Skip to main content
Log in

Quantum context-aware recommendation systems based on tensor singular value decomposition

  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

In this paper, we propose a quantum algorithm for recommendation systems which incorporates the contextual information of users to the personalized recommendation. The preference information of users is encoded in a third-order tensor of dimension N which can be approximated by the truncated tensor singular value decomposition (t-svd) of the subsample tensor. Unlike the classical algorithm that reconstructs the approximated preference tensor using truncated t-svd, our quantum algorithm obtains the recommended product under certain context by measuring the output quantum state corresponding to an approximation of a user’s dynamic preferences. The algorithm achieves the time complexity \(\mathcal {O}(\sqrt{k}N\mathrm{polylog}(N))\), compared to the classical counterpart with complexity \(\mathcal {O}(kN^3)\), where k is the truncated tubal rank.

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

Notes

  1. Webscope v1.0, http://research.yahoo.com/.

  2. grouplens.org/datasets/movielens.

References

  1. Achlioptas, D., McSherry, F.: Fast computation of low-rank matrix approximations. J. ACM (JACM) 54(2), 9 (2007)

    Article  MathSciNet  Google Scholar 

  2. Arrazola, J. M., Delgado, A., Bardhan, B. R., Lloyd, S.: Quantum-inspired algorithms in practice. Quantum 4, 307 (2020)

  3. Biamonte, J., Wittek, P., Pancotti, N., Rebentrost, P., Wiebe, N., Lloyd, S.: Quantum machine learning. Nature 549(7671), 195–202 (2017)

    Article  ADS  Google Scholar 

  4. Biswas, A.K., Hasan, M.M., Chowdhury, A.R., Babu, H.M.H.: Efficient approaches for designing reversible binary coded decimal adders. Microelectron. J. 39(12), 1693–1703 (2008)

    Article  Google Scholar 

  5. Chiang, C.-F.: Quantum phase estimation with an arbitrary number of qubits. Int. J. Quantum Inf. 11(01), 1350008 (2013)

    Article  MathSciNet  Google Scholar 

  6. Clader, B.D., Jacobs, B.C., Sprouse, C.R.: Preconditioned quantum linear system algorithm. Phys. Rev. Lett. 110(25), 150502 (2013)

    Article  Google Scholar 

  7. Comon, P.: Tensor decompositions. Mathematics in Signal Processing V, pp. 1–24 (2002)

  8. De Lathauwer, L., De Moor, B., Vandewalle, J.: A multilinear singular value decomposition. SIAM J. Matrix Anal. Appl. 21(4), 1253–1278 (2000)

    Article  MathSciNet  Google Scholar 

  9. Ely, G., Aeron, S., Hao, N., Kilmer, M.E.: 5d seismic data completion and denoising using a novel class of tensor decompositions. Geophysics 80(4), V83–V95 (2015)

    Article  ADS  Google Scholar 

  10. Giovannetti, V., Lloyd, S., Maccone, L.: Quantum random access memory. Phys. Rev. Lett. 100(16), 160501 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  11. Gu, L., Wang, X., Zhang, G.: Quantum higher order singular value decomposition. In: 2019 IEEE International Conference on Systems, Man and Cybernetics (SMC), pp. 1166–1171 (2019)

  12. Hao, N., Kilmer, M., Braman, K., Hoover, R.: Facial recognition using tensor-tensor decompositions. SIAM J. Imaging Sci. [electronic only], 6, 02 (2013)

  13. Harrow, A.W., Hassidim, A., Lloyd, S.: Quantum algorithm for linear systems of equations. Phys. Rev. Lett. 103(15), 150502 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  14. Hu, S., Qi, L., Zhang, G.: Computing the geometric measure of entanglement of multipartite pure states by means of non-negative tensors. Phys. Rev. A 93(1), 012304 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  15. Huggins, W., Patil, P., Mitchell, B., Whaley, K.B., Stoudenmire, E.M.: Towards quantum machine learning with tensor networks. Quantum Sci. Technol. 4(2), 024001 (2019)

    Article  ADS  Google Scholar 

  16. Karatzoglou, A., Amatriain, X., Baltrunas, L., Oliver, N.: Multiverse recommendation: n-dimensional tensor factorization for context-aware collaborative filtering. In: Proceedings of the Fourth ACM Conference on Recommender Systems, pp. 79–86 (2010)

  17. Kerenidis, I., Landman, J., Luongo, A., Prakash, A.: q-means: a quantum algorithm for unsupervised machine learning. In: Advances in Neural Information Processing Systems, pp. 4134–4144 (2019)

  18. Kerenidis, I., Prakash, A.: Quantum recommendation systems. In: Papadimitriou, C.H. (ed.), 8th Innovations in Theoretical Computer Science Conference (ITCS 2017), Volume 67 of Leibniz International Proceedings in Informatics (LIPIcs), pp. 49:1–49:21, Dagstuhl, Germany, 2017. Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik

  19. Kilmer, M.E., Martin, C.D.: Factorization strategies for third-order tensors. Linear Algebra Appl. 435(3), 641–658 (2011)

    Article  MathSciNet  Google Scholar 

  20. Kolda, T.G., Bader, B.W.: Tensor decompositions and applications. SIAM Rev. 51(3), 455–500 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  21. Liu, C., Zhou, J., He, K.: Image compression based on truncated hosvd. In: 2009 International Conference on Information Engineering and Computer Science, pp. 1–4. IEEE (2009)

  22. Lloyd, S., Mohseni, M., Rebentrost, P.: Quantum principal component analysis. Nat. Phys. 10(9), 631 (2014)

    Article  Google Scholar 

  23. Long, G.: General quantum interference principle and duality computer. Commun. Theor. Phys. 45(5), 825 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  24. Long, G.: Duality quantum computing and duality quantum information processing. Int. J. Theor. Phys. 50(4), 1305–1318 (2011)

    Article  MathSciNet  Google Scholar 

  25. Ma, Y., Wang, Y., Tresp, V.: Quantum machine learning algorithm for knowledge graphs (2020). arXiv preprint arXiv:2001.01077

  26. Martin, C.D., Shafer, R., LaRue, B.: An order-\(p\) tensor factorization with applications in imaging. SIAM J. Sci. Comput. 35(1), A474–A490 (2013)

    Article  MathSciNet  Google Scholar 

  27. Nielsen, M.A., Chuang, I.: Quantum Computation and Quantum Information (2002)

  28. Orús, R.: A practical introduction to tensor networks: matrix product states and projected entangled pair states. Ann. Phys. 349, 117–158 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  29. Oseledets, I.V.: Tensor-train decomposition. SIAM J. Sci. Comput. 33(5), 2295–2317 (2011)

    Article  MathSciNet  Google Scholar 

  30. Preskill, J.: Quantum computing in the NISQ era and beyond. Quantum 2, 79 (2018)

    Article  Google Scholar 

  31. Qi, L., Chen, H., Chen, Y.: Tensor Eigenvalues and Their Applications, vol. 39. Springer, Berlin (2018)

    Book  Google Scholar 

  32. Qi, L., Luo, Z.: Tensor Analysis: Spectral Theory and Special Tensors, Volume 151. SIAM (2017)

  33. Qi, L., Zhang, G., Braun, D., Bohnet-Waldraff, F., Giraud, O.: Regularly decomposable tensors and classical spin states. Commun. Math. Sci. 15, 1651 (2017)

    Article  MathSciNet  Google Scholar 

  34. Qi, L., Zhang, G., Ni, G.: How entangled can a multi-party system possibly be? Phys. Lett. A 382(22), 1465–1471 (2018)

    Article  ADS  MathSciNet  Google Scholar 

  35. Rafailidis, D., Nanopoulos, A.: Modeling users preference dynamics and side information in recommender systems. IEEE Trans. Syst. Man Cybern. Syst. 46(6), 782–792 (2015)

    Article  Google Scholar 

  36. Rebentrost, P., Mohseni, M., Lloyd, S.: Quantum support vector machine for big data classification. Phys. Rev. Lett. 113(13), 130503 (2014)

    Article  ADS  Google Scholar 

  37. Rebentrost, P., Schuld, M., Wossnig, L., Petruccione, F., Lloyd, S.: Quantum gradient descent and Newton’s method for constrained polynomial optimization. New J. Phys. 21(7), 073023 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  38. Rebentrost, P., Steffens, A., Marvian, I., Lloyd, S.: Quantum singular-value decomposition of nonsparse low-rank matrices. Phys. Rev. A 97(1), 012327 (2018)

    Article  ADS  MathSciNet  Google Scholar 

  39. Rendle, S., Balby Marinho, L., Nanopoulos, A., Schmidt-Thieme, L.: Learning optimal ranking with tensor factorization for tag recommendation. In: Proceedings of the 15th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pp. 727–736 (2009)

  40. Rendle, S., Schmidt-Thieme, L.: Pairwise interaction tensor factorization for personalized tag recommendation. In: Proceedings of the Third ACM International Conference on Web Search and Data Mining, pp. 81–90 (2010)

  41. Ross, S.M.: Introduction to Probability Models. Academic Press, ISE, Cambridge (2006)

    MATH  Google Scholar 

  42. Sarma, R., Jain, R.: Quantum gate implementation of a novel reversible half adder and subtractor circuit. In: 2018 International Conference on Intelligent Circuits and Systems (ICICS), pp. 72–76. IEEE (2018)

  43. Shao, C.: From linear combination of quantum states to Grover’s searching algorithm (2018). arXiv preprint arXiv:1807.09693

  44. Shao, C., Li, Y., Li, H.: Quantum algorithm design: techniques and applications. J. Syst. Sci. Complex. 32(1), 375–452 (2019)

    Article  MathSciNet  Google Scholar 

  45. Song, G., Ng, M.K., Zhang, X.: Robust tensor completion using transformed tensor singular value decomposition. Numer. Linear Algebra Appl. 27(3), e2299 (2020)

    Article  MathSciNet  Google Scholar 

  46. Symeonidis, P., Nanopoulos, A., Manolopoulos, Y.: Tag recommendations based on tensor dimensionality reduction. In: Proceedings of the 2008 ACM Conference on Recommender Systems, pp. 43–50 (2008)

  47. Tang, E.: A quantum-inspired classical algorithm for recommendation systems. In: Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, pp. 217–228 (2019)

  48. Teixeira, M., Rodriguez, D.: A class of fast cyclic convolution algorithms based on block pseudocirculants. IEEE Signal Process. Lett. 2(5), 92–94 (1995)

    Article  ADS  Google Scholar 

  49. Thapliyal, H., Ranganathan, N.: Design of reversible sequential circuits optimizing quantum cost, delay, and garbage outputs. ACM J. Emerg. Technol. Comput. Syst. (JETC) 6(4), 1–31 (2010)

    Article  Google Scholar 

  50. Thapliyal, H., Ranganathan, N.: A new design of the reversible subtractor circuit. In: 2011 11th IEEE International Conference on Nanotechnology, pp. 1430–1435. IEEE (2011)

  51. Tucker, L.R.: Some mathematical notes on three-mode factor analysis. Psychometrika 31(3), 279–311 (1966)

    Article  MathSciNet  Google Scholar 

  52. Wang, C., Wossnig, L.: A quantum algorithm for simulating non-sparse Hamiltonians. Quantum Inf. Comput. 20, 597–615 (2020)

  53. Wu, M., He, S., Zhang, Y., Chen, J., Sun, Y., Liu, Y.-Y., Zhang, J., Poor, H.V.: A tensor-based framework for studying eigenvector multicentrality in multilayer networks. Proc. Natl. Acad. Sci. 116(31), 15407–15413 (2019)

    Article  ADS  Google Scholar 

  54. Xiong, L., Chen, X., Huang, T.-K., Schneider, J., Carbonell, J.G.: Temporal collaborative filtering with Bayesian probabilistic tensor factorization. In: Proceedings of the 2010 SIAM International Conference on Data Mining, pp. 211–222. SIAM (2010)

  55. Zhang, G.: Dynamical analysis of quantum linear systems driven by multi-channel multi-photon states. Automatica 83, 186–198 (2017)

    Article  MathSciNet  Google Scholar 

  56. Zhang, J., Saibaba, A.K., Kilmer, M.E., Aeron, S.: A randomized tensor singular value decomposition based on the t-product. Numer. Linear Algebra Appl. 25(5), e2179 (2018)

    Article  MathSciNet  Google Scholar 

  57. Zhang, M., Ni, G., Zhang, G.: Iterative methods for computing U-eigenvalues of non-symmetric complex tensors with application in quantum entanglement. Comput. Optim. Appl. 75, 779–798 (2020)

    Article  ADS  MathSciNet  Google Scholar 

  58. Zhang, Z., Aeron, S.: Exact tensor completion using t-svd. IEEE Trans. Signal Process. 65(6), 1511–1526 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  59. Zhang, Z., Ely, G., Aeron, S., Hao, N., Kilmer, M.: Novel methods for multilinear data completion and de-noising based on tensor-svd. In: Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, pp. 3842–3849 (2014)

  60. Zhou, P., Lu, C., Lin, Z., Zhang, C.: Tensor factorization for low-rank tensor completion. IEEE Trans. Image Process. 27(3), 1152–1163 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  61. Zhou, X.-Q., Ralph, T.C., Kalasuwan, P., Zhang, M., Peruzzo, A., Lanyon, B.P., O’Brien, J.L.: Adding control to arbitrary unknown quantum operations. Nat. Commun. 2(1), 1–8 (2011)

    Google Scholar 

Download references

Acknowledgements

This research is supported in part by Hong Kong Research Grant council (RGC) Grants (No. 15208418, No. 15203619, No. 15506619) and Shenzhen Fundamental Research Fund, China, under Grant No. JCYJ20190813165207290.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Guofeng Zhang.

Additional information

Publisher's Note

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

Appendices

A The proof of Theorem 4

Before proving Theorem 4, we would like to sketch the proof first. According to Assumption 1, each frontal slice \(T^{(k)}\) is stored in the binary tree structure. Hence, based on the proof of Lemma 3 in [18], the states \(| \mathcal {T}(i,:,k) \rangle \), corresponding to the i-th row of \(T^{(k)}\), can be prepared efficiently by operators \(P_k\). Based on these operators, two new isometries \(\hat{P}_m\) and \(\hat{Q}_m\) are constructed in order to perform QSVE on \(\hat{T}^{(m)}\). The detail of the proof is given below.

Proof

Since every \(T^{(k)}\), \(k=0,\ldots ,N-1\), is stored in the binary tree structure, the quantum computer can perform the following mapping in \(\mathcal {O}(\mathrm{polylog}(N))\) time, as shown in Theorem 5.1 in [18]:

$$\begin{aligned}&U_{P_k}: | i \rangle | 0 \rangle \nonumber \\&\quad \rightarrow | i \rangle | \mathcal {T}(i,:,k) \rangle =\frac{1}{||\mathcal {T}(i,:,k)||_2}\sum _{j=0}^{N-1}\mathcal {T}_{ijk}| i \rangle | j \rangle , \end{aligned}$$
(29)

where \(\mathcal {T}(i,:,k)\) is the i-th row of \(T^{(k)}\).

Define the degenerate operator \(P_k \in \mathbb {R}^{N^2 \times N}\) related to \(U_{P_k}\) as

$$\begin{aligned} P_k: | i \rangle \rightarrow | i \rangle | \mathcal {T}(i,:,k) \rangle . \end{aligned}$$
(30)

That is,

$$\begin{aligned} P_k=\sum _{i=0}^{N-1}| i \rangle | \mathcal {T}(i,:,k) \rangle |\langle i |. \end{aligned}$$
(31)

Based on the efficiently implemented operator \(U_{P_k}\), we define another operator

$$\begin{aligned} U_{\hat{P}_m}\triangleq \frac{1}{\sqrt{N}}\sum _{k=0}^{N-1}\omega ^{km}U_{P_k.} \end{aligned}$$
(32)

It can be easily seen that the operator \(U_{\hat{P}_m}\) achieves the state preparation of the rows of the matrix \(\hat{T}^{(m)}\), i.e., \(U_{\hat{P}_m}=\frac{1}{\sqrt{N}}\sum _{k=0}^{N-1}\sum _{i=0}^{N-1}\omega ^{km}| i \rangle | \mathcal {T}(i,:,k) \rangle |\langle i ||\langle 0 |=\sum _{i}| i \rangle | \hat{\mathcal {T}}(i,:,m) \rangle |\langle i ||\langle 0 |\), where \(| \hat{\mathcal {T}}(i,:,m) \rangle \) is the state of the i-th row of \(\hat{T}^{(m)}\). Similarly, the isometry corresponding to \(U_{\hat{P}_m}\) is \(\hat{P}_m=\sum _{i}| i \rangle | \hat{\mathcal {T}}(i,:,m) \rangle |\langle i |\). It can be easily shown that \(\hat{P}_m\) is an isometry since \(\hat{P}_m^{\dagger }\hat{P}_m=I_{N}\).

Since \(U_{P_k}\) can be implemented in time \(\mathcal {O}(\mathrm{polylog}(N))\), \(U_{\hat{P}_m}\) can be implemented in time \(\mathcal {O}(N\mathrm{polylog}N)\) using the linear combination of unitaries (LCU) technique [6, 23, 24, 43, 44].

The LCU technique was first proposed by Long in their work [23] in a more general form, and Shao summarized this result in [44]. The problem of LCU can be formulated as follows: Given \(\alpha _i \in \mathbb {C}\) and unitary operators \(U_i\), \(i = 0, 1, \ldots , N-1\), implement linear operator \(L=\sum _{j=0}^{N-1}\alpha _jU_j\). The algorithm stated in [6] implements L in time \(\mathcal {O}((T_\mathrm{in} +\mathrm{log}N)N\max _j |\alpha _j|/||L| \psi \rangle ||)\), where \(| \psi \rangle \) is any given initial state and \(T_\mathrm{in}\) is the time to implement \(U_0, U_1, \ldots , U_{N-1}\). In our case, \(T_\mathrm{in}=N\mathrm{polylog}N\), \(\alpha _j=\omega ^{km}/\sqrt{N}\) and the input state is chosen as \(| \psi \rangle =\sum _{i=0}^{N}| i \rangle | 0 \rangle \). Thus, \(||L| \psi \rangle ||=||\frac{1}{\sqrt{N}}\sum _{k=0}^{N-1}\omega ^{km}U_{P_k}| \psi \rangle ||=||\frac{1}{\sqrt{N}}\sum _{k=0}^{N-1}\sum _{i=0}^{N-1}\omega ^{km}| i \rangle | \mathcal {T}(i,:,k) \rangle ||=\sqrt{N}\). Therefore, the complexity to implement \(U_{\hat{P}_m}\) is \(\mathcal {O}(N\mathrm{polylog}N)\).

Next, we define the mapping

$$\begin{aligned}&U_{\hat{Q}_m}: | 0 \rangle | j \rangle \rightarrow | \varvec{s}_{\hat{T}^{(m)}} \rangle | j \rangle \nonumber \\&\quad =\frac{1}{||\hat{T}^{(m)}||_F}\sum _{i}||\hat{\mathcal {T}}(i,:,m)||_2| i \rangle | j \rangle , \end{aligned}$$
(33)

where \(\varvec{s}_{\hat{T}^{(m)}}\) is a vector whose i-th entry is \(\frac{||\hat{\mathcal {T}}(i,:,m)||}{||\hat{T}^{(m)}||_F}\). Similar with \(U_{\hat{P}_m}\), the corresponding isometry is defined as \(\hat{Q}_m=\sum _{j}| \varvec{s}_{\hat{T}^{(m)}} \rangle | j \rangle |\langle j |\) satisfying \(\hat{Q}_m^{\dagger }\hat{Q}_m=I_{N}\), which can be verified easily.

Now we perform QSVE on the matrix \(\hat{T}^{(m)}\). First, the factorization \(\frac{\hat{T}^{(m)}}{||\hat{T}^{(m)}||_F}=\hat{P}_m^{\dagger }\hat{Q}_m\) can be easily verified.

Second, we can prove that \(2\hat{P}_m\hat{P }_m^{\dagger }-I_{N^2}\) is a reflection and it be implemented through \(U_{\hat{P}_m}\). Actually,

$$\begin{aligned}&2\hat{P}_m\hat{P }_m^{\dagger }-I_{N^2} \nonumber \\&\quad =2\sum _{i}| i \rangle | \hat{\mathcal {T}}(i,:,m) \rangle |\langle i ||\langle \hat{\mathcal {T}}(i,:,m) |-I_{N^2} \nonumber \\&\quad =U_{\hat{P}_m}\left[ 2\sum _{i}| i \rangle | 0 \rangle |\langle i ||\langle 0 |-I_{N^2} \right] U_{\hat{P}_m}^{\dagger }, \end{aligned}$$
(34)

where \(2\sum _{i}| i \rangle | 0 \rangle |\langle i ||\langle 0 |-I_{N^2}\) is a reflection. The similar result holds for \(2\hat{Q}_m\hat{Q}_m^{\dagger }-I_{N^2}\).

Now denote

$$\begin{aligned} W_m=\left( 2\hat{P}_m\hat{P}_m^{\dagger }-I_{N^2}\right) \left( 2\hat{Q}_m\hat{Q}_m^{\dagger }-I_{N^2}\right) . \end{aligned}$$
(35)

Let \(\hat{T}^{(m)} = \sum _{i=0}^{r-1}\hat{\sigma }_i^{(m)}\hat{u}_i^{(m)}\hat{v}_i^{(m)\dagger }\) be the singular value decomposition of \(\hat{T}^{(m)}\). We can prove that the subspace spanned by \(\{\hat{Q}_m| \hat{v}_i^{(m)} \rangle , \hat{P}_m| \hat{u}_i^{(m)} \rangle \}\) is invariant under the unitary transformation \(W_m\):

$$\begin{aligned}&W_m\hat{Q}_m| \hat{v}_i^{(m)} \rangle =\frac{2\hat{\sigma }_i^{(m)}}{||\hat{T}^{(m)}||_F}\hat{P}_m| \hat{u}_i^{(m)} \rangle -Q| \hat{v}_i^{(m)} \rangle , \\&W_m\hat{P}_m| \hat{u}_i^{(m)} \rangle = \\&\quad \left( \frac{4\hat{\sigma }_{i}^2}{||\hat{T}^{(m)}||_F}-1 \right) \hat{P}_m| \hat{u}_i^{(m)} \rangle -\frac{2\hat{\sigma }_{i}}{||\hat{T}^{(m)}||_F}\hat{Q}_m| \hat{v}_i^{(m)} \rangle . \end{aligned}$$

The matrix \(W_m\) can be calculated under an orthonormal basis using the Schmidt orthogonalization. It is a rotation in the subspace spanned by its eigenvectors \(| \omega ^{(m)}_{i \pm } \rangle \) with correspondent eigenvalues \(\mathrm{e}^{\pm i\theta _i^{(m)}}\), where \(\theta _i^{(m)}\) is the rotation angle satisfying

$$\begin{aligned} \cos (\theta _i^{(m)}/2)=\frac{\hat{\sigma }_i^{(m)}}{||\hat{T}^{(m)}||_F}, \end{aligned}$$
(36)

that is,

$$\begin{aligned}&\hat{Q}_m| \hat{v}_i^{(m)} \rangle =\sqrt{2}\left( | \omega ^{(m)}_{i +} \rangle +| \omega ^{(m)}_{i -} \rangle \right) , \\&\hat{P}_m| \hat{u}_i^{(m)} \rangle =\sqrt{2}\left( \mathrm{e}^{ i\theta _i/2}| \omega ^{(m)}_{i +} \rangle +\mathrm{e}^{ -i\theta _i/2}| \omega ^{(m)}_{i -} \rangle \right) . \end{aligned}$$

Here, we choose the input state \(| \hat{\mathcal {T}}(i,:,m) \rangle \) represented in (12), then

$$\begin{aligned}&\hat{Q}_m| \hat{\mathcal {T}}(i,:,m) \rangle =\sum _{j=0}^{N-1}\sqrt{2}\beta _{j}^{(im)}\left( | \omega ^{(m)}_{i +} \rangle +| \omega ^{(m)}_{i -} \rangle \right) . \end{aligned}$$
(37)

We perform the phase estimation on \(W_m\) with running time \(\mathcal {O}\left( N\mathrm{polylog}N/{\epsilon _\mathrm{SVE}^{(m)}}\right) \) and compute the estimated singular value of \(\hat{T}^{(m)}\) through oracle with a computable function \(f(x)=||\hat{T}^{(m)}||_F\cos (x/2)\). According to the relations between \(\theta _i^{(m)}\) and \(\hat{\sigma }_i^{(m)}\) in (36), we obtain

$$\begin{aligned} \sum _{j=0}^{N-1}\sqrt{2}\beta _{j}^{(im)}\left( | \omega ^{(m)}_{i +} \rangle | \overline{\theta }_i^{(m)} \rangle + | \omega ^{(m)}_{i -} \rangle | -\overline{\theta }_i^{(m)} \rangle \right) | \overline{\sigma }_{i}^{(m)} \rangle . \end{aligned}$$
(38)

We next uncompute the phase estimation procedure and then apply the inverse of \(U_{\hat{Q}_m}\) to obtain the desired state (13) in Theorem 4. \(\square \)

B The proof of Lemma 5

Proof

Let \(\sigma _i\) denote the singular value of A and l be the largest integer for which \(\sigma _{l} \ge \frac{\epsilon ||A||_F}{\sqrt{k}}.\) By the triangle inequality, \(||A-A_{\ge \sigma }||_F \le ||A-A_k||_F +||A_k-A_{\ge \sigma }||_F.\) If \(k \le l\), it’s easy to conclude that \(||A_k-A_{\ge \sigma }||_F \le ||A-A_k||_F \le \epsilon ||A||_F\). If \(k > l\), \(||A_k-A_{\ge \sigma }||_F^2=\sum _{i=l+1}^{k}\sigma _i^2 \le k\sigma _{l+1}^2 \le k\left( \frac{\epsilon ||A||_F}{\sqrt{k}}\right) ^2 \le \left( \epsilon ||A||_F\right) ^2\). In either case, we have \(||A-A_{\ge \sigma }||_F \le 2\epsilon ||A||_F\). \(\square \)

The proof of Theorem 5

Proof

Based on Lemma 5 in the main text, if the best rank-k approximation satisfies \(||\hat{T}^{(m)}-\hat{T}^{(m)}_{k}||_F \le \epsilon ^{(m)}||\hat{T}^{(m)}||_F, \) then

$$\begin{aligned} ||\hat{T}^{(m)}-\hat{T}^{(m)}_{\ge \tau _m}||_F \le 2 \epsilon ^{(m)}||\hat{T}^{(m)}||_F \le \epsilon _0||\hat{T}^{(m)}||_F \end{aligned}$$
(39)

for \(m=0, \ldots , N-1.\) By summing both sides of (39) over m, we get

$$\begin{aligned} ||\hat{\mathcal {T}}-\hat{\mathcal {T}}_{\ge \varvec{\tau }}||_F^2 = \sum _{m=0}^{N-1}||\hat{T}^{(m)}-\hat{T}^{(m)}_{\ge \tau _m}||_F^2 \le \epsilon _0^2||\hat{\mathcal {T}}||_F^2. \end{aligned}$$
(40)

Since the inverse QFT along the third mode of the tensor \(\mathcal {T}\) cannot change the Frobenius norm of its horizontal slice, (40) can be rewritten as

$$\begin{aligned} ||\mathcal {T}-\mathcal {T}_{\ge \varvec{\tau }}||_F^2 \le \epsilon _0^2||\mathcal {T}||_F^2. \end{aligned}$$
(41)

Moreover, noticing that \(||\mathcal {T}-\mathcal {T}_{\ge \varvec{\tau }}||_F^2=\sum _{i=0}^{N-1}||\mathcal {T}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F^2\), we have \({E}\left( ||\mathcal {T}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F^2\right) \le \frac{\epsilon _0^2||\mathcal {T}||_F^2}{N}.\) Due to Markov’s Inequality [41, Proposition 2.6], for \(\delta \in (0,1)\),

$$\begin{aligned}&\Pr \left( ||\mathcal {T}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F^2 > \frac{\epsilon _0^2||\mathcal {T}||_F^2}{\delta N}\right) \nonumber \\&\quad \le \frac{{E}\left( ||\mathcal {T}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F^2\right) \delta N}{\epsilon _0^2||\mathcal {T}||_F^2} \le \delta \end{aligned}$$
(42)

holds. This means at least \((1-\delta )N\) users i satisfy

$$\begin{aligned} ||\mathcal {T}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F^2 \le \frac{\epsilon _0^2||\mathcal {T}||_F^2}{\delta N}. \end{aligned}$$
(43)

Notice \({E}\left( ||\mathcal {T}||_F^2\right) =||\mathcal {A}||_F^2/p\). Using the Chernoff bound, we have \(\Pr \left( ||\mathcal {T}||_F^2 >(1+\theta )||\mathcal {A}||_F^2/p\right) \le \mathrm{e}^{-\theta ^2||\mathcal {A}||_F^2/3p}\) for \(\theta \in [0,1]\), which is exponentially small. Here, we choose \(\theta =1/10\), then \(||\mathcal {T}||_F^2 \le 11||\mathcal {A}||_F^2/10p\).

Based on the second assumption in Assumption 1, we sum both sides of (6) for m and i, respectively, obtaining

$$\begin{aligned} \frac{1}{1+\gamma }\frac{||\mathcal {A}||_F^2}{N} \le ||\mathcal {A}(i,:,:)||_F^2 \le (1+\gamma )\frac{||\mathcal {A}||_F^2}{N}, \end{aligned}$$
(44)

and

$$\begin{aligned} \frac{1}{1+\gamma }\frac{||\mathcal {A}||_F^2}{N} \le ||A^{(m)}||_F^2 \le (1+\gamma )\frac{||\mathcal {A}||_F^2}{N}. \end{aligned}$$
(45)

Then, (43) becomes

$$\begin{aligned} ||\mathcal {T}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F^2 \le \frac{11\epsilon _0^2(1+\gamma )}{10\delta p}||\mathcal {A}(i,:,:)||_F^2. \end{aligned}$$
(46)

Meanwhile, since

$$\begin{aligned} {E}\left( ||\mathcal {A}(i,:,:)-\mathcal {T}(i,:,:)||_F^2\right) =\left( \frac{1}{p}-1\right) ||\mathcal {A}(i,:,:)||_F^2, \end{aligned}$$

then

$$\begin{aligned}&\Pr \left( ||\mathcal {A}(i,:,:)-\mathcal {T}(i,:,:)||_F^2 > \nu ||\mathcal {A}(i,:,:)||_F^2\right) \nonumber \\&\quad \le \mathrm{e}^{-\zeta ^2\left( \frac{1}{p}-1\right) \frac{||\mathcal {A}||_F^2}{3N(1+\gamma )}}, \end{aligned}$$
(47)

where \(\nu =(1+\zeta )\left( \frac{1}{p}-1\right) \) and \(\zeta \in (0,1)\). This means with probability at least \(p_1\),

$$\begin{aligned} ||\mathcal {A}(i,:,:)-\mathcal {T}(i,:,:)||_F^2 \le \nu ||\mathcal {A}(i,:,:)||_F^2. \end{aligned}$$
(48)

Combining (46) and (48) together and by triangle inequality, we obtain

$$\begin{aligned} ||\mathcal {A}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F \le \epsilon ||\mathcal {A}(i,:,:)||_F. \end{aligned}$$
(49)

According to Lemma 4, the probability that sampling according to \(\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)\) provides a bad recommendation is

$$\begin{aligned} \Pr \limits _{t\sim \mathcal {U}_N, j\sim \mathcal {T}_{\ge \varvec{\tau }}(i,:,t)} [(i,j,t) \mathrm{bad}] \le \left( \frac{\epsilon }{1-\epsilon }\right) ^2. \end{aligned}$$
(50)

\(\square \)

D. The proof of Theorem 6

Proof

According to Theorem 4, the modified QSVE algorithm performed on the frontal slice \(\hat{T}^{(m)}\) takes time \(\mathcal {O}\left( N\mathrm{polylog}(N)/{\epsilon _\mathrm{SVE}^{(m)}}\right) \).

In Step 5 of Algorithm 2, we need to repeat the measurement \(\mathcal {O}\left( \frac{||\mathcal {T}(i,:,:)||_F}{||\hat{\mathcal {T}}_{\ge \varvec{\tau }}(i,:,:)||_F}\right) \) times in order to ensure the probability of getting the outcome \(| 0 \rangle \) in this step is close to 1. For most users, we can prove that \(\frac{||\mathcal {T}(i,:,:)||_F}{||\hat{\mathcal {T}}_{\ge \varvec{\tau }}(i,:,:)||_F}\) is bounded and the upper bound is a constant for appropriate parameters. The proof is in the following.

Since \({E}\left( ||\mathcal {T}(i,:,:)||_F^2\right) =\frac{||\mathcal {T}(i,:,:)||_F^2}{p} \le (1+\gamma )\frac{||\mathcal {A}||_F^2}{pN},\) then by Chernoff bound,

$$\begin{aligned} ||\mathcal {T}(i,:,:)||_F^2 \le \frac{2(1+\gamma )||\mathcal {A}||_F^2}{pN} \end{aligned}$$
(51)

holds with probability close to 1. Moreover, by Theorem 5, there are at least \((1-\delta )N\) users satisfying \(||\mathcal {A}(i,:,:)-\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F\le \epsilon ||\mathcal {A}(i,:,:)||_F\), then \((1+\epsilon )||\mathcal {A}(i,:,:)||_F \le ||\mathcal {T}_{\ge \varvec{\tau }}(i,:,:)||_F \le (1+\epsilon )||\mathcal {A}(i,:,:)||_F.\) Since the Frobenius norm is unchanged under the Fourier transform, we get

$$\begin{aligned} (1+\epsilon )||\hat{T}_{(i)}||_F \le ||\hat{\mathcal {T}}_{\ge \varvec{\tau }}(i,:,:)||_F \le (1+\epsilon )||\hat{T}_{(i)}||_F. \end{aligned}$$
(52)

Therefore,

$$\begin{aligned} ||\hat{\mathcal {T}}_{\ge \varvec{\tau }}(i,:,:)||_F^2 \ge (1+\epsilon )^2||\hat{T}_{(i)}||_F^2\ge \frac{(1+\epsilon )^2}{1+\gamma }\frac{||\mathcal {A}||_F^2}{N}. \end{aligned}$$
(53)

Combining (51) and (53) together, we can conclude that for at least \((1-\delta )N\) users, \(\frac{||\mathcal {T}(i,:,:)||_F}{||\hat{\mathcal {T}}_{\ge \varvec{\tau }}(i,:,:)||_F}\) is bounded, that is,

$$\begin{aligned} \frac{||\mathcal {T}(i,:,:)||_F}{||\hat{\mathcal {T}}_{\ge \varvec{\tau }}(i,:,:)||_F} \le \left( \frac{(1+\gamma )\frac{2||\mathcal {A}||_F^2}{pN}}{\frac{(1+\epsilon )^2}{1+\gamma }\frac{||\mathcal {A}||_F^2}{N}}\right) ^{1/2} = \frac{\sqrt{2}(1+\gamma )}{(1+\epsilon )\sqrt{p}}. \end{aligned}$$
(54)

The precision for the singular value estimation algorithm on the matrix \(||\hat{T}^{(m)}||_F\) can be chosen as \(\epsilon _\mathrm{SVE}^{(m)}=\frac{\tau _m}{||\hat{T}^{(m)}||_F}.\) Therefore, the total time complexity of Algorithm 2 is

$$\begin{aligned}&({\mathrm{log}N})^4 \cdot \frac{N\mathrm{polylog}(N)}{\min \limits _{m}{\epsilon _\mathrm{SVE}^{(m)}}} \cdot \frac{||\mathcal {T}(i,:,:)||_F}{||\hat{\mathcal {T}}_{\ge \varvec{\tau }}(i,:,:)||_F} \\&\quad \le ({\mathrm{log}N})^4{N\mathrm polylog}(N) \max \limits _{m}{\frac{||\hat{T}^{(m)}||_F}{\tau _m}}\cdot \frac{\sqrt{2}(1+\gamma )}{(1+\epsilon )\sqrt{p}} \\&\quad \approxeq \frac{\sqrt{k}N\mathrm{polylog}(N) (1+\gamma )}{\min \limits _{m}{\epsilon ^{(m)}(1+\epsilon )\sqrt{p}}}, \end{aligned}$$

where \(\epsilon ^{(m)}\) and \(\epsilon \) are defined in Theorem 5. \(\square \)

E The quantum cost

Defining the cost of quantum circuits is not an easy task due to the fact that each quantum computer model may have a different cost for a given quantum gate. Here, the quantum cost of a reversible gate is defined to be the number of \(1\times 1\) and \(2\times 2\) reversible gates or quantum logic gates required in its design. The quantum costs of all reversible \(1\times 1\) and \(2\times 2\) gates are taken as unity [4, 42, 49, 50]. The cost of a circuit is calculated by summing up the costs of the gates composing the circuit. First, we analyze the quantum cost of QFT and QSVE circuits and then we analyze the cost of each gate in Fig. 3.

1.1 E.1 The cost of QFT

The QFT under an orthonormal basis \(|x\rangle \in \{|0\rangle , \ldots ,|N-1\rangle \}\) is defined as the linear operator with the following action on the basis vectors:

$$\begin{aligned} \mathrm {QFT}:|x\rangle \rightarrow \frac{1}{\sqrt{N}} \sum _{k=0}^{N-1} \omega ^{x \cdot k}|k\rangle , \end{aligned}$$

where \(\omega =\mathrm{e}^{\frac{2\pi \mathrm {i}}{N}}\). The inverse QFT is then defined as

$$\begin{aligned} \mathrm {QFT}^{\dagger }:|k\rangle \rightarrow \frac{1}{\sqrt{N}} \sum _{x=0}^{N-1} \omega ^{-k \cdot x}|x\rangle . \end{aligned}$$

The circuit of QFT, shown in Fig. 7, is composed of a total number of \(\mathcal {O}\left( \lceil \mathrm{log}N \rceil ^2\right) \) H gates, CNOT gates and controlled phase gate \(R_{m}\) (see [27, Section 5.1]), where

$$\begin{aligned} H=\frac{1}{\sqrt{2}}\left( \begin{array}{cc} 1 &{} 1 \\ 1 &{} -1 \end{array}\right) , \quad \quad R_{m}=\left( \begin{array}{cc} 1 &{} 0 \\ 0 &{} \mathrm{e}^{\frac{2 \pi \mathrm {i}}{2^{m}}} \end{array}\right) \end{aligned}$$

with \(m=2, \ldots , n\), \(n=\lceil \mathrm{log}N \rceil \).

Fig. 7
figure 7

The circuit of QFT

1.2 E.2 The cost of QSVE

The QSVE algorithm, originally proposed in [18], is introduced in the preliminary part of the manuscript. The QSVE algorithm and its circuit are, respectively, given in Algorithm 3 and Fig. 8. In what follows, we focus on the quantum cost of this algorithm.

Let \(A \in \mathbb {R}^{N \times N}\) be a matrix with singular value decomposition \(A=\sum _{i} \sigma _{i} u_{i} v_{i}^\mathrm{T}\) stored in the binary tree data structure introduced in Lemma 2. In Step 2 of Algorithm 3, the unitary \(U_Q\) corresponding to Q is defined as

$$\begin{aligned} U_Q: \left| 0 \right\rangle \left| j \right\rangle \rightarrow |\widetilde{A}, j\rangle =\frac{1}{\Vert A\Vert _{F}} \sum _{i \in [N]}\left\| A_{i}\right\| |i, j\rangle \end{aligned}$$
(55)

for \(j =0,\ldots ,N-1\), where \(\widetilde{A}\) is a vector whose i-th row is \(\widetilde{A}_{i}=\left\| A_{i}\right\| \) for \(i \in [N].\) Since \(\widetilde{A}\) is stored in a classical binary tree with depth \(\lceil \mathrm{log}N \rceil \), \(U_Q\) can be implemented by performing \(\lceil \mathrm{log}N \rceil \) controlled rotations on \(| 0 \rangle ^{\otimes \lceil \mathrm{log}N \rceil }\) (see [18, Appendix A]). To implement each controlled rotation \(\sum _{\tilde{\theta } \in \{0,1\}^{\otimes \lceil \mathrm{log}N \rceil }}|\widetilde{\theta }\rangle \langle \widetilde{\theta }| \otimes \mathrm{e}^{-\mathrm {i} Y \tilde{\theta }}\) with Pauli Y matrix \(Y=\begin{bmatrix} 0 &{} -\mathrm {i} \\ \mathrm {i} &{} 0 \\ \end{bmatrix}\), we use one rotation controlled on each qubit of the first register which can be implemented with cost \({{\mathcal {O}}}(\lceil \mathrm{log}N \rceil )\) (see [52, Lemma 2]). Therefore, the cost of implementing \(U_Q\) in Step 2 is \(\sum _{\lceil \mathrm{log}N \rceil }{\lceil \mathrm{log}N \rceil }^2\), i.e., \(\mathcal {O}({\lceil \mathrm{log}N \rceil }^3)\).

figure c
Fig. 8
figure 8

The circuit of QSVE algorithm. \(U_{f}\) is a unitary operator implemented through oracle with a computable function \(f(x)=||A||_F\cos (x/2)\). \(w=\mathrm{log}N-1\)

In Step 3, we can prove that the quantum cost of implementing the unitary W is \(\mathcal {O}(N{\lceil \mathrm{log}N \rceil }^3)\). To be specific, notice that \(W=U \cdot V,\) where \(V=2 Q Q^\mathrm{T}-I_{N^2}\), \(U=2 P P^\mathrm{T}-I_{N^2}=U_{P}\left( 2 \sum _{i}|i\rangle |0\rangle \langle i|\langle 0|-I_{N^2}\right) U_{P}^{\dagger }\), and \(U_P: \left| i\right\rangle \left| 0\right\rangle \rightarrow \left| i, A_{i}\right\rangle \) for \(i \in [N]\).

The unitary operator \(2 \sum _{i}|i\rangle |0\rangle \langle i|\langle 0|-I_{N^2}\) can be realized by the circuit shown in Fig. 9. Since each row of the matrix A is stored in the classical binary tree structure, \(U_P\) can be implemented with cost \(\mathcal {O}(N{\lceil \mathrm{log}N \rceil }^3)\), as compared to the cost \(\mathcal {O}({\lceil \mathrm{log}N \rceil }^3)\) of implementing \(U_Q\). To summarize, a total cost of \(\mathcal {O}(N{\lceil \mathrm{log}N \rceil }^3)\) is required to implement W if we ignore less significant cost. Next, we perform phase estimation on W. When the singular values of W is precise to the \(\lceil \mathrm{log}N \rceil \)-th bit, the total number of rotation gate invocations is \(\mathcal {O}\left( \lceil \mathrm{log}N \rceil ^2\right) \) (See [5, Section 5.2]). The sequence of controlled-\(W^{2^j}\), \(j=0, \ldots , \lceil \mathrm{log}N \rceil -1\) operations in the phase estimation procedure can be implemented using \(\mathcal {O}(\lceil \mathrm{log}N \rceil ^3)\) gates by the modular exponentiation technique (see [27, Box 5.2]). Thus, the cost of Step 3 is \(\mathcal {O}(N{\lceil \mathrm{log}N \rceil }^3)\), which dominates the cost of QSVE algorithm.

Fig. 9
figure 9

Quantum circuit to run \(2 \sum _{i}| i \rangle ^{\otimes n}| 0 \rangle |\langle i |^{\otimes n}\langle 0|-I^{\otimes (n+1)}\) for the input state with \(n+1\) qubits. The block Z is the Pauli Z gate

To summarize, the quantum cost of QSVE is \(\mathcal {O}(N{\lceil \mathrm{log}N \rceil }^3)\).

1.3 E.3 The cost of the circuit in Fig. 3 for Algorithm 2

As discussed in Appendix E.1, the circuit of QFT in Step 1 of Algorithm 2 is composed of \(\mathcal {O}(\lceil \mathrm{log}N \rceil ^2)\) H gates, CNOT gates and two-qubit controlled phase gates.

In Step 2, we give a rough estimation of the cost of the controlled-unitary operator U defined in (10). Current standard methods for realizing controlled-unitary gates rely on the decomposition of \(U_\mathrm{SVE}^{(m)}\) into a set of \(1\times 1\) and \(2\times 2\) reversible gates. Specifically, based on the method proposed in [61], the number of additional operations required to add a control to each \(U_{\mathrm {SVE}}^{(m)}\) will generally be far less than the cost of the constructed \(U_\mathrm{SVE}^{(m)}\), so we next focus on estimating the cost of achieving \(U_\mathrm{SVE}^{(m)}, m =0,\ldots ,N-1.\) The most costly step of this process lies in constructing the operator \(U_{\hat{P}_m}\) in (32), \(m =0,\ldots ,N-1\). According to the analysis in the paragraph above (33), implementing all \(U_{P_k}\) in (29) for \(k=0, \ldots , N-1\) occupies the major cost of achieving \(U_{\hat{P}_m}, m =0,\ldots ,N-1\), which takes a total of \(\mathcal {O}(N^2{\lceil \mathrm{log}N \rceil }^3)\) \(1\times 1\) and \(2\times 2\) reversible gates, since the cost of each \(U_{P_k}\) is \(\mathcal {O}(N{\lceil \mathrm{log}N \rceil }^3)\) based on the cost of \(U_P\) analyzed in Appendix E.2. To summarize, the quantum cost of implementing the operator U is \(\mathcal {O}(N^2{\lceil \mathrm{log}N \rceil }^3)\).

In conclusion, the quantum cost of implementing Algorithm 2 is \(\mathcal {O}(N^2{\lceil \mathrm{log}N \rceil }^3)\), which is mainly concentrated in achieving the operator U if we ignore the insignificant cost, such as the cost of implementing the operator V in (29). Actually, it is reasonable that the cost of Algorithm 2 is N times more expensive than that of the QSVE algorithm proposed in [18]. The cost of the QSVE algorithm on matrix \(A \in \mathbb {C}^{N \times N}\) mainly concentrates on the cost of quantum access to classical data with N rows of A stored in N binary trees. By contrast, our algorithm deals with tensor \(\mathcal {T}\in \mathbb {R}^{N \times N \times N}\) whose \(N^2\) tubes are assumed to be stored in \(N^2\) binary trees, so our algorithm needs N times more gates than the QSVE algorithm to quantum access to the data structure.

Since implementing the physical operations depends on many factors, such as the initial quantum circuit, the quantum computer to be performed on, or the Hamiltonian of the system, the cost presented here is not necessarily the true cost but it provides a reference value. Also, it is not necessarily minimal because it can be decreased by finding more efficient quantum circuits.

The number of \(1\times 1\) and \(2\times 2\) reversible gates required by our quantum algorithm scales polynomially with the dimension of the preference tensor, which presents a big obstacle to practical realization. In fact, not only our algorithm and the QSVE algorithm, but also some benchmark algorithms, such as the HHL algorithm [13], appear too expensive to be executed efficiently by a quantum computer and are not likely to be feasible in the Noisy Intermediate-Scale Quantum (NISQ) era; see [30, Sections 6.6 and 6.7], [44, Section 1.10]. Finally, we point out most quantum machine learning algorithms focus on the time complexity rather than their quantum cost; see, e.g., [6, 13, 17, 36, 37].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, X., Gu, L., Lee, Hw. et al. Quantum context-aware recommendation systems based on tensor singular value decomposition. Quantum Inf Process 20, 190 (2021). https://doi.org/10.1007/s11128-021-03131-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11128-021-03131-y

Keywords

Navigation