Weyl Semimetallic State in the Rashba-Hubbard Model
Abstract
We investigate the Hubbard model with the Rashba spin-orbit coupling on a square lattice. The Rashba spin-orbit coupling generates two-dimensional Weyl points in the band dispersion. In a system with edges along [11] direction, zero-energy edge states appear, while no edge state exists for a system with edges along an axis direction. The zero-energy edge states with a certain momentum along the edges are predominantly in the up-spin state on the right edge, while they are predominantly in the down-spin state on the left edge. Thus, the zero-energy edge states are helical. By using a variational Monte Carlo method for finite Coulomb interaction cases, we find that the Weyl points can move toward the Fermi level by the correlation effects. We also investigate the magnetism of the model by the Hartree-Fock approximation and discuss weak magnetic order in the weak-coupling region.
1 Introduction
In a two-dimensional system without inversion symmetry, such as in an interface of a heterostructure, a momentum-dependent spin-orbit coupling is allowed. It is called the Rashba spin-orbit coupling [1]. The Rashba spin-orbit coupling lifts the spin degeneracy and affects the electronic state of materials.
Several interesting phenomena originating from the Rashba spin-orbit coupling have been proposed and investigated. By considering the spin precession by the Rashba spin-orbit coupling, Datta and Das proposed the spin transistor [2], in which electron transport between spin-polarized contacts can be modulated by the gate voltage. After this proposal, the tunability of the Rashba spin-orbit coupling by the gate voltage has been experimentally demonstrated [3, 4, 5]. Such an effect may be used in a device in spintronics. The possibility of the intrinsic spin Hall effect, which is also important in the research field of spintronics, by the Rashba spin-orbit coupling has been discussed for a long time [6, 7, 8, 9, 10, 11, 12]. Another interesting phenomenon with the Rashba spin-orbit coupling is superconductivity. When the Rashba spin-orbit coupling is introduced in a superconducting system, even- and odd-parity superconducting states are mixed due to the breaking of the inversion symmetry [13, 14, 15]. This mixing affects the magnetic properties of the superconducting state, such as the Knight shift.
While the above studies have mainly focused on the one-electron states in the presence of the Rashba spin-orbit coupling, the effects of the Coulomb interaction between electrons have also been investigated [16, 17]. The Hubbard model with the Rashba spin-orbit coupling on a square lattice called the Rashba-Hubbard model is one of the simplest models to investigate such effects. In this study, we investigate the ground state of this model at half-filling, i.e., electron number per site , by the variational Monte Carlo method and the Hartree-Fock approximation.
In the strong coupling limit, an effective localized model is derived and the possibility of long-period magnetic order is discussed [18, 19, 20]. The long-period magnetism is a consequence of the Dzyaloshinskii-Moriya interaction caused by the Rashba spin-orbit coupling.
Such magnetic order is also discussed by the Hartree-Fock approximation for the Rashba-Hubbard model [21, 22, 23]. However, there is a contradiction among these studies even within the Hartree-Fock approximation. In the weak-coupling region with a finite Rashba spin-orbit coupling, an antiferromagnetic order is obtained in Ref. \citenMinar2013, but a paramagnetic phase is obtained in Refs. \citenKennedy2022 and \citenKawano2023. We will discuss this point in Sect. 5.
The knowledge of the electron correlation beyond the Hartree-Fock approximation is limited. The electron correlation in the Rashba-Hubbard model is studied by a dynamical mean-field theory mainly focusing on magnetism [24] and by a cluster perturbation theory investigating the Mott transition in the paramagnetic state [25]. We will study the electron correlation in the paramagnetic phase by using the variational Monte Carlo method in Sect. 4. The results concerning the Mott transition are consistent with Ref. \citenBrosco2020. In addition, we find a transition to a Weyl semimetallic state by the electron correlation.
Even without the Coulomb interaction, the band structure of this model is intriguing. When the Rashba spin-orbit coupling is finite, the upper and lower bands touch each other at Weyl points. In the large Rashba spin-orbit coupling limit, all the Weyl points locate at the Fermi level for half-filling. Topological aspects of the Weyl points and corresponding edge states of this simple model are discussed in Sect. 3.
2 Model
The model Hamiltonian is given by . The kinetic energy term is given by
(1) |
where is the annihilation operator of the electron at site with spin and is the Fourier transform of it. denotes a pair of nearest-neighbor sites, is the hopping integral, and the kinetic energy is , where the lattice constant is set as unity.
The Rashba spin-orbit coupling term is given by [26]
(2) |
where () is the unit vector along the () direction, are the Pauli matrices, is the coupling constant of the Rashba spin-orbit coupling, , and . We can assume and without loss of generality. We parametrize them as and .
The band dispersion of is
(3) |
where . The bandwidth is . Due to the electron-hole symmetry of the model, the Fermi level is zero at half-filling. For , that is, without the Rashba spin-orbit coupling, the band is doubly degenerate [Fig. 1(a)].

For a finite , the spin degeneracy is lifted except at the time-reversal invariant momenta , , , and [Figs. 1(b) and 1(c)]. These are two-dimensional Weyl points. The energies at the Weyl points and are always zero. By increasing to (), the energies at the other Weyl points and also move to zero. In Fig. 1(d), we show the energy dispersion in the entire Brillouin zone for . We can see the linear dispersions around the Weyl points.
The Coulomb interaction term is given by
(4) |
where and is the coupling constant of the Coulomb interaction.
3 Topology and edge states of the non-interacting Hamiltonian
The energy bands degenerate when , i.e., at the Weyl points. In the vicinity of these points, we set and obtain
(5) |
The chirality of each Weyl point is defined as [27] and we obtain and .
The winding number of a normalized two-component vector field is [28, 27]
(6) |
where is a loop enclosing . We obtain . Figure 2 shows around and as examples.

We can recognize the winding numbers 1 and , respectively, from this figure.
These topological numbers are related to the Berry phase [29]. The eigenvector of with eigenvalue is . The Berry connection is
(7) |
Then, the Berry phase is
(8) |
From the existence of such topological defects like the Weyl points, we expect edge states as in graphene with Dirac points [30, 31, 32]. We consider two types of edges: the edges along an axis direction [straight edges, Fig. 3(a)] and the edges along [11] direction [zigzag edges, Fig. 3(b)].

We denote the momentum along the edges as and the momentum perpendicular to the edges as . To discuss the existence of the edge states, the chiral symmetry and the winding number for a fixed are important [31, 32]. The Rashba term has a chiral symmetry: and with being the unit matrix. The winding number for a fixed is given by
(9) |
For the straight edges, we find and we expect that the edge states are probably absent. For the zigzag edges, and , where we have set times the bond length as unity, and we find except for , , and (projected Weyl points). At the projected Weyl points, . Thus, the edge states should exist except for the projected Weyl points at least without .
We note that the edge states can be understood as those of a one-dimensional topological insulator. The model only with the Rashba term with fixed is a one-dimensional model. When this one-dimensional system has a gap with a non-zero topological number, the system can be regarded as a one-dimensional topological insulator and has edge states. This one-dimensional system is of symmetry class BDI and can possess a topological number of [33, 34, 35].
To explicitly demonstrate the existence of the edge states, we numerically evaluate the band energy for lattices with finite widths. We denote the number of lattice sites perpendicular to the edges as (see Fig. 3) and obtain bands. The obtained energy bands are shown in Fig. 4.

For the straight edges [Figs. 4(a)–(c)], we do not find the edge states. It is consistent with . For the zigzag edges [Figs. 4(d)–(f)], we obtain isolated zero-energy states except for [Fig. 4(d)]. In particular, for , the zero-energy states appear at all the points except for the projected Weyl points as is expected from . We find that the zero-energy states remain even for finite as shown in Fig. 4(e). For an even number of , the energy of the zero-energy states shifts from zero around the projected Weyl points when is small. For an odd number of , we obtain zero energy even for a small . Thus, we set in the calculations.
We discuss the characteristics of the zero-energy edge states. We define as the Fourier transform of along the edges, where labels the site perpendicular to the edges (see Fig. 3). For the lattice with the zigzag edges, we can show that the states and do not have matrix elements of , where is the vacuum state. Thus, these states are the zero-energy states for completely localized on the left and right edges, respectively, with opposite spins. This helical character of the edge states is natural since the system lacks inversion symmetry [36] due to the Rashba spin-orbit coupling. For other momenta and , we calculate the spin density of the zero-energy edge states , where denotes the zero-energy state at momentum . The zero-energy states are doubly degenerate, and we take the average of the two states. We show for , as an example, in Fig. 5.

At where the bulk band gap is sufficiently large, the zero-energy states are localized well on the edges [Figs. 5(c) and 5(d)]. As the bulk band gap becomes small, the zero-energy states penetrate inner sites [Figs. 5(b) and 5(e)] and the zero-energy states extend in the entire lattice when the gap closes [Figs. 5(a) and 5(f)]. The spin components are opposite between the edges. For example, for and , the up-spin state dominates on the right edge while the down-spin state dominates on the left edge. Thus, the edge states are helical. The spin components are exchanged between states at and [compare Fig. 5(d) with Fig. 5(g) and Fig. 5(e) with Fig. 5(h)]. In Fig. 5(i), we show a schematic view of the spin density corresponding to on the real-space lattice.
4 Weyl semimetallic state induced by the correlation effects
In this section, we investigate the effects of the Coulomb interaction at half-filling, i.e., the electron number per site , within the paramagnetic phase by applying the variational Monte Carlo method [37]. To achieve this objective, it is necessary to select a wave function capable of describing the Mott insulating state, as a Mott transition is anticipated, at least in the ordinary Hubbard model without the Rashba spin-orbit coupling. In this study, we employ a wave function with doublon-holon binding factors [doublon-holon binding wave function (DHWF)] [38, 39]. A doublon means a doubly occupied site and a holon means an empty site. Such intersite factors like doublon-holon binding factors are essential to describe the Mott insulating state [40, 41]. Indeed, the DHWF has succeeded in describing the Mott transition for the single-orbital [40, 42, 43, 44] and two-orbital [45, 46, 47, 48] Hubbard models.
The DHWF is given by
(10) |
The Gutzwiller projection operator
(11) |
describes onsite correlations, where is the projection operator onto the doublon state at and is a variational parameter. The parameter tunes the population of the doubly occupied sites. When the onsite Coulomb interaction is strong and , most sites should be occupied by a single electron each. In this situation, if a doublon is created, a holon should be around it to reduce the energy by using singly occupied virtual states. and describe such doublon-holon binding effects. is an operator to include intersite correlation effects concerning the doublon states. This is defined as follows [47, 49, 48]:
(12) |
where is the projection operator onto the holon state at and denotes the vectors connecting the nearest-neighbor sites. gives factor when site is in the doublon state and there is no holon at nearest-neighbor sites . Similarly, describing the intersite correlation effects on the holon state is defined as
(13) |
Factor appears when a holon exists without a nearest-neighboring doublon. For the half-filled case, we can use the relation due to the electron-hole symmetry of the model.
The one-electron part of the wave function is given by the ground state of the non-interacting Hamiltonian in which in is replaced by . We can choose different from the original in the model Hamiltonian. Such a band renormalization effect of the one-electron part is discussed for a Hubbard model with next-nearest-neighbor hopping [50]. We define the normal state as , i.e., remains the bare value. We also define the Weyl semimetallic state as , i.e., all the Weyl points are at the Fermi level and the Fermi surface disappears. In addition, we can choose other values of , but in a finite-size lattice, a slight change of does not change the set of the occupied wave numbers and the wave function . Thus, we have limited choices for as the band renormalization in the Hubbard model with the next-nearest-neighbor hopping [50].
We use the antiperiodic-periodic boundary conditions since the closed shell condition is satisfied, i.e., no point is exactly on the Fermi surface for a finite-size lattice and there is no ambiguity to construct . The calculations are done for lattices with , 14, and 16.
We evaluate the expectation value of energy by the Monte Carlo method. We optimize the variational parameters and to minimize the energy. We denote the optimized energy of as . In particular, we denote and . By using the Monte Carlo method, we also evaluate the momentum distribution function , where represents the expectation value in the optimized wave function.
In Fig. 6(a), we show in the normal state at for .

For , has clear discontinuities at the Fermi momenta. On the other hand, for , does not have such a discontinuity; that is, the system is insulating and a Mott metal-insulator transition takes place between and . To determine the Mott metal-insulator transition point , we evaluate the quasiparticle renormalization factor , which is inversely proportional to the effective mass and becomes zero in the Mott insulating state, by the jump in . Except for , we evaluate by the jump between and as shown in Fig. 6(a). For , the above path does not intersect the Fermi surface and we use the jump between and instead. In Fig. 6(b), we show the dependence of for and . By extrapolating to zero, we determine . We note that for a small with a large , the Mott transition becomes first-order consistent with a previous study for [43].
We have also evaluated energies for some values of . Figure 7(a) shows energies for and measured from the normal state energy at for .

The normal state has the lowest energy, at least for . Thus, the renormalization of , even if it exists, is weak for a system distant from the Weyl semimetallic state (). A similar conclusion is obtained for a small intersite spin-orbit coupling case of the Kane-Mele-Hubbard model [51]. It is in contrast to the onsite spin-orbit coupling case [52, 51, 49, 48, 53], where the effective spin-orbit coupling is enhanced by the Coulomb interaction even when the bare spin-orbit coupling is small. On the other hand, the renormalization of becomes strong around . In Fig. 7(b), we show the energy of the Weyl semimetallic state measured from that of the normal state for for . becomes lower than the normal state energy at . There is a possibility that the normal state changes to the Weyl semimetallic state gradually by changing continuously. However, for a finite lattice, the choices of are limited between and . For example, at , there is no choice for and and only one choice for . For this reason, we evaluate by comparing the energies of the normal and the Weyl semimetallic states to show the tendency toward the Weyl semimetallic state by the renormalization effect on . Such a renormalization effect is also discussed for a Dirac semimetal system [54].
Figure 8 shows a phase diagram without considering magnetic order.

The size dependence of the phase boundaries is weak. For a weak Rashba spin-orbit coupling region, i.e., for a small , the Rashba spin-orbit coupling stabilizes the metallic phase. It is consistent with a previous study by a cluster perturbation theory [25]. Around , we obtain a wide region of the Weyl semimetallic phase. Thus, we expect phenomena originating from the Weyl points can be realized even away from with the aid of electron correlations. In the Weyl semimetallic state, the density of states at the Fermi level vanishes, and thus, energy gain is expected similar to the energy gain by a gap opening in an antiferromagnetic transition. We note that such a renormalization effect on cannot be expected within the Hartree-Fock approximation and is a result of the electron correlations beyond the Hartree-Fock approximation.
5 Hartree-Fock approximation for magnetism
In this section, we discuss the magnetism of the model by the Hartree-Fock approximation. The energy dispersion given in Eq. (3) has the following property: for . When , in particular, . Thus, the Fermi surface is perfectly nested for half-filling (the Fermi energy is zero) with the nesting vector [see Figs. 9(a)–(c)].

Due to this nesting, the magnetic susceptibility at of the non-interacting system diverges at zero temperature [21]. It indicates that the magnetic order occurs with an infinitesimally small value of the Coulomb interaction at zero temperature. However, some recent Hartree-Fock studies argue the existence of the paramagnetic phase with finite [22, 23]. To resolve this contradiction and gain insights into magnetism, we apply the Hartree-Fock approximation to the model within two-sublattice magnetic order, i.e., with ordering vector of or .
The Hartree-Fock Hamiltonian is given by
(14) |
where -summation runs over the folded Brillouin zone of the antiferromagnetic state, , , and . Here, , where represents the expectation value in the ground state of . We solve the gap equation self-consistently.
First, we consider the magnetic order for . Without the Rashba spin-orbit coupling, the asymptotic form for the weak-coupling region was obtained by Hirsch analyzing the gap equation [55]. If we take into consideration the fact that the asymptotic form of the density of states for [56] is a good approximation even up to the band edge [see Fig. 9(d)], we obtain . Indeed, this approximate form reproduces the numerical data well in the weak-coupling region as shown in Fig. 10(a).

For a finite , we find numerically that is parallel to the or direction. It is expected from the effective Hamiltonian in the strong coupling limit we will discuss later. By assuming and , we obtain for a finite , where is the density of states at the Fermi level. The coefficient to is determined by the entire behavior of the density of states up to the band edge [see Figs. 9(e) and 9(f)] and we cannot obtain it analytically in general. Figures 10(b) and 10(c) show the numerically obtained for and , respectively, along with the fitted curves of , where is the fitting parameter. The fitted curves reproduce well the numerical data in the weak-coupling region.
From the obtained asymptotic form and the numerical data supporting it, we conclude that the magnetic order occurs by an infinitesimally small for consistent with the divergence of the magnetic susceptibility [21]. We cannot apply this asymptotic form for since there. The numerical result shown in Fig. 10(d) indicates a first-order transition for .
Note that it is possible to include magnetic order in the variational Monte Carlo calculations [57, 58, 59, 60, 61, 62, 63, 64, 47, 48]. We have also performed such variational Monte Carlo calculations incorporating magnetic order for the present model and observed the emergence of weak magnetic order for small values of , consistent with the results obtained via the Hartree-Fock approximation.
Here, we discuss previous papers indicating the existence of the paramagnetic phase with finite . In Ref. \citenKennedy2022, the authors introduced a threshold for the magnetization . Then, the authors determined the magnetic transition point when becomes smaller than . However, becomes exponentially small in the weak-coupling region, as understood from the above analysis. In Ref. \citenKennedy2022, is not sufficiently small to discuss the exponentially small value of and a finite region of the paramagnetic phase was obtained. In Ref. \citenKawano2023, the authors calculated the energy difference between the paramagnetic state and the antiferromagnetic state. Then, the authors introduced a scaling between and , where is the antiferromagnetic transition point. They tuned to collapse the data with different onto a single curve in a large- region. Then, they obtained finite for . However, this scaling analysis does not have a basis. In particular, if such a scaling holds for critical behavior, the data collapse should occur for , not for a large- region.
We have also solved the gap equation for and obtained parallel to the direction. By comparing energies for and , we construct a phase diagram shown in Fig. 11.

As noted, the antiferromagnetic state with occurs at infinitesimally small except for . The Weyl semimetallic state remains for at . The antiferromagnetic state with appears at large for .
This phase boundary can be understood from the effective Hamiltonian in the strong coupling limit. The effective Hamiltonian is derived from the second-order perturbation theory concerning and and is given by [18, 19, 21, 20]
(15) |
where or , , , or , is the spin operator at site , , , , and the other components of are zero. From the anisotropy in the interaction, we expect the ordered moments along the or direction for and along the direction for . Thus, the directions of the ordered moments obtained with the Hartree-Fock approximation are in accord with the effective Hamiltonian. For (), the magnetic order with is stable as in the ordinary Heisenberg model. For (), the magnetic order with has lower energy than that with due to the anisotropic interaction. For (), if we ignore the Dzyaloshinskii-Moriya interaction , the model is reduced to the compass model [65]. It is known as a highly frustrated model. The condition corresponds to . Thus, the phase boundary obtained with the Hartree-Fock approximation at a large- region is corresponding to the highly frustrated region of the model.
However, in a large- region, we expect longer-period magnetic order due to the Dzyaloshinskii-Moriya interaction. It is out of the scope of the present study and has already been investigated by previous studies using the effective Hamiltonian [18, 19, 20]. Our important finding in this section is the absence of the paramagnetic phase except for in the weak-coupling region. However, the ordered moment and the energy gain of the antiferromagnetic state in the weak-coupling region are exponentially small. Thus, the transition temperature should be very low, and the effects of this magnetic order should be weak even at zero temperature. In addition, this magnetic order would be easily destroyed by perturbations such as the next-nearest-neighbor hopping breaking the nesting condition [15]. Thus, the discussions in the previous sections without considering magnetic order are still meaningful.
6 Summary
We have investigated the Rashba-Hubbard model on a square lattice. The Rashba spin-orbit coupling generates the two-dimensional Weyl points, which are characterized by non-zero winding numbers. We have investigated lattices with edges and found zero-energy states on a lattice with zigzag edges. The zero-energy states are localized around the edges and have a helical character. The large density of states due to the flat zero-energy band may result in magnetic polarization at edges, similar to graphene [30].
We have also examined the effects of the Coulomb interaction . The Coulomb interaction renormalizes the ratio of the coupling constant of the Rashba spin-orbit coupling to the hopping integral effectively. As a result, the Weyl points can move to the Fermi level by the correlation effects. Thus, the Coulomb interaction can enhance the effects of the Weyl points and assist in observing phenomena originating from the Weyl points even if the bare Rashba spin-orbit coupling is not large.
We have also investigated the magnetism of the model by the Hartree-Fock approximation. We have found that the antiferromagnetic state with the ordering vector occurs at infinitesimally small due to the perfect nesting of the Fermi surface even for a finite . However, the density of states at the Fermi level becomes small for a large and as a result, the energy gain by the antiferromagnetic order is small in the weak-coupling region. Therefore, the transition temperature is very low, and the effects of the magnetic order should be weak in such a region. In addition, this magnetic order would be unstable against perturbations, such as the inclusion of next-nearest-neighbor hopping [15]. Thus, we conclude that the discussions on the Weyl semimetal without assuming magnetism are still meaningful.
This work was supported by JSPS KAKENHI Grant Number JP23K03330.
References
- [1] Y. A. Bychkov and E. I. Rashba, JETP Lett. 39, 78 (1984).
- [2] S. Datta and B. Das, Appl. Phys. Lett. 56, 665 (1990).
- [3] M. Schultz, F. Heinrichs, U. Merkt, T. Colin, T. Skauli, and S. Løvold, Semicond. Sci. Technol. 11, 1168 (1996).
- [4] J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys. Rev. Lett. 78, 1335 (1997).
- [5] G. Engels, J. Lange, T. Schäpers, and H. Lüth, Phys. Rev. B 55, R1958 (1997).
- [6] J. Sinova, D. Culcer, Q. Niu, N. A. Sinitsyn, T. Jungwirth, and A. H. MacDonald, Phys. Rev. Lett. 92, 126603 (2004).
- [7] J.-i. Inoue, G. E. W. Bauer, and L. W. Molenkamp, Phys. Rev. B 70, 041303(R) (2004).
- [8] O. Chalaev and D. Loss, Phys. Rev. B 71, 245318 (2005).
- [9] O. V. Dimitrova, Phys. Rev. B 71, 245327 (2005).
- [10] N. Sugimoto, S. Onoda, S. Murakami, and N. Nagaosa, Phys. Rev. B 73, 113305 (2006).
- [11] V. K. Dugaev, M. Inglot, E. Y. Sherman, and J. Barnaś, Phys. Rev. B 82, 121310(R) (2010).
- [12] A. Shitade and G. Tatara, Phys. Rev. B 105, L201202 (2022).
- [13] L. P. Gor’kov and E. I. Rashba, Phys. Rev. Lett. 87, 037004 (2001).
- [14] Y. Yanase and M. Sigrist, J. Phys. Soc. Jpn. 77, 124711 (2008).
- [15] J. Beyer, J. B. Hauck, L. Klebl, T. Schwemmer, D. M. Kennes, R. Thomale, C. Honerkamp, and S. Rachel, Phys. Rev. B 107, 125115 (2023).
- [16] G.-H. Chen and M. E. Raikh, Phys. Rev. B 60, 4826 (1999).
- [17] D. Maryenko, M. Kawamura, A. Ernst, V. K. Dugaev, E. Y. Sherman, M. Kriener, M. S. Bahramy, Y. Kozuka, and M. Kawasaki, Nat Commun 12, 3180 (2021).
- [18] D. Cocks, P. P. Orth, S. Rachel, M. Buchhold, K. Le Hur, and W. Hofstetter, Phys. Rev. Lett. 109, 205303 (2012).
- [19] J. Radić, A. Di Ciolo, K. Sun, and V. Galitski, Phys. Rev. Lett. 109, 085303 (2012).
- [20] M. Gong, Y. Qian, M. Yan, V. W. Scarola, and C. Zhang, Sci Rep 5, 10050 (2015).
- [21] J. Minář and B. Grémaud, Phys. Rev. B 88, 235130 (2013).
- [22] W. Kennedy, S. dos Anjos Sousa-Júnior, N. C. Costa, and R. R. dos Santos, Phys. Rev. B 106, 165121 (2022).
- [23] M. Kawano and C. Hotta, Phys. Rev. B 107, 045123 (2023).
- [24] X. Zhang, W. Wu, G. Li, L. Wen, Q. Sun, and A.-C. Ji, New J. Phys. 17, 073036 (2015).
- [25] V. Brosco and M. Capone, Phys. Rev. B 101, 235149 (2020).
- [26] F. Mireles and G. Kirczenow, Phys. Rev. B 64, 024426 (2001).
- [27] J.-M. Hou, Phys. Rev. Lett. 111, 130403 (2013).
- [28] K. Sun, W. V. Liu, A. Hemmerich, and S. Das Sarma, Nat. Phys. 8, 67 (2012).
- [29] M. V. Berry, Proc. R. Soc. London, Ser. A 392, 45 (1984).
- [30] M. Fujita, K. Wakabayashi, K. Nakada, and K. Kusakabe, J. Phys. Soc. Jpn. 65, 1920 (1996).
- [31] S. Ryu and Y. Hatsugai, Phys. Rev. Lett. 89, 077002 (2002).
- [32] Y. Hatsugai, Solid State Commun. 149, 1061 (2009).
- [33] A. P. Schnyder, S. Ryu, A. Furusaki, and A. W. W. Ludwig, Phys. Rev. B 78, 195125 (2008).
- [34] A. Kitaev, AIP Conf. Proc. 1134, 22 (2009).
- [35] S. Ryu, A. P. Schnyder, A. Furusaki, and A. W. W. Ludwig, New J. Phys. 12, 065010 (2010).
- [36] J. D. Mella and L. E. F. Foa Torres, Phys. Rev. B 105, 075403 (2022).
- [37] H. Yokoyama and H. Shiba, J. Phys. Soc. Jpn. 56, 1490 (1987).
- [38] T. A. Kaplan, P. Horsch, and P. Fulde, Phys. Rev. Lett. 49, 889 (1982).
- [39] H. Yokoyama and H. Shiba, J. Phys. Soc. Jpn. 59, 3669 (1990).
- [40] H. Yokoyama, Prog. Theor. Phys. 108, 59 (2002).
- [41] M. Capello, F. Becca, S. Yunoki, and S. Sorella, Phys. Rev. B 73, 245116 (2006).
- [42] T. Watanabe, H. Yokoyama, Y. Tanaka, and J.-i. Inoue, J. Phys. Soc. Jpn. 75, 074707 (2006).
- [43] H. Yokoyama, M. Ogata, and Y. Tanaka, J. Phys. Soc. Jpn. 75, 114706 (2006).
- [44] S. Onari, H. Yokoyama, and Y. Tanaka, Physica C 463–465, 120 (2007).
- [45] A. Koga, N. Kawakami, H. Yokoyama, and K. Kobayashi, AIP Conf. Proc. 850, 1458 (2006).
- [46] Y. Takenaka and N. Kawakami, J. Phys.: Conf. Ser. 400, 032099 (2012).
- [47] K. Kubo, Phys. Rev. B 103, 085118 (2021).
- [48] K. Kubo, J. Phys. Soc. Jpn. 91, 124707 (2022).
- [49] K. Kubo, JPS Conf. Proc. 38, 011161 (2023).
- [50] R. Sato and H. Yokoyama, J. Phys. Soc. Jpn. 85, 074701 (2016).
- [51] M. Richter, J. Graspeuntner, T. Schäfer, N. Wentzell, and M. Aichhorn, Phys. Rev. B 104, 195107 (2021).
- [52] Z. Liu, J.-Y. You, B. Gu, S. Maekawa, and G. Su, Phys. Rev. B 107, 104407 (2023).
- [53] K. Jiang, Chin. Phys. Lett. 40, 017102 (2023).
- [54] J. Fujioka, R. Yamada, M. Kawamura, S. Sakai, M. Hirayama, R. Arita, T. Okawa, D. Hashizume, M. Hoshino, and Y. Tokura, Nat Commun 10, 362 (2019).
- [55] J. E. Hirsch, Phys. Rev. B 31, 4403 (1985).
- [56] P. Fazekas: Lecture Notes on Electron Correlation and Magnetism (World Scientific, January 1999), Vol. 5 of Series in Modern Condensed Matter Physics.
- [57] H. Yokoyama and H. Shiba, J. Phys. Soc. Jpn. 56, 3582 (1987).
- [58] K. Kubo, J. Phys.: Conf. Ser. 150, 042101 (2009).
- [59] K. Kubo, Phys. Rev. B 79, 020407(R) (2009).
- [60] K. Kubo and P. Thalmeier, J. Phys. Soc. Jpn. 80, SA121 (2011).
- [61] K. Kubo, J. Phys.: Conf. Ser. 592, 012039 (2015).
- [62] K. Kubo, J. Phys. Soc. Jpn. 84, 094702 (2015).
- [63] K. Kubo, Physics Procedia 75, 214 (2015).
- [64] K. Kubo and H. Onishi, J. Phys. Soc. Jpn. 86, 013701 (2017).
- [65] K. I. Kugel and D. I. Khomskii, Sov. Phys. Usp. 25, 231 (1982).