Skip to main content
Log in

Time-consistency of risk measures: how strong is such a property?

  • Published:
Decisions in Economics and Finance Aims and scope Submit manuscript

Abstract

Quite recently, a great interest has been devoted to time-consistency of risk measures in its different formulations (see Delbaen in Memoriam Paul-André Meyer, Lecture notes in mathematics, vol 1874, pp 215–258, 2006; Föllmer and Penner in Stat Decis 14(1):1–15, 2006; Bion-Nadal in Stoch Process Appl 119:633–654, 2009; Delbaen et al. in Finance Stoch 14(3):449–472, 2010; Laeven and Stadje in Math Oper Res 39:1109–1141, 2014, among many others). However, almost all the papers address to coherent or convex risk measures satisfying cash-additivity. In the present work, we study time-consistency for more general dynamic risk measures where either only cash-invariance or both cash-invariance and convexity are dropped. This analysis is motivated by the recent papers of El Karoui and Ravanelli (Math Finance 19:561–590, 2009) and Cerreia-Vioglio et al. (Math Finance 21(4):743–774, 2011) who discussed and weakened the axioms above by introducing cash-subadditivity and quasi-convexity. In particular, we investigate and discuss whether the notion of time-consistency is too restrictive, when considered in the general framework of quasi-convex and cash-subadditive risk measures. Finally, we provide some conditions guaranteeing time-consistency in this more general framework.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Although the dual representation of dynamic convex and cash-subadditive risk measures has been proved in Mastrogiacomo and Rosazza Gianin (2015) under the further assumption of normalization, the representation in (2) can be recovered easily for not necessarily normalized risk measures.

  2. Although Cheridito and Kupper (2011) considered normalized risk measures, the same construction holds true also for non-normalized ones.

References

  • Acciaio, B., Penner, I.: Dynamic Risk Measures. Advanced Mathematical Methods for Finance, pp. 1–34. Springer, Berlin (2011)

    Book  Google Scholar 

  • Acciaio, B., Föllmer, H., Penner, I.: Risk assessment for uncertain cash flows: model ambiguity, discounting ambiguity, and the role of bubbles. Finance Stoch. 16, 669–709 (2012)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Artzner, P., Delbaen, F., Eber, J.M., Heath, D., Ku, H.: Coherent multiperiod risk adjusted values and Bellman’s principle. Preprint. ETH Zurich, Switzerland (2004)

  • Bartl, D. (2016): Pointwise dual representation of dynamic convex expectation. Working paper. arXiv:1612.09103v1

  • Bereanu, B.: Quasi-convexity, strictly quasi-convexity and pseudo-convexity of composite objective functions. Revue francaise d’automatique, informatique, recherche opérationnelle. Mathématique 6, 15–26 (1972)

    Google Scholar 

  • Biagini, S., Bion-Nadal, J.: Dynamic quasi concave performance measures. J. Math. Econ. 55, 143–153 (2014)

    Article  Google Scholar 

  • Bielecki, T.R., Cialenco, I., Pitera, M.: A survey of time consistency of dynamic risk measures and dynamic performance measures in discrete time: LM-measure perspective. Probab. Uncertain. Quant. Risk 2(3), 1–52 (2017)

    Google Scholar 

  • Bion-Nadal, J.: Dynamic risk measures: time consistency and risk measures from BMO martingales. Finance Stoch. 12, 219–244 (2008)

    Article  Google Scholar 

  • Bion-Nadal, J.: Time consistent dynamic risk processes. Stoch. Process. Appl. 119, 633–654 (2009)

    Article  Google Scholar 

  • Cheridito, P., Kupper, M.: Recursiveness of indifference prices and translation-invariant preferences. Math. Finan. Econ. 2, 173–188 (2009)

    Article  Google Scholar 

  • Cheridito, P., Kupper, M.: Composition of time-consistent dynamic monetary risk measures in discrete time. Int. J. Theor. Appl. Finance 14(1), 137–162 (2011)

    Article  Google Scholar 

  • Cerreia-Vioglio, S., Maccheroni, F., Marinacci, M., Montrucchio, L.: Risk measures: rationality and diversification. Math. Finance 21(4), 743–774 (2011)

    Google Scholar 

  • Cheridito, P., Delbaen, F., Kupper, M.: Coherent and convex monetary risk measures for bounded càdlàg processes. Stoch. Process. Appl. 112(1), 1–22 (2004)

    Article  Google Scholar 

  • Cheridito, P., Delbaen, F., Kupper, M.: Coherent and convex monetary risk measures for unbounded càdlàg processes. Finance Stoch. 9(3), 369–387 (2005)

    Article  Google Scholar 

  • Cheridito, P., Delbaen, F., Kupper, M.: Dynamic monetary risk measures for bounded discrete time processes. Electron. J. Probab. 11, 57–106 (2006)

    Article  Google Scholar 

  • de Finetti, B.: Sulle stratificazioni convesse. Ann. Mat. Pura Appl. 30, 173–183 (1949)

    Article  Google Scholar 

  • Delbaen, F.: Coherent risk measures. Cattedra Galileiana. Lecture Notes, Scuola Normale Superiore, Pisa, Italy (2000)

  • Delbaen, F.: Coherent risk measures on general probability spaces. In: Sandmann, K., Schönbucher, P.J. (eds.) Advances in Finance and Stochastics, pp. 1–37. Springer, Berlin (2002)

    Google Scholar 

  • Delbaen, F.: The structure of \(m\)-stable sets and in particular of the set of risk neutral measures. In: Memoriam Paul-André Meyer, Lecture Notes in Mathematics, vol. 1874, pp. 215–258 (2006)

  • Delbaen, F., Peng, S., Rosazza Gianin, E.: Representation of the penalty term of dynamic concave utilities. Finance Stoch. 14(3), 449–472 (2010)

    Article  Google Scholar 

  • Deprez, O., Gerber, H.U.: On convex principles of premium calculation. Insur. Math. Econ. 4, 179–189 (1985)

    Article  Google Scholar 

  • Detlefsen, K., Scandolo, G.: Conditional and dynamic convex risk measures. Finance Stoch. 9(4), 539–561 (2005)

    Article  Google Scholar 

  • Drapeau, S., Kupper, M.: Risk preferences and their robust representation. Math. Oper. Res. 38(1), 28–62 (2013)

    Article  Google Scholar 

  • Drapeau, S., Kupper, M., Tangpi, L., Rosazza Gianin, E.: Dual representation of minimal supersolutions of convex BSDEs. Annales de l’Institut Henri Poincaré (B) Probabilités et Statistiques 52(2), 868–887 (2016)

    Article  Google Scholar 

  • El Karoui, N., Ravanelli, C.: Cash sub-additive risk measures and interest rate ambiguity. Math. Finance 19, 561–590 (2009)

    Article  Google Scholar 

  • Fabella, R.V.: Quasi-concave (composition) functions with nonconcave argument functions. Int. Econ. Rev. 33, 473–477 (1992)

    Article  Google Scholar 

  • Föllmer, H., Penner, I.: Convex risk measures and the dynamics of their penalty functions. Stat. Decis. 14(1), 1–15 (2006)

    Google Scholar 

  • Föllmer, H., Schied, A.: Convex measures of risk and trading constraints. Finance Stoch. 6(4), 429–447 (2002)

    Article  Google Scholar 

  • Föllmer, H., Schied, A.: Stochastic Finance, 2nd edn. de Gruyter, Berlin (2008)

    Google Scholar 

  • Frittelli, M., Maggis, M.: Dual representation of quasiconvex conditional maps. SIAM J. Financ. Math. 2, 357–382 (2011a)

    Article  Google Scholar 

  • Frittelli, M., Maggis, M.: Conditional certainty equivalent. Int. J. Theor. Appl. Finance 14(1), 41–59 (2011b)

    Article  Google Scholar 

  • Frittelli, M., Rosazza Gianin, E.: Putting order in risk measures. J. Bank. Finance 26(7), 1473–1486 (2002)

    Article  Google Scholar 

  • Frittelli, M., Rosazza Gianin, E.: Dynamic convex risk measures. In: Szegö, G. (ed.) Risk Measures for the 21st Century, pp. 227–248. Wiley, New York (2004)

    Google Scholar 

  • Goovaerts, M.J., Kaas, R., Laeven, R.J.A.: Decision principles derived from risk measures. Insur. Math. Econ. 47, 294–302 (2010)

    Article  Google Scholar 

  • Greenberg, H.J., Pierskalla, W.P.: A review of quasi-convex functions. Oper. Res. 19, 1553–1570 (1971)

    Article  Google Scholar 

  • Klöppel, S., Schweizer, M.: Dynamic utility indifference valuation via convex risk measures. Math. Finance 17(4), 599–627 (2007)

    Article  Google Scholar 

  • Laeven, R.J.A., Stadje, M.A.: Robust portfolio choice and indifference valuation. Math. Oper. Res. 39, 1109–1141 (2014)

    Article  Google Scholar 

  • Mangasarian, O.L.: Convexity, pseudo-convexity and quasi-convexity of composite functions. Cahiers du Centre d’Etudes de Recherche Opérationelle 12, 114–122 (1970)

    Google Scholar 

  • Mastrogiacomo, E., Rosazza Gianin, E.: Time-consistency of cash-subadditive risk measures. Working paper (2015). arXiv:1512.03641

  • Neveu, J.: Discrete Parameter Martingales. North Holland Publishing Company, Amsterdam (1975)

    Google Scholar 

  • Riedel, F.: Dynamic coherent risk measures. Stoch. Process. Appl. 112(2), 185–200 (2004)

    Article  Google Scholar 

  • Roorda, B., Schumacher, J.M.: Weakly time consistent concave valuations and their dual representations. Finance Stoch. 20(1), 123–151 (2016)

    Article  Google Scholar 

  • Rosazza Gianin, E.: Risk measures via g-expectations. Insur. Math. Econ. 39, 19–34 (2006)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emanuela Rosazza Gianin.

Additional information

Publisher's Note

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

The present paper contains (partially) and extends a preliminary working paper (Mastrogiacomo and Rosazza Gianin 2015) (focusing only on the dual representation approach of cash-subadditive risk measures) to the quasi-convex case and also by covering the approach based on generators.

The authors wish to thank the anonymous referee for her/his comments and suggestions that contributed to improve the paper.

Appendix: Proof of Theorem 2

Appendix: Proof of Theorem 2

In the following, we will restrict to a discrete-time setting. To prove Theorem 2, the idea is to proceed by steps by using an approach similar to El Karoui and Ravanelli (2009) and Acciaio et al. (2012), that is to define a suitable setting where the original risk measure \(\pi \) can be transformed into a cash-additive risk measure \(\hat{\pi }\). We will then apply to \(\hat{\pi }\) the results on time-consistency already known for cash-additive risk measures and then, after having shown the relation between the penalty terms of \(\pi \) and \(\hat{\pi }\), we will formulate the necessary condition in terms of c. Differently from El Karoui and Ravanelli (2009) and Acciaio et al. (2012), respectively, we will consider a dynamic setting and a finite time horizon.

The main difficulties and reasons why we consider a discrete-time setting are that in continuous time, there are some additional technicalities concerning the Itô-Watanabe decomposition [with consequences on (23)] and, moreover, it is more difficult to relate the expected values between the initial and the extended probability spaces [see (22)].

Before proving Theorem 2, we introduce the setting and recall some results that will be used in the proof.

1.1 A.1 Setting

We consider a discrete-time setting with finite time horizon T. We denote by \(\mathbb {T}\) the set of times, i.e. \(\mathbb {T}=\left\{ 0,1,\ldots ,T\right\} \), and by \(\hat{\Omega }\triangleq \Omega \times \mathbb {T}\) the enlarged sample space. Any discrete-time stochastic process \(\hat{X}=(X_t)_{t\in \mathbb {T}}\) with \(X_t\in L^{\infty }(\mathcal {F}_t)\) can be therefore understood as defined on the enlarged space \(\hat{\Omega }= \Omega \times \mathbb {T}\), that is

$$\begin{aligned} \hat{X}(\omega ,t)=X_t(\omega ). \end{aligned}$$

Let

$$\begin{aligned} \hat{\mathcal {F}}= \sigma (\left\{ A_t \times \left\{ t\right\} : A_t \in \mathcal {F}_t, t \in \mathbb {T}\right\} ) \end{aligned}$$

be the \(\sigma \)-algebra \(\hat{\mathcal {F}}\) considered on \(\hat{\Omega }\).

Furthermore, let \((\hat{\Omega },\hat{\mathcal {F}})\) be endowed with the filtration \((\hat{\mathcal {F}}_t)_{t\in \mathbb {T}}\) given by

$$\begin{aligned} \hat{\mathcal {F}}_t=\sigma (\left\{ A_j \times \left\{ j\right\} , A_t \times \left\{ t,\ldots ,T\right\} : A_j \in \mathcal {F}_j,j<t,A_t \in \mathcal {F}_t\right\} ), \quad t \in \mathbb {T}, \end{aligned}$$

and with the probability measure \(\hat{P}\triangleq P\otimes \mu \), where \(\mu =(\mu _t)_{t\in \mathbb {T}}\) is some adapted process satisfying \(\sum _{t\in \mathbb {T}} \mu _t=1\), \(\mu _t >0\) for all \(t\in \mathbb {T}\) and

$$\begin{aligned} E_{\hat{P}}[\hat{X}]\triangleq E_P \left[ \sum \limits _{t\in \mathbb {T}} X_t \mu _t\right] \end{aligned}$$

for any \(\hat{X} \in L^{\infty } (\hat{\mathcal {F}})\). A random variable \(\hat{X}=(X_t)_{t\in \mathbb {T}}\) on \((\hat{\Omega },\hat{\mathcal {F}},\hat{P})\) is then \(\hat{\mathcal {F}}_t\)-measurable if and only if \(X_s\) is \(\mathcal {F}_s\)-measurable for all \(s=0,1\ldots ,t\) and \(X_s=X_t\) for all \(s>t\).

Notice that the present setting is exactly the same as in Acciaio et al. (2012) with the only difference that we use a finite time horizon.

Denote by \(\hat{\mathcal {P}}\) (respectively, \(\mathcal {P}\)) the set of all probability measures that are absolutely continuous with respect to \(\hat{P}\) (resp. P). For any \(Q\in \mathcal {P}\), denote by \(\Gamma (Q)\) the set of positive random measures \(\gamma \) on \(\mathbb {T}\) such that

$$\begin{aligned} \sum _{t\in \mathbb {T}} \gamma _t =1, \quad Q{\hbox {-}}a.s. \end{aligned}$$

and by \(\Delta (Q)\) the set of predictable non-increasing processes \(\delta =(\delta _t)_{t\in \mathbb {T}}\) with \(\delta _0=1\) and \(\delta _T= \lim _{t \rightarrow T} \delta _t\), Q-a.s.. We recall from Acciaio et al. (2012) the following result guaranteeing that any \(\hat{Q}\) in \(\hat{\mathcal {P}}\) admits a decomposition \(\hat{Q}(\mathrm{d}\omega , t)= Q(\mathrm{d}\omega ) \otimes \gamma (\omega , t)\) for some probability measure \(Q \in \mathcal {P}\) and some random measure \(\gamma \in \Gamma (Q)\).

Theorem 15

(see Lemma 3.2, Theorem 3.4 and Corollary B.3 in Acciaio et al. 2012)

  1. (i)

    Fix any probability measure \(Q\in \mathcal {P}\). To any \(\gamma \in \Gamma (Q)\) it corresponds a process \(\delta \in \Delta (Q)\) and vice versa, by means of

    $$\begin{aligned} \delta _t&=1-\sum _{s=0}^{t-1}\gamma _s= \sum _{s=t}^T \gamma _s, \quad Q{\hbox {-}}a.s., \, \forall t\in \mathbb {T}\\ \gamma _t&=\delta _t-\delta _{t+1}, \quad t\in \mathbb {T}. \end{aligned}$$
  2. (ii)

    Any probability measure \(\hat{Q}\in \hat{\mathcal {P}}\) admits a decomposition \(\hat{Q}= Q \otimes \gamma = Q \otimes \delta \) by a probability measure \(Q\in \mathcal {P}\) and a measure \(\gamma \in \Gamma (Q)\) (resp. \(\delta \in \Delta (Q)\)), meaning that

    $$\begin{aligned} E_{\hat{Q}}[\hat{X}] = E_Q\left[ \sum _{t\in \mathbb {T}} X_t \gamma _t\right] = E_Q\left[ \sum _{t\in \mathbb {T}} (X_t-X_{t+1})\delta _t\right] \end{aligned}$$
    (22)

    holds for any \(\hat{X} \in L^{\infty } (\hat{\mathcal {F}})\). Conversely, given any \(Q\in \mathcal {P}\) and any \(\gamma \in \Gamma (Q)\) (resp. any \(\delta \in \Delta (Q)\)), then

    $$\begin{aligned} \hat{Q}= Q\otimes \gamma = Q \otimes \delta \end{aligned}$$
    (23)

    is a probability measure in \(\hat{\mathcal {P}}\).

  3. (iii)

    For any \(\hat{Q}\in \hat{\mathcal {P}}\) with decomposition \(\hat{Q}= Q \otimes \gamma = Q \otimes \delta \), the conditional expectation of \(\hat{X}\in L^\infty (\hat{\mathcal {F}})\) given \(\hat{\mathcal {F}}_t\) is

    $$\begin{aligned} E_{\hat{Q}}[ \hat{X}|\hat{\mathcal {F}}_t] = X_0 \mathbb {1}_{\left\{ 0 \right\} } + \cdots +X_{t-1} \mathbb {1}_{\left\{ t-1 \right\} } +E_{Q}\left[ \sum _{s=t}^T \frac{\gamma _s}{\delta _t}X_s \bigg |\mathcal {F}_t\right] \mathbb {1}_{\left\{ t,\ldots ,T\right\} }, \end{aligned}$$
    (24)

    where the conditional expectation under Q is well defined Q-a.s. on \(\left\{ \delta _t>0\right\} \).

1.2 A.2 Proof of Theorem 2

In order to prove Theorem 2, we proceed by steps as follows. First, we transform our dynamic convex cash-subadditive risk measure \(\pi \) into a convex and cash-additive one \(\hat{\pi }\). Then, we apply to \(\pi \) the results already known for cash-additive risk measures. Finally, we translate such results by going back to \(\pi \).

Given a dynamic risk measure \(\pi _{t,T} : L^\infty (\mathcal {F}) \rightarrow L^\infty (\mathcal {F}_t)\), we define the dynamic risk measure \(\hat{\pi }_{t,T}: L^\infty (\hat{\mathcal {F}}) \rightarrow L^\infty (\hat{\mathcal {F}}_t)\) as

$$\begin{aligned} \hat{\pi }_{t,T}(\hat{X}) \triangleq X_0 \mathbb {1}_{\left\{ 0 \right\} } + X_1 \mathbb {1}_{\left\{ 1 \right\} } +\cdots + X_{t-1} \mathbb {1}_{\left\{ t-1 \right\} } + \big (\pi _{t,T}(X_T-X_t)+X_t \big ) \mathbb {1}_{\left\{ t,t+1,\ldots ,T\right\} }\nonumber \\ \end{aligned}$$
(25)

for any \(\hat{X}=(X_t)_{t\in \mathbb {T}} \in L^{\infty } (\hat{\mathcal {F}})\). Similarly, \(\hat{\pi }_{t,v}: L^\infty (\hat{\mathcal {F}}_v) \rightarrow L^\infty (\hat{\mathcal {F}}_t)\) can be defined starting from \(\pi _{t,v}\) as

$$\begin{aligned} \hat{\pi }_{t,v}(\hat{X}) \triangleq X_0 \mathbb {1}_{\left\{ 0 \right\} } + X_1 \mathbb {1}_{\left\{ 1 \right\} } +\cdots + X_{t-1} \mathbb {1}_{\left\{ t-1 \right\} } + \big (\pi _{t,v}(X_v-X_t)+X_t \big ) \mathbb {1}_{\left\{ t,t+1,\ldots ,T\right\} }\nonumber \\ \end{aligned}$$
(26)

for any \(\hat{X}=(X_t)_{t\in \mathbb {T}} \in L^{\infty } (\hat{\mathcal {F}})\).

The following proposition shows how the properties on \(\pi _{t,v}\) will be translated in those of \(\hat{\pi }_{t,v}\).

Proposition 16

Let \(\hat{\pi }_{t,T}: L^\infty (\hat{\mathcal {F}}) \rightarrow L^\infty (\hat{\mathcal {F}}_t)\) be defined starting from \(\pi _{t,T} : L^\infty (\mathcal {F}) \rightarrow L^\infty (\mathcal {F}_t)\) as in (25).

  1. (i)

    If \((\pi _{t,T})_{t}\) is a normalized, cash-subadditive and convex dynamic risk measure, then \((\hat{\pi }_{t,T})_{t}\) is a normalized, cash-additive and convex dynamic risk measure.

  2. (ii)

    If \(\pi _{t,T}\) is regular, then \(\hat{\pi }_{t,T}\) is regular.

  3. (iii)

    If \(\pi _{t,T}\) is continuous from below, then \(\hat{\pi }_{t,T}\) is continuous from below.

  4. (iv)

    Items (i)–(iii) hold true also when \(\hat{\pi }_{t,T}\) (resp. \(\pi _{t,T}\)) is replaced by \(\hat{\pi }_{t,v}\) (resp. \(\pi _{t,v}\)).

  5. (v)

    \((\hat{\pi }_{s,t})_{s,t \in \mathbb {T},s \le t}\) is time-consistent if and only if

    $$\begin{aligned} (\pi _{s,t}(\cdot - X_s) +X_s)_{0\le s\le t\le T} \end{aligned}$$

    is time-consistent for all \(X_s \in L^\infty (\mathcal {F}_s)\), that is: if and only if

    $$\begin{aligned} \pi _{s,t}(\pi _{t,v}(X_v -X_t) +X_t -X_s)+X_s = \pi _{s,t}(X_v-X_s)+X_s, \end{aligned}$$
    (27)

    for all \(X_v \in L^{\infty }(\mathcal {F}_v),X_t \in L^{\infty }(\mathcal {F}_t), X_s \in L^{\infty }(\mathcal {F}_s)\).

In the following, \((\pi _{s,t})_{s,t \in \mathbb {T},s \le t}\) satisfying (27) will be called X-time-consistent. Notice that such a property implies the classical time-consistency and is implied by time-consistency and cash-additivity together. The main reason why we assume such a stronger condition on \(\pi \) is that, by definition of \(\hat{\pi }\), we need to manage the term \(\pi _{t,v}(X_v-X_t)+X_t \) making \(\hat{\pi }\) cash-additive.

Proof

  1. (i)

    First of all, we notice that \(\hat{\pi }_{t,T}\) is \(\hat{\mathcal {F}}_t\)-measurable. Take, indeed, an arbitrary \(\hat{X}\in L^\infty (\hat{\mathcal {F}})\) and set \(\hat{Y}=\hat{\pi }_{t,T}(\hat{X})\). By definition of \(\hat{\pi }\), it then follows that, for any \(u<t\), \(Y_u= X_u\) (hence \(\mathcal {F}_u\)-measurable), while for any \(u\ge t\) it holds that \(Y_u=Y_t=\pi _{t,T}(X_T-X_t)+X_t\) is \(\mathcal {F}_t\)-measurable. Thus \(\hat{\pi }_{t,T}\) is well defined as a mapping from \(L^\infty (\hat{\mathcal {F}})\) to \(L^\infty (\hat{\mathcal {F}}_t)\). It remains to check the other properties. Normalization of \(\hat{\pi }_{t,T}\) follows immediately by that of \(\pi _{t,T}\). Cash-additivity. Let \(\hat{X}\in L^\infty (\hat{\mathcal {F}})\) and \(\hat{m}_t\in \hat{\mathcal {F}}_t\). Then

    $$\begin{aligned} \hat{m}_t=m_0 \mathbb {1}_{\left\{ 0 \right\} } + \cdots + m_{t-1} \mathbb {1}_{\left\{ t-1\right\} } + m_t \mathbb {1}_{\left\{ t,t+1,\ldots ,T\right\} }, \end{aligned}$$

    with \(m_u \in \mathcal {F}_u\) for all \(u\le t\) and \(m_u=m_t \in \mathcal {F}_t\) for all \(u> t\). By definition of \(\hat{\pi }\), it follows that

    $$\begin{aligned} \hat{\pi }_{t,T}(\hat{X}+ \hat{m}_t) = \,&(X_0+m_0) \mathbb {1}_{\left\{ 0 \right\} } + \cdots + (X_{t-1}+m_{t-1} )\mathbb {1}_{\left\{ t-1\right\} } \\&+ (\pi _{t,T}(X_T+m_t -X_t-m_t )+X_t+m_t)\mathbb {1}_{\left\{ t,t+1,\ldots ,T\right\} }\\ = \,&\hat{\pi }_{t,T}(\hat{X})+\hat{m}_t. \end{aligned}$$

    Convexity of \(\hat{\pi }_{t,T}\) is a direct consequence of convexity of \(\pi _{t,T}\). Monotonicity Consider any \(\hat{X}, \hat{Y}\in L^\infty (\hat{\mathcal {F}})\) such that \(\hat{X}\le \hat{Y}\), that is \(X_u\le Y_u\)P-a.s. for any \(u\in \left\{ 0,1,\ldots ,T \right\} \). It follows that

    $$\begin{aligned} \hat{\pi }_{t,T}(\hat{X})&=X_0 \mathbb {1}_{\left\{ 0 \right\} } +\cdots + X_{t-1} \mathbb {1}_{\left\{ t-1 \right\} } + (\pi _{t,T}(X_T-X_t)+X_t) \mathbb {1}_{\left\{ t,\ldots ,T\right\} }\\&\le Y_0 \mathbb {1}_{\left\{ 0 \right\} } +\cdots + Y_{t-1} \mathbb {1}_{\left\{ t-1 \right\} } + (\pi _{t,T}(X_T-X_t)+X_t) \mathbb {1}_{\left\{ t,\ldots ,T\right\} }\\&\le Y_0 \mathbb {1}_{\left\{ 0 \right\} } +\cdots + Y_{t-1} \mathbb {1}_{\left\{ t-1 \right\} } + (\pi _{t,T}(X_T-Y_t)+Y_t-X_t+X_t) \mathbb {1}_{\left\{ t,\ldots ,T\right\} }\\&\le \hat{\pi }_{t,T}(\hat{Y}), \end{aligned}$$

    where the last two inequalities are due, respectively, to cash-subadditivity (together with \(Y_t \ge X_t\)) and to monotonicity of \(\pi _{t,T}\).

  2. (ii)

    Regularity Because of normalization of \(\hat{\pi }_{t,T}\), it is sufficient to verify that \(\hat{\pi }_{t,T}(\hat{X}\mathbb {1}_{{\hat{A}}}) = \hat{\pi }_{t,T}(\hat{X}) \mathbb {1}_{{\hat{A}}}\) for any \(\hat{X}\in L^\infty (\hat{\mathcal {F}})\) and \({\hat{A}}\in \hat{\mathcal {F}}_t\). By definition of \(\hat{\mathcal {F}}_t\), it is sufficient to prove regularity of \(\hat{\pi }_{t,T}\) on events of the form

    $$\begin{aligned} A_s\times \left\{ s\right\} \quad \text{ with } A_s\in \mathcal {F}_s \, (s<t), \quad \quad \text{ or } \quad \quad A_t\times \left\{ t,\ldots ,T \right\} \quad \text{ with } A_t\in \mathcal {F}_t. \end{aligned}$$

    Let us start by considering a set \({\hat{A}}= A_s\times \left\{ s\right\} \in \hat{\mathcal {F}}_t\) with \(s<t\) and \(A_s\in \mathcal {F}_s\). Then \(\hat{X}\mathbb {1}_{{\hat{A}}} = X_s\mathbb {1}_{A_s \times \left\{ s\right\} }\) and, consequently,

    $$\begin{aligned} \hat{\pi }_{t,T}(\hat{X}\mathbb {1}_{{\hat{A}}}) = X_s\mathbb {1}_{A_s \times \left\{ s\right\} } + (\pi _{t,T}(0 ) +0) \mathbb {1}_{\left\{ t, \ldots , T\right\} } = \hat{\pi }_{t,T}(\hat{X}) \mathbb {1}_{{\hat{A}}}. \end{aligned}$$

    Consider now a set \({\hat{A}}=A_t\times \left\{ t,\ldots ,T \right\} \) with \(A_t \in \mathcal {F}_t\). Hence

    $$\begin{aligned} \hat{X}\mathbb {1}_{{\hat{A}}} = 0 \mathbb {1}_{\left\{ 0, \ldots , t-1 \right\} }+X_u\mathbb {1}_{A_t \times \left\{ t,\ldots ,T\right\} } \end{aligned}$$

    and, because of the regularity of \(\pi _{t,T}\),

    $$\begin{aligned} \hat{\pi }_{t,T}(\hat{X}\mathbb {1}_{{\hat{A}}})&= 0 \mathbb {1}_{\left\{ 0, \ldots , t-1 \right\} }+ \big (\pi _{t,T}(X_T \mathbb {1}_{A_t }-X_t \mathbb {1}_{A_t } ) +X_t \mathbb {1}_{A_t }\big ) \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \\&= 0 \mathbb {1}_{\left\{ 0, \ldots , t-1\right\} } +(\pi _{t,T}(X_T-X_t ) +X_t)\mathbb {1}_{A_t } \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \\&= \hat{\pi }_{t,T}(\hat{X}) \mathbb {1}_{{\hat{A}}}. \end{aligned}$$
  3. (iii)

    Continuity from below Let \((\hat{X}^n)_{n \ge 0}\) be an arbitrary sequence with \(\hat{X}^n \nearrow _n \hat{X}\), i.e. \(X_u^n (\omega ) \nearrow _n X_u(\omega )\) for \(\hat{P}\)-a.e. \((\omega ,u) \in \hat{\Omega }\). Then

    $$\begin{aligned} 0 \le \,&\hat{\pi }_{t,T}(\hat{X}) - \hat{\pi }_{t,T}(\hat{X}^n) \\ = \,&(X_0-X_0^n) \mathbb {1}_{\left\{ 0\right\} } + \cdots + (X_{t-1}-X_{t-1}^n) \mathbb {1}_{\left\{ t-1\right\} } \\&+ \big ( \pi _{t,T}(X_T-X_t ) + X_t - \pi _{t,T}(X_T^n- X_t^n ) -X_t^n\big ) \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \\ \le \,&(X_0-X_0^n) \mathbb {1}_{\left\{ 0\right\} } + \cdots + (X_{t-1}-X_{t-1}^n) \mathbb {1}_{\left\{ t-1\right\} } \\&+ \big ( \pi _{t,T}(X_T-X_t ) + X_t - (\pi _{t,T}(X_T^n- X_t ) +X_t^n)\big ) \mathbb {1}_{\left\{ t,\ldots ,T\right\} }, \end{aligned}$$

    where the last inequality is due to monotonicity of \(\pi _{t,T}\) and to \(X^n_t \le X_t\) for any n. By continuity from below of \(\pi _{t,T}\) and by the assumption on \(X_u^n\), we can conclude that

    $$\begin{aligned} 0 \le \hat{\pi }_{t,T}(\hat{X}) - \hat{\pi }_{t,T}(\hat{X}^n) \rightarrow _n 0, \end{aligned}$$

    hence continuity from below of \(\hat{\pi }_{t,T}\).

  4. (iv)

    can be checked similarly as in (i)–(iii).

  5. (v)

    Time-consistency First, assume that \((\pi _{s,t}(\cdot - Z_s)+Z_s)_{0\le s\le t\le T}\) is time-consistent for any random variable \(Z_s\in L^{\infty }(\mathcal {F}_s)\). Consider an arbitrary \(\hat{X}\in L^\infty (\hat{\mathcal {F}}_v)\) and set \(\hat{Y}= \hat{\pi }_{t,v}(\hat{X})\). Hence, \(Y_u=X_u\) for \(u<t\), and \(Y_u=\pi _{t,v}(X_v-X_t)+X_t\) for \(u\ge t\). Then

    $$\begin{aligned} \hat{\pi }_{s,t}(\hat{\pi }_{t,v}(\hat{X}))&= \hat{\pi }_{s,t}(\hat{Y})\\&= Y_u \mathbb {1}_{\{0,1,\ldots ,s-1\}} + \big (\pi _{s,t}( Y_t-Y_s)+Y_s\big )\mathbb {1}_{\left\{ s,\ldots ,T\right\} } \\&=X_u \mathbb {1}_{\{0,1,\ldots ,s-1\}} + \big (\pi _{s,t}( \pi _{t,v}(X_v-X_t)+X_t-Y_s)+Y_s\big )\mathbb {1}_{\left\{ s,\ldots ,T\right\} } \\&= \hat{\pi }_{s,v}(\hat{X}), \end{aligned}$$

    where the last equality is due to X-time consistency of \(\pi \). Conversely, it is easy to check that time-consistency of \((\hat{\pi }_{s,t})_{0\le s\le t\le T}\) implies

    $$\begin{aligned} \pi _{s,t}\big (\pi _{t,v}(X_v-X_t)+X_t-X_s \big )=\pi _{t,v}(X_v-X_s)+X_s, \end{aligned}$$

    for all \((X_s)_{s\in \mathbb {T}}\) with \(X_s\) being \(\mathcal {F}_s\)-measurable, that is X-time-consistency of \((\pi _{s,t}(\cdot -X_s)+X_s)_{0\le s\le t\le T}\).

\(\square \)

In the next proposition, we show that the minimal penalty function of \((\pi _{s,t})_{s,t \in \mathbb {T},s \le t}\) and the minimal penalty function of \((\hat{\pi }_{s,t})_{s,t \in \mathbb {T},s \le t}\) coincide. We recall that the minimal penalty term \((c_{s,t})_{s,t \in \mathbb {T},s \le t}\) of a dynamic convex and cash-additive (respectively, cash-subadditive) risk measure \((\pi _{s,t})_{s,t \in \mathbb {T},s \le t}\) which is regular and continuous from below is given by

$$\begin{aligned} c_{s,t}(Q)= {\text {ess.sup}}_{X \in L^{\infty }(\mathcal {F}_t)} \{E_Q [X \big | \mathcal {F}_t] - \pi _{s,t}(X) \}, \end{aligned}$$
(28)

respectively

$$\begin{aligned} c_{s,t}(DQ)= {\text {ess.sup}}_{X \in L^{\infty }(\mathcal {F}_t)} \{D_{s,t} E_Q [X \big | \mathcal {F}_t] - \pi _{s,t}(X) \}. \end{aligned}$$
(29)

See Detlefsen and Scandolo (2005) and Mastrogiacomo and Rosazza Gianin (2015), respectively.

Proposition 17

Let \(c_{t,s}\) (resp. \(\hat{c}_{t,s}\)) be the minimal penalty function of \(\pi _{t,s}\) (resp. \(\hat{\pi }_{t,s}\)). For any probability measure \(\hat{Q}\) with decomposition \(\hat{Q}=Q\otimes \delta =Q \otimes \gamma \) where \(Q\in \mathcal {P}\), \(\delta \in \Delta (Q)\) and \(\gamma \in \Gamma (Q)\), it holds that

$$\begin{aligned} \hat{c}_{t,t+1}(\hat{Q}) =c_{t,t+1}(\tilde{\delta }_t Q)\mathbb {1}_{\{t,\ldots ,T\}} \end{aligned}$$
(30)

and, under the assumption that \(\gamma _{t+1}= \cdots = \gamma _{T-1}=0\),

$$\begin{aligned} \hat{c}_{t,T}(\hat{Q}) =c_{t,T}(\tilde{\delta }_t Q)\mathbb {1}_{\{t,\ldots ,T\}} \end{aligned}$$
(31)

where \(\tilde{\delta }_t= 1- \tilde{\gamma }_t =1- \frac{\gamma _t}{\delta _t}\).

Proof

Let \(Q\in \mathcal {P}\), \(\delta \in \Delta (Q)\) and \(\hat{Q}=Q\otimes \delta \). Since \( \hat{\pi }_{t,t+1}\) is a convex and cash-additive risk measure, its minimal penalty function is given by

$$\begin{aligned}&\hat{c}_{t,t+1}(\hat{Q}) \\&\quad = {\text {ess.sup}}_{\hat{X}\in L^\infty (\hat{\mathcal {F}}_{t+1})} \left\{ E_{\hat{Q}}[\hat{X}|\hat{\mathcal {F}}_t]- \hat{\pi }_{t,t+1}(\hat{X}) \right\} \\&\quad = {\text {ess.sup}}_{\hat{X}\in L^\infty (\hat{\mathcal {F}}_{t+1})} \nonumber \\&\qquad \times \left\{ E_{\hat{Q}}[\hat{X}|\hat{\mathcal {F}}_t]- X_u \mathbb {1}_{\left\{ 0, \ldots , t-1 \right\} } -\bigg (\pi _{t,t+1}(X_{t+1}-X_t)+X_t \bigg ) \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \right\} \\&\quad = {\text {ess.sup}}_{X_t \in L^{\infty }(\mathcal {F}_t), X_{t+1} \in L^{\infty }(\mathcal {F}_{t+1})} \nonumber \\&\qquad \times \left\{ E_Q\left[ \frac{\gamma _t}{\delta _t}X_t + \sum _{s=t+1}^T\frac{\gamma _{s}}{\delta _t}X_{t+1} \bigg |\mathcal {F}_t\right] - \bigg (\pi _{t,t+1}(X_{t+1}-X_t) +X_t\bigg ) \right\} \mathbb {1}_{\left\{ t,\ldots ,T\right\} }\\&\quad ={\text {ess.sup}}_{X_t \in L^{\infty }(\mathcal {F}_t), X_{t+1} \in L^{\infty }(\mathcal {F}_{t+1})} \nonumber \\&\qquad \times \left\{ (1- \tilde{\gamma }_t) E_Q\left[ X_{t+1} - X_t\big |\mathcal {F}_t\right] - \pi _{t,t+1}(X_{t+1}-X_t) \right\} \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \\&\quad ={\text {ess.sup}}_{Y \in L^{\infty }(\mathcal {F}_{t+1})} \left\{ (1- \tilde{\gamma }_t) E_Q\left[ Y \big | \mathcal {F}_t\right] - \pi _{t,t+1}(Y) \right\} \mathbb {1}_{\left\{ t,\ldots ,T\right\} }\\&\quad =c_{t,t+1}(\tilde{\delta }_t Q) \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \end{aligned}$$

where the third equality follows from (24), the forth follows by the definition of \(\hat{\mathcal {F}}_{t+1}\) and where \(\tilde{\gamma _s}\triangleq \frac{\gamma _s}{\delta _t}\ge 0\) for any \(s \ge t\) and \(\sum _{s=t}^T \tilde{\gamma }_s = 1\).

Similarly, the minimal penalty function of \(\hat{\pi }_{t,T}\) is given by

$$\begin{aligned} \hat{c}_{t,T}(\hat{Q})&= {\text {ess.sup}}_{\hat{X}\in L^\infty (\hat{\mathcal {F}})} \left\{ E_{\hat{Q}}[\hat{X}|\hat{\mathcal {F}}_t]- \hat{\pi }_{t,T}(\hat{X}) \right\} \\&= {\text {ess.sup}}_{\hat{X}\in L^\infty (\hat{\mathcal {F}})}\nonumber \\&\qquad \times \left\{ E_{\hat{Q}}[\hat{X}|\hat{\mathcal {F}}_t]- X_u \mathbb {1}_{\left\{ 0, \ldots , t-1 \right\} } -\bigg (\pi _{t,T}(X_T-X_t)+X_t \bigg ) \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \right\} \\&= {\text {ess.sup}}_{X_t \in L^{\infty }(\mathcal {F}_t), \ldots , X_T \in L^{\infty }(\mathcal {F})} \nonumber \\&\qquad \times \left\{ E_Q\left[ \sum _{s=t}^T \frac{\gamma _s}{\delta _t}X_s\bigg |\mathcal {F}_t\right] \mathbb {1}_{\left\{ t,\ldots ,T\right\} }- \bigg (\pi _{t,T}(X_T-X_t) +X_t\bigg )\mathbb {1}_{\left\{ t,\ldots ,T\right\} } \right\} \\&={\text {ess.sup}}_{X_t \in L^{\infty }(\mathcal {F}_t), \ldots , X_T \in L^{\infty }(\mathcal {F})} \nonumber \\&\qquad \times \left\{ E_Q\left[ \sum _{s=t}^T \tilde{\gamma }_sX_s\bigg |\mathcal {F}_t\right] - \bigg (\pi _{t,T}(X_T-X_t) +X_t\bigg ) \right\} \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \end{aligned}$$

By the assumption on \(\gamma \), it follows that \({\tilde{\gamma }}_{t+1}=\cdots ={\tilde{\gamma }}_{T-1}=0\) and, consequently,

$$\begin{aligned} \hat{c}_{t,T}(\hat{Q})&= {\text {ess.sup}}_{X_t \in L^{\infty }(\mathcal {F}_t), X_T \in L^{\infty }(\mathcal {F})} \\&\qquad \times \left\{ \tilde{\gamma }_t X_t + E_Q\left[ {\tilde{\gamma }}_T X_{T}\big |\mathcal {F}_t\right] - \bigg (\pi _{t,T}(X_{T}-X_t) +X_t\bigg )\right\} \mathbb {1}_{\left\{ t,\ldots ,T\right\} } \\&={\text {ess.sup}}_{X_t \in L^{\infty }(\mathcal {F}_t), X_T \in L^{\infty }(\mathcal {F})} \\&\qquad \left\{ (1- \tilde{\gamma }_t )E_Q\left[ X_{T}-X_t\big |\mathcal {F}_t\right] - \pi _{t,T}(X_{T}-X_t) \right\} \mathbb {1}_{\left\{ t,\ldots ,T\right\} }\\&={\text {ess.sup}}_{Y \in L^\infty (\mathcal {F})} \left\{ \tilde{\delta }_t E_Q\left[ Y\big |\mathcal {F}_t\right] - \pi _{t,T}(Y)\right\} \mathbb {1}_{\left\{ t,\ldots ,T\right\} }\\&= c_{t,T}(\tilde{\delta }_tQ) \mathbb {1}_{\left\{ t,\ldots ,T\right\} }. \end{aligned}$$

\(\square \)

We are now ready to prove that the generalized cocycle is a necessary condition for a convex and cash-subadditive dynamic risk measure (in a discrete-time setting) to be X-time-consistent.

Proof (Proof of Theorem 2)

Assume first that \((\pi _{t,v})_{t, v\in \mathbb {T}, t\le v}\) is normalized. From Proposition 16 it follows that \((\hat{\pi }_{t,v})_{t, v\in \mathbb {T}, t\le v}\) associated with \((\pi _{t,v})_{t, v\in \mathbb {T}, t\le v}\) as in (26) is a normalized, cash-additive and convex dynamic risk measure which is regular, continuous from below and time-consistent. Consequently, by Bion-Nadal (2008, Theorem 3.3) \((\hat{c}_{t,T})_{t\in \mathbb {T}}\) associated with \((\hat{\pi }_{t,T})_{t \in \mathbb {T}}\) satisfies the cocycle condition

$$\begin{aligned} \hat{c}_{t,T}(\hat{Q})= \hat{c}_{t,t+1}(\hat{Q})+ E_{\hat{Q}}\left[ \hat{c}_{t+1,T}(\hat{Q})\big |\hat{\mathcal {F}}_t\right] , \end{aligned}$$
(32)

for any \(\hat{Q}\in \hat{\mathcal {P}}\). Take then any \(\hat{Q}\in \hat{\mathcal {P}}\) such that \(\hat{Q}= Q \otimes \gamma \) with \(\gamma \) as in Proposition 17 and any \(Q \in \mathcal {P}\). Then by

$$\begin{aligned} \hat{c}_{t,t+1}(\hat{Q})&= c_{t,t+1}(\tilde{\delta }_t Q)\mathbb {1}_{\{t,\ldots ,T\}}\nonumber \\ \hat{c}_{t,T}(\hat{Q})&= c_{t,T}(\tilde{\delta }_t Q)\mathbb {1}_{\{t,\ldots ,T\}} \end{aligned}$$
(33)

and

$$\begin{aligned}&E_{\hat{Q}}\left[ \hat{c}_{t+1,T}(\hat{Q})\big |\hat{\mathcal {F}}_t\right] \\&\quad = 0 \mathbb {1}_{\left\{ 0\right\} } +\cdots + 0 \mathbb {1}_{\left\{ t-1\right\} } + E_{Q}\left[ \sum _{s=t}^T {\tilde{\gamma }}_s \hat{c}_{t+1,T}(\hat{Q})\big |_s\bigg |\mathcal {F}_t\right] \mathbb {1}_{\{t,\ldots ,T\}}\\&\quad = E_{Q}\left[ \underbrace{{\tilde{\gamma }}_t \cdot 0 + {\tilde{\gamma }}_{t+1} c_{t+1,T}(\tilde{\delta }_t Q) + \cdots + }_{=0}{\tilde{\gamma }}_{T} c_{t+1,T}(\tilde{\delta }_t Q)\bigg |\hat{\mathcal {F}}_t\right] \mathbb {1}_{\{t,\ldots ,T\}} \\&\quad = E_{Q}\left[ (1-{\tilde{\gamma }}_{t}) c_{t+1,T}(\tilde{\delta }_t Q)\big |\mathcal {F}_t\right] \mathbb {1}_{\{t,\ldots ,T\}}\\&\quad =\tilde{\delta }_t E_{Q}\left[ c_{t+1,T}(\tilde{\delta }_t Q)\big |\mathcal {F}_t\right] \mathbb {1}_{\{t,\ldots ,T\}}, \end{aligned}$$

it follows that

$$\begin{aligned} E_{\hat{Q}}\left[ \hat{c}_{t+1,T}(\hat{Q})\big |\hat{\mathcal {F}}_t\right] =\tilde{\delta }_t E_{Q}\left[ c_{t+1,T}(\tilde{\delta }_t Q)\big |\mathcal {F}_t\right] \mathbb {1}_{\{t,\ldots ,T\}} \end{aligned}$$
(34)

Combining (32)–(34) we obtain

$$\begin{aligned} c_{t,T}(\tilde{\delta }_t Q)= c_{t,t+1}(\tilde{\delta }_t Q) + \tilde{\delta }_t E_{Q}\left[ c_{t+1,T}(\tilde{\delta }_t Q)\big |\mathcal {F}_t\right] . \end{aligned}$$
(35)

Since Q and \(\gamma \) were arbitrarily chosen, condition (35) holds for any \(Q \in \mathcal {P}\) and \(\tilde{\delta }_t \in \mathcal {D}\). Suppose now that \((\pi _{t,u})_{t, u\in \mathbb {T}, t\le u}\) is not necessarily normalized. Since \(\bar{\pi }_{s,u}(X) \triangleq \pi _{s,u}(X) - \pi _{s,u}(0)\) is a normalized, convex and cash-subadditive risk measure with minimal penalty term \(\bar{c}_{s,u}(DQ) = c_{s,u}(DQ) + \pi _{s,u}(0) = c_{s,u}(DQ) - {\text {ess.inf}}_{D ,Q} c_{s,u}(DQ)\), the last statement of Theorem 2 then follows by applying the first part of the result.\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mastrogiacomo, E., Rosazza Gianin, E. Time-consistency of risk measures: how strong is such a property?. Decisions Econ Finan 42, 287–317 (2019). https://doi.org/10.1007/s10203-019-00233-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10203-019-00233-2

Keywords

Mathematics Subject Classification

JEL Classification

Navigation