1 Introduction

Chess tournaments are often organized in the Swiss system when there are too many participants to play a round-robin tournament. They go for a predetermined number of rounds, in each round two players compete head-to-head. None of them are eliminated during the tournament, but there are pairs of players without a match between them. Let us denote the number of rounds by c and the number of participants by n.

Two emerging issues in Swiss system tournaments are how to pair the players and how to rank the participants on the basis of their respective results. The pairing algorithm aims to pair players with a similar performance, measured by the number of their wins and draws (see FIDE 2015 for details).

The ranking involves two main challenges. The first one is the possible appearance of circular triads when player i has won against player j, player j has won against player k, but player k has won against player i. The second issue arises as the consequence of incomplete comparisons since \(c < n-1\). For example, if player i has played only against player j, then its rank probably should depend on the results of player j.

The final ranking of the players is usually determined by the aggregated number of points scored: the winner of a match gets one point, the loser gets zero points, and a draw means half-half points for both players. However, it usually does not result in a linear order (a complete, transitive and antisymmetric binary relation) of the participants.Footnote 1 Ties are eliminated by the sequential application of various tie-breaking rules (FIDE 2015).

These ranking(s), based on lexicographical orders, will be called official ranking(s). They can differ in the tie-breaking rules.

Official rankings are not able to solve the problem caused by different schedules as players with weaker opponents can score the same number of points more easily (Brozos-Vázquez et al. 2010; Csató 2012, 2013; Forlano 2011; Jeremic and Radojicic 2010; Redmond 2003). It turns out that players with an improving performance during the tournament are preferred contrary to players with a declining one. Consider two players i and j with an equal number of points after playing some rounds. Player i is said to be on the inner circle if it scored more points in some of the first rounds relative to player j who is said to be on the outer circle. Since they have played against opponents with a similar number of points in each round, it is probable that player j has met with weaker opponents. Tie-breaking rules may take the performance of opponents into account but a similar problem arises if player j has marginally more points than player i as a lexicographical order is not continuous.

This is known to be an inherent defect of these systems. In fact, players sometimes may deliberately seek for a draw or defeat in the first (Forlano 2011). It is tolerated because the case concerns very rarely the winner of the tournament, at least with an adequate number of rounds.

Nevertheless, some works have aimed to improve on ranking in Swiss system tournaments. Redmond (2003) presents a generalization of win-loss ratings by accounting for the strength of schedule. Brozos-Vázquez et al. (2010) argue for the use of recursive methods (as a tie-breaking rule) in Swiss system chess tournaments. Forlano (2011) shows a way to correct the points for wins and draws in order to derive a more legitimate ranking.

The current paper will attempt to solve the problem through the use of paired comparison-based ranking procedures. In a sense, it is a return to the origins of this line of research as early works were often inspired by chess tournaments (Landau 1895, 1914; Zermelo 1929).

We use a parametric family of scoring procedures based on linear algebra, the generalized row sum (Chebotarev 1989, 1994) as well as the least squares method, which was extensively used for sport rankings (Leeflang and Praag 1971; Stefani 1980). Despite the issue has been touched by Csató (2013), here a deeper methodological foundation will be given for the problem and the evaluation of rankings will be revisited.

The analysis is based on some recent results. González-Díaz et al. (2014) have discussed the axiomatic properties of generalized row sum and least squares, Csató (2015) has given an interpretation for the least squares method, while Can (2014) has contributed to the choice of distance functions between rankings.

In order to avoid the prominent role of colour allocation in individual championships, the discussion focuses on team tournaments, where a match between two teams is played on 2t boards such that t players of a team play with white and the other t players of the team play with black. The winner of a game on a board gets 1 point, the loser gets 0 points, and the draw yields 0.5 points for both teams, thus 2t board points are allocated in a given match. The winner team achieving more (at least \(t + 0.5\)) board points scores 2 match points, the loser 0, while a draw results in 1 match point for both teams. Therefore one can choose between the board points and the match points as the basis of the official ranking. Recently the use of match points is preferred in chess olympiads and team European championships.

The paper is structured as follows. Section 2 shortly outlines the ranking problem, ranking methods and their relevant properties. It also aims to incorporate Swiss system chess team tournaments into this framework. The proposed model is applied in Sect. 3 to rank the participants in the 2011 and 2013 European Team Chess Championship open tournaments. Twelve rankings, distinguished by the influence of opponents’ performance and match versus table points, are introduced and compared on the basis of their distances. They are visualized with multidimensional scaling (MDS), while differences to the official ranking are revealed by the decomposition of the least squares method.

On the basis of these results, we argue for the use of least squares method with a generalized result matrix favouring match points. It is supported by several arguments, variability of the ranking with respect to the role of match and board points as well as retrodictive performance (the ability to match the outcomes of matches already played) and robustness (stability of the ranking between two subsequent rounds).

Finally, Sect. 4 summarizes the findings and review possible extensions of the model. A reader familiar with ranking problems (González-Díaz et al. 2014; Csató 2015) may skip Sects. 2.1 and 2.2.

2 Modelling of the tournament as a ranking problem

In the following some fundamental concepts of the paired comparison-based ranking methodology are presented. A detailed discussion can be found in Csató (2015).

2.1 The ranking problem

Let \(N = \{ 1,2, \dots ,n \}\), \(n \in \mathbb {N}\) be a set of objects. The matches matrix \(M = (m_{ij}) \in \mathbb {N}^{n \times n}\) contains the number of comparisons between the objects, and is symmetric (\(M^\top = M\)).Footnote 2 Diagonal elements \(m_{ii}\) are supposed to be 0 for all \(i = 1,2, \dots ,n\). Let \(d_i = \sum _{j=1}^n m_{ij}\) be the total number of comparisons of object i and \(\mathfrak {d} = \max \{ d_i: i \in N \}\) be the maximal number of comparisons with the other objects. Let \(m = \max \{ m_{ij}: i,j \in N \}\).

The results matrix \(R = (r_{ij}) \in \mathbb {R}^{n \times n}\) contains the outcome of comparisons between the objects, and is skew-symmetric (\(R^\top = -R\)). All elements are limited by \(r_{ij} \in \left[ -m_{ij},\, m_{ij} \right] \). \((r_{ij} + m_{ij})/(2m_{ij}) \in \left[ 0,1 \right] \) may be regarded as the likelihood that object i defeats j. Then \(r_{ij} = m_{ij}\) means that i is perfectly better than j, and \(r_{ij} = 0\) corresponds to an undefined relation (if \(m_{ij} = 0\)) or to the lack of preference (if \(m_{ij} > 0\)) between the two objects. A ranking problem is given by the triplet (NRM). Let \(\mathcal {R}^n\) be the class of ranking problems with \(|N| = n\).

A ranking problem is called round-robin if \(m_{ij} = 1\) for all \(i \ne j\), that is, every object has been compared exactly once with all of the others. A ranking problem is called balanced if \(d_{i} = d_{j}\) for all \(i,j = 1,2, \dots ,n\), that is, every object has the same number of comparisons.

2.2 Rankings derived from rating functions

Matches matrix M can be represented by an undirected multigraph \(G := (V,E)\) where vertex set V corresponds to the object set N, and the number of edges between objects i and j is equal to \(m_{ij}\). The number of edges adjacent to i is the degree \(d_i\) of node i. A path from object \(k_1\) to object \(k_t\) is a sequence of objects \(k_1, k_2, \dots , k_t\) such that \(m_{k_\ell k_{\ell +1}} > 0\) for all \(\ell = 1,2, \dots ,t-1\). Two vertices are connected if G contains a path between them. A graph is said to be connected if every pair of vertices is connected.

Graph G is called the comparison multigraph associated with the ranking problem (NRM), and is independent of the results of paired comparisons. The Laplacian matrix \(L = (\ell _{ij}) \in \mathbb {R}^{n \times n}\) of graph G is given by \(\ell _{ij} = -m_{ij}\) for all \(i \ne j\) and \(\ell _{ii} = d_i\) for all \(i = 1,2, \dots ,n\).

Vectors are denoted by bold fonts, and assumed to be column vectors. Let \(\mathbf {e} \in \mathbb {R}^n\) be given by \(e_i = 1\) for all \(i = 1,2, \dots ,n\) and \(I \in \mathbb {R}^{n \times n}\) be the matrix with \(I_{ij} = 1\) for all \(i,j = 1,2, \dots ,n\).

A rating (scoring) method f is an \(\mathcal {R}^n \rightarrow \mathbb {R}^n\) function, \(f_i = f_i(N,R,M)\) is the rating of object i. It defines a ranking method by i weakly above j in the ranking problem (NRM) if and only if \(f_i(N,R,M) \ge f_j(N,R,M)\). Throughout the paper, the notions of rating and ranking methods will be used analogously since all ranking procedures discussed are based on rating vectors. Rating methods \(f^1\) and \(f^2\) are called equivalent if they result in the same ranking for any ranking problem (NRM).

Let us introduce some rating methods.

Definition 1

Row sum method: \(\mathbf {s}(N,R,M): \mathcal {R}^n \rightarrow \mathbb {R}^n\) such that \(\mathbf {s} = R \mathbf {e}\).

Row sum will also be referred to as scores, \(\mathbf {s}\) is sometimes called the scores vector. It does not take the comparison multigraph into account.

Definition 2

Generalized row sum method: \(\mathbf {x}(\varepsilon )(N,R,M): \mathcal {R}^n \rightarrow \mathbb {R}^n\) such that

$$\begin{aligned} \left( I+ \varepsilon L\right) \mathbf {x}(\varepsilon ) = \left( 1 + \varepsilon m n\right) \mathbf {s}, \end{aligned}$$

where \(\varepsilon > 0\) is a parameter.

This parametric rating procedure was constructed axiomatically by Chebotarev (1989) and thoroughly analysed in Chebotarev (1994). Generalized row sum adjusts the standard row sum by accounting for the performance of objects compared with it, and so on. \(\varepsilon \) indicates the importance attributed to this correction. It follows from the definition that \(\lim _{\varepsilon \rightarrow 0} \mathbf {x}(\varepsilon ) = \mathbf {s}\) for all ranking problems (NRM).

Both the row sum and the generalized row sum ratings are well-defined and can be obtained from a system of linear equations for all ranking problems since \((I+ \varepsilon L)\) is positive definite for any \(\varepsilon \ge 0\).

In our model the outcome of paired comparisons is restricted by \(-m \le r_{ij} \le m\) for all \(i,j \in N\). Then Chebotarev (1994, Proposition 5.1) argues that the reasonable upper bound of \(\varepsilon \) is \(1 / \left[ m(n-2) \right] \).

\(h_{ij} = r_{ij} / \min \{ m_{ij}; 1 \} \in \left[ -1, 1 \right] \) Footnote 3 may be identified as the normalized difference between the latent ratings \(q_i\) and \(q_j\) of objects i and j. Then it makes sense to choose \(\mathbf {q}\) in order to minimize the error according to an appropriate objective function.

Definition 3

Least squares method: \(\mathbf {q}(N,R,M): \mathcal {R}^n \rightarrow \mathbb {R}^n\) such that it is the solution to the problem

$$\begin{aligned} \min _{\mathbf {q} \in \mathbb {R}^n} m_{ij} \left[ h_{ij} - \left( q_i - q_j \right) \right] ^2 \end{aligned}$$

satisfying \(\mathbf {e}^\top \mathbf {q} = 0\).

The normalization \(\mathbf {e}^\top \mathbf {q} = 0\) is necessary because the value of the objective function is the same for \(\mathbf {q}\) and \(\mathbf {q} + \beta \mathbf {e}\), \(\beta \in \mathbb {R}\).

The least squares ranking method is well-known in a lot of fields, a review about its origin is given by González-Díaz et al. (2014) and Csató (2015). It has strong connections to generalized row sum.

Proposition 1

The least squares rating can be obtained as a solution of the linear system of equations \(L \mathbf {q} = \mathbf {s}\) and \(\mathbf {e}^\top \mathbf {q} = 0\) for all ranking problems (NRM).

Proof

See Csató (2015, p. 57). \(\square \)

Lemma 1

For all ranking problems (NRM), the least squares method is equivalent to the limit of generalized row sum if \(\varepsilon \rightarrow \infty \) since \(\lim _{\varepsilon \rightarrow \infty } \mathbf {x}(\varepsilon ) = m n \mathbf {q}\).

Proof

See Chebotarev and Shamis (1998, p. 326) and Csató (2016a). \(\square \)

Proposition 2

The least squares rating \(\mathbf {q}(N,R,M)\) is unique if and only if comparison multigraph G of the ranking problem is (NRM) connected.

Proof

See Csató (2015, p. 59). \(\square \)

Note that in the case of an unconnected comparison multigraph there are independent ranking problems.

A graph-theoretic interpretation of the generalized row sum method is given by Shamis (1994). Csató (2015) provides the following iterative decomposition of least squares.

Proposition 3

Let the comparison multigraph of a ranking problem (NRM) be connected and not regular bipartite. Then the unique solution of the least squares problem is \(\mathbf {q} = \lim _{k \rightarrow \infty } \mathbf {q}^{(k)}\) where

$$\begin{aligned} \mathbf {q}^{(0)}= & {} \left( 1/ \mathfrak {d}\right) \mathbf {s},\\ \mathbf {q}^{(k)}= & {} \mathbf {q}^{(k-1)} + \frac{1}{\mathfrak {d}} \left[ \frac{1}{\mathfrak {d}} \left( \mathfrak {d}I - L \right) \right] ^k \mathbf {s} \quad (k = 1,2, \dots ). \end{aligned}$$

2.3 Two properties of scoring procedures

In order to argue for the use of these methods we discuss some axioms.

Definition 4

Admissible transformation of the results (Csató 2014): Let \((N,R,M) \in \mathcal {R}^n\) be a ranking problem. An admissible transformation of the results provides a ranking problem \((N,kR,M) \in \mathcal {R}^n\) such that \(k > 0\), \(k \in \mathbb {R}\) and \(k r_{ij} \in \left[ -m_{ij},m_{ij} \right] \) for all \(i \in N\).

Multiplier k cannot be too large since \(-m_{ij} \le k r_{ij} \le m_{ij}\) should be satisfied for all \(i,j \in N\) according to the definition of the results matrix. \(k \le 1\) is always allowed.

Definition 5

Scale invariance (SI) (Csató 2014): Let \((N,R,M), (N,kR,M) \in \mathcal {R}^n\) be two ranking problems such that (NkRM) is obtained from (NRM) through an admissible transformation of the results. Scoring procedure \(f: \mathcal {R}^n \rightarrow \mathbb {R}^n\) is scale invariant if \(f_i(N,R,M) \ge f_j(N,R,M) \Leftrightarrow f_i(N,kR,M) \ge f_j(N,kR,M)\) for all \(i,j \in N\).

Scale invariance implies that the ranking is invariant to a proportional modification of wins (\(r_{ij} > 0\)) and losses (\(r_{ij} < 0\)). It seems to be important for applications. If the outcomes of paired comparisons cannot be measured on a continuous scale, it is not trivial how to transform them into \(r_{ij}\) values. SI provides that it is not a problem in certain cases. For example, if only three outcomes are possible, the coding (\(r_{ij} = \kappa \) for wins; \(r_{ij} = 0\) for draws; \(r_{ij} = -\kappa \) for losses) makes the ranking independent from \(0 < \kappa \le 1\). It may also be advantageous when relative intensities are known such as a regular win is two times better than an overtime triumph.

Lemma 2

The row sum, generalized row sum and least squares methods satisfy SI.

Proof

See Csató (2014, Lemma 4.3). \(\square \)

One disadvantage of the row sum procedure is its independence of irrelevant matches (González-Díaz et al. 2014; Csató 2016b). However, it causes no problem in the round-robin case, so it makes sense to preserve the attributes of row sum on this set.

Definition 6

Score consistency (SCC) (González-Díaz et al. 2014): Scoring procedure \(f: \mathcal {R}^n \rightarrow \mathbb {R}^n\) is score consistent if \(f_i(N,R,M) \ge f_j(N,R,M) \Leftrightarrow s_i(N,R,M) \ge s_j(N,R,M)\) for all \(i,j \in N\) and round-robin ranking problem \((N,R,M) \in \mathcal {R}^n\).

A score consistent method is equivalent to the row sum method in the case of round-robin ranking problems. A similar requirement is mentioned by Zermelo (1929) and David (1987, Property 3).

Remark 1

Regarding the generalized row sum method, Chebotarev (1994, Property 3) introduces a more general axiom called agreement: if \((N,R,M) \in \mathcal {R}^n\) is a round-robin ranking problem, then \(\mathbf {x}(\varepsilon )(N,R,M) = \mathbf {s}(N,R,M)\).

Lemma 3

Row sum, generalized row sum and least squares methods satisfy SCC.

Proof

For generalized row sum, see Remark 1. In the case of least squares the proof is given by González-Daz et al. (2014, Proposition 5.3). \(\square \)

Further properties of the scoring procedures are discussed by González-Díaz et al. (2014) and Csató (2014).

2.4 Interpretation of Swiss system chess team tournaments as a ranking problem

In order to use the scoring procedures presented above, the chess tournament should be formulated as a ranking problem:

  • Set of objects N consists of the teams of the competition;

  • Matches matrix M is given by \(m_{ij} = 1\) if teams i and j have played against each other and \(m_{ij} = 0\) otherwise.

For the sake of simplicity it is assumed that n is even, thus all teams play exactly c matches (there are no byes).Footnote 4

The results matrix should be skew-symmetric. It excludes the incorporation of some individual competitions where a win results in three points and a draw gives one point since then a win and a loss is not equal to two draws. Furthermore, the model is not able to reflect that the first-mover with white have an inherent advantage in the game. In order to eliminate the role of colour allocation, only team tournaments are discussed.

First two extreme possibilities are suggested for the choice of results matrix.

Notation 1

\(MP_{ij}\) and \(BP_{ij}\) are the numbers of match points and board points of team i against team j, respectively.

\(\mathbf {mp}\) and \(\mathbf {gp}\) are the vectors of match points and board points, respectively.

Definition 7

Match points ranking: The ranking derived from the vector \(\mathbf {mp}\).

Definition 8

Board points ranking: The ranking derived from the vector \(\mathbf {bp}\).

Note that match and board points rankings are usually not linear orders in a Swiss system tournament. Ties are broken according to the rules given by the official ranking.

Definition 9

Match points based results matrix: Results matrix \(R^{MP}\) is based on match points if \(r_{ij}^{MP} = MP_{ij}-1\) for all \(i,j \in N\).

Definition 10

Board points based results matrix: Results matrix \(R^{BP}\) is based on board points if \(r_{ij}^{BP} = BP_{ij}-t\) for all \(i,j \in N\).

Note that match and board points based results matrices are skew-symmetric (a match between two teams is played on 2t boards). The two concepts can be integrated.

Definition 11

Generalized results matrix: Results matrix \(R^{P}(\lambda )\) is generalized if \(r_{ij}^{P}(\lambda ) = (1-\lambda ) \left( MP_{ij}-1 \right) + \lambda \left( BP_{ij}-t \right) / t\) for all \(i,j \in N\) such that \(\lambda \in \left[ 0,1 \right] \).

Lemma 4

\(R^{P}(\lambda = 0) = R^{MP}\) and \(R^{P}(\lambda = 1) = R^{BP}\).

The row sum method is closely related to the match and board points rankings.

Lemma 5

Row sum method applied on the match points based results matrix is equivalent to the match points ranking: \(s_i(R^{MP}) \ge s_i(R^{MP}) \iff mp_i \ge mp_j\).

Proof

\(d_i = c\) for all \(i \in N\), hence \(\mathbf {s}(N,R^{MP},M) = \mathbf {mp} - c \mathbf {e}\). \(\square \)

Lemma 6

Row sum method applied on the board points based results matrix is equivalent to the board points ranking: \(s_i(R^{BP}) \ge s_i(R^{BP}) \iff bp_i \ge bp_j\).

Proof

\(d_i = c\) for all \(i \in N\), hence \(\mathbf {s}(N,R^{BP},M) = \mathbf {bp} - ct \mathbf {e}\). \(\square \)

A crucial argument for the application of paired comparison-based ranking methodology is provided by the following result.

Theorem 1

Let \((N,R,M) \in \mathcal {R}^n\) be a round-robin ranking problem. Then:

  • Generalized row sum and least squares methods applied on the match points based results matrix are equivalent to the match points ranking: \(x_i(\varepsilon )(R^{MP}) \ge x_i(\varepsilon )(R^{MP}) \iff q_i(R^{MP}) \ge q_j(R^{MP}) \iff mp_i \ge mp_j\).

  • Generalized row sum and least squares methods applied on the board points based results matrix are equivalent to the board points ranking: \(x_i(\varepsilon )(R^{BP}) \ge x_i(\varepsilon )(R^{BP}) \iff q_i(R^{BP}) \ge q_j(R^{BP}) \iff bp_i \ge bp_j\).

Proof

In the case of round-robin problems, generalized row sum and least squares are equivalent to the row sum method due to axiom SCC (Lemma 3), hence Lemmata 5 and 6 provide the statement. \(\square \)

Generalized row sum and least squares methods take the opponents of each team into account. Due to Theorem 1, they result in the official ranking without tie-breaking rules in the ideal round-robin case. When the official ranking is based on match points, the transformation \(R^{MP}\) is recommended. Generalized results matrix with a small (i.e. close to 0) parameter \(\lambda \) gives a similar outcome but it reflects the number of board points, the magnitude of wins and losses. This effect becomes more significant as \(\lambda \) increases. \(R^{BP}\) extends the board points ranking to Swiss system competitions.

Proposition 4

Let \((N,R,M) \in \mathcal {R}^n\) be a ranking problem, and \(k \in \left( 0,1 \right] \). Generalized row sum and least squares methods give the same ranking if they are applied on \(R^{MP}\) and \(k R^{MP}\), on \(R^{BP}\) and \(k R^{BP}\) as well as on \(R^{P}(\lambda )\) and \(k R^{P}(\lambda )\).

Proof

It is the consequence of property SI (Lemma 1). \(\square \)

Due to Proposition 4, there exists only one ranking on the basis of match points if wins are more valuable than losses and draws correspond to an indifference relation. Analogously, there is a unique ranking based on board points. Without scale invariance, the ranking may depend on the results matrix chosen such as wins are represented by \(r_{ij} = 0.5\) or \(r_{ij} = 1\), for example.

Generalized row sum and least squares methods use all information of the tournament (about the opponents, opponents of opponents and so on) to break the ties. Consequently, it is very unlikely that teams remain tied, unless they have exactly the same opponents (and in such a case it seems reasonable not to break the tie). The elimination of arbitrary tie-breaking rules is a substantial advantage over official rankings.

3 Application: European chess team championships

In the following section, the theoretical model suggested in Sect. 3 will be scrutinized in practice.

3.1 Examples and implementation

The method proposed in Sect. 3 is illustrated with an analysis of two chess team tournaments:

In both tournaments the number of competing teams was \(n = 38\), playing on \(2t = 4\) tables during \(c = 9\) rounds. Results are known for about one quarter of possible pairs, \(9 \times 19 = 171\) from \(n(n-1)/2 = 703\).Footnote 5

The official ranking was based on the number of match points in both cases but tie-breaking rules were different. In the 2013 competition application of the first tie-breaking rule (Olympiad–Sonneborn–Berger points) was enough, while in 2011 two tie-breaking rules (board points and Buchholz points—aggregated board points of the opponents) should be used in some cases.

Fig. 1
figure 1

Distribution of (173) match results, ETCC 2013

Distribution of match results for ETCC 2013 is drawn in Fig. 1. Minimal victory (2.5:1.5) is the mode, so incorporating board points will not influence the rankings much.

There are two exogenous rankings called Official according to the tournament rules and Start based on Élő points of players, reflecting the past performance of team members. Further 12 rankings have been calculated from the ranking problem representation. Four results matrices have been considered: \(R^{MP}\), \(R^{MB} = R^P(1/4) = 3/4 \, R^{MP} + 1/4 \, R^{BP}\), \(R^{BM} = R^P(2/3) = 1/3 \, R^{MP} + 2/3 \, R^{BP}\) and \(R^{BP}\). They were plugged into three methods, least squares (LS) and generalized row sum with \(\varepsilon _1 = 1/324\) (\(GRS_1\)) and \(\varepsilon _2 = 1/6\) (\(GRS_2\)). Note that \(\varepsilon _1\) is smaller and \(\varepsilon _2\) is larger than the reasonable upper bound of 1 / 36.

Existence of a unique least squares solution requires connectedness of the comparison multigraph (Proposition 2), which is provided after the third round. Rankings in the first two rounds are highly unreliable, therefore they were eliminated. From the third round all methods give one, thus \(7 \times 13 + 1 = 92\) rankings will be analysed as Start remains unchanged.Footnote 6

Start and Official rankings are strict, that is, they do not allow for ties by definition. It can be checked that the other rankings also give a linear order of teams in all cases.

Notation 2

The 14 final rankings are denoted by Start, Official; \(GRS_1(R^{MP})\), \(GRS_1(R^{MB})\), \(GRS_1(R^{BM})\), \(GRS_1(R^{BP})\); \(GRS_2(R^{MP})\), \(GRS_2(R^{MB})\), \(GRS_2(R^{BM})\), \(GRS_2(R^{BP})\); and \(LS(R^{MP})\), \(LS(R^{MB})\), \(LS(R^{BM})\), \(LS(R^{BP})\). In the figures they are abbreviated by Start, Off; G1, G2, G3, G4; S1, S2, S3, S4; and L1, L2, L3, L4, respectively.

3.2 Visualisation of the rankings

For the comparison of final rankings their distances have been calculated according to the well-known Kemeny distance (Kemeny 1959) and its weighted version proposed by Can (2014). Both distances are defined on the domain of strict rankings, i.e. ties are not allowed. Our rankings satisfy this condition.

Kemeny distance is the number of pairs of alternatives ranked oppositely in the two rankings examined.

Example 1

The Kemeny distance of \(a \succ b \succ c\) and \(b \succ a \succ c\) is 1, because they only disagree on how to order a and b.

The Kemeny distance of \(a \succ b \succ c\) and \(a \succ c \succ b\) is 1 because of the sole disagreement on how to order b and c.

Kemeny distance was characterized by Kemeny and Snell (1962), however, Can and Storcken (2013) achieved the same result without one condition. Can and Storcken (2013) also provides an extensive overview about the origin of this measure.

According to Example 1, the dissimilarity between \(a \succ b \succ c\) and \(b \succ a \succ c\) and between \(a \succ b \succ c\) and \(a \succ c \succ b\) by the Kemeny distance is identical. However, in our chess example a disagreement at the top of the rankings may be more significant than a disagreement at the bottom since the audience is usually interested in the first three, five or ten places but people are not bothered much whether a team is the 31st or 34th.

For this purpose, Can (2014) proposes some functions on strict rankings in the spirit of Kemeny metric. They are respectful to the number of swaps but allow for variation in the treatment of different pairs of disagreements by weighting them according to an exogenous weight vector. It has some price since the calculation will depend on the order of swaps between the two rankings. Can (2012, Theorem 1) shows that only the path-minimizing function satisfies the triangular inequality condition for all possible weight vectors. Finding the path-minimizing metric is not trivial, it is equivalent to solving a short-path problem in general, but the solution is known if the weights are monotonically decreasing (increasing) from the upper parts of a ranking to the lower parts.Footnote 7

These results have inspired us to choose a monotonically decreasing weight vector meaning that swaps in the first places are more important than changes at the bottom of the rankings.

Definition 12

The weight vector of our weighted distance is given by \(\omega _i = 1/i\) for all \(i = 1,2, \dots ,n-1\).

Example 2

The weighted distance of \(a \succ b \succ c\) and \(b \succ a \succ c\) is 1 (a swap at the first position).

The weighted distance of \(a \succ b \succ c\) and \(a \succ c \succ b\) is 1 / 2 (a swap at the second position).

Lemma 7

The maximum of Kemeny distance is \(n(n-1)/2 (= 703)\) and the maximum of weighted distance is \(n-1 (= 37)\) if and only if the two rankings are entirely opposite.

Proof

The maximal number of swaps between two rankings is \(n(n-1)/2\) in the case of two entirely opposite rankings, which is also their Kemeny distance.

Take the ranking \(a_1 \succ a_2 \succ \dots \succ a_n\). The winners’ decomposition (Can 2014, Example 2) first permutes \(a_n\) to the first place, which involves one swap in each position from the first to the \((n-1)\)th, contributing by \(1 + 1/2 + \dots 1/(n-1)\) to the weighted distance. Thereafter, it permutes \(a_{n-1}\) to the second place, which involves one swap in each position from the second to the \((n-1)\)th, contributing by \(1/2 + \dots 1/(n-1)\) to the weighted distance, and so on. Thus the total weighted distance of two entirely opposite rankings is \(1 \times 1 + 2 \times 1/2 + 3 \times 1/3 + \dots + (n-1) \times 1/(n-1) = n-1\). \(\square \)

We do not know about any other application of Can (2014)’s novel method.

Table 1 Distances of rankings, ETCC 2011

Distances of rankings of ETCC 2011 is presented in Table 1. All Kemeny distances are significantly smaller than its maximum of 703 for entirely opposite rankings. Largest values usually occur in comparison with Start since the latter is not influenced by the results. However, rankings based on match points and board points are also relatively far from each other. Official coincides with \(GRS_1(R^{MB})\).

Weighted distances are presented in Table A.1.b. Its maximum is 37. Ratio of Kemeny and weighted distances are between 8.73 and 17.44 for ETCC 2011, and between 5.81 and 18.73 for ETCC 2013. In the second case accounting for swaps’ positions has a larger effect but the discrepancy of the two distances remains smaller than expected, that is, variations are more or less equally distributed along the rankings.

The ranking from \(GRS_1(R^{MP})\) means a kind of tie-breaking rule for match points both in ETCC 2011 and ETCC 2013: generalized row sum gives the match points ranking for \(\varepsilon = 0\), while a small increase in the parameter breaks ties among teams according to the strength of their opponents. The official ranking also aims to eliminate ties, although it uses a different approach.

The pairwise distances of 14 rankings can be plotted in a 13-dimensional space without loss of information but it still seems to be unmanageable. Therefore multidimensional scaling (Kruskal and Wish 1978) has been applied, similarly to Csató (2013). It is a statistical method in information visualization for exploring similarities or dissimilarities in data, a textbook application of MDS is to draw cities on a map from the matrix consisting of their air distances.

Kemeny and weighted distances are measured on a ratio scale due to the existence of a natural minimum and maximum. Then discrepancies of the reduced dimensional map are linear functions of the original distances. Both Stress and RSQ tests for validity strengthen that two dimensions are sufficient to plot the 14 rankings, however, one dimension is too restrictive. The method produces a map where only the position of objects count, more similar rankings are closer to each other. Only the distances of points representing the rankings yield information, the meaning of the axes remains obscure.

MDS maps reinforce the conjecture from Table 1 that Start is far away from all other rankings (see Csató 2016a, Fig. 2). Thus Start ranking is omitted from further analysis, which improves the mapping, too.

Fig. 2
figure 2

MDS maps of rankings, ETCC 2013 (see Notation 2 for explanation) a Kemeny distance, without Start. b Weighted distance, without Start

There is not much difference between the four charts (ETCC 2011 vs. 2013, Kemeny vs. weighted distances). MDS procedures of ETCC 2013 and Kemeny distances have more favourable validity measures than MDS procedures of ETCC 2011 and weighted distances. They reveal the following results shown by Fig. 2:

  1. 1.

    Start significantly differs from all other rankings since it does not depend on the results of the tournament;

  2. 2.

    Generalized row sum rankings (with low \(\lambda \)) are more similar to the official one than least squares;

  3. 3.

    The order of results matrices by the variability of rankings for a given scoring method is \(R^{MP}< R^{MB}< R^{BM} < R^{BP}\), a greater role of match points (smaller \(\lambda \)) stabilizes the rankings;

  4. 4.

    The order of scoring procedures by the variability of rankings for a given results matrix is \(LS< GRS_2 < GRS_1\), a greater influence of opponents’ results stabilizes the rankings;

  5. 5.

    The effect of tie-breaking rule for match points is not negligible (Off and G1 are not very close to each other).

On the basis of these observations, the application of least squares with a generalized results matrix favouring match points (a low \(\lambda \), for example, 1 / 4 as in \(R^{MB}\)) is proposed for ranking in Swiss system chess team tournaments. It gives an incentive to score more board points but still prefers match points.

3.3 Analysis of a ranking

The decomposition of the least squares rating (Csató 2015) offers another approach to compare the rankings. The ranking problem is balanced and the comparison multigraph is regular, therefore Proposition 3 can be applied. In the zeroth step (\(\mathbf {q}^{(0)}\)) it gives the match points ranking, the official ranking without the application of tie-breaking rules. After that, the iterated ratings reflect the strength of opponents, opponents of opponents and so on by accounting for their average match points as \(\mathfrak {d}I-L = M\). A ranking equivalent to \(\mathbf {q}(R^{MP})\) is obtained after the seventh (from \(\mathbf {q}^{(7)}(R^{MP})\)) and after the twelfth step (from \(\mathbf {q}^{(12)}(R^{MP})\)) in the case of ETCC 2011 and ETCC 2013, respectively.

Table 2 shows the changes of teams’ positions in each step of the decomposition of the ranking \(LS(R^{MP})\) for ETCC 2013. In the second column (\(\mathbf {q}^{(0)}\)), ties are broken according to the official rules, so it coincides with the official ranking. In subsequent steps there are no ties. The last change is a swap of Turkey and Montenegro in the twelfth step of the iteration. The least squares method is far from being only a tie-breaking rule for match points (contrary to generalized row sum with \(\varepsilon _1 = 1/324\)), a team may overtake another one despite its disadvantage of two match points.

Correction according to opponents’ strength results in seven positions improvement for Slovenia together with a four positions decline for Romania and six for Netherlands. Hence Slovenia overtakes Netherlands despite it has a two match points disadvantage.Footnote 8 Subsequent steps of the iteration usually result in a similar direction of swaps, however, in a more moderated extent. A notable exception is Romania, regaining some positions due to indirect opponents. The monotonic decrease of absolute adjustments is violated only by Lithuania.

Table 2 Positional changes in the decomposition of \(LS(R^{MP})\), ETCC 2013 Rank improvements and declines between the rankings from the corresponding \({\mathbf{q}}^{(k)}\) rating are indicated by the arrows
figure b
and
figure c
, respectively. A number in brackets represent the same number of
figure d
or
figure e
arrows. Lack of change is indicated by –

There are two changes among the top six teams. France becomes the winner of the tournament after \(k=2\) instead of Azerbaijan. It can be debated since the latter team has no loss, however, the schedule of France was more difficult. The swap of Russia and Armenia may be explained by the advance on an outer circle of the former team (i.e. Russia had a worse performance than Armenia during the first rounds of the tournament).

Table 3 Positional changes in the decomposition of \(LS(R^{MP})\), ETCC 2011 Rank improvements and declines between the rankings from the corresponding \(\mathbf {q}^{(k)}\) rating are indicated by the arrows
figure f
and
figure g
, respectively. A number in brackets represents the same number of
figure h
or
figure i
arrows. Lack of change is indicated by–

Imperfection of the official ranking is further highlighted by ETCC 2011, for which Table 3 contains the positional changes according to the iterative decomposition of \(LS(R^{MP})\). Here France scored three wins and three draws in the first six rounds but it has been defeated three times after that, presenting an extreme example of advance on an inner circle. Thus France had a more challenging schedule compared to teams with the same number of match points, reflected in the significant adjustment by the least squares method.

On the other side, Serbia loses nine, and Georgia loses 14 positions. They had luck with the opponents, for example, Georgia had not played against a better team according to the official ranking, which is quite strange for a team at the 13th place. Consequently, both Serbia and Georgia significantly benefit from decreasing \(\varepsilon \) or increasing the role of board points.

3.4 Assessment of the rankings

The 14 rankings are evaluated from three aspects:

  • Predictive performance: ability to forecast the outcomes of future matches;

  • Retrodictive performance: ability to match the results of contests already played;

  • Robustness: stability between subsequent rounds.

The first two are standard aspects for the classification of mathematical ranking models (Pasteur 2010). However, for the ranking of a Swiss system tournament, the second is much more important: the aim is to get a meaningful ranking on the basis of matches already played, shown by in-sample fit.

The third measure, stability, seems to be important because of (at least) two causes. First, both the participants and the audience may dislike if the rankings are volatile. Naturally, extreme stability is not favourable, too, but it is usually not a problem in a Swiss system tournament. The second argument for robustness may be that the number of rounds is often determined arbitrarily, for instance, it was 13 in the 2006 and 11 in the 2013 chess olympiads with 148 and 146 teams, respectively.

Predictive and retrodictive performances are measured by the number of match and board points scored by an underdog against a better team. It does not take into account the difference of positions, only its sign.

Prediction power has a meaning only after the third round, when the comparison multigraph becomes connected. Start has the most favourable forecasting performance for the remaining matches, especially in the first rounds, that is, match outcomes are determined by teams’ ability rather than by their results in the competition (Csató 2016a, Figure A.1). There is also no difference among the methods in prediction power if only the next round is scrutinized (Csató 2016a, Figure A.2).

The fact that Start ranking is the best for forecasting match results reflects the insignificance of prediction precision for a Swiss system tournament ranking: after all, what is the meaning to organize a contest if its final result is determined by teams’ ability?

Retrodictive performance has a meaning after the third round, too, however, it is also defined after the last round when prediction power cannot be interpreted. Least squares method seems to be the best from this point of view, despite its statistical significance remains dubious (Csató 2016a, Figure A.3). Generalized row sum is placed between the least squares and official rankings. Choice of the results matrix and the tournament does not influence these findings.

Fig. 3
figure 3

Robustness (distance between subsequent rounds), ETCC 2011 a Kemeny distance, results matrix \(R^{MP}\). b Weighted distance, results matrix \(R^{MP}\)

Stability is defined as the distance of rankings in subsequent rounds. It has no meaning for Start but can be calculated for all other rankings from the third round. Figure 3 illustrates the robustness of some rankings in ETCC 2011. Variability does not decrease monotonically, but a solid decline is observed as the actual round gives relatively fewer and fewer information. Ranking \(LS(R^{MP})\) is the most robust according to both Kemeny and weighted distances, followed by \(GRS_2(R^{MP})\), then \(GRS_1(R^{MP})\) and Official: rankings become less volatile by taking into account the performance of opponents. Difference of absolute values seems to be more significant in the case of weighted distance, the least squares method is robust especially in the first, critical places. The order of variability \(LS< GRS_2 < GRS_1\) is valid for all other result matrices, however, \(GRS_1\) is sometimes more volatile than the official ranking.

In the case of ETCC 2013, these conclusions are more uncertain but least squares remains the most stable with the exception of first rounds (Csató 2016a, Figure A.4). Readers interested in a somewhat more detailed analysis of the two tournaments are encouraged to study Csató (2016a).

To summarize, the least squares method gives the most robust and legitimate ranking. Therefore, its application is also recommended if the organizers want to mitigate the effects of the (predetermined) number of rounds on the ranking.

4 Discussion

The paper has given an axiomatic analysis of ranking in Swiss system chess team tournaments. The framework is flexible with respect to the role of the opponents (parameter \(\varepsilon \)) and the influence of match and board points (choice of the results matrix). The suggested methods are close to the concept of official rankings (they coincide in the case of round-robin tournaments), can be calculated iteratively or by solving a system of linear equations and have a clear interpretation on the comparison multigraph. They also do not call for arbitrary tie-breaking rules.

The model is tested on the results of the 2011 and 2013 European Team Chess Championship open tournaments, which supports the application of least squares method due to its relative insensitivity to the choice between match and board points, retrodictive accuracy and stability. There is an opportunity to take into account the number of board points scored by using a generalized results matrix favouring match points (small \(\lambda \) close to zero). The findings confirm that the official rankings have significant failures, therefore recursive methods, similar to generalized row sum and least squares, are worth to consider for ranking purposes.

Naturally, the framework may have some disadvantages (Brozos-Vázquez et al. 2010): a computer is needed in order to calculate the ranking of the tournament, and it will be difficult for the players to verify and understand the whole procedure. However, we agree with Forlano (2011) that ’The fact that players are not able to foresee the final standing should not be considered a disadvantage but a way to force the players to play each round as the decisive one.’ as well as ’The fact that the the players cannot replicate the method manually should be seen of no significance.’ While the least squares method is more complicated than usual tie-breaking rules, its simple graph interpretation (Csató 2015) and its similarity to an ’infinite Buchholz’ may help in the understanding.

Anyway, there usually exists a trade-off between simplicity and other favourable properties (sample fit, robustness), and the use of more developed methods is worth to consider in the case of Swiss system tournaments in order to avoid anomalies of the ranking,Footnote 9 such as when a Hungarian commentator speaks about ’the curse of the Swiss system’.Footnote 10 It is not necessarily the mistake of Swiss system rather a failure of the official ranking, which can be improved significantly by accounting for the strength of opponents.

Nevertheless, the choice between simplicity and more plausible rankings is not a modelling issue. An alternative may be to use these methods only for tie-breaking purposes.

There are some obvious areas of future research. In the analysis several complications observed have been neglected like matches played with black or white (an unavoidable problem in individual tournaments) or different number of matches due to byes or unplayed games. The choice of parameter \(\varepsilon \) also requires further investigation. Our findings can be strengthened or falsified by the examination of other competitions and simulations of Swiss system tournaments.

Finally, two possible uses of the proposed ranking method are worth to mention. First, it can be incorporated into the pairing algorithm, resulting in a more balanced schedules. Second, extensive analysis of the stability of a ranking between subsequent rounds may contribute to the choice of the number of rounds, which can be made endogenous as a function of the number of participants and other restrictions.