Skip to main content
Log in

Effort complementarity and sharing rules in group contests

  • Original Paper
  • Published:
Social Choice and Welfare Aims and scope Submit manuscript

Abstract

In this paper, we consider a prize-sharing rule design problem in a group contest with effort complementarities within groups by employing a CES effort aggregator function. We derive the conditions for a monopolization rule that dominates an egalitarian rule if the objective of the rule design is to maximize the group’s winning probability. We find conditions under which the monopolization rule maximizes the group’s winning probability, while the egalitarian rule is strictly preferred by all members of the group. Without effort complementarity, there cannot be such a conflict of interest.

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. This is called PNC (potential non-contributor condition) in Nitzan and Ueda (2014). In addition, this result needs a mild technical condition called RC (regularity condition) for the cost function around zero effort.

  2. They also show that the egalitarian sharing rule maximizes the winning probability if the marginal cost starts from zero and is convex (their anti-Olson theorem).

  3. Esteban and Ray (2001) clarify that Olson’s (1965) results do not depend on whether the prize is public or private. They explain that after fixing the equal allocation of the private good among all members in a group, whether its winning probability increases or not depends on the elasticity of the marginal effort cost when the population increases.

  4. Choi et al. (2016) analyze a two-group contest when the prize is allocated by an intra-group contest between the group members with asymmetric powers. Cheikbossian and Fayat (2018) analyze a contest between two groups by introducing a rivalry parameter of an impure public good prize.

  5. Kolmar and Rommeswinkel (2013) are the first in the literature to introduce group members’ effort complementarity in group contests by using a CES production function. They call this CES function a group impact function.

  6. We employ the Tullock-form contest success function (Tullock 1980).

  7. Nitzan and Ueda (2011) assume that individual effort levels are observable by the group leader and analyze the case of the convex combination between the egalitarian rule and a relative-effort sharing rule, which allocates the winning prize proportionally to their effort levels.

  8. Baik (2008) calls this Nash equilibrium a group-i-specific equilibrium.

  9. In a group contest for multiple (homogenous) indivisible prizes, Crutzen et al. (2020) obtain a closely related result. When \(2r<\beta\), allocating prizes according to a fair lottery maximizes group success, while when \(2r>\beta\), allocating prizes according to a predetermined (priority) list is better than the fair lottery allocation rule.

  10. See also the “anti-Olson theorem” in Nitzan and Ueda (2014).

  11. In Proposition 3, condition \(1-\frac{n_{i}}{n_{i}+1}\beta >P_{iM}\) appears to be a condition for an endogenous variable \(P_{iM}\) because \(P_{iM}\) is explicitly unsolvable. However, \(P_{iM}\) is uniquely led by

    $$\begin{aligned} P_{iM}=\frac{P_{iM}^{\frac{1}{\beta }}(1-P_{iM})^{\frac{1}{\beta }}}{P_{iM}^{ \frac{1}{\beta }}(1-P_{iM})^{\frac{1}{\beta }}+X_{-i}}. \end{aligned}$$

    This is determined by the exogenous variables \(\beta\) and \(X_{-i}\) only. Thus, we can confirm if condition \(1-\frac{n_{i}}{n_{i}+1}\beta >P_{iM}\) is satisfied with economic data \(\beta\), \(n_{i}\), and \(X_{-i}\).

  12. Readers may think that it is unrealistic to assume that groups can observe other groups’ sharing rules. Nitzan and Ueda (2011) assume that sharing rules are the private information of each group and use perfect Bayesian equilibrium with the same beliefs for other groups’ sharing rules at every information set. Since the model does not involve a real asymmetric information problem, their perfect Bayesian equilibrium coincides with our subgame perfect equilibrium under complete information.

  13. We thank Kaoru Ueda for suggesting that we use the share function approach.

  14. When the group leader chooses the monopolization rule at Stage 1, effort complementarity is irrelevant on the equilibrium path. Effort complementarity is in effect only off the equilibrium path.

References

  • Baik KH (2008) Contests with group-specific public-good prizes. Soc Choice Welf 30:103–117

    Article  Google Scholar 

  • Cheikbossian G, Fayat R (2018) Group size, collective action and complementarities in efforts. Econ Lett 168:77–81

    Article  Google Scholar 

  • Choi JP, Chowdhury SM, Kim J (2016) Group contests with internal conflict and power asymmetry. Scand J Econ 118:816–840

    Article  Google Scholar 

  • Cornes R, Hartley R (2005) Asymmetric contests with general technologies. Econ Theor 26–4:923–946

    Article  Google Scholar 

  • Crutzen BSY, Flamand S, Sahuguet N (2020) A model of a team contest, with an application to incentives under list proportional representation. J Public Econ 182:104109

    Article  Google Scholar 

  • Epstein GS, Mealem Y (2009) Group specific public goods, orchestration of interest groups with free riding. Public Choice 139(3–4):357–369

    Article  Google Scholar 

  • Esteban J, Ray D (2001) Collective action and the group size paradox. Am Polit Sci Rev 95–3:663–672

    Article  Google Scholar 

  • Kolmar M, Rommeswinkel H (2013) Contests with group-specific public goods and complementarities in efforts. J Econ Behav Organ 89:9–22

    Article  Google Scholar 

  • Nitzan S, Ueda K (2011) Prize sharing in collective contests. Eur Econ Rev 55:678–687

    Article  Google Scholar 

  • Nitzan S, Ueda K (2014) Intra-group heterogeneity in collective contests. Soc Choice Welf 43:219–238

    Article  Google Scholar 

  • Olson M (1965) The logic of collective action. Harvard University Press, Cambridge

    Google Scholar 

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

    Google Scholar 

  • Ueda K (2002) Oligopolization in collective rent-seeking. Soc Choice Welf 19–3:613–626

    Article  Google Scholar 

Download references

Acknowledgements

We thank Editor François Maniquet, Associate Editor, and two anonymous referees for their helpful comments and suggestions. Special thanks are due to Kaoru Ueda for his valuable comments on an earlier version of the paper. This paper was completed when Kobayashi was visiting Boston College on his sabbatical. Kobayashi thanks Hosei University for their financial support and Boston College for their hospitality.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hideo Konishi.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1

Here, we collect all proofs.

Proof of Lemma 1

Recalling \(X_{i}=( \sum _{j=1}^{n_{i}}e_{ij}^{r})^{\frac{1}{r}}\) and given \(X_{-i}\), maximizing the winning probability of group i means that \(X_{i}\) becomes as large as possible at Nash equilibrium in group i. If \(X_{i}\) is a strictly increasing function of \(A_{i}\), we can maximize \(X_{i}\) by maximizing \(A_{i}\) subject to \(\sum _{j=1}^{n_{i}}a_{ij}=1\).

From (2), let \(\phi (X_{i},A_{i})=X_{i}^{\beta }-P_{i}(1-P_{i})A_{i}=0\). Recalling that \(\beta \ge 1\), \(0<r\le 1\) and that \(P_{i}\) is a function of \(X_{i}\) through \(P_{i}=\frac{X_{i}}{ X_{i}+X_{-i}}\), and by differentiating \(\phi\) with respect to \(A_i\), we get

$$\begin{aligned} \frac{dX_{i}}{dA_{i}}=-\frac{\phi _{A_{i}}}{\phi _{X_{i}}}=\frac{ P_{i}(1-P_{i})}{X_{i}^{\beta -1}(\beta -1+2P_{i})}>0 \end{aligned}$$

for \(\beta >r\) by using \(A_{i}=\frac{X_{i}^{\beta }}{P_{i}(1-P_{i})}\) from (2). Thus, \(X_{i}\) is a strictly increasing function in \(A_{i}\) for \(\beta >r\). \(\square\)

Proof of Proposition 1

From Lemma 1, it is enough to maximize \(A_{i}\). It is also enough to maximize the contents in parentheses in \(A_{i}\) because \(\frac{\beta -r}{r}>0\). Note that \(A_{i}^{\frac{r}{\beta -r}}=\sum _{j=1}^{n_{i}}a_{ij}^{\frac{r}{\beta -r}}\) is an additively separable function. Since \(r>0\), our maximization problem boils down to

$$\begin{aligned} \max \sum _{j=1}^{n_{i}}a_{ij}^{\frac{r}{\beta -r}}\text { subject to (i) } \sum _{j=1}^{n_{i}}a_{ij}=1\text { and (ii) }a_{ij}\ge 0 \quad \text { for all } \quad j=1,\ldots ,n_{i}. \end{aligned}$$

Thus, it is easy to see that \(\frac{r}{\beta -r}\lesseqgtr 1\) dictates the optimal sharing rule. We obtain three cases:

  • Case 1: If \(2r<\beta\), \(A_{i}\) is maximized when \(a_{1}=a_{2}=\ldots =a_{n_{i}}=1/n_{i}\).

  • Case 2: If \(2r=\beta\), \(A_{i}\) is constant for any sharing rule.

  • Case 3: If \(2r>\beta\), \(A_{i}\) is maximized when \(a_{ij}=1\) for a single j, and \(a_{i\ell }=0\) for all other \(\ell\).

\(\square\)

Proof of Lemma 2

Let \(\Delta \equiv n_{i}^{-\frac{\beta -2r}{r\beta }}\) in relation to \(n_{i}\) in (4). Rewriting (4), we have

$$\begin{aligned} \left( 1-P_{iE}\right) \left( P_{iE}(1-P_{iE})\right) ^{\frac{1}{\beta } }=P_{iE}X_{-i}\Delta . \end{aligned}$$
(10)

By totally differentiating the above, we obtain

$$\begin{aligned}&\left[ -\left( P_{iE}(1-P_{iE})\right) ^{\frac{1}{\beta }}-X_{-i}\Delta + \frac{1}{\beta }\left( 1-P_{iE}\right) \left( P_{iE}(1-P_{iE})\right) ^{ \frac{1}{\beta }-1}\left( 1-2P_{iE}\right) \right] dP_{iE}\\&\quad =P_{iE}X_{-i}d \Delta . \end{aligned}$$

After solving (10) for \(X_{-i}\), we substitute it into the above and obtain

$$\begin{aligned}&\frac{1}{P_{iE}}\left( P_{iE}(1-P_{iE})\right) ^{\frac{1}{\beta }}\left[ -1+ \frac{1}{\beta }\left( 1-2P_{iE}\right) \right] dP_{iE}\\&\quad =\left( 1-P_{iE}\right) \left( P_{iE}(1-P_{iE})\right) ^{\frac{1}{\beta }}\frac{ d\Delta }{\Delta } \end{aligned}$$

or

$$\begin{aligned} \frac{dP_{iE}}{d\Delta }=\frac{P_{iE}\left( 1-P_{iE}\right) }{\Delta \left[ -1+\frac{1}{\beta }\left( 1-2P_{iE}\right) \right] }. \end{aligned}$$

Since \(-1+\frac{1}{\beta }\left( 1-2P_{iE}\right) <0\) from \(\beta \ge 1\), we have \(\frac{dP_{iE}}{d\Delta }<0\). By differentiating \(\Delta\) with respect to \(n_{i}\) and r, we have

$$\begin{aligned} \frac{d\Delta }{dn_{i}}=-\frac{\beta -2r}{r\beta }\times \frac{\Delta }{n_{i} } \end{aligned}$$

and

$$\begin{aligned} \frac{d\Delta }{dr}=\frac{1}{r^{2}}(\log n_{i})\Delta >0, \end{aligned}$$

respectively. We obtain the results using the chain rule. \(\square\)

Proof of Lemma 3

Recall \(U_{iE}(n_{i},r)=\frac{P_{iE}(n_{i},r)}{n_{i}}\left( 1-\frac{1}{\beta }\left( 1-P_{iE}(n_{i},r)\right) \frac{1}{n_{i}}\right) \equiv {\tilde{U}} (n_{i},P_{iE}(n_{i},r))\). This implies

$$\begin{aligned} \frac{\partial {\tilde{U}}}{\partial P_{iE}}=\frac{1}{n_{i}}-\frac{1}{\beta n_{i}^{2}}(1-2P_{iE})=\frac{1}{n_{i}^{2}\beta }(n_{i}\beta +2P_{iE}-1). \end{aligned}$$
(11)

Thus, by totally differentiating \({\tilde{U}}(n_{i},P_{iE}(n_{i},1))\) with respect to \(n_{i}\) using (6), we obtain

$$\begin{aligned}&\frac{dU_{iE}(n_{i},1)}{dn_{i}} \\&\quad =\frac{\partial {\tilde{U}}}{\partial n_{i}}+\frac{\partial {\tilde{U}}}{ \partial P_{iE}}\frac{dP_{iE}}{dn_{i}} \\&\quad =\frac{P_{iE}}{n_{i}^{2}}\left[ -1+\frac{2}{n_{i}\beta }(1-P_{iE})\right] + \frac{\left( 2-\beta \right) P_{iE}\left( 1-P_{iE}\right) }{n_{i}\left( 1-2P_{iE}-\beta \right) }\left[ \frac{1}{n_{i}^{2}\beta }(n_{i}\beta +2P_{iE}-1)\right] \\&\quad =\frac{P_{iE}}{n_{i}^{3}}\left[ -n_{i}+\frac{2}{\beta }\left( 1-P_{iE}\right) +\left( 1-P_{iE}\right) \frac{2-\beta }{1-2P_{iE}-\beta } \left( n_{i}+\frac{1}{\beta }\left( 2P_{iE}-1\right) \right) \right] \\&\quad =\frac{P_{iE}}{n_{i}^{3}\beta \left( 1-2P_{iE}-\beta \right) } \\&\qquad \times \Big [ -n_{i}\beta \left( 1-2P_{iE}-\beta \right) +2\left( 1-P_{iE}\right) \left( 1-2P_{iE}-\beta \right) \\&\qquad + \,\left( 1-P_{iE}\Big ) \left( 2-\beta \right) \left( \beta n_{i}-\left( 1-2P_{iE}\right) \right) \right] . \end{aligned}$$

Since \(1-2P_{iE}-\beta <0\), we can focus on the sign of the contents of the brackets:

$$\begin{aligned} \left[ \ \cdot \ \right]= & {} -n_{i}\beta \left( 1-2P_{iE}-\beta \right) +n_{i}\beta \left( 2-\beta -2P_{iE}+P_{iE}\beta \right) \\&+2\left( 1-2P_{iE}-\beta \right) \left( 1-P_{iE}\right) -\left( 1-P_{iE}\right) \left( 2-\beta \right) \left( 1-2P_{iE}\right) \\= & {} \beta [(n_{i}-1)+P_{iE}(n_{i}\beta -1+2P_{iE})]>0 \end{aligned}$$

for any \(n_{i}\ge 1\). Thus, we conclude that \(\frac{dU_{iE}}{dn_{i}}<0\) for any \(n_{i}\ge 1\). This implies that \(U_{iE}(n_{i},1)<U_{iM}\) for any \(n_{i}\ge 2\) when \(r=1\). We have completed the proof. \(\square\)

Proof of Lemma 4

First note that the assumptions \(n_{i}\ge 2\) and \(1+\frac{1}{n_{i}}>\beta\) imply \(2>\beta\). Consider the case of \(r=\frac{\beta }{2}\). Since \(P_{iE}(n_{i},\frac{\beta }{ 2})=P_{iE}(1,\frac{\beta }{2})=P_{iM}\) by Proposition 1, we have

$$\begin{aligned} U_{iE}\left( n_{i},\frac{\beta }{2}\right)= & {} \frac{P_{iE}\left( n_{i},\frac{\beta }{2}\right) }{n_{i} }\left( 1-\frac{1}{\beta }\left( 1-P_{iE}\left( n_{i},\frac{\beta }{2}\right) \right) \frac{1}{n_{i}} \right) \\= & {} \frac{P_{iM}}{n_{i}}\left( 1-\frac{1}{\beta }(1-P_{iM})\frac{1}{n_{i}} \right) . \end{aligned}$$

By subtracting \(U_{iM}=P_{iM}\left[ 1-\frac{1}{\beta }(1-P_{iM})\right]\) from \(U_{iE}(n_{i},\frac{\beta }{2})\), we obtain

$$\begin{aligned} U_{iE}\left( n_{i},\frac{\beta }{2}\right) -U_{iM}= & {} P_{iM}\left[ -1+\frac{1}{\beta } (1-P_{iM})+\frac{1}{n_{i}}\left( 1-\frac{1}{\beta }(1-P_{iM})\frac{1}{n_{i}} \right) \right] \\= & {} P_{iM}\left[ -\left( 1-\frac{1}{n_{i}}-\frac{1}{\beta }\left( 1-\frac{1}{ n_{i}^{2}}\right) \right) -\frac{1}{\beta }P_{iM}\left( 1-\frac{1}{n_{i}^{2}} \right) \right] \\= & {} P_{iM}\left( 1-\frac{1}{n_{i}^{2}}\right) \left[ -\frac{n_{i}}{n_{i}+1}+ \frac{1}{\beta }-\frac{1}{\beta }P_{iM}\right] . \end{aligned}$$

Then, the condition of \(U_{iE}(n_{i},\frac{\beta }{2})>U_{iM}\) is

$$\begin{aligned} 1-\frac{n_{i}}{n_{i}+1}\beta >P_{iM}. \end{aligned}$$
(12)

That is, if (12) is satisfied, \(U_{iE}(n_{i},\frac{\beta }{2} )>U_{iM}\) holds, while \(U_{iE}(n_{i},1)<U_{iM}\). Since \(\frac{dP_{iE}}{dr}<0\) holds by (5) in Lemma 2 and from (11), \(\frac{\partial {\tilde{U}}}{\partial P_{iE}}=\frac{1}{n_{i}^{2}\beta }(n_{i}\beta +2P_{iE}-1)>0\), we have \(\frac{dU_{iE}(P_{iE}(r))}{dr}=\frac{\partial \tilde{ U}_{iE}}{\partial P_{iE}}\frac{dP_{iE}}{dr}<0\), which is \(U_{iE}\) monotonically decreasing in r. Considering the above facts and given that \(U_{iE}\) is continuous in r, there is a unique \({\hat{r}}\in (\frac{\beta }{2} ,1)\), such that \(U_{iE}(n_{i},r)<U_{iM}\) holds for all \(r\in ({\hat{r}},1]\) and \(U_{iE}(n_{i},r)>U_{iM}\) holds for all \(r\in [\frac{\beta }{2}, {\hat{r}})\). \(\square\)

Fig. 1
figure 1

Share function \(P_i(X;A_i)\) of group i and aggregate share function f(XA)

Proof of Lemma 5

First, focus on the \(P_{i}(X;A_{i})\) function. Starting from the original \(A_{i}\) and equilibrium \(X^{*}\), \(A_{i}\) is increased by \(\Delta A_{i}>0\). Since \(\frac{\partial P_{i}}{\partial A_{i}}>0\) for all X from (9 ), the \(P_{i}\) function shifts up vertically. Let \({\tilde{X}}\) be such that \(P_{i}(X^{*};A_{i})=P_{i}({\tilde{X}};A_{i}+\Delta A_{i})\) (see Fig. 1). Since \(\frac{\partial P_{i}}{\partial X}<0\) from (8), \({\tilde{X}} >X^{*}\) holds, and for any \(X\in (X^{*},{\tilde{X}})\), we have \(P_{i}(X;A_{i}+\Delta A_{i})>P_{i}(X^{*};A_{i})\). Recall that the equilibrium \(X^{*}\) is described by the aggregate share function

$$\begin{aligned} f(X^{*};A)=\sum _{i^{\prime }\ne i}P_{i^{\prime }}(X^{*};A_{i^{\prime }})+P_{i}(X^{*};A_{i})=1. \end{aligned}$$

Let \(A_{-i}\) be a vector that removes \(A_{i}\) from A. By increasing \(A_{i}\) by \(\Delta A_{i}\), the equilibrium aggregate effort \(X^{**}\) satisfies

$$\begin{aligned} f(X^{**};A_{i}+\Delta A_{i},A_{-i})=\sum _{i^{\prime }\ne i}P_{i^{\prime }}(X^{**};A_{i^{\prime }})+P_{i}(X^{**};A_{i}+\Delta A_{i})=1. \end{aligned}$$

Since \(\frac{\partial P_{i^{\prime }}}{\partial X}<0\) for all \(i^{\prime }=1,\ldots ,m\), we have \(X^{**}>X^{*}\) and

$$\begin{aligned} f({\tilde{X}};A_{i}+\Delta A_{i},A_{-i})= & {} \sum _{i^{\prime }\ne i}P_{i^{\prime }}({\tilde{X}};A_{i^{\prime }})+P_{i}({\tilde{X}};A_{i}+\Delta A_{i}) \\= & {} \sum _{i^{\prime }\ne i}P_{i^{\prime }}({\tilde{X}};A_{i^{\prime }})+P_{i}(X^{*};A_{i})<1. \end{aligned}$$

By the intermediate value theorem, \(X^{**}\in (X^{*},{\tilde{X}})\) holds. We conclude \(P_{i}(X^{**};A_{i}+\Delta A_{i})>P_{i}(X^{*};A_{i})\).

This implies that as \(A_{i}\) increases, \(P_{i}(X^*; A_{i})\) increases. That is, maximizing \(A_{i}\) achieves the maximum winning probability for group i. \(\square\)

Appendix 2

Here, we repeat our analysis by using the Epstein and Mealem’s generalized Tullock contest, and show that Lemma 1 and Proposition 1 hold. We confirm this first. The expected payoff of member j in group i is \(U_{ij}=\frac{\sum _{j=1}^{n_{i}}e_{ij}^{r}}{ \sum _{j=1}^{n_{i}}e_{ij}^{r}+X_{-i}}a_{ij}-\frac{1}{\beta }e_{ij}^{\beta }\). The first order condition is

$$\begin{aligned} \frac{\partial U_{ij}}{\partial e_{ij}}=\frac{re_{ij}^{r-1}X_{-i}}{ \left( \sum _{j=1}^{n_{i}}e_{ij}^{r}+X_{-i}\right) ^{2}}a_{ij}-e_{ij}^{\beta -1}=0. \end{aligned}$$

This can be rewritten as

$$\begin{aligned} P_{i}(1-P_{i})\frac{re_{ij}^{r}}{\sum _{j=1}^{n_{i}}e_{ij}^{r}} a_{ij}-e_{ij}^{\beta }=0. \end{aligned}$$
(13)

We process a procedure similar to the one at the end of Sect. 2 and get \(e_{ij}^{r}=\left( \frac{rP_{i}(1-P_{i})}{X_{i}}\right) ^{\frac{r}{\beta -r} }a_{ij}^{\frac{r}{\beta -r}}\) from (13). By summing up each \(e_{ij}^{r}\), we have \(\sum _{j=1}^{n_{i}}e_{ij}^{r}=X_{i}=\left( \frac{ rP_{i}(1-P_{i})}{X_{i}}\right) ^{\frac{r}{\beta -r}}\hat{A_i}\) where \(\hat{ A_i} = \sum _{j=1}^{n_i}a_{ij}^{\frac{r}{\beta -r}}\). Let \({\hat{\phi }}(X_{i}, \hat{A_i})=X_{i}-\left( \frac{rP_{i}(1-P_{i})}{X_{i}}\right) ^{\frac{r}{ \beta -r}}\hat{A_i}=0.\) By differentiating \({\hat{\phi }}\) with respect to \(\hat{A_i}\) and noting that \(P_{i}\) is a function of \(X_{i}\), we have

$$\begin{aligned} \frac{dX_{i}}{d\hat{A_i}}=-\frac{{\hat{\phi }}_{\hat{A_i}}}{{\hat{\phi }}_{X_{i}}}= \frac{(rP_{i}(1-P_{i})/X_{i})^{\frac{r}{\beta -r}}}{(\beta -r+2rP_{i})/(\beta -r)}>0 \end{aligned}$$

for \(\beta >r\) by using \(\hat{A_i}=X_{i}\left( \frac{rP_{i}(1-P_{i})}{X_{i}} \right) ^{-\frac{r}{\beta -r}}\). Therefore, since Lemma 1 holds, Proposition 1 also holds in this case. Proposition 2 holds as well. However, Proposition 3 does not hold. We check this second point. Under the egalitarian rule, since \(X_{i}= \sum _{j=1}^{n_{i}}e_{ij}^{r}=n_{i}e_{i}^{r}\), we have \(e_{i}=n_{i}^{-\frac{2 }{\beta }}r^{\frac{1}{\beta }}P_{iE}^{\frac{1}{\beta }}(1-P_{iE})^{\frac{1}{ \beta }}\) from (13). Using this, we have

$$\begin{aligned} P_{iE}=\frac{n_{i}e_{i}^{r}}{n_{i}e_{i}^{r}+X_{-i}}=\frac{n_{i}^{\frac{\beta -2r}{\beta }}r^{\frac{r}{\beta }}P_{iE}^{\frac{r}{\beta }}(1-P_{iE})^{\frac{r }{\beta }}}{n_{i}^{\frac{\beta -2r}{\beta }}r^{\frac{r}{\beta }}P_{iE}^{ \frac{r}{\beta }}(1-P_{iE})^{\frac{r}{\beta }}+X_{-i}} \end{aligned}$$
(14)

and

$$\begin{aligned} U_{iE}=P_{iE}\frac{1}{n_{i}}-\frac{1}{\beta }e_{i}^{\beta }=\frac{P_{iE}}{ n_{i}}\left( 1-\frac{1}{\beta }(1-P_{iE})\frac{r}{n_{i}}\right) . \end{aligned}$$

Let \({\hat{\Delta }} \equiv n_i^{-\frac{\beta -2r}{\beta }}\) in relation to \(n_i\) in (14). We process the same procedure as in the proof of Lemma 2. Rewriting (14), we have

$$\begin{aligned} r^{\frac{r}{\beta }} \left( 1-P_{iE}\right) \left( P_{iE}(1-P_{iE})\right) ^{ \frac{r}{\beta }}=P_{iE}X_{-i} {\hat{\Delta }}. \end{aligned}$$
(15)

By totally differentiating the above expression and conducting the same operations as the proof of Lemma 2, we obtain

$$\begin{aligned} \frac{dP_{iE}}{d {\hat{\Delta }} }=\frac{P_{iE}\left( 1-P_{iE}\right) }{ \hat{ \Delta } \left[ -1+\frac{r}{\beta }\left( 1-2P_{iE}\right) \right] }<0 \end{aligned}$$

because of \(\frac{r}{\beta }(1-2P_{iE}) < 1\). Differentiating \({\hat{\Delta }}\) with respect to \(n_i\), we have

$$\begin{aligned} \frac{d {\hat{\Delta }} }{d n_{i}}=- \frac{\beta -2r}{\beta }\times \frac{ \hat{ \Delta } }{n_{i}}. \end{aligned}$$

We then obtain \(\frac{dP_{iE}}{dn_{i}}=\frac{2r-\beta }{n_i} \frac{ P_{iE}(1-P_{iE})}{r(1-2P_{iE}) -\beta }.\) The sign of this formula depends only on the sign of \(2r - \beta\), as well as (6) from Lemma 2. Therefore, Proposition 2 holds in this case.

By processing the same procedure as the proof of Lemma 4, we obtain

$$\begin{aligned} U_{iE}\left( n_{i},\frac{\beta }{2}\right) -U_{iM}= & {} P_{iM}\left[ -1+\frac{1}{\beta } (1-P_{iM})r+\frac{1}{n_{i}}\left( 1-\frac{1}{\beta }(1-P_{iM})\frac{r}{n_{i}} \right) \right] \\= & {} P_{iM}\left( 1-\frac{1}{n_{i}^{2}}\right) \left[ -\frac{n_{i}}{n_{i}+1}+ \frac{r}{\beta }-\frac{r}{\beta }P_{iM}\right] \\= & {} P_{iM}\left( 1-\frac{1}{n_{i}^{2}}\right) \left[ -\frac{n_{i}}{n_{i}+1}+ \frac{1}{2}-\frac{1}{2}P_{iM}\right] \end{aligned}$$

at \(r=\frac{\beta }{2}\). For the above expression to be positive, the sign in the brackets needs to be positive. Thus,

$$\begin{aligned} P_{iM}<\frac{1-n_{i}}{n_{i}+1}<0. \end{aligned}$$

However, this condition contradicts the definition of the probability. Lemma 4 does not hold. Therefore, Proposition 3 also fails to hold in the generalized Tullock contest. \(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kobayashi, K., Konishi, H. Effort complementarity and sharing rules in group contests. Soc Choice Welf 56, 205–221 (2021). https://doi.org/10.1007/s00355-020-01277-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00355-020-01277-9

Navigation