You are on page 1of 20

CHAPTER 16

Weighted Voting Games

Georgios Chalkiadakis and Michael Wooldridge

16.1 Introduction

In this chapter, we consider weighted voting games: a form of social choice system
that seems much simpler than most schemes considered in this handbook, but which
is, nevertheless, widely used for many important real-world social choice problems.
There are at least two very good reasons for studying weighted voting systems: first,
as we have already mentioned, they are widely used in practice; and second, for all
their apparent simplicity, they possess interesting mathematical and computational
properties, making them interesting objects from the point of view of scientific study.
Weighted voting games originated in the domain of cooperative game the-
ory (Chalkiadakis et al., 2011). They model decision-making situations in which a
set of voters must make a binary (yes/no) decision on some particular issue; each
voter is allocated a numeric weight, and the decision is carried if the sum of weights
of voters in favour of it meets or exceeds some specific given threshold, called the
quota. Weighted voting games have many applications beyond social choice theory.
For example, they can be used to model settings where each player has a certain amount
of a given resource (say, time, money, or manpower), and there is a goal that can be
reached by any coalition that possesses a sufficient amount of this resource.
The remainder of this chapter is structured as follows.
r In Section 16.2, we present the basic models and solution concepts that will be used
throughout the remainder of the chapter. We start by defining cooperative games in
characteristic function form, and then introduce weighted voting games. We go on to
describe some key solution concepts for cooperative games: the core, the Shapley value,
and the Banzhaf index.
r In Section 16.3, we consider computational properties of weighted voting games: in
particular, the complexity of computing the core, the Shapley value, and the Banzhaf
index.
r In Section 16.4, we consider the (sometimes unintuitive) relationship between the weight
that a voter is assigned in a weighted voting game and the power that this voter then
wields.

377

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


378 16 weighted voting games

r In Section 16.5, we consider the extent to which weighted voting games can be considered
as a representation scheme for yes/no voting systems (i.e., simple cooperative games).
We give a condition on yes/no voting systems that is both necessary and sufficient for
such a system to be representable as a weighted voting game.

Throughout the chapter, we assume familiarity with the basic notation and terminology
of computational complexity theory (big-O notation, the classes P, NP, coNP, and
#P) (Garey and Johnson, 1979; Papadimitriou, 1994).

16.2 Basic Definitions

Weighted voting games are a special case of a class of cooperative games, and the solu-
tion concepts that we consider for weighted voting games are in fact those developed
within cooperative game theory. In this section, we begin by presenting the key game
model from cooperative game theory, and then introduce weighted voting games as a
special case of such game. We then define the key solution concepts for such games.

16.2.1 Cooperative Games


We start by defining the game model that underpins weighted voting games: formally,
these are cooperative games with transferable utility in characteristic function form,
but we will refer to them as cooperative games.

Definition 16.1. A cooperative game, G, is given by a pair G = (N, v), where N =


{1, . . . , n} is the set of players of the game, and v : 2N → R is the characteristic
function of the game. Unless otherwise stated, we assume that v(∅) = 0, and moreover
that v({i}) = 0 for all i ∈ N. We will say a cooperative game G = (N, v) is simple if
v(C) ∈ {0, 1} for all C ⊆ N, in which case we say C ⊆ N are winning if v(C) = 1,
and losing otherwise. A simple game is nontrivial if v(N) = 1. We will usually assume
games are nontrivial.

The basic idea behind this model is that any subset C ⊆ N of the players can cooperate
with each other, and, by joining forces, they can obtain the value v(C). The model
does not specify how the players cooperate; it only specifies what value they could
obtain through cooperation. It is conventional to refer to (sub)sets of players C ⊆ N
as coalitions: in everyday use the term “coalition” implies a collection of people with
some common cause or commitment to joint action, but we will mean it simply in the
sense of a set of players, who may or may not choose to cooperate. We will refer to the
set of all players N as the grand coalition.

16.2.2 Weighted Voting Games


We are now ready to define weighted voting games. A weighted voting game is a simple
cooperative game (i.e., a game in which every coalition has a value of either 0 or 1),
which is defined by each player in the game having a weight, and where a coalition
obtains the value 1 if the sum of their values meets or exceeds a given quota. Formally:

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .2 basic definitions 379

Definition 16.2. A weighted voting game G with a set of players N = {1, . . . , n} is


given by a list of weights w = (w1 , . . . , wn ) ∈ Rn and a quota q ∈ R; we will write
G = [N; w; q]. The characteristic function v : 2N → {0, 1} of the game is defined as
follows:


1 if i∈C wi  q
v(C) =
0 otherwise.

 We write w(C) to denote the total weight of a coalition C, that is, we set w(C) =
i∈C wi . It is usually assumed that all weights and the quota are nonnegative; in what
follows, we will make this assumption as well. Under this assumption, weighted voting
games are monotone: if a coalition wins, then adding players to the coalition cannot turn
it into a losing coalition. It is common to assume that 0 < q ≤ w(N); this condition
ensures that the empty coalition is losing and the grand coalition is winning (and hence
the game is nontrivial).
Our definition of weighted voting games allows for arbitrary real number weights.
However, it is not clear how to efficiently store and manipulate such weights, which
presents a difficulty from the algorithmic point of view. Fortunately, it turns out that
any weighted voting game is equivalent to a game with fairly small integer weights.
More precisely, we have the following theorem, which follows from results on linear
threshold functions (Muroga, 1971).

Theorem 16.1. Let us say two weighted voting games G = [N; w; q] and G =
[N; w ; q  ] are equivalent iff for all coalitions C ⊆ N we have w(C)  q iff w (C)  q  .
Then, for any weighted voting game G = [N; w; q] with |N| = n, there exists an equiv-
alent weighted voting game G = [N; w ; q  ] with w = (w1 , . . . , wn ) such that all wi ,
i = 1, . . . , n, and q  are nonnegative integers, and wi = O(2n log n ) for all i = 1, . . . , n.

We can therefore assume without loss of generality that all weights and the quota
are integers given in binary. We remark that, even though the entries of the weight
vector w are exponential in n, they can be represented using O(n log n) bits—that is,
a weighted voting game with n players can be described using poly(n) bits. Thus, any
weighted voting game has an equivalent “compact” representation: this fact is useful
when considering questions relating to computational complexity.
Before proceeding, let us say a few words about the applications of weighted voting
games. These games can be used to describe many real-world situations. In particular,
they are very well suited to model coalition formation in legislative bodies. In more
detail, each party in a parliament can be associated with a player in a weighted voting
game; the weight of the player is given by the total number of the representatives of this
party. The quota corresponds to the number of votes needed to pass a bill; while in most
cases a simple majority q = !w(N)/2" + 1 suffices, in some circumstances a bill can
only be passed if it is supported by two thirds of the legislators (i.e., q = 2w(N)/3),
or even the entire legislative body (q = w(N)). Another example of a weighted voting
game is shareholder voting: the weight of each voter is proportional to the number of
shares she holds. A weighted voting game also arises in a setting where there is a task
that requires q hours of work, and there is a group of agents N = {1, . . . , n} such that
each agent i can dedicate wi of his time to this task.

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


380 16 weighted voting games

16.2.3 Solution Concepts


The basic assumption in cooperative game theory is that players will make strategic
decisions about who they will cooperate with. The best-known solution concept in
cooperative game theory, the core, formalizes an answer to the question of which
players will cooperate. The idea of the core is that a stability is a necessary condition
for coalition formation: a coalition will not form if some sub-coalition can do better by
defecting and working together as a team. To formulate the core, we need some further
definitions.

Definition 16.3. An imputation for a cooperative game G = (N, v) is a tuple of real


numbers x = (x1 , . . . , xn ) ∈ Rn satisfying the following conditions:
r n x = v(N )
i=1 i
r x  v({i}) for all i ∈ N
i

Where x = (x1 ,
. . . , xn ) is an imputation and C ⊆ N is a set of players, we denote by
x(C) the value i∈C xi .

We think of an imputation as an indication of how the value of the grand coalition should
be divided among players in the game. The first requirement relates to efficiency: it
simply says that the total value available should be distributed. The second requirement
relates to individual rationality: it says that no player should receive a payoff that is
lower than it could obtain alone.
The core then attempts to characterize a set of “acceptable” imputations: imputations
to which no coalition could realistically object. We will see the formal definition first,
and then discuss it.

Definition 16.4. The core, C(G), of a cooperative game G = (N, v) is the set:

C(G) = {x | ∀C ⊆ N : x(C)  v(C)}.

Thus, if an imputation x is not in the core, then there exists some coalition C ⊆ N
such that v(C) > x(C). Such a coalition would object to the imputation x, because they
could do better by working together as a team: the value they would obtain in this way
could be distributed among the members of C in such a way that all the players in C
receive a higher payoff than they do under the imputation x. Note that the core of a
game may be empty, as the following example illustrates.

Example 16.2. Consider a game G with N = {1, 2, 3} and the characteristic function
v defined as follows:
1 if |C|  2
v(C) =
0 otherwise.
Thus, a coalition obtains a value of 1 iff the coalition contains at least 2 members.
Now consider any imputation for the game. The imputation (1/3, 1/3, 1/3) is not in
the core, because any pair of agents could defect, obtaining the value 1 which can
be shared among themselves. No other imputation x = (x1 , x2 , x3 ) can be in the core
because two players will always be able to defect and share the value assigned to the

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .2 basic definitions 381

third player among themselves. For example, if x = (0, 0.5, 0.5) then players 1 and 2
could benefit by defecting.
The core formalizes the idea of the grand coalition being stable, in the sense that no
coalition can profitably defect from it. However, it is easy to construct examples where
the core contains imputations that seem unreasonable:
Example 16.3. Suppose G = ({1, 2}, v) is such that:
v({1}) = v({2}) = 5
v({1, 2}) = 20.
The reader may easily verify that the imputation (5, 15) is in the core, but from the
point of view of player 1, this seems unreasonable: it gives all the surplus obtained by
cooperation to player 2, while there seems to be nothing in the game to distinguish the
contribution that player 1 makes from the contribution that player 2 makes.
Thus, the core is not appropriate as a framework for deciding how to divide coali-
tional value. The Shapley value is the best known solution concept for this purpose.
Formulated by Nobel Laureate Lloyd Shapley in the early 1950s (Shapley, 1953), the
Shapley value is defined in terms of the marginal contribution that players make in
games. Formally, where C ⊆ N \ {i}, the marginal contribution that player i makes to
C is denoted by δi (C), and is simply the value that i would add to the coalition C by
joining it:

δi (C) = v(C ∪ {i}) − v(C).


We will say a player is a dummy if δi (C) = 0 for all C ⊆ N \ {i}. Thus, a dummy player
is a player that never adds any value to a coalition. We will say players i = j ∈ N are
symmetric if δi (C) = δj (C) for all C ⊆ N \ {i, j }. Thus, symmetric players are players
who always make identical contributions to coalitions. We let  denote all possible
orderings of the players N, and denote members of  by π, π  , . . . and so on. Where
π ∈  and i ∈ N, we denote by πi the set of players that precede i in the ordering π.
Then:

Definition 16.5. The Shapley value for a game G is the imputation ϕ(G) =
(ϕ1 (G), . . . , ϕn (G)) defined as follows:
1 
ϕi (G) = δi (πi ). (16.1)
n! π∈

Thus, player i’s Shapley value is the average marginal contribution that player i
makes, over all possible orderings of the players, to the set of players preceding i in
the ordering.1 Before proceeding, we note that the Shapley value can be presented in
several different ways; one important equivalent formulation is as follows:
 (n − |S| − 1)!|S|!
ϕi (G) = δi (S). (16.2)
S⊆N\{i}
n!

1 Put otherwise, pick a value k = 0 . . . n − 1 uniformly at random, and then select a set S of size k uniformly at
random from the possible subsets of N \ i. The expected marginal contribution of i to S is the Shapley value.

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


382 16 weighted voting games

This latter formulation is less common than the former, but we will make use of it in
one of our proofs.
Now, much of the interest in the Shapley value derives from the fact that it can be
shown to be the unique solution to a set of axioms characterizing fair imputations—that
is, fair ways to divide the value v(N) among the players N. To define these axioms,
we need a little more notation. Where G = (N, v) and G = (N, v  ) are games with
the same set of players, the game G + G is the game with the same player set, and
characteristic function v  : 2N → R defined by v  (C) = v(C) + v  (C). We can now
state Shapley’s characterization of the value.

Theorem 16.4 (Shapley). For all games G, the ϕ(G) payoff division scheme satisfies
the following properties:
r Efficiency: n ϕ (G) = v(N )
i=1 i
r Dummy Player: If i is a dummy player in G then ϕ (G) = 0
i
r Symmetry: If i and j are symmetric in G then ϕ (G) = ϕ (G)
i j
r Additivity: The value is additive over games: For all games G = (N, v) and G =
(N, v  ), and for all players i ∈ N , we have ϕi (G + G ) = ϕi (G) + ϕi (G )

Moreover, ϕ(G) is the only payoff division scheme that simultaneously satisfies all
these properties.

In weighted voting game settings, the Shapley value has an important interpretation: it
essentially measures the probability that a player will be able to turn a losing coalition
into a winning one—that is, the probability that a player is pivotal for a coalition. As
such, the Shapley value can be interpreted as a measure of the power wielded by a
voter: the greater the Shapley value of a player, the more power that player wields.
When used to analyze the power of voters in social choice settings, the Shapley value
is called the Shapley-Shubik voting index.
A closely related, but simpler measure of voting power is the Banzhaf index (Banzhaf,
1965; Dubey and Shapley, 1979). The Banzhaf index for a player i in game G is denoted
βi (G), and is defined as follows:
1 
βi (G) = δi (C).
2n−1 C⊆N\{i}

Thus, βi (G) measures the probability that i will be able to turn a losing coalition into
a winning one, assuming that all members of that coalition have already assembled
(the more “refined” Shapley value considers all possible ways that the coalition could
form). The Banzhaf index has properties similar to the Shapley value, but does not
satisfy efficiency (Chalkiadakis et al., 2011, p. 22). Two axiomatisations for this index
are provided in Dubey and Shapley (1979) and in Lehrer (1988).
We end this section by noting that the Shapley value and the Banzhaf index are
not the only measures of power used in (weighted) voting games. Examples of other
power indices include the Deegan-Packel index (Deegan and Packel, 1978), which
attempts to measure a player’s expected gain from participating in minimal-sized
winning coalitions; the Public Good index (Holler, 1982), which measures a player’s
power as the number of all minimal winning coalitions in which the player participates,

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .3 basic computational properties 383

divided by the sum of all such numbers over all players; and the Coleman initiative
and preventive power indices (Coleman, 1971), which, respectively, correspond to the
fraction of losing coalitions that a given player can turn into winning, and vice versa.

16.3 Basic Computational Properties

In this section, we consider computational questions associated with weighted voting


games. The main questions relate to the solution concepts for cooperative games—
the core and the Shapley value—that we discussed earlier. However, we will start by
considering some simpler properties. First, consider the following definition:
Definition 16.6. Let G = (N, v) be a cooperative game, and let i be a player in G.
Then i is said to be a veto player in G if v(C) = 0 for all C ⊆ N \ {i}.
Thus, a veto player is one whose presence is necessary for a coalition to be winning.
The obvious decision problem associated with veto players is as follows:
Name: Veto Player.
Instance: Weighted voting game G = [N ; w; q], player i ∈ N .
Question: Is i a veto player in G?

We start our survey of computational problems with some good news:


Lemma 16.5. The Veto Player problem is polynomial time solvable for weighted
voting games.
Proof. Given a player i, it suffices check whether the coalition N \ {i} is losing, which
is simple arithmetic.
The significance of this result is made apparent by the following:
Theorem 16.6. A simple cooperative game G = (N, v) has a nonempty core iff it
has a veto player. It follows that it is possible to check in polynomial time whether a
weighted voting game has a nonempty core.
Proof. For the left-to-right implication of the first statement, suppose for sake of
contradiction that G has a nonempty core but no veto player. Take an imputation x in
the core. Because x(N) = 1, we have xi > 0 for some i ∈ N, and hence x(N \ {i}) =
1 − xi < 1. However, because i is not a veto player then v(N \ {i}) = 1 > x(N \
{i}); contradiction. Now for the right-to-left implication. Suppose i is a veto player,
and consider the imputation in which xj = 0 for all j = i. This imputation is in the
core: because i is a veto player, any deviating coalition would have value 0, and so
no such coalition could profitably deviate. The second statement then follows from
Lemma 16.5.
Now consider the following problem.
Name: Dummy Player.
Instance: Weighted voting game G = [N ; w; q], player i ∈ N .
Question: Is i a dummy player in G?

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


384 16 weighted voting games

In contrast to checking for veto players, identifying dummy players is compu-


tationally hard. To show this, we provide a reduction from the classic Partition
problem (Garey and Johnson, 1979, p. 223).

Theorem 16.7. Dummy Player is coNP-complete.

Proof. If a player i is not a dummy player, this can be proved by exhibiting a coalition
C such that w(C) < q, w(C ∪ {i}) ≥ q. Thus, Dummy Player is in coNP. To show
coNP-hardness, we will transform an instance I = (a1 , . . . , ak , K) of Partition into
a weighted voting game G = [N; w; q], which is constructed as follows. We set N =
{1, . . . , k, k + 1}, wi = 2ai for i = 1, . . . , k, wk+1 = 1, and q = 2K + 1. It is easy to
see that I is a “yes”-instance of Partition if and only if player k + 1 is not a dummy
in G. Indeed, suppose that there is a subset of indices J such that i∈J ai = K.
Then we have w(J ) = 2K, w(J ∪ {k + 1}) = 2K + 1 = q, that is, k + 1 is pivotal for
J ∪ {k + 1}. Conversely, suppose that k + 1 is pivotal for some coalition C. Then we
have w(C \ {k + 1}) < q, w(C) ≥ q. Becausewk+1 = 1 and all weights are integer,
this means that w(C \ {k + 1}) = 2K, that is, i∈C\{k+1} ai = K. We conclude that I
is a “yes”-instance of Partition if and only if (G, k) is a “no”-instance of Dummy
Player; and thus, Dummy Player is coNP-hard.

Another important observation is that in our hardness reduction, the weights are
derived from the numbers in the instance of Partition. Thus, our hardness result is
relevant only if the weights are fairly large. Put differently, we have shown that Dummy
Player is unlikely to admit an algorithm that runs in time polynomial in the input
size, that is, poly(n, log wmax ). Now, in some applications of weighted voting games,
such as shareholder voting, the weights (i.e., numbers of shares held by individual
shareholders) can be huge. However, in parliamentary voting the weight of each party
is usually fairly small: for instance, at the time of writing the U.K. House of Commons
has 650 seats, and the Hellenic Parliament has 300 seats. In such settings, we might
be satisfied with a pseudopolynomial algorithm, that is, an algorithm that runs in time
poly(n, wmax ), or, equivalently, runs in polynomial time if all numbers in the input are
given in unary. It turns out that Dummy Player admits such an algorithm. This will
follow from Theorem 16.9, which shows that each player’s Shapley value in a weighted
voting game can be computed in pseudopolynomial time. Indeed, because weighted
voting games are monotone, a player is a dummy in a weighted voting game if and
only if her Shapley value is 0. Therefore, an algorithm for computing players’ Shapley
values can be used to identify dummies.
In order to use the Shapley value and the Banzhaf index to measure the agents’ power
in weighted voting games, we would like to have an efficient algorithm for computing
these indices. However, such an algorithm is unlikely to exist: Theorem 16.7, combined
with the dummy player axiom and the fact that weighted voting games are monotone,
directly implies that checking whether an agent’s Shapley value is 0 is coNP-hard. In
fact, computing the Shapley value and the Banzhaf index in weighted voting games is
computationally hard, and we will now investigate the complexity of this problem in
more detail (see also Deng and Papadimitriou, 1994; Faliszewski and Hemaspaandra,
2009; Matsui and Matsui, 2001; Prasad and Kelly, 1990).

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .3 basic computational properties 385

To understand just how hard the problem is, we need to recall some concepts from
computational complexity that are not quite as well known as notions such as NP-
hardness. We want to know how hard it is to compute the Shapley value in a weighted
voting game. Such a problem is not a decision problem like Dummy Player, because
the output of the problem is not “yes” or “no”: we want a numeric answer. In fact, it is
a counting problem. To understand what we mean by a counting problem, consider the
following problem: We are given a propositional logic formula ψ, and we are asked
how many satisfying assignments the formula has. This problem is known as #Sat.
Notice that the output of #Sat is not “yes” or “no”: it is a natural number. Also notice
that #Sat can be seen to trivially subsume the canonical NP-complete problem Sat,
because if we can count the number of satisfying assignments a formula has, we can
certainly tell whether it has at least one (which is what the Sat problem asks). The
relevant complexity class through which to understand the complexity of problems
like #Sat is called #P (Papadimitriou, 1994, p. 442). A problem is in #P if there is a
nondeterministic polynomial time Turing machine T such that the number of accepting
computations of T on a given input x gives the answer to the problem. For the case of
#Sat, for example, such a Turing machine simply guesses a valuation for the problem
instance ψ, accepts if that valuation satisfies ψ, and rejects otherwise; clearly, the
number of accepting runs is exactly the number of satisfying assignments. Hardness
can be shown in the usual way for computational complexity classes, for example by
showing how we can reduce #Sat. We will now show that computing the Shapley
value in weighted voting games is #P-complete.

Theorem 16.8 (Deng and Papadimitriou, 1994). The problem of computing the
Shapley value of a player i in a given weighted voting game G is #P-complete.

Proof. For membership, consider a Turing machine such that the computations of the
machine each correspond to an ordering π of the players; a computation accepts if
δi (πi ) = 1. Clearly, the number of accepting computations is n!ϕi (G).
For hardness, we reduce the counting version of the well-known Knapsack problem.
In this problem, we are given positive integers a1 , . . . , am and a further positive integer
 and we are asked to compute the number of subsets S of {1, . . . , m} such that
K,
i∈S ai = K. Computing the number of such sets is known to be #P-hard. In fact,
we work with a simplified version of this problem, which is also known to be #P-
hard (Papadimitriou,
m 1994). In this version it is first assumed that K = M2 , where
M = i=1 ai , and further, that all solutions have equal cardinality.
We create a weighted voting game with n = m + 1 players, as follows. For  players
w
1  i  m we set wi = ai , and we set wn = 1. The quota is set to be q = i∈N2 i .
Now, it is easy to see that for all S ⊆ N, we have v(S) − v(S \ {n}) = 1 iff the
following conditions hold:

1. n ∈ S

2. j ∈S wj > (M + 1)/2

3. j ∈S\{n} wj < (M + 1)/2

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


386 16 weighted voting games

But, because wn = 1, this is the same as saying that


 M
wj = = K.
j ∈S\{n}
2

But in this case, S \ {n} is clearly a solution to the original instance of Knapsack.
It follows from the alternate form of the Shapley value—recall Equation 16.2—that
ϕi (G) is exactly
(n − k)!(k − 1)!
n!
times the number of solutions to the instance of knapsack (recall that k is the common
size of any solution instance).
However, just as in the case of Dummy Player, these hardness results are only relevant
when the weights can be assumed to be large. For small weights both the Shapley value
and the Banzhaf index can be computed by a pseudopolynomial algorithm, as shown
by Matsui and Matsui (Matsui and Matsui, 2000).
Theorem 16.9. Given an n-player weighted voting game G = [N; w; q] and a player
i ∈ N, we can compute βi (G) and ϕi (G) in time O(n2 wmax ) and O(n3 wmax ), respec-
tively.
Proof. We will first describe the algorithm for the Shapley value; later, we will explain
how to simplify it for the Banzhaf index.
By renumbering the players if necessary, we can assume that i = n, that is, our
goal is to compute the Shapley value of the last player. We can assume without loss
of generality that wn > 0, because otherwise we clearly have ϕn (G) = βn (G) = 0.
Observe first that it suffices to determine, for each s = 0, . . . , n − 1, the number Ns
of s-element subsets of N \ {n} that have weight W ∈ {q − wn , . . . , q − 1}. Indeed,
whenever i is pivotal for a coalition C, |C| = s + 1, it is pivotal for all permutations
in which the agents in C \ {i} appear in the first s positions, followed by i; there are
exactly s!(n − s − 1)! such permutations (where we use the convention that 0! = 1).
Therefore, the formula for the Shapley value can be rewritten as

1 
n−1
ϕi (G) = s!(n − s − 1)!Ns . (16.3)
n! s=0
To compute Ns , we use dynamic programming. Specifically, we define X[j, W, s]
to be the number of s-element subsets of {1, . . . , j } that have weight W ; here, j ranges
from 1 to n − 1, s ranges from 0 to n − 1, and W ranges from 0 to w(N). For s = 0,
j = 1, . . . , n − 1, we have

1 if W = 0
X[j, W, 0] =
0 otherwise.
Furthermore, for j = 1, s = 1, . . . , n − 1 we have

1 if W = w1 and s = 1
X[1, W, s] =
0 otherwise.

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .3 basic computational properties 387

Now, having computed the values X[j  , W  , s  ] for all j  < j , all W  = 0, . . . , w(N),
and all s  = 0, . . . , n − 1, we can compute X[j, W, s] for W = 0 . . . , w(N) and s =
1, . . . , n − 1 as follows:

X[j, W, s] = X[j − 1, W, s] + X[j − 1, W − wj , s − 1].

In the preceding equation, the first term counts the number of subsets that have weight
W and size s and do not contain j , whereas the second term counts the number of
subsets of this weight and size that do contain j .
Thus, we can inductively compute X[n − 1, W, s] for all W = 0, . . . , w(N) and all
s = 0, . . . , n − 1. Now, Ns , s = 0, . . . , n − 1, can be computed as

Ns = X[n − 1, q − wn , s] + · · · + X[n − 1, q − 1, s].

By substituting this expression into Equation 16.3, we can compute the Shapley value
of player n.
The running time of this algorithm is dominated by the time needed to fill out the
table X[j, W, s]. The size of this table can be bounded by n × nwmax × n, and each
of its entries can be computed in O(1) steps, which proves our bound on the running
time.
For the Banzhaf index, the dynamic program can be simplified by omitting the third
index, s: indeed, to compute the Banzhaf index, we simply need to know how many
subsets of N \ {n} have weight that is at least q − wn and at most q − 1. This allows
us to reduce the running time from O(n3 wmax ) to O(n2 wmax ).

When the weights are large, Theorem 16.9 is not very useful, and we may want to
resort to heuristics and/or approximation algorithms for computing the power indices.
Now, Theorem 16.7 implies that, unless P= NP, no efficient algorithm can approximate
the Shapley value within a constant factor on all instances. However, it does not rule
out the existence of randomized algorithms that are probably approximately correct
(PAC), that is, produce a good solution with high probability. In fact, the good news is
that such algorithms do actually exist.
Take the recent algorithm by Bachrach et al. (2010a), for instance. The main idea
of this algorithm, dating back to Mann and Shapley (1960), is to randomly sample
a coalition and check whether the given player is pivotal for it. It is not hard to see
that the fraction of coalitions for which i is pivotal provides an unbiased estimator of
i’s Banzhaf index. Bachrach et al. show that, by averaging over poly(n, ln 1/δ, 1/)
samples, we can obtain an estimate that is within a distance  from i’s true Banzhaf
index with probability at least 1 − δ; this approach can be generalized to the Shapley
value. The accuracy and confidence achieved by use of the randomized algorithm are
typically very high, and can be achieved with only a fraction of the sample size that
is theoretically required by the approach—as verified by empirical results (Bachrach
et al., 2010a).
Many other papers consider the problem of computing power indices, either approx-
imately or exactly; an (incomplete) list includes Alonso-Meijide et al. (2012), Fatima
et al. (2008), Leech (2003), Mann and Shapley (1962), Merrill (1982), and Owen
(1975). Moreover, there are several recent attempts to efficiently compute power

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


388 16 weighted voting games

indices in restricted settings, taking advantage of specific domain characteristics (see,


e.g., Durán et al., 2003; Bachrach and Shah, 2013; See et al., 2014).

16.4 Voter Weight versus Voter Power

In weighted voting games, one might expect that there is a close correlation between
the weight wi that a player i is assigned, and the power that player i wields. After
all, a player’s ability to turn a losing coalition into a winning coalition derives from
their weight. This intuition is not entirely incorrect: it is not hard to show that power
is monotone in weight—that is, for any weighted voting game G = [N; w; q] and
any two players i, j ∈ N we have ϕi (G) ≤ ϕj (G) if and only if wi ≤ wj . However,
the power of an agent cannot be simply represented by its weight, in general: in that
sense,“weighted voting doesn’t work” (Banzhaf, 1965). Indeed, two agents may have
identical voting power even if their weights differ dramatically.
Example 16.10. After the May 2010 elections in the United Kingdom, the Conservative
Party had 307 seats, the Labour Party had 258 seats, the Liberal Democrats (LibDems)
had 57 seats, and all other parties shared the remaining 28 seats (with the most powerful
of them getting 8 seats). The U.K. Parliament is a simple majority system, so the quota
is 326. To keep things simple in this analysis, we will treat the “other” parties as a
single block (the other parties would be unlikely to act as a block in practice).
Now, it is easy to see that in this weighted voting game there are two two-party
coalitions (Conservatives+Labour and Conservatives+LibDems) that can get a majority
of seats. Moreover, if Labour or LibDems want to form a coalition that does not include
Conservatives, they need each other and the others. Thus, Labour and LibDems have the
same Shapley value, despite being vastly different in size. In more detail, the Shapley
values are as follows:

i ϕi
Conservative 1/2
Labour 1/6
LibDem 1/6
Other 1/6

The phenomenon illustrated in Example 16.10 has been observed in many decision-
making bodies. It explains why major parties often end up making concessions to
smaller parties in order to form a winning coalition: the small parties may wield
substantial voting power. Example 16.10 also indicates that to determine an agent’s
power, we have to take into account the distribution of the other players’ weights as
well as the quota. In particular, if we keep the weights fixed, but alter the quota, an
agent’s power can change considerably.
Example 16.11. Consider a weighted voting game with two players of weight 4 and
two players of weight 1. If the quota is set to 10, the only winning coalition is the grand
coalition, so each player’s Shapley value is 1/4. If the quota is set to 8, the smaller

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .4 voter weight versus voter power 389

players are dummies, so their Shapley value is 0. Finally, if the quota is set to 5, a player
of weight 1 is pivotal only if it appears in the second position, and a player of weight
4 appears in the first position. There are four permutations that satisfy this condition,
so for q = 5 the Shapley value of each of the smaller players is 1/6.
The role of the quota in determining the agents’ power in weighted voting games
was studied in detail by Zuckerman et al. (2008), and then by Zick et al. (2011).
Zuckerman et al. considered how a central authority might manipulate the quota of a
game from a worst-case point of view, and also from an algorithmic points of view.
They demonstrated that given a collection of voter weights and a specific undesirable
voter, it was possible to compute in polynomial time the quota that would minimize
the Banzhaf index of that voter. They also showed that checking whether a player was
more powerful with one quota than another, with respect to both the Shapley value
and Banzhaf index, is complete with respect to probabilistic polynomial time (i.e.,
PP-complete (Papadimitriou, 1994, p. 256)).

16.4.1 Paradoxes of Power


An agent’s Shapley value and his Banzhaf index may behave in an unexpected way if
we modify the game. For example, we might naively expect that adding players to a
game would reduce the power of players already present in the game, but this is not
necessarily the case: when a new player joins the game, the power of some existing
players may in fact increase! Consider the following example.
Example 16.12. Consider a weighted voting game G = [{1, 2, 3}; (2, 2, 1); 4]. Clearly,
player 3 is a dummy in this game, so ϕ3 (G) = 0. Now, suppose that a new player with
weight 1 joins this game. In the resulting game G , player 3 is pivotal for the coalition
consisting of himself, the new player and one of the other two players, so ϕ3 (G ) > 0.
Another interesting observation is that, when a player i in a game G splits into two
different players—that is, distributes his weight between two identities i  and i  —the
sum of the new identities’ Shapley values in the resulting game can be considerably
different from i’s Shapley value in the original game.
Example 16.13. Consider an n-player weighted voting game G = [N; w; q] with
w = (2, 1, . . . , 1) and q = n + 1. In this game the only winning coalition is the grand
coalition, so ϕi (G) = n1 for all i = 1, . . . , n. Now, suppose that player 1 decides to
split into two unit-weight players 1 and 1 . In the resulting game G = [N \ {1} ∪
{1 , 1 }; (1, . . . , 1); n + 1] all players are symmetric, and therefore have equal Shapley
1 2
value, namely, n+1 . Thus, the joint power of the two new identities of player 1 is n+1 ,
that is, almost twice his original power!
However, weight-splitting may also lead to a reduction in power. To see this, consider
an n-player weighted voting game G = [N; w; q] with w = (2, 2, . . . , 2) and q =
2n − 1. By symmetry, we have ϕi (G) = n1 for all i = 1, . . . , n. However, it can be
shown that if player 1 splits into two unit-weight players 1 and 1 , the sum of their
Shapley values in the new game G = [N \ {1} ∪ {1 , 1 }; (1, 1, 2, . . . , 2); 2n − 1] is
2
only n(n+1) ; the proof of this fact is left as an exercise for the reader. Thus, weight-
splitting lowers the agent’s power by a factor of (n + 1)/2.

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


390 16 weighted voting games

The phenomena illustrated in Examples 16.12 and 16.13 are known as the paradox
of new members and the paradox of size, respectively. There are several other forms of
counterintuitive behavior exhibited by the power indices; jointly, they are referred to as
the paradoxes of power. These paradoxes are discussed by Felsenthal and Machover
(1998) and subsequently by Laruelle and Valenciano (2005) (see also the references
therein). The paradox of size is studied in detail by Aziz et al. (2011), who show
that the games described in Example 16.13 exhibit the strongest form of this paradox
possible: in an n-player weighted voting game, splitting into two identities can increase
(respectively, decrease) an agent’s Shapley value by at most a factor of 2n/(n + 1)
(respectively, (n + 1)/2). Aziz et al. (2011) also show that deciding whether a given
player can split so as to increase his power is NP-hard. Rey and Rothe (2014) show
similar hardness results for coalitional splitting, merging and annexation in weighted
voting games.

16.5 Simple Games and Yes/No Voting Systems

At this point, let us step back from considering weighted voting games specifically,
and consider the wider class of social choice systems of which weighted voting games
are an instance. Yes/No voting systems are voting systems in which a proposal (e.g.,
a new law, or a change to tax regulations) is pitted against the status quo (Taylor and
Zwicker, 1999). Decision making in many political bodies can be understood as a
yes/no voting system (e.g., in the United Kingdom, the voting system of the House
of Commons; in the European Union, the voting system in the enlarged EU; in the
United States, the U.S. federal system; in the United Nations, the voting system of
the Security Council (Taylor, 1995)). Formally, yes/no voting systems have a very
simple structure:

Definition 16.7. A yes/no voting system is a pair Y = (N, W ), where N = {1, . . . , n}


is the set of voters, and W ⊆ 2N is the set of winning coalitions, with the intended
interpretation that, if C ∈ W , then C would be able to determine the outcome (either
“yes” or “no”) to the question at hand, should they collectively choose to.

Notice that yes/no voting systems can alternatively be understood as simple coopera-
tive games. Formally, we can understand a simple game G = (N, v : 2N → {0, 1}) as
defining a yes/no voting system YG = (N, WG ) in which

WG = {C ⊆ N | v(C) = 1}.

Results relating to simple games can thus be applied directly to yes/no voting sys-
tems, and we can therefore choose to work with whichever model we find most
convenient.
Several possible conditions on yes/no voting games suggest themselves as being
appropriate for some (though of course not all) scenarios:
r Nontriviality: There are some winning coalitions, but not all coalitions are winning—
formally, ∅ ⊂ W ⊂ 2N .

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .5 simple games and yes/no voting systems 391

r Monotonicity: If C wins, then every superset of C also wins—formally, if C ⊆ C and


1 2
C1 ∈ W then C2 ∈ W .
r Zero-sum: If a coalition C wins, then the agents outside C do not win—formally, if
C ∈ W then N \ C ∈ W .
r Empty coalition loses: ∅ ∈ W .
r Grand coalition wins: N ∈ W .

Of course, these conditions are not independent: the final two conditions imply the
first, for example.
Now, we can think of weighted voting games as being a compact representation for
yes/no voting systems; the set W is simply
W = {C ⊆ N | w(C)  q}.
The representation is compact because we do not need to explicitly list all winning
coalitions; we simply specify the weights and quota. This suggests an interesting
question: Are weighted voting games a complete representation for yes/no voting
systems? Is it the case that, for every yes/no voting system Y = (N, W ) we can find a
weighted voting game G with player set N such that W is exactly the set of winning
coalitions in G? The answer is no:
Theorem 16.14. There are yes/no voting systems for which there exist no weighted
voting game with the same set of winning coalitions.
Proof. Consider a yes/no voting system Y = (N, W ) such that C ∈ W iff C contains
an odd number of players. Take an odd number k and consider the coalition C =
{1, . . . , k}. By definition we have
1. C∈W
2. C ∪ {k + 1} ∈ W
3. C ∪ {k + 2} ∈ W
4. C ∪ {k + 1, k + 2} ∈ W
But now consider the weight wk+2 : the first and third conditions imply wk+2 < 0,
while the second and fourth imply wk+2 > 0. There can thus be no weighted voting
game representing the yes/no voting system in which the winning coalitions are those
containing an odd number of players.
This raises a further interesting question: Can we identify a condition on yes/no
voting systems that is necessary and sufficient to ensure that they can be represented
by a weighted voting game? The answer is yes. The property of trade robustness was
identified by Taylor and Zwicker as a necessary and sufficient condition for a yes/no
voting system to be representable as a weighted voting game (Taylor and Zwicker,
1999). “Trade robustness” is easily explained. Suppose we have some collection of
disjoint coalitions C1 , . . . , Ck with members drawn from the set of voters N. Now
discard the losing coalitions from this list—that is, the coalitions Ci such that Ci ∈ W —
leaving us with winning coalitions only. Let us then say a trade is an exchange of players
between two of the winning coalitions. For example, it may be that coalition C1 gives
players {3, 4} to coalition C2 , and in return coalition C2 gives player 7 to C1 . Of course,
for such a trade to be “legal,” we would have to have {3, 4} ⊆ C1 and 7 ∈ C2 . (We will

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


392 16 weighted voting games

permit the possibility of empty sets of players being transferred in a trade.) Then we
have:
Definition 16.8. A yes/no voting game is said to be trade robust if, after an any legal
sequence of trades starting with a set of winning coalitions, at least one of the resulting
coalitions is still winning.
We have:
Theorem 16.15 (Taylor and Zwicker, 1999). For every yes/no voting system Y =
(N, W ), there exists a weighted voting game that is equivalent to Y iff Y is trade robust.
To further illustrate trade robustness, let us consider a richer example.
Example 16.16. Consider a game G = (N, v) with N = {1, 2, 3, 4} given by
v(C) = 0, if |C| ≤ 1
v(C) = 1, if |C| ≥ 3
v({1, 2}) = v({3, 4}) = v({1, 4}) = v({2, 3}) = 1,
v({1, 3}) = v({2, 4}) = 0.
This game is not trade robust, because when two winning coalitions C1 = {1, 2} and
C2 = {3, 4} trade members (2 moves from C1 to C2 and 3 moves from C2 to C1 ), both
of the resulting coalitions are losing.
Notice, however, G can be represented as an intersection of two weighted voting
games, in the following sense. Let w1 = (1, 0, 1, 0), w2 = (0, 1, 0, 1), q 1 = q 2 = 1,
and set G1 = [N; w1 ; q 1 ], G2 = [N; w2 ; q 2 ]. Observe that a coalition C is winning
in both G1 and G2 if and only if it contains both an even-numbered player and an
odd-numbered player—that is, if and only if v(C) = 1.
The construction presented in Example 16.16 can be used to describe other simple
games in the language of weighted voting games. That is, we can take k ≥ 1 weighted
voting games G1 , . . . , Gk over the same set of players N, and define a new simple
game G = (N, v) by setting v(C) = 1 if and only if C is a winning coalition in each of
the underlying games G1 , . . . , Gk . It turns out that any yes/no game can be obtained
in this manner.
Theorem 16.17. For any simple game G = (N, v), there exists a list of weighted
voting games G1 , . . . , Gk , where Gj = [N; wj ; q j ] for j = 1, . . . , k such that for any
coalition C ⊆ N it holds that v(C) = 1 if and only if wj (C) ≥ q j for each j = 1, . . . , k.
Proof. Let C 1 , . . . , C k be the list of losing coalitions in G. We define the j th weighted
j j j
voting game Gj by setting wi = 0 if i ∈ C j and wi = 1 if i ∈ C j (here wi denotes
the weight of the ith player in Gj ) and q j = 1. That is, a coalition C is winning in
Gj if and only if it contains an agent i ∈ C j , or, equivalently, if and only if it is not a
subset of C j .
Consider a coalition C with v(C) = 1. By monotonicity, C is not a subset of any
losing coalition, so we have wj (C) ≥ q j for any j = 1, . . . , k. On the other hand, if

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


1 6 .5 simple games and yes/no voting systems 393

v(C) = 0, then C = C j for some j = 1, . . . , k, so wj (C) = 0 < q j . Thus, the theorem


is proved.

Games that are represented as intersections of k weighted voting games are


known as k-weighted voting games or vector weighted voting games. More for-
mally, a k-weighted voting game is given by a set of players N, |N| = n, for
each player i ∈ N, a k-dimensional weight vector wi = (wi1 , . . . , wik ) whose entries
are nonnegative integers, and k nonnegative integer quotas q 1 , . . . , q k ; we write
, wn ; q 1 , . . . , q k ]. A coalition C ⊆ N is deemed to be winning in
G = [N; w1 , . . .
j
G if and only if i∈C wi ≥ q j for all j = 1, . . . , k.
Observe that G can be associated with k weighted voting games G1 , . . . , Gk , where
j j
G = [N; (w1 , . . . , wn ); q j ]; these games are called the component games of G, and
j

the weight vector of the j th game is denoted by wj . Clearly, any vector voting game is
a simple game, and Theorem 16.17 shows that the converse is also true.
It is important to note that vector weighted voting games are not theoretical con-
structs: they feature quite prominently in our lives. For example, the following political
systems can be understood as vector weighted voting games (Bilbao et al., 2002; Taylor,
1995; Taylor and Zwicker, 1999):
r The U.S. federal system is a 2-weighted voting game, in which the components corre-
spond to the two chambers (the House of Representatives and the Senate). The players
are the president, vice president, senators, and representatives. In the game correspond-
ing to the House of Representatives, senators have zero weight, while in the game
corresponding to the Senate, representatives have zero weight. The president is the only
player to have nonzero weight in both games.
r The voting system of the European Union is a 3-weighted voting game (Bilbao et al.,
2002). Specifically, in the Council of the European Union, a law requires the support
of 50% of the countries, 62% of the population of the European Union, and 74% of the
“commissioners” of the EU. Each member state is a player, so (as of 2011) the players
are as follows:

Germany, United Kingdom, France, Italy, Spain, Poland, Romania, The Netherlands, Greece,
Czech Republic, Belgium, Hungary, Portugal, Sweden, Bulgaria, Austria, Slovak Repub-
lic, Denmark, Finland, Ireland, Lithuania, Latvia, Slovenia, Estonia, Cyprus, Luxembourg,
Malta.

The three component games in the EU voting system are shown in Figure 16.1 (we omit
the player set N = {1, . . . , 27}, as well as all brackets and parentheses, from the nota-
tion). Weights in the first game are assigned according to the number of commissioners
the respective member state has. The second game is a simple majority game: every
member state gets one vote, and a law must have the support of at least 14 member
states. In the third game, weights are assigned in proportion to the population of the
respective member state.

Theorem 16.17 allows us to represent any simple game G as a vector weighted


voting game; however, the number k of the component games can be quite large (and,
in particular, exponential in the number of players n). The number of the component

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


394 16 weighted voting games

G1 : 29 29 29 29 27 27 14 13 12 12 12 12 12 10 10 10 7 7 7 7 7 4 4 4 4 4 3 255
G2 : 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 14
G3 : 170 123 122 120 82 80 47 33 22 21 21 21 21 18 17 17 11 11 11 8 8 5 4 3 2 1 1 620

Figure 16.1. Voting in the Council of the European Union is a 3-weighted voting game.

games in a minimal such representation can be interpreted as the inherent complexity


of the game. Therefore, given a simple game G, we may be interested in finding the
smallest value of k such that G can be represented as a k-weighted voting game. This
value of k is called the dimension of G; we write dim(G) = k. We emphasize that even
if we are given a representation of G as a k-weighted voting game, this does not mean
that dim(G) = k: indeed, there may exist a k  < k such that G can be represented as a
k  -weighted voting game, so the only conclusion that we can derive is that dim(G) ≤ k.
It turns out that there exist simple games of exponentially large dimension. More
precisely, for any odd value of n there exists an n-player simple game G such that
dim(G) ≥ 2n/2−1 (see Taylor and Zwicker (1999) for a proof).
Vector weighted voting games can be interpreted as conjunctions of weighted voting
games. One can also combine weighted voting games according to more complex
Boolean formulas: for instance, to win in the game (G1 ∨ G2 ) ∧ G3 , a coalition must
win in one of the games G1 and G2 as well as in G3 . This representation for simple
games is studied in detail by Faliszewski et al. (2009a); in particular, Faliszewski et
al. show that it can be considerably more compact than the representation via vector
weighted voting games.

16.6 Conclusions

Despite their simple mathematical structure, weighted voting games are surprisingly
rich with respect to their computational and mathematical properties. In this chapter, we
have presented a survey of the key weighted voting games properties, with particular
emphasis on computing solution concepts.

16.7 Further Reading

For an outstanding introduction to cooperative game theory set in the wider context of
game theory, see Maschler et al. (2013). There seem to be relatively few texts that focus
specifically on cooperative games; some examples are Brânzei et al. (2005), Curiel
(1997), and Driessen (1988). A good introduction to the mathematics of cooperative
games is Peleg and Sudhölter (2007). Simple games in general are discussed in detail
in Taylor and Zwicker (1999). For an introduction to the computational aspects of
cooperative game theory, see Chalkiadakis et al. (2011). For a detailed survey of the
computational complexity of solution concepts for weighted voting games, see Elkind
et al. (2009b).

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


acknowledgments 395

Acknowledgments

We would like to thank Edith Elkind for her permission to adapt co-authored material
for this chapter, and in particular, material from our book (Chalkiadakis et al., 2011).
We would also like to thank our editors (in particular Jérôme Lang and Ariel Proccacia)
and our reviewers, Yair Zick and Yoram Bachrach, for their invaluable corrections and
comments. Wooldridge was supported by the European Research Council under grant
291528 (“RACE”).

https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press


https://doi.org/10.1017/CBO9781107446984.017 Published online by Cambridge University Press

You might also like