Skip to main content
Log in

Statistical intrusion detection and eavesdropping in quantum channels with coupling: multiple-preparation and single-preparation methods

  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

Classical, i.e., non-quantum, communications include configurations with multiple-input multiple-output (MIMO) channels. Some associated signal processing tasks consider these channels in a symmetric way, i.e., by assigning the same role to all channel inputs, and similarly to all channel outputs. These tasks especially include channel identification/estimation and channel equalization, tightly connected with source separation. Their most challenging version is the blind one, i.e., when the receivers have (almost) no prior knowledge about the emitted signals. Other signal processing tasks consider classical communication channels in an asymmetric way. This especially includes the situation when data are sent by Emitter 1 to Receiver 1 through a main channel, and an “intruder” (including Receiver 2) interferes with that channel so as to extract information, thus performing so-called eavesdropping, while Receiver 1 may aim at detecting that intrusion, which leads to a decision problem (existence of intrusion/no intrusion). Part of the above processing tasks have been extended to quantum channels, including those that have several quantum bits (qubits) at their input and output. For such quantum channels, beyond previously reported work for symmetric scenarios, we here address asymmetric (blind and non-blind) ones, with emphasis on intrusion detection and additional comments about eavesdropping. To develop fundamental concepts, we first consider channels with exchange coupling as a toy model. We especially use the general quantum information processing framework that we recently developed, to derive new attractive intrusion detection methods based on a single preparation of each state. Finally, we discuss how the proposed methods might be extended, beyond the specific class of channels analyzed here.

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.

Similar content being viewed by others

Data availability

Data sharing not applicable to this article as no datasets were generated or analyzed during the current study.

Notes

  1. See also [29, p. 398] for the other earliest references.

  2. The terms “mixing” and “mixtures” should be considered with care when dealing with quantum data: In this paper, when speaking of random pure states, we implicitly refer to some statistical mixtures, as defined in quantum mechanics, but, except in the present note, we do not explicitly use the expression “statistical mixture.”

  3. These vectors \( | + {\rangle } \) and \( | - {\rangle } \) are often, respectively, denoted as \(|0 {\rangle } \) and \(|1 {\rangle } \) (see e.g., [29]), especially when considering an abstract view of qubits. When having in mind the physical implementation of qubits as electron spins, as in most of the present paper, the notations \( | + {\rangle } \) and \( | - {\rangle } \) are also widely used, with a reference to spin component measurements along the quantization axis, as detailed further in this paper.

  4. It should be noted that the observed signals involved in this QSS problem have a specific nature, as compared with standard non-quantum BSS problems. In the latter problems, each value of an observed signal is usually the value of a measured physical quantity, such as the value of a voltage measured at a given time. In contrast, as shown by (22), each value of an observed signal is here the value of a probability (which is estimated in practice). The overall signal composed of all successive values of a given observation (e.g., all values of \(x_{1}\)) therefore consists of a set of values of probabilities (e.g., all values of \({p_1}\) ), which depend on the values of the coefficients used for initializing the qubit states.

References

  1. Abed-Meraim, K., Qiu, W., Hua, Y.: Blind system identification. Proc. IEEE 85(8), 1310–1322 (1997)

    Article  Google Scholar 

  2. Baldwin, C.H., Kalev, A., Deutsch, I.: Quantum process tomography of unitary and near-unitary maps. Phys. Rev. A 90, 012110–1 to 012110–10 (2014)

  3. Blume-Kohout, R., Gamble, J.K., Nielsen, E., Mizrahi, J., Sterk, J.D., Maunz, P.: Robust, self-consistent, closed-form tomography of quantum logic gates on a trapped ion qubit. arXiv:1310.4492v1 (2013)

  4. Boyd, R.: Non-linear Optics. Academic Press, Cambridge (2008)

    Google Scholar 

  5. Branderhorst, M.P.A., Nunn, J., Walmsley, I.A., Kosut, R.L.: Simplified quantum process tomography. New J. Phys. 11, 115010+12 (2009)

  6. Chuang, I.L., Nielsen, M.A.: Prescription for experimental determination of the dynamics of a quantum black box. J. Mod. Opt. 44(11–12), 2455–2467 (1997)

    Article  ADS  Google Scholar 

  7. Comon, P., Jutten, C.: Handbook of Blind Source Separation. Independent Component Analysis and Applications, Academic Press, Oxford (2010)

    Google Scholar 

  8. Deville, Y.: Blind source separation and blind mixture identification methods. In: Webster, J. (ed.) Wiley Encyclopedia of Electrical and Electronics Engineering, pp. 1–33. Wiley (2016)

  9. Deville, Y., Deville, A.: Blind separation of quantum states: estimating two qubits from an isotropic Heisenberg spin coupling model. In: Proceedings of the 7th International Conference on Independent Component Analysis and Signal Separation (ICA 2007), vol. LNCS 4666, pp. 706–713. London, UK, Springer (2007). ISSN 0302-9743. Erratum: replace two terms \( E \{ r_i \} E \{ q_i \} \) in (33) of [9] by \( E \{ r_i q_i \} \), since \( q_i \) depends on \( r_i\)

  10. Deville, Y., Deville, A.: Classical-processing and quantum-processing signal separation methods for qubit uncoupling. Quantum Inf. Process. 11(6), 1311–1347 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  11. Deville, Y., Deville, A.: A quantum-feedforward and classical-feedback separating structure adapted with monodirectional measurements; blind qubit uncoupling capability and links with ICA. In: Proceedings of the 23rd IEEE International Workshop on Machine Learning for Signal Processing (MLSP 2013). Southampton, United Kingdom (2013)

  12. Deville, Y., Deville, A.: Blind qubit state disentanglement with quantum processing: principle, criterion and algorithm using measurements along two directions. In: Proceedings of the 2014 IEEE International Conference on Acoustics, Speech, and Signal Processing (ICASSP 2014), pp. 6262–6266. Florence, Italy (2014)

  13. Deville, Y., Deville, A.: Quantum-source independent component analysis and related statistical blind qubit uncoupling methods. In: Naik, G.R., Wang, W. (eds.) Blind Source Separation: Advances in Theory, Algorithms and Applications, vol. 1, pp. 3–37. Springer, Berlin (2014)

  14. Deville, Y., Deville, A.: From blind quantum source separation to blind quantum process tomography. In: Proceedings of the 12th International Conference on Latent Variable Analysis and Signal Separation (LVA/ICA 2015), pp. 184–192. Liberec, Czech Republic. Springer International Publishing Switzerland, LNCS 9237 (2015)

  15. Deville, A., Deville, Y.: Concepts and criteria for blind quantum source separation and blind quantum process tomography. Entropy 19, paper no. 311 (2017)

  16. Deville, Y., Deville, A.: Blind quantum source separation: quantum-processing qubit uncoupling systems based on disentanglement. Digital Signal Process. 67, 30–51 (2017)

    Article  MathSciNet  Google Scholar 

  17. Deville, Y., Deville, A.: Stochastic quantum information processing, with applications to blind quantum system identification and source separation. In: Proceedings of the 2018 IEEE 28th International Workshop on Machine Learning for Signal Processing (MLSP 2018). Aalborg, Denmark (2018)

  18. Deville, Y., Deville, A.: Quantum process tomography with unknown single-preparation input states: Concepts and application to the qubit pair with internal exchange coupling. Phys. Rev. A 101(4), 042332–1 to 042332–18 (2020)

  19. Deville, Y., Deville, A.: New single-preparation methods for unsupervised quantum machine learning problems. IEEE Transactions on Quantum Engineering 2(Article Sequence Number: 3104224), 1–24 (2021). https://doi.org/10.1109/TQE.2021.3121797

  20. Ding, Z., Li, Y.: Blind equalization and identification. Marcel Dekker, New York (2001)

    Google Scholar 

  21. Dirac, P.: The Principles of Quantum Mechanics. Clarendon Press, Oxford (1930)

    MATH  Google Scholar 

  22. Hyvarinen, A., Karhunen, J., Oja, E.: Independent Component Analysis. Wiley, New York (2001)

    Book  Google Scholar 

  23. Krenn, M., Malik, M., Scheidl, T., Ursin, R., Zeilinger, A.: Optics in our time. In: Al-Amri, M.D., El-Gomati, M.M., Zubairy, M.S. (eds.) Quantum Communication with Photons, pp. 485–482. Springer, Cham (2016)

  24. Landau, L., Lifchitz, E.: Electrodynamique des milieux continus (French Edition, Physique théorique, Tome 8). Editions Mir, Moscow, USSR (1969)

  25. Mansour, A., Barros, A.K., Ohnishi, N.: Blind separation of sources: methods, assumptions and applications. IEICE Trans. Fundam. Electron. Commun. Comput. Sci. E83-A(8), 1498–1512 (2000)

  26. Mansour, A., Mesleh, R., Abaza, M.: New challenges in wireless & free space optical communications. Elsevier Opt. Lasers Eng. 89, 95–108 (2017)

    Article  Google Scholar 

  27. Merkel, S.T., Gambetta, J.M., Smolin, J.A., Poletto, S., Córcoles, A.D., Johnson, B.R., Ryan, C.A., Steffen, M.: Self-consistent quantum process tomography. Phys. Rev. A 87, 062119–1 to 062119–9 (2013)

  28. Navon, N., Akerman, N., Kotler, S., Glickman, Y., Ozeri, R.: Quantum process tomography of a Mølmer-Sørensen interaction. Phys. Rev. A 90, 010103–1 to 010103–5 (2014)

  29. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge University Press, Cambridge (2000)

    MATH  Google Scholar 

  30. Preskill, J.: Lecture notes for ph219/cs219: quantum information and computation, ch. 3. http://www.theory.caltech.edu/preskill/ph219/chap3_15.pdf (2018 (fall term))

  31. Proakis, J.G.: Digital Communications. McGraw-Hill, Boston (2001)

    MATH  Google Scholar 

  32. Scharf, L.L.: Statistical Signal Processing. Detection, Estimation and Time Series Analysis, Addison-Wesley, Reading (1991)

    MATH  Google Scholar 

  33. Shukla, A., Mahesh, T.S.: Single-scan quantum process tomography. Phys. Rev. A 90, 052301–1 to 052301–6 (2014)

  34. Takahashi, M., Bartlett, S.D., Doherty, A.C.: Tomography of a spin qubit in a double quantum dot. Phys. Rev. A 88, 022120–1 to 022120–9 (2013)

  35. Wang, Y., Dong, D., Petersen, I.R., Zhang, J.: An approximate algorithm for quantum Hamiltonian identification with complexity analysis. In: Proceedings of the 20th World Congress of the International Federation of Automatic Control (IFAC 2017), pp. 12241–12245. Toulouse, France (2017)

  36. White, A.G., Gilchrist, A.: Measuring two-qubit gates. J. Opt. Soc. Am. B 24(2), 172–183 (2007)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Razvan Scripcaru for his participation in the early stages of this investigation.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yannick Deville.

Ethics declarations

Conflict of interest

The authors declare that they have 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

Appendix A: Definition of a single qubit

Qubits are widely used instead of classical bits for performing computations in the field of QIP [29]. Whereas a classical bit can only take two values, usually denoted as 0 and 1, at an initial time \(t_0\) a qubit with index \(j\) has a quantum state expressed, for a pure state, as

$$\begin{aligned} | \psi _ j( t_0) {\rangle } = \alpha _ j| + {\rangle } + \beta _ j| - {\rangle } \end{aligned}$$
(19)

in the basis defined by the two orthonormal vectors that we hereafterFootnote 3 denote \( | + {\rangle } \) and \( | - {\rangle } \), where \( \alpha _ j\) and \( \beta _ j\) are two complex-valued coefficients constrained to meet the condition

$$\begin{aligned} | \alpha _ j| ^2 + | \beta _ j| ^2 = 1 \end{aligned}$$
(20)

which expresses that the state \( | \psi _ j( t_0) {\rangle } \) is normalized. In most of the literature, \( \alpha _ j\) and \( \beta _ j\) are deterministic, i.e., fixed, values so that \( | \psi _ j( t_0) \rangle \) is a deterministic pure state. In part of our investigations dealing with BQSS and related tasks, we also considered the case when \( \alpha _ j\) and \( \beta _ j\) are random variables (RVs), so that \( | \psi _ j( t_0) \rangle \) is a random pure state (see e.g., [10, 13, 15, 18]).

From a Quantum Physics (QP) point of view, the above abstract mathematical model especially applies to electron spins 1/2, which are quantum (i.e., non-classical) objects. The component of such a spin, with index \(j\), along a given arbitrary axis Oz defines a two-dimensional linear operator \( s_{ jz}. \) The two eigenvalues of this operator are equal to \( + \frac{1}{2} \) and \( - \frac{1}{2} \) in normalized units, and the corresponding eigenvectors are therefore denoted as \( | + {\rangle } \) and \( | - {\rangle } \). The value obtained when measuring this spin component can only be \( + \frac{1}{2} \) or \( - \frac{1}{2} \). Moreover, let us assume this spin is in the state \( | \psi _{ j} ( t_0) {\rangle } \) defined by (19) when performing such a measurement. Then, the probability that the measured value is equal to \( + \frac{1}{2} \) (respectively, \( - \frac{1}{2} \)) is equal to \( | \alpha _{ j} | ^2 \) (respectively \( | \beta _{ j} | ^2 \)), i.e., to the squared modulus of the coefficient in (19) of the associated eigenvector \( | + {\rangle } \) (respectively \( | - {\rangle } \)).

The above discussion concerns the state of the considered spin at a given initial time \(t_0\). This state then evolves with time. The spin is here supposed to be placed in a static magnetic field and thus coupled to it. The time interval when it is considered is assumed to be short enough for the coupling between the spin and its environment to be negligible. In these conditions, the spin has a Hamiltonian [29]. Therefore, if the spin state \( | \psi _{ j} ( t_0) {\rangle } \) at time \( t_0 \) is defined by (19), it then evolves according to Schrödinger’s equation and its value at any subsequent time t is

$$\begin{aligned} | \psi _{ j} (t) {\rangle } = \alpha _{j } e^{- i \omega _p ( t - t_0 ) } | + {\rangle } + \beta _{j } e^{- i\omega _m ( t - t_0 ) } | - {\rangle } \end{aligned}$$
(21)

where the real (angular) frequencies \( \omega _p \) and \( \omega _m \) depend on the considered physical setup and i is the imaginary unit.

Appendix B: Quantum source separation problem associated with the considered quantum channels

We here detail the quantum source separation problem associated with the “mixing model” (5)–(7) . That model involves the following items. The observations are the probabilities \(p_1\), \({p_2}\) and \({p_4}\) measured for each choice of the initial states (1) of the qubits. More precisely, these probabilities are not known exactly but estimated in practice. The procedure that we used to this end, e.g., in [10, 13], and that is also widely employed in the QIP literature [3, 6], operates as follows for each choice of the initial states (1) of the qubits. We repeatedly perform two operations: (i) we first initialize these qubits according to (1) and (ii) after a fixed time interval when coupling occurs, we measure the two spin components along Oz associated with the system composed of these two coupled qubits. The relative frequencies of occurrence of all four possible couples of values of spin components (i.e., \(\left( +\frac{1}{2},+\frac{1}{2}\right) \) to \(\left( -\frac{1}{2},-\frac{1}{2}\right) \)) then yield estimates of the corresponding probabilities. This approach therefore requires a large number (typically from a few thousand up to a few hundred thousand [10, 16]) of copies of the considered two-qubit state. At this stage, we ignore the resulting estimation errors and therefore consider the exact mixing model (5)–(7). Using standard BSS notations, the observation vector is therefore \(x= [ x_{1}, x_{2}, x_{3}] ^{T} \) , where \( ^T \) stands for transpose andFootnote 4

$$\begin{aligned} x_{1}= {p_1},\quad x_{2}= {p_2} , \quad x_{3}= {p_4} . \end{aligned}$$
(22)

Equations (5)–(7) show that the source vector to be retrieved from these observations turns out to be \( s= [ s_{1}, s_{2}, s_{3}] ^{T} \) with \( s_{1}= r_1, s_{2}= {r_2}\) and \( s_{3}= \varDelta _I \). The parameters \({q_j}\) are then derived from (4). The four phase parameters in (3) cannot be individually extracted from their combination \(\varDelta _I\) (anyway, only the phase differences \( ( \phi _j - \theta _j ) \) have a physical meaning [18]). The transform from the sources to the observations defined by the nonlinear mixing model (5)–(7) involves a single “mixing parameter,” namely v. As shown by (10), this parameter always meets the condition \( 0 \le v^2 \le 1 \).

Appendix C: Communications based on photons

Communications based on photons deserve the following comments. First considering the classical framework, everyday communications use electromagnetic waves propagating either in free space or in a solid medium, e.g., an optical fiber [26]. Such media are non-magnetic, and their electric properties are classically described by the induction vector \(\overrightarrow{D}\) (H. Lorentz), representing a local mean of the microscopic electric vector [24]. \(\overrightarrow{E }\) being the applied field, in vacuum, \(\overrightarrow{D}=\varepsilon _{0} \overrightarrow{E}\) (SI units, \(\varepsilon _{0}\): vacuum permittivity), and in a dielectric medium \(\overrightarrow{D}=\varepsilon _{0}\overrightarrow{E} +\overrightarrow{P}.\) The polarization \(\overrightarrow{P}\) may be seen as the response to the excitation \(\overrightarrow{E}\), ferroelectrics, with a spontaneous polarization, being an exception. Most dielectric media are linear, i.e., \(\overrightarrow{P}\) increases linearly with the excitation (description using a scalar or more generally a tensor not depending upon the excitation). Moreover, the appearance of the laser in 1960, i.e., of intense coherent electromagnetic sources, allowed the development of nonlinear optics. Turning now to the quantum behavior associated with these phenomena, one should again make a distinction between linear and nonlinear setups. One first thinks of an electromagnetic wave propagating in vacuum space (or possibly in a linear medium): Already in 1930 Dirac [21] considered a weak electromagnetic beam and its associated photons; a device separates this beam into two partial beams, which are then made to interfere. One could then a priori think that two distinct photons possibly interfere. But, in such conditions, according to the general principles of quantum mechanics “each photon then interferes only with itself. Interference between two different photons can never occur ”. With respect to photon-based quantum communications addressed in the present paper, this entails that, if only considering communications through vacuum space or a linear medium, no entanglement is created in the transmission channel itself. This should be contrasted with the scenarios considered above, where entanglement is created by the channel itself (here with exchange coupling), whereas the original two-qubit state associated with Emitter 1 and Emitter 2 is unentangled.

The above manifestation of the superposition principle for photons, or the presence of (previously prepared) entangled states when more than one photon are implied may also be found in dielectrics, but other quantum phenomena may also be found in some optically nonlinear dielectric materials: (1) two intense laser beams at frequencies \(\omega _{1}\) and \( \omega _{2}\) may allow fluorescence at \(\omega _{1}+\omega _{2};\) there, two photons with respective frequencies \(\omega _{1}\) and \(\omega _{2}\) generate a photon with frequency \(\omega _{1}+\omega _{2}\ \) [4]. (2) spontaneous emission may sometimes allow emission at the difference frequency: The material receives a laser beam with frequency \(\omega _{p}\) (p: pump), and emits at both \(\omega \) (with a material dependent value) and \(\omega _{p}-\omega \) (so-called optical parametric fluorescence); here, a photon with frequency \(\omega _{p}\) generates two photons, with respective frequencies \(\omega \) and \(\omega _{p}-\omega \) [4]. Quantum communications take profit of the superposition principle, e.g. when involving two photons in an entangled state, and of the no-cloning theorem, both specific to quantum mechanics. Future quantum communication networks should make use of quantum teleportation—which allows transport of information, presently using entangled photon qubits—and of quantum repeaters interconnecting quantum nodes [23]. An eavesdropper, trying to access the information circulating within such a network could, e.g., try and operate by interacting with a repeater. Besides, one may imagine a scenario involving transmission through quantum channels, by means of photons, be they initially entangled or not, mainly with free propagation (linear medium), but now also with a nonlinear medium inserted (e.g., by an eavesdropper) in part of the overall transmission path forming what we called the “main channel” between Emitter 1 and Receiver 1 above. One might then investigate to which extent an “intruder,” composed of Emitter 2 and Receiver 2, would thus be able to interact with the main channel so as to extract information from it (eavesdropping), and to which extent Receiver 1 would be able to detect this intrusion. The relevance and attractiveness of this scenario need to be further investigated.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Deville, Y., Deville, A., Mansour, A. et al. Statistical intrusion detection and eavesdropping in quantum channels with coupling: multiple-preparation and single-preparation methods. Quantum Inf Process 21, 94 (2022). https://doi.org/10.1007/s11128-022-03436-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11128-022-03436-6

Keywords

Navigation