Next Article in Journal
An Entropy Measure of Non-Stationary Processes
Previous Article in Journal
Applied Thermodynamics: Grain Boundary Segregation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Symmetric Informationally-Complete Quantum States as Analogues to Orthonormal Bases and Minimum-Uncertainty States

1
School of Physics, University of Sydney, Sydney, NSW 2006, Australia
2
University of Waterloo, 200 University Ave West, Waterloo, ON N2L 3G1, Canada
3
Quantum Information Processing Group, Raytheon BBN Technologies, 10 Moulton Street, Cambridge, MA 02138, USA
*
Author to whom correspondence should be addressed.
Entropy 2014, 16(3), 1484-1492; https://doi.org/10.3390/e16031484
Submission received: 21 January 2014 / Revised: 4 February 2014 / Accepted: 5 March 2014 / Published: 14 March 2014

Abstract

: Recently there has been much effort in the quantum information community to prove (or disprove) the existence of symmetric informationally complete (SIC) sets of quantum states in arbitrary finite dimension. This paper strengthens the urgency of this question by showing that if SIC-sets exist: (1) by a natural measure of orthonormality, they are as close to being an orthonormal basis for the space of density operators as possible; and (2) in prime dimensions, the standard construction for complete sets of mutually unbiased bases and Weyl-Heisenberg covariant SIC-sets are intimately related: The latter represent minimum uncertainty states for the former in the sense of Wootters and Sussman. Finally, we contribute to the question of existence by conjecturing a quadratic redundancy in the equations for Weyl-Heisenberg SIC-sets.

1. Introduction

Recently, there has been significant interest in the quantum-information community to prove or disprove the general existence of so-called symmetric informationally complete (SIC) quantum measurements [110]. The question is simple enough: For a Hilbert space d of arbitrary dimension d, is there always a set of d2 quantum states i〉 such that |ψij|2 = constant for all i ≠= j? If so, then the operators E i = 1 d ψ i ψ i can be shown to form the elements of a tomographically complete positive operator-valued measure, often dubbed a SIC-POVM. Interestingly, despite the elementary feel to this question—i.e., it seems the sort of thing one might find as an exercise in a linear-algebra textbook—and the considerable efforts to solve it, the answer remains elusive. One should ask: why is there such interest in the question in the first place? Unfortunately it is hard to deny the impression that much of the motivation stems from little more than the mathematical challenge—not really a good physical reason for months of effort. In this paper, we address the issue of whether there are other, more physical motivations for seeking an answer to the existence of such states.

Indeed, there are already notable motivations from applied physics—e.g., uses of these states in quantum cryptography [1114]. Here, however, we go further by deriving a few previously unobserved properties that single out SIC-sets as particularly relevant to the elementary structure of Hilbert space. These properties make no direct use of the original defining symmetry for SICs, and instead imply them. It is our hope that a better understanding of SIC-sets will lead to new and powerful tool for studying the structure of quantum mechanics [1517].

Specifically, the paper is as follows. We first demonstrate a geometric sense in which SIC-sets of projectors Πi = i〉〈ψi|, if they exist, are as close as possible to being an orthonormal basis on the cone of nonnegative operators. This complements the frame theoretic version of the same question proved by Scott [18]. The two results together clinch the idea that SIC-sets are promising candidates for a stand-in for orthonormal bases on the space of density operators, and we take the opportunity to express the structure of pure states with respect to these preferred bases. Thereafter, we focus on the case of Weyl-Heisenberg (WH) covariant SIC-sets in prime dimensions. In prime dimensions, complete sets of mutually unbiased bases always exist [19] and here we demonstrate a simple expression for them in terms of the WH unitary operators. We then define a notion of minimum uncertainty state with respect to the measurement of these bases [20] and find that the Weyl-Heisenberg SIC-sets, whenever they exist, consist solely of minimum uncertainty states. This provides a strong motivation for considering WH SIC-sets as a special finite-dimensional analogue of coherent states and stresses the further interest of writing the quantum mechanical state space in terms of such operator bases. Finally, we conclude with some general remarks and conjecture a significant reduction in the defining equations for WH SIC-sets.

2. Quasi-Orthonormal Bases and the Space of Density Operators

When equipped with the usual Hilbert-Schmidt inner product 〈A,B〉 = tr(AB), the set of operators acting on a d-dimensional complex vector space becomes a d2-dimensional Hilbert space. Suppose we are given d2 operators Ai normalized so that tr ( A i A i ) = 1. Under what conditions can the Ai be orthogonal to each other? For instance, it is known that one can require the Ai to be Hermitian or unitary and it is not too restrictive for orthogonality to obtain [21]. However, requiring the operators to be positive semi-definite is another matter.

We would like to measure “how orthonormal” a set of positive semi-definite Ai can be. A natural class of measures for this is

K t = i j A , B t = i j ( tr ( A i A j ) ) t

for any real number t ≥ 1. Clearly, one will have an orthonormal basis if and only if Kt = 0 for this sum of N = d4 – d2 terms. But for positive semi-definite Ai, as we will prove, it turns out that

K t d 2 ( d - 1 ) ( d + 1 ) t - 1

In other words, it is impossible to choose an orthonormal basis all of whose elements lie in the positive cone of operators. The natural question to ask is, is there a set of d2 Ai for which Kt achieves the lower bound for t > 1 (for t = 1, any set of Ai/d that forms a POVM would saturate the bound for K1)? If so, we will refer to such a set as a quasi-orthonormal basis.

Note that the linear independence of the Ai does not have to be imposed as a separate requirement. An orthonormal set of vectors is automatically linearly independent. Similarly with quasi-orthonormal sets: requiring Kt to achieve its lower bound forces linear independence. This will follow from one condition for achieving equality—that 〈Ai, Aj〉 = constant when i ≠ = j [2].

We will first prove Equation (2) for the case t = 1, using a special instance of the Cauchy-Schwarz inequality: namely, if λr is any set of n real numbers, n r λ r 2 = ( r 1 2 ) ( r λ r 2 ) ( r λ r ) 2, with equality if and only if λ1 = · · · = λn. Let G be a positive semi-definite operator defined by G = ∑i Ai, and note that trAi ≥ 1 because tr  A i 2 = 1. Applying the Schwarz inequality to the eigenvalues of G, we find tr  G 2 1 d ( tr  G ) 2 d 3 or equivalently, K1d3 – d2. Equality is obtained if and only if G = dI and tr Ai = 1 for all i, i.e., Ai are all rank-1 projectors.

Next, let f(x) = xt, then f(x) is a strictly convex function for t > 1. By writing Kt = ∑i≠j f tr(AiAj)), and applying Jensen inequality, we complete the proof for Equation (2). In Jensen inequality, the equality holds if and only if tr(AiAj) is a constant c for all i ≠= j. One can determine this constant c by taking the trace of both sides of the equation (∑Ai)2 = G2 = d2I, which comes from the equality condition for the previous Cauchy-Schwarz inequality, to get c = ( d 3 - tr  A i 2 ) / ( d 4 - d 2 ) = 1 / ( d + 1 ).

Putting all this together, we conclude that the necessary and sufficient conditions for the Ai to constitute a quasi-orthonormal basis are:

(1)

Each Ai is a rank-1 projector.

(2)

tr ( A i A j ) = 1 d + 1 for all ij.

These are the same as the conditions for the operators Ai to constitute a SIC-set. In particular they imply ∑i Ai = dI and, via [2], that the Ai must be linearly independent.

As an aside, let us point out that this derivation sheds light on the meaning of the “second frame potential” Φ = ∑i,j|ψij|4 introduced by Renes et al. [3] to aid in finding SIC-POVMs numerically. Beforehand its motivation seemed to be only the heuristic of constrained particles distancing themselves from each other because of an interaction potential—an idea that has its roots in Reference [22]. Now, one sees that whenever Ai = i〉〈ψi|, Φ = K2 + d2i.e., Φ is essentially our own orthonormality measure in this case. Equation (2) implies that Φ ≥ 2d3/(d + 1), which was proved by different means in Reference [3].

Returning to the general development, what we have shown is that SIC-sets Πi = i〉〈ψi|, i = 1,. . ., d2, play a special role in the geometry of the cone of positive operators, and by implication, the convex set of quantum states in general—i.e., the density operators. For the purpose of foundational studies—particularly ones of the quantum Bayesian variety [1517,2325]—it is worthwhile recording what this convex set looks like from this perspective.

Imagine introducing E i = 1 d Π i as a canonically given quantum measurement. For a quantum state ρ, the outcomes will occur with probabilities p(i) = tr ρEi. Using the fact that ρ has a unique expansion in terms of the Πi, one can work out that

ρ = i ( ( d + 1 ) p ( i ) - 1 d ) Π i

On the other hand, one might imagine taking the vector of probabilities p(i) as the more basic specification, and the density operator ρ as a convenient, but derivative, specification of the quantum state. Requisite to doing this, one must have an understanding of the allowed probabilities p(i)—not every choice of probabilities in Equation (3) will give rise to positive semi-definite ρ. This can be done conveniently by taking note of the characterization of pure quantum states demonstrated in Reference [26]. Remarkably, it is enough to specify that any Hermitian operator M satisfy only two trace conditions, trM2 = 1 and trM3 = 1, to insure that it be a rank-1 projection operator—i.e., a pure state. One can show that by letting ri be the real eigenvalues of M and noticing that the first condition implies |ri| ≤ 1 for all i, and that tr  M 2 - tr  M 3 = i r i 2 ( 1 - r i ) 0, with equality if and only if one of the ri is 1 and the rest are 0. In terms of the p(i) these two conditions become:

i p ( i ) 2 = 2 d ( d + 1 )
i , j , k c i j k p ( i ) p ( j ) p ( k ) = d + 7 ( d + 1 ) 3

where cijk = Re tr(ΠiΠjΠk). The full set of quantum states is thus the convex hull of the solutions to Equations (4) and (5). One nice thing about Equation (5) is that it makes transparent the algebraic structure lying behind these special probability vectors: For, one can use the coefficients cijk as structure coefficients in defining the anticommutator on the space of operators.

3. Weyl-Heisenberg SIC-Sets and Minimum Uncertainty States

A particularly important class of SIC-sets are those covariant under the action of the Weyl-Heisenberg (WH) group. Most of the SIC-sets constructed to date are of this variety [1,38,10], and these structures have many pleasing properties. In this Section, we derive a few new properties of such sets and show a sense in which their elements can be called minimum uncertainty states when d is prime.

Choose a standard basis |0〉,. . ., |d – 1〉 in d. Let Z|j〉 = ωj |j〉 and X|j〉 = |j + 1〉, where ω = e2πi/d and the addition in |j + 1〉 is modulo d. The WH displacement operators are defined by Dr = τ r1r2Xr1Zr2for r = (r1, r2) and τ = −eiπ/d. A WH covariant SIC-set is constructed by taking a single normalized vector 〉 and acting on it with the unitaries Dr. As r1, r2 range over the integers 0,. . ., d – 1 this gives d2 vectors r〉 = Dr〉. If a 〉 exists such that

ψ D r ψ = { 1 r = 0 e i θ r d + 1 r 0

the r〉〈ψr| form a SIC-set and we call 〉 fiducial.

We can write Equation (6) in a more convenient form by working things out directly in terms of components. We have ψ D r ψ = τ r 1 r 2 j ω j r 2 ψ j + r 1 * ψ j where ψj = 〈j|ψ〉. Consequently, taking a Fourier transform,

1 d r 2 ω k r 2 ψ D r ψ 2 = j ψ j ψ j + k * ψ j + r 1 * ψ j + k + r 1

Replacing |ψ|Dr|2 with (r,0 + 1)/(d + 1) we deduce that 〉 is a fiducial vector if and only if

j ψ j ψ j + k * ψ j + l * ψ j + k + l = 1 d + 1 ( δ k 0 + δ l 0 )

In the equations |ψ|Dr|2 = (r,0 + 1)/(d + 1), the left hand sides are sums of terms quartic in the components, multiplied by powers of ω. In Equation (8) the powers of ω have disappeared; this makes them easier to work with. Similarly, note that the frame potential Φ can be written in an ω-free form. Starting from Equation (7), it follows that

Φ = d 3 k , l | j ψ j ψ j + k * ψ j + l * ψ j + k + l | 2

Finally, let us note the important case of Equation (8) where l = 0. The only terms appearing are then, in terms of probabilities pj = j |2 :

j p j p j + k = 1 d + 1 ( 1 + δ k 0 )

Let us now specialize to prime dimension and work toward expressing the fiduciality condition Equation (8) in terms of probabilities of measurement outcomes for a complete set of d + 1 mutually unbiased bases (MUBs). Assume d > 2 (the d = 2 case requires special treatment, though our conclusions continue to hold).

Consider the matrix F = ( α β γ δ ) where α, β, γ, δ are nonnegative integers < d such that det F = 1 mod d. We will refer to such matrices as symplectic matrices. Let U be the unitary

U = { 1 d j , k τ β - 1 ( α k 2 - 2 j k + δ j 2 ) j k β 0 j τ α γ j 2 α j j β = 0

Then it can be shown [4] that

U D r U = D F r

for all r. Moreover, if U′ is any other unitary with this property, then U′ = eU for some phase e. In these expressions all arithmetical operations on the indices are mod d. In particular β1 is the unique positive integer less than d such that β1β = 1 mod d.

If β ≠= 0, it can be seen that |j|U|k| = d1/2 for all j, k. So we can use symplectic transformations to generate MUBs. For definiteness take V, W to be the unitaries corresponding to ( 1 1 0 1 ) and ( 0 1 - 1 0 ) and define

m , j = { V m j m = 0 , 1 , , d - 1 W j m =

This gives us a full set of d + 1 MUBs labeled by m.

These bases are essentially the only ones obtainable from the standard basis by symplectic transformations. To see this let F be an arbitrary symplectic matrix with β ≠= 0, and let U be the corresponding unitary. We have

F = ( α β γ δ ) = { ( 1 1 0 1 ) β δ - 1 ( δ - 1 0 γ δ ) δ 0 ( 0 1 - 1 0 ) ( β - 1 0 α β ) δ = 0

But, unitaries corresponding to matrices with a 0 in the top right-hand position merely permute and re-phase the elements of the standard basis. So the basis U|0〉,. . ., U|d – 1〉 coincides with one of the bases |m, 0〉,. . . |m, (d – 1)〉 up to permutation and re-phasing. We have |j〉〈j| = (1/d)∑r ω−jr D(0,r). In view of Equations (12) and (13) this means

m , j m , j = { 1 d r ω - j r D ( m r , r ) m 1 d r ω - j r D ( r , 0 ) m =

Now consider an arbitrary 〉, and let pm,j = |m, j|ψ|2. It follows from the above that

j p m , j p m , j + k = { 1 d r ω k r ψ D ( m r , r ) ψ 2 m 1 d r ω k r ψ D ( r , 0 ) ψ 2 m =

from which one sees that 〉 is fiducial if and only if

j p m , j p m , j + k = 1 d + 1 ( 1 + δ k 0 )

Thus, instead of imposing all of the Equation (8) on the components in the standard basis, we can impose just the special case Equation (10) for each MUB separately.

We are now ready to derive the minimum uncertainty property. Consider an arbitrary state 〉 and a measurement of one of the bases m in a complete set of MUBs. This will give rise to outcomes with a probability distribution pm,j. We will quantify the uncertainty in the outcomes by the quadratic Rényi entropy [20,27]

H m = - log 2 ( j p m , j 2 )

This measure and the measure related to it by deleting the logarithm seem to be playing a widening role in quantum information studies [28]. In particular, it is one of the most important measures for quantifying an eavesdropper’s information in quantum cryptography [29].

A minimum uncertainty state is one that minimizes the total uncertainty, T = ∑m Hm. To see the conditions for this, we first appeal to the fact that [28,30]

m , j p m , j 2 = 2

for any 〉. From the convexity of the logarithm, it follows that for any sequence of positive numbers λ1,. . ., λn, (1/n)∑j log2 λj ≤ log2 ((1/n)∑j λj) with equality if and only if λ1 = · · · = λn. Thus, in view of Equation (19)

T ( d + 1 ) log 2 ( d + 1 2 )

with equality if and only if j p m , j 2 = 2 / ( d + 1 ) for all m. Comparing this with Equation (17), one sees that every WH fiducial vector achieves the lower bound and is therefore a minimum uncertainty state. Unfortunately, the theorem does not go the other way: it is not always the case that every minimum uncertainty state is a fiducial state.

4. Discussion

We have argued that the SIC-sets in general dimensions, and more specifically the Weyl-Heisenberg covariant ones (at least in prime dimensions), are particularly interesting structures in Hilbert space geometry. Thus they call out for a better understanding, and we hope this paper is an advertisement for that cause.

There are clearly many more things that can be asked. For instance, are SIC fiducial states also minimum uncertainty states in prime-power dimensions? In non-prime-power dimensions, we face the problem that complete sets of MUBs may not even exist. In that case, can one generalize the definition of a MUB in such a way that SIC fiducial states continue to be minimum uncertainty states? With regard to the property of quasi-orthonormality, is it possible to find non-Weyl-Heisenberg SIC-sets for which Equation (5) takes a nicer form? For instance, if the Πi were actually orthogonal, then one would have tr (ΠiΠjΠk) = δijδjkδki. How close can one come to a form like this with a SIC-set? Many questions like this loom.

But most looming is the question of whether SIC-sets exist in arbitrary dimension. In this regard, let us close with a conjecture. The necessary and sufficient conditions for 〉 to be a Weyl-Heisenberg fiducial vector are captured in Equation (8), which represents d2 simultaneous equations. However, we have noted numerically (up to d = 28) that satisfying roughly 3 2 d of these equations is enough to imply the rest. For instance, it appears to be sufficient to satisfy Equation (8) only for the cases k = 0, 1, 2 with l = 0,... d/2. Perhaps this always holds true.

Acknowledgments

This research was supported in part by the U. S. Office of Naval Research (Grant No. N00014-09-1-0247). H. B. Dang is supported by Vanier Canada Graduate Scholarship and Natural Sciences and Engineering Research Council of Canada.

Conflicts of Interest

The authors declare no conflicts of interest.

  • Author ContributionsAll 3 authors were fully involved at all 3 stages of the research: (1) conception and design, (2) drafting and revising, (3) final approval of the version to be published. The work progressed by a process of ongoing discussions and interactions between the authors.

References

  1. Zauner, G. Quantum Designs: Foundations of a non-commutative Design Theory. Ph.D. Thesis, University of Vienna, Vienna, Austria, 1999. [Google Scholar]
  2. Caves, C.M. Symmetric informationally complete POVMs, Available online: http://info.phys.unm.edu/caves/reports/infopovm.pdf accessed on 18/11/2013.
  3. Renes, J.M.; Blume-Kohout, R.; Scott, A.J.; Caves, C.M. Symmetric informationally complete quantum measurements. J. Math. Phys 2004, 45, 2171–2180. [Google Scholar]
  4. Appleby, D.M. Symmetric informationally complete positive operator valued measures and the extended Clifford group. J. Math. Phys 2005, 46, 052107. [Google Scholar]
  5. Grassl, M. On SIC-POVMs and MUBs in dimension 6 2004. arXiv:quant-ph/0406175.
  6. Wootters, W.K. Quantum Measurements and finite geometry. Found. Phys 2006, 36, 112–126. [Google Scholar]
  7. Klappenecker, A.; Rötteler, M.; Shparlinski, I.E.; Winterhof, A. On approximately symmetric informationally complete positive operator-valued measures and related systems of quantum states. J. Math. Phys 2005, 46, 082104. [Google Scholar]
  8. Flammia, S.T. On SIC-POVMs in prime dimensions. J. Phys. Math. Gen 2006, 39, 13483–13493. [Google Scholar]
  9. Scott, A.J.; Grassl, M. Symmetric informationally complete positive-operator-valued measures: A new computer study. J. Math. Phys 2010, 51, 042203. [Google Scholar]
  10. Appleby, D.M. Symmetric informationally complete measurements of arbitrary rank. Opt. Spectrosc 2007, 103, 416–428. [Google Scholar]
  11. Fuchs, C.A.; Sasaki, M. Squeezing quantum information through a classical channel: Measuring the “quantumness” of a set of quantum states. Quant. Inform. Comput 2003, 3, 377–404. [Google Scholar]
  12. Řeháček, J.; Englert, B.G.; Kaszlikowski, D. Minimal qubit tomography. Phys. Rev. A 2004, 70, 052321. [Google Scholar]
  13. Renes, J.M. Equiangular spherical codes in quantum cryptography. Quant. Inform. Comput 2005, 5, 81–92. [Google Scholar]
  14. Kim, I.H. Quamtumness, Generalized Spherical 2-Design and Symmetric Informationally Complete POVM 2007. arXiv:quant-ph/0608024.
  15. Fuchs, C.A.; Sasaki, M. On the quantumness of a Hilbert space. Quant. Inform. Comput 2004, 4, 467–478. [Google Scholar]
  16. Fuchs, C.A. Quantum Mechanics as Quantum Information (and only a little more) 2002. arXiv:quant-ph/0205039.
  17. Fuchs, C.A.; Schack, R. Quantum-Bayesian coherence. Rev. Mod. Phys 2013, 85, 1693–1715. [Google Scholar]
  18. Scott, A.J. Tight informationally complete quantum measurements. J. Phys. Math. Gen 2006, 39, 13507–13530. [Google Scholar]
  19. Ivonovic, I.D. Geometrical description of quantal state determination. J. Phys. Math. Gen 1981, 14, 3241–3245. [Google Scholar]
  20. Wootters, W.K.; Sussman, D.M. Discrete phase space and minimum-uncertainty states 2007. arXiv:0704.1277 [quant-ph].
  21. Schwinger, J. Unitary operator bases. Proc. Natl. Acad. Sci. USA 1960, 46, 570–579. [Google Scholar]
  22. Benedetto, J.J.; Fickus, M. Finite Normalized Tight Frames. Adv. Comput. Math 2003, 18, 357–385. [Google Scholar]
  23. Caves, C.M.; Fuchs, C.A.; Schack, R. Subjective probability and quantum certainty. Stud. Hist. Philos. Mod. Phys 2007, 38, 255–274. [Google Scholar]
  24. Fuchs, C.A.; Schack, R. Unknown Quantum States and Operations, a Bayesian View; Lecture Notes in Physics; Springer: Berlin, Germany, 2004; pp. 147–187. [Google Scholar]
  25. Appleby, D.M. Facts, Values and Quanta. Found. Phys 2005, 35, 627–668. [Google Scholar]
  26. Jones, N.S.; Linden, N. Parts of quantum states. Phys. Rev. A 2005, 71, 012324. [Google Scholar]
  27. Aczél, J.; Daróczy, Z. On Measures of Information and Their Characterizations; Academic Press: New York, NY, USA, 1975; Volume 115. [Google Scholar]
  28. Brukner, C.; Zeilinger, A. Conceptual inadequacy of the Shannon information in quantum measurements. Phys. Rev. A 2001, 63, 022113. [Google Scholar]
  29. Lutkenhaus, N. Estimates for practical quantum cryptography. Phys. Rev. A 1999. [Google Scholar] [CrossRef]
  30. Larsen, U. Superspace geometry: The exact uncertainty relationship between complementary aspects. J. Phys. Math Gen 1990. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Appleby, D.M.; Dang, H.B.; Fuchs, C.A. Symmetric Informationally-Complete Quantum States as Analogues to Orthonormal Bases and Minimum-Uncertainty States. Entropy 2014, 16, 1484-1492. https://doi.org/10.3390/e16031484

AMA Style

Appleby DM, Dang HB, Fuchs CA. Symmetric Informationally-Complete Quantum States as Analogues to Orthonormal Bases and Minimum-Uncertainty States. Entropy. 2014; 16(3):1484-1492. https://doi.org/10.3390/e16031484

Chicago/Turabian Style

Appleby, D. Marcus, Hoan Bui Dang, and Christopher A. Fuchs. 2014. "Symmetric Informationally-Complete Quantum States as Analogues to Orthonormal Bases and Minimum-Uncertainty States" Entropy 16, no. 3: 1484-1492. https://doi.org/10.3390/e16031484

Article Metrics

Back to TopTop