Next Article in Journal
Graph Theoretic Analyses of Tessellations of Five Aperiodic Polykite Unitiles
Previous Article in Journal
Performance Analysis and Cost Optimization of the M/M/1/N Queueing System with Working Vacation and Working Breakdown
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Probing Chirality of the Quantum Hall Effect via the Landauer–Büttiker Formalism with Two Current Sources

Department of Physics and Astronomy, Sejong University, Seoul 05006, Republic of Korea
Mathematics 2025, 13(18), 2981; https://doi.org/10.3390/math13182981
Submission received: 12 August 2025 / Revised: 11 September 2025 / Accepted: 12 September 2025 / Published: 15 September 2025
(This article belongs to the Special Issue Mathematics Methods in Quantum Physics and Its Applications)

Abstract

The quantum Hall effect is a paradigmatic example of topological order, characterized by precisely quantized Hall resistance and dissipationless edge transport. These edge states are chiral, propagating unidirectionally along the boundary, and their directionality is determined by the external magnetic field. While chirality is a central feature of the quantum Hall effect, directly probing it remains experimentally nontrivial. In this study, we introduce a simple and effective method to probe the chirality of edge transport using two independently controlled current sources in a Hall bar geometry. The system under investigation is monolayer epitaxial graphene grown on a silicon carbide substrate, exhibiting robust quantum Hall states. By varying the configurations of the two current sources, we measure terminal voltages and analyze the transport characteristics. Our results demonstrate that the observed behavior can be understood as a linear superposition of chiral contributions to the edge transport. This superposition enables tunable combinations of longitudinal and Hall resistances and enables additive or canceling behavior of Hall voltages depending on current source configuration. The Landauer–Büttiker formalism provides a quantitative framework to describe these observations, capturing the interplay between edge state chirality and the measurement configuration. This research offers a simple yet effective experimental and analytical approach for probing chiral edge currents and highlights the linear superposition principle in the quantum Hall effect.

1. Introduction

The quantum Hall effect (QHE) is characterized by the precise quantization of Hall resistance and the vanishing of longitudinal resistance in the presence of a strong perpendicular magnetic field [1]. Depending on the system, the quantized values can correspond to integer [2], half-integer [3], or fractional multiples of the fundamental resistance quantum [4]. This quantization arises from the formation of topologically non-trivial insulating bulk states, which give rise to robust and dissipationless edge channels [5,6,7,8]. These topological features are believed to ensure the robustness and precision of quantized transport [9,10] and serve as the foundation for related phenomena such as the quantum anomalous Hall effect [11] and the quantum spin Hall effect [12], which can occur even in the absence of an external magnetic field. The exactness of quantized Hall resistance has enabled its use in quantum resistance standards [9,10]. The QHE remains at the research frontier, with topics including the incompressible nature of the bulk [13], the appearance of hot spots [14], and the intriguing behavior of snake states [15].
One of the defining features of the QHE is the presence of chiral edge currents coexisting with an insulating bulk [1]. Classically, this behavior can be understood as arising from skipping cyclotron orbits along the sample boundaries [16]. Quantum mechanically, it originates from the formation of discrete Landau levels and their bending near the sample edge [16]. From a topological perspective, these edge states arise from the bulk–edge correspondence [5,6,7,8]. The chiral edge current explains the peculiar features of the QHE and the quantization of Hall voltages and the zero longitudinal resistance. However, the quantization of Hall resistance and zero longitudinal resistance does not directly prove the chiral edge current. The QHE without edge current has been studied [5] and has partly motivated the exploration of the QHE in a variety of geometries, including three-dimensional structures [17], series and parallel connections [18], and anti-Hall bar- or Corbino-type configurations [19,20,21,22,23,24,25]. Recently, the chiral edge states were directly probed by local probe techniques [26,27] and shot noise measurements [28,29]. These local probe techniques are based on microwave impedance microscopy [26] and a single-electron transistor [27]. These techniques require a rather special and demanding experimental setup.
In this work, we investigate the chirality of edge states in the QHE using two current source (TCS) excitations in a standard Hall bar geometry. The use of TCSs enables a variety of excitation and measurement configurations, offering new approaches to probing the chirality of edge currents generated by each source. All measurements across different configurations are analyzed using the Landauer–Büttiker formalism (LBF) [30,31]. The results are consistent with the presence of chiral edge currents and a bulk that remains incompressible: the total current contributed by each source flows in the same direction regardless of the configuration. This behavior contrasts sharply with that of conventional diffusive conductors, where oppositely directed currents would be expected to partially or fully cancel out. Compared with the local probe and shot noise techniques, this TCS-based method is more accessible and practical and yet provides compelling additional evidence for the chiral nature of edge states in the QHE.

2. Materials and Methods

For this study, we used monolayer epitaxial graphene (epigraphene) grown on silicon carbide substrates as the two-dimensional channel [9,10]. Epigraphene is an almost ideal two-dimensional system and thus serves as a perfect platform for the study of the chirality of the QHE. It is a single carbon atom thick, and its experimentally reported thickness is around 0.35 nm, which agrees with the interlayer distance of graphite [32]. The epigraphene was patterned into a Hall bar geometry (Figure 1a, width = 30 µm, length = 180 µm) using electron-beam lithography and was doped near the charge neutrality point (~5 × 1010 electrons/cm2) through a molecular doping method (See Figure 1f) [33]. This low carrier density enables high mobility, resulting in a robust QHE. After doping, the sample exhibited a carrier mobility of 32,000 cm2/V·s and a sheet resistance of 4 kΩ/square at a temperature of 2 K. A perpendicular magnetic field was applied to the sample surface to induce a QH state with a filling factor of ν = 2, corresponding to the first plateau in the half-integer QHE of epigraphene [3]. For the robustness of the QHE, all measurements were performed at T = 2 K, which is the lowest temperature of our experimental setup, where the thermal energy, kBT (kB is the Boltzmann constant), is sufficiently lower than the Landau level spacings for the magnetic field of this study. We used two Keithley current sources and nanovoltmeters. The systematic investigation of the effect of thickness and temperature on the QHE in the TCS configuration will be the subject of future work.
Before performing the main measurements, we evaluated the contact resistances of all eight terminals on the sample using the three-terminal method under quantum Hall conditions [34]. In the three-terminal method, contact resistance is estimated by measuring the longitudinal resistance with one of the voltage probes placed on the drain contact. Then the measured resistance is the sum of the contact resistance of the drain contact and the lead resistance. All contact resistances were found to be below 10 Ω, except for contact 2 (Rc = 60 Ω), contact 7 (Rc = 60 Ω), and contact 8 (Rc = 40 kΩ) (see Figure 1 for contact numbering). The carrier density extracted from the Hall voltage measured between terminals 2 and 8 was 3.5 × 1010 electrons/cm2, which deviates from the values obtained using other voltage probes (e.g., terminals 3–7 and 4–6), where the carrier density was approximately 5 × 1010 electrons/cm2. This discrepancy could be related to the high contact resistance of terminal 8. For the carrier density measurements, current was sourced through contacts 1 and 5. The robustness of the QHE in our sample is further confirmed by the observation of a critical current as high as 5 µA at a magnetic field of 2 T and a temperature of 2 K, where the longitudinal resistance starts to deviate from zero. Our bias currents are of the order of 100 nA, well below the critical current. This high critical current in epigraphene compared with GaAs in part arises from the large cyclotron Landau level spacings and high electron–phonon energy relaxation rates in epigraphene [35]. Critical current is an important phenomenon for resistance metrology [35]. However, a comprehensive breakdown theory that accounts for all its features has yet to be made available [36,37]. The proposed theories include an electron avalanche [38], field-induced tunneling between Landau levels [39], and spontaneous emission of phonons [40]. The TCS method would provide additional experimental conditions for the study of breakdown mechanisms in the future.

3. Results

Figure 1a shows a schematic of the TCS configuration, where the current sources A and B supply currents I a and I b , which are directed horizontally and vertically, respectively. The voltage measured between terminals 8 and 6 when both current sources are active ( I a = I 15 , I b = I 37 ) is denoted by V 15,37,86 . Due to the chirality of the edge currents in the QH regime, we expect the principle of superposition to hold in the TCS configuration: the voltage when both current sources are on should equal the sum of the voltages measured when each source is applied individually. To test this, we compare V 15,37,86 with V 15,86 = I 15 R 15,86 and V 37,86 = I 37 R 37,86 , where each is measured with the other current source turned off. As shown in Figure 1b, we find that V 15,37,86 = V 15,86 + V 37,86 , consistent with the quantized Hall resistance R = h / 2 e 2 or zero in the QH state. We further verify this superposition behavior by fixing I 15 = 100 nA and varying I 37 as 100 nA, 200 nA, or 300 nA. The resulting voltage V 15,37,86 changes accordingly, with V 15,86 remaining constant and V 37,86 varying linearly, as expected (Figure 1c,d). The superposition principle is also confirmed for another voltage terminal pair: V 15,37,28   = V 15,28 + V 37,28 , where V 15,28 = I 15 · ( h / 2 e 2 ) in the QH regime with ν = 2 , as shown in Figure 1e.
We explain the linear superposition behavior observed in the QHE with TCSs, as shown in Figure 1, using the LBF. In the QH regime at filling factor ν = 2 , the LBF accounts for chiral edge channels that contribute a quantum conductance of 2 e 2 / h , while the bulk remains incompressible. This quantization 2 e 2 / h arises from the half-integer QH effect in monolayer graphene, where the Hall conductivity is given by σ x y = g ( N + 1 2 ) e 2 / h , with degeneracy g = 4 reflecting the spin and valley degeneracy, and the zeroth Landau level ( N = 0 ) [3]. Within the LBF framework, the current and voltage at each terminal can be expressed as
I i = ( 2 e 2 / h ) j T j i V i T i j V j = ( 2 e 2 / h ) V i V i 1 .
Here, I i denotes the current flowing out of the i th terminal into the sample, V i is the voltage at the i th terminal, and T j i is the transmission probability for an electron injected from terminal i to reach terminal j . Charge conservation requires i I i = 0 . For dissipationless chiral transport in the QH regime, the transmission probabilities are defined as T j i = 1 if j   =   i   + 1 and T j i = 0 otherwise [30,31]. When both current sources are on, the LBF allows for linear superposition:
I i a , b = I i a + I i b
V i a , b = V i a + V i b ,
where I i a , b and V i a , b are the total current and voltage at terminal i with both sources are on, and I i a and V i a (or, respectively, I i b and V i b ) are those when only source a (or, respectively, b ) is on. In the specific configuration shown in Figure 1, where I a = I 15 and I b = I 37 , the LBF gives
I 1 a = I a = I 5 a
I 3 b = I b = I 7 b
with zero currents at all other terminals. For the voltages, we find
V i a = V 1 a = V a for   i = 1,2 , 3,4
V i b = V 3 b = V b for   i = 3,4 , 5,6
and the voltage is zero for the other terminals. From this, we obtain
V 8 a , b V 6 a , b = V 3 b = V b = ( h / 2 e 2 ) I b
V 2 a , b V 8 a , b = V 1 a = V a = ( h / 2 e 2 ) I a
Each voltage drop is determined solely by the corresponding current source. These results are in excellent agreement with the experimental observations in Figure 1b–e.
We investigate additional configurations and conclude that the QHE with TCSs consistently follows the additive nature of the LBF, regardless of the configuration. In Figure 2, the positions of the current sources are fixed as in Figure 1, but the voltage is measured between terminal 2 and 6. In this setup, both V 15,26 and V 37,26 contain contributions from both the longitudinal and transverse conductance arising from currents I a and I b , respectively. According to the LBF, the superposition of contributions from both sources indicates that both currents contribute to the Hall voltage. This prediction is verified in Figure 2. In Figure 2b, the Hall voltage is completely canceled out when I a = I b , consistent with the expected result. Figure 2c,d demonstrate that the Hall voltages can be tuned and made asymmetric by varying the relative magnitudes of the two current sources.
V 2 a , b V 6 a , b = V a V b = ( h / 2 e 2 ) ( I a I b )
In Figure 3, we explore configurations where the two current sources I a and I b are applied in parallel (Figure 3a) and antiparallel (Figure 3d) directions, such that the currents either add or cancel out. According to the LBF, for the parallel configuration in Figure 3a, the currents are distributed as
I 1 a = I a = I 5 a
I 2 b = I b = I 4 b
with zero currents at all other terminals. The corresponding voltages are
V i a = V 1 a = V a   for   i = 1,2 , 3,4
V i b = V 2 b = V b for   i = 2,3
The other terminals have zero voltage. In this configuration, the expected voltage differences are
V 8 a , b V 6 a , b = 0
V 3 a , b V 7 a , b = V a + V b = ( h / 2 e 2 ) ( I a + I b )
The experimental results in Figure 3b,c confirm these predictions.
In the antiparallel configuration shown in Figure 3d, only the direction of current for source B is reversed.
I 4 b = I b = I 2 b
V i b = V 4 b = V b for   i = 4,5 , 6,7 , 8,1 .
The voltages at the other terminals remain zero. The resulting voltage differences are
V 8 a , b V 6 a , b = V b V b = 0
V 3 a , b V 7 a , b = V a V b = ( h / 2 e 2 ) ( I a I b )
which are in excellent agreement with the experimental data shown in Figure 3e,f. These results clearly demonstrate that even when I b is applied from terminal 4 to terminal 2 in the antiparallel configuration, the current does not flow directly from the short path 4➔3➔2. Instead, due to the chirality of the edge states, it follows a longer clockwise path: 4➔5➔6➔7➔8➔1➔2.
We further examine a configuration in which the two current sources share a common drain contact, as shown in Figure 4. According to the LBF, the current distribution in this setup is I 1 a = I a = I 4 a , I 8 b = I b = I 4 b , with zero currents at all other terminals. The corresponding voltages are V i a = V 1 a = V a for terminals i = 1,2 , 3 and V i b = V 8 b = V b for i = 8,1 , 2,3 . All other terminals have zero voltages. From this, we find the following voltage differences:
V 2 a , b V 3 a , b = ( V a + V b ) ( V a + V b ) = 0
V 2 a , b V 7 a , b = V a + V b = ( h / 2 e 2 ) ( I a + I b )
These predictions are in excellent agreement with the experimental results shown in Figure 4b,c.

4. Discussion

We explored additional current source configurations and confirmed that the LBF consistently describes transport behavior in all cases. The chiral edge currents along the sample boundary are schematically illustrated in Figure 5 for a configuration where I a and I b are nominally applied in opposite directions. Unlike in conventional conductors, where opposing currents may cancel each other, the zero quantum Hall voltages observed in Figure 3e,f do not result from a vanishing net current. Instead, they arise from the specific voltage distribution dictated by chiral edge transport, as described by Equation (1).

5. Conclusions

We have experimentally demonstrated the linear superposition principle of QH transport under TCS excitations in a conventional Hall bar geometry. Our device is based on epitaxial monolayer graphene, doped near the charge neutrality point to enhance mobility and stabilize the QH state. The observed superposition reveals a combinatorial transport behavior in which both longitudinal and transverse QH resistances coexist in a tunable way. This TCS-based approach provides a powerful and flexible method for probing chirality and edge transport in QH systems.

Funding

This work was partly supported by the Institute of Information & Communications Technology Planning & Evaluation(IITP)-ITRC(Information Technology Research Center) grant funded by the Korea government(MSIT)(IITP-2025-RS-2024-00437191), the faculty research fund of Sejong University in 2025, and the Korean-Swedish Basic Research Cooperative Program of the NRF (No. NRF-2017R1A2A1A18070721), and the Swedish Foundation for Strategic Research (SSF) (No. IS14-0053, GMT14-0077, and RMA15-0024).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

K.H.K. thanks Hans He for fabricating the sample.

Conflicts of Interest

The author declares no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
QHEQuantum Hall Effect
TCSTwo Current Sources
LBFLandauer–Büttiker formalism
QHQuantum Hall

References

  1. Tong, D. Lectures on the Quantum Hall Effect. arXiv 2016, arXiv:1606.06687. [Google Scholar] [CrossRef]
  2. von Klitzing, K.; Dorda, G.; Pepper, M. New method for high-accuracy determination of the fine-structure constant based on quantized Hall resistance. Phys. Rev. Lett. 1980, 45, 494–497. [Google Scholar] [CrossRef]
  3. Zhang, Y.; Tan, Y.-W.; Stormer, H.L.; Kim, P. Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 2005, 438, 201–204. [Google Scholar] [CrossRef] [PubMed]
  4. Stormer, H.L.; Tsui, D.C.; Gossard, A.C. The fractional quantum Hall effect. Rev. Mod. Phys. 1999, 71, S298–S305. [Google Scholar] [CrossRef]
  5. von Klitzing, K. The quantum Hall effect—An edge phenomenon? Physical B 1993, 184, 1–6. [Google Scholar] [CrossRef]
  6. Avron, J.E.; Osadchy, D.; Seiler, R. A topological look at the quantum Hall effect. Phys. Today 2003, 56, 38–43. [Google Scholar] [CrossRef]
  7. Thouless, D.J.; Kohmoto, M.; Nightingale, M.P.; den Nijs, M. Quantized Hall conductance in a two-dimensional periodic potential. Phys. Rev. Lett. 1982, 49, 405–408. [Google Scholar] [CrossRef]
  8. Hatsugai, Y. Chern number and edge states in the integer quantum Hall effect. Phys. Rev. Lett. 1993, 71, 3697–3700. [Google Scholar] [CrossRef]
  9. Tzalenchuk, A.; Lara-Avila, S.; Kalaboukhov, A.; Paolillo, S.; Syväjärvi, M.; Yakimova, R.; Kazakova, O.; Janssen, T.J.B.M.; Fal’ko, V.; Kubatkin, S. Towards a quantum resistance standard based on epitaxial graphene. Nat. Nanotechnol. 2010, 5, 186–189. [Google Scholar] [CrossRef]
  10. He, H.; Cedergren, K.; Shetty, N.; Lara-Avila, S.; Kubatkin, S.; Bergsten, T.; Eklund, G. Accurate graphene quantum Hall arrays for the new International System of Units. Nat. Commun. 2022, 13, 6933. [Google Scholar] [CrossRef]
  11. Chang, C.-Z.; Liu, C.-X.; MacDonald, A.H. Colloquium: Quantum anomalous Hall effect. Rev. Mod. Phys. 2023, 95, 011002. [Google Scholar] [CrossRef]
  12. Bernevig, B.A.; Zhang, S.-C. Quantum Spin Hall Effect. Phys. Rev. Lett. 2006, 96, 106802. [Google Scholar] [CrossRef] [PubMed]
  13. Kendirlik, E.M.; Sirt, S.; Kalkan, S.B.; Ofek, N.; Umansky, V.; Siddiki, A. The local nature of incompressibility of quantum Hall effect. Nat. Commun. 2017, 8, 14082. [Google Scholar] [CrossRef] [PubMed]
  14. Komiyama, S.; Sakuma, H.; Ikushima, K.; Hirakawa, K. Electron temperature of hot spots in quantum Hall conductors. Phys. Rev. B 2006, 73, 045333. [Google Scholar] [CrossRef]
  15. Rickhaus, P.; Makk, P.; Liu, M.-H.; Tóvári, E.; Weiss, M.; Maurand, R.; Richter, K.; Schönenberger, C. Snake trajectories in ultraclean graphene p–n junctions. Nat. Commun. 2015, 6, 6470. [Google Scholar] [CrossRef]
  16. Haug, R.J. Edge-state transport and its experimental consequences in high magnetic fields. Semicond. Sci. Technol. 1993, 8, 131–153. [Google Scholar] [CrossRef]
  17. Tang, F.; Ren, Y.; Wang, P.; Zhong, R.; Schneeloch, J.; Yang, S.A.; Yang, K.; Lee, P.A.; Gu, G.; Qiao, Z.; et al. Three-dimensional quantum Hall effect and metal–insulator transition in ZrTe5. Nature 2019, 569, 537–541. [Google Scholar] [CrossRef]
  18. Delahaye, F. Series and parallel connection of multiterminal quantum Hall-effect devices. J. Appl. Phys. 1993, 73, 7914–7920. [Google Scholar] [CrossRef]
  19. Mani, R.G.; von Klitzing, K. Hall effect under null current conditions. Appl. Phys. Lett. 1994, 64, 1262–1264. [Google Scholar] [CrossRef]
  20. Mani, R.G. Transport study of GaAs/AlGaAs heterostructure- and n-type GaAs-devices in the anti Hall bar within a Hall bar configuration. J. Phys. Soc. Jpn. 1996, 65, 1751–1759. [Google Scholar] [CrossRef]
  21. Mani, R.G. Experimental technique for realizing dual and multiple Hall effects in a single specimen. Europhys. Lett. 1996, 34, 139–144. [Google Scholar] [CrossRef]
  22. Mani, R.G. Steady-state bulk current at high magnetic fields in Corbino-type GaAs/AlGaAs heterostructure devices. Europhys. Lett. 1996, 36, 203–208. [Google Scholar] [CrossRef]
  23. Oswald, M.; Oswald, J.; Mani, R.G. Voltage and current distribution in a doubly connected two-dimensional quantum Hall system. Phys. Rev. B 2005, 72, 035334. [Google Scholar] [CrossRef]
  24. Oswald, J.; Oswald, M. Magnetotransport in a doubly connected two-dimensional quantum Hall system in the low magnetic field regime. Phys. Rev. B 2006, 74, 153315. [Google Scholar] [CrossRef]
  25. Uiberacker, C.; Stecher, C.; Oswald, J. Microscopic details of the integer quantum Hall effect in an anti-Hall bar. Phys. Rev. B 2012, 86, 045304. [Google Scholar] [CrossRef]
  26. Ji, Z.; Park, H.; Barber, M.E.; Hu, C.; Watanabe, K.; Taniguchi, T.; Chu, J.-H.; Xu, X.; Shen, Z.-X. Local probe of bulk and edge states in a fractional Chern insulator. Nature 2024, 635, 578–583. [Google Scholar] [CrossRef]
  27. Ilani, S.; Martin, J.; Teitelbaum, E.; Smet, J.H.; Mahalu, D.; Umansky, V.; Yacoby, A. The microscopic nature of localization in the quantum Hall effect. Nature 2004, 427, 328–332. [Google Scholar] [CrossRef]
  28. de Picciotto, R.; Reznikov, M.; Heiblum, M.; Umansky, V.; Bunin, G.; Mahalu, D. Direct observation of a fractional charge. Nature 1997, 389, 162–164. [Google Scholar] [CrossRef]
  29. Garg, M.; Maillet, O.; Samuelson, N.L.; Wang, T.; Feng, J.; Cohen, L.A.; Watanabe, K.; Taniguchi, T.; Roulleau, P.; Sassetti, M.; et al. Enhanced shot noise in graphene quantum point contacts with electrostatic reconstruction. arXiv 2025, arXiv:2503.17209. [Google Scholar] [CrossRef]
  30. Büttiker, M. Absence of backscattering in the quantum Hall effect in multiprobe conductors. Phys. Rev. B 1988, 38, 9375–9389. [Google Scholar] [CrossRef]
  31. Datta, S. Electronic Transport in Mesoscopic Systems; Cambridge University Press: Cambridge, UK, 1997. [Google Scholar]
  32. Marks, S.; Pinard, P.; Burgess, S.; Bithell, J.; Beanland, R. Measuring the Thickness of 2D Materials Using EDS. Microsc. Microanal. 2020, 26 (Suppl. S2), 1212. [Google Scholar] [CrossRef]
  33. He, H.; Kim, K.H.; Danilov, A.; Montemurro, D.; Yu, L.; Park, Y.W.; Lombardi, F.; Bauch, T.; Moth-Poulsen, K.; Iakimov, T.; et al. Uniform doping of graphene close to the Dirac point by polymer-assisted assembly of molecular dopants. Nat. Commun. 2018, 9, 3956. [Google Scholar] [CrossRef]
  34. Yager, T.; Lartsev, A.; Cedergren, K.; Yakimova, R.; Panchal, V.; Kazakova, O.; Tzalenchuk, A.; Kim, K.-H.; Park, Y.-W.; Lara-Avila, S.; et al. Low contact resistance in epitaxial graphene devices for quantum metrology. AIP Adv. 2015, 5, 087134. [Google Scholar] [CrossRef]
  35. Janssen, T.J.B.M.; Rozhko, S.; Tzalenchuk, A.; Alexander-Webber, J.A.; Nichoas, R.J. Breakdown of the quantum Hall effect in epitaxial graphene. In Proceedings of the 29th Conference on Precision Electromagnetic Measurement (CPEM 2014), Rio de Janeiro, Brazil, 24–29 August 2014; IEEE: New York, NY, USA, 2014. [Google Scholar] [CrossRef]
  36. Haremski, P.; Mausser, M.; Gauß, A.; von Klitzing, K.; Weis, J. Electrically induced breakdown of the quantum Hall effect at different Hall bar widths: Visualizing the edge- and bulk-dominated regimes within a quantum Hall plateau. Phys. Rev. B 2020, 102, 205306. [Google Scholar] [CrossRef]
  37. Nachtwei, G. Breakdown of the quantum Hall effect. Physical E 1999, 4, 79–101. [Google Scholar] [CrossRef]
  38. Chida, K.; Hata, T.; Arakawa, T.; Matsuo, S.; Nishihara, Y.; Tanaka, T.; Ono, T.; Kobayashi, K. Avalanche electron bunching in a Corbino disk in the quantum Hall effect breakdown regime. Phys. Rev. B 2014, 89, 235318. [Google Scholar] [CrossRef]
  39. Eaves, L.; Sheard, F.W. Size-dependent quantized breakdown of the dissipationless quantum Hall effect in narrow channels. Semicond. Sci. Technol. 1986, 1, 346. [Google Scholar] [CrossRef]
  40. Streda, P.; von Klitzing, K. Critical non-dissipative current of the quantum Hall regime. J. Phys. C Solid State Phys. 1984, 17, L483. [Google Scholar] [CrossRef]
Figure 1. (a) Measurement configuration with two current sources in a Hall bar geometry. (bd) Voltage between terminals 8 and 6 measured for different current combinations: (b) I a = 100 nA and I b = 100 nA; (c) I a = 100 nA and I b = 200 nA; (d) I a = 100 nA and I b = 300 nA. (e) Voltage between terminals 2 and 8 for I a = 100 nA and I b = 100 nA. I 0 = 100 nA denotes the reference current unit. (f) Schematic of molecular doping, showing multiple layers of poly(methyl methacrylate) (PMMA) and a PMMA+ 2,3,5,6-Tetrafluoro-tetracyano-quino-dimethane (F4TCNQ) blend on top of the epitaxial graphene (denoted by G).
Figure 1. (a) Measurement configuration with two current sources in a Hall bar geometry. (bd) Voltage between terminals 8 and 6 measured for different current combinations: (b) I a = 100 nA and I b = 100 nA; (c) I a = 100 nA and I b = 200 nA; (d) I a = 100 nA and I b = 300 nA. (e) Voltage between terminals 2 and 8 for I a = 100 nA and I b = 100 nA. I 0 = 100 nA denotes the reference current unit. (f) Schematic of molecular doping, showing multiple layers of poly(methyl methacrylate) (PMMA) and a PMMA+ 2,3,5,6-Tetrafluoro-tetracyano-quino-dimethane (F4TCNQ) blend on top of the epitaxial graphene (denoted by G).
Mathematics 13 02981 g001
Figure 2. (a) Measurement configuration with two current sources. (bd) Voltage between terminals 2 and 6 measured under different conditions: (b) I a = 100 nA and I b = 100 nA; (c) I a = 100 nA and I b = 200 nA; (d) I a = 200 nA and I b = 100 nA. The reference current I 0 = 100 nA.
Figure 2. (a) Measurement configuration with two current sources. (bd) Voltage between terminals 2 and 6 measured under different conditions: (b) I a = 100 nA and I b = 100 nA; (c) I a = 100 nA and I b = 200 nA; (d) I a = 200 nA and I b = 100 nA. The reference current I 0 = 100 nA.
Mathematics 13 02981 g002
Figure 3. (a) Measurement configuration for the parallel injection case, corresponding to (b,c). (b) Voltage between terminals 8 and 6 measured with I a = 100 nA and I b = 100 nA. (c) Voltage between terminals 3 and 7 under the same parallel current conditions. (d) Measurement configuration for the antiparallel current injection case, corresponding to (e,f). (e) Voltage between terminals 8 and 6 with I a = 100 nA and I b = 100 nA. (f) Voltage between terminals 3 and 7 under the same antiparallel conditions. The reference current I 0 = 100 nA.
Figure 3. (a) Measurement configuration for the parallel injection case, corresponding to (b,c). (b) Voltage between terminals 8 and 6 measured with I a = 100 nA and I b = 100 nA. (c) Voltage between terminals 3 and 7 under the same parallel current conditions. (d) Measurement configuration for the antiparallel current injection case, corresponding to (e,f). (e) Voltage between terminals 8 and 6 with I a = 100 nA and I b = 100 nA. (f) Voltage between terminals 3 and 7 under the same antiparallel conditions. The reference current I 0 = 100 nA.
Mathematics 13 02981 g003
Figure 4. Measurement configuration where two current sources share a common drain. (a) Setup corresponding to the measurement shown in (b). (b) Voltage between terminals 2 and 3 with I a = 100 nA and I b = 100 nA. (c) Setup corresponding to the measurement shown in (d). (d) Voltage between terminals 2 and 7 with I a = 100 nA and I b = 100 nA. The reference current I 0 = 100 nA.
Figure 4. Measurement configuration where two current sources share a common drain. (a) Setup corresponding to the measurement shown in (b). (b) Voltage between terminals 2 and 3 with I a = 100 nA and I b = 100 nA. (c) Setup corresponding to the measurement shown in (d). (d) Voltage between terminals 2 and 7 with I a = 100 nA and I b = 100 nA. The reference current I 0 = 100 nA.
Mathematics 13 02981 g004
Figure 5. Schematic of current distribution in multiple current source excitation configurations. Each current from the respective source follows the same chiral edge direction, resulting in the additive behavior of the QHE.
Figure 5. Schematic of current distribution in multiple current source excitation configurations. Each current from the respective source follows the same chiral edge direction, resulting in the additive behavior of the QHE.
Mathematics 13 02981 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kim, K.H. Probing Chirality of the Quantum Hall Effect via the Landauer–Büttiker Formalism with Two Current Sources. Mathematics 2025, 13, 2981. https://doi.org/10.3390/math13182981

AMA Style

Kim KH. Probing Chirality of the Quantum Hall Effect via the Landauer–Büttiker Formalism with Two Current Sources. Mathematics. 2025; 13(18):2981. https://doi.org/10.3390/math13182981

Chicago/Turabian Style

Kim, Kyung Ho. 2025. "Probing Chirality of the Quantum Hall Effect via the Landauer–Büttiker Formalism with Two Current Sources" Mathematics 13, no. 18: 2981. https://doi.org/10.3390/math13182981

APA Style

Kim, K. H. (2025). Probing Chirality of the Quantum Hall Effect via the Landauer–Büttiker Formalism with Two Current Sources. Mathematics, 13(18), 2981. https://doi.org/10.3390/math13182981

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

Article Metrics

Back to TopTop