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

Ab initio exploration of short-pitch skyrmion materials: Role of orbital frustration

Takuya Nomoto nomoto@ap.t.u-tokyo.ac.jp Research Center for Advanced Science and Technology, University of Tokyo, Komaba, Meguro-ku, Tokyo 153-8904, Japan    Ryotaro Arita arita@riken.jp Research Center for Advanced Science and Technology, University of Tokyo, Komaba, Meguro-ku, Tokyo 153-8904, Japan RIKEN Center for Emergent Matter Science (CEMS), Wako 351-0198, Japan
Abstract

In recent years, the skyrmion lattice phase with a short lattice constant has attracted attention due to its high skyrmion density, making it a promising option for achieving high-density storage memory and for observing novel phenomena like the quantized topological Hall effect. Unlike conventional non-centrosymmetric systems where the Dzyaloshinsky-Moriya interaction plays a crucial role, the short pitch skyrmion phase requires a quadratic magnetic interaction J(𝒒)J(\bm{q}) with a peak at finite-𝑸\bm{Q}, and weak easy-axis magnetic anisotropy is also critical. Thus, conducting first-principles evaluations is essential for understanding the formation mechanism as well as for promoting the discovery of new skyrmion materials. In this Perspective, we focus on recent developments of the first-principles evaluations of these properties and apply them to the prototype systems GdT2X2T_{2}X_{2} and EuT2X2T_{2}X_{2}, where TT denotes a transition metal and XX represents Si or Ge. In particular, based on the spin density functional theory with the Hubbard correction combined with the Liechtenstein method in the Wannier tight-binding model formalism, we first show that the Hubbard UU and Hund’s coupling is essential to stabilize a skyrmion lattice state by enhancing the easy-axis anisotropy. We then discuss mechanisms of finite-𝑸{\bm{Q}} instability and show that competition among Gd-5dd orbitals determines whether ferromagnetism or a finite-𝑸{\bm{Q}} structure is favored in GdT2T_{2}Si2 with T=T= Fe and Ru. Our systematic calculations reveal that GdRu2X2X_{2}, GdOs2X2X_{2}, and GdReX22{}_{2}X_{2} are promising, while GdAgX22{}_{2}X_{2}, GdAuX22{}_{2}X_{2}, and EuAgX22{}_{2}X_{2} are possible candidates as the skyrmion host materials. Analysis based on a spin spiral calculation for the candidate materials is also presented.

preprint: paper

I Introduction

A magnetic skyrmion is a particle-like topological spin texture realized in condensed matter systems. Its magnetization distribution is characterized by vorticity, helicity, and a winding number, and the coupling to an itinerant electron provides a unique platform for exploring the physics under the emergent gauge field [1, 2, 3]. The skyrmion is also considered a candidate element in future magnetoelectric devices due to its stability against an external perturbation and high controllability by an electric current [4, 5, 6, 7, 8]. Particularly, a racetrack memory [9], a magnetic random access memory [10, 11], and an artificial synapse for neuromorphic computing [12, 13] are promising applications of the topologically stable magnetic object.

The magnetic skyrmions form a two-dimensional lattice in a ferromagnetic background. While an array of magnetic bubbles or skyrmions has been observed in a magnetic thin film since the 1970s [14, 15], the first observation in a bulk magnet was accomplished in 2009 by the neutron scattering measurement for MnSi with B20 crystal structure [16]. In the chiral magnet, an early theory suggested that a Dzyaloshinsky-Moriya (DM) interaction is the key to realizing the skyrmion lattice phase [17, 18]. Namely, the competition between ferromagnetic and DM interaction induces a non-collinear spin spiral structure, whose modulation pitch is determined by A/DA/D with AA being magnetic stiffness and DD being a DM constant. Once the spiral state is stabilized, a triple-𝑸\bm{Q} state, which is identical to the triangular skyrmion lattice, is expected to appear under a magnetic field due to the momentum conservation of the quartic term in the Ginzburg-Landau functional. Today, there are many materials showing the DM-induced skyrmion phase in bulk, thin film, and heterostructures. In these systems, a typical modulation pitch is around 10-100nm, much larger than the lattice constant since DAD\ll A is satisfied in most cases [19, 20, 21, 22, 23].

On the other hand, the magnetic skyrmion observed in the centrosymmetric crystals has recently attracted much attention, where the DM interaction should not play any role. The first observation was reported in Gd2PdSi3 with Gd forming a two-dimensional triangular lattice network [24, 25], followed by Gd3Ru4Al12 with a Kagome lattice [26], GdRu2Si2 [27], EuAl4 [28, 29, 30, 31], and EuGa4 [30, 32] with a square lattice network. A remarkable feature of the skyrmion lattice in centrosymmetric systems is its relatively shorter modulation pitch, typically given by a few nanometers. Since this does not originate from the DM interaction, its formation mechanism has been intensively studied recently. In terms of the shortness of the modulation pitch, the skyrmion lattices observed in EuPtSi [33, 34] and Mn1-xFexGe [35, 36] might be classified into the same category as the centrosymmetric systems even though their crystal structures do not have the space inversion symmetry. Since a short-pitch skyrmion lattice is favorable for realizing high-density memory bits in a storage device, revealing the key quantity to determine the modulation pitch will promote further materials design and engineering of the skyrmion phase. Furthermore, the high skyrmion density of the short-pitch skyrmion is of great theoretical importance, as it provides a unique platform to observe fascinating phenomena, such as a quantized topological Hall effect [37] and a crossover behavior of the topological Hall effect where both the real-space and momentum-space Berry curvature play a crucial role [38].

To date, there have been a lot of theoretical works on the formation of a short-pitch skyrmion lattice. They include calculations based on a classical spin model, an itinerant electron model, and an effective Ginzburg-Landau functional. These theories enable us to grasp a physical insight into the skyrmion lattice formation and, indeed, the importance of the magnetic frustrations [39, 40], compass anisotropy of the magnetic interactions [41, 42], higher-order spin interactions [43, 44], and Fermi surface nesting [45, 46] have been recognized as an essential property to stabilize the skyrmion lattice phase (for more details, see Ref. [47]). However, these model calculations are usually performed with a given modulation vector and include adjustable parameters. Thus, they have no predictive power for the modulation pitch itself in real materials. On the other hand, in spite of the expensive numerical cost, one can investigate the stability of finite-𝑸\bm{Q} spin spiral structures from first-principles, which is helpful to see the origin of the modulation pitch and its material dependence. Indeed, the calculations for GdRu2Si2 [48, 49], Gd2PdSi3 [48, 49], and MnGe [50, 51] have successfully unveiled microscopic mechanisms in stabilizing their short-pitch skyrmion phase. Thus, broadening the target compounds along this line is of great importance, helping us understand the material dependence more deeply and promoting the discovery of new skyrmion materials.

Based on the theory for short-pitch skrymion phase, a promising system for observing skyrmion phases should possess weak easy-axis anisotropy and spin-spin interactions favoring finite-𝑸\bm{Q} modulations [47]. Since these effects on skyrmion formation are well-established, in this Perspective, we aim to highlight recent progress in first-principles evaluations of these properties and offer predictive insights to guide experimental efforts in discovering new skyrmion materials. To achieve this, based on spin density functional theory (SDFT) and SDFT with the Hubbard correction UU (SDFT+UU), we perform systematic first-principles calculations for GdT2X2T_{2}X_{2} and EuT2X2T_{2}X_{2} with TT being a transition metal element and XX being Si or Ge. Since the largest number of skyrmion compounds have been found in this ThCr2Si2-type crystal structure, these compounds would be ideal target materials. First, we show that the inclusion of UU and Hund’s coupling JJ is essential to obtain easy-axis anisotropy, which is necessary for stabilizing a skyrmion lattice than a spin spiral state. Then, we evaluate magnetic interactions J(𝒒)J(\bm{q}) based on the so-called Liechtenstein method with the Wannier tight-binding model formalism. The calculated J(𝒒)J(\bm{q}) is in good agreement with E(𝒒)E(\bm{q}) directly evaluated by the spin spiral calculations within SDFT+UU, supporting the validity of the present Liechtenstein calculations. Based on the results, we argue that the finite-𝑸{\bm{Q}} structure is determined not only by the shape of the Fermi surface but also by the details of the electronic structure. Especially, we show that competition among the Gd-5dd orbital contributions determines whether a ferromagnetism or finite-𝑸{\bm{Q}} structure is favored in GdT2T_{2}Si2 with T=T= Fe and Ru. According to our systematic calculations, GdRu2X2X_{2}, GdOs2X2X_{2}, GdWX22{}_{2}X_{2}, GdReX22{}_{2}X_{2}, GdAgX22{}_{2}X_{2}, GdAuX22{}_{2}X_{2}, EuCoX22{}_{2}X_{2}, and EuAgX22{}_{2}X_{2} with X=X= Si and Ge are possible candidates for showing the skyrmion phase. The present study provides a firm ground to discover and design the short-pitch skyrmion phase from systematic calculations, and the extension to the other crystal structures will be a promising future development.

II Methods

In this section, we summarize the methods used in this Perspective. First, we explain the evaluation of J(𝒒)J(\bm{q}) based on the Liechtenstein method. Although this method was first formulated in the multiple scattering theory with the Green’s function techniques, it has been applied to the SDFT Hamiltonian with various spatially localized bases, including the Wannier functions [52, 53]. Here, we show the formulation following Ref. [53]. We begin with the following tight-binding Hamiltonian,

H=12(t12+v12)c1c2,\displaystyle H=\sum_{12}(t_{12}+v_{12})c_{1}^{\dagger}c_{2}, (1)

where indices 1 and 2 run over all degrees of freedom that specify the Wannier functions, including atomic sites ii, orbitals \ell, and spins σ\sigma. c1c_{1}^{\dagger} (c1c_{1}) represents an electron creation (annihilation) operator in this basis. t12t_{12} and v12v_{12} denote spin-independent and spin-dependent hopping integral matrices, respectively. These parameters are extracted from the SDFT/SDFT+UU calculation with ferromagnetic reference states through the Wannier construction process [54, 55].

Based on the Hamiltonian (1), we can evaluate the magnetic interaction JijJ_{ij} in the classical Heisenberg model as follows,

Jij=12Trnσ[Gij(ωn)ΣiGji(iωn)Σj].\displaystyle J_{ij}=-\frac{1}{2}{\rm Tr}_{n\ell\sigma}[G_{ij}(\omega_{n})\Sigma^{i}G_{ji}(i\omega_{n})\Sigma^{j}]. (2)

Here, Trnσ=TnTrσ{\rm Tr}_{n\ell\sigma}=T\sum_{n}{\rm Tr}_{\ell\sigma}, where Trσ{\rm Tr}_{\ell\sigma} denotes the trace over the orbital and spin spaces. ωn=πT(2n+1)\omega_{n}=\pi T(2n+1) denotes the Matsubara frequency, and TT is the temperature. The Green’s function G(ωn)G(\omega_{n}) and the magnetic perturbation matrix at ii site Σi\Sigma^{i} are defined by,

[G(ωn)]121\displaystyle[G(\omega_{n})]^{-1}_{12} =iωnδ12t12v12\displaystyle=i\omega_{n}\delta_{12}-t_{12}-v_{12} (3)
Σ12i\displaystyle\Sigma_{12}^{i} =v~i1,i2σσ1σ2x\displaystyle=\tilde{v}_{i\ell_{1},i\ell_{2}}\sigma^{x}_{\sigma_{1}\sigma_{2}} (4)

where we approximate Σ12i\Sigma^{i}_{12} as a local quantity. σμ\sigma^{\mu} (μ=x,y,z\mu=x,y,z) is the Pauli matrix, and v~i11,i22\tilde{v}_{i_{1}\ell_{1},i_{2}\ell_{2}} is defined by vi11σ1,i22σ2=v~i11,i2iσσ1σ2zv_{i_{1}\ell_{1}\sigma_{1},i_{2}\ell_{2}\sigma_{2}}=\tilde{v}_{i_{1}\ell_{1},i_{2}\ell_{i}}\sigma^{z}_{\sigma_{1}\sigma_{2}}. Then, Gij(ωn)G_{ij}(\omega_{n}) is defined as a sub-matrix of G12(ωn)G_{12}(\omega_{n}) having (i,j)(i,j) sites component. The evaluation of the Matsubara summation in Eq. (2) is performed by the sparse sampling technique with the intermediate representation [56, 57]. More details about the implementation are found in Ref. [53]. Based on JijJ_{ij} evaluated by Eq. (2), its Fourier transform J(𝒒)J(\bm{q}) is obtained as follows:

J(𝒒)=1Nijei𝒒(𝑹i𝑹j)JijJ0,\displaystyle J(\bm{q})=\frac{1}{N}\sum_{ij}e^{-i\bm{q}\cdot(\bm{R}_{i}-\bm{R}_{j})}J_{ij}-J_{0}, (5)

where 𝑹i\bm{R}_{i} denotes the real space coordinate of the site ii including the relative coordinate of the sublattice. Here, we have introduced J0J_{0}, the on-site magnetic interaction, to guarantee the absence of the self-interaction term in the Heisenberg model. Note that, although JijJ_{ij} evaluated by Eq. (2) will be finite for non-magnetic sites such as a transition metal TT and X=X= Si, Ge in GdT2X2T_{2}X_{2} and EuT2X2T_{2}X_{2}, the i,ji,j summation in Eq. (5) should be taken only for the magnetic site, namely, Gd or Eu site in our cases. Indeed, JijJ_{ij} connecting non-magnetic elements can take a large value when both the density of states (DOS) and exchange splitting near the Fermi level are finite, although the mapping into the classical spin model for these orbitals is totally inadequate. This procedure can remove this artificial contribution in the evaluation of J(𝒒)J(\bm{q}), which is essential for the SDFT+UU calculations. Based on the obtained J(𝒒)J(\bm{q}), we can see whether the system favors the finite-𝑸\bm{Q} modulation.

On the other hand, we can evaluate the energy of the spin spiral states with the modulation vector 𝑸\bm{Q}, denoted by E(𝒒)E(\bm{q}), within SDFT and SDFT+UU. The calculation is based on the generalized Bloch theorem, in which the energy of the spin spiral state is evaluated with specific twisted boundary conditions for the Bloch spinors [58]. A clear advantage of this method is that the calculated E(𝒒)E(\bm{q}) is exact within the SDFT/SDFT+UU level, and no assumption of the mapping to the classical spin model is introduced. Moreover, it is not necessary to expand the unit cell even when calculating the finite-𝑸\bm{Q} structure, which makes the calculation with small 𝑸\bm{Q} tractable. However, despite these advantages, it requires a calculation for every 𝑸\bm{Q}, and thus, is much more expensive than the Liechtenstein method if one wants to see the detailed structures of E(𝒒)E(\bm{q}). Thus, we use these two methods in a complementary manner.

III Calculation details

For the evaluations of J(𝒒)J(\bm{q}) and E(𝒒)E(\bm{q}), we use Vienna ab initio simulation package (VASP) [59] for the electronic structure calculations. Here, the exchange-correlation functional proposed by Perdew, Burke, and Ernzerhof [60], and pseudopotentials with the projector augmented wave (PAW) basis [61, 62] are used. The spin orbit interaction is neglected except for the magnetocrystalline anisotropy calculations.

The structure optimization is performed for all GdT2X2T_{2}X_{2} and EuT2X2T_{2}X_{2} compounds by the SDFT+UU method assuming ferromagnetic states. The convergence criteria of the structure optimization is set to 10510^{-5} eV and the corresponding electronic self-consistent loop is 10610^{-6} eV with the accurate precision mode. Here, we use a 16×16×1616\times 16\times 16 Monkhorst-Pack 𝒌\bm{k}-mesh, and a cutoff of the plane wave basis set is set to be twice the recommended maximum value of the cutoff. For the magnetocrystalline anisotropy calculations, we employ 10810^{-8} eV as the convergence criteria for electronic calculations and 21×21×2121\times 21\times 21 Monkhorst-Pack 𝒌\bm{k}-mesh. The cutoff is set to be twice the recommended value. For the energy calculations of the spin spiral states, we employ 10810^{-8} eV as the convergence criteria for electronic calculations, 16×16×1616\times 16\times 16 Monkhorst-Pack 𝒌\bm{k}-mesh, and the cutoff is 2.5 times as large as the recommended values.

For constructing the Wannier tight-binding model, we use 156 Bloch states evaluated on 10×10×1010\times 10\times 10 uniform 𝒌\bm{k}-grid in the disentanglement procedure. The trial orbitals are set to the Gd/Eu-dd and ff orbitals, transition metal TT-ss, pp, and dd orbitals, and XX = Si/Ge-ss and pp orbitals. The outer window is set as it includes all 156 Bloch states, and the inner window is set as it covers from the bottom of the bands to the bands at 4 eV above the Fermi level. The obtained tight-binding model consists of 38 orbitals per spin. Based on the obtained tight-binding model, JijJ_{ij} and J(𝒒)J(\bm{q}) are evaluated by Eqs. (2) and  (5), respectively. Here, we set the temperature T=58T=58 K and employ 48×48×4848\times 48\times 48 uniform 𝒌\bm{k}-grid.

IV Results and Discussion

IV.1 Magnetocrystalline anisotropy

Refer to caption
Figure 1: Magnetocrystalline anisotropy of GdRu2Si2 and GdRu2Ge2. The label “without UU” stands for the result of the standard SDFT calculation. The others correspond to SDFT+UU calculations with U=6.7U=6.7 eV.

According to the theory of skyrmion lattice formation [47], the system should possess two properties to realize a skyrmion lattice phase in a broad region of parameter space. Namely, it is necessary that the spin interaction is compatible with a finite-QQ modulation, and the magnetocrystalline anisotropy is the easy axis type as well as sufficiently weak. It should be noted that, with the exception of Gd3+ and Eu2+, rare earth elements possess a finite orbital moment with strong spin-orbit coupling, leading to strong magnetocrystalline anisotropy in most cases. As a result, compounds containing these elements are unlikely to exhibit the weak magnetocrystalline anisotropy required for the formation of skyrmion lattice phases. Therefore, in this Perspective, we only consider Gd3+- and Eu2+-based compounds, where the orbital angular momentum of their 4ff-orbitals are almost quenched in the ionic ground states. On the other hand, there is no general trend for these elements, whether the anisotropy is the easy axis or easy plane type. Thus, it is crucial to see whether the system shows the correct magnetic anisotropy before discussing the modulation vectors based on the electronic structure.

Refer to caption
Figure 2: Magnetocrystalline anisotropy of (a) GdT2T_{2}Si2, (b) GdT2T_{2}Ge2, (c) EuT2T_{2}Si2 and (d) EuT2T_{2}Ge2. The SDFT+UU results stand for the calculations with U=6.7U=6.7 eV and J=0.7J=0.7 eV. Note that the data for EuMn2Si2, EuZr2Si2, EuZr2Ge2, EuTc2Si2, EuTc2Ge2, EuHf2Ge2, EuTa2Si2 and EuRe2Ge2 in SDFT, and GdCr2Ge2 in SDFT+UU are not shown. This is because the fully polarized ferromagnetic solutions along the xx and/or zz axis are unstable for these materials.

As an example, let us first see in Fig. 1 the calculated magnetocrystalline anisotropy,

ΔE:=E(𝒎//𝒙)E(𝒎//𝒛)\displaystyle\Delta E:=E(\bm{m}//\bm{x})-E(\bm{m}//\bm{z}) (6)

for the skyrmion compound GdRu2Si2 and a reference compound GdRu2Ge2 [63]. We can see that the energy scale of the magnetic anisotropy is quite small, |ΔE|<5|\Delta E|<5 meV per formula unit, as expected. However, the sign of ΔE\Delta E is negative in the SDFT results, indicating that a helical magnetic structure is expected to have a lower energy than the skyrmion state. This situation drastically changes by including the Hubbard correction UU. In the case of Gd3+, since UU enhances the energy difference between the lowest energy state with L=0L=0 and the first excited state with L0L\neq 0, we can expect that the magnetocrystalline anisotropy decreases by increasing the value of UU, which is consistent with the results in Fig. 1. Notably, ΔE\Delta E becomes small but positive with U=6.7U=6.7 eV even without Hund’s coupling correction JJ. In addition, we can also see that the anisotropy becomes larger positive with increasing the strength of JJ, which is compatible with the observed skyrmion lattice phases. This result clearly indicates that the inclusion of UU and JJ is essential to obtain the correct magnetocrystalline anisotropy in the Gd and Eu-based skyrmion materials.

Figures 2(a-d) summarize the results of ΔE\Delta E in all Gd and Eu-based 122 compounds that we discuss for this Perspective. We can see that, for most transition metal TT except for Ta, the SDFT+UU results show the same or lower |ΔE||\Delta E| than SDFT, which is similar to the cases of GdRu2Si2 and GdRu2Ge2. Although the values of ΔE\Delta E highly depend on the host ff-element and Eu-based compounds tend to have considerable negative ΔE\Delta E in SDFT, these trends are strongly suppressed in SDFT+UU. In SDFT+UU, ΔE\Delta E is no longer sensitive to the transition metal TT and the group 14 element X=X= Si and Ge. Also, the amplitude |ΔE||\Delta E| is sufficiently small and |ΔE|<3|\Delta E|<3 meV/f.u. is satisfied in all compounds. It is worth noting that only one (seven) in GdT2X2T_{2}X_{2} (EuT2X2T_{2}X_{2}) among all 48 compounds considered show negative ΔE\Delta E, implying that the weak easy-axis anisotropy is an intrinsic property inherent in this ThCr2Si2-type crystal structure.

IV.2 Fermi surface and magnetic instability

Next, let us move on to the modulation vector of GdT2X2T_{2}X_{2} and EuT2X2T_{2}X_{2}. According to the previous discussion, it is essential to include the local Coulomb repulsion UU and Hund’s coupling JJ to reproduce the correct magnetocrystalline anisotropy observed in the experiments. Thus, in the present study, we discuss the modulation vector 𝑸\bm{Q} based on the electronic structures in the SDFT+UU calculations. Apparently, the effect of UU and JJ on 𝑸\bm{Q} should not be significant if 𝑸\bm{Q} is determined only by the shape of the Fermi surface, which is often supposed in the RKKY scenario for the skyrmion lattice formation. Indeed, since the magnetic moments of Gd3+ and Eu2+ are fully polarized even in the absence of UU and JJ, the Fermi surface is expected to be insensitive to the values of UU and JJ. We can directly confirm that magnetic moments are fully polarized in the ferromagnetic states in most compounds except for EuMn2Si2, EuZr2Si2, EuZr2Ge2, EuTc2Si2, EuTc2Ge2, EuHf2Ge2, EuTa2Si2 and EuRe2Ge2. Thus, although the electronic structure above the Fermi energy should be modified to some extent because of the change of the exchange splitting, this will not affect the modulation vector 𝑸\bm{Q} in the skyrmion phase.

Refer to caption
Figure 3: Band structures of (a) majority spin of GdRu2Si2, (b) minority spin of GdRu2Si2, (c) majority spin of GdFe2Si2 and (d) minority spin of GdFe2Si2. Solid lines correspond to SDFT and dashed lines correspond to SDFT+UU results with U=6.7U=6.7 eV and J=0.7J=0.7 eV.

The insensitivity of the Fermi surface against UU and JJ can be seen directly by the band structure calculations. For example, we show the band structures near the Fermi levels of GdRu2Si2 and GdFe2Si2 in Figs. 3(a-d). Here, the majority and minority bands calculated with SDFT and SDFT+UU are plotted, respectively. We can see from the figures that, although they differ in detail especially above the Fermi levels, UU and JJ do not change the Fermi surface, which means that SDFT and SDFT+UU results share the same Fermi surface instability. Note that this trend is valid for all compounds when SDFT correctly captures the ionic state of Gd3+ and Eu2+.

In the continuum model, it is well-known that the Lindhard function describes the Fermi surface instability. In a real material with many bands, the nesting function,

χnn(𝒒)=1N𝒌f(εn𝒌+𝒒)f(εn𝒌)εn𝒌+𝒒εn𝒌,\displaystyle\chi_{nn^{\prime}}({\bm{q}})=-\frac{1}{N}\sum_{\bm{k}}\frac{f(\varepsilon_{n{\bm{k}}+{\bm{q}}})-f(\varepsilon_{n^{\prime}{\bm{k}}})}{\varepsilon_{n{\bm{k}}+{\bm{q}}}-\varepsilon_{n^{\prime}{\bm{k}}}}, (7)

plays the same role. Here, nn and 𝒌\bm{k} denote the band index and crystal momentum, respectively. εn𝒌\varepsilon_{n{\bm{k}}} is the eigenvalues of the Bloch states having nn and 𝒌\bm{k}, and f(ε)f(\varepsilon) is the Fermi distribution function. The nesting function is sometimes considered to describe not only the Fermi surface instability but also the spin instability, such as spin-density-wave and helical 𝑸\bm{Q} magnetic structures. However, as is known in the fluctuation theory for multi-orbital systems, the orbital character of the bands is essential to describe the fluctuation. In this case, it is necessary to consider the (zeroth-order) spin correlation function χμν(𝒒)\chi_{\mu\nu}(\bm{q}) defined as follows,

χμν(𝒒)\displaystyle\chi_{\mu\nu}({\bm{q}}) =1N𝑹1234eiϕ13𝒒(𝑹)s13μs24νχ13,24(𝑹),\displaystyle=\frac{1}{N}\sum_{\bm{R}}\sum_{1234}e^{i\phi_{13}^{\bm{q}}(\bm{R})}s^{\mu}_{13}s^{\nu}_{24}\chi_{13,24}(\bm{R}), (8)
χ13,24(𝑹)\displaystyle\chi_{13,24}(\bm{R}) =TωnG12(iωn,𝑹)G43(iωn,𝑹).\displaystyle=-T\sum_{\omega_{n}}G_{12}(i\omega_{n},\bm{R})G_{43}(i\omega_{n},-\bm{R}). (9)

Here, the phase ϕ12𝒒(𝑹)\phi_{12}^{\bm{q}}(\bm{R}) is defined by ϕ12𝒒(𝑹)=𝒒(𝑹i3𝑹i1)\phi_{12}^{\bm{q}}(\bm{R})=\bm{q}\cdot(\bm{R}_{i_{3}}-\bm{R}_{i_{1}}), and ωn\omega_{n} and ωq\omega_{q} denote the Fermionic and Bosonic Matsubara frequency, respectively. Since Eq. (9) leads to the nesting function with a modification due to the transformation from local orbital to band basis, the 𝒒\bm{q}-dependence of χμν(𝒒)\chi_{\mu\nu}({\bm{q}}) is mainly determined by Eq. (7). However, it should be noted that the basis transformation gives an additional source of the 𝒒\bm{q}-dependence, and χμν(𝒒)\chi_{\mu\nu}({\bm{q}}) is sensitive not only to the shape of the Fermi surface but also to its orbital character. Moreover, in the strongly correlated electron systems, Eqs. (8) and (9) are not sufficient to describe the correct spin fluctuation. In this case, the effect of UU must be taken more accurately using a diagrammatic technique, which is practically impossible in most cases. In that sense, the Liechtenstein method presented provides a simple but reliable approximation in taking the correlation effects. Indeed, it is known that this method successfully reproduces the result of t/Ut/U expansion in the strong correlation limit and reproduces Eqs. (8) and (9) in the weak correlation limit. It is worth noting that in the intermediate and strong UU regime, not only the Fermi surface but also the electronic structure far from the Fermi level can affect the spin instability.

IV.3 J(𝒒)J(\bm{q}) in GdRu2Si2 and GdFe2Si2 evaluated by the Liechtenstein method

Refer to caption
Figure 4: Modulation vector 𝑸\bm{Q} dependence of the spin exchange interaction J(𝒒)J(\bm{q}) in (a) GdRu2Si2 and (b) GdFe2Si2 calculated based on the Liechtenstein formula. A pink (cyan) line with open (filled) circle corresponds to the SDFT (SDFT+UU with U=6.7U=6.7 eV and J=0.7J=0.7 eV) result.

Figures 4(a) and (b) show J(𝒒)J(\bm{q}) for GdRu2Si2 and GdFe2Si2 calculated based on the Liechtenstein formula, Eq. (5). The result in Fig. 4(a) shows good agreement with that in Ref. [49] but is slightly different from that in Ref. [48] due to an erroneous treatment of the phase factor in Eq. (8). According to the results, since the peak position 𝑸\bm{Q} in J(𝒒)J(\bm{q}) is the signature of magnetic instability with the modulation vector 𝑸\bm{Q}, we can say that GdRu2Si2 favors the spin spiral state with 𝑸(0.25,0,0)\bm{Q}\sim(0.25,0,0) both in SDFT and SDFT+UU. On the other hand, in GdFe2Si2, finite-𝑸\bm{Q} instability is seen only in SDFT while SDFT and SDFT+UU give the same Fermi surface (see Figs. 3(c) and (d)). This result indicates that the finite-𝑸\bm{Q} structure is determined not only by the shape of the Fermi surface but also by other factors, such as an orbital character and electronic structure far from the Fermi level.

The ferromagnetic instability observed in the SDFT+UU calculation for GdFe2Si2 is also found in the spin spiral calculation (which will be discussed in Section IV.6). However, it seems to be inconsistent with the observed spin flop transition in GdFe2Si2 [64], which implies that we may need to treat the correlation effect more precisely to understand the magnetisn in GdFe2Si2. On the other hand, GdRu2Si2 has a peak at 𝑸(0.25,0,0)\bm{Q}\sim(0.25,0,0) both in SDFT+UU and SDFT calculation, similar to the previous results based on SDFT (without UU[48, 49], which is consistent with the experiment.

Note that, in principle, it is necessary to perform a spin model simulation to find out a correct magnetic ground state, especially for multiple-𝑸\bm{Q} states like a skyrmion lattice phase. For a spin model of GdRu2Si2, an analysis based on the Landau-Lifshitz-Gilbert (LLG) equation was performed in Ref. [49], using the following model for the internal energy,

E=12ijJij𝒎i𝒎jKi(𝒎i𝒆z)2𝑩i𝒎i,\displaystyle E=-\frac{1}{2}\sum_{ij}J_{ij}{\bm{m}}_{i}\cdot{\bm{m}}_{j}-K\sum_{i}(\bm{m}_{i}\cdot\bm{e}_{z})^{2}-\bm{B}\cdot\sum_{i}\bm{m}_{i}, (10)

where KK is a magnetocrystalline anisotropy constant, and 𝑩\bm{B} is a external magnetic field applied along the cc axis. They solved the LLG equations for Eq. (10) with JijJ_{ij} evaluated from first-principles, which is consistent with Fig. 4(a), and obtained a magnetic phase diagram in terms of KK and BB, as shown in Fig. 5. According to Fig. 5, the skyrmion lattice phase appears when the easy-axis anisotropy KK is in the range from 0.2 to 0.4 meV. On the other hand, as discussed in the previous sections, the SDFT+UU calculation reproduces the easy-axis anisotropy of GdRu2Si2. The value of KK in GdRu2Si2 is 0.43\sim 0.43 meV, which is in good agreement with the spin model simulation Note that the significance of weak easy-axis anisotropy can be straightforwardly understood: if the anisotropy is easy-plane type, a helical single-𝑸\bm{Q} state is favored over multiple-𝑸\bm{Q} states like a skyrmion phase, and if the easy-axis anisotropy is too strong, Ising-type magnetic order is favored over finite-𝑸\bm{Q} structures. Thus, we conclude that the SDFT+UU result in GdRu2Si2 is fully consistent with the observation of the skyrmion lattice phase, and the present method correctly captures the essential physics behind the magnetic structure.

Refer to caption
Figure 5: (a) A magnetic phase diagram of GdRu2Si2 in the (K,B)(K,B) parameter space, where KK is the magnetocrystalline anisotropy and BB is the magnetic field. (b)-(d) Representative magnetic textures described by the Gd moments. (b) Helical state for (K,B)=(0.2,0.6)(K,B)=(0.2,0.6), (c) Skyrmion state for (K,B)=(0.2,1.0)(K,B)=(0.2,1.0), and (d) cycloidal state for (K,B)=(0.1,1.0)(K,B)=(0.1,1.0). Reproduced from Ref. [49].

IV.4 Orbital decomposition of J(𝒒)J(\bm{q}) calculated based on the Liechtenstein method

Refer to caption
Figure 6: Orbital decomposed J(𝒒)J(\bm{q}) calculated by the Liechtenstein method. (a) The contributions from the Gd-5dd and 4ff orbitals. (b) The diagonal contribution to in the Gd-5dd manifold. Here, the indices 1 to 5 correspond to the dz2,dxz,dyz,dx2y2d_{z^{2}},d_{xz},d_{yz},d_{x^{2}-y^{2}}, and dxyd_{xy} orbitals, respectively. (c) The off-diagonal contributions in the Gd-5dd manifold.

One of the advantages of evaluating J(𝒒)J(\bm{q}) based on the tight-binding formalism is that it is easy to discuss the microscopic origin of the magnetic interaction [65, 48]. Although J(𝒒)J(\bm{q}) was decomposed into the Gd-4ff and 5dd components and a possible frustration mechanism between these two was discussed in Ref. [48], one may think that the situation will be different in a real material since their calculations were based on the SDFT electronic structures. Indeed, with the SDFT+UU results, we can easily show that the contribution from Gd-4ff orbital is strongly suppressed and never affect the modulation vector of the helical states, as shown in Fig. 6(a). Naively, this indicates the absence of conventional RKKY interactions in this compound since it is included in the Gd-4ff contribution as discussed in Ref. [48]. On the other hand, as seen in the previous subsections, the orbital character and the electronic structures not at the Fermi level are essential factors in determining the finite-𝑸\bm{Q} instability. Since the 𝒒\bm{q} dependence of J(𝒒)J(\bm{q}) mainly comes from the Gd-5dd manifold according to Fig. 6(a), it would be interesting to decompose J(𝒒)J(\bm{q}) into each 5dd orbital based on the SDFT+UU electronic structures.

Refer to caption
Figure 7: Orbital decomposed J(𝒒)J(\bm{q}) calculated by the Liechtenstein method. The total (dotted line with open square), diagonal dx2x2d_{x^{2}-x^{2}} orbital (solid line with filled square), and the other contribution (dotted like with filled circle) are shown. The pink and cyan lines correspond to the results for GdRu2Si2 and GdFe2Si2, respectively.

Figure 6(b) and (c) show the diagonal and off-diagonal contributions to J(𝒒)J(\bm{q}) in the Gd-5dd manifold, respectively. Among the five 5dd orbitals, the contributions from the dz2d_{z^{2}} and dxyd_{xy} orbitals are negligibly small, and the 𝒒\bm{q} dependence of J(𝒒)J(\bm{q}) originates from the remaining three orbitals. Among the three, we can see that only the diagonal contribution from the dx2y2d_{x^{2}-y^{2}} orbital has a peak structure at the Γ\Gamma point, favoring the ferromagnetic ground state. In contrast, the other diagonal and off-diagonal ones show the peak at 𝑸(0.25,0,0)\bm{Q}\sim(0.25,0,0) or almost flat along the Γ\Gamma-X line. These results indicate that the orbital frustration exists in the 5dd orbital manifold, and whether the system favors the ferromagnetic or finite-𝑸\bm{Q} spin spiral state depends on the competition between the contributions from dx2y2d_{x^{2}-y^{2}} and the other orbitals. Notably, this situation is realized not only in GdRu2Si2 but also in GdFe2Si2 as is shown in Fig. 7. The figure shows why these two compounds favor different magnetic structures though they share almost the same Fermi surface. Namely, the contribution from the dx2y2d_{x^{2}-y^{2}} orbital relative to that from the other dd orbitals is larger in GdFe2Si2 than GdRu2Si2. This is mainly due to the suppression of the finite-𝑸\bm{Q} peak in the contribution from the other dd orbitals in GdFe2Si2.

Here, let us look into the detail of the microscopic origin of J(𝒒)J(\bm{q}) based on the following simple model Hamiltonian H=Hd+Hf+HdfH=H_{d}+H_{f}+H_{df}, where,

Hd\displaystyle H_{d} =tdi,jσdiσdjσ+Udinidnid,\displaystyle=t_{d}\sum_{\braket{i,j}\sigma}d^{\dagger}_{i\sigma}d_{j\sigma}+U_{d}\sum_{i}n_{i\uparrow}^{d}n_{i\downarrow}^{d}, (11)
Hf\displaystyle H_{f} =tfi,jσfiσfjσ+Ufinifnif,\displaystyle=t_{f}\sum_{\braket{i,j}\sigma}f^{\dagger}_{i\sigma}f_{j\sigma}+U_{f}\sum_{i}n_{i\uparrow}^{f}n_{i\downarrow}^{f}, (12)
Hdf\displaystyle H_{df} =Viσ(diσfiσ+fiσdiσ)+Jcfi𝒔d𝒔f.\displaystyle=V\sum_{i\sigma}(d^{\dagger}_{i\sigma}f_{i\sigma}+f^{\dagger}_{i\sigma}d_{i\sigma})+J_{cf}\sum_{i}{\bm{s}}^{d}\cdot{\bm{s}}^{f}. (13)

Here, Hd,HfH_{d},H_{f} and HdfH_{df} represent terms for the Gd-5dd orbitals, 4ff orbitals and their hybridization terms, respectively. Our calculations for GdT2X2T_{2}X_{2} show that, in the presence of UU, the 5dd contribution always overcomes that of 4ff since the latter is strongly suppressed. This clearly indicates that both the super-exchange interaction Jexftf2/UfJ_{\rm ex}^{f}\sim t_{f}^{2}/U_{f} and the conventional RKKY interaction JRKKY(V2/Uf)2χdJ_{\rm RKKY}\sim(V^{2}/U_{f})^{2}\chi^{d}, where χd\chi^{d} is the spin susceptibility in the 5dd orbitals, are negligibly small in these compounds. Conversely, the strong magnetic interaction from the 5dd orbitals can be easily understood since they have large DOS near the Fermi level, and have a finite exchange splitting due to the magnetic ordering although it is not as large as that of the 4ff orbitals. Here, we note that there are two possibilities for the origin of the exchange splitting in the 5dd states. One is the Coulomb interaction inherent in the 5dd orbitals, i.e., the second term in eq. (11), in which case the spin instability is described purely by the 5dd orbitals. The other is the magnetic coupling between the 5dd and 4ff orbitals, i.e., the second term in eq. (13), which one may call ”generalized” RKKY interaction since it comes from the coupling between conduction bands and local spins [66]. It should be noted that in SDFT+UU, the correlation effect is included in the exchange-correlation functional, and the many-body problem turns into an effective one-body calculation. Then, we can not distinguish these two contributions from the resulting exchange splittings. Investigating the microscopic origin of J(𝒒)J(\bm{q}) more precisely requires more elaborated calculations, such as a calculation combined with dynamical mean-field approximation [67]. This direction of development is an interesting and important future task in this field.

Refer to caption
Figure 8: Spin interactions J(𝒒)J(\bm{q}) in GdT2T_{2}X2X_{2} calculated by the Liechtenstein method based on the SDFT+UU electronic structures. The pink, green, and cyan lines with square, circle, and triangle points correspond to 3dd, 4dd, and 5dd elements as TT, respectively. The solid (dashed) lines with filled (open) symbols stand for X=X= Si (Ge). Note that the high symmetry points Z, Γ\Gamma, and X correspond to (0,0,2πc),(0,0,0)(0,0,\frac{2\pi}{c}),(0,0,0) and (2πa,0,0)(\frac{2\pi}{a},0,0) in the cartesian frame of the reciprocal lattice space, respectively.
Refer to caption
Figure 9: Spin interactions J(𝒒)J(\bm{q}) in EuT2T_{2}X2X_{2} calculated by the Liechtenstein method based on the SDFT+UU electronic structures. Color, line and point types are the same as Fig. 9.

IV.5 Systematic evaluation of J(𝒒)J(\bm{q}) based on the Liechtenstein method

Let us move on to the results of J(𝒒)J(\bm{q}) calculated with SDFT+UU for GdT2T_{2}X2X_{2} and EuT2T_{2}X2X_{2} in Fig. 9 and Fig. 9, respectively. From these figures, we see that the general trend of J(𝒒)J(\bm{q}) is not sensitive to the choice of group 14 elements X=X= Si or Ge. On the other hand, the choices of different ff-elements and different periods (i.e., 3dd, 4dd, or 5dd) of the transition metals TT dramatically change the structure of J(𝒒)J({\bm{q}}). For example, for the GdT2X2T_{2}X_{2} compounds, the TT dependence of group 5 (T=T= V, Nb, Ta), 8 (T=T= Fe, Ru, Os), and 9 (T=T= Co, Rh, Ir) are not so significant. On the other hand, for EuT2X2T_{2}X_{2}, the TT dependence is more noticeable. These dependencies may come from the chemical pressure effects on the lattice structures or the spreads and local energy levels of the dd-orbitals.

According to Figs. 9 and 9, the most promising candidates showing skyrmion lattice phase would be compounds with TT being the group 8 transition metal. In particular, GdRu2Si2, GdRu2Ge2, GdOs2Si2, GdOs2Ge2, and EuOs2Si2 have the highest peaks at finite 𝑸\bm{Q} along the Γ\Gamma-X line. Together with the small easy-axis anisotropy shown in Figs. 2(a-d), they should show the skyrmion lattice phase, and indeed, it is already found in GdRu2Si2. Except for these compounds, we can also see that GdWX22{}_{2}X_{2}, GdReX22{}_{2}X_{2}, GdAgX22{}_{2}X_{2}, GdAuX22{}_{2}X_{2}, EuCoX22{}_{2}X_{2}, and EuAgX22{}_{2}X_{2} with XX = Si and Ge have the highest peaks along the Γ\Gamma-X line. However, the peaks in GdWX22{}_{2}X_{2} and EuCoX22{}_{2}X_{2} may be too small to stabilize the magnetic order.

IV.6 Modulation vector based on spin spiral calculations

Up to now, we have discussed the magnetic interaction J(𝒒)J(\bm{q}) and the corresponding modulation vectors 𝑸\bm{Q} based on the Liechtenstein method. The calculations are performed for the Wannier tight-binding models derived from first-principles for the ferromagnetic ground states. Thus, in principle, the reliability of the 𝒒\bm{q}-dependence far from the Γ\Gamma point, as well as the validity of the mapping to the classical spin model cannot be justified. Here, we present spin spiral calculations for the materials expected to be a good candidate for showing the skyrmion phase. Since the calculation is exact within the SDFT+UU level, the results can be used in a way complementary to that of the Liechtenstein method.

Refer to caption
Figure 10: Modulation vector 𝒒\bm{q} dependence of the ground state energy E(𝒒)E(\bm{q}) calculated based on the Frozen magnon method. (a) The results based on the SDFT+UU with U=6.7U=6.7 eV and J=0.7J=0.7 eV for GdT2X2T_{2}X_{2} with T=T= Fe, Ru, Os and X=X= Si, Ge. (b) The results based on the SDFT for GdT2X2T_{2}X_{2} with T=T= Fe, Ru, Os and X=X= Si, Ge. (c) The results based on the SDFT+UU with U=6.7U=6.7 eV and J=0.7J=0.7 eV for GdT2X2T_{2}X_{2} with T=T= Re, Ag, Au and X=X= Si, Ge.

Figure 10(a) shows the 𝒒\bm{q}-dependence of the total energies for GdT2T_{2}X2X_{2} with T=T= Fe, Ru, Os. The calculations are performed by the spin spiral calculations based on SDFT+UU with U=6.7U=6.7 and J=0.7J=0.7 eV. Since the minimum position of E(𝒒)E(\bm{q}) represents the most stable modulation vector, we can see that GdRu2Si2, GdRu2Ge2, and GdOs2Si2 favor the finite-QQ state with 𝑸(0.20,0,0)\bm{Q}\sim(0.20,0,0)-(0.25,0,0)(0.25,0,0). On the other hand, GdFe2Si2 and GdFe2Ge2 have the minimum at the Γ\Gamma point, indicating that the ferromagnetic state is the most stable. These features are in good agreement with the calculations based on the Liechtenstein method. Although the GdOs2Ge2 does not show the finite-𝑸\bm{Q} instability in contrast with the Liechtenstein result, the peak of J(𝒒)J(\bm{q}) in GdOs2Si2 is more fragile than the others as is shown in Fig. 9(e). Thus, the disagreement with the Liechtenstein method is not serious. For the comparison, we also show E(𝒒)E(\bm{q}) of GdT2X2T_{2}X_{2} with T=T= Fe, Ru, Os and X=X= Si, Ge based on SDFT in Fig. 10(b). We can see that the finite-𝑸\bm{Q} instability becomes stronger in all cases than in SDFT+UU. Here, GdFe2Si2 shows the finite-𝑸\bm{Q} instability in SDFT while it does not in SDFT+UU, which again agrees well with the Liechtenstein calculations. In Fig. 10(c), we show E(𝒒)E(\bm{q}) for the other candidates for showing the skyrmion phase predicted by the Liechtenstein calculations, namely, GdT2X2T_{2}X_{2} with T=T= Re, Ag, Au and X=X= Si, Ge, EuAg2Si2 and EuAg2Ge2. We can see that GdRe2Si2, and GdRe2Ge2 show nearly flat 𝒒\bm{q} dependence near the Γ\Gamma point, and thus, it is possible that some subtle perturbations select a finite-𝑸\bm{Q} state. On the other hand, GdAg2X2X_{2}, GdAu2X2X_{2}, and EuAg2X2X_{2} have a broad peak at 𝑸(12,0,0)\bm{Q}\sim(\frac{1}{2},0,0) corresponding to the 90-degree rotating structure with the nearest neighboring spins, which seems to be too large to realize a skyrmion phase. However, due to its broad nature, we may still have a chance to achieve a skyrmion phase with smaller 𝑸\bm{Q} in experiments by using, for example, chemical substitution and external pressure. We leave the possible magnetic structures favored in these compounds as a future study.

V Conclusion and outlook

In this Perspective, we present systematic first-principles calculations for GdT2X2T_{2}X_{2} and EuT2X2T_{2}X_{2} with TT being a transition metal element and XX being Si or Ge. From the magnetocrystalline anisotropy calculations based on SDFT and SDFT+UU, we show that the inclusion of Coulomb interaction UU and Hund’s coupling JJ is essential to obtain the easy-axis anisotropy. Then, based on the SDFT+UU electronic structures, we evaluate magnetic interactions J(𝒒)J(\bm{q}) based on the Liechtenstein method and show that the obtained J(𝒒)J(\bm{q}) is in good agreement with E(𝒒)E(\bm{q}) by the spin spiral calculations. Our calculations indicate that the finite-𝑸{\bm{Q}} structure is determined not only by the Fermi surface topology but also by the details of the electronic structure, and the competition of each Gd-5d5d orbital contribution determines whether a ferromagnetic spin configuration or finite-𝑸{\bm{Q}} structure is favored in GdT2T_{2}Si2 with T=T= Fe and Ru. According to our calculations, GdRuX22{}_{2}X_{2}, GdOsX22{}_{2}X_{2}, and GdReX22{}_{2}X_{2} are promising candidates, while GdAgX22{}_{2}X_{2}, GdAuX22{}_{2}X_{2}, and EuAgX22{}_{2}X_{2} are possible candidates for showing the skyrmion lattice phase. Since the systematic calculations based on the Liechtenstein method is shown to be helpful in evaluating the finite-𝑸\bm{Q} structure, the extension to other crystal structure would be a possible direction to discover and engineer new skyrmion compounds. On the other hand, developing new methods to evaluate transport properties in the short pitch skyrmion phase is another crucial direction for its practical applications. Since most of the novel phenomena associated with its high skyrmion density have been studied based on simple models, further development is necessary to achieve a better understanding of material dependence and quantitative evaluations.

VI Acknowledgments

This work was supported by JSPS-KAKENHI (No. 22H00290, 21H04437, 21H04990, and 19H05825) and JST-PRESTO (No. JPMJPR20L7). We also acknowledge the Center for Computational Materials Science, Institute for Materials Research, Tohoku University, for the use of MASAMUNE-IMR (Project No. 202112-SCKXX-0010).

VII AUTHOR DECLARATIONS

Conflict of Interest

The authors have no conflicts to disclose.

Author Contributions

Takuya Nomoto:Formal analysis (lead); Writing - original draft. Ryotaro Arita: Formal analysis (supporting); Writing - review & editing.

DATA AVAILABILITY

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • Nagaosa and Tokura [2013] N. Nagaosa and Y. Tokura, Topological properties and dynamics of magnetic skyrmions, Nature Nanotechnology 8, 899 (2013).
  • Göbel et al. [2021] B. Göbel, I. Mertig, and O. A. Tretiakov, Beyond skyrmions: Review and perspectives of alternative magnetic quasiparticles, Physics Reports 895, 1 (2021), beyond skyrmions: Review and perspectives of alternative magnetic quasiparticles.
  • Tokura and Kanazawa [2021] Y. Tokura and N. Kanazawa, Magnetic skyrmion materials, Chemical Reviews 121, 2857 (2021), pMID: 33164494.
  • Finocchio et al. [2016] G. Finocchio, F. B. ttner, R. Tomasello, M. Carpentieri, and M. Kläui, Magnetic skyrmions: from fundamental to applications, Journal of Physics D: Applied Physics 49, 423001 (2016).
  • Fert et al. [2017] A. Fert, N. Reyren, and V. Cros, Magnetic skyrmions: advances in physics and potential applications, Nature Reviews Materials 2, 17031 (2017).
  • Luo and You [2021] S. Luo and L. You, Skyrmion devices for memory and logic applications, APL Materials 9, 050901 (2021).
  • Marrows and Zeissler [2021] C. H. Marrows and K. Zeissler, Perspective on skyrmion spintronics, Applied Physics Letters 119, 250502 (2021).
  • Wang et al. [2022] H. Wang, Y. Dai, G.-M. Chow, and J. Chen, Topological hall transport: Materials, mechanisms and potential applications, Progress in Materials Science 130, 100971 (2022).
  • Parkin et al. [2008] S. S. P. Parkin, M. Hayashi, and L. Thomas, Magnetic domain-wall racetrack memory, Science 320, 190 (2008).
  • Sampaio et al. [2013] J. Sampaio, V. Cros, S. Rohart, A. Thiaville, and A. Fert, Nucleation, stability and current-induced motion of isolated magnetic skyrmions in nanostructures, Nature Nanotechnology 8, 839 (2013).
  • Zhang et al. [2018] X. Zhang, W. Cai, X. Zhang, Z. Wang, Z. Li, Y. Zhang, K. Cao, N. Lei, W. Kang, Y. Zhang, H. Yu, Y. Zhou, and W. Zhao, Skyrmions in magnetic tunnel junctions, ACS Applied Materials & Interfaces 10, 16887 (2018).
  • Huang et al. [2017] Y. Huang, W. Kang, X. Zhang, Y. Zhou, and W. Zhao, Magnetic skyrmion-based synaptic devices, Nanotechnology 28, 08LT02 (2017).
  • Song et al. [2020] K. M. Song, J.-S. Jeong, B. Pan, X. Zhang, J. Xia, S. Cha, T.-E. Park, K. Kim, S. Finizio, J. Raabe, J. Chang, Y. Zhou, W. Zhao, W. Kang, H. Ju, and S. Woo, Skyrmion-based artificial synapses for neuromorphic computing, Nature Electronics 3, 148 (2020).
  • Lin et al. [1973] Y. S. Lin, P. J. Grundy, and E. A. Giess, Bubble domains in magnetostatically coupled garnet films, Applied Physics Letters 23, 485 (1973).
  • Garel and Doniach [1982] T. Garel and S. Doniach, Phase transitions with spontaneous modulation-the dipolar ising ferromagnet, Phys. Rev. B 26, 325 (1982).
  • Mühlbauer et al. [2009] S. Mühlbauer, B. Binz, F. Jonietz, C. Pfleiderer, A. Rosch, A. Neubauer, R. Georgii, and P. Böni, Skyrmion lattice in a chiral magnet, Science 323, 915 (2009).
  • Bogdanov and Yablonskii [1989] A. N. Bogdanov and D. A. Yablonskii, Thermodynamically stable “vortices” in magnetically ordered crystals. the mixed state of magnets, Sov. Phys. JETP 68, 101 (1989).
  • Rößler et al. [2006] U. K. Rößler, A. N. Bogdanov, and C. Pfleiderer, Spontaneous skyrmion ground states in magnetic metals, Nature 442, 797 (2006).
  • Fert and Levy [1980] A. Fert and P. M. Levy, Role of anisotropic exchange interactions in determining the properties of spin-glasses, Phys. Rev. Lett. 44, 1538 (1980).
  • Heide et al. [2008] M. Heide, G. Bihlmayer, and S. Blügel, Dzyaloshinskii-moriya interaction accounting for the orientation of magnetic domains in ultrathin films: Fe/w(110), Phys. Rev. B 78, 140403 (2008).
  • Freimuth et al. [2014] F. Freimuth, S. B. gel, and Y. Mokrousov, Berry phase theory of dzyaloshinskii– moriya interaction and spin– orbit torques, Journal of Physics: Condensed Matter 26, 104202 (2014).
  • Kikuchi et al. [2016] T. Kikuchi, T. Koretsune, R. Arita, and G. Tatara, Dzyaloshinskii-moriya interaction as a consequence of a doppler shift due to spin-orbit-induced intrinsic spin current, Phys. Rev. Lett. 116, 247201 (2016).
  • Koretsune et al. [2018] T. Koretsune, T. Kikuchi, and R. Arita, First-principles evaluation of the dzyaloshinskii– moriya interaction, Journal of the Physical Society of Japan 87, 041011 (2018).
  • Saha et al. [1999] S. R. Saha, H. Sugawara, T. D. Matsuda, H. Sato, R. Mallik, and E. V. Sampathkumaran, Magnetic anisotropy, first-order-like metamagnetic transitions, and large negative magnetoresistance in single-crystal gd2pdsi3{\mathrm{gd}}_{2}{\mathrm{pdsi}}_{3}Phys. Rev. B 60, 12162 (1999).
  • Kurumaji et al. [2019] T. Kurumaji, T. Nakajima, M. Hirschberger, A. Kikkawa, Y. Yamasaki, H. Sagayama, H. Nakao, Y. Taguchi, T. hisa Arima, and Y. Tokura, Skyrmion lattice with a giant topological hall effect in a frustrated triangular-lattice magnet, Science 365, 914 (2019).
  • Hirschberger et al. [2019] M. Hirschberger, T. Nakajima, S. Gao, L. Peng, A. Kikkawa, T. Kurumaji, M. Kriener, Y. Yamasaki, H. Sagayama, H. Nakao, K. Ohishi, K. Kakurai, Y. Taguchi, X. Yu, T.-h. Arima, and Y. Tokura, Skyrmion phase and competing magnetic orders on a breathing kagomé lattice, Nature Communications 10, 5831 (2019).
  • Khanh et al. [2020] N. D. Khanh, T. Nakajima, X. Yu, S. Gao, K. Shibata, M. Hirschberger, Y. Yamasaki, H. Sagayama, H. Nakao, L. Peng, K. Nakajima, R. Takagi, T.-h. Arima, Y. Tokura, and S. Seki, Nanometric square skyrmion lattice in a centrosymmetric tetragonal magnet, Nature Nanotechnology 15, 444 (2020).
  • Shang et al. [2021] T. Shang, Y. Xu, D. J. Gawryluk, J. Z. Ma, T. Shiroka, M. Shi, and E. Pomjakushina, Anomalous hall resistivity and possible topological hall effect in the eual4{\mathrm{eual}}_{4} antiferromagnet, Phys. Rev. B 103, L020405 (2021).
  • Kaneko et al. [2021] K. Kaneko, T. Kawasaki, A. Nakamura, K. Munakata, A. Nakao, T. Hanashima, R. Kiyanagi, T. Ohhara, M. Hedo, T. Nakama, and Y. Ōnuki, Charge-density-wave order and multiple magnetic transitions in divalent europium compound eual4, Journal of the Physical Society of Japan 90, 064704 (2021).
  • Zhu et al. [2022] X. Y. Zhu, H. Zhang, D. J. Gawryluk, Z. X. Zhen, B. C. Yu, S. L. Ju, W. Xie, D. M. Jiang, W. J. Cheng, Y. Xu, M. Shi, E. Pomjakushina, Q. F. Zhan, T. Shiroka, and T. Shang, Spin order and fluctuations in the eual4{\mathrm{eual}}_{4} and euga4{\mathrm{euga}}_{4} topological antiferromagnets: A μSR\mu\mathrm{SR} study, Phys. Rev. B 105, 014423 (2022).
  • Takagi et al. [2022] R. Takagi, N. Matsuyama, V. Ukleev, L. Yu, J. S. White, S. Francoual, J. R. L. Mardegan, S. Hayami, H. Saito, K. Kaneko, K. Ohishi, Y. Ōnuki, T.-h. Arima, Y. Tokura, T. Nakajima, and S. Seki, Square and rhombic lattices of magnetic skyrmions in a centrosymmetric binary compound, Nature Communications 13, 1472 (2022).
  • Zhang et al. [2021] H. Zhang, X. Y. Zhu, Y. Xu, D. J. Gawryluk, W. Xie, S. L. Ju, M. Shi, T. Shiroka, Q. F. Zhan, E. Pomjakushina, and T. Shang, Giant magnetoresistance and topological hall effect in the euga4 antiferromagnet, Journal of Physics: Condensed Matter 34, 034005 (2021).
  • [33] Y. Ōnuki, M. Kakihana, W. Iha, K. Nakaima, D. Aoki, A. Nakamura, F. Honda, M. Nakashima, Y. Amako, J. Gouchi, Y. Uwatoko, S. Nakamura, T. Sakakibara, T. Takeuchi, Y. Haga, H. Ikeda, H. Harima, M. Hedo, and T. Nakama, Unique skyrmion phases and conduction electrons in cubic chiral antiferromagnet euptsi and related compounds, in Proceedings of the International Conference on Strongly Correlated Electron Systems (SCES2019).
  • Sakakibara et al. [2021] T. Sakakibara, S. Nakamura, S. Kittaka, M. Kakihana, M. Hedo, T. Nakama, and Y. Ōnuki, Magnetic phase transitions of the 4f skyrmion compound euptsi studied by magnetization measurements, Journal of the Physical Society of Japan 90, 064701 (2021).
  • Kanazawa et al. [2011] N. Kanazawa, Y. Onose, T. Arima, D. Okuyama, K. Ohoyama, S. Wakimoto, K. Kakurai, S. Ishiwata, and Y. Tokura, Large topological hall effect in a short-period helimagnet mnge, Phys. Rev. Lett. 106, 156603 (2011).
  • Shibata et al. [2013] K. Shibata, X. Z. Yu, T. Hara, D. Morikawa, N. Kanazawa, K. Kimoto, S. Ishiwata, Y. Matsui, and Y. Tokura, Towards control of the size and helicity of skyrmions in helimagnetic alloys by spin– orbit coupling, Nature Nanotechnology 8, 723 (2013).
  • Hamamoto et al. [2015] K. Hamamoto, M. Ezawa, and N. Nagaosa, Quantized topological hall effect in skyrmion crystal, Phys. Rev. B 92, 115417 (2015).
  • Matsui et al. [2021] A. Matsui, T. Nomoto, and R. Arita, Skyrmion-size dependence of the topological hall effect: A real-space calculation, Phys. Rev. B 104, 174432 (2021).
  • Okubo et al. [2012] T. Okubo, S. Chung, and H. Kawamura, Multiple-qq states and the skyrmion lattice of the triangular-lattice heisenberg antiferromagnet under magnetic fields, Phys. Rev. Lett. 108, 017206 (2012).
  • Leonov and Mostovoy [2015] A. O. Leonov and M. Mostovoy, Multiply periodic states and isolated skyrmions in an anisotropic frustrated magnet, Nature Communications 6, 8275 (2015).
  • Chen et al. [2016] J. P. Chen, D.-W. Zhang, and J.-M. Liu, Exotic skyrmion crystals in chiral magnets with compass anisotropy, Scientific Reports 6, 29126 (2016).
  • Chen et al. [2018] J. Chen, D.-W. Zhang, Y. Chen, X. Gao, and J.-M. Liu, Compass-anisotropy-modulated helical states and skyrmion crystals in chiral magnets, Physics Letters A 382, 2944 (2018).
  • Heinze et al. [2011] S. Heinze, K. von Bergmann, M. Menzel, J. Brede, A. Kubetzka, R. Wiesendanger, G. Bihlmayer, and S. Blü gel, Spontaneous atomic-scale magnetic skyrmion lattice in two dimensions, Nature Physics 7, 713 (2011).
  • Hayami et al. [2017] S. Hayami, R. Ozawa, and Y. Motome, Effective bilinear-biquadratic model for noncoplanar ordering in itinerant magnets, Phys. Rev. B 95, 224424 (2017).
  • Ozawa et al. [2016] R. Ozawa, S. Hayami, K. Barros, G.-W. Chern, Y. Motome, and C. D. Batista, Vortex crystals with chiral stripes in itinerant magnets, Journal of the Physical Society of Japan 85, 103703 (2016).
  • Ozawa et al. [2017] R. Ozawa, S. Hayami, and Y. Motome, Zero-field skyrmions with a high topological number in itinerant magnets, Phys. Rev. Lett. 118, 147205 (2017).
  • Hayami and Motome [2021] S. Hayami and Y. Motome, Topological spin crystals by itinerant frustration, Journal of Physics: Condensed Matter 33, 443001 (2021).
  • Nomoto et al. [2020a] T. Nomoto, T. Koretsune, and R. Arita, Formation mechanism of the helical 𝒒\bm{q} structure in gd-based skyrmion materials, Phys. Rev. Lett. 125, 117204 (2020a).
  • Bouaziz et al. [2022] J. Bouaziz, E. Mendive-Tapia, S. Blügel, and J. B. Staunton, Fermi-surface origin of skyrmion lattices in centrosymmetric rare-earth intermetallics, Phys. Rev. Lett. 128, 157206 (2022).
  • Grytsiuk et al. [2020] S. Grytsiuk, J.-P. Hanke, M. Hoffmann, J. Bouaziz, O. Gomonay, G. Bihlmayer, S. Lounis, Y. Mokrousov, and S. Blü gel, Topological– chiral magnetic interactions driven by emergent orbital magnetism, Nature Communications 11, 511 (2020).
  • Mendive-Tapia et al. [2021] E. Mendive-Tapia, M. dos Santos Dias, S. Grytsiuk, J. B. Staunton, S. Blügel, and S. Lounis, Short period magnetization texture of b20-mnge explained by thermally fluctuating local moments, Phys. Rev. B 103, 024410 (2021).
  • Korotin et al. [2015] D. M. Korotin, V. V. Mazurenko, V. I. Anisimov, and S. V. Streltsov, Calculation of exchange constants of the heisenberg model in plane-wave-based methods using the green’s function approach, Phys. Rev. B 91, 224405 (2015).
  • Nomoto et al. [2020b] T. Nomoto, T. Koretsune, and R. Arita, Local force method for the ab initio tight-binding model: Effect of spin-dependent hopping on exchange interactions, Phys. Rev. B 102, 014444 (2020b).
  • Mostofi et al. [2008] A. A. Mostofi, J. R. Yates, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, wannier90: A tool for obtaining maximally-localised wannier functions, Computer Physics Communications 178, 685 (2008).
  • Marzari et al. [2012] N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Maximally localized wannier functions: Theory and applications, Rev. Mod. Phys. 84, 1419 (2012).
  • Chikano et al. [2019] N. Chikano, K. Yoshimi, J. Otsuki, and H. Shinaoka, irbasis: Open-source database and software for intermediate-representation basis functions of imaginary-time green’ s function, Computer Physics Communications 240, 181 (2019).
  • Li et al. [2020] J. Li, M. Wallerberger, N. Chikano, C.-N. Yeh, E. Gull, and H. Shinaoka, Sparse sampling approach to efficient ab initio calculations at finite temperature, Phys. Rev. B 101, 035144 (2020).
  • Sandratskii [1998] L. M. Sandratskii, Noncollinear magnetism in itinerant-electron systems: Theory and applications, Advances in Physics 47, 91 (1998).
  • Kresse and Furthmüller [1996] G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54, 11169 (1996).
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett. 77, 3865 (1996).
  • Blöchl [1994] P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50, 17953 (1994).
  • Kresse and Joubert [1999] G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B 59, 1758 (1999).
  • Garnier et al. [1995] A. Garnier, D. Gignoux, N. Iwata, D. Schmitt, T. Shigeoka, and F. Zhang, Anisotropic metamagnetism in gdru2si2, Journal of Magnetism and Magnetic Materials 140-144, 899 (1995), international Conference on Magnetism.
  • Mihalik et al. [2005] M. Mihalik, P. Svoboda, M. Mihalik, J. Vejpravová, and V. Sechovský, Magnetic and transport properties of gdfe2si2 single-crystal, Journal of Magnetism and Magnetic Materials 290-291, 606 (2005), proceedings of the Joint European Magnetic Symposia (JEMS’ 04).
  • Kvashnin et al. [2016] Y. O. Kvashnin, R. Cardias, A. Szilva, I. Di Marco, M. I. Katsnelson, A. I. Lichtenstein, L. Nordström, A. B. Klautau, and O. Eriksson, Microscopic origin of heisenberg and non-heisenberg exchange interactions in ferromagnetic bcc fe, Phys. Rev. Lett. 116, 217202 (2016).
  • [66] J. Bouaziz and S. Blügel, in private communication.
  • Katsnelson and Lichtenstein [2000] M. I. Katsnelson and A. I. Lichtenstein, First-principles calculations of magnetic interactions in correlated systems, Phys. Rev. B 61, 8906 (2000).