Skip to main content
Log in

Evaluating three levels of quantum metrics on quantum-inspire hardware

  • Published:
Quantum Information Processing Aims and scope Submit manuscript

Abstract

With the rise of quantum computing, many quantum devices have been developed and many more devices are being developed as we speak. This begs the question of which device excels at which tasks and how to compare these different quantum devices with one another. The answer is given by quantum metrics, of which many exist today already. Different metrics focus on different aspects of (quantum) devices and choosing the right metric to benchmark one device against another is a difficult choice. In this paper, we aim to give an overview of this zoo of metrics by grouping established metrics in three levels: component level, system level and application level. With this characterization, we also mention what the merits and uses are for each of the different levels. In addition, we evaluate these metrics on the Starmon-5 device of Quantum Inspire through the cloud access, giving the most complete benchmark of a quantum device from an user experience to date.

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
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Data availability

All data generated or analyzed during this study are included in this published article (and its supplementary information files).

References

  1. Chuang, I.L., Gershenfeld, N., Kubinec, M.: Experimental implementation of fast quantum searching. Phys. Rev. Lett. 80, 3408–3411 (1998). https://doi.org/10.1103/PhysRevLett.80.3408

    Article  ADS  Google Scholar 

  2. IBM: IBM Quantum Computing: Roadmap (2015)

  3. IBM: IBM Unveils 400 qubit-plus quantum processor and next-generation IBM quantum system two (2022)

  4. D-wave systems Inc.: the advantage\(^{{\rm TM}}\) quantum computer: D-wave

  5. Moore, G.E.: Cramming more components onto integrated circuits. Proc. IEEE 86(1), 82–85 (1998) https://doi.org/10.1109/JPROC.1998.658762

  6. Jurcevic, P., Javadi-Abhari, A., Bishop, L.S., Lauer, I., Bogorin, D.F., Brink, M., Capelluto, L., Günlük, O., Itoko, T., Kanazawa, N., Kandala, A., Keefe, G.A., Krsulich, K., Landers, W., Lewandowski, E.P., McClure, D.T., Nannicini, G., Narasgond, A., Nayfeh, H.M., Pritchett, E., Rothwell, M.B., Srinivasan, S., Sundaresan, N., Wang, C., Wei, K.X., Wood, C.J., Yau, J.-B., Zhang, E.J., Dial, O.E., Chow, J.M., Gambetta, J.M.: Demonstration of quantum volume 64 on a superconducting quantum computing system Quantum Sci. Technol. 6, (2021). https://doi.org/10.1088/2058-9565/abe519

  7. Schoot, W., Leermakers, D., Wezeman, R.S., Neumann, N.M.P., Phillipson, F.: Evaluating the Q-score of quantum annealers. In: 2022 IEEE International Conference on Quantum Software (QSW), pp 9–16 (2022)

  8. Last, T., Samkharadze, N., Eendebak, P., Versluis, R., Xue, X., Sammak, A., Brousse, D., Loh, K., Polinder, H., Scappucci, G., Veldhorst, M., Vandersypen, L., Maturová, K., Veltin, J., Alberts, G.: Quantum Inspire: QuTech’s platform for co-development and collaboration in quantum computing. In: Sanchez, M.I., Panning, E.M. (eds.) Novel Patterning Technologies for Semiconductors, MEMS/NEMS and MOEMS 2020, vol. 11324, p. 113240. SPIE, Delft, The Netherlands (2020). https://doi.org/10.1117/12.2551853 . International Society for Optics and Photonics

  9. QuTech: STARMON-5: operational specifics. https://www.quantum-inspire.com/kbase/starmon-5-operational-specifics/. Accessed 11 Dec 2023

  10. Magesan, E., Gambetta, J.M., Emerson, J.: Scalable and robust randomized benchmarking of quantum processes. Phys. Rev. Lett. 106, 180504 (2011). https://doi.org/10.1103/PhysRevLett.106.180504

    Article  ADS  Google Scholar 

  11. Nielsen, E., Gamble, J.K., Rudinger, K., Scholten, T., Young, K., Blume-Kohout, R.: Gate set tomography. Quantum 5, 557 (2021). https://doi.org/10.22331/q-2021-10-05-557

    Article  Google Scholar 

  12. Medford, J., Cywinski, L., Barthel, C., Marcus, C.M., Hanson, M.P., Gossard, A.C.: Scaling of dynamical decoupling for spin qubits. Phys. Rev. Lett. 108, 086802 (2012). https://doi.org/10.1103/PhysRevLett.108.086802

    Article  ADS  Google Scholar 

  13. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information, p. 676. Cambridge University Press, Cambridge (2000)

    Google Scholar 

  14. Berg, E.v.d., Minev, Z.K., Kandala, A., Temme, K.: Probabilistic error cancellation with sparse Pauli-Lindblad models on noisy quantum processors. arXiv (2022). https://doi.org/10.48550/ARXIV.2201.09866

  15. Giurgica-Tiron, T., Hindy, Y., LaRose, R., Mari, A., Zeng, W.J.: Digital zero noise extrapolation for quantum error mitigation. In: 2020 IEEE International Conference on Quantum Computing and Engineering (QCE), pp. 306–316 (2020). https://doi.org/10.1109/QCE49297.2020.00045

  16. QuTech: measurement error mitigation on quantum inspire

  17. Cross, A.W., Bishop, L.S., Sheldon, S., Nation, P.D., Gambetta, J.M.: Validating quantum computers using randomized model circuits. Phys. Rev. A 100, 032328 (2019). https://doi.org/10.1103/PhysRevA.100.032328

    Article  ADS  Google Scholar 

  18. Wack, A., Paik, H., Javadi-Abhari, A., Jurcevic, P., Faro, I., Gambetta, J.M., Johnson, B.R.: Quality, Speed, and Scale: three key attributes to measure the performance of near-term quantum computers. arXiv (2021). https://doi.org/10.48550/ARXIV.2110.14108

  19. Lubinski, T., Johri, S., Varosy, P., Coleman, J., Zhao, L., Necaise, J., Baldwin, C.H., Mayer, K., Proctor, T.: Application-oriented performance benchmarks for quantum computing (2021). arXiv preprint arXiv:2110.03137

  20. Deutsch, D., Jozsa, R.: Rapid solution of problems by quantum computation. In: Proceedings of the Royal Society of London. Series A: Mathematical and Physical Sciences vol 439, pp 553–558 (1992) https://doi.org/10.1098/rspa.1992.0167https://royalsocietypublishing.org/doi/pdf/10.1098/rspa.1992.0167

  21. Coppersmith, D.: An approximate Fourier transform useful in quantum factoring. arXiv (1994). https://doi.org/10.48550/ARXIV.QUANT-PH/0201067

  22. Feynman, R.P.: Simulating physics with computers. International Journal of Theoretical Physics 21(6–7), 467–488 (1982). https://doi.org/10.1007/bf02650179

    Article  ADS  MathSciNet  Google Scholar 

  23. Lanyon, B.P., Whitfield, J.D., Gillett, G.G., Goggin, M.E., Almeida, M.P., Kassal, I., Biamonte, J.D., Mohseni, M., Powell, B.J., Barbieri, M., Aspuru-Guzik, A., White, A.G.: Towards quantum chemistry on a quantum computer. Nat. Chem. 2(2), 106–111 (2010). https://doi.org/10.1038/nchem.483

    Article  Google Scholar 

  24. Martiel, S., Ayral, T., Allouche, C.: Benchmarking quantum coprocessors in an application-centric, hardware-agnostic, and scalable way. IEEE Trans. Quantum Eng. 2, 1–11 (2021). https://doi.org/10.1109/tqe.2021.3090207

    Article  Google Scholar 

  25. Schoot, W., Wezeman, R., Neumann, N.M.P., Phillipson, F., Kooij, R.: Q-score Max-Clique: the first quantum metric evaluation on multiple computational paradigms (2023). arXiv preprint arXiv:2302.00639

  26. QuTech: Quantum Inspire Home. https://www.quantum-inspire.com/. Accessed 21 Dec 2022

  27. Traficante, D.D.: Relaxation. Can t2, be longer than t1? Concepts Magn. Reson. 3(3), 171–177 (1991). https://doi.org/10.1002/cmr.1820030305

    Article  Google Scholar 

  28. Levitt, M.H.: Spin Dynamics: Basics of Nuclear Magnetic Resonance, 2nd edn. Wiley, Hoboken (2008)

    Google Scholar 

  29. QuTech: Quantum Inspire Starmon-5 Fact Sheet. Technical report, QuTech (2022)

  30. Qiskit: Qiskit/qiskit-aer: Aer is a high performance simulator for quantum circuits that includes noise models. https://github.com/Qiskit/qiskit-aer. Accessed 19 Dec 2022

  31. Martiel, S., Ayral, T., Allouche, S.: Qscore. https://github.com/myQLM/qscore/tree/master/qat/qscore. Accessed 11 Dec 2023

  32. Farhi, E., Goldstone, J., Gutmann, S.: A quantum approximate optimization algorithm. arXiv (2014). https://doi.org/10.48550/ARXIV.1411.4028

  33. IBM: Compute Resources. https://quantum-computing.ibm.com/services/resources. Accessed 21 Dec 2023

  34. Watson, T.F., Philips, S.G.J., Kawakami, E., Ward, D.R., Scarlino, P., Veldhorst, M., Savage, D.E., Lagally, M.G., Friesen, M., Coppersmith, S.N., Eriksson, M.A., Vandersypen, L.M.K.: A programmable two-qubit quantum processor in silicon (2017) arXiv:1708.04214

Download references

Acknowledgements

This project has received funding from the European Union’s Horizon 2020 research and innovation programme under grant agreement No. 951852.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ward van der Schoot.

Ethics declarations

Conflict of interest

The research leading to these results received funding from the European Union’s Horizon 2020 research and innovation programme under grant agreement No. 951852.

Additional information

Publisher's Note

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

A Component-level metrics details

A Component-level metrics details

In this appendix we discuss the specific experiments used for the component-level metrics. All experiments are performed by executing a quantum circuit while varying one or more of the circuit parameters.

1.1 A.1 Experiment settings

Unless stated otherwise, each data point was run with 4, 096 shots. In addition, the following setting were used for each experiment:

  • \({T_1}\) Number of different wait-durations: 32, linearly spaced between 0 and the maximum wait-duration of 60 \(\mu s\);

  • \({T_2^*}\) Number of different wait-durations: 32, linearly spaced between 0 and the maximum wait-duration of 24 \(\mu s\), artificial oscillation frequency 0.125 MHz;

  • \({T_2^\textrm{Hahn}}\) Number of different wait-durations (total wait-duration of circuit): 32, linearly spaced between 0 and the maximum wait-duration of 120 \(\mu s\);

  • Single-qubit RB Lengths of Cliffords sequences: \(N \in \left[ 1, 20, 40, 80, 120\right] \). Number of sequences per length: 10.

1.2 A.2 Randomized benchmarking

For each qubit of the Starmon-5 device, the single-qubit fidelity is computed via standard randomized benchmarking. The standard randomized benchmarking protocol [10] consists of circuits that contain a sequence \(C_j\) of N random single-qubit Clifford gates. For each sequence of random Clifford gates \(C_j\), there exists a single-qubit Clifford gate \(C_{N+1}=(\prod C_j)^{-1}\), due to the definition of Clifford gates. For an ideal quantum system, the combined circuit with gates \(C_j\), \(j=1, \ldots , N+1\) acts as the identity. In practice, due to errors in the gates, errors will increase exponentially with the number of Cliffords N, resulting in incorrect measurement results. For each value of N we generate multiple sequences of random Cliffords and apply them to the qubit initialized in state \(|0\rangle \). For each sequence, the resulting qubit state is measured. The following model can then be fitted to the data

$$\begin{aligned} F(N) = A \alpha ^N + B, \end{aligned}$$
(3)

where F(N) is the fraction of \(|1\rangle \) measurements and A and B are constants to be determined. Notice that F(N) exactly counts the fraction of incorrect measurements. From the parameter \(\alpha \) one can calculate the error per Clifford r (EPC) using the equation

$$\begin{aligned} r = \frac{1}{2}(1-\alpha ). \end{aligned}$$
(4)

The single-qubit gate fidelity is determined from the average number of native gates required for a random Clifford, which is 1.875 on Starmon-5. The single-qubit gate fidelity can hence be computed as \(\textrm{F}_{\textrm{1Q}}=(1-r)^{(1/1.875)}\).

Fig. 10
figure 10

Quantum circuit diagrams for the single-qubit coherence measurements

1.3 A.3 Single-qubit coherence time experiments

In this work, three different single-qubit coherence times are computed: the \(T_1\), \(T_2^\textrm{Hahn}\) and \(T_2^*\) time. These values are constants used in a function fitted to the data, computed as the number of \(|1\rangle \) measurements after applying a certain quantum circuit. The quantum circuit for each coherence time experiment can be found in Fig. 10.

For the \(T_1\) measurement, the qubit is prepared in the \(|1\rangle \) state by applying an X-gate to the \(|0\rangle \) state. The qubit is then measured after a variable waiting duration, called the wait-duration. Typically, a wait-duration of up to a few times the \(T_1\) time is used. The system remains idle during the waiting stage, but the qubit can decay from the \(|1\rangle \) state to the \(|0\rangle \) state. This decay results in an exponential decay of the number of measured \(|1\rangle \) states. The number of measured \(|1\rangle \) states for variable wait-durations t is fitted to the function \(F(t)=A +B e^{-t/T_1}\), from which the \(T_1\) time can be computed.

To measure the \(T_2^*\) time of the system, an \(\sqrt{X}\)-gate is applied to the \(|0\rangle \)-state to bring the system to an equal superposition of the 0- and 1-state. Afterward, an \(R_Z(\phi )\)-rotation is applied, again after a waiting stage with variable waiting time. The \(R_Z(\phi )\) is applied with an angle of \(\phi (t) = \sin (\phi _R t)\) that depends on the wait-duration t and the artificial frequency shift \(\phi _R\). Lastly, a \(\sqrt{X}\)-gate is applied to map the system back. The result of the circuit is a damped oscillation with oscillation frequency \(\phi _R\). The oscillation frequency is chosen such that in the total measurement a few oscillations occur so that we can properly fit the model. The artificial frequency shift [34] is added to the circuit to prevent confounding between the damping of the signal (due to the dephasing) and a very low frequency oscillation (due to a frequency mismatch between the driving signal and the qubit resonance frequency). After performing all the experiments, the number of \(|1\rangle \) states for variable wait-durations is fitted to the functions \(F(t) = A + B e^{-t/T_2^*} \sin (\omega t + \phi )\), from which the \(T_2^*\) time can be computed.

The \(T_2^\textrm{Hahn}\) time is measured in a similar fashion as the \(T_2^*\) time. Again, a \(\sqrt{X}\)-gate is applied to the \(|0\rangle \) state, after which the system remains idle for a variable waiting time. Then, an X gate is applied as a single refocusing pulse. Then, the system remains idle for the same variable wait-duration, after which another \(\sqrt{X}\) is finally applied. An noiseless device would always measure the resulting state as \(|0\rangle \), but due to noise the number of \(|1\rangle \) measurements for increasing wait-durations t follows an exponential increase. By fitting this number to the function \(F(t)=A +B e^{-t/T_2^\textrm{Hahn}}\), the \(T_2^\textrm{Hahn}\) time can be computed.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

van der Schoot, W., Wezeman, R., Eendebak, P.T. et al. Evaluating three levels of quantum metrics on quantum-inspire hardware. Quantum Inf Process 22, 451 (2023). https://doi.org/10.1007/s11128-023-04184-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11128-023-04184-x

Keywords

Navigation