Previous Article in Journal
Ab Initio Investigation on the Magnetic Moments, Magnetocrystalline Anisotropy and Curie Temperature of Fe2P-Based Magnets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Toroidal Monopole in a Single-Site System

Graduate School of Science, Hokkaido University, Sapporo 060-0810, Japan
Magnetism 2025, 5(3), 15; https://doi.org/10.3390/magnetism5030015
Submission received: 9 April 2025 / Revised: 12 June 2025 / Accepted: 23 June 2025 / Published: 25 June 2025

Abstract

A magnetic toroidal monopole, which characterizes time-reversal-odd polar-charge quantity, manifests itself not only in antiferromagnetism but also in time-reversal switching physical responses. We theoretically investigate an atomic-scale description of the magnetic toroidal monopole based on multipole representation theory, which consists of four types of multipoles. We show that the magnetic toroidal monopole degree of freedom is activated as the off-diagonal imaginary hybridization between the single-site orbitals with the same orbital angular momentum but different principal quantum numbers. We demonstrate that the expectation value of the magnetic toroidal monopole becomes nonzero when both electric and magnetic fields are applied to the system.

1. Introduction

Time-reversal symmetry is one of the most fundamental concepts in physics. When the time-reversal symmetry is present, the laws of physics are invariant when the direction of time is reversed. In condensed matter physics, the presence and absence of time-reversal symmetry affects not only microscopic electronic band structures but also macroscopic physical phenomena. For example, energy levels are at least twofold degenerate, which is the so-called Kramers degeneracy, in a time-reversal symmetric nonmagnetic system with half-integer spin. Another example is the robustness of gapless surface states in topological insulators against nonmagnetic disorder when the time-reversal symmetry is preserved [1,2,3,4,5,6].
On the other hand, when the time-reversal symmetry is broken, magnetic properties and phenomena emerge. The most typical example is a ferromagnetic state with a net magnetization, which becomes a source of the anomalous Hall effect owing to a finite Berry curvature in momentum space [7,8,9,10,11,12,13,14,15,16]. Recently, it was recognized that such an anomalous Hall effect can be induced even in antiferromagnets without the net magnetization when a time-reversal symmetry breaking occurs in antiferromagnets with collinear [17,18,19,20,21,22,23,24], coplanar [25,26,27,28,29], and noncoplanar spin textures [30,31,32,33] as found in ferromagnets. When the spatial inversion symmetry is further broken in addition to the time-reversal symmetry, the system exhibits multiferroic phenomena, such as the magnetoelectric effect [34,35,36,37]. Furthermore, the breaking of the time-reversal symmetry leads to unconventional superconductivity, such as chiral p-wave superconductivity [38,39,40,41,42,43,44].
In crystals, the systems usually have other point-group symmetries that preserve crystal structures, such as rotational and mirror symmetries [45]. When the time-reversal symmetry is broken under magnetic phase transitions, some point-group symmetries are often lost simultaneously. For example, in the case of the ferromagnets, the mirror symmetries parallel to the magnetic-moment direction are broken as well as the time-reversal symmetry. With such symmetry breaking in mind, there is a natural question arising as follows: what happens when only the time-reversal symmetry is broken while the other symmetries are preserved? The recent theoretical study has revealed that such a situation occurs when magnetic orderings in the system are characterized by magnetic toroidal monopole ordering, which corresponds to a time-reversal-odd polar-charge quantity [46]. It is noted that the present magnetic toroidal monopole ordering is qualitatively different from the concept of altermagnetism [47], which has been recently recognized as the time-reversal symmetry breaking antiferromagnets. Altermagnetism mainly covers the following two topics: one is the emergence of the anomalous Hall effect in antiferromagnets without a net magnetization, and the other is the emergence of the momentum-dependent spin-split band structure without the relativistic spin–orbit coupling. From the multipole viewpoint, the former is explained by the activation of the magnetic dipole (axial vector quantity), while the latter is explained by the activation of the magnetic toroidal quadrupole (rank-2 polar tensor quantity), magnetic octupole (rank-3 axial tensor quantity), and so on. Meanwhile, the present magnetic toroidal monopole (rank-0 polar scalar) itself exhibits neither the anomalous Hall effect nor momentum-dependent spin-split band structure; the magnetic toroidal monopole does not couple to the momentum with the spin degree of freedom, which means that the magnetic toroidal monopole is not related to the spin splitting. However, under the magnetic toroidal monopole ordering, unconventional time-reversal switching responses are expected, such as magnetic field-induced rotational distortion, electric field-induced spin vortex, and electromagnetic field-induced chirality, which do not occur in conventional ferromagnets. Recently, such an antiferromagnet with a magnetic toroidal monopole has been identified in Co2SiO4, where the electric field-induced nonreciprocal dichroism [48,49] was observed [50]. Although physical phenomena driven by the magnetic toroidal monopole have been observed, their microscopic description is still lacking. In particular, it is unclear when and how such a magnetic toroidal monopole degree of freedom emerges in an atomic-scale wave function.
In the present study, we theoretically investigate the atomic-scale description of the magnetic toroidal monopole based on the multipole representation theory [51,52]. We show that the magnetic toroidal monopole degree of freedom corresponds to the off-diagonal imaginary component between the orbitals with the same orbital angular momentum but different principal quantum numbers. In other words, the magnetic toroidal monopole can be activated as the imaginary hybridization between the 1s–2s orbitals, 2p–3p orbitals, 3d–4d orbitals, 4f–5f orbitals, and so on. We demonstrate that such an imaginary hybridization can occur by applying both electric and magnetic fields in appropriate directions. The present results indicate that the time-reversal-odd polar-charge degree of freedom can be described by the atomic-scale electronic degrees of freedom, which enables us to express all the multipole degrees of freedom by atomic wave functions.
The rest of this paper is organized as follows. In Section 2, we introduce the magnetic toroidal monopole degree of freedom that characterizes the time-reversal-odd polar-charge quantity. Then, after presenting a cluster description of the magnetic toroidal monopole when the system exhibits antiferromagnetic orderings in Section 3, we show a symmetry relationship between the magnetic toroidal monopole and atomic-scale electronic degrees of freedom in Section 4. In Section 5, we demonstrate that the expectation value of the magnetic toroidal monopole becomes nonzero when the symmetry of the system satisfies its emergence condition. Finally, Section 6 is devoted to the conclusion of the present paper.

2. Magnetic Toroidal Monopole

Let us start by introducing the magnetic toroidal dipole, which has been extensively studied in the context of multiferroics [53,54,55,56,57,58,59,60,61]. Since the magnetic toroidal dipole is represented by a vector product between the position vector r i and the spin vector S i (the subscript i stands for the site index), i.e., r i × S i , so as to form the vortex-type spin configuration [Figure 1a], it breaks both the spatial inversion and time-reversal symmetries. Accordingly, intriguing physical phenomena have been explored in both theory and experiments, such as the linear magnetoelectric effect [48,62,63,64,65,66,67,68], asymmetric magnon dispersions [69,70,71,72,73], directional-dependent nonlinear transport [74,75,76,77,78,79], nonlinear spin Hall effect [80,81], and intrinsic nonlinear Hall effect [82,83,84,85]. Various antiferromagnetic materials hosting the magnetic toroidal dipole have been observed in collinear antiferromagnetic structures like Cr2O3 [62,86,87], GaFeO3 [88,89,90,91], LiCoPO4 [92,93,94,95], Ba2CoGe2O7 [96,97,98,99], LiFeSi2O6 [100,101,102], CuMnAs [82,103,104,105,106], BaCoSiO4 [107,108], and PbMn2Ni6Te3O18 [109] and noncollinear antiferromagnetic structures like UNi4B [110,111,112].
The monopole object is generally constructed by taking the inner product of the dipole moment and the position vector. In other words, the magnetic toroidal monopole can be constructed by T · r , where T stands for the magnetic toroidal dipole. The schematic picture of the magnetic toroidal monopole is presented in Figure 1b. Since T and r correspond to the time-reversal-odd and time-reversal-even polar vectors, respectively, the magnetic toroidal monopole is characterized by the time-reversal-odd polar charge. Thus, the emergence of the magnetic toroidal monopole itself breaks the time-reversal symmetry but does not break the point-group symmetry. This indicates that the magnetic toroidal monopole belongs to the identity irreducible representation of the 32 crystal point groups without the time-reversal operation, such as m 3 ¯ m and 4 / m m m [113]. In addition to Co2SiO4, which was experimentally identified as the magnetic toroidal monopole ordering, there are several candidate antiferromagnetic materials hosting the magnetic toroidal monopole, such as KMnF3 [114], Ca2RuO4 [115], MnV2O4 [116], Mn3As2 [117], MnO3 [118], Er2Cu2O5 [119], Ho2Ge2O7 [120], Mn3IrGe [121], ScMnO3 [118], and Mn2FeMoO6 [122]. In these materials, unconventional cross correlations between the physical quantities with different time-reversal parities can be expected, such as magnetic field-induced rotational distortion, electric field-induced spin vortex, and electromagnetic field-induced chirality.

3. Cluster Description

In this section, we show the relationship between antiferromagnetic structures and magnetic toroidal monopole based on cluster multipole representation [28,123,124,125,126]. As discussed in Section 2, there are various antiferromagnetic structures accompanying the magnetic toroidal monopole irrespective of cluster or periodic crystal structures [46]. Although we focus on the atomic-scale magnetic toroidal monopole without antiferromagnetic ordering in the present study, we here present two examples in order to gain a deeper insight into the magnetic toroidal monopole degree of freedom intuitively.
Figure 2a shows a noncollinear spin configuration under the magnetic point group 4 / m with the 8c Wyckoff sites. The vortex-type noncollinear spin configuration gives rise to the local magnetic toroidal dipole; the positive (negative) z component of the magnetic toroidal dipole is induced in the plane HGCD (EFBA), which is denoted by the red arrows in Figure 2a. Since the magnitude of the induced magnetic toroidal dipole for two planes is equivalent, there is no net magnetic toroidal dipole in the whole cluster. Instead, one can easily find the monopole component of the magnetic toroidal dipole distribution, i.e., T · r 0 , where the origin of r is taken at the center of the square prism. Thus, the noncollinear antiferromagnetic structure in Figure 2a is regarded as the magnetic toroidal monopole ordering.
Figure 2b represents another example of the antiferromagnetic structure hosting the magnetic toroidal monopole. The collinear-type magnetic structure under the magnetic point group m m m with the 8g Wyckoff sites satisfies the symmetry of the magnetic toroidal monopole. Indeed, when we calculate the magnetic toroidal dipole on each plaquette in the rectangular parallelepiped, one can find the two-in and two-out structure of the magnetic toroidal dipole, as shown in Figure 2b. It is noted that the distribution of the magnetic toroidal dipole has the monopole component, since the magnitude of the magnetic toroidal dipole on the plane ABCD or EFGH is not equivalent to that on the plane BCGF or ADHE, owing to the rectangular symmetry. In other words, the distribution of the magnetic toroidal dipole is decomposed into the monopole and quadrupole components. Thus, the collinear magnetic structure in Figure 2b is also regarded as the magnetic toroidal monopole ordering, which exhibits physical phenomena driven by the magnetic toroidal monopole. A similar discussion holds for periodic systems as long as the same magnetic symmetry is considered. In this way, cluster antiferromagnetic structures consisting of multiple sites can host the magnetic toroidal monopole.

4. Atomic-Scale Description

In order to discuss the possibility of the atomic-scale magnetic toroidal monopole, we introduce the atomic-scale multipole operators consisting of electric, magnetic, magnetic toroidal, and electric toroidal multipoles, where each multipole is characterized by two quantum numbers as follows: the orbital angular momentum l and its z component m ( l m l ) [127]. The four types of multipoles are characterized by different spatial inversion and time-reversal parities; electric (electric toroidal) multipoles are characterized by the time-reversal-even polar (axial) tensor, while magnetic toroidal (magnetic) multipoles are characterized by the time-reversal-odd polar (axial) tensor.
From the microscopic viewpoint, whether multipoles are activated in the physical space is determined by the addition rule of the angular momenta in terms of the wave function. When the active multipoles are considered for the physical space spanned by two wave functions with the orbital angular momenta L 1 and L 2 , only the rank-l multipoles satisfying | L 1 L 2 | l L 1 + L 2 are active. By taking the real wave function as the basis for s, p, d, and f orbitals, here and hereafter, the electric and electric toroidal multipoles correspond to the real-number matrix, while the magnetic and magnetic toroidal multipoles correspond to the imaginary-number matrix. For example, in the s-p orbital system without the spin degree of freedom, the active multipoles are a rank-1 electric dipole and a rank-1 magnetic toroidal dipole. Such a correspondence between the active multipoles and atomic s, p, d, and f wave functions is summarized in Ref. [52], where it was shown that all the multipoles except for the magnetic toroidal monopole can become active in the atomic scale.
In the present study, we focus on the multipole degrees of freedom between the wave functions with the same orbital angular momentum but different principal quantum numbers n. In other words, we consider the multipole degrees of freedom in the physical space spanned by 1s–2s orbitals, 2p–3p orbitals, 3d–4d orbitals, 4f–5f orbitals, and so on. By adopting two orbitals with different principal quantum numbers, one can find another electronic degree of freedom, which corresponds to the imaginary off-diagonal matrix elements. Indeed, such a degree of freedom can express the magnetic toroidal monopole degree of freedom, as detailed below.
Let us consider the active multipole degrees of freedom in the 1s–2s orbitals. Since there is only one degree of freedom in the s orbital, there are one real and one imaginary matrix elements in the physical space spanned by the 1 s and 2 s orbitals. When we adopt the real wave function, the real (imaginary) matrix component corresponds to the time-reversal-even (time-reversal-odd) multipoles from the time-reversal operation. In addition, the rank of the active multipoles must be zero because both 1 s and 2 s orbitals are characterized by zero orbital angular momentum from the addition rule of the angular momenta. Thus, one identifies that the off-diagonal real matrix element between the 1s and 2s orbitals corresponds to the electric monopole, while the off-diagonal imaginary matrix element corresponds to the magnetic toroidal monopole. These results indicate that the magnetic toroidal monopole can be described by the atomic-scale wave functions invoking the imaginary hybridization between the s orbitals with different principal quantum numbers.
Similarly, the magnetic toroidal monopole can be active for the other orbitals with different principal quantum numbers, such as p, d, and f orbitals. In these orbitals with nonzero orbital angular momenta, unconventional higher-rank multipole degrees of freedom are also active, which include the rank-1 electric toroidal dipole related to the ferroaxial (ferrorotational) ordering [128,129,130,131,132,133,134] and rank-2 magnetic toroidal quadrupole. All the active multipoles in these atomic wave functions are summarized in Table 1.

5. Model Analysis

We demonstrate that the expectation values of the magnetic toroidal monopole become nonzero by performing the microscopic model calculations. We introduce the lattice model, where each atom is loosely connected by effective exchange interactions. Each site consists of 2 s - and 3 s -orbital wave functions ( ψ s 1 , ψ s 2 ) and three 2 p -orbital wave functions ( ψ x , ψ y , ψ z ) . The model Hamiltonian within the single-site unit cell is given by
H = Δ 1 Q 0 p + Δ 2 Q 0 s 2 + λ L · S h Q z Q z h T z T z h Q z Q z h T z T z ,
where ( Q 0 p , Q 0 s 2 , L , S , Q z , T z , Q z , T z ) are 10 × 10 matrices, which are obtained by the formula in Ref. [127]. The first and second terms represent the atomic energy levels, which arise from the different atomic energies of s and p orbitals. The first term stands for the atomic energy level difference between the 2s and 2p orbitals by Δ 1 , while the second term stands for that between the 2s and 3s orbitals by Δ 2 . The expressions of Q 0 p and Q 0 s 2 are explicitly given by
Q 0 p = 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 ,
Q 0 s 2 = 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 ,
where the basis is represented by { ψ s 1 , ψ x , ψ y , ψ z , ψ s 2 , ψ s 1 , ψ x , ψ y , ψ z , ψ s 2 } . We set Δ 1 = 1 and Δ 2 = 2 . The second term represents the relativistic spin–orbit coupling, which works on the p orbital; L and S stand for the orbital and spin angular momentum operators of the p orbital, respectively. We set λ = 0.3 .
The fourth to seventh terms are introduced to induce the magnetic toroidal monopole. Since the magnetic toroidal monopole is related to T · r , as discussed in Section 2, we consider the molecular fields corresponding to T and r in an atomic scale. One is the electric dipole field corresponding to r in the fourth and sixth terms, where Q z and Q z are given by
Q z = 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 ,
Q z = 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 1 3 0 .
Thus, the fourth and sixth terms describe the real hybridization between 2 s 2 p z and 3 s 2 p z orbitals, respectively. We set the different molecular fields for the 2 s 2 p z orbital space and 3 s 2 p z orbital space, h Q z and h Q z , respectively. Such an electric dipole degree of freedom is activated by the polar crystalline electric field or the electric field. Meanwhile, the fifth and seventh terms mimic the effect of the magnetic toroidal dipole field corresponding to T , whose matrices are given by
T z = 0 0 0 i 3 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 3 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 3 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 3 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 ,
T z = 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 3 3 0 0 0 0 0 0 0 0 i 3 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 3 3 0 0 0 0 0 0 0 0 i 3 3 0 .
These matrices indicate that the fifth and seventh terms describe the imaginary hybridization between 2 s 2 p z and 3 s 2 p z orbitals, respectively. We set the different molecular fields for the 2 s 2 p z orbital space and 3 s 2 p z orbital space, h T z and h T z , respectively. The model in Equation (1) possesses neither time-reversal symmetry nor product symmetry of spatial inversion and time-reversal symmetries so that the magnetic toroidal monopole can be active. Since the Hamiltonian is hermitian, all the eigenvalues of the Hamiltonian are real.
Figure 3a represents the h T z dependence of the expectation values in terms of relevant multipoles as follows: the magnetic toroidal monopole T 0 and two magnetic toroidal dipoles T z and T z . Here, the magnetic toroidal monopole is defined as the imaginary hybridization between 2s and 3 s orbitals, whose matrix is given by
T 0 = 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 .
The symbol stands for the expectation values at the temperature T = 0.1 . The other model parameters are taken at h Q z = h Q z = 0.2 and h T z = 0.1 .
The results in Figure 3a show that the magnetic toroidal monopole is induced irrespective of h T z for nonzero h Q z , h Q z , and h T z , which means that the atomic-scale magnetic toroidal monopole becomes the atomic-scale order parameter once the irreducible representation of the magnetic toroidal monopole corresponds to the identity irreducible representation. In order to understand the microscopic essence to induce the magnetic toroidal monopole, we perform an expansion method to T 0 [135,136]. As a result, we find that T 0 is proportional to ( h Q z h T z h Q z h T z ) , which means that the spin–orbit coupling λ and atomic energy levels Δ 1 and Δ 2 play no significant role in inducing the magnetic toroidal monopole, whereas both the electric dipole and magnetic toroidal dipole fields are important.
Next, we show that the magnetic toroidal monopole can be induced by applying both external electric and magnetic fields without the magnetic toroidal dipole field. We consider the model Hamiltonian as follows:
H = Δ 1 Q 0 p + Δ 2 Q 0 s 2 + λ L · S h Q z Q z h Q x Q x h Q z Q z h Q x Q x h M y M y ,
where the first to third terms are the same as those in the model in Equation (1). Instead of the magnetic toroidal dipole field in the fifth and seventh terms in Equation (1), we introduce two different dipole fields. One is the x component of the electric dipole field appearing in the fifth and seventh terms, where the matrix forms of Q x and Q x are given by
Q x = 0 1 3 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 ,
Q x = 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 3 0 0 0 .
This indicates the hybridization between s and p x orbitals. The other is the magnetic dipole field appearing in the last term, where M y is given by
M y = 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 i 0 0 0 0 0 0 0 0 0 0 0 0 0 ,
where we neglect the effect of the magnetic field on the spin moments for simplicity. Thus, we consider the situation where the electric field is applied in the x z plane and the magnetic field is applied along the y direction. In the following, we parameterize h Q z = 0.2 cos θ , h Q x = 0.1 sin θ , h Q z = 0.2 cos θ , and h Q x = 0.05 sin θ .
Figure 3b represents the θ dependence of T 0 , T z , and T z , where T 0 becomes nonzero except for θ = 0 , π / 2 , and π . The disappearance of the magnetic toroidal monopole at θ = 0 , π / 2 , and π is due to the absence of x or z components of the electric dipole field. This means that the application of the electric field along the high-symmetry directions, such as the x and z directions, is not enough to induce the magnetic toroidal monopole. Such behavior is qualitatively understood from the fact that the magnetic toroidal monopole is induced when both the electric dipole and magnetic toroidal dipole are induced, as demonstrated in Figure 3a; the z component of the magnetic toroidal dipole in Equation (1) corresponds to the product of the x component of the electric dipole and the y component of the magnetic dipole in Equation (9). Indeed, the essential model parameters for T 0 in the present model, which are obtained by the expansion method [135,136], are given by h M y ( h Q z h Q x h Q z h Q x ) ; the two components of the electric field as well as the magnetic field perpendicular to the electric field are necessary.

6. Conclusions

In conclusion, we have investigated the possibility of the atomic-scale magnetic toroidal monopole. Based on the symmetry, multipole representation, and model analyses, we have shown that the atomic-scale magnetic toroidal monopole corresponds to the off-diagonal imaginary hybridization between the orbitals with the same orbital angular momentum but different principal quantum numbers. The present results indicate that the atomic-scale magnetic toroidal monopole can be a potential order parameter in antiferromagnets for the magnetic phase transitions that break only the time-reversal symmetry.
Let us comment on the active magnetic toroidal monopole for other degrees of freedom. The magnetic toroidal monopole can also be activated through the bond degree of freedom as the imaginary hybridization between the different sites with the same orbital, depending on the spatial distribution of the imaginary hybridizations among multiple sites. In such a situation, since the imaginary hybridization occurs within the same orbital, e.g., 1s–1s orbital, the larger hybridization effect compared to the atomic-scale case might be expected.

Funding

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

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Kane, C.L.; Mele, E.J. Z2. Phys. Rev. Lett. 2005, 95, 146802. [Google Scholar] [CrossRef] [PubMed]
  2. Moore, J.E. The birth of topological insulators. Nature 2010, 464, 194–198. [Google Scholar] [CrossRef]
  3. Hasan, M.Z.; Kane, C.L. Colloquium: Topological insulators. Rev. Mod. Phys. 2010, 82, 3045–3067. [Google Scholar] [CrossRef]
  4. Hasan, M.Z.; Moore, J.E. Three-dimensional topological insulators. Annu. Rev. Condens. Matter Phys. 2011, 2, 55–78. [Google Scholar] [CrossRef]
  5. Qi, X.L.; Zhang, S.C. Topological insulators and superconductors. Rev. Mod. Phys. 2011, 83, 1057. [Google Scholar] [CrossRef]
  6. Senthil, T. Symmetry-protected topological phases of quantum matter. Annu. Rev. Condens. Matter Phys. 2015, 6, 299–324. [Google Scholar] [CrossRef]
  7. Karplus, R.; Luttinger, J.M. Hall Effect in Ferromagnetics. Phys. Rev. 1954, 95, 1154–1160. [Google Scholar] [CrossRef]
  8. Smit, J. The spontaneous Hall effect in ferromagnetics II. Physica 1958, 24, 39–51. [Google Scholar] [CrossRef]
  9. Maranzana, F.E. Contributions to the Theory of the Anomalous Hall Effect in Ferro- and Antiferromagnetic Materials. Phys. Rev. 1967, 160, 421–429. [Google Scholar] [CrossRef]
  10. Berger, L. Side-Jump Mechanism for the Hall Effect of Ferromagnets. Phys. Rev. B 1970, 2, 4559–4566. [Google Scholar] [CrossRef]
  11. Nozieres, P.; Lewiner, C. A simple theory of the anomalous Hall effect in semiconductors. J. Phys. 1973, 34, 901–915. [Google Scholar] [CrossRef]
  12. Ye, J.; Kim, Y.B.; Millis, A.J.; Shraiman, B.I.; Majumdar, P.; Tešanović, Z. Berry Phase Theory of the Anomalous Hall Effect: Application to Colossal Magnetoresistance Manganites. Phys. Rev. Lett. 1999, 83, 3737–3740. [Google Scholar] [CrossRef]
  13. Jungwirth, T.; Niu, Q.; MacDonald, A.H. Anomalous Hall Effect in Ferromagnetic Semiconductors. Phys. Rev. Lett. 2002, 88, 207208. [Google Scholar] [CrossRef]
  14. Gosálbez-Martínez, D.; Souza, I.; Vanderbilt, D. Chiral degeneracies and Fermi-surface Chern numbers in bcc Fe. Phys. Rev. B 2015, 92, 085138. [Google Scholar] [CrossRef]
  15. Nagaosa, N.; Sinova, J.; Onoda, S.; MacDonald, A.H.; Ong, N.P. Anomalous Hall effect. Rev. Mod. Phys. 2010, 82, 1539–1592. [Google Scholar] [CrossRef]
  16. Xiao, D.; Chang, M.C.; Niu, Q. Berry phase effects on electronic properties. Rev. Mod. Phys. 2010, 82, 1959–2007. [Google Scholar] [CrossRef]
  17. Solovyev, I.V. Magneto-optical effect in the weak ferromagnets LaMO3 (M = Cr, Mn, and Fe). Phys. Rev. B 1997, 55, 8060–8063. [Google Scholar] [CrossRef]
  18. Sivadas, N.; Okamoto, S.; Xiao, D. Gate-Controllable Magneto-optic Kerr Effect in Layered Collinear Antiferromagnets. Phys. Rev. Lett. 2016, 117, 267203. [Google Scholar] [CrossRef]
  19. Li, X.; MacDonald, A.H.; Chen, H. Quantum Anomalous Hall Effect through Canted Antiferromagnetism. arXiv 2019, arXiv:1902.10650. [Google Scholar]
  20. Šmejkal, L.; González-Hernández, R.; Jungwirth, T.; Sinova, J. Crystal time-reversal symmetry breaking and spontaneous Hall effect in collinear antiferromagnets. Sci. Adv. 2020, 6, eaaz8809. [Google Scholar] [CrossRef]
  21. Samanta, K.; Ležaić, M.; Merte, M.; Freimuth, F.; Blügel, S.; Mokrousov, Y. Crystal Hall and crystal magneto-optical effect in thin films of SrRuO3. J. Appl. Phys. 2020, 127, 213904. [Google Scholar] [CrossRef]
  22. Naka, M.; Hayami, S.; Kusunose, H.; Yanagi, Y.; Motome, Y.; Seo, H. Anomalous Hall effect in κ-type organic antiferromagnets. Phys. Rev. B 2020, 102, 075112. [Google Scholar] [CrossRef]
  23. Hayami, S.; Kusunose, H. Essential role of the anisotropic magnetic dipole in the anomalous Hall effect. Phys. Rev. B 2021, 103, L180407. [Google Scholar] [CrossRef]
  24. Feng, Z.; Zhou, X.; Šmejkal, L.; Wu, L.; Zhu, Z.; Guo, H.; González-Hernández, R.; Wang, X.; Yan, H.; Qin, P.; et al. An anomalous Hall effect in altermagnetic ruthenium dioxide. Nat. Electron. 2022, 5, 735–743. [Google Scholar] [CrossRef]
  25. Tomizawa, T.; Kontani, H. Anomalous Hall effect in the t2g orbital kagome lattice due to noncollinearity: Significance of the orbital Aharonov-Bohm effect. Phys. Rev. B 2009, 80, 100401. [Google Scholar] [CrossRef]
  26. Chen, H.; Niu, Q.; MacDonald, A.H. Anomalous Hall Effect Arising from Noncollinear Antiferromagnetism. Phys. Rev. Lett. 2014, 112, 017205. [Google Scholar] [CrossRef] [PubMed]
  27. Nakatsuji, S.; Kiyohara, N.; Higo, T. Large anomalous Hall effect in a non-collinear antiferromagnet at room temperature. Nature 2015, 527, 212. [Google Scholar] [CrossRef]
  28. Suzuki, M.T.; Koretsune, T.; Ochi, M.; Arita, R. Cluster multipole theory for anomalous Hall effect in antiferromagnets. Phys. Rev. B 2017, 95, 094406. [Google Scholar] [CrossRef]
  29. Chen, H.; Wang, T.C.; Xiao, D.; Guo, G.Y.; Niu, Q.; MacDonald, A.H. Manipulating anomalous Hall antiferromagnets with magnetic fields. Phys. Rev. B 2020, 101, 104418. [Google Scholar] [CrossRef]
  30. 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]
  31. Shindou, R.; Nagaosa, N. Orbital Ferromagnetism and Anomalous Hall Effect in Antiferromagnets on the Distorted fcc Lattice. Phys. Rev. Lett. 2001, 87, 116801. [Google Scholar] [CrossRef] [PubMed]
  32. Taguchi, Y.; Oohara, Y.; Yoshizawa, H.; Nagaosa, N.; Tokura, Y. Spin chirality, Berry phase, and anomalous Hall effect in a frustrated ferromagnet. Science 2001, 291, 2573–2576. [Google Scholar] [CrossRef]
  33. 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]
  34. Fiebig, M. Revival of the magnetoelectric effect. J. Phys. D Appl. Phys. 2005, 38, R123. [Google Scholar] [CrossRef]
  35. Katsura, H.; Nagaosa, N.; Balatsky, A.V. Spin Current and Magnetoelectric Effect in Noncollinear Magnets. Phys. Rev. Lett. 2005, 95, 057205. [Google Scholar] [CrossRef] [PubMed]
  36. Mostovoy, M. Ferroelectricity in Spiral Magnets. Phys. Rev. Lett. 2006, 96, 067601. [Google Scholar] [CrossRef]
  37. Tokura, Y.; Seki, S.; Nagaosa, N. Multiferroics of spin origin. Rep. Prog. Phys. 2014, 77, 076501. [Google Scholar] [CrossRef] [PubMed]
  38. Ott, H.R.; Rudigier, H.; Rice, T.M.; Ueda, K.; Fisk, Z.; Smith, J.L. p-Wave Superconductivity in UBe13. Phys. Rev. Lett. 1984, 52, 1915–1918. [Google Scholar] [CrossRef]
  39. Sigrist, M.; Rice, T.M. Phenomenological theory of the superconductivity phase diagram of U1−xThxBe13. Phys. Rev. B 1989, 39, 2200–2216. [Google Scholar] [CrossRef]
  40. Ghosh, H. Time-reversal symmetry-breaking superconductivity. Phys. Rev. B 1999, 59, 3357–3360. [Google Scholar] [CrossRef]
  41. Sigrist, M. Time-reversal symmetry breaking states in high-temperature superconductors. Prog. Theor. Phys. 1998, 99, 899–929. [Google Scholar] [CrossRef]
  42. Mackenzie, A.P.; Maeno, Y. The superconductivity of Sr2RuO4 and the physics of spin-triplet pairing. Rev. Mod. Phys. 2003, 75, 657–712. [Google Scholar] [CrossRef]
  43. Schemm, E.; Gannon, W.; Wishne, C.; Halperin, W.; Kapitulnik, A. Observation of broken time-reversal symmetry in the heavy-fermion superconductor UPt3. Science 2014, 345, 190–193. [Google Scholar] [CrossRef] [PubMed]
  44. Ghosh, S.K.; Smidman, M.; Shang, T.; Annett, J.F.; Hillier, A.D.; Quintanilla, J.; Yuan, H. Recent progress on superconductors with time-reversal symmetry breaking. J. Phys. Condens. Matter 2020, 33, 033001. [Google Scholar] [CrossRef]
  45. Dresselhaus, M.S.; Dresselhaus, G.; Jorio, A. Group Theory: Application to the Physics of Condensed Matter; Springer: Berlin/Heidelberg, Germany, 2008. [Google Scholar]
  46. Hayami, S.; Kusunose, H. Time-reversal switching responses in antiferromagnets. Phys. Rev. B 2023, 108, L140409. [Google Scholar] [CrossRef]
  47. Šmejkal, L.; Sinova, J.; Jungwirth, T. Beyond Conventional Ferromagnetism and Antiferromagnetism: A Phase with Nonrelativistic Spin and Crystal Rotation Symmetry. Phys. Rev. X 2022, 12, 031042. [Google Scholar] [CrossRef]
  48. Schmid, H. On ferrotoroidics and electrotoroidic, magnetotoroidic and piezotoroidic effects. Ferroelectrics 2001, 252, 41–50. [Google Scholar] [CrossRef]
  49. Schmid, H. Some symmetry aspects of ferroics and single phasemultiferroics. J. Phys. Condens. Matter 2008, 20, 434201. [Google Scholar] [CrossRef]
  50. Hayashida, T.; Matsumoto, K.; Kimura, T. Electric Field-Induced Nonreciprocal Directional Dichroism in a Time-Reversal-Odd Antiferromagnet. Adv. Mater. 2024, 20, 2414876. [Google Scholar] [CrossRef]
  51. 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] [PubMed]
  52. 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]
  53. Dubovik, V.; Cheshkov, A. Multipole expansion in classical and quantum field theory and radiation. Sov. J. Part. Nucl. 1975, 5, 318–337. [Google Scholar]
  54. Dubovik, V.; Tugushev, V. Toroid moments in electrodynamics and solid-state physics. Phys. Rep. 1990, 187, 145–202. [Google Scholar] [CrossRef]
  55. Gorbatsevich, A.; Kopaev, Y.V. Toroidal order in crystals. Ferroelectrics 1994, 161, 321–334. [Google Scholar] [CrossRef]
  56. Kopaev, Y.V. Toroidal ordering in crystals. Physics-Uspekhi 2009, 52, 1111–1125. [Google Scholar] [CrossRef]
  57. Spaldin, N.A.; Fiebig, M.; Mostovoy, M. The toroidal moment in condensed-matter physics and its relation to the magnetoelectric effect. J. Phys. Condens. Matter 2008, 20, 434203. [Google Scholar] [CrossRef]
  58. Papasimakis, N.; Fedotov, V.; Savinov, V.; Raybould, T.; Zheludev, N. Electromagnetic toroidal excitations in matter and free space. Nat. Mater. 2016, 15, 263. [Google Scholar] [CrossRef]
  59. Talebi, N.; Guo, S.; van Aken, P.A. Theory and applications of toroidal moments in electrodynamics: Their emergence, characteristics, and technological relevance. Nanophotonics 2018, 7, 93–110. [Google Scholar] [CrossRef]
  60. Cheong, S.W.; Talbayev, D.; Kiryukhin, V.; Saxena, A. Broken symmetries, non-reciprocity, and multiferroicity. NPJ Quantum Mater. 2018, 3, 19. [Google Scholar] [CrossRef]
  61. Xu, X.; Huang, F.T.; Cheong, S.W. Magnetic toroidicity. J. Phys. Condens. Matter 2024, 36, 203002. [Google Scholar] [CrossRef]
  62. Popov, Y.F.; Kadomtseva, A.; Belov, D.; Vorob’ev, G.; Zvezdin, A. Magnetic-field-induced toroidal moment in the magnetoelectric Cr2O3. J. Exp. Theor. Phys. Lett. 1999, 69, 330–335. [Google Scholar] [CrossRef]
  63. Ederer, C.; Spaldin, N.A. Towards a microscopic theory of toroidal moments in bulk periodic crystals. Phys. Rev. B 2007, 76, 214404. [Google Scholar] [CrossRef]
  64. Yanase, Y. Magneto-Electric Effect in Three-Dimensional Coupled Zigzag Chains. J. Phys. Soc. Jpn. 2014, 83, 014703. [Google Scholar] [CrossRef]
  65. Thöle, F.; Spaldin, N.A. Magnetoelectric multipoles in metals. Philos. Trans. R. Soc. A 2018, 376, 20170450. [Google Scholar] [CrossRef] [PubMed]
  66. Saito, H.; Uenishi, K.; Miura, N.; Tabata, C.; Hidaka, H.; Yanagisawa, T.; Amitsuka, H. Evidence of a New Current-Induced Magnetoelectric Effect in a Toroidal Magnetic Ordered State of UNi4B. J. Phys. Soc. Jpn. 2018, 87, 033702. [Google Scholar] [CrossRef]
  67. Yanagi, Y.; Hayami, S.; Kusunose, H. Manipulating the magnetoelectric effect: Essence learned from Co4Nb2O9. Phys. Rev. B 2018, 97, 020404. [Google Scholar] [CrossRef]
  68. Hayami, S.; Kusunose, H.; Motome, Y. Emergent odd-parity multipoles and magnetoelectric effects on a diamond structure: Implication for the 5d transition metal oxides AOsO4 (A = K, Rb, and Cs). Phys. Rev. B 2018, 97, 024414. [Google Scholar] [CrossRef]
  69. Miyahara, S.; Furukawa, N. Nonreciprocal Directional Dichroism and Toroidalmagnons in Helical Magnets. J. Phys. Soc. Jpn. 2012, 81, 023712. [Google Scholar] [CrossRef]
  70. Miyahara, S.; Furukawa, N. Theory of magneto-optical effects in helical multiferroic materials via toroidal magnon excitation. Phys. Rev. B 2014, 89, 195145. [Google Scholar] [CrossRef]
  71. Hayami, S.; Kusunose, H.; Motome, Y. Asymmetric Magnon Excitation by Spontaneous Toroidal Ordering. J. Phys. Soc. Jpn. 2016, 85, 053705. [Google Scholar] [CrossRef]
  72. Okuma, N. Nonreciprocal superposition state in antiferromagnetic optospintronics. Phys. Rev. B 2019, 99, 094401. [Google Scholar] [CrossRef]
  73. Matsumoto, T.; Hayami, S. Nonreciprocal magnons due to symmetric anisotropic exchange interaction in honeycomb antiferromagnets. Phys. Rev. B 2020, 101, 224419. [Google Scholar] [CrossRef]
  74. Sawada, K.; Nagaosa, N. Optical Magnetoelectric Effect in Multiferroic Materials: Evidence for a Lorentz Force Acting on a Ray of Light. Phys. Rev. Lett. 2005, 95, 237402. [Google Scholar] [CrossRef] [PubMed]
  75. Kawaguchi, H.; Tatara, G. Effective Hamiltonian theory for nonreciprocal light propagation in magnetic Rashba conductor. Phys. Rev. B 2016, 94, 235148. [Google Scholar] [CrossRef]
  76. Watanabe, H.; Yanase, Y. Nonlinear electric transport in odd-parity magnetic multipole systems: Application to Mn-based compounds. Phys. Rev. Res. 2020, 2, 043081. [Google Scholar] [CrossRef]
  77. Watanabe, H.; Yanase, Y. Photocurrent response in parity-time symmetric current-ordered states. Phys. Rev. B 2021, 104, 024416. [Google Scholar] [CrossRef]
  78. Suzuki, Y. Tunneling spin current in systems with spin degeneracy. Phys. Rev. B 2022, 105, 075201. [Google Scholar] [CrossRef]
  79. Yatsushiro, M.; Oiwa, R.; Kusunose, H.; Hayami, S. Analysis of model-parameter dependences on the second-order nonlinear conductivity in P T -symmetric collinear antiferromagnetic metals with magnetic toroidal moment on zigzag chains. Phys. Rev. B 2022, 105, 155157. [Google Scholar] [CrossRef]
  80. Kondo, H.; Akagi, Y. Nonlinear magnon spin Nernst effect in antiferromagnets and strain-tunable pure spin current. Phys. Rev. Res. 2022, 4, 013186. [Google Scholar] [CrossRef]
  81. Hayami, S.; Yatsushiro, M.; Kusunose, H. Nonlinear spin Hall effect in P T -symmetric collinear magnets. Phys. Rev. B 2022, 106, 024405. [Google Scholar] [CrossRef]
  82. Wang, C.; Gao, Y.; Xiao, D. Intrinsic Nonlinear Hall Effect in Antiferromagnetic Tetragonal CuMnAs. Phys. Rev. Lett. 2021, 127, 277201. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, H.; Zhao, J.; Huang, Y.X.; Wu, W.; Sheng, X.L.; Xiao, C.; Yang, S.A. Intrinsic Second-Order Anomalous Hall Effect and Its Application in Compensated Antiferromagnets. Phys. Rev. Lett. 2021, 127, 277202. [Google Scholar] [CrossRef]
  84. Kirikoshi, A.; Hayami, S. Microscopic mechanism for intrinsic nonlinear anomalous Hall conductivity in noncollinear antiferromagnetic metals. Phys. Rev. B 2023, 107, 155109. [Google Scholar] [CrossRef]
  85. Ota, K.; Shimozawa, M.; Muroya, T.; Miyamoto, T.; Hosoi, S.; Nakamura, A.; Homma, Y.; Honda, F.; Aoki, D.; Izawa, K. Zero-field current-induced Hall effect in ferrotoroidic metal. arXiv 2022, arXiv:2205.05555. [Google Scholar]
  86. Folen, V.J.; Rado, G.T.; Stalder, E.W. Anisotropy of the Magnetoelectric Effect in Cr2O3. Phys. Rev. Lett. 1961, 6, 607–608. [Google Scholar] [CrossRef]
  87. Krotov, S.; Kadomtseva, A.; Popov, Y.F.; Zvezdin, A.; Vorob’ev, G.; Belov, D. Magnetoelectric interactions and induced toroidal ordering in Cr2O3. J. Magn. Magn. Mater. 2001, 226, 963–964. [Google Scholar] [CrossRef]
  88. Arima, T.; Jung, J.H.; Matsubara, M.; Kubota, M.; He, J.P.; Kaneko, Y.; Tokura, Y. Resonant magnetoelectric X-ray scattering in GaFeO3: Observation of ordering of toroidal moments. J. Phys. Soc. Jpn. 2005, 74, 1419–1422. [Google Scholar] [CrossRef]
  89. Staub, U.; Bodenthin, Y.; Piamonteze, C.; García-Fernández, M.; Scagnoli, V.; Garganourakis, M.; Koohpayeh, S.; Fort, D.; Lovesey, S.W. Parity- and time-odd atomic multipoles in magnetoelectric GaFeO3 as seen via soft x-ray Bragg diffraction. Phys. Rev. B 2009, 80, 140410. [Google Scholar] [CrossRef]
  90. Staub, U.; Piamonteze, C.; Garganourakis, M.; Collins, S.P.; Koohpayeh, S.M.; Fort, D.; Lovesey, S.W. Ferromagnetic-type order of atomic multipoles in the polar ferrimagnetic GaFeO3. Phys. Rev. B 2012, 85, 144421. [Google Scholar] [CrossRef]
  91. Nie, Y.M. First-principles approach to investigate toroidal property of magnetoelectric multiferroic GaFeO3. J. Appl. Phys. 2016, 119. [Google Scholar] [CrossRef]
  92. Van Aken, B.B.; Rivera, J.P.; Schmid, H.; Fiebig, M. Observation of ferrotoroidic domains. Nature 2007, 449, 702–705. [Google Scholar] [CrossRef] [PubMed]
  93. Zimmermann, A.S.; Meier, D.; Fiebig, M. Ferroic nature of magnetic toroidal order. Nat. Commun. 2014, 5, 4796. [Google Scholar] [CrossRef]
  94. Fogh, E.; Zaharko, O.; Schefer, J.; Niedermayer, C.; Holm-Dahlin, S.; Sørensen, M.K.; Kristensen, A.B.; Andersen, N.H.; Vaknin, D.; Christensen, N.B.; et al. Dzyaloshinskii-Moriya interaction and the magnetic ground state in magnetoelectric LiCoPO4. Phys. Rev. B 2019, 99, 104421. [Google Scholar] [CrossRef]
  95. Tóth, B.; Kocsis, V.; Tokunaga, Y.; Taguchi, Y.; Tokura, Y.; Bordács, S. Imaging antiferromagnetic domains in LiCoPO4 via the optical magnetoelectric effect. Phys. Rev. B 2024, 110, L100405. [Google Scholar] [CrossRef]
  96. Murakawa, H.; Onose, Y.; Miyahara, S.; Furukawa, N.; Tokura, Y. Ferroelectricity Induced by Spin-Dependent Metal-Ligand Hybridization in Ba2CoGe2O7. Phys. Rev. Lett. 2010, 105, 137202. [Google Scholar] [CrossRef]
  97. Yamauchi, K.; Barone, P.; Picozzi, S. Theoretical investigation of magnetoelectric effects in Ba2CoGe2O7. Phys. Rev. B 2011, 84, 165137. [Google Scholar] [CrossRef]
  98. Toledano, P.; Khalyavin, D.D.; Chapon, L.C. Spontaneous toroidal moment and field-induced magnetotoroidic effects in Ba2CoGe2O7. Phys. Rev. B 2011, 84, 094421. [Google Scholar] [CrossRef]
  99. Hutanu, V.; Sazonov, A.; Murakawa, H.; Tokura, Y.; Náfrádi, B.; Chernyshov, D. Symmetry and structure of multiferroic Ba2CoGe2O7. Phys. Rev. B 2011, 84, 212101. [Google Scholar] [CrossRef]
  100. Baum, M.; Schmalzl, K.; Steffens, P.; Hiess, A.; Regnault, L.P.; Meven, M.; Becker, P.; Bohatý, L.; Braden, M. Controlling toroidal moments by crossed electric and magnetic fields. Phys. Rev. B 2013, 88, 024414. [Google Scholar] [CrossRef]
  101. Tolédano, P.; Ackermann, M.; Bohatý, L.; Becker, P.; Lorenz, T.; Leo, N.; Fiebig, M. Primary ferrotoroidicity in antiferromagnets. Phys. Rev. B 2015, 92, 094431. [Google Scholar] [CrossRef]
  102. Lee, C.; Kang, J.; Hong, J.; Shim, J.H.; Whangbo, M.H. Analysis of the difference between the pyroxenes LiFeSi2O6 and LiFeGe2O6 in their spin order, spin orientation, and ferrotoroidal order. Chem. Mater. 2014, 26, 1745–1750. [Google Scholar] [CrossRef]
  103. Wadley, P.; Novák, V.; Campion, R.; Rinaldi, C.; Martí, X.; Reichlová, H.; Železnỳ, J.; Gazquez, J.; Roldan, M.; Varela, M.; et al. Tetragonal phase of epitaxial room-temperature antiferromagnet CuMnAs. Nat. Commun. 2013, 4, 2322. [Google Scholar] [CrossRef] [PubMed]
  104. Wadley, P.; Howells, B.; Železnỳ, J.; Andrews, C.; Hills, V.; Campion, R.P.; Novák, V.; Olejník, K.; Maccherozzi, F.; Dhesi, S.; et al. Electrical switching of an antiferromagnet. Science 2016, 351, 587–590. [Google Scholar] [CrossRef] [PubMed]
  105. Godinho, J.; Reichlová, H.; Kriegner, D.; Novák, V.; Olejník, K.; Kašpar, Z.; Šobáň, Z.; Wadley, P.; Campion, R.; Otxoa, R.; et al. Electrically induced and detected Néel vector reversal in a collinear antiferromagnet. Nat. Commun. 2018, 9, 4686. [Google Scholar] [CrossRef]
  106. Manchon, A.; Železný, J.; Miron, I.M.; Jungwirth, T.; Sinova, J.; Thiaville, A.; Garello, K.; Gambardella, P. Current-induced spin–orbit torques in ferromagnetic and antiferromagnetic systems. Rev. Mod. Phys. 2019, 91, 035004. [Google Scholar] [CrossRef]
  107. Ding, L.; Xu, X.; Jeschke, H.O.; Bai, X.; Feng, E.; Alemayehu, A.S.; Kim, J.; Huang, F.T.; Zhang, Q.; Ding, X.; et al. Field-tunable toroidal moment in a chiral-lattice magnet. Nat. Commun. 2021, 12, 5339. [Google Scholar] [CrossRef] [PubMed]
  108. Xu, X.; Huang, F.T.; Admasu, A.S.; Kratochvílová, M.; Chu, M.W.; Park, J.G.; Cheong, S.W. Multiple ferroic orders and toroidal magnetoelectricity in the chiral magnet BaCoSiO4. Phys. Rev. B 2022, 105, 184407. [Google Scholar] [CrossRef]
  109. Nakamura, R.; Aoki, I.; Kimura, K. Ferrotoroidic State Induced by Structural Rotation and Falsely Chiral Antiferromagnetism in PbMn2Ni6Te3O18. J. Phys. Soc. Jpn. 2024, 93, 063703. [Google Scholar] [CrossRef]
  110. Mentink, S.A.M.; Drost, A.; Nieuwenhuys, G.J.; Frikkee, E.; Menovsky, A.A.; Mydosh, J.A. Magnetic Ordering and Frustration in Hexagonal UNi4B. Phys. Rev. Lett. 1994, 73, 1031–1034. [Google Scholar] [CrossRef]
  111. Oyamada, A.; Kondo, M.; Fukuoka, K.; Itou, T.; Maegawa, S.; Li, D.X.; Haga, Y. NMR studies of the partially disordered state in a triangular antiferromagnet UNi4B. J. Phys. Condens. Matter 2007, 19, 145246. [Google Scholar] [CrossRef]
  112. Yanagisawa, T.; Matsumori, H.; Saito, H.; Hidaka, H.; Amitsuka, H.; Nakamura, S.; Awaji, S.; Gorbunov, D.I.; Zherlitsyn, S.; Wosnitza, J.; et al. Electric Quadrupolar Contributions in the Magnetic Phases of UNi4B. Phys. Rev. Lett. 2021, 126, 157201. [Google Scholar] [CrossRef] [PubMed]
  113. 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]
  114. Knight, K.S.; Khalyavin, D.D.; Manuel, P.; Bull, C.L.; McIntyre, P. Nuclear and magnetic structures of KMnF3 perovskite in the temperature interval 10 K–105 K. J. Alloys Compd. 2020, 842, 155935. [Google Scholar] [CrossRef]
  115. Porter, D.G.; Granata, V.; Forte, F.; Di Matteo, S.; Cuoco, M.; Fittipaldi, R.; Vecchione, A.; Bombardi, A. Magnetic anisotropy and orbital ordering in Ca2RuO4. Phys. Rev. B 2018, 98, 125142. [Google Scholar] [CrossRef]
  116. Garlea, V.O.; Jin, R.; Mandrus, D.; Roessli, B.; Huang, Q.; Miller, M.; Schultz, A.J.; Nagler, S.E. Magnetic and Orbital Ordering in the Spinel MnV2O4. Phys. Rev. Lett. 2008, 100, 066404. [Google Scholar] [CrossRef] [PubMed]
  117. Karigerasi, M.H.; Lam, B.H.; Avdeev, M.; Shoemaker, D.P. Two-step magnetic ordering into a canted state in ferrimagnetic monoclinic Mn3As2. J. Solid State Chem. 2021, 294, 121901. [Google Scholar] [CrossRef]
  118. Muñoz, A.; Alonso, J.A.; Martínez-Lope, M.J.; Casáis, M.T.; Martínez, J.L.; Fernández-Díaz, M.T. Magnetic structure of hexagonal RMnO3 (R= Y, Sc): Thermal evolution from neutron powder diffraction data. Phys. Rev. B 2000, 62, 9498–9510. [Google Scholar] [CrossRef]
  119. Garca-Muoz, J.L.; Rodrguez-Carvajal, J.; Obradors, X.; Vallet-Reg, M.; González-Calbet, J.; Parras, M. Complex magnetic structures of the rare-earth cuprates R2Cu2O5 (R = Y, Ho, Er, Yb, Tm). Phys. Rev. B 1991, 44, 4716–4719. [Google Scholar] [CrossRef]
  120. Morosan, E.; Fleitman, J.A.; Huang, Q.; Lynn, J.W.; Chen, Y.; Ke, X.; Dahlberg, M.L.; Schiffer, P.; Craley, C.R.; Cava, R.J. Structure and magnetic properties of the Ho2Ge2O7 pyrogermanate. Phys. Rev. B 2008, 77, 224423. [Google Scholar] [CrossRef]
  121. Eriksson, T.; Bergqvist, L.; Nordblad, P.; Eriksson, O.; Andersson, Y. Structural and magnetic characterization of Mn3IrGe and Mn3Ir (Si1−xGex): Experiments and theory. J. Solid State Chem. 2004, 177, 4058–4066. [Google Scholar] [CrossRef]
  122. Li, M.R.; Retuerto, M.; Walker, D.; Sarkar, T.; Stephens, P.W.; Mukherjee, S.; Dasgupta, T.S.; Hodges, J.P.; Croft, M.; Grams, C.P.; et al. Magnetic-Structure-Stabilized Polarization in an Above-Room-Temperature Ferrimagnet. Angew. Chem. Int. Ed. 2014, 53, 10774. [Google Scholar] [CrossRef]
  123. Hayami, S.; Kusunose, H.; Motome, Y. Spontaneous parity breaking in spin–orbital coupled systems. Phys. Rev. B 2014, 90, 081115. [Google Scholar] [CrossRef]
  124. Hayami, S.; Kusunose, H.; Motome, Y. Spontaneous Multipole Ordering by Local Parity Mixing. J. Phys. Soc. Jpn. 2015, 84, 064717. [Google Scholar] [CrossRef]
  125. Hayami, S.; Kusunose, H.; Motome, Y. Emergent spin-valley-orbital physics by spontaneous parity breaking. J. Phys. Condens. Matter 2016, 28, 395601. [Google Scholar] [CrossRef] [PubMed]
  126. Suzuki, M.T.; Ikeda, H.; Oppeneer, P.M. First-principles theory of magnetic multipoles in condensed matter systems. J. Phys. Soc. Jpn. 2018, 87, 041008. [Google Scholar] [CrossRef]
  127. 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]
  128. Hlinka, J. Eight Types of Symmetrically Distinct Vectorlike Physical Quantities. Phys. Rev. Lett. 2014, 113, 165502. [Google Scholar] [CrossRef]
  129. Hlinka, J.; Privratska, J.; Ondrejkovic, P.; Janovec, V. Symmetry Guide to Ferroaxial Transitions. Phys. Rev. Lett. 2016, 116, 177602. [Google Scholar] [CrossRef]
  130. 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]
  131. 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]
  132. Nasu, J.; Hayami, S. Antisymmetric thermopolarization by electric toroidicity. Phys. Rev. B 2022, 105, 245125. [Google Scholar] [CrossRef]
  133. 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]
  134. Cheong, S.W.; Lim, S.; Du, K.; Huang, F.T. Permutable SOS (symmetry operational similarity). NPJ Quantum Mater. 2021, 6, 58. [Google Scholar] [CrossRef]
  135. Hayami, S.; Yanagi, Y.; Kusunose, H. Spontaneous antisymmetric spin splitting in noncollinear antiferromagnets without spin–orbit coupling. Phys. Rev. B 2020, 101, 220403. [Google Scholar] [CrossRef]
  136. Oiwa, R.; Kusunose, H. Systematic Analysis Method for Nonlinear Response Tensors. J. Phys. Soc. Jpn. 2022, 91, 014701. [Google Scholar] [CrossRef]
Figure 1. Schematics of (a) the magnetic toroidal dipole and (b) magnetic toroidal monopole. The blue and red arrows represent the magnetic dipole and magnetic toroidal dipole moments, respectively. The magnetic toroidal monopole is characterized by the emanating structure of the magnetic toroidal dipole.
Figure 1. Schematics of (a) the magnetic toroidal dipole and (b) magnetic toroidal monopole. The blue and red arrows represent the magnetic dipole and magnetic toroidal dipole moments, respectively. The magnetic toroidal monopole is characterized by the emanating structure of the magnetic toroidal dipole.
Magnetism 05 00015 g001
Figure 2. Antiferromagnetic clusters hosting the magnetic toroidal monopole under the magnetic point groups (a) 4 / m and (b) m m m . The blue and red arrows represent the spin and magnetic toroidal dipole, respectively.
Figure 2. Antiferromagnetic clusters hosting the magnetic toroidal monopole under the magnetic point groups (a) 4 / m and (b) m m m . The blue and red arrows represent the spin and magnetic toroidal dipole, respectively.
Magnetism 05 00015 g002
Figure 3. (a) h T z dependence of the expectation values of the magnetic toroidal monopole T 0 and two magnetic toroidal dipoles T z and T z in the model in Equation (1). (b) θ dependence of the same values as (a) in the model in Equation (9).
Figure 3. (a) h T z dependence of the expectation values of the magnetic toroidal monopole T 0 and two magnetic toroidal dipoles T z and T z in the model in Equation (1). (b) θ dependence of the same values as (a) in the model in Equation (9).
Magnetism 05 00015 g003
Table 1. Active multipoles in the atomic s, p, d, and f wave functions with different principal quantum numbers. E, M, MT, and ET represent electric, magnetic, magnetic toroidal, and electric toroidal multipoles, respectively. # represents the number of independent multipoles.
Table 1. Active multipoles in the atomic s, p, d, and f wave functions with different principal quantum numbers. E, M, MT, and ET represent electric, magnetic, magnetic toroidal, and electric toroidal multipoles, respectively. # represents the number of independent multipoles.
LOrbital# l = 0 123456
0s-s2E/MT
1p-p18E/MTM/ETE/MT
2d-d50E/MTM/ETE/MTM/ETE/MT
3f-f98E/MTM/ETE/MTM/ETE/MTM/ETE/MT
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. Magnetic Toroidal Monopole in a Single-Site System. Magnetism 2025, 5, 15. https://doi.org/10.3390/magnetism5030015

AMA Style

Hayami S. Magnetic Toroidal Monopole in a Single-Site System. Magnetism. 2025; 5(3):15. https://doi.org/10.3390/magnetism5030015

Chicago/Turabian Style

Hayami, Satoru. 2025. "Magnetic Toroidal Monopole in a Single-Site System" Magnetism 5, no. 3: 15. https://doi.org/10.3390/magnetism5030015

APA Style

Hayami, S. (2025). Magnetic Toroidal Monopole in a Single-Site System. Magnetism, 5(3), 15. https://doi.org/10.3390/magnetism5030015

Article Metrics

Back to TopTop