1 Introduction

The Shapley (1953) value is one of the basic solution concepts of cooperative game theory. Axiomatically, it was characterized by the linearity, efficiency, symmetry, and null player axioms. In various fields of economics and political science it is necessary to consider games that involve a large number of “insignificant” players. In such cases it is fruitful to model the game as a game with a continuum of players. Aumann and Shapley (1974) modified the value axioms and extended the definition of the value to games with a continuum of players. They proved the existence and uniqueness of the value on important spaces of nonatomic games. In computational terms, both the Shapley value and the Aumann–Shapley value assign each agent his average marginal contribution to coalitions he may join, in certain models of random ordering of the agents.

On general classes of nonatomic games the value may not exist. Nevertheless, this is not the case when market games, derived from transferable utility (TU) perfectly competitive economies, are concerned. The core of a market game is nonempty, and if all agents have differentiable utility functions, then the Aumann–Shapley value is also well defined, and it constitutes the unique element of the core.

Without the differentiability assumption, the core of a market game is usually multivalued. But even in this scenario, all known values have the property of being elements of the core. This is the case with (Hart 1980) measure-based values of a nonatomic market game, and also with the (Mertens 1988) value (which is defined on a large space of games containing the linear space spanned by all nonatomic market games). In fact, in the special case of market games, the Mertens value is given by an explicitFootnote 1 formula as a barycenter of the core (Mertens 1988).

The important and natural question that arises is whether the properties of the value can uniquely determine it, and make sure it is a member of the core. Arbitrary core selections can lead to distortions, e.g., significant differences in payoffs in “nearby” economies, discrimination between agents, and inconsistencies between choices. However, if the value indeed uniquely selects an element of the core, then the different choices of that element in different games are linked to each other via the value axioms, in a consistent and economically meaningful way.

This problem has proved to be extremely difficult. The first advancement in that direction was made by Haimanko (2002), essentiallyFootnote 2 proving that if the games are functions of finitely many mutually singular probability measuresFootnote 3 then the standard value axioms together with the continuity axiom uniquely determine the Mertens value. Nevertheless, (Haimanko 2002) analysis heavily relies on the mutual-singularity assumption. It seems that a milestone in the way to characterize the value on the space of all market games would be to first eliminate the mutual singularity assumption.

One of the main tools developed by Haimanko (2002) is a representation of the value as an “average” of marginals of the underlying game. In this paper we prove a generalization of this result for the case of general vector measure market games. We then use this result to characterize the Mertens value as the unique continuous value on the space of vector measure market games. A subsequent result is a sufficient condition for the uniqueness of the value on other spaces of vector measure market games.

2 Definitions, axioms, and the main result

Let \((T, {\mathcal {C}})\) be a standard measure space. We shall call the members of T players and the members of \({\mathcal {C}}\) coalitions. A game is a real valued function \(v:{\mathcal {C}}\rightarrow {\mathbb {R}}\) s.t. \(v(\emptyset )=0\). A game v is monotonic iff \(v(S)\le v(S^\prime )\) whenever \(S\subseteq S^\prime \). If Q is a set of games then \(Q^+\) denotes the subset of monotonic games in Q, and \(Q^1\) denotes the subset \(\{v\in Q^+:v(T)=1\}\). A game v is of bounded variation iff it is the difference between two monotonic games. The space of all games of bounded variation will be denoted by BV. The variation of a game \(v\in BV\) is the supremum of the variation of v over all increasing chains \(S_0\subseteq S_1\subseteq ...\subseteq S_m\) in \({\mathcal {C}}\), or equivalently,

$$\begin{aligned} \Vert v\Vert =\inf \left\{ u(T)+w(T): \textit{u,w are monotonic games s.t.}\; v=u-w\right\} . \end{aligned}$$
(1)

The variation defines a norm on BV (see Aumann and Shapley 1974). FA denotes the subspace of BV consisting of all finitely additive games, and NA denotes its subspace consisting of nonatomic and countably additive measures.

Let \(\varTheta \) denote the group of measurable automorphismsFootnote 4 of \((T,{\mathcal {C}})\). Each \(\theta \in \varTheta \) induces a linear mapping of BV onto itself byFootnote 5 \((\theta v)(S)=v(\theta S)\). A linear subspace \(Q\subseteq BV\) is symmetric iff \(\theta Q=Q\) for each \(\theta \in \varTheta \).

Let \(Q\subseteq BV\) be a symmetric space. A map \(\varphi : Q\rightarrow BV\) is positive iff \(\varphi (Q^+)\subseteq BV^+\), symmetric iff \(\theta \varphi =\varphi \theta \) for every \(\theta \in \varTheta \), and efficient iff \(\varphi (v)(T)=v(T)\) for every \(v\in Q\).

Definition 1

Let \(Q\subseteq BV\) be a symmetric linear space. A value on Q is a symmetric, positive and efficient linear map \(\varphi :Q\longrightarrow FA\). A value is continuous iff it is continuous w.r.t. \(\Vert \cdot \Vert _{BV}\).

A market function is a concave, continuous, nondecreasing, and homogeneous of degree 1 function on \({\mathbb {R}}_+^k\). Denote by \(M_+^k\) the cone of market functions on \({\mathbb {R}}_+^k\), and let \(M^k\) be the vector space of differences of functions in \(M_+^k\). Denote by \(LM_+^k\) the set of Lipschitz market functions, i.e., the subset of \(M_+^k\) consisting of Lipschitz functions. Denote by \(LM^k\) the vector space of differences of Lipschitz market functions. Denote by \({\mathcal {M}}\) the space of games of the form \(f\circ \mu \) where \(f\in M^k\) and \(\mu \in \left( NA^1\right) ^k\) for some \(k\ge 2\). We refer to \({\mathcal {M}}\) as the space of vector measure market games. Denote by \(\mathcal {LM}\) the linear subspace of \({\mathcal {M}}\) consisting of games of the form \(f\circ \mu \) where \(f\in LM^k\) and \(\mu \in \left( NA^1\right) ^k\) for some \(k\ge 2\).

In this work we shall prove a representation theorem for values on \(\mathcal {LM}\). This result will be formulated and developed in Sect. 3.3. It will be then employed to characterize the continuous values on \({\mathcal {M}}\) and the values on \(\mathcal {LM}\). It is known that the (Mertens 1988) value, \(\Psi _M\), is continuous. Additionally, the Mertens value is entirely determined by the core of a market game, and is itself an element of the core (Mertens 1988). We shall prove that this selection is uniquely determined by the value and continuity axioms.

Theorem 1

The unique continuous value on \({\mathcal {M}}\) is the Mertens value.

Theorem 2

The unique value on \(\mathcal {LM}\) is the Mertens value.

3 Preparations

3.1 Directional derivatives of market functions

Given \(f\in M^k\), \(x\in {\mathbb {R}}_{++}^k\), and \(y\in {\mathbb {R}}^k\), the directional derivative df(xy) of f at x in the direction y is given by

$$\begin{aligned} df(x,y)=\lim \limits _{\varepsilon \searrow 0}\frac{f(x+\varepsilon y)-f(x)}{\varepsilon }. \end{aligned}$$
(2)

The limit exists for every \(f\in M_+^k\) by concavity.

If \(f\in M_+^k\) then for every \(x\in {\mathbb {R}}_{++}^k\) the function \(df(x,\cdot ):{\mathbb {R}}^k\rightarrow {\mathbb {R}}\) is concave. Thus the directional derivative of \(df(x,\cdot )\) at \(y\in {\mathbb {R}}^k\) in the direction \(z\in {\mathbb {R}}^k\) which is given by

$$\begin{aligned} df(x,y,z)=\lim \limits _{\varepsilon \searrow 0}\frac{df(x,y+\varepsilon z)-df(x,y)}{\varepsilon }, \end{aligned}$$
(3)

exists. By linearity, Eqs. (2)–(3) extend to every \(f\in M^k\).

Denote by \(\mathbf 1 _k\in {\mathbb {R}}^k\) the vector with \((\mathbf 1 _k)_1=\ldots =(\mathbf 1 _k)_k=1\), and let \(D^k=\left\{ t\mathbf 1 _k:t\in {\mathbb {R}}_+\right\} \). Notice that if \(f\in M^k\) then for every \(x\in D^k{\setminus }\left\{ 0_k\right\} \), every \(y,z\in {\mathbb {R}}^k\), every \(a\in {\mathbb {R}}\), and every \(b>0\),

$$\begin{aligned} df(x,a\mathbf 1 _k+by,z)=df(\mathbf 1 _k,y,z). \end{aligned}$$
(4)

For every \(\chi :T\rightarrow {\mathbb {R}}^k\), \(x\in {\mathbb {R}}_{++}^k\), and \(y\in {\mathbb {R}}^k\) let \(df(x,y,\chi ):T\rightarrow {\mathbb {R}}\) be given by

$$\begin{aligned} \forall t\in T,\; df(x,y,\chi )(t)=df(x,y,\chi (t)). \end{aligned}$$
(5)

3.2 The direction space \(X_\lambda \)

For every \(k\ge 1\) and \(\mu \in \left( NA^1\right) ^k\) let \(AF(\mu )\) be the vector space generated by the range of \(\mu \), denoted \(\mathcal R(\mu )\). Further denoteFootnote 6 \({\mathbb {S}}_\perp ^\mu =\left\{ \frac{x}{\Vert x\Vert _2}:x\in AF(\mu ),\overline{x}=0\right\} \), and \(\varDelta (\mu )=\left\{ \frac{x}{k\overline{x}}:x\in \mathcal R(\mu ){\setminus }D^k\right\} \), with \(\overline{x}=\frac{1}{k}\sum \nolimits _{i=1}^kx_i\). For every \(\mu \in \left( NA^1\right) ^k\) endow the set \(\Lambda _\mu =\varDelta (\mu )\sqcup {\mathbb {S}}_\perp ^\mu \) with the topology \(\mathfrak T_\mu \) whose restriction to either \(\varDelta (\mu )\) or \({\mathbb {S}}_\perp ^\mu \) is equivalent to the Euclidean topology, and if a sequence \(\left( x^n\right) _{n=1}^\infty \subset \varDelta (\mu )\) converges, in the Euclidean topology, to some point in \(D^k\) and satisfies \(\frac{x^n-\overline{x^n}\mathbf 1 _k}{\Vert x^n-\overline{x^n}\mathbf 1 _k\Vert _2}\underset{n\rightarrow \infty }{\longrightarrow }y\in {\mathbb {S}}_\perp ^\mu \) then \(x^n\underset{n\rightarrow \infty }{\longrightarrow }y\) in \(\mathfrak T_\mu \). The topological space \((\Lambda _\mu ,\mathfrak T_\mu )\) is thus a compact metrizable space.

For every \(\lambda \in NA^1\) and \(k\ge 1\) denote \(\mathcal Z_\lambda ^k=\left\{ \mu \in \left( NA^1\right) ^k:\overline{\mu }\ll \lambda ,\;\frac{d\overline{\mu }}{d\lambda }\in L^\infty (\lambda )\right\} \), and \(\mathcal Z_\lambda ^*=\bigcup \nolimits _{k=1}^\infty \mathcal Z_\lambda ^k\). Let \(B_+^1(T,{\mathcal {C}})\) be the set of bounded measurable functions \(\chi :T\rightarrow {\mathbb {R}}\) with \(0\le \chi \le 1\). The direction space with perspective \(\lambda \), denoted \(X_\lambda \), is the closure of the image of \(B_+^1(T,{\mathcal {C}})\) in \(\prod \nolimits _{\mu \in \mathcal Z_\lambda ^*}\Lambda _\mu \), under the mapping

$$\begin{aligned} y\mapsto \left( y(\mu )\right) _{\mu \in \mathcal Z_\lambda ^*}, \end{aligned}$$
(6)

where for every \(y\in B_+^1(T,{\mathcal {C}})\) and \(\mu \in \mathcal Z_\lambda ^k\) \(y(\mu )={\left\{ \begin{array}{ll} \frac{\mu (y)}{k\overline{\mu (y)}}, &{} \mu (y)\not \in D^k\\ 0_k, &{} \mu (y)\in D^k \end{array}\right. }\). \(X_\lambda \) is thus compact and Hausdorff, and every \(x\in X_\lambda \) has the form \(x=\left( x(\mu )\right) _{\mu \in \mathcal Z_\lambda ^*}\) with \(x(\mu )\in \Lambda _\mu \) for every \(\mu \in \mathcal Z_\lambda ^*\).

3.3 Representations of values on \(\mathcal {LM}\).

Fix \(\lambda \in NA^1\). For every \(k\ge 2\), \(f\in LM^k\) and \(\mu \in \mathcal Z_\lambda ^k\) the marginal \(\partial (f,\mu ):X_\lambda \rightarrow L^\infty (\lambda )\) is given by

$$\begin{aligned} \partial (f,\mu )(x)={\left\{ \begin{array}{ll} df\left( x(\mu ),\frac{d\mu }{d\lambda }\right) , &{} x(\mu )\in \varDelta (\mu )\\ df\left( \mathbf 1 _k,x(\mu )+\mathbf 1 _k,\frac{d\mu }{d\lambda }\right) , &{} x(\mu )\in {\mathbb {S}}_\perp ^\mu \end{array}\right. }. \end{aligned}$$
(7)

Remark 1

Let \(y\in B_+^1(T,{\mathcal {C}})\), \(f\in LM^k\) and \(\mu \in \mathcal Z_\lambda ^k\), and suppose \(\mu (y)\ne 0_k\). Then

$$\begin{aligned} \partial (f,\mu )(y)=df\left( \mu (y),\frac{d\mu }{d\lambda }\right) \end{aligned}$$
(8)

with \(y\in X_\lambda \) on the left hand side of Eq. (8) and \(y\in B_+^1(T,{\mathcal {C}})\) on the right hand side of the equation. Indeed, for \(\mu (y)\not \in D^k\) we have

$$\begin{aligned} df\left( \mu (y),\frac{d\mu }{d\lambda }\right) =df\left( \frac{\mu (y)}{k\overline{\mu }(y)}, \frac{d\mu }{d\lambda }\right) =df\left( y(\mu ), \frac{d\mu }{d\lambda }\right) . \end{aligned}$$

and for \(\mu (y)\in D^k{\setminus }\left\{ 0_k\right\} \) we have

$$\begin{aligned} df\left( \mu (y),\frac{d\mu }{d\lambda }\right) =df\left( \mathbf 1 _k,\mathbf 1 _k,\frac{d\mu }{d\lambda }\right) =df\left( \mathbf 1 _k,y(\mu )+\mathbf 1 _k,\frac{d\mu }{d\lambda }\right) \end{aligned}$$

The following representation theorem will play an important role in the proof of our main result:

Theorem 3

Let \(\varphi \) be a value on \(\mathcal {LM}\). For every \(\lambda \in NA^1\) there is a finitely additive, positive vector measure \(P_\lambda \) of bounded semi-variation (i.e., \(|P_\lambda |(X_\lambda )<\infty \). See Appendix 3 for details.) on the Borel sets of \(X_\lambda \) with values in \(\mathcal L \left( L^\infty (\lambda ), L^2(\lambda )\right) \) s.t. for every coalition \(S\in {\mathcal {C}}\) the vector measure \(P_\lambda ^S=\langle P_\lambda ,\chi _S\rangle \) is positive, regular, and countably additive of bounded variation, and for every \(f\in LM^k\) and \(\mu \in \mathcal Z_\lambda ^k\) we have for every \(S\in {\mathcal {C}}\)

$$\begin{aligned} \varphi \left( f\circ \mu \right) (S)= \int _{X_\lambda }\partial (f,\mu )(x)dP_\lambda ^S(x) \end{aligned}$$
(9)

and in case that alsoFootnote 7 \(\partial (f,\mu )\in C(X_\lambda ,L^\infty (\lambda ))\) then

$$\begin{aligned} \varphi \left( f\circ \mu \right) (S)= \int _{X_\lambda }\partial (f,\mu )(x) dP_\lambda ^S(x)=\int _S\left( \int _{X_\lambda } \partial (f,\mu )(x)dP_\lambda (x)\right) d\lambda \end{aligned}$$
(10)

Theorem 3 is proved in Appendix 1.

Lemma 1

For every \(\lambda \in NA^1\) and \(\chi \in L^\infty (\lambda )\) the vector measure \(P_\lambda \) in Theorem 3 satisfies

$$\begin{aligned} \langle \chi ,P_\lambda \rangle (E)=\chi \langle 1,P_\lambda \rangle (E). \end{aligned}$$
(11)

for every Borel set \(E\subseteq X_\lambda \).

Lemma 1 is proved in Appendix 1.

3.4 Values of exact market games

For every \(k\ge 2\) denote by \(\varDelta ^k\) the \(k-1\) dimensional simplex in \({\mathbb {R}}_+^k\). The space of exact market games, \(\mathcal {EM}\), is the linear space spanned by games of the form \(v(S)=\min \nolimits _{c\in C}c\cdot \mu (S)\), for a compact and convex \(C\subseteq \varDelta ^k\) and \(\mu \in \left( NA^1\right) ^k\), for some \(k\ge 2\). The following Theorem was in fact proved in Edhan (2015):

Theorem 4

The Mertens value is the unique value on \(\mathcal {EM}\).

Remark 2

As \(\mathcal {EM}\subseteq {\mathcal {M}}\) the Mertens value is well defined on \(\mathcal {EM}\) by Eq. (61) (in the Appendix). Note that in Theorem 4 we do not require the continuity axiom.

4 The proof

Let \(\varphi \) be a continuous value on \({\mathcal {M}}\). By (Edhan 2015, Corollary 1), \(\varphi \) is uniquely determined by its values on \(\mathcal {LM}\). The restriction of \(\varphi \) to \(\mathcal {LM}\) is a continuous value on \(\mathcal {LM}\). By abuse of notation we shall denote it by \(\varphi \). Then:

Proposition 1

\(\varphi \) is the Mertens value.

To prove Proposition 1, which will in fact prove Theorems 1 and 2, we shall need the following Lemmata.

Lemma 2

For every \(w\in {\mathbb {S}}_\perp ^\mu {\setminus }\{0_k\}\) and every sufficiently small \(\epsilon >0\) there is a continuous and positively homogeneous of degree 1 function \(h_\epsilon ^w:{\mathbb {R}}_+^k\rightarrow {\mathbb {R}}\) which is twice continuously differentiable on \({\mathbb {R}}_+^k{\setminus }\{0_k\}\), vanishes on the conical diagonal neighborhood \(N_\epsilon =\{x:\sum \nolimits _{\ell =1}^k|x_\ell -\overline{x}|<\epsilon k\overline{x}\}\), and for every \(z\in {\mathbb {R}}_+^k\) with \(w\cdot z\ne 0\) and every \(1\le \ell \le k\) we have \(\frac{\partial h_\epsilon ^w}{\partial x_\ell }(z)\underset{\epsilon \rightarrow 0^+}{\longrightarrow }w_\ell \), where the convergence is uniformly bounded in \(\epsilon \).

Proof

Choose a continuously twice differentiable function \(g_{\epsilon }:{\mathbb {R}}\rightarrow {\mathbb {R}}\) satisfying \(g_{\epsilon }=0\) on \([-\epsilon ,\epsilon ]\), \(g_{\epsilon }^\prime =1\) on \([-2\epsilon ,2\epsilon ]^c\), and \(0\le g_{\epsilon }^\prime \le 1\) on \({\mathbb {R}}\). Define \(h_\epsilon ^w:{\mathbb {R}}_+^k\rightarrow {\mathbb {R}}\) by

$$\begin{aligned} h_\epsilon ^w(x)={\left\{ \begin{array}{ll} k\overline{x}g_\epsilon \left( \frac{w\cdot x}{k\overline{x}}\right) , &{} x\ne 0_k\\ 0, &{} x=0_k. \end{array}\right. } \end{aligned}$$
(12)

The function \(h_\epsilon ^w\) is continuous and positively homogeneous of degree 1 on \({\mathbb {R}}_+^k\), and it is twice continuously differentiable on \({\mathbb {R}}_+^k{\setminus }\left\{ 0_k\right\} \). It also vanishes on \(N_\epsilon \); indeed if \(x\in N_\epsilon {\setminus }\left\{ 0_k\right\} \) then

$$\begin{aligned} \left| \frac{w\cdot x}{k\overline{x}}\right| =\left| w\cdot \left( \frac{x}{k\overline{x}}-\frac{1}{k}\mathbf 1 _k\right) \right| \le \sum \limits _{i=1}\left| \frac{x_i}{k\overline{x}} -\frac{1}{k}\right| <\epsilon , \end{aligned}$$
(13)

hence \(h_\epsilon ^w(x)=g_\epsilon \left( \frac{w\cdot x}{k\overline{x}}\right) =0\).

As to the convergence of the partial derivative for \(z\in {\mathbb {R}}_+^k\) with \(w\cdot z\ne 0\), notice that

$$\begin{aligned} \frac{\partial h_\epsilon ^w}{\partial x_\ell }(z)= g_\epsilon \left( \frac{w\cdot z}{k\overline{z}}\right) + \frac{k\overline{z}w_\ell -w\cdot z}{k\overline{z}} g_\epsilon ^\prime \left( \frac{w\cdot z}{k\overline{z}}\right) \underset{\epsilon \rightarrow 0^+}{\longrightarrow }w_\ell . \end{aligned}$$
(14)

Also notice that the convergence is indeed uniformly bounded w.r.t. \(\epsilon \).\(\square \)

Lemma 3

Let \(\mu \in \mathcal Z_\lambda ^k\) with \(dim(AF(\mu ))\ge 2\), and denote \(S_\mu =\{s\in T:\frac{d\mu }{d\lambda }(s)\not \in D^k\}\). Then (in the vector lattice \(L^2(\lambda )\))

$$\begin{aligned} \langle \chi _{S_\mu },P_\lambda \rangle (\{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}=0. \end{aligned}$$
(15)

Proof

There is a Borel set \(\Omega \subset {\mathbb {S}}_\perp ^\mu \) of Haar measure 1 s.t. for every \(w\in \Omega \) we have \(\langle 1,P_\lambda ^T\rangle (\{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu ,w\cdot x(\mu )=0\})=0\). Hence for every \(S\in {\mathcal {C}}\) and every \(w\in \Omega \) we have \(\langle 1,P_\lambda ^S\rangle (\{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu ,w\cdot x(\mu )=0\})=0\). Choose \(w\in \Omega \) and given \(\epsilon >0\) let choose the function \(h_\epsilon ^w\) as in Lemma 2. Then for any sufficiently small \(\epsilon >0\), \(\Vert \partial (h_\epsilon ^w,\mu )\Vert _\infty \) is uniformly bounded w.r.t. \(\epsilon \), and for every \(x\in X_\lambda \) with \(x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \) and \(w\cdot x(\mu )\ne 0\) we have \(\partial (h_\epsilon ^w,\mu )(x)\underset{\epsilon \rightarrow 0^+}{\longrightarrow }\frac{d(w\cdot \mu )}{d\lambda }\) in the \(L^\infty (\lambda )\) norm. Thus

$$\begin{aligned} \varphi (h_\epsilon ^w\circ \mu )(S)= & {} \int _{X_\lambda }\partial (h_\epsilon ^w,\mu )(x)dP_\lambda ^S(x)\nonumber \\= & {} \int _{\{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}dh_\epsilon ^w\left( x(\mu ), \frac{d\mu }{d\lambda }\right) dP_\lambda ^S(x) \underset{\epsilon \rightarrow 0^+}{\longrightarrow }\nonumber \\&\quad \left\langle \frac{d(w\cdot \mu )}{d\lambda },P_\lambda ^S\right\rangle \left( \{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}\right) , \end{aligned}$$
(16)

where the convergence in line (16) follows from the bounded convergence theorem 6 (in the Appendix). By (Edhan 2015, Theorem 1) we have \(\varphi (h_\epsilon ^w\circ \mu )(S)=0\), and by combining that with Eq. (16) we obtain for every \(w\in \Omega \)

$$\begin{aligned} \forall S\in {\mathcal {C}},\;\;\;\; 0= & {} \left\langle \frac{d(w\cdot \mu )}{d\lambda },P_\lambda ^S\right\rangle \left( \{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}\right) \Rightarrow \end{aligned}$$
(17)
$$\begin{aligned} 0= & {} \left\langle \frac{d(w\cdot \mu )}{d\lambda },P_\lambda \right\rangle \left( \{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}\right) , \end{aligned}$$
(18)

where the equality in line (18) holds in \(L^2(\lambda )\) and follows by taking the Radon–Nikodym derivative in line (17). By combining Eq. (11) in Lemma 1 together with Eq. (18) we obtain for every \(w\in \Omega \) and \(\lambda \)-a.e. \(s\in T\) with \(\frac{d(w\cdot \mu )}{d\lambda }(s)\ne 0\) that

$$\begin{aligned} 0=\langle 1,P_\lambda \rangle \left( \left\{ x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \right\} \right) (s). \end{aligned}$$
(19)

As \(\Omega \) is of full Lebesgue measure in \({\mathbb {S}}_\perp ^\mu \) we thus conclude that

$$\begin{aligned} 0=\langle 1,P_\lambda \rangle \left( \{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}\right) (s) \end{aligned}$$
(20)

for \(\lambda \)-a.e. \(s\in T\) with \(\frac{d\mu }{d\lambda }(s)\not \in D^k\), hence, by Lemma 1, we have (in the vector lattice \(L^2(\lambda )\))

$$\begin{aligned} 0=\langle \chi _{S_\mu },P_\lambda \rangle \left( \{x\in X_\lambda :x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}\right) . \end{aligned}$$
(21)

\(\square \)

It is sufficient to verify Proposition 1 for games \(f\circ \mu \) with \(f\in LM_+^k\) and \(\mu \in \left( NA^1\right) ^k\) with \(dim(AF(\mu ))\ge 2\). For such f let \(h(f):{\mathbb {R}}_+^k\rightarrow {\mathbb {R}}\) be given by \(h(f)(x)=df(\mathbf 1 _k,x)\). Then \(h(f)\circ \mu \in \mathcal {EM}\). Hence, for every \(S\in {\mathcal {C}}\)

$$\begin{aligned} \varphi (f\circ \mu )(S)&=\int \limits _{X_{\overline{\mu }}}\partial (f,\mu )(x)dP_{\overline{\mu }}^S(x)\nonumber \\&=\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}df\left( x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)\nonumber \\&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\in {\mathbb {S}}_\perp ^\mu \}}df\left( \mathbf 1 _k,x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)\nonumber \\&=\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}df\left( x(\mu ),\frac{d\mu }{d\overline{\mu }}\chi _{S_\mu ^c}\right) dP_{\overline{\mu }}^S(x)\nonumber \\&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}df\left( x(\mu ),\frac{d\mu }{d\overline{\mu }}\chi _{S_\mu }\right) dP_{\overline{\mu }}^S(x) \end{aligned}$$
(22)
$$\begin{aligned}&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\in {\mathbb {S}}_\perp ^\mu \}}df\left( \mathbf 1 _k,x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x) \nonumber \\&=\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}df\left( x(\mu ),\chi _{S_\mu ^c}\mathbf 1 _k\right) dP_{\overline{\mu }}^S(x) \end{aligned}$$
(23)
$$\begin{aligned}&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( \mathbf 1 _k,x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)\nonumber \\&\ge \int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( x(\mu ),\chi _{S_\mu ^c}\mathbf 1 _k\right) dP_{\overline{\mu }}^S(x) \end{aligned}$$
(24)
$$\begin{aligned}&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( \mathbf 1 _k,x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)\nonumber \\&=\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( x(\mu ),\frac{d\mu }{d\overline{\mu }}\chi _{S_\mu }\right) dP_{\overline{\mu }}^S(x)\nonumber \\&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( x(\mu ),\chi _{S_\mu ^c}\mathbf 1 _k\right) dP_{\overline{\mu }}^S(x) \end{aligned}$$
(25)
$$\begin{aligned}&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( \mathbf 1 _k,x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)\nonumber \\&=\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\not \in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)\nonumber \\&\quad +\int \limits _{\{x\in X_{\overline{\mu }}:x(\mu )\in {\mathbb {S}}_\perp ^\mu \}}dh(f)\left( \mathbf 1 _k,x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)\end{aligned}$$
(26)
$$\begin{aligned}&=\int \limits _{X_{\overline{\mu }}}dh(f)\left( \mathbf 1 _k,x(\mu ),\frac{d\mu }{d\overline{\mu }}\right) dP_{\overline{\mu }}^S(x)= \varphi (h(f)\circ \mu )(S)\nonumber \\&=\Psi _M(h(f)\circ \mu )(S)=\Psi _M(f\circ \mu )(S), \end{aligned}$$
(27)

where the first equality in line (22) follows from Theorem (3), the equality in line (23) follows by combining Lemma 3 with the definition of h(f), the inequality in line (24) follows as \(df(z,\mathbf 1 _k)\ge f(\mathbf 1 _k)=h(f)(\mathbf 1 _k)=dh(f)(z,\mathbf 1 _k)\) for every \(z\in {\mathbb {R}}_{+}^k{\setminus } D^k\), the equality in line (25) follows from Lemma 3, the equality in line (26) follows as \(dh(f)(z,x)+dh(f)(z,a\mathbf 1 _k)=dh(f)(z,x+a\mathbf 1 _k)\) for every \(z\in {\mathbb {R}}_+^k{\setminus } D^k\), \(x\in {\mathbb {R}}_+^k\), and \(a\in {\mathbb {R}}_+\), the first equality in line (27) follows from Theorem 3, and the second equality in line (27) follows from Theorem 4.

Now, by the efficiency axiom we obtain for every \(S\in {\mathcal {C}}\)

$$\begin{aligned} \varphi (f\circ \mu )(S)=\Psi _M(f\circ \mu )(S), \end{aligned}$$
(28)

and we are done.

5 An alternative axiomatization

As mentioned in the introduction, the axioms of positivity and continuity [axioms (4) and (5) respectively] can be replaced by the axiom of contraction [axiom (\(4^\prime \))]. Indeed, suppose that a map \(\Psi :{\mathcal {M}}\rightarrow FA_+\) obeys the axioms of efficiency and contraction. Then \(\Psi \) is obviously continuous. Furthermore, by (Aumann and Shapley 1974, Proposition 4.6) \(\Psi \) also obeys the positivity axiom.