Abstract
Antiferromagnetic spintronics have attracted wide attention due to its great potential in constructing ultradense and ultrafast antiferromagnetic memory that suits modern high-performance information technology. The electrical 180° switching of Néel vector is a long-term goal for developing electrical-controllable antiferromagnetic memory with opposite Néel vectors as binary “0” and “1.” However, the state-of-art antiferromagnetic switching mechanisms have long been limited for 90° or 120° switching of Néel vector, which unavoidably require multiple writing channels that contradict ultradense integration. Here, we propose a deterministic switching mechanism based on spin-orbit torque with asymmetric energy barrier and experimentally achieve electrical 180° switching of spin-splitting antiferromagnet Mn5Si3. Such a 180° switching is read out by the Néel vector–induced anomalous Hall effect. On the basis of our writing and readout methods, we fabricate an antiferromagnet device with electrical-controllable high- and low-resistance states that accomplishes robust write and read cycles. Besides fundamental advance, our work promotes practical spin-splitting antiferromagnetic devices based on spin-splitting antiferromagnet.
The electrical 180° switching and detection of the Néel vector in spin-splitting antiferromagnet Mn5Si3 is achieved.
INTRODUCTION
One of the most well-known applications of spintronics is the electrical-controllable nonvolatile magnetic random-access memory (MRAM), where the information states “1” and “0” are stored by the up and down directions of the order parameter in ferromagnet (FM), i.e., the magnetic moment (1–3). For the original MRAM, the electrical read is based on tunneling magnetoresistance (4–6) (TMR), and the electrical write is realized by applying current along single channel to switch the magnetic moment due to spin transfer torques (7, 8). Over the past decades, although many emerging effects such as spin-orbit torques (SOTs) switching (9, 10) have improved the performance of MRAM, intrinsic large stray fields and gigahertz dynamics of FM fundamentally limit MRAM from achieving future breakthroughs in high-density integration and high-speed operation (11, 12). Besides, a very small magnetic field may be able to switch the direction of magnetic moment in MRAM, challenging its data storage stability (11, 12). Supreme to FM counterparts, antiferromagnet (AFM) exhibits tremendous potential for building ultradense and ultrafast AFM-RAM with high immunity to magnetic field disturbance thanks to their intrinsic advantages of zero stray field, terahertz dynamics, and compensated moment (13–17). With TMR predicted (18, 19) and found (20, 21) recently in many AFMs, the long-desired electrical-controllable AFM-RAM becomes possible as long as the electrical switching of the AFM order parameter (i.e., the Néel vector) can be achieved.
However, the state-of-art reorientation of Néel vectors are all in-plane 90° or 120° switching by SOTs, which inevitably require multiple writing channels, leading to huge integration difficulties for AFM-RAM (13, 22–24). Similar to FM MRAM, practical AFM-RAM undoubtably needs electrical current along single writing channel to switch the Néel vector back and forth, which generates high and low resistance to represent binary “1” and “0” (11). An ideal solution is the 180° switching of Néel vectors back and forth (n+ ↔ n−) in AFMs by electrical current with opposite polarities along single writing channel. Moreover, 180° switching is truly needed to generate substantial antiferromagnetic TMR (18, 19). Therefore, electrical 180° switching of Néel vector is very important but remains unveiled.
The core difficulty of electrical 180° switching of Néel vector lies in equal energy barriers for n+ → n− and n− → n+, defined by uniaxial magnetic anisotropy. In this case, if SOTs drive the Néel vector to rotate from n+ to n− by overcoming the energy barrier for n+ → n−, it can also continuously drive the Néel vector to rotate from n− to n+ (23), which makes the 180° switching of Néel vector nondeterministic. That is why the more promising 180° switching of Néel vector remains unrealized for a long time since the first discovery of 90° switching of Néel vector in 2016 (13). Note that the 180° switching of CuMnAs by staggered spin-orbit fields is actually due to antiferromagnetic domain wall motion instead of the most desired deterministic 180° switching of Néel vector (25). Besides, the 180° switching of perpendicular cluster magnetic octuple has been achieved in noncollinear AFM (ncAFM) Mn3Sn, but it is fundamentally different from 180° switching of Néel vector as the order parameter in collinear AFM (cAFM) (26, 27). Another important difficulty is to read out the 180° switching of Néel vector. Antiferromagnetic anisotropic magnetoresistance has been widely adopted, but it cannot read out 180° switching (13, 22–24). Harmonic measurements need specific sublattice symmetry breaking, which is not general and is hard to implement (25). Fortunately, the 180° switching of Néel vector may be feasibly read out by the anomalous Hall effect (AHE) (28–31) in some recently found special cAFMs with spin-splitting band structure (32–37), also termed as altermagnets (38, 39). Moreover, their spin-splitting band structure and related properties, such as unconventional spin current generation (40–42) and piezomagnetism (37), can all be manipulated through 180° switching of the Néel vector, which makes it very promising to investigate the electrical 180° switching of Néel vector in spin-splitting AFM.
Here, we achieve deterministic electrical 180° switching of the Néel vector in spin-splitting AFM Mn5Si3 films based on SOTs with asymmetric switching barrier that can be easily generalized to other AFM materials. We prove theoretically and experimentally that 180° switching of the Néel vector indeed tunes the spin-splitting bands and flips the anomalous Hall conductivity (AHC), which can be used as an unconventional electrical readout approach of 180° switching. With these write and read methods ready, we successfully fabricate an AFM device with high and low resistance that realizes robust write and readout cycles, paving the way for the long-desired electrical-controllable AFM-RAM.
RESULTS
Spin splitting manipulated by reversing the Néel vector
We start with introducing basic crystal structure and magnetic phase characteristics of Mn5Si3. The space group of Mn5Si3 at room temperature is P63/mcm with the unit cell composing of four Mna and six Mnb atoms at two inequivalent Wyckoff positions and six Si atoms (Fig. 1A) (43–45). As temperature decreases, Mn5Si3 undergoes a magnetic phase transition from paramagnetic to cAFM and then to ncAFM. In the cAFM phase, the G-type AFM ordering occurs on two-thirds of Mnb atoms, where adjacent magnetic Mnb atoms have opposite spin orientations to form the Néel vector (Fig. 1A and fig. S1 for other perspectives) (43–47). The PT symmetry is broken in cAFM Mn5Si3 thin film (29), and, hence, its energy state at a generic momentum is spin-polarized, i.e., , because of the exchange coupling between itinerant electrons and local magnetic moments. Moreover, the states at mirror symmetry–related momenta must have contrasting spin polarizations to form the C-paired spin-momentum locking, further suppressing spin-flipping (37), which is confirmed by our first-principles calculations as shown in Fig. 1B. These unique features provide possibilities for the existence of AHE in cAFM Mn5Si3 thin film (29).
Next, it is demonstrated that through reversing the Néel vector, we can manipulate spin-resolved and Berry curvature–dependent phenomena, which, in turn, serve as potential readout approach of 180° Néel vector switching in cAFM Mn5Si3. Although spin-orbit coupling has negligible effect on the band structure (fig. S1), it can break the strict SU (2) symmetry and, hence, induce nonzero Berry curvature for spin-splitting bands of cAFM Mn5Si3 with C-paired spin-momentum locking as presented in Fig. 1C. When the Néel vector n is switched 180° from n+ to n–, equivalent to the time-reversal operation, the spin orientations will be reversed (Fig. 1A). Accordingly, both spin splitting (Fig. 1B) and Berry curvature (Fig. 1C) at (kx, ky) will switch their signs at the T-paired momentum (−kx, −ky), as displayed in Fig. 1 (D and E, respectively). Therefore, through 180° Néel vector switching, we can control the spin-splitting bands and Berry curvature and, hence, manipulate all spin-resolved and Berry curvature–dependent phenomena such as AHE. Correspondingly, AHE may serve as a potential readout method of 180° Néel vector switching.
Demonstration of AHE as a readout method
As expected, AHE is observed distinctly in the cAFM temperature range (60 to 230 K) of sputtered Mn5Si3(0001) thin films (fig. S2 and text S1) by Hall resistivity measurements (fig. S3) and polar magneto-optical Kerr effect measurements (fig. S4). Typical hysteresis of Hall resistivity ρyx under out-of-plane magnetic field at 150 K is shown in Fig. 2A, exhibiting a nonvolatile characteristic. Notably, the corresponding hysteresis of magnetization M in Fig. 2B exhibits two abrupt changes around zero field and 4 kOe: The one around zero field (black line) is due to the net moment from unavoidable local defects in the sputtering process, which does not contribute to ρyx near zero field and can be markedly suppressed (fig. S5 and text S2). The other around 4 kOe (blue line) has consistent coercive field Hc with that of the ρyx hysteresis, which can be more clearly illustrated by similar peak locations in field-derivative hysteresis of both ρyx and M around 4 kOe (fig. S3 and text S1). The net moment around 4 kOe, denoted to be m, most likely originates from the relativistic Dzyaloshinskii-Moriya interaction (DMI) due to interfacial symmetry breaking (48).
A natural question is that whether the AHE is simply due to this net moment m. The answer is that AHE comes from Néel vector–dependent Berry curvature instead of m. First, the measured m can be as small as ~1 emu cm−3, corresponding to only 2.5 mμB per magnetic Mnb atom, which can hardly induce the measured large ρyx of 0.17 microhm cm with comparable magnitude to typical FMs (49, 50). m varies for samples of different thicknesses, but ρyx almost remains unchanged, which further rules out that ρyx comes from m (Fig. 2C, fig. S6, and text S2). Moreover, the coercive field of AHE markedly decreases when the width of the Hall bar is reduced below 500 nm, which is quite different from FMs (fig. S7 and text S2). Besides, ρyx has opposite temperature dependence compared to longitudinal resistivity ρxx (inset of fig. S3), which excludes out extrinsic mechanisms and indicates intrinsic mechanism (49, 50). Hence, the experimentally observed AHE should be attributed to the intrinsic Néel vector–dependent Berry curvature as indicated by first-principles calculations in Fig. 1. Note that there is no detectable net moment m in a previous result on cAFM Mn5Si3 thin film, which may be due to different deposition methods and substrates (29). It is worth emphasizing that it is impossible to measure AHE with absolutely zero net moment because both Hall vector and magnetization follow the same symmetry rules (31, 49, 50).
We further performed a systematic theoretical study through atomic spin simulations and first-principles calculations (Materials and Methods) to explicitly explain why the AHE indeed arises from n-dependent Berry curvature instead of the negligible net moment m. We first demonstrate that the 180° switching of m by reversing magnetic field will simultaneously drive the Néel order n to reverse, which provides prerequisites for n-dependent AHE measured by sweeping magnetic field. It is investigated from the responses of m and n to external magnetic field by simulations, considering the realistic crystal and magnetic structure of cAFM Mn5Si3 with all of the parameters extracted from first-principles calculations (figs. S8 and S9 and text S3). The simulated magnetic hysteresis is shown in Fig. 2D, where the coercive field and the magnitude of simulated m match very well with experimental m in Fig. 2B (text S3). Figure 2E presents the simulated trajectories of both m and n when sweeping the out-of-plane magnetic field H from 15 to −15 kOe, where the 180° switching of m from m+ to m− by H drives the 180° switching of n from n+ to n− simultaneously. This process only requires overcoming the magnetic anisotropy energy (MAE) because m and n remain the same chirality with unchanged DMI energy. The reverse of m can also be achieved without reversing n by only switching the tilting direction of cAFM moment, which does not overcome the MAE but changes the chirality between m and n, and thus increases DMI energy. However, this scenario can be excluded for cAFM Mn5Si3 because simply reversing m without reversing n cannot bring about the measured AHE, as we discussed in the previous paragraph. In addition, first-principles calculations also reveal that MAE is one order of magnitude smaller than DMI energy (text S3), further illustrating that the reverse of m and n will appear simultaneously in cAFM Mn5Si3.
With m and n reverse at the same time demonstrated, we move on to demonstrate that it is the reverse of n instead of m that determines the AHE by first-principles calculations. The dependence of AHC σxy on both m and n in cAFM Mn5Si3 is calculated at 150 K, as shown in Fig. 2F. σxy at the Fermi level is calculated to be approximately 400 S m−1, matching the experimentally obtained σxy of around 300 S m−1. Notably, σxy values for n+ with m+ or m− are almost the same, but both are opposite to those for n− with m+ or m−. Thus, the influence of m on σxy is negligible, and the sign reversal of σxy is determined by the 180° switching of n, consistent with aforementioned experimental results. Moreover, σxy at the Fermi level for n along any direction is calculated and presented as a spherical image in Fig. 2G. It turns out that σxy always reverses once n is 180° switched, irrespective of any specific direction n aligning. As a result, both our experimental and theoretical results reveal that the sign of σxy is uniquely determined by the direction of n, and thus, the AHE can serve as a reliable readout method for 180° switching of n in cAFM Mn5Si3.
Asymmetric energy barrier mechanism for electrical 180° switching of Néel vector
After solving the problem of reading the 180° switching of Néel vector, the remaining question is how to realize the deterministic electrical 180° switching of Néel vector, i.e., the information write process. There are three basic requirements in general: (i) Both n+ and n− should be stable with minimum energy to realize binary information states “0” and “1”; (ii) n needs to be driven into motion electrically; (iii) transition from n+ to n− and its time-reversal counterpart from n− to n+ must have unequal energy barriers to achieve the deterministic switching. The requirement (i) is naturally satisfied in cAFMs since both n+ and n− are stable along the easy axis due to MAE, and thus, we will focus on how to fulfill requirements (ii) and (iii) in the following.
According to the definition of m and n, they are always perpendicular to each other in canted AFMs, and a strong DMI vector d enforces a fixed chirality between m and n as demonstrated above in cAFM Mn5Si3. Without losing the generality, we start by assuming d to be perpendicular to the plane of m and n, which forms an orthogonal coordinate system as presented in Fig. 3A. To achieve requirement (ii), n needs to be driven into rotation along the n × p direction through the AFM exchange torque induced by the spin polarization p, which can be usually generated from the spin Hall effect of heavy metals such as Pt adjacent to Mn5Si3. We first consider p perpendicular to the plane of m and n, i.e., along d, and n will roughly rotate circularly in this plane (Fig. 3A) for this scenario. However, because of equal transition barriers of n+ to n− and n− to n+ (H = 0 case of Fig. 3B), n will rotate continuously driven by p and, lastly, nondeterministically relaxes to n+ or n− after p is withdrawn. To address this issue, we propose to construct asymmetric switching barriers through a fixed magnetic field H acting on nonzero m as shown in Fig. 3B (H > 0 case), and the asymmetry will be maximized when H is perpendicular to m as shown in Fig. 3A. In this case, through choosing a positive p with suitable magnitude, n can rotate from n+ to n− through the lower barrier but cannot further climb over the higher barrier from n− to n+ and thus stays at the intermediate state n'–. When p is withdrawn, n will naturally stabilize as n−. Similarly, by applying another p with opposite direction, n can switch from n− back to n+ through the same lower energy barrier. Thus, the desired deterministic 180° switching between n+ and n− can be electrically realized by damping-like SOT with a tiny external magnetic field.
This SOT mechanism with asymmetric energy barrier for 180° switching between n+ and n− can further be confirmed explicitly by atomic spin simulations. Under a positive p, n will rotate clockwise accompanied by m moving toward the hemisphere favored by positive H and then settles at an intermediate state n'– (Fig. 3C). After SOT is turned off, n will naturally relax to the local energy minimum n− to achieve the deterministic 180° switching from n+ to n− (fig. S10 for other perspectives). Similarly, n will switch back to n+ from n− deterministically with a reversing p through anticlockwise rotation (Fig. 3D). Reversing the direction of H also gives rise to deterministic 180° switching with reversing switching polarity (fig. S11). The same conclusion can be drawn through similar mechanisms for arbitrary orientations of p and H revealed by atomic spin simulations, even when d is not perpendicular to m and n (text S4 and figs. S12 and S13). Note that this mechanism is not restricted to Mn5Si3 but can be easily generalized to other AFMs with tiny intrinsic net moment.
Experimental demonstration of electrical 180° switching of the Néel vector
We carried out SOT switching experiments on Mn5Si3(0001)/Pt crossbars of 6 μm by 4 μm (Fig. 4A) based on the mechanism above, where Pt serves as the spin source. Electrical pulses of different magnitude were introduced into the write channel WI+-WI− with an assistant field H along the same direction, and the Hall voltage was collected simultaneously along RI+-RI− to read the direction of n through AHE. As expected, 180° switching of n in cAFM Mn5Si3 are observed under various temperatures (fig. S14), and the results at 180 K are representatively presented in Fig. 4B. The largest switching ratio reaches 41%, and such an incomplete switching is most likely due to the existence of microscopical multidomain and local defects pinning (fig. S15). Opposite switching polarity can be observed if the polarity of current pulse or H is reversed, consistent with aforementioned theoretical results and matching basic symmetry requirements for deterministic SOT switching (9, 10). No switching can be observed with a high H of 8 kOe, which confines the direction of m and thus also confines the direction of n, further verifying that the change of Hall voltage is due to the switching of n instead of thermal artifacts. Moreover, AHE can also be observed in the temperature range of ncAFM Mn5Si3 (below 60 K), but electrical 180° switching cannot be achieved, further demonstrating that it is n that can be switched electrically (fig. S16).
We now show that the 180° SOT switching of n in cAFM Mn5Si3 is fundamentally different from FMs and ncAFMs such as Mn3Sn (26, 27, 51–55). In FMs, the assistant field H can effectively assist the SOT to overcome MAE for switching, as characterized by a linear decrease of the critical switching current density Jc with increasing H (44). In contrast, for cAFM Mn5Si3, H only eliminates the degeneracy of energy barriers between n+ to n− and n− to n+ but barely helps n to overcome MAE. That is because it is the large AFM exchange energy that determines the tilting of n toward p as a prerequisite to obtain nonzero AFM exchange torque for overcoming MAE, where AFM exchange energy is much larger than the Zeeman energy induced by H acting on the negligible m. Thus, Jc should be very insensitive to H for cAFM Mn5Si3. As presented in Fig. 4C and fig. S17, we carried out SOT switching measurements under different assistant field H, and the experimentally extracted Jc (text S5) remains almost unchanged with the increase of H for cAFM Mn5Si3. Besides, unlike uniaxial Néel vector n as order parameter perpendicular to m in cAFM Mn5Si3, cluster magnetic octuple as the order parameter with complex magnetic anisotropy in ncAFM Mn3Sn is parallel to its intrinsic m, leading to different dynamic behaviors under p and H of specific symmetry (26, 27, 51–53).
To investigate the efficiency for 180° switching of n in cAFM Mn5Si3, we calculate the ratio of AHE coercivity to critical switching current density μ0Hc/Jc as a figure of merit (52). For the averaged Jc of 49 MA cm−2 at 180 K (fig. S15), it turns out to be about 5 mT cm2 MA−1, which is five times higher than traditional ferromagnetic systems such as Pt/Co/AlOx of approximately 1 mT cm2 MA−1 (52, 56), indicating that the 180° SOT switching of n in cAFM Mn5Si3 shows high efficiency. Moreover, we test the endurance for 180° switching of n by applying electrical pulses with opposite directions alternately. The Hall voltage can be switched between high and low values corresponding to information states between “0” and “1” (Fig. 4D), which is reversible and reproducible. The device does not show degradation after 500 cycles, manifesting its robustness. As a result, spin-splitting Mn5Si3 shows unique advantages to construct electrical-controllable AFM-RAM with low power consumption, high reliability, and exceptional abundance in addition to its intrinsic high density and high speed.
DISCUSSION
We achieve electrical 180° switching of the Néel vector in spin-splitting cAFM Mn5Si3. The demonstration of electrical 180° switching of the Néel vector involves two steps. One is to prove that it is the 180° switching of the Néel vector instead of net moment that tunes the spin-splitting band structure and brings about opposite AHC, which makes AHE as a unique readout method of 180° switching of Néel vector. It is supported by both systematic theoretical calculations and several control experiments including thickness-dependent and channel width–dependent measurements. The other is to design SOT switching mechanism with asymmetric energy barriers between n+ and n−. It is experimentally implemented and confirmed by realizing current and assistant field polarity–dependent 180° Néel vector switching and further verified by assistant field magnitude-dependent measurements and nonswitching of ncAFM Mn5Si3. On the basis of electrical 180° switching and electrical readout of the Néel vector in Mn5Si3, electrical-controllable AFM devices with “0” and “1” resistance states were fabricated, which show high efficiency and high reproducibility. Our results establish the groundwork for practical applications of spin-splitting AFMs in next-generation information technology and serve as the basis of exploring other intriguing properties in spin-splitting AFMs arising from interactions among different degrees of freedom such as spin, valley, and charge.
MATERIALS AND METHODS
Sample preparation
Six nanometers of Mn5Si3(0001) films was grown on Al2O3(0001) substrates by cosputtering of Mn and Si at 465°C with a rate of 0.4 Å s−1 under a base pressure below 5 × 10−8 torr, followed by annealing at 600°C for 2 hours. After that, 10-nm Pt was deposited in situ by magnetron sputtering at room temperature. Crossbars (6 μm by 4 μm) for SOT switching measurements were patterned using optical lithography combined with Ar ion milling. Then, Cr(10 nm)/Cu(70 nm) electrode was electron beam (e-beam)–evaporated for wire bonding. For room temperature and low-temperature x-ray diffraction (XRD) measurements, Mn5Si3(0001) films (80 nm) were prepared to obtain diffraction peaks with substantial intensity for precisely determining the lattice parameter. Besides, Mn5Si3(0001) films (130 nm) were grown for control measurements. For size-dependent measurements of the AHE, e-beam lithography was used to fabricate crossbars with width varying from 100 nm to 5 μm. All of the samples were kept in a glove box with O2 and H2O < 0.01 parts per million to prevent degradation or oxidation.
Characterizations
Cross-sectional transmission electron microscopy (TEM) images for Mn5Si3(6 nm)/Pt(10 nm) were collected at room temperature using a commercial TEM system (JEM-2100F). Magnetic hysteresis curves were collected in commercial superconducting quantum interference device (SQUID; Quantum Design), where the diamagnetic contribution of Al2O3(0001) substrate was subtracted. Magnetization-temperature curves were also collected from SQUID without subtracting diamagnetic contribution of Al2O3(0001) substrate. θ-2θ XRD measurement at room temperature was carried out at Rigaku SmartLab. In situ θ-2θ XRD at low temperatures was performed at BL02U2 Beamline from Shanghai Synchrotron Radiation Facility. Atomic force microscopy images were collected at Bruker Dimension FastScan.
Transport measurements
ρxx and ρyx of Mn5Si3 films were measured using a standard four-terminal method in commercial Physical Property Measurement System (Quantum Design). Ordinary Hall resistivity that is linear to magnetic field was subtracted from ρyx. For SOT switching measurements, writing pulses (1 ms) within the range of 46 mA were added in the WI+-WI− channel with an assistant field H applied along the WI+-WI− channel, followed by waiting for 10 s before collecting the Hall voltage in the RV+-RV− channel by a nanovoltmeter with a base current of 2 mA (Fig. 4A).
Time-resolved reflectivity measurements
Time-resolved reflectivity measurements were carried out by a Ti:sapphire oscillator (center wavelength of 800 nm, repetition rate of 80 MHz, and pulse width of 20 fs) in a typical wavelength-degenerate pump-probe setup. Both the pump and probe beams were linearly polarized in a cross-polarization configuration to block the pump scattering. The pump beam was focused to 30 μm on the sample with a fixed fluence at 35.4 μJ/cm2. The focal spot size and fluence of the probe beam were about 2 and 40 times smaller than those of the pump beam, respectively. The intensity of the pump beam was modulated by an optical chopper at 3 kHz to facilitate lock-in detection. The reflectivity signal was detected by a balanced detector to suppress laser power fluctuations. The sample was kept in a cryostat under vacuum of better than 1 × 10−3 mbar during experiments.
Magneto-optical Kerr effect measurements
The polar magneto-optical Kerr effect measurements were carried out using a power-stabilized 633-nm HeNe laser. After transmitting through a linear polarizer, the light (25 μW) was focused to an ~1-μm spot on the sample by a 40× reflective objective at normal incidence to avoid the large backgrounds that occur when a typical lens is used. The sample was mounted in the vacuum chamber of an optical superconducting magnet system with the magnetic field applied perpendicular to the sample plane (Faraday geometry). The reflected beam was modulated at ~50 kHz by a photoelastic modulator, split by a Wollaston prism and detected using a balanced photodiode with a standard optical bridge arrangement. The resulting 50- and 100-kHz modulations detected by lock-in amplifiers then correspond to the ellipticity and rotation angle of the beam, respectively. We additionally modulated the intensity of the beam with a lower-frequency (~1319 Hz) chopper to measure the dc signal for normalization using a third lock-in amplifier. The background from the optic window and a linear background were subtracted.
First-principles calculations
First-principles calculations were performed on the basis of density functional theory (DFT) (57) as implemented in Vienna ab initio simulation package (58). Exchange-correlation interaction was described by the Perdew-Burke-Ernzerhof parametrization of generalized gradient approximation (59). We also used Perdew-Wang (60) and local density approximation exchange-correlation functional, where similar results were obtained. Structures were relaxed until the force on each atom is less than 0.01 eV/Å. The cutoff energy and electronic iteration convergence criterion were set to 400 and 10−5 eV, respectively. To model the Brillouin zone, a Monkhorst-Pack (MP) k-grid mesh (61) of 7 × 7 × 9 was used. For MAE calculations, self-consistent charge density was read to get a converged result. Four-states method was used to evaluate the strength of DMI. Berry curvature and AHC were calculated on the basis of ab initio tight-binding models with all parameters extracted from first-principles calculations using the FPLO software (62).
Atomic spin simulations of magnetic hysteresis
We carried out atomic spin simulations on VAMPIRE software (63), which can define specific lattice parameters a and c consistent with our hexagonal cAFM phase Mn5Si3 thin film of 6.902 and 4.795 Å, respectively. Discretized 50 × 50 × 10 units cells in the x-y-z direction were used to model the Mn5Si3 thin film. The Hamiltonian of the system at 0 K can be written as following
Here, i and j denote unit cell and p and q represent sublattice (p, q = 1, 2, 3, 4). Htotal is the total Hamiltonian, and mip (mjq) is the unit magnetic moment at sublattice Mnp (Mnq) of unit cell i (j) with the magnitude of 2.4 μB. Jip,jq, dip,jq, and K are the exchange interaction constant, the DMI vector, and the uniaxial anisotropy constant, respectively. HZ is the Zeeman energy under magnetic field, equaling , where μ0, ms, and H are the vacuum permeability, the saturation magnetization of magnetic Mn atom, and the external field, respectively. k is the direction of easy axis, which was set to be (1,1,1) direction in the cartesian coordinate system with the uniaxial anisotropy constant K of 0.1 meV from DFT estimations (text S3). Four exchange interactions were considered, namely J1 = −12.23 meV, J2 = −2.16 meV, J3 = 3.98 meV, and J4 = 2.89 meV (fig. S8A) with magnitude consistent with former DFT calculation results (64, 65). Note that J4 interaction was set to be ferromagnetic to eliminate the 𝓉𝒯 symmetry for consistency with experimentally measured nonzero σxy (text S3). The DMI vector was set along x axis with a magnitude of 1.35 meV from DFT calculations and the experimentally measured out-of-plane AHE-related net moment m (text S2). These settings of basic material parameters give a simulated magnetic hysteresis that is analogous with SQUID-derived m under 150 K (Fig. 2C and text S2), illustrating that these parameters are suitable for modeling the magnetic switching behavior of cAFM Mn5Si3. Specifically speaking, the Landau-lifshitz-Gilbert (LLG) equation for each magnetic Mn atom is written as
where Heff,ip is the effective magnetic field acting on the magnetic Mn atom at sublattice p of unit cell i. , where Hth,ip is the Gaussian stochastic magnetic field to describe the effect of finite temperature (66). γ and α are the gyromagnetic ratio of the electron and the Gilbert damping coefficient, equaling 176 s−1 T−1 and 0.1, respectively. A fine iteration time step Δt of 0.1 fs was used to precisely capture the dynamics of atomic magnetic moments. First, the system was initialized with random orientations of atomic magnetic moments and then relaxed into a perfect G-type AFM ordering at 0 K (fig. S8B). After applying negative magnetic field, each atomic magnetic moment can be switched by 180° (fig. S8C). The simulated m and n for each unit cell should be defined as m = (m1 + m2 + m3 + m4)/4 and n = (m1 − m2 + m3 − m4)/4, which further equaling (m1 + m2)/2 and (m1 − m2)/2 because m1 = m3 and m2 = m4 due to J3. Thus, the 180° switching of each atomic magnetic moment is consistent with the 180° switching of m and n. Then, the hysteresis was simulated at 150 K (Fig. 2D), and 180° switching of m and n can also be observed (Fig. 2E), where m and n on all of unit cells were averaged.
Atomic spin simulations of SOT switching
To understand the mechanism for SOT 180° switching of the Néel vector n of cAFM Mn5Si3, we consider the damping-like effective field HDL of the spin polarization p generated by the spin Hall effect of Pt. Then, the LLG equation under SOT can be written as
where HDL,ip is the damping-like effective field of SOT acting on the magnetic Mn atom at sublattice p (p = 1, 2, 3, 4) of unit cell i. p is the unit vector for the spin polarization. ξDL is the magnitude of damping-like SOT, which has linear relationship with the magnitude of HDL,ip and current density J (text S3). Note that for the simulation of SOT switching, 5 × 5 × 5 unit cells were used to save calculation time, which gives consistent results with that of 50 × 50 × 10 unit cells. In analogous to the simulations for magnetic hysteresis, the system was initialized with random orientations of atomic magnetic moments and relaxed into G-type AFM ordering before adding a 0.1-ns-long electrical pulse with transverse p under assistant field H. The magnitude of assistant field H was fixed at 10 kOe. This magnitude of H is reasonable because the simulations of SOT switching were carried out at 0 K, and the coercive field of simulated magnetic hysteresis at 0 K is around 150 kOe. It means that H of no more than 10% of the coercive field can assist the deterministic 180° switching of n, consistent with experimental results. m and n on all of unit cells were averaged to plot trajectories of m and n under p. The settings of pulse lengths and random seeds do not influence simulation results, proving their reliability (text S6). The influence of field-like SOTs was also investigated by adding field-like effective field HFL as a term of ∣γ∣mip × HFL,ip. It turns out that HFL has a negative influence on SOT switching, which increases the critical switching current density (text S7).
Acknowledgments
We acknowledge BL02U2 of Shanghai Synchrotron Radiation Facility and the support from Beijing Innovation Center for Future Chip (ICFC), Tsinghua University. Some devices were fabricated via an Ultraviolet Maskless Lithography machine (model: UV Litho-ACA, TuoTuo Technology).
Funding: This work was supported by the National Key R&D Program of China (grant nos. 2022YFA1402603, 2021YFB3601301, 2021YFA1401500, 2020YFA0308800, and 2021YFA1400100), National Natural Science Foundation of China (grant nos. 52225106, 12241404, 12074212, and 12022416), Hong Kong Research Grants Council (grant nos. 16304523, 16303821, 16306722), and Natural Science Foundation of Beijing, China (grant no. JQ20010).
Author contributions: Conceptualization: L.H., X.F., X. Cheng, J.D., L.L., Yichi Zhang, W.Z., L.Y., C.S., J.L., and F.P. Data curation: X.F., X. Cheng, L.L., Y.L., Y.Zho., Q.W., C.S., J.L., and F.P. Formal analysis: L.H., X.F., R.P., X. Cheng, J.D., L.L., Y.L., Yang Zhang, C.S., J.L., and F.P. Funding acquisition: X. Cheng, L.Y., C.S., J.L., and F.P. Investigation: L.H., X.F., J.D., L.L., Y.L., S.L., Yang Zhang, C.S., J.L., and F.P. Methodology: L.H., X.F., X. Cheng, J.D., L.L., C.C., L.Y., C.S., J.L., and F. P. Project administration: L.H., L.Y., C.S., J.L., and F.P. Resources: J.D., Y.Zho., Q.W., L.Y., C.S., J.L., and F.P. Software: L.H., X.F., X. Cheng, Yang Zhang, C.S., J.L., and F.P. Supervision: L.H., Q.W., L.Y., C.S., J.L., and F.P. Validation: L.H., X.F., R.P., J.D., L.L., Y.L., Q.W., X. Chen, L.Y., Yang Zhang, C.S., J.L., and F.P. Visualization: L.H., X.F., X. Cheng, J.D., Y.L., W.Z., H.B., Y.Zho., C.S., J.L., and F.P. Writing—original draft: L.H., X.F., X. Cheng, J.D., Y.L., Y.L., Y.Zho., C.C., C.S., J.L., and F.P. Writing—review and editing: L.H., X.F., X. Cheng, J.D., L.L., Y.L., W.Z., Y.Zho., S.L., L.Y., C.S., J.L., and F.P.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.
Supplementary Materials
This PDF file includes:
Other Supplementary Material for this manuscript includes the following:
REFERENCES AND NOTES
- 1.Wolf S. A., Awschalom D. D., Buhrman R. A., Daughton J. M., von Molnár S., Roukes M. L., Chtchelkanova A. Y., Treger D. M., Spintronics: A spin-based electronics vision for the future. Science 294, 1488–1495 (2001). [DOI] [PubMed] [Google Scholar]
- 2.Bhatti S., Sbiaa R., Hirohata A., Ohno H., Fukami S., Piramanayagam S. N., Spintronics based random access memory: A review. Mater. Today 20, 530–548 (2017). [Google Scholar]
- 3.Song C., Zhang R. Q., Liao L. Y., Zhou Y. J., Zhou X. F., Chen R. Y., You Y. F., Chen X. Z., Pan F., Spin-orbit torques: Materials, mechanisms, performances, and potential applications. Prog. Mater. Sci. 118, 100761 (2021). [Google Scholar]
- 4.Julliere M., Tunneling between ferromagnetic films. Phys. Lett. 54, 225–226 (1975). [Google Scholar]
- 5.Miyazaki T., Tezuka N., Giant magnetic tunneling effect in Fe/Al2O3/Fe junction. J. Magn. Magn. Mater. 139, L231–L234 (1995). [Google Scholar]
- 6.Moodera J. S., Kinder L. R., Wong T. M., Meservey R., Large magnetoresistance at room temperature in ferromagnetic thin film tunnel junctions. Phys. Rev. Lett. 74, 3273–3276 (1995). [DOI] [PubMed] [Google Scholar]
- 7.Slonczewski J. C., Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater. 159, L1–L7 (1996). [Google Scholar]
- 8.Berger L., Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev. B 54, 9353–9358 (1996). [DOI] [PubMed] [Google Scholar]
- 9.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 476, 189–193 (2011). [DOI] [PubMed] [Google Scholar]
- 10.Liu L., Pai C.-F., Li Y., Tseng H. W., Ralph D. C., Buhrman R. A., Spin-torque switching with the giant spin Hall effect of tantalum. Science 336, 555–558 (2012). [DOI] [PubMed] [Google Scholar]
- 11.Jungwirth T., Marti X., Wadley P., Wunderlich J., Antiferromagnetic spintronics. Nat. Nanotechnol. 11, 231–241 (2016). [DOI] [PubMed] [Google Scholar]
- 12.Baltz V., Manchon A., Tsoi M., Moriyama T., Ono T., Tserkovnyak Y., Antiferromagnetic spintronics. Rev. Mod. Phys. 90, (2018). [Google Scholar]
- 13.Wadley P., Howells B., Zelezny J., Andrews C., Hills V., Campion R. P., Novak V., Olejnik K., Maccherozzi F., Dhesi S. S., Martin S. Y., Wagner T., Wunderlich J., Freimuth F., Mokrousov Y., Kunes J., Chauhan J. S., Grzybowski M. J., Rushforth A. W., Edmonds K. W., Gallagher B. L., Jungwirth T., Electrical switching of an antiferromagnet. Science 351, 587–590 (2016). [DOI] [PubMed] [Google Scholar]
- 14.Vaidya P., Morley S. A., van Tol J., Liu Y., Cheng R., Brataas A., Lederman D., del Barco E., Subterahertz spin pumping from an insulating antiferromagnet. Science 368, 160–165 (2020). [DOI] [PubMed] [Google Scholar]
- 15.Park B. G., Wunderlich J., Martí X., Holý V., Kurosaki Y., Yamada M., Yamamoto H., Nishide A., Hayakawa J., Takahashi H., Shick A. B., Jungwirth T., A spin-valve-like magnetoresistance of an antiferromagnet-based tunnel junction. Nat. Mater. 10, 347–351 (2011). [DOI] [PubMed] [Google Scholar]
- 16.Wang Y. Y., Song C., Cui B., Wang G. Y., Zeng F., Pan F., Room-temperature perpendicular exchange coupling and tunneling anisotropic magnetoresistance in an antiferromagnet-based tunnel junction. Phys. Rev. Lett. 109, 137201 (2012). [DOI] [PubMed] [Google Scholar]
- 17.Marti X., Fina I., Frontera C., Liu J., Wadley P., He Q., Paull R. J., Clarkson J. D., Kudrnovský J., Turek I., Kuneš J., Yi D., Chu J. H., Nelson C. T., You L., Arenholz E., Salahuddin S., Fontcuberta J., Jungwirth T., Ramesh R., Room-temperature antiferromagnetic memory resistor. Nat. Mater. 13, 367–374 (2014). [DOI] [PubMed] [Google Scholar]
- 18.Shao D.-F., Zhang S.-H., Li M., Eom C.-B., Tsymbal E. Y., Spin-neutral currents for spintronics. Nat. Commun. 12, 7061 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Šmejkal L., Hellenes A. B., González-Hernández R., Sinova J., Jungwirth T., Giant and tunneling magnetoresistance in unconventional collinear antiferromagnets with nonrelativistic spin-momentum coupling. Phys. Rev. X 12, 011028 (2022). [Google Scholar]
- 20.Qin P., Yan H., Wang X., Chen H., Meng Z., Dong J., Zhu M., Cai J., Feng Z., Zhou X., Liu L., Zhang T., Zeng Z., Zhang J., Jiang C., Liu Z., Room-temperature magnetoresistance in an all-antiferromagnetic tunnel junction. Nature 613, 485–489 (2023). [DOI] [PubMed] [Google Scholar]
- 21.Chen X., Higo T., Tanaka K., Nomoto T., Tsai H., Idzuchi H., Shiga M., Sakamoto S., Ando R., Kosaki H., Matsuo T., Nishio-Hamane D., Arita R., Miwa S., Nakatsuji S., Octupole-driven magnetoresistance in an antiferromagnetic tunnel junction. Nature 613, 490–495 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Bodnar S. Y., Smejkal L., Turek I., Jungwirth T., Gomonay O., Sinova J., Sapozhnik A. A., Elmers H. J., Klaui M., Jourdan M., Writing and reading antiferromagnetic Mn2Au by Néel spin-orbit torques and large anisotropic magnetoresistance. Nat. Commun. 9, 348 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Chen X. Z., Zarzuela R., Zhang J., Song C., Zhou X. F., Shi G. Y., Li F., Zhou H. A., Jiang W. J., Pan F., Tserkovnyak Y., Antidamping-torque-induced switching in biaxial antiferromagnetic insulators. Phys. Rev. Lett. 120, 207204 (2018). [DOI] [PubMed] [Google Scholar]
- 24.Cheng Y., Yu S., Zhu M., Hwang J., Yang F., Electrical switching of tristate antiferromagnetic Néel order inα−Fe2O3 epitaxial films. Phys. Rev. Lett. 124, 027202 (2020). [DOI] [PubMed] [Google Scholar]
- 25.Godinho J., Reichlova H., Kriegner D., Novak V., Olejnik K., Kaspar Z., Soban Z., Wadley P., Campion R. P., Otxoa R. M., Roy P. E., Zelezny J., Jungwirth T., Wunderlich J., Electrically induced and detected Néel vector reversal in a collinear antiferromagnet. Nat. Commun. 9, 4686 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Higo T., Kondou K., Nomoto T., Shiga M., Sakamoto S., Chen X., Nishio-Hamane D., Arita R., Otani Y., Miwa S., Nakatsuji S., Perpendicular full switching of chiral antiferromagnetic order by current. Nature 607, 474–479 (2022). [DOI] [PubMed] [Google Scholar]
- 27.Yoon J.-Y., Zhang P., Chou C.-T., Takeuchi Y., Uchimura T., Hou J. T., Han J., Kanai S., Ohno H., Fukami S., Liu L., Handedness anomaly in a non-collinear antiferromagnet under spin–orbit torque. Nat. Mater. 22, 1106–1113 (2023). [DOI] [PubMed] [Google Scholar]
- 28.Feng Z., Zhou X., Šmejkal L., Wu L., Zhu Z., Guo H., González-Hernández R., Wang X., Yan H., Qin P., Zhang X., Wu H., Chen H., Meng Z., Liu L., Xia Z., Sinova J., Jungwirth T., Liu Z., An anomalous Hall effect in altermagnetic ruthenium dioxide. Nat. Electron. 5, 735–743 (2022). [Google Scholar]
- 29.Kounta I., Reichlova H., Kriegner D., Lopes Seeger R., Bad'ura A., Leiviska M., Boussadi A., Heresanu V., Bertaina S., Petit M., Schmoranzerova E., Smejkal L., Sinova J., Jungwirth T., Baltz V., Goennenwein S. T. B., Michez L., Competitive actions of MnSi in the epitaxial growth of Mn5Si3 thin films on Si(111). Phys. Rev. Mater. 7, 024416 (2023). [Google Scholar]
- 30.Gonzalez Betancourt R. D., Zubac J., Gonzalez-Hernandez R., Geishendorf K., Soban Z., Springholz G., Olejnik K., Smejkal L., Sinova J., Jungwirth T., Goennenwein S. T. B., Thomas A., Reichlova H., Zelezny J., Kriegner D., Spontaneous anomalous Hall effect arising from an unconventional compensated magnetic phase in a semiconductor. Phys. Rev. Lett. 130, 036702 (2023). [DOI] [PubMed] [Google Scholar]
- 31.Šmejkal L., MacDonald A. H., Sinova J., Nakatsuji S., Jungwirth T., Anomalous Hall antiferromagnets. Nat. Rev. Mater. 7, 482–496 (2022). [Google Scholar]
- 32.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). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Naka M., Hayami S., Kusunose H., Yanagi Y., Motome Y., Seo H., Spin current generation in organic antiferromagnets. Nat. Commun. 10, 4305 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Ahn K.-H., Hariki A., Lee K.-W., Kuneš J., Antiferromagnetism inRuO2asd-wave pomeranchuk instability. Phys. Rev. B 99, 184432 (2019). [Google Scholar]
- 35.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). [Google Scholar]
- 36.Gonzalez-Hernandez R., Smejkal L., Vyborny K., Yahagi Y., Sinova J., Jungwirth T., Zelezny J., Efficient electrical spin splitter based on nonrelativistic collinear antiferromagnetism. Phys. Rev. Lett. 126, 127701 (2021). [DOI] [PubMed] [Google Scholar]
- 37.Ma H. Y., Hu M., Li N., Liu J., Yao W., Jia J. F., Liu J., Multifunctional antiferromagnetic materials with giant piezomagnetism and noncollinear spin current. Nat. Commun. 12, 2846 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.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). [Google Scholar]
- 39.Šmejkal L., Sinova J., Jungwirth T., Emerging research landscape of altermagnetism. Phys. Rev. X 12, 040501 (2022). [Google Scholar]
- 40.Bose A., Schreiber N. J., Jain R., Shao D.-F., Nair H. P., Sun J., Zhang X. S., Muller D. A., Tsymbal E. Y., Schlom D. G., Ralph D. C., Tilted spin current generated by the collinear antiferromagnet ruthenium dioxide. Nat. Electron. 5, 267–274 (2022). [Google Scholar]
- 41.Bai H., Han L., Feng X. Y., Zhou Y. J., Su R. X., Wang Q., Liao L. Y., Zhu W. X., Chen X. Z., Pan F., Fan X. L., Song C., Observation of spin splitting torque in a collinear antiferromagnetRuO2. Phys. Rev. Lett. 128, 197202 (2022). [DOI] [PubMed] [Google Scholar]
- 42.Karube S., Tanaka T., Sugawara D., Kadoguchi N., Kohda M., Nitta J., Observation of spin-splitter torque in collinear antiferromagneticRuO2. Phys. Rev. Lett. 129, 137201 (2022). [DOI] [PubMed] [Google Scholar]
- 43.Brown P. J., Forsyth J. B., Nunez V., Tasset F., The low-temperature antiferromagnetic structure of Mn5Si3revised in the light of neutron polarimetry. J. Phys. Condens. Matter 4, 10025–10036 (1992). [Google Scholar]
- 44.Brown P., Forsyth J., Antiferromagnetism in Mn5Si3: The magnetic structure of the AF2 phase at 70 K. J. Phys. Condens. Matter 7, 7619–7628 (1995). [Google Scholar]
- 45.Gottschilch M., Gourdon O., Persson J., de la Cruz C., Petricek V., Brueckel T., Study of the antiferromagnetism of Mn5Si3: An inverse magnetocaloric effect material. J. Mater. Chem. 22, 15275–15284 (2012). [Google Scholar]
- 46.Surgers C., Fischer G., Winkel P., Lohneysen H. V., Large topological Hall effect in the non-collinear phase of an antiferromagnet. Nat. Commun. 5, 3400 (2014). [DOI] [PubMed] [Google Scholar]
- 47.Biniskos N., Schmalzl K., Raymond S., Petit S., Steffens P., Persson J., Bruckel T., Spin fluctuations drive the inverse magnetocaloric effect in Mn5Si3. Phys. Rev. Lett. 120, 257205 (2018). [DOI] [PubMed] [Google Scholar]
- 48.Dzyaloshinsky I., A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics. J. Phys. Chem. Solid 4, 241–255 (1958). [Google Scholar]
- 49.Nagaosa N., Sinova J., Onoda S., MacDonald A. H., Ong N. P., Anomalous Hall effect. Rev. Mod. Phys. 82, 1539–1592 (2010). [Google Scholar]
- 50.Xiao D., Chang M.-C., Niu Q., Berry phase effects on electronic properties. Rev. Mod. Phys. 82, 1959–2007 (2010). [Google Scholar]
- 51.Tsai H., Higo T., Kondou K., Nomoto T., Sakai A., Kobayashi A., Nakano T., Yakushiji K., Arita R., Miwa S., Otani Y., Nakatsuji S., Electrical manipulation of a topological antiferromagnetic state. Nature 580, 608–613 (2020). [DOI] [PubMed] [Google Scholar]
- 52.Takeuchi Y., Yamane Y., Yoon J. Y., Itoh R., Jinnai B., Kanai S., Ieda J., Fukami S., Ohno H., Chiral-spin rotation of non-collinear antiferromagnet by spin-orbit torque. Nat. Mater. 20, 1364–1370 (2021). [DOI] [PubMed] [Google Scholar]
- 53.Pal B., Hazra B. K., Göbel B., Jeon J.-C., Pandeya A. K., Chakraborty A., Busch O., Srivastava A. K., Deniz H., Taylor J. M., Meyerheim H., Mertig I., Yang S.-H., Parkin S. S. P., Setting of the magnetic structure of chiral kagome antiferromagnets by a seeded spin-orbit torque. Sci. Adv. 8, eabo5930 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Nayak A. K., Fischer J. E., Sun Y., Yan B., Karel J., Komarek A. C., Shekhar C., Kumar N., Schnelle W., Kübler J., Felser C., Parkin S. S. P., Large anomalous Hall effect driven by a nonvanishing Berry curvature in the noncolinear antiferromagnet Mn3Ge. Sci. Adv. 2, e1501870 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Železný J., Zhang Y., Felser C., Yan B., Spin-polarized current in noncollinear antiferromagnets. Phys. Rev. Lett. 119, 187204 (2017). [DOI] [PubMed] [Google Scholar]
- 56.Liu L., Lee O. J., Gudmundsen T. J., Ralph D. C., Buhrman R. A., Current-induced switching of perpendicularly magnetized magnetic layers using spin torque from the spin Hall effect. Phys. Rev. Lett. 109, 096602 (2012). [DOI] [PubMed] [Google Scholar]
- 57.Kohn W., Sham L. J., Self-consistent equations including exchange and correlation effects. Phys. Rev. 140, A1133–A1138 (1965). [Google Scholar]
- 58.Kresse G., Furthmüller J., Efficient iterative schemes forab initiototal-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
- 59.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
- 60.Perdew J. P., Chevary J. A., Vosko S. H., Jackson K. A., Pederson M. R., Singh D. J., Fiolhais C., Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 46, 6671–6687 (1992). [DOI] [PubMed] [Google Scholar]
- 61.Monkhorst H. J., Pack J. D., Special points for Brillouin-zone integrations. Phys. Rev. B 13, 5188–5192 (1976). [Google Scholar]
- 62.Koepernik K., Eschrig H., Full-potential nonorthogonal local-orbital minimum-basis band-structure scheme. Phys. Rev. B 59, 1743–1757 (1999). [Google Scholar]
- 63.Evans R. F. L., Fan W. J., Chureemart P., Ostler T. A., Ellis M. O. A., Chantrell R. W., Atomistic spin model simulations of magnetic nanomaterials. J. Phys. Condens. Matter 26, 103202 (2014). [DOI] [PubMed] [Google Scholar]
- 64.dos Santos F. J., Biniskos N., Raymond S., Schmalzl K., dos Santos Dias M., Steffens P., Persson J., Blügel S., Lounis S., Brückel T., Spin waves in the collinear antiferromagnetic phase ofMn5Si3. Phys. Rev. B 103, 024407 (2021). [Google Scholar]
- 65.Biniskos N., dos Santos F. J., Schmalzl K., Raymond S., dos Santos Dias M., Persson J., Marzari N., Blügel S., Lounis S., Brückel T., Complex magnetic structure and spin waves of the noncollinear antiferromagnet Mn5Si3. Phys. Rev. B 105, 104404 (2022). [Google Scholar]
- 66.Evans R. F. L., Atxitia U., Chantrell R. W., Quantitative simulation of temperature-dependent magnetization dynamics and equilibrium properties of elemental ferromagnets. Phys. Rev. B 91, 144425 (2015). [Google Scholar]
- 67.Sato H., Yamanouchi M., Miura K., Ikeda S., Gan H. D., Mizunuma K., Koizumi R., Matsukura F., Ohno H., Junction size effect on switching current and thermal stability in CoFeB/MgO perpendicular magnetic tunnel junctions. Appl. Phys. Lett. 99, 042501 (2011). [Google Scholar]
- 68.Sato H., Enobio E. C. I., Yamanouchi M., Ikeda S., Fukami S., Kanai S., Matsukura F., Ohno H., Properties of magnetic tunnel junctions with a MgO/CoFeB/Ta/CoFeB/MgO recording structure down to junction diameter of 11 nm. Appl. Phys. Lett. 105, 062403 (2014). [Google Scholar]
- 69.Wernsdorfer W., Doudin B., Mailly D., Hasselbach K., Benoit A., Meier J., Ansermet J. P., Barbara B., Nucleation of magnetization reversal in individual nanosized nickel wires. Phys. Rev. Lett. 77, 1873–1876 (1996). [DOI] [PubMed] [Google Scholar]
- 70.Cayssol F., Ravelosona D., Chappert C., Ferré J., Jamet J. P., Domain wall creep in magnetic wires. Phys. Rev. Lett. 92, 107202 (2004). [DOI] [PubMed] [Google Scholar]
- 71.Bai H., Zhu W., You Y., Chen X., Zhou X., Pan F., Song C., Size-dependent anomalous Hall effect in noncollinear antiferromagnetic Mn3Sn films. Appl. Phys. Lett. 117, 052404 (2020). [Google Scholar]
- 72.Sato Y., Takeuchi Y., Yamane Y., Yoon J.-Y., Kanai S., Ieda J. I., Ohno H., Fukami S., Thermal stability of non-collinear antiferromagnetic Mn3Sn nanodot. Appl. Phys. Lett. 122, 122404 (2023). [Google Scholar]
- 73.Dong S., Xiang H., Dagotto E., Magnetoelectricity in multiferroics: A theoretical perspective. Natl. Sci. Rev. 6, 629–641 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.He X., Helbig N., Verstraete M. J., Bousquet E., TB2J: A python package for computing magnetic interaction parameters. Comput. Phys. Commun. 264, 107938 (2021). [Google Scholar]
- 75.Liechtenstein A. I., Katsnelson M. I., Antropov V. P., Gubanov V. A., Local spin density functional approach to the theory of exchange interactions in ferromagnetic metals and alloys. J. Magn. Magn. Mater. 67, 65–74 (1987). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.