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

Antiferro skyrmion crystal phases in a synthetic bilayer antiferromagnet
under an in-plane magnetic field

Satoru Hayami Graduate School of Science Graduate School of Science Hokkaido University Hokkaido University Sapporo 060-0810 Sapporo 060-0810 Japan Japan
Abstract

We theoretically propose a generation of an antiferro skyrmion crystal (AF-SkX) in a synthetic bilayer antiferromagnetic system under an external magnetic field. By performing the simulated annealing for a spin model on a centrosymmetric bilayer triangular lattice, we find that the interplay among the layer-dependent Dzyaloshinskii-Moriya interaction, the easy-plane single-ion anisotropy, the interlayer exchange interaction, and the in-plane magnetic field leads to instability toward the AF-SkX. The obtained AF-SkX consists of the skyrmion layer and anti-skyrmion layer with opposite skyrmion numbers, which results in no skyrmion number as well as spin scalar chirality in the whole system. We also show that a ferri-type skyrmion crystal, where one of the layers has the skyrmion number of one and the other does not, appears when the magnitude of the magnetic field is varied. Our results provide a guideline to experimentally search for the AF-SkX in bulk and layered materials.

1 Introduction

Noncoplanar magnets, where the spins align neither in a line nor on a plane, have attracted growing interest in various fields of condensed matter physics, such as spintronics, topological magnets, and multiferroics [1]. When noncoplanar spin textures are coupled to itinerant electrons, unconventional quantum transport, such as the topological Hall effect and nonreciprocal transport, occurs through a fictitious magnetic field based on the Berry phase mechanism, which does not rely on relativistic spin-orbit coupling [2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12]. As such a fictitious field reaches as large as 10310^{3}10410^{4} T, noncoplanar magnets may be useful to realize high-speed and low-power devices for spintronic applications [13, 14, 15, 16, 17].

One of the most typical examples to have noncoplanar spin textures is a magnetic skyrmion, which is characterized by an integer topological number (skyrmion number) [18, 19, 20, 21]. Once the skyrmion is periodically aligned under the lattice structures, which is so-called the skyrmion crystal (SkX), a large magnetoelectric response triggered by the Berry phase is expected [22, 23, 24]. In spite of complicated spin textures in the SkX, it has been observed in various materials under both noncentrosymmetric [25, 26, 27, 28, 29, 30, 31, 32] and centrosymmetric [33, 34, 35, 36, 37, 38, 39] lattice structures [40]. Simultaneously, theoretical studies have revealed the stabilization mechanisms of the SkX: the Dzyaloshinskii-Moriya (DM) interaction [41, 42, 43, 44, 45], dipolar interaction [46, 47, 48], frustrated exchange interaction [49, 50, 51, 52, 53, 54, 55, 56, 57], long-range spin interaction [58, 59, 60, 61, 62, 63, 64], multiple spin interaction [65, 66, 67, 68, 69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79], anisotropic exchange interaction [80, 81, 82, 83, 84, 85, 86, 87, 88, 89], and their combination [90, 91, 92, 93]. Thus, many potential situations exist so as to engineer and design the SkX irrespective of the lattice structures and microscopic mechanisms.

On the other hand, an SkX in synthetic antiferromagnets, which is characterized by an antiferroic alignment of the magnetic skyrmions under multi-sublattice structures, has been still rare compared to the SkX [94, 95, 96]. We here refer to such an SkX as antiferro SkX (AF-SkX), since the AF-SkX consists of the skyrmions with an opposite skyrmion number in different sublattices as a consequence of the staggered alignment of spins like 𝑺(A)=𝑺(B)\bm{S}^{(\rm A)}=-\bm{S}^{(\rm B)} (𝑺(η)\bm{S}^{(\eta)} is the spins for sublattice η\eta). In such a situation, the emergent magnetic field is canceled out in the whole system; the topological Hall effect vanishes [97]. Meanwhile, the spin textures in the AF-SkX are topologically protected, which results in other topological phenomena, such as the topological spin Hall effect [98, 97, 99]. Moreover, it can be promising for spintronic applications owing to the absence of the skyrmion Hall effect; the isolated AFM skyrmion can move parallel to an applied current without deviating [100, 101, 102, 103, 104, 105, 106, 107]. Thus, it is desired to find out the stabilization mechanism of the AF-SkX, which is one of the challenges in this active research area. However, such a condition especially for the case in the presence of an external magnetic field has not been clarified yet, since the magnetic field usually breaks the relationship of 𝑺(A)=𝑺(B)\bm{S}^{(\rm A)}=-\bm{S}^{(\rm B)} in the spin textures.

In the present study, we theoretically investigate the realization of the AF-SkX in centrosymmetric layered magnets under the external magnetic field. The results are obtained by performing the simulated annealing for an effective spin model with the intralayer momentum-resolved interaction and interlayer exchange interaction on a bilayer triangular lattice. By taking into account the effect of the staggered DM interaction, easy-plane anisotropy, and in-plane magnetic field, we show that the AF-SkX is robustly stabilized in the ground state. The AF-SkX exhibits the opposite skyrmion number for different layers so that the spin scalar chirality in the whole system vanishes. This AF-SkX is qualitatively different from the previous antiferromagnetic SkX stabilized at zero field [100, 101, 108, 109, 110, 111] and sublattice-dependent SkX with a uniform scalar chirality [112, 113, 114, 115, 116, 117, 118, 119, 120, 121]. We also find that two ferri-type SkX (Ferri-SkX) phases, which consist of the skyrmion layer and topologically-trivial magnetic layer, appear in the vicinity of the AF-SkX phase. Our results provide a way of realizing the AF-SkX by applying the magnetic field far from zero magnetic field, which will stimulate further experimental observations of exotic topological spin crystals beyond the conventional SkX.

The remainder of this paper is organized as follows. In Sec. 2, we introduce the spin model on the bilayer triangular lattice. We also outline simulated annealing to obtain the ground-state spin configuration. Then, we show the magnetic phase diagram while changing the interlayer exchange interaction and the in-plane magnetic field in Sec. 3. We discuss the spin and scalar chirality configurations in the AF-SkX as well as the other multiple-QQ and single-QQ states. Finally, we summarize the paper in Sec. 4.

2 Model and method

Refer to caption
Figure 1: (Color online) (a) Bilayer triangular lattice consisting of layer A and layer B; JJ_{\parallel} represents the interlayer exchange coupling. The black circle denotes the inversion center. (b) The layer-dependent DM interaction for (left) layer A and (right) layer B in the presence of the local crystalline electric field 𝑬\bm{E}. The green arrows stand for the DM vectors, which are opposite for two layers.

We investigate the multiple-QQ instability in the bilayer triangular-lattice system. The bilayer structure consists of two triangular layers on the xyxy plane, which are denoted as layer A and layer B, as shown in Fig. 1(a). The two layers are separated along the zz axis by c=1c=1, where the lattice sites are located at the same xyxy position; the lattice constant of the triangular lattice is set as unity. The effect of the difference along the zz and inplane directions is taken into account as the different magnitudes of the interactions [see JJ and JJ_{\parallel} in Eqs. (2) and (3)]. It is noted that an inversion center is present at the nearest-neighbor bond between layers A and B denoted as the black circle in Fig. 1(a), although there is no local inversion center in each layer. Specifically, we consider the situation where the total bilayer system belongs to the D6hD_{\rm 6h} point group, and the individual layer has the C6vC_{\rm 6v} symmetry; the local crystalline electric field 𝑬\bm{E} with the same magnitude but the opposite direction is present in the zz direction [Fig. 1(b)], which is a source of the layer-dependent DM interaction, as described below. In such a situation, the DM vector lies in the xyxy plane like the Rashba-induced case [122, 123, 124].

In such a layered lattice structure without local inversion symmetry, we consider the spin Hamiltonian \mathcal{H}, which is given by

=\displaystyle\mathcal{H}= ηη++loc,\displaystyle\sum_{\eta}\mathcal{H}^{\perp}_{\eta}+\mathcal{H}^{\parallel}+\mathcal{H}^{{\rm loc}}, (1)

where \mathcal{H}^{\perp}, \mathcal{H}^{\parallel}, and loc\mathcal{H}^{{\rm loc}} are explicitly represented as

η=\displaystyle\mathcal{H}^{\perp}_{\eta}= ν[J𝑺𝑸ν(η)𝑺𝑸ν(η)i𝑫𝑸ν(η)(𝑺𝑸ν(η)×𝑺𝑸ν(η))],\displaystyle\sum_{\nu}\Big{[}-J\bm{S}^{(\eta)}_{{\bm{Q}_{\nu}}}\cdot\bm{S}^{(\eta)}_{-\bm{Q}_{\nu}}-i\bm{D}^{(\eta)}_{\bm{Q}_{\nu}}\cdot(\bm{S}^{(\eta)}_{\bm{Q}_{\nu}}\times\bm{S}^{(\eta)}_{-\bm{Q}_{\nu}})\Big{]}, (2)
=\displaystyle\mathcal{H}^{\parallel}= Ji,j𝑺i𝑺j,\displaystyle J_{\parallel}\sum_{\langle i,j\rangle}\bm{S}_{i}\cdot\bm{S}_{j}, (3)
loc=\displaystyle\mathcal{H}^{{\rm loc}}= Aioni(Siz)2HiSix.\displaystyle A^{\rm ion}\sum_{i}(S_{i}^{z})^{2}-H\sum_{i}S_{i}^{x}. (4)

Here, 𝑺i\bm{S}_{i} is the classical localized spin at site ii with the magnitude of |𝑺i|=1|\bm{S}_{i}|=1. The intralayer Hamiltonian \mathcal{H}^{\perp} in Eq. (2) consists of the Heisenberg-type isotropic exchange interaction JJ in the first term and the antisymmetric DM interaction 𝑫𝑸ν(η)\bm{D}_{\bm{Q}_{\nu}}^{(\eta)} in the second term. Both interaction terms are represented in momentum space; 𝑺𝑸ν(η)\bm{S}^{(\eta)}_{\bm{Q}_{\nu}} with layer η\eta and wave vector 𝑸ν\bm{Q}_{\nu} is the Fourier transform of 𝑺i\bm{S}_{i}.

To focus on the multiple-QQ states that arise from a superposition of finite-QQ spiral states, we consider the ±𝑸1=(Q,0)\pm\bm{Q}_{1}=(Q,0), ±𝑸2=(Q/2,3Q/2)\pm\bm{Q}_{2}=(-Q/2,\sqrt{3}Q/2), and ±𝑸3=(Q/2,3Q/2)\pm\bm{Q}_{3}=(-Q/2,-\sqrt{3}Q/2) components of the interaction, which are related by threefold rotational symmetry of the triangular lattice [125]. We set Q=π/3Q=\pi/3 without loss of generality. Such a finite-QQ instability is obtained by considering the effect of magnetic frustration arising from the competing exchange interaction in real and/or momentum space. JJ is independent of 𝑸ν\bm{Q}_{\nu}, while 𝑫𝑸ν(η)\bm{D}_{\bm{Q}_{\nu}}^{(\eta)} depends on 𝑸ν\bm{Q}_{\nu}. We assume that the position of 𝑸ν\bm{Q}_{\nu} is fixed against the magnetic field for simplicity. Owing to the polar-type crystalline electric field along the zz direction, the DM vector is perpendicular to both the intralayer nearest-neighbor bond direction and the zz direction, as shown in Fig. 1(b); the magnitude of the DM interaction is set as |𝑫𝑸ν(η)|=D|\bm{D}_{\bm{Q}_{\nu}}^{(\eta)}|=D. In addition, the sign of the DM interaction is opposite for layers A and B due to the staggered crystalline electric field. Thus, 𝑫𝑸1(A)=𝑫𝑸1(B)=(0,D,0)\bm{D}_{\bm{Q}_{1}}^{(\rm A)}=-\bm{D}_{\bm{Q}_{1}}^{(\rm B)}=(0,-D,0), 𝑫𝑸2(A)=𝑫𝑸2(B)=(3D/2,D/2,0)\bm{D}_{\bm{Q}_{2}}^{(\rm A)}=-\bm{D}_{\bm{Q}_{2}}^{(\rm B)}=(\sqrt{3}D/2,D/2,0), and 𝑫𝑸3(A)=𝑫𝑸3(B)=(3D/2,D/2,0)\bm{D}_{\bm{Q}_{3}}^{(\rm A)}=-\bm{D}_{\bm{Q}_{3}}^{(\rm B)}=(-\sqrt{3}D/2,D/2,0). It is noted that DD can take an arbitrary value depending on the real-space interactions; we treat DD as the model parameter.

The interlayer Hamiltonian \mathcal{H}^{\parallel} in Eq. (3) includes the antiferromagnetic nearest-neighbor interaction between layers A and B, J>0J_{\parallel}>0; i,j\langle i,j\rangle represents the nearest-neighbor pair between layers A and B. The local Hamiltonian loc\mathcal{H}^{{\rm loc}} in Eq. (4) consists of the easy-plane single-ion anisotropy Aion>0A^{\rm ion}>0 and the in-plane magnetic field along the xx direction with the magnitude of HH. It is noted that the in-plane magnetic field breaks the threefold rotational symmetry of the lattice structure, which makes 𝑸1\bm{Q}_{1} and 𝑸2,3\bm{Q}_{2,3} inequivalent. The recent theoretical studies have revealed that the interplay between a layer-dependent DM interaction and interlayer exchange interaction stabilizes the SkX with the net scalar chirality rather than the AF-SkX without the net scalar chirality under the out-of-plane magnetic field irrespective of the sign of the interlayer exchange interaction [126, 127, 128, 129].

The model in Eq. (1) has five independent parameters: JJ, DD, JJ_{\parallel}, AionA^{\rm ion}, and HH. Among them, we take J=1J=1 as the energy unit of the model. Besides, we fix D=0.05D=0.05 and Aion=0.2A^{\rm ion}=0.2 so that the SkX is stabilized under the in-plane magnetic field in the single-layer limit (J=0J_{\parallel}=0); the SkX tends to be destabilized without AionA^{\rm ion} or easy-axis anisotropy Aion<0A^{\rm ion}<0. Especially, only the single-QQ conical spiral state appears for Aion=D=0A^{\rm ion}=D=0. We discuss the stability of the AF-SkX when JJ_{\parallel} and HH are varied. For the ferromagnetic interlayer exchange interaction, i.e., J<0J_{\parallel}<0, a ferro-type SkX with the same skyrmion number in both layers is stabilized as found in previous studies [126, 127, 128, 129].

We construct the magnetic phase diagram of the model in Eq. (1) at low temperatures by performing simulated annealing. The simulations are carried out with the standard Metropolis local updates for real-space spins, where the direction of spins is randomly chosen on the sphere with |𝑺i|=1|\bm{S}_{i}|=1. In each simulation, we gradually reduce the temperature with a rate Tn+1=αTnT_{n+1}=\alpha T_{n}, where TnT_{n} is the nnth-step temperature, to find the low-energy spin configurations. We set the initial temperature T0/J=1T_{0}/J=1–10, final temperature Tf/J=0.001T_{f}/J=0.001, and α=0.999995\alpha=0.999995. The initial spin configuration is randomly chosen. At the final temperature, 10510^{5}10610^{6} Monte Carlo sweeps are performed for measurements. We also start the simulations from the spin configurations obtained by the above procedure in the vicinity of the phase boundaries. The total number of spins is taken as N=2×L2=2×242N=2\times L^{2}=2\times 24^{2} under the periodic boundary conditions, where LL is the length of the triangular-lattice system along the 𝒂1=(1,0)\bm{a}_{1}=(1,0) and 𝒂2=(1/2,3/2)\bm{a}_{2}=(-1/2,\sqrt{3}/2) directions (𝒂1\bm{a}_{1} and 𝒂2\bm{a}_{2} are the primitive lattice vectors).

We calculate the spin- and chirality-related quantities to identify the magnetic phases. In the spin sector, we calculate the α=x,y,z\alpha=x,y,z component of the spin structure factor Ssηα(𝒒)S_{s\eta}^{\alpha}(\bm{q}), which is given by

Ssηα(𝒒)=1L2i,jηSiαSjαei𝒒(𝒓i𝒓j),\displaystyle S_{s\eta}^{\alpha}(\bm{q})=\frac{1}{L^{2}}\sum_{i,j\in\eta}S^{\alpha}_{i}S^{\alpha}_{j}e^{i\bm{q}\cdot(\bm{r}_{i}-\bm{r}_{j})}, (5)

where the site indices ii and jj are taken for the spins on layers η=A\eta={\rm A} and B. The magnetic moments at wave vector 𝒒\bm{q} and layer η\eta are given by mη𝒒α=Ssηα(𝒒)/L2m^{\alpha}_{\eta\bm{q}}=\sqrt{S_{s\eta}^{\alpha}(\bm{q})/L^{2}}. We also calculate mη𝒒xy=(mη𝒒x)2+(mη𝒒y)2m^{xy}_{\eta\bm{q}}=\sqrt{(m^{x}_{\eta\bm{q}})^{2}+(m^{y}_{\eta\bm{q}})^{2}}. The uniform magnetization per layer is given by Mηα=(1/L2)iηSiαM^{\alpha}_{\eta}=(1/L^{2})\sum_{i\in\eta}S^{\alpha}_{i}.

In the scalar chirality sector, we compute the uniform spin scalar chirality for layer η\eta, which is represented by

χηsc=1L2𝑹η𝑺i(𝑺j×𝑺k),\displaystyle\chi^{\rm sc}_{\eta}=\frac{1}{L^{2}}\sum_{\bm{R}\in\eta}\bm{S}_{i}\cdot(\bm{S}_{j}\times\bm{S}_{k}), (6)

where 𝑹\bm{R} represents the position vector at the centers of triangles; the sites ii, jj, and kk form the triangle at 𝑹\bm{R} in the counterclockwise order. It is noted that there are upward and downward triangle plaquettes. The local scalar chirality is given by χ𝑹=𝑺i(𝑺j×𝑺k)\chi_{\bm{R}}=\bm{S}_{i}\cdot(\bm{S}_{j}\times\bm{S}_{k}).

3 Results

Refer to caption
Figure 2: (Color online) Magnetic phase diagram in the plane of the interlayer exchange interaction, JJ_{\parallel}, and the magnetic field along the xx direction, HH, obtained by the simulated annealing at T/J=0.001T/J=0.001. 1QQ and 3QQ denote the single-QQ and triple-QQ states, respectively. Ferri-SkX, AF-SkX, and FP stand for the ferromagnetic skyrmion crystal, the antiferroic skyrmion crystal, and the fully-polarized state, respectively. The rhombus symbols at (J,H)=(0.1,0.75)(J_{\parallel},H)=(0.1,0.75) and (J,H)=(0.225,0.9)(J_{\parallel},H)=(0.225,0.9) denote the AF-SkX’ phase, which is characterized by the opposite skyrmion number for two layers like the AF-SkX but χAscχBsc\chi^{\rm sc}_{\rm A}\neq-\chi^{\rm sc}_{\rm B}.
Refer to caption
Figure 3: (Color online) Left: Real-space spin configurations on layer A of (a) the 1Q1Q I state at J=0.1J_{\parallel}=0.1 and H=0.5H=0.5, (b) the Ferri-SkX I at J=0.1J_{\parallel}=0.1 and H=0.7H=0.7, (c) the AF-SkX at J=0.1J_{\parallel}=0.1 and H=0.9H=0.9, and (d) the Ferri-SkX II at J=0.1J_{\parallel}=0.1 and H=1.2H=1.2. The arrows and colors represent the in-plane and out-of-plane components of the spin moment, respectively [130]. Middle left: Real-space distribution of the spin scalar chirality at each triangle plaquette. Middle right and right: The spin and scalar chirality configurations on layer B corresponding to the left and middle left panels.
Refer to caption
Figure 4: (Color online) The same plots as in Fig. 3 for (a) the 3Q3Q I state at J=0.1J_{\parallel}=0.1 and H=1.3H=1.3, (b) the 1QQ II state at J=0.1J_{\parallel}=0.1 and H=1.5H=1.5, (c) the 3QQ II state at J=0.4J_{\parallel}=0.4 and H=1H=1, (d) the 1QQ III state at J=0.4J_{\parallel}=0.4 and H=1.2H=1.2, and (e) the 1QQ IV state at J=0.4J_{\parallel}=0.4 and H=1.5H=1.5.
Refer to caption
Figure 5: (Color online) Square root of the spin structure factor in (left) the xx, (second left) yy, and (third left) zz components for layer A. The right three panels represent the data for layer B corresponding to the left three ones. The data are (a) the 1Q1Q I state at J=0.1J_{\parallel}=0.1 and H=0.5H=0.5, (b) the Ferri-SkX I at J=0.1J_{\parallel}=0.1 and H=0.7H=0.7, (c) the AF-SkX at J=0.1J_{\parallel}=0.1 and H=0.9H=0.9, and (d) the Ferri-SkX II at J=0.1J_{\parallel}=0.1 and H=1.2H=1.2. The black hexagons represent the first Brillouin zone.
Refer to caption
Figure 6: (Color online) The same plots as in Fig. 5 for (a) the 3Q3Q I state at J=0.1J_{\parallel}=0.1 and H=1.3H=1.3, (b) the 1QQ II state at J=0.1J_{\parallel}=0.1 and H=1.5H=1.5, (c) the 3QQ II state at J=0.4J_{\parallel}=0.4 and H=1H=1, (d) the 1QQ III state at J=0.4J_{\parallel}=0.4 and H=1.2H=1.2, and (e) the 1QQ IV state at J=0.4J_{\parallel}=0.4 and H=1.5H=1.5.
Refer to caption
Figure 7: (Color online) HH dependence of (a) MηxM^{x}_{\eta}, (b) Mηy,zM^{y,z}_{\eta}, (c) (mη𝑸νxy)2(m^{xy}_{\eta\bm{Q}_{\nu}})^{2}, (d) (mη𝑸νz)2(m^{z}_{\eta\bm{Q}_{\nu}})^{2}, and (e) χηsc\chi^{\rm sc}_{\eta} for η=\eta=A and B at J=0.1J_{\parallel}=0.1. The vertical dashed lines represent the phase boundary. The region drawn in blue represents the AF-SkX.
Refer to caption
Figure 8: (Color online) The same plots as in Fig. 7 at J=0.4J_{\parallel}=0.4.

Figure 2 shows the magnetic phase diagram against JJ_{\parallel} and HH at D=0.05D=0.05 and Aion=0.2A^{\rm ion}=0.2, which is obtained by performing simulated annealing down to T/J=0.001T/J=0.001. There are nine different phases in the phase diagram in addition to the fully-polarized state and the AF-SkX’, the latter of which is denoted as the rhombus symbol in Fig. 2 [see the caption in Fig. 2]. As the AF-SkX’ state appears only at two points in the phase diagram, we omit the detailed spin configurations on this phase. The spin and scalar chirality in real space for nine phases are shown in Figs. 3 and 4 and the corresponding spin structure factor in momentum space is shown in Figs. 5 and 6. In the left and middle right panels of Figs. 3 and 4, the arrows represent (Six,Siy)(S_{i}^{x},S_{i}^{y}), while the colors represent SizS_{i}^{z}. In the middle left and right panels of Figs. 3 and 4, the color represents χ𝑹\chi_{\bm{R}}. In Figs. 5 and 6, the square root of Ssηα(𝒒)S_{s\eta}^{\alpha}(\bm{q}) for α=x,y,z\alpha=x,y,z and η=\eta=A, B is shown. We also show the HH dependence of the uniform magnetization MηαM^{\alpha}_{\eta}, QνQ_{\nu}-resolved magnetic moments 𝒎η𝑸ν\bm{m}_{\eta\bm{Q}_{\nu}}, and scalar chirality χηsc\chi^{\rm sc}_{\eta} per layer at J=0.1J_{\parallel}=0.1 in Fig. 7 and J=0.4J_{\parallel}=0.4 in Fig. 8. In Figs. 7 and 8, the data of 𝒎η𝑸ν\bm{m}_{\eta\bm{Q}_{\nu}} in each ordered state are appropriately sorted for better readability.

For H=0H=0, the single-QQ (1QQ) I state appears for J0J_{\parallel}\geq 0. As shown in the real-space spin and chirality configurations in each layer in Fig. 3(a), the spin configurations in both layers are characterized by the single-QQ spiral state along the 𝑸3\bm{Q}_{3} direction. Due to the antiferromagnetic interlayer exchange interaction, the xyxy spins in layers A and B align antiparallel to each other. The spiral plane is fixed by DD and AionA^{\rm ion}: The former tends to favor the out-of-plane cycloidal spiral along the 𝑸ν\bm{Q}_{\nu} direction, while the latter tends to favor the in-plane cycloidal spiral. As DD is smaller than AionA^{\rm ion} in the present model parameter, the mη𝑸νzm^{z}_{\eta\bm{Q}_{\nu}} is much small than mη𝑸νxm^{x}_{\eta\bm{Q}_{\nu}} and mη𝑸νym^{y}_{\eta\bm{Q}_{\nu}}, as shown in Fig. 5(a). In other words, the spiral plane almost lies in the xyxy plane at zero field. The helicity of the out-of-plane cycloidal spiral is opposite for two layers due to the opposite sign of DD, while that of the in-plane one is the same.

Let us discuss a phase sequence against the in-plane magnetic field. We first examine the case for the small interlayer exchange interaction J=0.1J_{\parallel}=0.1. When the in-plane magnetic field is applied along the xx direction, the spiral states with 𝑸2\bm{Q}_{2} and 𝑸3\bm{Q}_{3} show smaller energy than that with 𝑸1\bm{Q}_{1}. This is because the magnetic field tends to favor the spiral, whose plane is perpendicular to the field direction. Indeed, mη𝑸νzm^{z}_{\eta\bm{Q}_{\nu}} is slightly enhanced owing to the spin flop against the field direction in Fig. 7(d). In addition to the xx component of the uniform magnetization MAx=MBxM^{x}_{\rm A}=M^{x}_{\rm B} in Fig. 7(a), the yy and zz components of the magnetization, i.e., MηyM^{y}_{\eta} and MηzM^{z}_{\eta}, are induced in both layers; MηyM^{y}_{\eta} is induced to have the uniform component as MAy=MByM^{y}_{\rm A}=M^{y}_{\rm B}, while MηzM^{z}_{\eta} is induced to have the staggered component as MAz=MBzM^{z}_{\rm A}=-M^{z}_{\rm B}, as shown in Fig. 7(b). Meanwhile, the 1QQ I state does not have a uniform scalar chirality χηsc\chi^{\rm sc}_{\eta} in both zero and finite fields, as shown in Figs. 3(a) and 7(e).

With an increase of HH, the 1QQ I state turns into the Ferri-SkX I state through the first-order phase transition, whose spin and chirality configurations in real space per layer are shown in Fig. 3(b). Although the spin configurations in both layers are characterized by the triple-QQ modulations, they show layer dependence, as shown in Fig. 5(b). Thus, there are both uniform and staggered components in the magnetization along the yy and zz directions, as shown in Fig. 7(b). In addition, both layers exhibit nonzero scalar chirality with different magnitudes in Fig. 7(e). Especially, one finds that the spin configuration on layer B corresponds to the SkX when calculating the skyrmion number. The skyrmion number in layer B becomes 1-1, while that in layer A becomes zero. Thus, this state is regarded as the ferri-type SkX consisting of the skyrmion layer and the topologically-trivial magnetic layer. It is noted that the sign of the skyrmion number in the SkX layer is arbitrary, which indicates that the skyrmion and anti-skyrmion are energetically degenerate. The similar states consisting of the SkX layer and the other magnetic layer have been realized in the trilayer system [127] and the threefold-screw symmetric system [131].

When HH is further increased, both layers exhibit the SkX spin configuration, as shown in Fig. 3(c). The spin structure factor in this state shows the same triple-QQ peak structures for layers A and B, as shown in Fig. 5(c). Meanwhile, there are two clear differences in terms of the spin and chirality configurations between the two layers: One is the sign of the zz-spin component and the other is the sign of the scalar chirality. In the case of Fig. 3(c), the skyrmion numbers in layers A and B are 1-1 and +1+1, respectively; the total skyrmion number in the whole system becomes zero. Similar to the skyrmion number, the total scalar chirality becomes zero owing to the cancellation between the two layers, χAsc=χBsc\chi^{\rm sc}_{\rm A}=-\chi^{\rm sc}_{\rm B} in Fig. 7(e). Thus, this state well corresponds to the AF-SkX, where the topological spin Hall effect is expected without the topological Hall effect. In this sense, this state is qualitatively different from the sublattice-dependent SkXs in previous studies, which have nonzero scalar chirality in the whole system [112, 113, 114, 115, 116, 117, 118, 119, 120, 121]. In addition, the present SkX is also different from the antiferromagnetic SkX parametrized by the Nèel order describing the bipartite magnetic lattice [100, 101, 108, 109, 110, 111].

The appearance of the AF-SkX is due to the synergy among DD, AionA^{\rm ion}, JJ_{\parallel}, and HH in the model in Eq. (1). The easy-plane single-ion anisotropy AionA^{\rm ion} and the in-plane magnetic field HH play an important role in stabilizing the SkX in each layer. In fact, these two ingredients become sources of the SkX when D=0D=0 and J=0J_{\parallel}=0, although the skyrmion and anti-skyrmion are energetically degenerate in this case  [132]. In other words, the skyrmion number in the whole system takes the values from +2+2 to 2-2 depending on initial spin configurations in the simulations. By introducing DD and JJ_{\parallel}, such degeneracy is lifted and the relative relationship of the skyrmion number in each layer is determined; one of the layers takes the skyrmion number of +1+1, while the other takes that of 1-1.

The increase of HH in the AF-SkX region leads to the transition to another ferri-type SkX denoted as the Ferri-SkX II in Fig. 2. Similar to the Ferri-SkX I, one of the layers is characterized by the SkX with the skyrmion number of ±1\pm 1, while the other is characterized by the topologically trivial triple-QQ state without the skyrmion number. The real-space spin and scalar chirality configurations in Fig. 3(d) and the spin structure factor in Fig. 5(d) are also similar to those in the Ferri-SkX I. The difference between them is found in the quantitative value of the scalar chirality, as shown in Fig. 7(e). In the Ferri-SkX II, the scalar chirality in the trivial triple-QQ layer is negligibly small; it takes at most 0.004. On the other hand, the trivial triple-QQ layer in the Ferri-SkX I exhibits relatively large scalar chirality from 0.020.02 to 0.10.1.

The Ferri-SkX II state is replaced by the 3QQ I state with jumps of MηαM^{\alpha}_{\eta}, mη𝑸ναm^{\alpha}_{\eta\bm{Q}_{\nu}}, and χηsc\chi^{\rm sc}_{\eta} by increasing HH, as shown in Figs. 2 and 7. In this state, the spin configurations in the two layers are the same except for the relative phase of the spin density waves, as shown in Fig. 4(a); they are mainly characterized by the single-QQ spiral modulation along the 𝑸3\bm{Q}_{3} direction. As shown in the spin structure factor in Fig. 6(a), the dominant peaks appear at 𝑸3\bm{Q}_{3} in Ssηy(𝒒)S_{s\eta}^{y}(\bm{q}) and Ssηz(𝒒)S_{s\eta}^{z}(\bm{q}), while the sub-dominant peaks appear at 𝑸1\bm{Q}_{1} in Ssηx(𝒒)S_{s\eta}^{x}(\bm{q}) and Ssηz(𝒒)S_{s\eta}^{z}(\bm{q}) and 𝑸2\bm{Q}_{2} in Ssηy(𝒒)S_{s\eta}^{y}(\bm{q}) and Ssηz(𝒒)S_{s\eta}^{z}(\bm{q}). Although this triple-QQ spin texture accompanies the scalar chirality density waves shown in Fig. 4(a), it does not have a uniform component of the spin chirality in Fig. 7(e)

With increasing HH, the intensities at the sub-dominant 𝑸ν\bm{Q}_{\nu} gradually become small, and the 1QQ II state appears. In contrast to the 1QQ I state, the spin configuration is characterized by the spiral on the yzyz plane to gain the energy by the in-plane magnetic field, as shown in Figs. 4(b) and 6(b). This state continuously turns into the fully-polarized state, as shown in Fig. 7.

When JJ_{\parallel} is increased, the Ferri-SkX I, AF-SkX, Ferri-SkX II, and the 3QQ I state vanish, which indicates that the small energy scale of JJ_{\parallel} compared to JJ is important to stabilize these states. For relatively large JJ_{\parallel}, additional three phases appear in the phase diagram in Fig. 2: 3QQ II, 1QQ III, and 1QQ IV states. The HH dependence of spin- and chirality-related quantities in these phases is shown in the case of J=0.4J_{\parallel}=0.4 in Fig. 8.

The 3QQ II state appears in the intermediate-field region, which is obtained by increasing HH in the 1QQ I state for J0.32J_{\parallel}\gtrsim 0.32. The phase transition is of first order with jumps of MηαM^{\alpha}_{\eta}, mη𝑸ναm^{\alpha}_{\eta\bm{Q}_{\nu}}, and χηsc\chi^{\rm sc}_{\eta}; see Fig. 8(a) for example. Similar to the Ferri-SkX I and II, this state consists of layers with different spin configurations accompanying the uniform scalar chirality, as shown in Figs. 4(c) and 8(e). However, there are no skyrmion numbers in both layers. As shown in the spin structure factor in Fig. 6(c), both spin configurations are characterized by the dominant peak at 𝑸3\bm{Q}_{3} and the sub-dominant peaks at 𝑸1\bm{Q}_{1} and 𝑸2\bm{Q}_{2}, but their amplitudes are different for different layers, as shown in Figs. 8(c) and 8(d).

The 3QQ II state is replaced by the 1QQ III state at H1.05H\simeq 1.05. The spin configuration in this state is characterized by the single-QQ spirals along the 𝑸1\bm{Q}_{1} direction, as shown in Figs. 4(d) and 6(d). Compared to the spin configurations for layers A and B, the yy and zz components of the spin align antiparallel, while the xx component of the spin aligns parallel. This state exhibits a staggered magnetization parallel to the yy direction, MAy=MByM^{y}_{\rm A}=-M^{y}_{\rm B}, as shown in Fig. 8(b).

The further increase of HH changes the single-QQ ordering vector from the 𝑸1\bm{Q}_{1} to 𝑸2\bm{Q}_{2} or 𝑸3\bm{Q}_{3} for H1.3H\gtrsim 1.3, which indicates the phase transition to the 1QQ IV state. As shown in the real-space spin configuration in Fig. 4(e) and the momentum-space spin structure factor in Fig. 6(e), this state is similar to the 1QQ II state shown in Figs. 4(b) and 6(b). Meanwhile, the behavior of mη𝑸νzm^{z}_{\eta\bm{Q}_{\nu}} is different from each other: mA𝑸3zmB𝑸3zm^{z}_{{\rm A}\bm{Q}_{3}}\neq m^{z}_{{\rm B}\bm{Q}_{3}} for the 1QQ IV state, while mA𝑸3z=mB𝑸3zm^{z}_{{\rm A}\bm{Q}_{3}}=m^{z}_{{\rm B}\bm{Q}_{3}} for the 1QQ II state, as shown in Figs. 7(d) and 8(d). In addition, the 1QQ IV state has the yy-spin component of the staggered magnetization in Fig. 8(b), while the 1QQ II state has that of the uniform magnetization in Fig. 7(b). When HH is further increased, this state continuously turns into the 1QQ II state. It is noted that there is no almost zz-spin modulation at 𝑸3\bm{Q}_{3} in the 1QQ II state for large JJ_{\parallel}; this spin configuration is almost identified as the single-QQ fan state. Finally, the fully-polarized state appears for H2.85H\gtrsim 2.85.

4 Summary

We have investigated the possibility of generating the AF-SkX without the uniform scalar chirality (skyrmion number) in synthetic layered antiferromagnets. By focusing on the centrosymmetric bilayer structure with the local DM interaction, we find that the AF-SkX can be stabilized in the external magnetic field. In addition, we obtain the instability toward the ferri-type SkX and other multiple-QQ states depending on the model parameters. Our study indicates that the rich topological spin textures emerge from the competing exchange interactions including magnetic anisotropy. The necessary ingredients for the AF-SkX are (1) the staggered DM interaction, (2) the easy-plane single-ion anisotropy, (3) the interlayer exchange interaction, and (4) the in-plane magnetic field.

Based on the above necessary ingredients, let us discuss an important experimental situation to realize the AF-SkX. One is the lattice structure with the staggered DM interaction. As the sublattice-dependent feature of the DM interaction is essential, not only the bilayer systems but also the bulk systems can be a candidate. For example, the hexagonal system, where the space group belongs to P6/mmmP6/mmm and the magnetic sites lie at the 2e2e site, is one of the candidate systems to possess the staggered DM interaction. In addition, a large distance between the different sublattices would be desired, since the small interlay exchange interaction tends to stabilize the AF-SkX. For magnetic ions, heavier elements would be better to make the magnitudes of the easy-plane single-ion anisotropy and the DM interaction larger. Once such a situation is satisfied, one can obtain the AF-SkX when an external magnetic field is applied along the in-plane direction.

Acknowledgements.
This research was supported by JSPS KAKENHI Grants Numbers JP21H01037, JP22H04468, JP22H00101, JP22H01183, and by JST PRESTO (JPMJPR20L8). Parts of the numerical calculations were performed in the supercomputing systems in ISSP, the University of Tokyo.

References

  • [1] C. D. Batista, S.-Z. Lin, S. Hayami, and Y. Kamiya, Rep. Prog. Phys. 79, 084504 (2016).
  • [2] D. Loss and P. M. Goldbart, Phys. Rev. B 45, 13544 (1992).
  • [3] J. Ye, Y. B. Kim, A. J. Millis, B. I. Shraiman, P. Majumdar, and Z. Tešanović, Phys. Rev. Lett. 83, 3737 (1999).
  • [4] F. D. M. Haldane, Phys. Rev. Lett. 93, 206602 (2004).
  • [5] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Rev. Mod. Phys. 82, 1539 (2010).
  • [6] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys. 82, 1959 (2010).
  • [7] K. Ohgushi, S. Murakami, and N. Nagaosa, Phys. Rev. B 62, R6065 (2000).
  • [8] G. Tatara and H. Kawamura, J. Phys. Soc. Jpn. 71, 2613 (2002).
  • [9] R. Shindou and N. Nagaosa, Phys. Rev. Lett. 87, 116801 (2001).
  • [10] I. Martin and C. D. Batista, Phys. Rev. Lett. 101, 156402 (2008).
  • [11] S. Hayami and M. Yatsushiro, Phys. Rev. B 106, 014420 (2022).
  • [12] S. Hayami and M. Yatsushiro, J. Phys. Soc. Jpn. 91, 094704 (2022).
  • [13] I. Žutić, J. Fabian, and S. Das Sarma, Rev. Mod. Phys. 76, 323 (2004).
  • [14] T. Jungwirth, X. Marti, P. Wadley, and J. Wunderlich, Nat. Nanotechnol. 11, 231 (2016).
  • [15] V. Baltz, A. Manchon, M. Tsoi, T. Moriyama, T. Ono, and Y. Tserkovnyak, Rev. Mod. Phys. 90, 015005 (2018).
  • [16] L. Šmejkal, Y. Mokrousov, B. Yan, and A. H. MacDonald, Nat. Phys. 14, 242 (2018).
  • [17] M. B. Jungfleisch, W. Zhang, and A. Hoffmann, Phys. Lett. A 382, 865 (2018).
  • [18] T. H. R. Skyrme, Nucl. Phys. 31, 556 (1962).
  • [19] A. N. Bogdanov and D. A. Yablonskii, Sov. Phys. JETP 68, 101 (1989).
  • [20] A. Bogdanov and A. Hubert, J. Magn. Magn. Mater. 138, 255 (1994).
  • [21] N. Nagaosa and Y. Tokura, Nat. Nanotechnol. 8, 899 (2013).
  • [22] P. Bruno, V. K. Dugaev, and M. Taillefumier, Phys. Rev. Lett. 93, 096806 (2004).
  • [23] A. Neubauer, C. Pfleiderer, B. Binz, A. Rosch, R. Ritz, P. G. Niklowitz, and P. Böni, Phys. Rev. Lett. 102, 186602 (2009).
  • [24] N. Kanazawa, Y. Onose, T. Arima, D. Okuyama, K. Ohoyama, S. Wakimoto, K. Kakurai, S. Ishiwata, and Y. Tokura, Phys. Rev. Lett. 106, 156603 (2011).
  • [25] S. Mühlbauer, B. Binz, F. Jonietz, C. Pfleiderer, A. Rosch, A. Neubauer, R. Georgii, and P. Böni, Science 323, 915 (2009).
  • [26] X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han, Y. Matsui, N. Nagaosa, and Y. Tokura, Nature 465, 901 (2010).
  • [27] X. Z. Yu, N. Kanazawa, Y. Onose, K. Kimoto, W. Zhang, S. Ishiwata, Y. Matsui, and Y. Tokura, Nat. Mater. 10, 106 (2011).
  • [28] S. Seki, X. Z. Yu, S. Ishiwata, and Y. Tokura, Science 336, 198 (2012).
  • [29] T. Adams, A. Chacon, M. Wagner, A. Bauer, G. Brandl, B. Pedersen, H. Berger, P. Lemmens, and C. Pfleiderer, Phys. Rev. Lett. 108, 237204 (2012).
  • [30] S. Seki, J.-H. Kim, D. S. Inosov, R. Georgii, B. Keimer, S. Ishiwata, and Y. Tokura, Phys. Rev. B 85, 220406 (2012).
  • [31] A. K. Nayak, V. Kumar, T. Ma, P. Werner, E. Pippel, R. Sahoo, F. Damay, U. K. Rößler, C. Felser, and S. S. Parkin, Nature 548, 561 (2017).
  • [32] T. Kurumaji, T. Nakajima, V. Ukleev, A. Feoktystov, T.-h. Arima, K. Kakurai, and Y. Tokura, Phys. Rev. Lett. 119, 237201 (2017).
  • [33] T. Kurumaji, T. Nakajima, M. Hirschberger, A. Kikkawa, Y. Yamasaki, H. Sagayama, H. Nakao, Y. Taguchi, T.-h. Arima, and Y. Tokura, Science 365, 914 (2019).
  • [34] 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, Nat. Commun. 10, 5831 (2019).
  • [35] M. Hirschberger, S. Hayami, and Y. Tokura, New J. Phys. 23, 023039 (2021).
  • [36] 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, Nat. Nanotechnol. 15, 444 (2020).
  • [37] Y. Yasui, C. J. Butler, N. D. Khanh, S. Hayami, T. Nomoto, T. Hanaguri, Y. Motome, R. Arita, T. h. Arima, Y. Tokura, and S. Seki, Nat. Commun. 11, 5925 (2020).
  • [38] N. D. Khanh, T. Nakajima, S. Hayami, S. Gao, Y. Yamasaki, H. Sagayama, H. Nakao, R. Takagi, Y. Motome, Y. Tokura, T.-h. Arima, and S. Seki, Adv. Sci. 9, 2105452 (2022).
  • [39] 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, Nat. Commun. 13, 1472 (2022).
  • [40] Y. Tokura and N. Kanazawa, Chem. Rev. 121, 2857 (2021).
  • [41] I. Dzyaloshinsky, J. Phys. Chem. Solids 4, 241 (1958).
  • [42] T. Moriya, Phys. Rev. 120, 91 (1960).
  • [43] U. K. Rößler, A. N. Bogdanov, and C. Pfleiderer, Nature 442, 797 (2006).
  • [44] S. D. Yi, S. Onoda, N. Nagaosa, and J. H. Han, Phys. Rev. B 80, 054416 (2009).
  • [45] S. Hayami and R. Yambe, Phys. Rev. B 105, 224423 (2022).
  • [46] O. I. Utesov, Phys. Rev. B 103, 064414 (2021).
  • [47] O. I. Utesov, Phys. Rev. B 105, 054435 (2022).
  • [48] Q. Tong, F. Liu, J. Xiao, and W. Yao, Nano Lett. 18, 7194 (2018).
  • [49] T. Okubo, S. Chung, and H. Kawamura, Phys. Rev. Lett. 108, 017206 (2012).
  • [50] A. O. Leonov and M. Mostovoy, Nat. Commun. 6, 8275 (2015).
  • [51] S.-Z. Lin and S. Hayami, Phys. Rev. B 93, 064430 (2016).
  • [52] S. Hayami, S.-Z. Lin, and C. D. Batista, Phys. Rev. B 93, 184413 (2016).
  • [53] S. Hayami, S.-Z. Lin, Y. Kamiya, and C. D. Batista, Phys. Rev. B 94, 174420 (2016).
  • [54] S.-Z. Lin and C. D. Batista, Phys. Rev. Lett. 120, 077202 (2018).
  • [55] S. Hayami, J. Magn. Magn. Mater. 553, 169220 (2022).
  • [56] S. Hayami, Phys. Rev. B 105, 174437 (2022).
  • [57] K. Okigami, R. Yambe, and S. Hayami, J. Phys. Soc. Jpn. 91, 103701 (2022).
  • [58] M. A. Ruderman and C. Kittel, Phys. Rev. 96, 99 (1954).
  • [59] T. Kasuya, Prog. Theor. Phys. 16, 45 (1956).
  • [60] K. Yosida, Phys. Rev. 106, 893 (1957).
  • [61] Z. Wang, Y. Su, S.-Z. Lin, and C. D. Batista, Phys. Rev. Lett. 124, 207201 (2020).
  • [62] K. Mitsumoto and H. Kawamura, Phys. Rev. B 104, 184432 (2021).
  • [63] K. Mitsumoto and H. Kawamura, Phys. Rev. B 105, 094427 (2022).
  • [64] K. Kobayashi and S. Hayami, Phys. Rev. B 106, L140406 (2022).
  • [65] S. Heinze, K. von Bergmann, M. Menzel, J. Brede, A. Kubetzka, R. Wiesendanger, G. Bihlmayer, and S. Blügel, Nat. Phys. 7, 713 (2011).
  • [66] R. Ozawa, S. Hayami, and Y. Motome, Phys. Rev. Lett. 118, 147205 (2017).
  • [67] S. Hayami, R. Ozawa, and Y. Motome, Phys. Rev. B 95, 224424 (2017).
  • [68] M. H. Christensen, B. M. Andersen, and P. Kotetes, Phys. Rev. X 8, 041022 (2018).
  • [69] S. Hayami and Y. Motome, Phys. Rev. B 99, 094420 (2019).
  • [70] S. Hayami, J. Magn. Magn. Mater. 513, 167181 (2020).
  • [71] R. Eto and M. Mochizuki, Phys. Rev. B 104, 104425 (2021).
  • [72] P. Nikolić, Phys. Rev. B 103, 155151 (2021).
  • [73] S. Hayami and Y. Motome, J. Phys.: Condens. Matter 33, 443001 (2021).
  • [74] S. Hayami, New J. Phys. 23, 113032 (2021).
  • [75] S. Hayami, T. Okubo, and Y. Motome, Nat. Commun. 12, 6927 (2021).
  • [76] Z. Wang and C. D. Batista, arXiv:2111.13976 , (2021).
  • [77] R. Eto, R. Pohle, and M. Mochizuki, Phys. Rev. Lett. 129, 017201 (2022).
  • [78] S. Hayami, J. Phys. Soc. Jpn. 91, 023705 (2022).
  • [79] S. Hayami and Y. Kato, J. Magn. Magn. Mater. 571, 170547 (2023).
  • [80] S. Hayami and Y. Motome, Phys. Rev. Lett. 121, 137202 (2018).
  • [81] D. Amoroso, P. Barone, and S. Picozzi, Nat. Commun. 11, 5784 (2020).
  • [82] S. Hayami and Y. Motome, Phys. Rev. B 103, 024439 (2021).
  • [83] S. Hayami and Y. Motome, Phys. Rev. B 103, 054422 (2021).
  • [84] R. Yambe and S. Hayami, Sci. Rep. 11, 11184 (2021).
  • [85] Z. Wang, Y. Su, S.-Z. Lin, and C. D. Batista, Phys. Rev. B 103, 104408 (2021).
  • [86] D. Amoroso, P. Barone, and S. Picozzi, Nanomaterials 11, 1873 (2021).
  • [87] S. Hayami and R. Yambe, Phys. Rev. B 105, 104428 (2022).
  • [88] Y. Kato and Y. Motome, Phys. Rev. B 105, 174413 (2022).
  • [89] R. Yambe and S. Hayami, Phys. Rev. B 106, 174437 (2022).
  • [90] D. S. Kathyat, A. Mukherjee, and S. Kumar, Phys. Rev. B 103, 035111 (2021).
  • [91] S. Hayami and R. Yambe, Phys. Rev. B 104, 094425 (2021).
  • [92] S. Hayami and R. Yambe, J. Phys. Soc. Jpn. 90, 073705 (2021).
  • [93] S. Hayami, J. Phys.: Materials , (2022).
  • [94] A. N. Bogdanov, U. K. Rößler, M. Wolf, and K.-H. Müller, Phys. Rev. B 66, 214410 (2002).
  • [95] B. Göbel, I. Mertig, and O. A. Tretiakov, Phys. Rep. 895, 1 (2021).
  • [96] R. Yambe and S. Hayami, Phys. Rev. B 107, 014417 (2023).
  • [97] B. Göbel, A. Mook, J. Henk, and I. Mertig, Phys. Rev. B 96, 060406 (2017).
  • [98] P. M. Buhl, F. Freimuth, S. Blügel, and Y. Mokrousov, Phys. Status Solidi RRL 11, 1700007 (2017).
  • [99] C. A. Akosa, O. A. Tretiakov, G. Tatara, and A. Manchon, Phys. Rev. Lett. 121, 097204 (2018).
  • [100] X. Zhang, Y. Zhou, and M. Ezawa, Sci. Rep. 6, 24795 (2016).
  • [101] J. Barker and O. A. Tretiakov, Phys. Rev. Lett. 116, 147203 (2016).
  • [102] X. Zhang, Y. Zhou, and M. Ezawa, Nat. Commun. 7, 10293 (2016).
  • [103] C. Jin, C. Song, J. Wang, and Q. Liu, Appl. Phys. Lett. 109, 182404 (2016).
  • [104] L. Shen, J. Xia, G. Zhao, X. Zhang, M. Ezawa, O. A. Tretiakov, X. Liu, and Y. Zhou, Phys. Rev. B 98, 134448 (2018).
  • [105] L. Shen, J. Xia, G. Zhao, X. Zhang, M. Ezawa, O. A. Tretiakov, X. Liu, and Y. Zhou, Appl. Phys. Lett. 114, 042402 (2019).
  • [106] Z. Jin, T. T. Liu, W. H. Li, X. M. Zhang, Z. P. Hou, D. Y. Chen, Z. Fan, M. Zeng, X. B. Lu, X. S. Gao, M. H. Qin, and J.-M. Liu, Phys. Rev. B 102, 054419 (2020).
  • [107] L. Shen, X. Li, Y. Zhao, J. Xia, G. Zhao, and Y. Zhou, Phys. Rev. Applied 12, 064033 (2019).
  • [108] P. F. Bessarab, D. Yudin, D. R. Gulevich, P. Wadley, M. Titov, and O. A. Tretiakov, Phys. Rev. B 99, 140411 (2019).
  • [109] V. P. Kravchuk, O. Gomonay, D. D. Sheka, D. R. Rodrigues, K. Everschor-Sitte, J. Sinova, J. van den Brink, and Y. Gaididei, Phys. Rev. B 99, 184429 (2019).
  • [110] R. Zarzuela, S. K. Kim, and Y. Tserkovnyak, Phys. Rev. B 100, 100408 (2019).
  • [111] W. Legrand, D. Maccariello, F. Ajejas, S. Collin, A. Vecchiola, K. Bouzehouane, N. Reyren, V. Cros, and A. Fert, Nat. Mater. 19, 34 (2020).
  • [112] H. D. Rosales, D. C. Cabra, and P. Pujol, Phys. Rev. B 92, 214439 (2015).
  • [113] S. A. Díaz, J. Klinovaja, and D. Loss, Phys. Rev. Lett. 122, 187203 (2019).
  • [114] S. A. Osorio, H. D. Rosales, M. B. Sturla, and D. C. Cabra, Phys. Rev. B 96, 024404 (2017).
  • [115] Z. Liu and H. Ian, Superlattices and Microstructures 126, 25 (2019).
  • [116] S. Gao, H. D. Rosales, F. A. G. Albarracín, V. Tsurkan, G. Kaur, T. Fennell, P. Steffens, M. Boehm, P. Čermák, A. Schneidewind, E. Ressouche, D. C. Cabra, C. Rüegg, and Z. Oksana, Nature 586, 37 (2020).
  • [117] Z. Liu, M. dos Santos Dias, and S. Lounis, J. Phys.: Condens. Matter 32, 425801 (2020).
  • [118] M. Tomé and H. D. Rosales, Phys. Rev. B 103, L020403 (2021).
  • [119] A. Mukherjee, D. S. Kathyat, and S. Kumar, Sci. Rep. 11, 9566 (2021).
  • [120] A. Mukherjee, D. S. Kathyat, and S. Kumar, Phys. Rev. B 103, 134424 (2021).
  • [121] H. D. Rosales, F. A. G. Albarracín, K. Guratinder, V. Tsurkan, L. Prodan, E. Ressouche, and O. Zaharko, Phys. Rev. B 105, 224402 (2022).
  • [122] K.-W. Kim, H.-W. Lee, K.-J. Lee, and M. D. Stiles, Phys. Rev. Lett. 111, 216601 (2013).
  • [123] A. Kundu and S. Zhang, Phys. Rev. B 92, 094434 (2015).
  • [124] S. Hayami, H. Kusunose, and Y. Motome, J. Phys.: Condens. Matter 28, 395601 (2016).
  • [125] We here neglect the contribution from the interaction at higher-harmonic wave vectors, which might play an important role in stabilizing the SkX in the case of the square-lattice system [78].
  • [126] S. Hayami, Phys. Rev. B 105, 014408 (2022).
  • [127] S. Hayami, Phys. Rev. B 105, 184426 (2022).
  • [128] S. Hayami, J. Phys.: Condens. Matter 34, 365802 (2022).
  • [129] S.-Z. Lin, arXiv:2112.12850 , (2021).
  • [130] The color in each pixel is plotted by taking the average of SizS_{i}^{z} for four neighboring spins located at 𝒓i\bm{r}_{i}, 𝒓i+(1,0,0)\bm{r}_{i}+(1,0,0), 𝒓i+(1/2,3/2,0)\bm{r}_{i}+(1/2,\sqrt{3}/2,0), and 𝒓i+(3/2,3/2,0)\bm{r}_{i}+(3/2,\sqrt{3}/2,0) for better visualization, where 𝒓1\bm{r}_{1} represent the position vector at site ii.
  • [131] S. Hayami, Phys. Rev. B 105, 224411 (2022).
  • [132] S. Hayami, Phys. Rev. B 103, 224418 (2021).