14
Magnetic Anisotropy in Spin-3/2 with Heavy Ligand in Honeycomb Mott Insulators: Application to CrI 3 P. Peter Stavropoulos, 1 Xiaoyu Liu, 1 and Hae-Young Kee 1, 2, * 1 Department of Physics and Center for Quantum Materials, University of Toronto, 60 St. George St., Toronto, Ontario, M5S 1A7, Canada 2 Canadian Institute for Advanced Research, Toronto, Ontario, M5G 1Z8, Canada Ferromagnetism in the two-dimensional CrI3 has generated a lot of excitement, and it was re- cently proposed that the spin-orbit coupling (SOC) in Iodine may generate bond-dependent spin interactions leading to magnetic anisotropy. Here we derive a microscopic spin model of S=3/2 on transition metals surrounded by heavy ligands in honeycomb Mott insulators using a strong-coupling perturbation theory. For ideal octahedra we find Heisenberg and Kitaev interactions, which favor the magnetic moment along the cubic axis via quantum fluctuations. When a slight trigonal distortion of the octahedra is present together with the SOC, three additional interactions arise, comprised of the off-diagonal symmetric Γ and Γ 0 , and single-ion anisotropy. The resulting magnetic anisotropy pins the moment perpendicular to the honeycomb plane as observed in a single-layer of CrI3, suggesting the significance of SOC and trigonal distortion in understanding magnetism of two dimensional Mott insulators. Comparison to the spin-orbit coupled J eff = 1/2 and S=1 models is also presented. I. INTRODUCTION Transition metal trihalides (TMT) are layered materi- als composed of transition metals (M) and halides (X) of the group 9 in a 1:3 ratio. They have a honeycomb layered structure, and depending on the filling of the d- orbitals in the transition metals, some are semiconductors and some are metals.[1] Among them, RuCl 3 , VI 3 and CrI 3 are Mott insulators. Magnetic orderings in these systems further establish the importance of electronic correlations and call for a microscopic understanding of spin models. For example, based on a strong-coupling perturbation the- ory of the generic spin model[24], it was shown that α- RuCl 3 described by the effective spin J eff =1/2 has dom- inant bond-dependent Kitaev and off-diagonal symmetric Γ interactions.[5, 6] RuCl 3 has become an emergent can- didate of spin-1/2 Kitaev spin liquid[7]. Intense research activities on various properties of RuCl 3 have been carried out[6, 814] and recently a magnetic-field induced spin liq- uid was suggested.[1521] In parallel theoretical interest in the ground state of higher-spin Kitaev models was initiated by classical model studies.[22, 23] The classical Kitaev model has a macro- scopic degeneracy named a classical spin liquid[22], but the higher-spin quantum Kitaev model is not exactly solvable, and the ground state is currently unknown. Various numer- ical studies such as exact diagonalization on S=1 suggested that the ground state is possibly a spin liquid with gapless excitations.[24] These studies were mainly of theoretical in- terest, until a microscopic derivation of the S=1 Kitaev- Heisenberg model in multi-orbital systems was found.[25] Heavy ligand spin-orbit coupling (SOC) and strong Hund’s coupling in e g orbitals is a way to generate S=1 bond- dependent Kitaev interaction. The magnetic field effects on the S=1 Kitaev model have also been investigated.[2628] The bond-dependent interactions have recently been adopted into TMT systems, because the nearest neighbor * [email protected] (n.n.) Heisenberg J , Kitaev K and Γ interactions are al- lowed based on the symmetry of the lattice.[4, 29, 30] In particular, ferromagnetism in a single-layer CrI 3 has gener- ated excitement in recent years. [3145] CrI 3 is a ferromag- netic (FM) insulator with T c 61K for bulk samples.[4648] Single-layer CrI 3 was successfully synthesized, which showed an FM ordering with T c 45K. [49] The two- dimensional FM Heisenberg model is insufficient to explain finite T c , i.e., the Mermin-Wagner theorem[50], and several theoretical models were proposed to explain the magnetic anisotropy. They include the XXZ model[51, 52], single- ion anisotropy and Kitaev[53], and large Kitaev and small symmetric off-diagonal Γ interactions[54]. While the Heisenberg, Kitaev, and Γ interactions are al- lowed by the symmetry, and found to be significant in the earlier derivations for lower-spins[3, 4, 25], their strengths may not be significant in S=3/2 systems. Thus, a micro- scopic derivation of S=3/2 model is necessary to find the sources of the magnetic anisotropy. Here we derive a n.n. spin model for S=3/2 with three electrons in t 2g orbitals of transition metal sites and strong SOC in p-orbitals of ligands. We take into account strong electron-electron in- teractions in multi-orbital systems including Hund’s cou- pling, and effects of trigonal distortions present in R ¯ 3 rhombohedral lattice. Contributions from e g orbitals are important as shown below. The minimal n.n. model in- cludes J , K, Γ, another symmetric off-diagonal Γ 0 [55], and single-ion anisotropy A c along the ˆ c-axis, denoted as the J - K - Γ - Γ 0 - A c model. The rest of the paper is organized as follows. In Sec. II, the on-site Kanamori interaction and tight binding Hamil- tonian are presented. In Sec. III, we derive the n.n. spin model consisting of Kitaev and Heisenberg interactions for ideal octahedra environment using standard perturbation theory. In Sec. IV, we study the spin model with trigonal distortion present in R ¯ 3 rhombohedral lattice. This in- cludes the distortion-induced hopping matrix elements and three additional spin interactions generated via combined effects of SOC and distortion. In Sec. V, we apply the the- ory to CrI 3 and present the exchange interaction strengths using tight binding parameter sets obtained by density functional theory. The effects of the resulting magnetic arXiv:2009.04475v2 [cond-mat.str-el] 16 Mar 2021

Application to CrI - arXiv · 2020. 9. 11. · Magnetic Anisotropy in Spin-3/2 with Heavy Ligand in Honeycomb Mott Insulators: Application to CrI 3 P. Peter Stavropoulos1 and Hae-Young

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • Magnetic Anisotropy in Spin-3/2 with Heavy Ligand in Honeycomb Mott Insulators:Application to CrI3

    P. Peter Stavropoulos,1 Xiaoyu Liu,1 and Hae-Young Kee1, 2, ∗

    1Department of Physics and Center for Quantum Materials,University of Toronto, 60 St. George St., Toronto, Ontario, M5S 1A7, Canada

    2Canadian Institute for Advanced Research, Toronto, Ontario, M5G 1Z8, Canada

    Ferromagnetism in the two-dimensional CrI3 has generated a lot of excitement, and it was re-cently proposed that the spin-orbit coupling (SOC) in Iodine may generate bond-dependent spininteractions leading to magnetic anisotropy. Here we derive a microscopic spin model of S=3/2 ontransition metals surrounded by heavy ligands in honeycomb Mott insulators using a strong-couplingperturbation theory. For ideal octahedra we find Heisenberg and Kitaev interactions, which favor themagnetic moment along the cubic axis via quantum fluctuations. When a slight trigonal distortion ofthe octahedra is present together with the SOC, three additional interactions arise, comprised of theoff-diagonal symmetric Γ and Γ′, and single-ion anisotropy. The resulting magnetic anisotropy pinsthe moment perpendicular to the honeycomb plane as observed in a single-layer of CrI3, suggestingthe significance of SOC and trigonal distortion in understanding magnetism of two dimensional Mottinsulators. Comparison to the spin-orbit coupled Jeff= 1/2 and S=1 models is also presented.

    I. INTRODUCTION

    Transition metal trihalides (TMT) are layered materi-als composed of transition metals (M) and halides (X)of the group 9 in a 1:3 ratio. They have a honeycomblayered structure, and depending on the filling of the d-orbitals in the transition metals, some are semiconductorsand some are metals.[1] Among them, RuCl3, VI3 and CrI3are Mott insulators. Magnetic orderings in these systemsfurther establish the importance of electronic correlationsand call for a microscopic understanding of spin models.For example, based on a strong-coupling perturbation the-ory of the generic spin model[2–4], it was shown that α-RuCl3 described by the effective spin Jeff = 1/2 has dom-inant bond-dependent Kitaev and off-diagonal symmetricΓ interactions.[5, 6] RuCl3 has become an emergent can-didate of spin-1/2 Kitaev spin liquid[7]. Intense researchactivities on various properties of RuCl3 have been carriedout[6, 8–14] and recently a magnetic-field induced spin liq-uid was suggested.[15–21]

    In parallel theoretical interest in the ground state ofhigher-spin Kitaev models was initiated by classical modelstudies.[22, 23] The classical Kitaev model has a macro-scopic degeneracy named a classical spin liquid[22], but thehigher-spin quantum Kitaev model is not exactly solvable,and the ground state is currently unknown. Various numer-ical studies such as exact diagonalization on S=1 suggestedthat the ground state is possibly a spin liquid with gaplessexcitations.[24] These studies were mainly of theoretical in-terest, until a microscopic derivation of the S=1 Kitaev-Heisenberg model in multi-orbital systems was found.[25]Heavy ligand spin-orbit coupling (SOC) and strong Hund’scoupling in eg orbitals is a way to generate S=1 bond-dependent Kitaev interaction. The magnetic field effectson the S=1 Kitaev model have also been investigated.[26–28]

    The bond-dependent interactions have recently beenadopted into TMT systems, because the nearest neighbor

    [email protected]

    (n.n.) Heisenberg J , Kitaev K and Γ interactions are al-lowed based on the symmetry of the lattice.[4, 29, 30] Inparticular, ferromagnetism in a single-layer CrI3 has gener-ated excitement in recent years. [31–45] CrI3 is a ferromag-netic (FM) insulator with Tc ∼ 61K for bulk samples.[46–48] Single-layer CrI3 was successfully synthesized, whichshowed an FM ordering with Tc ∼ 45K. [49] The two-dimensional FM Heisenberg model is insufficient to explainfinite Tc, i.e., the Mermin-Wagner theorem[50], and severaltheoretical models were proposed to explain the magneticanisotropy. They include the XXZ model[51, 52], single-ion anisotropy and Kitaev[53], and large Kitaev and smallsymmetric off-diagonal Γ interactions[54].

    While the Heisenberg, Kitaev, and Γ interactions are al-lowed by the symmetry, and found to be significant in theearlier derivations for lower-spins[3, 4, 25], their strengthsmay not be significant in S=3/2 systems. Thus, a micro-scopic derivation of S=3/2 model is necessary to find thesources of the magnetic anisotropy. Here we derive a n.n.spin model for S=3/2 with three electrons in t2g orbitalsof transition metal sites and strong SOC in p-orbitals ofligands. We take into account strong electron-electron in-teractions in multi-orbital systems including Hund’s cou-pling, and effects of trigonal distortions present in R3̄rhombohedral lattice. Contributions from eg orbitals areimportant as shown below. The minimal n.n. model in-cludes J , K, Γ, another symmetric off-diagonal Γ′[55], andsingle-ion anisotropy Ac along the ĉ-axis, denoted as theJ −K − Γ− Γ′ −Ac model.

    The rest of the paper is organized as follows. In Sec. II,the on-site Kanamori interaction and tight binding Hamil-tonian are presented. In Sec. III, we derive the n.n. spinmodel consisting of Kitaev and Heisenberg interactions forideal octahedra environment using standard perturbationtheory. In Sec. IV, we study the spin model with trigonaldistortion present in R3̄ rhombohedral lattice. This in-cludes the distortion-induced hopping matrix elements andthree additional spin interactions generated via combinedeffects of SOC and distortion. In Sec. V, we apply the the-ory to CrI3 and present the exchange interaction strengthsusing tight binding parameter sets obtained by densityfunctional theory. The effects of the resulting magnetic

    arX

    iv:2

    009.

    0447

    5v2

    [co

    nd-m

    at.s

    tr-e

    l] 1

    6 M

    ar 2

    021

    mailto:[email protected]

  • 2

    Figure 1. Edge-shared octahedra honeycomb structure unit cella, b, in global coordinates xyz. Transition metal sites M ingray and non-magnetic ligands X in purple. The n.n. bonds X,Y, Z are related by C3 symmetry. The sites M1, M2, X1, and X2are involved in the second order strong-coupling expansion onthe Z bond. Indirect hopping integrals t0, t1 and t2 are shown.

    anisotropy on the moment direction are found in Sec. VI. InSec. VII we discuss the origin of the spin gap, finite Tc, andspin wave spectrum within the J −K−Γ−Γ′−Ac model.The effects of the the second n.n. Dzyaloshinskii-Moriya(DM) interaction is also discussed. Finally in Sec. VIII,we summarize our results and compare with Jeff=1/2 andS=1 spin models. The detailed calculations are presentedin the Appendix.

    II. KANAMORI INTERACTION AND TIGHTBINDING HAMILTONIAN

    The honeycomb network is made of metal (M) d-orbitalsites with half filled t2g orbitals and octahedra cages of non-magnetic ligand (X) sites with fully occupied p-orbitals.The full Hamiltonian is composed of the on-site Kanamoriinteraction and the tight binding Hamiltonian between twosites.

    The on-site Hamiltonian of the M sites is described by theKanamori interaction [56] as well as crystal field spitting:

    Hee = U∑α

    nα↑nα↓ +U ′

    2

    ∑α6=β,σ,σ′

    nασnβσ′

    −JH2

    ∑α6=β,σ,σ′

    c†ασc†βσ′cβσcασ′ + JH

    ∑α 6=β

    c†α↑c†α↓cβ↓cβ↑

    +∆c∑α∈eg,σ

    c†ασcασ, (1)

    where the density operator nασ is given by c†ασcασ, and c

    †ασ

    is the creation operator with α orbital and spin σ. U and

    U ′ are the intra-orbital and inter-orbital Hubbard inter-action respectively, and JH is the Hund’s coupling for thespin-exchange and pair-hopping terms. ∆c is a crystal fieldsplitting on the M sites, originated from the surroundingoctahedra, leading to the splitting of the d-orbitals into t2gand eg orbitals. In a d

    3 system one has half-filled t2g or-bitals, where the Hund’s coupling JH selects for the S=3/2configuration as the ground state, and the angular momen-tum is quenched. A table of the excited state energy spec-trum is show in Appendix 1. The energies of the exitedstates are larger than the hopping integrals, which allowsus to treat the tight binding hopping integrals as a pertur-bation.

    In the edge shared octahedra structure, each bond be-tween n.n. M sites involves two adjacent ligands, as shownin Fig. 1. A tight binding Hamiltonian between two tran-sition metal sites M1 and M2 including the two adjacentligands X1 and X2 is given below

    HTB =

    05×5 TM1M2 TM1X1 TM1X2

    TM1M2† 05×5 TM2X1 TM2X2

    TM1X1† TM2X2

    † 03×3 03×3

    TM1X2† TM2X2

    † 03×3 03×3

    , (2)

    where 0n×n refers the n × n null matrix. The basisis chosen as (C†M1,d, C

    †M2,d

    , C†X1,p, C†X2,p

    ), where C†Mi,d =(c†i,x2−y2 , c

    †i,3z2−r2 , c

    †i,yz, c

    †i,xz, c

    †i,xy,

    )are five d-orbitals at

    site Mi, and C†Xm,p

    =(c†m,px , c

    †m,py , c

    †m,pz

    )are three p-

    orbitals at ligand site Xm. Each block of indirect hoppingbetween Mi and Xm is denoted by TMiXm and the directhopping between M sites by TM1M2 . The details of eachblock matrix will be presented latter.

    To account for the indirect d to p hoppings we integrateout the p-orbitals through a perturbative procedure trun-cated at second order, leading to an effective d to d hoppingmodel:

    TeffMiMj =∑

    (a,m)

    TMiXm |a〉〈a|TXmMj∆Ea

    , (3)

    where (a,m) represent a sum over all single hole statesa of all sites Xm. The hole states are SOC states, thuscreating two energy costs ∆Ea = ∆−λp/2 or ∆+λp, where∆ = �d− �p is the atomic energy difference between M andX sites, and λp is the SOC in p-orbitals. The SOC willintroduce explicit spin dependence in the effective d to dhopping. The total effective hopping between the two Msites now reads

    HeffTB =

    010x10 TM1M2 + TeffM1M2TM1M2

    † + TeffM1M2†

    010x10

    ,(4)

    where we still retain the bare direct hoppings TM1M2 . Be-low we focus on the ideal honeycomb structure and firstexamine the effects of TeffM1M2 before adding the direct hop-pings TM1M2 and summarizing the resulting spin model.

  • 3

    III. IDEAL HONEYCOMB STRUCTURE

    To understand the microscopic origin of the spin model,we start with the ideal honeycomb network surrounded byperfect edge shared octahedra. It was shown that the sym-metry of the edge-shared octahedra Z bond allows Heisen-berg J , Kitaev K, and symmetric off-diagonal Γ interac-tions [4, 29, 30]. However, since their strengths depend onvarious exchange processes, we perform the strong couplingperturbation theory to determine the exchange terms.

    Truncating at second order in perturbation theory, we ar-rive at the following Heisenberg-Kitaev (J−K) spin modelfor the ideal honeycomb octahedra.

    H =∑〈ij〉∈γ

    J0Si · Sj +K0Sγi Sγj , (5)

    where γ = x, y, z bond, and J0 and K0 refer to Heisenbergand Kitaev interactions for the ideal octahedra.

    Below we present the details of the derivation of Heisen-berg and Kitaev interactions. An explanation of the ab-sence of the Γ interaction within the second order pertur-bation theory is also discussed. The exchange processes in-clude the contributions from both indirect and direct hop-pings. We focus on the Z bond of the honeycomb Fig. 1,as the other two bonds are related by C3 symmetry.

    A. Superexchange path: indirect hopping

    We first consider indirect hoppings between the M and Xsites, which are the largest hopping integrals. They enterthrough the effective hoppings of Eq. (4). The non-zero in-direct hopping between M1 and X1 sites t0, t1, t2 are shownin Fig. 1. These hoppings are incorporated in TeffM1M2 ,which can be simplified using the Slater-Koster decompo-sition and symmetry related M −X bonds as shown in theAppendix. There are two contributions to both Heisenbergand Kitaev interactions, i.e., each interaction is composedof two exchange terms; one is from t2g − t2g hoppings andthe other is from eg − t2g hoppings

    J0 = Jt2g0 + J

    eg0 , K0 = K

    t2g0 +K

    eg0 , (6)

    where the superscript t2g and eg refer to its correspondinghopping processes. Below we present each exchange path

    leading to Jt2g0 , K

    t2g0 , J

    eg0 , and K

    eg0 .

    1. t2g − t2g contributions

    Introducing the effective hopping integral teff =t203

    (2

    ∆− λp/2+

    1

    ∆ + λp

    )between M1 and M2 via p-

    orbitals, and the ratio r =2λp

    2∆ + λpbetween SOC λp and

    the atomic energy difference ∆, the hopping matrix involv-ing only t2g orbitals, denoted by t0 in Fig. 1, can be sim-

    plified in block form to

    TeffM1M2(t2g ⊗ t2g) = teff

    02×2 σo i

    r

    2σx

    σo 02×2 −ir

    2σy

    −i r2σx i

    r

    2σy 02×2

    ,(7)

    where t2g = {dyz, dxz, dxy} and σi with i = x, y, z are thePauli matrices carrying the spin degrees of freedom and σois the 2 × 2 identity matrix. The holes in the intermedi-ate sates at the X site and their indirect hopping TMiXmdetermine the type of σ matrices in the effective hoppingFor example in the limit λp → 0, only the dyz − dxz termteffσ0 → t20/∆ is present which contributes to the directhopping channel. The new terms present for non-zero λpare the spin-flip (SF) terms between dyz/dxz and dxy. Suchterms would normally not appear in the second order per-turbation process, as they involve a dyz/dxz − pz hoppingfollowed by a px/py−dxy hopping, which will only occur ifpz is entangled with px/py. The SOC among the px, py, pzgenerates such entanglement. Furthermore, when SOC isthe dominant energy scale of the hole states, the wavefunc-tions are inevitably mixtures of p-orbitals and their spin,leading to σi dependence proportional to the r ratio of theSOC and atomic energy difference.

    The superexchange process involving only t2g orbitalsEq. (7) result in

    Jt2g0 =

    8t2eff9 (U + 2JH)

    , Kt2g0 = −

    4(rteff)2

    9 (U + 2JH). (8)

    The spin-dependent hoppings have generated a Szi Szj Ki-

    taev interaction. This can be rudimentarily understood bythe following steps. To simplify the steps, we focus onone spin 1/2 electron hopping along the Z bond, throughonly a SF hopping. Imagining two sites starting in (↑, ↑)state, the SF hopping can lower the energy by the pro-cess: (↑, ↑) − SF − (0, ↓↑) − SF − (↑, ↑), at a energy costof −(rteff)2/U . On the other hand, if the two sites startin (↑, ↓) the SF process is forbidden from Pauli exclusionprinciple. Thus the exchange path starting from (↑, ↑) andending in (↑, ↑) lowers the energy, which generates the FMSzi S

    zj interaction. Carrying out the details for the S=3/2

    leads to the expression of the spin exchange terms shownin Eq. (8).

    Among the symmetry allowed terms, the symmet-ric off-diagonal term with operator Sxi S

    yj + S

    yi S

    xj =

    i

    2

    (S−i S

    −j − S

    +i S

    +j

    )does not occur in the second order per-

    turbation results. This operator would connect (↑, ↑) to(↓, ↓), which is definitely possible from a SF process, how-ever, there is a subtle cancellation. Having a single Paulimatrix in the SF term creates such cancellation among thetwo paths within the spin block, resulting in a null Γ term.Thus it becomes finite only when higher order perturbationterms are included, or when the octahedra are no longerideal, as we will show in the later Sec. IV.

  • 4

    2. eg − t2g contributions

    Given that the p-orbital’s hybridization with eg is largerthan with t2g, denoted by t1 and t2 in Fig. 1,this con-tribution is essential. The final form of the effectiveTeffM1M2(eg⊗t2g) hopping includes SF terms in the dyz, dxzto eg blocks, as well as spin-independent −2teff(t2/t0)σ0hopping between dxy and d3z2−r2 . The effective hoppingmatrix involving eg − t2g hoppings, in block matrix form,reads

    TeffM1M2(eg⊗t2g) = teff

    ir

    2

    t1t0σy i

    r

    2

    t1t0σx 02×2

    −i r2

    t2t0σy i

    r

    2

    t2t0σx −2

    t2t0σo

    ,(9)

    where eg = {dx2−y2 , d3z2−r2} and t2g = {dyz, dxz, dxy}.Carrying out the strong coupling expansion, these hoppingslead to the additional contribution to the Heisenberg andKitaev interactions

    Jeg0 = −

    16 JH t2eff

    3 (∆c + U ′ − JH) (∆c + U ′ + 3JH)t22t20,

    Keg0 =

    2 JH(r teff)2

    3 (∆c + U ′ − JH) (∆c + U ′ + 3JH)t21 + t

    22

    t20.

    (10)

    Similar to the t2g − t2g case, the SF terms contribute tothe Kitaev interaction, while the spin-independent hop-ping, −2teff(t2/t0)σ0, generates a FM Heisenberg term.The FM Heisenberg interaction originates from the com-petition of two exited states separated by Hund’s coupling,i.e, eg paths consistent with the earlier findings obtainedby the first principle calculations.[45, 57]

    B. Direct hopping

    The direct hopping, denoted by td1, td2, td3 and t̃d0 inTM1M2 between M1 and M2 is given by

    TM1M2 (t2g ⊗ t2g) =

    td1 td2 0td2 td1 00 0 td3

    ⊗ σo,TM1M2 (eg ⊗ t2g) =

    (0 0 00 0 t̃d0

    )⊗ σo.

    (11)

    The Slater-Koster decomposition is explained in the Ap-pendix. Since there is no SF hopping terms, this does notgenerate the Kitaev interaction, but changes the Heisen-berg interaction.

    C. Summary and Comments

    Combining both indirect and direct hopping contribu-tions, the two exchange interactions for the ideal octahedra

    environment Eq. (5) are found to be

    J0 =4(2t2d1 + 2(teff + td2)

    2 + t2d3)

    9 (U + 2JH)

    −4JH

    (2teff(t2/t0)− t̃d0

    )23 (∆c + U ′ − JH) (∆c + U ′ + 3JH)

    ,

    K0 = −4 (r teff)

    2

    9 (U + 2JH)

    +2 JH(r teff)

    2

    3 (∆c + U ′ − JH) (∆c + U ′ + 3JH)t21 + t

    22

    t20. (12)

    Thus for the ideal honeycomb structure, the J −K modelis derived within the second order perturbation theory.

    Some comments are useful, which will also motivate fur-ther investigation on the effects of trigonal distortion pre-sented in the next section. Kitaev and Heisenberg inter-actions have been generated while the Γ interaction hasnot appeared at second order due to a subtle cancella-tion. The Kitaev has been generated purely from SF hop-pings and has a prefactor of r2 = (2λp/(2∆ + λp))

    2, whilethe Heisenberg includes spin-independent hopping contri-butions. Nevertheless both K and J have two contributionswhich come with opposite signs: one from t2g only pathsand the other from the paths involving eg. This leads to asmaller Kitaev compared to Heisenberg interaction, unlessthe contribution from eg paths reduces the overall strengthof the Heisenberg interaction. So long as the crystal fieldspitting is not excessively large, the naturally larger eg hop-pings will drive the system to a FM Heisenberg interaction.

    The FM J − K model pins the moment along the cu-bic axis when the quantum fluctuations are taken intoaccount[58, 59], not along the observed [111] direction inCrI3. We thus investigate if other interaction terms maybe generated. The Γ interaction allowed by the symme-try can be finite if higher order perturbation terms areincluded. However, one may ask if there are other interac-tions allowed by a slight distortion of the lattice within thesecond order perturbation theory without invoking higherorder terms. Indeed TMT materials do not have ideal oc-tahedra, but have either rhombohedral R3̄ or monoclinicC2/m structures, and their magnetism strongly depends onstructural differences and number of layers.[60–63] Belowwe study the effects of distorted octahedra, which induceadditional hopping integrals which were forbidden withoutthe distortion.

    IV. EFFECTS OF DISTORTION: DISTORTEDOCTAHDERA

    CrI3 goes through a structural transition from C2/m toR3̄ structure at low temperature.[48] In the rhombohedralstructure, there are two types of X ligand position devi-ations from the ideal octahedra structure. As shown inFig. 2(a), a single octahedron can be viewed as two shadedtriangles. One distortion is the staggered rotations of thetwo triangles denoted by δx blue arrows, with displace-ments of X sites perpendicular to the ĉ = [1, 1, 1]/

    √3 di-

    rection. The other distortion is the compression of the dis-tance between these two triangles along the ĉ-axis denoted

  • 5

    a)

    b)

    Figure 2. a) The distorted octahedra in R3̄ is shown. The octa-hedron made of X ligand can be viewed as two yellow trianglesnormal to the ĉ = [1, 1, 1]/

    √3 direction. The blue arrows rep-

    resent new positions of X due to the staggered rotations of thetwo yellow triangular faces. The change of position X due to thestaggered rotations is parameterized by δx. In addition, thereis a compression of the two yellow triangles, squeezed as shownby the orange arrows parallel to the ĉ-axis. The change of posi-tion due to the compression is parameterized by δx′. b) A topview of distortions in a unit cell is shown. The dotted circles �indicate the new position of X moving out of the page along +ĉwhile the circled cross ⊗ indicate a new position of X movinginto the page toward −ĉ. The exact positions of X(1,2,3,4,5,6) asa function of δx, δx′ are found in the Appendix.

    by δx′ orange arrows, with displacements along the ĉ di-rection. Here the dimensionless parameters δx and δx′ arein units of the distance between the n.n. M sites dM . An-alytic formulas of the new positions of X sites under stag-gered rotations and compression are found in Appendix 3.In the R3̄ space group there are other types of distortions,namely the M1−X2,4,6 bond length can be different fromthe M1−X1,3,5 bond length. This type of distortion is gen-erally exceedingly small compared to δx and δx′ and weneglect it in the following analysis. Fig. 2(b) shows a topview of the the honeycomb unit cell with two such distortedoctahedra forming the Z bond, and the staggered rotationof the right octahedron is the mirror image of the left oc-tahedron.

    The distortion-induced hopping matrices have all ele-ments non-zero in TMiXm as a result of lowering the lo-cal symmetry of the octahedron from Oh to D3. We de-note the new distortion allowed hopping integrals as δti.Starting with the distortion-induced hoppings, we followthe procedure described in Sec. III, namely we use thedistortion-induced TMiXm matrices to derive the effectiveTeffM1M2(t2g ⊗ t2g) and T

    effM1M2

    (eg ⊗ t2g). Details of theirform is deferred to Appendix 4. Treating the distortion-induced effective hoppings as a perturbation against theon-site interactions Eq. (1) the minimal n.n. spin model isfinally given by

    H =∑

    〈ij〉∈αβ(γ)

    [JSi · Sj +KSγi S

    γj + Γ(S

    αi S

    βj + S

    βi S

    αj )

    + Γ′(Sαi Sγj + S

    βi S

    γj + S

    γi S

    αj + S

    γi S

    βj )]

    +∑i

    Ac(Si · ĉ)2,

    (13)where α, β, (γ) refers to the γ bond taking α and β spin

    components [4, 55], and ĉ = [1, 1, 1]/√

    3. In addition tothe J −K terms, two symmetric off-diagonal terms Γ andΓ′ have been generated, as well as the single-ion term Acallowed by the C3 symmetry present on every site. Γ, Γ

    and Ac are proportional to distortion induced hoppings δtas well as r, thus both distortion-induced hoppings andSOC is needed to bring rise to these terms. The form ofthe spin model J −K −Γ−Γ′ −Ac to leading order in δtias well as eg contributions are listed in Table VI in the Ap-pendix. Note that both Heisenberg and Kitaev interactionsare renormalized by the distortion, but the Heisenberg in-teraction has a linear term in δt, while Kitaev interactiondoes not.

    V. APPLICATION TO CRI3

    To apply the above model to understand the magnetismin CrI3, we use ab initio calculations to obtain the mi-croscopic parameters. The calculation is performed withVienna ab initio simulation package (VASP)[64]. We usethe Perdew-Burke-Ernzerhof (PBE) functional[65] in ourcalculations. The experimental bulk structure[48] is usedfor the calculations. We use WANNIER90[66] to extractthe hopping integrals.

    From the density functional theory results, we estimatethe crystal field splitting ∆c = 1253 meV from on-site Crd-orbitals as well as a crystal field splitting from the on-site I p-orbitals of ∆p = 528 meV. From Cr d- and I p-orbitals we extract the atomic energy difference betweenCr and I ∆ = 2070 meV. Finally we find the dominantindirect p−d hopping integrals t0, t1 and t2 and the directd − d hoppings parameters, listed in Table I, as well asthe distortion induced p − d hoppings, listed in Table II.We verified that t0, t1 and t2 obtained by the ab initiocalculation match well with the Slater-Koster expectationsof Eq. (A.2).

    In the ab initio calculations, we find a sizable crystalfield spitting ∆p on the ligand I sites, which we have nottaken into account in the earlier analysis. It is about

  • 6

    t0 t1 t2 td1 td2 td3 t̃d0590.1 -992.13 -558.3 44.67 -41.26 -147.44 -20.75

    TABLE I. Indirect and direct hoppings in units of meV obtainedby ab initio calculations. The hopping integrals are defined inEq. (A.1) and Eq. (11).

    δt1 δt2 δt3 δt4 δt5 δt′1 δt

    ′2 δt

    ′3 δt

    ′4

    2.34 21.76 -22.81 -61.74 65.86 70.97 -43.50 -49.41 29.65

    TABLE II. Distortion-induced indirect hoppings in units ofmeV obtained by ab initio calculations. The hopping integralsare defined in Eq. (A.5).

    528 meV, which is comparable to an estimated SOC pa-rameter of λp = 630 meV. To capture its effects we re-visit the effective hopping derivation, and add the crystalfield spitting ∆p on the X sites. Then the energy level

    in Eq. (3) split into three levels as ∆Ea = ∆ − λ2 ,∆ +14

    (2∆p + λ±

    √(2∆p − λ)2 + 8λ2

    ). While the holes in I

    sites are no longer pure total angular momentum states,the SOC is sizable enough to carry the spin entanglementthrough to the spin model. We found that second orderperturbation theory, including the spitting ∆p in the effec-tive hoppings, does not change the form of the spin modelEq. (13). The analytic formulas would be vastly more com-plex in this case, so we proceed to a direct numerical eval-uation of J −K − Γ− Γ′ −Ac exchange terms using effec-tive hoppings TeffM1M2 obtained by the ab-inito calculationslisted in the Tables.

    Assuming the spherical symmetry, i.e., U ′ = U − 2JH ,we are left with two unknown parameters U and JH . Weplot J , K, Γ, Γ′, and Ac as a function of JH/U for severalU values, with results shown in Fig. 3. The dominant termis Heisenberg except near the range 0.11 . JH/U . 0.13where J is almost zero before it changes the sign. Thesign of J(K) are sensitive to the ratio JH/U and an ad-equately large Hund’s coupling allows the FM Heisenberg(AFM Kitaev) interaction to persevere. This reflects thecompetition of the t2g vs eg terms seen in Eq. (12), witheg eventually wining over t2g leading to the FM Heisen-berg interaction. The distortion induced Γ− Γ′ −Ac comein with FM sign, and the single-ion anisotropy Ac is thesizable term. Adopting the ratio of JH/U ∼ 0.24 ob-tained by cRPA clculations [67], CrI3 sits in the shadedarea in Fig. 3. CrI3 is a ferromagnet with non-negligibleK − Γ − Γ′ − Ac, and we now examine the implication ofthese exchange terms on the observed moment.

    VI. MAGNETIC ANISOTROPY

    We have shown how the ideal honeycomb structure leadsto J −K model up to second order in perturbation, whiledistortions additionally generate Γ−Γ′−Ac. The magneticmoment of the ferromagnetic state obtained with the FMJ−K model is pinned along the cubic axis such as [100] andC3 equivalent directions via quantum fluctuations[58, 59].

    Figure 3. Spin model parameters J , K, Γ, Γ′, and Ac estimatedfrom ab initio parameters and plotted against JH/U . Solid,dashed, and dotted lines correspond to U = 3eV, 4eV and 5eVrespectively. Shaded region corresponds to JH/U ∼ 0.24 whichis relevant to CrI3 microscopics.

    We take a closer look at the effects of Γ − Γ′ − Ac on themoment direction. We show three examples in Fig. 4 toillustrate the moment pinning direction.

    Following the method in Ref. [59], we perform exact diag-onalization calculations on an eight site honeycomb cluster,with the cluster setup shown in the Appendix. Once theground state wavefunction |GS〉 is obtained, the probabilitydistribution P = |〈ΨFM (θ, φ)| GS〉|2 is computed, where|ΨFM (θ, φ)〉 is a ferromagnetic ansatz with moment direc-tion pointing at (θ, φ) on the sphere. Results are show inFig. 4 for different values of J , K, and Γ. Independent of ra-tio of J and K, the moment is along the cubic axis as shownin the panel (a) and (b). On the other hand, when a smallFM Γ is introduced, the moment is along the ĉ direction asshown in panel (c). This effect can be anticipated from the

    a) b) c)

    Figure 4. Moment pinning calculation for difrent values of(J,K,Γ) in the ferromagnetic phase: a) (−1.00,−0.20, 0.00),b) (−0.20,−1.00, 0.00), c) (−0.20,−1.00,−0.02). Kitaev inter-action always prefers the cubic axis in panel a) and b). A smallinteraction like Γ is necessary to pin along the ĉ direction asshow in panel c).

  • 7

    classical analysis. The classical J−K−Γ model in the FMstate with moment S0 = (S

    x0 , S

    y0 , S

    z0 ) has an energy density

    per unit cell uc = (3J+K)+2Γ(Sx0Sy0 +S

    y0S

    z0+S

    z0S

    x0 ), which

    demands min [uc] ⇒ min [sgn(Γ)(Sx0Sy0 + S

    y0S

    z0 + S

    z0S

    x0 )].

    When Γ < 0 we have min [−(Sx0Sy0 + S

    y0S

    z0 + S

    z0S

    x0 )] ⇒

    S0 = ĉ leading to a moment pined on the [1,1,1]. We chooseΓ/K � 1 to show that a tiny FM Γ anisotropy results inthe ĉ-axis moment shown in Fig. 4 (c).

    FM Γ′ has the identical effect as FM Γ with respect tothe moment direction of the ferromagnet. Clearly FM Acalso promotes the ĉ-axis moment direction. All three inter-actions individually prefer the moment on the ĉ-axis. Thiseffect is bigger than the quantum fluctuation effect of cubicaxis pinning from the Kitaev term resulting in the observedĉ-axis moment direction.

    VII. SPIN GAP AND FINITE TRANSITIONTEMPERATURE

    The preceding results suggest that it is likely that CrI3has dominant FM Heisenberg and smaller Kitaev interac-tion, and the magnetic anistoropy originates from the dis-tortion of the octahedra together with the SOC of heavy lig-ands. The magnetic ordering pattern changes from bulk tofilms [49, 63, 67–70], suggesting a strong coupling betweenmagnetism and crystal structures, which further impliesthe importance of the distorted octahedra in the presenceof SOC. Our microscopic model has an interesting connec-tion to the previous studies. The Kitaev and single-ionanistoropy in addition to the Heisenberg interaction werefound in Ref.[53] using density functional theory. If Γ = Γ′,the J−Γ−Γ′ model maps to the XXZ model[51, 52] in thea− b− c crystallographic coordinate

    JSi · Sj+Γ(Sxi Syj + S

    yi S

    xj + S

    xi S

    zj + S

    zi S

    xj + S

    yi S

    zj + S

    zi S

    yj )

    =⇒ JabSi · Sj + JcSciScj , (14)

    where Jab = J − Γ and Jc = 3Γ.Since the exchange parameters strongly depend on

    JH/U , ∆c/U , and tight binding parameters, it is use-ful to obtain experimental inputs to determine some pa-rameters. The inelastic neutron scattering experimentsand magneto-Raman spectroscopy have reported a spingap of approximately 0.36 meV at the Brillouin zone(BZ) center Γ-point.[71, 72] This is also consistent witha small anistoropy found in the ferromagnetic resonanceexperiment[54], which is about 0.07 meV leading to thespin gap of 0.3 meV.

    Based on our spin wave analysis using the J −K − Γ−Γ′ − Ac model including the second n.n. DM interaction,the spin wave dispersion ωk is expressed as ω0+ρ k

    2 aroundthe Γ-point. Here ω0 and ρ are the spin gap and stiffness,respectively, and they are given by

    ωo = −S(3Γ + 6Γ′ + 2Ac),

    ρ =S

    12

    ∣∣∣3J +K − Γ− 2Γ′ − (K+2Γ−2Γ′)22(2Γ+4Γ′+2Ac+3J+K) ∣∣∣ (15)The details of the linear spin wave theory (LSWT) is pre-sented in Appendix 7.

    Kitaev and DM interactions do not generate a gap atthe Γ-point within the LSWT. The classical ferromag-netic ground state under the Kitaev and DM terms have acontinuous degeneracy, and as a result expanding aroundthis ground state in the LSWT will not result in a spingap from these terms. The spin gap is rather small, asexpected because it is originated from a combination ofslightly distorted octahedra and SOC, i.e., Γ, Γ′, andAc. While small, it is essential for a finite Tc in a sin-gle layer CrI3. In the FM ordered phase, at low tem-peratures the magnons are excited and their number isgiven by Ns(T ) =

    ∫d2k 1

    eβωk−1 =πβρ

    ∫βω0

    dxex−1 . Without

    the spin gap ω0, Ns diverges in two-dimension at any tem-perature except T = 0, i.e., the celebrated Mermin-Wagnertheorem[50]. Thus one can understand the essential role ofω0 which cuts the divergence, and allows the FM orderingat finite temperatures, as long as ω0(T ) remains finite forT ≤ Tc. While quantifying the transition temperature re-quires further analysis [73] , the temperature dependenceof ρ(T ) and ω0(T ) from the inelastic neutron scatteringmeasurement[71] indicates the crucial role of ω0(T ) whichvanishes at Tc.

    Another important parameter is the Kitaev interac-tion K, which leads to a gap at the the BZ corner K-point known as the Dirac gap[54], reported in the neutronscattering.[71, 74] However, the second n.n. DM term alsogenerates the Dirac gap.[71, 74] We would like to point outthat Γ and Γ′ also play a part in the Dirac gap as shownin Eq. (A.14). Estimating the Kitaev interaction by an in-dependent experimental measurement and further analysison the individual role of the Kitaev and DM interactionsremain to be resolved in future studies.

    VIII. SUMMARY AND DISCUSSION

    We have shown a microscopic derivation of the n.n. spinmodel for honeycomb Mott insulators with three electronsin t2g orbitals at M sites surrounded by octahedral cagesof ligands X with strong SOC. Using the standard strong-coupling perturbation theory, we found that there are onlyHeisenberg and Kitaev interactions for the ideal honey-comb lattice among the three symmetry allowed interac-tions (J,K,Γ), because Γ is zero up to the second order per-turbation. The exchange paths between t2g and t2g vs. t2gand eg via ligands generate opposite signs for both Heisen-berg and Kitaev interactions. The Heisenberg interactionis of order t2eff/U , while the Kitaev is smaller by a factor ofr2 ∼ (λp/∆)2. The FM Heisenberg interaction originatesfrom the eg paths, with the hopping integral between egand p-orbitals being larger compared to t2g and p-orbitals.

    The FM Heisenberg and Kitaev interaction leads to FMordering, but the moment direction is pinned along thecubic x, y, or z axis, e.g., [100] (and C3 equivalent direc-tions) via quantum fluctuations. The ĉ-axis, [111], momentpinning found in CrI3 should thus originate from other in-teractions, which are also responsible for the spin gap atthe Γ-point in the neutron scattering measurement[71]. In-cluding the distorted octahedra present in the rhombohe-dral structure, three additional spin interactions are found.Inspecting the linear order in the distortion-induced hop-

  • 8

    ping paths within the second order perturbation theory, Γ,Γ′ and single-ion anisotropy Ac contain terms linear in thedistortion-induced hopping integrals. The Heisenberg in-teraction also contains such additional linear term, but theKitaev does not, implying that it is possible to fine-tune asystem closer to the Kitaev dominant regime via trigonaldistortions.

    In this work we have focused on the nature of the mono-layer, however, some comments are in place when consider-ing the physics of the bulk. We showed that the t2g−t2g vs.eg − t2g contributions to the intra-layer Heisenberg term Jare opposite in sign. This property should hold for inter-layer Heisenberg interaction J⊥ as well, because it is deter-mined from superexchange processes. Thus the magneticordering pattern between layers depends on the details oforbital compositions. From the ab initio calculation, wefound that the inter-layer hopping ranges from 10meV to30meV. This leads to J⊥ of order 0.1 meV. While it doesnot affect the intra-layer magnetism presented here, it isimportant for the ordering pattern in the bulk[49, 63, 67–70].

    Comparison to Jeff = 1/2, S=1, and 3/2 spin systemswould be useful. The SOC is necessary to generate thebond-dependent interaction, as spin and orbitals should beentangled to get such a directional dependent spin interac-tion. However, the presence of SOC is not enough to findan exotic phase like a spin liquid, because the dominant in-teraction is often the Heisenberg interaction. To comparedifferent spin cases, a summary of the ideal honeycomb ex-change interactions, J0, K0, and Γ0 including the effectiveindirect hopping (teff) and direct hopping (td) integrals,only up to the second order perturbation, is shown in thefollowing Table III.

    Spin Heisenberg J0 Kitaev K0symm.

    off-diagonal Γ0

    Jeff = 1/2 (d5) O(

    t2dt2eff

    ) O(JHU

    ) O( tdteff

    JHU

    )

    S = 1 (d8) O(r2) O(r2) 0

    S = 3/2 (d3) O(1) O(r2) 0

    TABLE III. The leading term of the exchange interactions forthe ideal honeycomb structure in unit of t2eff/U for different spinS including only up to the second order perturbation terms.See the Appendix for the full expression of JKΓΓ′Ac for S=3/2including the trigonal distortion-induced hopping contributions.

    Focusing on the ideal octahedra and n.n. model via sec-ond order superexchange processes, Jeff = 1/2 is uniquebecause the Heisenberg term is absent. On the other hand,for the S=1 model from d2 in eg-orbitals, the heavy ligandSOC generates the Kitaev interaction, which has the sameorder of magnitude with J . In fact, K = −2J , if only egpaths are considered.[25] For S=3/2 case, we found that Jis order 1 in units of roughly t2eff/U , while K is smaller by

    r2. Thus it is hard to compete with the Heisenberg inter-action. We speculate that this is valid for spins equal orhigher than 3/2. Unlike S=1/2 and S=1 cases, the Heisen-berg interaction is dominant in S=3/2, but a delicate can-cellation among different contributions to the Heisenberginteraction, may let the Kitaev interaction overtake a ma-jor place. In particular, the Heisenberg interaction is moresensitive to the distortion-induced hopping integrals thanthe Kitaev term as shown in Table VI in the Appendix,manipulating ocathedra may be a way to tune the systemto a desired Kitaev dominant regime.

    In summary, in the ideal octahedra environment, we findthat there are only two spin interactions, Heisenberg andKitaev interactions. Kitaev interaction is generally weakercompared to the Heisenberg interaction in contrast to thelower spin models. Indirect hoppings among t2g and t2gvs. eg and t2g have opposite contributions. A detailedbalance between the two indirect and direct hopping con-tributions highly depends on the hopping integrals, Hund’scoupling strength, and crystal field spitting. Γ interactionis absent up to the second order due to a subtle cancella-tion. We further show that trigonal distortion allows threeadditional interactions, two symmetric off-diagonal inter-actions Γ and Γ′, and single-ion anisotropy Ac along theĉ-axis. They are all linearly proportional to a distortion-induced hopping integral. While they are much smallerthan the Heisenberg interaction, they are essential for aspin gap in the FM phase of CrI3 leading to a finite Tc.Our study offers a microscopic route to the n.n. spin mod-els, J − K − Γ − Γ′ − Ac interactions. Given that thereare five exchange terms within the n.n. model, and secondn.n. interactions including DM may be comparable to Γ,Γ′ and Ac, further theoretical and experimental studies arerequired to determine the microscopic parameters of CrI3beyond the n.n. model.

    -¿

    ACKNOWLEDGMENTS

    We acknowledge I. Lee, C. Hammel, N. Trivedi, P. Dai,and R. Valenti for useful discussion. H.Y.K. acknowl-edges the funding from the Canada Research Chair Pro-gram. This work was supported by the Natural Sciencesand Engineering Research Council of Canada, the Centerfor Quantum Materials at the University of Toronto, andthe Canadian Institute for Advanced Research. Compu-tations were performed on the Niagara supercomputer atthe SciNet HPC Consortium. SciNet is funded by: theCanada Foundation for Innovation under the auspices ofCompute Canada; the Government of Ontario; Ontario Re-search Fund - Research Excellence; and the University ofToronto.

    [1] M. A. McGuire, Crystals 7, 121 (2017).[2] G. Khaliullin, Prog. Theor. Phys. Supp. 160, 155 (2005).[3] G. Jackeli and G. Khaliullin, Phys. Rev. Lett. 102, 017205

    (2009).[4] J. G. Rau, E. K.-H. Lee, and H.-Y. Kee, Phys. Rev. Lett.

    112, 077204 (2014).

    http://dx.doi.org/10.3390/cryst7050121http://dx.doi.org/10.1143/PTPS.160.155http://dx.doi.org/10.1103/PhysRevLett.102.017205http://dx.doi.org/10.1103/PhysRevLett.102.017205http://dx.doi.org/10.1103/PhysRevLett.112.077204http://dx.doi.org/10.1103/PhysRevLett.112.077204

  • 9

    [5] K. W. Plumb, J. P. Clancy, L. J. Sandilands, V. V. Shankar,Y. F. Hu, K. S. Burch, H.-Y. Kee, and Y.-J. Kim, Phys.Rev. B 90, 041112(R) (2014).

    [6] H.-S. Kim, V. Shankar V., A. Catuneanu, and H.-Y. Kee,Phys. Rev. B 91, 241110(R) (2015).

    [7] A. Kitaev, Ann. Phys. (N. Y.) 321, 2 (2006), JanuarySpecial Issue.

    [8] L. J. Sandilands, Y. Tian, K. W. Plumb, Y.-J. Kim, andK. S. Burch, Phys. Rev. Lett. 114, 147201 (2015).

    [9] J. A. Sears, M. Songvilay, K. W. Plumb, J. P. Clancy,Y. Qiu, Y. Zhao, D. Parshall, and Y.-J. Kim, Phys. Rev.B 91, 144420 (2015).

    [10] R. D. Johnson, S. C. Williams, A. A. Haghighirad, J. Sin-gleton, V. Zapf, P. Manuel, I. I. Mazin, Y. Li, H. O. Jeschke,R. Valent́ı, and R. Coldea, Phys. Rev. B 92, 235119 (2015).

    [11] A. Banerjee, C. A. Bridges, J.-Q. Yan, A. A. Aczel, L. Li,M. B. Stone, G. E. Granroth, M. D. Lumsden, Y. Yiu,J. Knolle, S. Bhattacharjee, D. L. Kovrizhin, R. Moessner,D. A. Tennant, D. G. Mandrus, and S. E. Nagler, Nat.Materials 15, 733 (2016), article.

    [12] H. B. Cao, A. Banerjee, J.-Q. Yan, C. A. Bridges, M. D.Lumsden, D. G. Mandrus, D. A. Tennant, B. C. Chak-oumakos, and S. E. Nagler, Phys. Rev. B 93, 134423(2016).

    [13] H.-S. Kim and H.-Y. Kee, Phys. Rev. B 93, 155143 (2016).[14] L. Janssen, E. C. Andrade, and M. Vojta, Phys. Rev. B

    96, 064430 (2017).[15] R. Yadav, N. A. Bogdanov, V. M. Katukuri, S. Nishimoto,

    J. van den Brink, and L. Hozoi, Sci. Rep. 6, 37925 (2016).[16] S.-H. Baek, S.-H. Do, K.-Y. Choi, Y. S. Kwon, A. U. B.

    Wolter, S. Nishimoto, J. van den Brink, and B. Büchner,Phys. Rev. Lett. 119, 037201 (2017).

    [17] A. U. B. Wolter, L. T. Corredor, L. Janssen, K. Nenkov,S. Schönecker, S.-H. Do, K.-Y. Choi, R. Albrecht,J. Hunger, T. Doert, M. Vojta, and B. Büchner, Phys.Rev. B 96, 041405(R) (2017).

    [18] J. Zheng, K. Ran, T. Li, J. Wang, P. Wang, B. Liu, Z.-X.Liu, B. Normand, J. Wen, and W. Yu, Phys. Rev. Lett.119, 227208 (2017).

    [19] N. Janša, A. Zorko, M. Gomiľsek, M. Pregelj, K. W.Krämer, D. Biner, A. Biffin, C. Rüegg, and M. Klanǰsek,Nat. Phys. 14, 786 (2018).

    [20] Y. Kasahara, T. Ohnishi, Y. Mizukami, O. Tanaka, S. Ma,K. Sugii, N. Kurita, H. Tanaka, J. Nasu, Y. Motome,T. Shibauchi, and Y. Matsuda, Nature 559, 227 (2018).

    [21] M. Yamashita, J. Gouchi, Y. Uwatoko, N. Kurita, andH. Tanaka, Phys. Rev. B 102, 220404(R) (2020).

    [22] G. Baskaran, D. Sen, and R. Shankar, Phys. Rev. B 78,115116 (2008).

    [23] J. Oitmaa, A. Koga, and R. R. P. Singh, Phys. Rev. B 98,214404 (2018).

    [24] A. Koga, H. Tomishige, and J. Nasu, J. Phys. Soc. Jpn.87, 063703 (2018).

    [25] P. P. Stavropoulos, D. Pereira, and H.-Y. Kee, Phys. Rev.Lett. 123, 037203 (2019).

    [26] C. Hickey, C. Berke, P. P. Stavropoulos, H.-Y. Kee, andS. Trebst, Phys. Rev. Research 2, 023361 (2020).

    [27] Z. Zhu, Z.-Y. Weng, and D. N. Sheng, Phys. Rev. Research2, 022047(R) (2020).

    [28] I. Khait, P. P. Stavropoulos, H.-Y. Kee, and Y. B. Kim,arXiv:2001.06000 [cond-mat.str-el].

    [29] Y. Yamaji, Y. Nomura, M. Kurita, R. Arita, andM. Imada, Phys. Rev. Lett. 113, 107201 (2014).

    [30] V. M. Katukuri, S. Nishimoto, V. Yushankhai, A. Stoy-anova, H. Kandpal, S. Choi, R. Coldea, I. Rousochatzakis,L. Hozoi, and J. van den Brink, New J. Phys. 16, 013056(2014).

    [31] W.-B. Zhang, Q. Qu, P. Zhu, and C.-H. Lam, J. Mater.Chem. C 3, 12457 (2015).

    [32] D. Torelli and T. Olsen, 2D Mater. 6, 015028 (2018).[33] S. S. Pershoguba, S. Banerjee, J. C. Lashley, J. Park,

    H. Ågren, G. Aeppli, and A. V. Balatsky, Phys. Rev. X 8,011010 (2018).

    [34] F. Zheng, J. Zhao, Z. Liu, M. Li, M. Zhou, S. Zhang, andP. Zhang, Nanoscale 10, 14298 (2018).

    [35] L. Webster and J.-A. Yan, Phys. Rev. B 98, 144411 (2018).[36] J. Liu, M. Shi, P. Mo, and J. Lu, AIP Adv. 8, 055316

    (2018).[37] M. Wu, Z. Li, T. Cao, and S. G. Louie, Nat. Commun. 10,

    2371 (2019).[38] T. Olsen, MRS Commun. 9, 1142 (2019).[39] O. Besbes, S. Nikolaev, N. Meskini, and I. Solovyev, Phys.

    Rev. B 99, 104432 (2019).[40] V. K. Gudelli and G.-Y. Guo, New J. Phys. 21, 053012

    (2019).[41] C. Xu, J. Feng, M. Kawamura, Y. Yamaji, Y. Nahas,

    S. Prokhorenko, Y. Qi, H. Xiang, and L. Bellaiche, Phys.Rev. Lett. 124, 087205 (2020).

    [42] M. Pizzochero and O. V. Yazyev, J. Phys. Chem. 124, 7585(2020).

    [43] M. Pizzochero, R. Yadav, and O. V. Yazyev, 2D Mater. 7,035005 (2020).

    [44] E. Aguilera, R. Jaeschke-Ubiergo, N. Vidal-Silva, L. E. F.Foa Torres, and A. S. Nunez, Phys. Rev. B 102, 024409(2020).

    [45] D. Soriano, M. I. Katsnelson, and J. Fernández-Rossier,Nano Lett. 20, 6225 (2020).

    [46] L. L. Handy and N. W. Gregory, J. Am. Chem. Soc. 74,891 (1952).

    [47] J. F. Dillon and C. E. Olson, J. Appl. Phys. 36, 1259 (1965).[48] M. A. McGuire, H. Dixit, V. R. Cooper, and B. C. Sales,

    Chem. Mater. 27, 612 (2015).[49] B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein,

    R. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A.McGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo-Herrero, and X. Xu, Nature 546, 270 (2017).

    [50] N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133(1966).

    [51] J. L. Lado and J. Fernández-Rossier, 2D Mater. 4, 035002(2017).

    [52] D.-H. Kim, K. Kim, K.-T. Ko, J. Seo, J. S. Kim, T.-H.Jang, Y. Kim, J.-Y. Kim, S.-W. Cheong, and J.-H. Park,Phys. Rev. Lett. 122, 207201 (2019).

    [53] C. Xu, J. Feng, H. Xiang, and L. Bellaiche, npj Comput.Materials 4, 57 (2018).

    [54] I. Lee, F. G. Utermohlen, D. Weber, K. Hwang, C. Zhang,J. van Tol, J. E. Goldberger, N. Trivedi, and P. C. Hammel,Phys. Rev. Lett. 124, 017201 (2020).

    [55] J. G. Rau and H.-Y. Kee, arXiv:1408.4811 [cond-mat.str-el].

    [56] J. Kanamori, Prog. Theor. Phys. 30, 275 (1963).[57] I. V. Kashin, V. V. Mazurenko, M. I. Katsnelson, and

    A. N. Rudenko, 2D Mater. 7, 025036 (2020).[58] G. Jackeli and A. Avella, Phys. Rev. B 92, 184416 (2015).[59] J. Chaloupka and G. Khaliullin, Phys. Rev. B 94, 064435

    (2016).[60] D. R. Klein, D. MacNeill, J. L. Lado, D. Sori-

    ano, E. Navarro-Moratalla, K. Watanabe, T. Taniguchi,S. Manni, P. Canfield, J. Fernández-Rossier, and P. Jarillo-Herrero, Science 360, 1218 (2018).

    [61] T. Song, X. Cai, M. W.-Y. Tu, X. Zhang, B. Huang, N. P.Wilson, K. L. Seyler, L. Zhu, T. Taniguchi, K. Watanabe,M. A. McGuire, D. H. Cobden, D. Xiao, W. Yao, andX. Xu, Science 360, 1214 (2018).

    http://dx.doi.org/ 10.1103/PhysRevB.90.041112http://dx.doi.org/ 10.1103/PhysRevB.90.041112http://dx.doi.org/10.1103/PhysRevB.91.241110http://dx.doi.org/https://doi.org/10.1016/j.aop.2005.10.005http://dx.doi.org/ 10.1103/PhysRevLett.114.147201http://dx.doi.org/10.1103/PhysRevB.91.144420http://dx.doi.org/10.1103/PhysRevB.91.144420http://dx.doi.org/10.1103/PhysRevB.92.235119http://dx.doi.org/10.1038/nmat4604http://dx.doi.org/10.1038/nmat4604http://dx.doi.org/ 10.1103/PhysRevB.93.134423http://dx.doi.org/ 10.1103/PhysRevB.93.134423http://dx.doi.org/10.1103/PhysRevB.93.155143http://dx.doi.org/10.1103/PhysRevB.96.064430http://dx.doi.org/10.1103/PhysRevB.96.064430http://dx.doi.org/ https://doi.org/10.1038/srep37925http://dx.doi.org/ 10.1103/PhysRevLett.119.037201http://dx.doi.org/10.1103/PhysRevB.96.041405http://dx.doi.org/10.1103/PhysRevB.96.041405http://dx.doi.org/ 10.1103/PhysRevLett.119.227208http://dx.doi.org/ 10.1103/PhysRevLett.119.227208http://dx.doi.org/ 10.1038/s41567-018-0129-5http://dx.doi.org/10.1038/s41586-018-0274-0http://dx.doi.org/ 10.1103/PhysRevB.102.220404http://dx.doi.org/10.1103/PhysRevB.78.115116http://dx.doi.org/10.1103/PhysRevB.78.115116http://dx.doi.org/10.1103/PhysRevB.98.214404http://dx.doi.org/10.1103/PhysRevB.98.214404http://dx.doi.org/10.7566/JPSJ.87.063703http://dx.doi.org/10.7566/JPSJ.87.063703http://dx.doi.org/10.1103/PhysRevLett.123.037203http://dx.doi.org/10.1103/PhysRevLett.123.037203http://dx.doi.org/10.1103/PhysRevResearch.2.023361http://dx.doi.org/10.1103/PhysRevResearch.2.022047http://dx.doi.org/10.1103/PhysRevResearch.2.022047http://arxiv.org/abs/2001.06000http://dx.doi.org/ 10.1103/PhysRevLett.113.107201http://dx.doi.org/10.1088/1367-2630/16/1/013056http://dx.doi.org/10.1088/1367-2630/16/1/013056http://dx.doi.org/ 10.1039/C5TC02840Jhttp://dx.doi.org/ 10.1039/C5TC02840Jhttp://dx.doi.org/10.1088/2053-1583/aaf06dhttp://dx.doi.org/ 10.1103/PhysRevX.8.011010http://dx.doi.org/ 10.1103/PhysRevX.8.011010http://dx.doi.org/10.1039/C8NR03230Khttp://dx.doi.org/10.1103/PhysRevB.98.144411http://dx.doi.org/10.1063/1.5030441http://dx.doi.org/10.1063/1.5030441http://dx.doi.org/ 10.1038/s41467-019-10325-7http://dx.doi.org/ 10.1038/s41467-019-10325-7http://dx.doi.org/10.1557/mrc.2019.117http://dx.doi.org/10.1103/PhysRevB.99.104432http://dx.doi.org/10.1103/PhysRevB.99.104432http://dx.doi.org/10.1088/1367-2630/ab1ae9http://dx.doi.org/10.1088/1367-2630/ab1ae9http://dx.doi.org/10.1103/PhysRevLett.124.087205http://dx.doi.org/10.1103/PhysRevLett.124.087205http://dx.doi.org/10.1021/acs.jpcc.0c01873http://dx.doi.org/10.1021/acs.jpcc.0c01873http://dx.doi.org/10.1088/2053-1583/ab7cabhttp://dx.doi.org/10.1088/2053-1583/ab7cabhttp://dx.doi.org/10.1103/PhysRevB.102.024409http://dx.doi.org/10.1103/PhysRevB.102.024409http://dx.doi.org/10.1021/acs.nanolett.0c02381http://dx.doi.org/10.1021/ja01124a009http://dx.doi.org/10.1021/ja01124a009http://dx.doi.org/10.1063/1.1714194http://dx.doi.org/10.1021/cm504242thttp://dx.doi.org/ 10.1038/nature22391http://dx.doi.org/10.1103/PhysRevLett.17.1133http://dx.doi.org/10.1103/PhysRevLett.17.1133http://dx.doi.org/10.1088/2053-1583/aa75edhttp://dx.doi.org/10.1088/2053-1583/aa75edhttp://dx.doi.org/10.1103/PhysRevLett.122.207201http://dx.doi.org/ 10.1038/s41524-018-0115-6http://dx.doi.org/ 10.1038/s41524-018-0115-6http://dx.doi.org/ 10.1103/PhysRevLett.124.017201http://arxiv.org/abs/1408.4811http://arxiv.org/abs/1408.4811http://dx.doi.org/10.1143/PTP.30.275http://dx.doi.org/10.1088/2053-1583/ab72d8http://dx.doi.org/10.1103/PhysRevB.92.184416http://dx.doi.org/10.1103/PhysRevB.94.064435http://dx.doi.org/10.1103/PhysRevB.94.064435http://dx.doi.org/ 10.1126/science.aar3617http://dx.doi.org/10.1126/science.aar4851

  • 10

    [62] D. R. Klein, D. MacNeill, Q. Song, D. T. Larson, S. Fang,M. Xu, R. A. Ribeiro, P. C. Canfield, E. Kaxiras, R. Comin,and P. Jarillo-Herrero, Nat. Phys. 15, 1255 (2019).

    [63] H. H. Kim, B. Yang, S. Li, S. Jiang, C. Jin, Z. Tao,G. Nichols, F. Sfigakis, S. Zhong, C. Li, S. Tian, D. G.Cory, G.-X. Miao, J. Shan, K. F. Mak, H. Lei, K. Sun,L. Zhao, and A. W. Tsen, Proc. Natl. Acad. Sci. U. S. A.116, 11131 (2019).

    [64] G. Kresse and J. Hafner, Phy. Rev. B 47, 558(R) (1993).[65] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.

    Lett. 77, 3865 (1997).[66] A. A. Mostofi, J. R. Yates, Y.-S. Lee, I. Souza, D. Vander-

    bilt, and N. Marzari, Comput. Phys. Commun. 178, 685(2008).

    [67] S. W. Jang, M. Y. Jeong, H. Yoon, S. Ryee, and M. J.Han, Phys. Rev. Mater. 3, 031001(R) (2019).

    [68] N. Sivadas, S. Okamoto, X. Xu, C. J. Fennie, and D. Xiao,Nano Lett. 18, 7658 (2018).

    [69] P. Jiang, C. Wang, D. Chen, Z. Zhong, Z. Yuan, Z.-Y. Lu,and W. Ji, Phys. Rev. B 99, 144401 (2019).

    [70] D. Soriano, C. Cardoso, and J. Fernández-Rossier, SolidState Commun. 299, 113662 (2019).

    [71] L. Chen, J.-H. Chung, B. Gao, T. Chen, M. B. Stone, A. I.Kolesnikov, Q. Huang, and P. Dai, Phys. Rev. X 8, 041028(2018).

    [72] J. Cenker, B. Huang, N. Suri, P. Thijssen, A. Miller,T. Song, T. Taniguchi, K. Watanabe, M. A. McGuire,D. Xiao, and X. Xu, Nat. Phys. 17, 20 (2021).

    [73] P. A. Maksimov and A. L. Chernyshev, Phys. Rev. Research2, 033011 (2020).

    [74] L. Chen, J.-H. Chung, T. Chen, C. Duan, A. Schneidewind,I. Radelytskyi, D. J. Voneshen, R. A. Ewings, M. B. Stone,A. I. Kolesnikov, B. Winn, S. Chi, R. A. Mole, D. H. Yu,B. Gao, and P. Dai, Phys. Rev. B 101, 134418 (2020).

    [75] J. C. Slater and G. F. Koster, Phys. Rev. 94, 1498 (1954).[76] T. Holstein and H. Primakoff, Phys. Rev. 58, 1098 (1940).[77] J. Colpa, Physica A 93, 327 (1978).

    APPENDIX

    1. Energy levels of on-site Kanamori Hamiltonian

    To obtain the n.n. spin interaction, we used the sec-ond order perturbation theory, where the dominant inter-action is Eq. (1). First we note that in the lowest energystate there are three electrons at each metal site M. Theexchange processes then involve one electron hopping be-tween M sites. Thus the intermediate states have two elec-trons in t2g-orbitals on one M site and four electrons ineither t2g- or eg-orbitals on the other M site. On the otherhand the single-ion anisotropy would follow from a singleM site, where the three electrons in the ground state inter-act with an exited three electron state. The energy levelsof all states involved in these exchange processes are givenin Table IV.

    2. Ideal hoppings

    In the ideal octahedron environment the allowed indirecthoppings are show in Fig. 5. On the M1−X1 bond we have

    Degen. Energy Microscopics

    2 electrons (only in t2g)

    1 EA,2,1 = U + 2JH + 2�d

    2 EA,2,2 = U − JH + 2�d3 EA,2,3 = U

    ′ + JH + 2�d

    9 EA,2,4 = U′ − JH + 2�d

    3 electrons (GS)

    4 EA,3,gs = 3U′ − 3JH + 3�d

    3 electrons (only in t2g)

    4 EA,3,1 = 3U′ + 3�d

    6 EA,3,2 = U + 2U′ − 2JH + 3�d

    6 EA,3,3 = U + 2U′ + 3�d

    3 electrons (2-t2g, 1-eg)

    8 EA,3,e1 = U + 2U′ − 2JH + ∆c + 3�d

    4 EA,3,e2 = U + 2U′ + JH + ∆c + 3�d

    24 EA,3,e3 = 3U′ − 3JH + ∆c + 3�d

    24 EA,3,e4 = 3U′ + ∆c + 3�d

    4 electrons (only in t2g)

    1 EA,4,1 = 2U + 4U′ + 4�d

    2 EA,4,2 = 2U + 4U′ − 3JH + 4�d

    3 EA,4,3 = U + 5U′ − JH + 4�d

    9 EA,4,4 = U + 5U′ − 3JH + 4�d

    4 electrons (3-t2g, 1-eg)

    4 EA,4,e1 = 6U′ + ∆c + 4�d

    6 EA,4,e2 = U + 5U′ + ∆c + 4�d

    10 EA,4,e3 = 6U′ − 6JH + ∆c + 4�d

    18 EA,4,e4 = U + 5U′ − 4JH + ∆c + 4�d

    18 EA,4,e5 = 6U′ − 2JH + ∆c + 4�d

    24 EA,4,e6 = U + 5U′ − 2JH + ∆c + 4�d

    TABLE IV. Spectrum of Hamiltonian Eq. (1). Listed states cancontribute to the second order strong coupling perturbation.

    the hopping matrix

    px py pz

    TM1X1 =

    t1 0 0−t2 0 00 0 00 0 t00 t0 0

    dx2−y2d3z2−r2dyzdxzdxy ,

    (A.1)

    The other indirect bonds TM2X2 , TM2X1 , TM1X2 are ob-tained from TM1X2 by consecutively applying the sym-metry operation of the octahedron C4(0, 0, 1). Use ofEq. (A.1) and symmetry related bonds are then used inEq. (4) to arrive at effective hoppings Eq. (7) and Eq. (9).Slater-Koster analysis[75] allows us to decompose t0, t1, t2and t3 into σ- and π-bonding integrals between d- and p-

    http://dx.doi.org/10.1038/s41567-019-0651-0http://dx.doi.org/10.1073/pnas.1902100116http://dx.doi.org/10.1073/pnas.1902100116http://dx.doi.org/10.1103/PhysRevB.47.558http://dx.doi.org/10.1103/PhysRevLett.77.3865http://dx.doi.org/10.1103/PhysRevLett.77.3865http://dx.doi.org/ 10.1016/j.cpc.2007.11.016http://dx.doi.org/ 10.1016/j.cpc.2007.11.016http://dx.doi.org/ 10.1103/PhysRevMaterials.3.031001http://dx.doi.org/ 10.1021/acs.nanolett.8b03321http://dx.doi.org/10.1103/PhysRevB.99.144401http://dx.doi.org/ https://doi.org/10.1016/j.ssc.2019.113662http://dx.doi.org/ https://doi.org/10.1016/j.ssc.2019.113662http://dx.doi.org/10.1103/PhysRevX.8.041028http://dx.doi.org/10.1103/PhysRevX.8.041028http://dx.doi.org/ 10.1038/s41567-020-0999-1http://dx.doi.org/10.1103/PhysRevResearch.2.033011http://dx.doi.org/10.1103/PhysRevResearch.2.033011http://dx.doi.org/10.1103/PhysRevB.101.134418http://dx.doi.org/10.1103/PhysRev.94.1498http://dx.doi.org/10.1103/PhysRev.58.1098http://dx.doi.org/https://doi.org/10.1016/0378-4371(78)90160-7

  • 11

    Figure 5. Indirect hoppings from X site to M site, in the globalcoordinates xyz.

    orbitals

    t0 = tpdπ, t1 =

    √3tpdσ2

    , t2 =tpdσ

    2, t3 = tpdσ,

    tpdσ < 0, tpdπ > 0, |tpdπ| < |tpdσ| .(A.2)

    The direct hoppings Eq. (11) include δ-bonding in additionto σ- and π-bondings

    td1 =tddπ + tddδ

    2, td2 =

    −tddπ + tddδ2

    td3 =3tddσ + tddδ

    4, t̃d0 =

    √3 (tddσ − tddδ)

    4,

    tddσ < 0, tddπ > 0, tddδ < 0,

    |tddδ| < |tddπ| < |tddσ| ,

    (A.3)

    with the Anderse prediction tddσ : tddπ : tddδ = −6 : 4 : −1.

    3. Positions of X sites under distortions

    In the R3̄ space group, octahedra made of two trianglesas shown in Fig. 2 are rotated around ĉ-axis denoted by δxblue arrows. Furthermore there is a slight change in thedistance between these two triangles denoted by δx′ yellowarrows. The new positions of ligands are parameterizedby dimensionless δx and δx′, in units of the distance dMbetween M1 and M2. One can track the positions rm ofXm, as it is related to the undistorted position ro,m, δxand δx′:

    rm = ro,m + δx dM (ξam â + ξ

    bm b̂) + δx

    ′ dMζm ĉ

    m ro,m(ξam, ξ

    bm

    )ζm

    X1 (−dM√

    20 0 ) ( −1 , −1 ) 1

    X2 ( 0dM√

    20 ) ( 1 , 1 ) −1

    X3 ( 0−dM√

    20 ) ( 1 , 0 ) 1

    X4 ( 0 0dM√

    2) ( −1 , 0 ) −1

    X5 ( 0 0−dM√

    2) ( 0 , 1 ) 1

    X6 (dM√

    20 0 ) ( 0 , −1 ) −1

    ,

    (A.4)

    where Xm lable sites as in Fig. 2(a) and â, b̂ and ĉ are unitvectors along the lattice vectors of Fig. 2(b).

    4. Distortion-induced hoppings

    To get the distortion-induced hoppings one needs to findthe directional cosines of the indirect d−p bonds using theligand positions Eq. (A.4). With the metal site located atrM1 and ligand Xm located at rm, the required directionalcosines follow from rm − rM1 . They are the inputs in theSlater-Koster formula for rotated bonds Ref. [75]. Here weonly show the leading term in each matrix element gen-erated from the above procedure. Note that all elementsbecome finite, and the distortion-induced hopping betweent2g and p-oribtials are denoted by δti and eg and p-orbitalsby δt′i:

    TM1X1 =

    t1 δt

    ′1 δt

    ′2

    −t2 δt′3 δt′4δt1 δt2 δt3δt4 δt1 t0δt5 t0 δt1

    , (A.5)

    where X1 labels the site as in Fig. 2(a). The dis-torted octahedron realizes the D3 point group, which con-tains C3(1, 1, 1) and C

    ′2(−1, 1, 0) rotations. The hoppings

    TM1X3 and TM1X5 are recovered by applying C3(1, 1, 1)to TM1X1 . Further TM1X1 relates to TM1X2 by aC ′2(−1, 1, 0). Finally TM1X2 relates to TM1X4 and TM1X6by C3(1, 1, 1). The direct hopping integrals denoted byTM1M2 is same the as Eq. (11). Making use of Eq. (A.5)we derive effective hoppings TeffM1M2 following the methodin the main text Eq. (4). The distortion induced effectivehoppings to leading order in δt and listed in Table V.

    5. Single-ion anisotropy term

    Due to the distortion and SOC, the single-ion anisotropicterm is also generated via the hopping to anions which canhop back to create spin-dependent on-site terms denotedby TeffM1M1 = T

    effM2M2

    . Without the trigonal distortion,the effective hopping integrals between t2g at M1 are given

  • 12

    TABLE V. Effective hoppings to leading order in distortion-induced hoppings, where δt are defined in Eq. (A.5). Distortion inducedhoppings are grouped in: δta = δt1 − δt4, δtb = 2δt1 + δt4, δtc = δt3 + δt5, δtd = δt1 − δt2, δte = δt1 + δt4, δtf = δt1 + δt2.

    TeffM1M2(t2g ⊗ t2g) = teff

    TA TC TDT †C TA UTDU†T †D U

    †T †DU TB

    TeffM1M2(eg ⊗ t2g) = teff ( T̃A −UT̃AU† T̃CT̃B UT̃BU† T̃D)

    U = eiπ 1√

    2(σx2 −

    σy2 ) =

    i√2

    (σx − σy)

    TA =

    (2δt3t0

    +td1teff

    )σo

    TB =

    (4δt5t0

    +td3teff

    )σo

    TC =

    (1 +

    td2teff

    )σo + i

    r

    2

    δtat0

    (σx + σy)

    TD =δtbt0σo + i

    r

    2

    (σx −

    δtct0σy −

    δtdt0σz

    )

    T̃A =

    (δt′2t0

    +δtdt1t20

    )σo + i

    r

    2

    ((δt3t1t20− δt

    ′1

    t0

    )σx +

    t1t0σy −

    δtet1t20

    σz

    )T̃B =

    (δt′4t0− δtf t2

    t20

    )σo + i

    r

    2

    ((δt3t2t20− δt

    ′3

    t0

    )σx −

    t2t0σy −

    δtat2t20

    σz

    )T̃C = i

    r

    2

    ((δt1t1t20− δt

    ′2

    t0

    )(σx + σy)− 2

    (δt5t1t20− δt

    ′1

    t0

    )σz

    )T̃D =

    (−2 t2

    t0+t̃d0teff

    )σo + i

    r

    2

    (δt1t2t20

    +δt′4t0

    )(σx − σy) .

    TABLE VI. Spin model terms, under distortions, to leading order in δt. Distortion-induced δt defined in Eq. (A.5), and charactersubscripted δt defined in caption of Table V and Eq. (A.8)

    t2g ⊗ t2g eg ⊗ t2g

    J =4(2t2d1 + 2(teff + td2)

    2 + t2d3)

    9 (U + 2JH)−

    4JH(2teff(t2/t0)− t̃d0

    )23 (∆c + U ′ − JH) (∆c + U ′ + 3JH)

    +32teff

    9(U + 2JH)

    (td1

    δt3t0

    + td3δt5t0

    )K = − 4 (r teff)

    2

    9 (U + 2JH)

    2 JH(r teff)2

    3 (∆c + U ′ − JH) (∆c + U ′ + 3JH)t21 + t

    22

    t20

    Γ = − 8(rteff)2

    9 (U + 2JH)

    (δtct0

    )− 4JH(rteff)

    2

    3 (∆c + U ′ − JH) (∆c + U ′ + 3JH)

    (t1(t1δt3 − t0δt′1)

    t30− t2(t2δt3 − t0δt

    ′3)

    t30

    )Γ′ = − 4(rteff)

    2

    9 (U + 2JH)

    (δtdt0

    )− 2JH(rteff)

    2

    3 (∆c + U ′ − JH) (∆c + U ′ + 3JH)

    (− t

    21δtet30

    +t22δtat30

    )Ac = −

    4(rteff)2

    (U + 3JH − U ′)

    (δtαt0

    )− 4JH(rteff)

    2

    ∆c (∆c + 3JH)

    (t1t0

    (δtβt1t20

    +δt′αt0

    )− t2 + t3

    t0

    (δtβt2 + 2δt1t3

    t20+δt′βt0

    ))

    by

    TeffM1M1(t2g ⊗ t2g) =

    teff

    4σo irσz −irσy−irσz 4σo irσxirσy −irσx 4σo

    . (A.6)Similarly, effective hopping between t2g and eg is found as

    TeffM1M1(eg ⊗ t2g) =

    teff

    irt1t0σx ir

    t1t0σy −i2r

    t1t0σz

    irt2 + t3t0

    σx −irt2 + t3t0

    σy 02×2

    .(A.7)

    They lead to an Si ·Si term, which is just equal to S(S+1),an irrelevant constant to the spin model. Thus there is no

    single-ion anisotropy without the trigonal distortion. Withthe trigonal distortion, the single-ion anisotropy along ĉ-axis is generated:

    Ac = −4(rteff)

    2

    (U + 3JH − U ′)

    (δtαt0

    )− 4JH(rteff)

    2

    ∆c (∆c + 3JH)

    (t1t0

    (δtβt1t20

    +δt′αt0

    )− t2 + t3

    t0

    (δtβt2 + 2δt1t3

    t20+δt′βt0

    )),

    (A.8)

    where δtα = −2δt1 + δt2 + δt3 + δt4 + δt5, δtβ = δt2 + δt3,

    δt′α = −δt′1 +√

    3δt′4 − δt′22

    , and δt′β = −δt′3 +√

    3δt′2 + δt′4

    2.

  • 13

    6. Spin interaction with trigonal distortion

    Putting them all together, we have Heisenberg, Kitaev,Γ, Γ′, and single-ion anisotropy Ac term. Their form toleading order in distortion induced hoppings δt are summa-rized in Table VI bellow. Note that the Kitaev interactiondoes not have any linear term of distortion, and Γ, Γ′ andsingle-ion anisotropy Ac along ĉ-axis are finite due to bothSOC and distortion.

    7. Spin Wave Theory

    Figure 6. X Y and Z bonds of the J −K − Γ− Γ′ − Ac shownin red, green and blue respectively. Second n.n are shown inorange. The orange arrows indicate when sgn(ij) = +1 in theDM term.

    We consider the J−K−Γ−Γ′−Ac as well as the secondn.n. DM term (Dc)

    H =∑

    〈i,j〉∈αβ(γ)

    [JSi · Sj +KSγi S

    γj + Γ

    (Sαi S

    βj + S

    βi S

    αj

    )+Γ′

    (Sαi S

    γj + S

    βi S

    γj + S

    γi S

    αj + S

    γi S

    βj

    ) ]+∑〈〈i,j〉〉

    Dc · (Si × Sj) +∑i

    Ac (Si · ĉ)2 ,

    (A.9)where Dc = Dc sgn(ij)ĉ and sgn(ij) = +1 when i to jpoints along the orange arrows in Fig. 6. The standardHolstein-Primakoff transformation [76] expanded to linearorder in S read

    S+ =√

    2S

    (1− a

    †a

    2S

    ) 12

    a '√

    2Sa,

    S− =√

    2Sa†(

    1− a†a

    2S

    ) 12

    '√

    2Sa†,

    Sz = S − a†a.

    (A.10)

    Using the above and Fourier transforming Eq. (A.9) leadsto

    H = ECL +∑

    k∈BZx†khkxk,

    x†k = (a†k, b

    †k, a−k, b−k),

    hk =

    ho−(k) h1(k) 0 h2(k)h1(k)∗ ho+(k) h2(−k) 00 h2(−k)∗ ho+(k) h1(k)h2(k)

    ∗ 0 h1(k)∗ ho−(k)

    ,(A.11)

    where the two species of bosons ak and bk correspond tothe two sublattices of the unit cell. The h(k) terms are

    ho±(k) = ho ± hDM (k),

    ho = −S (2Ac + 2Γ + 4Γ′ + 3J +K) ,

    hDM (k) = 2SDc (sin(a · k) + sin(b · k)− sin((a + b) · k)) ,

    h1(k) = −S (Γ + 2Γ′ − 3J −K)

    3(1 + e−ia·k + eib·k),

    h2(k) =S (2Γ− 2Γ′ +K)

    6

    ((1− i

    √3)e−ia·k

    +(1 + i√

    3)eib·k − 2),(A.12)

    where a and b are the lattice vectors in Fig. 6. Followingstandard methods of diagonalizing BdG Hamiltonians[77]we find the lowest eigenvalue around the Γ point in the BZ,and upon expanding to orders of k, we get the spin gap ωoand spin stiffness ρ:

    ωk = ωo + ρk2

    ωo = S |3Γ + 6Γ′ + 2Ac|

    ρ =S

    12

    ∣∣∣∣∣3J +K − Γ− 2Γ′− (K + 2Γ− 2Γ

    ′)2

    2 (2Γ + 4Γ′ + 2Ac + 3J +K)

    ∣∣∣∣∣.(A.13)

    At the K point in the BZ the Dirac gap is ωK+ − ωK− where

    ωK+ = S{

    (6Γ′ + 2Ac + 3J) (4Γ + 2Γ′ + 2Ac + 3J + 2K)

    +6√

    3Dc (2Γ + 4Γ′ + 2Ac + 3J +K) + 27D

    2c

    }1/2,

    ωK− = S∣∣2Γ + 4Γ′ + 2Ac + 3J +K − 3√3Dc∣∣ .

    (A.14)

    8. Exact diagonalization calculations

    The moment pinning calculations are carried out by find-ing the ground state from exact diagonalization on an 8 site(2x2) honeycomb cluster with periodic conditions, shownin Fig. 7.

  • 14

    Figure 7. X,Y and Z bonds show in red, green and blue respec-tively. Dashed bonds represent the periodic boundary condi-tions.

    Magnetic Anisotropy in Spin-3/2 with Heavy Ligand in Honeycomb Mott Insulators: Application to CrI3AbstractI IntroductionII Kanamori interaction and tight binding HamiltonianIII Ideal honeycomb structureA Superexchange path: indirect hopping1 t2g-t2g contributions2 eg-t2g contributions

    B Direct hoppingC Summary and Comments

    IV Effects of distortion: distorted octahderaV Application to CrI3VI Magnetic AnisotropyVII Spin gap and finite transition temperatureVIII Summary and Discussion Acknowledgments References APPENDIX1 Energy levels of on-site Kanamori Hamiltonian2 Ideal hoppings3 Positions of X sites under distortions4 Distortion-induced hoppings5 Single-ion anisotropy term6 Spin interaction with trigonal distortion7 Spin Wave Theory8 Exact diagonalization calculations