Skip to main content
Log in

The weighted-egalitarian Shapley values

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

Abstract

We propose a new class of allocation rules for cooperative games with transferable utility (TU-games), weighted-egalitarian Shapley values, where each rule in this class takes into account each player’s contributions and heterogeneity among players to determine each player’s allocation. We provide an axiomatic foundation for the rules with a given weight profile (i.e., exogenous weights) and the class of rules (i.e., endogenous weights). The axiomatization results illustrate the differences among our class of rules, the Shapley value, the egalitarian Shapley values, and the weighted Shapley values.

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. For recent studies, see Casajus (2011; 2014) and Casajus and Yokote (2017a).

  2. Abe and Nakada (2017) extends a monotonic redistribution rule (Casajus 2015, 2016; Casajus and Yokote 2017b) to exhibit agents’ heterogeneity. See Roth (1979), Kalai and Samet (1987), Chun (1988; 1991), and Nowak and Radzik (1995) for other examples.

  3. Weak monotonicity without weights was introduced by van den Brink et al. (2013) to characterize egalitarian Shapley values. We discuss this topic later.

  4. For other characterizations of the weighted Shapley values, see Kalai and Samet (1987), Chun (1991), Hart and Mas-Colell (1989), and Yokote (2014).

  5. In Sect. 5, we discuss the future direction to unify these two classes.

  6. This is also known as the equal treatment property.

  7. Note that (WDMSP) is weaker than (SYM). It is also a weaker version of differential marginality defined by Casajus (2010, 2011). Formally, differential marginality is defined as follows: for any \(i,j \in N\) and \(v, v' \in \mathcal {G}\), if \(v(S \cup \{i\})-v(S \cup \{j\}) \)\(= v'(S \cup \{i\})-v'(S \cup \{j\})\) for all \(S \subseteq N{\setminus }\{i,j\}\), then \(f_i(v)-f_j(v)=f_i(v')-f_j(v')\).

  8. A game v is monotonic if \(v(T) \ge v(S)\) for any \(S \subseteq T\).

  9. This is a corollary implied by their more general result. They consider more general weight systems.

  10. If \(f_i(\mathbf 0 )=\delta <0\) for all \(i \in N\), it contradicts to (E). Hence, supposing that \(f_i(\mathbf 0 )=\delta >0\) for some i, we consider \(v \in \mathcal {G}_N\) such that \(v(N)=\varepsilon \in (0,\delta )\) and \(v(S)=0\) for \(S\ne N\). By (M\(^{-}\)), \(f_i(v)\ge f_i(\mathbf 0 )=\delta \). Also, by (SP), \(f_i(v)>0\) for all \(i \in N\). This contradicts (E) because \(v(N)=\varepsilon \in (0,\delta )\).

References

  • Abe T, Nakada S (2017) Monotonic redistribution: reconciling performance-based allocation and weighted division. Int Game Theory Rev 19:175022

    Article  Google Scholar 

  • Béal S, Casajus A, Huettner F, Rémila E, Solal P (2016) Characterizations of weighted and equal division values. Theory Decis 80:649–667

    Article  Google Scholar 

  • van den Brink R, Funaki Y, Ju Y (2013) Reconciling marginalism with egalitarianism: consistency, monotonicity, and implementation of egalitarian Shapley values. Soc Choice Welf 40:693–714

    Article  Google Scholar 

  • Casajus A (2010) Another characterization of the Owen value without the additivity axiom. Theory Decis 69:523–536

    Article  Google Scholar 

  • Casajus A (2011) Differential marginality, van den Brink fairness, and the Shapley value. Theory Decis 71:163–174

    Article  Google Scholar 

  • Casajus A (2014) The Shapley value without efficiency and additivity. Math Soc Sci 68:1–4

    Article  Google Scholar 

  • Casajus A (2015) Monotonic redistribution of performance-based allocations: a case for proportional taxation. Theor Econ 10:887–892

    Article  Google Scholar 

  • Casajus A (2016) Differentially monotonic redistribution of income. Econ Lett 141:112–115

    Article  Google Scholar 

  • Casajus A, Huettner F (2013) Null players, solidarity, and the egalitarian Shapley values. J Math Econ 49:58–61

    Article  Google Scholar 

  • Casajus A, Huettner F (2014) Weakly monotonic solutions for cooperative games. J Econ Theory 154:162–172

    Article  Google Scholar 

  • Casajus A, Yokote K (2017a) Weak differential marginality and the Shapley value. J Econ Theory 167:274–284

    Article  Google Scholar 

  • Casajus A, Yokote K (2017b) Weak differential monotonicity, flat tax, and basic income. Econ Lett 151:100–103

    Article  Google Scholar 

  • Chun Y (1988) The proportional solution for rights problems. Math Soc Sci 15:231–246

    Article  Google Scholar 

  • Chun Y (1989) A new axiomatization of the Shapley value. Games Econ Behav 1:119–130

    Article  Google Scholar 

  • Chun Y (1991) On the symmetric and weighted Shapley values. Int J Game Theory 20:183–190

    Article  Google Scholar 

  • Hart S, Mas-Colell A (1989) Potential, value, and consistency. Econometrica 57:589–614

    Article  Google Scholar 

  • Joosten R, Peters H, Thuijsman F (1994) Socially acceptable values for transferable utility games. Report M94-03, Maastricht University

  • Joosten R (1996) Dynamics, equilibria and values. Dissertation, Maastricht University

  • Joosten R (2016) More on linear-potential values and extending the ‘Shapley family’ for TU-games, presented in Stony Brook, see the conference website http://www.gtcenter.org/?page=Archive/2016/ConfTalks.html

  • Kalai E, Samet D (1987) On weighted Shapley values. Int J Game Theory 16:205–222

    Article  Google Scholar 

  • Nowak A, Radzik T (1994) A solidarity value for n-person transferable utility games. Int J Game Theory 23:43–48

    Article  Google Scholar 

  • Nowak A, Radzik T (1995) On axiomatizations of the weighted Shapley values. Games Econ Behav 8:389–405

    Article  Google Scholar 

  • Roth A (1979) Proportional solutions to the bargaining problem. Econometrica 47:775–778

    Article  Google Scholar 

  • Shapley L (1953a) Additive and non-additive set functions. PhD Thesis, Department of Mathematics, Princeton University

  • Shapley L (1953b) A value for n-person games. In: Kuhn HW, Tucker AW (eds) Contributions to the theory of games II (Annals of Mathematics Studies 28). Princeton University Press, Princeton, pp 307–317

    Google Scholar 

  • Shapley L, Shubik M (1969) On the core of an economic system with externalities. Am Econ Rev 59:678–684

    Google Scholar 

  • Young P (1985) Monotonic solutions of cooperative games. Int J Game Theory 14:65–72

    Article  Google Scholar 

  • Yokote K (2015) Weak addition invariance and axiomatization of the weighted Shapley value. Int J Game Theory 44:275–293

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Satoshi Nakada.

Additional information

This paper was previously circulated as “Priority-regarding Shapley values”. The authors are grateful to an associate editor and anonymous referees for many helpful suggestions. We thank Yukihiko Funaki, Hideshi Itoh, Jingyi Xue, Shinsuke Kambe, Takashi Ui and Koji Yokote for constructive comments. We also thank seminar participants in the Contract Theory Workshop East and Graduate Summer Workshop on Game Theory 2016 at Soul National University. Abe and Nakada acknowledge financial support from the Japan Society for Promotion of Science (JSPS). All remaining errors are our own.

Appendices

Appendix A: Proof of Theorem 2

Let \(\mathcal {G}^{c} \subseteq \mathcal {G}\) denote the set of games such that \(v(N)=c\). Also, for any player \(i\in N\) and \(c\in \mathbb {R}\), let \(\mathcal {G}^{c,i}\) denote the set of games v such that \(v(N)=c\) and i is null player. Note that \(\mathcal {G}^{c,i} \subseteq \mathcal {G}^{c}\) for all \(i\in N\) and \(c\in \mathbb {R}\). Let \(\varDelta _i(v)=(v(S\cup \{i\})-v(S))_{S\subseteq N{\setminus } \{i\}} \in \mathbb {R}^{2^{(N-1)}}\) be a vector of marginal contributions of i in v. Therefore player \(i\in N\) is a null player in v if \(\varDelta _i(v)= \mathbf{0}\). Let \(\varLambda ^i\) be the set of all vectors of marginal contribution of i: \(\varLambda ^i=\{ \varDelta _i(v) | v\in \mathcal {G}\}\).

For each \(x \in \mathbb {R}^N\), let \(m_x \in \mathcal {G}\) be an additive game, \(m_x(S)=\sum _{i \in S}x_i\) for all \(S \subseteq N\). Let \(\mathcal {G}^{add}\) be the set of additive games. Since there is a one-to-one correspondence between \(x \in \mathbb {R}^N\) and an additive game \(m_x\), we can identify \(\mathcal {G}^{add}\) with \(\mathbb {R}^N\). Abe and Nakada (2017) provide the following result, which will be useful later.

Theorem A.1

(Abe and Nakada 2017) Let \(n \ne 2\). \(f:\mathcal {G}^{add} \rightarrow \mathbb {R}^n\) satisfies (E), (\(\mathrm {M}^{-}\)), (NY), and (RIN) if and only if there exists some \(\delta \in [0,1]\) and \(w \in \mathcal {W}\) such that \(f_i(x)=\delta \cdot x_i+(1-\delta ) \cdot w_i \cdot \sum _{l \in N}x_l\) for all \(x \in \mathbb {R}^n\) and \(i \in N\).

Now, we offer the proof of Theorem 2. It is clear that the rule satisfies all the axioms. We suppose that a rule \(f: \mathcal {G}\rightarrow \mathbb {R}^N\) satisfies (E), (\(\mathrm {M}^{-}\)), (RIN), (WDMSP), and (NY).

Claim 1

For each \(i \in N\), there exist functions \(\phi _i: \varLambda ^i \times \mathbb {R}\rightarrow \mathbb {R}\) and \(\alpha _i: \mathbb {R}\rightarrow \mathbb {R}\) such that \(f_i(v)=\phi _i(\varDelta _i(v), v(N))+\alpha _i(v(N))\).

We first take any \(c\in \mathbb {R}\). For any \(i\in N\) and \(v \in \mathcal {G}^c\), we have the following equation: for any \(\bar{v}\in \mathcal {G}^c\) such that \(\varDelta _i(v)=\varDelta _i(\bar{v})\),

$$\begin{aligned} f_i(v)\overset{\text {(M}^-)}{=}f_i(\bar{v})\ =:\alpha _i(c,\varDelta _i(v)). \end{aligned}$$
(A.1)

Specifically, we denote

$$\begin{aligned} \alpha _i(c)=\alpha _i(c,\mathbf{0}). \end{aligned}$$
(A.2)

Moreover, for any \(i\in N\) and \(v,v' \in \mathcal {G}^c\), we have

$$\begin{aligned}&f_i(v)-f_i(v') \overset{(A.1)}{=} \alpha _i(c,\varDelta _i(v))-\alpha _i(c,\varDelta _i(v'))\nonumber \\&\quad =: \phi _i(\varDelta _i(v),\varDelta _i(v'),c). \end{aligned}$$
(A.3)

Hence, for any \(i\in N\) and \(v \in \mathcal {G}^c\), we obtain the following equation: for any \(v' \in \mathcal {G}^{c,i}\),

$$\begin{aligned} \phi _i(\varDelta _i(v),\varDelta _i(v'),c) \overset{(A.3)}{=} f_i(v)-f_i(v') \overset{(A.1)}{=} f_i(v)-\alpha _i(c). \end{aligned}$$
(A.4)

Note that \(f_i(v)-\alpha _i(c)\) is independent from \(v' \in \mathcal {G}^{c,i}\). For any \(i\in N\) and \(v \in \mathcal {G}^c\) let

$$\begin{aligned} \phi _i(\varDelta _i(v),c):=f_i(v)-\alpha _i(c). \end{aligned}$$
(A.5)

Hence, for any \(i\in N\) and \(v \in \mathcal {G}\), we obtain

$$\begin{aligned} f_i(v)\overset{(A.5)}{=}\phi _i(\varDelta _i(v),v(N))+\alpha _i(v(N)). \end{aligned}$$
(A.6)

This completes Claim 1.

Claim 2

For any \(i\in N\), and \(c \in \mathbb {R}\), the function \(\phi _i(\cdot , c): \varLambda ^i \rightarrow \mathbb {R}\) satisfies (M) within \(\mathcal {G}^c\): for any \(v,v'\in \mathcal {G}^c\), if \(v(S\cup \{i\})-v(S)\ge v'(S\cup \{i\})-v'(S)\) for any \(S\subset N{\setminus } \{i\}\), then \(\phi (\varDelta _i(v), c)\ge \phi (\varDelta _i(v'), c)\). Moreover, \(\phi _i(\mathbf{0},c)=0\).

Let \(c=v(N)=v'(N)\). We have

$$\begin{aligned} \phi (\varDelta _i(v),c)- \phi (\varDelta _i(v'),c)&= \phi (\varDelta _i(v),c) + \alpha _i(c)- (\phi (\varDelta _i(v'),v'(N))+\alpha _i(c))\\&\overset{\text {C1}}{=} f_i(v) - f_i(v')\\&\overset{\text {(M}^-)}{\ge } 0. \end{aligned}$$

Moreover, for any \(c\in \mathbb {R}\),

$$\begin{aligned} \phi _i(\mathbf{0},c) \overset{(A.5),(A.4)}{=}\phi _i(\mathbf{0},\mathbf{0},c) \overset{(A.3),(A.1)}{=}\alpha _i(c,\mathbf{0})-\alpha _i(c,\mathbf{0})=0. \end{aligned}$$
(A.7)

Claim 3

The function \(\phi \) is symmetric: for any \(v \in \mathcal {G}\) and \(i, j \in N\), if ij is symmetric in v, then \(\phi _i(\varDelta _i(v), v(N))=\phi _j(\varDelta _j(v), v(N))\).

For any \(i,j\in N\) and \(v \in \mathcal {G}\) such that \(v(S \cup \{i\})-v(S) = v(S \cup \{j\})-v(S)\) for all \(S \subseteq N{\setminus }\{i,j\}\), let \(v'=v(N)u_{N \backslash \{i,j\}}\). Then, we have

$$\begin{aligned} \phi _i(\varDelta _i(v), v(N)) \overset{(A.6)}{=} f_i(v)-f_i(v') \overset{{\mathrm{(WDMSP)}}}{=}f_j(v)-f_j(v') \overset{(A.6)}{=} \phi _j(\varDelta _j(v), v(N)). \end{aligned}$$
(A.8)

This completes Claim 3.

Claim 4

The function \(\phi \) satisfies \(\delta \)-efficiency (\(\delta \)-E): there is a \(\delta \in [0,1]\) such that

\(\sum _{i \in N}\phi _i(\varDelta _i(v), v(N))=\delta v(N)\) for any \(v \in \mathcal {G}\).

Let \({\tilde{f}}: \mathcal {G}^{add} \rightarrow \mathbb {R}\) be the restriction of f on \(\mathcal {G}^{add}\). Then, by Theorem A.1, for each \(m_x \in \mathcal {G}^{add}\), we have

$$\begin{aligned} {\tilde{f}}_i(m_x)= & {} \phi _i(\varDelta _i(m_x), \sum _{l \in N}x_l)+\alpha _i\left( \sum _{l \in N}x_l\right) \nonumber \\= & {} \delta \cdot x_i+(1-\delta ) \cdot w_i \cdot \sum _{l \in N}x_l. \end{aligned}$$
(A.9)

for some \(\delta \in [0,1]\) and \(w \in \mathcal {W}\). In particular, for \(x_i=0\), (A.7) implies \(\alpha _i(\sum _{l \in N}x_l)\)\(= (1-\delta ) \cdot w_i \cdot \sum _{l \in N}x_l\). Hence, from Claim 1, it follows that for each \(v \in \mathcal {G}\),

$$\begin{aligned} f_i(v)=\phi _i(\varDelta _i(v), v(N))+(1-\delta ) \cdot w_i \cdot v(N). \end{aligned}$$
(A.10)

Since f satisfies (E), \(\sum _{i \in N}f_i(v)=\sum _{i \in N}\phi _i(\varDelta _i(v), v(N))+(1-\delta ) v(N)=v(N)\), which implies that \(\phi \) satisfies the following property:

$$\begin{aligned} (\delta -\text {E}): \sum _{i \in N}\phi _i(\varDelta _i(v), v(N))=\delta v(N). \end{aligned}$$

This completes Claim 4.

Claim 5

There is a \(\delta \in [0,1]\) and a \(w \in \mathcal {W}\) such that \(f_i(x)=\delta \cdot Sh_i(v)+(1-\delta ) \cdot w_i \cdot v(N)\) for any \(v \in \mathcal {G}\).

Fixing \(c \in \mathbb {R}\), we write \(\phi _i^c(v)=\phi _i(\varDelta _i(v), c)\) for each \(v \in \mathcal {G}^c\). The function \(\phi ^c(v): \mathcal {G}^c \rightarrow \mathbb {R}^N\) satisfies (\(\delta \)-E), (M) and (SYM) within \(\mathcal {G}^c\) by Claims 2, 3 and 4. Hence, by the same argument of Theorem 1, we have

$$\begin{aligned} \phi _i^c(v)=\delta Sh_i(v). \end{aligned}$$

Since \(c \in \mathbb {R}\) is arbitrarily chosen, for any \(v\in \mathcal {G}\), we have

$$\begin{aligned} \phi _i(\varDelta _i(v), v(N))=\phi _i^{v(N)}(v)=\delta Sh_i(v) \end{aligned}$$
(A.11)

Finally, by (A.10) and (A.11), we obtain \(f_i(x)=\delta \cdot Sh_i(v)+(1-\delta ) \cdot w_i \cdot v(N)\), which completes the proof. \(\square \)

Appendix B: Proof of Theorem 3

We first show that (E\(^*\)), (M\(^{-*}\)), (SYM\(^{*}\)) characterizes the egalitarian Shapley value when we fix \(w=(\frac{1}{n},\ldots ,\frac{1}{n})\), which can be useful in later.

Lemma 1

Suppose that \(\mathcal {W}=\{(\frac{1}{n},\ldots ,\frac{1}{n})\}\) and \(n \ne 2\). Then, an allocation rule \(f: \mathcal {G}\times \mathcal {W}\rightarrow \mathbb {R}^N\) satisfies (E\(^*\)), (M\(^{-*}\)), (SYM\(^{*}\)) if and only if it is an egalitarian-Shapley value.

Proof

This follows from the axioms and arguments in Casajus and Huettner (2014) if \(w=(\frac{1}{n},\ldots ,\frac{1}{n})\). \(\square \)

Now, we offer the proof of Theorem 3. It is clear that the rule satisfies all the axioms. We suppose that a rule \(f: \mathcal {G}\rightarrow \mathbb {R}^N\) satisfies (E\(^*\)), (\(\mathrm {M}^{-*}\)), (RIN\(^*\)), (SYM\(^{*}\)), and (FEC\(^*\)). Claim 1 can be thought of as an analog of that of Theorem 2. The differences lie in Claims 2, 3 and 4. In this proof, we first specify the form of the Shapley value, while we first specify the weighted division in Theorem 2.

Claim 1

For each \(i \in N\), there exists functions \(\phi _i(v): \varLambda ^i \times \mathbb {R}\)\( \rightarrow \mathbb {R}\) and \(\alpha _i: \mathcal {W} \times \mathbb {R} \rightarrow \mathbb {R}\) such that \(f_i(v,w)=\phi _i(\varDelta _i(v), v(N))+\alpha _i(w, v(N))\).

We first take any \(c \in \mathbb {R}\). For any \(i \in N\), \(v \in \mathcal {G}\) and \(w \in \mathcal {W}\), we have the following equation: for any \({\bar{v}} \in \mathcal {G}^c\) such that \(\varDelta _i(v)=\varDelta _i({\bar{v}})\),

$$\begin{aligned} f_i(v, w)\overset{(\text {M}^{-*})}{=}f_i({\bar{v}},w)=: \alpha _i(w, c, \varDelta _i(v)). \end{aligned}$$
(B.1)

Specifically, we denote

$$\begin{aligned} \alpha _i(w, c)=\alpha _i(w, c, \mathbf 0 ). \end{aligned}$$
(B.2)

By (FEC\(^*\)), for any \(c\in \mathbb {R}\) and \(i\in N\), there is a function \(\phi ^c_i:\mathcal {G}^c \rightarrow \mathbb {R}\) such that

$$\begin{aligned} \phi ^c_i(v)&=f_i(v,w)-f_i(cu_{N{\setminus } \{i\}},w)\\&\quad \overset{(B.2)}{=}f_i(v,w)-\alpha _i(w, c). \end{aligned}$$

By (B.1), we know that \(\phi ^c_i(v)=\phi ^c_i({\bar{v}})\) if \(v(N)={\bar{v}}(N)=c\) and \(\varDelta _i(v)=\varDelta _i({\bar{v}})\). Hence, we can define \(\phi _i(\varDelta _i(v), c): \varLambda ^i \times \mathbb {R} \rightarrow \mathbb {R}\) as \(\phi _i(\varDelta _i(v), c)=:\phi ^c_i(v)\). Therefore, for any \(i \in N\), \(v \in \mathcal {G}\) and \(w \in \mathcal {W}\), we obtain \(f_i(v,w)=\phi _i(\varDelta _i(v), v(N))\)\(+\alpha _i(w, v(N))\). This completes Claim 1.

Claim 2

For any \(v \in \mathcal {G}\), there exists a \(\delta \in [0,1]\) and \(d^{v(N)}_i \in \mathbb {R}\) such that \(\phi _i(\varDelta _i(v), v(N))=\delta Sh_i(v)+d^{v(N)}_i\).

Let \(w^*=(1/n,\ldots ,1/n) \in \mathcal {W}\), i.e., the equal weight. For any \(c \in \mathbb {R}\) and any \(v \in \mathcal {G}^c\), by Claim 1, we have

$$\begin{aligned} f_i(v, w^*)=\phi _i(\varDelta _i(v), v(N))+\alpha _i(w^*, v(N)), \end{aligned}$$
(B.3)

and, by Lemma 1, there exists \(\delta \in [0,1]\) such that

$$\begin{aligned} f_i(v, w^*)=\delta Sh_i(v)+(1-\delta )\frac{1}{n}c. \end{aligned}$$
(B.4)

Note that \(\delta \) does not depend on \(c \in \mathbb {R}\). For any \(v'\in \mathcal {G}^{c,i}\), we have

$$\begin{aligned}&\phi _i(\varDelta _i(v'), c)+\alpha _i(w^*, c) \overset{(B.3)}{=} f_i(v', w^*) \overset{(B.4)}{=} \delta Sh_i(v')+(1-\delta )\frac{1}{n}c\nonumber \\&\quad =(1-\delta )\frac{1}{n}c. \end{aligned}$$
(B.5)

Note that player i is a null player in game \(v'\in \mathcal {G}^{c,i}\). Hence, for any \(v',v''\in \mathcal {G}^{c,i}\), we have \(\phi _i(\varDelta _i(v'), c)+\alpha _i(w^*, c)\overset{(B.5)}{=}(1-\delta )\frac{1}{n}c\overset{(B.5)}{=}\phi _i(\varDelta _i(v''), c)+\alpha _i(w^*, c)\) and, so, denote \(d^c_i:=\phi _i(\varDelta _i(v'), c)=\phi _i(\varDelta _i(v''), c)\). We obtain

$$\begin{aligned} \alpha _i(w^*, c)\overset{(B.5), d^c_i}{=}(1-\delta )\frac{1}{n}c-d^c_i. \end{aligned}$$
(B.6)

Therefore, for every \(v \in \mathcal {G}^c\), we must have

$$\begin{aligned} \phi _i(\varDelta _i(v), c) \overset{(B.3)(B.4)(B.6)}{=} \delta Sh_i(v)+d^c_i. \end{aligned}$$
(B.7)

Since \(c \in \mathbb {R}\) is arbitrary chosen, we obtain \(\phi _i(\varDelta _i(v'), v(N))=\delta Sh_i(v)+d^{v(N)}_i\) for all \(v \in \mathcal {G}\).

Claim 3

\(\alpha _i(w, v(N))=(1-\delta )\cdot w_iv(N)-d^{v(N)}_i\) for each \(w \in \mathcal {W}\).

Consider any \(w \in \mathcal {W}\) and player \(k^* \in N\) such that \(k^* \in \mathrm{argmin}_{i \in N, w_i>0} w_i\). Note that \(k^*\) is well-defined because \(\sum _{i \in N}w_i=1\) and \(w_i \ge 0\) for any \(i \in N\). By Claim 2, for any player \(i \ne k^*\) and any \(c\in \mathbb {R}\),

$$\begin{aligned} f_k(cu_{\{i\}},w)= \left\{ \begin{array}{ll} \delta c+d^{c}_k+\alpha _k(w, c) &{} \quad \mathrm{if}~~ k=i,\\ d^{c}_k+\psi ^{c}_k(w) &{}\quad \mathrm{otherwise}.\\ \end{array} \right. \end{aligned}$$

Hence, we have

$$\begin{aligned} \sum _{k \in N}(\alpha _k(w, c)+d^{c}_k)\overset{(\text {E}^*)}{=}(1-\delta )c. \end{aligned}$$
(B.8)

Moreover, for any \(i\ne k^*\), j\((j\ne i,\ j\ne k^*)\) and, by considering a game \(cu_{\{j\}}\), we have

$$\begin{aligned} \alpha _i(w, c)+d^{c}_i \overset{(\text {RIN}^*)}{=}\frac{w_i}{w_{k^*}}(\alpha _{k*}(w, c)+d^{c}_{k^*}), \end{aligned}$$
(B.9)

because i and \(k^*\) are null players in \(cu_{\{j\}}\). Therefore, for any \(i\in N\), we have

$$\begin{aligned}&\alpha _{i}(w, c)+d^{c}_i-(1-\delta ) w_ic\\&\quad \overset{(B.9)}{=} w_i\cdot \left[ \frac{1}{w_{k^*}}(\alpha _{k*}(w, c)+d^{c}_{k^*})-(1-\delta )c\right] \\&\quad \overset{(B.8)}{=} w_i\cdot \left[ \frac{1}{w_{k^*}}(\alpha _{k*}(w, c)+d^{c}_{k^*})-\sum _{k \in N}(\alpha _{k}(w, c)+d^{c}_k)\right] \\&\quad \overset{(B.9)}{=} \frac{w_i}{w_{k^*}}\cdot \left[ (\alpha _{k*}(w, c)+d^{c}_{k^*})-\sum _{k \in N}w_k(\alpha _{k*}(w, c)+d^{c}_{k^*})\right] \\&\quad \overset{\sum _{k}w_k=1}{=} \frac{w_i}{w_{k^*}}\cdot \Bigl [(\alpha _{k*}(w, c)+d^{c}_{k^*})-(\alpha _{k*}(w, c)+d^{c}_{k^*})\Bigr ]\\&\quad = 0. \end{aligned}$$

Since \(c \in \mathbb {R}\) is arbitrary chosen, we obtain \(\alpha _{i}(w, v(N))=(1-\delta )\cdot w_iv(N)-d^{v(N)}_i\) for all \(v \in \mathcal {G}\).

Claim 4

For any \(v \in \mathcal {G}\) and \(w\in \mathcal {W}\), there exists a \(\delta \in [0,1]\) such that \(f_i(v,w)=\delta \cdot Sh_i(v) + (1-\delta )\cdot w_i v(N)\).

For any \(v \in \mathcal {G}\) and \(w\in \mathcal {W}\), we have

$$\begin{aligned} f_i(v,w)&\overset{\text {C1}}{=}\phi _i(\varDelta _i(v), v(N))+\alpha _i(w, v(N))\\&\quad \overset{\text {C2}}{=}\delta Sh_i(v)+d^{v(N)}_i +\alpha _i(w, v(N))\\&\quad \overset{\text {C3}}{=}\delta Sh_i(v)+d^{v(N)}_i +(1-\delta )\cdot w_i v(N)-d^{v(N)}_i\\&\quad =\delta Sh_i(v)+(1-\delta )\cdot w_iv(N). \end{aligned}$$

This completes the proof. \(\square \)

Appendix C: Independence of axioms and a counterexample for \(n=2\)

1.1 Independence of axioms for Theorem 2

The independence of the axioms is shown in the examples listed below.

Example C.1

Consider the following function: for any \(i\in N\) and \(v\in \mathcal {G}\),

$$\begin{aligned} f^{\text {E}}_i(v)=0. \end{aligned}$$

This function satisfies all axioms except (E).

Example C.2

Consider the following function: for any \(i\in N\) and \(v\in \mathcal {G}\),

$$\begin{aligned} f^\mathrm{{M}^{-}}_i(v)=2 Sh_i(v)- \frac{v(N)}{n}. \end{aligned}$$

This function satisfies all axioms except (\(\mathrm {M}^{-}\)).

Example C.3

Consider the following function: for any \(i\in N\) and \(v\in \mathcal {G}\),

$$\begin{aligned} f_i^\mathrm{{RIN}}(v)=\delta Sh_i+ (1-\delta ) \frac{i+v(N)^2}{{\bar{N}}+n(v(N))^2}v(N), \end{aligned}$$

where \({\bar{N}}=\sum _{i \in N}i=\frac{n(n-1)}{2}\) and i is the natural number representing player i. This rule satisfies (E), (\(\mathrm {M}^{-}\)), (WMDSP) and (NY) but not (RIN). To check (\(\mathrm {M}^{-}\)), let \(h_i(a)= \frac{i+a^2}{{\bar{N}}+a^2}a=\frac{ia+a^3}{{\bar{N}}+a^2}\). Then, we have \(\frac{d h_i(a)}{da}=\frac{na^4+(3{\bar{N}}-ni)a^2+i{\bar{N}}}{({\bar{N}}+a^2)^2}>0\) for all \(a \in \mathbb {R}\) because \(na^4+i{\bar{N}}>0\) for all i and \(3{\bar{N}}-ni \ge \frac{n(n-3)}{2} \ge 0\).

Example C.4

Consider the following function: for any \(i\in N\) and \(v\in \mathcal {G}\),

$$\begin{aligned} f^{\text {WMDSP}}_i(v)=v(P^\sigma _i\cup \{i\})-v(P^\sigma _i) \end{aligned}$$

where \(P^\sigma _i\) is the set of predecessors of i in \(\sigma \). This function satisfies all the axioms except (WMDSP).

Example C.5

Consider the following function: for any \(i\in N\) and \(v\in \mathcal {G}\),

$$\begin{aligned} f_i^\mathrm{{NY}}(v)= \left\{ \begin{array}{ll} Sh_1+10 &{}\quad \mathrm{if}~~ i=1,\\ Sh_i-\frac{10}{n-1} &{} \quad \mathrm{if}~~i \ne 1.\\ \end{array} \right. \end{aligned}$$

This rule satisfies all the axioms except (NY).

1.2 Independence of axioms for Theorem 3

The independence of the axioms is shown in the examples listed below.

Example C.6

Consider the following function: for any \(i\in N\), \(v \in \mathcal {G}\) and \(w \in \mathcal {W}\),

$$\begin{aligned} f^{\text {E}^*}_i(v,w)=0. \end{aligned}$$

Then, the function satisfies all axioms except (E\(^*\)).

Example C.7

Consider the following function: for any \(i\in N\), \(v \in \mathcal {G}\) and \(w \in \mathcal {W}\),

$$\begin{aligned} f^{\text {M}^{-*}}_i(v,w)=2Sh_i(v) -w_i v(N). \end{aligned}$$

Then, the function satisfies all axioms except (M\(^{-*}\)).

Example C.8

Consider the following function: for any \(i\in N\), \(v \in \mathcal {G}\) and \(w \in \mathcal {W}\),

$$\begin{aligned} f^{\text {RIN}^*}_i(v,w)=\delta \cdot \frac{v(N)}{|N|} + (1-\delta )\cdot w_i v(N). \end{aligned}$$

The function satisfies all axioms except (RIN\(^*\)).

Example C.9

Consider the following function: for any \(i\in N\), \(v \in \mathcal {G}\) and \(w \in \mathcal {W}\),

$$\begin{aligned} f^{\text {SYM}^*}_i(v,w)=\delta \cdot Sh^z_i(v) + (1-\delta ) \cdot w_i v(N), \end{aligned}$$

where \(Sh^z_i(v)\) is the weighted Shapley value for a given weight \(z\in R^N_{++}\). Since anonymity is defined over \(\mathcal {G}\) and \(\mathcal {W}\), the function satisfies all axioms except (SYM\(^*\)).

Example C.10

Consider the following function: for any \(i\in N\), \(v \in \mathcal {G}\) and \(w \in \mathcal {W}\),

$$\begin{aligned} f^{\text {FEC}^*}_i(v,w)=w_{\text {min}} \cdot Sh_i(v) + (1-w_{\text {min}}) w_i v(N), \end{aligned}$$

where \(w_{\text {min}}=\min _{j\in N} w_j\). This function satisfies all axioms except (FEC\(^*\)).

1.3 A counterexample to Theorems 2 and 3 for \(n=2\)

Theorems 2 and 3 fail for \(n=2\). Consider the following allocation rule \(f^{\heartsuit }\) on \(N=\{1,2\}\):

$$\begin{aligned} (f_1^{\heartsuit }(v,w),f_2^{\heartsuit }(v,w))= \left\{ \begin{array}{ll} (Sh_1(v),Sh_2(v)), &{} \quad Sh_1(v)\ge 0\text { and }Sh_2(v)\ge 0, \\ (0,v(N)), &{} \quad Sh_1(v)< 0\text { and }Sh_2(v)>0 \wedge v(N)\ge 0, \\ (v(N),0), &{} \quad Sh_1(v)< 0\text { and }Sh_2(v)>0 \wedge v(N)< 0, \\ (Sh_1(v),Sh_2(v)), &{} \quad Sh_1(v)\le 0\text { and }Sh_2(v)\le 0, \\ (0,v(N)), &{}\quad Sh_1(v)> 0\text { and }Sh_2(v)<0 \wedge v(N)\le 0, \\ (v(N),0), &{}\quad Sh_1(v)> 0\text { and }Sh_2(v)<0 \wedge v(N) > 0, \end{array} \right. \end{aligned}$$

for any \(v\in \mathcal {G}\) and \(w\in \mathcal {W}\).

Note that this function does not depend on w. It is clear that \(f^{\heartsuit }\) satisfies (E\(^*\)) and (M\(^{-*}\)). It satisfies (SYM\(^{*}\)) because if the players 1 and 2 are symmetric in the sense of marginal contribution and have the same weight, they receive \((Sh_1(v),Sh_2(v))\). It satisfies (RIN\(^*\)) because if the players 1 and 2 are null players, the game v is the null game: \(v(12)=v(1)=v(2)=0\). Since \(f^{\heartsuit }\) does not depend on w, it clearly satisfies (FEC\(^*\)). The same argument applies to Theorem 2.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Abe, T., Nakada, S. The weighted-egalitarian Shapley values. Soc Choice Welf 52, 197–213 (2019). https://doi.org/10.1007/s00355-018-1143-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00355-018-1143-3

Navigation