Skip to main content
Log in

A simple framework for the axiomatization of exponential and quasi-hyperbolic discounting

  • Published:
Theory and Decision Aims and scope Submit manuscript

Abstract

The main goal of this paper is to investigate which normative requirements, or axioms, lead to exponential and quasi-hyperbolic forms of discounting. Exponential discounting has a well-established axiomatic foundation originally developed by Koopmans (Econometrica 28(2):287–309, 1960, 1972) and Koopmans et al. (Econometrica 32(1/2):82–100, 1964) with subsequent contributions by several other authors, including Bleichrodt et al. (J Math Psychol 52(6):341–347, 2008). The papers by Hayashi (J Econ Theory 112(2):343–352, 2003) and Olea and Strzalecki (Q J Econ 129(3):1449–1499, 2014) axiomatize quasi-hyperbolic discounting. The main contribution of this paper is to provide an alternative foundation for exponential and quasi-hyperbolic discounting, with simple, transparent axioms and relatively straightforward proofs. Using techniques by Fishburn (The foundations of expected utility. Reidel Publishing Co, Dordrecht, 1982) and Harvey (Manag Sci 32(9):1123–1139, 1986), we show that Anscombe and Aumann’s (Ann Math Stat 34(1):199–205, 1963) version of Subjective Expected Utility theory can be readily adapted to axiomatize the aforementioned types of discounting, in both finite and infinite horizon settings.

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. Pan et al. (2015) have recently provided a generalization of quasi-hyperbolic discounting to continuous time. The proposed generalization is called two-stage exponential discounting. An axiomatic foundation for this discount function is given for single dated outcomes.

  2. It should also be mentioned that our setting considers a discrete time space. A continuous time framework can be found, for example, in the above-mentioned paper by Fishburn and Rubinstein (1982) and in a generalized model of hyperbolic discounting introduced by Loewenstein and Prelec (1992). Harvey (1986) analyses discrete sequences of timed outcomes with a continuous time space.

  3. It is worth mentioning that Fishburn’s motivation for the convergence axiom B6 looks somewhat contrived in the context of acts (Fishburn 1982, p. 113). However, it becomes very natural in the context where states of the world are re-interpreted as periods of time.

  4. As pointed out above, mixture independence stated for n periods implies joint independence for n periods. Hence, this raises the obvious question of whether it is possible to use an n-period version of the subjective mixture independence axiom to obtain a time separable discounted utility representation without the need for the Debreu-type independence conditions.

References

  • Anscombe, F. J., & Aumann, R. J. (1963). A definition of subjective probability. Annals of Mathematical Statistics, 34(1), 199–205.

    Article  Google Scholar 

  • Bleichrodt, H., Rohde, K. I. M., & Wakker, P. P. (2008). Koopmans constant discounting for intertemporal choice: a simplification and a generalization. Journal of Mathematical Psychology, 52(6), 341–347.

    Article  Google Scholar 

  • Chark, R., Chew, S. H., & Zhong, S. (2015). Extended present bias: a direct experimental test. Theory and Decision, 79(1), 151–165.

    Article  Google Scholar 

  • Debreu, G. (1960). Topological methods in cardinal utility theory. In K. J. Arrow, S. Karlin, & P. Suppes (Eds.), Mathematical methods in the social sciences (pp. 16–26). Stanford: Stanford University Press.

    Google Scholar 

  • Epstein, L. G. (1983). Stationary cardinal utility and optimal growth under uncertainty. Journal of Economic Theory, 31(1), 133–152.

    Article  Google Scholar 

  • Fishburn, P. C. (1970). Utility theory for decision making. New York: Wiley.

    Google Scholar 

  • Fishburn, P. C. (1982). The foundations of expected utility. Dordrecht: Reidel Publishing Co.

    Book  Google Scholar 

  • Fishburn, P. C., & Rubinstein, A. (1982). Time preference. International Economic Review, 23(3), 677–694.

    Article  Google Scholar 

  • Gilboa, I., & Schmeidler, D. (1989). Maxmin expected utility with non-unique prior. Journal of Mathematical Economics, 18(2), 141–153.

    Article  Google Scholar 

  • Gorman, W. M. (1968). The structure of utility functions. The Review of Economic Studies, 35(4), 367–390.

    Article  Google Scholar 

  • Grandmont, J.-M. (1972). Continuity properties of a von Neumann–Morgenstern utility. Journal of Economic Theory, 4(1), 45–57.

    Article  Google Scholar 

  • Grant, S., & Van Zandt, T. (2009). Expected utility theory. In P. Anand, P. Pattanaik, & C. Puppe (Eds.), The handbook of rational and social choice (pp. 21–68). Oxford: Oxford University Press.

    Chapter  Google Scholar 

  • Gul, F. (1992). Savage’s theorem with a finite number of states. Journal of Economic Theory, 57(1), 99–110.

    Article  Google Scholar 

  • Harvey, C. M. (1986). Value functions for infinite-period planning. Management Science, 32(9), 1123–1139.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Koopmans, T. C. (1960). Stationary ordinal utility and impatience. Econometrica, 28(2), 287–309.

    Article  Google Scholar 

  • Koopmans, T. C. (1972). Representations of preference orderings over time. In C. B. McGuire & R. Radner (Eds.), Decision and organization (pp. 79–100). Amsterdam: North Holland.

    Google Scholar 

  • Koopmans, T. C., Diamond, P. A., & Williamson, R. E. (1964). Stationary utility and time perspective. Econometrica, 32(1/2), 82–100.

    Article  Google Scholar 

  • Laibson, D. (1997). Golden eggs and hyperbolic discounting. Quarterly Journal of Economics, 112(2), 443–477.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Olea, J. L. M., & Strzalecki, T. (2014). Axiomatization and measurement of quasi-hyperbolic discounting. Quarterly Journal of Economics, 129(3), 1449–1499.

    Article  Google Scholar 

  • Pan, J., Webb, C. S., & Zank, H. (2015). An extension of quasi-hyperbolic discounting to continuous time. Games and Economic Behavior, 89, 43–55.

    Article  Google Scholar 

  • Phelps, E. S., & Pollak, R. A. (1968). On second-best national saving and game-equilibrium growth. The Review of Economic Studies, 35(2), 185–199.

    Article  Google Scholar 

  • Ryan, M. J. (2009). Generalizations of SEU: A geometric tour of some non-standard models. Oxford Economic Papers, 61(2), 327–354.

    Article  Google Scholar 

  • Samuelson, P. A. (1937). A note on measurement of utility. The Review of Economic Studies, 4(2), 155–161.

    Article  Google Scholar 

  • Savage, L. J. (1954). The foundations of statistics. New York: Wiley.

    Google Scholar 

  • von Neumann, J., & Morgenstern, O. (1947). Theory of games and economic behavior. Princeton: Princeton University Press.

    Google Scholar 

  • Wakai, K. (2009). A model of utility smoothing. Econometrica, 76(1), 137–153.

    Article  Google Scholar 

  • Young, E. R. (2007). Generalized quasi-geometric discounting. Economics Letters, 96(3), 343–350.

    Article  Google Scholar 

Download references

Acknowledgments

I would like to thank my supervisor Matthew Ryan for thoughtful advice and support. I am grateful to Arkadii Slinko and Simon Grant for helpful discussions. I would also like to thank the Australian National University for their hospitality while working on the paper. Helpful comments from seminar participants at Australian National University, The University of Auckland and participants at Auckland University of Technology Mathematical Sciences Symposium (2014), Centre for Mathematical Social Sciences Summer Workshop (2014), the joint conferences “Logic, Game theory, and Social Choice 8” and “The 8th Pan-Pacific Conference on Game Theory” (2015) Academia Sinica are also gratefully acknowledged. Financial support from the University of Auckland is gratefully acknowledged. Finally, I would like to thank the anonymous reviewer for their effort and the constructive comments provided.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nina Anchugina.

Appendix: Proof of Theorem 2

Appendix: Proof of Theorem 2

Proof

Necessity of the axioms is straightforward to verify. Therefore, we will focus on the proof of sufficiency.

Step 1. Applying Theorem 1 of Fishburn (1982) to the mixture set X, it follows from Axioms I1, I3, I4 that there exists a mixture linear utility function u preserving the order on X (unique up to positive affine transformations). Normalize u so that \(u(x_0)=0\). Note that by non-triviality \(u(x_0)\) is in the interior of the non-degenerate interval u(X).

Convert streams into their utility vectors by replacing the outcomes in each period by their utility values. Define the following order: \((v_1, v_2, \ldots ) \succcurlyeq ^* (u_1, u_2, \ldots ) \Leftrightarrow \) there exist \(\mathbf{x}, \mathbf{y} \in X^{\infty }\) such that \(\mathbf{x} \succcurlyeq \mathbf{y}\) and \(u(x_t)=v_t\) and \(u(y_t)=u_t\) for every t. This order is unambiguously defined because of the monotonicity assumption, i.e., if \(x_i \sim x^{\prime }_i\) then \((x_1, \ldots , x_i, \ldots ) \sim (x_1, \ldots , x^{\prime }_i, \ldots )\).

The preference order \(\succcurlyeq ^*\) inherits the properties of weak order, mixture independence and mixture continuity from \(\succcurlyeq \). Note that \(u(X)^{\infty }\) is a mixture set under the standard operation of taking convex combinations: if \(\mathbf{v}, \mathbf{u} \in u(X)^{\infty }\) then

$$\begin{aligned} \mathbf{v}\lambda \mathbf{u}=\lambda \mathbf{v} + (1-\lambda ) \mathbf{u} \ \text {for every} \ \lambda \in (0,1). \end{aligned}$$

Therefore, by Theorem 1 of Fishburn (1982) we obtain a mixture linear representation \(U :\) \(u(X)^{\infty } \rightarrow \mathbb {R}\), where U is unique up to positive affine transformations.

Hence, \(\mathbf{v} \succcurlyeq ^* \mathbf{u}\) if and only if \(U(\mathbf{v}) \ge U(\mathbf{u})\).

Step 2. Normalize U so that \(U(0, 0, \ldots )=U(\mathbf{0})=0\). Since 0 is in the interior of u(X), and since \(U(\mathbf{v} \lambda \mathbf{0})=\lambda U(\mathbf{v})\) for any \(\mathbf{v}\in \mathbb {R}^\infty \) and for every \(\lambda \in (0, 1)\), we can assume that U is defined on \(\mathbb {R}^{\infty }\).

Mixture linearity of U implies standard linearity of U on \(\mathbb {R}^{\infty }\). To prove this, we need to show that \(U(k \mathbf{v})= kU(\mathbf{v})\) for any k and \(U(\mathbf{v}+\mathbf{u})=U(\mathbf{v})+U(\mathbf{u})\) for any \(\mathbf{u}, \mathbf{v} \in \mathbb {R}^{\infty }\).

As \(u(X)^{\infty }\) is a mixture set under the operation of taking convex combinations, \(U(\mathbf{v}k\mathbf{0})=U(k\mathbf{v}+(1-k)\mathbf{0})=U(k \mathbf{v})=k U(\mathbf{v})\) for any \(k \in (0, 1)\). If \(k>1\) then \(U(\mathbf{v})=U(\frac{k}{k}{} \mathbf{v})=\frac{1}{k}U(k\mathbf{v})\). Multiplying both parts of this equation by k, we obtain \(U(k \mathbf{v})=k U(\mathbf{v})\) for all \(k>1\). Therefore, \(U(k \mathbf{v})= kU(\mathbf{v})\) for any \(k>0\).

To prove that \(U(\mathbf{v}+\mathbf{u})=U(\mathbf{v})+U(\mathbf{u})\), consider the mixture \(\mathbf{v}\frac{1}{2}{} \mathbf{u}\). By mixture linearity of U we have:

$$\begin{aligned} U\left( \mathbf{v}\frac{1}{2}{} \mathbf{u}\right) =\frac{1}{2}U(\mathbf{v})+\frac{1}{2}U(\mathbf{u})=\frac{1}{2}\left( U(\mathbf{v})+U(\mathbf{u})\right) . \end{aligned}$$
(7)

On the other hand, \(\mathbf{v}\frac{1}{2}{} \mathbf{u}=\frac{1}{2}\mathbf{v}+\frac{1}{2}{} \mathbf{u}=\frac{1}{2}(\mathbf{v}+\mathbf{u})\). Therefore,

$$\begin{aligned} U\left( \mathbf{v}\frac{1}{2}{} \mathbf{u}\right) =U\left( \frac{1}{2}(\mathbf{v}+\mathbf{u})\right) = \frac{1}{2} U(\mathbf{v}+\mathbf{u}) \end{aligned}$$
(8)

Comparing (7) and (8), we conclude that \(U(\mathbf{v}+\mathbf{u})=U(\mathbf{v})+U(\mathbf{u})\).

Finally, note that

$$\begin{aligned} U(\mathbf{0})=U\left( \mathbf{v}+(\mathbf{-v})\right) =U(\mathbf{v})+U(\mathbf{-v})=0, \end{aligned}$$

hence \(U(\mathbf{-v})=-U(\mathbf{v})\). Therefore, if \(k<0\), then \(U(k \mathbf{v})= -kU(\mathbf{-v})=kU(\mathbf{v})\).

For each T, consider the function \(f :\mathbb {R}^T \rightarrow \mathbb {R}\) defined as follows:

$$\begin{aligned} f(v_1, \ldots , v_T)=U(v_1, \ldots , v_T, 0, 0, \ldots ). \end{aligned}$$

This function is linear on \(\mathbb {R}^T\) and it satisfies \(f(\mathbf{0})=0\); therefore,

$$\begin{aligned} f(v_1, \ldots , v_T)=\displaystyle \sum _{t=1}^T w_t^T v_t, \end{aligned}$$

where \(\mathbf{w}^T=(w^T_1, \ldots , w^T_T)\). By monotonicity \(w^T_t \ge 0\) for all \(t \le T\).

Note that \(w^T_t=U([1]_t)\), where \([1]_t\) is the vector with 1 in period t and 0 elsewhere. It follows that \(w^{T}_t=w^{T^{\prime }}_t\) for any T and \(T^{\prime }\). Hence, there is a vector \(\mathbf{w} \in \mathbb {R}^\infty \) such that \(U(v_1, \ldots , v_T, 0, 0, \ldots )=\sum \nolimits _{t=1}^{\infty } w_t v_t\) for any \((v_1, \ldots , v_T)\in \mathbb {R}^T\).

Recalling that \(v_t=u(x_t)\), we obtain

$$\begin{aligned} U(u(x_1), \ldots , u(x_T), 0, 0, \ldots )=\sum _{t=1}^T w_t u(x_t)\quad \text { for all } \mathbf{x} \in X^{*}. \end{aligned}$$

Therefore, for every \(\mathbf{x}, \mathbf{y} \in X^*\) we have \(\mathbf{x} \succcurlyeq \mathbf{y}\) if and only if

$$\begin{aligned} \sum _{t=1}^T w_t u(x_t) \ge \sum _{t=1}^T w_t u(y_t). \end{aligned}$$

By slightly abusing the notation, re-define U so that:

$$\begin{aligned} U(x_1, \ldots , x_T, x_0, x_0, \ldots )=\sum _{t=1}^T w_t u(x_t) \quad \text { for all } \mathbf{x} \in X^{*}. \end{aligned}$$

Hence, \(U(\mathbf{x})=\sum \nolimits _{t=1}^{\infty } w_t u(x_t)\) represents preferences on \(X^*\).

Step 3. Next, we show that \(U(x_1, x_2, \ldots )\) converges for any \((x_1, x_2, \ldots )\). Define \(U_T:X^\infty \rightarrow {\mathbb R}\) as follows: \(U_T(\mathbf{x}) = \sum \nolimits _{t=1}^T u_t(x_t)\), where \(u_t(x_t)=w_t u(x_t)\). Consider the sequence of functions \(U_1, U_2, \ldots , U_T, \ldots \) According to the Cauchy Criterion, a sequence of functions \(U_T(\mathbf{x})\) defined on \(X^\infty \) converges on \(X^\infty \) if and only if for any \(\varepsilon >0\) and any \(\mathbf{x} \in X^\infty \) there exists \(T \in \mathbb {N}\) such that \(|U_N(\mathbf{x}) - U_M(\mathbf{x}) |< \varepsilon \) for any \(N, M \ge T\).

Fix some \(\mathbf{x}\in X^\infty \) and \(\varepsilon >0\). Suppose that for some k it is possible to choose \(x^+, x^-\) such that \([x^+]_k \succ [x_k]_k \succ [x^-]_k\). By Step 2 the preference \([x^+]_k \succ [x_k]_k \succ [x^-]_k\) implies that \(w_k>0\). Therefore, as u is a continuous function, it is without loss of generality to assume that

$$\begin{aligned} u_k(x^+)-u_k(x_k)<\varepsilon /2 \text { and } u_k(x_k)-u_k(x^-)<\varepsilon /2. \end{aligned}$$

It follows that \(u_k(x^+)-u_k(x^-)<\varepsilon , \) or \(u_k(x^-)-u_k(x^+)>-\varepsilon .\) By Axiom I6 there exist \(T^+\) and \(T^-\) satisfying \(k\le \min \{T^-, T^+\}\) such that

$$\begin{aligned} \mathbf{x}_{k, N}^+ \succcurlyeq \mathbf{x} \succcurlyeq \mathbf{x}_{k, M}^-, \ \text {for all} \ N\ge T^+, \ M\ge T^-. \end{aligned}$$

Let \(T^*=\max \{T^-, T^+\}\). It is necessary to demonstrate that \(|U_N(\mathbf{x}) - U_M(\mathbf{x}) |< \varepsilon \) for any \(N, M \ge T^*\). If \(N=M\) the result is obviously true. If \(N\ne M\) then it is without loss of generality to assume that \(N>M\). By the additive representation:

$$\begin{aligned} U(\mathbf{x}_{k,N}^+) \ge U(\mathbf{x}_{k,M}^-). \end{aligned}$$

Expanding

$$\begin{aligned} u_k(x^+)+\sum _{t=1, t \ne k}^N u_t(x_t) \ge u_k(x^-)+\sum _{t=1, t \ne k}^M u_t(x_t). \end{aligned}$$

By rearranging this inequality

$$\begin{aligned} \sum _{t=M+1}^N u_t(x_t) \ge u_k(x^-)-u_k(x^+) > -\varepsilon . \end{aligned}$$

As \(N>M \ge T^*\) it is also true that \(U(\mathbf{x}_{k, M}^+) \ge U(\mathbf{x}_{k, N}^-\)), hence

$$\begin{aligned} \sum _{t=M+1}^N u_t(x_t) \le u_k(x^+) - u_k(x^-) < \varepsilon . \end{aligned}$$

Note that

$$\begin{aligned} \sum _{t=M+1}^N u_t(x_t) = U_N(\mathbf{x})-U_M(\mathbf{x}). \end{aligned}$$

Hence, \(|U_N(\mathbf{x}) - U_M(\mathbf{x}) |< \varepsilon \) and it follows that \(U(\mathbf{x})\) converges by the Cauchy criterion.

Suppose now that it is not possible to find such k that \([x^+]_k \succ [x_k]_k \succ [x^-]_k \) for some \(x^+, x^- \in X\). If \(w_t=0\) for all t then the result is trivial. Suppose that \(w_t>0\) for some t. Then for every period t for which \(w_t>0\) we have

$$\begin{aligned} x_t \in X^e \equiv \{z \in X \ :\ z\succcurlyeq z' \ \text {for all} \ z'\in X \, \text {or} \ z'\succcurlyeq z \ \text {for all} \ z'\in X \}. \end{aligned}$$

For some \(\lambda \in (0,1)\) replace \(x_t\) with the mixture \(x_t \lambda x_0\) for each t. Call the resulting stream \(\mathbf{x}^*\). Then

$$\begin{aligned} U_T(\mathbf{x})-U_T(\mathbf{x}^*)=\sum _{t=1}^T u_t(x_t)-\sum _{t=1}^T u_t(x_t \lambda x_0)=(1-\lambda )\sum _{t=1}^T u_t(x_t)=(1-\lambda )U_T(\mathbf{x}). \end{aligned}$$

By rearranging this equation it follows that \(U_T(\mathbf{x}^*)=\lambda U_T(\mathbf{x})\). By the previous argument \(U_T(\mathbf{x}^*)\) converges; therefore, \(U_T(\mathbf{x})\) converges.

Step 4. Show that \(U(\mathbf{x})\) represents the order on \(X^\infty \). Suppose that \(\mathbf{x} \succcurlyeq \mathbf{y}\), where \(\mathbf{x}, \mathbf{y} \in X^\infty \). If for some kj it is possible to find \(x^+, y^-\) such that \([x^+]_k \succ [x_k]_k\) and \([y^-]_j \prec [y_j]_j\), then \([x^+ \lambda x_k]_k \succ [x_k]_k\) for every \(\lambda \in (0, 1)\) and \([y^- \mu y_j]_j \prec [y_j]_j\) for every \(\mu \in (0,1)\). Let \(x^*=x^+ \lambda x_k\) and \(y^*=y^- \mu y_j\) for some \(\lambda , \mu \in (0, 1)\). Denote

$$\begin{aligned} \mathbf{x}_{k,N}^*=(x_1, \ldots , x_{k-1}, x^*, x_{k+1}, \ldots , x_N, x_0, x_0, \ldots ), \end{aligned}$$

and

$$\begin{aligned} \mathbf{y}_{j,M}^*=(y_1, \ldots , y_{j-1}, y^*, y_{j+1}, \ldots , y_M, x_0, x_0, \ldots ). \end{aligned}$$

Then by Axiom I6, there exist \(T^-, T^+\) such that

$$\begin{aligned} \mathbf{x}_{k,N}^*\succcurlyeq \mathbf{x}\succcurlyeq \mathbf{y} \succcurlyeq \mathbf{y}_{j,M}^*\ \end{aligned}$$

for all \(N\ge T^+\) and for all \(M\ge T^{-}\). Since \(\mathbf{x}_{k,N}^* \succcurlyeq \mathbf{y}_{j,M}^*\) and U represents \(\succcurlyeq \) on \(X^*\) we have:

$$\begin{aligned} U(\mathbf{x}_{k,N}^*)\ge U(\mathbf{y}_{j,M}^*). \end{aligned}$$

By Step 3 we know that \(U(x_1, \ldots , x_{k-1}, x^*, x_{k+1} \ldots )\) and \(U(y_1, \ldots , y_{j-1}, y^*, y_{j+1}, \ldots )\) converge, so

$$\begin{aligned} U(x_1, \ldots , x_{k-1}, x^*, x_{k+1}, \ldots ) \ge U(y_1, \ldots , y_{j-1}, y^*, y_{j+1}, \ldots ). \end{aligned}$$

Recall that \(x^*=x^+ \lambda x_k\) and \(y^*=y^- \mu y_j\) for some \(\lambda \in (0,1)\) and some \(\mu \in (0,1)\). Since \(\lambda \) and \(\mu \) are arbitrary, it follows that \(U(\mathbf{x}) \ge U(\mathbf{y})\).

If it is not possible to find \(x^+\), \(y^-\) such that \([x^+]_k \succ [x_k]_k\) and \([y^-]_j \prec [y_j]_j\), then either \(w_t=0\) for all t, in which case \(U(\mathbf{x})=U(\mathbf{y})\); or \(x_t \succcurlyeq z'\) for all \(z'\in X\) and all t with \(w_t>0\), in which case \(U(\mathbf{x}) \ge U(\mathbf{y})\); or \(z' \succcurlyeq y_t\) for all \(z'\in X\) and all t with \(w_t>0\) in which case \(U(\mathbf{x}) \ge U(\mathbf{y})\).

It is worth noting that as \(\mathbf{x} \succcurlyeq \mathbf{y}\) implies \(U(\mathbf{x}) \ge U(\mathbf{y})\), then by Axiom I2 it follows that \(w_t>0\) for at least one t. Therefore, \(\sum _{t=1}^\infty w_t>0\). Normalizing by \(1/\sum _{t=1}^\infty w_t\), we can assume that \(\sum _{t=1}^\infty w_t=1\).

Next, assume that \(U(\mathbf{x}) \ge U(\mathbf{y})\). Suppose that it is possible to find k and \(x^+, x^- \in X\) such that \(\mathbf{x}_{k, N}^+\succcurlyeq \mathbf{x}\succcurlyeq \mathbf{x}_{k, N}^-\) for some fixed N. By mixture continuity, the set \(\{ \alpha :\mathbf{x}_{k,N}^+ \alpha \mathbf{x}_{k, N}^- \succcurlyeq \mathbf {x}\}\) is closed. By assumption \(\mathbf{x}_{k, N}^+\succcurlyeq \mathbf{x}\), so it follows that \(\alpha = 1\) is included into the set. Analogously, the set \(\{ \beta :\mathbf{x} \succcurlyeq \mathbf{x}_{k,N}^+ \beta \mathbf{x}_{k,N}^-\}\) is closed. In fact, \(\beta = 0\) belongs to the set, as \( \mathbf{x} \succcurlyeq \mathbf{x}_{k, N}^-\). Therefore, as both sets are closed, nonempty and form the unit interval, their intersection is nonempty. Hence, there exists \(\lambda \) such that \(\mathbf{x} \sim \mathbf{x}_{k, N}^+ \lambda \mathbf{x}_{k, N}^-\). Note that \(\mathbf{x}_{k, N}^+ \lambda \mathbf{x}_{k,N}^- = (x_1, \ldots , x_{k-1}, x^+ \lambda x^-, x_{k+1} \ldots , x_N, x_0, x_0, \ldots )\). Let \(x^+ \lambda x^- = x^*\). Define \(\mathbf{x}_{k, N}^*= (x_1, \ldots ,x_{k-1}, x^*, x_{k+1}, \ldots , x_N, x_0, x_0, \ldots )\). Therefore, if there exist periods kj and outcomes \(x^+, x^-, y^+, y^-\in X\) such that \(\mathbf{x}_{k, N}^+ \succcurlyeq \mathbf{x} \succcurlyeq \mathbf{x}_{k, N}^-\) and \(\mathbf{y}_{j, M}^+ \succcurlyeq \mathbf{y} \succcurlyeq \mathbf{y}_{j, M}^-\) for some N and some M, we can find \(\lambda , \mu \in [0, 1]\) such that \(\mathbf{x} \sim \mathbf{x}_{k, N}^*\) and

$$\begin{aligned} \mathbf{y} \sim \mathbf{y}_{j, M}^* = (y_1, \ldots , y_{j-1}, y^*,y_{j+1}, \ldots , y_M, x_0, x_0, \ldots ), \end{aligned}$$

where \(y^*=y^+\mu y^-\). We have already shown that if \(\mathbf{x}\succcurlyeq \mathbf{y}\) then \(U(\mathbf{x}) \ge U(\mathbf{y})\). From \(\mathbf{x} \sim \mathbf{x}_{k, N}^*\) and \(\mathbf{y} \sim \mathbf{y}_{j, M}^*\) it, therefore, follows that:

$$\begin{aligned} U(\mathbf{x}_{k, N}^*)= U(\mathbf{x}) \text { and } U(\mathbf{y}_{j, M}^*)= U(\mathbf{y}). \end{aligned}$$

Hence, from the assumption \(U(\mathbf{x}) \ge U(\mathbf{y})\) we obtain:

$$\begin{aligned} U(\mathbf{x}_{k, N}^*) \ge U(\mathbf{y}_{j, M}^*). \end{aligned}$$

Recall that U is an order-preserving function on \(X^*\). Thus, \(\mathbf{x}_{k, N}^* \succcurlyeq \mathbf{y}_{j, M}^*\). Since \(\mathbf{x} \sim \mathbf{x}_{k, N}^*\) and \(\mathbf{y} \sim \mathbf{y}_{j, M}^*\), we obtain \(\mathbf{x} \succcurlyeq \mathbf{y}\).

Suppose now that there is no such kj or outcomes \(x^+, x^-, y^+, y^-\) such that \(\mathbf{x}_{k, N}^+\succcurlyeq \mathbf{x}\succcurlyeq \mathbf{x}_{k, N}^-\) and \(\mathbf{y}_{j, M}^+\succcurlyeq \mathbf{y}\succcurlyeq \mathbf{y}_{j, M}^-\) for some N and some M. Then, using Axiom I6, we can conclude that either \(x_t \in X^e\) for every t with \(w_t>0\) or \(y_t\in X^e\) for every t with \(w_t>0\). Assume that there is only an upper bound to preferences; i.e., \(X^e \equiv \{z \in X \ :\ z \succcurlyeq z' \ \text {for every} \ z'\in X\}\). Then \(U(\mathbf{x}) \ge U(\mathbf{y})\) means that \(x_t\in X^e\) whenever \(w_t>0\). Therefore, \(U(\mathbf{x})=U({\overline{\mathbf{x}}})\), where \({\overline{\mathbf{x}}}=(\overline{x}, \overline{x}, \ldots )\) and \(\overline{x} \in X^e\). Hence, it follows by monotonicity that \(\mathbf{x}\succcurlyeq \mathbf{y}\). In the case when there is only a lower bound, i.e., \(\underline{x} \in X^e \equiv \{z \in X \ :\ z' \succcurlyeq z \ \text {for every} \ z'\in X\}\), the argument is similar.

Next, suppose that X is preference bounded above and below, i.e., there exist \(\underline{x},\overline{x} \in X^e\) with \(\overline{x}\succcurlyeq x \succcurlyeq \underline{x}\) for every \(x\in X\). Assume that \(U(\mathbf{x}) \ge U(\mathbf{y})\). We need to demonstrate that \(\mathbf{x} \succcurlyeq \mathbf{y}\). By monotonicity and continuity there exist \(\lambda , \mu \in [0, 1]\) such that \(\mathbf{x} \sim {\overline{\mathbf{x}} }\lambda {\underline{\mathbf{x}}}\) and \(\mathbf{y} \sim {\overline{\mathbf{x}}}\mu {\underline{\mathbf{x}}}\). Since by assumption \(U(\mathbf{x}) \ge U(\mathbf{y})\) and U represents the preference order on constant streams, we have \(U({\overline{\mathbf{x}} }\lambda { \underline{\mathbf{x}}}) \ge U({\overline{\mathbf{x}} }\mu {\underline{\mathbf{x}}})\). By rearranging this inequality \((\lambda -\mu )(U({ \overline{\mathbf{x}}})-U({\underline{\mathbf{x}}}))\), and using \(U({ \overline{\mathbf{x}}}) > U({\underline{\mathbf{x}}})\) it follows that \(\lambda \ge \mu \). Therefore, as \(\mathbf{x} \sim {\overline{\mathbf{x}}}\lambda {\underline{\mathbf{x}}}\) and \(\mathbf{y} \sim {\overline{\mathbf{x}} }\mu { \underline{\mathbf{x}}}\) and \(\lambda \ge \mu \), we conclude that \(\mathbf{x} \succcurlyeq \mathbf{y}\).

Thus, \((\mathbf{w}, u)\) is an AA representation for \(\succcurlyeq \).

Step 5. Uniqueness of \(w_t\). Assume that \((\mathbf{w}^\prime , u^\prime )\) is another AA representation. Then, for any t we have \(w_t>0\) if and only if \(w^\prime _t>0\). Consider the set of all constant programs \(\{\mathbf{x}\in X^\infty : \mathbf{x} = (a, a, \ldots ), \text { where } a\in X\}\), which is a mixture set. Applying \((\mathbf{w}^\prime , u^\prime )\) and \((\mathbf{w}, u)\) to this set we conclude that \(u(a)>u(b)\) if and only if \(u^\prime (a)>u^\prime (b)\) for every \(a, b \in X\). By Theorem 1 Fishburn (1982) it implies that \(u=A u^\prime +B\) for some \(A>0\) and some B. Hence,

$$\begin{aligned} \sum _{t=1}^ \infty w_t u(x_t) \ge \sum _{t=1}^ \infty w_t u(y_t) \ \text {if and only if} \ \sum _{t=1}^ \infty w^\prime _t u(x_t) \ge \sum _{t=1}^ \infty w^\prime _t u(y_t). \end{aligned}$$

For any ts with \(t \ne s\) and any \(x', x'' \in X\), let \([x',x'']_{t,s}\) denote the stream with \(x'\) in the tth position, \(x''\) in the sth position and \(x_0\) elsewhere. Fix ts with \(w_t>0\) and \(w_s>0\). Using non-triviality, choose some \(x^+, x^- \in X\) such that \(x^+ \succ x^-\). Define \(\mathbf{x}=[x^+, x^+]_{t,s}, \ \mathbf{y}= [x^+, x^-]_{t,s}, \ \mathbf{z}= [x^-, x^-]_{t,s}\). From the AA representation it follows that \(\mathbf{x} \succ \mathbf{y} \succ \mathbf{z}\). By continuity of the AA representation there exists \(\lambda \in (0, 1)\) such that \(\mathbf{y} \sim \mathbf{x} \lambda \mathbf{z}\). Applying the AA representation to \(\mathbf{y} \sim \mathbf{x} \lambda \mathbf{z}\) we obtain

$$\begin{aligned} w_t u(x^+)+w_s u(x^-) =\lambda (w_t+w_s) u(x^+)+(1-\lambda ) (w_t+w_s) u(x^-). \end{aligned}$$

It follows that \((1-\lambda )w_t = \lambda w_s\). Similarly, \((1-\lambda )w^\prime _t = \lambda w^\prime _s\). Therefore, \(w_t/w_s =w^\prime _t/w^\prime _s\). As this is true for any ts, we obtain that \(\mathbf{w}=C {\mathbf{w}}^\prime \) for some \(C>0\). The sufficiency of the uniqueness conditions follows by routine arguments. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Anchugina, N. A simple framework for the axiomatization of exponential and quasi-hyperbolic discounting. Theory Decis 82, 185–210 (2017). https://doi.org/10.1007/s11238-016-9566-8

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11238-016-9566-8

Keywords

JEL Classification

Navigation