruszczyński 2010) is the following: for all \(Z,W\in \mathscr {Z}_T\)if \(\rho _{t,T}(Z) \le \rho _{t,T}(W)\) then  \(\rho _{s,T}(Z) \le \rho _{s,T}(W)\) for all \(s<t\). This can be used to derive recursive relations \(\rho _{t,T}(Z) = \rho _t\big (\rho _{t+1,T}(Z)\big )\), with simpler one-step conditional risk mappings \(\rho _t:\mathscr {Z}_{t+1}\rightarrow \mathscr {Z}_t\), \(t=1,\ldots ,T-1\).

When applied to processes described by controlled kernels, like controlled Markov processes, the theory of dynamic measures of risk encounters difficulties. The domain and range spaces of the one-step mappings \(\rho _t\) change, when t increases. With \(\mathscr {Z}_t\) containing all \(\mathscr {F}_t\)-measurable random variables, arbitrary dependence of \(\rho _t\) on the past is allowed. These difficulties are compounded by the fact that a control policy changes the probability measure on the space of paths of the process. Risk measurement of the entire family of processes defined by control policies is needed.

In the extant literature, three basic approaches to introduce risk aversion in Markov decision processes have been employed: utility functions (see, e.g., Jaquette 1973; Denardo and Rothblum 1979; Bäuerle and Rieder 2013; Jaśkiewicz et al. 2013), mean-variance models (see, e.g., White 1988; Filar et al. 1989; Mannor and Tsitsiklis 2013; Arlotto et al. 2014), and entropic (exponential) models (see, e.g., Howard and Matheson 1971/72; Marcus et al. 1997; Coraluppi and Marcus 1999; Di Masi and Stettner 1999; Levitt and Ben-Israel 2001). Our approach complements the utility and exponential models; the mean-variance models do not satisfy, in general, the monotonicity and time consistency conditions, except the version of Chen et al. (2014).

In Ruszczyński (2010), we introduced Markov risk measures, in which the one-step conditional risk mappings \(\rho _t\) have a special form, which allowed for the development of dynamic programming equations and corresponding solution methods. This is related to the expected utility models with an aggregator, considered in Jaśkiewicz et al. (2013), but considers models which are nonlinear in probability (do not have an expected value representation). The aggegator is replaced by a more general transition risk mapping. Our ideas were successfully extended to transient models in Çavus and Ruszczyński (2014a, b) and Lin and Marcus (2013) and to problems with unbounded costs in Chu and Zhang (2014) and Shen et al. (2013). These works were further generalized in Fan and Ruszczyński (2016), where we introduced so-called process-based measures and described them by a sequence of transition risk mappings: static law invariant risk measures on a space \(\mathscr {V}\) of measurable functions on the state space \(\mathscr {X}\). In the special case of controlled Markov processes, we derived the structure postulated in Ruszczyński (2010).

In this paper, we develop risk theory for partially observable controlled Markov processes. In the expected-value case, this classical topic is covered in many monographs (see, e.g., Hinderer 1970; Bertsekas and Shreve 1978; Bäuerle and Rieder 2011 and the references therein). The standard approach is to consider the belief state: the conditional probability distribution of the unobserved part of the state. The recent article (Feinberg et al. 2016) provides the state-of-the-art setting. The risk-averse case has been dealt, so far, with the use of the entropic risk measure (James et al. 1994; Fernández-Gaucherand and Marcus 1997). A more general partially-observable utility model was recently analyzed in Bäuerle and Rieder (2017).

Our main result is that in partially observable systems the dynamic risk measures can be equivalently modeled by special forms of transition risk mappings: static risk measures on the space of functions defined on the observable part of the state only. We also derive dynamic programming equations for risk-averse partially observable Markov models. In these equations, the state space comprises belief states and observable states, as in the expected value model, but the conditional expectation is replaced by a transition risk mapping.

The paper is organized as follows. In Sect. 2, we briefly describe the partially observable Markov process and introduce relevant notation. In Sect. 3, we recall the concept of a dynamic risk measure and review its properties. We introduce the key property of stochastic dynamic time consistency and use it to derive a special structure of risk measures. In Sect. 4 we specialize these results to Markov systems with costs dependent on both components of the current state: observable and unobservable. In Sect. 5, we prove dynamic programming equations for risk-averse partially observable models. Finally, in Sect. 6, we illustrate our approach on a machine replacement problem.

2 Partially observable Markov decision processes

We consider a partially observable Markov process \(\{X_t,Y_t\}_{t=1,\ldots , T}\), in which \(\{X_t\}\) is observable and \(\{Y_t\}\) is not. We use the term “partially observable Markov decision process” (POMDP) in a more general way than the extant literature, because we consider dynamic risk measures of the cost sequence, rather than just the expected value.

In order to develop our subsequent theory, it is essential to define the model in a clear and rigorous way (cf.Bäuerle and Rieder 2011, Ch. 5). The state space of the model is defined as \(\mathscr {X}\times \mathscr {Y}\) where \((\mathscr {X}, \mathscr {B}(\mathscr {X}))\) and \((\mathscr {Y}, \mathscr {B}(\mathscr {Y}))\) are two Borel spaces (Borel subsets of Polish spaces). From the modeling perspective, \(x\in \mathscr {X}\) is the part of the state that we can observe at each step, while \(y\in \mathscr {Y}\) is unobservable. The measurable space that we will work with is then given by \(\varOmega =(\mathscr {X}\times \mathscr {Y})^T\) endowed with the canonical product \(\sigma \)-field \(\mathscr {F}\), and we use \(x_t\) and \(y_t\) to denote the canonical projections at time t.

Let \(\{\mathscr {F}^{X,Y}_t\}_{t=1,\ldots ,T}\) denote the natural filtration generated by the process (XY) and \(\{\mathscr {F}^X_t\}_{t=1,\ldots ,T}\) be the filtration generated by the process X.

The control space is given by a Borel space \(\mathscr {U}\), and since only \(\mathscr {X}\) is observable, the set of admissible controls at step t is given by a measurable multifunction \(\mathscr {U}_t : \mathscr {X}\rightrightarrows \mathscr {U}\) with nonempty values. The transition kernel at time t is \(K_t: \mathscr {X}\times \mathscr {Y}\times \mathscr {U}\rightarrow \mathscr {P}(\mathscr {X}\times \mathscr {Y})\), where \(\mathscr {P}(\mathscr {X}\times \mathscr {Y})\) is the space of probability measures on \(\mathscr {X}\times \mathscr {Y}\). In other words, if the state at time t is (xy) and we apply control u, the distribution of the next state is \(K_t(x,y,u)\).

At time t, the history of observed states is \(h_t = (x_1, x_2, \ldots , x_t)\), while all the information available for making a decision is \(g_t = (x_1, u_1, x_2, u_2, \ldots , x_t)\). We use \( \mathscr {H}_t = \mathscr {X}^t = \underbrace{\mathscr {X}\times \cdots \times \mathscr {X}}_{t\text { times}} \) to denote the spaces of possible histories \(h_t\). We make distinction of \(g_t\) and \(h_t\) because we should make decision of \(u_t\) based on \(g_t\) as the past controls \(u_1, \ldots , u_{t-1}\) are also taken into consideration when estimating the conditional distribution of \(Y_t\) (see Sect. 4). We write \(H_t\) for \((X_1,\ldots ,X_t)\).

For this controlled process, a (deterministic) history-dependent admissible policy \(\pi =(\pi _1,\ldots ,\pi _T)\) is a sequence of measurable functions \(\pi _t(g_t)\) such that \(\pi _t(g_t) \in \mathscr {U}_t(x_t)\) for all possible \(g_t\) [such a policy exists, due to the measurable selector theorem of Kuratowski and Ryll-Nardzewski (1965)]. We can easily prove by induction on t that for such an admissible policy \(\pi \), each \(\pi _t\) reduces to a measurable function of \(h_t=(x_1, x_2,\ldots , x_t)\), as \(u_s = \pi _s(x_1,\ldots ,x_s)\) for all \(s=1 ,\ldots , t-1\). We are still using \(\pi _s\) to denote the decision rule; it will not lead to any misunderstanding. Therefore the set of admissible policies is

$$\begin{aligned} \varPi = \ \big \{ \, \pi = (\pi _1,\ldots ,\pi _T) \, | \, \; \pi _t(x_1,\ldots ,x_t) \in \mathscr {U}_t(x_t),\ t=1,\ldots ,T \,\big \}. \end{aligned}$$

For a random \(Y_1\), any policy \(\pi \in \varPi \) defines a process \(\{X_t,Y_t,U_t\}_{t=1,\ldots ,T}\) on the probability space \((\varOmega , \mathscr {F}, P^\pi )\), with \(U_t=\pi _t(X_1,\ldots ,X_t)\).

We assume that the cost process \(Z^\pi _t\), \(t=1,\ldots ,T,\) is bounded and adapted to \(\{\mathscr {F}_t^{X}\}\), i.e., \(Z^\pi _t \in \mathscr {Z}_t\) for all \(\pi \) and t, where

$$\begin{aligned} \mathscr {Z}_t = \left\{ \, Z: \varOmega \rightarrow {\mathbb {R}} \, \big | \, Z \text { is } \mathscr {F}^{X}_t\text {-measurable and bounded} \right\} , \quad t=1,\ldots ,T. \end{aligned}$$

We allow the cost process to depend on the policy \(\pi \) in order to cover the case of control-dependent costs, such as \(Z_t^\pi =c_t(X_t,\pi _t(X_t))\), or more general cases discussed in section. For any \(Z \in \mathscr {Z}_t\), a measurable and bounded functional \(\overline{Z} : \mathscr {X}^t \rightarrow \mathbb {R}\) exists such that \(Z = \overline{Z}(X_1,\ldots ,X_t)\). With an abuse of notation, we still use Z to denote this function.

3 Risk measures for partially observable systems

3.1 Dynamic risk measures

For any policy \(\pi \in \varPi \), our objective is to evaluate at each time t the sequence of costs \(Z^\pi _t, \ldots , Z^\pi _T\) in such a way that the evaluation is \(\mathscr {F}^X_t\)-measurable. We denote \({\mathscr {Z}}_{t,T} = {\mathscr {Z}}_t \times \cdots \times {\mathscr {Z}}_T\), \(t=1,\ldots ,T\).

In what follows, all equality and inequality relations between random variables are understood in the “everywhere” sense.

Definition 1

A mapping \(\rho _{t,T}: {\mathscr {Z}}_{t,T} \rightarrow {\mathscr {Z}}_t\), where \(1 \le t \le T\), is called a conditional risk measure, if it satisfies the monotonicity property: for all \((Z_t,\ldots ,Z_T)\) and \((W_t,\ldots , W_T)\) in \({\mathscr {Z}}_{t,T}\), if \(Z_s \le W_s\) for all \(s=t, \ldots , T\), then \( \rho _{t,T}(Z_t,\ldots ,Z_T) \le \rho _{t,T}(W_t,\ldots ,W_T)\).

Definition 2

A conditional risk measure \(\rho _{t,T}: {\mathscr {Z}}_{t,T} \rightarrow {\mathscr {Z}}_t\)

  1. (i)

    is normalized if \(\rho _{t,T}(0,\ldots ,0)=0\);

  2. (ii)

    is translation invariant if \(\forall (Z_t,\ldots ,Z_T) \in {\mathscr {Z}}_{t,T}\), \(\qquad \rho _{t,T}(Z_t,\ldots ,Z_T)=Z_t+\rho _{t,T}(0,Z_{t+1},\ldots ,Z_T)\);

  3. (iii)

    has the local property if for any event \(A \in \mathscr {F}^X_t\) and all \((Z_t,\ldots ,Z_T) \in {\mathscr {Z}}_{t,T}\) we have \(\mathbb {1}_A \rho _{t,T}(Z_t,\ldots ,Z_T)=\rho _{t,T}(\mathbb {1}_A Z_t,\ldots ,\mathbb {1}_A Z_T).\)

From now on, we assume all conditional risk measures to be at least normalized.

Definition 3

A dynamic risk measure \(\big \{\rho _{t,T}\big \}_{t = 1, \ldots , T} \) is a sequence of conditional risk measures \(\rho _{t,T}: {\mathscr {Z}}_{t,T} \rightarrow {\mathscr {Z}}_t\). We say that it is normalized, translation-invariant, decomposable, or has the local property, if all \(\rho _{t,T}\), \(t=1,\ldots ,T\), satisfy the respective conditions of Definition 2.

3.2 Stochastic conditional time consistency

For a partially observable controlled process defined in Sect. 2, we have to use a family of risk measures \(\big \{\rho ^\pi _{t,T}\big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\), because the policy affects the probability measure on the space \(\varOmega \). When two policies \(\pi \) and \(\pi '\) are compared, even if the resulting costs were pointwise equal, \(\rho _{t,T}^{\pi }(Z_t,\ldots ,Z_T)\) and \(\rho _{t,T}^{\pi '}(Z_t,\ldots ,Z_T)\) may differ, because the probability measures \(P^\pi \) and \(P^{\pi '}\) could differ. The concept of stochastic conditional time consistency allows us to relate the whole family of risk measures.

Definition 4

A family of dynamic risk measures \(\big \{\rho ^\pi _{t,T}\big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\) is stochastically conditionally time consistent if for any \(\pi ,\pi ' \in \varPi \), for any \(1 \le t < T\), for all \(h_t \in \mathscr {X}^t\), all \((Z_{t+1},\ldots ,Z_T)\in \mathscr {Z}_{t+1,T}\) and all \( (W_{t+1},\ldots ,W_T)\in \mathscr {Z}_{t+1,T}\), the condition

$$\begin{aligned} \big ( \rho ^\pi _{t+1,T}(Z_{t+1},\ldots ,Z_T)\mid H^\pi _t=h_t \big ) \preceq _{\mathrm{st}} \big ( \rho ^{\pi '}_{t+1,T}(W_{t+1},\ldots ,W_T)\mid H_t^{\pi '}=h_t \big ), \end{aligned}$$

implies

$$\begin{aligned} \rho ^\pi _{t,T}(0,Z_{t+1},\ldots ,Z_T)(h_t) \le \rho ^{\pi '}_{t,T}(0,W_{t+1},\ldots ,W_T)(h_t). \end{aligned}$$

Remark 1

The conditional stochastic order “\(\preceq _{\mathrm{st}}\)” means that for all \(\eta \in \mathbb {R}\) we have

$$\begin{aligned}&\mathbb {P}^\pi \big [ \rho ^\pi _{t+1,T}(Z_{t+1},\ldots ,Z_T)(H_{t+1}) \le \eta \big | H_t = h_t \big ]\\&\quad \le \mathbb {P}^{\pi '} \big [ \rho ^{\pi '}_{t+1,T}(W_{t+1},\ldots ,W_T)(H_{t+1}) \le \eta \big | H_t = h_t \big ]. \end{aligned}$$

Proposition 1

If a family of dynamic risk measures \(\big \{\rho ^\pi _{t,T}\big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\) is normalized, has the translation property, and is stochastically conditionally time consistent, then it has the local property.

Proof

We use induction on t from T down to 1. At the final time, for all \(A \in \mathscr {F}_T\) and all \(Z_T\in \mathscr {Z}_T\) we have \(\rho ^\pi _{T,T}(\mathbb {1}_A Z_T) = \mathbb {1}_A Z_T = \mathbb {1}_A \rho ^\pi _{T,T}( Z_T)\).

Suppose \(\rho _{t+1,T}^\pi \) has the local property. Then for all \(A\in \mathscr {F}_t\), by translation,

$$\begin{aligned} \rho ^\pi _{t,T}(\mathbb {1}_A Z_t,\ldots ,\mathbb {1}_A Z_T) = \mathbb {1}_A Z_t + \rho ^\pi _{t,T}(0,\mathbb {1}_A Z_{t+1},\ldots ,\mathbb {1}_A Z_T) \end{aligned}$$

and

$$\begin{aligned} \mathbb {1}_A \rho ^\pi _{t,T}(Z_t,\ldots ,Z_T) = \mathbb {1}_A Z_t + \mathbb {1}_A \rho ^\pi _{t,T}(0,Z_{t+1},\ldots ,Z_T). \end{aligned}$$

To verify the local property of \(\rho ^\pi _{t,T}\) we need to show that both right hand sides are equal. For any \(h_t \in \mathscr {X}^t\) we have

$$\begin{aligned}{}[(\mathbb {1}_A \rho ^\pi _{t,T}(0,Z_{t+1},\ldots ,Z_T))](h_t) = \mathbb {1}_A(h_t) \rho ^\pi _{t,T}(0,Z_{t+1},\ldots ,Z_T))(h_t). \end{aligned}$$

The local property of \(\rho ^\pi _{t+1,T}\) yields

$$\begin{aligned} \rho ^\pi _{t+1,T}(\mathbb {1}_A Z_{t+1},\mathbb {1}_A Z_{t+2},\ldots ,\mathbb {1}_A Z_T) (h_t, \cdot ) = \mathbb {1}_A(h_t) \rho ^\pi _{t+1,T}(Z_{t+1},\ldots ,Z_T) (h_t,\cdot ), \end{aligned}$$

so by stochastic conditional time consistency,

$$\begin{aligned} \rho ^\pi _{t,T}(0,\mathbb {1}_A Z_{t+1},\ldots ,\mathbb {1}_A Z_T)(h_t) = {\left\{ \begin{array}{ll} 0 &{} \; \text {if } \mathbb {1}_A(h_t) = 0,\\ \rho ^\pi _{t,T}(0, Z_{t+1},\ldots , Z_T)(h_t) &{} \; \text {if } \mathbb {1}_A(h_t) = 1. \end{array}\right. } \end{aligned}$$

Thus,

$$\begin{aligned} \rho ^\pi _{t,T}(0,\mathbb {1}_A Z_{t+1},\ldots ,\mathbb {1}_A Z_T)(h_t) = \mathbb {1}_A \rho ^\pi _{t,T}(0,Z_{t+1},\ldots ,Z_T)(h_t), \; \forall h_t \in \mathscr {X}^t, \end{aligned}$$

which proves the local property of \(\rho ^\pi _{t,T}\). \(\square \)

The following theorem shows that the stochastic conditional time consistency implies that one-step risk mappings can be represented by static law-invariant risk measures on \(\mathscr {V}\), the set of all bounded measurable functions on \(\mathscr {X}\). We first recall the definition of a risk measure and slightly refine the standard concept of law invariance.

Definition 5

A measurable functional \(r:\mathscr {V}\rightarrow \mathbb {R}\) is called a risk measure.

  1. (i)

    It is monotonic, if \(V \le W\) implies \(r(V) \le r(W)\);

  2. (ii)

    It is normalized if \(r(0)=0\);

  3. (iii)

    It is translation invariant if for all \(V \in \mathscr {V}\) and all \(a\in \mathbb {R}\), \(r(a+V)=a + r(V)\);

  4. (iv)

    It is law invariant with respect to the probability measure q on \((\mathscr {X},\mathscr {B}(\mathscr {X}))\), if , where means that \(q\{ V \le \eta \} = q\{ W \le \eta \}\) for all \(\eta \in \mathbb {R}\).

The conditional distribution of \(\rho ^\pi _{t+1,T}(Z_{t+1},\ldots ,Z_T)(H_{t+1})\) given \(H_t=h_t\) under \(P^\pi \) plays an important role in the stochastic conditional time consistency, so does the conditional distribution of \(X_{t+1}\), given \(h_t\). We denote the latter by \(Q_t^\pi (h_t) \in \mathscr {P}(\mathscr {X})\):

$$\begin{aligned} Q_t^\pi (h_t)(C) = \mathbb {P}^\pi [X_{t+1} \in C \, | \, H_t = h_t],\quad \forall C \in \mathscr {B}(\mathscr {X}). \end{aligned}$$
(1)

Later in Sect. 4 we show that \(Q_t^\pi \) can be computed in a recursive way with the help of belief states and Bayes operators. We can now state the main result of this section.

Theorem 1

A family of dynamic risk measures \(\big \{\rho _{t,T}^\pi \big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\) is normalized, translation invariant, and stochastically conditionally time consistent if and only if transition risk mappings

$$\begin{aligned} \sigma _t: \bigg \{\bigcup _{\pi \in \varPi } \mathrm{graph}(Q_t^\pi )\bigg \} \times \mathscr {V}\rightarrow {\mathbb {R}}, \quad t=1\ldots T-1, \end{aligned}$$

exist, such that

  1. (i)

    For all \(t=1 \ldots T-1\) and all \(h_t \in \mathscr {X}^t\), \(\sigma _t(h_t,\cdot ,\cdot )\) is normalized and has the following property of strong monotonicity with respect to stochastic dominance:

    $$\begin{aligned}&\forall q^1,q^2 \in \big \{ \, Q_t^\pi (h_t):\pi \in \varPi \, \big \}, \; \forall \, V^1,V^2 \in \mathscr {V}, \;\\&(V^1;q^1) \preceq _{\mathrm{st}} (V^2;q^2) \implies \sigma _t(h_t,q^1,V^1) \le \sigma _t(h_t,q^2,V^2), \end{aligned}$$

    where \((V;q) = q \circ V^{-1}\) means “the distribution of V under q;”

  2. (ii)

    For all \(\pi \in \varPi \), for all \(t=1 \ldots T-1\), for all \((Z_t,\ldots ,Z_T)\in {\mathscr {Z}}_{t,T}\), and for all \(h_t \in \mathscr {X}^t\),

    $$\begin{aligned} \rho ^\pi _{t,T}(Z_t,Z_{t+1},\ldots ,Z_T)(h_t) = Z_t + \sigma _t(h_t,Q_t^\pi (h_t),\rho _{t+1,T}^\pi (Z_{t+1},\ldots ,Z_T) (h_t,\cdot )). \end{aligned}$$
    (2)

Moreover, for all \(t=1 \ldots T-1\), \(\sigma _t\) is uniquely determined by \(\{\rho _{t,T}^\pi \}^{\pi \in \varPi }\) as follows: for every \(h_t \in \mathscr {X}^t\), for every \(q \in \big \{ \, Q_t^\pi (h_t):\pi \in \varPi \, \big \}\), and for every \(V\in \mathscr {V}\),

$$\begin{aligned} \sigma _t(h_t,q,V)=\rho _{t,T}^\pi (0,v,0,\ldots ,0)(h_t), \end{aligned}$$

where \(\pi \) is any admissible policy such that \(q=Q_t^\pi (h_t)\), and \(v\in \mathscr {S}_{t+1}\) satisfies the equation \(v(h_t,\cdot ) = V(\cdot )\), and can be arbitrary elsewhere.

Proof

Assume \(\big \{\rho _{t,T}^\pi \big \}^{\pi \in \varPi }_{t=1,\ldots ,T}\) is translation invariant and stochastically conditionally time consistent. For any \(V\in \mathscr {V}\) and any \(h_t\in \mathscr {X}^t\) we define \(v(h_t,\cdot ) = V(\cdot )\). The function v is an element of \(\mathscr {S}_{t+1}\). Then the formula \( \sigma _t^\pi (h_t,q,V)=\rho _{t,T}^\pi (0,v,0,\ldots ,0)(h_t), \) defines for each \(\pi \) a normalized and monotonic risk measure on the space \(\mathscr {V}\). For any \((Z_t,\ldots ,Z_T)\in {\mathscr {Z}}_{t,T}\), setting

$$\begin{aligned} W(x)&= \rho _{t+1,T}^\pi (Z_{t+1},\ldots ,Z_T)(h_t,x), \quad \forall \,x\in \mathscr {X},\\ w(h_{t+1})&= {\left\{ \begin{array}{ll} W(x), &{} \text {if } h_{t+1} = (h_t,x),\\ 0, &{} \text {otherwise}, \end{array}\right. } \end{aligned}$$

we obtain, by translation invariance and normalization,

$$\begin{aligned} \rho _{t+1,T}^\pi (w,0,\ldots ,0)(h_t,\cdot ) = w(h_t,\cdot ) = \rho _{t+1,T}^\pi (Z_{t+1},\ldots ,Z_T)(h_t,\cdot ). \end{aligned}$$

Thus, by translation invariance and stochastic conditional time consistency,

$$\begin{aligned}&\rho _{t,T}^\pi (Z_t,\ldots ,Z_T)(h_t) = Z_t(h_t)+\rho _{t,T}^\pi (0,Z_{t+1},\ldots ,Z_T)(h_t)\nonumber \\&\quad = Z_t(h_t)+\rho _{t,T}^\pi (0,w,0,\ldots ,0)(h_t) = Z_t(h_t)+\sigma _t^\pi (h_t,q,W). \end{aligned}$$
(3)

This chain of relations proves also the uniqueness of \(\sigma _t^\pi \) for each \(\pi \).

We can now verify the strong monotonicity of \(\sigma _t^\pi (h_t,\cdot ,\cdot )\) with respect to stochastic dominance. Suppose

$$\begin{aligned} (V^1;Q_t^{\pi _1}(h_t)) \preceq _{\mathrm{st}} (V^2;Q_t^{\pi _2}(h_t)), \end{aligned}$$
(4)

where \(V^1,V^2\in \mathscr {V}\) and \(h_t\in \mathscr {X}^t\). Define \(v^1(h_t,\cdot ) = V^1(\cdot )\) and \(v^2(h_t,\cdot ) = V^2(\cdot )\). Then Definition 4 implies that \(\rho ^{\pi _1}_{t,T}(0,v^1,0,\ldots ,0)(h_t) \le \rho ^{\pi _2}_{t,T}(0,v^2,0,\ldots ,0)(h_t)\). This combined with (3) yields

$$\begin{aligned} \sigma _t^{\pi _1}(h_t,Q_t^{\pi _1}(h_t),V^1) \le \sigma _t^{\pi _2}(h_t,Q_t^{\pi _2}(h_t),V^2). \end{aligned}$$
(5)

Suppose \(Q_t^{\pi _1}(h_t) = Q_t^{\pi _2}(h_t)\) and \(V^1=V^2\). Then both \(\preceq _{\mathrm{st}}\) and \(\succeq _{\mathrm{st}}\) are true in (4) and thus (5) becomes an equality. This proves that in fact \(\sigma _t^\pi \) does not depend on \(\pi \), and all dependence on \(\pi \) is carried by the controlled kernel \(Q_t^\pi \). Moreover, the function \(\sigma _t(h_t,\cdot ,\cdot )\) is indeed strongly monotonic with respect to stochastic dominance.

On the other hand, if such transition risk mappings \(\sigma _t\) exist, then \(\big \{\rho _{t,T}\big \}^{\pi \in \varPi }_{t=1,\ldots ,T}\) is stochastically conditionally time consistent by the monotonicity and law invariance of \(\sigma _t(h_t,\cdot ,\cdot )\). We can now use (2) to obtain for any \(t=1,\ldots ,T-1\), and for all \(h_t\in \mathscr {X}^t\) the translation invariance of \(\rho _{t,T}^\pi \). \(\square \)

The following transition risk mappings satisfy the condition of Theorem 1 and correspond to stochastically conditionally time consistent risk measures.

Example 1

The entropic transition risk mapping,

$$\begin{aligned} \sigma _t(h_t,q,v) = \frac{1}{\gamma }\ln \big ( \mathbb {E}_q[e^{\gamma v}]\big ) = \frac{1}{\gamma }\ln \bigg ( \int _\mathscr {X}e^{\gamma v(x)}\;q(dx)\bigg ),\quad \gamma >0. \end{aligned}$$

We could make \(\gamma \) dependent on the time t, the current state \(x_t\), or the entire history \(h_t\), and still obtain a stochastically conditionally time consistent risk measure (with \(q=Q_t(h_t)\)).

Example 2

The mean-semideviation transition risk mapping:

$$\begin{aligned} \sigma _t(h_t,q,v) = \int _{\mathscr {X}} v(s)\;q(ds) + \varkappa _t(h_t) \left( \int _{\mathscr {X}} \Big [\Big ( v(s)- \int _{\mathscr {X}} v(s')\;q(ds')\Big )_+\Big ]^p\;q(ds)\right) ^{1/p}, \end{aligned}$$

where \(\varkappa _t:\mathscr {X}^t\rightarrow [0,1]\) is a measurable function, and \(p\in [1,+\infty )\). It is an analogue of the static mean-semideviation measure of risk, whose consistency with stochastic dominance is well-known (Ogryczak and Ruszczyński 1999, 2001). In the risk measure, we use \(q=Q_t(h_t)\).

In fact, all coherent or convex law invariant risk measures may be used to construct \(\sigma _t\); just the dependence on the probability measure q must be explicit.

4 Markov models

4.1 Bayes operator

At each time t, the conditional distribution of the next observable state \(Q_t^\pi (h_t)\) defined in (1) can be easily computed if we know the conditional distribution of the current unobservable state, called the belief state:

$$\begin{aligned} \varXi _t^\pi (h_t) \in \mathscr {P}(\mathscr {Y}): \; \varXi _t^\pi (h_t)(D) = \mathbb {P}^\pi [Y_t \in D \, | \, H_t = h_t], \quad \forall D \in \mathscr {B}(\mathscr {Y}), \end{aligned}$$
(6)

as we have

$$\begin{aligned} Q_t^\pi (h_t) = \int _{\mathscr {Y}} K_t^X(x_t, y ,\pi _t(h_t)) \; \varXi _t^\pi (h_t)(dy), \end{aligned}$$
(7)

where \(K_t^X(x_t, y ,\pi _t(h_t))\) is the marginal distribution of \(K_t(x_t, y ,\pi _t(h_t))\) on \(\mathscr {X}\).

In a POMDP, the Bayes operator provides a way to update from prior belief to posterior belief. Suppose the current state observation is x, the action is u, and the conditional distribution of the unobservable state, given the history of the process, is \(\xi \). After a new observation \(x'\) of the observable part of the state, we can find a formula to determine the posterior distribution of the unobservable state.

Let us start with a fairly general construction of the Bayes operator. Assuming the above setup, for given \((x, \xi , u) \in \mathscr {X}\times \mathscr {P}(\mathscr {Y}) \times \mathscr {U}\), define a new measure \(m_t(x,\xi ,u)\) on \(\mathscr {X}\times \mathscr {Y}\), initially on all measurable rectangles \(A \times B\), as

$$\begin{aligned} m_t(x,\xi ,u) (A \times B) = \int _\mathscr {Y}K_t(A \times B \,\vert \, x, y, u) \;\xi (dy). \end{aligned}$$

We verify readily that this uniquely defines a probability measure on \(\mathscr {X}\times \mathscr {Y}\). If the measurable space \((\mathscr {Y}, \mathscr {B}(\mathscr {Y}))\) is standard Borel, i.e., isomorphic to a Borel subspace of \(\mathbb {R}\), we can disintegrate \(m_t(x,\xi ,u)\) into its marginal \(\lambda _t(x,\xi ,u) (dx')\) on \(\mathscr {X}\) and a transition kernel \(\varGamma _t(x,\xi ,u)(x', dy')\) from \(\mathscr {X}\) to \(\mathscr {Y}\):

$$\begin{aligned} m_t(x,\xi ,u) (dx', dy') = \lambda _t(x,\xi ,u) (dx') \, \varGamma _t(x,\xi ,u)(x', dy'). \end{aligned}$$

For all \(C \in \mathscr {B}(\mathscr {Y})\), we define the Bayes operator of the POMDP as follows:

$$\begin{aligned} \varPhi _t(x, \xi , u, x')(C) = \varGamma _t(x,\xi ,u)(x', C). \end{aligned}$$

The above argument shows that the Bayes operator exists and is unique as long as the space \(\mathscr {Y}\) is standard Borel, which is almost always the case in applications of POMDP. In the following, we always assume the existence of the Bayes operator.

Example 3

Assume that each transition kernel \(K_t(x,y,u)\) has a density \(q_t(\cdot , \cdot \mid x,y,u)\) with respect to a finite product measure \(\mu _X \otimes \mu _Y\) on \(\mathscr {X}\times \mathscr {Y}\). Then the Bayes operator has the form

$$\begin{aligned} \big [ \varPhi _t(x,\xi ,u,x')\big ] (A) = \frac{\int _A \int _\mathscr {Y}q_t(x', y' \mid x,y,u)\; \xi (dy) \; \mu _Y(dy')}{\int _\mathscr {Y}\int _\mathscr {Y}q_t(x', y' \mid x,y,u)\; \xi (dy) \; \mu _Y(dy')},\quad \forall \, A \in \mathscr {B}(\mathscr {Y}). \end{aligned}$$

If the formula above has a zero denominator for some \((x,\xi ,u,x')\), we can formally define \(\varPhi _t(x,\xi ,u,x')\) to be an arbitrarily selected distribution on \(\mathscr {Y}\).

Thus, we can calculate \(Q_t\) and \(\varXi _t\) (defined in (1) and (6)) recursively with the help of Bayes operators:

  • The initial belief \(\varXi _1(x_1)\) is the conditional distribution of \(Y_1\) given \(X_1=x_1\);

  • The Bayes operator provides us the following formula to update the belief states:

    $$\begin{aligned} \varXi _{t+1}^\pi (h_{t+1}) = \varPhi _t \big ( x_t,\varXi _{t}^\pi (h_{t}),\pi _t(h_t), x_{t+1} \big ), \end{aligned}$$

    and, by induction on t, \( \varXi _t^\pi (h_t) = \varXi _t \big ( x_1,\pi _1(x_1),\ldots ,x_{t-1},\pi _{t-1}(h_{t-1}),x_t \big )\);

  • The conditional distribution of \(X_{t+1}\) can be calculated by (7), and

    $$\begin{aligned} Q_t^\pi (h_t) = Q_t\big ( x_1,\pi _1(x_1),\ldots ,x_t,\pi _t(h_t) \big ). \end{aligned}$$

4.2 Markov risk measures

We make the following additional assumptions:

Assumption 1

The costs \(Z_1,\ldots ,Z_T\) are only dependent on the current observable state, the current belief state, and the current control, that is,

$$\begin{aligned} Z_t^\pi (h_t) = r_t(x_t, \varXi _t^\pi (h_t), \pi _t(h_t)),\quad t=1,\ldots ,T. \end{aligned}$$
(8)

For example, in expected value models, we have \(r_t(x,\xi ,u) = \int _\mathscr {Y}c_t(x,y,u)\,\xi (dy)\), where \(c_t:\mathscr {X}\times \mathscr {Y}\times \mathscr {U}\rightarrow \mathbb {R}\) is the running cost function, but more general functionals can be used here instead of the expectation with respect to the belief state.

Definition 6

In POMDP, a policy \(\pi \in \varPi \) is Markov if \(\pi _t(h_t)=\pi _t(h_t')\) for all \(t = 1 ,\ldots , T\) and all \(h_t, \, h_t' \in \mathscr {X}^t\) such that \(x_t=x_t'\) and \(\varXi _t^\pi (h_t)=\varXi _t^\pi (h_t')\).

For a fixed Markov policy \(\pi \), the future evolution of the process \(\{(X_\tau ,\varXi _\tau ^\pi )\}_{\tau =t ,\ldots , T}\) is solely dependent on the current \((x_t,\varXi _t^\pi (h_t))\), and so is the distribution of the future risk functions \(r_\tau (X_\tau ,\varXi _\tau ^\pi ,\pi _\tau (X_\tau ,\varXi _\tau ^\pi ))\), \(\tau =t,\ldots ,T\). Therefore, we can define the Markov property of risk measures for POMDP. To alleviate notation, for all \(\pi \in \varPi \) and for a measurable and bounded \(r=(r_1, \ldots , r_T)\), we write

$$\begin{aligned} v_t^{\pi }(h_t) := \rho ^\pi _{t,T}\big (r_t(X_t,\varXi _t^\pi ,\pi _t(H_t)), \ldots ,r_T(X_T,\varXi _T^\pi ,\pi _T(H_T))\big )(h_t). \end{aligned}$$
(9)

Definition 7

A family of dynamic risk measures \(\big \{\rho _{t,T}^\pi \big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\) for a POMDP is Markov if for all Markov policies \(\pi \in \varPi \), for all bounded measurable \(r=(r_1,\ldots ,r_T)\), and for all \(h_t=(x_1,\ldots ,x_t)\) and \(h_t'=(x_1',\ldots ,x_t')\) in \(\mathscr {X}^t\) such that \(x_t=x_t'\) and \(\varXi _t^\pi (h_t)=\varXi _t^\pi (h_t')\), we have

$$\begin{aligned} v^{\pi }_t(h_t)=v^{\pi }_t(h_t'). \end{aligned}$$

Proposition 2

A normalized, translation invariant, and stochastically conditionally time consistent family of risk measures \(\big \{\rho _{t,T}^\pi \big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\) is Markov if and only if the dependence of \(\sigma _t\) on \(h_t\) is carried by \((x_t,\varXi _t^\pi (h_t))\) only, for all \(t=1 ,\ldots , T-1\).

Proof

Fix \(t = 1,\ldots ,T-1\) and \(W\in \mathscr {V}\). Let \(\pi \in \varPi \) be an arbitrary policy. Consider \(h_t,h_t' \in \mathscr {X}^t\) such that \(x_t=x_t'\), \(\varXi _t^\pi (h_t)=\varXi _t^\pi (h_t')=\xi _t\), and \(Q_t^\pi (h_t)=Q_t^\pi (h_t')\). We construct a Markov policy \(\lambda \in \varPi \), such that \(\pi _t(h_t)=\lambda _t(x_t,\xi _t)\) for this particular t and \(h_t\). By construction, \(Q_t^\pi (h_t) = Q_t^\lambda (x_t,\xi _t)\). We also construct a sequence of costs \(r=(0,\ldots , 0,r_{t+1}, 0 ,\ldots , 0)\) with \(r_{t+1}(x',\xi ',u')\equiv W(x')\). We obtain the following chain of equations, in which we use the construction of \(\lambda \), the construction of r, the Markov property of Definition 7, and the assumed equality of \(Q_t^\pi (h_t)\) and \(Q_t^\pi (h_t')\):

$$\begin{aligned} \sigma _t(h_t,Q_t^\pi (h_t),W)&=\sigma _t(h_t,Q_t^\lambda (x_t,\xi _t),W)=v^{\lambda }_t(h_t) =v^{\lambda }_t(h_t')\\&= \sigma _t(h_t',Q_t^\lambda (x_t,\xi _t),W) = \sigma _t(h_t',Q_t^\pi (h_t),W)\\&= \sigma _t(h_t',Q_t^\pi (h_t'),W). \end{aligned}$$

Therefore, \(\sigma _t\) is its direct dependence on \(h_t\) is carried by \((x_t,\xi _t)\) only.

If \(\sigma _t, \, t=1 ,\ldots , T-1\), are all memoryless, we can prove by induction backward in time that for all \(t=T ,\ldots , 1\), \(v^{\pi }_t(h_t)=v^{\pi }_t(h_t')\) for all Markov \(\pi \) and all \(h_t,h_t' \in \mathscr {X}^t\) such that \(x_t=x_t'\) and \(\xi _t=\xi _t'\). \(\square \)

The following theorem summarizes our observations.

Theorem 2

A family of dynamic risk measures \(\big \{\rho _{t,T}^\pi \big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\) for a POMDP is normalized, translation-invariant, stochastically conditionally time consistent, and Markov if and only if transition risk mappings

$$\begin{aligned} \sigma _t: \left\{ \big (x_t, \varXi ^\pi _t(h_t), Q_t^\pi (h_t)\big ) : \pi \in \varPi , \, h_t \in \mathscr {X}^t \right\} \times \mathscr {V}\rightarrow {\mathbb {R}}, \quad t=1\ldots T-1, \end{aligned}$$

exist, such that

  1. (i)

    for all \(t=1,\ldots ,T-1\) and all \((x,\xi ) \in \left\{ \big (x_t, \varXi ^\pi _t(h_t) \big ) : \pi \in \varPi , \, h_t \in \mathscr {X}^t \right\} \), \(\sigma _t(x,\xi ,\cdot ,\cdot )\) is normalized and strongly monotonic with respect to stochastic dominance on \(\big \{ \, Q_t^\pi (h_t): \pi \in \varPi , \, h_t \in \mathscr {X}^t \text { such that } x_t=x, \, \varXi _t^\pi (h_t)=\xi \, \big \}\);

  2. (ii)

    for all \(\pi \in \varPi \), for all measurable bounded r, for all \(t=1,\ldots ,T-1\), and for all \(h_t \in \mathscr {X}^t\),

    $$\begin{aligned} v^{\pi }_t(h_t)=r_t(x_t,\varXi ^\pi _t(h_t),\pi _t(h_t))+\sigma _t\big (x_t, \varXi ^\pi _t(h_t),Q_t^\pi (h_t),v^{\pi }_{t+1}(h_t,\cdot )\big ). \end{aligned}$$
    (10)

This allows us to evaluate risk of Markov policies in a recursive way.

Corollary 1

Under the conditions of Theorem 2, for any Markov policy \(\pi \), the function (9) depends on r, \(\pi _t\),...,\(\pi _{T}\), and \((x_t,\xi _t)\) only, and the following relation is true:

$$\begin{aligned} v^{\pi _t,\ldots ,\pi _T}_t(x_t,\xi _t)= & {} r_t(x_t,\xi _t,\pi _t(x_t,\xi _t))\nonumber \\&+\sigma _t\big (x_t,\xi _t,\int _\mathscr {Y}K_t^X(x_t,y,\pi _t(x_t,\xi _t))\;\xi _t(dy),x'\nonumber \\&\mapsto v^{\pi _{t+1},\ldots ,\pi _T}_{t+1}(x',\varPhi _t(x_t,\xi _t,\pi _t(x_t,\xi _t),x'))\big ). \end{aligned}$$
(11)

Proof

We use induction backward in time. For \(t=T\) we have \(v^{\pi }_T(h_T)= r_T(x_T,\xi _T, \pi _T(x_T,\xi _T))\) and our assertion is true. If it is true for \(t+1\), formula (10) reads

$$\begin{aligned} v^{\pi }_t(h_t)= & {} r_t(x_t,\xi _t,\pi _t(x_t,\xi _t))\\&+\sigma _t\big (x_t,\xi _t,Q_t(x_t,\xi _t,\pi _t(x_t,\xi _t)),\\&x'\mapsto v^{\pi _{t+1},\ldots ,\pi _T}_{t+1}(x',\varPhi _t(x_t,\xi _t,\pi _t(x_t,\xi _t),x'))\big ). \end{aligned}$$

Substitution of (7) proves our assertion. \(\square \)

5 Dynamic programming

We consider a family of dynamic risk measures \(\big \{\rho _{t,T}^\pi \big \}_{t=1,\ldots ,T}^{\pi \in \varPi }\) which is normalized, translation-invariant, stochastically conditionally time consistent, and Markov. Our objective is to analyze the risk minimization problem:

$$\begin{aligned} \min _{\pi \in \varPi } v_1^\pi (x_1,\varXi _1(x_1)),\quad x_1\in \mathscr {X}. \end{aligned}$$

For this purpose, we introduce the family of value functions:

$$\begin{aligned} v_t^*(h_t)=\inf _{\pi \in \varPi _{t,T}} v_t^\pi (h_t), \quad t= 1,\ldots ,T, \quad h_t \in \mathscr {X}^t, \end{aligned}$$
(12)

where \(\varPi _{t,T}\) is the set of feasible deterministic policies \(\pi =\{\pi _t,\ldots ,\pi _T\}\). By Theorem 2, transition risk mappings \(\big \{\sigma _t\big \}_{t=1, \ldots , T-1}\) exist, such that Eq. (10) hold.

We assume that the spaces \(\mathscr {P}(\mathscr {X})\) and \(\mathscr {P}(\mathscr {Y})\) are equipped with the topology of weak convergence, and the space \(\mathscr {V}\) is equipped with the topology of pointwise convergence. All continuity statements are made with respect to the said topologies.

We also assume that the kernels \(K_t(x,y,u)\) have densities \(q_t(\cdot , \cdot \mid x,y,u)\) with respect to a finite product measure \(\mu _X \otimes \mu _Y\) on \(\mathscr {X}\times \mathscr {Y}\), as in Example 3. In this case,

$$\begin{aligned} \Bigg [ \int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy)\Bigg ](dx') = \Bigg [\int _\mathscr {Y}\int _\mathscr {Y}q_t(x',y' \mid x,y,u)\, \xi (dy)\,\mu _Y(dy')\Bigg ]\;\mu _X(dx'). \end{aligned}$$
(13)

Our main result is that the value functions (12) are memoryless, that is, they depend on \((x_t,\xi _t)\) only, and that they satisfy a generalized form of a dynamic programming equation. The equation also allows us to identify the optimal policy.

Theorem 3

We assume the following conditions:

  1. (i)

    The functions \((x,u) \mapsto q_t(x',y'|x,y,u)\) are continuous at all \((x',y',x,y,u)\), uniformly over \((x',y',y)\);

  2. (ii)

    The transition risk mappings \(\sigma _t(\cdot ,\cdot ,\cdot ,\cdot )\), \(t=1,\ldots ,T\), are lower semi-continuous;

  3. (iii)

    The functions \(r_t(\cdot ,\cdot ,\cdot )\), \(t=1,\ldots ,T\), are lower semicontinuous;

  4. (iv)

    The multifunctions \(\mathscr {U}_t(\cdot )\), \(t=1,\ldots ,T\), are compact-valued and upper-semicontinuous.

Then the functions \(v_t^*\), \(t=1,\ldots ,T\) are memoryless, lower semicontinuous, and satisfy the following dynamic programming equations:

$$\begin{aligned} v_T^*(x,\xi )= & {} \min _{u \in \mathscr {U}_T(x)} r_T(x,\xi ,u), \quad x \in \mathscr {X}, \quad \xi \in \mathscr {P}(\mathscr {X}), \\ v_t^*(x,\xi )= & {} \min _{u \in \mathscr {U}_t(x)} \bigg \{ r_t(x,\xi ,u) {}\\&+\,\sigma _t\Big (x,\xi ,\int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy), x' \mapsto v_{t+1}^*\big (x',\varPhi _t(x,\xi ,u,x')\big )\Big ) \bigg \},\\&x \in \mathscr {X}, \quad \xi \in \mathscr {P}(\mathscr {Y}), \quad t = T-1,\ldots ,1. \end{aligned}$$

Moreover, an optimal Markov policy \(\hat{\pi }\) exists and satisfies the equations:

$$\begin{aligned} \hat{\pi }_T(x,\xi )\in & {} \mathop {\mathrm{argmin}}_{u \in \mathscr {U}_T(x)}\, r_T(x,\xi ,u), \quad x \in \mathscr {X}, \quad \xi \in \mathscr {P}(\mathscr {Y}),\\ \hat{\pi }_t(x,\xi )\in & {} \mathop {\mathrm{argmin}}_{u \in \mathscr {U}_t(x)} \bigg \{ r_t(x,\xi ,u) \\&+\, \sigma _t\Big (x,\xi ,\int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy), x' \mapsto v_{t+1}^*\big (x',\varPhi _t(x,\xi ,u,x')\big )\Big ) \bigg \},\\&x \in \mathscr {X}, \quad \xi \in \mathscr {P}(\mathscr {Y}), \quad t = T-1,\ldots ,1. \end{aligned}$$

Proof

For all \(h_T \in \mathscr {X}^T\) we have

$$\begin{aligned} v_T^*(h_T)= \inf _{\pi _T} r_T(x_T,\xi _T,\pi _T(h_T)) =\inf _{u\in \mathscr {U}_T(x_T)} r_T(x_T,\xi _T,u). \end{aligned}$$
(14)

By assumptions (iii) and (iv), owing to the Berge theorem [see (Aubin and Frankowska 2009, Theorem 1.4.16)], the infimum in (14) is attained and is a lower semicontinuous function of \((x_T,\xi _T)\). Hence, \(v_T^*\) is memoryless. Moreover, the optimal solution mapping \(\varPsi _T(x,\xi )=\big \{u\in \mathscr {U}_T(x): r_T(x,\xi ,u)=v_T^*(x,\xi )\big \}\) has nonempty and closed values and is measurable. Therefore, a measurable selector \(\hat{\pi }_T\) of \(\varPsi _T\) exists (see, Kuratowski and Ryll-Nardzewski 1965; Aubin and Frankowska 2009, Thm. 8.1.3), and

$$\begin{aligned} v_T^*(h_T)=v_T^*(x_T,\xi _T) = v_T^{\hat{\pi }_T}(x_T,\xi _T). \end{aligned}$$

We prove the theorem by induction backward in time. Suppose \(v^*_{t+1}(\cdot )\) is memoryless, lower semicontinuous, and Markov decision rules \(\{\hat{\pi }_{t+1},\ldots ,\hat{\pi }_T\}\) exist such that

$$\begin{aligned} v^*_{t+1}(h_{t+1}) = v^*_{t+1}(x_{t+1},\xi _{t+1}) = v_{t+1}^{\{\hat{\pi }_{t+1},\ldots ,\hat{\pi }_T\}}(x_{t+1},\xi _{t+1}), \quad \forall h_{t+1}\in {\mathscr {X}}^{t+1}. \end{aligned}$$

Then for any \(h_t \in {\mathscr {X}}^t\) formula (10), after substituting (7), yields

$$\begin{aligned} v_t^*(h_t)&=\inf _{\pi \in \varPi _{t,T}} v_t^\pi (h_t) \\&=\inf _{\pi \in \varPi _{t,T}} \left\{ r_t(x_t,\xi _t,\pi _t(h_t)) \right. \\&\left. +\, \sigma _t\Big (x_t,\xi _t, \int _\mathscr {Y}K_t^X(x_t,y,\pi _t(h_t))\;\xi _t(dy),v_{t+1}^{\pi }(h_t,\cdot )\Big )\right\} . \end{aligned}$$

Since \(v_{t+1}^\pi (h_t,x') \ge v_{t+1}^*\big (x',\varPhi _t\big (x_t,\xi _t,\pi _t(h_t),x'\big )\big )\) for all \(x' \in \mathscr {X}\), and \(\sigma _t\) is non-decreasing with respect to the last argument, we obtain

$$\begin{aligned}&v_t^*(h_t) \ge \inf _{\pi \in \varPi _{t,T}} \bigg \{ r_t(x_t,\xi _t,\pi _t(h_t))\nonumber \\&{\quad } + \sigma _t\Big (x_t,\xi _t, \int _\mathscr {Y}K_t^X(x_t,y,\pi _t(h_t))\;\xi _t(dy),x' \mapsto v_{t+1}^*(x',\varPhi _t(x_t,\xi _t,\pi _t(h_t),x'))\Big )\bigg \}\nonumber \\&\quad = \inf _{u \in \mathscr {U}_t(x_t)} \bigg \{ r_t(x_t,\xi _t,u) \nonumber \\&\quad +\, \sigma _t\Big (x_t,\xi _t, \int _\mathscr {Y}K_t^X(x_t,y,u)\;\xi _t(dy),x' \mapsto v_{t+1}^*(x',\varPhi _t(x_t,\xi _t,u,x'))\Big ) \bigg \}. \end{aligned}$$
(15)

In order to complete the induction step, we need to establish lower semicontinuity of the mapping

$$\begin{aligned} (x,\xi ,u) \mapsto \sigma _t\Big (x,\xi ,\int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy), x' \mapsto v_{t+1}^*\big (x',\varPhi _t(x,\xi ,u,x')\big )\Big ). \end{aligned}$$
(16)

To this end, suppose \(x^{(k)}\rightarrow x\), \(\xi ^{(k)}\rightarrow \xi \) (weakly), \(u^{(k)}\rightarrow u\), as \(k\rightarrow \infty \).

First, we verify that the mapping \((x,\xi ,u)\mapsto \int _\mathscr {Y}K_t^X(x,y,u)\,\xi (dy)\) appearing in the third argument of \(\sigma _t\) is weakly continuous. By formula (13), for any bounded continuous function \(f:\mathscr {X}\rightarrow \mathbb {R}\) we have

$$\begin{aligned}&\int _{\mathscr {X}} f(x')\;\Bigg [ \int _\mathscr {Y}K_t^X(x^{(k)},y,u^{(k)})\;\xi ^{(k)}(dy)\Bigg ](dx')\nonumber \\&\quad = \int _{\mathscr {X}} f(x')\;\Bigg [\int _\mathscr {Y}\int _\mathscr {Y}q_t(x',y' \mid x^{(k)},y,u^{(k)})\, \xi ^{(k)}(dy)\,\mu _Y(dy')\Bigg ]\;\mu _X(dx')\nonumber \\ \end{aligned}$$
(17)

By assumption (i),

$$\begin{aligned} \lim _{k\rightarrow \infty }\int _\mathscr {Y}\big [q_t(x',y' \mid x^{(k)},y,u^{(k)})- q_t(x',y' \mid x,y,u)\big ] \xi ^{(k)}(dy) = 0, \end{aligned}$$
(18)

uniformly over \(x',y'\). Moreover, by Lebesgue theorem, the function

$$\begin{aligned} y \mapsto \int _{\mathscr {X}} f(x')\;\int _\mathscr {Y}q_t(x',y' \mid x,y,u)\, \,\mu _Y(dy')\;\mu _X(dx') \end{aligned}$$
(19)

is continuous. Therefore, combining (17) and (18), we obtain the chain of equations:

$$\begin{aligned}&{\lim _{k\rightarrow \infty }\int _{\mathscr {X}} f(x')\;\Bigg [ \int _\mathscr {Y}K_t^X(x^{(k)},y,u^{(k)})\;\xi ^{(k)}(dy)\Bigg ](dx')}\\&\quad =\lim _{k\rightarrow \infty } \int _{\mathscr {X}} f(x')\;\Bigg [\int _\mathscr {Y}\int _\mathscr {Y}q_t(x',y' \mid x,y,u)\, \xi ^{(k)}(dy)\,\mu _Y(dy')\Bigg ]\;\mu _X(dx')\\&\quad =\lim _{k\rightarrow \infty } \int _\mathscr {Y}\Bigg [\int _{\mathscr {X}} f(x')\;\int _\mathscr {Y}q_t(x',y' \mid x,y,u)\, \,\mu _Y(dy')\;\mu _X(dx')\Bigg ]\xi ^{(k)}(dy)\\&\quad = \int _\mathscr {Y}\Bigg [\int _{\mathscr {X}} f(x')\;\int _\mathscr {Y}q_t(x',y' \mid x,y,u)\, \,\mu _Y(dy')\;\mu _X(dx')\Bigg ]\xi (dy)\\&\quad = \int _{\mathscr {X}} f(x')\;\Bigg [ \int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy)\Bigg ](dx'). \end{aligned}$$

The last by one equation follows from the weak convergence of \(\xi ^{(k)}\) to \(\xi \) and from the continuity of the function (19). Thus, the third argument of \(\sigma _t\) in (16) is continuous with respect to \((x,\xi ,u)\).

Let us examine the last argument of \(\sigma _t\) in (16). By (18), for every continuous bounded function \(f(\cdot )\) on \(\mathscr {Y}\), and for each fixed \(x'\in \mathscr {X}\),

$$\begin{aligned}&\lim _{k\rightarrow \infty } \int _\mathscr {Y}f(y') \;\varPhi _t(x^{(k)},\xi ^{(k)},u^{(k)},x')(dy') \\&\quad = \lim _{k\rightarrow \infty } \frac{\int _\mathscr {Y}f(y') \int _\mathscr {Y}q_t(x', y' \mid x^{(k)},y,u^{(k)})\; \xi ^{(k)}(dy) \; \mu _Y(dy')}{\int _\mathscr {Y}\int _\mathscr {Y}q_t(x', y' \mid x^{(k)},y,u^{(k)})\; \xi ^{(k)}(dy) \; \mu _Y(dy')} \\&\quad = \frac{\int _\mathscr {Y}f(y') \int _\mathscr {Y}q_t(x', y' \mid x,y,u)\; \xi (dy) \; \mu _Y(dy')}{\int _\mathscr {Y}\int _\mathscr {Y}q_t(x', y' \mid x,y,u)\; \xi (dy) \; \mu _Y(dy')}, \end{aligned}$$

provided that \((x,\xi ,u,x')\) is such that

$$\begin{aligned} \int _\mathscr {Y}\int _\mathscr {Y}q_t(x', y' \mid x,y,u)\; \xi (dy) \; \mu _Y(dy') > 0. \end{aligned}$$
(20)

Therefore, the operator \(\varPhi _t(\cdot ,\cdot ,\cdot ,x')\) is weakly continuous at these points. Let \(x,\xi ,u\) be fixed. Consider the sequence of functions \(V^{(k)}:\mathscr {X}\rightarrow \mathbb {R}\), \(k=1,2,\ldots \), and the function \(V:\mathscr {X}\rightarrow \mathbb {R}\), defined as follows:

$$\begin{aligned} V^{(k)}(x')&= v_{t+1}^*\big (x',\varPhi _t(x^{(k)},\xi ^{(k)},u^{(k)},x')\big ),\\ V(x')&= v_{t+1}^*\big (x',\varPhi _t(x,\xi ,u,x')\big ). \end{aligned}$$

Since \(v_{t+1}^*(\cdot ,\cdot )\) is lower-semicontinuous and \(\varPhi _t(\cdot ,\cdot ,\cdot ,x')\) is continuous, whenever condition (20) is satisfied, we infer that \(V(x') \le \liminf _{k\rightarrow \infty }\; V^{(k)}(x')\), at all \(x'\in \mathscr {X}\) at which (20) holds. As \(v_{t+1}^*\) and \(\varPhi _t\) are measurable, both V and \(\liminf _{k\rightarrow \infty } V^{(k)}\) are measurable as well.

By Theorem 2, the mapping \(\sigma _t\) is preserving the stochastic order \(\preceq _{\text {st}}\) of the last argument with respect to the measure \(\int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy)\). Since

$$\begin{aligned} \left( \int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy)\right) \left\{ x'\in \mathscr {X}: \int _\mathscr {Y}\int _\mathscr {Y}q_t(x', y' \mid x,y,u)\; \xi (dy) \; \mu _Y(dy') = 0\right\} = 0, \end{aligned}$$

the value of \(\liminf _{k\rightarrow \infty }\; V^{(k)}(x')\) at the set of \(x'\) at which (20) is violated, is irrelevant. Consequently, by assumption (ii), with the view at the already established continuity of the third argument, we obtain the following chain of relations:

$$\begin{aligned} \sigma _t\Big (x,\xi ,\int _\mathscr {Y}&K_t^X(x,y,u)\;\xi (dy), V \Big ) \le \sigma _t\Big (x,\xi ,\int _\mathscr {Y}K_t^X(x,y,u)\;\xi (dy), \liminf _{k\rightarrow \infty }\, V^{(k)} \Big )\\&= \sigma _t\Big (x,\xi ,\lim _{k\rightarrow \infty }\int _\mathscr {Y}K_t^X(x^{(k)},y,u^{(k)})\;\xi ^{(k)}(dy), \liminf _{k\rightarrow \infty }\, V^{(k)} \Big )\\&\le \liminf _{k\rightarrow \infty } \sigma _t\Big (x^{(k)},\xi ^{(k)},\int _\mathscr {Y}K_t^X(x^{(k)},y,u^{(k)})\;\xi ^{(k)}(dy), V^{(k)} \Big ). \end{aligned}$$

Consequently, the mapping (16) is lower semicontinuous.

Using assumptions (ii) and (iv) and invoking the Berge theorem again (see, e.g., Aubin and Frankowska 2009, Theorem 1.4.16), we deduce that the infimum in (15) is attained and is a lower semicontinuous function of \((x_t,\xi _t)\). Moreover, the optimal solution mapping, that is, the set of \( u\in \mathscr {U}_T(x)\) at which the infimum in (15) is attained, is nonempty, closed-valued, and measurable. Therefore, a minimizer \(\hat{\pi }_t\) in (15) exists and is a measurable function of \((x_t,\xi _t)\) (see, e.g., Kuratowski and Ryll-Nardzewski 1965; Aubin and Frankowska 2009, Thm. 8.1.3). Substituting this minimizer into (15), we obtain

$$\begin{aligned} v_t^*(h_t)\ge & {} r_t\big (x_t,\xi _t,\hat{\pi }_t(x_t,\xi _t)\big ) \\&{\quad } + \sigma _t\Big (x_t,\xi _t, \int _\mathscr {Y}K_t^X\big (x_t,y,\hat{\pi }_t(x_t,\xi _t)\big )\;\xi _t(dy), \\&\quad x'\mapsto v_{t+1}^*\big (x',\varPhi _t(x,\xi ,\hat{\pi }_t(x_t,\xi _t),x')\big )\Big )\\&\quad = v_t^{\{\hat{\pi }_t,\ldots ,\hat{\pi }_T\}}(x_t,\xi _t). \end{aligned}$$

In the last equation, we used Corollary 1. On the other hand, we have

$$\begin{aligned} v_t^*(h_t)=\inf _{\pi \in \varPi _{t,T}} v_t^\pi (h_t) \le v_t^{\{\hat{\pi }_t,\ldots ,\hat{\pi }_T\}}(x_t,\xi _t). \end{aligned}$$

Therefore \(v_t^*(h_t)=v_t^{\{\hat{\pi }_t,\ldots ,\hat{\pi }_T\}}(x_t,\xi _t)\) is memoryless, lower semicontinuous, and

$$\begin{aligned}&v_t^*(x_t,\xi _t) = \min _{u \in \mathscr {U}_t(x_t)} \bigg \{ r_t(x_t,\xi _t,u)\\&\quad {} + \sigma _t\Big (x_t,\xi _t,\int _\mathscr {Y}K_t^X(x_t,y,u)\;\xi _t(dy), x' \mapsto v_{t+1}^*\big (x',\varPhi _t(x_t,\xi _t,u,x')\big )\Big ) \bigg \}\\&\quad = r_t\big (x_t,\xi _t,\hat{\pi }_t(x_t,\xi _t)\big )\\&{\quad } + \sigma _t\bigg (x_t,\xi _t,\int _\mathscr {Y}K_t^X\big (x_t,y,\hat{\pi }_t(x_t,\xi _t)\big )\;\xi _t(dy), \\&\quad x' \mapsto v_{t+1}^*\big (x',\varPhi _t(x_t,\xi _t,\hat{\pi }_t(x_t,\xi _t),x')\big )\bigg ). \end{aligned}$$

This completes the induction step. \(\square \)

The most essential assumption of Theorem 3 is assumption (ii) of the lower semicontinuity of the transition risk mappings \(\sigma _t(\cdot ,\cdot ,\cdot ,\cdot )\). If these mappings are derived from convex or coherent risk measures, as illustrated in Fan and Ruszczyński (2016), their lower semicontinuity with respect to the last argument follows from the corresponding property of the risk measure. In particular, (Ruszczyński and Shapiro 2006a, Cor. 3.1) derives continuity from monotonicity on Banach lattices. The semicontinuity with respect to the third argument, the probability measure, is a more complex issue. The reference Fan and Ruszczyński (2016), Lemmas 4.7 and 4.8, verifies this condition for two popular risk measures: the Average Value at Risk and the mean-semideviation measure. Similar remarks apply to the assumption (iii) about the state-control risk functions. The assumptions (i) and (iv) are the same as in the utility models of Bäuerle and Rieder (2017).

We could have made the sets \(\mathscr {U}_t\) depend on \(\xi _t\), but this is hard to justify.

6 Illustration: machine deterioration

6.1 Description of the process

Consider the problem of minimizing costs of using a machine in T periods. The condition of the machine can deteriorate over time, but is not known with certainty. The only information available is the operating cost. The control in any period is to continue using the machine, or to replace it.

At the beginning of period \(t=1, \ldots , T\), the condition of the machine is denoted by \(Y_t \in \{1,2\}\), with 1 denoting the “good” and 2 the “bad” state. The controls are denoted by \(u_t \in \{0,1\}\), with 0 meaning “continue”, and 1 meaning “replace”.

The dynamics is Markovian, with the following transition matrices \(K^{[u]} \), \({u \in \{0,1\}}\):

$$\begin{aligned} K^{[0]}=\begin{pmatrix} 1-p &{} p\\ 0 &{} 1 \end{pmatrix}, \quad K^{[1]}=\begin{pmatrix} 1-p &{} p\\ 1-p &{} p \end{pmatrix}. \end{aligned}$$
(21)

We can observe the cost incurred during period t, denoted by \(X_{t+1}\). The increment of the time index is due to the fact that the cost becomes known at the end of the period, and provides information for the decision in the next period. The conditional distribution of \(X_{t+1}\), given \(y_t\) and \(u_t\), is described by two density functions \(f_1\) and \(f_2\):

$$\begin{aligned} \begin{aligned} \mathbb {P}\big [X_{t+1} \le C \,\big |\, y_t=i,u_t=0\big ] = \int _{-\infty }^{C} f_i(x)\;dx, \quad i=1,2,\\ \mathbb {P}\big [X_{t+1} \le C \,\big |\, y_t=i,u_t=1\big ] = \int _{-\infty }^{C} f_1(x)\;dx, \quad i=1,2. \end{aligned} \end{aligned}$$
(22)

Assumption 2

The functions \(f_1\) and \(f_2\) are uniformly bounded and the conditional distribution of \(X_{t+1}\) given that the machine is in “good” condition is stochastically smaller than the conditional distribution of \(x_{t+1}\) given that the machine is in “bad” condition, i.e.,

$$\begin{aligned} \int _{-\infty }^{C} f_1(x)\;dx \ge \int _{-\infty }^{C} f_2(x)\;dx,\quad \forall \, C \in \mathbb {R}; \end{aligned}$$

with a slight abuse of notation, we write it \(f_1 \preceq _{\mathrm{st}} f_2\).

Thus the relations (21) and (22) define \(\big \{X_t,Y_t\big \}_{t=1,\ldots ,T+1}\) as a partially observable Markov process controlled by \(\big \{u_t\big \}_{t=1,\ldots ,T}\). Based on observations \((x_1,\ldots ,x_t)\), the belief state \(\xi _t \in [0,1]\) denotes the conditional probability that \(Y_t=1\). We can update the posterior belief state as follows:

$$\begin{aligned} \xi _{t+1} = {\left\{ \begin{array}{ll} \varPhi (\xi _t,x_{t+1}), &{} \text { if } u_t=0;\\ 1-p, &{} \text { if } u_t=1,\\ \end{array}\right. } \end{aligned}$$

where \(\varPhi \) is the Bayes operator,

$$\begin{aligned} \varPhi (\xi ,x') = \frac{(1-p) \xi f_1(x')}{\xi f_1(x') + (1-\xi )f_2(x')}. \end{aligned}$$
(23)

We assume that the initial probability \(\xi _0 \in [0,1]\) is known; then \(\xi _1(x_1)=\varPhi (\xi _0,x_1)\). From (23) we see that \(\varPhi (0,\,\cdot \,)=0\), \(\varPhi (1,\,\cdot \,)=1-p\), and \(\varPhi (\,\cdot \,,x')\) is non-decreasing.

6.2 Risk modeling

At the beginning of period t, if we replace the machine (\(u_t=1\)), there is an additional fixed replacement cost R. Then the costs incurred are

$$\begin{aligned} \left\{ \begin{aligned}&r_t(x_t,u_t)=R\cdot u_t + x_t,\quad t=1 ,\ldots ,T;\\&r_{T+1}(x_{T+1})=x_{T+1}. \end{aligned} \right. \end{aligned}$$
(24)

We denote the history of observations by \(h_t = (x_1 ,\ldots ,x_t)\) and the set of all history-dependent policies by \(\varPi := \big \{ \, \pi = (\pi _1,\ldots ,\pi _T) \, | \, \forall t, \; \pi _t(x_1,\ldots ,x_t) \in \{0,1\} \,\big \}\). We want to evaluate the risk of costs (24) for any \(\pi \in \varPi \), and find an optimal policy.

As shown in Theorem 2, construction of Markovian risk measures is equivalent to specifying transition risk mappings \(\sigma _t : \mathbb {R}\times \mathscr {P}(\mathbb {R}) \times \mathscr {P}(\mathbb {R}) \times \mathscr {V}\rightarrow \mathbb {R}\), where\(\mathscr {V}\) is the space of all bounded and measurable functions from \(\mathbb {R}\) to \(\mathbb {R}\). For simplicity, we assume that \(\sigma _t(\cdot ,\cdot ,\cdot ,\cdot )\) is the same for all t and does not depend on the current state \((x_t,\xi _t)\), that is, \(\sigma _t(x,\xi ,q,v) = \sigma (q,v)\).

Remark 2

For a probability measure \(q \in \mathscr {P}(\mathbb {R})\) that has f as the density function, with slight abuse of notation, we also write \(\sigma (f,\cdot )\) instead of \(\sigma (q,\cdot )\).

6.3 Value and policy monotonicity

We assume that the transition risk mapping \(\sigma : \mathscr {P}(\mathbb {R}) \times \mathscr {V}\rightarrow \mathbb {R}\) satisfy all assumptions of Theorem 3. Then the optimal value functions \(v_t^*,\,t=1,\ldots ,T+1\) are memoryless and satisfy the following dynamic programming equations:

$$\begin{aligned} v_t^*(x,\xi )= & {} x + \min \Big ( R + \sigma \big (f_1, x' \mapsto v_{t+1}^*(x',1-p) \big ) ;\nonumber \\&\sigma \big (\xi f_1 + (1-\xi ) f_2, x' \mapsto v_{t+1}^*(x',\varPhi (\xi ,x'))\big ) \Big ),\nonumber \\&x \in \mathbb {R}, \quad \xi \in [0,1],\quad t=1,\ldots ,T, \end{aligned}$$
(25)

with the final stage value \(v_{T+1}^*(x,\xi )=x\). Moreover, an optimal Markov policy exists, which is defined by the minimizers in the above dynamic programming equations.

Directly from (25) we see that \(v_t^*(x,\xi ) = x + w_t^*(\xi )\), \( t=1 ,\ldots ,T+1\). The dynamic programming Eq. (25) simplify as follows:

$$\begin{aligned} w_t^*(\xi ) = \min \Big \{ R + \sigma \big (f_1, x' \mapsto x' + w_{t+1}^*(1-p) \big ) ;\nonumber \\ \sigma \big (\xi f_1 + (1-\xi ) f_2, x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x'))\big ) \Big \}, \end{aligned}$$
(26)

with the final stage value \(w_{T+1}^*(\,\cdot \,)=0\). We can establish monotonicity of \(w^*(\cdot )\).

Theorem 4

If \(\frac{f_1}{f_2}\) is non-increasing, then the functions \(w_t^*: [0,1]\rightarrow \mathbb {R},\; t=1 ,\ldots ,T+1\) are non-increasing.

Proof

Clearly, \(w_{T+1}^*\) is non-increasing. Assume by induction that \(w_{t+1}^*\) is non-increasing. For any \(\xi _1 \le \xi _2\) we have:

1. :

\(\xi _1 f_1 + (1-\xi _1) f_2 \succeq _{\text {st}} \xi _2 f_1 + (1-\xi _2) f_2\), because \(f_1 \preceq _{\text {st}} f_2\).

2. :

For all \(x'\), we have \(x'+w_{t+1}^*(\varPhi (\xi _1,x')) \ge x'+w_{t+1}^*(\varPhi (\xi _2,x'))\), as \(w_{t+1}^*\) is non-increasing and \(\varPhi (\cdot ,x')\) is non-decreasing.

3. :

the mapping \(x' \mapsto x'+w_{t+1}^*(\varPhi (\xi ,x'))\) is non-decreasing for all \(\xi \). To show that, it is sufficient to establish that \(x' \mapsto \varPhi (\xi ,x')\) is non-increasing, and this can be seen from the formula (for \(0 \le p < 1\)):

$$\begin{aligned} \frac{1}{\varPhi (\xi ,x')}=\frac{1}{1-p}\left( 1+\frac{f_2(x')}{f_1(x')}\left( \frac{1}{\xi }-1 \right) \right) . \end{aligned}$$

Thus

$$\begin{aligned} \begin{aligned}&{\sigma \big (\xi _1 f_1 + (1-\xi _1) f_2, x' \mapsto x'+w_{t+1}^*(\varPhi (\xi _1,x'))\big )}\qquad \\&\quad \ge \sigma \big (\xi _1 f_1 + (1-\xi _1) f_2, x' \mapsto x'+w_{t+1}^*(\varPhi (\xi _2,x'))\big )&\qquad \text {(because of 2.)}\\&\quad \ge \sigma \big (\xi _2 f_1 + (1-\xi _2) f_2, x' \mapsto x'+w_{t+1}^*(\varPhi (\xi _2,x'))\big )&\qquad \text {(because of 1. and 3.)} \end{aligned} \end{aligned}$$

which completes the induction step. \(\square \)

The monotonicity assumption on \(\frac{f_1}{f_2}\) is in fact a sufficient condition for \(f_1 \preceq _{\text {st}} f_2\).

From Theorem 4 we obtain the following threshold property of the policy.

Theorem 5

Under the assumptions of Theorem 4, there exist thresholds \(\xi ^*_t \in [0,1],\; t=1,\ldots ,T\) such that the policy

$$\begin{aligned} u_t^* = {\left\{ \begin{array}{ll} 0 &{} \text {if } \xi _t > \xi _t^*,\\ 1 &{} \text {if } \xi _t \le \xi _t^*, \end{array}\right. } \end{aligned}$$

is optimal.

Proof

Suppose \(\xi \) is such that replacement at time t is optimal:

$$\begin{aligned} R + \sigma \big (f_1, x' \mapsto x' + w_{t+1}^*(1-p) \big ) \le \sigma \big (\xi f_1 + (1-\xi ) f_2, x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x'))\big ). \end{aligned}$$

Then for any \(\zeta \le \xi \), we have \(\xi f_1 + (1-\xi ) f_2 \preceq _{\text {st}} \zeta f_1 + (1-\zeta ) f_2\) and \(\varPhi (\xi ,x') \ge \varPhi (\zeta ,x')\). Consequently,

$$\begin{aligned} \begin{aligned}&{R + \sigma \big (f_1, x' \mapsto x' + w_{t+1}^*(1-p) \big )}\qquad \\&\quad \le \sigma \big (\xi f_1 + (1-\xi ) f_2, x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x'))\big )\\&\quad \le \sigma \big (\zeta f_1 + (1-\zeta ) f_2, x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x'))\big )\\&\quad \le \sigma \big (\zeta f_1 + (1-\zeta ) f_2, x' \mapsto x' + w_{t+1}^*(\varPhi (\zeta ,x'))\big ), \end{aligned} \end{aligned}$$

and replacement is optimal for \(\zeta \) as well. \(\square \)

6.4 Numerical illustration

In this section, we solve the problem in the special case where \(f_1\) and \(f_2\) are density functions \(\mathbb {U}(m_1,M_1)\) and \(\mathbb {U}(m_2,M_2)\) with \(m_1 \le m_2 \le M_1 \le M_2\). Then the Bayes operator is piece-wise constant with respect to \(x'\):

$$\begin{aligned} \varPhi (\xi ,x')=\left\{ \begin{aligned}&1-p,&\text {if } m_1 \le x' \le m_2;\\&\frac{(1-p)\xi (M_2-m_2)}{\xi (M_2-m_2)+(1-\xi )(M_1-m_1)}:=\hat{\phi }(\xi ),&\text {if } m_2 \le x' \le M_1;\\&0,&\text {if } M_1 \le x' \le M_2.\\ \end{aligned} \right. \end{aligned}$$

The conditional distribution of \(x'\) given \(\xi \) is described by the density function \(\xi f_1 + (1-\xi ) f_2\), which is also constant in each of the three intervals \([m_1,m_2)\), \([m_2,M_1]\) and \((M_1,M_2]\), with the following probabilities amassed in each of the three intervals:

$$\begin{aligned} \begin{aligned}&q_1(\xi ) = \frac{\xi (m_2-m_1)}{M_1-m_1},\\&q_2(\xi ) = (M_1-m_2)\left( \frac{\xi }{M_1-m_1} +\frac{1-\xi }{M_2-m_2} \right) ,\\&q_3(\xi ) = \frac{(1-\xi )(M_2-M_1)}{M_2-m_2}. \end{aligned} \end{aligned}$$

We use the mean-semideviation transition risk mapping of Example 2, with \(p=1\) and constant \(\varkappa \). It is strongly monotonic with respect to stochastic order and lower semi-continuous with respect to (qv). Then the dynamic programming Eq. (26) for \(t=1,\ldots ,T\) become:

$$\begin{aligned} w_t^*(\xi ) = \min \Big \{ R + E_t^*(1) + \mathbb {E}_{f_1} \big ( x' \mapsto x' + w_{t+1}^*(1-p) - E_t^*(1) \big )_+ ;\nonumber \\ E_t^*(\xi ) + \mathbb {E}_{\xi f_1 + (1-\xi ) f_2} \big ( x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x')) - E_t^*(\xi ) \big )_+ \Big \}, \end{aligned}$$
(27)

where

$$\begin{aligned} E_t^*(\xi ) := & {} \mathbb {E}_{\xi f_1 + (1-\xi ) f_2} \big ( x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x')) \big )\nonumber \\= & {} q_1(\xi )\left( \frac{m_1+m_2}{2} + w_{t+1}^*(1-p) \right) \nonumber \\&\quad +\, q_2(\xi )\left( \frac{m_2+M_1}{2} + w_{t+1}^*(\hat{\phi }(\xi )) \right) \nonumber \\&\quad +\, q_3(\xi )\left( \frac{M_1+M_2}{2} + w_{t+1}^*(0) \right) . \end{aligned}$$
(28)

As \(x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x')) - E_t^*(\xi )\) is linear in each of the intervals \([m_1,m_2]\), \([m_2,M_1]\) and \([M_1,M_2]\), we have

$$\begin{aligned} \begin{aligned}&{\mathbb {E}_{\xi f_1 + (1-\xi ) f_2} \big ( x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x')) - E_t^*(\xi ) \big )_+}\qquad \\&\quad = q_1(\xi ) \, \theta \big (m_1,m_2,E_t^*(\xi )-w_{t+1}^*(1-p)\big )\\&\qquad +q_2(\xi ) \, \theta \big (m_2,M_1,E_t^*(\xi )-w_{t+1}^*(\hat{\phi }(\xi ))\big )\\&\qquad +q_3(\xi ) \, \theta \big (M_1,M_2,E_t^*(\xi )-w_{t+1}^*(0)\big ), \end{aligned} \end{aligned}$$
(29)

where, for \(a_1 \le a_2\),

$$\begin{aligned} \theta (a_1,a_2,a_3):= \frac{\int _{a_1}^{a_2}(\,\cdot \,-a_3)_+}{a_2-a_1}= {\left\{ \begin{array}{ll} \frac{1}{2}(a_1+a_2), &{}\text { if }a_3 \le a_1;\\ \frac{1}{2}(a_3+a_2)\frac{a_2-a_3}{a_2-a_1}, &{}\text { if }a_1<a_3<a_2;\\ 0, &{}\text { if }a_3 \ge a_2.\\ \end{array}\right. } \end{aligned}$$

For any t and any \(\xi \), if we know \(w_{t+1}^*(1-p)\), \(w_{t+1}^*(\hat{\phi }(\xi ))\) and \(w_{t+1}^*(0)\), then the computation of \(w_t^*(\xi )\) can be accomplished in three steps:

1. :

Compute \(E_t^*(1)\) and \(E_t^*(\xi )\) by (28).

2. :

Compute \(\mathbb {E}_{f_1} \big ( x' \mapsto x' + w_{t+1}^*(1-p) - E_t^*(1) \big )_+ \) and \(\mathbb {E}_{\xi f_1 + (1-\xi ) f_2} \big ( x' \mapsto x' + w_{t+1}^*(\varPhi (\xi ,x')) - E_t^*(\xi ) \big )_+\) by (29).

3. :

Compute \(w_t^*(\xi )\) using the dynamic programming Eq. (27).

Since \(w_{T+1}^*=0\), all \(w_t^*(\xi )\) can be calculated by recursion backward in time.

In Fig. 1, we display the distribution of the total cost obtained by simulating 100,000 runs of the system with two policies: risk-neutral and risk-averse for an example with \(m_1=0\), \(m_2=80\), \(M_1=100\), \(M_2=500\), \(p=0.2\), \(T=6\), \(R=50\), and \(\varkappa =0.9\).

The application of the risk-averse model increases the threshold values \(\xi _t^*\) of the optimal policies and results in a significantly less dispersed distribution of the total cost. For additional details, see Fan (2017).

Fig. 1
figure 1

Empirical distribution of the total cost for the risk-neutral model (blue) and the risk-averse model (orange)