Next Article in Journal
Challenges and Modification Strategies of Ni-Rich Cathode Materials Operating at High-Voltage
Next Article in Special Issue
Self-Templating Synthesis of N/P/Fe Co-Doped 3D Porous Carbon for Oxygen Reduction Reaction Electrocatalysts in Alkaline Media
Previous Article in Journal
Review of Photothermal Technique for Thermal Measurement of Micro-/Nanomaterials
Previous Article in Special Issue
Structure and Magnetic Properties of ErFexMn12−x (7.0 ≤ x ≤ 9.0, Δx = 0.2)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comprehensive Study of the Current-Induced Spin–Orbit Torque Perpendicular Effective Field in Asymmetric Multilayers

1
Songshan Lake Materials Laboratory, Dongguan 523808, China
2
Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
3
Key Laboratory for Magnetism and Magnetic Materials of Ministry of Education, School of Physical Science and Technology, Lanzhou University, Lanzhou 730000, China
4
Faculty of Physics and Electronic Science, Hubei University, Wuhan 430062, China
5
School of Physics & Optoelectric Engineering, Guangdong University of Technology, Guangzhou 510006, China
6
College of Physics, Sichuan University, Chengdu 610064, China
7
Department of Electrical Engineering, University of California, Los Angeles, CA 90095, USA
8
Nanofabrication Facility, Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences, Suzhou 215123, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2022, 12(11), 1887; https://0-doi-org.brum.beds.ac.uk/10.3390/nano12111887
Submission received: 5 May 2022 / Revised: 29 May 2022 / Accepted: 30 May 2022 / Published: 31 May 2022

Abstract

:
The spin–orbit torques (SOTs) in the heavy metal (HM)/ferromagnetic metal (FM) structure hold promise for next-generation low-power and high-density spintronic memory and logic applications. For the SOT switching of a perpendicular magnetization, an external magnetic field is inevitable for breaking the mirror symmetry, which is not practical for high-density nanoelectronics applications. In this work, we study the current-induced field-free SOT switching and SOT perpendicular effective field ( H z e f f ) in a variety of laterally asymmetric multilayers, where the asymmetry is introduced by growing the FM layer in a wedge shape. We show that the design of structural asymmetry by wedging the FM layer is a universal scheme for realizing field-free SOT switching. Moreover, by comparing the FM layer thickness dependence of ( H z e f f ) in different samples, we show that the efficiency (β = H z e f f /J, J is the current density) is sensitive to the HM/FM interface and the FM layer thickness. The sign of β for thin FM thicknesses is related to the spin Hall angle (θSH) of the HM layer attached to the FM layer. β changes its sign with the thickness of the FM layer increasing, which may be caused by the thickness dependence of the work function of FM. These results show the possibility of engineering the deterministic field-free switching by combining the symmetry breaking and the materials design of the HM/FM interface.

Current-induced spin–orbit torque (SOT) provides an energy-efficient and fast way to electrically manipulate the magnetization [1,2,3,4] and dynamics of spin textures (such as chiral domain wall (DW) [5,6,7,8,9] and magnetic skyrmions [10,11,12], etc.) in the heavy metal (HM)/ferromagnetic metal (FM) multilayers. In such a structure, an in-plane current (I) flowing through the HM layer is converted to a pure spin current (Js) due to the spin Hall effect [1,13,14] and/or interfacial Rashba effect [15]. The Js injects into the adjacent FM layer and thus exerts the SOTs. To enable the SOT-driven perpendicular magnetization switching, an external magnetic field is inevitable to break the mirror symmetry [2], which is impractical for high-density nanoelectronics applications. Until 2014, the field-free SOT switching of a perpendicular magnetization was achieved by introducing a laterally asymmetric structure [16], providing a new pathway to realize all-electric deterministic switching. After that, many other strategies have been proposed for realizing field-free SOT switching [17,18,19,20,21,22,23,24,25,26,27,28,29,30,31]. For the case of laterally asymmetric structure, the field-free SOT switching is driven by the current-induced out-of-plane effective magnetic field ( H z e f f ). The magnitude and sign of H z e f f determine the switching efficiency and switching polarity at zero external field, respectively. However, the key factors that affect the magnitude and sign of H z e f f are still elusive.
In this work, we aim to explore the key factors that affect the current-induced H z e f f and the resulting field-free SOT switching in a variety of laterally asymmetric structures. We find that the H z e f f is generally introduced in various laterally asymmetric structures. By comparing the FM thickness dependence of the efficiency (β) of H z e f f (i.e., β = H z e f f /J, where J is the current density), we show that β is closely related to the HM/FM interface and the FM layer thickness. Our results advance the understanding of the current-induced out-of-plane effective magnetic field in the laterally asymmetric structures.
The film stacks consisting of (i) Ta(5)/Gd(1)/CoFeB(w)/MgO(2), (ii) Pt(5)/CoFeB(w)/MgO(2), (iii) IrMn(5)/CoFeB(w)/MgO(2), (iv) Ta(5)/CoFeB(w)/MgO(2), (v) Ta(5)/Mo(1)/CoFeB(w)/MgO(2), and (vi) W(5)/CoFeB(w)/MgO(2) (thickness in nm) were prepared by magnetron sputtering at room temperature on Si substrates capped with a 100 nm thermal oxide under a base pressure of <1 × 10−8 Torr. The CoFeB layer was grown by the oblique sputtering method and hence has a wedge-sharp structure (w). The CoFeB layer thickness (denoted as tCoFeB) varies from 0.50 nm to 1.20 nm within the lateral length of ~5 cm. It is worth noting that we calibrate the wedged thickness in a large lateral scale, therefore, the several nm-scale thickness difference can be detected precisely. The other layers were uniformly grown by rotating the substrate during the deposition. The stacks were annealed at 250 °C for 30 min to enhance the perpendicular magnetic anisotropy (PMA). The basic magnetic properties of the different samples are similar, therefore, only the results of the Ta(5)/Gd(1)/CoFeB(w)/MgO(2) multilayer are presented. The schematic illustration of the Ta/Gd/CoFeB/MgO structure is shown in Figure 1a. The films were patterned into Hall bar devices with the dimension of 130 × 20 μm2 (see Figure 1b) via standard photolithography and dry etching techniques for anomalous Hall effect (AHE) and magneto-optical Kerr effect (MOKE) microscopy measurements. For the Hall bar device, there could be some thickness variation, however, the wedged trend should be kept.
Figure 1c shows the AHE loops of the devices with a series of tCoFeB, in which the RH and Hz are the Hall resistance and out-of-plane external magnetic field, respectively. The sharp-square loops indicate the existence of a PMA for the devices. The dynamics of the domain wall driven by Hz for the whole Hall bar device with tCoFeB = 0.70 nm is shown in Figure 1d. In image ①, the red dotted line shows the current channel of the Hall bar device. At first, a large Hz along +z direction was applied to saturate the sample, and the picture was chosen as the reference as shown in image ①. As Hz increases in the −z direction and reaches the switching field, a reversed domain is nucleated at the bottom edge of the device (see image ②). As the field increases, the domain expands to the whole Hall bar device, as shown in images ③–⑤. These results show that the switching is accomplished by domain nucleation and the domain wall motion. We also measured the perpendicular anisotropy energy density Ku (Ku = μ0HkMs/2). Ms is the saturation magnetization, which is measured by the superconducting quantum interference device (SQUID) and has a magnitude of ~710 emu/cm3, μ0 is the vacuum permeability. Hk is the effective anisotropy field, which is measured by the in-plane AHE loops, as shown in Figure 1e. It is known that RH is only proportional to the z-axis component of magnetization (M) in a system with PMA. As the in-plane magnetic field (Hx) increases, M will be rotated from the z direction (easy axis) to the x direction (hard axis). Consequently, there is a reduction of the RH at high fields, as shown in Figure 1e. Figure 1f summarizes the tCoFeB dependence of Ku. Ku increases first when tCoFeB < 0.77 nm, which has been attributed to the change of the CoFeB/MgO interface (i.e., the interfacial anisotropy) caused by B diffusion [32]. With further increasing the CoFeB thickness, Ku starts to decrease since the PMA has an interfacial origin [33].
Next, we show the field-free SOT switching in the devices with different tCoFeB performed by the Keithley 2612A source/measure unit. In the measurements, the writing pulses with a width of 1 ms were injected along the device channel. To avoid applying a zero-writing current, the step number was set as 101 for scanning the writing pulse between −35 (−30) mA and 35 (30) mA. After each writing pulse, the Hall resistance was measured by a reading current of 3 mA. Figure 2a shows the field-free SOT switching for devices of 0.66 nm < tCoFeB < 0.91 nm. The current density (Je) is calculated by assuming a uniform current distribution across the film stack. As a reference, we have measured the magnetization switching curve in the absence of external magnetic fields loop in the structure without the wedge structure, as shown in Figure 2a. One can see that there is no deterministic SOT-induced magnetization switching. The previous work has demonstrated that the field-free SOT switching was driven by the current-induced H z e f f that originates from the lateral asymmetry [16]. In detail, the wedged layer is deposited at an oblique angle with respect to the substrate surface (along the y axis) without rotating the substrate, so it is grown in a tilted direction away from the substrate normal. Consequently, it breaks the mirror symmetry with respect to the x–z plane and allows for the creation of a built-in effective electric field (E) along the y axis. Consequently, a current induced H z e f f is expected due to the Rashba spin–orbit coupling (SOC), which is expressed by H z e f f = α ⸱ (p × E) [15,34]. Here, α is the Rashba SOC constant depends on the materials, and p represents the electron’s momentum. If H z e f f is larger than the coercivity, the magnetization switching can be achieved, although the thickness gradient is very tiny. Figure 2b shows the SOT switching loops under different Hx for the device with tCoFeB = 0.81 nm. We found that the switching polarity changes from a clockwise mode to an anticlockwise mode when the field is increased to Hx = 50 Oe. In this case, the switching is not dominated by the H z e f f anymore, and the conventional damping-like SOT dominates the switching with the assistance of the in-plane magnetic field. It is worth noting that the switching polarity under a large positive Hx is consistent with the case in the Ta/CoFeB/MgO system [2]. Nevertheless, the Gd has a positive spin Hall angle (θSH) [35], which is opposite to Ta. In this regard, we conclude that the conventional damping-like SOT originates from the spin current that is generated in the Ta layer and diffuses through the Gd layer even it is partially compensated by the spin current in the Gd layer.
During the current-driven magnetization switching, the MOKE measurements were performed simultaneously. Figure 2c shows the MOKE images of SOT switching by using the same Hall bar device as shown in Figure 1d. During the measurements, a large Hz along +z direction was first applied to saturate the magnetization, and the picture was chosen as the reference, as shown in image ①. Then, the current pulses were applied to drive the magnetization switching. Interestingly, the nucleation position is different from that in Figure 1d. For the SOT-driven switching, the initial reversal of domain occurs on the right side of the device channel, as shown in image ②, which is likely due to the presence of Dzyaloshinskii–Moriya interaction (DMI) [36,37] that tilts the magnetization at the device boundary [38], lowering the SOT switching barrier for the right edge and thus to induce the domain nucleation. As the current increases, the DW is subsequently driven to the left side of the device channel. For the observed DW, the neighboring magnetizations on its two sides point ↑↓. When a current is applied along +x axis, the spin orientation (σ) of the net spin current is along +y direction. The dimpling-like field (HDL) as a driven force of the DW can be expressed as HDL = m × σ, pointing respectively along +z or −z direction when the magnetic moment (m) in the DW along +x (→) or −x (←) directions. Here, “→” and “←” refer to the in-plane component of m in the center of DW. Obviously, the current along +x drives the DW to move along −x direction, namely, an effective HDLalong −z direction is generated. Thus, the magnetic configuration in the domain wall is ↑←↓, i.e., a left-handed chirality. The magnitude of DMI can be obtained by measuring the AHE loops of switching a Hall cross with combined Hx and Hz under a series of DC current densities [39,40]. Figure 2d summarizes the SOT efficiency (χ) as a function of Hx. The saturated field with the maximum χ corresponds to the effective DMI field (|HDMI|). Then, the DMI exchange constant (|D|) can be obtained by using |D| = μ0MsΔ|HDMI| [41], where Δ is the DW width and is related to exchange stiffness constant A ≈ 1.5 × 10−11 J/m and Ku, with the form of Δ = (A/Ku)1/2 [6,39]. The value of the |D| scales linearly with the inverse of tCoFeB, as shown in Figure 2e, indicating its interfacial origin.
In the following, we extract the current-induced H z e f f . Figure 3a shows the hysteresis AHE loops under currents with opposite polarities. The AHE loop shifts to the left under a negative current, indicating the existence of a perpendicular effective field along +z direction ( H z + ). Similarly, a positive current generates an effective perpendicular field along the −z direction ( H z ). The averaged perpendicular effective field H z e f f can be obtained by H z e f f = ( H z H z + )/2 = βJ, where β and J refer to the efficiency of H z e f f and the current density, respectively. Figure 3b shows the tCoFeB dependence of β, where β decreases firstly and changes its sign at tCoFeB ≈ 0.66 nm, after that β increases negatively. The sign change of β is consistent with our previous work [16]. It is worth noting that the β at 0.63 nm < tCoFeB < 0.66 nm shows small magnitudes, which are likely responsible for the partial switching and same polarity of switching, as shown in Figure 2a.
To explore the key factors that affect the β, samples ii-vi were measured. Figure 4 summarizes the CoFeB thickness dependence of β for these samples. We found that H z e f f exists in all laterally asymmetric structures, indicating this is a universal phenomenon. The Pt/CoFeB, IrMn/CoFeB, and Gd/CoFeB samples have a similar thickness dependence, while Ta/CoFeB, Mo/CoFeB, and W/CoFeB samples show an opposite dependence. To better compare these samples, the β, θSH, and Ds = D·tFM are extracted for all the samples, as shown in Table 1. We note that some parameters are obtained from the literatures. We found that the interfacial HM layer attached to the FM has a significant relation to the β. First, the sign of β for the thin CoFeB region is determined by the sign of θSH of the interfacial HM layer attached to the FM layer. The β is positive (negative) for the interfacial HM that has a positive (negative) θSH. For example, the β values are positive when the interfacial HM are Pt, IrMn, and Gd, which have a positive θSH. Similarly, β values are negative when the interfacial HM are Ta, Mo, and W, which have a negative θSH. Consequently, β is negative for Ta/CoFeB, Mo/CoFeB, and W/CoFeB samples. As the thickness of CoFeB increases, β changes its sign at thick CoFeB side except for the Pt/CoFeB and IrMn/CoFeB samples. Our previous work pointed out that the sign of β likely depends on the work functions of interfacial HM and FM [34]. Thus, the sign-change of β in our systems may be ascribed to the thickness dependence of work function of CoFeB layer. The sign-change was not observed for Pt and IrMn samples, which may be attributed to the fact that the PMA regions in these two systems are narrow and the sign reversal thickness is not reached. For all the samples, the magnitude of β strongly depends on the CoFeB thickness, which may be caused by the thickness dependences of the work function of FM layer, Ku, interfacial DMI, and the oblique deposition induced crystal structure asymmetry of CoFeB layer. Further work is required to elucidate the microscopic origin of the thickness dependence.
In conclusion, we have demonstrated that the current-induced SOT perpendicular effective field is universal for a variety of laterally asymmetric multilayers with a wedged FM layer. The efficiency β is sensitive to the HM/FM interface and the FM layer thickness. The sign of β in a laterally asymmetric structure at thin FM thickness position is determined by the sign of the θSH of interfacial HM layer attached to the FM layer. As the thickness of FM increases, the sign reversal of β is observed, which may be related to the thickness dependence of the work function of FM. Our work advances the understanding of the out-of-plane effective field in the laterally asymmetric device and provides a pathway in engineering the perpendicular effective field. However, additional advantages may be added to the field-free SOT devices. For example, Pt usually introduces a large DMI [3], IrMn provides antiferromagnetic coupling [20,44], Mo improves the sample’s thermal stability [49], and W has a larger θSH [48].

Author Contributions

Investigation, formal analysis, data collection, B.C., Z.Z. (Zengtai Zhu), C.W., X.G., Z.N., T.G., P.C. and D.Z.; data analysis, data interpretation, B.C., Z.Z. (Zengtai Zhu), C.W., X.G., Z.N., T.G., P.C., D.Z. and T.Y.; conceptualization, methodology, supervision, H.W., L.X., Z.Z. (Zhongming Zeng), S.L., G.Z., G.Y. and K.L.W.; writing—original draft preparation, B.C., Z.Z. (Zengtai Zhu), C.W., H.W. and G.Y.; writing—review and editing, B.C., Z.Z. (Zengtai Zhu), C.W., H.W. and G.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (Grants No. 2021YFB3601300), the Guangdong Basic and Applied Basic Research Foundation (Grant No. 2020A1515110553), the Beijing Natural Science Foundation (Grant No. Z190009), the National Natural Science Foundation of China (NSFC, Grant Nos. 11874409, 11904056, 91963201, 5167109), the Science Center of the National Science Foundation of China (No. 52088101), the K. C. Wong Education Foundation (Grant No. GJTD-2019-14), the China Postdoctoral Science Foundation (Grant No. 2020M670499), the 111 Project (Grant No. B20063), and Guangzhou basic and applied basic research project (No. 202102020053). T.Y. acknowledges the International Visiting Program for Excellent Young Scholars of SCU. H.W. acknowledges the start-up funding from Songshan Lake Materials Laboratory (Y1D1071S511).

Data Availability Statement

All the data present in this paper will be made available upon reasonable request. Please contact the corresponding author for further information.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miron, I.M.; Garello, K.; Gaudin, G.; Zermatten, P.-J.; Costache, M.V.; Auffret, S.; Bandiera, S.; Rodmacq, B.; Schuhl, A.; Gambardella, P. Perpendicular switching of a single ferromagnetic layer induced by in-plane current injection. Nature 2011, 476, 189–193. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, L.; Pai, C.-F.; Li, Y.; Tseng, H.; Ralph, D.; Buhrman, R. Spin-torque switching with the giant spin Hall effect of tantalum. Science 2012, 336, 555–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Liu, L.; Lee, O.; Gudmundsen, T.; Ralph, D.; Buhrman, R. Current-induced switching of perpendicularly magnetized magnetic layers using spin torque from the spin Hall effect. Phys. Rev. Lett. 2012, 109, 096602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kim, J.; Sinha, J.; Hayashi, M.; Yamanouchi, M.; Fukami, S.; Suzuki, T.; Mitani, S.; Ohno, H. Layer thickness dependence of the current-induced effective field vector in Ta/CoFeB/MgO. Nat. Mater. 2013, 12, 240–245. [Google Scholar] [CrossRef] [PubMed]
  5. Yamanouchi, M.; Chiba, D.; Matsukura, F.; Ohno, H. Current-induced domain-wall switching in a ferromagnetic semiconductor structure. Nature 2004, 428, 539. [Google Scholar] [CrossRef] [PubMed]
  6. Metaxas, P.; Jamet, J.; Mougin, A.; Cormier, M.; Ferré, J.; Baltz, V.; Rodmacq, B.; Dieny, B.; Stamps, R. Creep and flow regimes of magnetic domain-wall motion in ultrathin Pt/Co/Pt films with perpendicular anisotropy. Phys. Rev. Lett. 2007, 99, 217208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Parkin, S.S.; Hayashi, M.; Thomas, L. Magnetic domain-wall racetrack memory. Science 2008, 320, 190–194. [Google Scholar] [CrossRef]
  8. Miron, I.M.; Moore, T.; Szambolics, H.; Buda-Prejbeanu, L.D.; Auffret, S.; Rodmacq, B.; Pizzini, S.; Vogel, J.; Bonfim, M.; Schuhl, A. Fast current-induced domain-wall motion controlled by the Rashba effect. Nat. Mater. 2011, 10, 419. [Google Scholar] [CrossRef]
  9. Ryu, K.-S.; Thomas, L.; Yang, S.-H.; Parkin, S. Chiral spin torque at magnetic domain walls. Nat. Nanotechnol. 2013, 8, 527. [Google Scholar] [CrossRef]
  10. Nagaosa, N.; Tokura, Y. Topological properties and dynamics of magnetic skyrmions. Nat. Nanotechnol. 2013, 8, 899–911. [Google Scholar] [CrossRef]
  11. Jiang, W.; Upadhyaya, P.; Zhang, W.; Yu, G.; Jungfleisch, M.B.; Fradin, F.Y.; Pearson, J.E.; Tserkovnyak, Y.; Wang, K.L.; Heinonen, O.; et al. Blowing magnetic skyrmion bubbles. Science 2015, 349, 283. [Google Scholar] [CrossRef] [Green Version]
  12. Jiang, W.; Zhang, X.; Yu, G.; Zhang, W.; Wang, X.; Jungfleisch, M.B.; Pearson, J.E.; Cheng, X.; Heinonen, O.; Wang, K.L.; et al. Direct observation of the skyrmion Hall effect. Nat. Phys. 2016, 13, 162–169. [Google Scholar] [CrossRef]
  13. Dyakonov, M.; Perel, V. Current-induced spin orientation of electrons in semiconductors. Phys. Lett. A 1971, 35, 459–460. [Google Scholar] [CrossRef]
  14. Hirsch, J.E. Spin hall effect. Phys. Rev. Lett. 1999, 83, 1834. [Google Scholar] [CrossRef] [Green Version]
  15. Bychkov, Y.A.; Rashba, E.I. Properties of a 2D electron gas with lifted spectral degeneracy. JETP Lett. 1984, 39, 78–81. [Google Scholar]
  16. Yu, G.; Upadhyaya, P.; Fan, Y.; Alzate, J.G.; Jiang, W.; Wong, K.L.; Takei, S.; Bender, S.A.; Chang, L.T.; Jiang, Y.; et al. Switching of perpendicular magnetization by spin–orbit torques in the absence of external magnetic fields. Nat. Nanotechnol. 2014, 9, 548–554. [Google Scholar] [CrossRef]
  17. You, L.; Lee, O.; Bhowmik, D.; Labanowski, D.; Hong, J.; Bokor, J.; Salahuddin, S. Switching of perpendicularly polarized nanomagnets with spin orbit torque without an external magnetic field by engineering a tilted anisotropy. Proc. Natl. Acad. Sci. USA 2015, 112, 10310–10315. [Google Scholar] [CrossRef] [Green Version]
  18. Oh, Y.W.; Chris Baek, S.H.; Kim, Y.M.; Lee, H.Y.; Lee, K.D.; Yang, C.G.; Park, E.S.; Lee, K.S.; Kim, K.W.; Go, G.; et al. Field-free switching of perpendicular magnetization through spin–orbit torque in antiferromagnet/ferromagnet/oxide structures. Nat. Nanotechnol. 2016, 11, 878–884. [Google Scholar] [CrossRef]
  19. Fukami, S.; Zhang, C.; DuttaGupta, S.; Kurenkov, A.; Ohno, H. Magnetization switching by spin–orbit torque in an antiferromagnet–ferromagnet bilayer system. Nat. Mater. 2016, 15, 535. [Google Scholar] [CrossRef]
  20. Lau, Y.C.; Betto, D.; Rode, K.; Coey, J.M.; Stamenov, P. Spin–orbit torque switching without an external field using interlayer exchange coupling. Nat. Nanotechnol. 2016, 11, 758–762. [Google Scholar] [CrossRef]
  21. Cai, K.; Yang, M.; Ju, H.; Wang, S.; Ji, Y.; Li, B.; Edmonds, K.W.; Sheng, Y.; Zhang, B.; Zhang, N.; et al. Electric field control of deterministic current-induced magnetization switching in a hybrid ferromagnetic/ferroelectric structure. Nat. Mater. 2017, 16, 712–716. [Google Scholar] [CrossRef]
  22. Ma, Q.; Li, Y.; Gopman, D.B.; Kabanov, Y.P.; Shull, R.D.; Chien, C.L. Switching a Perpendicular Ferromagnetic Layer by Competing Spin Currents. Phys. Rev. Lett. 2018, 120, 117703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Wang, M.; Cai, W.; Zhu, D.; Wang, Z.; Kan, J.; Zhao, Z.; Cao, K.; Wang, Z.; Zhang, Y.; Zhang, T.; et al. Field-free switching of a perpendicular magnetic tunnel junction through the interplay of spin–orbit and spin-transfer torques. Nature Electronics 2018, 1, 582–588. [Google Scholar] [CrossRef]
  24. Liu, Y.; Zhou, B.; Zhu, J.-G. Field-free Magnetization Switching by Utilizing the Spin Hall Effect and Interlayer Exchange Coupling of Iridium. Sci. Rep. 2019, 9, 325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Chang, M.; Yun, J.; Zhai, Y.; Cui, B.; Zuo, Y.; Yu, G.; Xi, L. Field free magnetization switching in perpendicularly magnetized Pt/Co/FeNi/Ta structure by spin orbit torque. Appl. Phys. Lett. 2020, 117, 142404. [Google Scholar] [CrossRef]
  26. Cao, Y.; Sheng, Y.; Edmonds, K.W.; Ji, Y.; Zheng, H.; Wang, K. Deterministic Magnetization Switching Using Lateral Spin–Orbit Torque. Adv. Mater. 2020, 32, 1907929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Wei, J.; Wang, X.; Cui, B.; Guo, C.; Xu, H.; Guang, Y.; Wang, Y.; Luo, X.; Wan, C.; Feng, J.; et al. Field-Free Spin–Orbit Torque Switching in Perpendicularly Magnetized Synthetic Antiferromagnets. Adv. Funct. Mater. 2021, 32, 2109455. [Google Scholar] [CrossRef]
  28. Wu, H.; Nance, J.; Razavi, S.A.; Lujan, D.; Dai, B.; Liu, Y.; He, H.; Cui, B.; Wu, D.; Wong, K.; et al. Chiral Symmetry Breaking for Deterministic Switching of Perpendicular Magnetization by Spin–Orbit Torque. Nano Lett. 2021, 21, 515–521. [Google Scholar] [CrossRef]
  29. Zheng, Z.; Zhang, Y.; Lopez-Dominguez, V.; Sanchez-Tejerina, L.; Shi, J.; Feng, X.; Chen, L.; Wang, Z.; Zhang, Z.; Zhang, K.; et al. Field-free spin–orbit torque-induced switching of perpendicular magnetization in a ferrimagnetic layer with a vertical composition gradient. Nat. Commun. 2021, 12, 4555. [Google Scholar] [CrossRef]
  30. Chen, R.; Cui, Q.; Liao, L.; Zhu, Y.; Zhang, R.; Bai, H.; Zhou, Y.; Xing, G.; Pan, F.; Yang, H.; et al. Reducing Dzyaloshinskii-Moriya interaction and field-free spin–orbit torque switching in synthetic antiferromagnets. Nat. Commun. 2021, 12, 3113. [Google Scholar] [CrossRef]
  31. Xie, X.; Zhao, X.; Dong, Y.; Qu, X.; Zheng, K.; Han, X.; Han, X.; Fan, Y.; Bai, L.; Chen, Y.; et al. Controllable field-free switching of perpendicular magnetization through bulk spin–orbit torque in symmetry-broken ferromagnetic films. Nat. Commun. 2021, 12, 2473. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, T.; Cai, J.W.; Sun, L. Large enhanced perpendicular magnetic anisotropy in CoFeB/MgO system with the typical Ta buffer replaced by an Hf layer. AIP Adv. 2012, 2, 032151. [Google Scholar] [CrossRef] [Green Version]
  33. Ikeda, S.; Miura, K.; Yamamoto, H.; Mizunuma, K.; Gan, H.D.; Endo, M.; Kanai, S.; Hayakawa, J.; Matsukura, F.; Ohno, H. A perpendicular-anisotropy CoFeB-MgO magnetic tunnel junction. Nat. Mater. 2010, 9, 721–724. [Google Scholar] [CrossRef] [PubMed]
  34. Cui, B.; Wu, H.; Li, D.; Razavi, S.A.; Wu, D.; Wong, K.L.; Chang, M.; Gao, M.; Zuo, Y.; Xi, L.; et al. Field-Free spin–orbit Torque Switching of Perpendicular Magnetization by the Rashba Interface. ACS Appl. Mater. Interfaces 2019, 11, 39369–39375. [Google Scholar] [CrossRef] [PubMed]
  35. Reynolds, N.; Jadaun, P.; Heron, J.T.; Jermain, C.L.; Gibbons, J.; Collette, R.; Buhrman, R.A.; Schlom, D.G.; Ralph, D.C. Spin Hall torques generated by rare-earth thin films. Phys. Rev. B 2017, 95, 064412. [Google Scholar] [CrossRef] [Green Version]
  36. Dzyaloshinsky, I. A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics. Sov. Phys. JETP 1957, 5, 1259. [Google Scholar] [CrossRef]
  37. Moriya, T. Anisotropic Superexchange Interaction and Weak Ferromagnetism. Phys. Rev. 1960, 120, 91–98. [Google Scholar] [CrossRef]
  38. Yu, G.; Upadhyaya, P.; Shao, Q.; Wu, H.; Yin, G.; Li, X.; He, C.; Jiang, W.; Han, X.; Amiri, P.K.; et al. Room-Temperature Skyrmion Shift Device for Memory Application. Nano Lett. 2017, 17, 261–268. [Google Scholar] [CrossRef]
  39. Pai, C.-F.; Mann, M.; Tan, A.J.; Beach, G.S.D. Determination of spin torque efficiencies in heterostructures with perpendicular magnetic anisotropy. Phys. Rev. B 2016, 93, 144409. [Google Scholar] [CrossRef] [Green Version]
  40. Cui, B.; Yu, D.; Shao, Z.; Liu, Y.; Wu, H.; Nan, P.; Zhu, Z.; Wu, C.; Guo, T.; Chen, P.; et al. Néel-type elliptical skyrmions in a laterally asymmetric magnetic multilayer. Adv. Mater. 2021, 33, 2006924. [Google Scholar] [CrossRef]
  41. Thiaville, A.; Rohart, S.; Jué, É.; Cros, V.; Fert, A. Dynamics of Dzyaloshinskii domain walls in ultrathin magnetic films. Europhys. Lett. 2012, 100, 57002. [Google Scholar] [CrossRef] [Green Version]
  42. Lee, J.W.; Oh, Y.-W.; Park, S.-Y.; Figueroa, A.I.; van der Laan, G.; Go, G.; Lee, K.-J.; Park, B.-G. Enhanced spin–orbit torque by engineering Pt resistivity in Pt/Co/AlOx structures. Phys. Rev. B 2017, 96, 064405. [Google Scholar] [CrossRef]
  43. Ma, X.; Yu, G.; Tang, C.; Li, X.; He, C.; Shi, J.; Wang, K.L.; Li, X. Interfacial Dzyaloshinskii-Moriya interaction: Effect of 5d band filling and correlation with spin mixing conductance. Phys. Rev. Lett. 2018, 120, 157204. [Google Scholar] [CrossRef] [Green Version]
  44. Wu, D.; Yu, G.; Chen, C.-T.; Razavi, S.A.; Shao, Q.; Li, X.; Zhao, B.; Wong, K.L.; He, C.; Zhang, Z.; et al. Spin–orbit torques in perpendicularly magnetized Ir22Mn78/Co20Fe60B20/MgO multilayer. Appl. Phys. Lett. 2016, 109, 222401. [Google Scholar] [CrossRef]
  45. Ma, X.; Yu, G.; Razavi, S.A.; Sasaki, S.S.; Li, X.; Hao, K.; Tolbert, S.H.; Wang, K.L.; Li, X. Dzyaloshinskii-Moriya interaction across an antiferromagnet-ferromagnet interface. Phys. Rev. Lett. 2017, 119, 027202. [Google Scholar] [CrossRef] [Green Version]
  46. Sagasta, E.; Omori, Y.; Vélez, S.; Llopis, R.; Tollan, C.; Chuvilin, A.; Hueso, L.E.; Gradhand, M.; Otani, Y.; Casanova, F. Unveiling the mechanisms of the spin Hall effect in Ta. Phys. Rev. B 2018, 98, 060410. [Google Scholar] [CrossRef] [Green Version]
  47. Chen, T.-Y.; Chan, H.-I.; Liao, W.-B.; Pai, C.-F. Current-induced spin–orbit torque and field-free switching in Mo-based magnetic heterostructures. Phys. Rev. Appl. 2018, 10, 044038. [Google Scholar] [CrossRef] [Green Version]
  48. Zhang, C.; Fukami, S.; Watanabe, K.; Ohkawara, A.; DuttaGupta, S.; Sato, H.; Matsukura, F.; Ohno, H. Critical role of W deposition condition on spin–orbit torque induced magnetization switching in nanoscale W/CoFeB/MgO. Appl. Phys. Lett. 2016, 109, 192405. [Google Scholar] [CrossRef]
  49. Wu, D.; Yu, G.; Shao, Q.; Li, X.; Wu, H.; Wong, K.L.; Zhang, Z.; Han, X.; Khalili Amiri, P.; Wang, K.L. In-plane current-driven spin–orbit torque switching in perpendicularly magnetized films with enhanced thermal tolerance. Appl. Phys. Lett. 2016, 108, 212406. [Google Scholar] [CrossRef]
Figure 1. (a) Sketch of the multilayer stack of Ta(5)/Gd(1)/CoFeB(w)/MgO(2) (layer thickness in nm). (b) Hall bar device and the measurement configuration. (c) AHE loops for the devices with different CoFeB thicknesses under a current of I = 1 mA. (d) Representative MOKE images of the perpendicular magnetic field induced domain wall motion for the whole Hall bar device with tCoFeB = 0.70 nm. (e) Hall resistance (RH) as a function of the in-plane magnetic field (Hx). (f) The CoFeB thickness dependence of perpendicular magnetic anisotropy energy (Ku).
Figure 1. (a) Sketch of the multilayer stack of Ta(5)/Gd(1)/CoFeB(w)/MgO(2) (layer thickness in nm). (b) Hall bar device and the measurement configuration. (c) AHE loops for the devices with different CoFeB thicknesses under a current of I = 1 mA. (d) Representative MOKE images of the perpendicular magnetic field induced domain wall motion for the whole Hall bar device with tCoFeB = 0.70 nm. (e) Hall resistance (RH) as a function of the in-plane magnetic field (Hx). (f) The CoFeB thickness dependence of perpendicular magnetic anisotropy energy (Ku).
Nanomaterials 12 01887 g001
Figure 2. (a) Field-free SOT switching loops for the devices with different tCoFeB. (b) SOT switching at zero field and in-plane magnetic fields of Hx = ±50 Oe for the device with tCoFeB = 0.81 nm. (c) Representative MOKE images of pulsed current−driven magnetization switching for the whole Hall bar device with tCoFeB = 0.70 nm. (d) The measured SOT efficiency χ as a function of the in-plane magnetic field for different tCoFeB. (e) The relationship between the estimated DMI constant |D| and 1/tCoFeB.
Figure 2. (a) Field-free SOT switching loops for the devices with different tCoFeB. (b) SOT switching at zero field and in-plane magnetic fields of Hx = ±50 Oe for the device with tCoFeB = 0.81 nm. (c) Representative MOKE images of pulsed current−driven magnetization switching for the whole Hall bar device with tCoFeB = 0.70 nm. (d) The measured SOT efficiency χ as a function of the in-plane magnetic field for different tCoFeB. (e) The relationship between the estimated DMI constant |D| and 1/tCoFeB.
Nanomaterials 12 01887 g002
Figure 3. Current-induced out-of-plane effective magnetic fields measured using loops-shift methods. (a) Out-of-plane hysteresis loops under opposite current polarities for tCoFeB = 0.70 nm. (b) β as a function of CoFeB thickness, where the full field-free SOT switching only can be found in the green region.
Figure 3. Current-induced out-of-plane effective magnetic fields measured using loops-shift methods. (a) Out-of-plane hysteresis loops under opposite current polarities for tCoFeB = 0.70 nm. (b) β as a function of CoFeB thickness, where the full field-free SOT switching only can be found in the green region.
Nanomaterials 12 01887 g003
Figure 4. The summarized β as a function of wedged CoFeB thickness in X/CoFeB systems, where X are Pt, IrMn, Gd, Ta, Mo, and W. The inset shows the enlarged IrMn/CoFeB case.
Figure 4. The summarized β as a function of wedged CoFeB thickness in X/CoFeB systems, where X are Pt, IrMn, Gd, Ta, Mo, and W. The inset shows the enlarged IrMn/CoFeB case.
Nanomaterials 12 01887 g004
Table 1. Room temperature β, θSH, and Ds in this work.
Table 1. Room temperature β, θSH, and Ds in this work.
HMβ (Oe/(106 A/cm2))θSHDs (10−15 J/m)
Thin CoFeBThick CoFeB
Pt8.32.20.05~0.15 [42]−965 [43]
IrMn0.200.057 [44]−172 [45]
Gd0.24−1.440.04 [35]−146
Ta−5.51.2−0.05~−0.35 [2,46]36 [43]
Mo−5.11.1−0.003 [47]490 [47]
W−3.52.5−0.14~−0.49 [48]73 [43]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cui, B.; Zhu, Z.; Wu, C.; Guo, X.; Nie, Z.; Wu, H.; Guo, T.; Chen, P.; Zheng, D.; Yu, T.; et al. Comprehensive Study of the Current-Induced Spin–Orbit Torque Perpendicular Effective Field in Asymmetric Multilayers. Nanomaterials 2022, 12, 1887. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12111887

AMA Style

Cui B, Zhu Z, Wu C, Guo X, Nie Z, Wu H, Guo T, Chen P, Zheng D, Yu T, et al. Comprehensive Study of the Current-Induced Spin–Orbit Torque Perpendicular Effective Field in Asymmetric Multilayers. Nanomaterials. 2022; 12(11):1887. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12111887

Chicago/Turabian Style

Cui, Baoshan, Zengtai Zhu, Chuangwen Wu, Xiaobin Guo, Zhuyang Nie, Hao Wu, Tengyu Guo, Peng Chen, Dongfeng Zheng, Tian Yu, and et al. 2022. "Comprehensive Study of the Current-Induced Spin–Orbit Torque Perpendicular Effective Field in Asymmetric Multilayers" Nanomaterials 12, no. 11: 1887. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12111887

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop