Abstract
Quantized longitudinal voltage drops are observed along a length of a GaAs/AlGaAs heterostructure quantum Hall effect device at applied currents large enough for the device to be in the breakdown regime. The range of currents is extensive enough to demonstrate that it is the longitudinal voltage that is quantized, rather than the longitudinal resistance. A black-box and a quasi-elastic inter-Landau level scattering (QUILLS) model are then employed to calculate the fraction of electrons making transitions into higher Landau levels, the transition rates, and the maximum electric field across the device.
Keywords: breakdown, quantized dissipation, quantized resistance states, quantized voltage states, quantum Hall effect, quasi-elastic inter-Landau level scattering, two-dimensional electron gas
1. Introduction
The integer quantum Hall effect [1–3] requires a fully quantized two-dimensional electron gas (2DEG). At low currents there is negligible dissipation within the interior of the 2DEG in the Hall plateau regions of high-quality devices. Within these regions the Hall resistance RH of the i th plateau has the value RH(i) = h/(e2i), where h is the Planck constant, e is the elementary charge, and i is an integer. At high currents, however, energy dissipation can suddenly appear [4, 5]. This is often referred to as the breakdown regime of the quantum Hall effect.
Dissipative breakdown signals can be detected by measuring longitudinal voltage differences Vx between potential probes placed along the side of a device, where x is the direction of current flow. Cage et al. [6–9] found examples where the curves of breakdown voltages Vx plotted versus magnetic flux density B were definitely quantized. It could be, however, that it is the longitudinal resistance Rx that is quantized, rather than Vx, since Rx = Vx/Ix. Indeed, Bliek et al. [10] assumed a quantized Rx in a phenomenological model to explain breakdown structures in their curves of Vx versus B for samples with narrow constrictions. Also, although not quantized, Sachrajda et al. [11] assumed magnetic field and current dependent resistive channels along the sample to explain their breakdown data.
Knowing whether Rx or Vx is quantized does matter because it can help determine what mechanism best describes the breakdown phenomena. For example, the quasi-elastic inter-Landau level scattering (QUILLS) models of Heinonen, Taylor, and Girvin [12] and Eaves and Sheard [13] assume that Vx is quantized.
It was not possible to determine which entity, Rx or Vx, was quantized in our previous experiments [6–9] because the critical currents for the advent of breakdown were high, and the range of source-drain currents ISD = Ix over which quantized breakdown occurred varied by only a few percent. It will be possible, however, to show in the present experiment that it is Vx which is quantized because the range of currents is more extensive.
A black-box model [7–9] that is based on the conservation of energy will then be used to determine the fraction of electrons making transitions between Landau levels and the transition rates. In addition, the maximum electric field across the sample will be deduced from the QUILLS model of Eaves and Sheard [13].
2. Experiment
2.1 Device
The device is a GaAs/AlxGa1−xAs heterostructure1 grown by molecular beam epitaxy at AT&T Bell Laboratories, with x = 0.29 being the fraction of aluminum atoms replacing gallium atoms in the crystal. It is designated as GaAs(7), has a zero magnetic field mobility of about 100 000 cm2/(Vs) at 1.2 K, and exhibits excellent integral quantum Hall effect properties.
The inset of Fig. 1 shows the device geometry. It is 4.6 mm long and has a width w of 0.4 mm. The two outer Hall potential probe pairs are displaced from the central pair by ±1 mm. The magnetic flux density B is perpendicular to the device and points into the figure. Electrons enter at the upper left hand corner of the device and exit at the lower right hand corner for this magnetic field direction and current. Potential probes 2, 4, and 6 are near the potential of the source S, which is grounded. Probes 1, 3, and 5 are near the drain potential D, and have a positive potential relative to the source.
Fig. 1.
Ten sweeps of Vx(6,S) versus B for the i = 2 plateau at 1.2 K with applied currents between +36 μA and +45 μA in 1 μA increments. The sweeps are in the direction of increasing B. The data have different characteristics in regions a through e. Upward arrows indicate magnetic flux density values for which calculations are made in Figs. 2 and 3. The inset shows the sample geometry and the magnetic field direction.
2.2 Longitudinal Voltage Versus Magnetic Flux Density
The dissipative voltages Vx were measured between probes 6 and S, hereafter denoted as Vx(6,S) ≡ V(6)−V(S). These two probes were chosen because breakdown occurred over a wide range of source-drain currents in this region of the device. The contact resistances were negligible, so they did not contribute to the Vx signals. Normally we would also monitor the longitudinal voltage Vx(5, S) on the opposite side of the device to assure that they were the same as Vx(6, S), but in this case the Vx(5, S) signal corresponds to integer quantum Hall voltages VH = RHISD, which were also observed on probe set VH(5, 6).
Figure 1 shows ten sweeps of Vx(6, S) versus the magnetic flux density B for the i =2 (12 906.4 Ω) quantized Hall resistance plateau at a temperature of 1.2 K for injected electron currents ISD of +36 μA to +45 μA in 1 μA increments, where positive current corresponds to electrons entering the source and exiting the drain. Distinct changes in character of the Vx signals occur in the five regions, a to e, in the figure. The signals in regions a and e have some structure, but mainly depend on the magnitude of the current. This current dependence was determined at magnetic flux density values of 10.64 T and 12.61 T (indicated in Fig. 1 by the upward arrows at the beginning and end of the sweeps) by plotting Vx versus ISD in Fig. 2. Vx increases linearly with current, and ΔVx/ΔISD is 3.38 kΩ and 2.18 kΩ at 10.64 T and 12.61 T, respectively; so regions a and e basically exhibit an ohmic behavior. The data in region b clearly show discrete, well-defined voltage states, with some switching between states, but they have very little correlation with current. The data in region c are quantized, and have a current dependence. The signals in region d are quantized, and the quantization is current-independent to within 1 % over a current range that varies by 25 %. Therefore, dissipation in the middle region of the quantized Hall resistance, which happens to be the magnetic flux density regime that has the best developed breakdown quantization, involves a quantized voltage Vx rather than a quantized resistance Rx.
Fig. 2.
Plots of Vx(6,S) versus ISD at magnetic flux density values corresponding to the beginning and the end of the sweeps if Fig. 1. The resistance values in parentheses are the slopes of the straight lines fitted to the data. The sample has an ohmic behavior at these two magnetic fields.
2.3 Critical Current
The critical current Icr for which Vx(6, S) is never zero across the magnetic field sweep is +40 μA. It was ±230 μA for Vx(2, 4) on the same device [7, 8]. This could imply that the reduced critical current for Vx(6, S) is due to an influence from the current emerging from the corner of the source, such as the heating-induced current instabilities proposed by Komiyama et al. [14]. Such a mechanism is not the reason for the reduced value of Icr for Vx(6, S), however, because Icr was still only about −44 μA for the opposite current direction. We have not found a correlation of the critical current value with the location of Vx. For instance, in a device designated as GaAs(2), Icr was about ±20 μA at one end of the device for Vx(6, S), ±87 μA for Vx(4, 6), only ±21 μA for Vx(2, 4), and the largest value ±134 μA for the other end, Vx(D,1).
3. Analysis
3.1 Transition Rates
A portion of Fig. 1 where the Vx signals are quantized is enlarged in Fig. 3, and a family of shaded curves is also displayed. These curves have equal (quantized) voltage separations at each value of magnetic flux density, but the voltage separations are allowed to vary with B in order to obtain smooth curves that provide the best fit to the data. The five shaded curves correspond to a Vx = 0.0 mV ground state and four excited states. Quantum numbers M of the voltage states are labeled in brackets. The data are current-dependent for B less than about 12.2 T, so the shaded curves are arbitrarily fitted to the 41 μA data, which is about midway in the current range.
Fig. 3.
An enlarged view of part of the data shown in Fig. 1. A family of shaded curves having equal voltage spacing at each magnetic flux density value is fitted to the 41 mA data. Voltage quantization numbers are shown in brackets. The percentages of conducting electrons making transitions to higher Landau levels are indicated for three magnetic flux density values.
We use a simple black-box model [7–9] based on energy conservation arguments to interpret some aspects of the voltage quantization displayed in Fig. 3. The dissipation detected by the Vx(6,S) signal is assumed to arise from transitions in which electrons occupying states of the originally full ground state Landau level are excited to states in higher Landau levels and then return to the lowest Landau level. There is an electrical energy loss per carrier for M Landau level transitions of Mħωc, where ωc = eB/m* is the cyclotron angular frequency and m* is the reduced mass of the electron (0.068 times the free electron mass me in GaAs). The power loss is ISDVx, and ISDVx = r(2/i)Mħωc, where r is the transition rate from the ground state to the excited state and then back to the ground state, and i is the Hall plateau number. Thus
(1) |
where f is the ratio of the transition rate r within the breakdown region to the rate ISD/e that electrons transit the device; f can also be interpreted as the fraction of conducting electrons that undergo transitions.
The black-box model predicts that about 49.4 %, 27.4 %, and 28.3 % of the conducting electrons are making inter-Landau transitions for the three magnetic flux densities selected in Fig. 3, with an uncertainty of about ±1 %. The 49.4 % value is for ISD = 41 μA, but whatever the current, large numbers of electrons seem to be making these transitions. The transition rates at 41 μA are 1.3 × 1014/s, 7.0 × 1013/s, and 7.2 × 1013/s, respectively for these three percentage values.
3.2 Maximum Electric Field
To predict the maximum value of the electric field Emax within the sample when breakdown is occurring we use the quasi-elastic inter-Landau level scattering (QUILLS) model of Heinonen, Taylor, and Girvin [12] and Eaves and Sheard [13], and the notation and coordinate system of Cage and Lavine [15]. The conducting electrons completely fill the maximum allowed number of states of the first (N = 0) Landau level. Wavefunctions of these states are represented in the Landau gauge as normalized products of Hermite polynomials across the sample in the y direction multiplied by plane waves propagating down the sample in the x direction. Each state undergoes cycloidal motion down the device and occupies a unique center of mass position y0 somewhere across the device width.
The confining potential and the applied current create an electric field distribution E(y) across the device [16]. If E(y) becomes sufficiently large in some portion of the device width then the Landau levels tilt enough to allow a population inversion, and electrons occupying eigenstates at positions y0 in the lowest Landau level N can make transitions to states of lower total energy at positions y0′ in a higher Landau level N′. Acoustic phonons are emitted in the x direction during these transitions in order to conserve energy and momentum. The electrons then emit optical phonons of total energy (N′ – N)ħωc and return to eigenstates of the initial ground state Landau level N.
We can obtain a reasonable estimate of the maximum electric field by noting that the spatial extent of the y-axis motion of the wavefunction in Ref. [15] decays rapidly beyond the turning points of a classical harmonic oscillator whose amplitude of motion is [13], where lB = (ħ/eB)1/2 is the magnetic length and the cyclotron radius of the lowest Landau level. The matrix elements of the acoustic phonon transitions become significant only when the initial and final state wavefunctions overlap [13, 17]. Transitions between the N and N′ eigenstates therefore commence when
(2) |
where N = 0 in our case for the i = 2 plateau, and N′ = M. The maximum electric field is then
(3) |
where the small contribution of the acoustic phonon transition in the numerator of Eq. (3) has been neglected.
We can use Eqs. (2) and (3) to calculate Emax at 12.25 T for the M = 1 transition of Fig. 3 (which is first excited at 40 μA). Emax is 1.1 × 106 V/m. That happens to be the same value predicted for the M = 1 transition of Vx(4,6) for the GaAs(8) device at 12.3 T and 215 μA [15]. Note that Eqs. (2) and (3) at first appear to be independent of current, but in reality the current must be increased to a unique value before M = 1 transitions are induced. Also, note that it was possible to obtain this value of Emax only because the M values could be uniquely identified in the breakdown data of Fig. 3.
The confining potential has large gradients near the device periphery, and the logarithmic charge redistribution potential, which arises from the applied current, also increases dramatically at one side of the device [16]. This side is determined by the magnetic field direction interacting with the conducting electrons. Therefore in our case, Emax is likely to be located somewhere along the side, between the source S and potential probe 6.
An Emax value of 1.1 × 106 V/m generates a large local current density Jx = σxyEmax = Emax/12 906.4 Ω = 85 A/m at ISD = 40 μA. The electron drift velocity in this region of the device is then vx = Emax/B = 8.9 × 104 m/s. This electron velocity is 36 times faster than the acoustic phonon velocity vs = 2.47 × 103 m/s [18]. The value of the acoustic phonon energy ħωs = ħMωcvs/(vx−vs) is 2.9 % of the total optical phonon energy Mħωc (which is 3.4 × 10−21 J).
4. Conclusions
It is the longitudinal voltage, Vx, rather than the longitudinal resistance, Rx, that is quantized in breakdown of the quantum Hall effect at large applied currents. Proposed mechanisms and models for the breakdown phenomena must account for, or at least not contradict, this fact. The black-box model [6–9,15] and the quasi-elastic inter-Landau level scattering model [12,13,15,17] are consistent with this observation, and lead to predictions of the transition rates and maximum electric fields within the device.
Acknowledgments
The author thanks A. C. Gossard of the University of California at Santa Barbara who made the MBE-grown GaAs/AlGaAs heterostructures while at AT&T Bell Laboratories, D. C. Tsui of Princeton University who defined the device geometries and made ohmic contacts to the 2DEG, and M. D. Stiles, K. C. Lee, R. E. Elmquist, and E. R. Williams for their discussions and comments. This work was supported in part by the Calibration Coordination Group of the Department of Defense.
Biography
About the author: Marvin E. Cage is a physicist in the Electricity Division at NIST. The National Institute of Standards and Technology is an agency of the Technology Administration, U.S. Department of Commerce.
Footnotes
Certain commercial equipment, instruments, or materials are identified in this paper to foster understanding. Such identification does not imply recommendation or endorsement by the National Institute of Standards and Technology, nor does it imply that the materials or equipment identified are necessarily the best available for the purpose.
5. References
- 1.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]
- 2.Prange RE, Girvin SM, editors. The Quantum Hall Effect. Springer-Verlag; New York: 1987. pp. 1–419. [Google Scholar]
- 3.Van Degrift CT, Cage ME, Girvin SM, editors. The Integral and Fractional Quantum Hall Effects. American Association of Physics Teachers; College Park, Maryland: 1991. pp. 1–116. [Google Scholar]
- 4.Ebert G, von Klitzing K, Ploog K, Weimann G. Two-Dimensional Magneto-Quantum Transport on GaAs/AlGaAs Heterostructures Under Non-Ohmic Conditions. J Phys C. 1983;16:5441–5448. [Google Scholar]
- 5.Cage ME, Dziuba RF, Field BF, Williams ER, Girvin SM, Gossard AC, Tsui DC, Wagner RJ. Dissipation and Dynamic Nonlinear Behavior in the Quantum Hall Regime. Phys Rev Lett. 1983;51:1374–1377. [Google Scholar]
- 6.Cage ME, Marullo Reedtz G, Yu DY, Van Degrift CT. Quantized Dissipative States at Breakdown of the Quantum Hall Effect. Semicond Sci Technol. 1990;5:L351–L354. [Google Scholar]
- 7.Cage ME. Magnetic Field Dependence of Quantized Hall Effect Breakdown Voltages. Semicond Sci Technol. 1992;7:L1119–L1122. [Google Scholar]
- 8.Cage ME. Dependence of Quantized Hall Effect Breakdown Voltage on Magnetic Field and Current. J Res Natl Inst Stand Technol. 1993;98:361–373. doi: 10.6028/jres.098.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Lavine CF, Cage ME, Elmquist RE. Spectroscopic Study of Quantized Breakdown Voltage States of the Quantum Hall Effect. J Res Natl Inst Stand Technol. 1994;99:757–764. doi: 10.6028/jres.099.068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Bliek L, Hein G, Kose V, Niemeyer J, Weimann G, Schlapp W. A New Quantum Effect in the Transverse Magnetoresistance of Two-Dimensional Conductors with a Narrow Constriction in the Conducting Channel. In: Landwehr G, editor. Proc Int Conf on the Application of High Magnetic Fields in Semiconductor Phys. Vol. 71. Springer-Verlag; Berlin: 1987. pp. 113–117. [Google Scholar]
- 11.Sachrajda AS, Landheer D, Boulet R, Moore T. Evidence for an Inhomogeneity Size Effect in Narrow GaAs/Al-GaAs Constrictions. Phys Rev B. 1989;39:10460–10463. doi: 10.1103/physrevb.39.10460. [DOI] [PubMed] [Google Scholar]
- 12.Heinonen O, Taylor PL, Girvin SM. Electron-Phonon Interactions and the Breakdown of the Dissipationless Quantum Hall Effect. Phys Rev. 1984;30:3016–3019. [Google Scholar]
- 13.Eaves L, Sheard FW. Size-Dependent Quantised Breakdown of the Dissipationless Quantum Hall Effect in Narrow Channels. Semicond Sci Technol. 1986;1:L346–L349. [Google Scholar]
- 14.Komiyama S, Takamasu T, Hiyamizu S, Sasa S. Breakdown of the Quantum Hall Effect Due to Electron Heating. Solid State Commun. 1985;54:479–484. [Google Scholar]
- 15.Cage ME, Lavine CF. Using Quantized Breakdown Voltage Signals to Determine the Maximum Electric Fields in a Quantum Hall Effect Sample. J Res Natl Inst Stand Technol. 1995;100:269–276. doi: 10.6028/jres.100.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Cage ME, Lavine CF. Potential and Current Distributions Calculated Across a Quantum Hall Effect Sample at Low and High Currents. J Res Natl Inst Stand Technol. 1995;100:529–541. doi: 10.6028/jres.100.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Cage ME, Yu DY, Marullo G. Reedtz, Observation and an Explanation of Breakdown of the Quantum Hall Effect. J Res Natl Inst Stand Technol. 1990;95:93–99. doi: 10.6028/jres.095.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.McSkimin HJ, Jayaraman A, Andreatch PJ. Elastic Moduli of GaAs at Moderate Pressures and the Evaluation of Compression to 250 kbar. J Appl Phys. 1967;38:2362–2364. [Google Scholar]