Next Article in Journal
A New High-Order Fractional Parallel Iterative Scheme for Solving Nonlinear Equations
Next Article in Special Issue
Exploring the Multiplication of Resonant Modes in Off-Center-Driven Chladni Plates from Maximum Entropy States
Previous Article in Journal
Hesitant Fuzzy Monotonic Dependent OWA Operator and Its Application in Symmetric Group Decision-Making
Previous Article in Special Issue
Searching for Extra Higgs Boson Effects in General Two-Higgs Doublet Model (2HDM)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multipole Moments Under Square Vortex and Skyrmion Crystals

Graduate School of Science, Hokkaido University, Sapporo 060-0810, Japan
Symmetry 2024, 16(11), 1451; https://doi.org/10.3390/sym16111451
Submission received: 16 October 2024 / Revised: 29 October 2024 / Accepted: 31 October 2024 / Published: 1 November 2024
(This article belongs to the Special Issue Feature Papers in 'Physics' Section 2024)

Abstract

:
Non-coplanar spin textures such as magnetic vortices and skyrmions manifest themselves in unusual physical phenomena owing to their topologically nontrivial properties. Here, we investigate emergent multipole moments under vortex and skyrmion crystals in the centrosymmetric tetragonal system. Depending on the vorticity and helicity of the vortex or skyrmion, various multipole moments, including magnetic toroidal and electric toroidal multipoles, are induced on the atomic scale. In particular, the vortex and skyrmion spin textures consisting of multiple spin density waves give rise to density waves in terms of other multipole moments. Our results reveal a close relationship between non-coplanar multiple-Q spin textures and multipole moments.

1. Introduction

The interplay among internal electronic degrees of freedom such as charge, spin, and orbit has attracted much attention in condensed matter physics. For example, coupling between the charge and spin degrees of freedom leads to magnetoresistance [1,2,3,4,5,6,7], multiferroic properties [8,9,10,11,12,13,14,15,16], and the topological Hall/Nernst effect [17,18,19,20,21,22,23,24,25,26,27,28], whereas coupling between the spin and orbital degrees of freedom leads to the spin Hall effect [29,30,31,32,33,34,35,36,37,38,39,40], the Edelstein effect [41,42,43,44,45,46,47,48,49], and noncentrosymmetric superconductivity [50,51,52,53,54,55,56]. Thus, understanding the relationship between microscopic electronic degrees of freedom and macroscopic physical phenomena has been a central issue for years.
The concept of multipoles can connect such microscopic and macroscopic properties, as the symmetry-adapted basis consisting of multipoles, referred to as the symmetry-adapted multipole basis (SAMB), corresponds to a complete basis set in physical space [57,58]. The SAMB is constituted of four multipoles: the electric multipole, characterized by a time-reversal-even polar tensor; the magnetic multipole, characterized by a time-reversal-odd axial tensor; the magnetic toroidal multipole, characterized by a time-reversal-odd polar tensor; and the electric toroidal multipole, characterized by a time-reversal-even axial tensor. Reflecting its completeness, the SAMB can describe various order parameters under complicated electronic, magnetic, and lattice structures, such as chiral charge ordering with the electric toroidal monopole [59,60,61], ferroaxial ordering with the electric toroidal dipole [62,63,64,65,66,67,68,69], and non-coplanar magnetic ordering with the magnetic toroidal octupole [70]. Furthermore, emergent multipole moments can account for various physical properties, such as band-dispersion modulation, cross-correlation, and nonlinear transport from the microscopic viewpoint [71].
In the present study, we focus on the multipole moments induced by multiple-Q spin textures. A multiple-Q state is defined as a superposition of multiple spin density waves [72,73,74,75,76,77]. These states describe various topologically nontrivial spin textures, including vortex and skyrmion crystals [78,79,80,81,82,83,84,85]. Because these multiple-Q states often exhibit non-coplanar spin textures, various types of symmetry reduction are expected; hence, various multipole moments can be induced. To demonstrate this, we consider the double-Q vortex and skyrmion spin textures in the sp orbital model on a two-dimensional square lattice, where the sp orbital basis includes multiple monopolar, dipolar, quadrupolar, and octupolar degrees of freedom in its Hilbert space. We show the multipole moments induced by the vortex and skyrmion spin textures with different vorticities and helicities. Our results provide rich multipole physics brought about by multiple-Q states.
The rest of this paper is organized as follows. In Section 2, we introduce a tight-binding model with s and p orbitals. We list all the multipole degrees of freedom in s-p orbital space. We also introduce the vortex and skyrmion spin textures consisting of a superposition of double-Q spiral waves. Then, we show the multipole moments induced under vortex and skyrmion crystals in Section 3, where both uniform and finite-q components of the multipole moments appear according to the type of vortex or skyrmion crystal. Finally, Section 4 concludes the paper.

2. Model

We start by introducing multipole moments. There are four types of multipoles, according to their spatial inversion ( P ) and time-reversal ( T ) parities: electric multipoles Q l m , with ( P , T ) = [ ( 1 ) l , + 1 ] ; magnetic multipoles M l m , with ( P , T ) = [ ( 1 ) l + 1 , 1 ] ; magnetic toroidal multipoles T l m , with ( P , T ) = [ ( 1 ) l , 1 ] ; and electric toroidal multipoles G l m , with ( P , T ) = [ ( 1 ) l + 1 , + 1 ] , where l and m stand for the rank of the multipole and its component, respectively. Hereafter, we denote the monopole ( l = 0 ) as X 0 , the dipole ( l = 1 ) as ( X x , X y , X z ) , the quadrupole ( l = 2 ) as ( X u , X v , X y z , X z x , X x y ) , and the octupole ( l = 3 ) as ( X x y z , X x α , X y α , X z α , X x β , X y β , X z β ) for X = Q , M , T , G .
When we consider the four orbitals, consisting of one s orbital and three p orbitals, i.e., p x , p y , and p z , the Hilbert space spanned by these four orbitals can be described by an 8 × 8 matrix with the spin degree of freedom; in other words, there are 64 independent matrix elements in the sp hybridized system. By using the multipole operators defined in [86], the correspondence between the multipoles and matrix elements is obtained; see Appendix A for the specific expressions. In the spinless Hilbert space, the 16 active multipoles are as follows: the electric monopoles Q s , 0 and Q p , 0 , electric dipoles Q = ( Q x , Q y , Q z ) , electric quadrupoles ( Q u , Q v , Q y z , Q z x , Q x y ) , magnetic dipoles M = ( M x , M y , M z ) , and magnetic toroidal dipoles T = ( T x , T y , T z ) . In the spinful Hilbert space, the other 48 multipoles with spin dependence are the electric monopole Q 0 ( s ) , the electric dipoles ( Q x ( s ) , Q y ( s ) , Q z ( s ) ) , the electric quadrupoles ( Q u ( s ) , Q v ( s ) , Q y z ( s ) , Q z x ( s ) , Q x y ( s ) ) , the magnetic monopole M 0 ( s ) , the magnetic dipoles ( M s , x ( s ) , M s , y ( s ) , M s , z ( s ) ) , ( M p , x ( s ) , M p , y ( s ) , M p , z ( s ) ) , and ( M a , x ( s ) , M a , y ( s ) , M a , z ( s ) ) , the magnetic quadrupoles ( M u ( s ) , M v ( s ) , M y z ( s ) , M z x ( s ) , M x y ( s ) ) , the magnetic octupoles ( M x y z ( s ) , M x α ( s ) , M y α ( s ) , M z α ( s ) , M x β ( s ) , M y β ( s ) , M z β ( s ) ) , the magnetic toroidal dipoles ( T x ( s ) , T y ( s ) , T z ( s ) ) , the magnetic toroidal quadrupoles ( T u ( s ) , T v ( s ) , T y z ( s ) , T z x ( s ) , T x y ( s ) ) , the electric toroidal monopole G 0 ( s ) , the electric toroidal dipoles ( G x ( s ) , G y ( s ) , G z ( s ) ) , and the electric toroidal quadrupoles ( G u ( s ) , G v ( s ) , G y z ( s ) , G z x ( s ) , G x y ( s ) ) , where the superscript ( s ) denotes a spin-dependent multipole. The active multipoles in the sp hybridized orbital system are summarized in Table 1. It is noteworthy that similar multipoles are also active for different orbitals, such as the pd and df orbitals; the following results for the sp model can be straightforwardly applied to the other orbital cases, as the qualitative results are unaltered by of the choice of orbitals.
We consider a tight-binding model on a two-dimensional square lattice with spatial inversion symmetry under the D 4 h symmetry, as there are no odd-parity multipoles or other unconventional multipole moments in the paramagnetic state, which enables us to investigate the relationship between multipole moments and magnetic structures. The Hamiltonian is provided by
H = i j α α σ ( t i j α α c i α σ c j α σ + H . c . ) + λ 2 i α ˜ α ˜ σ σ c i α ˜ σ H SOC c i α ˜ σ + J i α σ σ c i α σ σ σ σ c i α σ · M i ,
where c i α σ and c i α σ represent the creation and annihilation Fermion operators at site i, orbital α = s , p x , p y , and p z (or α ˜ = p x , p y , and p z ), and spin σ , respectively. The first term represents the kinetic energy of electrons, which includes nearest-neighbor hopping t s between the s orbitals, hopping t x y between the p x and p y orbitals, hopping t z between the p z orbitals, and hopping t s p between the s and ( p x , p y ) orbitals, all of which are taken so as to preserve the symmetry of the tetragonal square-lattice structure. We ignore the orbital-dependent onsite potential for simplicity. The second term represents the atomic spin–orbit coupling for the p orbitals, where H SOC is represented by the 6 × 6 matrix provided by
H SOC = 0 i σ z i σ y i σ z 0 i σ x i σ y i σ x 0 ,
where σ μ is the μ = x , y , z component of the Pauli matrix in spin space. In the following calculations, we set t s = 1 , t x y = 0.7 , t z = 0.4 , t s p = 0.6 , and λ = 0.5 , where t s is set as the energy unit of the model.
The third term represents the effect of magnetic mean fields arising from the vortex and skyrmion spin textures; J = 1 is the coupling constant, and the orbital component is neglected for simplicity. We suppose that the effective site-dependent magnetic field M i is represented in the following forms. For the Néel-type vortex and skyrmion spin textures, M i is represented by
M i = N i cos Q 1 · r i cos Q 2 · r i M ˜ z sin Q 1 · r i sin Q 2 · r i T ,
where the double-Q ordering wave vectors are provided by Q 1 = ( π / 3 , 0 ) and Q 2 = ( 0 , π / 3 ) ; r i is the position vector, N i represents the local constraint to satisfy | M i | = 1 , and M ˜ z is the variational parameter corresponding to a net magnetization. For the vortex (skyrmion) crystal, M ˜ z = 0 ( M ˜ z = 0.5 ) in the following calculations. The real-space spin configurations of the Néel-type vortex and skyrmion crystals are shown in Figure 1a and Figure 1b, respectively. In the vortex crystal in Figure 1a, the skyrmion number is zero owing to the cancellation of the spin scalar chirality between the cores with M z = + 1 and M z = 1 , where the vortex cores are located at the center of the square plaquette [87]. This state is also referred to as a meron–antimeron crystal. Meanwhile, the skyrmion number is 1 in the skyrmion crystal in Figure 1b.
The different types of vortex and skyrmion crystals are obtained by changing the in-plane spin components in Equation (3). The Bloch-type vortex and skyrmion spin textures with different helicities from the Néel-type ones are provided by
M i = N i cos Q 2 · r i cos Q 1 · r i M ˜ z sin Q 1 · r i sin Q 2 · r i T ,
where the skyrmion number is also characterized by 1 owing to the same vorticity. The spin configurations of the Bloch-type vortex and skyrmion crystals are shown in Figure 2a and Figure 2b, respectively. Similarly, we can define the spin textures for two types of anti-type vortex and skyrmion spin textures with skyrmion number + 1 , i.e., with different vorticities from the Néel-type and Bloch-type ones. The spin ansatz of type-I anti-vortex and anti-skyrmion crystals is provided by
M i = N i cos Q 2 · r i cos Q 1 · r i M ˜ z sin Q 1 · r i sin Q 2 · r i T ,
while that of type-II anti-vortex and anti-skyrmion crystals is provided by
M i = N i cos Q 1 · r i cos Q 2 · r i M ˜ z sin Q 1 · r i sin Q 2 · r i T .
The spin configurations corresponding to these anti-type vortex and skyrmion crystals are presented in Figure 2c–f. Furthermore, we can define hybrid vortex and skyrmion crystals which consist of a linear combination of the Néel-type and Bloch-type spin textures. The spin configurations of such hybrid vortex and skyrmion crystals are shown in Figure 2g and Figure 2h, respectively.
From the viewpoint of stability, the double-Q vortex and skyrmion crystals are stabilized when taking into account the effect of the Dzyaloshinskii–Moriya interaction in non-centrosymmetric lattice structures [88,89] or frustrated and multiple-spin interactions in centrosymmetric lattice structures [81,90,91,92]. Especially in the latter mechanism, the Néel-type, Bloch-type, and anti-type vortex and skyrmion crystals are energetically degenerated when the effect of the spin–orbit coupling is ignored [93]. This degeneracy is lifted by the magnetic anisotropy arising from the interplay between the spin–orbit coupling and the crystalline electric field [94]. For example, the magnetic anisotropy under C 4 h symmetry favors hybrid vortex and skyrmion crystals [95]. Double-Q skyrmion crystals have been found in both non-centrosymmetric materials such as Co 10 x / 2 Zn 10 x / 2 Mn x [96,97,98,99,100] and Cu2OSeO3 [101,102] and in centrosymmetric materials such as GdRu2Si2 [103,104,105,106,107].

3. Results

3.1. Néel-Type Vortex Crystal

First, we discuss the result in the case of the Néel-type vortex crystal by setting M ˜ z = 0 in Equation (3). We diagonalize the Hamiltonian in Equation (1) for the N = 6 × 6 spins under the periodic boundary conditions; we also consider the 200 × 200 supercells in order to reduce the finite-size effect. Figure 3 and Figure 4 show the electric-type and magnetic-type multipole structure factors at μ = 0 ( μ represents the chemical potential), respectively. The multipole structure factor is defined as
X ˜ ( q ) = 1 N i j X i X j e i q · ( r i r j ) ,
where X represents all the multipoles activated in the sp hybridized system. Because the multipoles belonging to the same irreducible representation under the D 4 h point group show qualitatively similar profiles in the multipole structure factor, we plot one of the multipole structure factors in each irreducible representation in Figure 3 and Figure 4: the data in Figure 3a–l correspond to those in the A 1 g + , A 2 g + , B 1 g + , B 2 g + , two E g + , A 1 u + , A 2 u + , B 1 u + , B 2 u + , and two E u + irreducible representations, respectively, the data in Figure 4a–l correspond to those in the A 1 g , A 2 g , B 1 g , B 2 g , two E g , A 1 u , A 2 u , B 1 u , B 2 u , and two E u irreducible representations, respectively, and the superscript of the irreducible representation stands for the parity in terms of the time-reversal operations. The correspondence between the irreducible representations and multipoles is summarized in Table 2. In addition, we show the real-space electric-type and magnetic-type multipole configurations in Figure A1 and Figure A2, respectively, in Appendix B for reference.
Under the Néel-type vortex crystal, the magnetic point group reduces from 4 / m m m 1 to 4 m m 1 . It is notable that the time-reversal symmetry is not broken under the Néel-type vortex crystal, as the magnetic texture is invariant for the product of the time-reversal and translational operations. Thus, the uniform component of the magnetic-type multipoles is not induced in the Néel-type vortex crystal. Meanwhile, the symmetries with respect to the spatial inversion, the mirror in the horizontal plane, and the C 2 rotation around the x axis are broken, which indicates that the A 2 u + irreducible representation corresponds to the order parameter. Because the electric dipoles Q z and Q z ( s ) belong to the A 2 u + irreducible representation, they are regarded as the multipole order parameter of the Néel-type vortex crystal. According to symmetry reduction, electric dipole-related physical phenomena such as Rashba-type antisymmetric spin splitting in the band structure [108,109,110,111] and the transverse Edelstein effect [41] are expected.
In addition, we find that finite-q modulations, i.e., density waves, are induced in each multipole channel, as shown in Figure 3 and Figure 4. The nonzero modulations for the wave vector q are understood from the effective coupling between multipoles and spin configurations [112]. For example, the q component of the electric quadrupole Q v x 2 y 2 is coupled to the spins as ( S q 1 x S q 2 x S q 1 y S q 2 y ) δ q 1 + q 2 + q , 0 in the lowest-order approximation, where δ is the Kronecker delta. From Equation (3), the above quantity becomes nonzero for q 1 = Q 1 , q 2 = Q 1 , and q = 2 Q 1 (or q 1 = Q 2 , q 2 = Q 2 , and q = 2 Q 2 ). This is why the intensities at 2 Q 1 and 2 Q 2 appear in Q ˜ v ( q ) . Similarly, the q component of the electric quadrupole Q x y x y is coupled to S q 1 x S q 2 y δ q 1 + q 2 + q , 0 , which leads to the intensities at Q 1 ± Q 2 in Q ˜ x y ( q ) . In this way, the finite-q peak positions in the multipole structure factor are accounted for by the two or higher-order spin-scattering processes.

3.2. Néel-Type Skyrmion Crystal

Next, we consider the Néel-type skyrmion crystal by setting M ˜ z = 0.5 in Equation (3). Figure 5 and Figure 6 represent the electric-type and magnetic-type multipole structure factors, respectively. Owing to the presence of uniform magnetization, the symmetry in the Néel-type skyrmion crystal is further reduced from 4 m m 1 in the Néel-type vortex crystal to 4 m m . In addition to A 2 u + , the A 2 g and A 1 u irreducible representations belong to the identity irreducible representation, as shown in Table 3. In other words, the uniform component of the magnetic monopole M 0 is induced, as shown in Figure 6g. Thus, the longitudinal magnetoelectric effect, which is caused by M 0 , is expected in the Néel-type skyrmion crystal. Compared to the result in the Néel-type vortex crystal, the number of peak positions in the multipole structure factor increases in the Néel-type skyrmion crystal. This is because of the additional coupling between the multipoles and spin textures owing to the q = 0 component. For example, in the case of Q ˜ v ( q ) , the Q 1 and Q 2 components are additionally induced, as ( S q 1 x S q 2 x S q 1 y S q 2 y ) δ q 1 + q 2 + q , 0 becomes nonzero for q 1 = Q 1 + Q 2 , q 2 = Q 2 , and q = Q 1 (or q 1 = Q 1 + Q 2 , q 2 = Q 1 , and q = Q 2 ). It is notable that S Q 1 + Q 2 x and S Q 1 + Q 2 y become nonzero only for M ˜ z 0 , as shown in Figure 4e,f and Figure 6e,f. The emergent multipole moments in the vortex and skyrmion crystal are consistent with the symmetry argument under magnetic point groups [71].

3.3. Other Vortex and Skyrmion Types

In this subsection, we discuss the multipole moments induced by other types of vortex and skyrmion crystals. For the Bloch-type vortex crystal in Figure 2a, the electric toroidal monopole G 0 and electric toroidal quadrupole G u are induced instead of Q z in the Néel-type vortex crystal. Because these multipoles are related to chirality, the Bloch-type vortex crystal exhibits the chirality-related physical phenomena, such as antisymmetric spin splitting in the form of c 1 ( k x σ x + k y σ y ) + c 2 k z σ z ( c 1 and c 2 are numerical coefficients) and the longitudinal Edelstein effect. When the effect of the magnetic field is taken into account, the Bloch-type vortex crystal turns into the Bloch-type skyrmion crystal in Figure 2b, where the magnetic toroidal dipole T z is additionally induced. Thus, the nonreciprocal transport along the z direction and the linear transverse magnetoelectric effect are expected to occur.
In the cases of the type-I and type-II anti-vortex crystals in Figure 2c,e, the rank-2 electric toroidal quadrupoles G v and G x y [113] are induced, which exhibit antisymmetric spin splittings in the functional forms of k x σ x k y σ y and k x σ y + k y σ x , respectively. When the type-I and type-II anti-skyrmion crystals in Figure 2d,f are realized under the external magnetic field, the magnetic quadrupoles M x y and M v are induced, both of which become the origin of the linear magnetoelectric effect, although the tensor components are different from those under M 0 and T z .
Finally, further multipole moments can be induced when the hybrid vortex and skyrmion crystals in Figure 2g,h are considered. In this case, the electric toroidal dipole G z is additionally induced owing to the simultaneous appearance of M 0 and T z , which becomes the origin of the rotational responses between conjugate fields and physical quantities with the same symmetry, such as the spin current generation [68]. A similar situation happens for anti-hybrid vortex (skyrmion) crystals consisting of a superposition of type-I and type-II anti-vortex (anti-skyrmion) crystals. We summarize the induced multipole moments under different types of vortex and skyrmion crystals in Table 4.

4. Conclusions

In this paper, we have investigated the relationship between multipole moments and topologically nontrivial spin textures. By introducing four types of multipoles (electric, magnetic, magnetic toroidal, and electric toroidal), we have shown that different types of multipoles are induced by various types of vortex and skyrmion crystals with different vorticities and helicities under tetragonal symmetry. In addition, we have classified the induced multipole moments in terms of the uniform and finite-q components. Because different multipoles give rise to different physical properties, rich physics can be expected in non-coplanar multiple-q states with various multipole moments, which can stimulate the future exploration of functional magnetic materials.

Funding

This research was supported by JSPS KAKENHI under Grant Numbers JP21H01037, JP22H00101, JP22H01183, JP23H04869, JP23K03288, and JP23K20827, as well as by JST CREST (JPMJCR23O4) and JST FOREST (JPMJFR2366).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The author declares no conflicts of interest.

Appendix A. Matrix Elements of Multipoles

Below, we show the expressions of each multipole operator activated in the sp hybridized system in matrix form [86]. By considering the atomic wave function for the s orbital as
ϕ 0 = 1 4 π
and those for the three p orbitals as
ϕ p x = 3 4 π x r , ϕ p y = 3 4 π y r , ϕ p z = 3 4 π z r ,
we can evaluate the multipole moments one by one.
In the spinless space, the matrix for the 16 multipoles, for which the basis is set for ( ϕ 0 , ϕ x , ϕ y , ϕ z ) , is provided by
Q s , 0 = 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 , Q p , 0 = 0 0 0 0 0 1 0 0 0 0 1 0 0 0 0 1 , Q x = 1 3 0 1 0 0 1 0 0 0 0 0 0 0 0 0 0 0 , Q y = 1 3 0 0 1 0 0 0 0 0 1 0 0 0 0 0 0 0 , Q z = 1 3 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 , Q = ( Q x , Q y , Q z ) , M x = 0 0 0 0 0 0 0 0 0 0 0 i 0 0 i 0 , M y = 0 0 0 0 0 0 0 i 0 0 0 0 0 i 0 0 , M z = 0 0 0 0 0 0 i 0 0 i 0 0 0 0 0 0 , M = ( M x , M y , M z ) , T x = 1 3 3 0 i 0 0 i 0 0 0 0 0 0 0 0 0 0 0 , T y = 1 3 3 0 0 i 0 0 0 0 0 i 0 0 0 0 0 0 0 , T z = 1 3 3 0 0 0 i 0 0 0 0 0 0 0 0 i 0 0 0 , T = ( T x , T y , T z ) , Q u = 1 5 0 0 0 0 0 1 0 0 0 0 1 0 0 0 0 2 , Q v = 3 5 0 0 0 0 0 1 0 0 0 0 1 0 0 0 0 0 , Q y z = 3 5 0 0 0 0 0 0 0 0 0 0 0 1 0 0 1 0 , Q z x = 3 5 0 0 0 0 0 0 0 1 0 0 0 0 0 1 0 0 , Q x y = 3 5 0 0 0 0 0 0 1 0 0 1 0 0 0 0 0 0 .
In the spinful space, the matrix for the other 48 multipoles is obtained by the addition rule between the above spinless multipoles and the Pauli matrix σ = ( σ x , σ y , σ z ) , which is provided by
Q 0 ( s ) = 1 3 ( M x σ x + M y σ y + M z σ z ) ,
Q ( s ) = 1 2 ( σ × T ) ,
Q u ( s ) = 1 6 ( M x σ x M y σ y + 2 M z σ z ) ,
Q v ( s ) = 1 2 ( M x σ x M y σ y ) ,
Q y z ( s ) = 1 2 ( M z σ y + M y σ z ) ,
Q z x ( s ) = 1 2 ( M z σ x + M x σ z ) ,
Q x y ( s ) = 1 2 ( M y σ x + M x σ y ) ,
M 0 ( s ) = 1 3 ( Q x σ x + Q y σ y + Q z σ z ) ,
M s ( s ) = Q s , 0 σ ,
M p ( s ) = Q p , 0 σ ,
M a , x ( s ) = 3 10 1 3 Q u ( ) σ x + Q x y σ y + Q z x σ z ,
M a , y ( s ) = 3 10 Q x y σ x 1 3 Q u ( + ) σ y + Q y z σ z ,
M a , z ( s ) = 3 10 Q z x σ x + Q y z σ y + 2 3 Q u σ z ,
M u ( s ) = 1 6 ( Q x σ x Q y σ y + 2 Q z σ z ) ,
M v ( s ) = 1 2 ( Q x σ x Q y σ y ) ,
M y z ( s ) = 1 2 ( Q z σ y + Q y σ z ) ,
M z x ( s ) = 1 2 ( Q z σ x + Q x σ z ) ,
M x y ( s ) = 1 2 ( Q y σ x + Q x σ y ) ,
M x y z ( s ) = 1 3 ( Q y z σ x + Q z x σ y + Q x y σ z ) ,
M x α ( s ) = 1 5 3 2 Q u ( ) σ x + Q x y σ y + Q z x σ z ,
M y α ( s ) = 1 5 Q x y σ x + 3 2 Q u ( + ) σ y + Q y z σ z ,
M z α ( s ) = 1 5 ( Q z x σ x + Q y z σ y 3 Q u σ z ) ,
M x β ( s ) = 1 3 1 2 Q v ( + ) σ x + Q x y σ y Q z x σ z ,
M y β ( s ) = 1 3 Q x y σ x 1 2 Q v ( ) σ y + Q y z σ z ,
M z β ( s ) = 1 3 ( Q z x σ x Q y z σ y + Q v σ z ) ,
T ( s ) = 1 2 ( σ × Q ) ,
T u ( s ) = 1 2 ( Q y z σ x Q z x σ y ) ,
T v ( s ) = 1 6 ( Q y z σ x + Q z x σ y 2 Q x y σ z ) ,
T y z ( s ) = 1 6 ( Q v ( + ) σ x Q x y σ y + Q z x σ z ) ,
T z x ( s ) = 1 6 ( Q x y σ x Q v ( ) σ y Q y z σ z ) ,
T x y ( s ) = 1 6 ( Q z x σ x + Q y z σ y + 2 Q v σ z ) ,
G 0 ( s ) = 1 3 ( T x σ x + T y σ y + T z σ z ) ,
G ( s ) = 1 2 ( σ × M ) ,
G u ( s ) = 1 6 ( T x σ x T y σ y + 2 T z σ z ) ,
G v ( s ) = 1 2 ( T x σ x T y σ y ) ,
G y z ( s ) = 1 2 ( T z σ y + T y σ z ) ,
G z x ( s ) = 1 2 ( T z σ x + T x σ z ) ,
G x y ( s ) = 1 2 ( T y σ x + T x σ y ) ,
where Q u ( ± ) = Q u ± 3 Q v and Q v ( ± ) = ± 3 Q u + Q v .

Appendix B. Real-Space Multipole Configurations

Below, we show the real-space multipole configurations of the Néel-type vortex crystal and the Néel-type skyrmion crystal. Figure A1 and Figure A2 show the electric-type and magnetic-type multipole configurations of the Néel-type vortex crystal in real space, which correspond to Figure 3 and Figure 4, respectively. Similarly, the electric-type and magnetic-type multipole configurations of the Néel-type skyrmion crystal in real space are shown in Figure A3 and Figure A4, which correspond to Figure 5 and Figure 6, respectively.
Figure A1. The real-space electric-type multipole configurations in the Néel-type vortex crystal for (a) Q u , (b) G z ( s ) , (c) Q v , (d) Q x y , (e) G x ( s ) , (f) G y ( s ) , (g) G u ( s ) , (h) Q z , (i) G v ( s ) , (j) G x y ( s ) , (k) Q x , and (l) Q y .
Figure A1. The real-space electric-type multipole configurations in the Néel-type vortex crystal for (a) Q u , (b) G z ( s ) , (c) Q v , (d) Q x y , (e) G x ( s ) , (f) G y ( s ) , (g) G u ( s ) , (h) Q z , (i) G v ( s ) , (j) G x y ( s ) , (k) Q x , and (l) Q y .
Symmetry 16 01451 g0a1
Figure A2. The real-space magnetic-type multipole configurations in the Néel-type vortex crystal for (a) T u ( s ) , (b) M z , (c) T v ( s ) , (d) T x y ( s ) , (e) M x , (f) M y , (g) M 0 ( s ) , (h) T z ( s ) , (i) M v ( s ) , (j) M x y ( s ) , (k) T x , and (l) T y .
Figure A2. The real-space magnetic-type multipole configurations in the Néel-type vortex crystal for (a) T u ( s ) , (b) M z , (c) T v ( s ) , (d) T x y ( s ) , (e) M x , (f) M y , (g) M 0 ( s ) , (h) T z ( s ) , (i) M v ( s ) , (j) M x y ( s ) , (k) T x , and (l) T y .
Symmetry 16 01451 g0a2
Figure A3. The real-space electric-type multipole configurations in the Néel-type skyrmion crystal for (a) Q u , (b) G z ( s ) , (c) Q v , (d) Q x y , (e) G x ( s ) , (f) G y ( s ) , (g) G u ( s ) , (h) Q z , (i) G v ( s ) , (j) G x y ( s ) , (k) Q x , and (l) Q y .
Figure A3. The real-space electric-type multipole configurations in the Néel-type skyrmion crystal for (a) Q u , (b) G z ( s ) , (c) Q v , (d) Q x y , (e) G x ( s ) , (f) G y ( s ) , (g) G u ( s ) , (h) Q z , (i) G v ( s ) , (j) G x y ( s ) , (k) Q x , and (l) Q y .
Symmetry 16 01451 g0a3
Figure A4. The real-space magnetic-type multipole configurations in the Néel-type skyrmion crystal for (a) T u ( s ) , (b) M z , (c) T v ( s ) , (d) T x y ( s ) , (e) M x , (f) M y , (g) M 0 ( s ) , (h) T z ( s ) , (i) M v ( s ) , (j) M x y ( s ) , (k) T x , and (l) T y .
Figure A4. The real-space magnetic-type multipole configurations in the Néel-type skyrmion crystal for (a) T u ( s ) , (b) M z , (c) T v ( s ) , (d) T x y ( s ) , (e) M x , (f) M y , (g) M 0 ( s ) , (h) T z ( s ) , (i) M v ( s ) , (j) M x y ( s ) , (k) T x , and (l) T y .
Symmetry 16 01451 g0a4

References

  1. Baibich, M.N.; Broto, J.; Fert, A.; Van Dau, F.N.; Petroff, F.; Etienne, P.; Creuzet, G.; Friederich, A.; Chazelas, J. Giant magnetoresistance of (001) Fe/(001) Cr magnetic superlattices. Phys. Rev. Lett. 1988, 61, 2472. [Google Scholar] [CrossRef] [PubMed]
  2. Kusters, R.; Singleton, J.; Keen, D.; McGreevy, R.; Hayes, W. Magnetoresistance measurements on the magnetic semiconductor Nd0.5Pb0.5MnO3. Phys. B Condens. Matter 1989, 155, 362–365. [Google Scholar] [CrossRef]
  3. von Helmolt, R.; Wecker, J.; Holzapfel, B.; Schultz, L.; Samwer, K. Giant negative magnetoresistance in perovskitelike La2/3Ba1/3MnOx ferromagnetic films. Phys. Rev. Lett. 1993, 71, 2331. [Google Scholar] [CrossRef] [PubMed]
  4. Urushibara, A.; Moritomo, Y.; Arima, T.; Asamitsu, A.; Kido, G.; Tokura, Y. Insulator-metal transition and giant magnetoresistance in La1−xSrxMnO3. Phys. Rev. B 1995, 51, 14103. [Google Scholar] [CrossRef]
  5. Ramirez, A. Colossal magnetoresistance. J. Phys. Condens. Matter 1997, 9, 8171. [Google Scholar] [CrossRef]
  6. Tokura, Y.; Tomioka, Y. Colossal magnetoresistive manganites. J. Magn. Magn. Mat. 1999, 200, 1–23. [Google Scholar] [CrossRef]
  7. Tokura, Y. (Ed.) Colossal Magnetoresistive Oxides; Gordon & Breach Science Publisher: London, UK, 1999. [Google Scholar]
  8. Kimura, T.; Kawamoto, S.; Yamada, I.; Azuma, M.; Takano, M.; Tokura, Y. Magnetocapacitance effect in multiferroic BiMnO3. Phys. Rev. B 2003, 67, 180401. [Google Scholar] [CrossRef]
  9. Katsura, H.; Nagaosa, N.; Balatsky, A.V. Spin Current and Magnetoelectric Effect in Noncollinear Magnets. Phys. Rev. Lett. 2005, 95, 057205. [Google Scholar] [CrossRef]
  10. Mostovoy, M. Ferroelectricity in Spiral Magnets. Phys. Rev. Lett. 2006, 96, 067601. [Google Scholar] [CrossRef]
  11. Eerenstein, W.; Mathur, N.; Scott, J.F. Multiferroic and magnetoelectric materials. Nature 2006, 442, 759–765. [Google Scholar] [CrossRef]
  12. Sushkov, A.B.; Aguilar, R.V.; Park, S.; Cheong, S.W.; Drew, H.D. Electromagnons in Multiferroic YMn2O5 and TbMn2O5. Phys. Rev. Lett. 2007, 98, 027202. [Google Scholar] [CrossRef] [PubMed]
  13. Cheong, S.W.; Mostovoy, M. Multiferroics: A magnetic twist for ferroelectricity. Nat. Mater. 2007, 6, 13–20. [Google Scholar] [CrossRef] [PubMed]
  14. Scott, J. Data storage: Multiferroic memories. Nat. Mater. 2007, 6, 256–257. [Google Scholar] [CrossRef]
  15. Ramesh, R.; Spaldin, N.A. Multiferroics: Progress and prospects in thin films. Nat. Mater. 2007, 6, 21. [Google Scholar] [CrossRef] [PubMed]
  16. Tokura, Y.; Seki, S.; Nagaosa, N. Multiferroics of spin origin. Rep. Prog. Phys. 2014, 77, 076501. [Google Scholar] [CrossRef]
  17. Ohgushi, K.; Murakami, S.; Nagaosa, N. Spin anisotropy and quantum Hall effect in the kagomé lattice: Chiral spin state based on a ferromagnet. Phys. Rev. B 2000, 62, R6065–R6068. [Google Scholar] [CrossRef]
  18. Tatara, G.; Kawamura, H. Chirality-driven anomalous Hall effect in weak coupling regime. J. Phys. Soc. Jpn. 2002, 71, 2613–2616. [Google Scholar] [CrossRef]
  19. Neubauer, A.; Pfleiderer, C.; Binz, B.; Rosch, A.; Ritz, R.; Niklowitz, P.G.; Böni, P. Topological Hall Effect in the A Phase of MnSi. Phys. Rev. Lett. 2009, 102, 186602. [Google Scholar] [CrossRef]
  20. Kanazawa, N.; Onose, Y.; Arima, T.; Okuyama, D.; Ohoyama, K.; Wakimoto, S.; Kakurai, K.; Ishiwata, S.; Tokura, Y. Large Topological Hall Effect in a Short-Period Helimagnet MnGe. Phys. Rev. Lett. 2011, 106, 156603. [Google Scholar] [CrossRef]
  21. Shiomi, Y.; Kanazawa, N.; Shibata, K.; Onose, Y.; Tokura, Y. Topological Nernst effect in a three-dimensional skyrmion-lattice phase. Phys. Rev. B 2013, 88, 064409. [Google Scholar] [CrossRef]
  22. Hamamoto, K.; Ezawa, M.; Nagaosa, N. Quantized topological Hall effect in skyrmion crystal. Phys. Rev. B 2015, 92, 115417. [Google Scholar] [CrossRef]
  23. Göbel, B.; Mook, A.; Henk, J.; Mertig, I. Unconventional topological Hall effect in skyrmion crystals caused by the topology of the lattice. Phys. Rev. B 2017, 95, 094413. [Google Scholar] [CrossRef]
  24. Leroux, M.; Stolt, M.J.; Jin, S.; Pete, D.V.; Reichhardt, C.; Maiorov, B. Skyrmion lattice topological Hall effect near room temperature. Sci. Rep. 2018, 8, 15510. [Google Scholar] [CrossRef]
  25. Nakazawa, K.; Bibes, M.; Kohno, H. Topological Hall effect from strong to weak coupling. J. Phys. Soc. Jpn. 2018, 87, 033705. [Google Scholar] [CrossRef]
  26. Kurumaji, T.; Nakajima, T.; Hirschberger, M.; Kikkawa, A.; Yamasaki, Y.; Sagayama, H.; Nakao, H.; Taguchi, Y.; Arima, T.h.; Tokura, Y. Skyrmion lattice with a giant topological Hall effect in a frustrated triangular-lattice magnet. Science 2019, 365, 914–918. [Google Scholar] [CrossRef]
  27. Park, H.; Heinonen, O.; Martin, I. First-principles study of magnetic states and the anomalous Hall conductivity of MNb3S6(M = Co, Fe, Mn, and Ni). Phys. Rev. Mater. 2022, 6, 024201. [Google Scholar] [CrossRef]
  28. Oike, H.; Ebino, T.; Koretsune, T.; Kikkawa, A.; Hirschberger, M.; Taguchi, Y.; Tokura, Y.; Kagawa, F. Topological Nernst effect emerging from real-space gauge field and thermal fluctuations in a magnetic skyrmion lattice. Phys. Rev. B 2022, 106, 214425. [Google Scholar] [CrossRef]
  29. Hirsch, J.E. Spin Hall Effect. Phys. Rev. Lett. 1999, 83, 1834–1837. [Google Scholar] [CrossRef]
  30. Zhang, S. Spin Hall Effect in the Presence of Spin Diffusion. Phys. Rev. Lett. 2000, 85, 393–396. [Google Scholar] [CrossRef]
  31. Murakami, S.; Nagaosa, N.; Zhang, S.C. Dissipationless quantum spin current at room temperature. Science 2003, 301, 1348–1351. [Google Scholar] [CrossRef]
  32. Murakami, S.; Nagaosa, N.; Zhang, S.C. Spin-Hall Insulator. Phys. Rev. Lett. 2004, 93, 156804. [Google Scholar] [CrossRef] [PubMed]
  33. Sinova, J.; Culcer, D.; Niu, Q.; Sinitsyn, N.A.; Jungwirth, T.; MacDonald, A.H. Universal Intrinsic Spin Hall Effect. Phys. Rev. Lett. 2004, 92, 126603. [Google Scholar] [CrossRef] [PubMed]
  34. Kane, C.L.; Mele, E.J. Quantum Spin Hall Effect in Graphene. Phys. Rev. Lett. 2005, 95, 226801. [Google Scholar] [CrossRef] [PubMed]
  35. Bernevig, B.A.; Hughes, T.L.; Zhang, S.C. Quantum Spin Hall Effect and Topological Phase Transition in HgTe Quantum Wells. Science 2006, 314, 1757. [Google Scholar] [CrossRef] [PubMed]
  36. Xu, C.; Moore, J.E. Stability of the quantum spin Hall effect: Effects of interactions, disorder, and Z2 topology. Phys. Rev. B 2006, 73, 045322. [Google Scholar] [CrossRef]
  37. Fujimoto, S. Emergent Nodal Excitations due to Coexistence of Superconductivity and Antiferromagnetism: Cases with and without Inversion Symmetry. J. Phys. Soc. Jpn. 2006, 75, 083704. [Google Scholar] [CrossRef]
  38. Bernevig, B.A.; Zhang, S.C. Quantum Spin Hall Effect. Phys. Rev. Lett. 2006, 96, 106802. [Google Scholar] [CrossRef]
  39. Sinova, J.; Valenzuela, S.O.; Wunderlich, J.; Back, C.H.; Jungwirth, T. Spin Hall effects. Rev. Mod. Phys. 2015, 87, 1213. [Google Scholar] [CrossRef]
  40. Hayami, S.; Yatsushiro, M.; Kusunose, H. Nonlinear spin Hall effect in PT-symmetric collinear magnets. Phys. Rev. B 2022, 106, 024405. [Google Scholar] [CrossRef]
  41. Edelstein, V.M. Spin polarization of conduction electrons induced by electric current in two-dimensional asymmetric electron systems. Solid State Commun. 1990, 73, 233–235. [Google Scholar] [CrossRef]
  42. Yip, S.K. Two-dimensional superconductivity with strong spin-orbit interaction. Phys. Rev. B 2002, 65, 144508. [Google Scholar] [CrossRef]
  43. Fujimoto, S. Magnetoelectric effects in heavy-fermion superconductors without inversion symmetry. Phys. Rev. B 2005, 72, 024515. [Google Scholar] [CrossRef]
  44. Shen, K.; Vignale, G.; Raimondi, R. Microscopic Theory of the Inverse Edelstein Effect. Phys. Rev. Lett. 2014, 112, 096601. [Google Scholar] [CrossRef]
  45. Song, Q.; Zhang, H.; Su, T.; Yuan, W.; Chen, Y.; Xing, W.; Shi, J.; Sun, J.; Han, W. Observation of inverse Edelstein effect in Rashba-split 2DEG between SrTiO3 and LaAlO3 at room temperature. Sci. Adv. 2017, 3, e1602312. [Google Scholar] [CrossRef] [PubMed]
  46. Yoda, T.; Yokoyama, T.; Murakami, S. Orbital Edelstein Effect as a Condensed-Matter Analog of Solenoids. Nano Lett. 2018, 18, 916–920. [Google Scholar] [CrossRef]
  47. Zhou, C.; Liu, Y.P.; Wang, Z.; Ma, S.J.; Jia, M.W.; Wu, R.Q.; Zhou, L.; Zhang, W.; Liu, M.K.; Wu, Y.Z.; et al. Broadband Terahertz Generation via the Interface Inverse Rashba-Edelstein Effect. Phys. Rev. Lett. 2018, 121, 086801. [Google Scholar] [CrossRef]
  48. Massarelli, G.; Wu, B.; Paramekanti, A. Orbital Edelstein effect from density-wave order. Phys. Rev. B 2019, 100, 075136. [Google Scholar] [CrossRef]
  49. Xu, H.; Zhou, J.; Wang, H.; Li, J. Light-induced static magnetization: Nonlinear Edelstein effect. Phys. Rev. B 2021, 103, 205417. [Google Scholar] [CrossRef]
  50. Bauer, E.; Sigrist, M. (Eds.) Non-Centrosymmetric Superconductors: Introduction and Overview (Lecture Notes in Physics), 2012nd ed.; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  51. Yip, S. Noncentrosymmetric superconductors. Annu. Rev. Condens. Matter Phys. 2014, 5, 15–33. [Google Scholar] [CrossRef]
  52. Sigrist, M.; Agterberg, D.F.; Fischer, M.H.; Goryo, J.; Loder, F.; Rhim, S.H.; Maruyama, D.; Yanase, Y.; Yoshida, T.; Youn, S.J. Superconductors with staggered non-centrosymmetricity. J. Phys. Soc. Jpn. 2014, 83, 061014. [Google Scholar] [CrossRef]
  53. Yuan, N.F.Q.; Mak, K.F.; Law, K.T. Possible Topological Superconducting Phases of MoS2. Phys. Rev. Lett. 2014, 113, 097001. [Google Scholar] [CrossRef] [PubMed]
  54. Singh, R.P.; Hillier, A.D.; Mazidian, B.; Quintanilla, J.; Annett, J.F.; Paul, D.M.; Balakrishnan, G.; Lees, M.R. Detection of Time-Reversal Symmetry Breaking in the Noncentrosymmetric Superconductor Re6Zr Using Muon-Spin Spectroscopy. Phys. Rev. Lett. 2014, 112, 107002. [Google Scholar] [CrossRef] [PubMed]
  55. Ali, M.N.; Gibson, Q.D.; Klimczuk, T.; Cava, R.J. Noncentrosymmetric superconductor with a bulk three-dimensional Dirac cone gapped by strong spin-orbit coupling. Phys. Rev. B 2014, 89, 020505. [Google Scholar] [CrossRef]
  56. Barker, J.A.T.; Singh, D.; Thamizhavel, A.; Hillier, A.D.; Lees, M.R.; Balakrishnan, G.; Paul, D.M.; Singh, R.P. Unconventional Superconductivity in La7Ir3 Revealed by Muon Spin Relaxation: Introducing a New Family of Noncentrosymmetric Superconductor That Breaks Time-Reversal Symmetry. Phys. Rev. Lett. 2015, 115, 267001. [Google Scholar] [CrossRef] [PubMed]
  57. Kusunose, H.; Hayami, S. Generalization of microscopic multipoles and cross-correlated phenomena by their orderings. J. Phys. Condens. Matter 2022, 34, 464002. [Google Scholar] [CrossRef]
  58. Hayami, S.; Kusunose, H. Unified description of electronic orderings and cross correlations by complete multipole representation. J. Phys. Soc. Jpn. 2024, 93, 072001. [Google Scholar] [CrossRef]
  59. Kishine, J.i.; Kusunose, H.; Yamamoto, H.M. On the definition of chirality and enantioselective fields. Isr. J. Chem. 2022, 62, e202200049. [Google Scholar] [CrossRef]
  60. Hayami, S.; Kusunose, H. Chiral charge as hidden order parameter in URu2Si2. J. Phys. Soc. Jpn. 2023, 92, 113704. [Google Scholar] [CrossRef]
  61. Inda, A.; Oiwa, R.; Hayami, S.; Yamamoto, H.M.; Kusunose, H. Quantification of chirality based on electric toroidal monopole. J. Chem. Phys. 2024, 160, 184117. [Google Scholar] [CrossRef]
  62. Hlinka, J. Eight Types of Symmetrically Distinct Vectorlike Physical Quantities. Phys. Rev. Lett. 2014, 113, 165502. [Google Scholar] [CrossRef]
  63. Hlinka, J.; Privratska, J.; Ondrejkovic, P.; Janovec, V. Symmetry Guide to Ferroaxial Transitions. Phys. Rev. Lett. 2016, 116, 177602. [Google Scholar] [CrossRef] [PubMed]
  64. Jin, W.; Drueke, E.; Li, S.; Admasu, A.; Owen, R.; Day, M.; Sun, K.; Cheong, S.W.; Zhao, L. Observation of a ferro-rotational order coupled with second-order nonlinear optical fields. Nat. Phys. 2020, 16, 42–46. [Google Scholar] [CrossRef]
  65. Hayashida, T.; Uemura, Y.; Kimura, K.; Matsuoka, S.; Morikawa, D.; Hirose, S.; Tsuda, K.; Hasegawa, T.; Kimura, T. Visualization of ferroaxial domains in an order-disorder type ferroaxial crystal. Nat. Commun. 2020, 11, 4582. [Google Scholar] [CrossRef] [PubMed]
  66. Hayashida, T.; Uemura, Y.; Kimura, K.; Matsuoka, S.; Hagihala, M.; Hirose, S.; Morioka, H.; Hasegawa, T.; Kimura, T. Phase transition and domain formation in ferroaxial crystals. Phys. Rev. Materials 2021, 5, 124409. [Google Scholar] [CrossRef]
  67. Cheong, S.W.; Huang, F.T.; Kim, M. Linking emergent phenomena and broken symmetries through one-dimensional objects and their dot/cross products. Rep. Prog. Phys. 2022, 85, 124501. [Google Scholar] [CrossRef]
  68. Hayami, S.; Oiwa, R.; Kusunose, H. Electric Ferro-Axial Moment as Nanometric Rotator and Source of Longitudinal Spin Current. J. Phys. Soc. Jpn. 2022, 91, 113702. [Google Scholar] [CrossRef]
  69. Nagai, T.; Kimura, T. Chemical Switching of Ferroaxial and Nonferroaxial Structures Based on Second-Order Jahn–Teller Activity in (Na,K)2Hf (BO3)2. Chem. Mater. 2023, 35, 4109–4115. [Google Scholar] [CrossRef]
  70. Hayami, S.; Yatsushiro, M. Nonlinear nonreciprocal transport in antiferromagnets free from spin-orbit coupling. Phys. Rev. B 2022, 106, 014420. [Google Scholar] [CrossRef]
  71. Yatsushiro, M.; Kusunose, H.; Hayami, S. Multipole classification in 122 magnetic point groups for unified understanding of multiferroic responses and transport phenomena. Phys. Rev. B 2021, 104, 054412. [Google Scholar] [CrossRef]
  72. Bak, P.; Lebech, B. “Triple-q→” Modulated Magnetic Structure and Critical Behavior of Neodymium. Phys. Rev. Lett. 1978, 40, 800–803. [Google Scholar] [CrossRef]
  73. McEwen, K.A.; Walker, M.B. Free-energy analysis of the single-q and double-q magnetic structures of neodymium. Phys. Rev. B 1986, 34, 1781–1783. [Google Scholar] [CrossRef] [PubMed]
  74. Zochowski, S.; McEwen, K. Thermal expansion study of the magnetic phase diagram of neodymium. J. Magn. Magn. Mater. 1986, 54, 515–516. [Google Scholar] [CrossRef]
  75. Forgan, E.; Rainford, B.; Lee, S.; Abell, J.; Bi, Y. The magnetic structure of CeAl2 is a non-chiral spiral. J. Phys. Condens. Matter 1990, 2, 10211. [Google Scholar] [CrossRef]
  76. Longfield, M.J.; Paixão, J.A.; Bernhoeft, N.; Lander, G.H. Resonant x-ray scattering from multi-k magnetic structures. Phys. Rev. B 2002, 66, 054417. [Google Scholar] [CrossRef]
  77. Bernhoeft, N.; Paixão, J.A.; Detlefs, C.; Wilkins, S.B.; Javorský, P.; Blackburn, E.; Lander, G.H. Resonant x-ray scattering from UAs0.8Se0.2: Multi-k configurations. Phys. Rev. B 2004, 69, 174415. [Google Scholar] [CrossRef]
  78. Rößler, U.K.; Bogdanov, A.N.; Pfleiderer, C. Spontaneous skyrmion ground states in magnetic metals. Nature 2006, 442, 797–801. [Google Scholar] [CrossRef] [PubMed]
  79. Mühlbauer, S.; Binz, B.; Jonietz, F.; Pfleiderer, C.; Rosch, A.; Neubauer, A.; Georgii, R.; Böni, P. Skyrmion lattice in a chiral magnet. Science 2009, 323, 915–919. [Google Scholar] [CrossRef]
  80. Yu, X.Z.; Onose, Y.; Kanazawa, N.; Park, J.H.; Han, J.H.; Matsui, Y.; Nagaosa, N.; Tokura, Y. Real-space observation of a two-dimensional skyrmion crystal. Nature 2010, 465, 901–904. [Google Scholar] [CrossRef]
  81. Yi, S.D.; Onoda, S.; Nagaosa, N.; Han, J.H. Skyrmions and anomalous Hall effect in a Dzyaloshinskii-Moriya spiral magnet. Phys. Rev. B 2009, 80, 054416. [Google Scholar] [CrossRef]
  82. Binz, B.; Vishwanath, A.; Aji, V. Theory of the Helical Spin Crystal: A Candidate for the Partially Ordered State of MnSi. Phys. Rev. Lett. 2006, 96, 207202. [Google Scholar] [CrossRef]
  83. Binz, B.; Vishwanath, A. Theory of helical spin crystals: Phases, textures, and properties. Phys. Rev. B 2006, 74, 214408. [Google Scholar] [CrossRef]
  84. Nagaosa, N.; Tokura, Y. Topological properties and dynamics of magnetic skyrmions. Nat. Nanotechnol. 2013, 8, 899–911. [Google Scholar] [CrossRef] [PubMed]
  85. Tokura, Y.; Kanazawa, N. Magnetic Skyrmion Materials. Chem. Rev. 2021, 121, 2857. [Google Scholar] [CrossRef]
  86. Kusunose, H.; Oiwa, R.; Hayami, S. Complete Multipole Basis Set for Single-Centered Electron Systems. J. Phys. Soc. Jpn. 2020, 89, 104704. [Google Scholar] [CrossRef]
  87. Hayami, S.; Yambe, R. Locking of skyrmion cores on a centrosymmetric discrete lattice: Onsite versus offsite. Phys. Rev. Res. 2021, 3, 043158. [Google Scholar] [CrossRef]
  88. Dzyaloshinsky, I. A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics. J. Phys. Chem. Solids 1958, 4, 241–255. [Google Scholar] [CrossRef]
  89. Moriya, T. Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 1960, 120, 91. [Google Scholar] [CrossRef]
  90. Hayami, S.; Motome, Y. Néel- and Bloch-Type Magnetic Vortices in Rashba Metals. Phys. Rev. Lett. 2018, 121, 137202. [Google Scholar] [CrossRef] [PubMed]
  91. Utesov, O.I. Thermodynamically stable skyrmion lattice in a tetragonal frustrated antiferromagnet with dipolar interaction. Phys. Rev. B 2021, 103, 064414. [Google Scholar] [CrossRef]
  92. Hayami, S. Multiple skyrmion crystal phases by itinerant frustration in centrosymmetric tetragonal magnets. J. Phys. Soc. Jpn. 2022, 91, 023705. [Google Scholar] [CrossRef]
  93. Okubo, T.; Chung, S.; Kawamura, H. Multiple-q States and the Skyrmion Lattice of the Triangular-Lattice Heisenberg Antiferromagnet under Magnetic Fields. Phys. Rev. Lett. 2012, 108, 017206. [Google Scholar] [CrossRef] [PubMed]
  94. Hayami, S.; Yambe, R. Degeneracy Lifting of Néel, Bloch, and Anti-Skyrmion Crystals in Centrosymmetric Tetragonal Systems. J. Phys. Soc. Jpn. 2020, 89, 103702. [Google Scholar] [CrossRef]
  95. Hayami, S.; Yambe, R. Helicity locking of a square skyrmion crystal in a centrosymmetric lattice system without vertical mirror symmetry. Phys. Rev. B 2022, 105, 104428. [Google Scholar] [CrossRef]
  96. Tokunaga, Y.; Yu, X.; White, J.; Rønnow, H.M.; Morikawa, D.; Taguchi, Y.; Tokura, Y. A new class of chiral materials hosting magnetic skyrmions beyond room temperature. Nat. Commun. 2015, 6, 7638. [Google Scholar] [CrossRef]
  97. Karube, K.; White, J.; Reynolds, N.; Gavilano, J.; Oike, H.; Kikkawa, A.; Kagawa, F.; Tokunaga, Y.; Rønnow, H.M.; Tokura, Y.; et al. Robust metastable skyrmions and their triangular–square lattice structural transition in a high-temperature chiral magnet. Nat. Mater. 2016, 15, 1237. [Google Scholar] [CrossRef]
  98. Karube, K.; White, J.S.; Morikawa, D.; Dewhurst, C.D.; Cubitt, R.; Kikkawa, A.; Yu, X.; Tokunaga, Y.; Arima, T.H.; Rønnow, H.M.; et al. Disordered skyrmion phase stabilized by magnetic frustration in a chiral magnet. Sci. Adv. 2018, 4, eaar7043. [Google Scholar] [CrossRef]
  99. Karube, K.; White, J.S.; Ukleev, V.; Dewhurst, C.D.; Cubitt, R.; Kikkawa, A.; Tokunaga, Y.; Rønnow, H.M.; Tokura, Y.; Taguchi, Y. Metastable skyrmion lattices governed by magnetic disorder and anisotropy in β-Mn-type chiral magnets. Phys. Rev. B 2020, 102, 064408. [Google Scholar] [CrossRef]
  100. Henderson, M.E.; Bleuel, M.; Beare, J.; Cory, D.G.; Heacock, B.; Huber, M.G.; Luke, G.M.; Pula, M.; Sarenac, D.; Sharma, S.; et al. Skyrmion alignment and pinning effects in the disordered multiphase skyrmion material Co8Zn8Mn4. Phys. Rev. B 2022, 106, 094435. [Google Scholar] [CrossRef]
  101. Chacon, A.; Heinen, L.; Halder, M.; Bauer, A.; Simeth, W.; Mühlbauer, S.; Berger, H.; Garst, M.; Rosch, A.; Pfleiderer, C. Observation of two independent skyrmion phases in a chiral magnetic material. Nat. Phys. 2018, 14, 936–941. [Google Scholar] [CrossRef]
  102. Takagi, R.; Yamasaki, Y.; Yokouchi, T.; Ukleev, V.; Yokoyama, Y.; Nakao, H.; Arima, T.; Tokura, Y.; Seki, S. Particle-size dependent structural transformation of skyrmion lattice. Nat. Commun. 2020, 11, 5685. [Google Scholar] [CrossRef]
  103. Khanh, N.D.; Nakajima, T.; Yu, X.; Gao, S.; Shibata, K.; Hirschberger, M.; Yamasaki, Y.; Sagayama, H.; Nakao, H.; Peng, L.; et al. Nanometric square skyrmion lattice in a centrosymmetric tetragonal magnet. Nat. Nanotechnol. 2020, 15, 444. [Google Scholar] [CrossRef] [PubMed]
  104. Khanh, N.D.; Nakajima, T.; Hayami, S.; Gao, S.; Yamasaki, Y.; Sagayama, H.; Nakao, H.; Takagi, R.; Motome, Y.; Tokura, Y.; et al. Zoology of Multiple-Q Spin Textures in a Centrosymmetric Tetragonal Magnet with Itinerant Electrons. Adv. Sci. 2022, 9, 2105452. [Google Scholar] [CrossRef] [PubMed]
  105. Matsuyama, N.; Nomura, T.; Imajo, S.; Nomoto, T.; Arita, R.; Sudo, K.; Kimata, M.; Khanh, N.D.; Takagi, R.; Tokura, Y.; et al. Quantum oscillations in the centrosymmetric skyrmion-hosting magnet GdRu2Si2. Phys. Rev. B 2023, 107, 104421. [Google Scholar] [CrossRef]
  106. Wood, G.D.A.; Khalyavin, D.D.; Mayoh, D.A.; Bouaziz, J.; Hall, A.E.; Holt, S.J.R.; Orlandi, F.; Manuel, P.; Blügel, S.; Staunton, J.B.; et al. Double-Q ground state with topological charge stripes in the centrosymmetric skyrmion candidate GdRu2Si2. Phys. Rev. B 2023, 107, L180402. [Google Scholar] [CrossRef]
  107. Eremeev, S.; Glazkova, D.; Poelchen, G.; Kraiker, A.; Ali, K.; Tarasov, A.V.; Schulz, S.; Kliemt, K.; Chulkov, E.V.; Stolyarov, V.; et al. Insight into the electronic structure of the centrosymmetric skyrmion magnet GdRu2Si2. Nanoscale Adv. 2023, 5, 6678–6687. [Google Scholar] [CrossRef]
  108. Rashba, E.I. Properties of semiconductors with an extremum loop. 1. Cyclotron and combinational resonance in a magnetic field perpendicular to the plane of the loop. Sov. Phys. Solid State 1960, 2, 1109–1122. [Google Scholar]
  109. Ishizaka, K.; Bahramy, M.; Murakawa, H.; Sakano, M.; Shimojima, T.; Sonobe, T.; Koizumi, K.; Shin, S.; Miyahara, H.; Kimura, A.; et al. Giant Rashba-type spin splitting in bulk BiTeI. Nat. Mater. 2011, 10, 521–526. [Google Scholar] [CrossRef]
  110. Bahramy, M.S.; Arita, R.; Nagaosa, N. Origin of giant bulk Rashba splitting: Application to BiTeI. Phys. Rev. B 2011, 84, 041202(R). [Google Scholar] [CrossRef]
  111. Sunko, V.; Rosner, H.; Kushwaha, P.; Khim, S.; Mazzola, F.; Bawden, L.; Clark, O.; Riley, J.; Kasinathan, D.; Haverkort, M.; et al. Maximal Rashba-like spin splitting via kinetic-energy-coupled inversion-symmetry breaking. Nature 2017, 549, 492–496. [Google Scholar] [CrossRef]
  112. Hayami, S.; Motome, Y. Charge density waves in multiple-Q spin states. Phys. Rev. B 2021, 104, 144404. [Google Scholar] [CrossRef]
  113. Hayami, S.; Yanagi, Y.; Kusunose, H.; Motome, Y. Electric Toroidal Quadrupoles in the Spin-Orbit-Coupled Metal Cd2Re2O7. Phys. Rev. Lett. 2019, 122, 147602. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The real-space spin configurations in (a) the Néel-type vortex crystal and (b) the Néel-type skyrmion crystal. The arrows represent the direction of the spin moments while the contour shows the z-spin component M z .
Figure 1. The real-space spin configurations in (a) the Néel-type vortex crystal and (b) the Néel-type skyrmion crystal. The arrows represent the direction of the spin moments while the contour shows the z-spin component M z .
Symmetry 16 01451 g001
Figure 2. The real-space spin configurations in (a) the Bloch-type vortex crystal, (b) Bloch-type skyrmion crystal, (c) type-I anti-vortex crystal, (d) type-I anti-skyrmion crystal, (e) type-II anti-vortex crystal, (f) type-II anti-skyrmion crystal, (g) hybrid vortex crystal, and (h) hybrid skyrmion crystal. The arrows represent the direction of the spin moments, while the contour shows the z-spin component M z .
Figure 2. The real-space spin configurations in (a) the Bloch-type vortex crystal, (b) Bloch-type skyrmion crystal, (c) type-I anti-vortex crystal, (d) type-I anti-skyrmion crystal, (e) type-II anti-vortex crystal, (f) type-II anti-skyrmion crystal, (g) hybrid vortex crystal, and (h) hybrid skyrmion crystal. The arrows represent the direction of the spin moments, while the contour shows the z-spin component M z .
Symmetry 16 01451 g002
Figure 3. The square root of the electric-type multipole structure factor in the Néel-type vortex crystal for (a) Q ˜ u ( q ) , (b) G ˜ z ( s ) ( q ) , (c) Q ˜ v ( q ) , (d) Q ˜ x y ( q ) , (e) G ˜ x ( s ) ( q ) , (f) G ˜ y ( s ) ( q ) , (g) G ˜ u ( s ) ( q ) , (h) Q ˜ z ( q ) , (i) G ˜ v ( s ) ( q ) , (j) G ˜ x y ( s ) ( q ) , (k) Q ˜ x ( q ) , and (l) Q ˜ y ( q ) .
Figure 3. The square root of the electric-type multipole structure factor in the Néel-type vortex crystal for (a) Q ˜ u ( q ) , (b) G ˜ z ( s ) ( q ) , (c) Q ˜ v ( q ) , (d) Q ˜ x y ( q ) , (e) G ˜ x ( s ) ( q ) , (f) G ˜ y ( s ) ( q ) , (g) G ˜ u ( s ) ( q ) , (h) Q ˜ z ( q ) , (i) G ˜ v ( s ) ( q ) , (j) G ˜ x y ( s ) ( q ) , (k) Q ˜ x ( q ) , and (l) Q ˜ y ( q ) .
Symmetry 16 01451 g003
Figure 4. The square root of the magnetic-type multipole structure factor in the Néel-type vortex crystal for (a) T ˜ u ( s ) ( q ) , (b) M ˜ z ( q ) , (c) T ˜ v ( s ) ( q ) , (d) T ˜ x y ( s ) ( q ) , (e) M ˜ x ( q ) , (f) M ˜ y ( q ) , (g) M ˜ 0 ( s ) ( q ) , (h) T ˜ z ( s ) ( q ) , (i) M ˜ v ( s ) ( q ) , (j) M ˜ x y ( s ) ( q ) , (k) T ˜ x ( q ) , and (l) T ˜ y ( q ) .
Figure 4. The square root of the magnetic-type multipole structure factor in the Néel-type vortex crystal for (a) T ˜ u ( s ) ( q ) , (b) M ˜ z ( q ) , (c) T ˜ v ( s ) ( q ) , (d) T ˜ x y ( s ) ( q ) , (e) M ˜ x ( q ) , (f) M ˜ y ( q ) , (g) M ˜ 0 ( s ) ( q ) , (h) T ˜ z ( s ) ( q ) , (i) M ˜ v ( s ) ( q ) , (j) M ˜ x y ( s ) ( q ) , (k) T ˜ x ( q ) , and (l) T ˜ y ( q ) .
Symmetry 16 01451 g004
Figure 5. The square root of the electric-type multipole structure factor in the Néel-type skyrmion crystal for (a) Q ˜ u ( q ) , (b) G ˜ z ( s ) ( q ) , (c) Q ˜ v ( q ) , (d) Q ˜ x y ( q ) , (e) G ˜ x ( s ) ( q ) , (f) G ˜ y ( s ) ( q ) , (g) G ˜ u ( s ) ( q ) , (h) Q ˜ z ( q ) , (i) G ˜ v ( s ) ( q ) , (j) G ˜ x y ( s ) ( q ) , (k) Q ˜ x ( q ) , and (l) Q ˜ y ( q ) .
Figure 5. The square root of the electric-type multipole structure factor in the Néel-type skyrmion crystal for (a) Q ˜ u ( q ) , (b) G ˜ z ( s ) ( q ) , (c) Q ˜ v ( q ) , (d) Q ˜ x y ( q ) , (e) G ˜ x ( s ) ( q ) , (f) G ˜ y ( s ) ( q ) , (g) G ˜ u ( s ) ( q ) , (h) Q ˜ z ( q ) , (i) G ˜ v ( s ) ( q ) , (j) G ˜ x y ( s ) ( q ) , (k) Q ˜ x ( q ) , and (l) Q ˜ y ( q ) .
Symmetry 16 01451 g005
Figure 6. The square root of the magnetic-type multipole structure factor in the Néel-type skyrmion crystal for (a) T ˜ u ( s ) ( q ) , (b) M ˜ z ( q ) , (c) T ˜ v ( s ) ( q ) , (d) T ˜ x y ( s ) ( q ) , (e) M ˜ x ( q ) , (f) M ˜ y ( q ) , (g) M ˜ 0 ( s ) ( q ) , (h) T ˜ z ( s ) ( q ) , (i) M ˜ v ( s ) ( q ) , (j) M ˜ x y ( s ) ( q ) , (k) T ˜ x ( q ) , and (l) T ˜ y ( q ) .
Figure 6. The square root of the magnetic-type multipole structure factor in the Néel-type skyrmion crystal for (a) T ˜ u ( s ) ( q ) , (b) M ˜ z ( q ) , (c) T ˜ v ( s ) ( q ) , (d) T ˜ x y ( s ) ( q ) , (e) M ˜ x ( q ) , (f) M ˜ y ( q ) , (g) M ˜ 0 ( s ) ( q ) , (h) T ˜ z ( s ) ( q ) , (i) M ˜ v ( s ) ( q ) , (j) M ˜ x y ( s ) ( q ) , (k) T ˜ x ( q ) , and (l) T ˜ y ( q ) .
Symmetry 16 01451 g006
Table 1. Classification of multipole moments in the sp hybridized orbital system: the upper panel shows the multipoles activated in spinless space, while the lower panel shows those activated in spinful space. Here, l means the rank of the multipole; see the main text for the details of the multipole notation.
Table 1. Classification of multipole moments in the sp hybridized orbital system: the upper panel shows the multipoles activated in spinless space, while the lower panel shows those activated in spinful space. Here, l means the rank of the multipole; see the main text for the details of the multipole notation.
- Without spin dependence
basis l = 0 l = 1 l = 2
s-s Q s , 0
p-p Q p , 0 M x , M y , M z Q u , Q v , Q y z , Q z x , Q x y
s-p Q x , Q y , Q z
T x , T y , T z
- With spin dependence
basis l = 0 l = 1 l = 2 l = 3
s-s M s , x ( s ) , M s , y ( s ) , M s , z ( s )
p-p Q 0 ( s ) M p , x ( s ) , M p , y ( s ) , M p , z ( s ) Q u ( s ) , Q v ( s ) , Q y z ( s ) , Q z x ( s ) , Q x y ( s ) M x y z ( s ) , M x α ( s ) , M y α ( s ) , M z α ( s ) , M x β ( s ) , M y β ( s ) , M z β ( s )
M a , x ( s ) , M a , y ( s ) , M a , z ( s ) T u ( s ) , T v ( s ) , T y z ( s ) , T z x ( s ) , T x y ( s )
G x ( s ) , G y ( s ) , G z ( s )
s-p G 0 ( s ) Q x ( s ) , Q y ( s ) , Q z ( s ) G u ( s ) , G v ( s ) , G y z ( s ) , G z x ( s ) , G x y ( s )
M 0 ( s ) T x ( s ) , T y ( s ) , T z ( s ) M u ( s ) , M v ( s ) , M y z ( s ) , M z x ( s ) , M x y ( s )
Table 2. Symmetry reduction from the magnetic point group 4 / m m m 1 to 4 m m 1 when the double-Q Néel-type vortex crystal occurs in Equation (3). The correspondence among the irreducible representation (Irrep.), multipole (MP), and main peak positions in the multipole structure factor is presented. The upper (lower) panel represents the electric-type (magnetic-type) multipoles. In the spinful basis, the irreducible representations ( A 1 g , A 2 g , B 1 g , B 2 g , E g , A 1 u , A 2 u , B 1 u , B 2 u , E u ) are replaced by ( E 1 / 2 g , E 1 / 2 g , E 3 / 2 g , E 3 / 2 g , E 1 / 2 g E 3 / 2 g , E 1 / 2 u , E 1 / 2 u , E 3 / 2 u , E 3 / 2 u , E 1 / 2 u E 3 / 2 u ) .
Table 2. Symmetry reduction from the magnetic point group 4 / m m m 1 to 4 m m 1 when the double-Q Néel-type vortex crystal occurs in Equation (3). The correspondence among the irreducible representation (Irrep.), multipole (MP), and main peak positions in the multipole structure factor is presented. The upper (lower) panel represents the electric-type (magnetic-type) multipoles. In the spinful basis, the irreducible representations ( A 1 g , A 2 g , B 1 g , B 2 g , E g , A 1 u , A 2 u , B 1 u , B 2 u , E u ) are replaced by ( E 1 / 2 g , E 1 / 2 g , E 3 / 2 g , E 3 / 2 g , E 1 / 2 g E 3 / 2 g , E 1 / 2 u , E 1 / 2 u , E 3 / 2 u , E 3 / 2 u , E 1 / 2 u E 3 / 2 u ) .
Irrep.MPPeak Positions
A 1 g + A 1 + Q s , 0 , Q p , 0 , Q 0 ( s ) , Q u , Q u ( s ) 0 , Q 1 ± Q 2 , 2 Q 1 , 2
A 2 g + A 2 + G z ( s ) 3 Q 1 ± Q 2 , 3 Q 2 ± Q 1
B 1 g + B 1 + Q v , Q v ( s ) 2 Q 1 , 2
B 2 g + B 2 + Q x y , Q x y ( s ) Q 1 ± Q 2 , 2 Q 1 ± 2 Q 2
E g + E + ( Q y z , Q z x ) , ( Q y z ( s ) , Q z x ( s ) ) , ( G x ( s ) , G y ( s ) ) Q 1 ± Q 2 , 2 Q 1 , 2
A 1 u + A 2 + G 0 ( s ) , G u ( s ) 3 Q 1 ± Q 2 , 3 Q 2 ± Q 1
A 2 u + A 1 + Q z , Q z ( s ) 0 , Q 1 ± Q 2 , 2 Q 1 , 2
B 1 u + B 2 + G v ( s ) Q 1 ± Q 2 , 2 Q 1 ± 2 Q 2
B 2 u + B 1 + G x y ( s ) 2 Q 1 , 2
E u + E + ( Q x , Q y ) , ( Q x ( s ) , Q y ( s ) ) , ( G y z ( s ) , G z x ( s ) ) Q 1 ± Q 2 , 2 Q 1 , 2
A 1 g A 1 T u ( s ) 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
A 2 g A 2 M z , M s , z ( s ) , M p , z ( s ) , M a , z ( s ) , M z α ( s ) Q 1 , 2 , 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
B 1 g B 1 T v ( s ) , M x y z ( s ) 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
B 2 g B 2 T x y ( s ) , M z β ( s ) Q 1 , 2 , 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
E g E ( M x , M y ) , ( M s , x ( s ) , M s , y ( s ) ) , ( M p , x ( s ) , M p , y ( s ) ) , ( M a , x ( s ) , M a , y ( s ) ) , Q 1 , 2 , 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
( T y z ( s ) , T z x ( s ) ) , ( M x α ( s ) , M y α ( s ) ) , ( M x β ( s ) , M y β ( s ) )
A 1 u A 2 M 0 ( s ) , M u ( s ) Q 1 , 2 , 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
A 2 u A 1 T z , T z ( s ) 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
B 1 u B 2 M v ( s ) Q 1 , 2 , 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
B 2 u B 1 M x y ( s ) 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
E u E ( T x , T y ) , ( T x ( s ) , T y ( s ) ) , ( M y z ( s ) , M z x ( s ) ) Q 1 , 2 , 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
Table 3. Symmetry reduction from the magnetic point group 4 / m m m 1 to 4 m m when the double-Q Néel-type skyrmion crystal occurs in Equation (3). The correspondence among the irreducible representation (Irrep.), multipole (MP), and main peak positions in the multipole structure factor is presented. The upper (lower) panel represents the electric-type (magnetic-type) multipoles.
Table 3. Symmetry reduction from the magnetic point group 4 / m m m 1 to 4 m m when the double-Q Néel-type skyrmion crystal occurs in Equation (3). The correspondence among the irreducible representation (Irrep.), multipole (MP), and main peak positions in the multipole structure factor is presented. The upper (lower) panel represents the electric-type (magnetic-type) multipoles.
Irrep.MPPeak Positions
A 1 g + A + Q s , 0 , Q p , 0 , Q 0 ( s ) , Q u , Q u ( s ) 0 , Q 1 , 2 , Q 1 ± Q 2
A 2 g + A G z ( s ) 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
B 1 g + B + Q v , Q v ( s ) Q 1 , 2 , 2 Q 1 , 2
B 2 g + B Q x y , Q x y ( s ) Q 1 ± Q 2 , 2 Q 1 ± 2 Q 2
E g + E ( 1 , 2 ) + ( Q y z , Q z x ) , ( Q y z ( s ) , Q z x ( s ) ) , ( G x ( s ) , G y ( s ) ) Q 1 , 2 , Q 1 ± Q 2
A 1 u + A G 0 ( s ) , G u ( s ) 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
A 2 u + A + Q z , Q z ( s ) 0 , Q 1 , 2 , Q 1 ± Q 2
B 1 u + B G v ( s ) Q 1 ± Q 2 , 2 Q 1 ± 2 Q 2
B 2 u + B + G x y ( s ) Q 1 , 2 , 2 Q 1 , 2
E u + E ( 1 , 2 ) ( Q x , Q y ) , ( Q x ( s ) , Q y ( s ) ) , ( G y z ( s ) , G z x ( s ) ) Q 1 , 2 , Q 1 ± Q 2
A 1 g A T u ( s ) 2 Q 1 ± Q 2 , 2 Q 1 ± Q 2
A 2 g A + M z , M s , z ( s ) , M p , z ( s ) , M a , z ( s ) , M z α ( s ) 0 , Q 1 , 2 , Q 1 ± Q 2
B 1 g B T v ( s ) , M x y z ( s ) Q 1 ± Q 2 , 2 Q 2 ± 2 Q 1
B 2 g B + T x y ( s ) , M z β ( s ) Q 1 , 2 , 2 Q 1 , 2
E g E ( 1 , 2 ) ( M x , M y ) , ( M s , x ( s ) , M s , y ( s ) ) , ( M p , x ( s ) , M p , y ( s ) ) , ( M a , x ( s ) , M a , y ( s ) ) , Q 1 , 2 , Q 1 ± Q 2
( T y z ( s ) , T z x ( s ) ) , ( M x α ( s ) , M y α ( s ) ) , ( M x β ( s ) , M y β ( s ) )
A 1 u A + M 0 ( s ) , M u ( s ) 0 , Q 1 , 2 , Q 1 ± Q 2
A 2 u A T z , T z ( s ) 2 Q 1 ± Q 2 , 2 Q 2 ± Q 1
B 1 u B + M v ( s ) Q 1 , 2 , 2 Q 1 , 2
B 2 u B M x y ( s ) Q 1 ± Q 2 , 2 Q 1 ± 2 Q 2
E u E ( 1 , 2 ) + ( T x , T y ) , ( T x ( s ) , T y ( s ) ) , ( M y z ( s ) , M z x ( s ) ) Q 1 , 2 , Q 1 ± Q 2
Table 4. Uniform component of multipole moments induced by vortex and skyrmion crystals with different vorticities and helicities; H represents the magnetic field and MPG stands for the magnetic point group. The multipoles are classified according to their spatial inversion ( P ) and time-reversal ( T ) parities.
Table 4. Uniform component of multipole moments induced by vortex and skyrmion crystals with different vorticities and helicities; H represents the magnetic field and MPG stands for the magnetic point group. The multipoles are classified according to their spatial inversion ( P ) and time-reversal ( T ) parities.
( P , T )
Type H MPG ( + 1 , + 1 ) ( 1 , + 1 ) ( + 1 , 1 ) ( 1 , 1 )
para = 0 4 / m m m 1 Q 0 , Q u
Bloch = 0 4221 G 0 , G u
0 42 2 G 0 , G u M z , M z α T z
Néel = 0 4 m m 1 Q z
0 4 m m Q z M z , M z α M 0 , M u
anti-I = 0 4 ¯ 2 m 1 G v
0 4 ¯ 2 m G v M z , M z α M x y
anti-II = 0 4 ¯ m 21 G x y
0 4 ¯ m 2 G x y M z , M z α M v
hybrid = 0 41 G z G 0 , Q z , G u
0 4 G z G 0 , Q z , G u M z , T u , M z α M 0 , T z , M u
anti-hybrid = 0 4 ¯ 1 G z G v , G x y
0 4 ¯ G z G v , G x y M z , T u , M z α M v , T x y
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hayami, S. Multipole Moments Under Square Vortex and Skyrmion Crystals. Symmetry 2024, 16, 1451. https://doi.org/10.3390/sym16111451

AMA Style

Hayami S. Multipole Moments Under Square Vortex and Skyrmion Crystals. Symmetry. 2024; 16(11):1451. https://doi.org/10.3390/sym16111451

Chicago/Turabian Style

Hayami, Satoru. 2024. "Multipole Moments Under Square Vortex and Skyrmion Crystals" Symmetry 16, no. 11: 1451. https://doi.org/10.3390/sym16111451

APA Style

Hayami, S. (2024). Multipole Moments Under Square Vortex and Skyrmion Crystals. Symmetry, 16(11), 1451. https://doi.org/10.3390/sym16111451

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

Article Metrics

Back to TopTop