Skip to main content
Log in

The class of ASN-position values

Centrality and consequences of connection failure

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

Abstract

This paper introduces a class of cooperative allocation rules for TU-games with a network structure that account for an agent’s centrality in the network. In contrast to existing centrality measures, this approach analyzes the consequences of tie failure (rather than node failure). Though not directly applied to the centrality issue, tie failures are the idea underlying the Position value (the Shapley value of the arc-game). In contrast, we allow for a whole class of allocation rules as a basis since the Shapley approach might be unreasonable in political applications, especially in networks with incompatibilities. Requiring additivity (A), equal treatment of equals (Symmetry S), and irrelevance of unproductive agents (nullplayer irrelevance N) for the underlying basis, we define the class of ASN-position values and provide axiomatic characterizations. To emphasize the crucial role of links, we refine this class to multiplicative ASN-position values and provide a monotonicity result with respect to links. We apply our approach to the case of the 2001 state parliament elections in Hamburg, Germany to use our allocation rule as a power index and further provide an example where we use our approach as a centrality measure to identify top key nodes. In both applications, the position value taking the Banzhaf value as a basis turns out to be more convincing than the Shapley-based approach.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. The Banzhaf-Power-Index is also known as Banzhaf-Coleman index (cf. Coleman 1971) or Penrose-Banzhaf index (cf. Penrose 1946).

  2. There also exist generalizations in the Jackson and Wolinsky (1996) framework of networks such as Kamijo (2009).

  3. In the setting of cooperative games on networks, links from an individual to itself, that is, links ii are excluded. Therefore, we add \(i\ne j\) to the definition.

  4. Imposing for example Efficiency would uniquely determine the Shapley value. Then, single-valuedness is obsolete and NI could be replaced by either N or NPO.

  5. This axiom is also know as arc anonymity (e.g. van den Nouweland 1993), or link anonymity (e.g. van den Nouweland and Slikker 2012) which might be more convincing terms if no relation to centrality is of interest. We stick to the original term due to Borm et al. (1992) in order to emphasize the centrality issue.

  6. Note that such direct characterizations for general networks may be possible within the framework of Jackson and Wolinsky (1996) if one further demands superadditive coalition functions as in van den Nouweland and Slikker (2012) or Belau (2016). However, the Jackson and Wolinsky (1996) framework explicitly differs from the framework of communication situations we use here such that findings cannot be transferred. Furthermore, superadditivity is quite a restrictive property, especially for political applications and in presence of incompatibilities.

  7. Note that Slikker (2005b) uses reward games which differ from TU-games with a network structure, such that we need to slightly modify the definition.

  8. An axiomatization of the multiplicative normalized Banzhaf value is, for example, proposed by van den Brink and van der Laan (1998).

  9. For completeness, one should note that this coalition broke 2 years later due to personal issues between the leaders of “CDU” and “Schill.” However, these issues were absent at the time of coalition formation, and hence, cannot be taken into account for forecasting issues.

  10. For further details see the appendix, Table 8.

References

  • Banzhaf J (1952) Weighted voting doesn’t work: a mathematical analysis. Rutgers Law Rev 19(2):317–343

    Google Scholar 

  • Barrat A, Barthélémy M, Pastor-Satorras R, Vespignani A (2004) The architecture of complex weighted networks. Proc Natl Acad Sci 101(11):3747–3752

    Article  Google Scholar 

  • Belau J (2013) A new outside option value for networks: the kappa-value. Econ Theory Bull 1(2):175–188

    Article  Google Scholar 

  • Belau J (2016) Outside option values for network games. Math Soc Sci 84:76–86

    Article  Google Scholar 

  • Bergantiños G, Carreras F, García-Jurado I (1993) Cooperation when some players are incompatible. Zeitschrift für Oper Res 38(2):187–201

    Google Scholar 

  • Bonacich P (1972) Factoring and weighting approaches to status scores and clique identification. J Math Sociol 2(1):113–120

    Article  Google Scholar 

  • Borm P, Owen G, Tijs S (1992) On the position value for communication situations. SIAM J Discret Math 5:305–320

    Article  Google Scholar 

  • Coleman JS (1971) Control of collectivities and the power of a collectivity to act. In: Lieberman B (ed) Social Choice. Gordon and Breach, New York, pp 269–300

    Google Scholar 

  • Derks JM, Haller H (1999) Null players out? Linear values for games with variable supports. Int Game Theory Rev 1:301–314

    Article  Google Scholar 

  • Davis M, Maschler M (1965) The kernel of a cooperative game. Naval Res Logist Q 12(3):223–259

    Article  Google Scholar 

  • Dubey P, Shapley L (1979) Mathematical properties of the banzhaf power index. Math Oper Res 4:99–131

    Article  Google Scholar 

  • Edgeworth FY (1881) Mathematical psychics: an essay on the application of mathematics to the moral sciences. Paul, C. K., London

  • Freeman LC (1978) Centrality in social networks: conceptual clarification. Soc Netw 1:215–239

    Article  Google Scholar 

  • Ghintran A (2013) Weighted position values. Math Soc Sci 65:157–163

    Article  Google Scholar 

  • Ghintran A, González-Arangüena E, Manuel C (2012) A probabilistic position value. Ann Oper Res 201(1):183–196

    Article  Google Scholar 

  • Gillies DB (1959) Solutions to general non-zero-sum games. In: Tukcer AW, Luce RD (eds) Contributions to the theory of games, vol 4. Princeton University Press, Princeton

    Google Scholar 

  • Gómez D, González-Arangüena E, Manuel C, Owen G, Del Pozo M, Tejada J (2003) Centrality and power in social networks: a game theoretic approach. Math Soc Sci 146(1):27–54

    Article  Google Scholar 

  • Grofman B, Owen G (1982) A game theoretic approach to measuring degree of centrality in social networks. Soc Netw 4(3):213–224

    Article  Google Scholar 

  • Haeringer G (2006) A new weight scheme for the shapley value. Math Soc Sci 52(1):88–98

    Article  Google Scholar 

  • Jackson M, Wolinsky A (1996) A strategic model of social and economic networks. J Econ Theory 71:44–74

    Article  Google Scholar 

  • Kamijo Y (2009) A linear proportional effort allocation rule. Math Soc Sci 58:341–353

    Article  Google Scholar 

  • Kamijo Y, Kongo T (2009) Weighted position value. GLOPE II working paper series no. 16. Waseda University, Tokyo

  • Meessen R (1988) Communication games. Master’s thesis, Department of Mathematics, University of Nijmegen, The Netherlands

  • Myerson R (1977) Graphs and cooperations in games. Math Oper Res 2:225–229

    Article  Google Scholar 

  • Newman MEJ (2004) Analysis of weighted networks. Phys Rev E 70:056131

    Article  Google Scholar 

  • Opsahl T, Colizza V, Panzarasa P, Ramasco JJ (2008) Prominence and control: the weighted rich-club effect. Phys Rev Lett 10:168702

    Article  Google Scholar 

  • Owen G (1975) Multilinear extensions and the banzhaf value. Naval Res Logist Q 22:741–750

    Article  Google Scholar 

  • Owen G (1986) Values of graph restricted games. SIAM J Algebraic Discret Methods 7:210–220

    Article  Google Scholar 

  • Penrose L (1946) The elementary statistics of majority voting. J R Stat Soc 109(1):53–57

    Article  Google Scholar 

  • Ruiz LM, Valenciano F, Zarzuelo JM (1998) The family of least square values for transferable utility games. Games Econ Behav 24(1):109–130

    Article  Google Scholar 

  • Shapley L (1953) A value for n-person games. In: Kuhn H, Tucker A (eds) Contributions to the theory of games, vol II. Princeton University Press, Princeton , pp 307–317

  • Shapley L, Shubik M (1969) Pure competition, coalitional power and fair devision. Int Econ Rev 10(3):337–362

    Article  Google Scholar 

  • Slikker M (2005a) A characterization of the position value. Int J Game Theory 33(4):504–514

    Article  Google Scholar 

  • Slikker M (2005b) Link monotonic allocation schemes. Int Game Theory Rev 7(4):473–489

    Article  Google Scholar 

  • Suri N, Narahari Y (2008) Determining the top-k nodes in social networks using the shapley value. AAMAS 2008: proceedings of the seventh international joint conference on autonomous agents and multi-agent systems, pp 1509–1512

  • van den Brink R, van der Laan G (1998) Axiomatizations of the normalized banzhaf value and the shapley value. Soc Choice Welf 15(4):567–582

    Article  Google Scholar 

  • van den Nouweland A (1993) Games and graphs in economic situations. Ph.d. thesis, Tilburg University, Tilburg, The Netherlands

  • van den Nouweland A, Slikker M (2012) An axiomatic characterization of the position value for network situations. Math Soc Sci 64(3):266–271

    Article  Google Scholar 

  • Wasserman S, Faust K (1994) Social network analysis: methods and applications. Cambridge University Press, Cambridge

    Book  Google Scholar 

Download references

Acknowledgements

Thanks to André Casajus, Sebastian Garmann, Linda Hirt-Schierbaum, Eliane Lambertz, Walter Trockel and Wolfgang Leininger for helpful comments and discussions, and to the participants of the Dortmund Brown Bag Seminar in 2014, the 7th RGS Doctoral Conference in Economics 2014 in Dortmund, the 10th Spain-Italy-Netherlands Meeting on Game Theory (SING10) 2014 in Krakow and the 29th Annual Congress of the European Economic Association 2014 in Toulouse. Further thanks to two anonymous reviewers and the associate editor for their valuable comments and suggestions. All remaining errors are mine.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Julia Belau.

Appendices

Appendix A: Proofs

Proof of Lemma 1: If \(Y\in \mathcal {Y}\) satisfies S, the corresponding Y-Position value \(\pi ^Y\) satisfies DEG.

Proof

Consider any (Nvg) which is link anonymous, that is, there exists a function \(f:\{0,1,\ldots ,|g|\}\longrightarrow {\varvec{\mathbb {R}}}\) such that \(v^N(g')=f(|g'|)\) for all \(g'\subseteq g\).Then, for any two links \(\lambda ,\tilde{\lambda }\in g\), we have

$$\begin{aligned} v^N(g'\cup \lambda )= & {} f(|g'\cup \lambda |)=f(|g'|+1)=f(|g'\cup \tilde{\lambda }|)\\= & {} v^N(g'\cup \tilde{\lambda })\quad \forall \,g'\subseteq g{\setminus }\{\lambda ,\tilde{\lambda }\} \end{aligned}$$

that is, all \(\lambda ,\tilde{\lambda }\in g\) are (pairwise) symmetric in \((g,v^N)\). Therefore, for any allocation rule \(Y\in \mathcal {Y}\) which satisfies S, we have \(Y_{\lambda }(g,v^N)=Y_{\tilde{\lambda }}(g,v^N)=:\alpha \in {\varvec{\mathbb {R}}}\) (constant due to S). Hence, for the corresponding Y-Position value we have for any \(i\in N\)

$$\begin{aligned} \pi ^Y_i(N,v,g)=\sum \limits _{\lambda \in g_i}\frac{1}{2}Y_{\lambda }(g,v^N)=\alpha \frac{|g_i|}{2} \end{aligned}$$

which proves DEG. \(\square \)

Proof of Lemma 2: If \(Y\in \mathcal {Y}\) satisfies NI, the corresponding Y-Position value \(\pi ^Y\) satisfies SLP.

Proof

Consider any (Nvg). For all \(\tilde{\lambda }\in g\) which are superfluous in (Nvg), that is, \(v^N(g'\cup \tilde{\lambda })=v^N(g')\) for all \(g'\subseteq g{\setminus }\{\tilde{\lambda }\}\), we have that \(\tilde{\lambda }\) is a Nullplayer in \((g,v^N)\). By Y satisfying NI=NPO+N we have for all \(i\in N\)

$$\begin{aligned} \pi _i^Y(N,v,g)&=\sum \limits _{\lambda \in g_i{\setminus }\tilde{\lambda }}\frac{1}{2}Y_{\lambda }(g,v^N)+{\left\{ \begin{array}{ll}\frac{1}{2}Y_{\tilde{\lambda }}(g,v^N),&{} \text {if }\tilde{\lambda }\in g_i\\ 0,&{} \text {otherwise }\end{array}\right. }\\&\mathop {=}\limits ^{\text {NPO}}\sum \limits _{\lambda \in g_i{\setminus }\tilde{\lambda }}\frac{1}{2}Y_{\lambda }(g{\setminus }\tilde{\lambda },v^N)+{\left\{ \begin{array}{ll}\frac{1}{2}Y_{\tilde{\lambda }}(g,v^N),&{} \text {if }\tilde{\lambda }\in g_i\\ 0,&{} \text {otherwise }\end{array}\right. }\\&\mathop {=}\limits ^{\text {N}}\sum \limits _{\lambda \in g_i{\setminus }\tilde{\lambda }}\frac{1}{2}Y_{\lambda }(g{\setminus }\tilde{\lambda },v^N)=\pi _i^Y(N,v,g{\setminus }\tilde{\lambda }) \end{aligned}$$

which proves SLP. \(\square \)

Proof of Theorem 1: If an allocation rule for TU games with a network structure satisfies A, DEG, SLP and Y-CLP, \(Y\in ASN\), it is uniquely determined on cycle-free networks and it is given by the ASN-Position value \(\pi ^Y\in \Pi (ASN)\) corresponding to Y.

Proof

Existence: Since \(\pi ^Y\in \Pi (ASN)\), that is, the corresponding Y satisfies A, S, NI, we know that \(\pi ^Y\) satisfies A, DEG and SLP by Corollary 1, Lemmas 1 and 2, respective ly. Y-CLP follows by Corollary 2.

Uniqueness: We follow the idea of the proof for the (Shapley-) Position value of Borm et al. (1992). Let Y in ASN and let \(W\in \mathcal {Y}_G \) satisfy A, DEG, SLP and Y-CLP. Consider any TU-game with a network structure (Nvg) such that g is cycle free. The unanimity games \(\{u_T\}_{T\in 2^N{\setminus } \{\emptyset \}}\) with \(u_T(K)=1\), if \(T\subseteq K\) and 0, otherwise, build a basis of V. Hence, every \(v\in V\) can be written as

$$\begin{aligned} v=\sum \limits _{T\in 2^N{\setminus } \{\emptyset \}}\mu _T(v)u_T \end{aligned}$$

where the coefficients \(\{\mu _T\}_{T\in 2^N{\setminus } \{\emptyset \}}\) are called the Harsanyi dividends of v. Therefore, by A, it is sufficient to show that \( W(N,\beta u_T,g) \) is uniquely determined for all \(\beta \in {\varvec{\mathbb {R}}}\) and \(T\in 2^N\) such that \(|T|\ge 2\). Let such \(\beta , T\) be arbitrary but fixed.

Case 1: \(\not \exists \) \(C\in \mathcal {C}(N,g)\) such that \(T\subseteq C\).

That is, there exists \(i,j\in T\) being unconnected in g and hence, \(\beta u_T^N(g')=0\) for all \(g'\subseteq g\). Therefore, every \(\lambda \in g\) is superfluous, and hence, by SLP, we have \(W(N,\beta u_T,g)=W(N,\beta u_T,g{\setminus }\lambda _1)=W(N,\beta u_T,g{\setminus }\{\lambda _1,\lambda _2\})=\cdots =W(N,\beta u_T,\emptyset )\).

Trivially, the game \((N,\beta u_t,\emptyset )\) is link anonymous and hence, by DEG, there exists a constant \(\alpha \in {\varvec{\mathbb {R}}}\) such that

$$\begin{aligned} W_i(N,\beta u_T,g)=W_i(N,\beta u_t,\emptyset )=\alpha \cdot C^d_i(\emptyset )=0\,\forall \, i\in N\quad \text {(uniquely determined)} \end{aligned}$$

Case 2: \(\exists \) \(C\in \mathcal {C}(N,g)\) such that \(T\subseteq C\).

Consider the (unique) connected hull (cf. Owen 1986) of TH(T), given by

$$\begin{aligned} H(T):=\bigcap \big \{S|T\subseteq S\subseteq C\text { such that }g|_{S}\text { is connected subgraph}\big \} \end{aligned}$$

As g is cycle-free, H(T) is the minimal set of nodes that are essential to connect T. Note that cycle-freeness is essential here: if there is more than one path connecting T, the intersection is empty on the disjoint parts of the connecting paths. We have

$$\begin{aligned} \beta u_T^N(g')={\left\{ \begin{array}{ll}\beta ,&{}\text { if }g|_{H(T)}\subseteq g'\\ 0,&{}\text { otherwise}\end{array}\right. } \end{aligned}$$

All links \(\lambda \notin g|_{H(T)}\) are superfluous in \((N,\beta u_T,g)\), hence, by SLP,\(W(N,\beta u_T,g)=W(N,\beta u_T,g|_{H(T)})\).

The game \((N,\beta u_T,g|_{H(T)})\) is link anonymous (all links have the same number of swings, namely, one) with

$$\begin{aligned} f:\{0,1,\ldots ,|g|_{H(T)}|\}\longrightarrow {\varvec{\mathbb {R}}}, f(x):={\left\{ \begin{array}{ll} \beta ,&{}\text { if }x=|g|_{H(T)}|\\ 0,&{}\text { otherwise} \end{array}\right. } \end{aligned}$$

hence, by DEG, there exists a constant \(\alpha \in {\varvec{\mathbb {R}}}\) such that

$$\begin{aligned} W_i(N,\beta u_T,g)\mathop {=}\limits ^{\text {SLP}}W_i(N,\beta u_T, g|_{H(T)})\mathop {=}\limits ^{\text {DEG}}\alpha \cdot C^d_i(g|_{H(T)}) \end{aligned}$$
(1)

It directly follows that

$$\begin{aligned} W_i(N,\beta u_T,g)=0\quad \forall \, i\in N{\setminus } H(T) \end{aligned}$$
(2)

By (2) and Y-CLP, we have

$$\begin{aligned} \sum \limits _{i\in H(T)}W_i(N,\beta u_T,g)&\mathop {=}\limits ^{(2)}\sum \limits _{i\in C}W_i(N,\beta u_T,g) \mathop {=}\limits ^{\text {Y-CLP}}\sum \limits _{\lambda \in g|_C}Y_{\lambda }(g,\beta u_T^N). \end{aligned}$$

Note that all \(\lambda \in g{\setminus } g|_{H(T)}=g|_{N{\setminus } H(T)}\) are Nullplayers in \((g,\beta u_T^N)\) which yields and \(Y_{\lambda }(g,\beta u_T^N)=0\) since Y satisfies N. Furthermore, all \(\lambda ,\tilde{\lambda }\in g|_{H(T)}\) are pairwise symmetric in \((g,\beta u_T^N)\) and since Y satisfies S, there exist \(B^Y_T\in {\varvec{\mathbb {R}}}\) such that

$$\begin{aligned} Y_{\lambda }(g,\beta u_T^N)={\left\{ \begin{array}{ll} B^Y_T,&{}\text {if }\lambda \in g|_{H(T)}\\ 0,&{}\lambda \notin g|_{H(T)} \end{array}\right. }\Rightarrow \sum \limits _{\lambda \in g|_C}Y_{\lambda }(g,\beta u_T^N)=B_T^Y\cdot |g|_{H(T)}| \end{aligned}$$

where \(B^Y_T\in {\varvec{\mathbb {R}}}\) is uniquely determined by \(Y, \beta \) and T due to single-valuedness of Y and independent on g by NPO.

On the other hand, by (1), we have

$$\begin{aligned} \sum \limits _{i\in C}W_i(N,\beta u_T,g)=\alpha \sum \limits _{i\in C}C^d_i(g|_{H(T)})=\alpha \cdot 2\cdot |g|_{H(T)}| \end{aligned}$$

Combining this, we get

$$\begin{aligned} \alpha \cdot 2\cdot |g|_{H(T)}|=B_T^Y\cdot p|g|_{H(T)}|\Leftrightarrow \alpha =\frac{B_T^Y}{2} \end{aligned}$$

and hence

$$\begin{aligned} W_i(N,\beta u_T,g)=\frac{B_T^Y}{2}C^d_i(g|_{H(T)})=\frac{B_T^Y}{2}|g_i|_{H(T)}| \end{aligned}$$

which is uniquely determined, given \(Y\in ASN \). \(\square \)

Proof of Lemma 3: If \(Y\in ASN \), the corresponding Y-Position value \(\pi ^Y\) satisfies BLC.

Proof

Let (Nvg) be a TU game with a network structure. Following Slikker (2005a), there exists a unique linear combination of unanimity games (on link sets) representing the link game \(v^N\), that is \(v^N\) exists \(\{\beta ^{v}_{g'}\}_{g'\subseteq g}\) such that

$$\begin{aligned} v^N=\sum \limits _{g'\subseteq g}\beta ^v_{g'}u_{g'}. \end{aligned}$$

Note that the dividends \(\beta ^v\) do not depend on the choice set g (cf. Slikker 2005a) and that \(\beta ^v_{\emptyset }=0\) for zero-normalized v.

Now let \(Y\in ASN\) and consider \((g,\beta ^v_{g'}u_{g'})\) with \(g'\subseteq g\) arbitrary but fixed. All \(\lambda \in g{\setminus } g'\) are Nullplayers in this game which yields \(Y_\lambda (g,\beta ^v_{g'}u_{g'})=0\) for all \(\lambda \in g{\setminus } g'\) by N and \(Y_\lambda (g,\beta ^v_{g'}u_{g'})=Y_\lambda (g',\beta ^v_{g'}u_{g'})\) for all \(\lambda \in g'\) by NPO. All \(\lambda ,\tilde{\lambda }\in g'\) are (pairwise) symmetric which yields \(Y_\lambda (g,\beta ^v_{g'}u_{g'})=Y_\lambda (g',\beta ^v_{g'}u_{g'})=B_{g'}^{v,Y}\) by S where \(B_{g'}^{v,Y}\in \mathbb {R}\) is unique by single-valuedness and does not depend on the choice set g (NPO).

By A we obtain

$$\begin{aligned} Y_\lambda (g,v^N)=\sum \limits _{g'\subseteq g}Y_\lambda (g,\beta ^v_{g'}u_{g'})=\sum \limits _{\begin{array}{c} g'\subseteq g:\\ \lambda \in g' \end{array}}B_{g'}^{v,Y}\quad \forall \,\lambda \in g. \end{aligned}$$

For the corresponding Y-Position value we hence have

$$\begin{aligned} \pi ^{Y}_i(N,v,g)&=\sum \limits _{\lambda \in g_i}\frac{1}{2}\sum \limits _{\begin{array}{c} g'\subseteq g:\\ \lambda \in g' \end{array}}B_{g'}^{v,Y}=\sum \limits _{\begin{array}{c} g'\subseteq g,\\ g'\ne \emptyset \end{array}}\sum \limits _{\lambda \in g_i'}\frac{1}{2}B_{g'}^{v,Y}=\sum \limits _{\begin{array}{c} g'\subseteq g,\\ g'\ne \emptyset \end{array}}B_{g'}^{v,Y}\frac{|g_i'|}{2} \end{aligned}$$

and therefore (using that \(B_{g'}^{v,Y}\) does not depend on g) we obtain

which proves BLC. \(\square \)

Proof of Theorem 2: If an allocation rule for TU games with a network structure satisfies BLC and Y-CLP, \(Y\in ASN \), it is uniquely determined for general networks and it is given by the ASN-Position value \(\pi ^Y\in \Pi (ASN)\) corresponding to Y.

Proof

Existence: Since \(\pi ^Y\in \Pi (ASN)\), that is, the corresponding Y is an ASN-value, we know that \(\pi ^Y\) satisfies BLC by Lemma 3. Y-CLP follows by Corollary 2. Uniqueness: Suppose VW being two allocation rules for network structures satisfying BLC and Y-CLP, \(Y\in ASN \). We proceed by induction over |g|.

Induction basis [IB]: For \(|g|=0\), that is, \(g=\emptyset \), we have that \(\mathcal {C}(N,g)=\{i\}_{i\in N}\) and hence, by Y-CLP, we have

$$\begin{aligned} V_i(N,v,g)&=\sum \limits _{i\in C}V_i(N,v,g)= \sum \limits _{\lambda \in \emptyset }Y_{\lambda }(g,v^N)=0\\&=\sum \limits _{i\in C}W_i(N,v,g)=W_i(N,v,g). \end{aligned}$$

Now suppose that \(V(N,v,g)=W(N,v,g)\) for all g such that \(|g|=k, k\ge 0\) (induction hypothesis [IH]).

Consider g such that \(|g|=k+1\). As \(k+1\ge 1\), there exists \(i,j\in N, i\ne j\) such that \(j\in C_i(N,g)\) (for all l with \(|C_l(N,g)|=1\), we have \(V_l(N,v,g)=W_l(N,v,g)\) by Y-CLP). By BLC, we have for all \(i,j\in C_i(N,g)\):

$$\begin{aligned} \sum \limits _{\lambda \in g_j}V_i(N,v,g)-\sum \limits _{\lambda \in g_i}V_j(N,v,g)&\mathop {=}\limits ^{\text {BLC}}\sum \limits _{\lambda \in g_j}V_i(N,v,g{\setminus }\lambda )-\sum \limits _{\lambda \in g_i}V_j(N,v,g{\setminus }\lambda )\\&\mathop {=}\limits ^{\text {[IH]}}\sum \limits _{\lambda \in g_j}W_i(N,v,g{\setminus }\lambda )-\sum \limits _{\lambda \in g_i}W_j(N,v,g{\setminus }\lambda )\\&\mathop {=}\limits ^{\text {BLC}}\sum \limits _{\lambda \in g_j}W_i(N,v,g)-\sum \limits _{\lambda \in g_i}W_j(N,v,g) \end{aligned}$$

and hence

$$\begin{aligned} |g_j|V_i(N,v,g)-|g_i|V_j(N,v,g)=|g_j|W_i(N,v,g)-|g_i|W_j(N,v,g) \end{aligned}$$
(3)

Summing up (3) over all \(j\in C\) yields:

$$\begin{aligned}&\displaystyle \sum \limits _{j\in C}\left[ |g_j|V_i(N,v,g)-|g_i|V_j(N,v,g)\right] =\sum \limits _{j\in C}\left[ |g_j|W_i(N,v,g)-|g_i|W_j(N,v,g)\right] \\&\displaystyle \Leftrightarrow V_i(N,v,g)\underbrace{\sum \limits _{j\in C}|g_j|}_{\begin{array}{c} :=A(j)>0 \\ \text {by }|C_i|\ge 1 \end{array}}-|g_i|\sum \limits _{j\in C}V_j(N,v,g) =W_i(N,v,g)\underbrace{\sum \limits _{j\in C}|g_j|}_{\begin{array}{c} :=A(j)>0 \\ \text {by }|C_i|\ge 1 \end{array}}-|g_i|\sum \limits _{j\in C}W_j(N,v,g)\\&\displaystyle \mathop {\Rightarrow }\limits ^{\text {Y-CLP}} A(j)\cdot V_i(N,v,g)-|g_i|\sum \limits _{\lambda \in g|_C}Y_{\lambda }(g,v^N) =A(j)\cdot W_i(N,v,g)-|g_i|\sum \limits _{\lambda \in g|_C}Y_{\lambda }(g,v^N)\\&\displaystyle \Leftrightarrow V_i(N,v,g)=W_i(N,v,g)\quad \forall \, i\in N \end{aligned}$$

which finishes the proof. \(\square \)

Proof of Theorem 3: Any multiplicative Y-Position value \(\pi ^Y\in \Pi (ASN_m)\) can be written as

$$\begin{aligned} \pi _i^Y(N,v,g)=\sum \limits _{\lambda \in g_i}\sum \limits _{g'\subseteq g{\setminus }\lambda }\frac{w^Y_{|g|,|g'|}}{2} \left( v^N(g'\cup \lambda )-v^N(g')\right) \end{aligned}$$

where the weighting scheme \(\left\{ w^Y_{|g|,|g'|}\right\} _{|g'|=0,\ldots ,|g|-1}\) satisfies

  1. 1.

    \(w^Y_{|g|,|g'|}=Y_{\lambda }(g,1_{g'\cup \lambda })\) for any \(\lambda \in g\) and \(g'\subseteq g{\setminus }\lambda \)

  2. 2.

    \(w^Y_{|g|,|g'|}+w^Y_{|g|,|g'|+1}=w^Y_{|g|-1,|g'|}\quad \forall \,|g'|=0,\ldots ,|g|-1\)

Proof

Let \(Y\in ASN, K\subseteq N\) be arbitrary but fixed and consider the game \((N,1_K)\). All players in K as well as all players in \(N{\setminus } K\) are (pairwise) symmetric in this game, hence, by S, there must exist \(w^Y(N,K)\) and \(\tilde{w}^Y(N,K)\) (uniquely determined by single-valuedness) such that we have

$$\begin{aligned} Y_i(N,1_K)={\left\{ \begin{array}{ll} w^Y(N,K),&{}i\in K\\ \tilde{w}^Y(N,K),&{}i\notin K \end{array}\right. } \end{aligned}$$
(4)

for all \(i\in N\). Let \(i\in N\) be arbitrary but fixed and, for all \(K\subseteq N{\setminus }\{i\}\), consider the game \((N,1_{K\cup \{i\}}+1_K)\). One can easily check that i is a Nullplayer in this game and using N and A we observe

$$\begin{aligned} \tilde{w}^Y(N,K)=-w^Y(N,K\cup \{i\})\text { for all }i\in N\text { and all }K\subseteq N{\setminus }\{i\}. \end{aligned}$$
(5)

Now, let \(i,j\in N, j\ne i\) be arbitrary but fixed and, for all \(K\subseteq N{\setminus }\{i,j\}\), consider the game \((N,1_{K\cup \{i\}}+1_{K\cup \{j\}})\). One can easily check that i and j are symmetric in this game and using S, A and (5) we observe

$$\begin{aligned} w^Y(N,K\cup \{i\})=w^Y(N,K\cup \{j\})\text { for all }i,j\in N,i\ne j,\text { and all }K\subseteq N{\setminus }\{i,j\}. \end{aligned}$$
(6)

For any \(K,K'\subseteq N\) with \(|K|=|K'|\), we can build a sequence\(\{K=K_0,K_1,\ldots , K_{l}=K'\}\) where every set can be obtained from its predecessor by replacing exactly one player. Using (6) we hence obtain that the weights only depend on sizes of coalitions (cf. also Ruiz et al. (1998), proof of Lemma 9). We obtain

$$\begin{aligned} Y_i(N,1_K)=\left\{ \begin{array}{lll} w^Y(|N|,|K|),&{}i\in K\\ -w^Y(|N|,|K\cup \{i\}|),&{}i\notin K\end{array}\right. \quad \text { for all }K\subseteq N\text { with }|K|=k. \end{aligned}$$
(7)

Now let \(Y\in ASN_m\) and (Nv) be any TU-game. Note that any \(v\in V\) can be written as \( v=\sum _{K\subseteq N}v(K)\cdot 1_K. \) Using (7) and that Y is not only additive but also multiplicative (m), we have

$$\begin{aligned} Y_i(N,v)\mathop {=}\limits ^{\mathbf{A }}\sum \limits _{K\subseteq N}Y_i(N,v(K)\cdot 1_K)&\mathop {=}\limits ^{\mathbf{m }}\sum \limits _{K\subseteq N}v(K)\cdot Y_i(N,1_K)\nonumber \\&=\sum \limits _{K\subseteq N:i\in K}v(K)\cdot w^Y(|N|,|K|)\nonumber \\&\quad +\sum \limits _{K\subseteq N:i\notin K}v(K)\cdot [-w^Y(|N|,|K\cup \{i\}|)]\nonumber \\&=\sum \limits _{K\!\subseteq \! N{\setminus }\{i\}}w^Y(|N|,|K\cup \{i\}|)[v(K\cup \{i\})\!-\!v(K)]. \end{aligned}$$
(8)

Now let (Nvg) be any TU-game with a network structure. Using (8), we obtain

$$\begin{aligned} \pi _i^Y(N,v,g)&=\sum \limits _{\lambda \in g_i}\frac{1}{2}\sum \limits _{g'\subseteq g{\setminus }\lambda }w^Y_{|g|,|g'|}\left( v^N(g'\cup \lambda )-v^N(g')\right) \end{aligned}$$

where \(w^Y_{|g|,|g'|}=Y_{\lambda }(g,1_{g'\cup \lambda })\) for any \(\lambda \in g\), and \(g'\subseteq g{\setminus }\lambda \).

To show the second part of the Theorem, we will make use of NPO. Let \(\lambda ,\lambda '\in g\) with \(\lambda '\ne \lambda \) being a Nullplayer in \((g,v^N)\). Following the same steps as in Derks and Haller (1999) we obtain

$$\begin{aligned} w^Y_{|g|,|g'|}+w^Y_{|g|,|g'|+1}=w^Y_{|g|-1,|g'|}\,\forall \,|g'|=0,\ldots ,|g|-1 \end{aligned}$$
(9)

\(\square \)

Proof of Theorem 4: We will make use of the following expression:

Lemma 6

The weights from Theorem 3 satisfy

$$\begin{aligned} w^Y_{m,k}=\sum \limits _{l=0}^{n-m}\left( {\begin{array}{c}n-m\\ l\end{array}}\right) w^Y_{n,k+l} \quad \text { for all }n\ge m>k. \end{aligned}$$

The proof follows by induction over n (starting at \(n=m\)).

Proof of Theorem 4.1: Any multiplicative ASN-Position value \(\pi ^Y\in \Pi (ASN_m)\) satisfies LMON for all TU-games with a network structure such that the corresponding link-game is link-convex.

Proof

Let \(Y\in ASN_m\) and let (Nvg) be any TU-game with a network structure such that the corresponding link-game \(v^N\) is link-convex. By Theorem 3, for any \(i\in N\) we have

$$\begin{aligned} \pi _i^Y(N,v,g)&=\sum \limits _{\lambda \in g_i}\frac{1}{2}\sum \limits _{g'\subseteq g{\setminus }\lambda }w^Y_{|g|,|g'|}\left( v^N(g'\cup \lambda )-v^N(g')\right) \\&=\frac{1}{2}\sum \limits _{\lambda \in g_i}\sum \limits _{\begin{array}{c} g'\subseteq g:\\ \lambda \in g' \end{array}}w^Y_{|g|,|g'|-1}\left( v^N(g')-v^N(g'{\setminus }\lambda )\right) \\&=\frac{1}{2}\sum \limits _{\begin{array}{c} g'\subseteq g:\\ g'\ne \emptyset \end{array}}w^Y_{|g|,|g'|-1}\sum \limits _{\lambda \in g'_i}\left( v^N(g')-v^N(g'{\setminus }\lambda )\right) \end{aligned}$$
(*)

Now consider any \(\tilde{g}\subseteq g\). We can rewrite any \(g'\subseteq g\) as \(g'=h_1\cup h_2\) with \(h_1\subseteq \tilde{g}\) and \(h_2\subseteq g{\setminus }\tilde{g}\). Note that we either have \(h_1\ne \emptyset \) or \(h_2\ne \emptyset \). From (\(*\)) and using link convexity of \(v^N\) we obtain

$$\begin{aligned} \pi _i^Y(N,v,g)&=\frac{1}{2}\sum \limits _{h_1\subseteq \tilde{g}}\sum \limits _{h_2\subseteq g{\setminus }\tilde{g}}w^Y_{|g|,|h_1\cup h_2|-1}\sum \limits _{\lambda \in (h_1\cup h_2)_i}\\&\quad \times \left( v^N(h_1\cup h_2)-v^N((h_1\cup h_2){\setminus }\lambda )\right) \\&\ge \frac{1}{2}\sum \limits _{h_1\subseteq \tilde{g}}\sum \limits _{h_2\subseteq g{\setminus }\tilde{g}}w^Y_{|g|,|h_1\cup h_2|-1}\sum \limits _{\lambda \in (h_1)_i}\left( v^N(h_1)-v^N(h_1{\setminus }\lambda )\right) \\&=\frac{1}{2}\sum \limits _{h_1\subseteq \tilde{g}}\sum \limits _{\lambda \in (h_1)_i}\left( v^N(h_1)-v^N(h_1{\setminus }\lambda )\right) \sum \limits _{h_2\subseteq g{\setminus }\tilde{g}}\underbrace{w^Y_{|g|,|h_1|+|h_2|-1}}_{\begin{array}{c} \text {only depends on}\\ \text {sizes of link-sets} \end{array}}\\&=\frac{1}{2}\sum \limits _{h_1\subseteq \tilde{g}}\sum \limits _{\lambda \in (h_1)_i}\left( v^N(h_1)-v^N(h_1{\setminus }\lambda )\right) \\&\quad \times \sum \limits _{k=0}^{|g|-|\tilde{g}|}\left( {\begin{array}{c}|g|-|\tilde{g}|\\ k\end{array}}\right) w^Y_{|g|,(|h_1|-1)+k}\\&{\mathop {=}\limits ^{\text {Lemma }6}}\frac{1}{2}\sum \limits _{h_1\subseteq \tilde{g}}w^Y_{|\tilde{g}|,|h_1|-1}\sum \limits _{\lambda \in (h_1)_i}\left( v^N(h_1)-v^N(h_1{\setminus }\lambda )\right) \\&\mathop {=}\limits ^{(*)}\pi _i^Y(N,v,\tilde{g}) \end{aligned}$$

\(\square \)

Proof of Theorem 4.2: Any multiplicative ASN-Position value \(\pi ^Y\in \Pi (ASN_m)\) satisfies Component Decomposability CD:

$$\begin{aligned} \pi _i^Y(N,v,g)=\pi _i^Y(C_i(N,g),v|_{C_i(N,g)},g|_{C_i(N,g)}). \end{aligned}$$

Proof

Since \(g_i\subseteq (g|_{C_i(N,g)})_i\), it is sufficient to show that any ASN value Y satisfies

$$\begin{aligned} Y_{\lambda }(g,v^N)=Y_{\lambda }(g|_C,v^N|_{C})\quad \forall \, \lambda \in g|_C,\, C\in \mathcal {C}(N,g)\text { and all }v\in V^0. \end{aligned}$$

Let \(Y\in ASN_m, (N,v)\) be a standard TU-game and let g be a network. Consider \(C\in \mathcal {C}(N,g)\). For any \(\lambda \in g|_C\), marginal contributions are not effected by connections outside C due to the form of \(v^N\). Using this and following the same decomposition method as in the proof of Theorem 4.1, we obtain

$$\begin{aligned} Y_{\lambda }(g,v^N)&=\sum \limits _{g'\subseteq g{\setminus }\lambda }w^Y_{|g|,|g'|}\left[ v^N((g'\cap g|_C)\cup \lambda )-v^N(g'\cap g|_C)\right] \\&=\sum \limits _{g'\subseteq g{\setminus }\lambda }w^Y_{|g|,|g'|}\underbrace{\left[ v^N(g'|_C\cup \lambda )-v^N(g'|_C)\right] }_{\begin{array}{c} \text {the same for all }g',g''\subseteq g{\setminus }\lambda \\ \text {such that } g'|_C=g''|_C \end{array}}\\&=\sum \limits _{\tilde{g}\subseteq (g|_C){\setminus }\lambda }\left[ v^N(\tilde{g}\cup \lambda )-v^N(\tilde{g})\right] \sum \limits _{h\subseteq g|_{N{\setminus } C}}w^Y_{|g|,|\tilde{g}|+|h|}\\&=\sum \limits _{\tilde{g}\subseteq (g|_C){\setminus }\lambda }\left[ v^N(\tilde{g}\cup \lambda )-v^N(\tilde{g})\right] \sum \limits _{k=0}^{|g|_{N{\setminus } C}|}\left( {\begin{array}{c}|g|_{N{\setminus } C}|\\ k\end{array}}\right) w^Y_{|g|,|\tilde{g}|+k}\\&\mathop {=}\limits ^{\begin{array}{c} |g|_{N{\setminus } C}|=|g|-|g|_C|\text { and}\\ \text {Lemma 6} \end{array}}\sum \limits _{\tilde{g}\subseteq (g|_C){\setminus }\lambda }w^Y_{|g|_C|,|\tilde{g}|}\left[ v^N(\tilde{g}\cup \lambda )- v^N(\tilde{g})\right] \\&=Y_{\lambda }(g|_C,v^N|_C) \end{aligned}$$

which finishes the proof. \(\square \)

Proof of Theorem 4.3: Any multiplicative ASN-Position value \(\pi ^Y\in \Pi (ASN_m)\) can be written as

$$\begin{aligned} \pi _i^Y(N,v,g)=\sum \limits _{\lambda \in g_i}\sum \limits _{\begin{array}{c} g'\subseteq g:\\ \lambda \in g' \end{array}}\beta _{g'}^v\frac{w^Y_{|g'|,|g'|-1}}{2} \end{aligned}$$

where \(\{\beta _{g'}^v\}_{g'\subseteq g}\) are the Harsanyi dividends representing \(v^N\) and \(w^Y\) is the weight defined in Theorem 3.

Proof

Let \(Y\in ASN_m\) and (Nvg) be any TU game with a network structure. Recall from the proof of Theorem 3, that the link game \(v^N\) can be written as a (unique) linear combination of (link set) unanimity games where the corresponding scalars \(\{\beta ^v_{g'}\}_{g'\subseteq g}\) are called Harsanyi dividends. Let \(g'\subseteq g\) be arbitrary but fixed. By N, we have \(Y_{\lambda }(g,u_{g'})=0\) for all \(\lambda \notin g'\). Consider any \(\lambda \in g'\). Note that any (link set) unanimity game can be written as

$$\begin{aligned} u_{g'}=\sum \limits _{\bar{g}\subseteq g{\setminus } g'}1_{g'\cup \bar{g}}. \end{aligned}$$

Using the weights from Theorem 3 and by Lemma 6, we obtain

$$\begin{aligned} Y_{\lambda }(g,u_{g'})= & {} \sum \limits _{\bar{g}\subseteq g{\setminus } g'}Y_{\lambda }(g,1_{g'\cup \bar{g}})\mathop {=}\limits ^{\lambda \in g'}\sum \limits _{\bar{g}\subseteq g{\setminus } g'}w^Y_{|g|,|g'|+|\bar{g}|-1}\\= & {} \sum \limits _{l=0}^{|g|-|g'|}\left( {\begin{array}{c}|g|-|g'|\\ l\end{array}}\right) w^Y_{|g|,|g'|-1+l}=w^Y_{|g'|,|g'|-1} \end{aligned}$$

which yields

$$\begin{aligned} \pi _i^Y(N,v,g)= & {} \sum \limits _{\lambda \in g_i}\frac{1}{2}Y_{\lambda }(g,v^N)=\sum \limits _{\lambda \in g_i}\frac{1}{2}\sum \limits _{\begin{array}{c} g'\subseteq g:\\ \lambda \in g' \end{array}}\beta _{g'}^vY_{\lambda }(g,u_{g'})\\= & {} \sum \limits _{\lambda \in g_i}\sum \limits _{\begin{array}{c} g'\subseteq g:\\ \lambda \in g' \end{array}}\beta _{g'}^v\frac{w^Y_{|g'|,|g'|-1}}{2} \end{aligned}$$

\(\square \)

Proof of Lemma 5: The two characterizing axioms of the ASN-Position values are independent.

Proof

To show independence of the axioms, we need to find elements in \(\mathcal {Y}_G\) which satisfy one but not the other property. The Shapley-Position value satisfies BLC but (trivially) does not satisfy CLP w. r. t. the Banzhaf value. Now consider the allocation rule which shares the Banzhaf-link power of a component equally within this component, that is, the rule given by

$$\begin{aligned}W_i(N,v,g):=\frac{\sum \limits _{\lambda \in g|_{\mathcal {C}_{i}}}Ba_{\lambda }(g|_{\mathcal {C}_{i}},v^N|_{\mathcal {C}_{i}})}{|\mathcal {C}_{i}|}\in \mathcal {Y}_G.\end{aligned}$$

By definition and using Theorem 4.2, W satisfies CLP w. r. t. the Banzhaf value. To see that W does not satisfy BLC, consider the following TU-game with a network structure:

$$\begin{aligned}N = \{1, 2, 3, 4\},\quad v(K)={\left\{ \begin{array}{ll} 1,&{}\{1,2\}\subseteq K\\ 1,&{}\{1,3\}\subseteq K\\ 1,&{}\{2,3,4\}\subseteq K\\ 0,&{}\text { otherwise} \end{array}\right. },\quad g=\{12,23,34\}\end{aligned}$$

All players are connected in (Ng) resulting in \(C_i=N\) for all \(i\in N\). For the Banzhaf values of links we obtain \(Ba_{\{12\}}=\frac{3}{2^{3-1}}=\frac{3}{4}\) (3 possibilities of creating 1 when materializing) and \(Ba_{\{23\}}=Ba_{\{34\}}=\frac{1}{4}\) (1 possibility of creating 1 when materializing) resulting in a component power of . Equal sharing implies for \(i=1,2,3,4\). For BLC consider players 1 and 2. We have

$$\begin{aligned}&\sum \limits _{\lambda \in g_1}\left[ W_2(N,v,g)-W_2(\mathcal {C}(2)(N,g-\lambda ),v|_{\mathcal {C}(2)(N,g-\lambda )},g|_{\mathcal {C}(2)(N,g-\lambda )})\right] \\&\quad =\frac{5}{16}-W_2(\{2,3,4\},v|_{\{2,3,4\}}, \{23,34\})\\ \text { and}&\\&\sum \limits _{\lambda \in g_2}\left[ W_1(N,v,g)-W_1(\mathcal {C}(1)(N,g-\lambda ),v|_{\mathcal {C}(1)(N,g-\lambda )},g|_{\mathcal {C}(1)(N,g-\lambda )})\right] \\&\quad =\frac{5}{16}-W_1(\{1\},v|_{\{1\}},\emptyset )+\frac{5}{16}-W_1(\{1,2\},v|_{\{1,2\}},\{12\}) \end{aligned}$$

In the restricted game \((\{2,3,4\},v|_{\{2,3,4\}}, \{23,34\})\), each link has a Banzhaf power of \(\frac{1}{2^{2-1}}\) and equal sharing implies for \(i=2,3,4\). In the restricted game \((\{1,2\},v|_{\{1,2\}},\{12\})\), the link has a Banzhaf power of \(\frac{1}{2^{1-1}}\) and equal sharing implies for \(i=1,2\). We hence obtain \(-\frac{1}{48}\) in the first case but \(\frac{1}{8}\ne -\frac{1}{48}\) in the second case. \(\Rightarrow \) BLC is violated. \(\square \)

Appendix B: Supplementary material political example

See Tables 7, 8 and 9.

Table 7 Swings & Shapley/Banzhaf values (to 100%) for Parties (2nd line expresses less restrictive case)
Table 8 Swings & Shapley/Banzhaf values (to 100%) for connections
Table 9 Classic centrality approaches

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Belau, J. The class of ASN-position values. Soc Choice Welf 50, 65–99 (2018). https://doi.org/10.1007/s00355-017-1074-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00355-017-1074-4

Navigation