Phonon downconversion to suppress correlated errors in superconducting qubits
Abstract
Quantum error correction can preserve quantum information in the presence of local errors, but correlated errors are fatal. For superconducting qubits, high-energy particle impacts from background radioactivity produce energetic phonons that travel throughout the substrate and create excitations above the superconducting ground state, known as quasiparticles, which can poison all qubits on the chip. We use normal metal reservoirs on the chip back side to downconvert phonons to low energies where they can no longer poison qubits. We introduce a pump-probe scheme involving controlled injection of pair-breaking phonons into the qubit chips. We examine quasiparticle poisoning on chips with and without back-side metallization and demonstrate a reduction in the flux of pair-breaking phonons by over a factor of 20. We use a Ramsey interferometer scheme to simultaneously monitor quasiparticle parity on three qubits for each chip and observe a two-order of magnitude reduction in correlated poisoning due to background radiation.
Qubits formed from superconducting integrated circuits are one of the leading systems for implementation of a fault-tolerant quantum computer Kjaergaard2020 . For sufficiently high gate fidelity, error correction schemes such as the surface code Fowler2012 can mitigate local errors. However, recent work has shown that high-energy particle impacts from low-level radioactivity and cosmic-ray muons will generate nonequilibrium quasiparticles (QPs) Cardani2021 ; Catelani2011 ; Vepsalainen2020 that can lead to correlated errors across a multiqubit array Wilen2021 ; Martinis2021 ; McEwen2021 . Such correlated errors cannot be mitigated by current error correction schemes, thus posing a significant challenge to realization of a fault-tolerant quantum computer.
Particle impacts deposit energy of order 100 keV in the device substrate, leading to the generation of large numbers of electron-hole pairs and a cascade of high-energy phonons Wilen2021 . These phonons travel throughout the chip and break Cooper pairs with high probability when they scatter off superconducting structures on the device layer, thus generating QPs at arbitrary locations relative to the particle impact site Martinez2019 ; Martinis2021 ; Wilen2021 . Prior work has explored low-gap superconducting structures for phonon downconversion to protect superconducting resonators and detectors with a higher gap energy Henriques2019 ; Karatsu2019 . Another scheme involves placing superconducting detectors on thin suspended membranes deVisser2021 . The use of normal metal layers on the back side of superconducting qubit chips was proposed in Ref. Martinis2021 to downconvert energetic phonons below the superconducting gap. Because this downconversion process is based on the scattering of phonons with conduction electrons in the metal, the low rate of electron-phonon scattering at low temperatures Wellstood1994 dictates large volumes of normal metal for efficient phonon downconversion. A calculation in Ref. Martinis2021 based on the phonon scattering rate in the normal metal on the back side and the pair-breaking rate in the superconducting film on the device layer indicates that achieving a 100-fold improvement in the qubit energy relaxation time in the aftermath of a particle impact requires a 6-m-thick normal metal layer.
Here we implement this idea using thick electroplated Cu reservoirs to promote downconversion of phonons below the superconducting gap edge. To test this approach in a controlled way, we integrate Josephson junctions around the chip perimeter and controllably bias the junctions above the superconducting gap to generate pair-breaking phonons on demand. With explicit phonon injection, we find that the phonon-downcoversion structures reduce the flux of pair-breaking phonons by more than a factor of 20. We also examine the correlated errors in multiqubit chips with and without Cu reservoirs and find a two order of magnitude reduction in correlated error rate, to the point where these errors no longer pose a limit to fault-tolerant operation.

Results
Experimental design. The experimental geometry is shown in Fig. 1. We study two nominally identical chips, one with back-side normal metallization (Cu chip) and one without (non-Cu chip). Each chip incorporates an array of charge-sensitive transmon qubits, with Josephson injector junctions arrayed around the perimeter of the chip. Each qubit has a readout resonator that is inductively coupled to a common feedline that can be used for multiplexed readout. We measure both chips in the same low-temperature environment on the same cooldown. For our qubits, we target a somewhat low ratio of Josephson energy to single-electron charging energy to produce a peak-to-peak charge dispersion between 1-5 MHz. This allows us to monitor QP parity switching for each qubit, which is a sensitive measure of QP poisoning Riste2013 ; Serniak2018 ; Serniak2019 . For the experiments presented here, we focus on three of the qubits on each chip: , , [Fig. 1(b)]. Details of the qubit parameters and experimental configuration are given in Methods and Supplementary Note 3-4.
The normal metal reservoirs on the Cu chip consist of 10- thick islands patterned from Cu films grown by electrodeposition onto the back side of a high-resistivity double-side polished Si wafer following electron-beam evaporation of a Ti/Cu seed layer. The 10- thickness was chosen based on the estimate from Ref. Martinis2021 . The islands are defined with a lattice of partial dicing saw cuts through the Cu film into the back side of the wafer, resulting in island areas of [Fig. 1(c)] (see Methods and Supplementary Note 1-2). This is done to suppress damping from coupling to the transmission line mode formed by a continuous metal layer on the back side of the chip and the ground plane on the device layer that would otherwise degrade qubit coherence Martinis2021 . Metallic losses due to capacitive coupling between the qubit and Cu island are projected to limit the qubit quality factor to M for the qubit design considered in Ref. Martinis2021 ; the smaller qubit island size for our qubits reduces this coupling and raises this quality factor limit, thus making a negligible impact on .
The injector junctions are fabricated at the same time as the qubit junctions with a standard Al-AlOx-Al process. The ground plane, qubit capacitor islands, readout resonators, and injector junction pads are all fabricated from Nb. There is no direct galvanic connection from the injector junction pads to ground, so the QP poisoning proceeds via phonon emission. By biasing the injector junction above , where is the superconducting energy gap for Al, we break Cooper pairs and generate local QPs that subsequently recombine, emitting phonons. These phonons then travel through the Si, and, upon encountering Al junction electrodes for a qubit, a phonon will break a Cooper pair with high probability and generate two QPs. A similar injection scheme was used in Ref. Patel2017 ; Mannila2021 . Although the phonons injected by the tunnel junction will be lower in energy than those generated by a particle impact, the tunnel junction gives us the ability to control the timing, duration, and location of the phonon injection, in contrast to phonons from particle impacts, which occur at random times and locations.
Enhanced relaxation from phonon injection. In a first series of experiments, we measure the energy relaxation time of all three qubits on each chip following pulsed QP injection. Here we focus on () for the non-Cu (Cu) chip, but we observe similar behavior for the other two qubits on each chip. We use a standard inversion recovery measurement to probe . To quantify degradation in , we plot , where is the baseline relaxation time from an average of several measurements with no injector junction bias (see Methods). The change in reduced QP density in the qubit junction leads can be calculated from Wang2014 (see Methods).
We start by applying a 10- injection pulse with amplitude = 1 mV, well beyond , so that we expect significant QP poisoning in the absence of any mitigation. We vary the delay time between the end of the injection pulse and the pulse at the start of the sequence. For the non-Cu chip, increases substantially following the injection pulse and reaches a maximum poisoning level about after the end of the injection pulse [Fig. 2(a)]. This delayed onset of poisoning is consistent with the propagation timescale for the injected phonons to diffuse through the substrate to the qubit junction. In the absence of phonon downconversion structures, phonons travel throughout the substrate following boundary-limited diffusion, where they scatter randomly off the top and bottom surfaces of the Si chip.

This leads to an effective diffusivity , where is the speed of sound in Si and is the chip thickness. Thus, the timescale for phonons to diffuse from the injector junction to each qubit (at a separation of 4-6 mm from the injector) is of the order of . Following this peak, recovers towards the unpoisoned baseline level following an exponential decay with a characteristic time of 60 (see Supplementary Note 5). This corresponds to the timescale for phonons to exit the substrate at the chip perimeter where the sample is acoustically anchored to the device enclosure. The chip is attached to the machined Al enclosure using a small amount of low-temperature adhesive (GE varnish) at the corners of the chip.
For the Cu chip, is difficult to distinguish from the baseline level for all delays. For a 30-s delay, for the non-Cu chip is over a factor of 35 larger than for the Cu chip. This is our first key result demonstrating the effectiveness of Cu reservoirs in reducing phonon-mediated QP poisoning.
We then explore the variation of with for a fixed delay of and a 10-s pulse width. For the non-Cu chip, we observe a significant increase in when the pulse amplitude exceeds 0.4 mV. For the Cu chip, doesn’t change significantly at ; however, we observe a gradual rise in to a peak at a pulse amplitude of 0.56 mV, followed by a reduction to the baseline level for larger pulse amplitudes. We understand the peak in , which is also visible for the non-Cu chip on top of the overall poisoning curve, to be due to photon-assisted poisoning from absorption of Josephson radiation emitted by the injector junction mediated by a spurious mm-wave resonance in the qubit (see Supplementary Note 5-6). Such antenna effects in qubit structures can lead to resonant absorption of electromagnetic radiation, which can drive high-frequency currents through the qubit junction and generate QPs Houzet2019 ; Rafferty2021 ; liu2022 . We would not expect the Cu reservoirs to have any effect on this photon-based QP poisoning mechanism.

QP parity switching. In a separate series of experiments, we exploit the non-negligible charge dispersion of our qubits to probe the charge parity of the qubit islands as a sensitive probe of QP poisoning. We employ a Ramsey pulse sequence to map QP parity onto qubit 1-state occupation Riste2013 ; Serniak2018 ; Christensen2019 (see Methods). We first perform the QP parity switching measurement on each of the qubits at a repetition period of 10 ms and then compute the power spectral density of parity switches. We fit a Lorentzian to the measured spectrum for each chip to extract the characteristic parity switching rate , as in Ref. Riste2013 . In Fig. 3(a), we plot typical spectra from one qubit on each chip. The resulting values for for both chips are low: = 0.360 s-1 (0.023 s-1) for the non-Cu (Cu) chip. To the best of our knowledge, for the non-Cu chip is consistent with the lowest rates for QP poisoning reported in the literature Pan2022 ; Kurter2021 , while for the Cu chip our measured poisoning rate is an order of magnitude lower.
The low QP poisoning rates on both chips are likely due to a combination of best practices for shielding, filtering, and thermalization (see Supplementary Note 3). In addition, the relatively compact qubit design results in a rather high fundamental antenna resonance frequency, 270 GHz (see Supplementary Note 6), which is likely above the cutoff of the spectrum of blackbody radiation from higher temperature stages of the cryostat. Following the analysis in Ref. liu2022 applied to the geometry of our qubits, we calculate an effective blackbody temperature of 330 mK (280 mK) for on the non-Cu (Cu) chip from our measured values. In addition to photon absorption by the spurious antenna resonance, the residual QP poisoning is likely due to high-energy particle impacts or other radiation sources that generate pair-breaking phonons. We attribute the even lower QP parity switching rate for the Cu chip to absorption of a significant fraction of the phonons generated from these poisoning events by the Cu islands on the back side of the chip, which we will subsequently quantify.
The low QP baseline poisoning rates allow us to directly investigate QP poisoning in the presence of controlled injection of pair-breaking phonons into the chip. Here, we sample QP parity on the qubits with a 100- repetition period, while pulsing the injector junction at an amplitude of 1 mV and a fixed rate of 20 Hz. Since this experimental duty cycle is much faster than our background switching rate, we can apply a moving average over 100 time steps for both the non-Cu and Cu chip to improve the signal-to-noise ratio for these time traces. We then perform a hidden Markov model analysis to identify the parity switches (see Methods).
As we vary the pulse length from 20 ns to , we increase the injected energy and thus the number of pair-breaking phonons coupled to the chip. Longer injection pulses result in a higher rate of parity switches, with almost all of the parity switches synced with the phonon injection pulses [Fig. 3(c,d)]. In Fig. 3(b) we plot the ratio of the measured switching rate to the rate of phonon injection as a function of the injection pulse duration; this quantity corresponds to the probability of a measured parity switch per phonon injection pulse. For sufficiently high injected energy, we expect that pair-breaking phonons will randomize the parity on every qubit island on a timescale much shorter than our 100- sampling period. Because we observe a change of QP parity only for an odd number of switches, we expect our measured parity switching rate to saturate at half the injection rate for long injection pulses. As expected, our measured probabilities saturate around 0.5; however, the Cu chip requires roughly 20 times the injection energy to achieve the same level of poisoning as the non-Cu chip. If we assume that each injection pulse generates a number of pair-breaking phonons in the Si that is proportional to the pulse duration, this indicates that the Cu islands on the back side of the Cu chip downconvert 95% of the injected phonons.
Multi-qubit correlated parity switching. We next perform simultaneous parity measurement of all three qubits on each chip and we analyze the resulting time series to identify coincidences (see Supplementary Note 9-10). We first apply this approach to the measurements with periodic QP poisoning from controlled phonon injection. For sufficiently long injection pulses, we expect the probability of double coincidences to saturate at and for triple coincidences to saturate at ; this is indeed what we observe (see Supplementary Note 8).

After confirming that our analysis successfully detects coincidences induced by controlled injection, we next apply this same approach to measurement of correlated poisoning induced by environmental radiation. Since the background poisoning rates are quite low for these devices, we reduce the experimental duty cycle to 10 ms and acquire simultaneous parity data over several days to build up sufficient statistics to detect coincidences. Figure 4 presents our results from these measurements. For all three qubits on each chip, the single-qubit parity switching rate is consistent with our previously measured values. Based on the observed parity switching rates, we calculate the expected random double- and triple-coincidence rates (see Methods). For the non-Cu chip, the expected random rates for double (triple) coincidences are less than the observed coincidence rates by nearly a factor of 2 (3), indicating the presence of significant correlated switching.
Based on the analysis in Refs. Wilen2021 ; Martinis2021 , in the absence of mitigation, we expect correlated poisoning events to be dominated by impacts that broadcast high-energy phonons throughout the entire chip. For a given impact rate , we thus expect a rate of individual qubit parity switches , a rate of two-fold coincidences , and a rate of three-fold coincidences . We solve a system of equations for the observed coincidence probabilities to obtain the actual exclusive rates for single-, double-, and triple-qubit poisoning events (see Supplementary Note 11). If there were no other poisoning mechanisms and no phonon loss in the chip so that each impact poisoned all qubits with 100% probability, we would expect an extracted event rate for equal to , while all single- and two-fold poisoning rates would be zero, since all particle impacts are expected to couple to all qubits via high-energy phonons.
Figure 4(b) presents the observed parity rates and extracted poisoning event rates for both chips. For the non-Cu chip, the extracted three-qubit event rate is high [0.064(9) s-1], indicating the presence of significant correlated poisoning between widely separated qubits. However, the event rates for double- and single-qubit poisoning are also significant; we note that for the three qubit pairs with different physical separations, there is no clear dependence of two-fold correlated poisoning on the distance between qubits. For any practical implementation, there will always be some degree of phonon loss, for example, from the anchoring points where the chip is attached to the sample enclosure or through wirebonds, so that even in the absence of phonon downconversion structures, not all qubits are poisoned by each impact. In this case, could be estimated as 1.1 s-1, the sum of all the poisoning rates for the non-Cu chip in Fig. 4(b). For the Cu chip, all of the extracted correlated event rates are two orders of magnitude lower than for the non-Cu chip; the sum of all poisoning rates on the Cu chip is 0.15 s-1, which is dominated by the single-qubit poisoning rates. This indicates that the Cu reservoirs greatly reduce the footprint of the phonon burst following a high-energy particle impact.
Discussion
We have separately performed repeated charge tomography for one qubit on each chip and observed a rate of large offset charge jumps of 0.0012(1) [0.0011(1)] s-1 for the non-Cu [Cu] chip. Following the detailed modeling and analysis in Ref. Wilen2021 , we estimate the rate of impacts on our chips to be 0.083(8) s-1 (see Supplementary Note 13). Thus, the higher total poisoning rate, particularly on the non-Cu chip, compared to the estimated from the offset-charge measurements suggests the presence of additional phonon-mediated poisoning mechanisms in our device. THz photons above from warmer portions of the cryostat could break pairs in the Nb ground plane and couple phonons into the substrate from recombination, thus poisoning nearby qubits, but without the chip-wide burst of phonons from a high-energy impact. Additionally, the cryogenic dark matter detection community has observed heat-only events that are attributed to mechanical cracking processes in the device and sample enclosure that release stresses, typically at the attachment points Aastrom2006 ; Armengaud2016 ; recent work reported heat-only events in superconducting transition edge sensors on a Si chip attached to a sample holder with GE varnish Anthony2022 . Such events can produce large bursts of phonons that are detected by the sensors in these experiments, but with no accompanying charge signal. The dynamics of such heat-only events will depend on the details of the device and enclosure design, but could potentially occur in our qubit chip and sample enclosure and serve as another phonon-mediated QP poisoning mechanism. The overall reduced poisoning rates on the Cu chip indicate that the normal metal structures reduce phonon-mediated poisoning from other mechanisms in our system, such as THz photons or heat-only events, as well.
Excess QPs cause both enhanced parity switching [Fig. 3(b)] and reduced (Fig. 2), thus resulting in enhanced bit-flip errors McEwen2021 . Thus, our demonstrated suppression of correlated QP poisoning from phonon downconversion provides a strategy for reducing correlated errors in large qubit arrays. For robust error detection, we require single-qubit errors below the 10-4 level, which will correspond to random error coincidences between pairs of qubits at the level. Thus, any correlated two-qubit errors must be below the 10-8 level Wilen2021 . If we assume a surface code duty cycle of 1 MHz and take our largest extracted two-fold poisoning event rate of 0.002 s-1, we find a two-fold error probability of . Thus, our initial attempt at correlated error suppression by phonon downconversion already yields a correlated error rate below the threshold necessary for fault-tolerant operation. Further optimization, including an investigation of the dependence of the downconversion efficiency on metal film thickness and composition, and incorporation of additional mitigation strategies should guarantee the robust operation of error-corrected quantum processors in the presence of low-level pair-breaking radiation.
Methods
Device fabrication. Both the non-Cu and Cu chips are fabricated from high-resistivity (10 k-cm) Si wafers. For the Cu chip, the wafer is double-side polished to allow for fabrication of the Cu reservoirs. Deep-UV photolithography is used to pattern the ground plane, feedline, readout resonators, qubit islands, charge-bias lines, and injector junction pads, followed by reactive ion etching of the Nb film. After the base-layer processing, the Cu reservoir fabrication on the wafer with the Cu chip is started by first preparing a protective resist layer on the device surface, then evaporating a seed layer of Ti and Cu on the opposite side. We electroplate a 10- thick film of Cu on top of the seed layer; we pattern the Cu reservoirs by dicing islands with partial dicing saw cuts that extend into the back surface of the Si (see Supplementary Note 1). The qubit and injector junctions on both chips are Al-AlOx-Al junctions made by double-angle evaporation, producing qubit frequencies in the range of 4.7-5.3 GHz (Supplementary Table 1).
Measurement setup. Measurements on both the non-Cu and Cu chips are performed on the same dilution refrigerator cooldown running at a temperature below 15 mK. The Al sample boxes for both chips are mounted on the same cold-finger inside a single Cryoperm magnetic shield. A Radiall relay switch on the output lines of the two devices allows us to switch between measurements of one chip or the other. Supplementary Fig. 2 details the configuration of cabling, attenuation, filtering, and shielding inside the cryostat, as well as the room-temperature electronics hardware for control and readout. The inner surfaces of the Cryoperm magnetic shield and the mixing chamber shield were both coated with an infrared-absorbent layer Barends2011 . For the charge-biasing of the qubits, wiring limitations on our dilution refrigerator prevented us from connecting to all of the bias traces on the chips. For the non-Cu chip, charge-bias lines are connected to and ; for the Cu chip, there is only a bias connection to .
Relaxation and injection measurements. Phonon injection experiments are done by pulsing the bias on one of the Josephson junctions near the edge of each chip [Fig. 1 and Supplementary Fig. 1(c)] followed by a measurement of qubit , from which we compute (see Supplementary Note 5). In addition to analyzing the response of with bias-pulse amplitude and delay between the pulse and measurement, we compute the change in reduced QP density, Wang2014 , where is the qubit transition frequency.
Single-qubit parity measurements. The Ramsey pulse sequence that we use for mapping QP parity onto qubit 1-state occupation is as follows: apply a pulse, idle for a time corresponding to a quarter of a qubit precession period, then apply a pulse, followed by a qubit measurement Riste2013 ; Serniak2018 ; Christensen2019 . If the offset charge is at maximal charge dispersion, the final pulse projects the qubit to either the 0 or 1 state, dependent on the QP parity. In order to have an uninterrupted measurement sequence, active stabilization of the offset charge is not performed. The power spectral densities of the QP parity switching are computed from records of 20,000 single shots of the parity-mapping pulse sequence measured at a repetition period of 10 ms. We apply a simple thresholding scheme based on the 0/1 readout calibration for each qubit to generate a digital time trace of the QP parity. We then compute the PSD of this digital signal and average 20-160 of these curves to obtain Figure 3(a) for on each chip and Supplementary Fig. 7 for all three qubits on both chips. Since the offset charge is not actively stabilized, when the offset charge randomly jumps near the degeneracy point, the parity readout fidelity vanishes. This results in an enhancement of the white noise floor, but the characteristic QP parity switching rate can still be extracted (see Supplementary Note 7).
In addition to the PSD measurements, we also study single-qubit QP parity switching with periodic phonon injection [Fig. 3(b-d)]. Here, we simultaneously produce phonons by pulsing the injector junction to an amplitude of 1 mV at a frequency of 20 Hz while recording single shots of the QP parity-mapping pulse sequence at a repetition period of 100 s for a duration of 400 s. As with the PSD measurements, we do not actively stabilize offset charge, and thus the offset charge will occasionally jump randomly to near degeneracy where the QP parity cannot be discriminated. In order to process the data, we apply a moving average of 100 time steps to the QP parity traces. The portions of the averaged parity traces where the peak-to-peak amplitude is below a threshold determined by the 0/1 readout calibration levels are masked off and not analyzed further. Next, a hidden Markov model (HMM) is used to identify the QP parity. We assign a probability for the parity signal to have an odd- or even-parity state based on Gaussian fits to the 0/1 readout calibration measurements for each qubit. The probability for the states to transition is set by the from the corresponding PSD for each qubit. We then use the Viterbi algorithm to fit a digital signal to the averaged QP parity data (see Supplementary Note 9).
Multi-qubit parity measurements. For measurements of multi-qubit QP parity switching due to background radioactivity, we perform the QP parity-mapping pulse sequence for all three qubits on a chip simultaneously at a repetition period of 10 ms. We use the previously described HMM to identify QP parity switching from the time trace for each qubit using a moving average of 40 points. This results in the averaged QP parity switching events having a sloped step, with the width of each parity switch approximately equal to the number of points used in the moving average. Following the HMM extraction of digital parity switching traces, we identify a coincidence switching event between qubits to occur when the digital time traces switch within a window of 40 data points (see Supplementary Fig. 10). The coincident events are indexed with the relevant qubits involved in the switching event: , , , or . We restrict each switch of a given qubit to participate in only one event per coincidence type. For example, a switch cannot be used for two coincidences, but could be used for a coincidence and a coincidence. The switching rate for each type of coincidence event , where , is given by , where is the total number of events and is the total duration of unmasked data for event type . Note that double coincidences between qubits and are only counted during the period when both qubits are unmasked; similarly, triple coincidences require that all three qubits are unmasked. The uncertainty in comes from the standard Poisson counting errors (see Supplementary Note 10).
Extraction of correlated poisoning rates. For a set of observed single-qubit parity switching rates with a particular non-zero window for identifying coincidences, one would expect a rate of random uncorrelated coincidence switching given by the product of the probabilities for observing a parity switch for each constituent qubit during the interval . These expected background coincidence parity switching rates are listed in Supplementary Table 2. The error bars for these random background coincidence rates were computed by summing the fractional uncertainty for each observed rate in quadrature. While the quantities we measure in our simultaneous parity measurements are the observed parity switching rates, we would like to compute the actual poisoning event rates for each qubit, or group of qubits, exclusively. For example, a single poisoning event that couples to both and will contribute to but will not contribute to or . Based on these criteria, we use the observed parity switching rates to compute the probability for observing each type of parity switching event in a window interval as . We then derive expressions for the probability of observing each type of parity switching event in terms of the actual probability for each type of poisoning event, as listed in Supplementary Eq. (2). We numerically solve the system of equations to obtain the actual poisoning probabilities and then calculate the extracted poisoning rates reported in Fig. 4(b) and Supplementary Table 2. The error bars on each actual poisoning probability are calculated by numerically computing the derivative with respect to each of the observed switching probabilities, then multiplying by the corresponding Poisson error bar for the observed switching probability and summing these together in quadrature (see Supplementary Note 11).
Data availability
Data used in this work is available on 10.5281/zenodo.7114229. Supplementary data is available upon reasonable request.
Code availability
Code used in this work is available upon reasonable request.
References
References
- (1) Kjaergaard, M. et al. Superconducting qubits: Current state of play. Annual Review of Condensed Matter Physics 11, 369–395 (2020).
- (2) Fowler, A. G., Mariantoni, M., Martinis, J. M. & Cleland, A. N. Surface codes: Towards practical large-scale quantum computation. Physical Review A 86, 032324 (2012).
- (3) Cardani, L. et al. Reducing the impact of radioactivity on quantum circuits in a deep-underground facility. Nature Communications 12, 1–6 (2021).
- (4) Catelani, G., Schoelkopf, R. J., Devoret, M. H. & Glazman, L. I. Relaxation and frequency shifts induced by quasiparticles in superconducting qubits. Physical Review B 84, 064517 (2011).
- (5) Vepsäläinen, A. P. et al. Impact of ionizing radiation on superconducting qubit coherence. Nature 584, 551–556 (2020).
- (6) Wilen, C. et al. Correlated charge noise and relaxation errors in superconducting qubits. Nature 594, 369–373 (2021).
- (7) Martinis, J. M. Saving superconducting quantum processors from decay and correlated errors generated by gamma and cosmic rays. npj Quantum Information 7, 1–9 (2021).
- (8) McEwen, M. et al. Resolving catastrophic error bursts from cosmic rays in large arrays of superconducting qubits. Nature Physics 1–5 (2021).
- (9) Martinez, M. et al. Measurements and simulations of athermal phonon transmission from silicon absorbers to aluminum sensors. Physical Review Applied 11, 064025 (2019).
- (10) Henriques, F. et al. Phonon traps reduce the quasiparticle density in superconducting circuits. Applied Physics Letters 115, 212601 (2019).
- (11) Karatsu, K. et al. Mitigation of cosmic ray effect on microwave kinetic inductance detector arrays. Applied Physics Letters 114, 032601 (2019).
- (12) De Visser, P. J., De Rooij, S. A., Murugesan, V., Thoen, D. J. & Baselmans, J. J. Phonon-trapping-enhanced energy resolution in superconducting single-photon detectors. Physical Review Applied 16, 034051 (2021).
- (13) Wellstood, F., Urbina, C. & Clarke, J. Hot-electron effects in metals. Physical Review B 49, 5942 (1994).
- (14) Ristè, D. et al. Millisecond charge-parity fluctuations and induced decoherence in a superconducting transmon qubit. Nature Communications 4, 1–6 (2013).
- (15) Serniak, K. et al. Hot nonequilibrium quasiparticles in transmon qubits. Physical Review Letters 121, 157701 (2018).
- (16) Serniak, K. et al. Direct dispersive monitoring of charge parity in offset-charge-sensitive transmons. Physical Review Applied 12, 014052 (2019).
- (17) Patel, U., Pechenezhskiy, I. V., Plourde, B. L. T., Vavilov, M. & McDermott, R. Phonon-mediated quasiparticle poisoning of superconducting microwave resonators. Physical Review B 96, 220501 (2017).
- (18) Mannila, E., Maisi, V. & Pekola, J. Self-calibrating superconducting pair-breaking detector. Physical Review Letters 127, 147001 (2021).
- (19) Wang, C. et al. Measurement and control of quasiparticle dynamics in a superconducting qubit. Nature Communications 5, 1–7 (2014).
- (20) Houzet, M., Serniak, K., Catelani, G., Devoret, M. & Glazman, L. Photon-assisted charge-parity jumps in a superconducting qubit. Physical Review Letters 123, 107704 (2019).
- (21) Rafferty, O. et al. Spurious antenna modes of the transmon qubit. arXiv preprint arXiv:2103.06803 (2021).
- (22) Liu, C.-H. et al. Quasiparticle poisoning of superconducting qubits from resonant absorption of pair-breaking photons. arXiv preprint arXiv:2203.06577 (2022).
- (23) Christensen, B. et al. Anomalous charge noise in superconducting qubits. Physical Review B 100, 140503 (2019).
- (24) Pan, X. et al. Engineering superconducting qubits to reduce quasiparticles and charge noise. arXiv preprint arXiv:2202.01435 (2022).
- (25) Kurter, C. et al. Quasiparticle tunneling as a probe of Josephson junction quality and capacitor material in superconducting qubits. arXiv:2106.11488 (2021).
- (26) Åström, J. et al. Fracture processes observed with a cryogenic detector. Physics Letters A 356, 262–266 (2006).
- (27) Armengaud, E. et al. Constraints on low-mass WIMPs from the EDELWEISS-III dark matter search. Journal of Cosmology and Astroparticle Physics 2016, 019 (2016).
- (28) Anthony-Petersen, R. et al. A stress induced source of phonon bursts and quasiparticle poisoning. arXiv preprint arXiv:2208.02790 (2022).
- (29) Barends, R. et al. Minimizing quasiparticle generation from stray infrared light in superconducting quantum circuits. Applied Physics Letters 99, 113507 (2011).
Acknowledgements
This work is supported by the U.S. Government under ARO grant W911NF-18-1-0106. Fabrication was performed in part at the Cornell NanoScale Facility, a member of the National Nanotechnology Coordinated Infrastructure (NNCI), which is supported by the National Science Foundation (Grant NNCI-2025233).
Author contributions
V.I. and J.K. took and analysed the data. V.I. and A.B. designed and fabricated the devices. J.K., C.P.L., and E.Y. developed code for analysing parity switching data. C.H.L., S.P., and R.M. performed the CST Microwave simulations. J.K., V.I., and A.B. helped to develop the measurement and fabrication infrastructure. B.L.T.P. designed the experiment and directed data-taking and analysis. V.I., J.K., C.P.L., E.Y., R.M., and B.L.T.P. co-wrote the manuscript.
Competing interests
Authors B.L.T.P and R.M. declare a competing interest in the form of a utility patent application, FABRICATION OF NORMAL CONDUCTING OR LOW-GAP ISLANDS FOR DOWNCONVERSION OF PAIR-BREAKING PHONONS IN SUPERCONDUCTING QUANTUM CIRCUITS (application number 17/469,380), filed by the Wisconsin Alumni Research Foundation on September 8, 2021. B.L.T.P and R.M. are the sole inventors on the application, which is currently pending. The invention covered by the patent application concerns the technique for fabricating the normal metal structures for phonon downconversion. This article describes experimental work by our two research groups demonstrating the effectiveness of this technique for phonon downconversion. The remaining authors declare no competing interests.
Additional Information
Correspondence and requests for materials should be addressed to B.L.T.Plourde.
Supplementary Information: Phonon downconversion to suppress correlated errors in superconducting qubits
Supplementary Note 1: Device Fabrication
Both the non-Cu and Cu chips are fabricated on high-resistivity (-cm) 100-mm Si wafers; for the Cu chip, the wafer is double-side polished to facilitate the deposition and patterning of the Cu reservoirs. Initially the wafer is put through a standard RCA clean process and then submerged in a buffered-2 per vol. HF bath to remove native oxides immediately before sputter-deposition for the non-Cu (Cu) chip of a 55-nm (80-nm) thick Nb film on the top surface of each wafer. We pattern the Nb films using a deep-UV photostepper to define the ground plane, feedline, readout resonators, qubit islands, charge-bias lines, and injector junction pads followed by a dry etch using BCl3, Cl2, and Ar in an ICP etcher. In the case of the non-Cu devices, we proceed with the wafer to the Josephson junction definition step, while for the Cu devices, we next fabricate the Cu reservoirs.
After stripping the base layer resist with a TMAH hot strip bath, we then coat the surface with the Nb pattern using a thick photoresist layer (SPR-220-3.0) to protect the Nb during the subsequent backside processing for the Cu reservoirs. We then deposit a metal seed layer on the back side of the wafer using electron-beam evaporation of Ti (20 nm) at a deposition rate of 1Å/s followed by Cu (100 nm) deposited at 2Å/s.
For the Cu reservoirs, we deposit Cu on the wafer back side with an electrodeposition process by submerging our wafer into a copper sulfate and sulfuric acid solution. We grow a 10- thick Cu film on top of the seed layers at a rate of 3.3 using an alternating current deposition mode. A test film grown with the same parameters and patterned into a narrow strip using Kapton tape was measured to have RRR 42. The islands were defined with a lattice of partial 50--wide dicing saw cuts through the Cu film into the back side of the wafer, with the cuts extending into the back surface of the Si, resulting in island areas of . After the Cu islands are fabricated, all resist is stripped in a TMAH hot strip bath.
For both wafers, the Josephson junctions are then defined with a conventional double-angle shadow-evaporation process using 100 keV electron-beam lithography of a PMMA/MMA bilayer resist stack. After an in situ ion mill cleaning step to remove native oxide from the Nb surface at the contact points to the junction electrodes, the junctions are formed with electron-beam evaporation of Al. The bottom (top) junction electrode is 40 (80) nm thick.

Supplementary Note 2: Device Layout
Following the fabrication, the wafers are diced into chips that are (8 mm)2. The coplanar waveguide feedline runs across the middle of the wafer, with the 1/4-wave readout resonators for each qubit inductively coupled to the feedline. A full-chip layout can be seen in Supplementary Fig. 1, along with close-up views of each qubit, the non-Cu injector junction, and the Cu island pattern on the back side of the Cu chip. Editing of the individual micrographs to obtain the full-chip image is described in Supplementary Note 7. The locations of the qubits measured in the experiment relative to the injector junctions used for controlled QP poisoning, as well as the inter-qubit separations, are indicated in Supplementary Fig. 1(g-i).

Supplementary Note 3: Device and Measurement Setup
Measurements on both the non-Cu and Cu chips are performed on the same dilution refrigerator cooldown running at a temperature below 15 mK. The Al sample boxes for both chips are mounted on the same cold-finger inside a single Cryoperm magnetic shield. A Radiall relay switch on the output lines of the two devices allows us to switch between measurements of one chip or the other. Supplementary Fig. 2 details the configuration of cabling, attenuation, filtering, and shielding inside the cryostat, as well as the room-temperature electronics hardware for control and readout. The inner surfaces of the Cryoperm magnetic shield and the mixing chamber shield were both coated with an infrared-absorbent layer sBarends2011 . For the charge biasing of the qubits, wiring limitations on our dilution refrigerator prevented us from connecting to all of the bias traces on the chips. For the non-Cu chip, charge bias lines are connected to and ; for the Cu chip, there is only a bias connection to .
Supplementary Note 4: Device Parameters
Supplementary Table 1 lists relevant qubit parameters for both chips, including the qubit transition frequency , the readout resonator frequency , the peak-to-peak maximum charge dispersion , the mean and standard deviation from repeated baseline measurements, and the ratios. During the junction fabrication, the same double-angle evaporation process is used for the injector and qubit junctions, and thus all junctions on a chip have nominally the same critical current density. For each device, one of the junctions around the perimeter of the chip is connected to a 50- bias lead to use as the injector junction (indicated by color highlighting in Supplementary Fig. 1); the injector junction for each chip is 3 times the area of the qubit junctions. Because the junctions on the Cu and non-Cu chips were processed separately, the critical current densities on the two chips are slightly different. For the injector junctions, for the non-Cu (Cu) chips. The qubit junctions were all designed to have the same area and, based on witness junctions on the same chip written with the same area, had normal resistance of for the non-Cu (Cu) chips.
Qubit Parameters | |||||||
Device | Qubit | ||||||
non-Cu | 4.6555 | 6.0431 | 34 | 10 | 3.743 | 24 | |
4.7363 | 6.1506 | 20 | 2 | 3.201 | 26 | ||
4.8408 | 6.229 | 16 | 2 | 4.631 | 25 | ||
Cu | 4.9959 | 6.3977 | 16 | 3 | 1.878 | 29 | |
5.2536 | 6.4868 | 21 | 5 | 1.146 | 32 | ||
5.3190 | 6.5963 | 13 | 4 | 1.938 | 31 |
Supplementary Note 5: Details of measurements
For measurements of enhancements to the qubit relaxation rate following pulsing of the injector junction, in the main paper we present measurements of for on the non-Cu chip and for the Cu chip. In this section, we compile these measurements for the other qubits, and show that the response of the other qubits on each chip is consistent with the representative measurements in the main paper.
Supplementary Fig. 3(a) contains measurements of vs. the delay between the 10-s injection pulse and the pulse for the relaxation measurement for all three qubits on both the non-Cu and Cu chips. In Supplementary Fig. 3(b), we plot the same data for the three qubits on the non-Cu chip on a semilog scale. The black dashed line corresponds to a characteristic timescale of 60 s for injected phonons to leave the chip following the phonon arrival peak. Error bars on values, here and in Fig. 2 in the main paper, are calculated from fit errors with 95 confidence intervals from fits with contributions added in quadrature.
Supplementary Fig. 4 contains plots of vs. for a delay of 30 s for all three qubits on both chips. In Supplementary Fig. 5, we plot the same type of measurements but with a delay of 100 ns. In this second case, the antenna-resonance peaks from the photonic coupling to the Josephson radiation emitted by the injector junction are enhanced, while the remaining phononic poisoning is somewhat lower, as not all of the injected phonons have reached the qubit yet. The change in reduced QP density in the qubit junction leads, , that is plotted on the right axes can be calculated from as sWang2014 , where is the qubit transition frequency.
When the injection pulse amplitude is below , we observe only minimal reduction in ; there is still some non-zero, but small, poisoning in this regime because our junction biasing scheme still permits the injector junction to undergo relaxation oscillations for small bias voltages sVernon1968 , where the junction can momentarily switch out to the gap before retrapping.




Supplementary Note 6: Antenna-mode simulations
As described in the main paper, voltage bias of the injector junction will also induce poisoning from the emission of Josephson radiation. For a pulse amplitude , the Josephson radiation will have frequency , where is the magnetic flux quantum; is Planck’s constant and is the electron charge. Such electromagnetic radiation can be resonantly absorbed by qubit structures acting as antennas, with typical resonant frequencies in the hundreds of GHz range. The absorbed radiation can then drive high-frequency currents through the qubit junction and generate QPs, as described recently in Ref. sRafferty2021 ; sliu2022 . A related photon-based QP poisoning mechanism was considered in Ref. sHouzet2019 .
In order to model the spurious qubit antenna resonances on our devices, we follow the analysis in Ref. sRafferty2021 ; sliu2022 and compute the radiation impedances of the injector junction and the qubit structure with a finite-element simulation using CST Microwave sCST .
With the critical current values for the injector and qubit junctions extracted from the on-chip witness junction measurements, we calculate the product of the coupling efficiencies to free space for the injector junction and the qubit junction . In Supplementary Fig. 6, we plot this product as a function of the injector junction pulse amplitude for both on the Cu chip and on the non-Cu chip. The fundamental peaks in the simulation for both qubits match the measured antenna resonances from in Fig. 2(b) in the main paper and Supplementary Fig. 5 in the supplement.
Supplementary Note 7: Parity switching power spectra for all qubits

We implement a Ramsey pulse sequence that has been used previously to map QP parity onto qubit 1-state occupation sRiste2013 ; sSerniak2018 ; sChristensen2019 ). We apply an pulse, idle for a time corresponding to a quarter of a qubit precession period, then apply a pulse, followed by a qubit measurement. If the offset charge corresponds to the point of maximum charge dispersion, the final pulse will rotate the state vector to the north/south poles of the Bloch sphere dependent on the QP parity state. Although some of the qubits on each chip have connections to the charge-bias line, we have chosen to perform our QP parity switching measurements without active stabilization of the offset charge. This allows the QP parity measurements to proceed without interruptions from periodic charge-tomography sequences sChristensen2019 . However, when the offset charge jumps to near (for integer ), where the bands cross, the fidelity of the QP parity-mapping sequence approaches zero.
In order to compute the power spectral density of the QP parity switching, we perform the QP parity switching measurement on each qubit with 20,000 single shots at a repetition period of 10 ms (although PSD measurements for the Cu chip on the second cooldown used a 25-ms repetition period). For each single-shot measurement stream, we apply a simple thresholding based on the 0/1 readout calibration levels for each respective qubit to produce a digital time trace of QP parity. We then compute the PSD from the resulting digital trace and average several such PSD traces together (between 20-160) to obtain the curves in Fig. 3(a) in the main paper and Supplementary Fig. 7. Because we are not actively stabilizing the offset charge at the point of maximum dispersion, some of the PSD traces that are being averaged will have the environmental offset charge near the degeneracy point, where the QP parity readout fidelity vanishes. This results in an enhancement of the white noise floor, but still allows for a clear extraction of the characteristic QP parity switching rate.
We are able to fit the resulting power spectra with a single Lorentzian using the form described in Ref. sRiste2013 :
(1) |
where is the parity switching rate, is the parity sequence mapping fidelity, and is the parity measurement repetition period.
Supplementary Fig. 7 shows the PSD for all three qubits on each chip. During our experiment, after collecting a majority of our data once the dilution refrigerator had been cold for several months, an unplanned power outage caused our dilution refrigerator to warm up to room temperature. Upon immediately cooling the same two devices back down, without making any changes to the wiring or shielding, we remeasured the PSD for each qubit within a few weeks of the start of this second cooldown. The plots in Supplementary Fig. 7 contain the PSD for each qubit on both chips measured on the first and second cooldowns. For all qubits, the QP parity switching rates increase on the second cooldown, likely because some elements in the qubit environment, for example, the isolators, attenuators, or shields, have not yet fully cooled to the base temperature (see Supplementary Note 12 for further discussion). Nonetheless, the values on the Cu chip remain at least one order of magnitude lower compared to the non-Cu chip.
Supplementary Note 8: QP Parity switching with pulsed injection for all qubits

To complement Fig. 3(b) in the main paper, in this section, we plot the measured switching probability for each qubit on the non-Cu chip [Supplementary Fig. 8(a)] and Cu chip [Supplementary Fig. 8(b)] as a function of the injector junction pulse length. For each chip, all three qubits exhibit a similar behavior. Supplementary Fig. 8(c,d) contain the double- and triple-coincidence switching probabilities for the non-Cu and Cu chips, along with comparisons to the square and cube of the single-qubit switching probabilities for one of the qubits on each chip, as discussed in the main paper.
Supplementary Note 9: Identification of QP parity switching events
In order to locate the parity switching steps from the simultaneous QP parity measurements, we apply the following data processing steps. First, because the offset charge was not actively stabilized, we need to identify the portions of the data stream for each qubit where the environmental offset charge jumped to near the degeneracy point, where the parity mapping fidelity approaches zero. This involves finding the envelope of the peak-to-peak signal for the parity time trace with an applied moving average of 100 time steps. If the envelope is below a threshold determined by the qubit 0/1 readout calibration levels, the portion of data until the envelope extends above the threshold is masked off and not analyzed further when digitizing the parity time traces. We next digitize the parity time traces by applying a moving average to the unmasked raw parity data to improve the signal-to-noise ratio. We then use a hidden Markov model (HMM) to identify the parity states. For the QP parity data without junction injection presented here, we use a moving average of 40 time steps. After fitting Gaussians to the qubit 0/1 single-shot readout calibration measurements, we use these distributions to assign a probability for the parity signal to have a value along the signal axis corresponding to an odd- or even-parity state. For the HMM, we also set the probability for the system to transition from odd to even parity and vice versa based on the repetition time of the single shots and the extracted separately from the QP parity PSDs for each qubit. With this information, we then use the Viterbi algorithm to fit a digital signal to the averaged data, thus extracting the most probable parity state given the readout value along the signal axis. In a few instances, we use a modified HMM scheme for the parity analysis. This involves implementing a simple threshold method which assigns the parity of the state based on the data with an applied moving average relative to the total mean of the data. With a parity value assigned at each time index, we derive the statistics for the value of the parity signal given its state. We then use these parameters and the transition probabilities described previously to augment the HMM approach and fit a digital signal to the averaged data. Supplementary Fig. 9 shows an example of ths parity switching analysis for on the non-Cu chip.

We then use the digital signal that was found through the HMM scheme to locate parity switches. We take the absolute value of the difference of adjacent points of the digital signal, which results in a peak at the location of each parity switch. The parity switching rate for each qubit , where , is given by , where is the total number of parity switches for that particular qubit and is the total duration of unmasked data for the qubit. The uncertainty in comes from the standard Poisson counting errors .
Supplementary Note 10: Extraction of QP parity switching coincidences
Measuring the parity of all three qubits on either chip simultaneously allows us to track correlated events between qubits. In order to identify parity switching coincidences, we must process the digital parity traces obtained as described in the previous section to look for simultaneous switching between qubits. Because the moving averages that are applied to the raw parity measurement data to improve the signal-to-noise ratio also cause the switching events to have a shallower step, we must implement a windowing process to find coincidences. Because the effective width of the switching steps is approximately equal to the number of moving average time steps, we set our window size to match the number of moving averages, thus, coincident switches should occur no farther apart than the width of the falling/rising edges.

Supplementary Fig. 10 shows the effects of different window sizes for the same example parity data trace. We sweep our window through the simultaneous digital signals, and if multiple switches occur within our window size, they are identified as coincidences. For a window size well below the number of moving averages, the code misses a double coincidence [Supplementary Fig. 10(a)], while for a window size much greater than the number of moving averages, switches from separate events are misidentified as a coincidence.
Following the coincidence switching identification, the events are indexed with the appropriate type (, , , or ). With this approach, every triple coincidence is also counted as three double coincidences. We also restrict each switch of a given qubit to participate in only one event per coincidence type. For example, a switch cannot be used for two coincidences, but could be used for a coincidence and a coincidence. In Supplementary Fig. 11, we present example simultaneous parity traces for all three qubits for both chips. We also represent the locations of extracted coincidences with vertical dashed lines.
The switching rate for each type of coincidence event , where , is given by , where is the total number of events and is the total duration of unmasked data for event type . Note that double coincidences between qubits and are only counted during the period when both qubits are unmasked; similarly, triple coincidences require that all three qubits are unmasked. The uncertainty in comes from the standard Poisson counting errors .
In Supplementary Table 2, we explore the effect of different window sizes and moving averages on the observed parity switching rates for both chips. For higher averages, we observe a moderate decrease (10%) in the single-qubit switching rate, which we attribute to occasional narrow features with two closely spaced switches that get averaged below the threshold for larger numbers of moving averages. At the same time, the double- and triple-coincidence rates increase somewhat as the window size increases. Nonetheless, we still observe the same overall trend between the two chips: the single-qubit parity switching rates for the Cu chip remain 1 order of magnitude lower than for the non-Cu chip, and the double- and triple-coincidence rates are still 2 orders of magnitude lower.

Observed, random background, and extracted parity switching rates () for different window sizes and moving averages | |||||||||
Device | window size, moving average | Type | |||||||
non-Cu | 20, 20 | observed | 320(3) | 333(4) | 272(3) | 42(2) | 38(2) | 36(1) | 6.1(8) |
background | - | - | - | 21.3(4) | 18.1(3) | 17.4(3) | 1.16(2) | ||
extracted | 410(20) | 430(20) | 320(20) | 110(10) | 100(10) | 100(10) | 23(8) | ||
20, 40 | observed | 299(3) | 301(4) | 252(3) | 37(2) | 35(2) | 33(1) | 5.2(8) | |
background | - | - | - | 18.0(3) | 15.2(3) | 15.0(2) | 0.91(2) | ||
extracted | 390(20) | 390(20) | 300(20) | 100(10) | 100(10) | 90(10) | 20(7) | ||
40, 40 | observed | 299(3) | 301(4) | 252(3) | 65(2) | 60(2) | 57(2) | 12(1) | |
background | - | - | - | 36.0(6) | 30.3(5) | 30.1(5) | 3.62(8) | ||
extracted | 200(20) | 190(20) | 120(20) | 180(10) | 170(10) | 150(10) | 64(9) | ||
Cu | 20, 20 | observed | 25.9(4) | 36.5(5) | 32.7(7) | 0.56(8) | 0.6(2) | 0.5(1) | 0.06(6) |
background | - | - | - | 0.189(4) | 0.239(6) | 0.169(4) | 0.00124(4) | ||
extracted | 49(1) | 70(1) | 63(2) | 1.1(7) | 1.2(9) | 0.7(7) | 0.5(5) | ||
20, 40 | observed | 22.1(3) | 33.6(5) | 23.0(5) | 0.57(8) | 0.4(1) | 0.36(9) | 0.06(6) | |
background | - | - | - | 0.149(3) | 0.155(4) | 0.102(3) | 0.00069(2) | ||
extracted | 42(1) | 65(1) | 44(1) | 1.2(7) | 0.7(8) | 0.6(7) | 0.5(5) | ||
40, 40 | observed | 22.1(3) | 33.6(5) | 23.0(5) | 0.8(1) | 0.8(2) | 0.5(1) | 0.06(6) | |
background | - | - | - | 0.298(6) | 0.310(9) | 0.204(6) | 0.00274(9) | ||
extracted | 41(1) | 63(1) | 43(2) | 1.9(7) | 1.6(9) | 1.0(7) | 0.4(6) |
Supplementary Note 11: Identification of correlated QP poisoning rates
For a set of observed single-qubit parity switching rates with a non-zero window for identifying double- and triple-coincidence switching, one would expect a rate for random uncorrelated coincidence switching given by the product of the probabilities for observing a parity switch for each of the constituent qubits in the coincidence event during the interval . Thus, the expected random background double-concidence rate for qubits and is given by ; similarly the expected random background triple-coincidence rate for qubits , , and is given by . These expected random coincidence parity switching rates are listed in Supplementary Table 2 for different numbers of moving averages and window sizes. The error bars for these random background coincidence rates were computed by summing the fractional uncertainty for each observed rate in quadrature. We note that these random background coincidence rates remain well below the observed rates as we vary the windowing and averaging.
While the quantities we measure in our simultaneous parity measurements are the observed parity switching rates, we would like to compute the actual poisoning event rates for each qubit, or group of qubits in the case of correlated poisoning. This calculation requires accounting for the random background coincidence switching described above, as well as the probability for recording a parity switch for a given poisoning event: 1/2 in the case of single-qubit poisoning, 1/4 for double-qubit correlated poisoning, and 1/8 for triple-qubit poisoning, as described in the main paper.
For the observed parity switching, each double-coincidence event will also be recorded as two single-qubit switching events; similarly, each triple-coincidence event will also be recorded as three double-qubit switching events and three single-qubit switching events. Here, we define the extracted poisoning event rates to be exclusive; for example, a single poisoning event that couples to both and will contribute to but will not contribute to or .
Based on these criteria, we can use the observed parity switching rates to compute the probability for observing each type of parity switching event in a window interval as . We can then derive expressions for the probability of observing each type of parity switching event in terms of the actual probability for each type of poisoning event:
(2) |
With the experimentally measured switching probabilities , we numerically solve the system of equations to obtain the actual poisoning probabilities . We then calculate the actual poisoning rates . We compute the error bars on each actual poisoning probability by numerically computing the derivative with respect to each of the observed switching probabilities, then multiplying by the corresponding Poisson error bar for the observed switching probability and summing these together in quadrature.
In Supplementary Table 3, we list the values for each single qubit parity switch and coincidence event, as well as the total unmasked duration for the particular type of event. For the right three columns, the observed parity switching rates , expected random background coincidence rates , and extracted actual poisoning rates correspond to the values presented in Fig. 4(b) in the main paper.
Observed parity switching and extracted poisoning event rates | |||||||
Device | Qubit(s) | Separation | (s) | (s-1) | (s-1) | (s-1) | |
non-Cu | - | 8528 | 28,557 | 0.299(3) | - | 0.20(2) | |
- | 5202 | 17,272 | 0.301(4) | - | 0.19(2) | ||
- | 7959 | 31,609 | 0.252(3) | - | 0.12(2) | ||
5.3 mm | 832 | 12,851 | 0.065(2) | 0.0360(6) | 0.18(1) | ||
4.5 mm | 670 | 11,124 | 0.060(2) | 0.0303(5) | 0.17(1) | ||
2.0 mm | 1078 | 18,941 | 0.057(2) | 0.0301(5) | 0.15(1) | ||
- | 109 | 8,842 | 0.012(1) | 0.00362(8) | 0.064(9) | ||
Cu | - | 4031 | 182,103 | 0.0221(3) | - | 0.041(1) | |
- | 4515 | 134,192 | 0.0336(5) | - | 0.063(1) | ||
- | 1779 | 77,322 | 0.0230(5) | - | 0.043(2) | ||
5.3 mm | 66 | 78,936 | 0.0008(1) | 0.000298(6) | 0.0019(7) | ||
4.5 mm | 20 | 25,376 | 0.0008(2) | 0.000310(9) | 0.0016(9) | ||
2.0 mm | 22 | 41,277 | 0.0005(1) | 0.000204(6) | 0.0010(7) | ||
- | 1 | 15,389 | 0.00006(6) | 0.00000274(9) | 0.0004(6) |
Supplementary Note 12: QP Parity Switching Rates on Different Cooldowns
As described in Supplementary Note 14, an unplanned power outage caused our experiment to be split between two cooldowns. Most of the data was collected during the first cooldown, after the dilution refrigerator had been cold for several months. Data measured during the second cooldown was taken within a few weeks of the start of the cooldown. Supplementary Table 4 compares the observed parity switching rates and extracted poisoning rates for the non-Cu chip on the two cooldowns. Supplementary Table 5 makes the same comparison for the Cu chip. Although the cryostat was not opened in between the cooldowns and nothing was changed in the wiring, filtering, or shielding, the shorter time period after the start of the second cooldown likely resulted in incomplete thermalization of the radiative environment of the qubit, potentially involving amorphous, non-metallic elements in some of the microwave components or qubit packaging that could slowly release heat over long timescales. This would lead to higher effective blackbody temperatures and a larger flux of THz photons or enhancements to other sources of pair-breaking phonons, such as heat-only events sAastrom2006 ; sArmengaud2016 ; sAnthony2022 , thus resulting in the higher poisoning rates observed on both chips. We note that Ref. sMannila2022 also reported a slow decay in QP poisoning rates of a mesoscopic superconducting island over a timescale of several weeks with no clear mechanism for the source of the poisoning.
Observed parity switching rates (s-1) for the non-Cu chip | |||||||
Cooldown | |||||||
1 | 0.299(3) | 0.301(4) | 0.252(3) | 0.065(2) | 0.060(2) | 0.057(2) | 0.012(1) |
2 | 0.505(5) | 0.508(4) | 0.495(8) | 0.170(4) | 0.161(6) | 0.162(8) | 0.042(5) |
Extracted poisoning event rates (s-1) | |||||||
1 | 0.20(2) | 0.19(2) | 0.12(2) | 0.18(1) | 0.17(1) | 0.15(1) | 0.064(9) |
2 | 0.01(4) | 0.02(4) | 0.03(5) | 0.35(3) | 0.31(3) | 0.32(4) | 0.33(3) |
spacing | - | - | - | 5.3 mm | 4.5 mm | 2.0 mm | - |
Observed parity switching rates (s-1) for the Cu chip | |||||||
Cooldown | |||||||
1 | 0.0221(3) | 0.0336(5) | 0.0230(5) | 0.0008(1) | 0.0008(2) | 0.0005(1) | 0.00006(6) |
2 | 0.056(2) | 0.053(2) | 0.039(1) | 0.0051(9) | 0.005(1) | 0.0047(7) | 0.0003(3) |
Extracted poisoning event rates (s-1) | |||||||
1 | 0.041(1) | 0.063(1) | 0.043(2) | 0.0019(7) | 0.0016(9) | 0.0010(7) | 0.0004(6) |
2 | 0.082(8) | 0.07(1) | 0.047(8) | 0.015(6) | 0.017(6) | 0.016(5) | 0.000(3)* |
spacing | - | - | - | 5.3 mm | 4.5 mm | 2.0 mm | - |
Supplementary Note 13: Offset Charge Measurements
For a charge-sensitive qubit, besides the parity-mapping sequence, one can also perform a charge tomography sequence to measure the environmental offset charge, provided the qubit has a charge-bias line sChristensen2019 ; sWilen2021 . The Ramsey sequence involves two pulses with an idle time , where is the maximum charge dispersion for the qubit. A qubit measurement at the end of the sequence results in a 1-state probability:
(3) |
where is the sum of the externally applied gate charge and the environmental offset charge ; and are fitting parameters. Supplementary Fig. 12(a) shows an example charge tomography trace for on the non-Cu chip and a fit to Supplementary Eq. (3).
The charge tomography measurement sequence takes 28 (28.8) s and we repeat this sequence 2000 (2250) times for () on the non-Cu (Cu) chip. ( on the non-Cu chip; on the Cu chip). From the fit to each tomography scan, we extract , which we plot as a function of time over 16 (18) hours for the non-Cu (Cu) chips [Supplementary Fig. 12(b)]. From these traces, we find that large charge jumps () occur at a rate of 0.0012(1) s-1 and 0.0011(1) s-1 for a qubit on the Cu and non-Cu chips, respectively.

Based on this rate of offset charge jumps, we can estimate the rate of impacts on the chip by following the detailed analysis in Ref. sWilen2021 . In this case, the authors obtained s-1 from similar measurements of offset charge jump rates, combined with detailed modeling of the effective charge sensing area of their qubits (19,902 m2) and simulations of the charge dynamics in the Si substrate. We can approximate the charge sensing area for our qubits by taking this to be the area of the qubit shunt capacitor island extended out to half of the distance between the island and ground plane pocket (6612 m2). To estimate for our experiment, we scale the corresponding value in Ref. sWilen2021 by the ratio of the charge sensing area in Ref. sWilen2021 to that for our qubit, the ratio of our measured offset charge jump rate to that in Ref. sWilen2021 (0.00135 s-1), and the ratio of our qubit chip area [] to that in Ref. sWilen2021 [], leading to the estimate s-1 in our system.
Supplementary Note 14: Editing of Device Images

The image presented in Supplementary Fig. 13(a) was made by stitching together 20 optical micrographs to achieve a full chip picture. Once each micrograph was aligned, the composite image was converted to grayscale and the contrast increased. At this step, minor surface contamination was removed digitally to limit distraction from important device features. Finally, to remove the vignetting present in each individual picture, the image was processed using MATLAB, which identified the range of pixel values for the Nb background and altered each pixel to reflect the average value with some random noise. The result can be seen in Supplementary Fig. 13(b). The images presented in Fig. 1 and Supplementary Fig. 1 have been false-colored to highlight different parts of the chip.
References
- (1) Barends, R. et al. Minimizing quasiparticle generation from stray infrared light in superconducting quantum circuits. Applied Physics Letters 99, 113507 (2011).
- (2) Wang, C. et al. Measurement and control of quasiparticle dynamics in a superconducting qubit. Nature Communications 5, 1–7 (2014).
- (3) Vernon Jr, F. L. & Pedersen, R. J. Relaxation oscillations in Josephson junctions. Journal of Applied Physics 39, 2661–2664 (1968).
- (4) Rafferty, O. et al. Spurious antenna modes of the transmon qubit. arXiv preprint arXiv:2103.06803 (2021).
- (5) Liu, C.-H. et al. Quasiparticle poisoning of superconducting qubits from resonant absorption of pair-breaking photons. arXiv preprint arXiv:2203.06577 (2022).
- (6) Houzet, M., Serniak, K., Catelani, G., Devoret, M. & Glazman, L. Photon-assisted charge-parity jumps in a superconducting qubit. Physical Review Letters 123, 107704 (2019).
- (7) CST Studio Suite. www.3ds.com.
- (8) Ristè, D. et al. Millisecond charge-parity fluctuations and induced decoherence in a superconducting transmon qubit. Nature Communications 4, 1–6 (2013).
- (9) Serniak, K. et al. Hot nonequilibrium quasiparticles in transmon qubits. Physical Review Letters 121, 157701 (2018).
- (10) Christensen, B. et al. Anomalous charge noise in superconducting qubits. Physical Review B 100, 140503 (2019).
- (11) Åström, J. et al. Fracture processes observed with a cryogenic detector. Physics Letters A 356, 262–266 (2006).
- (12) Armengaud, E. et al. Constraints on low-mass WIMPs from the EDELWEISS-III dark matter search. Journal of Cosmology and Astroparticle Physics 2016, 019 (2016).
- (13) Anthony-Petersen, R. et al. A stress induced source of phonon bursts and quasiparticle poisoning. arXiv preprint arXiv:2208.02790 (2022).
- (14) Mannila, E. T. et al. A superconductor free of quasiparticles for seconds. Nature Physics 18, 145–148 (2022).
- (15) Wilen, C. et al. Correlated charge noise and relaxation errors in superconducting qubits. Nature 594, 369–373 (2021).