Skip to main content
Log in

Encrypted quantum state tomography with phase estimation for quantum Internet

  • Regular Article
  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

Quantum state tomography (QST) is a fundamental tool requiring privacy in future distributed systems where unknown states are measured for extracting information. Gentle measurement and differential privacy (DP)-based privacy solutions minimize the damage on unknown state and leakage about the quantum information, respectively. In this article, we propose a fundamentally different design for privacy-preserving QST in a multi-party setting. We assume that Alice delegates QST task of a distant source for which she has no access to a third-party player Bob accessing to the source while preserving the source privacy against the operations realized by Bob. Encrypted QST algorithm is proposed which encodes or maps source computational basis states by exploiting phase estimation and feature mapping concept of quantum machine learning (QML). Bob maps basis states to eigenvalues of a specially designed unitary operator in an entangled manner with his ancillary qubits while teleporting the source qubits back to Alice before applying conventional QST. Encoding mechanism is conjectured as having NP-hard decoding complexity based on difficulty of subset-sum problem combined with Hadamard transform. Linear optical design and quantum circuit implementations are presented for future experiments in noisy intermediate-scale quantum (NISQ) devices. Theoretical and numerical supporting evidences are proposed supporting the proposed eigenstructure. EQST promises further applications for multiple source classification tasks and as a novel feature mapping method for future data embedding tasks in QML.

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

Similar content being viewed by others

Data availability

All data generated or analyzed during this study are included in this published article.

References

  1. D’Ariano, G.M., Paris, M.G., Sacchi, M.F.: Quantum tomography. Adv. Imaging Electron Phys. 128, 206–309 (2003)

    Google Scholar 

  2. Cramer, M., Plenio, M.B., Flammia, S.T., Somma, R., Gross, D., Bartlett, S.D., Landon-Cardinal, O., Poulin, D., Liu, Y.-K.: Efficient quantum state tomography. Nat. Commun. 1(1), 1–7 (2010)

    Article  Google Scholar 

  3. Christandl, M., Renner, R.: Reliable quantum state tomography. Phys. Rev. Lett. 109(12), 120403 (2012)

    Article  ADS  Google Scholar 

  4. Du, Y., Hsieh, M.-H., Liu, T., You, S., Tao, D.: Quantum differentially private sparse regression learning. arXiv preprint arXiv:2007.11921 (2020)

  5. Cacciapuoti, A.S., Caleffi, M., Tafuri, F., Cataliotti, F.S., Gherardini, S., Bianchi, G.: Quantum internet: networking challenges in distributed quantum computing. IEEE Netw. 34(1), 137–143 (2019)

    Article  Google Scholar 

  6. Barz, S., Kashefi, E., Broadbent, A., Fitzsimons, J.F., Zeilinger, A., Walther, P.: Demonstration of blind quantum computing. Science 335(6066), 303–308 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Fitzsimons, J.F.: Private quantum computation: an introduction to blind quantum computing and related protocols. npj Quantum Inf 3(1), 1–11 (2017)

    Article  Google Scholar 

  8. Kashefi, E., Pappa, A.: Multiparty delegated quantum computing. Cryptography 1(2), 12 (2017)

    Article  Google Scholar 

  9. Zhou, L., Ying, M.: Differential privacy in quantum computation. In: 2017 IEEE 30th Computer Security Foundations Symposium (CSF), pp. 249–262. IEEE (2017)

  10. Aaronson, S., Rothblum, G.N.: Gentle measurement of quantum states and differential privacy. In: Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, pp. 322–333 (2019)

  11. Aaronson, S.: Shadow tomography of quantum states. SIAM J. Comput. 49(5), 18–368 (2019)

    MathSciNet  Google Scholar 

  12. Arunachalam, S., Grilo, A.B., Yuen, H.: Quantum statistical query learning. arXiv preprint arXiv:2002.08240 (2020)

  13. Angrisani, A., Doosti, M., Kashefi, E.: Differential privacy amplification in quantum and quantum-inspired algorithms. arXiv preprint arXiv:2203.03604 (2022)

  14. Kitaev, A.Y.: Quantum measurements and the Abelian stabilizer problem. arXiv preprint quant-ph/9511026 (1995)

  15. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge University Press, Cambridge (2010)

    MATH  Google Scholar 

  16. Schuld, M., Petruccione, F.: Machine Learning with Quantum Computers. Springer, Cham (2021)

    Book  MATH  Google Scholar 

  17. Havlíček, V., Córcoles, A.D., Temme, K., Harrow, A.W., Kandala, A., Chow, J.M., Gambetta, J.M.: Supervised learning with quantum-enhanced feature spaces. Nature 567(7747), 209–212 (2019)

    Article  ADS  Google Scholar 

  18. Dwork, C., McSherry, F., Nissim, K., Smith, A.: Calibrating noise to sensitivity in private data analysis. In: Theory of Cryptography Conference, pp. 265–284. Springer (2006)

  19. Dwork, C., Roth, A., et al.: The algorithmic foundations of differential privacy. Found. Trends Theor. Comput. Sci. 9(3–4), 211–407 (2014)

    MathSciNet  MATH  Google Scholar 

  20. Arunachalam, S., Quek, Y., Smolin, J.: Private learning implies quantum stability. arXiv preprint arXiv:2102.07171 (2021)

  21. Ying, S., Ying, M., Feng, Y.: Quantum privacy-preserving perceptron. arXiv preprint arXiv:1707.09893 (2017)

  22. Gross, D., Liu, Y.-K., Flammia, S.T., Becker, S., Eisert, J.: Quantum state tomography via compressed sensing. Phys. Rev. Lett. 105(15), 150401 (2010)

    Article  ADS  Google Scholar 

  23. Kyrillidis, A., Kalev, A., Park, D., Bhojanapalli, S., Caramanis, C., Sanghavi, S.: Provable compressed sensing quantum state tomography via non-convex methods. npj Quantum Inf. 4(1), 1–7 (2018)

    Article  Google Scholar 

  24. Englert, B.-G.: Fringe visibility and which-way information: an inequality. Phys. Rev. Lett. 77(11), 2154 (1996)

    Article  ADS  Google Scholar 

  25. Gulbahar, B.: Strong simulation of tracking single photons with which-way-detectors in linear optics. Quantum Sci. Technol. 8(1), 015003 (2022)

    Article  ADS  Google Scholar 

  26. Schuld, M., Killoran, N.: Quantum machine learning in feature Hilbert spaces. Phys. Rev. Lett. 122(4), 040504 (2019)

    Article  ADS  Google Scholar 

  27. Lloyd, S., Schuld, M., Ijaz, A., Izaac, J., Killoran, N.: Quantum embeddings for machine learning. arXiv preprint arXiv:2001.03622 (2020)

  28. Gulbahar, B.: Quantum path computing: computing architecture with propagation paths in multiple plane diffraction of classical sources of fermion and boson particles. Quantum Inf. Process. 18(6), 1–40 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  29. Gulbahar, B.: Theory of quantum path computing with Fourier optics and future applications for quantum supremacy, neural networks and nonlinear Schrödinger equations. Sci. Rep. 10(1), 1–23 (2020)

    Article  Google Scholar 

  30. Amico, M., Dittel, C.: Simulation of wave-particle duality in multipath interferometers on a quantum computer. Phys. Rev. A 102(3), 032605 (2020)

    Article  ADS  Google Scholar 

  31. Tadej, W., Życzkowski, K.: A concise guide to complex Hadamard matrices. Open Syst. Inf. Dyn. 13(2), 133–177 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  32. Moré, J.J., Sorensen, D.C.: Computing a trust region step. SIAM J. Sci. Stat. Comput. 4(3), 553–572 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  33. Powell, M.J.: A Fortran subroutine for solving systems of nonlinear algebraic equations. Technical report, Atomic Energy Research Establishment, Harwell, England (United Kingdom) (1968)

  34. Conn, A.R., Gould, N.I., Toint, P.L.: Trust-Region Methods. SIAM, Optimization. Society for Industrial and Applied Mathematics, Philadelphia (2000)

    Book  MATH  Google Scholar 

Download references

Acknowledgements

This work was supported by TUBITAK (The Scientific and Technical Research Council of Turkey) under Grant \(\#\)119E584.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Burhan Gulbahar.

Ethics declarations

Conflict of interest

The authors have no conflict of interest to declare that are relevant to the content of this article.

Additional information

Publisher's Note

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

Appendices

Appendix A: Evolution of single photon state in linear optical set-up with multiple WPDs

We assume that \(BS_k\) in Fig. 4a applies the following operator:

$$\begin{aligned} U_{BS,k} \, = \, \begin{bmatrix} \cos (\theta _k) &{} \quad \imath \, \sin (\theta _k) \\ \imath \, \sin (\theta _k) &{} \quad \cos (\theta _k)\\ \end{bmatrix} \end{aligned}$$
(A1)

The standard phase shift operator shifts the phase of the photon in the path with the following for the chosen path index \(j = 0\) or 1:

$$\begin{aligned} \Phi _{k,j} \, \equiv \, \begin{bmatrix} e^{\imath \, \phi _{k,0}} &{} \quad 0 \\ 0 &{} \quad e^{\imath \, \phi _{k,1}}\\ \end{bmatrix} \end{aligned}$$
(A2)

WPD with the index k applies either \(U_{k,0}\) or \(U_{k,1}\) depending on the selected path of the photon passing through BS. Although there is no constraint on the unitaries and the dimension of Hilbert space in WPDs, we assume that each WPD state is two dimensional with the corresponding single qubit unitary gates \(U_{k,0}\) and \(U_{k,1}\). We assume that the initial states in WPDs at time \(t_0\) are pure states represented by \(|D_n\rangle \, \otimes \, |D_{n-1}\rangle \, \ldots \otimes |D_1\rangle \) in order to better observe the evolution of each path through consecutive WPDs. At time \(t_i\), the state is denoted as \(|\Psi _i\rangle \). Then, at time \(t_0\), the initial state is as follows:

$$\begin{aligned} |\Psi _0\rangle \, = \,|Q\rangle \, \otimes \, |D_n\rangle \, \otimes \, |D_{n-1}\rangle \, \ldots \, \otimes \, |D_1\rangle \end{aligned}$$
(A3)

where \(|Q\rangle = \alpha _0 \, |0\rangle \, + \, \alpha _1 \, |1\rangle \) denotes the initial state of the single photon. Assume that \(|Q\rangle = |0\rangle \) is chosen to analyze the evolution of the WPD states more simply. In the following, we remove the tensor product \(\otimes \) in order to simplify the expressions which can be understood from the context. Furthermore, we reorder positions of WPD states for the following modeling in order to better characterize the evolution since the interaction of WPDs starts with the first WPD. At the end, we re-correct again.

After the first beam splitter \(BS_1\), the state is transformed into the following:

$$\begin{aligned} U_{BS,1} \, |0\rangle \,|D_1\rangle \, |D_{2}\rangle \, \ldots \, |D_n\rangle \end{aligned}$$
(A4)

The state becomes \(\big (\cos (\theta _1) |0\rangle \, + \, \imath \, \sin (\theta _1) \, |1\rangle \big )\, |D_1 \, D_{2} \, \ldots \, D_n\rangle \). Then, unitary operator is applied on the first WPD state depending on the path chosen by the photon and the state becomes entangled with the photon path as follows:

$$\begin{aligned} \big (\cos (\theta _1) |0\rangle U_{1,0} |D_1\rangle \, + \, \imath \, \sin (\theta _1) |1\rangle U_{1,1} |D_1\rangle \big )\,| D_2 \, \ldots \, D_n \rangle \ \end{aligned}$$
(A5)

Finally, phase shift operator \(\Phi _{1,j}\) is applied depending on the path resulting in \(|\Psi _1\rangle \) as a superposition of two paths entangling the photon and the first WPD state:

$$\begin{aligned} |\Psi _1\rangle \,= & {} \, \big ( \cos (\theta _1) \, e^{\imath \, \phi _{1,0}} \, |0\rangle U_{1,0} |D_1\rangle \, + \, \imath \, \sin (\theta _1) \, e^{\imath \, \phi _{1,1}} \, |1\rangle U_{1,1} |D_1\rangle \big ) \, \nonumber \\{} & {} \otimes | D_2 \, \ldots \, D_n \rangle \end{aligned}$$
(A6)

Defining \(\widetilde{c}_{k,j} \equiv \cos (\theta _k) \, e^{\imath \, \phi _{k,j}}\), \(\widetilde{s}_{k,j} \equiv \imath \, \sin (\theta _k) \, e^{\imath \, \phi _{k,j}} \) and \( |d_j\rangle _{k} \equiv U_{k,j} |D_k\rangle \) simplifies the expression as \(|\Psi _1\rangle = \, \big ( \widetilde{c}_{1,0} \, |0\rangle |d_0\rangle _{1} + \widetilde{s}_{1,1} \, |1\rangle |d_1\rangle _{1} \big ) \, | D_2 \, \ldots \, D_n \rangle \). Similarly, entanglement of the photon with the next WPD detector results in \(|\Psi _2\rangle \) consisting of the superposition of four paths:

$$\begin{aligned} |\Psi _2\rangle= & {} \big ( |0\rangle ( \widetilde{c}_{2,0}\, \widetilde{c}_{1,0} \, |d_0\rangle _{1} \, + \, \widetilde{s}_{2,0}\, \widetilde{s}_{1,1} \, |d_1\rangle _{1} ) |d_0\rangle _{2} \, \nonumber \\{} & {} + |1\rangle ( \widetilde{c}_{1,0}\, \widetilde{s}_{2,1} \, |d_0\rangle _{1} \, + \, \widetilde{c}_{2,1}\, \widetilde{s}_{1,1} \, |d_1\rangle _{1} ) |d_1\rangle _{2} \big ) \nonumber \\{} & {} \otimes \, | D_3 \, \ldots \, D_n \rangle \, \hspace{0.2in} \end{aligned}$$
(A7)

It can be simply iterated until the \((n+1)\)th step to compute \(|\Psi _{n+1}\rangle \) as follows after re-ordering the positions of WPD states:

$$\begin{aligned} \sum _{j_{1} =0}^{1} \,\sum _{j_{2} =0}^{1} \, \ldots \, \sum _{j_{n} =0}^{1} \, \sum _{j_{n+1} =0}^{1} \, A(j_1, \ldots , j_n, j_{n+1}) \, \, |j_{n+1}\rangle _{n+1} \, |d_{j_n}\rangle _n \, \ldots \, |d_{j_1}\rangle _1\hspace{0.15in}{} & {} \end{aligned}$$
(A8)

where we denote the final state of the photon as \(|j_{n+1}\rangle _{n+1}\) and \(A(j_1, \ldots , j_n, j_{n+1})\) is defined as follows:

$$\begin{aligned} A(j_1, \ldots , j_{n+1})\equiv & {} K_{j_1} \prod _{k=1}^{n} \chi _{k+1, j_{k}, j_{k+1}} \end{aligned}$$
(A9)

where the following amplitudes are defined:

$$\begin{aligned} \chi _{k+1, j_{k}, j_{k+1}}= & {} {\left\{ \begin{array}{ll} \widetilde{c}_{k+1, j_{k+1}}, &{} \text {if } j_{k} \, = \, 0, \, j_{k+1}\,=\,0 \\ \widetilde{s}_{k+1, j_{k+1}}, &{} \text {if } j_{k} \, = \, 1, \, j_{k+1}\,=\,0 \\ \widetilde{s}_{k+1, j_{k+1}}, &{} \text {if } j_{k} \, = \, 0, \, j_{k+1}\,=\,1 \\ \widetilde{c}_{k+1, j_{k+1}}, &{} \text {if } j_{k} \, = \,1, \, j_{k+1}\,=\,1 \\ \end{array}\right. }\end{aligned}$$
(A10)
$$\begin{aligned} K_{j_1}= & {} {\left\{ \begin{array}{ll} \widetilde{c}_{1,0}, &{} \text {if } j_{1} \, = \, 0 \\ \widetilde{s}_{1,1}, &{} \text {if } j_{1} \, = \, 1 \\ \end{array}\right. } \end{aligned}$$
(A11)

Appendix B: Duality of the solution for ancillary qubit state

Assume that the following equalities are satisfied for some unknown \(\omega \), \(\gamma _{0, 0}\), \(\gamma _{0, 1}\) and \(\gamma _{1, 0}\) depending on \(\phi _k\) for \(k \in [1, n+1]\) for varying \(\vec {s}\):

$$\begin{aligned} \varrho _{0,0}= & {} \cos (\omega ) \, e^{\imath \, \gamma _{00}} \end{aligned}$$
(B12)
$$\begin{aligned} \varrho _{1,0}= & {} \sin (\omega )\, e^{\imath \, \gamma _{10}} \end{aligned}$$
(B13)
$$\begin{aligned} \varrho _{0,1}= & {} \sin (\omega )\, e^{\imath \, \gamma _{01}} \end{aligned}$$
(B14)
$$\begin{aligned} \varrho _{1,1}= & {} \cos (\omega )\, e^{\imath \, (\gamma _{01} \, + \, \gamma _{10} \,- \, \gamma _{00}\,- \, \pi )} \end{aligned}$$
(B15)

where \(\varrho _{j_0, k_0}\) are expressed for \(j_0, \, k_0 \, \in [0,1]\) as defined in Sect. 5. Furthermore, assume that equalities (3033) have a solution with \(a_0 = \mu _0\), \(b_0 = \mu _1\), \(a_1 = \mu _2\) and \(b_1 \, = \,0\). Then, by inserting the solutions and using the assumptions in (B12B15), the satisfied equalities in (3033) are converted to the following:

$$\begin{aligned} E_1\equiv & {} \mu _2 \,\sin (2 \, \omega ) \big (\mu _0 \, g_1 \, - \, \mu _1 \, g_2 \big ) \, - \, \sin ^2(\omega ) \big ( \mu _0^2 \, + \, \mu _1^2 \, - \, \mu _2^2 \big ) \end{aligned}$$
(B16)
$$\begin{aligned} E_2\equiv & {} \sin ^2(\omega ) \big (\mu _0^2 \,+ \, \mu _1^2 \,- \, \mu _2^2 \big ) \, + \, \mu _2 \, \sin (2 \,\omega ) \big (\mu _1 \, g_2 \, - \, \mu _0 \, g_1 \big ) \end{aligned}$$
(B17)
$$\begin{aligned} E_3\equiv & {} \sin (\omega ) \Big ( \big ( \mu _1^2 \,- \, \mu _0^2 \big ) \, g_3 \, + \, 2 \, \mu _0 \, \mu _1 \, g_4 \, + \, \mu _2^2 \, g_5\Big ) \nonumber \\{} & {} + \, \mu _2 \, \cos (\omega ) \, \big ( \mu _0 f_4 \, - \, \mu _1 \, f_1 \big ) \end{aligned}$$
(B18)
$$\begin{aligned} E_4\equiv & {} \sin (\omega ) \Big ( \big ( \mu _1^2 \, - \, \mu _0^2 \big ) \, g_4 \, - \, 2 \, \mu _0 \, \mu _1 \, g_3 \, + \, \mu _2^2 \, g_6 \Big ) \nonumber \\{} & {} + \mu _2 \, \cos (\omega ) \big ( \mu _0 \, f_1 \, + \, \mu _1 \, f_4 \big ) \end{aligned}$$
(B19)
$$\begin{aligned} E_5\equiv & {} \mu _0^2 + \mu _1^2 + \mu _2^2 - 1 \end{aligned}$$
(B20)

where \(g_1 = \cos (\gamma _{00} \,- \,\gamma _{10})\), \(g_2 = \sin (\gamma _{00} \, - \, \gamma _{10}) \), \(g_3 = \cos (\gamma _{00} \, + \, \gamma _{01})\), \(g_4 = \sin (\gamma _{00} \, + \, \gamma _{01})\), \(g_5 = \cos (\gamma _{00} \,+ \, \gamma _{10})\), \(g_6 = \sin (\gamma _{00} \, + \, \gamma _{10})\), \(f_1 = \big (\sin (2 \, \gamma _{00}) \, + \, \sin (\gamma _{01} \, + \, \gamma _{10}) \big )\), \(f_2 = \, \mu _2^2 \, \cos (\gamma _{00} \, + \, \gamma _{10})\), \(f_3 = \, 2 \, \mu _0 \, \mu _1 \, \sin (\gamma _{00} \, + \, \gamma _{01})\), \(f_4 = \big (\cos (2 \, \gamma _{00}) \, + \, \cos (\gamma _{01} \, + \, \gamma _{10}) \big )\), \(E_k \, = \, 0\) for \(k \in [1, 5]\). Similarly, if another solution, i.e., the dual solution, is \(a_0 = \mu _2\), \(b_0 \, = \,0\), \(a_1 = - \, \mu _0\) and \(b_1 = \mu _1\), then (B12B15) are utilized to obtain following which should give \(E_k \, = \, 0\) for \(k \in [6, 10]\):

$$\begin{aligned} E_6\equiv & {} - \, E_1 \end{aligned}$$
(B21)
$$\begin{aligned} E_7\equiv & {} - \, E_2 \end{aligned}$$
(B22)
$$\begin{aligned} E_8\equiv & {} \sin (\omega ) \Big ( \big ( \mu _1^2 \, - \, \mu _0^2 \big ) \, g_5 \, - \, 2 \, \mu _0 \, \mu _1 \, g_6 \, + \, \mu _2^2 \,g_3 \Big ) \nonumber \\{} & {} + \, \mu _2 \cos (\omega ) \big ( \mu _0 \, f_4 \, + \, \mu _1 \, f_1 \big ) \end{aligned}$$
(B23)
$$\begin{aligned} E_9\equiv & {} \sin (\omega ) \Big ( \big ( \mu _1^2 \, - \, \mu _0^2 \big ) \, g_6 \, + \, 2 \, \mu _0 \, \mu _1 \, g_5 \, + \, \mu _2^2 \, g_4 \Big ) \nonumber \\{} & {} + \, \mu _2 \cos (\omega ) \big ( \mu _0 \, f_1 \, - \, \mu _1 \, f_4 \big ) \end{aligned}$$
(B24)
$$\begin{aligned} E_{10}\equiv & {} E_5 \end{aligned}$$
(B25)

Therefore, for both the solutions to exist, \(E_k\) should be zero for \(k \in [1, 10]\). Equating \(E_{1}\), \(E_2\) and \(E_3\) to zero gives the following solutions for \(\omega \) and \(\mu _0\):

$$\begin{aligned} \omega \,= & {} \, \tan ^{-1}\Bigg (-\, \frac{\mu _2 \,\Big ( \mu _0 \big (\cos (2 \, \gamma _{00}) \, + \, \cos (\gamma _{01} \,+ \, \gamma _{10}) \big ) \, - \, \mu _1 \, f_1 \Big )}{\left( \mu _1^2 \,- \, \mu _0^2\right) \, g_3 \, + \, f_2 \,+ f_3} \bigg ) \end{aligned}$$
(B26)
$$\begin{aligned} \mu _0\,= & {} \, - \, \mu _1 \, \cot \left( \frac{\gamma _{01} \,- \, \gamma _{10}}{2}\right) \end{aligned}$$
(B27)

Substituting \(\mu _0\) into \(E_k\) for \(k \in \lbrace 3, 4, 8, 9\rbrace \) results in the following equalities:

$$\begin{aligned} \frac{E_3}{\cos (\gamma _{00} \, + \, \gamma _{10})} \, - \, \frac{E_8}{\cos (\gamma _{00} \, + \, \gamma _{01})}= & {} 0 \end{aligned}$$
(B28)
$$\begin{aligned} \frac{E_4}{\sin (\gamma _{00} \, + \, \gamma _{10})} \, - \, \frac{E_9}{\sin (\gamma _{00} \, + \, \gamma _{01})}= & {} 0 \end{aligned}$$
(B29)

If \(\cos (\gamma _{00} \, + \, \gamma _{01})\), \(\cos (\gamma _{00} \, + \, \gamma _{10})\), \(\sin (\gamma _{00} \, + \, \gamma _{01})\) and \(\sin (\gamma _{00} \, + \, \gamma _{10})\) are not equal to zero, i.e., \(\gamma _{00} \, + \, \gamma _{01}\) and \(\gamma _{00} \, + \, \gamma _{10}\) are not multiples of \( \pi \, / \, 2\), then \(E_3 = 0\) and \(E_4 = 0\) imply that \(E_8 = 0\) and \(E_{9} = 0\), respectively. It is also easily seen that \(E_1 \, + \, E_2 = 0\). Since \(E_6 \, = \, - \, E_1\), \(E_7 \, = \, - \, E_2\) and \(E_{10} = E_5\), the equalities \(E_{1} = E_2 = E_5 = E_6 = E_ 7 = E_{10}= 0\) are satisfied together. Therefore, all the equalities are satisfied together for being equal to zero by providing the existence of the duality solution where \(\gamma _{00}\), \(\gamma _{01}\) and \(\gamma _{10}\) depend on \(\phi _k\) for \(k \in [1, n+1]\) for varying \(\vec {s}\) as defined in (B12B15). They should also satisfy the conditions of \(\cos (\gamma _{00} \, + \, \gamma _{01})\), \(\cos (\gamma _{00} \, + \, \gamma _{10})\), \(\sin (\gamma _{00} \, + \, \gamma _{01})\) and \(\sin (\gamma _{00} \, + \, \gamma _{10})\) being not equal to zero.

We have utilized the assumptions about the forms of \(\varrho _{k_0, j_0}\) in (B12B15) while defining \(\gamma _{00}\), \(\gamma _{01}\) and \(\gamma _{10}\). In order to satisfy these equalities, \(E_{k}\) values should be equal to zero for \(k \in [11, 14]\) defined as follows:

$$\begin{aligned} E_{11}\equiv & {} \vert \varrho _{0,0} \vert ^2 \, + \, \vert \varrho _{1,0} \vert ^2 \, -\, 1 \end{aligned}$$
(B30)
$$\begin{aligned} E_{12}\equiv & {} \vert \varrho _{0,0} \vert ^2 \, + \, \vert \varrho _{0,1} \vert ^2 \, -\, 1 \end{aligned}$$
(B31)
$$\begin{aligned} E_{13}\equiv & {} \vert \varrho _{0,1} \vert ^2 \, + \, \vert \varrho _{1,1} \vert ^2 \, -\, 1 \end{aligned}$$
(B32)
$$\begin{aligned} E_{14}\equiv & {} \varrho _{0,0}^* \, \varrho _{1,0} + \varrho _{0,1}^* \, \varrho _{1,1} \end{aligned}$$
(B33)

\(E_{k}\) values for \(k \in [11, 14]\) are calculated by using \(\varrho _{j_0, k_0} \equiv \vec {u}_{k_0}^T \, \big (\prod _{l=1}^{m-1} \mathbf {\widehat{M}}_{m-l, s_{m-l}} \big ) \, \textbf{K}_{n-m+1} \, \vec {v}_{j_0} \, / \, \sqrt{2^{n+1}}\) with the modeling in Sect. 5. If the most general case is chosen with \(m = n+1\), then \(\varrho _{j_0, k_0}\) becomes equal to \( \vec {u}_{k_0}^T \, \big (\prod _{l=1}^{n} \mathbf {\widehat{M}}_{n+1-l, s_{n+1-l}} \big ) \, \vec {v}_{j_0} \, / \, \sqrt{2^{n+1}}\). One can perform the calculations with symbolic variables \(\phi _k\) to check whether \(E_{k} \, = \, 0\) for \(k \in [11, 14]\). We perform iterations of symbolic calculations until \(n \, = \, 6\) for small values of n to observe that \(E_{k} \, = \, 0\) for \(k \in [11, 14]\). We also perform numerical analysis in Sect. B.1 until \(n = 13\) satisfying (B12B15) with great accuracy. It is an open issue to show that the equalities are satisfied for all n by using trigonometry and by exploiting the rotation matrices in the definition of \(\mathbf {\widehat{M}}_{k, s_{k}}\) for a more rigorous proof.

As a result, there exist dual solutions depending on whether \(\phi _k\) for \(k \in [1, n+1]\) satisfies \(E_k = 0\) for \(k \in [1, 10]\) while \(\gamma _{00} \, + \, \gamma _{01}\) and \(\gamma _{00} \, + \, \gamma _{10}\) are not multiples of \( \pi \, / \, 2\). It is an open issue to determine the class of \(\phi _k\) for \(k \in [1, n+1]\) providing the favorable solution and the eigenstructure. Uniformly distributed values of \(\phi _k\) for \(k \in [1, n+1]\) provide promising solutions in numerical analysis as a supporting numerical evidence as shown next.

1.1 B.1. Numerical Analysis

We obtain the unitary matrix \(U_{\vec {\Phi }}\) for varying n and phase shift values by using (18) and then calculate its eigenvalues and eigenvectors. We numerically analyze that it has distinct eigenvalues and also the coefficients \(\alpha _0\) and \(\alpha _1\) defined for the eigenvectors in (21) and (22) satisfy the duality relation. The main assumptions to be satisfied for the proposed eigenstructure are given in (B12B15) and the following equalities \(E_1\) to \(E_{10}\) in (B16B25). We solve (B12B15) for \(\omega \), \(\gamma _{00}\), \(\gamma _{01}\) and \(\gamma _{10}\) given the uniformly distributed values of \(\phi _i\) for \(i \in [1, n+1]\) and define the following performance metric for the solution corresponding to a specific \(\vec {s} = [s_1 \, \ldots \, s_{m-1}]\) for \(m \in [1, n+1]\) indexed with \(k \in [1, 2^n]\):

$$\begin{aligned} \epsilon _{1}[k]\equiv & {} \vert \varrho _{0,0} \, - \, \cos (\omega ) \, e^{\imath \, \gamma _{00}} \vert \, + \, \vert \varrho _{1,0} \, - \, \sin (\omega )\, e^{\imath \, \gamma _{10}} \vert \, \nonumber \\{} & {} + \, \vert \varrho _{0,1} \, - \, \sin (\omega )\, e^{\imath \, \gamma _{01}} \vert + \, \vert \varrho _{1,1} \, - \, \cos (\omega )\, e^{\imath \, (\gamma _{01} \, + \, \gamma _{10} \,- \, \gamma _{00}\,- \, \pi )} \vert \qquad \quad \end{aligned}$$
(B34)

We use a classical computer with finite precision, i.e., use MATLAB with 16 digits of precision by default, to solve the nonlinear equalities by using fsolve function in the optimization toolbox. We set tolerance levels for the optimization toolbox equal to \(10^{-30}\). It solves nonlinear equations based on trust region methods [32,33,34]. The second error metric we define is based on solutions of (B16B25) to find \(a_{0,k}\), \(b_{0,k}\), \(a_{1,k}\) and \(b_{1,k}\) for the \(k^{th}\) eigenvector by using calculated values of \(\omega \), \(\gamma _{00}\), \(\gamma _{01}\) and \(\gamma _{10}\). We define the error in satisfying the equalities by also considering that \(E_6 = - E_1\), \(E_7 = - E_2\) and \(E_{10} = E_5\) for \(k^{\text{ th }}\) vector \(\vec {s}\) for \(k \in [1, 2^n]\) as follows:

$$\begin{aligned} \epsilon _{2}[k]\equiv & {} \, \vert E_1 \vert \, + \, \vert E_3 \vert \, + \, \vert E_4 \vert \, + \,\vert E_5 \vert \, + \, \vert E_8 \vert \, + \, \vert E_9 \vert \end{aligned}$$
(B35)

We desire that \(\epsilon _1\) and \(\epsilon _2\) are small by satisfying the proposed eigenstructure.

Finally, the distinctness of eigenvalues is checked by calculating differences among neighbor pairs of the angles of eigenvalues after sorting them in ascending order. If we define \( \equiv e^{\imath \, 2\, \pi \, \lambda _k }\) as the eigenvalue corresponding to kth eigenvector for \(k \in [1, 2^{n+1}]\), then eigenvalue distinctness vector \(\vec {d}_{E}\) with elements denoted by \(d_{E}[k]\) for \(k \in [1, 2^{n+1}-1 ]\) is defined as follows:

$$\begin{aligned} d_{E}[k] \equiv \vert \lambda _{k+1} \, - \, \lambda _{k} \vert \end{aligned}$$
(B36)

The minimum value in the vector \(\vec {d}_{E}\) determines the minimum distance between angle parameters of neighbor eigenvalues. It determines the number of ancillary qubits in Bob, i.e., \(\widetilde{t}\), in order to approximate the eigenvalue with enough precision so that different basis states will be reliably mapped to different eigenvalues.

A numerical analysis example is provided for the following \(\vec {\Phi }\):

$$\begin{aligned} \vec {\Phi } = \begin{bmatrix} \phi _1&\ldots&\phi _{n+1} \end{bmatrix} = \pi \begin{bmatrix} 1.755&0.648&1.754&0.615&0.880&1.061&1.290 \end{bmatrix} \end{aligned}$$
(B37)

where \(n = 6\) and \(\phi _i \, / \, \pi \) has three digits of precision. The values of \(\epsilon _1[k]\) and \(\epsilon _2[k]\) are shown for \(k \in [1, 2^{n}]\) in Fig. 7a and b, respectively. In Fig. 7c, \(d_{E}[k]\) for \(k \in [1, 2^{n+1} -1 ]\) is shown. It is observed in Fig. 7a and b that error metrics are significantly small satisfying the proposed eigenstructure with accuracy. Eigenvalues are all different with the minimum difference of the angle parameters \(\approx \, 5\, \times 10^{-6}\) as shown in Fig. 7c where there are a large number of neighbor pairs with much larger differences reaching \(\approx 3\, \times 10^{-2}\).

Fig. 7
figure 7

The eigenvalues and eigenvectors of \(U_{\vec {\Phi }}\) are calculated for \(\vec {\Phi } = \begin{bmatrix} \phi _1&\phi _2&\ldots&\phi _{n+1} \end{bmatrix} = \pi \, \begin{bmatrix} 1.755&0.648&1.754&0.615&0.880&1.061&1.290 \end{bmatrix}\) where \(n = 6\). The performance metric values satisfying the theoretically modeled architecture are shown for a \(\epsilon _1[k]\) and b \(\epsilon _2[k]\) for \(k \in [1, 2^n]\) by numerically analyzing the accuracy of satisfying (B12B15) and (B16B25), respectively. c \(d_{E}[k]\) showing the distance between the angular parameters of neighbor pairs of the sorted eigenvalues for \(k \in [1, 2^{n+1} -1]\)

Next, we calculate performance metrics for many trials of uniformly distributed phase shift values and n values. We denote performance metrics indexed by using also the trial number l and n as \(\epsilon _{1,n,l}[k]\), \(\epsilon _{2,n,l}[k]\) and \(d_{E,n, l}[k]\) for a uniformly distributed \(\vec {\Phi }\) in each trial. Then, we define average errors in order to characterize how much the proposed eigenstructure is satisfied for increasing values of n on average of all \(\vec {s}\) for \(\epsilon _{1}\) or \(\epsilon _{2}\). In addition, we find the minimum distance for the angle parameters of the eigenvalues for each trial. We define the following:

$$\begin{aligned} \overline{\epsilon }_{i}[n,l]\equiv & {} \frac{1}{2^{n+1}} \sum _{k = 1}^{2^{n+1}} \epsilon _{i,n,l}[k] \end{aligned}$$
(B38)
$$\begin{aligned} \overline{d}_{E}[n,l]\equiv & {} \min _{k} \lbrace d_{E, n, l}[k] \rbrace \end{aligned}$$
(B39)

where \( i \in [1, 2]\). The number of digits of precision in phase shift values is set to 16 by conjecturing that it increases the likelihood to obtain distinct eigenvalues. We are motivated based on the complexity of calculating \(\textbf{V}_{\vec {s}}\) in (27) with consecutive multiplication of rotation operators.

Fig. 8
figure 8

The average performance metrics of \(U_{\vec {\Phi }}\) for a \(\overline{\epsilon }_{1}[n,l]\) and b \(\overline{\epsilon }_{2}[n,l]\) both sorted in ascending order with respect to the value of \(\overline{\epsilon }_{2}[n,l]\) for trial indices l. c \(\overline{\epsilon }_{1}[n,l]\) sorted in ascending order for l. d \(\overline{d}_{E}[n]\) for varying values of n

We sort the values of \(\overline{\epsilon }_{2}[n,l]\) in ascending order with respect to l in order to observe the performance of \(\overline{\epsilon }_{1}[n,l]\) for specific \(\vec {\Phi }\) encountered in different trials where the equalities in (B16B25) are not satisfied with accuracy. The performance results sorted with respect to the values of \(\overline{\epsilon }_{2}[n,l]\) are shown in Fig. 8a and b for \(\overline{\epsilon }_{1}[n,l]\) and \(\overline{\epsilon }_{2}[n,l]\), respectively, for varying \(n \in [3, 13]\) and \(l \in [1, 150]\). It is observed in Fig. 8a and b that the proposed eigenstructure is maintained with enough accuracy for many values of uniformly distributed \(\vec {\Phi }\) for varying values of n. The value of \(\overline{\epsilon }_{1}[n,l]\) for varying n is approximately \(\approx 10^{-15}\) as also shown in Fig. 8c where the values of \(\overline{\epsilon }_{1}[n,l]\) are sorted in ascending order. while the error occurs due to \(\overline{\epsilon }_{2}[n,l]\). The total count of samples providing good results decreases as we increase n, i.e., it becomes more difficult to find a good random \(\vec {\Phi }\) providing small \(\overline{\epsilon }_{2}[n,l]\). It is observed that the fraction of bad \(\vec {\Phi }\) values increases almost linearly with increasing n requiring to get more samples to find a good \(\vec {\Phi }\) with increasing n. It is an open issue to determine the conditions on \(\vec {\Phi }\) satisfying the proposed eigenstructure for large values of n.

On the other hand, we define the average minimum distance for M trials where the specific trials are selected producing significantly small \(\overline{\epsilon }_{1}[n,l]\) and \(\overline{\epsilon }_{2}[n,l]\), i.e., smaller than \(10^{-15}\). Then, \(\overline{d}_{E}[n] \) is defined as follows for the specific set of trials producing good \(\vec {\Phi }\) values and the proposed eigenstructure:

$$\begin{aligned} \overline{d}_{E}[n]\equiv & {} \frac{1}{M} \sum _{l = 1}^{M} \overline{d}_{E}[n,l] \end{aligned}$$
(B40)

Numerical analysis of \(\overline{d}_{E}[n]\) allows to observe how the minimum distance increases for increasing n values. Angular parameters of the eigenvalues are exponentially closer with increasing n due to the increasing size of the Hilbert space with \(2^{n+1}\) and the number of eigenvalues as shown in Fig. 8d. This will increase the number of ancillary qubits of Bob \(\widetilde{t}\) as \(\mathcal {O}(n)\) with a linear increase in order to increase the precision for detecting different eigenvalues.

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

Gulbahar, B. Encrypted quantum state tomography with phase estimation for quantum Internet. Quantum Inf Process 22, 288 (2023). https://doi.org/10.1007/s11128-023-04034-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11128-023-04034-w

Keywords

Navigation