Skip to main content
Log in

Optimizing political influence: a jury theorem with dynamic competence and dependence

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

Abstract

The purpose of this paper is to illustrate, formally, an ambiguity in the exercise of political influence. To wit: A voter might exert influence with an eye toward maximizing the probability that the political system (1) obtains the correct (e.g. just) outcome, or (2) obtains the outcome that he judges to be correct (just). And these are two very different things. A variant of Condorcet’s Jury Theorem which incorporates the effect of influence on group competence and interdependence is developed. Analytic and numerical results are obtained, the most important of which is that it is never optimal—from the point-of-view of collective accuracy—for a voter to exert influence without limit. He ought to either refrain from influencing other voters or else exert a finite amount of influence, depending on circumstance. Philosophical lessons are drawn from the model, to include a solution to Wollheim’s “paradox in the theory of democracy”.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. Some may hold a stronger view, under which (i) people have a right to participate in politics, regardless of their epistemic state; or (ii) unrestrained influence is an essential element of the deliberation which justifies a political process.

  2. Technically, dependence is bad from the point-of-view of aggregation quality only if voters are competent; that is, more likely than a coin flip to vote correctly (see §2). No assumptions about voter competence are imposed in this paper’s model.

  3. For an introduction to the epistemology of disagreement, see Christensen (2009) and Feldman (2007).

  4. On the steadfast position, see, e.g., Bergmann (2009), Kelly (2005), and Van Inwagen (2010). On conciliationism, see Christensen (2007), Elga (2007), and Feldman (2007). I offer my (generally conciliatory) thoughts in Mulligan (2021).

  5. Example: There is a three-member civil jury. Two jurors believe, with credence 0.4, that the facts and the law favor the plaintiff. The third juror believes this with credence 0.9. So they “conciliate”, and all adopt the average credence of 0.57. Before conciliating, the respondent would prevail, two votes to one; after conciliating, the plaintiff wins unanimously. It is unclear that this is an epistemic improvement, to say nothing of justice. (N.B. in American civil cases the standard of proof is a “preponderance of the evidence” and so the two possible outcomes are symmetric. For criminal cases, with the much higher “beyond a reasonable doubt” standard, this is not the case.)

  6. See also Austin-Smith (1990).

  7. This is not to say that the other assumptions may never be violated in practice. But they are often complied with. Many real-world decisions are naturally dichotomous, or they are not naturally dichotomous but are broken down into dichotomous steps for practical reasons. And simple majority rule is ubiquitous [I offer an alternative in Mulligan (2018b)]. In any case, work has been done extending Condorcet’s theorem to deal with these limitations; on the relaxation of dichotomous choice, e.g., see Hummel (2010), Lam and Suen (1996), List and Goodin (2001), Miller (1996), and Paroush (1990).

  8. In particular, the optimal decision rule may be identified by weighting each vote in proportion to the log-odds of that voter’s competence. Other relevant literature includes Grofman and Feld (1983), Nitzan and Paroush (1984, 1985), and Paroush (1998).

  9. One might think that this problem can be bypassed by shifting from a brute independence assumption to independence conditional upon common information (like evidence). Unfortunately, such a shift imperils other necessary assumptions (like the minimal competence assumption)–see Dietrich and Spiekermann (2013a, 2013b, 2020) and Ladha (1993).

  10. For other models and discussions of dependent voting, see Berend and Sapir (2005, 2007), Berg (19931996), Dietrich and List (2004), Kaniovski (2010), Kaniovski and Zaigraev (2011), Ladha (1992, 19931995), List and Pettit (2004), Nitzan and Paroush (1985), Peleg and Zamir (2012), Shapley and Grofman (1984), and Zaigraev and Kaniovski (2013).

  11. A natural question is whether the core results of this paper can be obtained with general functions c and \(\rho\), without choosing specific functional forms. General functions would be constrained as follows: Both at least C1 and strictly increasing in i; \(c(i)=c_{init}\) for \(i=0\), \(c(i)=1\) in the limit as \(i \rightarrow \infty\); \(\rho (i)=0\) for \(i=0\), \(\rho (i)=1\) in the limit as \(i \rightarrow \infty\). The answer is no. Some functions approximating certain fixed curves, which satisfy these constraints, may yield an infinite number of optima rather than the unique optimum obtained here. I thank Willie Wong for identifying this counterexample.

  12. As one example of the difficulties dependence may introduce, consider a simple case of three voters whose competence and pairwise correlation are known precisely. How likely is it that this electorate will choose correctly? It is impossible to know. These data fail to specify a unique joint distribution (although they do constrain potential distributions). The most common representation of the joint distribution is that given by Bahadur (1961) and Lazarsfeld (1956), which includes \(2^n-n-1\) correlation parameters. The challenge of applying this representation is that some of these parameters (often the higher-order correlations) are unavailable, and generally cannot be ignored (i.e. set to 0). To deal with these limitations, Van Der Geest (2005) uses a maximum entropy method, which can be implemented numerically, to infer higher-order correlations. Kaniovski (2008a) obtains the relevant probabilities by identifying the distribution that is “closest” (in a least squares sense) to that in which voters are assumed to be independent. This approach also may be implemented numerically, and Kaniovski obtains an analytic solution for the special case of homogeneous competence. (Numerical examples of this approach are provided in Kaniovski 2008b.)

  13. Note that we continue to assume aggregation via simple majority rule here. When an expert exists, this may no longer be optimal.

  14. See https://www.opensecrets.org/overview/topindivs.php?cycle=2016&view=fc, retrieved 6 August 2020.

  15. Technically, the game only has a prisoner’s dilemma structure if Adelson and Soros are interpreted as opinion leaders who seek justice as they see it rather than as the superior, process-optimizing type. The reason is that, so interpreted, each would prefer most of all to exert influence and have his opponent exert no influence, followed by refraining from influence in the face of his opponent’s restraint, followed by exerting influence in the face of his opponent’s exertion of influence. In any case, the dynamic described for the two relevant cases—each exerts influence and each refrains from exerting influence—holds under both interpretations.

  16. E.g. Brennan (2016), Estlund (2008), and Landemore (2013).

  17. Including Foot (2002), Schueler (2007), and Wilcox (1968).

  18. For critical commentary on Wollheim’s paradox, see Honderich (1974), Paris and Reynolds (1978), and Weiss (1973).

  19. See, e.g., Althaus (2003), Brennan (2016), Converse (1964), Delli Carpini and Keeter (1996), Neuman (1986), Shenkman (2008), and Somin (2013).

References

  • Althaus SL (2003) Collective preferences in democratic politics: opinion surveys and the will of the people. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Austin-Smith D (1990) Information transmission in debate. Am J Polit Sci 34:124–52

    Article  Google Scholar 

  • Bahadur RR (1961) A representation of the joint distribution of responses to n dichotomous items. In: Solomon H (ed) Studies in item analysis and prediction. Stanford University Press, Stanford, pp 158–68

    Google Scholar 

  • Berend D, Sapir L (2005) Monotonicity in Condorcet Jury Theorem. Soc Choice Welf 24:83–92

    Article  Google Scholar 

  • Berend D, Sapir L (2007) Monotonicity in Condorcet’s Jury Theorem with dependent voters. Soc Choice Welf 28:507–28

    Article  Google Scholar 

  • Berg S (1993) Condorcet’s Jury Theorem, dependency among voters. Soc Choice Welf 10:87–95

    Article  Google Scholar 

  • Berg S (1996) Condorcet’s Jury Theorem and the reliability of majority voting. Group Decis Negot 5:229–38

    Article  Google Scholar 

  • Bergmann M (2009) Rational disagreement after full disclosure. Episteme 6:336–53

    Article  Google Scholar 

  • Boland PJ (1989) Majority systems and the Condorcet Jury Theorem. The Statistician 38:181–89

    Article  Google Scholar 

  • Boland PJ, Proschan F, Tong YL (1989) Modelling dependence in simple and indirect majority systems. J Appl Probab 26:81–89

    Article  Google Scholar 

  • Brennan J (2016) Against Democracy. Princeton University Press, Princeton

    Book  Google Scholar 

  • Brennan G, Lomasky L (1993) Democracy & decision: the pure theory of electoral preference. Cambridge University Press, New York

    Book  Google Scholar 

  • Christensen D (2007) Epistemology of disagreement: the good news. Philos Rev 119:187–217

    Article  Google Scholar 

  • Christensen D (2009) Disagreement as evidence: the epistemology of controversy. Philos Compass 4:756–67

    Article  Google Scholar 

  • Condorcet (1785) Essai sur l’application de l’analyse à la probabilité des décisions rendues à la pluralité des voix. Imprimerie Royale, Paris

    Google Scholar 

  • Converse PE (1964) The nature of belief systems in mass publics. In: Apter DE (ed) Ideology and discontent. Free Press, New York, pp 164–93

    Google Scholar 

  • Delli Carpini MX, Keeter S (1996) What Americans know about politics and why it matters. Yale University Press, New Haven

    Google Scholar 

  • Dietrich F (2008) The premises of Condorcet’s Jury Theorem are not simultaneously justified. Episteme 5:56–73

    Article  Google Scholar 

  • Dietrich F, List C (2004) A model of jury decision where all jurors have the same evidence. Synthese 142:175–202

    Article  Google Scholar 

  • Dietrich F, Spiekermann K (2013a) Epistemic democracy with defensible premises. Econ Philos 89:87–120

    Article  Google Scholar 

  • Dietrich F, Spiekermann K (2013b) Independent opinions? On the causal foundations of belief formation and jury theorems. Mind 122:655–85

    Article  Google Scholar 

  • Dietrich F, Spiekermann K (2020) Jury theorems. In: Fricker M, Graham PJ, Henderson D, Pedersen NJLL (eds) The Routledge handbook of social epistemology. Routledge, New York, pp 386–96

    Google Scholar 

  • Elga A (2007) Reflection and disagreement. Noûs 41:478–502

    Article  Google Scholar 

  • Estlund D (2008) Democratic authority: a philosophical framework. Princeton University Press, Princeton

    Google Scholar 

  • Ewin RE (1967) Wollheim’s paradox of democracy. Australas J Philos 45:356–57

    Article  Google Scholar 

  • Feddersen T, Pesendorfer W (1996) The swing voter’s curse. Am Econ Rev 86:408–24

    Google Scholar 

  • Feddersen T, Pesendorfer W (1997) Voting behavior and information aggregation in elections with private information. Econometrica 65:1029–58

    Article  Google Scholar 

  • Feddersen T, Pesendorfer W (1999) Abstention in elections with asymmetric information and diverse preferences. Am Polit Sci Rev 93:381–98

    Article  Google Scholar 

  • Feldman R (2007) Reasonable religious disagreements. In: Antony LM (ed) Philosophers without gods: meditations on atheism and the secular life. Oxford University Press, New York, pp 194–214

    Google Scholar 

  • Foot P (2002) Moral relativism. In: Foot P (ed) Moral dilemmas: and other topics in moral philosophy. Oxford University Press, New York, pp 20–36

    Chapter  Google Scholar 

  • Goldman A (1999) Knowledge in a social world. Oxford University Press, Oxford

    Book  Google Scholar 

  • Grofman B, Feld SL (1983) Determining optimal weights for expert judgment. In: Grofman B, Owen G (eds) Information pooling and group decision making: proceedings of the Second University of California, Irvine, Conference on Political Economy, 2:167–72. Greenwich, CT: JAI Press

  • Grofman B, Owen G, Feld SL (1983) Thirteen theorems in search of the truth. Theory Decis 15:261–78

    Article  Google Scholar 

  • Honderich T (1974) A difficulty with democracy. Philos Public Affairs 3:221–26

    Google Scholar 

  • Hummel P (2010) Jury theorems with multiple alternatives. Soc Choice Welf 34:65–103

    Article  Google Scholar 

  • Kaniovski S (2008a) The exact bias of the Banzhaf measure of power when votes are neither equiprobable nor independent. Soc Choice Welf 31:281–300

    Article  Google Scholar 

  • Kaniovski S (2008b) Straffin meets Condorcet: what can a voting power theorist learn from a jury theorist? Homo Oeconomicus 25:181–202

    Google Scholar 

  • Kaniovski S (2010) Aggregation of correlated votes and Condorcet’s Jury Theorem. Theor Decis 69:43–68

    Article  Google Scholar 

  • Kaniovski S, Zaigraev A (2011) Optimal jury design for homogenous juries with correlated votes. Theor Decis 71:439–59

    Article  Google Scholar 

  • Kelly T (2005) The epistemic significance of disagreement. In: Gendler TS, Hawthorne J (eds) Oxford studies in epistemology, vol 1. Oxford University Press, Oxford, pp 167–96

    Google Scholar 

  • Ladha KK (1992) The Condorcet Jury Theorem, free speech, and correlated votes. Am J Polit Sci 36:617–34

    Article  Google Scholar 

  • Ladha KK (1993) Condorcet’s Jury Theorem in light of de Finetti’s theorem. Soc Choice Welf 10:69–85

    Article  Google Scholar 

  • Ladha KK (1995) Information pooling through majority-rule voting: Condorcet’s Jury Theorem with correlated votes. J Econ Behav Organ 26:353–72

    Article  Google Scholar 

  • Laguerre M (1883) Mémoire sur la théorie des équations numériques. J Math Pures Appl 9:99–146

    Google Scholar 

  • Lam L, Suen CY (1996) Majority vote of even and odd experts in a polychotomous choice situation. Theor Decis 41:13–36

    Article  Google Scholar 

  • Landemore H (2013) Democratic reason: politics, collective intelligence, and the rule of the many. Princeton University Press, Princeton

    Google Scholar 

  • Lazarsfeld PF (1956) Some observations on dichotomous systems. Columbia University Sociology Department, New York

    Google Scholar 

  • List C, Goodin RE (2001) Epistemic democracy: generalizing the Condorcet Jury Theorem. J Polit Philos 9:277–306

    Article  Google Scholar 

  • List C, Pettit P (2004) An epistemic free-riding problem? In: Catton P, Macdonald G (eds) Karl Popper: critical appraisals. Routledge, Abingdon, pp 128–58

    Google Scholar 

  • List C, Spiekermann K (2016) The Condorcet Jury Theorem and voter-specific truth. In: McLaughlin BP, Kornblith H (eds) Goldman and his critics. Wiley, Malden, MA, pp 219–31

    Chapter  Google Scholar 

  • Miller NR (1996) Information, individual errors, and collective performance: empirical evidence on the Condorcet Jury Theorem. Group Decis Negot 5:211–28

    Article  Google Scholar 

  • Mulligan T (2018a) Justice and the meritocratic state. Routledge, New York

    Google Scholar 

  • Mulligan T (2018b) Plural voting for the twenty-first century. Philos Q 68:286–306

    Article  Google Scholar 

  • Mulligan T (2021) The epistemology of disagreement: why not Bayesianism? Episteme 18:587–602

    Article  Google Scholar 

  • Nelson M (2019) Propositional attitude reports. In: Zalta EN (ed) Stanford Encyclopedia of Philosophy (Spring 2019 Edition), https://plato.stanford.edu/archives/spr2019/entries/prop-attitude-reports/. Accessed 27 Apr 2022

  • Neuman WR (1986) The paradox of mass politics: knowledge and opinion in the american electorate. Harvard University Press, Cambridge

    Google Scholar 

  • Nitzan S, Paroush J (1982) Optimal decision rules in uncertain dichotomous choice situations. Int Econ Rev 23:289–97

    Article  Google Scholar 

  • Nitzan S, Paroush J (1984) The significance of independent decisions under uncertain dichotomous choice situations. Theor Decis 17:47–60

    Article  Google Scholar 

  • Nitzan S, Paroush J (1985) Collective decision making: an economic outlook. Cambridge University Press, Cambridge

    Google Scholar 

  • Paris DC, Reynolds JF (1978) Paradox, rationality, and politics: Wollheim’s democracy. J Polit 40:956–83

    Article  Google Scholar 

  • Paroush J (1990) Multi-choice problems and the essential order among decision rules. Econ Lett 32:121–25

    Article  Google Scholar 

  • Paroush J (1998) Stay away from fair coins: a Condorcet Jury Theorem. Soc Choice Welf 15:15–20

    Article  Google Scholar 

  • Peleg B, Zamir S (2012) Extending the Condorcet Jury Theorem to a general dependent jury. Soc Choice Welf 39:91–125

    Article  Google Scholar 

  • Rawls J (1999) A theory of justice, Revised. Harvard University Press, Cambridge

    Book  Google Scholar 

  • Reichenbach H (1956) The direction of time. University of California Press, Berkeley

    Book  Google Scholar 

  • Schueler GF (2007) Is it possible to follow one’s conscience? Am Philos Q 44:51–60

    Google Scholar 

  • Shapley L, Grofman B (1984) Optimizing group judgmental accuracy in the presence of interdependencies. Public Choice 43:329–43

    Article  Google Scholar 

  • Shenkman R (2008) Just how stupid are we? Facing the truth about the American voter. Basic Books, New York

    Google Scholar 

  • Simmons AJ (1979) Moral principles and political obligations. Princeton University Press, Princeton

    Google Scholar 

  • Somin I (2013) Democracy and political ignorance: why smaller government is better. Stanford University Press, Stanford

    Google Scholar 

  • Van Der Geest PAG (2005) The binomial distribution with dependent Bernoulli trials. J Stat Comput Simul 75:141–54

    Article  Google Scholar 

  • Van Inwagen P (2010) We’re right. They’re wrong. In: Feldman R, Warfield TA (eds) Disagreement. Oxford University Press, New York, pp 10–28

    Chapter  Google Scholar 

  • Weiss DD (1973) Wollheim’s paradox: survey and solution. Polit Theory 1:154–70

    Article  Google Scholar 

  • Wilcox JT (1968) Is it always right to do what you think is right? J Value Inquiry 2:95–107

    Article  Google Scholar 

  • Wollheim R (1962) A paradox in the theory of democracy. In: Laslett P, Runciman WG (eds) Philosophy, politics and society. Barnes and Noble, New York, pp 71–87

    Google Scholar 

  • Zaigraev A, Kaniovski S (2013) A note on the probability of at least k successes in n correlated binary trials. Oper Res Lett 41:116–20

    Article  Google Scholar 

Download references

Funding

None.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Mulligan.

Ethics declarations

Conflict of interest / competing interests

None.

Additional information

Publisher's Note

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

I wish to acknowledge, with thanks, generous feedback from David Faraci, Nick Geiser, John Hasnas, Dmitrii Karp, Raphael Lehrer, Iosif Pinelis, Kirun Sankaran, Willie Wong, and Maryam Yashtini. Audiences at the 2021 International Conference on Social Choice and Voting Theory, Virginia Tech, the Georgetown Institute for the Study of Markets and Ethics Workshop, and the 2018 Philosophy, Politics, and Economics Society annual meeting (which heard an early draft of this paper) contributed as well. Two anonymous reviewers for Social Choice and Welfare provided a number of helpful comments.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mulligan, T. Optimizing political influence: a jury theorem with dynamic competence and dependence. Soc Choice Welf (2022). https://doi.org/10.1007/s00355-022-01407-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00355-022-01407-5

Navigation