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

Pseudo-spin rotation symmetry breaking by
Coulomb interaction terms in spin-orbit coupled systems

Shubhajyoti Mohapatra    Avinash Singh avinas@iitk.ac.in Department of Physics, Indian Institute of Technology, Kanpur - 208016, India
Abstract

By transforming from the pure-spin-orbital (t2gt_{\rm 2g}) basis to the spin-orbital entangled pseudo-spin-orbital basis, the pseudo-spin rotation symmetry of the different Coulomb interaction terms is investigated under SU(2) transformation in pseudo-spin space. While the Hubbard and density interaction terms are invariant, the Hund’s coupling and pair-hopping interaction terms explicitly break pseudo-spin rotation symmetry systematically. The form of the symmetry-breaking terms obtained from the transformation of the Coulomb interaction terms accounts for the easy xx-yy plane anisotropy and magnon gap for the out-of-plane mode, highlighting the importance of mixing with the nominally non-magnetic JJ=3/2 sector, and providing a physically transparent approach for investigating magnetic ordering and anisotropy effects in perovskite (Sr2IrO4\rm Sr_{2}IrO_{4}) and other d5d^{5} pseudo-spin compounds.

pacs:
75.30.Ds, 71.27.+a, 75.10.Lp, 71.10.Fd

I Introduction

Arising from a novel interplay between crystal field, spin-orbit coupling (SOC) and intermediate-strength Coulomb interactions, the emergent quantum states which essentially determine the electronic and magnetic properties of the iridium based transition-metal oxides involve correlated motion of electrons in spin-orbital entangled states.krempa_AR_2014 ; rau_AR_2016 ; bertinshaw_AR_2018 In the spin-orbit Mott insulator Sr2IrO4\rm Sr_{2}IrO_{4} with d5d^{5} configuration, electronic states near the Fermi energy have dominantly JJ=1/2 character, and important magnetic properties such as in-plane canted antiferromagnetic (AFM) order and magnon excitations have been extensively discussed in terms of the effectively single pseudo orbital (JJ=1/2) picture.kim_PRL_2008 ; kim_Science_2009 ; watanabe_PRL_2010 ; kim1_PRL_2012 Finite-interaction and finite-SOC effects are responsible for the strong zone-boundary magnon dispersion measured in resonant inelastic X-ray scattering (RIXS) studies, highlighting the observable effect of mixing between JJ=1/2 and 3/2 sectors.iridate_one

The Dzyaloshinskii-Moriya (DM) and pseudo-dipolar (PD) anisotropic interactions in the J=1/2J=1/2 sector, although weakly affected by the tetragonal splitting,kim_PRL_2012 are not the source of true anisotropy in Sr2IrO4\rm Sr_{2}IrO_{4}, as they yield pseudo-spin canting with no magnon gap due to compensation. True anisotropy has been ascribed to the Hund’s coupling term (JHJ_{\rm H}) using strong-coupling expansion (including virtual excitations to JJ=3/2 states) and numerical self-consistent calculation.jackeli_PRL_2009 ; igarashi_PRB_2013 ; perkins_PRB_2014 ; vale_PRB_2015 ; igarashi_JPSJ_2014 While Coulomb interactions were considered within the pure-spin-orbital basis (t2gt_{\rm 2g} orbitals, pure spins) in above approaches, their treatment within a pseudo-spin-orbital basis, and the role of weak magnetism in the other two pseudo orbitals (JJ=3/2 sector) on the JHJ_{\rm H}-induced easy-plane magnetic anisotropy and magnon gap (40\sim 40 meV), as measured in recent resonant inelastic X-ray scattering (RIXS) studies,kim_NATCOMM_2014 ; pincini_PRB_2017 ; porras_PRB_2019 have not been elucidated. Furthermore, the pseudo-spin-orbital based approach can allow for a unified study of both intra-orbital (magnon) and inter-orbital (spin-orbit exciton) excitations within a single formalism.

Magnetic anisotropy is generally associated with spin rotation symmetry breaking. Therefore, a general pseudo-spin rotation symmetry analysis of the different Coulomb interaction terms, treating all three pseudo orbitals on the same footing, can provide additional insight into the origin of true magnetic anisotropy in Sr2IrO4\rm Sr_{2}IrO_{4} arising from the interplay of spin-orbital entanglement and Coulomb interaction. Due to the spin-orbital entanglement, the same pseudo-spin rotation for all three pseudo orbitals (l=1,2,3l=1,2,3) corresponds to different pure-spin rotations for the three pure (t2gt_{\rm 2g}) orbitals. This follows directly from the relation ψμ=σμlcμlψl\psi_{\mu}=\sigma_{\mu}\sum_{l}c_{\mu l}\psi_{l} between the fermionic field operators in the pure-spin-orbital basis (μ=yz,xz,xy\mu=yz,xz,xy) and pseudo-spin-orbital basis (l=1,2,3l=1,2,3), where the Pauli matrices σμ=σx,σy,σz\sigma_{\mu}=\sigma_{x},\sigma_{y},\sigma_{z} corresponding to the three pure orbitals μ=yz,xz,xy\mu=yz,xz,xy. The same SU(2) transformation ψlψl=[U]ψl\psi_{l}\rightarrow\psi_{l}^{\prime}=[U]\psi_{l} for all three pseudo orbitals corresponds to different SU(2) transformations ψμψμ=[Uμ]ψμ\psi_{\mu}\rightarrow\psi_{\mu}^{\prime}=[U_{\mu}]\psi_{\mu} for the three pure orbitals due to the orbital dependence of [Uμ]=σμ[U]σμ[U_{\mu}]=\sigma_{\mu}[U]\sigma_{\mu}.

Therefore, the question of how the different Coulomb interaction terms transform under same pseudo-spin rotation for all three pseudo orbitals assumes importance. In other words, while all Coulomb interaction terms are invariant under same pure-spin rotation for all three pure orbitals, does this invariance hold under same pseudo-spin rotation in the pseudo-spin-orbital basis? Pseudo-spin rotation symmetry breaking by any Coulomb interaction term would imply true magnetic anisotropy and gapped magnon spectrum.

In this paper, we will show that while the Hubbard (U)(U) and density (U)(U^{\prime}) interaction terms do preserve pseudo-spin rotation symmetry, the Hund’s coupling (JHJ_{\rm H}) and pair-hopping (JPJ_{\rm P}) interaction terms explicitly break this symmetry systematically. This symmetry breaking results in (on-site) anisotropic interactions between moments in the JJ=1/2 and JJ=3/2 sectors, highlighting the importance of the weak magnetism in the nominally filled JJ=3/2 sector as well as the mixing between the two sectors. Magnetic anisotropy will not survive in the large SOC limit when the two sectors become effectively decoupled. The physically transparent approach for investigating magnetic ordering and anisotropy effects will be illustrated for the perovskite compound Sr2IrO4\rm Sr_{2}IrO_{4}, where the sign of JJ=3/2 sector moments directly yields easy xx-yy plane anisotropy.

The structure of this paper is as below. After introducing the transformation between the pure-spin-orbital and pseudo-spin-orbital bases in Sec. II, a symmetry analysis of the individual Coulomb interaction terms is carried out in Sec. III, where the Hund’s coupling and pair hopping interaction terms are shown to explicitly break pseudo-spin rotation symmetry. The transformed Coulomb interaction terms are obtained in Sec. IV, explicitly showing that the symmetry breaking terms account for the easy xx-yy plane anisotropy in Sr2IrO4\rm Sr_{2}IrO_{4}, as illustrated for the ground state energy in Sec. V, and for magnon excitations and anisotropy gap in Sec. VI. Comparison with an independent t2gt_{\rm 2g} orbital based approach is discussed in Sec. VII, which provides confirmation of the above symmetry analysis and also illustrates the significant simplification achieved in the pseudo-spin-orbital basis. After a critical comparison (Sec. VIII) of different but equivalent approaches for studying magnetic anisotropy in Sr2IrO4\rm Sr_{2}IrO_{4}, some conclusions are finally presented in Sec. IX.

II Pseudo-spin-orbital basis

Due to large crystal-field splitting (\sim3 eV) in the IrO6\rm IrO_{6} octahedra, low-energy physics in d5d^{5} iridates is effectively described by projecting out the empty eg{\rm e_{g}} levels which are well above the t2gt_{\rm 2g} levels. Spin-orbit coupling (SOC) further splits the t2gt_{\rm 2g} states into JJ=1/2 doublet (mJ=±1/2m_{J}=\pm 1/2) and JJ=3/2 quartet (mJ=±1/2,±3/2m_{J}=\pm 1/2,\pm 3/2), with an energy gap of 3λ/23\lambda/2 (Fig. 1). Nominally, four of the five electrons fill the JJ=3/2 states, leaving one electron for the JJ=1/2 sector, rendering it magnetically active in the ground state.

In the (yzyz\downarrow,xzxz\downarrow,xyxy||yz\uparrow||\;yz\uparrow,xzxz\uparrow,xyxy\downarrow) basis, the SOC Hamiltonian becomes block diagonal with two [3×33\times 3] blocks. There is no mixing between the (,,\downarrow,\downarrow,\uparrow) and (,,\uparrow,\uparrow,\downarrow) sectors, which can therefore be treated as pseudo spins \uparrow and \downarrow for the three doubly degenerate eigenstates (pseudo orbitals).watanabe_PRL_2010 Corresponding to the three Kramers pairs |J,mj|J,m_{j}\rangle above, the pseudo-spin-orbital basis states |l,τ|l,\tau\rangle for the three pseudo orbitals (l=1,2,3l=1,2,3), with pseudo spins (τ=,\tau=\uparrow,\downarrow) each, have the form:

|l=1,τ=σ\displaystyle\ket{l=1,\tau=\sigma} =\displaystyle= |12,±12=[|yz,σ¯±i|xz,σ¯±|xy,σ]/3\displaystyle\Ket{\frac{1}{2},\pm\frac{1}{2}}=\left[\Ket{yz,\bar{\sigma}}\pm i\Ket{xz,\bar{\sigma}}\pm\Ket{xy,\sigma}\right]/\sqrt{3}
|l=2,τ=σ\displaystyle\ket{l=2,\tau=\sigma} =\displaystyle= |32,±12=[|yz,σ¯±i|xz,σ¯2|xy,σ]/6\displaystyle\Ket{\frac{3}{2},\pm\frac{1}{2}}=\left[\Ket{yz,\bar{\sigma}}\pm i\Ket{xz,\bar{\sigma}}\mp 2\Ket{xy,\sigma}\right]/\sqrt{6}
|l=3,τ=σ¯\displaystyle\ket{l=3,\tau=\bar{\sigma}} =\displaystyle= |32,±32=[|yz,σ±i|xz,σ]/2\displaystyle\Ket{\frac{3}{2},\pm\frac{3}{2}}=\left[\Ket{yz,\sigma}\pm i\Ket{xz,\sigma}\right]/\sqrt{2} (1)

where |yz,σ\Ket{yz,\sigma}, |xz,σ\Ket{xz,\sigma}, |xy,σ\Ket{xy,\sigma} are the t2g basis states and the signs ±\pm correspond to spins σ=/\sigma=\uparrow/\downarrow. The coherent superposition of different-symmetry t2gt_{\rm 2g} orbitals, with opposite spin polarization between xzxz/yzyz and xyxy levels implies spin-orbital entanglement, and also imparts unique extended 3D shape to the pseudo-orbitals l=1,2,3l=1,2,3, as shown in Fig 1. The pseudo-spin dynamics in iridate heterostructures are gaining interest as their magnetic properties are much more sensitive to structural distortion compared to pure spin systems due to spin-orbital entanglement.meyers_SREP_2019 ; mohapatra_PRB_2019

Refer to caption
Figure 1: The pseudo-spin-orbital energy level scheme for the three Kramers pairs along with their orbital shapes. The colors represent the weights of pure spins (\uparrow-red, \downarrow-blue) in each pair.

Taking the conjugate to express the above basis transformation in terms of the l,τ|\langle l,\tau| and μ,σ|\langle\mu,\sigma| states, and rewriting in terms of the corresponding fermionic field operators:

ψl=(alal)andψμ=(aμaμ)\psi_{l}=\left(\begin{array}[]{c}a_{l\uparrow}\\ a_{l\downarrow}\end{array}\right)\;\;\;{\rm and}\;\;\;\psi_{\mu}=\left(\begin{array}[]{c}a_{\mu\uparrow}\\ a_{\mu\downarrow}\end{array}\right) (2)

involving the annihilation operators for the pseudo orbitals (l=1,2,3l=1,2,3, τ=,\tau=\uparrow,\downarrow) and the t2gt_{2g} orbitals (μ=yz,xz,xy\mu=yz,xz,xy, σ=,\sigma=\uparrow,\downarrow), we obtain (using Pauli matrices):

ψ1\displaystyle\psi_{1} =\displaystyle= 13[σxψyz+σyψxz+σzψxy]\displaystyle\frac{1}{\sqrt{3}}\left[\sigma_{x}\psi_{yz}+\sigma_{y}\psi_{xz}+\sigma_{z}\psi_{xy}\right]
ψ2\displaystyle\psi_{2} =\displaystyle= 16[σxψyz+σyψxz2σzψxy]\displaystyle\frac{1}{\sqrt{6}}\left[\sigma_{x}\psi_{yz}+\sigma_{y}\psi_{xz}-2\sigma_{z}\psi_{xy}\right]
ψ3\displaystyle\psi_{3} =\displaystyle= 12[σxψyzσyψxz].\displaystyle\frac{1}{\sqrt{2}}\left[\sigma_{x}\psi_{yz}-\sigma_{y}\psi_{xz}\right]. (3)

Inverting the above transformation yields the t2gt_{2g} basis states represented in terms of the pseudo-spin-orbital basis states:

ψyz\displaystyle\psi_{yz} =\displaystyle= σx[13ψ1+16ψ2+12ψ3]\displaystyle\sigma_{x}\left[\frac{1}{\sqrt{3}}\psi_{1}+\frac{1}{\sqrt{6}}\psi_{2}+\frac{1}{\sqrt{2}}\psi_{3}\right]
ψxz\displaystyle\psi_{xz} =\displaystyle= σy[13ψ1+16ψ212ψ3]\displaystyle\sigma_{y}\left[\frac{1}{\sqrt{3}}\psi_{1}+\frac{1}{\sqrt{6}}\psi_{2}-\frac{1}{\sqrt{2}}\psi_{3}\right]
ψxy\displaystyle\psi_{xy} =\displaystyle= σz[13ψ123ψ2].\displaystyle\sigma_{z}\left[\frac{1}{\sqrt{3}}\psi_{1}-\sqrt{\frac{2}{3}}\psi_{2}\right]. (4)

The above equations are convenient for transforming the hopping and Coulomb interaction terms to the pseudo-spin-orbital basis, and can be expressed in the compact form:

ψμ=σμl=1,2,3cμlψl\psi_{\mu}=\sigma_{\mu}\sum_{l=1,2,3}c_{\mu l}\,\psi_{l} (5)

where σμ=σx,σy,σz\sigma_{\mu}=\sigma_{x},\sigma_{y},\sigma_{z} for the three orbitals μ=yz,xz,xy\mu=yz,xz,xy, respectively, and the (real) transformation coefficients cμlc_{\mu l} are explicitly shown in Eq. (4).

III Pseudo-spin rotation symmetry breaking

We consider the on-site Coulomb interaction terms in the t2gt_{2g} basis (μ,ν=yz,xz,xy\mu,\nu=yz,xz,xy):

int\displaystyle\mathcal{H}_{\rm int} =\displaystyle= Ui,μniμniμ+Ui,μ<ν,σniμσniνσ¯+(UJH)i,μ<ν,σniμσniνσ\displaystyle U\sum_{i,\mu}{n_{i\mu\uparrow}n_{i\mu\downarrow}}+U^{\prime}\sum_{i,\mu<\nu,\sigma}{n_{i\mu\sigma}n_{i\nu\overline{\sigma}}}+(U^{\prime}-J_{\mathrm{H}})\sum_{i,\mu<\nu,\sigma}{n_{i\mu\sigma}n_{i\nu\sigma}} (6)
+\displaystyle+ JHi,μνaiμaiνaiμaiν+JPi,μνaiμaiμaiνaiν\displaystyle J_{\mathrm{H}}\sum_{i,\mu\neq\nu}{a_{i\mu\uparrow}^{\dagger}a_{i\nu\downarrow}^{\dagger}a_{i\mu\downarrow}a_{i\nu\uparrow}}+J_{\mathrm{P}}\sum_{i,\mu\neq\nu}{a_{i\mu\uparrow}^{\dagger}a_{i\mu\downarrow}^{\dagger}a_{i\nu\downarrow}a_{i\nu\uparrow}}
=\displaystyle= Ui,μniμniμ+U′′i,μ<νniμniνJHi,μν𝐒iμ𝐒iν+JPi,μνaiμaiμaiνaiν\displaystyle U\sum_{i,\mu}{n_{i\mu\uparrow}n_{i\mu\downarrow}}+U^{\prime\prime}\sum_{i,\mu<\nu}n_{i\mu}n_{i\nu}-J_{\mathrm{H}}\sum_{i,\mu\neq\nu}{\bf S}_{i\mu}\cdot{\bf S}_{i\nu}+J_{\mathrm{P}}\sum_{i,\mu\neq\nu}a_{i\mu\uparrow}^{\dagger}a_{i\mu\downarrow}^{\dagger}a_{i\nu\downarrow}a_{i\nu\uparrow}

including the intra-orbital (U)(U) and inter-orbital (U)(U^{\prime}) density interaction terms, the Hund’s coupling term (JH)(J_{\rm H}), and the pair hopping interaction term (JP(J_{\rm P}=JH)J_{\rm H}). Here aiμσa_{i\mu\sigma}^{\dagger} and aiμσa_{i\mu\sigma} are the creation and annihilation operators for site ii, orbital μ\mu, spin σ=,\sigma=\uparrow,\downarrow, the density operator niμσ=aiμσaiμσn_{i\mu\sigma}=a_{i\mu\sigma}^{\dagger}a_{i\mu\sigma}, the total density operator niμ=niμ+niμ=ψiμψiμn_{i\mu}=n_{i\mu\uparrow}+n_{i\mu\downarrow}=\psi_{i\mu}^{\dagger}\psi_{i\mu}, and U′′=UJH/2U^{\prime\prime}=U^{\prime}-J_{\rm H}/2. All interaction terms above are SU(2) invariant and thus possess pure-spin rotation symmetry. In the following, we consider the transformation of individual Coulomb interaction terms to the pseudo-spin-orbital basis using Eq. (5), and examine their SU(2) transformation behavior in pseudo-spin space.

III.1 Total density operator

For the total density operator (for site ii), we obtain using Eq. (5):

nμ=ψμψμ=l,mcμlcμmψlψmn_{\mu}=\psi_{\mu}^{\dagger}\psi_{\mu}=\sum_{l,m}c_{\mu l}c_{\mu m}\psi_{l}^{\dagger}\psi_{m} (7)

where we have used σμ=σμ\sigma_{\mu}^{\dagger}=\sigma_{\mu} and σμ2=𝟏\sigma_{\mu}^{2}={\bf 1}. Now, under the SU(2) transformation in pseudo-spin space (same for all three pseudo orbitals ll):

ψlψl=[U]ψl\psi_{l}\rightarrow\psi_{l}^{\prime}=[U]\psi_{l} (8)

the ψlψm\psi_{l}^{\dagger}\psi_{m} terms are invariant, and the total density operator is therefore SU(2) invariant. Therefore, the density interaction terms (UU^{\prime}) in Eq. (6) and the Hubbard interaction terms (UU) (using (n+n)2=n+n+2nn(n_{\uparrow}+n_{\downarrow})^{2}=n_{\uparrow}+n_{\downarrow}+2n_{\uparrow}n_{\downarrow}) are SU(2) invariant and possess spin rotation symmetry in pseudo-spin space.

III.2 Pair hopping interaction term

For the pair hopping interaction term (for site ii), we obtain:

JHaμaμaνaν\displaystyle J_{\rm H}\;a_{\mu\uparrow}^{\dagger}a_{\mu\downarrow}^{\dagger}a_{\nu\downarrow}a_{\nu\uparrow} =\displaystyle= JHaμaνaμaν=JH2(aμaν+aμaν)(aμaν+aμaν)\displaystyle J_{\rm H}\;a_{\mu\uparrow}^{\dagger}a_{\nu\uparrow}a_{\mu\downarrow}^{\dagger}a_{\nu\downarrow}=\frac{J_{\rm H}}{2}\Big{(}a_{\mu\uparrow}^{\dagger}a_{\nu\uparrow}+a_{\mu\downarrow}^{\dagger}a_{\nu\downarrow}\Big{)}\Big{(}a_{\mu\uparrow}^{\dagger}a_{\nu\uparrow}+a_{\mu\downarrow}^{\dagger}a_{\nu\downarrow}\Big{)} (9)
=\displaystyle= JH2(ψμψν)2,\displaystyle\frac{J_{\rm H}}{2}\left(\psi_{\mu}^{\dagger}\psi_{\nu}\right)^{2},

which is SU(2) invariant and possesses spin-rotation symmetry in pure-spin space. However, SU(2) invariance is lost in pseudo-spin space, as shown below. Again, using Eq. (5) to transform to the pseudo-spin-orbital basis, we obtain:

ψμψν=l,mcμlcνmψl(σxσy)ψm=l,mcμlcνmψl(iσz)ψm,\psi_{\mu}^{\dagger}\psi_{\nu}=\sum_{l,m}c_{\mu l}c_{\nu m}\;\psi_{l}^{\dagger}(\sigma_{x}\sigma_{y})\psi_{m}=\sum_{l,m}c_{\mu l}c_{\nu m}\;\psi_{l}^{\dagger}(i\sigma_{z})\psi_{m}, (10)

where we have taken μ=yz\mu=yz, ν=xz\nu=xz to illustrate the operations with Pauli matrices. Now, under the SU(2) transformation in pseudo-spin space (Eq. 8), the last term in Eq. (10):

ψl(iσz)ψmψl[U](iσz)[U]ψmψl(iσz)ψm,\psi_{l}^{\dagger}\left(i\sigma_{z}\right)\psi_{m}\rightarrow\psi_{l}^{\dagger}\left[U\right]^{\dagger}\left(i\sigma_{z}\right)\left[U\right]\psi_{m}\neq\psi_{l}^{\dagger}\left(i\sigma_{z}\right)\psi_{m}, (11)

showing that ψμψν\psi_{\mu}^{\dagger}\psi_{\nu} is not SU(2) invariant. The pair hopping interaction term therefore explicitly breaks pseudo-spin rotation symmetry.

III.3 Hund’s coupling term

For this term involving the pure-spin rotationally symmetric interaction 𝐒iμ𝐒iν{\bf S}_{i\mu}\cdot{\bf S}_{i\nu}, we consider the spin density operator (for site ii), and obtain using Eq. (5):

2𝐒μ=ψμ𝝈ψμ=lmcμlcμmψl(σμ𝝈σμ)ψm2{\bf S}_{\mu}=\psi_{\mu}^{\dagger}\makebox{\boldmath$\sigma$}\psi_{\mu}=\sum_{lm}c_{\mu l}c_{\mu m}\psi_{l}^{\dagger}(\sigma_{\mu}\makebox{\boldmath$\sigma$}\sigma_{\mu})\psi_{m} (12)

which transforms under the SU(2) transformation (Eq. 8) to:

2𝐒μ2𝐒μ=ψμ𝝈ψμ=lmcμlcμmψl([U]σμ𝝈σμ[U])ψm2{\bf S}_{\mu}\rightarrow 2{\bf S}_{\mu}^{\prime}=\psi_{\mu}^{\prime\dagger}\makebox{\boldmath$\sigma$}\psi_{\mu}^{\prime}=\sum_{lm}c_{\mu l}c_{\mu m}\psi_{l}^{\dagger}\left([U]^{\dagger}\sigma_{\mu}\makebox{\boldmath$\sigma$}\sigma_{\mu}[U]\right)\psi_{m} (13)

We now consider the term in brackets above for the case σμ=σx\sigma_{\mu}=\sigma_{x} (yzyz orbital) and represent it in terms of a rotation operation in spin space:

[U]σx(σxσyσz)σx[U]=[U](σxσyσz)[U]=(σxσyσz)=Rx(π)R(U)(σxσyσz)[U]^{\dagger}\sigma_{x}\left(\begin{array}[]{c}\sigma_{x}\\ \sigma_{y}\\ \sigma_{z}\end{array}\right)\sigma_{x}[U]=[U]^{\dagger}\left(\begin{array}[]{r}\sigma_{x}\\ -\sigma_{y}\\ -\sigma_{z}\end{array}\right)[U]=\left(\begin{array}[]{r}\sigma_{x}^{\prime}\\ -\sigma_{y}^{\prime}\\ -\sigma_{z}^{\prime}\end{array}\right)=R_{x}(\pi)R(U)\left(\begin{array}[]{c}\sigma_{x}\\ \sigma_{y}\\ \sigma_{z}\end{array}\right) (14)

where

R(U)(σxσyσz)=(σxσyσz)=[U](σxσyσz)[U]R(U)\left(\begin{array}[]{c}\sigma_{x}\\ \sigma_{y}\\ \sigma_{z}\end{array}\right)=\left(\begin{array}[]{r}\sigma_{x}^{\prime}\\ \sigma_{y}^{\prime}\\ \sigma_{z}^{\prime}\end{array}\right)=[U]^{\dagger}\left(\begin{array}[]{c}\sigma_{x}\\ \sigma_{y}\\ \sigma_{z}\end{array}\right)[U] (15)

shows the spin rotation by the rotation matrix R(U)R(U) corresponding to the SU(2) transformation [U][U], and

Rx(π)=(100010001)R_{x}(\pi)=\left(\begin{array}[]{ccc}1&0&0\\ 0&-1&0\\ 0&0&-1\end{array}\right) (16)

is the rotation matrix corresponding to π\pi rotation about the xx axis.

Similarly, for the 𝐒ν{\bf S}_{\nu} operator with ν=y\nu=y (xzxz orbital), we will obtain the product Ry(π)R(U)R_{y}(\pi)R(U). Therefore, the 𝐒μ𝐒ν{\bf S}_{\mu}\cdot{\bf S}_{\nu} interaction term will yield the matrix product:

[Rx(π)R(U)~]Ry(π)R(U)=R~(U)Rx(π)Ry(π)R(U)[\widetilde{R_{x}(\pi)R(U)}]R_{y}(\pi)R(U)=\widetilde{R}(U)R_{x}(\pi)R_{y}(\pi)R(U) (17)

where R~(U)\widetilde{R}(U) is the transpose of R(U)R(U) and we have used R~x(π)=Rx(π)\widetilde{R}_{x}(\pi)=R_{x}(\pi) for the diagonal matrix. Finally, since

R~(U)Rx(π)Ry(π)R(U)Rx(π)Ry(π)\widetilde{R}(U)R_{x}(\pi)R_{y}(\pi)R(U)\neq R_{x}(\pi)R_{y}(\pi) (18)

as Rx(π)Ry(π)𝟏R_{x}(\pi)R_{y}(\pi)\neq{\bf 1}, the Hund’s coupling term 𝐒μ𝐒ν{\bf S}_{\mu}\cdot{\bf S}_{\nu} is not pseudo-spin SU(2) invariant and therefore does not possess pseudo-spin rotation symmetry.

The above symmetry analysis shows that the Hund’s coupling and pair hopping interaction terms explicitly break pseudo-spin rotation symmetry. It is important to note here that the spin rotation symmetry is broken systematically. In other words, it is broken for each term in 𝐒μ𝐒ν{\bf S}_{\mu}\cdot{\bf S}_{\nu} involving the summations over (l,m)(l,m) and (l,m)(l^{\prime},m^{\prime}) for 𝐒μ{\bf S}_{\mu} and 𝐒ν{\bf S}_{\nu} in Eq. (12), and similarly for the pair hopping interaction term in Eq. (10).

IV Transformed Coulomb interaction terms

In the following, we illustrate the transformation of the different Coulomb interaction terms to the pseudo-spin-orbital basis, starting with ll=1 (JJ=1/2) sector of the intra-pseudo-orbital interaction terms. Similar transformation to the JJ basis has been discussed recently, focussing only on the density interaction terms.martins_JPCM_2017 Considering first the pair-hopping interaction term (Eq. 9), and retaining only the ll=mm=ll^{\prime}=mm^{\prime}=1 terms (indicated by \leadsto below), we obtain:

JH2μν(ψμψν)2=JH2μν[l,mcμlcνmψl(σμσν)ψm][l,mcμlcνmψl(σμσν)ψm]\displaystyle\frac{J_{\rm H}}{2}\sum_{\mu\neq\nu}(\psi_{\mu}^{\dagger}\psi_{\nu})^{2}=\frac{J_{\rm H}}{2}\sum_{\mu\neq\nu}\left[\sum_{l,m}c_{\mu l}c_{\nu m}\psi_{l}^{\dagger}(\sigma_{\mu}\sigma_{\nu})\psi_{m}\right]\left[\sum_{l^{\prime},m^{\prime}}c_{\mu l^{\prime}}c_{\nu m^{\prime}}\psi_{l^{\prime}}^{\dagger}(\sigma_{\mu}\sigma_{\nu})\psi_{m^{\prime}}\right] (19)
\displaystyle\leadsto JH[cyz,12cxz,12{ψ1(iσz)ψ1}2+cxz,12cxy,12{ψ1(iσx)ψ1}2+cxy,12cyz,12{ψ1(iσy)ψ1}2]\displaystyle J_{\rm H}\left[c_{yz,1}^{2}c_{xz,1}^{2}\left\{\psi_{1}^{\dagger}(i\sigma_{z})\psi_{1}\right\}^{2}+c_{xz,1}^{2}c_{xy,1}^{2}\left\{\psi_{1}^{\dagger}(i\sigma_{x})\psi_{1}\right\}^{2}+c_{xy,1}^{2}c_{yz,1}^{2}\left\{\psi_{1}^{\dagger}(i\sigma_{y})\psi_{1}\right\}^{2}\right]
=\displaystyle= 4JH[cyz,12cxz,12{S1z2}+cxz,12cxy,12{S1x2}+cxy,12cyz,12{S1y2}]=4JH9𝐒1𝐒1\displaystyle-4J_{\rm H}\left[c_{yz,1}^{2}c_{xz,1}^{2}\left\{S_{1z}^{2}\right\}+c_{xz,1}^{2}c_{xy,1}^{2}\left\{S_{1x}^{2}\right\}+c_{xy,1}^{2}c_{yz,1}^{2}\left\{S_{1y}^{2}\right\}\right]=-\frac{4J_{\rm H}}{9}{\bf S}_{1}\cdot{\bf S}_{1}

Similarly, for the Hund’s coupling term in Eq. (6), we obtain:

2JHμ<ν𝐒μ.𝐒ν\displaystyle-2J_{\rm H}\sum_{\mu<\nu}{\bf S}_{\mu}.{\bf S}_{\nu} =\displaystyle= JH2μ<ν[l,mcμlcμmψl(σμ𝝈σμ)ψm].[l,mcνlcνmψl(σν𝝈σν)ψm]\displaystyle-\frac{J_{\rm H}}{2}\sum_{\mu<\nu}\left[\sum_{l,m}c_{\mu l}c_{\mu m}\psi_{l}^{\dagger}(\sigma_{\mu}\makebox{\boldmath$\sigma$}\sigma_{\mu})\psi_{m}\right].\left[\sum_{l^{\prime},m^{\prime}}c_{\nu l^{\prime}}c_{\nu m^{\prime}}\psi_{l^{\prime}}^{\dagger}(\sigma_{\nu}\makebox{\boldmath$\sigma$}\sigma_{\nu})\psi_{m^{\prime}}\right] (23)
\displaystyle\leadsto JH2μ<ν[cμ12cν12{ψ1(Rμ𝝈)ψ1}.{ψ1(Rν𝝈)ψ1}]\displaystyle-\frac{J_{\rm H}}{2}\sum_{\mu<\nu}\left[c_{\mu 1}^{2}c_{\nu 1}^{2}\left\{\psi_{1}^{\dagger}(R_{\mu}\makebox{\boldmath$\sigma$})\psi_{1}\right\}.\left\{\psi_{1}^{\dagger}(R_{\nu}\makebox{\boldmath$\sigma$})\psi_{1}\right\}\right]
=\displaystyle= 2JHμ<ν[cμ12cν12(S1xS1yS1z)R~μRν(S1xS1yS1z)]=2JH9𝐒1𝐒1\displaystyle-2J_{\rm H}\sum_{\mu<\nu}\left[c_{\mu 1}^{2}c_{\nu 1}^{2}(S_{1x}\;\;S_{1y}\;\;S_{1z})\widetilde{R}_{\mu}R_{\nu}\left(\begin{array}[]{c}S_{1x}\\ S_{1y}\\ S_{1z}\end{array}\right)\right]=\frac{2J_{\rm H}}{9}{\bf S}_{1}\cdot{\bf S}_{1}

as the product cμ12cν12c_{\mu 1}^{2}c_{\nu 1}^{2} is identical for all three orbital pairs, and for the rotation matrices (Eq. 16), we have RxRy+RyRz+RzRx=𝟏R_{x}R_{y}+R_{y}R_{z}+R_{z}R_{x}=-{\bf 1}. Although the pair-hopping and Hund’s coupling interaction terms generally break pseudo-spin rotation symmetry, within the magnetically active ll=1 (JJ=1/2) sector, they individually yield classically isotropic terms of the form 𝐒1𝐒1{\bf S}_{1}\cdot{\bf S}_{1}. This follows from the special symmetry within this sector, as reflected by the identical coefficients cμlc_{\mu l} for all three orbitals μ=yz,xz,xy\mu=yz,xz,xy for l=1l=1 (Eq. 4).

Finally, from the remaining JHJ_{\rm H} term (in the U′′U^{\prime\prime} term of Eq. 6), again retaining only the l=m=l=m=1l=m=l^{\prime}=m^{\prime}=1 term, we obtain (using Eq. 7 for site ii):

JH2μ<νnμnν\displaystyle-\frac{J_{\rm H}}{2}\sum_{\mu<\nu}n_{\mu}n_{\nu} =\displaystyle= JH2μ<ν(l,mcμlcμmψlψm)(l,mcνlcνmψlψm)\displaystyle-\frac{J_{\rm H}}{2}\sum_{\mu<\nu}\left(\sum_{l,m}c_{\mu l}c_{\mu m}\psi_{l}^{\dagger}\psi_{m}\right)\left(\sum_{l^{\prime},m^{\prime}}c_{\nu l^{\prime}}c_{\nu m^{\prime}}\psi_{l^{\prime}}^{\dagger}\psi_{m^{\prime}}\right) (24)
\displaystyle\leadsto JH2μ<νcμ12cν12(ψ1ψ1)2=JH2μ<νcμ12cν12(n1+n1)(n1+n1)\displaystyle-\frac{J_{\rm H}}{2}\sum_{\mu<\nu}c_{\mu 1}^{2}c_{\nu 1}^{2}(\psi_{1}^{\dagger}\psi_{1})^{2}=-\frac{J_{\rm H}}{2}\sum_{\mu<\nu}c_{\mu 1}^{2}c_{\nu 1}^{2}(n_{1\uparrow}+n_{1\downarrow})(n_{1\uparrow}+n_{1\downarrow})
=\displaystyle= 2JH9𝐒1𝐒1JH3n1\displaystyle\frac{2J_{\rm H}}{9}{\bf S}_{1}\cdot{\bf S}_{1}-\frac{J_{\rm H}}{3}n_{1}

where we have used n1σ2=n1σn_{1\sigma}^{2}=n_{1\sigma} and 2n1n1=n1(4/3)𝐒1𝐒12n_{1\uparrow}n_{1\downarrow}=n_{1}-(4/3){\bf S}_{1}\cdot{\bf S}_{1}. Collecting all the 𝐒1𝐒1{\bf S}_{1}\cdot{\bf S}_{1} interaction terms resulting from the pair-hopping, Hund’s coupling, and density interaction terms corresponding to JHJ_{\rm H}, as obtained in Eqs. 19, 23, 24, yields an exact cancellation.

Similarly, considering the other two intra-pseudo-orbital cases (l=m=l=m=2,3l=m=l^{\prime}=m^{\prime}=2,3) in the J=3/2J=3/2 sector, we obtain from the three explicitly JHJ_{\rm H} interaction terms:

JH2μν(ψμψν)2\displaystyle\frac{J_{\rm H}}{2}\sum_{\mu\neq\nu}(\psi_{\mu}^{\dagger}\psi_{\nu})^{2} \displaystyle\leadsto 4JH9𝐒2𝐒2+JH3S2zS2zJHS3zS3z\displaystyle-\frac{4J_{\rm H}}{9}{\bf S}_{2}\cdot{\bf S}_{2}+\frac{J_{\rm H}}{3}S_{2z}S_{2z}-J_{\rm H}S_{3z}S_{3z}
2JHμ<ν𝐒μ𝐒ν\displaystyle-2J_{\rm H}\sum_{\mu<\nu}{\bf S}_{\mu}\cdot{\bf S}_{\nu} \displaystyle\leadsto JH18𝐒2𝐒2+JH3S2zS2z+JH2𝐒3𝐒3JHS3zS3z\displaystyle\frac{J_{\rm H}}{18}{\bf S}_{2}\cdot{\bf S}_{2}+\frac{J_{\rm H}}{3}S_{2z}S_{2z}+\frac{J_{\rm H}}{2}{\bf S}_{3}\cdot{\bf S}_{3}-J_{\rm H}S_{3z}S_{3z}
JH2μ<νnμnν\displaystyle-\frac{J_{\rm H}}{2}\sum_{\mu<\nu}n_{\mu}n_{\nu} \displaystyle\leadsto JH6𝐒2𝐒2+JH6𝐒3.𝐒3JH4(n2+n3)\displaystyle\frac{J_{\rm H}}{6}{\bf S}_{2}\cdot{\bf S}_{2}+\frac{J_{\rm H}}{6}{\bf S}_{3}.{\bf S}_{3}-\frac{J_{\rm H}}{4}(n_{2}+n_{3}) (25)

explicitly showing the symmetry breaking contributions (SlzSlz)(S_{lz}S_{lz}) from the pair-hopping and Hund’s coupling interaction terms, as expected from the SU(2) transformation analysis in Sec. III, in the absence of the special symmetry in the J=3/2J=3/2 sector. However, since SαSα=(1/4)[(n+n)2nn]=(1/3)𝐒𝐒S_{\alpha}S_{\alpha}=(1/4)[(n_{\uparrow}+n_{\downarrow})-2n_{\uparrow}n_{\downarrow}]=(1/3){\bf S}\cdot{\bf S} for all three components α=x,y,z\alpha=x,y,z of S=1/2S=1/2 quantum spin operators Sα=(1/2)ψσαψS_{\alpha}=(1/2)\psi^{\dagger}\sigma_{\alpha}\psi, there is no true magnetic anisotropy even if classically anisotropic terms such as SlzSlzS_{lz}S_{lz} are present, as in Eq. (25).

Substituting SlzSlz=(1/3)𝐒l𝐒lS_{lz}S_{lz}=(1/3){\bf S}_{l}\cdot{\bf S}_{l} in Eq. (25) yields an exact cancellation of the three explicitly JHJ_{\rm H} interaction terms in the J=3/2J=3/2 sector also. Therefore, the Hubbard like (or equivalently 𝐒l𝐒l{\bf S}_{l}\cdot{\bf S}_{l}) intra-pseudo-orbital interaction terms in all three sectors result only from the UU and UU^{\prime} terms in Eq. (6). Using similar analysis as above, and dropping one-particle density terms as in Eq. (24), one obtains (for site ii):

Uμnμnμ+Uμ<νnμnν\displaystyle U\sum_{\mu}{n_{\mu\uparrow}n_{\mu\downarrow}}+U^{\prime}\sum_{\mu<\nu}n_{\mu}n_{\nu} \displaystyle\leadsto (U+2U3)n1n1+(U+U2)(n2n2+n3n3)\displaystyle\left(\frac{U+2U^{\prime}}{3}\right)n_{1\uparrow}n_{1\downarrow}+\left(\frac{U+U^{\prime}}{2}\right)(n_{2\uparrow}n_{2\downarrow}+n_{3\uparrow}n_{3\downarrow}) (26)

From the above discussion it follows that true magnetic anisotropy results only from the inter-orbital anisotropic interaction terms such as SlzSlzS_{lz}S_{l^{\prime}z} with lll^{\prime}\neq l. Including inter-pseudo-orbital cases such as ll=mm, ll^{\prime}=mm^{\prime} and ll=mm^{\prime}, ll^{\prime}=mm in Eqs. (19) and (23), and keeping all interaction terms relevant for the present study (Hubbard, Hund’s coupling, and density), we obtain (for site ii):

int(i)\displaystyle\mathcal{H}_{\rm int}(i) =\displaystyle= (U+2U3)n1n1+(U+U2)[n2n2+n3n3]\displaystyle\left(\frac{U+2U^{\prime}}{3}\right)n_{1\uparrow}n_{1\downarrow}+\left(\frac{U+U^{\prime}}{2}\right)\left[n_{2\uparrow}n_{2\downarrow}+n_{3\uparrow}n_{3\downarrow}\right] (27)
\displaystyle- (UU3)2𝐒1𝐒2(UU2JH3)[(2𝐒1+𝐒2)𝐒3]+2JH[S1zS2zS1zS3z]\displaystyle\left(\frac{U-U^{\prime}}{3}\right)2{\bf S}_{1}\cdot{\bf S}_{2}-\left(\frac{U-U^{\prime}-2J_{\rm H}}{3}\right)\big{[}(2{\bf S}_{1}+{\bf S}_{2})\cdot{\bf S}_{3}\big{]}+2J_{\rm H}[S_{1}^{z}S_{2}^{z}-S_{1}^{z}S_{3}^{z}]
+\displaystyle+ (U+5U3JH6)[n1n2+n1n3]+(U+11U6JH12)n2n3\displaystyle\left(\frac{U+5U^{\prime}-3J_{\rm H}}{6}\right)[n_{1}n_{2}+n_{1}n_{3}]+\left(\frac{U+11U^{\prime}-6J_{\rm H}}{12}\right)n_{2}n_{3}

where 𝐒m=ψm𝝉2ψm{\bf S}_{m}=\psi_{m}^{\dagger}\frac{\mbox{\boldmath$\tau$}}{2}\psi_{m} and nm=ψm𝟏ψm=nm+nmn_{m}=\psi_{m}^{\dagger}{\bf 1}\psi_{m}=n_{m\uparrow}+n_{m\downarrow} are the spin and charge density operators. Using the spherical symmetry condition (UU^{\prime}=UU-2JH2J_{\mathrm{H}}), the transformed interaction Hamiltonian (27) simplifies to:

int(i)\displaystyle{\mathcal{H}}_{\rm int}(i) =\displaystyle= (U43JH)n1n1+(UJH)[n2n2+n3n3]\displaystyle\left(U-\frac{4}{3}J_{\rm H}\right)n_{1\uparrow}n_{1\downarrow}+\left(U-J_{\rm H}\right)\left[n_{2\uparrow}n_{2\downarrow}+n_{3\uparrow}n_{3\downarrow}\right] (28)
\displaystyle- 43JH𝐒1𝐒2+2JH[𝒮1z𝒮2z𝒮1z𝒮3z]\displaystyle\frac{4}{3}J_{\rm H}{\bf S}_{1}\cdot{\bf S}_{2}+2J_{\rm H}\left[\mathcal{S}_{1}^{z}\mathcal{S}_{2}^{z}-\mathcal{S}_{1}^{z}\mathcal{S}_{3}^{z}\right]
+\displaystyle+ (U136JH)[n1n2+n1n3]+(U73JH)n2n3.\displaystyle\left(U-\frac{13}{6}J_{\rm H}\right)\left[n_{1}n_{2}+n_{1}n_{3}\right]+\left(U-\frac{7}{3}J_{\rm H}\right)n_{2}n_{3}.

In the above equation, the Hubbard-like terms 𝒰mnmnm𝒰m𝐒m𝐒m{\cal U}_{m}n_{m\uparrow}n_{m\downarrow}\sim-{\cal U}_{m}{\bf S}_{m}\cdot{\bf S}_{m}, the Hund’s-coupling-like term 𝐒1𝐒2{\bf S}_{1}\cdot{\bf S}_{2}, and the density terms nlnmn_{l}n_{m}, are all invariant under pseudo-spin rotation and the corresponding SU(2) transformation ψmψm=[U]ψm\psi_{m}\rightarrow\psi_{m}^{\prime}=[U]\psi_{m}. Therefore, only the interaction terms S1zS2zS_{1}^{z}S_{2}^{z} and S1zS3zS_{1}^{z}S_{3}^{z} between moments in the JJ=1/2 and JJ=3/2 sectors are responsible for the magneto-crystalline anisotropy in Sr2IrO4\rm Sr_{2}IrO_{4}, highlighting the importance of the weak magnetism in the nominally filled JJ=3/2 sector due to the mixing between the two sectors. Magnetic anisotropy will not survive in the large SOC limit when the two sectors become effectively decoupled. As shown below, an easy xx-yy plane anisotropy is obtained from the dominant term (2JHS1zS3z-2J_{\rm H}S_{1}^{z}S_{3}^{z}) due to opposite sign of JJ=3/2 sector moment.

V Easy x-y plane anisotropy

We consider the various interaction terms in Eq. (28) in the Hartree-Fock (HF) approximation, focussing on the staggered field terms corresponding to (π,π\pi,\pi) ordered AF state on the square lattice. For general ordering direction with components 𝚫𝒍\Delta_{l}= (Δlx,Δly,Δlz)(\Delta_{l}^{x},\Delta_{l}^{y},\Delta_{l}^{z}), the staggered field term for sector ll in the pseudo-orbital basis is given by:

sf(l)=𝐤sψ𝐤ls(s𝝉𝚫𝒍)ψ𝐤ls=𝐤ssψ𝐤ls(ΔlzΔlxiΔlyΔlx+iΔlyΔlz)ψ𝐤ls\mathcal{H}_{\rm sf}(l)=\sum_{{\bf k}s}\psi_{{\bf k}ls}^{\dagger}\begin{pmatrix}-s\makebox{\boldmath$\tau\cdot\Delta_{l}$}\end{pmatrix}\psi_{{\bf k}ls}=\sum_{{\bf k}s}-s\psi_{{\bf k}ls}^{\dagger}\begin{pmatrix}\Delta_{l}^{z}&\Delta_{l}^{x}-i\Delta_{l}^{y}\\ \Delta_{l}^{x}+i\Delta_{l}^{y}&-\Delta_{l}^{z}\\ \end{pmatrix}\psi_{{\bf k}ls} (29)

where ψ𝐤ls=(a𝐤lsa𝐤ls)\psi_{{\bf k}ls}^{\dagger}=(a_{{\bf k}ls\uparrow}^{\dagger}\;\;a_{{\bf k}ls\downarrow}^{\dagger}), s=±1s=\pm 1 for the two sublattices A/B, and the staggered field components Δl=1,2,3α=x,y,z\Delta_{l=1,2,3}^{\alpha=x,y,z} are self-consistently determined from:

2Δ1α\displaystyle 2\Delta_{1}^{\alpha} =\displaystyle= 𝒰1m1α+2JH3m2α+JH(m3αm2α)δαz\displaystyle{\mathcal{U}}_{1}m_{1}^{\alpha}+\frac{2J_{\rm H}}{3}m_{2}^{\alpha}+J_{\rm H}(m_{3}^{\alpha}-m_{2}^{\alpha})\delta_{\alpha z}
2Δ2α\displaystyle 2\Delta_{2}^{\alpha} =\displaystyle= 𝒰2m2α+2JH3m1αJHm1αδαz\displaystyle{\mathcal{U}}_{2}m_{2}^{\alpha}+\frac{2J_{\rm H}}{3}m_{1}^{\alpha}-J_{\rm H}m_{1}^{\alpha}\delta_{\alpha z}
2Δ3α\displaystyle 2\Delta_{3}^{\alpha} =\displaystyle= 𝒰3m3α+JHm1αδαz\displaystyle{\mathcal{U}}_{3}m_{3}^{\alpha}+J_{\rm H}m_{1}^{\alpha}\delta_{\alpha z} (30)

in terms of the staggered pseudo-spin magnetization components ml=1,2,3α=x,y,zm_{l=1,2,3}^{\alpha=x,y,z}. In practice, it is easier to choose set of 𝚫l=1,2,3{\bf\Delta}_{l=1,2,3} and self-consistently determine the Hubbard-like interaction strengths 𝒰l=1,2,3{\mathcal{U}}_{l=1,2,3} such that 𝒰1=U43JH{\mathcal{U}}_{1}=U-\frac{4}{3}J_{\rm H} and 𝒰2=𝒰3=UJH{\mathcal{U}}_{2}={\mathcal{U}}_{3}=U-J_{\rm H} using Eq. (30). The interaction strengths are related by 𝒰l=2,3=𝒰l=1+JH/3{\mathcal{U}}_{l=2,3}={\mathcal{U}}_{l=1}+J_{\rm H}/3.

Transforming the staggered-field term back to the three-orbital basis (yzσ,xzσ,xyσ¯)(yz\sigma,xz\sigma,xy\bar{\sigma}), and including the crystal field, SOC and band terms,watanabe_PRL_2010 we have considered HF=SO+band+cf+sf\mathcal{H}_{\rm HF}=\mathcal{H}_{\rm SO}+\mathcal{H}_{\rm band+cf}+\mathcal{H}_{\rm sf} in our band structure and spin fluctuation analysis, where

band+cf\displaystyle\mathcal{H}_{\rm band+cf} =\displaystyle= 𝐤σsψ𝐤σs[(ϵ𝐤yz000ϵ𝐤xz000ϵ𝐤xy+ϵxy)δss+(ϵ𝐤yzϵ𝐤yz|xz0ϵ𝐤yz|xzϵ𝐤xz000ϵ𝐤xy)δs¯s]ψ𝐤σs\displaystyle\sum_{{\bf k}\sigma s}\psi_{{\bf k}\sigma s}^{\dagger}\left[\begin{pmatrix}{\epsilon_{\bf k}^{yz}}^{\prime}&0&0\\ 0&{\epsilon_{\bf k}^{xz}}^{\prime}&0\\ 0&0&{\epsilon_{\bf k}^{xy}}^{\prime}+\epsilon_{xy}\end{pmatrix}\delta_{ss^{\prime}}+\begin{pmatrix}\epsilon_{\bf k}^{yz}&\epsilon_{\bf k}^{yz|xz}&0\\ -\epsilon_{\bf k}^{yz|xz}&\epsilon_{\bf k}^{xz}&0\\ 0&0&\epsilon_{\bf k}^{xy}\end{pmatrix}\delta_{\bar{s}s^{\prime}}\right]\psi_{{\bf k}\sigma s^{\prime}} (31)

in the composite three-orbital, two-sublattice basis. The crystal field induced tetragonal splitting is included as the xyxy orbital energy offset ϵxy\epsilon_{xy} from the degenerate yz/xzyz/xz orbitals. The different hopping terms in Eq. (31) connecting the same (s=ss=s^{\prime}) and opposite (sss\neq s^{\prime}) sublattice(s) are given by:

ϵ𝐤xy\displaystyle\epsilon_{\bf k}^{xy} =\displaystyle= 2t1(coskx+cosky)\displaystyle-2t_{1}(\cos{k_{x}}+\cos{k_{y}})
ϵ𝐤xy\displaystyle{\epsilon_{\bf k}^{xy}}^{\prime} =\displaystyle= 4t2coskxcosky 2t3(cos2kx+cos2ky)\displaystyle-4t_{2}\cos{k_{x}}\cos{k_{y}}-\>2t_{3}(\cos{2{k_{x}}}+\cos{2{k_{y}}})
ϵ𝐤yz\displaystyle\epsilon_{\bf k}^{yz} =\displaystyle= 2t5coskx2t4cosky\displaystyle-2t_{5}\cos{k_{x}}-2t_{4}\cos{k_{y}}
ϵ𝐤xz\displaystyle\epsilon_{\bf k}^{xz} =\displaystyle= 2t4coskx2t5cosky\displaystyle-2t_{4}\cos{k_{x}}-2t_{5}\cos{k_{y}}
ϵ𝐤yz|xz\displaystyle\epsilon_{\bf k}^{yz|xz} =\displaystyle= 2tm(coskx+cosky).\displaystyle-2t_{m}(\cos{k_{x}}+\cos{k_{y}}). (32)

Here t1t_{1}, t2t_{2}, t3t_{3} are respectively the first, second, and third neighbor hopping terms for the xyxy orbital. For the yzyz (xzxz) orbital, t4t_{4} and t5t_{5} are the NN hopping terms in yy (x)(x) and xx (y)(y) directions, respectively. Mixing between xzxz and yzyz orbitals is represented by the NN hopping term tmt_{m}. We have taken values of the tight-binding parameters (t1t_{1}, t2t_{2}, t3t_{3}, t4t_{4}, t5t_{5}, tmt_{\rm m}, ϵxy\epsilon_{xy}, λ\lambda) = (1.0, 0.5, 0.25, 1.028, 0.167, 0.2, -0.7, 1.35) in units of t1t_{1}, where the energy scale t1t_{1} = 280 meV. Using above parameters, the calculated electronic band structure shows AFM insulating state and mixing between pseudo-orbital sectors.watanabe_PRL_2010 ; iridate_one

For the band dispersion terms in Eq. (31) and henceforth, we have used the crystal (global) coordinate axes referred to as x,y,zx,y,z for convenience. The spin coordinate axes are chosen to align with the crystal axes. If required, a site-dependent spin rotation allows one to transform back to the local spin coordinate axes common with the octahedral axes due to SOC.

Canted AFM state

The octahedral-rotation-induced orbital mixing hopping term (tmt_{\rm m}) between yzyz and xzxz orbitals generates PD (SizSjzS_{i}^{z}S_{j}^{z}) and DM [z^(𝐒i×𝐒j){\hat{z}}\cdot({\bf S}_{i}\times{\bf S}_{j})] anisotropic interactions in the strong coupling limit.iridate_one However, the AFM-state energy is invariant with respect to change of ordering direction from zz axis to xx-yy plane provided spins are canted at the optimal canting angle, thus preserving the gapless Goldstone mode. Fig. 2(a) shows the variation of AFM-state energy with canting angle (ϕ\phi) for ordering in the xx-yy plane. The energy minimum at the optimal canting angle is exactly degenerate with the energy for zz-direction ordering. This absence of true anisotropy in Sr2IrO4\rm Sr_{2}IrO_{4} due to octahedral rotation alone is consistent with the general gauge transformation analysis showing that the spin-dependent hopping terms arising from the orbital-mixing terms in the three-orbital model can be gauged away.

Refer to caption
Refer to caption
Figure 2: Variation of AFM state energy per state (in units of t1t_{1}) (a) for xx-yy plane ordering with canting angle ϕ\phi including finite yzxzyz-xz orbital mixing hopping showing degeneracy at the optimal canting angle with the zz-ordered AFM state, and (b) with the staggered field polar angle θ\theta showing easy xx-yy plane anisotropy for finite Hund’s coupling.

JHJ_{\rm H}-induced easy-plane anisotropy

The Hund’s-coupling-induced easy-plane magnetic anisotropy is explicitly shown in Fig.2(b) by the variation of AFM-state energy with polar angle θ\theta corresponding to staggered field orientation in the xx-zz plane, with Δ1z=(Δ+Δani)cosθ\Delta_{1}^{z}=(\Delta+\Delta_{\rm ani})\cos\theta and Δ1x=Δsinθ\Delta_{1}^{x}=\Delta\sin\theta. Here Δ\Delta represents the spin-rotationally-symmetric part (𝒰1m1+2JH3m2)/2({\mathcal{U}}_{1}m_{1}+\frac{2J_{\rm H}}{3}m_{2})/2 of the staggered field and Δani=JH(m3zm2z)/2\Delta_{\rm ani}=J_{\rm H}(m_{3}^{z}-m_{2}^{z})/2 is the symmetry-breaking term, as seen from Eq. (30). We have taken Δ\Delta=0.9, Δani\Delta_{\rm ani}=0.01-0.01, and the orbital mixing hopping term tmt_{\rm m} has been set to zero for simplicity. Converting from energy per state as shown in Fig.2 to simply energy per site (×\times10 occupied states for each 𝐤{\bf k}), yields the magnetic anisotropy energy EgHF(z)EgHF(x)0.012E_{\rm g}^{\rm HF}(z)-E_{\rm g}^{\rm HF}(x)\approx 0.012, which is comparable to the Δani\Delta_{\rm ani} magnitude. The simplified analysis presented in this section, with staggered field only for the ll=1 orbital, serves to explicitly illustrate the magnetic anisotropy features within our three-band-model calculation.

VI Magnon excitations and anisotropy gap

In view of the Hund’s-coupling-induced easy xx-yy plane anisotropy as discussed above, we consider the xx-ordered AFM state. The magnon propagator corresponding to transverse spin fluctuations should therefore yield one gapless mode (yy direction) and one gapped mode (zz direction). Accordingly, we consider the time-ordered magnon propagator:

χ(𝐪,ω)=𝑑tieiω(tt)ei𝐪.(𝐫i𝐫j)Ψ0|T[Simα(t)Sjnβ(t)]|Ψ0\chi({\bf q},\omega)=\int dt\sum_{i}e^{i\omega(t-t^{\prime})}e^{-i{\bf q}.({\bf r}_{i}-{\bf r}_{j})}\langle\Psi_{0}|T[S_{im}^{\alpha}(t)S_{jn}^{\beta}(t^{\prime})]|\Psi_{0}\rangle (33)

involving the transverse α,β=y,z\alpha,\beta=y,z components of the pseudo-spin operators SimαS_{im}^{\alpha} and SjnβS_{jn}^{\beta} for pseudo orbitals mm and nn at lattice sites ii and jj.

In the random phase approximation (RPA), the magnon propagator is obtained as:

[χ(𝐪,ω)]=[χ0(𝐪,ω)]12[𝒰][χ0(𝐪,ω)][\chi({\bf q},\omega)]=\frac{[\chi^{0}({\bf q},\omega)]}{1-2[\mathcal{U}][\chi^{0}({\bf q},\omega)]} (34)

where the bare particle-hole propagator:

[χ0(𝐪,ω)]abαβ=14𝐤[φ𝐤𝐪|τα|φ𝐤aφ𝐤|τβ|φ𝐤𝐪bE𝐤𝐪+E𝐤+ωiη+φ𝐤𝐪|τα|φ𝐤aφ𝐤|τβ|φ𝐤𝐪bE𝐤+E𝐤𝐪ωiη][\chi^{0}({\bf q},\omega)]_{ab}^{\alpha\beta}=\frac{1}{4}\sum_{{\bf k}}\left[\frac{\langle\varphi_{\bf k-q}|\tau^{\alpha}|\varphi_{\bf k}\rangle_{a}\langle\varphi_{\bf k}|\tau^{\beta}|\varphi_{\bf k-q}\rangle_{b}}{E^{+}_{\bf k-q}-E^{-}_{\bf k}+\omega-i\eta}+\frac{\langle\varphi_{\bf k-q}|\tau^{\alpha}|\varphi_{\bf k}\rangle_{a}\langle\varphi_{\bf k}|\tau^{\beta}|\varphi_{\bf k-q}\rangle_{b}}{E^{+}_{\bf k}-E^{-}_{\bf k-q}-\omega-i\eta}\right] (35)

was evaluated in the composite spin-orbital-sublattice basis (2 spin components α,β=y,z\alpha,\beta=y,z \otimes 3 pseudo orbitals m=1,2,3m=1,2,3 \otimes 2 sublattices s,s=s,s^{\prime}= A,B) by integrating out the fermions in the (π,π)(\pi,\pi) ordered state. Here E𝐤E_{\bf k} and φ𝐤\varphi_{\bf k} are the eigenvalues and eigenvectors of the Hamiltonian matrix in the pseudo-orbital basis, the indices a,b=1,6a,b=1,6 correspond to the orbital-sublattice subspace, and the superscript +()+(-) refers to particle (hole) energies above (below) the Fermi energy. The amplitudes φ𝐤τm\varphi^{m}_{{\bf k}\tau} were obtained by projecting the 𝐤{\bf k} states in the three-orbital basis on to the pseudo-orbital basis states |m,τ=,|m,\tau=\uparrow,\downarrow\rangle corresponding to the J=1/2J=1/2 and 3/23/2 sector states, as given below:

φ𝐤1=13(ϕ𝐤yziϕ𝐤xz+ϕ𝐤xy)\displaystyle\varphi_{{\bf k}\uparrow}^{1}=\frac{1}{\sqrt{3}}\left(\phi^{yz}_{{\bf k}\downarrow}-i\phi^{xz}_{{\bf k}\downarrow}+\phi^{xy}_{{\bf k}\uparrow}\right)\;\;\;\;\;\; φ𝐤1=13(ϕ𝐤yz+iϕ𝐤xzϕ𝐤xy)\displaystyle\varphi_{{\bf k}\downarrow}^{1}=\frac{1}{\sqrt{3}}\left(\phi^{yz}_{{\bf k}\uparrow}+i\phi^{xz}_{{\bf k}\uparrow}-\phi^{xy}_{{\bf k}\downarrow}\right)
φ𝐤2=16(ϕ𝐤yziϕ𝐤xz2ϕ𝐤xy)\displaystyle\varphi_{{\bf k}\uparrow}^{2}=\frac{1}{\sqrt{6}}\left(\phi^{yz}_{{\bf k}\downarrow}-i\phi^{xz}_{{\bf k}\downarrow}-2\phi^{xy}_{{\bf k}\uparrow}\right)\;\;\;\; φ𝐤2=16(ϕ𝐤yz+iϕ𝐤xz+2ϕ𝐤xy)\displaystyle\varphi_{{\bf k}\downarrow}^{2}=\frac{1}{\sqrt{6}}\left(\phi^{yz}_{{\bf k}\uparrow}+i\phi^{xz}_{{\bf k}\uparrow}+2\phi^{xy}_{{\bf k}\downarrow}\right)
φ𝐤3=12(ϕ𝐤yz+iϕ𝐤xz)\displaystyle\varphi_{{\bf k}\uparrow}^{3}=\frac{1}{\sqrt{2}}\left(\phi^{yz}_{{\bf k}\downarrow}+i\phi^{xz}_{{\bf k}\downarrow}\right)\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; φ𝐤3=12(ϕ𝐤yziϕ𝐤xz)\displaystyle\varphi_{{\bf k}\downarrow}^{3}=\frac{1}{\sqrt{2}}\left(\phi^{yz}_{{\bf k}\uparrow}-i\phi^{xz}_{{\bf k}\uparrow}\right) (36)

in terms of the amplitudes ϕ𝐤σμ\phi^{\mu}_{{\bf k}\sigma} in the three-orbital basis (μ=yz,xz,xy)(\mu=yz,xz,xy).

The rotationally invariant Hubbard- and Hund’s coupling-like terms having the form SimαSinβδαβS_{im}^{\alpha}S_{in}^{\beta}\delta_{\alpha\beta} are diagonal in spin components (α=β\alpha=\beta). The on-site Coulomb interaction terms are also diagonal in the sublattice basis (s=ss=s^{\prime}). The interaction matrix [𝒰][\mathcal{U}] in Eq. (34) is therefore obtained as:

[𝒰]=(𝒰123JH023JH𝒰2000𝒰3)δαβδss+(0JHJHJH00JH00)δαzδβzδss[\mathcal{U}]=\begin{pmatrix}\mathcal{U}_{1}&\frac{2}{3}J_{\rm H}&0\\ \frac{2}{3}J_{\rm H}&\mathcal{U}_{2}&0\\ 0&0&\mathcal{U}_{3}\end{pmatrix}\delta_{\alpha\beta}\delta_{ss^{\prime}}+\begin{pmatrix}0&-J_{\rm H}&J_{\rm H}\\ -J_{\rm H}&0&0\\ J_{\rm H}&0&0\end{pmatrix}\delta_{\alpha z}\delta_{\beta z}\delta_{ss^{\prime}} (37)

in the pseudo-orbital basis. While the first interaction term above preserves spin rotation symmetry, the second interaction term (corresponding to the SimzSinzS_{im}^{z}S_{in}^{z} terms in Eq. 28) breaks rotation symmetry and is responsible for easy xx-yy plane anisotropy. The spin wave energies are calculated from the pole condition 12ξ𝐪(ω)=01-2\xi_{\bf q}(\omega)=0 of Eq. 34, where ξ𝐪(ω)\xi_{\bf q}(\omega) are the eigenvalues of the [𝒰][χ0(𝐪,ω)][{\cal U}][\chi^{0}({\bf q},\omega)] matrix. The 12×1212\times 12 [χ0(𝐪,ω)][\chi^{0}({\bf q},\omega)] matrix was evaluated by performing the 𝐤\bf k sum over the 2D Brillouin zone divided into a 300 ×\times 300 mesh.

Refer to caption
Figure 3: The calculated magnon dispersion in the three-orbital model with staggered field in the xx direction. The easy xx-yy plane anisotropy arising from Hund’s coupling results in one gapless mode and one gapped mode corresponding to transverse fluctuations in the yy and zz directions, respectively.

The calculated magnon energies in the xx-ordered AFM state are shown in Fig. 3. Here we have taken staggered field values Δl=1,2,3x=(0.92,0.08,0.06)\Delta_{l=1,2,3}^{x}=(0.92,0.08,-0.06) in units of t1t_{1}, which ensures self-consistency for all three orbitals, with the given relations 𝒰2{\mathcal{U}}_{2}=𝒰3{\mathcal{U}}_{3}=𝒰1{\mathcal{U}}_{1}+JH/3J_{\rm H}/3. Using the calculated sublattice magnetization values ml=1,2,3xm_{l=1,2,3}^{x}=(0.65,0.005,-0.038), we obtain 𝒰l=1,2,3{\mathcal{U}}_{l=1,2,3}=(0.80,0.83,0.83) eV, which finally yields UU=𝒰1\mathcal{U}_{1}+43JH\frac{4}{3}J_{\rm H}=0.93 eV for JHJ_{\rm H}=0.1 eV.

The magnon dispersion clearly shows the Goldstone mode and the gapped mode, corresponding to transverse spin fluctuations in the yy and zz directions, respectively. The easy xx-yy plane anisotropy arising from Hund’s coupling results in energy gap 40\approx 40 meV for the out-of-plane (zz) mode. The two modes are degenerate at (π,0)(\pi,0) and (π/2,π/2)(\pi/2,\pi/2). The excitation energy at (π,0)(\pi,0) is approximately twice that at (π/2,π/2)(\pi/2,\pi/2), and the strong zone-boundary dispersion in this iridate compound was ascribed to finite-UU and finite-SOC effects.iridate_one The calculated magnon dispersion and energy gap are in very good agreement with RIXS measurements.kim1_PRL_2012 ; kim_NATCOMM_2014 ; pincini_PRB_2017 ; porras_PRB_2019

The electron fillings in the different pseudo orbitals are obtained as nl=1,2,3(1.064,1.99,1.946)n_{l=1,2,3}\approx(1.064,1.99,1.946). Finite mixing between the JJ=1/2 and 3/2 sectors is reflected in the small deviations from ideal fillings (1,2,2) and also in the very small magnetic moment values for l=2,3l=2,3 as given above, which play a crucial role in the expression of magnetic anisotropy and magnon gap in view of the anisotropic JHJ_{\rm H} interaction terms in Eq. (28). The values λ\lambda=0.38 eV, UU=0.93 eV, and JHJ_{\rm H}=0.1 eV taken above lie well within the estimated parameter range for Sr2IrO4\rm Sr_{2}IrO_{4}.igarashi_PRB_2014 ; zhou_PRX_2017

Refer to caption
Figure 4: Variation of magnon gap with magnetic moment |m3x||m_{3}^{x}| in the ll=3 orbital (blue curve). The magnitude of |m3x||m_{3}^{x}| decreases with SOC strength (red curve) due to suppression of mixing between JJ=1/2 and 3/2 sectors. Here (U,JH)(U,J_{\rm H})=(0.93 eV, 0.1 eV).

We have investigated the crucial role of the small JJ=3/2 sector magnetic moment on the magnon gap by studying the variation with SOC strength which effectively controls the mixing between JJ=1/2 and 3/2 sectors. Fig. 4 shows that the magnon gap sharply increases with the dominant magnetic moment |m3x||m_{3}^{x}|, highlighting the finite-SOC effect on the experimentally observed out-of-plane magnon gap in Sr2IrO4\rm Sr_{2}IrO_{4}. The opposite sign of the magnetic moment m3xm_{3}^{x} as compared to m1xm_{1}^{x} (due to spin-orbital entanglement) plays a vital role in the easy-plane anisotropy. It should be noted that the tetragonal splitting ϵxy\epsilon_{xy} weakly affects the magnon gap through the JJ=3/2 sector magnetic moments.

The magnon gap in Fig. 3 is related to the anisotropy term Δani\Delta_{\rm ani} (see discussion of Fig. 2), and we consider here an analytical expression relating the two which can be derived for a simple model. For the anisotropic Hubbard model,chatterji_PRB_2007 using the spin fluctuation analysis for the AFM state on the square lattice,singh_PRB_1990 we obtain:

ωgapωmax=(Uanit)(U4t)\frac{\omega_{\rm gap}}{\omega_{\rm max}}=\sqrt{\left(\frac{U_{\rm ani}}{t}\right)\left(\frac{U}{4t}\right)} (38)

in the strong coupling limit (Ut(U\gg t), where ωmax=2(4t2/U)\omega_{\rm max}=2(4t^{2}/U), and Uani2ΔaniU_{\rm ani}\approx 2\Delta_{\rm ani} is the anisotropy term included in the Hubbard model. With Δani/t=0.01\Delta_{\rm ani}/t=0.01 (as in Fig. 2(b) discussion) and U/4t=1U/4t=1 for simplicity, the ratio is obtained as 1/7\approx 1/7, which is in good agreement with Fig. 3. It should be noted that a magnon gap of 4\sim 4 meV (one-tenth of that in Fig. 3) would correspond to an anisotropy term Δani/t104\Delta_{\rm ani}/t\sim 10^{-4} (one-hundredth of above), which is about 0.03 meV using the hopping energy scale t1=280t_{1}=280 meV.

VII Comparison with pure orbital based approach

A fully self-consistent approach within the pure (t2gt_{\rm 2g}) orbital basis, wherein all orbital off-diagonal spin and charge condensates are included (along with the diagonal condensates) in the Coulomb interaction contributions, has been applied recently to the Ca2RuO4\rm Ca_{2}RuO_{4} compound with electron filling n=4n=4 to investigate the complex interplay due to intimately intertwined roles of SOC, Coulomb interactions, and structural distortions.ruthenate_one We discuss below the relevant results obtained by extending this approach to the Sr2IrO4\rm Sr_{2}IrO_{4} compound with n=5n=5.

Table 1: Self consistently determined magnetization and density values for the three pure (μ=yz,xz,xy\mu=yz,xz,xy) and three pseudo (l=1,2,3l=1,2,3) orbitals on the two sublattices (ss=A/B), showing the simplified structure in the pseudo orbital basis. The parameter set is same as in previous section.
μ\mu (s) mμxm_{\mu}^{x} mμym_{\mu}^{y} mμzm_{\mu}^{z} nμn_{\mu}
yzyz (A/B) ±\pm0.24 0.0 0.0 1.62
xzxz (A/B) \mp0.24 0.0 0.0 1.62
xyxy (A/B) \mp0.18 0.0 0.0 1.75
ll (s) mlxm_{l}^{x} mlym_{l}^{y} mlzm_{l}^{z} nln_{l}
11 (A/B) ±\pm0.69 0.0 0.0 1.04
22 (A/B) ±\pm0.006 0.0 0.0 1.99
33 (A/B) \mp0.023 0.0 0.0 1.97

Most importantly, we have explicitly confirmed the JHJ_{\rm H} induced easy-plane anisotropy. The anisotropy is completely absent when JHJ_{\rm H}=0 even for finite ϵxy\epsilon_{xy}, and the fully self consistent calculation (with octahedral rotation turned off for simplicity) yields degenerate solutions corresponding to arbitrary orientation of the (ll=1) pseudo-spin moment, confirming that finite ϵxy\epsilon_{xy} is not the source of any anisotropy. Moreover, we do not find any indication of easy-axis anisotropy within the basal plane even in the combined presence of Hund’s coupling (JHJ_{\rm H}), octahedral rotation (tmt_{m}), and tetragonal distortion (ϵxy\epsilon_{xy}). When a small octahedral tilting about crystal aa axis is included, we obtain easy-axis anisotropy along crystal bb axis and pseudo-spin canting in the zz direction, which can be understood in terms of the induced DM interaction resulting from the interplay between octahedral tilting and SOC.ruthenate_one

The simplicity in the pseudo orbital basis is illustrated by comparing the magnetization and density values for the pure and pseudo orbitals. The projected amplitudes (Eq. 36) were used to convert from pure to pseudo orbitals. The results are presented in Table I, showing the much simpler structure in the pseudo orbital basis, with only ll=1 sector magnetically active and nearly nominal density values. Furthermore, while the orbital off-diagonal condensates in the pure orbital basis are finite and yield the Coulomb SOC renormalization, all off-diagonal condensates in the pseudo orbital basis were found to be negligible, which is expected as the SOC term is diagonal. The Coulomb renormalized SOC gap in this basis is directly obtained from the orbital diagonal terms.mohapatra_JMMM_2020 The reduced mlxm_{l}^{x} value for ll=3 in Table I compared with the previous section (see Fig. 4) is consistent with the Coulomb enhanced SOC value (from 0.4\sim 0.4 to 0.6 eV) in the fully self consistent calculation.

VIII Discussion

Finally, a critical comparison with other approaches for studying magnetic anisotropy in Sr2IrO4\rm Sr_{2}IrO_{4} is presented below. In Ref. [10], the IrO6\rm IrO_{6} octahedral rotation induced PD (JzSizSjz)(J_{z}S_{i}^{z}S_{j}^{z}) and DM [𝐃(𝐒i×𝐒j)][{\bf D}\cdot({\bf S}_{i}\times{\bf S}_{j})] terms were considered as the dominant anisotropic interactions, which were shown to be gauged out by a site-dependent rotation of the spin operators (𝐒𝐒~{\bf S}\rightarrow{\bf{\tilde{S}}}), resulting in no true magnetic anisotropy in the absence of Hund’s coupling. After including the JHJ_{\rm H} induced corrections, the resulting anisotropic interaction terms were obtained as:

~ani=Γ1S~izS~jz±Γ2(S~ixS~jxS~iyS~jy)\mathcal{\tilde{H}_{\rm ani}}=-\Gamma_{1}{\tilde{S}}_{i}^{z}{\tilde{S}}_{j}^{z}\pm\Gamma_{2}({\tilde{S}}_{i}^{x}{\tilde{S}}_{j}^{x}-{\tilde{S}}_{i}^{y}{\tilde{S}}_{j}^{y}) (39)

in terms of the rotated spin operator 𝐒~{\bf{\tilde{S}}}, where the ±\pm sign corresponds to bond along the x(y)x(y) direction. Presence of tetragonal distortion was found to affect the coefficient Γ1\Gamma_{1}.

Together with the Heisenberg AFM interaction, the first term with Γ1>0\Gamma_{1}>0 leads to easy aba-b plane magnetic ordering and magnon gap for out-of-plane fluctuations, whereas the much smaller easy-axis anisotropy and magnon gap for in-plane fluctuations were ascribed to the second term (Γ2Γ1\Gamma_{2}\ll\Gamma_{1}). In Ref. [17], the large (40 meV) magnon gap measured for out-of-plane fluctuations was explained in terms of the above JHJ_{\rm H} induced anisotropy term Γ1\Gamma_{1}, although JHJ_{\rm H} was not explicitly mentioned in the very brief discussion, as the main focus was on resolving the small (\sim 3 meV) magnon gap corresponding to the easy-axis anisotropy and basal-plane fluctuations via high-resolution RIXS and inelastic neutron scattering (INS). The proposed mechanism for the easy-axis anisotropy involve interlayer coupling and orthorhombic distortion which are beyond the scope of this work.

For the easy-plane anisotropy, our conclusions are in agreement with the above analysis, including no true anisotropy in the absence of JHJ_{\rm H}, and the tetragonal distortion term only affecting the magnitude and not being the source of anisotropy. In terms of symmetry breaking, the above JHJ_{\rm H} induced easy-plane anisotropy term Γ1\Gamma_{1} obtained via strong coupling expansion is equivalent to our local anisotropic interaction term (dominantly 2JHS1zS3z-2J_{\rm H}S_{1z}S_{3z}) derived using the transformation. The major difference is that while in Ref. [17] the Γ1\Gamma_{1} value was treated as a fitting parameter in the 40 meV magnon gap analysis, in our work the explicit form of the anisotropic interaction term [2JH(S1zS2zS1zS3z)2J_{\rm H}(S_{1z}S_{2z}-S_{1z}S_{3z})] has been derived, and the magnon gap is determined using the same microscopic Hamiltonian parameters which account for the high energy magnon feature (strong zone boundary dispersion) within a unified scheme. Furthermore, all physical terms are treated on the same footing, which is especially important for the weakly correlated 5d5d systems.iridate_one In general, the nature of magnetic anisotropy depends, through the sign and magnitude of the JJ=3/2 sector moments S2zS_{2z} and S3zS_{3z}, on the lattice, band structure, and magnetic ordering, which is of particular interest for the honeycomb lattice compounds Na2IrO3\rm Na_{2}IrO_{3} and α\alpha-RuCl3\rm RuCl_{3}.

IX Conclusion

While all Coulomb interaction terms are invariant under same pure-spin rotation for all three pure (t2gt_{\rm 2g}) orbitals, the Hund’s coupling and pair hopping interaction terms were shown to explicitly break pseudo-spin rotation symmetry systematically due to the spin-orbital entanglement. Transformation of the various Coulomb interaction terms to the pseudo-spin-orbital basis formed by the JJ=1/2 and 3/2 states therefore provides a physically transparent approach for investigating magnetic ordering and anisotropy effects in the perovskite (Sr2IrO4\rm Sr_{2}IrO_{4}) and other d5d^{5} pseudo-spin compounds. Explicitly pseudo-spin symmetry-breaking terms were obtained (dominantly 2JHS1zS3z-2J_{\rm H}S_{1}^{z}S_{3}^{z}), resulting in easy xx-yy plane anisotropy and magnon gap for the out-of-plane mode, highlighting the importance of mixing with the nominally non-magnetic JJ=3/2 sector in determining the magnetic properties of Sr2IrO4\rm Sr_{2}IrO_{4}.

References

  • (1) W. Witczak-Krempa, G. Chen, Y. B. Kim, and L. Balents, Correlated Quantum Phenomena in the Strong Spin-Orbit Regime, Annu. Rev. Condens. Matter Phys. 5, 57 (2014).
  • (2) J. G. Rau, E. Kin-Ho Lee, and H.-Y. Kee, Spin-Orbit Physics Giving Rise to Novel Phases in Correlated Systems: Iridates and Related Materials, Annu. Rev. Condens. Matter Phys. 7, 195 (2016).
  • (3) J. Bertinshaw, Y. K. Kim, G. Khaliullin, and B. J. Kim, Square Lattice Iridates, Annu. Rev. Condens. Matter Phys. 10, 315 (2019).
  • (4) B. J. Kim, H. Jin, S. J. Moon, J.-Y. Kim, B.-G. Park, C. S. Leem, J. Yu, T. W. Noh, C. Kim, S.-J. Oh, J.-H. Park, V. Durairaj, G. Cao, and E. Rotenberg, Novel Jeff=1/2{J}_{\mathrm{eff}}=1/2 Mott State Induced by Relativistic Spin-Orbit Coupling in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, Phys. Rev. Lett. 101, 076402 (2008).
  • (5) B. Kim, H. Ohsumi, T. Komesu, S. Sakai, T. Morita, H. Takagi, and T. Arima, Phase-Sensitive Observation of a Spin-Orbital Mott State in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, Science 323, 1329 (2009).
  • (6) H. Watanabe, T. Shirakawa, and S. Yunoki, Microscopic Study of a Spin-Orbit-Induced Mott Insulator in Ir Oxides, Phys. Rev. Lett. 105, 216410 (2010).
  • (7) J. Kim, D. Casa, M. H. Upton, T. Gog, Y.-J. Kim, J. F. Mitchell, M. van Veenendaal, M. Daghofer, J. van den Brink, G. Khaliullin, and B. J. Kim, Magnetic Excitation Spectra of Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4} Probed by Resonant Inelastic X-Ray Scattering: Establishing Links to Cuprate Superconductors, Phys. Rev. Lett. 108, 177003 (2012).
  • (8) S. Mohapatra, J. van den Brink, and A. Singh, Magnetic Excitations in a Three-Orbital Model for the Strongly Spin-Orbit Coupled Iridates: Effect of Mixing Between the J=12J=\frac{1}{2} and 32\frac{3}{2} Sectors, Phys. Rev. B 95, 094435 (2017).
  • (9) B. H. Kim, G. Khaliullin, and B. I. Min, Magnetic Couplings, Optical Spectra, and Spin-Orbit Exciton in 5d5d Electron Mott Insulator Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, Phys. Rev. Lett. 109, 167205 (2012).
  • (10) G. Jackeli and G. Khaliullin, Mott Insulators in the Strong Spin-Orbit Coupling Limit: From Heisenberg to a Quantum Compass and Kitaev Models, Phys. Rev. Lett. 102, 017205 (2009).
  • (11) J-i Igarashi and T. Nagao, Strong Coupling Theory of Spin and Orbital Excitations in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, Phys. Rev. B 88, 104406 (2013).
  • (12) N. B. Perkins, Y. Sizyuk, and P. Wölfle, Interplay of Many-Body and Single-Particle Interactions in Iridates and Rhodates, Phys. Rev. B 89, 035143 (2014).
  • (13) J. G. Vale, S. Boseggia, H. C. Walker, R. Springell, Z. Feng, E. C. Hunter, R. S. Perry, D. Prabhakaran, A. T. Boothroyd, S. P. Collins, H. M. Rønnow, and D. F. McMorrow, Importance of XYXY Anisotropy in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4} Revealed by Magnetic Critical Scattering Experiments, Phys. Rev. B 92, 020406(R) (2015).
  • (14) J-i Igarashi and T. Nagao, Interplay between Hund’s Coupling and Spin-Orbit Interaction on Elementary Excitations in Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, J. Phys. Soc. Jpn. 83, 053709 (2014).
  • (15) J. Kim, M. Daghofer, A. H. Said, T. Gog, J. van den Brink, G. Khaliullin, and B. J. Kim, Excitonic Quasiparticles in a Spin–Orbit Mott Insulator, Nat. Commun. 5, 4453 (2014).
  • (16) D. Pincini, J. G. Vale, C. Donnerer, A. de la Torre, E. C. Hunter, R. Perry, M. Moretti Sala, F. Baumberger, and D. F. McMorrow, Anisotropic Exchange and Spin-Wave Damping in Pure and Electron-Doped Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, Phys. Rev. B 96, 075162 (2017).
  • (17) J. Porras, J. Bertinshaw, H. Liu, G. Khaliullin, N. H. Sung, J.-W. Kim, S. Francoual, P. Steffens, G. Deng, M. Moretti Sala, A. Effimenko, A. Said, D. Casa, X. Huang, T. Gog, J. Kim, B. Keimer, and B. J. Kim, Pseudospin-Lattice Coupling in the Spin-Orbit Mott Insulator Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, Phys. Rev. B 99, 085125 (2019).
  • (18) D. Meyers, Y. Cao, G. Fabbris, N. J. Robinson, L. Hao, C. Frederick, N. Traynor, J. Yang, J. Lin, M. H. Upton, D. Casa, J.-W. Kim, T. Gog, E. Karapetrova, Y. Choi, D. Haskel, P. J. Ryan, L. Horak, X. Liu, J. Liu, and M. P. M. Dean, Magnetism in Iridate Heterostructures Leveraged by Structural Distortions, Sci. Rep. 9, 4263 (2019).
  • (19) S. Mohapatra, S. Aditya, R. Mukherjee, and A. Singh, Octahedral Tilting Induced Isospin Reorientation Transition in Iridate Heterostructures, Phys. Rev. B 100, 140409(R) (2019).
  • (20) S. Boseggia, Ph.D. Thesis, Magnetic order and excitations in perovskite iridates studied with resonant X-ray scattering techniques (2014).
  • (21) C. Martins, M. Aichhorn, and S. Biermann, Coulomb Correlations in 4d4d and 5d5d Oxides from First Principles— or How Spin– Orbit Materials Choose Their Effective Orbital Degeneracies, J. Phys.: Condens. Matter 29, 263001 (2017).
  • (22) J-i Igarashi and T. Nagao, Analysis of Resonant Inelastic X-ray Scattering from Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4} in an Itinerant-Electron Approach, Phys. Rev. B 90, 064402 (2014).
  • (23) S. Zhou, K. Jiang, H. Chen, and Z. Wang, Correlation Effects and Hidden Spin-Orbit Entangled Electronic Order in Parent and Electron-Doped Iridates Sr2IrO4{\mathrm{Sr}}_{2}{\mathrm{IrO}}_{4}, Phys. Rev. X 7, 041018 (2017).
  • (24) S. Mohapatra and A. Singh, Magnetic Reorientation Transition in a Three Orbital Model for Ca2RuO4\rm Ca_{2}RuO_{4} — Interplay of Spin-Orbit Coupling, Tetragonal Distortion, and Coulomb Interactions, J. Phys.: Condens. Matter 32, 485805 (2020).
  • (25) S. Mohapatra and A. Singh, Correlated Motion of Particle-Hole Excitations Across the Renormalized Spin-Orbit Gap in Sr2IrO4\rm Sr_{2}IrO_{4}, J. Magn. Magn. Mater 512, 166997 (2020).
  • (26) T. Chatterji, S. Ghosh, A. Singh, L. P. Regnault, M. Rheinstadter, G.J. Mcintyre, and T. Lonkai, Spin Dynamics of YMnO3\rm YMnO_{3} studied via Inelastic Neutron Scattering and the Anisotropic Hubbard Model, Phys. Rev. B 76, 144406 (2007).
  • (27) A. Singh and Z.Teššanović, Spin Fluctuations in an Itinerant Antiferromagnet, Phys. Rev. B 41, 11457 (1990).