Play all audios:
ABSTRACT As quantum processors grow in complexity, new challenges arise such as the management of device variability and the interface with supporting electronics. Spin qubits in silicon
quantum dots can potentially address these challenges given their control fidelities and potential for compatibility with large-scale integration. Here we report the integration of 1,024
independent silicon quantum dot devices with on-chip digital and analogue electronics, all operating below 1 K. A high-frequency analogue multiplexer provides fast access to all devices with
minimal electrical connections, allowing characteristic data across the quantum dot array to be acquired and analysed in under 10 min. This is achieved by leveraging radio-frequency
reflectometry with state-of-the-art signal integrity, characterized by a typical signal-to-noise voltage ratio in excess of 75 for an integration time of 3.18 μs. We extract key quantum dot
parameters by automated machine learning routines to assess quantum dot yield and understand the impact of device design. We find correlations between quantum dot parameters and
room-temperature transistor behaviour that could be used as a proxy for in-line process monitoring. SIMILAR CONTENT BEING VIEWED BY OTHERS QUBITS MADE BY ADVANCED SEMICONDUCTOR MANUFACTURING
Article Open access 29 March 2022 ON-DEMAND ELECTRICAL CONTROL OF SPIN QUBITS Article 12 January 2023 ASSESSMENT OF THE ERRORS OF HIGH-FIDELITY TWO-QUBIT GATES IN SILICON QUANTUM DOTS
Article Open access 20 August 2024 MAIN Semiconductor quantum dots (QDs) are a potential platform to implement a fault-tolerant quantum computer, a novel computing paradigm expected to
outperform classical high-performance computers in areas such as materials and drug discovery, database searches and factorization. This is due to their small footprint, ability to host
coherent and controllable spin qubits, and their potential compatibility with advanced semiconductor manufacturing. Spin qubits in isotopically enriched silicon have, in particular,
demonstrated control, preparation and read-out fidelities1,2,3,4,5 above the threshold to perform quantum error correction6. However, such fault tolerance—in which an error-correcting code
reduces the error rate to a negligible amount—is predicted to require millions of physical qubits to resolve practical problems7. As solid-state quantum processors scale up to useful levels
of complexity, two important challenges must be addressed. First, the number of connections between the room-temperature processor and the quantum processor cannot continue to grow in
proportion with the number of qubits8,9,10. Frequency-division multiplexing has been applied to allow multiple qubits to share measurement electronics. However, frequency crowding limits
this approach, so far, to eight qubits per line11. Crossbar approaches12,13, in which \(O(\sqrt{N})\) lines intersect at _N_ qubit locations, offer an elegant solution to reduce wiring,
although with stringent requirements on qubit variability and limitations in processor operation. Ultimately, the use of switches to achieve time-division multiple access (TDMA), as in
dynamic random-access memory, in both read-out and control lines to each qubit unit cell provides the greatest flexibility and scalability. For d.c. signals, off-chip and on-chip cryogenic
switches with ratios up to 1:36 and 1:64, respectively, have been used to address quantum device arrays12,14,15. Multiplexed control circuitry operating at 1 K with >100 MHz pulsing has
also recently been reported16. For radio-frequency (rf) signals used for high-speed read-out, up to 1:3 switching has been shown on chip17,18,19, and high-frequency 1:4 cryogenic
multiplexers (MUXs) have been used with superconducting qubits20. Developing TDMA at scale requires an efficient interface between classical electronics and the quantum processor. The second
challenge is managing and minimizing process variability between the qubits21, requiring each qubit to be independently characterized and tuned. Minimizing process variability is already an
integral component of modern semiconductor manufacturing. However, current process characterization is optimized for classical transistors, and extending this to quantum devices requires
substantial development in high-throughput cryogenic testing capabilities. State-of-the-art methods rely on newly developed cryogenic probing to enable wafer-scale testing22. However, this
method is currently limited to temperatures above 1.6 K and is unable to efficiently utilize the wafer space since nanometre-scale devices need to be directly contacted to macroscopic pads.
On-chip multiplexing techniques can provide access to large numbers of densely packed devices with minimal input/output connections, whereas testing can be performed under the optimal
temperature conditions of spin qubit devices (millikelvins)12,17. Large-scale on-chip switching of rf signals, therefore, addresses both short-term and long-term challenges in quantum
computer development: providing a way to characterize many quantum devices to address process variability, and to address many qubits in scaled-up quantum processors. In this Article, we
report the rapid characterization of 1,024 QD devices fabricated using a commercial foundry process by integrating high-speed classical multiplexing electronics to individually address each
device. We develop tools to automatically extract key indicators for QD performance and their suitability for use in qubit technologies. We perform a statistical analysis of different device
dimensions, and under varied operating conditions. Ultimately, we show that the electrostatic properties of these devices can be characterized at cryogenic temperatures in under 10 min of
measurement time (Supplementary Section I) by means of rf reflectometry techniques. We also establish a link between cryogenic and room-temperature device properties, opening a new avenue
for pre-cryogenic validation of silicon qubit technologies. CRYOGENIC MULTIPLEXED ACCESS TO 1,024 QDS We use an all-to-all MUX that enables the selection of a single cell given a 10-bit
input. It contains an analogue bus of three device control lines and ten digital address lines (five row-select lines and five column-select lines) to address each QD device under test in a
32 × 32 array. The devices are selectively connected to the analogue bus using complementary metal–oxide–semiconductor (CMOS) transmission gates integrated within the same silicon as the
quantum devices. The integrated circuit is designed using an ultrathin body and buried oxide fully depleted silicon-on-insulator process. QDs form in the undoped channel of the transistors
when a voltage difference _V_GS between the gate and source approaches the threshold voltage (Fig. 1b). To observe the discrete charging of QDs, a source and drain tunnelling resistance
larger than the resistance quantum is necessary. This occurs naturally in these devices due to the increase in resistivity of the undoped silicon region at low temperatures. This region lies
beneath the gate and spacers (Fig. 1b). In Fig. 1c, we show an example of the characteristic discrete charging, which is a diamond-shaped region of decreased conductance nestled within
regions of higher conductance. REFLECTOMETRY PERFORMANCE VIA A MUX To expedite device measurements, we use rf reflectometry23,24,25,26. Reflectometry can detect QD charge transitions with
high bandwidth by measuring changes in device impedance. The typically high device impedance (>100 kΩ) is matched to a 50 Ω line by embedding the integrated circuit in a matching network
containing a superconducting spiral inductor (Methods). This allows us to monitor the reflected voltage as a function of the device impedance. Here we characterize the performance of this
technique in terms of the signal-to-noise ratio (SNR) and bandwidth at a farm level. A peak in the |_S_11| trace corresponds to the electrochemical potential of the QD, aligning with the
Fermi level of the source or drain. We define the signal as the height of this peak relative to the mean background, and the noise as the standard deviation of the background signal. In Fig.
2a, we show the SNR2 value of a QD charge transition as a function of integration time _τ_. Through linear regression assuming a _y_ intercept of zero, we determine the minimum integration
time required to attain an SNR of 1, which amounts to _t_min = 556 ± 6 ps. This minimum integration time serves as a benchmark for evaluating the read-out performance26 and demonstrates that
the MUX does not compromise the signal quality. In fact, our apparatus outperforms the state of the art achieved with reflectometry in single-electron transistors27. Overall, in the TDMA
implementation presented here, the integration time is fixed at _t_int ≈ 11 μs (Supplementary Section I). Next, we test the bandwidth at the farm level by looking at the dependence of SNR
with respect to the probe frequency for nine example devices (Fig. 2b). For TDMA, it is critical that the frequency region of a high signal overlaps for each device in the farm. We
demonstrate that this holds, with the average SNR bandwidth (6.4 MHz), defined as the full-width at half-maximum of the SNR2 signal, which is similar to the resonator bandwidth (9.5 MHz).
Finally, we show how the cryogenic performance of a fully depleted silicon-on-insulator process28,29 is pivotal in creating our low-temperature multiplexing circuit. In particular,
back-gating through the buried oxide enables the compensation of the known transistor threshold voltage increase at low temperatures, which becomes important when delivering high-frequency
signals through the MUX (Fig. 3). In particular, we measure the on resistance _R_on of a single transmission gate over a wide back-gate-voltage range (_V_NW,PW for the n- and p-type
field-effect transistors, NFET and PFET, respectively) and for multiple common-mode drain/source voltages. During these measurements, the NFET gate is held at _V_DD = 0.8 V and the PFET gate
is held at _V_SS = 0 V. By applying a forward back-bias (positive for NFET and negative for PFET), we can reduce _R_on by more than an order of magnitude for a common-mode voltage _V_D =
_V_S = 0.4 V (Fig. 3a). We note that the combination of forward back-bias and the large dimensions of the MUX transistors prevents the undesirable occurrence of Coulomb blockade in the
transmission gate. The impact of back-biasing on rf performance is also evident from the quality of the reflectometry signal. Figure 3b shows two Coulomb oscillations as measured in
reflectometry as the MUX back-bias increases. The two oscillations are well resolved when _V_NW = −_V_PW > 1.5 V. However, for lower values, a substantial voltage drop occurs at the MUX,
shifting the position of the oscillations. The shift is accompanied by a reduction in SNR, which—along with the unstable signal behaviour—can be linked to the increasing MUX resistance. We
highlight that the back-gate potentials applied to the analogue circuitry are independent of those applied to the QD devices. ANALYSIS AND EXTRACTION OF QD FEATURES AT SCALE To characterize
the QD devices, we develop a method to automatically extract (Methods) the first observed electron loading voltage (_V_1e), the gate lever arm (_α_G) that describes the strength of the
electrostatic coupling of the gate electrode to the dot, and the source–drain lever arm difference (_α_D − _α_S) to measure the device asymmetry. More specifically, we define _V_1e as the
gate voltage at which we first detect a Coulomb oscillation at zero _V_DS, representing the loading of an electron to a QD that may already contain a number of electrons. These parameters
are the ones that can be extracted unambiguously, using the first measured Coulomb blockade oscillation. To extract further electrostatic properties, like the charging energy, further
oscillations are required; however, for many measured devices, the presence of additional QDs cannot be conclusively ruled out. Furthermore, a secondary charge sensor would be required to
verify that we have reached the single-electron regime; therefore, here, instead, we find the distribution of gate voltages corresponding to the first visible transition using our read-out
methods (reflectometry and d.c. transport) to establish the trends and variability between devices. Though we cannot guarantee reaching the single-electron regime, we have identified these
QDs as being in the few-electron regime30 (Supplementary Section II). As stated above, a single measurement is used to extract all of these parameters automatically, by monitoring the device
as _V_GS and _V_DS are varied (Methods). A measurement of this kind is shown in Fig. 1c. In the farm, eight transistor variants were tested with increasing gate lengths (_L_) of 28 nm, 40
nm, 60 nm and 80 nm and channel widths (_W_) of 80 nm and 100 nm. Our first observation shows that not all transistors can provide good QDs (Fig. 4a). For these devices, we, therefore,
cannot extract the QD parameters described above, and therefore, they must first be filtered out. We have trained a convolutional neural network (CNN) to categorize our devices into three
categories: clear Coulomb blockade (good), no Coulomb blockade (bad) and multiple series QDs (multi), which present as non-closing diamonds (Supplementary Section IV provides more
information regarding the classification criteria). Each device is manually labelled by two domain experts and these labels are used to train the CNN (Methods). The proportion of devices
that fall into each category as determined by the CNN is shown in Fig. 4b. To automatically extract the parameters, we developed tools to process data acquired through both transport (d.c.)
and reflectometry (rf) measurements. Transistor characteristic measurements are commonly performed in d.c.; thus, the d.c. data presented here serve as a reference for comparison with the rf
data, revealing good agreement between both techniques (Fig. 4c–e and Extended Data Fig. 1). Moreover, increasing the integration time does not result in a substantive change in the
extracted parameters (Extended Data Fig. 2). The parameters _V_1e, _α_G and _α_D − _α_S can be determined from the pair of intersecting lines in a charge map (Fig. 1c). The intercept gives
_V_1e, and the lever arms can be calculated from the gradients of the two lines (Methods). For each charge map, we perform a fitting routine that finds the best pair of lines, with higher
weight given to _V_1e at lower voltages, and intercept close to _V_DS = 0 V. VARIABILITY OF INDUSTRIALLY FABRICATED QDS To assess the inherent process variability, considerable effort was
focused on the design stage to suppress the known sources of semiconductor process variability, for example, layout effects (Methods). Therefore, the variability we measure is primarily
inherent to each device under test. As shown in Fig. 4b, in devices with greater gate lengths, multiple dots are generated, resulting in complex multi-dot structures. On the contrary,
devices with shorter gate lengths yield single QDs. However, in most devices, a small gate length leads to an early transistor turn-on, resulting in ‘bad’ dots. Overall, the device designs
with the highest proportion of good QD features have shorter gate lengths, namely, 28 nm and 40 nm. The two channel widths do not impart notable differences across the farm. The
automatically extracted parameters for these good devices are shown in Fig. 4c–e. With decreasing gate length, we see a lower threshold voltage due to the increasing effect of the electric
field produced by the source and drain, a well-known short-channel transistor effect, probably caused by drain-induced barrier lowering31. The gate lever arm and lever arm asymmetry remain
fairly constant, indicating that dots remain well controlled by the gate even at the smallest dimensions. For the _L_ = 28 nm case, we find the first observed electron voltages
(\(\overline{{V}_{{\rm{1e}}}}=387\pm 22\,{\rm{m}}\,{\rm{V}}\)) and the gate lever arms (\(\overline{{\alpha }_{{\rm{G}}}}=0.741 \pm 0.082\)) are narrowly distributed. We find the standard
deviation of _V_1e (~22 mV) is comparable to the spacing (25 ± 4 mV) between the first and second observed electron (_V_2e) loading voltages, measured from a subset of devices in which a
second transition is clearly visible. This suggests that the tight requirements for shared voltage control12 are within reach, but need further reduction. Alternatively, small variations may
be compensated with independent voltage trimming of each of the QD back-gates. However, this comes at the cost of a greater number of control lines per QD. It must be stressed that the
overarching challenge with the presented devices is QD quality, which must be addressed before voltage sharing can be considered. Determining the exact origin of the variability in QD
quality is beyond the scope of this work, but we hypothesize that irregularities in electrostatic potential caused by (1) an elevated number of two-level fluctuators near the Si/SiO2
interface30 and/or (2) gate metal workfunction inhomogeneities32 could be the cause. We note that the large gate lever arm will be beneficial when implementing gate-based dispersive
read-out26. The mean dot asymmetry \(\overline{{\alpha }_{{\rm{D}}}-{\alpha }_{{\rm{S}}}}=-0.040 \pm 150\) shows that, on average, the QDs are well centred in the channel. We note that since
all the lever arms must add to 1, a large asymmetry places an upper bound on the gate lever arm, that is, _α_G ≤ 1 − ∣_α_D − _α_S∣ (Supplementary Section V). This result highlights an
inverse relation between the asymmetric QD position within the channel and gate control over the QD, emphasizing the importance of QD location being central under the gate. We note that for
the devices here, the QD is placed between two large conducting leads; however, for larger QD arrays, most dots may only have a single lead, or only other dots nearby. In such devices, the
importance of the short-channel effect we see here is lessened. In light of this, our measurements provide a worst-case indication for dot asymmetry under a single gate. ROOM-TEMPERATURE
CORRELATIONS WITH QD PARAMETERS It would be ideal to determine the QD parameters without needing to cool the device to cryogenic temperatures. For a QD, once the thermal energy _k_B_T_
becomes much larger than the charging energy _E_C = _e_2/_C_Σ ≈ 18.5 meV (Methods), transport through the transistor does not exhibit blockade and the device behaves as a simple transistor.
However, we gain a new set of parameters used in the classical modelling of transistors, such as the threshold voltage (_V_th). We next establish a direct link between QD parameters and
classical transistor behaviour (Methods), which allows device yield and uniformity to be assessed without requiring expensive and time-intensive cooling in a dilution refrigerator. Recent
work has correlated classical transistor behaviour from room temperature to 4.2 K (ref. 33); here we extend this to the behaviour of QDs at lower temperatures. We highlight that
pre-cryogenic validation can be an invaluable tool in the life cycle of a silicon quantum computer. We explore the relation between _V_1e and _V_th obtained from devices with the shortest
gate length, _L_ = 28 nm, classified as ‘good’. To explore this, we use probabilistic programming, a technique that offers an insight into both systematic patterns and random fluctuations in
the data. We consider a simple linear model of the form $${V}_{{\rm{1e}}}({V}_{{\rm{th}}})=\alpha {V}_{{\rm{th}}}+\beta +{\mathcal{N}}(0,\,{\sigma }_{{{\rm{V}}}_{1{\rm{e}}}}).$$ (1) Here
_α_ and _β_ are coefficients of the linear model, and \({\mathcal{N}}(0,\,{\sigma }_{{{\rm{V}}}_{1{\rm{e}}}})\) is a normal distribution that represents the intrinsic random fluctuations in
_V_1e. In this approach, _V_1e, _V_th and the parameters _α_, _β_ and \({\sigma }_{{{\rm{V}}}_{1{\rm{e}}}}\) are treated as random variables rather than fixed quantities. We express our
initial expectations using prior distributions, which convey our preliminary knowledge about the parameters before any data are collected. For the priors, we opt for normal distributions
(\({\mathcal{N}}\)), as they offer a balanced representation of our initial understanding without being overly restrictive. Next, we refine these parameter distributions using Hamiltonian
Monte Carlo34, aiming to find the best fit to the observed data for _V_1e. Figure 5a shows the linear fit extracted by averaging the Hamiltonian Monte Carlo trial results. The extracted
slope (\(\overline{\alpha }=1.01\pm 0.02\)) indicates a clear relationship between room-temperature threshold voltage and the first observed QD electron loading voltage. It is well known
that transistor threshold voltages increase at cryogenic temperatures, and this is also apparent in the _V_1e offset voltage, \(\overline{\beta }=210\pm 30\,{\rm{m}}\,{\rm{V}}\). The result
from the model is the posterior distribution of _V_1e. The uncertainty of the posterior distribution (~22 mV) encompasses variations arising from both \(\overline{{V}_{{\rm{th}}}}\)
(\(\overline{{\sigma }_{{{\rm{V}}}_{{\rm{th}}}}}=15\,{\rm{m}}{\rm{V}}\)) and the intrinsic randomness in _V_1e (\(\overline{{\sigma }_{{{\rm{V}}}_{1{\rm{e}}}}}=16\pm 1\,{\rm{m}}{\rm{V}}\)).
Completely removing variations in the threshold voltage at the foundry level (\({\sigma }_{{{\rm{V}}}_{{\rm{th}}}}=0\)) would consequently decrease the variation in _V_1e to \({\sigma
}_{{{\rm{V}}}_{1{\rm{e}}}}=16\,{\rm{m}}{\rm{V}}\). CONCLUSIONS We have reported testing a device farm of 1,024 QDs based on simple transistor structures. Our approach can, however, be
extended to more complex unit cells, such as coupled QD systems—the basic building block of semiconductor-based quantum computers. Our rf read-out techniques can be leveraged to embed
compact dispersive spin qubit read-out with the unit cells of scaled-up QD architectures35,36. We developed an integrated CMOS three-channel 1:1,024 analogue MUX, but the capabilities of
this silicon technology reach further. Foundry-based platforms could, in particular, allow the co-integration of ultralow power electronic modules such as digital-to-analogue converters,
low-noise amplifiers and digital controllers alongside the qubit system. These technologies have been demonstrated in stand-alone processes37, but tightly integrating all of these modules
with semiconductor qubits—and retaining their qualities—remains an open challenge, especially given the limited cooling power of cryostats and thermal conductivity of silicon at low
temperatures. Our observation that cryogenic parameters of silicon QD devices can be predicted from room-temperature behaviour has implications for the time and resources required to monitor
process variations and optimize the design and production of future quantum devices. Further development of pre-cryogenic methods and analysis tools could allow wider industry engagement
and a substantial cost reduction in technology development, particularly if further correlations can be extracted when complex unit cells are studied. METHODS D.C. TRANSPORT MEASUREMENT With
a small bias of ~ 1 mV, the transport current formed by the tunnelling of electrons one by one through a QD is typically small (~1 pA to 1,000 pA). To record this level of current, we use a
transimpedance amplifier to convert the current to a voltage with a gain of 107 V A−1. To acquire a Coulomb diamond map, a triangular wave with a frequency of 20 Hz was applied to the gate
and the signal is acquired on the rising edge of the slope. Devices were measured with either five or ten averages; a single transport Coulomb diamond measurement takes approximately 15 s or
30 s. RF REFLECTOMETRY MEASUREMENT Measuring a high-impedance device in reflectometry requires an impedance-transforming circuit (Supplementary Section VI). This allows changes in an
otherwise large device impedance to be measured by a 50 Ω matched meter, for example, resistance changing from 1,000 kΩ to 100 kΩ. The complex impedance of the QD has contributions from the
resistance and capacitance of the tunnel barrier between the source lead and the dot. When the dot and source electrochemical potentials are equal, the Coulomb blockade is lifted and
electrons can tunnel elastically through the barrier. This results in an apparent change in the barrier resistance and capacitance, the latter of which may have both quantum and junction
contributions38,39. This impedance change can be detected at the 50 Ω output as a change in the reflected rf signal near the resonance frequency. To measure the Coulomb diamond maps, a
triangular waveform with a frequency of 203 Hz was applied to the gate. Signal acquisition occurs during the rising edge of the waveform. The frequency of the triangular waveform is
constrained by the _R__C_ low-pass filter on the printed circuit board with a cutoff frequency of _f_c = 16 kHz. Each Coulomb diamond map comprises 60 distinct traces, covering a range of
source–drain voltages (_V_DS) from –15 mV to 15 mV. The acquisition time required for measuring the 1,024 Coulomb diamond maps is 2 min 31 s. The sampling rate used for acquiring these maps
is 1 MSa s−1. However, the effective integration time is constrained to 10.68 μs due to limitations imposed by the room-temperature low-pass filters for the in-phase and quadrature signals
(Supplementary Section I). COULOMB DIAMOND CLASSIFICATION USING A NEURAL NETWORK We categorize devices into three groups based on their Coulomb blockade maps (Fig. 4a). * Good: these devices
display a clearly defined hourglass shape in their Coulomb blockade map, enabling parameter extraction (Fig. 1c). * Bad: these devices lack observable Coulomb blockade or exhibiting
classical transistor turn-on superimposed with Coulomb blockade. This behaviour is probably caused by low resistances of the tunnelling barriers. * Multi: these devices form several QDs in
series, identifiable by overlapping Coulomb diamonds, giving rise to extended regions of blockade. On the basis of the above criteria, domain experts manually classified the devices.
However, the boundaries between categories can be ambiguous, as evidenced by experts agreeing on the classification approximately 80% of the time for the same dataset. Examples of such
ambiguity are faint hourglass shapes mixed with noise or incomplete hourglass shapes. Although manual classification worked well with the current number of devices (approximately 1,000), the
expected increase in device volume calls for an automated solution. To tackle this challenge with our modest dataset, we applied strategies proven effective in image classification with
limited training data. Specifically, we utilize transfer learning by implementing the well-known ResNet26d architecture, pre-trained on the extensive ImageNet’s database. This approach
enables robust performance despite the scale of our dataset40. Alongside transfer learning, we used data augmentation techniques like image rotation, warping, zooming and changes in
saturation. These methods introduce variety into the training data. In addition, we incorporate mixup, a technique that generates new images by linearly combining pairs of original training
images41. The neural network processes Coulomb blockade data as greyscale images. The dataset was randomly partitioned, with 80% allocated to a training set and 20%, to a test set. To
optimize the training process, we used the one-cycle policy, which dynamically adjusts the learning rate, increasing it to the maximum and then gradually decreasing it42. This dynamic
learning rate schedule helps regulate the training process, accelerates convergence and reduces sensitivity to the hyperparameter of the learning rate, ensuring a more robust and efficient
training phase. In our neural network implementation, we focus on capturing label uncertainty. This is achieved through label smoothing43. In standard binary assignment, like entropy loss,
the correct class receives a probability of 1 and others, 0. On the contrary, label smoothing adjusts the ground-truth labels by assigning a probability slightly less than 1 to the correct
class and distributing the remaining probability uniformly across all classes. This adjustment helps mitigate over-reliance on specific training labels and encourages the model to generalize
better across different classes. The performance of the CNN is assessed using the confusion matrix (Extended Data Table 1), yielding an accuracy of 88%. This accuracy level is constrained
due to the ambiguous boundaries between different device classes. To demonstrate that, we evaluated the ChimeraMix architecture, a state-of-the-art approach, which achieved over 96%
classification accuracy using labels from a single expert. ChimeraMix relies on training a generative adversarial network to mix examples of the same class, augmenting the size of the
dataset. Finally, to optimize the neural network further, we suggest a non-binary scoring system for each Coulomb map (for example, using a scale from one to ten). This approach involves
training the network on labels that reflect the collective agreement among multiple human classifiers, which helps mitigate individual biases. Moreover, adopting this method enables the
neural network to better understand the nuances present in edge cases between good, multi and bad dots. PARAMETER EXTRACTION FROM COULOMB DIAMONDS Most key parameters that describe a single
QD can be obtained from the position and shape of the Coulomb diamond measurements44,45. When measuring the rf response at the drain, the positive (negative) edges of the Coulomb diamonds
appear when the QD electrochemical potential _μ_dot is aligned with the drain (source) electrochemical potential _μ_D (_μ_S). The first observed electron loading voltage _V_1e corresponds to
the crossing point of the first pair of these edges (nominally at _μ_S = _μ_D = 0 eV). The gate lever arm _α_G = _C_G/_C_Σ, where _C_G is the gate capacitance and _C_Σ is the sum of the dot
capacitance to each terminal, represents the coupling strength of the gate to the dot. This parameter can be calculated as $${\alpha }_{{\rm{G}}}={\left(\frac{1}{| {m}_{1}| }+\frac{1}{|
{m}_{2}| }\right)}^{-1},$$ (2) where _m_1 and _m_2 are the positive and negative gradients of the pair of edges that form the first visible hourglass (Fig. 1c). Moreover, when a voltage bias
is applied antisymmetrically (_V_D = −_V_S), the relative coupling capacitance of the source (_α_S = _C_S/_C_Σ) and drain (_α_D = _C_D/_C_Σ) can be directly obtained from the gradient of
the Coulomb diamond edges as $${\alpha }_{{\rm{D}}}-{\alpha }_{{\rm{S}}}=\frac{{m}_{1}+{m}_{2}}{{m}_{1}-{m}_{2}}.$$ (3) This is a measure of the asymmetry of dot formation under the gate.
The full step-by-step process for extracting the dot parameters from Coulomb blockade maps is detailed in Supplementary Section IV, but we provide a succinct summary here. We perform digital
filtering to reduce noise and enhance contrast in the acquired data, and then apply a Canny edge detection algorithm to digitize the charge stability map and identify the edges of the
Coulomb diamonds. We then use a Hough transform to convert the binary image to information about the edges parametrized by their length and angle. We identify good fits to the first visible
Coulomb oscillation with a pair of long line segments, one with a positive slope and the other with a negative one, which intersect near _V_DS = 0 V. ROOM-TEMPERATURE MEASUREMENTS AND
CRYOGENIC CORRELATIONS We analyse the transport measurements through transistors at room temperature with a source–drain bias of _V_DS = 50 mV. _V_th is determined by extrapolating the
_I_D–_V_GS curve, between the maximum transconductance point max(_g_m) = max(d_I_D/d_V_GS) and the maximum subthreshold slope point max(SS) = max(log[_I_D/_V_GS]) (ref. 46). _V_th is the
voltage at which this extrapolated line intersects _I_D = 0. To understand the systematic relationship between _V_1e and _V_th and their random variation, we use the probabilistic
programming framework NumPyro47. As stated in the main text, we use a model defined by three random variables: _α_, _β_ and _σ_. These variables are initially set with a suitable prior
distribution. The posterior distribution is then formed using Hamiltonian Monte Carlo34 sampling. CHARGING ENERGY We obtain a charge energy of _E_C = 18.5 ± 3 meV as _E_C = ∣_e_∣Δ_V_G_α_G,
where ∣_e_∣ ≈ 1.6 × 10−19 C is the elementary charge, _α_G is the gate lever arm and Δ_V_G is the difference between the first and second detected electron (_V_2e) loading voltages. Some
examples of Coulomb diamonds can be found in the main text showing slightly smaller charging energies than the average. DEVICE FABRICATION The die was fabricated using the GlobalFoundries
22FDX 22 nm fully depleted silicon-on-insulator process. A single die was used for this study. EXPERIMENTAL SETUP The integrated circuit die was glued to a carrier printed circuit board with
conducting silver paste. The die is wire bonded (17.5 μm AlSi) to gold-plated copper tracks. The low-frequency control lines are routed to a connector to attach to a motherboard, which has
passive first-order in-line filtering (_RC_ = 10 μs). The reflectometry line is directly routed to an SMP connector on the carrier printed circuit board. All the cryogenic measurements were
performed in a Bluefors XLD dilution refrigerator, where the device was mounted to the mixing chamber plate operating at 10 mK. When the chip is powered on, the die temperature was measured
as 600 mK using on-chip thermometry48 (value quoted for a nominally identical die). The reason for this elevated temperature is the static power draw of approximately 4 μW needed for the
digital and analogue support electronics. A QDevil QDAC II was used to supply the d.c. voltages to the QD device terminals, chip supplies and row/column address lines. To sweep the gate
voltage, a triangular wave with 50% duty cycle was supplied by a Keysight 33500B arbitrary waveform generator. The rf reference signal was generated by a Rohde & Schwarz SMB100B. The
reflected signal was amplified by a Low Noise Factory cryogenic amplifier (LNF-LNC0.2_3A s/n 2541z) at 4 K. This signal is then further amplified at room temperature using two Mini-Circuits
amplifiers (ZX60-P103LN+ and ZX60-33LNR-S+) before being separated into its in-phase and quadrature components using a Polyphase microwave quadrature demodulator (AD0540B). These signals are
finally amplified and filtered by a Stanford SR560 and then digitized via a Spectrum M4i.4421-x8 digitizer PCIe card. For d.c. measurements, we used two transimpedance amplifiers (Basel
Precision Instruments SP983c, IF3602 junction field-effect transistor) to simultaneously monitor the source (_I_S) and drain (_I_D) currents. The gain of the amplifiers was set to 107 V A−1
with a low-pass filter bandwidth of 1 kHz. For our measurements, we use _I_sig = (_I_D − _I_S)/2 to remove any offsets. During rf measurements, the transimpedance amplifiers are removed and
the source and drain are directly driven with the QDAC II voltage sources mentioned above. DATA AVAILABILITY The data that support the plots within this paper and other findings of this
study are available from the corresponding author upon request. REFERENCES * Xue, X. et al. Quantum logic with spin qubits crossing the surface code threshold. _Nature_ 601, 343–347 (2022).
Article MATH Google Scholar * Noiri, A. et al. Fast universal quantum gate above the fault-tolerance threshold in silicon. _Nature_ 601, 338 (2022). Article MATH Google Scholar *
Mills, A. R. et al. Two-qubit silicon quantum processor with operation fidelity exceeding 99%. _Sci. Adv._ 8, eabn5130 (2022). Article Google Scholar * Yoneda, J. et al. Quantum
non-demolition readout of an electron spin in silicon. _Nat. Commun_. 11, 1144 (2020). * Blumoff, J. Z. et al. Fast and high-fidelity state preparation and measurement in triple-quantum-dot
spin qubits. _PRX Quantum_ 3, 010352 (2022). Article Google Scholar * Fowler, A. G., Mariantoni, M., Martinis, J. M. & Cleland, A. N. Surface codes: towards practical large-scale
quantum computation. _Phys. Rev. A_ 86, 032324 (2012). Article Google Scholar * Beverland, M. E. et al. Assessing requirements to scale to practical quantum advantage. Preprint at
https://arxiv.org/abs/2211.07629 (2022). * Gonzalez-Zalba, M. F., Barraud, S., Ferguson, A. J. & Betz, A. C. Probing the limits of gate-based charge sensing. _Nat. Commun._ 6, 6084
(2015). Article MATH Google Scholar * Franke, D., Clarke, J., Vandersypen, L. & Veldhorst, M. Rent’s rule and extensibility in quantum computing. _Microprocess. Microsyst._ 67, 1
(2019). Article MATH Google Scholar * Vandersypen, L. M. K. et al. Interfacing spin qubits in quantum dots and donors—hot, dense, and coherent. _npj Quantum Inf._ 3, 34 (2017). Article
Google Scholar * George, R. E. et al. Multiplexing superconducting qubit circuit for single microwave photon generation. _J. Low Temp. Phys._ 189, 60 (2017). Article MATH Google Scholar
* Bavdaz, P. L. et al. A quantum dot crossbar with sublinear scaling of interconnects at cryogenic temperature. _npj Quantum Inf._ 8, 86 (2022). Article Google Scholar * Li, R. et al. A
crossbar network for silicon quantum dot qubits. _Sci. Adv._ 4, eaar3960 (2018). Article Google Scholar * Al-Taie, H. et al. Cryogenic on-chip multiplexer for the study of quantum
transport in 256 split-gate devices. _Appl. Phys. Lett._ 102, 243102 (2013). Article MATH Google Scholar * Puddy, R. K. et al. Multiplexed charge-locking device for large arrays of
quantum devices. _Appl. Phys. Lett._ 107, 143501 (2015). Article Google Scholar * Jadot, B. et al. A cryogenic active router for qubit array biasing from d.c. to 320 MHz at 100 nm gate
pitch. In _ESSCIRC 2023- IEEE 49th European Solid State Circuits Conference (ESSCIRC)_ 157–160 (IEEE, 2023). * Ruffino, A. et al. A cryo-CMOS chip that integrates silicon quantum dots and
multiplexed dispersive readout electronics. _Nat. Electron._ 5, 53 (2022). Article MATH Google Scholar * Schaal, S. et al. A CMOS dynamic random access architecture for radio-frequency
readout of quantum devices. _Nat. Electron._ 2, 236 (2019). Article MATH Google Scholar * Ward, D. R., Savage, D. E., Lagally, M. G., Coppersmith, S. N. & Eriksson, M. A. Integration
of on-chip field-effect transistor switches with dopantless Si/SiGe quantum dots for high-throughput testing. _Appl. Phys. Lett._ 102, 213107 (2013). Article Google Scholar * Acharya, R.
et al. Suppressing quantum errors by scaling a surface code logical qubit. _Nature_ 614, 676 (2023). Article MATH Google Scholar * Gonzalez-Zalba, M. F. et al. Scaling silicon-based
quantum computing using CMOS technology. _Nat. Electron._ 4, 872 (2021). Article MATH Google Scholar * Neyens, S. et al. Probing single electrons across 300-mm spin qubit wafers. _Nature_
629, 80–85 (2024). Article Google Scholar * Schoelkopf, R. J., Wahlgren, P., Kozhevnikov, A. A., Delsing, P. & Prober, D. E. The radio-frequency single-electron transistor (RF-SET): a
fast and ultrasensitive electrometer. _Science_ 280, 1238–1242 (1998). * Duty, T. et al. Observation of quantum capacitance in the Cooper-Pair transistor. _Phys. Rev. Lett._ 95, 206807
(2005). Article MATH Google Scholar * Wallraff, A. et al. Strong coupling of a single photon to a superconducting qubit using circuit quantum electrodynamics. _Nature_ 431, 162 (2004).
Article MATH Google Scholar * Vigneau, F. et al. Probing quantum devices with radio-frequency reflectometry. _Appl. Phys. Rev._ 10, 021305 (2023). * Keith, D. et al. Single-shot spin
readout in semiconductors near the shot-noise sensitivity limit. _Phys. Rev. X_ 9, 041003 (2019). MATH Google Scholar * Beckers, A. et al. Characterization and modeling of 28-nm FDSOI CMOS
technology down to cryogenic temperatures. _Solid-State Electron._ 159, 106 (2019). Article MATH Google Scholar * Cardoso Paz, B. et al. Performance and low-frequency noise of 22-nm
FDSOI down to 4.2 K for cryogenic applications. _IEEE Trans. Electron Devices_ 67, 4563 (2020). Article MATH Google Scholar * de Kruijf, M., Noah, G. M., Gomez-Saiz, A., Morton, J. J.
& Gonzalez-Zalba, M. Measurement of cryoelectronics heating using a local quantum dot thermometer in silicon. _Chip_ 3, 100097 (2024). * Troutman, R. & Fortino, A. Simple model for
threshold voltage in a short-channel IGFET. _IEEE Trans. Electron Devices_ 24, 1266 (1977). Article MATH Google Scholar * Zeng, Z., Triozon, F. & Niquet, Y.-M. Carrier scattering by
workfunction fluctuations and interface dipoles in high-_K_/metal gate stacks. In _2016 International Conference on Simulation of Semiconductor Processes and Devices (SISPAD)_ 369–372 (IEEE,
2016). * Paz, B. C. et al. Integrated variability measurements of 28 nm FDSOI MOSFETs down to 4.2 K for cryogenic CMOS applications. In _2020 IEEE 33rd International Conference on
Microelectronic Test Structures (ICMTS)_ 1–5 (IEEE, 2020). * Hoffman, M. D. & Gelman, A. The no-U-turn sampler: adaptively setting path lengths in Hamiltonian Monte Carlo. _J. Mach.
Learn. Res._ 15, 1593–1623 (2014). * Veldhorst, M., Eenink, H. G. J., Yang, C. H. & Dzurak, A. S. Silicon CMOS architecture for a spin-based quantum computer. _Nat. Commun._ 8, 1766
(2017). Article Google Scholar * Oakes, G. A. et al. Fast high-fidelity single-shot readout of spins in silicon using a single-electron box. _Phys. Rev. X_ 13, 011023 (2023). MATH Google
Scholar * Xue, X. et al. CMOS-based cryogenic control of silicon quantum circuits. _Nature_ 593, 205 (2021). Article MATH Google Scholar * Luryi, S. Quantum capacitance devices. _Appl.
Phys. Lett._ 52, 501 (1988). Article Google Scholar * Büttiker, M., Thomas, H. & Prêtre, A. Mesoscopic capacitors. _Phys. Lett. A_ 180, 364 (1993). Article MATH Google Scholar *
Yosinski, J., Clune, J., Bengio, Y. & Lipson, H. How transferable are features in deep neural networks? _Adv. Neural Inf. Process. Syst._ 27, 3320–3328 (2014). MATH Google Scholar *
Zhang, H., Cisse, M., Dauphin, Y. N. & Lopez-Paz, D. mixup: beyond empirical risk minimization. Preprint at https://arxiv.org/abs/1710.09412 (2017). * Smith, L. N. & Topin, N.
Super-convergence: very fast training of neural networks using large learning rates. In _Artificial Intelligence and Machine Learning for Multi-Domain Operations Applications_ 1100612 (SPIE,
2019). * Müller, R., Kornblith, S. & Hinton, G. E. When does label smoothing help? _Adv. Neural Inf. Process. Syst._ 32, 4694–4703 (2019). MATH Google Scholar * Hanson, R.,
Kouwenhoven, L. P., Petta, J. R., Tarucha, S. & Vandersypen, L. M. K. Spins in few-electron quantum dots. _Rev. Mod. Phys._ 79, 1217 (2007). Article Google Scholar * Yang, T. Y. et al.
Quantum transport in 40-nm MOSFETs at deep-cryogenic temperatures. _IEEE Electron Device Lett._ 41, 981 (2020). MATH Google Scholar * Pierre, M. et al. Single-donor ionization energies in
a nanoscale CMOS channel. _Nat. Nanotechnol._ 5, 133 (2010). Article MATH Google Scholar * Phan, D., Pradhan, N. & Jankowiak, M. Composable effects for flexible and accelerated
probabilistic programming in NumPyro. Preprint at https://arxiv.org/abs/1912.11554 (2019). * Noah, G. M. et al. CMOS on-chip thermometry at deep cryogenic temperatures. _Appl. Phys. Rev._
11, 021414 (2024). Article MATH Google Scholar Download references ACKNOWLEDGEMENTS E.J.T. acknowledges the Engineering and Physical Sciences Research Council (EPSRC) through the Centre
for Doctoral Training in Delivering Quantum Technologies (grant no. EP/S021582/1). M.F.G.-Z. acknowledges a UKRI Future Leaders Fellowship (grant no. MR/V023284/1). We thank N. Cave at
GlobalFoundries for helpful discussions. AUTHOR INFORMATION Author notes * These authors contributed equally: Edward J. Thomas, Virginia N. Ciriano-Tejel. AUTHORS AND AFFILIATIONS * Quantum
Motion, London, UK Edward J. Thomas, Virginia N. Ciriano-Tejel, David F. Wise, Domenic Prete, Mathieu de Kruijf, David J. Ibberson, Grayson M. Noah, Alberto Gomez-Saiz, M. Fernando
Gonzalez-Zalba, Mark A. I. Johnson & John J. L. Morton * Department of Electronic and Electrical Engineering, UCL, London, UK Edward J. Thomas & John J. L. Morton * London Centre for
Nanotechnology, UCL, London, UK Mathieu de Kruijf & John J. L. Morton Authors * Edward J. Thomas View author publications You can also search for this author inPubMed Google Scholar *
Virginia N. Ciriano-Tejel View author publications You can also search for this author inPubMed Google Scholar * David F. Wise View author publications You can also search for this author
inPubMed Google Scholar * Domenic Prete View author publications You can also search for this author inPubMed Google Scholar * Mathieu de Kruijf View author publications You can also search
for this author inPubMed Google Scholar * David J. Ibberson View author publications You can also search for this author inPubMed Google Scholar * Grayson M. Noah View author publications
You can also search for this author inPubMed Google Scholar * Alberto Gomez-Saiz View author publications You can also search for this author inPubMed Google Scholar * M. Fernando
Gonzalez-Zalba View author publications You can also search for this author inPubMed Google Scholar * Mark A. I. Johnson View author publications You can also search for this author inPubMed
Google Scholar * John J. L. Morton View author publications You can also search for this author inPubMed Google Scholar CONTRIBUTIONS E.J.T., V.N.C.-T. and D.P. performed the experiments
under the supervision of M.A.I.J. and M.F.G.-Z. M.d.K. and D.J.I. contributed to the preparation of the experiment and preliminary characterization. M.F.G.-Z. and J.J.L.M. conceived the
experiment. A.G.-S. designed the device with contributions from M.F.G.-Z. and J.J.L.M. The device analogue components were validated by G.M.N. D.F.W. and V.N.C.-T. designed and developed the
automated analysis tools. Further data analysis was performed by E.J.T., V.N.C.-T., D.F.W. and M.A.I.J. The manuscript was prepared by M.A.I.J., V.N.C.-T., D.F.W. and M.F.G.-Z., with
contributions from all authors. CORRESPONDING AUTHOR Correspondence to Mark A. I. Johnson. ETHICS DECLARATIONS COMPETING INTERESTS The authors declare no competing interests. PEER REVIEW
PEER REVIEW INFORMATION _Nature Electronics_ thanks the anonymous reviewers for their contribution to the peer review of this work. ADDITIONAL INFORMATION PUBLISHER’S NOTE Springer Nature
remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. EXTENDED DATA EXTENDED DATA FIG. 1 PARAMETER VARIATION WITH BACK-GATE VOLTAGE. A-C,
Estimated distributions of first observed electron loading voltage, gate lever arm and QD asymmetry as back-gate voltage is varied (L = 28 nm). As backgate increases the dc and rf datasets
have 206, 207, 178, 170, 175 and 176 data points. The box plot features are as follows: centre line, median, box edges 25th and 75th percentile, whiskers, minimum and maximum of the dataset
excluding outliers more than 1.5 × outside the interquartile range. EXTENDED DATA FIG. 2 PARAMETER EXTRACTION PERFORMANCE WITH INCREASING INTEGRATION TIME. A-C, Distributions of extracted
parameters (first detected electron loading voltage, gate lever arm, and source-drain asymmetry) using swarm plots (left) and a box plot (right), as the number of averages is varied in rf
measurements. The swarm plot reveals individual data points’ distribution, while the box plot displays central tendencies and quartiles for easy comparison and outlier detection. As
expected, we observe a decrease in the parameter standard deviation as averages increase. Each group contains 206 samples after removing outliers which are more than 1.5 × outside the
interquartile range. The box plot features are as follows: centre line, median, box edges 25th and 75th percentile, whiskers, minimum and maximum of the dataset excluding outliers more than
1.5x outside the interquartile range. SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION Supplementary Sections I–IX, Figs. 1–7 and Discussion. RIGHTS AND PERMISSIONS OPEN ACCESS This
article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as
you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party
material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s
Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. Reprints and permissions ABOUT THIS ARTICLE CITE THIS ARTICLE Thomas, E.J., Ciriano-Tejel, V.N., Wise, D.F.
_et al._ Rapid cryogenic characterization of 1,024 integrated silicon quantum dot devices. _Nat Electron_ 8, 75–83 (2025). https://doi.org/10.1038/s41928-024-01304-y Download citation *
Received: 13 February 2024 * Accepted: 05 November 2024 * Published: 03 January 2025 * Issue Date: January 2025 * DOI: https://doi.org/10.1038/s41928-024-01304-y SHARE THIS ARTICLE Anyone
you share the following link with will be able to read this content: Get shareable link Sorry, a shareable link is not currently available for this article. Copy to clipboard Provided by the
Springer Nature SharedIt content-sharing initiative