Skip to main content
Log in

Crafting consensus

  • Published:
Public Choice Aims and scope Submit manuscript

Abstract

The paper analyzes the problem of a committee chair using favors at her disposal to maximize the likelihood that her proposal gains committee support. The favors increase the probability of a given member approving the chair’s proposal via a smooth voting function. The decision-making protocol is any quota voting rule. The paper characterizes the optimal allocation of any given level of favors and the optimal expenditure-minimizing level of favors. The optimal allocation divides favors uniformly among a coalition of the committee members. At a low level of favors, the coalition comprises all committee members. At a high level, it is the minimum winning coalition. The optimal expenditure level guarantees the chair certain support of the minimum winning coalition if favors are abundant and uncertain support of all committee members if favors are scarce; elitist or egalitarian committees are compatible with a strategic chair. The results are robust to changing the chair’s objectives and to alternative voting functions, and reconcile theoretical predictions with empirical observations about legislative bargaining experiments, lobby vote buying and executive lawmaking.

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.

Fig. 1

Similar content being viewed by others

Notes

  1. Others, for example Dougherty and Edward (2012), use the term k-majority rule.

  2. Our voting functions are formally related to the contest success functions. See Amegashie (2002) on the use of contest success functions in a bargaining model and Corchon and Dahm (2010) on the connection between contests and bargaining. Moreover, the voting functions are related to probabilistic voting (Lindbeck and Weibull 1987) in electoral competition models. We study distribution (of favors), while electoral models typically study redistribution (of income) from one group of voters to another and predict no redistribution with homogeneous voters.

  3. Supermajorities arise in legislative bargaining in non-stationary equilibria (Norman 2002; Dahm et al. 2014; Dahm and Glazer 2015), when proposing is costly (Glazer and McMillan 1990, 1992) or while bargaining over public goods (Volden and Wiseman 2007).

  4. Supermajorities arise in government formation when proposers are uncertain about which coalitions will form (Weingast 1979; Shepsle and Weingast 1981), under preference for ideologically connected coalitions (Axelrod 1970), under uncertainty about the size of the voting body created by abstention (Koehler 1975), because supermajorities make legislative logrolls self-enforcing (Carrubba and Volden 2000), because supermajorities allow for extraction of rents from excess coalition members (Baron and Diermeier 2001) and in the presence of the agenda setter’s aspirations for political stability and greatness (Doron and Sherman 1995).

  5. Most of the vote-buying literature restricts attention either to vote (yes versus no) contingent payments or to outcome (acceptance versus rejection) contingent payments. Dal Bo (2007) introduces ‘pivotal bribes’—payments contingent on a vote being outcome-pivotal—and shows that these allow a single vote-buyer to purchase all votes almost for free. Felgenhauer and Gruner (2008) and Morgan and Vardy (2011) use pivotal bribes, at least in part of their analyses. The pivotal bribes are related but distinct from the observation that if voters can commit to vote for a certain action in return for a transfer, the resulting Bertrand competition among the voters benefits the vote-buyer (Ferejohn 1986).

  6. Tyutin and Zaporozhets (2017), assuming identical payments to the members of the paid coalition, generalize this insight to multiple players.

  7. We assume that the chair is not a voting member of the committee. This is not important. If the chair were to vote, we could relabel n and q, if necessary, and proceed without further alterations.

  8. The full specification is required for the discussion of alternative voting functions in Sect. 5.

  9. Section 5 discusses voting functions generated by different distributions of \((o_{i})_{i\in N}\).

  10. The upfront payments and campaign promises terminology comes from the vote-buying literature, specifically from Dekel et al. (2008).

  11. Section 5 discusses the addition of an outside option the chair receives if her proposal fails.

  12. This observation greatly simplifies the subsequent analysis, as it implies that the uncertainty remaining in the number of votes the chair receives has a Binomial distribution with \(n-r^{*}-s^{*}\) trials, not a Poisson Binomial distribution. The Poisson Binomial random variable represents the number of successes in n independent Bernoulli trials with success probabilities \(p_{1},\ldots ,p_{n}\). The Binomial distribution is the special case when all of the probabilities are equal.

  13. Universal coalitions, in which all members receive favors, can be seen as equally extreme and, hence, implausible prediction as the minimum winning coalitions. However, the model predicts universal coalitions among members who are willing to sell their votes, leaving members who are not willing to sell outside the model. Moreover, not every favor-receiver votes yes.

  14. We conjecture that \(r^{*}\) with increasing \(b\in [q-1,q)\) visits any integer in \(\{0,\ldots ,n-q\}\) and is non-decreasing. We have (numerically) confirmed the conjecture for \(n\le 13\) and any \(q\le n\). Formal proof, however, eludes us.

  15. An alternative intuition uses the fact that the mean number of votes given any \({\mathbf {p}}=(p_{i})_{i\in N}\) is \(\sum _{i\in N}p_{i}\le b\). When \(b\ge q\), the minimum winning coalition clearly is optimal. When \(b<q-1\), the chair’s proposal on average loses and her optimal favor allocation maximizes the variance in the number of votes, \(\sum _{i\in N}p_{i}(1-p_{i})\), which is done by setting \({\mathbf {p}}=(\frac{b}{n})_{i\in N}\). When \(b\in [q-1,q)\), the chair increases the variance only partially.

  16. A Gaussian or ordinary hypergeometric function \(\,_{2}F_{1}\) is a special function that includes many other (special) functions as special cases. See the preliminary section to the proofs in Appendix 1 or Olver et al. (2010) for details.

  17. Propositions 2 and 3 reflect a large increase in B. The former proposition shows that B and \(b^{*}\) are positively related in general.

  18. Because \(\frac{n}{2}+\frac{1}{2}\,_{2}F_{1}(1,1+n,1+\frac{n}{2},\frac{1}{2})\) equals \(\frac{n}{2}+\frac{1}{2}+\frac{\sqrt{\pi }}{2}z(n)\), where, using \(\Gamma \), the Euler gamma function, \(z(n)=\frac{\Gamma \left( 1+\frac{n}{2}\right) }{\Gamma \left( \frac{1+n}{2}\right) }\) and we have \(\lim _{n\rightarrow \infty }z(n)-\frac{\sqrt{n}}{\sqrt{2}}=0\).

  19. The objective \(B{\mathbb {R}}_{q|n}[b]-b\) assumes that, for all \(i\in N\), the chair pays \(x_{i}\) to i and i votes yes with probability \(p_{i}(x_{i})=x_{i}\). This requires a behavioral model of the voting functions, for example, i voting yes out of gratitude to the chair for having received \(x_{i}\). Generating the voting functions through vote-contingent payments and vote motivation implies that the chair pays only for the votes she actually receives.

  20. Assuming that the chair pays only for the votes she actually receives considerably complicates the analysis. In the \({{\mathbb {CE}}}\) problem, the chair now maximizes, by choosing \({\mathbf {p}}\in {\mathbb {R}}^{n}\) subject to the ex-post budget constraint \({\mathbf {p}}*{\mathbf {1}}\le B\) and \(p_{i}\in [0,1]\) for \(\forall i\in N\), either \((B-{\mathbf {p}}*{\mathbf {p}}){\mathbb {P}}_{q|n}[{\mathbf {p}}]\) or \(B{\mathbb {P}}_{q|n}[{\mathbf {p}}]-{\mathbf {p}}*{\mathbf {p}}\), depending on whether she pays for the votes only when her proposal passes or not. The complication arises because \({\mathbb {P}}_{q|n}[{\mathbf {p}}]\) and \(-{\mathbf {p}}*{\mathbf {p}}\) must be maximized jointly, so that the nested nature of the \({{\mathbb {FA}}}\) and \({{\mathbb {CE}}}\) problems is lost. The only observation that does not require extended additional analysis is that, for \(B<q-1\), any solution to the \({{\mathbb {CE}}}\) problem must involve uniform allocation of favors among committee members. This is because, for any \(b'<B\), \(p_{i}=\frac{b'}{n}\) for \(\forall i\in N\) maximizes both \({\mathbb {P}}_{q|n}[{\mathbf {p}}]\) and \(-{\mathbf {p}}*{\mathbf {p}}\), the former by Proposition 1 and the latter by a simple argument.

  21. Because B in \(B{\mathbb {R}}_{q|n}[b]-b\) captures the importance of the proposal for the chair, the second effect does not arise when the proposal is important for the chair.

  22. Setting \(n=q=1\) and \(B\in (0,1)\cup (1,2)\), \(b^{*}=\frac{B}{2}\) in the benchmark \({{\mathbb {CE}}}\) while \(b^{*}=0\) if \(B<1\) and \(b^{*}=1\) if \(B>1\) with upfront payments. The change in \(b^{*}\) under upfront payments illustrates the first effect when \(B>1\) and the second effect when \(B<1\).

  23. Upfront payments lower the amount of favors distributed for a small overall budget and increase the amount of favors distributed for a large overall budget. The first effect is similar to the effect of upfront payments in Dekel et al. (2008) and is driven by the all-pay nature of the upfront payments. The second effect is specific to our model and arises because the upfront payments can be considered as an investment, and the chair ensures that her investment returns a profit (results in acceptance) by increasing the investment itself.

  24. As an example, consider binary voting over alternatives yielding utility \(x_{i}\) and y in a quantal response model such as the Quantal Response Equilibrium of McKelvey and Palfrey (1995). Then the probability that i votes for \(x_{i}\) is \(p_{i}(x_{i})=\frac{\exp {\lambda x_{i}}}{\exp {\lambda x_{i}}+\exp {\lambda y}}\), where \(\lambda \ge 0\) measures the precision in i’s best response. Thus, \(p_{i}\) is increasing as a function of \(x_{i}\), convex for \(x_{i}\le y\) and concave otherwise.

  25. For the Quantal Response Equilibrium mentioned in footnote 24, we have constructed an example with \(n=3\) and \(q=2\) to illustrate how concavity creates incentives for \(x_{i}\approx x_{j}\) and convexity for \(x_{i}\gg x_{j}\). As y increases from 0 to 1, the voting functions change in a smooth manner from concave to convex, assuming \(b=1\). Correspondingly, the optimal favor allocation focuses first on all 3 players and then switches to focusing on 2, minimum winning, players.

  26. The jump in the \({{\mathbb {CE}}}\) solution is driven by points where \({\mathbb {R}}_{q|n}\) is not differentiable. These points correspond to the values of b at which the \({{\mathbb {FA}}}\) solution switches between different coalition sizes. This shows that analysis of the \({{\mathbb {CE}}}\) problem cannot be conducted using calculus since \({\mathbb {R}}_{q|n}\) is endogenous.

  27. The empirical tests of Grossman and Helpman (1994) by Goldberg and Maggi (1999) and of Groseclose and Snyder (1996) by Wiseman (2004) are notable exceptions.

  28. Campos and Giovannoni (2007), Richter et al. (2009), Dorsch (2013) and Kang (2016) study lobbying empirically (see Binderkrantz 2014 and Gregor 2017, for surveys).

  29. The theorem is stated for \(t\in (0,1-p)\), but its discussion clarifies that for \(t=1-p\) the bound is \(p^{n}\) and \(p^{n}\le \exp {[-2n(1-p)^2]}\) \(\forall n\in {\mathbb {N}}_{\ge 1}\) and \(\forall p\in (0,1)\) is standard to verify.

  30. Hoeffding (1956, Theorem 2) shows that if \({\mathbf {p}}^{*}\) has two interior coordinates \(p_{i}^{*}\ne p_{j}^{*}\), then there exists another solution to \({{\mathbb {FA}}}\), \({\mathbf {p}}'\), identical to \({\mathbf {p}}^{*}\) except that \(p_{i}'=p_{j}'\). Lemma 4 is stronger, it shows that \({\mathbf {p}}^{*}\) with interior \(p_{i}^{*}\ne p_{j}^{*}\) does not exist.

  31. Theorem 1 in Darroch (1964) relies on all coordinates of \({\mathbf {p}}\) to lie strictly between zero and unity. To see we can apply the result, note that \({\mathbb {P}}^{*}_{s|n-3}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}^{*}_{s-u|n-3-u-z}[{\mathbf {p}}^{\{ijkuz\},*}]\) for \(s\in \{q-3,q-2,q-1\}\). We have, \(\forall s\in \{q-3,q-2,q-1\}\), \(s-u\ge q-3-u\ge 0\) by \(u\le q-3\), \(n-3-u-z\ge 2\) by \(z+u\le n-5\) and \(s-u\le n-3-u-z\Leftrightarrow z\le n-3-s\le n-q-2\) by \(z\le n-q-2\).

References

  • Aliprantis, C. D., & Border, K. C. (2006). Infinite dimensional analysis. A Hitchhiker's guide. Berlin: Springer.

    Google Scholar 

  • Amegashie, J. A. (2002). Committees and rent-seeking effort under probabilistic voting. Public Choice, 112(3–4), 345–350.

    Article  Google Scholar 

  • Austen-Smith, D., & Banks, J. S. (1988). Elections, coalitions, and legislative outcomes. American Political Science Review, 82(2), 405–422.

    Article  Google Scholar 

  • Axelrod, R. M. (1970). Conflict of interest: A theory of divergent goals with applications to politics. Chicago, IL: Markham Publishing.

    Google Scholar 

  • Banks, J. S. (2000). Buying supermajorities in finite legislatures. American Political Science Review, 94(3), 677–681.

    Article  Google Scholar 

  • Banks, J. S., & Duggan, J. (2000). A bargaining model of collective choice. American Political Science Review, 94(1), 73–88.

    Article  Google Scholar 

  • Banks, J. S., & Duggan, J. (2006). A general bargaining model of legislative policy-making. Quarterly Journal of Political Science, 1(1), 49–85.

    Article  Google Scholar 

  • Baron, D. P. (1991). A spatial bargaining theory of government formation in parliamentary systems. American Political Science Review, 85(1), 137–164.

    Article  Google Scholar 

  • Baron, D. P. (1993). Government formation and endogenous parties. American Political Science Review, 87(1), 34–47.

    Article  Google Scholar 

  • Baron, D. P., & Diermeier, D. (2001). Elections, governments, and parliaments in proportional representation systems. Quarterly Journal of Economics, 116(3), 933–967.

    Article  Google Scholar 

  • Baron, D. P., & Ferejohn, J. A. (1989). Bargaining in legislatures. American Political Science Review, 83(4), 1181–1206.

    Article  Google Scholar 

  • Bassi, A. (2013). A model of endogenous government formation. American Journal of Political Science, 57(4), 777–793.

    Google Scholar 

  • Bernheim, B. D., & Whinston, M. D. (1986). Menu auctions, resource allocation, and economic influence. Quarterly Journal of Economics, 101(1), 1–31.

    Article  Google Scholar 

  • Binderkrantz, A. S. (2014). Legislatures, lobbying, and interest groups. In S. Martin, T. Saalfeld, & K. W. Strom (Eds.), The Oxford handbook of legislative studies (pp. 526–542). Oxford: Oxford University Press.

    Google Scholar 

  • Boyer, P. C., Konrad, K. A., & Roberson, B. (2017). Targeted campaign competition, loyal voters, and supermajorities. Journal of Mathematical Economics, 71, 49–62.

    Article  Google Scholar 

  • Camerer, C. F. (2003). Behavioral game theory: Experiments in strategic interaction. Princeton, NJ: Princeton University Press.

    Google Scholar 

  • Campos, N. F., & Giovannoni, F. (2007). Lobbying, corruption and political influence. Public Choice, 131(1), 1–21.

    Article  Google Scholar 

  • Cardona, D., & Ponsati, C. (2007). Bargaining one-dimensional social choices. Journal of Economic Theory, 137(1), 627–651.

    Article  Google Scholar 

  • Cardona, D., & Ponsati, C. (2011). Uniqueness of stationary equilibria in bargaining one-dimensional policies under (super) majority rules. Games and Economic Behavior, 73(1), 65–75.

    Article  Google Scholar 

  • Carrubba, C. J., & Volden, C. (2000). Coalitional politics and logrolling in legislative institutions. American Journal of Political Science, 44(2), 261–277.

    Article  Google Scholar 

  • Chen, Y., & Eraslan, H. (2014). Rhetoric in legislative bargaining with asymmetric information. Theoretical Economics, 9(2), 483–513.

    Article  Google Scholar 

  • Corchon, L., & Dahm, M. (2010). Foundations for contest success functions. Economic Theory, 43(1), 81–98.

    Article  Google Scholar 

  • Dahm, M., & Glazer, A. (2015). A carrot and stick approach to agenda-setting. Journal of Economic Behavior & Organization, 116, 465–480.

    Article  Google Scholar 

  • Dahm, M., Dur, R., & Glazer, A. (2014). How a firm can induce legislators to adopt a bad policy. Public Choice, 159(1–2), 63–82.

    Article  Google Scholar 

  • Dal Bo, E. (2007). Bribing voters. American Journal of Political Science, 51(4), 789–803.

    Article  Google Scholar 

  • Darroch, J. N. (1964). On the distribution of the number of successes in independent trials. Annals of Mathematical Statistics, 35(3), 1317–1321.

    Article  Google Scholar 

  • Dekel, E., Jackson, M. O., & Wolinsky, A. (2008). Vote buying: General elections. Journal of Political Economy, 116(2), 351–380.

    Article  Google Scholar 

  • Dekel, E., Jackson, M. O., & Wolinsky, A. (2009). Vote buying: Legislatures and lobbying. Quarterly Journal of Political Science, 4(2), 103–128.

    Article  Google Scholar 

  • Diaconis, P., & Zabell, S. (1991). Closed form summation for classical distributions: Variations on a theme of de Moivre. Statistical Science, 6(3), 284–302.

    Article  Google Scholar 

  • Diermeier, D., & Merlo, A. (2000). Government turnover in parliamentary democracies. Journal of Economic Theory, 94(1), 46–79.

    Article  Google Scholar 

  • Diermeier, D., & Myerson, R. B. (1999). Bicameralism and its consequences for the internal organization of legislatures. American Economic Review, 89(5), 1182–1196.

    Article  Google Scholar 

  • Doron, G., & Sherman, M. (1995). A comprehensive decision-making exposition of coalition politics: The framer’s perspective of size. Journal of Theoretical Politics, 7(3), 317–333.

    Article  Google Scholar 

  • Dorsch, M. (2013). Bailout for sale? The vote to save wall street. Public Choice, 155(3–4), 211–228.

    Article  Google Scholar 

  • Dougherty, K. L., & Edward, J. (2012). Voting for Pareto optimality: A multidimensional analysis. Public Choice, 151(3–4), 655–678.

    Article  Google Scholar 

  • Eraslan, H. (2002). Uniqueness of stationary equilibrium payoffs in the Baron–Ferejohn model. Journal of Economic Theory, 103(1), 11–30.

    Article  Google Scholar 

  • Felgenhauer, M., & Gruner, H. P. (2008). Committees and special interests. Journal of Public Economic Theory, 10(2), 219–243.

    Article  Google Scholar 

  • Ferejohn, J. (1986). Incumbent performance and electoral control. Public Choice, 50(1–3), 5–25.

    Article  Google Scholar 

  • Frechette, G. R., Kagel, J. H., & Lehrer, S. F. (2003). Bargaining in legislatures: An experimental investigation of open versus closed amendment rules. American Political Science Review, 97(2), 221–232.

    Article  Google Scholar 

  • Frechette, G. R., Kagel, J. H., & Morelli, M. (2005). Behavioral identification in coalitional bargaining: An experimental analysis of demand bargaining and alternating offers. Econometrica, 73(6), 1893–1937.

    Article  Google Scholar 

  • Gamson, W. A. (1961). A theory of coalition formation. American Sociological Review, 26(3), 373–382.

    Article  Google Scholar 

  • Glazer, A., & McMillan, H. (1990). Optimal coalition size when making proposals is costly. Social Choice and Welfare, 7(4), 369–380.

    Article  Google Scholar 

  • Glazer, A., & McMillan, H. (1992). Amend the old or address the new: Broad-based legislation when proposing policies is costly. Public Choice, 74(1), 43–58.

    Article  Google Scholar 

  • Goldberg, P. K., & Maggi, G. (1999). Protection for sale: An empirical investigation. American Economic Review, 89(5), 1135–1155.

    Article  Google Scholar 

  • Gregor, M. (2017). Lobbying mechanisms. In N. Schofield & G. Caballero (Eds.), State, institutions and democracy: Contributions of political economy. Cham: Springer.

    Google Scholar 

  • Groseclose, T., & Snyder, J. M, Jr. (1996). Buying supermajorities. American Political Science Review, 90(2), 303–315.

    Article  Google Scholar 

  • Grossman, G. M., & Helpman, E. (1994). Protection for sale. American Economic Review, 84(4), 833–850.

    Google Scholar 

  • Hoeffding, W. (1956). On the distribution of the number of successes in independent trials. Annals of Mathematical Statistics, 27(3), 713–721.

    Article  Google Scholar 

  • Hoeffding, W. (1963). Probability inequalities for sums of bounded random variables. Journal of the American Statistical Association, 58(301), 13–30.

    Article  Google Scholar 

  • Kang, K. (2016). Policy influence and private returns from lobbying in the energy sector. Review of Economic Studies, 83(1), 269–305.

    Article  Google Scholar 

  • Koehler, D. H. (1975). Legislative coalition formation: The meaning of minimal winning size with uncertain participation. American Journal of Political Science, 19(1), 27–39.

    Article  Google Scholar 

  • Kuhn, H. W., & Tucker, A. W. (1951). Nonlinear programming. In J. Neyman (Ed.), Proceedings of the second Berkeley symposium on mathematical statistics and probability (pp. 481–492). Berkeley, CA: University of California Press.

  • Laver, M. (1998). Models of government formation. Annual Review of Political Science, 1(1), 1–25.

    Article  Google Scholar 

  • Laver, M., Marchi, S., & Mutlu, H. (2011). Negotiation in legislatures over government formation. Public Choice, 147(3–4), 285–304.

    Article  Google Scholar 

  • Le Breton, M., & Salanie, F. (2003). Lobbying under political uncertainty. Journal of Public Economics, 87(12), 2589–2610.

    Article  Google Scholar 

  • Le Breton, M., & Zaporozhets, V. (2007). Legislative lobbying under political uncertainty. IDEI Working Paper Series No. 493

  • Le Breton, M., & Zaporozhets, V. (2010). Sequential legislative lobbying under political certainty. Economic Journal, 120(543), 281–312.

    Article  Google Scholar 

  • Le Breton, M., Sudholter, P., & Zaporozhets, V. (2012). Sequential legislative lobbying. Social Choice and Welfare, 39(2–3), 491–520.

    Article  Google Scholar 

  • Lindbeck, A., & Weibull, J. (1987). Balanced-budget redistribution as the outcome of political competition. Public Choice, 52(3), 273–297.

    Article  Google Scholar 

  • Mandler, M. (2013). How to win a large election. Games and Economic Behavior, 78(1), 44–63.

    Article  Google Scholar 

  • McKelvey, R. D., & Palfrey, T. R. (1995). Quantal response equilibria for normal form games. Games and Economic Behavior, 10(1), 6–38.

    Article  Google Scholar 

  • Miller, L., & Vanberg, C. (2013). Decision costs in legislative bargaining: An experimental analysis. Public Choice, 155(3–4), 373–394.

    Article  Google Scholar 

  • Miller, L., & Vanberg, C. (2015). Group size and decision rules in legislative bargaining. European Journal of Political Economy, 37, 288–302.

    Article  Google Scholar 

  • Montero, M. (2007). Inequity aversion may increase inequity. Economic Journal, 117(519), C192–C204.

    Article  Google Scholar 

  • Morgan, J., & Vardy, F. (2011). On the buyability of voting bodies. Journal of Theoretical Politics, 23(2), 260–287.

    Article  Google Scholar 

  • Morgan, J., & Vardy, F. (2012). Negative vote buying and the secret ballot. Journal of Law, Economics, & Organization, 28(4), 818–849.

    Article  Google Scholar 

  • Myerson, R. B. (1993). Incentives to cultivate favored minorities under alternative electoral systems. American Political Science Review, 87(4), 856–869.

    Article  Google Scholar 

  • Norman, P. (2002). Legislative bargaining and coalition formation. Journal of Economic Theory, 102(2), 322–353.

    Article  Google Scholar 

  • Olver, F. W. J., Lozier, D. W., Boisvert, R. F., & Clark, C. W. (Eds.). (2010). NIST handbook of mathematical functions. New York, NY: Cambridge University Press.

    Google Scholar 

  • Pardalos, P. M. (2009). Kuhn–Tucker optimality conditions. In C. A. Floudas & P. M. Pardalos (Eds.), Encyclopedia of optimization (2nd ed., pp. 1794–1797). New York, NY: Springer.

    Chapter  Google Scholar 

  • Richter, B. K., Samphantharak, K., & Timmons, J. F. (2009). Lobbying and taxes. American Journal of Political Science, 53(4), 893–909.

    Article  Google Scholar 

  • Riker, W. H. (1962). The theory of political coalitions. London: Yale University Press.

    Google Scholar 

  • Romer, T., & Rosenthal, H. (1978). Political resource allocation, controlled agendas, and the status quo. Public Choice, 33(4), 27–43.

    Article  Google Scholar 

  • Romer, T., & Rosenthal, H. (1979). Bureaucrats versus voters: On the political economy of resource allocation by direct democracy. Quarterly Journal of Economics, 93(4), 563–587.

    Article  Google Scholar 

  • Saiegh, S. M. (2011). Ruling by statute: How uncertainty and vote buying shape lawmaking. New York, NY: Cambridge University Press.

    Book  Google Scholar 

  • Saiegh, S. M. (2014). Lawmaking. In S. Martin, T. Saalfeld, & K. W. Strom (Eds.), The Oxford handbook of legislative studies (pp. 481–513). Oxford: Oxford University Press.

    Google Scholar 

  • Seidmann, D. J. (2011). A theory of voting patterns and performance in private and public committees. Social Choice and Welfare, 36(1), 49–74.

    Article  Google Scholar 

  • Shepsle, K. A., & Weingast, B. R. (1981). Political preferences for the pork barrel: A generalization. American Journal of Political Science, 25(1), 96–111.

    Article  Google Scholar 

  • Snyder, J. M. (1991). On buying legislatures. Economics & Politics, 3(2), 93–109.

    Article  Google Scholar 

  • Tsai, T. S., & Yang, C. C. (2010a). Minimum winning versus oversized coalitions in public finance: The role of uncertainty. Social Choice and Welfare, 34(2), 345–361.

    Article  Google Scholar 

  • Tsai, T. S., & Yang, C. C. (2010b). On majoritarian bargaining with incomplete information. International Economic Review, 51(4), 959–979.

    Article  Google Scholar 

  • Tyutin, A., & Zaporozhets, V. (2017). On legislative lobbying under political uncertainty. TSE Working Paper Series No. 17-807

  • Volden, C., & Wiseman, A. E. (2007). Bargaining in legislatures over particularistic and collective goods. American Political Science Review, 101(1), 79–92.

    Article  Google Scholar 

  • Weingast, B. R. (1979). A rational choice perspective on congressional norms. American Journal of Political Science, 23(2), 245–262.

    Article  Google Scholar 

  • Wiseman, A. E. (2004). Tests of vote-buyer theories of coalition formation in legislatures. Political Research Quarterly, 57(3), 441–450.

    Article  Google Scholar 

Download references

Acknowledgements

I would like to thank discussant Matthias Dahm, associate editor and two anonymous referees. I also thank Sourav Bhattacharya, Matthew Ellman, Martin Gregor, Sjaak Hurkens, Marie Hušková, Navin Kartik, Michel Le Breton, Ines Macho-Stadler, Salvatore Nunnari, Clara Ponsatí, Jakub Steiner, Miroslav Zelený and seminar and conference participants at Universitat Autonoma de Barcelona, Columbia University, the Priorat Workshop in Bargaining and Politics, University of Malaga, EPCS 2014, University of St Andrews and EEA-ESEM 2014 for their helpful comments and discussions. Financial support by the Czech Science Foundation, grant 14-27902P, is gratefully acknowledged. All remaining errors are my own.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jan Zápal.

Appendix 1: Proofs

Appendix 1: Proofs

1.1 Preliminaries

We present a series of auxiliary results that facilitate the proofs below. First, consider a random variable \(X_{j}\) representing success or failure in a single Bernoulli trial with a probability of success \(p\in [0,1]\). The number of successes in n independent identical Bernoulli trials follows a Binomial distribution B(np) with the probability mass function f(knp), giving a probability of exactly \(k\in \{0,\ldots ,n\}\) successes.

Lemma 1

Suppose \(n\in {\mathbb {N}}_{\ge 1}\), \(k\in \{0,\ldots ,n-1\}\) and \(p\in (0,1)\). Then \(f(k+1,n,p)\ge f(k,n,p)\) if and only if \(k+1\le (n+1)p\).

Proof

Fix \(n\in {\mathbb {N}}_{\ge 1}\), \(k\in \{0,\ldots ,n-1\}\) and \(p\in (0,1)\). We have \(f(k,n,p)=\left( {\begin{array}{l}n\\ k\end{array}}\right) p^{k}(1-p)^{n-k}\) and, hence,

$$\begin{aligned} \frac{f(k+1,n,p)}{f(k,n,p)}=\frac{(n-k)p}{(k+1)(1-p)}. \end{aligned}$$
(8)

Thus, \(f(k+1,n,p)\ge f(k,n,p)\Leftrightarrow \frac{(n-k)p}{(k+1)(1-p)}\ge 1\Leftrightarrow np\ge k+1-p\).

We denote the probability of exactly k or more successes by \(F(k,n,p)=\sum _{s=k}^{n}f(s,n,p)\). Below, we often differentiate F(knp) with respect to p. To this end, it is helpful to express F(knp) in terms of a regularized incomplete beta function (see Olver et al. 2010, for details):

$$\begin{aligned} F(k,n,p)=I_{p}(k,n-(k-1)) \end{aligned}$$
(9)

where

$$\begin{aligned} I_{x}(a,b)&=\frac{B(x,a,b)}{B(a,b)}\\ B(x,a,b)&=\int _{0}^{x}t^{a-1}(1-t)^{b-1}dt\\ B(a,b)&=\frac{(a-1)!(b-1)!}{(a+b-1)!}\\ \end{aligned}$$
(10)

are a regularized incomplete beta function, an incomplete beta function and a complete beta function, respectively. The derivative of F(knp) for \(p\in (0,1)\) and \(k\in \{0,\ldots ,n\}\) is thus equal to

$$\begin{aligned} \frac{\partial F(k,n,p)}{\partial p}=\frac{k}{p}f(k,n,p). \end{aligned}$$
(11)

Note that, for any \(n\in {\mathbb {N}}_{\ge 1}\) and \(k\in \{0,\ldots ,n\}\), \(f(k,n,p)>0\) when \(p\in (0,1)\).

Working with F(knp) is in general problematic. However, it can be bounded for large n. The following is a direct application of Theorem 1 in Hoeffding (1963) to the random variables \(X_{j}\) representing the success or failure in the Bernoulli trials.

Theorem 1

(Hoeffding 1963) Let random variable \(X_{j}\in \{0,1\}\) represent success (\(X_{j}=1\)) or failure (\(X_{j}=0\)) with \({\mathbb {P}}[X_{j}=1]=p\in (0,1)\) for \(\forall j\in \{1,\ldots ,n\}\). Then

$$\begin{aligned} {\mathbb {P}}\left[ \textstyle \sum _{j=1}^{n}X_{j}\ge n(p+t)\right] \le \exp {[-2nt^{2}]} \end{aligned}$$
(12)

for \(0<t\le 1-p\) and \(\forall n\in {\mathbb {N}}_{\ge 1}\).Footnote 29

The final result represents the bound on the upper tail of the Binomial distribution expressed as a fraction of its probability mass function.

Theorem 2

(Diaconis and Zabell 1991) For any integer k satisfying \(k>np\) where \(n\in {\mathbb {N}}_{\ge 1}\) and \(p\in (0,1)\)

$$\begin{aligned} \frac{F(k,n,p)}{f(k,n,p)}\le \frac{k(1-p)}{k-np}. \end{aligned}$$
(13)

We use the bound on the upper tail since an exact expression in terms of a hypergeometric function (for details see Olver et al. 2010, Section 8.17), that is,

$$\begin{aligned} \frac{F(k,n,p)}{f(k,n,p)}=(1-p)\,_{2}F_{1}(1,1+n,1+k,p) \end{aligned}$$
(14)

is difficult to work with. The hypergeometric function itself is defined as

$$\begin{aligned} \,_{2}F_{1}(a,b,c,z)=\sum _{s=0}^{\infty }\frac{a)_{s}b)_{s}}{c)_{s}}\frac{z^{s}}{s!} \end{aligned}$$
(15)

where \((a)_{n}\) is Pochhammer’s symbol defined as \(a(a+1)\ldots (a+n-1)\) if \(n\ge 1\) and 1 for \(n=0\). From Olver et al. (2010, Section 15.4), we have \(\lim _{p\rightarrow 1^{-}}(1-p)\,_{2}F_{1}(1,1+n,1+n,p)=1\).

Despite F(knp) / f(knp) in general being not amenable to manipulation, it is monotone in some variables.

Lemma 2

Suppose \(n\in {\mathbb {N}}_{\ge 1}\), \(k\in \{0,\ldots ,n\}\) and \(p\in (0,1)\). Then \(\frac{F(k,n,p)}{f(k,n,p)}\) is nondecreasing in p.

Proof

Fix \(n\in {\mathbb {N}}_{\ge 1}\), \(k\in \{0,\ldots ,n\}\) and \(p\in (0,1)\). We have

$$\begin{aligned} \frac{F(k,n,p)}{f(k,n,p)}=\frac{\sum _{s=k}^{n}f(s,n,p)}{f(k,n,p)}=\sum _{s=k}^{n}\frac{f(s,n,p)}{f(k,n,p)}. \end{aligned}$$
(16)

For any \(s\in \{k,\ldots ,n\}\), we have

$$\begin{aligned} \frac{f(s,n,p)}{f(k,n,p)}=\frac{\left( {\begin{array}{l}n\\ s\end{array}}\right) }{\left( {\begin{array}{l}n\\ k\end{array}}\right) }\frac{p^{s}}{p^{k}}\frac{(1-p)^{n-s}}{(1-p)^{n-k}}=\frac{\left( {\begin{array}{l}n\\ s\end{array}}\right) }{\left( {\begin{array}{l}n\\ k\end{array}}\right) }\left( \frac{p}{1-p}\right) ^{s-k} \end{aligned}$$
(17)

and the lemma follows since \(\frac{\partial }{\partial p}\frac{p}{1-p}=\frac{1}{(1-p)^{2}}\ge 0\).

Lemma 3

Suppose \(n\in {\mathbb {N}}_{\ge 1}\), \(k\in \{1,\ldots ,n\}\) and \(r\in \{0,\ldots ,n-k\}\). Then \(\frac{F(k,n-r,\frac{k}{n-r})}{f(k,n-r,\frac{k}{n-r})}\) is nonincreasing in r.

Proof

Fix \(n\in {\mathbb {N}}_{\ge 1}\), \(k\in \{1,\ldots ,n\}\) and \(r\in \{0,\ldots ,n-k\}\). Since \(r\in \{0,\ldots ,n-k\}\), \(n-r\in \{k,\ldots ,n\}\) and thus \(\frac{k}{n-r}\in (0,1)\) unless \(r=n-k\). We thus have

$$\begin{aligned} \frac{F(k,n-r,\frac{k}{n-r})}{f(k,n-r,\frac{k}{n-r})}=\sum _{s=k}^{n-r}\frac{f(s,n-r,\frac{k}{n-r})}{f(k,n-r,\frac{k}{n-r})}. \end{aligned}$$
(18)

When \(r\in \{0,\ldots ,n-k-1\}\) increases to \(r+1\), the sum in (18) is over fewer terms. It thus suffices to show that \(f(s,n-r,\frac{k}{n-r})/f(k,n-r,\frac{k}{n-r})\) is nonincreasing in r. For \(s\in \{k,\ldots ,n-r\}\) we have

$$\begin{aligned} \frac{f(s,n-r,\frac{k}{n-r})}{f(k,n-r,\frac{k}{n-r})}&=\frac{\left( {\begin{array}{l}n-r\\ s\end{array}}\right) }{\left( {\begin{array}{l}n-r\\ k\end{array}}\right) }\frac{\left( \frac{k}{n-r}\right) ^{s}}{\left( \frac{k}{n-r}\right) ^{k}}\frac{\left( \frac{n-r-k}{n-r}\right) ^{n-r-s}}{\left( \frac{n-r-k}{n-r}\right) ^{n-r-k}}\\&=\frac{k!}{s!}\frac{(n-r-k)!}{(n-r-s)!}\left( \frac{k}{n-r-k}\right) ^{s-k}\\&=\frac{k!}{s!}k^{s-k}\underbrace{\frac{n-r-k}{n-r-k}\frac{n-r-k-1}{n-r-k}\cdot \ldots \cdot \frac{n-r-s+1}{n-r-k}}_{s-k}.\\ \end{aligned}$$
(19)

Each of the last \(s-k\) terms writes \(\frac{n-r-k-i}{n-r-k}\) for \(i\in \{0,\ldots ,s-k-1\}\) and the lemma follows since \(\frac{\partial }{\partial r}\frac{n-r-k-i}{n-r-k}=\frac{-i}{(n-r-k)^2}\le 0\).

1.2 Proof of Proposition 1

First, note that the \({{\mathbb {FA}}}\) problem has a solution, as it involves maximization of the continuous objective function \({\mathbb {P}}_{q|n}\) over a compact region X(b). \({\mathbb {P}}_{q|n}\) is differentiable with respect to every coordinate of \({\mathbf {p}}\); hence, any solution to \({{\mathbb {FA}}}\) necessarily satisfies the standard Kuhn and Tucker (1951) conditions. No further constraint qualification is required, as the constraints of \({{\mathbb {FA}}}\) are cut out by affine functions (Pardalos 2009). The Lagrangian for \({{\mathbb {FA}}}\) is

$$\begin{aligned} L({\mathbf {p}},\lambda ,{\mathbf {m}}^{+},{\mathbf {m}}^{-})={\mathbb {P}}_{q|n}[{\mathbf {p}}]-\lambda \left[ \sum _{i=1}^{n}p_{i}-b\right] -{\mathbf {m}}^{+}*({\mathbf {p}}-{\mathbf {1}})+{\mathbf {m}}^{-}*{\mathbf {p}} \end{aligned}$$
(20)

where \(\lambda \), \({\mathbf {m}}^{+}=(m_{i}^{+})_{i\in N}\) and \({\mathbf {m}}^{-}=(m_{i}^{-})_{i\in N}\) are Lagrange multipliers and \({\mathbf {1}}\) is the unit vector in \({\mathbb {R}}^{n}\). Throughout, we denote the solution to \({{\mathbb {FA}}}\) by \({\mathbf {p}}^{*}\), \(\lambda ^{*}\), \({\mathbf {m}}^{+,*}\) and \({\mathbf {m}}^{-,*}\). The typical element of \({\mathbf {p}}^{*}\) is \(p_{i}^{*}\) and similarly for \({\mathbf {m}}^{+,*}\) and \({\mathbf {m}}^{-,*}\).

To prove part 1, note that, \(\forall {\mathbf {p}}\in X(b)\), \({\mathbb {P}}_{q|n}[{\mathbf {p}}]\le 1\). If \(b\ge q\), then for any \({\mathbf {p}}'\) with \(|\{i\in N|p_{i}'=1\}|\ge q\) we have \({\mathbb {P}}_{q|n}[{\mathbf {p}}']=1\) and, hence, \({\mathbf {p}}'\) is a solution to \({{\mathbb {FA}}}\). Conversely, for any \({\mathbf {p}}'\) with \(|\{i\in N|p_{i}'=1\}|<q\) we have \({\mathbb {P}}_{q|n}[{\mathbf {p}}']<1\) and, hence, \({\mathbf {p}}'\) is not a solution to \({{\mathbb {FA}}}\).

The proof of part 2 is significantly more involved. Suppose \(b<q\). Simple argument shows that the constraint presented by b has to be binding in any solution to \({{\mathbb {FA}}}\), so that \(\sum _{i=1}^{n}p_{i}^{*}=b\). The following lemma helps to prove the remaining claims.

Lemma 4

Suppose \(b<q\). Given \({\mathbf {p}}^{*}\), if \(p_{i}^{*}\in (0,1)\) and \(p_{j}^{*}\in (0,1)\) for some \(i\in N\) and \(j\in N{\setminus }\{i\}\), then \(p_{i}^{*}=p_{j}^{*}\).Footnote 30

Proof

Using (20), the first-order necessary condition for the optimality of \({\mathbf {p}}^{*}\) for any \(p_{i}^{*}\in (0,1)\) is

$$\begin{aligned} \frac{\partial L({\mathbf {p}}^{*},\lambda ^{*},{\mathbf {m}}^{+,*},{\mathbf {m}}^{-,*})}{\partial p_{i}}=\frac{\partial {\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]}{\partial p_{i}}-\lambda ^{*}=0. \end{aligned}$$
(21)

Let \({\mathbf {p}}^{\{ij\}}\) be obtained from \({\mathbf {p}}\) by dropping \(p_{i}\) and \(p_{j}\) and, similarly, \(N^{\{ij\}}=N{\setminus }\{i,j\}\). Recall that \({\mathbb {P}}_{s|n}^{*}[{\mathbf {p}}]\) denotes the probability of exactly s out of n committee members accepting the chair’s proposal when allocated favors \({\mathbf {p}}\). We have

$$\begin{aligned} {\mathbb {P}}_{q|n}[{\mathbf {p}}]&=\,{\mathbb {P}}_{q|n-2}[{\mathbf {p}}^{\{ij\}}](1-p_{i})(1-p_{j})\\&\quad +\,{\mathbb {P}}_{q-1|n-2}[{\mathbf {p}}^{\{ij\}}](p_{i}(1-p_{j})+(1-p_{i})p_{j})\\&\quad +\,{\mathbb {P}}_{q-2|n-2}[{\mathbf {p}}^{\{ij\}}]p_{i}p_{j}\\ &=\,{\mathbb {P}}_{q|n-2}[{\mathbf {p}}^{\{ij\}}]+(p_{i}+p_{j}){\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]\\&\quad +\, p_{i}p_{j}\left[ {\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]-{\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]\right] .\\ \end{aligned}$$
(22)

\({\mathbb {P}}_{q|n}^{*}\) in (22) is well defined if \(n\ge 1\) and \(n\ge q\ge 0\). For \(n\ge 1\) and \(q>n\) or \(q<0\), we use, by convention, \({\mathbb {P}}_{q|n}^{*}=0\). Similarly, for \(n=q=0\), we use \({\mathbb {P}}_{q|n}^{*}=1\) and for \(n=0\) and \(q\ne 0\) we use \({\mathbb {P}}_{q|n}^{*}=0\). We do not need to consider \(n<0\), as the lemma clearly holds for \(n=1\).

If \(p_{i}^{*}\in (0,1)\) and \(p_{j}^{*}\in (0,1)\) such that \(i\ne j\) exist, then from (21), we have \(\frac{\partial {\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]}{\partial p_{i}}=\frac{\partial {\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]}{\partial p_{j}}\) and thus, using (22),

$$\begin{aligned} \left( p_{i}^{*}-p_{j}^{*}\right) \left[ {\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]-{\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\right] =0. \end{aligned}$$
(23)

Suppose, towards a contradiction, \(p_{i}^{*}\in (0,1)\), \(p_{j}^{*}\in (0,1)\) and \(p_{i}^{*}\ne p_{j}^{*}\). From (23), \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]={\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\). This is impossible with \(n=2\). Hence, suppose \(n\ge 3\).

First, we claim that \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]={\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\) and \(n\ge 3\) implies \(2\le q\le n-1\). Suppose \(q=1\). Then \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0\) and \({\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=\prod _{k\in N^{\{ij\}}}(1-p_{k}^{*})\), so that \({\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0\) implies \(\sum _{i\in N}p_{i}^{*}\ge n-2\ge 1\), a contradiction to \(b<q=1\). Alternatively, suppose \(q=n\). Then \({\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0\), so that \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0\) implies that \({\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]=0\), a contradiction to optimality of \({\mathbf {p}}^{*}\) as \({\mathbb {P}}_{q|n}[{\mathbf {p}}(0,\frac{b}{n},0)]>0\). Hence, suppose \(q\in \{2,\ldots ,n-1\}\).

Second, we claim that \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]={\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]>0\). Suppose not. Then, unless \({\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]=0\), there exists \(s\in \{q,\ldots ,n-2\}\) such that \({\mathbb {P}}_{s|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]>0\). Since, for some \(s\in \{q,\ldots ,n-2\}\), \({\mathbb {P}}_{s|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]>0={\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]={\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\), we have \(\sum _{i\in N}p_{i}^{*}\ge q\), a contradiction to \(b<q\).

As \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]={\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]>0\), there exists at least one coordinate of \({\mathbf {p}}^{\{ij\},*}\), \(p_{k}^{*}\), such that \(p_{k}^{*}\in (0,1)\). If all coordinates of \({\mathbf {p}}^{\{ij\},*}\) were equal either to zero or unity, then we would have \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\in \{0,1\}\) and \({\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\in \{0,1\}\) with both probabilities equal to unity impossible. Rewriting, we have

$$\begin{aligned} {\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]-{\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]= & {} \,p_{k}^{*}\left[ {\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]-{\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\right] \\& +\,(1-p_{k}^{*})\left[ {\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]-{\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\right] .\\ \end{aligned}$$
(24)

Third, we claim that \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\ne {\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\). To see this, suppose that \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\). Then, by (24), \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\), since \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]-{\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0\). If \(n=3\), then \(q\in \{2,\ldots ,n-1\}\) implies \(q=2\) and we have \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\ne {\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=1\). If \(n=4\), then \(q\in \{2,\ldots ,n-1\}\) implies \(q\in \{2,3\}\). Suppose \(q=2\). Then \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\), \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=1-{\mathbf {p}}^{\{ijk\},*}\) and \({\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]={\mathbf {p}}^{\{ijk\},*}\), and thus either \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\ne {\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) or \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\ne {\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\). Suppose \(q=3\). Then \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=1-{\mathbf {p}}^{\{ijk\},*}\), \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]={\mathbf {p}}^{\{ijk\},*}\) and \({\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\), and again either \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\ne {\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) or \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\ne {\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\). If \(n\ge 5\), we argue first that \(q\in \{2,n-1\}\) is not possible. To see this, suppose \(q=2\). Then \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\) and \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=\prod _{l\in N^{\{ijk\}}}(1-p_{l}^{*})\) so that \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\) implies \(\sum _{i\in N}p_{i}^{*}\ge n-3\ge 2\), a contradiction to \(b<q=2\). Alternatively, suppose \(q=n-1\). Then \({\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\), so that \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\) implies that \({\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]=0\), a contradiction to optimality of \({\mathbf {p}}^{*}\) as \({\mathbb {P}}_{q|n}[{\mathbf {p}}(0,\frac{b}{n},0)]>0\).

For \(n\ge 5\) and \(q\in \{3,\ldots ,n-2\}\), we have \({\mathbb {P}}_{s|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) independent of s for \(s\in \{q-3,q-2,q-1\}\). If \({\mathbb {P}}_{s|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\) \(\forall s\in \{q-3,q-2,q-1\}\), then, unless \({\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]=0\), there exists \(s'\in \{q,\ldots ,n\}\) such that \({\mathbb {P}}_{s'|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]>0\) and, hence, as \({\mathbb {P}}_{s|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\) \(\forall s\in \{q-3,q-2,q-1\}\), \(\sum _{i\in N}p_{i}^{*}\ge q\), a contradiction to \(b<q\). If, \({\mathbb {P}}_{s|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=1\) for some \(s\in \{q-3,q-2,q-1\}\), then \({\mathbb {P}}_{s'|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=0\) \(\forall s'\in \{q-3,q-2,q-1\}{\setminus }\{s\}\) and, hence, \({\mathbb {P}}_{s|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) depends on s. Hence, if \({\mathbb {P}}_{s|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) is independent of s, then \({\mathbb {P}}_{s|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]=c\in (0,1)\) \(\forall s\in \{q-3,q-2,q-1\}\). This implies that the number of unit coordinates in \({\mathbf {p}}^{\{ijk\},*}\), u, satisfies \(u\le q-3\), that there are at least two interior coordinates in \({\mathbf {p}}^{\{ijk\},*}\) and thus that the number of zero or unit coordinates in \({\mathbf {p}}^{\{ijk\},*}\), \(z+u\), satisfies \(z+u\le n-5\).

Furthermore, we argue that \(z\le n-q-2\). Since the number of committee members receiving positive favors \(3+u+n-3-z-u=n-z\) has to be at least q, we have \(z\le n-q\) and it thus suffices to rule out \(z\in \{n-q-1,n-q\}\). Note that \({\mathbb {P}}^{*}_{s|n-3}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}^{*}_{s-u|n-3-u-z}[{\mathbf {p}}^{\{ijkuz\},*}]\) for \(s\in \{q-3,q-2,q-1\}\), where \({\mathbf {p}}^{\{ijkuz\},*}\) denotes \({\mathbf {p}}^{*}\) after dropping \(p_{i}^{*}\), \(p_{j}^{*}\), \(p_{k}^{*}\) and all unit and zero coordinates. If \(z=n-q\), then we have \({\mathbb {P}}^{*}_{q-2|n-3}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}^{*}_{q-2-u|q-3-u}[{\mathbf {p}}^{\{ijkuz\},*}]=0\), since \(q-3-u\ge 0\) by \(u\le q-3\), and \({\mathbb {P}}^{*}_{q-3|n-3}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}^{*}_{q-3-u|q-3-u}[{\mathbf {p}}^{\{ijkuz\},*}]\ne 0\), a contradiction. If \(z=n-q-1\), then we have \({\mathbb {P}}^{*}_{q-1|n-3}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}^{*}_{q-1-u|q-2-u}[{\mathbf {p}}^{\{ijkuz\},*}]=0\), since \(q-2-u\ge 0\) by \(u\le q-3\), and \({\mathbb {P}}^{*}_{q-2|n-3}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}^{*}_{q-2-u|q-2-u}[{\mathbf {p}}^{\{ijkuz\},*}]\ne 0\), a contradiction.

Thus we conclude that \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) and \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]={\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) is not possible since \({\mathbb {P}}_{q|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) is, as a function of q, first increasing and then decreasing, except possibly for two equal maxima (Darroch 1964).Footnote 31

Finally, we claim that \(p_{i}^{*}\ne p_{j}^{*}\) provides a contradiction. From \(p_{i}^{*}\ne p_{j}^{*}\), either \(p_{i}^{*}\ne p_{k}^{*}\) or \(p_{j}^{*}\ne p_{k}^{*}\). Without loss of generality, assume \(p_{i}^{*}\ne p_{k}^{*}\). Since \({\mathbb {P}}_{q-3|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\ne {\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\), \({\mathbb {P}}_{q-2|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\ne {\mathbb {P}}_{q-1|n-3}^{*}[{\mathbf {p}}^{\{ijk\},*}]\) and thus, replacing \(p_{k}^{*}\) with \(p_{i}^{*}\) in (24), we have

$$\begin{aligned} {\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{jk\},*}]-{\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{jk\},*}]\ne 0 \end{aligned}$$
(25)

for any \(n\ge 3\) and \(q\in \{2,\ldots ,n-1\}\). Hence, by (23), \(p_{j}^{*}=p_{k}^{*}\) as \(p_{k}^{*}\in (0,1)\). Moreover, \(p_{i}^{*}\ne p_{j}^{*}\) implies \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]-{\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0\) by (23) and, hence, by (22), \({\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]\) depends only on \(p_{i}^{*}+p_{j}^{*}\). Therefore, for all sufficiently small \(\epsilon \ne 0\), \({\mathbf {p}}^{\prime }\) identical to \({\mathbf {p}}^{*}\) except that \(p_{i}^{\prime }=p_{i}^{*}-\epsilon \) and \(p_{j}^{\prime }=p_{j}^{*}+\epsilon \) also solves \({{\mathbb {FA}}}\). Repeating the same argument that lead to \(p_{j}^{*}=p_{k}^{*}\), using \(p_{i}^{\prime }\) and \(p_{j}^{\prime }\) instead, we have \(p_{j}^{\prime }=p_{k}^{\prime }\), which is contradiction since \(p_{j}^{*}=p_{k}^{*}=p_{k}^{\prime }\) implies \(p_{j}^{*}=p_{j}^{\prime }\) but \(p_{j}^{*}\ne p_{j}^{\prime }\).

Fix \({\mathbf {p}}^{*}\). By Lemma 4, \({\mathbf {p}}^{*}={\mathbf {p}}(r^{*},p^{*},s^{*})\) where \(p^{*}=\frac{b-s^{*}}{n-r^{*}-s^{*}}\). What remains to be shown are \(s^{*}=0\), \(r^{*}\le n-q\) and, when \(b<q-1\), \(r^{*}=0\). That \(r^{*}\le n-q\) is obvious. If \(r^{*}>n-q\), then the number of committee members receiving non-zero favors is \(n-r^{*}<q\) and \({\mathbb {P}}_{q|n}[{\mathbf {p}}(r^{*},p^{*},s^{*})]=0\).

For \(s^{*}=0\) and \(r^{*}=0\) if \(b<q-1\), first note that both hold for \(n=1\), thus suppose \(n\ge 2\). We proceed by analyzing the first order necessary conditions for optimality derived from (20). These read

$$\begin{aligned}\frac{\partial {\mathbb {P}}_{q|n}({\mathbf {p}}^{*})}{\partial p_{i}}-\lambda ^{*}+m^{-,*}_{i}=0 &\quad {\text {if }}\,p_{i}^{*}=0\\ \frac{\partial {\mathbb {P}}_{q|n}({\mathbf {p}}^{*})}{\partial p_{i}}-\lambda ^{*}=0 & \quad {\text {if }}\,p_{i}^{*}\in (0,1)\\ \frac{\partial {\mathbb {P}}_{q|n}({\mathbf {p}}^{*})}{\partial p_{i}}-\lambda ^{*}-m^{+,*}_{i}=0&\quad {\text {if }}\,p_{i}^{*}=1\\ \end{aligned}$$
(26)

which, using the non-negativity of the Lagrange multipliers, can be rewritten as

$$\begin{aligned} \left. \frac{\partial {\mathbb {P}}_{q|n}({\mathbf {p}}^{*})}{\partial p_{i}}\right| _{p_{i}^{*}=1}\ge \left. \frac{\partial {\mathbb {P}}_{q|n}({\mathbf {p}}^{*})}{\partial p_{i}}\right| _{p_{i}^{*}\in (0,1)}\ge \left. \frac{\partial {\mathbb {P}}_{q|n}({\mathbf {p}}^{*})}{\partial p_{i}}\right| _{p_{i}^{*}=0} \end{aligned}$$
(27)

and implies that

$$\begin{aligned} {\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\ge {\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}] \end{aligned}$$
(28)

for any pair \(p_{i}^{*}>p_{j}^{*}\) by (22).

Consider \(q=1\). Then \(s^{*}=0\) holds since \(b<q\) and \(r^{*}=0\) if \(b<q-1\) holds vacuously as \(q-1=0\). Consider \(q=n\). Suppose, towards a contradiction, that \(s^{*}>0\). Then there exists \(p_{i}^{*}=1\) and, since \(b<q\), \(p_{j}^{*}<1\), so that (30) must hold. The left hand side of (30) equals zero and, hence, \({\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0\) so that \({\mathbb {P}}_{q|n}[{\mathbf {p}}(r^{*},p^{*},s^{*})]=0\), a contradiction. Suppose, towards another contradiction, that \(r^{*}>0\). Then \({\mathbb {P}}_{q|n}[{\mathbf {p}}(r^{*},p^{*},s^{*})]=0\) since \(q=n\), a contradiction. By covering \(q=1\) and \(q=n\), we have covered \(n=2\). Thus suppose \(n\ge 3\) and \(q\in \{2,\ldots ,n-1\}\).

Consider the number of entries in \({\mathbf {p}}^{*}={\mathbf {p}}(r^{*},p^{*},s^{*})\) that differ from zero and unity, \(n-r^{*}-s^{*}\). We argue that \(n-r^{*}-s^{*}\ge 2\). If \(n-r^{*}-s^{*}=0\), then b is an integer and \(b<q\) implies \(b\le q-1\), so that \(s^{*}\le q-1\) and, hence, \({\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]=0\). To see that \(n-r^{*}-s^{*}\ne 1\), suppose, towards a contradiction, that \(n-r^{*}-s^{*}=1\). Then, unless \({\mathbb {P}}_{q|n}[{\mathbf {p}}^{*}]=0\), we have \(s^{*}=q-1\). Pick \(p_{i}^{*}=1\) and \(p_{j}^{*}\in (0,1)\). Then \({\mathbf {p}}^{\{ij\},*}\) in (30) is \(n-2\) vector with \(q-2\) unit coordinates and \(n-q\) zero coordinates, and, hence, \({\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0<1={\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\), a contradiction. Thus \(n-r^{*}-s^{*}\ge 2\).

Moreover, we argue that \(r^{*}=n-q\) when \(s^{*}\ge 1\) is not possible. Suppose, towards a contradiction, that \(r^{*}=n-q\) and \(s^{*}\ge 1\). Pick \(p_{i}^{*}=1\) and \(p_{j}^{*}\in (0,1)\). Then \({\mathbf {p}}^{\{ij\},*}\) in (30) is \(n-2\) vector with \(n-r^{*}-2=q-2\) positive coordinates, and, hence, \({\mathbb {P}}_{q-1|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]=0<{\mathbb {P}}_{q-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]\), a contradiction.

We now show that \(s^{*}=0\). Suppose, towards a contradiction, that \(s^{*}\ge 1\). Since \(s^{*}\ge 1\) and \(n-r^{*}-s^{*}\ge 2\), we can pick \(p_{i}^{*}=1\) and \(p_{j}^{*}\in (0,1)\). Then \({\mathbf {p}}^{\{ij\},*}\) has \(r^{*}\) zero coordinates, \(s^{*}-1\) unit coordinates and \(n-2-r^{*}-(s^{*}-1)\) of the \(p^{*}\) coordinates. In (30), we have \({\mathbb {P}}_{s|n-2}[{\mathbf {p}}^{\{ij\},*}]={\mathbb {P}}_{s-(s^{*}-1)|n-2-r^{*}-(s^{*}-1)}[{\mathbf {p}}^{\{ijuz\},*}]\) for \(s\in \{q-2,q-1\}\), where \({\mathbf {p}}^{\{ijuz\},*}\) denotes \({\mathbf {p}}^{*}\) after dropping \(p_{i}^{*}\), \(p_{j}^{*}\), \(r^{*}\) zero coordinates and \(s^{*}-1\) unit coordinates. We have \(n-2-r^{*}-(s^{*}-1)=n-r^{*}-s^{*}-1\ge 1\) by \(n-r^{*}-s^{*}\ge 2\), \(s-(s^{*}-1)\ge 0\) \(\forall s\in \{q-2,q-1\}\) by \(s^{*}\le q-1\), and \(s-(s^{*}-1)\le n-2-r^{*}-(s^{*}-1)\Leftrightarrow r^{*}\le n-2-s\) \(\forall s\in \{q-2,q-1\}\) by \(r^{*}\le n-q-1\) when \(s^{*}\ge 1\). Therefore, by Lemma 1, we can rewrite (30) as

$$\begin{aligned} \left( n-2-r^{*}-(s^{*}-1)+1\right) p^{*}&=b-s^{*}\\&\ge q-2-(s^{*}-1)+1\\ \end{aligned}$$
(29)

which simplifies to \(b\ge q\), a contradiction.

We now show that \(r^{*}=0\) if \(b<q-1\). Suppose, towards a contradiction, that \(r^{*}\ge 1\) and \(b<q-1\). Since \(r^{*}\ge 1\) and \(n-r^{*}\ge 2\), we can pick \(p_{j}^{*}=0\) and \(p_{i}^{*}\in (0,1)\). Then \({\mathbf {p}}^{\{ij\},*}\) has \(r^{*}-1\) zero coordinates and \(n-2-(r^{*}-1)\) of the \(p^{*}\) coordinates. In (30), we have \({\mathbb {P}}_{s|n-2}[{\mathbf {p}}^{\{ij\},*}]={\mathbb {P}}_{s|n-2-(r^{*}-1)}[{\mathbf {p}}^{\{ijz\},*}]\) for \(s\in \{q-2,q-1\}\), where \({\mathbf {p}}^{\{ijz\},*}\) denotes \({\mathbf {p}}^{*}\) after dropping \(p_{i}^{*}\), \(p_{j}^{*}\) and \(r^{*}-1\) zero coordinates. We have \(n-2-(r^{*}-1)=n-r^{*}-1\ge 1\) by \(n-r^{*}\ge 2\), \(s\ge 0\) \(\forall s\in \{q-2,q-1\}\) by \(q\ge 2\), and \(s\le n-2-(r^{*}-1)\Leftrightarrow r^{*}\le n-1-s\) \(\forall s\in \{q-2,q-1\}\) by \(r^{*}\le n-q\). Therefore, by Lemma 1, we can rewrite (30) as

$$\begin{aligned} \left( n-2-(r^{*}-1)+1\right) p^{*}&=b\\&\ge q-2+1\\ \end{aligned}$$
(30)

which simplifies to \(b\ge q-1\), a contradiction. \(\square \)

1.3 Proof of Proposition 2

We first observe that \({{\mathbb {CE}}}\), \(\max _{b\in [0,B]}(B-b){\mathbb {R}}_{q|n}[b]\), admits a solution by the Weierstrass theorem since \({\mathbb {R}}_{q|n}[b]\) is continuous in b, which in turn follows from the Theorem of the Maximum (Aliprantis and Border 2006, Theorem 17.31). We also observe that the solution to \({{\mathbb {CE}}}\), \(b^{*}\), has to satisfy \(b^{*}\le q\) for any B. For \(b\ge q\), \({\mathbb {R}}_{q|n}[b]=1\), hence, the objective function of \({{\mathbb {CE}}}\) is decreasing in b for \(b>q\).

For part 1, we must show that \(b^{*}\ne q\) when \(B<q+1\). Suppose, towards a contradiction, that \(b^{*}=q\) and \(B<q+1\). Since \(b^{*}=q\), then \({\mathbb {R}}_{q|n}[b^{*}]=1\), and the value of the objective function in \({{\mathbb {CE}}}\) equals \(B-q\). Since \(b^{*}=q\), by Proposition 1 part 1, \({\mathbf {p}}^{*}\) that solves \({{\mathbb {FA}}}\) given \(b^{*}=q\) satisfies \(|\{i\in N|p_{i}^{*}=1\}|\ge q\). Consider \({\mathbf {p}}'\) identical to \({\mathbf {p}}^{*}\) except that for i such that \(p_{i}^{*}=1\), \(p_{i}'=1-\epsilon \) for some small \(\epsilon >0\). The objective function in \({{\mathbb {CE}}}\) evaluated at \({\mathbf {p}}'\) is \((B-q+\epsilon )(1-\epsilon )\). Since \(b^{*}=q\), we must have

$$\begin{aligned} B-q&\ge (B-q+\epsilon )(1-\epsilon )\\ B&\ge q+1-\epsilon \\ \end{aligned}$$
(31)

for any \(\epsilon >0\). However, since \(B<q+1\), there always exists \(\epsilon >0\) such that \(B<q+1-\epsilon \), a contradiction.

For part 2, suppose, towards a contradiction, that \(b^{*}\) implies \(r^{*}=n-q\) in \({{\mathbb {FA}}}\) and \(B<q-\frac{1}{q}\). Since the number of committee members receiving positive favors is \(n-r^{*}=q\), the objective function in \({{\mathbb {CE}}}\) under \(r^{*}=n-q\) is

$$\begin{aligned} (B-b)\left( \frac{b}{q}\right) ^{q} \end{aligned}$$
(32)

and standard argument shows that \(b^{*}=B\frac{q}{q+1}\). However, since \(B<q-\frac{1}{q}\), we have \(b^{*}=B\frac{q}{q+1}<q-1\Leftrightarrow B<q-\frac{1}{q}\), so that, by Proposition 1 part 2, \(r^{*}=0\), a contradiction.

For part 3, when \(B<q-1\), \(b\in [0,B]\) in \({{\mathbb {CE}}}\) implies \(b^{*}<q-1\), which implies \(r^{*}=0\) in \({{\mathbb {FA}}}\).

To see that \(\tilde{b}^{*}\ge b^{*}\) if \(\tilde{B}>B\), where \(\tilde{b}^{*}\) solves \({{\mathbb {CE}}}\) given \(\tilde{B}>B\), suppose, towards a contradiction, that \(\tilde{b}^{*}<b^{*}\). From \(b^{*}\le q\), we have \(\tilde{b}^{*}<q\). Because \(b^{*}\) solves \({{\mathbb {CE}}}\) given B and \(\tilde{b}^{*}\) solves \({{\mathbb {CE}}}\) given \(\tilde{B}\), we have

$$\begin{aligned} (B-b^{*}){\mathbb {R}}_{q|n}[b^{*}]&\ge (B-\tilde{b}^{*}){\mathbb {R}}_{q|n}[\tilde{b}^{*}]\\ (\tilde{B}-\tilde{b}^{*}){\mathbb {R}}_{q|n}[\tilde{b}^{*}]&\ge (\tilde{B}-b^{*}){\mathbb {R}}_{q|n}[b^{*}].\\ \end{aligned}$$
(33)

Summing the two inequalities, we have \({\mathbb {R}}_{q|n}[b^{*}](B-\tilde{B})\ge {\mathbb {R}}_{q|n}[\tilde{b}^{*}](B-\tilde{B})\), which, by \(\tilde{B}>B\), implies \({\mathbb {R}}_{q|n}[b^{*}]\le {\mathbb {R}}_{q|n}[\tilde{b}^{*}]\). However, from the structure of the \({{\mathbb {FA}}}\) problem, we have \({\mathbb {R}}_{q|n}[b^{*}]>{\mathbb {R}}_{q|n}[\tilde{b}^{*}]\) when \(\tilde{b}^{*}<b^{*}\) and \(\tilde{b}^{*}<q\). \(\square \)

1.4 Proof of Proposition 3

As in the proof of Proposition 2, \(b^{*}\le q\). Moreover, since \({\mathbb {R}}_{q|n}[0]=0\), \(b^{*}\ne 0\). Thus it suffices to show that \(b^{*}\notin (0,q)\). Moreover, if we show that \(b^{*}=q\) for B, then \(b^{*}=q\) for any \(B'\ge B\). To see this, if \(b^{*}=q\) under B, we have

$$\begin{aligned} B-q\ge (B-b){\mathbb {R}}_{q|n}[b] \end{aligned}$$
(34)

\(\forall b\in [0,q]\). Therefore, \(\forall b\in [0,q]\) and any \(c\ge 0\),

$$\begin{aligned} B+c-q&\ge (B-b){\mathbb {R}}_{q|n}[b]+c\\&\ge (B+c-b){\mathbb {R}}_{q|n}[b]\\ \end{aligned}$$
(35)

where the second inequality follows from \({\mathbb {R}}_{q|n}[b]\in [0,1]\), and, hence, \(b^{*}=q\) under \(B+c\ge B\).

In order to show \(b^{*}\notin (0,q)\), it suffices to show that \((B-b)F(q,n-r,\frac{b}{n-r})\) is increasing for any \(b\in (0,q)\) and \(r\in \{0,\ldots ,n-q\}\). To see this, we know that \({\mathbb {R}}_{q|n}[b]\) is continuous in b on (0, q) and, by Proposition 1, \(\forall b\in (0,q)\), \(R_{q|n}[b]\in \{F(q,n-r,\frac{b}{n-r})|r\in \{0,\ldots ,n-q\})\).

Fix \(n\in {\mathbb {N}}_{\ge 1}\), \(q\in \{1,\ldots ,n\}\), \(B\ge q+(1-\frac{q}{n})\,_{2}F_{1}(1,1+n,1+q,\frac{q}{n})\), \(b\in (0,q)\) and \(r\in \{0,\ldots ,n-q\}\). Because \(q+(1-\frac{q}{n})\,_{2}F_{1}(1,1+n,1+q,\frac{q}{n})=q+\frac{F(q,n,\frac{q}{n})}{f(q,n,\frac{q}{n})}\ge q+1\), we have \(B\ge q+1\). We claim that \(\frac{\partial }{\partial b}(B-b)F(q,n-r,\frac{b}{n-r})>0\).

Suppose first that \(r=n-q\). Then \(F(q,n-r,\frac{b}{n-r})=F(q,q,\frac{b}{q})=\left( \frac{b}{q}\right) ^{q}\) and we have \(\frac{\partial }{\partial b}(B-b)F(q,n-r,\frac{b}{n-r})=\left( \frac{b}{q}\right) ^{q-1}[B-b-\frac{b}{q}]\), where \(B-b-\frac{b}{q}>B-q-1\ge 0\). Since we have proven our claim for \(r=n-q\), suppose \(r\in \{0,\ldots ,n-q-1\}\), so that we can also suppose \(q\in \{1,\ldots ,n-1\}\). This implies \(n-r\in \{q+1,\ldots ,n\}\) and, hence, \(\frac{b}{n-r}\in (0,\frac{q}{q+1})\subset (0,1)\).

Since \(\frac{\partial }{\partial b}(B-b)F(q,n-r,\frac{b}{n-r})=(B-b)\frac{q}{b}f(q,n-r,\frac{b}{n-r})-F(q,n-r,\frac{b}{n-r})\), we have

$$\begin{aligned} \frac{\frac{\partial }{\partial b}(B-b)F(q,n-r,\frac{b}{n-r})}{f(q,n-r,\frac{b}{n-r})}&=(B-b)\frac{q}{b}-\frac{F(q,n-r,\frac{b}{n-r})}{f(q,n-r,\frac{b}{n-r})}\\&\ge (B-b)\frac{q}{b}-\frac{F(q,n-r,\frac{q}{n-r})}{f(q,n-r,\frac{q}{n-r})}\\&\ge (B-b)\frac{q}{b}-\frac{F(q,n,\frac{q}{n})}{f(q,n,\frac{q}{n})}\\&>B-q-\frac{F(q,n,\frac{q}{n})}{f(q,n,\frac{q}{n})}\ge 0.\\ \end{aligned}$$
(36)

The first equality uses \(f(q,n-r,\frac{b}{n-r})>0\). The first and the second inequality follows from Lemma 2 and 3 respectively. The third inequality follows by \(b<q\) and the last inequality by \(B\ge q+(1-\frac{q}{n})\,_{2}F_{1}(1,1+n,1+q,\frac{q}{n})\), or, equivalently, \(B\ge q+\frac{F(q,n,\frac{q}{n})}{f(q,n,\frac{q}{n})}\). This proves the main claim of the proposition.

What remains is \(q+1+q(1-\frac{q}{n})\ge q+(1-\frac{q}{n})\,_{2}F_{1}(1,1+n,1+q,\frac{q}{n})\) for any \(n\in {\mathbb {N}}_{\ge 1}\) and \(q\in \{1,\ldots ,n\}\). For \(q=n\), because \(\lim _{q\rightarrow n}(1-\frac{q}{n})\,_{2}F_{1}(1,1+n,1+n,\frac{q}{n})=1\), the inequality writes \(q+1\ge q+1\). Suppose \(q\in \{1,\ldots ,n-1\}\). Then, we have

$$\begin{aligned} q+\left(1-\frac{q}{n}\right)_{2}F_{1}\left(1,1+n,1+q,\frac{q}{n}\right)&=q+\frac{F(q,n,\frac{q}{n})}{f(q,n,\frac{q}{n})} \\&= q+1+\frac{F(q+1,n,\frac{q}{n})}{f(q+1,n,\frac{q}{n})}\frac{f(q+1,n,\frac{q}{n})}{f(q,n,\frac{q}{n})}\\&=q+1+\frac{F(q+1,n,\frac{q}{n})}{f(q+1,n,\frac{q}{n})}\frac{q}{q+1} \leq q+1+\frac{(q+1)(1-\frac{q}{n})}{q+1-q}\frac{q}{q+1} \\&= q+1+q(1-\frac{q}{n}). \end{aligned}$$
(37)

The first equality follows by (14). The second equality uses \(F(q,n,\frac{q}{n})=\sum _{s=q}^{n}f(q,n,\frac{q}{n})=f(q,n,\frac{q}{n})+\sum _{s=q+1}^{n}f(s,n,\frac{q}{n})=f(q,n,\frac{q}{n})+F(q+1,n,\frac{q}{n})\). The third equality uses \(\frac{f(q+1,n,\frac{q}{n})}{f(q,n,\frac{q}{n})}=\frac{(n-q)\frac{q}{n}}{(q+1)(1-\frac{q}{n})}=\frac{q}{q+1}\), similarly to the proof of Lemma 1. The inequality follows from Diaconis and Zabell’s Theorem 2. The last equality is algebra. \(\square \)

1.5 Proof of Proposition 4

To prove the proposition, we use Hoeffding’s Theorem 1. Let \(b'=\frac{b}{n}\), \(q'=\frac{q}{n}\), \(B'=\frac{B}{n}\) and \(r'=\frac{r}{n}\in [0,1-q']\). Given n and r, \(n-r=n(1-r')\) of the committee members receive positive favors and vote yes with probability \(p=\frac{b}{n-r}=\frac{b'}{1-r'}\). Set \(t=\frac{q-b}{n-r}=\frac{q'-b'}{1-r'}\) in the statement of Theorem 1. Then \(t\in (0,1-p]\) as \(q'-b'>0\) and \(t\le 1-p\Leftrightarrow \frac{q'-b'}{1-r'}\le 1-\frac{b'}{1-r'}\Leftrightarrow q'\le 1-r'\), which holds since \(r'\le 1-q'\).

Using Theorem 1, we have

$$\begin{aligned} F(q,n-r,b/(n-r))&=\,F(n(1-r')(p+t),n(1-r'),b'/(1-r'))\\ &\le\,\exp {[-2n(1-r')t^{2}]}\\ &\le\,\exp {\left[ -2n\frac{(q'-b')^{2}}{1-r'}\right] }\\ \end{aligned}$$
(38)

where \(\frac{(q'-b')^{2}}{1-r'}>0\) for any \(r'\in [0,1-q']\).

Now, \(b'\) is optimal in \({{\mathbb {CE}}}\) if \((B-b)F(q,n-r,b/(n-r))\ge B-q\), or, equivalently,

$$\begin{aligned} F(q,n(1-r'),b'/(1-r'))\ge \frac{B'-q'}{B'-b'}. \end{aligned}$$
(39)

Fixing \(b'\) such that \(\epsilon =q'-b'>0\) and assuming \(B'>q'\), there exists \(\underline{n}_{\,\epsilon }\) such that for any \(n\ge \underline{n}_{\,\epsilon }\)

$$\begin{aligned} F(q,n(1-r'),b'/(1-r'))\le \exp {\left[ -2n\frac{(q'-b')^{2}}{1-r'}\right] }<\frac{B'-q'}{B'-b'} \end{aligned}$$
(40)

as \(\lim _{n\rightarrow \infty }{\exp {\left[ -2n\frac{(q'-b')^{2}}{1-r'}\right] }}=0\), which concludes the proof. \(\square \)

1.6 Proof of Proposition 5

Suppose first that g is linear. The objective function in \({{\mathbb {FA}}_{g}}\) rewrites as

$$\begin{aligned}\sum _{s=0}^{n}{\mathbb {P}}_{s|n}^{*}[{\mathbf {p}}]g(s)&=\sum _{s=1}^{n}{\mathbb {P}}_{s-1|n-1}^{*}[{\mathbf {p}}^{\{i\}}]\left[ g(s)p_{i}+g(s-1)(1-p_{i})\right] \\ &=\sum _{s=1}^{n}{\mathbb {P}}_{s-1|n-1}^{*}[{\mathbf {p}}^{\{i\}}]\left[ g(s-1)p_{i}+kp_{i}+g(s-1)(1-p_{i})\right] \\ &=kp_{i}+\sum _{s=1}^{n}{\mathbb {P}}_{s-1|n-1}^{*}[{\mathbf {p}}^{\{i\}}]g(s-1)\\ \end{aligned} $$
(41)

where the first and the third equality follow by algebra and the second equality uses linearity of g. Proceeding iteratively, \(\sum _{s=0}^{n}{\mathbb {P}}_{s|n}^{*}[{\mathbf {p}}]g(s)=k\sum _{i\in N}p_{i}+\sum _{s=n}^{n}{\mathbb {P}}_{s-n|n-n}^{*}[{\mathbf {p}}^{\{1\ldots n\}}]g(s-n)=k\sum _{i\in N}p_{i}+g(0)\) and, hence, any \({\mathbf {p}}\in X(b)\) such that \(\sum _{i\in N}p_{i}=b\) constitutes a solution to \({{\mathbb {FA}}_{g}}\).

Suppose now that g is either convex or concave. The objective function in \({{\mathbb {FA}}_{g}}\) rewrites as

$$\begin{aligned}\sum _{s=0}^{n}{\mathbb {P}}_{s|n}^{*}[{\mathbf {p}}]g(s)&=\sum _{s=0}^{n}g(s)\left[ \begin{array}{l}(p_{i}(1-p_{j})+(1-p_{i})p_{j}){\mathbb {P}}_{s-1|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]\\ +\,p_{i}p_{j}{\mathbb {P}}_{s-2|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]+(1-p_{i})(1-p_{j}){\mathbb {P}}_{s|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]\end{array}\right] \\ &=\sum _{s=2}^{n}{\mathbb {P}}_{s-2|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]\left[ \begin{array}{l}g(s-1)(p_{i}(1-p_{j})+(1-p_{i})p_{j})\\ +\, g(s)p_{i}p_{j}+g(s-2)(1-p_{i})(1-p_{j})\end{array}\right] \\ &=\sum _{s=2}^{n}{\mathbb {P}}_{s-2|n-2}^{*}[{\mathbf {p}}^{\{ij\}}]\left[ \begin{array}{l}p_{i}p_{j}\left[ g(s)-2g(s-1)+g(s-2)\right] \\ +\, (p_{i}+p_{j})\left[ g(s-1)-g(s-2)\right] +g(s-2)\end{array}\right] .\\ \end{aligned}$$
(42)

Suppose \({\mathbf {p}}^{*}\) constitutes a solution to \({{\mathbb {FA}}_{g}}\). Then \((p_{i}^{*},p_{j}^{*})\) has to constitute a solution to

$$\begin{aligned} \max _{(p_{i},p_{j})\in {\mathbb {R}}^{2}}&\quad \sum _{s=2}^{n}{\mathbb {P}}_{s-2|n-2}^{*}[{\mathbf {p}}^{\{ij\},*}]z(p_{i},p_{j},s)\\ {\text {s.t.}}&\quad (p_{i},p_{j})\in X(p_{i}^{*}+p_{j}^{*}). \end{aligned} \quad\quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \ \ \ ({{\mathbb {FA}}_{g}'})$$

where

$$\begin{aligned} z(p_{i},p_{j},s)=\left[ \begin{array}{l}p_{i}p_{j}\left[ g(s)-g(s-1)-(g(s-1)-g(s-2))\right] \\ +(p_{i}+p_{j})\left[ g(s-1)-g(s-2)\right] +g(s-2)\end{array}\right] . \end{aligned}$$
(43)

Clearly, solution to \({{\mathbb {FA}}}_{g}'\), \((p_{i}',p_{j}')\) satisfies \(p_{i}'+p_{j}'=p_{i}^{*}+p_{j}^{*}=k^{*}\). Substituting \(p_{i}+p_{j}=k^{*}\) into \(z(p_{i},p_{j},s)\) gives

$$\begin{aligned} \frac{\partial }{\partial p_{i}}z(p_{i},k^{*}-p_{i},s)&=(k^{*}-2p_{i})\Delta ^{2}g(s)\\ \frac{\partial ^{2}}{\partial ^{2}p_{i}}z(p_{i},k^{*}-p_{i},s)&=-2\Delta ^{2}g(s)\\ \end{aligned} $$
(44)

where \(\Delta ^{2}g(s)=g(s)-g(s-1)-(g(s-1)-g(s-2))\).

To prove part 1, suppose \(p_{i}^{*}\ne p_{j}^{*}\), so that \(k^{*}>0\). Since g is convex, \(\Delta ^{2}g(s)>0\) and, hence, \(\frac{\partial ^{2}}{\partial ^{2}p_{i}}z(p_{i},k^{*}-p_{i},s)<0\), so that the objective function in \({{\mathbb {FA}}}_{g}'\) is strictly concave in \(p_{i}\). Therefore, \((p_{i}',p_{j}')=(\frac{k^{*}}{2},\frac{k^{*}}{2})\), a contradiction to \(p_{i}^{*}\ne p_{j}^{*}\).

To prove part 2, suppose \(p_{i}^{*}\in (0,1)\) and \(p_{j}^{*}\in (0,1)\) so that \(k^{*}>0\). Since g is concave, \(\Delta ^{2}g(s)<0\) and, hence, \(\frac{\partial ^{2}}{\partial ^{2}p_{i}}z(p_{i},k^{*}-p_{i},s)>0\), so that the objective function in \({{\mathbb {FA}}}_{g}'\) is strictly convex in \(p_{i}\). Therefore, either \((p_{i}',p_{j}')\in \{(k^{*},0),(0,k^{*})\}\) if \(k^{*}\in (0,1]\) or \((p_{i}',p_{j}')\in \{(1,k^{*}-1),(k^{*}-1,1)\}\) if \(k^{*}\in (1,2]\). In both cases, \((p_{i}^{*},p_{j}^{*})\) cannot constitute a solution to \({{\mathbb {FA}}}_{g}'\), a contradiction.\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zápal, J. Crafting consensus. Public Choice 173, 169–200 (2017). https://doi.org/10.1007/s11127-017-0470-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11127-017-0470-8

Keywords

JEL Classification

Navigation