A Noncommutative Nullstellensatz for Perfect Two-Answer Quantum Nonlocal Games
Abstract.
This paper introduces a noncommutative version of the Nullstellensatz, motivated by the study of quantum nonlocal games. It has been proved that a two-answer nonlocal game with a perfect quantum strategy also admits a perfect classical strategy. We generalize this result to the infinite-dimensional case, showing that a two-answer game with a perfect commuting operator strategy also admits a perfect classical strategy. This result induces a special case of noncommutative Nullstellensatz.
1. Introduction
Quantum nonlocal games have been a vibrant area of research across mathematics, physics, and computer science in recent decades. They help understand quantum nonlocality, which was famously verified by the violation of Bell inequalities (Bell, 1964). In 1969, Clauser et al. first introduced quantum nonlocal games (Clauser et al., 1969). A nonlocal game typically involves two or more players and a verifier. The verifier sends questions to the players independently, and each player responds without any communication between them. A predefined scoring function determines whether the players win based on the given questions and their answers. The distinction between classical and quantum strategies lies in whether players can share quantum entanglement. For instance, in the CHSH game, the classical strategy limits the winning probability to at most . In contrast, quantum strategies using shared entangled states can achieve a success probability of .
The mathematical models of quantum nonlocal games are often described using algebraic structures (Dykema et al., 2019; Bene Watts et al., 2023; Watts and Helton, 2020; Lupini et al., 2020). algebras, noncommutative Nullstellensatz (see (Cimprič et al., 2013, 2014, 2015)) and Positivstellensatz (see (Helton, 2002; Helton and Mccullough, 2004)) are used for characterizing the different types of strategies for nonlocal games. Our previous work also gave an algebraic characterization for perfect strategies of mirror games using the universal game algebra, Nullstellensatz, and sums of squares (Yan et al., 2023).
Nonlocal games with two answers are games in which the set of possible responses consists of only two options (Bene Watts et al., 2023) (also called binary games in (Cleve et al., 2004)). This paper proposes a noncommutative Nullstellensatz inspired by the perfect commuting operator strategies for two-answer nonlocal games. Specifically, we proved that a two-answer game that admits a perfect commuting operator strategy also has a perfect classical strategy, a generalization of the work (Cleve et al., 2004, Theorem 3). Combined with the algebraic characterization of perfect commuting operator strategy (Bene Watts et al., 2023), we get a new form of noncommutative Nullstellensatz. Although our problem is motivated by nonlocal games, our proofs are presented in a purely algebraic form, allowing readers unfamiliar with quantum nonlocal games to directly engage with the algebraic versions of the theorems.
2. Preliminaries
2.1. Motivations
If the readers are familiar with this field, they can skip the content of this subsection.
A quantum nonlocal game can be described as a scoring function from the finite set to , where the player Alice has a question set and an answer set , while the player Bob has a question set and an answer set . In a round of the game, Alice would receive the question and answer according to and her strategy; similarly, Bob would receive the question and answer . The players cannot communicate during the game but can make arrangements before playing it. The players are considered to win the game when , and they lose in all other cases.
A (deterministic) classical strategy involves two mappings
when Alice receives a question , she responds with , and similarly, Bob responds with when he receives .
If the players share a quantum state on a (perhaps infinite-dimensional) Hilbert space , and for every question pair , Alice and Bob perform commuting projection-valued measurements (PVMs)
respectively to determine their answers, then the game is said to have a commuting operator strategy.
The PVMs satisfy the following relations:
These relations can be abstracted to obtain the universal game algebra for the nonlocal game (Bene Watts et al., 2023, Section 3).
Furthermore, if we restrict the quantum state to be a tensor , where and are in finite-dimensional Hilbert space and respectively, then we get a (finite-dimensional) quantum strategy.
By defining the three types of strategies, we know the classical strategies are contained in the quantum strategies, which are included in the commuting operator strategies. We call a strategy perfect if the players can always win the game using this strategy. Therefore, a game that admits a perfect classical strategy also has a perfect commuting operator strategy. However, the converse does not hold. For example, the famous Magic Square game admits a perfect quantum strategy but has no perfect classical strategy (Cleve et al., 2004). However, in certain exceptional cases, these strategies may be equivalent.
For a two-answer game, that is, one whose answer sets are both , if it admits a perfect quantum strategy, then Cleve, Hoyer, Toner, and Watrous showed that the two-answer game must have a perfect classical strategy (Cleve et al., 2004, Theorem 3). We contribute to extending this theorem to the infinite-dimensional case, proving that a two-answer game with a perfect commuting operator strategy also admits a perfect classical strategy. This result, combined with the work of Watts, Helton, and Klep (Bene Watts et al., 2023, Theorem 4.3), derive a version of the noncommutative Nullstellensatz using a sum of squares (SOS) expression.
2.2. Universal Game Algebra for Two-Answer Games
Let be finite sets, where , and be the free algebra generated by .
Define the two-sided ideal
and let
(2.1) |
Since
we have
Similarly, one can show that
The elements in are the relationships the generators satisfy. We can also equip with the natural involution induced by
Then is a complex algebra.
The relations in are precisely those satisfied by the PVMs of a two-answer game. Thus, this algebra can characterize the commuting operator strategies of a two-answer game. serves as the universal game algebra for two-answer games, as discussed in (Bene Watts et al., 2023, Section 3). Furthermore, is a group algebra.
Let
(2.2) |
for any , we have
(2.3) | |||
(2.4) |
Let be the group generated by elements and . Equip the group algebra of with the natural involution :
Then, we observe that
We denote
It is well known that is Archimedean (see (Cimprič, 2009, example 3) or (Netzer and Thom, 2013, Remark 4.1)), that is, for every , it can be shown that
where .
We also need to introduce the concept of representation.
Definition 2.1.
A -representation of is a unital homomorphism
where denotes the set of bounded linear operators on a Hilbert space and satisfies .
3. Main Results
Let be finite sets, where , and be the free algebra generated by . Let be the complex algebra defined in the previous Subsection 2.2. Our main result is stated below:
Theorem 3.1.
Let denote the universal game algebra for two-answer games. Let
(3.1) |
be the index set of , where
(3.2) |
Let be the left ideal generated by . Then
if and only if there exists a representation
such that
Remark 1.
We can interpret the set as the invalid determining set of a two-answer game, where the scoring function
See (Bene Watts et al., 2023, Definition 3.4).
We prove this theorem by the following propositions.
Proposition 3.2.
((Bene Watts et al., 2023, Theorem 4.3)) Let denote the universal game algebra for two-answer games. If
there exists a representation
and , where is a separable Hilbert space, such that
for all .
We emphasize that is a separable Hilbert space, which will be used in the proof of Proposition 3.3. For completeness, we briefly outline the proof given by Watts, Helton, and Klep in (Bene Watts et al., 2023, Theorem 4.3).
Proof Sketch.
By the Hahn-Banach theorem (Barvinok, 2002, Theorem III.1.7) and Archimedeanity of , there exists a functional which strictly separate and , i.e
We list the properties of as follows:
-
•
-
•
for every .
Now, the GNS construction yields the desired *-representation and a cyclic vector . Define the sesquilinear form on
and
(3.3) |
By Cauchy-Schwarz inequality, is a left ideal of . Form the quotient space , and equip it with the inner product . We can complete to the Hilbert space .
It is worth mentioning that we can assume to be a separable Hilbert space, as this assumption holds because has only a finite number of generators, allowing us to generate a countable dense subset of using these generators with rational coefficients. By applying this to the quotient space, we establish the separability of .
Define the quotient map
the cyclic vector
and the left regular representation
By Archimedeanity of , it is easy to verify that is bounded for every , and thus is a representation. Finally, the result
follows from
∎
Proposition 3.3.
Let denote the universal game algebra for two-answer games. Suppose there exists a representation
and , where is a separable Hilbert space, such that
for all ( is defined in equation (3.2)). Then there exists a one-dimensional representation such that
Remark 2.
The proof below extends the argument in (Cleve et al., 2004, Theorem 4), which was originally stated for the tensor product of two finite-dimensional Hilbert spaces, to the more general setting of infinite-dimensional Hilbert spaces. In fact, the condition in Proposition 3.3 implies that the tuple
defines a perfect commuting operator strategy for the two-answer game with an invalid determining set . Furthermore, the conclusion of Proposition 3.3 demonstrates that the mappings induced by :
for , provide a perfect deterministic strategy for this game. Therefore, this proposition also implies that a two-answer game with a perfect commuting operator strategy also admits a perfect classical strategy.
Proof.
We construct the one-dimensional representation as follows. Since
for every fixed pair , we know that there exist such that
Let
(3.4) |
we have
since
and thus
for any , where is the index of the invalid determining set (3.2), see Remark 1.
Using the generators and defined in (2.2), we can rewrite:
(3.5) | ||||
Since is separable, we can choose an orthogonal basis of named
where . Define
Unlike the finite-dimensional case considered in the proof of (Cleve et al., 2004, Theorem 4), here we need to show that for every , , and are well-defined.
As and , there must exist a such that (otherwise, , which contradicts the assumption that and ). Similarly, we can also prove that is well-defined.
Let
(3.6) | ||||
(3.7) |
We have the following claim:
Claim 3.4.
We construct the one-dimensional representation as follows. For every , define
and for every , define
Then, by linearity and homogeneity, we extend to the entire game algebra .
Since and are either or , they are naturally commutative. It is straightforward to check that satisfies:
and
for all , , i.e., and satisfy the same relations as and in . Therefore, is indeed a representation of .
Remark 3.
For quantum nonlocal games, the value is the probability that the players provide the answer pair for the question pair . Since we start with a perfect commuting operator strategy, implies that the scoring function . In other words, Claim 3.4 indicates that
That is, the mappings and defined in equations (3.6) and (3.7) actually define a perfect deterministic classic strategy for the two-answer game.
Now, we provide the proof of Claim 3.4.
Proof of Claim 3.4.
We set and in equations (3.5), and compute
(3.9) | ||||
Notice that and are commutative self-adjoint operators, so and are all real numbers.
If , and since , we conclude that . Moreover, according to (3.6), if , we have
if , we have
Therefore, the value below is always positive:
Similarly, if , we have
Therefore, if either or is nonzero, we have
Since , we have
Hence, we only need to consider the case
Since we are working with infinite-dimensional separable Hilbert spaces, we modify the argument in (Cleve et al., 2004, Theorem 4) by incorporating Cauchy-Schwarz inequality and Parseval’s identity for proving that
Conversely, suppose
(3.10) |
holds. By Cauchy-Schwarz’s inequality, we know that
Since is a unit vector and the eigenvalues of can only be , we know
Applying the equality condition of the Cauchy-Schwarz inequality, and the assumption (3.10), we obtain
(3.11) |
By Parseval’s identity, we have
and
Then equation (3.11) gives
which implies that
(3.12) |
holds for every . However, equation (3.12) must fail to hold for . It is clear that equation (3.12) fails when . Assume , we find that the arguments of and both lie in the range , which contradicts equation (3.12) once again!
Therefore, when , we have shown that
That is,
which always holds, thereby proving the claim. ∎
Finally, we prove Theorem 3.1.
4. Some Discussions
Here are some remarks and discussions about our results.
Remark 4.
Watts, Helton, and Klep demonstrated that for a torically determined game, the question of whether the game has a perfect commuting operator strategy can be translated into a subgroup membership problem (Bene Watts et al., 2023, Section 5). However, this result cannot be used to prove our theorem. The reason is that if we regard as the determining set of the game, the elements in may not necessarily be expressible in the form . In other words, a two-answer game is not necessarily a torically determined game.
Remark 5.
Suppose the answer set or contains three or more elements. In that case, it is well known that our main result (Theorem 3.1) fails to hold, as there exists a nonlocal game that has a perfect commuting operator strategy but no perfect classical strategies(Cleve et al., 2004; Slofstra, 2019). From another perspective, equation (3.5) no longer holds in this case, which prevents us from reaching a similar conclusion.
Remark 6.
The algebra is finitely generated, and the set is also finite. However, the proof of our theorem relies on infinite-dimensional space. It is currently unclear whether the proof can be simplified to avoid the use of infinite-dimensional spaces.
Acknowledgements.
The authors would like to thank Sizhuo Yan, Jianting Yang, and Yuming Zhao for their helpful discussions and suggestions.References
- (1)
- Barvinok (2002) Alexander Barvinok. 2002. A course in convexity. Vol. 54. American Mathematical Soc.
- Bell (1964) John S Bell. 1964. On the einstein podolsky rosen paradox. Physics Physique Fizika 1, 3 (1964), 195.
- Bene Watts et al. (2023) Adam Bene Watts, J William Helton, and Igor Klep. 2023. Noncommutative Nullstellensätze and Perfect Games. In Annales Henri Poincaré, Vol. 24. Springer, 2183–2239.
- Cimprič (2009) Jakob Cimprič. 2009. A representation theorem for Archimedean quadratic modules on rings. Canad. Math. Bull. 52, 1 (2009), 39–52.
- Cimprič et al. (2014) Jakob Cimprič, J William Helton, Igor Klep, Scott McCullough, and Christopher Nelson. 2014. On real one-sided ideals in a free algebra. Journal of Pure and Applied Algebra 218, 2 (2014), 269–284.
- Cimprič et al. (2013) Jakob Cimprič, J William Helton, Scott McCullough, and Christopher Nelson. 2013. A noncommutative real nullstellensatz corresponds to a noncommutative real ideal: Algorithms. Proceedings of the London Mathematical Society 106, 5 (2013), 1060–1086.
- Cimprič et al. (2015) J Cimprič, J Helton, S McCullough, and C Nelson. 2015. Real nullstellensatz and*-ideals in*-algebras. The Electronic Journal of Linear Algebra 30 (2015), 19–50.
- Clauser et al. (1969) John F Clauser, Michael A Horne, Abner Shimony, and Richard A Holt. 1969. Proposed experiment to test local hidden-variable theories. Physical review letters 23, 15 (1969), 880.
- Cleve et al. (2004) Richard Cleve, Peter Hoyer, Benjamin Toner, and John Watrous. 2004. Consequences and limits of nonlocal strategies. In Proceedings. 19th IEEE Annual Conference on Computational Complexity, 2004. IEEE, 236–249.
- Dykema et al. (2019) Ken Dykema, Vern I Paulsen, and Jitendra Prakash. 2019. Non-closure of the set of quantum correlations via graphs. Communications in Mathematical Physics 365 (2019), 1125–1142.
- Helton (2002) J. William Helton. 2002. “Positive” Noncommutative Polynomials Are Sums of Squares. Annals of Mathematics 156, 2 (2002), 675–694.
- Helton and Mccullough (2004) J. William Helton and Scott Mccullough. 2004. A Positivstellensatz for Non-commutative Polynomials. Trans. Amer. Math. Soc. 356, 9 (01 2004), 3721–3737.
- Lupini et al. (2020) Martino Lupini, Laura Mančinska, Vern I Paulsen, David E Roberson, Giannicola Scarpa, Simone Severini, Ivan G Todorov, and Andreas Winter. 2020. Perfect strategies for non-local games. Mathematical Physics, Analysis and Geometry 23, 1 (2020), 7.
- Netzer and Thom (2013) Tim Netzer and Andreas Thom. 2013. Real closed separation theorems and applications to group algebras. Pacific J. Math. 263, 2 (2013), 435–452.
- Slofstra (2019) William Slofstra. 2019. The set of quantum correlations is not closed. In Forum of Mathematics, Pi, Vol. 7. Cambridge University Press, e1.
- Watts and Helton (2020) Adam Bene Watts and J William Helton. 2020. 3XOR Games with Perfect Commuting Operator Strategies Have Perfect Tensor Product Strategies and are Decidable in Polynomial Time. arXiv preprint arXiv:2010.16290 (2020).
- Yan et al. (2023) Sizhuo Yan, Jianting Yang, Tianshi Yu, and Lihong Zhi. 2023. A Characterization of Perfect Strategies for Mirror Games. In Proceedings of the 2023 International Symposium on Symbolic and Algebraic Computation. 545–554.