Skip to main content
Log in

Paired comparisons analysis: an axiomatic approach to ranking methods

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

Abstract

In this paper we present an axiomatic analysis of several ranking methods for general tournaments. We find that the ranking method obtained by applying maximum likelihood to the (Zermelo-)Bradley-Terry model, the most common method in statistics and psychology, is one of the ranking methods that perform best with respect to the set of properties under consideration. A less known ranking method, generalised row sum, performs well too. We also study, among others, the fair bets ranking method, widely studied in social choice, and the least squares method.

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. In matrix form, a binary tournament corresponds to a binary matrix \(A\in \{0,1\}^{n\times n}\) such that for each pair of different alternatives \(i\) and \(j,\, A_{ij}+A_{ji}=1\).

  2. This is specially important, for instance, in testing objects, where each pair of objects may be tested by several experts.

  3. Dutta and Laslier (1999) consider a setting where this last restriction is generalised to allow for intensities, but completeness is still a requirement.

  4. In recent years, the related issue of ranking scientific journals has received a lot of attention in economics (see, for instance, Liebowitz and Palmer (1984) and Palacios-Huerta and Volij (2004)). In this setting, the rankings are defined on the basis of citation matrices, which contain information regarding the number of times each journal has been cited by any other journal. There is a fundamental difference between the two settings. In our setting, a victory of \(i\) over \(j\) should be seen as something good for \(i\) and bad for \(j\). However, when looking at scientific journals, a citation from journal \(j\) to journal \(i\) should be good for journal \(i\), but not necessarily bad for journal \(j\). Clearly, this cannot be ignored when defining properties of a ranking method and, therefore, it would be inappropriate to include in our axiomatic analysis ranking methods that are based on citation matrices.

  5. More precisely, they generalise the aggregate net scores, which we define in Sect. 2.

  6. We do not restrict the non-zero entries in \(A\) to be natural numbers as in, e.g., Slutzki and Volij (2005). This choice of domain has no impact on the analysis, but allows for simpler definitions of the properties under study.

  7. This ensures that all the rankings methods we discuss in this paper are well defined.

  8. It is worth noting that, although the approach developed by Slutzki and Volij is quite appealing normatively, there may be settings where one might desire to define rankings directly on reducible tournaments. In particular, consider the following situation (we thank P. Chebotarev for suggesting this example). There are \(n+1\) players, with player \(i\) having beaten player \(j\) whenever \(i<j<n+1\). Further, player \(n+1\) has beaten player \(n\). According to the approach in Slutzki and Volij (2005), players \(1\) and \(n+1\) are incomparable (both of them have a perfect score). However, it might be argued that player \(1\), who has beaten \(n-1\) other players, should be regarded as stronger than player \(n+1\), who has only beaten the weakest player.

  9. Note that, although in our setting the number of matches \(M_{ij}(A)\) may not be an integer number, we always have in mind situations in which as many (possibly non-integer) points are divided between two players as matches they play. This extra generality could be useful in settings where partial comparisons are possible or, as we mentioned in the Introduction, different matches have different weights.

  10. Despite being very special, round-robin ranking problems are still more general than binary tournaments, since they allow for intensities (\(A_{ij}\) needs not be \(0\) or \(1\)) and ties (\(A_{ij}\) may equal \(A_{ji}\)).

  11. Sometimes we use the notation \(r(A)\) to indicate that the rating vector is derived from matrix \(A\).

  12. We postpone a discussion on the role of \(\varepsilon \) and this particular choice for this parameter to Sect. 3. Yet, it is worth noting that \(\varepsilon =\frac{1}{\hat{m} (n-2)}\) is not well defined for the (trivial) class of two-player ranking problems. Thus, \(n>2\) is implicitly assumed when the generalised row sum method is discussed.

  13. The reader interested in a deeper discussion of these and other ranking methods for this and other settings may refer to David (1988); Laslier (1997); Chebotarev and Shamis (1998, 1999) and Brozos-Vázquez et al. (2008).

  14. Refer, for instance, to Ford (1957) or David (1988).

  15. This (tie-breaking) rule, which originally was only defined for round-robin ranking problems, is commonly referred to as Sonneborn-Berger. In fact, Sonneborn and Berger criticised Neustadtl’s method by arguing that the players’s scores should be added to the Neustadtl ranking vector.

  16. Similar ideas have been also used in slightly different settings in papers such as Borm et al. (2002); Herings et al. (2005) and Slikker et al. (2010).

  17. Although Bruno Buchholz is widely recognised to have developed the method named after him (see World Chess Federation (2012)) in 1932, we were unfortunately not able to find a document on this by the inventor himself.

  18. Actually, Buchholz is commonly used as a tie-breaker in non round-robin chess tournaments and is computed as the average of the scores of the opponents of each player, i.e., \(\bar{M}s\). For players with an equal score, this definition results in the same relative ranking as the definition we present here.

  19. Although recursive Buchholz was not defined in Brozos-Vázquez et al. (2008), it can be seen as a variation of recursive performance where \(F_L\) is replaced by the identity. Thus, the existence and uniqueness of \(r^{\text{ rb}}\) follows from Theorem 2 in that paper.

  20. We thank P. Chebotarev for pointing this out.

  21. Given the similarity in the definitions of recursive Buchholz and recursive performance, one may wonder whether the latter can also be rewritten as a least squares method for some adequately chosen \(D_{ij}\). For recursive Buchholz, one of the key elements in the proof of Proposition 3.1 is that the \(D_{ij}\) entries can be seen as a disaggregation of the scores of the players, which are then recovered through the sums \(\sum _{j\in N} M_{ij}D_{ij}\). In the recursive performance, the role of the scores vector is played by the vector \(\hat{c}\). Yet, this vector aggregates the pairwise results of each player via the nonlinear function \(F_L\), which makes it impossible to disaggregate in a natural way the components \(\hat{c}\) to obtain the \(D_{ij}\) values.

  22. In Chebotarev (1994) this property is referred to as transposability.

  23. It is worth noting that the coincidence of the rankings proposed by maximum likelihood and the scores was already established in Zermelo (1929).

  24. To what extent it is important to work with ranking methods that satisfy this property depends on how the matches matrix is constructed; for instance, on whether or not the players can choose their own opponents.

  25. Two very similar properties are discussed in Chebotarev and Shamis (1998) and Conner and Grant (2009).

References

  • Arrow KJ (1963) Social choice and individual values. Wiley, New York

    Google Scholar 

  • Berman A, Plemmons R (1994) Nonnegative matrices in the mathematical sciences, 2nd edn. SIAM, Philadelphia

    Book  Google Scholar 

  • Birkhoff G (1946) Tres observaciones sobre el álgebra lineal. Univ Nac Tucumán Revista A 5:147–150

    Google Scholar 

  • Borm P, Fiestras-Janeiro G, Hamers H, Sánchez-Rodríguez E, Voorneveld M (2002) On the convexity of games corresponding with sequencing situations with due dates. Eur J Oper Res 136:616–634

    Article  Google Scholar 

  • Bouyssou D (2004) Monotonicity of ranking by choosing: a progress report. Soc Choice Welf 23:249–273

    Article  Google Scholar 

  • Bradley R, Terry M (1952) Rank analysis of incomplete block designs. I. The method of paired comparisons. Biometrika 39:324–345

    Google Scholar 

  • Brozos-Vázquez M, Campo-Cabana MA, Díaz-Ramos JC, González-Díaz J (2008) Ranking participants in tournaments by means of rating functions. J Math Econ 44:1246–1256

    Article  Google Scholar 

  • Chebotarev P (1989) Generalization of the row sum method for incomplete paired comparisons. Autom Remote Control 50:1103–1113

    Google Scholar 

  • Chebotarev P (1994) Aggregation of preferences by the generalized row sum method. Mathe Soc Sci 27:293–320

    Article  Google Scholar 

  • Chebotarev P, Shamis E (1997) Econometric Decision Models: Constructing Scalar-Valued Objective Functions, of Lecture Notes in Economics and Mathematical Systems. In: Tangian A, Gruber J (eds) Constructing an objective function for aggregating incomplete preferences, vol 453. Springer-Verlag, Berlin, pp 100–124

    Google Scholar 

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

    Article  Google Scholar 

  • Chebotarev P, Shamis E (1999) Preference fusion when the number of alternatives exceeds two: indirect scoring procedures. J Frankl Inst 336:205–226

    Article  Google Scholar 

  • Conner GR, Grant CP (2009) Neighborhood monotonicity, the extended Zermelo model, and symmetric knockout tournaments. Discret Math 309:3998–4010

    Article  Google Scholar 

  • Copeland AH (1951) A reasonable social welfare function. In: Seminar on Applications of Mathematics to the Social sciences. University of Michigan

  • Daniels H (1969) Round-robin tournament scores. Biometrika 56:295–299

    Article  Google Scholar 

  • David HA (1988) The method of paired comparisons. Chapman and Hall, London

    Google Scholar 

  • Dutta B, Laslier J (1999) Comparison functions and choice correspondences. Soc Choice Welf 16:513–552

    Article  Google Scholar 

  • Dykstra O (1956) A note on the rank analysis of incomplete block designs: applications beyond the scope of existing tables. Biometrics 3:301–306

    Article  Google Scholar 

  • Ford LR (1957) Solution of a ranking problem from binary comparisons. Am Math Mon 64:28–33

    Article  Google Scholar 

  • Gulliksen H (1956) A least squares solution for paired comparisons with incomplete data. Psychometrika 21:125–134

    Article  Google Scholar 

  • Herings P, van der Laan G, Talman D (2005) The positional power nodes in digraphs. Soc Choice Welf 24:439–454

    Article  Google Scholar 

  • Horst P (1932) A method for determining the absolute affective value of a series of stimulus situations. J Educational Psychol 23:418–440

    Article  Google Scholar 

  • Kaiser HF, Serlin RC (1978) Contributions to the method of paired comparisons. Appl Psychol Meas 2:421–430

    Article  Google Scholar 

  • Kendall M (1955) Further contributions to the theory of paired comparisons. Biometrics 11:43–62

    Article  Google Scholar 

  • Langville AN, Meyer CD (2012) The science of rating and ranking. Princeton University Press, Princeton

    Google Scholar 

  • Laslier J (1992) Solutions de tournois. Technical report, Habilitation à diriger les recherches en Economie, Université de Cergy-Pontoise

  • Laslier J (1997) Tournament solutions and majority voting. Springer-Verlag, New York

    Book  Google Scholar 

  • Levchenkov V (1992) Social choice theory: a new insight. Technical report, Institute of Systems Analysis, Moscow

  • Liebowitz S, Palmer J (1984) Assesing the relative impacts of economic journals. J Econ Lit 22:77–88

    Google Scholar 

  • Moon JW, Pullman N (1970) On generalized tournament matrices. SIAM Rev 12:384–399

    Article  Google Scholar 

  • Mosteller F (1951) Remarks on the method of paired comparisons. I. The least squares solution assuming equal standard deviations and equal correlations. Psychometrika 16:3–11

    Article  Google Scholar 

  • Neustadtl H (1882) Letter to The Chess Monhtly, London. In: L. Hoffer and J. Zukertort eds

  • Palacios-Huerta I, Volij O (2004) The measurement of intellectual influence. Econometrica 72:963–977

    Article  Google Scholar 

  • Rubinstein A (1980) Ranking the participants in a tournament. SIAM J Appl Math 38:108–111

    Article  Google Scholar 

  • Slikker M, Borm P, van den Brink R (2010) Internal slackening scoring methods. Beta working paper, WP-306. Technische Universiteit Eindhoven, Eindhoven

    Google Scholar 

  • Slutzki G, Volij O (2005) Ranking participants in generalized tournaments. Int J Game Theory 33:255–270

    Article  Google Scholar 

  • Slutzki G, Volij O (2006) Scoring of web pages and tournaments: axiomatizations. Soc Choice Welf 26: 75–92

    Article  Google Scholar 

  • Wei T (1952) The algebraic foundations of ranking theory. Ph. D. thesis, Cambridge University.

  • World Chess Federation (2012) Handbook. C: General rules and recommendations for tournaments. Annex to the FIDE tournament regulations regarding tiebreaks. http://www.fide.com/component/handbook

  • Zermelo E (1929) Die Berechnung der Turnier-Ergebnisse als ein Maximumproblem der Wahrscheinlichkeitsrechnung. Mathematische Zeitschrift 29:436–460

    Article  Google Scholar 

Download references

Acknowledgments

We thank Miguel Brozos-Vázquez and José Carlos Díaz-Ramos for helpful discussions. We also thank Pavel Chebotarev and an anonymous referee for their comments on earlier versions of the manuscript. Julio González Díaz acknowledges the financial support of the Spanish Ministry for Science and Innovation through project MTM2011-27731-c03, and from the Xunta de Galicia through project INCITE09-207-064-PR.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Julio González-Díaz.

A Proofs of Propositions 6.4 and 6.5

A Proofs of Propositions 6.4 and 6.5

We start with an auxiliary result that is crucial in the proof for both fair bets and least squares.

Lemma 8.1

Let \(B\in \mathbb R ^{n\times n}\) be such that

  1. (i)

    \(B\) is invertible,

  2. (ii)

    for all \(i\ne j,\, B_{ij}\le 0\), and

  3. (iii)

    \(\sum _{j=1}^n B_{ji}\ge 0\) for all \(i\in \{1,\dots ,n\}\).

If \(\gamma ,\uplambda \in \mathbb R ^n\) are two vectors such that \(\uplambda \) is nonnegative and \(B\gamma =\lambda \), then \(\gamma \) is nonnegative.

Proof

It follows from the assumptions that \(B\) is an M-matrix (see, for instance, Berman and Plemmons (1994)). As a result, \(B^{-1}\) is non-negative, from which the result follows immediately. \(\square \)

Proof of Proposition 6.4

Let \(i,j,k\in N\) be as in the statement. Below we explicitly characterise how the recursive Buchholz ratings vary as a function of \(A_{ik}\) and \(A_{ki}\), provided that \(M_{ik}\) stays constant. Recall that \(r^{\text{ rb}}\) is the unique solution of \(\bar{M}x+\hat{s}=x\) such that \((r^{\text{ rb}})^{\top }e=0\). Hence, \((I-\bar{M})r^{\text{ rb}}=\hat{s}\). Define \(B=I-\bar{M}\) and \(\breve{N}=N\backslash \{i,k\}\). Then, equation \(\ell \) of the system \(Br^{\text{ rb}}=\hat{s}\) can be written as

$$\begin{aligned} \sum _{h\in \breve{N}}B_{\ell h}r^{\text{ rb}}_h=\hat{s}_{\ell }-B_{\ell i}r^{\text{ rb}}_i-B_{\ell k}r^{\text{ rb}}_k. \end{aligned}$$
(8.1)

with \(\ell \in \breve{N}\). Define \(\breve{B}\in \mathbb R ^{(n-2)\times (n-2)}\) to be the matrix obtained from \(B\) by deleting the rows and columns corresponding to players \(i\) and \(k\).

We prove now that \(\breve{B}\) is invertible. Suppose, on the contrary, that there is an \(y\in \mathbb R ^{n-2},\, y\ne 0\), such that \(y^{\top }\breve{B}^{\top }=0\). Let \(\ell \in \breve{N}\) be such that \(y_{\ell }=\max _{h\in \breve{N}}y_h\). We assume, without loss of generality, that \(y_{\ell }>0\). For each \(h\ne \ell ,\, B^{\top }_{h\ell }\le 0\) and, hence, \(-y_hB^{\top }_{h\ell }\le -y_{\ell }B^{\top }_{h\ell }\), with equality only if \(y_h=y_{\ell }\) or \(B^{\top }_{h\ell }=0\). Since \(y^{\top }\breve{B}^{\top }=0,\, \sum _{h\in \breve{N}\backslash \{\ell \}} -y_hB^{\top }_{h\ell }=y_{\ell }B^{\top }_{\ell \ell }\). Further, since \(\sum _{h\in N}B^{\top }_{h\ell }=0\), we have \(\sum _{h\in \breve{N}\backslash \{\ell \}}-B^{\top }_{h\ell }=B^{\top }_{\ell \ell }+B^{\top }_{i\ell }+B^{\top }_{k\ell }\le B^{\top }_{\ell \ell }\), with equality only if \(B^{\top }_{i\ell }=B^{\top }_{k\ell }=0\). Then, we have

$$\begin{aligned} y_{\ell }B^{\top }_{\ell \ell }=\sum _{h\in \breve{N}\backslash \{\ell \}}-y_hB^{\top }_{h\ell }\le y_{\ell }\sum _{h\in \breve{N}\backslash \{\ell \}}-B^{\top }_{h\ell }\le y_{\ell }B^{\top }_{\ell \ell } \end{aligned}$$

and, hence, all the inequalities are indeed equalities. Therefore, \(B^{\top }_{i\ell }=B^{\top }_{k\ell }=0\) and, for each \(h\in \breve{N}\backslash \{\ell \},\, y_h=y_{\ell }\) or \(B^{\top }_{h\ell }=0\). Define \(\bar{N}=\{m\in \breve{N}\,|\,y_m=\max _{h\in \breve{N}}y_h\}\). Now, for each \(m\in \bar{N}\), we have \(B^{\top }_{im}=B^{\top }_{km}=0\) and, further, for each \(h\in \breve{N}\backslash \bar{N},\, B^{\top }_{hm}=0\). That is, no player outside \(\bar{N}\) has played against players inside \(\bar{N}\), which contradicts the irreducibility of \(A\).

Define \(C=(\breve{B})^{-1},\, \breve{r}^{\text{ rb}}=(r^{\text{ rb}}_h)_{h\in \breve{N}},\, B^i=(B_{hi})_{h\in \breve{N}}\) and \(B^k=(B_{hk})_{h\in \breve{N}}\). Then, using (8.1) we have \(\breve{B}\breve{r}^{\text{ rb}}=\breve{s}-B^ir^{\text{ rb}}_i-B^kr^{\text{ rb}}_k\) and hence, \(\breve{r}^{\text{ rb}}=C(\breve{s}-B^ir^{\text{ rb}}_i-B^kr^{\text{ rb}}_k)\). So, for all \(\ell \in \breve{N}\),

$$\begin{aligned} r^{\text{ rb}}_{\ell }=\breve{r}^{\text{ rb}}_{\ell }=\sum _{h\in \breve{N}}C_{\ell h}(\breve{s}_h-B_{hi}r^{\text{ rb}}_i-B_{hk}r^{\text{ rb}}_k). \end{aligned}$$

Define \(\gamma ^{s}_{\ell }=\sum _{h\in \breve{N}}C_{\ell h}\breve{s}_h,\, \gamma ^i_{\ell }=-\sum _{h\in \breve{N}}C_{\ell h} B_{hi}\) and \(\gamma ^k_{\ell }=-\sum _{h\in \breve{N}}C_{\ell h}B_{hk}\). Then, for each \(\ell \in \breve{N}\),

$$\begin{aligned} r^{\text{ rb}}_{\ell }=\gamma ^{s}_{\ell }+\gamma ^i_{\ell }r^{\text{ rb}}_i+\gamma ^k_{\ell }r^{\text{ rb}}_k. \end{aligned}$$
(8.2)

Furthermore, equation \(i\) in \(Br^{\text{ rb}}=\hat{s}\) is

$$\begin{aligned} B_{ii}r^{\text{ rb}}_i+B_{ik}r^{\text{ rb}}_k+\sum _{\ell \in \breve{N}}B_{i\ell }r^{\text{ rb}}_{\ell }=\hat{s}_i. \end{aligned}$$
(8.3)

Define \(\Gamma ^{i,i}=-\sum _{\ell \in \breve{N}}B_{i\ell }\gamma ^i_{\ell }\) and \(\Gamma ^{i,k}=-\sum _{\ell \in \breve{N}}B_{i\ell }\gamma ^k_{\ell }\). Then, plugging in the expression of each \(r^{\text{ rb}}_{\ell }\) (8.2) into (8.3) we get

$$\begin{aligned} (B_{ii}-\Gamma ^{i,i})r^{\text{ rb}}_i+(B_{ik}-\Gamma ^{i,k})r^{\text{ rb}}_k=\hat{s}_i-\sum _{\ell \in \breve{N}}\gamma ^{s}_{\ell }. \end{aligned}$$
(8.4)

Now, adding up (8.2) over all \(\ell \in \breve{N}\) and using that \(\sum _{h\in N}r^{\text{ rb}}_h=0\),

$$\begin{aligned} (1+\sum _{\ell \in \breve{N}}\gamma ^i_{\ell })r^{\text{ rb}}_i+(1+\sum _{\ell \in \breve{N}}\gamma ^k_{\ell })r^{\text{ rb}}_k=-\sum _{\ell \in \breve{N}} \gamma ^{s}_{\ell }. \end{aligned}$$
(8.5)

Define \(\sigma _i=\sum _{\ell \in \breve{N}}\gamma ^i_{\ell }\) and \(\sigma _k=\sum _{\ell \in \breve{N}}\gamma ^k_{\ell }\). Then, solving Eqs. (8.4) and (8.5), we get

$$\begin{aligned} r^{\text{ rb}}_i=\frac{\hat{s}_i-(1-\frac{B_{ik}-\Gamma ^{i,k}}{1+\sigma _k})\sum _{\ell \in \breve{N}} \gamma ^{s}_{\ell }}{(B_{ii}-\Gamma ^{i,i})-(B_{ik}-\Gamma ^{i,k})\frac{1+\sigma _k}{1+\sigma _i}} \quad \text{ and} \quad r^{\text{ rb}}_k=\frac{-\sum _{\ell \in \breve{N}}\gamma ^{s}_{\ell }}{1+\sigma _k}-\frac{1+\sigma _i}{1+\sigma _k}r^{\text{ rb}}_i.\nonumber \\ \end{aligned}$$
(8.6)

To understand how \(r^{\text{ rb}}_i\) and \(r^{\text{ rb}}_k\) vary with \(\hat{s}_i\), it is convenient to know the signs of \(\gamma ^i\) and \(\gamma ^k\). We claim that both \(\gamma ^i\) and \(\gamma ^k\) are nonnegative vectors. By definition, \(\gamma ^i=-CB^i\) and, since \(C^{-1}=\breve{B},\, \breve{B}\gamma =-B^i\). Furthermore, \(-B^i\ge 0\). Since matrix \(\breve{B}\) and vectors \(\gamma ^i\) and \(-B^i\) satisfy the conditions of Lemma 8.1, \(\gamma ^i\) is nonnegative. The argument for \(\gamma ^k\) is analogous using \(-B^k\) instead of \(-B^i\). The nonnegativity of \(\gamma ^i\) and \(\gamma ^k\) implies that \(\sigma _i\) and \(\sigma _k\) are also nonnegative. Since \(\gamma ^k\) is nonnegative, also \(\Gamma ^{i,k}\) is nonnegative and \(B_{ik}-\Gamma ^{i,k}\) is negative. Furthermore,

$$\begin{aligned} B_{ii}-\Gamma ^{i,i}=B_{ii}+\sum _{\ell \in \breve{N}}B_{i\ell }\gamma ^i_{\ell }\ge B_{ii}+\sum _{\ell \in \breve{N}} B_{i\ell }\ge 0. \end{aligned}$$

We reexamine now Eq. (8.6). Note that \(\gamma ^{s},\, \gamma ^i,\, \gamma ^k,\, \Gamma ^{i,i},\, \Gamma ^{i,k},\, B_{ii}\) and \(B_{ik}\) only depend on \(\breve{B}\). Then, the denominator of the expression for \(r^{\text{ rb}}_i\) is positive and so \(r^{\text{ rb}}_i\) is strictly increasing in \(\hat{s}_i\). Further, since \(r^{\text{ rb}}_k\) is strictly decreasing in \(r^{\text{ rb}}_i\), it is strictly decreasing in \(\hat{s}_i\).

Now, because of (8.2), \(r^{\text{ rb}}_{\ell }\) is weakly increasing in \(r^{\text{ rb}}_i\) and \(r^{\text{ rb}}_k\). Yet, since \(r^{\text{ rb}}_i\) and \(r^{\text{ rb}}_k\) are strictly increasing and decreasing, respectively, in \(\hat{s}_i\), some extra work is needed to understand how \(r^{\text{ rb}}_{\ell }\) varies with \(\hat{s}_i\). To do so, we first show that all the components of \(\gamma ^i\) and \(\gamma ^k\) are no larger than 1. We prove it for \(\gamma ^i\), the proof for \(\gamma ^k\) being analogous.

$$\begin{aligned} \breve{B}(e-\gamma ^i)=\breve{B}e-\breve{B}\gamma ^i=\breve{B}e- B^i, \end{aligned}$$

and, for each \(\ell \in \breve{N}\),

$$\begin{aligned} (\breve{B}e-B^i)_{\ell }=\sum _{h\in \breve{N}}B_{\ell h}+B_{\ell i}\ge \sum _{h\in \breve{N}}B_{\ell h}+B_{\ell i}+B_{ki}=0. \end{aligned}$$

Then, since \(\breve{B}e-B^i\) is a nonnegative vector, matrix \(\breve{B}\) and vectors \(e-\gamma ^i\) and \(\breve{B}e- B^i\) are in the conditions of Lemma 8.1 and, hence, \(e-\gamma ^i\) is nonnegative.

Therefore, we know that all the components of \(\gamma ^i\) and \(\gamma ^k\) are no larger than 1. Looking again at Eq. (8.2), we have that \(r^{\text{ rb}}_{\ell }\) cannot increase with \(\hat{s}_i\) faster than \(r^{\text{ rb}}_i\) so \(r^{\text{ rb}}_{\ell }/r^{\text{ rb}}_i\) is weakly decreasing in \(\hat{s}_i\). Similarly, \(r^{\text{ rb}}_{\ell }/r^{\text{ rb}}_k\) is weakly increasing in \(\hat{s}_i\). From this, the statement follows. \(\square \)

Proof of Proposition 6.5

The proof of is analogous to the proof of Proposition 6.4, with \(B=L_A-A\) instead of \(B=I-\bar{M}\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

González-Díaz, J., Hendrickx, R. & Lohmann, E. Paired comparisons analysis: an axiomatic approach to ranking methods. Soc Choice Welf 42, 139–169 (2014). https://doi.org/10.1007/s00355-013-0726-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00355-013-0726-2

Keywords

Navigation