Quantum pseudo-telepathy

Last updated

Quantum pseudo-telepathy describes the use of quantum entanglement to eliminate the need for classical communications. [1] [2] A nonlocal game is said to display quantum pseudo-telepathy if players who can use entanglement can win it with certainty while players without it can not. The prefix pseudo refers to the fact that quantum pseudo-telepathy does not involve the exchange of information between any parties. Instead, quantum pseudo-telepathy removes the need for parties to exchange information in some circumstances.

Contents

Quantum pseudo-telepathy is generally used as a thought experiment to demonstrate the non-local characteristics of quantum mechanics. However, quantum pseudo-telepathy is a real-world phenomenon which can be verified experimentally. It is thus an especially striking example of an experimental confirmation of Bell inequality violations.

The magic square game

When attempting to construct a 3x3 table filled with the numbers +1 and -1, such that each row has an even number of negative entries and each column an odd number of negative entries, a conflict is bound to emerge. Mermin-Peres magic square.svg
When attempting to construct a 3×3 table filled with the numbers +1 and −1, such that each row has an even number of negative entries and each column an odd number of negative entries, a conflict is bound to emerge.

A simple magic square game demonstrating nonclassical correlations was introduced by P.K. Aravind [3] based on a series of papers by N. David Mermin [4] [5] and Asher Peres [6] and Adán Cabello [7] [8] that developed simplifying demonstrations of Bell's theorem. The game has been reformulated to demonstrate quantum pseudo-telepathy. [9]

Game rules

This is a cooperative game featuring two players, Alice and Bob, and a referee. The referee asks Alice to fill in one row, and Bob one column, of a 3×3 table with plus and minus signs. Their answers must respect the following constraints: Alice's row must contain an even number of minus signs, Bob's column must contain an odd number of minus signs, and they both must assign the same sign to the cell where the row and column intersects. If they manage they win, otherwise they lose.

Alice and Bob are allowed to elaborate a strategy together, but crucially are not allowed to communicate after they know which row and column they will need to fill in (as otherwise the game would be trivial).

Classical strategy

It is easy to see that if Alice and Bob can come up with a classical strategy where they always win, they can represent it as a 3×3 table encoding their answers. But this is not possible, as the number of minus signs in this hypothetical table would need to be even and odd at the same time: every row must contain an even number of minus signs, making the total number of minus signs even, and every column must contain an odd number of minus signs, making the total number of minus signs odd.

With a bit further analysis one can see that the best possible classical strategy can be represented by a table where each cell now contains both Alice and Bob's answers, that may differ. It is possible to make their answers equal in 8 out of 9 cells, while respecting the parity of Alice's rows and Bob's columns. This implies that if the referee asks for a row and column whose intersection is one of the cells where their answers match they win, and otherwise they lose. Under the usual assumption that the referee asks for them uniformly at random, the best classical winning probability is 8/9.

Pseudo-telepathic strategies

Use of quantum pseudo-telepathy would enable Alice and Bob to win the game 100% of the time without any communication once the game has begun.

This requires Alice and Bob to possess two pairs of particles with entangled states. These particles must have been prepared before the start of the game. One particle of each pair is held by Alice and the other by Bob, so they each have two particles. When Alice and Bob learn which column and row they must fill, each uses that information to select which measurements they should make to their particles. The result of the measurements will appear to each of them to be random (and the observed partial probability distribution of either particle will be independent of the measurement performed by the other party), so no real "communication" takes place.[ citation needed ]

However, the process of measuring the particles imposes sufficient structure on the joint probability distribution of the results of the measurement such that if Alice and Bob choose their actions based on the results of their measurement, then there will exist a set of strategies and measurements allowing the game to be won with probability 1.

Note that Alice and Bob could be light years apart from one another, and the entangled particles will still enable them to coordinate their actions sufficiently well to win the game with certainty.

Each round of this game uses up one entangled state. Playing N rounds requires that N entangled states (2N independent Bell pairs, see below) be shared in advance. This is because each round needs 2-bits of information to be measured (the third entry is determined by the first two, so measuring it isn't necessary), which destroys the entanglement. There is no way to reuse old measurements from earlier games.

The trick is for Alice and Bob to share an entangled quantum state and to use specific measurements on their components of the entangled state to derive the table entries. A suitable correlated state consists of a pair of entangled Bell states:

here and are eigenstates of the Pauli operator Sx with eigenvalues +1 and −1, respectively, whilst the subscripts a, b, c, and d identify the components of each Bell state, with a and c going to Alice, and b and d going to Bob. The symbol represents a tensor product.

Observables for these components can be written as products of the Pauli matrices:

Products of these Pauli spin operators can be used to fill the 3×3 table such that each row and each column contains a mutually commuting set of observables with eigenvalues +1 and −1, and with the product of the observables in each row being the identity operator, and the product of observables in each column equating to minus the identity operator. This is a so-called Mermin–Peres magic square. It is shown in below table.

Effectively, while it is not possible to construct a 3×3 table with entries +1 and −1 such that the product of the elements in each row equals +1 and the product of elements in each column equals −1, it is possible to do so with the richer algebraic structure based on spin matrices.

The play proceeds by having each player make one measurement on their part of the entangled state per round of play. Each of Alice's measurements will give her the values for a row, and each of Bob's measurements will give him the values for a column. It is possible to do that because all observables in a given row or column commute, so there exists a basis in which they can be measured simultaneously. For Alice's first row she needs to measure both her particles in the basis, for the second row she needs to measure them in the basis, and for the third row she needs to measure them in an entangled basis. For Bob's first column he needs to measure his first particle in the basis and the second in the basis, for second column he needs to measure his first particle in the basis and the second in the basis, and for his third column he needs to measure both his particles in a different entangled basis, the Bell basis. As long as the table above is used, the measurement results are guaranteed to always multiply out to +1 for Alice along her row, and −1 for Bob down his column. Of course, each completely new round requires a new entangled state, as different rows and columns are not compatible with each other.

Current research

It has been demonstrated that the above-described game is the simplest two-player game of its type in which quantum pseudo-telepathy allows a win with probability one. [10] Other games in which quantum pseudo-telepathy occurs have been studied, including larger magic square games, [11] graph colouring games [12] giving rise to the notion of quantum chromatic number, [13] and multiplayer games involving more than two participants. [14]

In July 2022 a study reported the experimental demonstration of quantum pseudotelepathy via playing the nonlocal version of Mermin-Peres magic square game. [15]

Greenberger–Horne–Zeilinger game

The Greenberger–Horne–Zeilinger (GHZ) game is another interesting example of quantum pseudo-telepathy. Classically, the game has 75% winning probability. However, with a quantum strategy, the players will always win with winning probability equals to 1.

There are three players, Alice, Bob, and Carol playing against a referee. The referee poses a question to each of the players. The three players each respond with an answer . The referee draws three questions x, y, z uniformly from the 4 options . As a clarification, if question triple is chosen, then Alice receives bit 0, Bob receives bit 1, and Carol receives bit 1 from the referee. Based on the question bit received, Alice, Bob, and Carol each respond with an answer a, b, c also in the form of 0 or 1. The players can formulate a strategy together prior to the start of the game. However, no communication is allowed during the game itself.

The players win if , where indicates OR condition and indicates summation of answers modulo 2. In other words, the sum of three answers has to be even if . Otherwise, the sum of answers has to be odd.

Winning condition of GHZ game
0000 mod 2
1101 mod 2
1011 mod 2
0111 mod 2

Classical strategy

Classically, Alice, Bob, and Carol can employ a deterministic strategy that always end up with odd sum (e.g. Alice always output 1. Bob and Carol always output 0). The players win 75% of the time and only lose if the questions are .

In fact, this is the best winning strategy classically. We can only satisfy a maximum of 3 out of 4 winning conditions. Let be Alice's response to question 0 and 1 respectively, be Bob's response to question 0, 1, and be Carol's response to question 0, 1. We can write all constraints that satisfy winning conditions as

Suppose that there is a classical strategy that satisfies all four winning conditions, all four conditions hold true. Through observation, each term appears twice on the left hand side. Hence, the left side sum = 0 mod 2. However, the right side sum = 1 mod 2. The contradiction shows that all four winning conditions cannot be simultaneously satisfied.

Quantum strategy

Now we have come to the interesting part where Alice, Bob, and Carol decided to adopt a quantum strategy. The three of them now share a tripartite entangled state , known as the GHZ state.

If question 0 is received, the player makes a measurement in the X basis . If question 1 is received, the player makes a measurement in the Y basis . In both cases, the players give answer 0 if the result of the measurement is the first state of the pair, and answer 1 if the result is the second state of the pair.

It is easy to check that with this strategy the players win the game with probability 1.

See also

Notes

  1. Brassard, Gilles; Broadbent, Anne; Tapp, Alain (2003). Dehne, Frank; Sack, Jörg-Rüdiger; Smid, Michiel (eds.). Multi-party Pseudo-Telepathy. Vol. 2748. Berlin, Heidelberg: Springer Berlin Heidelberg. pp. 1–11. doi:10.1007/978-3-540-45078-8_1. ISBN   978-3-540-40545-0.
  2. Brassard, Gilles; Cleve, Richard; Tapp, Alain (1999). "Cost of Exactly Simulating Quantum Entanglement with Classical Communication". Physical Review Letters. 83 (9): 1874–1877. arXiv: quant-ph/9901035 . Bibcode:1999PhRvL..83.1874B. doi:10.1103/PhysRevLett.83.1874. S2CID   5837965.
  3. Aravind, P.K. (2004). "Quantum mysteries revisited again". American Journal of Physics. 72 (10): 1303–1307. arXiv: quant-ph/0206070 . Bibcode:2004AmJPh..72.1303A. CiteSeerX   10.1.1.121.9157 . doi:10.1119/1.1773173.
  4. Mermin, N. David (1990-08-01). "Quantum mysteries revisited". American Journal of Physics. 58 (8): 731–734. doi:10.1119/1.16503. ISSN   0002-9505.
  5. Mermin, N. David (1990-12-31). "Simple unified form for the major no-hidden-variables theorems". Physical Review Letters. 65 (27): 3373–3376. doi:10.1103/PhysRevLett.65.3373. ISSN   0031-9007.
  6. Peres, Asher (December 1990). "Incompatible results of quantum measurements". Physics Letters A. 151 (3–4): 107–108. doi:10.1016/0375-9601(90)90172-K.
  7. Cabello, A. (2001). "Bell's theorem without inequalities and without probabilities for two observers". Physical Review Letters. 86 (10): 1911–1914. arXiv: quant-ph/0008085 . Bibcode:2001PhRvL..86.1911C. doi:10.1103/PhysRevLett.86.1911. PMID   11289818. S2CID   119472501.
  8. Cabello, A. (2001). "All versus nothing inseparability for two observers". Physical Review Letters. 87 (1): 010403. arXiv: quant-ph/0101108 . Bibcode:2001PhRvL..87a0403C. doi:10.1103/PhysRevLett.87.010403. PMID   11461451. S2CID   18748483.
  9. Brassard, Gilles; Broadbent, Anne; Tapp, Alain (2005-06-16). "Recasting Mermin's multi-player game into the framework of pseudo-telepathy". Quantum Info. Comput. 5 (7): 538–550. doi:10.48550/arXiv.quant-ph/0408052.
  10. Gisin, N.; Methot, A. A.; Scarani, V. (2007). "Pseudo-telepathy: Input cardinality and Bell-type inequalities". International Journal of Quantum Information. 5 (4): 525–534. arXiv: quant-ph/0610175 . doi:10.1142/S021974990700289X. S2CID   11386567.
  11. Kunkri, Samir; Kar, Guruprasad; Ghosh, Sibasish; Roy, Anirban (2006). "Winning strategies for pseudo-telepathy games using single non-local box". arXiv: quant-ph/0602064 .
  12. Avis, D.; Hasegawa, Jun; Kikuchi, Yosuke; Sasaki, Yuuya (2006). "A Quantum Protocol to Win the Graph Colouring Game on All Hadamard Graphs". IEICE Transactions on Fundamentals of Electronics, Communications and Computer Sciences. 89 (5): 1378–1381. arXiv: quant-ph/0509047 . Bibcode:2006IEITF..89.1378A. doi:10.1093/ietfec/e89-a.5.1378.
  13. Cameron, Peter J.; Montanaro, Ashley; Newman, Michael W.; Severini, Simone; Winter, Andreas (2007). "On the quantum chromatic number of a graph". Electronic Journal of Combinatorics. 14 (1). arXiv: quant-ph/0608016 . doi:10.37236/999. S2CID   6320177.
  14. Brassard, Gilles; Broadbent, Anne; Tapp, Alain (2005). "Recasting Mermin's multi-player game into the framework of pseudo-telepathy". Quantum Information and Computation. 5 (7): 538–550. arXiv: quant-ph/0408052 . Bibcode:2004quant.ph..8052B. doi:10.26421/QIC5.7-2.
  15. Xu, Jia-Min; Zhen, Yi-Zheng; Yang, Yu-Xiang; Cheng, Zi-Mo; Ren, Zhi-Cheng; Chen, Kai; Wang, Xi-Lin; Wang, Hui-Tian (2022-07-26). "Experimental Demonstration of Quantum Pseudotelepathy". Physical Review Letters. 129 (5): 050402. arXiv: 2206.12042 . Bibcode:2022PhRvL.129e0402X. doi:10.1103/PhysRevLett.129.050402. PMID   35960591. S2CID   250048711.

Related Research Articles

<span class="mw-page-title-main">Einstein–Podolsky–Rosen paradox</span> Historical critique of quantum mechanics

The Einstein–Podolsky–Rosen (EPR) paradox is a thought experiment proposed by physicists Albert Einstein, Boris Podolsky and Nathan Rosen which argues that the description of physical reality provided by quantum mechanics is incomplete. In a 1935 paper titled "Can Quantum-Mechanical Description of Physical Reality be Considered Complete?", they argued for the existence of "elements of reality" that were not part of quantum theory, and speculated that it should be possible to construct a theory containing these hidden variables. Resolutions of the paradox have important implications for the interpretation of quantum mechanics.

In physics, the no-cloning theorem states that it is impossible to create an independent and identical copy of an arbitrary unknown quantum state, a statement which has profound implications in the field of quantum computing among others. The theorem is an evolution of the 1970 no-go theorem authored by James Park, in which he demonstrates that a non-disturbing measurement scheme which is both simple and perfect cannot exist. The aforementioned theorems do not preclude the state of one system becoming entangled with the state of another as cloning specifically refers to the creation of a separable state with identical factors. For example, one might use the controlled NOT gate and the Walsh–Hadamard gate to entangle two qubits without violating the no-cloning theorem as no well-defined state may be defined in terms of a subsystem of an entangled state. The no-cloning theorem concerns only pure states whereas the generalized statement regarding mixed states is known as the no-broadcast theorem.

<span class="mw-page-title-main">Quantum teleportation</span> Physical phenomenon

Quantum teleportation is a technique for transferring quantum information from a sender at one location to a receiver some distance away. While teleportation is commonly portrayed in science fiction as a means to transfer physical objects from one location to the next, quantum teleportation only transfers quantum information. The sender does not have to know the particular quantum state being transferred. Moreover, the location of the recipient can be unknown, but to complete the quantum teleportation, classical information needs to be sent from sender to receiver. Because classical information needs to be sent, quantum teleportation cannot occur faster than the speed of light.

<span class="mw-page-title-main">Qubit</span> Basic unit of quantum information

In quantum computing, a qubit or quantum bit is a basic unit of quantum information—the quantum version of the classic binary bit physically realized with a two-state device. A qubit is a two-state quantum-mechanical system, one of the simplest quantum systems displaying the peculiarity of quantum mechanics. Examples include the spin of the electron in which the two levels can be taken as spin up and spin down; or the polarization of a single photon in which the two spin states can also be measured as horizontal and vertical linear polarization. In a classical system, a bit would have to be in one state or the other. However, quantum mechanics allows the qubit to be in a coherent superposition of multiple states simultaneously, a property that is fundamental to quantum mechanics and quantum computing.

<span class="mw-page-title-main">Quantum entanglement</span> Correlation between quantum systems

Quantum entanglement is the phenomenon of a group of particles being generated, interacting, or sharing spatial proximity in such a way that the quantum state of each particle of the group cannot be described independently of the state of the others, including when the particles are separated by a large distance. The topic of quantum entanglement is at the heart of the disparity between classical and quantum physics: entanglement is a primary feature of quantum mechanics not present in classical mechanics.

Bell's theorem is a term encompassing a number of closely related results in physics, all of which determine that quantum mechanics is incompatible with local hidden-variable theories, given some basic assumptions about the nature of measurement. "Local" here refers to the principle of locality, the idea that a particle can only be influenced by its immediate surroundings, and that interactions mediated by physical fields cannot propagate faster than the speed of light. "Hidden variables" are putative properties of quantum particles that are not included in quantum theory but nevertheless affect the outcome of experiments. In the words of physicist John Stewart Bell, for whom this family of results is named, "If [a hidden-variable theory] is local it will not agree with quantum mechanics, and if it agrees with quantum mechanics it will not be local."

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 the interpretation of quantum mechanics, a local hidden-variable theory is a hidden-variable theory that satisfies the principle of locality. These models attempt to account for the probabilistic features of quantum mechanics via the mechanism of underlying, but inaccessible variables, with the additional requirement that distant events be statistically independent.

In quantum information science, the Bell's states or EPR pairs are specific quantum states of two qubits that represent the simplest examples of quantum entanglement. The Bell's states are a form of entangled and normalized basis vectors. This normalization implies that the overall probability of the particle being in one of the mentioned states is 1: . Entanglement is a basis-independent result of superposition. Due to this superposition, measurement of the qubit will "collapse" it into one of its basis states with a given probability. Because of the entanglement, measurement of one qubit will "collapse" the other qubit to a state whose measurement will yield one of two possible values, where the value depends on which Bell's state the two qubits are in initially. Bell's states can be generalized to certain quantum states of multi-qubit systems, such as the GHZ state for three or more subsystems.

In physics, the no-communication theorem or no-signaling principle is a no-go theorem from quantum information theory which states that, during measurement of an entangled quantum state, it is not possible for one observer, by making a measurement of a subsystem of the total state, to communicate information to another observer. The theorem is important because, in quantum mechanics, quantum entanglement is an effect by which certain widely separated events can be correlated in ways that, at first glance, suggest the possibility of communication faster-than-light. The no-communication theorem gives conditions under which such transfer of information between two observers is impossible. These results can be applied to understand the so-called paradoxes in quantum mechanics, such as the EPR paradox, or violations of local realism obtained in tests of Bell's theorem. In these experiments, the no-communication theorem shows that failure of local realism does not lead to what could be referred to as "spooky communication at a distance".

<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.

<span class="mw-page-title-main">Greenberger–Horne–Zeilinger state</span> "Highly entangled" quantum state of 3 or more qubits

In physics, in the area of quantum information theory, a Greenberger–Horne–Zeilinger state is a certain type of entangled quantum state that involves at least three subsystems. The four-particle version was first studied by Daniel Greenberger, Michael Horne and Anton Zeilinger in 1989, and the three-particle version was introduced by N. David Mermin in 1990. Extremely non-classical properties of the state have been observed. GHZ states for large numbers of qubits are theorized to give enhanced performance for metrology compared to other qubit superposition states.

Relational quantum mechanics (RQM) is an interpretation of quantum mechanics which treats the state of a quantum system as being relational, that is, the state is the relation between the observer and the system. This interpretation was first delineated by Carlo Rovelli in a 1994 preprint, and has since been expanded upon by a number of theorists. It is inspired by the key idea behind special relativity, that the details of an observation depend on the reference frame of the observer, and uses some ideas from Wheeler on quantum information.

Quantum game theory is an extension of classical game theory to the quantum domain. It differs from classical game theory in three primary ways:

  1. Superposed initial states,
  2. Quantum entanglement of initial states,
  3. Superposition of strategies to be used on the initial states.

In theoretical physics, quantum nonlocality refers to the phenomenon by which the measurement statistics of a multipartite quantum system do not allow an interpretation with local realism. Quantum nonlocality has been experimentally verified under a variety of physical assumptions. Any physical theory that aims at superseding or replacing quantum theory should account for such experiments and therefore cannot fulfill local realism; quantum nonlocality is a property of the universe that is independent of our description of nature.

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.

Quantum refereed game in quantum information processing is a class of games in the general theory of quantum games. 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.

<span class="mw-page-title-main">Quantum complex network</span> Notion in network science of quantum information networks

Quantum complex networks are complex networks whose nodes are quantum computing devices. Quantum mechanics has been used to create secure quantum communications channels that are protected from hacking. Quantum communications offer the potential for secure enterprise-scale solutions.

Consider two remote players, connected by a channel, that don't trust each other. The problem of them agreeing on a random bit by exchanging messages over this channel, without relying on any trusted third party, is called the coin flipping problem in cryptography. Quantum coin flipping uses the principles of quantum mechanics to encrypt messages for secure communication. It is a cryptographic primitive which can be used to construct more complex and useful cryptographic protocols, e.g. Quantum Byzantine agreement.

Quantum secret sharing (QSS) is a quantum cryptographic scheme for secure communication that extends beyond simple quantum key distribution. It modifies the classical secret sharing (CSS) scheme by using quantum information and the no-cloning theorem to attain the ultimate security for communications.