Next Article in Journal
An Efficient Deep Learning Approach for Colon Cancer Detection
Next Article in Special Issue
Quasi-Shor Algorithms for Global Benchmarking of Universal Quantum Processors
Previous Article in Journal
VOC-Net: A Deep Learning Model for the Automated Classification of Rotational THz Spectra of Volatile Organic Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Measurement of Quasiparticle Diffusion in a Superconducting Transmon Qubit

National Laboratory of Solid State Microstructures, School of Physics, Nanjing University, Nanjing 210093, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2022, 12(17), 8461; https://doi.org/10.3390/app12178461
Submission received: 28 July 2022 / Revised: 18 August 2022 / Accepted: 22 August 2022 / Published: 24 August 2022
(This article belongs to the Special Issue Superconducting Quantum Computing and Devices)

Abstract

:
Quasiparticles, especially the ones near the Josephson junctions in the superconducting qubits, are known as an important source of decoherence. By injecting quasiparticles into a quantum chip, we characterized the diffusion feature by measuring the energy relaxation time and the residual excited-state population of a transmon qubit. From the extracted transition rates, we phenomenologically modeled the quasiparticle diffusion in a superconducting circuit that contained “hot” nonequilibrium quasiparticles in addition to low-energy ones.

1. Introduction

The superconducting quantum circuit has become a promising candidate for realizing universal quantum computation for rapid development in recent years. At this stage, the prospective superconducting qubits are aimed at longer coherence times and higher readout fidelities. It has been known that some loss channels—such as dielectric loss from two-level defects [1]—and quasiparticles [2,3,4,5,6,7,8] can set limits on qubit coherence. Among them, the quasiparticles that may absorb the qubit energy when tunneling through the junction are regarded as the dominant loss for the microsecond limit on the energy relaxation time T 1 [9,10].
While the quasiparticle density, which is assumed to be in thermal equilibrium, is suppressed exponentially as the bath temperature decreases, recently observed experimental results have shown that the normalized nonequilibrium quasiparticle density is in a range of 10 8 10 6 , which is several orders of magnitude larger than the “cold” equilibrium one at a 20 mK bath temperature [9,11,12,13,14,15,16]. The nonequilibrium quasiparticles are attributed as a source of the loss mechanism and residual excited-state population. As illustrated in Figure 1a, quasiparticles near the junction with energies E q p of all ranges can absorb the qubit-transition energy and transfer the qubit from the excited state to the ground state, and the “hot” ones that have energies greater than Δ + E g e may release energy and excite the qubit [17,18].
Several techniques have been proposed to depict the nonequilibrium quasiparticles in superconducting qubits, such as treating the transmon qubit as a real-time quasiparticle-tunneling detector [9,19], the charge-parity correlation measurement [10], and direct dispersive monitoring [20]. However, most of these measurements require high-fidelity dispersive readouts or the transmon qubit in a low E J / E C regime. In this letter, we propose a simple and practical protocol, which is demonstrated on a normally anharmonic ( E J / E C 50 ) single-junction transmon qubit. We study the features of quasiparticles in the sample with the contactless quasiparticle injection technique [21] and read out with the routine called “high-power readout” [22], in which the high demand for dispersive readout fidelities is not necessary. In the experiment, the energy relaxation time T 1 and the residual excited-state population P e r are measured with different parameters of injected quasiparticles. From the down-transition rates of the qubit and the energy distribution of nonequilibrium quasiparticles extracted from the experimental data, we phenomenologically model the quasiparticle diffusion [23,24].

2. Device and Quasiparticle Injection

The single-junction transmon qubit that we used in the experiment was centered in a 3D cavity that was made of aluminum (Al 6061-T6 alloy) and had a size of 35.56 × 5.08 × 17.78 mm. The bare frequency of the cavity ω b a r e / 2 π was 9.0524 GHz. The Josephson junction was located in the middle of two thin-film electrodes that had the size of 250 × 500   μ m, as illustrated in Figure 1b. The qubit was cooled down to 30 mK in an Oxford DR200 cryogen-free dilution refrigerator with the transition frequencies of ω g e / 2 π = 7.0905 GHz and ω e f / 2 π = 6.748 GHz and the ratio of Josephson coupling energy to charging energy of E J / E C = 58 . The average measured energy relaxation time T 1 was 7.29 μ s. The data were obtained by repeating the measurement 1000 times with a visibility of 50 % . More details about the sample can be found in Appendix A.
In this work, we modified the technique of quasiparticle injection reported in [21] by changing the injection duration time to 3–5 μ s from a hundred microseconds. This allowed us to study the diffusion of the quasiparticles in the qubit. The injection pulse was input at the frequency of ω b a r e with the power [21,22] corresponding to the one on the right side of the red line in Figure 1c. For convenience, we mark the power of 0 dB with the red line. The gap between the injection pulse and the measurement pulse in Figure 1d is the diffusion time of the quasiparticles.

3. Quasiparticle Probe

In the high-frequency regime ω g e E q p Δ , the down-transition rate induced by quasiparticles Γ q p for a single-junction transmon qubit is [5,7]:
Γ q p = ω p 2 ω g e π 2 Δ E g e χ q p ,
where ω p = 8 E C E J is the plasma frequency and χ q p is the quasiparticle density per Cooper pair.
When the qubit is only excited thermally, the residual excited-state population is P e r = e ω g e k B T / ( e ω g e k B T + 1 ) = 1.2 × 10 5 . The measured residual excited-state population is P e r = 0.1208 , corresponding to an effective qubit temperature of 174 mK, which is much higher than that of the bath environment of T = 30 mK. This phenomenon indicates that the qubit is also partly excited by other sources, such as the “hot” quasiparticles. Hence, despite the uncharted energy distribution of “hot” quasiparticles, it is deemed that the qubit excitation rate can be written in the implicit function of the quasiparticle density χ q p , that is, Γ ( χ q p ) .
Consequently, we measure the energy relaxation time T 1 and the residual excited-state population P e r to study the quasiparticles in the superconducting qubit. Since the longitudinal relaxation rate Γ l = 1 / T 1 , we can obtain both the up- and down-transition rates Γ ( Γ ) from the measured T 1 and P e r , which can be written as [25]:
T 1 = 1 Γ Γ . P e r = Γ Γ + Γ .
In the T = 30 mK environment, the qubit excitation rate in thermal equilibrium can be neglected, since Γ = e ω g e / ( k B T ) Γ 10 23 Γ ; as a result, the relaxation rate can simplified as Γ = 1 / T 1 .
In this work, we measure T 1 and P e r with different densities of quasiparticles by adjusting the parameters of the injection pulses. The power and the pulse length of the applied microwave correspond to the density of the quasiparticles, while the time gap between the injection and measurement denotes the diffusion time t d i f f . As shown in Figure 2, the energy relaxation time and the residual excited-state population vary according to different parameters of injection. Figure 2a,b show the experimental data of T 1 and P e r , respectively. We measured the Rabi oscillation between | 1 and | 2 . As shown in Figure 2b, the reference Rabi oscillation, Rabi r e f , is measured after a π g e pulse is applied to the qubit in the ground state, while the signal one, Rabi s i g , is measured with no front operation. With both traces in Figure 2b, we can obtain the accurate P e r [11]. In Figure 2c,d, T 1 and P e r are plotted as a function of diffusion time, from which we can capture some features of the quasiparticle dynamics. With the increase in the injected microwave power, the qubit is affected by more quasiparticles, resulting in a larger decrease in T 1 . After sufficient diffusion time, T 1 converges to the original value, which means that the energy relaxation time is influenced by quasiparticles. The experimental data in Figure 2 indicate that T 1 and P e r are sensitive to quasiparticles in the superconducting qubit and can be used to characterize the features of quasiparticles.

4. Quasiparticle Diffusion in the Superconducting Qubit

The transition rates Γ and Γ presented in Figure 3a were calculated with Equation (2). Note that Γ are an order of magnitude smaller than Γ , which also almost overlap, indicating that the density of “hot” quasiparticles is independent of the injection power. Meanwhile, there is a positive correlation between Γ and the injection power, demonstrating that most of the generated quasiparticles are “cold” in this injection protocol.
Now, we study the phenomenological energy distribution of the quasiparticles injected into the superconducting qubit. To simplify the calculation, the diffusion, which is clearly related to the energy of the quasiparticles, is taken into account to model the quasiparticles near the junction [21]
χ q p t d i f f = D 2 χ q p .
In Equation (4), D = 60 v q p cm 2 /s is the diffusion constant in aluminum, where the normalized velocity v q p = ( 1 Δ 2 / E q p 2 ) 1 / 2 depends on the quasiparticle’s energy E q p [5,23]. The Gaussian solution of this diffusion equation can be written as
χ q p ( x , t d i f f ) = C 0 π D t d i f f e x p ( x 2 4 D t d i f f ) ,
where C 0 is the initial value of χ q p and t d i f f represents the diffusion time. x is the distance over which the created quasiparticles diffuse from the outer side of the electrode to the junction [23]. For simplicity, diffusion is considered in only one dimension, and x = 0.5 mm in our sample. With Equations (1) and (4), we obtain the relationship between the diffusion time and Γ .
As shown in Figure 3a, the transition rates of the qubit extracted from the experiment change with the diffusion time. The four colored curves point to the four injection powers of 4, 7, 10, and 13 dB, respectively, and the maximum values, which are denoted as Γ m a x , corresponding to four injection powers are 2.586 × 10 4 , 3.454 × 10 4 , 4.468 × 10 4 , and 5.22 × 10 4 . With the increase in the injection power, Γ of the quasiparticle takes more time to reach Γ m a x . This contradicts the previous theoretical model, in which only the “cold” quasiparticles are considered. In this original model, the power is only related to the density of the quasiparticles. With the bath temperature of T = 30 mK, quasiparticles with the average energy E q p = Δ + k T have the same velocity, resulting in the same diffusion feature. Therefore, the diffusion time corresponding to Γ m a x should be the same with different injection powers (see the blue line in Figure 3c). Therefore, we modify the original model by introducing the “hot” quasiparticles with the average energy of E q p = Δ + E g e . Considering the phenomenon in which Γ is independent of the injection power, the density of quasiparticles can be phenomenologically written as:
χ q p ( p , t d i f f ) = χ q p c ( p , t d i f f ) e Γ d c · t d i f f + χ q p h ( t d i f f ) e Γ d h · t d i f f + χ q p 0 .
where p represents the injection power and χ q p 0 represents the quasiparticles that are present when no quasiparticles are injected, and the system is in thermal equilibrium. We assume that the “hot” quasiparticles are independent of the injection power. With the presence of “hot” quasiparticles, the fitted diffusion time corresponding to the Γ m a x changes with different injection powers (see the gray line in Figure 3c), which is consistent with the experimental data in Figure 3a. Furthermore, the decay rates of the quasiparticles Γ d are also introduced to fit the sharply descending trace on the right side of Γ m a x . Here, we assume that recombination and vortex trapping [26,27] are important channels for the decay of quasiparticles. From our model, with the measurement errors taken into account, we extracted the decay rates of the “cold” and “hot” quasiparticles, which were Γ d c = 1 / 140 and Γ d h = 1 / 350 μ s, respectively. With the injection powers of 4, 7, 10, and 13 dB, the fitting C 0 c for “cold” quasiparticles in Equation (4) was 0.75 × C i n i , 1.8 × C i n i , 2.9 × C i n i ,   a n d   4.1 × C i n i , where C i n i = [ 2.8 × 10 3 , 2.95 × 10 3 ] for the injection durations of 3 and 4.5 μ s, and C 0 h of the “hot” quasiparticles is 2 × 10 3 . Figure 3d demonstrates that the fitting results are in good agreement at different injection powers. The deviation between the experimental data and the theoretical fitting could be caused by the following reasons: (1) the transitions between “hot” and “cold” quasiparticles during diffusion; (2) the use of the quasiparticles’ average energies instead of the integral of the energies when calculating the normalized quasiparticle diffusion velocities; (3) recombination and vortex trapping were not taken into account in our model in order to simplify the dynamic features of quasiparticles.

5. Conclusions

The energy decay time T 1 and the residual excited-state population P e r of a qubit were measured to characterize the features of quasiparticles in a superconducting circuit. Additionally, we phenomenologically modeled quasiparticle diffusion with “hot” and “cold” quasiparticles by using the extracted up- and down-transition rates.

Author Contributions

Conceptualization, W.Z.; Data curation, Y.D. and Y.L.; Investigation, Y.Z. and Z.M.; Writing—original draft, Y.D.; Writing—review & editing, X.T. and Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Key R&D Program of Guangdong Province (Grant No. 2018B030326001), NSFC (Grants No. 11474152, No. 12074179, No. U21A20436, and No. 61521001), and NSF of Jiangsu Province (Grant No. BE2021015-1).

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Parameters of the Sample

The transmon was patterned with a 30 KeV e-beam lithography, followed by double-angle shadow evaporation of aluminum (the thicknesses of the layers were 30 and 80 nm, respectively) and lift-off procedures to form a junction on a high-resistivity silicon substrate.
Table A1. Parameters of the sample.
Table A1. Parameters of the sample.
ParameterValue
Bare frequency of the cavity9.0524 GHz
Dispersive frequency of the cavity9.0905 GHz
Qubit frequency7.0905 GHz
Anharmonicity−343 MHz
Cavity–qubit coupling66 MHz
Purcell decay1/630 μ s

References

  1. Lisenfeld, J.; Bilmes, A.; Ustinov, A.V. Enhancing the Coherence of Superconducting Quantum Bits with Electric Fields. arXiv 2022, arXiv:2208.01570. [Google Scholar] [CrossRef]
  2. Martinis, J.M.; Cooper, K.B.; McDermott, R.; Steffen, M.; Ansmann, M.; Osborn, K.D.; Cicak, K.; Oh, S.; Pappas, D.P.; Simmonds, R.W.; et al. Decoherence in Josephson Qubits from Dielectric Loss. Phys. Rev. Lett. 2005, 95, 210503. [Google Scholar] [CrossRef] [PubMed]
  3. Gambetta, J.M.; Murray, C.E.; Fung, Y.K.K.; McClure, D.T.; Dial, O.; Shanks, W.; Sleight, J.W.; Steffen, M. Investigating Surface Loss Effects in Superconducting Transmon Qubits. IEEE Trans. Appl. Supercond. 2017, 27, 1–5. [Google Scholar] [CrossRef]
  4. de Graaf, S.E.; Faoro, L.; Ioffe, L.B.; Mahashabde, S.; Burnett, J.J.; Lindström, T.; Kubatkin, S.E.; Danilov, A.V.; Tzalenchuk, A.Y. Two-level systems in superconducting quantum devices due to trapped quasiparticles. Sci. Adv. 2020, 6, eabc5055. [Google Scholar] [CrossRef] [PubMed]
  5. Martinis, J.M.; Ansmann, M.; Aumentado, J. Energy Decay in Superconducting Josephson-Junction Qubits from Nonequilibrium Quasiparticle Excitations. Phys. Rev. Lett. 2009, 103, 097002. [Google Scholar] [CrossRef]
  6. Córcoles, A.D.; Chow, J.M.; Gambetta, J.M.; Rigetti, C.; Rozen, J.R.; Keefe, G.A.; Beth Rothwell, M.; Ketchen, M.B.; Steffen, M. Protecting superconducting qubits from radiation. Appl. Phys. Lett. 2011, 99, 181906. [Google Scholar] [CrossRef]
  7. Catelani, G.; Schoelkopf, R.J.; Devoret, M.H.; Glazman, L.I. Relaxation and frequency shifts induced by quasiparticles in superconducting qubits. Phys. Rev. B 2011, 84, 064517. [Google Scholar] [CrossRef]
  8. Gustavsson, S.; Yan, F.; Catelani, G.; Bylander, J.; Kamal, A.; Birenbaum, J.; Hover, D.; Rosenberg, D.; Samach, G.; Sears, A.P.; et al. Suppressing relaxation in superconducting qubits by quasiparticle pumping. Science 2016, 354, 1573–1577. [Google Scholar] [CrossRef]
  9. Serniak, K.; Hays, M.; de Lange, G.; Diamond, S.; Shankar, S.; Burkhart, L.D.; Frunzio, L.; Houzet, M.; Devoret, M.H. Hot Nonequilibrium Quasiparticles in Transmon Qubits. Phys. Rev. Lett. 2018, 121, 157701. [Google Scholar] [CrossRef]
  10. Ristè, D.; Bultink, C.C.; Tiggelman, M.J.; Schouten, R.N.; Lehnert, K.W.; DiCarlo, L. Millisecond charge-parity fluctuations and induced decoherence in a superconducting transmon qubit. Nat. Commun. 2013, 4, 1913. [Google Scholar] [CrossRef] [Green Version]
  11. Jin, X.Y.; Kamal, A.; Sears, A.P.; Gudmundsen, T.; Hover, D.; Miloshi, J.; Slattery, R.; Yan, F.; Yoder, J.; Orlando, T.P.; et al. Thermal and Residual Excited-State Population in a 3D Transmon Qubit. Phys. Rev. Lett. 2015, 114, 240501. [Google Scholar] [CrossRef] [PubMed]
  12. de Visser, P.J.; Goldie, D.J.; Diener, P.; Withington, S.; Baselmans, J.J.A.; Klapwijk, T.M. Evidence of a Nonequilibrium Distribution of Quasiparticles in the Microwave Response of a Superconducting Aluminum Resonator. Phys. Rev. Lett. 2014, 112, 047004. [Google Scholar] [CrossRef] [PubMed]
  13. Lenander, M.; Wang, H.; Bialczak, R.C.; Lucero, E.; Mariantoni, M.; Neeley, M.; O’Connell, A.D.; Sank, D.; Weides, M.; Wenner, J.; et al. Measurement of energy decay in superconducting qubits from nonequilibrium quasiparticles. Phys. Rev. B 2011, 84, 024501. [Google Scholar] [CrossRef]
  14. Sun, L.; DiCarlo, L.; Reed, M.D.; Catelani, G.; Bishop, L.S.; Schuster, D.I.; Johnson, B.R.; Yang, G.A.; Frunzio, L.; Glazman, L.; et al. Measurements of Quasiparticle Tunneling Dynamics in a Band-Gap-Engineered Transmon Qubit. Phys. Rev. Lett. 2012, 108, 230509. [Google Scholar] [CrossRef]
  15. Nsanzineza, I.; Plourde, B.L.T. Trapping a Single Vortex and Reducing Quasiparticles in a Superconducting Resonator. Phys. Rev. Lett. 2014, 113, 117002. [Google Scholar] [CrossRef]
  16. Barends, R.; Wenner, J.; Lenander, M.; Chen, Y.; Bialczak, R.C.; Kelly, J.; Lucero, E.; O’Malley, P.; Mariantoni, M.; Sank, D.; et al. Minimizing quasiparticle generation from stray infrared light in superconducting quantum circuits. Appl. Phys. Lett. 2011, 99, 113507. [Google Scholar] [CrossRef]
  17. Catelani, G.; Basko, D.M. Non-equilibrium quasiparticles in superconducting circuits: Photons vs. phonons. SciPost Phys. 2019, 6, 13. [Google Scholar] [CrossRef]
  18. Wenner, J.; Yin, Y.; Lucero, E.; Barends, R.; Chen, Y.; Chiaro, B.; Kelly, J.; Lenander, M.; Mariantoni, M.; Megrant, A.; et al. Excitation of Superconducting Qubits from Hot Nonequilibrium Quasiparticles. Phys. Rev. Lett. 2013, 110, 150502. [Google Scholar] [CrossRef]
  19. Houzet, M.; Serniak, K.; Catelani, G.; Devoret, M.H.; Glazman, L.I. Photon-Assisted Charge-Parity Jumps in a Superconducting Qubit. Phys. Rev. Lett. 2019, 123, 107704. [Google Scholar] [CrossRef]
  20. Serniak, K.; Diamond, S.; Hays, M.; Fatemi, V.; Shankar, S.; Frunzio, L.; Schoelkopf, R.; Devoret, M. Direct Dispersive Monitoring of Charge Parity in Offset-Charge-Sensitive Transmons. Phys. Rev. Appl. 2019, 12, 014052. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, C.; Gao, Y.Y.; Pop, I.M.; Vool, U.; Axline, C.; Brecht, T.; Heeres, R.W.; Frunzio, L.; Devoret, M.H.; Catelani, G.; et al. Measurement and control of quasiparticle dynamics in a superconducting qubit. Nat. Commun. 2014, 5, 5836. [Google Scholar] [CrossRef] [PubMed]
  22. Reed, M.D.; DiCarlo, L.; Johnson, B.R.; Sun, L.; Schuster, D.I.; Frunzio, L.; Schoelkopf, R.J. High-Fidelity Readout in Circuit Quantum Electrodynamics Using the Jaynes-Cummings Nonlinearity. Phys. Rev. Lett. 2010, 105, 173601. [Google Scholar] [CrossRef] [PubMed]
  23. Friedrich, S.; Segall, K.; Gaidis, M.C.; Wilson, C.M.; Prober, D.E.; Szymkowiak, A.E.; Moseley, S.H. Experimental quasiparticle dynamics in a superconducting, imaging x-ray spectrometer. Appl. Phys. Lett. 1997, 71, 3901–3903. [Google Scholar] [CrossRef]
  24. Ullom, J.N.; Fisher, P.A.; Nahum, M. Magnetic field dependence of quasiparticle losses in a superconductor. Appl. Phys. Lett. 1998, 73, 2494–2496. [Google Scholar] [CrossRef]
  25. Krantz, P.; Kjaergaard, M.; Yan, F.; Orlando, T.P.; Gustavsson, S.; Oliver, W.D. A quantum engineer’s guide to superconducting qubits. Appl. Phys. Rev. 2019, 6, 021318. [Google Scholar] [CrossRef]
  26. Bilmes, A.; Zanker, S.; Heimes, A.; Marthaler, M.; Schön, G.; Weiss, G.; Ustinov, A.V.; Lisenfeld, J. Electronic decoherence of two-level systems in a Josephson junction. Phys. Rev. B 2017, 96, 064504. [Google Scholar] [CrossRef]
  27. Rothwarf, A.; Taylor, B.N. Measurement of Recombination Lifetimes in Superconductors. Phys. Rev. Lett. 1967, 19, 27–30. [Google Scholar] [CrossRef]
Figure 1. (a) Quasiparticle-induced transitions with quasiparticles’ energies near the superconducting gap Δ and above Δ + E g e . The blue dotted line indicates that the quasiparticles absorb energies, while the red one indicates that they emit energies. (b) Diagram of the Josephson junction centered between the two electrodes and placed in the 3D cavity. (c) Sketch of the response at ω b a r e versus the high input power. To inject quasiparticles, the injected microwave power needs to be above 0 dB, which is marked by the red line. (d) Pulse sequence of measurement for injecting quasiparticles.
Figure 1. (a) Quasiparticle-induced transitions with quasiparticles’ energies near the superconducting gap Δ and above Δ + E g e . The blue dotted line indicates that the quasiparticles absorb energies, while the red one indicates that they emit energies. (b) Diagram of the Josephson junction centered between the two electrodes and placed in the 3D cavity. (c) Sketch of the response at ω b a r e versus the high input power. To inject quasiparticles, the injected microwave power needs to be above 0 dB, which is marked by the red line. (d) Pulse sequence of measurement for injecting quasiparticles.
Applsci 12 08461 g001
Figure 2. Experimental impacts of injected quasiparticles. (a) State evolution of the energy relaxation time T 1 measured by injecting the power of 7 dB. The green data were measured with a diffusion time of 50 μ s, while the red data were measured when t d i f f = 200 μ s. (b) Rabi oscillations observed with 7 dB of injection power and 50 μ s of diffusion time. The green line represents the reference Rabi oscillation, while the red line represents the signal Rabi oscillation. From the amplitudes of the Rabi oscillation, we can calculate the residual excited-state population [11]. (c) The energy relaxation time T 1 and (d) residual excited-state population P e r were measured as a function of t d i f f . The four curves in each figure represent, respectively, the responses of the qubit induced by quasiparticle injection with injection powers of 4, 7, 10, and 13 dB. The yellow line represents the data without the quasiparticle injection. The data were measured with an injection duration time of 3 μ s.
Figure 2. Experimental impacts of injected quasiparticles. (a) State evolution of the energy relaxation time T 1 measured by injecting the power of 7 dB. The green data were measured with a diffusion time of 50 μ s, while the red data were measured when t d i f f = 200 μ s. (b) Rabi oscillations observed with 7 dB of injection power and 50 μ s of diffusion time. The green line represents the reference Rabi oscillation, while the red line represents the signal Rabi oscillation. From the amplitudes of the Rabi oscillation, we can calculate the residual excited-state population [11]. (c) The energy relaxation time T 1 and (d) residual excited-state population P e r were measured as a function of t d i f f . The four curves in each figure represent, respectively, the responses of the qubit induced by quasiparticle injection with injection powers of 4, 7, 10, and 13 dB. The yellow line represents the data without the quasiparticle injection. The data were measured with an injection duration time of 3 μ s.
Applsci 12 08461 g002
Figure 3. Quasiparticle diffusion in the superconducting qubit. (a) Transition rate as a function of diffusion time. The solid circles on the dotted (solid) line represent the up- and down-transition rates Γ ( Γ ) extracted from T 1 and P e r . The four colored curves point to the four injection powers. These data were measured with an injection duration of 3 μ s. (b) Γ as a function of diffusion time fitted by the modified model with different injection powers. (c) The diffusion time corresponding to the fitted Γ m a x with (without) the existence of “hot” quasiparticles. (d) Γ m a x corresponding to different injection powers when measured with injection durations of 3 and 4.5 μ s. The experimental data and simulated results are represented by circles and dotted lines, respectively.
Figure 3. Quasiparticle diffusion in the superconducting qubit. (a) Transition rate as a function of diffusion time. The solid circles on the dotted (solid) line represent the up- and down-transition rates Γ ( Γ ) extracted from T 1 and P e r . The four colored curves point to the four injection powers. These data were measured with an injection duration of 3 μ s. (b) Γ as a function of diffusion time fitted by the modified model with different injection powers. (c) The diffusion time corresponding to the fitted Γ m a x with (without) the existence of “hot” quasiparticles. (d) Γ m a x corresponding to different injection powers when measured with injection durations of 3 and 4.5 μ s. The experimental data and simulated results are represented by circles and dotted lines, respectively.
Applsci 12 08461 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dong, Y.; Li, Y.; Zheng, W.; Zhang, Y.; Ma, Z.; Tan, X.; Yu, Y. Measurement of Quasiparticle Diffusion in a Superconducting Transmon Qubit. Appl. Sci. 2022, 12, 8461. https://doi.org/10.3390/app12178461

AMA Style

Dong Y, Li Y, Zheng W, Zhang Y, Ma Z, Tan X, Yu Y. Measurement of Quasiparticle Diffusion in a Superconducting Transmon Qubit. Applied Sciences. 2022; 12(17):8461. https://doi.org/10.3390/app12178461

Chicago/Turabian Style

Dong, Yuqian, Yong Li, Wen Zheng, Yu Zhang, Zhuang Ma, Xinsheng Tan, and Yang Yu. 2022. "Measurement of Quasiparticle Diffusion in a Superconducting Transmon Qubit" Applied Sciences 12, no. 17: 8461. https://doi.org/10.3390/app12178461

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop