Skip to main content
Log in

Uncertain discount and hyperbolic preferences

  • Published:
Theory and Decision Aims and scope Submit manuscript

Abstract

This paper studies the interaction between savagean uncertainty and time preferences. We introduce a variation of the discounted subjective expected utility model, where time preferences are state dependent. Before uncertainty is resolved, the individual is unsure about the discount factor that will be used, even when evaluating certain payoffs. The model can account for the present bias and diminishing impatience, even if the future is discounted geometrically. The present bias disappears when the immediate payoff becomes uncertain. Although preferences are not stationary, choices may be time consistent.

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. See Thaler (1981), Benzion et al. (1989) and more recently, Epper et al. (2011), Andreoni and Sprenger (2012), Halevy (2015), for evidence of diminishing impatience.

  2. Section 2 contains an alternative example.

  3. See Halevy (2005) for references.

  4. A relevant corrigendum of Halevy (2008) and Saito (2011) is Chakraborty and Halevy (2016), where the correct relations between violations of stationarity and violations of expected utility are presented.

  5. The positive relation between income and patience dates back to Fisher (1930): “The degree of his impatience depends on his entire income stream, beginning at the present instant and stretching indefinitely into the future.”

  6. \(4(0.5\times 0.99^0+0.5\times 0.3^0)>7(0.5\times 0.3^7+0.5\times 0.99^7)\).

  7. Gollier (2004) proposed an alternative rationale giving the opposite result, impatience should be increasing. This is called “Weitzman–Gollier puzzle” (Gollier and Weitzman 2010).

  8. In his experiment, up to 20% of the subjects made non-stationary but time consistent choices.

  9. The meaning of invariance is the following (see Halevy 2015): \((x',x'',x''',\ldots )\sim _\tau (y',y'',y''',\ldots )\), if and only if \((x',x'',x''',\ldots )\sim _{\tau +1} (y',y'',y''',\ldots )\).

  10. The meaning of consistency is the following (see Halevy 2015): \((z,x',x'',\ldots )\sim _\tau (z,y',y'',\ldots )\), if and only if \((x',x'',\ldots )\sim _{\tau +1} (y',y'',\ldots )\).

  11. A similar condition has been proposed by Millner and Heal (2016) in a different choice problem: time consistency of a social planner that aggregates preferences of individuals with heterogeneous discount rates.

  12. As noted in Karni (2007, Footnote 7) his results hold for any product of connected, separable, topological spaces. Our outcome space \(X^\infty \) satisfies the previous properties. Few conditions characterize the model: \(\succcurlyeq \) is a continuous weak order, there exist a best and a worst outcome (A.0 in Karni 2007) and cardinal coherence (A.1 in Karni 2007).

References

  • Andreoni, J., & Sprenger, C. (2012). Estimating time preferences from convex budgets. The American Economic Review, 102(7), 3333–3356.

    Article  Google Scholar 

  • Azfar, O. (1999). Rationalizing hyperbolic discounting. Journal of Economic Behavior & Organization, 38(2), 245–252.

    Article  Google Scholar 

  • Baucells, M., & Heukamp, F. H. (2010). Common ratio using delay. Theory and Decision, 68(1), 149–158.

    Article  Google Scholar 

  • Baucells, M., & Heukamp, F. H. (2012). Probability and time trade-off. Management Science, 58(4), 831–842.

    Article  Google Scholar 

  • Benzion, U., Rapoport, A., & Yagil, J. (1989). Discount rates inferred from decisions: An experimental study. Management Science, 35(3), 270–284.

    Article  Google Scholar 

  • Bleichrodt, H., Rohde, K. I. M., & Wakker, P. P. (2009). Non-hyperbolic time inconsistency. Games and Economic Behavior, 66(1), 27–38.

    Article  Google Scholar 

  • Chakraborty, A., & Halevy, Y. (2016). Strotz meets Allais: remarks on the relation between present bias and the certainty effect. Unpublished Manuscript

  • Dasgupta, P., & Maskin, E. (2005). Uncertainty and hyperbolic discounting. American Economic Review, 95, 1290–1299.

    Article  Google Scholar 

  • Ellsberg, D. (1961). Risk, ambiguity, and the savage axioms. The Quarterly Journal of Economics, 75(4), 643–669.

  • Epper, T., & Fehr-Duda, H. (2015). The missing link: Unifying risk taking and time discounting. Unpublished manuscript

  • Epper, T., Fehr-Duda, H., & Bruhin, A. (2011). Viewing the future through a warped lens: Why uncertainty generates hyperbolic discounting. Journal of Risk and Uncertainty, 43(3), 169–203.

  • Epstein, L., & Schneider, M. (2003). Recursive multiple-priors. Journal of Economic Theory, 113(1), 1–31.

    Article  Google Scholar 

  • Ergin, H., & Gul, F. (2009). A theory of subjective compound lotteries. Journal of Economic Theory, 144, 899–929.

    Article  Google Scholar 

  • Farmer, J. D., & Geanakoplos, J. (2009). Hyperbolic discounting is rational: Valuing the far future with uncertain discount rates. Tech. Rep. 1719, Cowles Foundation for Research in Economics, Yale University.

  • Fisher, I. (1930). The theory of interest. New York: Macmillan.

    Google Scholar 

  • Ghirardato, P. (2002). Revisiting savage in a conditional world. Economic Theory, 20(1), 83–92.

    Article  Google Scholar 

  • Gollier, C. (2004). Maximizing the expected net future value as an alternative strategy to gamma discounting. Finance Research Letters, 1(2), 85–89.

    Article  Google Scholar 

  • Gollier, C., & Weitzman, M. L. (2010). How should the distant future be discounted when discount rates are uncertain? Economics Letters, 107(3), 350–353.

    Article  Google Scholar 

  • Halevy, Y. (2005). Diminishing impatience: Disentangling time preference from uncertain lifetime. Unpublished manuscript.

  • Halevy, Y. (2008). Strotz meets Allais: Diminishing impatience and the certainty effect. American Economic Review, 98(3), 1145–62.

    Article  Google Scholar 

  • Halevy, Y. (2015). Time consistency: Stationarity and time invariance. Econometrica, 83(1), 335–352.

    Article  Google Scholar 

  • Harris, C., & Laibson, D. (2013). Instantaneous gratification. The Quarterly Journal of Economics, 128(1), 205–248.

    Article  Google Scholar 

  • Hayashi, T. (2003). Quasi-stationary cardinal utility and present bias. Journal of Economic Theory, 112(2), 343–352.

    Article  Google Scholar 

  • Higashi, Y., Hyogo, K., & Takeoka, N. (2009). Subjective random discounting and intertemporal choice. Journal of Economic Theory, 144(3), 1015–1053.

    Article  Google Scholar 

  • Jackson, M. O., & Yariv, L. (2014). Present bias and collective dynamic choice in the lab. American Economic Review, 104(12), 4184–4204.

    Article  Google Scholar 

  • Karni, E. (2007). Foundations of Bayesian theory. Journal of Economic Theory, 132(1), 167–188.

    Article  Google Scholar 

  • Keren, G., & Roelofsma, P. (1995). Immediacy and certainty in intertemporal choice. Organizational Behavior and Human Decision Processes, 63(3), 287–297.

    Article  Google Scholar 

  • Koopmans, T. C. (1960). Stationary ordinal utility and impatience. Econometrica, 28, 287.

    Article  Google Scholar 

  • Laibson, D. (1997). Golden eggs and hyperbolic discounting. The Quarterly Journal of Economics, 112(2), 443–77.

    Article  Google Scholar 

  • Loewenstein, G., & Prelec, D. (1992). Anomalies in intertemporal choice: Evidence and an interpretation. The Quarterly Journal of Economics, 107(2), 573–597.

    Article  Google Scholar 

  • Luttmer, E. G. J., & Mariotti, T. (2003). Subjective discounting in an exchange economy. Journal of Political Economy, 111(5), 959–989.

    Article  Google Scholar 

  • Maccheroni, F., Marinacci, M., & Rustichini, A. (2006). Dynamic variational preferences. Journal of Economic Theory, 128(1), 4–44.

    Article  Google Scholar 

  • Millner, A., & Heal, G. (2016). Collective intertemporal choice: The possibility of time consistency. Working Paper 22524, National Bureau of Economic Research. doi:10.3386/w22524.

  • Montiel Olea, J. L., & Strzalecki, T. (2014). Axiomatization and measurement of quasi-hyperbolic discounting. The Quarterly Journal of Economics, 129, 1449–1499.

    Article  Google Scholar 

  • Myerson, J., Green, L., Scott Hanson, J., Holt, D. D., & Estle, S. J. (2003). Discounting delayed and probabilistic rewards: Processes and traits. Journal of Economic Psychology, 24(5), 619–635.

    Article  Google Scholar 

  • Nau, R. (2006). Uncertainty aversion with second-order utilities and probabilities. Management Science, 52(1), 136–145.

    Article  Google Scholar 

  • Saito, K. (2011). Strotz meets allais: Diminishing impatience and the certainty effect: Comment. American Economic Review, 101(5), 2271–2275.

    Article  Google Scholar 

  • Saito, K. (2015). A relationship between risk and time preferences. Unpublished manuscript.

  • Sozou, P. D. (1998). On hyperbolic discounting and uncertain hazard rates. Proceedings of the Royal Society of London Series B: Biological Sciences, 265(1409), 2015–2020.

    Article  Google Scholar 

  • Thaler, R. (1981). Some empirical evidence on dynamic inconsistency. Economics Letters, 8(3), 201–207.

    Article  Google Scholar 

  • Weber, J. B., & Chapman, G. B. (2005). The combined effects of risk and time on choice: Does uncertainty eliminate the immediacy effect? Does delay eliminate the certainty effect? Organizational Behavior and Human Decision Processes, 96(2), 104–118.

    Article  Google Scholar 

  • Weitzman, M. L. (1998). Why the far-distant future should be discounted at its lowest possible rate. Journal of Environmental Economics and Management, 36(3), 201–208.

    Article  Google Scholar 

  • Weitzman, M. L. (2001). Gamma discounting. American Economic Review, 91(1), 260–271. doi:10.1257/aer.91.1.260.

    Article  Google Scholar 

  • Yaari, M. E. (1965). Uncertain lifetime, life insurance, and the theory of the consumer. The Review of Economic Studies, 32(2), 137–150.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniele Pennesi.

Appendix: Proofs

Appendix: Proofs

Proof of Theorem 1

We derive I(t) with respect to t (assuming I(t) as function from \({\mathbb {R}}_+\rightarrow {\mathbb {R}}_+\)) and we show its derivative is negative.

$$\begin{aligned} \frac{\partial I(t)}{\partial t}=\frac{E_p[\delta (\omega )^t\ln \delta (\omega )] E_p[\delta (\omega )^{t+1}]-E_p[\delta (\omega )^t]E_p[\delta (\omega )^{t+1}\ln \delta (\omega )]}{(E_p[\delta (\omega )^{t+1}])^2} \end{aligned}$$

Since \(\delta (\omega )\) are all smaller or equal than one, \(E_p[\delta (\omega )^t]\ge E_p[\delta (\omega )^{t+1}]\), hence

$$\begin{aligned}&E_p[\delta (\omega )^t\ln \delta (\omega )] E_p[\delta (\omega )^{t+1}]-E_p[\delta (\omega )^t]E_p[\delta (\omega )^{t+1}\ln \delta (\omega )]\\&\quad \le E_p[\delta (\omega )^{t}](E_p[\delta (\omega )^t\ln \delta (\omega )]-E_p[\delta (\omega )^{t+1}\ln \delta (\omega )]) \end{aligned}$$

the left-hand side of the previous inequality is equal to

$$\begin{aligned} E_p[\delta (\omega )^{t}](E_p[\delta (\omega )^t\ln \delta (\omega )(1-\delta (\omega ))]) \end{aligned}$$

Since \((1-\delta (\omega ))\ge 0\) and \(\ln \delta (\omega )\le 0\), the sign of the expression is negative. \(\square \)

Proof of Theorem 2

Time invariance and time consistency imply stationarity. By time consistency \((z,x)\sim _\tau (z,y)\), if and only if, \(x\sim _{\tau +1}y\). By time invariance \(x\sim _{\tau +1}y\), if and only if, \(x\sim _{\tau }y\).

Stationarity and time consistency imply time invariance. Without loss of generality, let \(\tau '>\tau \). By stationarity, \(x\sim _\tau y\), if and only if, \((z,x)\sim _{\tau }(z,y)\). By time consistency, \((z,x)\sim _{\tau }(z,y)\), if and only if, \(x\sim _{\tau +1}y\). Repeating the same argument until \(\tau '\) gives the result.

Time invariance and stationarity imply time consistency. By stationarity \((z,x)\sim _\tau (z,y)\), if and only if, \(x\sim _{\tau }y\). By time invariance, \(x\sim _{\tau }y\), if and only if, \(x\sim _{\tau +1}y\).   

\(\square \)

Proof of Proposition 1

If \(u_\tau \equiv u_{\tau '}\) and \(E_{p_\tau }[\delta _\tau (\omega )^{t-\tau }]=E_{p_{\tau '}}[\delta _{\tau '}(\omega )^{t-\tau '}]\), for all \(\tau ,\tau '\in {\mathbb {N}}\) time invariance follows immediately. To see the opposite implication, consider the following state-independent acts \(h=(x_0,x_1,z,z,z,\ldots )\) and \(h'=(y_0,y_1,z,z,z,\ldots )\), who differs in the firs two periods only and assume that \(h\sim _\tau h'\). Then by time invariance, \(h\sim _\tau h'\), if and only if, \(h\sim _{\tau '}h'\). Since \(u_\tau \equiv u_{\tau '}\), rescale the utilities to be equal. Then, \(u(x_0)+u(x_{1})E_{p_\tau }[\delta _\tau (\omega )]= u(y_0)+u(y_{1})E_{p_\tau }[\delta _\tau (\omega )]\), if and only if, \(u(x_0)+u(x_{1})E_{p_{\tau '}}[\delta _\tau (\omega )]= u(y_0)+u(y_{1})E_{p_{\tau '}}[\delta _{\tau '}(\omega )]\), or \(u(x_0)-u(y_0)+E_{p_\tau }[\delta _\tau (\omega )](u(x_1)-u(y_1))=u(x_0)-u(y_0)+E_{p_{\tau '}}[\delta _{\tau '}(\omega )](u(x_1)-u(y_1))\) and it follows that \(E_{p_{\tau }}[\delta _{\tau }(\omega )]=E_{p_{\tau '}}[\delta _{\tau '}(\omega )]\). \(\square \)

Proof of Theorem 3

Suppose that \(\delta _\tau (\omega )=\delta _{\tau '}(\omega )=\delta (\omega )\) for all \(\tau ,\tau '\) and all \(\omega \in \varOmega \) and \(u_\tau \equiv u_{\tau '}\) for all \(\tau ,\tau '\), and we normalize the utilities to be all equal. Then, time consistency holds if

$$\begin{aligned} \sum _{\omega \in \varOmega }p_\tau (\omega )\left[ u(z) +\sum _{s=1}^\infty \delta (\omega )^{s}u(x_{\tau +s})\right] =\sum _{\omega \in \varOmega }p_\tau (\omega )\left[ u(z) +\sum _{s=1}^\infty \delta (\omega )^{s}u(y_{\tau +s})\right] \end{aligned}$$

or

$$\begin{aligned} \sum _{s=1}^\infty u(x_{\tau +s})\left[ \,\sum _{\omega \in \varOmega }p_\tau (\omega ) \delta (\omega )^{s}\right] =\sum _{s=1}^\infty u(y_{\tau +s})\left[ \,\sum _{\omega \in \varOmega }p_\tau (\omega ) \delta (\omega )^{s}\right] \end{aligned}$$
(2)

is equivalent to:

$$\begin{aligned}&\sum _{\omega \in \varOmega }p_{\tau +1}(\omega ) \left[ u(x_{\tau +1})+\sum _{s=1}^\infty \delta (\omega )^{s}u (x_{\tau +1+s})\right] \\&\quad =\sum _{\omega \in \varOmega }p_{\tau +1} (\omega )\left[ u(y_{\tau +1})+\sum _{s=1}^\infty \delta (\omega )^{s}u(y_{\tau +s})\right] \end{aligned}$$

or

$$\begin{aligned} \sum _{s=0}^\infty u(x_{\tau +1+s})\left[ \,\sum _{\omega \in \varOmega }p_{\tau +1}(\omega )\delta (\omega )^{s}\right] =\sum _{s=0}^\infty u(y_{\tau +1+s})\left[ \,\sum _{\omega \in \varOmega }p_{\tau +1}(\omega )\delta (\omega )^{s}\right] \quad \end{aligned}$$
(3)

The left-hand side of Eq. (2) is equal to \(u(x_{\tau +1})E_{p_\tau }[\delta (\omega )]+u(x_{\tau +2})E_{p_\tau }[\delta ^2_\omega ]+\cdots \) where the left-hand side of Eq. (3) is equal to \(u(x_{\tau +1})+u(x_{\tau +2})E_{p_{\tau +1}}[\delta (\omega )]+\cdots \) Let

$$\begin{aligned} p_{\tau +1}(\omega )=\frac{p_\tau (\omega )\delta (\omega )}{\sum _{\omega '\in \varOmega }p_\tau (\omega ')\delta (\omega ')} \end{aligned}$$

then equality (3) becomes

$$\begin{aligned}&\sum _{s=0}^\infty u(x_{\tau +1+s})\left[ \,\sum _{\omega \in \varOmega }\frac{p_\tau (\omega ) \delta (\omega )^{s+1}}{\sum _{\omega '\in \varOmega }p_\tau (\omega ') \delta ({\omega '})}\right] \\&\quad =\sum _{s=0}^\infty u(y_{\tau +1+s})\left[ \,\sum _{\omega \in \varOmega }\frac{p_\tau (\omega ) \delta (\omega )^{s+1}}{\sum _{\omega '\in \varOmega }p_\tau (\omega ') \delta ({\omega '})}\right] \end{aligned}$$

that corresponds to \(u(x_{\tau +1})+u(x_{\tau +2})E_{p_\tau }[\delta (\omega )^{2}]\cdot M+\ldots =u(y_{\tau +1})+u(y_{\tau +2})E_{p_\tau }[\delta (\omega )^{2}]\cdot M+\ldots \), where \(M=[E_{p_\tau }[\delta (\omega )]]^{-1}\). Multiplying both sides by \(M^{-1}\), we have \(u(x_{\tau +1})E_{p_\tau }[\delta (\omega )]+u(x_{\tau +2})E_{p_\tau }[\delta (\omega )^{2}]+\ldots =u(y_{\tau +1})E_{p_\tau }[\delta (\omega )]+u(y_{\tau +2})E_{p_\tau }[\delta (\omega )^{2}]+\ldots \) which is equality (2). \(\square \)

Proof of Theorem 4

Suppose that \(\delta _\tau (\omega )=\delta _{\tau '}(\omega )=\delta (\omega )\) for all \(\tau ,\tau '\) and all \(\omega \in \varOmega \) and \(u_\tau \equiv u_{\tau '}\) for all \(\tau ,\tau '\), and we normalize the utilities to be all equal. Then, conditional time consistency holds if

$$\begin{aligned} \sum _{\omega \in E}p_\tau (\omega )\left[ u(z)+\sum _{s=1}^\infty \delta (\omega )^{s}u(x_{\tau +s})\right] =\sum _{\omega \in E}p_\tau (\omega )\left[ u(z)+\sum _{s=1}^\infty \delta (\omega )^{s}u(y_{\tau +s})\right] \end{aligned}$$

or

$$\begin{aligned} \sum _{s=1}^\infty u(x_{\tau +s})\left[ \,\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )^{s}\right] =\sum _{s=1}^\infty u(y_{\tau +s})\left[ \,\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )^{s}\right] \end{aligned}$$
(4)

is equivalent to:

$$\begin{aligned}&\sum _{\omega \in \varOmega }p_{\tau +1,E}(\omega )\left[ u(x_{\tau +1}) +\sum _{s=1}^\infty \delta (\omega )^{s}u(x_{\tau +1+s})\right] \\&\quad =\sum _{\omega \in \varOmega }p_{\tau +1,E}(\omega )\left[ u(y_{\tau +1})+\sum _{s=1}^\infty \delta (\omega )^{s}u(y_{\tau +s})\right] \end{aligned}$$

or

$$\begin{aligned} \sum _{s=0}^\infty u(x_{\tau +1+s})\left[ \,\sum _{\omega \in \varOmega }p_{\tau +1,E}(\omega )\delta (\omega )^{s}\right] =\sum _{s=0}^\infty u(y_{\tau +1+s})\left[ \,\sum _{\omega \in \varOmega }p_{\tau +1,E}(\omega )\delta (\omega )^{s}\right] \end{aligned}$$
(5)

The left-hand side of Eq. (4) is equal to \(u(x_{\tau +1})E_{p_\tau }[\delta (\omega )]+u(x_{\tau +2})E_{p_\tau }[\delta ^2_\omega ]+\cdots \) where the left-hand side of Eq. (5) is equal to \(u(x_{\tau +1})+u(x_{\tau +2})E_{p_{\tau +1,E}}[\delta (\omega )]+\cdots \) Let

$$\begin{aligned} p_{\tau +1,E}(\omega )=\frac{p_\tau (\omega )\delta (\omega )}{\sum _{\omega '\in E}p_\tau (\omega ')\delta (\omega ')} \end{aligned}$$

if \(\omega \in E\) and 0 otherwise. Then equality (5) becomes

$$\begin{aligned}&\sum _{s=0}^\infty u(x_{\tau +1+s})\left[ \,\sum _{\omega \in E}\frac{p_\tau (\omega )\delta (\omega )^{s+1}}{\sum _{\omega '\in E}p_\tau (\omega ')\delta ({\omega '})}\right] \\&\quad =\sum _{s=0}^\infty u(y_{\tau +1+s})\left[ \,\sum _{\omega \in E}\frac{p_\tau (\omega )\delta (\omega )^{s+1}}{\sum _{\omega '\in E}p_\tau (\omega ')\delta ({\omega '})}\right] \end{aligned}$$

that corresponds to \(u(x_{\tau +1})+u(x_{\tau +2})[\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )^{2}]\cdot M_E+\ldots =u(y_{\tau +1})+u(y_{\tau +2})[\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )^{2}]\cdot M_E+\cdots \), where \(M_E=[\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )]^{-1}\). Multiplying both sides by \(M_E^{-1}\), we have \(u(x_{\tau +1})[\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )]+u(x_{\tau +2})[\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )^2]+\cdots =u(y_{\tau +1})[\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )]+u(y_{\tau +2})[\sum _{\omega \in E}p_\tau (\omega )\delta (\omega )^2]+\cdots \) which is equality (4). \(\square \)

Proof of Theorem 1

By standard arguments and the cardinal uniqueness of each \(v_\omega (\cdot )\), the Conditional Temporal axiom implies that each conditional preference is represented by a geometrically discounted utility, hence

$$\begin{aligned} x\succcurlyeq _\omega y \Longleftrightarrow \sum _{t=0}^\infty \delta (\omega )^tu_\omega (x_t)\ge \sum _{t=0}^\infty \delta (\omega )^tu_\omega (y_t) \end{aligned}$$

for some \(\delta (\omega )\in (0,1]\) and \(u_\omega (\cdot ):[0,1]\rightarrow {\mathbb {R}}\). Now take two constant act xy with \(x\succcurlyeq _\omega y\), by Certainty consistency, \(x\succcurlyeq _{\omega '} y\) for all \(\omega '\in \varOmega \). Then, \(u_\omega (x)\frac{1}{1-\delta (\omega )}\ge u_\omega (y)\frac{1}{1-\delta (\omega )}\), if and only if, \(u_{\omega '}(x)\frac{1}{1-\delta (\omega ')}\ge u_{\omega '}(y)\frac{1}{1-\delta (\omega ')}\), or \(u_{\omega }(x)\ge u_{\omega }(y)\), if and only if, \(u_{\omega '}(x)\ge u_{\omega '}(y)\). Since it holds for any xy, \(u_{\omega }(\cdot )\) and \(u_{\omega '}(\cdot )\) represent the same preferences over X, by uniqueness up to positive affine transformation of both we can equate all the \(u_{\omega }(\cdot )=u(\cdot )\). \(\square \)

Proof of the Example 2 in Sect. 7. Suppose that \((x,t)\sim (y,t+1)\), then

$$\begin{aligned} V_S(h)=\sum _{\omega \in \varOmega }p(\omega )\ln (\delta ^t_\omega u(x))=\sum _{\omega \in \varOmega }p(\omega )\ln (\delta ^{t+1}_\omega u(y))=V_S(h') \end{aligned}$$

equivalently

$$\begin{aligned}&\sum _{\omega \in \varOmega }p(\omega )(t\ln (\delta _\omega ))+\ln u(x)=\sum _{\omega \in \varOmega }p(\omega )\left( (t+1)\ln (\delta _\omega )\right) +\ln u(y) \\&\quad \Longleftrightarrow t\sum _{\omega \in \varOmega }p(\omega )(\ln (\delta _\omega ))+\ln u(x)=(t+1)\sum _{\omega \in \varOmega }p(\omega )(\ln (\delta _\omega ))+\ln u(y)\\&\quad \Longleftrightarrow \ln u(x)=\sum _{\omega \in \varOmega }p(\omega )(\ln (\delta _\omega ))+\ln u(y)\\&\quad \Longleftrightarrow V_S(g)=V_S(g') \end{aligned}$$

I would like to thank the editor and two anonymous referees for helpful comments.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pennesi, D. Uncertain discount and hyperbolic preferences. Theory Decis 83, 315–336 (2017). https://doi.org/10.1007/s11238-017-9603-2

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11238-017-9603-2

Keywords

Navigation