A Parallel Repetition Theorem for the GHZ Game
Abstract
We prove that parallel repetition of the (3-player) GHZ game reduces the value of the game polynomially fast to 0. That is, the value of the GHZ game repeated in parallel times is at most . Previously, only a bound of , where is the inverse Ackermann function, was known [Ver96].
The GHZ game was recently identified by Dinur, Harsha, Venkat and Yuen as a multi-player game where all existing techniques for proving strong bounds on the value of the parallel repetition of the game fail. Indeed, to prove our result we use a completely new proof technique. Dinur, Harsha, Venkat and Yuen speculated that progress on bounding the value of the parallel repetition of the GHZ game may lead to further progress on the general question of parallel repetition of multi-player games. They suggested that the strong correlations present in the GHZ question distribution represent the “hardest instance” of the multi-player parallel repetition problem [DHVY17].
Another motivation for studying the parallel repetition of the GHZ game comes from the field of quantum information. The GHZ game, first introduced by Greenberger, Horne and Zeilinger [GHZ89], is a central game in the study of quantum entanglement and has been studied in numerous works. For example, it is used for testing quantum entanglement and for device-independent quantum cryptography. In such applications a game is typically repeated to reduce the probability of error, and hence bounds on the value of the parallel repetition of the game may be useful.
1 Introduction
In a -player game, players are given correlated “questions” sampled from a distribution and must produce corresponding “answers” such that satisfy a fixed predicate . Crucially, the players are not allowed to communicate amongst themselves after receiving their questions (but they may agree upon a strategy beforehand). The value of the game is the probability with which the players can win with an optimal strategy. Multi-player games play a central role in theoretical computer science due to their intimate connection with multi-prover interactive proofs (MIPs) [BGKW88], hardness of approximation [FGL+91], communication complexity [PRW97, BJKS04], and the EPR paradox and non-local games [EPR35, CHTW04]
One basic operation on multi-player games is parallel repetition. In the -wise parallel repetition of a game, question tuples are sampled independently for . The player is given , and is required to produce . The players win if for every , is a winning answer for questions . Parallel repetition was first proposed in [FRS94] as an intuitive attempt to reduce the value of a game from to , but in general this is not what happens [For89, Fei91, FV96, Raz11]. The actual effect is far more subtle and a summary of some of the known results is given in Table 1.
Two-player games | -player games | |
---|---|---|
Classical | [Raz98] | [Ver96] |
Entangled | [Yue16] | (trivial) |
Non-Signaling | [Hol09] | [HY19] |
Much less is known about games with three or more players than about two-player games. Only very weak bounds are known on how -wise parallel repetition decreases the value of a three-player game (as a function of ). There is a similar gap in our understanding when players are allowed to share entangled state; in fact, no bounds here are known whatsoever in the three-player case. If players are more generally allowed to use any no-signaling strategy, then there are in fact counterexamples (lower bounds) showing that parallel repetition may utterly fail to reduce the (no-signaling) value of a three-player game.
1.1 The GHZ Game
The GHZ game, which we will denote by , is a three-player game with query distribution that is uniform on . To win, players are required on input to produce such that . It is easily verified that the value of is .
Dinur et al. [DHVY17] identified the GHZ game as a simple example of a game for which we do not know exponential decay bounds, writing
“We suspect that progress on bounding the value of the parallel repetition of the GHZ game will lead to further progress on the general question.”
and
“We believe that the strong correlations present in the GHZ question distribution represent the “hardest instance” of the multiplayer parallel repetition problem. Existing techniques from the two-player case (which we leverage in this paper) appear to be incapable of analyzing games with question distributions with such strong correlations.”
The GHZ game also plays an important role in quantum information theory and in particular in entanglement testing and device-independent quantum cryptography. Its salient properties are that it is an XOR game for which quantum (entangled) players can play perfectly, but classical players can win only with probability strictly less than [MS13]. No such two-player game is known. Moreover, the GHZ game has the so called, self testing property, that all quantum strategies that achieve value 1 are essentially equivalent. This property is important for entanglement testing and device-independent quantum cryptography.
Prior to our work, the best known parallel repetition bound for the GHZ game was due to Verbitsky [Ver96], who observed a connection between parallel repetition and the density Hales-Jewett theorem from Ramsey theory [FK91]. Using modern quantitative versions of this theorem [Pol12], Verbitsky’s result implies a bound of approximately , where is the inverse Ackermann function.
We prove a bound of .
2 Technical Overview
To prove our parallel repetition theorem for the GHZ game we show that for an arbitrary strategy, even if we condition on that strategy winning in several coordinates , there still exists some coordinate in which that strategy loses with significant probability. We consider the finer-grained event that also specifies specific queries and answers in coordinates , and abstract it out as a sufficiently dense product event over the three players’ inputs.
Given an arbitrary product event that occurs with sufficiently high probability, we show that some coordinate of is hard. We do this in three high-level steps:
-
1.
We first prove this for the simpler case in which is an affine subspace of . In fact, we show in this case that many coordinates of are hard.
-
2.
We then prove that when is arbitrary, can be written as a convex combination of components , where is a large affine subspace, with most such components “indistinguishable” from . Specifically, our main requirement is that for all sufficiently compressing linear functions on , the KL divergence of from is small, where we sample and .
-
3.
With this notion of indistinguishability, we prove that if is indistinguishable from , then the GHZ game (or any game with a constant-sized answer alphabet) is roughly as hard in every coordinate with query distribution as with .
We conclude that for many coordinates , there is a significant portion of for which the coordinate is hard. We emphasize that unlike all previous parallel repetition bounds, our proof does not construct a local embedding of into for general .
Local Embeddability in Affine Subspaces
We first show that if is any affine subspace of sufficiently low codimension in , then there exist many coordinates for which is locally embeddable in the coordinate of the conditional distribution . In fact, it will suffice for us to consider only affine “power” subspaces, i.e. of the form for some linear subspace in and vector . Let denote the queries in each of the repetitions.
Our observation is that when is affine there exists a subset of coordinates with such that for any , depends on only via the differences . Indeed, if and if each is given by an affine equation for a sufficiently “skinny” matrix , then by the pigeonhole principle there must exist two distinct subset row-sums of with equal values. By considering the symmetric difference of these subsets, and using the fact that we are working over , there is a set such that the -subset row-sum of is . Thus the value of is unchanged if is subtracted from for every .
As a result, the players can all sample and , which are independent of , using shared randomness. On input , the player can locally compute from and .
Pseudo-Affine Decompositions
At a high level, we next show that if is an arbitrary product event (with sufficient probability mass) then has a “pseudo-affine decomposition”. That is, there is a partition of into affine subspaces such that if is a random part of (as weighted by ), then any strategy for can be extended to a strategy for that is similarly successful in expectation.
To construct , we prove the following sufficient conditions for to be a pseudo-affine decomposition:
-
•
When is a random part of (as weighted by ), the distributions and are indistinguishable to all sufficiently compressing linear distinguishers. That is, if is an affine shift of , then for all subspaces of sufficiently small co-dimension, the distributions and are statistically close modulo .
-
•
Each part of is in fact an affine shift of a product space for some linear space .
We construct satisfying these conditions iteratively. Starting with the singleton partition, as long as a random part of has some subspace for which and are distributed differently mod , we replace each part of by all the affine shifts of in . We show that this process cannot be repeated too many times when has sufficient density.
Pseudorandomness Preserves Hardness
The high-level reason these conditions suffice is because for any strategy , they enable us to refine to a partition such that when is sampled from for a random part in , the distribution of is as if were sampled uniformly from (i.e. with , , and mutually independent). Moreover, when we construct we partition each part of into all affine shifts of some linear space where the codimension of in is not too large. Thus the strategy on effectively has the players acting as independent (randomized) functions of their inputs modulo . Such strategies generalize to by the first property of pseudo-affine decompositions stated above.
To construct , we ensure that is uncorrelated with every affine function on when is a random part of , and then prove the desired independence by Fourier analysis. We construct by iterative refinement of . Start by considering a random part of . Whenever is correlated with an affine -valued function , replace in by and , and do this in parallel for all parts of . We show that this cannot be repeated too many times, and thus we quickly arrive at our desired .
3 Preliminaries
In this section we describe some preliminary definitions that are somewhat specific to this work. More standard preliminaries are given in Appendices A and B.
3.1 Set Theory
Definition 3.1.
For any set , a partition of is a pairwise disjoint set of subsets of , whose union is all of .
If is a partition of and is an element of , we write to denote the (unique) element of that contains .
3.2 Linear Algebra
If is a linear subspace of , we write rather than to emphasize that is a subspace rather than an unstructured subset.
We crucially rely on the Cauchy-Schwarz inequality:
Definition 3.2 (Inner Product Space).
A real inner product space is a vector space over together with an operation satisfying the following axioms for all :
-
•
Symmetry: .
-
•
Linearity in the first111Because of symmetry, this implies also linearity in the second argument, aka bilinearity. argument: .
-
•
Positive Definiteness: if .
Theorem 3.3 (Cauchy-Schwarz).
In any inner product space, it holds for all vectors and that .
3.3 Multi-Player Games
In parallel repetition we often work with Cartesian product sets of the form . For these sets, we will use superscripts to index the outer product and subscripts to index the inner product. That is, we view elements of as tuples , where the component of is . We will also write to denote the vector . If is a collection of subsets indexed by subscripts, we write or to denote the set . Similarly, if is a product set , we say is a product function if for .
Definition 3.4 (Multi-player Games).
A -player game is a tuple , where and are finite sets, is a probability measure on , and is a “winning probability” predicate.
Definition 3.5 (Parallel Repetition).
Given a -player game , its -fold parallel repetition, denoted , is defined as the -player game , where .
Definition 3.6.
The success probability of a function in a -player game is
Definition 3.7.
The value of a -player game , denoted , is the maximum, over all functions , of .
Fact 3.8.
Randomized strategies are no better than deterministic strategies.
Definition 3.9 (Value in coordinate).
If is a game (with a product winning predicate), the value of in the coordinate, denoted , is the value of the game , where .
Definition 3.10 (Game with Modified Query Distribution).
If is a game, and is a probability measure on , we write to denote the game .
4 Key Lemmas
In this section, we give some Fourier-analytic conditions (see Appendix B for the basics of Fourier analysis) that imply independence of random variables under the (parallel repeated) GHZ query distribution.
It will be convenient for us to work with probability distributions in terms of their densities (see Appendix A for basic probability definitions and notation).
Definition 4.1 (Probability Densities).
If is a probability distribution with support contained in , then the density of in is
If is unspecified, then by default it is taken to be .
Lemma 4.2.
Let be a (finite) vector space over , let be uniform on , and let be uniform on .
For any subset of ,
where denotes the density in of the uniform distribution on .
Proof.
Let denote the density in of . That is,
Then
(Plancherel) | (1) |
We now compute and . We start by noting that the dual space is isomorphic to . That is, each character is of the form for some (uniquely determined) and conversely, each choice of gives rise to some .
The Fourier transform of is given by
(2) |
Corollary 4.3.
Proof.
For any probability density function , we have , so
Lemma 4.4.
Let be a (finite) vector space over , let be uniform on , let be uniform on , and let denote the identity222Specifically, with the formalism of random variables as functions on a sample space, we mean that is the identity function, mapping to . random variable on . Let be a -valued random variable for each , let , and let .
Let be a subspace of . If for all ,
(4) |
then
Proof.
For , we will write to denote the set . Recall that denotes the set of all cosets . For every , every , and every , define to be the indicator for the set . Define to be the density (in ) of the uniform distribution on . That is,
is easily seen to be related to as
With this notation, our assumption that Eq. 4 holds (for all ) is equivalent to assuming that for all ,
(5) |
This is because for all , the distribution is uniform on .
In general for , we have (by Corollary 4.3) that for any ,
(6) |
because:
-
•
the event is a product event , where each depends only on or equivalently on ,
-
•
the distribution is uniform on , and
-
•
the distribution is uniform on .
Thus we have
Now, we apply Cauchy-Schwarz on the inner product space whose elements are real-valued functions of , and where the inner product is defined by . This bounds the above by
By the independence of and under for , this is equal to
But the function is just , so the above is
which by the definition of expectation is
We use Eq. 5 to bound this by
(Parseval’s Theorem) | |||
But for , the supports of and are disjoint, so this is at most . ∎
5 Local Embeddability in Affine Subspaces
In this section we show that the parallel repeated GHZ query distribution has many coordinates in which the GHZ query distribution can be locally embedded, even conditioned on any affine event of low co-dimension. We first recall the notion of a local embedding.
Definition 5.1.
Let be a finite set, let and be positive integers, let be a probability distribution on , and let be a probability distribution on .
We say that is locally embeddable in the coordinate of if there exists a probability distribution on a set and functions such that when sampling , , if denotes the random variable
then:
-
1.
The probability law of is exactly .
-
2.
It holds with probability that .
Proposition 5.2.
Let and be positive integers with . Let denote the GHZ query distribution (uniform on the set ), and let be an affine shift of for a subspace of codimension with .
Then there exist at least distinct values of for which is locally embeddable in the coordinate of .
Proof.
Suppose otherwise. Without loss of generality, suppose that the coordinates that are not locally embeddable include the first coordinates (otherwise, can be permuted to make this so). That is, for each , is not locally embeddable in the coordinate of .
Let the defining equations for be written as
for some choice of , and let be such that .
Because , the pigeonhole principle implies that there exist two distinct sets such that
where recall that denotes the row of . Thus, there is a non-empty subset such that
(7) |
Fix some such . We will show that for any , is locally embeddable in the coordinate of , which is a contradiction. Let denote the -valued random variable given by the identity function.
Claim 5.3.
For any , the distribution is identical to (i.e., uniformly random on ).
Proof.
Let be given. It suffices to show that for every , there is a bijection such that satisfies if and only if satisfies . Such a bijection can be constructed by defining, for all ,
clearly is an injective map from to and satisfies , so the only remaining thing to check is that it indeed maps into . This is true because it preserves . Indeed, for any ,
For any , let denote the random variable .
Claim 5.4.
For any , it holds in that and are independent.
Proof.
Equivalently (using the definition of ), let denote the event that , i.e. for all ,
We need to show that in , the random variables and are conditionally independent given . To show this, we rely on the following fact:
Fact 5.5.
If and are any independent random variables, and if is any event that depends only on (and occurs with non-zero probability), then and are conditionally independent given .
It is clear that and are independent in . It is also the case that depends only on : is defined by the constraint that for all ,
(by Eq. 7) | ||||
We now put everything togther. Fix any . We construct a local embedding of into the coordinate of . For each , we define such that for each :
Define the distribution to be the distribution on obtained by independently sampling and , then defining
It clearly holds with probability that .
Claim 5.6.
.
Proof.
By definition, it is immediate that: and .
Finally, is fully determined by and , which are independent in both (because and are sampled independently in the definition of ) and (by Claim 5.4). ∎
We have constructed an embedding of into one of the first coordinates of , which is the desired contradiction. ∎
6 Decomposition Into Pseudorandom Affine Components
In this section we show that if is an arbitrary event with sufficient probability mass under , then can be decomposed into components with affine support that are “similar” to corresponding components of . We will call such components pseudorandom.
We say that is an affine partition of to mean that:
-
•
Each part of has the form where is a subspace of , and
-
•
Each has the same dimension, which we refer to as the dimension of and denote by . The codimension of is defined to be .
Definition 6.1.
If is an affine shift of a vector space (for ), we say that a -valued random variable is -close to if for all linear functions we have , where denotes the function mapping
We write to denote the minimum for which is -close to .
We remark that is a non-decreasing function of .
Lemma 6.2.
Let denote the distribution , let be the identity random variable, let be an event with , and let . For any and any , there exists an affine partition of , of codimension at most , such that:
(8) |
Proof.
We construct the claimed partition iteratively. Start with the trivial -dimensional affine partition . Whenever is a partition for which Eq. 8 does not hold, there exists a function that:
-
•
When restricted to any part of , is of the form for some linear function , and
-
•
(9)
Without loss of generality, we additionally assume that each is “full rank” when restricted to . That is, if is an affine shift of , where has dimension , then the restriction of to is a linear map of rank . It is clear that any may be modified to be full rank without decreasing the KL divergence of Eq. 9.
Then by the chain rule for KL divergences,
(10) |
The left-hand side of Eq. 10 is equivalent to
with , which is an affine partition of dimension at least .
Thus with the non-negative potential function
we have . But , so there must exist for which Eq. 8 holds with , which has co-dimension at most . ∎
7 Pseudorandomness Preserves Hardness
Proposition 7.1.
Let be an affine shift of a linear subspace and let be a the uniform distribution on , which we assume to be non-empty. Let denote the identity random variable, let be an event with , and define . Suppose that is -close to as in Definition 6.1, for satisfying .
Then for each , we have .
Proof.
Fix to be any coordinate, and let be an arbitrary strategy. Let denote .
Claim 7.2.
There exists a subspace of codimension at most such that:
-
•
The coordinate of depends only on .
-
•
For all ,
where denotes the uniform distribution on .
Proof.
Start with (this ensures that any subspace satisfies the first desired property). Define a potential function
which is clearly non-negative. Additionally, (and in particular ) is at most because for any subspace and any , the entropy chain rule implies
(in the first step we used the fact that is a function of .
For , define to maximize
and define . By the entropy chain rule, we have .
Since the initial potential is at most , and all potentials are at least , there must be some for which . The corresponding is the desired subspace of . ∎
By assumption of Proposition 7.1 (together with Pinsker’s inequality), and are -close in total variational distance. We thus have that
(11) |
by the general fact that if and are two distributions that are -close in total variational distance, and if is a -bounded random variable, then .
We now obtain a probabilistic lower bound on . We first lower bound its log-expectation:
(Fact A.17) | ||||
Markov’s inequality then implies that for any ,
(12) |
Combining Eq. 12 with Eq. 11 and Fact A.18, we get
Since this holds for all and because , Corollary C.2 implies that
(13) |
where the last inequality follows from our assumption that .
Putting everything together, we have
where denotes closeness in total variational distance.
But is just the distribution on obtained by sampling , , where is the following randomized strategy. On input , uses local randomness to sample and output . By Fact 3.8, the probability that (which is well-defined because is a function of ) is at most .
We thus have
Since this holds for arbitrary , we have . ∎
8 Proof of Main Theorem
Theorem 8.1.
If denotes the GHZ game, then .
Proof.
Recall .
Let denote ; that is is uniform on . Let be any product event in with (where is a parameter we will specify later), and let denote .
Let be a parameter we will specify later, and let . Recall our definition of (Definition 6.1). Lemma 6.2 states that there exists an affine partition of , of codimension at most , such that:
Moreover,
Markov’s inequality thus implies that with probability at least when sampling , it holds that and . Call such a pseudorandom, and let denote the set of pseudorandom .
By Proposition 7.1, for each pseudorandom we have
(14) |
as long as
(15) |
where is a parameter we will specify later.
By Proposition 5.2, for each (with ), it holds for all but values of , we have . By averaging, there exists some such that
which is at most if
(16) |
Putting everything together, we have
if Eqs. 15 and 16 are satisfied and if . Setting , , , ensures that these constraints are all satisfied for sufficiently large .
Applying Lemma 8.2 below with and completes the proof. ∎
Lemma 8.2 (Parallel Repetition Criterion).
Let be a game, and let denote . Suppose is a function with and is a constant such that for all with there exists such that . Then
Proof.
Fix any . Consider the probability space defined by sampling , and let . We define additional random variables and where is an arbitrary fixed value, for all , and depends deterministically on as follows. When , is defined to be a value that minimizes . With these definitions, each event is a product event. In particular, if then .
Let denote the event that , let denote the event , and let denote . Since is the union of some subset of the disjoint product events , we have
Moreover, for all for which , we know that . Thus as long as , we have
Iterating this inequality as long as the condition is satisfied, we find such that . This is minimized for or and gives . ∎
Appendix A Probability Theory
We recall the notions of probability theory that we will need.
Definition A.1.
A probability distribution on a finite set is a function satisfying for all and . We extend the domain of to by writing to denote for any “event” .
Definition A.2.
The support of is the set .
Definition A.3.
A -valued random variable on a sample space is a function .
Definition A.4 (Expectations).
If is a probability distribution and is a random variable, the expectation of under , denoted , is defined to be .
We refer to subsets of as events. We use standard shorthand for denoting events. For instance, if is a -valued random variable and , we write to denote the event .
Definition A.5 (Indicator Random Variables).
For any event , we write to denote a random variable defined as
Definition A.6 (Independence).
Events are said to be independent under a probability distribution if . Random variables are said to be independent if the events are independent for any choice of .
Definition A.7 (Conditional Probabilities).
If is a probability distribution and is an event with , then the conditional distribution of given is denoted and is defined to be
If is a random variable and is a probability distribution, we write to denote the induced distribution of under . That is, .
If is an event, we write as shorthand for .
Definition A.8 (Entropy).
If is a probability distribution, the entropy (in nats) of is
When is a random variable associated with a probability distribution , we sometimes write as shorthand for .
Definition A.9 (Conditional Entropy).
If is a probability measure with random variables and , we write
Fact A.10 (Chain Rule of Conditional Entropy).
For any probability measure and any random variables , , it holds that
A.1 Divergences
Definition A.11 (Total Variational Distance).
If are two probability distributions, then the total variational distance between and , denoted , is
An equivalent definition is
Definition A.12 (Kullback-Leibler (KL) Divergence).
If are probability distributions, the Kullback-Leibler divergence of from is
where terms of the form are treated as if and otherwise, and terms of the form are treated as .
The following relation between total variational distance and Kullback-Leiber divergence, known as Pinsker’s inequality, is of fundamental importance.
Theorem A.13 (Pinsker’s Inequality).
For any probability distributions , it holds that .
Definition A.14 (Conditional KL Divergence).
If are probability distributions and if , , , and are random variables on , we write
which is taken to be if there exists with but .
KL divergence obeys a chain rule analogous to that for entropy.
Fact A.15 (Chain Rule for KL Divergence).
If are probability distributions and are random variables on , then
A.2 Conditional KL Divergence
Fact A.16.
If is a probability distribution and is an event, then
Fact A.17.
Let be probability distributions and let , be random variables on with a function of . Then
Proof.
This is well known, but for completeness:
(chain rule) | ||||
( is a function of ) | ||||
A.3 Conditional Statistical Distance
Fact A.18.
Let be probability distributions, and let be an arbitrary event. Then
Proof.
Suppose for the sake of contradiction that for some , we have
Multiplying on both sides by , we obtain
Since and , we have
which is a contradiction. ∎
Corollary A.19.
Let be a probability distribution, let , and be random variables on , and let be an event such that , and let denote . Then
Proof.
Appendix B Fourier Analysis
For any (finite) vector space over , the character group of , denoted , is the set of group homomorphisms mapping (viewed as an additive group) to (viewed as a multiplicative group). Each such homomorphism is called a character of .
We will distinguish the spaces of functions mapping from and functions mapping and view them as two different inner product spaces. For functions mapping , we define the inner product
and for functions mapping , we define the inner product
If there is danger of ambiguity, we use to denote the latter inner product, and to denote its corresponding norm.
Fact B.1.
Given a choice of basis for , there is a canonical isomorphism between and . Specifically, if , then the characters of are the functions of the form
for .
Definition B.2.
For any function , its Fourier transform is the function defined by
One can verify that the characters of are orthonormal. Together with the assumption that is finite, we can deduce that is equal to .
Theorem B.3 (Plancherel).
For any ,
An important special case of Plancherel’s theorem is Parseval’s theorem:
Theorem B.4 (Parseval).
For any ,
Appendix C Bound on Optimization Problem
Let denote the inverse of the function ( is known in the literature as the (principal branch of the) Lambert W function). We rely on the following theorem:
Theorem C.1 ([HH00, Corollary 2.4]).
There exists a constant (in particular, works) such that for all ,
The following corollary is more directly suited to our needs.
Corollary C.2.
For any satisfying ,
Proof.
The minimum is achieved (up to a factor of two) when because is monotonically increasing with while is monotonically decreasing. Making the change of variables , this is equivalent to , i.e. . This choice of (or equivalently ) gives
(Definition of ) | ||||
(Theorem C.1) | ||||
References
- [BGKW88] Michael Ben-Or, Shafi Goldwasser, Joe Kilian, and Avi Wigderson. Multi-prover interactive proofs: How to remove intractability assumptions. In STOC, pages 113–131. ACM, 1988.
- [BJKS04] Ziv Bar-Yossef, T. S. Jayram, Ravi Kumar, and D. Sivakumar. An information statistics approach to data stream and communication complexity. J. Comput. Syst. Sci., 68(4):702–732, 2004.
- [CHTW04] Richard Cleve, Peter Høyer, Benjamin Toner, and John Watrous. Consequences and limits of nonlocal strategies. In CCC, pages 236–249. IEEE Computer Society, 2004.
- [DHVY17] Irit Dinur, Prahladh Harsha, Rakesh Venkat, and Henry Yuen. Multiplayer parallel repetition for expanding games. In ITCS, volume 67 of LIPIcs, pages 37:1–37:16. Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 2017.
- [EPR35] Albert Einstein, Boris Podolsky, and Nathan Rosen. Can quantum-mechanical description of physical reality be considered complete? Physical review letters, 47(10):777, 1935.
- [Fei91] Uriel Feige. On the success probability of the two provers in one-round proof systems. In Structure in Complexity Theory Conference, pages 116–123. IEEE Computer Society, 1991.
- [FGL+91] Uriel Feige, Shafi Goldwasser, László Lovász, Shmuel Safra, and Mario Szegedy. Approximating clique is almost NP-complete (preliminary version). In FOCS, pages 2–12. IEEE Computer Society, 1991.
- [FK91] H. Furstenberg and Y. Katznelson. A density version of the Hales-Jewett theorem. Journal d’Analyse Mathématique, 57(1):64–119, December 1991.
- [For89] Lance Jeremy Fortnow. Complexity-theoretic aspects of interactive proof systems. PhD thesis, MIT, 1989.
- [FRS94] Lance Fortnow, John Rompel, and Michael Sipser. On the power of multi-prover interactive protocols. Theor. Comput. Sci., 134(2):545–557, 1994.
- [FV96] Uriel Feige and Oleg Verbitsky. Error reduction by parallel repetition - a negative result. In Steven Homer and Jin-Yi Cai, editors, CCC, pages 70–76. IEEE Computer Society, 1996.
- [GHZ89] Daniel M. Greenberger, Michael A. Horne, and Anton Zeilinger. Going Beyond Bell’s Theorem, pages 69–72. Springer Netherlands, Dordrecht, 1989.
- [HH00] Abdolhossein Hoorfar and Mehdi Hassani. Inequalities on the Lambert W function and hyperpower function. J. Inequal. Pure and Appl. Math, 2000.
- [Hol09] Thomas Holenstein. Parallel repetition: Simplification and the no-signaling case. Theory Comput., 5(1):141–172, 2009.
- [HY19] Justin Holmgren and Lisa Yang. The parallel repetition of non-signaling games: counterexamples and dichotomy. In STOC, pages 185–192. ACM, 2019.
- [MS13] Carl A. Miller and Yaoyun Shi. Optimal robust self-testing by binary nonlocal XOR games. In TQC, volume 22 of LIPIcs, pages 254–262. Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 2013.
- [Pol12] D.H.J. Polymath. A new proof of the density Hales-Jewett theorem. Annals of Mathematics, 175(3):1283–1327, May 2012.
- [PRW97] Itzhak Parnafes, Ran Raz, and Avi Wigderson. Direct product results and the GCD problem, in old and new communication models. In Frank Thomson Leighton and Peter W. Shor, editors, STOC, pages 363–372. ACM, 1997.
- [Raz98] Ran Raz. A parallel repetition theorem. SIAM J. Comput., 27(3):763–803, 1998.
- [Raz11] Ran Raz. A counterexample to strong parallel repetition. SIAM J. Comput., 40(3):771–777, 2011.
- [Ver96] Oleg Verbitsky. Towards the parallel repetition conjecture. Theor. Comput. Sci., 157(2):277–282, 1996.
- [Yue16] Henry Yuen. A parallel repetition theorem for all entangled games. In ICALP, volume 55 of LIPIcs, pages 77:1–77:13. Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 2016.