LOCC

Last updated

LOCC paradigm: the parties are not allowed to exchange particles coherently. Only local operations and classical communication is allowed LOCC.png
LOCC paradigm: the parties are not allowed to exchange particles coherently. Only local operations and classical communication is allowed

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.

Contents

Mathematical properties

The formal definition of the set of LOCC operations is complicated due to the fact that later local operations depend in general on all the previous classical communication and due to the unbounded number of communication rounds. For any finite number one can define , the set of LOCC operations that can be achieved with rounds of classical communication. The set becomes strictly larger whenever is increased and care has to be taken to define the limit of infinitely many rounds. In particular, the set LOCC is not topologically closed, that is there are quantum operations that can be approximated arbitrarily closely by LOCC but that are not themselves LOCC. [1]

A one-round LOCC is a quantum instrument , for which the trace-non-increasing completely positive maps (CPMs) are local for all measurement results , i.e., and there is one site such that only at the map is not trace-preserving. This means that the instrument can be realized by the party at site applying the (local) instrument and communicating the classical result to all other parties, which then each perform (conditioned on ) trace-preserving (deterministic) local quantum operations .

Then are defined recursively as those operations that can be realized by following up an operation with a -operation. Here it is allowed that the party which performs the follow-up operations depends on the result of the previous rounds. Moreover, we also allow "coarse-graining", i.e., discarding some of the classical information encoded in the measurement results (of all rounds).

The union of all operations is denoted by and contains instruments that can be approximated better and better with more LOCC rounds. Its topological closure contains all such operations.

It can be shown that all these sets are different: [1]

The set of all LOCC operations is contained in the set of all separable operations. contains all operations that can be written using Kraus operators that have all product form, i.e.,

with . Not all operations in are LOCC,

i.e., there are examples that cannot be implemented locally even with infinite rounds of communication. [1]

LOCC are the "free operations" in the resource theories of entanglement: Entanglement cannot be produced from separable states with LOCC and if the local parties in addition to being able to perform all LOCC operations are also furnished with some entangled states, they can realize more operations than with LOCC alone.

Examples

LOCC operations are useful for state preparation, state discrimination, and entanglement transformations.

State preparation

Alice and Bob are given a quantum system in the product state . Their task is to produce the separable state . With local operations alone this cannot be achieved, since they cannot produce the (classical) correlations present in . However, with LOCC (with one round of communication) can be prepared: Alice throws an unbiased coin (that shows heads or tails each with 50% probability) and flips her qubit (to ) if the coin shows "tails", otherwise it is left unchanged. She then sends the result of the coin-flip (classical information) to Bob who also flips his qubit if he receives the message "tails". The resulting state is . In general, all separable states (and only these) can be prepared from a product states with LOCC operations alone. [1]

State discrimination

Given two quantum states on a bi- or multipartite Hilbert space , the task is to determine which one of two (or more) possible states it is. As a simple example, consider the two Bell states

Let's say the two-qubit system is separated, where the first qubit is given to Alice and the second is given to Bob. Without communication, Alice and Bob cannot distinguish the two states, since for all local measurements all measurement statistics are exactly the same (both states have the same reduced density matrix). E.g., assume that Alice measures the first qubit, and obtains the result 0. Since this result is equally likely to occur (with probability 50%) in each of the two cases, she does not gain any information on which Bell pair she was given and the same holds for Bob if he performs any measurement. But now let Alice send her result to Bob over a classical channel. Now Bob can compare his result to hers and if they are the same he can conclude that the pair given was , since only this allows for a joint measurement outcome . Thus with LOCC and two measurements these two states can be distinguished perfectly. Note that with global (nonlocal or entangled) measurements, a single measurement (on the joint Hilbert space) is sufficient to distinguish these two (mutually orthogonal) states.

There are quantum states that cannot be distinguished with LOCC operations. [2]

Entanglement transformations

While LOCC cannot generate entangled states out of product states, they can be used to transform entangled states into other entangled states. The restriction to LOCC severely limits which transformations are possible.

Entanglement conversion

Nielsen [3] has derived a general condition to determine whether one pure state of a bipartite quantum system may be transformed into another using only LOCC. Full details may be found in the paper referenced earlier, the results are sketched out here.

Consider two particles in a Hilbert space of dimension with particle states and with Schmidt decompositions

The 's are known as Schmidt coefficients. If they are ordered largest to smallest (i.e., with ) then can only be transformed into using only local operations if and only if for all in the range

In more concise notation:

This is a more restrictive condition than that local operations cannot increase entanglement measures. It is quite possible that and have the same amount of entanglement but converting one into the other is not possible and even that conversion in either direction is impossible because neither set of Schmidt coefficients majorises the other. For large if all Schmidt coefficients are non-zero then the probability of one set of coefficients majorising the other becomes negligible. Therefore, for large the probability of any arbitrary state being convertible into another via LOCC becomes negligible.

The operations described so far are deterministic, i.e., they succeed with probability 100%. If one is satisfied by probabilistic transformations, many more transformations are possible using LOCC. [4] These operations are called stochastic LOCC (SLOCC). In particular for multi-partite states the convertibility under SLOCC is studied to gain a qualitative insight into the entanglement properties of the involved states. [5]

Going beyond LOCC: Catalytic conversion

If entangled states are available as a resource, these together with LOCC allow a much larger class of transformations. This is the case even if these resource states are not consumed in the process (as they are, for example, in quantum teleportation). Thus transformations are called entanglement catalysis. [6] In this procedure, the conversion of an initial state to a final state that is impossible with LOCC is made possible by taking a tensor product of the initial state with a "catalyst state" and requiring that this state is still available at the end of the conversion process. I.e., the catalyst state is left unchanged by the conversion and can then be removed, leaving only the desired final state. Consider the states,

These states are written in the form of Schmidt decomposition and in a descending order. We compare the sum of the coefficients of and

00.40.5
10.80.75
20.91.0
31.01.0

In the table, red color is put if , green color is put if , and white color is remained if . After building up the table, one can easily to find out whether and are convertible by looking at the color in the direction. can be converted into by LOCC if the color are all green or white, and can be converted into by LOCC if the color are all red or white. When the table presents both red and green color, the states are not convertible.

Now we consider the product states and

Similarly, we make up the table:

00.240.30
10.480.50
20.640.65
30.800.80
40.860.90
50.921.00
60.961.00
71.001.00

The color in the direction are all green or white, therefore, according to the Nielsen's theorem, is possible to be converted into by the LOCC. The catalyst state is taken away after the conversion. Finally we find by the LOCC.

If correlations between the system and the catalyst are allowed, catalytic transformations between bipartite pure states are characterized via the entanglement entropy. [7] In more detail, a pure state can be converted into another pure state via catalytic LOCC if and only if

,

where is the von Neumann entropy, and and are the reduced states of and , respectively. In general, the conversion is not exact, but can be performed with an arbitrary accuracy. The amount of correlations between the system and the catalyst can also be made arbitrarily small.

See also

Related Research Articles

Bra–ket notation, also called Dirac notation, is a notation for linear algebra and linear operators on complex vector spaces together with their dual space both in the finite-dimensional and infinite-dimensional case. It is specifically designed to ease the types of calculations that frequently come up in quantum mechanics. Its use in quantum mechanics is quite widespread.

<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.

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.

<span class="mw-page-title-main">Quantum decoherence</span> Loss of quantum coherence

Quantum decoherence is the loss of quantum coherence, the process in which a system's behaviour changes from that which can be explained by quantum mechanics to that which can be explained by classical mechanics. In quantum mechanics, particles such as electrons are described by a wave function, a mathematical representation of the quantum state of a system; a probabilistic interpretation of the wave function is used to explain various quantum effects. As long as there exists a definite phase relation between different states, the system is said to be coherent. A definite phase relationship is necessary to perform quantum computing on quantum information encoded in quantum states. Coherence is preserved under the laws of quantum physics.

<span class="mw-page-title-main">Bloch sphere</span> Geometrical representation of the pure state space of a two-level quantum mechanical system

In quantum mechanics and computing, the Bloch sphere is a geometrical representation of the pure state space of a two-level quantum mechanical system (qubit), named after the physicist Felix Bloch.

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 3 or more subsystems.

<span class="mw-page-title-main">Superdense coding</span> Two-bit quantum communication protocol

In quantum information theory, superdense coding is a quantum communication protocol to communicate a number of classical bits of information by only transmitting a smaller number of qubits, under the assumption of sender and receiver pre-sharing an entangled resource. In its simplest form, the protocol involves two parties, often referred to as Alice and Bob in this context, which share a pair of maximally entangled qubits, and allows Alice to transmit two bits to Bob by sending only one qubit. This protocol was first proposed by Charles H. Bennett and Stephen Wiesner in 1970 and experimentally actualized in 1996 by Klaus Mattle, Harald Weinfurter, Paul G. Kwiat and Anton Zeilinger using entangled photon pairs. Superdense coding can be thought of as the opposite of quantum teleportation, in which one transfers one qubit from Alice to Bob by communicating two classical bits, as long as Alice and Bob have a pre-shared Bell pair.

<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.

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.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

Quantum walks are quantum analogues of classical random walks. In contrast to the classical random walk, where the walker occupies definite states and the randomness arises due to stochastic transitions between states, in quantum walks randomness arises through: (1) quantum superposition of states, (2) non-random, reversible unitary evolution and (3) collapse of the wave function due to state measurements.

In the case of systems composed of subsystems, the classification of quantum-entangledstates is richer than in the bipartite case. Indeed, in multipartite entanglement apart from fully separable states and fully entangled states, there also exists the notion of partially separable states.

<span class="mw-page-title-main">Helium atom</span> Atom of helium

A helium atom is an atom of the chemical element helium. Helium is composed of two electrons bound by the electromagnetic force to a nucleus containing two protons along with two neutrons, depending on the isotope, held together by the strong force. Unlike for hydrogen, a closed-form solution to the Schrödinger equation for the helium atom has not been found. However, various approximations, such as the Hartree–Fock method, can be used to estimate the ground state energy and wavefunction of the atom. Historically, the first such helium spectrum calculation was done by Albrecht Unsöld in 1927. Its success was considered to be one of the earliest signs of validity of Schrödinger's wave mechanics.

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 the theory of quantum communication, an amplitude damping channel is a quantum channel that models physical processes such as spontaneous emission. A natural process by which this channel can occur is a spin chain through which a number of spin states, coupled by a time independent Hamiltonian, can be used to send a quantum state from one location to another. The resulting quantum channel ends up being identical to an amplitude damping channel, for which the quantum capacity, the classical capacity and the entanglement assisted classical capacity of the quantum channel can be evaluated.

The entropy of entanglement is a measure of the degree of quantum entanglement between two subsystems constituting a two-part composite quantum system. Given a pure bipartite quantum state of the composite system, it is possible to obtain a reduced density matrix describing knowledge of the state of a subsystem. The entropy of entanglement is the Von Neumann entropy of the reduced density matrix for any of the subsystems. If it is non-zero, i.e. the subsystem is in a mixed state, it indicates the two subsystems are entangled.

<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.

In quantum information theory and quantum optics, the Schrödinger–HJW theorem is a result about the realization of a mixed state of a quantum system as an ensemble of pure quantum states and the relation between the corresponding purifications of the density operators. The theorem is named after physicists and mathematicians Erwin Schrödinger, Lane P. Hughston, Richard Jozsa and William Wootters. The result was also found independently by Nicolas Gisin, and by Nicolas Hadjisavvas building upon work by Ed Jaynes, while a significant part of it was likewise independently discovered by N. David Mermin. Thanks to its complicated history, it is also known by various other names such as the GHJW theorem, the HJW theorem, and the purification theorem.

In quantum physics, the "monogamy" of quantum entanglement refers to the fundamental property that it cannot be freely shared between arbitrarily many parties.

Bell diagonal states are a class of bipartite qubit states that are frequently used in quantum information and quantum computation theory.

References

  1. 1 2 3 4 Chitambar, E.; Leung, D.; Mancinska, L.; Ozols, M.; Winter, A. (2012). "Everything You Always Wanted to Know About LOCC (But Were Afraid to Ask)". Commun. Math. Phys. 328 (1): 303. arXiv: 1210.4583 . Bibcode:2014CMaPh.328..303C. doi:10.1007/s00220-014-1953-9. S2CID   118478457.
  2. Charles H. Bennett; David P. DiVincenzo; Christopher A. Fuchs; Tal Mor; Eric Rains; Peter W. Shor; John A. Smolin; William K. Wootters (1999). "Quantum nonlocality without entanglement". Phys. Rev. A. 59 (2): 1070–1091. arXiv: quant-ph/9804053 . Bibcode:1999PhRvA..59.1070B. doi:10.1103/PhysRevA.59.1070. S2CID   15282650.
  3. M. A. Nielsen (1999). "Conditions for a Class of Entanglement Transformations". Phys. Rev. Lett. 83 (2): 436–439. arXiv: quant-ph/9811053 . Bibcode:1999PhRvL..83..436N. doi:10.1103/PhysRevLett.83.436. S2CID   17928003.
  4. Guifré Vidal (2000). "Entanglement Monotones". J. Mod. Opt. 47 (2–3): 355. arXiv: quant-ph/9807077 . Bibcode:2000JMOp...47..355V. doi:10.1080/09500340008244048. S2CID   119347961.
  5. G. Gour; N. R. Wallach (2013). "Classification of Multipartite Entanglement of All Finite Dimensionality". Phys. Rev. Lett. 111 (6): 060502. arXiv: 1304.7259 . Bibcode:2013PhRvL.111f0502G. doi:10.1103/PhysRevLett.111.060502. PMID   23971544. S2CID   1570745.
  6. D. Jonathan; M. B. Plenio (1999). "Entanglement-assisted local manipulation of pure quantum states". Phys. Rev. Lett. 83 (17): 3566–3569. arXiv: quant-ph/9905071 . Bibcode:1999PhRvL..83.3566J. doi:10.1103/PhysRevLett.83.3566. S2CID   392419.
  7. Kondra, Tulja Varun; Datta, Chandan; Streltsov, Alexander (2021-10-05). "Catalytic Transformations of Pure Entangled States". Physical Review Letters. 127 (15): 150503. arXiv: 2102.11136 . Bibcode:2021PhRvL.127o0503K. doi:10.1103/PhysRevLett.127.150503. PMID   34678004. S2CID   237532098.

Further reading