Low-power non-volatile spintronic memory: STT-RAM and beyond

Similar documents
Perpendicular MTJ stack development for STT MRAM on Endura PVD platform

NONVOLATILE SPINTRONICS: PERSPECTIVES ON INSTANT-ON NONVOLATILE NANOELECTRONIC SYSTEMS

10. Magnetoelectric Switching

A Technology-Agnostic MTJ SPICE Model with User-Defined Dimensions for STT-MRAM Scalability Studies

MRAM: Device Basics and Emerging Technologies

Nonvolatile CMOS Circuits Using Magnetic Tunnel Junction

Lecture 6 NEW TYPES OF MEMORY

Gate voltage modulation of spin-hall-torque-driven magnetic switching. Cornell University, Ithaca, NY 14853

Wouldn t it be great if

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

Electric-Field-Controlled Magnetoelectric RAM: Progress, Challenges, and Scaling

MSE 7025 Magnetic Materials (and Spintronics)

Compact Modeling of STT-RAM and MeRAM A Verilog-A model of Magnetic Tunnel Junction Behavioral Dynamics

An Overview of Spin-based Integrated Circuits

Author : Fabrice BERNARD-GRANGER September 18 th, 2014

MTJ-Based Nonvolatile Logic-in-Memory Architecture and Its Application

New Approaches to Reducing Energy Consumption of MRAM write cycles, Ultra-high efficient writing (Voltage-Control) Spintronics Memory (VoCSM)

Two-terminal spin orbit torque magnetoresistive random access memory

SPIN TRANSFER TORQUES IN HIGH ANISOTROPY MAGNETIC NANOSTRUCTURES

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

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

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

A Perpendicular Spin Torque Switching based MRAM for the 28 nm Technology Node

Page 1. A portion of this study was supported by NEDO.

Switching Properties in Magnetic Tunnel Junctions with Interfacial Perpendicular Anisotropy: Micromagnetic Study

Spin-transfer-torque efficiency enhanced by edge-damage. of perpendicular magnetic random access memories

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

Spin orbit torque driven magnetic switching and memory. Debanjan Bhowmik

Magnetization Dynamics in Spintronic Structures and Devices

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

Mesoscopic Spintronics

Enhanced spin orbit torques by oxygen incorporation in tungsten films

Advanced Lab Course. Tunneling Magneto Resistance

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

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

This document is an author-formatted work. The definitive version for citation appears as:

Embedded MRAM Technology For logic VLSI Application

Spin-torque nano-oscillators trends and challenging

9. Spin Torque Majority Gate

Optical studies of current-induced magnetization

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 63, NO. 11, NOVEMBER

Advanced Topics In Solid State Devices EE290B. Will a New Milli-Volt Switch Replace the Transistor for Digital Applications?

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

0.002 ( ) R xy

Nanoelectronics 12. Atsufumi Hirohata Department of Electronics. Quick Review over the Last Lecture

SUPPLEMENTARY INFORMATION

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

V High frequency magnetic measurements

CHAPTER 2 MAGNETISM. 2.1 Magnetic materials

introduction: what is spin-electronics?

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

Spintronics. Seminar report SUBMITTED TO: SUBMITTED BY:

Spin Hall effect clocking of nanomagnetic logic without a magnetic field

Magnetic tunnel junction beyond memory from logic to neuromorphic computing WANJUN PARK DEPT. OF ELECTRONIC ENGINEERING, HANYANG UNIVERSITY

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

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

SUPPLEMENTARY INFORMATION

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

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

SPICE Modeling of STT-RAM for Resilient Design. Zihan Xu, Ketul Sutaria, Chengen Yang, Chaitali Chakrabarti, Yu (Kevin) Cao School of ECEE, ASU

Voltage effects in poly and single-crystal 3d ferromagnetic metal/mgo systems

Influence of exchange bias on magnetic losses in CoFeB/MgO/CoFeB tunnel junctions

Shape anisotropy revisited in single-digit

X-Ray Spectro-Microscopy Joachim Stöhr Stanford Synchrotron Radiation Laboratory

Emerging spintronics-based logic technologies

From Hall Effect to TMR

Ultrafast switching of a nanomagnet by a combined out-of-plane and in-plane polarized spin-current pulse

arxiv:cond-mat/ v1 4 Oct 2002

arxiv: v1 [physics.app-ph] 1 May 2017

MSE 7025 Magnetic Materials (and Spintronics)

Introduction to magnetic recording + recording materials

Spin-transfer switching and thermal stability in an FePt/Au/FePt nanopillar prepared by alternate monatomic layer deposition

MESL: Proposal for a Non-volatile Cascadable Magneto-Electric Spin Logic

TRANSVERSE SPIN TRANSPORT IN GRAPHENE

Current-Induced Magnetization Switching in MgO Barrier Based Magnetic Tunnel. Junctions with CoFeB/Ru/CoFeB Synthetic Ferrimagnetic Free Layer

Solid-State Electronics

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

AN ABSTRACT OF THE THESIS OF

Magnetic Race- Track Memory: Current Induced Domain Wall Motion!

The contribution of hot-electron spin polarization to the spin-dependent magnetotransport in a spin-valve transistor at finite temperatures

The exchange interaction between FM and AFM materials

Supplementary Figure 1. Magnetic domain configuration under out-of-plane field application. (a), (b) MTXM images showing magnetic domain state

Advanced Flash and Nano-Floating Gate Memories

Magnetic Tunnel Junction for Integrated Circuits: Scaling and Beyond

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

Room Temperature Planar Hall Transistor

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

Injecting, Controlling, and Storing Magnetic Domain Walls in Ferromagnetic Nanowires

GaAs and InGaAs Single Electron Hex. Title. Author(s) Kasai, Seiya; Hasegawa, Hideki. Citation 13(2-4): Issue Date DOI

Improving STT-MRAM Density Through Multibit Error Correction

Challenges for Materials to Support Emerging Research Devices

Adaptive Compact Magnetic Tunnel Junction Model

Spin Torque and Magnetic Tunnel Junctions

A Universal Memory Model for Design Exploration. Ketul Sutaria, Chi-Chao Wang, Yu (Kevin) Cao School of ECEE, ASU

SUPPLEMENTARY INFORMATION

Spin-Based Computing: Device Concepts, Current Status, and a Case Study on a High-Performance Microprocessor

Ferromagnetism and Electronic Transport. Ordinary magnetoresistance (OMR)

On the Ultimate Speed of Magnetic Switching

JUNCTION LEAKAGE OF A SiC-BASED NON-VOLATILE RANDOM ACCESS MEMORY (NVRAM) K. Y. Cheong ABSTRACT INTRODUCTION

Contents. Acknowledgments

Transcription:

IOP PUBLISHING JOURNAL OF PHYSICS D: APPLIED PHYSICS J. Phys. D: Appl. Phys. 46 (2013) 074003 (10pp) doi:10.1088/0022-3727/46/7/074003 Low-power non-volatile spintronic memory: STT-RAM and beyond K L Wang, J G Alzate and P Khalili Amiri Electrical Engineering Department, University of California, Los Angeles, CA 90095, USA Received 15 August 2012, in final form 18 October 2012 Published 31 January 2013 Online at stacks.iop.org/jphysd/46/074003 Abstract The quest for novel low-dissipation devices is one of the most critical for the future of semiconductor technology and nano-systems. The development of a low-power, universal memory will enable a new paradigm of non-volatile computation. Here we consider STT-RAM as one of the emerging candidates for low-power non-volatile memory. We show different configurations for STT memory and demonstrate strategies to optimize key performance parameters such as switching current and energy. The energy and scaling limits of STT-RAM are discussed, leading us to argue that alternative writing mechanisms may be required to achieve ultralow power dissipation, a necessary condition for direct integration with CMOS at the gate level for non-volatile logic purposes. As an example, we discuss the use of the giant spin Hall effect as a possible alternative to induce magnetization reversal in magnetic tunnel junctions using pure spin currents. Further, we concentrate on magnetoelectric effects, where electric fields are used instead of spin-polarized currents to manipulate the nanomagnets, as another candidate solution to address the challenges of energy efficiency and density. The possibility of an electric-field-controlled magnetoelectric RAM as a promising candidate for ultralow-power non-volatile memory is discussed in the light of experimental data demonstrating voltage-induced switching of the magnetization and reorientation of the magnetic easy axis by electric fields in nanomagnets. (Some figures may appear in colour only in the online journal) 1. Introduction The continuous scaling and development of CMOS has spurred a technological revolution, allowing for faster and more powerful microprocessors and electronics with an increased number of functionalities [1, 2]. CMOS is currently the dominating technology for logic circuits, but is quickly approaching its scaling limits due to increased problems with power dissipation at scaled technology nodes. The increase in power dissipation results from the increase in static leakage (standby) power, as well as from the increase in density as the device size is scaled down. The integration of a fast, energyefficient non-volatile memory technology with CMOS can help alleviate this problem. Non-volatile logic circuits consisting of volatile CMOS, combined with non-volatile memory, can make electronic products non-volatile at the device, gate, circuit and system levels, hence allowing for continued scaling with improved energy efficiency by eliminating static power dissipation. In the memory hierarchy, typically the information is temporally stored in static random access memories (SRAMs) as cache memory next to the CMOS logic, and in dynamic random access memories (DRAMs) as the principal working memory, and then permanently stored in Flash or hard drives, used as non-volatile memories for long-term retention. Table 1 shows the most important metrics for SRAM, DRAM and NOR Flash. SRAM is the fastest of the three (operating in the GHz range), needs little dynamic power ( 100 fj per switch), has unlimited endurance, and can be incorporated directly on the logic chip. However, SRAM is volatile, has a very low density (resulting in a higher cost per bit) and the standby power consumption becomes a problem as the transistor dimensions are scaled down, leading to higher leakage currents. DRAM has a higher density compared with SRAM, but it is also volatile and needs a periodic refresh, which results in power consumption for maintaining the memory state. Also, the conventional DRAM fabrication process is not compatible with standard CMOS logic. The NOR flash memories have the highest density among the three, and are non-volatile. However, they have very slow write speed and very limited endurance. They also use high power for writing data and high internal voltages are needed for their operation. 0022-3727/13/074003+10$33.00 1 2013 IOP Publishing Ltd Printed in the UK & the USA

Table 1. Comparison of existing non-spintronic and emerging spintronic memory technologies, highlighting STT-RAM (current-induced switching) and MeRAM (voltage-controlled switching). Production processes In development Technology SRAM DRAM NOR Flash STT-RAM MeRAM Energy/bit (fj) 100 1000 10 6 100 <10 Write speed (ns) 1 20 1000 1 10 1 10 Read speed (ns) 1 30 10 1 10 1 10 Density (area in F 2 ) >30 6 10 4 8 10 30 4 8 Endurance (cycles) Very High Very High Low Very High Very High Non-volatile No No Yes Yes Yes Standby power Leakage Current Refresh Current None None None Cost overhead versus CMOS Large area (6T) Separate process Separate process Back-end (BEOL) process Back-end (BEOL) process Non-volatile logic capability No No No Very limited due to power Yes Spintronic devices, i.e. those which exploit the exchange interaction of the electron spins, where magnetic and transport properties are coupled, are strong candidates for non-volatile memory due to the inherent hysteresis in ferromagnetic materials, and the compatibility of some of these materials with the standard CMOS process [3 5]. In fact, magnetoresistive RAM (MRAM) has exhibited significant advantages as a fast, fairly low-power, high-endurance, radiation-resistant non-volatile memory, which can be integrated to the CMOS as a back-end of line (BEOL) process [6]. Usually, the read-out of the MRAM bits is reliably performed via the tunnelling magnetoresistance (TMR) [7, 8] effect in magnetic tunnel junctions (MTJs). As for the writing process, the first generations of MRAM utilized the Oersted fields generated by running currents in adjacent conducting lines of memory cells to switch the magnetization. However, this strategy has proven to be highly inconvenient in terms of energy efficiency, scalability and density. Therefore, the manipulation of the magnetic moments by currents or electric voltages is required to overcome the shortcomings of Oersted-fieldswitched MRAM. In this work, we will first concentrate on the switching of magnetization of MRAM bits using spin-polarized currents via the spin transfer torque (STT) effect [7, 9, 10] in order to achieve low power operation. As observed in table 1, the use of spin-polarized currents instead of Oersted fields allows for switching energies and speed close to SRAM, while the density can be better than SRAM and potentially approach close to DRAM levels, making STT-RAM a suitable candidate for embedded as well as potentially standalone applications. More importantly, STT-RAM has better scalability as compared with Oersted-field-switched MRAM, where the switching current increases as the MTJ dimensions are reduced. We will examine strategies to optimize the performance for STT-RAM devices in this paper. In the second part of this paper, we will study beyond STT mechanisms, targeting for ultralow power performance (switching energies <1 fj). This is a critical challenge since the switching energies of STT-RAM are still around two orders of magnitude higher compared with CMOS switching energy ( 1 fj for the 32 nm node), limiting the possibility of integration of MTJ devices at the gate level with CMOS logic for non-volatile logic architectures [11]. Additionally, the ultralow power MTJs could be used to relieve the increasing static leakage power, as the transistors are scaled, due to the continuous energy that needs to be delivered to retain the information in the volatile CMOS [12]. Hence, ultralow-power voltage-controlled MRAM could impart a significant overall energy advantage over STT-RAM based circuits. For ultra-low power spin-based memories beyond STT-RAM, this paper will discuss two approaches: first, we will consider magnetoelectric effects which manipulate the magnetization by electric fields (or voltages) instead of the current-driven mechanisms [13, 14]. In particular, the ultimate scalability of STT-RAM in terms of density and energy may be limited by the large currents needed to switch the device; thus, a voltage control scheme is expected to translate into a reduction in the switching energy and increased densities in a future magnetoelectric RAM (MeRAM), while keeping the advantages of STT-RAM (See table 1) [15]. Finally, we will discuss the giant spin Hall effect as an alternative mechanism to reduce the switching currents for STT-RAM by driving pure spin currents through MTJ devices [16]. 2. STT as switching mechanism for MTJs The discovery of the possibility to manipulate and induce switching of the magnetization by spin-polarized currents via the STT effect [9, 10] in nanomagnets has been the key motivating force behind the recent rise of interest in MRAM, and particularly STT-RAM [17, 18]. STT-RAM conserves the advantages of Oersted-field-switched (toggle) MRAM (high speed, very high endurance, non-volatility), but the currentdriven switching mechanism makes STT-RAM more scalable and allows for smaller switching energies (on the order of 100 fj) compared with Oersted-field-switched MRAM, as will be discussed later. A typical simplified STT-RAM structure is shown in figure 1: it consists of a free layer which can take two states (parallel or antiparallel) to a pinned (fixed) layer, both of the layers separated by a tunnelling oxide, which is usually MgO to allow for read-out via the TMR (tunnelling magnetoresistance) effect [6, 19]. The writing process is performed by passing a spin-polarized current, which transfers some of its momentum to the nanomagnet, inducing a torque that can result in switching depending on the direction of the current. 2

stability factor of = 40 translates into a mean retention time 10 years for a discrete device. Specifically, the thermal stability factor is given by = M s H k V/2kT, where M s is the saturation magnetization, H k is the magnetic anisotropy field, V is the volume of the free layer, k is the Boltzmann constant and T is the temperature. For I-STT and C-STT devices, H k is mainly determined by the in-plane shape anisotropy (i.e. the energetic preference for the magnetization to align along a preferred direction as determined by the shape of the magnetic bit). Therefore, as the volume of the MTJ is decreased in I-STT and C-STT devices, H k needs to be increased, for example, by increasing the aspect ratio of the ellipses or the thickness of the free layer to retain stability. For the I-STT configuration, the switching current density is given by [20] J c0 = (2eαM s t/ hη)(h k + H d /2), (1) Figure 1. Typical architecture for a 1 transistor 1 MTJ (1T-1R) memory cell. The MTJ is composed of a bi-stable free layer and a pinned layer separated by a tunnelling oxide. The device is fabricated as part of the back-end (BEOL) process, compatible with CMOS logic processes. Therefore, the combination of STT and TMR gives rise to the memory loops observed in the quasi-static I V relation as well as the pulsed R V characteristics of figure 2 (left) and (right), respectively. In this section we will consider three different designs for STT-RAM devices (referred to as I-STT, C-STT and P-STT standing for in-plane, combined and perpendicular), which offer different characteristics and performances due to their configuration of magnetization direction and anisotropies of the free and fixed layers. In our discussion, we will consider MTJ devices due to their high magnetoresistance ratio and the possibility of tuning the MgO thickness for impedance optimization for energy efficiency and compatibility with the access transistors used in a memory implementation. We will particularly emphasize strategies to increase speed and enhance energy efficiency during switching. 2.1. In-plane free layers Two of the device configurations where the magnetization of the free layer lies in the plane of the thin film (inplane configuration) are shown in figures 3(a) and (b). The equilibrium states are determined by the shape anisotropy along the long (easy) axis of the nanopillar. In the I-STT (inplane STT) configuration, the polarizer (pinned layer) will be in the same plane as the free layer, while the C-STT (combination- STT) structure (also referred to as orthogonal STT-RAM) includes an additional polarizer for reasons to be discussed later in this section. The simplest MTJ structure to realize is I-STT, and has been thus studied the most over the past years. In I-STT devices, one of the key challenges is to reduce the switching current density J c0 while keeping a reasonable thermal stability factor for non-volatility. The value of defines the magnetic bit s stability against false switching events due to thermal activation (i.e. retention time); for example, a thermal where α is the free layer damping factor, η is the spin-transfer efficiency, M s and t are the free layer saturation magnetization and thickness, H k is the in-plane shape-induced anisotropy field and H d 4πM s H k is the out-of-plane demagnetizing field. Scaling of I-STT devices generally requires a tradeoff between the switching current density J c0 and the thermal stability factor, where the goal is to minimize the ratio J c0 / while preserving a given goal for the thermal stability factor. I-STT suffers from a small spin torque during the initial incubation stage of magnetization switching, a consequence of its parallel-magnetized equilibrium states. Its switching speed is generally limited to 1ns [21, 22]. Figure 4 shows the typical switching voltage and energy dependence on pulse width in an I-STT device for pulses >1 ns. An alternative to eliminate the initial incubation of I-STT devices is to include an additional out-of-plane polarizer (see figures 3(c) and (d)) [21 23]. This C-STT configuration uses a combination of inplane and perpendicular polarizers, where only the in-plane polarizer provides TMR for read-out of the in-plane free layer state. In C-STT devices, spin torque from the perpendicular polarizer causes the free layer magnetization to go out-ofplane, where it precesses around the demagnetizing field [24, 25], allowing for ultrafast precessional switching limited only by the ferromagnetic resonance (FMR) frequency of the bit. As observed in the macrospin simulations of figures 3(c) and (d), the torque provided by the out-of-plane polarizer eliminates the initial incubation resulting in precessional switching on the order of 100 ps. Further, switching with pulses down to 50 ps has been demonstrated also in spin-valve structures [26]. Note that the in-plane polarizer is still required for read-out and also provides additional torque. If we take into account that the write energy is given by E = V 2 t/r, the reduction in the required pulse time t outpaces the potential increase in the required voltage for C-STT, resulting in a reduction in the switching energy larger than 50% as compared with an I-STT device with an otherwise similar layered structure (see figure 4). To further reduce the write energy in STT-RAM, one can perform further optimization on both the free layer as well as the MgO barrier. If the switching energy is written as E = V 2 t/r = JC 2A2 Rt, it is evident that it can be further decreased in I-STT devices by (1) reducing the resistance 3

Figure 2. (Left) Example of current-induced switching results in STT-RAM. Quasi-static V I curves show the existence of two available resistance states, defining the memory window for STT-RAM devices. The free layer of the MTJ can be switched parallel or anti-parallel to the pinned layer depending on the direction of the current. (Right) The pulsed R V curve demonstrates switching of the MTJ with voltage pulses of 1 ns with voltages 0.5 V. The read-out of the state is performed via the TMR effect with a typical TMR ratio >100%. (Reprinted with permission from [27]. Copyright 2011, American Institute of Physics.) Figure 3. (a) The layered structure for I-STT (IST) device, including a synthetic antiferromagnet (SAF) to compensate for the dipole fields from the pinned layer acting on the free layer. (b) The inclusion of an additional perpendicular polarizer results in a C-STT (OST) configuration, where the additional polarizer exerts a large initial torque for the free layer switching. (c) Macrospin simulations showing the intrinsic limitation of I-STT (IST) due to the incubation time required for a switching event. (d) The perpendicular polarizer in C-STT (OST) results in precessional switching, hence eliminating the initial incubation, allowing for switching with pulses of the order of 100 ps. (Reprinted with permission from [22]. Copyright 2011, American Institute of Physics.) R of the MTJ, or by (2) decreasing the switching current density. The first strategy is limited by practical reasons, since reducing the resistance of the MTJ requires thinner MgO barriers, which may result in reduction in the TMR ratio, endurance or breakdown voltage, as well as increases in undesired device-to-device variability. Furthermore, scaling of STT-RAM generally requires moving to thinner MgO barriers to maintain CMOS-compatible voltage levels (i.e. reducing the RA product to prevent R from increasing), and hence MgO thickness is generally set by considerations other than energy. Figure 5 (left) shows the dependences of resistance area (RA) product and TMR on the MgO barrier thickness in CoFeB MgO MTJs, measured on a single wafer with varying MgO thicknesses [27, 28]. It is observed that the TMR rapidly decreases when the thickness of the MgO becomes less than 0.8 nm, whereas the variability of the RA product increases for the same range of tunnelling barrier thicknesses. Figure 5 (right) shows the write energy and switching current density as a function of RA product in an optimized I-STT CoFeB/MgO/CoFeB tunnel junction. Although the expected reduction in the switching energy with decreasing RA product is observed, the switching current also increases (possibly as a result of higher effective damping in the free layer due to the reduced MgO quality for low RA devices), further limiting the strategy of reducing the resistance of the MTJ to obtain smaller switching energies. 4

cancelled and therefore, the switching current density would be given by J c0 = 2eαM s H k t/ hη, while the switching current over thermal stability figure of merit will be given by I c0 = 4eαkT/ hη. (4) Figure 4. (a) I-STT switching is limited to pulses above 1 ns. The C-STT configuration allows for switching on the timescales <1 ns. (b) Although larger voltages are needed for C-STT switching in this experiment due to the second MgO barrier, the energy per switch at the bit level can still be lowered compared with I-STT due to the use of ultra fast ( hundreds of ps) pulses. (Reprinted with permission from [22]. Copyright 2011, American Institute of Physics.) Another strategy to decrease the switching energy requires the reduction in the switching current of the I-STT devices. Specifically, the switching current to thermal stability figure of merit is given by I c0 = (4eαkT/ hη)(1+h d/2h k ), (2) where it can be observed that the switching current is dominated by the out-of-plane demagnetizing field H d, which typically does not determine the thermal stability, given that H k H d. Hence, a promising method to decrease the switching current without sacrificing the thermal stability is to decrease the demagnetizing field H d by the introduction of perpendicular anisotropy in the free layer [6, 29, 30]. If the perpendicular anisotropy creates an anisotropy field H k, the switching current over thermal stability ratio can become much smaller I c0 = (4eαkT/ hη)(1+(h d H k )/2H k ). (3) A significant interface-induced perpendicular anisotropy has been observed in Fe-rich CoFeB/MgO junctions [31], while keeping large TMR values. Using Co 20 Fe 60 B 20 free layers, our recent work [30] has demonstrated the reduction in the average quasi-static switching current density by >40% (from 2.8 to 1.6 MA cm 2 ) due to the presence of the perpendicular anisotropy. Figure 6 (left) shows the dependence of the switching current (at 10 ns) on perpendicular anisotropy, where the fast decrease in switching current with thickness fits with the interfacial origin of the anisotropy. 2.2. Perpendicular free layers If the perpendicular anisotropy of the free and fixed layers are large enough to overcome their respective demagnetizing fields (i.e. H k > H d ), the magnetizations of the layers become perpendicular, giving rise to a fully perpendicular (P-STT) configuration [31]. Given that the thermal stability in P-STT is determined by the perpendicular anisotropy of the free layer (and not by the elliptical shape as in I-STT), the P-STT structures can be made in circular shape [32, 33]. In the P-STT configuration, the demagnetizing field is fully It can be noted that the elimination of the demagnetizing field leads to a potentially smaller switching current over thermal stability ratio, if damping can be controlled. Several works have recently demonstrated P-STT structures utilizing the interfacial perpendicular anisotropy of Fe-rich CoFeB layers in MgO MTJs [31, 34, 35], with switching currents as low as <2MAcm 2 [34]. Figure 6 (right) shows the dependence of the switching current on the thickness of the free layer, where it can be clearly observed that, as the free layer becomes thinner, the increase in the perpendicular anisotropy in the free layer leads to a larger switching current despite its reduced volume. Finally, it is worth mentioning that there are on-going efforts to reduce the switching current in STT-RAM bits by other approaches such as bias-assisted [36] and thermally assisted switching [37, 38]. In the latter scenario, it has been experimentally demonstrated that the temperature rise of the MTJ while the current pulse is applied can reduce the perpendicular anisotropy in P-STT bits to assist the reversal process, resulting in a reduction in the required switching current without sacrificing the thermal stability of the memory cell in the idle state [38]. 3. Beyond STT-RAM: magnetoelectric memory Despite the promising advantages of STT-RAM over other competing volatile and non-volatile technologies, it faces a number of fundamental and technological challenges to be addressed. In particular, the scaling of the MTJ dimensions will be limited by the super-paramagnetic limit for the inplane configuration, while the development of materials with stronger perpendicular anisotropies is required to maintain the stability in smaller P-STT MTJs [39]. But even for P-STT, the ratio of switching current to thermal stability (hence retention time) is determined by fundamental constants and parameters with a limited tuning range (i.e. damping and spin transfer efficiency in equation (4)). Therefore, if we assume a constant thermal stability scaling rule with small variations in α and η, the switching current will remain constant across the different technology nodes, independent of the MTJ size, and consequently, the STT-RAM memory cell size will be mostly limited by the size of the transistor required to drive the current and not by the MTJ itself. In addition to scalability issues, the switching energy will also not scale proportionally if the resistance of the MTJ is kept constant (by decreasing the RA product as the MTJ size is scaled down). Considering that the STT-RAM switching energy is still close to two orders of magnitude larger than CMOS switching energies ( 100 fj for STT-RAM, compared with 1 fj for the 32 nm CMOS node respectively), the range of applications for hybrid CMOS- MTJ circuits would be limited by the relatively large energy 5

Figure 5. (Left) Dependence of MTJ resistance area (RA) product and TMR ratio on the MgO barrier thickness in a CoFeB MgO MTJ with RA = 3.5 µm 2. For MgO thicknesses below 0.8 nm, a noticeable increase on the spread of the RA product and a decrease in the TMR ratio is observed. (Reprinted with permission from [28]. Copyright 2011, Institute of Electrical and Electronics Engineers.) (Right) The reduction in the RA product is shown to decrease the write energy in an in-plane MTJ. However, as the RA product is reduced, the switching current also increases, limiting this strategy for decreasing write energy. (Reprinted with permission from [27]. Copyright 2011, American Institute of Physics.) Figure 6. (Left) The interfacial anisotropy in Fe-rich CoFeB can be exploited to reduce the switching current in I-STT devices. As the free layer thickness is decreased, the perpendicular anisotropy increases due to the interfacial origin of the anisotropy. When the switching current density at 10 ns is measured as a function of free layer thickness, we observe a noticeable reduction, much larger than what would be expected from the reduction in the free layer volume alone. (Reprinted with permission from [30]. Copyright 2011, American Institute of Physics.) (Right) When the free layer thickness is reduced further, the perpendicular anisotropy surpasses the demagnetizing field and the free layer becomes perpendicular. In the P-STT configuration, we observe the inverse effect, where decreasing the free layer thickness increases H k, resulting in larger current densities. (Reprinted with permission from [34]. Copyright 2012, American Institute of Physics.) dissipation of the memory elements, even when scaling is taken into consideration. In order to reach or surpass the energy efficiency of CMOS, leading to an ultra low-power non-volatile memory with switching energies 1 fj or below at the bit level, we will consider in this section (1) magnetoelectric effects, in which voltage, instead of current, is used to manipulate the magnetization of the memory device [13, 14, 40 42, 45 47], and (2) the possibility of exploiting the giant spin Hall effect to switch the magnetization in MTJs [16]. 3.1. Magnetoelectric MTJ devices A possible approach for developing an ultra-low-power spinbased memory is replacing the spin-polarized currents used to manipulate and switch the magnetization by voltage-driven effects [13, 14, 40, 41]. A particularly interesting approach for the realization of a magnetoelectric memory is based on the observation that the interfacial perpendicular anisotropy in Fe-rich CoFeB/MgO junctions can be modulated by voltage [13, 43, 44]. When a bias voltage is applied, the accumulated 6

Figure 7. Voltage-induced switching of high resistance MTJ with 70 nm 180 nm dimensions for AP to P (top) and P to AP (bottom) [42, 47]. The switching process is assisted by a small magnetic field H Bias, which determines the switching direction. Note that the same voltage polarity was used in both cases, confirming the voltage-induced mechanism. The switching mechanism can be described as follows: due to the voltage-controlled interfacial anisotropy, the coercivity in the hysteresis loops is reduced. Therefore, the initial equilibrium states (A,D) are forced to relax to the only available state (B,E) with the voltage pulse applied. charge near the CoFeB/MgO interface can modify the orbital occupancies of different symmetries, which at the same time control the surface anisotropy due to spin orbit coupling. Therefore, the interface anisotropy can be controlled by the amount of charges accumulated at the interface, i.e. the electric field applied to the junction. The voltage control of the perpendicular anisotropy has been exploited to induce switching in MTJ devices [42, 45 47]. When ultra-fast (<1 ns) voltage pulses are applied, the voltage-induced magnetization dynamics can be used to generate toggle switching of the MTJ by proper timing of the pulse [45, 48, 49], in an analogous fashion to the C-STT switching discussed previously. In a second approach, the voltage control of the interfacial anisotropy has been used to modulate the coercivity of the MTJ devices, allowing for electric-field-assisted switching [46, 47]. As demonstrated in figure 7, the applied voltage induces a reduction in coercivity due to the modification of the interface anisotropy, translating into a single available state where the system is forced to thermally relax for a given applied bias magnetic field H Bias. Once the voltage pulse is released, the system will thus have switched to the opposite direction. Note that the direction of the switching therefore will be determined by the effective field H Bias, while the switching is obtained for the same voltage polarity in both directions (see figure 7). We have recently demonstrated that this approach allows for purely voltage induced (non-stt) switching with voltage pulses of 1 V in amplitude and down to t = 10 ns in duration [42, 47]. In this experiment, MTJs with high resistance (>100 k ) were used, leading to leakage currents below 10 µa. This leads to a small energy dissipation V 2 t/r (due to the leakage current) which is already one order of magnitude lower compared with STT-RAM. If the leakage current is further decreased, the energy per switch will eventually be limited to CV 2, where C is the cell capacitance. This would be a clear path towards reaching switching energies 10 aj while keeping the non-volatility of the memory elements. One of the parameters relevant to the characterization of a magnetoelectric-based memory (e.g. MeRAM) is the amount of effective magnetic field H eff generated per unit of applied voltage V or electric field E. Obviously, a larger magnetoelectric coefficient H eff /V or H eff /E could result in smaller switching voltage and energy for the memory cell. In addition, the voltage required for switching in a MeRAM architecture should be small enough compared with the breakdown voltage of the junction for reliable operation. Currently, devices with an RA product of 3.5 µm 2 have been measured to sustain >10 16 pulses 0.5 V at 5 ns [28]. Although the RA product to be used in MeRAM cells would be orders of magnitude larger (with increased breakdown voltage of the junction), further increase in the magnetoelectric coefficient is desirable to reduce the switching voltage (to below 0.5 V) and thus, improve the endurance of the memory bits, as well as ensuring their compatibility with scaled CMOS logic voltage levels. Finally, as the MeRAM bits are scaled down, H k will have to be increased to keep the thermal stability 7

constant. Therefore, a larger modulation of the effective fields, comparable to H k, would be required to switch the scaled bits. In our MTJ devices with Fe-rich CoFeB free layer, the voltage-controlled interfacial effect can generate effective fields as large as 600 Oe per volt [43]. A magnetoelectric coefficient close to the one observed in CoFeB/MgO systems has also been observed for the FePt/MgO interface experimentally [50]; however, this is still half of the theoretically expected values for this system. Enhancement of the magnetoelectric coefficient has been demonstrated by stacking MgO next to a high-k dielectric [51, 52]. The observed enhancement is shown to be proportional to the increase in the effective dielectric constant of the stack, as expected (i.e. 30% in MgO/HfO 2 Al 2 /O 3 stacks [51]). Further, a 6.5 times enhancement of the magnetoelectric coefficient, compared with Fe-rich CoFeB/MgO systems, has been obtained recently in experiments using FePd next to MgO [53]. Finally, modulation of anisotropy by electric fields has also been observed in GdO x [54], AlO x and TaO x [55], the latter with a magnetoelectric coefficient similar to Fe-rich CoFeB/MgO. The challenge of finding a ferromagnet material (FM)/oxide system with an enhanced voltage-controlled magnetic anisotropy (VCMA) effect and simultaneously large magnetoresistance is an important topic of research in this area. A second possible route to realize magnetoelectric effects for MeRAM is to use single or multi-phase (artificial heterostructure) multiferroic materials [14, 41]. In the case of multiferroic heterostructures, a voltage applied to the material generates a mechanical strain in a piezoelectric material. If the ferromagnetic phase has a magnetostrictive property, the strain will result in an effective magnetic field, which can then be exploited to induce switching. It has been reported that multiferroic heterostructures offer larger magnetoelectric coefficients as compared with single phase materials. Figure 8 shows an experimental demonstration of this principle, where an electric field as small as 0.14 MV m 1 (or 1.4 kv cm 1 applied across a PMT-PT substrate) can generate a 90 reorientation of a Ni film easy axis (see figure 8). This experiment demonstrates voltagecontrolled reorientation of magnetization, as well as nonvolatility requiring no continuously applied voltage to keep the magnetization in the reoriented state, thus demonstrating an electric field controlled non-volatile magnetic memory operation. The read-out operation of this kind of memories could still be performed via the GMR or TMR effects, as proposed in [14, 56]. However, implementation of MeRAM arrays based on this device will require addressing challenges in terms of integration of multiferroic materials and oxides with CMOS processing. Moreover, the energy dissipation in this scenario also includes the piezoelectric hysteresis (poling) energy loss, as well as the large capacitance due to the increased dielectric constant of the piezoelectric. 3.2. Giant spin Hall-based MTJ devices Recently, it was demonstrated that a current flowing through a metallic layer with large spin orbit coupling, such as Ta 8 Figure 8. Experimental evidence of voltage control of magnetization for a 30 nm magnetostrictive Ni film deposited on a piezoelectric PMN-PT substrate. By measuring the state of the magnetic film using the magneto-optical Kerr effect (MOKE), it is shown that a 90 reorientation of the easy axis can be observed with fields 0.14 MV m 1 (or 1.4 kv cm 1 ). This experiment demonstrates voltage-controlled reorientation of magnetization, as well as non-volatility requiring no continuously applied voltage to keep the magnetization in the reoriented state, thus demonstrating an electric field controlled non-volatile magnetic memory operation. (Reprinted with permission from [14]. Copyright 2011, American Institute of Physics.) adjacent to a CoFeB free layer, can generate a spin current through the MTJ (via the giant spin Hall effect), which is large enough to induce switching in the device [16]. The device structure is shown in figure 9, where the electrical current J C flowing between terminals A and C induces a spin current J S = θ SH J C through the MTJ device, where θ SH is referred to as the spin Hall angle. Note that the direction of the current J C will also control the spin polarization of J S and therefore, the switching direction will be determined by the sign of J C. Using the β-phase of Ta, the spin Hall angle θ SH can be as large as 12% without further increase in the damping of the adjacent CoFeB free layer [16]. However, it has also been demonstrated that the spin Hall angle can be as large as 30% in tungsten [57], comparable to the spin-current-efficiency for regular two terminal MTJ devices (η 60%). Although this three-terminal device potentially results in a penalty in terms of density as compared with a twoterminal MTJ, it also offers advantages, which address specific challenges of STT-RAM. First, the switching current density required to switch the MTJ can be tuned by engineering the dimensions of the metal (e.g. Ta) wire, without disturbing the

Figure 9. Three-terminal device configuration exploiting the giant spin Hall effect in β-ta/cofeb [16]. By running a current J C through the tantalum wire (between terminals A and C), a pure spin current J S through the MTJ is induced, which can be used for switching the free layer. The spin polarization of the spin current J S will depend on the direction of J C. One of the advantages of this three-terminal configuration is that the required current density can be tuned by engineering the lateral area of the wire A Ta, while A MTJ can be designed to account for a given thermal stability factor. thermal stability of the MTJ. In particular, the ratio of the switching current due to spin Hall (I SH ) to the required current for STT (I STT ) can be approximated by I SH I STT η θ SH A Ta A MTJ, (5) where A Ta and A MTJ are the cross-sectional area of the underlying metal layer, and the MTJ device, respectively (see figure 9). Thus, considering that the area of the metal layer (interconnect) A Ta could be engineered to be as much as one order of magnitude smaller as compared with A MTJ by reducing its thickness to <5 nm, the switching current can be lowered by nearly one order of magnitude compared with STT-based switching. Second, the read (terminal B to A or C in figure 9) and write paths (A to C) would be separated, allowing disturbfree read operations. Finally, in terms of switching energy, the smaller switching current would flow through a lower resistance metallic wire (e.g. 0.3 µm 2 for a L = 200 nm Ta wire as compared with an MTJ), leading to switching energy levels on the order of CMOS devices ( 1 fj). However, the need for ultrathin metal interconnects can limit the overall size of memory arrays in this approach, given the high resistance associated with them. 4. Conclusions The intrinsic properties of spin-based systems make them advantageous as a prospective candidate for fast, dense and energy efficient non-volatile memory devices. STT-RAM, where a spin-polarized current is used to manipulate the magnetization state of an MTJ, has become one of the emerging candidates particularly for embedded memory, since it can be directly integrated with CMOS, matching or outperforming SRAM, DRAM or Flash in most respects (except for density). The compatibility of CMOS and STT-RAM could spur 9 new paradigms for non-volatile computation. Even so, the future scalability of STT-RAM is limited by its currentbased mechanism and also relies on the development of novel materials with higher perpendicular anisotropies and lower damping. However, the recent demonstration of the giant spin Hall effect in Ta/CoFeB interfaces provides an attractive alternative for reduction in the switching current, opening a possible path for lowering the energies required for currentbased switching. Alternative voltage-based writing mechanisms for MTJs may also provide a viable route towards switching energies close to or smaller compared with CMOS ( 1 fj). Replacing the STT-based manipulation of magnetization with voltagebased mechanisms has been predicted to take the switching energies to the sub-femto-joule regime. Two different principles of electric field control of magnetism were discussed, namely, voltage control of interface anisotropy (VCMA) and multiferroic heterostructures, along with memory devices based on them, demonstrating magnetic switching and reorientation of the easy axis by electric fields. So far, voltage-induced switching assisted by a magnetic field in the thermally activated regime and electric field based toggle switching have been demonstrated. Voltage-induced, magnetic field independent switching with improved energy efficiency and similar reliability and read-out characteristics as compared with STT-RAM, is still to be demonstrated. However, it is expected that these novel mechanisms could lead to a MeRAM, with dramatically higher energy efficiency and scalability than those of purely current-controlled memories such as STT-RAM. Acknowledgments This work was supported by the DARPA STT-RAM and NV Logic programs under program manager Dr D Shenoy, and by the Nanoelectronics Research Initiative through the Western Institute of Nanoelectronics. References [1] Mayberry M 2010 Emerging Technologies and Moore s Law: Prospects for the Future. Available: csg5.nist.gov/pml/ div683/upload/mayberry March 2010.pdf [2] Moore G E 1965 Cramming more components onto integrated circuits Electron. Mag. 38 114 7 [3] Lin C J et al 2009 45nm low power CMOS logic compatible embedded STT MRAM utilizing a reverse-connection 1T/1MTJ cell Electron Devices Meeting (IEDM), 2009 IEEE Int. (Baltimore, MA) pp 1 4 [4] Assefa S et al 2007 Fabrication and characterization of MgO-based magnetic tunnel junctions for spin momentum transfer switching J. Appl. Phys. 102 063901 [5] Huai Y M et al 2004 Observation of spin-transfer switching in deep submicron-sized and low-resistance magnetic tunnel junctions Appl. Phys. Lett. 84 3118 20 [6] Chen E et al 2010 Advances and future prospects of spin-transfer torque random access memory IEEE Trans. Magn. 46 1873 8 [7] Katine J A et al 2000 Current-driven magnetization reversal and spin-wave excitations in Co/Cu/Co pillars Phys. Rev. Lett. 84 3149 52 [8] Parkin SSPet al 2004 Giant tunnelling magnetoresistance at room temperature with MgO (100) tunnel barriers Nature Mater. 3 862 7

[9] Slonczewski J C 1996 Current-driven excitation of magnetic multilayers J. Magn. Magn. Mater. 159 L1 7 [10] Berger L 1996 Emission of spin waves by a magnetic multilayer traversed by a current Phys. Rev. B 54 9353 8 [11] Wang K L and Khalili Amiri P 2012 Nonvolatile spintronics: perspectives on instant-on nonvolatile nanoelectronic systems SPIN 2 1250009 [12] Paul S et al 2011 A circuit and architecture codesign approach for a hybrid CMO&-STTRAM nonvolatile FPGA IEEE Trans. Nanotechnol. 10 385 94 [13] Maruyama T et al 2009 Large voltage-induced magnetic anisotropy change in a few atomic layers of iron Nature Nanotechnol. 4 158 61 [14] Wu T et al 2011 Electrical control of reversible and permanent magnetization reorientation for magnetoelectric memory devices Appl. Phys. Lett. 98 262504 [15] Khalili Amiri P and Wang K L 2012 Voltage-controlled magnetic anisotropy in spintronic devices SPIN 2 1240002 [16] Liu L et al 2012 Spin-torque switching with the giant spin hall effect of tantalum Science 336 555 8 [17] International technology Roadmap for Semiconductors 2005 [18] Wang K L et al 2011 From nanoelectronics to nano-spintronics J. Nanosci. Nanotechnol. 11 306 13 [19] Lee K and Kang S H 2011 Development of embedded STT-MRAM for mobile system-on-chips IEEE Trans. Magn. 47 131 6 [20] Sun J Z 2000 Spin-current interaction with a monodomain magnetic body: a model study Phys. Rev. B 62 570 8 [21] Kent A D et al 2004 Spin-transfer-induced precessional magnetization reversal Appl. Phys. Lett. 84 3897 9 [22] Rowlands G E et al 2011 Deep subnanosecond spin torque switching in magnetic tunnel junctions with combined in-plane and perpendicular polarizers Appl. Phys. Lett. 98 102509 [23] Liu H et al 2010 Ultrafast switching in magnetic tunnel junction based orthogonal spin transfer devices Appl. Phys. Lett. 97 242510 [24] Bedau D et al 2010 Ultrafast spin-transfer switching in spin valve nanopillars with perpendicular anisotropy Appl. Phys. Lett. 96 022514 [25] Bedau D et al 2010 Spin-transfer pulse switching: from the dynamic to the thermally activated regime Appl. Phys. Lett. 97 262502 [26] Lee O J et al 2011 Spin-torque-driven ballistic precessional switching with 50ps impulses Appl. Phys. Lett. 99 102507 [27] Zeng Z M et al 2011 Effect of resistance area product on spin-transfer switching in MgO-based magnetic tunnel junction memory cells Appl. Phys. Lett. 98 072512 [28] Amiri P K et al 2011 Low write-energy magnetic tunnel junctions for high-speed spin-transfer-torque MRAM IEEE Electron Device Lett. 32 57 9 [29] Liu L Q et al 2009 Reduction of the spin-torque critical current by partially canceling the free layer demagnetization field Appl. Phys. Lett. 94 122508 [30] Amiri P K et al 2011 Switching current reduction using perpendicular anisotropy in CoFeB MgO magnetic tunnel junctions Appl. Phys. Lett. 98 112507 [31] Ikeda S et al 2010 A perpendicular-anisotropy CoFeB MgO magnetic tunnel junction Nature Mater. 9 721 4 [32] Huai Y 2008 Spin-transfer Torque MRAM (STT-MRAM): challenges and prospects AAPPS Bull. 18 33 40 [33] Apalkov D et al 2010 Comparison of scaling of in-plane and perpendicular spin transfer switching technologies by micromagnetic simulation IEEE Trans. Magn. 46 2240 3 [34] Rahman M T et al 2012 Reduction of switching current density in perpendicular magnetic tunnel junctions by tuning the anisotropy of the CoFeB free layer J. Appl. Phys. 111 07C907 [35] Worledge D C et al 2011 Spin torque switching of perpendicular Ta CoFeB MgO-based magnetic tunnel junctions Appl. Phys. Lett. 98 022501 [36] Deac A et al 2006 Current-induced magnetization switching in exchange-biased spin valves for current-perpendicular-to-plane giant magnetoresistance heads Phys. Rev. B 73 064414 [37] Hatami M et al 2007 Thermal spin-transfer torque in magnetoelectronic devices Phys. Rev. Lett. 99 066603 [38] Bandiera S et al 2011 Spin transfer torque switching assisted by thermally induced anisotropy reorientation in perpendicular magnetic tunnel junctions Appl. Phys. Lett. 99 202507 [39] Driskill-Smith A et al 2011 Latest advances and roadmap for in-plane and perpendicular STT-RAM Memory Workshop (IMW), 2011 3rd IEEE Int. (Baltimore, MD) pp 1 3 [40] Endo M et al 2010 Electric-field effects on thickness dependent magnetic anisotropy of sputtered MgO/Co(40)Fe(40)B(20)/Ta structures Appl. Phys. Lett. 96 212503 [41] Ramesh R and Spaldin N A 2007 Multiferroics: progress and prospects in thin films Nature Mater. 6 21 9 [42] Alzate J G et al 2013 Voltage-induced switching of nanoscale magnetic tunnel junctions IEDM Tech Digest at press [43] Zhu J et al 2012 Voltage-induced ferromagnetic resonance in magnetic tunnel junctions Phys. Rev. Lett. 108 197203 [44] Duan C G et al 2008 Tailoring magnetic anisotropy at the ferromagnetic/ferroelectric interface Appl. Phys. Lett. 92 122905 [45] Shiota Y et al 2012 Induction of coherent magnetization switching in a few atomic layers of FeCo using voltage pulses Nature Mater. 11 39 43 [46] Wang W-G et al 2012 Electric-field-assisted switching in magnetic tunnel junctions Nature Mater. 11 64 8 [47] Alzate J G et al 2011 Voltage-induced switching of CoFeB MgO magnetic tunnel junctions 56th Conf. on Magnetism and Magnetic Materials (Scottsdale, AZ) EG-11 [48] Kanai S et al 2012 Electric field-induced magnetization reversal in a perpendicular-anisotropy CoFeB-MgO magnetic tunnel junction Appl. Phys. Lett. 101 122403 [49] Shiota Y et al 2012 Pulse voltage-induced dynamic magnetization switching in magnetic tunneling junctions with high resistance area product Appl. Phys. Lett. 101 102406 [50] Seki T et al 2011 Coercivity change in an FePt thin layer in a Hall device by voltage application Appl. Phys. Lett. 98 212505 [51] Kita K et al 2012 Electric-field-control of magnetic anisotropy of Co 0.6 Fe 0.2 B 0.2 /oxide stacks using reduced voltage J. Appl. Phys. 112 033919 [52] Endo M et al 2010 Electric-field effects on thickness dependent magnetic anisotropy of sputtered MgO/Co 40 Fe 40 B 20 /Ta structures Appl. Phys. Lett. 96 212503 [53] Bonell F et al 2011 Large change in perpendicular magnetic anisotropy induced by an electric field in FePd ultrathin films Appl. Phys. Lett. 98 232510 [54] Bauer U et al 2012 Electric field control of domain wall propagation in Pt/Co/GdOx films Appl. Phys. Lett. 100 192408 [55] Schellekens A J et al 2012 Electric-field control of domain wall motion in perpendicularly magnetized materials Nature Commun. 3 847 [56] Hu J-M et al 2011 High-density magnetoresistive random access memory operating at ultralow voltage at room temperature Nature Commun. 2 553 [57] Pai C-F et al 2012 Spin transfer torque devices utilizing the giant spin Hall effect of tungsten Appl. Phys. Lett. 101 122404 10