Probabilistic view of voting, paradoxes, and manipulation

By Elchanan Mossel

Abstract

The Marquis de Condorcet, a French philosopher, mathematician, and political scientist, studied mathematical aspects of voting in the eighteenth century. Condorcet was interested in studying voting rules as procedures for aggregating noisy signals and in the paradoxical nature of ranking three or more alternatives. We survey some of the main mathematical models, tools, and results in a theory that studies probabilistic aspects of social choice. Our journey will take us through major results in mathematical economics from the second half of the twentieth century, through the theory of Boolean functions and their influences and through recent results in Gaussian geometry and functional inequalities.

1. Introduction

The Marquis de Condorcet, a French philosopher, mathematician, and political scientist, studied mathematical aspects of voting in the eighteenth century. It is remarkable that already in the eighteenth century Condorcet was an advocate of equal rights for women and people of all races and of free and equal public education Reference 73. His applied interest in democratic processes led him to write an influential paper in 1785 Reference 16, where in particular he was interested in voting as an aggregation procedure and where he pointed out the paradoxical nature of voting in the presence of three or more alternatives.

1.1. The law of large numbers and Condorcet’s jury theorem

In what is known as Condorcet’s jury theorem, Condorcet considered the following setup. There are voters and two alternatives denoted (which stands for ) and (which stands for ). Each voter obtains a signal which indicates which of the alternatives is preferable. The assumption is that there is an a priori better alternative and that each voter independently obtains the correct information with probability and incorrect information with probability . The voters then take a majority vote to decide the winner. Without loss of generality, we may assume that the correct alternative is and therefore the individual signals are i.i.d. (independent and identically distributed) random variables where and . Let denote the Majority function, i.e., the function that returns the most popular values among its inputs. Condorcet’s jury theorem is the following.

Theorem 1.1.

For every :

.

If are odd, then .

The first part of the theorem is immediate from the law of large numbers (which was known at the time), so the novel contribution was the second part. In the early days of modern democracy, Condorcet used his model to argue that the more people participating in decision making, the more likely that the correct decision is arrived at. We leave the proof of the second part of the theorem as an exercise.

1.2. Condorcet’s paradox and Arrow’s theorem

As hinted earlier, things are more interesting when there are three or more alternatives. In the same 1785 paper, Condorcet proposed the following paradox. Consider three voters named , , and , and three alternatives named , , and . Each voter ranks the three alternatives in one of six linear orders. While it is tempting to represent the orders as elements of the permutation group , it will be more useful for us to use the following representation. Voter preference is given by , where if she prefers to , and otherwise; if she prefers to , and otherwise; and if she prefers to , and otherwise. Each of the six rankings corresponds to one of the vectors in .

Condorcet considered three voters, with rankings given by , , , or in our notation by the rows of the following matrix:

How should we decide how to aggregate the individual rankings? If we use the majority rule to decide between each pair of preferences, then we apply the majority rule on each of the columns of the matrix and conclude that overall preference is . In other words, the overall preference is that , the overall preference is that , and the overall preference is that . This does not correspond to an order! This is what is known as Condorcet’s paradox.

Almost two hundred years later, Ken Arrow asked if perhaps the paradox is the result of using the majority function to decide between every pair of alternatives? Can we avoid paradoxes if we aggregate pairwise preferences using a different function?

One function that never results in paradoxes is the dictator function as the aggregate ranking is .

Arrow in his famous theorem Reference 1Reference 2 proved the following.

Theorem 1.2.

Let and suppose that never results in a paradox, so for all it holds that . Then is a dictator: there exists an such that for all , or there exists an such that for all .

With the right notation and formulation, the proof of Arrow’s theorem is very short (see Reference 3Reference 50).

Proof.

First note that if is a constant function, then the outcome is always . Suppose that is not a dictator and not a constant, then depends on at least two coordinates. Without loss of generality, let these coordinates be and . Therefore,

We now choose and for . This guarantees that for all no matter what the values of and are. This can be also verified from the matrix in Equation 1. Now choose and so that results in a paradox:

Arrow also considered a more general setting where , , and are allowed to be different functions. In this case, there are other functions that never result in a paradox. For example if and for all inputs, then can be arbitrary. This corresponds to the case where the second alternative is ignored (always ranked last) and the choice between and is determined by . Arrow’s theorem in this setting can be stated as follows:

Theorem 1.3.

Let and suppose that never results in a paradox. So for all it holds that . Then either two of the functions are constants of opposite signs or there exists an such that , , and are dictators on voter , so or .

Proof.

If two of the functions, say and , take the same constant value and the third function is not constant, then clearly one can find such that and . So without loss of generality we may assume at least two of the functions, say and are not constant. Let denote the set of variables that may change the value of and similarly and . Since and are not constant, it follows that and are not empty. If there exists a variable and a variable , then by the same argument as in Theorem 1.2, there exist resulting in a paradox. Thus, the only case remaining is where and or . In either case, the functions and are all functions of variable only. It is now easy to verify that it must be the case that is a dictator on voter .

1.3. Manipulation and the Gibbard–Satterthwaite theorem

A naturally desirable property of a voting system is strategy-proofness (a.k.a. nonmanipulability): no voter should benefit from voting strategically, i.e., voting not according to her true preferences. However, Gibbard Reference 28 and Satterthwaite Reference 71 showed that no reasonable voting system can be strategy proof. Before stating the result, let us specify the model more formally.

The setting here is different than the setup of Arrow’s theorem: We consider voters electing a winner among alternatives. The voters specify their opinion by ranking the alternatives, and the winner is determined according to some predefined social choice function (SCF) of all the voters’ rankings, where denotes the set of all possible total orderings of the alternatives. We call a collection of rankings by the voters a ranking profile. We say that an SCF is manipulable if there exists a ranking profile where a voter who knows how all other voters vote, can achieve a more desirable outcome of the election according to her true preferences by voting in a way that does not reflect her true preferences.

For example, consider Borda voting, where each candidate receives a score which is the sum of its ranks, and the candidate with the lowest score wins. If the individual rankings are , then is the winner, but if the second voter were to vote instead, then will become the winner, so the second voter is incentivized to cast a vote that is different from her true ranking.

Theorem 1.4 (Gibbard and Satterthwaite).

Any SCF which is not a dictatorship (i.e., not a function of a single voter) and which allows at least three alternatives to be elected is manipulable.

This theorem has contributed to the realization that it is unlikely to expect truthfulness in voting. There are many proofs of the Gibbard–Satterthwaite theorem, but all are more complex than the proof of Arrow’s theorem given above. We will not provide proof of the theorem here.

1.4. Judgment aggregation

A recent line of work is devoted to the problem of judgment aggregation. In the legal literature, Kornhauser and Sager Reference 43 discuss a situation where three cases are considered in court, and by law, one should rule against if and only if there is a ruling against both and . When several judges are involved, their opinions should be aggregated using a function that preserves this law, that is, satisfies

where . List and Pettit Reference 45Reference 46 showed that the only nonconstant aggregation functions that satisfy Equation 2 are the AND functions, known in the social choice literature as oligarchies, i.e., functions of the form for some .

1.5. Modern perspectives

Work since the 1980s addressed novel aspects of aggregation of votes. Condorcet’s jury theorem assumes a probability distribution over the voters but is restricted to a specific aggregation function (majority) while Arrow’s theorem considers general aggregation functions but involves no probability model. There are many interesting questions that can be asked by combining the two perspectives. First, it is natural to ask about the aggregation properties of Boolean functions. The study of aggregation properties of Boolean functions was fundamental to the development of the area of analysis of Boolean functions since the 1980s, starting with the work of Ben-Or and Linial Reference 6 and of Kahn, Kalai, and Linial Reference 36. Second, we can ask questions regarding the probability of manipulation and paradoxes, questions that were analyzed since the early 2000s, starting with the works by Kalai, Nisan, Friedgut, and collaborators Reference 26Reference 37Reference 38.

The theory that was developed is intimately connected to the area of property testing in theoretical computer science and to additive combinatorics. Moreover, we will see that some of the main results and techniques have a discrete isoperimetric flavor. We discuss some of these topics and their connections as follows.

In Section 2, we will start by studying the question of noise stability of Boolean functions that were originally studied by Benjamini, Kalai, and Schramm in the context of percolation Reference 7 and which later played an important role in analyzing the probability of paradoxes, starting with the work of Kalai Reference 37. This work as well as motivation from theoretical computer science Reference 41, led to the proof of the Majority Is Stablest theorem by Mossel, O’Donnell, and Oleszkiewicz Reference 57Reference 58. The proof of these results will require some of the main analytical tools in the area, including, notably, hypercontraction Reference 4Reference 10Reference 29Reference 66, the invariance principle Reference 58, and Borell’s Gaussian noise-stability result Reference 12.

In Section 3, we sketch proofs of quantitative versions of Arrow’s theorem. We will follow Reference 52 by first proving a Gaussian version of Arrow’s theorem, as well as a quantitive version using reverse hypercontraction Reference 11. Combining the two, we will prove a general quantitive Arrow’s theorem Reference 52. We will also present Kalai’s original proof of a quantitative Arrow’s theorem Reference 37, which is less general but uses only hypercontraction via Reference 25.

Different tools were used in proving different quantitive versions of the manipulation theorem. The first proof, which applies only to three alternatives, uses a reduction to a quantitive Arrow’s theorem Reference 24Reference 26. In Section 4 we discuss later approaches that apply for any number of candidates and use reverse hypercontraction as a major tool Reference 32Reference 62. The classical proofs of manipulation theorems often use long paths of voting profiles. The most general proof in Reference 62 will quantify such arguments using geometric tools from the theory of Markov chains.

Recent work Reference 23 established quantitative versions of the result by List and Pettit by showing that if is -close to satisfying judgment aggregation, then it is -close to an oligarchy; this improved upon prior work by Nehama Reference 64 in which decays polynomially with . These results are based on the analysis of a variant of the noise operator, named the one-sided noise operator.

1.6. Probability of paradox for the majority

As a preview of what’s to come, we will compute the asymptotic probability of a nontransitive outcome in the Condorcet setup with three alternatives and where voters vote uniformly at random. The first reference to this computation is by Guilbaud (1952); see Reference 14 and Reference 5Reference 27.

Let us denote the Majority function by and assume that the number of voters is odd. Paradoxes seem more likely when there is no bias towards a particular candidate, so we will consider voters who vote independently and where voter votes uniformly at random from the six possible rankings. Recall that we encode the six possible rankings by vectors . Here is if a voter ranks above/below , is if a voter ranks above/below , is if a voter ranks above/below .

How do we analyze the probability of a paradox? The following simple fact was used in Reference 37: Since the binary predicate , can be expressed as

we can write

which due to symmetry can be written as

Moreover, the uniform distribution over satisfies and the coordinates are independent. As we will see shortly, the quantity is called the noise stability of with noise parameter . Its asymptotic value as is easy to compute using a two-dimensional central limit theorem (CLT) to obtain

where and we can see that

and

In particular, the probability of paradox does not vanish as .

2. Noise stability

2.1. Boolean noise stability

Consider the following thought experiment. Suppose the voters in binary voting obtain independent uniform signals: or with probability . This is the same setting as in Condorcet’s jury theorem except the voters are completely uninformed.

Now consider the following process that produces a vector as a noisy version of . For each independently, let with probability , and let with probability , where . We chose the parametrization so that .

How should we interpret ? A simple interpretation is as a noise process of voting machines. Suppose that when each voter votes, there is a small probability, say , that the voting machine records the opposite vote (independently for all voters and independently of the intended vote). In this case . Given a voting aggregation function , ideally we would like the quantity

to be as large as possible if and as small as possible if . The quantity is called the noise stability of . More generally, following Reference 7 we define the following.

Definition 2.1.

For two functions the (-)noisy inner product of and denoted by is defined by , where are i.i.d. mean () and -correlated (). The noise stability of is its noisy inner product with itself: .

We can also write the noisy inner product in terms of the noise operator :

Definition 2.2.

The Markov operator maps functions to functions . It is defined by

This section discusses noise stability of various families of functions and analytical properties of the operator , which will be used later.

The noise operator plays a key role in the theory of hypercontraction Reference 10. Note that

and that

Basic properties of this operator can be revealed using its eigenfunctions, i.e., the Fourier basis. The following proposition is straightforward to prove.

Proposition 2.3.

For , write , so . Then:

is an orthonormal basis for the space of all functions .

is an eigenfunction of which corresponds to the eigenvalue : .

The following folklore result is easily derived by explicitly writing the Fourier expression of in terms of the basis .

Theorem 2.4.

For every , for every , and for every with , it holds that

Moreover, the only optimizers are dictator functions, i.e., functions of the form or .

Theorem 2.4 allows a quick proof of a version of Arrow’s theorem by Kalai Reference 37:

Corollary 2.5.

In the context of Arrow’s theorem if and , then , , and are all the same dictator.

Proof.

Use the previous theorem and

From the voting perspective, there is something a little disappointing about Theorem 2.4. It says that if we want to maximize robustness, then among all balanced functions dictator is optimal. However, in the theory of voting dictators are usually not considered good voting schemes. From a mathematical perspective, it is disappointing that there is something special about . In particular the following problem is open.

Problem 2.6.

For a generic , what is the value of

To the best of our knowledge, the value is not known for all ; see Reference 75.

Here are some additional examples,

A similar argument to the theorem shows that if , then for . The parity function achieves equality.

The asymptotic noise stability of Majority is given by the Sheppard formula (see Reference 72), i.e., , where are -correlated Gaussian random variables,

In particular if , then is of order . Compare this to a dictator where it is of order .

If we consider where is odd and the function implements electoral college, i.e.,

then it is easy to see that asymptotically the noise stability is given by . In particular if for small , then is of order .

Let be the Majority function on voters, define , and by induction,

This function is called the recursive majority function. The paper Reference 56 shows that for every , if is large enough and and if , then

where are -correlated. In other words, this function is very far from being noise stable.

2.2. Gaussian noise stability

We will now take a detour and consider analogous quantities defined in Gaussian space. We will later see that this is quite useful in the Boolean setting.

Definition 2.7.

In Gaussian space, the (-)noisy inner product of and denoted by is defined as

where are i.i.d. two-dimensional Gaussian vectors, such that are standard (mean , variance ) Gaussian random variables and . The noise stability of is its noisy inner product with itself: .

We will generally use , , etc., to denote functions over the Boolean cube and , , etc., for functions in . In particular, for we write for the indicator of the interval whose Gaussian measure is .

We can now state Borell’s Reference 12 noise stability result.

Theorem 2.8.

For all , , and , it holds that

Borell Reference 12 was interested in more general functionals of the heat equations, and he showed that these functionals increase with respect to nonincreasing spherical rearrangement. The fact that half-spaces are the unique optimizers of -noisy inner products was proven in Reference 55, where a robust version of the theorem is also proven. Tighter robust versions were later proven by Eldan Reference 21. Other alternative proofs and generalizations of Borell’s result include Reference 33Reference 42.

2.3. Gaussian and Boolean noise stability

By applying the CLT, it is easy to check that Gaussian noise stability provides bounds on Boolean noise stability.

Proposition 2.9.

For every , , and for every

there exists sequence of Boolean functions such that and

Moreover, by Theorem 2.8 and Theorem 2.4 it follows that the extreme Gaussian noise stability is bounded away from the extreme Boolean noise stability at and ; see Figure 1.

The proof of the proposition is standard using approximation of Gaussian random variables in terms of sums of independent Bernoullis:

Proof.

To show that we can obtain the right endpoint of the interval, let

denote the indicator that the normalized sums of the lie in the intervals and , and apply the CLT. The proof of achievability of the left endpoint is similar where now we take the intervals and .

To obtain an intermediate point , take the inputs of and to be defined on overlapping blocks of bits, e.g, by varying from to in

we get all values in . Taking overlapping intervals for and gives all the values in .

2.4. Smooth Boolean functions

To better understand the connection between Boolean and Gaussian stability, we define two notions of smoothness, termed low influences and resilience for Boolean functions. We begin with the notion of influence. This notion measures the power of a voter Reference 20. It plays a crucial role in the analysis of Boolean functions Reference 6 and Reference 36. When it comes to general probability spaces, there are several possible definitions; see, e.g., Reference 13Reference 40. We choose the definition, which is closely related to the notion of local variance in statistics. To simplify notation, we will often omit the sigma algebra and probability measure defined over a probability space .

Definition 2.10.

Consider a probability space . For a function , we define the th influence of as

where the expected value is with respect to the product measure on . In the Boolean case with the uniform measure , the influence is equivalently defined as

or as

where

is the discrete th directional derivative.

An easy corollary of the definition is that for , it holds that is small in the sense that the sum of its influences is bounded as a function of only.

Lemma 2.11.

Let and . Then

Proof.

The proof follows.

Hypercontractivity is a key feature of many of the proofs in the analysis of Boolean functions starting with the Kahn–Kalai–Linial paper Reference 36. Many of the results in analysis of Boolean functions in general and in the study of quantitative social choice in particular, use the famous hypercontractive theorem:

Theorem 2.12 (Reference 10).

Let and . Then if , then

Theorem 2.12 has a long history. First, Gaussian versions of it were proven by Nelson Reference 65Reference 66 and in the Boolean setting by Bonami Reference 9Reference 10 and by Gross Reference 29; see, e.g., Reference 69 for more detailed discussion.

The easy Lemma 2.11 states that is smooth in a local sense—as on average, it bounds the sum of its discrete derivatives. Theorem 2.12 proves global smoothness, as it shows that higher norms of are bounded by lower norms of .

Interestingly, in quantitative social choice, a reverse inequality proved by Borell Reference 11 also plays an important role.

Theorem 2.13 (Reference 11).

Let and . Then, for any ,

and for any ,

Recall that for , and we write

While the inequalities may seem like a curiosity, as and “norms” for are rarely used in analysis (nor are they norms), the second inequality is quite helpful in some social choice proofs. In more general settings, reverse hypercontraction is implied by standard hypercontraction and in fact by a weaker inequality, called the modified log-Sobolev inequality. For a more general discussion of reverse hypercontraction and its applications, see Reference 59Reference 60.

For a voting function , is the probability that voter is the deciding voter, given all other votes. A stronger notion of the power of a voter or a small set of voters is that their vote affects the expected outcome on average. A function whose expectation is not affected by any small set of voters is called resilient. More formally,

Definition 2.14.

We say that a function is -resilient if

for all sets with and all .

Proposition 2.15.

If satisfies

then is -resilient. In particular if has all influences bounded by , then is -resilient.

Proof.

The second statement follows from the first one immediately as for every nonempty , we may choose , and then

as needed. For the first statement, assume Equation 4. Then

Resilient functions have long been studied in the context of pseudo-randomness; see, e.g., Reference 15.

Thus, the statement that a function has a high-influence variable means that there exists a voter that can have a noticeable effect on the outcome if voter has access to all other votes cast. The statement that a function is not resilient implies that there is a bounded set of voters who have noticeable effect on the outcome on average, i.e., with no access to other votes cast. Consider the following examples:

Dictator has maximal influence of (and all other ). It is also not resilient for .

Majority has all influences of order and is also -resilient.

An example of a resilient function with a high-influence variable is the function

Here coordinate has influence but the function is resilient. In terms of voting, voter has a lot of power if she has access to all other votes cast (or the majority of the votes), but without access to this information, she is powerless. Moreover, every small set of voters can change the expected value of (by conditioning on their vote) by . Another simple example is the parity function , which is -resilient for every , but where all influences are .

The Majority Is Stablest theorem states that the extremal noise stability of low-influence/resilient functions on the discrete cube is captured by Gaussian noise stability. Here are three increasingly stronger statements along this line.

Theorem 2.16 (Reference 57Reference 58).

For every , there exists a for which the following holds. Let satisfy for all . Then

This theorem is called “Majority Is Stablest” ,since where and .

It turns out that for two functions, it is in fact enough that one of them is low influence to obtain the same results, i.e.:

Theorem 2.17 (Reference 51, Prop. 1.15).

For every and , there exists a for which the following holds. Let be such that for all . Then

where one can take

In particular the statement above holds when and is any Boolean function bounded between and .

Moreover, one can replace the low-influence condition by the condition that the function is resilient:

Theorem 2.18 (Reference 53).

For every , there exist for which the following holds. Let be -resilient, and let be an arbitrary function. Then

One can take

where is given by Equation 6.

Note in particular that for our current bounds for and for fixed , is exponential in a polynomial in and is doubly exponential in a polynomial in . Similar statements for one function were proven before by Reference 70 and appeared in Reference 35.

There are two known proof strategies for Theorem 2.16. In Reference 57Reference 58, the authors applied a nonlinear invariance principle, which states that the distribution of low-degree polynomials with low influences is nearly identical when the variables are independent Bernoulli and when they are independent Gaussians. Thus, it is possible to apply Theorem 2.8 (Reference 12) with error that diminishes with the influences.

A second approach Reference 18Reference 19 does not use Borell’s result and is based on induction on dimension in the discrete cube. It is inspired by Bobkov’s inductive proof of the Gaussian isoperimetric inequality Reference 8.

Obtaining the stronger Theorems 2.17 and 2.18 from the weaker statements is more straightforward using simple averaging arguments Reference 51 and applying Boolean regularity lemmas Reference 35Reference 53Reference 70.

2.5. Are pluralities stablest?

Given the asymptotic optimality of the stability of Majority, it is natural to ask if a similar statement holds for low-influence or resilient functions for , where the conjectured most-stable functions are now plurality functions. This was first asked in Reference 41. Similar to the case of , there is an equivalent Gaussian question Reference 33, which was conjectured to be true in Reference 33. The shape of the conjectured optimal partition of to pasts has a number of names, including the “Gaussian double-bubble”, the “standard ”, and the “peace-sign”.

For the isoperimetric problem, corresponding to , the optimality of such a partition for was obtained in Reference 17, under mild conditions and for all and no additional conditions, in recent work Reference 47Reference 48.

The case of noise stability (i.e., constant ) turned out to be much more subtle. Recall that in the binary case, low-influence functions cannot be asymptotically more stable than majorities with the same expectation. In Reference 30 the authors showed that for any probability measure that has full support there exists low-influence functions that are more stable than all plurality functions with the same expected values. Thus low influence functions that are not balanced can be asymptotically more stable in all pluralities with the same expectation. On the other hand, in the balanced case, for , and for small positive values of , the conjectured Gaussian double-bubble was very recently shown to be optimal Reference 31.

3. Paradoxes, noise stability, and reverse hypercontraction

3.1. Probability of paradox

Our next goal is to prove a quantitative version of Arrow’s theorem following Reference 52. We will only discuss the case of three alternatives but will allow different functions to determine different pairwise selections. Recall that we consider voters who vote independently and where voter votes uniformly at random from the six possible rankings. Recall that we encode the six possible rankings by vectors . Here is if a voter ranks above/below , is if voter ranks above/below , is if voter ranks above/below . We will assume further that are the aggregation functions for the vs. , vs. , and vs. preferences. We will again use the following observation used in Reference 37: Since the binary predicate for can be expressed as

we can write

where the last equality follows from the fact that the uniform distribution over satisfies and similarly for other pairs of coordinates.

To state a quantitive version, we will say that a function is -close to a function if . The quantitative version we we wish to prove is the following.

Theorem 3.1.

For every , there exists such that the following holds for every . If

then either two of the functions are -close to constant functions of the opposite sign, or there exists a variable such that , , and are all -close to the same dictator on voter .

The main significance of Theorem 3.1 is that it is dimension independent. We get the same bound no matter what the dimension number of voters is. This shows that one cannot avoid the curse of paradoxes in voting by assuming the probability of a paradox vanishes as the number of voters grows. The dependence of on proven in Reference 52 is worse than exponential, i.e., . Using the results of Reference 52, hypercontractivity and reverse-hypercontractivity, Keller Reference 39 obtained the optimal dependency, .

Before discussing how to prove Theorem 3.1, we give a direct implication of the Majority Is Stablest theorem in the case where the functions are all balanced so . Using the Majority Is Stablest theorem, we obtain the following.

Theorem 3.2 (Reference 37Reference 58).

For every , there exists a such that if satisfy and have all influences bounded above by , then

Again, the right hand side of equation Equation 8 is the asymptotic probability that when are all given by the same Majority function. Theorem 3.2 provides a surprising counter argument to Condorcet’s arguments. Condorcet argued that pairwise ranking by Majority is problematic as it results in a paradox and Theorem 3.2 shows that in fact Majority asymptotically minimizes the probability of a paradox among low-influence functions.

We also have the following strengthening of Theorem 3.2.

Theorem 3.3.

For every , there exist such that if satisfy and , , and are all -resilient, then

3.2. Kalai’s proof for the balanced case

The special case of Arrow’s theorem, where all the functions are balanced, was the first where a quantitative Arrow’s theorem was proved by Kalai Reference 37. The function is balanced if which is equivalent to . In terms of voting this means that a priori both outcomes of are equally likely.

In this short section we provide the statement and the proof of this special case.

Theorem 3.4 (Reference 37).

There exists a constant such that if satisfy and

then there exists a dictator function , such that

The proof of the theorem will use the Friedgut–Kalai–Naor theorem Reference 25, which will be proved at the end of the section.

Theorem 3.5 (Reference 25).

There exists a constant such that if satisfies and

Then there exists a dictator such that .

We now prove Theorem 3.4.

Proof.

Recalling

and Theorem 2.4, it is clear that to prove Theorem 3.4 it suffices to prove the following. There exists a constant such that if satisfy and

then there exists a dictator , such that

Note that

where . Thus if Equation 10 holds, then

It therefore suffices to show that if Equation 12 holds, then Equation 11 holds. By the Cauchy–Schwarz inequality,

Therefore, by the Friedgut–Kalai–Naor theorem, Theorem 3.5, is -close to a dictator . Similarly, is -close to a dictator . Note that the statement of the theorem is trivial if and , so assume . It remains to show that . Note that if , then

However, by Equation 12 and using the fact that ,

and therefore . The assertion follows.

We will now prove the Friedgut–Kalai–Naor theorem, Theorem 3.5, so that the results of this section are self-contained. The proof gives a flavor of some of the reasoning applied in analysis of Boolean function but is not needed for any of the material in the following sections. The proof is similar to the proof in Reference 69. For an alternative proof see Reference 34. We will use the following corollary of hypercontraction.

Lemma 3.6.

Let . Then

Proof.

By hypercontraction if , then

On the other hand,

so by choosing , we get

The proof will also use the Paley–Zygmund inequality stating that for a positive random variable and for it holds that

Applying this inequality for and using Lemma 3.6 implies the following.

Corollary 3.7.

For ,

We now prove Theorem 3.5.

Proof.

Let

Note that and that

Note that implies that . Moreover using the fact that , for all and sufficiently small ,

Therefore,

In particular for , we obtain

On the other hand, applying Equation 13 with , implies that

and therefore , with . Now

so we obtain that

as needed.

3.3. Sketch of proof of the general case

The case where the functions are not balanced goes through a longer route Reference 52. We will sketch the proof of Theorem 3.1, which we restate here.

Theorem 3.8.

For every , there exists such that the following holds for every . If

then either two of the functions are -close to constant functions of the opposite sign, or there exists a variable such that , , and are all -close to the same dictator on voter .

The main steps of the proof are the following.

I.

First a Gaussian version of the theorem is formulated and proved. One advantage of Gaussian space is that it has no dictators, and therefore, the statement is simpler: that unless some choices are almost fixed, there is a good probability of paradox.

II.

Once a Gaussian version is proven and using the Majority Is Stablest theorem, one can deduce the same statement as long as all of the influences are small. In fact, due to Theorem 2.17, it is sufficient that each variable is influential for at most one of the functions .

III.

Using the reverse hypercontractivity inequality by Borell Reference 11, Theorem 2.13, it is easy to show that if two voters have high influence for two different functions , then the probability of paradox is high.

IV.

The remaining case is where there is only one voter who is influential. In this case, by conditioning on the vote of this voter and applying the low-influence result in II, it is possible to conclude that the function is either close to a dictator or has a high probability of paradox.

To provide more of a taste of the proof, we state the Gaussian version of Arrow’s theorem in I and give more details on the application of the reverse hypercontractivity inequality in III.

The Gaussian version corresponds to a situation where the functions can only “see” averages of large subsets of the voters. We thus define a three-dimensional normal vector . The first coordinate of is supposed to represent the deviation of the number of voters where ranks above from the mean. The second coordinate is for ranking above , and the last coordinate for ranking above .

Since averaging maintains the expected value and covariances, we define

We let , …, be independent copies of . We write , and for we write . The Gaussian version of Arrow’s theorem states the following.

Theorem 3.9.

For every there exists a such that the following hold. Let . Assume that for all and all it holds that

Then

Moreover, one may take .

Interestingly it is not hard to prove Theorem 3.9 either by Borell’s half-space result, Theorem 2.8, or using a Gaussian variant of his reverse hypercontraction result, Theorem 2.13.

We next apply reverse hypercontraction in the setting of III. We say that coordinate is pivotal for at if there exists a value of such that . We first prove the following.

Lemma 3.10.

Suppose that and . Let

Then

Proof.

Let

Then and , and our goal is to obtain a bound on . Note that the event is determined by and the event is determined by . So the proof follows reverse hypercontraction with .

We can therefore conclude that:

Theorem 3.11.

Suppose that there exist voters and such that

Then .

Proof.

Without loss of generality assume that and . Let , where and are the events from Lemma 3.10. By the lemma we have . Note that conditioned on any , the functions , , and on coordinates and satisfy the condition of Arrow’s theorem, Theorem 1.3. Thus with probability at least , the outcome is not transitive. Therefore,

3.4. More general statements

In this subsection we discuss a more general statement of Arrow’s theorem which is closer to its original formulation and its quantitative counterpart. This requires that we introduce a number of additional definitions. The reduction from the more general statements of Arrow’s theorem to the three candidate case discussed above will be carried out in this subsection.

3.4.1. General setup

Consider , a set of alternatives. A transitive preference over is a ranking of the alternatives from top to bottom where ties are not allowed. Such a ranking corresponds to a permutation of the elements of where is the rank of alternative . The set of all rankings will be denoted by .

A constitution is a function that associates to every -tuple of transitive preferences (also called a profile), and every pair of alternatives a preference between and . Some key properties of constitutions include the following.

Transitivity. The constitution is transitive if is transitive for all . In other words, for all and for all three alternatives , , and , if prefers to , and prefers to , it also prefers to . Thus is transitive if and only if its image is a subset of the permutations on elements.

Independence of irrelevant alternatives (IIA). The constitution satisfies the IIA property if for every pair of alternatives and , the social ranking of vs. (higher or lower) depends only on their relative rankings by all voters. The IIA condition implies that the pairwise preference between any pair of outcomes depends only on the individual pairwise preferences. Thus, if satisfies the IIA property, then there exists functions for every pair of candidates and such that

Unanimity. The constitution satisfies unanimity if the social outcome ranks above whenever all individuals rank above .

The constitution is a dictator on voter if for all or if for all , where is the ranking by reversing the ranking .

Arrow’s theorem states Reference 1Reference 2 the following.

Theorem 3.12.

Any constitution on three or more alternatives which satisfies transitivity, IIA, and unanimity is a dictatorship.

It is possible to give a characterization of all constitutions satisfying IIA and transitivity. Results of Wilson Reference 74 provide a partial characterization for the case where voters are allowed to rank some alternatives as equal. In order to obtain a quantitative version of Arrow’s theorem, we give an explicit and complete characterization of all constitutions satisfying IIA and transitivity in the case where all voters vote using a strict preference order. Write for the set of all constitutions on alternatives and voters satisfying IIA and transitivity. For the characterization it is useful write for the statement that for all it holds that ranks all alternatives in above all alternatives in . We will further write for the constitution restricted to the alternatives in . The IIA condition implies that depends only on the individual rankings of the alternatives in the set . The characterization of we prove is the following.

Theorem 3.13.

The class consist exactly of all constitutions satisfying the following. There exists a partition of the set of alternatives into disjoint sets , …, such that the following hold.

.

For all such that , there exists a voter such that is a dictator on voter .

For all such that , the constitution is a nonconstant function of the preferences on the alternatives in .

We note that for all all elements of are not desirable as constitutions. Indeed elements of either have dictators whose vote is followed with respect to some of the alternatives or they always rank some alternatives on top some other. For a related discussion see Reference 74. The statement above follows easily from Reference 74, Theorem 3. The exact formulation is taken from Reference 50.

The main goal of the current section is to provide a quantitative version of Theorem 3.13 assuming voters vote independently and uniformly at random. Note that Theorem 3.13 above implies that if , then the probability of a paradox, , satisfies . However if is large and the probability of a nontransitive outcome is indeed as small as , one may argue that a nontransitive outcome is so unlikely that in practice Arrow’s theorem is irrelevant.

Theorem 3.14.

For every number of alternatives and , there exists a , such that for every , if is a constitution on voters and alternatives satisfying

IIA and

,

then there exists satisfying .

We therefore obtain the following.

Corollary 3.15.

For any number of alternatives and , there exists a , such that for every , if is a constitution on voters and alternatives satisfying

IIA, and

is -far from any dictator, so for any dictator ,

for every pair of alternatives and , the probability that ranks above is at least ,

then the probability of a nontransitive outcome, , is at least .

Proof.

Assume by contradiction that . Then by Theorem 3.14 there exists a function satisfying . Note that for every pair of alternatives and it holds that

Therefore for every pair of alternatives there is a positive probability that ranks above . Thus by Theorem 3.14 it follows that is a dictator which is a contradiction.

Remark 3.16.

Note that if and is any constitution satisfying , then .

Remark 3.17.

The bounds stated in Theorem 3.14 and Corollary 3.15 in terms of and are not optimal. We expect that the true dependency has which is some fixed power of . Moreover, we expect that the bound should be improved to .

3.4.2. Nisan’s argument

Noam Nisan argued in his blog Reference 68 that the natural way to study quantitative versions of Arrow’s theorem is to look at functions from to and check to what extent they satisfy the IIA property. That is, while so far we insisted on the IIA condition and checked how far we are from transitivity, Nisan’s suggested that it is more natural to insist on transitivity and ask how far we are from IIA. The point of view of quantitative versions of Arrow’s theorem was also taken in Reference 22.

To formalize his approach, Nisan defines a function to be -IIA if for every two alternatives and , it holds that , where is uniformly chosen and is uniformly chosen conditioned on the ranking at being identical to that of for all voters. In his blog Nisan sketches how a quantitative Arrow’s theorem proven for the definition used here implies a quantitative Arrow’s theorem for his definition. We briefly repeat the argument with some minor modifications and corrections.

Fix alternatives and write for the probability that, given a vector of binary preferences between and , ranks above . If satisfies the IIA property, then almost surely. If is -IIA, then , and therefore .

Assume a quantitative Arrow’s theorem such as the one proven here with parameters , and suppose by contradiction that is -IIA and -far from for some small to be determined later. Define a function as follows. Let rank above if for the majority of which agree with in the orderings it holds that ranks above . We note that for every pair of alternatives it holds that

By taking a union bound on all pairs of alternatives, this implies that . Note further that satisfies the IIA property by definition. Since is transitive and from the quantitative Arrow’s theorem proven here we conclude that

and a contradiction is implied unless . Thus the Arrow’s theorem for the -IIA definition holds with . (We briefly note that moving from to does not preserve the property of the function being balanced so in the setting of Kalai’s theorem an additional argument is needed.)

3.4.3. Proof of Theorem 3.14

Proof.

The proof follows by applying Theorem 3.1 to triplets of alternatives. Assume .

Note that if are two different functions, each of which is either a dictator or a constant function, then . Therefore, for all it holds that for at most one function which is either a dictator or a constant function. In case there exists such function, we let ; otherwise, we let .

Let be the social choice function defined by the functions . Clearly,

The proof would follow if we could show and, therefore, .

To prove that it suffices to show that for every set of three alternatives, it holds that . Since implies , Theorem 3.1 implies that there exists a function such that . There are two cases to consider:

is a dictator. This implies that is -close to a dictator for each and therefore for all pairs , so .

There exists an alternative (say ) that always ranks at the top/bottom. In this case we have that and are at most -far from the constant functions and (or and ). The functions and have to take the same constant values, and therefore again we have that .

The proof follows.

Remark 3.18.

Note that this proof is generic in the sense that it takes the quantitative Arrow’s result for three alternatives as a black box and produces a quantitative Arrow’s result for any alternatives.

3.4.4. Other probability measures

Given the right analytic tools, it is not hard to generalize the proof of Theorems 3.1 and 3.14 to other product distributions. This is done in Reference 60 where some of the tools related to reverse hypercontractivity are developed. In Reference 60 the authors obtain the following extension.

Theorem 3.19 (Quantitative Arrow’s theorem for general distribution).

Let be general distribution on with assigning positive probability to each element of . Let denote the distribution on . Then for any number of alternatives and , there exists , such that for every , if satisfies

IIA and

.

Then there exists a function which is transitive and satisfies the IIA property and .

3.5. Other variants

In concluding this section, we discuss the optimal low-influence function for voting in the case of alternatives. When we are considering alternatives, we want to define more formally the possible outcome in Arrow’s voting. Since for every two alternatives a winner is decided, the aggregation results in a tournament on the set . Recall that is a tournament on if it is a directed graph on the vertex set such that for all either or . Given individual rankings the tournament is defined as follows.

Let if , and let if . Note that .

The binary decision between each pair of candidates is performed via an antisymmetric function so that for all . Here we assume that each two candidates are compared using the same function . Moreover, since we require that the comparison between and and and will be the same, we require that is antisymmetric. The tournament is then defined by letting if and only if .

Note that there are tournaments while there are only linear rankings. For the purposes of social choice, some tournaments make more sense than others.

Definition 3.20.

We say that a tournament is linear if it is acyclic. We will write for the logical statement that is acyclic. Nonlinear tournaments are often referred to as nonrational in economics as they represent an order where there are three candidates , , and such that is preferred to , is preferred to , and is preferred to .

We say that the tournament is a unique max tournament if there is a candidate such that for all it holds that . We write for the logical statement that has a unique max. Note that the unique max property is weaker than linearity. It corresponds to the fact that there is a candidate that dominates all other candidates.

A generalization of Borell’s result along with a general invariance principle Reference 49Reference 51 allows us to prove the following Reference 33.

Theorem 3.21.

For any and there exists a such that for any antisymmetric satisfying ,

An alternative proof can be derived via multidimensional generalization of the inductive Majority Is Stablest theorem using a general notion of -concavity Reference 44Reference 63.

It is not hard to show that

To prove Equation 16 one can use a multidimensional CLT to represent the advantage of candidate over candidate as

where , , …, are i.i.d. random variables.

Other than the case , where the notions of unique-max and linear tournaments coincide, very little is known about which function maximizes the probability of a linear order. Even computing this probability for Majority provides a surprising result Reference 51:

Proposition 3.22.

We have

We find this asymptotic behavior quite surprising. Indeed, given the previous results that the probability that there is a unique max is , one may expect that the probability that the order is linear would be

However, it turns out that there is a strong negative correlation between the event that there is a unique maximum among the candidates and that among the other candidates there is a unique max.

Proof.

We use the multidimensional CLT. Let

By the CLT the collection of variables converges to a joint Gaussian vector satisfying for all distinct ,

and for all and .

We are interested in providing bounds on

as the probability that the resulting tournament is an order is obtained by multiplying this quantity by a factor.

We claim that there exist independent random variables for and for such that

where . This follows from the fact that the joint distribution of Gaussian random variables is determined by the covariance matrix (this is noted in the literature in Reference 67).

We now prove the upper bound. Let be a constant to be chosen later. Note that for all and large enough it holds that

Therefore the probability that for at least half of the ’s in the interval it holds that is at most

Let’s assume that at least half of the ’s in the interval satisfy that . We claim that in this case the number of pairs such that and is .

For the last claim, partition the interval into subintervals of length and note that at least of the points belong to subintervals which contain at least points. This implies that the number of pairs satisfying is .

Note that, for such pair in order that , we need that which happens with constant probability.

We conclude that, given that half of the ’s fall in , the probability of a linear order is bounded by

Thus overall we have bounded the probability by

The optimal exponent is giving the desired upper bound.

For the lower bound we condition on taking value in . Each probability is at least and therefore the probability that all take such values is

Moreover, conditioned on taking such values, the probability that

for all is at least

This proves the required result.

4. Manipulation and isoperimetry

4.1. Quantitive manipulation

The Gibbard–Satterthwaite theorem, Theorem 1.4, states that under natural conditions, there exist profiles of voters such that at least one voter can manipulate the vote. We are interested in quantitative versions of the statement. A natural approach is to view quantitative statements as isoperimetric results. In classical isoperimetric theory, the goal is to find conditions that establish a large boundary between sets. In the context of manipulation we can consider a voter who can manipulate as a special boundary point, and our goal is to prove that there are many boundary points.

It is natural to consider the following graph where the vertex set is —the set of all voting profiles and there are edges between voting profiles that differ at a single voter. The statement of the Gibbard–Satterthwaite theorem can be interpreted in terms of this graph: for certain natural partitions of into parts, there is an edge of the graph between two different parts that corresponds to a manipulation. It is not hard to see that the existence of many edges between different parts of the graph follows from classical isoperimetric theory.

Thus one may consider quantitative statements of the Gibbard–Satterthwaite theorem as isoperimetric statements: It is not only the case that there are many edges between different parts of the partition, but it is also the case that many of these edges correspond to manipulation by one of the voters.

In the classical setup, isoperimetry and concentration of measure are closely related. In particular, standard concentration of measure results imply that for any set of fractional size at least in , the set of profiles at graph distance at most contains almost the whole graph. However, it is not known under what conditions typically a small coalition can manipulate.

It may be useful to consider the following examples:

Consider the plurality function with alternatives and voters. For a voter to be able to manipulate, it must hold that the difference between the top candidate and the second to top is at most . This implies that the probability that there exists a voter who can manipulate is .

A second question we can ask is what is the minimal size of a coalition that can manipulate with probability close to . Since the difference between the top candidate and the second candidate is typically of order , it is clear that the size has to be at least order , and in fact it is easy to see that if that for every fixed set of size , with probability there exists a subset that can change the outcome of the elections by manipulating. Indeed if is the top candidate, is the second from the top, and is a candidate different than and , we may take to be all the voters in that rank above above . If these voters would rank as their top candidate, the outcome will be which is more favorable to them.

Consider the case and the function where is the tribes function. The tribes function Reference 6 is a balanced monotone function, where all influences are . In this case a voter can manipulate if and only if they are influential. Therefore, in general, we cannot expect that the probability that an individual voter can manipulate is higher than .

4.1.1. A quantitive Gibbard–Satterthwaite theorem

Our goal in this section is to sketch the proof of a quantitative version of the manipulation theorem. We will mostly follow Reference 61Reference 62 who proved a pretty general average manipulation theorem for a single voter. Some special cases of the theorem were known before: in particular in the case of three alternatives, this was proved by Friedgut, Kalai, Keller, and Nisan Reference 24Reference 26. The original proofs of Gibbard–Satterthwaite theorem were carried out by reduction to Arrow’s theorem. The arguments of Reference 24Reference 26 succeed in providing a quantitive reduction to the quantitive Arrow’s theorem, but only in the case of alternatives.

Here our goal is to sketch the following result: if and the SCF is -far from the family of nonmanipulable functions, then the probability of a ranking profile being manipulable is bounded from below by a polynomial in , , and . We continue by stating the results and their implications, and sketching the main steps of the proof.

4.1.2. Definitions and formal statements

Recall that our basic setup consists of voters electing a winner among alternatives via an SCF . We now define manipulability in more detail.

Definition 4.1 (Manipulation points).

Let be a ranking profile. Write to denote that alternative is preferred over by voter . An SCF is manipulable at the ranking profile if there exists a and an such that and only differ in the th coordinate and

In this case we also say that is a manipulation point of , and that is a manipulation pair for . We say that is manipulable if it is manipulable at some point . We also say that is an -manipulation point of if has a manipulation pair such that is obtained from by permuting (at most) adjacent alternatives in one of the coordinates of . (We allow —any manipulation point is an -manipulation point for .)

Let denote the set of manipulation points of the SCF , and for a given let denote the set of -manipulation points of . When the SCF is obvious from the context, we write simply and .

We first recall the Gibbard–Satterthwaite theorem (stated as Theorem 1.4 in the Introduction).

Theorem 4.2 (Gibbard and Satterthwaite Reference 28Reference 71).

Any SCF which takes at least three values and is not a dictator (i.e., not a function of only one voter) is manipulable.

This theorem is tight in the sense that monotone SCFs, which are dictators or only have two possible outcomes, are indeed nonmanipulable (a function is nonmonotone, and clearly manipulable, if for some ranking profile a voter can change the outcome from, say, to by moving ahead of in her preference). It is useful to introduce a refined notion of a dictator before defining the set of nonmanipulable SCFs.

Definition 4.3 (Dictator on a subset).

For a subset of alternatives , let be the SCF on one voter whose output is always the top-ranked alternative among those in .

Definition 4.4 (Nonmanipulable SCFs).

We denote by

the set of nonmanipulable SCFs, which is the following:

When the parameters and are obvious from the context, we omit them.

Another useful class of functions, which is larger than but which has a simpler description, is the following.

Definition 4.5.

Define, for parameters and that remain implicit,

The notation should be thought of as “closure” rather than “complement”.

As discussed previously, our goal is to study manipulability from a quantitative viewpoint, and in order to do so we need to define the distance between SCFs.

Definition 4.6 (Distance between SCFs).

The distance between two SCFs is defined as the fraction of inputs on which they differ: , where is uniformly selected. For a class of SCFs, we write .

The concepts of anonymity and neutrality of SCFs will be important to us, so we define them here.

Definition 4.7 (Anonymity).

An SCF is anonymous if it is invariant under changes made to the names of the voters. More precisely, an SCF is anonymous if for every and every ,

Definition 4.8 (Neutrality).

An SCF is neutral if it commutes with changes made to the names of the alternatives. More precisely, an SCF is neutral if for every and every ,

Our goal is to sketch the proof of the following theorem.

Theorem 4.9.

Suppose we have voters, alternatives, and an SCF satisfying . Then

for some polynomial , where is selected uniformly.

An immediate consequence is that

for some polynomial , where is uniformly selected, and is obtained from by uniformly selecting a coordinate , uniformly selecting , and then uniformly randomly permuting the following four adjacent alternatives in , and .

4.1.3. Proof ideas

We first present our techniques that achieve a lower bound for the probability of manipulation that involves factors of and then describe how a refined approach leads to a lower bound which has inverse polynomial dependence on .

Rankings graph and applying the original Gibbard–Satterthwaite theorem

Consider the graph having vertex set , the set of all ranking profiles, and let if and only if and differ in exactly one coordinate. The SCF naturally partitions into subsets. Since every manipulation point must be on the boundary between two such subsets, we are interested in the size of such boundaries.

For two alternatives and and for voter , denote by the boundary between and in voter . A simple lemma tells us that at least two of the boundaries are large. In the following assume that these are and . The case where the boundaries are and , where are distinct, is easier due to the independence between the relative ranking of vs. and vs. .

Now if a ranking profile lies on both of these boundaries, then applying the original Gibbard–Satterthwaite theorem to the restricted SCF on two voters where we fix all coordinates of except the first two, we get that there must exist a manipulation point which agrees with in all but the first two coordinates. Consequently, if we can show that the intersection of the boundaries and is large, then we have many manipulation points.

Fibers and reverse hypercontractivity

In order to have more “control” over what is happening at the boundaries, we partition the graph further—this idea is due to Friedgut et al. Reference 24Reference 26. Given a ranking profile and two alternatives and , induces a vector of preferences between and . For a vector , we define the fiber with respect to preferences between and , denoted by , to be the set of ranking profiles for which the vector of preferences between and is . We can then partition the vertex set into such fibers, and work inside each fiber separately. Working inside a specific fiber is advantageous, because it gives us the extra knowledge of the vector of preferences between and .

We distinguish two types of fibers: large and small. We say that a fiber with respect to preferences between and is large if almost all of the ranking profiles in this fiber lie on the boundary , and is small otherwise. Now since the boundary is large, either there is big mass on the large fibers with respect to preferences between and or big mass on the small fibers. This holds analogously for the boundary and fibers with respect to preferences between and .

Consider the case when there is big mass on the large fibers of both and . Notice that for a ranking profile , being in a fiber with respect to preferences between and only depends on the vector of preferences between and , , which is a uniform bit vector. Similarly, being in a fiber with respect to preferences between and only depends on . Moreover, we know the exact correlation between the coordinates of and , and it is in exactly this setting where reverse hypercontractivity applies and shows that the intersection of the large fibers of and is also large. Finally, by the definition of a large fiber it follows that the intersection of the boundaries and is large as well, and we can finish the argument using the Gibbard–Satterthwaite theorem as above.

To deal with the case when there is big mass on the small fibers of , we use various isoperimetric techniques, including the canonical path method. In particular, we use the fact that for a small fiber for , the relative size of the boundary of in the small fiber is comparable to the size of in the small fiber itself, up to polynomial factors.

A refined geometry

Using this approach with the rankings graph above, our bound includes factors. In order to obtain inverse polynomial dependence on , we use a refined approach. Instead of the rankings graph outlined above, we use an underlying graph with a different edge structure: if and only if and differ in exactly one coordinate, and in this coordinate they differ by a single adjacent transposition. In order to prove the refined result, we need to show that the geometric and combinatorial quantities, such as boundaries and manipulation points, are roughly the same in the refined graph as in the original rankings graph.

Acknowledgments

My interest in voting was strongly influenced by Gil Kalai. Our conversations over the years have played a major role in my research in this area as well as in related areas, such as analysis of Boolean functions. Thanks to Ryan O’Donnell, Krzysztof Oleszkiewicz, Joe Neeman, Miki Racz, Anindya De, Noam Nisan, Nathan Keller, Oded Regev, Ronen Eldan, Michel Ledoux, Nati Linial, and Yuval Peres for fruitful collaborations and fascinating conversations. Thanks to anonymous reviewers who provided detailed and insightful feedback and comments. This survey is based in part on lecture series given at the 2019 St. Flour Summer School in Probability. Readers who are interested in details of many of the proofs sketched here, can find many of the detailed proofs and other related results in the draft Reference 54.

About the author

Elchanan Mossel is professor of mathematics at the Massachusetts Institute of Technology. His research spans a number of topics across probability, statistics, economics, computer science, and mathematical biology. He is known for his work in discrete Fourier analysis and its applications to computational complexity and social choice theory and for his research of information flow in biological, economic, and inferential networks. Mossel held a Sloan Fellowship and is a fellow of the American Mathematical Society, a Simons Fellow and a Vannevar Bush Fellow.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. 1.1. The law of large numbers and Condorcet’s jury theorem
    2. Theorem 1.1.
    3. 1.2. Condorcet’s paradox and Arrow’s theorem
    4. Theorem 1.2.
    5. Theorem 1.3.
    6. 1.3. Manipulation and the Gibbard–Satterthwaite theorem
    7. Theorem 1.4 (Gibbard and Satterthwaite).
    8. 1.4. Judgment aggregation
    9. 1.5. Modern perspectives
    10. 1.6. Probability of paradox for the majority
  3. 2. Noise stability
    1. 2.1. Boolean noise stability
    2. Definition 2.1.
    3. Definition 2.2.
    4. Proposition 2.3.
    5. Theorem 2.4.
    6. Corollary 2.5.
    7. Problem 2.6.
    8. 2.2. Gaussian noise stability
    9. Definition 2.7.
    10. Theorem 2.8.
    11. 2.3. Gaussian and Boolean noise stability
    12. Proposition 2.9.
    13. 2.4. Smooth Boolean functions
    14. Definition 2.10.
    15. Lemma 2.11.
    16. Theorem 2.12 (10).
    17. Theorem 2.13 (11).
    18. Definition 2.14.
    19. Proposition 2.15.
    20. Theorem 2.16 (5758).
    21. Theorem 2.17 (51, Prop. 1.15).
    22. Theorem 2.18 (53).
    23. 2.5. Are pluralities stablest?
  4. 3. Paradoxes, noise stability, and reverse hypercontraction
    1. 3.1. Probability of paradox
    2. Theorem 3.1.
    3. Theorem 3.2 (3758).
    4. Theorem 3.3.
    5. 3.2. Kalai’s proof for the balanced case
    6. Theorem 3.4 (37).
    7. Theorem 3.5 (25).
    8. Lemma 3.6.
    9. Corollary 3.7.
    10. 3.3. Sketch of proof of the general case
    11. Theorem 3.8.
    12. Theorem 3.9.
    13. Lemma 3.10.
    14. Theorem 3.11.
    15. 3.4. More general statements
    16. Theorem 3.12.
    17. Theorem 3.13.
    18. Theorem 3.14.
    19. Corollary 3.15.
    20. Theorem 3.19 (Quantitative Arrow’s theorem for general distribution).
    21. 3.5. Other variants
    22. Definition 3.20.
    23. Theorem 3.21.
    24. Proposition 3.22.
  5. 4. Manipulation and isoperimetry
    1. 4.1. Quantitive manipulation
    2. Definition 4.1 (Manipulation points).
    3. Theorem 4.2 (Gibbard and Satterthwaite 2871).
    4. Definition 4.3 (Dictator on a subset).
    5. Definition 4.4 (Nonmanipulable SCFs).
    6. Definition 4.5.
    7. Definition 4.6 (Distance between SCFs).
    8. Definition 4.7 (Anonymity).
    9. Definition 4.8 (Neutrality).
    10. Theorem 4.9.
    11. Rankings graph and applying the original Gibbard–Satterthwaite theorem
    12. Fibers and reverse hypercontractivity
    13. A refined geometry
  6. Acknowledgments
  7. About the author

Figures

Figure 1.

The noise stability of dictator and Gaussian half-space of measure , i.e., functions and . Note that for every , the dictator is more stable than the corresponding half-spaces, and for every it is less stable than the corresponding half-space.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{axis}[ domain=-1:1, samples=101, smooth, no markers, xlabel = $\rho$, ylabel = Stability, axis lines = middle, legend pos = south east ] \addplot[ domain=-1:1, samples=100, color=red, ] {x}; \addplot[ domain = -1:1, samples = 1000, color = blue, ] {1- acos(x)/90}; \end{axis} \end{tikzpicture}

Mathematical Fragments

Theorem 1.2.

Let and suppose that never results in a paradox, so for all it holds that . Then is a dictator: there exists an such that for all , or there exists an such that for all .

Equation (1)
Theorem 1.3.

Let and suppose that never results in a paradox. So for all it holds that . Then either two of the functions are constants of opposite signs or there exists an such that , , and are dictators on voter , so or .

Theorem 1.4 (Gibbard and Satterthwaite).

Any SCF which is not a dictatorship (i.e., not a function of a single voter) and which allows at least three alternatives to be elected is manipulable.

Equation (2)
Theorem 2.4.

For every , for every , and for every with , it holds that

Moreover, the only optimizers are dictator functions, i.e., functions of the form or .

Theorem 2.8.

For all , , and , it holds that

Lemma 2.11.

Let and . Then

Theorem 2.12 (Reference 10).

Let and . Then if , then

Theorem 2.13 (Reference 11).

Let and . Then, for any ,

and for any ,

Proposition 2.15.

If satisfies

then is -resilient. In particular if has all influences bounded by , then is -resilient.

Theorem 2.16 (Reference 57Reference 58).

For every , there exists a for which the following holds. Let satisfy for all . Then

Theorem 2.17 (Reference 51, Prop. 1.15).

For every and , there exists a for which the following holds. Let be such that for all . Then

where one can take

In particular the statement above holds when and is any Boolean function bounded between and .

Theorem 2.18 (Reference 53).

For every , there exist for which the following holds. Let be -resilient, and let be an arbitrary function. Then

One can take

where is given by Equation 6.

Theorem 3.1.

For every , there exists such that the following holds for every . If

then either two of the functions are -close to constant functions of the opposite sign, or there exists a variable such that , , and are all -close to the same dictator on voter .

Theorem 3.2 (Reference 37Reference 58).

For every , there exists a such that if satisfy and have all influences bounded above by , then

Theorem 3.4 (Reference 37).

There exists a constant such that if satisfy and

then there exists a dictator function , such that

Theorem 3.5 (Reference 25).

There exists a constant such that if satisfies and

Then there exists a dictator such that .

Equation (10)
Equation (11)
Equation (12)
Lemma 3.6.

Let . Then

Corollary 3.7.

For ,

Theorem 3.9.

For every there exists a such that the following hold. Let . Assume that for all and all it holds that

Then

Moreover, one may take .

Lemma 3.10.

Suppose that and . Let

Then

Theorem 3.13.

The class consist exactly of all constitutions satisfying the following. There exists a partition of the set of alternatives into disjoint sets , …, such that the following hold.

.

For all such that , there exists a voter such that is a dictator on voter .

For all such that , the constitution is a nonconstant function of the preferences on the alternatives in .

Theorem 3.14.

For every number of alternatives and , there exists a , such that for every , if is a constitution on voters and alternatives satisfying

IIA and

,

then there exists satisfying .

Corollary 3.15.

For any number of alternatives and , there exists a , such that for every , if is a constitution on voters and alternatives satisfying

IIA, and

is -far from any dictator, so for any dictator ,

for every pair of alternatives and , the probability that ranks above is at least ,

then the probability of a nontransitive outcome, , is at least .

Equation (16)

References

[1]
K. Arrow, A difficulty in the theory of social welfare, J. of Political Economy 58 (1950), 328–346., Show rawAMSref\bib{Arrow:50}{article}{ author={Arrow, K.}, title={A difficulty in the theory of social welfare}, date={1950}, journal={J. of Political Economy}, volume={58}, pages={328\ndash 346}, } Close amsref.
[2]
K. J. Arrow, Social choice and individual values, Cowles Commission Monograph No. 12, John Wiley & Sons, Inc., New York, N. Y.; Chapman & Hall, Ltd., London, 1951. MR0039976, Show rawAMSref\bib{Arrow:63}{book}{ author={Arrow, Kenneth J.}, title={Social choice and individual values}, series={Cowles Commission Monograph No. 12}, publisher={John Wiley \& Sons, Inc., New York, N. Y.; Chapman \& Hall, Ltd., London}, date={1951}, pages={xi+99}, review={\MR {0039976}}, } Close amsref.
[3]
S. Barberá, Pivotal voters: a new proof of Arrow’s theorem, Econom. Lett. 6 (1980), no. 1, 13–16, DOI 10.1016/0165-1765(80)90050-6. MR614478, Show rawAMSref\bib{Barbera:80}{article}{ author={Barber\'{a}, Salvador}, title={Pivotal voters: a new proof of Arrow's theorem}, journal={Econom. Lett.}, volume={6}, date={1980}, number={1}, pages={13--16}, issn={0165-1765}, review={\MR {614478}}, doi={10.1016/0165-1765(80)90050-6}, } Close amsref.
[4]
W. Beckner, Inequalities in Fourier analysis, Ann. of Math. (2) 102 (1975), no. 1, 159–182, DOI 10.2307/1970980. MR385456, Show rawAMSref\bib{Beckner:75}{article}{ author={Beckner, William}, title={Inequalities in Fourier analysis}, journal={Ann. of Math. (2)}, volume={102}, date={1975}, number={1}, pages={159--182}, issn={0003-486X}, review={\MR {385456}}, doi={10.2307/1970980}, } Close amsref.
[5]
C. E. Bell, A random voting graph almost surely has a Hamiltonian cycle when the number of alternatives is large, Econometrica 49 (1981), no. 6, 1597–1603, DOI 10.2307/1911423. MR636170, Show rawAMSref\bib{Bell:81}{article}{ author={Bell, Colin E.}, title={A random voting graph almost surely has a Hamiltonian cycle when the number of alternatives is large}, journal={Econometrica}, volume={49}, date={1981}, number={6}, pages={1597--1603}, issn={0012-9682}, review={\MR {636170}}, doi={10.2307/1911423}, } Close amsref.
[6]
M. Ben-Or and N. Linial, Collective coin flipping, Randomness and computation, 1990., Show rawAMSref\bib{BenorLinial:90}{incollection}{ author={Ben-Or, M.}, author={Linial, N.}, title={Collective coin flipping}, date={1990}, booktitle={Randomness and computation}, editor={Micali, S.}, publisher={Academic Press}, address={New York}, } Close amsref.
[7]
I. Benjamini, G. Kalai, and O. Schramm, Noise sensitivity of Boolean functions and applications to percolation, Inst. Hautes Études Sci. Publ. Math. 90 (1999), 5–43 (2001). MR1813223, Show rawAMSref\bib{BeKaSc:99}{article}{ author={Benjamini, Itai}, author={Kalai, Gil}, author={Schramm, Oded}, title={Noise sensitivity of Boolean functions and applications to percolation}, journal={Inst. Hautes \'{E}tudes Sci. Publ. Math.}, number={90}, date={1999}, pages={5--43 (2001)}, issn={0073-8301}, review={\MR {1813223}}, } Close amsref.
[8]
S. G. Bobkov, An isoperimetric inequality on the discrete cube, and an elementary proof of the isoperimetric inequality in Gauss space, Ann. Probab. 25 (1997), no. 1, 206–214, DOI 10.1214/aop/1024404285. MR1428506, Show rawAMSref\bib{Bobkov:97b}{article}{ author={Bobkov, S. G.}, title={An isoperimetric inequality on the discrete cube, and an elementary proof of the isoperimetric inequality in Gauss space}, journal={Ann. Probab.}, volume={25}, date={1997}, number={1}, pages={206--214}, issn={0091-1798}, review={\MR {1428506}}, doi={10.1214/aop/1024404285}, } Close amsref.
[9]
A. Bonami, Ensembles dans le dual de (French), Ann. Inst. Fourier (Grenoble) 18 (1968), no. fasc. 2, 193–204 (1969). MR249940, Show rawAMSref\bib{Bonami:68}{article}{ author={Bonami, Aline}, title={Ensembles $\Lambda (p)$ dans le dual de $D^{\infty }$}, language={French}, journal={Ann. Inst. Fourier (Grenoble)}, volume={18}, date={1968}, number={fasc. 2}, pages={193--204 (1969)}, issn={0373-0956}, review={\MR {249940}}, } Close amsref.
[10]
A. Bonami, Étude des coefficients de Fourier des fonctions de (French, with English summary), Ann. Inst. Fourier (Grenoble) 20 (1970), no. fasc. 2, 335–402 (1971). MR283496, Show rawAMSref\bib{Bonami:70}{article}{ author={Bonami, Aline}, title={\'{E}tude des coefficients de Fourier des fonctions de $L^{p}(G)$}, language={French, with English summary}, journal={Ann. Inst. Fourier (Grenoble)}, volume={20}, date={1970}, number={fasc. 2}, pages={335--402 (1971)}, issn={0373-0956}, review={\MR {283496}}, } Close amsref.
[11]
C. Borell, Positivity improving operators and hypercontractivity, Math. Z. 180 (1982), no. 2, 225–234, DOI 10.1007/BF01318906. MR661699, Show rawAMSref\bib{Borell:82}{article}{ author={Borell, Christer}, title={Positivity improving operators and hypercontractivity}, journal={Math. Z.}, volume={180}, date={1982}, number={2}, pages={225--234}, issn={0025-5874}, review={\MR {661699}}, doi={10.1007/BF01318906}, } Close amsref.
[12]
C. Borell, Geometric bounds on the Ornstein-Uhlenbeck velocity process, Z. Wahrsch. Verw. Gebiete 70 (1985), no. 1, 1–13, DOI 10.1007/BF00532234. MR795785, Show rawAMSref\bib{Borell:85}{article}{ author={Borell, Christer}, title={Geometric bounds on the Ornstein-Uhlenbeck velocity process}, journal={Z. Wahrsch. Verw. Gebiete}, volume={70}, date={1985}, number={1}, pages={1--13}, issn={0044-3719}, review={\MR {795785}}, doi={10.1007/BF00532234}, } Close amsref.
[13]
J. Bourgain, J. Kahn, G. Kalai, Y. Katznelson, and N. Linial, The influence of variables in product spaces, Israel J. Math. 77 (1992), no. 1-2, 55–64, DOI 10.1007/BF02808010. MR1194785, Show rawAMSref\bib{BKKKL:92}{article}{ author={Bourgain, Jean}, author={Kahn, Jeff}, author={Kalai, Gil}, author={Katznelson, Yitzhak}, author={Linial, Nathan}, title={The influence of variables in product spaces}, journal={Israel J. Math.}, volume={77}, date={1992}, number={1-2}, pages={55--64}, issn={0021-2172}, review={\MR {1194785}}, doi={10.1007/BF02808010}, } Close amsref.
[14]
C. D Campbell and G. Tullock, The paradox of voting—a possible method of calculation, American Political Science Review 60 (1966), no. 3, 684–685., Show rawAMSref\bib{CampbellTullock:66}{article}{ author={Campbell, Colin~D}, author={Tullock, Gordon}, title={The paradox of voting---a possible method of calculation}, date={1966}, journal={American Political Science Review}, volume={60}, number={3}, pages={684\ndash 685}, } Close amsref.
[15]
B. Chor, O. Goldreich, J. Håsted, J. Freidmann, S. Rudich, and R. Smolensky, The bit extraction problem or t-resilient functions, 26th Annual Symposium on Foundations of Computer Science, 1985, pp. 396–407., Show rawAMSref\bib{Chor_etal:85}{inproceedings}{ author={Chor, Benny}, author={Goldreich, Oded}, author={H{\aa }sted, Johan}, author={Freidmann, Joel}, author={Rudich, Steven}, author={Smolensky, Roman}, title={The bit extraction problem or t-resilient functions}, organization={IEEE}, date={1985}, booktitle={26th Annual Symposium on Foundations of Computer Science}, pages={396\ndash 407}, } Close amsref.
[16]
J.-A.-N. Condorcet, Essai sur l’application de l’analyse à la probabilité des décisions rendues à la pluralité des voix, De l’Imprimerie Royale, 1785., Show rawAMSref\bib{Condorcet:85}{book}{ author={Condorcet, Jean-Antoine-Nicolas}, title={Essai sur l'application de l'analyse \`a la probabilit\'e des d\'ecisions rendues \`a la pluralit\'e des voix}, publisher={De l'Imprimerie Royale}, date={1785}, } Close amsref.
[17]
J. Corneli, I. Corwin, S. Hurder, V. Sesum, Y. Xu, E. Adams, D. Davis, M. Lee, R. Visocchi, and N. Hoffman, Double bubbles in Gauss space and spheres, Houston J. Math. 34 (2008), no. 1, 181–204. MR2383703, Show rawAMSref\bib{CCHSXADLV:08}{article}{ author={Corneli, J.}, author={Corwin, I.}, author={Hurder, S.}, author={Sesum, V.}, author={Xu, Y.}, author={Adams, E.}, author={Davis, D.}, author={Lee, M.}, author={Visocchi, R.}, author={Hoffman, N.}, title={Double bubbles in Gauss space and spheres}, journal={Houston J. Math.}, volume={34}, date={2008}, number={1}, pages={181--204}, issn={0362-1588}, review={\MR {2383703}}, } Close amsref.
[18]
A. De, E. Mossel, and J. Neeman, Majority is stablest: discrete and SoS, STOC’13—Proceedings of the 2013 ACM Symposium on Theory of Computing, ACM, New York, 2013, pp. 477–486, DOI 10.1145/2488608.2488668. MR3210809, Show rawAMSref\bib{DeMoNe:13}{article}{ author={De, Anindya}, author={Mossel, Elchanan}, author={Neeman, Joe}, title={Majority is stablest: discrete and SoS}, conference={ title={STOC'13---Proceedings of the 2013 ACM Symposium on Theory of Computing}, }, book={ publisher={ACM, New York}, }, date={2013}, pages={477--486}, review={\MR {3210809}}, doi={10.1145/2488608.2488668}, } Close amsref.
[19]
A. De, E. Mossel, and J. Neeman, Majority is stablest: Discrete and sos, Theory of Computing 12 (2016), no. 4, 1–50., Show rawAMSref\bib{DeMoNe:16}{article}{ author={De, Anindya}, author={Mossel, Elchanan}, author={Neeman, Joe}, title={Majority is stablest: Discrete and sos}, date={2016}, journal={Theory of Computing}, volume={12}, number={4}, pages={1\ndash 50}, url={http://www.theoryofcomputing.org/articles/v012a004}, } Close amsref.
[20]
P. Dubey and L. S. Shapley, Mathematical properties of the Banzhaf power index, Math. Oper. Res. 4 (1979), no. 2, 99–131, DOI 10.1287/moor.4.2.99. MR543924, Show rawAMSref\bib{DubeyShapley:79}{article}{ author={Dubey, Pradeep}, author={Shapley, Lloyd S.}, title={Mathematical properties of the Banzhaf power index}, journal={Math. Oper. Res.}, volume={4}, date={1979}, number={2}, pages={99--131}, issn={0364-765X}, review={\MR {543924}}, doi={10.1287/moor.4.2.99}, } Close amsref.
[21]
R. Eldan, A two-sided estimate for the Gaussian noise stability deficit, Invent. Math. 201 (2015), no. 2, 561–624, DOI 10.1007/s00222-014-0556-6. MR3370621, Show rawAMSref\bib{Eldan:15}{article}{ author={Eldan, Ronen}, title={A two-sided estimate for the Gaussian noise stability deficit}, journal={Invent. Math.}, volume={201}, date={2015}, number={2}, pages={561--624}, issn={0020-9910}, review={\MR {3370621}}, doi={10.1007/s00222-014-0556-6}, } Close amsref.
[22]
D. Falik and E. Friedgut, Between Arrow and Gibbard–Satterthwaite; a representation theoretic approach, Israel J. Math. 201 (2014), no. 1, 247–297, DOI 10.1007/s11856-014-1064-5. MR3265286, Show rawAMSref\bib{FalikFriedgut:14}{article}{ author={Falik, Dvir}, author={Friedgut, Ehud}, title={Between Arrow and Gibbard--Satterthwaite; a representation theoretic approach}, journal={Israel J. Math.}, volume={201}, date={2014}, number={1}, pages={247--297}, issn={0021-2172}, review={\MR {3265286}}, doi={10.1007/s11856-014-1064-5}, } Close amsref.
[23]
Y. Filmus, N. Lifshitz, D. Minzer, and E. Mossel, AND testing and robust judgement aggregation, Proceedings of the 52nd Annual ACM SIGACT Symposium on Theory of Computing, 2020, pp. 222–233., Show rawAMSref\bib{FLMM:20}{inproceedings}{ author={Filmus, Yuval}, author={Lifshitz, Noam}, author={Minzer, Dor}, author={Mossel, Elchanan}, title={{AND} testing and robust judgement aggregation}, date={2020}, booktitle={Proceedings of the 52nd Annual ACM SIGACT Symposium on Theory of Computing}, pages={222\ndash 233}, } Close amsref.
[24]
E. Friedgut, G. Kalai, N. Keller, and N. Nisan, A quantitative version of the Gibbard–Satterthwaite theorem for three alternatives, SIAM J. Comput. 40 (2011), no. 3, 934–952, DOI 10.1137/090756740. MR2823513, Show rawAMSref\bib{FKKN:11}{article}{ author={Friedgut, Ehud}, author={Kalai, Gil}, author={Keller, Nathan}, author={Nisan, Noam}, title={A quantitative version of the Gibbard--Satterthwaite theorem for three alternatives}, journal={SIAM J. Comput.}, volume={40}, date={2011}, number={3}, pages={934--952}, issn={0097-5397}, review={\MR {2823513}}, doi={10.1137/090756740}, } Close amsref.
[25]
E. Friedgut, G. Kalai, and A. Naor, Boolean functions whose Fourier transform is concentrated on the first two levels, Adv. in Appl. Math. 29 (2002), no. 3, 427–437, DOI 10.1016/S0196-8858(02)00024-6. MR1942632, Show rawAMSref\bib{FrKaNa:02}{article}{ author={Friedgut, Ehud}, author={Kalai, Gil}, author={Naor, Assaf}, title={Boolean functions whose Fourier transform is concentrated on the first two levels}, journal={Adv. in Appl. Math.}, volume={29}, date={2002}, number={3}, pages={427--437}, issn={0196-8858}, review={\MR {1942632}}, doi={10.1016/S0196-8858(02)00024-6}, } Close amsref.
[26]
E. Friedgut, G. Kalai, and N. Nisan, Elections can be manipulated often, Proceedings of the 49th annual ieee symposium on foundations of computer science (focs), 2009, pp. 243–249., Show rawAMSref\bib{FrKaNi:08}{inproceedings}{ author={Friedgut, E.}, author={Kalai, G.}, author={Nisan, N.}, title={Elections can be manipulated often}, date={2009}, booktitle={Proceedings of the 49th annual ieee symposium on foundations of computer science (focs)}, pages={243\ndash 249}, } Close amsref.
[27]
W. V Gehrlein, Condorcet’s paradox, Theory and Decision Library C, vol. 40, Springer-Verlag Berlin Heidelberg, 2006, DOI 10.1007/3-540-33799-7., Show rawAMSref\bib{Gehrlein:06}{book}{ author={Gehrlein, William~V}, title={Condorcet's paradox}, publisher={Springer-Verlag Berlin Heidelberg}, series={Theory and Decision Library C}, volume={40}, date={2006}, doi={10.1007/3-540-33799-7}, } Close amsref.
[28]
A. Gibbard, Manipulation of voting schemes: a general result, Econometrica 41 (1973), 587–601, DOI 10.2307/1914083. MR441407, Show rawAMSref\bib{Gibbard:73}{article}{ author={Gibbard, Allan}, title={Manipulation of voting schemes: a general result}, journal={Econometrica}, volume={41}, date={1973}, pages={587--601}, issn={0012-9682}, review={\MR {441407}}, doi={10.2307/1914083}, } Close amsref.
[29]
L. Gross, Logarithmic Sobolev inequalities, Amer. J. Math. 97 (1975), no. 4, 1061–1083, DOI 10.2307/2373688. MR420249, Show rawAMSref\bib{Gross:75}{article}{ author={Gross, Leonard}, title={Logarithmic Sobolev inequalities}, journal={Amer. J. Math.}, volume={97}, date={1975}, number={4}, pages={1061--1083}, issn={0002-9327}, review={\MR {420249}}, doi={10.2307/2373688}, } Close amsref.
[30]
S. Heilman, E. Mossel, and J. Neeman, Standard simplices and pluralities are not the most noise stable (abstract), Proceedings of the 2015 Conference on Innovations in Theoretical Computer Science, 2015, pp. 255–255., Show rawAMSref\bib{HeMoNe:15a}{inproceedings}{ author={Heilman, Steven}, author={Mossel, Elchanan}, author={Neeman, Joe}, title={Standard simplices and pluralities are not the most noise stable (abstract)}, organization={ACM}, date={2015}, booktitle={Proceedings of the 2015 Conference on Innovations in Theoretical Computer Science}, pages={255\ndash 255}, } Close amsref.
[31]
S. Heilman and A. Tarter, Three candidate plurality is stablest for small correlations, Preprint, arXiv:2011.05583, 2020., Show rawAMSref\bib{HeilmanTarter:20}{eprint}{ author={Heilman, Steven}, author={Tarter, Alex}, title={Three candidate plurality is stablest for small correlations}, date={2020}, arxiv={2011.05583}, } Close amsref.
[32]
M. Isaksson, G. Kindler, and E. Mossel, The geometry of manipulation—a quantitative proof of the Gibbard–Satterthwaite theorem, Combinatorica 32 (2012), no. 2, 221–250, DOI 10.1007/s00493-012-2704-1. MR2927640, Show rawAMSref\bib{IsKiMo:12}{article}{ author={Isaksson, Marcus}, author={Kindler, Guy}, author={Mossel, Elchanan}, title={The geometry of manipulation---a quantitative proof of the Gibbard--Satterthwaite theorem}, journal={Combinatorica}, volume={32}, date={2012}, number={2}, pages={221--250}, issn={0209-9683}, review={\MR {2927640}}, doi={10.1007/s00493-012-2704-1}, } Close amsref.
[33]
M. Isaksson and E. Mossel, Maximally stable Gaussian partitions with discrete applications, Israel J. Math. 189 (2012), 347–396, DOI 10.1007/s11856-011-0181-7. MR2931402, Show rawAMSref\bib{IsakssonMossel:12}{article}{ author={Isaksson, Marcus}, author={Mossel, Elchanan}, title={Maximally stable Gaussian partitions with discrete applications}, journal={Israel J. Math.}, volume={189}, date={2012}, pages={347--396}, issn={0021-2172}, review={\MR {2931402}}, doi={10.1007/s11856-011-0181-7}, } Close amsref.
[34]
J. Jendrej, K. Oleszkiewicz, and J. O. Wojtaszczyk, On some extensions of the FKN theorem, Theory Comput. 11 (2015), 445–469, DOI 10.4086/toc.2015.v011a018. MR3446023, Show rawAMSref\bib{JeOlWo:15}{article}{ author={Jendrej, Jacek}, author={Oleszkiewicz, Krzysztof}, author={Wojtaszczyk, Jakub O.}, title={On some extensions of the FKN theorem}, journal={Theory Comput.}, volume={11}, date={2015}, pages={445--469}, review={\MR {3446023}}, doi={10.4086/toc.2015.v011a018}, } Close amsref.
[35]
C. Jones, A noisy-influence regularity lemma for Boolean functions, Preprint, arXiv:1610.06950, 2016., Show rawAMSref\bib{Jones:16}{eprint}{ author={Jones, Chris}, title={A noisy-influence regularity lemma for Boolean functions}, date={2016}, arxiv={1610.06950}, } Close amsref.
[36]
J. Kahn, G. Kalai, and N. Linial, The influence of variables on Boolean functions, Proceedings of the 29th Annual Symposium on Foundations of Computer Science, 1988, pp. 68–80., Show rawAMSref\bib{KaKaLi:88}{inproceedings}{ author={Kahn, J.}, author={Kalai, G.}, author={Linial, N.}, title={The influence of variables on Boolean functions}, date={1988}, booktitle={Proceedings of the 29th Annual Symposium on Foundations of Computer Science}, pages={68\ndash 80}, } Close amsref.
[37]
G. Kalai, A Fourier-theoretic perspective on the Condorcet paradox and Arrow’s theorem, Adv. in Appl. Math. 29 (2002), no. 3, 412–426, DOI 10.1016/S0196-8858(02)00023-4. MR1942631, Show rawAMSref\bib{Kalai:02}{article}{ author={Kalai, Gil}, title={A Fourier-theoretic perspective on the Condorcet paradox and Arrow's theorem}, journal={Adv. in Appl. Math.}, volume={29}, date={2002}, number={3}, pages={412--426}, issn={0196-8858}, review={\MR {1942631}}, doi={10.1016/S0196-8858(02)00023-4}, } Close amsref.
[38]
G. Kalai, Social indeterminacy, Econometrica 72 (2004), no. 5, 1565–1581, DOI 10.1111/j.1468-0262.2004.00543.x. MR2078213, Show rawAMSref\bib{Kalai:04}{article}{ author={Kalai, Gil}, title={Social indeterminacy}, journal={Econometrica}, volume={72}, date={2004}, number={5}, pages={1565--1581}, issn={0012-9682}, review={\MR {2078213}}, doi={10.1111/j.1468-0262.2004.00543.x}, } Close amsref.
[39]
N. Keller, A tight quantitative version of Arrow’s impossibility theorem, J. Eur. Math. Soc. (JEMS) 14 (2012), no. 5, 1331–1355, DOI 10.4171/JEMS/334. MR2966653, Show rawAMSref\bib{Keller:12}{article}{ author={Keller, Nathan}, title={A tight quantitative version of Arrow's impossibility theorem}, journal={J. Eur. Math. Soc. (JEMS)}, volume={14}, date={2012}, number={5}, pages={1331--1355}, issn={1435-9855}, review={\MR {2966653}}, doi={10.4171/JEMS/334}, } Close amsref.
[40]
N. Keller, E. Mossel, and A. Sen, Geometric influences, Annals of Probability 40 (2012), no. 3, 1135–1166., Show rawAMSref\bib{KeMoSe:12}{article}{ author={Keller, N.}, author={Mossel, E.}, author={Sen, A.}, title={Geometric influences}, date={2012}, journal={Annals of Probability}, volume={40}, number={3}, pages={1135\ndash 1166}, } Close amsref.
[41]
S. Khot, G. Kindler, E. Mossel, and R. O’Donnell, Optimal inapproximability results for MAX-CUT and other 2-variable CSPs?, SIAM J. Comput. 37 (2007), no. 1, 319–357, DOI 10.1137/S0097539705447372. MR2306295, Show rawAMSref\bib{KKMO:07}{article}{ author={Khot, Subhash}, author={Kindler, Guy}, author={Mossel, Elchanan}, author={O'Donnell, Ryan}, title={Optimal inapproximability results for MAX-CUT and other 2-variable CSPs?}, journal={SIAM J. Comput.}, volume={37}, date={2007}, number={1}, pages={319--357}, issn={0097-5397}, review={\MR {2306295}}, doi={10.1137/S0097539705447372}, } Close amsref.
[42]
G. Kindler, N. Kirshner, and R. O’Donnell, Gaussian noise sensitivity and Fourier tails, Israel J. Math. 225 (2018), no. 1, 71–109, DOI 10.1007/s11856-018-1646-8. MR3805643, Show rawAMSref\bib{KiKiOd:18}{article}{ author={Kindler, Guy}, author={Kirshner, Naomi}, author={O'Donnell, Ryan}, title={Gaussian noise sensitivity and Fourier tails}, journal={Israel J. Math.}, volume={225}, date={2018}, number={1}, pages={71--109}, issn={0021-2172}, review={\MR {3805643}}, doi={10.1007/s11856-018-1646-8}, } Close amsref.
[43]
L. A Kornhauser and L. G Sager, Unpacking the court, Yale Law Jour. 96 (1986), 82., Show rawAMSref\bib{KornhauserSager:86}{article}{ author={Kornhauser, Lewis~A}, author={Sager, Lawrence~G}, title={Unpacking the court}, date={1986}, journal={Yale Law Jour.}, volume={96}, pages={82}, } Close amsref.
[44]
M. Ledoux, Remarks on Gaussian noise stability, Brascamp-Lieb and Slepian inequalities, Geometric aspects of functional analysis, Lecture Notes in Math., vol. 2116, Springer, Cham, 2014, pp. 309–333, DOI 10.1007/978-3-319-09477-9_20. MR3364694, Show rawAMSref\bib{Ledoux:14}{article}{ author={Ledoux, Michel}, title={Remarks on Gaussian noise stability, Brascamp-Lieb and Slepian inequalities}, conference={ title={Geometric aspects of functional analysis}, }, book={ series={Lecture Notes in Math.}, volume={2116}, publisher={Springer, Cham}, }, date={2014}, pages={309--333}, review={\MR {3364694}}, doi={10.1007/978-3-319-09477-9\_20}, } Close amsref.
[45]
C. List and P. Pettit, Aggregating sets of judgments: An impossibility result, Economics and Philosophy 18 (2002), no. 1, 89–110., Show rawAMSref\bib{ListPettit:02}{article}{ author={List, Christian}, author={Pettit, Philip}, title={Aggregating sets of judgments: An impossibility result}, date={2002}, journal={Economics and Philosophy}, volume={18}, number={1}, pages={89\ndash 110}, } Close amsref.
[46]
C. List and P. Pettit, Aggregating sets of judgments: two impossibility results compared, Synthese 140 (2004), no. 1-2, 207–235, DOI 10.1023/B:SYNT.0000029950.50517.59. With a comment by Isaac Levi. MR2077375, Show rawAMSref\bib{ListPettit:04}{article}{ author={List, Christian}, author={Pettit, Philip}, title={Aggregating sets of judgments: two impossibility results compared}, note={With a comment by Isaac Levi}, journal={Synthese}, volume={140}, date={2004}, number={1-2}, pages={207--235}, issn={0039-7857}, review={\MR {2077375}}, doi={10.1023/B:SYNT.0000029950.50517.59}, } Close amsref.
[47]
E. Milman and J. Neeman, The Gaussian double-bubble conjecture, Preprint, arXiv:1801.09296, 2018., Show rawAMSref\bib{MilmanNeeman:18a}{eprint}{ author={Milman, Emanuel}, author={Neeman, Joe}, title={The {G}aussian double-bubble conjecture}, date={2018}, arxiv={1801.09296}, } Close amsref.
[48]
E. Milman and J. Neeman, The Gaussian multi-bubble conjecture, Preprint, arXiv:1805.10961, 2018., Show rawAMSref\bib{MilmanNeeman:18b}{eprint}{ author={Milman, Emanuel}, author={Neeman, Joe}, title={The {G}aussian multi-bubble conjecture}, date={2018}, arxiv={1805.10961}, } Close amsref.
[49]
E. Mossel, Gaussian bounds for noise correlation of functions and tight analysis of long codes, Foundations of Computer Science, 2008 (FOCS 08), 2008, pp. 156–165., Show rawAMSref\bib{Mossel:08}{inproceedings}{ author={Mossel, E.}, title={{G}aussian bounds for noise correlation of functions and tight analysis of long codes}, date={2008}, booktitle={Foundations of Computer Science, 2008 (FOCS 08)}, publisher={IEEE}, pages={156\ndash 165}, url={http://www.stat.berkeley.edu/~mossel/publications/MosselE-Gaussian.pdf}, } Close amsref.
[50]
E. Mossel, Arrow’s impossibility theorem without unanimity, Preprint, arXiv:0901.4727, 2009., Show rawAMSref\bib{Mossel:09a}{eprint}{ author={Mossel, E.}, title={Arrow's impossibility theorem without unanimity}, date={2009}, arxiv={0901.4727}, } Close amsref.
[51]
E. Mossel, Gaussian bounds for noise correlation of functions, Geom. Funct. Anal. 19 (2010), no. 6, 1713–1756, DOI 10.1007/s00039-010-0047-x. MR2594620, Show rawAMSref\bib{Mossel:10}{article}{ author={Mossel, Elchanan}, title={Gaussian bounds for noise correlation of functions}, journal={Geom. Funct. Anal.}, volume={19}, date={2010}, number={6}, pages={1713--1756}, issn={1016-443X}, review={\MR {2594620}}, doi={10.1007/s00039-010-0047-x}, } Close amsref.
[52]
E. Mossel, A quantitative Arrow theorem, Probab. Theory Related Fields 154 (2012), no. 1-2, 49–88, DOI 10.1007/s00440-011-0362-7. MR2981417, Show rawAMSref\bib{Mossel:12}{article}{ author={Mossel, Elchanan}, title={A quantitative Arrow theorem}, journal={Probab. Theory Related Fields}, volume={154}, date={2012}, number={1-2}, pages={49--88}, issn={0178-8051}, review={\MR {2981417}}, doi={10.1007/s00440-011-0362-7}, } Close amsref.
[53]
E. Mossel, Gaussian bounds for noise correlation of resilient functions, Israel J. Math. 235 (2020), no. 1, 111–137, DOI 10.1007/s11856-019-1951-x. MR4068779, Show rawAMSref\bib{Mossel:20resilient}{article}{ author={Mossel, Elchanan}, title={Gaussian bounds for noise correlation of resilient functions}, journal={Israel J. Math.}, volume={235}, date={2020}, number={1}, pages={111--137}, issn={0021-2172}, review={\MR {4068779}}, doi={10.1007/s11856-019-1951-x}, } Close amsref.
[54]
E. Mossel, Probabilistic aspects of voting, intransitivity and manipulation, Preprint, arXiv:2012.10352, 2020., Show rawAMSref\bib{Mossel:20LN}{eprint}{ author={Mossel, Elchanan}, title={Probabilistic aspects of voting, intransitivity and manipulation}, date={2020}, arxiv={2012.10352}, } Close amsref.
[55]
E. Mossel and J. Neeman, Robust optimality of Gaussian noise stability, J. Eur. Math. Soc. (JEMS) 17 (2015), no. 2, 433–482, DOI 10.4171/JEMS/507. MR3317748, Show rawAMSref\bib{MosselNeeman:15b}{article}{ author={Mossel, Elchanan}, author={Neeman, Joe}, title={Robust optimality of Gaussian noise stability}, journal={J. Eur. Math. Soc. (JEMS)}, volume={17}, date={2015}, number={2}, pages={433--482}, issn={1435-9855}, review={\MR {3317748}}, doi={10.4171/JEMS/507}, } Close amsref.
[56]
E. Mossel and R. O’Donnell, On the noise sensitivity of monotone functions, Random Structures Algorithms 23 (2003), no. 3, 333–350, DOI 10.1002/rsa.10097. MR1999039, Show rawAMSref\bib{MosselOdonnell:03}{article}{ author={Mossel, Elchanan}, author={O'Donnell, Ryan}, title={On the noise sensitivity of monotone functions}, journal={Random Structures Algorithms}, volume={23}, date={2003}, number={3}, pages={333--350}, issn={1042-9832}, review={\MR {1999039}}, doi={10.1002/rsa.10097}, } Close amsref.
[57]
E. Mossel, R. O’Donnell, and K. Oleszkiewicz, Noise stability of functions with low influences: invariance and optimality (extended abstract), 46th annual ieee symposium on foundations of computer science (focs 2005), 23-25 october 2005, pittsburgh, pa, usa, proceedings, 2005, pp. 21–30., Show rawAMSref\bib{MoOdOl:05}{inproceedings}{ author={Mossel, E.}, author={O'Donnell, R.}, author={Oleszkiewicz, K.}, title={Noise stability of functions with low influences: invariance and optimality (extended abstract)}, date={2005}, booktitle={46th annual ieee symposium on foundations of computer science (focs 2005), 23-25 october 2005, pittsburgh, pa, usa, proceedings}, publisher={IEEE Computer Society}, pages={21\ndash 30}, url={http://www.stat.berkeley.edu/~mossel/publications/moo-focs-final.pdf}, } Close amsref.
[58]
E. Mossel, R. O’Donnell, and K. Oleszkiewicz, Noise stability of functions with low influences: invariance and optimality, Annals of Mathematics 171 (2010), no. 1, 295–341., Show rawAMSref\bib{MoOdOl:10}{article}{ author={Mossel, E.}, author={O'Donnell, R.}, author={Oleszkiewicz, K.}, title={Noise stability of functions with low influences: invariance and optimality}, date={2010}, journal={Annals of Mathematics}, volume={171}, number={1}, pages={295\ndash 341}, url={http://front.math.ucdavis.edu/0503.5503}, } Close amsref.
[59]
E. Mossel, R. O’Donnell, O. Regev, J. E. Steif, and B. Sudakov, Non-interactive correlation distillation, inhomogeneous Markov chains, and the reverse Bonami-Beckner inequality, Israel J. Math. 154 (2006), 299–336. MR2254545, Show rawAMSref\bib{MORSS:06}{article}{ author={Mossel, E.}, author={O'Donnell, R.}, author={Regev, O.}, author={Steif, J.~E.}, author={Sudakov, B.}, title={Non-interactive correlation distillation, inhomogeneous {M}arkov chains, and the reverse {B}onami-{B}eckner inequality}, date={2006}, journal={Israel J. Math.}, volume={154}, pages={299\ndash 336}, url={http://front.math.ucdavis.edu/0410.5560}, review={\MR {MR2254545}}, } Close amsref.
[60]
E. Mossel, K. Oleszkiewicz, and A. Sen, On reverse hypercontractivity, Geom. Funct. Anal. 23 (2013), no. 3, 1062–1097, DOI 10.1007/s00039-013-0229-4. MR3061780, Show rawAMSref\bib{MoOlSe:13}{article}{ author={Mossel, Elchanan}, author={Oleszkiewicz, Krzysztof}, author={Sen, Arnab}, title={On reverse hypercontractivity}, journal={Geom. Funct. Anal.}, volume={23}, date={2013}, number={3}, pages={1062--1097}, issn={1016-443X}, review={\MR {3061780}}, doi={10.1007/s00039-013-0229-4}, } Close amsref.
[61]
E. Mossel and M. Z. Rácz, A quantitative Gibbard–Satterthwaite theorem without neutrality [extended abstract], STOC’12—Proceedings of the 2012 ACM Symposium on Theory of Computing, ACM, New York, 2012, pp. 1041–1060, DOI 10.1145/2213977.2214071. MR2961564, Show rawAMSref\bib{MosselRacz:12}{article}{ author={Mossel, Elchanan}, author={R\'{a}cz, Mikl\'{o}s Z.}, title={A quantitative Gibbard--Satterthwaite theorem without neutrality [extended abstract]}, conference={ title={STOC'12---Proceedings of the 2012 ACM Symposium on Theory of Computing}, }, book={ publisher={ACM, New York}, }, date={2012}, pages={1041--1060}, review={\MR {2961564}}, doi={10.1145/2213977.2214071}, } Close amsref.
[62]
E. Mossel and M. Z. Rácz, A quantitative Gibbard–Satterthwaite theorem without neutrality, Combinatorica 35 (2015), no. 3, 317–387, DOI 10.1007/s00493-014-2979-5. MR3367129, Show rawAMSref\bib{MosselRacz:15}{article}{ author={Mossel, Elchanan}, author={R\'{a}cz, Mikl\'{o}s Z.}, title={A quantitative Gibbard--Satterthwaite theorem without neutrality}, journal={Combinatorica}, volume={35}, date={2015}, number={3}, pages={317--387}, issn={0209-9683}, review={\MR {3367129}}, doi={10.1007/s00493-014-2979-5}, } Close amsref.
[63]
J. Neeman, A multidimensional version of noise stability, Electron. Commun. Probab. 19 (2014), no. 72, 10, DOI 10.1214/ECP.v19-3005. MR3274518, Show rawAMSref\bib{Neeman:14}{article}{ author={Neeman, Joe}, title={A multidimensional version of noise stability}, journal={Electron. Commun. Probab.}, volume={19}, date={2014}, pages={no. 72, 10}, review={\MR {3274518}}, doi={10.1214/ECP.v19-3005}, } Close amsref.
[64]
I. Nehama, Approximately classic judgement aggregation, Ann. Math. Artif. Intell. 68 (2013), no. 1-3, 91–134, DOI 10.1007/s10472-013-9358-6. MR3145873, Show rawAMSref\bib{Nehama:2013}{article}{ author={Nehama, Ilan}, title={Approximately classic judgement aggregation}, journal={Ann. Math. Artif. Intell.}, volume={68}, date={2013}, number={1-3}, pages={91--134}, issn={1012-2443}, review={\MR {3145873}}, doi={10.1007/s10472-013-9358-6}, } Close amsref.
[65]
E. Nelson, A quartic interaction in two dimensions, Mathematical Theory of Elementary Particles (Proc. Conf., Dedham, Mass., 1965), M.I.T. Press, Cambridge, Mass., 1966, pp. 69–73. MR0210416, Show rawAMSref\bib{Nelson:66}{article}{ author={Nelson, Edward}, title={A quartic interaction in two dimensions}, conference={ title={Mathematical Theory of Elementary Particles}, address={Proc. Conf., Dedham, Mass.}, date={1965}, }, book={ publisher={M.I.T. Press, Cambridge, Mass.}, }, date={1966}, pages={69--73}, review={\MR {0210416}}, } Close amsref.
[66]
E. Nelson, The free Markoff field, J. Functional Analysis 12 (1973), 211–227, DOI 10.1016/0022-1236(73)90025-6. MR0343816, Show rawAMSref\bib{Nelson:73}{article}{ author={Nelson, Edward}, title={The free Markoff field}, journal={J. Functional Analysis}, volume={12}, date={1973}, pages={211--227}, review={\MR {0343816}}, doi={10.1016/0022-1236(73)90025-6}, } Close amsref.
[67]
R. G. Niemi and H. F. Weisberg, A mathematical solution for the probability of paradox of voting, Behavioral Science 13 (1968), 317–323., Show rawAMSref\bib{NiemiWeisberg:68}{article}{ author={Niemi, R.~G.}, author={Weisberg, H.~F.}, title={A mathematical solution for the probability of paradox of voting}, date={1968}, journal={Behavioral Science}, volume={13}, pages={317\ndash 323}, } Close amsref.
[68]
N. Nisan, Algorithmic game theory / economics. postname: From Arrow to Fourier (2009), http://agtb.wordpress.com/2009/03/31/from-arrow-to-fourier/., Show rawAMSref\bib{Nisan:FromArrowtoFourier}{webpage}{ author={Nisan, N.}, title={Algorithmic game theory / economics. postname: {F}rom {A}rrow to {F}ourier}, date={2009}, url={http://agtb.wordpress.com/2009/03/31/from-arrow-to-fourier/}, } Close amsref.
[69]
R. O’Donnell, Analysis of Boolean functions, Cambridge University Press, New York, 2014, DOI 10.1017/CBO9781139814782. MR3443800, Show rawAMSref\bib{ODonnell:14}{book}{ author={O'Donnell, Ryan}, title={Analysis of Boolean functions}, publisher={Cambridge University Press, New York}, date={2014}, pages={xx+423}, isbn={978-1-107-03832-5}, review={\MR {3443800}}, doi={10.1017/CBO9781139814782}, } Close amsref.
[70]
R. O’Donnell, R. Servedio, L.-Y. Tan, and A. Wan, A regularity lemma for low noisy-influences, 2010. Unpublished manuscript., Show rawAMSref\bib{OSTW:10}{misc}{ author={O'Donnell, Ryan}, author={Servedio, Rocco}, author={Tan, Li-Yang}, author={Wan, Andrew}, title={A regularity lemma for low noisy-influences}, date={2010}, note={Unpublished manuscript}, } Close amsref.
[71]
M. A. Satterthwaite, Strategy-proofness and Arrow’s conditions: Existence and correspondence theorems for voting procedures and social welfare functions, J. Econom. Theory 10 (1975), no. 2, 187–217, DOI 10.1016/0022-0531(75)90050-2. MR414051, Show rawAMSref\bib{Satterthwaite:75}{article}{ author={Satterthwaite, Mark Allen}, title={Strategy-proofness and Arrow's conditions: Existence and correspondence theorems for voting procedures and social welfare functions}, journal={J. Econom. Theory}, volume={10}, date={1975}, number={2}, pages={187--217}, issn={0022-0531}, review={\MR {414051}}, doi={10.1016/0022-0531(75)90050-2}, } Close amsref.
[72]
W. Sheppard, On the application of the theory of error to cases of normal distribution and normal correlation, Phil. Trans. Royal Soc. London 192 (1899), 101–168., Show rawAMSref\bib{Sheppard:99}{article}{ author={Sheppard, W.}, title={On the application of the theory of error to cases of normal distribution and normal correlation}, date={1899}, journal={Phil.\ Trans.\ Royal Soc.\ London}, volume={192}, pages={101\ndash 168}, } Close amsref.
[73]
Wikipedia contributors, Marquis de Condorcet—Wikipedia, the free encyclopedia, 2019. [Online; accessed 2019]., Show rawAMSref\bib{wiki:Condorcet}{misc}{ author={{Wikipedia contributors}}, title={Marquis de Condorcet---{W}ikipedia{,} the free encyclopedia}, date={2019}, url={https://en.wikipedia.org/w/index.php?title=Plagiarism&oldid=5139350}, note={[Online; accessed 2019]}, } Close amsref.
[74]
R. Wilson, Social choice theory without the Pareto principle, J. Econom. Theory 5 (1972), no. 3, 478–486, DOI 10.1016/0022-0531(72)90051-8. MR449494, Show rawAMSref\bib{Wilson:72}{article}{ author={Wilson, Robert}, title={Social choice theory without the Pareto principle}, journal={J. Econom. Theory}, volume={5}, date={1972}, number={3}, pages={478--486}, issn={0022-0531}, review={\MR {449494}}, doi={10.1016/0022-0531(72)90051-8}, } Close amsref.
[75]
L. Yu and V. Y. Tan, On non-interactive simulation of binary random variables, IEEE Transactions on Information Theory 67 (2021), no. 4, 2528–2538., Show rawAMSref\bib{YuTan:21}{article}{ author={Yu, Lei}, author={Tan, Vincent~YF}, title={On non-interactive simulation of binary random variables}, date={2021}, journal={IEEE Transactions on Information Theory}, volume={67}, number={4}, pages={2528\ndash 2538}, } Close amsref.

Article Information

MSC 2020
Primary: 60-02 (Research exposition (monographs, survey articles) pertaining to probability theory)
Secondary: 91B12 (Voting theory), 91B14 (Social choice)
Author Information
Elchanan Mossel
Massachusetts Institute of Technology, Cambridge, Massachusetts
elmos@mit.edu
ORCID
MathSciNet
Additional Notes

The author was partially supported by NSF Grant CCF 1665252, DMS-1737944, DOD ONR grant N00014-17-1-2598, Simons Investigator award (622132), and Vannevar Bush Faculty Fellowship ONR-N00014-20-1-2826.

Journal Information
Bulletin of the American Mathematical Society, Volume 59, Issue 3, ISSN 1088-9485, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2021 by Elchanan Mossel
Article References

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.