Log in

A Neyman–Pearson problem with ambiguity and nonlinear pricing

  • Published:
Mathematics and Financial Economics Aims and scope Submit manuscript

Abstract

We consider a problem of the Neyman–Pearson type arising in the theory of portfolio choice in the presence of probability weighting, such as in markets with Choquet pricing (as in Araujo et al. in Econ Theory 49(1):1–35, 2011; Cerreia-Vioglio et al. in J Econ Theory 157(1):730–762, 2015; Chateauneuf and Cornet in Submodular financial markets with frictions. Working Paper, 2015; Chateauneuf et al. in Math Finance 6(3):323–330, 1996) and ambiguous beliefs about the payoffs of contingent claims (see Gilboa and Marinacci, in: Acemoglu, Arellano, Dekel (eds) Advances in economics and econometrics: theory and applications, tenth world congress of the econometric society, Cambridge University Press, Cambridge, 2013). Specifically, we consider a problem of optimal choice of a contingent claim so as to minimize a non-linear pricing functional (or a distortion risk measure), subject to a minimum expected performance measure (or a minimum expected return or utility), where expectations with respect to distorted probabilities are taken in the sense of Choquet. Such contingent claims are called cost-efficient. We give an analytical characterization of cost-efficient contingent claims under very mild assumptions on the probability weighting functions, thereby extending some of the results of Ghossoub (Math Financ Econ 10(1):87–111, 2016), and we provide examples of some special cases of interest. In particular, we show how a cost-efficient contingent claim exhibits a desirable monotonicity property: It is anti-comonotonic with the random mark-to-market value (or return, etc.) of the underlying financial position, and it is hence a hedge against such variability.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Canada)

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. A map** \(\rho : B^{+}\left( \Sigma \right) \rightarrow \mathbb {R}\) is said to preserve uniformly bounded pointwise convergence if for any \(Y^{*} \in B^{+}\left( \Sigma \right) \) and for any sequence \(\{Y_{n}\}_{n \geqslant 1} \in B^{+}\left( \Sigma \right) \), one has \(\underset{n \rightarrow + \infty }{\lim }\rho \left( Y_{n}\right) = \rho \left( Y^{*}\right) \) whenever (i) \(\underset{n \rightarrow +\infty }{\lim }Y_{n} = Y^{*}\) (pointwise); and, (ii) there exists some \(N \in \left( 0,+\infty \right) \) such that \(Y_{n} \leqslant N\), for each \(n \geqslant 1\). For instance, the Lebesgue integral preserves uniformly bounded pointwise convergence [10, Th. 2.4.4], as well as the Choquet integral with respect to any continuous capacity [33, Th. 7.16].

  2. See Gilboa and Marinacci [21] for an overview of models of decision under ambiguity.

  3. A finite nonnegative measure \(\eta \) on a measurable space \(\left( \Omega , \mathcal {A}\right) \) is said to be nonatomic if for any \(A \in \mathcal {A}\) with \(\eta \left( A\right) > 0\), there is some \(B \in \mathcal {A}\) such that \(B \subsetneq A\) and \(0< \eta \left( B\right) < \eta \left( A\right) \).

  4. This assumption can be dropped, but one would have to use the Distributional Transform approach of Rüschendorf [35]. All the results of this paper would still hold, with adequate modifications.

  5. Any \(Y \in B\left( \Sigma \right) \) is bounded, and we define its supnorm by \(\Vert Y\Vert _{sup} := \sup \{ |Y\left( s\right) |: s \in S \} < +\infty \).

  6. Concavity of the performance function \(\phi \) reflects the idea that diversification is beneficial.

  7. This assumption is purely for mathematical convenience, and it could be relaxed.

  8. See, for instance, He et al. [24, Appendix B].

  9. That is, \(P \circ \phi _{1}^{-1}\left( B\right) = P \circ \phi _{2}^{-1}\left( B\right) \), for any Borel set B.

  10. Any \(q \in \mathcal {Q}^{**}\) is differentiable a.s., being monotone [10, Th. 6.3.3].

  11. Following **a and Zhou [39], we write \(\phi \circ q\left( t\right) = \int _{0}^{t} d\phi \circ q\left( x\right) \) and apply Fubini’s Theorem.

References

  1. Aliprantis, C.D., Border, K.C.: Infinite Dimensional Analysis, 3rd edn. Springer, Berlin (2006)

    MATH  Google Scholar 

  2. Araujo, A., Chateauneuf, A., Faro, J.H.: Pricing rules and Arrow–Debreu ambiguous valuation. Econ. Theory 49(1), 1–35 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  3. Campi, L., Jouini, E., Porte, V.: Efficient portfolios in financial markets with proportional transaction costs. Math. Finance Econ. 7(3), 281–304 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  4. Carlier, G., Dana, R.A.: Core of convex distortions of a probability. J. Econ. Theory 113(2), 199–222 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  5. Carlier, G., Dana, R.A.: Two-persons efficient risk-sharing and equilibria for concave law-invariant utilities. Econ. Theory 36(2), 189–223 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cerreia-Vioglio, S., Maccheroni, F., Marinacci, M.: Put–Call parity and market frictions. J. Econ. Theory 157(1), 730–762 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  7. Chateauneuf, A., Cornet, B.: Submodular Financial Markets with Frictions. Working Paper (2015)

  8. Chateauneuf, A., Kast, R., Lapied, A.: Choquet pricing for financial markets with frictions. Math. Finance 6(3), 323–330 (1996)

    Article  MATH  Google Scholar 

  9. Choquet, G.: Theory of capacities. Ann. Inst. Fourier 5, 131–295 (1954)

    Article  MathSciNet  MATH  Google Scholar 

  10. Cohn, D.L.: Measure Theory. Birkhauser, Boston (1980)

    Book  MATH  Google Scholar 

  11. Dana, R.A.: Market behavior when preferences are generated by second-order stochastic dominance. J. Math. Econ. 40(6), 619–639 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  12. Dana, R.A.: A representation result for concave Schur concave functions. Math. Finance 15(4), 613–634 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  13. Dana, R.A., Meilijson, I.: Modelling Agents’ Preferences in Complete Markets by Second Order Stochastic Dominance. Mimeo, New York (2003)

    Google Scholar 

  14. Denneberg, D.: Non-Additive Measure and Integral. Kluwer Academic Publishers, Dordrecht (1994)

    Book  MATH  Google Scholar 

  15. Dunford, N., Schwartz, J.T.: Linear Operators, Part 1: General Theory. Wiley, Hoboken (1958)

    MATH  Google Scholar 

  16. Dybvig, P.H.: Distributional analysis of portfolio choice. J. Bus. 61(3), 369–393 (1988)

    Article  Google Scholar 

  17. Dybvig, P.H.: Inefficient dynamic portfolio strategies or how to throw away a million dollars in the stock market. Rev. Financ. Stud. 1(1), 67–88 (1988)

    Article  Google Scholar 

  18. Föllmer, H., Schied, A.: Stochastic Finance: An Introduction in Discrete Time, 3rd edn. Walter de Gruyter, Berlin (2011)

    Book  MATH  Google Scholar 

  19. Ghossoub, M.: Equimeasurable rearrangements with capacities. Math. Oper. Res. 40(2), 429–445 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  20. Ghossoub, M.: Cost-efficient contingent claims with market frictions. Math. Finance Econ. 10(1), 87–111 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  21. Gilboa, I., Marinacci, M.: Ambiguity and the Bayesian Paradigm. In: Acemoglu, D., Arellano, M., Dekel, E. (eds.) Advances in Economics and Econometrics: Theory and Applications, Tenth World Congress of the Econometric Society. Cambridge University Press, Cambridge (2013)

    Google Scholar 

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

    Article  MATH  Google Scholar 

  23. Harrison, J.M., Kreps, D.M.: Martingales and arbitrage in multiperiod securities markets. J. Econ. Theory 20(3), 381–408 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  24. He, X., Kouwenberg, R., Zhou, X.Y.: Rank-dependent utility and risk taking in complete markets. SIAM J. Financ. Math. 8(1), 214–239 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  25. Huber, P.J., Strassen, V.: Minimax tests and the Neyman–Pearson lemma for capacities. Ann. Stat. 1(2), 251–263 (1973)

    Article  MathSciNet  MATH  Google Scholar 

  26. Inada, K.: On a two-sector model of economic growth: comments and a generalization. Rev. Econ. Stud. 30(2), 119–127 (1963)

    Article  Google Scholar 

  27. Jouini, E., Kallal, H.: Efficient trading strategies in the presence of market frictions. Rev. Financ. Stud. 14(2), 343–369 (2001)

    Article  Google Scholar 

  28. Jouini, E., Porte, V.: Efficient Trading Strategies. Mimeo, New York (2005)

    Google Scholar 

  29. Kusuoka, S.: On law invariant coherent risk measures. In: Kusuoka, S., Maruyama, T. (eds.) Advances in Mathematical Economics. Advances in Mathematical Economics, vol. 3, pp. 83–95. Springer, Tokyo (2001)

  30. Lehrer, E.: A new integral for capacities. Econ. Theor. 39(1), 157–176 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  31. Marinacci, M., Montrucchio, L.: Introduction to the Mathematics of Ambiguity. In: Gilboa, I. (ed.) Uncertainty in Economic Theory: Essays in Honor of David Schmeidlers 65th Birthday, pp. 46–107. Routledge, London (2004)

    Chapter  Google Scholar 

  32. Markowitz, H.: Portfolio Selection. J. Finance 7(1), 77–91 (1952)

    Google Scholar 

  33. Pap, E.: Null-Additive Set Functions. Kluwer Academic Publishers, Dordrecht (1995)

    MATH  Google Scholar 

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

    Article  Google Scholar 

  35. Rüschendorf, L.: On the distributional transform, Sklar’s theorem, and the empirical copula process. J. Stat. Plan. Inference 139(11), 3921–3927 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  36. Schied, A.: On the Neyman–Pearson problem for law-invariant risk measures and robust utility functionals. Ann. Appl. Probab. 14(3), 1398–1423 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  37. Schmeidler, D.: Integral representation without additivity. Proc. Am. Math. Soc. 97(2), 255–261 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  38. Schmeidler, D.: Subjective probability and expected utility without additivity. Econometrica 57(3), 571–587 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  39. **a, J., Zhou, X.Y.: Arrow–Debreu equilibria for rank-dependent utilities. Math. Finance 26(3), 589–601 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  40. Xu, Z.Q.: A note on the quantile formulation. Math. Finance 26(3), 558–588 (2016)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mario Ghossoub.

Additional information

I am grateful to Frank Riedel and Alexander Schied for comments and suggestions.

Appendices

Appendix A: Choquet integration and quantile-based risk measures

1.1 The Choquet integral

Consider a probability space \(\left( S, \Sigma , P\right) \), and let \(B\left( \Sigma \right) \) denote the linear space of all bounded, real-valued, and \(\Sigma \)-measurable functions on \(\left( S,\Sigma \right) \).

Definition A.1

A (normalized) capacity on the measurable space \(\left( S, \Sigma \right) \) is a set function \(\nu : \Sigma \rightarrow \left[ 0,1\right] \) such that \(\nu \left( \varnothing \right) =0\), \(\nu \left( S\right) = 1\), and \(\nu \) is monotone, that is, for any \(A,B \in \Sigma , \ A \subseteq B \Rightarrow \nu \left( A\right) \leqslant \nu \left( B\right) \).

The capacity \(\nu \) is said to be:

  1. (1)

    supermodular if \(\nu \left( A \cup B\right) + \nu \left( A \cap B\right) \geqslant \nu \left( A\right) + \nu \left( B\right) \), for all \(A,B \in \Sigma \); and,

  2. (2)

    submodular if \(\nu \left( A \cup B\right) + \nu \left( A \cap B\right) \leqslant \nu \left( A\right) + \nu \left( B\right) \), for all \(A,B \in \Sigma \).

Remark A.2

For instance, if \(T : \left[ 0,1\right] \rightarrow \left[ 0,1\right] \) is an increasing function, such that \(T(0)=0\) and \(T(1)=1\), then the set function \(\nu := T \circ P\) is a capacity on \(\left( S, \Sigma \right) \) called a distorted probability measure. The function T is usually called a probability distortion. If, moreover, the distortion function T is convex (resp. concave), then the capacity \(\nu = T \circ P\) is supermodular (resp. submodular) ([9, p. 287] or [14, Ex. 2.1]).

Definition A.3

A capacity \(\nu \) on \(\left( S,\Sigma \right) \) is said to be

  1. (1)

    Continuous from above if for any sequence \(\{A_{n}\}_{n \geqslant 1}\subseteq \Sigma \) such that \(A_{n+1}\subseteq A_{n}\) for each n,

    $$\begin{aligned} \underset{n \rightarrow +\infty }{\lim }\nu \left( A_{n}\right) = \nu \left( \bigcap _{n = 1}^{+\infty } A_{n}\right) ; \end{aligned}$$
  2. (2)

    Continuous from below if for any sequence \(\{A_{n}\}_{n \geqslant 1}\subseteq \Sigma \) such that \(A_{n+1}\supseteq A_{n}\) for each n,

    $$\begin{aligned} \underset{n \rightarrow +\infty }{\lim }\nu \left( A_{n}\right) = \nu \left( \bigcup _{n = 1}^{+\infty } A_{n}\right) ; \end{aligned}$$
  3. (3)

    Continuous if it is continuous both from above and below.

For instance, if \(\nu \) is a distorted probability measure of the form \(T \circ P\) where T is a continuous function, then \(\nu \) is a continuous capacity.

Definition A.4

Let \(\nu \) be a capacity on \(\left( S, \Sigma \right) \). The Choquet integral of \(Y \in B\left( \Sigma \right) \) with respect to \(\nu \) is defined by

$$\begin{aligned} \int Y \ d\nu := \int _{0}^{+\infty } \nu \left( \{ s \in S: Y\left( s\right) \geqslant t \}\right) \ dt + \int _{-\infty }^{0} \left[ \nu \left( \{ s \in S: Y\left( s\right) \geqslant t \}\right) - 1\right] \ dt, \end{aligned}$$

where the integrals are taken in the sense of Riemann.

Definition A.5

Two functions \(Y_{1},Y_{2} \in B\left( \Sigma \right) \) are said to be comonotonic (resp. anti-comonotonic) if

$$\begin{aligned} \Big [ Y_{1}\left( s\right) - Y_{1}\left( s^{\prime }\right) \Big ] \Big [ Y_{2}\left( s\right) - Y_{2}\left( s^{\prime }\right) \Big ] \geqslant 0 \ (\hbox {resp.} \leqslant 0), \hbox { for all } s, s^{\prime } \in S. \end{aligned}$$

Note that \(Y_{1},Y_{2} \in B\left( \Sigma \right) \) are anti-comonotonic if and only if \(Y_{1}\) and \(-Y_{2}\) are comonotonic. Moreover, any \(Y \in B\left( \Sigma \right) \) is both comonotonic and anti-comonotonic with any \(c \in \mathbb {R}\). Also, if \(Y_{1},Y_{2} \in B\left( \Sigma \right) \), and if \(Y_{2}\) is of the form \(Y_{2} = I \circ Y_{1}\), for some Borel-measurable function I, then \(Y_{2}\) is comonotonic (resp. anti-comonotonic) with \(Y_{1}\) if and only if the function I is nondecreasing (resp. nonincreasing).

Intuitively, two functions that are anti-comonotonic are a hedge against each other. The following result formalizes this fact.

Proposition A.6

(Prop. 4.5 of Denneberg [14]) Two functions \(Y_{1},Y_{2}: S \rightarrow \mathbb {R}\) are anti-comonotonic if and only if there is a function \(Z: S \rightarrow \mathbb {R}\), a non-decreasing function \(u: \mathbb {R} \rightarrow \mathbb {R}\), and a nonincreasing function \(v: \mathbb {R} \rightarrow \mathbb {R}\) such that

$$\begin{aligned} Y_{1} = u \circ Z \ \ \hbox {and} \ \ Y_{2} = v \circ Z. \end{aligned}$$

The Choquet integral with respect to a (countably additive) measure is the usual Lebesgue integral with respect to that measure [31, p. 59]. Unlike the Lebesgue integral, the Choquet integral is not an additive operator on \(B\left( \Sigma \right) \). However, the Choquet integral is additive over comonotonic functions.

Proposition A.7

Let \(\nu \) be a capacity on \(\left( S, \Sigma \right) \).

  1. (1)

    Comonotonic Additivity: If \(\phi _{1}, \phi _{2} \in B\left( \Sigma \right) \) are comonotonic, then \(\int \left( \phi _{1} + \phi _{2}\right) \ d\nu = \int \phi _{1} \ d\nu + \int \phi _{2} \ d\nu \);

  2. (2)

    Monotonicity: If \(\phi _{1}, \phi _{2} \in B\left( \Sigma \right) \) are such that \(\phi _{1} \leqslant \phi _{2}\), then \(\int \phi _{1} \ d\nu \leqslant \int \phi _{2} \ d\nu \);

  3. (3)

    Positive homogeneity: For all \(\phi \in B\left( \Sigma \right) \) and all \(c \geqslant 0\), \(\int c\phi \ d\nu = c\int \phi \ d\nu \);

  4. (4)

    If \(\nu \) is submodular, then for any \(\phi _{1}, \phi _{2} \in B\left( \Sigma \right) \), \(\int \left( \phi _{1} + \phi _{2}\right) \ d\nu \leqslant \int \phi _{1} \ d\nu + \int \phi _{2} \ d\nu \);

  5. (5)

    If \(\nu \) is supermodular, then for any \(\phi _{1}, \phi _{2} \in B\left( \Sigma \right) \), \(\int \left( \phi _{1} + \phi _{2}\right) \ d\nu \geqslant \int \phi _{1} \ d\nu + \int \phi _{2} \ d\nu \).

We refer to Marinacci and Montrucchio [31] or Denneberg [14] for more about capacities and Choquet integrals.

1.2 Strongly diffuse capacities

If \(\nu \) is a capacity on \(\left( S, \Sigma \right) \) and X a random variable on \(\left( S, \Sigma \right) \), then the set function \(\nu \circ X^{-1}\) defined on the Borel \(\sigma \)-algebra on the range of X is a capacity.

Definition A.8

(Ghossoub [19]) Let \(\nu \) be a capacity on \(\left( S, \Sigma \right) \) and let X be a random variable on \(\left( S, \Sigma \right) \). Then the capacity \(\nu \circ X^{-1}\) is said to be:

  1. (1)

    Diffuse if \(\nu \circ X^{-1} \Big (\{t\}\Big ) = 0\), for all \(t \in \mathbb {R}\);

  2. (2)

    Strongly diffuse if \(\nu \circ X^{-1}\Big (\left( a,b\right) \Big ) = \nu \circ X^{-1}\Big (\left[ a,b\right] \Big )\), for all \(a,b \in \mathbb {R}\) such that \(a \leqslant b\).

If \(\nu \circ X^{-1}\) is strongly diffuse, we will also say that \(\nu \) is strongly diffuse with respect to X (or simply, strongly diffuse if the context is clear). Strong diffuseness implies diffuseness. For capacities that are distortions of a probability measure, we have the following stronger result.

Proposition A.9

(Ghossoub [19]) Let \(\nu \) be a capacity on \(\left( S, \Sigma \right) \) and let X be a random variable on \(\left( S, \Sigma \right) \), and suppose that \(\nu \) is a distorted probability measure of the form \(\nu = T \circ P\), for some probability measure P on \(\left( S, \Sigma \right) \) and some distortion function \(T: \left[ 0,1\right] \rightarrow \left[ 0,1\right] \), strictly increasing with \(T\left( 0\right) = 0\) and \(T\left( 1\right) = 1\). Then the following are equivalent.

  1. (1)

    \(\nu \circ X^{-1}\) is strongly diffuse;

  2. (2)

    \(\nu \circ X^{-1}\) is diffuse; and,

  3. (3)

    \(P \circ X^{-1}\) is diffuse (i.e., nonatomic).

Definition A.10

Let \(\nu \) be a capacity on the measurable space \(\left( S, \Sigma \right) \) and let \(\phi \in B\left( \Sigma \right) \). Define the upper-distribution of \(\phi \) with respect to \(\nu \) as the function

$$\begin{aligned} \begin{aligned} G_{\nu ,\phi }:&\ \mathbb {R} \rightarrow \left[ 0,1\right] \\&\ t \mapsto G_{\nu ,\phi }\left( t\right) := \nu \big ( \{s \in S: \phi \left( s\right) > t\} \big ). \end{aligned} \end{aligned}$$

If \(\phi _{1}, \phi _{2} \in B\left( \Sigma \right) \), we write \(\phi _{1} \overset{\nu }{\sim }\phi _{2}\) to mean that \(\phi _{1}\) and \(\phi _{2}\) have the same upper-distribution with respect to \(\nu \). Then a map** \(V: B\left( \Sigma \right) \rightarrow \mathbb {R}\) is said to be \(\nu \)-upper-law-invariant if for any \(\phi _{1}, \phi _{2} \in B\left( \Sigma \right) \),

$$\begin{aligned} \phi _{1} \overset{\nu }{\sim }\phi _{2} \ \Longrightarrow \ V\left( \phi _{1}\right) = V\left( \phi _{2}\right) . \end{aligned}$$

The Choquet integral is an example of a \(\nu \)-upper-law-invariant functional on \(B\left( \Sigma \right) \). Note that \(G_{\nu ,\psi }\) is nonincreasing, and if \(\nu \) is continuous from below then \(G_{\nu ,\psi }\) is right-continuous [14, p. 46]. Moreover, if \(\nu = T \circ P\), for some probability measure P on \(\left( S, \Sigma \right) \) and some distortion function \(T : \left[ 0,1\right] \rightarrow \left[ 0,1\right] \), then for any \(\phi _{1}, \phi _{2} \in B\left( \Sigma \right) \), if \(\phi _{1}\) and \(\phi _{2}\) are identically distributedFootnote 9 according to P, then they have the same upper-distribution with respect to \(\nu \). Finally, if \(\nu \) is a bone fide additive measure, then two functions have the same upper-distribution with respect to \(\nu \) if and only if they are identically distributed according to \(\nu \).

1.3 Robust representations of the Choquet integral

Let \(ba\left( \Sigma \right) \) denote the linear space of all bounded finitely additive set functions on \(\left( S,\Sigma \right) \), endowed with the usual mixing operations. When endowed with the variation norm \(\Vert . \Vert _{v}\), \(ba\left( \Sigma \right) \) is a Banach space. By a classical result [15, IV.5.1], \(\left( ba\left( \Sigma \right) , \Vert .\Vert _{v}\right) \) is isometrically isomorphic to the norm-dual of the Banach space \(\left( B\left( \Sigma \right) ,\Vert .\Vert _{sup} \right) \) via the duality \(<\phi , \lambda \>= \int \phi \ d \lambda , \ \forall \lambda \in ba\left( \Sigma \right) , \ \forall \phi \in B\left( \Sigma \right) \). Consequently, we can endow \(ba\left( \Sigma \right) \) with the weak\(^{*}\) topology \(\sigma \left( ba\left( \Sigma \right) , B\left( \Sigma \right) \right) \). If \(ca\left( \Sigma \right) \) denotes the collection of all countably additive elements of \(ba\left( \Sigma \right) \), then \(ca\left( \Sigma \right) \) is a \(\Vert .\Vert _{v}\)-closed linear subspace of \(ba\left( \Sigma \right) \). Hence, \(ca\left( \Sigma \right) \) is \(\Vert .\Vert _{v}\)-complete, i.e., \(\left( ca\left( \Sigma \right) , \Vert .\Vert _{v}\right) \) is a Banach space. By a classical result of Huber and Strassen [25] and Schmeidler [37, 38], we have the following representations of the Choquet integral with respect to a given capacity.

Proposition A.11

Let \(\nu \) be a capacity on \(\left( S, \Sigma \right) \).

  1. (1)

    If \(\nu \) is supermodular (e.g., a convex distortion of a probability measure), then there exists a non-empty, convex, and weak\(^{*}\)-compact collection \(\Pi \subset ca\left( \Sigma \right) \) of probability measures, called the core of \(\nu \) such that for all \(Y \in B\left( \Sigma \right) \),

    $$\begin{aligned} \int Y d\nu = \underset{\mu \in \Pi }{\min }\int Y d\mu . \end{aligned}$$
  2. (2)

    If \(\nu \) is submodular (e.g., a concave distortion of a probability measure), then there exists a non-empty, convex, and weak\(^{*}\)-compact collection \(\mathcal {A} \subset ca\left( \Sigma \right) \) of probability measures, called the anti-core of \(\nu \) such that for all \(Y \in B\left( \Sigma \right) \),

    $$\begin{aligned}\int Y d\nu = \underset{\mu \in \mathcal {A}}{\max }\int Y d\mu .\end{aligned}$$

1.4 Quantiles and quantile-based risk measures

For \(Y \in B\left( \Sigma \right) \), let \(F_{Y}\left( t\right) := P\left( \{s \in S : Y\left( s\right) \leqslant t\}\right) \) denote the cumulative distribution function of Y with respect to the probability measure P, and let \(F^{-1}_{Y}\left( t\right) \) denote the left-continuous inverse of the \(F_{Y}\) (i.e., the quantile function of Y), defined by

$$\begin{aligned} F^{-1}_{Y}\left( t\right) := \inf \Big \{z \in \mathbb {R} \ \Big | \ F_{Y}\left( z\right) \geqslant t\Big \}, \ \forall t \in \left[ 0, 1\right] . \end{aligned}$$

If the space \(\left( S, \Sigma , P\right) \) is nonatomic, then there exists a random variable U that is uniformly distributed on \(\left( 0,1\right) \) [18, Proposition A.27]. For all \(Y \in B\left( \Sigma \right) \) and all distortion functions T, we then have

$$\begin{aligned} \int Y dT \circ P = \int T^{\prime }\left( 1-U\right) F^{-1}_{Y}\left( U\right) dP = \int _{0}^{1} T^{\prime }\left( 1-t\right) F^{-1}_{Y}\left( t\right) dt. \end{aligned}$$

Letting \(k\left( t\right) = T^{\prime }\left( 1-t\right) \), it follows that

$$\begin{aligned} \int Y dT \circ P = \int _{0}^{1} k\left( t\right) F^{-1}_{Y}\left( t\right) dt, \end{aligned}$$
(A.1)

and \(\int _{0}^{1} k\left( t\right) dt = 1\). Risk measures \(\rho : B\left( \Sigma \right) \rightarrow \mathbb {R}\) of the form given in Eq. (A.1) are referred to as quantile-based risk measures (Kusuoka [29]). By Proposition A.11, we obtain the following result.

Proposition A.12

Let \(\rho : B\left( \Sigma \right) \rightarrow \mathbb {R}\) be a quantile-based risk measures of the form \(\rho \left( Y\right) = \int _{0}^{1} k\left( t\right) F^{-1}_{Y}\left( t\right) dt\), where \(\int _{0}^{1} k\left( t\right) dt = 1\).

  1. (1)

    If k is increasing, then there exists a non-empty, convex, and weak\(^{*}\)-compact collection of probability measures \(\mathcal {A} \subset ca\left( \Sigma \right) \) such that \(\rho \) has the following robust representation:

    $$\begin{aligned} \rho \left( Y\right) = \underset{\mu \in \mathcal {A}}{\max }\int Y d\mu , \hbox { for all } Y \in B\left( \Sigma \right) . \end{aligned}$$
  2. (2)

    If k is decreasing, then there exists a non-empty, convex, and weak\(^{*}\)-compact collection of probability measures \(\Pi \subset ca\left( \Sigma \right) \) such that \(\rho \) has the following robust representation:

    $$\begin{aligned} \rho \left( Y\right) = \underset{\mu \in \Pi }{\min }\int Y d\mu , \hbox { for all } Y \in B\left( \Sigma \right) . \end{aligned}$$

Remark A.13

Given a quantile-based risk measures of the form \(\rho \left( Y\right) = \int _{0}^{1} k\left( t\right) F^{-1}_{Y}\left( t\right) dt\), where \(\int _{0}^{1} k\left( t\right) dt = 1\), defining an absolutely continuous function \(T: \left[ 0,1\right] \rightarrow \left[ 0,1\right] \) by \(T^{\prime }\left( 1-t\right) = k\left( t\right) \) yields a representation of quantile-based risk measures as Choquet integrals with respect to the distorted probability measure \(T \circ P\).

Appendix B: Proof of Theorem 3.1

1.1 Quantile re-formulation

For \(Y \in B\left( \Sigma \right) \), let \(F_{Y}\left( t\right) := P\{s \in S : Y\left( s\right) \leqslant t\}\) denote the cumulative distribution function of Y with respect to the probability measure P, and let \(F^{-1}_{Y}\left( t\right) \) denote the left-continuous inverse of the \(F_{Y}\) (i.e., the quantile function of Y), defined by

$$\begin{aligned} F^{-1}_{Y}\left( t\right) := \inf \Big \{z \in \mathbb {R} \ \Big | \ F_{Y}\left( z\right) \geqslant t\Big \}, \ \forall t \in \left[ 0, 1\right] . \end{aligned}$$

Let \(U := F_{X}\left( X\right) \). By assumption of nonatomicity of \(P \circ X^{-1}\), U is a uniformly distributed random variable on \(\left( 0,1\right) \) [18, Lemma A.21]. For all \(Y \in B\left( \Sigma \right) \), the fact that \(\phi \) is increasing and U is uniformly distributed on \(\left( 0,1\right) \) implies that

$$\begin{aligned} \begin{aligned} \int \phi \left( Y\right) d T_{2}\circ P&= \int T_{2}^{\prime }\left( 1-U\right) F^{-1}_{\phi \left( Y\right) }\left( U\right) dP = \int T_{2}^{\prime }\left( 1-U\right) \phi \left( F^{-1}_{Y}\left( U\right) \right) dP \\&= \int _{0}^{1} T_{2}^{\prime }\left( 1-t\right) \phi \left( F^{-1}_{Y}\left( t\right) \right) dt = \int _{0}^{1} T_{2}^{\prime }\left( t\right) \phi \left( F^{-1}_{Y}\left( 1-t\right) \right) dt. \\ \end{aligned} \end{aligned}$$

Moreover,

$$\begin{aligned} \int Y dT_{1} \circ P= & {} \int T_{1}^{\prime }\left( 1-U\right) F^{-1}_{Y}\left( U\right) dP \\= & {} \int _{0}^{1} T_{1}^{\prime }\left( 1-t\right) F^{-1}_{Y}\left( t\right) dt = \int _{0}^{1} T_{1}^{\prime }\left( t\right) F^{-1}_{Y}\left( 1-t\right) dt, \end{aligned}$$

and \(0 \leqslant Y \leqslant N\) whenever \(0 \leqslant F_{Y}^{-1}\left( t\right) \leqslant N\), for all \(t \in \left( 0,1\right) .\)

Let \(\mathcal {Q}\) denote the collection of all quantile functions, and let \(\mathcal {Q}^{*}\) denote the collection of all quantile functions f that satisfy \(0 \leqslant f\left( t\right) \leqslant N\), for all \(t \in \left( 0,1\right) \). Then

$$\begin{aligned} Q = \Big \{f: \left( 0,1\right) \rightarrow \mathbb {R} \ \Big | \ f \hbox { is nondecreasing and left-continuous}\Big \}, \end{aligned}$$

and

$$\begin{aligned} \mathcal {Q}^{*} = \Big \{ f \in \mathcal {Q}: 0 \leqslant f\left( t\right) \leqslant N, \hbox { for each } 0< t < 1\Big \}. \end{aligned}$$
(B.1)

Consider the following problem:

Problem B.1

For \(P_{0}\) as in Assumption 2.4,

$$\begin{aligned} \inf _{f \in \mathcal {Q}^{*}} \left\{ \int _{0}^{1} T_{1}^{\prime }\left( 1-t\right) f\left( t\right) dt : \int _{0}^{1} T_{2}^{\prime }\left( 1-t\right) \phi \left( f\left( t\right) \right) dt \geqslant P_{0} \right\} . \end{aligned}$$
(B.2)

Lemma B.2

If \(f^{*}\) is optimal for Problem (B.1), then \(Y^{*} := f^{*}\left( 1-F_{X}\left( X\right) \right) \) is optimal for Problem (2.6) and anti-comonotonic with X.

Proof

Let \(f^{*}\) be optimal for Problem (B.1). Then, by definition of \(\mathcal {Q}^{*}\), \(f^{*}\) is the quantile function of some \(Z \in B\left( \Sigma \right) \) such that \(0 \leqslant Z \leqslant N\). Therefore, since \(U := F_{X}\left( X\right) \) is a uniformly distributed random variable on \(\left( 0,1\right) \), \(Y^{*} = f^{*}\left( 1-U\right) = F_{Z}^{-1}\left( 1-U\right) \) is the nondecreasing equimeasurable rearrangement of Z with respect to X, and hence \(0 \leqslant Y^{*} \leqslant N\) and \(F_{Y^{*}} = F_{Z}\) (see Ghossoub [19] and references therein). Thus, by law-invariance of the Choquet integral with respect to a distortion of P,

$$\begin{aligned} \int \phi \left( Y^{*}\right) \ dT_{2} \circ P = \int \phi \left( Z\right) \ dT_{2} \circ P= \int _{0}^{1} T_{2}^{\prime }\left( 1-t\right) \phi \left( f\left( t\right) \right) dt \geqslant P_{0}, \end{aligned}$$

where the last inequality follows from the feasibility of \(f^{*}\) for Problem (B.1). Hence, \(Y^{*}\) is feasible for Problem (2.6).

To show optimality of \(Y^{*}\) for Problem (2.6), let Y by any other feasible solution for Problem (2.6) and \(F_{Y}^{-1}\) its quantile function. Then \(F_{Y}^{-1}\) is feasible for Problem (B.1), and hence

$$\begin{aligned} \int Y d T_{1}\circ P= & {} \int _{0}^{1} T_{1}^{\prime }\left( 1-t\right) F^{-1}_{R}\left( t\right) dt \\\geqslant & {} \int _{0}^{1} T_{1}^{\prime }\left( 1-t\right) f^{*}\left( t\right) dt =\int Z d T_{1}\circ P =\int Y^{*} d T_{1}\circ P. \end{aligned}$$

Therefore, \(Y^{*}\) is optimal or Problem (2.6). \(\square \)

Now, letting \(v\left( t\right) = T_{1}^{-1}\left( t\right) \) and using the change of variable \(z = v^{-1}\left( t\right) \) gives, for all \(f \in \mathcal {Q}^{**}\),

$$\begin{aligned} \int _{0}^{1} T_{1}^{\prime }\left( t\right) f\left( 1-t\right) dt= & {} \int _{0}^{1} f\left( 1-t\right) dT_{1}\left( t\right) \\= & {} \int _{0}^{1} f\left( 1-t\right) dv^{-1}\left( t\right) =\int _{0}^{1} f\left( 1-v\left( z\right) \right) dz =\int _{0}^{1} q\left( t\right) dt, \end{aligned}$$

where \(q\left( t\right) := f\left( 1-v\left( t\right) \right) \), for all \(t \in \left( 0,1\right) \). Moreover,

$$\begin{aligned} \begin{aligned} \int _{0}^{1} T_{2}^{\prime }\left( t\right) \phi \left( f\left( 1-t\right) \right) dt&=\int _{0}^{1} T_{2}^{\prime }\left( v\left( z\right) \right) \phi \left( f\left( 1-v\left( z\right) \right) \right) dv\left( z\right) \\&=\int _{0}^{1} T_{2}^{\prime }\left( T_{1}^{-1}\left( z\right) \right) \phi \left( q\left( z\right) \right) v^{\prime }\left( z\right) dz\\&=\int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) } \phi \left( q\left( t\right) \right) dt. \\ \end{aligned} \end{aligned}$$

Now, define the set \(\mathcal {Q}^{**}\) by:

$$\begin{aligned} \mathcal {Q}^{**}:= & {} \Big \{q: \left( 0,1\right) \rightarrow \mathbb {R} \ \Big | \ q \hbox { is nonincreasing } \hbox { and left-continuous, } \nonumber \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \hbox {and } 0 \leqslant q\left( t\right) \leqslant N, \hbox { for each } 0< t < 1\Big \},\nonumber \\ \end{aligned}$$
(B.3)

and consider the following problem:

Problem B.3

For \(P_{0}\) as in Assumption 2.4,

$$\begin{aligned} \inf _{q \in \mathcal {Q}^{**}} \left\{ \int _{0}^{1} q\left( t\right) dt: \int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) } \phi \left( q\left( t\right) \right) dt \geqslant P_{0} \right\} . \end{aligned}$$

Lemma B.4

If \(q^{*}\) is optimal for Problem (B.3), then the function \(f^{*}\) defined by \(f^{*}\left( t\right) := q^{*}\left( T_{1}\left( 1-t\right) \right) \) is optimal for Problem (B.1). Moreover, \(Y^{*} := f^{*}\left( 1-F_{X}\left( X\right) \right) = q^{*} \left( T_{1} \left( F_{X}\left( X\right) \right) \right) \) is optimal for Problem (2.6) and anti-comonotonic with X.

Proof

Suppose \(q^{*}\) is optimal for Problem (B.3), and let \(f^{*}\left( t\right) := q^{*}\left( T_{1}\left( 1-t\right) \right) \). Then

  • \(0 \leqslant f^{*} \leqslant N\) since \(q^{*} \in \mathcal {Q}^{**}\);

  • By continuity of \(T_{1}\) and left-continuity of \(q^{*}\), \(f^{*}\) is left-continuous;

  • Since \(q^{*}\) is nonincreasing and \(T_{1}\) is increasing, \(f^{*}\) is nondecreasing;

  • Morevoer,

    $$\begin{aligned} \int _{0}^{1} T_{2}^{\prime }\left( 1-t\right) \phi \left( f^{*}\left( t\right) \right) dt= & {} \int _{0}^{1} T_{2}^{\prime }\left( t\right) \phi \left( f^{*}\left( 1-t\right) \right) dt\\= & {} \int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( z\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( z\right) \right) } \phi \left( f^{*}\left( 1-v\left( z\right) \right) \right) dz\\= & {} \int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( z\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( z\right) \right) } \phi \left( q^{*}\left( T_{1}\left( 1-\left( 1-v\left( z\right) \right) \right) \right) \right) dz\\= & {} \int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( z\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( z\right) \right) } \phi \left( q^{*}\left( T_{1}\left( v\left( z\right) \right) \right) \right) dz\\= & {} \int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) } \phi \left( q^{*}\left( t\right) \right) dt \geqslant P_{0}, \end{aligned}$$

where the last equality follows from the feasibility of \(q^{*}\) for Problem (B.3).

Therefore, \(f^{*}\) is feasible for Problem (B.1). To show optimality of \(f^{*}\) for Problem (B.1), let f be any other feasible solution for Problem (B.1) and define q by \(q\left( t\right) := f\left( 1-v\left( t\right) \right) \). Then:

  • \(0 \leqslant q \leqslant N\) since \(f^{*} \in \mathcal {Q}^{*}\);

  • q is left-continuous and nonincreasing, by the Inverse Function Theorem;

  • Moreover,

    $$\begin{aligned} \begin{aligned} \int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) } \phi \left( q\left( t\right) \right) dt&=\int _{0}^{1} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) } \phi \left( f\left( 1-v\left( t\right) \right) \right) dt\\&=\int _{0}^{1} T^{\prime }_{2}\left( t\right) \phi \left( f\left( 1-t\right) \right) dt \\&=\int _{0}^{1} T^{\prime }_{2}\left( 1-t\right) \phi \left( f\left( t\right) \right) dt \geqslant P_{0}, \end{aligned} \end{aligned}$$

where the last equality follows from the feasibility of f for Problem (B.1).

Hence, q is feasible for Problem (B.3), and so \(\int _{0}^{1} q^{*}\left( t\right) dt \leqslant \int _{0}^{1} q\left( t\right) dt.\) But,

$$\begin{aligned} \begin{aligned} \int _{0}^{1} T^{\prime }\left( 1-t\right) f^{*}\left( t\right) dt&= \int _{0}^{1}T^{\prime }\left( t\right) f^{*}\left( 1-t\right) dt \\&= \int _{0}^{1} f^{*}\left( 1-v\left( t\right) \right) dt = \int _{0}^{1} q^{*}\left( T_{1}\left( v\left( t\right) \right) \right) dt = \int _{0}^{1} q^{*}\left( t\right) dt, \end{aligned} \end{aligned}$$

and \(\int _{0}^{1} T^{\prime }\left( 1-t\right) f\left( t\right) dt = \int _{0}^{1} q\left( t\right) dt\). Therefore,

$$\begin{aligned} \int _{0}^{1} T^{\prime }\left( 1-t\right) f^{*}\left( t\right) dt \leqslant \int _{0}^{1} T^{\prime }\left( 1-t\right) f\left( t\right) dt. \end{aligned}$$

Hence, \(f^{*}\) is optimal for Problem (B.1), and the rest follows from Lemma B.2. \(\square \)

1.2 Solving problem (B.3)

In light of Lemma B.4, we turn our attention to solving Problem (B.3). In order to do that, we will use a similar methodology to the one used by Xu [40], but modified and adapted to the present setting.

Now, define the function \(\psi : \left[ 0,1\right] \rightarrow \mathbb {R}^{+}\) by

$$\begin{aligned} \psi \left( t\right) := \int _{0}^{t} \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( x\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( x\right) \right) } dx = \int _{0}^{T_{1}^{-1}\left( t\right) } T_{2}^{\prime }\left( x\right) dx = T_{2}\left( T_{1}^{-1}\left( t\right) \right) , \end{aligned}$$
(B.4)

so that \(\psi ^{\prime }\left( t\right) = \frac{T_{2}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) }{T_{1}^{\prime }\left( T_{1}^{-1}\left( t\right) \right) }\).

Lemma B.5

Let \(\delta \) be the concave envelope of the function \(\psi \) on \(\left[ 0,1\right] \). Then for any \(q \in \mathcal {Q}^{**}\),

$$\begin{aligned} \int _{0}^{1} \left( \phi \circ q\right) \left( t\right) \psi ^{\prime }\left( t\right) dt \leqslant \int _{0}^{1} \left( \phi \circ q\right) \left( t\right) \delta ^{\prime }\left( t\right) dt. \end{aligned}$$

Proof

Let \(\delta \) be the concave envelope of the function \(\psi \) on \(\left[ 0,1\right] \). Since \(\phi \) is increasing and each \(q \in \mathcal {Q}^{**}\) is nonincreasing, and since \(\delta \left( t\right) \geqslant \psi \left( t\right) \), for all \(t \in \left[ 0,1\right] \), it follows thatFootnote 10 for all \(q \in \mathcal {Q}^{**}\),

$$\begin{aligned} \int _{0}^{1} \left[ \delta \left( y\right) - \psi \left( y\right) \right] d \phi \circ q\left( y\right) \leqslant 0. \end{aligned}$$

Therefore, since \(\psi \left( 0\right) = \delta \left( 0\right) \) and \(\psi \left( 1\right) = \delta \left( 1\right) \), Fubini’s TheoremFootnote 11 gives

$$\begin{aligned} \begin{aligned} 0&\leqslant \int _{0}^{1} \left[ \left( \delta \left( 1\right) - \psi \left( 1\right) \right) - \left( \delta \left( y\right) - \psi \left( y\right) \right) \right] d \phi \circ q\left( y\right) \\&= \int _{0}^{1} \int _{y}^{1} \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] dx \ d \phi \circ q\left( y\right) \\&= \int _{0}^{1} \int _{0}^{x} \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] d \phi \circ q\left( y\right) \ dx = \int _{0}^{1} \left[ \int _{0}^{x} d \phi \circ q\left( y\right) \right] \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] dx \\&= \int _{0}^{1} \left[ \left( \phi \circ q\right) \left( x\right) - \left( \phi \circ q\right) \left( 0\right) \right] \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] dx \\&= \int _{0}^{1} \delta ^{\prime }\left( x\right) \left( \phi \circ q\right) \left( x\right) dx - \int _{0}^{1} \psi ^{\prime }\left( x\right) \left( \phi \circ q\right) \left( x\right) dx - \left( \phi \circ q\right) \left( 0\right) \left[ \delta \left( 1\right) - \delta \left( 0\right) \right] \\&\quad + \left( \phi \circ q\right) \left( 0\right) \left[ \psi \left( 1\right) - \psi \left( 0\right) \right] \\&= \int _{0}^{1} \delta ^{\prime }\left( x\right) \left( \phi \circ q\right) \left( x\right) dx - \int _{0}^{1} \psi ^{\prime }\left( x\right) \left( \phi \circ q\right) \left( x\right) dx - \left( \phi \circ q\right) \left( 0\right) \left[ \psi \left( 1\right) - \psi \left( 0\right) \right] \\&\quad + \left( \phi \circ q\right) \left( 0\right) \left[ \psi \left( 1\right) - \psi \left( 0\right) \right] \\&= \int _{0}^{1} \delta ^{\prime }\left( x\right) \left( \phi \circ q\right) \left( x\right) dx - \int _{0}^{1} \psi ^{\prime }\left( x\right) \left( \phi \circ q\right) \left( x\right) dx. \end{aligned} \end{aligned}$$

That is, \( \int _{0}^{1} \left( \phi \circ q\right) \left( t\right) \psi ^{\prime }\left( t\right) dt \leqslant \int _{0}^{1} \left( \phi \circ q\right) \left( t\right) \delta ^{\prime }\left( t\right) dt\). \(\square \)

Now consider the following problem:

Problem B.6

For \(P_{0}\) as in Assumption 2.4,

$$\begin{aligned} \inf _{q \in \mathcal {Q}^{**}} \left\{ \int _{0}^{1} q\left( t\right) dt: \int _{0}^{1} \delta ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) dt \geqslant P_{0} \right\} . \end{aligned}$$

We first solve Problem (B.6) and then show that the solution is also optimal for Problem (B.3).

Lemma B.7

If \(q^{*} \in \mathcal {Q}^{**}\) satisfies:

  1. (1)

    \(\int _{0}^{1} \delta ^{\prime }\left( t\right) \left( \phi \circ q^{*}\right) \left( t\right) dt = P_{0}\); and,

  2. (2)

    There exists some \(\lambda \geqslant 0\) such that for all \(t \in \left( 0,1\right) \),

    $$\begin{aligned} q^{*}\left( t\right) = \underset{0 \leqslant y \leqslant N}{{{\mathrm{\arg \min }}}}\left[ y - \lambda \delta ^{\prime }\left( t\right) \phi \left( y\right) \right] , \end{aligned}$$

Then \(q^{*}\) is optimal for Problem (B.6).

Proof

Let \(q^{*} \in \mathcal {Q}^{**}\) be such that the two conditions above are satisfied. Then \(q^{*}\) is feasible for Problem (B.6). To show optimality, let \(q \in \mathcal {Q}^{**}\) be any feasible solution for Problem (B.6). Then, by definition of \(q^{*}\), it follws that for each t,

$$\begin{aligned} q^{*}\left( t\right) - q\left( t\right) \leqslant \lambda \delta ^{\prime }\left( t\right) \left[ \phi \left( q^{*}\left( t\right) \right) - \phi \left( q\left( t\right) \right) \right] . \end{aligned}$$

Hence,

$$\begin{aligned} \int _{0}^{1} q^{*}\left( t\right) dt - \int _{0}^{1} q\left( t\right) dt\leqslant & {} \lambda \left[ \int _{0}^{1} \delta ^{\prime }\left( t\right) \phi \left( q^{*}\left( t\right) \right) dt - \int _{0}^{1} \delta ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) dt\right] \\= & {} \lambda \left( P_{0} - \int _{0}^{1} \delta ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) dt\right) \leqslant 0. \end{aligned}$$

Therefore, \(\int _{0}^{1} q^{*}\left( t\right) dt \leqslant \int _{0}^{1} q\left( t\right) dt\). \(\square \)

Lemma B.8

For each \(\lambda > 0\), define the function \(q^{*}_{\lambda }\) by

$$\begin{aligned} q^{*}_{\lambda }\left( t\right) := \min \left[ N, \max \left( 0, \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) \right) \right] . \end{aligned}$$
(B.5)

Then:

  1. (1)

    For each \(\lambda > 0\), \(q^{*}_{\lambda } \in \mathcal {Q}^{**}\);

  2. (2)

    There exists \(\lambda ^{*} \geqslant 0\) such that \(\int _{0}^{1} \delta ^{\prime }\left( t\right) \left( \phi \circ q^{*}_{\lambda ^{*}}\right) \left( t\right) dt = P_{0}\); and

  3. (3)

    For all \(t \in \left( 0,1\right) \), \(q^{*}_{\lambda ^{*}}\left( t\right) = \underset{0 \leqslant y \leqslant N}{{{\mathrm{\arg \min }}}}\left[ y - \lambda ^{*} \delta ^{\prime }\left( t\right) \phi \left( y\right) \right] .\)

Proof

Follows from convexity of \(\phi \) and the monotonicity and continuity properties of \(\phi \) and \(\delta ^{\prime }\), as well as from Assumption 2.4 and the Intermediate Value Theorem. \(\square \)

Therefore, lemmata B.5, B.7, and B.8 imply that for any \(\lambda > 0\) and any \(q \in \mathcal {Q}^{**}\),

$$\begin{aligned} \begin{aligned} \int _{0}^{1}\left[ q\left( t\right) - \lambda \psi ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) \right] dt&= \int _{0}^{1}q\left( t\right) dt - \lambda \int _{0}^{1} \psi ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) dt\\&\geqslant \int _{0}^{1}q\left( t\right) dt - \lambda \int _{0}^{1} \delta ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) dt \\&= \int _{0}^{1}\left[ q\left( t\right) dt - \lambda \delta ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) \right] dt \\&\geqslant \int _{0}^{1}\left[ q^{*}_{\lambda }\left( t\right) dt - \lambda \delta ^{\prime }\left( t\right) \phi \left( q^{*}_{\lambda }\left( t\right) \right) \right] dt, \end{aligned} \end{aligned}$$

where \(q^{*}_{\lambda }\) is as in Eq. (B.5). Now, for all \(\lambda > 0\), since \(q^{*}_{\lambda }\) is monotone, it is differentiable a.e., and we have:

$$\begin{aligned} q^{*}_{\lambda }\left( t\right) = \left\{ \begin{array}{ll} 0 &{} \quad \hbox {if }\ \ \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) \leqslant 0,\\ \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) &{} \quad \hbox {if }\ \ 0< \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) < N,\\ N &{} \quad \hbox {if }\ \ \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) \geqslant N,\\ \end{array} \right. \end{aligned}$$

and

$$\begin{aligned} dq^{*}_{\lambda }\left( t\right) = \left\{ \begin{array}{l l} 0 &{} \quad \hbox {if }\ \ \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) \leqslant 0,\\ \xi _{\lambda }\left( t\right) d\delta ^{\prime }\left( t\right) &{} \quad \hbox {if }\ \ 0< \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) < N,\\ 0 &{} \quad \hbox {if }\ \ \left( \phi ^{\prime }\right) ^{-1}\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) \geqslant N,\\ \end{array} \right. \end{aligned}$$
(B.6)

where \(\xi _{\lambda }\left( t\right) := \frac{- \left( \left( \phi ^{\prime }\right) ^{-1}\right) ^{\prime }\left( \frac{1}{\lambda \delta ^{\prime }\left( t\right) }\right) }{\lambda \left( \delta ^{\prime }\right) ^{2}\left( t\right) }\).

Now, define the subsets \(\mathcal {A}\) and \(\mathcal {B}\) of \(\left[ 0,1\right] \) by:

$$\begin{aligned} \mathcal {A}:= & {} \left\{ t \in \left[ 0,1\right] : \delta \left( t\right) = \psi \left( t\right) \right\} \ \hbox {and }\\ \mathcal {B}:= & {} \left\{ t \in \left[ 0,1\right] : \delta \left( t\right) \ne \psi \left( t\right) \right\} = \left\{ t \in \left[ 0,1\right] : \delta \left( t\right) > \psi \left( t\right) \right\} . \end{aligned}$$

Then for any \(\lambda > 0\),

$$\begin{aligned} \begin{aligned} \int _{0}^{1} \left[ \delta \left( t\right) - \psi \left( t\right) \right] d \phi \circ q^{*}_{\lambda }\left( t\right)&= \int _{0}^{1} \phi ^{\prime }\left( q^{*}_{\lambda }\left( t\right) \right) \left[ \delta \left( t\right) - \psi \left( t\right) \right] dq^{*}_{\lambda }\left( t\right) \\&= \int _{\mathcal {A}} \phi ^{\prime }\left( q^{*}_{\lambda }\left( t\right) \right) \left[ \delta \left( t\right) - \psi \left( t\right) \right] dq^{*}_{\lambda }\left( t\right) \\&\quad + \int _{\mathcal {B}} \phi ^{\prime }\left( q^{*}_{\lambda }\left( t\right) \right) \left[ \delta \left( t\right) - \psi \left( t\right) \right] dq^{*}_{\lambda }\left( t\right) \\&= \int _{\mathcal {B}} \phi ^{\prime }\left( q^{*}_{\lambda }\left( t\right) \right) \left[ \delta \left( t\right) - \psi \left( t\right) \right] dq^{*}_{\lambda }\left( t\right) . \\ \end{aligned} \end{aligned}$$

But, since \(\delta \) is affine on \(\mathcal {B}\), it follows from Eq. (B.6) that \(dq^{*}_{\lambda }\left( t\right) = 0\) on \(\mathcal {B}\). Consequently,

$$\begin{aligned} \int _{0}^{1} \left[ \delta \left( t\right) - \psi \left( t\right) \right] d \phi \circ q^{*}_{\lambda }\left( t\right) = 0. \end{aligned}$$

Therefore, applying Fubini’s theorem, as in the proof of Lemma B.5, gives

$$\begin{aligned} \begin{aligned} 0&=\int _{0}^{1} \left[ \delta \left( t\right) - \psi \left( t\right) \right] d \phi \circ q^{*}_{\lambda }\left( t\right) = \int _{0}^{1} \left( \phi \circ q^{*}_{\lambda }\right) ^{\prime }\left( t\right) \left[ \delta \left( t\right) - \psi \left( t\right) \right] dt \\&= - \int _{0}^{1} \left( \phi \circ q^{*}_{\lambda }\right) ^{\prime }\left( t\right) \left[ \left( \delta \left( 1\right) - \psi \left( 1\right) \right) - \left( \delta \left( t\right) - \psi \left( t\right) \right) \right] dt, \end{aligned} \end{aligned}$$

and hence

$$\begin{aligned} 0= & {} \int _{0}^{1} \left( \phi \circ q^{*}_{\lambda }\right) ^{\prime }\left( t\right) \left[ \int _{t}^{1}\left( \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right) dx\right] dt \\= & {} \int _{0}^{1} \int _{t}^{1} \left( \phi \circ q^{*}_{\lambda }\right) ^{\prime }\left( t\right) \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] dx \ dt\\= & {} \int _{0}^{1} \int _{0}^{x} \left( \phi \circ q^{*}_{\lambda }\right) ^{\prime } \left( t\right) \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] dt \ dx \\= & {} \int _{0}^{1} \left[ \int _{0}^{x} \left( \phi \circ q^{*}_{\lambda }\right) ^{\prime } \left( t\right) dt\right] \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] dx\\= & {} \int _{0}^{1} \left[ \left( \phi \circ q^{*}_{\lambda }\right) \left( x\right) - \left( \phi \circ q^{*}_{\lambda }\right) \left( 0\right) \right] \left[ \delta ^{\prime }\left( x\right) - \psi ^{\prime }\left( x\right) \right] dx \\= & {} \int _{0}^{1} \delta ^{\prime }\left( x\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( x\right) dx - \int _{0}^{1} \psi ^{\prime }\left( x\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( x\right) dx - \left( \phi \circ q^{*}_{\lambda }\right) \left( 0\right) \left[ \delta \left( 1\right) - \delta \left( 0\right) \right] \\&+ \left( \phi \circ q^{*}_{\lambda }\right) \left( 0\right) \left[ \psi \left( 1\right) - \psi \left( 0\right) \right] \\= & {} \int _{0}^{1} \delta ^{\prime }\left( x\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( x\right) dx - \int _{0}^{1} \psi ^{\prime }\left( x\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( x\right) dx - \left( \phi \circ q^{*}_{\lambda }\right) \left( 0\right) \left[ \psi \left( 1\right) - \psi \left( 0\right) \right] \\&+ \left( \phi \circ q^{*}_{\lambda }\right) \left( 0\right) \left[ \psi \left( 1\right) - \psi \left( 0\right) \right] \\= & {} \int _{0}^{1} \delta ^{\prime }\left( x\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( x\right) dx - \int _{0}^{1} \psi ^{\prime }\left( x\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( x\right) dx. \end{aligned}$$

Consequently, for each \(\lambda > 0\),

$$\begin{aligned} \int _{0}^{1} \delta ^{\prime }\left( t\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( t\right) dt = \int _{0}^{1} \psi ^{\prime }\left( t\right) \left( \phi \circ q^{*}_{\lambda }\right) \left( t\right) dt, \end{aligned}$$

and so, for all \(q \in \mathcal {Q}^{**}\),

$$\begin{aligned} \begin{aligned} \int _{0}^{1}\left[ q\left( t\right) - \lambda \psi ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) \right] dt&= \int _{0}^{1}q\left( t\right) dt - \lambda \int _{0}^{1} \psi ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) dt \\&\geqslant \int _{0}^{1}q\left( t\right) dt - \lambda \int _{0}^{1} \delta ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) dt\\&= \int _{0}^{1}\left[ q\left( t\right) dt - \lambda \delta ^{\prime }\left( t\right) \phi \left( q\left( t\right) \right) \right] dt \\&\geqslant \int _{0}^{1}\left[ q^{*}_{\lambda }\left( t\right) dt - \lambda \delta ^{\prime }\left( t\right) \phi \left( q^{*}_{\lambda }\left( t\right) \right) \right] dt \\&= \int _{0}^{1}\left[ q^{*}_{\lambda }\left( t\right) dt - \lambda \psi ^{\prime }\left( t\right) \phi \left( q^{*}_{\lambda }\left( t\right) \right) \right] dt. \end{aligned} \end{aligned}$$

Therefore, \(q^{*}_{\lambda ^{*}}\) is optimal for Problem (B.3). Lemma B.4 concludes the proof of Theorem 3.1. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ghossoub, M. A Neyman–Pearson problem with ambiguity and nonlinear pricing. Math Finan Econ 12, 365–385 (2018). https://doi.org/10.1007/s11579-017-0207-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11579-017-0207-y

Keywords

Mathematics Subject Classification

JEL Classification

Navigation