Вы находитесь на странице: 1из 14

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO.

5, MAY 2007

907

Spintronics
Michael E. Flatt, Member, IEEE
(Invited Paper)

AbstractInvestigations of the dynamics of spin-polarized electronic current through and near materials with spin-dependent electronic structures have created a rich new eld dubbed spintronics. The implications of spintronics research extend deep into the realm of fundamental material properties, yet spintronics applications have also revolutionized the magnetic-storage industry by providing efcient room-temperature magnetic sensors. Control of nonequilibrium spin-polarized populations of electrons through and near magnets has led to the dominance of linear (resistive) spintronic devices for magnetic readout in commercial magnetic storage. Rapid progress in understanding the fundamental physics of nonlinear spin-polarized electronic transport in metals and semiconductors suggests new applications for spintronic devices in fast nonvolatile memory as well as logic devices, with or without magnetic materials or magnetic elds. Ongoing study of the interaction between such spintronic elements and optical elds, particularly in semiconductors, promises the future development of optical spintronic devices. Index TermsMagnetoelectronics, Rashba elds, spin coherence, spin electronics, Spin Hall effect, spin relaxation times, spin transistor, spin transport, spintronics, spin valve.

I. INTRODUCTION N 1988, during an investigation of the properties of multilayers of magnetic and nonmagnetic metallic materials, a dramatic dependence of the electrical resistance on the magnetization orientation of neighboring layers (parallel or antiparallel) was reported and named giant magnetoresistance (GMR) [1], [2]. Within a couple of years the spin valve had been introduced [3], with two metallic magnetic layers separated by a nonmagnetic spacer. Shortly thereafter, roomtemperature magnetic-eld sensors had been made [4] from spin valves which were superior to previous state-of-the-art devices based on anisotropic MR (AMR). IBM introduced GMR-based magnetic media read heads into its commercial disk-drive products in 1997, and soon, all disk-drive companies were offering GMR-based read heads instead of AMR-based heads. Out of a eld treating the effects of magnetic elds on electronic transport (magnetoelectronics, which includes AMR and many other effects) was rapidly born a eld focusing

Fig. 1. (a) High-resistance and (b) low-resistance geometry of CIP GMR. The spacer region is a nonmagnetic metal. (c) High-resistance and (d) low-resistance geometry of CPP GMR (if the spacer region is a nonmagnetic metal) or TMR (if the spacer region is an insulator). Arrows indicate the magnetization direction of the magnetic layers. For applications, one ferromagnetic layer is typically pinned through shape anisotropy and may be exchange-biased [10] and the other layer is reoriented by the external magnetic eld.

Manuscript received December 1, 2006; revised January 24, 2007. This work was supported by the Ofce of Naval Research. The review of this paper was arranged by Editor H. Morkoc. The author is with the Department of Physics and Astronomy, Department of Electrical and Computer Engineering, and Optical Science and Technology Center, The University of Iowa, Iowa City, IA 52242-1479 USA. Digital Object Identier 10.1109/TED.2007.894376

on spin-dependent electronics, which came to be called spintronics. Within this eld of spintronics (in its area of overlap with magnetoelectronics), a fundamental physics discovery came to drive a large commercial sector (2005 sales exceeding $3 billion) within a decade of discovery. The initial discovery of GMR was for a conguration (shown in Fig. 1(a) and (b) for a spin valve) called current-in-plane GMR (CIP GMR). Shortly thereafter, a simpler experimental geometry was investigated (Fig. 1(c) and (d), called the current perpendicular-to-plane (CPP) conguration [5], [6]), which yielded a much larger value than in CIP GMR of the MR (the percent difference in resistance for parallel and antiparallel orientations of the two ferromagnetic regions in the spin valve). In each of these congurations, the two ferromagnetic regions are separated by a region of nonmagnetic metal, typically with a higher conductivity than either of the two ferromagnetic layers. Motivated by earlier low-temperature measurements of small magnetic-eld effects on the tunneling resistance between ferromagnetic layers [7], it was found that replacing the nonmagnetic metal in the CPP-GMR geometry with an insulating tunnel barrier can yield large values of the MR at

0018-9383/$25.00 2007 IEEE

908

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO. 5, MAY 2007

room temperature [8]. This effect, referred to as tunneling MR (TMR), can exceed a factor of three at room temperature [9]. Applications of such structures in magnetic-eld sensors rely on the reorientation of the magnetization of one of the two layers in a small magnetic eld (the other layer is pinned through shape anisotropy and may be exchange-biased [10]) and electrical detection of the change in resistance. The sensor can be placed in a magnetic-storage read head, and the small magnetic eld can come from an encoded bit on magnetic media (such as in a hard disk drive). Alternatively (such as in magnetic random-access memory, or MRAM), the spin valve itself can be used as a magnetic bit, with the two-memory states corresponding to antiparallel and parallel congurations. In the initial MRAM designs, a small magnetic eld generated by a current pulse in an integrated wire (or pair of wires) is used to write the memory state, and readout is performed by measuring the resistance of the spin valve. The discovery of these spin-dependent effects in metallic electrical transport has spawned a wide range of further investigations and has united previously distinct subelds of physics [11][21] into the new eld of spintronics. This includes spindependent transport in semiconductors, spin-dependent tunneling through insulators, and spin-transport effects driven by the spinorbit interaction in semiconductors, oxides, and metals. The investigation of equilibrium and transient spin-dependent phenomena using optical techniques, a rich eld in the 1970s [12], has blossomed again, particularly for magnetic insulators and nonmagnetic semiconductors. The discovery of semiconductors that can become ferromagnetic upon dilute doping with magnetic atoms [15], [22], [23] may connect all these areas by providing electrical control of magnetic properties, spindependent transport, and strong magnetooptical effects, such as Faraday rotation. The new areas of semiconductor spintronics have yet to have a commercial impact, although the properties found in these systems are unique and numerous theoretical device proposals have been put forward. Here, some of the unifying themes of spintronics will be explored, including various regimes of spin transport in magnetic and nonmagnetic metals, semiconductors and oxides, spin-dependent optical properties such as Faraday rotation, and electrical control of magnetic properties and spin lifetimes. The treatment begins with a discussion in Section II of the spindependent electronic properties of materials, including those characterizing spin transport and spin persistence (spin conductivities and spin lifetimes), then continues with the coupling of optics to spin (Section III), proceeds to the behavior of spin-dependent electrical transport (Section IV), and concludes with some discussion of the roadblocks remaining for some of the more long-term device applications (Section V). For more extensive treatments of these issues, reviews treating spintronics can be found in [17] and [18], and topical treatments of individual aspects can be found in [13][16], [19][21]. II. SPIN-DEPENDENT ELECTRONIC PROPERTIES OF MATERIALS As with ordinary charge-transport problems, an understanding of the eigenenergies and eigenstates of a material is es-

sential for calculations of spin transport and magnetooptical effects. For ordinary calculations of charge transport, the spindependent character of these eigenstates is neglected. The discovery of GMR itself, however, is evidence that, in some circumstances, the spin-dependent character of electrical transport can have a sufcient effect on ordinary charge transport to be easily detected. To identify the principal characteristics of spin-dependent electronic structure important to spintronics, the relevant materials can be grouped into several classes, based on whether they are magnetic or nonmagnetic and whether they are metals or insulators (including semiconductors). A. Magnetic Metals The common starting point for most analyses of spin transport is the two-channel conduction approximation, originally developed for magnetic metals [24], [25] and, subsequently, applied to heterostructures of magnetic and nonmagnetic metals [26], [27]. This approach models the current as carried independently by spin-up and spin-down carriers, with a separate conductivity for each. The spin-up and spin-down carriers can independently reach a local equilibrium but may be out of equilibrium with each other. The rate of spin-ip processes, which couple the two channels, is much smaller than the ordinary scattering rate characterizing the spin-conserving momentum relaxation of nonequilibrium spin-up or spin-down distributions. Thus, even though metals often have mean free paths much shorter than a nanometer, the spin-diffusion length can range from several nanometers in strongly scattering magnetic materials to several micrometers in metals lacking magnetic impurities or strong spinorbit scattering. In this two-channel model, the bulk conductivity of spin-up and spin-down carriers is determined separately by the density of states at the Fermi energy EF for spin-up and spin-down carriers, so () = e2 N() (EF )D() (1)

where D() is the diffusion constant for spin-up (spin-down) carriers. The response of such a material to an applied electric eld is to ow a spin-polarized current, with a polarization (dened below) that is determined by the conductivities Pcurrent = J J = . J + J + (2)

Another important quantity describing the properties of magnetic metals is the change in chemical potential associated with an excess of spin-polarized carriers () = N() (EF ). n() (3)

One of the successes of spin-transport theory and experiment is the ability to generate regions of spin accumulation in magnetic and nonmagnetic materials. In metals, the density of excess carriers of one spin direction is usually several orders

FLATT: SPINTRONICS

909

of magnitude lower than the background density of carriers, so the resulting change in the conductivity (which is dependent on the density of states at the Fermi energy), for either spin-up or spin-down carriers, is negligible. The spin-dependent conductivity, the spin-dependent density of states at the Fermi energy, and the spin-ip time sf phenomenologically describe most bulk transport phenomena in metals. The applicability of this simple phenomenological model is due to the central role that the Fermi surface plays in metallic electronic transport. The electronic structure deep within the band has little effect on the properties of charge transport or spin transport through these materials. In spintronic structures such as the CIP GMR spin valves, however, in which the mean free path of carriers is longer than the thickness of the nonmagnetic metal layer separating the ferromagnetic layers, the spin-up and spin-down conductivities must be determined for the structure as a whole and cannot be constructed simply from the spin-resolved conductivities of the bulk-constituent materials. The spin-ip time sf , in combination with the conductivities above, determines the diffusion length for an accumulation of one spin direction or the other relative to the background. These spin-diffusion lengths can be quite short (a few nanometers in permalloy [28]) but as long as micrometers in clean noble metals such as gold [29], copper, or silver [30]. The addition of impurities that increase the spinorbit coupling or that increase mutual spin-ip scattering events with an impurity magnetic moment can decrease the spin-diffusion length to a few nanometers in copper or silver [30] and likely limit the spin-diffusion lengths in magnetic materials as well. The majority spin direction of carriers is commonly dened in magnetic metals as the spin direction parallel to the bulk magnetization. In a Stoner model [31], the electronic structure of spin-up carriers can be determined from that of spin-down carriers by displacing the electronic structure of the spindown carriers by a momentum-independent energy. For a freeelectron carrier dispersion relation the Stoner model implies that the carriers with high conductivity have a spin direction parallel to the magnetic moment, which is not always the case in real structures with more complex band structures (particularly in magnetic semiconductors). The direction of the magnetization usually does not play a direct role in metallic spin transport, except in how it affects the relative orientation of high- or lowconductivity spin channels to an applied magnetic eld. Other electronic properties of the metal often have signicantly less of an effect on transport. As an example, the Fermi momentum (which has not been specied above) does not signicantly inuence bulk electronic transport independently of the transport parameters introduced above, although it can play an important role in determining the tunneling through an interface. During tunneling, the wave functions of the metal must be matched with those of the insulating material, and thus, the Fermi momentum (and band character) of the carriers at the Fermi level in the metal may play an important role. Another exception is the mean free path, which plays an important role in CIP GMR as it determines the range over which the two ferromagnetic layers of the spin valve can communicate with each other.

The transport of carriers within a magnetic metal is also inuenced by the spinorbit interaction. In any material with a spinorbit interaction, the eigenstates are not factorable into a spinor and an orbital wavefunction. As a consequence, when carriers move in a preferred direction, such as through the ow of charge current, there is an inuence on the trajectory of the carriers depending on whether their spin is oriented in one direction or another. The effect is a transverse force with opposite sign for spin-up and spin-down carriers. In a magnetic metal, as the conductivities of spin-up and spin-down carriers are different, the ordinary charge current is composed of different numbers of spin-up and spin-down carriers moving with different velocities. As carriers of one spin orientation are deected toward one side of the wire, and of the other are deected toward the other side, the consequence of spinorbit entanglement is a Hall voltage across the wire and a spin accumulation (of opposite signs) on the two sides of the wire. This transverse charge current, called the anomalous Hall effect, originates either from a difference in the scattering process from an impurity for spin-up and spin-down carriers (so-called skew scattering [32]) or from the change in the form of the position and velocity operators in a material with spinorbit interaction (so-called side-jump [33][36]). B. Nonmagnetic Metals The properties of nonmagnetic metals can be described with the same phenomenological quantities as magnetic metals, although nonmagnetic metals will have identical conductivities and densities of states at the Fermi energy for spin-up and spindown carriers. This does not, however, preclude the existence of signicant spin-dependent transport in nonmagnetic metals. The clearest example of this effect is the Spin Hall effect. The origin of the effect can be the same as the anomalous Hall effect described above for magnetic metals, but as an equal number of spin-up and spin-down carriers are moving in the nonmagnetic material, there is no Hall voltage, only two counterpropagating spin currents of opposite polarization [37], [38]. The spin Hall effect was rst identied in transport in nonmagnetic semiconductors [39][41] but has since been demonstrated in aluminum [42]. The phenomenon often is seen as an accumulation of spins pointing along one direction on one side of a wire and of spins pointing along the other direction on the other side of the same wire. The size of the effect in aluminum [42] is similar to that predicted in [43] and [44]. As this effect has been observed at room temperature, now in ZnSe materials [45], perhaps the Spin Hall effect will prove useful eventually for routing packets of spin, for communications, or for logic. Another important consequence of the correlation between spin and orbit is spin relaxation in nonmagnetic metals. The nonmagnetic metals of interest are predominately centrosymmetric elemental crystals or random alloys. In the absence of an applied magnetic eld, all states in the electronic spectrum of such materials appear in (at least) doubly degenerate pairs. As the spin and orbital degrees of freedom are entangled for these states, it is proper to refer to this degeneracy as pseudospin degeneracy, although it is often referred to as spin degeneracy

910

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO. 5, MAY 2007

in the literature. Pseudospin degeneracy can be derived from the dual requirements of time-reversal (Kramers) symmetry and inversion symmetry Es (k) = Es (k) Es (k) = Es (k) imply Es (k) = Es (k) (Pseudospin degeneracy) (6) (Time reversal symmetry) (Inversion symmetry) (4) (5)

where s is the pseudospin direction. The carriers in a metal are scattered from one momentum state to another on femtosecond timescales. As the entanglement of spin and orbit differs slightly for different momentum states, each time a transition from one momentum state to another occurs, there is a nite probability of ipping the spin. This process, called the ElliottYafet process [46], [47], limits spin lifetimes in sufciently pure copper [48] and many other nonmagnetic metals. C. Magnetic Insulators and Semiconductors Some of the largest MR effects measured to date have been found at low temperature in tunnel junctions involving ferromagnetic insulators [49][51]. These materials exhibit spinsplit band-edge energies of the order of a couple hundred of millielectronvolts. Thus, the barrier for spin-up and spin-down carriers to tunnel through the insulating barrier differs by much more than kB T and may be highly magnetic-eld-dependent. Although these rare earth magnetic insulators are not magnetically ordered at room temperature, their existence has prompted considerable exploratory research looking for roomtemperature magnetic insulators that permit spin-preserving transport through a conduction-band barrier (spin-preserving FowlerNordheim tunneling). Such a room-temperature magnetic insulator would very likely be a magnetic semiconductor. The ideal magnetic semiconductor would be a material that was ferromagnetic above room temperature, exhibited a spinsplit band structure, had high-mobility carriers, and could be controllably doped (preferably from n-type to p-type). One approach to look for such materials is to add magnetic dopants to nonmagnetic semiconducting hosts [22], [23]. Success in growing ferromagnetic In1x Mnx As and Ga1x Mnx As (which are not ferromagnetic at room temperature) and the prediction of room-temperature ferromagnetism in other doped semiconductor hosts [52], strongly motivated researchers to explore other potential hosts. Many of the subsequent reports of hightemperature ferromagnetic semiconductors in the literature, however, have not been conrmed, probably due to the potential formation of different combinations of the dopants and host elements than intended. Many of these different phases have ferromagnetic transition temperatures well above room temperature but are metal inclusions or clusters embedded in the material, instead of true magnetic semiconductors. It has proven very challenging, in some cases, to differentiate the signal of a sample with small regions of metallic phases embedded in nonmagnetic semiconductors from that of a sample with a

true ferromagnetic semiconductor. In the meantime, substantial progress on elevating the Curie temperature of Ga1x Mnx As (to above 170 K) was achieved by subjecting the samples to extended low-temperature postgrowth anneals [53], [54], which appear to substantially reduce the density of undesirable interstitial Mn. Extensive characterization studies have been done on some materials that are clearly ferromagnetic semiconductors, such as Ga1x Mnx As, and include magnetotransport, bulk magnetization, and magnetooptics (magnetic circular dichroism) [17], and as a result, the mechanism of ferromagnetism has begun to become clear. When the magnetic dopant Mn is substituted for Ga, the Mn acts as a relatively shallow acceptor (113-meV binding energy). Due to the exchange interaction between the core spin and the hole, the bound hole has its spin aligned antiparallel to the core spin of the Mn. When the density of Mn dopants becomes high enough, the wave functions of neighboring acceptor states can overlap. Through the double-exchange mechanism [55], the energy is substantially lowered if the hole spins are oriented parallel, which drives the ferromagnetic state. Recently, the anisotropic energetics of the interaction between Mn pairs were visualized directly through scanning tunneling microscope experiments [56], verifying predictions made earlier [57] that also explained the anisotropy of the individual Mn acceptor state [58], [59]. When the density of Mn becomes large enough, the atomic details of the alloy can be smoothed away and the true spatial distribution of the Mn dopants replaced by a mean-eld exchange interaction. Calculations based on such mean-eld theories [52], [60], [61], in which the doubleexchange model approaches a Zener limit [52], for such hole-mediated ferromagnetic semiconductors, have proved successful at explaining the Curie temperatures, easy axes, magnetooptical properties, and magnetoelastic constants. Some properties, such as the optical conductivity [62], still appear challenging to explain based on either the single-Mn models or the mean-eld models. The importance of holes in mediating the ferromagnetic state suggests that it should be possible to modify the properties of the magnetic material by reducing the number of holes (through gating). Such an effect would be extremely challenging (if not impossible) to achieve in a magnetic metal because of the very high carrier concentration but has been demonstrated for these hole-mediated ferromagnetic semiconductors. Both the Curie temperature [63] and the coercive eld [64] have been changed with a gate eld. Very large MRs have also been found when tunneling through regions of depleted Ga1x Mnx As, a magnetic semiconductor [65]. Here, the electronic structure of the depletion edge of a doped magnetic semiconductor was assumed to remain spin-split, as that of a magnetic insulator, for in bulk when the carriers are depleted from Ga1x Mnx As; the material does not stay ferromagnetic. The current limitations of Ga1x Mnx As and other holemediated ferromagnets for technological applications are the relatively high level of doping required to get a high Curie temperature (corresponding to Mn concentrations near 10%), the extremely short lifetimes of optically injected carriers [66], and the Curie temperatures below room temperature. As these materials must be grown at relatively low temperatures (com-

FLATT: SPINTRONICS

911

pared to optimal temperatures for nonmagnetic semiconductors), control of the doping prole sufcient to grow magnetic bipolar devices has proved a great challenge, although there is a recent report of success in fabricating a magnetic p-n diode [67]. D. Nonmagnetic Insulators and Semiconductors Just as the correlation of spin properties and orbital ones led to the Spin Hall effect and nite spin lifetimes in nonmagnetic metals, these same phenomena can arise in nonmagnetic insulators and semiconductors. Transport through or within these materials, however, can also involve the electronic structure far from the Fermi surface, so the number of potential parameters required to describe spin transport can be far greater than those required for metals. Recently, it was predicted that the tunneling process through MgO barriers would be extraordinarily spin-selective, and subsequently, the MR measured in tunnel junctions involving this material greatly exceeded the expected result from the spin polarization of the metallic contacts [9], [68]. The prediction was based on rst-principles calculations of the electronic structure of FeMgOFe structures. The very large magnetoresistive effect in such structures (exceeding a factor of three change in resistance [9]) persists even to room temperature and can also be used as an efcient spin injector from a magnetic metal into GaAs at room temperature [69], [70]. As most semiconductors do not have a center of symmetry, the consequences of the spinorbit interaction can be quite different than in nonmagnetic metals. Even in the absence of an applied magnetic eld, the pseudospin states of nonmagnetic zincblende semiconductors, such as GaAs and other IIIV semiconductors, are not degenerate [71]. Pseudospin states of the electron are split at nite k by the relativistic transformation of internal electric crystal elds into magnetic elds in the rest frame of the moving electron. The splitting is described by the Hamiltonian H = gB(k) S = (k) S (7)

would have an innite lifetime from this mechanism (again, ignoring the pseudospin nature of the states). For a zincblende crystal, however, it is impossible to choose a global quantization axis, because the direction of the crystal magnetic eld varies with k. The effective eld for k (110) is perpendicular to the eld for k (110), so no global quantization axis exists [71]. An unusual special case exists to this rule, for there is a global quantization axis for (110)-grown zincblende quantum wells. The presence of this effective internal magnetic eld implies that the spin polarization of a population of carriers will dephase due to the variability of the internal eld with momentum. Dyakonov and Perel [72] have developed a theory for the spin lifetime in bulk zincblende semiconductors based on dephasing by the effective crystal magnetic eld [B(k) in (7)]. A derivation based on the density matrix formalism is viewable in [12]. Thus, the spin lifetime in a material is closely connected to the structure of B(k), which itself is closely connected to the nature of the inversion asymmetry of the material. For example, the presence of the global quantization axis for (110)-grown zincblende quantum wells produces an exceptionally long spin lifetime for spins oriented parallel to that direction [73][76]. As the inversion asymmetry of a material can be controlled, either in growth through the introduction of composition gradients or after growth through the application of electric (gate) elds, this mechanism of spin relaxation provides a direct handle for tuning the spin lifetime. The effective magnetic eld B arising from these effects is known simply as the Rashba eld [77], [78] and has the form B = E k (9)

where B(k) is the effective crystal magnetic eld, (k) is the resulting Larmor precession vector, g is the g factor, and is the Bohr magneton. For direct-gap semiconductors, the effective crystal magnetic eld vanishes at the conduction minimum (k = 0), because Kramers degeneracy (4) requires that B(k) = B(k). (8)

where E is an effective electric eld arising from the combination of the effects above (compositional gradients or asymmetry and applied external electric eld), k is the carriers crystal momentum, and depends on the strength of the spinorbit interaction in the material. For (110) zincblende quantum wells, this Rashba eld destroys the global quantization direction causing the spin lifetime to shorten [79], potentially by several orders of magnitude. Reductions of a factor of ten in the spin lifetime in a GaAs/AlGaAs quantum well [80] and of a factor of four in the spin lifetime in an InAs/AlSb quantum well [81] have been observed experimentally through the application of an electric eld. III. OPTICAL COUPLING TO THE SPIN The photon itself couples only weakly to the spin; however, the spinorbit interaction permits spin-selective optical transitions to occur even through electric-dipole interactions. The details of such transitions depend on the electronic structure of the specic material, and most optical experiments to probe spin dynamics have been performed on direct-gap zincblende semiconductors. For these materials, the conduction band has s-like symmetry and the valence band has p-like symmetry. The crystal eld of the semiconductor does not further split the electronic states in the conduction or valence band at the point. Instead, the splitting that occurs is due to the spin

Despite this, within 100 meV of the band edge in GaAs, this internal magnetic eld approaches 1000 Tesla, which far exceeds the ability of a magnetic eld applied from a typical laboratory magnet to split the electronic spin states. From (8) it is apparent that an ensemble of spins in such a material will experience a varying effective magnetic eld, depending on the carrier momentum, that will cause spins in the ensemble to precess differently from one another (a phenomenon referred to as dephasing). As a spin only precesses in a transverse eld, if B(k) were parallel to a xed direction for all k, then a macroscopic magnetization oriented along that axis

912

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO. 5, MAY 2007

orbit interaction, which splits the six p states into a fourfold degenerate J = 3/2 multiplet and a J = 1/2 doublet. The valence-band edge corresponds to the J = 3/2 multiplet, and the individual states are labeled heavy hole and light hole up and down. The heavy-hole states correspond to orbital angular momentum parallel to the spin angular momentum and are product states of spin and orbit. The light-hole states, however, are not product states and consist of one-third spin antiparallel to orbital momentum along with two-thirds spin perpendicular to orbital momentum. An electric-dipole transition will be a spin-conserving transition and, if it involves circularly polarized light, will either change the orbital angular momentum by +1(+ ) or 1( ) along the axis of propagation (taken to be +). A + photon z will thus generate a heavy hole of Jz = +3/2 and an electron of Sz = 1/2 or a light hole of Jz = +1/2 and an electron of Sz = +1/2, although the probability of the rst process is three times larger than the probability of the second. Thus, circularly polarized light will generate an optically excited density of spin-polarized conduction electrons [12], with a maximum polarization of 50% Pdensity = n n . n + n (10)

Similarly, spin-polarized electrons recombining with unpolarized holes will generate circularly polarized light with a polarization of 50%. In a quantum well, the light-hole states are typically of higher energy than the heavy-hole states due to quantum connement (and possibly strain). In this situation, in which the light-hole states are not optically accessed, the polarization generated or detected can approach 100%. These spin-dependent selection rules are further modied in quantum dots, where they can be very sensitive to the size, shape, and strain of the dot [82]. Time-resolved nonlinear optical techniques have provided direct measurements of the evolution of a population of spins, once they have been injected into a nonmagnetic semiconductor. In most bulk and quantum-well structures, the spin relaxation times for holes are very short [83] due to the much greater spinorbit entanglement in the p-type valence bands than in the s-type conduction bands. In time-resolved photoluminescence the electrons recombine with unpolarized holes, and the degree of circular polarization of the light determines the remaining polarization of the electrons [84][89]. Another approach relies on the saturation of the optical transition used to generate the spin-polarized carriers once the spin-polarized carriers are present. A second pulse (the probe) incident on the material with the same circular polarization (SCP) as the initial pulse (the pump) will be transmitted through the sample more than it would have been without the pump [90][92]. To distinguish spin-dependent properties from ordinary band-lling effects, the experiment is also done with a probe of the opposite circular polarization (OCP) than the pump. For bulk materials with selection rules near 50%, the transmission probability of the probe with OCP will still be modied by the pump but less than the probe with the SCP. For quantum wells, the transmission of an OCP pump can be negligible right after the pump pulse. As

the carriers lose their spin polarization through spin relaxation or decoherence, the transmission of the OCP pump will rise and that of the SCP pump will decay, so the spin lifetime can be determined from the decay time (T1 ) of the SCPOCP signal. One elegant method of probing the spin dynamics of optically injected carriers is Faraday rotation. In this technique [93], the real index of refraction is measured, rather than the transmission coefcient. The difference in transmission coefcients implies that the real indexes for SCP and OCP light will be different. Thus a linearly polarized probe pulse will have its polarization axis rotated (Faraday rotation), and this rotation can be sensitively detected with a crossed polarizer. By applying a magnetic eld transverse to the light-propagation direction, the spin population can be forced to precess, yielding oscillatory Faraday rotation angles which directly indicate the Larmor precession frequency of the spin as well as the trans verse spin ensemble decoherence time T2 . Measurements via this technique identied very long spin-coherence times, even at room temperature, in IIVI semiconductor materials [94]. The spin-coherence times at low temperature, in commercial wafers of GaAs, were found to exceed 100 ns [95]. Spatially resolved Faraday-rotation measurements also clearly showed that packets of spin-polarized carriers in semiconductors could be moved from one region of the semiconductor to another [96] or from one semiconductor material (such as GaAs) to another (such as ZnSe) [97]. Observations of the decay of an optically induced spin grating, created by interfering two cross-polarized pulses of light propagating at an angle to each other through the sample, also indicated that spin-polarized carriers could move without losing their spin orientation [98]. The very rapid spin relaxation possible in some semiconductor quantum wells has led to the proposal of spin-based optical switches [99]. Spin relaxation times shorter than 1 ps have been found in GaSb quantum wells [92], whose optical transitions have a wavelengths near 1.55 m. Recently, even faster optically switched transitions have been achieved using the excitation of virtual spin-polarized excitons [100]. Potential optical applications of ferromagnetic semiconductors include switchable magnetooptical elements, such as Faraday isolators. A Faraday isolator uses the time-reversalsymmetry breaking of the Faraday effect to generate one-way transmission. These devices are often used to isolate laser cavities from the remainder of an optical circuit. For communications wavelengths, the material commonly used is yttrium iron garnet, which provides a large Faraday rotation and low optical loss. Use of a switchable ferromagnet would permit the one-directional transmission of a Faraday isolator to be turned ON or OFF, or the direction of the allowed transmission to be changed dynamically. IV. TRANSPORT For most cases of simple macroscopic-charge-current ow, such as current ow in metals, transport is analyzed assuming that the charge carriers are in a local equilibrium. In this framework, all the complex dynamics of an ensemble of charge carriers are ignored, and the local properties are summarized by a small number of parameters: local chemical potential; electric

FLATT: SPINTRONICS

913

potential; conductivity; and temperature. Charge transport is then analyzed by relating the current J directly to the local change in the chemical potential by J = (

A. Incoherent Spin Transport 1) Formalism: A spin-polarized electron, through interaction with external elds or its surroundings, can ip its spin, so the spin polarization of a current is not conserved in a spintronic circuit. Thus, the ow of spin-polarized current should be characterized by parameters, such as the spin-ip time (and related spin-diffusion length), describing aspects of the steady-state transport that are not subject to the same strict conservation rules as charge current. The treatment here begins with some general considerations that apply to both metals and semiconductors, and then species what approximations are commonly introduced to treat spin transport in metals or semiconductors. Incoherent spin transport is usually considered either by introducing separate quasichemical potentials for spin-up and spin-down carriers (an approach common in the metallic community) or by introducing drift-diffusion equations that separately track spin-up and spin-down carrier densities and currents. These two approaches are, however, very closely related [105], as illustrated here for charge ow. In the steady-state transport of a single charge species, the continuity equation and conservation of charge lead to J=0 (12)

(11)

where is the chemical potential for electrons measured relative to the electric potential, is the electric potential, and is the conductivity. In systems such as semiconductors that can sustain large deviations of the local charge density from equilibrium, the conductivity is separated into the carrier density and mobility and must be found self-consistently along with the solution of the Poisson equation. Conservation of charge requires that, in steady-state transport, every electron passing into an element of a circuit must pass out. Solution of (11) for the boundary conditions appropriate for a given voltage drop will produce a full description of the current distribution in response to an applied voltage. Spin transport requires a new descriptive framework. The spin of an electron can be altered through interaction with external elds or the surroundings and, hence, is not a conserved quantity for the carriers. Furthermore, the spin, as a quantummechanical quantity, cannot be fully described according to its projection along a single quantization axis. An ensemble of spins with no spin-polarization along the z axis could simply be an unpolarized ensemble or could be an ensemble that is fully polarized in the + direction. x The theoretical treatment of spin transport simplies considerably if a single quantization axis can be selected and the spin polarization in transverse directions can be neglected. In this situation, referred to as incoherent spin transport, a quasi-chemical potential can be introduced for each of the two spin directions of a spin-1/2 particle, a spin-up quasi-chemical potential, and a spin-down quasi-chemical potential. The term quasi-chemical potential refers to a model where the spinup carriers are considered to be in local equilibrium with other spin-up carriers (and the same for spin-down carriers) but spinup and spin-down carriers are not in equilibrium with each other (the two-channel model of Section II). This is a very similar model to that used to describe nonequilibrium transport of electrons and holes in semiconductors. Considerable progress has been made in analyzing incoherent spin transport, for it is remarkably similar to the two-band model for transport in semiconductors [101], [102]. Thus, equations that apply to the transport of electrons and holes can at times be simply converted into equations that apply to the transport of spin-up and spin-down carriers. If a single quantization axis cannot be dened, then the system is in the realm of coherent spin transport. One approach to the theory of coherent spin transport is to consider the evolution of the spin-density matrix through the system instead of merely considering the evolution of the diagonal (spin-up and spin-down) elements. In some limiting situations, evolution equations for the macroscopic spin (treated as a classical vector quantity) of the ensemble have been developed and successfully applied to situations of interest [96], [103], [104].

and thus the current can be written as the gradient of a scalar potential J = = ( + For nondegenerate carriers be written as
).

(13)

= kB T ln(n/no ), and (13) can

J = E + (kT /n)n = E + eDn

(14)

where n is the carrier density (no is the equilibrium carrier density), kB is Boltzmanns constant, T is the temperature, E is the electric eld, e is the magnitude of the electric charge, and D is the diffusion constant. The far right expression of 1. Thus (14) also holds in the degenerate regime, if n/no one can use either the chemical-potential expression or the drift-diffusion equations to calculate the current in a general situation. When either of these equations is combined with the Poisson equation relating the local deviation from equilibrium of the charge density to the electric eld e E = (n no ) (15)

a full solution is possible. Frequently, the expressions in terms of quasi-chemical potentials are used when the carrier density is sufciently high that the conductivity can be considered independent of the nonequilibrium spin-polarized carriers (as is the case in metals). Under these conditions, (13) can be solved for with the appropriate boundary conditions (continuity of current and a given voltage drop across the structure), and it is not necessary to consider

914

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO. 5, MAY 2007

the carrier density as a separate variable to be determined selfconsistently. When the carrier density is low enough that the spin-polarized carriers markedly change the conductivity of the material (as can be the case in semiconductors), then in (13) must be determined self-consistently, requiring solving for the carrier density in addition to the chemical potential and the electric eld. In these situations, it is easier to use drift-diffusion equations, which can be written without the chemical potential as a separate quantity to be determined. The essential modication to the charge transport expressions of (12)(14) for spin transport is to consider the possibility of spins ipping from up to down or down to up as follows: n = je + e n e n t n = je e n + e n . t (16) (17)

[17]. To include the possibility of spin imbalance in both the conduction and valence bands, four currents are required as follows: je = en e E + eDe n je = en e E + eDe n jh = ep h E eDh p jh = ep h E eDh p . (21) (22) (23) (24)

The evolution in time and space of these four currents and the electric eld comes from the four continuity equations e n = je ee n+e e n t eGe + eR n p + eR n p n = je + ee n ee n t eGe + eR n p + eR n p p = jh + eh p ep p t + eGh eR n p eR n p p = jh eh p + eh p t + eGh eR n p eR n p

(25)

In steady state the left-hand sides vanish. Within metals the expressions can be further simplied by considering chargeneutral perturbations to the carrier density. Such perturbations naturally satisfy the Poisson equation with E = 0, as n = n . Substituting into (16), (17) yields n (e + e ) n = . D
2

(26)

(27)

(18)

For small perturbations n is proportional to , so we can dene a chemical-potential difference between spin-up and spin-down carriers s such that [106], [107] 2 s = s L2 (19)

(28)

and the Poisson equation relating the local deviation from equilibrium of the charge densities to the electric eld e E = (n + n p n ). (29)

where L2 = D/(e + e ). The chemical potential for each spin direction is continuous across an interface in the absence of an interfacial resistance [106]. The presence of an interfacial resistance can be absorbed into an effective chemical-potential discontinuity determined by the spin-resolved current through the interface. For spin-dependent transport in semiconductors, however, the density dependence of the conductivity can have a dramatic effect on the transport [108][110]. Equation (19) is modied by the presence of an electric-eld-dependent term describing carrier drift and is simpler to express in terms of the carrier density. For a nondegenerate distribution of carriers 2 (n n ) + eE (n n ) (n n ) = 0. (20) kB T L2

For degenerate carriers the energy scale appearing in the new drift term of (20) is related to the Fermi energy rather than the thermal energy. The consequence of this new drift term is a different upstream and downstream diffusion length for carriers, similar to the different upstream and downstream diffusion lengths for packets of minority carriers in semiconductors. Fully general drift-diffusion equations for carrier motion in a semiconductor describe the combined motion of electrons and holes in the presence of electric and quasi-electric elds, including the effects of space-charge elds on carrier motion

Here, are rates for spin relaxation of electrons and holes, G are generation rates for electronhole pairs, R are their recombination rates, and is the dielectric constant. Varying magnetic-band edges can be considered within this formalism by permitting the effective electric eld to differ for different bands and spin directions [111], [112]. This can produce unusual drift effects in magnetic heterostructures, including effects analogous to grading the base-doping concentration in a bipolar transistor [113]. 2) Applications of Incoherent Spin-Transport Theory: The mechanism for GMR in the CIP and CPP geometries (Fig. 1) is quite different. For the CIP geometry, the MR comes from enhanced spin scattering of carriers moving through the nonmagnetic spacer when they can interact with two regions of oppositely oriented ferromagnetic material. If the spin scattering for carriers of one spin direction is greater than for carriers of the opposite spin when those carriers are within either the ferromagnet or at the ferromagnetnonmagnetic-metal interface, then for antiparallel ferromagnetic layers and thin spacer layers, the carriers of both spin directions will experience a strong spin scattering. This occurs when the spacer layer is thinner than a mean free path, so a carrier in the nonmagnetic metal spacer can sample the spin-dependent scattering coming from both layers. When the top and bottom layer of ferromagnet are oriented parallel to each other, however, the scattering of one

FLATT: SPINTRONICS

915

spin direction is even larger but that of the other spin direction is very small. In the two-channel model, the high conductivity of the weakly scattered spin direction for the parallel-layer geometry dominates over the two lower conductivity channels for the antiparallel-layer geometry. Various origins of the spindependent scattering have been considered, rst in approximate models that achieved qualitative agreement [26], [27], [114], [115], and then, in rst-principle calculations that achieved quantitative agreement [116]. For the CPP geometry the MR comes instead from the matching conditions of (19) at the two interfaces [106], [117], [118]. Trying to force a spin-polarized current into the nonmagnetic metal produces a spin-accumulation region in the metal (corresponding to spin injection [119]) and that accumulation decays away from the interface over a length scale given by the spin-diffusion length. As a result, the current owing in the nonmagnetic metal is still spin polarized. If the current reaches the second ferromagnetic layer prior to the spin-diffusion length, then the current, still spin polarized, will nd it easier to ow into a ferromagnetic layer oriented parallel to the rst. Thus, the voltage drop across the sandwich will be smaller for parallel ferromagnetic layers than for antiparallel layers. For TMR, the resistance of each channel can be independently determined assuming that the resistance is proportional to the product of the densities of states on the two sides. Then for parallel layers of the same ferromagnetic material, the resistance is proportional to [N (EF )2 + N (EF )2 ]1 , and for antiparallel layers, the resistance is proportional to [2N (EF )N (EF )]1 . The mechanism for MR in the CPP geometry requires the ow of a spin-polarized current in the spacer region, which implies an accumulation of spin-polarized carriers within the nonmagnetic spacer (spin injection). The amount of spin injection, however, is sensitively dependent on the relative conductivity of the magnetic and the nonmagnetic materials [120]. The high resistance of a semiconductor material, relative to a magnetic metal, leads to a negligible spin-injection efciency. Initially, this obstacle was overcome by using a high-resistance magnetic metal, either a magnetically polarizable semiconductor [121] or a ferromagnetic semiconductor [122], and detecting the circular polarization of the optical illumination in a so-called spin light-emitting diode. Optical techniques have also been used to demonstrate incoherent spin transport under large electric elds in nonmagnetic-semiconductor structures [123]. Subsequently, a high-resistance spin-selective barrier was proposed as a route around the spin-injection obstacle [124][126]. The size of the required barrier is substantially reduced relative to the expressions in the study in [124][126] by nonlinear drift-diffusion effects in semiconductors [109], [110]. Convincing success in spin injection through a highresistance barrier into a semiconductor was achieved a couple of years later [127][129] and has since reached 70% efciency at room temperature with an MgO barrier [69]. B. Coherent Spin Transport The formalism for coherent transport is much less developed than for incoherent transport; however, an intense research

effort is underway, motivated by the discovery of coherent spin-transport phenomena in both magnetic multilayers and nonmagnetic semiconductors. Recently, the linear nature of metallic transport in magnetic multilayers was given a new twist from the discovery of spin torque. In this effect, the ow of current through a CPP spin-valve structure with nonparallel ferromagnetic layers can cause the magnetization of the layers to precess or reorient [130], [131]. In one conguration, the rst ferromagnetic layer injects a spin-polarized distribution into the nonmagnetic metal and that spin-polarized distribution retains its polarization transverse to the orientation of the second ferromagnetic layer. Upon scattering from the second ferromagnetic layer, or entering into it, the transverse component is changed, and as a result, the magnetization of the second ferromagnetic layer experiences a torque [132]. Switching of small magnetic domains via this effect has been achieved, although the currents required are large [133], [134]. This mechanism may replace the applied magnetic elds required to reorient memory elements in MRAM with spin-polarized electrical currents. Coherent transport within nonmagnetic semiconductors in a magnetic eld has been directly visualized using spatially resolved pumpprobe measurements [96]. Coherent spin precession is also possible without an external magnetic eld through the effect of the internal spinorbit eld that produces spin decoherence in noncentrosymmetric semiconductors. These internal magnetic elds can cause spin precession during the coherent propagation of spin packets over micrometers in nonmagnetic semiconductors [103], [104], [135], [136]. Some of these results can be understood by a natural generalization of the two-channel spin-transport theory to precessing geometries [101], [103]. A more complete diffusive transport theory that tracks properly how spin polarization along one axis can evolve into a spin polarization along a perpendicular axis is grounded best as the evolution of a spin density matrix [137]. In the diffusive regime within such a theory, there naturally arise distinct length scales for the dynamics of the spin polarization parallel and perpendicular to an applied magnetic eld. C. Transistor Structures Once spin-polarized distributions can be generated, moved through magnetic and nonmagnetic semiconductor materials, manipulated, and detected, it is natural to explore the potential to use such distributions in multiterminal devices such as transistors. The rst such three-terminal device proposed was a semiconductor device using coherent spin transport that had incoherent charge inputs and outputs [138] and subsequent devices have extended the proposed range of device mechanism and device functionality. Since the proposal in [138], metallic three-terminal devices (where a third contact is made to the spacer layer in a CPP GMR conguration) [139] and hybrid devices, in which the base of a silicon transistor is replaced by a spin valve [140], have been built. Purely semiconductor proposals include unipolar spin transistors [102], [113], [141], magnetic p-n diodes [67], [111], [112], magnetic bipolar transistors [142], [143], and other spin eld-effect transistors [76], [144], [145].

916

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO. 5, MAY 2007

Fig. 2. Schematic comparison of a MOSFET and a spin-based FET. The MOSFET works by controlling the height of the barrier, which is up in (a) and down in (b). The barrier height and width is largely determined by the desired ONOFF current ratio and leakage current. The spin-based FET considered here works by controlling the nature of the initial state moving past xed barriers; if the carriers in the base are fully spin polarized, the transistor is (c) OFF, otherwise the transistor is (d) ON.

It is reasonable to question, however, whether there is any advantage to using spin-based devices to perform transistor action rather than charge-based devices. After all, several decades of extensive worldwide research and development have been devoted to optimizing MOSFETs. Is it realistic to expect that spin-based devices could compete? This question is best addressed separately, depending on whether the inputs of the spin device are incoherent charge inputs or not. For devices with incoherent charge inputs, the minimum switching energy identied initially by Landauer [146], Ebit = kT ln2 23 meV, still applies. Although the device itself does not function by modifying the pathways for charge motion, Landauers argument itself analyzes the modication of a charge on the gate of a transistor with incoherent charge inputs and outputs, independent of the mechanism for transistor action once that charge is established. The conditions for deriving this universal form for the switching energy, however, require that the switching is done slowly. As the size shrinks and the required speed increases for charge-based devices along the semiconductor roadmap [147], it is not clear that this fundamental limit will be reached. At the nal node of the 2003 roadmap, in 2018, the switching energy required for low-standby-power (LSTP) devices is 1500 eV/m. Even for a gate width comparable to the 10-nm gate length, the switching energy would be approximately three orders of magnitude larger than the fundamental limit. The static-power dissipation is another challenge in improving transistor-device performance that is related to the dynamic-power dissipation (from gate switching) in CMOS devices. Static-power dissipation in CMOS originates from sourcedrain leakage, which can be minimized by making the energetic barrier for the ow of carriers high and thick. This solution for static-power dissipation, however, degrades the dynamic-power-dissipation performance. Power dissipation from sourcedrain leakage is intrinsic to the MOSFET mechanism and, thus, would be considerably different for spin transistors. Using the spin polarization of the input and output current to encode information and the ability to generate an effective gate bias that is spin-dependent (through the generation of spin accumulation) to manipulate that information appears

likely to permit much more exibility in the device design than is possible using only incoherent charge inputs and outputs for a spin transistor and a charge potential on the gate. Although specic device congurations have not been extensively proposed, some capabilities that should prove useful include the ability to generate spin-polarized distributions in the absence of magnetic elds and magnetic materials [148], [149], to manipulate [103], [104], [150] and interfere [151] those distributions through transport, and to route spins in different directions based on their spin polarization [39][41]. Thus, many of the required elements have already been experimentally demonstrated for a logical architecture based on the coherent motion and manipulation of spin. As the most developed and analyzed proposals for competitive spin-logic devices are based on spin transistors with incoherent charge input, output, and gates, a deeper comparison of these with charge transistors is in order. Perhaps spin transistors will be able to approach this fundamental limit in smaller and faster transistors than charge-based transistors. To analyze this hypothesis, Fig. 2 shows a schematic comparison of the two transistor mechanisms. On the left is the mechanism for operation of a charge transistor. Here, a barrier is raised and lowered to turn the drain current OFF and ON. The barrier has to be sufciently high and thick to maintain a large ONOFF ratio for the drain current, however the higher and thicker it is, the larger the threshold voltage and the capacitance, and thus, the larger the gate switching energy will be. The spin eld-effect transistor, however, does not function by raising or lowering a barrier. As shown in Fig. 2(c) and (d) for one version of the spin eld-effect transistor [76], the barrier is identical in both the ON and OFF congurations. The key is that the barrier is thick and high for carriers of one spin polarization and is nonexistent for carriers of the opposite spin polarization. Carriers of one orientation are injected into the base region and, then, cannot proceed until they ip their spin. If the spin lifetime in the base is very long, then the amount of current owing out the drain is very small, whereas if the spin lifetime in the base is short, then the draincurrent is large. Modication of the spin lifetime to a length appropriate for the ON conguration requires a substantially smaller electric eld than raising or lowering the barrier for the MOSFET. A

FLATT: SPINTRONICS

917

1-meV spin splitting can cause a spin to completely reorient by precession in only 1 ps. Generating this spin relaxation indirectly via applying an electric eld and producing a Rashba eld to relax the spins through the spinorbit interaction implies that a larger voltage is required; for the structure in [76], a Vth 100 mV is sufcient to reduce the spin lifetime to 10 ps. The threshold voltage of an LSTP-CMOS MOSFET in 2018 is four times larger. The capacitance of the structure in [76] can also be considerably smaller than the capacitance of LSTP-CMOS, yielding an estimated power-delay product 500 times smaller. Estimates of the leakage current based on the long spin lifetimes observed in semiconductor quantum wells yield a leakage current six times smaller than LSTP-CMOS. Thus a spin-transistor design, with incoherent charge inputs, outputs, and gates, could be signicantly superior [152] to LSTP-CMOS even at the end of the roadmap node in 2018. Of course, signicant material advances would be required, including the development of highly efcient spin-selective barriers and the demonstration of a much larger range of spinlifetime tuning than achieved to date. V. CONCLUSION The new eld of spintronics was born in the intersection of magnetism, electronic transport, and optics. It has achieved commercial success in some areas and is advancing toward additional applications that rely on recent fundamental discoveries. The eld is sufciently broad that there is no single central obstacle to the application of these fundamental physical principles to new devices. Some of the advances that might be most helpful would be room-temperature demonstrations of: 1) a carrier-mediated ferromagnetic semiconductor whose density could be modied over a broad range by a gate eld; 2) the injection of nearly 100% spin-polarized current from a ferromagnetic metal into a nonmagnetic semiconductor; 3) a nonmagnetic material with a gate-tunable or optically tunable Spin Hall effect, such that the Spin Hall conductivity could be tuned from positive to negative; 4) gate-tunable spin lifetimes over several orders of magnitude; 5) highefciency minority carrier transport through a ferromagnetic semiconductor (such as electrons through Ga1x Mnx As); 6) a ferromagnetic semiconductor with very low optical loss; and 7) spin transport in Si or SiGe devices. These are, of course, only a small selection of the possible areas that would have a tremendous effect on spintronics research and on achieving the devices described here (and others). ACKNOWLEDGMENT The author would like to thank D. D. Awschalom, N. Samarth, and G. Vignale for helpful conversations. R EFERENCES
[1] M. N. Baibich, J. M. Broto, A. Fert, F. N. Van Dau, F. Petroff, P. Eitenne, G. Creuzet, A. Friederich, and J. Chazelas, Giant magnetoresistance of (001)Fe/(001)Cr magnetic superlattices, Phys. Rev. Lett., vol. 61, no. 21, pp. 24722475, Nov. 1988. [2] G. Binasch, P. Grnberg, F. Saurenbach, and W. Zinn, Enhanced magnetoresistance in layered magnetic structures with antiferromagnetic

[3] [4] [5] [6] [7] [8] [9]

[10] [11] [12] [13] [14] [15] [16] [17] [18] [19]

[20] [21] [22] [23] [24] [25] [26] [27]

[28]

[29]

interlayer exchange, Phys. Rev. B, Condens. Matter, vol. 39, no. 7, pp. 48284830, Mar. 1989. B. Dieny, V. S. Speriosu, S. S. P. Parkin, B. A. Gurney, D. R. Wilhoit, and D. Mauri, Giant magnetoresistive in soft ferromagnetic multilayers, Phys. Rev. B, Condens. Matter, vol. 43, no. 1, pp. 12971300, Jan. 1991. J. Daughton, J. Brown, E. Chen, R. Beech, A. Pohm, and W. Kude, Magnetic eld sensors using GMR multilayer, IEEE Trans. Magn., vol. 30, no. 6, pp. 46084610, Nov. 1994. W. P. Pratt, S.-F. Lee, J. M. Slaughter, R. Loloee, P. A. Schroeder, and J. Bass, Perpendicular giant magnetoresistances of Ag/Co multilayers, Phys. Rev. Lett., vol. 66, no. 23, pp. 30603063, Jun. 1991. S. F. Lee, W. P. Pratt, R. Loloee, P. A. Schroeder, and J. Bass, Fielddependent interface resistance of Ag/Co multilayers, Phys. Rev. B, Condens. Matter, vol. 46, no. 1, pp. 548551, Jul. 1992. M. Julliere, Tunneling between ferromagnetic lms, Phys. Lett. A, vol. 54, no. 3, pp. 225226, Sep. 1975. J. S. Moodera, L. R. Kinder, T. M. Wong, and R. Meservey, Large magnetoresistance at room temperature in ferromagnetic thin lm tunnel junctions, Phys. Rev. Lett., vol. 74, no. 16, pp. 32733276, Apr. 1995. S. S. P. Parkin, C. Kaiser, A. Panchula, P. M. Rice, B. Hughes, M. Samant, and S.-H. Yang, Giant tunnelling magnetoresistance at room temperature with MgO (100) tunnel barriers, Nat. Mater., vol. 3, no. 12, pp. 862867, Dec. 2004. W. H. Meiklejohn and C. P. Bean, New magnetic anisotropy, Phys. Rev., vol. 102, no. 5, pp. 14131414, Jun. 1956. C. P. Slichter, Principles of Magnetic Resonance. New York: Harper & Row, 1963. F. Meier and B. P. Zachachrenya, Optical Orientation: Modern Problems in Condensed Matter Science, vol. 8. Amsterdam, The Netherlands: North Holland, 1984. G. Prinz, Magnetoelectronics, Phys. Today, vol. 48, no. 4, pp. 5863, Apr. 1995. G. Prinz, Magnetoelectronics, Science, vol. 282, no. 5394, pp. 1660 1663, Nov. 1998. H. Ohno, Making nonmagnetic semiconductors ferromagnetic, Science, vol. 281, no. 5379, pp. 951955, Aug. 1998. D. D. Awschalom, M. E. Flatt, and N. Samarth, Spintronics, Sci. Amer., vol. 286, no. 6, pp. 6673, Jun. 2002. D. D. Awschalom, N. Samarth, and D. Loss, Eds., Semiconductor Spintronics and Quantum Computation, Heidelberg, Germany: Springer-Verlag, 2002. M. Ziese and M. J. Thornton, Eds., Spin Electronics, ser. Lecture Notes in Physics, vol. 569. Heidelberg, Germany: Springer-Verlag, 2001. S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von Molnr, M. L. Roukes, A. Y. Chtchelkanova, and D. M. Treger, Spintronics: A spin-based electronics vision for the future, Science, vol. 294, no. 5546, pp. 14881495, Nov. 2001. A. H. MacDonald, P. Schiffer, and N. Samarth, Ferromagnetic semiconductors: Moving beyond (Ga,Mn)As, Nat. Mater., vol. 4, no. 3, pp. 195202, Mar. 2005. T. Jungwirth, J. Sinova, J. Maek, J. Kuera, and A. H. MacDonald, Theory of ferromagnetic (III,Mn)V semiconductors, Rev. Mod. Phys., vol. 78, no. 3, pp. 809864, 2006. H. Munekata, H. Ohno, S. von Molnr, A. Segmller, L. L. Chang, and L. Esaki, Diluted magnetic IIIV semiconductors, Phys. Rev. Lett., vol. 63, no. 17, pp. 18491852, Oct. 1989. H. Ohno, A. Shen, F. Matsukura, A. Oiwa, A. Endo, S. Katsumoto, and Y. Iye, (Ga,Mn)As: A new diluted magnetic semiconductor based on GaAs, Appl. Phys. Lett., vol. 69, no. 3, pp. 363365, Jul. 1996. N. F. Mott, The electrical conductivity of transition metals, Proc. R. Soc. Lond. A, Math. Phys. Sci., vol. 153, no. 880, pp. 699717, 1936. A. Fert and I. A. Campbell, Two-current conduction in nickel, Phys. Rev. Lett., vol. 21, no. 16, pp. 11901192, Oct. 1968. R. E. Camley and J. Barna, Theory of giant magnetoresistance effects s in magnetic layered structures with antiferromagnetic coupling, Phys. Rev. Lett., vol. 63, no. 6, pp. 664667, Aug. 1989. J. Barna, A. Fuss, R. E. Camley, P. Grnberg, and W. Zinn, Novel s magnetoresistance effect in layered magnetic structures: Theory and experiment, Phys. Rev. B, Condens. Matter, vol. 42, no. 13, pp. 8110 8120, Nov. 1990. S. Dubois, L. Piraux, J. M. George, K. Ounadjela, J. L. Duvail, and A. Fert, Evidence for a short spin diffusion length in permalloy from the giant magnetoresistance of multilayered nanowires, Phys. Rev. B, Condens. Matter, vol. 60, no. 1, pp. 477484, Jul. 1999. M. Johnson, Spin accumulation in gold lms, Phys. Rev. Lett., vol. 70, no. 14, pp. 21422145, Apr. 1993.

918

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO. 5, MAY 2007

[30] Q. Yang, P. Holody, S.-F. Lee, L. L. Henry, R. Loloee, P. A. Schroeder, W. P. Pratt, and J. Bass, Spin ip diffusion length and giant magnetoresistance at low temperatures, Phys. Rev. Lett., vol. 72, no. 20, pp. 3274 3277, May 1994. [31] E. C. Stoner, Collective electron ferromagnetism. II. Energy and specic heat, Proc. R. Soc. Lond. A, Math. Phys. Sci., vol. 169, no. 938, pp. 339371, 1939. [32] J. Smit, The spontaneous Hall effect in ferromagnetics I, Physica, vol. 21, no. 610, pp. 877887, 1955. [33] J. M. Luttinger, Theory of the Hall effect in ferromagnetic substances, Phys. Rev., vol. 112, no. 3, pp. 739751, Nov. 1958. [34] L. Berger, Side-jump mechanism for the Hall effect of ferromagnets, Phys. Rev. B, Condens. Matter, vol. 2, no. 11, pp. 45594566, Dec. 1970. [35] L. Berger, Application of the side-jump model to the Hall effect and Nernst effect in ferromagnets, Phys. Rev. B, Condens. Matter, vol. 5, no. 5, pp. 18621870, Mar. 1972. [36] S. K. Lyo and T. Holstein, Side-jump mechanism for ferromagnetic Hall effect, Phys. Rev. Lett., vol. 29, no. 7, pp. 423425, Aug. 1972. [37] M. I. Dyakonov and V. I. Perel, Current-induced spin orientation of electrons in semiconductors, Phys. Lett. A, vol. 35, no. 6, pp. 459460, Jul. 1971. [38] J. E. Hirsch, Spin Hall effect, Phys. Rev. Lett., vol. 83, no. 9, pp. 1834 1837, Aug. 1999. [39] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Observation of the Spin Hall effect in semiconductors, Science, vol. 306, no. 5703, pp. 19101913, Dec. 2004. [40] J. Wunderlich, B. Kaestner, J. Sinova, and T. Jungwirth, Experimental observation of the Spin Hall effect in a two-dimensional spin-orbit coupled semiconductor system, Phys. Rev. Lett., vol. 94, no. 4, p. 047204, 2005. [41] V. Sih, R. C. Myers, Y. K. Kato, W. H. Lau, A. C. Gossard, and D. D. Awschalom, Spatial imaging of the Spin Hall effect and currentinduced polarization in two-dimensional electron gases, Nat. Phys., vol. 1, no. 1, pp. 3135, Oct. 2005. [42] S. O. Valenzuela and M. Tinkham, Direct electronic measurement of the Spin Hall effect, Nature, vol. 442, no. 7099, pp. 176179, Jul. 2006. [43] S. Zhang, Spin Hall effect in the presence of spin diffusion, Phys. Rev. Lett., vol. 85, no. 2, pp. 393396, Jul. 2000. [44] R. V. Shchelushkin and A. Brataas, Spin Hall effects in diffusive normal metals, Phys. Rev. B, Condens. Matter, vol. 71, no. 4, p. 045 123, Jan. 2005. [45] N. P. Stern, S. Ghosh, G. Xiang, M. Zhu, N. Samarth, and D. D. Awschalom, Current-induced polarization and the Spin Hall effect at room temperature, Phys. Rev. Lett., vol. 97, no. 12, p. 126 603, Sep. 2006. [46] R. J. Elliott, Theory of the effect of spin-orbit coupling on magnetic resonance in some semiconductors, Phys. Rev., vol. 96, no. 2, pp. 266 279, Oct. 1954. [47] Y. Yafet, g factors and spin-lattice relaxation of conduction electrons, Solid State Phys., vol. 14, pp. 298, 1963. [48] S. Schultz and C. Latham, Observation of electron spin resonance in copper, Phys. Rev. Lett., vol. 15, no. 4, pp. 148151, Jul. 1965. [49] L. Esaki, P. J. Stiles, and S. von Molnr, Magnetointernal eld emission in junctions of magnetic insulators, Phys. Rev. Lett., vol. 19, no. 15, pp. 852854, Oct. 1967. [50] J. S. Moodera, X. Hao, G. A. Gibson, and R. Meservey, Electron-spin polarization in tunnel junctions in zero applied eld with ferromagnetic EuS barriers, Phys. Rev. Lett., vol. 61, no. 5, pp. 637640, Aug. 1988. [51] X. Hao, J. S. Moodera, and R. Meservey, Spin-lter effect of ferromagnetic europium sulde tunnel barriers, Phys. Rev. B, Condens. Matter, vol. 42, no. 13, pp. 82358243, Nov. 1990. [52] T. Dietl, H. Ohno, F. Matsukura, J. Cibert, and D. Ferrand, Zener model description of ferromagnetism in zinc-blende magnetic semiconductors, Science, vol. 287, no. 5455, pp. 10191022, 2000. [53] K. W. Edmonds, K. Y. Wang, R. P. Campion, A. C. Neumann, N. R. S. Farley, B. L. Gallagher, and C. T. Foxon, High-Curietemperature Ga1x Mnx As obtained by resistance-monitored annealing, Appl. Phys. Lett., vol. 81, no. 26, pp. 49914993, Dec. 2002. [54] K. C. Ku, S. J. Potashnik, R. F. Wang, S. H. Chun, P. Schiffer, N. Samarth, M. J. Seong, A. Mascarenhas, E. Johnston-Halperin, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Highly enhanced Curie temperature in low-temperature annealed [Ga,Mn]As epilayers, Appl. Phys. Lett., vol. 82, no. 14, pp. 23022304, Apr. 2003. [55] P. W. Anderson, Theory of magnetic exchange interactions: Exchange in insulators and semiconductors, Solid State Phys., vol. 14, pp. 99214, 1963.

[56] D. Kitchen, A. Richardella, J.-M. Tang, M. E. Flatt, and A. Yazdani, Atom-by-atom substitution of Mn in GaAs and visualization of their hole-mediated interactions, Nature, vol. 442, no. 7101, pp. 436439, Jul. 2006. [57] J.-M. Tang and M. E. Flatt, Multiband tight-binding model of local magnetism in Ga1x Mnx As, Phys. Rev. Lett., vol. 92, no. 4, p. 047 201, 2004. [58] A. M. Yakunin, A. Yu. Silov, P. M. Koenraad, W. V. Roy, J. D. Boeck, J. H. Wolter, J.-M. Tang, and M. E. Flatt, Spatial structure of an individual Mn acceptor in GaAs, Phys. Rev. Lett., vol. 92, no. 21, p. 216 806, 2004. [59] A. M. Yakunin, A. Yu. Silov, P. M. Koenraad, J.-M. Tang, M. E. Flatt, W. V. Roy, J. D. Boeck, and J. H. Wolter, Spatial structure of MnMn acceptor pairs in GaAs, Phys. Rev. Lett., vol. 95, no. 25, p. 256 402, 2005. [60] M. Abolfath, T. Jungwirth, J. Brum, and A. H. MacDonald, Theory of magnetic anisotropy in III1x Mnx V ferromagnets, Phys. Rev. B, Condens. Matter, vol. 63, no. 5, p. 054 418, Feb. 2001. [61] T. Dietl, H. Ohno, and F. Matsukura, Hole-mediated ferromagnetism in tetrahedrally coordinated semiconductors, Phys. Rev. B, Condens. Matter, vol. 63, no. 19, p. 195 205, May 2001. [62] K. S. Burch, D. B. Shrekenhamer, E. J. Singley, J. Stephens, B. L. Sheu, R. K. Kawakami, P. Schiffer, N. Samarth, D. D. Awschalom, and D. N. Basov, Impurity band conduction in a high temperature ferromagnetic semiconductor, Phys. Rev. Lett., vol. 97, no. 8, p. 087 208, 2006. [63] H. Ohno, D. Chiba, F. Matsukura, T. Omiya, E. Abe, T. Dietl, Y. Ohno, and K. Ohtani, Electric-eld control of ferromagnetism, Nature, vol. 408, no. 6815, pp. 944946, Dec. 2000. [64] D. Chiba, M. Yamanouchi, F. Matsukura, and H. Ohno, Electrical manipulation of magnetization reversal in a ferromagnetic semiconductor, Science, vol. 301, no. 5635, pp. 943945, Aug. 2003. [65] C. Rster, T. Borzenko, C. Gould, G. Schmidt, L. W. Molenkamp, X. Liu, T. J. Wojtowicz, J. K. Furdyna, Z. G. Yu, and M. E. Flatt, Very large magnetoresistance in lateral ferromagnetic (Ga,Mn)As wires with nanoconstrictions, Phys. Rev. Lett., vol. 91, no. 21, p. 216 602, Nov. 2003. [66] B. Beschoten, P. A. Crowell, I. Malajovich, D. D. Awschalom, F. Matsukura, A. Shen, and H. Ohno, Magnetic circular dichroism studies of carrier-induced ferromagnetism in Ga1x Mnx As, Phys. Rev. Lett., vol. 83, no. 15, pp. 30733076, Oct. 1999. [67] P. Chen, J. Moser, P. Kotissek, J. Sadowski, M. Zenger, D. Weiss, and W. Wegscheider, All-electrical measurement of spin injection in a magnetic p-n junction diode, Phys. Rev. B, vol. 74, no. 24, p. 241 302 (R), Dec. 2006. [68] W. H. Butler, X.-G. Zhang, T. C. Schulthess, and J. M. MacLaren, Spindependent tunneling conductance of Fe|MgO|Fe sandwiches, Phys. Rev. B, Condens. Matter, vol. 63, no. 5, p. 054 416, Jan. 2001. [69] X. Jiang, R. Wang, R. M. Shelby, R. M. Macfarlane, S. R. Bank, J. S. Harris, and S. S. P. Parkin, Highly spin-polarized roomtemperature tunnel injector for semiconductor spintronics using MgO(100), Phys. Rev. Lett., vol. 94, no. 5, p. 056 601, 2005. [70] G. Salis, R. Wang, X. Jiang, R. M. Shelby, S. S. P. Parkin, S. R. Bank, and J. S. Harris, Temperature independence of the spin-injection efciency of a MgO-based tunnel spin injector, Appl. Phys. Lett., vol. 87, no. 26, p. 262 503, 2005. [71] G. Dresselhaus, Spinorbit coupling effects in zinc blende structures, Phys. Rev., vol. 100, no. 2, pp. 580586, Oct. 1955. [72] M. I. Dyakonov and V. I. Perel, Spin relaxation of conduction electrons in noncentrosymmetric semiconductors, Sov. Phys.Solid State, vol. 13, no. 12, pp. 30233026, 1972. [73] M. I. Dyakonov and V. Y. Kachorovskii, Spin relaxation of twodimensional electrons in noncentrosymmetric semiconductors, Sov. Phys.Semicond., vol. 20, no. 1, pp. 110112, 1986. [74] R. Winkler, Spin orientation and spin precession in inversionasymmetric quasi-two-dimensional electron systems, Phys. Rev. B, Condens. Matter, vol. 69, no. 4, p. 045 317, Jan. 2004. [75] Y. Ohno, R. Terauchi, T. Adachi, F. Matsukura, and H. Ohno, Spin relaxation in GaAs(110) quantum wells, Phys. Rev. Lett., vol. 83, no. 20, pp. 41964199, Nov. 1999. [76] K. Hall, W. H. Lau, K. Gndo du, M. E. Flatt, and T. F. Boggess, g Nonmagnetic semiconductor spin transistor, Appl. Phys. Lett., vol. 83, no. 14, pp. 29372939, Oct. 2003. [77] E. I. Rashba, Properties of semiconductors with an extremum loop I. Cyclotron and combinational resonance in a magnetic eld perpendicular to the plane of the loop, Sov. Phys.Solid State, vol. 2, no. 6, pp. 11091122, 1960.

FLATT: SPINTRONICS

919

[78] Y. A. Bychkov and E. I. Rashba, Oscillatory effects and the magnetic susceptibility of carriers in inversion layers, J. Phys. C, Solid State Phys., vol. 17, no. 33, pp. 60396045, 1984. [79] W. H. Lau and M. E. Flatt, Tunability of electron spin coherence in IIIV quantum wells, J. Appl. Phys., vol. 91, no. 10, pp. 86828684, May 2002. [80] O. Z. Karimov, G. H. John, R. T. Harley, W. H. Lau, M. E. Flatt, M. Henini, and R. Airey, High temperature gate control of quantum well spin memory, Phys. Rev. Lett., vol. 91, no. 24, p. 246 601, 2003. [81] K. Hall, K. Gndo du, J. L. Hicks, A. N. Kocbay, M. E. Flatt, g T. F. Boggess, K. Holabird, A. Hunter, D. H. Chow, and J. J. Zinck, Room-temperature electric-eld controlled spin dynamics in (110) InAs quantum wells, Appl. Phys. Lett., vol. 86, no. 20, p. 202114, May 2005. [82] C. E. Pryor and M. E. Flatt, Accuracy of circular polarization as a measure of spin polarization in quantum dot qubits, Phys. Rev. Lett., vol. 91, no. 25, p. 257 901, 2003. [83] T. Uenoyama and L. J. Sham, Hole relaxation and luminescence polarization in doped and undoped quantum wells, Phys. Rev. Lett., vol. 64, no. 25, pp. 30703073, Jun. 1990. [84] R. J. Seymour and R. R. Alfano, Time-resolved measurement of the electron-spin relaxation kinetics in GaAs, Appl. Phys. Lett., vol. 37, no. 2, pp. 231233, Jul. 1980. [85] T. C. Damen, K. Leo, J. Shah, and J. E. Cunningham, Spin relaxation and thermalization of excitons in GaAs quantum wells, Appl. Phys. Lett., vol. 58, no. 17, pp. 19021904, Apr. 1991. [86] M. Kohl, M. R. Freeman, D. D. Awschalom, and J. M. Hong, Femtosecond spectroscopy of carrier-spin relaxation in GaAsAlx Ga1x As quantum wells, Phys. Rev. B, Condens. Matter, vol. 44, no. 11, pp. 59235926, Sep. 1991. [87] S. Bar-Ad and I. Bar-Joseph, Exciton spin dynamics in GaAs heterostructures, Phys. Rev. Lett., vol. 68, no. 3, pp. 349352, Jan. 1992. [88] J. Wagner, H. Schneider, D. Richards, A. Fischer, and K. Ploog, Observation of extremely long electron-spin-relaxation times in p-type -doped GaAs/Alx Ga1x As double heterostructures, Phys. Rev. B, Condens. Matter, vol. 47, no. 8, pp. 47864789, Feb. 1993. [89] M. Oestreich, S. Hallstein, A. P. Heberle, K. Eberl, E. Bauser, and W. W. Rhle, Temperature and density dependence of the electron Land g factor in semiconductors, Phys. Rev. B, Condens. Matter, vol. 53, no. 12, pp. 79117916, Mar. 1996. [90] A. Tackeuchi, S. Muto, T. Inata, and T. Fuji, Direct observation of picosecond spin relaxation of excitons in GaAs/AlGaAs quantum wells using spin-dependent optical nonlinearity, Appl. Phys. Lett., vol. 56, no. 22, pp. 22132215, May 1990. [91] R. S. Britton, T. Grevatt, A. Malinowski, R. T. Harley, P. Perozzo, A. R. Cameron, and A. Miller, Room temperature spin relaxation in GaAs/AlGaAs multiple quantum wells, Appl. Phys. Lett., vol. 73, no. 15, pp. 21402142, 1998. [92] K. C. Hall, S. W. Leonard, H. M. van Driel, A. R. Kost, E. Selvig, and D. H. Chow, Subpicosecond spin relaxation in GaAsSb multiple quantum wells, Appl. Phys. Lett., vol. 75, no. 26, pp. 41564158, 1999. [93] S. A. Crooker, J. J. Baumberg, F. Flack, N. Samarth, and D. D. Awschalom, Terahertz spin precession and coherent transfer of angular momenta in magnetic quantum wells, Phys. Rev. Lett., vol. 77, no. 13, pp. 28142817, Sep. 1996. [94] J. M. Kikkawa, I. P. Smorchkova, N. Samarth, and D. D. Awschalom, Room-temperature spin memory in two-dimensional electron gases, Science, vol. 277, no. 5330, pp. 12841287, Aug. 1997. [95] J. M. Kikkawa and D. D. Awschalom, Resonant spin amplication in ntype GaAs, Phys. Rev. Lett., vol. 80, no. 19, pp. 43134316, May 1998. [96] J. M. Kikkawa and D. D. Awschalom, Lateral drag of spin coherence in gallium arsenide, Nature, vol. 397, no. 6715, pp. 139141, Jan. 1999. [97] I. Malajovich, J. J. Berry, N. Samarth, and D. D. Awschalom, Persistent sourcing of coherent spins for multifunctional semiconductor spintronics, Nature, vol. 411, no. 6839, pp. 770772, Jun. 2001. [98] A. R. Cameron, P. Riblet, and A. Miller, Spin gratings and the measurement of electron drift mobility in multiple quantum well semiconductors, Phys. Rev. Lett., vol. 76, no. 25, pp. 47934796, Jun. 1996. [99] Y. Nishikawa, A. Tackeuchi, S. Nakamura, S. Muto, and N. Yokoyama, All-optical picosecond switching of a quantum well etalon using spinpolarization relaxation, Appl. Phys. Lett., vol. 66, no. 7, pp. 839841, Feb. 1995. [100] E. J. Gansen, K. Jarasiunas, and A. Smirl, Femtosecond all-optical polarization switching based on the virtual excitation of spin-polarized excitons in quantum wells, Appl. Phys. Lett., vol. 80, no. 6, pp. 971 973, Feb. 2002. [101] M. E. Flatt and J. M. Byers, Spin diffusion in semiconductors, Phys. Rev. Lett., vol. 84, no. 18, pp. 42204223, May 2000.

[102] M. E. Flatt and G. Vignale, Unipolar spin diodes and transistors, Appl. Phys. Lett., vol. 78, no. 9, pp. 12731275, Feb. 2001. [103] S. A. Crooker and D. L. Smith, Imaging spin ows in semiconductors subject to electric, magnetic and strain elds, Phys. Rev. Lett., vol. 94, no. 23, p. 236 601, 2005. [104] S. A. Crooker, M. Furis, X. Lou, C. Adelmann, D. L. Smith, C. J. Palmstrm, and P. A. Crowell, Imaging spin transport in lateral ferromagnet/semiconductor structures, Science, vol. 309, no. 5744, pp. 21912195, Sep. 2005. [105] C. Kittel and H. Kroemer, Thermal Physics. New York: Freeman, 1980. [106] P. C. van Son, H. van Kempen, and P. Wyder, Boundary resistance of the ferromagneticnonferromagnetic metal interface, Phys. Rev. Lett., vol. 58, no. 21, pp. 22712273, May 1987. [107] S. Hersheld and H. L. Zhao, Charge and spin transport through a metallic ferromagneticparamagneticferromagnetic junction, Phys. Rev. B, Condens. Matter, vol. 56, no. 6, pp. 32963305, Aug. 1997. [108] A. G. Aronov and G. E. Pikus, Spin injection into semiconductors, Sov. Phys.Semicond., vol. 10, no. 6, pp. 698700, 1976. [109] Z. G. Yu and M. E. Flatt, Electric-eld dependent spin diffusion and spin injection into semiconductors, Phys. Rev. B, Condens. Matter, vol. 66, no. 20, p. 201202 (R), Nov. 2002. [110] Z. G. Yu and M. E. Flatt, Spin diffusion and injection in semiconductor structures: Electric eld effects, Phys. Rev. B, Condens. Matter, vol. 66, no. 23, p. 235 302, Dec. 2002. [111] I. uti , J. Fabian, and S. Das Sarma, Spin-polarized transport c in inhomogeneous magnetic semiconductors: Theory of magnetic/ nonmagnetic p-n junctions, Phys. Rev. Lett., vol. 88, no. 6, p. 066 603, Jan. 2002. [112] J. Fabian, I. uti , and S. D. Sarma, Theory of spin-polarized bipolar c transport in magnetic p-n junctions, Phys. Rev. B, Condens. Matter, vol. 66, no. 16, p. 165 301, Oct. 2002. [113] M. E. Flatt and G. Vignale, Heterostructure unipolar spin transistors, J. Appl. Phys., vol. 97, no. 10, p. 104508, May 2005. [114] P. M. Levy, S. Zhang, and A. Fert, Electrical conductivity of magnetic multilayered structures, Phys. Rev. Lett., vol. 65, no. 13, pp. 16431646, Sep. 1990. [115] A. Barthlmy and A. Fert, Theory of the magnetoresistance in magnetic multilayers: Analytical expressions from a semiclassical approach, Phys. Rev. B, Condens. Matter, vol. 43, no. 16, pp. 13 124 13 129, Jun. 1991. [116] W. H. Butler, X.-G. Zhang, T. C. Schulthess, D. M. C. Nicholson, J. M. MacLaren, V. S. Speriosu, and B. A. Gurney, Conductance and giant magnetoconductance of Co|Cu|Co spin valves: Experiment and theory, Phys. Rev. B, Condens. Matter, vol. 56, no. 22, pp. 14 574 14 582, Dec. 1997. [117] M. Johnson, Analysis of anomalous multilayer magnetoresistance within the thermomagnetoelectric system, Phys. Rev. Lett., vol. 67, no. 25, pp. 35943597, Dec. 1991. [118] T. Valet and A. Fert, Theory of the perpendicular magnetoresistance in magnetic multilayers, Phys. Rev. B, Condens. Matter, vol. 48, no. 10, pp. 70997113, Sep. 1993. [119] M. Johnson and R. H. Silsbee, Spin-injection experiment, Phys. Rev. B, Condens. Matter, vol. 37, no. 10, pp. 53265335, Apr. 1988. [120] G. Schmidt, D. Ferrand, L. W. Molenkamp, A. T. Filip, and B. J. van Wees, Fundamental obstacle for electrical spin injection from a ferromagnetic metal into a diffusive semiconductor, Phys. Rev. B, Condens. Matter, vol. 62, no. 8, pp. R4790R4793, Aug. 2000. [121] R. Fiederling, M. Keim, G. Reuscher, W. Ossau, G. Schmidt, A. Waag, and L. W. Molenkamp, Injection and detection of a spin-polarized current in a light-emitting diode, Nature, vol. 402, no. 6763, pp. 787 790, 1999. [122] Y. Ohno, D. K. Young, B. Beschoten, F. Matsukura, H. Ohno, and D. D. Awschalom, Electrical spin injection in a ferromagnetic semiconductor heterostructure, Nature, vol. 402, no. 6763, pp. 790792, Dec. 1999. [123] D. Hgele, M. Oestreich, W. W. Rhle, N. Nestle, and K. Eberl, Spin transport in GaAs, Appl. Phys. Lett., vol. 73, no. 11, pp. 15801582, Sep. 1998. [124] E. I. Rashba, Theory of electrical spin injection: Tunnel contacts as a solution of the conductivity mismatch problem, Phys. Rev. B, Condens. Matter, vol. 62, no. 24, pp. R16 267R16 270, Dec. 2000. [125] D. L. Smith and R. N. Silver, Electrical spin injection into semiconductors, Phys. Rev. B, Condens. Matter, vol. 64, no. 4, p. 045 323, Jul. 2001. [126] A. Fert and H. Jaffrs, Conditions for efcient spin injection from a ferromagnetic metal into a semiconductor, Phys. Rev. B, Condens. Matter, vol. 64, no. 18, p. 184 420, Oct. 2001.

920

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 54, NO. 5, MAY 2007

[127] A. T. Hanbicki, B. T. Jonker, G. Itskos, G. Kioseoglou, and A. Petrou, Efcient electrical spin injection from a magnetic metal/tunnel barrier contact into a semiconductor, Appl. Phys. Lett., vol. 80, no. 7, pp. 1240 1242, Feb. 2002. [128] V. F. Motsnyi, J. D. Boeck, J. Das, W. V. Roy, G. Borghs, E. Goovaerts, and V. I. Safarov, Electrical spin injection in a ferromagnet/tunnel barrier/semiconductor heterostructure, Appl. Phys. Lett., vol. 81, no. 2, pp. 265267, Jul. 2002. [129] C. Adelmann, X. Lou, J. Strand, C. J. Palmstrm, and P. A. Crowell, Spin injection and relaxation in ferromagnetsemiconductor heterostructures, Phys. Rev. B, Condens. Matter, vol. 71, no. 12, p. 121 301 (R), 2005. [130] L. Berger, Emission of spin waves by a magnetic multilayer traversed by a current, Phys. Rev. B, Condens. Matter, vol. 54, no. 13, pp. 9353 9358, Oct. 1996. [131] J. Slonczewski, Current-driven excitation of magnetic multilayers, J. Magn. Magn. Mater., vol. 159, no. 1/2, pp. L1L7, Jun. 1996. [132] M. Tsoi, A. G. M. Jansen, J. Bass, W.-C. Chiang, M. Seck, V. Tsoi, and P. Wyder, Excitation of a magnetic multilayer by an electric current, Phys. Rev. Lett., vol. 80, no. 19, pp. 42814284, May 1998. [133] E. B. Myers, D. C. Ralph, J. A. Katine, R. N. Louie, and R. A. Buhrman, Current-induced switching of domains in magnetic multilayer devices, Science, vol. 285, no. 5429, pp. 867870, Aug. 1999. [134] J. A. Katine, F. J. Albert, R. A. Buhrman, E. B. Myers, and D. C. Ralph, Current-driven magnetization reversal and spin-wave excitations in Co/Cu/Co pillars, Phys. Rev. Lett., vol. 84, no. 14, pp. 31493152, Apr. 2000. [135] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Coherent spin manipulation without magnetic elds in strained semiconductors, Nature, vol. 427, no. 6969, pp. 5053, Jan. 2004. [136] B. A. Bernevig and S.-C. Zhang, Spin splitting and spin current in strained bulk semiconductors, Phys. Rev. B, Condens. Matter, vol. 72, no. 11, p. 115 204, Sep. 2005. [137] Y. Qi and S. Zhang, Spin diffusion at nite electric and magnetic elds, Phys. Rev. B, Condens. Matter, vol. 67, no. 5, p. 052407, Feb. 2003. [138] S. Datta and B. Das, Electronic analog of the electro-optic modulator, Appl. Phys. Lett., vol. 56, no. 7, pp. 665667, Feb. 1990. [139] M. Johnson, Bipolar spin switch, Science, vol. 260, no. 5106, pp. 320 323, Apr. 1993. [140] D. J. Monsma, R. Vlutters, and J. C. Lodder, Room temperatureoperating spin-valve transistors formed by vacuum bonding, Science, vol. 281, no. 5375, pp. 407409, 1998. [141] M. Deutsch, G. Vignale, and M. E. Flatt, Effect of electrical bias on spin transport across a magnetic domain wall, J. Appl. Phys., vol. 96, no. 12, pp. 74247427, Dec. 2004. [142] M. E. Flatt, Z. G. Yu, E. Johnston-Halperin, and D. D. Awschalom, Theory of semiconductor magnetic bipolar transistors, Appl. Phys. Lett., vol. 82, no. 26, pp. 47404742, Jun. 2003.

[143] J. Fabian, I. uti , and S. D. Sarma, Magnetic bipolar transistor, Appl. c Phys. Lett., vol. 84, no. 1, pp. 8587, Jan. 2004. [144] J. Schliemann, J. C. Egues, and D. Loss, Nonballistic spin-eld-effect transistor, Phys. Rev. Lett., vol. 90, no. 14, p. 146 801, Apr. 2003. [145] X. Cartoi` a, D. Z.-Y. Ting, and Y.-C. Chang, A resonant spin lifetime x transistor, Appl. Phys. Lett., vol. 83, no. 7, pp. 14621464, Aug. 2003. [146] R. Landauer, Irreversibility and heat generation in the computing process, IBM J. Res. Develop., vol. 5, no. 3, pp. 183191, 1961. [147] International Technology Roadmap for Semiconductors. (2003). Semicond. Ind. Assoc., San San Jose, CA. [Online]. Available: http://public.itrs.net [148] Y. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Current induced spin polarization in strained semiconductors, Phys. Rev. Lett., vol. 93, no. 17, p. 176 601, 2004. [149] A. Y. Silov, P. A. Blajnov, J. H. Wolter, R. Hey, K. H. Ploog, and N. S. Averkiev, Current-induced spin polarization at a single heterojunction, Appl. Phys. Lett., vol. 85, no. 24, pp. 59295931, Dec. 2004. [150] Y. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Electrical initialization and manipulation of electron spins in an L-shaped strained n-InGaAs channel, Appl. Phys. Lett., vol. 87, no. 2, p. 022 503, Jul. 2005. [151] Y. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Electron spin interferometry using a semiconductor ring structure, Appl. Phys. Lett., vol. 86, no. 16, p. 162107, Apr. 2005. [152] K. Hall and M. E. Flatt, Performance of a spin-based insulated gate eld effect transistor, Appl. Phys. Lett., vol. 88, no. 16, p. 162 503, Apr. 2006.

Michael E. Flatt (M01) received the A.B. degree in physics from Harvard University, Cambridge, MA, in 1988, and the Ph.D. degree in physics from the University of California, Santa Barbara (UCSB), in 1992. He was a Scholar in Residence at iQUEST, UCSB, during the academic year 20002001. After postdoctoral positions in the Institute for Theoretical Physics, UCSB, and in the Division of Applied Sciences, Harvard University, he joined the faculty of The University of Iowa, Iowa City, where he is currently a Professor in the Department of Physics and Astronomy and the Department of Electrical and Computer Engineering. He is also currently the Associate Director of the Optical Science and Technology Center, The University of Iowa.

Вам также может понравиться