Quantum refereed game

Last updated

Quantum refereed game in quantum information processing is a class of games in the general theory of quantum games. [1] It is played between two players, Alice and Bob, and arbitrated by a referee. The referee outputs the pay-off for the players after interacting with them for a fixed number of rounds, while exchanging quantum information.

Contents

Definition

An -turn quantum referee performs rounds of interaction with the player Alice and Bob. Each interaction involves receiving some quantum states from Alice and Bob, processing the quantum states together with the "left-over" state from the previous interaction, producing some output state, and sending part of the output state to the players. At the end of the rounds, the referee processes the final state received from the players and decides the pay-off for Alice and Bob. The role of the referee is to pass along qubits to players Alice and Bob. It is the referee's job to entangle the qubits, which is argued to be essential in quantum games. When Alice and Bob return the qubits to the referee, the referee then checks their final states. [2]

Quantum Refereed Games QuantumRefereedGames.svg
Quantum Refereed Games

Mathematically, an n-turn referee is a measuring co-strategy whose input spaces and output spaces are of the form

and

for complex Euclidean spaces and .

represent the message sent by the referee to Alice and Bob during turn , and correspond to their responses. At the end of turns, the referee produces an output

An -turn quantum refereed game consists of an n-turn referee along with functions that maps each measurement output to Alice's and Bob's pay-off.

Individual quantum refereed games may place specific restrictions on strategies Alice and Bob can choose from. For example, in nonlocal games [3] and pseudo-telepathy games, [4] Alice and Bob are allowed to share entanglement but are forbidden from communicating. In general, such restrictions may not apply in quantum refereed games.

A language L is considered to have a refereed game with error ε if it has a polynomial time verifier satisfying these conditions: for each string x∈L Alice (the yes prover) can convince the referee to accept x with probability of at least 1-ε regardless of Bob's strategy (the no prover) and for each string x∉L Bob can convince the referee to reject x with a probability of at least 1-ε regardless of Alice's strategy. [5]

Zero-sum games

Similar to a classical zero-sum game, a zero-sum quantum refereed game [1] is a quantum refereed game with the additional constraint .

It is natural to assume Alice and Bob play independent strategies in a zero-sum quantum refereed game, since it cannot simultaneously be to both players' advantage to communicate directly with one another or to initially share an entanglement state {reference}. In this case, Alice's and Bob's strategy can be represented by

and

where is the set of all n-turn strategies having input space and output space .

The combined strategy is then .

Min-max theorem

Define , and , then Alice's expected pay-off is

The optimal strategy for Alice then lies in the min-max problem

.

The above equality holds because are drawn from compact and convex sets and . It is called the min-max theorem for zero-sum quantum games.

One turn games

One turn quantum refereed games are a sub set of quantum refereed games (QRG) where there are two unbounded players (Alice and Bob) and a computationally bounded referee. They are called one turn games or QRG1 because there is only one turn per game. The game works by having each player send a density matrix to the referee who then plugs those states into his quantum circuit. The winner of the game is decided by the outcome of the circuit where, Alice wins the majority of times when a "yes" or |1> state is produced by the circuit and Bob wins the majority of the time when a "no" or |0> state is produced by the circuit. [6]  A turn consists of the referee sending a message to the prover (Alice or Bob) and then Alice or Bob sending a response back to the referee. [5] The order of the game goes as follows: Alice sends the referee her density matrix, then the referee processes Alice's state and sends a state to Bob, Bob then measures the state and sends a classical result back to the referee, the referee then checks Bob's measurement and either produces a "yes" meaning Alice wins or produces a "no" and Bob wins. [5]

Bell state games

In a Bell State quantum refereed game, there are three participants, Alice, Bob, and the Referee. The game consists of three doors. Behind each door is either an x or an o (spin up state or spin down state). The referee gives Alice and Bob three conditions about what is behind each of the doors. For example, the conditions could be: 1) Doors1 and 2 have the same. 2) Doors 2 and 3 have the same. 3) Door 1 and 3 are different.

The aim of this game is for Alice and Bob to find a matching pair behind the doors. In quantum terms, this means that Alice and Bob produce matching density states. During the game, Alice and Bob are not allowed to communicate, but they are allowed to strategize before the game begins. They do this by sharing an entangled pair of photons. Strategizing allows for Alice and Bob to maximize their changes of winning. Without strategizing, Alice and Bob have a 2/3 chance of winning. By strategizing, Alice and Bob's probability of producing matching quantum states increases from 2/3 to 3/4. [7]

Quantum interactive proof with competing provers

A quantum interactive proof with two competing provers is a generalization of the single prover quantum interactive proof system. [8] [9] It can be modelled by zero-sum refereed games where Alice and Bob are the competing provers, and the referee is the verifier. The referee is assumed to be computationally bounded (polynomial size quantum circuit), whereas Alice and Bob can be computationally unrestricted. Alice, Bob and the referee receive a common string, and after fixed rounds of interactions (exchanging quantum information between the provers and the referee), the referee decides whether Alice wins or Bob wins.

Classical games

In the classical setting, RG can be viewed as the following problem. Alice, Bob, and the referee is given some statement. Alice is trying to convince the referee that the statement is true while Bob is trying to convince the referee that the statement is false. The referee, who has limited computing power, will look at the proofs provided by Alice and Bob, ask them questions, and at the end of the day decide which player is correct (wins). The goal is for the referee to find an algorithm such that if the statement is true, there is a way for Alice to win with probability greater than 3/4, and if the statement is false, there is a way for Bob to win with probability greater than 3/4. This probability is equal to 1-ε. [5]

In the language of complexity theory, a promise problem has a classical refereed game (classical RG) if there exists a referee described by polynomial time randomized computation, such that

1. for each , there is a strategy for Alice to win with probability ≥ 3/4, and
2. for each , there is a strategy for Bob to win with probability ≥ 3/4.

It is known that RG = EXP. [10] [11]

Quantum games

Quantum interactive proof systems with competing provers is a generalization of the classical RG where the referee is now restricted to polynomial-time generated quantum circuits and may exchange quantum information with the players. [1] Therefore, QRG can be seen as the following problem. Alice, Bob and the referee is given some statement (it may involve a quantum state). Alice is trying to convince the referee the statement is true while Bob is trying to convince the referee the statement is false. The referee can ask the provers questions via quantum states, receive answers in quantum states, and analyse the received quantum states using a quantum computer. After communicating with Alice and Bob for rounds, the referee decides whether the statement is true or false. If there is a way for the referee to make a correct decision with probability ≥ 3/4, the game is in QRG. This probability is ≥ 1-ε. [5]

More formally, QRG denotes the complexity class for all promise problems having quantum refereed games defined as follows. Given a string , a promise problem is in QRG if there is a referee represented by a polynomial time generated quantum circuit such that

1. if , there exists a strategy for Alice to win with probability ≥ 3/4, and
2. if , there exists a strategy for Bob to win with probability ≥ 3/4.

It turns out that QRG = EXP — allowing the referee to use quantum circuit and send or receive quantum information does not give the referee any extra power. EXP ⊆ QRG follows from the fact that EXP = RG ⊆ QRG. proved QRG ⊆ EXP by a formulation of QRG using semidefinite programs (SDP).

Semidefinite program formulation

For a quantum refereed game, at the end of all the interactions, the referee outputs one of the two possible outcomes to indicate whether Alice wins or Bob wins .

Setting results in a quantum refereed game whose value is the maximum winning probability for Alice.

Using the same notation as the zero sum quantum refereed game as above, the referee is represented by operators , Alice may pick a strategy from , and Bob from . Define

, and
,

where is the partial trace operator.

The referee outputs with probability , and with probability . can be considered as a co-strategy that merges Alice's strategy with the referee's.

For any given strategy Alice chooses, the maximum winning probability for Bob is

,

which, by the property of the strategy representation, is equal to

.

Therefore, to maximize Alice's winning probability, , the maximum winning probability for Bob, needs to be minimized over all possible strategies. The goal is then to compute

.

This minimization problem can be expressed by the following SDP problem: [1]

.

The dimension of the input and output space of this SPD is exponential (from the tensor product states) in , and the SDP has a size polynomial in the dimension of its input and output space. Since there are efficient algorithms that can solve SDP in polynomial-time, [12] [13] [14] it follows that QRG ⊆ EXP.

See also

Related Research Articles

In physics, the CHSH inequality can be used in the proof of Bell's theorem, which states that certain consequences of entanglement in quantum mechanics cannot be reproduced by local hidden-variable theories. Experimental verification of the inequality being violated is seen as confirmation that nature cannot be described by such theories. CHSH stands for John Clauser, Michael Horne, Abner Shimony, and Richard Holt, who described it in a much-cited paper published in 1969. They derived the CHSH inequality, which, as with John Stewart Bell's original inequality, is a constraint—on the statistical occurrence of "coincidences" in a Bell test—which is necessarily true if an underlying local hidden-variable theory exists. In practice, the inequality is routinely violated by modern experiments in quantum mechanics.

In mathematics, specifically functional analysis, a trace-class operator is a linear operator for which a trace may be defined, such that the trace is a finite number independent of the choice of basis used to compute the trace. This trace of trace-class operators generalizes the trace of matrices studied in linear algebra. All trace-class operators are compact operators.

In physics, a partition function describes the statistical properties of a system in thermodynamic equilibrium. Partition functions are functions of the thermodynamic state variables, such as the temperature and volume. Most of the aggregate thermodynamic variables of the system, such as the total energy, free energy, entropy, and pressure, can be expressed in terms of the partition function or its derivatives. The partition function is dimensionless.

In linear algebra and functional analysis, the partial trace is a generalization of the trace. Whereas the trace is a scalar valued function on operators, the partial trace is an operator-valued function. The partial trace has applications in quantum information and decoherence which is relevant for quantum measurement and thereby to the decoherent approaches to interpretations of quantum mechanics, including consistent histories and the relative state interpretation.

A Tsirelson bound is an upper limit to quantum mechanical correlations between distant events. Given that quantum mechanics violates Bell inequalities, a natural question to ask is how large can the violation be. The answer is precisely the Tsirelson bound for the particular Bell inequality in question. In general, this bound is lower than the bound that would be obtained if more general theories, only constrained by "no-signalling", were considered, and much research has been dedicated to the question of why this is the case.

<span class="mw-page-title-main">LOCC</span> Method in quantum computation and communication

LOCC, or local operations and classical communication, is a method in quantum information theory where a local (product) operation is performed on part of the system, and where the result of that operation is "communicated" classically to another part where usually another local operation is performed conditioned on the information received.

In functional analysis and quantum information science, a positive operator-valued measure (POVM) is a measure whose values are positive semi-definite operators on a Hilbert space. POVMs are a generalization of projection-valued measures (PVM) and, correspondingly, quantum measurements described by POVMs are a generalization of quantum measurement described by PVMs.

In cryptography, XTR is an algorithm for public-key encryption. XTR stands for 'ECSTR', which is an abbreviation for Efficient and Compact Subgroup Trace Representation. It is a method to represent elements of a subgroup of a multiplicative group of a finite field. To do so, it uses the trace over to represent elements of a subgroup of .

Holevo's theorem is an important limitative theorem in quantum computing, an interdisciplinary field of physics and computer science. It is sometimes called Holevo's bound, since it establishes an upper bound to the amount of information that can be known about a quantum state. It was published by Alexander Holevo in 1973.

In quantum mechanics, notably in quantum information theory, fidelity is a measure of the "closeness" of two quantum states. It expresses the probability that one state will pass a test to identify as the other. The fidelity is not a metric on the space of density matrices, but it can be used to define the Bures metric on this space.

In mathematics, Hochschild homology (and cohomology) is a homology theory for associative algebras over rings. There is also a theory for Hochschild homology of certain functors. Hochschild cohomology was introduced by Gerhard Hochschild (1945) for algebras over a field, and extended to algebras over more general rings by Henri Cartan and Samuel Eilenberg (1956).

In physics, a quantum instrument is a mathematical abstraction of a quantum measurement, capturing both the classical and quantum outputs. It combines the concepts of measurement and quantum operation. It can be equivalently understood as a quantum channel that takes as input a quantum system and has as its output two systems: a classical system containing the outcome of the measurement and a quantum system containing the post-measurement state.

Amplitude amplification is a technique in quantum computing which generalizes the idea behind Grover's search algorithm, and gives rise to a family of quantum algorithms. It was discovered by Gilles Brassard and Peter Høyer in 1997, and independently rediscovered by Lov Grover in 1998.

A quantum t-design is a probability distribution over either pure quantum states or unitary operators which can duplicate properties of the probability distribution over the Haar measure for polynomials of degree t or less. Specifically, the average of any polynomial function of degree t over the design is exactly the same as the average over Haar measure. Here the Haar measure is a uniform probability distribution over all quantum states or over all unitary operators. Quantum t-designs are so called because they are analogous to t-designs in classical statistics, which arose historically in connection with the problem of design of experiments. Two particularly important types of t-designs in quantum mechanics are projective and unitary t-designs.

Entanglement distillation is the transformation of N copies of an arbitrary entangled state into some number of approximately pure Bell pairs, using only local operations and classical communication.

In quantum mechanics, and especially quantum information theory, the purity of a normalized quantum state is a scalar defined as

A Werner state is a × -dimensional bipartite quantum state density matrix that is invariant under all unitary operators of the form . That is, it is a bipartite quantum state that satisfies

The min-entropy, in information theory, is the smallest of the Rényi family of entropies, corresponding to the most conservative way of measuring the unpredictability of a set of outcomes, as the negative logarithm of the probability of the most likely outcome. The various Rényi entropies are all equal for a uniform distribution, but measure the unpredictability of a nonuniform distribution in different ways. The min-entropy is never greater than the ordinary or Shannon entropy and that in turn is never greater than the Hartley or max-entropy, defined as the logarithm of the number of outcomes with nonzero probability.

Generalized relative entropy is a measure of dissimilarity between two quantum states. It is a "one-shot" analogue of quantum relative entropy and shares many properties of the latter quantity.

In quantum information, the diamond norm, also known as completely bounded trace norm, is a norm on the space of quantum operations, or more generally on any linear map that acts on complex matrices. Its main application is to measure the "single use distinguishability" of two quantum channels. If an agent is randomly given one of two quantum channels, permitted to pass one state through the unknown channel, and then measures the state in an attempt to determine which operation they were given, then their maximal probability of success is determined by the diamond norm of the difference of the two channels.

References

  1. 1 2 3 4 Gutoski, G; Watrous J (2007). "Toward a general theory of quantum games". Proceedings of the thirty-ninth annual ACM symposium on Theory of computing. pp. 565–574. arXiv: quant-ph/0611234 . Bibcode:2006quant.ph.11234G. doi:10.1145/1250790.1250873. ISBN   9781595936318. S2CID   2329605.
  2. "Let the quantum games begin". Physics World. 2002-10-02. Retrieved 2020-11-11.
  3. Cleve, R; Hoyer P.; Toner B.; Watrous J. (2004). "Consequences and limits of nonlocal strategies". Proceedings of the 19th Annual IEEE Conference on Computational Complexity: 236–249. arXiv: quant-ph/0404076 . Bibcode:2004quant.ph..4076C.
  4. Brassard, G.; Broadbent A.; Tapp A. (2005). "Quantum pseudo-telepathy". Foundations of Physics. 35 (11): 1877–1907. arXiv: quant-ph/0407221 . Bibcode:2005FoPh...35.1877B. doi:10.1007/s10701-005-7353-4. S2CID   7395322.
  5. 1 2 3 4 5 Gutoski, Gus; Watrous, John (2005). "Quantum Interactive Proofs with Competing Provers". Stacs 2005. Lecture Notes in Computer Science. Vol. 3404. pp. 605–616. arXiv: cs/0412102 . doi:10.1007/978-3-540-31856-9_50. ISBN   978-3-540-24998-6. S2CID   15662983.
  6. Ghosh, Soumik (2020). "A study of one-turn quantum refereed games" (PDF). U Waterloo. Retrieved 2020-10-11.
  7. Web.Stanford.Edu, 2020, http://web.stanford.edu/~oas/SI/QM/notes/SIQMWeek3.pdf .
  8. Kitaev, A; Watrous J (2000). "Parallelization, amplification, and exponential time simulation of quantum interactive proof system". Proceedings of the 32nd AMC Symposium on Theory of Computing: 608–617.
  9. Watrous, J (2003). "PSPACE has constant-round quantum interactive proof systems". Theoretical Computer Science. 292 (3): 575–588. doi: 10.1016/s0304-3975(01)00375-9 .
  10. Koller, D; Megiddo N (1992). "The complexity of two-person zero-sum games in extensive form". Games and Economic Behavior. 4 (4): 528–552. CiteSeerX   10.1.1.30.7625 . doi:10.1016/0899-8256(92)90035-q.
  11. Feige, U; Kilian J (1997). "Making games short (Extended abstract)". Proceedings of the twenty-ninth annual ACM symposium on Theory of computing - STOC '97. pp. 506–516. CiteSeerX   10.1.1.5.1990 . doi:10.1145/258533.258644. ISBN   978-0897918886. S2CID   15664449.
  12. KHACHIYAN, L (1979). "A polynomial time algorithm in linear programming". Soviet Mathematics - Doklady. 20: 191–194.
  13. Grötschel, M; Lovász L.; Schrijver, A. (1988). Geometric Algorithms and Combinatorial Optimization. Algorithms and Combinatorics. Springer. ISBN   978-3-642-97883-8.
  14. Nesterov, Yurii; Nemirovskii, Arkadii (1994). Interior-Point Polynomial Algorithms in Convex Programming (PDF). SIAM Studies in Applied Mathematics. Vol. 13. doi:10.1137/1.9781611970791. ISBN   978-0-89871-319-0. S2CID   117194167.