Decay of spin polarized hot carrier current in a quasi. one-dimensional spin valve structure arxiv:cond-mat/ v1 [cond-mat.mes-hall] 10 Oct 2003

Similar documents
Datta-Das type spin-field effect transistor in non-ballistic regime

Spin Transport in III-V Semiconductor Structures

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

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

UTILIZATION of spin-dependent phenomena in electronic

Supplementary material: Nature Nanotechnology NNANO D

Anisotropic spin splitting in InGaAs wire structures

Transient grating measurements of spin diffusion. Joe Orenstein UC Berkeley and Lawrence Berkeley National Lab

Spin Lifetime Enhancement by Shear Strain in Thin Silicon-on-Insulator Films. Dmitry Osintsev, Viktor Sverdlov, and Siegfried Selberherr

Spring 2009 EE 710: Nanoscience and Engineering

Effect of Spin-Orbit Interaction and In-Plane Magnetic Field on the Conductance of a Quasi-One-Dimensional System

Spin-Orbit Interactions in Semiconductor Nanostructures

Optically induced Hall effect in semiconductors

Persistent spin helix in spin-orbit coupled system. Joe Orenstein UC Berkeley and Lawrence Berkeley National Lab

Semiclassical Monte Carlo Model for In-Plane Transport of Spin-Polarized Electrons in III-V Heterostructures

Physics of Semiconductors

Citation for published version (APA): Filip, A. T. (2002). Spin polarized electron transport in mesoscopic hybrid devices s.n.

arxiv: v2 [cond-mat.stat-mech] 17 Dec 2011

Non-equilibrium Green s functions: Rough interfaces in THz quantum cascade lasers

Monte Carlo modeling of spin FETs controlled by spin orbit interaction

Semiclassical Transport Models for Semiconductor Spintronics

Hot-phonon effects on electron transport in quantum wires

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

Theory of electrical spin injection: Tunnel contacts as a solution of the conductivity mismatch problem

QUANTUM INTERFERENCE IN SEMICONDUCTOR RINGS

Lecture I. Spin Orbitronics

arxiv:cond-mat/ v1 25 Mar 2004

Limitations in the Tunability of the Spin Resonance of 2D Electrons in Si by an Electric Current

arxiv: v2 [cond-mat.mes-hall] 14 Jul 2009

Conserved Spin Quantity in Strained Hole Systems with Rashba and Dresselhaus Spin-Orbit Coupling

Chapter 3 Properties of Nanostructures

Carrier Dynamics in Quantum Cascade Lasers

Spin-orbit Effects in Semiconductor Spintronics. Laurens Molenkamp Physikalisches Institut (EP3) University of Würzburg

wire z axis Under these assumptions, if we model the electrons by plane waves in the z direction we get n E, n, 1,2,

QUANTUM WELLS, WIRES AND DOTS

arxiv:cond-mat/ v1 [cond-mat.mes-hall] 14 Mar 2006

arxiv:cond-mat/ v2 [cond-mat.mtrl-sci] 6 Dec 2002

Tunable All Electric Spin Polarizer. School of Electronics and Computing Systems University of Cincinnati, Cincinnati, Ohio 45221, USA

Spins and spin-orbit coupling in semiconductors, metals, and nanostructures

ELECTRONS AND PHONONS IN SEMICONDUCTOR MULTILAYERS

Narrow-Gap Semiconductors, Spin Splitting With no Magnetic Field and more.. Giti Khodaparast Department of Physics Virginia Tech

Spin Filtering: how to write and read quantum information on mobile qubits

Electronic and Optoelectronic Properties of Semiconductor Structures

All-electrical measurements of direct spin Hall effect in GaAs with Esaki diode electrodes.

Spin Dynamics in Semiconductors, Chapter 4 of Semiconductor Spintronics and Quantum Computation edited by D. D. Awschalom, D. Loss, and N. Samarth.

SEMICONDUCTOR SPINTRONICS FOR QUANTUM COMPUTATION

Fundamental concepts of spintronics

Spin scattering by dislocations in III-V semiconductors

A -SiC MOSFET Monte Carlo Simulator Including

Classification of Solids

Directions for simulation of beyond-cmos devices. Dmitri Nikonov, George Bourianoff, Mark Stettler

Spin injection and the local Hall effect in InAs quantum wells

Spin-resolved Hall effect driven by spin-orbit coupling. Physical Review B - Condensed Matter And Materials Physics, 2005, v. 71 n.

Spin transverse force on spin current in an electric field

arxiv:cond-mat/ v2 [cond-mat.mes-hall] 11 Jul 2001

arxiv:cond-mat/ v1 [cond-mat.mes-hall] 27 Nov 2001

White Rose Research Online URL for this paper:

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

Spin Transport in Organic Semiconductors: A Brief Overview of the First Eight Years

MONTE CARLO MODELING OF HEAT GENERATION IN ELECTRONIC NANOSTRUCTURES

PHYSICAL REVIEW B 71,

Magnetostatic modulation of nonlinear refractive index and absorption in quantum wires

Modeling for Semiconductor Spintronics

Mesoscopic Spintronics

Electro-Thermal Transport in Silicon and Carbon Nanotube Devices E. Pop, D. Mann, J. Rowlette, K. Goodson and H. Dai

SUPPLEMENTARY INFORMATION

arxiv:cond-mat/ v1 [cond-mat.mes-hall] 3 Feb 1999

Surfaces, Interfaces, and Layered Devices

NOVEL GIANT MAGNETORESISTANCE MODEL USING MULTIPLE BARRIER POTENTIAL

Detectors of the Cryogenic Dark Matter Search: Charge Transport and Phonon Emission in Ge 100 Crystals at 40 mk

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

Energy dispersion relations for holes inn silicon quantum wells and quantum wires

Electric current-induced spin orientation in quantum well structures

Interference of magnetointersubband and phonon-induced resistance oscillations in single GaAs quantum wells with two populated subbands

ELECTRONS AND PHONONS IN SEMICONDUCTOR MULTILAYERS

Terahertz Lasers Based on Intersubband Transitions

Rashba coupling in quantum dots: exact solution

Spin orbit torque driven magnetic switching and memory. Debanjan Bhowmik

Spintronics: a step closer to the "The Emperor's New Mind" Ferenc Simon

NUMERICAL CALCULATION OF THE ELECTRON MOBILITY IN GaAs SEMICONDUCTOR UNDER WEAK ELECTRIC FIELD APPLICATION

Semiconductor Physics and Devices Chapter 3.

arxiv:cond-mat/ v1 [cond-mat.mes-hall] 20 Jan 2004

Spin and charge transport in the presence of spin-orbit interaction

Mesoscopic Spin Hall Effect in Multiprobe Semiconductor Bridges

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

Transverse spin-orbit force in the spin Hall effect in ballistic semiconductor wires

Solid-State Electronics

Minimal Update of Solid State Physics

Spin caloritronics in magnetic/non-magnetic nanostructures and graphene field effect devices Dejene, Fasil

The effect of light illumination in photoionization of deep traps in GaN MESFETs buffer layer using an ensemble Monte Carlo simulation

Max-Planck-Institut für Metallforschung Stuttgart. Towards Spin Injection into Silicon. Saroj Prasad Dash. Dissertation an der Universität Stuttgart

ELECTRIC FIELD AND STRAIN EFFECTS ON SURFACE ROUGHNESS INDUCED SPIN RELAXATION IN SILICON FIELD-EFFECT TRANSISTORS

arxiv: v2 [cond-mat.mes-hall] 6 Dec 2018

arxiv: v2 [cond-mat.mes-hall] 6 Apr 2011

NONLINEAR TRANSPORT IN BALLISTIC SEMICONDUCTOR DIODES WITH NEGATIVE EFFECTIVE MASS CARRIERS

Supporting Information for Quantized Conductance and Large g-factor Anisotropy in InSb Quantum Point Contacts

Electron Spin for Classical Information Processing: A Brief Survey of Spin-Based Logic. Devices, Gates and Circuits

Orbital Mechanisms of Electron- Spin Manipulation by an Electric Field

THERMOELECTRIC PROPERTIES OF ULTRASCALED SILICON NANOWIRES. Edwin Bosco Ramayya

Spin and Charge transport in Ferromagnetic Graphene

Transcription:

Decay of spin polarized hot carrier current in a quasi one-dimensional spin valve structure arxiv:cond-mat/0310245v1 [cond-mat.mes-hall] 10 Oct 2003 S. Pramanik and S. Bandyopadhyay Department of Electrical Engineering, Virginia Commonwealth University, Richmond, Virginia 23284 M. Cahay Department of Electrical and Computer Engineering and Computer Science, University of Cincinnati, Cincinnati, Ohio 45221 Abstract We study the spatial decay of spin polarized hot carrier current in a spin-valve structure consisting of a semiconductor quantum wire flanked by half-metallic ferromagnetic contacts. The current decays because of D yakonov-perel spin relaxation in the semiconductor caused by Rashba spin orbit interaction. The associated relaxation length is found to decrease with increasing lattice temperature (in the range 30-77 K) and exhibit a non-monotonic dependence on the electric field driving the current. The relaxation lengths are several tens of microns which are at least an order of magnitude larger than what has been theoretically calculated for two-dimensional structures at comparable temperatures, Rashba interaction strengths and electric fields. This improvement is a consequence of one-dimensional carrier confinement that does not necessarily suppress carrier scattering, but nevertheless suppresses D yakonov-perel spin Corresponding author. e-mail: sbandy@vcu.edu 1

relaxation. 72.25.Dc, 72.25.Mk, 72.25.Hg, 72.25.Rb Typeset using REVTEX 2

Ever since the discovery of the spin-valve effect [1], there has been considerable interest in studying spin transport in non-magnetic materials in which spin polarized carriers are injected from a ferromagnetic contact and detected by another ferromagnetic contact. The spin valve structure has also been employed to devise novel spintronic devices, such as the so-called spin field effect transistor [2], where an electron s spin (rather than its charge) is employed to elicit transistor action. The basic spin valve geometry is shown in the top panel of Fig. 1. It consists of a semiconductor channel (assumed to be quasi one-dimensional for this study) flanked by two half-metallic ferromagnetic contacts. One contact (called the source ) injects spin polarized current into the channel and thus acts as a spin polarizer. The other contact acts as a spin analyzer and is termed the drain. Carriers drift from the source to the drain under the influence of a driving electric field. When they arrive at the drain, they are transmitted with a probability T 2 = cos 2 (θ/2) where θ is the angle between the electron s spin polarization at the drain end and the drain s magnetization [2]. With increasing degree of spin depolarization in the channel (caused by spin relaxation), the average misalignment angle θ (for the electron ensemble) increases and consequently the transmitted current decreases. Ultimately, when there is no residual spin polarization in the current (i.e. carriers are equally likely to have their spins aligned parallel or anti-parallel to the drain s magnetization), the transmitted current will fall to 50% of its maximum value. We are interested in finding how the (transmitted) spin polarized current falls off with distance along the channel at different driving electric fields and temperatures. Spins depolarize in the channel primarily because of spin-orbit interactions caused by bulk inversion asymmetry (Dresselhaus spin-orbit coupling) [3] and structural inversion asymmetry (Rashba spin-orbit coupling) [4]. These spin-orbit couplings are momentum dependent, and because different electrons have different momenta that change randomly due to scattering, the spins become randomized by scattering and the ensemble averaged spin and spin polarized current decay with distance. This mechanism of spin relaxation is the D yakonov- Perel mechanism [5] which is overwhelmingly dominant in quasi one-dimensional structures 3

over the Elliott-Yafet [6] or Bir-Aronov-Pikus [7] mechanisms. The spatial decay of spin due to D yakonov-perel mechanism was studied in the past by Bournel, et. al. [11] and Saikin, et. al. [12] in two-dimensional channels. They mostly dealt with low driving electric fields so that transport is linear or quasi-linear. In contrast, we have studied the spatial decay in quasi one-dimensional structures of both spin and spin polarized current at high driving electric fields of 1-10 kv/cm, which result in hot carrier transport and non-linear effects. In a one-dimensional structure, the spin polarized current due to one electron is proportional to qv x T 2 where v x is the ensemble averaged velocity of the electrons along the channel. As stated before, the quantity T 2 depends on the component of the electron s spin polarization along the magnetization of the drain. We will assume that the source and drain are both magnetized along the channel s axis (x-axis). This results in the initial spin orientation to be along the channel axis. Accordingly, T 2 = cos 2 (θ/2) cos(θ) = S x / S 2 x +S 2 y +S 2 z = S x, (1) where S n is the spin component along the n-axis and S x is the normalized value of S x. The ensemble averaged spin polarized current at any position x is given by I s (x) = q v x, S x f(v x, S x,x)v x T( S x ) 2, (2) where the velocity (v x )- and spin ( S x )- dependent distribution function f(v x, S x,x) at any position x is found directly from the Monte Carlo simulator described in ref. [13] (all pertinent details of the simulator can be found in ref. [13] and will not be repeated here). We only mention that in the simulator, we use a parabolic energy versus velocity dispersion relation E = ( h 2 /2m )(nπ/w z ) 2 + ( h 2 /2m )(π/w y ) 2 + (1/2)m v 2 x (n is the subband index in the z-direction), neglecting any band structure non-parabolicity which is not important in the energy ranges encountered [8]. This dispersion relation allows us to calculate the velocity v x from the carrier energy E and subband index n (which are tracked in the simulator) very easily. If instead we used the energy versus wavevector relation (which is traditional) 4

and then attempted to find v x from the velocity versus wavevector relation, it would have been immensely complicated. The reason is that the velocity (or energy) versus wavevector relation is spin-dependent in the presence of the Rashba effect [9] and becomes even more complicated if the Rashba effect is strong which leads to spin mixing effects [10]. These complications would have been overwhelming in our case since we have a continuous distribution of spin and hence would have been faced with a denumerably infinite number of energy versus wavevector relations. The way to avoid this daunting complication (and the associated numerical cost) is to use the energy-velocity relation which is spin-independent instead of the energy-wavevector relation which is spin-dependent. In the simulation, carriers are injected into a quasi one-dimensional GaAs channel of rectangular cross section 30 nm 4 nm (see top panel of Fig. 1). We have assumed that there is a transverse electric field of 100 kv/cm (in the y-direction) that gives rise to a structural inversion asymmetry and induces a Rashba effect in the channel. This field perturbs the subband energies in the channel but only slightly. The transverse voltage drop over a 4 nm wide channel due to this field is 40 mev while the lowest subband energy is 355 mev. Therefore, the perturbation is 11% for the lowest subband and progressively decreases for higher subbands. Consequently, we neglect this perturbation. Electrons are injected from a Fermi Dirac distribution with their spins all aligned along the channel axis (x-axis) in order to simulate the spin polarizer. At any given position x, we find the spin vector S x and compute the quantity T( S x ) 2 for every electron using Equation (1). We also find the velocity v x for every electron at position x and then compute the spin polarized current I s by performing the ensemble averaging given by Equation (2). We have found I s versus position x for four different channel electric fields of 1, 2, 4 and 10 kv/cm and two different temperatures of 30 and 77 K. In the Monte Carlo simulation, we find that typically the lowest 4 subbands are occupied by electrons. The Dresselhaus interaction strength is different in different subbands because the interaction strength is proportional to (nπ/w) 2 where W is the transverse dimension (4 nm in our case) and n is the subband index. As electrons are scattered between various 5

subbands, they witness varying spin orbit interaction which results in a decaying envelope of the ensemble averaged spin (or spin polarized current). This is the cause of D yakonov-perel relaxation in quasi 1-d structures. In Fig. 1, we show the spatial decay of the normalized spin polarized current I s for the four different (x-directed) channel electric fields at a temperature of 30 K. In Fig. 2, we show the same quantity (along with the spatial decay of the ensemble averaged spin component S x ) at an electric field of 2 kv/cm at temperatures of 30 and 77 K. Spin depolarization is complete when I s reaches a value of 0.5. At this point, an electron is equally likely to have its spin aligned parallel or anti-parallel to the drain s magnetization (and therefore it is equally likely to be transmitted or blocked). We can define a relaxation length as the distance over which the injected spin polarized current decays to 50% of its value (i.e. becomes completely depolarized). Table I gives the relaxation lengths at different electric field strengths and different temperatures. As expected, the relaxation length decreases with increasing carrier temperature because of increased scattering that causes increased spin depolarization. The electric field, on the other hand, has two opposing effects. The scattering rate increases slowly with the electric field, but so does the ensemble averaged carrier drift velocity until the saturation velocity is reached. A larger drift velocity makes the carriers travel a greater distance before getting depolarized. Consequently, the relaxation length at first increases with increasing electric field, but once the drift velocity begins to saturate, the increased scattering takes over and the relaxation length starts to decrease with increasing electric field. The dependence of relaxation length on the electric field is therefore non-monotonic. Based on the data in Table I, we find that the relaxation length for spin polarized current is very large (between 20 and 100 µm for the cases considered). This is at least an order of magnitude larger than what was calculated for two-dimensional structures [11,12] at comparable temperatures and driving electric fields. This difference is not due to any suppression of scattering. In fact, even though elastic scattering is suppressed in quasi one-dimensional structures [14], inelastic scattering is not [15], and the calculated mobility in one-dimensional 6

structures in this temperature range is less than that in bulk [16]. The true origin of the difference lies in the fact that Dresselhaus and Rashba interactions cause a carrier s spin to precess slowly (during free flight) about a so-called spin precession vector that is defined by the carrier s momentum [11]. In a one-dimensional structure, a carrier is free to move only along one direction, and therefore the Rashba or the Dresselhaus spin precession vector always points along one particular direction. Scattering can change its magnitude, but not its direction. This leads to slow spin relaxation. In contrast, scattering can change both the magnitude and the direction of the spin precession vector in two- or three-dimensional structures. Therefore, spin depolarizes much faster in multi-dimensional structures. Before concluding this Letter, we should mention that in the type of structures considered here, there is always a magnetic field in the channel caused by the ferromagnetic contacts. This field, however weak, ensures that the eigenstates in the channel are not spin eigenstates [17]. Therefore, even non-magnetic scatterers can cause spin relaxation[18]. This mechanism has not been considered here, since we have not considered the channel magnetic field. In the absence of this mechanism, we have shown that spin relaxation length of carriers is very large in quasi one-dimensional structures, even at elevated temperatures and high electric fields. Large spin relaxation lengths have been observed before in multi-dimensional structures, but only at low driving electric fields and low temperatures [19]. One dimensional confinement can extend the range to high electric fields and elevated temperatures, which are required for realistic device applications. The work of S. P and S. B. were supported by the US National Science Foundation under grant ECS-0196554. 7

REFERENCES [1] B. Dieny, V. S. Speriosu, S. S. P. Parkin, B. A. Gurney, D. R. Wilhoit and D. Mauri, Phys. Rev. B, 43, 1297 (1991). [2] S. Datta and B. Das, Appl. Phys. Lett., 56, 665 (1990). [3] G. Dresselhaus, Phys. Rev., 100, 580 (1955). [4] E. I. Rashba, Sov. Phys. Semicond., 2, 1109 (1960); Y. A. Bychkov and E. I. Rashba, J. Phys. C, 17, 6039 (1984). [5] M. I. D yakonov and V. I. Perel, Sov. Phys. Solid State, 13, 3023 (1972). [6] R. J. Elliott, Phys. Rev., 96, 266 (1954). [7] G. L. Bir, A. G. Aronov and G. E. Pikus, Sov. Phys. JETP, 42, 705 (1976). [8] It is possible to introduce some band structure non-parabolicity by using an energydependent effective mass. However, this is not important. The carrier kinetic energies remain small in every subband so that non-parabolicity effects are never significant. If the energy of a carrier in the lowest subbands begins to increase, the carrier suffers intersubband transition to a higher subband by absorbing or emitting phonons. This process keeps the kinetic energy in every subband small and the carrier temperature remains very close to the lattice temperature. The intersubband scattering (not intervalley transfer) is also responsible for velocity saturation in the quantum wire. [9] L.W.Molenkamp, G.Schmidt andg.e.w.bauer, Phys. Rev.B,64, 121202(R)(2001). [10] M. Governale and U. Zülicke, Phys. Rev. B, 66, 073311 (2002). [11] A. Bournel, P. Dollfus, S. Galdin, F-X Musalem and P. Hesto, Solid State Commun., 104, 85 (1997); A. Bournel, P. Dollfus, P. Bruno and P. Hesto, Mat. Sci. Forum, 297-298, 205(1999); A. Bournel, V. Delmouly, P. Dollfus, G. Tremblay and P. Hesto, Physica E, 10, 86 (2001). 8

[12] S. Saikin, M. Shen, M. C. Cheng and V. Privman, www.arxiv.org/cond-mat/0212610; M. Shen. S. Saikin, M. C. Cheng and V. Privman, www.arxiv.org/cond-mat/0302395. [13] S. Pramanik, S. Bandyopadhyay and M. Cahay, Phys. Rev. B, 68, 075313 (2003).. [14] H. Sakaki, Jpn. J. Appl. Phys., 19, L735 (1980). [15] N. Telang and S. Bandyopadhyay, Phys. Rev. B, 48, 18002 (1993). [16] N. Telang and S. Bandyopadhyay, Appl. Phys. Lett., 66, 1623 (1995); Phys. Rev. B, 51, 9728 (1995). [17] M. Cahay and S. Bandyopadhyay, Phys. Rev. B, 68, 115316 (2003). [18] M. Cahay and S. Bandyopadhyay, submitted to Phys. Rev. B. [19] D. D. Awschalom and J. M. Kikkawa, Phys. Today, 52, 33 (1999); Phys. Rev. Lett., 80, 4313 (1998); Nature (London), 397, 139 (1999). 9

FIGURES FIG. 1. Spatial decay of the normalized spin polarized current in a GaAs quantum wire channel of rectangular cross-section 30 nm 4 nm. The results are shown for four different channel electric fields of 1, 2, 4 and 10 kv/cm at an electron temperature of 30 K. The top panel shows schematic of a spin valve with a quasi one-dimensional channel. The half-metallic ferromagnetic source and drain contacts act as spin polarizers and analyzers, while the gate terminal is used to apply a transverse electric field on the channel to induce a Rashba effect. FIG. 2. Spatial decay of the normalized spin polarized current and the injected spin vector in the channel of Fig. 1. The results are shown for two different temperatures of 30 K and 77 K at a channel electric field of 2 kv/cm. 10

TABLES TABLE I. Spin relaxation length dependence on temperature and driving electric field. Electric field (kv/cm) Temperature (K) Spin relaxation (µm) 1.0 30 20 2.0 30 60 4.0 30 100 10.0 30 50 2.0 77 30 11

Gate x y Spin polarizer (source) Spin analyzer (drain) Normalized spin polarized current I s 1.0 0.8 0.6 0.4 0.2 0 4 kv/cm 2 kv/cm 1 kv/cm 10 kv/cm 0 20 40 60 80 100 120 140 160 position x (µm)

1.0 Spin component S x or normalized spin polarized current I s 0.8 0.6 0.4 0.2 I (30 K) s I (77 K) s S (30 K) x S (77 K) x 0 0 20 40 60 80 100 120 140 160 position (µm)