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:

A solid-state quantum microscope for wavefunction control of an atom-based quantum dot device in silicon

Abstract

Quantum states of atomic systems can be directly addressed using quantum optical microscopes. However, solid-state microscopy techniques cannot typically achieve both local measurements and control of the state due to their measurement mechanism or the presence of a globally conductive substrate that impedes local gate control. Here we report a solid-state quantum microscope that can control and locally probe the wavefunctions of atomic quantum dots in silicon. Our microscope consists of a scanning tunnelling microscope tip, source and gate electrodes defined on an insulating silicon substrate by subsurface antimony implantation and phosphorus dopants incorporated with atomic precision. In contrast to conventional semiconductor qubit devices, the macroscopic electrodes are fabricated before patterning the nanoscale elements. A light-assisted method is designed to make the substrate conductive to stabilize the microscope tip close to the quantum dots, before reversing to an insulator for local gating and spectroscopy. We show that the microscope can be used to tune and map the charge states of single and double quantum dots, as well as control the relative electrochemical potential between two dots.

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: Quantum microscopy of an atomic device in silicon.
Fig. 2: Spatially and spectroscopically resolving an atomically engineered QD.
Fig. 3: Tuning the strength of the STM probe.
Fig. 4: Aligning two QDs in space and energy.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

References

  1. Koenraad, P. & Flatté, M. E. Single dopants in semiconductors. Nat. Mater. 10, 91–100 (2011).

    Article  Google Scholar 

  2. Georgescu, I. M., Ashhab, S. & Nori, F. Quantum simulation. Rev. Mod. Phys. 86, 153–185 (2014).

    Article  Google Scholar 

  3. Gross, C. & Bakr, W. S. Quantum gas microscopy for single atom and spin detection. Nat. Phys. 17, 1316–1323 (2021).

    Article  Google Scholar 

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

    Article  Google Scholar 

  5. Zhang, J. et al. Observation of a many-body dynamical phase transition with a 53-qubit quantum simulator. Nature 551, 601–604 (2017).

    Article  Google Scholar 

  6. Kindem, J. M. et al. Control and single-shot readout of an ion embedded in a nanophotonic cavity. Nature 580, 201–204 (2020).

    Article  Google Scholar 

  7. Yin, C. et al. Optical addressing of an individual erbium ion in silicon. Nature 497, 91–94 (2013).

    Article  Google Scholar 

  8. Thiele, S. et al. Electrically driven nuclear spin resonance in single-molecule magnets. Science 344, 1135–1138 (2014).

    Article  Google Scholar 

  9. Feynman, R. P. Simulating physics with computers. Int. J. Theor. Phys. 21, 467–488 (1982).

    Article  MathSciNet  Google Scholar 

  10. Arute, F. et al. Quantum supremacy using a programmable superconducting processor. Nature 574, 505–510 (2019).

    Article  Google Scholar 

  11. Muhonen, J. et al. Storing quantum information for 30 seconds in a nanoelectronic device. Nat. Nanotechnol. 9, 986–991 (2014).

    Article  Google Scholar 

  12. Hanson, R. et al. Spins in few-electron quantum dots. Rev. Mod. Phys. 79, 1217–1265 (2007).

    Article  Google Scholar 

  13. Veldhorst, M. et al. A two-qubit logic gate in silicon. Nature 526, 410–414 (2015).

    Article  Google Scholar 

  14. Koiller, B., Hu, X. & Das Sarma, S. Exchange in silicon-based quantum computer architecture. Phys. Rev. Lett. 88, 027903 (2001).

    Article  Google Scholar 

  15. 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 

  16. Zimmerman, N. M., Huang, P. & Culcer, D. Valley phase and voltage control of coherent manipulation in Si quantum dots. Nano Lett. 17, 4461–4465 (2017).

    Article  Google Scholar 

  17. Voisin, B. et al. Valley interference and spin exchange at the atomic scale in silicon. Nat. Commun. 11, 6124 (2020).

    Article  Google Scholar 

  18. Mazurenko, A. et al. A cold-atom Fermi-Hubbard antiferromagnet. Nature 545, 462–466 (2017).

    Article  Google Scholar 

  19. Bernien, H. et al. Probing many-body dynamics on a 51-atom quantum simulator. Nature 551, 579–584 (2017).

    Article  Google Scholar 

  20. Dehollain, J. P. et al. Nagaoka ferromagnetism observed in a quantum dot plaquette. Nature 579, 528–533 (2020).

    Article  Google Scholar 

  21. Crommie, M. F. et al. Confinement of electrons to quantum corrals on a metal surface. Science 262, 218–220 (1993).

    Article  Google Scholar 

  22. Nadj-Perge, S. et al. Observation of Majorana fermions in ferromagnetic atomic chains on a superconductor. Science 346, 602–607 (2014).

    Article  Google Scholar 

  23. Martínez-Blanco, J. et al. Gating a single-molecule transistor with individual atoms. Nat. Phys. 11, 640–644 (2015).

    Article  Google Scholar 

  24. Paul, W. et al. Control of the millisecond spin lifetime of an electrically probed atom. Nat. Phys. 13, 403–407 (2017).

    Article  Google Scholar 

  25. Drost, R. et al. Topological states in engineered atomic lattices. Nat. Phys. 13, 668–671 (2017).

    Article  Google Scholar 

  26. Kempkes, S. N. et al. Robust zero-energy modes in an electronic higher-order topological insulator. Nat. Mater. 18, 1292–1297 (2019).

    Article  Google Scholar 

  27. Topinka, M. A. et al. Imaging coherent electron flow from a quantum point contact. Science 289, 2323–2326 (2000).

    Article  Google Scholar 

  28. Woodside, M. T. & McEuen, P. L. Scanned probe imaging of single-electron charge states in nanotube quantum dots. Science 296, 1098–1101 (2002).

    Article  Google Scholar 

  29. Shapir, I. et al. Imaging the electronic Wigner crystal in one dimension. Science 364, 870–875 (2019).

    Article  Google Scholar 

  30. Bleszynski-Jayich, A. C. et al. Imaging a one-electron InAs quantum dot in an InAs/InP nanowire. Phys. Rev. B 77, 245327 (2008).

    Article  Google Scholar 

  31. Tachizaki, T., Hayashi, K., Kanemitsu, Y. & Hirori, H. On the progress of ultrafast time-resolved THz scanning tunneling microscopy. APL Mater. 9, 060903 (2021).

    Article  Google Scholar 

  32. Yang, K. et al. Coherent spin manipulation of individual atoms on a surface. Science 366, 509–512 (2019).

    Article  Google Scholar 

  33. LeRoy, B. J. et al. Three-terminal scanning tunneling spectroscopy of suspended carbon nanotubes. Phys. Rev. B 72, 075413 (2005).

    Article  Google Scholar 

  34. Tsai, H. Z. et al. A molecular shift register made using tunable charge patterns in one-dimensional molecular arrays on graphene. Nat. Electron. 3, 598–603 (2020).

    Article  Google Scholar 

  35. Fuechsle, M. et al. A single-atom transistor. Nat. Nanotechnol. 7, 242–246 (2012).

    Article  Google Scholar 

  36. Ng, K. S. H. et al. Scanned single-electron probe inside a silicon electronic device. ACS Nano 14, 9449–9455 (2020).

    Article  Google Scholar 

  37. Kim, J. C., Kline, J. S. & Tucker, J. R. Fabrication of contact electrodes in Si for nanoelectronic devices using ion implantation. Appl. Surf. Sci. 239, 335–341 (2005).

    Article  Google Scholar 

  38. Ward, D. R. et al. All-optical lithography process for contacting nanometer precision donor devices. Appl. Phys. Lett. 111, 193101 (2017).

    Article  Google Scholar 

  39. Ramanayaka, A. N. et al. STM patterned nanowire measurements using photolithographically defined implants in Si(100). Sci. Rep. 8, 1790 (2018).

    Article  Google Scholar 

  40. Nylandsted Larsen, A. et al. The nature of electrically inactive antimony in silicon. J. Appl. Phys. 59, 1908–1917 (1986).

    Article  Google Scholar 

  41. Wang, Y. et al. Highly tunable exchange in donor qubits in silicon. npj Quantum Inf. 2, 16008 (2016).

    Article  Google Scholar 

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

    Article  Google Scholar 

  43. Mądzik, M. T. et al. Precision tomography of a three-qubit donor quantum processor in silicon. Nature 601, 348–353 (2022).

  44. Wyrick, J. et al. Atom-by-atom fabrication of single and few dopant quantum devices. Adv. Funct. Mater. 29, 1903475 (2019).

    Article  Google Scholar 

  45. Roche, B. et al. A tunable, dual mode field-effect or single electron transistor. Appl. Phys. Lett. 100, 032107 (2012).

    Article  Google Scholar 

  46. Weber, B. et al. Spin blockade and exchange in Coulomb-confined silicon double quantum dots. Nat. Nanotechnol. 9, 430–435 (2014).

    Article  Google Scholar 

  47. Weber, B. et al. Engineering independent electrostatic control of atomic-scale (~4 nm) silicon double quantum dots. Nano Lett. 12, 4001–4006 (2012).

    Article  Google Scholar 

  48. Wijnheijmer, A. P. et al. Single Si dopants in GaAs studied by scanning tunneling microscopy and spectroscopy. Phys. Rev. B 84, 125310 (2011).

    Article  Google Scholar 

  49. Rontani, M. & Molinari, E. Imaging quasiparticle wave functions in quantum dots via tunneling spectroscopy. Phys. Rev. B 71, 233106 (2005).

    Article  Google Scholar 

  50. Salfi, J. et al. Quantum simulation of the Hubbard model with dopant atoms in silicon. Nat. Commun. 7, 11342 (2016).

    Article  Google Scholar 

  51. Usman, M. et al. Spatial metrology of dopants in silicon with exact lattice site precision. Nat. Nanotechnol. 11, 763–768 (2016).

    Article  Google Scholar 

  52. Salfi, J. et al. Spatially resolving valley quantum interference of a donor in silicon. Nat. Mater. 13, 605–610 (2014).

    Article  Google Scholar 

  53. Keith, D. et al. Benchmarking high fidelity single-shot readout of semiconductor qubits. New J. Phys. 21, 063011 (2019).

    Article  Google Scholar 

  54. Bonet, E., Deshmukh, M. M. & Ralph, D. C. Solving rate equations for electron tunneling via discrete quantum states. Phys. Rev. B 65, 045317 (2002).

    Article  Google Scholar 

  55. Broome, M. A. et al. Two-electron spin correlations in precision placed donors in silicon. Nat. Commun. 9, 980 (2018).

    Article  Google Scholar 

  56. Hallam, T. et al. Effective removal of hydrogen resists used to pattern devices in silicon using scanning tunneling microscopy. Appl. Phys. Lett. 86, 143116 (2005).

    Article  Google Scholar 

  57. Schofield, S. et al. Quantum engineering at the silicon surface using dangling bonds. Nat. Commun. 4, 1649 (2013).

    Article  Google Scholar 

  58. Van der Wiel, W. G. et al. Electron transport through double quantum dots. Rev. Mod. Phys. 75, 1–22 (2002).

    Article  Google Scholar 

  59. Kiczynski, M. et al. Engineering topological states in atom-based semiconductor quantum dots. Nature 606, 694–699 (2022).

    Article  Google Scholar 

  60. Le, N. H. et al. Topological phases of a dimerized Fermi-Hubbard model for semiconductor nano-lattices. npj Quantum Inf. 6, 24 (2020).

    Article  Google Scholar 

  61. Škereň, T. et al. Bipolar device fabrication using a scanning tunnelling microscope. Nat. Electron. 3, 524–530 (2020).

    Article  Google Scholar 

  62. Mannini, M. et al. Magnetic behaviour of TbPc2 single-molecule magnets chemically grafted on silicon surface. Nat. Commun. 5, 4582 (2014).

    Article  Google Scholar 

  63. Nickel, A. et al. Electronically driven single-molecule switch on silicon dangling bonds. J. Phys. Chem. C 120, 27027–27032 (2016).

    Article  Google Scholar 

  64. Huff, T. et al. Binary atomic silicon logic. Nat. Electron. 1, 636–643 (2018).

    Article  Google Scholar 

Download references

Acknowledgements

We acknowledge support from the ARC Centre of Excellence for Quantum Computation and Communication Technology (CE170100012), an ARC Discovery Project (DP180102620), Silicon Quantum Computing Pty Ltd, the US Army Research Office (W911NF-17-1-0202) and from the NSW and ACT Nodes of the Australian National Fabrication Facility. J.S. acknowledges support from an ARC DECRA fellowship (DE160101490). B.C.J. and J.C.M. acknowledge the AFAiiR node of the NCRIS Heavy Ion Capability for access to the ion implantation facilities.

Author information

Authors and Affiliations

Authors

Contributions

B.V., J.S. and S.R. designed the project. B.C.J. and J.C.M. designed and performed the stopping and range of ions in matter simulations, antimony implantation and Rutherford backscattering measurements. B.V. performed the atomic fabrication, low-temperature measurements and data analysis, with J.S., M.Y.S. and S.R. providing inputs. D.D.S.M. performed the electrostatic simulation of the device, with J.S., B.V. and S.R. providing inputs. B.V. wrote the manuscript with contribution of all authors.

Corresponding authors

Correspondence to B. Voisin or S. Rogge.

Ethics declarations

Competing interests

M.Y.S. is a director of the company Silicon Quantum Computing Pty Ltd.

Peer review

Peer review information

Nature Electronics thanks the anonymous reviewers for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Implantation procedure and simulation.

a, Implantation process. The implantation energy is chosen to overcome a 22.5 nm sacrificial oxide barrier. The sample presented here was implanted with a double dose of 60 keV and 100 keV with respective doses of 4 × 1014 and 1 × 1015 cm2. These energies are less than that of the sample presented in the manuscript (70 and 110 keV) to compensate for a thinner oxide (22.5 vs 27 nm). b, Stopping and Range of ions in matter (SRIM) simulations of the implantation profile, for 60 keV (black), 100 keV (red) and total (blue). The concentration peaks around 40 nm in the silicon substrate, and is largely above the metal/insulator transition at the SiO2/Si interface. The lateral range of these implants is less than 20 nm while the diffusion upon flash annealing is less than 25 nm. This is sufficient to ensure Sb is confined to the electrodes and does not diffuse into the regions where the STM is performed.

Extended Data Fig. 2 Implantation profile analysis.

a, Comparison of the SRIM simulation and (black) and of Rutherford back-scattering (RBS) measurements of a sample annealed at 660C for 15h. The depth of the SRIM data was offset to match the onset of the RBS signal, as the sacrificial oxide was etched for the RBS samples. The random data (red) correspond to the total amount of Sb, the channeled data (blue) correspond to interstitial and unactivated Sb atoms. b, Random RBS data for different flash anneal conditions: 660C for 15h (red), followed by a 1100C/20 s (dark red) or a 1200C/20 s flash (brown). The 1100C used for the device leaves a high concentration of antimony atoms in the top few nanometers of the silicon substrate. c, Sb sheet density, determined as the area of the random RBS signal, as a function of the flash anneal conditions. The density remains above the metal-insulator transition despite a reduction due to deep diffusion in the substrate. d, Sb activated fraction, determined as 1 − Ac/Ar where Ar and Ac are the area under the random and channeled RBS signals, respectively. The first point (no annealing) is close to zero as annealing is necessary to activate the Sb atoms, it is slightly negative due to a statistical difference between the random and channeled measurements. The activated fraction is above 80% once the sample is annealed, the error bars increase with annealing temperature due to lower Sb counts in the channeled data. The mean values and error bars shown for the sheet density and activated fraction are obtained from least-square Gaussian fits to the RBS measurements, with details given in Supplementary Information Section 1.

Extended Data Fig. 3 Measurement protocol.

a, Measurement protocol. The transition between the topography mode (feedback loop and light on, Vb = Vg = − 1.6 V) and the spectroscopy mode is defined as follows. The light and feedback loop are first turned off, and then the bias and gate voltages are swept to the initial values of the spectrum to be acquired, which are closer to zero with respect to topography settings. The tip is then brought closer to the sample. This order, gate and bias voltages first before tip height offset, ensures for the tunnel current not to exceed 1 nA, which could create spurious hydrogen desorption at the surface. The spectroscopy diagram shown here is designed to record the tunnel current as a function of the bias voltage, for a fixed gate voltage and tip position. Once the spectroscopy is completed, a transition is implemented to revert back to topography settings. A similar protocol is implemented for spatially resolved spectroscopy. b, Current Isg between the source and the gate when the light is off. In this configuration, the transimpedance amplifier is connected to the gate electrode and the tip was left floating. Minimal leakage current is observed for a large bias range Vb − Vg from -1.5 to +1.5 V, with a linear fit of the region around Vb − Vg = 0 V (black dotted line) yielding Rsg = 9.45 GΩ, similar to other dopant-based devices.

Extended Data Fig. 4 Evolution of the QD charge distribution upon increasing filling factor.

a-c STM images of first three charge transitions for QD#1, same as main text. d Average line cut of the normalised STM images, taken along the horizontal grey section #1 shown in a. The third transition shows more charge distribution in the bottom left corner of the QD designed area. e Average line cut taken along the horizontal grey section #2 shown in a. The third transition shows more charge distribution around the top right section of thew QD designed area.

Extended Data Fig. 5 Wavefunction and spectroscopy data for QD#2.

a, Stability diagram Isd vs Vb, Vg. This QD shows more resonances than QD#1 indicating that more donors have been incorporated. b, Wavefunction image taken at Vg = − 0.35 V and Vb = − 0.23 V. The dashed rectangle highlights the size of the lithography area. Beside drift correction, an additional Fourier filtering was performed for this image as the tunnel current was close to the noise background. c, Wavefunction image taken at Vg = 0 V and Vb = − 0.8 V. At such large occupation number and large source-drain bias, the wavefunction extends outside the lithography area. d, Spectroscopy line Isd vs Vb and x taken across the reservoir and the QD at Vg = 0 V. e, Same as b taken at Vg = -1 V. centred on the QD. The plateaux in tunnel current vs Vb indicate the addition of electrons on the QD. f Spectroscopy line cuts extracted from b and c, see arrows for colour code. There is no tunnel current within the silicon gap away from the QD and the reservoir. The Coulomb plateaux are visible when the tip is located over the QD. The silicon gap completely closes when the tip is above the source reservoir (black line), which evidences the continuity in the metallic contact between the antimony implanted electrode and its atomic-scale phosphorus extension.

Extended Data Fig. 6 Large scale image of the double dot device.

a, Topography image of the double dot device lithography, same as main text. b, Quantum state image taken at Vb = Vg = − 0.8 V, same scale as the lithography image. As in the main text, we take as reference to align the furthest dot from the source (called LD). The two dots are barely visible at this lower gate and bias voltages, but direct tunnel current over the reservoir can be evidenced. A diffusion of the source lead Δdlead = 2.0 nm towards the quantum dots can be detected in this image, attributed to the high concentration of dopants in this element. c Large scale topography image (Vb = Vg = − 2.1 V) around the DQD (associated topography image to the quantum state image shown in b), where the dangling bonds can be evidenced as bright features. d Small scale topography image (Vb = Vg = − 2.1 V) around the DQD, corresponding to the dashed square in b and c. A difference in the dangling bond brightness can be observed between the topography and the quantum state images. The silicon substrate is illuminated and conductive in topography mode, resulting in all dangling bonds having a similar brightness. On the contrary, only the dangling bonds close to incorporated dopants (DQD or source) are bright in the quantum state image, as the sample is insulating in this mode and transport must occur through conductive elements.

Supplementary information

Supplementary Information

Supplementary Sections 1–3 and Figs. 1–3.

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

Voisin, B., Salfi, J., St Médar, D.D. et al. A solid-state quantum microscope for wavefunction control of an atom-based quantum dot device in silicon. Nat Electron 6, 409–416 (2023). https://doi.org/10.1038/s41928-023-00979-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41928-023-00979-z

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