Skip to main content
Log in

Relating an entanglement measure with statistical correlators for two-qudit mixed states using only a pair of complementary observables

  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

We focus on characterising entanglement of high-dimensional bipartite states using various statistical correlators for two-qudit mixed states. The salient results obtained are as follows: (a) A scheme for determining the entanglement measure given by negativity is explored by analytically relating it to the widely used statistical correlators, viz. mutual predictability, mutual information and Pearson correlation coefficient, for different types of bipartite arbitrary-dimensional mixed states. Importantly, this is demonstrated using only a pair of complementary observables pertaining to the mutually unbiased bases. (b) The relations thus derived provide the separability bounds for detecting entanglement obtained for a fixed choice of the complementary observables, while the bounds per se are state-dependent. Such bounds are compared with the earlier suggested separability bounds. (c) We also show how these statistical correlators can enable distinguishing between the separable, distillable and bound entanglement domains of the one-parameter Horodecki two-qutrit states. Further, the relations linking negativity with the statistical correlators have been derived for such Horodecki states in the domain of distillable entanglement. Thus, this entanglement characterisation scheme based on statistical correlators and harnessing complementarity of the observables opens up a potentially rich direction of study which is applicable for both distillable and bound entangled states.

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

Similar content being viewed by others

Data Availability

There are no data associated with the manuscript as it is a theoretical work. Analysis details are available with the corresponding author on reasonable request.

References

  1. Bennett, C.H., Shor, P.W., Smolin, J.A., Thapliyal, A.V.: Entanglement-assisted classical capacity of noisy quantum channels. Phys. Rev. Lett. 83, 3081–3084 (1999). https://doi.org/10.1103/PhysRevLett.83.3081

    Article  ADS  Google Scholar 

  2. Bechmann-Pasquinucci, H., Tittel, W.: Quantum cryptography using larger alphabets. Phys. Rev. A 61, 062308 (2000). https://doi.org/10.1103/PhysRevA.61.062308

    Article  ADS  MathSciNet  Google Scholar 

  3. Cerf, N.J., Bourennane, M., Karlsson, A., Gisin, N.: Security of quantum key distribution using \( d\)-level systems. Phys. Rev. Lett. 88, 127902 (2002). https://doi.org/10.1103/PhysRevLett.88.127902

    Article  ADS  Google Scholar 

  4. Vértesi, T., Pironio, S., Brunner, N.: Closing the detection loophole in bell experiments using qudits. Phys. Rev. Lett. 104, 060401 (2010). https://doi.org/10.1103/PhysRevLett.104.060401

    Article  ADS  Google Scholar 

  5. Sheridan, L., Scarani, V.: Security proof for quantum key distribution using qudit systems. Phys. Rev. A 82, 030301 (2010). https://doi.org/10.1103/PhysRevA.82.030301

    Article  ADS  Google Scholar 

  6. Bruß, D., Christandl, M., Ekert, A., Englert, B.-G., Kaszlikowski, D., Macchiavello, C.: Tomographic quantum cryptography: equivalence of quantum and classical key distillation. Phys. Rev. Lett. 91, 097901 (2003). https://doi.org/10.1103/PhysRevLett.91.097901

    Article  ADS  Google Scholar 

  7. Wang, C., Deng, F.-G., Li, Y.-S., Liu, X.-S., Long, G.L.: Quantum secure direct communication with high-dimension quantum superdense coding. Phys. Rev. A 71, 044305 (2005). https://doi.org/10.1103/PhysRevA.71.044305

    Article  ADS  Google Scholar 

  8. Huang, Y.: Computing quantum discord is NP-complete. New J. Phys. 16(3), 033027 (2014). https://doi.org/10.1088/1367-2630/16/3/033027

    Article  ADS  MathSciNet  Google Scholar 

  9. Harrow, A.W., Natarajan, A., Wu, X.: An improved semidefinite programming hierarchy for testing entanglement. Commun. Math. Phys. 352(3), 881–904 (2017). https://doi.org/10.1007/s00220-017-2859-0

    Article  ADS  MathSciNet  Google Scholar 

  10. Bavaresco, J., Herrera Valencia, N., Klöckl, C., Pivoluska, M., Erker, P., Friis, N., Malik, M., Huber, M.: Measurements in two bases are sufficient for certifying high-dimensional entanglement. Nat. Phys. 14(10), 1032–1037 (2018). https://doi.org/10.1038/s41567-018-0203-z

    Article  Google Scholar 

  11. Schneeloch, J., Howland, G.A.: Quantifying high-dimensional entanglement with Einstein–Podolsky–Rosen correlations. Phys. Rev. A 97, 042338 (2018). https://doi.org/10.1103/PhysRevA.97.042338

    Article  ADS  Google Scholar 

  12. Erker, P., Krenn, M., Huber, M.: Quantifying high dimensional entanglement with two mutually unbiased bases. Quantum 1, 22 (2017). https://doi.org/10.22331/q-2017-07-28-22

    Article  Google Scholar 

  13. Jebarathinam, C., Home, D., Sinha, U.: Pearson correlation coefficient as a measure for certifying and quantifying high-dimensional entanglement. Phys. Rev. A 101, 022112 (2020). https://doi.org/10.1103/PhysRevA.101.022112

    Article  ADS  Google Scholar 

  14. Ghosh, D., Jennewein, T., Sinha, U.: Direct determination of arbitrary dimensional entanglement monotones using statistical correlators and minimal complementary measurements. Quant. Sci. Technol. 7(4), 045037 (2022). https://doi.org/10.1088/2058-9565/ac8e28

    Article  ADS  Google Scholar 

  15. Spengler, C., Huber, M., Brierley, S., Adaktylos, T., Hiesmayr, B.C.: Entanglement detection via mutually unbiased bases. Phys. Rev. A 86, 022311 (2012). https://doi.org/10.1103/PhysRevA.86.022311

    Article  ADS  Google Scholar 

  16. Maccone, L., Bruß, D., Macchiavello, C.: Complementarity and correlations. Phys. Rev. Lett. 114, 130401 (2015). https://doi.org/10.1103/PhysRevLett.114.130401

    Article  ADS  Google Scholar 

  17. Zyczkowski, K., Horodecki, P., Sanpera, A., Lewenstein, M.: Volume of the set of separable states. Phys. Rev. A 58, 883–892 (1998). https://doi.org/10.1103/PhysRevA.58.883

    Article  ADS  MathSciNet  Google Scholar 

  18. Zyczkowski, K.: Volume of the set of separable states. II. Phys. Rev. A 60, 3496–3507 (1999). https://doi.org/10.1103/PhysRevA.60.3496

    Article  ADS  MathSciNet  Google Scholar 

  19. Vidal, G., Werner, R.F.: Computable measure of entanglement. Phys. Rev. A 65, 032314 (2002). https://doi.org/10.1103/PhysRevA.65.032314

    Article  ADS  Google Scholar 

  20. Leggio, B., Napoli, A., Nakazato, H., Messina, A.: Bounds on mixed state entanglement. Entropy 22(1), 62 (2020). https://doi.org/10.3390/e22010062

    Article  ADS  MathSciNet  Google Scholar 

  21. Gray, J., Banchi, L., Bayat, A., Bose, S.: Machine-learning-assisted many-body entanglement measurement. Phys. Rev. Lett. 121, 150503 (2018). https://doi.org/10.1103/PhysRevLett.121.150503

    Article  ADS  Google Scholar 

  22. Zhou, Y., Zeng, P., Liu, Z.: Single-copies estimation of entanglement negativity. Phys. Rev. Lett. 125, 200502 (2020). https://doi.org/10.1103/PhysRevLett.125.200502

    Article  ADS  MathSciNet  Google Scholar 

  23. Elben, A., Kueng, R., Huang, H.-Y.R., Bijnen, R., Kokail, C., Dalmonte, M., Calabrese, P., Kraus, B., Preskill, J., Zoller, P., Vermersch, B.: Mixed-state entanglement from local randomized measurements. Phys. Rev. Lett. 125, 200501 (2020). https://doi.org/10.1103/PhysRevLett.125.200501

    Article  ADS  Google Scholar 

  24. He, H., Vidal, G.: Disentangling theorem and monogamy for entanglement negativity. Phys. Rev. A 91, 012339 (2015). https://doi.org/10.1103/PhysRevA.91.012339

    Article  ADS  MathSciNet  Google Scholar 

  25. Khasin, M., Kosloff, R., Steinitz, D.: Negativity as a distance from a separable state. Phys. Rev. A 75, 052325 (2007). https://doi.org/10.1103/PhysRevA.75.052325

    Article  ADS  MathSciNet  Google Scholar 

  26. Akhtarshenas, S.J., Farsi, M.: Negativity as entanglement degree of the Jaynes–Cummings model. Phys. Scr. 75(5), 608 (2007). https://doi.org/10.1088/0031-8949/75/5/003

    Article  ADS  MathSciNet  Google Scholar 

  27. Adesso, G., Serafini, A., Illuminati, F.: Extremal entanglement and mixedness in continuous variable systems. Phys. Rev. A 70, 022318 (2004). https://doi.org/10.1103/PhysRevA.70.022318

    Article  ADS  Google Scholar 

  28. Eltschka, C., Siewert, J.: Negativity as an estimator of entanglement dimension. Phys. Rev. Lett. 111, 100503 (2013). https://doi.org/10.1103/PhysRevLett.111.100503

    Article  ADS  Google Scholar 

  29. Terhal, B.M.: Bell inequalities and the separability criterion. Phys. Lett. A 271(5), 319–326 (2000). https://doi.org/10.1016/S0375-9601(00)00401-1

    Article  ADS  MathSciNet  Google Scholar 

  30. Edwards, R.E.: Functional Analysis: Theory and Applications. Courier Corporation (2012)

  31. Gharibian, S.: Strong NP-Hardness of the Quantum Separability Problem (2009)

  32. Tiranov, A., Designolle, S., Cruzeiro, E.Z., Lavoie, J., Brunner, N., Afzelius, M., Huber, M., Gisin, N.: Quantification of multidimensional entanglement stored in a crystal. Phys. Rev. A 96, 040303 (2017). https://doi.org/10.1103/PhysRevA.96.040303

    Article  ADS  Google Scholar 

  33. Martin, A., Guerreiro, T., Tiranov, A., Designolle, S., Fröwis, F., Brunner, N., Huber, M., Gisin, N.: Quantifying photonic high-dimensional entanglement. Phys. Rev. Lett. 118, 110501 (2017). https://doi.org/10.1103/PhysRevLett.118.110501

    Article  ADS  Google Scholar 

  34. Sperling, J., Vogel, W.: The schmidt number as a universal entanglement measure. Phys. Scr. 83(4), 045002 (2011). https://doi.org/10.1088/0031-8949/83/04/045002

    Article  ADS  Google Scholar 

  35. Wyderka, N., Chesi, G., Kampermann, H., Macchiavello, C., Bruß, D.: Construction of efficient Schmidt-number witnesses for high-dimensional quantum states. Phys. Rev. A 107, 022431 (2023). https://doi.org/10.1103/PhysRevA.107.022431

    Article  ADS  MathSciNet  Google Scholar 

  36. Buscemi, F.: All entangled quantum states are nonlocal. Phys. Rev. Lett. 108, 200401 (2012). https://doi.org/10.1103/PhysRevLett.108.200401

    Article  ADS  Google Scholar 

  37. Branciard, C., Rosset, D., Liang, Y.-C., Gisin, N.: Measurement-device-independent entanglement witnesses for all entangled quantum states. Phys. Rev. Lett. 110, 060405 (2013). https://doi.org/10.1103/PhysRevLett.110.060405

    Article  ADS  Google Scholar 

  38. Guo, Y., Yu, B.-C., Hu, X.-M., Liu, B.-H., Wu, Y.-C., Huang, Y.-F., Li, C.-F., Guo, G.-C.: Measurement-device-independent quantification of irreducible high-dimensional entanglement. NPJ Quant. Inf. 6(1), 52 (2020). https://doi.org/10.1038/s41534-020-0282-4

    Article  ADS  Google Scholar 

  39. Durt, T., Englert, B.-G., Bengtsson, I., Życzkowski, K.: On mutually unbiased bases. Int. J. Quant. Inf. 08(04), 535–640 (2010). https://doi.org/10.1142/s0219749910006502

    Article  Google Scholar 

  40. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge Series on Information and the Natural Sciences. Cambridge University Press (2000). https://books.google.co.in/books?id=aai-P4V9GJ8C

  41. Klappenecker, A., Roetteler, M.: Constructions of Mutually Unbiased Bases (2003)

  42. Keyl, M.: Fundamentals of quantum information theory. Phys. Rep. 369(5), 431–548 (2002). https://doi.org/10.1016/S0370-1573(02)00266-1

    Article  ADS  MathSciNet  Google Scholar 

  43. Tsokeng, A.T., Tchoffo, M., Fai, L.C.: Dynamics of entanglement and quantum states transitions in spin-qutrit systems under classical dephasing and the relevance of the initial state. J. Physics Commun. 2(3), 035031 (2018). https://doi.org/10.1088/2399-6528/aab51b

    Article  ADS  Google Scholar 

  44. Derkacz, L., Jakóbczyk, L.: Quantum interference and evolution of entanglement in a system of three-level atoms. Phys. Rev. A 74, 032313 (2006). https://doi.org/10.1103/PhysRevA.74.032313

    Article  ADS  Google Scholar 

  45. Huang, Z., Maccone, L., Karim, A., Macchiavello, C., Chapman, R.J., Peruzzo, A.: High-dimensional entanglement certification. Sci. Rep. 6(1), 27637 (2016). https://doi.org/10.1038/srep27637

    Article  ADS  Google Scholar 

  46. Xiang, G.-Y., Li, J., Yu, B., Guo, G.-C.: Remote preparation of mixed states via noisy entanglement. Phys. Rev. A 72, 012315 (2005). https://doi.org/10.1103/PhysRevA.72.012315

    Article  ADS  Google Scholar 

  47. Sentís, G., Eltschka, C., Gühne, O., Huber, M., Siewert, J.: Quantifying entanglement of maximal dimension in bipartite mixed states. Phys. Rev. Lett. 117, 190502 (2016). https://doi.org/10.1103/PhysRevLett.117.190502

    Article  ADS  Google Scholar 

  48. Horodecki, P., Horodecki, M., Horodecki, R.: Bound entanglement can be activated. Phys. Rev. Lett. 82, 1056–1059 (1999). https://doi.org/10.1103/PhysRevLett.82.1056

    Article  ADS  MathSciNet  Google Scholar 

  49. Horodecki, P., Horodecki, M., Horodecki, R.: Bound entanglement can be activated. Phys. Rev. Lett. 82, 1056–1059 (1999). https://doi.org/10.1103/PhysRevLett.82.1056

    Article  ADS  MathSciNet  Google Scholar 

  50. Hiesmayr, B.C., Löffler, W.: Complementarity reveals bound entanglement of two twisted photons. New J. Phys. 15(8), 083036 (2013). https://doi.org/10.1088/1367-2630/15/8/083036

    Article  ADS  Google Scholar 

  51. Vértesi, T., Brunner, N.: Disproving the peres conjecture by showing bell nonlocality from bound entanglement. Nat. Commun. 5(1), 5297 (2014). https://doi.org/10.1038/ncomms6297

    Article  ADS  Google Scholar 

  52. Moroder, T., Gittsovich, O., Huber, M., Gühne, O.: Steering bound entangled states: a counterexample to the stronger peres conjecture. Phys. Rev. Lett. 113, 050404 (2014). https://doi.org/10.1103/PhysRevLett.113.050404

    Article  ADS  Google Scholar 

  53. Horodecki, K., Horodecki, M., Horodecki, P., Oppenheim, J.: Secure key from bound entanglement. Phys. Rev. Lett. 94, 160502 (2005). https://doi.org/10.1103/PhysRevLett.94.160502

    Article  ADS  MathSciNet  Google Scholar 

  54. Epping, M., Brukner, I.C.V.: Bound entanglement helps to reduce communication complexity. Phys. Rev. A 87, 032305 (2013). https://doi.org/10.1103/PhysRevA.87.032305

    Article  ADS  Google Scholar 

  55. Yu, S., Oh, C.H.: Family of nonlocal bound entangled states. Phys. Rev. A 95, 032111 (2017). https://doi.org/10.1103/PhysRevA.95.032111

    Article  ADS  Google Scholar 

Download references

Acknowledgements

US acknowledges partial support provided by the Ministry of Electronics and Information Technology (MeitY), Government of India, under grant for Centre for Excellence in Quantum Technologies with Ref. No. 4(7)/2020–ITEA, partial support of the QuEST-DST Project Q-97 of the Government of India and the QuEST-ISRO research grant. DH thanks NASI for the Senior Scientist Platinum Jubilee Fellowship and acknowledges support of the QuEST-DST Project Q-98 of the Government of India.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Urbasi Sinha.

Ethics declarations

Conflict of interest

The authors declare no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Joint probabilities of measurement outcomes for the chosen states

Statistical correlators like MP, MI and PCC are functions of the joint probabilities of outcomes of the measurements performed on the bipartite state considered. Here we explicitly derive the relevant expressions for the classes of states that we have considered.

1.1 Noisy Bell state

The density operators for the three types of noises are given in Eqs. (9), (10) and (11). The joint probabilities of measurement outcomes for all the three types with respect to the Z basis are

$$\begin{aligned} p_{\textrm{iso},Z} =&\langle i,j|\frac{\mathbbm {1}}{d^2}|i,j\rangle = \frac{1}{d^2} \end{aligned}$$
(36)
$$\begin{aligned} p_{\textrm{cna},Z} =&\frac{1}{d}\sum \limits _{a=0}^{d-1} \langle i,j |a,a\rangle \langle a,a |i,j\rangle = \frac{1}{d}\delta _{ij} \end{aligned}$$
(37)
$$\begin{aligned} p_{\textrm{cnb},Z} =&\frac{1}{d(d-1)} \sum \limits _{\begin{array}{c} a,b=0\\ a\ne b \end{array}} \langle i,j|a,b\rangle \langle a,b|i,j\rangle = \frac{1}{d(d-1)} (1 - \delta _{ij}) \end{aligned}$$
(38)

and therefore, the mutual predictabilities of all these three states are given by

$$\begin{aligned} P_{\textrm{iso},Z} =&\frac{1}{d} \end{aligned}$$
(39)
$$\begin{aligned} P_{\textrm{cna},Z} =&1 \end{aligned}$$
(40)
$$\begin{aligned} P_{\textrm{cnb},Z} =&0 \end{aligned}$$
(41)

Similarly, with respect to the X basis, one obtains

$$\begin{aligned} p_{\textrm{iso},X} =&\langle i,j|\frac{\mathbbm {1}}{d^2}|i,j\rangle = \frac{1}{d^2} \end{aligned}$$
(42)
$$\begin{aligned} p_{\textrm{cna},X} =&\frac{1}{d}\sum \limits _{a=0}^{d-1} \langle i,j |a,a\rangle \langle a,a |i,j\rangle = \frac{1}{d^2} \end{aligned}$$
(43)
$$\begin{aligned} p_{\textrm{cnb},X} =&\frac{1}{d(d-1)} \sum \limits _{\begin{array}{c} a,b=0\\ a\ne b \end{array}} \langle i,j|a,b\rangle \langle a,b|i,j\rangle = \frac{1}{d^2} \end{aligned}$$
(44)

and, therefore, the mutual predictability of all these noises separately is 1/d.

1.2 Werner State

$$\begin{aligned} \rho _\textrm{w} =&a \frac{2}{d(d+1)} P_\textrm{sym}+ (1-a) \frac{2}{d(d-1)} P_\textrm{as} \end{aligned}$$
(45)

where

$$\begin{aligned} P_\textrm{sym} =&\frac{1}{2} \left( \mathbbm {1} + P \right) \end{aligned}$$
(46)
$$\begin{aligned} P_\textrm{as} =&\frac{1}{2} \left( \mathbbm {1} - P \right) \end{aligned}$$
(47)
$$\begin{aligned} P =&\sum \limits _{i,j} |i\rangle \langle j| \otimes |j\rangle \langle i| \end{aligned}$$
(48)

The joint probability of outcomes in the Z basis for the operator P is

$$\begin{aligned} p_{P,Z} =&\sum \limits _{a,b} \langle i,j|a,b\rangle \langle b,a|i,j\rangle = \delta _{ij} \end{aligned}$$
(49)

Note that due to the \(U \otimes U\) invariance of Werner states, the probability of measurement outcomes with respect to the X basis is the same as that with respect to the Z basis.

State-dependent separability bounds

As discussed in §4, with the fixed choice of measurement bases, which in this work are the X and Z bases, our separability bounds are state-dependent. This is because the relations between the negativity and the statistical correlators are state-dependent. On the other hand, with the chosen bases in this paper, the separability criteria of [15, 16] are not suitable to detect all the entangled states (among the classes of states considered here). We show this by using the sums of statistical correlators with respect to the Z and X bases in the figures described below. For comparison, we also indicate in these figures the respective bounds of [15, 16].

In Figs. 131415, we have plotted the negativity of the Noisy Bell state given by Eq. (14) with respect to the sums of MPs, MIs and PCCs for the X and Z bases, respectively, in the case of \(d=3\). Note that, for all the three statistical correlators, the corresponding state-dependent separability bounds depend upon the mixedness parameter a as indicated by Eqs. (15) and (16). In Figs. 1617 and 18, we have plotted the negativity of Werner state with respect to the sums of MPs, MIs and PCCs for the X and Z bases, respectively, in the case of \(d=3\). For these observables, there exists entangled states such that the sums of the statistical correlators are less than the state-independent bounds [15, 16]. In other words, the use of the observables X and Z is not suitable to detect certain entangled states using the state-independent bounds but which can be detected using the state-dependent bounds.

Fig. 13
figure 13

Negativity \((\mathcal {N}\)) of the noisy Bell state versus the sum of MPs \((P_{Z}+P_{X})\) with respect to the X and Z bases for \(d=3\). The vertical dotted line corresponds to the state-independent but observable dependent separability bound \(1 + 1/d\). The line for each value of the state parameter a intersects the horizontal line corresponding to \(\mathcal {N}=0\), with the point of intersection indicating the state-dependent separability bound for a given value of a

Fig. 14
figure 14

Negativity \((\mathcal {N}\)) of the noisy Bell state versus the sum of MIs \((I_{Z}+I_{X})\) with respect to the X and Z bases for \(d=3\). The vertical dotted line corresponds to the state-independent but observable dependent separability bound \(\log _2 d\). The line for each value of the state parameter a intersects the horizontal line corresponding to \(\mathcal {N}=0\), with the point of intersection indicating the state-dependent separability bound for a given value of a

Fig. 15
figure 15

Negativity \((\mathcal {N})\) of the noisy Bell state versus the sum of PCCs \((PCC_{X}+PCC_{Z})\) with respect to the X and Z bases for \(d=3\). The vertical dotted line corresponds to the state-independent but observable dependent separability bound 1. The line for each value of the state parameter a intersects the horizontal line corresponding to \(\mathcal {N}=0\), with the point of intersection indicating the state-dependent separability bound for a given value of a

Fig. 16
figure 16

Negativity \((\mathcal {N})\) of the Werner state versus the sum of MPs \((P_{Z}+P_{X})\) with respect to the X and Z bases for \(d=3\). The vertical dotted line corresponds to the state-independent but observable dependent separability bound \(1 + 1/d\)

Fig. 17
figure 17

Negativity \((\mathcal {N})\) of the Werner state versus the sum of MIs \((I_{Z}+I_{X})\) with respect to the X and Z bases for \(d=3\). The vertical dotted line corresponds to the state-independent but observable dependent separability bound \(\log _2 d\)

Fig. 18
figure 18

Negativity \((\mathcal {N})\) of the Werner state versus the sum of PCCs \((PCC_{X}+PCC_{Z})\) with respect to the X and Z bases. The vertical dotted line corresponds to the state-independent but observable dependent separability bound given by 1

Proof of Maccone et al. Conjecture for pure and coloured noise A states

Here we deal with the problem of finding pairs of complementary observables \(\lbrace \hat{A},\hat{C} \rbrace \) and \(\lbrace \hat{B},\hat{D} \rbrace \) such that the sum of PCCs \(|PCC_{AB}|+|PCC_{CD}|>1\) for a bipartite arbitrary-dimensional entangled state. For two given observables, say, X and Y, \(PCC_{XY}\) is given by

$$\begin{aligned} PCC_{XY} = \frac{\langle X\otimes Y\rangle -\langle X\otimes \mathbbm {1}\rangle \langle \mathbbm {1}\otimes Y\rangle }{\sqrt{\langle X^{2}\otimes \mathbbm {1}\rangle -\langle X\otimes \mathbbm {1}\rangle ^{2}}\sqrt{\langle \mathbbm {1}\otimes Y^{2}\rangle -\langle \mathbbm {1}\otimes Y^2\rangle }} \end{aligned}$$
(50)

Taking the negativity (\(\mathcal {N}\)) as the measure of entanglement, we show that, using suitable pairs of complementary observables \(\{A,C\}\) and \(\{B,D\}\), \(|PCC_{AB}|+|PCC_{CD}|-1\propto \mathcal {N}\) for certain types of states, viz. pure and coloured noise A state.

We proceed as follows. First, we construct the relevant observables and demonstrate their mutual unbiasedness. Next, we obtain the values of the relevant Pearson correlators for a d-dimensional bipartite state. Finally, we derive the desired relationship between the sum of Pearson correlators and the negativity for pure and coloured noise A states.

1.1 Construction of the required observables and their properties

The complementary observables for each subsystem are the generalised Z observable and a modification of the generalised X observable which is given by

$$\begin{aligned} W = \sum _{i,j:\langle i|j\rangle = 0}|j\rangle \langle i| \end{aligned}$$
(51)

In Eq. (51), the summation is over both i and j. For a given i, we sum over only those values of j which are orthogonal to i, i.e. \(\langle i|j\rangle =0\). Another way to express the above observable is to consider generalised Pauli basis \(\hat{\sigma }_{j,k}=|j\rangle \langle k|+|k\rangle \langle j|\) with \(j>k\). We can then write \(W=\underset{j,k}{\sum }\hat{\sigma }_{j,k}\).

The observable defined in Eq. (51) projects each computational basis vector \(|j\rangle \) to \(\sum _{k:\langle k|j\rangle =0}|k\rangle \), i.e. to its complete orthogonal subspace, \(|i\rangle \longrightarrow \sum _{j:\langle i|j\rangle =0}|j\rangle \).

Now, we show that the observables Z and W are complementary to each other, i.e. the corresponding eigenstates are mutually unbiased.

1.1.1 Demonstration of maximum complementarity of the W and Z observables

Note that the eigenstates of Z form the computational basis \(\{|j\rangle \}\). Therefore, to show the maximum complementarity of Z and W we only need to show that the eigenvectors of W are mutually unbiased to the states of the computational basis. For this purpose, we define a suitable basis, mutually unbiased with respect to the computational basis \(\{|j\rangle \}\) given by \(\lbrace |k\rangle =\frac{1}{\sqrt{d}}\sum _{j}\textrm{e}^{\frac{2\textrm{i}\pi jk}{d}}|j\rangle \rbrace \).

For \(k=0\), we can show that

$$\begin{aligned} W|k=0\rangle&= \frac{1}{\sqrt{d}}\sum _{j}W|j\rangle \end{aligned}$$
(52)
$$\begin{aligned}&=\frac{1}{\sqrt{d}}\sum _{j,k:\langle k|j\rangle =0}|k\rangle \end{aligned}$$
(53)
$$\begin{aligned}&=\frac{(d-1)}{\sqrt{d}}\sum _{k}|k\rangle \end{aligned}$$
(54)
$$\begin{aligned}&=(d-1)|k=0\rangle \end{aligned}$$
(55)

The above calculation shows us that \(|k=0\rangle \) is one of the eigenstates of W with eigenvalue \((d-1)\). For \(1 \le k \le (d-1)\), taking \(\omega =\textrm{e}^{\frac{2\textrm{i}\pi }{d}}\) we obtain

$$\begin{aligned} W|k\rangle&= \frac{1}{\sqrt{d}}\sum _{j}\omega ^{jk}W|j\rangle \nonumber \\&=\frac{1}{\sqrt{d}}\sum _{j,l:\langle l|j\rangle =0}\omega ^{jk}|l\rangle \end{aligned}$$
(56)
$$\begin{aligned}&=\frac{1}{\sqrt{d}}\sum _{l}|l\rangle \sum _{j:\langle l|j\rangle =0}\omega ^{jk} \end{aligned}$$
(57)

Note that for any k which is not a multiple of d we have

$$\begin{aligned} \sum _{j=0}^{d-1}\omega ^{jk}=\frac{1-\textrm{e}^{2\textrm{i}\pi k}}{1-\textrm{e}^{\frac{2\textrm{i}\pi k}{d}}}=0. \end{aligned}$$
(58)

Using Eq. (58), we can rewrite Eq. (57) as

$$\begin{aligned} W|k\rangle&=-\frac{1}{\sqrt{d}}\sum _{l}|l\rangle \omega ^{lk}\\&=-|k\rangle \end{aligned}$$

In other words, the states \(|k\rangle =\frac{1}{\sqrt{d}}\sum _{j}\omega ^{jk}|j\rangle \) with \(1\le k\le (d-1)\) are the eigenstates of W with eigenvalues \(-1\).

Next, we outline the steps used in proving a relation which will be crucial for calculating the denominator of the expression of PCC given by Eq. (50).

  • The following relation holds for the observable given by Eq. (51)

    $$\begin{aligned} W^{2}=\left[ (d-1)\mathbbm {1}+(d-2)W\right] \end{aligned}$$
    (59)

    Note that from Eq. (51) it is evident that we can write the matrix form of the observable W as follows:

    $$\begin{aligned} W=\begin{bmatrix} 0 &{} 1 &{} 1 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 0 &{} 1 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 0 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 1 &{} 0 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 1 &{} 1 &{} 0 &{} \dots &{} 1 \\ \vdots &{} \ddots &{} \\ 1 &{} 1 &{} 1 &{} 1 &{} 1 &{} \dots &{} 0 \end{bmatrix} \end{aligned}$$
    (60)

    Consequently, for \(W^{2}\) we have the following expression

    $$\begin{aligned} W^{2}&=\begin{bmatrix} 0 &{} 1 &{} 1 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 0 &{} 1 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 0 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 1 &{} 0 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 1 &{} 1 &{} 0 &{} \dots &{} 1 \\ \vdots &{} \ddots &{} \\ 1 &{} 1 &{} 1 &{} 1 &{} 1 &{} \dots &{} 0 \end{bmatrix} \begin{bmatrix} 0 &{} 1 &{} 1 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 0 &{} 1 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 0 &{} 1 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 1 &{} 0 &{} 1 &{} \dots &{} 1 \\ 1 &{} 1 &{} 1 &{} 1 &{} 0 &{} \dots &{} 1 \\ \vdots &{} \ddots &{} \\ 1 &{} 1 &{} 1 &{} 1 &{} 1 &{} \dots &{} 0 \end{bmatrix}\nonumber \\&= \begin{bmatrix} (d-1) &{} (d-2) &{} (d-2) &{} (d-2) &{} \dots &{} (d-2) \\ (d-2) &{} (d-1) &{} (d-2) &{} (d-2) &{} \dots &{} (d-2) \\ (d-2) &{} (d-2) &{} (d-1) &{} (d-2) &{} \dots &{} (d-2) \\ (d-2) &{} (d-2) &{} (d-2) &{} (d-1) &{} \dots &{} (d-2) \\ \vdots &{} \vdots &{} \vdots &{} \vdots &{} \ddots &{} \vdots \\ (d-2) &{} (d-2) &{} (d-2) &{} (d-2) &{} \dots &{} (d-1) \\ \end{bmatrix}\nonumber \\&=(d-1)\mathbbm {1}+(d-2)W \end{aligned}$$
    (61)

1.2 Pearson correlators for a d-dimensional bipartite state

Here we find the values of \(PCC_{Z}\) and \(PCC_{W}\) for a general bipartite qudit state.

To set the stage, we will define a few mathematical notations. The vector \(|\bar{i}\rangle \) denotes an arbitrary vector from the computational basis spanning the orthogonal subspace of \(|i\rangle \). For example, for \(d=3\), and the computational basis \(\{|0\rangle ,|1\rangle ,|2\rangle \}\), the vector \(|\bar{0}\rangle \) would denote an arbitrary vector from the set \(\{|1\rangle ,|2\rangle \}\). We use this notation in Eq. (51). For a fixed i, the summation \(\sum _{j:\langle i|j\rangle =0}\) is equivalent to \(\sum _{\bar{i}}\). In this notation, we can rewrite Eq. (51) as follows:

$$\begin{aligned} W =\sum _{i,\bar{i}}|\bar{i}\rangle \langle i| \end{aligned}$$
(62)

Given a \(d_{a}\times d_{b}\) bipartite state

$$\begin{aligned} \rho _{ab}=\underset{i,j;k,l}{\sum } \epsilon _{i,j;k,l}|i\rangle \langle j|\otimes |k\rangle \langle l|, \end{aligned}$$
(63)

for \(W_{a}=\sum _{i,\bar{i}}|\bar{i}\rangle \langle i|\) and \(W_{b}=\sum _{k,\bar{k}}|\bar{k}\rangle \langle k|\), the following relationships hold

$$\begin{aligned} \langle W_{a}\otimes W_{b}\rangle&= \sum _{i,\bar{i};k,\bar{k}}\epsilon _{i,\bar{i};k,\bar{k}} \end{aligned}$$
(64)
$$\begin{aligned} \langle W_{a}\otimes \mathbbm {1}_{b}\rangle&= \sum _{i,\bar{i};k,k}\epsilon _{i,\bar{i};k,k} \end{aligned}$$
(65)
$$\begin{aligned} \langle \mathbbm {1}_{a}\otimes W_{b}\rangle&= \sum _{i,i;k,\bar{k}}\epsilon _{i,i;k,\bar{k}} \end{aligned}$$
(66)

1.2.1 Proofs of Eqs. (64)–(66)

For

$$\begin{aligned} \rho =\underset{i,j;k,l}{\sum }\ \epsilon _{i,j;k,l}|i\rangle \langle j|\otimes |k\rangle \langle l| \end{aligned}$$

given by Eq. (63), we have

$$\begin{aligned} W_{a}\otimes W_{b}\rho =\underset{\bar{i},i,j;\bar{k},k,l}{\sum } \epsilon _{i,j;k,l}|\bar{i}\rangle \langle j|\otimes |\bar{k}\rangle \langle l|.\end{aligned}$$

Note that \(\bar{i}\ne i\) and \(\bar{k} \ne k\) by definition. Consequently, for the expectation value, we have

$$\begin{aligned} \langle W_{a}\otimes W_{b}\rho \rangle =\underset{i,\bar{i};k,\bar{k}}{\sum } \epsilon _{i,\bar{i};k,\bar{k}} \end{aligned}$$

thus proving Eq. (64).

Similarly, we can prove Eqs. (65) and (66). Then, combining Eqs. (64)–(66), we can obtain the following expression for the numerator of the expression for \(PCC_{W}\) as follows:

$$\begin{aligned} \sum _{i,\bar{i};k,\bar{k}}\epsilon _{i,\bar{i};k,\bar{k}}-\left( \sum _{i,\bar{i};k}\epsilon _{i,\bar{i};k,k}\right) \left( \sum _{i;k,\bar{k}}\epsilon _{i,i;k,\bar{k}}\right) \end{aligned}$$
(67)

Similarly, from Eq. (63) using \(Z=\sum _{k}e_{k}|k\rangle \langle k|\), we can obtain the expression of the numerator of \(PCC_{Z}\) as follows

$$\begin{aligned} \sum _{j,k}e_{j}e_{k}\epsilon _{j,j;k,k} -\left( \sum _{j,k}e_{j}\epsilon _{j,j;k,k}\right) \left( \sum _{j,k}e_{k}\epsilon _{j,j;k,k}\right) \end{aligned}$$
(68)

Note that \(Tr(W)=0\) as well as \(Tr(Z)=0\) in the computational basis. Using Eqs. (68) and (67), one can then seek to obtain the sum of PCCs for the various types of states and relate them to the negativity for the respective states.

In what follows, we will show that for both pure and coloured noise A states, the following relation holds good

$$\begin{aligned} PCC_{Z}+PCC_{W}-1 \propto \mathcal {N} \end{aligned}$$
(69)

1.3 The sum of Pearson correlators for the pure and coloured noise A states

1.3.1 Pure state

For pure entangled states, we have the Schmidt bases \(\lbrace |i\rangle \rbrace \) and \(\lbrace |\alpha _{i}\rangle \rbrace \) such that

$$\begin{aligned} |\psi \rangle _{d}&= \sum _{i}\sqrt{\lambda _{i}}|i\rangle |\alpha _{i}\rangle \end{aligned}$$
(70)
$$\begin{aligned} \rho _{d}&= \sum _{i;j}\sqrt{\lambda _{i}\lambda _{j}}|i\rangle \langle \alpha _{j}|\otimes |i\rangle \langle \alpha _{j}|. \end{aligned}$$
(71)

The negativity for a pure state is given by

$$\begin{aligned} \mathcal {N}_{p}=\frac{1}{2}\sum _{i,j:i\ne j}\sqrt{\lambda _{i}\lambda _{j}}=\frac{1}{2}\sum _{i,\bar{i}}\sqrt{\lambda _{i}\lambda _{\bar{i}}}. \end{aligned}$$
(72)

Writing Eq. (71) in the form of Eq. (63), we obtain

$$\begin{aligned} \epsilon _{i,j;k,l}= {\left\{ \begin{array}{ll} \sqrt{\lambda _{i}\lambda _{j}} &{} \text {for }i=k\hbox { and }j=l, \epsilon _{i,j;i,j},\\ 0 &{} \text {otherwise.} \end{array}\right. } \end{aligned}$$
(73)

Equation (73) implies that for \(j=\bar{i}\) the nonzero terms are \(\epsilon _{i,\bar{i};i,\bar{i}}=\sqrt{\lambda _{i}\lambda _{\bar{i}}}\). From Eq. (64), we obtain \(\langle W_{a}\otimes W_{b}\rangle =\sum _{i,\bar{i}}\epsilon _{i,\bar{i};i,\bar{i}}=\sum _{i,\bar{i}}\sqrt{\lambda _{i}\lambda _{\bar{i}}}=2\mathcal {N}_{p}\). From Eqs. (65) and (66), we have \(\langle W_{a}\otimes \mathbbm {1}\rangle =\langle \mathbbm {1}\otimes W_{b}\rangle =0\). Consequently, the following relationship is obtained

$$\begin{aligned} PCC_{W}=\frac{2\mathcal {N}_{p}}{\Delta _{a}\Delta _{b}} \end{aligned}$$
(74)

Here the variances \(\Delta _{a},\Delta _{b}\) are given by

$$\begin{aligned}{} & {} \Delta _{a}=\sqrt{\langle W_{a}^{2}\otimes \mathbbm {1}\rangle -\langle W_{a}\otimes \mathbbm {1}\rangle }\\{} & {} \Delta _{b}=\sqrt{\langle \mathbbm {1}\otimes W_{b}^{2}\rangle -\langle \mathbbm {1}\otimes W_{b}\rangle }. \end{aligned}$$

Using Eq. (61) and the result that \(\langle W_{a}\otimes \mathbbm {1}\rangle =\langle \mathbbm {1}\otimes W_{b}\rangle =0\), we can obtain

$$\begin{aligned} \Delta _{a}=\Delta _{b}=\sqrt{d-1} \end{aligned}$$

Thus, Eq. (74) can be rewritten as

$$\begin{aligned} PCC_{W}=\frac{2\mathcal {N}_{p}}{(d-1)} \end{aligned}$$
(75)

Now, considering \(PCC_{Z}\), note that since with respect to the Schmidt decomposition, the outcomes are perfectly correlated, and it follows that

$$\begin{aligned} PCC_{Z}=1 \end{aligned}$$
(76)

Combining Eqs. (75) and (76), we then obtain

$$\begin{aligned} PCC_{Z}+PCC_{W} = 1 + \frac{2\mathcal {N}_{p}}{(d-1)} \end{aligned}$$
(77)

Note that for the pure product state, the above sum is 1.

1.3.2 Coloured noise A state

First, we consider the coloured noise A state defined as follows

$$\begin{aligned} \rho _{cna}&= \frac{p}{d}\sum _{j=0,k=0}^{d-1}|j\rangle \langle k|\otimes |j\rangle \langle k|+\frac{(1-p)}{d}\sum _{j=0}^{d-1}|j\rangle \langle j|\otimes |j\rangle \langle j|\\&= \frac{1}{d}\sum _{j}|j\rangle \langle j|\otimes |j\rangle \langle j|+\frac{p}{d}\sum _{j,\bar{j}}|j\rangle \langle \bar{j}|\otimes |j\rangle \langle \bar{j}| \nonumber \end{aligned}$$
(78)

The negativity \((\mathcal {N}_{a})\) of a colored noise A state is given by

$$\begin{aligned} \mathcal {N}_{a}=\frac{p(d-1)}{2} \end{aligned}$$
(79)

Writing Eq. (78) in the form of Eq. (63), we obtain

$$\begin{aligned} \epsilon _{i,j;k,l}= {\left\{ \begin{array}{ll} \frac{1}{d} &{} \text {if }i=j=k=l, \epsilon _{j,j;j,j} \\ \frac{p}{d} &{} \text {if }i=k\ne j=l, \epsilon _{j,\bar{j};j,\bar{j}}\\ \text {zero} &{} \text {otherwise} \end{array}\right. } \end{aligned}$$
(80)

It is evident from Eq. (80) that the terms \(\epsilon _{j,j;k,\bar{k}}=\epsilon _{j,\bar{j};k,k}=0\). Consequently, R.H.S of Eqs. (65) and (66) are zero and we have \(\underset{j,\bar{j},k,\bar{k}}{\sum }\epsilon _{j,\bar{j};k,\bar{k}}=p(d-1)\).

Using Eq. (64), we then have \(\langle W_{a}\otimes W_{a}\rangle =p(d-1)=2\mathcal {N}_{a}\), whence the numerator of the expression of \(PCC_{W}\) becomes \(2\mathcal {N}_{a}\). Using Eq. (59), the denominator of \(PCC_{W}\) is obtained as \((d-1)\). Thus, for the coloured noise A we obtain

$$\begin{aligned} PCC_{W} = \frac{2\mathcal {N}_{a}}{(d-1)} \end{aligned}$$
(81)

Next, in order to obtain the numerator of the expression for \(PCC_{Z}\) for the coloured noise A state, we can use Eq. (68) along with Eq. (80) which yield

$$\begin{aligned} \langle Z_{a}\otimes Z_{b}\rangle -\langle Z_{a}\otimes \mathbbm {1}_{b}\rangle \langle \mathbbm {1}_{a}\otimes Z_{b}\rangle = \sum _{j}e_{j}^{2} \end{aligned}$$
(82)

The denominator of the expression for \(PCC_{Z}\) can then be obtained as follows

$$\begin{aligned} \langle Z_{a}^{2}\otimes \mathbbm {1}\rangle&= p\sum _{j}e_{j}^{2}+(1-p)\sum _{j}e_{j}^{2}\nonumber \\&= \sum _{j}e_{j}^{2} = \langle \mathbbm {1}\otimes Z_{b}^{2}\rangle \end{aligned}$$
(83)

Using \(Tr(Z_{a})=Tr(Z_{b})=\sum _{j}j=0\), one can verify that

$$\begin{aligned} \langle Z_{a}\otimes \mathbbm {1}\rangle&= \langle \mathbbm {1}\otimes Z_{b}\rangle \nonumber \\&= \sum _{j}e_{j}\nonumber \\&= 0 \end{aligned}$$
(84)

Combining Eqs. (82)–(84), the value of \(PCC_{Z}\) is obtained as

$$\begin{aligned} PCC_{Z}=1 \end{aligned}$$
(85)

Combining Eqs. (85) and (81), the final relation is then derived as

$$\begin{aligned} PCC_{Z}+PCC_{W}=1+\frac{2\mathcal {N}_{a}}{(d-1)} \end{aligned}$$
(86)

To sum up, the sums of the two PCCs, \(PCC_{Z}\) and \(PCC_{W}\), as functions of the negativity of the states considered, given by Eqs. (77) and (86) justify the conjecture made by Maccone et al. [16] for the bipartite arbitrary-dimensional pure and coloured noise A states, respectively.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sadana, S., Kanjilal, S., Home, D. et al. Relating an entanglement measure with statistical correlators for two-qudit mixed states using only a pair of complementary observables. Quantum Inf Process 23, 138 (2024). https://doi.org/10.1007/s11128-024-04348-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11128-024-04348-3

Keywords

Navigation