An Intrinsic Spin Orbit Torque Nano-Oscillator

Similar documents
Mesoscopic Spintronics

Spatiotemporal magnetic imaging at the nanometer and picosecond scales

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

Lecture I. Spin Orbitronics

0.002 ( ) R xy

Optical studies of current-induced magnetization

MSE 7025 Magnetic Materials (and Spintronics)

Observation of the intrinsic inverse spin Hall effect in Ni 80 Fe 20. Yuichiro Ando, Teruya Shinjo and Masashi Shiraishi * #

Spin orbit torque driven magnetic switching and memory. Debanjan Bhowmik

Focused-ion-beam milling based nanostencil mask fabrication for spin transfer torque studies. Güntherodt

Magnetic oscillations driven by the spin Hall effect in 3-terminal magnetic tunnel junction. devices. Cornell University, Ithaca, NY 14853

Mesoscopic Spintronics

Temporal Evolution of Auto-Oscillations in an Yttrium- Iron-Garnet/Platinum Microdisk Driven by Pulsed Spin Hall Effect-Induced Spin-Transfer Torque

TRANSVERSE SPIN TRANSPORT IN GRAPHENE

Spin-torque nano-oscillators trends and challenging

Electric power transfer in spin pumping experiments

Spin transfer torque devices utilizing the giant spin Hall effect of tungsten

arxiv: v1 [cond-mat.mtrl-sci] 28 Jul 2008

SPIN TRANSFER TORQUES IN HIGH ANISOTROPY MAGNETIC NANOSTRUCTURES

Magneto-Seebeck effect in spin-valve with in-plane thermal gradient

arxiv:cond-mat/ v1 4 Oct 2002

Magnon-drag thermopile

High-frequency measurements of spin-valve films and devices invited

Expecting the unexpected in the spin Hall effect: from fundamental to practical

4.5 YIG thickness dependence of the spin-pumping effect in YIG/Pt heterostructures

Center for Spintronic Materials, Interfaces, and Novel Architectures. Voltage Controlled Antiferromagnetics and Future Spin Memory

Lecture I. Spin Orbitronics

Spin-orbit torque in Pt/CoNiCo/Pt symmetric devices

Supplementary figures

Spin Funneling for Enhanced Spin Injection into Ferromagnets: Supplementary Information

An Overview of Spintronics in 2D Materials

Introduction to Spintronics and Spin Caloritronics. Tamara Nunner Freie Universität Berlin

Spin pumping in Ferromagnet-Topological Insulator-Ferromagnet Heterostructures Supplementary Information.

Recent developments in spintronic

Observation of the intrinsic inverse spin Hall effect from ferromagnet

Enhanced spin orbit torques by oxygen incorporation in tungsten films

Determination of the Interfacial Dzyaloshinskii-Moriya Interaction (idmi) in the Inversion Symmetry Broken Systems

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION

Strong Spin Hall Effect in the Antiferromagnet PtMn. Abstract

Exchange Splitting of Backward Volume Spin Wave Configuration Dispersion Curves in a Permalloy Nano-stripe

S. Mangin 1, Y. Henry 2, D. Ravelosona 3, J.A. Katine 4, and S. Moyerman 5, I. Tudosa 5, E. E. Fullerton 5

Ferromagnetism and Electronic Transport. Ordinary magnetoresistance (OMR)

Current-induced switching in a magnetic insulator

Magnetic droplet solitons generated by pure spin currents

Cover Page. The handle holds various files of this Leiden University dissertation

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

Room-temperature perpendicular magnetization switching through giant spin-orbit torque from sputtered Bi x Se (1-x) topological insulator material

Unidirectional spin-wave heat conveyer

Current-induced asymmetric magnetoresistance due to energy transfer via

Current-driven ferromagnetic resonance, mechanical torques and rotary motion in magnetic nanostructures

Spin dynamics in Bi 2 Se 3 /ferromagnet heterostructures

Low Energy Spin Transfer Torque RAM (STT-RAM / SPRAM) Zach Foresta April 23, 2009

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

SUPPLEMENTARY INFORMATION

Magnetoresistance due to Domain Walls in Micron Scale Fe Wires. with Stripe Domains arxiv:cond-mat/ v1 [cond-mat.mes-hall] 9 Mar 1998.

Spin-orbit driven ferromagnetic resonance: A nanoscale magnetic characterisation technique. Abstract

THEORY OF SPIN HALL EFFECT AND INTERFACE SPIN-ORBIT COUPLING

Mesoscopic Spintronics

Two-dimensional mutual synchronization in spin Hall nano-oscillator arrays

Room temperature spin-orbit torque switching induced by a

Spin injection and absorption in antiferromagnets

From Spin Torque Random Access Memory to Spintronic Memristor. Xiaobin Wang Seagate Technology

Spin wave assisted current induced magnetic. domain wall motion

Mutual synchronization of nano-oscillators driven by pure spin current

Spin-transfer-torque through antiferromagnetic IrMn

Chapter 103 Spin-Polarized Scanning Tunneling Microscopy

Resonance Measurement of Nonlocal Spin Torque in a Three-Terminal Magnetic Device

Spin Orbit Coupling (SOC) in Graphene

SUPPLEMENTARY INFORMATION

Controlled enhancement of spin current emission by three-magnon splitting

Mon., Feb. 04 & Wed., Feb. 06, A few more instructive slides related to GMR and GMR sensors

Time resolved transport studies of magnetization reversal in orthogonal spin transfer magnetic tunnel junction devices

Fermi level dependent charge-to-spin current conversion by Dirac surface state of topological insulators

Current-induced Domain Wall Dynamics

Quantitative characterization of spin-orbit torques in Pt/Co/Pt /Co/Ta/BTO heterostructure on the magnetization azimuthal angle dependence

Supplementary material: Nature Nanotechnology NNANO D

Spin pumping in magnetic trilayer structures with an MgO barrier Supplementary Information.

arxiv: v1 [cond-mat.mtrl-sci] 21 Dec 2017

Ferromagnetic resonance in Yttrium Iron Garnet

Mutual synchronization of spin torque nano-oscillators through a non-local and tunable electrical coupling

Gauge Concepts in Theoretical Applied Physics

Observation of an anti-damping spin-orbit torque originating from the Berry curvature

Dynamically-generated pure spin current in single-layer graphene

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

Nonlinear Electrodynamics and Optics of Graphene

Spin Current and Spin Seebeck Effect

Anisotropic Current-Controlled Magnetization Reversal in the Ferromagnetic Semiconductor (Ga,Mn)As

Cover Page. The handle holds various files of this Leiden University dissertation

Magnetic resonance studies of the fundamental spin-wave modes in individual submicron Cu/NiFe/Cu perpendicularly magnetized disks.

Systèmes Hybrides. Norman Birge Michigan State University

Supplementary material for : Spindomain-wall transfer induced domain. perpendicular current injection. 1 ave A. Fresnel, Palaiseau, France

Putting the Electron s Spin to Work Dan Ralph Kavli Institute at Cornell Cornell University

Magnetism (Spins) Seen in a New Light. Muhammad Sabieh Anwar

Spin pumping and spin transport in magne0c metal and insulator heterostructures. Eric Montoya Surface Science Laboratory Simon Fraser University

Large-amplitude coherent spin waves excited by spin-polarized current in nanoscale spin valves

Damping of magnetization dynamics

Switching of perpendicular magnetization by spin-orbit torques in the absence of. external magnetic fields. California 90095, United States

Spintronics KITP, UCSB. Interfacial spin-orbit coupling and chirality. Hyun-Woo Lee (POSTECH, KOREA)

Transcription:

arxiv:188.933v1 [cond-mat.mes-hall] 28 Aug 218 An Intrinsic Spin Orbit Torque Nano-Oscillator M. Haidar, 1 A. A. Awad, 1 M. Dvornik, 1 R. Khymyn, 1 A. Houshang, 1 J. Åkerman 1,2 1 Physics Department, University of Gothenburg, 412 96 Gothenburg, Sweden 2 Material Physics, School of Engineering Sciences, KTH Royal Institute of Technology, Electrum 229, 164 4 Kista,Sweden To whom correspondence should be addressed; E-mail: johan.akerman@physics.gu.se. Spin torque and spin Hall effect nano-oscillators generate high intensity spin wave auto-oscillations on the nanoscale enabling novel microwave applications in spintronics, magnonics, and neuromorphic computing. For their operation, these devices require externally generated spin currents either from an additional ferromagnetic layer or a material with a high spin Hall angle. Here we demonstrate highly coherent field and current tunable microwave signals from nano-constrictions in single 15 2 nm thick permalloy layers. Using a combination of spin torque ferromagnetic resonance measurements, scanning micro-brillouin light scattering microscopy, and micromagnetic simulations, we identify the auto-oscillations as emanating from a localized edge mode of the nano-constriction driven by spin-orbit torques. Our results pave the way for greatly simplified designs of auto-oscillating nano-magnetic systems only requiring a single ferromagnetic layer. 1

Introduction Spin transfer torque (1, 2) (STT) the transfer of angular momentum from a spin current to the local spins can act as negative spin wave damping and drive the magnetization of a ferromagnet (FM) into a state of sustained large-angle precessionfor STT to occur, one first needs to generate a spin current and then inject this spin current into a ferromagnetic layer with a magnetization direction that is non-collinear with the polarization axis of the spin current. In so-called spin torque nano-oscillators (STNOs) these conditions can be fulfilled either in giant magnetoresistance trilayers, or in magnetic tunnel junctions, if the two ferromagnetic layers are in a non-collinear state (3). More recently, the spin Hall effect (4 6) was instead used to create pure spin currents injected into an adjacent ferromagnetic layer, in devices known as spin Hall nano-oscillators (SHNOs) (3). These are both much easier to fabricate (7), and exhibit superior synchronization properties and therefore much higher signal coherence (8). Even so, SHNOs still suffer from a number of drawbacks and conflicting requirements: i) the non-magnetic layer generating the spin current dramatically increases the zero-current spin wave damping of the ferromagnetic layer (up to 3x for NiFe/Pt (9)), ii) the surface nature of the spin Hall effect requires ultrathin ferromagnetic layers for reasonable threshold currents, which further increases the spin wave damping, and iii) since the current is shared between the driving layer and the magnetoresistive ferromagnetic layer, neither driving nor signal generation benefits from the total current, leading to non-optimal threshold currents, low output power, unnecessary dissipation, and heating. It would therefore be highly advantageous if one could do away with the additional metal layers for operation. It was recently realized that spin-orbit torque can emerge from Berry curvature due to a broken inversion symmetry (1), either in the bulk of the magnetic material (11, 12) or at its interfaces. In particular, interfaces to insulating oxides (13 17) can lead to a non-uniform spin 2

distribution over the film thickness (18 2). As oxide spin-orbitronics has seen considerable development in recent years (21,22), oxide based spin-orbit torque nano-oscillators with a single metallic layer could be envisioned. If so, the current would be confined to the ferromagnetic metal layer, avoiding the detrimental current sharing between adjacent metal layers and leading to an overall lower power consumption. Here we propose, and experimentally demonstrate, an intrinsic spin orbit torque (SOT) nano-oscillator based on a single permalloy layer with thickness in the 15 2 nm range, grown on an Al 2 O 3 substrate and capped with SiO 2. The magnetodynamics of the unpatterned stacks is characterized using ferromagnetic resonance (FMR) spectroscopy and the linear spin wave modes of the final devices are determined using spin-torque FMR. Spin wave auto-oscillations are observed both electrically, via a microwave voltage from the anisotropic magnetoresistance, and optically, using micro-brillouin Light Scattering (µ-bls) microscopy. The current-field symmetry of the observed auto-oscillations agree with the expected symmetry for spin orbit torques, either intrinsic to the Py layer itself or originating from the Al 2 O 3 /Py and Py/SiO 2 interfaces. Fig. 1A shows a scanning electron microscopy image of a w = 3 nm wide nano-constriction etched out of t = 15 nm thick single permalloy (Ni 8 Fe 2 ; Py) layer, magnetron sputtered onto a sapphire substrate. On top of the nano-constriction a coplanar waveguide is fabricated for electrical measurements (not shown). During operation, an external magnetic field (H), applied at an out-of-plane angle θ = 75 and an in-plane angle φ = 22, tilts the Py magnetization out of plane. Fig. 1B shows the thickness dependence of the Gilbert damping (α) of the unprocessed Py films as well as the resistance (R) and the anisotropic magnetoresistance (AMR) of the final devices (inset). As the thickness of the Py layer increases, we measure a noticeable decrease of α and R, and a corresponding increase of AMR (23). Thus, using thicker permalloy films could 3

be advantageous as compared to thinner films where by reducing α one can reduce the threshold current for driving spin wave auto-oscillations and by increasing AMR one can achieve larger output power as the electrical read-out is based on AMR in these devices. To determine the magnetodynamic properties of the nano-constrictions, we first characterized them in the linear regime using spin torque ferromagnetic resonance spectroscopy (ST- FMR) (24) in a field applied at θ = 75 and φ = 22. A typical ST-FMR spectrum as a function of field magnitude, and at constant frequency f = 1 GHz, reveals two well-separated resonances that can both be accurately fit with Lorentzians (Fig. 1C). The frequency vs. field dependence of the lower field peak agrees perfectly with the FMR dispersion at the specific angles used, from which we can extract an effective magnetization µ M eff =.72 T. We can directly image the spatial distribution of the two linear modes using micro-brillouin light scattering (µ-bls) microscopy of the thermally excited SWs: the amplitude of the FMR mode is the largest well outside the nano-constriction region (Fig. 1D) whereas the amplitude of the higher-field (lower energy) mode is localized to the center of the nano-constriction (Fig. 1E). The localized mode is confined to the nano-constriction due to a local reduction of the internal magnetic field in this region. The nature of the observed modes is further corroborated by carrying out micromagnetic simulations where we calculate the spatial profile of the magnetization within the nano-constriction. The spatial maps for the simulated FMR mode and the localized mode (Fig. 1F, G) closely agree with the µ-bls maps. We then investigated the impact of a direct current on the linewidths ( H) of both the FMR and the localized modes (Fig. 2A). While the linewidths of both modes decreases with current magnitude for an even current-field symmetry (the orientation of the field and current satisfy H I > ) and increases for an odd symmetry (H I < ), the localized mode (blue circles) is three times more strongly affected than the FMR mode. This is again consistent with the localized mode residing in the nano-constriction region where the current density is 4

B SEM Damping, R, & MR 2 AMR R α (x 1-3 ) w = 3 nm z H θ φ x y 15 1 D ST-FMR 3.5 2 1 2 3 t (nm) 4 1 nm C 4.8.4 NiFe 6 AMR (%) 1.2 R (Ω) A 1 BLS 2 3 t (nm) F 4 Simula on Vmix (μv) FMR 2.5 G E Localized 1.5.5.6.7.8.9 2 nm Figure 1: Nano-constriction fabrication, characterization and simulation. (A) Scanning electron microscope image of a permalloy nano-constriction showing its width (w = 3 nm) and the direction of the applied field. (B) Gilbert damping of blanket permalloy films as a function of film thickness. Inset: AMR (closed symbol) and resistance (open symbol) vs. thickness of 3 nm wide nano-constrictions. (C) ST-FMR spectrum vs. field at f = 1 GHz for a w = 3 nm and t = 15 nm nano-constriction showing its two linear modes; red and blue lines are Lorentzian fits to the FMR mode and the localized mode, respectively. The spatial distribution of the FMR and the localized modes, imaged by µ-bls (D,E), and calculated using micromagnetic simulations (F,G). 5

the largest. Naively extrapolating the H vs. current data to zero linewidth then predicts that auto-oscillations on this mode should be possible at a current magnitude of 2 ma. That auto-oscillations are indeed possible to achieve at such currents is experimentally confirmed by sweeping the field magnitude from in Fig. 2B&C. The electrical microwave power spectral density (PSD) is measured at constant currents of I = 2 ma (Fig. 2B) and I = 2 ma (Fig. 2C) as the field is swept from.8 T to -.8 T. Auto-oscillations are detected under positive applied fields between +1 koe and +8 koe with a frequency tunable from 3.5 to 12 GHz. Somewhat unexpectedly, a measurable PSD is also observed under negative fields between -1 koe and -3 koe where the frequency increases from 3.5 to 5.5 GHz. Then, by switching the direction of the current, (Fig. 2C), we measure a corresponding continuous branch of auto-oscillations under negative fields between -1 koe and -8 koe and, again, a noticeable PSD appears under a positive field between +1 koe and +3 koe, this time at an essentially constant frequency of about 4 GHz. We categorize the measured PSD into two regions depending on the strength of the magnetic field: region (I) for low fields H 3 koe and region (II) for high fields H > 3 koe. In region (I) the PSD is independent of the field-current symmetry, i.e. a measureable PSD is always present regardless of the field and current polarities. The microwave signal in this region is rather weak with a large linewidth. In region (II) the auto-oscillations are both sharper and more powerful and are detected only when the orientation of the field and current satisfy H I >. In this region, the spin current creates a spin-torque that is aligned either anti-parallel (H I > ) or parallel (H I < ) to the damping and hence it either suppresses or enhances the damping depending on symmetry. To better understand the origin of the two different regions and their distinctly different symmetries we performed µ-bls measurements vs. field and current. Fig. 2D shows the frequency vs. field of the thermally populated linear modes at I = ma, where we observe the localized 6

A C B12 ΔH (Oe) 15 1 -H 5 +H ΔH = 94.5-41.9*I -1 D Frequency (GHz) ΔH = 133.5-15.1*I I (ma) II I 4 II==-2 +2mA ma -.8 -.4 1 E I = ma 8 II. II I I = -2 ma.8 -.8 -.4.4 F I = 1.2 ma..8.4 I = -1.2 ma x13 6 II 8 x12 4 2. I 4-2.2.4.6. I (ma).4.2 6 4 2 2.6 BLS counts Frequency (GHz) 12..2-2.4 I (ma) 2.6 Figure 2: Current induced spin wave auto-oscillations. (A) Current dependent ST-FMR linewidth of the FMR (red circles) and localized (blue circles) modes for positive field (filled symbols) and negative field (open symbols). (B, C) Electrical measurements of the power spectral densities of spin wave auto-oscillations vs. applied field measured under a direct current of (B) +2 ma and (C) -2 ma. Label I and II indicate low-field and high-field auto-oscillations, respectively. (D, E, F) µ-bls measurements of the frequency vs. applied field showing the intensity of the FMR and the localized mode measured at ma (D), +1.2 ma (E), and -1.2 ma (F). The insets in (E, F) show the change of the BLS counts as a function of current at µ H =.7 T and µ H =.2 T, respectively. 7

mode and the continuous SW band above the FMR frequency. When we increase the current to 1.2 ma (Fig. 2E) a noticeable increase in the intensity of the localized mode can be observed for high positive fields, i.e. in region (II). It is noteworthy that the SW amplitude in this region depends exponentially on current magnitude, as shown in the inset of (Fig. 2E), consistent with spin-torque driven auto-oscillations. In contrast, by switching the current to 1.2 ma the intensity of the localized mode in region (II) becomes weaker than at zero current, consistent with additional damping from spin-torque, while it increases slightly a low fields (region (I)). The current dependence in region (I) is however much weaker and essentially linear, suggesting thermal activation, but not auto-oscillations. We can hence conclude from the µ-bls characterization that the observed auto-oscillations originate in a continuous manner from the localized mode inside the nano-constriction and that the auto-oscillations require a certain magnetic state reached only for a field above.3 T. These results were reproducible from device to device and auto-oscillations were observed for both t = 15 and 2 nm. Next, we discuss the possible mechanism that could be driving the auto-oscillations, i.e. for which, the spin current should exert an opposite torque to the Gilbert damping with the observed current-field symmetry. Since our devices have a strongly non-uniform magnetization distribution in the vicinity of the constriction and should have a relatively strong temperature gradient during operation, we might have anti-damping originating from (i) Zhang-Li torque (25), and (ii) spin torque from the spin dependent Seebeck effect (26). However, neither of these torques follows the experimentally observed H I > symmetry, as their sign would not depend on the direction of the applied current. The observed symmetry instead strongly suggests that spin-orbit torque (SOT) is the main source of anti-damping. Recent results obtained in ferromagnetic/oxide heterostructures suggest that SOT can result from non-equilibrium spin accumulation due to inversion symmetry breaking and enhanced spin-orbit coupling at the interfaces (1, 27, 28). In principle, intrinsic 8

anomalous (29, 3) and spin Hall effects in Py (19, 2) and interfacial Rashba effect (13) might all contribute to the SOT in our system. In fact, our micromagnetic simulations, assuming a spin Hall/Rashba-like SOT, require a spin Hall angle of about.13 to reproduce the auto-oscillation threshold currents observed in the experiments, which is in the range of values reported for Py/oxide stacks (22). Besides the fundamental interest of generating, controlling, and optimizing SOT in single ferromagnetic layers, e.g. using ultra-low damping metals and different oxide interfaces, our results will have a direct impact on a wide range of applications. A magnetic tunnel junction (MTJ) in the nano-constriction region only requires an additional ferromagnetic layer, which will allow for separate optimization of the SOT drive and the MTJ based high-power read-out. Arrays of SOT driven nano-constrictions can also be used for oscillator based neuromorphic computing (31). In the sub-threshold regime, the intrinsic damping in magnonic crystals can be greatly reduced, solving their issues with high transmission losses (32). Acknowledgments: This work was supported by the European Research Council (ERC) under the European Communitys Seventh Framework Programme (FP/27-213)/ERC Grant 37144 MUSTANG. This work was also supported by the Swedish Research Council (VR), the Swedish Foundation for Strategic Research (SSF), and the Knut and Alice Wallenberg Foundation, and the Wenner-Gren Foundation. 9

References and Notes 1. L. Berger, Phys. Rev. B 54, 9353 (1996). 2. J. Slonczewski, J. Magn. Magn. Mater. 159, L1 (1996). 3. T. Chen, et al., Proc. IEEE 14, 1919 (216). 4. J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999). 5. M. I. D yakonov, V. I. Perel, J. Exp. Theor. Phys. Lett. 13, 467 (1971). 6. J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, T. Jungwirth, Rev. Mod. Phys. 87, 1213 (215). 7. V. E. Demidov, S. Urazhdin, A. Zholud, A. V. Sadovnikov, S. O. Demokritov, Appl. Phys. Lett. 15, 17241 (214). 8. A. A. Awad, et al., Nat. Phys. 13, 292 (217). 9. T. Nan, et al., Phys. Rev. B 91, 214416 (215). 1. H. Kurebayashi, et al., Nature nanotechnology 9, 211 (214). 11. D. Fang, et al., Nature nanotechnology 6, 413 (211). 12. C. Safranski, E. A. Montoya, I. N. Krivorotov, arxiv preprint arxiv:1712.7335 (217). 13. I. M. Miron, et al., Nat. Mat. 1, 419 (211). 14. H. An, Y. Kageyama, Y. Kanno, N. Enishi, K. Ando, Nature communications 7, 1369 (216). 15. S. Emori, et al., Phys. Rev. B 93, 1842 (216). 1

16. K.-U. Demasius, et al., Nature communications 7, 1644 (216). 17. T. Gao, et al., Phys. Rev. Lett. 121, 1722 (218). 18. M. Haidar, M. Bailleul, Phys. Rev. B 88, 54417 (213). 19. A. Tsukahara, et al., Phys. Rev. B 89, 235317 (214). 2. A. Azevedo, et al., Phys. Rev. B 92, 2442 (215). 21. D. C. Vaz, A. Barthlmy, M. Bibes, Japanese Journal of Applied Physics 57, 92A4 (218). 22. A. Manchon, et al., arxiv preprint arxiv:181.9636 (218). 23. S. Ingvarsson, et al., Phys. Rev. B 66, 214416 (22). 24. L. Liu, T. Moriyama, D. C. Ralph, R. A. Buhrman, Phys. Rev. Lett. 16, 3661 (211). 25. S. Zhang, Z. Li, Phys. Rev. Lett. 93, 12724 (24). 26. A. Slachter, F. L. Bakker, J.-P. Adam, B. J. van Wees, Nature Physics 6, 879 (21). 27. K. M. D. Hals, A. Brataas, Y. Tserkovnyak, EPL (Europhysics Letters) 9, 472 (21). 28. K.-W. Kim, K.-J. Lee, J. Sinova, H.-W. Lee, M. D. Stiles, Phys. Rev. B 96, 14438 (217). 29. K. S. Das, W. Y. Schoemaker, B. J. van Wees, I. J. Vera-Marun, Phys. Rev. B 96, 2248 (217). 3. A. Bose, et al., Phys. Rev. Applied 9, 6426 (218). 31. J. Torrejon, et al., Nature 547, 428 (217). 32. V. V. Kruglyak, S. O. Demokritov, D. Grundler, Journal of Physics D: Applied Physics 43, 2641 (21). 11