This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Magnetic domain wall dynamics under external electric field in bilayer CrI3

Sahar Izadi Vishkayi School of Physics, Institute for Research in Fundamental Sciences (IPM), Tehran 19395-5531, Iran    Reza Asgari [email protected] School of Physics, Institute for Research in Fundamental Sciences (IPM), Tehran 19395-5531, Iran School of Physics, University of New South Wales, Kensington, NSW 2052, Australia
Abstract

Motivated by manipulating the magnetic order of bilayer CrI3, we carry out microscopic calculations to find the magnetic order and various magnetic domains of the system in the presence of an electric field. Making use of density functional simulations, a spin model Hamiltonian is introduced consisting of isotropic exchange couplings, Dzyaloshinskii-Moriya (DM) interaction, and on-site magnetic anisotropy. The spin dynamics of two well-known states of bilayer CrI3, low temperature (LT) and high temperature (HT) phases, are obtained by solving the Landau-Lifshitz-Gilbert equation. We show that the magnetic texture is stacking-dependent in bilayer CrI3 and stable magnetic domains can appear in the HT stack which are tunable by external electric and magnetic fields. Therefore, we suggest that the HT phase represents a promising candidate for data storage in the modern generation of spintronic devices working on magnetic domain engineering.

I Introduction

Two-dimensional (2D) magnetic materials have been investigated by experimental Huang et al. (2017); Deng et al. (2018); Gong et al. (2017); Gibertini et al. (2019); Gong and Zhang (2019); Meseguer-Sánchez et al. (2021) and theoretical Lado and Fernández-Rossier (2017); Morell et al. (2019); Torelli and Olsen (2020); Frey et al. (2019); Menichetti et al. (2019) researchers to analyze their attractive properties from the initial report of the successful synthesis of thin layer CrI3 Huang et al. (2017). The presence of a substantial magnetocrystalline anisotropy was discovered, thus breaking the invariance under spin rotations, and the crystal magnetic ordering was observed to be a layer-dependent phenomenon. Because of the existence of metastable magnetic domains with ultra-thin domain walls (like permanent nanoparticle magnets Hubert and Schäfer (2008)) which possess spin dynamics Wahab et al. (2021); Chen et al. (2021); Akram et al. (2021); Xu et al. (2021), CrI3, in monolayer and multilayer forms, has recently been described as a viable contender for the modern generation of memory devices. The microscopic nature of these magnetic domains in the presence of external fields is still poorly understood, offering a significant challenge for theoretical materials science.

Two geometrically stable stacks of bilayer CrI3 structures are known as low-temperature (LT) and high-temperature (HT) phases (in more than 210210 K) Morell et al. (2019); Jiang et al. (2019); Kim et al. (2019). The LT phase is a ferromagnetic (FM) ground state in which the magnetic moments of the Cr atoms are aligned in both layers, whereas the magnetic moments of the Cr atoms in distinct layers of the HT phase are in an antiferromagnetic (AFM) configuration Morell et al. (2019); Lei et al. (2021); Song et al. (2019); Li et al. (2019); Sivadas et al. (2018). The competition between orbitally dependent interlayer AFM super-super-exchange and interlayer FM super-super-exchange causes the magnetic ordering. In the HT bilayer CrI3, a vertical external electric field generates an AFM to an FM magnetic transition Jiang et al. (2018a); Morell et al. (2019); Jiang et al. (2018b). Because of their sensitivity to the electric field, the HT phase of the bilayer CrI3 can be more efficient in ultrafast memory devices, and it could be the first device for the engineering of magnetic domains in the world of 2D magnets. For a monolayer up to bulk structures, the Curie temperature of a thin layer CrI3 ranges from 4545 K to 6161 K Huang et al. (2017); Kashin et al. (2020). The magnetic ordering temperature of a thin layer CrI3 is raised by the number of layers, according to experiments Huang et al. (2017). The Néel temperature of 4545 K is reported for the HT bilayer Huang et al. (2018) and it is expected that the LT bilayer maintains a larger Curie temperature due to strong FM ordering between the layers Wang and Sanyal (2021).

Although the Ising Hamiltonian model is a classic model for explaining the magnetic behavior of 2D CrI3, it cannot explain the experimentally observed spin wave Chen et al. (2018). To explain the spin canting physics, an extended Heisenberg Hamiltonian containing Kitaev Lee et al. (2020), biquadratic anisotropy exchange term Wahab et al. (2021), or Dzyaloshinskii-Moria (DM) Vishkayi et al. (2020); Jaeschke-Ubiergo et al. (2021) interaction terms was proposed. The spin wave in a CrI3 structure Chen et al. (2020) is described by a Heisenberg Hamiltonian with Kitaev or DM interactions, and a small portion of the DM interactions in 2D CrI3 is responsible for a topological gap Jaeschke-Ubiergo et al. (2021).

The ability to alter and adjust magnetic domains of 2D bilayer magnets to make them useful in technology and spintronics is a key question here. We focus on the bilayer CrI3 in the presence of an external electric field Huang et al. (2018); Jiang et al. (2018a, b); Morell et al. (2019); Lei et al. (2021) to answer this specific topic.

In this study, we use a spin model Hamiltonian in the presence of external fields, with isotropic exchange and DM interactions. We use density functional theory (DFT) computations to derive a suitable spin model Hamiltonian for the bilayer CrI3 that accurately reproduces the magnetic characteristics of the system. The isotropic exchange coupling parameters of the bilayer CrI3 have been investigated in the presence of an external electric field Lei et al. (2021), but the Dzyaloshinskii-Moria (DM) interaction has yet to be investigated. In contrast to what we found for a monolayer CrI3 Vishkayi et al. (2020), our calculations suggest that the DM interaction in the HT bilayer CrI3 is substantially connected to the applied electric field. The spin dynamics of the system are calculated using the Landau-Lifshitz-Gilbert formalism Ellis et al. (2015) with DFT-derived exchange parameters as starting inputs. To obtain a temperature dependent magnetization, we also employ the Monte Carlo Metropolis technique with 10610^{6} time steps for a 50×5050\times 50 nm2 layer.

The weak strength of the isotropic intralayer exchange coupling signifies the magnetic ordering of the layers, which can be modified by an external electrical field, according to our findings. The findings also reveal that an electric field, like a magnetic field, alters the magnetic texture of the system. Furthermore, by applying an electric field resulting from an increase in the intralayer isotropic exchange coefficient, the temperature of the magnetic order of the HT bilayer is raised. The electric field also affects the magnetic texturing on the domain walls. The results reveal that the magnetic textures in the studied systems can be manipulated using an external electric field. To be more precise, in the LT bilayer, we show that the DM interactions lead to the creation of a quasi-circular domain shape in the layers, but its value is extremely smaller than the intralayer exchange coupling to stabilize it in the system, therefore, the obtained magnetic domain is unstable and disappearing quickly.

Refer to caption
Figure 1: (Color online) The top view ((a) and (c)) and side view ((b) and (d)) of the HT and LT bilayer CrI3 structures, respectively. The black lozenges show the unit cell of bilayers. The blue (pink) circles represent chromium (iodine) atoms. In (b) and (d), chromium atoms are labeled by 1-4 in the unit cell. In (b) the bonding angles between the atoms of each layer are shown by α\alpha, β\beta and θ\theta. The crystal parameters are given in Table I.

This paper is organized as follows. We commence with a description of our theoretical formalism in Sec. II, followed by the details of the DFT simulations, the spin model Hamiltonian and spin dynamics. Numerical results for the spin-spin interaction parameters in the presence of an electric field, the exchange coefficients in both the HT and LT phases and spin dynamics in two cases are reported in Sec. III. We summarize our main findings in Sec. IV.

II Theory and Models

II.1 DFT calculations

We use Quantum Espresso package Giannozzi et al. (2009) to simulate density-functional based calculations. The generalized gradient exchange-correlation within Perdew-Burke-Ernzerhof Perdew et al. (1996) functional is considered to expose the electronic and magnetic ground states of two bilayer CrI3 structures. An energy cut-off of 5050 Ry for wave vector and an 8×8×18\times 8\times 1 grid mesh of kk-points within the first Brillouin zone are defined as the converged input parameters. To avoid repeats effects, we consider a vacuum of 2020 Å  along the zz-direction. The van der Waals Grimme-D2 Grimme (2006) correction is used to consider the interaction between layers. DFT+U+U calculations Cococcioni and De Gironcoli (2005) are also needed to find correct magnetic ground states of CrI3 bilayers. At this point, we consider U=3U=3 eV as the on-site Hubbard parameter similar to Refs. Jiang et al. (2019); Lei et al. (2021); Leon et al. (2020); Akram et al. (2021). To obtain a reliable total ground-state energy, we maintain a high degree of accuracy of 101010^{-10} eV. The bilayers are relaxed until the maximum force on each atom is 0.010.01 eV/Å  even under the electric field conditions. It should be noted that the consideration of the spin-orbit coupling (SOC) and noncollinear spin-polarization are essential to obtain the DM interaction and magnetic anisotropic energy (MAE).

II.2 Spin-model Hamiltonian

A spin model Hamiltonian for a 2D magnetic hexagonal lattice in each layer is utilized to provide spin-spin interactions in a magnetic 2D bilayer CrI3 system:

=\displaystyle{\cal H}= i,j[12Jij𝐒i𝐒j+γi|Siz|2+\displaystyle\sum_{i,j}[\frac{1}{2}J_{ij}{\bf S}_{i}\cdot{\bf S}_{j}+{\gamma}_{i}|S_{iz}|^{2}+
12𝐃ij(𝐒i×𝐒j)+iμB𝐁ext𝐒i],\displaystyle\frac{1}{2}{\bf D}_{ij}\cdot({\bf S}_{i}\times{\bf S}_{j})+\sum_{i}\mu_{B}{{\bf B}_{ext}}\cdot{\bf{S}}_{i}], (1)

where JijJ_{ij} is the isotropic exchange coupling parameter between ii and jj atoms, γi\gamma_{i} denotes a magnetic anisotropy coefficient, 𝐃ij{\bf D}_{ij} is the DM interaction vector and the last term shows the effect of an external magnetic field, 𝐁ext{\bf B}_{ext}, and μB\mu_{B} is the magnetic Bohr. The magnetic moment of ith Cr atom is indicated by 𝐒i{\bf S}_{i} in Eq. (1). It is intriguing to compare the strength of the interlayer and intralayer exchange coefficient parameters which are respectively referred by vv and tt indices in this paper.

II.3 Spin Dynamics

The Landau-Lifshitz-Gilbert (LLG) equation Ellis et al. (2015) is used to atomistic model the spin dynamics of the system after the coefficients of the spin Hamiltonian are established. We follow the method introduced by Evans et al. Evans et al. (2014) implemented in VAMPIRE code. The time evolution of the spin direction is given by LLG equation according to

𝐒it=γ1+λ2[𝐒i×𝐇effi+λ𝐒i×(𝐒i×𝐇effi)],\frac{\partial{\bf S}_{i}}{\partial t}=-\frac{\gamma}{1+\lambda^{2}}[{\bf S}_{i}\times{\bf H}_{eff}^{i}+\lambda{\bf S}_{i}\times({\bf S}_{i}\times{\bf H}_{eff}^{i})], (2)

where γ\gamma is the gyromagnetic ration and λ\lambda is microscopic damping equal to unity for finding the equilibration states quickly. The effective field applied on each spin is defined by 𝐇effi=/𝐒i+𝐇thi{\bf H}_{eff}^{i}=-{\partial\cal H}/{\partial{\bf S}_{i}}+{\bf H}_{th}^{i} and the effective thermal field, 𝐇thi{\bf H}_{th}^{i}, is counted by Langevin dynamic approach Brown (1979) to consider thermal spin fluctuations.

In order to obtain the spin dynamics simulations, a 100 nm ×\times 100 nm square bilayer CrI3 was kept in thermal equilibrium above the critical (Néel or Curie) temperature and then linearly cooled to 0 K in 2 ns. The magnetic moments of atoms are randomized at a temperature above the critical temperature, so we chose random initial spin directions for the spin dynamics simulation. The time step of 1 fs is considered in the integration scheme of Heun’s method García-Palacios and Lázaro (1998) for solving LLG equations, suitable for materials with low critical temperature and large dimensions to obtain stable results Evans et al. (2014).

The long-range demagnetization field is also considered in the spin dynamics calculations. The magnetic dipole-dipole interaction induces local spin-flip in the magnetic materials and leading to the creation of complicated spin texture even in the materials with a FM ground state Kawaguchi et al. (2007); Ezawa (2010). We discretize the layer to macrocells with 11 nm3 volume to avoid expenses of the demagnetization field computing at the atomistic level. The magnetic moments of macrocells, 𝐦mc{\bf m}_{mc}, are defined by the sum of atomic magnetic moments within the macrocell and the demagnetization field of ith macrocell is given by

𝐇dmci=μ04π(ij𝐌ij𝐦mcj)μ03𝐦mciVmci,{\bf H}_{d}^{mci}={\frac{{\mu}_{0}}{4\pi}}(\sum_{i\neq j}{\bf M}_{ij}\cdot{\bf m}_{mc}^{j})-{\frac{\mu_{0}}{3}}{\frac{{\bf m}_{mc}^{i}}{V_{mc}^{i}}}, (3)

where VmciV_{mc}^{i} is the volume of ith macrocell, μ0=4π×107\mu_{0}=4\pi\times 10^{-7} is the vacuum permeability and 𝐌ij{\bf M}_{ij} is the dipole-dipole interaction matrix between ii and jj macrocells defined as Evans et al. (2014),

𝐌ij=[3rxrx1rij3133rxry3rxrz3rxry3ryry1rij3133ryrz3rxrz3ryrz3rzrz1rij313],{\bf M}_{ij}=\begin{bmatrix}\frac{3r_{x}r_{x}-1}{r_{ij}^{3}}-\frac{1}{3}&3r_{x}r_{y}&3r_{x}r_{z}\\ 3r_{x}r_{y}&\frac{3r_{y}r_{y}-1}{r_{ij}^{3}}-\frac{1}{3}&3r_{y}r_{z}\\ 3r_{x}r_{z}&3r_{y}r_{z}&\frac{3r_{z}r_{z}-1}{r_{ij}^{3}}-\frac{1}{3}\end{bmatrix}, (4)

where the positions of macrocells are obtained by the magnetic center of the mass relation and the vector between the position of ii and jj macrocells is shown by 𝐫ij=rij(rxi^+ryj^+rzk^){\bf r}_{ij}=r_{ij}(r_{x}\hat{i}+r_{y}\hat{j}+r_{z}\hat{k}). We consider that the sheet is linearly cooled in 22 ns from an equilibrium temperature, upper the Curie temperature, to zero Kelvin. To obtain temperature dependent magnetization, we use the Monte Carlo Metropolis algorithm with 10610^{6} time steps for a 5050×50\times 50 nm2 sheet.

III Numerical Results and Discussions

The magnetic ground state, symmetric and asymmetric exchange coefficients, and temperature-dependent magnetization of bilayer CrI3 in both the HT and LT phases in the ground state are presented in this section. The effects of an external electric field on the coefficients are then studied. Finally, we investigate the spin dynamics of both phases in the presence of a perpendicular electric field.

The HT phase of bilayer CrI3 is in the form of the monoclinic structure, while the LT bilayer possesses a rhombohedral crystal structure as displayed in Figs. 1(a)-(d). The different crystal structures leads to the unique magnetic behaviors. Both the HT and LT phases are relaxed by spin-dependent DFT calculations and their optimized lattice constant, bonding lengths and angles are reported in Table 1. The results reveal that the bonding angles and lengths of both layers are the same, and that the geometry of the bilayer is unaffected by the electric field, which is in good agreement with those described in Ref. Morell et al. (2019). To obtain a formation of the energy, we use the straight formula of EF=Et(bilayer)nCrE(Cr)nIE(I)E_{\rm F}=E_{t}(bilayer)-n_{Cr}E(Cr)-n_{I}E(I) where Et(bilayer)E_{t}(bilayer) is the total energy of the bilayer CrI3, nCr(nI)n_{Cr}(n_{I}) is the number of the chromium (iodine) atoms in the unit cell, and E(Cr)E(Cr) and E(I)E(I) are the energy of free Cr and I atoms, respectively. The formation energies of the HT and LT bilayers are given in Table 1 indicating that the both phases are energetically favored to occur.

Table 1: Lattice constant, aa, Cr-Cr (I-I) distance between the layers, hCrCr(hII)h_{Cr-Cr}(h_{I-I}), Cr-I (Cr-Cr) bonding length, dCrI(dCrCr)d_{Cr-I}(d_{Cr-Cr}), two different I-Cr-I bonding angles, βICrI\beta_{I-Cr-I} and αICrI\alpha_{I-Cr-I}, and Cr-I-Cr bonding angle, θCrICr\theta_{Cr-I-Cr}, and formation energy, EfE_{f} of the HT and LT bilayer CrI3 structures. The bonding angles are shown in Fig. 1 (b).
  Phase    a(Å)a(\AA) hCrCr(Å)h_{Cr-Cr}(\AA) hII(Å)h_{I-I}(\AA) dCrI(Å)d_{Cr-I}(\AA) dCrCr(Å)d_{Cr-Cr}(\AA) βICrI(o)\beta_{I-Cr-I}(^{o}) αICrI(o)\alpha_{I-Cr-I}(^{o}) θCrICr(o)\theta_{Cr-I-Cr}(^{o}) EFE_{F} (eV/per Cr)
HT 6.977 7.24 3.99 2.81 4.03 177.4 89.9 91.6 -8.81
LT 6.979 6.59 3.35 2.81 4.03 177.6 89.9 91.7 -8.86

III.1 Exchange coefficients in the HT phase

In the HT phase of bilayer CrI3, there are four Cr atoms in the unit cell; two in each layer. At zero external magnetic field, if the nearest neighbor interlayer and intralayer interactions are considered, the extended first two terms of the model Hamiltonian (Eq. (1)) in the unit cell can be written as

=\displaystyle{\cal H}= 12(ηtJt[(𝐒1𝐒2)+(𝐒2𝐒1)+(𝐒3𝐒4)+(𝐒4𝐒3)]+\displaystyle\frac{1}{2}(\eta_{t}J_{t}[({\bf S}_{1}\cdot{\bf S}_{2})+({\bf S}_{2}\cdot{\bf S}_{1})+({\bf S}_{3}\cdot{\bf S}_{4})+({\bf S}_{4}\cdot{\bf S}_{3})]+
η1vJ1v[(𝐒1𝐒3)+(𝐒3𝐒1)+(𝐒2𝐒4)+(𝐒4𝐒2)]+\displaystyle\eta_{1v}J_{1v}[({\bf S}_{1}\cdot{\bf S}_{3})+({\bf S}_{3}\cdot{\bf S}_{1})+({\bf S}_{2}\cdot{\bf S}_{4})+({\bf S}_{4}\cdot{\bf S}_{2})]+
η2vJ2v[(𝐒1𝐒4)+(𝐒4𝐒1)+(𝐒2𝐒3)+(𝐒3𝐒2)])+\displaystyle\eta_{2v}J_{2v}[({\bf S}_{1}\cdot{\bf S}_{4})+({\bf S}_{4}\cdot{\bf S}_{1})+({\bf S}_{2}\cdot{\bf S}_{3})+({\bf S}_{3}\cdot{\bf S}_{2})])+
ηγ|Sz|2\displaystyle\eta\gamma|S_{z}|^{2} (5)

where ηt=3\eta_{t}=3 is the number of the nearest neighbors in each layer and η1v=η2v=1\eta_{1v}=\eta_{2v}=1 are the number of type-11 (between atoms numbered by 11 (22) and 33 (44) in Fig. 1 (b)) and type-22 (between atoms numbered by 11 (22) and 44 (33) in Fig. 1 (b)) neighbors in the interlayer. η=4\eta=4 is the number of Cr atoms in the unit cell. The exchange coupling coefficients are obtained by mapping the model Hamiltonian onto the DFT-based calculations which are computed for various magnetic configurations. For calculation of the isotropic exchange coefficient, 44-distinct spin configurations are considered as; FM (𝐒1=Sk^,𝐒2=Sk^,𝐒3=Sk^,𝐒4=Sk^{\bf S}_{1}=S\hat{k},{\bf S}_{2}=S\hat{k},{\bf S}_{3}=S\hat{k},{\bf S}_{4}=S\hat{k}), AFM1 (𝐒1=Sk^,𝐒2=Sk^,𝐒3=Sk^,𝐒4=Sk^{\bf S}_{1}=S\hat{k},{\bf S}_{2}=S\hat{k},{\bf S}_{3}=-S\hat{k},{\bf S}_{4}=-S\hat{k}), AFM2 (𝐒1=Sk^,𝐒2=Sk^,𝐒3=Sk^,𝐒4=Sk^{\bf S}_{1}=S\hat{k},{\bf S}_{2}=-S\hat{k},{\bf S}_{3}=-S\hat{k},{\bf S}_{4}=S\hat{k}) and AFM3 (𝐒1=Sk^,𝐒2=Sk^,𝐒3=Sk^,𝐒4=Sk^{\bf S}_{1}=S\hat{k},{\bf S}_{2}=-S\hat{k},{\bf S}_{3}=S\hat{k},{\bf S}_{4}=-S\hat{k}) in which all the magnetic moments of Cr atoms are aligned in the zz-direction and S=3/2S=3/2. In Fig. 2, the considered magnetic orders are represented in bilayer CrI3. In these configurations, the DM terms of Eq. (1) are vanished owing to the parallel magnetic moments of Cr atoms, and their total energies are obtained by

EFM=2S2(ηtJt+η1vJ1v+η2vJ2v)+ηγS2,\displaystyle E_{FM}=2S^{2}(\eta_{t}J_{t}+\eta_{1v}J_{1v}+\eta_{2v}J_{2v})+\eta\gamma S^{2}, (6)
EAFM1=2S2(ηtJtη1vJ1vη2vJ2v)+ηγS2,\displaystyle E_{AFM1}=2S^{2}(\eta_{t}J_{t}-\eta_{1v}J_{1v}-\eta_{2v}J_{2v})+\eta\gamma S^{2},
EAFM2=2S2(ηtJtη1vJ1v+η2vJ2v)+ηγS2,\displaystyle E_{AFM2}=2S^{2}(-\eta_{t}J_{t}-\eta_{1v}J_{1v}+\eta_{2v}J_{2v})+\eta\gamma S^{2},
EFM3=2S2(ηtJt+η1vJ1vη2vJ2v)+ηγS2.\displaystyle E_{FM3}=2S^{2}(-\eta_{t}J_{t}+\eta_{1v}J_{1v}-\eta_{2v}J_{2v})+\eta\gamma S^{2}.

We carry out spin-dependent DFT calculations to find the total energies of the considered magnetic configurations and as a consequence the exchange coupling parameters of the HT bilayer are given by

Jt=(EFM+EAFM1)(EAFM2+EAFM3)8S2ηt,\displaystyle J_{t}=\frac{(E_{FM}+E_{AFM1})-(E_{AFM2}+E_{AFM3})}{8S^{2}\eta_{t}}, (7)
Jv=EFMEAFM14S2ηv,\displaystyle J_{v}=\frac{E_{FM}-E_{AFM1}}{4S^{2}\eta_{v}}, (8)

where Jv=J1v+J2vJ_{v}=J_{1v}+J_{2v} is the interlayer exchange coupling and ηv=η1v=η2v\eta_{v}=\eta_{1v}=\eta_{2v}. The intralayer exchange coupling is Jt=7.00J_{t}=-7.00 meV which presents the strong ferromagnetic coupling of Cr atoms in the layers. Jv=+0.08J_{v}=+0.08 meV indicates the layers are coupled in a form of antiferromagnetic. Furthermore, the weak strength of the isotropic interlayer exchange coupling denotes the magnetic order of the layers can be tuned by the external electrical field Morell et al. (2019).

Refer to caption
Figure 2: (Color online) The magnetic ordering of Cr atoms in (a) FM, (b) AFM1, (c) AFM2 and (d) AFM3 spin configurations of HT bilayer CrI3. The dark (light) balls represent the Cr atoms in the top (bottom) layer and arrows show the magnetic moment vectors of Cr atoms in the z-direction. The dashed lines connect the atoms with intralayer and interlayer exchange coupling which are indicated, respectively, by t and v indices.

We observe the variation of the isotropic interlayer and intralayer exchange coupling by DFT-based calculations, shown in Fig. 3. It is observed that the sign of the interlayer exchange coupling is changed between 22-2.52.5 V/nm, hence a magnetic transition from the AFM to the FM occurs between layers while the strength of the intralayer exchange coefficient is no longer altered by the electric field. This means that the ferromagnetic coupling of the Cr atoms in the layer is extremely strong.

Refer to caption
Figure 3: (Color online) (a) The intralayer (squares) and interlayer (lozenges) isotropic exchange coefficients as a function of the external electric field, ElE_{l}. The change of the sign of the interlayer exchange shows a magnetic transition from the AFM to a FM between the layers. The ferromagnetic coupling of Cr atoms in the layer is extremely strong. (b) The non-zero component of interlayer and intralayer DM vectors of the bilayer CrI3 structures as a function of ElE_{l}. The intralayer exchange coupling is changed significantly by the electric field. Although the zz and xx components of the interlayer DM interactions are independent of the electric field, the sign and the value of DvyD_{v}^{y} vary.

The extended Hamiltonian for the DM term is given by

DM=\displaystyle{\cal H}_{DM}= 12[𝐃t12(𝐒1×𝐒2)+𝐃t12(𝐒1×𝐒2)+𝐃t12′′\displaystyle\frac{1}{2}[{\bf D}_{t12}\cdot({\bf S}_{1}\times{\bf S}_{2})+{\bf D}_{t12^{\prime}}\cdot({\bf S}_{1}\times{\bf S}_{2^{\prime}})+{\bf D}_{t12^{\prime\prime}}\cdot
(𝐒1×𝐒2′′)+𝐃t21(𝐒2×𝐒1)+𝐃t21(𝐒2×𝐒1)\displaystyle({\bf S}_{1}\times{\bf S}_{2^{\prime\prime}})+{\bf D}_{t21}\cdot({\bf S}_{2}\times{\bf S}_{1})+{\bf D}_{t21^{\prime}}\cdot({\bf S}_{2}\times{\bf S}_{1^{\prime}})
+𝐃t21′′(𝐒2×𝐒1′′)+𝐃t34(𝐒3×𝐒4)+𝐃t34\displaystyle+{\bf D}_{t21^{\prime\prime}}\cdot({\bf S}_{2}\times{\bf S}_{1^{\prime\prime}})+{\bf D}_{t34}\cdot({\bf S}_{3}\times{\bf S}_{4})+{\bf D}_{t34^{\prime}}\cdot
(𝐒3×𝐒4)+𝐃t34′′(𝐒3×𝐒4′′)+𝐃t43(𝐒4×𝐒3)\displaystyle({\bf S}_{3}\times{\bf S}_{4^{\prime}})+{\bf D}_{t34^{\prime\prime}}\cdot({\bf S}_{3}\times{\bf S}_{4^{\prime\prime}})+{\bf D}_{t43}\cdot({\bf S}_{4}\times{\bf S}_{3})
+𝐃t43(𝐒4×𝐒3)+𝐃t43′′(𝐒4×𝐒3′′)+\displaystyle+{\bf D}_{t43^{\prime}}\cdot({\bf S}_{4}\times{\bf S}_{3^{\prime}})+{\bf D}_{t43^{\prime\prime}}\cdot({\bf S}_{4}\times{\bf S}_{3^{\prime\prime}})+
𝐃v13(𝐒1×𝐒3)+𝐃v31(𝐒3×𝐒1)+\displaystyle{\bf D}_{v13}\cdot({\bf S}_{1}\times{\bf S}_{3})+{\bf D}_{v31}\cdot({\bf S}_{3}\times{\bf S}_{1})+
𝐃v14(𝐒1×𝐒4)+𝐃v41(𝐒4×𝐒1)+\displaystyle{\bf D}_{v14}\cdot({\bf S}_{1}\times{\bf S}_{4})+{\bf D}_{v41}\cdot({\bf S}_{4}\times{\bf S}_{1})+
𝐃v23(𝐒2×𝐒3)+𝐃v32(𝐒3×𝐒2)+\displaystyle{\bf D}_{v23}\cdot({\bf S}_{2}\times{\bf S}_{3})+{\bf D}_{v32}\cdot({\bf S}_{3}\times{\bf S}_{2})+
𝐃v24(𝐒2×𝐒4)+𝐃v42(𝐒4×𝐒2)],\displaystyle{\bf D}_{v24}\cdot({\bf S}_{2}\times{\bf S}_{4})+{\bf D}_{v42}\cdot({\bf S}_{4}\times{\bf S}_{2})], (9)

where 𝐒i{\bf S}_{i^{\prime}} and 𝐒i′′{\bf S}_{i^{\prime\prime}} are equal to the magnetic moment of ith Cr atom, and 𝐒i=𝐒i′′=𝐒i{\bf S}_{i^{\prime}}={\bf S}_{i^{\prime\prime}}={\bf S}_{i}. On the other hand, the DM vector between two atoms can be obtained by 𝐃ij=Dij𝐮^ij{\bf D}_{ij}=D_{ij}{\hat{\bf u}}_{ij}, where Dij=DjiD_{ij}=D_{ji} is a scalar and 𝐮^ij{\hat{\bf u}}_{ij} is a unit-vector in the direction of the line connecting ii and jj atoms Yang et al. (2015). Therefore, 𝐃ij=𝐃ji{\bf D}_{ij}=-{\bf D}_{ji}, 𝐃t12=𝐑z(120)𝐃t12{\bf D}_{t12^{\prime}}={\bf R}_{z}(120){\bf D}_{t12}, 𝐃t12′′=𝐑z(240)𝐃t12{\bf D}_{t12^{\prime\prime}}={\bf R}_{z}(240){\bf D}_{t12} and 𝐑z(θ){\bf R}_{z}(\theta) is a θ\theta degree rotation around the zz-axis. Three first terms of Eq. (9) can be written as

\displaystyle{\cal H} =DM12{}_{DM12}= (10)
(𝐃t12+𝐑z(120)𝐃t12+𝐑z(240)𝐃t12)(𝐒1×𝐒2).\displaystyle({\bf D}_{t12}+{\bf R}_{z}(120){\bf D}_{t12}+{\bf R}_{z}(240){\bf D}_{t12})\cdot({\bf S}_{1}\times{\bf S}_{2}).

If 𝐃t12=(Dt12xi^+Dt12yj^+Dt12zk^){\bf D}_{t12}=(D_{t12}^{x}\hat{i}+D_{t12}^{y}\hat{j}+D_{t12}^{z}\hat{k}), then (𝐃t12+𝐑z(120)𝐃t12+𝐑z(240)𝐃t12)=3Dt12zk^({\bf D}_{t12}+{\bf R}_{z}(120){\bf D}_{t12}+{\bf R}_{z}(240){\bf D}_{t12})=3D_{t12}^{z}\hat{k}, hence only the zz-component of the DM vector between two Cr atoms in the layers of the bilayer CrI3 can be entered in the model Hamiltonian of Eq. (1). In the similar way, we can write (𝐃t34+𝐑z(120)𝐃t34+𝐑z(240)𝐃t34)=3Dt34zk^({\bf D}_{t34}+{\bf R}_{z}(120){\bf D}_{t34}+{\bf R}_{z}(240){\bf D}_{t34})=3D_{t34}^{z}\hat{k} for the DM interaction between Cr atoms in the top layer. Due to the symmetries between the layers, it is possible to consider Dt12z=Dt34z=DtzD_{t12}^{z}=D_{t34}^{z}=D_{t}^{z}. On the other hand, 𝐃t12=𝐃t21{\bf D}_{t12}=-{\bf D}_{t21} and 𝐃t13=𝐃t31{\bf D}_{t13}=-{\bf D}_{t31}, hence for the first five lines of Eq. (9), we have the intralayer terms of the DM interaction as,

DMt=12[6Dtzk^(𝐒1×𝐒2)+6Dtzk^(𝐒3×𝐒4)].{\cal H}_{DMt}=\frac{1}{2}[6D_{t}^{z}\hat{k}\cdot({\bf S}_{1}\times{\bf S}_{2})+6D_{t}^{z}\hat{k}\cdot({\bf S}_{3}\times{\bf S}_{4})]. (11)

Therefore, the symmetries of the HT phase dictate that just the zz-component of the intralayer DM interaction vector, DtzD_{t}^{z}, has a non-zero value.

Refer to caption
Figure 4: (Color online) (a) The normalized magnetization of the HT bilayers as a function of temperature under various external electric field. (b) The magnetic order temperature, TT^{*} (T=TNT^{*}=T_{N} for El2E_{l}\leqslant 2 and T=TCT^{*}=T_{C} for El2.5E_{l}\geqslant 2.5), (lozenges) and magnetic anisotropy energy (squares) of the HT bilayer CrI3 versus the external electric field. The magnetic order temperature of the HT bilayer is increased by applying the electric field that arises from the increasing of the intralayer isotropic exchange coefficient. Furthermore, the magnetic moments of the Cr atoms are in out-of-plane direction and it is unchangeable by the external electric field.

Moreover, the geometry of the bilayer crystal indicates that 𝐮^13=𝐮^24\hat{\bf u}_{13}=\hat{\bf u}_{24} and 𝐮^32=𝐑y(180)𝐮^14\hat{\bf u}_{32}={\bf R}_{y}(180)\hat{\bf u}_{14}, then we can assume 𝐃v13=𝐃v24{\bf D}_{v13}={\bf D}_{v24} and 𝐃v32=𝐑y(180)𝐃v14{\bf D}_{v32}={\bf R}_{y}(180){\bf D}_{v14}. To obtain the interlayer contribution of the DM vector, we have to focus on the last four lines of Eq. (9) which can be summarized in the following form,

DMv\displaystyle{\cal H}_{DMv} =𝐃v13(𝐒1×𝐒3)+𝐃v13(𝐒2×𝐒4)+𝐃v14\displaystyle={\bf D}_{v13}\cdot({\bf S}_{1}\times{\bf S}_{3})+{\bf D}_{v13}\cdot({\bf S}_{2}\times{\bf S}_{4})+{\bf D}_{v14}\cdot
(𝐒1×𝐒4)+𝐑y(180)𝐃v14(𝐒3×𝐒2)\displaystyle({\bf S}_{1}\times{\bf S}_{4})+{\bf R}_{y}(180){\bf D}_{v14}\cdot({\bf S}_{3}\times{\bf S}_{2}) (12)

where 𝐃v13=(Dv1x,Dv1y,Dv1z){\bf D}_{v13}=(D_{v1}^{x},D_{v1}^{y},D_{v1}^{z}), 𝐃v14=(Dv2x,Dv2y,Dv2z){\bf D}_{v14}=(D_{v2}^{x},D_{v2}^{y},D_{v2}^{z}) and 𝐑y(180)𝐃v14=(Dv2x,Dv2y,Dv2z){\bf R}_{y}(180){\bf D}_{v14}=(-D_{v2}^{x},D_{v2}^{y},-D_{v2}^{z}). Eight different spin configurations are needed to obtain enough information about the interlayer and intralayer DM vectors given in Table 2.

Table 2: Eight spin configurations for calculation of the interlayer and intralayer DM vectors of the HT bilayer.
Config. 1st1^{st} 2nd2^{nd} 3rd3^{rd} 4th4^{th} 5th5^{th} 6th6^{th} 7th7^{th} 8th8^{th}
𝐒1{\bf S}_{1} Si^S\hat{i} Si^S\hat{i} Si^S\hat{i} Si^S\hat{i} Sk^S\hat{k} Sk^S\hat{k} Sk^S\hat{k} Sk^S\hat{k}
𝐒2{\bf S}_{2} S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}- Si^S\hat{i} Si^S\hat{i} Sk^S\hat{k} Sk^S\hat{k} Sk^S\hat{k} Sk^S\hat{k}
cos(π6)j^)\cos(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j})
𝐒3{\bf S}_{3} Si^S\hat{i} Si^S\hat{i} S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}- S(cos(π6)k^+S(\cos(\frac{\pi}{6})\hat{k}+ S(cos(π6)k^S(\cos(\frac{\pi}{6})\hat{k}- S(cos(π6)k^+S(\cos(\frac{\pi}{6})\hat{k}+ S(cos(π6)k^S(\cos(\frac{\pi}{6})\hat{k}-
sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)i^)\sin(\frac{\pi}{6})\hat{i}) sin(π6)i^)\sin(\frac{\pi}{6})\hat{i}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j})
𝐒4{\bf S}_{4} S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}- S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}- S(cos(π6)k^+S(\cos(\frac{\pi}{6})\hat{k}+ S(cos(π6)k^S(\cos(\frac{\pi}{6})\hat{k}- S(cos(π6)k^+S(\cos(\frac{\pi}{6})\hat{k}+ S(cos(π6)k^S(\cos(\frac{\pi}{6})\hat{k}-
sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)i^)\sin(\frac{\pi}{6})\hat{i}) sin(π6)i^)\sin(\frac{\pi}{6})\hat{i}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j})

The DM interactions are calculated by mapping the considered spin configurations on the model Hamiltonian and comparing the spin-dependent DFT-obtained total energies. At this place, we want to focus on the DM terms described in Eqs. (11) and (12).

DMt(1st)=6S[(Dtzk^)(sin(π/6)k^)],\displaystyle{\cal H}_{DMt(1^{st})}=6S[(D_{t}^{z}\hat{k})\cdot(\sin(\pi/6)\hat{k})],
DMt(2nd)=6S[(Dtzk^)(sin(π/6)k^)],\displaystyle{\cal H}_{DMt(2^{nd})}=-6S[(D_{t}^{z}\hat{k})\cdot(\sin(\pi/6)\hat{k})],
DMv(1st)=DMv(2nd)=0.\displaystyle{\cal H}_{DMv(1^{st})}={\cal H}_{DMv(2^{nd})}=0. (13)

By imposing the above relations into Eq. (1), the total energy equations are defined and the zz-component of the intralayer DM vector is given by

Dtz=E1stE2nd12S2sin(π/6).D_{t}^{z}=\frac{E_{1st}-E_{2nd}}{12S^{2}\sin(\pi/6)}. (14)

In a similar way, we can find the DM terms of model Hamiltonian for the other configurations and the interlayer DM vector of the HT bilayer is achieved by DFT-obtained total energies according to

Dvz\displaystyle D_{v}^{z} =Dv1z+Dv2z=E3rdE4th4S2sin(π/6),\displaystyle=D_{v1z}+D_{v2z}=\frac{E_{3rd}-E_{4th}}{4S^{2}\sin(\pi/6)},
Dvy\displaystyle D_{v}^{y} =Dv1y=E5thE6th4S2sin(π/6),\displaystyle=D_{v1y}=\frac{E_{5th}-E_{6th}}{4S^{2}\sin(\pi/6)},
Dvx\displaystyle D_{v}^{x} =Dv1x+Dv2x=E8thE7th4S2sin(π/6).\displaystyle=D_{v1x}+D_{v2x}=\frac{E_{8th}-E_{7th}}{4S^{2}\sin(\pi/6)}. (15)

We obtain Dtz=1.1D_{t}^{z}=1.1 μ\mueV, Dvz=3.5D_{v}^{z}=3.5 μ\mueV , Dvy=38.8D_{v}^{y}=-38.8 μ\mueV and Dvx=3.1D_{v}^{x}=3.1 μ\mueV for the HT bilayer. Accordingly, the intralayer DM interaction is ignorable in comparison with the interlayer one. The results show the strongest component of the interlayer DM vector is in the yy-direction; this means the magnetic moments of Cr atoms in the layers tend to rotate along the xx-zz plane, as long as they align in the zz-direction because of the high MAE of the bilayer.

The interlayer and intralayer DM interactions can be tuned by an external electric field, very similar to the isotropic exchange coupling. In Fig. 3 (b), it is observed that the intralayer exchange coupling is changed significantly by the electric field. Although the zz and xx components of the interlayer DM interactions are independent of the electric field, the sign and the value of DvyD_{v}^{y} vary. The sign of the yy-component of the DM changes from negative to positive values between 22-2.52.5 V/nm where a transition from the AFM to FM is achieved for the HT bilayer. In fact, a phase transition needs a spin-canting between layers in the xx-zz plane satisfied by the non-zero DvyD_{v}^{y} which is in the same order of the JvJ_{v} in the HT bilayer. The application of the electric field shifts the atomic positions slightly, resulting in a change in the orbital couplings between atomic neighbors. Indeed, if the egt2ge_{g}-t_{2g} coupling, the resulting FM exchange will larger than that causing the AFM exchange (egege_{g}-e_{g} orbital coupling), the bilayer is in the FM ground states and vice versa Sivadas et al. (2018); Akram et al. (2021). Upon increasing the electric field, an AFM to FM transition is observed in the HT bilayer CrI3, showing the dominant effect of egt2ge_{g}-t_{2g} coupling. We expect these changes to be accompanied by spin tilt, and consequently, the DM interactions representing the spin tilting ability are tuned by the electric field.

To calculate the magnetic order temperature, Curie TCT_{C}, or Néel TNT_{N} temperature of the FM or AFM bilayer, a sample of 50×\times50 nm2 is considered and the metropolis Monte Carlo algorithms are used for the calculation of temperature-dependent magnetization (see Fig. 4 (a)). According to Fig. 4 (b), the magnetic order temperature of the HT bilayer is increased by applying the electric field that arises from the increasing of the intralayer isotropic exchange coefficient.

The variation of MAE=(EE)/ηS2MAE=(E_{\perp}-E_{\parallel})/\eta S^{2} as a function of the electric field is shown in Fig. 4 (b). It shows that the magnetic moments of the Cr atoms are in out-of-plane direction, and it is unchangeable by the external electric field.

III.2 Exchange coefficients in the LT phase

Now we turn our attention to another phase of the bilayer system. Four Cr atoms in a unit cell of the LT-phase bilayer CrI3 prefer to determine a parallel spin orientation in which the minimized total energy belongs to the FM magnetic configuration Morell et al. (2019). Here, similar to the HT phase, we are interested in obtaining the symmetric and antisymmetric exchange coefficients of the LT bilayer. We should note that the intralayer nearest neighbors of the LT is equivalent to that in the HT phase, while there is only an interlayer nearest neighbor bond between the Cr atoms numbered by 22 and 33 in Fig. 1 (d). In the case of the LT bilayer CrI3, the first two terms of Eq. (1) are extended to be as

=\displaystyle{\cal H}= 12(ηtJt[(𝐒1𝐒2)+(𝐒2𝐒1)+(𝐒3𝐒4)+(𝐒4𝐒3)]+\displaystyle\frac{1}{2}(\eta_{t}J_{t}[({\bf S}_{1}\cdot{\bf S}_{2})+({\bf S}_{2}\cdot{\bf S}_{1})+({\bf S}_{3}\cdot{\bf S}_{4})+({\bf S}_{4}\cdot{\bf S}_{3})]+
ηvJv[(𝐒2𝐒3)+(𝐒3𝐒2)])+ηγ|Sz|2,\displaystyle\eta_{v}J_{v}[({\bf S}_{2}\cdot{\bf S}_{3})+({\bf S}_{3}\cdot{\bf S}_{2})])+\eta\gamma|S_{z}|^{2}, (16)

where ηt=3\eta_{t}=3 and ηv=1\eta_{v}=1 are the number of the intralayer and interlayer nearest neighbors. We consider four different magnetic configurations of Cr atoms, as a FM (𝐒1=Sk^{\bf S}_{1}=S\hat{k}, 𝐒2=Sk^{\bf S}_{2}=S\hat{k}, 𝐒3=Sk^{\bf S}_{3}=S\hat{k}, 𝐒4=Sk^{\bf S}_{4}=S\hat{k}), AFM1 (𝐒1=Sk^{\bf S}_{1}=S\hat{k}, 𝐒2=Sk^{\bf S}_{2}=S\hat{k}, 𝐒3=Sk^{\bf S}_{3}=-S\hat{k}, 𝐒4=Sk^{\bf S}_{4}=-S\hat{k}), AFM2 (𝐒1=Sk^{\bf S}_{1}=S\hat{k}, 𝐒2=Sk^{\bf S}_{2}=-S\hat{k}, 𝐒3=Sk^{\bf S}_{3}=-S\hat{k}, 𝐒4=Sk^{\bf S}_{4}=S\hat{k}) and AFM3 (𝐒1=Sk^{\bf S}_{1}=S\hat{k}, 𝐒2=Sk^{\bf S}_{2}=-S\hat{k}, 𝐒3=Sk^{\bf S}_{3}=S\hat{k}, 𝐒4=Sk^{\bf S}_{4}=-S\hat{k}), and thus able to find interlayer and intralayer symmetric exchange coefficients. By applying the spin vectors on Eq. (16), the relations for the total energies of the magnetic configurations are given by,

EFM=E0+2S2ηtJt+S2ηvJv+ηγS2,\displaystyle E_{FM}=E_{0}+2S^{2}\eta_{t}J_{t}+S^{2}\eta_{v}J_{v}+\eta\gamma S^{2},
EAFM1=E0+2S2ηtJtS2ηvJv+ηγS2,\displaystyle E_{AFM1}=E_{0}+2S^{2}\eta_{t}J_{t}-S^{2}\eta_{v}J_{v}+\eta\gamma S^{2},
EAFM2=E02S2ηtJt+S2ηvJv+ηγS2,\displaystyle E_{AFM2}=E_{0}-2S^{2}\eta_{t}J_{t}+S^{2}\eta_{v}J_{v}+\eta\gamma S^{2},
EAFM3=E02S2ηtJtS2ηvJv+ηγS2.\displaystyle E_{AFM3}=E_{0}-2S^{2}\eta_{t}J_{t}-S^{2}\eta_{v}J_{v}+\eta\gamma S^{2}. (17)

The DFT-obtained total energies give us the symmetric exchange coefficients as

Jt\displaystyle J_{t} =(EFM+EAFM1)(EAFM2+EAFM3)8S2ηt,\displaystyle=\frac{(E_{FM}+E_{AFM1})-(E_{AFM2}+E_{AFM3})}{8S^{2}\eta_{t}},
Jv\displaystyle J_{v} =(EFMEAFM1)2S2ηv.\displaystyle=\frac{(E_{FM}-E_{AFM1})}{2S^{2}\eta_{v}}. (18)

The results displayed in Fig. 5 (a) show the interlayer and intralayer symmetric exchange couplings possess negative and consistent values with the FM ground state of the LT bilayers. The sign of JtJ_{t} and JvJ_{v} remain negative by applying the electric field, while their amounts are slightly increased. In fact, the electric field develops the ferromagnetic coupling of layers increase in the LT bilayer.

Refer to caption
Figure 5: (Color online) (a) The intralayer (squares) and interlayer (lozenges) isotropic exchange coefficients and (b) the zz-component of interlayer and intralayer DM vectors of the LT bilayer CrI3 as a function of the external electric field, ElE_{l}. The sign of the exchange coefficients remain negative by applying the electric field, while their amounts are slightly increased showing that the electric field makes the ferromagnetic coupling of layers stronger. Furthermore, the zz- component of the DM interaction of the LT bilayer increases by the electric field meaning that the spin-canting are in the plane of the Cr atoms.

The third term of the Hamiltonian in Eq. (1) can be written for the LT bilayer as,

DM\displaystyle{\cal H}_{DM} =12[𝐃t12(𝐒1×𝐒2)+𝐃t12(𝐒1×𝐒2)+𝐃t12′′\displaystyle=\frac{1}{2}[{\bf D}_{t12}\cdot({\bf S}_{1}\times{\bf S}_{2})+{\bf D}_{t12^{\prime}}\cdot({\bf S}_{1}\times{\bf S}_{2}^{\prime})+{\bf D}_{t12^{\prime\prime}}
(𝐒1×𝐒2′′)+𝐃t21(𝐒2×𝐒1)+𝐃t21(𝐒2×𝐒1)\displaystyle\cdot({\bf S}_{1}\times{\bf S}_{2}^{\prime\prime})+{\bf D}_{t21}\cdot({\bf S}_{2}\times{\bf S}_{1})+{\bf D}_{t21^{\prime}}\cdot({\bf S}_{2}\times{\bf S}_{1}^{\prime})
+𝐃t21′′(𝐒2×𝐒1′′)+𝐃t34(𝐒3×𝐒4)+𝐃t34(𝐒3\displaystyle+{\bf D}_{t21^{\prime\prime}}\cdot({\bf S}_{2}\times{\bf S}_{1}^{\prime\prime})+{\bf D}_{t34}\cdot({\bf S}_{3}\times{\bf S}_{4})+{\bf D}_{t34^{\prime}}\cdot({\bf S}_{3}
×𝐒4)+𝐃t34′′(𝐒3×𝐒4′′)+𝐃t43(𝐒4×𝐒3)+𝐃t43\displaystyle\times{\bf S}_{4}^{\prime})+{\bf D}_{t34^{\prime\prime}}\cdot({\bf S}_{3}\times{\bf S}_{4}^{\prime\prime})+{\bf D}_{t43}\cdot({\bf S}_{4}\times{\bf S}_{3})+{\bf D}_{t43^{\prime}}
(𝐒4×𝐒3)+𝐃t43′′(𝐒4×𝐒3′′)+𝐃v23(𝐒2×𝐒3)\displaystyle\cdot({\bf S}_{4}\times{\bf S}_{3}^{\prime})+{\bf D}_{t43^{\prime\prime}}\cdot({\bf S}_{4}\times{\bf S}_{3}^{\prime\prime})+{\bf D}_{v23}\cdot({\bf S}_{2}\times{\bf S}_{3})
+𝐃v32(𝐒3×𝐒2)].\displaystyle+{\bf D}_{v32}\cdot({\bf S}_{3}\times{\bf S}_{2})]. (19)

Similar to the HT phase, the symmetric of atomic positions dictates to have only the zz-component of the intralayer DM interaction in the Hamiltonian which is given by Eq. (11). For the interlayer DM interaction of the LT phase

DMv=𝐃v23(𝐒2×𝐒3).{\cal H}_{DMv}={\bf D}_{v23}\cdot({\bf S}_{2}\times{\bf S}_{3}). (20)

It should be noticed that there is a C3C_{3} rotation axis on the bonding length between Cr atoms numbered by 22 and 33 in the unit cell of the LT bilayers (see Fig. 1 (d)). According to the Moriya symmetry rules Moriya (1960), the DM interaction between atoms 22 and 33 is parallel to the C3C_{3} axis which is along the zz-direction. Therefore, we need to find the zz-component of the interlayer DM interaction which can be obtained by two spin configurations (see Table 3). Finally, the zz-component of intralayer and interlayer DM interactions are respectively given by

Dtz=\displaystyle D_{t}^{z}= (E1stE2nd)+(E3rdE4th)12S2sin(π/6),\displaystyle\frac{(E_{1st}-E_{2nd})+(E_{3rd}-E_{4th})}{12S^{2}\sin(\pi/6)},
Dvz=\displaystyle D_{v}^{z}= E3rdE4th2S2sin(π/6).\displaystyle\frac{E_{3rd}-E_{4th}}{2S^{2}\sin(\pi/6)}. (21)
Table 3: Four spin configurations for calculation of the interlayer and intralayer DM vectors of the LT bilayer CrI3.
Config. 1st1^{st} 2nd2^{nd} 3rd3^{rd} 4th4^{th}
𝐒1{\bf S}_{1} Si^S\hat{i} Si^S\hat{i} Si^S\hat{i} Si^S\hat{i}
𝐒2{\bf S}_{2} S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}- Si^S\hat{i} Si^S\hat{i}
sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j})
𝐒3{\bf S}_{3} Si^S\hat{i} Si^S\hat{i} S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}-
sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j})
𝐒4{\bf S}_{4} S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}- S(cos(π6)i^+S(\cos(\frac{\pi}{6})\hat{i}+ S(cos(π6)i^S(\cos(\frac{\pi}{6})\hat{i}-
sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j}) sin(π6)j^)\sin(\frac{\pi}{6})\hat{j})

The negligible values of Dtz=0.4D_{t}^{z}=-0.4 μ\mueV and Dvz=0.3D_{v}^{z}=-0.3 μ\mueV are obtained using our spin-dependent DFT-based calculations. It should be noted that there is an inversion symmetry in the center of the 22 and 33 atomic bond in the LT bilayer, so DvzD_{v}^{z} should be zero, and therefore, there is no any DM interaction between the layers. We investigate the effect of the external electric field on the interlayer and intralayer DM interactions of the LT phase, and the results are reported in Fig. 5 (b). It is observed that the interlayer and intralayer DM interactions of the LT phase are increased by developing the electric field. It is interesting that the order of DM interaction variation is similar for both LT and HT phases, but their behavior is different. In the presence of an electric field, just the zz-component of the DM interaction of the LT bilayer increases; this means that the spin-canting are in the plane of the Cr atoms. On the other hand, the interlayer symmetric exchange coupling, JvJ_{v}, is much stronger than DvzD_{v}^{z}, subsequently, the FM coupling survives in a competition between FM coupling of layers and in-plane canting of spins.

In the HT bilayer, the amount of the interlayer DM vector and interlayer symmetric exchange coupling are in the same order. In fact, they are in a close competition with each other and it remains in the external electric field. It is intriguing that the both of them leads to a transition from the AFM to FM in the HT phase; the interlayer DM interaction is dominant in the yy-direction which leads to the spin canting in the xx-zz-plane and, as a consequence, helps to the rotation of the spin-direction from the AFM to FM and also the sign of JvJ_{v} is changed by applying the electric field.

It should be noticed that the interlayer and intralayer isotropic exchange coupling in the Heisenberg Hamiltonian of the bilayer CrI3 have been calculated in Ref. Lei et al. (2021) which agrees well with our work. They obtained Jt=6.49J_{t}=-6.49 meV (by DFT) and Jv=+0.10J_{v}=+0.10 meV (by experimental results from Ref. Huang et al. (2017)) which are close to our findings (Jt=7.00J_{t}=-7.00 meV and Jv=+0.8J_{v}=+0.8 meV) in order of magnitude and sign for HT bilayer CrI3. They have also shown that the isotropic exchange couplings are tuned by the electric field. There are DFT-based calculations Jiang et al. (2019); Morell et al. (2019); Lei et al. (2021) and experimental results Huang et al. (2018); Jiang et al. (2018b, a) showing that the interlayer exchange coupling sign of HT phase of bilayer CrI3 is changed by the electric field. One can find the DFT-obtained exchange couplings of different stacks of bilayer CrI3 in the absence of an electric field in Refs. Sivadas et al. (2018); Jiang et al. (2019); Leon et al. (2020); Akram et al. (2021). In contrast, the interlayer and intralayer DM interactions of the bilayer CrI3 have not been calculated in previous works, and we report here their electric field tunable behavior.

The temperature dependence of the magnetic moments of the Cr atoms, M(T)M(T) in the LT bilayer is illustrated in Fig. 6 (a). It is shown that M(T)M(T) is nearly independent of the electric field. The variation of the Currie temperature, TCT_{C}, by the electric field arises directly from the variation of JtJ_{t} and it is gradually increased in a more intense electric field (see Fig. 6 (b)). The MAE of the LT phase as a function of the electric field is shown in Fig. 6 (b). The MAE remains negative by varying the electric field so the easy axis of the LT bilayer lies in the zz-direction and its value does not change significantly.

We obtained a Néel temperature of 44 K for the AFM-HT bilayer and a Curie temperature of 59 K for the FM-LT bilayer CrI3, which are tuned by the electric field due to the variation in the parameters of the spin model Hamiltonian. Increasing the number of layers from monolayer to bulk increases the critical temperature between 45 K and 61 K Huang et al. (2017); Kashin et al. (2020). A Néel temperature of 45 K was reported experimentally for the HT bilayer Huang et al. (2018). The larger Curie temperature is due to the stronger FM interlayer exchange coupling in LT bilayer CrI3 Wang and Sanyal (2021).

Refer to caption
Figure 6: (Color online) (a) The magnetization of the LT bilayers as a function of temperature under different external electric field. (b) The Curie temperature (lozenges) and magnetic anisotropy energy (squares) of LT bilayer CrI3 versus external electric field. Notice that the MAE remains negative by changing the electric field, therefore, the easy axis of the LT bilayer is in the zz-direction and its value does not change significantly.

Although the magnetic coupling coefficients of both the LT and HT bilayers are tuned, the magnetic ground state of the HT bilayer can be manipulated by the electric field. To perceive the possibility of manipulating magnetic textures in the bilayer CrI3, we simulate their spin dynamics in the presence of the external electric and magnetic fields and the emergence of the magnetic domains and skyrmion patterns are explored.

To obtain the spin dynamics of the bilayer, the atomic position of the system is constant while different atomic magnetic moments are appeared by time evolutions Evans et al. (2014). Here, we obtained that the formation energy is 1000 times more than the exchange energy in the bilayer CrI3 which guarantees the stability of the atomic structure under spin evolutions. In fact, we want to mention that the spin evolution happens in a constant atomic structure in any electric field. Under the electric field, the atomic positions are relaxed by DFT-based calculations and leading variation of the interband and intraband isotropic exchange and DM coefficients, so the spin dynamics are calculated in the new structure.

III.3 Spin dynamics of the HT phase

In the HT bilayer, the AFM coupling between layers is dictated in the spin dynamics of the system. The spin dynamics of the top and bottom layers are plotted in Fig. 7. In T=0.3T=0.3 K, it is observed that the bilayer is divided into two areas in which layers are coupled in form of an anti-ferromagnetic. While the bottom layer has a magnetic moment in the +z+z-direction, the top layer has a magnetic moment in the z-z-direction and vice versa; this means that the AFM order is constant between the layers. This regularity is also established on the domain walls, where the perpendicular components of the magnetic moments, MM_{\perp}, of two layers are in opposite directions. The results of the simulations show that the created magnetic texture is stable in size, but MM_{\perp} is slowly changed to be perpendicular to the domain wall by the time evolution as shown in Figs. 7 (a)-(c). Thus, an ultrathin Néel-type domain wall appears on both the top and bottom layers. It is clear that the zero magnetic moments are obtained for the HT bilayer, but there is an interesting microscopic magnetic pattern on each layer.

Refer to caption
Figure 7: (Color online) The magnetic texture of the top and bottom layers of the HT bilayer in El=0E_{l}=0 V/nm, B=0B=0, (a) T=0.3T=0.3 K, time=2{\text{time}}=2 ns, (b) T=0T=0, time=2.2{\text{time}}=2.2 ns and (c) T=0T=0, time=3{\text{time}}=3 ns. The blue (red) areas show where the magnetic moments of Cr atoms are in the -z(+z)-direction. M{M}_{\perp} is represented by black arrows on the domain walls. The created magnetic texture is stable in size, but M{M}_{\perp} is slowly changed to be perpendicular to the domain wall by the time evolution.

We obtain the symmetric and asymmetric exchange coupling coefficients of the spin Hamiltonian depending on the external electric field, therefore, we explore the variation of the spin dynamics of the bilayer by the electric field shown in Figs. 8 (a)-(c). In El=1E_{l}=1 V/nm, the magnetic order of the HT bilayer is the AFM and JtJ_{t} and JvJ_{v} are near to their values in the absence of the electric field case, while the value of DtzD_{t}^{z} is increased to 25.525.5 μ\mueV, therefore the new spin texture is created (see Fig. 8 (b)). If we focus on one of the layers, we see that the wider magnetic domains in El=0E_{l}=0 case (Fig. 8 (a)) changes to two smaller quasi-circle magnetic domains with finite diameters (nearly 2525 nm and 1010 nm). Skyrmion patterns occur around the magnetic moments, while their chirality is different for the top and bottom layers. The places and sizes of magnetic domains are the same in the top and bottom layers, however, the direction of the magnetic moments are exactly opposite for inside and outside of the domain walls. We obtain these stable quasi-circle magnetic domains under the time evolution.

Refer to caption
Figure 8: (Color online) The magnetic texture of the top and bottom layers of the HT bilayer in (a) El=0E_{l}=0, (b) El=1E_{l}=1 V/nm and (c) El=2E_{l}=2 V/nm, T=0.3T=0.3 K, B=0B=0 and time=2{\text{time}}=2 ns. The skyrmion patterns occur around the magnetic moments, while their chirality is different for the top and bottom layers. The places and sizes of magnetic domains are the same in the top and bottom layers, however, the direction of the magnetic moments are exactly opposite for inside and outside of the domain walls.

For the HT bilayer, the amount of the intralayer DM interaction reaches to 4141 μ\mueV by increasing the electric field to El=2E_{l}=2 V/nm, and the interlayer isotropic exchange coupling is equal to 20-20 μ\mueV. Although the value of JvJ_{v} is reduced, the negative sign shows the bilayer is still in the AFM phase. These changes lead to a new spin texture in the bilayer. In Fig. 8 (c), 30 nm-diameter quasi-circle magnetic domains are created in the both top and bottom layers with opposite spin orientation, so the AFM configuration of the bilayer is conserved. Most importantly, the diameter, position, number of the magnetic domain and the domain wall chirality are related to the external electric field. In fact, the magnetic texture of the HT bilayer can be manipulated by an external electric field owing to the variation of the DM and isotropic exchange coupling interactions. Therefore, stable magnetic domains are constituted in the HT bilayer that possess fascinating microscopic information, while the total magnetism is zero in the macroscopic level.

Refer to caption
Figure 9: (Color online) The magnetic texture of the top and bottom layers of HT bilayer in El=3E_{l}=3 V/nm, T=0.3T=0.3 K, B=0B=0 and time=2{\text{time}}=2 ns. Notice that the the magnetic domain in the FM bilayer is stable under the time evolution.

In El=3E_{l}=3 V/nm, the layers are coupled in a form of a ferromagnetic. This phase transition is obviously observed in the spin dynamics simulation which both layers have the same spin orientation (see Fig. 9) and originates from the positive sign of JvJ_{v}. On the other hand, the perpendicular-component of the magnetic moments of the top and bottom layers are similar on the domain walls. The spin dynamics simulations show that the obtained magnetic domain in the FM bilayer is still stable under the time evolution.

Refer to caption
Figure 10: (Color online) The spin dynamic results of the HT bilayer in El=0E_{l}=0 V/nm, (a) B=0B=0, T=0.3T=0.3 K and time=2time=2 ns, (b) B=0.5B=0.5 T, T=0.3T=0.3 K and time=2{\text{time}}=2 ns and (c) B=0.5B=0.5 T, T=0T=0 K and time=3{\text{time}}=3 ns. the external magnetic field is in competition with the interlayer exchange coupling and desire to bring the spin of the Cr atoms parallel to the external magnetic field.

According to Figs. 10 (a)-(c), our simulation results show that the external magnetic field changes the spin dynamic of the bilayer. In the absence of an external electric field, Fig.10 (b) shows that there is an area with FM configurations in the AFM background of the bilayer. Indeed, the external magnetic field is in competition with the interlayer exchange coupling and desire to accompany the spin of the Cr atoms parallel to the external magnetic field. It is not completely winner in the competition when BextB_{ext} is less than 0.50.5 T. By increasing the external magnetic field to Bext=2B_{ext}=2 T, the layers gain parallel magnetic moments in the direction of the applied magnetic field.

Refer to caption
Figure 11: (Color online) The spin dynamic results of the HT bilayer in El=2E_{l}=2 V/nm, B=0.01B=0.01 T (a) T=0.3T=0.3 K, time=2time=2 ns, (b) T=0T=0 K, time=2.2{\text{time}}=2.2 ns and (c) T=0T=0 K, time=3{\text{time}}=3 ns. Notice, here, the applied magnetic field can not create a phase transition in the magnetic order of the HT bilayer.

Another example is the system when exposed to an external electric field, El=2E_{l}=2 V/nm. In Bext=0.5B_{ext}=0.5 T, the AFM pattern is completely disappeared and the magnetic moments of the Cr atoms are parallel. In the competition between exchange coupling coefficients and the Zeeman energy, the Zeeman energy of the strong magnetic field has conquered. In the presence of 0.010.01 T external magnetic field, the simulation results (Figs. 11(a)-(c)) show that the quasi-circular magnetic domain form converts to a quasi-elliptical shape, but the AFM configuration between the top and bottom layers is conserved because the applied magnetic field is not large enough to create a phase transition in the magnetic ordering of the bilayer. Interestingly, the obtained magnetic domain and almost domain wall do not change by the time evolution after cooling time (2 ns). Accordingly, a stable magnetic texture can be obtained and tuned by a weak external magnetic field, while the strong magnetic field leads to the magnetic phase transition in the bilayer.

III.4 Spin dynamic of the LT phase

In the LT bilayer, a strong negative interlayer exchange coupling leads to the creation of the FM magnetic configuration. The spin dynamic simulation, in Fig. 12 (a) shows two wide areas with opposite spin directions in each layer while the magnetic moments of the layers are actually analogous. In the boundary of the two areas, the perpendicular magnetic moments of the Cr atoms are dominated and their chirality is the same in both layers. In fact, a 1D spin-wave is created at the boundary and the results show that the chirality of the spin-wave is changed smoothly in time (compare Fig. 12 (a) and (b)). On the other hand, the created areas sizes are stable by the time evolution, so we can consider that the 1D in-plane spin-wave is metastable (due to the change of chirality). In the absence of an electric field, the ignorable DM interaction cannot create a magnetic domain, and the observed spin-wave is due to the demagnetization field. For clarity, we report the results of the spin dynamic simulation in the absence of a demagnetization field in Fig. 12 (c). It is obviously observed that the magnetic domain is disappeared and all the Cr atoms of both layers have parallel magnetic moments due to strong interlayer and intralayer isotropic exchange couplings of the LT bilayer CrI3.

Refer to caption
Figure 12: (Color online) The spin dynamic results of the LT bilayer in El=0E_{l}=0 V/nm, Bext=0B_{ext}=0 T (a) T=0.3T=0.3 K and time=2time=2 ns, (b) T=0T=0 K and time=3{\text{time}}=3 ns and (c) T=0.3T=0.3 K and time=2{\text{time}}=2 ns in the absence of a demagnetization field. A 1D metastable spin-wave is created at the boundary and the chirality of the spin wave is changed smoothly in time. In this case, the Cr atoms have parallel magnetic moments due to strong interlayer and intralayer isotropic exchange couplings.

The interlayer and intralayer DM interactions are increased by applying the external electric field, while the sign of the exchange coupling and their coefficients values remain practically constant in the LT bilayer CrI3. The spin dynamic simulations show the existence of a similar quasi-circular magnetic domain shape in the top and bottom layers in the presence of the electric field. Hence, the meta-stable magnetic domain is disturbed by increasing the DM interaction in El=1E_{l}=1 V/nm and El=2E_{l}=2 V/nm, see Figs. 13 (a)-(c). The results show that this quasi-circular magnetic domains are disappeared by cooling the system and the magnetic moments of the Cr atoms gained parallelized, see Fig. 14 for the LT bilayer under El=1E_{l}=1 V/nm. We can assume that the DM interactions leads to the creation of a quasi-circular domain shape in the layers, but its value is extremely smaller than the exchange coupling to stabilize it in the system, therefore, the obtained magnetic domain is unstable and disappears rapidly in the LT bilayer.

Refer to caption
Figure 13: (Color online) The magnetic texture of the top and bottom layers of the LT bilayer in (a) El=0E_{l}=0, (b) El=1E_{l}=1 V/nm and (c) El=2E_{l}=2 V/nm, T=0.3T=0.3 K, B=0B=0 and time=2{\text{time}}=2 ns. The existence of a similar quasi-circle magnetic domain in the top and bottom layers in the presence of the electric field are seen and the meta-stable magnetic domain is disturbed by increasing the DM interaction.
Refer to caption
Figure 14: (Color online) The time evolution of the top layer of the LT phase under El=1E_{l}=1 V/nm, T=0T=0 K and B=0B=0. The bottom layer has a similar spin dynamic results due to the FM coupling of the layers.

The spin dynamics in the twisted bilayer CrI3 was recently calculated using an arbitrary DM interaction and a local moiré field instead of an isotropic exchange coupling between the layers Akram et al. (2021) while ignoring the demagnetization field. They showed that the twisted texture depends on the twist angle. It is worth noting that applying an electric field is easier than twisting the double layer CrI3 in experiments. It is experimentally shown that the moiré magnetic domain in AFM twisted bilayer magnetic crystals can be controlled by gate voltages Xu et al. (2021). Despite a recent interest in studying spin dynamics in 2D magnets Akram et al. (2021); Xu et al. (2021); Wahab et al. (2021); Tong et al. (2018); Khoshlahni et al. (2019); Behera et al. (2019); Abdul-Wahab et al. (2021) and their promising application in new generation magnetic domain-based devices Allwood et al. (2005); Vélez et al. (2019); Luo et al. (2020), there is no report of spin texture formation in the bilayer CrI3 by considering all DFT-obtained isotropic and anisotropic exchange couplings in the presence of an electric field. Here we calculate the spin dynamics of a large bilayer CrI3 as an experimental domain by solving the LLG equations where the demagnetizing field is important to find the best magnetic ground state of the system. The results show that the magnetic domain in the HT bilayer CrI3 can be tuned by the electric field, which is consistent with the experimental results Xu et al. (2021).

IV Conclusion

The interlayer and intralayer isotropic exchange coupling coefficients and DM interactions are obtained by DFT-based calculations for bilayer CrI3. We have shown that in the presence of an external electric field, interlayer DM interactions have a critical impact on the magnetic phase transition of the HT bilayer. The effects of the external electric field on the parameters of the spin-model Hamiltonian and temperature-dependent magnetic moments of bilayers are explored. Most importantly, a stable magnetic domain in the HT bilayer CrI3 which can be manipulated by the electric and magnetic fields can be constituted owing to the tunable interlayer and intralayer exchange coupling and DM interactions. The spin dynamics simulations show that the HT bilayer CrI3 can represent a promising candidate for spintronic and logic memory applications. The interlayer exchange coupling, on the other hand, is only weakly affected by the external electric field, hence there is no stable magnetic domain in the LT bilayer CrI3. Current experiments can investigate these findings. Furthermore, when disorder and magnetic impurity are taken into account, our results can be extended in a system.

Acknowledgements.
We thank Z. Torbatian and S. Heydari for fruitful discussions. R. A thanks for support from the Australian Research Council Centre of Excellence in Future Low-Energy Electronics Technologies (project number CE170100039).

References

  • Huang et al. (2017) B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A. McGuire, D. H. Cobden, et al., “Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit,” Nature 546, 270 (2017).
  • Deng et al. (2018) Y. Deng, Y. Yu, Y. Song, J. Zhang, N. Z. Wang, Z. Sun, Y. Yi, Y. Z. Wu, S. Wu, J. Zhu, et al., “Gate-tunable room-temperature ferromagnetism in two-dimensional fe 3 gete 2,” Nature 563, 94 (2018).
  • Gong et al. (2017) C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao, W. Bao, C. Wang, Y. Wang, et al., “Discovery of intrinsic ferromagnetism in two-dimensional van der waals crystals,” Nature 546, 265 (2017).
  • Gibertini et al. (2019) M. Gibertini, M. Koperski, A. F. Morpurgo,  and K. S. Novoselov, “Magnetic 2d materials and heterostructures,” Nat. nanotechnol. 14, 408 (2019).
  • Gong and Zhang (2019) C. Gong and X. Zhang, “Two-dimensional magnetic crystals and emergent heterostructure devices,” Science 363, 4450 (2019).
  • Meseguer-Sánchez et al. (2021) J. Meseguer-Sánchez, C. Popescu, J. L. García-Muñoz, H. Luetkens, G. Taniashvili, E. Navarro-Moratalla, Z. Guguchia,  and E. J. Santos, “Coexistence of structural and magnetic phases in van der waals magnet cri3,” Nat. commun. 12, 1 (2021).
  • Lado and Fernández-Rossier (2017) J. L. Lado and J. Fernández-Rossier, “On the origin of magnetic anisotropy in two dimensional CrI3{\mathrm{CrI}}_{3},” 2D Mater. 4, 035002 (2017).
  • Morell et al. (2019) E. S. Morell, A. León, R. H. Miwa,  and P. Vargas, “Control of magnetism in bilayer CrI3{\mathrm{CrI}}_{3} by an external electric field,” 2D Mater. 6, 025020 (2019).
  • Torelli and Olsen (2020) D. Torelli and T. Olsen, “First principles heisenberg models of 2d magnetic materials: the importance of quantum corrections to the exchange coupling,” J. Condens. Matter Phys. 32, 335802 (2020).
  • Frey et al. (2019) N. C. Frey, A. Bandyopadhyay, H. Kumar, B. Anasori, Y. Gogotsi,  and V. B. Shenoy, “Surface-engineered mxenes: electric field control of magnetism and enhanced magnetic anisotropy,” ACS Nano 13, 2831 (2019).
  • Menichetti et al. (2019) G. Menichetti, M. Calandra,  and M. Polini, “Electronic structure and magnetic properties of few-layer cr2ge2te6: the key role of nonlocal electron–electron interaction effects,” 2d Mater. 6, 045042 (2019).
  • Hubert and Schäfer (2008) A. Hubert and R. Schäfer, Magnetic domains: the analysis of magnetic microstructures (Springer Science & Business Media, 2008).
  • Wahab et al. (2021) D. A. Wahab, M. Augustin, S. M. Valero, W. Kuang, S. Jenkins, E. Coronado, I. V. Grigorieva, I. J. Vera-Marun, E. Navarro-Moratalla, R. F. Evans, et al., “Quantum Rescaling, Domain Metastability, and Hybrid Domain-Walls in 2D CrI3 Magnets,” Adv. Mater. 33, 2004138 (2021).
  • Chen et al. (2021) L. Chen, J.-H. Chung, M. B. Stone, A. I. Kolesnikov, B. Winn, V. O. Garlea, D. L. Abernathy, B. Gao, M. Augustin, E. J. G. Santos,  and P. Dai, “Magnetic field effect on topological spin excitations in cri3{\mathrm{cri}}_{3},” Phys. Rev. X 11, 031047 (2021).
  • Akram et al. (2021) M. Akram, H. LaBollita, D. Dey, J. Kapeghian, O. Erten,  and A. S. Botana, “Moiré skyrmions and chiral magnetic phases in twisted crx3 (x= i, br, and cl) bilayers,” Nano Lett. 21, 6633 (2021).
  • Xu et al. (2021) Y. Xu, A. Ray, Y.-T. Shao, S. Jiang, D. Weber, J. E. Goldberger, K. Watanabe, T. Taniguchi, D. A. Muller, K. F. Mak, et al., “Emergence of a noncollinear magnetic state in twisted bilayer cri3,” arXiv preprint arXiv:2103.09850  (2021), https://doi.org/10.48550/arXiv.2103.09850.
  • Jiang et al. (2019) P. Jiang, C. Wang, D. Chen, Z. Zhong, Z. Yuan, Z.-Y. Lu,  and W. Ji, “Stacking tunable interlayer magnetism in bilayer CrI3,” Phys. Rev. B 99, 144401 (2019).
  • Kim et al. (2019) H. H. Kim, B. Yang, S. Li, S. Jiang, C. Jin, Z. Tao, G. Nichols, F. Sfigakis, S. Zhong, C. Li, et al., “Evolution of interlayer and intralayer magnetism in three atomically thin chromium trihalides,” PNAS 116, 11131 (2019).
  • Lei et al. (2021) C. Lei, B. L. Chittari, K. Nomura, N. Banerjee, J. Jung,  and A. H. MacDonald, “Magnetoelectric Response of Antiferromagnetic CrI3 Bilayers,” Nano Lett. 21, 1948 (2021).
  • Song et al. (2019) T. Song, Z. Fei, M. Yankowitz, Z. Lin, Q. Jiang, K. Hwangbo, Q. Zhang, B. Sun, T. Taniguchi, K. Watanabe, et al., “Switching 2D magnetic states via pressure tuning of layer stacking,” Nat. Mater. 18, 1298 (2019).
  • Li et al. (2019) T. Li, S. Jiang, N. Sivadas, Z. Wang, Y. Xu, D. Weber, J. E. Goldberger, K. Watanabe, T. Taniguchi, C. J. Fennie, et al., “Pressure-controlled interlayer magnetism in atomically thin CrI3,” Nat. Mater. 18, 1303 (2019).
  • Sivadas et al. (2018) N. Sivadas, S. Okamoto, X. Xu, C. J. Fennie,  and D. Xiao, “Stacking-dependent magnetism in bilayer CrI3{\mathrm{CrI}}_{3},” Nano Lett. 18, 7658 (2018).
  • Jiang et al. (2018a) S. Jiang, J. Shan,  and K. F. Mak, “Electric-field switching of two-dimensional van der Waals magnets,” Nat. Mater. 17, 406 (2018a).
  • Jiang et al. (2018b) S. Jiang, L. Li, Z. Wang, K. F. Mak,  and J. Shan, “Controlling magnetism in 2d cri 3 by electrostatic doping,” Nat. nanotechnol. 13, 549 (2018b).
  • Kashin et al. (2020) I. Kashin, V. Mazurenko, M. Katsnelson,  and A. Rudenko, “Orbitally-resolved ferromagnetism of monolayer cri3,” 2D Mater. 7, 025036 (2020).
  • Huang et al. (2018) B. Huang, G. Clark, D. R. Klein, D. MacNeill, E. Navarro-Moratalla, K. L. Seyler, N. Wilson, M. A. McGuire, D. H. Cobden, D. Xiao, et al., “Electrical control of 2D magnetism in bilayer CrI3{\mathrm{CrI}}_{3},” Nat. nanotechnol. 13, 544 (2018).
  • Wang and Sanyal (2021) D. Wang and B. Sanyal, “Systematic study of monolayer to trilayer cri3: Stacking sequence dependence of electronic structure and magnetism,” J. Phys. Chem. C . 125, 18467 (2021).
  • Chen et al. (2018) L. Chen, J.-H. Chung, B. Gao, T. Chen, M. B. Stone, A. I. Kolesnikov, Q. Huang,  and P. Dai, “Topological Spin Excitations in Honeycomb Ferromagnet CrI3{\mathrm{CrI}}_{3},” Phys. Rev. X 8, 041028 (2018).
  • Lee et al. (2020) I. Lee, F. G. Utermohlen, D. Weber, K. Hwang, C. Zhang, J. van Tol, J. E. Goldberger, N. Trivedi,  and P. C. Hammel, “Fundamental spin interactions underlying the magnetic anisotropy in the Kitaev ferromagnet CrI3,” Phys. Rev. Lett. 124, 017201 (2020).
  • Vishkayi et al. (2020) S. I. Vishkayi, Z. Torbatian, A. Qaiumzadeh,  and R. Asgari, “Strain and electric-field control of spin-spin interactions in monolayer CrI3,” Phys. Rev. Mater. 4, 094004 (2020).
  • Jaeschke-Ubiergo et al. (2021) R. Jaeschke-Ubiergo, E. S. Morell,  and A. S. Nunez, “Theory of magnetism in the van der Waals magnet CrI3,” Phys. Rev. B 103, 174410 (2021).
  • Chen et al. (2020) L. Chen, J.-H. Chung, T. Chen, C. Duan, A. Schneidewind, I. Radelytskyi, D. J. Voneshen, R. A. Ewings, M. B. Stone, A. I. Kolesnikov, et al., “Magnetic anisotropy in ferromagnetic CrI3{\mathrm{CrI}}_{3},” Phys. Rev. B 101, 134418 (2020).
  • Ellis et al. (2015) M. O. Ellis, R. F. Evans, T. A. Ostler, J. Barker, U. Atxitia, O. Chubykalo-Fesenko,  and R. W. Chantrell, “The Landau–Lifshitz equation in atomistic models,” Low Temp. Phys. 41, 705 (2015).
  • Giannozzi et al. (2009) P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, et al., “QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials,” J. Phys. Condens. Matter. 21, 395502 (2009).
  • Perdew et al. (1996) J. P. Perdew, K. Burke,  and M. Ernzerhof, “Generalized gradient approximation made simple,” Phys. Rev. Lett. 77, 3865 (1996).
  • Grimme (2006) S. Grimme, “Semiempirical GGA-type density functional constructed with a long-range dispersion correction,” J. Comput. Chem 27, 1787 (2006).
  • Cococcioni and De Gironcoli (2005) M. Cococcioni and S. De Gironcoli, “Linear response approach to the calculation of the effective interaction parameters in the LDA+ U method,” Phys. Rev. B 71, 035105 (2005).
  • Leon et al. (2020) A. Leon, J. González, F. D. C. de Lima, J. Mejia,  and E. S. Morell, “Strain-induced phase transition in CrI3{\mathrm{CrI}}_{3} bilayers,” 2D Mater. 7, 035008 (2020).
  • Evans et al. (2014) R. F. Evans, W. J. Fan, P. Chureemart, T. A. Ostler, M. O. Ellis,  and R. W. Chantrell, “Atomistic spin model simulations of magnetic nanomaterials,” J. Phys. Condens. Matter 26, 103202 (2014).
  • Brown (1979) W. Brown, “Thermal fluctuation of fine ferromagnetic particles,” IEEE Trans. Magn. 15, 1196 (1979).
  • García-Palacios and Lázaro (1998) J. L. García-Palacios and F. J. Lázaro, “Langevin-dynamics study of the dynamical properties of small magnetic particles,” Phys. Rev. B 58, 14937 (1998).
  • Kawaguchi et al. (2007) Y. Kawaguchi, H. Saito,  and M. Ueda, “Can spinor dipolar effects be observed in Bose-Einstein condensates?” Phys. Rev. Lett. 98, 110406 (2007).
  • Ezawa (2010) M. Ezawa, “Giant skyrmions stabilized by dipole-dipole interactions in thin ferromagnetic films,” Phys. Rev. Lett. 105, 197202 (2010).
  • Yang et al. (2015) H. Yang, A. Thiaville, S. Rohart, A. Fert,  and M. Chshiev, “Anatomy of dzyaloshinskii-moriya interaction at Co/Pt interfaces,” Phys. Rev. Lett. 115, 267210 (2015).
  • Moriya (1960) T. Moriya, “Anisotropic superexchange interaction and weak ferromagnetism,” Phys. Rev. 120, 91 (1960).
  • Tong et al. (2018) Q. Tong, F. Liu, J. Xiao,  and W. Yao, “Skyrmions in the moiré of van der waals 2d magnets,” Nano Lett. 18, 7194 (2018).
  • Khoshlahni et al. (2019) R. Khoshlahni, A. Qaiumzadeh, A. Bergman,  and A. Brataas, “Ultrafast generation and dynamics of isolated skyrmions in antiferromagnetic insulators,” Phys. Rev. B 99, 054423 (2019).
  • Behera et al. (2019) A. K. Behera, S. Chowdhury,  and S. R. Das, “Magnetic skyrmions in atomic thin CrI3{\mathrm{CrI}}_{3} monolayer,” Appl. Phys. Lett. 114, 232402 (2019).
  • Abdul-Wahab et al. (2021) D. Abdul-Wahab, E. Iacocca, R. F. Evans, A. Bedoya-Pinto, S. Parkin, K. S. Novoselov,  and E. J. Santos, “Domain wall dynamics in two-dimensional van der waals ferromagnets,” Appl. Phys. Rev. 8, 041411 (2021).
  • Allwood et al. (2005) D. A. Allwood, G. Xiong, C. Faulkner, D. Atkinson, D. Petit,  and R. Cowburn, “Magnetic domain-wall logic,” Science 309, 1688 (2005).
  • Vélez et al. (2019) S. Vélez, J. Schaab, M. S. Wörnle, M. Müller, E. Gradauskaite, P. Welter, C. Gutgsell, C. Nistor, C. L. Degen, M. Trassin, et al., “High-speed domain wall racetracks in a magnetic insulator,” Nat. Commun. 10, 1 (2019).
  • Luo et al. (2020) Z. Luo, A. Hrabec, T. P. Dao, G. Sala, S. Finizio, J. Feng, S. Mayr, J. Raabe, P. Gambardella,  and L. J. Heyderman, “Current-driven magnetic domain-wall logic,” Nature 579, 214 (2020).