Abstract
As first recognized in 2010, epitaxial graphene on SiC(0001) provides a platform for quantized Hall resistance (QHR) metrology unmatched by other two-dimensional structures and materials. Here we report graphene parallel QHR arrays, with metrologically precise quantization near 1000 Ω. These arrays have tunable carrier densities, due to uniform epitaxial growth and chemical functionalization, allowing quantization at the robust ν = 2 filling factor in array devices at relative precision better than 10−8. Broad tunability of the carrier density also enables investigation of the ν = 6 plateau. Optimized networks of QHR devices described in this work suppress Ohmic contact resistance error using branched contacts and avoid crossover leakage with interconnections that are superconducting for quantizing magnetic fields up to 13.5 T. Our work enables more direct scaling of resistance for quantized values in arrays of arbitrary network geometry.
I. INTRODUCTION
The von Klitzing constant, RK = h/e2, is defined in the 2019 redefinition of the SI base units as a universal reference of electrical resistance [1], and it is accessed through the integer quantum Hall effect (QHE) [2]. A single device used as a quantized Hall resistance (QHR) standard can provide only one or two highly precise resistance plateaus. Resistance scaling to artifact standards at the highest precision requires cryogenic ratio bridges, so access to a broad and useful spectrum of traceable resistance measurements is out of reach for many laboratories. This has raised the interest in constructing accessible quantum-based resistance standards using arrays of semiconductor QHE devices [3–7], and in the case of epitaxial graphene (EG), using arrays [8,9] or devices with multiple regions of opposite charge separated by sharp p-n junctions [10–12].
The magnetoresistance in EG devices is dominated by a robust plateau [13,14] at Landau level (LL) filling factor ν = 2, which allows precise measurements for a wide range of magnetic field B, source-drain current I, and temperature T. Here we describe parallel QHE arrays based on N interconnected EG devices. Parallel arrays of resistance (RK/Nν) bring practical benefits, since they provide broad areal distribution of power dissipation to increase the QHE breakdown current. This allows precise measurements with non-cryogenic measurement electronics, which are limited by current sensitivity for higher-resistance, lower-current QHR standards.
In our devices, external voltage and current terminals are connected to the source and drain of 13 QHE elements, as shown in Fig. 1(a). NbTiN films form parallel superconducting interconnections to the 13 elements, with resistances rc ≈ 1 Ω at the 25-nm-thick Pd/Au underlayer which forms the contact interface to the two-dimensional electron gas (2DEG) [see Fig. 1(b)]. Multiple connections to the equipotential edges are central to QHR arrays, as described by Delahaye [15] and investigated by Jeffery and co-workers [16]. Multiple contacts are a topological solution that allows precise two-terminal QHE measurements as if the device had ideal contacts (rc = 0) for devices with longitudinal resistivity ρxx ≈ 0.
FIG. 1.

Device layout and features in a quantizing magnetic field. (a) Each of 13 array elements is 0.4 mm × 0.2 mm monolayer EG and has multiple contact branches. (b) Single array element contact layout, with a Pd/Au underlayer that provides EG surface adhesion and ensures a normal metal interface with NbTiN superconductor. (c) With strong quantization, boundaries of high and low electrochemical potential surround the EG, beginning at the two hotspots created by the power dissipation at the source and drain. These are labeled equipotential, although small changes may occur where current enters or leaves the device at intermediate contacts. The positions of hotspots depend on the orientation of B. In one field direction B+ (magnetic vector pointing into the page; equipotential boundary shown by the green line), the Hall voltage is nearly perpendicular to the source-drain current vector. This voltage is labeled VH+xx to account for Rxx contributions due to the small diagonal offset of the voltage probes, as described in the text. The opposite field direction (B−) (equipotential boundary shown by the dashed orange line) causes the voltage and current contact points to exchange positions and increases the sensitivity to longitudinal resistance due to the width of the regions labeled source and drain.
In Fig. 1(c), the superconductor wraps around from the source to the top Hall contact, and from the drain to the opposing Hall contact. For fixed total current I, most current flows into the 2DEG near the corners of the source-drain contacts (hot spots) as selected by the magnetic field vector. If the contact resistances rc are too large, then some current could flow through the final Hall contacts; however, the error decreases by a factor ϵ ≈ (ρxx + rc)/RK for each added edge connection to a quantized 2DEG element [15,16] provided that the metallic reservoir contacts are well separated [17].
Before all measurements, these devices were functionalized using chemical treatment to obtain η6-coordination of Cr(CO)3 to EG [18,19]. This forms equidistant bonds to three C atoms, preserving the planarity and superior electronic properties while providing a large drop in the carrier density of EG. This treatment provides p-type surface functionalization to counteract n-type doping of ≈1 × 1013 cm−2 induced in pristine EG on SiC(0001) by the substrate interface. Gentle heating at 40 °C–100 °C provides fine adjustment of the charge functionalization by removing weakly bound molecules adsorbed from the atmosphere.
II. QUANTIZATION OF THE HALL RESISTANCE
We will present measurements at several carrier concentrations n0 produced by gentle heating in vacuum for device 1 and device 2, which share the same 4H-SiC(0001) substrate. In the parallel arrays, we cannot obtain n0 from the low-field slope of the Hall resistance due to the Rxx contribution at low B. At sufficiently high carrier densities, we can reliably derive n0 when the ν = 6 plateau near RK/78 ≈ 331 Ω is resolved as described in Sec. II D. We then use the slope from the origin to the approximate ν = 6 plateau center to find n0 and estimate the mobility μ.
A. Initial characterization
Figure 2 shows low resolution results for the array resistance at T = 1.6 K while sweeping the magnetic field. The symmetric form of the resistance profile is caused by reciprocity between voltage and current, as shown by the green and orange lines in Fig. 1(c). Onsager-Casimir reciprocity [20] with magnetic field reversal exists for linear electronic circuits and is well-known in QHE systems [6,15,17]. The observed magnetoresistance peaks centered at B = 0 are ascribed to weak localization resistance in monolayer graphene [21]. The decreased peak height implies electron heating in the high-current regime [22].
FIG. 2.

Resistance of array device 2 as a function of magnetic field B measured using a current source and a digital voltmeter. The symmetric resistance profile for B− and B+ typifies the combined voltage and current interconnections of a parallel array.
Cryogenic current comparator (CCC) bridge [23] measurements of two similar 13-element arrays are shown in Fig. 3. Device 1 is not perfectly quantized for I = 0.3 mA at T = 1.6 K, with relative deviation near 10−8 at ±9 T. Device 2 achieves full quantization above |7| T. For B+ and B− the deviations are (−0.65 ± 6.32) and (−0.25 ± 6.32) nΩ/Ω, respectively, at SI resistance value RK/26 ≈ 992.800 287 Ω. CCC ratio uncertainty is below 10−9, and most uncertainty originates from our 100 Ω artifact references. All expanded uncertainty values given here are for a 2σ confidence interval.
FIG. 3.

Symmetric cryogenic current comparator bridge data for two devices at 1.6 K, showing precise reciprocity with magnetic field reversal. Device 1 does not reach full quantization at |B| = 9 T. The results obtained on device 2 are quantized at 7.5 and 9 T, with deviation from the quantized value (−0.45 ± 4.47) n Ω/Ω when averaged for both field directions. Standard deviations are smaller than the size of the markers.
The room-temperature direct current comparator (DCC) bridge is more affordable than the CCC, requires much less training, and is widely used by calibration laboratories for low and moderate resistance values (<100 kΩ). DCC ratios are stable at the level of 10−8, and we regularly calibrate the DCC ratios using the same ratios of the CCC. Our choice of 13 QHR elements in parallel provides optimum DCC ratio sensitivity in automated bridge designs. All measurements that follow were made with a calibrated DCC, with the sample immersed in liquid 3He.
B. Atmospheric doping
As described by Chuang et al. [24], we use atmospheric chemical doping to reversibly modify the electronic properties of graphene, including carrier density and mobility. The initial growth of graphene on hexagonal, basal plane SiC(0001) creates an interface layer of carbon with covalent bonds to some Si atoms on the SiC silicon-face surface [25,26]. The electronic configuration of the carbon atoms that bind to Si is sp3-hybridized rather than sp2 as in free-standing graphene [27]. The interface layer is nonconducting but the electronic sp3 hybridization induces n-type doping in the conducting epitaxial graphene layers that form later and lie above the interface layer [28,29]. Atmospheric molecules adsorbed on EG tend to act as electron acceptors, and thus reduce the n-type doping. This doping process involves several atmospheric gases with various rates of adsorption [19], as has been observed over periods of hours to months [30,31]. When Cr(CO)3 η6-graphene hexahapto functionalization is produced by reaction with heated Cr(CO)6 vapor, some of the atmospheric dopant molecules are replaced by Cr(CO)3, but many carbon atoms remain exposed to readsorb these dopants when exposed to ambient air. The Cr(CO)3 functionalization is stable for gentle heating up to about 150 °C, and atmospheric doping is reduced by heating at lower temperatures, thus heating in vacuum can be used to drive off atmospheric dopants and adjust the carrier density. Our results with Cr(CO)3 show that the carrier density can be adjusted from p-type to n-type, or from lower n-type to higher n-type levels by such heating, as reported in several prior publications [32–34].
C. Device fabrication
Our lithographic fabrication process is shown in Fig. 4. First, a protection layer (10 nm Pd/15 nm Au) is deposited on the graphene with an e-beam evaporator. The protection layer is used to prevent the contamination of graphene during the fabrication process [34]. Pd is used here to uniformly cover graphene [35], and gold is used to prevent the oxidation of the Pd. After depositing the protection layer, we deposited 80 nm gold with a Hall bar pattern which acts as a sacrificial layer. The sacrificial layer is used as a metal mask for the following processing. Then, we use ion milling to define the EG regions on the surface of the device. Ar ions will uniformly etch the surface of the device so that only the regions covered with the sacrificed layer will remain after ion milling. The superconducting electrodes (NbTiN) are then sputtered on the device on top of the Pd/Au contact regions, and they do not directly touch graphene to prevent the interference between the quantum Hall state in graphene and the superconducting state in NbTiN [36,37]. Finally, we use dilute aqua regia to remove the protection layer on the graphene and then functionalize the graphene surface with Cr(CO)3.
FIG. 4.

(a) As-grown graphene. (b) Protection layer deposition. (c) Sacrificed layer deposition. (d) Definition of the device pattern with ion milling. (e) Superconductor electrode deposition in a sputter chamber. (f) Removal of the sacrificial layer with dilute aqua regia.
D. Carrier density tuning
We characterized the QHR arrays at the plateaus shown in Fig. 5, where both ν = 2 and 6 regions are present for some carrier densities. Heating in vacuum [24] was used to increase the n-type carrier density in our devices by ≈40%, with n0 ≈ 7 × 1011 cm−2 and μ ≈ 4000 cm2/V s for both device 1 and device 2, in agreement with values derived from the low-field Hall slope for large-area single EG devices produced by similar methods [38,39]. The broad magnetoresistance features are given by the black curve in Fig. 5, and precise DCC ratios were measured based on the same group of 100 Ω standard resistors (see Fig. 6). B+ yielded a ν = 2 resistance plateau value ≈(7.3 ± 6.9) × 10−9 higher than RK/26 at T = 0.35 K, starting from B = 12.9 T to maximum field B = 13.5 T (see Fig. 7). This offset in resistance quantization is consistent with earlier CCC data (Fig. 3, green curve in Fig. 5) at T = 1.6 K and B = 9 T, where carrier density n0 can now be estimated as 4 × 1011 cm−2 by comparing the ν = 2 plateau onsets.
FIG. 5.

Measured QHARS resistance using a current source and digital voltmeter, for several carrier density levels (n0). Starting with the highest resistance at B = 0 (green), heating in vacuum was used to increase n0 to around 7 × 1011 cm−2 (black). Subsequent heating cycles in vacuum were used to produce the blue curve with n0 ≈ 1.6 × 1012 cm−2 (not discussed in the text), and the yellow curve with n0 ≈ 1.7 × 1012 cm−2, where a broad plateau is observed near 330.93 Ω.
FIG. 6.

The region near 330.93 Ω for the black curve of Fig. 5, with n0 ≈ 7 × 1011 cm−2. The ν = 6 plateau is not visible, but these points show the minimum slope region measured using the DCC bridge at resistance levels near RK/78. The slope of the fitted line is ≈54.5 (mΩ/Ω)/T and the plateau center was estimated as B ≈ 4.8 T.
FIG. 7.

The ν = 2 plateau measured at selected values of B for device 1 at two temperatures (0.35 and 1.6 K) in a 3He cryostat using a room-temperature direct current comparator (DCC). The carrier density was determined as n0 ≈ 7 × 1011 cm−2 as shown by the black curve in Fig. 5. The inset shows data with a much finer scale, where the quantized plateau is seen to begin around 12.8 T for B+ and 13 T for B−. In general, the resistance values for B+ are closer to the quantized value for field values shown in the inset, but at somewhat lower fields the B− results are displaced closer to the plateau by increased longitudinal resistance.
We then heated the devices in vacuum again to obtain a much higher carrier density, yielding results shown in Figs. 8(a) and 8(b). Here n0 ≈ 1.7 × 1012 cm−2, and we estimate μ ≈ 2300 cm2/V s. The ν = 6 plateau, now centered near 11.4 T, was investigated using both devices with the DCC bridge for I = 300, 100, 50, and 20 μA (see the orange curve in Fig. 5). Device 1 displayed an approach to quantization near RK/78, however we see a strong dependence on the device current, much stronger than expected from the current dependence for the ν = 2 plateau (see Fig. 9). The current dependence is reduced for the low current range between 50 and 20 μA, and a reasonably flat and broad plateau appears for B+ at 20 μA as shown in Fig. 8(b). The deviation of the central plateau at ν = 6 for 20 μA is (0.17 ± 0.08) μΩ/Ω. The approach to reciprocity at the plateau centers is demonstrated in our devices at the two carrier densities where the ν = 6 and 2 plateaus in Fig. 8(c) begin at similar ranges of B, although the reciprocity range where R(B+) − R(B−) ≈ 0 is small for the ν = 6 plateau.
FIG. 8.

Resistance quantization measurements for device 1 at T ≈ 0.35 K. Standard deviations are smaller than the size of the markers, as derived from DCC bridge ratios. (a) The B+ and B− resistance for the ν = 6 plateau near RK/78 ≈ 330.933 429 Ω, where fitted lines are offset from the quantized value over a range of 1 T, with the scale in mΩ/Ω. (b) The ν = 6 resistance differs from the quantized value by (0.247 ± 0.054) μΩ/Ω for I = 20 μA and by (0.402 ± 0.027) μΩ/Ω for I = 50 μA over a range of 0.6 T in B+, from 11 to 11.6 T; the scale is in μΩ/Ω. For the plateau center of 11.2–11.4 T, the measured deviation is (0.171 ± 0.076) μΩ/Ω at 20 μA and (0.264 ± 0.078) μΩ/Ω at 50 μA. The expanded uncertainties are for a 2σ confidence interval. (c) Comparing the magnetic field reciprocity at the ν = 2 and 6 filling factors.
FIG. 9.

Current dependence. (a) The ν = 6 region, for device 1 and device 2, measured at two carrier density values. Quantization at the ν = 6 plateau was not attained for device 2, even at the higher carrier density. (b) The ν = 2 plateau resistance in device 1 measured at B = −13 T with source-drain current levels from 100 to 1000 μA for the carrier density n0 ≈ 7 × 1011 cm−2. Inset: The values obtained at and below 300 μA are consistent with full quantization at RK/26 with error bars showing the standard uncertainty of the DCC bridge data. The increase in the deviation at higher currents is proportional to the increase in power dissipation, or I2R.
III. DISCUSSION
A. Metrological requirements of quantized Hall array resistance standards
GaAs/AlGaAs heterostructure-based quantized Hall array resistance standards have demonstrated relative precision of a few parts in 109, approaching that of single devices, both at large and small values relative to RK [5–7]. The device elements (typically of quantized resistance value RK/2) in a parallel array with N elements share the current nearly equally allowing an increase in the measurement sensitivity with higher total current. Joule heat is dissipated in two small areas of each element adjacent to the contact points where current enters and exits the two-dimensional electron gas (2DEG). For small relative resistance differences Δn (n = 1, 2, … , N), the N-element array resistance is given by Taylor expansion as
| (1) |
When the ideal quantized plateau state is obtained for the QHE elements, all Δn should be too small to be measurable. In single QHR devices, one can follow international guidelines [40] to verify the ideal behavior of the QHR standard. Specifically, precise four-terminal measurements can verify that longitudinal resistance Rxx is negligible in the region of measurement, and three-terminal measurements can show that the resistance in the contacts is small compared to RK. These tests are impossible for arrays, because resistance measurements at the individual Hall contacts are inaccessible due to the permanent array interconnections between them.
As described earlier, arrays studied here have superconducting, crossover-free interconnections, and split contact design [37]. These new techniques successfully eliminate the accumulation of internal resistances and leakage currents that typically occur at interconnections and lead crossings between interconnected devices. The high critical magnetic field of the NbTiN superconductor ensures superconductivity up to T ≈ 6 K for magnetic flux density of 13.5 T. At B = 0, the typical critical transition temperature is 12.5 K [37]. Thus, in the ideal QHE condition where ρxx ≈ 0 for all elements, the EG array closely approximates a condition where the chemical potentials at the 2DEG Hall terminals are equal to the respective potentials of the superconducting source and drain.
Empirical studies show that the longitudinal resistivity, ρxx, in the 2DEG of GaAs-based heterostructures for fixed magnetic flux contributes to the deviation in Hall resistivity, Δρxy, as [41]
| (2) |
Here, s represents the effect of geometrical contact misalignment and microscopic 2DEG disorder. Equation (2) was shown to be applicable over three or four orders of magnitude in ρxx for conditions of variable temperature and power dissipation close to the plateau center in the QHE regime.
B. Landau-level broadening
Landau-level quantization in monolayer graphene is fourfold-degenerate; at moderate magnetic fields, only the quantum occupation numbers ν = (2, 6, 10, … .) appear in a sweep of R(B), and only the ν = 2 plateau is fully quantized. To our knowledge, no metrological measurements have been attempted previously with graphene in the ν = 6 quantized state, although both the ν = 4 and 2 plateaus were often used for metrology of GaAs/AlGaAs heterostructure-based devices in the 1980s and 1990s [41]. In semiconductor-based 2DEGs, the separation of the LLs is linear with the magnetic flux density B. In contrast, the energies of the LL states in graphene are given by
| (3) |
Here e is the elementary charge, ħ is the reduced Planck constant, νF is the Fermi velocity, n is the LL index, and E0 is the Dirac point energy. Thus, the LL center energy separation (E2–E1) above n = 1 is about 0.41 times as wide compared to the separation above the n = 0 LL, which is occupied equally by both electrons and holes. However, the degree of quantization is determined by the broadening of the LLs as well as their spacing.
Theoretical and experimental reports have raised questions concerning the broadening of LLs in monolayer graphene. Yang, Peeters, and Wu [42] included electron-impurity (e–i) and electron-electron (e-e) interactions, LL coupling, Fermi energy, and magnetic-field-dependent contributions. Funk, Knorr, Wendler, and Malic [43] found that the LL broadening is mainly due to e-i interactions and proportional to B1/2, counteracting the similar dependence on separation between graphene LLs. They note experimental evidence by Orlita et al. [44] indicating that the number of resolvable LL optical transitions is constant for a range of magnetic field strengths in EG, suggesting a dependence on (nB)1/2 in the LL broadening that cancels the increased LL spacing. For conventional 2DEG systems, the e–i broadening also scales as B1/2, but the spacing of the LLs increases linearly with B, increasing the separation and reducing the overlap in the density of states. Our results extend the study of LL broadening in EG, although we do see evidence from the plateau slope in Figs. 5, 8(a), and 8(b) that the LL overlap at ν = 6 decreases for higher n0 and produces a stronger plateau at higher B in the same device.
The deviation from quantization and the plateau evolution for the observed filling factors in electrical measurements is dependent on carrier density (in part due to e–e interactions and screening effects), mobility (e–i interactions and screening), temperature and source-drain current (phonons and thermal activation), and other factors. Up to now, detailed experimental studies of the interplateau transitions by varying these parameters have not been possible at a high level of precision.
Modifying the carrier density in EG causes a reciprocal change in the mobility, as shown by several studies [38,45]. It is likely that the ν = 6 plateau quantization may be degraded by the reduction in mobility that occurs along with the higher carrier density needed to observe this plateau.
It is well known that the presence of impurities broadens and obscures fractional QHE states [46–48] that lie close in energy, but also that some disorder is needed to produce broad LL plateaus for the integer QHE.
The scaling properties of condensed matter systems are of great interest, and such effects have been the basis of much theoretical and experimental work [49,50]. We note that most theoretical studies of LL broadening are on graphene on SiO2. Most studies of EG have focused on the ν = 2 quantum Hall plateau, which extends to high B due to charge transfer from the SiC substrate, and has also been observed in graphene covered by InSe [14,51]. Therefore, additional experimental and theoretical work focused on LL broadening specifically for graphene on SiC will be highly desirable in the search for improved and accessible electronic standards as well as a better understanding of the underlying interactions.
IV. CONCLUSION
In summary, we have described the QHR realized in EG devices constructed with 13 QHE elements in parallel, measured at the two plateaus that lie between the zeroth and second LLs. Carrier density adjustment is demonstrated using simple low-vacuum techniques. The carrier density is quite uniform over large sample areas of 1 mm × 5 mm, and the density of impurities that scatter electrons is also uniform for these EG devices, as shown by the width of the ν = 6 plateau at higher carrier densities and mobility of order μ ≈ 4000 cm2/V s at relatively high carrier density n0 ≈ 7 × 1011 cm−2.
These are the first graphene QHR arrays suitable as primary SI references, at a useful resistance level near 1 kΩ where room-temperature DCC bridges are most precise. The arrays have quantized resistance RK/26 with ν = 2 filling factor in each element. Quantized results for ν = 6 at resistance ≈RK/78 are less precise, but the measured offset of device 1 is approximately 2 × 10−7 near the plateau center for I = 20 μA. Magnetic field reciprocity gives evidence on the longitudinal resistivity in our QHE arrays, but it has not been carefully studied in epitaxial graphene. Future work on interplateau transitions and reciprocity, using similarly fabricated QHE single devices as well as arrays, may provide a better understanding of theoretical and experimental aspects of the QHE.
ACKNOWLEDGMENT
The authors thank Klaus von Klitzing, Stephan Schlamminger, and Maicol Ochoa for helpful discussions.
References
- [1].von Klitzing K, Chakraborty T, Kim P, Madhavan V, Dai X, McIver J, Tokura Y, Savary L, Smirnova D, Rey AM et al. , 40 years of the quantum Hall effect, Nat. Rev. Phys 2, 397 (2020). [Google Scholar]
- [2].von Klitzing K, Dorda G, and Pepper M, New Method for High Accuracy Determination of the Fine-Structure Constant Based on Quantized Hall Resistance, Phys. Rev. Lett 45, 494 (1980). [Google Scholar]
- [3].Piquemal FPM, Blanchet J, Genevès G, and André JP, A first attempt to realize (multiple-QHE devices)-series array resistance standards, IEEE Trans. Instrum. Meas 48, 296 (1999). [Google Scholar]
- [4].Poirier W, Bounouh A, Piquemal F, and André JP, A new generation of QHARS: Discussion about the technical criteria for quantization, Metrologia 41, 285 (2004). [Google Scholar]
- [5].Poirier W, Bounouh A, Hayashi K, Fhirna H, Piquemal F, Genevès G, and André JP, RK/100 and RK/200 quantum Hall array resistance standards, J. Appl. Phys 92, 2844 (2002). [Google Scholar]
- [6].Könemann J, Ahlers F-J, Pesel E, Pierz K, and Schumacher HW, Magnetic field reversible serial quantum Hall arrays, IEEE Trans. Instrum. Meas 60, 2512 (2011). [Google Scholar]
- [7].Oe T, Gorwadkar S, Itatani T, and Kaneko N-H, Development of 1 MΩ quantum Hall array resistance standards, IEEE Trans. Instrum. Meas 66, 1475 (2017). [Google Scholar]
- [8].Park J, Kim WS, and Chae DS, Realization of 5h/e2 with graphene quantum Hall resistance array, Appl. Phys. Lett 116, 093102 (2020). [Google Scholar]
- [9].Kruskopf M, Rigosi AF, Panna AR, Marzano M, Patel D, Jin H, Newell DB, and Elmquist RE, Next-generation crossover-free quantum Hall arrays with superconducting interconnections, Metrologia 56, 065002 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [10].Lohmann T, von Klitzing K, and Smet JH, Four-terminal magneto-transport in graphene p-n junctions created by spatially selective doping, Nano Lett 9, 1973 (2009). [DOI] [PubMed] [Google Scholar]
- [11].Woszczyna M, Friedemann M, Dziomba T, Weimann Th., and Ahlers FJ, Graphene p-n junction arrays as quantum-Hall resistance standards, Appl. Phys. Lett 99, 022112 (2011). [Google Scholar]
- [12].Hu J, Rigosi AF, Kruskopf M, Yang Y, Wu B-Y, Tian J, Panna AR, Lee H-Y, Payagala SU, Jones GR et al. , Towards epitaxial graphene p-n junctions as electrically programmable quantum resistance standards, Sci. Rep 8, 15018 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Tzalenchuk A, Lara-Avila S, Kalaboukhov A, Paolilo S, Syväjärvi M, Yakimova R, Kazakova O, Janssen TJBM, Fal’ko VI, and Kubatkin S, Towards a quantum resistance standard based on epitaxial graphene, Nat. Nanotech 5, 186 (2010). [DOI] [PubMed] [Google Scholar]
- [14].Janssen TJBM, Tzalenchuk A, Yakimova R, Kubatkin S, Lara-Avila S, Kopylov S, and Fal’ko VI, Anomalously strong pinning of the filling factor ν = 2 in epitaxial graphene, Phys. Rev. B 83, 233402 (2011). [Google Scholar]
- [15].Delahaye F, Series and parallel connection of multiterminal quantum Hall-effect devices, J. Appl. Phys 73, 7914 (1993). [Google Scholar]
- [16].Jeffery A, Elmquist RE, and Cage ME, Precision tests of a QHE device DC equivalent circuit using double-series and triple-series connections, J. Res. Nat. Inst. Stand. Technol 100, 677 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [17].Büttiker M, Absence of backscattering in the quantum Hall effect in multiprobe conductors, Phys. Rev. B 38, 9375 (1988). [DOI] [PubMed] [Google Scholar]
- [18].Sarkar S, Zhang H, Huang J-W, Wang F, Bekyarova E, Lau CN, and Haddon RC, Organometallic hexahapto functionalization of single layer graphene as a route to high mobility graphene devices, Adv. Mater 25, 1131 (2013). [DOI] [PubMed] [Google Scholar]
- [19].Rigosi AF, Kruskopf M, Hill HM, Jin H, Wu B-Y, Johnson PE, Zhang S, Berilla M, Hight-Walker AR, Hacker CA et al. , Gateless and reversible carrier density tunability in epitaxial graphene devices functionalized with chromium tricarbonyl, Carbon 142, 468 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Casimir H, On Onsager’s principle of microscopic reversibility, Rev. Mod. Phys 17, 343 (1945). [Google Scholar]
- [21].McCann E, Kechedzhi K, Fal’ko VI, Suzuura H, Ando T, and Altshuler BL, Weak-Localization Magnetoresistance and Valley Symmetry in Graphene, Phys. Rev. Lett 97, 146805 (2006). [DOI] [PubMed] [Google Scholar]
- [22].Baker AMR, Alexander-Webber JA, Altebaeumer T, Janssen TJBM, Tzalenchuk A, Lara-Avila S, Kubatkin S, Yakimova R, Lin C-T, Li L-J et al. , Weak localization scattering lengths in epitaxial and CVD graphene, Phys. Rev. B 86, 235441 (2012). [Google Scholar]
- [23].Williams JM, Cryogenic current comparators and their application to electrical metrology, IET Sci. Meas. Technol 5, 211 (2011). [Google Scholar]
- [24].Chuang C, Yang Y, Pookpanratana S, Hacker C, Liang C-T, and Elmquist RE, Chemical-doping-driven crossover from graphene to “ordinary metal” in epitaxial graphene grown on SiC, Nanoscale 9, 11537 (2017). [DOI] [PubMed] [Google Scholar]
- [25].Lampin E, Priester C, Krzeminski C, and Magaud L, Graphene buffer layer on Si-terminated SiC studied with an empirical interatomic potential, J. Appl. Phys 107, 103514 (2010). [Google Scholar]
- [26].Riedl C, Starke U, Bernhardt J, Franke M, and Heinz K, Structural properties of the graphene-SiC(0001) interface as a key for the preparation of homogeneous large-terrace graphene surfaces, Phys. Rev. B 76, 245406 (2007). [Google Scholar]
- [27].Goler S, Coletti C, Piazza V, Pingue P, Colangelo F, Pellegrini V, Emtsev KV, Forti S, Starke U, Beltram F et al. , Revealing the atomic structure of the buffer layer between SiC(0001) and epitaxial graphene, Carbon 51, 249 (2013). [Google Scholar]
- [28].Nicotra G, Ramasse QM, Deretzis I, La Magna A, Spinella C, and Giannazzo F, Delaminated graphene at silicon carbide facets: Atomic scale imaging and spectroscopy, ACS Nano 7, 3045 (2013). [DOI] [PubMed] [Google Scholar]
- [29].Hannon JB, Copel M, and Tromp RM, Direct Measurement of the Growth Mode of Graphene on SiC(0001) and SiC(000–1), Phys. Rev. Lett 107, 166101 (2011). [DOI] [PubMed] [Google Scholar]
- [30].Pearce R, Eriksson J, Iakimov T, Hultman L, Spetz AL, and Yakimova R, On the differing sensitivity to chemical gating of single and double layer epitaxial graphene explored using scanning kelvin probe microscopy, ACS Nano 7, 4647 (2013). [DOI] [PubMed] [Google Scholar]
- [31].Knight S, Hofmann T, Bouhafs C, Armakavicius N, Kühne P, Stanishev V, Ivanov IG, Yakimova R, Wimer S, Schubert M et al. , In-situ terahertz optical Hall effect measurements of ambient effects on free charge carrier properties of epitaxial graphene, Sci. Rep 7, 5151 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [32].Rigosi AF, Liu C-I, Glavin NR, Yang Y, Hill HM, Hu J, Hight-Walker AR, Richter CA, Elmquist RE, and Newell DB, Electrical stabilization of surface resistivity in epitaxial graphene systems by amorphous boron nitride encapsulation, ACS Omega 2, 2326 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [33].Chuang C, Liu C-W, Yang Y, Syong W-R, Liang C-T, and Elmquist RE, Magnetoresistance of ultralow-hole-density monolayer epitaxial graphene grown on SiC, Materials 12, 2696 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [34].Yang Y, Huang L-I, Fukuyama Y, Liu F-H, Real MA, Barbara P, Liang C-T, Newell DB, and Elmqust RE, Low carrier density epitaxial graphene devices on SiC, Small 11, 90 (2015). [DOI] [PubMed] [Google Scholar]
- [35].Watanabe E, Conwill A, Tsuya D, and Koide Y, Low contact resistance metals for graphene based devices, Diam. Relat. Mater 24, 171 (2012). [Google Scholar]
- [36].Rickhaus P, Weiss M, Marot L, and Schönenberger C, Quantum Hall effect in graphene with superconducting electrodes, Nano Lett 12, 1942 (2012). [DOI] [PubMed] [Google Scholar]
- [37].Kruskopf M, Rigosi AF, Panna AR, Patel DK, Jin H, Marzano M, Berilla M, Newell DB, and Elmquist RE, Two-terminal and multi-terminal designs for next-generation quantized Hall resistance standards: Contact material and geometry, IEEE Trans. Electron Dev 66, 3973 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Yang Y, Cheng G, Mende P, Calizo IG, Feenstra RM, Chuang C, Liu C-W, Liu C-I, Jones GR, Hight-Walker AR et al. , Epitaxial graphene homogeneity and quantum Hall effect in millimeter-scale devices, Carbon 115, 229 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [39].Kruskopf M and Elmquist RE, Epitaxial graphene for quantum resistance metrology, Metrologia 55, R27 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [40].Delahaye F and Jeckelmann B, Revised technical guidelines for reliable dc measurements of the quantized Hall resistance, Metrologia 40, 217 (2003). [Google Scholar]
- [41].Jeckelmann B and Jeanneret B, The quantum Hall effect as an electrical resistance standard, Rep. Prog. Phys 64, 1603 (2001). [Google Scholar]
- [42].Yang CH, Peeters FM, and Xu W, Landau-level broadening due to electron-impurity interaction in graphene in strong magnetic fields, Phys. Rev. B 82, 075401 (2010). [Google Scholar]
- [43].Funk H, Knorr A, Wendler F, and Malic E, Microscopic view on Landau level broadening mechanisms in graphene, Phys. Rev. B 92, 205428 (2015). [Google Scholar]
- [44].Orlita M, Faugeras C, Grill R, Wysmolek A, Strupinski W, Berger C, de Heer WA, Martinez G, and Potemski M, Carrier Scattering from Dynamical Magnetoconductivity in Quasineutral Epitaxial Graphene, Phys. Rev. Lett 107, 216603 (2011). [DOI] [PubMed] [Google Scholar]
- [45].Tedesco JL, VanMil BL, Myers-Ward RL, McCrate JM, Kitt SA, Campbell PM, Jernigan GG, Culbertson JC, Eddy CR Jr., and Gaskill DK, Hall effect mobility of epitaxial graphene grown on silicon carbide, Appl. Phys. Lett 95, 122102 (2009). [Google Scholar]
- [46].Tsui DC, Stormer HL, and Gossard AC, Two-Dimensional Magnetotransport in the Extreme Quantum Limit, Phys. Rev. Lett 48, 1559 (1982). [Google Scholar]
- [47].Zhang Y, Jiang Z, Small JP, Purewal MS, Tan Y-W, Fazlollahi M, Chudow JD, Jaszczak JA, Stormer HL, and Kim P, Landau-Level Splitting in Graphene in High Magnetic Fields, Phys. Rev. Lett 96, 136806 (2006). [DOI] [PubMed] [Google Scholar]
- [48].Tőke C, Lammert PE, Crespi VH, and Jain JK, Fractional quantum Hall effect in graphene, Phys. Rev. B 74, 235417 (2006). [Google Scholar]
- [49].Fischer M and Barber M, Scaling Theory for Finite-Size Effects in the Critical Region, Phys. Rev. Lett 28, 1516 (1972). [Google Scholar]
- [50].Kivelson S, Lee D-H, and Zhang S-C, Global phase diagram in the quantum Hall effect, Phys. Rev. B 46, 2223 (1992). [DOI] [PubMed] [Google Scholar]
- [51].Kudrynskyi ZR, Bhuiyan MA, Makarovsky O, Greener JDG, Vdovin EE, Kovalyuk ZD, Cao Y, Mishchenko A, Novoselov KS, Beton PH et al. , Giant Quantum Hall Plateau in Graphene Coupled to an InSe van der Waals Crystal, Phys. Rev. Lett 119, 157701 (2017). [DOI] [PubMed] [Google Scholar]
