Skip to main content
Log in

Economic consequences of Nth-degree risk increases and Nth-degree risk attitudes

  • Published:
Journal of Risk and Uncertainty Aims and scope Submit manuscript

Abstract

We study comparative statics of Nth-degree risk increases within a large class of problems that involve bidimensional payoffs and additive or multiplicative risks. We establish necessary and sufficient conditions for unambiguous impact of Nth-degree risk increases on optimal decision making. We develop a simple and intuitive approach to interpret these conditions : novel notions of directional Nth-degree risk aversion that are characterized via preferences over lotteries.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Other work evaluating economic consequences of Nth-degree risk increases includes Baiardi and Menegatti’s (2011) analysis of the trade-off between (dirty) production and environmental-quality and Courbage and Rey’s (2012) analysis of the optimal allocation of a national health budget.

  2. It is more common to define Nth-degree risk increases in terms of a utility function u (α) as follows: 2 is an increase in Nth-degree risk over 1 if and only if E[u( 2)] < E[u( 1)] for all N times continuously differentiable real valued utility function u such that (−1) (N+1) u (N) > 0. Our equivalent characterization is somewhat more natural when dealing with the effect of risk increases on optimally chosen variables (i.e., on marginal utility).

  3. Ekern’s (1980) result, but not Lemma 1, also implies that 2 N 1 is necessary to have unanimity of ranking within the class of utility functions q such that (−1)N q (N) > 0, so the information provided by the two results is somewhat different. We thank an anonymous referee for pointing this out.

  4. Denuit et al.’s (1999) result is expressed in terms of maximal generators for s-convex orders and established for non-necessarily regular functions by the introduction of the concept of Nth degree divided difference. The authors remark that for regular functions the conditions on Nth-degree divided difference are equivalent to conditions on the Nth derivative. In Appendix A we propose a direct and simpler proof of the result when the functions are assumed to be regular.

  5. The canonical portfolio problem arises as a special case of this framework in which the attributes are perfect substitutes (setting U(y, z) = v(y + z) and redefining the variables), making it, in essence, a univariate problem. Assumption A2 is always satisfied in this setting.

  6. Similarly, in the context of problems with non-separable utility and an additive risk, Chiu and Eeckhoudt (2010) and Baiardi and Menegatti (2011) established the sufficiency of the condition in Proposition 1.

  7. For example, Rothschild and Stiglitz (1971), Keenan et al. (2006), Eeckhoudt and Schlesinger (2008), Chiu and Eeckhoudt (2010), Baiardi and Menegatti (2011), Chiu et al. (2012), and Courbage and Rey (2012).

  8. Note that the condition R N(y, z) ≤ N is coupled with ( − 1)N U (0, N)(y, z) ≥ 0 which, for N = 2, excludes risk averse consumers. In the special case that α = 0 this additional condition is no longer necessary. Therefore, assuming risk aversion, ( − 1)2 U (0, 2)(y, z) < 0, leads to a change in the direction of the inequality of the first condition, which becomes R 2(y, z) ≥ 2, as established by Rothschild and Stiglitz (1971) and Eeckhoudt and Schlesinger (2008) in the context of the 2-date saving problem with rate-of-return risk.

  9. Suppose, for example, that relative risk aversion is constant: v(z) = (1 − γ)− 1 z 1 − γ. Then, if γ < 1 an Nth-degree risk increase in the rate-of-return will decrease savings. For γ > 1, an Nth-degree risk increase in the rate-of-return will increase savings for some wealth levels and it will decrease savings for other wealth levels.

  10. Eeckhoudt et al. (2007) restrict themselves to the case x 1 = 0. All our results could be generalized further, along the lines of Eeckhoudt et al.’s (2009) results in a univariate framework, by considering the case in which x 1 is a first order stochastically dominated shift over x 2.

  11. Eeckhoudt et al. (2009) showed that if \(\widetilde {\alpha }_{2}\succcurlyeq _{N}\widetilde {\alpha }_{1}\) and \(\tilde {x}_{1}\succcurlyeq _{M}\tilde {x} _{2} \), then the 50-50 lottery [\(\widetilde {\alpha }_{2}+ \tilde {x}_{1}, \widetilde {\alpha }_{1}+ \tilde {x}_{2}\)] is an (N+M)th-degree risk increase over [\(\widetilde {\alpha }_{2}+ \tilde {x}_{2}, \widetilde { \alpha }_{1}+ \tilde {x}_{1}\)] in the sense of Ekern (1980), i.e., the second lottery is preferred by all decision makers with ( − 1)N + M U (0, N + M)(y, z) ≤ 0. When ρ y = 0 and ρ z > 0, and for a fixed value of y, our Proposition deals with the special case in which M = 1. Denuit et al. (2010a) proved a multivariate version of Eeckhoudt et al. (2009)’s results.

  12. The case with ρ y ≤ 0 and ρ z ≥ 0 is the relevant scenario to interpret the conditions found in Section 2.

References

  • Baiardi, D., & Menegatti, M. (2011). Pigouvian tax, abatement policies and uncertainty on the environment. Journal of Economics, 103(3), 221–251.

    Article  Google Scholar 

  • Block, M.K., & Heineke, J.M. (1973). The allocation of effort under uncertainty: the case of risk-averse behavior. Journal of Political Economy, 81, 376–385.

    Article  Google Scholar 

  • Chiu, W.H., & Eeckhoudt, L. (2010). The effects of stochastic wages and non-labor income on labor supply: update and extensions. Journal of Economics, 100, 69–83.

    Article  Google Scholar 

  • Chiu, W.H., Eeckhoudt, L., Rey, B. (2012). On relative and partial risk attitudes: theory and implications. Economic Theory, 50, 151–167.

    Article  Google Scholar 

  • Courbage, C., & Rey, B. (2012). Priority setting in health care and higher order degree change in risk. Journal of Health Economics, 31, 484–489.

    Article  Google Scholar 

  • Dardanoni, V. (1988). Optimal choices under uncertainty: the case of two-argument utility functions. Economic Journal, 98, 429–450.

    Article  Google Scholar 

  • Dardanoni, V., & Wagstaff, A. (1990). Uncertainty and the demand for medical care. Journal of Health Economics, 9, 23–38.

    Article  Google Scholar 

  • Denuit, M., De Vylder, E., Lefevre, C. (1999). Extremal generators and extremal distributions for the continuous s-convex stochastic orderings. Insurance: Mathematics and Economics, 24, 201–217.

    Article  Google Scholar 

  • Denuit, M., Eeckhoudt, L., Tsetlin, I., Winkler, R.L. (2010a). Multivariate concave and convex stochastic dominance. INSEAD working paper, 2010/29.

  • Denuit, M., Eeckhoudt, L., Rey, B. (2010b). Some consequences of correlation aversion in decision science. Annals of Operations Research, 176, 259–269.

    Article  Google Scholar 

  • Denuit, M., Eeckhoudt, L., Menegatti, M. (2011). Correlated risks, bivariate utility and optimal choices. Economic Theory, 46, 39–54.

    Article  Google Scholar 

  • Eeckhoudt, L., & Schlesinger, H. (2006). Putting risk in its proper place. American Economic Review, 96, 280–289.

    Article  Google Scholar 

  • Eeckhoudt, L., & Schlesinger, H. (2008). Changes in risk and the demand for saving. Journal of Monetary Economics, 55, 1329–1336.

    Article  Google Scholar 

  • Eeckhoudt, L., Schlesinger, H., Tselin, I. (2009). Apportioning of risks via stochastic dominance. Journal of Economic Theory, 144, 994–1003.

    Article  Google Scholar 

  • Eeckhoudt, L., Rey, B., Schlesinger, H. (2007). A good sign for multivariate risk taking. Management Science, 53, 117–124.

    Article  Google Scholar 

  • Ekern, S. (1980). Increasing Nth-degree risk. Economics Letters, 6, 329–333.

    Article  Google Scholar 

  • Gollier, C. (2001). The economics of risk and time. MIT Press.

  • Hadar, J., & Seo, T. K. (1990). The effects of shifts in a return distribution on optimal portfolios. International Economic Review, 31, 721–736.

    Article  Google Scholar 

  • Jean, W.H. (1980). The geometric mean and stochastic dominance. Journal of Finance, 35, 151–158.

    Article  Google Scholar 

  • Keenan, D.C., Kim, I., Warren, R.S. (2006). The private provision of public goods under uncertainty: a symmetric-equilibrium approach. Journal of Public Economic Theory, 8, 863–873.

    Article  Google Scholar 

  • Kimball, M. (1990). Precautionary saving in the small and in the large. Econometrica, 58, 58–73.

    Article  Google Scholar 

  • Leland, H.E. (1968). Saving and uncertainty: the precautionary demand for saving. Quarterly Journal of Economics, 82, 465–73.

    Article  Google Scholar 

  • Menezes, C., Geiss, C., Tressler, J. (1980). Increasing downside risk. American Economic Review, 70, 921–932.

    Google Scholar 

  • Menezes, C., & Wang, X.H. (2005). Increasing outer risk. Journal of Mathematical Economics, 41, 875–886.

    Article  Google Scholar 

  • Rothschild, M., & Stiglitz, J. (1970). Increasing risk I: a definition. Journal of Economic Theory, 2, 225–243.

    Article  Google Scholar 

  • Rothschild, M., & Stiglitz, J. (1971). Increasing risk II: its economic consequences. Journal of Economic Theory, 3, 66–84.

    Article  Google Scholar 

  • Sandler, T., Sterbenz, F.P., Posnett, J. (1987). Free riding and uncertainty. European Economic Review, 31, 1605–1617.

    Article  Google Scholar 

  • Sandmo, A. (1970). The effect of uncertainty on saving decisions. Review of Economic Studies, 37, 353–60.

    Article  Google Scholar 

  • Tressler, J.H., & Menezes, C.F. (1980). Labor supply and wage rate uncertainty. Journal of Economic Theory, 23, 425–436.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Diego Nocetti.

Additional information

The authors thank the participants of the conference “Risk and choice: A conference in honor of Louis Eeckhoudt” (Toulouse, July, 2012) as well as those of EGRIE conference (Palma, September, 2012) and of the Risk Seminar (Paris-Dauphine, January 2013) and in particular Christophe Courbage, Harris Schlesinger and Bertrand Villeneuve. Special thanks to Louis Eeckhoudt who discussed the paper and provided very useful hints and comments as well as to two anonymous referees who permitted us to greatly improve the paper in particular by bringing our attention to the possibility to generalize the increasing risk results to a stochastic dominance framework. The first two authors gratefully acknowledge the financial support of the ANR (RISK project) as well as of the Risk Foundation (Groupama Chair).

Appendices

Appendix A: Increases in Nth-degree risk

Proof of Lemma 1

The fact that 2. implies 1. results directly from Ekern (1980). For the sake of completeness we rederive it. Let \((\widetilde {\alpha }_{1},\widetilde { \alpha }_{2})\) be such that \(\widetilde {\alpha }_{2}\succcurlyeq _{N} \widetilde {\alpha }_{1}.\) We have

$$\begin{array}{@{}rcl@{}} E[ q(\widetilde{\alpha }_{2})] -E\left[ q(\widetilde{\alpha }_{1}) \right] &=&\int_{0}^{B}q(x)dF_{\widetilde{\alpha }_{2}}(x)- \int_{0}^{B}q(x)dF_{\widetilde{\alpha }_{1}}(x) \\ &=&\sum\limits_{k=1}^{N}(-1)^{k-1}q^{(k-1)}(B)\left[ F_{\widetilde{\alpha }_{2}}^{[ k] }(B)-F_{\widetilde{\alpha }_{1}}^{[ k] }(B)\right] \\ &&+\int_{0}^{B}(-1)^{N}q^{(N)}(x)\left[ F_{\widetilde{\alpha }_{2}}^{\left[ N \right] }(x)-F_{\widetilde{\alpha }_{1}}^{[ N] }(x)\right] dx. \end{array} $$

Since q and its N first derivatives are continuous on \(\mathbb {R}_{+}\), they are bounded on all compact subsets and all our integrals are well defined. □

By definition, all the terms in the sum are equal to 0 and \(F_{\widetilde { \alpha }_{2}}^{[ N] }(x)-F_{\widetilde {\alpha }_{1}}^{\left [ N \right ] }(x)\geq 0\) on [0, B]. By 2., the integral is then nonnegative and \(E[ q(\widetilde {\alpha }_{2})] \geq E\left [ q( \widetilde {\alpha }_{1})\right ] \).

Let us now prove that 1. implies 2. Let q be an N times continuously differentiable function on \(\mathbb {R}_{+}\) such that \(E\left [ q(\widetilde { \alpha }_{2})\right ] \geq E[ q(\widetilde {\alpha }_{1})] \) for all pair \((\widetilde {\alpha }_{1},\widetilde {\alpha }_{2})\) such that \( \widetilde {\alpha }_{2}\succcurlyeq _{N}\widetilde {\alpha }_{1}.\) Let, if it exists, α be a nonnegative real number such that q (N)(α) ≠ 0 and let ε denote a positive real number such that q (N)(t) ≠ 0 for t ∈ [α, α + ε]. The sign of q (N)remains then constant on [α, α + ε]. Let \(\widetilde { \beta }_{1}\) and \(\widetilde {\beta }_{2}\) be two nonnegative bounded above random variables such that \(\widetilde {\beta }_{2}\succcurlyeq _{N} \widetilde {\beta }_{1}.\) Let B be a common upper bound for \(\widetilde { \beta }_{1}\) and \(\widetilde {\beta }_{2}\) and let \(\widetilde {\alpha }_{1}= \frac {\varepsilon }{B}\widetilde {\beta }_{1}+\alpha \) and \(\widetilde {\alpha }_{2}=\frac {\varepsilon }{B}\widetilde {\beta }_{2}+\alpha .\) The random variables \(\widetilde {\alpha }_{1}\) and \(\widetilde {\alpha }_{2}\) take their values in [α, α + ε] and it is easy to check that \(F_{\widetilde {\alpha }_{i}}^{[ k] }( t) =\left ( \frac {\varepsilon }{B}\right )^{k-1}F_{\widetilde {\beta }_{i}}^{[ k] }\left ( \frac {t-\alpha }{\varepsilon }B\right ) \) for k = 1, 2. . . and i = 1, 2. Therefore, \(\widetilde {\alpha }_{2}\succcurlyeq _{N} \widetilde {\alpha }_{1}\) which implies, by 1., that \(E\left [ q(\widetilde { \alpha }_{2})\right ] \geq E[ q(\widetilde {\alpha }_{1})] .\) We have

$$\begin{array}{@{}rcl@{}} E[ q(\widetilde{\alpha }_{2})] -E\left[ q(\widetilde{\alpha }_{1}) \right] &=&\int_{\alpha }^{\alpha +\varepsilon }q(x)dF_{\widetilde{\alpha } _{2}}(x)-\int_{\alpha }^{\alpha +\varepsilon }q(x)dF_{\widetilde{\alpha } _{1}}(x) \\ &=&\sum\limits_{k=1}^{N}(-1)^{k-1}q^{(k-1)}(\alpha +\varepsilon )\left[ F_{ \widetilde{\alpha }_{2}}^{[ k] }(\alpha +\varepsilon )-F_{ \widetilde{\alpha }_{1}}^{[ k] }(\alpha +\varepsilon )\right] \\ &&+\int_{\alpha }^{\alpha +\varepsilon }(-1)^{N}q^{(N)}(t)\left[ F_{ \widetilde{\alpha }_{2}}^{[ N] }(t)-F_{\widetilde{\alpha }_{1}}^{[ N] }(t)\right] dt. \end{array} $$

By definition, we have \(F_{\widetilde {\alpha }_{2}}^{[ k] }(\alpha +\varepsilon )=F_{\widetilde {\alpha }_{1}}^{[ k] }(\alpha +\varepsilon )\) for k = 1, . . . , N and \(F_{\widetilde {\alpha }_{2}}^{[ N] }(t)-F_{\widetilde {\alpha }_{1}}^{[ N] }(t)\geq 0\) and the inequality is strict for some t in [α, α + ε], and even on a neighborhood of t by continuity of \( F_{\widetilde {\alpha }_{1}}^{[ N] }\) and \(F_{\widetilde {\alpha } _{2}}^{[ N] }.\) Since the sign of q (N)remains constant on [α, α + ε], this gives that ( − 1)N q (N)(t) > 0 on α, α + ε and ( − 1)N q (N)(x) > 0 for all x such that q (N)(x) ≠ 0 which completes the proof.

1.1 Proof of Proposition 1

Let us prove that 2. implies 1. Let \(\left ( \widetilde {\alpha }_{1}, \widetilde {\alpha }_{2}\right ) \) be such that \(\widetilde {\alpha } _{2}\succcurlyeq _{N}\widetilde {\alpha }_{1}\) and let \(q(\alpha )=g(x_{1}^{\ast },\alpha ,\mu ).\) We have \(\frac {\partial ^{N}g}{\partial \alpha ^{N}}(x_{1}^{\ast },\alpha ,\mu )=-pU^{( 1,N) }\left ( K-x_{1}^{\ast }p,x_{1}^{\ast }\mu +\alpha \right ) +\mu U^{\left ( 0,N+1\right ) }( K-x_{1}^{\ast }p,x_{1}^{\ast }\mu +\alpha ) .\) By 2., we have \(( -1)^{N}\frac {\partial ^{N}g}{\partial \alpha ^{N}} (x_{1}^{\ast },\alpha ,\mu )\geq 0\) and ( − 1)N q (N)(α) ≥ 0. Since \(\widetilde {\alpha }_{i}\) is bounded above ( i = 1, 2), \( x_{1}^{\ast }\mu +\widetilde {\alpha }_{i}\) is bounded and bounded away from zero and U and its N + 1 first derivatives are bounded on the convex hull of the set of values taken by \(\left ( K-x_{1}^{\ast }p,x_{1}^{\ast }\mu + \widetilde {\alpha }_{i}\right ) \). The same applies to q on the convex hull of the set of values taken by \(\widetilde {\alpha }_{i}\), i = 1, 2. By Lemma 1, this leads to \(E[ q(\widetilde {\alpha }_{1})] \leq E [ q(\widetilde {\alpha }_{2})] .\) By definition, we have \(E\left [ q(\widetilde {\alpha }_{1})\right ] =0,\) which gives \(E\left [ q(\widetilde { \alpha }_{2})\right ] \geq 0\) or \(E\left [ g(x_{1}^{\ast },\widetilde {\alpha } _{2},\mu )\right ] \geq 0.\) By concavity of U, it is easy to check that \(E [ g(x,\widetilde {\alpha }_{2},\mu )] \) is a decreasing function of x. Since \(x_{2}^{\ast }\) is characterized by \(E\left [ g(x_{2}^{\ast }, \widetilde {\alpha }_{2},\mu )\right ] =0,\) we obtain that \(x_{2}^{\ast }\geq x_{1}^{\ast }.\)

Let us prove 1. implies 2. As in the proof of Lemma 1, we consider \(\widetilde {\beta }_{1}\) and \(\widetilde {\beta }_{2}\) two nonnegative random variables with a common upper bound B such that \(\widetilde {\beta } _{2}\succcurlyeq _{N}\widetilde {\beta }_{1}.\) As above, we introduce the random variables \(\widetilde {\alpha }_{1,\varepsilon }=\frac {\varepsilon }{B} \widetilde {\beta }_{1}+\alpha ^{\ast }\) and \(\widetilde {\alpha } _{2,\varepsilon }=\frac {\varepsilon }{B}\widetilde {\beta }_{2}+\alpha ^{\ast }\) for some α > 0 and some ε > 0. The random variables \(\widetilde {\alpha }_{1,\varepsilon }\) and \(\widetilde {\alpha } _{2,\varepsilon }\) take their values in [α , α + ε] and \(\widetilde {\alpha }_{2,\varepsilon }\succcurlyeq _{N}\widetilde {\alpha }_{1,\varepsilon }.\) Let us consider a given real number γ and let us define the random variables \( \widetilde {\gamma }_{i,\varepsilon }\), i = 1, 2, as lotteries giving \( \widetilde {\alpha }_{i,\varepsilon }\) with probability ε and γ with probability 1 − ε. The constant γ provides an additional degree of freedom that will prove useful in order to control the pair \(( K-x_{1}^{\ast }p,x_{1}^{\ast }\mu +\alpha ) .\) We have \(\widetilde {\gamma }_{2,\varepsilon }\succcurlyeq _{N}\widetilde {\gamma }_{1,\varepsilon }\) as an immediate consequence of the stability of this order relation under probability mixtures. Let \(x_{1,\varepsilon }^{\ast }\) and \(x_{2,\varepsilon }^{\ast }\) be respectively the solutions of \(P_{U,K,p}( \widetilde {\gamma }_{1,\varepsilon },\mu )\) and \(P_{U,K,p}(\widetilde {\gamma }_{2,\varepsilon },\mu ).\) By 1., we have \(x_{2,\varepsilon }^{\ast }\geq x_{1,\varepsilon }^{\ast }.\) By definition, we have \(E\left [ g(x_{1,\varepsilon }^{\ast },\widetilde {\gamma }_{1,\varepsilon },\mu ) \right ] =0\) and \(E\left [ g(x_{2,\varepsilon }^{\ast },\widetilde {\gamma } _{2,\varepsilon },\mu )\right ] =0.\) By concavity of U, g is a decreasing function of x and we have \(E\left [ g(x_{1,\varepsilon }^{\ast },\widetilde { \gamma }_{2,\varepsilon },\mu )\right ] \geq 0.\) Let q be defined by \( q(\alpha )=g(x_{1,\varepsilon }^{\ast },\alpha ,\mu ).\) We have

$$\begin{array}{@{}rcl@{}} E[ q(\widetilde{\gamma }_{2,\varepsilon })] -E\left[ q(\widetilde{ \gamma }_{1,\varepsilon })\right] &=&\left( E\left[ q(\widetilde{\alpha } _{2,\varepsilon })\right] -E\left[ q(\widetilde{\alpha }_{1,\varepsilon }) \right] \right) \varepsilon \\ &=&\sum_{k=1}^{N}(-1)^{k-1}q^{(k-1)}(\alpha^{\ast }+\varepsilon ) \notag \\ &&\times\left[ F_{ \widetilde{\alpha }_{2},P}^{[ k] }(\alpha^{\ast }+\varepsilon )-F_{\widetilde{\alpha }_{1},P}^{[k] }(\alpha^{\ast }+\varepsilon )\right] \varepsilon \\ &&+\varepsilon \int_{\alpha^{\ast }}^{\alpha^{\ast }+\varepsilon }(-1)^{N}q^{(N)}(t)\left[ F_{\widetilde{\alpha }_{2},P}^{\left[ N\right] }(t)-F_{\widetilde{\alpha }_{1},P}^{[ N] }(t)\right] dt. \end{array} $$

By construction, the left side of the equality is nonnegative, all the terms in the sum are equal to zero and \(F_{\widetilde {\alpha }_{2},P}^{\left [ N \right ] }(t)-F_{\widetilde {\alpha }_{1},P}^{[ N] }(t)\) is nonnegative and nonzero. Note that \(\widetilde {\gamma }_{i,\varepsilon }\) is bounded and bounded away from zero for i = 1, 2. Therefore q and its first N derivatives are bounded on the convex hull of the set of values taken by \( \widetilde {\gamma }_{i,\varepsilon }\), i = 1, 2. Therefore, ( − 1)N q (N)(t) is nonnegative at least on a given subinterval of α , α + ε. Let \(\alpha _{\varepsilon }^{\ast }\ \)be in α , α + ε such that \((-1)^{N}q^{(N)}(\alpha _{\varepsilon }^{\ast })\geq 0.\) We have then

$$ (-1)^{N+1}pU^{( 1,N) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\mu +\alpha_{\varepsilon }^{\ast }\right) +(-1)^{N}\mu U^{( 0,N+1) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\mu +\alpha_{\varepsilon }^{\ast }\right) \geq 0 $$
(6)

where \(x_{1,\varepsilon }^{\ast }\) satisfies

$$E\left[ -pU^{( 1,0) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\mu +\widetilde{\gamma }_{1,\varepsilon }\right) +\mu U^{( 0,1) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\mu +\widetilde{\gamma }_{1,\varepsilon }\right) \right] =0 $$

or

$$ \varepsilon E\left[ g(x_{1,\varepsilon }^{\ast },\frac{\varepsilon }{B} \widetilde{\beta }_{1}+\alpha^{\ast },\mu )\right] +(1-\varepsilon )g(x_{1,\varepsilon }^{\ast },\gamma ,\mu )=0. $$
(7)

Remark that, until now, α , K and γ have been arbitrarily chosen. Let us now choose them carefully in order to derive our result. Let (Y, Z) be arbitrary in \(( \mathbb {R}_{+}^{\ast })^{2}\). By our Inada condition, \(\lim _{z\rightarrow 0}\frac {U^{\left ( 0,1\right ) }( Y,z) }{U^{( 1,0) }( Y,z) } =\infty \) and \(\lim _{z\rightarrow \infty }\frac {U^{( 0,1) }\left ( Y,z\right ) }{U^{( 1,0) }( Y,z) }=0\) which gives that there exists some z > 0 such that \(\frac {U^{( 0,1) }( Y,z^{\ast }) }{U^{( 1,0) }( Y,z^{\ast }) }=\frac {p}{\mu }\) or − p U (1, 0)(Y, z ∗) + μ U (0, 1)(Y, z ∗) = 0. We choose x > 0 such that μ x < inf(Z, z ∗) and α and K are taken such that α = Zμ x > 0 and K = Y + p x . Let γ be given by γ = z μ x > 0. We have

$$\begin{array}{@{}rcl@{}} g(x^{\ast },\widetilde{\gamma },\mu ) &=&-pU^{( 1,0) }\left( K-x^{\ast }p,x^{\ast }\mu +\gamma \right) +\mu U^{( 0,1) }\left( K-x^{\ast }p,x^{\ast }\mu +\gamma \right) \\ &=&-pU^{( 1,0) }( Y,z^{\ast }) +\mu U^{\left( 0,1\right) }( Y,z^{\ast }) =0. \end{array} $$

The solution of Eq. 7 for ε = 0 is then given by x . In a well chosen neighborhood of x , xKx p and \(x\longmapsto x\mu +\frac {\varepsilon }{B}\widetilde {\beta }_{1}+\alpha ^{\ast }\) are bounded and bounded away from 0. The functions U 1, 0 and U (0, 1) being continuously differentiable, the function \((x,\varepsilon )\longmapsto \varepsilon E\left [ g(x,\frac {\varepsilon }{B}\widetilde {\beta }_{1}+\alpha ^{\ast },\mu )\right ] +(1-\varepsilon )g(x,\widetilde {\gamma },\mu )\) is then differentiable with respect to x at (x , 0). Furthermore, the derivative of this last function with respect to x at (x , 0) is nonzero (concavity of U ). The solution \(x_{1,\varepsilon }^{\ast }\) of Eq. 7 is then continuous with respect to ε in a neighborhood of 0 which gives \(\lim _{\varepsilon \rightarrow 0}x_{1,\varepsilon }^{\ast }=x^{\ast }.\) Furthermore, we clearly have \(\lim _{\varepsilon \rightarrow 0}\alpha _{\varepsilon }^{\ast }=\alpha ^{\ast }.\) Taking the limit in Eq. 6 when ε tends to 0, we obtain

$$(-1)^{N+1}pU^{( 1,N) }\left(K\,-\,x^{\ast }p,x^{\ast }\mu \,+\,\alpha^{\ast }\right) +(-1)^{N}\mu U^{( 0,N+1) }\left( K\,-\,x^{\ast}p,x^{\ast }\mu +\alpha^{\ast }\right) \geq 0 $$

or, by construction

$$(-1)^{N+1}pU^{( 1,N) }( Y,Z) +(-1)^{N}\mu U^{\left( 0,N+1\right) }( Y,Z) \geq 0 $$

1.2 Proof of Proposition 2

Let us prove that 2. implies 1. Let \(\left (\widetilde {\mu }_{1},\widetilde { \mu }_{2}\right ) \) be such that \(\widetilde {\mu }_{2}\succcurlyeq _{N} \widetilde {\mu }_{1}\) and let \(q(\mu )=g(x_{1}^{\ast },\alpha ,\mu ).\) We have \(q^{(N)}(\mu )=-p(x_{1}^{\ast })^{N}U^{\left ( 1,N\right )}(K-x_{1}^{\ast }p,x_{1}^{\ast }\mu +\alpha )+\mu (x_{1}^{\ast })^{N}U^{(0,N+1)}\left (K-x_{1}^{\ast }p,x_{1}^{\ast }\mu +\alpha \right )+N(x_{1}^{\ast })^{N-1}U^{(0,N)}\left (K-x_{1}^{\ast }p,x_{1}^{\ast }\mu +\alpha \right ) \). Since we assumed that \(x_{1}^{\ast }\geq 0\), by 2, we have

\((-1)^{N}p(x_{1}^{\ast })^{N}U^{(1,N) }(K-x_{1}^{\ast }p,x_{1}^{\ast }\mu +\alpha )\leq 0\) and

$$\begin{array}{@{}rcl@{}} &&(-1)^{N}\left(x_{1}^{\ast}\mu U^{(0,N+1)}(K-x_{1}^{\ast}p,x_{1}^{\ast}\mu+\alpha)+NU^{\left( 0,N\right)}(K-x_{1}^{\ast}p,x_{1}^{\ast}\mu+\alpha)\right)\\ &=&(-1)^{N}\frac{x_{1}^{\ast}\mu}{x_{1}^{\ast}\mu+\alpha} \left(\left( x_{1}^{\ast }\mu +\alpha\right)U^{(0,N+1)}\left(K-x_{1}^{\ast }p,x_{1}^{\ast}\mu +\alpha\right)\right.\\ &&\left.\qquad\;\;\;\,\,\qquad\qquad+NU^{\left( 0,N\right)}\left( K-x_{1}^{\ast}p,x_{1}^{\ast}\mu +\alpha\right)\right)\\ &&+(-1)^{N}\frac{\alpha N}{x_{1}^{\ast}\mu +\alpha}U^{\left( 0,N\right)}(K-x_{1}^{\ast }p,x_{1}^{\ast}\mu +\alpha) \\ &\geq &0 \end{array} $$

which gives ( − 1)N q (N)(μ) ≥ 0. Since \( \widetilde {\mu }_{i}\) is bounded above ( i = 1, 2), \(x_{1}^{\ast }\widetilde { \mu }_{i}+\alpha \) is bounded and bounded away from zero and U and its N + 1 first derivatives are bounded on the convex hull of the set of values taken by \(\left ( K-x_{1}^{\ast }p,x_{1}^{\ast }\widetilde {\mu }_{i}+\alpha \right ) \). The same applies to q on the convex hull of the set of values taken by \(\widetilde {\mu }_{i}\), i = 1, 2. By Lemma 1, this leads to \(E[ q(\widetilde {\mu }_{1})] \leq E\left [ q(\widetilde {\mu } _{2})\right ] .\) By definition, we have \(E\left [ q(\widetilde {\mu }_{1}) \right ] =0\) which gives \(E[ q(\widetilde {\mu }_{2})] =E\left [ g(x_{1}^{\ast },\alpha ,\widetilde {\mu }_{2})\right ] \geq 0.\) By concavity of U, it is easy to check that g(x, α, μ) is a decreasing function of x. From there we derive that \(x_{2}^{\ast }\geq x_{1}^{\ast }.\)

Let us prove that 1. implies 2. As in the proof of Lemma 1 and Proposition 1, we consider \(\widetilde {\beta }_{1}\) and \( \widetilde {\beta }_{2}\) two nonnegative random variables with a common upper bound B such that \(\widetilde {\beta }_{2}\succcurlyeq _{N}\widetilde { \beta }_{1}.\) As above, we introduce the random variables \(\widetilde {\mu } _{1,\varepsilon }=\frac {\varepsilon }{B}\widetilde {\beta }_{1}+\mu ^{\ast }\) and \(\widetilde {\mu }_{2,\varepsilon }=\frac {\varepsilon }{B}\widetilde { \beta }_{2}+\mu ^{\ast }\) for some μ > 0 and some ε > 0. The random variables \(\widetilde {\mu }_{1,\varepsilon }\) and \( \widetilde {\mu }_{2,\varepsilon }\) take their values in [μ , μ + ε] and \(\widetilde {\mu }_{2,\varepsilon }\succcurlyeq _{N}\widetilde {\mu }_{1,\varepsilon }.\) Let us consider a given γ and let us define the random variables \(\widetilde {\gamma } _{i,\varepsilon }\), i = 1, 2, as a lottery that takes the value \(\widetilde { \mu }_{i,\varepsilon }(\omega )\) with probability ε and the value γ with probability 1 − ε. As previously, we have \( \widetilde {\gamma }_{2,\varepsilon }\succcurlyeq _{N}\widetilde {\gamma } _{1,\varepsilon }.\) Let \(x_{1,\varepsilon }^{\ast }\) and \(x_{2,\varepsilon }^{\ast }\) be respectively the solutions of \(P_{U,K,p}(\alpha ,\widetilde { \gamma }_{1,\varepsilon })\) and \(P_{U,K,p}(\alpha ,\widetilde {\gamma } _{2,\varepsilon }).\) If \(x_{1,\varepsilon }^{\ast }\) is nonnegative then, by 1., we have \(x_{2,\varepsilon }^{\ast }\geq x_{1,\varepsilon }^{\ast }.\) By definition, we have \(E\left [ g(x_{1,\varepsilon }^{\ast },\alpha ,\widetilde { \gamma }_{1,\varepsilon })\right ] =0\) and \(E\left [ g(x_{2,\varepsilon }^{\ast },\alpha ,\widetilde {\gamma }_{2,\varepsilon })\right ] =0.\) By concavity of U, g is a decreasing function of x and we have \(E\left [ g(x_{1,\varepsilon }^{\ast },\alpha ,\widetilde {\gamma }_{2,\varepsilon }) \right ] \geq 0.\) Let q be defined by \(q(\mu )=g(x_{1,\varepsilon }^{\ast },\alpha ,\mu ).\) We have

$$\begin{array}{@{}rcl@{}} E[ q(\widetilde{\gamma }_{2,\varepsilon })] -E\left[ q(\widetilde{ \gamma }_{1,\varepsilon })\right] &=&\left( E\left[ q(\widetilde{\mu } _{2,\varepsilon })\right] -E\left[ q(\widetilde{\mu }_{1,\varepsilon }) \right] \right) \varepsilon \\ &=&\sum_{k=1}^{N}(-1)^{k-1}q^{(k-1)}(\mu^{\ast }+\varepsilon )\\ &&\times\left[ F_{ \widetilde{\mu }_{2},P}^{[ k] }(\mu^{\ast }+\varepsilon )-F_{ \widetilde{\mu }_{1},P}^{[ k] }(\mu^{\ast }+\varepsilon )\right] \varepsilon \\ &&+\varepsilon \int_{\mu^{\ast }}^{\mu^{\ast }+\varepsilon }(-1)^{N}q^{(N)}(t)\left[ F_{\widetilde{\mu }_{2},P}^{\left[ N\right] }(t)-F_{\widetilde{\mu }_{1},P}^{[ N] }(t)\right] dt. \end{array} $$

By construction, the left side of the equality is nonnegative, all the terms in the sum are equal to zero and \(F_{\widetilde {\mu }_{2},P}^{[ N] }(t)-F_{\widetilde {\mu }_{1},P}^{[ N] }(t)\) is nonnegative and nonzero. Note that \(\widetilde {\gamma }_{i,\varepsilon }\) is bounded and bounded away from zero for i = 1, 2. Therefore q and its first N derivatives are bounded on the convex hull of the set of values taken by \( \widetilde {\gamma }_{i,\varepsilon }\), i = 1, 2. Therefore, ( − 1)N q (N) is nonnegative at least on a given subinterval of μ , μ + ε. Let \(\mu _{\varepsilon }^{\ast }\ \)be in μ , μ + ε such that \( (-1)^{N}q^{(N)}(\mu _{\varepsilon }^{\ast })\geq 0.\) We have then

$$\begin{array}{@{}rcl@{}} &&(-1)^{N+1}p( x_{1,\varepsilon }^{\ast })^{N}U^{\left( 1,N\right) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\mu_{\varepsilon }^{\ast }+\alpha \right) \\ &&+(-1)^{N}\mu_{\varepsilon }^{\ast }\left( x_{1,\varepsilon }^{\ast }\right)^{N}U^{( 0,N+1) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\mu_{\varepsilon }^{\ast }+\alpha \right) \\ &&+(-1)^{N}N( x_{1,\varepsilon }^{\ast })^{N-1}U^{\left( 0,N\right) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\mu_{\varepsilon }^{\ast }+\alpha \right) \\ &\geq &0 \end{array} $$
(8)

where \(x_{1,\varepsilon }^{\ast }\) satisfies

$$E\left[-pU^{( 1,0) }\left( K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast }\widetilde{\gamma }_{1,\varepsilon }+\alpha \right) +\widetilde{\gamma }_{1,\varepsilon }U^{( 0,1) }\left(K-x_{1,\varepsilon }^{\ast }p,x_{1,\varepsilon }^{\ast}\widetilde{\gamma} _{1,\varepsilon }+\alpha \right)\right]\,=\,0 $$

or

$$ \varepsilon E\left[ g\left(x_{1,\varepsilon }^{\ast },\alpha ,\frac{\varepsilon }{ B}\widetilde{\beta }_{1}+\mu^{\ast }\right)\right] +(1-\varepsilon )g(x_{1,\varepsilon }^{\ast },\alpha ,\gamma )=0. $$
(9)

Remark that, until now p, μ , K, α and γ have been arbitrarily chosen. Let us now choose them carefully in order to derive our result. We assume first that γ is equal to 1. Let x > 0 be given and let (M, Y, Z) in \(( \mathbb {R}_{+})^{3} \) such that M < Z. We take \(p=\frac {U^{( 0,1) }(Y,x^{\ast }+Z-M)}{U^{( 1,0) }(Y,x^{\ast }+Z-M)}\), K = Y + p x , α = ZM and \(\mu ^{\ast }=\frac {M}{x^{\ast }}.\) By construction, we have α > 0 and we have

$$\begin{array}{@{}rcl@{}} g(x^{\ast },\alpha ,\gamma ) &=&-pU^{( 1,0) }\left( K-x^{\ast }p,x^{\ast }\gamma +\alpha \right) +\widetilde{\gamma }U^{( 0,1) }( K-x^{\ast }p,x^{\ast }\gamma +\alpha ) \\ &=&-pU^{( 1,0) }( K-x^{\ast }p,x^{\ast }+\alpha ) +U^{( 0,1) }( K-x^{\ast }p,x^{\ast }+\alpha ) \\ &=&-pU^{( 1,0) }( Y,x^{\ast }+Z-M) +U^{\left( 0,1\right) }( Y,x^{\ast }+Z-M) \\ &=&0. \end{array} $$

The solution of Eq. 9 for ε = 0 is then given by x . In a well chosen neighborhood of x , xKx p and \(x\longmapsto x\left (\mu ^{\ast }+\frac {\varepsilon }{B}\widetilde { \beta }_{1}\right ) +\alpha \) are bounded and bounded away from 0. The functions U (1, 0) and U (0, 1) being continuously differentiable, the function \((x,\varepsilon )\longmapsto \varepsilon E\left [ g(x,\alpha ,\frac {\varepsilon }{B}\widetilde {\beta }_{1}+\mu ^{\ast })\right ] +(1-\varepsilon )g(x,\alpha ,\gamma )\) is then differentiable with respect to x at (x , 0). Furthermore, the derivative with respect to x at (x , 0) is nonzero. The solution \( x_{1,\varepsilon }^{\ast }\) of Eq. 9 is then continuous with respect to ε in a neighborhood of 0 which gives \( \lim _{\varepsilon \rightarrow 0}x_{1,\varepsilon }^{\ast }=x^{\ast }\) and guarantees that \(x_{1,\varepsilon }^{\ast }>0\) for ε small enough . Since we clearly have \(\lim _{\varepsilon \rightarrow 0}\mu _{\varepsilon }^{\ast }=\mu ^{\ast },\) taking the limit in Eq. 8 when ε tends to 0, we obtain

$$(-1)^{N}(x^{\ast})^{N-1}\left(-px^{\ast}U^{(1,N)}(Y,Z)+\mu^{\ast}x^{\ast}U^{(0,N+1)}(Y,Z)+NU^{(0,N)}(Y,Z)\right)\geq 0 $$

or

$$ (-1)^{N}\left(-px^{\ast}U^{(1,N)}(Y,Z) +MU^{(0,N+1)}(Y,Z)+NU^{(0,N)}(Y,Z)\right) \geq 0 $$
(10)

This result being true for all (Y, Z) in \(\left (\mathbb {R}_{+}^{\ast }\right )^{2},\) all M ∈ (0, Z) all x > 0 and for \(p=\frac {U^{(0,1)}(Y,x^{\ast }+Z-M)}{U^{(1,0)}(Y,x^{\ast }+Z-M)}.\) When x goes to 0, Y, M and Z being fixed, \(x^{\ast }\frac {U^{(0,1)}\left (Y,x^{\ast }+Z-M\right )}{U^{( 1,0) }\left (Y,x^{\ast }+Z-M\right )}\) goes to 0 and we obtain that

$$(-1)^{N}\left(MU^{(0,N+1)}(Y,Z)+NU^{(0,N)}(Y,Z)\right) \geq 0 $$

for all (Y, Z) in \(( \mathbb {R}_{+}^{\ast })^{2}\) and all M ∈ (0, Z) or equivalently

$$\begin{array}{@{}rcl@{}} (-1)^{N}&&\left(ZU^{(0,N+1)}(Y,Z)+NU^{(0,N)}(Y,Z)\right)\geq 0\\ &&\text{and}\,\,(-1)^{N}U^{(0,N)}(Y,Z)\geq 0\,\,\text{for all}\,\,(Y,Z)\,\,\text{in}\,\,\left(\mathbb{R}_{+}^{\ast}\right)^{2}. \end{array} $$

We assumed that \(z\frac {U^{( 0,1) }(Y,z)}{U^{( 1,0)}(Y,z)}\) is unbounded. We then have either \(\lim _{z\rightarrow 0}z\frac { U^{( 0,1) }(Y,z)}{U^{( 1,0) }(Y,z)}=\infty \) or \( \lim _{z\rightarrow \infty }z\frac {U^{( 0,1) }(Y,z)}{U^{\left ( 1,0\right ) }(Y,z)}=\infty \). If \(\lim _{z\rightarrow 0}z\frac {U^{\left ( 0,1\right ) }(Y,z)}{U^{( 1,0) }(Y,z)}=\infty \), it suffices to take \(x^{\ast }=\frac {\zeta }{2}\) and \(M=Z-\frac {\zeta }{2}\) for ζ arbitrarily small to make the quantity \(px^{\ast }=\frac {U^{\left ( 0,1\right ) }(Y,x^{\ast }+Z-M)}{U^{( 1,0) }(Y,x^{\ast }+Z-M)} x^{\ast }=\frac {1}{2}\zeta \frac {U^{( 0,1) }(Y,\zeta )}{U^{\left ( 1,0\right ) }(Y,\zeta )}\) arbitrarily large. Since M, Y and Z are bounded, Eq. 10 gives then ( − 1)N U (1, N)(Y, Z) ≤ 0. If \(\lim _{z\rightarrow \infty }z\frac {U^{\left ( 0,1\right ) }(Y,z)}{U^{( 1,0) }(Y,z)}=\infty ,\) it suffices to take x sufficiently large to make the quantity x + ZM sufficiently large and \(\frac {U^{( 0,1) }(Y,x^{\ast }+Z-M)}{ U^{( 1,0) }(Y,x^{\ast }+Z-M)}( x^{\ast }+Z-M) \) arbitrarily large. Since Z is kept fixed, \(\frac {x^{\ast }}{x^{\ast }+Z-M}\) is arbitrarily close to 1 and p x arbitrarily large. Since M, Y and Z are bounded, Eq. 10 gives then ( − 1)N U (1, N)(Y, Z) ≤ 0.The Proof of Corollary 1 follows directly by reversing the inequalities above.

1.3 Proof of Proposition 3

In an expected utility framework, Nth-degree risk aversion in the direction of (ρ y , ρ z ) is equivalent to

$$\begin{array}{@{}rcl@{}} &&\frac{1}{2}E\left[ U\left( y+x_{2}\rho_{y},z+x_{2}\rho_{z}+\widetilde{ \alpha }_{2}\right) \right] -\frac{1}{2}E\left[ U\left( y+x_{2}\rho_{y},z+x_{2}\rho_{z}+\widetilde{\alpha }_{1}\right) \right] \\ &>&\frac{1}{2}E\left[ U\left( y+x_{1}\rho_{y},z+x_{1}\rho_{z}+\widetilde{ \alpha }_{2}\right) \right] -\frac{1}{2}E\left[ U\left( y+x_{1}\rho_{y},z+x_{1}\rho_{z}+\widetilde{\alpha }_{1}\right) \right] . \end{array} $$

for all \(( y,z,x_{1},x_{2}) \in \mathbb {R}_{+}^{4}\) with x 2 > x 1. The previous inequation is satisfied for all x 2 > x 1 if and only if \(E\left [ U\left ( y+t\rho _{y},z+t\rho _{z}+\widetilde {\alpha } _{2}\right ) \right ] -E\left [ U\left ( y+t\rho _{y},z+t\rho _{z}+\widetilde { \alpha }_{1}\right ) \right ] \) is increasing in t or, equivalently, if and only if \(E\left [ f\left ( y+t\rho _{y},z+t\rho _{z}+\widetilde {\alpha } _{2}\right ) \right ] \) is larger than \(E\left [ f\left ( y+t\rho _{y},z+t\rho _{z}+\widetilde {\alpha }_{1}\right ) \right ] \) for all \((y,z,t)\in \mathbb {R} _{+}^{3}\). This is satisfied if and only if \(E\left [ f\left ( y,z+\widetilde { \alpha }_{2}\right ) \right ] \) is larger than \(E\left [ f\left ( y,z+\widetilde { \alpha }_{1}\right ) \right ] \) for all \((y,z)\in \mathbb {R}_{+}^{2}.\) Since we want this inequality to be true for all \((\widetilde {\alpha }_{1}, \widetilde {\alpha }_{2})\) such that \(\widetilde {\alpha }_{2}\) is an increase in Nth-degree risk over \(\widetilde {\alpha }_{1},\) this inequality is equivalent, by Lemma 1, to ( − 1)N f (0, N) ≥ 0.

1.4 Proof of Proposition 4

In an expected utility model, Nth-degree multiplicative-risk attraction in the direction of (ρ y , ρ z ) is equivalent to

$$\begin{array}{@{}rcl@{}} &&\frac{1}{2}E\left[ U\left( y+x_{2}\rho_{y},z+x_{2}\rho_{z}\widetilde{ \alpha }_{2}\right) \right] -\frac{1}{2}E\left[ U\left( y+x_{2}\rho_{y},z+x_{2}\rho_{z}\widetilde{\alpha }_{1}\right) \right] \\ &>&\frac{1}{2}E\left[ U\left( y+x_{1}\rho_{y},z+x_{1}\rho_{z}\widetilde{ \alpha }_{2}\right) \right] -\frac{1}{2}E\left[ U\left( y+x_{1}\rho_{y},z+x_{1}\rho_{z}\widetilde{\alpha }_{1}\right) \right] \end{array} $$
(11)

for all \((\widetilde {\alpha }_{1},\widetilde {\alpha }_{2})\) such that \( \widetilde {\alpha }_{1}\preccurlyeq _{N}\widetilde {\alpha }_{2}\) and all (x 1, x 2) such that 0 < x 1 < x 2 and y + x i ρ y ≥ 0, i = 1, 2. This inequality is satisfied for all x 2 > x 1 if and only if \(E \left [ U( y+t\rho _{y},z+t\rho _{z}\widetilde {\alpha }_{2}) \right ] -E\left [ U\left ( y+t\rho _{y},z+t\rho _{z}\widetilde {\alpha } _{1}\right ) \right ] \) is increasing on \(T(y)=\left \{t\in \mathbb {R} _{+}:y+t\rho _{y}\geq 0\right \} \) or, equivalently, if and only if \(E\left [ h(\widetilde {\alpha }_{2},y,z,t) \right ] \) is larger than \(E [ h( \tilde {\alpha }_{1},y,z,t) ] \) for all \((y,z)\in \mathbb {R}_{+}^{2}\) and for tT(y) where h(α, y, z, t) = ρ y U (1, 0) (y + t ρ y , z + t ρ z α) + α ρ z U (0, 1) (y + t ρ y , z + t ρ z α). By Lemma 1, this last inequality is satisfied for every pair \((\widetilde {\alpha }_{1},\widetilde {\alpha }_{2})\) such that \(\widetilde {\alpha }_{2}\) is an increase in Nth-degree risk over \( \widetilde {\alpha }_{1}\) if and only if \((-1)^{N}\frac {\partial ^{N}h}{\partial \alpha ^{N}}\geq 0.\)

Simple calculus gives \(\frac {\partial ^{N}h}{\partial \alpha ^{N}}=t^{N}\rho _{y}\rho _{z}^{N}U^{(1,N)}+\alpha t^{N}\rho _{z}^{N+1}U^{\left ( 0,N+1\right ) }+Nt^{N-1}\rho _{z}^{N}U^{( 0,N) }\) where derivatives of U are taken at (y + t ρ y , z + α t ρ z ). Since t ≥ 0 and ρ z ≥ 0, our necessary and sufficient condition can then be rewritten as ( − 1)N t ρ y U (1, N) + α t ρ z U (0, N + 1) + N U (0, N) ≥ 0 for all \((\alpha ,y,z)\in \mathbb {R}_{+}^{3}\ \)and for tT(y). Denoting y + t ρ y by Y and z + α t ρ z by Z, our necessary and sufficient condition is equivalent to ( − 1)N(Yy)U (1, N) + (Zz)U (0, N + 1) + N U (0, N) ≥ 0 for all \((Y,Z)\in \mathbb {R}_{+}^{2}\) and all (y, z) such that yY and 0 ≤ zZ and where the derivatives are taken at (Y, Z). Taking successively y = Y and z = Z, y = Y and z = 0 we obtain that (− 1)N U (0, N) ≥ 0 and ( − 1)N Z U (0, N + 1) + N U (0, N) ≥ 0. Letting y go to , we obtain ( − 1)N U (1, N) ≤ 0. Conversely, if these 3 conditions are satisfied and for a given yY, we have ( − 1)N(Yy)U (1, N) ≥ 0 and

$$\begin{array}{@{}rcl@{}} &&(-1)^{N}\left((Z-z)U^{(0,N+1)}+NU^{(0,N)}\right)\\ &=&\frac{Z-z}{Z}(-1)^{N}\left(ZU^{(0,N+1)}+NU^{(0,N)}\right)+\frac{z}{Z}(-1)^{N}NU^{(0,N)}\\ &\geq &0 \end{array} $$

which gives that our necessary and sufficient condition is satisfied.

A far as Nth-degree multiplicative-risk aversion is concerned, it is characterized by the fact that \(E\left [ U\left ( y+t\rho _{y},z+t\rho _{z} \widetilde {\alpha }_{2}\right ) \right ] -E\left [ U\left ( y+t\rho _{y},z+t\rho _{z}\widetilde {\alpha }_{1}\right ) \right ] \) is decreasing on \(T(y)=\left \{ t\in \mathbb {R}_{+}:y+t\rho _{y}\geq 0\right \} \) or, equivalently, by the fact that \(E[ -h( \widetilde {\alpha }_{2},y,z,t) ] \) is larger than \(E[ -h( \tilde {\alpha }_{1},y,z,t) ] \) for all \((y,z)\in \mathbb {R}_{+}^{2}\) and for tT(y) or, finally, by \( ( -1)^{N+1}\frac {\partial ^{N}h}{\partial \alpha ^{N}}\geq 0.\) The rest of the proof is then directly adapted from the multiplicative-risk attraction setting.

Appendix B: Nth-degree stochastic dominance

In this appendix we generalize Lemma 1, Propositions 1 and 2, and Corollary 1. We begin by formalizing the concept of Nth-degree stochastic dominance.

Definition 4 (Jean)

Let \(\widetilde {\alpha }_{1}\) and \(\widetilde {\alpha }_{2}\) denote two random variables with values in [0, B]. We say that \( \widetilde {\alpha }_{2}\) is dominated in the sense of Nth-degree stochastic dominance by \(\widetilde {\alpha }_{1},\) and we denote it by \(\widetilde { \alpha }_{2}\succcurlyeq _{NSD}\widetilde {\alpha }_{1},\) if \(F_{\widetilde { \alpha }_{2}}^{[ N] }(x)\geq F_{\widetilde {\alpha }_{1}}^{\left [ N \right ] }(x)\) for all x ∈ [0, B] where the inequality is strict for some x and \(F_{\widetilde {\alpha }_{2}}^{[ k] }(B)\geq F_{\widetilde {\alpha }_{1}}^{[ k] }(B)\) for k = 1, . . . , N − 1.

Jean (1980) characterizes Nth-degree stochastic dominance: He establishes that \(\widetilde {\alpha }_{2}\) is dominated in the sense of Nth-degree stochastic dominance by \(\widetilde {\alpha }_{1}\) if and only if \(E\left [ q( \widetilde {\alpha }_{2})\right ] >E[ q(\widetilde {\alpha }_{1})] \) for all N times continuously differentiable real valued function q such that ( − 1)k q (k) > 0 for k = 1, . . . , N where \(q^{(k)}=\frac { d^{k}q}{d\alpha ^{k}}.\) The following Lemma characterizes the set of N times continuously differentiable functions for which \(E\left [ q(\widetilde { \alpha }_{2})\right ] \geq E[ q(\widetilde {\alpha }_{1})] \) for all pair \((\widetilde {\alpha }_{1},\widetilde {\alpha }_{2})\) where \(\widetilde {\alpha }_{2}\succcurlyeq _{NSD}\widetilde {\alpha }_{1}.\)

Lemma 2

Let q be a given real valued function that is N times continuously differentiable on+. The following are equivalent.

  1. 1.

    For all pair \((\widetilde {\alpha }_{1},\widetilde {\alpha }_{2})\) such that \(\widetilde {\alpha }_{2}\succcurlyeq _{NSD}\widetilde {\alpha }_{1},\) we have \(E[ q(\widetilde {\alpha }_{2})] \geq E\left [ q(\widetilde { \alpha }_{1})\right ] .\)

  2. 2.

    For all x ≥ 0, we have ( − 1)k q (k) ≥ 0 for k = 1, . . . , N.

Proof

The fact that 2. implies 1. results directly from Jean (1980). Let us now prove that 1. implies 2. Let q be an N times continuously differentiable function on \(\mathbb {R}_{+}\) such that \(E\left [ q(\widetilde {\alpha }_{2})\right ] \geq E\left [ q(\widetilde {\alpha }_{1}) \right ] \) for all pair \((\widetilde {\alpha }_{1},\widetilde {\alpha }_{2})\) such that \(\widetilde {\alpha }_{2}\succcurlyeq _{NSD}\widetilde {\alpha } _{1}. \) Let k be in {1, . . . , N} and let \((\widetilde {\alpha }_{1},\widetilde {\alpha }_{2})\) such that \(\widetilde {\alpha } _{2}\succcurlyeq _{k}\widetilde {\alpha }_{1}.\) By definition, we have \(F_{ \widetilde {\alpha }_{2}}^{[ k] }(x)\geq F_{\widetilde {\alpha } _{1}}^{[ k] }(x)\) for all x ∈ [0, B] where the inequality is strict for some x and \(F_{\widetilde {\alpha }_{2}}^{\left [ i \right ] }(B)=F_{\widetilde {\alpha }_{1}}^{[ i] }(B)\) for i = 1, . . . , k − 1. By integration, we obtain \(F_{\widetilde {\alpha }_{2}}^{\left [ j\right ] }(x)\geq F_{\widetilde {\alpha }_{1}}^{[ j] }(x)\) for all jk and all x ∈ [0, B] and, in particular, \( \widetilde {\alpha }_{2}\succcurlyeq _{NSD}\widetilde {\alpha }_{1}\) hence \(E [ q(\widetilde {\alpha }_{2})] \geq E\left [ q(\widetilde {\alpha } _{1})\right ] .\) By Lemma 2 we have ( − 1)k q (k) ≥ 0 on 0, B for k = 1, . . . , N. □

The following Proposition establishes the effect on optimal choice of an Nth-degree stochastically dominated shift on the initial endowment α.

Proposition 5

Let U be a given increasing, strictly concave and infinitely differentiable utility function satisfying Assumption A1. Let us consider p and μ as given. The following properties are equivalent:

  1. 1.

    For all initial endowment \((K,\widetilde {\alpha }_{1})\), a switch from \(\widetilde {\alpha }_{1}\) to \(\widetilde {\alpha }_{2}\) such that \(\widetilde {\alpha }_{2}\succcurlyeq _{NSD}\widetilde {\alpha }_{1}\) increases the optimal level of the choice variable, i.e., \(x_{2}^{\ast }\geq x_{1}^{\ast }\).

  2. 2.

    For all (y, z), we have ( − 1)k (− p U (1, k)(y, z) + μ U (0, k + 1)(y, z)) ≥ 0 for k = 1, . . . , N.

Proof

Let us prove that 2. implies 1. Let \(\left ( \widetilde { \alpha }_{1},\widetilde {\alpha }_{2}\right ) \) be such that \(\widetilde { \alpha }_{2}\succcurlyeq _{NSD}\widetilde {\alpha }_{1}\) and let \(q(\alpha )=g(x_{1}^{\ast },\alpha ,\mu ).\) By 2., we have ( − 1)k q (k)(α) ≥ 0 for k = 1, . . . , N. By Lemma 2, this leads to \(E[ q(\widetilde {\alpha }_{1})] \leq E\left [ q( \widetilde {\alpha }_{2})\right ] .\) By definition, we have \(E\left [ q( \widetilde {\alpha }_{1})\right ] =0,\) which gives \(E\left [ q(\widetilde { \alpha }_{2})\right ] \geq 0\) or \(E\left [ g(x_{1}^{\ast },\widetilde {\alpha } _{2},\mu )\right ] \geq 0.\) By concavity of U, it is easy to check that \(E [ g(x,\widetilde {\alpha }_{2},\mu )] \) is a decreasing function of x. Since \(x_{2}^{\ast }\) is characterized by \(E\left [ g(x_{2}^{\ast }, \widetilde {\alpha }_{2},\mu )\right ] =0,\) we obtain that \(x_{2}^{\ast }\geq x_{1}^{\ast }.\)

Let us prove 1. implies 2. As in the proof of the Lemma 2, it suffices to remark that, for k = 1, . . . , N, \(\widetilde {\alpha } _{2}\succcurlyeq _{k}\widetilde {\alpha }_{1}\) implies \(\widetilde {\alpha } _{2}\succcurlyeq _{NSD}\widetilde {\alpha }_{1}.\)

Finally, the following Proposition establishes the effect on optimal choice of an Nth-degree stochastically dominated shift on the multiplicative variable μ.

Proposition 6

Let U be a given increasing, strictly concave and infinitely differentiable utility function satisfying Assumption A2. The following properties are equivalent:

  1. 1.

    For all initial endowment (K, α) and all asset cost and payoff \( (p,\widetilde {\mu }_{1})\) such that \(x_{1}^{\ast }\geq 0\), and any switch from \(\widetilde {\mu }_{1}\) to \(\widetilde {\mu }_{2}\) with \(\widetilde {\mu } _{2}\succeq _{NSD}\widetilde {\mu }_{1}\) increases (decreases) the optimal level of the choice variable, i.e., \(x_{2}^{\ast }\geq x_{1}^{\ast }\).

  2. 2.

    For all (y, z), we have ( − 1)k U 1, k(y, z) ≤ 0(r e s p. ≥ 0)R k(y, z) ≤ N and( − 1)k U 0, k(y, z) ≥ 0(r e s p. ≤ 0) for k = 1, . . . , N.

Proof

Follows the proof of Proposition 5. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Jouini, E., Napp, C. & Nocetti, D. Economic consequences of Nth-degree risk increases and Nth-degree risk attitudes. J Risk Uncertain 47, 199–224 (2013). https://doi.org/10.1007/s11166-013-9176-6

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11166-013-9176-6

Keywords

JEL Classification

Navigation