Skip to main content
Log in

Risk lovers and the rent over-investment puzzle

  • Published:
Public Choice Aims and scope Submit manuscript

Abstract

In this paper, we prove existence and uniqueness of equilibrium in a rent-seeking contest given a class of heterogeneous risk-loving players. We explore the role third-order risk attitude plays in equilibrium and find that imprudence is sufficient for risk lovers to increase rent-seeking investment above the risk-neutral outcome. Moreover, we show that rent can be fully dissipated in a standard Tullock contest played by a large number of risk-lovers.

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 Congleton et al. (2008) or Konrad (2009) for an extensive review.

  2. The term “over-investment” used in this paper is different from the winner’s curse common to auction literature. See also footnote 4.

  3. Menezes et al. (1980) make the initial reference to this third-order effect and name it downside risk aversion. Their description suggests that downside-risk-averse agents dislike a transfer of risk from higher to lower levels of wealth.

  4. Note the difference between the terms “over-investment” and “over-dissipation”. In this paper, over-investment refers to an individual’s rent-seeking investment that exceeds the risk-neutral prediction, while over-dissipation refers to an outcome where participants spend more in the aggregate than the value of the available rent. Using simple numerical analysis, Tullock (1980) shows the conditions under which rents are over-dissipated, namely non-linear rent-seeking production functions and the number of risk-neutral players.

  5. Recent experimental studies suggest that risk-lovingness may help explain over-bidding in auctions; see Gneezy and Smorodinsky (2006) and Platt et al. (2013).

  6. Also known as the self-protection problem, first studied by Ehrlich and Becker (1972).

  7. Cornes and Hartley (2012) conjecture that DARA players may induce a unique equilibrium. Yamazaki (2009) also considers the DARA case without the assumption of small rent value. However, we discover a flaw in his proof that restricts the existence of equilibrium to some special cases. See Appendix 2 for further discussion.

  8. See Abbink et al. (2010) for an example.

  9. In a standard Tullock contest, the production function is linear and the number of players and the value of the prize, which is common knowledge, are fixed. See Tullock (1980) and Higgins et al (1985) for a treatment of both linear and non-linear Tullock contests.

  10. For a graphical representation of this approach, see Tullock (1975).

  11. In fact, convexity of indifference curves is not necessary for the existence of player i’s best response function given a concave constraint as in (5). For a risk averter with CARA utility functions, indifference curves are concave yet Cornes and Hartley (2003) prove that the objective function is strictly quasiconcave and therefore a global maximum exists. In the proof of Lemma 2, we show, despite the fact that u i is strictly convex in wealth, that Eu i is strictly quasiconcave in y i and player i’s best response function exists.

  12. This is a special case of the first-order condition for each player’s maximization problem derived below in (26) with s i (Y e) = 1/n for i = 1,…,n.

  13. Consider (10). If we plot u′ and draw a chord connecting u′(I − g(y)) and u′(I − g(y) + R), we find that the area under the chord, i.e., the denominator in the brackets, is smaller than the area under u′, i.e., the numerator in the brackets for any y if and only if u′ is concave, i.e., u′′′ < 0, regardless of the sign of u″.

  14. See Konrad and Schlesinger (1997) for a detailed discussion. Treich (2010) considers only prudent risk-averters.

  15. See Crainich et al. (2013) and Ebert (2013) for properties of prudent risk-lovers.

  16. This result still holds in some generalized contests. If each player’s production function has \({f_{i}}^{\prime} \,(0) > 1\), then we will find g i ′(0) < 1 and the corresponding threshold value κ will be larger than in (12).

References

  • Abbink, K., Brandts, J., Herrmann, B., & Orzen, H. (2010). Intergroup conflict and intra-group punishment in an experimental contest game. American Economic Review, 100, 420–447.

    Article  Google Scholar 

  • Congleton, R., Hillman, A., & Konrad, K. (Eds.). (2008). 40 years of research on rent seeking 1: Theory of rent seeking. Heidelberg: Springer.

    Google Scholar 

  • Cornes, R., & Hartley, R. (2003). Risk aversion, heterogeneity and contests. Public Choice, 117, 1–25.

    Article  Google Scholar 

  • Cornes, R., & Hartley, R. (2012). Risk aversion in symmetric and asymmetric contests. Economic Theory, 51, 247–275.

    Article  Google Scholar 

  • Crainich, D., Eeckhoudt, L., & Trannoy, A. (2013). Even (mixed) risk lovers are prudent. American Economic Review, 103, 1529–1535.

    Article  Google Scholar 

  • Deck, C., & Schlesinger, H. (2010). Exploring higher order risk effects. Review of Economic Studies, 77, 1403–1420.

    Article  Google Scholar 

  • Ebert, S. (2013). Even (mixed) risk lovers are prudent: Comment. American Economic Review, 103, 1536–1537.

    Article  Google Scholar 

  • Eeckhoudt, L., & Gollier, C. (2005). The impact of prudence on optimal prevention. Economic Theory, 26, 989–994.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Ehrlich, I., & Becker, G. (1972). Market insurance, self-insurance, and self-protection. Journal of Political Economy, 80, 623–648.

    Article  Google Scholar 

  • Gneezy, U., & Smorodinsky, R. (2006). All-pay auctions–an experimental study. Journal of Economic Behavior & Organization, 61, 255–275.

    Article  Google Scholar 

  • Higgins, R., Shughart, W., & Tollison, R. (1985). Free entry and efficient rent seeking. Public Choice, 46, 247–258.

    Article  Google Scholar 

  • Hillman, A. L., & Katz, E. (1984). Risk-averse rent seekers and the social cost of monopoly power. Economic Journal, 94, 104–110.

    Article  Google Scholar 

  • Houser, D., & Stratmann, T. (2012). Gordon Tullock and experimental economics. Public Choice, 152, 211–222.

    Article  Google Scholar 

  • Jindapon, P. (2013). Do risk lovers invest in self-protection? Economics Letters, 121, 290–293.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Konrad, K. A. (2009). Strategy and dynamics in contests. Oxford: Oxford University Press.

    Google Scholar 

  • Konrad, K., & Schlesinger, H. (1997). Risk aversion in rent seeking and rent augmenting games. Economic Journal, 107, 1671–1683.

    Article  Google Scholar 

  • Lim, W., Matros, A., & Turocy, T. L. (2014). Bounded rationality and group size in Tullock contests: Experimental evidence. Journal of Economic Behavior & Organization, 99, 155–167.

    Article  Google Scholar 

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

    Google Scholar 

  • Noussair, C., Trautmann, S., & Van de Kuilen, G. (2014). Higher order risk attitudes, demographics, and financial decisions. Review of Economic Studies, 81, 325–355.

    Article  Google Scholar 

  • Pe˘carić, J., Proschan, F., & Tong, Y. (1992). Convex functions, partial orderings, and statistical applications. New York: Academic Press.

    Google Scholar 

  • Platt, B. C., Price, J., & Teppan, H. (2013). The role of risk preferences in pay-to-bid auctions. Management Science, 59, 2117–2134.

    Article  Google Scholar 

  • Sheremeta, R. (2013). Overbidding and heterogeneous behavior in contest experiments. Journal of Economic Surveys, 27, 491–514.

    Article  Google Scholar 

  • Szidarowszky, F., & Okuguchi, K. (1997). On the existence and uniqueness of pure Nash equilibrium in rent-seeking games. Games and Economics Behavior, 18, 135–140.

    Article  Google Scholar 

  • Tarazona-Gomez, M. (2004). Are individuals prudent? An experimental approach using lotteries. University of Toulouse Working Paper.

  • Treich, N. (2010). Risk-aversion and prudence in rent-seeking games. Public Choice, 145, 339–349.

    Article  Google Scholar 

  • Tullock, G. (1975). On the efficient organization of trials. Kyklos, 28, 745–762.

    Article  Google Scholar 

  • Tullock, G. (1980). Efficient rent-seeking. In J. M. Buchanan, R. D. Tollison, & G. Tullock (Eds.), Toward a theory of the rent-seeking society (pp. 97–112). College Station: Texas A & M University Press.

    Google Scholar 

  • Yamazaki, T. (2009). The uniqueness of pure-strategy Nash equilibrium in rent-seeking games with risk-averse players. Public Choice, 139, 335–342.

    Article  Google Scholar 

Download references

Acknowledgments

We thank Klaus Abbink, Paul Pecorino, Ray Rees, Harris Schlesinger, Richard Watt, and seminar participants at University of Alabama, University of Canterbury, and the 2013 Public Choice Society Annual Meeting for valuable comments. We acknowledge Culverhouse College of Commerce for financial support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paan Jindapon.

Appendices

Appendix 1: Proofs

1.1 Proof of Lemma 1

The slope of indifference curve q i (y i | \(\bar{U}\)) is given by

$$q_{i} '\left( {y_{i} |\bar{U}} \right) = [q_{i} \phi_{i} (I_{i} - g_{i} \left( {y_{i} )} \right) + (1 - q_{i} )\psi_{i} \left( {I_{i} - g_{i} \left( {y_{i} )} \right)} \right]g_{i}^{\prime} \left( {y_{i} } \right),$$
(14)

where

$$\phi_{i} = \frac{{u_{i}^{\prime} \left( {w + R} \right)}}{{u_{i} (w + R) - u_{i} (w)}}$$
(15)

and

$$\psi_{i} = \frac{{u_{i}^{\prime} (w)}}{{u_{i} (w + R) - u_{i} (w)}} .$$
(16)

Thus, under general conditions, the indifference curve is always upward sloping. By letting v i : = I i  − g i (y i ), we derive the second derivative of q i (y i | \(\bar{U}\)) with respect to y i from (14) as

$$q_{i} ''(y_{i} |\bar{U}) \, = \, \left[ {q_{i} \phi_{i} \left( {v_{i} } \right) \, + \, \left( { 1- q_{i} } \right)\psi_{i} \left( {v_{i} } \right)} \right]g_{i} ''\left( {y_{i} } \right) \, + \, \left[ {q_{i} \varPhi_{i} \left( {v_{i} } \right) \, + \, \left( { 1- q_{i} } \right)\varPsi_{i} \left( {v_{i} } \right)} \right]\left[ {g_{i} '\left( {y_{i} } \right)} \right]^{ 2} ,$$
(17)

where

$$\varPhi_{i} \left( w \right) = \phi_{i} \left( w \right)\left[ {\phi_{i} \left( w \right) - \psi_{i} \left( w \right)} \right] - \phi_{i} '(w)$$
(18)

and

$$\varPsi_{i} \left( w \right) = \psi_{i} \left( w \right)\left[ {\phi_{i} \left( w \right) - \psi_{i} \left( w \right)} \right] - \psi_{i} '(w) .$$
(19)

The first term on the right-hand side of (17) is positive; however, the sign of the second term is uncertain. To guarantee that it is positive we assume (i) φ i (w) ≤ φ i (w)[φ i (w) − ψ i (w)] and (ii) ψ i (w) ≤ ψ i (w)[φ i (w) − ψ i (w)] for all w ∈ (I i  − R,I i ). Even though Conditions (i) and (ii) are stated in terms of φ i and ψ i instead of u i , their intuition is quite simple. Using (15) and (16), we can write conditions (i) and (ii) as

$$\frac{{u_{i}^{''} (w + R)}}{{u_{i}^{\prime} (w + R)}} \le 2\frac{{u_{i}^{\prime} \left( {w + R} \right) - u_{i}^{\prime} (w)}}{{u_{i} \left( {w + R} \right) - u_{i} (w)}}$$
(20)

and

$$\frac{{u_{i}^{''} (w)}}{{u_{i}^{\prime} (w)}} \le 2\frac{{u_{i}^{\prime} \left( {w + R} \right) - u_{i}^{\prime} (w)}}{{u_{i} \left( {w + R} \right) - u_{i} (w)}}$$
(21)

for all w ∈ (I i  − R,I i ) respectively. Thus, both inequalities can be satisfied if

$$\frac{{u_{i}^{''} (w)}}{{u_{i}^{\prime} (w)}} \le 2\frac{{u_{i}^{''} (z)}}{{u_{i}^{\prime} (z)}}$$
(22)

for all w, z ∈ (I i  − R,I i  + R).

1.2 Proof of Lemma 2

Part (i) Given the optimization problem in (3), an interior solution y * i satisfies the first-order condition:

$$\begin{aligned} &F\left( {y_{i}^{*} } \right)\text{ := }\frac{{Y_{ - i} }}{{\left( {y_{i}^{*} + Y_{ - i} } \right)^{2} }}[u_{i} \left( {I_{i} - g_{i} (y_{i}^{*} ) + R} \right) - u_{i} \left( {I_{i} - g_{i} (y_{i}^{*} )} \right)] \hfill \\& - \left[ {\left( {\frac{{y_{i}^{*} }}{{y_{i}^{*} + Y_{ - i} }}} \right)u^{\prime}_{i} \left( {I_{i} - g_{i} \left( {y_{i}^{*} } \right) + R} \right) + \left( {\frac{{Y_{ - i} }}{{y_{i}^{*} + Y_{ - i} }}} \right)u^{\prime}_{i} \left( {I_{i} - g_{i} \left( {y_{i}^{*} } \right)} \right)} \right]g'_{i} \left( {y_{i}^{*} } \right) = 0. \hfill \\ \end{aligned}$$
(23)

For the equivalent constrained-optimization problem given by (4) and (5), an interior solution (\(y_{i}^{*},p_{i}^{*}\)) satisfies the tangency condition and the constraint:

$$\frac{{Y_{ - i} }}{{\left( {y_{i}^{*} + Y_{ - i} } \right)^{2} }} - [p_{i}^{*} \phi_{i} \left( {v_{i}^{*} } \right) + \left( {1 - p_{i}^{*} } \right)\psi_{i} \left( {v_{i}^{*} )} \right]g_{i}^{\prime} \left( {y_{i} } \right) = 0$$
(24)
$$p_{i}^{*} - \frac{{y_{i}^{*} }}{{y_{i}^{*} + Y_{ - i} }} = 0 ,$$
(25)

where \(v_{i}^{*}\) = I i g i (\(y_{i}^{*}\)) and φ i and ψ i are defined in (15) and (16), respectively. Substituting \(p_{i}^{*}\) from (25) into (24) yields

$$G\left( {y_{i}^{*} } \right)\text{ := }\frac{{Y_{ - i} }}{{\left( {y_{i}^{*} + Y_{ - i} } \right)^{2} }} - \left[ {\left( {\frac{{y_{i}^{*} }}{{y_{i}^{*} + Y_{ - i} }}} \right)\phi_{i} \left( {v_{i}^{*} } \right) + \left( {\frac{{Y_{ - i} }}{{y_{i}^{*} + Y_{ - i} }}} \right)\psi_{i} (v_{i}^{*} )} \right]g'_{i} \left( {y_{i}^{*} } \right) = 0 .$$
(26)

Note that G(y i ) is the difference between the slope of the constraint and the slope of indifference curve at (y i , p(y i |Y i )). Since the former is decreasing (in fact, strictly decreasing when Y i  > 0) and the latter is strictly increasing (see Lemma 1), G(y i ) is strictly decreasing in y i . If Y i is less than a threshold (derived in part (ii) below), G(y i ) > 0 when y i  = 0. Moreover, we find that G(y i ) < 0 when y i is large enough. There consequently exists only one value \(y_{i}^{*}\) such that \(G\left( {y_{i}^{*} } \right) = 0\).

Next, we define

$$\Delta (y_{i}^{ * } ) \, : = u_{i} (I_{i} - g_{i} (y_{i}^{ * } ) \, + R) - u_{i} (I_{i} - g_{i} (y_{i}^{ * } )).$$
(27)

Since u′ and u″ are strictly positive, we know that ∆(y i ) is strictly positive and strictly decreasing in y i . Given the definitions of G(y i ) and ∆(y i ), we can write the first-order condition in (23) as

$$F(y_{i}^{ * } ) \, = G(y_{i}^{ * } )\varDelta (y_{i}^{ * } ) \, = \, 0.$$
(28)

The properties of G(y i ) and ∆(y i ) described above imply that F(y i ) is strictly decreasing for all \(y_{i} \le y_{i}^{*}\) and that \(y_{i}^{*}\) is the only value of y i such that the first-order condition holds. Thus we can say that F(y i ) ≥ (≤) 0 if and only if y i  ≤ (≥) \(y_{i}^{*}\); it follows that Eu i in (3) is strictly quasiconcave in y i . Since budgets do not constrain each player’s investment, the only possible corner solution would be at zero, which occurs only when Y i is so large that \(y_{i}^{*}\) satisfying the first-order condition is negative. If this is the case, then we know that a positive solution to the first-order condition in (3) does not exist because F(y i ) < 0 for all y i  ≥ 0, so player i’s optimal decision is setting y i  = 0.

Part (ii) We find the status quo level of player i’s utility with no investment and no chance of winning to be u i (I i ). The constrained optimization problem yields an interior solution, i.e., \(y^{*}\) > 0 given Y i , whenever the slope of the indifference curve q i (y i | \(\bar{U}\) = u i (I i )) at y i  = 0 is less than the slope of the constraint p(y i |Y i ) at y i  = 0. Since \(q_{i} '(y_{i} |\bar{U})\)) is given by (14) and \(p^{\prime} \left( {y_{i} |Y_{ - i} } \right) = \frac{{Y_{ - i} }}{{\left( {y_{i} + Y_{ - i} } \right)^{2} }}\), we have \(q_{i} '(y_{i} |\bar{U} = u_{i} \left( {I_{i} } \right)) \, < p'\left( {y_{i} |Y_{ - i} } \right)\) at y i  = 0 if and only if

$$\psi_{i} (I_{i} )g_{i}^{\prime} \left( 0 \right) < \frac{1}{{Y_{ - i} }} .$$
(29)

By letting \(\kappa_{i} = { 1}/[\psi_{i} (I_{i} )g_{i} '\left( 0 \right)]\), we find that \(y^{*}\) > 0 if and only if Y i  < κ i .

Part (iii) Define player i’s share function, the probability that player i wins the contestgiven his optimal investment as s i (\(Y^{*}\)) where \(Y^{*}\) =  (Y i ) + Y i . Since \(y_{i}^{*}\)(0) > 0, \(Y^{*}\) = 0 is never an equilibrium and we do not need to define s i (0). Using the constraint (25), we rewrite the tangency condition (24) as

$$\frac{{1 - p_{i}^{*} }}{{Y^{*} }} - [p_{i}^{*} \phi_{i} \left( {v_{i}^{*} } \right) + \left( {1 - p_{i}^{*} } \right)\psi_{i} \left( {v_{i}^{*} )} \right]g_{i}^{\prime} \left( {y_{i} } \right) = 0,$$
(30)

which can be rearranged as

$$p_{i}^{*} = \frac{{[1 - \psi_{i} \left( {v_{i}^{*} } \right)g_{i}^{\prime} \left( {y_{i}^{*} } \right)Y^{*} ]}}{{\left[ {1 - \psi_{i} \left( {v_{i}^{*} } \right)g_{i}^{\prime} \left( {y_{i}^{*} } \right)Y^{*} } \right] + \phi_{i} \left( {v_{i}^{*} } \right)Y^{*} }} .$$
(31)

Therefore, we have

$$s_{i} (Y^{*} ) = \hbox{max} \left\{ {\frac{{[1 - \psi_{i} \left( {v_{i}^{*} } \right)g_{i}^{\prime} \left( {y_{i}^{*} } \right)Y^{*} ]}}{{\left[ {1 - \psi_{i} \left( {v_{i}^{*} } \right)g_{i}^{\prime} \left( {y_{i}^{*} } \right)Y^{*} } \right] + \phi_{i} \left( {v_{i}^{*} } \right)Y^{*} }},0} \right\} .$$
(32)

We find that s i is a continuous function of \(Y^{*}\) for \(Y^{*}\) > 0 and that 1 is its upper bound. In addition, as \(Y^{*}\) approaches 0, s i converges to 1. From part (ii) we know that \(y^{*}\) = 0 if Y i  ≥ κ i, where κ i  = 1/[ψ i (I i )g i ′(0)]. Thus we can say that s i  = 0 if \(Y^{*}\) ≥ κ i , which is consistent with the numerator of (32). To show (d) we totally differentiate the system of Eqs. (24) and (25) with respect to Y i .

$$\frac{{dp_{i}^{*} }}{{dY_{ - i} }} = \frac{{ - \frac{{Y_{ - i} }}{{Y^{*4} }} - \frac{{y_{i}^{*} }}{{Y^{*2} }}\left[ {p_{i}^{*} \phi_{i} \left( {v_{i}^{*} } \right) + (1 - p_{i}^{*} )\psi_{i} \left( {v_{i}^{*} } \right)} \right]g_{i}'' \left( {y_{i}^{*} } \right) + \frac{{y_{i}^{*} }}{{Y^{*2} }}[p_{i}^{*} \phi^{\prime}_{i} \left( {v_{i}^{*} } \right) + \left( {1 - p_{i}^{*} } \right)\psi^{\prime}_{i} \left( {v_{i}^{*} )} \right]\left[ {g_{i}^{\prime} \left( {y_{i} } \right)} \right]^{2} }}{{q_{i}{''} \left( {y_{i}^{*} |u_{i}^{*} } \right) - p{''} (y_{i}^{*} |Y_{ - i} )}}$$
(33)
$$\frac{{dy_{i}^{*} }}{{dY_{ - i} }} = \frac{{\frac{{y_{i}^{*} - Y_{ - i} }}{{Y^{*3} }} + \frac{{y_{i}^{*} }}{{Y^{*2} }}\left[ {\phi_{i} \left( {v_{i}^{*} } \right) - \psi_{i} \left( {v_{i}^{*} } \right)} \right]g_{i}^{\prime} \left( {y_{i}^{*} } \right)}}{{q_{i}^{\prime\prime} \left( {y_{i}^{*} |u_{i}^{*} } \right) - p^{\prime\prime} (y_{i}^{*} |Y_{ - i} )}}$$
(34)

where \(u_{i}^{*}\) is the optimal EU corresponding to \(y_{i}^{*}\). Since \(q_{i}^{\prime\prime} (y_{i}^{*} |u_{i}^{*} ) \ge 0\) and

$$p^{\prime\prime} \left( {y_{i}^{*} |Y_{ - i} } \right) = \frac{{ - 2Y_{ - i} }}{{\left( {y_{i}^{*} + Y_{ - i} } \right)^{3} }} < 0 ,$$
(35)

the denominator of each equation is always positive. Consider the numerator of (33). The middle term is nonnegative because u i is convex and g i is concave. Thus \(\frac{{dp_{i}^{*} }}{{dY_{ - i} }} < 0\) if

$$\frac{{y_{i}^{*} }}{{Y^{*2} }}[p_{i}^{*} \phi^{\prime}_{i} \left( {v_{i}^{*} } \right) + \left( {1 - p_{i}^{*} } \right)\psi^{\prime}_{i} \left( {v_{i}^{*} )} \right]\left[ {g_{i}^{\prime} \left( {y_{i} } \right)} \right]^{2} < \frac{{Y_{ - i} }}{{Y^{*4} }}.$$
(36)

Substituting g i ′(y i ) from (24) and \(\frac{{y_{i}^{*} }}{{Y^{*} }}\) from (24) in (36) yields

$$p_{i}^{*} (1 - p_{i}^{*} )[p_{i}^{*} \phi^{\prime}_{i} \left( {v_{i}^{*} } \right) + \left( {1 - p_{i}^{*} } \right)\psi^{\prime}_{i} \left( {v_{i}^{*} )} \right] < \left[ {p_{i}^{*} \phi_{i} \left( {v_{i}^{*} } \right) + \left( {1 - p_{i}^{*} } \right)\psi_{i} \left( {v_{i}^{*} } \right)} \right]^{2} .$$
(37)

Conditions (i) and (ii) in the proof of Lemma 1 guarantee that (37) holds, so \(\frac{{dp_{i}^{*} }}{{dY_{ - i} }} < 0\) for all Y i ∈ (0 i ). Now consider (34). Substituting \(q_{i}^{\prime\prime} (y_{i}^{*} |u_{i}^{*} )\) from (17) and \(p^{\prime\prime} (y_{i}^{*} |Y_{ - i} )\) from (35) into (34) yields \(\frac{{dy_{i}^{*} }}{{dY_{ - i} }} > - 1\). Given the definition of \(Y^{*}\) above, it follows that \(\frac{{dY^{*} }}{{dY_{ - i} }} = \frac{{dy_{i}^{*} }}{{dY_{ - i} }} + 1\).

Since \(\frac{{dy_{i}^{*} }}{{dY_{ - i} }} > - 1\), then \(\frac{{dY^{*} }}{{dY_{ - i} }} > 0\). Given \(\frac{{dp_{i}^{*} }}{{dY_{ - i} }} < 0\), it follows that \(\frac{{ds_{i} }}{{dY_{ - i} }} < 0\) for all \(Y^{*}\) ∈ (0 i ).

1.3 Proof of Proposition 2

Using a generalization of Hermite–Hadamard’s inequality given by Theorem 5.11 in Pe˘carić et al. (1992), we have

$$\left( \frac{1}{n} \right)u^{\prime}\left( {I - x + R} \right) + \left( {\frac{n - 1}{n}} \right)u^{\prime}\left( {I - x} \right) < \frac{n}{2R}\mathop \smallint \limits_{I - x}^{I - x + 2R/n} u^{\prime}\left( w \right)dw$$
(38)

when u′′′ < 0. Using Steffensen’s inequality given by Theorem 6.19 in Pe˘carić et al. (1992), we find

$$\mathop \smallint \limits_{I - x}^{I - x + 2R/n} u^{\prime}\left( w \right)dw < \frac{2}{n}\mathop \smallint \limits_{I - x}^{I - x + R} u^{\prime}\left( w \right)dw$$
(39)

when u″ > 0. Thus, (38) and (39) jointly imply

$$\left( \frac{1}{n} \right)u^{\prime}\left( {I - x + R} \right) + \left( {\frac{n - 1}{n}} \right)u^{\prime}\left( {I - x} \right) < \frac{1}{R}\mathop \smallint \limits_{I - x}^{I - x + R} u^{\prime}\left( w \right)dw$$
(40)

and the bracketed term in (9) is greater than 1. Proposition 2 clearly follows.

Appendix 2: A note on equilibrium in Yamazaki (2009)

In the proof of Lemma 1 found in Yamazaki (2009), the author attempts to confirm the convexity of an indifference curve G by examining its first partial derivative with respect to y i :

$$\frac{{\partial G_{i} }}{{\partial y_{i} }} = \frac{{g_{i}^{\prime} (y_{i} )}}{{u_{i} \left( {z_{i}^{H} } \right) - u_{i} \left( {z_{i}^{L} } \right)}}\left[ {u_{i}^{\prime} \left( {z_{i}^{L} } \right) + \frac{{\bar{U} - u_{i} \left( {z_{i}^{L} } \right)}}{{u_{i} \left( {z_{i}^{H} } \right) - u_{i} \left( {z_{i}^{L} } \right)}}(u_{i}^{\prime} \left( {z_{i}^{H} } \right) - u_{i}^{\prime} \left( {z_{i}^{L} } \right))} \right] ,$$
(41)

where z H i denotes income with the rent and z L i is income without it, with all other variables being defined previously. The author incorrectly claims u i (z H i ) − u i (z L i ) to be non-increasing in y i when the difference is actually increasing in y i under the assumption of risk aversion. Therefore we cannot conclude that indifference curves are convex and the existence of a unique equilibrium under DARA breaks down.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jindapon, P., Whaley, C.A. Risk lovers and the rent over-investment puzzle. Public Choice 164, 87–101 (2015). https://doi.org/10.1007/s11127-015-0270-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11127-015-0270-y

Keywords

JEL Classification

Navigation