The Aldous–Lyons Conjecture I: Subgroup Tests
Abstract.
This paper, and its companion [Tailored_MIPRE], are devoted to a negative resolution of the Aldous–Lyons Conjecture [Aldous_Lyons_Conj, Aldous--Lyons_conj_blogpost]. This conjecture, originated in probability theory, is well known (cf. [Gelander_ICM2018]) to be equivalent to the statement that every invariant random subgroup of the free group is co-sofic. We disprove this last statement.
In this part we introduce subgroup tests. These tests are finite distributions over continuous functions from the space of subgroups of the free group to . Subgroup tests provide a general framework in which one can study invariant random subgroups of the free group. Classical notions such as group soficity and group stability arise naturally in this framework. By the correspondence between subgroups of the free group and Schreier graphs, one can view subgroup tests as a property testing model for certain edge-labeled graphs. This correspondence also provides the connection to random networks.
Subgroup tests have values, which are their asymptotic optimal expectations when integrated against co-sofic invariant random subgroups. Our first main result is that, if every invariant random subgroup of the free group is co-sofic, then one can approximate the value of a subgroup test up to any positive additive constant.
Our second main result is an essentially value preserving correspondence between certain non-local games and subgroup tests. By composing this correspondence with a stronger variant of the reduction in [MIPRE], proved in the companion paper [Tailored_MIPRE], we deduce that approximating the sofic value of a subgroup test is as hard as the Halting Problem, and in particular, undecidable. The combination of our two main results proves the existence of non co-sofic invariant random subgroups of the free group.
Contents
1. Introduction
In their seminal paper [Aldous_Lyons_Conj], Aldous and Lyons ask whether every unimodular network is sofic. Conditional on a certain result which will be addressed in a companion paper [Tailored_MIPRE], we prove the answer is ‘no’. This conjecture, which originated in probability theory, can also be expressed in the language of invariant random subgroups, and this is the one we will use in this paper. We explain and motivate this conjecture before discussing our approach. The formal definitions and results will start in Section 1.1.
The Aldous–Lyons Conjecture
What is soficity?
Roughly speaking, a countable group is sofic if it admits a sequence of partial actions on finite sets which approximates the action of on itself by left-translations. If is finitely presented then this is equivalent to the existence of a sequence of finite graphs which converge to the Cayley graph of in the sense of Benjamini–Schramm, which is a kind of local-on-average convergence.
To be more precise, a rooted graph is a pair where is a graph and is a distinguished vertex called the root. A random rooted graph is sofic if there exists a sequence of finite graphs such that, if is a uniformly random vertex of , then converges to in distribution (this means: for every , the radius neighborhood of the root in converges in distribution to the radius neighborhood of the root in ). This notion of convergence is due to Benjamini and Schramm [MR1873300]: They proved that if each finite graph is planar and there is a uniform degree bound, then the limit is almost surely recurrent.
The notions of Benjamini–Schramm convergence and soficity naturally generalize from graphs to labeled graphs or networks [Aldous_Lyons_Conj], simplicial complexes [MR2797963], manifolds [MR4520306], measured equivalence relations [MR2566316], measured groupoids [dykema-2014] and most generally, measured metric spaces [khezeli2023unimodular].
Again, a finitely presented group is sofic if one of its Cayley graphs is sofic.111If one of the Cayley graphs is sofic, then all those that use finitely many generators are. For example, amenable groups are sofic. In fact, the Cayley graph of an amenable group admits a sequence of finite sub-graphs in which the isoperimetric ratio (number of boundary vertices to number of vertices) tends to zero, which implies this sequence Benjamini–Schramm converges to the Cayley graph. Also, residually finite groups are sofic because they admit sequences of finite quotient groups whose Cayley graphs approximate the Cayley graph of the given group.
Gromov implicitly introduced sofic groups in [MR1694588]. He was motivated by Gottshalk’s Conjecture: If is a countable group and , then any -equivariant continuous map which is injective is necessarily surjective. This is obviously true when is finite. Gromov proved that it holds when is sofic. Benjy Weiss gave another proof and coined the term ‘sofic’ from the Hebrew word for finite [weiss-2000].
The soficity property has been useful in obtaining positive results about and spectral theory invariants [MR3664810], group rings [MR2089244, MR2417890], invariant couplings of random fields [MR3503036], and in constructing dynamical invariants of groups actions [MR2552252, MR2854085, MR3077882, MR3132735, MR3993930]. See [Cap_Lup_Sofic_Hyperlinear_book, MR3821628] for more background on sofic groups. The basic idea is roughly the same in each case: One transfers properties of the finite approximating objects to properties of the limit.
It is a major open problem to determine whether all countable groups are sofic. We do not directly address this problem because the class of objects we study, described next, is more general than the class of groups.
What is unimodularity?
The main tool for analyzing Benjamini–Schramm limits is called the Mass Transport Principle which makes precise the intuitive notion of statistical homogeneity. To explain, a doubly-rooted graph is an ordered triple , where are vertices of . A transport function is a function , satisfying some measurability condition, which takes as input a doubly rooted graph and outputs a non-negative real number. The interpretation is that is the amount of mass the first root sends to the second root . So if is a random rooted graph then is the average amount of mass sent out of the root. Symmetrically, is the average amount of mass sent into the root. If these two quantities are equal for every transport function , then we say satisfies the Mass Transport Principle. Equivalently, we say it is unimodular (this is the term used by Aldous and Lyons in [Aldous_Lyons_Conj]). Alternatively, unimodularity can be formulated in terms of graphings or as invariance with respect to the root-changing equivalence relation on the space of rooted graphs. The former arises from the ergodic theory of measured equivalence relations (e.g. [MR1164598]). The latter point-of-view was introduced in [MR1631732] and further developed in [MR3504507].
The term unimodular comes from the following special case. Suppose is a connected transitive graph; this means that its automorphism group acts transitively on its vertex set. Then is unimodular if and only if the automorphism group is unimodular in the sense that its left and right Haar measures agree.
The Mass Transport Principle arose in Häggström’s study of percolation clusters on trees [MR1457624]. In fact, it is an exercise to show that if is unimodular and is given, then the Bernoulli -percolation cluster containing is also unimodular. Unimodularity was further developed in [blps-group-perc] (see also [MR2883390] where it was used to construct a version of the Euler characteristic for unimodular planar maps).
It is straightforward to verify that (1) unimodularity is closed under weak limits and (2) if is a finite graph and is a uniformly random vertex of then is unimodular. It follows from these observations that soficity implies unimodularity. Question 10.1 of [Aldous_Lyons_Conj] asks whether unimodularity implies soficity.
As above, unimodularity has been generalized to random rooted labeled graphs, simplicial complexes, manifolds and so on. Benjamini and Schramm note that many results which are known to hold in the deterministic setting of unimodular transitive graphs can be generalized to unimodular random graphs [MR1873300]. Aldous and Lyons demonstrate this with results about random walks, amenability, ends of graphs, spanning forests, percolation, and so on [Aldous_Lyons_Conj].
What is an invariant random subgroup?
As a first step towards our goal, let us describe the algebraic formulation of the Aldous–Lyons Conjecture in terms of invariant random subgroups, which were introduced in [MR2749291], [abert2014kesten], [MR3193754] (and implicitly in [stuck1994stabilizers]). The concept of invariant random subgroups has been profitably studied in the context of locally compact groups (e.g. [stuck1994stabilizers, MR3664810, Gelander_ICM2018]). However, we will restrict our focus to the case of most relevance to this paper: finitely generated groups. Already in this discrete setup, invariant random subgroups play an important role in graph convergence [Hatami_Lovasz_Szegedi_Graph_limits], stability properties of groups [BLT] and the analysis of dynamical systems [stuck1994stabilizers, MR3193754] for example.
So, fix a finitely generated group . An invariant random subgroup (or IRS) of is a random variable taking values in the space of all subgroups of , whose law is invariant under the conjugation action of . For example, every normal subgroup of is an IRS. More generally, if has only finitely many conjugates, then a uniformly random sample of its conjugates is an IRS. As finite index subgroups have only finitely many conjugates, they induce IRSs which are called elementary. To generate more examples, one can take convex combinations and weak* limits of elementary IRSs, and the IRSs that arise this way are called co-sofic. It turns out that for some groups every IRS is co-sofic, and for others there are non co-sofic IRSs (cf. [BLT]). The Aldous–Lyons Conjecture can be formulated as follows:
Are all IRSs of a (non-commutative) free group co-sofic?
For the rest of the paper, this is the formulation that we tackle, and it is known — and we explain it in the following paragraphs — to be equivalent to the probability theoretic formulation on Benjamini–Schramm limits of finite graphs.
The law (or distribution) of an IRS is a Borel probability measure on , the space of subgroups of . We let denote the space of all such measures. By abuse of language, we say that is an IRS if . To see the relation between the two formulations of the Aldous–Lyons Conjecture, note that if is a subgroup of , then the Schreier coset graph of with respect to some generating set is a rooted (directed, edge-labeled) graph, with the root being the coset of the identity. This gives a map from the space of subgroups to the space of isomorphism classes of rooted (edge-labeled, directed) graphs.
If then pushes forward under this map to a unimodular measure [abert2014kesten, Proposition 14]. Moreover, if is elementary, then its pushforward is induced by uniformly choosing a root of some finite connected graph. Therefore, if is co-sofic, then its pushforward is sofic.
Now, consider the special case in which is the free group generated by a finite set . Suppose is an invariant random subgroup of , and that the (random) Schreier coset graph of is sofic (in the directed, edge-labeled rooted graph category). This means there exists a sequence of finite directed, edge-labeled graphs which Benjamini–Schramm converge to the (random) Schreier coset graph of . After perturbing the given sequence if necessary, we may assume that each is itself a Schreier coset graph with respect to some finite-index subgroup of — the existence of such perturbations is exactly where the freeness of the group plays a role. Moreover, if is a uniformly random conjugate of , then the distribution of converges to . Therefore, a positive solution to the Aldous–Lyons Conjecture implies that every IRS of is co-sofic.222There is a standard “decoration of graphs” technique that allows one to show that the directed, edge labeled (with finitely many labels) version of the Aldous–Lyons conjecture is equivalent to the non-directed, non-edge labeled version. As Bálint Virág pointed to us, the same technique is used, e.g., to show that every group is the automorphism group of some graph. As this is standard, we do not elaborate on it anymore.
Our approach
What is a subgroup test?
The set of all IRSs of the free group and the set of co-sofic IRSs of the free group are both convex. Thus, the Aldous–Lyons Conjecture is asking whether certain convex sets can be separated. A natural way to distinguish between two convex sets is to use (continuous) linear functionals. So, in order to separate the above, one should search for a rich enough collection of functionals. Subgroup tests and their values will play this exact role.
A challenge is a continuous map from to . A subgroup test is a probability distribution on a finite set of challenges. One can integrate any subgroup test against any probability measure over . If is in , then this integral is called the value of the strategy against the test . Many natural and well studied properties of groups, specifically soficity and pointwise permutation stability, can be formulated as properties of subgroup tests.
The terminology of “tests” and “values” originates in the field of interactive proofs in computer science. An interactive proof designates an interaction between two entities, a “verifier” and a “prover”. The goal of the prover is to convince the verifier of the validity of a certain claim, and the goal of the verifier is to test the prover so that it accepts the interaction only if the claim is indeed correct. Examples of “claims” studied in computer science are the -colorability of a given input graph, or that a Turing machine whose description is passed as input halts. In our approach, the methods of the verifier and the prover are restricted. The prover is required to sample an element of . The sampling method itself is restricted to some sub-class of IRSs (e.g., elementary IRSs); the choice of distribution according which to sample is the only degree of freedom the prover has, and is called its strategy. The verifier gets access to (the indicator function of) the prover’s sampled subgroup , and it makes a decision whether to accept or reject this subgroup. To decide, the verifier holds some finite set — known beforehand to the prover — and according to which combinations of elements of are contained in , it accepts or rejects; the rules that control which combinations of elements result in acceptance and which combinations result in rejection are also know to the prover beforehand. The value of a strategy in a game is the probability, taken over the prover’s sampling and the verifier’s probabilistic choices, that the verifier makes the decision to accept when interacting with a prover using this strategy.
In spite of these restrictions, it is possible to verify hard computational problems in this setup.333For the expert, we note that restrictions on the prover and verifier can affect both the completeness and soundness properties of a proof system. Hence, a priori, they may either lower or raise the complexity of the associated class of interactive proofs compared to classical single-prover interactive proofs. For example, it is not difficult to design a subgroup test such that there is a value- strategy for this test if and only if a given graph is -colorable. Jumping ahead, the correct analogy is with the theory of multiprover interactive proofs, and even more specifically interactive proofs with quantum provers sharing entanglement, also known as non-local games. We explain below this connection, which plays an essential role in the second part of this paper.
What is a non-local game?
Morally, non-local games originated in Bell’s resolution [bell1964einstein] of the Einstein–Podolsky–Rosen paradox [einstein1935can]. Einstein was famously concerned by the ability of space-time separated particles to correlate in ways that might violate relativity, a possibility that seemed to be suggested by the mathematical modeling of quantum mechanics. Einstein argued for a local hidden variable theory — namely, that the particles did share some information beforehand, which allowed them to correlate — and that physicists should search for this hidden information. Bell provided a thought experiment, that in modern jargon is an instance of a non-local game, which proves that the kinds of correlations that arise from isolated quantum mechanical systems are intrinsically different from those that could be generated in any hidden variable model. Bell’s work lay the foundation for the subsequent design, and execution, of experiments which verify the existence of quantum entanglement (and thus refute Einstein’s approach; see also [clauser1969proposed]). The 2022 Nobel prize in physics was awarded to Aspect, Clauser and Zeilinger, partly for performing non-local games as experiments and verifying that the winning statistics in the games exceeds what local hidden variable models allow.
A correlation is a function , where and are finite sets, such that for every pair the function is a probability distribution over — the quantity should be thought of as the answer to “what is the probability and are provided as answers given that and were asked as questions?”; this cryptic phrase will soon be clarified. A correlation is said to be deterministic if there are two functions such that only if and . The convex hull of deterministic correlations is the set of local or classical correlations.444An example of a non-local correlation is only when and . Intuitively, this is because this correlation implies signaling from one system to the other. On the other hand, the quantum correlations are those that can be generated by performing quantum measurements on a (finite-dimensional) bipartite physical system, where the measurement on one part depends only on and generates , while the measurement on the other part depends only on and generates . For a formal description, see Remark 6.8. The set is convex, but interestingly it is not closed [slofstra2019set].
A non-local game is specified by a probability distribution on the set , and a decision function .555As can be seen in Section 1.5, our definition of non-local games is slightly different. This difference is mainly cosmetic, and is driven by our motivation to relate non-local games to subgroup tests. The interpretation of this combinatorial object as a “game” comes from thinking of it (similarly to subgroup tests) as an interactive proof, but this time with two provers, a la [ben1988multi]. The common dramatization goes as follows: Two provers are spatially separated. The verifier samples a pair of “questions” , and “asks” one prover and the other . The provers then apply some local procedure to choose their “answers” and respectively. Finally, the verifier accepts if and rejects otherwise.
Since the process by which the provers generate their answers is hidden to us, we can only “observe” (samples from) the distribution of answers given questions, namely, the underlying correlation. So, the correlation can be thought of as the “strategy” used by the provers (similar to the way IRSs are seen as strategies for the prover in a subgroup test). Bell’s separation of from amounts to devising a game that is “easy” for players that use quantum correlations, yet “hard” for players that use only classical ones.
Now, there is more than one candidate mathematical model for entanglement in quantum mechanics. A slightly generalized model, suggested by Tsirelson [tsirelson1993some], gives birth to a set of correlations known as quantum commuting correlations . This set is closed and contains the quantum correlations. Tsirelson famously asked whether the closure of is equal to [tsirelson2006bell]. At this point, we hope the type of problem already resonates with the reader, as this is again a separation of convex sets type of problem.666Connes’ embedding problem (CEP) is also a separation of convex sets type of problem, specifically, separating all characters of the free group from those that are limits of finite dimensional characters. This problem was shown [fritz2012tsirelson, junge2011connes, ozawa2013connes] to be equivalent to Tsirelson’s problem, and also has close connections with invariant random subgroups — as every IRS induces a character. We omit discussions on the relations between this work and CEP from the text, as they are not particularly helpful for understanding our approach.
Harnessing undecidability
It was known that if the quantum correlations are dense in the quantum commuting correlations, then the complexity class of multiprover interactive proofs with entangled provers () contains only decidable languages. By proving that the Halting Problem, which is undecidable, is in , the authors of [MIPRE] were able to resolve Tsirelson’s problem in the negative.
In this paper a similar path is followed: Rather than directly find a functional that separates the co-sofic and general IRSs of the free group, we describe a computational problem — approximating the sofic value of a subgroup test — whose undecidability implies a negative resolution of the Aldous–Lyons conjecture. Our approach can be described in two steps: First, show that a positive solution to the Aldous–Lyons Conjecture implies that approximating the sofic value of a subgroup test is decidable. Second, prove that this computational task is undecidable.
Elaborating on the first, given a subgroup test , we prove that its optimal value against co-sofic IRSs — which we call the sofic value of — can be approximated from below, and its optimal value against any IRS — which we call the ergodic value of — can be approximated from above. Thus, if the Aldous–Lyons Conjecture has a positive answer, then the sofic and ergodic value of a test always agree, and using the aforementioned approximation procedures they can be calculated to any predetermined accuracy.
Elaborating on the second, we provide a mechanism for translating certain non-local games (see Section 1.5), which we call tailored non-local games, to subgroup tests. This translation is essentially value preserving. Namely, if the non-local game had a (certain kind) of perfect quantum strategy — which we call a -aligned permutation strategy that commutes along edges — then the analogous subgroup test has a perfect (co-sofic) strategy. In the other direction, if the subgroup test has an almost perfect (co-sofic) strategy, then the non-local game has an almost perfect strategy. Hence, a certain strengthening of the reduction in [MIPRE], which states that small additive constant approximations to the value of tailored non-local game are undecidable (Theorem 7.4), implies that the sofic value of a game cannot be approximated to arbitrary precision. Thus, the Aldous–Lyons Conjecture has a negative solution.
The rest of the introduction provides a deeper dive into the definitions, results and ideas of this paper.
1.1. Subgroup Tests
Let be a finite set, be the free group with basis , the collection of subgroups of and the set of all Borel probability measures on . A challenge is a pair such that is a finite subset of , and is a function from subsets of to . A subset passes the challenge if , and fails it otherwise. A test is a finite collection of challenges , where is a finite index set, together with a probability distribution over the set . A strategy is a (Borel) probability distribution over subgroups of the free group, namely an element . One can run the test against , as follows. First the verifier samples according to . Next, the prover samples a subgroup according to . Finally, the verifier makes the decision to accept if passes the challenge , and reject if fails the challenge.
The value of using the strategy against the test is its acceptance probability, namely
For a fixed , the value is a continuous linear functional from probability distributions over subgroups of to . Hence, it enables us to study convex subsets of by their optimal value against a given test.
Remark 1.1.
As mentioned earlier, a challenge is, essentially, a continuous map from the space of subgroups of to . One can extract from any such continuous map a challenge that describes it and vice versa. Therefore, a test is a convex combination of continuous maps from to , and the value is integration of this map against a (probability) measure.
1.2. Invariant random subgroups
As inherits the product topology from the power set , it is compact. Every action of a group on a set extends naturally to an action of on the power set by
Hence, the conjugation action of on itself is inherited by . Also, is preserved by this action. Hence, the conjugation action extends to its power set by
The conjugation action of further extends to : For every Borel set and , we have
Definition 1.2.
The set of invariant random subgroups consists of all probability measures in that are invariant under the action of .
The ergodic value of a test is the best performance of a strategy against it, namely
(1.1) |
Remark 1.3.
The extremal points in the convex set are commonly referred to as ergodic IRSs. As we take the supremum of a linear functional over a convex compact set, the optimum is obtained on an extremal point, which explains the phrase ergodic value.
1.3. Finitely described invariant random subgroups
Let be a finite set, and let be the symmetric group acting on . Every map extends uniquely to an action of on . For a vertex , we can define its stabilizer to be
(1.2) |
which is a subgroup of . Hence, we can associate with an IRS of via the following sampling procedure: Choose uniformly at random. Output . We denote by the map from finite actions of to defined by this procedure, namely
(1.3) |
where is the Dirac measure concentrated on the subgroup and is the expectation according to the uniform measure over . The set of finitely described invariant random subgroups is the image of the map in .
Definition 1.4.
The sofic value of a test is the (asymptotic) best performance of a strategy against it, namely
(1.4) |
Remark 1.5.
The name sofic value is appropriate, as demonstrated in Section 4.2. Furthermore, we often abuse notation and use instead of , even though the action itself is not an IRS.
Since , we have for every test . The Aldous–Lyons conjecture [Aldous_Lyons_Conj] can be formulated as follows:
Conjecture 1.6 (Aldous–Lyons, cf. Section 6 of [Gelander_ICM2018]).
Is weak* dense in ?
Remark 1.7.
In their paper [Aldous_Lyons_Conj], Aldous and Lyons ask this as a question. In the passing of time, the positive form of this problem, namely that is dense in , received the name The Aldous–Lyons Conjecture (cf. [Aldous--Lyons_conj_blogpost]). Thus, we use the more common “Conjecture” phrasing instead of Problem.
The algebraic formulation we provided earlier, which can also be found in Section 6 of [Gelander_ICM2018], may seem different from the above. The question was whether every invariant random subgroup of is co-sofic, where co-sofic means an IRS which is the weak* limit of convex combinations of uniform distributions over finite index subgroups. Since the finitely described IRSs are dense in the co-sofic ones, the formulation in Conjecture 1.6 is equivalent to this standard (algebraic) formulation.
Lastly, the term co-sofic is closely related to the better known term of a sofic group (cf. Definition 4.4). A finitely presented group is sofic if and only if the Dirac measure concentrated on is a co-sofic IRS (cf. Proposition 6.1 in [Gelander_ICM2018]). Thus, a positive solution to the Aldous–Lyons Conjecture 1.6 implies in particular that every group is sofic.777The way we formulate it, a positive solution to the Aldous–Lyons Conjecture only implies that all finitely presented groups are sofic. It is standard to deduce it for all groups from the finitely presented ones [MR2089244].
Corollary 1.8.
Since is a continuous functional, a positive solution to the Aldous–Lyons Conjecture 1.6 implies that for every test , .
Remark 1.9.
Due to the relation between subgroups of the free group and Schreier graphs, one can view subgroup tests as a new property testing model for (edge-labeled) graphs: Let be a finite set and a transitive action (the non-transitive case is a convex combination of transitive ones). For , the stabilizer is exactly the set of labeled paths in the Schreier graph of with respect to and that begin and end at , namely, closed paths originating from . Thus, when the finitely described IRS is put against a test , we actually test the aforementioned Schreier graph in the following way: First, choose a vertex uniformly at random and sample . List the paths from and for each one check whether it is closed or open. According to this check, decide using whether to accept or reject.
1.4. Decidability of approximating the sofic value
If the challenge distribution in a subgroup test is rational, then can be encoded as a finite bit string. There are many (non-equivalent) ways of doing so. Assume we fixed some encoding for tests with rational challenge distributions such that, given the encoding, all underlying combinatorial data of the test can be calculated in finite time. Namely, the set of generators , the indexing set of challenges , the collection of challenges and the distribution can all be read from the encoding in finite time.
Theorem 1.10 (Main Theorem I).
For every fixed encoding of tests such that all combinatorial data of the test can be read from it in finite time:
-
(1)
There is a Turing machine that takes as an input (the encoding of) a test , and outputs an infinite sequence of non-decreasing numbers such that .
-
(2)
There is a Turing machine that takes as an input (the encoding of) a test , and outputs an infinite sequence of non-increasing numbers such that .
Remark 1.11.
Clause of Theorem 1.10 should be compared to the Navascués–Pironio–Acín (NPA) hierarchy in the study of non-local games [navascues2008convergent]. See also the introduction of [MIPRE].
Corollary 1.12.
If the Aldous–Lyons Conjecture 1.6 has a positive solution, then for every , there is a Turing machine which accepts a subgroup test as input and outputs its sofic value up to an additive error of at most .
1.5. Tailored non-local games and their associated subgroup tests
As described before, the standard definition of a (non-local) game consists of two finite sets, a questions set and an answers set , together with a probability distribution over and a decision function . In this paper, we use a slightly modified definition, that makes the connection with subgroup tests clearer. Before diving into our alternative definition, note first that the distribution defines a graph structure on , by associating edges with its support. Furthermore, as the exact choice of is irrelevant, we can assume it is the set of all bit strings up to some length .
A (synchronous, non-local) game consists of a graph , a distribution over , a length function , formal sets of generators — the union of which is denoted by — and decision functions for every edge , where .888So, the main difference in our definition, is that the decision function rejects automatically answers that are not of the appropriate length according to . A (synchronous, quantum) strategy for is a map , where the images are involutions, and commute for every fixed .999This is again not the standard definition of a quantum strategy. Unpacking our definition, associates an observable (Hermitian operator on a finite-dimensional Hilbert space that squares to the identity) with each bit of the answer when asked for . Since the answer must be totally measured, all the observables associated with a specific vertex must commute. The fact there is a single map , and not one for each prover, is the synchronicity of this strategy. Such a strategy defines for every a probability distribution over maps in the following way: For every , let be the projection on the -eigenspace of , and the projection on its -eigenspace. Furthermore, let be the dimension normalized trace on matrices. Then, the probability that is sampled according to is by definition
(1.5) |
A map sampled that way is said to be sampled according to , and we denote it by .101010Note that this sampling depends on the chosen edge . Namely, though implicit in the notation, is defined from for some edge . It is important to note that is not globally defined on all of . Relating to standard notations, we remark that by letting and , the correlation induced by the quantum strategy (as in Remark 6.8) agrees with formula (1.5). In a similar manner to subgroup tests, one can run the game against the strategy : First, the verifier samples an edge according to . Then, the prover samples according to . Finally, the verifier accepts if , and rejects otherwise.
As for subgroup tests, the value of the strategy against the non-local game is its acceptance probability when ran against the game. Namely,
The (synchronous) quantum value of , , is the supremum of over all possible quantum strategies against .
The main result in [MIPRE] (recalled in detail in Theorem 6.10 herein) is a reduction from the Halting Problem to approximating the quantum value of a game. Our main contribution in the second half of this paper (Sections 6,7 and 8) is a mapping from a specific sub-class of non-local games — which we term tailored non-local games and describe further below — to subgroup tests, that (almost) preserves their values.
Theorem 1.13 (Main Theorem II, informal. For the formal version see Theorem 7.3).
There is a transformation that takes as input a tailored non-local game (see Definition 6.22) and outputs a subgroup test (which we call the associated synchronous subgroup test, see Definition 7.1), such that:
- (1)
-
(2)
Every almost perfect finitely described strategy for the associated subgroup test can be transformed into an almost perfect quantum strategy to the original tailored non-local game.
Informally, a tailored game is one where the decision function takes the following form. First, examines the value taken by on a subset of the generators (we refer to these as readable variables). Based on this, determines a collection of linear equations on the entire string . Finally, accepts if and only if all these equations are satisfied. This form for the game can be seen as a generalization of binary linear constraint system games (Example 6.26), which corresponds to the case where the set of readable variables is empty.
A permutation strategy, as its name suggests, is a quantum strategy which is induced by permutations. Such a strategy is said -aligned when the observables associated with the readable variables (i.e. the bits of the answer that control the parity checks in the tailored game) are all diagonal matrices (with respect to some natural basis). The restriction to -aligned strategies is crucial for item (1) in Theorem 7.3 above to hold.
In a companion paper [Tailored_MIPRE], we prove that the reduction of the Halting Problem to approximating the quantum value of a game presented in [MIPRE] can be modified such that the resulting non-local game is tailored and satisfies the required stronger assumptions (Theorem 7.4): If the input Turing machine halts, then the game has a perfect -aligned permutation strategy that commutes along edges, and if the input Turing machine does not halt, then the value of the game is bounded from above by . Therefore, together with our mapping from non-local games to subgroup tests, we deduce that approximating the sofic value of a game is undecidable. In turn, by Corollary 1.12, the Aldous–Lyons Conjecture 1.6 has a negative solution.
Remark 1.14.
In [MIPRE, Section 12.3], the authors were able to specify a non-local game whose quantum value is bounded above by but has a perfect quantum commuting strategy. Using the results of [kim2018synchronous], which we elaborate on further in the next remark, this implies the existence of a finitely presented -algebra which can be endowed with a tracial state and cannot be embedded into an ultrapower of the hyperfinite factor. This is a somewhat constructive counter example, but the state itself is only proven to exist and does not have a constructive form (at this point). By the various formulations of Connes’ embedding problem, this proves that there is a non-hyperlinear character of the free group, which is the analogue of a non co-sofic IRS of the free group, but it is provided in a non-constructive form.
By applying the “same” construction as in [MIPRE, Section 12.3] with the algorithm constructed in Theorem 1.10 replacing the NPA hierarchy algorithm, one can pin down an explicit subgroup test with perfect ergodic strategy but all sofic strategies bounded above by some constant (which can be approximated by following the proofs). This is the most constructive we can get in this case, namely, the specific non co-sofic IRS which is the perfect strategy for this subgroup test is proven only to exist, and is not constructed. Thus, we decided not to specify this test.
It would be very interesting to find an explicit non co-sofic IRS, namely separate between and in a constructive manner.
Remark 1.15.
We comment on a natural attempt to extend the construction of [MIPRE] so as to directly deduce the existence of a non-sofic group, a result which would imply a negative answer to the Aldous–Lyons Conjecture.
The idea (which is folklore, and explicitly mentioned in [paddock2023satisfiability]) is as follows. It is well-known that to every synchronous game one can associate an algebra such that there is a tight connection between the -representation theory of this algebra and perfect values for the game in various correlation sets [paulsen2016estimating, kim2018synchronous, helton2019algebras]. In the special case where is a linear constraint system (LCS) game, there is a group such that is (isomorphic to the completion of) the complex group algebra [kim2018synchronous, goldberg2021synchronous].111111Actually, it is isomorphic to the completion of , where is a special generator in and is the two-sided ideal generated by in .
Now, if the reduction given in [MIPRE] returned only LCS games, then one would obtain the following result: there exists an LCS game with a perfect quantum commuting strategy, but no (asymptotically) perfect quantum strategy. By the connection mentioned above this implies that the group associated with is non-hyperlinear. As every sofic group is hyperlinear, one would thus obtain an example of a non-sofic group.
Since the reduction in [MIPRE] does not return LCS games, a workaround could be to find an embedding from the algebra , where is returned by [MIPRE], into where is an LCS game. Obstacles to this “black-box” approach are discussed in [paddock2023satisfiability].
One can view our approach as finding a “sweet spot” that is in some sense half-way between synchronous games, which were used in [MIPRE], and LCS games. This is because tailored games are a natural generalization of LCS games. In the companion paper we show that [MIPRE] can be adapted to return tailored games, and in the present paper we show that this suffices to resolve the Aldous–Lyons Conjecture. The existence of non-sofic groups (and non-hyperlinear groups) remains an elusive open problem.
1.6. Structure of the paper
In Section 2, we prove structural results about invariant random subgroups, some of which are standard and well known to experts. In Section 3, we prove our first main theorem, Theorem 1.10, and draw corollaries from it. In Section 4, we provide several examples of subgroup tests, inspired by both group theory and complexity theory. In Section 5, we discuss a rigidity property of tests that generalizes group stability and which is analogous to robustness of non-local games. In Section 6, we define tailored non-local games and analyze their properties. In Section 7, we describe the (almost) value preserving correspondence between tailored non-local games and synchronous subgroup tests, informally outlined in Theorem 1.13. This allows us in Corollary 7.5 to provide a negative solution to the Aldous–Lyons Conjecture 1.6. The proof of the aforementioned value preserving correspondence, Theorem 7.3, is proved in Section 8. Theorem 7.4, which is the other main ingredient in our refutation, is proved in a companion paper [Tailored_MIPRE]. There is no open problems section, but further research directions can be found in Remarks 1.14, 1.15 and 8.13.
Acknowledgements
We thank Alon Dogon for reading an early draft of this paper and providing many useful improvements to the presentation. We would also like to thank Mikael de la Salle for various discussions along the way. Finally, we want to thank Peter Sarnak for organizing a seminar on “strong convergence” in Princeton and asking the second author to give a talk, as it organized his thoughts and led to the current presentation of ideas.
Lewis Bowen is supported by NSF grant DMS-2154680. Michael Chapman acknowledges with gratitude the Simons Society of Fellows and is supported by a grant from the Simons Foundation (N. 965535). Alex Lubotzky is supported by the European Research Council (ERC) under the European Union’s Horizon 2020 (N. 882751), and by a research grant from the Center for New Scientists at the Weizmann Institute of Science. Thomas Vidick is supported by a research grant from the Center for New Scientists at the Weizmann Institute of Science, a Simons Investigator award, AFOSR Grant No. FA9550-22-1-0391, and ERC Consolidator Grant VerNisQDevS (101086733).
2. Group theoretic preliminaries
2.1. Topology (and other properties) of invariant random subgroups
121212The content of this section is standard. For example, it appears in Section 3 of [abert2014kesten].We first go through some properties of the space (induced with the product topology). For every , let
Then,
-
(1)
When and are finite, the set is open and closed. Moreover, the collection
is a basis for the topology of . In addition, is compact.
-
(2)
For every and subsets we have
The following is a standard fact. A proof is added because elements of it are later used in Lemma 2.12.
Claim 2.1.
The set of subgroups is closed in the collection of all subsets .
Proof.
A subset is a subgroup if it satisfies the following three conditions:
For every , define the following open sets: The subsets not containing , . The subsets containing but not , . The subsets containing but not , . Then,
which proves that is a closed set. ∎
Recall the definition of , the invariant random subgroups of the free group (Definition 1.2).
Fact 2.2.
Let . Then,
-
(1)
if and only if for every finite .
-
(2)
is conjugation invariant if and only if for every we have . Hence, if and only if for all and finite we have .
A sequence of measures is said to weak* converge to if for all continuous ,
We denote weak convergence by .
Remark 2.3.
In our context, there is a more “down to earth” way of viewing weak* convergence. For every finite subset , and every probability measure , we can define a probability measure by
Then,
where the convergence on the right is the standard one in the Euclidean space .
Claim 2.4.
The space is compact in the weak* topology.
Proof.
Since is compact, the fact can be deduced from the Banach–Alaoglu Theorem (cf. Theorem 3.15 in [Rudin_Functional_Analysisi_BOOK]). ∎
2.2. Pseudo subgroups
For every , the restriction map is
When we remove it from the notation and write instead of . For sets we can define
Note that .
Claim 2.5.
Let . Then and
Proof.
Let and such that . As , the intersection of with and depends only on . Hence
(2.1) |
Adding this observation to the fact is onto proves the desired claim. ∎
Now, we can pushforward measures along the inverse system of restrictions. Namely, given and a probability measure on we can define the probability measure on by
Corollary 2.6.
Claim 2.5 implies that is continuous. Therefore, is continuous.
Definition 2.7.
Let . A subset is said to be a pseudo subgroup of if there is a subgroup such that . Denote by the collection of pseudo subgroups of .131313Note that this collection is a closed subset of the subsets of .
Claim 2.8.
Let . Then, is a pseudo subgroup of if and only if , where is the subgroup of generated by .
Proof.
If there is a subgroup for which , then in particular . Because is always contained in , we can conclude. ∎
Corollary 2.9.
Let . If is not a pseudo subgroup of , and restricts to (i.e., ), then is not a pseudo subgroup of .
Proof.
By Claim 2.8, if is not a pseudo subgroup of , then there is a such that while . Since , and , the same satisfies that while , proving that is not a pseudo subgroup of . ∎
Definition 2.10.
Let be a subset.
-
(1)
The set is the collection of all that are concentrated on pseudo subgroups of , namely . Equivalently, .
-
(2)
The set is the subset of of distributions that are -invariant when defined. Namely, assuming are finite subsets such that
we have . We call probability distributions in pseudo invariant random subgroups over , or pseudo IRSs in short.
-
(3)
Let and .
Remark 2.11.
Note that and .
Lemma 2.12.
Let be subsets.
-
(1)
The restriction to of a distribution over pseudo subgroups of , is a distribution over pseudo subgroups of . Namely,
Similarly, the restriction to of a pseudo IRS of is a pseudo IRS of . Namely,
-
(2)
The (Borel) probability distributions over subgroups of are the intersection of pull-backs of all distributions over pseudo subgroups in finite subsets. Similarly, the IRSs of are the intersection of pull-backs of all pseudo IRSs of finite subsets. Namely,
Proof.
Assume further that . By Claim 2.5, for every finite ,
Thus
If such that for every , then the same applies to since it contains , and hence
Therefore is in and is proved.
By clause , for every we have and . In particular, when we have and . Hence
We are left to prove the reverse containments. Recall the open cover of the complement of the subgroups described in the proof of Claim 2.1. Given such that , we have
Hence, by the union bound (Boole’s inequality), there is either a pair such that , or a single element such that , or . Thus, if , then . In particular , proving the first reverse containment.
Remark 2.13.
The system of finite subsets of forms a directed system induced by inclusion. The pushforwards along restriction maps thus form an inverse system of maps between the compact sets for . In this language, Lemma 2.12 says that
2.3. Computational aspects of pseudo invariant random subgroups
Claim 2.14.
Given where , there is an algorithm deciding whether is a pseudo subgroup of .
Proof.
Given a finite set of elements , and another element , there is an algorithm to decide whether . Indeed, one can use the following:
-
(1)
Construct the core graph of via the method of Stallings’ foldings [Stallings_Foldings_of_G_trees, KAPOVICH2002608].
-
(2)
Check whether the path that begins at the basepoint of and traverses is closed.
-
(3)
If the path is closed, then . Otherwise .
By Claim 2.8, is not a pseudo subgroup of if and only if there exists a such that but . Hence, a sufficient and necessary condition for to be a pseudo subgroup of is to be accepted by the following algorithm:
-
(1)
Construct .
-
(2)
For every satisfying , reject if is a closed path in .
-
(3)
Accept if passed for every .
∎
Remark 2.15.
Recall that a free group is LERF (locally extended residually finite), i.e., for every finite set and every such that , there exists a finite index subgroup for which
This was originally proved by M. Hall [hall_1949], and can be proven using the method of Stallings’ foldings [Stallings_Foldings_of_G_trees, KAPOVICH2002608]. It follows that for every a pseudo subgroup of a finite , there exists a finite index subgroup for which .
Lemma 2.16.
Let be a finite set. Then, the set of distributions over pseudo subgroups of and the set of pseudo IRSs over are computable polytopes in . Namely, they are defined by (computable) integral linear equations and inequalities.
Proof.
Recall that . Namely, all these distributions are vectors which are indexed by subsets of . We describe the defining system of equations and inequalities forcing a vector to belong to . First, it needs to be a probability measure, hence
Further, for every , we can run the algorithm in Claim 2.14 to decide whether is a pseudo subgroup, and if not to add the constraint . This is the defining system of . To get the defining system of , we need to add the following: For every , we can check whether for all , the sets and are also contained in , and if so add the constraints
∎
3. Main Theorem I: Approximating the values of subgroup tests
Recall the definitions of challenges, tests and values from Section 1.1. Let be a test over the generating set , with challenges and rational distribution over . Let , which is a finite subset of . We can extend the definition of in the following way: For every satisfying , and for every distribution , let
The next claim is formalizing the following intuitive idea: Since depends only on the restriction of to , then replacing by its push-forward along any restriction will not change its value against .
Claim 3.1.
Under the extended definition of the value, for every and such that , and for every , we have
Proof.
By definition, sampling is the same as sampling and outputting . Also, since . Hence,
∎
Recall the first clause of Theorem 1.10: There is a Turing machine that takes as an input the encoding of a test , and outputs an infinite sequence of non-decreasing numbers such that .
Proof of Theorem 1.10 .
Let be a finite action. The value of against is
which is a rational combination of evaluations of . Since the sets , functions and coefficients can be calculated in finite time from the encoding of , and can be calculated from for every , the value of against can be computed in finite time.
There is a computable countable enumeration of all (equivalence classes of) finite -actions. To see that, note that a map is defined by the tuple . Since the elements of can be ordered (e.g., lexicographically), as well as the elements of , the infinite set of tuples inherits a countable enumeration. Denote by the action according to this enumeration. For , let
By our discussion so far, the number can be computed in finite time. Moreover, is a non-decreasing sequence by definition. Since , the sequence tends to the supremum of over , which is, by (1.4), . ∎
We now recall the second clause of Theorem 1.10: There is a Turing machine that takes as an input the encoding of a test , and outputs an infinite sequence of non-increasing numbers such that .
Proof of Theorem 1.10 .
Let as before, and let be a finite subset containing . Recall the definition of the pseudo invariant random subgroups over (Definition 2.10), and its pre-image . Recall also Claim 3.1 and the discussion before it, where we extended the notion of the value to probability distributions over subsets of that contain . The -approximation to the ergodic value of is defined to be
Let be a sequence of finite subsets of , all containing , such that and . We now claim the following:
-
(1)
is computable.
-
(2)
The sequence is non-increasing.
-
(3)
.
By Lemma 2.16, is defined by a computable integral system of equations and inequalities. Thus, maximizing a linear functional over it can be done using linear programming in finite time, proving .
By Lemma 2.12, whenever we have . Hence , which implies , proving .
Moreover, which implies . On the other hand, by Claim 2.4 the space is compact, and since are closed sets they are also compact. Hence there is a distribution satisfying . By the compactness of , there is a weak limit to some subsequence of , which we denote by . Without loss of generality the subsequence is the original sequence. Since for every , and since are closed, we deduce that . By the definition of a weak limit, since is a continuous functional, we have
proving . Therefore, choosing finishes the proof. ∎
4. Examples of tests
4.1. Relations verification test
Given a finite subset , the -verification test checks whether is contained in the sampled subgroup . It contains a single challenge with , namely
The sofic value of is always , e.g., using the Dirac measure concentrated on the group itself (which associates via with the trivial action of on a singleton). Moreover, if the conjugation orbit of a finite index subgroup passes with probability , then .
Remark 4.1.
Though the value of the -verification test does not give us much information about , we will see in Section 5 that the potential robustness of this test is an important property of . Moreover, this test has variations where you sample according to some distribution and check only that and not that all of is in . See for example [CL_part1, CVY_efficient] for more on that.
4.2. Separation test
Given two finite subsets , the -separation test checks whether is contained in the sampled subgroup but is not. It contains a single challenge with , namely
Definition 4.2.
A group is residually finite if for every finite subset where , there is a finite index normal subgroup such that .
Claim 4.3.
A finitely presented group is residually finite if and only if for every finite such that 141414The subgroup is the smallest normal subgroup containing . there is a finitely described IRS passing with probability .
Proof.
Let be residually finite, and a finite set disjoint from . By Definition 4.2, there is a finite index normal subgroup such that . Using the sequence of epimorphisms
we can define the left action of on which passes with probability .
On the other hand, given a finite subset , let be any set of representatives for . Since , we have . Hence, by our assumption, there is an action with . Without loss of generality, is transitive (otherwise, its restriction to any orbit will do). The fact implies in particular that for any . Thus, factors through and defines an action by . By the fact , the permutation has no fixed points for every , and thus has no fixed points for every . Hence is a normal subgroup of satisfying , as required to deduce is residually finite. ∎
Let be a finite set. Given two permutations , we define their normalized Hamming distance to be
(4.1) |
Definition 4.4.
A finitely presented group is sofic if for every such that , and for every , there is a finite set and an action such that
Remark 4.5.
It is straightforward to see that soficity is a relaxation of residual finiteness.
Proposition 4.6.
A finitely presented group is sofic if and only if for every finite such that we have .
Proof.
Assume is sofic, and let be a finite set such that . For every , there is an action , such that
Hence,
Since and are fixed, and is arbitrarily small, we have .
On the other hand, if for every we have , then for every there is an action with . Now, since
we deduce that
and is sofic. ∎
4.3. Formula satisfaction test
Let be a boolean formula. The test is defined over the formal set of generators . It has a single challenge , where . Note that the characteristic function of every subset is already in the form , and thus induces an assignment for the formal variables. The decision predicate accepts if induces a satisfying assignment for , i.e.,
It is straightforward that if and only if is satisfiable.
There is a more interesting version of this test, when the formula is a 3CNF. Recall that a 3CNF is an and of clauses of the form , where are taken from a fixed set of formal generators (as before), , and we interpret and . Now, checking that is satisfied by is the same as going over all the clauses and checking that they are all satisfied. Thus, we can define a randomized version of this check by associating a challenge with each clause , by letting and . Then, by choosing a uniform distribution over challenges, the associated test checks whether a uniformly chosen clause is satisfied by . The famous PCP theorem [PCP_thm] says that any 3CNF can be (efficiently) transformed into a not much larger 3CNF , such that if , then , and if , then The scaled up version of this argument shows that any problem (see [BFL91]) can be reduced to approximating the sofic value of a test.151515This last remark is meaningless as long as we do not specify the exact way we encode tests. The point is that there is a natural way to encode tests such that can be reduced to approximating the sofic value immediately from [BFL91]. In any case, this paper is devoted to proving a much stronger conclusion, which is that approximating the sofic value is as hard as the Halting Problem.
5. Robustness
In this section we define a rigidity property of tests called robustness (see Definition 5.7). Loosely, a test is robust if every almost optimal strategy against it is close to an optimal one. Though this notion is natural, one needs to clarify what is the measure of distance between strategies.
5.1. Edit distance
We first define a generalized version of the normalized Hamming distance (4.1). Let and be two finite sets, and assume . Given two permutations and , we define their normalized Hamming distance with errors to be
(5.1) |
If for every we let then we can define equivalently
which is the way we defined the normalized Hamming distance in (4.1). Since the generalized version is the same as the usual Hamming distance when , we use the same notation for both and just call them the Hamming distance from now on.
Let be a function, which we call the significance function. Given two actions and , we define their -weighted distance as follows:
For every , let be . Now, the -weighted edit distance171717 There is a more graph theoretical way of viewing the edit distance, in the spirit of the section 5.1.1. It essentially measures how much one Schreier graph needs to be changed to get to the other graph. For more on that, see Section 4 of [CL_part1] and the -graph notion in [BC22]. of and is the following minima:
Note that the edit distance is independent of the specific embedding of in . The edit distance induces a metric on . For every , let
(5.2) |
where recall the definition of in (1.3).
5.1.1. A graph theoretic perspective
Definition 5.1.
Let be a group with a finite generating set. Let be a set and an action of on . The generalized Schreier graph is a directed graph with vertex set and edges labeled by . The edges of are defined as follows: For every and , there is a directed edge from to labeled by , namely
We may say that an edge is labeled by some inverse of , such as . By that we mean the edge labeled by and oriented in the other direction, namely . This is natural as . Moreover, when generates , it includes words with inverses, and so when one associates words in the generators to paths in a labeled graph, this interpretation is natural (cf. [KAPOVICH2002608]).
Remark 5.2.
This generalizes the usual notion of a Schreier coset graph by choosing to be the natural action of on right -cosets.
The Schreier graph does not provide any new data about , but it provides a graph theoretic perspective for approaching IRSs. For example, a fact that we repeatedly use is that orbits of the action are in correspondence with the connected components of the Schreier graph. Moreover, the edit distance is better understood as a distance between graphs: Given two labeled graphs, one wants to transform one to the other. Whenever the endpoint of an edge labeled by is changed, there is an associated cost . The same cost is incurred when deleting completely or adding such an -labeled edge. The goal is to minimize the cost of moving between the graphs (up to isomorphism). The edit distance is exactly the minimal cost of such a transformation.
5.1.2. Topology induced by the edit distance
It turns out that regardless of the significance function , the topology induced by the edit distance is strictly stronger than the weak∗ topology:
Proposition 5.3 (Topology induced by the edit distance).
We have the following:
-
(1)
Given two sequences such that converges to in the weak∗ topology and , then converge to the same in the weak∗ topology.
-
(2)
There are sequences that converge in the weak∗ topology but are not Cauchy sequences in the edit distance topology.
Remark 5.4.
The above proposition should be contrasted with the amenable case, studied in [BLT, Proposition 6.8], where the authors showed that the converse of is true when the free group is replaced by an amenable group. See [BLT, Remark 6.9] for a different proof of the above clause .
Proof of Proposition 5.3.
Let be the minimum value of . Since is finite, . So where is the edit distance with significance function equal to the constant . So . This reduces the problem to the ordinary edit distance.
Let be finite subsets of . By Remark 2.3, we have
So it suffices to prove
Let be a large enough radius so that is contained in the ball of radius with respect to the word metric induced by .
Because and are finitely described, there are actions and which induce and respectively. We may assume that for each , either or . After conjugating if necessary, we may further assume
Given an integer , let be the set of all such that there exists in the ball with . The definition of edit distance gives
Suppose for some . Then there is a such that . Let with . Since , if , then and . Hence and so . Therefore,
Since is independent of , and , we can deduce that
for all .
Because , if , then if and only if . Hence
as . This finishes the proof of (1).
To prove item (2), recall that if is a graph then a subset is independent if no edge in has both endpoints in . If is finite then the independence ratio of is where is an independent set of maximum cardinality. If is an action then we let where is the associated action graph.
For any actions and , we have . So it suffices to show there exist action sequences , whose corresponding IRSs weak∗ converge to the same IRS and for some constant .
By using a first moment argument, it can be shown that if is chosen uniformly at random and if is the associated action graph then with high probability where is a constant depending only on (which we assume is at least 2). Moreover, there are sets with as such that if for all and are the corresponding IRSs then weak∗ converges to the trivial subgroup . This can be derived by estimating the expected number of short cycles in the graph [MR1725006]. A more precise estimate on is obtained in [MR4383230].
Therefore, there exists a sequence of actions such that if are the corresponding IRSs then weak∗ converges to the trivial subgroup and .
On the other hand, there exist whose action graphs are bi-partite such that if is the corresponding IRS then also weak∗ converges to the trivial subgroup . Since the independence ratio in this case is 1/2, it follows that the sequence cannot be Cauchy in the edit distance.
∎
5.2. Robustness
Given , we can define a map that given a reduced word counts the number of appearances of or in . Namely, if , where and , is a reduced word, then .
Let be a test with its usual associated data , and . We define , the significance function associated with , as follows:
(5.3) |
Claim 5.5.
Let , and be a test. Let and be finite sets, and and be actions. Assume . Then .
Proof.
Assume without loss of generality that and that . Let . For every , we have
(5.4) |
Inequality (5.4) is a consequence of the following implication: If acts on the same way as , then in particular is either in both and or in none of them. For , let be the suffix of from position , namely . Moreover, let be the empty word . Then
(5.5) |
Inequality (5.5) is a consequence of the following implication: If the endpoint of the path beginning at and labeled by in is different than the endpoint of the same path in , then the two paths diverged at some point. Now,
(5.6) |
Inequality (5.6) is a consequence of the union bound and the logic tautology , where . But, since
and since is uniformly distributed given that is uniformly distributed, we can deduce that
(5.7) |
All in all,
(5.8) |
Hence, by the union bound and inequality (5.8), we have
(5.9) |
Lastly, by combining
with (5.9), we deduce the claim. ∎
Remark 5.6.
For those familiar with robustness of non-local games, plays a somewhat analogous role to the state-dependent distance (cf. Section 4 of [coladangelo2017robust], for example).
Robustness is a reverse implication to the one in Claim 5.5.
Definition 5.7.
Let be a test and a function satisfying . We say that is -robust if:
-
(1)
There exists an optimal strategy, namely such that .
-
(2)
Almost optimal strategies are close to optimal ones, namely, if satisfies , then there is an optimal strategy such that .
Definition 5.8 (See [CL_part1]).
Let be a finitely presented group. Then, is said to be -homomorphism stable, where , if for every where there is a -action , where and such that
Remark 5.9.
This notion is more commonly known as flexible pointwise group stability in permutations. See [GlebskyRivera, ArzhantsevaPaunescu, BLT, BeckerLubotzky, CL_part1, CVY_efficient] for various results in this theory.
Fact 5.10.
The group is -homomorphism stable if and only if the -verification test is -robust. The exact relation between and can be calculated explicitly — they are constant multiples of one another, where the constants depend only on and .
6. Tailored non-local games
We commonly use both and for the set with -modular arithmetic.
6.1. Non-local games preliminaries
We repeat the definitions which appeared in Section 1.5 of the introduction. Since this paper is not focused on non-local games, we treat them somewhat technically. For formal definitions associated with the complexity class , we refer to [watrous2009quantum, vidick2016quantum]. For the connection between and nonlocal games, a good starting point is [cleve2004consequences].
Definition 6.1.
A (synchronous) non-local game consists of a finite graph , a length function , formal sets of generators for every vertex , a distribution over the edges , and decision functions for every edge , where . We denote by the set .
Remark 6.2.
Anyone familiar with the definition of a non-local game will immediately notice that the standard formalism is different than the one we chose here. Usually, a (synchronous) non-local game is defined as a pair of finite sets , commonly referred to as the question and answer sets respectively, together with a distribution over pairs of questions and a decision predicate satisfying for any and .
It is quite straightforward to move between the definitions. A way one can extract the data of Definition 6.1 from the above is as follows: Let be the constant function , and fix an embedding of into . The vertices of the underlying graph will be , and the support of will be the edge set . The formal generator corresponds to the bit of the answer when is asked as a question. Then, given that were asked, a pair of answers can be encoded as a map , where
(6.1) |
The notation (6.1) will be used repeatedly in the text. Lastly, , where is the aforementioned encoding.
In the other direction, we can define to be all bit strings shorter than , to be the vertex set , and stays the same. Finally, when are of the correct lengths, i.e. , and is when either of them is of the incorrect length. Since we consider only synchronous strategies, the condition is never checked in practice for and we can assume it was satisfied beforehand.
Because of this correspondence, we may refer to as the answer length function.
Non-local games are often dramatized as two-prover interactive proofs with one round [ben1988multi]. This perspective may help some readers to absorb the upcoming technicalities better. Two players, that can share some resources beforehand (random bits in the classical case and entangled qubits in the quantum case), are separated spatially — e.g., they are seated in far away rooms. A referee samples a pair of questions and sends one to each player. The players then use their shared resources to come up with answers, and send them back to the referee. The referee then decides, using the decision predicate, whether the players won or lost. The decision predicate as well as the distribution over possible questions are assumed to be known to the players beforehand. We demonstrate this dramatization in Example 6.26.
The following is the quantum mechanics’ analogue of a probability distribution. Similar to the way probability distributions (over finite sets) are collections of non-negative real numbers that add up to , the quantum analogue would be “non-negative” matrices that add up to the identity.
Definition 6.3.
A positive operator valued measure (POVM) of dimension with outcomes in a set is a mapping such that is a positive (semi-definite) matrix for every and . It is called projective, or a projection valued measure (PVM), if in addition is an orthogonal projection for every , namely , where is the conjugate transpose operation. Given that is a PVM and , we define for the marginal of to be by
(6.2) |
Thus, . Furthermore, the matrix is an order unitary, which we call the binary observable. All in all, a PVM can be given either as a map , or as a map such that the images of are commuting involutions — i.e., they square to the identity.
As its name suggests, every POVM defines a probability distribution over its answer set as follows
(6.3) |
where is the dimension normalized trace on matrices. Such an answer is said to be sampled according to and we denote it by .
Remark 6.4.
In the case of a PVM with outcome set , the observables generate a unitary representation of . The ’s in this case can be read from the unitary representation using the Fourier transform of the representation. Hence, for the case of , a PVM can be given as a collection of mutually perpendicular ortohgonal projections that sum up to the identity, or as a unitary representation of , and one can move from one perspective to the other without losing information.
Definition 6.5.
A (synchronous quantum) -dimensional strategy for is a map , where the images are involutions, and commutes for every fixed — by saying that a set of matrices (or later permutations) commutes, we mean that every pair of elements in this set commute. Another way of viewing strategies is by saying that they associate with every vertex an -dimensional PVM (in observable form). The strategy is said to be commuting along edges if commutes for every edge .
Such a strategy defines for every edge a probability distribution over functions as follows:
(6.4) |
where is again the dimension normalized trace on matrices, is the projection on the -eigenspace of , and is the projection on its -eigenspace.181818Note that , which exactly shows how to move from this observable form of the PVMs induced by back to its projection based form. We say that was sampled according to if it has this distribution, and we denote it by .
Remark 6.6.
The images of according to are usually called binary observables, and the value is usually called the measurement outcome.
Claim 6.7.
There is a procedural (i.e., algorithmic) way of sampling according to : Since commutes, these matrices have a mutual orthonormal basis of eigenvectors . Similarly, have a mutual orthonormal basis of eigenvectors . First, sample uniformly at random. Then, sample according to the squared length of its projection on , namely
where is the standard inner product on , i.e., . Finally, is determined as follows — for , , and for , . All in all,
which is independent of the ordering between and . Note also that the probability a specific is measured is independent of the chosen bases and .
Proof.
The function is sampled in the above procedure if for the resulting pair we have and . Denote by the product , and similarly . So,
By the fact that is a mutual orthonormal basis of eigenvectors for , every is either in the image of or in its kernel. Therefore, the above sum can be written as
Since is an orthonormal basis and are orthogonal projections, for every we have
Thus,
where the last equality uses the cyclicity of the trace and the fact is a projection. All in all,
as needed. ∎
Remark 6.8.
The formula (6.4) for the probability a specific is sampled is motivated by the Born measurement rule from quantum mechanics. In quantum mechanics, a (projective) measurement is specified by choosing a PVM acting on a separable Hilbert space (Definition 6.3 provided the finite dimensional case). Given a quantum state (unit vector) , a measurement of returns the outcome with probability
If there are two systems (e.g. two “players”), then a Hilbert space , is associated to each of them respectively, such that the joint Hilbert space is . The joint distribution on outcomes is then given by probabilities
(6.5) |
where is transposition with no complex conjugation, as opposed to .191919It is common to drop the transposition on the POVM in (6.5). But, in this case, a transposition will appear in the calculation using traces (6.6). So, we decided on this equivalent formulation. The case where and have the same (finite) dimension , and for orthonormal bases and of and respectively, termed “maximally entangled,” yields
(6.6) |
Definition 6.9.
The value of the strategy against the non-local game is
The (synchronous) quantum value of , , is the supremum over all possible strategies against .
The main result in [MIPRE] is a reduction from the Halting Problem to approximating the quantum value of a game:
Theorem 6.10 ().
There exists a polynomial time algorithm that takes as input (the encoding of) a Turing machine and outputs (the encoding of) a non-local game such that:
-
(1)
Sampling according to and evaluating from the encoding of the game can be done in time , where is the bit-length of the encoding of .
-
(2)
If halts, then there exists a quantum strategy that commutes along edges, for which . In particular, .
-
(3)
If never halts, then .
6.2. Permutation strategies
Our goal in Section 7 is to relate values of subgroup tests to values of non-local games and vice versa. The strategies for subgroup tests are finitely described IRSs, which are (induced by) maps of the form , while quantum strategies for non-local games are maps of the form adhering to certain conditions. These already seem quite related, as embedding into as permutation matrices is natural, and indeed transforms every (nice enough) finitely described IRS into a quantum strategy. But, there is a problem with doing this naively: Say that we want to construct a quantum strategy, such that for a certain , regardless of anything else, will always be . Then, we can achieve it by choosing . But, is not a permutation matrix. Therefore, any that will be naively transformed into a quantum strategy by embedding into as permutation matrices cannot ensure a specific is always evaluated to . To amend that, we use a tool developed in the study of solution groups of linear constraint system games (cf. [slofstra2019set]): We add a special generator that plays the role of . Since is a central involution without fixed points, the image of under needs to behave the same. This forces to act on an even sized set. Now, if we want to translate into being always , we will need to restrict ourselves to the -eigenspace of . This leads us to the following two definitions.
Definition 6.11.
A permutation strategy for a non-local game is a map with the following properties:
-
(1)
The image of is a central involution with no fixed points. Namely,
-
(2)
The rest of the images are involutions as well. Namely,
-
(3)
The image of commutes. Namely, for every fixed vertex ,
Remark 6.12.
Let be a permutation strategy for . Embed in as permutation matrices. Note that in this viewpoint, is acting naturally on functions , and the matrices are represented with respect to the standard basis, which consists of the indicators
Moreover, for a permutation and function , we have
(6.7) |
By of Definition 6.11, is a fixed point free involution, and thus its -eigenspace and -eigenspace are of the same dimension . Note that consists of all functions that are constant along -labeled edges in the Schreier graph202020Generalized Schreier graphs were defined in Section 5.1.1. , while consists of all functions that flip their signs along -labeled edges. Let be the orthogonal projection on . Note that and .
We can choose bases for and as follows. If we denote
(6.8) |
for every , then is the perfect matching induced by -labeled edges in . So, an orthogonal basis of may be taken to be , while for one takes — contains for each vector also its opposite, so this is a basis when choosing a representative for each such pair. The union of these bases is an orthogonal basis for . Furthermore, we later refer to as the standard basis of .
Remark 6.13.
Given that is a permutation strategy, there are orbits of , and acts on them in a well-defined manner. If we arbitrarily denote each orbit as , instead of a la (6.8), then acts naturally on the set in a way that commutes with the sign flip. Hence, a permutation strategy can be defined as a map into signed permutations, where is mapped to the sign flip, and such that the variables at each vertex are mapped to commuting involutions. This point of view may be helpful for understanding some of the next claims and remarks. We decided not to use the language of signed permutations in this paper, though in the companion paper [Tailored_MIPRE] it is more pronounced.
Definition 6.14.
Let be a permutation strategy for . Recall from Remark 6.12 that is the -eigenspace of , and that is the orthogonal projection on . Then, the quantum strategy induced by is the mapping to the -corners
Claim 6.15.
Proof.
The criterion for being a quantum strategy as in Definition 6.5 is to be a map into unitaries such that the image of the generators associated with any specific vertex commute, and all the images are involutions. First,
(6.9) |
Now, the fact that commutes with for any implies that . Therefore,
and the image of is contained in . Item of Definition 6.11 states that for all , and combined with (6.9) gives
i.e., the images of are involutions. Finally, since commutes for every fixed , we have
which finishes the proof. ∎
Example 6.16 (Classical permutation strategies).
Let be the group of permutations acting on two elements. For every fixed , we can define an action as follows
It is straightforward to see that the quantum strategy associated with satisfies , and thus that is sampled according to is deterministically . Strategies of this form are usually called deterministic. By taking direct sums of such actions (for potentially different ’s), we can get any (rational) distribution over deterministic strategies. These strategies are usually called classical. So, every (rational) classical strategy can be obtained as a permutation strategy.
Claim 6.17.
Let be a permutation strategy, and the associated quantum strategy. Then, if for a word we have , then .
Proof.
Recall from Remark 6.12 that the images of act naturally, and in the same way as , on , the -eigenspace of , which consist of functions that flip along -edges. Recall also the notation from (6.8). So, the fact that means that for any function , as well. This is true in particular for , the basis of described in the same remark. Note that for every we have . Hence, for every , we have
By evaluating the functions on both sides of the above equation on the point , we get
which means . Since was any point in , and we are done. ∎
Definition 6.18.
A non-local game is said to have a perfect permutation strategy if there exists a permutation strategy such that the quantum strategy induced by it, as in Definition 6.14, satisfies .
In the rest of this subsection we further analyze permutation strategies, and specifically discuss a procedural sampling — similar to Claim 6.7 — of a according to which is induced by a permutation strategy .
6.3. Interlude — Fourier bases of actions
Definition 6.19.
Let be the standard action of on itself, namely for every . The Fourier basis of is the following collection of functions, denoted by : For every , let be
(6.10) |
where is the standard bilinear form on .212121We try to keep for the bilinear form over a finite vector space, and for the complex inner product.
Claim 6.20.
The collection is an orthonormal basis of eigenvectors for the action .
Proof.
The action acts as permutations on the standard basis of , and thus extends to act on the whole vector space of functions from to . Using the standard inner product on as before, we can check that
Furthermore,
where the last equality is derived from the fact . Lastly,
and is indeed a collection of mutual eigenvectors for . ∎
Recall basic group actions definitions. Let be a group, and be two homomorphisms, which induce actions of on and respectively. A map is a factor (of -actions) — also known as a morphism of actions or a -equivariance — if for all and , we have Note that for every set , such a map induces a morphism of oriented edge-labeled graphs between and . The actions and are said to be equivalent if there exists a bijective factor between them, which in the graph case would be a (edge-labeled oriented) graph isomorphism between and . Given two actions, we can define their sum , where is the disjoint union, to be
If is a transitive action and , then it is equivalent to the action of on , where is defined (as in Section 1.3) to be The bijective factor in this case is associating with the coset , and recalling that by transitivity every is of the form for some . So, every action is equivalent to the sum of actions of on quotients of it , for some collection of subgroups .
Assume now that is for some positive integer . Let be a transitive action. Since every subgroup of is normal, this action factors through some for . Namely, we can find a bijective factor between and the action of on itself: Fix some ; is a subspace of , and we implicitly assumed it is isomorphic to ; for every , there is an element of for which , and thus for every ; so we can associate with and with . For every , let be its image in the bijection between the quotient space and . Then,
In particular, for , the action of on can be read from what happens along edges labeled by in — if , then is in the -eigenspace of , and otherwise it is in its -eigenspace.
All in all, given a not necessarily transitive action of on , its orbits have Fourier bases as in Definition 6.19 (extended to being zero on all vertices outside of the orbit), and the disjoint union of these will be called the Fourier basis of . Note that since the bijections between the orbits and can be chosen in many ways, the Fourier basis of is determined only up to the sign of each basis vector.
6.4. Procedural sampling according to permutation strategies
Let be a permutation strategy for , and recall Remark 6.12 and the notation therein. By Definition 6.11, consists of commuting involutions. Therefore, the group generated by is a quotient of — this can be acquired by linearly extending the map for every , where is the indicator of . This means that
(6.11) |
induces an action of on . Choose to be the Fourier basis of , as was defined in the Interlude Section 6.3. Let be the collection of eigenvectors in on which acts as , namely all such that . These are the Fourier basis elements that flip along edges labeled by in . Then, is an orthonormal basis of eigenvectors for .
Claim 6.21.
Let be a permutation strategy for , and the quantum strategy associated to . Then, the sampling procedure of according to can be described as follows:
-
•
Sample uniformly at random. Then, it is part of an orbit of (defined in (6.11)) and of .
-
•
Choose a pair and such that is supported on and is supported on , with probability
-
•
For , let if and otherwise. Similarly, for , let if and otherwise.
We may use the notation instead of to refer to sampling according to a permutation strategy.
Proof.
The probability of sampling which is in is . Thus, the probability of sampling a specific pair whose intersection of supports is is
This shows that the above sampling procedure provides the same distribution as the one in Claim 6.7.
∎
6.5. Tailored games
As permutation strategies allow us to translate finitely described strategies of subgroup tests to quantum strategies of non-local games, the soon to be defined -aligned permutation strategies of tailored non-local games will allow us to move in the other direction. It is harder to motivate our choices beforehand in this case, but we hope that Section 7 clarifies what may seem cryptic in this section. We begin by restricting the family of non-local games that we focus on — for the reader familiar with linear constraint system games, we note that this family is a generalization of them.
Definition 6.22.
Colloquially, a tailored non-local game is one where reads part of , and decides according to this partial view which parity checks to apply on the whole of .222222We considered calling such games controlled linear, since it is more informative. But, since conditionally linear is a term used in [MIPRE, Tailored_MIPRE], and terms containing linear are generally overused, we decided to use a less informative notion.
Formally, a tailored non-local game is equipped with an extra structure, described shortly, and its decision functions behave canonically with respect to this extra data. Instead of a single length function , has two length functions and , and . Before, the length function described the size of the formal generating set at each vertex. Now, the formal set of generators at will be a disjoint union of the sets and , where and . The elements of are called readable variables at and the elements of the linear or unreadable variables at . Let be the readability function, namely, the indicator of whether is readable or not. In addition, is equipped with a collection of controlled linear constraints functions that take as input a function , and outputs a collection of subsets of . Namely,
The image of is interpreted as a collection of linear constraints that will be verified by the decision function. Finally, the decision function behaves as follows: It restricts to the readable variables, namely looks at , and calculates . Then, it extends such that . Finally, for every , we have , and verifies that
Namely, consists of linear constraints that needs to satisfy.
Remark 6.23.
Though being tailored seems to be quite a restrictive form for a non-local game, every game can be tailored in a trivial manner. First, all variables are declared readable, namely and . Then, if the decision function decided to accept according to the original game, then it lets be empty (and thus all linear conditions will be satisfied regardless of what is). And, if decided to reject according to the original game, then it chooses to contain the singleton as the single subset appearing in . Note that represents the linear equation , which is , and thus cannot be satisfied.
The last remark raises the question: What have we gained by defining tailored non-local games, if any game can be tailored in a straightforward manner?
Definition 6.24.
A permutation strategy for a tailored non-local game is said to be -aligned if the assignments for readable variables act on each point in either like the identity or like . This is equivalent to being a diagonal matrix when presented according to the standard basis of .232323 The standard basis is usually called the -basis in quantum information theory, and thus the name -aligned strategies.
Remark 6.25.
The classical permutation strategies described in Example 6.16 are -aligned. But, one can construct permutation strategies that induce a classical strategy in the usual sense (i.e., whose images are all commuting) without being -aligned.
It is clearer now why the way one tailors a non-local game matters: The existence of a perfect -aligned permutation strategy for the game depends on it. Let us demonstrate this with binary linear constraint system (LCS) games, and specifically with the magic square game. For a full description of the magic square game we refer to [aravind2002simple] or [Tailored_MIPRE, Example 2.19], and for a general introduction to LCS games see [cleve2017perfect].
Example 6.26 (Linear constraint system games).
Let be an matrix with coefficients, and let be a column vector in . Classically, such a pair defines a system of linear equations over . It also defines a certain non-local game which is the quantum counterpart of this classical system of equations. Let us provide a dramatization of this game, in the spirit of Remark 6.2. In , the referee chooses a random linear constraint in (i.e., a row) and sends it to one of the players, and chooses a random variable that appears in the chosen constraint (i.e., a column whose intersection with the chosen row is non-zero) to the other player. It then expects the row player to respond with an assignment to the variables appearing in this constraint, and from the column player to respond with an assignment to its single variable. The referee declares this to be a winning condition if the row player’s assignment satisfied the constraint, and both players agree on the value of the variable given to the column player. The column player is thought to hold a global assignment and answer according to it, while the row player is assumed to have a satisfying assignment to each constraint — intuitively, by checking consistency between the players, the referee verifies that the global assignment of the column player indeed satisfies all the constraints.
Let us define this game rigorously. The vertices in the underlying graph of will be indexed by the rows and columns of the matrix , namely and . There is an edge between a row and a column if and only if . The length of every column vertex is , and we denote by the variable associated with the column . The length of each row vertex is the number of ’s in the row, and we associate variables to every row . The decision function gets as input an assignment to and , and accepts if and only if
(6.12) |
Though for our discussion the distribution over edges in this game is not important, one can consider the following standard sampling scheme: 1) Choose a row uniformly at random. 2) Choose a uniform column out of the support of the chosen row.
Let us describe a non-trivial tailoring of . First, all variables are unreadable, namely and . Given that the edge was sampled, the system will consist of two checks, which are derived from (6.12):242424note that since there are no readable variables, is constant.
Then, will force to check that , and will force it to check that , and thus it acts the same way as before.
The difference between the above tailored form of and the one suggested in Remark 6.23 may seem technical. But, here all the variables are unreadable, and in the version of Remark 6.23 version all variables are readable. If all variables of a tailored game are readable, for it to have a perfect -aligned permutation strategy is the same as having a perfect classical strategy — which in turn is the same as for the linear system to have a solution. But, when all the variables are unreadable, there could be a perfect -aligned permutation strategy without having a solution. For example, the perfect strategy for the Mermin–Peres magic square game (see for example [aravind2002simple]) can be derived from a permutation strategy (see our companion paper [Tailored_MIPRE, Example 2.19]), and since there are no readable variables in this case, it is automatically -aligned. Furthermore, the magic square game has no perfect classical strategy. This shows that if we care about perfect -aligned permutation strategies, the way we tailor a game matters.
7. Main Theorem II: Tests associated with tailored games
The goal of this section is to define a mapping from tailored non-local games to subgroup tests such that:
-
(1)
Completeness: Perfect -aligned permutation strategies that commute along edges for the tailored game are translated to perfect finitely described strategies of the subgroup test.
-
(2)
Soundness: Almost perfect finitely described strategies for the subgroup test are close to almost perfect -aligned permutation strategies for the tailored non-local game.
Throughout this section, is a tailored non-local game with underlying graph , length functions with and , formal sets of generators for every vertex , where and , a readability function where , a distribution over edges , and decision functions for every edge with a controlled linear constraints map .
Definition 7.1.
The synchronous subgroup test associated with the tailored non-local game is defined as follows: The set of generators in the test is . The set which parametrizes the challenges of is the edge set of , and the distribution over challenges is of . The decision function works as follows (and the subsets can be derived from this description): Given a subgroup of ,
-
Check 1.
It verifies that
-
Check 2.
It verifies that
-
Check 3.
It verifies that
-
Check 4.
If passed all the previous checks, then the following function can be calculated: if and if . After recovering this , the subset
(7.1) can be calculated. So, the decision function verifies that
where the product is according to some pre-fixed ordering of and .
If any of the above checks did not go through, rejects , and otherwise it accepts it.
Remark 7.2.
Let us motivate the checks that the decision function is applying on . Let be a finitely described strategy. If always passes Check 1, it means that acts as a fixed point free central involution on , which is part of the requirement of a permutation strategy for (see Definition 6.11). If always passes Check 2, then the images of all other variables in are involutions, and for any fixed , commutes, which are again requirements for to be a permutation strategy of . If always passes Check 3, then the readable variables always act as either or , which implies is a -aligned permutation strategy for (see Definition 6.24). So, the goal of the first three Checks is to force any finitely described strategy of to be a permutation -aligned strategy for . Lastly, Check 4 verifies “the same” linear relations as , which means that the winning probability of against and are similar.
We are ready to formulate our main theorem:
Theorem 7.3.
Let be a tailored non-local game with , and let be its associated synchronous subgroup test, as in Definition 7.1.
-
(1)
Completeness: If has a perfect -aligned permutation strategy that commutes along edges, then has a perfect finitely described strategy.
-
(2)
Soundness: If has a finitely described strategy with , then there exists a quantum strategy such that , where is a universal constant.
The next section is devoted to the proof of Theorem 7.3.
Theorem 7.4 (, see [Tailored_MIPRE]).
There exists a polynomial time algorithm that takes as input (the encoding of) a Turing machine and outputs (the encoding of) a tailored non-local game such that:
-
(1)
Sampling according to and evaluating from the encoding of the game can be done in time , where is the bit-length of the encoding of .
-
(2)
If halts, then has a perfect -aligned permutation strategy that commutes along edges.
-
(3)
If never halts, then .
Theorem 7.4 is proved in [Tailored_MIPRE].
Corollary 7.5.
The Aldous–Lyons Conjecture 1.6 has a negative solution.
Proof.
The idea is as follows. Given a Turing machine , we can define the tailored non-local game calculated in Theorem 7.4, and then calculate the associated subgroup test from Definition 7.1. In the next paragraph, the following two facts are proved:
-
•
If halts, then has sofic value .
-
•
On the other hand, if does not halt, then the sofic value of is smaller than , where for some universal constant independent of .
Recall that by Theorem 1.10, there is a computable sequence approaching from above and another computable sequence approaching from below. If the Aldous–Lyons Conjecture 1.6 has a positive solution, then these values are the same, and as Corollary 1.12 states, one can approximate up to any predetermined additive error . As can be computed directly from , one can choose , and thus deduce whether or . This in turn allows one to decide whether halts or not. Since the Halting Problem is undecidable, this is a contradiction, which implies that Conjecture 1.6 must have a negative solution.
Let us prove the two bullets above. Since the running time of the verification procedure of bounds , we can use (1) of Theorem 7.4 to deduce that . If halts, then by (2) of Theorem 7.4 and (1) of Theorem 7.3, we have . If does not halt, then by (3) of Theorem 7.4, we have . Hence, by (2) of Theorem 7.3, we have
Rearranging this inequality, we get
The quantity can be bounded from below by , where is a universal constant that depends on the constants implicit in the notation (guaranteed by Theorem 7.4), as well as the constant appearing in Theorem 7.3. This finishes the proof. ∎
8. Proving Theorem 7.3
8.1. Proving the completeness in Theorem 7.3
This is the easy direction, we have done most of the work towards it, and here we use the notation that we have established before. Recall that . Let be a perfect -aligned permutation strategy that commutes along edges for the tailored non-local game . Further recall that the finitely described IRS samples a subgroup of as follows — it takes a uniform and outputs . By Definition 6.11,
Therefore, for every ,
and when runs against it always passes Check 1. Also by Definition 6.11,
Therefore, for every ,
and when runs against it always passes Check 2. By Definition 6.24, for every we have
Namely, for every ,
and when runs against it always passes Check 3.
We are left to show that always passes Check 4. To that end, we need a couple of claims. The first is a well known and commonly used observation in quantum information theory. For the convenience of the reader, we provide a proof.
Claim 8.1.
Let be a finite set, such that commutes252525This claim can be generalized so that a weaker assumption is used. See Section 3 of [Tailored_MIPRE]. and all the images are involutions. Then, given a fixed , we have with probability when if and only if . Namely,
Proof.
Assume without loss of generality that is fully supported. Otherwise, we can focus on and proceed in the same manner. As before, we can write , where is the projection on the -eigenspace of , and is the projection on its -eigenspace. By Definition 6.5,
Note that , and thus is always . Therefore,
Thus, if and only if if and only if , which finishes the proof. ∎
Claim 8.2.
Let be a -aligned permutation strategy for a tailored game , and let be the quantum strategy associated with (as in Definition 6.14). Then, the restriction of to the readable variables in (which we denoted by ) depends only on the vertex sampled in the beginning of the procedure outlined in Claim 6.21. Furthermore, if we denote the sampled by , emphasizing its dependence on the sampled vertex, then , where was defined in (7.1).
Proof.
Let be some fixed vertex, be the orbit of , and be a readable variable. Since is -aligned, either or . Now, as discussed in Section 6.4, acting on is equivalent to an action of on itself. Furthermore, checking whether that is supported on is in the or eigenspace of depends only on whether flips along an -labeled edge in for some vertex in . Since , it must flip along -labeled edges. So, if then and , and if then and . Since this was independent of the specific that we considered, the first part of the claim is deduced. This proves that , where was defined in Check 4 of Definition 7.1, and therefore which is the second part of the claim. ∎
We can now come back to our proof. Let us, in the spirit of previous notations, denote . Let , and let be the orbit of under . Since commutes along edges (see Definition 6.5), induces an action of . Since is commuting, the action of readable variables in is constant, namely if a readable variable acts as on some , it will act that way on all , and similarly for acting as . Furthermore, by Claim 8.2, the linear constraints sampled in the associated game at a specific vertex are the same as the ones sampled by the tailored non-local game. Joining these observations, there is a set of constraints such that for every we have . So, we can focus on the orbit , and study the corner of , which was the quantum strategy induced by , with respect to the projection . Since is a perfect strategy, for every that was sampled (according to Claim 6.21) by first choosing , and for every , we have . By Claim 8.1, this means that is the identity when restricted to functions supported on that flip along -labeled edges. By Claim 6.17, we can deduce that acts as the identity on . Since this was true for every in , and for every , we can conclude that always passes Check 4 when run against . All in all, is a perfect strategy for , proving the completeness clause of Theorem 7.3.
8.2. Proving the soundness in Theorem 7.3
The plan is as follows:
- •
-
•
Then, we can bound from below the value of this perturbed strategy using the mechanisms of Section 5.
-
•
Since the resulting perturbed strategy satisfies the first three checks of , it is a -aligned permutation strategy for . For such strategies, we prove that their value against induces a lower-bound on their value against . This will conclude the proof.
The following three propositions are the manifestation of the above plan. We abuse notations and denote by the marginal distribution on vertices of , namely
Also, recall the notion of the significance function associated with a test , defined in (5.3), and the notion of the edit distance (5.2) between finitely described IRSs. Finally, recall that , and assume .
Proposition 8.3.
Let be a finitely described strategy for , with . Denote by the losing probability in given that an edge containing was sampled as the challenge — note that . Then, there is a strategy , where , that passes with probability Check 1, Check 2 and Check 3 in the definition of the associated test , and which is close to in the following sense:
for some universal constant .
Proposition 8.4.
Every finitely described strategy for that passes with probability Check 1, Check 2 and Check 3 in the definition of the associated test is a -aligned permutation strategy for and vice versa. Moreover, there is a universal constant such that if , then where is the quantum strategy associated with (Definition 6.14).
Proposition 8.5.
Let be the test associated with the tailored non-local game . Then, there is a universal constant , such that the significance associated with the test satisfies
(8.1) | ||||
(8.2) |
Remark 8.6.
We did not try to optimise the parameters in the three preceding propositions. Actually, we decided to choose a simpler version of the associated synchronous test (Definition 7.1) that provides an deterioration in the value, as seen in Proposition 8.3. We have variations of tailored games and the associated synchronous test which allow for only a deterioration in the value, but we could not find any reduction that is independent of .
Now, we can deduce the soundness clause of Theorem 7.3 as a corollary of the preceding propositions.
Proof of (2) in Theorem 7.3.
Let be the given finitely described strategy with value . By Proposition 8.3 and Proposition 8.5, there is a strategy passing Check 1, Check 2 and Check 3 with probability , and such that the significance weighted distance between and is
Therefore, by Claim 5.5,
By applying Proposition 8.4 on , its associated quantum strategy satisfies
By choosing , we deduce the claim. ∎
Proof of Proposition 8.5.
Note that participates in most of the checks in regardless of the specific challenge. In Check 1, it appears times, and since , this is bounded by . In Check 2 it is not used. In Check 3 it is used at most times, which is bounded from above by . Lastly, the number of times appears in is bounded by the number of subsets of that contain it, which is . All in all, the significance of is at most .
Now, a variable in appears only if we sampled an edge in with as one of its endpoints. If this happened, it will appear twice in Check 1, times in Check 2, at most twice in Check 3, and at most times in Check 4. Again, and , and thus appears at most times when an edge containing is sampled. This means that the significance of is at most . Choosing completes the proof. ∎
Proof of Proposition 8.4.
Let be a finitely described strategy that passes Check 1, Check 2 and Check 3 of with probability . Then, from perfectly passing Check 1 and Check 2, is a permutation strategy, and by perfectly passing Check 3 it is -aligned. The other direction was proved in the completeness analysis in Section 8.1. We are going to prove that
(8.3) |
and thus satisfies the claim.
We start by giving expressions for the success probability of in , and of the associated in respectively. Fix an edge , and denote by and the appropriate restrictions of the permutation strategy . The intersection of orbits , where is an orbit of and is an orbit of , partitions into disjoint sets. Fixing a pair of orbits such that , by Claim 8.2, there is a set of subsets of such that
For , define
(8.4) |
Let be the set of all vertices such that , namely all vertices in the intersection of orbits that do not satisfy the subgroup test constraint induced by . Let
(8.5) |
and let
where the union is over all orbits of and that have a non-empty intersection. Then is the set of all such that , hence
(8.6) |
where
(8.7) |
and and are the unique and orbits for which .
Recall that according to the procedural sampling described in Claim 6.21, any pair of Fourier basis elements supported on -orbit and -orbit respectively, is sampled with probability (conditional on the sampled being in ), and it induces an “assignment” for which . Let be the set of all such for which does not satisfy a specific linear constraint , namely . Let
be the collection of Fourier basis pairs supported on , such that the assignment induced by the pair is rejected by . Finally let
which is exactly the collection of possible sampled pairs by the quantum strategy induced by , given that the edge was asked, that produce an assignment satisfying . Hence,
(8.8) |
where
(8.9) |
and are the unique orbits for which . Observe the following claim.
Claim 8.7.
For every , -orbit and -orbit such that , we have
(8.10) |
Let us show how to deduce our conclusion assuming the validity of this claim, and then we will address the claim itself. Given , we have
Now, for every , we have by (8.5) that , which using the above analysis means
where the last inequality uses (8.7) and the fact . Therefore,
proving (8.3) and thus the proposition. ∎
Proof of Claim 8.7.
Let be the complement of in , namely
Recall the notation and from (8.4). For , we have and by the fact is a -orbit and is a -orbit. If , then by its definition we have , and . As and are involutions, we have , which implies . All in all, is invariant under the actions of and — graphically, if we add and labels to the Schrier graph of the action on , then they agree on every and are contained in it.
For we know that while . This means that for every -labeled edge between vertices in , and agree on one of its endpoints and evaluate to opposites on the other endpoint. This implies
where is the indicator function of the set , and is the pointwise product of functions. Thus,
and since is supported on , we have
Since there are many which are supported on , we deduce that , as desired. ∎
A short detour into (quite elementary) group stability results is needed for proving Proposition 8.3.
8.3. Interlude — Group stability
Claim 8.8 (Almost involutions are close to involutions).
Let . Then, there is such that and .
Proof.
Let . Note that the restriction is an involution . Define
Then and . ∎
Claim 8.9 (Almost fixed point free are close to fixed point free).
Let be an involution. Assume . Then, there is an involution with no fixed points , where , such that .
Proof.
Let . Since , we can deduce that . If is odd, add a vertex to it and make it even. Now that is even sized, we can choose any perfect matching on it and define to be the involution induced by this perfect matching. Extend to the rest of by letting it act as on the vertices out of . Then the resulting is an involution with no fixed points, and
∎
Claim 8.10 (Almost commuting involutions are close to commuting involutions).
Given two involutions , there is an involution such that
Proof.
Let , and let . Then . Moreover, for every ,
Hence, is invariant under the actions of and . Thus, the following is well defined
We therefore can conclude that , and
which implies as required. ∎
Remark 8.11.
Claim 8.10 can be proved even without assuming that the permutations are involutions (while allowing both permutations to change and not only ). This was originally proved in [ArzhantsevaPaunescu] and was later reproved with effective bounds in [BeckerMosheiff].
The following is a special case of item of Theorem 2 in [GlebskyRivera].
Claim 8.12 (Almost actions of finite groups are close to actual actions).
Let be a finite group, and let be a function. Assume that for every , . Then, there is a homomorphism such that
Proof.
Let be a vertex such that
(8.11) |
Then, if is the orbit of with respect to the action of (the group generated by) , then for every we have . To prove that, it is enough to prove it for a single step, namely for for some (the rest is by induction). So, for every , we have
where all equalities are either or (8.11). This means that the restriction of to its action on induces a homomorphism from to .
We assumed for every , that . Therefore, there are at most many vertices in for which there is some such that . So, the union of orbits as consists of at least of the vertices in . So, if we define whenever satisfies (8.11), and otherwise, then is an action of it satisfies for every , . ∎
8.4. Proof of Proposition 8.3
Throughout this proof, we use the term triangle inequality for iterative applications of the Hamming distance’s triangle inequality, as was used in the proof of Claim 5.5.
Let be a finitely described strategy for , with . By Check 1, . By Claim 8.8, there is a function that agrees with on all , but is an involution and . Again by Check 1, , and by the triangle inequality . Let . Then, by Claim 8.9, there is a function such that for every , we have , and is a fixed point free involution with . All in all, for every we have .
Now, let be the losing probability when an edge containing is sampled in . Then, . Furthermore, by Check 2, for every we have . Therefore, by the triangle inequality, . By Claim 8.8, there is a function such that , and for every and both and . Now, by Check 1, for every we have . Thus, by the triangle inequality, . Since and are involutions, we may apply Claim 8.10 and find a function such that , and . Again by the triangle inequality, we deduce that , and that for we have
Recall that (ignoring readability). We extend to in the following way: For , let
where the product is ordered according to the index . By Check 2, for every we have
Thus, by the triangle inequality, . Note also that the images of are involutions. Therefore, for every we have
By Claim 8.12 applied to each individually, we have a function such that commutes and consists only of involutions, and
Furthermore, because of the way is constructed in the proof of Claim 8.12, by letting , the image of is still a central involution with no fixed points. So, always passes Check 1 and Check 2.
Since passes Check 3 with probability when an edge with is sampled, we deduce that every readable variable satisfies
Hence, by the triangle inequality,
By the fact that commutes for every , the property implies for every , where is the orbit of induced by (which we usually denoted as beforehand). Since commutes with all images, the same is true for the property . Hence, in each orbit of , either all vertices satisfy or none of them. We can thus define a function that agrees with on the orbits in which all readable variables act appropriately, and on the other orbits we just define for every that (note that we keep acting the same no matter what). So, we have
and now always passes Check 1, Check 2 and Check 3. By the triangle inequality, and for every we have
Choosing and , we deduce the claim.
Remark 8.13.
How better can the parameters in Propositions 8.3, 8.4 and 8.5 be? We know how to make the parameters in Proposition 8.3 polynomial in by implementing a Hadmamard code in every vertex and using [BC22] or [GowersHatami] in the analysis instead of [GlebskyRivera] — since we did not see any theoretical gain by doing that, and it was a lengthier proof, we decided not to implement this parameter improvement. We are also not sure whether the dependence in Proposition 8.4 is needed or not. Finally, the significance analysis performed in Section 5 is very crude. Can a better analysis allow to remove the dependence on altogether, at least for some rich enough family of games?