Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Quantum tomography of an entangled three-qubit state in silicon

Abstract

Quantum entanglement is a fundamental property of coherent quantum states and an essential resource for quantum computing1. In large-scale quantum systems, the error accumulation requires concepts for quantum error correction. A first step toward error correction is the creation of genuinely multipartite entanglement, which has served as a performance benchmark for quantum computing platforms such as superconducting circuits2,3, trapped ions4 and nitrogen-vacancy centres in diamond5. Among the candidates for large-scale quantum computing devices, silicon-based spin qubits offer an outstanding nanofabrication capability for scaling-up. Recent studies demonstrated improved coherence times6,7,8, high-fidelity all-electrical control9,10,11,12,13, high-temperature operation14,15 and quantum entanglement of two spin qubits9,11,12. Here we generated a three-qubit Greenberger–Horne–Zeilinger state using a low-disorder, fully controllable array of three spin qubits in silicon. We performed quantum state tomography16 and obtained a state fidelity of 88.0%. The measurements witness a genuine Greenberger–Horne–Zeilinger class quantum entanglement that cannot be separated into any biseparable state. Our results showcase the potential of silicon-based spin qubit platforms for multiqubit quantum algorithms.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Device and experimental setup.
Fig. 2: Single-qubit and controlled phase operations.
Fig. 3: Three-qubit entanglement generation and measurement.

Similar content being viewed by others

Data availability

All data in this study are available from the Zenodo repository at https://doi.org/10.5281/zenodo.4722605. Source data are provided with this paper.

References

  1. Nielsen, M. A. & Chuang, I. L. Quantum Computation and Quantum Information (Cambridge Univ. Press, 2000).

  2. Dicarlo, L. et al. Preparation and measurement of three-qubit entanglement in a superconducting circuit. Nature 467, 574–578 (2010).

    Article  CAS  Google Scholar 

  3. Neeley, M. et al. Generation of three-qubit entangled states using superconducting phase qubits. Nature 467, 570–573 (2010).

    Article  CAS  Google Scholar 

  4. Häffner, H. et al. Scalable multiparticle entanglement of trapped ions. Nature 438, 643–646 (2005).

    Article  CAS  Google Scholar 

  5. Neumann, P. et al. Multipartite entanglement among single spins in diamond. Science 323, 1326–1330 (2009).

    Article  Google Scholar 

  6. Pla, J. J. et al. A single-atom electron spin qubit in silicon. Nature 489, 541–544 (2012).

    Article  CAS  Google Scholar 

  7. Pla, J. J. et al. High-fidelity readout and control of a nuclear spin qubit in silicon. Nature 496, 334–338 (2013).

    Article  CAS  Google Scholar 

  8. Veldhorst, M. et al. An addressable quantum dot qubit with fault-tolerant control-fidelity. Nat. Nanotechnol. 9, 981–985 (2014).

    Article  CAS  Google Scholar 

  9. Zajac, D. M. et al. Resonantly driven CNOT gate for electron spins. Science 359, 439–442 (2018).

    Article  CAS  Google Scholar 

  10. Yoneda, J. et al. A quantum-dot spin qubit with coherence limited by charge noise and fidelity higher than 99.9%. Nat. Nanotechnol. 13, 102–106 (2018).

    Article  CAS  Google Scholar 

  11. Watson, T. F. et al. A programmable two-qubit quantum processor in silicon. Nature 555, 633–637 (2018).

    Article  CAS  Google Scholar 

  12. Huang, W. et al. Fidelity benchmarks for two-qubit gates in silicon. Nature 569, 532–536 (2019).

    Article  CAS  Google Scholar 

  13. Yang, C. H. et al. Silicon qubit fidelities approaching incoherent noise limits via pulse engineering. Nat. Electron. 2, 151–158 (2019).

    Article  Google Scholar 

  14. Yang, C. H. et al. Silicon quantum processor unit cell operation above one Kelvin. Nature 580, 350–354 (2020).

    Article  CAS  Google Scholar 

  15. Petit, L. et al. Universal quantum logic in hot silicon qubits. Nature 580, 355–359 (2020).

    Article  CAS  Google Scholar 

  16. James, D. F. V., Kwiat, P. G., Munro, W. J. & White, A. G. Measurement of qubits. Phys. Rev. A 64, 052312 (2001).

    Article  CAS  Google Scholar 

  17. Vandersypen, L. M. K. et al. Interfacing spin qubits in quantum dots and donors—hot, dense, and coherent. npj Quantum Inf. 3, 34 (2017).

    Article  Google Scholar 

  18. Veldhorst, M., Eenink, H. G. J., Yang, C. H. & Dzurak, A. S. Silicon CMOS architecture for a spin-based quantum computer. Nat. Commun. 8, 1766 (2017).

    Article  CAS  Google Scholar 

  19. Yoneda, J. et al. Quantum non-demolition readout of an electron spin in silicon. Nat. Commun. 11, 1144 (2020).

    Article  CAS  Google Scholar 

  20. Xue, X. et al. Repetitive quantum non-demolition measurement and soft decoding of a silicon spin qubit. Phys. Rev. X 10, 021006 (2020).

    CAS  Google Scholar 

  21. Qiao, H. et al. Coherent multi-spin exchange coupling in a quantum-dot spin chain. Phys. Rev. X 10, 31006 (2020).

    CAS  Google Scholar 

  22. Zajac, D. M., Hazard, T. M., Mi, X., Nielsen, E. & Petta, J. R. Scalable gate architecture for densely packed semiconductor spin qubits. Phys. Rev. Appl. 6, 054013 (2016).

    Article  CAS  Google Scholar 

  23. Angus, S. J., Ferguson, A. J., Dzurak, A. S. & Clark, R. G. Gate-defined quantum dots in intrinsic silicon. Nano Lett. 7, 2051–2055 (2007).

    Article  CAS  Google Scholar 

  24. Elzerman, J. M. et al. Single-shot read-out of an individual electron spin in a quantum dot. Nature 430, 431–435 (2004).

    Article  CAS  Google Scholar 

  25. Sigillito, A. J., Gullans, M. J., Edge, L. F., Borselli, M. & Petta, J. R. Coherent transfer of quantum information in silicon using resonant SWAP gates. npj Quantum Inf. 5, 110 (2019).

    Article  Google Scholar 

  26. Takeda, K., Noiri, A., Yoneda, J., Nakajima, T. & Tarucha, S. Resonantly driven singlet-triplet spin qubit in silicon. Phys. Rev. Lett. 124, 117701 (2020).

    Article  CAS  Google Scholar 

  27. Tokura, Y., Van Der Wiel, W. G., Obata, T. & Tarucha, S. Coherent single electron spin control in a slanting Zeeman field. Phys. Rev. Lett. 96, 047202 (2006).

    Article  CAS  Google Scholar 

  28. Takeda, K. et al. A fault-tolerant addressable spin qubit in a natural silicon quantum dot. Sci. Adv. 2, e1600694 (2016).

    Article  CAS  Google Scholar 

  29. Borjans, F., Zajac, D. M., Hazard, T. M. & Petta, J. R. Single-spin relaxation in a synthetic spin–orbit field. Phys. Rev. Appl. 11, 044063 (2018).

    Article  Google Scholar 

  30. Knill, E. et al. Randomized benchmarking of quantum gates. Phys. Rev. A 77, 012307 (2008).

    Article  CAS  Google Scholar 

  31. Meunier, T., Calado, V. E. & Vandersypen, L. M. K. Efficient controlled-phase gate for single-spin qubits in quantum dots. Phys. Rev. B 83, 121403(R) (2011).

    Article  CAS  Google Scholar 

  32. Martins, F. et al. Noise suppression using symmetric exchange gates in spin qubits. Phys. Rev. Lett. 116, 116801 (2016).

    Article  CAS  Google Scholar 

  33. Reed, M. D. et al. Reduced sensitivity to charge noise in semiconductor spin qubits via symmetric operation. Phys. Rev. Lett. 116, 110402 (2016).

    Article  CAS  Google Scholar 

  34. Barends, R. et al. Superconducting quantum circuits at the surface code threshold for fault tolerance. Nature 508, 500–503 (2014).

    Article  CAS  Google Scholar 

  35. Reed, M. D. et al. Realization of three-qubit quantum error correction with superconducting circuits. Nature 482, 382–385 (2012).

    Article  CAS  Google Scholar 

  36. Kelly, J. et al. State preservation by repetitive error detection in a superconducting quantum circuit. Nature 519, 66–69 (2015).

    Article  CAS  Google Scholar 

  37. Mermin, N. D. Extreme quantum entanglement in a superposition of macroscopically distinct states. Phys. Rev. Lett. 65, 1838 (1990).

    Article  CAS  Google Scholar 

  38. Connors, E. J., Nelson, J., Qiao, H., Edge, L. F. & Nichol, J. M. Low-frequency charge noise in Si/SiGe quantum dots. Phys. Rev. B 100, 165305 (2019).

    Article  CAS  Google Scholar 

  39. Hendrickx, N. W. et al. A four-qubit germanium quantum processor. Nature 591, 580–585 (2021).

    Article  CAS  Google Scholar 

  40. Mueller, F. et al. Printed circuit board metal powder filters for low electron temperatures. Rev. Sci. Instrum. 84, 044706 (2013).

    Article  CAS  Google Scholar 

  41. Reilly, D. J., Marcus, C. M., Hanson, M. P. & Gossard, A. C. Fast single-charge sensing with a rf quantum point contact. Appl. Phys. Lett. 91, 162101 (2007).

    Article  CAS  Google Scholar 

  42. Noiri, A. et al. Radio-frequency detected fast charge sensing in undoped silicon quantum dots. Nano Lett. 20, 947–952 (2020).

    Article  CAS  Google Scholar 

  43. Andrews, R. W. et al. Quantifying error and leakage in an encoded Si/SiGe triple-dot qubit. Nat. Nanotechnol. 14, 747–750 (2019).

    Article  CAS  Google Scholar 

  44. Hensgens, T. et al. Quantum simulation of a Fermi–Hubbard model using a semiconductor quantum dot array. Nature 548, 70–73 (2017).

    Article  CAS  Google Scholar 

  45. Jones, A. M. et al. Spin-blockade spectroscopy of Si/Si–Ge quantum dots. Phys. Rev. Appl. 12, 014026 (2019).

    Article  CAS  Google Scholar 

  46. Newville, M., Stensitzki, T., Allen, D. & Ingargiola, A. LMFIT: non-linear least-square minimization and curve-fitting for Python. Zenodo https://zenodo.org/record/11813#.YH6fbej7SUl (2014).

  47. Jones, E., Oliphant, T. & Peterson, P. SciPy: open source scientific tools for Python. Science Open https://www.scienceopen.com/document?vid=ab12905a-8a5b-43d8-a2bb-defc771410b9 (2001).

Download references

Acknowledgements

We thank the Microwave Research Group at Caltech for technical support. This work was supported financially by Core Research for Evolutional Science and Technology (CREST), Japan Science and Technology Agency (JST) (JPMJCR15N2 and JPMJCR1675), MEXT Quantum Leap Flagship Program (MEXT Q-LEAP) grant no. JPMXS0118069228, JST Moonshot R&D grant no. JPMJMS2065 and JSPS KAKENHI grant nos. 16H02204, 17K14078, 18H01819, 19K14640 and 20H00237. T.N. acknowledges support from the Murata Science Foundation Research Grant and JST, PRESTO grant no. JPMJPR2017.

Author information

Authors and Affiliations

Authors

Contributions

K.T. and A.N. fabricated the device and performed the measurements. T.N., J.Y. and T.K. contributed the data acquisition and discussed the results. K.T. wrote the paper with inputs from all the co-authors. S.T. supervised the project.

Corresponding authors

Correspondence to Kenta Takeda or Seigo Tarucha.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Nanotechnology thanks Joseph Kerckhoff and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Initialization and measurement protocol.

Initialization and readout procedure. The spin readout and initialization for Q1 is performed near the (111)–(011) boundary, whereas for Q2 and Q3 it is performed near the (111)–(110) boundary. Note that Q2 cannot be directly read out through the reservoir as the co-tunnelling rate between the (111) and (101) states is too small compared to the spin relaxation rate \(T_1^{ - 1}\). The dwell times are 450 μs for Q3 initialization, 750 μs for Q1 initialization, 300 μs for Q3 readout and 750 μs for Q1 readout. The resonant SWAP pulses are 0.25 μs long and it corresponds to an exchange Rabi frequency of 2 MHz. The resonance frequency is typically around 410 MHz. The readout dwell time of Q3 is compromised for the finite T1 relaxation times of Q1 and Q2 and it results in imperfect spin-down initialization after the readout stage. Therefore, in order to increase the initialization fidelities, we use an explicit initialization stage which is redundant in the ideal case where all the three spins are initialized to spin-down after the readout stage.

Extended Data Fig. 2 Single-qubit characterization.

All qubits are initialized to spin-down before the manipulation stage and only one of the qubits is read out right after the manipulation stage unless noted. The exchange interactions are turned off. All errors represent the estimated standard errors for the best-fit values. ad, T1 measurements. First, a spin-up state is prepared using an X pulse. Then we vary the waiting time of tw at the single-qubit manipulation point before performing single-shot measurement. In this T1 measurement, all three spins are sequentially read out and therefore the visibilities of Q1 and Q2 are decreased by T1 relaxation during the readout stage. Note that the visibility of Q3 is unaffected by the sequential readout. eh, Ramsey interferometry measurements. First, a π/2 pulse (+2 MHz detuned from the resonance frequency) is applied to rotate the spin state to the xy-plane in the Bloch sphere. After an evolution time of tevol, another π/2 pulse is applied to project the spin state in the z-axis for measurement. The black solid curves are the fit with Gaussian decay. The integration time is 75.8 sec for all qubits. il, Hahn echo measurements. Each fitting curve is given by P(tevol) = Aexp(−(tevol/T2echo)γ) + B, where A and B are the constants to account for the readout fidelities and γ is an exponent. The exponents are found to be γ = 1.79 ± 0.12 (Q1), 2.75 ± 0.10 (Q2) and 2.61 ± 0.09 (Q3).

Source data

Extended Data Fig. 3 Additional randomized benchmarking measurements.

a, Measurement result with an asymmetric readout condition. F1 (F0) is the spin-up probability when the recovery Clifford gate is chosen so that the ideal final spin state is spin-up (-down). The data points at m = 0 are measured without any random Clifford gates applied. Only an X pulse is applied in the case of F1 and no pulse is applied in the case of F0. The dashed line shows a constant 0.462, the expected saturation value derived from the readout asymmetry. b, Measurement result with a more symmetric readout condition.

Source data

Extended Data Fig. 4 Randomized benchmarking with detuned microwave frequency.

a, Randomized benchmarking sequence for Q2 fidelity measurement. b, Randomized benchmarking measurement result of Q2 with a frequency detuning of 0.6 MHz. For each of the control bit (Q3) states, the measurement is performed for 16 sets of random Clifford gate sequences. The sequence fidelity shows an average of the results for the two control bit configurations. The errors represent the estimated standard errors for the best-fit values. c, d, Similar randomized benchmarking measurement performed for Q3. The errors represent the estimated standard errors for the best-fit values.

Source data

Extended Data Fig. 5 Measurements of exchange interactions.

All errors represent the estimated standard errors for the best-fit values. a, b, Controlled-rotation for Q1 and Q2. The measurement is performed to probe \(J_{12}^\prime\). First, Q1 and Q2 are initialized to spin-down. To prepare a spin-up control qubit (Q2) state, an X pulse is applied. After tuning on \(J_{12}^\prime\) by a gate voltage pulse, a low-power Gaussian microwave pulse (truncated at ±2σ) is applied to induce a controlled-rotation. The filled (open) circles show the measured spin-up probabilities with the control qubit spin-down (up). The solid lines are Gaussian fitting curves. From the separation of the two peaks, we obtain \(J_{12}^\prime\) = 2.75 ± 0.02 MHz. c, d, Similar controlled rotation measurement for Q2 and Q3. We obtain \(J_{23}^\prime\) = 12.50 ± 0.02 MHz from this measurement. e, Ramsey experiment to extract \(J_{12}^{{\mathrm{off}}}\). We perform two Ramsey measurements of Q1 with different control qubit (Q2) states. The difference of qubit frequency detuning is equivalent to \(J_{12}^{{\mathrm{off}}}\). f, Ramsey measurement result when Q2 is spin-down. The red circles are the measured Q1 spin-up probabilities and the black solid curve shows a fit with Gaussian decay. From the oscillation frequency of the decay curve, we extract δf = 2.28 ± 0.01MHz. g, Measurement similar to the one in f when Q2 is spin-up. We extract δf = 2.21 ± 0.01MHz. Since the difference between δf and δf is below the stochastic fluctuation of the frequency detuning, we conclude that \(J_{12}^{{\mathrm{off}}}\) is below our detection limit. Note that each frequency error shows one standard deviation of the fitting parameter. hj, Ramsey experiments to extract \(J_{23}^{{\mathrm{off}}}\). We obtain \(J_{23}^{{\mathrm{off}}}\) = δf−δf = 1.17 ± 0.01 MHz from these measurements.

Source data

Extended Data Fig. 6 Bell state tomography using Q2 and Q3.

As a benchmark of our two-qubit CZ gate, we perform Bell state tomography on Q2 and Q3. The experiment is a reduced version of the three-qubit GHZ state tomography. The readout errors are removed using the measured readout fidelities and maximum likelihood estimation is used to reconstruct the density matrices. a, Quantum gate sequence for Bell state creation and state tomography. By modifying the phase gates after the second CZ/2 pulse, we can create all four Bell states. be, Real parts of the measured density matrices for four Bell states, \(\Phi ^ + = (|{\uparrow \uparrow}\rangle + |{\downarrow \downarrow}\rangle )/\sqrt 2\) (b), \(\Phi ^ - = (|{\uparrow \uparrow}\rangle - |{\downarrow \downarrow} \rangle )/\sqrt 2\) (c), \(\Psi ^ + = (|{\downarrow \uparrow}\rangle + |{\downarrow \uparrow}\rangle)/\sqrt 2\) (d), and \(\Psi ^ - = (|{\downarrow \uparrow}\rangle - |{\uparrow \downarrow}\rangle)/\sqrt 2\) (e). We obtain the state fidelities relative to the target states of 0.942 (Φ+), 0.933 (Φ), 0.950 (Ψ+) and 0.940 (Ψ), and the concurrences of 0.950 (Φ+), 0.906 (Φ), 0.923 (Ψ+) and 0.935 (Ψ).

Source data

Extended Data Fig. 7 Imaginary part of the experimental GHZ state.

The imaginary part is all zero for an ideal GHZ state. Here, the maximum absolute value of the matrix elements is 0.09.

Source data

Source data

Source Data Fig. 1

Numerical data used to generate Fig. 1.

Source Data Fig. 2

Numerical data used to generate Fig. 2.

Source Data Fig. 3

Numerical data used to generate Fig. 3.

Source Data Extended Data Fig. 2

Numerical data used to generate Extended Data Fig. 2.

Source Data Extended Data Fig. 3

Numerical data used to generate Extended Data Fig. 3.

Source Data Extended Data Fig. 4

Numerical data used to generate Extended Data Fig. 4.

Source Data Extended Data Fig. 5

Numerical data used to generate Extended Data Fig. 5.

Source Data Extended Data Fig. 6

Numerical data used to generate Extended Data Fig. 6.

Source Data Extended Data Fig. 7

Numerical data used to generate Extended Data Fig. 7.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Takeda, K., Noiri, A., Nakajima, T. et al. Quantum tomography of an entangled three-qubit state in silicon. Nat. Nanotechnol. 16, 965–969 (2021). https://doi.org/10.1038/s41565-021-00925-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41565-021-00925-0

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing