Next Article in Journal
Novel MAGL Inhibitors Alleviate LPS-Induced Acute Kidney Injury by Inhibiting NLRP3 Inflammatory Vesicles, Modulating Intestinal Flora, Repairing the Intestinal Barrier, and Interfering with Serum Metabolism
Previous Article in Journal
Effective Reinforcement of Visible Light Photocatalytic and Gas Sensing Characteristics of Nanocrystalline TiO2: Gd-Based Nb and Mo Dopants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress and Prospects in Metallic FexGeTe2 (3 ≤ x ≤ 7) Ferromagnets

1
School of Materials Science and Engineering, Liaocheng University, Liaocheng 252000, China
2
College of Optoelectronic Engineering, Chengdu University of Information Technology, Chengdu 610225, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(21), 7244; https://doi.org/10.3390/molecules28217244
Submission received: 6 September 2023 / Revised: 5 October 2023 / Accepted: 21 October 2023 / Published: 24 October 2023
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
Thermal fluctuations in two-dimensional (2D) isotropy systems at non-zero finite temperatures can destroy the long-range (LR) magnetic order due to the mechanisms addressed in the Mermin-Wanger theory. However, the magnetic anisotropy related to spin–orbit coupling (SOC) may stabilize magnetic order in 2D systems. Very recently, 2D FexGeTe2 (3 ≤ x ≤ 7) with a high Curie temperature (TC) has not only undergone significant developments in terms of synthetic methods and the control of ferromagnetism (FM), but is also being actively explored for applications in various devices. In this review, we introduce six experimental methods, ten ferromagnetic modulation strategies, and four spintronic devices for 2D FexGeTe2 materials. In summary, we outline the challenges and potential research directions in this field.

1. Introduction

The Mermin-Wanger theory [1,2] asserts that thermal fluctuations occur in 2D isotropy systems at non-zero finite temperatures, which destroy the long-range magnetic order (LRMO). Specifically, exchange interactions alone should not generate magnetic order in 2D systems, and magnetic anisotropy [3,4,5] is also needed to maintain the LRMO. Surprisingly, it was found experimentally that low-temperature long-range ferromagnetic order (LRFO) can exist in the Cr2Ge2Te6 monolayer [4] and CrI3 monolayer [5,6]. Soon after, a vast range of 2D magnetic systems, including metallic (Fe3GeTe2 (FGT) [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23]), semiconductors (Cr2Ge2Te6 [4,24,25,26,27,28,29,30,31,32], CrI3 [5,33]), and topological insulators (MnBi2Te4 [34]), were successively implemented to promote the development of spintronics.
Recently, FexGeTe2 (3 ≤ x ≤ 7) has received intense attention as a metallic and high Curie temperature (TC) ferromagnet. Six synthesis methods, including solid-state reaction (SSR) [35,36], chemical vapor transport (CVT) [8,13,37], the flux method [11,21,38,39,40,41,42,43,44,45], exfoliation [14,15,34,46,47,48,49,50], chemical vapor deposition (CVD) [51,52], and molecular beam epitaxy (MBE) [7,53,54,55,56,57,58], have been used to attempt to obtain wafer-scale FexGeTe2 (3 ≤ x ≤ 7) materials with room-temperature ferromagnetism (RTFM). However, the TC of the MBE-prepared FexGeTe2 (3 ≤ x ≤ 7) samples (see Figure 1) ranges from 390 to 530 K, with FGT (TC ≈ 400 K) [56], Fe4GeTe2 (F4GT; TC ≈ 530 K) [59], and Fe5GeTe2 (F5GT; TC ≈ 390 K) [60]. Furthermore, RTFM has also been tuned with ten strategies: Fe stoichiometry [9,39,51,59,61,62,63,64,65], strain engineering [46,48,66,67,68,69,70,71,72,73,74,75], hydrostatic pressure [76,77,78,79,80,81], light control [53,82], electrical control [83,84], proximity effects [56,57,85,86,87,88,89], doping engineering [14,20,38,43,44,62,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106], intercalation [107,108] or irradiation [109], twisting [110,111], and patterning [16]. So far, twisting (see Figure 1) has only regulated the magnetic order theoretically and has not been achieved experimentally. Moreover, four typical devices have also been fabricated based on FGT, magnetic tunnel junctions (MTJ) [112,113,114,115], tunneling spin valves [18,99,116,117], nonlocal spin valves [118], and spin–orbit torque devices [20,119], in order to enrich their physical properties and develop their spintronic applications.
In this review, we introduce the developments in and structures of metallic FexGeTe2 ferromagnets. Subsequently, we summarize six experimental methods, as shown in Figure 1; the early samples prepared using SSR, CVT, and the flux method are mainly bulk single crystals. In addition, magnetic skyrmions in FexGeTe2 are summarized. Finally, we outline the challenges and potential research directions in this field.

2. Crystal Structure of Ferromagnetic FexGeTe2

The Fe3GeTe2 monolayer [46] comprises five atomic layers, as shown in Figure 2A. Specifically, Te atoms are located in the bottom and top layers, while Fe (I) atoms are located in the second and fourth layers. Notably, the intermediate layer is composed of Fe (II) atoms and Ge atoms. The local magnetic moments of Fe atoms for the FGT monolayer were determined by DFT-LDA (density functional theory–local density approximation) to be 1.723 µB and 1.005 µB, with the out-of-plane direction being its easy axis. They may be related to the several partially occupied d-bands passing through the Fermi level. In addition, the number ratio of Fe3+ to Fe2+ in 2D FexGeTe2 [64] is related to the x value, as shown in Figure 2B. When the value of x is 3, the ratio of Fe3+/Fe2+ is 2:1. However, when the x value is 5, only Fe3+ is present (Figure 2C).

3. Synthesis of Metallic FexGeTe2 with FM

3.1. Solid-State Reaction (SSR)

The solid-state reaction (SSR) is an experimental method for preparing bulk FGT crystals. As early as 2006, Deiseroth et al. [35] successfully prepared FGT crystals with hexagonal plates using SSR, which exhibited novel air stability and black metallic properties. Through magnetic testing, it was found that below 230 K, the crystals exhibited FM. Meanwhile, above 230 K, they exhibited Curie–Weiss paramagnetic behavior. After increased annealing, black Fe3−δGeTe2 (0 < δ < 0.3) polycrystalline powders could be easily obtained with SSR. Fe1 and Ge1 atoms are shown in detail in different coordination environments in Figure 3A,B, and their unit cell consist of two layers. The lattice parameters increase monotonically with decreasing δ (it represents the degree of iron deficiency in FGT) (Figure 3C), but, when δ exceeds 0.3, FeTe2 will appear as an impurity phase. Its magnetic phase transition temperature (Figure 3D) is about 240 K. Furthermore, its saturation behavior (see inset in Figure 3D) slows down in high magnetic fields, which is different from ordinary ferromagnets.
In order to obtain large quantities of high-quality FGT single crystals, Li et al. [120] designed a new experimental method of solid-phase sintering followed by recrystallization (Figure 3E–G). The as-grown plate-like sample (~10 g) is a layered single crystal with a smooth and complete surface, and its size can reach up to 8.5 mm. By intercalating sodium into as-grown FGT, Weber et al. [107] raised its TC to 350 K. After intercalation, the sample retained obvious layered features, with edge lengths of a grain size ranging from 10–50 μm.

3.2. Chemical Vapor Transport (CVT)

One main difference from SSR is that CVT often uses iodine [8,13,14,15,17,37,39,61,119] or TeCl4 [12] as the transport agent, as shown in Figure 4. However, the samples obtained via SSR and CVT were both bulk single crystals, as shown in the inset of Figure 4A,B. Previous studies have mainly focused on the magnetic microstructures of quasi-2D FGT. Based on the prediction that FGT monolayer could be mechanically exfoliated [46], soon after, Chu et al. [15] and Zhang et al. [14], respectively, obtained FGT monolayer samples with the assistance of Au film and Al2O3, respectively. Actually, Zhang et al. [121] exfoliated FGT monolayer from the most possible cleaving planes (001), with a thickness of 1.75 nm and a nearest neighbor atomic spacing of 0.338 nm, which was highly consistent with the lattice constant (a = 0.399 nm; c = 1.63 nm) of the FGT crystal. However, thin layer FGT was highly prone to deteriorate in air, and the device fabrication processes (Figure 4C–E) needed to be carried out in a glove box [49]. Notably, many novel physics-related effects, such as patterning-induced RFTM [16], gate-tunable FM [14], and layer-dependent FM [15], have been discovered.

3.3. Flux Growth

The flux method [122,123,124] is commonly used to prepare single crystals. For example, Canfield et al. [122,123] grew a wide variety of single-crystal binary or ternary intermetallic compounds from molten flux solutions. However, the thickness and lateral size of the samples could not be accurately controlled, with mechanical exfoliation still required to obtain thinner samples when fabricating FGT devices.
Recently, Gong et al. [44,45] proposed a universal flux-assisted growth (FAG) method (Figure 5A,B) to synthesize FexGeTe2 and MyFe5−yGeTe2 (M = Co, Ni) nanosheets on various substrates (Figure 5C). In addition, the sample thickness and lateral size (Figure 5D) of FGT could be precisely controlled by the growth temperature (Figure 5E) or cosolvents (Figure 5F). Although the FGT samples with a thickness of 5–10 nm (Figure 5G) were prepared on various substrates, in order to obtain atomically thin materials (ATMs), a confinement environment must be provided through two substrates. Up to 80 layered and non-layered ATMs [45] have also been successfully synthesized using FAG, which provides a new strategy for preparing wafer-scale 2D materials.

3.4. Exfoliation

3.4.1. Mechanical Exfoliation

Conventional mechanical exfoliation [125,126] can cleave thin FGT flakes onto SiO2/Si substrates, but its thinnest thickness is around 4.8 nm. After depositing Au onto SiO2/Si substrate, thinner samples can be obtained, and the Au substrate improves the yield to grow various thin layers of materials, including graphene [127,128], MoS2 [47,129,130,131,132], WSe2 [47,129,133], Bi2Te3 [129], and FGT [15,47]. Nevertheless, only a small amount of material can be exfoliated to a monolayer, which hinders the development of 2D magnetic materials. Notably, an Al2O3-assisted exfoliation method was also designed to produce monolayer FGT [14] and MnBi2Te4 [34] single crystals. When the sample was thinned from bulk to a monolayer, its TC decreased from 180 K to 20 K.

3.4.2. Liquid-Phase Exfoliation

Although many methods including SSR [107], CVT [14,15,17,19,61,119,121,134], flux [39,43,44,45], and MBE [7,53,54,55,56,57,58,59,60] have been used to prepare 2D FGT, an economical method for the large-scale preparation of few- or single-layer FGT nanoflakes is still lacking. As a typical example, Ma et al. [50] developed three-stage sonication-assisted liquid-phase exfoliation (TS-LPE) (Figure 6A) to produce large semiconductive FGT nanoflakes. After ball milling, the sample size and thickness (Figure 6B,C) are reduced by the milling time (Figure 6G), exposing more boundaries. Stirring causes the interlayer spacing to expand (Figure 6C,D), weakening the interlayer force to facilitate detachment and obtain high-integrity nanoflakes. In addition, XRD analysis [135] (Figure 6H,I) reflects the evolution of interlayer spacing. The expansion of interlayer spacing causes the FGT unit cell to move away from the equilibrium state in the c-direction (Figure 6J), making them unstable and prone to spall. In practice, the oxidation on the surface layer altered the electronic structure of the FGT system, making the FGT sample semiconductive and different from the metallic FGT prepared using other methods.

3.5. Chemical Vapor Deposition (CVD)

So far, researchers have mainly used CVT to prepare 2D magnetic bulk single crystals, which are then exfoliated into atomic layers to prepare devices. However, poor control of the number of layers and a limited sample size have hindered the development of 2D magnets. As a typical example, Liu et al. [51] designed a confined space chemical vapor deposition (CS-CVD) method for preparing 2D FGT or F5GT ferromagnets. They found that the optimal growth temperature was 570–580 °C, with an optimal distance of 10 cm between the Fe/Ge precursor and the Te precursor. When the thickness of the F5GT flakes changed from 4 nm to 1 nm, the TC value decreased by 100K. Very recently, Liu et al. [52] also introduced a general competitive-chemical-reaction-controlled CVD method for producing FGT crystals. The sample was a single layer with a grain size of ~50 μm.

3.6. Molecular Beam Epitaxy (MBE)

Wafer-scale single crystalline FGT thin films were grown on various substrates using the molecular beam epitaxial (MBE) [7,54,55,56,57,136] technique. After heterointegration with the topological insulator Bi2Te3 (Figure 7A), the TC of FGT can be increased to 400 K. This enhancement may be related to the interface exchange coupling. Remarkably, when the thickness of F4GT decreases, its TC is increased from 270 K to 530 K (Figure 7B). For F4GT thin films with a thickness of 10 nm, increasing the dosage of Fe can enhance their TC (Figure 7C). Although MBE can alone prepare wafer-scale FexGeTe2 (3 ≤ x ≤ 7) materials with RTFM, this method requires a high vacuum environment, which makes it expensive and limits its industrial applications.

4. Controlling FM in Metallic FexGeTe2

4.1. Fe Stoichiometry

In this research area, the earliest discovery was that the FM in polycrystalline FGT bulk structures [9] was related to the Fe content (Figure 8). The higher the Fe content, the larger the lattice constant of the a-axis and the smaller the lattice constant of the c-axis (Figure 8A). Single crystal samples show similar results to the polycrystalline samples. Moreover, the TC (Figure 8B) and MS decreased with the decrease in Fe content. Subsequently, ferromagnetic F4GT [44,59,61] and F5GT [39,42,44,51,62,137,138,139,140,141,142,143,144] materials were also obtained in experiments.
However, most previous reports have focused on FGT materials with a single Fe stoichiometry, and there have been few studies on FexGeTe2 materials using the same experimental method. In addition, theoretical calculations [63] revealed that as the Fe content increased, the interlayer gap gradually increased, and the magnetic anisotropy of its monolayer changed from out-of-plane (FGT) to in-plane (F4GT and F5GT).
Although previous reports have made significant progress in 2D FexGeTe2 systems, the mediation of magnetic anisotropy and their magnetic nature remains unresolved. Very recently, Liu et al. [64] preliminary reported a valence-dependent magnetic exchange model to explain the complex magnetic phase in FexGeTe2 systems. Furthermore, the magnetic moment and MAE (Figure 9A) were almost linearly correlated with Fe2+/Fe3+. Specifically, Fe3+ had a greater impact on magnetism, reducing the magnetic anisotropy energy in F5GT. Based on MAE and J (it represents the exchange coupling constant), the TC could be estimated using the 2D Heisenberg model. When x was greater than 4, the TC was much higher than RT (Figure 9B).
Moreover, the results obtained via different calculation methods [65] exhibit significant differences, especially when compared with experimental results. As x increases, its easy axis direction (shown by the black arrows) and the highest exchange interaction change from out-plane to in-plane (Figure 9C–F). Moreover, there were significant differences in the magnitude of the exchange interaction obtained via different calculation methods (GGA + DMFT, GGA, and GGA + U); specifically, the results calculated using GGA + U were overestimated. In addition, MAE exhibits a similar evolution from out-plane to in-plane. However, for different F5GT (UUU or UDU) configurations, the calculation results varied significantly. It was obvious that the TC calculated via GGA + DMFT was underestimated, while the result calculated by GGA was overestimated, compared to Figure 9B.

4.2. Strain Engineering

Strain engineering is an efficient strategy for modulating the FM of 2D materials [67,68,145]. However, previous theoretical works have focused on applying strain to FGT supercells by changing the lattice constants [46,70,71,73,146] and calculating the exchange coupling, magnetic anisotropy, and magnetic moment of strain through ab initio DFT. Furthermore, the TC could be estimated according to mean field theory (MFT) [10,61,65,147,148,149], random phase approximation (PRA) [147,149], or Monte Carlo (MC) [148,149,150,151] simulation. Recently, Miao et al. [48] and Yan et al. [72] loaded FGT nanoflakes into a three-point-bending experimental setup and applied uniaxial tensile strain to the sample on a polyimide (PI) or polyvinyl alcohol (PVA)/polyethylene terephthalate (PET) flexible polymer substrate by moving the needle. Moreover, the magnitude of the applied strain could be calculated using the following formula [48,72,152]:
ε = T 2 R
where T and R are the film thickness and bending radius, respectively. Surprisingly, after 0.32% strain was applied, the coercivity increased by 150% [48], a far greater increase than the improvement in HC generated by other traditional magnetic materials [153]. Most notably, its TC could be increased to 400 K [72] via uniaxial strain, further promoting the development of wearable spintronic devices with low-energy consumption.

4.3. Hydrostatic Pressure

Tuning the exchange coupling and magneto-crystalline anisotropy by applying hydrostatic pressure is another commonly used method for regulating 2D magnetism, which has been achieved in Cr2Gr2Te6 [154,155], CrI3 [33,156], and FGT [77,78,79] systems (Figure 10). The ferromagnetic evolution of FGT nanosheets under different pressures can be revealed through in situ magnetic circular dichroism (MCD) spectroscopy (Figure 10B). Furthermore, the magnetic hysteresis loop at 30 K exhibited a rectangular shape below 7 GPa, while its loop presented an eight-shaped skewed shape above 7.3 GPa. Moreover, TC increases as the pressure further decreases (Figure 10C), which may be related to the strengthening of the exchange interactions.
As another typical example, TC and the magnetic moment also increased with the decrease in the hydrostatic pressure (Figure 10D,E). It is evident that the increasing pressure reduced the length of the Fe–Fe bond, which inhibited magnetization through modification of the exchange interactions. In addition, a monotonic relationship between TC, the magnetic moment, and pressure was also found in the Fe-deficient FGT sample [77], similar to the FGT system. Overall, pressure modifies the metallic form of Fe3−xGeTe2 to its nonmetallic form.

4.4. Light Control

The continuous modulation of monolayer transition-metal dichalcogenides (TMDs) without intrinsic magnetism, including MoS2 [157], WS2 [157], and WSe2 [158], has been achieved using the optical approach. Recently, Tengdin et al. [82] demonstrated that spin polarization was transferred from Mn sublattices to Co on the Heusler compound Co2MnGe via femtosecond laser pulse, which is closely related to the wave function of electrons before and after being excited by light. The ultrafast spin transfer caused by the instantaneous incident light on the material does not only occur in Co2MnGe, but is also a common feature of many materials. Notably, Xu et al. [53] reported that the magnetic anisotropy energy (MAE) and TC were mediated with a femtosecond laser pulse, as shown in Figure 11. The optical doping effect alters the electronic structure of FGT, thereby affecting exchange interactions, TC, and MAE. According to Figure 11B, the TC of FGT was estimated to be ~200 K. Under the excitation of a femtosecond laser, electrons transitioned from an occupied state to an unoccupied state, causing the Fermi level EF to shift downwards and crossing the enhanced density of states (DOS) shown in Figure 11D. Furthermore, some clear magnetic hysteresis loops at room temperature (RT) can be observed in FGT samples with different thicknesses, according to Polar-MOKE measurements (Figure 11C). The TC of FGT can be increased to above RT through light control, providing many opportunities for the development of spintronic applications for 2D magnets.

4.5. Electrical Control

Previous studies have shown that electric fields modify the magnetism of metal films [159,160,161] and Fe/MgO junctions [162] by influencing the behavior of the electrons. Recently, Wang et al. [83] calculated the effect of the electric field on the magnetic anisotropy of the FGT monolayer, as shown in Figure 12A–C. The effect of orbital splitting caused by electron doping on magnetic anisotropy was more pronounced; meanwhile, the influence of hole doping related to orbital occupation was relatively weak. In addition, the change in magnetic anisotropy (Figure 12C) was more obvious in the single-gate configuration (Figure 12B).
Additionally, the generation of negative differential conductance (NDC) [84] can also be driven by a local electric field in FGT (Figure 12D). Furthermore, the three peaks in the Fe d orbits underwent significant shifts under the electric field. As the electric field was enhanced, the off-plane FM of FGT weakened, resulting in a decrease in MAE. Remarkably, in single-layer FGT, the electric field induces charge transfer in the FGT monolayer in the field direction. Therefore, applying an electric field has become an effective way to mediate 2D FM.

4.6. Proximity Effects

Proximity effects [85,163,164,165,166,167] are another dominant area of the research into 2D materials. For example, by using 2D magnetic materials adjacent to a bulk semiconductor substrate [168] or 2D materials with strong spin–orbit coupling [169], their magnetism can be enhanced. Intriguingly, Zhang et al. [85] fabricated antiferromagnetic FePS3(FPS)/ferromagnetic Fe3GeTe2(FGT) heterostructures and detected the enhancement of TC and HC through proximity coupling effects. Furthermore, FPS/FGT/FPS exhibits a slightly different modulation of HC compared to FPS/FGT, which is related to AFM-FM coupling. Moreover, the long-range magnetic order induced by topology triggered by femtosecond laser pulses [57] could also be maintained at room temperature.
The aforementioned methods for mediating magnetism require additional equipment, such as a three-point-bending experimental setup [48,72], DAC, [78], femtosecond laser excitations [53], and the scanning tunneling microscopy (STM) tip [84]. Directly exfoliating FGT nanosheets [88] onto different substrates could also tailor magnetism, as shown in Figure 13. Intriguingly, FGT samples with different thicknesses all exhibited magnetism, while samples on different substrates exhibited different TC, indicating that the substrate has a modulation effect on TC values. Furthermore, the lattice distortion and charge redistribution at the interface were related to substrate-induced FM, and this mechanism requires further exploration. Actually, substrates did not only affect the growth of 2D materials, but also determined their performance [170,171].

4.7. Doping Engineering

4.7.1. Doping with 3d Transition-Metals

Doping 3d-transition metal atoms is an effective strategy for controlling magnetism [66,69,90,172,173,174,175,176]. Theoretical calculations have shown that almost all 3d-transition metal atoms (except for Co atoms) [97] are more inclined to replace Fe1 atoms (Figure 2A). The charge transfer generated by doping atoms weakens the magnetic moment of Fe atoms, while the weakening effect of Fe1 atomic magnetic moment is more significant. However, the magnetism increases after doping with Co atoms, which may be related to the shrinking of the a-axis lattice constant. In experiments, doping 3d-transition metal atoms in bulk single-crystal samples were usually achieved via CVT [93] or self-flux [38]. Doping Ni atoms suppressed the ferromagnetic order, which rapidly decreased with the increase in the doping amount. The TC decreased from 212 K to 50 K, and after reaching 0.44, the magnetic moment remained almost constant. Furthermore, the long-range magnetic order was suppressed and subsequently transformed into a glassy magnetic phase (Figure 14). However, doping Co atoms may cause an increase in HC and the appearance of hard magnetic phases; this is related to the movement of pinned domain walls [8].
Bulk F5GT single crystals were also doped with Co atoms via CVT [95,96], as shown in Figure 15. As the amount of Co used for doping increases, it can drive the evolution of the lattice and of magnetism (Figure 15B). However, the nominal doping concentration was slightly different from the measured one, with only a specific concentration being more consistent. Afterward, Co atoms were doped into the lattice, resulting in a slight increase in their interlayer spacing. Indeed, a phase transition from FM to AFM occurred at high doses of doping in Figure 15C. However, Tian et al. [96] found that doping with 20% Co could increase its TC to 337 K and induce complex magnetic phase transitions at higher Co doping levels. Furthermore, hexagonal 2D CoyFe5−yGeTe2 (Figure 15E) and NiyFe5−yGeTe2 nanoflakes (Figure 15E) were prepared via flux-assisted growth (Figure 15D) [44]. Various elements in the nanosheets were evenly distributed via energy dispersive spectroscopy (EDS) elements mapping, and there were significant differences in the energy spectra of samples doped with different amounts of Co, with a doping amount of up to 66.7%. Furthermore, the doping of Co atoms caused a decrease in TC, and as the doping amount increased, its TC decreased even further (Figure 15F). In addition, the magnetic anisotropy underwent significant changes. Similarly, Ni doping can also cause a decrease in TC. In other words, the higher the content of Fe, the higher its TC.

4.7.2. Doping with Non-Metallic Atoms

Not only can Fe atoms be substituted with Co or Ni atoms [38,44,93,95,96,97], but doping can also be achieved by replacing Ge atoms with As atoms [92,98]. The doping of As atoms caused a decrease in the a-axis lattice constant and an increase in the c-axis lattice constant, thereby reducing the density of spin states below the Fermi level, resulting in a decrease in TC [92]. Furthermore, its MS decreased linearly with the increase in the doping amount in polycrystalline Fe3−yGe1−xAsxTe2 (0 ≤ x ≤ 0.85). Similarly, the expansion of the F5GT unit cell [98] in the c-axis direction and the contraction in the ab plane was also observed after doping with the As atom. In addition, its TC and MS decreased in polycrystalline Fe5Ge1−yAsyTe2 (0 ≤ y ≤ 1), a phenomenon similar to that observed in the Fe3-yGe1−xAsxTe2 (0 ≤ x ≤ 0.85) samples. Moreover, the stacking disorder caused by the local AFM coupling can reduce its MS.

4.7.3. Electron Doping

Remarkably, Deng et al. [14] found that FGT devices could be operating in ionic gates, which provides a new approach for mediating 2D FM. Although they did not fully explain the relationship between it’s ferromagnetism and electron doping, the importance of this strategy was acknowledged. Soon after, gate-control was implemented to regulate magnetic resistance [99], magnetic phases [62,101], and interlayer coupling [100,102,177]. Furthermore, the TC and HC in FGT flakes [101] were decreased after Li+ doping from lithium-ion-conducting glass-ceramics (LICGC). In addition, electron doping influenced the Fe–Ge plane in the middle of the FGT monolayer, weakening it’s resistance and enhancing it’s TC [105]. As a typical example, the modulation of interlayer coupling was achieved by fabricating FGT hall devices on solid-state proton conductors (Figure 16A). Furthermore, they discovered a clear exchange bias (EB) when changing the gate voltage (Figure 16B), which may be related to the presence of an AFM phase at low temperatures (Figure 16C). However, a random exchange bias (Figure 16D) occurs after applying a higher gate voltage. Furthermore, the exchange bias and coercivity undergo a complex evolution with the measurement times, but there have also been cases of small EB values and large coercivity. In addition, the type of AFM–FM interface coupling determines the positive and negative exchange bias (Figure 16E).
Additionally, Tan et al. [62] achieved the high-electron-concentration doping of F5GT through a solid proton conductor (Figure 17A,B). When a positive bias voltage was applied, the transport properties were unchanged. When a negative bias voltage was applied, there was a significant change in the transport properties caused by proton intercalation, especially when it reached −5 V, and the anomalous Hall loop disappeared (Figure 17C), accompanied by the appearance of a magnetic phase transition (Figure 17D,E). Moreover, different calculation methods have revealed that electron doping achieves a reversal of Hall conductivity and phase transition (Figure 17E). Therefore, electron doping or protonic gating indeed represents an efficient method of controlling magnetic phase transitions. Furthermore, Tang et al. [102] found that magnetic anisotropy in F5GT was continuously mediated by electrolyte gating. Moreover, the screening effect of itinerant electrons drives magnetic anisotropy to switch from an off-plane easy axis to an in-plane easy axis.

4.7.4. Hole Doping

Inspired by the gate-mediated RTFM in FGT thin flakes [14], many attempts have been made to control its ferromagnetism through hole [43,94,106] or electron [43,94,105] doping. In particular, the magnetic anisotropy in exfoliated Fe2.75GeTe2 flakes (Figure 18A,B) was inhibited by hole doping, resulting in a decrease in HC (Figure 18C). The magnetic anisotropy could undergo a 93% attenuation, but the change in the magnetic moment was very small, as shown in Figure 18D. Furthermore, the electronic structure of Fe2.75GeTe2 single crystals changes due to hole doping, causing significant changes in magnetic anisotropy. In addition, another report [94] suggested that hole doping was beneficial for maintaining the long-range ferromagnetic order.
Remarkably, the intrinsic Fe vacancies [106] were probed via STM, as presented in Figure 19A–D. The peak near 20 mV originated from the Kondo lattice [106,121], and this Fe vacancy is called a Kondo hole (Figure 19C,D). Hole doping elevated the energy band, and the Fermi surface of FGT was shifted towards a lower energy level. After the formation of Fe vacancies at the Fe (II) site, the magnetic moment near the Fe (I) site decreased (Figure 19E,F), accompanied by the appearance of a higher charge density. Furthermore, the introduction of Fe vacancies reduced the magnetic moment near them, further strengthened the Kondo screening effect, and, thus, weakened the magnetism, as detailed in Figure 19G,H. The Kondo holes can affect the charge distribution of their own sites, and converting them into momentum space has a more significant impact. In other words, hole doping weakened the ferromagnetism of FGT.

4.8. Intercalation or Irradiation

Recently, inserting sodium into Fe2.78GeTe2 powders [107] can raise its TC to ~300 K, as shown in Figure 20. After intercalating Na, more exposed edges appeared, and their layered features remained unchanged in a single crystal structure (Figure 20A,B). More specifically, the Fe, Ge, and Te elements (Figure 20C) were evenly distributed in the sample, while the inserted Na was concentrated at the edge. A phase transition occurred from PM (Fe2.78GeTe2) to FM (NaFe2.78GeTe2) at 200 K (Figure 20D), and the MS was enhanced (Figure 20E). Furthermore, the magnetic hysteresis loops are also measured at 350 K. Notably, impurity phases, such as Fe or Fe2−xGe, dominated the RTFM in the NaFe2.78GeTe2 samples. Alternatively, the TC and exchange bias could be mediated with Fe-intercalation [108], which induces magnetic order by reinforcing magnetic coupling. However, the detected TeGe antisite defects had no modulation effect on the TC of different samples. Thus, Na intercalation provides a novel strategy for enhancing TC, which is related to the tensile strain.
Inspired by pattern-induced ferromagnetism [16], Yang et al. [109] also improved the TC of FGT to 450 K via Ga irradiation. Irradiation induces an amorphous structure on the surface of the sample, which facilitates the formation of magnetic vortex states in FGT micropatterns. Moreover, irradiation causes a magnetization transition.

4.9. Twisting

Twisting 2D materials can introduce some novel properties, such as magnetism [178,179] and superconductivity [180], which trigger the interaction topology with magnetism in 2D ferromagnets, resulting in the formation of skyrmions [181,182] or magnons [183,184] in the twisting system. In fact, the stacking order directly affects the magnetism of bilayer CrI3 by changing the crystal structure [178] or interlayer magnetic coupling [178]. Surprisingly, the magnetism was obtained in double bilayer CrI3 [185] at small twist angles. Although the phase transition from AFM to FM has been theoretically achieved in twist-stacking bilayer FGT [110,179], it has not yet been experimentally achieved [186].

4.10. Patterning

Magnetic domain patterns on FGT surfaces can be modulated with various mechanisms [8,13,187,188], one of which is the phase transition from FM to AFM related to interlayer coupling [13]. The photoemission electron microscopy (PEEM) image in Figure 21A–E clearly shows the magnetic domain structure of FGT nanosheets, and the stripe domain structure disappears after reaching the TC of 230 K. After patterning the FGT sample into diamond and rectangular shapes using a focused ion beam (FIB), striped magnetic domain structures, similar to those in the unpatterned FGT (Figure 21B,C), were also observed, as shown in Figure 21D. However, the striped domain structure did not completely disappear and was significantly weakened at 230 K (Figure 21C).
In addition, there were only in-plane magnetic domains in the patterned FGT at 240 K (Figure 19D); they disappeared at ~370 K, indicating that the TC of bulk FGT was increased to 370 K. Notably, a novel magnetic vortex state or magnetic multidomain state was developed in patterned FGT (Figure 19D) at 300 K. Furthermore, spin reorientation occurred with increasing temperatures (Figure 19E). When using FIB to construct FGT patterns, Ga ions were unintentionally implanted into the sample, which may have caused the increase in TC; however, to date, this hypothesis remains speculative.

5. Band Structure of Ferromagnetic FexGeTe2

Like its bulk form, the FGT monolayer is metallic, as shown in Figure 22A–C. Its band structures near the Fermi level can mainly be attributed to the contribution of the Fe 3d orbitals. Moreover, it was confirmed that the FGT monolayer has the itinerant FM order according to Stoner’s criterion [46,189]. Remarkably, the Stoner model related to itinerant electrons can be used to better elucidate the spontaneous magnetization in most 2D metallic ferromagnets. In addition, the electronic band structures of all the FexGeTe2 systems are metallic (Figure 22D–G), similar to the FGT monolayer.
Furthermore, the F4GT and F5GT bilayer [190] have band structures similar to the FGT bulk and FGT monolayer, exhibiting metallic magnetic properties. However, there is a significant difference in the polarizability of F4GT and F5GT near the Fermi level, which leads to their different transport characteristics. The unique nature of FGT gives its related devices many advantages, including nonvolatility, low reversal magnetic field, and the magnetic field reversal FexGeTe2 electrodes through the spin-polarized current.

6. FexGeTe2-based Devices

To the best of our knowledge, four typical devices have been constructed based on metallic FexGeTe2 ferromagnets, including magnetic tunnel junctions (MTJ) [112,113,114,115,136], tunneling spin valves [18,99,116,117], nonlocal spin valves [118] and spin–orbit torque (SOT) devices [20,119,191]. As shown in Figure 23A,B, a nonlinear behavior originating from tunneling characteristics [115] was exhibited in the IV curve. Furthermore, a typical spin-valve behavior was also identified in the hysteresis loops (Figure 23C,D). After applying a specific voltage, the spin-transfer torque (STT) generated by the current caused the bottom FGT electrode to switch, which was closely related to MAE.
Regarding another typical device, Wang et al. [18] observed tunneling spin-valve behavior (Figure 24A,B) in an FGT/hBN/FGT heterostructure (Figure 25A). Its TMR reached up to 160%. The spin of the tunneling electrons created a nonlinear bias-dependent IV curve related to Figure 24C. As the bias voltage increased, TMR exhibited a highly significant attenuation, as shown in Figure 24D. The inelastic tunneling channel related to bias led to spin relaxation, which may suppress TMR signals.
After applying a voltage to FGT/Pt hybrid devices (Figure 25A) [20], a current was generated between FGT and Pt, forming spin–orbit torques in Pt. Furthermore, a hard magnetic loop (Figure 25B) similar to an FGT device has also been observed in FGT/Pt devices. As the applied in-plane magnetic field Hx was increased, its transition current decreased (Figure 25C–F), regardless of the direction of Hx. This switch was related to the magnetic domain and domain walls. Moreover, the low switching current of the FGT monolayer was beneficial for exploring more effective devices. Very recently, Wang et al. [191] increased the TC of FGT to RT by the topological insulator Bi2Te3 and achieved the SOT-driven magnetization switching at RT.
More significantly, a nonlocal spin valve device [118] that can operate at RT has also been successfully constructed. In addition, highly efficient spin manipulations were observed at the F5GT/graphene interface.

7. Magnetic Skyrmions in Metallic FexGeTe2

As topological magnetic materials, the spin textures in FexGeTe2 are also regulated by the Fe stoichiometry. For instance, Bloch-type skyrmion bubbles [192] were observed in FGT by using Lorentz transmission electron microscopy (LTEM). However, in another study [193], Néel-type skyrmions were reported, which indicated the existence of Dzyaloshinskii-Moriya interaction (DMI) in FGT. Generally, FGT was considered as a centrosymmetric material in the space group P63/mmc [35], which should not possess the asymmetric DMI. A possible interpretation is that the asymmetric interface between the pristine and oxidized FGT may induce an interfacial DMI [193].
More recently, a comprehensive study was implemented to answer this question more clearly [194]. It is found that the crystal structure of Fe3−xGeTe2 can be tuned by the Fe stoichiometry, that is, Fe3−xGeTe2 changes from centrosymmetric P63/mmc space group to non-centrosymmetric P63mc space group when x is larger than ~0.2 [194]. Notably, the P63mc space group allows an in-plane DMI in Fe2.74GeTe2, leading to the formation of Néel-type skyrmions. Whereas in Fe3.00GeTe2, the centrosymmetric P63/mmc structure forbids the DMI. Thus, Bloch-type skyrmion bubbles were observed when the dipole–dipole interaction and magnetic anisotropy obtained a delicate balance.
Moreover, in the ab-plane of F5GT, Bloch-type skyrmion bubbles with unconventional helicity polarization were observed [195,196], which shows special application potential for data storage and processing using the helicity freedom of skyrmions in the future. Additionally, in the ac and bc-plane of F5GT, (anti)meron chains were observed [197]. The tunable topological spin textures pave a promising way for constructing novel spintronic devices using FexGeTe2.

8. Outlook

In this review, we introduced the developments in and structures of 2D metallic FexGeTe2 ferromagnets. Then, we summarized six experimental methods, ten FM modulation strategies, and four typical spintronic devices that use 2D FexGeTe2 materials. Finally, we outlined the challenges and potential research directions in this field.
The samples prepared using the SSR, CVT, and flux methods were mainly bulk single crystals. Since the successful exfoliation of mono- or few-layer FGT, 2D FexGeTe2 has only been obtained experimentally. Very recently, CVD and MBE have also been used to prepare these materials. However, the lateral dimension of samples obtained via CVD was small and it was very difficult to accurately control the number of layers. Although MBE can be used to prepare wafer-scale materials and achieve 2D mode growth, it requires a high vacuum environment and is costly. So far, there is still no cost-effective method for preparing wafer-scale materials with controllable layers. In addition, new experimental methods, such as substitution reactions [198], are also worth further exploration.
Although ten strategies for regulating ferromagnetism have been proposed, not all of them have been achieved experimentally. Among these methods, twisting was only achieved through theoretical regulation. It is thus necessary to develop more strategies that could be executed experimentally to obtain RFTM in 2D FexGeTe2 materials. Furthermore, it would also be highly valuable to obtain FGT materials with good air stability [199] via new methods, as this would provide more convenient conditions for fabricating devices [200].

Author Contributions

Conceptualization, H.R.; writing—original draft preparation, H.R.; writing—review and editing, H.R. and M.L.; supervision, H.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Shandong Province Natural Science Foundation (ZR2021MA042 to H.R.) and the Doctoral Scientific Research Foundation of Liaocheng University (318052054 to H.R.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

2DTwo-dimensional
3DThree-dimensional
AFMAntiferromagnetism
ATMsAtomically thin materials
CGTCr2Ge2Te6
CS-CVDConfined space chemical vapor deposition
CVDChemical vapor deposition
CVTChemical vapor transport
DFTDensity functional theory
DMIDzyaloshinskii–Moriya interaction
DOSDensity of states
EBExchange-bias
EDXSEnergy-dispersive X-ray spectroscopy
EDSEnergy dispersive spectroscopy
EDXX-ray spectroscopy
FAGFlux-assisted growth
FGTFe3GeTe2
F4GTFe4GeTe2
F5GTFe5GeTe2
FIBFocused ion beam
FMFerromagnetism
FPSFePS3
GGAGeneralized-gradient approximation
HAADFHigh-angle annular dark-field
HRTEMHigh resolution transmission electron microscopy
LDALocal density approximation
LDA + ULocal density approximation plus Hubbard U
LICGCLithium-ion conducting glass-ceramics
LRLong-range
LRFOLong-range ferromagnetic order
LRMOLong-range magnetic order
LTEMLorentz transmission electron microscopy
MAEMagnetic anisotropy energy
MBEMolecular beam epitaxy
MCMonte Carlo
MCDMagnetic circular dichroism
MFTMean field theory
MOKEMagneto-optical Kerr effect
MTJMagnetic tunnel junctions
NDCNegative differential conductance
PBEPerdew–Burke–Ernzerhof
PEEMPhotoemission electron microscopy
PETPolyethyleneterephthalate
RHEEDReflection high-energy electron diffraction
PIPolyimide
PMParamagnetism
PRARandom phase approximation
PVApolyvinyl alcohol
p-xrdPowder X-ray diffraction
RTRoom temperature
RTFMRoom-temperature ferromagnetism
SAEDSelected area electron diffraction
Sc-xrdSingle-crystal X-ray
SEMScanning electron microscopy
SQUIDSuperconducting quantum interference device magnetometry
SOCSpin-orbit coupling
SSRSolid-state reaction
STEMScanning transmission electron microscopy
STMScanning tunneling microscopy
STTSpin-transfer torque
TMDsTransition-metal dichalcogenides
TMRTunneling magnetoresistance
TS-LPEThree-stage sonication-assisted liquid-phase exfoliation
XRDX-ray diffraction
RBending radius
TFilm thickness
TCCurie temperature
JExchange coupling constant
Jxthe injected current density
ΔETotal energy difference
εthe applied strain

References

  1. Mermin, N.D.; Wagner, H. Absence of ferromagnetism or antiferromagnetism in one- or two-dimensional isotropic Heisenberg models. Phys. Rev. Lett. 1966, 17, 1133–1136. [Google Scholar] [CrossRef]
  2. Hohenberg, P.C. Existence of long-range order in one and two dimensions. Phys. Rev. 1967, 158, 383–386. [Google Scholar] [CrossRef]
  3. Miller, J. Ferromagnetism found in two-dimensional materials. Phys. Today 2017, 70, 16–19. [Google Scholar] [CrossRef]
  4. Gong, C.; Li, L.; Li, Z.L.; Ji, H.W.; Stern, A.; Xia, Y.; Cao, T.; Bao, W.; Wang, C.Z.; Wang, Y.A.; et al. Discovery of intrinsic ferromagnetism in two-dimensional van der Waals crystals. Nature 2017, 546, 265–269. [Google Scholar] [CrossRef]
  5. Huang, B.; Clark, G.; Navarro-Moratalla, E.; Klein, D.R.; Cheng, R.; Seyler, K.L.; Zhong, D.; Schmidgall, E.; McGuire, M.A.; Cobden, D.H.; et al. Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit. Nature 2017, 546, 270–273. [Google Scholar] [CrossRef] [PubMed]
  6. Jiang, S.W.; Li, L.Z.; Wang, Z.F.; Mak, K.F.; Jie Shan, J. Controlling magnetism in 2D CrI3 by electrostatic doping. Nat. Nanotechnol. 2018, 13, 549–553. [Google Scholar] [CrossRef]
  7. Liu, S.S.; Yuan, X.; Zou, Y.C.; Sheng, Y.; Huang, C.; Zhang, E.Z.; Ling, J.W.; Liu, Y.W.; Wang, W.Y.; Zhang, C.; et al. Wafer-scale two-dimensional ferromagnetic Fe3GeTe2 thin films grown by molecular beam epitaxy. npj 2D Mater. Appl. 2017, 1, 30. [Google Scholar] [CrossRef]
  8. Leon-Brito, N.; Bauer, E.D.; Ronning, F.; Thompson, J.D.; Movshovich, R. Magnetic microstructure and magnetic properties of uniaxial itinerant ferromagnet Fe3GeTe2. J. Appl. Phys. 2016, 120, 083903. [Google Scholar] [CrossRef]
  9. May, A.F.; Calder, S.; Cantoni, C.; Cao, H.B.; McGuire, M.A. Magnetic structure and phase stability of the van der Waals bonded ferromagnet Fe3−XGeTe2. Phys. Rev. B 2016, 93, 014411. [Google Scholar] [CrossRef]
  10. Zhu, J.X.; Janoschek, M.; Chaves, D.S.; Cezar, J.C.; Durakiewicz, T.; Ronning, F.; Sassa, Y.; Mansson, M.; Scott, B.L.; Wakeham, N.; et al. Electronic correlation and magnetism in the ferromagnetic metal Fe3GeTe2. Phys. Rev. B 2016, 93, 144404. [Google Scholar] [CrossRef]
  11. Liu, Y.; Ivanovski, V.N.; Petrovic, C. Critical behavior of the van der Waals bonded ferromagnet Fe3−XGeTe2. Phys. Rev. B 2017, 96, 144429. [Google Scholar] [CrossRef]
  12. Wang, Y.H.; Xian, C.; Wang, J.; Liu, B.J.; Ling, L.S.; Zhang, L.; Cao, L.; Qu, Z.; Xiong, Y.M. Anisotropic anomalous hall effect in triangular itinerant ferromagnet Fe3GeTe2. Phys. Rev. B 2017, 96, 134428. [Google Scholar] [CrossRef]
  13. Yi, J.Y.; Zhuang, H.L.; Zou, Q.; Wu, Z.M.; Cao, G.X.; Tang, S.W.; Calder, S.A.; Kent, P.R.C.; Mandrus, D.; Gai, Z. Competing antiferromagnetism in a quasi-2D itinerant ferromagnet: Fe3GeTe2. 2D Mater. 2017, 4, 011005. [Google Scholar] [CrossRef]
  14. Deng, Y.J.; Yu, Y.J.; Song, Y.C.; Zhang, J.Z.; Wang, N.Z.; Sun, Z.Y.; Yi, Y.F.; Wu, Y.Z.; Wu, S.W.; Zhu, J.Y.; et al. Gate-tunable room-temperature ferromagnetism in two-dimensional Fe3GeTe2. Nature 2018, 563, 94–99. [Google Scholar] [CrossRef]
  15. Fei, Z.Y.; Huang, B.; Malinowski, P.; Wang, W.B.; Song, T.C.; Sanchez, J.; Yao, W.; Xiao, D.; Zhu, X.Y.; May, A.F.; et al. Two-dimensional itinerant ferromagnetism in atomically thin Fe3GeTe2. Nat. Mater. 2018, 17, 778–782. [Google Scholar] [CrossRef]
  16. Li, Q.; Yang, M.M.; Gong, C.; Chopdekar, R.V.; N’Diaye, A.T.; Turner, J.; Chen, G.; Scholl, A.; Shafer, P.; Arenholz, E.; et al. Patterning-induced ferromagnetism of Fe3GeTe2 van der Waals materials beyond room temperature. Nano Lett. 2018, 18, 5974–5980. [Google Scholar] [CrossRef]
  17. Tan, C.; Lee, J.; Jung, S.G.; Park, T.; Albarakati, S.; Partridge, J.; Field, M.R.; McCulloch, D.G.; Wang, L.; Lee, C. Hard magnetic properties in nanoflake van der Waals Fe3GeTe2. Nat. Commun. 2018, 9, 1554. [Google Scholar] [CrossRef]
  18. Wang, Z.; Sapkota, D.; Taniguchi, T.; Watanabe, K.; Mandrus, D.; Morpurgo, A.F. Tunneling spin valves based on Fe3GeTe2/hBN/Fe3GeTe2 van der Waals heterostructures. Nano Lett. 2018, 18, 4303–4308. [Google Scholar] [CrossRef]
  19. Albarakati, S.; Tan, C.; Chen, Z.J.; Partridge, J.G.; Zheng, G.L.; Farrar, L.; Mayes, E.L.H.; Field, M.R.; Lee, C.G.; Wang, Y.H.; et al. Antisymmetric magnetoresistance in van der Waals Fe3GeTe2/graphene/Fe3GeTe2 trilayer heterostructures. Sci. Adv. 2019, 5, eaaw0409. [Google Scholar] [CrossRef]
  20. Alghamdi, M.; Lohmann, M.; Li, J.X.; Jothi, P.R.; Shao, Q.M.; Aldosary, M.; Su, T.; Fokwa, B.P.T.; Shi, J. Highly efficient spin-orbit torque and switching of layered ferromagnet Fe3GeTe2. Nano Lett. 2019, 19, 4400–4405. [Google Scholar] [CrossRef]
  21. Calder, S.; Kolesnikov, A.I.; May, A.F. Magnetic excitations in the quasi-two-dimensional ferromagnet Fe3GeTe2 measured with inelastic neutron scattering. Phys. Rev. B 2019, 99, 094423. [Google Scholar] [CrossRef]
  22. Johansen, O.; Risinggard, V.; Sudbo, A.; Linder, J.; Brataas, A. Current control of magnetism in two-dimensional Fe3GeTe2. Phys. Rev. Lett. 2019, 122, 217203. [Google Scholar] [CrossRef]
  23. Kim, D.; Park, S.; Lee, J.; Yoon, J.; Joo, S.; Kim, T.; Min, K.J.; Park, S.Y.; Kim, C.; Moon, K.W.; et al. Antiferromagnetic coupling of van der waals ferromagnetic Fe3GeTe2. Nanotechnology 2019, 30, 245701. [Google Scholar] [CrossRef]
  24. Idzuchi, H.; Llacsahuanga Allcca, A.E.; Pan, X.C.; Tanigaki, K.; Chen, Y.P. Increased curie temperature and enhanced perpendicular magneto anisotropy of Cr2Ge2Te6/NiO heterostructures. Appl. Phys. Lett. 2019, 115, 232403. [Google Scholar] [CrossRef]
  25. Khan, I.; Hong, J.S. High curie temperature and strain-induced semiconductor-metal transition with spin reorientation transition in 2D CrPbTe3 monolayer. Nanotechnology 2020, 31, 195704. [Google Scholar] [CrossRef] [PubMed]
  26. Selter, S.; Bastien, G.; Wolter, A.U.B.; Aswartham, S.; Büchner, B. Magnetic anisotropy and low-field magnetic phase diagram of the quasi-two-dimensional ferromagnet Cr2Ge2Te6. Phys. Rev. B 2020, 101, 014440. [Google Scholar] [CrossRef]
  27. Šiškins, M.; Kurdi, S.; Lee, M.; Slotboom, B.J.M.; Xing, W.Y.; l Mañas-Valero, S.; Coronado, E.; Jia, S.; Han, W.; Sar, T.V.D.; et al. Nanomechanical probing and strain tuning of the curie temperature in suspended Cr2Ge2Te6-based heterostructures. npj 2D Mater. Appl. 2022, 6, 41. [Google Scholar] [CrossRef]
  28. McCray, A.R.C.; Li, Y.; Qian, E.; Li, Y.; Wang, W.; Huang, Z.J.; Ma, X.M.; Liu, Y.Z.; Chung, D.Y.; Kanatzidis, M.G.; et al. Direct observation of magnetic bubble lattices and magnetoelastic effects in van der Waals Cr2Ge2Te6. Adv. Funct. Mater. 2023, 23, 2214203. [Google Scholar] [CrossRef]
  29. Noah, A.; Zur, Y.; Fridman, N.; Singh, S.; Gutfreund, A.; Herrera, E.; Vakahi, A.; Remennik, S.; Huber, M.E.; Gazit, S.; et al. Nano-patterned magnetic edges in CrGeTe3 for quasi 1-D spintronic devices. ACS Appl. Nano Mater. 2023, 6, 8627–8634. [Google Scholar] [CrossRef]
  30. O’Neill, A.; Rahman, S.; Zhang, Z.; Schoenherr, P.; Yildirim, T.J.; Gu, B.; Su, G.; Lu, Y.R.; Seidel, J. Enhanced room temperature ferromagnetism in highly strained 2D semiconductor Cr2Ge2Te6. ACS Nano 2023, 17, 735–742. [Google Scholar] [CrossRef]
  31. Spachmann, S.; Selter, S.; Büchner, B.; Aswartham, S.; Klingeler, R. Strong uniaxial pressure dependencies evidencing spin-lattice coupling and spin fluctuations in Cr2Ge2Te6. Phys. Rev. B 2023, 107, 184421. [Google Scholar] [CrossRef]
  32. Kritika, V.; Vavilapalli, D.S.; Arya, A.; Srivastava, S.K.; Singh, R.; Sagdeo, A.; Jha, S.N.; Kumar, K.; Soma Banik, S. Magneto-strain effects in 2D ferromagnetic van der Waal material CrGeTe3. Sci. Rep. 2023, 13, 8579. [Google Scholar]
  33. Li, T.X.; Jiang, S.W.; Sivadas, N.; Wang, Z.F.; Xu, Y.; Weber, D.; Goldberger, J.E.; Watanabe, K.; Taniguchi, T.; Fennie, C.J.; et al. Pressure-controlled interlayer magnetism in atomically thin CrI3. Nat. Mater. 2019, 18, 1303–1308. [Google Scholar] [CrossRef]
  34. Deng, Y.J.; Yu, Y.J.; Shi, M.Z.; Guo, Z.X.; Xu, Z.H.; Wang, J.; Chen, X.H.; Zhang, Y.B. Quantum anomalous hall effect in intrinsic magnetic topological insulator MnBi2Te4. Science 2020, 367, eaax8156. [Google Scholar] [CrossRef] [PubMed]
  35. Deiseroth, H.J.; Aleksandrov, K.; Reiner, C.; Kienle, L.; Kremer, R.K. Fe3GeTe2 and Ni3GeTe2—Two new layered transition-metal compounds: Crystal structures, HRTEM investigations, and magnetic and electrical properties. Eur. J. Inorg. Chem. 2006, 2006, 1561–1567. [Google Scholar] [CrossRef]
  36. Verchenko, V.Y.; Tsirlin, A.A.; Sobolev, A.V.; Presniakov, I.A.; Shevelkov, A.V. Ferromagnetic order, strong magnetocrystalline anisotropy, and magnetocaloric effect in the layered telluride Fe3−δGeTe2. Inorg. Chem. 2015, 54, 8598–8607. [Google Scholar] [CrossRef]
  37. Chen, B.; Yang, J.H.; Wang, H.D.; Imai, M.; Ohta, H.; Michioka, C.; Yoshimura, K.; Fang, M.H. Magnetic properties of layered itinerant electron ferromagnet Fe3GeTe2. J. Phys. Soc. Jpn. 2013, 82, 124711. [Google Scholar] [CrossRef]
  38. Drachuck, G.; Salman, Z.; Masters, M.W.; Taufour, V.; Lamichhane, T.N.; Lin, Q.S.; Straszheim, W.E.; Bud’ko, S.L.; Canfield, P.C. Effect of nickel substitution on magnetism in the layered van der Waals ferromagnet Fe3GeTe2. Phys. Rev. B 2018, 98, 144434. [Google Scholar] [CrossRef]
  39. May, A.F.; Ovchinnikov, D.; Zheng, Q.; Hermann, R.; Calder, S.; Huang, B.V.; Fei, Z.Y.; Liu, Y.H.; Xu, X.D.; McGuire, M.A. Ferromagnetism near room temperature in the cleavable van der Waals crystal Fe5GeTe2. ACS Nano 2019, 13, 4436–4442. [Google Scholar] [CrossRef]
  40. You, Y.R.; Gong, Y.Y.; Li, H.; Li, Z.F.; Zhu, M.M.; Tang, J.X.; Liu, E.K.; Yao, Y.; Xu, G.Z.; Xu, F.; et al. Angular dependence of the topological hall effect in the uniaxial van der Waals ferromagnet Fe3GeTe2. Phys. Rev. B 2019, 100, 134441. [Google Scholar] [CrossRef]
  41. Ke, J.Z.; Yang, M.; Xia, W.; Zhu, H.P.; Liu, C.B.; Chen, R.; Dong, C.; Liu, W.X.; Shi, M.Y.; Guo, Y.F.; et al. Magnetic and magneto-transport studies of two-dimensional ferromagnetic compound Fe3GeTe2. J. Phys. Condens. Matter 1995, 32, 405805. [Google Scholar] [CrossRef]
  42. Ohta, T.; Sakai, K.; Taniguchi, H.; Driesen, B.; Okada, Y.; Kobayashi, K.; Niimi, Y. Enhancement of coercive field in atomically-thin quenched Fe5GeTe2. Appl. Phys. Express 2020, 13, 043005. [Google Scholar] [CrossRef]
  43. Park, S.Y.; Kim, D.S.; Liu, Y.; Hwang, J.; Kim, Y.; Kim, W.; Kim, J.Y.; Petrovic, C.; Hwang, C.; Mo, S.K.; et al. Controlling the magnetic anisotropy of the van der Waals ferromagnet Fe3GeTe2 through hole doping. Nano Lett. 2020, 20, 95–100. [Google Scholar] [CrossRef] [PubMed]
  44. Jiang, H.N.; Zang, Z.H.; Wang, X.G.; Que, H.F.; Wang, L.; Si, K.P.; Zhang, P.; Ye, Y.; Gong, Y.J. Thickness-tunable growth of composition-controllable two-dimensional FexGeTe2. Nano Lett. 2022, 22, 9477–9484. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, P.; Wang, X.G.; Jiang, H.N.; Zhang, Y.W.; He, Q.Q.; Si, K.P.; Li, B.X.; Zhao, F.F.; Cui, A.Y.; Wei, Y.; et al. Flux-assisted growth of atomically thin materials. Nat. Synth. 2022, 1, 864–872. [Google Scholar] [CrossRef]
  46. Zhuang, H.L.; Kent, P.R.C.; Hennig, R.G. Strong anisotropy and magnetostriction in the two-dimensional stoner ferromagnet Fe3GeTe2. Phys. Rev. B 2016, 93, 134407. [Google Scholar] [CrossRef]
  47. Huang, Y.; Pan, Y.H.; Yang, R.; Bao, L.H.; Meng, L.; Luo, H.L.; Cai, Y.Q.; Liu, G.D.; Zhao, W.J.; Zhou, Z.; et al. Universal mechanical exfoliation of large-area 2D Crystals. Nat. Commun. 2020, 11, 2453. [Google Scholar] [CrossRef] [PubMed]
  48. Yu, W.; Wang, C.; Liang, S.J.; Ma, Z.C.; Xu, K.; Liu, X.W.; Zhang, L.L.; Admasu, A.S.; Cheong, S.W.; Wang, L.Z.; et al. Strain-sensitive magnetization reversal of a van der waals magnet. Adv. Mater. 2020, 23, 2004533. [Google Scholar]
  49. Niu, W.; Cao, Z.; Wang, Y.L.; Wu, Z.Q.; Zhang, X.Q.; Han, W.B.; Wei, L.J.; Wang, L.X.; Xu, Y.B.; Zou, Y.M.; et al. Antisymmetric magnetoresistance in Fe3GeTe2 nanodevices of inhomogeneous thickness. Phys. Rev. B 2021, 104, 125429. [Google Scholar] [CrossRef]
  50. Ma, S.P.; Li, G.H.; Li, Z.; Zhang, Y.W.; Lu, H.L.; Gao, Z.S.; Wu, J.X.; Long, G.K.; Huang, Y. 2D magnetic semiconductor Fe3GeTe2 with few and single layers with a greatly enhanced intrinsic exchange bias by liquid-phase exfoliation. ACS Nano 2022, 16, 19439–19450. [Google Scholar] [CrossRef]
  51. Nair, G.K.R.; Zhang, Z.W.; Hou, F.C.; Abdelaziem, A.; Xu, X.D.; Yang, S.W.Q.; Zhang, N.; Li, W.Q.; Zhu, C.; Wu, Y.; et al. Phase-pure two-dimensional FexGeTe2 magnets with near-room-temperature TC. Nano Res. 2021, 15, 457–464. [Google Scholar] [CrossRef]
  52. Zhou, J.D.; Zhu, C.; Zhou, Y.; Dong, J.C.; Li, P.L.; Zhang, Z.W.; Wang, Z.; Lin, Y.C.; Shi, J.; Zhang, R.W.; et al. Composition and phase engineering of metal chalcogenides and phosphorous chalcogenides. Nat. Mater. 2022, 22, 450–458. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, B.; Liu, S.S.; Yang, L.; Chen, Z.D.; Zhang, E.Z.; Li, Z.H.; Wu, J.; Ruan, X.Z.; Xiu, F.X.; Liu, W.Q.; et al. Light-tunable ferromagnetism in atomically thin Fe3GeTe2 driven by femtosecond laser pulse. Phys. Rev. Lett. 2020, 125, 267205. [Google Scholar] [CrossRef]
  54. Liu, S.S.; Yang, K.; Liu, W.Q.; Zhang, E.Z.; Li, Z.H.; Zhang, X.Q.; Liao, Z.M.; Zhang, W.; Sun, J.B.; Yang, Y.K.; et al. Two-dimensional ferromagnetic superlattices. Nat. Sci. Rev. 2020, 7, 745–754. [Google Scholar] [CrossRef] [PubMed]
  55. Roemer, R.; Liu, C.; Zou, K. Robust ferromagnetism in wafer-scale monolayer and multilayer Fe3GeTe2. npj 2D Mater. Appl. 2020, 4, 33. [Google Scholar] [CrossRef]
  56. Wang, H.Y.; Liu, Y.J.; Wu, P.C.; Hou, W.J.; Jiang, Y.H.; Li, X.H.; Pandey, C.; Chen, D.D.; Yang, Q.; Wang, H.T.; et al. Above room-temperature ferromagnetism in wafer-scale two-dimensional van der Waals Fe3GeTe2 tailored by a topological insulator. ACS Nano 2020, 14, 10045–10053. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, X.H.; Wang, H.T.; Liu, H.J.; Wang, C.; Wei, G.S.; Fang, C.; Wang, H.C.; Geng, C.Y.; Liu, S.J.; Li, P.Y.; et al. Generation and control of terahertz spin currents in topology-induced 2D ferromagnetic Fe3GeTe2|Bi2Te3 heterostructures. Adv. Mater. 2022, 34, 2106172. [Google Scholar] [CrossRef]
  58. Zhou, W.Y.; Bishop, A.J.; Zhu, M.L.; Lyalin, I.; Walko, R.; Gupta, J.A.; Hwang, J.; Kawakami, R.K. Kinetically controlled epitaxial growth of Fe3GeTe2 van der Waals ferromagnetic films. ACS Appl. Mater. Interfaces 2022, 4, 3190–3197. [Google Scholar] [CrossRef]
  59. Wang, H.T.; Lu, H.C.; Guo, Z.X.; Li, A.; Wu, P.C.; Li, J.; Xie, W.R.; Sun, Z.M.; Li, P.; Damas, H.; et al. Interfacial engineering of ferromagnetism in wafer-scale van der Waals Fe4GeTe2 far above room temperature. Nat. Commun. 2023, 14, 2483. [Google Scholar] [CrossRef]
  60. Lv, H.; Silva, A.; Figueroa, A.I.; Guillemard, C.; Aguirre, I.F.; Camosi, L.; Aballe, L.; Valvidares, M.; Valenzuela, S.O.; Schubert, J.; et al. Large-area synthesis of ferromagnetic Fe5−XGeTe2/graphene van der Waals heterostructures with curie temperature above room temperature. Small 2023, 19, 2302387. [Google Scholar] [CrossRef]
  61. Seo, J.; Kim, D.Y.; An, E.S.; Kim, K.; Kim, G.Y.; Hwang, S.Y.; Kim, D.W.; Jang, B.G.; Kim, H.; Eom, G.; et al. Nearly room temperature ferromagnetism in a magnetic metal-rich van der Waals metal. Sci. Adv. 2020, 6, eaay8912. [Google Scholar] [CrossRef] [PubMed]
  62. Tan, C.; Xie, W.Q.; Zheng, G.L.; Aloufi, N.; Albarakati, S.; Algarni, M.; Li, J.B.; Partridge, J.; Culcer, D.; Wang, X.L.; et al. Gate-controlled magnetic phase transition in a van der Waals magnet Fe5GeTe2. Nano Lett. 2021, 21, 5599–5605. [Google Scholar] [CrossRef] [PubMed]
  63. Yang, X.X.; Zhou, X.D.; Feng, W.X.; Yao, Y.G. Strong magneto-optical effect and anomalous transport in the two-dimensional van der Waals magnets FenGeTe2 (N = 3, 4, 5). Phys. Rev. B 2021, 104, 104427. [Google Scholar] [CrossRef]
  64. Liu, Q.X.; Xing, J.P.; Jiang, Z.; Guo, Y.; Jiang, X.; Qi, Y.; Zhao, J.J. Layer-dependent magnetic phase diagram in FenGeTe2 (3 ≤ N ≤ 7) ultrathin films. Commun. Phys. 2022, 5, 140. [Google Scholar] [CrossRef]
  65. Ghosh, S.; Ershadrad, S.; Borisov, V.; Sanyal, B. Unraveling effects of electron correlation in two-dimensional FenGeTe2 (N = 3, 4, 5) by dynamical mean field theory. npj Comput. Mater. 2023, 9, 86. [Google Scholar] [CrossRef]
  66. Ren, H.T.; Xiang, G. Morphology-dependent room-temperature ferromagnetism in undoped ZnO nanostructures. Nanomaterials 2021, 11, 3199. [Google Scholar] [CrossRef]
  67. Ren, H.T.; Xiang, G. Strain-modulated magnetism in MoS2. Nanomaterials 2022, 12, 1929. [Google Scholar] [CrossRef]
  68. Ren, H.T.; Xiang, G. Strain engineering of intrinsic ferromagnetism in 2D van der Waals materials. Nanomaterials 2023, 13, 2378. [Google Scholar] [CrossRef]
  69. Ren, H.T.; Zhong, J.; Xiang, G. The progress on magnetic material thin films prepared using polymer-assisted deposition. Molecules 2023, 28, 5004. [Google Scholar] [CrossRef]
  70. Joe, M.; Yang, U.; Lee, C.G. First-principles study of ferromagnetic metal Fe5GeTe2. Nano Mater. Sci. 2009, 1, 299–303. [Google Scholar] [CrossRef]
  71. Hu, X.H.; Zhao, Y.H.; Shen, X.D.; Krasheninnikov, A.V.; Chen, Z.F.; Sun, L.T. Enhanced ferromagnetism and tunable magnetism in Fe3GeTe2 monolayer by strain engineering. ACS Appl. Mater. Interfaces 2020, 12, 26367–26373. [Google Scholar] [CrossRef] [PubMed]
  72. Hu, L.; Zhou, J.; Hou, Z.P.; Su, W.T.; Yang, B.Z.; Li, L.W.; Yan, M. Polymer-buried van der Waals magnets for promising wearable room-temperature spintronics. Mater. Horiz. 2021, 8, 3306–3314. [Google Scholar] [CrossRef] [PubMed]
  73. Zhu, M.M.; You, Y.R.; Xu, G.Z.; Tang, J.X.; Gong, Y.Y.; Xu, F. Strain modulation of magnetic coupling in the metallic van der Waals magnet Fe3GeTe2. Intermetallics 2021, 131, 107085. [Google Scholar] [CrossRef]
  74. Lim, M.; Choi, B.; Ghim, M.; Park, J.; Lee, H.W. Robustness of the intrinsic anomalous hall effect in Fe3GeTe2 to a uniaxial strain. Phys. Rev. Mater. 2023, 7, 064003. [Google Scholar] [CrossRef]
  75. Miao, X.Y.; Li, S.; Jiang, Z.Y.; Zhang, C.M.; Du, A.J. A strain induced polar metal phase in a ferromagnetic Fe3GeTe2 monolayer. Phys. Chem. Chem. Phys. 2023, 25, 18826–18832. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, X.Q.; Li, Z.Y.; Zhang, M.; Hou, T.; Zhao, J.G.; Li, L.; Rahman, A.; Xu, Z.L.; Gong, J.B.; Chi, Z.H.; et al. Pressure-induced modification of the anomalous hall effect in layered Fe3GeTe2. Phys. Rev. B 2019, 100, 014407. [Google Scholar] [CrossRef]
  77. O’Hara, D.J.; Brubaker, Z.E.; Stillwell, R.L.; O’Bannon, E.F.; Baker, A.A.; Weber, D.; Aji, L.B.B.; Goldberger, J.E.; Kawakami, R.K.; Zieve, R.J.; et al. Suppression of magnetic ordering in Fe-deficient Fe3−xGeTe2 from application of pressure. Phys. Rev. B 2020, 102, 014407. [Google Scholar] [CrossRef]
  78. Wang, H.S.; Xu, R.Z.; Liu, C.; Wang, L.; Zhang, Z.; Su, H.M.; Wang, S.M.; Zhao, Y.S.; Liu, Z.J.; Yu, D.P.; et al. Pressure-dependent intermediate magnetic phase in thin Fe3GeTe2 flakes. J. Phys. Chem. Lett. 2020, 11, 7313–7319. [Google Scholar] [CrossRef]
  79. Ding, S.L.; Liang, Z.Y.; Yang, J.; Yun, C.; Zhang, P.J.; Li, Z.F.; Xue, M.Z.; Liu, Z.; Tian, G.; Liu, F.Y.; et al. Ferromagnetism in two-dimensional Fe3GeTe2; tunability by hydrostatic pressure. Phys. Rev. B 2021, 103, 094429. [Google Scholar] [CrossRef]
  80. Li, Z.Y.; Tang, M.; Huang, J.W.; Qin, F.; Ao, L.Y.; Shen, Z.W.; Zhang, C.R.; Chen, P.; Bi, X.Y.; Qiu, C.Y.; et al. Magnetic anisotropy control with curie temperature above 400 K in a van der Waals ferromagnet for spintronic device. Adv. Mater. 2022, 34, 2201209. [Google Scholar] [CrossRef]
  81. Dang, N.T.; Kozlenko, D.P.; Lis, O.N.; Kichanov, S.E.; Lukin, Y.V.; Golosova, N.O.; Savenko, B.N.; Duong, D.L.; Phan, T.L.; Tran, T.A.; et al. High pressure-driven magnetic disorder and structural transformation in Fe3GeTe2: Emergence of a magnetic quantum critical point. Adv. Sci. 2023, 10, 2206842. [Google Scholar] [CrossRef] [PubMed]
  82. Tengdin, P.; Gentry, C.; Blonsky, A.; Zusin, D.; Gerrity, M.; Hellbrück, L.; Hofherr, M.; Shaw, J.; Kvashnin, Y.; Delczeg-Czirjak, E.K.; et al. Direct light–induced spin transfer between different elements in a spintronic Heusler material via femtosecond laser excitation. Sci. Adv. 2020, 6, eaaz1100. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, Y.P.; Chen, X.Y.; Long, M.Q. Modifications of magnetic anisotropy of Fe3GeTe2 by the electric field effect. Appl. Phys. Lett. 2020, 116, 092404. [Google Scholar] [CrossRef]
  84. Zhao, M.T.; Zhao, Y.Y.; Xi, Y.L.; Xu, H.; Feng, H.F.; Xu, X.; Hao, W.C.; Zhou, S.; Zhao, J.J.; Dou, S.X.; et al. Electric-field-driven negative differential conductance in 2D van der Waals ferromagnet Fe3GeTe2. Nano Lett. 2021, 21, 9233–9239. [Google Scholar] [CrossRef]
  85. Zhang, L.M.; Huang, X.Y.; Dai, H.W.; Wang, M.S.; Cheng, H.; Tong, L.; Li, Z.; Han, X.T.; Wang, X.; Ye, L.; et al. Proximity-coupling-induced significant enhancement of coercive field and curie temperature in 2D van der Waals heterostructures. Adv. Mater. 2020, 34, 2002032. [Google Scholar] [CrossRef]
  86. Chen, Q.; Liang, J.; Fang, B.; Zhu, Y.H.; Wang, J.C.; Lv, W.M.; Lv, W.X.; Cai, J.L.; Huang, Z.C.; Zhai, Y.; et al. Proximity effect of a two-dimensional van der Waals magnet Fe3GeTe2 on nickel films. Nanoscale 2021, 13, 14688–14693. [Google Scholar] [CrossRef]
  87. Tu, Z.Y.; Zhou, T.; Ersevim, T.; Arachchige, H.S.; Hanbicki, A.T.; Friedman, A.L.; Mandrus, D.; Ouyang, M.; Žutić, I.; Gong, C. Spin–orbit coupling proximity effect in MoS2/Fe3GeTe2 heterostructures. Appl. Phys. Lett. 2022, 120, 043102. [Google Scholar] [CrossRef]
  88. Zhang, L.M.; Song, L.Y.; Dai, H.W.; Yuan, J.H.; Wang, M.S.; Huang, X.Y.; Qiao, L.; Cheng, H.; Wang, X.; Ren, W.; et al. Substrate-modulated ferromagnetism of two-dimensional Fe3GeTe2. Appl. Phys. Lett. 2020, 116, 042402. [Google Scholar] [CrossRef]
  89. Kim, S.J.; Choi, D.; Kim, K.M.; Lee, K.Y.; Kim, D.H.; Hong, S.; Suh, J.; Lee, C.; Kim, S.K.; Park, T.E.; et al. Interface engineering of magnetic anisotropy in van der Waals ferromagnet-based heterostructures. ACS Nano 2021, 15, 16395–16403. [Google Scholar] [CrossRef] [PubMed]
  90. Xu, M.Q.; Kendrick, L.H.; Kale, A.; Gang, Y.Q.; Ji, G.; Scalettar, R.T.; Lebrat, M.; Greiner, M. Frustration- and doping-induced magnetism in a fermi–hubbard simulator. Nature 2023, 620, 971–976. [Google Scholar] [CrossRef]
  91. Verchenko, V.Y.; Sokolov, S.S.; Tsirlin, A.A.; Sobolev, A.V.; Presniakov, I.A.; Bykov, M.A.; Kirsanova, M.A.; Shevelkov, A.V. New Fe-based layered telluride Fe3−δAs1−YTe2: Synthesis, crystal structure and physical properties. Dalton Trans. 2016, 45, 16938–16947. [Google Scholar] [CrossRef]
  92. Yuan, D.D.; Jin, S.F.; Liu, N.; Shen, S.J.; Lin, Z.P.; Li, K.K.; Chen, X.L. Tuning magnetic properties in quasi-two-dimensional ferromagnetic Fe3-YGe1−XAsxTe2 (0 ≤ X ≤ 0.85). Mater. Res. Express 2017, 4, 036103. [Google Scholar] [CrossRef]
  93. Tian, C.K.; Wang, C.; Ji, W.; Wang, J.C.; Xia, T.L.; Wang, L.; Liu, J.J.; Zhang, H.X.; Cheng, P. Domain wall pinning and hard magnetic phase in Co-doped bulk single crystalline Fe3GeTe2. Phys. Rev. B 2019, 99, 094429. [Google Scholar] [CrossRef]
  94. Jang, S.W.; Yoon, H.; Jeong, M.Y.; Ryee, S.; Kim, H.S.; Han, M.J. Origin of ferromagnetism and the effect of doping on Fe3GeTe2. Nanoscale 2020, 12, 13501–13506. [Google Scholar] [CrossRef] [PubMed]
  95. May, A.F.; Du, M.H.; Cooper, V.R.; McGuire, M.A. Tuning magnetic order in the van der Waals metal Fe5GeTe2 by cobalt substitution. Phys. Rev. Mater. 2020, 4, 074008. [Google Scholar] [CrossRef]
  96. Tian, C.K.; Pan, F.H.; Xu, S.; Ai, K.; Xia, T.L.; Cheng, P. Tunable magnetic properties in van der Waals crystals (Fe1−XCox)5GeTe2. Appl. Phys. Lett. 2020, 116, 202402. [Google Scholar] [CrossRef]
  97. Zhang, S.Z.; Liang, X.; Zhao, H.Y.; Chen, Y.H.; He, Q.; Liu, J.; Lv, L.; Yang, J.H.; Wu, H.L.; Chen, L. Tuning the magnetic properties of Fe3GeTe2 by doping with 3d transition-metals. Phys. Lett. A 2021, 396, 127219. [Google Scholar] [CrossRef]
  98. May, A.F.; Yan, J.Q.; Hermann, R.; Du, M.H.; McGuire, M.A. Tuning the room temperature ferromagnetism in Fe5GeTe2 by arsenic substitution. 2D Mater. 2022, 9, 015013. [Google Scholar] [CrossRef]
  99. Yang, J.; Quhe, R.; Liu, S.Q.; Peng, Y.X.; Sun, X.T.; Zha, L.; Wu, B.C.; Shi, B.W.; Yang, C.; Shi, J.J.; et al. Gate-tunable high magnetoresistance in monolayer Fe3GeTe2 spin valves. Phys. Chem. Chem. Phys. 2020, 22, 25730–25739. [Google Scholar] [CrossRef]
  100. Zheng, G.L.; Xie, W.Q.; Albarakati, S.; Algarni, M.; Tan, C.; Wang, Y.H.; Peng, J.Y.; Partridge, J.; Farrar, L.; Yi, J.B.; et al. Gate-tuned interlayer coupling in van der Waals ferromagnet Fe3GeTe2 nanoflakes. Phys. Rev. Lett. 2020, 125, 047202. [Google Scholar] [CrossRef] [PubMed]
  101. Chen, G.Y.; Zhang, Y.; Qi, S.M.; Chen, J.H. Gate-controlled magnetic transitions in Fe3GeTe2 with lithium ion conducting glass substrate. Chin. Phys. B 2021, 30, 097504. [Google Scholar] [CrossRef]
  102. Tang, M.; Huang, J.W.; Qin, F.; Zhai, K.; Ideue, T.; Li, Z.Y.; Meng, F.H.; Nie, A.M.; Wu, L.L.; Bi, X.Y.; et al. Continuous manipulation of magnetic anisotropy in a van der Waals ferromagnet via electrical gating. Nat. Electron. 2023, 6, 28–36. [Google Scholar] [CrossRef]
  103. Stahl, J.; Pomjakushin, V.; Johrendt, D. Ferromagnetism in Fe3−X−YNixGeTe2. Z. Naturforsch. 2016, 71, 273–276. [Google Scholar] [CrossRef]
  104. Hwang, I.; Coak, M.J.; Lee, N.; Ko, D.S.; Oh, Y.; Jeon, I.; Son, S.; Zhang, K.X.; Kim, J.; Park, J.G. Hard ferromagnetic van-der-Waals metal (Fe,Co)(3)GeTe2: A new platform for the study of low-dimensional magnetic quantum criticality. J. Phys. Condens. Matter. 2019, 31, 50LT01. [Google Scholar] [CrossRef] [PubMed]
  105. Shen, Z.X.; Bo, X.Y.; Cao, K.; Wan, X.G.; He, L.X. Magnetic ground state and electron-doping tuning of curie temperature in Fe3GeTe2: First-principles studies. Phys. Rev. B 2021, 103, 085102. [Google Scholar] [CrossRef]
  106. Zhao, M.T.; Chen, B.B.; Xi, Y.L.; Zhao, Y.Y.; Xu, H.; Zhang, H.R.; Cheng, N.Y.; Feng, H.F.; Zhuang, J.C.; Pan, F.; et al. Kondo holes in the two-dimensional itinerant ising ferromagnet Fe3GeTe2. Nano Lett. 2021, 21, 6117–6123. [Google Scholar] [CrossRef]
  107. Weber, D.; Trout, A.H.; McComb, D.W.; Goldberger, J.E. Decomposition-induced room-temperature magnetism of the Na-intercalated layered ferromagnet Fe3−xGeTe2. Nano Lett. 2019, 19, 5031–5035. [Google Scholar] [CrossRef]
  108. Wu, Y.S.; Hu, Y.X.; Wang, C.; Zhou, X.; Hou, X.F.; Xia, W.; Zhang, Y.W.; Wang, J.H.; Ding, Y.F.; He, J.D.; et al. Fe-intercalation dominated ferromagnetism of van der Waals Fe3GeTe2. Adv. Mater. 2023, 35, 2302568. [Google Scholar] [CrossRef]
  109. Yang, M.M.; Li, Q.; Chopdekar, R.V.; Stan, C.; Cabrini, S.; Choi, J.W.; Wang, S.; Wang, T.Y.; Gao, N.; Scholl, A.; et al. Highly enhanced curie temperature in Ga-implanted Fe3GeTe2 van der Waals material. Adv. Quantum Technol. 2020, 3, 2000017. [Google Scholar] [CrossRef]
  110. Chen, D.; Sun, W.; Wang, W.X.; Li, X.N.; Li, H.; Cheng, Z.X. Twist-stacked 2D bilayer Fe3GeTe2 with tunable magnetism. J. Mater. Chem. C 2022, 10, 12741–12750. [Google Scholar] [CrossRef]
  111. Ko, E. Hybridized bands and stacking-dependent band edges in ferromagnetic Fe3GeTe2/CrGeTe3 moiré heterobilayer. Sci. Rep. 2022, 12, 5101. [Google Scholar] [CrossRef] [PubMed]
  112. Li, X.L.; Lu, J.T.; Zhang, J.; You, L.; Su, Y.R.; Tsymbal, E.Y. Spin-dependent transport in van der Waals magnetic tunnel junctions with Fe3GeTe2 electrodes. Nano Lett. 2019, 19, 5133–5139. [Google Scholar] [CrossRef] [PubMed]
  113. Su, Y.R.; Li, X.L.; Zhu, M.; Zhang, J.; You, L.; Tsymbal, E.Y. Van der Waals multiferroic tunnel junctions. Nano Lett. 2021, 21, 175–181. [Google Scholar] [CrossRef] [PubMed]
  114. Lin, Z.Z.; Chen, X. Ultrathin scattering spin filter and magnetic tunnel junction implemented by ferromagnetic 2D van der Waals material. Adv. Electron. Mater. 2020, 6, 1900968. [Google Scholar] [CrossRef]
  115. Wang, Z.A.; Xue, W.S.; Yan, F.G.; Zhu, W.K.; Liu, Y.; Zhang, X.H.; Wei, Z.M.; Chang, K.; Yuan, Z.; Wang, K.Y. Selectively controlled ferromagnets by electric fields in van der Waals ferromagnetic heterojunctions. Nano Lett. 2023, 23, 710–717. [Google Scholar] [CrossRef]
  116. Hu, C.; Zhang, D.; Yan, F.G.; Li, Y.C.; Lv, Q.S.; Zhu, W.K.; Wei, Z.M.; Chang, K.; Wang, K.Y. From two- to multi-state vertical spin valves without spacer layer based on Fe3GeTe2 van der Waals homo-junctions. Sci. Bull. 2020, 65, 1072–1077. [Google Scholar] [CrossRef]
  117. Zhu, W.K.; Lin, H.L.; Yan, F.G.; Hu, C.; Wang, Z.A.; Zhao, L.X.; Deng, Y.C.; Kudrynskyi, Z.R.; Zhou, T.; Kovalyuk, Z.D.; et al. Large tunneling magnetoresistance in van der Waals ferromagnet/semiconductor heterojunctions. Adv. Mater. 2021, 7, 2104658. [Google Scholar] [CrossRef]
  118. Zhao, B.; Ngaloy, R.; Ghosh, S.; Ershadrad, S.; Gupta, R.; Ali, K.; Hoque, A.M.; Karpiak, B.; Khokhriakov, D.; Polley, C.; et al. A room-temperature spin-valve with van der Waals ferromagnet Fe5GeTe2/graphene heterostructure. Adv. Mater. 2023, 35, 2209113. [Google Scholar] [CrossRef] [PubMed]
  119. Wang, X.; Tang, J.; Xia, X.X.; He, C.L.; Zhang, J.W.; Liu, Y.Z.; Wan, C.H.; Fang, C.; Guo, C.Y.; Yang, W.L.; et al. Current-driven magnetization switching in a van der Waals ferromagnet Fe3GeTe2. Sci. Adv. 2019, 5, eaaw8904. [Google Scholar] [CrossRef]
  120. Li, G.H.; Ma, S.P.; Li, Z.; Zhang, Y.W.; Diao, J.L.; Xia, L.; Zhang, Z.W.; Huang, Y. High-quality ferromagnet Fe3GeTe2 for high-efficiency electromagnetic wave absorption and shielding with wideband radar cross section reduction. ACS Nano 2022, 16, 7861–7879. [Google Scholar] [CrossRef]
  121. Zhang, Y.; Lu, H.Y.; Zhu, X.G.; Tan, S.Y.; Feng, W.; Liu, Q.; Zhang, W.; Chen, Q.Y.; Liu, Y.; Luo, X.B.; et al. Emergence of kondo lattice behavior in a van der Waals itinerant ferromagnet Fe3GeTe2. Sci. Adv. 2018, 4, eaao6791. [Google Scholar] [CrossRef]
  122. Canfield, P.C.; Fisk, Z. Growth of single crystals from metallic fluxes. Philos. Mag. B 1992, 65, 1117–1123. [Google Scholar] [CrossRef]
  123. Canfield, P.C.; Fisher, I.R. High-temperature solution growth of intermetallic single crystals and quasicrystals. J. Cryst. Growth 2001, 225, 155–161. [Google Scholar] [CrossRef]
  124. Yan, J.Q.; Sales, B.C.; Susner, M.A.; McGuire, M.A. Flux growth in a horizontal configuration: An analog to vapor transport growth. Phys. Rev. Mater. 2017, 1, 023402. [Google Scholar] [CrossRef]
  125. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666. [Google Scholar] [CrossRef]
  126. Chahal, S.; Ranjan, P.; Motlag, M.; Yamijala, S.S.R.K.C.; Late, D.J.; Sadki, E.H.S.; Cheng, G.J.; Kumar, P. Borophene via micromechanical exfoliation. Adv. Mater. 2021, 33, 2102039. [Google Scholar] [CrossRef] [PubMed]
  127. Hsu, C.L.; Lin, C.T.; Huang, J.H.; Chu, C.W.; Wei, K.H.; Li, L.J. Layer-by-layer graphene/TCNQ stacked films as conducting anodes for organic solar cells. ACS Nano 2012, 6, 5031–5039. [Google Scholar] [CrossRef]
  128. Moon, J.Y.; Kim, M.; Kim, S.I.; Xu, S.G.; Choi, J.H.; Whang, D.; Watanabe, K.; Taniguchi, T.; Park, D.S.; Seo, J.; et al. Layer-engineered large-area exfoliation of graphene. Sci. Adv. 2020, 6, eabc6601. [Google Scholar] [CrossRef]
  129. Magda, G.Z.; Pető, J.; Dobrik, G.; Hwang, C.; Biró, L.P.; Tapasztó, L. Exfoliation of large-area transition metal chalcogenide single layers. Sci. Rep. 2015, 5, 14714. [Google Scholar] [CrossRef]
  130. Ding, D.; Wang, S.; Xia, Y.P.; Li, P.; He, D.L.; Zhang, J.Q.; Zhao, S.W.; Yu, G.H.; Zheng, Y.H.; Cheng, Y.; et al. Atomistic insight into the epitaxial growth mechanism of single-crystal two-dimensional transition-metal dichalcogenides on Au (111) substrate. ACS Nano 2022, 16, 17356–17364. [Google Scholar] [CrossRef]
  131. Yang, P.F.; Zhang, S.Q.; Pan, S.Y.; Tang, B.; Liang, Y.; Zhao, X.X.; Zhang, Z.P.; Shi, J.P.; Huan, Y.H.; Shi, Y.P.; et al. Epitaxial growth of centimeter-scale single-crystal MoS2 monolayer on Au (111). ACS Nano 2020, 14, 5036–5045. [Google Scholar] [CrossRef] [PubMed]
  132. Reidy, K.; Varnavides, G.; Thomsen, J.D.; Kumar, A.; Pham, T.; Blackburn, A.M.; Anikeeva, P.; Narang, P.; LeBeau, J.M.; Ross, F.M. Direct imaging and electronic structure modulation of moiré superlattices at the 2D/3D interface. Nat. Commun. 2021, 12, 1290. [Google Scholar] [CrossRef] [PubMed]
  133. Gao, Y.; Hong, Y.L.; Yin, L.C.; Wu, Z.T.; Yang, Z.Q.; Chen, M.L.; Liu, Z.B.; Ma, T.; Sun, D.M.; Ni, Z.H. Ultrafast growth of high-quality monolayer WSe2 on Au. Adv. Mater. 2017, 29, 1700990. [Google Scholar] [CrossRef]
  134. May, A.F.; Yan, J.Q.; McGuire, M.A. A practical guide for crystal growth of van der Waals layered materials. J. Appl. Phys. 2020, 128, 051101. [Google Scholar] [CrossRef]
  135. Wang, X.W.; Wu, P.Y. Aqueous phase exfoliation of two-dimensional materials assisted by thermoresponsive polymeric ionic liquid and their applications in stimuli-responsive hydrogels and highly thermally conductive films. ACS Appl. Mater. Interfaces 2018, 10, 2504–2514. [Google Scholar] [CrossRef]
  136. Liu, S.S.; Li, Z.H.; Yang, K.; Zhang, E.Z.; Narayan, A.; Zhang, X.Q.; Zhu, J.Y.; Liu, W.Q.; Liao, Z.M.; Kudo, M.; et al. Tuning 2D magnetism in Fe3+XGeTe2 films by element doping. Nat. Sci. Rev. 2021, 9, 117. [Google Scholar] [CrossRef] [PubMed]
  137. Ly, T.T.; Park, J.; Kim, K.; Ahn, H.B.; Lee, N.J.; Kim, K.; Park, T.E.; Duvjir, G.; Lam, N.H.; Jang, K.; et al. Direct observation of Fe-Ge ordering in Fe5−XGeTe2 crystals and resultant helimagnetism. Adv. Funct. Mater. 2021, 31, 2009758. [Google Scholar] [CrossRef]
  138. Lv, X.W.; Pei, K.; Yang, C.D.; Qin, G.; Liu, M.; Zhang, J.C.; Che, R.C. Controllable topological magnetic transformations in the thickness-tunable van der Waals ferromagnet Fe5GeTe2. ACS Nano 2022, 16, 19319–19327. [Google Scholar] [CrossRef]
  139. Yamagami, K.; Fujisawa, Y.; Driesen, B.; Hsu, C.H.; Kawaguchi, K.; Tanaka, H.; Kondo, T.; Zhang, Y.; Wadati, H.; Araki, K.; et al. Itinerant ferromagnetism mediated by giant spin polarization of the metallic ligand band in the van der Waals magnet Fe5GeTe2. Phys. Rev. B 2021, 103, L060403. [Google Scholar] [CrossRef]
  140. Wu, X.C.; Lei, L.; Yin, Q.W.; Zhao, N.N.; Li, M.; Wang, Z.L.; Liu, Q.X.; Song, W.H.; Ma, H.; Ding, P.F.; et al. Direct observation of competition between charge order and itinerant ferromagnetism in the van der Waals crystal Fe5−xGeTe2. Phys. Rev. B 2021, 104, 165101. [Google Scholar] [CrossRef]
  141. Ohta, T.; Tokuda, M.; Iwakiri, S.; Sakai, K.; Driesen, B.; Okada, Y.; Kobayashi, K.; Niimi, Y. Butterfly-shaped magnetoresistance in van der Waals ferromagnet Fe5GeTe2. AIP Adv. 2021, 11, 025014. [Google Scholar] [CrossRef]
  142. Zhang, H.R.; Chen, R.; Zhai, K.; Chen, X.; Caretta, L.; Huang, X.X.; Chopdekar, R.V.; Cao, J.H.; Sun, J.R.; Yao, J.; et al. Itinerant ferromagnetism in van der Waals Fe5−xGeTe2 crystals above room temperature. Phys. Rev. B 2020, 102, 144425. [Google Scholar] [CrossRef]
  143. Stahl, J.; Shlaen, E.; Johrendt, D. The van der Waals ferromagnets Fe5−δGeTe2 and Fe5−δ−xNixGeTe2—Crystal structure, stacking faults, and magnetic properties. Z. Anorg. Allg. Chem. 2018, 644, 1923–1929. [Google Scholar] [CrossRef]
  144. Li, Z.X.; Xia, W.; Su, H.; Yu, Z.H.; Fu, Y.P.; Chen, L.M.; Wang, X.; Yu, N.; Zou, Z.Q.; Guo, Y.F. Magnetic critical behavior of the van der Waals Fe5GeTe2 crystal with near room temperature ferromagnetism. Sci. Rep. 2020, 10, 15345. [Google Scholar] [CrossRef]
  145. Sun, Y.F.; Liu, K. Strain engineering in functional 2-dimensional materials. J. Appl. Phys. 2019, 125, 082402. [Google Scholar] [CrossRef]
  146. Ren, H.T.; Xiang, G.; Lu, J.T.; Zhang, X.; Zhang, L. Biaxial strain-mediated room temperature ferromagnetism of ReS2 web buckles. Adv. Electron. Mater. 2019, 5, 1900814. [Google Scholar] [CrossRef]
  147. Pajda, M.; Kudrnovský, J.; Turek, I.; Drchal, V.; Bruno, P. Ab initio calculations of exchange interactions, spin-wave stiffness constants, and curie temperatures of Fe, Co, and Ni. Phys. Rev. B 2001, 64, 174402. [Google Scholar] [CrossRef]
  148. Wang, Q.H.; Bedoya-Pinto, A.; Blei, M.; Dismukes, A.H.; Hamo, A.; Jenkins, S.; Koperski, M.; Liu, Y.; Sun, Q.C.; Telford, E.J.; et al. The magnetic genome of two-dimensional van der Waals materials. ACS Nano 2022, 16, 6960–7079. [Google Scholar] [CrossRef]
  149. Torelli, D.; Olsen, T. Calculating critical temperatures for ferromagnetic order in two-dimensional materials. 2D Mater. 2019, 6, 015028. [Google Scholar] [CrossRef]
  150. 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 2015, 91, 144425. [Google Scholar] [CrossRef]
  151. Asselin, P.; Evans, R.F.L.; Barker, J.; Chantrell, R.W.; Yanes, R.; Chubykalo-Fesenko, O.; Hinzke, D.; Nowak, U. Constrained Monte Carlo method and calculation of the temperature dependence of magnetic anisotropy. Phys. Rev. B 2010, 82, 054415. [Google Scholar] [CrossRef]
  152. Li, Z.W.; Lv, Y.W.; Ren, L.W.; Li, J.; Kong, L.G.; Zeng, Y.J.; Tao, Q.Y.; Wu, R.X.; Ma, H.F.; Zhao, B.; et al. Efficient strain modulation of 2D materials via polymer encapsulation. Nat. Commun. 2020, 11, 1151. [Google Scholar] [CrossRef]
  153. Zhang, Y.; Shen, L.K.; Liu, M.; Li, X.; Lu, X.L.; Lu, L.; Ma, C.R.; You, C.Y.; Chen, A.P.; Huang, C.W.; et al. Flexible quasi-two-dimensional CoFe2O4 epitaxial thin films for continuous strain tuning of magnetic properties. ACS Nano 2017, 11, 8002–8009. [Google Scholar] [CrossRef]
  154. Lin, Z.S.; Lohmann, M.; Ali, Z.A.; Tang, C.; Li, J.X.; Xing, W.Y.; Zhong, J.N.; Jia, S.; Han, W.; Coh, S.; et al. Pressure-induced spin reorientation transition in layered ferromagnetic insulator Cr2Ge2Te6. Phys. Rev. Mater. 2018, 2, 051004. [Google Scholar] [CrossRef]
  155. Sun, Y.; Xiao, R.C.; Lin, G.T.; Zhang, R.R.; Ling, L.S.; Ma, Z.W.; Luo, X.; Lu, W.J.; Sun, Y.P.; Sheng, Z.G. Effects of hydrostatic pressure on spin-lattice coupling in two-dimensional ferromagnetic Cr2Ge2Te6. Appl. Phys. Lett. 2018, 112, 072409. [Google Scholar] [CrossRef]
  156. Song, T.C.; Fei, Z.Y.; Yankowitz, M.; Lin, Z.; Jiang, Q.N.; Hwangbo, K.; Zhang, Q.; Sun, B.S.; Taniguchi, T.; Watanabe, K.; et al. Switching 2D magnetic states via pressure tuning of layer stacking. Nat. Mater. 2019, 18, 1298–1302. [Google Scholar] [CrossRef] [PubMed]
  157. Yang, L.Y.; Sinitsyn, N.A.; Chen, W.B.; Yuan, J.T.; Zhang, J.; Lou, J.; Crooker, S.A. Long-lived nanosecond spin relaxation and spin coherence of electrons in monolayer MoS2 and WS2. Nat. Phys. 2015, 11, 830–834. [Google Scholar] [CrossRef]
  158. Jones, A.M.; Yu, H.Y.; Ghimire, N.J.; Wu, S.F.; Aivazian, G.; Ross, J.S.; Zhao, B.; Yan, J.Q.; Mandrus, D.G.; Xiao, D.; et al. Optical generation of excitonic valley coherence in monolayer WSe2. Nat. Nanotechnol. 2013, 8, 634–638. [Google Scholar] [CrossRef]
  159. Weisheit, M.; Fähler, S.; Marty, A.; Souche, Y.; Poinsignon, C.; Givord, D. Electric field-induced modification of magnetism in thin-film ferromagnets. Science 2007, 315, 349–351. [Google Scholar] [CrossRef] [PubMed]
  160. Chiba, D.; Fukami, S.; Shimamura, K.; Ishiwata, N.; Kobayashi, K.; Ono, T. Electrical control of the ferromagnetic phase transition in cobalt at room temperature. Nat. Mater. 2011, 10, 853–856. [Google Scholar] [CrossRef] [PubMed]
  161. Ovchinnikov, I.V.; Wang, K.L. Theory of electric-field-controlled surface ferromagnetic transition in metals. Phys. Rev. B 2009, 79, 020402. [Google Scholar] [CrossRef]
  162. Maruyama, T.; Shiota, Y.; Nozaki, T.; Ohta, K.; Toda, N.; Mizuguchi, M.; Tulapurkar, A.A.; Shinjo, T.; Shiraishi, M.; Mizukami, S.; et al. Large voltage-induced magnetic anisotropy change in a few atomic layers of iron. Nat. Nanotechnol. 2009, 4, 158–161. [Google Scholar] [CrossRef]
  163. Tong, Q.J.; Chen, M.X.; Yao, W. Magnetic proximity effect in a van der Waals moire superlattice. Phys. Rev. Appl. 2019, 12, 024031. [Google Scholar] [CrossRef]
  164. Dolui, K.; Petrovic, M.D.; Zollner, K.; Plechac, P.; Fabian, J.; Nikolic, B.K. Proximity spin-orbit torque on a two-dimensional magnet within van der Waals heterostructure: Current-driven antiferromagnet-to-ferromagnet reversible nonequilibrium phase transition in bilayer CrI3. Nano Lett. 2020, 20, 2288–2295. [Google Scholar] [CrossRef] [PubMed]
  165. Karpiak, B.; Cummings, A.W.; Zollner, K.; Vila, M.; Khokhriakov, D.; Hoque, A.M.; Dankert, A.; Svedlindh, P.; Fabian, J.; Roche, S.; et al. Magnetic proximity in a van der Waals heterostructure of magnetic insulator and graphene. 2D Mater. 2020, 7, 015026. [Google Scholar] [CrossRef]
  166. Tang, C.L.; Zhang, Z.W.; Lai, S.; Tan, Q.H.; Gao, W.B. Magnetic proximity effect in graphene/CrBr3 van der Waals heterostructures. Adv. Mater. 2020, 32, 1908498. [Google Scholar] [CrossRef]
  167. Zhong, D.; Seyler, K.L.; Linpeng, X.Y.; Wilson, N.P.; Taniguchi, T.; Watanabe, K.; McGuire, M.A.; Fu, K.C.; Xiao, D.; Yao, W.; et al. Layer-resolved magnetic proximity effect in van der Waals heterostructures. Nat. Nanotechnol. 2020, 15, 187–191. [Google Scholar] [CrossRef]
  168. Liu, N.S.; Zhou, S.; Zhao, J.J. High-curie-temperature ferromagnetism in bilayer CrI3 on bulk semiconducting substrates. Phys. Rev. Mater. 2020, 4, 094003. [Google Scholar] [CrossRef]
  169. Dong, X.J.; You, J.Y.; Zhang, Z.; Gu, B.; Su, G. Great enhancement of curie temperature and magnetic anisotropy in two-dimensional van der Waals magnetic semiconductor heterostructures. Phys. Rev. B 2020, 102, 144443. [Google Scholar] [CrossRef]
  170. Qin, B.; Huifang Ma, H.F.; Hossain, M.; Zhong, M.Z.; Xia, Q.L.; Li, B.; Duan, X.D. Substrates in the synthesis of two-dimensional materials via chemical vapor deposition. Chem. Mater. 2020, 32, 10321–10347. [Google Scholar] [CrossRef]
  171. Sun, Y.H.; Wang, R.M.; Liu, K. Substrate induced changes in atomically thin 2-dimensional semiconductors: Fundamentals, engineering, and applications. Appl. Phys. Rev. 2017, 4, 011301. [Google Scholar] [CrossRef]
  172. Ren, H.T.; Xiang, G. Recent progress in research on ferromagnetic rhenium disulfide. Nanomaterials 2022, 19, 3451. [Google Scholar] [CrossRef] [PubMed]
  173. Ren, H.T.; Xiang, G.; Gu, G.X.; Zhang, X. Enhancement of ferromagnetism of ZnO: Co nanocrystals by post-annealing treatment: The role of oxygen interstitials and zinc vacancies. Mater. Lett. 2014, 122, 256–260. [Google Scholar] [CrossRef]
  174. Chen, M.Y.; Hu, C.; Luo, X.F.; Hong, A.J.; Yu, T.; Yuan, C.L. Ferromagnetic behaviors in monolayer MoS2 introduced by nitrogen-doping. Appl. Phys. Lett. 2020, 116, 073102. [Google Scholar] [CrossRef]
  175. Tan, H.; Hu, W.; Wang, C.; Ma, C.; Duan, H.L.; Yan, W.S.; Cai, L.; Guo, P.; Sun, Z.H.; Liu, Q.H.; et al. Intrinsic ferromagnetism in Mn-substituted MoS2 nanosheets achieved by supercritical hydrothermal reaction. Small 2017, 13, 1701389. [Google Scholar] [CrossRef]
  176. Wang, J.Q.; Sun, F.; Yang, S.; Li, Y.T.; Zhao, C.; Xu, M.W.; Zhang, Y.; Zeng, H. Robust ferromagnetism in Mn-doped MoS2 nanostructures. Appl. Phys. Lett. 2016, 109, 092401. [Google Scholar] [CrossRef]
  177. Gweon, H.K.; Lee, S.Y.; Kwon, H.Y.; Jeong, J.; Chang, H.J.; Kim, K.W.; Qiu, Z.Q.; Ryu, H.; Jang, C.; Choi, J.W. Exchange bias in weakly interlayer-coupled van der Waals magnet Fe3GeTe2. Nano Lett. 2021, 21, 1672–1678. [Google Scholar] [CrossRef]
  178. Jiang, P.; Wang, C.; Chen, D.; Zhong, Z.; Yuan, Z.; Lu, Z.-Y.; Ji, W. Stacking tunable interlayer magnetism in bilayer CrI3. Phys. Rev. B 2019, 99, 144401. [Google Scholar] [CrossRef]
  179. Sivadas, N.; Satoshi Okamoto, S.; Xu, X.D.; Fennie, C.J.; Xiao, D. Stacking-dependent magnetism in bilayer CrI3. Nano Lett. 2018, 18, 7658–7664. [Google Scholar] [CrossRef]
  180. Cao, Y.; Fatemi, V.; Fang, S.A.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; Jarillo-Herrero, P. Unconventional superconductivity in magic-angle graphene superlattices. Nature 2018, 556, 43–50. [Google Scholar] [CrossRef]
  181. Akram, M.; Erten, O. Skyrmions in twisted van der Waals magnets. Phys. Rev. B 2021, 103, L140406. [Google Scholar] [CrossRef]
  182. Tong, Q.J.; Liu, F.; Xiao, J.; Yao, W. Skyrmions in the moiré of van der Waals 2D magnets. Nano Lett. 2018, 18, 7194–7199. [Google Scholar] [CrossRef]
  183. Wang, C.; Gao, Y.; Lv, H.Y.; Xu, X.D.; Xiao, D. Stacking domain wall magnons in twisted van der Waals magnets. Phys. Rev. Lett. 2020, 125, 247201. [Google Scholar] [CrossRef] [PubMed]
  184. Ghader, D. Magnon magic angles and tunable hall conductivity in 2D twisted ferromagnetic bilayers. Sci. Rep. 2020, 10, 15069. [Google Scholar] [CrossRef] [PubMed]
  185. Xie, H.C.; Luo, X.P.; Ye, G.H.; Ye, Z.P.; Ge, H.W.; Sung, S.H.; Rennich, E.; Yan, S.H.; Fu, Y.; Tian, S.J.; et al. Twist engineering of the two-dimensional magnetism in double bilayer chromium triiodide homostructures. Nat. Phys. 2021, 18, 30–36. [Google Scholar] [CrossRef]
  186. Guo, H.W.; Hu, Z.; Liu, Z.B.; Tian, J.G. Stacking of 2D materials. Adv. Funct. Mater. 2020, 31, 2007810. [Google Scholar] [CrossRef]
  187. Nguyen, G.D.; Lee, J.; Berlijn, T.; Zou, Q.; Hus, S.M.; Park, J.; Gai, Z.; Lee, C.; Li, A.P. Visualization and manipulation of magnetic domains in the quasi-two-dimensional material Fe3GeTe2. Phys. Rev. B 2018, 97, 014425. [Google Scholar] [CrossRef]
  188. Yin, S.Q.; Zhao, L.; Song, C.; Huang, Y.; Gu, Y.D.; Chen, R.Y.; Zhu, W.X.; Sun, Y.M.; Jiang, W.J.; Zhang, X.Z.; et al. Evolution of domain structure in Fe3GeTe2. Chin. Phys. B 2021, 30, 027505. [Google Scholar] [CrossRef]
  189. Moriya, T. Theory of itinerant electron magnetism. J. Magn. Magn. Mater. 1991, 100, 261–271. [Google Scholar] [CrossRef]
  190. Wu, B.C.; Fang, S.B.; Yang, J.; Liu, S.Q.; Peng, Y.X.; Li, Q.H.; Lin, Z.C.; Shi, J.J.; Yang, W.Y.; Luo, Z.C.; et al. High-performance FexGeTe2-based (X = 4 or 5) van der Waals magnetic yunnel junctions. Phys. Rev. Appl. 2023, 19, 024037. [Google Scholar] [CrossRef]
  191. Wang, H.Y.; Wu, H.; Zhang, J.; Liu, Y.J.; Chen, D.D.; Pandey, C.; Yin, J.L.; Wei, D.H.; Lei, N.; Shi, S.Y.; et al. Room temperature energy-efficient spin-orbit torque switching in two-dimensional van der Waals Fe3GeTe2 induced by topological insulators. Nat. Commun. 2023, 14, 5173. [Google Scholar] [CrossRef] [PubMed]
  192. Ding, B.; Li, Z.F.; Xu, G.Z.; Li, H.; Hou, Z.P.; Liu, E.K.; Xi, X.K.; Xu, F.; Yao, Y.; Wang, W.H. Observation of magnetic skyrmion bubbles in a van der Waals ferromagnet Fe3GeTe2. Nano Lett. 2020, 20, 868–873. [Google Scholar] [CrossRef] [PubMed]
  193. Park, T.E.; Peng, L.C.; Liang, J.H.; Hallal, A.; Yasin, F.S.; Zhang, X.C.; Song, K.M.; Kim, S.J.; Kim, K.; Weigand, M.; et al. Neel-type skyrmions and their current-induced motion in van der Waals ferromagnet-based heterostructures. Phys. Rev. B 2021, 103, 104410. [Google Scholar] [CrossRef]
  194. Liu, C.; Jiang, J.W.; Zhang, C.H.; Wang, Q.P.; Zhang, H.; Zheng, D.X.; Li, Y.; Ma, Y.C.; Algaidi, H.; Gao, X.S.; et al. Controllable skyrmionic phase transition between Néel skyrmions and Bloch skyrmionic bubbles in van der Waals ferromagnet Fe3−δGeTe2. Adv. Sci. 2023, 10, 2303443. [Google Scholar] [CrossRef]
  195. Zhang, C.H.; Liu, C.; Zhang, S.F.; Zhou, B.J.; Guan, C.S.; Ma, Y.C.; Algaidi, H.; Zheng, D.X.; Li, Y.; He, X.; et al. Magnetic skyrmions with unconventional helicity polarization in a van der Waals ferromagnet. Adv. Mater. 2022, 34, 2204163. [Google Scholar] [CrossRef]
  196. Schmitt, M.; Denneulin, T.; Kovács, A.; Saunderson, T.G.; Rüßmann, P.; Shahee, A.; Scholz, T.; Tavabi, A.H.; Gradhand, M.; Mavropoulos, P.; et al. Skyrmionic spin structures in layered Fe5GeTe2 up to room temperature. Commun. Phys. 2022, 5, 254. [Google Scholar] [CrossRef]
  197. Gao, Y.; Yin, Q.W.; Wang, Q.; Li, Z.L.; Cai, J.W.; Zhao, T.Y.; Lei, H.C.; Wang, S.G.; Zhang, Y.; Shen, B.G. Spontaneous (anti)meron chains in the domain walls of van der Waals ferromagnetic Fe5−XGeTe2. Adv. Mater. 2022, 32, 2005228. [Google Scholar] [CrossRef]
  198. Zhu, W.X.; Song, C.; Wang, Q.; Bai, H.; Yin, S.Q.; Pan, F. Anomalous displacement reaction for synthesizing above-room-temperature and air-stable vdw ferromagnet PtTe2Ge1/3. Natl. Sci. Rev. 2022, 10, nwac173. [Google Scholar] [CrossRef]
  199. Tu, Z.Y.; Xie, T.; Lee, Y.; Zhou, J.L.; Admasu, A.S.; Gong, Y.; Valanoor, N.; Cumings, J.; Cheong, S.W.; Takeuchi, I.; et al. Ambient effect on the curie temperatures and magnetic domains in metallic two-dimensional magnets. npj 2D Mater. Appl. 2021, 5, 62. [Google Scholar] [CrossRef]
  200. Cao, W.; Bu, H.M.; Vinet, M.; Cao, M.; Takagi, S.; Hwang, S.; Ghani, T.; Banerjee, K. The future transistors. Nature 2023, 620, 501–515. [Google Scholar] [CrossRef]
Figure 1. Overview of six synthesis methods and ten strategies for ferromagnetic FexGeTe2 (3 ≤ x ≤ 7) materials. Black font represents the synthesis methods and tunable strategies for obtaining the RTFM; white font represents the synthesis methods and strategies that can tune magnetism in experiments; green font represents only theoretically achievable tunable strategies.
Figure 1. Overview of six synthesis methods and ten strategies for ferromagnetic FexGeTe2 (3 ≤ x ≤ 7) materials. Black font represents the synthesis methods and tunable strategies for obtaining the RTFM; white font represents the synthesis methods and strategies that can tune magnetism in experiments; green font represents only theoretically achievable tunable strategies.
Molecules 28 07244 g001
Figure 2. Crystal structure of FexGeTe2 (3 ≤ x ≤ 7). (A) Crystal structure of FGT monolayer. Inequivalent Fe sites are numbered by I and II. Reprinted with permission from [46]. Copyright 2016, American Physical Society. (B) Schematic illustration of Te–substituted Fe7Ge4 crystal and of five structures in the series FexGeTe2 (4 ≤ x ≤ 7). (C) Stacked plane views along the [001] direction of FexGeTe2. Reprinted with permission from [64]. Copyright 2022, Springer Nature.
Figure 2. Crystal structure of FexGeTe2 (3 ≤ x ≤ 7). (A) Crystal structure of FGT monolayer. Inequivalent Fe sites are numbered by I and II. Reprinted with permission from [46]. Copyright 2016, American Physical Society. (B) Schematic illustration of Te–substituted Fe7Ge4 crystal and of five structures in the series FexGeTe2 (4 ≤ x ≤ 7). (C) Stacked plane views along the [001] direction of FexGeTe2. Reprinted with permission from [64]. Copyright 2022, Springer Nature.
Molecules 28 07244 g002
Figure 3. SSR–prepared FGT single crystals and polycrystalline structures. Fe2.9GeTe2 crystal structure: (A) unit cell and (B) polyhedral representation. The Fe3 position is vacant and shown only for comparison. (C) X–ray diffraction (XRD) patterns and energy–dispersive X–ray spectroscopy (EDXS) composition maps. * represents a second Fe1+δTe phase; # represents an impurity FeTe2 phase. On the composition maps, Fe is presented in red, Ge in green, and Te in blue. (D) Magnetic susceptibility χ versus T plot in different applied fields. Reprinted with permission from [36]. Copyright 2015, American Chemical Society. (E) FGT crystal structure. (F) XRD. (G) Size distribution of the plate–like FGT single crystal. The inset is the optical photo of FGT. Reprinted with permission from [120]. Copyright 2022, American Chemical Society.
Figure 3. SSR–prepared FGT single crystals and polycrystalline structures. Fe2.9GeTe2 crystal structure: (A) unit cell and (B) polyhedral representation. The Fe3 position is vacant and shown only for comparison. (C) X–ray diffraction (XRD) patterns and energy–dispersive X–ray spectroscopy (EDXS) composition maps. * represents a second Fe1+δTe phase; # represents an impurity FeTe2 phase. On the composition maps, Fe is presented in red, Ge in green, and Te in blue. (D) Magnetic susceptibility χ versus T plot in different applied fields. Reprinted with permission from [36]. Copyright 2015, American Chemical Society. (E) FGT crystal structure. (F) XRD. (G) Size distribution of the plate–like FGT single crystal. The inset is the optical photo of FGT. Reprinted with permission from [120]. Copyright 2022, American Chemical Society.
Molecules 28 07244 g003
Figure 4. CVT-prepared FGT bulk single crystals. (A) XRD. The inset shows a photograph of FGT single crystals on a 1 mm grid. Reprinted with permission from [12]. Copyright 2017, American Physical Society. (B) XRD. The inset is the optical image of this FGT crystal. Reprinted with permission from [49]. Copyright 2021, American Physical Society. (C) Process of fabricating the nanodevices. (D) Thickness inhomogeneity in the FGT nanodevice (E). Reprinted with permission from [49]. Copyright 2021, American Physical Society.
Figure 4. CVT-prepared FGT bulk single crystals. (A) XRD. The inset shows a photograph of FGT single crystals on a 1 mm grid. Reprinted with permission from [12]. Copyright 2017, American Physical Society. (B) XRD. The inset is the optical image of this FGT crystal. Reprinted with permission from [49]. Copyright 2021, American Physical Society. (C) Process of fabricating the nanodevices. (D) Thickness inhomogeneity in the FGT nanodevice (E). Reprinted with permission from [49]. Copyright 2021, American Physical Society.
Molecules 28 07244 g004
Figure 5. FexGeTe2 single crystal prepared via self–flux. (A) Schematic diagram of the flux-assisted growth process of FexGeTe2. (B) Cross–sectional structures of FGT, F4GT, and F5GT. (C) Optical images. Scale bars: 10 μm (Mica); 5 μm. (D) Optical images of FGT nanosheets grown on sapphire at 660, 670, 680, 690, and 700 °C, respectively. Scale bars: 10 μm. (E) Thickness and size evolution of FGT with temperature increase. (F) Differences in thicknesses and lateral sizes using different cosolvents. (G) Comparison of FGT under the same growth conditions. Reprinted with permission from [44]. Copyright 2022, American Chemical Society.
Figure 5. FexGeTe2 single crystal prepared via self–flux. (A) Schematic diagram of the flux-assisted growth process of FexGeTe2. (B) Cross–sectional structures of FGT, F4GT, and F5GT. (C) Optical images. Scale bars: 10 μm (Mica); 5 μm. (D) Optical images of FGT nanosheets grown on sapphire at 660, 670, 680, 690, and 700 °C, respectively. Scale bars: 10 μm. (E) Thickness and size evolution of FGT with temperature increase. (F) Differences in thicknesses and lateral sizes using different cosolvents. (G) Comparison of FGT under the same growth conditions. Reprinted with permission from [44]. Copyright 2022, American Chemical Society.
Molecules 28 07244 g005
Figure 6. (A) Schematic diagram of TS–LPE. SEM images of FGT crystals (B) in a pristine state and after ball milling (C), stirring (D), and sonication (E). (F) TEM image of FGT nanoflakes after centrifugation. (G) Statistical graph of the influence of the ball–milling time on the center size and the yield. (H) XRD peaks and (I) enlarged view of (002) peaks. FGT−1, FGT−2, and FGT−3 are the samples after ball milling (C), stirring (D), and sonication (E), respectively. (J) Variation in the cell energy with the lattice constant in the c-axis direction. Reproduced with permission from [50]. Copyright 2022, American Chemical Society.
Figure 6. (A) Schematic diagram of TS–LPE. SEM images of FGT crystals (B) in a pristine state and after ball milling (C), stirring (D), and sonication (E). (F) TEM image of FGT nanoflakes after centrifugation. (G) Statistical graph of the influence of the ball–milling time on the center size and the yield. (H) XRD peaks and (I) enlarged view of (002) peaks. FGT−1, FGT−2, and FGT−3 are the samples after ball milling (C), stirring (D), and sonication (E), respectively. (J) Variation in the cell energy with the lattice constant in the c-axis direction. Reproduced with permission from [50]. Copyright 2022, American Chemical Society.
Molecules 28 07244 g006
Figure 7. MBE–prepared FGT thin films. (A) Magnetic phase diagram of FGT and Bi2Te3|FGT versus layer thickness and temperature. Reproduced with permission from [56]. Copyright 2020, American Chemical Society. (B) TC for F4GT thin films with the thickness. (C) TC versus the c–lattice constants of F4GT. Reproduced with permission from [59]. Copyright 2023, Springer Nature.
Figure 7. MBE–prepared FGT thin films. (A) Magnetic phase diagram of FGT and Bi2Te3|FGT versus layer thickness and temperature. Reproduced with permission from [56]. Copyright 2020, American Chemical Society. (B) TC for F4GT thin films with the thickness. (C) TC versus the c–lattice constants of F4GT. Reproduced with permission from [59]. Copyright 2023, Springer Nature.
Molecules 28 07244 g007
Figure 8. (A) Normalized lattice parameters as a function of refined Fe content for FGT samples (powder X-ray diffraction (p-xrd) at room temperature), including results obtained from single-crystal X-ray (sc-xrd) diffraction data collected at 173 K. (B) TC as a function of lattice parameters. Reprinted with permission from [9]. Copyright 2016, American Physical Society.
Figure 8. (A) Normalized lattice parameters as a function of refined Fe content for FGT samples (powder X-ray diffraction (p-xrd) at room temperature), including results obtained from single-crystal X-ray (sc-xrd) diffraction data collected at 173 K. (B) TC as a function of lattice parameters. Reprinted with permission from [9]. Copyright 2016, American Physical Society.
Molecules 28 07244 g008
Figure 9. (A) Calculated magnetic anisotropy energy (MAE) and magnetic moment per atom for various Fe2+/Fe3+ ratios (x). (B) Calculated normalized magnetization of Fe atoms in FenGeTe2 as a function of temperature from Monte Carlo simulation. Reprinted with permission from [64]. Copyright 2022, Springer Nature. Crystal structure (CF) of FexGeTe2 (3 ≤ x ≤ 5) monolayer. Side Views of the FGT (C), F4GT (D), UUU Fe5GT (E), and UDU F5GT monolayer (F). The half–colored circles in Figure 9F show the Fe1–Ge split sites present in the UDU configuration. The lower panel shows a 3D side view of the FexGeTe2 monolayers. Note that: M A E = E E . Reprinted with permission from [65]. Copyright 2023, Springer Nature.
Figure 9. (A) Calculated magnetic anisotropy energy (MAE) and magnetic moment per atom for various Fe2+/Fe3+ ratios (x). (B) Calculated normalized magnetization of Fe atoms in FenGeTe2 as a function of temperature from Monte Carlo simulation. Reprinted with permission from [64]. Copyright 2022, Springer Nature. Crystal structure (CF) of FexGeTe2 (3 ≤ x ≤ 5) monolayer. Side Views of the FGT (C), F4GT (D), UUU Fe5GT (E), and UDU F5GT monolayer (F). The half–colored circles in Figure 9F show the Fe1–Ge split sites present in the UDU configuration. The lower panel shows a 3D side view of the FexGeTe2 monolayers. Note that: M A E = E E . Reprinted with permission from [65]. Copyright 2023, Springer Nature.
Molecules 28 07244 g009
Figure 10. (A) Schematic of the lattice structure of FGT under pressure. (B) Schematic of in situ MCD experimental setup under hydrostatic pressure. The thin FGT sample was covered with a thick h BN to avoid degradation induced by pressure media. P: polarizer; PEM: photoelastic modulator; BS: beam splitter; DAC: diamond anvil cell; LO: 50× long-working distance objective; L: lens; FGT: thin FGT sample. (C) Pressure–dependent phase diagram. The gray region represents a single–domain (SD) ferromagnetic state, the yellow region represents a labyrinthine–domain (LD) ferromagnet, and the pink region shows the paramagnetic state. Reprinted with permission from [78]. Copyright 2019, American Chemical Society. (D) The pressure dependence of TC, where TC is estimated from the minimum point of dM/dT. (E) The variation in Ms for FGT is caused by increasing the pressure. Reprinted with permission from [79]. Copyright 2021, American Physical Society.
Figure 10. (A) Schematic of the lattice structure of FGT under pressure. (B) Schematic of in situ MCD experimental setup under hydrostatic pressure. The thin FGT sample was covered with a thick h BN to avoid degradation induced by pressure media. P: polarizer; PEM: photoelastic modulator; BS: beam splitter; DAC: diamond anvil cell; LO: 50× long-working distance objective; L: lens; FGT: thin FGT sample. (C) Pressure–dependent phase diagram. The gray region represents a single–domain (SD) ferromagnetic state, the yellow region represents a labyrinthine–domain (LD) ferromagnet, and the pink region shows the paramagnetic state. Reprinted with permission from [78]. Copyright 2019, American Chemical Society. (D) The pressure dependence of TC, where TC is estimated from the minimum point of dM/dT. (E) The variation in Ms for FGT is caused by increasing the pressure. Reprinted with permission from [79]. Copyright 2021, American Physical Society.
Molecules 28 07244 g010
Figure 11. (A) The atomic structure of FGT monolayer. (B) Temperature–dependent Hall resistance (Rxy) as a function of the perpendicular magnetic field. (C) Diagram of the experimental configuration for the polar magneto–optical Kerr effect (MOKE) measurements. BS and WS here represent the beam splitter and Wollaston splitter, respectively. (D) Schematic of the laser–excited DOS in few–layered FGT thin films. The photon energy of 3.1 eV causes electron transitions (vertical blue arrows) from occupied states below the Fermi level EF to the unoccupied states above EF. Reprinted with permission from [53]. Copyright 2020, American Physical Society.
Figure 11. (A) The atomic structure of FGT monolayer. (B) Temperature–dependent Hall resistance (Rxy) as a function of the perpendicular magnetic field. (C) Diagram of the experimental configuration for the polar magneto–optical Kerr effect (MOKE) measurements. BS and WS here represent the beam splitter and Wollaston splitter, respectively. (D) Schematic of the laser–excited DOS in few–layered FGT thin films. The photon energy of 3.1 eV causes electron transitions (vertical blue arrows) from occupied states below the Fermi level EF to the unoccupied states above EF. Reprinted with permission from [53]. Copyright 2020, American Physical Society.
Molecules 28 07244 g011
Figure 12. Schematics of the atomic structure (A,B) of the FGT monolayer. The dual–gate configuration for simulating the electrostatic gating is shown in (B). (C) The MAE per unit cell as a function of the charge doping concentration in the single–gate (squares) and dual–gate (dots) configurations. The results obtained using the force theorem are denoted by the black line. Reprinted from Ref. [83] with the permission of AIP Publishing. (D) A series of dI/dV curves taken from a set point of 50−1000 pA at the same position. Reprinted with permission from [84]. Copyright 2021 American Chemical Society.
Figure 12. Schematics of the atomic structure (A,B) of the FGT monolayer. The dual–gate configuration for simulating the electrostatic gating is shown in (B). (C) The MAE per unit cell as a function of the charge doping concentration in the single–gate (squares) and dual–gate (dots) configurations. The results obtained using the force theorem are denoted by the black line. Reprinted from Ref. [83] with the permission of AIP Publishing. (D) A series of dI/dV curves taken from a set point of 50−1000 pA at the same position. Reprinted with permission from [84]. Copyright 2021 American Chemical Society.
Molecules 28 07244 g012
Figure 13. The RMCD signals of FGT flakes with varying thicknesses on Au, SiO2, and Al substrates are shown in (AC), respectively. The saturation fields as a function of temperature for FGT flakes with varying thicknesses on three substrates are shown in (DF). The empty circles are fitted using the function α(1 − T/TC)β. The dotted line corresponds to zero RMCD signal and saturation field. (GI) TC was extracted as a function of the thickness of the FGT flakes on three substrates from RMCD measurements. Reprinted from [88], with the permission of AIP Publishing.
Figure 13. The RMCD signals of FGT flakes with varying thicknesses on Au, SiO2, and Al substrates are shown in (AC), respectively. The saturation fields as a function of temperature for FGT flakes with varying thicknesses on three substrates are shown in (DF). The empty circles are fitted using the function α(1 − T/TC)β. The dotted line corresponds to zero RMCD signal and saturation field. (GI) TC was extracted as a function of the thickness of the FGT flakes on three substrates from RMCD measurements. Reprinted from [88], with the permission of AIP Publishing.
Molecules 28 07244 g013
Figure 14. Phase diagram of a Ni-doped FGT single crystal determined by magnetization and µSR (TM), and resistivity measurements (TC) showing an FM region up to x = 0.3, which is smeared into an FM spin glass [9]. Muon spin relaxation and rotation measurements (µSR); Short-range order (SRO). Reprinted with permission from [38]. Copyright 2018, American Physical Society.
Figure 14. Phase diagram of a Ni-doped FGT single crystal determined by magnetization and µSR (TM), and resistivity measurements (TC) showing an FM region up to x = 0.3, which is smeared into an FM spin glass [9]. Muon spin relaxation and rotation measurements (µSR); Short-range order (SRO). Reprinted with permission from [38]. Copyright 2018, American Physical Society.
Molecules 28 07244 g014
Figure 15. (A) Schematic of the F5GT layer with atomic types and positions labeled. (B) Change in the lattice type (rhombohedral/primitive) with cobalt doping (gold spheres) and the corresponding evolution of the magnetic order and anisotropy (black arrows). (C) Curie and Néel temperatures, magnetization induced along [001] at 300 K, and saturation magnetization. The dashed vertical lines indicate the approximate region where the magnetic anisotropy inverts for the ferromagnetic compositions; the solid lines in the FM region are intended to facilitate viewing. The AFM region is likely characterized by FM planes that are coupled antiferromagnetically along [001]. Reprinted with permission from [95]. Copyright 2020, American Physical Society. (D) Schematic of the melt flux containing CoyFe5−yGeTe2. OM images of CoyFe5−yGeTe2 (E) and NiyFe5−yGeTe2 (E). (F) Comparison of TC values of F5GT, Co1Fe4GeTe2, and Co3Fe2GeTe2. Reprinted with permission from [44]. Copyright 2022, American Chemical Society.
Figure 15. (A) Schematic of the F5GT layer with atomic types and positions labeled. (B) Change in the lattice type (rhombohedral/primitive) with cobalt doping (gold spheres) and the corresponding evolution of the magnetic order and anisotropy (black arrows). (C) Curie and Néel temperatures, magnetization induced along [001] at 300 K, and saturation magnetization. The dashed vertical lines indicate the approximate region where the magnetic anisotropy inverts for the ferromagnetic compositions; the solid lines in the FM region are intended to facilitate viewing. The AFM region is likely characterized by FM planes that are coupled antiferromagnetically along [001]. Reprinted with permission from [95]. Copyright 2020, American Physical Society. (D) Schematic of the melt flux containing CoyFe5−yGeTe2. OM images of CoyFe5−yGeTe2 (E) and NiyFe5−yGeTe2 (E). (F) Comparison of TC values of F5GT, Co1Fe4GeTe2, and Co3Fe2GeTe2. Reprinted with permission from [44]. Copyright 2022, American Chemical Society.
Molecules 28 07244 g015
Figure 16. (A) Schematic of the Hall bar device on a solid proton conductor used for measurements in (B,C), in which the current density is J and a perpendicular magnetic field B is applied. (B) Gate–tuned FM in FGT nanoflakes. (C) Remanence Hall resistance Rxy as a function of temperature. Inset: Schematic of a possible AFM phase in pristine FGT. (D) Effect of the exchange bias after zero–field cooling at 2 K. (E) Schematic of differing AFM–FM interfaces. Reprinted with permission from [100], Copyright 2020, American Physical Society.
Figure 16. (A) Schematic of the Hall bar device on a solid proton conductor used for measurements in (B,C), in which the current density is J and a perpendicular magnetic field B is applied. (B) Gate–tuned FM in FGT nanoflakes. (C) Remanence Hall resistance Rxy as a function of temperature. Inset: Schematic of a possible AFM phase in pristine FGT. (D) Effect of the exchange bias after zero–field cooling at 2 K. (E) Schematic of differing AFM–FM interfaces. Reprinted with permission from [100], Copyright 2020, American Physical Society.
Molecules 28 07244 g016
Figure 17. Schematic diagram (A) and optical image (B) of an F5GT SP–FET, where an F5GT flake lies on the solid proton conductor (SPC). Scale bar: 10 μm. (C) ρxy loops. (D) Gate–voltage–dependent carrier densities and anomalous Hall ratios. (E) Evolution of the energy difference between FM and AFM (ΔE) with charge doping under LDA + U and PBE functionals. Reprinted with permission from [62]. Copyright 2021, American Chemical Society.
Figure 17. Schematic diagram (A) and optical image (B) of an F5GT SP–FET, where an F5GT flake lies on the solid proton conductor (SPC). Scale bar: 10 μm. (C) ρxy loops. (D) Gate–voltage–dependent carrier densities and anomalous Hall ratios. (E) Evolution of the energy difference between FM and AFM (ΔE) with charge doping under LDA + U and PBE functionals. Reprinted with permission from [62]. Copyright 2021, American Chemical Society.
Molecules 28 07244 g017
Figure 18. (A) Optical image of an FGT flake. (B) Height profile. (C) HC of Fe deficient Fe3−xGeTe2 and FGT. The Fe3GeTe2 data are taken from [15] [marked *] and [17] [marked †]. (D) MAE and doping-dependent magnetization. Reprinted with permission from [43]. Copyright 2019, American Chemical Society.
Figure 18. (A) Optical image of an FGT flake. (B) Height profile. (C) HC of Fe deficient Fe3−xGeTe2 and FGT. The Fe3GeTe2 data are taken from [15] [marked *] and [17] [marked †]. (D) MAE and doping-dependent magnetization. Reprinted with permission from [43]. Copyright 2019, American Chemical Society.
Molecules 28 07244 g018
Figure 19. (A) Large–scale STM image of FGT. (B) Atom–resolved STM topography. (C) dI/dV curves. (D) Crystal structure of FGT. (E,F) Side views of the crystal structures for these two situations. (G,H) Illustrate the Kondo screening in a pristine Kondo lattice and in a lattice with a Kondo hole, respectively. Reprinted with permission from [106]. Copyright 2021, American Chemical Society.
Figure 19. (A) Large–scale STM image of FGT. (B) Atom–resolved STM topography. (C) dI/dV curves. (D) Crystal structure of FGT. (E,F) Side views of the crystal structures for these two situations. (G,H) Illustrate the Kondo screening in a pristine Kondo lattice and in a lattice with a Kondo hole, respectively. Reprinted with permission from [106]. Copyright 2021, American Chemical Society.
Molecules 28 07244 g019
Figure 20. (AC) TEM image and representative SAED pattern of NaFe2.78GeTe2 as well as EDX maps of the constituting elements. (D) Temperature dependence of the specific magnetic susceptibility χg at μ0H = 0.01T. (E) MH curves. Reprinted with permission from [107]. Copyright 2019, American Chemical Society.
Figure 20. (AC) TEM image and representative SAED pattern of NaFe2.78GeTe2 as well as EDX maps of the constituting elements. (D) Temperature dependence of the specific magnetic susceptibility χg at μ0H = 0.01T. (E) MH curves. Reprinted with permission from [107]. Copyright 2019, American Chemical Society.
Molecules 28 07244 g020
Figure 21. (A) PEEM topography image of an FGT flake (golden color) on a silicon substrate (purple color). (B,C) Magnetic–stripe domains. (D) Micron–sized diamond–shaped and rectangular–patterned structures. The magnetic stripe contrast (out-of-plane magnetization component, labeled as A and the spatially averaged contrast (in-plane magnetization component, labeled as B from the two selected areas. (E) Temperature dependence of the magnetic stripe contrast (out–of–plane magnetization component) and the spatially averaged contrast (in–plane magnetization component) from the two selected areas. Reprinted with permission from [16]. Copyright 2019, American Chemical Society.
Figure 21. (A) PEEM topography image of an FGT flake (golden color) on a silicon substrate (purple color). (B,C) Magnetic–stripe domains. (D) Micron–sized diamond–shaped and rectangular–patterned structures. The magnetic stripe contrast (out-of-plane magnetization component, labeled as A and the spatially averaged contrast (in-plane magnetization component, labeled as B from the two selected areas. (E) Temperature dependence of the magnetic stripe contrast (out–of–plane magnetization component) and the spatially averaged contrast (in–plane magnetization component) from the two selected areas. Reprinted with permission from [16]. Copyright 2019, American Chemical Society.
Molecules 28 07244 g021
Figure 22. (A,B) DFT–LDA–calculated band structures of the FGT monolayer. Electronic density of states (DOS) for the Fe d states in the non–spin–polarized system in units of states/eV/Fe atom/spin of FGT monolayer. This non–spin–polarized electronic structure (C) is used to obtain the density of states at the Fermi level D(EF). Reprinted with permission from [46]. Copyright 2016, American Physical Society. (DG) DFT + U–calculated band structures of FexGeTe2 (4 ≤ x ≤ 7). The purple and orange lines are spin-up and spin-down bands, respectively. Reprinted with permission from [64]. Copyright 2022, Springer Nature. Spin–resolved band structures and DOS of F4GT (H) bilayer (BL) (H) and F5GT (I) BL. The Fermi level is set to zero. Reprinted with permission from [190]. Copyright 2023, American Physical Society.
Figure 22. (A,B) DFT–LDA–calculated band structures of the FGT monolayer. Electronic density of states (DOS) for the Fe d states in the non–spin–polarized system in units of states/eV/Fe atom/spin of FGT monolayer. This non–spin–polarized electronic structure (C) is used to obtain the density of states at the Fermi level D(EF). Reprinted with permission from [46]. Copyright 2016, American Physical Society. (DG) DFT + U–calculated band structures of FexGeTe2 (4 ≤ x ≤ 7). The purple and orange lines are spin-up and spin-down bands, respectively. Reprinted with permission from [64]. Copyright 2022, Springer Nature. Spin–resolved band structures and DOS of F4GT (H) bilayer (BL) (H) and F5GT (I) BL. The Fermi level is set to zero. Reprinted with permission from [190]. Copyright 2023, American Physical Society.
Molecules 28 07244 g022
Figure 23. (A) Schematic of an FGT/CGT/FGT MTJ and the effect of directional electric fields on the charge transfer. (B) IV curve. The inset is the optical image of this FGT/CGT/FGT MTJ. (C) Out–of–plane magnetic–field–dependent resistance, after subtracting the noise background. The gray–dotted lines indicate the positions of the switching fields. The red (blue) curves present the magnetic field sweeping from negative (positive) to positive (negative). (D) Numerical derivative (dR/dH) curves under negative V bias. Reprinted with permission from [115]. Copyright 2023, American Chemical Society.
Figure 23. (A) Schematic of an FGT/CGT/FGT MTJ and the effect of directional electric fields on the charge transfer. (B) IV curve. The inset is the optical image of this FGT/CGT/FGT MTJ. (C) Out–of–plane magnetic–field–dependent resistance, after subtracting the noise background. The gray–dotted lines indicate the positions of the switching fields. The red (blue) curves present the magnetic field sweeping from negative (positive) to positive (negative). (D) Numerical derivative (dR/dH) curves under negative V bias. Reprinted with permission from [115]. Copyright 2023, American Chemical Society.
Molecules 28 07244 g023
Figure 24. (A) Spin–valve effect in an FGT/hBN/FGT van der Waals heterostructure. (B) Tunneling resistance measured at T = 4.2 K with B applied parallel to the FGT c–axis. Very sharp–resistance jumps are observed for B ≈ ± 0.7 T. The variation in tunneling magnetoresistance (TMR) is ∼160%. Upper panels: zoomed–in view of the magnetoresistance. (C) IV curves measured with the magnetization in the two FGT electrodes pointing parallel (black curve, B = 0 T) and antiparallel (red curve, B = −0.68 T) to each other. The insets show the corresponding configuration in the density of states of majority and minority spins in the two electrodes. (D) Bias dependence of TMR. Reprinted with permission from [18]. Copyright 2018, American Chemical Society.
Figure 24. (A) Spin–valve effect in an FGT/hBN/FGT van der Waals heterostructure. (B) Tunneling resistance measured at T = 4.2 K with B applied parallel to the FGT c–axis. Very sharp–resistance jumps are observed for B ≈ ± 0.7 T. The variation in tunneling magnetoresistance (TMR) is ∼160%. Upper panels: zoomed–in view of the magnetoresistance. (C) IV curves measured with the magnetization in the two FGT electrodes pointing parallel (black curve, B = 0 T) and antiparallel (red curve, B = −0.68 T) to each other. The insets show the corresponding configuration in the density of states of majority and minority spins in the two electrodes. (D) Bias dependence of TMR. Reprinted with permission from [18]. Copyright 2018, American Chemical Society.
Molecules 28 07244 g024
Figure 25. (A) Schematic illustration of the effective field responsible for switching the magnetic state of FGT in FGT/Pt hybrid devices. Jx is the injected current density, Hx is the applied in-plane field, HDL is the effective field from damping–like spin–orbit torques (SOT), and M is the FGT’s magnetization. (B) Hall resistance for our FGT(15 nm)/Pt(5 nm) device with anisotropy field Hk labeled on the graph. (CF) Effective switching current as a function of the applied in–plane negative (E) and positive (F) bias field. The color scale represents the switching resistance as a percentage of the absolute value of the anomalous Hall resistance at zero current RH0. (C,D) correspond to the line cuts in (E,F). Reprinted with permission from Ref. [20]. Copyright 2019, American Chemical Society.
Figure 25. (A) Schematic illustration of the effective field responsible for switching the magnetic state of FGT in FGT/Pt hybrid devices. Jx is the injected current density, Hx is the applied in-plane field, HDL is the effective field from damping–like spin–orbit torques (SOT), and M is the FGT’s magnetization. (B) Hall resistance for our FGT(15 nm)/Pt(5 nm) device with anisotropy field Hk labeled on the graph. (CF) Effective switching current as a function of the applied in–plane negative (E) and positive (F) bias field. The color scale represents the switching resistance as a percentage of the absolute value of the anomalous Hall resistance at zero current RH0. (C,D) correspond to the line cuts in (E,F). Reprinted with permission from Ref. [20]. Copyright 2019, American Chemical Society.
Molecules 28 07244 g025
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

Ren, H.; Lan, M. Progress and Prospects in Metallic FexGeTe2 (3 ≤ x ≤ 7) Ferromagnets. Molecules 2023, 28, 7244. https://doi.org/10.3390/molecules28217244

AMA Style

Ren H, Lan M. Progress and Prospects in Metallic FexGeTe2 (3 ≤ x ≤ 7) Ferromagnets. Molecules. 2023; 28(21):7244. https://doi.org/10.3390/molecules28217244

Chicago/Turabian Style

Ren, Hongtao, and Mu Lan. 2023. "Progress and Prospects in Metallic FexGeTe2 (3 ≤ x ≤ 7) Ferromagnets" Molecules 28, no. 21: 7244. https://doi.org/10.3390/molecules28217244

Article Metrics

Back to TopTop