Mutually unbiased bases

Last updated

An example of three bases in a two-dimensional space, where bases B1 and B2 are mutually unbiased, whereas the vectors of basis B3 do not have equal overlap with the vectors of basis B1 (as well as B2) and so B3 is not mutually unbiased with B1 (and B2). MUB bases plot.svg
An example of three bases in a two-dimensional space, where bases B1 and B2 are mutually unbiased, whereas the vectors of basis B3 do not have equal overlap with the vectors of basis B1 (as well as B2) and so B3 is not mutually unbiased with B1 (and B2).

In quantum information theory, a set of bases in Hilbert space Cd are said to be mutually unbiased to mean, that, if a system is prepared in an eigen state of one of the bases, then all outcomes of the measurement with respect to the other basis are predicted to occur with an equal probability inexorably equal to 1/d.

Contents

Overview

The notion of mutually unbiased bases was first introduced by Schwinger in 1960, [1] and the first person to consider applications of mutually unbiased bases was Ivanovic [2] in the problem of quantum state determination.

MUBs and their existence problem, is now known to have several closely related problems and equivalent avatars in several other branches of mathematics and quantum sciences, such as SIC-POVMs, finite projective / affine planes, complex Hadamard matrices etc. [see: Related Problems and Constructions sections in the below for details]

MUBs are important for quantum key distribution, more specifically in secure quantum key exchange. [3] Mutually unbiased bases are used in many protocols since the outcome is random when a measurement is made in a basis unbiased to that in which the state was prepared. When two remote parties share two non-orthogonal quantum states, attempts by an eavesdropper to distinguish between these by measurements will affect the system and this can be detected. While many quantum cryptography protocols have relied on 1-qubit technologies, employing higher-dimensional states, such as qutrits, allows for better security against eavesdropping. [3] This motivates the study of mutually unbiased bases in higher-dimensional spaces.

Other uses of mutually unbiased bases include quantum state reconstruction, [4] quantum error correction codes, [5] [6] detection of quantum entanglement, [7] [8] and the so-called "mean king's problem". [9] [10]

Definition and Examples

A pair of orthonormal bases and in Hilbert space Cd are said to be mutually unbiased, if and only if the square of the magnitude of the inner product between any basis states and equals the inverse of the dimension d: [11]

These bases are unbiased in the following sense: if a system is prepared in a state belonging to one of the bases, then all outcomes of the measurement with respect to the other basis are predicted to occur with equal probability.

Example for d = 2

The three bases

provide the simplest example of mutually unbiased bases in C2. The above bases are composed of the eigenvectors of the Pauli spin matrices and their product , respectively.

Example for d = 4

For d = 4, an example of d + 1 = 5 mutually unbiased bases where each basis is denoted by Mj, 0 ≤ j ≤ 4, is given as follows: [12]

Existence problem

Unsolved problem in mathematics:

What is the maximum number of MUBs in any given non-prime-power dimension d?

Let denote the maximum number of mutually unbiased bases in the d-dimensional Hilbert space Cd. It is an open question [13] how many mutually unbiased bases, , one can find in Cd, for arbitrary d.

In general, if

is the prime-power factorization of d, where

then the maximum number of mutually unbiased bases which can be constructed satisfies [11]

It follows that if the dimension of a Hilbert space d is an integer power of a prime number, then it is possible to find d + 1 mutually unbiased bases. This can be seen in the previous equation, as the prime number decomposition of d simply is . Therefore,

Thus, the maximum number of mutually unbiased bases is known when d is an integer power of a prime number, but it is not known for arbitrary d.

The smallest dimension that is not an integer power of a prime is d = 6. This is also the smallest dimension for which the number of mutually unbiased bases is not known. The methods used to determine the number of mutually unbiased bases when d is an integer power of a prime number cannot be used in this case. Searches for a set of four mutually unbiased bases when d = 6, both by using Hadamard matrices [11] and numerical methods [14] [15] have been unsuccessful. The general belief is that the maximum number of mutually unbiased bases for d = 6 is . [11]

Unsolved problem in mathematics:

Do SIC-POVMs exist in all dimensions?

The MUBs problem seems similar in nature to the symmetric property of SIC-POVMs. Wootters points out that a complete set of unbiased bases yields a geometric structure known as a finite projective plane, while a SIC-POVM (in any dimension that is a prime power) yields a finite affine plane, a type of structure whose definition is identical to that of a finite projective plane with the roles of points and lines exchanged. In this sense, the problems of SIC-POVMs and of mutually unbiased bases are dual to one another. [16]

In dimension , the analogy can be taken further: a complete set of mutually unbiased bases can be directly constructed from a SIC-POVM. [17] The 9 vectors of the SIC-POVM, together with the 12 vectors of the mutually unbiased bases, form a set that can be used in a Kochen–Specker proof. [18] However, in 6-dimensional Hilbert space, a SIC-POVM is known, but no complete set of mutually unbiased bases has yet been discovered, and it is widely believed that no such set exists. [19] [20]

Methods for finding mutually unbiased bases

Weyl group method

Let and be two unitary operators in the Hilbert space Cd such that

for some phase factor . If is a primitive root of unity, for example then the eigenbases of and are mutually unbiased.

By choosing the eigenbasis of to be the standard basis, we can generate another basis unbiased to it using a Fourier matrix. The elements of the Fourier matrix are given by

Other bases which are unbiased to both the standard basis and the basis generated by the Fourier matrix can be generated using Weyl groups. [11] The dimension of the Hilbert space is important when generating sets of mutually unbiased bases using Weyl groups. When d is a prime number, then the usual d + 1 mutually unbiased bases can be generated using Weyl groups. When d is not a prime number, then it is possible that the maximal number of mutually unbiased bases which can be generated using this method is 3.

Unitary operators method using finite fields

When d = p is prime, we define the unitary operators and by

where is the standard basis and is a root of unity.

Then the eigenbases of the following d + 1 operators are mutually unbiased: [21]

For odd d, the t-th eigenvector of the operator is given explicitly by [13]

When is a power of a prime, we make use of the finite field to construct a maximal set of d + 1 mutually unbiased bases. We label the elements of the computational basis of Cd using the finite field: .

We define the operators and in the following way

where

is an additive character over the field and the addition and multiplication in the kets and is that of .

Then we form d + 1 sets of commuting unitary operators:

and for each

The joint eigenbases of the operators in one set are mutually unbiased to that of any other set. [21] We thus have d + 1 mutually unbiased bases.

Hadamard matrix method

Given that one basis in a Hilbert space is the standard basis, then all bases which are unbiased with respect to this basis can be represented by the columns of a complex Hadamard matrix multiplied by a normalization factor. For d = 3 these matrices would have the form

The problem of finding a set of k+1 mutually unbiased bases therefore corresponds to finding k mutually unbiased complex Hadamard matrices. [11]

An example of a one parameter family of Hadamard matrices in a 4-dimensional Hilbert space is

Entropic uncertainty relations and MUBs

There is an alternative characterization of mutually unbiased bases that considers them in terms of uncertainty relations. [22]

Entropic uncertainty relations are analogous to the Heisenberg uncertainty principle, and Maassen and Uffink [23] found that for any two bases and :

where and and is the respective entropy of the bases and , when measuring a given state.

Entropic uncertainty relations are often preferable [24] to the Heisenberg uncertainty principle, as they are not phrased in terms of the state to be measured, but in terms of c.

In scenarios such as quantum key distribution, we aim for measurement bases such that full knowledge of a state with respect to one basis implies minimal knowledge of the state with respect to the other bases. This implies a high entropy of measurement outcomes, and thus we call these strong entropic uncertainty relations.

For two bases, the lower bound of the uncertainty relation is maximized when the measurement bases are mutually unbiased, since mutually unbiased bases are maximally incompatible: the outcome of a measurement made in a basis unbiased to that in which the state is prepared in is completely random. In fact, for a d-dimensional space, we have: [25]

for any pair of mutually unbiased bases and . This bound is optimal: [26] If we measure a state from one of the bases then the outcome has entropy 0 in that basis and an entropy of in the other.

If the dimension of the space is a prime power, we can construct d + 1 MUBs, and then it has been found that [27]

which is stronger than the relation we would get from pairing up the sets and then using the Maassen and Uffink equation. Thus we have a characterization of d + 1 mutually unbiased bases as those for which the uncertainty relations are strongest.

Although the case for two bases, and for d + 1 bases is well studied, very little is known about uncertainty relations for mutually unbiased bases in other circumstances. [27] [28]

When considering more than two, and less than bases it is known that large sets of mutually unbiased bases exist which exhibit very little uncertainty. [29] This means merely being mutually unbiased does not lead to high uncertainty, except when considering measurements in only two bases. Yet there do exist other measurements that are very uncertain. [27] [30]

Mutually unbiased bases in infinite dimension Hilbert spaces

While there has been investigation into mutually unbiased bases in infinite dimension Hilbert space, their existence remains an open question. It is conjectured that in a continuous Hilbert space, two orthonormal bases and are said to be mutually unbiased if [31]

For the generalized position and momentum eigenstates and , the value of k is

The existence of mutually unbiased bases in a continuous Hilbert space remains open for debate, as further research in their existence is required before any conclusions can be reached.

Position states and momentum states are eigenvectors of Hermitian operators and , respectively. Weigert and Wilkinson [31] were first to notice that also a linear combination of these operators have eigenbases, which have some features typical for the mutually unbiased bases. An operator has eigenfunctions proportional to with and the corresponding eigenvalues . If we parametrize and as and , the overlap between any eigenstate of the linear combination and any eigenstate of the position operator (both states normalized to the Dirac delta) is constant, but dependent on :

where and stand for eigenfunctions of and .

See also

Related Research Articles

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

Quantum entanglement is the phenomenon that occurs when a duet of particles are generated, interact, or share 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.

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

The uncertainty principle, also known as Heisenberg's indeterminacy principle, is a fundamental concept in quantum mechanics. It states that there is a limit to the precision with which certain pairs of physical properties, such as position and momentum, can be simultaneously known. In other words, the more accurately one property is measured, the less accurately the other property can be known.

<span class="mw-page-title-main">Schrödinger equation</span> Description of a quantum-mechanical system

The Schrödinger equation is a linear partial differential equation that governs the wave function of a quantum-mechanical system. Its discovery was a significant landmark in the development of quantum mechanics. It is named after Erwin Schrödinger, who postulated the equation in 1925 and published it in 1926, forming the basis for the work that resulted in his Nobel Prize in Physics in 1933.

In quantum mechanics, a density matrix is a matrix that describes the quantum state of a physical system. It allows for the calculation of the probabilities of the outcomes of any measurement performed upon this system, using the Born rule. It is a generalization of the more usual state vectors or wavefunctions: while those can only represent pure states, density matrices can also represent mixed states. Mixed states arise in quantum mechanics in two different situations:

  1. when the preparation of the system is not fully known, and thus one must deal with a statistical ensemble of possible preparations, and
  2. when one wants to describe a physical system that is entangled with another, without describing their combined state; this case is typical for a system interacting with some environment.
<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.

In quantum physics, a measurement is the testing or manipulation of a physical system to yield a numerical result. A fundamental feature of quantum theory is that the predictions it makes are probabilistic. The procedure for finding a probability involves combining a quantum state, which mathematically describes a quantum system, with a mathematical representation of the measurement to be performed on that system. The formula for this calculation is known as the Born rule. For example, a quantum particle like an electron can be described by a quantum state that associates to each point in space a complex number called a probability amplitude. Applying the Born rule to these amplitudes gives the probabilities that the electron will be found in one region or another when an experiment is performed to locate it. This is the best the theory can do; it cannot say for certain where the electron will be found. The same quantum state can also be used to make a prediction of how the electron will be moving, if an experiment is performed to measure its momentum instead of its position. The uncertainty principle implies that, whatever the quantum state, the range of predictions for the electron's position and the range of predictions for its momentum cannot both be narrow. Some quantum states imply a near-certain prediction of the result of a position measurement, but the result of a momentum measurement will be highly unpredictable, and vice versa. Furthermore, the fact that nature violates the statistical conditions known as Bell inequalities indicates that the unpredictability of quantum measurement results cannot be explained away as due to ignorance about "local hidden variables" within quantum systems.

<span class="mw-page-title-main">Squeezed coherent state</span> Type of quantum state

In physics, a squeezed coherent state is a quantum state that is usually described by two non-commuting observables having continuous spectra of eigenvalues. Examples are position and momentum of a particle, and the (dimension-less) electric field in the amplitude and in the mode of a light wave. The product of the standard deviations of two such operators obeys the uncertainty principle:

In physics, complementarity is a conceptual aspect of quantum mechanics that Niels Bohr regarded as an essential feature of the theory. The complementarity principle holds that objects have certain pairs of complementary properties which cannot all be observed or measured simultaneously, for examples, position and momentum or wave and particle properties. In contemporary terms, complementarity encompasses both the uncertainty principle and wave-particle duality.

A qutrit is a unit of quantum information that is realized by a 3-level quantum system, that may be in a superposition of three mutually orthogonal quantum states.

<span class="mw-page-title-main">Quantum tomography</span> Reconstruction of quantum states based on measurements

Quantum tomography or quantum state tomography is the process by which a quantum state is reconstructed using measurements on an ensemble of identical quantum states. The source of these states may be any device or system which prepares quantum states either consistently into quantum pure states or otherwise into general mixed states. To be able to uniquely identify the state, the measurements must be tomographically complete. That is, the measured operators must form an operator basis on the Hilbert space of the system, providing all the information about the state. Such a set of observations is sometimes called a quorum. The term tomography was first used in the quantum physics literature in a 1993 paper introducing experimental optical homodyne tomography.

In quantum mechanics, separable states are multipartite quantum states that can be written as a convex combination of product states. Product states are multipartite quantum states that can be written as a tensor product of states in each space. The physical intuition behind these definitions is that product states have no correlation between the different degrees of freedom, while separable states might have correlations, but all such correlations can be explained as due to a classical random variable, as opposed as being due to entanglement.

<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.

<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.

The Born rule is a postulate of quantum mechanics that gives the probability that a measurement of a quantum system will yield a given result. In its simplest form, it states that the probability density of finding a system in a given state, when measured, is proportional to the square of the amplitude of the system's wavefunction at that state. It was formulated and published by German physicist Max Born in July, 1926.

The quantum Heisenberg model, developed by Werner Heisenberg, is a statistical mechanical model used in the study of critical points and phase transitions of magnetic systems, in which the spins of the magnetic systems are treated quantum mechanically. It is related to the prototypical Ising model, where at each site of a lattice, a spin represents a microscopic magnetic dipole to which the magnetic moment is either up or down. Except the coupling between magnetic dipole moments, there is also a multipolar version of Heisenberg model called the multipolar exchange interaction.

In physics, a quantum amplifier is an amplifier that uses quantum mechanical methods to amplify a signal; examples include the active elements of lasers and optical amplifiers.

<span class="mw-page-title-main">SIC-POVM</span> Type of measurement in quantum mechanics

In the context of quantum mechanics and quantum information theory, symmetric, informationally complete, positive operator-valued measures (SIC-POVMs) are a particular type of generalized measurement (POVM). SIC-POVMs are particularly notable thanks to their defining features of (1) being informationally complete; (2)having the minimal number of outcomes compatible with informational completeness, and (3) being highly symmetric. In this context, informational completeness is the property of a POVM of allowing to fully reconstruct input states from measurement data.

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

In quantum mechanics, weak measurements are a type of quantum measurement that results in an observer obtaining very little information about the system on average, but also disturbs the state very little. From Busch's theorem the system is necessarily disturbed by the measurement. In the literature weak measurements are also known as unsharp, fuzzy, dull, noisy, approximate, and gentle measurements. Additionally weak measurements are often confused with the distinct but related concept of the weak value.

References

  1. Schwinger, J. (1960). "Unitary Operator Bases, Harvard University". Proc. Natl. Acad. Sci. U.S.A. 46 (4): 570–9. Bibcode:1960PNAS...46..570S. doi: 10.1073/pnas.46.4.570 . PMC   222876 . PMID   16590645.
  2. Ivanovic, I. D. (1981). "Geometrical description of quantal state determination". J. Phys. A. 14 (12): 3241–3245. Bibcode:1981JPhA...14.3241I. doi:10.1088/0305-4470/14/12/019.
  3. 1 2 M. Planat et al, A Survey of Finite Algebraic Geometrical Structures Underlying Mutually Unbiased Quantum Measurements, http://hal.ccsd.cnrs.fr/docs/00/07/99/18/PDF/MUB_FP.pdf.
  4. Wootters, W. K.; Fields, B. D. (1989). "Optimal State-Determination by Mutually Unbiased Measurements". Ann. Phys. 191 (2): 363–381. Bibcode:1989AnPhy.191..363W. doi:10.1016/0003-4916(89)90322-9. hdl: 10338.dmlcz/141471 .
  5. Gottesman, D. (1996). "Class of quantum error-correcting codes saturating the quantum Hamming bound". Phys. Rev. A. 54 (3): 1862–1868. arXiv: quant-ph/9604038 . Bibcode:1996PhRvA..54.1862G. doi:10.1103/physreva.54.1862. PMID   9913672. S2CID   16407184.
  6. Calderbank, A. R.; et al. (1997). "Quantum Error Correction and Orthogonal Geometry". Phys. Rev. Lett. 78 (3): 405–408. arXiv: quant-ph/9605005 . Bibcode:1997PhRvL..78..405C. doi:10.1103/physrevlett.78.405. S2CID   15326700.
  7. Huang, Yichen (29 July 2010). "Entanglement criteria via concave-function uncertainty relations". Physical Review A. 82 (1): 012335. Bibcode:2010PhRvA..82a2335H. doi:10.1103/PhysRevA.82.012335.
  8. Spengler, C.; Huber, M.; Brierley, S.; Adaktylos, T.; Hiesmayr, B. C. (2012). "Entanglement detection via mutually unbiased bases". Phys. Rev. A. 86 (2): 022311. arXiv: 1202.5058 . Bibcode:2012PhRvA..86b2311S. doi:10.1103/physreva.86.022311. S2CID   34502667.
  9. Vaidman, L.; et al. (1987). "How to ascertain the values of and of a spin-1/2 particle". Phys. Rev. Lett. 58 (14): 1385–1387. Bibcode:1987PhRvL..58.1385V. doi:10.1103/PhysRevLett.58.1385. PMID   10034422.
  10. Englert, B.-G.; Aharonov, Y. (2001). "The mean king's problem: prime degrees of freedom". Phys. Lett. A. 284 (1): 1–5. arXiv: quant-ph/0101134 . Bibcode:2001PhLA..284....1E. doi:10.1016/s0375-9601(01)00271-7. S2CID   14848100.
  11. 1 2 3 4 5 6 Bengtsson, Ingemar (2007). "Three Ways to Look at Mutually Unbiased Bases". AIP Conference Proceedings. Vol. 889. pp. 40–51. arXiv: quant-ph/0610216 . doi:10.1063/1.2713445. S2CID   12395501.
  12. Klappenecker, Andreas; Roetteler, Martin (2003). "Constructions of Mutually Unbiased Bases". arXiv: quant-ph/0309120 . Bibcode:2003quant.ph..9120K.{{cite journal}}: Cite journal requires |journal= (help)
  13. 1 2 Durt, T.; Englert, B.-G.; Bengtsson, I.; Życzkowski, K. (2010). "On mutually unbiased bases". International Journal of Quantum Information. 8 (4): 535–640. arXiv: 1004.3348 . doi:10.1142/s0219749910006502. S2CID   118551747.
  14. P. Butterley, W. Hall "Numerical evidence for the maximum number of mutually unbiased bases in dimension six, 2007, https://arxiv.org/abs/quant-ph/0701122.
  15. Brierley, S.; Weigert, S. (2008). "Maximal sets of mutually unbiased quantum states in dimension six". Phys. Rev. A. 78 (4): 042312. arXiv: 0808.1614 . Bibcode:2008PhRvA..78d2312B. doi:10.1103/physreva.78.042312. S2CID   9295036.
  16. Wootters, William K. (2004). "Quantum measurements and finite geometry". arXiv: quant-ph/0406032 .
  17. Stacey, Blake C. (2016). "SIC-POVMs and Compatibility among Quantum States". Mathematics. 4 (2): 36. arXiv: 1404.3774 . doi: 10.3390/math4020036 .
  18. Bengtsson, Ingemar; Blanchfield, Kate; Cabello, Adán (2012). "A Kochen–Specker inequality from a SIC". Physics Letters A . 376 (4): 374–376. arXiv: 1109.6514 . Bibcode:2012PhLA..376..374B. doi:10.1016/j.physleta.2011.12.011. S2CID   55755390.
  19. Grassl, Markus (2004). "On SIC-POVMs and MUBs in Dimension 6". arXiv: quant-ph/0406175 .
  20. Bengtsson, Ingemar; Życzkowski, Karol (2017). Geometry of quantum states : an introduction to quantum entanglement (Second ed.). Cambridge, United Kingdom: Cambridge University Press. pp. 313–354. ISBN   9781107026254. OCLC   967938939.
  21. 1 2 Bandyopadhyay, Somshubhro; Oscar Boykin, P.; Roychowdhury, Vwani; Vatan, Farrokh (2002). "A new proof for the existence of mutually unbiased bases". Algorithmica. 32 (4): 512–528. arXiv: quant-ph/0103162 . Bibcode:2001quant.ph..3162B. doi:10.1007/s00453-002-0980-7. S2CID   1280557.
  22. Hirschman, I. I. Jr. (1957). "A note on entropy". American Journal of Mathematics. 1957 (1): 152–156. doi:10.2307/2372390. JSTOR   2372390.
  23. H. Maassen, J.B.M. Uffink: Generalized entropic uncertainty relations: Phys. Rev. Lett. 60, 1103–1106 (1988).
  24. Damgaard, Ivan B.; Fehr, Serge; Renner, Renato; Salvail, Louis; Schaffner, Christian (2006). "A Tight High-Order Entropic Quantum Uncertainty Relation with Applications". arXiv: quant-ph/0612014 . Bibcode:2006quant.ph.12014D.{{cite journal}}: Cite journal requires |journal= (help)
  25. Deutsch, D. (1982). "Uncertainty in Quantum Measurements". Physical Review Letters. 50 (9): 631–633. Bibcode:1983PhRvL..50..631D. doi:10.1103/physrevlett.50.631.
  26. Ambainis, Andris (2009). "Limits on entropic uncertainty relations for 3 and more MUBs". arXiv: 0909.3720 .{{cite journal}}: Cite journal requires |journal= (help)
  27. 1 2 3 S. Wehner and A. Winter, 2010 New J. Phys. 12 025009: http://iopscience.iop.org/1367-2630/12/2/025009/.
  28. Wu, S.; Yu, S.; Mølmer, K. (2009). "Entropic uncertainty relation for mutually unbiased bases". Phys. Rev. A. 79 (2): 022104. arXiv: 0811.2298 . Bibcode:2009PhRvA..79b2104W. doi:10.1103/physreva.79.022104. S2CID   119222014.
  29. Ballester, M.; S. Wehner (2007). "Entropic uncertainty relations and locking: tight bounds for mutually unbiased bases" (PDF). Physical Review A. 75 (1): 022319. arXiv: quant-ph/0606244 . Bibcode:2007PhRvA..75a2319C. doi:10.1103/PhysRevA.75.012319. S2CID   41654752.
  30. Wehner, S.; A. Winter (2008). "Higher entropic uncertainty relations for anti-commuting observables". Journal of Mathematical Physics. 49 (6): 062105. arXiv: 0710.1185 . Bibcode:2008JMP....49f2105W. doi:10.1063/1.2943685. S2CID   118268095.
  31. 1 2 Weigert, Stefan; Wilkinson, Michael (2008). "Mutually unbiased bases for continuous variables". Physical Review A. 78 (2): 020303. arXiv: 0802.0394 . Bibcode:2008PhRvA..78b0303W. doi:10.1103/PhysRevA.78.020303. S2CID   67784632.