Zeeman splitting of single semiconductor impurities in resonant tunneling heterostructures

Similar documents
Sequential tunneling and spin degeneracy of zero-dimensional states

Electron-spectroscopic study of vertical In 1 x Ga x As quantum dots

Resonant tunneling in double-quantum-well triple-barrier heterostructures

Formation of unintentional dots in small Si nanostructures

ERRATA. Observation of the Local Structure of Landau Bands in a Disordered Conductor [Phys. Rev. Lett. 78, 1540 (1997)]

Fano resonances in transport across a quantum well in a tilted magnetic field

2005 EDP Sciences. Reprinted with permission.

Surfaces, Interfaces, and Layered Devices

Universal valence-band picture of. the ferromagnetic semiconductor GaMnAs

Semiconductor device structures are traditionally divided into homojunction devices

Physics of Semiconductors

Ballistic Electron Spectroscopy of Quantum Mechanical Anti-reflection Coatings for GaAs/AlGaAs Superlattices

Graphene and Carbon Nanotubes

Saroj P. Dash. Chalmers University of Technology. Göteborg, Sweden. Microtechnology and Nanoscience-MC2

1. Binary III-V compounds 2 p From which atoms are the 16 binary III-V compounds formed?...column III B, Al, Ga and In...column V N, P, As and Sb...

A spin Esaki diode. Makoto Kohda, Yuzo Ohno, Koji Takamura, Fumihiro Matsukura, and Hideo Ohno. Abstract

Single Electron Tunneling Through Discrete Semiconductor Impurity States A Dissertation Presented to the Faculty of the Graduate School of Yale Univer

Supplementary Information for Pseudospin Resolved Transport Spectroscopy of the Kondo Effect in a Double Quantum Dot. D2 V exc I

SUPPLEMENTARY INFORMATION

Terahertz Lasers Based on Intersubband Transitions

Current-driven Magnetization Reversal in a Ferromagnetic Semiconductor. (Ga,Mn)As/GaAs/(Ga,Mn)As Tunnel Junction

Low-dimensional electron transport properties in InAs/AlGaSb mesoscopic structures

Correlated 2D Electron Aspects of the Quantum Hall Effect

Nuclear spin spectroscopy for semiconductor hetero and nano structures

Defense Technical Information Center Compilation Part Notice

Lecture 18: Semiconductors - continued (Kittel Ch. 8)

Spring Semester 2012 Final Exam

A Tunable Kondo Effect in Quantum Dots

Electron Energy, E E = 0. Free electron. 3s Band 2p Band Overlapping energy bands. 3p 3s 2p 2s. 2s Band. Electrons. 1s ATOM SOLID.

SUPPLEMENTARY INFORMATION

On depolarisation in 0D systems: Lamb-like level shift

File name: Supplementary Information Description: Supplementary Figures and Supplementary References. File name: Peer Review File Description:

Charging and Kondo Effects in an Antidot in the Quantum Hall Regime

Session 6: Solid State Physics. Diode

SUPPLEMENTARY INFORMATION

Determination of the tunnel rates through a few-electron quantum dot

Upper-barrier excitons: first magnetooptical study

Lecture contents. Burstein shift Excitons Interband transitions in quantum wells Quantum confined Stark effect. NNSE 618 Lecture #15

Chapter 3 Properties of Nanostructures

1 Name: Student number: DEPARTMENT OF PHYSICS AND PHYSICAL OCEANOGRAPHY MEMORIAL UNIVERSITY OF NEWFOUNDLAND. Fall :00-11:00

Laser Diodes. Revised: 3/14/14 14: , Henry Zmuda Set 6a Laser Diodes 1

Electronic states of self-organized InGaAs quantum dots on GaAs (3 1 1)B studied by conductive scanning probe microscope

The Semiconductor in Equilibrium

Lecture 20: Semiconductor Structures Kittel Ch 17, p , extra material in the class notes

Tunable Non-local Spin Control in a Coupled Quantum Dot System. N. J. Craig, J. M. Taylor, E. A. Lester, C. M. Marcus

single-electron electron tunneling (SET)

Semiconductor Physics fall 2012 problems

Large Storage Window in a-sinx/nc-si/a-sinx Sandwiched Structure

Few-electron molecular states and their transitions in a single InAs quantum dot molecule

Luminescence basics. Slide # 1

Tunneling transport. Courtesy Prof. S. Sawyer, RPI Also Davies Ch. 5

Effects of Quantum-Well Inversion Asymmetry on Electron- Nuclear Spin Coupling in the Fractional Quantum Hall Regime

Coulomb-interaction induced incomplete shell filling in the hole system of InAs quantum dots

Tunnelling spectroscopy of low-dimensional states

Physics of Semiconductors (Problems for report)

Electron spins in nonmagnetic semiconductors

MSE 310/ECE 340: Electrical Properties of Materials Fall 2014 Department of Materials Science and Engineering Boise State University

Kondo effect in multi-level and multi-valley quantum dots. Mikio Eto Faculty of Science and Technology, Keio University, Japan

Self-consistent analysis of the IV characteristics of resonant tunnelling diodes

Enhancement-mode quantum transistors for single electron spin

Defense Technical Information Center Compilation Part Notice

Self-Assembled InAs Quantum Dots

arxiv: v1 [cond-mat.mes-hall] 20 Sep 2013

UNIT - IV SEMICONDUCTORS AND MAGNETIC MATERIALS

Quantum confinement induced by strain relaxation in an elliptical double-barrier Si/Si x Ge 1 x resonant tunneling quantum dot

Electronic transport in low dimensional systems

PHYSICS OF SEMICONDUCTORS AND THEIR HETEROSTRUCTURES

Three-Dimensional Silicon-Germanium Nanostructures for Light Emitters and On-Chip Optical. Interconnects

The effect of surface conductance on lateral gated quantum devices in Si/SiGe heterostructures

Charge spectrometry with a strongly coupled superconducting single-electron transistor

Quantum Condensed Matter Physics Lecture 12

KATIHAL FİZİĞİ MNT-510

Supplementary Information

(b) Spontaneous emission. Absorption, spontaneous (random photon) emission and stimulated emission.

Semiconductor Physics Problems 2015

Spin relaxation of conduction electrons Jaroslav Fabian (Institute for Theoretical Physics, Uni. Regensburg)

Resonant Tunnel Diode (RTD) Stability

Electron energy levels in superconducting and magnetic nanoparticles

Surfaces, Interfaces, and Layered Devices

MOS CAPACITOR AND MOSFET

Defense Technical Information Center Compilation Part Notice

Quantum Condensed Matter Physics Lecture 17

Appendix 1: List of symbols

Semiconductor Nanowires: Motivation

SPINTRONICS. Waltraud Buchenberg. Faculty of Physics Albert-Ludwigs-University Freiburg

Self-assembled SiGe single hole transistors

Solving rate equations for electron tunneling via discrete quantum states

Experimental discovery of the spin-hall effect in Rashba spin-orbit coupled semiconductor systems

Anisotropic spin splitting in InGaAs wire structures

Semiconductor Physics and Devices

Spin Orbit Coupling (SOC) in Graphene

Fundamentals of the Metal Oxide Semiconductor Field-Effect Transistor

Ge/Si Photodiodes with Embedded Arrays of Ge Quantum Dots for the Near Infrared ( mm) Region

R. Ludwig and G. Bogdanov RF Circuit Design: Theory and Applications 2 nd edition. Figures for Chapter 6

Electronic and Optoelectronic Properties of Semiconductor Structures

solidi current topics in solid state physics InAs quantum dots grown by molecular beam epitaxy on GaAs (211)B polar substrates

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Electrical Engineering And Computer Science Semiconductor Optoelectronics Fall 2002

Negative differential conductance and current bistability in undoped GaAs/ Al, Ga As quantum-cascade structures

Electrical Control of Single Spins in Semiconductor Quantum Dots Jason Petta Physics Department, Princeton University

Classification of Solids

Transcription:

Superlattices and Microstructures, Vol. 2, No. 4, 1996 Zeeman splitting of single semiconductor impurities in resonant tunneling heterostructures M. R. Deshpande, J. W. Sleight, M. A. Reed, R. G. Wheeler Departments of Physics, Applied Physics, and Electrical Engineering, Yale University, P.O. Box 2812, New Haven, CT 652, U.S.A. R. J. Matyi Central Research Laboratories, Texas Instruments Incorporated, Dallas, TX 75265, U.S.A. (Received 2 May 1996) Zeeman splitting of the ground state of single impurities in the quantum wells of resonant tunneling heterostructures is reported. We determine the absolute magnitude of the effective magnetic spin splitting factor g for a single impurity in a 44 Å Al.27 Ga.73 As/GaAs/Al.27 Ga.73 As quantum well to be.28 ±.2. This system also allows for independent measurement of the electron tunneling rates through the two potential barriers and estimation of the occupation probability of the impurity state in the quantum well. c 1996 Academic Press Limited Key words: resonant tunneling heterostructures, spin g factor, electron tunneling rates. 1. Introduction The experimental realization of granular electronic systems, such as low dimensional semiconductor and ultra small metallic systems, has focused attention on the basic physical properties of discrete systems. Especially intriguing are the semiconductor quantum dot [1 5] and the physically similar localized impurity state tunneling systems [6 11]. The latter provides a unique laboratory for the study of a single impurity. We present here the first observation of Zeeman spin splitting of a single semiconductor impurity, studied by tunneling transport. The measured g factor provides an important test of the band theory of confined semiconductor systems, and our method is particularly important for well widths less than 5 Å where other methods are less precise [12 14]. Isolated donor impurities in the quantum well regions of large area resonant tunneling heterostructures form localized ( 1 Å) hydrogenic states bound to the quantum eigenstates. Figure 1 illustrates a band diagram of the structure used in this study with one impurity state in the well schematically noted. Under applied bias, as the impurity state aligns with the emitter Fermi level, the current exhibits a step-like increase (Fig. 2). In general, there may be multiple impurities giving rise to multiple, overlapping steps in the current voltage (I (V )) Present address: Digital Semiconductor, Hudson, MA, U.S.A. Present address: Department of Materials Science and Engineering, University of Wisconsin at Madison, WI, U.S.A. 749 636/96/8513 + 1 $25./ c 1996 Academic Press Limited

514 Superlattices and Microstructures, Vol. 2, No. 4, 1996 Energy (mev) 2 2 2 4 6 8 Position Z (nm) Fig. 1. Model conduction band diagram of the device, at an applied bias of 1 mv, showing the quantum eigenstate (long line) and a localized impurity state (short line) in the well. The dotted line represents the Fermi level in the leads. 1..8 35 mk Current (na).6.4.2. 144 146 148 15 152 154 156 158 Bias voltage (mv) Fig. 2. I (V ) characteristics (zero magnetic field, T mix = 35 mk) of the quantum well device (w = 8.5 nm) showing the pre-threshold region with sharp step-like structures. The various current steps are attributed to tunneling through separate impurity states in the well. characteristics. An appropriately dilute, unintentional doping concentration gives rise to isolated, uncorrelated impurities, allowing measurement of a single one. In a magnetic field the spin degeneracy of the impurity ground state is broken, which results in a splitting of the current step in the I (V ) characteristics. The magnitude of the spin g factor can be determined from the voltage difference between the two fragments of the spin split step. The current magnitudes of the two fragments enable us, for the first time, to independently determine the electron tunneling rates through the two potential barriers in a sequential tunneling picture [4, 15]. 2. Sample design The nominally symmetric resonant tunneling heterostructures are grown by molecular beam epitaxy (MBE) on a Si-doped GaAs (1) substrate [16]. The epitaxial layers consist of a 1.8 1 18 cm 3 Si-doped GaAs

Superlattices and Microstructures, Vol. 2, No. 4, 1996 515 contact, a 15 nm undoped GaAs spacer layer, an undoped Al.27 Ga.73 As bottom barrier of width w, a 4.4 nm undoped GaAs quantum well, an undoped Al.27 Ga.73 As top barrier of the same width w, a 15 nm undoped GaAs spacer layer, and a 1.8 1 18 cm 3 Si doped GaAs top contact. We study devices with two different barrier widths (w = 8.5 nm and 6.5 nm). Square mesas with lateral dimensions from 2 µm to64µm are fabricated using standard photolithography techniques. Two terminal I (V ) characteristics are measured in a dilution refrigerator with a mixing chamber temperature (T mix ) of 35 mk. 3. Observations 3.1. Tunneling through impurities Figure 2 shows the I (V ) characteristics for a typical device (64 µm 2, T mix = 35 mk, w = 8.5 nm) showing the sharp steps in the I (V ) characteristics. This step structure is observed to be sample specific, but for a given sample it is exactly reproducible. Similar steps are observed in other devices with different barrier thicknesses. The step current magnitudes are observed to be of the order of e/τ (e is the electronic charge and τ is the lifetime of the quantum well state as estimated from a bandprofile model [9]). This suggests that the current steps are due to single electron tunneling through individual, zero dimensional states in the quantum well. Similar features have been observed and reported previously [6, 7], and the various current steps are attributed to tunneling through bound states of separate impurities in the quantum well. 3.2. Spin splitting Figure 3 shows an expanded view of the first current step edge at 35 mk for forward and reverse bias at T and at 9 T. The magnetic field is oriented perpendicular to the current direction (parallel to the heterointerfaces). At zero field, the ground state of the impurity is spin degenerate leading to a single current step. Upon lifting of the degeneracy at finite field, a splitting of the current step is observed. Figure 4 shows the dependence of the I (V ) characteristics upon increasing magnetic field. Note that the spin splitting increases as the field is swept from T (bottom) to 9 T (top). Figure 5 shows the voltage separation between the corresponding two spin-split conductance peaks which increases linearly with the magnetic field strength as expected. Due to the finite widths of the conductance peaks, it is not possible to resolve the splitting for magnetic fields less than 5.5 T. The best line fit to the data (Fig. 5) closely intersects V = atb = and has a slope g µ B/α, where µ B is the Bohr magneton, g is the effective gyromagnetic ratio of the impurity with the magnetic field perpendicular to the growth direction of the quantum well and α is the voltage to energy conversion factor [4, 5, 8, 11]. 3.2.1. Determination of α To determine the spin g factor it is necessary to know the voltage to energy conversion factor, α, which is accurately determined by studying the thermal broadening of the current step. The current through a localized impurity state in the quantum well, at bias V and temperature T, depends upon the density of occupied electronic states in the emitter at the same energy as the impurity state, which is proportional to the Fermi distribution function f. As shown in Fig. 6, the current plateau edges exhibit characteristic Fermi level thermal broadening. We can express the current as I (V, T ) = 2I th f(v,t) = 2I th 1 + exp[αe(v th V)/kT] (1) Splitting is also observed when the magnetic field is parallel to the current direction. We restrict this paper to the case of magnetic field perpendicular to current.

516 Superlattices and Microstructures, Vol. 2, No. 4, 1996 Forward bias Reverse bias 6 T 9 T 6 I 2 Current (pa) 4 I 2 Current (pa) 4 I 1 2 I 1 2 93. 94. Voltage (mv) 95. 12. 13. 14. Voltage (mv) Fig. 3. I (V ) characteristics at T mix = 35 mk of the first current step edge in forward bias (left) and reverse bias (right) at T( ) and 9 T ( ). I 1 and I 2 mark the current values at9tasshown.i 1 gives the current of the first fragment while I 2 is the net current of both fragments of the split step edge. 5 4 Current (pa) 3 2 1 93. 93.5 94. Voltage (mv) 94.5 95. Fig. 4. I (V ) characteristics (T mix = 35 mk) of the first current step edge in forward bias of the 8.5 nm barrier device in increasing magnetic fields from T (bottom) to 9 T (top). The curves are vertically offset for clarity.

Superlattices and Microstructures, Vol. 2, No. 4, 1996 517.4 + Reverse bias Forward bias Spin spliting V (mv).3.2.1 6 8 Magnetic field (T) 1 Fig. 5. The experimental spin splitting versus magnetic field perpendicular to current for the first current step of the 8.5 nm barrier device in forward and reverse bias at T mix = 35 mk. The solid lines are linear fits to the data. where e is the electron charge, k is Boltzmann s constant, and V th and I th are the threshold voltage and current at the observed common intersection point of the various I (V ) curves at different temperatures. The only free parameter α, is determined from a fit (Fig. 6) of the above Fermi function to the I (V ) traces at zero magnetic field. The fits are done for data taken at different temperatures, from.5 to 5 K, and α is determined for each. The average value of α along with the rms error is reported. For the device with 8.5 nm barrier α =.48±.2 for forward bias and.42 ±.2 for reverse bias. For the device with 6.5 nm barrier α =.4 ±.2 for forward bias. The fits are done for the region V V th when the current is small and not affected by the occupancy of the impurity state. 3.2.2. Determination of g factor From these values of α and the measured spin splitting in voltage, the absolute magnitude of the spin g value of the impurities is determined. For the device with 8.5 nm barrier we determine the g factor for two separate impurity steps in both forward and reverse bias directions. We obtain for the first impurity, g = (.28 ±.2,.28 ±.2) (for forward and reverse bias, respectively) and for the second impurity, g = (.28 ±.2,.27 ±.2). For the device with 6.5 nm barrier we study only one impurity in forward bias for which we get g = (.27 ±.2). The g value of electrons in bulk GaAs (.44) is different from the free electron value (+2) because of the spin orbit interaction in the valence band and its effect on the conduction band as analyzed using k p theory. In confined quantum well regions, it changes further and also becomes anisotropic. This is partly caused by the change in the electron and hole energies due to confinement, and partly due to the electron wavefunction penetration in the barrier material where the electron has a higher g value (+.39). g is theoretically predicted to be a strong function of the quantum well width and changes from the bulk value

518 Superlattices and Microstructures, Vol. 2, No. 4, 1996 Current (pa) 5 4 3 2 +.8 K 1. K 1.4 K 2. K 3. K + 4. K 1 94 95 96 97 Voltage (mv) Fig. 6. I (V ) characteristics of the first impurity current step edge of the 8.5 nm barrier device at different temperatures showing the Fermi level broadening and the Fermi fit to these I (V ) traces. of.44 to greater than +.4 for well widths less than 3 Å [12]. Our measurement is consistent with this prediction and with recent experimental results [13, 14] (assuming here that the sign of g is positive). Due to the 3D nature of the emitter, at any given energy electrons of both spin orientations are available for tunneling (even with applied magnetic field). Figure 7 shows thermal broadening of the first current step edge split at 9 T. Separate Fermi broadening of the two fragments of the split step indicates that in both cases electrons involved in tunneling are the Fermi electrons in the emitter. The observed experimental voltage difference is thus entirely due to the spin splitting of the impurity state, and is not affected by the g -factor of the electrons in the emitter. We also note that electron electron interactions will not alter the relative energy of the two spin states and thus the measured g factor. Coulomb interaction between the tunneling electron and the electrons in the emitter causes a rise in the step current near the threshold at low temperatures, which is observed (as first reported by Geim et al. [8]), but it should not affect the energy separation of the two spin states. 3.3. Determination of tunneling rates Analysis of the tunneling current through the spin split system yields additional information about the tunneling process. Recalling Fig. 3, the two fragments of the spin split step at 9 T are observed not to have the same current magnitude ((I 2 I 1 ) I 1 ). This difference is more prominent in reverse bias. To analyze this, we define Ɣ b and Ɣ t to be the tunneling rates for an electron to tunnel through the bottom and the top (referred to growth direction) potential barriers, respectively, of the double barrier heterostructure. We also define Ɣ cl and Ɣ em to be the electron tunneling rates through the collector (downstream) and emitter (upstream) (referred to electron flow direction) potential barriers. In forward bias, electron injection is through the top barrier into If the Fermi energy in the emitter is small and weakly localized states are formed in the emitter, the g -factor of these weakly localized states would have to be taken into account as reported by Sakai et al. [1].

Superlattices and Microstructures, Vol. 2, No. 4, 1996 519 3 Current (pa) 2 1 35 mk 1 mk 2 mk 3 mk 93.8 94. 94.2 Voltage (mv) 94.4 94.6 Fig. 7. I (V ) characteristics at 9 T showing the spin split first current step edge in forward bias at mixing chamber temperatures ranging from 35 to 3 mk. the quantum well (Ɣ em Ɣ t,ɣ cl Ɣ b ), while in reverse bias, electron injection is through the bottom barrier into the quantum well (Ɣ em Ɣ b,ɣ cl Ɣ t ). We also define p to be the occupation probability for an electron in a localized state in the quantum well. In the sequential tunneling picture, p = f Ɣ em /(Ɣ em +Ɣ cl ). Since the spin splitting energy ( E 15 µev) is much smaller than the barrier potential energy ( 3 mev), we assume that the tunneling rates are the same for electrons tunneling through the spin up or the spin down states irrespective of the slight energy difference. We also assume that the emitter electrons are not spin polarized which is a good assumption since the Fermi energy ( 4 mev) is much larger than the spin splitting energy even at 1 T. In high magnetic fields and at low temperatures, when the Fermi level is sharp, (Fig. 7), we can adjust the bias to have the following two conditions for a given impurity. At lower bias, only the lower energy spin state channel is active for conduction, and the current is given by I 1 = peɣ cl. (2) At higher bias, the higher energy spin state channel is also active for conduction, and the current is given by I 2 = p eɣ cl = (2p p 2 )eɣ cl, (3) where p = (2p p 2 ) is the probability of occupying either the lower or the higher state, but not both of them. Both states cannot be occupied at the same time due to the large Coulomb energy required for another electron to simultaneously occupy the second state. For this system the single electron Coulomb charging energy (U C = e 2 /2C where C is the effective capacitance of the double barrier device ) is much larger than C = (ɛ kπr 2)(d 1 t + d 1 b ), where k is the dielectric constant, d t and d b are the top and bottom barrier thicknesses and r is the radius of the lateral region affected by the localized impurity state in the quantum well. We take r as the Bohr radius of hydrogenic impurities in GaAs ( 1 Å) and obtain U C 9 mev. Although taking r as the Bohr radius is only an approximation, we obtain U C E even for r seven times larger.

52 Superlattices and Microstructures, Vol. 2, No. 4, 1996 8 Γ cl Γ em Tunneling rate (MHz) 6 4 2 6 7 8 9 1 Magnetic field (T) 11 Fig. 8. Tunneling rates Ɣ em and Ɣ cl as a function of the magnetic field strength perpendicular to current for the 8.5 nm barrier device in forward bias orientation. the spin splitting energy E. Here we have assumed that the tunneling rates are the same at the two different biases which is a good assumption since the bias difference is much smaller than the barrier potential. In the extreme limits these equations indicate that for Ɣ cl Ɣ em, p and I 2 2I 1 while for Ɣ cl Ɣ em, p 1 and I 2 I 1. This qualitatively explains the behaviour observed in Fig. 3. To get a quantitative understanding, we solve equations (2) and (3), using I 1 and I 2, to determine p. For the first impurity at 9 T, p =.21 for forward bias and p =.62 for reverse bias. A high p value indicates that the electron tunneling rate through the collector (downstream) barrier (Ɣ cl ) is lower than that through the emitter (upstream) barrier (Ɣ em ) causing an accumulation in the well. A higher p value for reverse bias (as compared to forward bias) suggests an asymmetry in the heterostructure growth with one barrier being slightly thicker than the other barrier. In forward bias the top barrier is the emitter barrier while in reverse bias the top barrier is the collector barrier. This implies that the top barrier is slightly thicker than the bottom barrier. This is consistent with the difference in the measured α values, observed asymmetry in the I (V ) characteristics for this sample (the main resonance peak voltage and peak current values are higher for reverse bias as compared to forward bias), and is in agreement with previous characterization [16]. We also obtain the absolute magnitude of the electron tunneling rates through the two potential barriers and study their dependence upon the magnetic field perpendicular to the current direction. Figures 8 and 9 show the tunneling rates in forward bias and reverse bias orientations respectively. Note that Ɣ em is smaller than Ɣ cl in forward bias orientation while Ɣ cl is smaller than Ɣ em in reverse bias orientation. This is because p is smaller than.5 in forward bias while it is larger than.5 in reverse bias. As mentioned before this asymmetry suggests that the top barrier of the heterostructure is slightly thicker than the bottom barrier. Despite the asymmetry, we observe that in either orientation Ɣ em decreases with field strength while Ɣ cl is approximately constant. Ɣ em roughly decreases by a factor of two as the field increases from 6 to 11 T in both bias orientations. (The oscillations in the tunneling rate are probably due to the fine structure observed on the current plateaus and its movement in magnetic field [9] not reported in this paper.) The suppression of Ɣ em

Superlattices and Microstructures, Vol. 2, No. 4, 1996 521 8 Γ em Γ cl Tunneling rate (MHz) 6 4 2 6 7 8 9 1 Magnetic field (T) 11 Fig. 9. Tunneling rates Ɣ em and Ɣ cl as a function of the magnetic field strength perpendicular to current for the 8.5 nm barrier device in reverse bias orientation. results in the observed plateau current suppression with magnetic field [17]. This indicates that the magnetic field affects the emitter-to-well tunneling process (Ɣ em ) more than it affects the well-to-collector tunneling process (Ɣ cl ). We therefore expect more current suppression when the current is emitter barrier limited than when it is collector barrier limited, which is observed (Fig. 3). For forward bias (thicker emitter) the current at 9 T is suppressed by a factor of 41% (as compared to T), while the suppression is only 2% for reverse bias (thicker collector). We have presented a g measurement for a given epitaxial structure. This method can be used for a precise determination of g as a function of quantum well width and magnetic field orientation to compare to theory. This technique generalizes to other measurements of interest such as g (B) dependence, the investigation of magnetic impurities, different heterojunction material systems, and the determination of tunneling rates for bandgap engineered structures of interest. Acknowledgements We thank Prof. D. E. Prober and Dr. M. Amman for many useful discussions, C. L. Fernando and Prof. W. R. Frensley for help with the modelling, and A. Mittal for experimental assistance. This work is supported by NSF DMR-9112497 and 9216121. References [1] M. A Reed et al. Phys. Rev. Lett. 6, 535 (1988). [2] S. Tarucha, Y. Hirayama, T. Saku and T. Kimura, Phys. Rev. B41, 5459 (199). [3] M. Teowordt et al. Phys. Rev. B463948 (1992). [4] B. Su, V. J. Goldman and J. E. Cunningham, Phys. Rev. B46, 7644 (1992). [5] J. W. Sleight et al. Phys. Rev. B53, 15727 (1996). [6] M. W. Dellow et al. Phys. Rev. Lett. 68, 1754 (1992); J. W. Sakai et al. Appl. Phys. Lett. 64, 2563 (1994). [7] M. R. Deshpande et al. Proceedings of the IEEE/Cornell Conference on Advanced Concepts in High Speed Semiconductor Devices and Circuits, p. 177 (1993).

522 Superlattices and Microstructures, Vol. 2, No. 4, 1996 [8] A. K. Geim et al. Phys. Rev. Lett. 72, 261 (1994). [9] M. R. Deshpande et al. in 22 International Conference on the Physics of Semiconductors, Edited by D. J. Lockwood, Singapore: World Scientific p 1899 (1995). [1] J. W. Sakai et al. Solid-State Electronics 37, 965 (1994). [11] M. R. Deshpande et al. Phys. Rev. Lett. 76, 1328 (1996). [12] E. L. Ivchenko and A. A. Kiselev, Sov. Phys. Semicond. 26, 827 (1992). [13] M. J. Snelling et al. Phys. Rev. B44, 11345 (1991); M. J. Snelling et al. Phys. Rev. B45, 3922 (1992). [14] R. M. Hannak et al. Solid State Commun. 93, 313 (1995). [15] We assume sequential tunneling because the barriers are thick (85 Å). S. Luryi, Appl. Phys. Lett. 47, 49 (1985). [16] M. A. Reed et al. Appl. Phys. Lett. 54, 1256 (1989). [17] J. W. Sakai et al. Phys. Rev. B48, 5664 (1993).