Abstract
Due to nonrelativistic altermagnetic spin splitting effect (ASSE), altermagnets can generate time-reversal-odd spin current and spin splitting torque (SST) with spin polarization parallel to the Néel vector. Hence the effective manipulation of SST would provide plenty of opportunities for designable spintronic devices. Here, we achieve electrical control of SST in altermagnetic RuO2, based on controllable Néel vector of RuO2 and Néel vector-dependent generation of SST. To demonstrate the current-induced switching of Néel vector via spin-orbit torque in RuO2 films, we use electrical transport measurements and X-ray magnetic linear dichroism (XMLD) measurements. The switching of Néel vector to the current direction and resultantly enhanced spin polarization parallel to the Néel vector brings about stronger ASSE-induced spin current. These findings not only enrich our understanding of the properties of altermagnets but also pave the way for high-speed memories and nano-oscillators with excellent controllability and efficiency.
Similar content being viewed by others
Introduction
In the field of information storage, nonvolatile magnetic random-access memory (MRAM) with high speed, high density and low dissipation plays a key role, and it requires efficient approaches, especially electrical ones, to recording data1,2,3,4. During the past decades, great progress has been made in spintronics, and spin torques carrying angular momentum has replaced current-induced Oersted field as the state of the art for writing technology5,6,7,8. The widely used spin torques include spin transfer torque (STT) and spin-orbit torque (SOT), and corresponding STT-MRAM and SOT-MRAM possess the strengths of lower power consumption, higher storage density and better reliability9,10,11. On one hand, the STT usually has higher spin torque efficiency due to nonrelativistic ferromagnetic exchange splitting, but relevant devices are more fragile because of the writing current flowing directly through the magnetic tunneling junctions (MTJs)12,13. On the other hand, for SOT-MRAM, although the writing current does not pass through MTJs, the spin torque efficiency is restricted by spin-orbit coupling (SOC) of spin source materials14,15. Therefore, it will be much beneficial to the improvement of data writing technique if the advantages of both STT and SOT can be combined together.
Recently, a new spin splitting torque (SST) related with nonrelativistic altermagnetic spin splitting effect (ASSE) has been predicted theoretically in collinear antiferromagnets with crystal and spin-rotation symmetries (Supplementary Information Note S1), which are termed as altermagnets16,17,18,19,20,21. This SOC-independent ASSE is able to generate nontrivial transverse spin current or out-of-plane polarized spin current with higher efficiency compared to other mechanisms, such as low crystal symmetry22,23,24,25. Experimental evidence of SST has been reported in RuO2 films with rutile crystal structure26,27,28. The SST exhibits controllable spin polarization with its direction parallel to the Néel vector of the altermagnets, which can overcome the limitation of orthogonal relation among charge current, spin polarization and spin current. Not only does RuO2 possess desirable electrical conductivity despite being a metallic oxide, but it also has relatively high Néel temperature above 300 K as well as feasible preparation techniques, which are in favor of being applied in spintronic devices operating at room temperature29,30,31,32,33. While some recent studies raise doubts about the antiferromagnetism of RuO234,35, there are some other latest papers pointing out that the properties of RuO2 can be fragile and may depend on various synthesis conditions36,37,38,39, which highlights the potential uncertainty in present measurements.
Most notably, the Néel vector of RuO2 is reported to be along the [001] crystallographic axis, and hence various behaviors related with SST can be observed in RuO2 films with different crystallographic orientations29,30. For example, RuO2(101) films with tilted Néel vector are not only able to generate SST with out-of-plane polarization in favor of field-free switching of the magnetization in adjacent magnetic layer27,28, but they can also be utilized for spin detection40. To extend the functionalities of altermagnetic spin sources, SST is expected to be manipulated, which will be achieved once the Néel vector can be controlled efficiently in altermagnets. Meanwhile, the effective control of Néel vector is likely to change the band structure of altermagnets, which will also enrich the new physics of altermagnetic materials41. Taking into account the difficulty of using extremely large magnetic field (usually dozens of Tesla) to overcome the anisotropic field of RuO2, it is more applicable to adopt electrical approach as an alternative solution, which is also suitable for device integration42,43.
In this work, we achieve the effective manipulation of Néel vector to tune SST in altermagnetic RuO2 via electrical approach. More importantly, both (i) controllable Néel vector under electrical currents and (ii) Néel vector-dependent generation of spin current and SST are two crucial prerequisites to accomplish this task. As presented in Fig. 1a, a y-direction preset current (Jpreset) flowing inside heavy metal (HM) layer generates SOT and switches the Néel vector (N) in RuO2 layer to be along y-axis as well. The charge current (JC) in RuO2 layer produces SST with spin polarization parallel to N, which is along y-axis (σy). By comparison, if using another Jpreset along x-axis at first, we can align N along x-axis and then obtain adjustable SST with x-direction spin polarization (σx) by applying JC, as displayed in Fig. 1b.
a The y-direction Jpreset flowing in subjacent HM layer generates SOT, leading to the alignment of N to be parallel with y-axis in RuO2 layer, and the charge current in RuO2 layer produces SST with σy. b The x-direction Jpreset switches N to be along the x-direction, giving rise to tunable SST with σx.
Results
Growth and fundamental characterizations of film samples
According to the physical picture of ASSE (Supplementary Information Note S1), RuO2(100) can generate SST by applying charge current while RuO2(110) cannot, though the Néel vector lies in plane for both cases. To investigate SST manipulation in altermagnetic RuO2, we need to achieve controllable Néel vector in RuO2 films with in-plane easy axis. Hence we deposited Pt/RuO2 film samples on Al2O3(0001) and MgO(100) substrates, respectively. The X-ray diffraction (XRD) patterns indicate that RuO2 has strong (100) and (110) texture in Al2O3/Pt/RuO2 and MgO/Pt/RuO2 samples respectively, and they exhibit fine crystallization (Supplementary Information Note S2, Fig. S2). The straight lines of magnetic field-dependent magnetization (M-H) curves measured by superconducting quantum interference device (SQUID) at 300 K (Supplementary Information Note S2, Fig. S3) can be accounted for by the diamagnetism of substrates, and there is negligible net magnetization in Pt/RuO2(100) or Pt/RuO2(110). Besides, the temperature-dependent magnetization (M-T) curves measured by SQUID reveal the Néel temperature of RuO2 films on Al2O3 and MgO substrates to be 400 ~ 450 K (Supplementary Information Note S2, Fig. S4). The Al2O3/Pt/RuO2 and MgO/Pt/RuO2 film samples with desirable crystal quality and magnetization can be used for further depositing comparable Py layers, which are also the basis for conducting following experiments on various device samples (Supplementary Information Note S3).
Current-induced Néel vector switching of RuO2
First we investigate current-induced switching of Néel vector in altermagnetic RuO2 by electrical transport measurement at room temperature. We fabricated eight-terminal devices of Pt(5 nm)/RuO2(15 nm) on Al2O3 and MgO substrates. As shown in the inset of Fig. 2a, each device has a pair of vertical (Iwrite1) and horizontal (Iwrite2) channels with the width of 20 μm. After applying each writing current pulse with different magnitude along one certain channel, the change of transverse Hall resistance (ΔRHall) is plotted in Fig. 2a for RuO2(100) (red square) and RuO2(110) (blue diamond). When we increase the magnitude of applied current pulses, no remarkable ΔRHall signals can be observed until it reaches threshold value, corresponding with the current densities of 5.3 × 1011 A m–2 and 4.8 × 1011 A m–2 in Pt layer (JPt), for Pt/RuO2(100) and Pt/RuO2(110) samples respectively28,44. Considering the characteristics of the power source used in our experiments, we further calibrate the actual threshold values to be 2.4 × 1011 A m–2 and 1.9 × 1011 A m–2 for Pt/RuO2(100) and Pt/RuO2(110) samples respectively (Supplementary Information Note S4). These critical values can be taken as reference to conduct the following experiments. The slight discrepancy of critical switching current densities is mainly caused by the different anisotropy when the films were deposited on different substrates.
a ΔRHall as a function of current density in the Pt layer, with each data point being collected after applying one pulse current. The inset shows the schematic of eight-terminal device for electric transport measurement, where the read current and voltage paths are mutually perpendicular and are aligned 45° to the two writing channels. b Pulse number dependence of ΔRHall of Pt/RuO2(100) (red square) and Pt/RuO2(110) (blue diamond) devices under critical switching current densities. Ru M-edge XMLD results of Pt/RuO2 devices for c RuO2(100) and d RuO2(110). From top to bottom: before applying writing currents (Original), after applying a group of five writing currents along vertical (Iwrite1) channel and along horizontal (Iwrite2) channel.
Then we use critical switching currents to test each Pt/RuO2 device for three cycles: a group of five successive pulse currents with the same amplitude are applied alternatively along the two orthogonal channels. Figure 2b displays ΔRHall of RuO2(100) (red square) and RuO2(110) (blue diamond) as a function of pulse numbers. For the two Pt/RuO2 devices, the obvious ΔRHall is observed after applying the first pulse current, and it increases gradually regarding the second to the fifth pulses. By comparison, no apparent ΔRHall signals can be detected in Pt or RuO2 devices when 1 ms-wide current pulses are applied (Supplementary Information Note S5). Significantly, previous literature has reported that thermal artifacts could be observed in electrical Hall resistance measurements under specific circumstances45. Hence to study the Joule heating caused by applied current pulses46, we meticulously conducted both experiment measurement and finite element simulation, and the results distinctly demonstrate that the temperature of device samples remains under the Néel temperature from start to finish (Supplementary Information Note S6). Therefore, the recorded ΔRHall of Pt/RuO2 devices can be attributed to current-induced Néel vector switching. The Hall resistance measurements preliminarily prove the in-plane Néel vector switching with 90° in altermagnetic RuO2, and partial switching of domains is likely to exist here47,48,49,50.
To further verify the in-plane Néel vector switching in altermagnetic RuO2 films, we performed the measurement of Ru M-edge X-ray magnetic linear dichroism (XMLD) at room temperature. By this synchrotron radiation test technique which is sensitive to in-plane Néel vector, we can reveal the distribution of Néel vector in RuO2(100) and RuO2(110) and its dependence on writing currents51,52,53. Considering the compatibility with the size of detecting spot (80 × 80 μm2) for XMLD measurement, we fabricated the eight-terminal devices with 100 μm-wide writing channels for Pt(5 nm)/RuO2(15 nm) in which the crystalline orientations of RuO2 are (100) and (110). The detecting light is focused on the intersection area of the device, and the signals of linearly polarized X-ray absorption spectroscopy (XAS) are collected (Supplementary Information Note S7, Fig. S12 and Fig. S13). The peak position of XAS curves near 462 eV corresponds to the M3 3p3/2 binding energy of Ru element. The XMLD results presented in Fig. 2c and Fig. 2d are obtained by the difference of linearly vertical (V) and horizontal (H) polarized XAS signals.
Before applying writing currents, we collected XMLD signals of the initial state of both devices. Due to non-epitaxial growth mode, twin crystal structure with different in-plane alignment of Néel vector exists in RuO2 films26. Hence the in-plane Néel vectors without uniaxial orientation cancel out each other, leading to no peaks or valleys in XMLD results for the initial state of both switching devices, which is presented by the (black circle) curves at the top of Fig. 2c and Fig. 2d. For each eight-terminal device, writing currents are applied by following two schemes: a group of successive five pulse currents flow through either the vertical channel (Iwrite1) or the horizontal channel (Iwrite2). The nominal values of JPt are 5.4 × 1011 A m–2 and 4.9 × 1011 A m–2 for Pt/RuO2(100) and Pt/RuO2(110) devices, respectively. To ensure the repeatability of testing results, we have measured sufficient switching devices after applying writing currents. Noticeably, spikes appear in the XMLD signals after applying writing currents, as evidenced by the representative peaks and valleys in middle (red square) and bottom (blue diamond) curves in Fig. 2c, d. After Iwrite1 flows through the vertical channel of Pt/RuO2(100) and Pt/RuO2(110) devices, the XMLD curves undergo zero-negative-positive-zero trace as photon energy increases, as displayed by two (red square) curves in the middle of Fig. 2c and Fig. 2d. Meanwhile, after we applied Iwrite2 to these devices, the distinct XMLD signals can be seen in the two (blue diamond) curves at the bottom of Fig. 2c, d.
It is reasonable to deduce that after we used writing currents which reached the threshold, the Néel vector is capable of being switched towards the direction of charge current in Pt/RuO2 devices. In contrast to their counterparts in the middle, the two curves at the bottom of Fig. 2c, d reveal that the XMLD signals exhibit a zero-positive-negative-zero tendency with increasing photon energy, which indicates the reversed polarity of XMLD signals after we rotated the direction of writing currents with in-plane 90°. Of note that there is a delay time from the application of current pulses to the measurement of XAS (or XMLD) signals. We have prolonged the delay time to make another test and observed the same phenomenon in XMLD patterns, which proves the stability of XMLD measurement results (Supplementary Information Note S7, Fig. S14). Therefore, the onefold orientation of Néel vector after applying writing currents is determined by the direction of applied charge currents, which results from SOT-induced Néel vector switching. The above XMLD measurements of Ru M-edge support the current-induced switching of Néel vector with in-plane 90° in altermagnetic RuO2, which is an important basis for manipulating SST because of the close relation between SST and the state of Néel vector. The XMLD data are compatible with the presence of antiferromagnetic order39, besides the previous evidence by transport measurements such as the anomalous Hall effect and SST26,27,28,42.
Manipulation of spin splitting torque in RuO2
To study the electrical control of spin splitting torque related to spin current generation in altermagnetic RuO2, we fabricated Pt/RuO2(100)/Py and Pt/RuO2(110)/Py devices for spin torque ferromagnetic resonance (ST-FMR) measurements, where the thickness of each layer is indicated by Pt(5 nm)/RuO2(10 nm)/Py(15 nm). It is worth noting that this well-established ST-FMR technique is particularly appropriate for detecting and analyzing spin currents with various spin polarizations (σx, σy and σz)24,27,54,55.
Figure 3a illustrates the setup of ST-FMR measurement to characterize the SST in Pt(5 nm)/RuO2(10 nm)/Py(15 nm) devices at room temperature. A radio-frequency (RF) charge current flowing through the RuO2 layer is converted into transverse spin current, and then it is injected into the adjacent Py layer. According to the physical picture of ASSE (Supplementary Information Note S1), the spin-current generation is contributed by ASSE together with spin Hall effect (SHE) in Pt/RuO2(100)/Py, while it is solely due to SHE in Pt/RuO2(110)/Py. Under the applied magnetic field, the oscillating spin current exerts spin torques on ferromagnetic Py layer leading to magnetization precession. The external magnetic field is swept with φH (angle between charge current and applied magnetic field), which can be regarded the same as φM (angle between charge current and magnetization in Py) because of the relatively small in-plane anisotropy of Py.
a The measurement setup of ST-FMR. b A typical ST-FMR spectrum consisted of antisymmetric (VA, red curve) and symmetric (VS, blue curve) signals under φH = 45° after applying writing currents of 60 mA. The analysis of spin torques with various polarization directions by fitting the angle-dependent VS results after applying writing currents of (c) 60 mA and (d) 150 mA. All data in panels (b, c, d) are obtained by measuring Pt/RuO2(100)/Py device under 6 GHz and 19 dBm. The |Vσx/Vσy| data points of (e) Pt/RuO2(100)/Py and f Pt/RuO2(110)/Py devices are displayed in terms of different applied writing currents. The |Vσx/Vσy| indicates the ratio of dampinglike torques with x-direction polarization (σx) to those with y-direction polarization (σy). The data points with error bars in panels (e and f) are fitted by B-spline curves, and the insets refer to the state of SST before and after Néel vector switching in Pt/RuO2(100)/Py and Pt/RuO2(110)/Py devices.
Figure 3b displays a typical ST-FMR spectrum of Pt/RuO2(100)/Py device after applying a group of writing currents (Iwrite) of 60 mA, and it is measured with φH = 45° under the excitation of 6 GHz and 19 dBm microwave current. As shown in Fig. 3b (with more data of the same type shown in Supplementary Information Note S8, Fig. S15), the (black) dots are measured DC voltage data, and corresponding fitted Vmix curve (green one) can be decomposed into both symmetric (VS, blue one) and antisymmetric (VA, red one) Lorentz line shapes. From the voltage signals, we can see that the amplitude of VA (red line) is comparatively much larger than that of VS (blue line), for the Oersted field plays a remarkable role in the former resulting from the shunting effect.
Each time after applying a group of five successive writing currents with the same magnitude to Pt/RuO2/Py devices, we conducted an angle-dependent ST-FMR measurement to characterize the SST in altermagnetic RuO2. The magnetic field was applied in the plane of tested samples, and the angle between the Pt/RuO2/Py devices and magnetic field was defined as φH. We swept φH with a step size of 15° under 6 GHz and 19 dBm microwave currents. Of note is that RuO2(100) and RuO2(110) studied in the present case can only generate in-plane spin polarizations (σx and σy). Besides, VS and VA are related with in-plane and out-of-plane current-induced spin torques respectively (Supplementary Information Note S8, Table S2). Therefore, we are able to analyze spin torques with in-plane polarization directions in Pt/RuO2(100)/Py device by trigonometric function fitting, according to the acquired VS data. The fitting results after applying Iwrite = 60 mA and Iwrite = 150 mA (nominally) are presented in Fig. 3c, d, which are below and above the critical switching current density, respectively. As depicted by (purple and orange) curves in Fig. 3c, d, prominent \({S}_{{{{\rm{DL}}}}}^{X}\) can be observed in the latter, while \({S}_{{{{\rm{DL}}}}}^{Y}\) remains almost the same under the two circumstances.
Notably, both SST (by ASSE) and SOT (by SHE) give rise to σy, but σx can only be attributed to SST. Therefore, we extract the ratios of \({S}_{{{{\rm{DL}}}}}^{X}\) to \({S}_{{{{\rm{DL}}}}}^{Y}\) from the angle-dependent measurements as |Vσx/Vσy| in terms of the writing currents. Correspondingly, the data points of |Vσx/Vσy| connected by B-spline lines are presented in Fig. 3e, f for Pt/RuO2(100)/Py and Pt/RuO2(110)/Py, respectively. Taking into account the multiple steps of data fitting with possible propagation of fitting errors, error bars are also displayed accompanying the final data points in Fig. 3e, f. As the magnitude of writing currents rises, the |Vσx/Vσy| for Pt/RuO2(100)/Py fluctuates within a reasonable range, and then it increases abruptly at Iwrite = 130 mA, as displayed in Fig. 3e. Under this situation, the corresponding nominal writing current density is about 5.1 × 1011 A m–2, which reaches the nominal threshold writing current density to trigger Néel vector switching. This result is further verified by repeated experiment (Supplementary Information Note S8, Fig. S16), whose critical value of current densities is also in accordance with the previous results measured by SOT switching experiments. Hence this current-induced Néel vector switching in RuO2 layer leads to the alignment of Néel vector to be along x-axis (being parallel to the direction of writing currents). Considering the potential partial switching of Néel vector, further improvement of switching ratio of domains will probably enhance the value of |Vσx/Vσy| to a larger extent. By contrast, Fig. 3f shows that there is no apparent enhancement for the σx component in Pt/RuO2(110)/Py as the writing currents increase, and typical angular-dependent ST-FMR results of Pt/RuO2(110)/Py (Supplementary Information Note S8, Fig. S17) are also different from those of Pt/RuO2(100)/Py shown in Fig. 3c, d.
Discussion
In view of the situation where the Néel vector lies in plane, RuO2(100) films have higher spin torque efficiency than RuO2(110) films do because of the existence of SST in the former, and the resultant spin torque conductivity surpasses that of many heavy metals and topological insulators, which are potential candidates for efficient spin sources10,14,26,56. However, the spin polarization directions of SST are unable to be changed in previous studies. To address this issue, a feasible solution is to manipulate the Néel vector in altermagnets by electrical currents. Once the writing currents meet the threshold, the SST can be tuned by applying charge currents in altermagnetic RuO2(100) layer along x-axis: its spin polarization direction is parallel to the switched Néel vector, which is shifted from σy to σx (Fig. 3e). After the Néel vector is switched in RuO2(100), the proportion of σx enhances but σy still dominates, which can be ascribed to conventional spin Hall effect. By comparison, no SST can be generated in (110)-oriented RuO2 by applied charge currents, according to the physical picture of ASSE mechanism (Fig. 3f). Hence the generated transverse spin current comes entirely from the conventional SHE depending on spin-orbit coupling, and the spin torques regarding different in-plane coordinate axes cannot be tuned, which is irrelevant to the Néel vector switching in RuO2(110).
It has been studied that magnetic moment-dependent spin torques can be produced by magnetic SHE (e.g., in Mn3Sn) or antiferromagnetic SHE (e.g., in Mn2Au), abbreviated as MSHE or AFM-SHE respectively25,57. Therefore, altermagnetic materials including RuO2 with ASSE offers a new way to generate the spin torques besides the above two counterparts. Meanwhile, the electric field control of Néel vector in collinear antiferromagnet Mn2Au with AFM-SHE is also able to tune the Néel spin-orbit torque (NSOT), which is different from the case in non-collinear Mn3Sn57,58. Sharing the similarity of Néel vector-dependent generation of spin torques, the tunable spin polarization of SST is parallel with the Néel vector in altermagnetic RuO2 while that of NSOT is perpendicular to the Néel vector in Mn2Au. Furthermore, compared to Mn2Au relying on multi-step Rashba-like mechanism, our RuO2 depends on nonrelativistic altermagnetic spin splitting to generate spin current and spin splitting torque with tunable polarization and higher efficiency.
In summary, we successfully achieve the electrical manipulation of altermagnetic RuO2 to tune the spin splitting torque, which is accomplished based on (i) controllable Néel vector of RuO2 and (ii) Néel vector-dependent generation of spin current and SST due to ASSE. By applying electrical writing currents, the Néel vector in altermagnetic RuO2 with in-plane easy axis is switched by 90°. This current-induced switching is evidenced by electrical transport measurements of transverse Hall resistance and it is also manifested unambiguously by XMLD results in terms of the writing currents (and their directions). Moreover, the spin polarizations of SST along different coordinate axes are analyzed by the angular dependence of ST-FMR voltage signals, and it proves distinctly that the SST can be tuned based on the control of Néel vector in altermagnetic RuO2. The effective manipulation of spin splitting torque in RuO2 via a feasible electrical approach not only provides important additional piece of information for understanding the altermagnetism of RuO2, but it is also beneficial to expanding the potential applications of altermagnetic spin sources with controllable SST as the efficient writing technology in the field of information storage involving MRAM.
Methods
Sample preparation
By DC magnetron sputtering method, two groups of Pt(5 nm)/RuO2(15 nm) film samples were prepared on single-crystal Al2O3(0001) and MgO(100) substrates, and single Pt(5 nm) and RuO2(15 nm) films were also deposited directly on the two substrates. The Pt layers were deposited by sputtering Pt target with Ar flow of 20 at 773 K under a base pressure below 5 × 10–5 Pa, and then they were annealed at 973 K for 1 hour. The RuO2 layers were in-situ deposited by sputtering Ru target with Ar:O2 flow of 20:5 at 773 K. The deposition rates of Pt and RuO2 films were 0.69 nm/min and 3.3 nm/min, respectively. For ST-FMR measurement, we deposited Pt(5 nm)/RuO2(10 nm)/Py(15 nm) film samples on Al2O3(0001) and MgO(100) substrates. After depositing Pt and RuO2 layers with the parameters introduced above, we in-situ sputtered Py target with Ar flow of 20 at room temperature with the deposition rate of 1.33 nm/min. Also, the Pt(5 nm)/RuO2(10 nm)/Py(15 nm) film samples were in-situ covered by 1 nm-thick Ru to prevent them from oxidation.
Sample characterization
Crystal quality and phase composition were analyzed by X-ray diffraction (XRD) characterization utilizing a Rigaku Smartlab instrument with Cu-Kα radiation (wavelength = 0.154 nm). The working voltage and working current were 40 kV and 150 mA, respectively. The scanning rate was 10°/min for each XRD pattern. Magnetic field dependence of magnetization (M-H) curves were measured via a superconducting quantum interference device (SQUID) from Quantum Design at room temperature. Temperature dependence of magnetization (M-T) curves were measured via the same SQUID instrument from 300 to 600 K. Photographs of three kinds of devices were taken by a Nikon ECLIPSE LV150NL optical microscope.
Fabrication of devices
By standard photolithography and Ar-ion milling techniques, the deposited Pt(5 nm), RuO2(15 nm) and Pt(5 nm)/RuO2(15 nm) films on single-crystal Al2O3(0001) and MgO(100) substrates were patterned into eight-terminal devices. The width of two orthogonal writing channels was 20 μm (for electrical transport measurement) or 100 μm (for XAS and XMLD measurement), and the two diagonal detecting paths were 5 μm wide. By standard photolithography and Ar-ion milling techniques, the deposited Pt(5 nm)/RuO2(10 nm)/Py(15 nm)/Ru(1 nm) films on Al2O3(0001) and MgO(100) substrates were patterned into stripes whose length and width were 50 μm and 20 μm, respectively. Then the Cr(10 nm)/Cu(80 nm) electrode was deposited on the stripes via electron-beam evaporation method.
Electrical transport measurement
The two diagonal detecting paths of Pt(5 nm), RuO2(15 nm) and Pt(5 nm)/RuO2(15 nm) devices were connected to a Keitheley 2400 current source offering a reading current of 1 mA and a Keitheley 2182 nanovoltmeter recording transverse Hall voltage signals. A Keysight 2901 power source was adopted to output current pulses, which was also able to monitor the real-time resistance change of tested devices. A group of five successive pulse currents with the same amplitude were applied along the horizontal channel (red arrow Iwrite1 in the inset of Fig. 2a), and then another group of writing currents were applied along the vertical channel (blue arrow Iwrite2 in the inset of Fig. 2a), which constituted a testing cycle. After applying each current pulse, the calculated gap value of Hall resistance with respect to the initial value before applying this group of current pulses is defined as the ΔRHall.
XAS and XMLD measurements
XAS and XMLD measurements in total-electron-yield mode were carried out at the beam line BL08U1A in the Shanghai Synchrotron Radiation Facility at 300 K. We used this technique to probe into the in-plane distribution of Néel vector in altermagnetic RuO2. The XMLD curves were the differences of linearly vertical (V) and horizontal (H) polarized XAS signals. The detecting light was perpendicular to the sample plane, and the spot with the size of 80×80 μm2 was focused on the intersection of eight-terminal devices.
ST-FMR measurement
We adopted ST-FMR technique to characterize generated spin current and spin splitting torque with different polarizations in altermagnetic RuO2. The sample holder was connected to a Ceyear 1465 F microwave source (0.1 ~ 40 GHz) via a bias-T which a Keitheley 2182 nanovoltmeter was also linked to. External magnetic field was swept along φH (0° ≤ φH ≤ 360°) with step size of 15°, and microwave current with 19 dBm and 6 GHz was flowing through the stripe. Meanwhile, the direct current (DC) voltage signals resulting from the anisotropic magnetoresistance (AMR) rectification of Py were recorded by the nanovoltmeter, which were consisted of symmetric and antisymmetric Lorentz line shapes, with the expression V = VSLS + VALA. As for LA = ΔH(H-Hres)/[(H-Hres)2 + ΔH2] and LS = (ΔH)2/[(H-Hres)2 + ΔH2], ΔH, Hres and H represent line width, resonance field and external magnetic field, respectively. For the spin torques due to in-plane spin polarizations, the angular dependence of VS and VS could be described by the following equations23,26:
where \({S}_{{{{\rm{DL}}}}}^{X}\) and \({S}_{{{{\rm{DL}}}}}^{Y}\) were the coefficients of dampinglike (DL) torques, and \({A}_{{{{\rm{FL}}}}}^{X}\) and \({A}_{{{{\rm{FL}}}}}^{Y}\) were those of fieldlike (FL) torques, with \({A}_{{{{\rm{FL}}}}}^{Y}\) also including the contribution by Oersted field torque.
Finite element simulations
We utilized finite element simulations for analyzing the temperature rise due to applied current pulses in Pt/RuO2 devices. The simulated structures were Sub./Pt(5 nm)/RuO2(15 nm) where Sub. referred to single-crystal Al2O3(0001) and MgO(100) substrates. Meanwhile, both writing channels of the simulated devices were 20 μm wide, which was in line with the device structure in electrical transport measurement. For the two Pt/RuO2 devices, we applied a group of five 1 ms-wide current pulses with critical value, and the interval between two successive pulses was 10 s. The thermal parameters of RuO2, including thermal conductivities (0.50 W cm–1 K–1) and specific heat capacities (56.42 J mol-1 K-1), were obtained from literature59,60, and those of Pt, Al2O3 and MgO were acquired directly from the build-in database of simulation software. The simulated results (Supplementary Information Fig. S11) could be classified as two categories: (i) time-dependent temperature variation, and (ii) space-dependent temperature distribution. For category (i), the monitoring space point was located at the center of the Pt/RuO2 switching devices (distance = 0 μm). For category (ii), the monitoring time point was in the middle of one current pulse (t = 0.5 ms).
Data availability
The data supporting the findings of this study are included in the paper and its Supplementary Information file. Further data sets are available from the corresponding author on reasonable request. Source data are provided with this paper.
References
Bhatti, S. et al. Spintronics based random access memory: a review. Mater. Today 20, 530–548 (2017).
Parkin, S. S. P. et al. Exchange-biased magnetic tunnel junctions and application to nonvolatile magnetic random access memory. J. Appl. Phys. 85, 5828–5833 (1999).
Wang, W. G., Li, M., Hageman, S. & Chien, C. L. Electric-field-assisted switching in magnetic tunnel junctions. Nat. Mater. 11, 64–68 (2012).
Kimel, A. V. & Li, M. Writing magnetic memory with ultrashort light pulses. Nat. Rev. Mater. 4, 189–200 (2019).
Jungwirth, T., Marti, X., Wadley, P. & Wunderlich, J. Antiferromagnetic spintronics. Nat. Nanotechnol. 11, 231–241 (2016).
Chumak, A. V., Vasyuchka, V. I., Serga, A. A. & Hillebrands, B. Magnon spintronics. Nat. Phys. 11, 453–461 (2015).
Linder, J. & Robinson, J. W. Superconducting spintronics. Nat. Phys. 11, 307–315 (2015).
Železný, J., Wadley, P., Olejník, K., Hoffmann, A. & Ohno, H. Spin transport and spin torque in antiferromagnetic devices. Nat. Phys. 14, 220–228 (2018).
Slonczewski, J. C. Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater. 159, L1–L7 (1996).
Liu, L. et al. Spin-torque switching with the giant spin Hall effect of tantalum. Science 336, 555–558 (2012).
Baltz, V. et al. Antiferromagnetic spintronics. Rev. Mod. Phys. 90, 015005 (2018).
Fong, X. et al. Spin-transfer torque memories: Devices, circuits, and systems. Proc. IEEE 104, 1449–1488 (2016).
Zhao, W. S. et al. Failure and reliability analysis of STT-MRAM. Microelectron. Reliab. 52, 1848–1852 (2012).
Song, C. et al. Spin-orbit torques: Materials, mechanisms, performances, and potential applications. Prog. Mater. Sci. 118, 100761 (2021).
Han, X. F., Wang, X., Wan, C. H., Yu, G. Q. & Lv, X. R. Spin-orbit torques: Materials, physics, and devices. Appl. Phys. Lett. 118, 120502 (2021).
González-Hernández, R. et al. Efficient electrical spin splitter based on nonrelativistic collinear antiferromagnetism. Phys. Rev. Lett. 126, 127701 (2021).
Yuan, L. D., Wang, Z., Luo, J. W., Rashba, E. I. & Zunger, A. Giant momentum-dependent spin splitting in centrosymmetric low-Z antiferromagnets. Phys. Rev. B 102, 014422 (2020).
Naka, M., Motome, Y. & Seo, H. Perovskite as a spin current generator. Phys. Rev. B 103, 125114 (2021).
Naka, M. et al. Spin current generation in organic antiferromagnets. Nat. Commun. 10, 4305 (2019).
Šmejkal, L., Sinova, J. & Jungwirth, T. Beyond Conventional Ferromagnetism and Antiferromagnetism: A Phase with Nonrelativistic Spin and Crystal Rotation Symmetry. Phys. Rev. X 12, 031042 (2022).
Šmejkal, L., Sinova, J. & Jungwirth, T. Emerging research landscape of altermagnetism. Phys. Rev. X 12, 040501 (2022).
Šmejkal, L., González-Hernández, R., Jungwirth, T. & Sinova, J. Crystal time-reversal symmetry breaking and spontaneous Hall effect in collinear antiferromagnets. Sci. Adv. 6, eaaz8809 (2020).
Ma, H. Y. et al. Multifunctional antiferromagnetic materials with giant piezomagnetism and noncollinear spin current. Nat. Commun. 12, 2846 (2021).
MacNeill, D. et al. Control of spin-orbit torques through crystal symmetry in WTe2/ferromagnet bilayers. Nat. Phys. 13, 300–305 (2017).
Chen, X. Z. et al. Observation of the antiferromagnetic spin Hall effect. Nat. Mater. 20, 800–804 (2021).
Bai, H. et al. Observation of spin splitting torque in a collinear antiferromagnet RuO2. Phys. Rev. Lett. 128, 197202 (2022).
Bose, A. et al. Tilted spin current generated by the collinear antiferromagnet ruthenium dioxide. Nat. Electron. 5, 267–274 (2022).
Karube, S. et al. Observation of spin-splitter torque in collinear antiferromagnetic RuO2. Phys. Rev. Lett. 129, 137201 (2022).
Berlijn, T. et al. Itinerant antiferromagnetism in RuO2. Phys. Rev. Lett. 118, 077201 (2017).
Zhu, Z. H. et al. Anomalous antiferromagnetism in metallic RuO2 determined by resonant X-ray scattering. Phys. Rev. Lett. 122, 017202 (2019).
Ruf, J. P. et al. Strain-stabilized superconductivity. Nat. Commun. 12, 59 (2021).
Uchida, M., Nomoto, T., Musashi, M., Arita, R. & Kawasaki, M. Superconductivity in uniquely strained RuO2 films. Phys. Rev. Lett. 125, 147001 (2020).
Bai, H. et al. Antiferromagnetism: An efficient and controllable spin source. Appl. Phys. Rev. 9, 041316 (2022).
Hiraishi, M. et al. Nonmagnetic ground state in RuO2 revealed by Muon spin rotation. Phys. Rev. Lett. 132, 166702 (2024).
Keßler, P. et al. Absence of magnetic order in RuO2: insights from μSR spectroscopy and neutron diffraction. npj Spintronics 2, 50 (2024).
Guo, Y. Q. et al. Direct and inverse spin splitting effects in altermagnetic RuO2. Adv. Sci. 11, 2400967 (2024).
Smolyanyuk, A., Mazin, I. I., Garcia-Gassull, L. & Valentí, R. Fragility of the magnetic order in the prototypical altermagnet RuO2. Phys. Rev. B 109, 134424 (2024).
Lin, Z. H. et al. Bulk band structure of RuO2 measured with soft x-ray angle-resolved photoemission spectroscopy. Phys. Rev. B 111, 134450 (2025).
Zhang, Y. C. et al. Probing the Néel order in altermagnetic RuO2 films via X-ray magnetic linear dichroism. Chin. Phys. Lett. 42, 027301 (2025).
Bai, H. et al. Efficient spin-to-charge conversion via altermagnetic spin splitting effect in antiferromagnet RuO2. Phys. Rev. Lett. 130, 216701 (2023).
Han, L. et al. Electrical 180o switching of Néel vector in spin-splitting antiferromagnet. Sci. Adv. 10, eadn0479 (2024).
Feng, Z. X. et al. An anomalous Hall effect in altermagnetic ruthenium dioxide. Nat. Electron. 5, 735–743 (2022).
Song, C., Cui, B., Li, F., Zhou, X. J. & Pan, F. Recent progress in voltage control of magnetism: Materials, mechanisms, and performance. Prog. Mater. Sci. 87, 33–82 (2017).
Qiu, X. Y. et al. Oxygen-dependent epitaxial growth of Pt(001) thin films on MgO(001) by magnetron sputtering. Appl. Surf. Sci. 406, 212–217 (2017).
Chiang, C. C., Huang, S. Y., Qu, D., Wu, P. H. & Chien, C. L. Absence of evidence of electrical switching of the antiferromagnetic Néel vector. Phys. Rev. Lett. 123, 227203 (2019).
Bodnar, S. Y. et al. Writing and reading antiferromagnetic Mn2Au by Néel spin-orbit torques and large anisotropic magnetoresistance. Nat. Commun. 9, 348 (2018).
Moriyama, T., Oda, K., Ohkochi, T., Kimata, M. & Ono, T. Spin torque control of antiferromagnetic moments in NiO. Sci. Rep. 8, 14167 (2018).
Wadley, P. et al. Electrical switching of an antiferromagnet. Science 351, 587–590 (2016).
DuttaGupta, S. et al. Spin-orbit torque switching of an antiferromagnetic metallic heterostructure. Nat. Commun. 11, 5715 (2020).
Chen, X. Z. et al. Control of spin current and antiferromagnetic moments via topological surface state. Nat. Electron. 5, 574–578 (2022).
Wu, H. et al. Current-induced Néel order switching facilitated by magnetic phase transition. Nat. Commun. 13, 1629 (2022).
Huang, Y. H. et al. A Spin-Orbit Torque Ratchet at Ferromagnet/Antiferromagnet Interface via Exchange Spring. Adv. Funct. Mater. 32, 2111653 (2022).
Yang, M. et al. Current switching of the antiferromagnetic Néel vector in Pd/CoO/MgO(001). Phys. Rev. B 106, 214405 (2022).
Liu, L. Q., Moriyama, T., Ralph, D. C. & Buhrman, R. A. Spin-torque ferromagnetic resonance induced by the spin Hall effect. Phys. Rev. Lett. 106, 036601 (2011).
Pai, C. F. et al. Spin transfer torque devices utilizing the giant spin Hall effect of tungsten. Appl. Phys. Lett. 101, 122404 (2012).
Mellnik, A. R. et al. Spin-transfer torque generated by a topological insulator. Nature 511, 449–451 (2014).
Kimata, M. et al. Magnetic and magnetic inverse spin Hall effects in a non-collinear antiferromagnet. Nature 565, 627–630 (2019).
Chen, X. Z. et al. Electric field control of Néel spin-orbit torque in an antiferromagnet. Nat. Mater. 18, 931–935 (2019).
Cordfunke, E. H. P., Konings, R. J. M., Westrum, E. F. Jr & Shaviv, R. The thermophysical and thermochemical properties of RuO2 from 0 to 1000 K. J. Phys. Chem. Solids 50, 429–434 (1989).
Ferizovic, D., Hussey, L. K., Huang, Y. S. & Muñoz, M. Determination of the room temperature thermal conductivity of RuO2 by the photothermal deflection technique. Appl. Phys. Lett. 94, 131913 (2009).
Acknowledgements
Professor Wanjun Jiang and Le Zhao from Tsinghua University are thanked for the help with experiment. Professor Yi Wang, Tuo Zhang, Fanyu Meng and Pengzhen Li from Dalian University of Technology are thanked for the help with revising the paper. This work is supported by the National Key Research and Development Program of China (Grant No. 2021YFB3601301) (C.S.), the National Natural Science Foundation of China (Grant No. 52225106, 12241404, T2394471, and 523B1007) (C.S.), the Open Fund of the State Key Laboratory of Spintronics Devices and Technologies (Grant No. SPL-2401) (C.S.), and the Fundamental Research Funds for the Central Universities (Grant No. DUT25RC(3)066) (Y.Z. (Yichi Zhang)). We acknowledge the beam line BL08U1A in Shanghai Synchrotron Radiation Facility (SSRF) for the XAS and XMLD measurements and the nanofabrication by Ultraviolet Maskless lithography machine (model: Uv litho- AcA, tuotuo technology).
Author information
Authors and Affiliations
Contributions
F.P. and C.S. supervised the project. Y.Z. (Yichi Zhang) conceived the idea and designed the experiments. Y.Z. (Yichi Zhang) conducted experiments, and H.B., J.D., L.H., C.C., S.L., Y.C., Y.Z. (Yingying Zhang), Q.W. and W.Z. helped with the experiments and data analysis. Y.Z. (Yichi Zhang) wrote the original draft of the paper, and H.B., J.D., L.H., and C.S. revised and edited the paper.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Stanislav Bodnar, Francesco Maccherozzi and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Figure caption are provided twice so one of them has been deleted. So Please take care
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Zhang, Y., Bai, H., Dai, J. et al. Electrical manipulation of spin splitting torque in altermagnetic RuO2. Nat Commun 16, 5646 (2025). https://doi.org/10.1038/s41467-025-60891-2
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-60891-2
This article is cited by
-
Evidence for single variant in altermagnetic RuO2(101) thin films
Nature Communications (2025)