Skip to main content
NIST Author Manuscripts logoLink to NIST Author Manuscripts
. Author manuscript; available in PMC: 2024 Jun 24.
Published in final edited form as: Phys Rev X. 2018 Dec;8(4):10.1103/PhysRevX.8.041028. doi: 10.1103/PhysRevX.8.041028

Topological Spin Excitations in Honeycomb Ferromagnet CrI3

Lebing Chen 1, Jae-Ho Chung 2,1,*, Bin Gao 1, Tong Chen 1, Matthew B Stone 3, Alexander I Kolesnikov 3, Qingzhen Huang 4, Pengcheng Dai 1,
PMCID: PMC11194793  NIHMSID: NIHMS1588415  PMID: 38915421

Abstract

In two-dimensional honeycomb ferromagnets, bosonic magnon quasiparticles (spin waves) may either behave as massless Dirac fermions or form topologically protected edge states. The key ingredient defining their nature is the next-nearest-neighbor Dzyaloshinskii-Moriya interaction that breaks the inversion symmetry of the lattice and discriminates chirality of the associated spin-wave excitations. Using inelastic neutron scattering, we find that spin waves of the insulating honeycomb ferromagnet CrI3 (TC=61K) have two distinctive bands of ferromagnetic excitations separated by a ∼4 meV gap at the Dirac points. These results can only be understood by considering a Heisenberg Hamiltonian with Dzyaloshinskii-Moriya interaction, thus providing experimental evidence that spin waves in CrI3 can have robust topological properties potentially useful for dissipationless spintronic applications.

Keywords: Condensed Matter Physics, Magnetism, Topological Insulators

I. INTRODUCTION

When quantum particles such as electrons are confined in two-dimensional (2D) geometry, the reduced lattice dimensions and particle interactions can drive the system into novel behavior such as the quantum Hall state under a large magnetic field perpendicular to the 2D electron gas [1]. In 1988, Haldane showed that some solid-state systems, e.g., the 2D honeycomb lattice, can also have quantum Hall state without having to apply magnetic fields due to their inherently topological band structure [2]. It is well known that materials with strong spin-orbit coupling can host topological band structures [3,4]. For 2D honeycomb and kagome lattices, a diverse range of novel electronic band properties and magnetism have been observed [5,6]. For instance, a graphene as the simplest honeycomb exhibits linear electronic dispersions near the Fermi surface allowing exotic massless Dirac fermions to appear [7,8]. Such band structure is built upon the two equivalent and interconnected triangular sublattices, which result in topological band crossing at the Fermi surface. [See Figs. 1(a) and 1(b) for the real and reciprocal spaces, respectively, of the honeycomb lattice.]

FIG. 1.

FIG. 1.

(a) Top view of the honeycomb lattice and the out-of-plane components of the antisymmetric DM vectors. The triangular arrows mark the second nearest-neighbor bond orientations that share the common sign of DM vectors. (b) The BZ zone boundaries and the high symmetry points on the (h,k) plane. (c) Crystal and magnetic structures of CrI3 with the Heisenberg exchange paths. Iodine sites are not displayed for simplicity. The dotted lines denote the hexagonal unit cell boundary. (d) Inelastic neutron scattering results of spin waves with Ei=50meV along the high symmetry directions in the (h,k) plane, which are observed at 5 K and integrated over 5l5. The superimposed solid lines are the calculations by the Heisenberg-only model. (e) Low-energy spin-wave mode along the l direction (Ei=8meV). Note that the minima appear at l=3n in the hexagonal setting. The data are normalized to the absolute units of mbarn/meV/Sr/f.u. (f.u. is formula unit) by a vanadium standard.

Topological band structures are not unique to systems with odd half-integer spin electronlike quasiparticles (fermions) like graphene. In fact, many systems with integer spin quasiparticles (bosons) can also have topological band structures. For example, topological photon modes have been realized in photonic crystals [9,10]. In addition, anomalous thermal Hall effects from topological magnon band structures have been predicted in insulating quantum magnets [11] and observed in an insulating collinear ferromagnet Lu2V2O7 with a pyrochlore structure [12]. Theoretically, several classes of ferromagnetic insulators have been predicted to have interesting topological properties [1317]. In the case of 2D honeycomb ferromagnets with two magnetic atoms per unit cell, magnetic versions of Dirac particles have been predicted [1820]. The magnon (spin-wave) band structure of these ferromagnets is essentially identical to the electronic counterpart of graphene with two modes, acoustic and optical spin waves, for each state reflecting two sublattices. If the spins interact only via the Heisenberg exchange couplings, the two spin-wave modes will cross with each other at K/K points at the corner of the Brillouin zone (BZ) boundary and form Dirac cones with linear dispersion [1820]. The presence of these Dirac points are robust against finite next-nearest-neighbor exchanges, which will only shift positions of the Dirac points. Such spin-wave bands have experimentally been observed in 2D ferromagnets CrBr3 and Cr2Si2Te6 [21,22], thus confirming the presence of nondegenerate band-touching (Dirac) points in the magnon excitation spectrum and leading to a massless Dirac Hamiltonian [1820]. Similar magnon band crossings have also been observed in the three-dimensional (3D) antiferromagnet Cu3TeO6 [23,24].

In the case of graphene, a finite spin-orbit coupling produces a small bulk semiconducting gap (approximately 1 μeV), leaving only the edge states to be truly conducting at absolute zero temperature [3,25]. A magnetic version of such topological edge states may also be realized if strong spin-orbit coupled antisymmetric Dzyaloshinkii-Moriya (DM) exchange opens a gap at the spin-wave crossing Dirac points [2628]. The DM interactions are known to act as effective vector potential leading to anomalous magnon Hall effect that facilitates topological edge transports [11,12,29]. In contrast to electron spin current where dissipation can be large due to Ohmic heating, noninteracting topological magnons, which are quantized spin-1 excitations from an ordered magnetic ground state, are uncharged and can in principle propagate for a long time without dissipation [3033]. Since the DM interaction will cancel out upon space inversion, a finite DM term may appear only between the next-nearest neighbors on the honeycomb lattices [see Fig. 1(a)]. Whereas the possible orientations of these DM vectors may depend on local symmetries [34], only the term collinear with magnetic moments can induce the Dirac gap. Such DM-induced topological magnons have been predicted [13,35] and observed in 2D kagome ferromagnet compound Cu[1,3-benzenedicarboxylate(bdc)] [Cu(1,3-bdc)], where an out-of-plane external magnetic field applied to align the in-plane ferromagnetic ordered moments along the c axis is found to also induce a spin gap at the Dirac points [36,37].

In honeycomb ferromagnets, it is unclear whether such topological magnons can exist. In fact, the topology of the next-nearest-neighbor bonds on a honeycomb lattice is equivalent to the nearest-neighbor bonds of a kagome lattice. Recent experimental discoveries of intrinsic 2D ferromagnetism in van der Waals materials suggest that the topological spin excitations will probably be more robust in the honeycomb lattices [3840].

In this work, we use inelastic neutron scattering to map out energy and wave-vector dependence of spin-wave excitations in CrI3, one of the honeycomb ferromagnets where topological Dirac magnons are predicted to appear [18,26]. The honeycomb lattice, shown in Fig. 1(c), is essentially identical to those of another chromium trihalide CrBr3, in which spin-wave excitations have long been known [21]. The magnetism in CrI3 is commonly ascribed to Cr3+ ions surrounded by I6 octahedra, forming a 2D honeycomb network [Fig. 1(c)]. The CrI3 layers are stacked against each other by van der Waals interaction, and have a monoclinic crystal structure at room temperature. Upon cooling, the monoclinic crystal structure transforms to the rhombohedral structure (space group: R3¯) over a wide temperature range (100–220 K) via lateral sliding of the CrI3 planes with hysteresis [41]. At Curie temperature TC=61K, ferromagnetic ordering appears with Cr3+ spins oriented along the c axis [Fig. 1(c)] [42]. This is different from the in-plane moment of Cu(1,3-bdc) at zero field, providing the necessary condition for DM interactions to open a gap in CrI3 without the need for an external magnetic field [13,35]. Since CrI3 has similar structural and ferromagnetic transitions as that of CrBr3, albeit at different temperatures, one would expect that spin-wave excitations of CrI3 should be similar to that of CrBr3, which have Dirac points at the acoustic and optical spin-wave crossing points [21]. Surprisingly, we find that spin waves in CrI3 exhibit remarkably large gaps at the Dirac points, thus providing direct evidence for the presence of DM interactions in CrI3 [26]. Therefore, CrI3 is an insulating ferromagnet that can potentially host topological edge magnons predicted by the theories [2628].

II. RESULTS AND DISCUSSIONS

Thin single-crystal platelets of CrI3 were grown by the chemical vapor transport method using I2 as the transport agent [41]. The grown crystals are typically 1 cm by 1 cm in area and extremely thin and fragile [42]. Our results are reported using a honeycomb structure with in-plane Cr-Cr distance of approximately 3.96 Å and c-axis layer spacing of 6.62 Å in the low temperature rhombohedral structure [Fig. 1(c)] [43]. The in-plane momentum transfer q=ha*+kb* is denoted as q=(h,k) in hexagonal reciprocal lattice units (r.l.u.), as shown in Fig. 1(b). The temperature-dependent magnetization and neutron powder diffraction measurements confirmed that the ferromagnetic transition occurs at TC61K with an ordered moment of 3.0 ± 0.2 μB per Cr3+ at 4 K, and the magnetic anisotropy has an easy axis along the c axis [42]. To observe spin-wave excitations, we coaligned and stacked approximately 25 pieces of platelets with a total mass of approximately 0.3 g. Time-of-flight inelastic neutron scattering experiments were performed using the SEQUOIA spectrometer of the Spallation Neutron Source at the Oak Ridge National Laboratory using three different incident energies of Ei=50, 25, and 8 meV [44]. Neutron powder diffraction measurements were carried out using the BT-1 diffractometer of NIST Center for Neutron Research.

Figure 1(d) shows an overview of spin-wave dispersions along high symmetry directions in the (h,k) plane. A nearly isotropic spin-wave mode emerges from the Γ point at the ferromagnetic zone center and moves towards the zone boundary with increasing energy. This low-energy mode accounts for the in-phase oscillations between the two sublattice Cr spins within a unit cell. In the rest of the manuscript, we refer to this low energy mode as the “acoustic” magnon mode. Along the [h,0,0] direction towards the M point [Fig. 1(b)], the acoustic mode reaches its maximum energy around ω=10meV while another mode is visible at high energy between 15 and 19 meV. This high energy mode accounts for the out-of-phase oscillations between the two sublattice Cr spins, which we refer to as the “optical” mode. The large separation in energy between the two modes is consistent with the dominant ferromagnetic exchanges. Using a simple Heisenberg Hamiltonian with only the in-plane magnetic exchange couplings and without the DM interaction [26], we can fit to the overall momentum dependence of the spin-wave excitations as the solid lines in Fig. 1(d) [42]. While the overall agreement of the Heisenberg Hamiltonian is reasonably good, the calculation apparently fails to explain the observed spin-wave dispersions along the [h,h,0] direction going through the Dirac K point. As indicated by a thick red arrow in Fig. 1(d), the spin-wave intensity exhibits a clear discontinuity where the acoustic and optical modes are expected to cross each other. This observation strongly suggests that magnons at the Dirac points in CrI3 have a finite effective mass contrary to the Heisenberg-only Hamiltonian [26].

To accurately determine the spin-wave gap at Dirac points, we plot in Fig. 2 the constant-energy cuts at different spin-wave energies with an energy integration range of 1.0 meV, obtained by using the Ei=25meV data. Since CrI3 is a ferromagnet, spin-wave excitations stem from the Γ point at low energy transfer about ω=3meV [Fig. 2(f)]. Upon increasing energy to ω=6meV, spin waves form an isotropic ring pattern around the Γ point [Fig. 2(e)]. At ω=10meV, the ring breaks into six-folded patterns concentrated around K/K points, revealing the typical Heisenberg gap at M [Fig. 2(d)]. When the energy transfer is further increased, the six-folded pattern becomes invisible at ω12meV and only reappears for ω14meV. Consistent with Fig. 1(d), we find that the spin gap around ω12meV extends to the entire BZ including six equivalent Dirac points. These results suggest that the associated magnon modes obtain finite mass via interactions with each other or with additional degrees of freedom. One candidate may be the spin wave interacting with lattice excitations (phonons). In general, dynamic spin-lattice coupling can create energy gaps or broadening in the magnon dispersion at the nominal intersections of magnon and phonon modes [4549]. Although CrI3 has several phonon modes in the vicinity of the spin gap [50], the large magnitude (approximately 4 meV) and extension over the entire BZ of the gap suggest that magnon-phonon coupling is unlikely to be the origin of the gap. Alternatively, if one can construct a spin Hamiltonian that breaks the sublattice symmetry of the Cr honeycomb lattice, such a Hamiltonian will feature a magnon spectrum with spin gaps but without nontrivial topology and chiral edge states expected for topological spin excitations. However, the two Cr sublattice ions are chemically identical in CrI3, and therefore the sublattice asymmetry spin Hamiltonian can be ruled out for CrI3.

FIG. 2.

FIG. 2.

Constant-ħω cuts of spin waves at selected energy-transfer values within the (h,k) plane at 5 K. In all plots, data were subtracted by the empty-can background, integrated over the energy range of ħω±0.5meV as well as integrating over 5l5. The dashed lines are the BZ boundaries of the 2D reciprocal lattice. In (f), the Γ, K, and M points are marked along the first BZ boundary for clarity.

To qualitatively understand the observed spin-wave excitations, we fit the spin-wave spectra with the SpinW program [51]. By including the DM interaction (A) in the linear spin-wave Heisenberg Hamiltonian as [26]

H=i<j[JijSiSj+Aij(Si×Sj)]jDz(Sjz)2, (1)

where Jij is magnetic exchange coupling of the spin Si and Sj, Aij is the DM interaction between sites i and j, and Dz is the easy-axis anisotropy along the z(c) axis. We fit the data in the following two steps assuming a Cr spin of S=3/2. First, we integrate the data over l to improve the statistics, and fit the spin-wave dispersions in the (h,k) planes excluding Jc. Assuming that the nearest, next-nearest, and next-next-nearest-neighbor Cr-Cr magnetic exchange couplings are J1, J2, and J3 as shown in Fig. 1(c), our best-fit values yield J1=2.01, J2=0.16, and J3=0.08meV, comparable to the density functional theory calculations [5254]. Remarkably, we find |A|=0.31meV across the next-nearest neighbors, which is larger than J2. We note that the value of Dz=0.49meV obtained by this method is overestimated because there is a finite spin-wave bandwidth along the l direction (approximately 1.8 meV) as shown in Fig. 1(e). Both Jc and Dz were finally obtained by fitting the low-energy modes along the l direction while fixing the in-plane exchange constants. The best-fit values of these two parameters are Jc=0.59 and Dz=0.22meV, respectively, which are significantly larger than those in CrBr3 [21]. In particular, the anisotropy term in CrI3 is an order of magnitude larger than those of CrBr3. These results suggest that ferromagnetic order in CrI3 has much stronger c-axis exchange coupling compared with CrBr3, although both materials are van der Waals ferromagnets [55]. Since the single-layer CrI3 orders ferromagnetically below about TC45K [39] and not significantly different from the bulk of TC=61K, the magnetic ordering temperature of CrI3 must be mostly controlled by the in-plane magnetic exchange couplings as the c-axis exchange coupling of Jc=0.59meV in bulk is expected to vanish in the monolayer CrI3.

Since the fitted DM interaction |A| and the easy-axis anisotropy Dz are rather large, it is interesting to ask if there are other physical effects that may contribute to the observed large spin gap. Theoretically, the pseudodipolar interaction that can arise from superexchange and spin-orbit coupling may induce anisotropic bond-directional exchanges in the nearest neighbor of a honeycomb lattice and open a spin gap at the Dirac points [32]. However, the effect of pseudodipolar interaction on spin gap is much smaller than the comparable strength of DM interaction |A|. As shown in Fig. 1(b) of Ref. [32], the size of spin gap at Dirac points is rather small even when the pseudodipolar interaction is 5 times the nearest-neighbor exchange coupling. To induce a spin gap of approximately 4 meV as observed in CrI3, the pseudodipolar and single-ion anisotropy terms should be 3.4 and 2.3 times the nearest-neighbor exchange coupling, respectively, using the gap formula described in Ref. [32]. Therefore, it is highly unlikely that the pseudodipolar interaction in CrI3 induces the observed large spin gap at the Dirac points.

Figure 3 compares the calculated spin-wave spectra using these parameters with the experimentally observed dispersions. The left panel in Fig. 3(a) plots the calculated dispersion along the [h,0] direction, while the right panel shows the data. Similar spin-wave calculations and observed spectra along the [h,h], [h1/2,h+1/2], and [h+1/6,1/3] directions are shown in Figs. 3(b)3(d), respectively. In all cases, the spin-wave gap observed at Dirac K points is well reproduced by the calculation. The calculated spectra also reasonably reproduce the strong (weak) intensity of the acoustic (optical) spin-wave modes within the first BZ, which becomes weaker (stronger) in the second BZ. At M points, the ferromagnetic nearest-neighbor exchange couplings (J1>0) of the Heisenberg Hamiltonian ensures that the acoustic spin-wave mode is always lower in energy than the optical mode. The overall dispersion along the MΓM direction is not significantly affected by the DM interaction A [Fig. 3(e)]. On the other hand, we see clear splitting of the acoustic and optical spin-wave modes at K points along the ΓKM direction due to the large DM term, which also enhances the magnon density of states at Dirac points. Figure 3(f) shows the calculated overall spin-wave dispersion including the DM interaction.

FIG. 3.

FIG. 3.

(a)–(d) The neutron scattering intensities experimentally observed at 5 K are directly compared with the calculated intensities and dispersions discussed in the text. The experimental data on the right panels were integrated over 5l5. The calculations on the left panels used the best-fitting parameters with Jc=0 as discussed in the text, and convoluted by the effective energy resolution of δ(ħω)=1.8meV considering the band with along l. The displayed cuts are described in (e) within the (h,k) plane. (f) 3D view of the 2D spin-wave excitations of CrI3.

If we assume that the observed spin gap near Dirac points is due to the presence of DM interactions, spin waves in CrI3 should have topological edge states in place of massless Dirac magnons [26,36]. Such topological magnons emerge from localized spin-wave modes forming chiral vortices, among which the handedness may be chosen via local magnetic fields. From Γ towards K, the two sublattice spins are displaced from each other along the direction parallel to the wave front. At K=(1/3,1/3,0), the spins of one sublattice will precess by 120° along the direction of wave propagation. As a result, the excitations of the two sublattice spins will be decoupled from each other since the in-plane Heisenberg exchanges are frustrated. For instance, the two sublattice excitations illustrated in Figs. 4(a) and 4(b) have the mutually equivalent handedness when they propagate along K. When localized within a single hexagon, however, they exhibit opposite handedness from each other as depicted by thin arrows. As a result, these two localized spin-wave excitations can become potentially incompatible with each other. Although mutually degenerate via Heisenberg exchanges, their degeneracy is lifted when the next-nearest-neighbor DM vectors are introduced [26]. Interestingly, the apparent magnitude of the DM interactions in CrI3 is larger than J2 along the same next-nearest pair, and as large as 14% of J1. Given that no such feature has been observed in CrBr3 [21], this effect must arise from the larger spin-orbit coupling in CrI3 when bromine is replaced by the heavier iodine.

FIG. 4.

FIG. 4.

(a) The spin displacements of the two sublattice spins at the right-handed chiral wave at K points. The long arrow at the center denotes the direction of spin-wave propagation. The thin triangular arrows mark the direction of rotation for the right-handed chiral within a hexagon. (b) Spin-wave excitations at T=52K. The dashed lines are the calculations using the best-fitting parameters from the data at T=5K but with Dz set to be 0. (c) Temperature dependencies of the stiffness of the acoustic mode [D(T), circle] and the anisotropy gap (Eg, square) obtained by fitting the data integrated over 5l5. For this reason, the absolute values of Eg are overestimated by the factor of approximately 2. The open and closed symbols are the results obtained using Ei=8 and 25 meV, respectively. The blue solid and dashed lines are expected D(T) from 3D Heisenberg and Ising models, respectively.

Finally, we discuss temperature dependence of spin-wave excitations in CrI3. As the temperature is increased towards TC, the magnons gradually broaden and soften. At T=52K (T/TC=0.85), the overall spin-wave spectra remain unchanged with a spin gap at Dirac points [Fig. 4(b)]. In the hydrodynamic limit of long wavelengths and small q, spin-wave energy ħω has the quadratic q dependence via ħω=Δ(T)+D(T)q2, where D(T) is the temperature dependence of the spin-wave stiffness and Δ(T) is the small dipolar gap arising from the spin anisotropy [56]. For a simple 3D Heisenberg ferromagnet, temperature dependence of the spin-wave stiffness D(T) is expected to renormalize to zero at TC via [(TTC)/TC]νβ, where ν=0.707 and β=0.367 are critical exponents [57]. The blue solid and dashed lines in Fig. 4(c) show the resulting temperature dependence of D(T) for the 3D Heisenberg and 1D Ising model with ν=0.631 and β=0.326, respectively. For CrI3, magnetic critical exponent behavior was found to be 3D-like with β=0.26 [55]. While the temperature dependence of D(T) is finite approaching TC clearly different from the 3D Heisenberg or 1D Ising expectation [Fig. 4(c)], the results are rather similar to D(T) in ferromagnetic manganese oxides [58]. On the other hand, Δ(T) vanishes at TC [Fig. 4(c)], suggesting that the spin anisotropy fields play an important role in stabilizing the 2D ferromagnetic ordering [54].

III. CONCLUSIONS

In summary, our inelastic neutron scattering experiments reveal a large gap in the spin-wave excitations of CrI3 at the Dirac points. The acoustic and optical spin-wave bands are separated from each other by approximately 4 meV, most likely arising from the next-nearest-neighbor DM interaction that breaks inversion symmetry of the lattice. This may lead to a nontrivial topological magnon insulator with magnon edge states, analogous to topological insulators in electronic systems but without electric Ohmic heating. These properties make CrI3 appealing for high-efficiency and dissipationless spintronic applications, although there are also challenges in making real spintronic devices [3033]. Our analysis of the observed spin waves suggests that the DM interaction is stronger than the Heisenberg exchange coupling between the next-nearest-neighbor pairs in the 2D honeycomb lattice of CrI3. The observation of a large spin gap in magnons of CrI3 and its absence in CrBr3 suggests that spin-orbit coupling plays an important role in the physics of topological spin excitations in honeycomb ferromagnet CrI3.

ACKNOWLEDGMENTS

The neutron scattering work at Rice is supported by the U.S. NSF Grant No. DMR-1700081 (P. D.). The CrI3 single-crystal synthesis work was supported by the Robert A. Welch Foundation Grant No. C-1839 (P. D.). The work of J. H. C. was supported by the National Research Foundation of Korea (No. NRF-2016R1D1A1B03934157; No. NRF-2017K1A3A7A09016303). Research at Oak Ridge National Laboratory’s Spallation Neutron Source was supported by the Scientific User Facilities Division, Office of Basic Energy Sciences, U.S. Department of Energy. Oak Ridge National Laboratory is managed by UT-Battelle, LLC, for U.S. DOE under Contract No. DE-AC05-00OR22725.

References

  • [1].Klitzing K. v., Dorda G, and Pepper M, New Method for High-Accuracy Determination of the Fine-Structure Constant Based on Quantized Hall Resistance, Phys. Rev. Lett 45, 494 (1980). [Google Scholar]
  • [2].Haldane FDM, Model for a Quantum Hall Effect without Landau Levels: Condensed-Matter Realization of the Parity Anomaly, Phys. Rev. Lett 61, 2015 (1988). [DOI] [PubMed] [Google Scholar]
  • [3].Hasan MZ and Kane CL, Topological Insulators, Rev. Mod. Phys 82, 3045 (2010). [Google Scholar]
  • [4].Qi X-L and Zhang S-C, Topological Insulators and Superconductors, Rev. Mod. Phys 83, 1057 (2011). [Google Scholar]
  • [5].Bansil A, Lin H, and Das T, Topological Band Theory, Rev. Mod. Phys 88, 021004 (2016). [Google Scholar]
  • [6].Zhou Y, Kanoda K, and Ng T-K, Quantum Spin Liquid States, Rev. Mod. Phys 89, 025003 (2017). [Google Scholar]
  • [7].Neto AHC, Guinea F, Peres NMR, Novoselov KS, and Geim AK, The Electronic Properties of Graphene, Rev. Mod. Phys 81, 109 (2009). [Google Scholar]
  • [8].Wehling TO, Black-Schaffer AM, and Balatsky AV, Dirac Materials, Adv. Phys 63, 1 (2014). [Google Scholar]
  • [9].Raghu S. and Haldane FDM, Analogs of Quantum-Hall-Effect Edge States in Photonic Crystals, Phys. Rev. A 78, 033834 (2008). [Google Scholar]
  • [10].Wang Z, Chong Y, Joannopoulos JD, and Soljačić M, Observation of Unidirectional Back Scattering-Immune Topological Electromagnetic States, Nature (London) 461, 772 (2009). [DOI] [PubMed] [Google Scholar]
  • [11].Katsura H, Nagaosa N, and Lee PA, Theory of the Thermal Hall Effect in Quantum Magnets, Phys. Rev. Lett 104, 066403 (2010). [DOI] [PubMed] [Google Scholar]
  • [12].Onose Y, Ideue T, Katsura H, Shiomi Y, Nagaosa N, and Tokura Y, Observation of the Magnon Hall Effect, Science 329, 297 (2010). [DOI] [PubMed] [Google Scholar]
  • [13].Zhang L, Ren J, Wang J-S, and Li B, Topological Magnon Insulator in Insulating Ferromagnet, Phys. Rev. B 87, 144101 (2013). [Google Scholar]
  • [14].Shindou R, Ohe J.-i., Matsumoto R, Murakami S, and Saitoh E, Chiral Spin-Wave Edge Modes in Dipolar Magnetic Thin Films, Phys. Rev. B 87, 174402 (2013). [Google Scholar]
  • [15].Mook A, Henk J, and Mertig I, Edge States in Topological Magnon Insulators, Phys. Rev. B 90, 024412 (2014). [Google Scholar]
  • [16].Mook A, Henk J, and Mertig I, Tunable Magnon Weyl Points in Ferromagnetic Pyrochlores, Phys. Rev. Lett 117, 157204 (2016). [DOI] [PubMed] [Google Scholar]
  • [17].Li F-Y, Li Y-D, Kim YB, Balents L, Yu Y, and Chen G, Weyl Magnons in Breathing Pyrochlore Antiferromagnets, Nat. Commun 7, 12691 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [18].Pershoguba SS, Banerjee S, Lashley JC, Park J, Gren H, Aeppli G, and Balatsky AV, Dirac Magnons in Honeycomb Ferromagnets, Phys. Rev. X 8, 011010 (2018). [Google Scholar]
  • [19].Owerre SA, Dirac Magnon Nodal Loops in Quasi-2D Quantum Magnets, Sci. Rep 7, 6931 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Fransson J, Black-Schaffer AM, and Balatsky AV, Magnon Dirac Materials, Phys. Rev. B 94, 075401 (2016). [Google Scholar]
  • [21].Samuelsen EJ, Silberglitt R, Shirane G, and Remeika JP, Spin Waves in Ferromagnetic CrBr3 Studied by Inelastic Neutron Scattering, Phys. Rev. B 3, 157 (1971). [Google Scholar]
  • [22].Williams TJ, Aczel AA, Lumsden MD, Nagler SE, Stone MB, Yan J-Q, and Mandrus D, Magnetic Correlations in the Quasi-Two-Dimensional Semiconducting Ferromagnet CrSiTe3, Phys. Rev. B 92, 144404 (2015). [Google Scholar]
  • [23].Bao S, Wang J, Wang W, Cai Z, Li S, Ma Z, Wang D, Ran K, Dong Z-Y, Abernathy DL, Yu S-L, Wan X, Li J-X, and Wen J, Discovery of Coexisting Dirac and Triply Degenerate Magnons in a Three-Dimensional Antiferromagnet, Nat. Commun 9, 2591 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].Yao W, Li C, Wang L, Xue S, Dan Y, Iida K, Kamazawa K, Li K, Fang C, and Li Y, Topological Spin Excitations in a Three-Dimensional Antiferromagnet, Nat. Phys 14, 1011 (2018). [Google Scholar]
  • [25].Yao Y, Ye F, Qi X-L, Zhang S-C, and Fang Z, Spin-Orbit Gap of Graphene: First-Principles Calculations, Phys. Rev. B 75, 041401(R) (2007). [Google Scholar]
  • [26].Owerre SA, A First Theoretical Realization of Honeycomb Topological Magnon Insulator, J. Phys. Condens. Matter 28, 386001 (2016). [DOI] [PubMed] [Google Scholar]
  • [27].Kim SK, Ochoa H, Zarzuela R, and Tserkovnyak Y, Realization of the Haldane-Kane-Mele Model in a System of Localized Spins, Phys. Rev. Lett 117, 227201 (2016). [DOI] [PubMed] [Google Scholar]
  • [28].Rückriegel A, Brataas A, and Duine RA, Bulk and Edge Spin Transport in Topological Magnon Insulators, Phys. Rev. B 97, 081106(R) (2018). [Google Scholar]
  • [29].Cheng R, Okamoto S, and Xiao D, Spin Nernst Effect of Magnons in Collinear Antiferromagnets, Phys. Rev. Lett 117, 217202 (2016). [DOI] [PubMed] [Google Scholar]
  • [30].Chumak AV, Vasyuchka VI, Serga AA, and Hille-brands B, Magnon Spintronics, Nat. Phys 11, 453 (2015). [Google Scholar]
  • [31].Chernyshev AL and Maksimov PA, Damped Topological Magnons in the Kagome-Lattice Ferromagnets, Phys. Rev. Lett 117, 187203 (2016). [DOI] [PubMed] [Google Scholar]
  • [32].Wang XS, Su Y, and Wang XR, Topologically Protected Unidirectional Edge Spin Waves and Beam Splitter, Phys. Rev. B 95, 014435 (2017). [Google Scholar]
  • [33].Wang XS, Zhang HW, and Wang XR, Topological Magnonics: A Paradigm for Spin-Wave Manipulation and Device Design, Phys. Rev Applied 9, 024029 (2018). [Google Scholar]
  • [34].Moriya T, Anisotropic Superexchange Interaction and Weak Ferromagnetism, Phys. Rev 120, 91 (1960). [Google Scholar]
  • [35].Pereiro M, Yudin D, Chico J, Etz C, Eriksson O, and Bergman A, Topological Excitations in a Kagome Magnet, Nat. Commun 5, 4815 (2014). [DOI] [PubMed] [Google Scholar]
  • [36].Chisnell R, Helton JS, Freedman DE, Singh DK, Bewley RI, Nocera DG, and Lee YS, Topological Magnon Bands in a Kagome Lattice Ferromagnet, Phys. Rev. Lett 115, 147201 (2015). [DOI] [PubMed] [Google Scholar]
  • [37].Chisnell R, Helton JS, Freedman DE, Singh DK, Demmel F, Stock C, Nocera DG, and Lee YS, Magnetic Transitions in the Topological Magnon Insulator Cu(1,3-bdc), Phys. Rev. B 93, 214403 (2016). [Google Scholar]
  • [38].Gong C, Li L, Li Z, Ji H, Stern A, Xia Y, Cao T, Bao W, Wang C, Wang Y, Qiu ZQ, Cava RJ, Louie SG, Xia J, and Zhang X, Discovery of Intrinsic Ferromagnetism in Two-Dimensional Van der Waals Crystals, Nature (London) 546, 265 (2017). [DOI] [PubMed] [Google Scholar]
  • [39].Huang B, Clark G, Navarro-Moratalla E, Klein DR, Cheng R, Seyler KL, Zhong D, Schmidgall E, McGuire MA, Cobden DH, Yao W, Xiao D, Jarillo-Herrero P, and Xu X, Layer-Dependent Ferromagnetism in a Van der Waals Crystal Down to the Monolayer Limit, Nature (London) 546, 270 (2017). [DOI] [PubMed] [Google Scholar]
  • [40].Bonilla M, Kolekar S, Ma Y, Diaz HC, Kalappatti V, Das R, Eggers T, Gutierrez HR, Phan M-H, and Batzill M, Strong Room-Temperature Ferromagnetism in VSe2 Mmonolayers on Van der Waals Substrates, Nat. Nanotechnol 13, 289 (2018). [DOI] [PubMed] [Google Scholar]
  • [41].Mcguire MA, Dixit H, Cooper VR, and Sales BC, Coupling of Crystal Structure and Magnetism in the Layered, Ferromagnetic Insulator CrI3, Chem. Mater 27, 612 (2015). [Google Scholar]
  • [42].See Supplemental material at http://link.aps.org/supplemental/10.1103/PhysRevX.8.041028 for additional details.
  • [43].Mcgurie MA, Crystal and Magnetic Structures in Layered, Transition Metal Dihalides and Trihalides, Crystal 7, 121 (2017). [Google Scholar]
  • [44].Granroth GE, Kolesnikov AI, Sherline TE, Clancy JP, Ross KA, Ruff JPC, Gaulin BD, and Nagler SE, SEQUOIA: A Newly Operating Chopper Spectrometer at the SNS, J. Phys. Conf. Ser 251, 012058 (2010). [Google Scholar]
  • [45].Anda E, Effect of Phonon-Magnon Interaction on the Green Functions of Crystals and Their Light-Scattering Spectra, J. Phys. C 9, 1075 (1976). [Google Scholar]
  • [46].Guerreiro SC and Rezende SM, Magnon-Phonon Interconversion in a Dynamically Reconfigurable Magnetic Material, Phys. Rev. B 92, 214437 (2015). [Google Scholar]
  • [47].Oh J, Le MD, Nahm H-H, Sim H, Jeong J, Perring TG, Woo H, Nakajima K, Ohira-Kawamura S, Yamani Z, Yoshida Y, Eisaki H, Cheong S-W, Chernyshev AL, and Park J-G, Spontaneous Decays of Magneto-Elastic Excitations in Non-Collinear Antiferromagnet Y; Lu MnO3, Nat. Commun 7, 13146 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Dai PC, Hwang HY, Zhang J, Fernandez-Baca JA, Cheong S-W, Kloc C, Tomioka Y, and Tokura Y, Magnon Damping by Magnon-Phonon Coupling in Manganese Perovskites, Phys. Rev. B 61, 9553 (2000). [Google Scholar]
  • [49].Man H, Shi Z, Xu GY, Xu YD, Chen X, Sullivan S, Zhou JS, Xia K, Shi J, and Dai PC, Direct Observation of Magnon-Phonon Coupling in Yttrium Iron Garnet, Phys. Rev. B 96, 100406(R) (2017). [Google Scholar]
  • [50].Zhang W-B, Qu Q, Zhua P, and Lam C-H, Robust Intrinsic Ferromagnetism and Half Semiconductivity in Stable Two-Dimensional Single-Layer Chromium Trihalides, J. Mater. Chem. C 3, 12457 (2015). [Google Scholar]
  • [51].Toth S. and Lake B, Linear Spin Wave Theory for Single-Q Incommensurate Magnetic Structures, J. Phys. Condens. Matter 27, 166002 (2015). [DOI] [PubMed] [Google Scholar]
  • [52].Wang H, Eyert V, and Schwingenschlögl U, Electronic Structure and Magnetic Ordering of the Semiconducting Chromium Trihalides CrCl3, CrBr3, and CrI3, J. Phys. Condens. Matter 23, 116003 (2011). [DOI] [PubMed] [Google Scholar]
  • [53].Wang H, Fan F, Zhu S, and Wu H, Doping Enhanced Ferromagnetism and Induced Half-Metallicity in CrI3 Monolayer, Europhys. Lett 114, 47001 (2016). [Google Scholar]
  • [54].Lado JL and Fernández-Rossier J, On the Origin of Magnetic Anisotropy in Two Dimensional CrI3, 2D Mater. 4, 035002 (2017). [Google Scholar]
  • [55].Liu Y. and Petrovic C, Three-Dimensional Magnetic Critical Behavior in CrI3, Phys. Rev. B 97, 014420 (2018). [Google Scholar]
  • [56].Lovesey SW, Theory of Thermal Neutron Scattering from Condensed Matter (Clarendon, Oxford, 1984), Vol. 2, Chap. 9. [Google Scholar]
  • [57].Collins M, Magnetic Critical Scattering (Clarendon, Oxford, 1989), p. 29. [Google Scholar]
  • [58].Dai PC, Fernandez-Baca JA, Plummer EW, Tomioka Y, and Tokura Y, Magnetic Coupling in the Insulating and Metallic Ferromagnetic La1−xCaxMnO3, Phys. Rev. B 64, 224429 (2001). [Google Scholar]

RESOURCES