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.

  • Article
  • Published:

Precision tomography of a three-qubit donor quantum processor in silicon

Abstract

Nuclear spins were among the first physical platforms to be considered for quantum information processing1,2, because of their exceptional quantum coherence3 and atomic-scale footprint. However, their full potential for quantum computing has not yet been realized, owing to the lack of methods with which to link nuclear qubits within a scalable device combined with multi-qubit operations with sufficient fidelity to sustain fault-tolerant quantum computation. Here we demonstrate universal quantum logic operations using a pair of ion-implanted 31P donor nuclei in a silicon nanoelectronic device. A nuclear two-qubit controlled-Z gate is obtained by imparting a geometric phase to a shared electron spin4, and used to prepare entangled Bell states with fidelities up to 94.2(2.7)%. The quantum operations are precisely characterized using gate set tomography (GST)5, yielding one-qubit average gate fidelities up to 99.95(2)%, two-qubit average gate fidelity of 99.37(11)% and two-qubit preparation/measurement fidelities of 98.95(4)%. These three metrics indicate that nuclear spins in silicon are approaching the performance demanded in fault-tolerant quantum processors6. We then demonstrate entanglement between the two nuclei and the shared electron by producing a Greenberger–Horne–Zeilinger three-qubit state with 92.5(1.0)% fidelity. Because electron spin qubits in semiconductors can be further coupled to other electrons7,8,9 or physically shuttled across different locations10,11, these results establish a viable route for scalable quantum information processing using donor nuclear and electron spins.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: Operation of a one-electron–two-nuclei quantum processor.
Fig. 2: Tomography of nuclear Bell states.
Fig. 3: Precise tomographic characterization of one- and two-qubit gate quality.
Fig. 4: Creation and tomography of an electron–nuclear three-qubit GHZ state.

Similar content being viewed by others

Data availability

The experimental data that support the findings of this study are available in Figshare, https://doi.org/10.6084/m9.figshare.c.5471706Source data are provided with this paper.

Code availability

The GST analysis was performed using a developmental version of pyGSTi that requires expert-level knowledge of the software to install and run. A future official release of pyGSTi will support the type of analysis performed here using a simple and well documented Python script. Until this code is available, interested readers can contact the corresponding author to get help with accessing and running the existing code. Multivalley effective mass theory calculations, some of the results of which are illustrated in Fig. 1b, were performed using a fork of the code first developed in the production of ref. 60 that was extended to include multielectron interactions as reported in ref. 59. Requests for a license for and copy of this code will be directed to points of contact at Sandia National Laboratories and the University of New South Wales, through the corresponding author. The analysis code for Bell state tomography is in Figshare, https://doi.org/10.6084/m9.figshare.c.5471706.

References

  1. Kane, B. E. A silicon-based nuclear spin quantum computer. Nature 393, 133–137 (1998).

    Article  Google Scholar 

  2. Vandersypen, L. M. K. & Chuang, I. L. NMR techniques for quantum control and computation. Rev. Mod. Phys. 76, 1037–1069 (2005).

    Article  Google Scholar 

  3. Saeedi, K. et al. Room-temperature quantum bit storage exceeding 39 minutes using ionized donors in silicon-28. Science 342, 830–833 (2013).

    Article  ADS  Google Scholar 

  4. Filidou, V. et al. Ultrafast entangling gates between nuclear spins using photoexcited triplet states. Nat. Phys. 8, 596–600 (2012).

    Article  CAS  Google Scholar 

  5. Nielsen, E. et al. Gate set tomography. Quantum 5, 557 (2021).

    Article  Google Scholar 

  6. Fowler, A. G., Mariantoni, M., Martinis, J. M. & Cleland, A. N. Surface codes: Towards practical large-scale quantum computation. Phys. Rev. A 86, 032324 (2012).

    Article  ADS  Google Scholar 

  7. Harvey-Collard, P. et al. Coherent coupling between a quantum dot and a donor in silicon. Nat. Commun. 8, 1029 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  8. He, Y. et al. A two-qubit gate between phosphorus donor electrons in silicon. Nature 571, 371–375 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  9. Madzik, M. T. et al. Conditional quantum operation of two exchange-coupled single-donor spin qubits in a MOS-compatible silicon device. Nat. Commun. 12, 181 (2021).

    Article  CAS  PubMed Central  Google Scholar 

  10. Hensen, B. et al. A silicon quantum-dot-coupled nuclear spin qubit. Nat. Nanotechnol. 15, 13–17 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  11. Yoneda, J. et al. Coherent spin qubit transport in silicon. Nat. Commun. 12, 4114 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. Zhong, M. et al. Optically addressable nuclear spins in a solid with a six-hour coherence time. Nature 517, 177–180 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  13. Muhonen, J. T. et al. Quantifying the quantum gate fidelity of single-atom spin qubits in silicon by randomized benchmarking. J. Phys. Condens. Matter 27, 154205 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  14. Bradley, C. et al. A ten-qubit solid-state spin register with quantum memory up to one minute. Phys. Rev. 9, 031045 (2019).

    Article  CAS  Google Scholar 

  15. Bourassa, A. et al. Entanglement and control of single nuclear spins in isotopically engineered silicon carbide. Nat. Mater. 19, 1319–1325 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  16. Waldherr, G. et al. Quantum error correction in a solid-state hybrid spin register. Nature 506, 204–207 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  17. Bhaskar, M. K. et al. Experimental demonstration of memory-enhanced quantum communication. Nature 580, 60–64 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  18. Pompili, M. et al. Realization of a multinode quantum network of remote solid-state qubits. Science 372, 259–264 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  19. 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  ADS  Google Scholar 

  20. Morello, A. et al. Single-shot readout of an electron spin in silicon. Nature 467, 687–691 (2010).

    Article  ADS  CAS  PubMed  Google Scholar 

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

    Article  ADS  CAS  PubMed  Google Scholar 

  22. Pla, J. J. et al. A single-atom electron spin qubit in silicon. Nature 489, 541–545 (2012).

    Article  ADS  CAS  PubMed  Google Scholar 

  23. Ivie, J. A. et al. Impact of incorporation kinetics on device fabrication with atomic precision. Phys. Rev. Appl. 16, 054037 (2021).

    Article  ADS  CAS  Google Scholar 

  24. Hile, S. J. et al. Addressable electron spin resonance using donors and donor molecules in silicon. Sci. Adv. 4, eaaq1459 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  25. Anandan, J. The geometric phase. Nature 360, 307–313 (1992).

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

  27. Dehollain, J. P. et al. Optimization of a solid-state electron spin qubit using gate set tomography. New J. Phys. 18, 103018 (2016).

    Article  ADS  Google Scholar 

  28. Blume-Kohout, R. et al. Demonstration of qubit operations below a rigorous fault tolerance threshold with gate set tomography. Nat. Commun. 8, 14485 (2017).

    Article  CAS  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  Google Scholar 

  30. Xue, X. et al. Benchmarking gate fidelities in a Si/SiGe two-qubit device. Phys. Rev. 9, 021011 (2019).

    Article  CAS  Google Scholar 

  31. Kimmel, S., da Silva, M. P., Ryan, C. A., Johnson, B. R. & Ohki, T. Robust extraction of tomographic information via randomized benchmarking. Phys. Rev. 4, 011050 (2014).

    Article  Google Scholar 

  32. Carignan-Dugas, A., Wallman, J. J. & Emerson, J. Bounding the average gate fidelity of composite channels using the unitarity. New J. Phys. 21, 053016 (2019).

    Article  ADS  MathSciNet  Google Scholar 

  33. Blume-Kohout, R. et al. A taxonomy of small Markovian errors. Preprint at https://arxiv.org/abs/2103.01928 (2021).

  34. Proctor, T., Rudinger, K., Young, K., Sarovar, M. & Blume-Kohout, R. What randomized benchmarking actually measures. Phys. Rev. Lett. 119, 130502 (2017).

    Article  ADS  MathSciNet  PubMed  Google Scholar 

  35. Novais, E. & Mucciolo, E. R. Surface code threshold in the presence of correlated errors. Phys. Rev. Lett. 110, 010502 (2013).

    Article  ADS  CAS  PubMed  Google Scholar 

  36. Neumann, P. et al. Multipartite entanglement among single spins in diamond. Science 320, 1326–1329 (2008).

    Article  ADS  CAS  PubMed  Google Scholar 

  37. Takeda, K. et al. Quantum tomography of an entangled three-qubit state in silicon. Nat. Nanotechnol. 16, 965–969 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  38. Gullans, M. J. & Petta, J. R. Protocol for a resonantly driven three-qubit Toffoli gate with silicon spin qubits. Phys. Rev. B 100, 085419 (2019).

    Article  ADS  CAS  Google Scholar 

  39. Mehring, M., Mende, J. & Scherer, W. Entanglement between an electron and a nuclear spin 1/2. Phys. Rev. Lett. 90, 153001 (2003).

    Article  ADS  CAS  PubMed  Google Scholar 

  40. Sackett, C. A. et al. Experimental entanglement of four particles. Nature 404, 256–259 (2000).

    Article  ADS  CAS  PubMed  Google Scholar 

  41. Wei, K. X. et al. Verifying multipartite entangled Greenberger–Horne–Zeilinger states via multiple quantum coherences. Phys. Rev. A 101, 032343 (2020).

    Article  ADS  CAS  Google Scholar 

  42. Gross, J. A., Godfrin, C., Blais, A. & Dupont-Ferrier, E. Hardware-efficient error-correcting codes for large nuclear spins. Preprint at https://arxiv.org/abs/2103.08548 (2021).

  43. Asaad, S. et al. Coherent electrical control of a single high-spin nucleus in silicon. Nature 579, 205–209 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Tosi, G. et al. Silicon quantum processor with robust long-distance qubit couplings. Nat. Commun. 8, 450 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  45. Pica, G., Lovett, B. W., Bhatt, R. N., Schenkel, T. & Lyon, S. A. Surface code architecture for donors and dots in silicon with imprecise and nonuniform qubit couplings. Phys. Rev. B 93, 035306 (2016).

    Article  ADS  Google Scholar 

  46. Buonacorsi, B. et al. Network architecture for a topological quantum computer in silicon. Quantum Sci. Technol. 4, 025003 (2019).

    Article  ADS  Google Scholar 

  47. Tosi, G., Mohiyaddin, F. A., Tenberg, S., Laucht, A. & Morello, A. Robust electric dipole transition at microwave frequencies for nuclear spin qubits in silicon. Phys. Rev. B 98, 075313 (2018).

    Article  ADS  Google Scholar 

  48. Mielke, J., Petta, J. R. & Burkard, G. Nuclear spin readout in a cavity-coupled hybrid quantum dot-donor system. PRX Quantum 2, 020347 (2021).

    Article  ADS  Google Scholar 

  49. Xue, X. et al. Quantum logic with spin qubits crossing the surface code threshold. Nature 601, 343–347 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Noiri, A. et al. Fast universal quantum gate above the fault-tolerance threshold in silicon. Nature 601, 338–342 (2022).

    Article  PubMed  Google Scholar 

  51. Adambukulam, C. et al. An ultra-stable 1.5 T permanent magnet assembly for qubit experiments at cryogenic temperatures. Rev. Sci. Instrum. 92, 085106 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  52. Kalra, R. et al. Vibration-induced electrical noise in a cryogen-free dilution refrigerator: characterization, mitigation, and impact on qubit coherence. Rev. Sci. Instrum. 87, 073905 (2016).

    Article  ADS  PubMed  Google Scholar 

  53. Dehollain, J. et al. Nanoscale broadband transmission lines for spin qubit control. Nanotechnology 24, 015202 (2012).

    Article  ADS  PubMed  Google Scholar 

  54. Feher, G. Electron spin resonance experiments on donors in silicon. I. Electronic structure of donors by the electron nuclear double resonance technique. Phys. Rev. 114, 1219–1244 (1959).

    Article  ADS  CAS  Google Scholar 

  55. Steger, M. et al. Optically-detected NMR of optically-hyperpolarized 31P neutral donors in 28Si. J. Appl. Phys. 109, 102411 (2011).

    Article  ADS  Google Scholar 

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

    Article  ADS  CAS  PubMed  Google Scholar 

  57. Morello, A. et al. Architecture for high-sensitivity single-shot readout and control of the electron spin of individual donors in silicon. Phys. Rev. B 80, 081307 (2009).

    Article  ADS  Google Scholar 

  58. Braginsky, V. B. & Khalili, F. Ya. Quantum nondemolition measurements: the route from toys to tools. Rev. Mod. Phys. 68, 1–11 (1996).

    Article  ADS  MathSciNet  Google Scholar 

  59. Joecker, B. et al. Full configuration interaction simulations of exchange-coupled donors in silicon using multi-valley effective mass theory. New J. Phys. 23, 073007 (2021).

    Article  ADS  CAS  Google Scholar 

  60. Gamble, J. K. et al. Multivalley effective mass theory simulation of donors in silicon. Phys. Rev. B 91, 235318 (2015).

    Article  ADS  Google Scholar 

  61. Nielsen, E. et al. Python GST Implementation (PyGSTi) v. 0.9. Technical Report (Sandia National Lab, 2019).

  62. Nielsen, E. et al. Probing quantum processor performance with pyGSTi. Quantum Sci. Technol. 5, 044002 (2020).

    Article  ADS  Google Scholar 

  63. Wilks, S. S. The large-sample distribution of the likelihood ratio for testing composite hypotheses. Ann. Math. Stat. 9, 60–62 (1938).

    Article  MATH  Google Scholar 

  64. Nielsen, E., Rudinger, K., Proctor, T., Young, K. & Blume-Kohout, R. Efficient flexible characterization of quantum processors with nested error models. New J. Phys. 23, 093020 (2021).

    Article  ADS  Google Scholar 

  65. Akaike, H. Information theory and an extension of the maximum likelihood principle. In Selected Papers of Hirotugu Akaike (eds Parzen, E. et al.) 199–213 (Springer, 1998).

  66. Tenberg, S. B. et al. Electron spin relaxation of single phosphorus donors in metal-oxide-semiconductor nanoscale devices. Phys. Rev. B 99, 205306 (2019).

    Article  ADS  CAS  Google Scholar 

  67. Hsueh, Y.-L. et al. Spin-lattice relaxation times of single donors and donor clusters in silicon. Phys. Rev. Lett. 113, 246406 (2014).

    Article  ADS  PubMed  Google Scholar 

Download references

Acknowledgements

We acknowledge conversations with W. Huang, R. Rahman, S. Seritan and C. H. Yang and technical support from T. Botzem. The research was supported by the Australian Research Council (grant no. CE170100012), the US Army Research Office (contract no. W911NF-17-1-0200), and the Australian Department of Industry, Innovation and Science (grant no. AUSMURI000002). We acknowledge support from the Australian National Fabrication Facility (ANFF). This material is based upon work supported in part by the iHPC facility at the University of Technology Sydney (UTS), by the US Department of Energy, Office of Science, Office of Advanced Scientific Computing Research’s Quantum Testbed Pathfinder and Early Career Research Programs, and by the US Department of Energy, Office of Science, National Quantum Information Science Research Centers (Quantum Systems Accelerator). Sandia National Laboratories is a multimission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC, a wholly owned subsidiary of Honeywell International, Inc., for the US Department of Energy’s National Nuclear Security Administration under contract DE-NA0003525. All statements of fact, opinion or conclusions contained herein are those of the authors and should not be construed as representing the official views or policies of the US Department of Energy, or the US Government.

Author information

Authors and Affiliations

Authors

Contributions

M.T.M., V.S. and F.E.H. fabricated the device, with the supervision of A.M. and A.S.D., on an isotopically enriched 28Si wafer supplied by K.M.I. A.M.J., B.C.J. and D.N.J. designed and performed the ion implantation. M.T.M. and S.A. performed the experiments and analysed the data, with A.L. and A.M.’s supervision. B.J. and A.D.B. developed and applied computational tools to calculate the electron wavefunction and the Hamiltonian evolution. A.Y. designed the initial GST sequences, with C.F.’s supervision. K.M.R., E.N., K.C.Y., T.J.P. and R.B.-K. developed and applied the GST method. A.M., R.B.-K., M.T.M. and S.A. wrote the manuscript, with input from all coauthors.

Corresponding author

Correspondence to Andrea Morello.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review information

Nature thanks Christopher Wood and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Device layout.

Scanning electron micrograph of a device identical to the one used in this experiment. 31P donor atoms are implanted in the region marked by the orange rectangle, using a fluence of 1.4 × 1012 cm−2 which results in a most probable inter-donor spacing of approximately 8 nm. Four metallic gates are fabricated around the implantation region, and used to modify the electrochemical potential of the donors. A nearby SET, formed using the SET top gate and barrier gates, enables charge sensing of a single donor atom, as well as its electron spin through spin-to-charge conversion (Methods). The tunnel coupling between the donors and SET is tuned by the rate gate situated between the SET and donor implant region. A nearby microwave (MW) antenna is used for ESR and NMR of the donor electron and nuclear spins, respectively.

Extended Data Fig. 2 Electrical tunability of the hyperfine interaction and the electron gyromagnetic ratio.

a, Map of the SET current as a function of SET gate and fast donor gates (pulsed jointly). The white dashed line indicates the location in gate space where the 2P donor cluster changes its charge state. The third, hyperfine-coupled electron is present on the cluster in the region to the right of the line. Electron spin readout is performed at the location indicated by the pink star. b, ESR spectrum of the electron bound to the 2P cluster, acquired while the system was tuned within the blue dashed rectangle in a. The hyperfine couplings A1, A2 are extracted from ESR frequencies as shown, namely \({A}_{1}=({\nu }_{{\rm{e}}|\Uparrow \Downarrow }+{\nu }_{{\rm{e}}|\Uparrow \Uparrow })/2-({\nu }_{{\rm{e}}|\Downarrow \Downarrow }+{\nu }_{{\rm{e}}|\Downarrow \Uparrow })/2\); \({A}_{2}={\nu }_{{\rm{e}}|\Uparrow \Uparrow }-{\nu }_{{\rm{e}}|\Uparrow \Downarrow }\). c, d, Extracted hyperfine couplings within the marked area. The data show that A1 decreases and A2 increases upon moving the operation point towards higher gate voltages and away from the donor readout position. e, A small change is also observed in the sum of the two hyperfine interactions At = A1 + A2. f, Electrical modulation (Stark shift) of the electron gyromagnetic ratio γe, extracted from the shift of the average of the hyperfine-split electron resonances. The ESR frequencies can be tuned with fast donor gates at the rate of \(\Delta {\nu }_{{\rm{e}}|\Uparrow \Uparrow }=0.3{{\rm{MHzV}}}^{-1}\); \(\Delta {\nu }_{{\rm{e}}|\Uparrow \Downarrow }=5.2{{\rm{MHzV}}}^{-1}\); \(\Delta {\nu }_{{\rm{e}}|\Downarrow \Uparrow }=7.6{{\rm{MHzV}}}^{-1}\); \(\Delta {\nu }_{{\rm{e}}|\Downarrow \Downarrow }=2.4{{\rm{MHzV}}}^{-1}\).

Extended Data Fig. 3 Coherence metrics of the electron spin qubit.

The columns correspond to the nuclear configurations \(|\Downarrow \Downarrow \rangle \), \(|\Downarrow \Uparrow \rangle \), \(|\Uparrow \Downarrow \rangle \), \(|\Uparrow \Uparrow \rangle \), respectively. All measurements start with the electron spin initialized in the \(|\downarrow \rangle \) state. Error bars are 1σ confidence intervals. a, Electron Rabi oscillations. The measurements were performed by applying a resonant ESR pulse of increasing duration. The different Rabi frequencies fRabi on each resonance are probably due to a frequency-dependent response of the on-chip antenna and the cable connected to it. b, Electron spin-lattice relaxation times T1e. Measurements were obtained by first adiabatically inverting the electron spin to \(|\uparrow \rangle \), followed by a varying wait time τ before electron readout. The observed relaxation times are nearly three orders of magnitude shorter than typically observed in single-electron, single-donor devices66, and even shorter compared to 1e–2P clusters. This strongly suggests that the measured electron is the third one, on top of two more tightly-bound electrons which form a singlet spin state67. We also observe a strong dependence of T1e on nuclear spin configuration. c, Electron dephasing times \({T}_{2{\rm{e}}}^{\ast }\). The measurements were conducted by performing a Ramsey experiment—that is, by applying two π/2 pulses separated by a varying wait time τ, followed by electron readout. The Ramsey fringes are fitted to a function of the form \({P}_{\uparrow }(\tau )={C}_{0}+{C}_{1}\,\cos (\Delta \omega \tau +\Delta \phi )\exp [-{(\tau /{T}_{2e}^{\ast })}^{2}]\), where Δω is the frequency detuning and Δϕ is a phase offset. The observed \({T}_{2{\rm{e}}}^{\ast }\) times are comparable to previous values for electrons coupled to a single 31P nucleus. d, Electron Hahn-echo coherence times \({T}_{2{\rm{e}}}^{{\rm{H}}}\), obtained by adding a π refocusing pulse to the Ramsey sequence. We also varied the phase of the final π/2 pulse at a rate of one period per τ = (5 kHz)−1, to introduce oscillations in the spin-up fraction which help improve the fitting. The curves are fitted to the same function used to fit the Ramsey fringes, with fixed Δω = 5 kHz. The measured \({T}_{2{\rm{e}}}^{{\rm{H}}}\) times are similar to previous observations for electrons coupled to a single 31P nucleus.

Extended Data Fig. 4 Nuclear spin coherence times.

Panels in column 1 (2) correspond to nucleus Q1 (Q2). Error bars are 1σ confidence intervals. a, Nuclear dephasing times \({T}_{2{\rm{n}}}^{\ast }\), obtained from a Ramsey experiment. Results are fitted with a decaying sinusoid with fixed exponent factor 2 (see Extended Data Fig. 3). b, Nuclear Hahn-echo coherence times \({T}_{2{\rm{n}}}^{{\rm{H}}}\). To improve fitting, oscillations are induced by incrementing the phase of the final π/2 pulse with τ at a rate of one period per (3.5 kHz)−1. Results are fitted with a decaying sinusoid with fixed exponent factor 2 (see Extended Data Fig. 3). c, Dependence of \({T}_{2{\rm{n}}}^{{\rm{H}}}\) on the amplitude of an off-resonance pulse. We perform this experiment to study whether a qubit, nominally left idle (or, in quantum information terms, subjected to an identity gate) is affected by the application of an RF pulse to the other qubit, at a vastly different frequency. Here, during the idle times between NMR pulses, an RF pulse is applied at a fixed frequency 20 MHz—far off resonance from both qubits’ transitions—with varying amplitude VRF. The red dashed line indicates the applied RF amplitude for NMR pulses throughout the experiment. We observe a slow decrease of \({T}_{2{\rm{n}}}^{{\rm{H}}}\) with increasing VRF. This is qualitatively consistent with the observation of large stochastic errors on the idle qubit, as extracted by the GST analysis in Fig. 3.

Extended Data Fig. 5 Nuclear spin quantum jumps caused by ionization shock.

The electron and nuclear spin readout relies upon spin-dependent charge tunnelling between the donors and the SET island. If the electron tunnels out of the two-donor system, the hyperfine interactions A1, A2 suddenly drop to zero. If A1 and A2 include an anisotropic component (for example, due to the non-spherical shape of the electron wavefunction which results in nonzero dipolar fields at the nuclei), the ionization is accompanied by a sudden change in the nuclear spin quantization axes (‘ionization shock’), and can result in a flip of the nuclear spin state. We measure the nuclear spin flips caused by ionization shock by forcibly loading and unloading an electron from the 2P cluster every 0.8 ms. a, For qubit 1 with A1 = 95 MHz, the flip rate is \({{\Gamma }}_{1}=2.8\times {10}^{-6}\frac{{N}_{{\rm{flip}}}}{{N}_{{\rm{ion}}}}\). b, For qubit 2 with A2 = 9 MHz, the flip rate is \({{\Gamma }}_{2}=4.0\times {10}^{-7}\frac{{N}_{{\rm{flip}}}}{{N}_{{\rm{ion}}}}\). This means that the nuclear spin readout via the electron ancilla is almost exactly quantum nondemolition. From this data, we also extract an average time between random nuclear spin flips of 283 seconds for qubit 1, and 2,000 seconds for qubit 2. The extremely low values of Γ—comparable to those observed in single-donor systems—are the reason why we can reliably operate the two 31P nuclei as high-fidelity qubits.

Extended Data Fig. 6 CNOT and zero-CNOT nuclear two-qubit gates.

We perform Rabi oscillation on the control qubit followed by the application of a, zCNOT or b, CNOT gates. The two qubits are initialized in the \(|\Downarrow \Downarrow \rangle \equiv |11\rangle \) state. We observe the Rabi oscillations of both qubits in phase for zCNOT and out of phase for CNOT. At every odd multiple of π/2 rotation of the control qubit the Bell states are created.

Extended Data Fig. 7 Two-qubit GST.

a, Measurement circuit for the two-qubit GST. A modified version of this circuit has been used for Bell state tomography. The green box prepares the qubit 2 in the \(|\Uparrow \rangle \) state, then the orange box prepares the qubit 1 in the \(|\Uparrow \rangle \) state. The readout step in the blue box (see Methods) determines whether the \(|\Downarrow \Downarrow \rangle \) state initialization was successful. Only then the record will be saved. The electron spin is prepared in \(|\downarrow \rangle \) during the nuclear spin readout process. Subsequently, the GST sequence is executed. The red box indicates the Q1, Q2 readout step. The total duration of the pulse sequence is 120 ms, of which nuclear spin initialization is 8.6 ms (green and yellow), initial nuclear spin readout is 26.5 ms (blue), 3 ms delay is added for electron initialization (between blue and purple), GST circuit is 10 μs–300 μs (purple), and nuclear readout is 80 ms (orange). b, Measurement results for individual two-qubit GST circuit. The first 145 circuits estimate the preparation and measurement fiducials, and the subsequent circuits are ordered by increasing circuit depth. At the end of a circuit, there are three situations for the target state populations: 1) the population is entirely in one state, while all others are zero; 2) the population is equally spread over two states, while the other two are zero; 3) the population is equally spread over all four states. The measured state populations for the different circuits therefore congregate around the four bands 0, 0.25, 0.5, and 1, as indicated by black dashed lines.

Extended Data Fig. 8 Estimated gate set, from process matrices to error rates.

Experimental GST data were analysed using pyGSTi to obtain self-consistent maximum likelihood estimates of two-qubit process matrices for all six elementary gates. These are represented (‘Process Matrix’ column) in a gauge that minimizes their average total error, as superoperators in the two-qubit Pauli basis. Green columns indicate positive matrix elements, orange ones are negative. Wireframe sections indicate differences between estimated and ideal (target) process matrices. Those process matrices can be transformed to error generators (‘Error Generator’ column) that isolate those differences, and are zero if the estimated gate equals its target. Each gate’s error generator was decomposed into a sparse sum of Hamiltonian and stochastic elementary error generators33. Those rates are depicted (‘All Error Rates’ column) as contributions to the gate’s total error, with 1σ uncertainties indicated in parentheses. Each non-vanishing elementary error rate (error generators are denoted ‘H’ or ‘S’ followed by a Pauli operator) is listed, and identified with its role in the total error budget (reproduced from Fig. 3). Orange bars indicate stochastic errors, dark blue indicate coherent errors that are intrinsic to the gate, and light blue indicate relational coherent errors that were assigned to this gate. Total height of the blue region indicates the total coherent error, but because coherent error amplitudes add in quadrature, individual components’ heights are proportional to their quadrature.

Extended Data Fig. 9 Simulation of standard and interleaved randomized benchmarking.

All simulated randomized benchmarking experiments used two-qubit Clifford subroutines compiled from the six native gates, requiring (on average) 14.58 individual gate operations per two-qubit Clifford. a, Standard randomized benchmarking, simulated using the GST-estimated gate set, yields a ‘reference’ decay rate of rr = 22.2(2)%, suggesting an average per-gate error rate of rr/14.58 ≈ 1.5%. 1σ confidence intervals are indicated in parentheses. bf, Simulated interleaved randomized benchmarking for the CZ gate, and one-qubit Xπ/2 and Yπ/2 gates on each qubit, yielded interleaved decay rates rr + ri. For each experiment, 1,000 random Clifford sequences were generated, at each of 15 circuit depths m, and simulated using the GST process matrices. Exact probabilities (effectively infinitely many shots of each sequence) were recorded. Inset histograms show the distribution over 1000 random circuits at m = 4. Observed decays are consistent with each gate’s GST-estimated infidelities—for example, 1 − F = 0.79% for the CZ gate (b). Performing these exact randomized benchmarking experiments in the lab would have required running 90,000 circuits to estimate a single parameter (ri) for each gate to the given precision of ±0.25%. Using fewer (<1,000) random circuits at each m would yield lower precision. GST required only 1,500 circuits to estimate all error rates to the same precision.

Extended Data Table 1 Estimated state preparation and measurement (SPAM) error rates

Supplementary information

Supplementary Information

This file contains supplemental information supporting the main claims of the paper. The information covers the following three components: supplemental data; extended GST analysis; and an analysis of possible causes for the observed two-qubit entangling errors.

Peer Review File

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mądzik, M.T., Asaad, S., Youssry, A. et al. Precision tomography of a three-qubit donor quantum processor in silicon. Nature 601, 348–353 (2022). https://doi.org/10.1038/s41586-021-04292-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-04292-7

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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