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

Analysis of model parameter dependences on the second-order nonlinear conductivity in 𝒫𝒯\mathcal{PT}-symmetric collinear antiferromagnetic metals with magnetic toroidal moment on zigzag chains

Megumi Yatsushiro1,3, Rikuto Oiwa2, Hiroaki Kusunose2, and Satoru Hayami3 1Department of Physics, Hokkaido University, Sapporo 060-0810, Japan
2Department of Physics, Meiji University, Kawasaki 214-8571, Japan
3 Department of Applied Physics, The University of Tokyo, Tokyo 113-8656, Japan
Abstract

A magnetic toroidal moment is a fundamental electronic degree of freedom in the absence of both spatial inversion and time-reversal symmetries and gives rise to novel multiferroic and transport properties. We elucidate essential model parameters of the nonlinear transport in the space-time (𝒫𝒯\mathcal{PT}) symmetric collinear antiferromagnetic metals accompanying a magnetic toroidal moment. By analyzing the longitudinal and transverse components of the second-order nonlinear conductivity on a two-dimensionally stacked zigzag chain based on the nonlinear Kubo formula, we show that an effective coupling between the magnetic toroidal moment and the antisymmetric spin-orbit interaction is an essential source of the nonlinear conductivity. Moreover, we find that the nonreciprocal longitudinal current and nonlinear transverse current in a multi-band system are largely enhanced just below the transition temperature of the antiferromagnetic ordering. We also discuss the relevance of the nonlinear conductivity to the linear magnetoelectric coefficient and conductivity. Our result serves as a guide for exploring microscopic essence and clarifying the parameter dependence of the nonlinear conductive phenomena in ferrotoroidal metals.

preprint: APS/123-QED

I Introduction

Spontaneous time-reversal symmetry breaking has long been attracted much attention, as it leads to intriguing physical phenomena, such as the anomalous Hall effect and the magneto-optical Kerr effect. Modern understanding of these phenomena has been achieved based on the Berry phase mechanism [1, 2]. Although such phenomena were originally studied in the ferromagnetic state, it has recently been recognized that similar phenomena can occur in a certain class of antiferromagnetic (AFM) states without the uniform magnetization [3]. For example, the collinear AFM ordering with the mirror symmetry breaking as the uniform magnetization, results in the anomalous Hall effect [4, 5, 6, 7]. Thus, the AFM materials can also exhibit the same physical properties as ordinary ferromagnetic ones, which is advantageous for functional materials without leakage of a magnetic field.

The AFM state also exhibits multiferroic phenomena when both spatial inversion (𝒫)(\mathcal{P}) and time-reversal (𝒯)(\mathcal{T}) symmetries are broken simultaneously while their product (𝒫𝒯)(\mathcal{PT}) symmetry is preserved. The typical example is the linear magnetoelectric effect in the AFM insulators, e.g., Cr2O3 [8], Ga2-xFexO3 [9, 10], LiCoPO4 [11, 12], and Ba2CoGe2O7 [13], and in the AFM metals, e.g., UNi4[14, 15, 16] and Ce3TiBi5 [17, 18]. Moreover, the nonreciprocal optical and transport properties have been studied [19, 20, 21, 22, 23]. Among them, multiferroic phenomena within the linear response theory have been understood by regarding the fact that the AFM states accompany the uniform orderings of the electronic odd-parity magnetic-type multipoles [24, 14, 25, 26, 27, 28, 29, 30, 31, 32, 33, 30, 34, 35, 36, 37, 38, 39], such as the magnetic toroidal (MT) dipole  [40, 41, 42, 43, 14, 25, 26, 44, 33, 30, 35, 37].

Meanwhile, the microscopic understanding of the nonlinear transports in AFMs has not been fully achieved except for several works [22, 45, 46] and symmetry analyses [39, 47]. For example, it remains unclear which model parameters are essentially important to induce nonlinear transports and how the odd-parity magnetic-type multipoles are related to them. To be clear this point and obtain an intuitive understanding of the nonlinear transport, it is useful to extract the essential model parameters, without which the nonlinear transport coefficients vanish, from various hopping processes, spin-orbit coupling, and order parameters in the microscopic model Hamiltonian. Such an understanding provides a guideline to explore new functional AFM materials with a giant nonlinear transport, and its efficient bottom-up design in combination with the ab initio calculations.

In this paper, we elucidate the microscopic essential model parameters for the second-order nonlinear conductivity in the 𝒫𝒯\mathcal{PT}-symmetric collinear AFMs by focusing on the role of the MT moment. By analyzing a minimal model on a two-dimensionally-stacked zigzag chain based on the nonlinear Kubo formula, we show that the effective coupling between the MT moment and one of the antisymmetric spin-orbit interactions (ASOIs) plays an essential role in inducing the longitudinal and transverse components of the nonlinear conductivity. Moreover, we find that the nonlinear conductivities are highly enhanced near the transition temperature in the case that the AFM molecular field is comparable to the ASOI in a multi-band system. We also discuss the relevance between the transverse nonlinear conductivity and the linear magnetoelectric coefficient by comparing the ASOI and temperature dependences.

The organization of this paper is as follows. In Sec. II, we introduce a minimal model on a two-dimensionally stacked zigzag chain. After showing the relation of an MT moment to the nonlinear conductivity and the linear magnetoelectric coefficient in Sec. III, the numerical results are presented in Sec. IV. In Sec. V, we discuss the essential model parameters and the semi-quantitative evaluation of the nonlinear conductivity. We summarize this paper in Sec. VI. In Appendix A, we present the functional forms of the odd-parity magnetic and MT multipoles. In Appendix B, we show the analytic expressions for the essential model parameters in the asymmetric band modulation, nonlinear conductivities, and linear magnetoelectric coefficient. Finally, we present the numerical result of the nonlinear transverse conductivity in the presence of the additional interlayer hopping in Appendix C.

Refer to caption
Figure 1: (a), (b) Schematic pictures of (a) a two-sublattice zigzag chain and (b) its stacking along the zz direction. (c) The temperature (TT) dependence of the MT moment TxMFT_{x}^{\rm MF} at α1=0.4\alpha_{1}=0.4 and α2=0.1\alpha_{2}=0.1. The AFM structure with the MT moment along the xx direction TxT_{x} is shown in the inset. (d) The energy bands measured from the chemical potential μ\mu at kz=0k_{z}=0 for three temperatures.

II Model

We consider a minimal two-dimensional system where the zigzag chain along the xx direction [Fig. 1(a)] is stacked along the zz direction [Fig. 1(b)]. The tight-binding Hamiltonian is given by

=\displaystyle\mathcal{H}= hopAB+hop+ASOI+int,\displaystyle\mathcal{H}_{\rm hop}^{\rm AB}+\mathcal{H}_{\rm hop}+\mathcal{H}_{\rm ASOI}+\mathcal{H}_{\rm int}, (1)
hopAB=\displaystyle\mathcal{H}_{\rm hop}^{\rm AB}= 𝒌σ{εAB(𝒌)c𝒌Aσc𝒌Bσ+H.c.},\displaystyle\sum_{\bm{k}}\sum_{\sigma}\left\{\varepsilon^{\rm AB}({\bm{k}})c_{{\bm{k}}{\rm A}\sigma}^{\dagger}c_{{\bm{k}}{\rm B}\sigma}+{\rm H.c.}\right\}, (2)
hop=\displaystyle\mathcal{H}_{\rm hop}= 𝒌σε(𝒌)(c𝒌Aσc𝒌Aσ+c𝒌Bσc𝒌Bσ),\displaystyle\sum_{\bm{k}}\sum_{\sigma}\varepsilon({\bm{k}})(c_{{\bm{k}}{\rm A}\sigma}^{\dagger}c_{{\bm{k}}{\rm A}\sigma}+c_{{\bm{k}}{\rm B}\sigma}^{\dagger}c_{{\bm{k}}{\rm B}\sigma}), (3)
ASOI=\displaystyle\mathcal{H}_{\rm ASOI}= 𝒌σσ𝒈(𝒌)𝝈σσ(c𝒌Aσc𝒌Aσc𝒌Bσc𝒌Bσ),\displaystyle\sum_{\bm{k}}\sum_{\sigma\sigma^{\prime}}{\bm{g}}({\bm{k}})\cdot{\bm{\sigma}}^{\sigma\sigma^{\prime}}(c_{{\bm{k}}{\rm A}\sigma}^{\dagger}c_{{\bm{k}}{\rm A}\sigma^{\prime}}-c_{{\bm{k}}{\rm B}\sigma}^{\dagger}c_{{\bm{k}}{\rm B}\sigma^{\prime}}), (4)
int=\displaystyle\mathcal{H}_{\rm int}= JAFijM^iAzM^jBz,\displaystyle J_{\rm AF}\sum_{\braket{ij}}\hat{M}_{i{\rm A}}^{z}\hat{M}_{j{\rm B}}^{z}, (5)

where c𝒌lσc_{{\bm{k}}l\sigma}^{\dagger} (c𝒌lσc_{{\bm{k}}l\sigma}) is the creation (annihilation) operator of electrons at wave vector 𝒌{\bm{k}}, sublattice l=A,Bl={\rm A},{\rm B}, and spin σ=,\sigma=\uparrow,\downarrow. The hopping Hamiltonian hopAB\mathcal{H}_{\rm hop}^{\rm AB} in Eq. (2) includes the nearest-neighbor hopping between A and B sublattices as εAB(𝒌)=2t1cos(kxa/2)\varepsilon^{\rm AB}({\bm{k}})=-2t_{1}\cos(k_{x}a/2), while hop\mathcal{H}_{\rm hop} includes the hoppings within the same sublattices along the xx and zz directions as ε(𝒌)=2t2cos(kxa)2t3cos(kzc)\varepsilon({\bm{k}})=-2t_{2}\cos{(k_{x}a)}-2t_{3}\cos{(k_{z}c)}. ASOI\mathcal{H}_{\rm ASOI} in Eq. (4) represents the ASOI that arises from the relativistic spin-orbit coupling as 𝒈(𝒌)=[α2sin(kzc),0,α1sin(kxa)]{\bm{g}}({\bm{k}})=[-\alpha_{2}\sin{(k_{z}c)},0,\alpha_{1}\sin{(k_{x}a)}]. The ASOI in Eq. (4) has the sublattice-dependent staggered form satisfying the global inversion symmetry [48, 24]. int\mathcal{H}_{\rm int} in Eq. (5) represents the Ising-type AFM exchange interaction of the nearest-neighbor A-B bond with JAF>0J_{\rm AF}>0 where M^iA(B)z=σσciA(B)σσσσzciA(B)σ\hat{M}_{i{\rm A}({\rm B})}^{z}=\sum_{\sigma\sigma^{\prime}}c_{i{\rm A}({\rm B})\sigma}^{\dagger}\sigma^{z}_{\sigma\sigma^{\prime}}c_{i{\rm A}({\rm B})\sigma^{\prime}} is the zz component of the magnetic dipole operator and cilσc_{il\sigma}^{\dagger} and cilσc_{il\sigma} are the Fourier transforms of c𝒌lσc_{\bm{k}l\sigma}^{\dagger} and c𝒌lσc_{\bm{k}l\sigma}, respectively. We adopt the Hartree-type mean-field approximation as

JAFijM^iAzM^jBz\displaystyle J_{\rm AF}\sum_{\braket{ij}}\hat{M}_{i{\rm A}}^{z}\hat{M}_{j{\rm B}}^{z}
J~AFi(M^AzM^iBz+M^BzM^iAzM^AzM^Bz),\displaystyle\to\tilde{J}_{\rm AF}\sum_{i}\left(\braket{\hat{M}_{\rm A}^{z}}\hat{M}_{i{\rm B}}^{z}+\braket{\hat{M}_{\rm B}^{z}}\hat{M}_{i{\rm A}}^{z}-\braket{\hat{M}_{\rm A}^{z}}\braket{\hat{M}_{\rm B}^{z}}\right), (6)

where \braket{\cdots} represents the statistical average and J~AF=2JAF\tilde{J}_{\rm AF}=2J_{\rm AF} is the renormalized coupling constant taking into account the two nearest-neighbor atomic sites. We set the model parameters as (t1,t2,t3,JAF)=(0.1,1,0.5,2.5)(t_{1},t_{2},t_{3},J_{\rm AF})=(0.1,1,0.5,2.5), electron filling as 1/51/5, and the lattice constant as a=c=1a=c=1 in the following discussion; t2t_{2} is set as the energy unit.

The model in Eq. (1) exhibits the MT moment when the global inversion symmetry is broken under the staggered AFM ordering, as shown in the inset of Fig. 1(c) [24, 14]. In the present system, the staggered AFM moment along the zz direction is equivalent to the uniform MT moment along the xx direction; TxMF(M^AzM^Bz)/2T_{x}^{\rm MF}\equiv(\braket{\hat{M}_{\rm A}^{z}}-\braket{\hat{M}_{\rm B}^{z}})/2 [49]; see also Appendix A. The TT dependence of TxMFT_{x}^{\rm MF} at α1=0.4\alpha_{1}=0.4 and α2=0.1\alpha_{2}=0.1 is shown in Fig. 1(c), where TxMFT_{x}^{\rm MF} is self-consistently determined for the two-sublattice unit cell by taking over 2002200^{2} grid points in the Brillouin zone. TxMFT_{x}^{\rm MF} becomes nonzero below the transition temperature TNT_{\rm N} and saturates below T0.2TNT\simeq 0.2T_{\rm N}. Almost the same behavior is obtained for α1,α20.5\alpha_{1},\alpha_{2}\lesssim 0.5. Reflecting TxMF0T_{x}^{\rm MF}\neq 0, the electronic band structure is asymmetrically modulated along the kxk_{x} direction, as shown in Fig. 1(d) [24, 49]. This asymmetric band modulation is understood from the effective coupling between TxMFT_{x}^{\rm MF} and the ASOI α1\alpha_{1} in the doubly degenerate bands with the 𝒫𝒯\mathcal{PT} symmetry, i.e.,

ε±(𝒌)\displaystyle\varepsilon_{\pm}({\bm{k}}) =ε(𝒌)±X(𝒌),\displaystyle=\varepsilon({\bm{k}})\pm X({\bm{k}}), (7)
X(𝒌)\displaystyle X({\bm{k}}) =(α1sxT~xMF)2+α22sz2+4t12cx/22,\displaystyle=\sqrt{(\alpha_{1}s_{x}-\tilde{T}_{x}^{\rm MF})^{2}+\alpha^{2}_{2}s^{2}_{z}+4t^{2}_{1}c^{2}_{x/2}}, (8)

where sx=sinkxs_{x}=\sin k_{x}, sz=sinkzs_{z}=\sin k_{z}, cx/2=coskx/2c_{x/2}=\cos k_{x}/2, and T~xMF=J~AFTxMF\tilde{T}_{x}^{\rm MF}=\tilde{J}_{\rm AF}T_{x}^{\rm MF}. The factor (α1sxT~xMF)2(\alpha_{1}s_{x}-\tilde{T}_{x}^{\rm MF})^{2} includes the coupling between T~xMF\tilde{T}_{x}^{\rm MF} and α1\alpha_{1} with the odd function of kxk_{x}. This asymmetric band modulation due to the coupling between α1\alpha_{1} and T~xMF\tilde{T}_{x}^{\rm MF} becomes a source of the nonlinear transport as will be discussed in the following sections; see also Appendix B.

III Second-order nonlinear conductivity and linear response coefficient

III.1 Second-order nonlinear conductivity

The second-order nonlinear conductivity tensor σμνλ\sigma_{\mu\nu\lambda} defined as Jμ=σμνλEνEλJ_{\mu}=\sigma_{\mu\nu\lambda}E_{\nu}E_{\lambda} (μ,ν,λ=x,y,z\mu,\nu,\lambda=x,y,z) is calculated on the basis of the second-order Kubo formula [22]. In the clean limit, the intraband contribution is dominant, which is given by

σμνλ=e3τ231V𝒌n2εn(𝒌)kμkνεn(𝒌)kλf[εn(𝒌)]εn(𝒌),\displaystyle\sigma_{\mu\nu\lambda}=\frac{e^{3}\tau^{2}}{\hbar^{3}}\frac{1}{V}\sum_{\bm{k}}\sum_{n}\frac{\partial^{2}\varepsilon_{n}({\bm{k}})}{\partial k_{\mu}\partial k_{\nu}}\frac{\partial\varepsilon_{n}({\bm{k}})}{\partial k_{\lambda}}\frac{\partial f[\varepsilon_{n}({\bm{k}})]}{\partial\varepsilon_{n}({\bm{k}})}, (9)

where e(>0)e(>0), τ\tau, \hbar, and VV are the elementary charge, relaxation time, the reduced Planck constant, and the system volume, respectively 111There is no contribution from the Berry curvature dipole [75] because of the 𝒫𝒯\mathcal{PT} symmetry, while the interband contribution in the 𝒯\mathcal{T}-breaking system [76] is neglected by considering the clean limit.. f[εn(𝒌)]f[\varepsilon_{n}({\bm{k}})] is the Fermi distribution function for the nnth-band eigen energy εn(𝒌)\varepsilon_{n}({\bm{k}}). The intraband contribution in Eq. (9) represents the Drude-type one with the dissipation τ2\tau^{-2}, whose expression eventually coincides with that obtained by the Boltzmann formalism [51, 52, 22, 53]. Hereafter, we use the scaled σμνλ\sigma_{\mu\nu\lambda} as σ¯μνλ=σμνλ/(e3τ23)\bar{\sigma}_{\mu\nu\lambda}=\sigma_{\mu\nu\lambda}/(e^{3}\tau^{2}\hbar^{-3}).

From Eq. (9), one finds the relation σμνν=σνμν\sigma_{\mu\nu\nu}=\sigma_{\nu\mu\nu} by integration by parts. This indicates that the Drude-type nonlinear conductivity σμνλ\sigma_{\mu\nu\lambda} is the totally symmetric rank-3 tensor with 10 independent components: σxxx\sigma_{xxx}, σyyy\sigma_{yyy}, σzzz\sigma_{zzz}, σxyy\sigma_{xyy}, σyzz\sigma_{yzz}, σzxx\sigma_{zxx}, σxxy\sigma_{xxy}, σyyz\sigma_{yyz}, σzzx\sigma_{zzx}, and σxyz\sigma_{xyz}. As σμνλ\sigma_{\mu\nu\lambda} is a third-rank polar time-reversal-odd tensor, i.e., σμνλσμνλ\sigma_{\mu\nu\lambda}\to-\sigma_{\mu\nu\lambda} under 𝒫\mathcal{P} or 𝒯\mathcal{T} operation but σμνλσμνλ\sigma_{\mu\nu\lambda}\to\sigma_{\mu\nu\lambda} under 𝒫𝒯\mathcal{PT} operation, it becomes nonzero when both the spatial inversion and time-reversal symmetries are absent. From the multipole viewpoint, above symmetry requirement means that the nonzero tensor components are related to the active odd-parity MT multipoles [54, 55, 56, 42, 57]: three rank-1 MT dipoles (Tx,Ty,Tz)(T_{x},T_{y},T_{z}) and seven rank-3 MT octupoles (Txyz,Txα,Tyα,Tzα,Txβ,Tyβ,Tzβ)(T_{xyz},T_{x}^{\alpha},T_{y}^{\alpha},T_{z}^{\alpha},T_{x}^{\beta},T_{y}^{\beta},T_{z}^{\beta}), whose correspondence is given by [39]

σ\displaystyle\sigma =(σxxxσyxxσzxxσxyyσyyyσzyyσxzzσyzzσzzzσxyzσyyzσzyzσxzxσyzxσzzxσxxyσyxyσzxy)T\displaystyle=\begin{pmatrix}\sigma_{xxx}&\sigma_{yxx}&\sigma_{zxx}\\ \sigma_{xyy}&\sigma_{yyy}&\sigma_{zyy}\\ \sigma_{xzz}&\sigma_{yzz}&\sigma_{zzz}\\ \sigma_{xyz}&\sigma_{yyz}&\sigma_{zyz}\\ \sigma_{xzx}&\sigma_{yzx}&\sigma_{zzx}\\ \sigma_{xxy}&\sigma_{yxy}&\sigma_{zxy}\\ \end{pmatrix}^{\rm T}
(3Tx+2TxαTyTyαTyβTzTzα+TzβTxTxα+Txβ3Ty+2TyαTzTzαTzβTxTxαTxβTyTyα+Tyβ3Tz+2TzαTxyzTzTzαTzβTyTyα+TyβTzTzα+TzβTxyzTxTxαTxβTyTyαTyβTxTxα+TxβTxyz)T,\displaystyle\leftrightarrow\begin{pmatrix}3{T}_{x}+2T_{x}^{\alpha}&T_{y}-T_{y}^{\alpha}-T_{y}^{\beta}&T_{z}-T_{z}^{\alpha}+T_{z}^{\beta}\\ T_{x}-T_{x}^{\alpha}+T_{x}^{\beta}&3{T}_{y}+2T_{y}^{\alpha}&T_{z}-T_{z}^{\alpha}-T_{z}^{\beta}\\ T_{x}-T_{x}^{\alpha}-T_{x}^{\beta}&T_{y}-T_{y}^{\alpha}+T_{y}^{\beta}&3{T}_{z}+2T_{z}^{\alpha}\\ T_{xyz}&T_{z}-T_{z}^{\alpha}-T_{z}^{\beta}&{T}_{y}-T_{y}^{\alpha}+T_{y}^{\beta}\\ {T}_{z}-T_{z}^{\alpha}+T_{z}^{\beta}&T_{xyz}&{T}_{x}-T_{x}^{\alpha}-T_{x}^{\beta}\\ {T}_{y}-T_{y}^{\alpha}-T_{y}^{\beta}&{T}_{x}-T_{x}^{\alpha}+T_{x}^{\beta}&T_{xyz}\\ \end{pmatrix}^{\rm T}, (10)

where the functional forms of dipoles and octupoles are summarized in Appendix A. The correspondence in Eq. (III.1) is obtained by decomposing σμνλ\sigma_{\mu\nu\lambda} into the tensor components with the same rotational symmetry to the dipoles and octupoles (See also Appendix A). When the MT dipole and/or MT octupole in Eq. (III.1) are activated in an AFM metal, the corresponding tensor component of σμνλ\sigma_{\mu\nu\lambda} becomes nonzero. From Eq. (III.1), one finds that MT dipole TμT_{\mu} is relevant to the longitudinal component σμμμ\sigma_{\mu\mu\mu} and the transverse components σμνν\sigma_{\mu\nu\nu} and σνμν\sigma_{\nu\mu\nu} (νμ\nu\neq\mu). It means that both nonreciprocal conductivity and nonlinear transverse conductivity are expected to be realized in the presence of the MT dipole, i.e., ferrotoroidal metals [58, 59, 39].

In the present system under the magnetic point-group mmmm^{\prime}mm with the nonzero MT moment TxMFT_{x}^{\rm MF}, five components σxxx\sigma_{xxx}, σxyy\sigma_{xyy}, σyxy\sigma_{yxy}, σxzz\sigma_{xzz}, and σzzx\sigma_{zzx} can be nonzero, since TxT_{x}, TxαT_{x}^{\alpha}, and TxβT_{x}^{\beta} in Eq. (III.1) belong to the totally symmetric irreducible representation [39]. Among them, σxyy\sigma_{xyy} and σyxy\sigma_{yxy} vanish owing to ky=0k_{y}=0 in the present two-dimensional system. In addition to the nonzero contribution from the linear conductivity σxx\sigma_{xx}, σxxx\sigma_{xxx} results in the nonreciprocal current, while σxzz\sigma_{xzz} without linear σxz\sigma_{xz} leads to the pure second-order transverse current, respectively.

III.2 Linear response coefficient

In the presence of the MT moment TxMFT_{x}^{\rm MF}, the linear magnetoelectric tensor αμν\alpha_{\mu\nu} in Mμ=αμνEνM_{\mu}=\alpha_{\mu\nu}E_{\nu} (μ,ν=x,y,z\mu,\nu=x,y,z) is also finite. We calculate the linear magnetoelectric tensor by the linear response theory as

αμν=\displaystyle\alpha_{\mu\nu}= egμB2Vi𝒌nmf[εn(𝒌)]f[εm(𝒌)][εn(𝒌)εm(𝒌)]2+(δ)2σμ𝒌nm\varvν𝒌mn,\displaystyle\frac{eg\mu_{\rm B}\hbar}{2Vi}\sum_{\bm{k}}\sum_{n\neq m}\frac{f[\varepsilon_{n}({\bm{k}})]-f[\varepsilon_{m}({\bm{k}})]}{[\varepsilon_{n}({\bm{k}})-\varepsilon_{m}({\bm{k}})]^{2}+(\hbar\delta)^{2}}\sigma_{\mu{\bm{k}}}^{nm}\varv_{\nu{\bm{k}}}^{mn}, (11)

where gg and μB\mu_{\rm B} are the g factor (g=2g=2) and Bohr magneton, respectively. σμ𝒌nm=n𝒌|σμ|m𝒌\sigma_{\mu{\bm{k}}}^{nm}=\braket{n{\bm{k}}}{\sigma_{\mu}}{m{\bm{k}}} and \varvν𝒌mn=m𝒌|\varvν𝒌|n𝒌\varv_{\nu{\bm{k}}}^{mn}=\braket{m{\bm{k}}}{\varv_{\nu{\bm{k}}}}{n{\bm{k}}} are the matrix elements of spin σμ\sigma_{\mu} and velocity \varvν𝒌=/(kν)\varv_{\nu{\bm{k}}}=\partial\mathcal{H}/(\hbar\partial k_{\nu}) for the eigenstate |n𝒌\ket{n{\bm{k}}}. We use the scaled α¯μν=αμν/(eμB)\bar{\alpha}_{\mu\nu}=\alpha_{\mu\nu}/(e\mu_{\rm B}\hbar) in the following discussion.

As αμν\alpha_{\mu\nu} in a 𝒫𝒯\mathcal{PT} symmetric system is relevant to the rank-0–2 odd-parity multipoles: magnetic monopole M0M_{0}, MT dipoles (Tx,Ty,Tz)(T_{x},T_{y},T_{z}), and magnetic quadrupoles (Mu,Mv,Myz,Mzx,Mxy)(M_{u},M_{v},M_{yz},M_{zx},M_{xy}) (see also Appendix A), the relation is represented as follows [31, 32]:

α\displaystyle\alpha =(αxxαxyαxzαyxαyyαyzαzxαzyαzz)\displaystyle=\begin{pmatrix}\alpha_{xx}&\alpha_{xy}&\alpha_{xz}\\ \alpha_{yx}&\alpha_{yy}&\alpha_{yz}\\ \alpha_{zx}&\alpha_{zy}&\alpha_{zz}\\ \end{pmatrix} (12)
(M0Mu+MvMxy+TzMzxTyMxyTzM0MuMvMyz+TxMzx+TyMyzTxM0+2Mu).\displaystyle\leftrightarrow\begin{pmatrix}M_{0}-M_{u}+M_{v}&M_{xy}+T_{z}&M_{zx}-T_{y}\\ M_{xy}-T_{z}&M_{0}-M_{u}-M_{v}&M_{yz}+T_{x}\\ M_{zx}+T_{y}&M_{yz}-T_{x}&M_{0}+2M_{u}\\ \end{pmatrix}. (13)

Since TxT_{x} and MyzM_{yz} become active for TxAF0T_{x}^{\rm AF}\neq 0 in the present system, αyz\alpha_{yz} and αzy\alpha_{zy} are expected to be nonzero. As αzy\alpha_{zy} is zero due to the two dimensionality, we only consider αyz\alpha_{yz}.

For the following discussion, we also present the linear Hall conductivity

σxz=\displaystyle\sigma_{xz}= e2Vi𝒌nmf[εn(𝒌)]f[εm(𝒌)][εn(𝒌)εm(𝒌)]2+(δ)2\varvx𝒌nm\varvz𝒌mn.\displaystyle\frac{e^{2}\hbar}{Vi}\sum_{\bm{k}}\sum_{n\neq m}\frac{f[\varepsilon_{n}({\bm{k}})]-f[\varepsilon_{m}({\bm{k}})]}{[\varepsilon_{n}({\bm{k}})-\varepsilon_{m}({\bm{k}})]^{2}+(\hbar\delta)^{2}}\varv_{x{\bm{k}}}^{nm}\varv_{z{\bm{k}}}^{mn}. (14)

We use the scaled value σ¯xz=σxz/(e2Hy)\bar{\sigma}_{xz}=\sigma_{xz}/(e^{2}\hbar H_{y}) in the following, where HyH_{y} is the Zeeman field along the yy direction.

IV Numerical Result

IV.1 Longitudinal second-order conductivity σxxx\sigma_{xxx}

Refer to caption
Figure 2: (a) The longitudinal second-order conductivity σ¯xxx\bar{\sigma}_{xxx} for α1=0.1\alpha_{1}=0.10.50.5 as a function of TT at α2=0.1\alpha_{2}=0.1. The inset shows σ¯xxx/α1\bar{\sigma}_{xxx}/\alpha_{1}. (b) The upper- and lower-band contributions to σ¯xxx\bar{\sigma}_{xxx} at α1=0.4\alpha_{1}=0.4.

We first show the numerical result of the longitudinal nonlinear conductivity σ¯xxx\bar{\sigma}_{xxx}. Figure 2(a) shows σ¯xxx\bar{\sigma}_{xxx} as a function of TT for various α1=0.1\alpha_{1}=0.10.50.5 at α2=0.1\alpha_{2}=0.1. The TT dependence for different α1\alpha_{1} is qualitatively similar; σ¯xxx\bar{\sigma}_{xxx} is largely enhanced just below T=TNT=T_{{\rm N}}, and shows maximum with decrease of TT. While further decreasing TT, σ¯xxx\bar{\sigma}_{xxx} shows the sign change, and then reaches a negative value at the lowest TT.

The nonzero σxxx\sigma_{xxx} is closely related to the formation of the asymmetric band structure under TxMF0T_{x}^{\rm MF}\neq 0, since σxxx\sigma_{xxx} has the same symmetry as TxMFT_{x}^{\rm MF} [39]. As the asymmetric band modulation is caused by the coupling between T~xMF\tilde{T}_{x}^{\rm MF} and α1\alpha_{1}, they are indispensable for nonzero σxxx\sigma_{xxx}. Indeed, σ¯xxx\bar{\sigma}_{xxx} vanishes for α1=0\alpha_{1}=0 or T~xMF=0\tilde{T}_{x}^{\rm MF}=0. Moreover, σ¯xxx\bar{\sigma}_{xxx} is well scaled by σ¯xxx/α1\bar{\sigma}_{xxx}/\alpha_{1} at low temperatures T0.7TNT\lesssim 0.7T_{\rm N} for small α1\alpha_{1}. See Sec. V.1 for the essential model parameters in details.

Meanwhile, σ¯xxx\bar{\sigma}_{xxx} is not scaled by α1\alpha_{1} for 0.7T/TN10.7\lesssim T/T_{\rm N}\leq 1 in the region where σ¯xxx\bar{\sigma}_{xxx} is drastically enhanced. This is attributed to the rapid increase of T~xMF\tilde{T}_{x}^{\rm MF} and resultant drastic change of the electronic band structure near the Fermi level. As σ¯xxx\bar{\sigma}_{xxx} in Eq. (9) includes the factors 2εn(𝒌)/kx2\partial^{2}\varepsilon_{n}({\bm{k}})/\partial k_{x}^{2} and εn(𝒌)/kx\partial\varepsilon_{n}({\bm{k}})/\partial k_{x}, the small X(𝒌)X(\bm{k}) appearing in the denominator of 2εn(𝒌)/kx2\partial^{2}\varepsilon_{n}({\bm{k}})/\partial k_{x}^{2} and εn(𝒌)/kx\partial\varepsilon_{n}({\bm{k}})/\partial k_{x} gives a dominant contribution. When considering the small order parameter compared to the ASOI, i.e., T~xMFα1\tilde{T}_{x}^{\rm MF}\lesssim\alpha_{1}, X(𝒌)X(\bm{k}) can become small when the Fermi wavenumber kxFk^{\rm F}_{x} satisfies T~xMFα1sinkxF\tilde{T}_{x}^{\rm MF}\simeq\alpha_{1}\sin k^{\rm F}_{x}, which results in a large enhancement of σ¯xxx\bar{\sigma}_{xxx}. Such an enhancement is remarkable when the upper and lower bands are closely located in the paramagnetic state with small X(𝒌)X(\bm{k}) as shown in Fig. 1(d), which can be realized for small t1=0.1t_{1}=0.1 and α2=0.1\alpha_{2}=0.1. In short, there are two conditions for large σ¯xxx\bar{\sigma}_{xxx}: One is the large essential parameters, such as α1\alpha_{1}, TxMFT_{x}^{\rm MF}, and JAFJ_{\rm AF}, and the other is to satisfy T~xMFα1sinkxF\tilde{T}_{x}^{\rm MF}\simeq\alpha_{1}\sin k^{\rm F}_{x} in a multi-band system. These conditions can be experimentally controlled by electron/hole doping and temperature.

The sign change of σ¯xxx\bar{\sigma}_{xxx} in TT dependence is owing to the multiband effect. As shown in Fig. 1(d), the band bottom is shifted in the opposite direction for the upper and lower bands, which means that the opposite sign of the coupling α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF} results in the opposite contribution to σ¯xxx\bar{\sigma}_{xxx}. This is demonstrated by decomposing σ¯xxx\bar{\sigma}_{xxx} into the upper- and lower-band contributions, as shown in Fig. 2(b). The results indicate that the dominant contribution of σ¯xxx\bar{\sigma}_{xxx} arises from the upper band for 0.9T/TN10.9\lesssim T/T_{\rm N}\leq 1, while that arises from the lower band for T/TN0.9T/T_{\rm N}\lesssim 0.9. The suppression of the upper-band contribution for low TT is because it becomes away from the Fermi level by the development of TxMFT_{x}^{\rm MF}.

IV.2 Transverse second-order conductivity σxzz\sigma_{xzz}

Refer to caption
Figure 3: The transverse second-order nonlinear conductivity σ¯xzz\bar{\sigma}_{xzz} for several α1\alpha_{1} and α2\alpha_{2} with α1=α2\alpha_{1}=\alpha_{2}. The inset represents σ¯zxx/(α1α22)\bar{\sigma}_{zxx}/(\alpha_{1}\alpha_{2}^{2}).

Next, let us discuss the transverse nonlinear conductivity σ¯xzz\bar{\sigma}_{xzz}. Figure 3 shows the TT dependence of σ¯xzz\bar{\sigma}_{xzz} for 0.02α1,α20.10.02\leq\alpha_{1},\alpha_{2}\leq 0.1 with α1=α2\alpha_{1}=\alpha_{2}. The behavior of σ¯xzz\bar{\sigma}_{xzz} against TT is similar to σ¯xxx\bar{\sigma}_{xxx} except for the sign change; σ¯xzz\bar{\sigma}_{xzz} becomes nonzero below T=TNT=T_{\rm N} and shows the maximum near TNT_{\rm N}. While decreasing TT, σ¯xzz\bar{\sigma}_{xzz} is suppressed and shows an almost constant value.

Similar to σxxx\sigma_{xxx}, the origin of nonzero σxzz\sigma_{xzz} is the asymmetric band modulation under TxMF0T_{x}^{\rm MF}\neq 0 via the effective coupling T~xMFα1\tilde{T}_{x}^{\rm MF}\alpha_{1}. Besides, we find another contribution from α2\alpha_{2} for nonzero σxzz\sigma_{xzz} in contrast to σxxx\sigma_{xxx}, where σ¯xzz\bar{\sigma}_{xzz} is well scaled by α1α22\alpha_{1}\alpha_{2}^{2} as shown in the inset of Fig. 3, as discussed in Sec. V.1. The additional parameter dependence for α22\alpha_{2}^{2} is owing to an additional symmetry between kzk_{z} and kz+πk_{z}+\pi for α2=0\alpha_{2}=0, which gives the opposite-sign contribution to σxzz\sigma_{xzz} so that totally σxzz=0\sigma_{xzz}=0.

IV.3 Comparison to magnetoelectric coefficient αyz\alpha_{yz}

Refer to caption
Figure 4: (a) The magnetoelectric coefficient α¯yz\bar{\alpha}_{yz} and (b) the quantity σ¯xzα¯yz\bar{\sigma}_{xz}\bar{\alpha}_{yz} with the same parameters as Fig. 3. σ¯xz\bar{\sigma}_{xz} is calculated by supposing the magnetic field Hy=0.01H_{y}=0.01. The insets of (a) and (b) represent α¯yz/α2\bar{\alpha}_{yz}/\alpha_{2} and σ¯xzα¯yz/(α1α22)\bar{\sigma}_{xz}\bar{\alpha}_{yz}/(\alpha_{1}\alpha_{2}^{2}), respectively.

We also present another MT-moment-driven phenomena, the magnetoelectric response, and compare its parameter and TT dependence to the nonlinear conductivities obtained in the previous section. Figure 4(a) shows the TT dependence of α¯yz\bar{\alpha}_{yz} for 0.02α1,α20.10.02\leq\alpha_{1},\alpha_{2}\leq 0.1 with α1=α2\alpha_{1}=\alpha_{2}, whose behavior is similar to the transverse nonlinear conductivity σxzz\sigma_{xzz} in Fig. 3 except for the sign. α¯yz\bar{\alpha}_{yz} is nonzero even if α1=0\alpha_{1}=0 that is different from the nonlinear conductivities, whereas α2\alpha_{2} and T~xMF\tilde{T}_{x}^{\rm MF} are essential to obtain the finite α¯yz\bar{\alpha}_{yz}, as detailed in Sec. V.1. As shown in the inset of Fig. 3(a), α¯yz\bar{\alpha}_{yz} is well scaled as α¯yz/α2\bar{\alpha}_{yz}/\alpha_{2} for small α2\alpha_{2}.

Moreover, it is noteworthy to comment on the relation between the transverse nonlinear conductivity and a combination of the linear magnetoelectric and Hall coefficients, since the nonlinear transverse transport in the 𝒫𝒯\mathcal{PT}-symmetric AFMs can be understood as the Hall transport driven by the induced magnetization through the linear magnetoelectric response at the phenomenological level [14, 21].

We show the TT dependence of σ¯xzα¯yz\bar{\sigma}_{xz}\bar{\alpha}_{yz} in Fig. 4(b) for the same parameters in Fig. 3. The small magnetic field Hy=0.01H_{y}=0.01 is introduced to mimic the induced magnetization in αyz\alpha_{yz}. Compared to the results in Fig. 3 and 4(b), one finds the resemblance between the TT dependences of σ¯xzz\bar{\sigma}_{xzz} and σ¯xzα¯yz\bar{\sigma}_{xz}\bar{\alpha}_{yz}, both of which are scaled by α1α22\alpha_{1}\alpha_{2}^{2}. A good qualitative correspondence in these responses indicates that the interpretation of dividing subsequent two linear processes for nonlinear conductivity is reasonable in the present model. The overall quantitative difference σ¯xzα¯yz/σ¯xzz102\bar{\sigma}_{xz}\bar{\alpha}_{yz}/\bar{\sigma}_{xzz}\sim 10^{-2} may be ascribed to the magnitude of the used internal magnetic field (Hy=0.01H_{y}=0.01) that should be replaced by the true internal field. However, it is hard to estimate it quantitatively.

V Discussion

V.1 Essential model parameters

We discuss the parameter dependences of the asymmetric band modulation, nonlinear conductivity, and the linear magnetoelectric and Hall coefficients at the level of the microscopic model Hamiltonian. For this purpose, we try to extract the essential parameters for each response from various hoppings, spin-orbit coupling, and internal/external field in the model Hamiltonian based on the method in Refs. 60, 61. In the following, we discuss the important model parameters in each case one by one, and the results are summarized in Table 1. The derivation is shown in Appendix B.

First, the essential parameters for the asymmetric band modulation [60] are given by T~xMFα1\tilde{T}_{x}^{\rm MF}\alpha_{1}, as shown in Appendix B.1. The result is consistent with the eigenvalues in Eq. (7).

Next, the essential model parameters for σxxx\sigma_{xxx} [61] (see also Appendix B.2) are given by

σxxx=\displaystyle\sigma_{xxx}= α1T~xMF[t12F(t1,t2,t3,α1,α2,T~xMF)\displaystyle\alpha_{1}\tilde{T}_{x}^{\rm MF}\left[t_{1}^{2}F(t_{1},t_{2},t_{3},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right.
+t2F(t1,t2,t3,α1,α2,T~xMF)],\displaystyle\left.\qquad\quad+t_{2}F^{\prime}(t_{1},t_{2},t_{3},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right], (15)

where FF and FF^{\prime} represent the arbitrary functions. Note that only the even power of α1\alpha_{1} and T~xMF\tilde{T}_{x}^{\rm MF} appears in FF and FF^{\prime}. Thus, one finds that the coupling of α1\alpha_{1} and T~xMF\tilde{T}_{x}^{\rm MF} is always necessary to induce σxxx\sigma_{xxx}, which is consistent with the numerical result presented in Sec. IV.1. Moreover, σxxx\sigma_{xxx} is closely related to the asymmetric band modulation because both of them are characterized by the same essential model parameters.

Similarly, the essential model parameters of σxzz\sigma_{xzz} are given by

σxzz=α1T~xMF[α22t2F(t1,t2,t3,α1,α2,T~xMF)],\displaystyle\sigma_{xzz}=\alpha_{1}\tilde{T}_{x}^{\rm MF}\left[\alpha_{2}^{2}t_{2}F(t_{1},t_{2},t_{3},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right], (16)

where the even power of α1\alpha_{1}, α2\alpha_{2}, and T~xMF\tilde{T}_{x}^{\rm MF} appears in FF. Equation (16) shows that the coupling of α1\alpha_{1} and T~xMF\tilde{T}_{x}^{\rm MF} is essential to induce σxzz\sigma_{xzz} as similar to σxxx\sigma_{xxx}, which is consistent with the numerical result in Sec. IV.2. Moreover, Eq. (16) indicates that t2t_{2} and even power of α2\alpha_{2} are also necessary for σxzz\sigma_{xzz} in the present model in Eq. (1).

In a similar way, the essential model parameters to induce αyz\alpha_{yz} and σxz\sigma_{xz} are given by

αyz\displaystyle\alpha_{yz} =α2T~xMF[t3F(t1,t2,t3,α1,α2,T~xMF)],\displaystyle=\alpha_{2}\tilde{T}_{x}^{\rm MF}\left[t_{3}F(t_{1},t_{2},t_{3},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right], (17)
σxz\displaystyle\sigma_{xz} =α1α2Hy[t3F(t1,t2,t3,α1,α2,Hy,T~xMF)].\displaystyle=\alpha_{1}\alpha_{2}H_{y}\left[t_{3}F(t_{1},t_{2},t_{3},\alpha_{1},\alpha_{2},H_{y},\tilde{T}_{x}^{\rm MF})\right]. (18)

This indicates that nonzero αyzσxz\alpha_{yz}\sigma_{xz} needs nonzero α1α22T~xMF\alpha_{1}\alpha_{2}^{2}\tilde{T}_{x}^{\rm MF}, which shows a good agreement with the condition for σxzz\sigma_{xzz}. The common essential model-parameter dependence in small parameter region was already confirmed in Secs. IV.2 and IV.3.

It is noteworthy that the above approach to extract the essential model parameters can be straightforwardly applied even when introducing the other model parameters. For example, let us consider the additional interlayer A-B hopping t4t_{4} in the model Hamiltonian. In this situation, one finds that there is no longer simple correlation between σxzz\sigma_{xzz} and σxzαyz\sigma_{xz}\alpha_{yz}; the essential model parameters for the former are α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF} rather than α1α22T~xMF\alpha_{1}\alpha_{2}^{2}\tilde{T}_{x}^{\rm MF}, while those for the latter still remains the same as α1α22T~xMF\alpha_{1}\alpha_{2}^{2}\tilde{T}_{x}^{\rm MF} as discussed in Appendix B. In other words, the factor α22t2\alpha^{2}_{2}t_{2} in the square bracket in Eq. (16) is not truly the essential factor. Indeed, the numerical results in the presence of t4t_{4} give a different temperature dependence from each other, as shown in Appendix C. Thus, the correspondence between σxzz\sigma_{xzz} and σxzαyz\sigma_{xz}\alpha_{yz} occurs depending on the hopping in the effective model, which is clarified by performing a procedure in Appendix B.

Table 1: Model parameters necessary for the asymmetric band modulation and response tensors indicated by the checkmark (✓). In the last two columns, model parameters are decomposed into the essential and semi-essential parts.
t2t_{2} t3t_{3} α1\alpha_{1} α2\alpha_{2} T~xMF\tilde{T}_{x}^{\rm MF} HyH_{y} essential semi-essential
asymmetric α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF}
band modulation
σxxx(t4=0)\sigma_{xxx}\,\,(t_{4}=0) α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF} t12,t2t_{1}^{2},t_{2}
σxxx(t40)\sigma_{xxx}\,\,(t_{4}\neq 0) α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF} t12,t2,t4t_{1}^{2},t_{2},t_{4}
σxzz(t4=0)\sigma_{xzz}\,\,(t_{4}=0) α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF} α22t2\alpha_{2}^{2}t_{2}
σxzz(t40)\sigma_{xzz}\,\,(t_{4}\neq 0) α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF} α22t2,t4\alpha_{2}^{2}t_{2},t_{4}
αyz(t4=0)\alpha_{yz}\,\,(t_{4}=0) α2T~xMF\alpha_{2}\tilde{T}_{x}^{\rm MF} t3t_{3}
αyz(t40)\alpha_{yz}\,\,(t_{4}\neq 0) α2T~xMF\alpha_{2}\tilde{T}_{x}^{\rm MF} t3,t4t_{3},t_{4}
σxz(t4=0)\sigma_{xz}\,\,(t_{4}=0) α1α2Hy\alpha_{1}\alpha_{2}H_{y} t3t_{3}
σxz(t40)\sigma_{xz}\,\,(t_{4}\neq 0) α1α2Hy\alpha_{1}\alpha_{2}H_{y} t3,t4t_{3},t_{4}

V.2 Quantitative evaluation

Finally, we discuss the order estimate of the nonlinear conductivity for α1=0.5\alpha_{1}=0.5 and α2=0.1\alpha_{2}=0.1 by the ratio σxxx/(σxx)2\sigma_{xxx}/(\sigma_{xx})^{2} with being independent of the relaxation time in the clean limit. By putting the typical values as a0.5a\sim 0.5 [nm] and |t2|=0.2|t_{2}|=0.2 eV, we obtain σxxx/(σxx)2103a2e1|t2|11018\sigma_{xxx}/(\sigma_{xx})^{2}\sim 10^{-3}\hbar a^{2}e^{-1}|t_{2}|^{-1}\sim 10^{-18} [m3 A-1] for T0T\to 0 and 101710^{-17} [m3 A-1] near TNT_{\rm N}, which is comparable to the value in the 2D nonmagnetic Rashba system under the magnetic field [51]. Further enhancement can be achieved by tuning the model parameters and electron filling.

VI Summary

In summary, we investigated the microscopic essence for the second-order nonlinear conductivity in the 𝒫𝒯\mathcal{PT}-symmetric collinear AFM with the MT moment on a two-dimensionally stacked zigzag chain by focusing on the role of the MT moment. Based on the nonlinear Kubo formula in the clean limit, we found that the effective coupling between the ASOI and the MT moment is essential for the nonlinear conductivity. By analyzing both the longitudinal and transverse components of the nonlinear conductivity while changing the ASOI and the temperature, we showed that their large enhancement can be achieved near the transition temperature, provided that the AFM molecular field is comparable to the ASOI in a multi-band system. We also discussed the similarity and difference between the transverse nonlinear transport and the combined response consisting of the linear magnetoelectric and Hall coefficients.

The present result elucidates the essential model parameters for MT-related physical phenomena, such as the nonlinear conductivity and the linear magnetoelectric effect, in 𝒫𝒯\mathcal{PT}-symmetric collinear AFMs. The similar analysis can be applied to examine the role of the MT moment for any collinear AFMs with the MT moment in the zigzag structure, e.g., CeRu2Al10 [62, 63], Ce3TiBi5 [18, 17], and α\alpha-YbAl1-xMnxB4 [64], and other ferrotoroidal metals/semiconductors with locally noncentrosymmetric crystal structures, such as Mn2Au [65, 46], RRB4 (R=R=Dy, Er) [66, 67], CuMnAs [68, 45], PrMnSbO [69], NdMnAsO [70], and XyX_{y}Fe2-xSe2 (X=X=K, Tl, Rb) [71, 72, 73], once the model Hamiltonian is given. The measurements of the linear magnetoelectric effect and the nonlinear conductivity for these materials are also useful to investigate their microscopic mechanisms. Moreover, the analysis is straightforwardly extended to the AFMs with the other odd-parity magnetic-type multipole moments, such as the MT octupole, since they are characterized by the same spatial inversion and time-reversal symmetries. Our study will stimulate a further investigation of the multiferroic and conductive phenomena in the 𝒫𝒯\mathcal{PT}-symmetric AFM metals.

Acknowledgements.
We thank Y. Motome and Y. Yanagi for the fruitful discussion. This research was supported by JSPJ KAKENHI Grant Numbers JP19K03752, JP19H01834, JP21H01037, and by JST PRESTO (JPMJPR20L8). M.Y. and R.O. are supported by a JSPS research fellowship and JSPS KAKENHI (Grant No. JP20J12026 and JP20J21838).

Appendix A Expressions of multipoles

We show the functional form of multipoles with rank 0–3 except the normalization constant: the rank 0 (monopole) is

X01,\displaystyle X_{0}\propto 1, (19)

the rank 1 (dipole) is

(Xx,Xy,Xz)(x,y,z),\displaystyle(X_{x},X_{y},X_{z})\propto(x,y,z), (20)

the rank 2 (quadrupole) is

Xu\displaystyle X_{u} 3z2r2,\displaystyle\propto 3z^{2}-r^{2}, (21)
Xv\displaystyle X_{v} x2y2,\displaystyle\propto x^{2}-y^{2}, (22)
(Xyz,Xzx,Xxy)\displaystyle(X_{yz},X_{zx},X_{xy}) (yz,zx,xy),\displaystyle\propto(yz,zx,xy), (23)

the rank 3 (octupole) is

Xxyz\displaystyle X_{xyz} xyz,\displaystyle\propto xyz, (24)
(Xxα,Xyα,Xzα)\displaystyle(X_{x}^{\alpha},X_{y}^{\alpha},X_{z}^{\alpha}) (x(5x23r2),y(5y23r2),z(5z23r2)),\displaystyle\propto\left(x(5x^{2}-3r^{2}),y(5y^{2}-3r^{2}),z(5z^{2}-3r^{2})\right), (25)
(Xxβ,Xyβ,Xzβ)\displaystyle(X_{x}^{\beta},X_{y}^{\beta},X_{z}^{\beta}) (x(y2z2),y(z2x2),z(x2y2)),\displaystyle\propto\left(x(y^{2}-z^{2}),y(z^{2}-x^{2}),z(x^{2}-y^{2})\right), (26)

where XX represents the types of multipoles. When XX corresponds to the time-reversal-odd polar (axial) tensor, it stands for TT (MM) for MT (magnetic) multipole.

By using the multipole notation, the collinear AFM with 𝒒=𝟎{\bm{q}}={\bm{0}} on a zigzag chain are represented by the MT dipole TzT_{z} when the AFM moment is along the xx direction as

Tz\displaystyle T_{z} =12l=A,B(RlxσlyRlyσlx)12(σBxσAx),\displaystyle=\frac{1}{2}\sum_{l={\rm A},{\rm B}}\left({R}_{l}^{x}\sigma_{l}^{y}-{R}_{l}^{y}\sigma_{l}^{x}\right)\to\frac{1}{2}\left(\sigma_{\rm B}^{x}-\sigma_{\rm A}^{x}\right), (27)

where σlμ\sigma_{l}^{\mu} and Rlμ{R}_{l}^{\mu} (μ=x,y,z\mu=x,y,z) are the magnetic moment and the position vector at llth atom, respectively [74]. Similarly, the AFM with the moment along the yy direction is characterized by the magnetic quadrupole MuM_{u} as

Mu\displaystyle M_{u} =l=A,B(2RlzσlzRlxσlxRlyσly)12(σByσAy),\displaystyle=\sum_{l={\rm A},{\rm B}}\left(2{R}_{l}^{z}\sigma_{l}^{z}-{R}_{l}^{x}\sigma_{l}^{x}-{R}_{l}^{y}\sigma_{l}^{y}\right)\to\frac{1}{2}\left(\sigma_{\rm B}^{y}-\sigma_{\rm A}^{y}\right), (28)

and that along the zz direction is by the MT dipole TxT_{x} as

Tx\displaystyle T_{x} =12l=A,B(RlyσlzRlzσly)12(σAzσBz).\displaystyle=\frac{1}{2}\sum_{l={\rm A},{\rm B}}\left({R}_{l}^{y}\sigma_{l}^{z}-{R}_{l}^{z}\sigma_{l}^{y}\right)\to\frac{1}{2}\left(\sigma_{\rm A}^{z}-\sigma_{\rm B}^{z}\right). (29)

Moreover, the dipole and octupole components of σμνλ\sigma_{\mu\nu\lambda} in Eq. (III.1) are related to the MT dipoles in Eq. (20) and MT octupoles in Eqs. (24)–(26) as follows:

Tx\displaystyle T_{x} 115ν=x,y,z(σxνν+2σννx)(cyclic),\displaystyle\leftrightarrow\frac{1}{15}\sum_{\nu=x,y,z}(\sigma_{x\nu\nu}+2\sigma_{\nu\nu x})\ (\mbox{cyclic}), (30)
Txyz\displaystyle T_{xyz} σxyz,\displaystyle\leftrightarrow\sigma_{xyz}, (31)
Txα\displaystyle T_{x}^{\alpha} 110(5σxxx3ν=x,y,zσxνν)(cyclic),\displaystyle\leftrightarrow\frac{1}{10}\left(5\sigma_{xxx}-3\sum_{\nu=x,y,z}\sigma_{x\nu\nu}\right)\ (\mbox{cyclic}), (32)
Txβ\displaystyle T_{x}^{\beta} 12(σxyyσzzx)(cyclic).\displaystyle\leftrightarrow\frac{1}{2}\left(\sigma_{xyy}-\sigma_{zzx}\right)\ (\mbox{cyclic}). (33)

Appendix B Essential model parameters in response tensors

We show the essential model parameters for the asymmetric band modulation, the longitudinal and transverse nonlinear conductivities, and the linear Hall and magnetoelectric coefficients, by using the systematic analysis method in Refs. [60] and [61]. The results are summarized in Table 1.

B.1 Asymmetric band modulation

First, we give the essential model parameters for the asymmetric band modulation. Following the method for extracting the essential model parameters in the thermal average of an hermitian operator [60, 61], we obtain the momentum distribution of the band modulation and its parameter dependences by analytically evaluating the low-order contributions of the following quantity,

Ωi(𝒌)=Tr[hi+1(𝒌)].\displaystyle\Omega^{i}(\bm{k})=\mathrm{Tr}\left[h^{i+1}(\bm{k})\right]. (34)

Here hi+1(𝒌)h^{i+1}(\bm{k}) denotes the (i+1)(i+1)-th power of the Hamiltonian matrix at wave vector 𝒌\bm{k}, i.e., \mathcal{H} in Eq. (1) is represented as =𝒌h(𝒌)\mathcal{H}=\sum_{\bm{k}}h({\bm{k}}). The 0th- and 1st-order contributions Ω0(𝒌)\Omega^{0}(\bm{k}) and Ω1(𝒌)\Omega^{1}(\bm{k}) are explicitly given by

Ω0(𝒌)=8(t2coskx+t3coskz),\displaystyle\Omega^{0}(\bm{k})=-8\left(t_{2}\cos k_{x}+t_{3}\cos k_{z}\right), (35)
Ω1(𝒌)=8α1T~xMFsinkx+4[(T~xMF)2+α12sin2kx+α22sin2kz+2t12(1+coskx)+4(t2coskx+t3coskz)2].\displaystyle\Omega^{1}(\bm{k})=-8\alpha_{1}\tilde{T}_{x}^{\rm MF}\sin k_{x}+4\left[\left(\tilde{T}_{x}^{\rm MF}\right)^{2}+\alpha_{1}^{2}\sin^{2}k_{x}+\alpha_{2}^{2}\sin^{2}k_{z}+2t_{1}^{2}(1+\cos k_{x})+4\left(t_{2}\cos k_{x}+t_{3}\cos k_{z}\right)^{2}\right]. (36)

The odd function of kxk_{x} appears only in the first term of Eq. (36) in the form proportional to α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF}, which means that the asymmetric band structure is induced by the coupling between the nonzero T~xMF\tilde{T}_{x}^{\rm MF} and α1\alpha_{1}. It is confirmed at least to the 6th order. Note that the odd functions of kxk_{x} included in the higher-order terms in Eq. (34) are always proportional to α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF}. Thus, both α1\alpha_{1} and T~xMF\tilde{T}_{x}^{\rm MF} are the essential model parameters for the asymmetric band structure, and their coupling is also crucial for nonlinear conductivities.

B.2 Second-order nonlinear conductivity

Next, we elucidate the essential model parameters in the longitudinal and transverse nonlinear conductivities. The essential model parameters in the Drude-type nonlinear conductivities can be extracted by evaluating the following quantity [61],

Re[Γμνλijk]=𝒌Re{Tr[\varv^μ𝒌hi(𝒌)\varv^ν𝒌hj(𝒌)\varv^λ𝒌hk(𝒌)]},\displaystyle\mathrm{Re}\left[\Gamma^{ijk}_{\mu\nu\lambda}\right]=\sum_{\bm{k}}\mathrm{Re}\left\{\mathrm{Tr}\left[\hat{\varv}_{\mu\bm{k}}h^{i}(\bm{k})\hat{\varv}_{\nu\bm{k}}h^{j}(\bm{k})\hat{\varv}_{\lambda\bm{k}}h^{k}(\bm{k})\right]\right\}, (37)

where \varv^μ𝒌\hat{\varv}_{\mu\bm{k}} denotes the μ\mu component of the velocity operator at 𝒌\bm{k}.

Refer to caption
Figure 5: (a) Schematic picture of the interlayer hopping t4t_{4} between A and B sublattices. (b), (c) The TT dependence of (b) σ¯xzz\bar{\sigma}_{xzz} and (c) σ¯xzα¯yz\bar{\sigma}_{xz}\bar{\alpha}_{yz} for (t4,α2)=(0.1,0)(t_{4},\alpha_{2})=(0.1,0), (0.1,0.1)(0.1,0.1), and (0.05,0.1)(0.05,0.1).

Here, we introduce the interlayer hopping between the sublattices A and B [Fig. 5(a)]. The effect of the additional hopping is taken into account by replacing εAB(𝒌)\varepsilon^{\rm AB}({\bm{k}}) as 2t1cos(kxa/2)2[t1+2t4cos(kzc)]cos(kxa/2)-2t_{1}\cos{(k_{x}a/2)}\to-2[t_{1}+2t_{4}\cos{(k_{z}c)}]\cos{(k_{x}a/2)}. The results of the evaluations are given as follows.

  • -

    Longitudinal nonlinear conductivity σxxx\sigma_{xxx}
    As the essential model parameters are included in any pairs of (i,j,k)(i,j,k) in Eq. (37), we here show two low-order contributions to Eq. (37) in the (i,j,k)=(0,0,1)(i,j,k)=(0,0,1) and (0,1,3)(0,1,3) terms as representative examples, which are explicitly given by

    Re[Γxxx001]=α1T~xMF(t12+2t42),\displaystyle\mathrm{Re}\left[\Gamma^{001}_{xxx}\right]=\alpha_{1}\tilde{T}_{x}^{\rm MF}\left(t_{1}^{2}+2t_{4}^{2}\right), (38)
    Re[Γxxx013]=4α1T~xMF(t2{α12α22+t12[4(T~xMF)2+7α12+2α22+3t12]}\displaystyle\mathrm{Re}\left[\Gamma^{013}_{xxx}\right]=4\alpha_{1}\tilde{T}_{x}^{\rm MF}\left(t_{2}\left\{\alpha_{1}^{2}\alpha_{2}^{2}+t_{1}^{2}\left[4\left(\tilde{T}_{x}^{\rm MF}\right)^{2}+7\alpha_{1}^{2}+2\alpha_{2}^{2}+3t_{1}^{2}\right]\right\}\right. (39)
    +t4[4(T~xMF)2t1t3+5α12t1t3α22t1t316t13t312t1t22t312t1t33\displaystyle\hskip 85.35826pt+t_{4}\left[-4\left(\tilde{T}_{x}^{\rm MF}\right)^{2}t_{1}t_{3}+5\alpha_{1}^{2}t_{1}t_{3}-\alpha_{2}^{2}t_{1}t_{3}-16t_{1}^{3}t_{3}-12t_{1}t_{2}^{2}t_{3}-12t_{1}t_{3}^{3}\right. (40)
    +8(T~xMF)2t2t4+14α12t2t4+2α22t2t4+36t12t2t448t1t3t43+18t2t43]).\displaystyle\hskip 85.35826pt\quad\left.\left.+8\left(\tilde{T}_{x}^{\rm MF}\right)^{2}t_{2}t_{4}+14\alpha_{1}^{2}t_{2}t_{4}+2\alpha_{2}^{2}t_{2}t_{4}+36t_{1}^{2}t_{2}t_{4}-48t_{1}t_{3}t_{4}^{3}+18t_{2}t_{4}^{3}\right]\right). (41)

    Then, the essential model parameters in the longitudinal nonlinear conductivity σxxx\sigma_{xxx} are α1\alpha_{1} and T~xMF\tilde{T}_{x}^{\rm MF}, which is consistent with the fact that the nonzero σxxx\sigma_{xxx} is closely related to the asymmetric band structure under TxMF0T_{x}^{\rm MF}\neq 0. Since all the terms in Eq. (37) are always proportional to α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF}, σxxx\sigma_{xxx} is written in the form:

    σxxx=α1T~xMF[t12F(t1,t2,t3,t4,α1,α2,T~xMF)+t2F(t1,t2,t3,t4,α1,α2,T~xMF)+t4F′′(t1,t2,t3,t4,α1,α2,T~xMF)],\displaystyle\sigma_{xxx}=\alpha_{1}\tilde{T}_{x}^{\mathrm{MF}}\left[t_{1}^{2}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})+t_{2}F^{\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})+t_{4}F^{\prime\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right], (42)

    where the even power of α1\alpha_{1} and T~xAF\tilde{T}_{x}^{\rm AF} appears in FF, FF^{\prime}, and F′′F^{\prime\prime}, e.g., α12\alpha_{1}^{2} and (T~xMF)2(\tilde{T}_{x}^{\rm MF})^{2} in Eq. (41). By introducing t40t_{4}\neq 0, the additional contribution appears, which results in the alternative behavior of σxxx\sigma_{xxx}.

  • -

    Transverse nonlinear conductivity σxzz\sigma_{xzz}
    Similar to σxxx\sigma_{xxx}, we show two low-order contributions to Eq. (37) in the (i,j,k)=(0,1,0)(i,j,k)=(0,1,0) and (0,1,1)(0,1,1) terms for example. The expressions are given by

    Re[Γxzz010]=24225α1T~xMFt42,\displaystyle\mathrm{Re}\left[\Gamma^{010}_{xzz}\right]=-\frac{242}{25}\alpha_{1}\tilde{T}_{x}^{\rm MF}t_{4}^{2}, (43)
    Re[Γxzz011]=12125α1T~xMF[α22t2+t4(4t1t3+8t2t4)].\displaystyle\mathrm{Re}\left[\Gamma^{011}_{xzz}\right]=\frac{121}{25}\alpha_{1}\tilde{T}_{x}^{\rm MF}\left[\alpha_{2}^{2}t_{2}+t_{4}\left(4t_{1}t_{3}+8t_{2}t_{4}\right)\right]. (44)

    Similar to this result, we find that all the terms in Eq. (37) are always proportional to α1T~xMF\alpha_{1}\tilde{T}_{x}^{\rm MF}, then σxzz\sigma_{xzz} is expressed as

    σxzz=\displaystyle\sigma_{xzz}= α1T~xMF[α22t2F(t1,t2,t3,t4,α1,α2,T~xMF)\displaystyle\alpha_{1}\tilde{T}_{x}^{\rm MF}\left[\alpha_{2}^{2}t_{2}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right.
    +t4F(t1,t2,t3,t4,α1,α2,T~xMF)],\displaystyle\left.\qquad\quad+t_{4}F^{\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right], (45)

    where the second term proportional to t4t_{4} does not vanish even for α2=0\alpha_{2}=0.

B.3 Linear responses

We further clarify the essential model parameters for the linear Hall and magnetoelectric coefficients. The essential model parameters in the inter-band contribution of the electric-field induced response tensors can be extracted by evaluating the following quantity [61],

Im[Γμνij]=𝒌Im{Tr[A^μ𝒌hi(𝒌)\varv^ν𝒌hj(𝒌)]},\displaystyle\mathrm{Im}\left[\Gamma^{ij}_{\mu\nu}\right]=\sum_{\bm{k}}\mathrm{Im}\left\{\mathrm{Tr}\left[\hat{A}_{\mu\bm{k}}h^{i}(\bm{k})\hat{\varv}_{\nu\bm{k}}h^{j}(\bm{k})\right]\right\}, (46)

where A^μ𝒌\hat{A}_{\mu\bm{k}} denotes the μ\mu component of an arbitrary hermitian operator at 𝒌\bm{k}.

  • -

    Magnetoelectric coefficient αyz\alpha_{yz}
    The magnetoelectric coefficient αyz\alpha_{yz} corresponds to the case with A^μ𝒌=σy\hat{A}_{\mu\bm{k}}=\sigma_{y} in Eq. (46). Similar to the nonlinear conductivities, the essential model parameters are included in any pairs of (i,j)(i,j) in Eq. (46). We show two cases by taking (i,j)=(0,2)(i,j)=(0,2) and (1,3)(1,3), which are given by

    Im[Γyz02]=445α2T~xMFt3,\displaystyle\mathrm{Im}\left[\Gamma_{yz}^{02}\right]=-\frac{44}{5}\alpha_{2}\tilde{T}_{x}^{\rm MF}t_{3}, (47)
    Im[Γyz13]=115α2T~xMF{t3[4(T~xMF)2+6α12+α22+8t1224t2212t2]+t4(16t1t2+24t3t4)}.\displaystyle\mathrm{Im}\left[\Gamma^{13}_{yz}\right]=\frac{11}{5}\alpha_{2}\tilde{T}_{x}^{\rm MF}\left\{t_{3}\left[4\left(\tilde{T}_{x}^{\rm MF}\right)^{2}+6\alpha_{1}^{2}+\alpha_{2}^{2}+8t_{1}^{2}-24t_{2}^{2}-12t_{2}\right]+t_{4}\left(16t_{1}t_{2}+24t_{3}t_{4}\right)\right\}. (48)

    We also find that all the terms in Eq. (46) are always proportional to α2T~xMF\alpha_{2}\tilde{T}_{x}^{\rm MF}, then αyz\alpha_{yz} is expressed as

    αyz=\displaystyle\alpha_{yz}= α2T~xMF[t3F(t1,t2,t3,t4,α1,α2,T~xMF)\displaystyle\alpha_{2}\tilde{T}_{x}^{\rm MF}\left[t_{3}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right.
    +t4F(t1,t2,t3,t4,α1,α2,T~xMF)].\displaystyle\left.\qquad\quad+t_{4}F^{\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})\right]. (49)

    Therefore, the essential model parameters are α2\alpha_{2} and T~xMF\tilde{T}_{x}^{\rm MF}, while αyz\alpha_{yz} also depends on the spin-independent hopping process t3t_{3} or t4t_{4}.

  • -

    Hall coefficient σxz\sigma_{xz}
    In order to discuss σxz\sigma_{xz}, we introduce the small magnetic field along the yy direction HyH_{y}. Then, we evaluate the essential model parameters for the Hall coefficient σxz\sigma_{xz} with A^μ𝒌=\varv^x𝒌\hat{A}_{\mu\bm{k}}=\hat{\varv}_{x\bm{k}} in Eq. (46). We show two low-order contributions to Eq. (46) in the (i,j)=(0,3)(i,j)=(0,3) and (1,3)(1,3) terms for example, which are given by

    Im[Γxz03]=445α1α2Hy(3t2t3+5t1t4),\displaystyle\mathrm{Im}\left[\Gamma^{03}_{xz}\right]=\frac{44}{5}\alpha_{1}\alpha_{2}H_{y}\left(3t_{2}t_{3}+5t_{1}t_{4}\right), (50)
    Im[Γxz13]=885α1α2Hy[2t12t3+t4(8t1t2+7t3t4)].\displaystyle\mathrm{Im}\left[\Gamma^{13}_{xz}\right]=\frac{88}{5}\alpha_{1}\alpha_{2}H_{y}\left[2t_{1}^{2}t_{3}+t_{4}\left(8t_{1}t_{2}+7t_{3}t_{4}\right)\right]. (51)

    All the terms in Eq. (46) are always proportional to α1α2Hy\alpha_{1}\alpha_{2}H_{y}, then σxz\sigma_{xz} is expressed as

    σxz=\displaystyle\sigma_{xz}= α1α2Hy[t3F(t1,t2,t3,t4,α1,α2,Hy,T~xMF)\displaystyle\alpha_{1}\alpha_{2}H_{y}\left[t_{3}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},H_{y},\tilde{T}_{x}^{\rm MF})\right.
    +t4F(t1,t2,t3,t4,α1,α2,Hy,T~xMF)].\displaystyle\left.\qquad\quad+t_{4}F^{\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},H_{y},\tilde{T}_{x}^{\rm MF})\right]. (52)

Therefore, the essential model parameters are α1\alpha_{1}, α2\alpha_{2}, and HyH_{y}, while σxz\sigma_{xz} also depends on the spin-independent hopping along the zz direction, t3t_{3} or t4t_{4}.

By combining the results, Eqs. (49) and (52), σxzαyz\sigma_{xz}\alpha_{yz} has the form:

σxzαyz=α1α22T~xMFHy[t32F(t1,t2,t3,t4,α1,α2,Hy,T~xMF)+t42F(t1,t2,t3,t4,α1,α2,Hy,T~xMF)+t3t4F′′(t1,t2,t3,t4,α1,α2,Hy,T~xMF)],\displaystyle\sigma_{xz}\alpha_{yz}=\alpha_{1}\alpha_{2}^{2}\tilde{T}_{x}^{\rm MF}H_{y}\left[t_{3}^{2}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},H_{y},\tilde{T}_{x}^{\rm MF})+t_{4}^{2}F^{\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},H_{y},\tilde{T}_{x}^{\rm MF})+t_{3}t_{4}F^{\prime\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},H_{y},\tilde{T}_{x}^{\rm MF})\right], (53)

which clearly shows that σxzαyzα1α22T~xMFHy\sigma_{xz}\alpha_{yz}\propto\alpha_{1}\alpha_{2}^{2}\tilde{T}_{x}^{\rm MF}H_{y} irrespective of the additional parameter of t4t_{4}.

When t4=0t_{4}=0, we find that both σxzz\sigma_{xzz} and σxzαyz\sigma_{xz}\alpha_{yz} are proportional to α1α22T~xMF\alpha_{1}\alpha_{2}^{2}\tilde{T}_{x}^{\rm MF}. On the other hand, such relation does not hold when t40t_{4}\neq 0; σxzzα1T~xMF\sigma_{xzz}\propto\alpha_{1}\tilde{T}_{x}^{\rm MF}, whereas σxzαyzα1α22T~xMF\sigma_{xz}\alpha_{yz}\propto\alpha_{1}\alpha_{2}^{2}\tilde{T}_{x}^{\rm MF}.

Appendix C Effect of additional interlayer hopping

We compare the transverse component of the nonlinear conductivity σxzz\sigma_{xzz} and the quantity σxzαyz\sigma_{xz}\alpha_{yz} in the presence of the interlayer hopping t4t_{4} between the sublattices A and B.

Figures 5(b) and 5(c) show σ¯xzz\bar{\sigma}_{xzz} and σ¯xzα¯yz\bar{\sigma}_{xz}\bar{\alpha}_{yz} as functions of TT, respectively, for t4=0.1,0.05t_{4}=0.1,0.05 and α2=0,0.1\alpha_{2}=0,0.1, where α1=0.4\alpha_{1}=0.4 is used. As shown by the red dashed line in Fig. 5(b), σ¯xzz\bar{\sigma}_{xzz} still remains nonzero even for α2=0\alpha_{2}=0, while σ¯xzα¯yz\bar{\sigma}_{xz}\bar{\alpha}_{yz} in Fig. 5(c) vanishes. Furthermore, the nonzero t4t_{4} enhances σ¯xzz\bar{\sigma}_{xzz}, while it suppresses σ¯xzα¯yz\bar{\sigma}_{xz}\bar{\alpha}_{yz} while increasing t4t_{4}. This is because the essential model parameters discussed in the previous section are different for σxzz\sigma_{xzz} and σxzαyz\sigma_{xz}\alpha_{yz}. Indeed, in the presence of t4t_{4} and α2\alpha_{2}, the essential model parameter of σxzz\sigma_{xzz} is represented as α1T~xMF[α22t2F(t1,t2,t3,t4,α1,α2,T~xMF)+t4F(t1,t2,t3,t4,α1,α2,T~xMF)]\alpha_{1}\tilde{T}_{x}^{\rm MF}[\alpha_{2}^{2}t_{2}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})+t_{4}F^{\prime}(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF})], which clearly shows that σxzz\sigma_{xzz} has the additional contribution from t4t_{4} and does not vanish for α2=0\alpha_{2}=0. On the other hand, the essential model parameters of σxz\sigma_{xz} and αyz\alpha_{yz} does not show the change from σxzα1α2HyF(t1,t2,t3,t4,α1,α2,Hy,T~xMF)\sigma_{xz}\to\alpha_{1}\alpha_{2}H_{y}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},H_{y},\tilde{T}_{x}^{\rm MF}) and αyzα2T~xMFF(t1,t2,t3,t4,α1,α2,T~xMF)\alpha_{yz}\to\alpha_{2}\tilde{T}_{x}^{\rm MF}F(t_{1},t_{2},t_{3},t_{4},\alpha_{1},\alpha_{2},\tilde{T}_{x}^{\rm MF}), respectively; the hopping t4t_{4} is not the essential model parameter for σxz\sigma_{xz} and αyz\alpha_{yz}. Thus, there is no simple relation between them.

References

  • Nagaosa et al. [2010] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Rev. Mod. Phys. 82, 1539 (2010).
  • Xiao et al. [2010] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys. 82, 1959 (2010).
  • Šmejkal et al. [2022] L. Šmejkal, A. H. MacDonald, J. Sinova, S. Nakatsuji, and T. Jungwirth, Nat. Rev. Mater.  (2022).
  • Solovyev [1997] I. V. Solovyev, Phys. Rev. B 55, 8060 (1997).
  • Li et al. [2019] X. Li, A. H. MacDonald, and H. Chen, arXiv:1902.10650  (2019).
  • Naka et al. [2020] M. Naka, S. Hayami, H. Kusunose, Y. Yanagi, Y. Motome, and H. Seo, Phys. Rev. B 102, 075112 (2020).
  • Hayami and Kusunose [2021a] S. Hayami and H. Kusunose, Phys. Rev. B 103, L180407 (2021a).
  • Popov et al. [1999] Y. F. Popov, A. Kadomtseva, D. Belov, G. Vorob’ev, and A. Zvezdin, J. Exp. Theor. Phys. Lett. 69, 330 (1999).
  • Popov et al. [1998] Y. F. Popov, A. Kadomtseva, G. Vorob’ev, V. Timofeeva, D. Ustinin, A. Zvezdin, and M. Tegeranchi, J. Exp. Theor. Phys. 87, 146 (1998).
  • Arima et al. [2005] T.-h. Arima, J.-H. Jung, M. Matsubara, M. Kubota, J.-P. He, Y. Kaneko, and Y. Tokura, J. Phys. Soc. Jpn. 74, 1419 (2005).
  • Van Aken et al. [2007] B. B. Van Aken, J.-P. Rivera, H. Schmid, and M. Fiebig, Nature 449, 702 (2007).
  • Zimmermann et al. [2014] A. S. Zimmermann, D. Meier, and M. Fiebig, Nat. Commun. 5, 4796 (2014).
  • Toledano et al. [2011] P. Toledano, D. D. Khalyavin, and L. C. Chapon, Phys. Rev. B 84, 094421 (2011).
  • Hayami et al. [2014a] S. Hayami, H. Kusunose, and Y. Motome, Phys. Rev. B 90, 024432 (2014a).
  • Hayami et al. [2015a] S. Hayami, H. Kusunose, and Y. Motome, J. Phys.: Conf. Ser. 592, 012101 (2015a).
  • Saito et al. [2018] H. Saito, K. Uenishi, N. Miura, C. Tabata, H. Hidaka, T. Yanagisawa, and H. Amitsuka, J. Phys. Soc. Jpn. 87, 033702 (2018).
  • Shinozaki et al. [2020a] M. Shinozaki, G. Motoyama, T. Mutou, S. Nishigori, A. Yamaguchi, K. Fujiwara, K. Miyoshi, and A. Sumiyama, JPS Conf. Proc. 30, 011189 (2020a).
  • Shinozaki et al. [2020b] M. Shinozaki, G. Motoyama, M. Tsubouchi, M. Sezaki, J. Gouchi, S. Nishigori, T. Mutou, A. Yamaguchi, K. Fujiwara, K. Miyoshi, and Y. Uwatoko, J. Phys. Soc. Jpn. 89, 033703 (2020b).
  • Tokura and Nagaosa [2018] Y. Tokura and N. Nagaosa, Nat. Commun. 9, 3740 (2018).
  • Ma et al. [2021] Q. Ma, A. G. Grushin, and K. S. Burch, Nat. Mater.  (2021).
  • Oyamadaa et al. [2018] A. Oyamadaa, T. Inoharaa, E. Yamamotob, and Y. Hagab, Prog. Nucl. Sci. Techn. 5, 128 (2018).
  • Watanabe and Yanase [2020] H. Watanabe and Y. Yanase, Phys. Rev. Research 2, 043081 (2020).
  • Watanabe and Yanase [2021] H. Watanabe and Y. Yanase, Phys. Rev. X 11, 011001 (2021).
  • Yanase [2014] Y. Yanase, J. Phys. Soc. Jpn. 83, 014703 (2014).
  • Hayami et al. [2014b] S. Hayami, H. Kusunose, and Y. Motome, Phys. Rev. B 90, 081115(R) (2014b).
  • Hayami et al. [2016] S. Hayami, H. Kusunose, and Y. Motome, J. Phys. Condens. Matter 28, 395601 (2016).
  • Watanabe and Yanase [2017] H. Watanabe and Y. Yanase, Phys. Rev. B 96, 064432 (2017).
  • Yanagi et al. [2018a] Y. Yanagi, S. Hayami, and H. Kusunose, Phys. Rev. B 97, 020404(R) (2018a).
  • Yanagi et al. [2018b] Y. Yanagi, S. Hayami, and H. Kusunose, Physica B Condens. Matter 536, 107 (2018b).
  • Shitade et al. [2018] A. Shitade, H. Watanabe, and Y. Yanase, Phys. Rev. B 98, 020407(R) (2018).
  • Hayami et al. [2018] S. Hayami, M. Yatsushiro, Y. Yanagi, and H. Kusunose, Phys. Rev. B 98, 165110 (2018).
  • Watanabe and Yanase [2018] H. Watanabe and Y. Yanase, Phys. Rev. B 98, 245129 (2018).
  • Gao and Xiao [2018] Y. Gao and D. Xiao, Phys. Rev. B 98, 060402(R) (2018).
  • Thöle and Spaldin [2018] F. Thöle and N. A. Spaldin, Philos. Trans. R. Soc. A 376, 20170450 (2018).
  • Yatsushiro and Hayami [2019] M. Yatsushiro and S. Hayami, J. Phys. Soc. Jpn. 88, 054708 (2019).
  • Thöle et al. [2020] F. Thöle, A. Keliri, and N. A. Spaldin, J. Appl. Phys. 127, 213905 (2020).
  • Yatsushiro and Hayami [2020] M. Yatsushiro and S. Hayami, J. Phys. Soc. Jpn. 89, 013703 (2020).
  • Hayami and Kusunose [2021b] S. Hayami and H. Kusunose, Phys. Rev. B 104, 045117 (2021b).
  • Yatsushiro et al. [2021] M. Yatsushiro, H. Kusunose, and S. Hayami, Phys. Rev. B 104, 054412 (2021).
  • Gorbatsevich and Kopaev [1994] A. A. Gorbatsevich and Y. V. Kopaev, Ferroelectrics 161, 321 (1994).
  • Ederer and Spaldin [2007] C. Ederer and N. A. Spaldin, Phys. Rev. B 76, 214404 (2007).
  • Spaldin et al. [2008] N. A. Spaldin, M. Fiebig, and M. Mostovoy, J. Phys. Condens. Matter 20, 434203 (2008).
  • Kopaev [2009] Y. V. Kopaev, Phys.-Uspekhi 52, 1111 (2009).
  • Gao et al. [2018] Y. Gao, D. Vanderbilt, and D. Xiao, Phys. Rev. B 97, 134423 (2018).
  • Wang et al. [2021] C. Wang, Y. Gao, and D. Xiao, Phys. Rev. Lett. 127, 277201 (2021).
  • Liu et al. [2021] H. Liu, J. Zhao, Y.-X. Huang, W. Wu, X.-L. Sheng, C. Xiao, and S. A. Yang, Phys. Rev. Lett. 127, 277202 (2021).
  • Zhang et al. [2020] C.-P. Zhang, X.-J. Gao, Y.-M. Xie, H. C. Po, and K. T. Law, arXiv:2012.15628  (2020).
  • Maruyama et al. [2012] D. Maruyama, M. Sigrist, and Y. Yanase, J. Phys. Soc. of Jpn. 81, 034702 (2012).
  • Hayami et al. [2015b] S. Hayami, H. Kusunose, and Y. Motome, J. Phys. Soc. Jpn. 84, 064717 (2015b).
  • Note [1] There is no contribution from the Berry curvature dipole [75] because of the 𝒫𝒯\mathcal{PT} symmetry, while the interband contribution in the 𝒯\mathcal{T}-breaking system [76] is neglected by considering the clean limit.
  • Ideue et al. [2017] T. Ideue, K. Hamamoto, S. Koshikawa, M. Ezawa, S. Shimizu, Y. Kaneko, Y. Tokura, N. Nagaosa, and Y. Iwasa, Nat. Phys. 13, 578 (2017).
  • Parker et al. [2019] D. E. Parker, T. Morimoto, J. Orenstein, and J. E. Moore, Phys. Rev. B 99, 045121 (2019).
  • Gao [2019] Y. Gao, Front. Phys. 14, 33404 (2019).
  • Dubovik and Cheshkov [1974] V. Dubovik and A. Cheshkov, Sov. J. Part. Nucl. 5, 318 (1974).
  • Dubovik et al. [1986] V. Dubovik, L. Tosunyan, and V. Tugushev, Zh. Eksp. Teor. Fiz. 90, 590 (1986).
  • Dubovik and Tugushev [1990] V. Dubovik and V. Tugushev, Phys. Rep. 187, 145 (1990).
  • Hayami and Kusunose [2018] S. Hayami and H. Kusunose, J. Phys. Soc. Jpn. 87, 033709 (2018).
  • Litvin [2008] D. B. Litvin, Acta Cryst. A 64, 316 (2008).
  • Schmid [2008] H. Schmid, J. Phys. Condens. Matter 20, 434201 (2008).
  • Hayami et al. [2020] S. Hayami, Y. Yanagi, and H. Kusunose, Phys. Rev. B 102, 144441 (2020).
  • Oiwa and Kusunose [2022] R. Oiwa and H. Kusunose, J. Phys. Soc. Jpn. 91, 014701 (2022).
  • Tursina et al. [2005] A. I. Tursina, S. N. Nesterenko, E. V. Murashova, I. V. Chernyshev, H. Noël, and Y. D. Seropegin, Acta Cryst. E 61, i12 (2005).
  • Kato et al. [2011] H. Kato, R. Kobayashi, T. Takesaka, T. Nishioka, M. Matsumura, K. Kaneko, and N. Metoki, J. Phys. Soc. Jpn. 80, 073701 (2011).
  • Suzuki et al. [2021] S. Suzuki, K. Takubo, K. Kuga, W. Higemoto, T. U. Ito, T. Tomita, Y. Shimura, Y. Matsumoto, C. Bareille, H. Wadati, S. Shin, and S. Nakatsuji, Phys. Rev. Research 3, 023140 (2021).
  • Barthem et al. [2013] V. Barthem, C. Colin, H. Mayaffre, M.-H. Julien, and D. Givord, Nat. Commun. 4, 2892 (2013).
  • Will et al. [1981] G. Will, W. Schäfer, F. Pfeiffer, F. Elf, and J. Etourneau, J. LESS-COMMN MET. 82, 349 (1981).
  • Will and Schäfer [1979] G. Will and W. Schäfer, J. LESS-COMMN MET 67, 31 (1979).
  • Wadley et al. [2015] P. Wadley, V. Hills, M. R. Shahedkhah, K. W. Edmonds, R. P. Campion, V. Novák, B. Ouladdiaf, D. Khalyavin, S. Langridge, V. Saidl, P. Nemec, A. W. Rushforth, B. L. Gallagher, S. S. Dhesi, F. Maccherozzi, J. Z̆elezný, and J. Jungwirth, Sci Rep 5, 17079 (2015).
  • Kimber et al. [2010] S. A. J. Kimber, A. H. Hill, Y.-Z. Zhang, H. O. Jeschke, R. Valentí, C. Ritter, I. Schellenberg, W. Hermes, R. Pöttgen, and D. N. Argyriou, Phys. Rev. B 82, 100412(R) (2010).
  • Emery et al. [2011] N. Emery, E. J. Wildman, J. M. S. Skakle, A. C. Mclaughlin, R. I. Smith, and A. N. Fitch, Phys. Rev. B 83, 144429 (2011).
  • Bao et al. [2011] W. Bao, Q.-Z. Huang, G.-F. Chen, D.-M. Wang, J.-B. He, and Y.-M. Qiu, Chin. Phys. Lett. 28, 086104 (2011).
  • Pomjakushin et al. [2011] V. Y. Pomjakushin, E. V. Pomjakushina, A. Krzton-Maziopa, K. Conder, and Z. Shermadini, J. Condens. Matter Phys. 23, 156003 (2011).
  • May et al. [2012] A. F. May, M. A. McGuire, H. Cao, I. Sergueev, C. Cantoni, B. C. Chakoumakos, D. S. Parker, and B. C. Sales, Phys. Rev. Lett. 109, 077003 (2012).
  • Suzuki et al. [2019] M.-T. Suzuki, T. Nomoto, R. Arita, Y. Yanagi, S. Hayami, and H. Kusunose, Phs. Rev. B 99, 174407 (2019).
  • Sodemann and Fu [2015] I. Sodemann and L. Fu, Phys. Rev. Lett. 115, 216806 (2015).
  • Gao et al. [2014] Y. Gao, S. A. Yang, and Q. Niu, Phys. Rev. Lett. 112, 166601 (2014).