Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Noise reduction by bias cooling in gated \chSi/\chSi_x Ge_1-x quantum dots

Julian Ferrero Physikalisches Institut, Karlsruhe Institute of Technology, Karlsruhe, Germany    Thomas Koch Physikalisches Institut, Karlsruhe Institute of Technology, Karlsruhe, Germany    Sonja Vogel Physikalisches Institut, Karlsruhe Institute of Technology, Karlsruhe, Germany    Daniel Schroller Physikalisches Institut, Karlsruhe Institute of Technology, Karlsruhe, Germany    Viktor Adam Physikalisches Institut, Karlsruhe Institute of Technology, Karlsruhe, Germany Institute for Quantum Materials and Technologies, Karlsruhe Institute of Technology, Karlsruhe, Germany    Ran Xue JARA-FIT Institute for Quantum Information, Forschungszentrum Jülich GmbH and RWTH Aachen University, Aachen, Germany    Inga Seidler JARA-FIT Institute for Quantum Information, Forschungszentrum Jülich GmbH and RWTH Aachen University, Aachen, Germany    Lars R. Schreiber JARA-FIT Institute for Quantum Information, Forschungszentrum Jülich GmbH and RWTH Aachen University, Aachen, Germany    Hendrik Bluhm JARA-FIT Institute for Quantum Information, Forschungszentrum Jülich GmbH and RWTH Aachen University, Aachen, Germany    Wolfgang Wernsdorfer wolfgang.wernsdorfer@kit.edu Physikalisches Institut, Karlsruhe Institute of Technology, Karlsruhe, Germany Institute for Quantum Materials and Technologies, Karlsruhe Institute of Technology, Karlsruhe, Germany
(May 8, 2024)
Abstract

Silicon-Germanium heterostructures are a promising quantum circuit platform, but crucial aspects as the long-term charge dynamics and cooldown-to-cooldown variations are still widely unexplored quantitatively. In this letter we present the results of an extensive bias cooling study performed on gated silicon-germanium quantum dots with an \chAl2O3-dielectric. Over 80 cooldowns were performed in the course of our investigations. The performance of the devices is assessed by low-frequency charge noise measurements in the band of 200 µHztimes200microhertz200\text{\,}\mathrm{\SIUnitSymbolMicro Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_Hz end_ARG to 10 mHztimes10millihertz10\text{\,}\mathrm{mHz}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_mHz end_ARG. We measure the total noise power as a function of the applied voltage during cooldown in four different devices and find a minimum in noise at 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias cooling voltage for all observed samples. We manage to decrease the total noise power median by a factor of 6 and compute a reduced tunneling current density using Schrödinger-Poisson simulations. Furthermore, we show the variation in noise from the same device in the course of eleven different cooldowns performed under the nominally same conditions.

preprint: AIP/123-QED

Electron spins in silicon-germanium represent a promising implementation of the solid state quantum computer. Single and two-qubit gates have been reported to reach gate fidelities above the error correction threshold Yoneda et al. (2018); Noiri et al. (2022); Xue et al. (2024). Recently, quantum error correction was performed on a three qubit device Takeda et al. (2022) and intermediate range coupling opened new prospects regarding a scaled up quantum processorSeidler et al. (2022); Langrock et al. (2023); Xue et al. (2024); Künne et al. (2023). Since quantum-dot based qubits need to be tuned and re-tuned, the long term stability of the qubit working point has gained research interest Kranz et al. (2020); Struck et al. (2020); Connors et al. (2022). Furthermore it has been shown that the charge noise level in \chSiGe devices is strongly dependent on the applied global field Takeda et al. (2013). Bias cooling is a readily accessible method which only relies on equipment already present in the typical semiconductor quantum dot setup. To perform bias cooling, the same bias-cooling-voltage VBCsubscriptVBC\mathrm{V_{BC}}roman_V start_POSTSUBSCRIPT roman_BC end_POSTSUBSCRIPT is applied to all gates of the sample during cooldown from room-temperature to mKmillikelvin\mathrm{mK}roman_mK temperatures. This causes charges to be trapped in localized defects which interact with the global electric field, thereby changing the stability of gate defined quantum dots. In gallium arsenide, bias cooling has been proven to reduce switching noise and therefore improving the sample stability Pioro-Ladrière et al. (2005), by filling DX centers in the dopant layer and therefore reducing the effective leakage rate of electrons from the gate layers into the two-dimensional-electron-gas (2DEG). In silicon-germanium, no such centers are present and bias cooling has not been quantitatively investigated yet.

We present the results of a systematic bias-cooling study, quantitatively investigating the noise power in the voltage range in between 1 Vtimes-1volt-1\text{\,}\mathrm{V}start_ARG - 1 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG and 1 Vtimes1volt1\text{\,}\mathrm{V}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG for VBCsubscriptVBC\mathrm{V_{BC}}roman_V start_POSTSUBSCRIPT roman_BC end_POSTSUBSCRIPT. More then 80 cooldowns were performed, on four different samples. Depending on the applied gate voltage during cooldown, we reduce the total noise power in the frequency band of 200 µHztimes200microhertz200\text{\,}\mathrm{\SIUnitSymbolMicro Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_Hz end_ARG to 10 mHztimes10millihertz10\text{\,}\mathrm{mHz}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_mHz end_ARG by a factor of 6, in comparison to the zero volt case.

Refer to caption
Figure 1: (a) False colored scanning electron micrograph. Bottommost gate layer in blue, topmost layer in green. The path of the current ImeassubscriptImeas\mathrm{I_{meas}}roman_I start_POSTSUBSCRIPT roman_meas end_POSTSUBSCRIPT used in the measurements is denoted in red. (b) Schematic cross-section with layer thicknesses to scale. The 2DEG is defined in the 10 nmtimes10nanometer10\text{\,}\mathrm{nm}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG strained Si layer. The SiGe spacer layer is protected from oxidation by a 2 nmtimes2nanometer2\text{\,}\mathrm{nm}start_ARG 2 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG silicon cap layer. The dashed line in (a) represents the corresponding position of the schematic cross-section in (b).

We use four gated Si/SiGe devices nominally equal to the device used in Seidler et al. (2022); Struck et al. (2024) (see Fig. 1 (a)). It contains two single-electron transistors (SETs) operating as proximal charge detectors and nine central finger gates as well as two confinement gates.

The gatestack is fabricated upon a Si/SiGe heterostructure, which consists of a 1000 nmtimes1000nanometer1000\text{\,}\mathrm{nm}start_ARG 1000 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG \chSi_0.7Ge_0.3 graded buffer layer on a Si substrate, a 10 nmtimes10nanometer10\text{\,}\mathrm{nm}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG SiSi\mathrm{Si}roman_Si channel, followed by a 30 nmtimes30nanometer30\text{\,}\mathrm{nm}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG \chSi_0.7Ge_0.3 spacer layer and capped by 2 nmtimes2nanometer2\text{\,}\mathrm{nm}start_ARG 2 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG naturally oxidized Si (see Fig.1 (b)). The heterostructure confines electrons in the growth direction, effectively forming a 2DEG in the silicon channel, when accumulated. Ohmic contacts to the quantum well layer are realized by implantation of phosphorus and activated by rapid thermal processing at 700 CsuperscriptC{}^{\circ}\mathrm{C}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT roman_C for 15 stimes15second15\text{\,}\mathrm{s}start_ARG 15 end_ARG start_ARG times end_ARG start_ARG roman_s end_ARG. The gate stack is an overlay of three metal layers consisting of 15 nmtimes15nanometer15\text{\,}\mathrm{nm}start_ARG 15 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG, 22 nmtimes22nanometer22\text{\,}\mathrm{nm}start_ARG 22 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG and 29 nmtimes29nanometer29\text{\,}\mathrm{nm}start_ARG 29 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG Pt on top of a 5 nmtimes5nanometer5\text{\,}\mathrm{nm}start_ARG 5 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG Ti adhesion layer. The metal layers are electrically insulated from the substrate and from each other by 10 nmtimes10nanometer10\text{\,}\mathrm{nm}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG atomic layer deposited \chAl2O3.

To characterize the performance under different bias cooling conditions, an efficient thermal cycling mechanism is needed. A heater consisting of constantan wire was installed on the mixing chamber of the used dilution refrigerator. The sample is heated locally to 300 Ktimes300kelvin300\text{\,}\mathrm{K}start_ARG 300 end_ARG start_ARG times end_ARG start_ARG roman_K end_ARG, while the 4 K stage of the fridge is kept below 10 Ktimes10kelvin10\text{\,}\mathrm{K}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_K end_ARG. This is achieved by thermally insulating the mixing chamber plate by stopping the circulation of the mixture while a high flow of liquid helium is supplied to the 4K stage. An automated thermal cycle from 30 mKtimes30millikelvin30\text{\,}\mathrm{mK}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_mK end_ARG to 300 Ktimes300kelvin300\text{\,}\mathrm{K}start_ARG 300 end_ARG start_ARG times end_ARG start_ARG roman_K end_ARG and back to 30 mKtimes30millikelvin30\text{\,}\mathrm{mK}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_mK end_ARG takes 3.5 htimes3.5hour3.5\text{\,}\mathrm{h}start_ARG 3.5 end_ARG start_ARG times end_ARG start_ARG roman_h end_ARG.

We monitor the turn-on voltage VTsubscriptVT\mathrm{V_{T}}roman_V start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT after reaching base temperature. To do this, we accumulate the electron gas using all gates, effectively forming a conductive sheet underneath the gate structure. This way, we minimize the effect of singular defects on the conducting channel. We measure the conductance GG\mathrm{G}roman_G in between the ohmic contacts of the left SET. Once it reaches one third of the saturation value (Gsat/3subscriptGsat3\mathrm{G_{sat}/3}roman_G start_POSTSUBSCRIPT roman_sat end_POSTSUBSCRIPT / 3), which was measured in a separate cooldown, the accumulation voltage is not increased further. The first apparent effect of bias cooling is that the accumulation voltage of the device shifts nearly linearly with the applied bias cooling voltage (Fig. 2). Charge carriers are trapped in between the quantum well and the metal gates. We suspect electrons getting caught at the Si-\chAl2O3 interface Hoex et al. (2008); Gielis et al. (2008). This interface is known to form an \chSiO_x layer, in which silicon atoms are partially replaced by aluminium atoms Hiller et al. (2019, 2021). These form acceptor states, which trap electrons. By applying the bias cooling voltage VBCsubscriptVBC\mathrm{V_{BC}}roman_V start_POSTSUBSCRIPT roman_BC end_POSTSUBSCRIPT during cooldown, we change the electrochemical potential of the defects, allowing thermally excited electrons to change the population of the defect states. Applying positive voltages attracts additional electrons into acceptor-states, increasing the turn-on voltage. Negative biases reduce the population of the \chAl-vacancies below its equilibrium, creating an excess of positively charged defects, decreasing the accumulation voltage.

Refer to caption
Figure 2: Turn-on voltage of the device over applied bias voltage during cooldown for four nominally equal devices. In between 1 Vtimes-1volt-1\text{\,}\mathrm{V}start_ARG - 1 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG and 1 Vtimes1volt1\text{\,}\mathrm{V}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG all samples behave linearly showing similar slopes. Below 1 Vtimes-1volt-1\text{\,}\mathrm{V}start_ARG - 1 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG the behavior differs. This might be due to different defect densities in the heterostructure. The inset shows an exemplary accumulation curve. The red-dashed line indicates the turn-on voltage, which is defined by the channel conductivity GG\mathrm{G}roman_G crossing a value of 0.4 µStimes0.4microsiemens0.4\text{\,}\mathrm{\SIUnitSymbolMicro S}start_ARG 0.4 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_S end_ARG.

Next, we characterize the noise power for each bias voltage. To investigate the charge noise we form a single electron transistor in between barrier gates (e.g. LB1 and LB2). The gate layout of the used device is shown in Fig. 1 (b). The challenge hereby is tuning the SETs to comparable working points, although the sample provides a different electrostatic environment in each cooldown. Noise seen by SET devices depends on their working point Elsayed et al. (2022). Therefore it is essential to repeat a fixed measurement protocol. Our protocol consists of six steps summarized in Fig. 3 (a) and (b). After accumulating the sample to the reference conductance of Gsat/3subscriptGsat3\mathrm{G_{sat}/3}roman_G start_POSTSUBSCRIPT roman_sat end_POSTSUBSCRIPT / 3 (step 1), we define the accumulation voltage Vrefsubscript𝑉refV_{\mathrm{ref}}italic_V start_POSTSUBSCRIPT roman_ref end_POSTSUBSCRIPT which is used as a starting point for the device tuning. With all gates at VrefsubscriptVref\mathrm{V_{ref}}roman_V start_POSTSUBSCRIPT roman_ref end_POSTSUBSCRIPT the device is in a state where a 2DEG is accumulated in the silicon quantum well below every metal gate. In step 2, we lower the voltages of all gates except the topgates and barriers of the SETs by 500 mVtimes500millivolt500\text{\,}\mathrm{mV}start_ARG 500 end_ARG start_ARG times end_ARG start_ARG roman_mV end_ARG to deplete the sample and confine the 2DEG only below the top gates. In the first iteration of step 3, the top gate and barrier voltages are increased until a third of the saturation conductance is reached (not shown in Fig 3(b)). For every subsequent iteration of step 3, only the top gate is raised until accumulation (Fig 3 (b)). In step 3, the topgate voltage (and in the first iteration also the barrier voltages, not shown in the voltage chronogram in Fig. 3 (b)) are increased until a third of the saturation conductance is reached in the measured channel. In step 4 the ’cutoff’ voltage of the barrier gates is determined by lowering the voltages of the barrier gates to the point where the conducting channels are fully confined and the measured current is cut off. Slightly above this cutoff voltage, we perform a 100x100 mV2times100millivolt2100\text{\,}{\mathrm{mV}}^{2}start_ARG 100 end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_mV end_ARG start_ARG 2 end_ARG end_ARG sweep downwards with both barrier gates. We record the current through the channel to see whether Coulomb oscillations are present. If not we now lower the barrier gates by another 100 mVtimes100millivolt100\text{\,}\mathrm{mV}start_ARG 100 end_ARG start_ARG times end_ARG start_ARG roman_mV end_ARG and re-accumulate the conducting channel with the top gate again up to a conductance of Gsat/3subscriptGsat3\mathrm{G_{sat}/3}roman_G start_POSTSUBSCRIPT roman_sat end_POSTSUBSCRIPT / 3. By repeating steps 3 and 4 iteratively we find the lowest (granulated by the resolution of our voltage steps) topgate voltage that allows formation of quantum dots, which are identified by a 2D barrier sweep exhibiting the Coulomb oscillations (step 5). The barrier gates are tuned to the first Coulomb peak (step 6) and a plunger trace is recorded. If the recorded trace shows a secant-like Coulomb peak, the SET has been formed and the tuning routine came to a successful conclusion.

Refer to caption
Figure 3: (a) Tuning workflow which was performed after each cooldown. The red square in step 2 denotes where the cutout shown in steps 3 to 5 lies within sample. The sample is accumulated to the reference conductance of Gsat/3subscriptGsat3\mathrm{G_{sat}/3}roman_G start_POSTSUBSCRIPT roman_sat end_POSTSUBSCRIPT / 3 (step 1). The voltages applied to all but the SET-gates are reduced by 500 mVtimes500millivolt500\text{\,}\mathrm{mV}start_ARG 500 end_ARG start_ARG times end_ARG start_ARG roman_mV end_ARG to confine the 2DEG (step 2). Using the SET gates a closed conducting channel is formed (step 3). A 100x100 mV2times100millivolt2100\text{\,}{\mathrm{mV}}^{2}start_ARG 100 end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_mV end_ARG start_ARG 2 end_ARG end_ARG sweep is performed with both barrier gates (step 4). Step 3 and 4 are repeated iteratively until a Coulomb oscillation is observed (step 5). The barrier gates are tuned to the first Coulomb peak (step 6) and a plunger trace is recorded. (b) Simplified chronogram of the voltages applied to the SET-gates. Numbers 1-6 correspond to the tuning steps in panel (a).

With the quantum dot formed, we characterize its noise in the band of 200 µHztimes200microhertz200\text{\,}\mathrm{\SIUnitSymbolMicro Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_Hz end_ARG to 10 mHztimes10millihertz10\text{\,}\mathrm{mHz}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_mHz end_ARG using peaktrackingKranz et al. (2020). The lower bound on the measured frequencies is set by the total length of our peaktracking measurements (5 htimes5hour5\text{\,}\mathrm{h}start_ARG 5 end_ARG start_ARG times end_ARG start_ARG roman_h end_ARG) and the upper bound is set by the rate at which the plunger gate can be swept to record the position of the peak in gatespace. The number of points recorded per plunger sweep varies depending on the expected drift of the sample. Therefore, a trace can contain 150 to 250 individual data points and be measured in 30 stimes30second30\text{\,}\mathrm{s}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_s end_ARG to 70 stimes70second70\text{\,}\mathrm{s}start_ARG 70 end_ARG start_ARG times end_ARG start_ARG roman_s end_ARG depending on the number and sampling rate of the points. We continuously sweep the plunger and record the conductivity of the SET device (Fig. 4 (a)).

Refer to caption
Figure 4: (a) Peaktracking measurement, performed on sample A with a VBCsubscriptVBC\mathrm{V_{BC}}roman_V start_POSTSUBSCRIPT roman_BC end_POSTSUBSCRIPT of 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. The colorplot consists of Coulomb peak traces recorded back to back. The red line indicates the position of the peak maxima. b) Noise power spectral density, fitted with a β/fα𝛽superscript𝑓𝛼\beta/f^{\alpha}italic_β / italic_f start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT power spectrum. To calculate the total noise power, the spectrum is integrated over the highlighted frequency range. The conversion of VSETsubscriptVSET\mathrm{V_{SET}}roman_V start_POSTSUBSCRIPT roman_SET end_POSTSUBSCRIPT to the chemical potential of the SET is done by a static leverarm of 0.039 eV/V.
Refer to caption
Figure 5: (a) - (c) Peaktracking measurements performed on Sample B for 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG, 0.625 Vtimes0.625volt0.625\text{\,}\mathrm{V}start_ARG 0.625 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG and 0.825 Vtimes0.825volt0.825\text{\,}\mathrm{V}start_ARG 0.825 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. (b) Shows a vanishing side peak, which might indicate the presence of a parasitic dot. (d) Noise power spectral densities fitted with a β/fα𝛽superscript𝑓𝛼\beta/f^{\alpha}italic_β / italic_f start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT power spectrum. e) Integrated noise power versus bias cooling voltage. While the noise increases towards negative biases for most samples, a local minimum at 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG is found in each trace. The dashed line denotes the 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG noise level. (f) Boxplot of the integrated noise power of 22 peaktrackings performed in alternating 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG/0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG cooldowns. The box includes the interquartile (IQR) range and the whiskers extend up to 1.5 times the IQR.The orange solid line displays the median value. The median value of 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG is almost an order of magnitude lower than for 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG.

The shown result was recorded on sample A after being cooled down with 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. We extract the exact peak position by fitting each individual trace with a secant function and define the position of its maximum as the peak position. The conversion of VSETsubscriptVSET\mathrm{V_{SET}}roman_V start_POSTSUBSCRIPT roman_SET end_POSTSUBSCRIPT to eVelectronvolt\mathrm{eV}roman_eV is done by a static lever arm of 0.039 eV/V. This value is the average of multiple extracted lever arms measured (using Coulomb diamond measurements Hanson et al. (2007)) for different samples and bias cooling voltages spanning a range from 0.0350.0350.0350.035 to 0.040 eV/V.

Next, we perform a Welch-estimation of the power-spectral-density using the standard signal.welch method from the scipy python package. To exclude artificial frequencies arising from drift in between the start and the end point of a trace, we additionally employ a Hann-type windowing function. The estimation is always performed on the first measured peak. The result can be found in Fig. 4 (b). The resulting spectrum follows a 1/fα1superscript𝑓𝛼1/f^{\alpha}1 / italic_f start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT-noise distribution with α=1.57𝛼1.57\alpha=~{}1.57italic_α = 1.57. This lies in between α1𝛼1\alpha\approx 1italic_α ≈ 1, suggesting the presence of many two level fluctuators with a broad distribution of time scales Paladino et al. (2014), and α2𝛼2\alpha\approx 2italic_α ≈ 2, indicating the presence of random walk noise Barnes and Allan (1966). Following, we integrate the fitted spectra in the band of 200 µHztimes200microhertz200\text{\,}\mathrm{\SIUnitSymbolMicro Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_Hz end_ARG to 10 mHztimes10millihertz10\text{\,}\mathrm{mHz}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_mHz end_ARG, highlighted in purple in Fig. 4 (b). Quoting the total noise power instead of the e.g. 1 mHztimes1millihertz1\text{\,}\mathrm{mHz}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_mHz end_ARG-noise has the advantage that all frequencies in the measured band contribute to the noise value. This metric is chosen because we want to weigh both, the rarely occurring large jumps as well as the small displacments in peak position that happen in between every single datapoint.

The procedure is applied to all measured peaktracks. A selection measured on sample B is found in Fig. 5 (a) to (c). Peaktracks dominated by large jumps tend towards α2𝛼2\alpha\approx 2italic_α ≈ 2, while peaktracks which are dominated by fluctuations around the original working point tend towards α1𝛼1\alpha\approx 1italic_α ≈ 1. Resulting noise spectra are shown in Fig. 5 (d). The noise power versus bias cooling voltage can be seen in Fig. 5 (e). Four samples show a significant noise reduction at 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. Sample B shows an improvement in the integrated noise power by two orders of magnitude. Towards more positive voltages, samples B and C show an increase in noise.

We identify two types of noise on the SET peak position. A high frequency, low amplitude fluctuation and rarely occurring jumps with a large amplitude. These jumps occur on the timescale of hours and drastically affect the noise performance (Fig. 5). Since the individual peak-tracking measurements have a length of five hours, they may not lead to statistically significant data. To gain insight into the significance of our observations, we performed a measurement campaign consisting of 22 cooldowns of sample A. We measured eleven times the zerobias followed by the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias. The results of the noise power integrated from 200 µHztimes200microhertz200\text{\,}\mathrm{\SIUnitSymbolMicro Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_Hz end_ARG to 10 mHztimes10millihertz10\text{\,}\mathrm{mHz}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_mHz end_ARG are shown in the Fig. 5 (f). The orange line shows the median. The median value of the 0.70.70.70.7 V measurements is reduced by a factor of 6 in comparison to the median of the 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG measurements. To determine the statistical significance of the measured results, we performed an unpaired t-test comparing the measured data sets for 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG and 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG, which results in a probability of 0.143 %times0.143percent0.143\text{\,}\mathrm{\char 37\relax}start_ARG 0.143 end_ARG start_ARG times end_ARG start_ARG % end_ARG that the two measured sample means arise from the same normal distribution.

Three-dimensional Poisson-Schrödinger-simulations have been performed on the SET-area of the investigated devices. \chSiGe-heterostacks are known to induce tunneling currents from the \chSi-channelHuang et al. (2014) into the cap. Charge redistributions, as seen in Fig. 5 (a), could be caused by local metal-to-insulator-transitions, triggered by tunneling into the cap Huang et al. (2014); Huang (2015). This interpretation is plausible, since the cap-oxide interface itself is MOSFET-like. In MOSFET devices the lengthscales of localization length versus potential variation Wilamowski et al. (2001) are known to create a percolation induced metal-to-insulator-transition Tracy et al. (2009). In that case the 2DEG in the cap breaks down into charge-carrier-puddles containing mobile carriers. With the simulations we wish to verify that tunneling currents into the cap exist and can therefore source electrons which cause the charge redistributions. Bias cooling was included in simulation by placing interface-charges in the interstitial silicon-oxide layer under the metal gates. The interface-charge density under each gate-layer was calculated based on the turn-on voltage shift. We simulated the density needed to shift the turn-on voltage of gate layer nn\mathrm{n}roman_n (with n{1,2,3}n123\mathrm{n}\in\{1,2,3\}roman_n ∈ { 1 , 2 , 3 }) individually. The same charge density is assumed under each gate of a specific gate layer. To reflect the device’s real working point, the gate-voltage configurations from individual measurements have been reproduced in simulation. The spatially large regions are treated (substrate, buffer, spacer) in Thomas-Fermi Thomas (1927); Fermi (1928); Tang, O’Regan, and Wu (2004) approximation. The regions where quantum confinement plays a major role (channel and cap),on the other hand, are treated with a self-consistent Schrödinger-Poisson approach in effective mass approximation Kane (1959); Trellakis, Andlauer, and Vogl (2006). In Fig. 6, the simulated band diagram in growth direction for a sample cooled down with VBC=0.7 VsubscriptVBCtimes0.7volt\mathrm{V_{BC}=$0.7\text{\,}\mathrm{V}$}roman_V start_POSTSUBSCRIPT roman_BC end_POSTSUBSCRIPT = start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG is shown. The barrier in between the silicon channel and the silicon cap-layer is of triangular shape, meaning that the tunneling currents can be calculated using the Fowler-Nordheim Fowler and Nordheim (1928) model. With the resulting electric field and electrochemical potential, we calculate the tunneling rates from the SiGe-channel into the silicon cap. The spacially varying tunneling currents are shown in Fig. 7 (a) for the 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG and (b) for the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias cooling case. First, it is important to note that the maximum tunneling rate for the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG case is reduced by seven orders of magnitude in comparison to the 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG case. Furthermore, the dot-region itself does not act as a source of electrons in the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG case, meaning the number of tunneling electrons is not only reduced, but the tunneling events are mainly taking place further away from the region of interest. Generally, the bias cooled samples are operated at lower internal electric fields, as seen in Fig. 7 (c) and (d).

Refer to caption
Figure 6: Simulated band structure alignment at the working point of a sample cooled down with VBC=0.7 VsubscriptVBCtimes0.7volt\mathrm{V_{BC}=$0.7\text{\,}\mathrm{V}$}roman_V start_POSTSUBSCRIPT roman_BC end_POSTSUBSCRIPT = start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. EΔ3subscriptEΔ3\mathrm{E_{\Delta 3}}roman_E start_POSTSUBSCRIPT roman_Δ 3 end_POSTSUBSCRIPT is the energy of the lowest conduction band. The violet region indicates the assumed position of the interface charges. They were distributed in a 1 nmtimes1nanometer1\text{\,}\mathrm{nm}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG thick sheet located 1.2 nmtimes1.2nanometer1.2\text{\,}\mathrm{nm}start_ARG 1.2 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG above the Si-cap-oxide interface. The orange arrow indicates the direction of the Fowler-Nordheim tunneling current density JFNsubscriptJFN\mathrm{J_{FN}}roman_J start_POSTSUBSCRIPT roman_FN end_POSTSUBSCRIPT. The electron density n is plotted in blue.
Refer to caption
Figure 7: a) Simulated tunneling current from silicon channel into the silicon cap for a sample cooled with a 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias (a) and 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias (b). Simulated electric field distribution at the channel-spacer interface of a sample cooled with a 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias (c) and 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias (d).

In the bias-cooled-case the electric field difference in between the dot region and the barriers is larger. This means that a dot can be accumulated at lower accumulation gate fields, since the electric field of the barriers is more sharply defined due to the frozen-in charges, which are located at the Si-cap-\chAl2O3 interface.

We repeated the simulation for each working point in Fig. 5 (f), to verify an interrelation in between tunneling events and noise. In Fig. 8 the measured noise power is plotted versus the calculated maximum tunneling current. The 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG cases show a bunching towards high tunneling rates and high noise, whereas the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG cases show bunching towards low tunneling and low noise. One possible explanation for the correlation of high noise and high tunneling rates is that the electrons flowing from the channel into the cap cause local metal-insulator transitions in the disordered silicon cap Huang et al. (2014). The avalanche-like charge redistributions seen in the peaktracking measurements might be caused by a few excess electrons, locally exceeding the percolation density, triggering a large charge transfer, which is supported by the fact that the number of large charge redistributions is reduced in the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG dataset. At high bias cooling voltage, a high number of the acceptor defects located at the Si-cap \chAl2O3 interface are charged. We propose two main mechanisms to explain the increase in noise for voltages above 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. First, the high density of negatively charged defects leads to working points with high electric fields. Additional tunneling could be a result. Furthermore, the increase in charged defects at the interface could lead to the individual defects exchanging charges, therefore effectively raising the noise again. For bias cooling voltages exceeding 2 Vtimes2volt2\text{\,}\mathrm{V}start_ARG 2 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG the 2DEG is accumulated without the application of an accumulation voltage. Here, the trapped charges mimic the role of a dopant. For more negative voltages the noise is increasing in the case of sample A and D, not showing a global trend.

In conclusion, we found that bias cooling of undoped \chSi/\chSi_x Ge_1-x heterostacks causes charges to be trapped in between the silicon channel and the metal gates. This shifts the turn-on voltages linearly with the applied cooldown bias in the range of 1 Vtimes-1volt-1\text{\,}\mathrm{V}start_ARG - 1 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG to 1 Vtimes1volt1\text{\,}\mathrm{V}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. We investigated the low frequency charge noise of bias cooled devices. For this we used peaktracking measurements. To quantify the noise for each bias cooling voltage we computed the total noise power for every peak track, which has a minimum around 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias cooling voltage. In samples A and B, a global minimum is visible in the total noise power. Samples C and D show a local minimum around 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG. In case of Sample B we could reduce the total noise power by a factor of 120 during the 0.75 Vtimes0.75volt0.75\text{\,}\mathrm{V}start_ARG 0.75 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG cooldown in comparison to the 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG cooldown, and during a 22-cooldown campaign the measured median of the total noise power was reduced by a factor of 6 for the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG cooldown in comparison to the 0 Vtimes0volt0\text{\,}\mathrm{V}start_ARG 0 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG value. Our measurements show that a significant variation of the noise level can occur as a function of cooldown-bias and that this variation is not a random fluctuation from cooldown to cooldown. While a cooldown bias of 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG leads to good results on all four devices considered, more statistics would be needed to tell if this value reflects a device-independent optimum, and how the optimal cooldown strategy depends on the device design. Another important implication of the significant variation between cooldowns and bias voltages is that a lot of statistics is needed to draw reliable conclusions from noise measurements. In addition, we present the results of a three-dimensional Schrödinger-Poisson simulation, based on measured working points. Here, we find that samples cooled down with a 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG bias show a by seven orders of magnitude reduced tunneling current from the channel into the cap. While the direct proof remains elusive, we correlate the simulation to our noise measurement results, and find a bunching of datapoints in the high noise, high tunneling as well as in the low noise, low tunneling quadrants. As a next step, bias cooling could be extended to qubit samples, investigating the effect on coherence. Furthermore, tunneling in the cap has been proposed as one of the root causes of the quantum dot instabilities.

This work has been funded by the German Research Foundation (DFG) under Germany’s Excellence Strategy - Cluster of Excellence Matter and Light for Quantum Computing" (ML4Q) EXC 2004/1 – 390534769 and the Gottfried Wilhelm Leibniz-Award, ZVN-2020_WE 4458-5. Project Si-QuBus received funding from the QuantERA ERA-NET Cofund in Quantum Technologies implemented within the European Union’s Horizon 2020 Programme. The device fabrication has been done at HNF - Helmholtz Nano Facility, Research Center Juelich GmbH Albrecht, Moers, and Hermanns (2017). Furthermore the authors wish to thank Malte Neul for the organization of our meetings and for helpful discussions.

Refer to caption
Figure 8: Measured integrated noise density versus simulated peak tunneling current. The datapoint belonging to the 0.7 Vtimes0.7volt0.7\text{\,}\mathrm{V}start_ARG 0.7 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG dataset in the high-tunneling, high-noise quadrant shows an excess of telegraph noise which was only recorded once.

References

  • Yoneda et al. (2018) J. Yoneda, K. Takeda, T. Otsuka, T. Nakajima, M. R. Delbecq, G. Allison, T. Honda, T. Kodera, S. Oda, Y. Hoshi, N. Usami, K. M. Itoh,  and S. Tarucha, “A quantum-dot spin qubit with coherence limited by charge noise and fidelity higher than 99.9%,” Nature Nanotechnology 13, 102–106 (2018).
  • Noiri et al. (2022) A. Noiri, K. Takeda, T. Nakajima, T. Kobayashi, A. Sammak, G. Scappucci,  and S. Tarucha, “Fast universal quantum gate above the fault-tolerance threshold in silicon,” Nature 601, 338–342 (2022).
  • Xue et al. (2024) R. Xue, M. Beer, I. Seidler, S. Humpohl, J.-S. Tu, S. Trellenkamp, T. Struck, H. Bluhm,  and L. R. Schreiber, “Si/sige qubus for single electron information-processing devices with memory and micron-scale connectivity function,” Nature Communications 15, 2296 (2024).
  • Takeda et al. (2022) K. Takeda, A. Noiri, T. Nakajima, T. Kobayashi,  and S. Tarucha, “Quantum error correction with silicon spin qubits,” Nature 608, 682–686 (2022).
  • Seidler et al. (2022) I. Seidler, T. Struck, R. Xue, N. Focke, S. Trellenkamp, H. Bluhm,  and L. R. Schreiber, “Conveyor-mode single-electron shuttling in si/sige for a scalable quantum computing architecture,” npj Quantum Information 8, 100 (2022).
  • Langrock et al. (2023) V. Langrock, J. A. Krzywda, N. Focke, I. Seidler, L. R. Schreiber,  and L. Cywiński, “Blueprint of a scalable spin qubit shuttle device for coherent mid-range qubit transfer in disordered si/sige/sio2subscriptsi/sige/sio2{\text{si/sige/sio}}_{2}si/sige/sio start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT,” PRX Quantum 4, 020305 (2023).
  • Künne et al. (2023) M. Künne, A. Willmes, M. Oberländer, C. Gorjaew, J. D. Teske, H. Bhardwaj, M. Beer, E. Kammerloher, R. Otten, I. Seidler, R. Xue, L. R. Schreiber,  and H. Bluhm, “The spinbus architecture: Scaling spin qubits with electron shuttling,”   (2023), arXiv:2306.16348 [quant-ph] .
  • Kranz et al. (2020) L. Kranz, S. K. Gorman, B. Thorgrimsson, Y. He, D. Keith, J. G. Keizer,  and M. Y. Simmons, “Exploiting a single-crystal environment to minimize the charge noise on qubits in silicon,” Adv. Mater. 32, 2003361 (2020).
  • Struck et al. (2020) T. Struck, A. Hollmann, F. Schauer, O. Fedorets, A. Schmidbauer, K. Sawano, H. Riemann, N. V. Abrosimov, L. Cywiński, D. Bougeard,  and L. R. Schreiber, “Low-frequency spin qubit energy splitting noise in highly purified 28si/sige,” npj Quantum Information 6, 40 (2020).
  • Connors et al. (2022) E. J. Connors, J. Nelson, L. F. Edge,  and J. M. Nichol, “Charge-noise spectroscopy of si/sige quantum dots via dynamically-decoupled exchange oscillations,” Nature Communications 13, 940 (2022).
  • Takeda et al. (2013) K. Takeda, T. Obata, Y. Fukuoka, W. M. Akhtar, J. Kamioka, T. Kodera, S. Oda,  and S. Tarucha, “Characterization and suppression of low-frequency noise in Si/SiGe quantum point contacts and quantum dots,” Applied Physics Letters 102, 123113 (2013)https://pubs.aip.org/aip/apl/article-pdf/doi/10.1063/1.4799287/13222126/123113_1_online.pdf .
  • Pioro-Ladrière et al. (2005) M. Pioro-Ladrière, J. H. Davies, A. R. Long, A. S. Sachrajda, L. Gaudreau, P. Zawadzki, J. Lapointe, J. Gupta, Z. Wasilewski,  and S. Studenikin, “Origin of switching noise in GaAs/Al1xGaxAsGaAssubscriptAl1xsubscriptGaxAs\mathrm{GaAs}/\mathrm{Al_{1-x}Ga_{x}As}roman_GaAs / roman_Al start_POSTSUBSCRIPT 1 - roman_x end_POSTSUBSCRIPT roman_Ga start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT roman_As lateral gated devices,” PRB 72, 115331 (2005).
  • Struck et al. (2024) T. Struck, A. Hollmann, F. Schauer, O. Fedorets, A. Schmidbauer, K. Sawano, H. Riemann, N. V. Abrosimov, L. Cywiński, D. Bougeard,  and L. R. Schreiber, “Spin-epr-pair separation by conveyor-mode single electron shuttling in si/sige,” Nature Communications 15, 1325 (2024).
  • Hoex et al. (2008) B. Hoex, J. J. H. Gielis, M. C. M. van de Sanden,  and W. M. M. Kessels, “On the c-si surface passivation mechanism by the negative-charge-dielectric al2o3,” Journal of Applied Physics 104, 113703 (2008).
  • Gielis et al. (2008) J. J. H. Gielis, B. Hoex, M. C. M. van de Sanden,  and W. M. M. Kessels, “Negative charge and charging dynamics in al2o3 films on si characterized by second-harmonic generation,” Journal of Applied Physics 104, 073701 (2008).
  • Hiller et al. (2019) D. Hiller, P. M. Jordan, K. Ding, M. Pomaska, T. Mikolajick,  and D. König, “Deactivation of silicon surface states by al-induced acceptor states from al-o monolayers in sio2,” Journal of Applied Physics 125, 015301 (2019).
  • Hiller et al. (2021) D. Hiller, D. Tröger, M. Grube, D. König,  and T. Mikolajick, “The negative fixed charge of atomic layer deposited aluminium oxide–a two-dimensional sio2/alox interface effect,” Journal of Physics D: Applied Physics 54, 275304 (2021).
  • Elsayed et al. (2022) A. Elsayed, M. Shehata, C. Godfrin, S. Kubicek, S. Massar, Y. Canvel, J. Jussot, G. Simion, M. Mongillo, D. Wan, B. Govoreanu, I. P. Radu, R. Li, P. V. Dorpe,  and K. D. Greve, “Low charge noise quantum dots with industrial cmos manufacturing,”   (2022), arXiv:2212.06464 [cond-mat.mes-hall] .
  • Hanson et al. (2007) R. Hanson, L. P. Kouwenhoven, J. R. Petta, S. Tarucha,  and L. M. K. Vandersypen, “Spins in few-electron quantum dots,” RMP 79, 1217–1265 (2007).
  • Paladino et al. (2014) E. Paladino, Y. M. Galperin, G. Falci,  and B. L. Altshuler, “1/f1𝑓1/f1 / italic_f noise: Implications for solid-state quantum information,” RMP 86, 361–418 (2014).
  • Barnes and Allan (1966) J. A. Barnes and D. W. Allan, “A statistical model of flicker noise,” Proceedings of the IEEE 54, 176–178 (1966).
  • Huang et al. (2014) C.-T. Huang, J.-Y. Li, K. S. Chou,  and J. C. Sturm, “Screening of remote charge scattering sites from the oxide/silicon interface of strained si two-dimensional electron gases by an intermediate tunable shielding electron layer,” Appl. Phys. Lett. 104, 243510 (2014).
  • Huang (2015) C.-T. Huang, Electrical and Material Properties of Strained Silicon/Relaxed Silicon Germanium Heterostructures for Single-Electron Quantum Dot Applications, Ph.D. thesis, Princeton University (2015).
  • Wilamowski et al. (2001) Z. Wilamowski, N. Sandersfeld, W. Jantsch, D. Többen,  and F. Schäffler, “Screening breakdown on the route toward the metal-insulator transition in modulation doped si ///sige quantum wells,” PRL 87, 026401 (2001).
  • Tracy et al. (2009) L. A. Tracy, E. H. Hwang, K. Eng, G. A. Ten Eyck, E. P. Nordberg, K. Childs, M. S. Carroll, M. P. Lilly,  and S. Das Sarma, “Observation of percolation-induced two-dimensional metal-insulator transition in a si mosfet,” PRB 79, 235307 (2009).
  • Thomas (1927) L. H. Thomas, “The calculation of atomic fields,” Mathematical Proceedings of the Cambridge Philosophical Society 23, 542–548 (1927).
  • Fermi (1928) E. Fermi, “Eine statistische methode zur bestimmung einiger eigenschaften des atoms und ihre anwendung auf die theorie des periodischen systems der elemente,” Zeitschrift für Physik 48, 73–79 (1928).
  • Tang, O’Regan, and Wu (2004) T.-W. Tang, T. O’Regan,  and B. Wu, “Thomas-fermi approximation for a two-dimensional electron gas at low temperatures,” J. Appl. Phys. 95, 7990–7997 (2004).
  • Kane (1959) E. O. Kane, “The semi-empirical approach to band structure,” Journal of Physics and Chemistry of Solids 8, 38–44 (1959).
  • Trellakis, Andlauer, and Vogl (2006) A. Trellakis, T. Andlauer,  and P. Vogl, “Efficient solution of the schrödinger-poisson equations in semiconductor device simulations,” in Large-Scale Scientific Computing, edited by I. Lirkov, S. Margenov,  and J. Waśniewski (Springer Berlin Heidelberg, Berlin, Heidelberg, 2006) pp. 602–609.
  • Fowler and Nordheim (1928) R. H. Fowler and L. Nordheim, “Electron emission in intense electric fields,” Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character 119, 173–181 (1928)https://royalsocietypublishing.org/doi/pdf/10.1098/rspa.1928.0091 .
  • Albrecht, Moers, and Hermanns (2017) W. Albrecht, J. Moers,  and B. Hermanns, “HNF - Helmholtz Nano Facility,” Journal of large-scale research facilities JLSRF 3, A112 (2017).