Next Article in Journal
Highly Wear-Resistant Triboelectric Nanogenerators Based on Fluorocarbon-Graphene Hybrids
Previous Article in Journal
First-Principles Calculations of Plasmon-Induced Hot Carrier Properties of μ-Ag3Al
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Observation of Thickness-Modulated Out-of-Plane Spin–Orbit Torque in Polycrystalline Few-Layer Td-WTe2 Film

1
State Key Laboratory of Advanced Refractories, Wuhan University of Science and Technology, Wuhan 430081, China
2
School of Materials, Wuhan University of Science and Technology, Wuhan 430081, China
3
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China
4
Key Laboratory of Artificial Micro- and Nano-Structures of Ministry of Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(10), 762; https://doi.org/10.3390/nano15100762
Submission received: 11 April 2025 / Revised: 9 May 2025 / Accepted: 12 May 2025 / Published: 19 May 2025

Abstract

:
The low-symmetry Weyl semimetallic Td-phase WTe2 exhibits both a distinct out-of-plane damping torque ( τ DL ) and exceptional charge–spin interconversion efficiency enabled by strong spin-orbit coupling, positioning it as a prime candidate for spin–orbit torque (SOT) applications in two-dimensional transition metal dichalcogenides. Herein, we report on thickness-dependent unconventional out-of-plane τ DL in chemically vapor-deposited (CVD) polycrystalline Td-WTe2 (t)/Ni80Fe20/MgO/Ti (Td-WTN-t) heterostructures. Angle-resolved spin-torque ferromagnetic resonance measurements on the Td-WTN-12 structure showed significant spin Hall conductivities of σSH,y = 4.93 × 103 (ℏ/2e) Ω−1m−1 and σSH,z = 0.81 × 103 (ℏ/2e) Ω−1m−1, highlighting its potential for wafer-scale spin–orbit torque device applications. Additionally, a detailed examination of magnetotransport properties in polycrystalline few-layer Td-WTe2 films as a function of thickness revealed a marked amplification of the out-of-plane magnetoresistance, which can be ascribed to the anisotropic nature of charge carrier scattering mechanisms within the material. Spin pumping measurements in Td-WTN-t heterostructures further revealed thickness-dependent spin transport properties of Td-WTe2, with damping analysis yielding an out-of-plane spin diffusion length of λSD ≈ 14 nm.

1. Introduction

A fundamental requirement in developing next-generation magnetic random-access memory (MRAM) is the concurrent achievement of high-speed magnetization switching and ultra-low power consumption [1,2]. Generally, conventional spin-transfer torque (STT)-MRAM employs spin-transfer torque from polarized currents for memory operations, but suffers from low spin polarization efficiency and inseparable read/write paths that increase power dissipation [3,4]. Remarkably, the development of spin–orbit torque (SOT)-MRAM has demonstrated remarkable advantages, including fully isolated read/write paths and superior power efficiency compared to conventional STT-MRAM, establishing it as a promising candidate for future memory applications [5,6]. The SOT-MRAM utilizes spin-orbit coupling in nonmagnetic layers to convert charge current to spin current through the spin Hall effect (SHE) or the Rashba–Edelstein effect (REE) [7,8]. The resulting spin accumulation at the ferromagnetic layer (FL) interface generates SOT for reliable magnetization reversal. However, SOT devices employing conventional heavy metal (Pt, Ta, W, etc.)/FL heterostructures are fundamentally restricted to in-plane spin polarization, requiring external magnetic fields for magnetization switching [9,10,11]. This intrinsic limitation highlights the necessity for designed strong spin–orbit coupling materials with pronounced out-of-plane spin-polarized spin current to realize field-free magnetization reversal in next-generation spintronics.
Two-dimensional transition metal dichalcogenides (TMDs) demonstrate remarkable potential for SOT devices [12], featuring strong SOC [13], non-trivial band topology [14], and long spin diffusion lengths [15], while their low crystal symmetry enables deterministic out-of-plane damping-like torque ( τ DL )-an uncommon synergy that surpasses conventional heavy metals [16]. For instance, Husain et al. reported a superior charge-to-spin conversion efficiency (0.25) in TaS2 (0.88 nm)/Permalloy (Py) (7 nm) heterostructures, outperforming other TMD-based SOT devices and setting a new benchmark for SOT performance in this material class [17]. Stiehl et al. first identified out-of-plane antidamping torque in β-MoTe2-based heterostructures, challenging the conventional understanding of SOT systems [18]. Xu et al. demonstrated efficient perpendicular magnetization switching in PtTe2/Au/CoTb heterostructures, achieving spin Hall conductivities comparable to those of platinum (0.2–2 × 105 ℏ/2e Ω−1m−1) [19]. Guimaraes et al. employed spin-torque ferromagnetic resonance (ST-FMR) measurements on NbSe2/Py bilayers, revealing both out-of-plane field-like torque ( τ FL ) and in-plane τ DL [20]. Also, Lv et al. explored the electric field modulation of SOT in WS2/NiFe bilayers and demonstrated that applying a back-gate voltage effectively controls the ratio between τ FL and τ DL [21].
Among them, the semimetallic Td-phase WTe2 emerges as a preeminent candidate among TMDs for efficient charge-to-spin conversion, owing to its unique combination of strong SOC and structural inversion symmetry breaking characteristics stemming from its type II Weyl semimetal nature that promotes robust spin–momentum locking and topological surface states [22]. Therefore, these distinctive properties of Td-WTe2 enable field-free magnetization switching in perpendicularly magnetized systems. MacNeill et al. identified an out-of-plane τ DL in WTe2/Py bilayers under current applied along the low-symmetry axis, but absent along the high-symmetry axis, highlighting crystal symmetry as a means to control SOT [23]. Additionally, their analysis of current-induced torques revealed that the τ DL scales exclusively with WTe2 thickness, while τ FL also exhibits significant thickness dependence [24]. Peng et al. demonstrated that magnetron-sputtered amorphous WTe2-based heterostructures achieve a remarkably high damping-like SOT efficiency of approximately 0.20, rivaling crystalline WTe2 systems while offering superior fabrication scalability [25]. Li et al. reported a significant enhancement of spin conductivity in WTe2, where the spin-hole locking effect at low temperatures substantially amplifies the field-like torque in WTe2/Py bilayers [22]. Wang et al. studied the PtTe2/WTe2 bilayer system and achieved the first room-temperature observation of a strong out-of-plane τ DL , arising from the intrinsic crystalline asymmetry of WTe2, which facilitated perpendicular magnetization switching without the need for an external magnetic field [26]. According to the aforementioned research, SOT devices based on Weyl semimetal WTe2 can generate unconventional out-of-plane τ DL , enabling perpendicular magnetization switching, making them a novel and competitive candidate in the field of spintronics. Currently, SOT devices based on Td-WTe2 are primarily fabricated using mechanical exfoliation, which preserves crystallographic orientation integrity but fails to meet the scalability requirements for spintronic device production [23,24,27,28]. However, high-quality centimeter-scale Td-WTe2 films are typically synthesized via chemical vapor deposition (CVD) [29,30,31], establishing CVD-grown polycrystalline few-layer Td-WTe2 as an ideal platform for wafer-scale spin–orbit torque (SOT) device exploration.
In this study, high-quality, centimeter-scale polycrystalline few-layer Td-WTe2 thin films with tunable c-axis orientation were successfully fabricated through a two-step process involving initial magnetron sputtering, followed by tellurium-assisted CVD. We systematically characterized the anisotropic magnetoresistance (MR) in polycrystalline few-layer Td-WTe2 films across 10–300 K, observing an enhanced out-of-plane MR compared to in-plane MR at 10 K, which is consistent with anisotropic scattering associated with spin–orbit coupling. Spin pumping measurements in Td-WTe2 (4, 6, 8, 10, and 12 nm)/Ni80Fe20 (NiFe)/MgO/Ti heterostructures reveal thickness-dependent damping enhancement, yielding a spin diffusion length of λSD ≈ 14 nm via inverse spin Hall effect (ISHE) analysis. More importantly, the angle-resolved ST-FMR measurements demonstrate thickness-dependent in-plane and out-of-plane τ DL spin Hall conductivities and spin Hall angle in Td-WTe2/NiFe devices, enabling precise control of the unconventional out-of-plane τ DL through thickness modulation.

2. Experimental Methods

2.1. Preparation Methods

We first fabricated WOx (x < 3) films (3, 5, 7, 9, and 11 nm) on SiO2/Si substrates (HF-Kejing, Hefei, China) by RF magnetron sputtering at room temperature using a WO3 (99.99%, MAT-CN, Nanchang, China) target. The process maintained 1.2 Pa working pressure through controlled Ar flow, with 40 W sputtering power. Subsequently, a ceramic crucible loaded with 0.4 g of high-purity Te powder (99.999%, Aladdin, Wuhan, China) and a suitable quantity of molecular sieve (CanNa12-2n[(AlO2)12(SiO2)12]-xH2O (Aladdin, Wuhan, China)), which effectively regulates the release of Te vapor, was positioned at the center of the first heating zone in the CVD tube furnace. The temperature was then raised to 540 °C within 15 min under precisely controlled conditions. Simultaneously, the sputtered WOx (x < 3) thin films were positioned in the second heating zone of the CVD tube furnace, located 24 cm downstream from the Te powder, and the temperature was elevated to 600 °C within 15 min under precisely controlled conditions. The WTe2 films were synthesized using CVD at atmospheric pressure for 60 min using high-purity H2 (40 sccm) and Ar (30 sccm) as carrier gases, with the specific CVD setup illustrated in Figure S1a. Prior to thermal deposition, the system was purged for 15 min with Ar (500 sccm) and H2 (200 sccm) to maintain oxygen-free growth conditions.

2.2. Device Fabrication

Firstly, NiFe with a thickness of 6 nm was deposited onto Td-WTe2 films of varying thicknesses (4, 6, 8, 10, and 12 nm) via magnetron sputtering technology. Subsequently, a MgO/Ti bilayer structure (2 nm each) (MgO(2)/Ti(2)) was deposited sequentially using magnetron sputtering. For the fabrication of spin pumping devices, Td-WTe2 samples were deposited onto SiO2 (300 nm)/Si with dimensions of 2 × 7 mm. The NiFe(6)/MgO(2)/Ti(2) layers were sputter-deposited onto the central 2 × 5 mm area of the WTe2 sample, leaving a reserved area for electrode connection. For the ST-FMR devices, multiple microstrip devices measuring 20 × 4 μm at various angles were fabricated on 1 × 1 cm Td-WTe2(t)/NiFe(6)/MgO(2)/Ti(2) (Td-WTN-t) samples using a positive photoresist (AZ5214) photolithography process in conjunction with Ar-ion milling technology. Subsequently, the top electrode pattern was defined using a negative photoresist lithography process, followed by the deposition of Ti (5 nm)/Au (50 nm) layers onto the patterned regions via magnetron sputtering to fabricate the top electrodes for device measurement connections.

2.3. Characterization Methods and Instruments

The structural properties and thickness of the film were evaluated using a SmartLab 3 kW X-ray diffraction (XRD) instrument with Cu-Kα radiation and X-ray reflectivity (XRR). The crystallographic phase of the material was characterized with a Renishaw Invia Qontor confocal Raman spectrometer (532 nm laser). The surface morphologies of the samples were analyzed using a SINICO-XK-40 optical microscope (MO) and a Bruker atomic force microscope (AFM). The compositional analysis of the samples was conducted using X-ray photoelectron spectroscopy (XPS) with a monochromatic Al-Kα (1486.6 eV) X-ray source, employing an ESCALAB 250Xi system. The electrical resistance and magnetotransport properties of the material were measured using the Quantum Design Physical Property Measurement System (PPMS). The spin pumping system primarily consists of a PPMS for generating a magnetic field (−0.4 T to 0.4 T), an RF microwave source, an SR830 lock-in amplifier, and a 2182 nanovoltmeter. The RF alternating excitation is modulated by a sinusoidal signal (2 V) generated by the lock-in amplifier (171.4 Hz), with the magnetic field applied in the out-of-plane direction. The sample was placed at the center of a controllable angle rotation stage, and angle-dependent ST-FMR measurements were performed within a magnetic field range of −0.1 T to 0.1 T. By maintaining a fixed RF current (Irf) (5–9 GHz), the in-plane magnetic field was systematically scanned at a specific angle. The RF current modulation was produced by a microwave signal generator operating at an output power of 25 dBm, and the resulting voltage signal was measured using a lock-in amplifier with a frequency of 431.12 Hz.

3. Results and Discussions

The characteristic diffraction peak of the (002) crystal plane at 2θ = 12.61° for Td-WTe2 films with different thicknesses indicates that the polycrystalline few-layer Td-WTe2 grows along the c-axis direction of the crystal, as shown in Figure 1a. The thicknesses of Td-WTe2 films were accurately determined to be 3.96 nm, 6.12 nm, 8.05 nm, 9.94 nm, and 12.02 nm by fitting the XRR spectra using X’Pert Reflectivity software, showing excellent agreement with the nominal values (Figure 1b). The Raman spectrum in Figure 1c reveals the characteristic resonance modes A 1 3 , A 1 4 , A 1 7 , and A 1 9 at 114.8, 132.6, 162.1, and 210.1 cm−1, respectively, consistent with previous reports [32]. The dominant A 1 7 peak at 162.1 cm−1 unequivocally confirms that all synthesized WTe2 thin-film samples exhibit the Td-phase structure. Furthermore, Figure S1b,c displays the MO and AFM images of the polycrystalline few-layer Td-WTe2 film (4 nm) grown by CVD, illustrating its uniform coverage on the Si/SiO2 (1 × 1 cm2) substrate with a surface roughness of ~0.92 nm. The XPS spectrum of the 4 nm thick Td-WTe2 film exhibits characteristic peaks at 31.41 eV (W 4f7/2), 33.53 eV (W 4f5/2), 572.81 eV (Te 3d5/2), and 583.17 eV (Te 3d3/2), confirming the presence of W4+ and Te2−, as shown in Figure S1d,e.
Figure 2a shows the resistivity (ρ) of polycrystalline few-layer Td-WTe2 measured from 10 K to 300 K using the four-point probe technique, where ρ exhibits a linear relationship with temperature [33]:
ρ = ρ 0 + a T + b e x p ( Δ / T )
Here, ρ0 represents the residual resistivity, a is the coefficient for electron–phonon scattering, b corresponds to the nonlinear contribution from electron–electron interactions, and Δ denotes the activation temperature or energy gap associated with nonlinear excitations. The ρ of Td-WTe2 films increases with decreasing thickness, which is closely related to the enhanced effects of defects, impurities, and interface scattering in thinner films, as shown in Figure 2a. Moreover, all Td-WTe2 samples exhibit a metal–insulator transition, with ρ increasing at lower temperatures due to localization effects, consistent with their semimetallic character [34,35]. Additionally, Table S1 summarizes the ρ and the conductivity (σ) measurements of Td-WTe2 films with thicknesses of 4, 6, 8, 10, and 12 nm.
Figure 2b,c shows the magnetotransport properties of polycrystalline few-layer Td-WTe2 investigated by measuring the MR variations under in-plane and out-of-plane magnetic fields (B = 5 T). The MR values for Td-WTe2 samples with different thicknesses were calculated using Equation (2).
MR = R ( B ) R ( 0 ) R ( 0 ) × 100 %
Here, R(B) denotes the resistance under an applied magnetic field, whereas R(0) represents the resistance at a zero magnetic field. As shown in Figure 2b,c, all samples exhibit positive, non-saturating MR at 10 K, which is significantly smaller than that of single-crystal Td-WTe2, likely attributed to the increased defects and disorder in polycrystalline films [36]. Compared to thicker samples, the 4 nm thick Td-WTe2 demonstrates a higher MR value, which is likely due to the reduced carrier mobility in the thin layers of Td-WTe2 and the imbalance in charge compensation between electron and hole densities [37]. Furthermore, we observed that the MR is higher under an out-of-plane magnetic field compared to an in-plane magnetic field, which can be ascribed to the anisotropic scattering in Td-WTe2 [38,39]. Figure 2c demonstrates that the MR of Td-WTe2 (4 nm) increases as the temperature decreases, reaching a value of 3.46% at 10 K.
Figure 3a depicts the spin pumping measurement process conducted on the Td-WTN-t samples to characterize their spin transport properties, with the detailed experimental procedure available in our previous report [40]. We measured the mixed voltage (Vtotal) generated by the conversion of spin current, injected from the NiFe layer, into charge current within the Td-WTe2 layer via the ISHE, resulting from the spin pumping effect. Figure 3b illustrates the Vtotal of the Td-WTN-4 sample measured at 4 GHz and 25 dBm. By applying Equation (3), the symmetric Lorentzian profile (green line), associated with the bulk ISHE and Seebeck effects, and the asymmetric Lorentzian line (blue line), resulting from the anisotropic magnetoresistance and anomalous Hall effect in the NiFe layer, were fitted [41]:
V total = V s Δ H 2 Δ H 2 + ( H ext H 0 ) 2 + V a ( H ext H 0 ) Δ H Δ H 2 + ( H ext H 0 ) 2
V SE = ( V s ( + H 0 ) V s ( H 0 ) ) / 2
Among them, Vs represents the symmetric Lorentz voltage, Va represents the antisymmetric Lorentz voltage, ΔH denotes the linewidth, H0 is the resonance field, and Hext is the externally applied magnetic field. To extract the pure spin–charge conversion voltage (VSE), VS is corrected according to Equation (4) to eliminate the influence of the Seebeck effect [42]. Figure 3c displays VS of 4 nm thick Td-WTe2 under positive and negative magnetic fields in the frequency range of 4–16 GHz, with VSE values extracted from Vs using Equation (4), while also revealing that the Vs value of Td-WTN-4 decreases as the frequency (f) increases.
To determine the effective saturation magnetization (Meff) of Td-WTN-t, the Kittel formula was employed to fit the relationship between H0 and f in Figure 3d, as expressed by Equation (5) [43]:
f = γ μ 0 2 π H 0 ( H 0 + M eff )
Here, γ = g μ B represents the gyromagnetic ratio, g is the Landé splitting factor, which is 2.1, μ B is the Bohr magneton, ℏ is the reduced Planck constant, and μ 0 is the vacuum permeability. Table 1 summarizes the Meff values of the Td-WTe2-t and NiFe(6) samples. The results reveal that Meff decreases progressively with increasing Td-WTe2 thickness, reaching a maximum of 549 kA/m, which is close to the saturation magnetization (Ms). This trend is attributed to the modulation of interfacial magnetism in the NiFe(6) layer induced by the Td-WTe2 overlayer [25]. Table 1 also shows a slight decrease in Ms with increasing WTe2 thickness, which may be attributed to the influence of WTe2 on the WTe2/NiFe interface, leading to modifications in the electronic structure of the NiFe layer. Meanwhile, Figure S2a displays the VSM data for each layer of the Td-WTN-4 sample, demonstrating that Td-WTe2 is uniformly deposited on the NiFe layer. Furthermore, XRR analysis of the Td-WTe2(6)/NiFe(6) bilayer in Td-WTN-6 revealed a uniform 6.08 nm NiFe film atop a 5.86 nm Td-WTe2 layer, further confirming precise thickness control and a layer-by-layer growth mode (Figure S2b). Figure 3e demonstrates the linear dependence of ΔH on f for Td-WTN-t, indicating that Gilbert damping is the primary contributing factor [44]. The Gilbert damping (α) values for Td-WTN-t with varying thicknesses are extracted based on the Landau-Lifshitz-Gilbert (LLG) equation, as shown in Equation (6) [45].
μ 0 Δ H = μ 0 Δ H 0 + 4 π 3 γ α f
Here, ΔH0 is the inhomogeneous line-broadening factor. Table 1 demonstrates that as the thickness of WTe2 ( t WTe 2 ) in Td-WTN-t increases, α gradually rises, and H0 shifts towards higher magnetic fields (Figure S3), likely due to the enhanced SOC effects in thicker Td-WTe2. It is important to emphasize that the ΔH0 values obtained from fitting all samples using Equation (6) are minimal, demonstrating that the CVD-grown Td-WTe2/NiFe heterostructure exhibits exceptional quality. Furthermore, determining the effective mixing conductance ( g eff ) is also a critical parameter for evaluating the efficiency of spin pumping, as shown in Equation (7) [46]:
g eff = 4 π M s t NiFe g μ B ( α α 0 )
where α0 is the damping constant of NiFe (6), and Δα = αα0represents the change in damping relative to the NiFe layer, as shown in Table 1. Our measurements reveal that Td-WTN-t (t > 8) exhibits an exceptionally large g eff , exceeding typical values for high-SOC TMDs by an order of magnitude (Table 1) [47,48,49], while reaching magnitudes comparable to heavy metals like Pt and Ta [50,51]. Notably, the g eff in Td-WTN-t increases with t WTe 2 , likely due to enhanced spin pumping and suppressed spin backflow in thicker flakes [52].
Based on the ballistic transport theory, we determined the α mechanism of the Td-WTe2-t samples, enabling the derivation of the spin diffusion length (λSD) along the thickness direction of polycrystalline Td-WTe2 prepared using CVD, as shown in Equation (8) [53].
α = α 0 + g μ B g r 4 π M s 1 t NiFe ( 1 e 2 t WTe 2 λ SD )
Here, the spin backflow at the Td-WTe2/NiFe interface is reflected by the exponential term, and g r denotes the spin-mixing conductance of Td-WTN-t, incorporating the spin backflow. Figure 4a shows the exponential relationship between the t WTe 2 and α in the Td-WTN-t samples, indicating that the spin backflow recovers part of the angular momentum loss in the NiFe. By fitting the experimental data in Figure 4a using Equation (8), the spin diffusion length λSD = 14.05 ± 6.2 nm was determined, which provides direct evidence for the prominent role of spin backflow effects at the interface.
To rigorously validate the extracted λSD in polycrystalline Td-WTe2, we first quantify the spin current density ( J S ) in the Td-WTN-t samples using the following expression [54]:
J S = g eff 2 e ( γ μ 0 h rf ) 2 [ μ 0 M s γ + ( μ 0 M s γ ) 2 + 16 ( π f ) 2 ] 8 π α 2 [ ( μ 0 M s γ ) 2 + 16 ( π f ) 2 ]
Here, f is set at 4 GHz and hrf denotes the Irf field (0.017 mT) [40]. Table 1 summarizes the J S for all Td-WTN-t samples, demonstrating a gradual increase in J S with increasing Td-WTe2 thickness. Furthermore, by fitting the functional relationship between VSE/ J S and t WTe 2 using Equation (10), λSD is extracted, as shown below [55]:
V SE = w θ SHE λ SD tanh ( t WTe 2 2 λ SD ) t WTe 2 σ WTe 2 + t NiFe σ NiFe J S
where w is the sample width 7 mm, θSH is the spin Hall angle, and σ WTe 2 and σ NiFe denote the electrical conductivities of the Td-WTe2 and NiFe (see Table S1). The relationship between VSE/ J S and the thickness of Td-WTe2 for all Td-WTN-t samples was fitted using Equation (10), resulting in a spin diffusion length of λSD = 14.19 ± 1.12 nm, as shown in Figure 4a. This is consistent with our previous findings obtained by fitting the relationship between the α of Td-WTN-t and the thickness of Td-WTe2. Therefore, the λSD of polycrystalline Td-WTe2 prepared by CVD is approximately 14 nm, further confirming its accuracy in comparison to the reported λSD values of other TMDs along the thickness direction [40,49]. Table 1 shows that Td-WTe2 films with thicknesses of 4–12 nm exhibit θSH ranging from 0.0785 to 0.096, comparable to heavy metals like Ta and Pt [56,57]. This highlights the effective spin-to-charge conversion capability of polycrystalline few-layer Td-WTe2, where θSH exhibits a thickness-dependent enhancement, suggesting a dominant contribution from the bulk ISHE. Figure S3b shows the relationship between the VSE and input power (P) for Td-WTN-4 in the range of 18–25 dBm at 4 GHz. The results reveal that VSE decreases linearly with reduced power, as further demonstrated by the linear fitting in Figure S3c.
We employed the ST-FMR technique to systematically characterize SOT in a polycrystalline few-layer Td-WTe2 spin source layer, and the specific test equipment is shown in Figure 5a. At room temperature, when Irf is injected into the Td-WTe2 layer, a spin current is generated via the SHE or REE and injected into the adjacent NiFe layer. This induces an SOT, driving the precessional dynamics of the NiFe magnetic moments. When the f of Irf matches the resonance f of NiFe, the lock-in amplifier detects the resistance oscillation signal Vmix induced by ferromagnetic resonance, while the optimization of resonance conditions is achieved by scanning the magnetization direction at an angle of φ = 45° relative to the direction of Irf. The SOT generated by the Td-WTe2 layer comprises an in-plane damping-like torque ( τ DL = m × z × m) and an out-of-plane field-like torque ( τ FL = z × m), where m is the magnetization vector of the NiFe layer, and z denotes the induced spin polarization [58,59,60]. Additionally, the inset in Figure 5a presents the MO image of the fabricated lithographic device. Figure 5b presents the characteristic Vmix signals of Td-WTN-4 under both positive and negative external magnetic fields at a frequency of 4 GHz (25 dBm). Utilizing Equation (11), the symmetric Lorentzian voltage VS (green line) and asymmetric Lorentzian voltage VA (blue line) components were extracted from the Vmix signals through fitting, as demonstrated in the following analysis [25].
V mix = V S Δ H 2 Δ H 2 + ( H ext H 0 ) 2 + V A ( H ext H 0 ) Δ H Δ H 2 + ( H ext H 0 ) 2
In conventional heavy metals, VS and VA represent damping-like and field-like SOT components, while the reduced crystal symmetry of Td-WTe2 can induce additional torque contributions from symmetry-breaking spin-orbit interactions. Furthermore, the characteristic Vmix spectra of Td-WTN-4, obtained under magnetic field sweeps at 4–10 GHz (Figure 5c), demonstrate that the amplitude of Vmix decreases with increasing f. Figure 5d presents the Vmix signal of Td-WTN-t, revealing its significant dependence on the thickness of Td-WTe2. This phenomenon can be attributed to the key role of the bulk SHE in the charge–spin transition process. Figure S4a summarizes the H0 positions of Td-WTN-t at different f, with the Meff obtained by fitting with Equation (5), yielding a value close to Ms (see Table 2). As t WTe 2 increases, the bulk SHE in Td-WTN-t becomes increasingly dominant, which enhances magnetic anisotropy and consequently leads to a decrease in Meff. To extract the α parameter (Td-WTN-t) in ST-FMR tests, the following formula can be used for fitting [61].
μ 0 Δ H = μ 0 Δ H 0 + 2 π γ α f
We applied Equation (12) to fit the ΔH-f dependence of the Td-WTN-t samples shown in Figure S4b, extracted the corresponding α coefficients (see Table 2), and observed a marked increase in α with increasing Td-WTe2 thickness. This increased α suggests greater dissipation of spin angular momentum in thicker Td-WTe2 layers. Furthermore, the small linewidth of ΔH0 observed in all samples provides further evidence of the uniform deposition of polycrystalline few-layer Td-WTe2 films on the SiO2 (300 nm)/Si substrate.
To investigate the unconventional SOT induced by spin currents in the Td-WTe2 layer, we conducted φ-dependent measurements (φ = 0–360°) of the ST-FMR voltage signals for the Td-WTN-t samples under an in-plane magnetic field. The VS and antisymmetric VA components were analyzed by fitting them according to Equations (13) and (14), enabling the extraction of relevant parameters [62].
V S = ( V S , x sin ( φ + φ 0 ) + V S , y cos ( φ + φ 0 ) + V S , z ) sin ( 2 ( φ + φ 0 ) )
V A = ( V A , x sin ( φ + φ 0 ) + V A , y cos ( φ + φ 0 ) + V A , z ) sin ( 2 ( φ + φ 0 ) )
Here, φ0 denotes the angular calibration offset; VS,x and VS,y originate from the in-plane τ DL , whereas VS,z corresponds to the in-plane τ FL ; VA,x represents the out-of-plane τ FL ; VA,y originates from the out-of-plane torque induced by the Oersted field generated by the Irf; and VA,z corresponds to the out-of-plane τ DL . Figure 6a,b and Figure S5a–f present the angular-dependent behavior of the VA and VS components for Td-WTN-t samples with varying thicknesses at a frequency of f = 4 GHz, accompanied by the corresponding fitting curves derived from Equations (13) and (14). Also, the amplitudes of the relevant VS and VA components extracted from all samples were used to systematically study the efficiency of conventional and unconventional SOT. Since the fitting parameters VA,x and VS,x exhibit negligible magnitudes, the contribution of the spin currents generated along the x-direction can be considered insignificant.
Based on the out-of-plane τ FL effects primarily originating from the Oersted field contribution induced by the Irf flowing through the Td-WTe2 layer, this study employs the correlation between the VA,y and the current density vector in the Td-WTN-t heterostructure. Through standardized normalization, we quantitatively characterize θSH. Therefore, based on the extracted amplitudes of VS,y, VA,y and VA,z, the effective spin Hall angles for the in-plane θSH,y and out-of-plane θSH,z components were quantified using Equations (15) and (16), expressed as [59]
θ SH , y = ( 1 + M eff / H 0 ) 1 / 2 e μ 0 M s t WTe 2 t NiFe V S , y V A , y
θ SH , z = e μ 0 M s t WTe 2 t NiFe V A , z V A , y
where Meff represents the effective magnetization of the Td-WTN-t samples (see Table 2), while Ms = 605 kA/m denotes the saturation magnetization of the NiFe layer (see Table 1). Table 2 presents the in-plane θSH,y for Td-WTN-4, Td-WTN-6, Td-WTN-8, Td-WTN-10, and Td-WTN-12 as 0.050, 0.071, 0.074, 0.077, and 0.078, respectively. Notably, the unconventional out-of-plane θSH,z for Td-WTN-t (4, 6, 8, 10, and 12 nm) exhibits an increasing trend with Td-WTe2 thickness, with corresponding values of 0.002, 0.003, 0.005, 0.010, and 0.013. Figure 7a illustrates the dependence of θSH,y and θSH,z on t WTe 2 , revealing that both θSH,y and θSH,z increase with increasing Td-WTe2 thickness. This enhancement can be attributed to the dominant contribution of the bulk SHE, where the amplified bulk SHE facilitates greater spin current accumulation at the interface, thereby augmenting both θSH,y and θSH,z. Notably, when Irf is applied along the low-symmetry crystallographic direction of Td-WTe2, it typically induces an unconventional out-of-plane τ DL [23]. However, since the few-layer Td-WTe2 fabricated by CVD in our study is polycrystalline, the aforementioned examples cannot be directly invoked to explain the origin of the out-of-plane θSH,z. Nevertheless, recent studies by Guimaraes et al. and Bangar et al. on polycrystalline materials suggest that the observed out-of-plane τ DL may arise from stress-induced symmetry breaking during the sample fabrication process [20,63].
To determine the spin Hall conductivities (σSH) at the interface between polycrystalline few-layer Td-WTe2 and NiFe along the y and z components, the σSH can be obtained by normalizing the measured θSH with the ρ of the Td-WTe2 thin film, as expressed by [64]
σ SH , y ( z ) = θ SH , y ( z ) / ρ
Here, σSH,y and σSH,z represent the σSH along the y- and z-directions, respectively. The σSH,y and σSH,z for Td-WTN-t (t = 4, 6, 8, 10, and 12 nm) are summarized in Table 2. Notably, σSH,y = 4.93 × 103 (ℏ/2e) Ω−1m−1 and σSH,z = 0.81 × 103 (ℏ/2e) Ω−1m−1 are achieved at t WTe 2 = 12 nm. Due to the polycrystalline nature of the fabricated Td-WTe2, both σSH,y and σSH,z are smaller than those of single-crystal Td-WTe2 but remain within the same order of magnitude [65]. And these values are comparable to the σSH reported by Shi et al. for epitaxial Td-WTe2 thin films grown via CVD [66], demonstrating that our polycrystalline few-layer Td-WTe2 also exhibits a significant out-of-plane SOT efficiency. Moreover, the σSH,y and σSH,z of the Td-WTN-t samples increase with t WTe 2 , as shown in Figure 7b. This trend further supports the fact that in thicker Td-WTN-t devices (when the Td-WTe2 thickness is less than λSD), the bulk SHE plays a dominant role, enabling a more efficient spin current injection into the NiFe layer, which in turn enhances both σSH,y and σSH,z. Furthermore, these findings suggest the potential for tuning the unconventional SOT in polycrystalline Td-WTe2 thin films by precisely controlling their thickness.

4. Conclusions

In summary, we successfully obtained centimeter-scale polycrystalline Td-WTe2 films with controlled thicknesses using CVD. The anisotropic scattering of charge carriers in Td-WTe2 with varying thicknesses leads to consistently higher MR under out-of-plane magnetic fields compared to in-plane configurations. Thickness-resolved spin pumping in Td-WTN-t identifies bulk ISHE as the primary spin–charge conversion mechanism. By modeling α and VSE/ J S vs. t WTe 2 , we extract λSD ≈ 14 nm, a critical parameter for vertical spintronic devices. Through angle-resolved ST-FMR, we resolve the t WTe 2 dependence of both in-plane and out-of-plan τ DL spin Hall conductivities in Td-WTN-t heterostructures, establishing polycrystalline few-layer Td-WTe2 as a platform for tunable unconventional SOT via scalable thickness modulation. Thicker Td-WTe2 boosts bulk SHE-driven spin current injection into NiFe, enhancing both in-plane and out-of-plane τ DL . Consequently, achieving controllable out-of-plane spin polarization and field-free magnetization switching in wafer-scale CVD-grown polycrystalline few-layer Td-WTe2 films has emerged as a critical research direction for SOT device applications.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano15100762/s1, Figure S1: (a) Schematic diagram of the growth mechanism of Td-WTe2 utilizing WOx (x < 3) as the precursor. (b,c) The MO and AFM images of Td-WTe2 (4 nm). (d,e) The XPS spectrum of Td-WTe2 (4 nm). Figure S2: (a) The VSM diagrams for NiFe(6), Td-WTe2(4)/NiFe(6), Td-WTe2(4)/NiFe(6)/MgO(2) and Td-WTe2(4)/NiFe(6)/MgO(2)/Ti(2), respectively. (b) The XRR spectra (intensity vs. 2θ) of Td-WTe2(6)/NiFe(6) films. Figure S3: (a) The Vs of the Td-WTN-t samples at 4 GHz. (b) The VSE plot of Td-WTN-4 at 4 GHz, ranging from 18 to 25 dBm. (c) The linear relationship between VSE and P for Td-WTN-4. Figure S4: (a) H0 dependence of f for Td-WTN-t. (b) f dependence of ΔH for Td-WTN-t. Figure S5: (a–f) The angular dependence of the ST-FMR signals with plane magnetic field for VS and VA components of Td-WTN-t (6, 8 and 10 nm) at 4 GHz. Table S1: The resistivity (ρ0) and conductivity (σ) of Td-WTe2(t) and NiFe(6) samples.

Author Contributions

Methodology, M.Z.; Formal analysis, M.Z., W.Z., Y.L. (You Lv), Y.L. (Yong Liu) and R.X.; Investigation, M.Z.; Resources, Z.Z. and Z.L.; Data curation, M.Z., W.Z., Y.L. (You Lv), Y.L. (Yong Liu) and R.X.; Writing—original draft, M.Z.; Writing—review & editing, Z.Z. and Z.L.; Supervision, Z.Z. and Z.L.; Project administration, Z.Z. and Z.L.; Funding acquisition, Z.Z. and Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grants No. 12204364), the National Key Research and Development Program of China (2022YFA1602701), and the National Major Scientific Instrument and Equipment Development Project of the National Natural Science Foundation of China (12227806).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Miron, I.M.; Garello, K.; Gaudin, G.; Zermatten, P.-J.; Costache, M.V.; Auffret, S.; Bandiera, S.; Rodmacq, B.; Schuhl, A.; Gambardella, P. Perpendicular switching of a single ferromagnetic layer induced by in-plane current injection. Nature 2011, 476, 189–193. [Google Scholar] [CrossRef] [PubMed]
  2. Cubukcu, M.; Boulle, O.; Drouard, M.; Garello, K.; Onur Avci, C.; Mihai Miron, I.; Langer, J.; Ocker, B.; Gambardella, P.; Gaudin, G. Spin-orbit torque magnetization switching of a three-terminal perpendicular magnetic tunnel junction. Appl. Phys. Lett. 2014, 104, 042406. [Google Scholar] [CrossRef]
  3. Fong, X.; Kim, Y.; Venkatesan, R.; Choday, S.H.; Raghunathan, A.; Roy, K. Spin-transfer torque memories: Devices, circuits, and systems. Proc. IEEE 2016, 104, 1449–1488. [Google Scholar] [CrossRef]
  4. Asifuzzaman, K.; Verdejo, R.S.; Radojković, P. Performance and Power Estimation of STT-MRAM main memory with reliable system-level simulation. ACM Trans. Embed. Comput. Syst. 2022, 21, 6. [Google Scholar] [CrossRef]
  5. Manchon, A.; Železný, J.; Miron, I.M.; Jungwirth, T.; Sinova, J.; Thiaville, A.; Garello, K.; Gambardella, P. Current-induced spin-orbit torques in ferromagnetic and antiferromagnetic systems. Rev. Mod. Phys. 2019, 91, 035004. [Google Scholar] [CrossRef]
  6. Liu, Y.; Shao, Q. Two-dimensional materials for energy-efficient spin–orbit torque devices. ACS Nano 2020, 14, 9389–9407. [Google Scholar] [CrossRef]
  7. 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 2012, 336, 555–558. [Google Scholar] [CrossRef]
  8. Shao, Q.; Yu, G.; Lan, Y.-W.; Shi, Y.; Li, M.-Y.; Zheng, C.; Zhu, X.; Li, L.-J.; Amiri, P.K.; Wang, K.L. Strong rashba-edelstein effect-induced spin–orbit torques in monolayer transition metal dichalcogenide/ferromagnet bilayers. Nano Lett. 2016, 16, 7514–7520. [Google Scholar] [CrossRef]
  9. Pai, C.-F.; Liu, L.; Li, Y.; Tseng, H.W.; Ralph, D.C.; Buhrman, R.A. Spin transfer torque devices utilizing the giant spin Hall effect of tungsten. Appl. Phys. Lett. 2012, 101, 122404. [Google Scholar] [CrossRef]
  10. Oh, Y.-W.; Ryu, J.; Kang, J.; Park, B.-G. Material and thickness investigation in ferromagnet/Ta/CoFeB trilayers for enhancement of spin–orbit torque and field-free switching. Adv. Electron. Mater. 2019, 5, 1900598. [Google Scholar] [CrossRef]
  11. de la Venta, J.J.; Wang, S.; Ramirez, J.G.; Schuller, I.K. Control of magnetism across metal to insulator transitions. Appl. Phys. Lett. 2013, 102, 122404. [Google Scholar] [CrossRef]
  12. Feng, J.; Li, K.; Zheng, M.; Zhang, W.; Liu, Y.; Wang, D.; Zhang, Z.; Zuo, C.; Xiong, R.; Lu, Z. Excellent spin-filtering and giant tunneling magnetoresistance in a dual-electrode van der Waals magnetic tunnel junction based on ferromagnetic CrSe2. Appl. Surf. Sci. 2023, 611, 155588. [Google Scholar] [CrossRef]
  13. Ali, M.N.; Xiong, J.; Flynn, S.; Tao, J.; Gibson, Q.D.; Schoop, L.M.; Liang, T.; Haldolaarachchige, N.; Hirschberger, M.; Ong, N.P.; et al. Large, non-saturating magnetoresistance in WTe2. Nature 2014, 514, 205–208. [Google Scholar] [CrossRef] [PubMed]
  14. Das, P.K.; Di Sante, D.; Vobornik, I.; Fujii, J.; Okuda, T.; Bruyer, E.; Gyenis, A.; Feldman, B.E.; Tao, J.; Ciancio, R.; et al. Layer-dependent quantum cooperation of electron and hole states in the anomalous semimetal WTe2. Nat. Commun. 2016, 7, 10847. [Google Scholar] [CrossRef]
  15. Lin, X.; Zhu, L. Magnetization switching in van der Waals systems by spin-orbit torque. Mater. Today Electron. 2023, 4, 100037. [Google Scholar] [CrossRef]
  16. Feng, B.; Chan, Y.-H.; Feng, Y.; Liu, R.-Y.; Chou, M.-Y.; Kuroda, K.; Yaji, K.; Harasawa, A.; Moras, P.; Barinov, A.; et al. Spin texture in type-II Weyl semimetal WTe2. Phys. Rev. B 2016, 94, 195134. [Google Scholar] [CrossRef]
  17. Husain, S.; Chen, X.; Gupta, R.; Behera, N.; Kumar, P.; Edvinsson, T.; García-Sánchez, F.; Brucas, R.; Chaudhary, S.; Sanyal, B.; et al. Large damping-like spin–orbit torque in a 2D conductive 1T-TaS2 monolayer. Nano Lett. 2020, 20, 6372–6380. [Google Scholar] [CrossRef]
  18. Stiehl, G.M.; Li, R.; Gupta, V.; Baggari, I.E.; Jiang, S.; Xie, H.; Kourkoutis, L.F.; Mak, K.F.; Shan, J.; Buhrman, R.A.; et al. Layer-dependent spin-orbit torques generated by the centrosymmetric transition metal dichalcogenide β-MoTe2. Phys. Rev. B 2019, 100, 184402. [Google Scholar] [CrossRef]
  19. Xu, H.; Wei, J.; Zhou, H.; Feng, J.; Xu, T.; Du, H.; He, C.; Huang, Y.; Zhang, J.; Liu, Y.; et al. High Spin Hall Conductivity in Large-Area Type-II Dirac Semimetal PtTe2. Adv. Mater. 2020, 32, 2000513. [Google Scholar] [CrossRef]
  20. Guimarães, M.H.D.; Stiehl, G.M.; MacNeill, D.; Reynolds, N.D.; Ralph, D.C. Spin–orbit torques in NbSe2/permalloy bilayers. Nano Lett. 2018, 18, 1311–1316. [Google Scholar] [CrossRef]
  21. Lv, W.; Jia, Z.; Wang, B.; Lu, Y.; Luo, X.; Zhang, B.; Zeng, Z.; Liu, Z. Electric-field control of spin–orbit torques in WS2/permalloy bilayers. ACS Appl. Mater. Interfaces 2018, 10, 2843–2849. [Google Scholar] [CrossRef] [PubMed]
  22. Li, P.; Wu, W.; Wen, Y.; Zhang, C.; Zhang, J.; Zhang, S.; Yu, Z.; Yang, S.A.; Manchon, A.; Zhang, X. Spin-momentum locking and spin-orbit torques in magnetic nano-heterojunctions composed of Weyl semimetal WTe2. Nat. Commun. 2018, 9, 3990. [Google Scholar] [CrossRef] [PubMed]
  23. MacNeill, D.; Stiehl, G.M.; Guimaraes, M.H.D.; Buhrman, R.A.; Park, J.; Ralph, D.C. Control of spin–orbit torques through crystal symmetry in WTe2/ferromagnet bilayers. Nat. Phys. 2017, 13, 300–305. [Google Scholar] [CrossRef]
  24. MacNeill, D.; Stiehl, G.M.; Guimarães, M.H.D.; Reynolds, N.D.; Buhrman, R.A.; Ralph, D.C. Thickness dependence of spin-orbit torques generated by WTe2. Phys. Rev. B 2017, 96, 054450. [Google Scholar] [CrossRef]
  25. Peng, C.-W.; Liao, W.-B.; Chen, T.-Y.; Pai, C.-F. Efficient spin-orbit torque generation in semiconducting WTe2 with hopping transport. ACS Appl. Mater. Interfaces 2021, 13, 15950–15957. [Google Scholar] [CrossRef]
  26. Wang, F.; Shi, G.; Kim, K.-W.; Park, H.-J.; Jang, J.G.; Tan, H.R.; Lin, M.; Liu, Y.; Kim, T.; Yang, D.; et al. Field-free switching of perpendicular magnetization by two-dimensional PtTe2/WTe2 van der Waals heterostructures with high spin Hall conductivity. Nat. Mater. 2024, 23, 768–774. [Google Scholar] [CrossRef]
  27. Shi, S.; Liang, S.; Zhu, Z.; Cai, K.; Pollard, S.D.; Wang, Y.; Wang, J.; Wang, Q.; He, P.; Yu, J.; et al. All-electric magnetization switching and Dzyaloshinskii-Moriya interaction in WTe2/ferromagnet heterostructures. Nat. Nanotechnol. 2019, 14, 945–949. [Google Scholar] [CrossRef]
  28. Zhao, B.; Khokhriakov, D.; Zhang, Y.; Fu, H.; Karpiak, B.; Hoque, A.M.; Xu, X.; Jiang, Y.; Yan, B.; Dash, S.P. Observation of charge to spin conversion in Weyl semimetal WTe2 at room temperature. Phys. Rev. Res. 2020, 2, 013286. [Google Scholar] [CrossRef]
  29. Wang, Q.; Li, J.; Besbas, J.; Hsu, C.-H.; Cai, K.; Yang, L.; Cheng, S.; Wu, Y.; Zhang, W.; Wang, K.; et al. Room-temperature nanoseconds spin relaxation in WTe2 and MoTe2 thin films. Adv. Sci. 2018, 5, 1700912. [Google Scholar] [CrossRef]
  30. Zhou, Y.; Jang, H.; Woods, J.M.; Xie, Y.; Kumaravadivel, P.; Pan, G.A.; Liu, J.; Liu, Y.; Cahill, D.G.; Cha, J.J. Direct synthesis of large-Scale WTe2 thin films with low thermal conductivity. Adv. Funct. Mater. 2017, 27, 1605928. [Google Scholar] [CrossRef]
  31. Li, J.; Cheng, S.; Liu, Z.; Zhang, W.; Chang, H. Centimeter-Scale, large-area, few-Layer 1T’-WTe2 films by chemical vapor deposition and its long-term stability in ambient condition. J. Phys. Chem. C 2018, 122, 7005–7012. [Google Scholar] [CrossRef]
  32. Kim, Y.; Jhon, Y.I.; Park, J.; Kim, J.H.; Lee, S.; Jhon, Y.M. Anomalous Raman scattering and lattice dynamics in mono- and few-layer WTe2. Nanoscale 2016, 8, 2309–2316. [Google Scholar] [CrossRef] [PubMed]
  33. Pandey, L.; Kumar, N.; Khan, A.; Kumar Gupta, N.; Hait, S.; Barwal, V.; Mishra, V.; Sharma, N.; Chaudhary, S. Growth and characterization of the sputtered type-II topological semimetal PdTe2 thin films and PdTe2/Co60Fe20B20 heterostructures. J. Magn. Magn. Mater. 2023, 584, 171075. [Google Scholar] [CrossRef]
  34. Zhang, E.; Chen, R.; Huang, C.; Yu, J.; Zhang, K.; Wang, W.; Liu, S.; Ling, J.; Wan, X.; Lu, H.-Z.; et al. Tunable positive to negative magnetoresistance in atomically thin WTe2. Nano Lett. 2017, 17, 878–885. [Google Scholar] [CrossRef]
  35. Song, P.; Hsu, C.; Zhao, M.; Zhao, X.; Chang, T.R.; Teng, J.; Lin, H.; Loh, K.P. Few-layer 1T′ MoTe2 as gapless semimetal with thickness dependent carrier transport. 2D Mater. 2018, 5, 031010. [Google Scholar] [CrossRef]
  36. Adhikari, R.; Adhikari, S.; Faina, B.; Terschanski, M.; Bork, S.; Leimhofer, C.; Cinchetti, M.; Bonanni, A. Positive magnetoresistance and chiral anomaly in exfoliated type-II weyl semimetal Td-WTe2. Nanomaterials 2021, 11, 2755. [Google Scholar] [CrossRef]
  37. Wang, L.; Gutiérrez-Lezama, I.; Barreteau, C.; Ubrig, N.; Giannini, E.; Morpurgo, A.F. Tuning magnetotransport in a compensated semimetal at the atomic scale. Nat. Commun. 2015, 6, 8892. [Google Scholar] [CrossRef]
  38. Tian, Z.K.; Guo, J.J.; Luo, Z.y.; Nie, Y.Z.; Xia, Q.l.; Zhou, Y.; Guo, G.H. Abnormal sign change of angle-dependent magnetoresistance in polycrystalline WTe2 nanoplate. Phys. E Low-Dimens. Syst. Nanostruct. 2023, 150, 115699. [Google Scholar] [CrossRef]
  39. Nepal, R.; Sharma, V.; Pogue, L.; Drichko, N.; Budhani, R.C. Disorder driven variations in magnetoresistance and planar Hall effect in Bi2Te3 thin films. Thin Solid Films 2022, 761, 139520. [Google Scholar] [CrossRef]
  40. Zheng, M.; Zhang, W.; Lv, Y.; Liu, Y.; Xiong, R.; Zhang, Z.; Lu, Z. Low-temperature fabrication, magnetoresistance and spin pumping studies of polycrystalline few-layer 1T’-MoTe2 films. J. Alloys Compd. 2025, 1015, 178775. [Google Scholar] [CrossRef]
  41. Kondou, K.; Sukegawa, H.; Kasai, S.; Mitani, S.; Niimi, Y.; Otani, Y. Influence of inverse spin Hall effect in spin-torque ferromagnetic resonance measurements. Appl. Phys. Express 2016, 9, 023002. [Google Scholar] [CrossRef]
  42. Fan, Y.; Li, H.; DC, M.; Peterson, T.; Held, J.; Sahu, P.; Chen, J.; Zhang, D.; Mkhoyan, A.; Wang, J.P. Spin pumping and large field-like torque at room temperature in sputtered amorphous WTe2−x films. APL Mater. 2020, 8, 041102. [Google Scholar] [CrossRef]
  43. Kittel, C. On the Theory of Ferromagnetic Resonance Absorption. Phys. Rev. 1948, 73, 155–161. [Google Scholar] [CrossRef]
  44. Behera, N.; Kumar, A.; Chaudhary, S.; Pandya, D.K. Two magnon scattering and anti-damping behavior in a two-dimensional epitaxial TiN/Py(tPy)/β-Ta(tTa) system. RSC Adv. 2017, 7, 8106–8117. [Google Scholar] [CrossRef]
  45. Weber, R.; Han, D.-S.; Boventer, I.; Jaiswal, S.; Lebrun, R.; Jakob, G.; Kläui, M. Gilbert damping of CoFe-alloys. J. Phys. D Appl. Phys. 2019, 52, 325001. [Google Scholar] [CrossRef]
  46. Tao, X.; Liu, Q.; Miao, B.; Yu, R.; Feng, Z.; Sun, L.; You, B.; Du, J.; Chen, K.; Zhang, S.; et al. Self-consistent determination of spin Hall angle and spin diffusion length in Pt and Pd: The role of the interface spin loss. Sci. Adv. 2018, 4, eaat1670. [Google Scholar] [CrossRef]
  47. Hait, S.; Gupta, N.K.; Sharma, N.; Pandey, L.; Kumar, N.; Barwal, V.; Kumar, P.; Chaudhary, S. Spin pumping in nanolayers of WS2/Co2FeAl heterostructures: Large spin mixing conductance and spin transparency. J. Appl. Phys. 2022, 132, 133901. [Google Scholar] [CrossRef]
  48. Mendes, J.B.S.; Aparecido-Ferreira, A.; Holanda, J.; Azevedo, A.; Rezende, S.M. Efficient spin to charge current conversion in the 2D semiconductor MoS2 by spin pumping from yttrium iron garnet. Appl. Phys. Lett. 2018, 112, 242407. [Google Scholar] [CrossRef]
  49. Sun, W.; Chen, Y.; Zhuang, W.; Chen, Z.; Song, A.; Liu, R.; Wang, X. Sizable spin-to-charge conversion in PLD-grown amorphous (Mo, W)Te2−x films. Nanotechnology 2023, 34, 135001. [Google Scholar] [CrossRef]
  50. You, Y.; Sakimura, H.; Harumoto, T.; Nakamura, Y.; Shi, J.; Song, C.; Pan, F.; Ando, K. Study of spin mixing conductance of single oriented Pt in Pt/Ni81Fe19 heterostructure by spin pumping. AIP Adv. 2021, 11, 035211. [Google Scholar] [CrossRef]
  51. Paikaray, B.; Sahoo, S.K.; Manoj, T.; Sriram, K.; Basumatary, H.; Haldar, A.; Murapaka, C. Large spin pumping and inverse spin Hall effect in Ta/Py bilayer structures. Phys. Status Solidi A 2022, 219, 2100608. [Google Scholar] [CrossRef]
  52. Hait, S.; Husain, S.; Bangar, H.; Pandey, L.; Barwal, V.; Kumar, N.; Gupta, N.K.; Mishra, V.; Sharma, N.; Gupta, P.; et al. Spin pumping through different spin–orbit coupling interfaces in β-W/interlayer/Co2FeAl heterostructures. ACS Appl. Mater. Interfaces 2022, 14, 37182–37191. [Google Scholar] [CrossRef] [PubMed]
  53. Mosendz, O.; Pearson, J.E.; Fradin, F.Y.; Bauer, G.E.W.; Bader, S.D.; Hoffmann, A. Quantifying spin hall angles from spin pumping: Experiments and theory. Phys. Rev. Lett. 2010, 104, 046601. [Google Scholar] [CrossRef] [PubMed]
  54. Rogdakis, K.; Sud, A.; Amado, M.; Lee, C.M.; McKenzie-Sell, L.; Jeon, K.R.; Cubukcu, M.; Blamire, M.G.; Robinson, J.W.A.; Cohen, L.F.; et al. Spin transport parameters of NbN thin films characterized by spin pumping experiments. Phys. Rev. Mater. 2019, 3, 014406. [Google Scholar] [CrossRef]
  55. Azevedo, A.; Vilela-Leão, L.H.; Rodríguez-Suárez, R.L.; Lacerda Santos, A.F.; Rezende, S.M. Spin pumping and anisotropic magnetoresistance voltages in magnetic bilayers: Theory and experiment. Phys. Rev. B 2011, 83, 144402. [Google Scholar] [CrossRef]
  56. Feng, Z.; Hu, J.; Sun, L.; You, B.; Wu, D.; Du, J.; Zhang, W.; Hu, A.; Yang, Y.; Tang, D.M.; et al. Spin Hall angle quantification from spin pumping and microwave photoresistance. Phys. Rev. B 2012, 85, 214423. [Google Scholar] [CrossRef]
  57. Yu, R.; Miao, B.F.; Sun, L.; Liu, Q.; Du, J.; Omelchenko, P.; Heinrich, B.; Wu, M.; Ding, H.F. Determination of spin Hall angle and spin diffusion length in β-phase-dominated tantalum. Phys. Rev. Mater. 2018, 2, 074406. [Google Scholar] [CrossRef]
  58. Bai, H.; Zhou, X.F.; Zhang, H.W.; Kong, W.W.; Liao, L.Y.; Feng, X.Y.; Chen, X.Z.; You, Y.F.; Zhou, Y.J.; Han, L.; et al. Control of spin-orbit torques through magnetic symmetry in differently oriented noncollinear antiferromagnetic Mn3Pt. Phys. Rev. B 2021, 104, 104401. [Google Scholar] [CrossRef]
  59. Liu, L.; Moriyama, T.; Ralph, D.C.; Buhrman, R.A. Spin-torque ferromagnetic resonance induced by the spin hall effect. Phys. Rev. Lett. 2011, 106, 036601. [Google Scholar] [CrossRef]
  60. Amin, V.P.; Stiles, M.D. Spin transport at interfaces with spin-orbit coupling: Formalism. Phys. Rev. B 2016, 94, 104419. [Google Scholar] [CrossRef]
  61. Luo, Y.; Chen, Q.; Li, R.; Wang, Y.; Lv, W.; Zhang, B.; Fan, Y.; Wu, H.; Zeng, Z. Enhanced spin–orbit torque and field-free switching in Au/TMDs/Ni hybrid structures. Nanoscale 2023, 15, 3142–3149. [Google Scholar] [CrossRef] [PubMed]
  62. 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.; et al. Tilted spin current generated by the collinear antiferromagnet ruthenium dioxide. Nat. Electron. 2022, 5, 267–274. [Google Scholar] [CrossRef]
  63. Bangar, H.; Gupta, P.; Singh, R.; Muduli, P.K.; Dewan, S.; Das, S. Optimization of growth of large-area SnS thin films and heterostructures for spin pumping and spin-orbit torque. Phys. Rev. Mater. 2023, 7, 094406. [Google Scholar] [CrossRef]
  64. Bainsla, L.; Zhao, B.; Behera, N.; Hoque, A.M.; Sjöström, L.; Martinelli, A.; Abdel-Hafiez, M.; Åkerman, J.; Dash, S.P. Large out-of-plane spin–orbit torque in topological Weyl semimetal TaIrTe4. Nat. Commun. 2024, 15, 4649. [Google Scholar] [CrossRef]
  65. Zhang, Y.; Xu, H.; Jia, K.; Lan, G.; Huang, Z.; He, B.; He, C.; Shao, Q.; Wang, Y.; Zhao, M.; et al. Room temperature field-free switching of perpendicular magnetization through spin-orbit torque originating from low-symmetry type II Weyl semimetal. Sci. Adv. 2023, 9, eadg9819. [Google Scholar] [CrossRef]
  66. Shi, S.; Li, J.; Hsu, C.-H.; Lee, K.; Wang, Y.; Yang, L.; Wang, J.; Wang, Q.; Wu, H.; Zhang, W.; et al. Observation of the out-of-plane polarized spin current from CVD grown WTe2. Adv. Quantum Technol. 2021, 4, 2100038. [Google Scholar] [CrossRef]
Figure 1. (a,b) XRD diffraction patterns and XRR spectra (intensity vs. 2θ) of polycrystalline few-layer Td-WTe2 films with varying thicknesses. (c) Raman spectra of polycrystalline few-layer Td-WTe2 films with different thicknesses, along with the SiO2/Si substrate.
Figure 1. (a,b) XRD diffraction patterns and XRR spectra (intensity vs. 2θ) of polycrystalline few-layer Td-WTe2 films with varying thicknesses. (c) Raman spectra of polycrystalline few-layer Td-WTe2 films with different thicknesses, along with the SiO2/Si substrate.
Nanomaterials 15 00762 g001
Figure 2. (a) The ρ variation of polycrystalline few-layer Td-WTe2 (4, 6, 8, 10, and 12 nm) from 10 to 300 K. (b) The MR of polycrystalline few-layer Td-WTe2 (4, 6, 8, 10, and 12 nm) under out-of-plane and in-plane magnetic field configurations at 10 K. (c) The MR of Td-WTe2 (4 nm) under in-plane and out-of-plane magnetic fields from 10 to 300 K.
Figure 2. (a) The ρ variation of polycrystalline few-layer Td-WTe2 (4, 6, 8, 10, and 12 nm) from 10 to 300 K. (b) The MR of polycrystalline few-layer Td-WTe2 (4, 6, 8, 10, and 12 nm) under out-of-plane and in-plane magnetic field configurations at 10 K. (c) The MR of Td-WTe2 (4 nm) under in-plane and out-of-plane magnetic fields from 10 to 300 K.
Nanomaterials 15 00762 g002
Figure 3. (a) Schematic representation of the spin-pumping experimental mechanism. (b) Vtotal spectrum of the Td-WTN-4 sample at 4 GHz, with Vs and Va as fitted curves. (c) The Vs of Td-WTN-4 at different frequencies. (d,e) The H0 curves and ΔH versus f curves for Td-WTN-t and NiFe(6). The solid spheres represent the experimentally measured data, while the red line indicates the fitted curve.
Figure 3. (a) Schematic representation of the spin-pumping experimental mechanism. (b) Vtotal spectrum of the Td-WTN-4 sample at 4 GHz, with Vs and Va as fitted curves. (c) The Vs of Td-WTN-4 at different frequencies. (d,e) The H0 curves and ΔH versus f curves for Td-WTN-t and NiFe(6). The solid spheres represent the experimentally measured data, while the red line indicates the fitted curve.
Nanomaterials 15 00762 g003
Figure 4. (a) Dependence of the α in Td-WTN-t samples on the thickness of Td-WTe2. (b) The correlation between VSE/ J S and t WTe 2 in Td-WTN-t samples.
Figure 4. (a) Dependence of the α in Td-WTN-t samples on the thickness of Td-WTe2. (b) The correlation between VSE/ J S and t WTe 2 in Td-WTN-t samples.
Nanomaterials 15 00762 g004
Figure 5. (a) Schematic of the ST-FMR test equipment unit and the Td-WTN-4 device. (b) ST-FMR results of Td-WTN-4 at 4 GHz. VS and VA are symmetric and antisymmetric fitted voltages for Vmix. (c) ST-FMR spectrum of Td-WTN-4 in the 4–10 GHz range. (d) ST-FMR spectra of Td-WTN-t (4, 6, 8, 10, and 12 nm) at 4 GHz.
Figure 5. (a) Schematic of the ST-FMR test equipment unit and the Td-WTN-4 device. (b) ST-FMR results of Td-WTN-4 at 4 GHz. VS and VA are symmetric and antisymmetric fitted voltages for Vmix. (c) ST-FMR spectrum of Td-WTN-4 in the 4–10 GHz range. (d) ST-FMR spectra of Td-WTN-t (4, 6, 8, 10, and 12 nm) at 4 GHz.
Nanomaterials 15 00762 g005
Figure 6. (a,b) The angular dependence of the ST-FMR signals with plane magnetic field for VS and VA components of Td-WTN-4 at 4 GHz. (c,d) The angular dependence of the in-plane magnetic field on the VS and VA components of Td-WTN-12 at 4 GHz.
Figure 6. (a,b) The angular dependence of the ST-FMR signals with plane magnetic field for VS and VA components of Td-WTN-4 at 4 GHz. (c,d) The angular dependence of the in-plane magnetic field on the VS and VA components of Td-WTN-12 at 4 GHz.
Nanomaterials 15 00762 g006
Figure 7. (a) The θSH,y and θSH,z of the Td-WTN-t heterostructure vary as functions of t WTe 2 . (b) The σSH,y and σSH,z of the Td-WTN-t heterostructure as a function of t WTe 2 .
Figure 7. (a) The θSH,y and θSH,z of the Td-WTN-t heterostructure vary as functions of t WTe 2 . (b) The σSH,y and σSH,z of the Td-WTN-t heterostructure as a function of t WTe 2 .
Nanomaterials 15 00762 g007
Table 1. The Ms of the Td-WTN-t heterostructure and NiFe(6), along with the various parameters extracted from the spin pumping measurements of both the Td-WTN-t heterostructure and NiFe(6), are presented.
Table 1. The Ms of the Td-WTN-t heterostructure and NiFe(6), along with the various parameters extracted from the spin pumping measurements of both the Td-WTN-t heterostructure and NiFe(6), are presented.
SampleMs (kA/m)Meff (kA/m)αΔα/10−3 g eff /(1019 m−2)JS (A/m2)θSH
NiFe(6)6055220.01089
Td-WTN-45525490.012761.870.4003918.630.0785
Td-WTN-65445470.013572.680.5644965.470.0756
Td-WTN-85374880.015935.041.0486776.040.0744
Td-WTN-105334730.017426.531.3477341.570.0779
Td-WTN-125224540.018357.461.5087558.320.0960
Table 2. Various parameters of the Td-WTN-t samples extracted from the ST-FMR measurements.
Table 2. Various parameters of the Td-WTN-t samples extracted from the ST-FMR measurements.
SampleMeff (kA/m)ασSH, y
(103 ħ/2e
−1m–1)
σSH, z
(103 ħ/2e
−1m–1)
θSH, yθSH, z
Td-WTN-45170.012662.830.090.0500.002
Td-WTN-65100.012864.130.150.0710.003
Td-WTN-84980.015694.350.240.0740.005
Td-WTN-104700.022494.680.580.0770.010
Td-WTN-124410.024264.930.810.0780.013
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zheng, M.; Zhang, W.; Lv, Y.; Liu, Y.; Xiong, R.; Zhang, Z.; Lu, Z. Observation of Thickness-Modulated Out-of-Plane Spin–Orbit Torque in Polycrystalline Few-Layer Td-WTe2 Film. Nanomaterials 2025, 15, 762. https://doi.org/10.3390/nano15100762

AMA Style

Zheng M, Zhang W, Lv Y, Liu Y, Xiong R, Zhang Z, Lu Z. Observation of Thickness-Modulated Out-of-Plane Spin–Orbit Torque in Polycrystalline Few-Layer Td-WTe2 Film. Nanomaterials. 2025; 15(10):762. https://doi.org/10.3390/nano15100762

Chicago/Turabian Style

Zheng, Mingkun, Wancheng Zhang, You Lv, Yong Liu, Rui Xiong, Zhenhua Zhang, and Zhihong Lu. 2025. "Observation of Thickness-Modulated Out-of-Plane Spin–Orbit Torque in Polycrystalline Few-Layer Td-WTe2 Film" Nanomaterials 15, no. 10: 762. https://doi.org/10.3390/nano15100762

APA Style

Zheng, M., Zhang, W., Lv, Y., Liu, Y., Xiong, R., Zhang, Z., & Lu, Z. (2025). Observation of Thickness-Modulated Out-of-Plane Spin–Orbit Torque in Polycrystalline Few-Layer Td-WTe2 Film. Nanomaterials, 15(10), 762. https://doi.org/10.3390/nano15100762

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

Article Metrics

Back to TopTop