Skip to main content
Log in

Master Lovas–Andai and equivalent formulas verifying the \(\frac{8}{33}\) two-qubit Hilbert–Schmidt separability probability and companion rational-valued conjectures

  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

We begin by investigating relationships between two forms of Hilbert–Schmidt two-rebit and two-qubit “separability functions”—those recently advanced by Lovas and Andai (J Phys A Math Theor 50(29):295303, 2017), and those earlier presented by Slater (J Phys A 40(47):14279, 2007). In the Lovas–Andai framework, the independent variable \(\varepsilon \in [0,1]\) is the ratio \(\sigma (V)\) of the singular values of the \(2 \times 2\) matrix \(V=D_2^{1/2} D_1^{-1/2}\) formed from the two \(2 \times 2\) diagonal blocks (\(D_1, D_2\)) of a \(4 \times 4\) density matrix \(D= \left||\rho _{ij}\right||\). In the Slater setting, the independent variable \(\mu \) is the diagonal-entry ratio \(\sqrt{\frac{\rho _{11} \rho _ {44}}{\rho _ {22} \rho _ {33}}}\)—with, of central importance, \(\mu =\varepsilon \) or \(\mu =\frac{1}{\varepsilon }\) when both \(D_1\) and \(D_2\) are themselves diagonal. Lovas and Andai established that their two-rebit “separability function” \(\tilde{\chi }_1 (\varepsilon )\) (\(\approx \varepsilon \)) yields the previously conjectured Hilbert–Schmidt separability probability of \(\frac{29}{64}\). We are able, in the Slater framework (using cylindrical algebraic decompositions [CAD] to enforce positivity constraints), to reproduce this result. Further, we newly find its two-qubit, two-quater[nionic]-bit and “two-octo[nionic]-bit” counterparts, \(\tilde{\chi _2}(\varepsilon ) =\frac{1}{3} \varepsilon ^2 \left( 4-\varepsilon ^2\right) \), \(\tilde{\chi _4}(\varepsilon ) =\frac{1}{35} \varepsilon ^4 \left( 15 \varepsilon ^4-64 \varepsilon ^2+84\right) \) and \(\tilde{\chi _8} (\varepsilon )= \frac{1}{1287}\varepsilon ^8 \left( 1155 \varepsilon ^8-7680 \varepsilon ^6+20160 \varepsilon ^4-25088 \varepsilon ^2+12740\right) \). These immediately lead to predictions of Hilbert–Schmidt separability/PPT-probabilities of \(\frac{8}{33}\), \(\frac{26}{323}\) and \(\frac{44482}{4091349}\), in full agreement with those of the “concise formula” (Slater in J Phys A 46:445302, 2013), and, additionally, of a “specialized induced measure” formula. Then, we find a Lovas–Andai “master formula,” \(\tilde{\chi _d}(\varepsilon )= \frac{\varepsilon ^d \Gamma (d+1)^3 \, _3\tilde{F}_2\left( -\frac{d}{2},\frac{d}{2},d;\frac{d}{2}+1,\frac{3 d}{2}+1;\varepsilon ^2\right) }{\Gamma \left( \frac{d}{2}+1\right) ^2}\), encompassing both even and odd values of d. Remarkably, we are able to obtain the \(\tilde{\chi _d}(\varepsilon )\) formulas, \(d=1,2,4\), applicable to full (9-, 15-, 27-) dimensional sets of density matrices, by analyzing (6-, 9, 15-) dimensional sets, with not only diagonal \(D_1\) and \(D_2\), but also an additional pair of nullified entries. Nullification of a further pair still leads to X-matrices, for which a distinctly different, simple Dyson-index phenomenon is noted. C. Koutschan, then, using his HolonomicFunctions program, develops an order-4 recurrence satisfied by the predictions of the several formulas, establishing their equivalence. A two-qubit separability probability of \(1-\frac{256}{27 \pi ^2}\) is obtained based on the operator monotone function \(\sqrt{x}\), with the use of \(\tilde{\chi _2}(\varepsilon )\).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24

Similar content being viewed by others

References

  1. Lovas, A., Andai, A.: Invariance of separability probability over reduced states in \(4 \times 4\) bipartite systems. J. Phys. A: Math. Theor. 50(29), 295303 (2017). http://stacks.iop.org/1751-8121/50/i=29/a=295303

  2. Slater, P.B.: Dyson indices and Hilbert–Schmidt separability functions and probabilities. J. Phys. A 40, 14279 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Slater, P.B.: Extended studies of separability functions and probabilities and the relevance of Dyson indices. J. Geom. Phys. 58, 1101–1123 (2008)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Slater, P.B.: Eigenvalues, separability and absolute separability of two-qubit states. J. Geom. Phys. 59, 17–31 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Slater, P.B.: Ratios of maximal concurrence-parameterized separability functions, and generalized Peres–Horodecki conditions. J. Phys. A 42, 465305 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Dumitriu, I., Edelman, A.: Matrix models for beta ensembles. J. Math. Phys. 43, 5830–5847 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Strzeboński, A.: Cylindrical algebraic decomposition using local projections. J. Symb. Comput. 76, 36–64 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  8. Slater, P.B.: A concise formula for generalized two-qubit Hilbert–Schmidt separability probabilities. J. Phys. A 46, 445302 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Provost, S.B.: Moment-based density approximants. Math. J. 9, 727–756 (2005)

    Google Scholar 

  10. Paule, P., Schorn, M.: A mathematica version of Zeilberger’s algorithm for proving binomial coefficient identities. J. Symb. Comput. 20(5–6), 673–698 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  11. Caves, C.M., Fuchs, C.A., Rungta, P.: Entanglement of formation of an arbitrary state of two rebits. Found. Phys. Lett. 14, 199–212 (2001)

    Article  MathSciNet  Google Scholar 

  12. Fei, J., Joynt, R.: Numerical computations of separability probabilities. Rep. Math. Phys. 78(2), 177–182 (2016), ISSN 0034-4877. http://www.sciencedirect.com/science/article/pii/S0034487716300611

  13. Slater, P.B., Dunkl, C.F.: Moment-based evidence for simple rational-valued Hilbert–Schmidt generic \(2 \times 2\) separability probabilities. J. Phys. A 45, 095305 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Gamel, O.: Entangled Bloch spheres: Bloch matrix and two-qubit state space. Phys. Rev. A 93(6), 062320 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  15. Shang, J., Seah, Y.-L., Ng, H.K., Nott, D.J., Englert, B.-G.: Monte Carlo sampling from the quantum state space. I. New J. Phys. 17(4), 043017 (2015)

    Article  ADS  Google Scholar 

  16. Zhou, D., Chern, G.-W., Fei, J., Joynt, R.: Topology of entanglement evolution of two qubits. Int. J. Mod. Phys. B 26, 1250054 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Khvedelidze, A., Rogojin, I.: On the geometric probability of entangled mixed states. J. Math. Sci. 209, 988–1004 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  18. Slater, P.B.: Octonionic two-qubit separability probability conjectures. arXiv preprint arXiv:1612.02798 (2016)

  19. Slater, P.B., Dunkl, C.F.: Formulas for rational-valued separability probabilities of random induced generalized two-qubit States. Adv. Math. Phys. 2015, 621353 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  20. Yin, X., He, Y., Ling, C., Tian, L., Cheng, X.: Empirical stochastic modeling of multipath polarizations in indoor propagation scenarios. IEEE Trans. Antennas Propag. 63(12), 5799–5811 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  21. Życzkowski, K., Penson, K.A., Nechita, I., Collins, B.: Generating random density matrices. J. Math. Phys. 52, 062201 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Batle, J., Plastino, A.R., Casas, M., Plastino, A.: Understanding quantum entanglement: qubits, rebits and the quaternionic approach. Opt. Spectrosc. 94, 759 (2003)

    Article  Google Scholar 

  23. Singh, R., Kunjwal, R., Simon, R.: Relative volume of separable bipartite states. Phys. Rev. A 89(2), 022308 (2014)

    Article  ADS  Google Scholar 

  24. Milz, S., Strunz, W.T.: Volumes of conditioned bipartite state spaces. J. Phys. A 48, 035306 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Slater, P.B.: Invariance of bipartite separability and PPT-probabilities over Casimir invariants of reduced states. Quantum Inf. Process. 15(9), 3745–3760 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Slater, P.B.: Two-qubit separability probabilities and beta functions. Phys. Rev. A 75, 032326 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  27. Bloore, F.: Geometrical description of the convex sets of states for systems with spin-1/2 and spin-1. J. Phys. A: Math. Gen. 9, 2059 (1976)

    Article  ADS  MathSciNet  Google Scholar 

  28. Andai, A.: Volume of the quantum mechanical state space. J. Phys. A: Math. Gen. 39(44), 13641 (2006)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Bratley, P., Fox, B.L., Niederreiter, H.: Implementation and tests of low-discrepancy sequences. ACM Trans. Model. Comput. Simul. (TOMACS) 2(3), 195–213 (1992)

    Article  MATH  Google Scholar 

  30. Peres, A.: Separability criterion for density matrices. Phys. Rev. Lett. 77(8), 1413 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Horodecki, M., Horodecki, P., Horodecki, R.: Separability of mixed states: necessary and sufficient conditions. Phys. Lett. A 223, 1 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Augusiak, R., Demianowicz, M., Horodecki, P.: Universal observable detecting all two-qubit entanglement and determinant-based separability tests. Phys. Rev. A 77(3), 030301 (2008)

    Article  ADS  Google Scholar 

  33. Blumenson, L.E.: A derivation of n-dimensional spherical coordinates. Am. Math. Mon. 67, 63–66 (1960), ISSN 00029890, 19300972. http://www.jstor.org/stable/2308932

  34. Hildebrand, R.: Semidefinite descriptions of low-dimensional separable matrix cones. Linear Algebra Appl. 429(4), 901–932 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  35. Moore, E.H.: On the determinant of an hermitian matrix of quaternionic elements. Bull. Am. Math. Soc. 28, 161–162 (1922)

    MATH  Google Scholar 

  36. Arnold, B.C., Press, S.J.: Compatible conditional distributions. J. Am. Stat. Assoc. 84(405), 152–156 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  37. Gelman, A., Speed, T.: Characterizing a joint probability distribution by conditionals. J. R. Stat. Soc. B (Methodol) 55, 185–188 (1993)

    MathSciNet  MATH  Google Scholar 

  38. Osipov, V.A., Sommers, H.-J., Życzkowski, K.: Random Bures mixed states and the distribution of their purity. J. Phys. A 43, 055302 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Mittelbach, M., Matthiesen, B., Jorswieck, E.A.: Sampling uniformly from the set of positive definite matrices with trace constraint. IEEE Trans. Signal Process. 60(5), 2167–2179 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  40. Wang, M., Ma, W.: A structure-preserving algorithm for the quaternion Cholesky decomposition. Appl. Math. Comput. 223, 354–361 (2013)

    MathSciNet  MATH  Google Scholar 

  41. Fei, J., Joynt, R.: Numerical computations of separability probabilities. arXiv.1409.1993

  42. Mendonça, P., Marchiolli, M.A., Galetti, D.: Entanglement universality of two-qubit X-states. Ann. Phys. 351, 79–103 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Khvedelidze, A., Torosyan, A.: Spectrum and separability of mixed 2-qubit X-states. arXiv preprint arXiv:1609.06209 (2016)

  44. Dunkl, C.F., Slater, P.B.: Separability probability formulas and their proofs for generalized two-qubit X-matrices endowed with Hilbert–Schmidt and induced measures. Random Matrices: Theory Appl. 4(04), 1550018 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  45. Glöckner, H.: Functions operating on positive semidefinite quaternionic matrices. Monatshefte für Mathematik 132(4), 303–324 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  46. Koutschan, C.: A fast approach to creative telescoping. Math. Comput. Sci. 4(2), 259–266 (2010), ISSN 1661-8289. https://doi.org/10.1007/s11786-010-0055-0

  47. Slater, P.B.: Formulas for generalized two-qubit separability probabilities. arXiv:1609.08561 [quant-ph]

  48. Ozawa, M.: Entanglement measures and the Hilbert–Schmidt distance. Phys. Lett. A 268, 158–160 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  49. Dittmann, J.: Explicit formulae for the Bures metric. J. Phys. A: Math. Gen. 32, 2663 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  50. Šafránek, D.: Discontinuities of the quantum Fisher information and the Bures metric. Phys. Rev. A 95(5), 052320 (2017)

    Article  Google Scholar 

  51. Slater, P. B.: Bloch radii repulsion in separable two-qubit systems. arXiv preprint arXiv:1506.08739 (2015)

  52. Koutschan, C.: Creative Telescoping for Holonomic Functions. Springer Vienna, Vienna, pp. 171–194, (2013) ISBN 978-3-7091-1616-6. https://doi.org/10.1007/978-3-7091-1616-6_7

  53. Życzkowski, K., Sommers, H.-J.: Induced measures in the space of mixed quantum states. J. Phys. A 34, 7111–7125 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  54. Bengtsson, I., Życzkowski, K.: Geometry of Quantum States. Cambridge, Cambridge (2006)

    Book  MATH  Google Scholar 

  55. Slater, P.B.: Eigenvalues, separability and absolute separability of two-qubit states. J. Geom. Phys. 59(1), 17–31 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  56. Hildebrand, R.: Positive partial transpose from spectra. Phys. Rev. A 76, 052325 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  57. Johnston, N.: Non-positive-partial-transpose subspaces can be as large as any entangled subspace. Phys. Rev. A 87(6), 064302 (2013). https://doi.org/10.1103/PhysRevA.87.064302

    Article  ADS  Google Scholar 

  58. Mendonça, P.E., Marchiolli, M.A., Hedemann, S.R.: Maximally entangled mixed states for qubit-qutrit systems. Phys. Rev. A 95(2), 022324 (2017)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

A number of people provided interesting comments in regard to questions posted on the Mathematics, Mathematica, MathOverflow and Physics Stack Exchanges. I discussed the two-quaterbit PPT-probability problem—and other items—extensively with (the always helpful/insightful) Charles Dunkl. Christoph Koutschan, as noted, performed certain calculations laying the foundation for a formal proof that the Lovas–Andai and “concise” formulas yield the same set of results. Christian Krattenthaler also responded to certain queries.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul B. Slater.

Appendices

Appendix A: Absolute separability probabilities

Those separable states that cannot be entangled through unitary operations have been designated as absolutely separable [54, p. 392].

In [55], we reported exact (but now decidedly not rational-valued, and much smaller-valued) formulas for the Hilbert–Schmidt absolute separability probabilities for the two-rebit, two-qubit and two-quaterbit states. For the convenience and interest of the reader, we present them here, while simplifying the forms of the last two.

The two-rebit absolute separability probability is expressible as [55, eq. (32)]

$$\begin{aligned} \frac{6928-2205 \pi }{16 \sqrt{2}} \approx 0.0348338, \end{aligned}$$
(A1)

the two-qubit as [55, eq. (34)]

$$\begin{aligned}&1-\frac{3217542976-5120883075 \pi +16386825840 \tan ^{-1}\left( \sqrt{2}\right) }{32768 \sqrt{2}}-\frac{29901918259}{497664} \nonumber \\&\quad \approx 0.00365826 \end{aligned}$$
(A2)

and the two-quaterbit as [55, eq. (36)]

$$\begin{aligned}&\frac{13}{3043362286338048} \times \Bigg (806338156306739134839776\nonumber \\&\quad -658857590468226345222144 \sqrt{2}+1048604423167357891775325 \sqrt{2} \pi \nonumber \\&\quad -3355534154135545253681040 \sqrt{2} \tan ^{-1}\left( \sqrt{2}\right) \Bigg ) \approx 0.0000401326. \end{aligned}$$
(A3)

Let us note that the integer components of the denominators appearing above are either simply powers of 2, or involve high powers of 2. Also, \( \tan ^{-1}\left( \sqrt{2}\right) ) \approx 0.955317\)—the angle between the space diagonal of a cube and any of its three connecting angles—has been termed the “magic angle” (see the eponymous Wikipedia article). We have also been able to obtain absolute separability probabilities in the two-rebit case, again featuring this particular angle prominently, when the Hilbert–Schmidt (\(k=0\)) measure is replaced by random induced measures [53] for \(k=1,2,3\).

It appears to be a substantial challenge—using eq. (4) of [56] to find the \(6 \times 6\) counterparts to these three formulas for \(4 \times 4\) systems.

Appendix B: Rebit-retrit and qubit-qutrit analyses

Let us now attempt to extend the two-rebit and two-qubit line of analysis above to rebit-retrit and qubit-qutrit settings—now, of course, passing from consideration of \(4 \times 4\) density matrices to \(6 \times 6\) ones. Lovas and Andai, in their quite recent study, had not yet addressed such issues. In [2], candidate (Slater-type) separability functions had been proposed. Two dependent variables (cf. the use of \(\mu =\sqrt{\frac{\rho _ {11} \rho _ {44}}{\rho _ {22} \rho _ {33}}}\) in the lower-dimensional setting above) had been employed [2, eq. (44)]. Let us now refer to these two variables as \(\tau _1=\sqrt{\frac{\rho _ {11} \rho _{55}}{\rho _ {22} \rho _ {44}}}\) and \(\tau _2 =\sqrt{\frac{\rho _ {22} \rho _{66}}{\rho _ {33} d _{55}}}\). But, interestingly, it was argued that only a single dependent variable \(\tau = \tau _1 \tau _2 = \sqrt{\frac{\rho _ {11} \rho _{66}}{\rho _ {33} \rho _ {44}}}\) sufficed for modeling the corresponding separability functions. The separability function in the rebit-retrit case was proposed to be simply proportional to \(\tau \) [2, eq.(98)].

In our effort to extend the Lovas–Andai analyses [1] to this setting, we now took \(D_1\) and \(D_2\) to equal the upper and lower diagonal \(3 \times 3\) blocks of the \(6 \times 6\) density matrix in question. Then, we computed the three singular values (\(s_1 \ge s_2 \ge s_3\)) of \(D_2^{1/2} D_1^{-1/2}\), and took the ratio variables \(\varepsilon _1 = \frac{s_2}{s_1}\) and \(\varepsilon _2 = \frac{s_3}{s_2}\) as the dependent ones in question. (An issue of possible concern is that, unlike the \(4 \times 4\) case [32], positivity of the determinant of the partial transpose of a \(6 \times 6\) density matrix is only a necessary, but not sufficient condition for separability.) Also, in the case of diagonal D, again the two variables in the \(\mu \) framework are equal to those in the \(\varepsilon \) setting, or to their reciprocals.

Then, we generated 3436 million rebit-retrit and 2379 million qubit-qutrit density matrices, randomly with respect to Hilbert–Schmidt measure. (These sizes are much larger than those employed in 2007—for similar purposes—in [2].) We appraised the separability of the density matrices D by testing whether the partial transpose, using the four \(3 \times 3\) blocks, had all its six eigenvalues positive. The separability probability estimates were \(0.13180011 \pm 0.0000113109\) and \(0.02785302 \pm 6.6124281 \cdot 10^{-6}\), respectively. (We can reject the qubit-qutrit conjecture of \(\frac{32}{1199} \approx 0.0266889\) advanced in [2, sec. 10.2]. A possible alternative candidate is \(\frac{72}{2585} =\frac{5 \cdot 11 \cdot 47}{2^3 \cdot 3^2} \approx 0.027853\), while in the rebit-retrit case, we have \(\frac{298}{2261} =\frac{7 \cdot 17 \cdot 19}{2 \cdot 149} \approx 0.1318001\).) Further, our estimates of the probabilities that D had two (the most possible [57]) negative eigenvalues, and hence a positive determinant, although being entangled, were \(0.0334197 \pm 0.0000409506\) in the rebit-retrit case and \(0.0103211 \pm 0.000031321\) in the qubit-qutrit instance.)

In the two-variable settings, we partition the square \([0,1]^2\) of possible separability probability results into an \(80 \times 80\) grid, and in the one-variable setting, use a partitioning (as in the two-rebit and two-qubit analyses above) into 200 subintervals of [0,1]. In Fig. 25 we show the ratio of the square of the rebit-retrit separability probability to the qubit-qutrit separability probability as a function of \(\tau \), while in Fig. 26, we show a two-dimensional version. Figure 27 is the analog of this last plot using the singular-value ratios \(\varepsilon _1\) and \(\varepsilon _2\). As in Figs. 17, 18 and 19, we observe a gradual increase in these Dyson-index-oriented analyses. In Figs. 28 and 29, we show the highly linear (“diagonal”) rebit-retrit and qubit-qutrit separability probabilities, holding \(\tau _1=\tau _2\).

Fig. 25
figure 25

The ratio of the square of the rebit-retrit separability probability to the qubit-qutrit separability probability as a function of \(\tau = \tau _1 \tau _2 = \sqrt{\frac{\rho _ {11} \rho _{66}}{\rho _ {33} \rho _ {44}}}\)

Fig. 26
figure 26

The ratio of the square of the rebit-retrit separability probability to the qubit-qutrit separability probability as a function of \(\tau _1=\sqrt{\frac{\rho _ {11} \rho _{55}}{\rho _ {22} \rho _ {44}}}\) and \(\tau _2 =\sqrt{\frac{\rho _ {22} d _{66}}{\rho _ {33} d _{55}}}\)

Fig. 27
figure 27

The ratio of the square of the rebit-retrit separability probability to the qubit-qutrit separability probability as a function of the singular-value ratios \(\varepsilon _1\) and \(\varepsilon _2\)

Fig. 28
figure 28

Rebit-retrit separability probabilities for \(\tau _1=\tau _2\)

Fig. 29
figure 29

Qubit-qutrit separability probabilities for \(\tau _1=\tau _2\)

Let us note that Mendonça, and Marchiolli, and Hedemann have recently shown [58, App. A] that for qubit-qutrit X-states, the partial transposes can—in contrast to more general such \(6 \times 6\) systems—have no more than one negative eigenvalue. Therefore, positivity of the determinant of the partial transpose is both necessary and sufficient for separability, in this case. Nevertheless, Dunkl has been able to conclude that the Hilbert–Schmidt separability probabilities reported in [44] for two-qubit X-states, continue to hold in these higher-dimensional qubit-qutrit X-state systems.

Appendix C: Comparison of Ginibre and Cholesky methods for quaternion positive-definite matrices—by C. F. Dunkl

The calculations depend on integrating monomials over the unit sphere in \(\mathbb {R}^{N}\). We use the Pochhammer symbol \(\left( a\right) _{n} :=\prod _{i=1}^{n}\left( a+i-1\right) \). If \(a\ne 0,-1,-2,\ldots \) then \(\Gamma \left( a+n\right) /\Gamma \left( a\right) =\left( a\right) _{n}\).

Lemma C.1

Let \(S_{N-1}\) be the unit sphere in \(\mathbb {R}^{N}\) with the inherited rotation-invariant measure \(\hbox {d}\omega \) and let \(n_{1},n_{2} ,\ldots ,n_{N}\in \mathbb {N}_{0}\) (\(\left\{ 0,1,2,3\ldots \right\} \)) then

$$\begin{aligned} \int _{S_{N-1}}\prod _{i=1}^{N}\left| x_{i}\right| ^{n_{i}} \hbox {d}\omega \left( x\right) =2\prod _{i=1}^{N}\Gamma \left( \frac{n_{i}+1}{2}\right) \Bigg /\Gamma \left( \frac{1}{2}\left( \sum _{i=1}^{N}n_{i}+N\right) \right) . \end{aligned}$$

Proof

Let \(n=\sum _{i=1}^{N}n_{i}\) and \(f\left( x\right) =\prod _{i=1}^{N}\left| x_{i}\right| ^{n_{i}}\). In spherical polar coordinates

$$\begin{aligned} \int _{\mathbb {R}^{N}}f\left( x\right) \exp \left( -\frac{\left| x\right| ^{2}}{2}\right) \hbox {d}x=\int _{0}^{\infty }r^{n}\exp \left( -\frac{r^{2}}{2}\right) r^{N-1}\hbox {d}r\int _{S_{N-1}}f\left( x\right) \hbox {d}\omega \left( x\right) . \end{aligned}$$

The left-hand side equals (by the substitution \(x_{i}^{2}=2t\))

$$\begin{aligned} \prod _{i=1}^{N}\int _{-\infty }^{\infty }\left| x_{i}\right| ^{n_{i} }{\text {e}}^{-\left| x_{i}\right| ^{2}/2}\hbox {d}x_{i}=\prod _{i=1}^{N}2^{\left( n_{i}+1\right) /2}\Gamma \left( \frac{n_{i}+1}{2}\right) , \end{aligned}$$

and

$$\begin{aligned} \int _{0}^{\infty }r^{n}\exp \left( -\frac{r^{2}}{2}\right) r^{N-1} \hbox {d}r=2^{\left( N+n\right) /2-1}\Gamma \left( \frac{n+N}{2}\right) . \end{aligned}$$

Divide the left side by this to obtain \(\int _{S}f\left( x\right) \hbox {d}\omega \left( x\right) \). \(\square \)

Denote the right-hand side by \(I\left[ n_{1},n_{2},\ldots ,n_{N}\right] \). We use \(0\$n\) to denote 0 listed n times. Thus the surface measure of \(S_{N-1}\) is \(I[0\$N]\). We need another lemma for integrating powers of sums of squares.

Lemma C.2

Suppose \(1\le A\le B\) and \(n=1,2,3,\ldots \) then the normalized integral

$$\begin{aligned} \int _{S_{B-1}}\left( \sum _{i=1}^{A}y_{i}^{2}\right) ^{n}\hbox {d}\omega \left( y\right) \left\{ \int _{S_{B-1}}\hbox {d}\omega \left( y\right) \right\} ^{-1}=\dfrac{\left( A/2\right) _{n}}{\left( B/2\right) _{n}}. \end{aligned}$$

Proof

The argument is similar to the proof of Lemma C.1. An alternative approach would rely on Dirichlet integrals \(\square \)

The Cholesky method begins with a random point on \(S_{27}\subset \mathbb {R}^{28}\) to form an upper triangular matrix A such that \(A_{11},A_{22},A_{33},A_{44}\ge 0\) and \(A_{ij}\in \mathbb {H}\) for \(1\le i<j\le 4\). Then \(Q:=A^{*}A\) is positive-definite and \(trQ=1\). In particular \(Q_{11}=A_{11}^{2}\) and \(\det Q=\left( A_{11}A_{22}A_{33}A_{44}\right) ^{2} \). The Jacobian is \(A_{11}^{13}A_{22}^{9}A_{33}^{5}A_{44}\). For the measure \(\left( \det Q\right) ^{k}\) the \(n^{th}\) moment of \(Q_{11}\) (that is \(\mathcal {E}\left( Q_{11}^{n}\right) \)) is given by

$$\begin{aligned} \mu _{n}&=\frac{I\left[ 2n+2k+13,2k+9,2k+5,2k+1,0\$24\right] }{I[2k+13,2k+9,2k+5,2k+1,0\$24]}\\&=\frac{\Gamma \left( k+7+n\right) \Gamma \left( 4k+28\right) }{\Gamma \left( k+7\right) \Gamma \left( 4k+28+n\right) }=\frac{\left( k+7\right) _{n}}{\left( 4k+28\right) _{n}},\\ \mu _{1}&=\frac{1}{4},~\mu _{2}=\frac{k+8}{4\left( 4k+29\right) },\\ \mu _{3}&=\frac{\left( k+8\right) \left( k+9\right) }{8\left( 4k+29\right) \left( 2k+15\right) }. \end{aligned}$$

The Ginibre method for \(M\times 4\) begins with a random point on \(S_{16M-1}\subset \mathbb {R}^{16M}\) to form an \(M\times 4\) matrix H with \(H_{ij} \in \mathbb {H}\) for \(1\le i\le M,1\le j\le 4\) and \(\sum _{i=1}^{M}\sum _{j=1}^{4}\left| H_{ij}\right| ^{2}=1\). (For a quaternion \(q=x_{1}+x_{2}\varvec{i}+x_{3}\varvec{j}+x_{4}\varvec{k}\) define \(\overline{q}=x_{1}-x_{2}\varvec{i}-x_{3}\varvec{j}-x_{4} \varvec{k}\) then \(\left| q\right| ^{2}=\overline{q}q=\sum _{i=1}^{4}x_{i}^{2}\).) Then \(Q:=H^{*}H\) is positive-definite and \(trQ=1\). In particular \(Q_{11}=\sum _{i=1}^{M}\left( A^{*}\right) _{1i}A_{i1} =\sum _{i=1}^{M}\overline{A_{i1}}A_{i1}=\sum _{i=1}^{M}\left| A_{i1} \right| ^{2}\). In real terms each \(\left| A_{i1}\right| ^{2}\) is a sum of four squared real variables so rewrite \(Q_{11}=\sum _{j=1}^{4M}x_{j} ^{2}\) where \(\left( x_{j}\right) _{j=1}^{16M}\) is a random point on \(S_{16M-1}\). Apply Lemma C.2 with \(A=4M\) and \(B=16M\) to obtain

$$\begin{aligned} \nu _{n}=\mathcal {E}\left( Q_{11}^{n}\right) =\mathcal {E}\left( \sum _{j=1}^{4M}x_{j}^{2}\right) ^{n}=\frac{\left( 2M\right) _{n}}{\left( 8M\right) _{n}} \end{aligned}$$

for \(n=1,2,3,\ldots \). In particular \(\nu _{1}=\frac{1}{4}\) and \(\nu _{2} =\dfrac{2M+1}{4\left( 8M+1\right) }\).

Thus \(\nu _{n}=\dfrac{\left( 2M\right) _{n}}{\left( 8M\right) _{n}}\) and \(\mu _{n}=\dfrac{\left( k+7\right) _{n}}{\left( 4k+28\right) _{n}}\) are equal for all n exactly when \(k+7=2M\). In itself this is not a proof that the Ginibre method produces \(\left( \det Q\right) ^{2M-7}\) times the HS measure. This statement is a consequence of equation (4.6) in [53].

In particular the Ginibre method does not lead to the HS measure for any M since k is necessarily odd.

The above calculations can be adapted to other values of the parameter \(\alpha \) (with \(\alpha =\frac{1}{2}\) for \(\mathbb {R}\), \(\alpha =1\) for \(\mathbb {C}\), and \(\alpha =2\) for \(\mathbb {H}\)). The Cholesky method starts with the sphere in \(\mathbb {R}^{4+12\alpha }\) (4 on diagonal, \(12\alpha \) off-diagonal), and the Jacobian is \(A_{11}^{1+6\alpha }A_{22}^{1+4\alpha } A_{33}^{1+2\alpha }A_{44}\). The nth moment of \(Q_{11}\) is

$$\begin{aligned} \frac{I\left[ 2n+1+6\alpha +2k,1+4\alpha +2k,1+2\alpha +2k,1+2k,0\$12\alpha \right] }{I\left[ 1+6\alpha +2k,1+4\alpha +2k,1+2\alpha +2k,1+2k,0\$12\alpha \right] }=\frac{\left( k+3\alpha +1\right) _{n}}{\left( 4k+12\alpha +4\right) _{n}}. \end{aligned}$$

Similarly to above the Ginibre matrix size \(M\times 4\) comes from a random point on \(S_{8\alpha M-1}\subset \mathbb {R}^{8\alpha M}\) and \(Q_{11}\) is the sum of \(2\alpha M\) squares \(x_{i}^{2}\) and the nth moment of \(Q_{11}\) is

$$\begin{aligned} \frac{\left( \alpha M\right) _{n}}{\left( 4\alpha M\right) _{n}}, \end{aligned}$$

which agrees with the \(\left( \det Q\right) ^{k}\times HS\) when \(k=\alpha M-3\alpha -1=\left( M-3\right) \alpha -1\). To apply the formula:

$$\begin{aligned} \alpha&=\frac{1}{2},k=\frac{1}{2}(M-5),\\ \alpha&=1,k=M-4,\\ \alpha&=2,k=2M-7. \end{aligned}$$

Thus the Ginibre method does produce HS random density matrices with \(M=5\) for \(\mathbb {R}\) and \(M=4\) for \(\mathbb {C}\).

Appendix D: A further formula for the Lovas–Andai integral—by Charles Dunkl

We deal here with the evaluation of the integral for even d.

$$\begin{aligned} \int _{0}^{1}t^{-3\left( 1+d\right) }~_{2}F_{1}\left( \genfrac{}{}{0.0pt}{}{2+3d,2+3d}{4+6d} ;1-\frac{1}{t^{2}}\right) \left( 1-t^{2}\right) ^{d}~_{3}F_{2}\left( \genfrac{}{}{0.0pt}{}{-\frac{d}{2},\frac{d}{2},d}{1+\frac{d}{2},1+\frac{3d}{2}} ;t^{2}\right) \mathrm {d}t \end{aligned}$$
(D1)

The first step is to rewrite the \(_{2}F_{1}\) series as a power series in \(t^{2}\). By the use of the identity

$$\begin{aligned} _{2}F_{1}\left( \genfrac{}{}{0.0pt}{}{a,b}{c} ;x\right) =\left( 1-x\right) ^{-a}~_{2}F_{1}\left( \genfrac{}{}{0.0pt}{}{a,c-b}{c} ;\frac{x}{x-1}\right) \end{aligned}$$

we obtain

$$\begin{aligned} _{2}F_{1}\left( \genfrac{}{}{0.0pt}{}{2+3d,2+3d}{4+6d} ;1-\frac{1}{t^{2}}\right) =t^{2\left( 2+3d\right) }~_{2}F_{1}\left( \genfrac{}{}{0.0pt}{}{2+3d,2+3d}{4+6d} ;1-t^{2}\right) \end{aligned}$$

Thus the desired integral equals

$$\begin{aligned} \int _{0}^{1}t^{1+3d}~_{2}F_{1}\left( \genfrac{}{}{0.0pt}{}{2+3d,2+3d}{4+6d} ;1-t^{2}\right) \left( 1-t^{2}\right) ^{d}~_{3}F_{2}\left( \genfrac{}{}{0.0pt}{}{-\frac{d}{2},\frac{d}{2},d}{1+\frac{d}{2},1+\frac{3d}{2}} ;t^{2}\right) \mathrm {d}t \end{aligned}$$

By expanding the series we can integrate term by term. The typical term is

$$\begin{aligned} \int _{0}^{1}t^{1+3d+2j}\left( 1-t^{2}\right) ^{d+n}\mathrm {d}t&=\frac{1}{2}\int _{0}^{1}s^{j+3d/2}\left( 1-s\right) ^{d+n}\mathrm {d}s\\&=\frac{1}{2}B\left( j+\frac{3d}{2}+1,d+n+1\right) \\&=\frac{1}{2} \frac{\Gamma \left( j+\frac{3d}{2}+1\right) \Gamma \left( d+n+1\right) }{\Gamma \left( j+\frac{5d}{2}+2+n\right) }\\&=\frac{1}{2}\frac{\Gamma \left( \frac{3d}{2}+1\right) \Gamma \left( d+1\right) }{\Gamma \left( \frac{5d}{2}+2\right) }\frac{\left( d+1\right) _{n}}{\left( j+\frac{5d}{2}+2\right) _{n}}\frac{\left( \frac{3d}{2}+1\right) _{j}}{\left( \frac{5d}{2}+2\right) _{j}}, \end{aligned}$$

by the use of the change of variable \(s=t^{2}\) and the beta function. The result for integral (D1) is (with d even)

$$\begin{aligned} \frac{1}{2}\frac{\Gamma \left( \frac{3d}{2}+1\right) \Gamma \left( d+1\right) }{\Gamma \left( \frac{5d}{2}+2\right) }\sum _{j=0}^{d/2} \frac{\left( -\frac{d}{2}\right) _{j}\left( \frac{d}{2}\right) _{j}\left( d\right) _{j}}{\left( 1+\frac{d}{2}\right) _{j}\left( 2+\frac{5d}{2}\right) _{j}j!}~_{3}F_{2}\left( \genfrac{}{}{0.0pt}{}{2+3d,2+3d,d+1}{4+6d,j+\frac{5d}{2}+2} ;1\right) . \end{aligned}$$
(D2)

By the use of the Gauss sum and contiguous hypergeometric series we can produce finite expressions for the integral. We state the Gauss sum (with \(\dot{c}>a+b\)) and define utility functions.

$$\begin{aligned} S\left( a,b,c\right)&:=\frac{\Gamma \left( c-a-b\right) \Gamma \left( c\right) }{\Gamma \left( c-a\right) \Gamma \left( c-b\right) },\\ _{2}F_{1}\left( \genfrac{}{}{0.0pt}{}{a,b}{c} ;1\right)&=S\left( a,b,c\right) . \end{aligned}$$

For \(k=0,1,2\ldots \) and \(k<\min \left( c-a-b,g\right) \) define

$$\begin{aligned} S_{32}\left( a,b,c;g,k\right) :=~_{3}F_{2}\left( \genfrac{}{}{0.0pt}{}{a,b,g}{c,g-k} ;1\right) . \end{aligned}$$

Proposition D.1

For \(k=0,1,2\ldots \) and \(k<\min \left( c-a-b,g\right) \)

$$\begin{aligned} S_{32}\left( a,b,c;g,k\right) =\frac{\Gamma \left( c-a-b\right) \Gamma \left( c\right) }{\Gamma \left( c-a\right) \Gamma \left( c-b\right) }\sum _{j=0}^{k}\frac{\left( -k\right) _{j}\left( a\right) _{j}\left( b\right) _{j}}{j!\left( 1+a+b-c\right) _{j}\left( g-k\right) _{j}}. \end{aligned}$$

Proof

For any n there is the formula (easy to verify, by finite differences, or the Chu–Vandermonde sum)

$$\begin{aligned} \frac{\left( g+n-k\right) _{k}}{\left( g-k\right) _{k}}=\sum _{j=0} ^{k}\left( {\begin{array}{c}k\\ j\end{array}}\right) \frac{n\left( n-1\right) \left( n-2\right) \cdots \left( n-j+1\right) }{\left( g-k\right) _{j}}. \end{aligned}$$

Then

$$\begin{aligned} ~_{3}F_{2}\left( \genfrac{}{}{0.0pt}{}{a,b,g}{c,g-k} ;1\right) =\sum _{n=0}^{\infty }\frac{\left( a\right) _{n}\left( b\right) _{n}}{\left( c\right) _{n}n!}\frac{\left( g\right) _{n}}{\left( g-k\right) _{n}} \end{aligned}$$

and

$$\begin{aligned} \frac{\left( g\right) _{n}}{\left( g-k\right) _{n}}=\frac{\left( g\right) _{n-k}\left( g+n-k\right) _{k}}{\left( g-k\right) _{k}\left( g\right) _{n-k}}=\frac{\left( g+n-k\right) _{k}}{\left( g-k\right) _{k}}. \end{aligned}$$

Also

$$\begin{aligned}&\sum _{n=0}^{\infty }\frac{\left( a\right) _{n}\left( b\right) _{n}}{\left( c\right) _{n}n!}n\left( n-1\right) \left( n-2\right) \cdots \left( n-j+1\right) =\sum _{n=j}^{\infty }\frac{\left( a\right) _{n}\left( b\right) _{n}}{\left( c\right) _{n}\left( n-j\right) !}\\&\quad =\frac{\left( a\right) _{j}\left( b\right) _{j}}{\left( c\right) _{j} }\sum _{m=0}^{\infty }\frac{\left( a+j\right) _{m}\left( b+j\right) _{m} }{\left( c+j\right) _{m}m!}=\frac{\left( a\right) _{j}\left( b\right) _{j}}{\left( c\right) _{j}}S\left( a+j,b+j,c+j\right) . \end{aligned}$$

making the change of index \(m=n-j\) (observe that \(n\cdots \left( n-j+1\right) =0\) for \(0\le n\le j-1\)). The S-term equals

$$\begin{aligned} \frac{\Gamma \left( c-a-b-j\right) \Gamma \left( c+j\right) }{\Gamma \left( c-a\right) \Gamma \left( c-b\right) }=\frac{\left( c\right) _{j}}{\left( c-a-b-j\right) _{j}}\frac{\Gamma \left( c-a-b\right) \Gamma \left( c\right) }{\Gamma \left( c-a\right) \Gamma \left( c-b\right) } \end{aligned}$$

by the use of the relation \(\Gamma \left( t\right) \left( t\right) _{j} =\Gamma \left( t+j\right) \); also by reversal \(\left( c-a-b-j\right) _{j}=\left( -1\right) ^{j}\left( 1+a+b-c\right) _{j}\). The binomial coefficient \(\left( {\begin{array}{c}k\\ j\end{array}}\right) =\left( -1\right) ^{j}\frac{\left( -k\right) _{j}}{j!}\). Combine the ingredients and this proves the formula. \(\square \)

Consider the typical term in (D2)

$$\begin{aligned} _{3}F_{2}\left( \genfrac{}{}{0.0pt}{}{2+3d,2+3d,d+1}{4+6d,j+\frac{5d}{2}+2} ;1\right) . \end{aligned}$$

Set \(a=2+3d,b=1+d,c=4+6d,g=2+3d\), \(k=\frac{d}{2}-j\). Thus desired integral (D1) equals

$$\begin{aligned}&\frac{1}{2}\frac{\Gamma \left( \frac{3d}{2}+1\right) \Gamma \left( d+1\right) }{\Gamma \left( \frac{5d}{2}+2\right) }\\&\quad \times \sum _{j=0}^{d/2} \frac{\left( -\frac{d}{2}\right) _{j}\left( \frac{d}{2}\right) _{j}\left( d\right) _{j}}{\left( 1+\frac{d}{2}\right) _{j}\left( 2+\frac{5d}{2}\right) _{j}j!}S_{32}\left( 2+3d,1+d,4+6d;2+3d,\frac{d}{2}-j\right) \\&=\frac{1}{2}\frac{\Gamma \left( \frac{3d}{2}+1\right) \Gamma \left( d+1\right) }{\Gamma \left( \frac{5d}{2}+2\right) }\frac{\Gamma \left( 4+6d\right) \Gamma \left( 1+2d\right) }{\Gamma \left( 2+3d\right) \Gamma \left( 3+5d\right) }\\&\quad \times \sum _{j=0}^{d/2}\frac{\left( -\frac{d}{2}\right) _{j}\left( \frac{d}{2}\right) _{j}\left( d\right) _{j}}{\left( 1+\frac{d}{2}\right) _{j}\left( 2+\frac{5d}{2}\right) _{j}j!}\sum _{i=0}^{d/2-j}\frac{\left( j-\frac{d}{2}\right) _{i}\left( 2+3d\right) _{i}\left( 1+d\right) _{i} }{i!\left( -2d\right) _{i}\left( 2+\frac{5d}{2}+j\right) _{i}}. \end{aligned}$$

The double sum in the last line can be rewritten as

$$\begin{aligned} \sum _{i\ge 0,j\ge 0}^{i+j\le d/2}\frac{\left( -\frac{d}{2}\right) _{i+j}\left( \frac{d}{2}\right) _{j}\left( d\right) _{j}\left( 2+3d\right) _{i}\left( 1+d\right) _{i}}{\left( 2+\frac{5d}{2}\right) _{i+j}\left( 1+\frac{d}{2}\right) _{j}i!j!\left( -2d\right) _{i}}. \end{aligned}$$

Collect the (independent of t) prefactors [from the product of (82) and (83)]

$$\begin{aligned} 18\frac{\Gamma \left( \frac{1}{2}+\frac{d}{2}\right) ^{3}\Gamma \left( \frac{7}{6}+\frac{d}{2}\right) ^{2}\Gamma \left( \frac{5}{6}+\frac{d}{2}\right) ^{2}\Gamma \left( 1+2d\right) }{\pi ^{7/2}\Gamma \left( 1+\frac{d}{2}\right) \Gamma \left( 3+5d\right) }3456^{d}. \end{aligned}$$

Evaluate for even d, \(\Gamma \left( \frac{1}{2}+\frac{d}{2}\right) =\Gamma \left( \frac{1}{2}\right) \left( \frac{1}{2}\right) _{d/2}\) (and \(\Gamma \left( \frac{1}{2}\right) =\sqrt{\pi }\));

$$\begin{aligned} \Gamma \left( \frac{7}{6}+\frac{d}{2}\right) \Gamma \left( \frac{5}{6} +\frac{d}{2}\right)&=\left( \frac{7}{6}\right) _{d/2}\left( \frac{5}{6}\right) _{d/2}\Gamma \left( \frac{7}{6}\right) \Gamma \left( \frac{5}{6}\right) \\&=\left( \frac{7}{6}\right) _{d/2}\left( \frac{5}{6}\right) _{d/2} \frac{1}{6}\Gamma \left( \frac{1}{6}\right) \Gamma \left( \frac{5}{6}\right) , \end{aligned}$$

and \(\Gamma \left( \frac{1}{6}\right) \Gamma \left( \frac{5}{6}\right) =\dfrac{\pi }{\sin \frac{\pi }{6}}=2\pi \) (recall \(\Gamma \left( t\right) \Gamma \left( 1-t\right) =\dfrac{\pi }{\sin \pi t}\)). Put it all together (even d)

$$\begin{aligned}&\mathcal {P}_{\mathrm {sep/PPT}}\left( d\right) \\&\quad =3456^{d}\frac{\left( \frac{1}{2}\right) _{d/2} ^{3}\left( \frac{7}{6}\right) _{d/2}^{2}\left( \frac{5}{6}\right) _{d/2}^{2}\left( 2d\right) !}{\left( \frac{d}{2}\right) !\left( 3\right) _{5d}}\sum _{i\ge 0,j\ge 0}^{i+j\le d/2}\frac{\left( -\frac{d}{2}\right) _{i+j}\left( \frac{d}{2}\right) _{j}\left( d\right) _{j}\left( 2+3d\right) _{i}\left( 1+d\right) _{i}}{\left( 2+\frac{5d}{2}\right) _{i+j}\left( 1+\frac{d}{2}\right) _{j}i!j!\left( -2d\right) _{i}}. \end{aligned}$$

Appendix E: Remark on Lovas–Andai paper

It certainly appears that the work of Lovas and Andai [1]—inspired by that of Milz and Strunz [24]—is highly innovative and successful in finding the two-rebit separability function \(\tilde{\chi }_1(\varepsilon )\), and verifying the conjecture that the two-rebit Hilbert–Schmidt separability probability is \(\frac{29}{64}\). However, in our study of the Lovas–Andai paper, we remain unconvinced by the chain of arguments on page 13 leading to the result \(\frac{1}{4}\) and have posted a stack exchange question (https://mathematica.stackexchange.com/q/144277/29989) in this regard.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Slater, P.B. Master Lovas–Andai and equivalent formulas verifying the \(\frac{8}{33}\) two-qubit Hilbert–Schmidt separability probability and companion rational-valued conjectures. Quantum Inf Process 17, 83 (2018). https://doi.org/10.1007/s11128-018-1854-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11128-018-1854-5

Keywords

Navigation