Skip to main content
Log in

Preference robust state-dependent distortion risk measure on act space and its application in optimal decision making

  • Original Paper
  • Published:
Computational Management Science Aims and scope Submit manuscript

Abstract

Decision-making under uncertainty involves three fundamental components: acts, states of nature, and consequences, as first introduced by Savage (The foundations of statistics, Wiley, New York, 1954). An act is uniquely determined by a number of random consequences that are associated with different states of nature. If the consequences are identical across all states of nature, then the act is state-independent. Prior research on distortion risk measures (DRMs) has primarily focused on state-independent acts. In this paper, we extend the research to state-dependent acts by introducing a state-dependent DRM (SDRM) under the Anscombe–Aumann’s framework (Anscombe and Aumannin in Ann Math Stat 34(1):199–205, 1963). The proposed SDRM is the weighted average of DRMs at each state, where the weights are determined by the decision maker’s (DM’s) subjective probabilities of the states. In situations where there is incomplete information about the DM’s true distortion function and/or the true subjective probabilities of the states, we introduce a preference robust SDRM (PRSDRM) for acts. The PRSDRM is based on the worst-case state-dependent distortion function and the worst-case subjective probabilities over a dependent joint ambiguity set constructed with partially available information. To compute the PRSDRM numerically, we show that when the distortion functions are concave, it can be formulated as a biconvex program and further as a convex program by changing some variables. As a motivation and application, we use the PRSDRM for decision-making problems and propose an alternating iterative algorithm for solving it. Finally, we conduct numerical experiments to assess the performance of our proposed model and computational scheme.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. Strictly speaking, Yaari considers prospects of random gains as the value of a choice function is parallel to a utility value, but this can be applied to prospects of random losses straightforwardly.

  2. Yaari considers random variables taking values over [0, 1] and his representation is \(\int _0^1 g(1-F_X(x))dx \le \int _0^1 g(1-F_Y(x))dx\), where g is an increasing function with \(g(0)=0\) and \(g(1)=1\). But this formulation can be easily generalized to the case when the random variables take values over \({\mathrm{I\!R}}\) and the inequality is equivalent to (1.1), see Wang and Xu (2023, Appendix A.1). The dual expected utility theory is in the sense that the distortion function g may be viewed as a “utility function” defined in the space of decumulative distribution functions (DDF) and the integral is the average of the “utility” of the DDF. Moreover, if we view the integral as a function of random variables, then it is a law invariant monetary risk measure satisfying monotonicity, translation invariance and co-monotonic additivity, see, e.g. Wang (1996), Wang et al. (2020).

  3. The Lebesgue-Stieltjes integral \(\int _{{\mathrm{I\!R}}} f(x)dg(x)\) is defined, when f is Borel-measurable and finite and g is monotonically increasing, as \(\int _{{\mathrm{I\!R}}} f(x)d\mu _g(x)\) with \(\mu _g([a,b])=g(b+)-g(a-)\) for any \(a\le b\) where \(g(b+):=\lim _{x\downarrow b}g(x)\) and \(g(a-):=\lim _{x\uparrow a}g(x)\), see, e.g. Carter and Van Brunt (2000).

  4. In Anscombe and Aumann (1963), consequences refer to random profits. In this paper, consequences refer to random losses. Consequently we use terminology “penalties” instead of “prizes”.

  5. If Q is a modification of the true probability distribution of s, then SDRM is not actually risk neutral.

  6. To ease the exposition, we slightly abuse the notation by writing \(Q^0\) and \(g^0\) for \(\{Q^0\}\) and \(\{g^0\}\) when the ambiguity sets reduce to a singleton.

References

  • Abdellaoui M, Bleichrodt H, Kemel E, l’Haridon O (2021) Measuring beliefs under ambiguity. Oper Res 69(2):599–612

    Google Scholar 

  • Acerbi C (2002) Spectral measures of risk: a coherent representation of subjective risk aversion. J Bank Finance 26(7):1505–1518

    Google Scholar 

  • Allais M (1953) Le comportement de l’homme rationnel devant le risque: critique des postulats et axiomes de l’école américaine. Econom J Econom Soc 503–546

  • Anderson DR, Sweeney DJ, Williams TA, Camm JD, Cochran JJ (2018) An introduction to management science: quantitative approach. Cengage Learn

  • Anscombe FJ, Aumann RJ (1963) A definition of subjective probability. Ann Math Stat 34(1):199–205

    Google Scholar 

  • Armbruster B, Delage E (2015) Decision making under uncertainty when preference information is incomplete. Manage Sci 61(1):111–128

    Google Scholar 

  • Arrow KJ (1974) Optimal insurance and generalized deductibles. Scand Actuar J 1974(1):1–42

    Google Scholar 

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

    Google Scholar 

  • Baccelli J (2021) The problem of state-dependent utility: a reappraisal. Br J Philos Sci 72(2):617–634

    Google Scholar 

  • Baillon A, Bleichrodt H (2015) Testing ambiguity models through the measurement of probabilities for gains and losses. Am Econ J Microecon 7(2):77–100

    Google Scholar 

  • Belles-Sampera J, Guillén M, Santolino M (2014) Beyond value-at-risk: Gluevar distortion risk measures. Risk Anal 34(1):121–134

    Google Scholar 

  • Belles-Sampera J, Guillen M, Santolino M (2016) What attitudes to risk underlie distortion risk measure choices? Insur Math Econ 68:101–109

    Google Scholar 

  • Ben-Tal A, Teboulle M (1987) Penalty functions and duality in stochastic programming via φ-divergence functionals. Math Oper Res 12(2):224–240

    Google Scholar 

  • Berkhouch M, Lakhnati G, Righi M (2019) Spectral risk measures and uncertainty. Available at SSRN 3390654

  • Boonen TJ et al (2015) Competitive equilibria with distortion risk measures. Astin Bull 45(3):703–728

    Google Scholar 

  • Brown DB, Giorgi ED, Sim M (2012) Aspirational preferences and their representation by risk measures. Manage Sci 58(11):2095–2113

    Google Scholar 

  • Bühlmann H (1985) Premium calculation from top down. ASTIN Bull J IAA 15(2):89–101

    Google Scholar 

  • Campbell JY (2006) Household finance. J Finance 61(4):1553–1604

    Google Scholar 

  • Carter M, Van Brunt B (2000) The Lebesgue–Stieltjes integral. Springer, Berlin, pp 49–70

    Google Scholar 

  • Castaño-Martínez A, López-Blazquez F, Pigueiras G, Sordo MÁ (2020) A method for constructing and interpreting some weighted premium principles. ASTIN Bull. J. IAA 50(3):1037–1064

    Google Scholar 

  • Cerreia-Vioglio S, Dillenberger D, Ortoleva P (2015) Cautious expected utility and the certainty effect. Econometrica 83(2):693–728

    Google Scholar 

  • Csiszár I (1967) Information-type measures of difference of probability distributions and indirect observation. Stud Sci Math Hung 2:229–318

    Google Scholar 

  • Delage E, Li JY-M (2017) Minimizing risk exposure when the choice of a risk measure is ambiguous. Manage Sci 64(1):327–344

    Google Scholar 

  • Delage E, Guo S, Xu H (2022) Shortfall risk models when information on loss function is incomplete. Oper Res 70(6):3511–3518

    Google Scholar 

  • Denneberg D (1990a) Distorted probabilities and insurance premiums. Methods Oper Res 63(3):3–5

    Google Scholar 

  • Denneberg D (1990b) Premium calculation: why standard deviation should be replaced by absolute deviation. ASTIN Bull J IAA 20(2):181–190

    Google Scholar 

  • Dhaene J, Kukush A, Linders D, Tang Q (2012) Remarks on quantiles and distortion risk measures. Eur Actuar J 2(2):319–328

    Google Scholar 

  • Escobar DD, Pflug GC (2018) The distortion principle for insurance pricing: properties, identification and robustness. Ann Oper Res 292:771–794

    Google Scholar 

  • Fan K (1953) Minimax theorems. Proc Natl Acad Sci USA 39(1):42

    Google Scholar 

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

    Google Scholar 

  • French KR, Poterba JM (1991) Investor diversification and international equity markets. Am Econ Rev 81(2):222–226

    Google Scholar 

  • Furman E, Wang R, Zitikis R (2017) Gini-type measures of risk and variability: Gini shortfall, capital allocations, and heavy-tailed risks. J Bank Finance 83:70–84

    Google Scholar 

  • Gilboa I, Schmeidler D (1989) Maxmin expected utility with non-unique prior. J Math Econ 18(2):141–153

    Google Scholar 

  • Goovaerts MJ, de Vylder F, Haezendonck J (1984) Insurance premiums: theory and applications. Elsevier Science, New York

    Google Scholar 

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

    Google Scholar 

  • Guo S, Xu H (2021) Robust spectral risk optimization when the subjective risk aversion is ambiguous: a moment-type approach. Math Program 194:1–36

    Google Scholar 

  • Guo S, Xu H, Zhang S (2023) Utility preference robust optimization with moment-type information structure. Oper Res. https://doi.org/10.1287/opre.2023.2464

    Article  Google Scholar 

  • Hardy MR (2001) A regime-switching model of long-term stock returns. N Am Actuar J 5(2):41–53

    Google Scholar 

  • Haskell WB, Huang W, Xu H (2018) Preference elicitation and robust optimization with multi-attribute quasi-concave choice functions. arXiv preprint arXiv:1805.06632

  • He XD, Kou S, Peng X (2022) Risk measures: robustness, elicitability, and backtesting. Ann Rev Stat Appl 9:141–166

    Google Scholar 

  • Hill B (2019) A non-Bayesian theory of state-dependent utility. Econometrica 87(4):1341–1366

    Google Scholar 

  • Hu J, Mehrotra S (2015) Robust decision making over a set of random targets or risk-averse utilities with an application to portfolio optimization. IIE Trans 47(4):358–372

    Google Scholar 

  • Hu J, Homem-de Mello T, Mehrotra S (2011) Risk-adjusted budget allocation models with application in homeland security. IIE Trans 43(12):819–839

    Google Scholar 

  • Hu J, Bansal M, Mehrotra S (2018) Robust decision making using a general utility set. Eur J Oper Res 269(2):699–714

    Google Scholar 

  • Jarrow R, Li S (2021) Concavity, stochastic utility, and risk aversion. Finance Stoch 25(2):311–330

    Google Scholar 

  • Jiang W, Hong H, Ren J (2018) On pareto-optimal reinsurance with constraints under distortion risk measures. Eur Actuar J 8(1):215–243

    Google Scholar 

  • Jiang W, Escobar-Anel M, Ren J (2020) Optimal insurance contracts under distortion risk measures with ambiguity aversion. ASTIN Bull J IAA 50(2):619–646

    Google Scholar 

  • Karlin S (1959) Mathematical methods and theory in games, programming and economics. Addison-Wesley, Reading

    Google Scholar 

  • Karmarkar US (1978) Subjectively weighted utility: a descriptive extension of the expected utility model. Organ Behav Hum Perform 21(1):61–72

    Google Scholar 

  • Karni E (1983) Risk aversion for state-dependent utility functions: measurement and applications. Int Econ Rev 637–647

  • Karni E, Schmeidler D, Vind K (1983) On state dependent preferences and subjective probabilities. Econom J Econom Soc 1021–1031

  • Li JY-M (2021) Inverse optimization of convex risk functions. Manage Sci 67(11):7113–7141

    Google Scholar 

  • Li M, Tong X, Xu H (2022) Randomization of spectral risk measure and distributional robustness. arXiv preprint arXiv:2212.08871

  • Liu J, Chen Z, Xu H (2021) Multistage utility preference robust optimization. arXiv preprint arXiv:2109.04789

  • Lu J (2019) Bayesian identification: a theory for state-dependent utilities. Am Econ Rev 109(9):3192–3228

    Google Scholar 

  • Maccheroni F (2002) Maxmin under risk. Econ Theor 19(4):823–831

    Google Scholar 

  • Maccheroni F, Marinacci M, Rustichini A (2006) Ambiguity aversion, robustness, and the variational representation of preferences. Econometrica 74(6):1447–1498

    Google Scholar 

  • Merkle M, Marinescu D, Merkle MM, Monea M, Stroe M (2014) Lebesgue–Stieltjes integral and young’s inequality. Appl Anal Discrete Math 60–72

  • Mononen L (2020) State-dependent utility and ambiguity. Technical report, Working paper

  • Morimoto T (1963) Markov processes and the h-theorem. J Phys Soc Jpn 18(3):328–331

    Google Scholar 

  • Neth S (2019) Measuring belief and risk attitude. arXiv preprint arXiv:1907.09115

  • Pardo L (2005) Statistical inference based on divergence measures. CRC Press, Boca Raton

    Google Scholar 

  • Pesenti SM, Wang Q, Wang R (2020) Optimizing distortion riskmetrics with distributional uncertainty. Available at SSRN

  • Pflug GC (2000) Some remarks on the value-at-risk and the conditional value-at-risk. In: Probabilistic constrained optimization. Springer, pp 272–281

  • Pichler A, Shapiro A (2015) Minimal representation of insurance prices. Insur Math Econ 62:184–193

    Google Scholar 

  • Puppe C (2012) Distorted probabilities and choice under risk, vol 363. Springer, Berlin

    Google Scholar 

  • Quiggin J (1982) A theory of anticipated utility. J Econ Behav Organ 3(4):323–343

    Google Scholar 

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

    Google Scholar 

  • Savage LJ (1954) The foundations of statistics. Wiley, New York

    Google Scholar 

  • Schied A, Föllmer H, Weber S (2009) Robust preferences and robust portfolio choice. Handb Numer Anal 15:29–87

    Google Scholar 

  • Sereda EN, Bronshtein EM, Rachev ST, Fabozzi FJ, Sun W, Stoyanov SV (2010) Distortion risk measures in portfolio optimization. In: Handbook of portfolio construction. Springer, pp 649–673

  • Shapiro A, Dentcheva D, Ruszczynski A (2021) Lectures on stochastic programming: modeling and theory. SIAM

  • Slovic P (1999) Trust, emotion, sex, politics, and science: surveying the risk-assessment battlefield. Risk Anal 19(4):689–701

    Google Scholar 

  • Sordo MA, Castaño-Martínez A, Pigueiras G (2016) A family of premium principles based on mixtures of tvars. Insur Math Econ 70:397–405

    Google Scholar 

  • Tasche D (2002) Expected shortfall and beyond. J Bank Finance 26(7):1519–1533

    Google Scholar 

  • Thurstone LL (1927) A law of comparative judgment. Psychol Rev 34(4):273

    Google Scholar 

  • Tsanakas A, Desli E (2003) Risk measures and theories of choice. Br Actuar J 9(4):959–991

    Google Scholar 

  • Vayanos P, Ye Y, McElfresh D, Dickerson J, Rice E (2020) Robust active preference elicitation. arXiv preprint arXiv:2003.01899

  • Von Neumann J, Morgenstern O (1947) Theory of games and economic behavior, 2nd rev

  • Wakker P, Deneffe D (1996) Eliciting von Neumann–Morgenstern utilities when probabilities are distorted or unknown. Manage Sci 42(8):1131–1150

    Google Scholar 

  • Wang S (1995) Insurance pricing and increased limits ratemaking by proportional hazards transforms. Insur Math Econ 17(1):43–54

    Google Scholar 

  • Wang S (1996) Premium calculation by transforming the layer premium density. ASTIN Bull J IAA 26(1):71–92

    Google Scholar 

  • Wang W, Xu H (2020) Robust spectral risk optimization when information on risk spectrum is incomplete. SIAM J Optim 30(4):3198–3229

    Google Scholar 

  • Wang W, Xu H (2023) Preference robust distortion risk measure and its application. Math Finance 33:389–434

    Google Scholar 

  • Wang Q, Wang R, Wei Y (2020) Distortion riskmetrics on general spaces. ASTIN Bull J IAA 50(3):827–851

    Google Scholar 

  • Wang S, Dhaene J, et al (1997) Comonotonicity, correlation order and premium principles. Citeseer

  • Weber M (1987) Decision making with incomplete information. Eur J Oper Res 28(1):44–57

    Google Scholar 

  • Weber S (2006) Distribution-invariant risk measures, information, and dynamic consistency. Math Finance Int J Math Stat Financ Econ 16(2):419–441

    Google Scholar 

  • Wirch JL, Hardy MR (2001) Distortion risk measures: coherence and stochastic dominance. In: International congress on insurance: mathematics and economics, pp 15–17

  • Wu G, Gonzalez R (1996) Curvature of the probability weighting function. Manage Sci 42(12):1676–1690

    Google Scholar 

  • Yaari ME (1987) The dual theory of choice under risk. Econom J Econom Soc 95–115

  • Yitzhaki S (1982) Stochastic dominance, mean variance, and Gini’s mean difference. Am Econ Rev 72(1):178–185

    Google Scholar 

  • Zhang Y, Xu H, Wang W (2020) Preference robust models in multivariate utility-based shortfall risk minimization. Optim Methods Softw 37:712–752

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Huifu Xu.

Additional information

Publisher's Note

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

This work is supported by RGC Grant 14500620 and CUHK start-up grant.

Appendices

Appendix A Supplementary materials

1.1 Some examples of DRMs

Example A.1

In this example, we assume that X is a non-negative random variable.

(i) Value at Risk (VaR). Let \(g_{\nu }(t)={\textbf{1}}_{(1-\nu ,1]}(t)\), where \({\textbf{1}}_{[\nu ,1]}(t)\) is the indicator function and \(\nu \in (0,1)\). Observe that \(F_X(x)\in [0,\nu )\) iff \(S_X(x)\in (1-\nu ,1]\), that is, \(x\in [0,F_X^{\leftarrow }(\nu ))\) iff \(g_{\nu }(S_X(x))=1\). Consequently, it follows from (2.1) that

$$\begin{aligned} \rho _{g_{\nu }}(X)= \int _0^{F_X^{-1}(\nu )} dx=F_X^{-1}(\nu )=:\text{ VaR}_{\nu }(X). \end{aligned}$$

(ii) Conditional Value at Risk \(\text{(CVaR) }\) (also called Expected Shortfall (Tasche 2002)). Let \(g_{\alpha }(t)=\min \left\{ \frac{t}{1-\alpha },1\right\} \) with \(\alpha \in (0,1)\). From (2.1), using integration by parts of Lebesgue-Stieltjes integral (see, for instance, Merkle et al. (2014)), we have

$$\begin{aligned} \rho _{g_{\alpha }}(X)= & {} \int _0^{F_X^{\leftarrow }(\alpha )} 1 dx +\int _{F_X^{\leftarrow }(\alpha )}^{\infty }\frac{S_X(x)}{1-\alpha }dx\nonumber \\= & {} F_X^{\leftarrow }(\alpha ) + \frac{1}{1-\alpha }\left[ S_X(x)x\big |_{F_X^{\leftarrow }(\alpha )_-}^\infty -\int _{F_X^{\leftarrow }(\alpha )}^{\infty }xdS_X(x)\right] \nonumber \\= & {} \frac{1}{1-\alpha }\left\{ {\mathbb {E}}[X{\textbf{1}}_{X\ge F_{X}^{\leftarrow }(\alpha )}] -F_X^{\leftarrow }(\alpha )[ P(X<F_X^{\leftarrow }(\alpha ))-\alpha ] \right\} \nonumber \\=: & {} \text{ ES}_\alpha (X) =\text{ CVaR}_\alpha (X), \end{aligned}$$
(A.1)

where \((\cdot )_- = \lim _{\delta \downarrow 0} (\cdot -\delta )\) and the third equality is due to the fact that \(\lim _{x\rightarrow \infty } xg(S_X(x))=0\) and \(-\int _{F_X^{\leftarrow }(\alpha )}^{\infty }xdS_X(x)=\mathbb E[X{\textbf{1}}_{X\ge F_X^{\leftarrow }(\alpha )}]\). The last equality is perhaps known, we refer readers to Wang and Xu (2023) for the detailed discussion about this. The relation can also be derived from (2.2), where we have

$$\begin{aligned} \rho _{g_{\alpha }}(X)=\int _0^1 F_X^{\leftarrow }(t)d{\tilde{g}}_{\alpha }(t)=\frac{1}{1-\alpha } \int _{\alpha }^1F_X^{\leftarrow }(t)dt=:\text{ CVaR}_{\alpha }(X), \end{aligned}$$
(A.2)

which is shown in Acerbi (2002).

From an insurance premium point of view, it might be more advantageous to have a non-flat tail distortion function because the whole loss distribution will be utilized, see, for instance, Wang (1995, 1996) and Wang et al. (1997) on premium principles.

(iii) Proportional hazards transform risk measure (Wang 1995). Let \(g_{\gamma }(t)=t^{\frac{1}{\gamma }}\) for \(\gamma >1\). Then

$$\begin{aligned} \rho _{g_{\gamma }}(X)=\int _0^{\infty } S_X(x)^{\frac{1}{\gamma }}dx=:\rho _{\textrm{PH}}(X). \end{aligned}$$

(iv) Gini’s risk measure (Denneberg 1990b). Let \(g_s(t)=t-s(t^2-t)\) for \(s\in (0,1)\). Then from (2.2),

$$\begin{aligned} \rho _{g_s}(X)= & {} \int _0^1 F_X^{\leftarrow }(1-t)dg_s(t)=\int _0^1 F_X^{\leftarrow }(1-t)[1-s(2t-1)]dt\\= & {} \int _0^1 F_X^{\leftarrow }(1-t)dt - s \int _0^1 F_X^{\leftarrow }(1-t)(2t-1)dt\\= & {} {\mathbb {E}}[X]+s\int _0^1 F_X^{\leftarrow }(t)(2t-1)dt= {\mathbb {E}}[X]+\frac{1}{2}s{\mathbb {E}}[|X-X'|]=:\text{ Gini}_s(X), \end{aligned}$$

where \(X'\) is an independent copy of X. The second last equality is due to the fact that \({\mathbb {E}}[|X-X'|]=2\int _0^1 (2t-1)F_X^{\leftarrow }(t)dt\) (can be calculated easily with Fubini’s Theorem, or see, for instance, Furman et al. (2017)), which is called the Gini’s mean difference and is closely related to stochastic dominance, see, for instance, Yitzhaki (1982).

1.2 Proof of Theorem 4.1

Proof

We use Lemma 4.1 to prove the result and do so in four steps.

Step 1. By the definition of PRSDRM of a state-dependent act \({\varvec{X}}\) and the structure of \(({\mathcal {G}},{\mathcal {Q}}_{\epsilon })_1\), we have

$$\begin{aligned} {\mathcal {T}}_{({\mathcal {G}},{\mathcal {Q}}_{\epsilon })_1}({\varvec{X}})= \sup _{(g,Q)\in ({\mathcal {G}},{\mathcal {Q}}_{\epsilon })_1}{\mathcal {T}}_{Q}^g({\varvec{X}})=\sup _{Q\in {\mathcal {Q}}_{\epsilon }} \sup _{(g,Q)\in ({\mathcal {G}},Q)_1}{\mathcal {T}}_{Q}^g({\varvec{X}}), \end{aligned}$$

where \(({\mathcal {G}},Q)_1=({\mathcal {G}},Q)_{pair}\cap ({\mathcal {G}},Q)_{ce}\cap ({\mathcal {G}}_{coh}\times Q)\) and \(Q\in {\mathcal {Q}}_{\epsilon }\). Let \(\varphi (Q)=\sup _{(g,Q)\in ({\mathcal {G}},Q)_1} {\mathcal {T}}_{Q}^g({\varvec{X}})\). Then \({\mathcal {T}}_{{\mathcal {Q}}_{\epsilon }}^{{\mathcal {G}}_1}({\varvec{X}})=\sup _{Q\in {\mathcal {Q}}_{\epsilon }}\varphi (Q)\). It suffices to derive the tractable reformulation of \(\varphi (Q)\) for a specified probability distribution \(Q=(q_1,q_2,\ldots ,q_n)\in {\mathcal {Q}}_{\epsilon }\). Throughout the proof, we use \({\mathcal {G}}_1(Q)=\{g\in {\mathcal {G}}:(g,Q)\in ({\mathcal {G}},Q)_1\}\) to emphasize the dependence of the ambiguity set of state-dependent distortion functions on the probability distribution Q of the belief of the states of nature. Similar notation for \({\mathcal {G}}_{pair}(Q)\), \({\mathcal {G}}_{ce}(Q)\), and \({\mathcal {G}}_{coh}(Q)\) represent \(({\mathcal {G}},Q)_{pair}\), \(({\mathcal {G}},Q)_{ce}\) and \({\mathcal {G}}_{coh}\times Q\) respectively.

Step 2. Let \({\overline{g}}\in {\mathcal {G}}_1(Q)\), \(\xi _j^s\in \Xi ^s\) and \(v_j^s={\overline{g}}(1-\xi _j^s,s)\) for \(j=0,1,\ldots ,J_s\) and \(s=1,\ldots ,n\). Then by the definition of distortion function, we have \(v_0^s={\overline{g}}(1,s)=1\) and \(v_{J_s}^s={\overline{g}}(0,s)=0\) for \(s=1,\ldots ,n\). Let \(\tau _i^s={\overline{g}}(1-\pi _i^s,s)\) for \(i=1,\ldots ,n_s\) and \(s=1,\ldots ,n\). Then \(\tau _0^s={\overline{g}}(1,s)=1\) and \(\tau _{n_s}^s={\overline{g}}(0,s)=0\) for \(s=1,\ldots ,n\). For the fixed \(v^s=(v_0^s,v_1^s,\ldots ,v_{J_s}^s)\), define

$$\begin{aligned} {\mathcal {G}}^s(v^s):=\{g\in {\mathcal {G}}^s:g(1-\xi _j^s,s)=v_j^s\;\text {for}\; j=0,1,\ldots ,J_s\}, \end{aligned}$$

where \({\mathcal {G}}^s\) represents all distortion functions for the state s, and for the fixed \(v=(v^1,\ldots ,v^n)\), define

$$\begin{aligned} {\mathcal {G}}(v)=\{g\in {\mathcal {G}}:g(1-\xi _j^s,s)=v_j^s\;\text {for}\; j=0,1,\ldots ,J_s;s=1,\ldots ,n\}. \end{aligned}$$

Thus, \({\mathcal {G}}(v)=\times _{s=1}^n {\mathcal {G}}^s(v^s)\). By the definition, \({\mathcal {G}}(v)\in {\mathcal {G}}\) and \({\overline{g}}\in {\mathcal {G}}(v)\). Then, \({\mathcal {G}}(v)\ne \emptyset \). On the other hand, for any \(g\in {\mathcal {G}}\), there exists \({\tilde{v}}=({\tilde{v}}^1,\ldots ,{\tilde{v}}^n)\) with \({\tilde{v}}^s\in {\mathrm{I\!R}}^{J_s+1}\) for \(s=1,\ldots ,n\) such that \(g\in {\mathcal {G}}({\tilde{v}})\). Thus,

$$\begin{aligned} {\mathcal {G}}=\bigcup _{v} {\mathcal {G}}(v)\;\text {and}\; {\mathcal {G}}_1(Q)=\bigcup _{v} ({\mathcal {G}}_1(Q)\cap {\mathcal {G}}(v)) \end{aligned}$$

and consequently we can write \(\varphi (Q)\) as

$$\begin{aligned} \varphi (Q)={\mathcal {T}}_Q^{{\mathcal {G}}_1}({\varvec{X}}) :=\sup _{v} {\mathcal {T}}_Q^{{\mathcal {G}}(v)}({\varvec{X}}) \quad \text{ s.t. } \quad {\mathcal {G}}(v)\cap {\mathcal {G}}_1(Q)\ne \emptyset . \end{aligned}$$

Note that \({\mathcal {G}}(v)\) defines the set of all functions in \({\mathcal {G}}\) whose values on \(\Xi ^s\) are \(v^s=(v_0^s,v_1^s,\ldots ,v_{J_s}^s)\) for \(s=1,\ldots ,n\), whereas \({\mathcal {G}}_{ce}(Q)\) is a set of specific functions in \({\mathcal {G}}\) which satisfy certainty equivalent conditions and their values are determined on a subset of \(\Xi ^s\) for \(s=1,\ldots ,n\). Moreover, since \({\mathcal {G}}(v)\) is determined by v, then for fixed v, either v satisfies the certainty equivalent conditions or not, which implies that either \({\mathcal {G}}(v)\) is a subset of \({\mathcal {G}}_{ce}(Q)\) or is disjoint from it. The same is true for the set \({\mathcal {G}}_{pair}(Q)\). Since \({\mathcal {G}}_1(Q)={\mathcal {G}}_{pair}(Q)\cap {\mathcal {G}}_{ce}(Q)\cap {\mathcal {G}}_{coh}(Q)\), it follows that

$$\begin{aligned} \varphi (Q)=\sup _{v}{} & {} {\mathcal {T}}_Q^{{\mathcal {G}}(v)\cap {\mathcal {G}}_{coh}(Q)}({\varvec{X}}) \end{aligned}$$
(A.4a)
$$\begin{aligned} \text{ s.t. }{} & {} {\mathcal {G}}(v)\cap {\mathcal {G}}_{coh}(Q)\ne \emptyset , {\mathcal {G}}(v)\subset {\mathcal {G}}_{pair}(Q), {\mathcal {G}}(v)\subset {\mathcal {G}}_{ce}(Q). \end{aligned}$$
(A.4b)

Step 3. Let \({\varvec{Y}}\) be a state-dependent act with consequence \(Y_s\) being a finite discretely distributed non-negative random loss such that \({\mathbb {P}}(Y_s=y_i^s)=q_i^s\) for \(i=1,\ldots ,L_s\) where \(y_1^s<y_2^s<\cdots <y_{L_s}^s\) for \(s=1,\ldots ,n\). We want to represent \({\mathcal {T}}_Q^g({\varvec{Y}})\) via \(\{(1-\xi _j^s,v_j^s)\}_{j=0,1}^{J_s}\) for \(s=1,\ldots ,n\) if the breakpoints of the quantile function of \(Y_s\) are contained in \(\Xi ^s\) for \(s=1,\ldots ,n\). Since the quantile function of \(Y_s\) is a step-like function with

$$\begin{aligned} F_{Y_s}^{\leftarrow }(t)=y_{l+1}^s,\;\text{ for }\; t\in (\pi _l^s,\pi _{l+1}^s]\;\text{ and }\;l=0,1,\ldots ,L_s, \end{aligned}$$

where \(y_0^s=0\), \(\pi _0^s=0\), \(\pi _{L_s+1}^s=1\) and \(\pi _l^s=\sum _{j\le l}q_j^s\) for \(l=1,\ldots ,L_s\) and \(s=1,\ldots ,n\). Moreover, since \(\pi _l^s\in \Xi ^s\) for \(l=1,\ldots ,L_s+1\) and \(s=1,\ldots ,n\) and \(\xi _j^s\) is the j-th smallest element in set \(\Xi ^s\) for \(s=1,\ldots ,n\), then by (4.2) we have

$$\begin{aligned} {\mathcal {T}}_Q^g({\varvec{Y}})= & {} \sum _{s=1}^n\sum _{j=1}^{L_s} q_s(y_j^s-y_{j-1}^s)g(1-\pi _{j-1}^s,s)\nonumber \\= & {} \sum _{s=1}^n\sum _{j=1}^{J_s}q_s[F_{Y_s}^{\leftarrow }(\xi _j^s)-F_{Y_s}^{\leftarrow }(\xi _{j-1}^s)]v_{j-1}^s. \end{aligned}$$
(A.5)

Consequently, \({\mathcal {T}}_Q^g({\varvec{W}}_k),{\mathcal {T}}_Q^g({\varvec{G}}_m)\) and \({\mathcal {T}}_Q^g({\varvec{B}}_m)\) have the similar representations of (A.5) for \(k=1,\ldots ,K\) and \(m=1,\ldots ,M\).

Step 4. We are now ready to reformulate problem (A.4). Let \(h_j^s(1-\xi )=v_j^s+\beta _j^s[(1-\xi )-(1-\xi _j^s)]\) be a support function of the graph of \(g(\cdot ,s)\) at point \((1-\xi _j^s,v_j^s)\) for \(j=0,1,\ldots ,J_s\) and \(s=1,\ldots ,n\). Note that here \(v_0^s,v_1^s,\ldots ,v_{J_s}^s\) and \(1-\xi _0^s,1-\xi _1^s,\ldots ,1-\xi _{J_s}^s\) for \(s=1,\ldots ,n\) are both in nonincreasing order. Let \({\tilde{\xi }}_j^s=1-\xi _j^s\) for \(j=0,1,\ldots ,J_s\) and \(s=1,\ldots ,n\). Relabel the sequence \({\tilde{\xi }}_0^s,{\tilde{\xi }}_1^s,\ldots ,{\tilde{\xi }}_{J_s}^s\) in nondecreasing order and denote them by \({\hat{\xi }}_j^s\) for \(j=0,1,\ldots ,J\) and \(s=1,\ldots ,n\). Then \({\hat{\xi }}_j^s={\tilde{\xi }}_{J-j}^s\) for \(j=0,1,\ldots ,J_s\) and \(s=1,\ldots ,n\). Let \({\hat{v}}_j^s:=g({\hat{\xi }}_j^s,s)=g({\tilde{\xi }}_{J-j}^s,s)=g(1-\xi _{J-j}^s,s):=v_{J-j}^s\) for \(j=0,1,\ldots ,J\) and \(s=1,\ldots ,n\). Consequently, we can apply Lemma 4.1 with \(({\hat{\xi }}_j^s, {\hat{v}}_j^s)\) for \(j=0,1,\ldots ,J\) and \(t_{j-1}^s=1-\pi _{j-1}^s\) for\(j=1,\ldots ,L_s\) for each s. By Lemma 4.1 and (4.2), the objective function and constraint \({\mathcal {G}}(v)\cap {\mathcal {G}}_{coh}\ne \emptyset \) in program (A.4) can be reformulated as

$$\begin{aligned} {\mathcal {T}}_Q^{{\mathcal {G}}(v)\cap {\mathcal {G}}_{coh}(Q)}({\varvec{X}})=\sup _{\tau ,\beta }{} & {} \sum _{s=1}^n \sum _{i=1}^{n_s}(x_i^s-x_{i-1}^s)q_s\tau _{i-1}^s\\ \text{ s.t. }{} & {} (4.5d)-(4.5h). \end{aligned}$$

Constraint \({\mathcal {G}}(v)\subset {\mathcal {G}}_{pair}(Q)\) can be represented as (4.5c) and constraint \({\mathcal {G}}(v)\subset {\mathcal {G}}_{ce}(Q)\) by (4.5b). Consequently,

$$\begin{aligned} \varphi (Q)=\sup _{v,\beta ,\tau }{} & {} \sum _{s=1}^n \sum _{i=1}^{n_s}(x_i^s-x_{i-1}^s)q_s\tau _{i-1}^s \\ \text{ s.t. }{} & {} (4.5b)-(4.5h). \end{aligned}$$

Finally, based on the arguments in Step 1, \({\mathcal {T}}_{({\mathcal {G}},{\mathcal {Q}}_{\epsilon })_1}({\varvec{X}})\) is the optimal value of the following program:

$$\begin{aligned} \sup _{v,\beta ,\tau ,q}{} & {} \sum _{s=1}^n \sum _{i=1}^{n_s}(x_i^s-x_{i-1}^s)q_s\tau _{i-1}^s \\ \text{ s.t. }{} & {} (4.5b)-(4.5j). \end{aligned}$$

As for the bi-convexity of program (4.5), the bi-convexity of constraints is directly from Proposition 3.1 and the bi-linear property of the objective function is straightforward. \(\square \)

1.3 Construction of ambiguity set for the special case \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\)

In the case when the ambiguity set of g and the ambiguity set of Q are independent, the construction of joint ambiguity set \(({\mathcal {G}},{\mathcal {Q}}_{\epsilon })_1\) may be simplified. This is primarily because the distortion function at each state can be elicited only based on a DM’s preference information at such state.

For each state \(s\in S\), let \({\mathcal {G}}^s\) denote the ambiguity set of plausible distortion functions of the DM with partially available information at state s. The ambiguity set of state-dependent distortion function can be represented as \({\mathcal {G}}=\times _{s\in S}{\mathcal {G}}^s\). Specifically, we introduce the following classes of distortion functions for the specified state \(s\in S\). For each \(s\in S\), let \(\{G_m^s,B_m^s\}_{m=1}^{M_s}\) be a set of comparable lotteries. The set \({\mathcal {G}}^s_{pair}\) denotes the set of all state-dependent distortion functions at state s, for \(m=1,\ldots ,M_s\), i.e.

$$\begin{aligned} {\mathcal {G}}^s_{pair}:=\{g_s\in {\mathcal {G}}^s:\rho _{g_s}(G_m^s)\le \rho _{g_s}(B_m^s) \;\text {for}\; m=1,\ldots ,M_s \}. \end{aligned}$$

For each \(s\in S\), let \(\{W_k^s\}_{k=1}^{K_s}\) be a list of lotteries with an associated set of “confidence” intervals \([{\underline{w}}_k^s,{\overline{w}}_k^s]\) for the “certainty equivalent” of each \(W_k^s\). The set \({\mathcal {G}}_{ce}^s\) denotes the set of all state dependent distortion functions at state s, which evaluate the risk of each \(W_k^s\) to be larger than \({\underline{w}}_k^s\) and lower than \({\overline{w}}_k^s\), i.e.,

$$\begin{aligned} {\mathcal {G}}_{ce}^s:=\{g_s\in {\mathcal {G}}^s: {\underline{w}}_k^s\le \rho _{g_s}(W_k^s)\le {\overline{w}}_k^s ,\;\text{ for }\;k=1,\ldots ,K_s\}. \end{aligned}$$

Let \({\mathcal {G}}_{coh}^s\) be the set of all concave state-dependent distortion function at the specified state \(s\in S\). Then, we consider the following subset of the general ambiguity set \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\):

  • \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }:= \left[ \times _{s\in S}({\mathcal {G}}_{pair}^s\cap {\mathcal {G}}_{ce}^s\cap {\mathcal {G}}_{coh}^s)\right] \times {\mathcal {Q}}_{\epsilon }\).

The following example illustrates the fundamental difference between the general ambiguity set \(({\mathcal {G}}_1,{\mathcal {Q}}_{\epsilon })\) and the special ambiguity set \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\) by comparing the difference between \(({\mathcal {G}}_{pair},Q)\) and \((\times _{s\in S}{\mathcal {G}}_{pair}^s)\times Q\) for \(Q\in {\mathcal {Q}}_{\epsilon }\) based on Example 3.1.

Example A.2

We now reconsider Example 3.1 with the states of nature \(S=\{s_1,s_2\}\) and the number of the pairs of pairwise comparisons is \(M=2\). A DM is asked to compare the following two sets of state-dependent acts:

$$\begin{aligned} {\varvec{G}}_1= & {} {\left\{ \begin{array}{ll} G_1^1=[1,\frac{1}{4}; 10,\frac{3}{4}] &{} \text{ for } \; s=s_1,\\ G_1^2=[2,\frac{1}{2};\;\;8,\frac{1}{4}] &{} \text{ for } \; s=s_2, \end{array}\right. }\\ \quad \text {and}\quad {\varvec{B}}_1= & {} {\left\{ \begin{array}{ll} B_1^1=[2,\frac{1}{5};12,\frac{3}{5}] &{} \text{ for } \; s=s_1,\\ B_1^2=[2,\frac{1}{5};10,\frac{3}{5}] &{} \text{ for } \; s=s_2, \end{array}\right. } \end{aligned}$$

and

$$\begin{aligned} {\varvec{G}}_2= & {} {\left\{ \begin{array}{ll} G_2^1=[1,\frac{1}{5}; 10,\frac{4}{5}] &{} \text{ for } \; s=s_1,\\ G_2^2=[3,\frac{1}{2};\;\;9,\frac{1}{4}] &{} \text{ for } \; s=s_2, \end{array}\right. }\\ \quad \text {and}\quad {\varvec{B}}_2= & {} {\left\{ \begin{array}{ll} B_2^1=[2,\frac{1}{4};12,\frac{1}{2}] &{} \text{ for } \; s=s_1,\\ B_2^2=[2,\frac{1}{5};10,\frac{3}{5}] &{} \text{ for } \; s=s_2. \end{array}\right. } \end{aligned}$$

Assume that the specified probability distribution of the DM’s belief of the states of nature \(\{s_1,s_2\}\) is \(Q=(\frac{1}{4},\frac{3}{4})\). In the following, we consider the differences between the two cases of pairwise comparison discussed in the preceding section: one is to compare the state-dependent acts \({\varvec{G}}_m\) and \({\varvec{B}}_m\) for \(m=1,2\), the other is to compare the consequences \(G_m^s\) and \(B_m^s\) at each state \(s=1,2\) for \(m=1,2\).

Case 1. Assume that the DM prefers \({\varvec{G}}_m\) to \({\varvec{B}}_m\) for \(m=1,2\), i.e. \({\mathcal {T}}_Q^g({\varvec{G}}_m)\le {\mathcal {T}}_Q^g({\varvec{B}}_m),\;\text{ for }\; m=1,\ldots ,2\). Thus, the general ambiguity set

$$\begin{aligned}{} & {} ({\mathcal {G}}_{pair},Q) \\{} & {} \quad =\left\{ (g,Q): \begin{aligned}&g=(g_1,g_2)\;\text {such that}\; g_1(\cdot )\;\text {and}\;g_2(\cdot )\;\text {are distortion functions}\;\text {and}\;Q=(\tfrac{1}{4},\tfrac{3}{4})\;\\&\text {with}\;\tfrac{1}{4}g_1(1)+\tfrac{9}{4}g_1(\tfrac{3}{4})+\tfrac{3}{2}g_2(\tfrac{3}{4})+\tfrac{9}{2}g_2(\tfrac{1}{4})\le \tfrac{1}{2}g_1(\tfrac{4}{5})+\tfrac{5}{2}g_1(\tfrac{3}{5})+\tfrac{3}{2}g_2(\tfrac{4}{5})+6g_2(\tfrac{3}{5}),\\&\text {and}\;\;\tfrac{1}{4}g_1(1)+\tfrac{9}{4}g_1(\tfrac{4}{5}) +\tfrac{9}{4}g_2(\tfrac{3}{4})+\tfrac{9}{2}g_2(\tfrac{1}{4}) \le 2g_1(\tfrac{3}{4})+\tfrac{5}{2}g_1(\tfrac{1}{2})+\tfrac{3}{2}g_2(\tfrac{4}{5})+6g_2(\tfrac{3}{5}). \end{aligned} \right\} . \end{aligned}$$

Case 2. Assume that the DM prefers \(G_m^s\) to \(B_m^s\) at each state for \(m=1,2\), i.e. \(\rho _{g_s}(G_m^s)\le \rho _{g_s}(B_m^s)\) at \(s=1,2\) for \(m=1,2\). Thus, the special ambiguity set

$$\begin{aligned}{} & {} {\mathcal {G}}_{pair}\times Q \\{} & {} \quad =\left\{ g\times Q: \begin{aligned}&g=(g_1,g_2)\;\text {such that}\; g_1(\cdot )\;\text {and}\;g_2(\cdot )\;\text {are distortion functions}\;\text {and}\;Q=(\tfrac{1}{4},\tfrac{3}{4})\\&\text {with}\; g_1(1)+9g_1(\tfrac{3}{4})\le 2g_1(\tfrac{4}{5})+10g_1(\tfrac{3}{5}),\; 2g_2(\tfrac{3}{4})+6g_2(\tfrac{1}{4})\le 2g_2(\tfrac{4}{5})+8g_2(\tfrac{3}{5}),\; \\&\text {and}\;\;g_1(1)+9g_1(\tfrac{4}{5})\le 2g_1(\tfrac{3}{4})+10g_1(\tfrac{1}{2}), \; 3g_2(\tfrac{3}{4})+6g_2(\tfrac{1}{4})\le 2g_2(\tfrac{4}{5})+8g_2(\tfrac{3}{5}). \end{aligned} \right\} . \end{aligned}$$

It is easy to observe that \({\mathcal {G}}_{pair}\times Q\subset ({\mathcal {G}}_{pair},Q)\).

Before ending this subsection, we give some properties of the defined ambiguity sets \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\), which will be beneficial for us to analyze the properties of the optimization programs of the tractable reformulation for the proposed PRSDRM based on \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\).

Proposition A.1

\({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\) is a convex set.

Proof

Given \((g_i,Q_i)\in {\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\) with \(Q_i=(q_1^i,q_2^i,\ldots ,q_n^i)\) for \(i=1,2\), let

$$\begin{aligned} (g_{\lambda },Q_{\lambda }):=\lambda (g_1,Q_1)+(1-\lambda )(g_2,Q_2)=(\lambda g_1+(1-\lambda )g_2, \lambda Q_1 +(1-\lambda ) Q_2) \end{aligned}$$

for an arbitrarily chosen \(\lambda \in [0,1]\). It is easy to see that \(g_{\lambda }\) is a state-dependent distortion function and \(Q_{\lambda }\) is a probability distribution on S. Moreover, since \(\phi \) is convex, then

$$\begin{aligned} \sum _{i=1}^n \phi \left( \frac{\lambda q_i^1+(1-\lambda )q_i^2}{q_i^0}\right) \le \lambda \sum _{i=1}^n \phi \left( \frac{q_i^1}{q_i^0}\right) +(1-\lambda ) \sum _{i=1}^n \phi \left( \frac{q_i^2}{q_i^0}\right) \le \epsilon . \end{aligned}$$

Consequently, \(Q_{\lambda }\in {\mathcal {Q}}_{\epsilon }\). In the following, we show that \(g_{\lambda }\in {\mathcal {G}}_1\). Since \(g_i\in {\mathcal {G}}_1\), then by definition, we have that for each \(s\in S\),

$$\begin{aligned} (g_i)_s\in {\mathcal {G}}_{pair}^s,\;(g_i)_s\in {\mathcal {G}}_{ce}^s,\;\text {and}\;(g_i)_s\in {\mathcal {G}}_{coh}^s. \end{aligned}$$

By the definition of DRM, we have that for any lottery X and \(s\in S\)

$$\begin{aligned} \rho _{(g_{\lambda })_s}(X)=\rho _{\lambda (g_1)_s +(1-\lambda ) (g_2)_s}(X)=\lambda \rho _{(g_1)_s}(X)+(1-\lambda )\rho _{(g_2)_s}(X). \end{aligned}$$

Consequently, \((g_{\lambda })_s\in {\mathcal {G}}_{pair}^s\) and \((g_{\lambda })_s\in {\mathcal {G}}_{ce}^s\) for \(s\in S\). Moreover, since \((g_i)_s\) is a concave function on [0, 1] for \(i=1,2\), then \((g_{\lambda })_s\) is also a concave function on [0, 1], i.e. \((g_{\lambda })_s\in {\mathcal {G}}_{coh}^s\). To summarize, we have showed that for each \(s\in S\), \((g_{\lambda })_s\in {\mathcal {G}}_{pair}^s\cap {\mathcal {G}}_{ce}^s\cap {\mathcal {G}}_{coh}^s\), i.e. \(g_{\lambda }\in \times _{s\in S} \left( {\mathcal {G}}_{pair}^s \cap {\mathcal {G}}_{ce}^s\cap {\mathcal {G}}_{coh}^s\right) ={\mathcal {G}}_1\). Finally, by the independence of the ambiguity sets \({\mathcal {G}}_1\) and \({\mathcal {Q}}_{\epsilon }\), we conclude that \((g_{\lambda },Q_{\lambda })\in {\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\). \(\square \)

1.4 Tractable reformulations for the special case

The tractable reformulation of PRSDRM under the special construction of the ambiguity set \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\) can also be simplified. Here we give the details.

1.4.1 Tractable reformulation of PRSDRM under \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\)

In this subsection, we consider tractable reformulation of PRSDRM when the ambiguity set of state dependent distortion functions is constructed as follows:

$$\begin{aligned} {\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }= & {} ({\mathcal {G}}_{pair}\cap {\mathcal {G}}_{ce}\cap {\mathcal {G}}_{coh})\times {\mathcal {Q}}_{\epsilon } = \left[ \times _{s\in S}({\mathcal {G}}_{pair}^s\cap {\mathcal {G}}_{ce}^s\cap {\mathcal {G}}_{coh}^s)\right] \nonumber \\{} & {} \times {\mathcal {Q}}_{\epsilon }=:\left[ \times _{s\in S}{\mathcal {G}}_1^s\right] \times {\mathcal {Q}}_{\epsilon }, \end{aligned}$$
(A.9)

where \({\mathcal {G}}_1^s={\mathcal {G}}_{pair}^s\cap {\mathcal {G}}_{ce}^s\cap {\mathcal {G}}_{cave}^s\) for \(s\in S\). Then

$$\begin{aligned} {\mathcal {T}}_{{\mathcal {Q}}_{\epsilon }}^{{\mathcal {G}}_1}({\varvec{X}}): =\sup _{(g,Q)\in {\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }} {\mathbb {E}}_{Q}[\rho _{g_s}(X_s)]. \end{aligned}$$
(A.10)

The program (A.10) is a semi-infinite optimization program and hard to solve in general. However, the following proposition says that for such construction of the ambiguity set \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\), the PRSDRM for a state-dependent act can be calculated by a two-stage optimization program.

Proposition A.2

For each \(s\in S\), let \(\psi _1(s)=\sup _{g_s\in {\mathcal {G}}_1^s}\rho _{g_s}(X_s)\). Then \({\mathcal {T}}_{{\mathcal {Q}}_{\epsilon }}^{{\mathcal {G}}_1}({\varvec{X}})=\sup _{Q\in {\mathcal {Q}}_{\epsilon }}{\mathbb {E}}_Q[\psi _1(s)]\).

Proof

Since the ambiguity sets \({\mathcal {G}}_1\) and \({\mathcal {Q}}_{\epsilon }\) are independent, then the PRSDRM of a state-dependent act \({\varvec{X}}\) under \({\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }\) is equivalent to the optimal value of the following optimization problem:

$$\begin{aligned} {\mathcal {T}}_{{\mathcal {Q}}_{\epsilon }}^{{\mathcal {G}}_1}({\varvec{X}}) =\sup _{(g,Q)\in {\mathcal {G}}_1\times {\mathcal {Q}}_{\epsilon }} {\mathbb {E}}_{Q}[\rho _{g_s}(X_s)] =\sup _{Q\in {\mathcal {Q}}_{\epsilon }}\sup _{g\in {\mathcal {G}}_1}{\mathbb {E}}_{Q}[\rho _{g_s}(X_s)]. \end{aligned}$$

Moreover, since \({\mathcal {G}}_1=\times _{s\in S}{\mathcal {G}}_1^s\) and for each \(s\in S\) the ambiguity set \({\mathcal {G}}_1^s\) only depends on the information of state s, then

$$\begin{aligned} \sup _{g\in {\mathcal {G}}_1}{\mathbb {E}}_{Q}[\rho _{g_s}(X_s)]={\mathbb {E}}_Q\left[ \sup _{g\in {\mathcal {G}}_1^s}\rho _{g_s}(X_s)\right] . \end{aligned}$$

This completes the proof. \(\square \)

Note that for each \(s\in S\), \(\sup _{g_s\in {\mathcal {G}}_1^s}\rho _{g_s}(X_s)\) is also a semi-infinite program in general. However, when \(X_s\) is discretely distributed, it can be computed by solving a finite dimensional linear program, see, e.g. (Wang and Xu 2023, Theorem 4.1). Once \(\psi _1(s)\) is determined for each \(s\in S\), the program \(\sup _{Q\in {\mathcal {Q}}_{\epsilon }}{\mathbb {E}}_Q[\psi _1(s)]\) would be a convex program and can be easily solved by convex optimization solver, see, e.g. CVX in Matlab.

In the followup discussions, for each \(s\in S\), we will use \(\Xi ^s\) to denote the set of all breakpoints of the quantile functions of \(W_k^s\), \(G_m^s\) and \(B_m^s\) for \(k=1,\ldots ,K_s\) and \(m=1,\ldots ,M_s\) and label them in the increasing order of the values, i.e. we will use \(\xi _j^s\) to denote the j-th smallest element of set \(\Xi ^s\) and let \(\xi _0^s=0\). For clarity of exposition, scenarios in \(\Omega \) will be indexed by i and elements in \(\Xi ^s\) will be indexed by j. Thus, the size of the optimization problem \(\sup _{g_s\in {\mathcal {G}}_1^s}\rho _{g_s}(X_s)\) will be determined by the number of pairwise comparisons \(M_s\), the number of certainty equivalent lotteries \(K_s\), the size of scenarios \(|\Omega |=n_s\), and the size of the total breakpoints \(J_s=|\Xi ^s|\). The next proposition showed by Wang and Xu (2023) states that \(\sup _{g_s\in {\mathcal {G}}_1^s}\rho _{g_s}(X_s)\) can be computed by solving a finite dimensional linear program of reasonable size when \(X_s\) is finite discretely distributed (without loss of generality, we may assume that \(X_s\) has \(n_s\) different realizations) as it involves \(2J_s+n_s+3\) variables and \(K_s+M_s+n_sJ_s+n_s+3J_s+5\) constraints at most (not counting the non-negative constraints).

Proposition A.3

Let \(X_s\) be a finite discretely distributed non-negative random loss with \({\mathbb {P}}(X_s=x_i^s)=p_i^s\) where \(x_1^s<x_2^s<\cdots <x_{n_s}^s\). Then \(\psi _1(s)\) is the optimal value of the following linear program:

$$\begin{array}{llr}\sup \limits_{v,\beta ,\tau }{} & {} \sum \limits_{i=1}^{n_s} (x_i^s-x_{i-1}^s) \tau _{i-1} & ({\text{A}}.11{\text{a}})\\ \\ \text{ s.t. }{} & {} {\underline{w}}_k^s\le \sum \limits_{j=1}^{J_s}[F_{W_k}^{\leftarrow }(\xi _j^s)-F_{W_k}^{\leftarrow }(\xi _{j-1}^s)] v_{j-1} \le {\overline{w}}_k^s,\;\text{ for }\; k=1,\ldots ,K_s, & ({\text{A}}.11{\text{b}})\\ \\{} & {} \sum \limits_{j=1}^{J_s}[F_{G_m^s}^{\leftarrow }(\xi _j^s)-F_{G_m^s}^{\leftarrow }(\xi _{j-1}^s)] v_{j-1}\le \sum \limits_{j=1}^{J_s}[F_{B_m^s}^{\leftarrow }(\xi _j^s)\nonumber \\{} & {} \quad \quad -F_{B_m^s}^{\leftarrow }(\xi _{j-1}^s)]v_{j-1},\; m=1,\ldots ,M_s, & ({\text{A}}.11{\text{c}})\\ \\{} & {} v_j+\beta _j(\xi _j^s-\pi _i)\ge \tau _i,\;\text{ for }\; i=0,1,\ldots ,n_s;j=0,1,\ldots ,J_s, & ({\text{A}}.11{\text{d}}) \\ \\{} & {} v_j+\beta _j(\xi _j^s-\xi _{j+1}^s)\ge v_{j+1},\;\text{ for }\; j=0,1,\ldots ,J_s-1, & ({\text{A}}.11{\text{e}}) \\ \\{} & {} v_j+\beta _{j+1}(\xi _j^s-\xi _{j+1}^s)\le v_{j+1},\;\text{ for }\; j=0,1,\ldots ,J_s-1, & ({\text{A}}.11{\text{f}}) \\ \\{} & {} \beta _j\ge 0,\;\text{ for }\; j=0,1,\ldots ,J_s, & ({\text{A}}.11{\text{g}}) \\ \\{} & {} v_0=1,v_{J_s}=0,\tau _0=1,\tau _{n_s}=0, \\ \end{array}$$

where \(\pi _0=0\) and \(\pi _i=\sum _{l\le i}p_l^s\) for \(i=1,\ldots ,n_s\).

Proof

The proof is similar to that of Wang and Xu (2023, Theorem 4.1), we skip the details. \(\square \)

To summarize the above discussion, we have the following theorem:

Theorem A.1

Let \({\varvec{X}}\) be a state-dependent act with consequences \(X_s\) being discretely distributed non-negative random loss with \({\mathbb {P}}(X_s=x_i^s)=p_i^s\) where \(x_1^s<x_2^s<\cdots <x_{n_s}^s\) for \(s\in S\). Then \({\mathcal {T}}_{{\mathcal {Q}}_{\epsilon }}^{{\mathcal {G}}_1}({\varvec{X}})\) is the optimal value of the following convex program:

$$\begin{array}{llr}\sup \limits_{q}{} & {} \sum \limits_{i=1}^{n} q_i \psi _1(s_i) & ({\text{A}}.12{\text{a}})\\ \\ \mathrm{{s.t.}}{} & {} \sum \limits_{i=1}^n q_i=1, \sum \limits_{i=1}^n \phi \left( \frac{q_i}{q_i^0}\right) q_i^0\le \epsilon ,\;\text {and}\;q_i\ge 0,\;\text {for}\;i=1,\ldots ,n. & ({\text{A}}.12{\text{b}}) \end{array}$$

1.5 Proof of Proposition 5.1

Proof

Let \((g,Q)\in ({\mathcal {G}},{\mathcal {Q}}_{\epsilon })_1\) be fixed and \(\tau _i^s=g(1-\frac{i}{n_s},s)\) for \(i=0,1,\ldots ,n_s\) and \(s=1,\ldots ,n\). To ease the notation, we set \(\tau _{-1}^s:=\tau _0^s\) for \(s=1,\ldots ,n\). Based on (5.2), we define

$$\begin{aligned} v(z,\tau , q)= \sum _{s=1}^n\sum _{i=1}^{n_s} q_s \gamma _i^s\textrm{CVaR}_{\beta _i^s}(f_s(z,\xi _s)), \end{aligned}$$

where \(\beta _i^s=\frac{i-1}{n_s}\), \(\gamma _i^s=(\phi _i^s-\phi _{i-1}^s)(n_s-i+1)\), \(\phi _i^s=\tau _{i-1}^s-\tau _i^s\) for \(i=1,\ldots ,n_s\) and \(s=1,\ldots ,n\). In other words, we have

$$\begin{aligned} \gamma _i^s=(2\tau _{i-1}^s-\tau _{i}^s-\tau _{i-2}^s)(n_s-i+1),\;\text {for}\; i=1,\ldots ,n_s \; \text {and}\; s=1,\ldots ,n. \end{aligned}$$

Since \(f_s(z,\xi _s)\) is a convex function in z for every fixed \(\xi _s\) and the operator \(\textrm{CVaR}_{\beta _i^s}(\cdot )\) is non-decreasing and convex, then \(\textrm{CVaR}_{\beta _i^s}(f_s(z,\xi _s))\) is a convex function of z. Moreover, since \(g\in {\mathcal {G}}_{coh}\), then \(\gamma _i^s\ge 0\) for \(i=1,\ldots ,n_s\) and \(s=1,\ldots ,n\). Thus, \(v(z,\tau ,q)\) is a convex function in z for every fixed \(\tau \) ( depending on g) and q. On the other hand, for fixed \(z\in Z\), \(v(z,\tau , q)\) is linear in both \(\tau \) and q. Unfortunately, v is not linear in \((\tau ,q)\) jointly which prevents us from applying the existing minimax saddle point results in the proof. This motivates us to do some transformation of the variables. Specifically, we set \({\tilde{\tau }}_i^s=q_s\tau _i^s\) and \(\tilde{\gamma }_i^s=(2{\tilde{\tau }}_{i-1}^s-{\tilde{\tau }}_{i}^s-{\tilde{\tau }}_{i-2}^s)(n_s-i+1)\) for \(i=1,\ldots ,n_s\; \text {and}\; s=1,\ldots ,n\). Consequently, we can write \(v(z,\tau , q)\) as

$$\begin{aligned} {\tilde{v}}(z,{\tilde{\tau }})=\sum _{s=1}^n\sum _{i=1}^{n_s} {\tilde{\gamma }}_i^s\textrm{CVaR}_{\beta _i^s}(f_s(z,\xi _s)). \end{aligned}$$

Note that \({\tilde{v}}(z,{\tilde{\tau }})\) is a convex function in z for every fixed \({\tilde{\tau }}\) ( depending on g and q). On the other hand, for fixed \(z\in Z\), \({\tilde{v}}(z,{\tilde{\tau }})\) is linear in \({\tilde{\tau }}\). By Remark 4.1, we know that the joint ambiguity set \(({\mathcal {G}},{\mathcal {Q}}_{\epsilon })_1\) can be represented by the constraints in Problem (4.7). Let \(\Upsilon \) denote the feasible set of the problem and \(\Upsilon _{{\tilde{\tau }}}\) the projection of the set \(\Upsilon \) on the space of \({\tilde{\tau }}\) variable. Since \(\Upsilon \) is convex and compact, then \(\Upsilon _{{\tilde{\tau }}}\) is also convex and compact. Thus, we can recast problem (5.1) as

$$\begin{aligned} \min _{z\in Z}\max _{{\tilde{\tau }} \in \Upsilon _{{\tilde{\tau }}}} {\tilde{v}}(z,{\tilde{\tau }}). \end{aligned}$$
(A.13)

Note that set \(\Upsilon _{{\tilde{\tau }}}\) is independent of z. By Fan (1953, Theorem 1 (ii)),

$$\begin{aligned} \min _{z\in Z} \max _{{\tilde{\tau }}\in \Upsilon _{\tau }} {\tilde{v}}(z,{\tilde{\tau }}) = \max _{{\tilde{\tau }}\in \Upsilon _{\tau }} \min _{z\in Z} {\tilde{v}}(z,{\tilde{\tau }}), \end{aligned}$$
(A.14)

which, by Karlin (1959, Corollary 1.3.1), is sufficient and necessary for the existence of a saddle point. Let \((z^*,{\tilde{\tau }}^*)\) denote the saddle point. Then

$$\begin{aligned} {\tilde{v}}(z^*,{\tilde{\tau }}) \le {\tilde{v}}(z^*,{\tilde{\tau }}^*) \le {\tilde{v}}(z,{\tilde{\tau }}^*). \end{aligned}$$
(A.15)

We are now ready to show the convergence of the sequence \(\{(z^j,{\tilde{\tau }}^j)\}\) generated by Algorithm 5.1. For \(j=1,2,\ldots \), it follows from Step 2 of Algorithm 5.1,

$$\begin{aligned} {\tilde{v}}(z^{j-1},{\tilde{\tau }}^j)\ge {\tilde{v}}(z^{j-1},{\tilde{\tau }}),\;\text {for all}\; {\tilde{\tau }}\in \Upsilon _{{\tilde{\tau }}}. \end{aligned}$$
(A.16)

Likewise, it follows from Step 3 of the algorithm

$$\begin{aligned} {\tilde{v}}(z^j,{\tilde{\tau }}^j) \le {\tilde{v}}(z,{\tilde{\tau }}^j),\;\text {for all}\; z\in Z. \end{aligned}$$
(A.17)

In the case when the algorithm terminates in finite steps, we have \(z^{j-1}=z^j\) and \({\tilde{\tau }}^{j-1}={\tilde{\tau }}^j\) for some j and consequently \((z^j,{\tilde{\tau }}^j)\) satisfies (A.15).

Next, we consider the case that the algorithm generates an infinite sequence \(\{(z^j,{\tilde{\tau }}^j)\}\). Let \(({\hat{z}},\hat{{\tilde{\tau }}})\) be a cluster point of \(\{(z^j,{\tilde{\tau }}^j)\}\). By taking a subsequence if necessary, we assume for the simplicity of notation that \((z^j,{\tilde{\tau }}^j)\rightarrow ({\hat{z}},\hat{{\tilde{\tau }}})\) as \(j\rightarrow \infty \). Assume for the sake of a contradiction that \(({\hat{z}},\hat{{\tilde{\tau }}})\) is not a solution to the problem (A.13). Then \(({\hat{z}},\hat{{\tilde{\tau }}})\) would violate one of the inequalities in (A.15). Consider the case that the second inequality of (A.15) is violated. Then there exists \(z_0\in Z\) such that

$$\begin{aligned} {\tilde{v}}({\hat{z}},\hat{{\tilde{\tau }}}) > {\tilde{v}}(z_0,\hat{{\tilde{\tau }}}). \end{aligned}$$

Since v is continuous, the inequality means that for j sufficiently large,

$$\begin{aligned} {\tilde{v}}(z^j,{\tilde{\tau }}^j)> {\tilde{v}}(z_0,{\tilde{\tau }}^j), \end{aligned}$$

which contradicts (A.17). Likewise, we can show that \(({\hat{z}},\hat{{\tilde{\tau }}})\) satisfies the first inequality in (A.15). This completes the proof. \(\square \)

Appendix B Data generation

2.1 Sample data generation

In this part, we consider how to generate data set of the random loss of \(m=3\) cities in \(n=2\) states, i.e. to generate the random matrix \(\varvec{\xi }=(\xi _1,\xi _2)\in {\mathrm{I\!R}}^{3\times 2}\). Based on Zhang et al. (2020), we also assume that there are three possible underlying loss scenarios corresponding to different levels of terrorist attacks: reduced loss, standard loss and increased loss, which means the size of sample space is \(|\Omega |=3\). Moreover, we assume that the random loss for city j at state s is generated from a given distribution. Specifically, we assume that the random loss \(\xi _j^s\) is decomposable into a state-based risk factor \(\vartheta _s \sim U(0,100\times (2+s))\) common to all cities for \(s=1,2\) and a city-based risk factor \(\zeta _j \sim U(100\times (1+j\times 3\%), 200\times (1+j\times 5\%))\) specific to city j for \(j=1,2,3\), where U means the uniform distribution over a specified range. Thus, we set

$$\begin{aligned} \xi _j^s=\vartheta _s+\zeta _j, \;\text {for}\; j=1,2,3\;\text {and}\;s=1,2. \end{aligned}$$

Note that for each s and j, after generating three losses from \(\xi _j^s\), we may sort them in non-decreasing order and assign the smallest loss to the scenario reduced loss, the second smallest loss to the scenario standard loss, and the largest loss to the scenario increased loss. Moreover, we also assume that the probability to each scenario is 1/3. Table 3 presents the generated sample dataset.

Table 3 The random loss \(\xi _j^s(\omega )\) for \(j=1,2,3\), \(s=1,2\) and \(|\Omega |=3\)

2.2 Elicited data generation

Potential questionnaires for eliciting a DM’s risk preference are given as follows.

  1. Q1.

    Which risky state-dependent act in the m-th pair of acts does the DM prefer for \(m=1,\ldots ,M\)?

    The preferred act is denoted by \({\varvec{G}}_m\) and the other is denoted by \({\varvec{B}}_m\). Note that when generating such acts, we require that the difference between the SDRMs of each pair of acts is greater than the specified threshold denoted by \(\kappa \) in order to make sure that the compared acts having significant differences and the DM can easily choose which one he prefers to. In our setting, we set \(\kappa =1\). The decision is made based on the values of SDRMs for each acts induced by the true state-dependent distortion function \(g_1^*\) and the true probability distribution of the belief of the states of nature, and the smaller is preferred.

  2. Q2.

    What is the smallest amount of cash, denoted by \(w_k^+\), that the DM would decline to pay instead of being exposed to the risk of \({\varvec{W}}_k\) and what is the largest amount of cash, denoted by \(w_k^-\), that the DM would be willing to commit instead of being exposed to the risk of \({\varvec{W}}_k\) for \(k=1,\ldots ,K\)? In this kind of questionnaire, we assume that \(w_k^+\) and \(w_k^-\) are determined based on the values of SDRMs for each acts induced by the true distortion function \(g_1^*\) and the true probability distribution of the belief of the states of nature as follows:

    $$\begin{aligned} w_k^+=(1+r){\mathcal {T}}_{Q^*}^{g_1^*}({\varvec{W}}_k),\quad w_k^-=(1-r){\mathcal {T}}_{Q^*}^{g_1^*}({\varvec{W}}_k), \end{aligned}$$

    where r is randomly generated from the uniform distribution on \([0,1\%]\).

  3. Q3.

    Experts reach a consensus that the probability distribution of the states of nature is \(Q^*\), does the DM agree with such judgement to some extent?

    Based on the DM’s answers, we may set the level of \(\epsilon \) in \({\mathcal {Q}}_{\epsilon }\) based on the DM’s confidence level to experts’ judgement. For example, if the DM fully agrees with such a judgement, then \(\epsilon =0\); if the DM has \(\alpha \%\) confidence to agree with the experts’ judgement, then we may set \(\epsilon =1-\alpha \%\), for example, we may choose \(\alpha \in \{99,97.5,95\}\), i.e. \(\epsilon \in \{0.01,0.025,0.05\}\).

In order to make the elicited data comparable with the random loss of \(\xi _j^s\), we generate the loss \(x_1,x_2\) from the uniform distribution over

$$\begin{aligned} {}[r_1\times \min _{s,j,\omega }\xi _j^s(\omega ), r_2\times \max _{s,j,\omega }\xi _j^s(\omega )] \end{aligned}$$

where \(r_1<1<r_2\). In our experiment, we set \(r_1=0.5\) and \(r_2=1.2\) and then the range to generate the realization of \(x_1\) and \(x_2\) is [188.87, 695.95]. Moreover, \(p_1,p_2\) are also randomly generated from the uniform distribution on (0, 1). We give an example of the generated data for preference elicitation in Table 4.

Table 4 An example of generated data for preference elicitation

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, W., Xu, H. Preference robust state-dependent distortion risk measure on act space and its application in optimal decision making. Comput Manag Sci 20, 45 (2023). https://doi.org/10.1007/s10287-023-00475-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10287-023-00475-x

Keywords

Navigation