Skip to main content
Log in

Distance rationalization of voting rules

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

Abstract

The concept of distance rationalizability allows one to define new voting rules or rationalize existing ones via a consensus, i.e., a class of elections that have a unique, indisputable winner, and a distance over elections: A candidate is declared an election winner if she is the consensus candidate in one of the nearest consensus elections. Many classic voting rules are defined or can be represented in this way. In this paper, we focus on the power and the limitations of the distance rationalizability approach. Lerer and Nitzan (J Econ Theory 37(1):191–201, 1985) and Campbell and Nitzan (Soc Choice Welf 3(1):1–16, 1986) show that if we do not place any restrictions on the notions of distance and consensus then essentially all voting rules can be distance-rationalized. We identify a natural class of distances on elections—votewise distances—which depend on the submitted votes in a simple and transparent manner, and investigate which voting rules can be rationalized via distances of this type. We also study axiomatic properties of rules that can be defined via votewise distances.

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. We refer to this result, proved by Lerer and Nitzan (1985) and generalized by Campbell and Nitzan (1986), as the ‘universal distance rationalizability theorem’. Unaware of these papers, we have rediscovered the universal distance rationalizability theorem in a conference paper (Elkind et al. 2010). We would like to thank the reviewer who brought the papers of Lerer and Nitzan and of Campbell and Nitzan to our attention.

  2. We have rediscovered a slightly stronger version of this result (Elkind et al. 2010, 2011).

  3. We follow the terminology of Lerer and Nitzan (1985) here.

  4. We remark that Lerer and Nitzan (1985) refer to distances that are defined in this way as symmetric additively decomposable metrics; generalizing this class of distances is at the heart of this paper.

  5. One can also consider situations in which the voters reach a consensus that several candidates are equally well qualified to be elected, but we do not discuss this possibility.

  6. One might think that the term “\({\mathcal {K}}\)-consistent” would be more appropriate than “\({\mathcal {K}}\)-compatible.” Indeed, a voting rule that elects the Condorcet winner whenever one exists is usually referred to as Condorcet-consistent. Nonetheless, we decided to use the term “\({\mathcal {K}}\)-compatible” to avoid confusion with the normative axiom of consistency.

  7. Technically, a norm is defined for a fixed value of \(n\), whereas voting rules are usually defined for any number of voters, i.e., we require a family of norms, one for each value of \(n\). Note that each \(p\)-norm, \(p\in \mathbb {Z}^+\cup \{\infty \}\), is indeed defined for all values of \(n\).

  8. Known in statistics as Spearman’s footrule (Kendall and Gibbons 1990).

  9. The proof is available on request.

  10. This notion is due to Lerer and Nitzan (1985), who did not use the term “rank-monotonicity.” We introduce this term to distinguish this notion from relative monotonicity (Definition 12).

References

  • Aleskerov F, Chistyakov V, Kalyagin V (2010) The threshold aggregation. Econ Lett 107(2):261–262

    Article  Google Scholar 

  • Arrow K (1951; revised edition, 1963) Social choice and individual values, Wiley, New York

  • Baigent N (1987a) Metric rationalisation of social choice functions according to principles of social choice. Math Soc Sci 13(1):59–65

    Article  Google Scholar 

  • Baigent N (1987b) Preference proximity and anonymous social choice. Q J Econ 102(1):161–169

    Article  Google Scholar 

  • Baigent N, Klamler C (2004) Transitive closure, proximity and intransitivities. Econ Theory 23(1):175–181

    Article  Google Scholar 

  • Bartholdi J III, Tovey C, Trick M (1989) Voting schemes for which it can be difficult to tell who won the election. Soc Choice Welf 6(2):157–165

    Article  Google Scholar 

  • Bauer F, Stoer J, Witzgall C (1961) Absolute and monotonic norms. Numerische Matematic 3:257–264

    Article  Google Scholar 

  • Bogard K (1973) Preference structures I: distances between transitive preference relations. J Math Sociol 3:49–67

    Article  Google Scholar 

  • Bogard K (1975) Preference structures II: distances between transitive preference relations. SIAM J Appl Math 29:254–262

    Article  Google Scholar 

  • Boutilier C, Procaccia A (2012, July) A dynamic rationalization of distance rationalizability. In: Proceedings of the 26th AAAI conference on artificial intelligence. AAAI Press, pp 1278–1284

  • Brams S, Fishburn P (2002) Voting procedures. In: Arrow K, Sen A, Suzumura K (eds) Handbook of social choice and welfare, vol 1. Elsevier, Amsterdam, pp 173–236

    Chapter  Google Scholar 

  • Brandt F (2009) Some remarks on Dodgson’s voting rule. Math Logic Q 55(4):460–463

    Article  Google Scholar 

  • Campbell D, Nitzan S (1986) Social compromise and social metrics. Soc Choice Welf 3(1):1–16

    Article  Google Scholar 

  • Caragiannis I, Procaccia A, Shah N (2013) When do noisy votes reveal the truth?. In: Proceedings of the 13th ACM Conference on electronic commerce, pp 143–160

  • Chebotarev PY, Shamis E (1998) Characterizations of scoring methods for preference aggregation. Ann Oper Res 80:299–332

    Article  Google Scholar 

  • Condorcet J (1785) Essai sur l’application de l’analyse à la probabilité des décisions rendues à la pluralité des voix. (Facsimile reprint of original published in Paris, 1972, by the Imprimerie Royale)

  • Conitzer V, Rognlie M, Xia L (2009, July) Preference functions that score rankings and maximum likelihood estimation. In: Proceedings of the 21st international joint conference on artificial intelligence. AAAI Press, pp 109–115

  • Conitzer V, Sandholm T (2005, July) Common voting rules as maximum likelihood estimators. In: Proceedings of the 21st conference on uncertainty in artificial intelligence. AUAI Press, pp 145–152

  • Cook W, Seiford L (1978) Priority ranking and consensus information. Manag Sci 24:1721–1732

    Article  Google Scholar 

  • Cook W, Seiford L (1982) On the Borda-Kendall consensus method for priority ranking problems. Manag Sci 28:621–637

    Article  Google Scholar 

  • Deza MM, Deza E (2009) Encyclopedia of distances. Springer, Berlin

    Book  Google Scholar 

  • Eckert D, Klamler C (2011) Distance-based aggregation theory. In: Herrera-Viedma E, Garca-Lapresta JL, Kacprzyk J, Fedrizzi M, Nurmi H, Zadrozny S (eds) Consensual processes. Springer, Berlin, pp 3–22

    Chapter  Google Scholar 

  • Elkind E, Faliszewski P, Slinko A (2010, May) On the role of distances in defining voting rules. In: Proceedings of the 9th international conference on autonomous agents and multiagent systems, pp 375–382

  • Elkind E, Faliszewski P, Slinko A (2011) Homogeneity and monotonicity of distance-rationalizable voting rules. In: Proceedings of the 10th international conference on autonomous agents and multiagent systems, pp 821–828

  • Elkind E, Faliszewski P, Slinko A (2012) Rationalizations of Condorcet consistent rules via distances of hamming type. Soc Choice Welf 4(39):891–905

    Article  Google Scholar 

  • Elkind E, Shah N (2014, July) Electing the most probable without eliminating the irrational: Voting over intransitive domains. In: Proceedings of the 30th conference on uncertainty in artificial intelligence

  • Elkind E, Slinko A (2015) Rationalizations of voting rules. In: Brandt F, Conitzer V, Endriss U, Lang J, Procaccia AD (eds) Handbook of computational social choice, Chapt 8. Cambridge University Press, Cambridge

    Google Scholar 

  • Farkas D, Nitzan S (1979) The Borda rule and Pareto stability : a comment. Econometrica 47:1305–1306

    Article  Google Scholar 

  • Goldsmith J, Lang J, Mattei N, Perny P (2014) Voting with rank dependent scoring rules. In: Proceedings of the 28th AAAI conference on artificial intelligence. AAAI Press, pp 698–704

  • Hemaspaandra E, Hemaspaandra L, Rothe J (1997) Exact analysis of Dodgson elections: Lewis Carroll’s 1876 voting system is complete for parallel access to NP. J ACM 44(6):806–825

    Article  Google Scholar 

  • Hudry O, Monjardet B (2010) Consensus theories. An oriented survey. Math Soc Sci 190:139–167

    Google Scholar 

  • Kendall M, Gibbons J (1990) Rank correlation methods. Oxford University Press, Oxford

    Google Scholar 

  • Lerer E, Nitzan S (1985) Some general results on the metric rationalization for social decision rules. J Econ Theory 37(1):191–201

    Article  Google Scholar 

  • Litvak B (1982) Information given by the experts. Methods of acquisition and analysis. Radio and Communication, Moscow

    Google Scholar 

  • Litvak B (1983) Distances and consensus rankings. Cybernetics and systems analysis 19(1):71–81. (Translated from Kibernetika, No. 1, pp 57–63, January-February 1983)

  • Meskanen T, Nurmi H (2008) Closeness counts in social choice. In: Braham M, Steffen F (eds) Power, freedom, and voting. Springer, Berlin

    Google Scholar 

  • Miller M, Osherson D (2009) Methods for distance-based judgment aggregation. Soc Choice Welf 4(32):575–601

    Article  Google Scholar 

  • Moulin H (1991) Axioms of cooperative decision making. Cambridge University Press, Cambridge

    Google Scholar 

  • Nitzan S (1981) Some measures of closeness to unanimity and their implications. Theory Decis 13(2):129–138

    Article  Google Scholar 

  • Nitzan S (1989) More on preservation of preference proximity and anonymous social choice. Q J Econ 104(1):187–190

    Article  Google Scholar 

  • Pfingsten A, Wagener A (2003) Bargaining solutions as social compromises. Theory Decis 55(4):359–389

    Article  Google Scholar 

  • Pivato M (2013) Voting rules as statistical estimators. Soc Choice Welf 40(2):581–630

    Article  Google Scholar 

  • Schechter E (1997) Handbook of analysis and its foundations. Academic Press, New York

    Google Scholar 

  • Xia L, Conitzer V (2011, July) A maximum likelihood approach towards aggregating partial orders. In: Proceedings of the 22nd international joint conference on artificial intelligence, pp 446–451

  • Xia L, Conitzer V, Lang J (2010) Aggregating preferences in multiissue domains by using maximum likelihood estimators. In: Proceedings of the 9th international conference on autonomous agents and multiagent systems, pp 399–406

  • Young H (1975) Social choice scoring functions. SIAM J Appl Math 28(4):824–838

    Article  Google Scholar 

  • Young H (1977) Extending Condorcet’s rule. J Econ Theory 16(2):335–353

    Article  Google Scholar 

  • Young H, Levenglick A (1978) A consistent extension of Condorcet’s election principle. SIAM J Appl Math 35(2):285–300

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Edith Elkind.

Appendices

Appendix A: \(\varvec{{\mathcal {M}}}\)-Borda is not homogeneous

We show that Theorem 9 cannot be extended to \(\ell _1\)-votewise rules by showing that \({\mathcal {M}}\)-Borda is not homogeneous. Consider election \(E\):

\(v_1\)

\(v_2\)

\(v_3\)

\(v_4\)

\(v_5\)

\(v_6\)

\(b\)

\(a\)

\(c\)

\(c\)

\(c\)

\(d\)

\(a\)

\(b\)

\(b\)

\(b\)

\(d\)

\(a\)

\(d\)

\(d\)

\(a\)

\(a\)

\(a\)

\(b\)

\(e\)

\(e\)

\(e\)

\(e\)

\(e\)

\(e\)

\(c\)

\(c\)

\(d\)

\(d\)

\(b\)

\(c\)

A simple calculation shows that, to become a majority winner, \(a\) needs \(4\) swaps, \(b\) needs \(3\) swaps, \(c\) needs \(4\) swaps, and \(d\) needs \(5\) swaps. Thus, \(b\) is the unique winner of this election under \({\mathcal {M}}\)-Borda. However, in the election \(2E\) candidate \(b\) needs \(5\) swaps to become a majority winner, while \(c\) requires only \(4\) swaps.

Appendix B: Proof of theorem 3

We will show that Plurality is \(\ell _1\)-votewise rationalizable with respect to strong unanimity. Given a set of candidates \(C=\{c_1, \dots , c_m\}\) and two preference orders \(u, v\in P(C)\), we say that \(v\) can be obtained from \(u\) by a cyclic shift if there exists an \(i\in [m]\) and a permutation \(\pi :C\rightarrow C\) such that \(v=\pi (c_1)\succ \cdots \succ \pi (c_m)\), and \(u=\pi (c_i)\succ \cdots \succ \pi (c_m)\succ \pi (c_1)\succ \cdots \succ \pi (c_{i-1})\).

We partition \(P(C)\) into \(m\) groups \(L_1, \dots , L_m\) by setting \(L_i=\{v\in P(C)\mid {\mathrm {pos}}(v, c_i)=1\}\). Set \(s=(m-1)!\) and, for each \(i\in [m]\), enumerate the preference orders in \(L_i\) as \(v_i^1, \dots , v_i^s\) so that for every \(i, j\in [m]\) and every \(t=1, \dots , s\) the vote \(v_j^t\) can be obtained from the vote \(v_i^t\) by a cyclic shift. To see that this is possible, note that for each \(v_i^t\), \(i\in [m]\), \(t\in [s]\), and each \(j\in [m]\), there is exactly one vote in \(L_j\) that can be obtained from \(v_i^t\) by a cyclic shift.

The distance \(d\) is defined as follows. We set \(d(v_i^t, v_j^r)=0\), if \(i=j\) and \(t=r\); \(d(v_i^t, v_j^r)=1\), if \(i=j\) or \(t=r\) but \((i, t)\ne (j,r)\), and \(d(v_i^t, v_j^r)=2\), if \(i\ne j\text { and } t\ne r\). In other words, \(d(u,u)=0\) for all \(u\in P(C)\), \(d(u,v)=1\) if and only if \(u\ne v\) and \(u\) and \(v\) rank the same candidate first or \(v\) is obtained from \(u\) by a cyclic shift, and \(d(u,v)=2\) in all other cases. Observe that since \(d(u, v)\in \{1, 2\}\) for \(u\ne v\), the mapping \(d\) satisfies the triangle inequality; it is also obviously symmetric.

Consider an election \(E=(C, V)\). For every \(j\in [m]\) and \(r\in [s]\) let \(a_j^r\) be the number of voters in \(V\) with preference order \(v_j^r\). Let us calculate the distance from \(E\) to the \({\mathcal {S}}\)-consensus election \((C, X_i^t)\) with \(X_i^t = |V|\times v_i^t\):

$$\begin{aligned} \widehat{d}(V,X_i^t)= \sum _{r\in [s]\setminus \{t\}} \!\!\!\!\!a_i^r \ \ + \sum _{j\in [m]\setminus \{i\}} \!\!\!\!\!\!a_j^t \ +\ 2\sum _{j\ne i}\sum _{r\ne t} a_j^r =\sum _{j\ne i}\sum _{r\in [s]} a_j^r \ \ + \sum _{j\in [m]}\sum _{r\ne t} a_j^r. \end{aligned}$$

As the first summand does not depend on \(t\), the distance from \(E\) to the nearest \({\mathcal {S}}\)-consensus with winner \(c_i\) is \( \min _{t\in [s]} \widehat{d}(V,X_i^t)=\sum _{j\ne i} \sum _{r\in [s]} a_j^r + \min _{t\in [s]}\sum _{j\in [m]}\sum _{r\ne t} a_j^r. \) The second component of this expression does not depend on \(i\), while its first component counts the number of voters who do not rank \(c_i\) first. Thus, the nearest strong unanimity consensus to \(E\) has \(c_i\) as its winner if and only if \(i\) minimizes the sum \(\sum _{j\in [m]\setminus \{i\}}\sum _{r\in [s]} a_j^r\) over all \(i\in [m]\), i.e., \(c_i\) has the largest number of first-place votes. Thus, Plurality is \(({\mathcal {S}}, \widehat{d})\)-rationalizable.\(\square \)

Appendix C: Proof of theorem 5

As in the proof of Theorem 4, we set \(C = \{a,b,c\}\) and use the same notation as throughout Sect. 4.2. We focus on STV, and define \(V\) and \(W\) as in the proof of Theorem 4. Also, we set \(m_{sm}=\min \{S, M\}\), \(m_{sb}=\min \{S, B\}\).

We will consider the majority consensus first. For the sake of contradiction, suppose that STV is \(({\mathcal {M}}, \widehat{d}^N)\)-rationalizable for some neutral distance \(d\) over \(P(C)\) and a symmetric norm \(N\) that is monotonic in the positive orthant. Let \((C, M_a)\) be an \({\mathcal {M}}\)-consensus with winner \(a\) that is closest to \((C, V)\) and let \((C, M_b)\) be an \({\mathcal {M}}\)-consensus with winner \(b\) that is closest to \((C, V)\). We have \( \widehat{d}^N(V,M_a) = N(0,\ldots ,0,\min \{m_{sb},m_{sm}\}) = \widehat{d}^N(V,M_b). \) However, this is a contradiction as \(a\) is the unique winner of \((C, V)\). Thus, STV is not \(({\mathcal {M}},\widehat{d}^N)\)-rationalizable.

Similarly, suppose that STV is \(({\mathcal {U}},\widehat{d}^N)\)-rationalizable. Let \((C, U_a)\) be a \({\mathcal {U}}\)-consensus with winner \(a\) that is closest to \((C, V)\) and let \((C, U_b)\) be a \({\mathcal {U}}\)-consensus with winner \(b\) that is closest to \((C, V)\). We have

$$\begin{aligned} \widehat{d}^N(V, U_a)= & {} N(0, \dots , 0, m_{sm}, \dots , m_{sm}, m_{sb}). \\ \widehat{d}^N(V, U_b)= & {} N(m_{sb}, \dots , m_{sb}, 0, \dots , 0, m_{sm}). \end{aligned}$$

As \(a\) is the unique winner of \((C, V)\), we have \(\widehat{d}^N(V, U_a)<\widehat{d}^N(V, U_b)\), and therefore by symmetry and monotonicity \(m_{sm}<m_{sb}\).

Now, let \((C, U'_a)\) be a \({\mathcal {U}}\)-consensus with winner \(a\) that is closest to \((C, W)\) and let \((C, U'_b)\) be a \({\mathcal {U}}\)-consensus with winner \(b\) that is closest to \((C, W)\). We have

$$\begin{aligned} \widehat{d}^N(W, U'_a)= & {} N(0, \dots , 0, m_{sm}, \dots , m_{sm}, m_{sm}), \\ \widehat{d}^N(W, U'_b)= & {} N(m_{sb}, \dots , m_{sb}, 0, \dots , 0, m_{sb}). \end{aligned}$$

As \(b\) is the unique winner of \((C, W)\), we have \(\widehat{d}^N(W, U'_a) > \widehat{d}^N(W, U'_b)\), and hence \(m_{sm} > m_{sb}\), a contradiction. Thus, STV is not \(({\mathcal {U}},\widehat{d}^N)\)-rationalizable.

Appendix D: Condorcet consensus and homogeneity

The following example shows that the Condorcet consensus is not split-homogeneous. Consider the following election \(E=(C, V)\) with \(C=\{a, b, c, d, e\}\) and \(V=(v_1, \dots , v_{12})\).

\(v_1\)

\(v_2\)

\(v_3\)

\(v_4\)

\(v_5\)

\(v_6\)

\(v_7\)

\(v_8\)

\(v_9\)

\(v_{10}\)

\(v_{11}\)

\(v_{12}\)

\(a\)

\(b\)

\(c\)

\(d\)

\(e\)

\(c\)

\(e\)

\(a\)

\(b\)

\(c\)

\(d\)

\(c\)

\(b\)

\(c\)

\(d\)

\(e\)

\(a\)

\(a\)

\(d\)

\(e\)

\(a\)

\(b\)

\(c\)

\(a\)

\(c\)

\(d\)

\(e\)

\(a\)

\(b\)

\(b\)

\(c\)

\(d\)

\(e\)

\(a\)

\(b\)

\(b\)

\(d\)

\(e\)

\(a\)

\(b\)

\(c\)

\(d\)

\(b\)

\(c\)

\(d\)

\(e\)

\(a\)

\(d\)

\(e\)

\(a\)

\(b\)

\(c\)

\(d\)

\(e\)

\(a\)

\(b\)

\(c\)

\(d\)

\(e\)

\(e\)

Set \(V_1=(v_1, \dots , v_5)\), \(V_2=(v_7, \dots , v_{11})\), \(V'_1=(v_1, \dots , v_6)\), \(V_2=(v_7, \dots , v_{12})\). The profile \(V_1\) is a Condorcet cycle, and the profile \(V_2\) is obtained from \(V_1\) by reversing the preference order of each voter. Voters \(v_6\) and \(v_{12}\) are identical and rank candidate \(c\) first. Since voters from the groups \(V_1\) and \(V_2\) ‘cancel’ each other, \(c\) is the Condorcet winner in \(E\). On the other hand, consider elections \(E_1 = (C, V'_1)\) and \(E_2= (C, V'_2)\). In \(E_1\), \(b\) is ranked above \(c\) in \(4\) votes, so \(c\) is not a Condorcet winner in \(E_1\). Similarly, in \(E_2\), \(d\) is ranked above \(c\) in \(4\) votes, so \(c\) is not a Condorcet winner in \(E_2\) either.

While Dodgson\(^\infty \) is homogeneous, there are \(\ell _\infty \)-votewise distances that, paired with the Condorcet consensus, yield rules that fail homogeneity.

Proposition 10

There exists a set of candidates \(C\) and a distance \(d\) on \(P(C)\) such that the voting rule rationalized by \(({\mathcal {C}},\widehat{d}^\infty )\) is not homogeneous.

Proof

We will use two types of operations on preference profiles. Operations of the first type are the usual swaps of adjacent candidates. Let us now define an operation of the second type: a two-candidate shift. Given a preference order \(v\) and two candidates \(c\) and \(d\) that are ranked consecutively in \(v\), a two-candidate shift of \((c, d)\) pushes both \(c\) and \(d\) two positions upwards maintaining their relative ranking; if this is not possible, then the two-candidate shift is not defined. For example, given a preference order \(u = a\succ b\succ c\succ d\succ e\) over the candidate set \(C = \{a,b,c,d,e\}\), a two-candidate shift of \((c, d)\) transforms \(u\) into \(v = c\succ d\succ a\succ b\succ e\). Note that \(v\) can be transformed back into \(u\) by a two-candidate shift of \((a, b)\); generally, if a vote \(u'\) can be transformed into a vote \(v'\) by a two-candidate shift, then \(v'\) can be transformed into \(u'\) by a two-candidate shift as well.

We define the distance \(d\) between two preference orders \(u\) and \(v\) as the minimum number of operations of both types that transform \(u\) into \(v\). It is easy to see that \(d\) is indeed a distance, because it counts the number of reversible operations that are required to transform one preference order into the other. Let \(\mathcal{{R}}\) be the voting rule that outputs all \(({\mathcal {C}},\widehat{d}^\infty )\)-winners.

We will now build an election \(E = (C,V)\) such that \(\mathcal{{R}}(E) \ne \mathcal{{R}}(2E)\). Set \(C = \{a,b,c,d\} \cup S\), where \(S=\cup _{j=1}^{18} S_j\) is disjoint from \(\{a, b, c, d\}\) and \(S_1, \dots , S_{18}\) are pairwise disjoint sets of candidates of size \(4\) each. Let \(V = (v_1, v_2, \ldots , v_6)\); the voters’ preference orders are given in the following table, where the candidates in each set \(S_j\), \(j=1, \dots , 18\), are listed in an arbitrary order and each symbol \(\downarrow \) replaces all candidates that have not been listed explicitly in the vote (for instance, in the first vote \(\downarrow \) replaces \(S\setminus (S_1\cup S_6\cup S_{12})\)).

\(v_1\)

\(v_2\)

\(v_3\)

\(v_4\)

\(v_5\)

\(v_6\)

\(c\)

\(a\)

\(d\)

\(d\)

\(d\)

\(c\)

\(S_1\)

\(b\)

\(S_2\)

\(S_3\)

\(a\)

\(a\)

\(b\)

\(c\)

\(c\)

\(b\)

\(S_4\)

\(S_5\)

\(S_6\)

\(d\)

\(S_7\)

\(S_8\)

\(b\)

\(b\)

\(a\)

\(S_9\)

\(a\)

\(a\)

\(S_{10}\)

\(S_{11}\)

\(S_{12}\)

\(S_{13}\)

\(S_{14}\)

\(S_{15}\)

\(c\)

\(d\)

\(d\)

\(S_{16}\)

\(b\)

\(c\)

\(S_{17}\)

\(S_{18}\)

\(\downarrow \)

\(\downarrow \)

\(\downarrow \)

\(\downarrow \)

\(\downarrow \)

\(\downarrow \)

Note that each candidate in \(\{a,b,c,d\}\) is preferred to each candidate in \(S\) by at least five voters. Further, four voters prefer \(a\) to \(b\) and only two voters prefer \(b\) to \(a\), while all other head-to-head contests among the candidates in \(\{a,b,c,d\}\) are tied. Note also that one needs at least two operations (swaps or shifts) to change the relative order of two candidates separated by a set \(S_j\), \(j=1, \dots , 18\).

It is easy to verify that \(\mathcal{{R}}(E) = \{a\}\). Indeed, \(a\)’s \(({\mathcal {C}},\widehat{d}^\infty )\)-score equals \(1\) (to make \(a\) the Condorcet winner, it suffices to swap it with \(d\) in \(v_5\) and with \(c\) in \(v_6\)), whereas every other candidate requires at least two operations per vote to become the Condorcet winner.

On the other hand, we have \(d\in \mathcal{{R}}(2E)\). Indeed, it is easy to see that each candidate’s \(({\mathcal {C}},\widehat{d}^\infty )\)-score in \(2E\) is at least \(1\). Further, the \(({\mathcal {C}},\widehat{d}^\infty )\)-score of \(d\) in \(2E\) is \(1\): to make \(d\) the Condorcet winner, it suffices to swap \(d\) and \(c\) in one copy of \(v_2\), and perform a two-candidate shift of \((c, d)\) in the other copy of \(v_2\).\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Elkind, E., Faliszewski, P. & Slinko, A. Distance rationalization of voting rules. Soc Choice Welf 45, 345–377 (2015). https://doi.org/10.1007/s00355-015-0892-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00355-015-0892-5

Navigation