Skip to main content
Log in

Decreasing absolute risk aversion, prudence and increased downside risk aversion

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

Abstract

Downside risk increases have previously been characterized as changes preferred by all decision makers u(x) with u′′′(x) > 0. For risk averse decision makers, u′′′(x) > 0 also defines prudence. This paper finds that downside risk increases can also be characterized as changes preferred by all decision makers displaying decreasing absolute risk aversion (DARA) since those changes involve random variables that have equal means. Building on these findings, the paper proposes using “more decreasingly absolute risk averse” or “more prudent” as alternative definitions of increased downside risk aversion. These alternative definitions generate a transitive ordering, while the existing definition based on a transformation function with a positive third derivative does not. Other properties of the new definitions of increased downside risk aversion are also presented.

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. Menezes, Geiss and Tressler say that “It is natural to expect an individual to be averse to downside risk if he is decreasingly risk averse.”

  2. An example provided later in the paper also shows that the ∅'''(u) ≥ 0 based definition allows distinct u(x) and v(x) to each be more downside risk averse than the other. Thus, the definition is not asymmetric either.

  3. Keenan and Snow (2012) recognize that the definition of increased downside risk aversion that they proposed earlier leads to an intransitive order. They also conclude, however, that “If one insists on a condition that fully characterizes greater downside risk aversion for both large and small changes in risk preferences, one has no choice but to use the condition ∅''' > 0.”

  4. If risk averse functions are desired, v(x) = −e-dx, u(x) = −e-cx and ∅(u) = -(-u)d/c provides an almost identical example.

  5. Because not all functions have integrals of closed form, it may not be possible to determine the u(x) associated with marginal utility u'(x).

  6. All these measures of the intensity of downside risk aversion have some desirable as well as undesirable properties. See Crainich and Eeckhoudt (2008) for the desirable properties of d(x), Keenan and Snow (2002, 2009, 2012) for the desirable properties of s(x). An undesirable property shared by s(x) and -Au'(x) here is that the intensity measure of downside risk aversion may be inconsistent with the direction of downside risk aversion. For example, quadratic utility functions are downside risk neutral (because u''' = 0), but s(x) < 0 and -Au'(x) < 0. In the case of constant absolute risk averse utility functions, the direction of downside risk aversion is positive (meaning decision makers with these utility functions are downside risk averse) while the intensity measure is zero for -Au'(x) and negative for s(x).

  7. Notice that this same use of marginal utility can be made when discussing when one decision maker is more risk averse than another. It is the case that v'(x) is more risk averse than u'(x) if and only if Aδ(x) ≥ 0. This provides an additional characterization of the Arrow-Pratt definition of more risk averse.

  8. Denuit and Eeckhoudt (2010) extend Chiu’s analysis to higher order risk aversion.

References

  • Chiu, W. H. (2005). Skewness preferences, risk aversion, and the precedence relationships on stochastic changes. Management Science, 51, 1816–1828.

    Article  Google Scholar 

  • Crainich, D., & Eeckhoudt, L. (2008). On the intensity of downside risk aversion. Journal of Risk and Uncertainty, 36, 267–276.

    Article  Google Scholar 

  • Denuit, M., & Eeckhoudt, L. (2010). A general index of absolute risk attitude. Management Science, 56, 712–715.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Fishburn, P., & Vickson, R. (1978). Theoretical foundations of stochastic dominance. In G. Whitmore & M. Findlay (Eds.), Stochastic dominance: An approach to decision-making under risk. Lexington: Lexington Books, D.C. Heath and Company.

    Google Scholar 

  • Jindapon, P., & Neilson, W. (2007). Higher-order generalizations of Arrow-Pratt and Ross risk aversion: A comparative statics approach. Journal of Economic Theory, 136, 719–728.

    Article  Google Scholar 

  • Keenan, D., & Snow, A. (2002). Greater downside risk aversion. Journal of Risk and Uncertainty, 24, 267–277.

    Article  Google Scholar 

  • Keenan, D., & Snow, A. (2009). Greater downside risk aversion in the large. Journal of Economic Theory, 144, 1092–1101.

    Article  Google Scholar 

  • Keenan, D., & Snow, A. (2010). Greater prudence and greater downside risk aversion. Journal of Economic Theory, 145, 2018–2026.

    Article  Google Scholar 

  • Keenan, D., & Snow, A. (2012). The Schwarzian derivative as a ranking of downside risk aversion. Journal of Risk and Uncertainty, 44(2), in press.

  • Kimball, M. (1990). Precautionary saving in the small and in the large. Econometrica, 58, 53–73.

    Article  Google Scholar 

  • Leland, H. (1968). Saving and uncertainty: The precautionary demand for saving. Quarterly Journal of Economics, 82, 465–473.

    Article  Google Scholar 

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

    Google Scholar 

  • Meyer, J. (2010). Representing risk preferences in expected utility based decision models. Annals of Operations Research, 176, 179–190.

    Article  Google Scholar 

  • Meyer, D., & Meyer, J. (2010). A Diamond-Stiglitz approach to the demand for self-protection. Journal of Risk and Uncertainty, 42, 45–60.

    Article  Google Scholar 

  • Modica, S., & Scarsini, M. (2005). A note on comparative downside risk aversion. Journal of Economic Theory, 122, 267–271.

    Article  Google Scholar 

  • Pratt, J. (1964). Risk aversion in the small and in the large. Econometrica, 32, 122–136.

    Article  Google Scholar 

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

    Article  Google Scholar 

Download references

Acknowledgments

The authors thank Louis Eeckhoudt, Harris Schlesinger, the editor and an anonymous referee for helpful comments. Support from the Private Enterprise Research Center at Texas A&M University is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Liqun Liu.

Appendix

Appendix

1.1 A1 Lemma 1

Lemma 1

F(x) is preferred or indifferent to G(x) for all decision makers v′(x) = δ′(x)·u′(x) with δ′′′(x) > 0 if and only if G(x) is an expected utility preserving increase in downside risk from F(x) for u(x).

Proof

The proof follows the similar steps in MGT′s proof of Theorem 1. Using v′(x) = δ′(x)·u′(x),

$$ {{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = \int {_{\text{a}}^{\text{b}}{\text{v}}\prime \left( {\text{x}} \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = \int {_{\text{a}}^{\text{b}}\delta \prime \left( {\text{x}} \right){\text{u}}\prime \left( {\text{x}} \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}}.} } $$

(The “only if” part)

Suppose EFv(x) ≥ EGv(x) for all δ′′′(x) > 0.

Let δ1(x) = θx3/3 + x and δ2(x) = θx3/3−x where θ > 0. Since δ1′′′(x) = δ2′′′(x) > 0, we have

$$ \begin{array}{*{20}{c}} {\int {_a^b\left( {\theta {x^2} + 1} \right)u\prime (x)\left( {G(x) - F(x)} \right)dx \geqslant 0} } \hfill \\ {\int {_a^b\left( {\theta {x^2} - 1} \right)u\prime (x)\left( {G(x) - F(x)} \right)dx \geqslant 0} } \hfill \\ \end{array} $$

which continues to hold as θ approaches zero. So \( \int {_{\text{a}}^{\text{b}}{\text{u}}\prime \left( {\text{x}} \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = 0} \), which is a) in Definition 4.

Note that when \( \int {_{\text{a}}^{\text{b}}{\text{u}}\prime \left( {\text{x}} \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = 0}, \)

$$ {{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = \int {_{\text{a}}^{\text{b}}\delta \prime \left( {\text{x}} \right){\text{u}}\prime \left( {\text{x}} \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}}{.} = - \int {_{\text{a}}^{\text{b}}\delta \prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}}.} } } $$

Now let δ3(x) = θx3/3 + x 2/2 and δ4(x) = θx3/3−x 2/2 where θ > 0. Since δ3′′′(x) = δ4′′′(x) > 0, we have

$$ \begin{array}{*{20}{c}} { - \int {_a^b\left( {2\theta x + 1} \right)\int {_a^xu\prime (s)\left( {G(s) - F(s)} \right)dsdx \geqslant 0} } } \hfill \\ { - \int {_a^b\left( {2\theta x - 1} \right)\int {_a^xu\prime (s)\left( {G(s) - F(s)} \right)dsdx \geqslant 0} } } \hfill \\ \end{array} $$

which continue to hold as θ approaches zero. So \( \int {_{\text{a}}^{\text{b}}\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} = 0} } \), which is the equality part of b) in Definition 4.

With \( \int {_{\text{a}}^{\text{b}}{\text{u}}\prime \left( {\text{x}} \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = 0\;{\text{and}}\;\int {_{\text{a}}^{\text{b}}\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} = 0} } }, \)

$$ {{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = \int {_{\text{a}}^{\text{b}}\delta \prime \prime \prime \left( {\text{y}} \right)\int {_{\text{a}}^{\text{y}}\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \cdot {\text{dy}}.} } } $$

We prove the weak inequality of b) by contradiction. Assume \( \int {_{\text{a}}^{\text{y}}\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} < 0} } \) for some y0 in [a,b]. Then there exists an interval in [a, b], [α, β], such that \( \int {_{\text{a}}^{\text{y}}\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} < 0} } \) for all y in [α, β]. Therefore, for

$$ {\delta_5}(x) = \left\{ {\begin{array}{*{20}{c}} {{x^3}/6,} \hfill & {x \in \left[ {\alpha, \beta } \right]} \hfill \\ {\theta {x^3}/6,} \hfill & {otherwise} \hfill \\ \end{array} } \right., $$

EFv(x) - EGv(x) < 0 for sufficiently small θ. Because δ5′′′(x) > 0, that contradicts the original assumption that EFv(x) ≥ EGv(x) for all δ′′′(x) > 0.

Moreover, \( \int {_{\text{a}}^{\text{y}}\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} > 0} } \) for some y in (a, b) as long as G(x) and F(x) are not identical.

(The “if” part)

Suppose G(x) is an expected utility preserving increase in downside risk from F(x) for u(x). Then,

$$ \begin{array}{*{20}{c}} {{{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = \int {_{\text{a}}^{\text{b}}\delta \prime \prime \prime \left( {\text{y}} \right)\int {_{\text{a}}^{\text{y}}\int {_{\text{a}}^{\text{x}}{\text{u}}\prime \left( {\text{s}} \right)\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \cdot {\text{dy}} \geqslant 0\;{\text{for}}\;{\text{all}}\;\delta \prime \prime \prime \left( {\text{y}} \right) > 0.} } } } \hfill & {\text{QED}} \hfill \\ \end{array} $$

1.2 A2 Risk adjusted cumulative distribution functions

For continuously differentiable CDFs, note that the derivative of u′(x)(G(x)−F(x)) is u′(g–f) + u′′(G–F), where f and g are the densities associated with CDFs F and G. First assume u′′ ≥ 0. Then u′·g + u′′·G ≥ 0 and u′·f + u′′·F ≥ 0. These two expressions can be density functions if the probabilities sum to one. Integrate each over [a, b] to find the total value for each is u′(b) and thus the normalizing factor is t = 1/u′(b). Next assume u′′ < 0 and consider u′·g + u′′·F ≥ 0 and u′·f + u′′·G ≥ 0. Again these can be density functions if the probabilities add to one. Because \( \int {_{\text{a}}^{\text{b}}\left( {{\text{u}}\prime \cdot {\text{g}} + {\text{u}}\prime \prime {\text{G}} - {\text{u}}\prime \cdot {\text{f}} - {\text{u}}\prime \prime {\text{F}}} \right){\text{dx}} = 0{,}\int {_{\text{a}}^{\text{b}}\left( {{\text{u}}\prime \cdot {\text{g}} + {\text{u}}\prime \prime \cdot {\text{F}}} \right){\text{dx}} = \int {_{\text{a}}^{\text{b}}\left( {{\text{u}}\prime \cdot {\text{f}} + {\text{u}}\prime \prime {\text{G}}} \right){\text{dx}} = 1/{\text{t}}} } } \) is the normalizing factor.

The probability distributions for these random variables have been “adjusted” by the reference decision maker with marginal utility u′(x) so that the means of \( \widehat{G}\left( {\text{x}} \right) \) and \( \widehat{F}\left( {\text{x}} \right) \) are the same. That is, they are risk adjusted so that for this person the choice between them depends only on their mean values. It is also the case that the conditions in Definition 4 imply that \( \widehat{G}\left( {\text{x}} \right) \) is a downside increase in risk from \( \widehat{F}\left( {\text{x}} \right) \) by the original MGT definition. Thus, the function u′(x)(G(x)−F(x)), properly normalized, adjusts the random alternatives associated with G(x) and F(x) so that he or she can act as a risk neutral person and choose between them based only on the mean values. This leads to an interpretation of Theorem 5. This theorem indicates that for those who are more downside risk averse than the person doing this risk adjustment, that is, those who are more decreasingly absolute risk averse than u′(x), this risk adjustment is not sufficient and \( \widehat{F}\left( {\text{x}} \right) \) is preferred or indifferent to \( \widehat{G}\left( {\text{x}} \right) \).

Consider the following example for F(x) and G(x) defined on [0, 1].

$$ {\text{Let}}\;{\text{G}}\left( {\text{x}} \right) = \begin{array}{*{20}{c}} 0 \hfill & {{\text{x}} < 0} \hfill \\ {.4{\text{x}}} \hfill & {0 \leqslant {\text{x}} < .75} \hfill \\ {.4{\text{x}} + .6} \hfill & {.75 \leqslant {\text{x}} < 1} \hfill \\ 1 \hfill & {{\text{x}} \geqslant 1} \hfill \\ \end{array} $$

and

$$ {\text{F}}\left( {\text{x}} \right) = \begin{array}{*{20}{c}} 0 \hfill & {{\text{x}} < .25} \hfill \\ {.8{\text{x}}} \hfill & {25 \leqslant {\text{x}} < .75} \hfill \\ {.6} \hfill & {.75 \leqslant {\text{x}} < 1} \hfill \\ 1 \hfill & {{\text{x}} \geqslant 1} \hfill \\ \end{array} $$

This implies that

$$ {\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right) = \begin{array}{*{20}{c}} 0 \hfill & {{\text{x}} < 0} \hfill \\ {.4{\text{x}}} \hfill & {0 \leqslant {\text{x}} < .25} \hfill \\ { - .4{\text{x}}} \hfill & {.25 \leqslant {\text{x}} < .75} \hfill \\ {.4{\text{x}}} \hfill & {.75 \leqslant {\text{x}} < 1} \hfill \\ 0 \hfill & {{\text{x}} \geqslant 1} \hfill \\ \end{array} $$

Consider utility function u(x) = ln x so that u′(x) = 1/x. This implies that

$$ {\text{u}}\prime \left( {\text{x}} \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right) = \begin{array}{*{20}{c}} 0 \hfill & {x < 0} \hfill \\ {.4} \hfill & {0 \leqslant {\text{x}} < .25} \hfill \\ { - .4} \hfill & {.25 \leqslant {\text{x}} < .75} \hfill \\ {.4} \hfill & {.75 \leqslant {\text{x}} < 1} \hfill \\ 0 \hfill & {{\text{x}} \geqslant 1} \hfill \\ \end{array} $$

This sum of the increases (or decreases) for u′(x)(G(x)−F(x)) is 1.2 which is greater than 1 so a scaling factor less than or equal to 1/1.2 is needed to allow u′(x)(G(x)−F(x)) to represent the difference between two CDFs. These rescaled CDFs are given below and can be viewed as the difference between two risk adjusted CDFs where person with utility u(x) = ln x has adjusted CDFs F(x) and G(x) so that they have equal mean values. For that person the ranking of the two CDFs is the same, but for those more decreasing absolute risk averse, the adjustment is not sufficient, and \( \widehat{\text{F}}\left( {\text{x}} \right) \) is preferred or indifferent to \( \widehat{\text{G}}\left( {\text{x}} \right) \).

$$ \widehat{\text{G}}\left( {\text{x}} \right) - \widehat{\text{F}}\left( {\text{x}} \right) = \begin{array}{*{20}{c}} 0 \hfill & {{\text{x}} < 0} \hfill \\ {{1}/{3}} \hfill & {0 \leqslant {\text{x}} < .{25}} \hfill \\ { - {1}/{3}} \hfill & {.{25} \leqslant {\text{x}} < .{75}} \hfill \\ {{1}/{3}} \hfill & {.{75} \leqslant {\text{x}} < {1}} \hfill \\ 0 \hfill & {{\text{x}} \geqslant 1} \hfill \\ \end{array} $$

1.3 A3 Proof of Theorem 5′

Proof

(The “if” part) Suppose that G(x) is an expected utility preserving increase in downside risk from F(x) for u(x). Then for any utility function v(x),

$$ \begin{array}{*{20}{c}} {{{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = - \int {_{\text{a}}^{\text{b}}{\text{v}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{dsdx}}} } } \\ { = - \int {_{\text{a}}^{\text{b}}\phi \prime {\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{dsdx}}} } } \\ { = \int {_{\text{a}}^{\text{b}}\phi \prime \prime {\text{u}}\prime \prime \left( {\text{y}} \right)\int {_{\text{a}}^{\text{y}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \cdot {\text{dy}}{.}} } } } \\ \end{array} $$

where the first equality is due to a) and the third equality due to b) in Definition 4′. Also from b) in Definition 4′, \( \int {_{\text{a}}^{\text{y}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \cdot {\text{dy}} \leqslant 0\;{\text{for}}\;{\text{all}}\;{\text{x}}} } \). Therefore, v(x) being more downside risk averse than u(x)–∅′′u′′(y) < 0 according to Lemma 2–implies EFv(x)–EGv(x) ≥ 0.

(The “only if” part) Suppose u(x) is risk averse. The case in which u(x) is risk loving can be similarly dealt with. Now suppose that F(x) is preferred or indifferent to G(x) for all risk averse decision makers v(x) who are more downside risk averse than u(x). That is \( {{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = \int {_{\text{a}}^{\text{b}}{\text{v}}\prime \left( {\text{x}} \right){\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right){\text{)dx}} = \int {_{\text{a}}^{\text{b}}\phi \left( {{\text{u}}\prime } \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} \geqslant 0\;{\text{for}}\;{\text{all}}\;\phi \;{\text{such}}\;{\text{that}}\;\phi \prime \prime > 0} } \).

Let ∅(u′) be θ(u′)2 +1 and θ(u′)2 -1, respectively, where θ > 0. That the above inequality holds for both ∅(u′) and for all positive θ implies \( \int {_{\text{a}}^{\text{b}}\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = 0} \), which is a) in Definition 4′.

With \( \int {_{\text{a}}^{\text{b}}\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = 0} \), we have \( {{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = \int {_{\text{a}}^{\text{b}}\phi \left( {{\text{u}}\prime } \right)\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = - \int {_{\text{a}}^{\text{b}}\phi \prime {\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{dsdx}} \geqslant 0} } } \). Let ∅(u′) be θ(u′)2 + u′ and θ(u′)2–u′, respectively, where θ > 0. Again, that this inequality holds for these ∅(u′) and for all positive θ implies \( \int {_{\text{a}}^{\text{b}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{dsdx}} = 0} } \), which is the equality in b) in Definition 4′.

With \( \int {_{\text{a}}^{\text{b}}\left( {{\text{G}}\left( {\text{x}} \right) - {\text{F}}\left( {\text{x}} \right)} \right){\text{dx}} = 0\;{\text{and}}\;} \int {_{\text{a}}^{\text{b}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} = 0} } \), we have \( {{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = \int {_{\text{a}}^{\text{b}}\phi \prime \prime {\text{u}}\prime \prime \left( {\text{y}} \right)\int {_{\text{a}}^{\text{y}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \cdot {\text{dy}} \geqslant 0} } } \) for all ∅′′ > 0, which implies \( \int {_{\text{a}}^{\text{y}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \leqslant 0} } \), the weak inequality of b) in Definition 4′.

Moreover, \( \int {_{\text{a}}^{\text{y}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} < 0} } \) for some y in (a, b) as long as G(x) and F(x) are not identical. QED

1.4 A4 Proof of Corollary 1′

Proof

Suppose v(x) is more downside risk averse than u(x). Then for any expected utility preserving increase in downside risk from F(x) to G(x) with u(x) as the reference decision maker,

$$ \begin{array}{*{20}{c}} {{{\text{E}}_{\text{F}}}{\text{v}}\left( {\text{x}} \right) - {{\text{E}}_{\text{G}}}{\text{v}}\left( {\text{x}} \right) = - \int {_{\text{a}}^{\text{b}}{\text{v}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{dsdx}}} } } \\ { = - \int {_{\text{a}}^{\text{b}}\delta \prime \prime \left( {\text{x}} \right){\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{dsdx}}} } } \\ { = \int {_{\text{a}}^{\text{b}}\delta \prime \prime \prime \left( {\text{y}} \right)\int {_{\text{a}}^{\text{y}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \cdot {\text{dy}}{.}} } } } \\ \end{array} $$

where the first equality is due to a) and the third equality due to b) in Definition 4′.

If we further assume that v′′(x)·u′′(x) > 0 in [a,b], then v(x) is more downside risk averse than u(x) implies δ′′′(y) ≤ 0. Therefore EFv(x)–EGv(x) ≥ 0 since \( \int {_{\text{a}}^{\text{y}}{\text{u}}\prime \prime \left( {\text{x}} \right)\int {_{\text{a}}^{\text{x}}\left( {{\text{G}}\left( {\text{s}} \right) - {\text{F}}\left( {\text{s}} \right)} \right){\text{ds}} \cdot {\text{dx}} \leqslant 0} } \) according to b) in Definition 4′.

Conversely, if EFv(x)–EGv(x) ≥ 0 for all expected utility preserving downside risk increases with u(x) as the reference decision maker, then it has to be the case that δ′′′(y) ≤ 0 and v(x) is more downside risk averse than u(x) (with the additional assumption that v′′(x)·u′′(x) > 0 in [a, b]). The proof of this is by contradiction with arguments parallel to those in Keenan and Snow (2009). Suppose δ′′′(y) > 0 at some point y0 in [a, b]. Then there would exist a surrounding interval within [a, b] on which δ′′′(y) > 0, which implies by choosing G(x)–F(x) to be concentrated on the same interval, one would obtain EFv(x)–EGv(x) < 0, contradicting the initial assumption that EFv(x)–EGv(x) ≥ 0. QED

Rights and permissions

Reprints and permissions

About this article

Cite this article

Liu, L., Meyer, J. Decreasing absolute risk aversion, prudence and increased downside risk aversion. J Risk Uncertain 44, 243–260 (2012). https://doi.org/10.1007/s11166-012-9141-9

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11166-012-9141-9

Keywords

JEL Classification

Navigation