Next Article in Journal
Effect of Nano TiO2 Flux on Depth of Penetration and Mechanical Properties of TIG-Welded SA516 Grade 70 Steel Joints—An Experimental Investigation
Previous Article in Journal
The Effect of Ti and Mo Microalloying on Hydrogen Embrittlement Resistance of Ultra-High Strength Medium Mn Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Fold Enhancement of Curie Temperature in Monolayer CrI3 by High Pressure

1
Department of Physics, Yantai University, Yantai 264005, China
2
Micro-Electronics Research Institute and School of Electronics and Information, Hangzhou Dianzi University, Hangzhou 310018, China
3
School of Information Engineering, Yantai Institute of Technology, Yantai 264005, China
*
Author to whom correspondence should be addressed.
Metals 2025, 15(4), 398; https://doi.org/10.3390/met15040398
Submission received: 3 February 2025 / Revised: 26 March 2025 / Accepted: 27 March 2025 / Published: 2 April 2025

Abstract

:
In recent years, the discovery of the two-dimensional (2D) intrinsically ferromagnetic monolayer CrI3 has opened up promising avenues for the advancement of spintronic devices. Nevertheless, the relatively low Curie temperature poses a significant challenge for practical applications. Herein, we determine changes in the superexchange interaction of ferromagnetic coupling caused under pressure by using first-principles calculations and Monte Carlo simulations. Based on the superexchange interaction of ferromagnetic coupling, the effect of applying high pressure on the Curie temperature of monolayer CrI3 is investigated. With a pressure coefficient of 2.0%, the Curie temperature is enhanced to 97.3 K, which is nearly double that of the monolayer CrI3 without pressure. In addition, the direction of the easy magnetization axis changes from the out-of-plane to the in-plane one when the pressure coefficient is 1.2%. Meanwhile, the band gap of monolayer CrI3 can be transformed from indirect to direct by applying high pressure. Our work enriches the process of modulating the magnetic and electronic properties of 2D monolayer materials.

1. Introduction

The pursuit of intrinsic magnetism in two-dimensional (2D) monolayer materials has been an ongoing research focus for an extended period [1,2] due to the unique physical properties and the promising applications in spintronics [3,4,5,6,7], magneto-optoelectronics [8,9,10], and data storage [11]. In 2017, intrinsic ferromagnetism was first observed in atomically thin layers of CrI3 and Cr2Ge2Te6, which opened new avenues for exploring low-dimensional magnetic performance and spintronics devices [8,12]. After that, more and more monolayer 2D materials [13,14,15,16] with intrinsic magnetism were investigated experimentally and theoretically, such as CrBr3 [17], VBr3 [18], VI3 [19], NiBr3 [20], FePS3 [21], and VSe2 [22]. Moreover, many 2D intrinsic magnetic materials have been predicted with the assistance of high-throughput screening and machine learning [23,24,25,26,27,28]. However, most 2D monolayer magnetic materials have a common weakness, which is that the Curie temperature (TC) is much lower than room temperature, which greatly hinders their applications.
As a representative 2D intrinsically ferromagnetic (FM) materials, the monolayer CrI3 is a potential candidate for practical application because of its distinct spin-lattice coupling [29] and strong perpendicular magnetic anisotropy [30]. Nevertheless, the TC of monolayer CrI3 is only 45 K [8], which makes enhancing its TC a great challenge. Many attempts have been made to improve the TC of monolayer CrI3, such as constructing heterostructures [31,32], adsorbing small molecules [33,34], and doping foreign elements [17,35,36]. For example, the TC of 2D monolayer can be raised to 99.3 K by constructing a CrI3/α-In2Se3 heterojunction [15]. Constructing a heterojunction with GaN is another effective method, which can increase the TC of monolayer CrI3 to 75 K. Meanwhile, the easy magnetization axis transforms from an out-of-plane to an in-plane orientation, with electron doping levels exceeding 0.2 electrons per unit cell [31]. Moreover, the adsorption of gas molecules, including CO, H2, N2, and H2O, can substantially enhance FM coupling in monolayer CrI3 [33]. Specifically, the adsorption of N2 can significantly improve the TC of the pristine monolayer CrI3 by approximately 2.6 times [33]. It has been reported that doping with RE atoms (RE = Sc, Y, La, Ce, Pr, Nd, Pm, Eu, Gd, Tm, and Lu) enables a significant enhancement of the Curie temperature and the control of total magnetic moments of the CrI3 monolayer, resulting in an obvious increase in TC for monolayer CrGdI6 to 347.58 K, as shown in [37].
To further investigate alternative approaches to enhancing the TC in monolayer CrI3, it is essential to elucidate the fundamental mechanisms that are responsible for FM coupling. In monolayer CrI3, the magnetic properties originate from the partially filled Cr-3d orbitals. The Cr3+ ion resides within an octahedral crystal field formed by six surrounding ligands, leading to a splitting of the Cr-3d orbitals into higher-energy, doubly degenerate eg orbitals and lower-energy, triply degenerate t2g orbitals. The three 3d electrons of Cr3+ occupy the lower-energy t2g orbitals. Mediated by the intermediate I-5p orbital, the spin direction of the electrons in the Cr-3d orbitals is consistently aligned, resulting in a superexchange interaction between the two Cr ions adjacent to the I ion, which is driven by FM coupling. It is known that the strength of the superexchange interaction is decisively affected by the overlap degree of the wave functions for Cr-d and I-p orbits, which is reflected by the bond length l of Cr-I and the bond angle θ of Cr-I-Cr [38]. Generally, the shorter l is and the nearer θ is to 180°, the stronger the exchange interaction, is and the higher the TC is [39,40].
According to our recent work [41], introducing I-vacancies and interstitial H-atoms into pristine monolayer CrI3 can induce internal pressure, leading to obvious changes in the Cr-I bond length (l) and the Cr-I-Cr bond angle (θ), which remarkably improves the value of the TC. When an interstitial hydrogen atom is introduced into the 2 × 2 × 1 CrI3 supercell, the TC can be enhanced to 112.4 K.
Inspired by the aforementioned results, in this paper, we attempt to investigate the effect of applying external pressure on the physical properties and TC of monolayer CrI3 by using first-principles calculations and Monte Carlo (MC) simulations. We establish a 2 × 2 × 1 CrI3 supercell to simulate the magnetic and electronic properties of monolayer CrI3 under different pressures. One novelty of this study is the use of first-principles calculations and Monte Carlo simulations to determine changes in the superexchange interaction of ferromagnetic coupling caused by pressure, which helps with understanding the relationships between the pressure and the TC of monolayer CrI3. Our work illustrates the promising potential of structural changes induced by high pressure in modulating the magnetization orientation and enhancing magnetism in monolayer CrI3.

2. Materials and Methods

Our first-principles calculations were conducted using the Vienna Ab initio Simulation Package (VASP), employing density functional theory (DFT) [42]. The generalized gradient approximation (GGA) exchange correlation was formulated according to the Perdew–Burke–Ernzerhof (PBE) approach [43]. Taking into account the correlation effects of Cr-3d electrons, the GGA + U method [44] was adopted. The main reason was that there is a strong coulomb interaction between the 3d orbital electrons of Cr in CrI3, and the general DFT exchange correlation function is not sufficient to describe the above coulomb interaction, resulting in the close proximity or even overlap of the orbitals. The addition of U and J take into account the coulomb repulsion between the local electrons with opposite spin on the same atom, resulting in the splitting of the energy levels, so that the theoretical band gap value is closer to the experimental value. Respectively, the on-site Coulomb interaction U and exchange interaction J parameters were set to 3.0 and 0.9 eV [45]. The core electrons were treated by the projector-augmented wave (PAW) method [46]. A vacuum region of approximately 20 Å was introduced along the z direction to accurately model the 2D monolayer system. The kinetic energy cutoff was set to 500 eV for the plane-wave basis functions. For the structural optimizations, the Brillouin zone was sampled using a 9 × 9 × 1 Monkhorst–Pack k-point mesh.
In the previous study, high pressure was applied on the monolayer MoS2 to regulate the optoelectronic and band structural properties via a diamond anvil cell and DFT calculations [47]. The fundamental configuration for the DFT calculations involves relaxing all structures within the x-y plane, while the z coordinates are constrained using a conjugate gradient method [47]. As we know, in monolayer CrI3, the Cr atomic plane is sandwiched by two I atomic planes, which is similar to that of monolayer MoS2. Thus, in this paper, we used a similar setup to perform the DFT calculation for investigating the effect of pressure on the physical properties of monolayer CrI3.
To ensure computational efficiency, the convergence criteria for energy and force in the self-consistent field (SCF) calculations were set to 10−8 eV and 0.001 eV/Å, respectively. Meanwhile, spin–orbit coupling (SOC) interaction was included in the noncollinear calculation when the values of magnetic anisotropy energy (MAE) were calculated. The calculation of structural optimization and the static and electronic state densities were carried out using the GGA + U method. Given that the GGA(PBE) method typically underestimates the energy gap of semiconductor materials, in order to make the calculation more accurate, we employed the Heyd–Scuseria–Ernzerhof (HSE06) [48] screened hybrid functional to calculate their electronic properties and band information. The mixing parameters were AEXX and ALDAX. ALDAX was constrained to be equal to 0.75.
Using a 32 × 32 × 1 2D honeycomb lattice with periodic boundary conditions, the TC of the samples based on monolayer CrI3 were calculated using the MC simulations. For each temperature examined, the MC simulations entailed 104 MC steps per site to attain thermal equilibrium. All MC simulations were conducted using the open-source project MCSOLVER [49].

3. Results and Discussion

As shown in Figure 1a,b, the 2D monolayer CrI3 possesses a rhombohedral BiI3 structure (space group R 3 ¯ ) at a low temperature. Each Cr3+ ion is surrounded by six I ions, which build an edge-sharing octahedra. Under this environment, the spin direction of the three 3d electrons which occupy the lower energy Cr-t2g orbits is consistent, resulting in   S = 3 2 . Our spin-polarized calculations show that the magnetic moment on each Cr atom is 3.1 μB, which agrees well with the experimental value of ∼3.1 μB [50]. Furthermore, high pressure was applied on the monolayer CrI3 along the z direction. Here, the pressure coefficient ɛ is defined as follows:
ɛ = z z 0 z 0
where z0 and z are the coordinates of the I atom along the z direction for the system without and with pressure, respectively. During the calculation process, in order to reflect the influence of pressure, we fixed the z-axis coordinate of the I atom and the z-direction of the lattice constant after setting the z-direction coordinate of the I atom when performing the structure optimization. In accordance with the renowned Mermin–Wagner theorem [51], the MAE serves as a critical parameter in 2D materials, enabling stabilization of the long-range magnetic ordering against thermal fluctuations. Therefore, the values of the MAE of monolayer CrI3 under different pressures were calculated. Here, the MAE is obtained by calculating the total energy difference (MAE = E//E) of two different magnetization directions (in-plane (E//) and the perpendicular (E)). Due to its varying directions in between the x or y direction, while calculating the easy magnetization axis, the spin directions for the in-plane specific configuration were x and y. The results are listed in Table 1, in which the positive and negative values of MAE show the easy magnetization axis (EMA) along the out-of-plane and in-plane directions, respectively. The MAE of monolayer CrI3 under zero pressure is 778 μeV/Cr, which is in close agreement with previous reports [52,53]. With the pressure coefficient ɛ increasing to 1.2%, the value of the MAE of monolayer CrI3 decreases from 778 to −9.2 μeV/Cr. Simultaneously, the EMA of monolayer CrI3 transforms from the out-of-plane to the in-plane direction. As the pressure coefficient ɛ increases from 1.2% to 2.0%, the MAE notably changes from −9.2 to −592 μeV/Cr, and the direction of EMA is still in-plane. These results suggest that the MAE and EMA of monolayer CrI3 can be effectively optimized through the application of high pressure.
With the nonzero MAE, the exchange interaction, which is another fundamental parameter for determining the TC of 2D magnetic systems, was then studied. The magnetic exchange parameter Ji is obtained by the DFT calculations, and the Heisenberg spin Hamiltonian is given by the following [54]:
H = E 0 + i j J 1 S i S j + i K J 2 S i S K + A S i z S i z
where J 1 and J 2 denote the parameters for the nearest and next-nearest magnetic exchange interaction, respectively; S i is the spin vector of each magnetic atom; A   is the anisotropy energy parameter; and S i z is the z component of spin vector. During the calculation of J 1 and J 2 , three spin arrangements (FM, Stripy-AFM, and Néel-AFM) were considered. In the FM configuration, all magnetic moments were initialized to align in a uniform direction, which is shown in Figure 1c. The arrangements of the magnetic moments for the Stripy-AFM and Néel-AFM configurations are illustrated in Figure 1d,e, respectively. The magnetic energies of the three spin configurations of monolayer CrI3 are expressed as follows:
  E F M = E 0 + 3 J 1 + 6 J 2 S 2 + A | S | 2
E S t r i p y = E 0 + J 1 2 J 2 S 2 + A | S | 2
  E N é e l = E 0 + 3 J 1 + 6 J 2 S 2 + A | S | 2
Correspondingly, the expressions of the magnetic exchange interaction parameters J 1 and J 2 are
  J 1 = E F M E N e ´ e l 6 | S | 2  
J 2 = 2 E F M 3 E S t r i p y + E N é e l 24 S 2
Here, S = 3 2 .
After comparing the calculated results shown in Table 1, we find that applying high pressure on monolayer CrI3 can lead to remarkable changes in the J 1 and J 2 . The values of J 1 and J 2 of monolayer CrI3 under zero pressure are −1.84 and −0.31 meV, respectively. As the pressure coefficient ɛ increases, the absolute value of J 1 changes significantly from 1.89 to 3.23 meV, and that of J 2 decreases slightly from 0.31 to 0.12 meV, as shown in Table 1.
In order to quantitatively estimate the TC value of monolayer CrI3 under different pressures, MC simulation was conducted based on the Heisenberg model with the values of J 1 and J 2 . The temperature-dependent variations in heat capacity (CV) and average magnetic moment of the monolayer CrI3 are shown in Figure 2. As we know, the temperature at which the mean magnetic moment exhibits a sharp decline to nearly zero and the CV peaks can be identified as the TC. For monolayer CrI3, the TC is estimated to be 53.5 K, as illustrated in Figure 2a, which is consistent with previously reported results [8,52,53]. With the pressure increasing, the TC value of monolayer CrI3 enhances dramatically, and the maximum value is 97.3 K when the pressure coefficient ɛ is 2.0%. The TC values under different pressures are listed in Table 1.
Generally speaking, direct exchange and superexchange interactions are two major kinds of exchange interactions, which are often discussed in research on 2D semiconducting magnetic materials. In the monolayer CrI3, the FM coupling stems from the superexchange interaction, which can be investigated using the projected density of states (PDOS). As shown in Figure 3, the PDOS peaks of spin-up and spin-down in the valence band and those of spin-down in the conduction band gradually move toward the Fermi energy level as the pressure coefficient ɛ increases. Meanwhile, two more intense resonant states, located at the range of −2 to 2 eV, imply that the hybridization degree between the d-orbital of Cr and p-orbital of I becomes stronger with the increase in pressure. It is well established that the strength of a superexchange interaction is significantly influenced by the degree of overlap between wave functions. Figure 4 illustrates the distribution of wave functions in monolayer CrI3 under varying pressures. Under the same display standard, when no pressure is applied, the degree of overlap between the spin-up and spin-down electronic state densities of Cr-d and I-p atoms is relatively small. Compared to the monolayer CrI3 under zero pressure, the application of high pressure markedly enhances the overlap between Cr-d and I-p orbitals under pressure coefficients ɛ of 1.0% and 2.0%. This indicates that the superexchange interaction under ferromagnetic coupling is significantly enhanced under the action of pressure, which accounts for the observed increase in TC.
As mentioned above, the mechanism of FM coupling in monolayer CrI3 is a superexchange interaction, whose strength is decisively affected by the bond length l of Cr-I and the bond angle θ of Cr-I-Cr [38]. Moreover, the shorter l is and the nearer θ is to 180°, the stronger the exchange interaction is, and the higher the TC is [39,40]. Thus, it is essential to investigate the impact of pressure on the Cr-I bond length l and Cr-I-Cr bond angle θ in monolayer CrI3. Following the structural optimization, the Cr-I bond length l and Cr-I-Cr bond angle θ with zero pressure were determined to be 2.75 Å and 95.2°, which is in agreement with the previously reported values of 2.74 Å and 95.2°. When the pressure coefficients ɛ are 0.5%, 1.0%, 1.2%, 1.5%, and 2.0%, the corresponding values of l and θ are listed in Table 1. Compared with these calculated results, the Cr-I bond length l becomes shorter and the Cr-I-Cr bond angle θ becomes larger with the pressure coefficient ɛ increasing, which can make the eg levels move closer to the occupied t2g levels and enhance ferromagnetic exchange. Thus, the TC is enhanced.
To further investigate the effect of high pressure on the electronic properties of 2D monolayer materials, the band structures of monolayer CrI3 for the FM phase under different pressures were calculated based on the HSE06 functional. As shown in Figure 5a, the monolayer CrI3 without applying pressure is an FM insulator with an indirect band gap of 2.03 eV, which is consistent with previously reported results [45]. With the pressure coefficient ɛ increasing to 1.0%, the valence bands of the spin-up are hardly shifted, but the conduction bands gradually move away from the Fermi level, resulting in an increase in the band gap of monolayer CrI3. When the pressure coefficient ɛ increases to 2.0%, instead of an indirect band gap, a direct electronic band gap of 3.23 eV is observed, which is important for optoelectronic and magneto-optical applications.

4. Conclusions

As shown in the above calculation results, applying high pressure can significantly enhance the TC of monolayer CrI3. Moreover, this is also possible to achieve experimentally. According to earlier reports [47], high pressure has been applied to monolayer MoS2 to regulate the band structure and optoelectronic properties via a diamond anvil cell. Based on these experimental findings, we propose that high pressure can be applied on monolayer CrI3 via a diamond anvil cell.
In conclusion, the effect of high pressures on the magnetic and electronic properties of monolayer CrI3 were investigated by using the first-principle calculations and MC simulations. The calculated results indicate that applying high pressure to two-dimensional monolayer CrI3 along the z direction can transform the direction of EMA from the out-of-plane to the in-plane and remarkably enhance the TC. In particular, when ɛ is increased to 2.0%, the TC of monolayer CrI3 increases from 53.5 to 97.3 K, which is nearly double that of pristine monolayer CrI3. The changes in the Cr-I bond length and Cr-I-Cr bond angle indicate that the wave function overlap of Cr-d and I-p orbitals increases, which is the reason for the TC enhancement of monolayer CrI3. Moreover, as the applied pressure coefficient ɛ reaches 2.0%, an indirect-to-direct change of band gap can be observed in monolayer CrI3. These findings provide practical methods and strategies for enhancing ferromagnetism in 2D monolayer systems, which is useful for the further potential application of 2D materials in spintronic devices.

Author Contributions

Conceptualization, D.W. (Dunhui Wang) and D.W. (Dong Wei); Methodology, W.S.; Software, Z.D.; Formal analysis, W.S.; Investigation, W.S. and D.W. (Dong Wei); Resources, D.W. (Dunhui Wang); Data curation, W.S.; Writing—original draft, W.S.; Writing—review & editing, D.W. (Dunhui Wang); Project administration, D.W. (Dunhui Wang). All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Shandong Province (Grant Nos. ZR2023QA145 and ZR2023ZD09). We are grateful to the High-Performance Computing Center of Nanjing University for carrying out the numerical calculations in this paper on its blade cluster system.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, L.S.; Zhou, J.; Li, H.; Shen, L.; Feng, Y.P. Recent progress and challenges in magnetic tunnel junctions with 2D materials for spintronic applications. Appl. Phys. Rev. 2021, 8, 021308. [Google Scholar] [CrossRef]
  2. Ahn, Y.J.; Guo, X.Y.; Son, S.H.; Sun, Z.L.; Zhao, L.Y. Progress and prospects in two-dimensional magnetism of van der Waals materials. Prog. Quantum. Electron. 2024, 93, 100498. [Google Scholar] [CrossRef]
  3. Qu, Y.Q.; Liao, Y.; He, J.J.; Chen, Y.; Yao, G. High-Temperature Intrinsic Two-Dimensional-XY Ferromagnetism and Strong Magnetoelastic Coupling in Tetragonal Monolayer MnGe. J. Phys. Chem. C 2024, 128, 4631–4638. [Google Scholar] [CrossRef]
  4. Papavasileiou, A.V.; Menelaou, M.; Sarkar, K.J.; Sofer, Z.; Polavarapu, L.; Mourdikoudis, S. Ferromagnetic Elements in Two-Dimensional Materials: 2D Magnets and Beyond. Adv. Funct. Mater. 2024, 34, 2309046. [Google Scholar] [CrossRef]
  5. Chen, X.; Zhang, X.; Xiang, G. Recent advances in two-dimensional intrinsic ferromagnetic materials Fe3X (X = Ge and Ga) Te2 and their heterostructures for spintronics. Nanoscale 2024, 16, 527–554. [Google Scholar] [CrossRef]
  6. Gao, H.G.; Qian, Y.; Ye, S.; Kong, K.P. First-principles study on the electric control of ferromagnetic behaviour of two-dimensional BaTiO3 (0 0 1) ultrathin film doped with Cr. Appl. Surf. Sci. 2022, 601, 154240. [Google Scholar] [CrossRef]
  7. Rodriguez-Vega, M.; Lin, Z.X.; Leonardo, A.; Ernst, A.; Chaudhary, G.; Vergniory, M.G.; Fiete, G.A. Phonon-mediated dimensional crossover in bilayer CrI3. Phys. Rev. B 2020, 102, 081117. [Google Scholar] [CrossRef]
  8. 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]
  9. Zhao, Y.; Yang, R.; Yang, K.; Dou, J.R.; Gou, J.Z.; Zhou, G.W.; Xu, X.H. Anomalous magnetic property and broadband photodetection in ultrathin non-layered manganese selenide semiconductor. Nano Res. 2024, 17, 8578–8584. [Google Scholar] [CrossRef]
  10. Chen, J.M.; Cheng, Z.X.; Chen, J.H.; Li, M.L.; Jia, X.H.; Ran, Y.Q.; Zhang, Y.; Li, Y.P.; Yu, T.J.; Dai, L. Spin-Enhanced Self-Powered Light Helicity Detecting Based on Vertical WSe2-CrI3 p-n Heterojunction. ACS Nano 2024, 18, 26261–26270. [Google Scholar] [CrossRef]
  11. Siudzinska, A.; Gorantla, S.M.; Serafinczuk, J.; Kudrawiec, R.; Hommel, D.; Bachmatiuk, A. Electron Beam-Induced Reduction of Cuprite. Metals 2022, 12, 2151. [Google Scholar] [CrossRef]
  12. Gong, C.; Li, L.; Li, Z.L.; Ji, H.W.; Stern, A.; Xia, Y.; Cao, T.; Bao, W.; Wang, C.Z.; Wang, Y.; et al. Discovery of intrinsic ferromagnetism in two-dimensional van der Waals crystals. Nature 2017, 546, 265–269. [Google Scholar] [CrossRef] [PubMed]
  13. 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] [PubMed]
  14. Chen, W.J.; Sun, Z.Y.; Wang, Z.J.; Gu, L.H.; Xu, X.D.; Wu, S.W.; Gao, C.L. Direct observation of van der Waals stacking-dependent interlayer magnetism. Science 2019, 366, 983–987. [Google Scholar] [CrossRef]
  15. Hu, J.K.; Fan, Z.Q.; Yang, J.B. Modulating the electronic and optical properties of CrI3/In2Se3 van der Waals heterostructures by external fields. Mater. Today Commun. 2024, 39, 108891. [Google Scholar] [CrossRef]
  16. 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]
  17. Chen, L.; Jiang, C.; Yang, M.Y.; Hu, T.; Meng, Y.; Lei, J.; Zhang, M.J. Magnetism and electronic structures of bismuth (stannum) films at the CrI3 (CrBr3) interface. Phys. Chem. Chem. Phys. 2021, 23, 4255–4261. [Google Scholar] [CrossRef]
  18. Liu, L.; Yang, K.; Wang, G.Y.; Wu, H. Two-dimensional ferromagnetic semiconductor VBr3 with tunable anisotropy. J. Mater. Chem. C 2020, 8, 14782–14788. [Google Scholar] [CrossRef]
  19. Tian, S.J.; Zhang, J.F.; Li, C.H.; Ying, T.P.; Li, S.Y.; Zhang, X.; Liu, K.; Lei, H.C. Ferromagnetic van der Waals Crystal VI3. J. Am. Chem. Soc. 2019, 141, 5326–5333. [Google Scholar] [CrossRef]
  20. Sun, J.X.; Zhong, X.; Cui, W.W.; Shi, J.M.; Hao, J.; Xu, M.L.; Li, Y.W. The intrinsic magnetism, quantum anomalous Hall effect and Curie temperature in 2D transition metal trihalides. Phys. Chem. Chem. Phys. 2020, 22, 2429–2436. [Google Scholar] [CrossRef]
  21. Kim, K.; Lee, J.U.; Cheong, H. Raman spectroscopy of two-dimensional magnetic van der Waals materials. Nanotechnology 2019, 30, 452001. [Google Scholar] [CrossRef] [PubMed]
  22. Yu, W.; Li, J.; Herng, T.S.; Wang, Z.S.; Zhao, X.X.; Chi, X.; Fu, W.; Abdelwahab, I.; Zhou, J.; Dan, J.D.; et al. Chemically Exfoliated VSe2 Monolayers with Room-Temperature Ferromagnetism. Adv. Mater. 2019, 31, 1903779. [Google Scholar] [CrossRef]
  23. Torelli, D.; Thygesen, K.S.; Olsen, T. High throughput computational screening for 2D ferromagnetic materials: The critical role of anisotropy and local correlations. 2D Mater. 2019, 6, 045018. [Google Scholar] [CrossRef]
  24. McGuire, M.A. Crystal and Magnetic Structures in Layered, Transition Metal Dihalides and Trihalides. Crystals 2017, 7, 121. [Google Scholar] [CrossRef]
  25. Lu, S.H.; Zhou, Q.H.; Guo, Y.L.; Zhang, Y.H.; Wu, Y.L.; Wang, J.L. Coupling a Crystal Graph Multilayer Descriptor to Active Learning for Rapid Discovery of 2D Ferromagnetic Semiconductors/Half-Metals/Metals. Adv. Mater. 2020, 32, 2002658. [Google Scholar] [CrossRef]
  26. Mounet, N.; Gibertini, M.; Schwaller, P.; Campi, D.; Merkys, A.; Marrazzo, A.; Sohier, T.; Castelli, I.E.; Cepellotti, A.; Pizzi, G.; et al. Two-dimensional materials from high-throughput computational exfoliation of experimentally known compounds. Nat. Nanotech. 2018, 13, 246–252. [Google Scholar] [CrossRef]
  27. Liu, H.; Sun, J.T.; Liu, M.; Meng, S. Screening Magnetic Two-Dimensional Atomic Crystals with Nontrivial Electronic Topology. J. Phys. Chem. Lett. 2018, 9, 6709–6715. [Google Scholar] [CrossRef]
  28. Zhu, Y.; Kong, X.H.; Rhone, T.D.; Guo, H. Systematic search for two-dimensional ferromagnetic materials. Phys. Rev. Mater. 2018, 2, 081001. [Google Scholar] [CrossRef]
  29. Webster, L.; Liang, L.; Yan, J.A. Distinct spin-lattice and spin-phonon interactions in monolayer magnetic CrI3. Phys. Chem. Chem. Phys. 2018, 20, 23546–23555. [Google Scholar] [CrossRef]
  30. Jiang, P.H.; Li, L.; Liao, Z.L.; Zhao, Y.X.; Zhong, Z.C. Spin Direction-Controlled Electronic Band Structure in Two-Dimensional Ferromagnetic CrI3. Nano Lett. 2018, 18, 3844–3849. [Google Scholar] [CrossRef]
  31. Ye, H.S.; Wang, X.; Bai, D.M.; Zhang, J.T.; Wu, X.S.; Zhang, G.P.; Wang, J.L. Significant enhancement of magnetic anisotropy and conductivity in GaN/CrI3 van der Waals heterostructures via electrostatic doping. Phys. Rev. B 2021, 104, 075433. [Google Scholar] [CrossRef]
  32. Han, J.N.; Ding, Z.X.; Li, Z.H.; Cao, S.G.; Zhang, Z.H.; Deng, X.Q. CrI3/Arsenene vdW heterstructure: Outstanding physical properties and substantially enhanced magnetic stability. Appl. Surf. Sci. 2024, 669, 160443. [Google Scholar] [CrossRef]
  33. Zheng, Z.J.; Ren, K.; Huang, Z.M.; Zhu, Z.Y.; Wang, K.; Shen, Z.L.; Yu, J. Remarkably improved Curie temperature for two-dimensional CrI3 by gas molecular adsorption: A DFT study. Semicond. Sci. Technol. 2021, 36, 075015. [Google Scholar] [CrossRef]
  34. Tang, C.; Zhang, L.; Du, A.J. Tunable magnetic anisotropy in 2D magnets via molecular adsorption. J. Mater. Chem. C 2020, 8, 14948–14953. [Google Scholar] [CrossRef]
  35. Mo, Y.Y.; Huang, X.K.; Xu, J.L.; Jiang, X.A.; Chen, C.; Jiang, X.P.; Liu, J.M. Interfacial engineering of orbital orientation for perpendicular magnetic anisotropy in Co-implanted CrI3 monolayer. J. Appl. Phys. 2024, 136, 084305. [Google Scholar] [CrossRef]
  36. Yang, Q.; Hu, X.H.; Shen, X.D.; Krasheninnikow, A.V.; Chen, Z.F.; Sun, L.T. Enhancing Ferromagnetism and Tuning Electronic Properties of CrI3 Monolayers by Adsorption of Transition-Metal Atoms. ACS Appl. Mater. Interfaces 2021, 13, 21593–21601. [Google Scholar] [CrossRef]
  37. Chen, G.X.; Li, B.B.; Li, X.F.; Wang, D.D.; Liu, S.; Zhang, J.M. Electronic structure and magnetic properties of CrI3 monolayer doped by rare earth metal atoms. J. Phys. Chem. Solids 2024, 187, 111838. [Google Scholar] [CrossRef]
  38. Zhao, Y.H.; Lin, L.F.; Zhou, Q.H.; Li, Y.H.; Yuan, S.J.; Chen, Q.; Dong, S.; Wang, J.L. Surface Vacancy-Induced Switchable Electric Polarization and Enhanced Ferromagnetism in Monolayer Metal Trihalides. Nano Lett. 2018, 18, 2943–2949. [Google Scholar] [CrossRef]
  39. Bozorth, R.M.; Ferromagnetism, D. Ferromagnetism (Chapter XII); Van Nostrand Company. Inc.: New York, NY, USA, 1951; p. 867. [Google Scholar]
  40. Anderson, P.W. Antiferromagnetism. Theory of Superexchange Interaction. Phys. Rev. 1950, 79, 350. [Google Scholar] [CrossRef]
  41. Su, W.X.; Zhang, Z.M.; Cao, Q.Q.; Wang, D.H.; Lu, H.M.; Mi, W.B.; Du, Y.W. Enhancing the Curie temperature of two-dimensional monolayer CrI3 by introducing I-vacancies and interstitial H-atoms. Phys. Chem. Chem. Phys. 2021, 23, 22103–22109. [Google Scholar] [CrossRef]
  42. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef] [PubMed]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  44. Liechtenstein, A.I.; Anisimov, V.I.; Zaanen, J. Density-functional theory and strong interactions: Orbital ordering in Mott-Hubbard insulators. Phys. Rev. B 1995, 52, 5467. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, J.Y.; Zhao, B.; Zhou, T.; Xue, Y.; Ma, C.L.; Yang, Z.Q. Strong magnetization and Chern insulators in compressed graphene/CrI3 van der Waals heterostructures. Phys. Rev. B 2018, 97, 085401. [Google Scholar] [CrossRef]
  46. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  47. Nayak, A.P.; Pandey, T.; Voiry, D.; Liu, J.; Moran, S.T.; Sharma, A.; Tan, C.; Chen, C.H.; Li, L.J.; Chhowalla, M.; et al. Pressure-Dependent Optical and Vibrational Properties of Monolayer Molybdenum Disulfide. Nano Lett. 2015, 15, 346–353. [Google Scholar] [CrossRef]
  48. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
  49. Liang, L.; Zhang, X. Available online: https://github.com/golddoushi/mcsolver (accessed on 5 June 2024).
  50. Dillon, J.F.; Olson, C.E. Magnetization, Resonance, and Optical Properties of the Ferromagnet CrI3. J. Appl. Phys. 1965, 36, 1259–1260. [Google Scholar] [CrossRef]
  51. Mermin, N.D.; Wagner, H. Absence of Ferromagnetism or Antiferromagnetism in One- or Two-Dimensional Isotropic Heisenberg Models. Phys. Rev. Lett. 1966, 17, 1133. [Google Scholar] [CrossRef]
  52. Webster, L.; Yan, J.A. Strain-tunable magnetic anisotropy in monolayer CrCl3, CrBr3, and CrI3. Phys. Rev. B 2018, 98, 144411. [Google Scholar] [CrossRef]
  53. Zhang, W.B.; Qu, Q.; Zhu, P.; Lam, C.H. Robust intrinsic ferromagnetism and half semiconductivity in stable two-dimensional single-layer chromium trihalides. J. Mater. Chem. C 2015, 3, 12457–12468. [Google Scholar] [CrossRef]
  54. Liu, L.; Lin, Z.Z.; Hu, J.F.; Zhang, X. Full quantum search for high TC two-dimensional van der Waals ferromagnetic semiconductors. Nanoscale 2021, 13, 8137–8145. [Google Scholar] [CrossRef]
Figure 1. (a) Top view and (b) side view of atomic structure of monolayer CrI3. The magnetic configurations of (c) FM, (d) Stripy-AFM, and (e) Néel-AFM. The l and θ represent the bond length of Cr-I and the bond angle of Cr-I-Cr, respectively.
Figure 1. (a) Top view and (b) side view of atomic structure of monolayer CrI3. The magnetic configurations of (c) FM, (d) Stripy-AFM, and (e) Néel-AFM. The l and θ represent the bond length of Cr-I and the bond angle of Cr-I-Cr, respectively.
Metals 15 00398 g001
Figure 2. Evolution of the MC-averaged magnetic moment (red) and heat capacity (blue) per Cr of monolayer CrI3 under pressure coefficients ɛ of (a) 0%, (b) 0.5%, (c) 1.0%, and (d) 2.0%.
Figure 2. Evolution of the MC-averaged magnetic moment (red) and heat capacity (blue) per Cr of monolayer CrI3 under pressure coefficients ɛ of (a) 0%, (b) 0.5%, (c) 1.0%, and (d) 2.0%.
Metals 15 00398 g002
Figure 3. The projected density of states of monolayer CrI3 under pressure coefficients ɛ of (a) 0%, (b) 1.0%, and (c) 2.0%.
Figure 3. The projected density of states of monolayer CrI3 under pressure coefficients ɛ of (a) 0%, (b) 1.0%, and (c) 2.0%.
Metals 15 00398 g003
Figure 4. The distribution of spin wave functions for monolayer CrI3 under pressure coefficients ɛ of (a) 0%, (b) 1.0%, and (c) 2.0%. The yellow and blue colors represent the spin-up and spin-down, respectively.
Figure 4. The distribution of spin wave functions for monolayer CrI3 under pressure coefficients ɛ of (a) 0%, (b) 1.0%, and (c) 2.0%. The yellow and blue colors represent the spin-up and spin-down, respectively.
Metals 15 00398 g004
Figure 5. The band structures of monolayer CrI3 for the FM phase under different pressures (a) 0%, (b) 1.0% and (c) 2.0% based on the HSE06 functional. The spin-up and spin-down band structures are in-dicated using red and blue arrows respectively. The energy band gaps between the conduction band’s minimum and valence band’s maximum are indicated using pink arrows in each case.
Figure 5. The band structures of monolayer CrI3 for the FM phase under different pressures (a) 0%, (b) 1.0% and (c) 2.0% based on the HSE06 functional. The spin-up and spin-down band structures are in-dicated using red and blue arrows respectively. The energy band gaps between the conduction band’s minimum and valence band’s maximum are indicated using pink arrows in each case.
Metals 15 00398 g005
Table 1. J 1 and J2, as well as the Curie temperature TC, for monolayer CrI3 under different pressures.
Table 1. J 1 and J2, as well as the Curie temperature TC, for monolayer CrI3 under different pressures.
ɛl (Å)θ (deg)J1 (meV)J2 (meV)TC (K)EMAMAE (μeV/Cr)
0%2.7595.2−1.89−0.3153.5z778
0.5%2.7396.0−2.16−0.2768.8z437
1.0%2.7097.0−2.47−0.2382.8z124
1.2%2.6997.3−2.60−0.1983.1x−9.2
1.5%2.6897.8−2.85−0.1690.6x−234
2.0%2.6698.7−3.23−0.1297.3x−592
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

Su, W.; Wang, D.; Wei, D.; Dai, Z. Two-Fold Enhancement of Curie Temperature in Monolayer CrI3 by High Pressure. Metals 2025, 15, 398. https://doi.org/10.3390/met15040398

AMA Style

Su W, Wang D, Wei D, Dai Z. Two-Fold Enhancement of Curie Temperature in Monolayer CrI3 by High Pressure. Metals. 2025; 15(4):398. https://doi.org/10.3390/met15040398

Chicago/Turabian Style

Su, Wenxia, Dunhui Wang, Dong Wei, and Zhenhong Dai. 2025. "Two-Fold Enhancement of Curie Temperature in Monolayer CrI3 by High Pressure" Metals 15, no. 4: 398. https://doi.org/10.3390/met15040398

APA Style

Su, W., Wang, D., Wei, D., & Dai, Z. (2025). Two-Fold Enhancement of Curie Temperature in Monolayer CrI3 by High Pressure. Metals, 15(4), 398. https://doi.org/10.3390/met15040398

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

Article Metrics

Back to TopTop