Linear Magnetoelectric Phase in Ultrathin MnPS3 Probed by Optical Second Harmonic Generation Hao Chu,1,2 Chang Jae Roh,3 Joshua O. Island,4 Chen Li,1,2 Sungmin Lee,5,6 Jingjing Chen,7 Je-Geun Park,5,6 Andrea F. Young,4 Jong Seok Lee,3 and David Hsieh 1,2,* 1Department of Physics, California Institute of Technology, Pasadena, California 91125, USA 2Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, California 91125, USA 3Department of Physics and Photon Science, Gwangju Institute of Science and Technology, Gwangju 61005, Republic of Korea 4Department of Physics, University of California, Santa Barbara, California 93106, USA 5Center for Correlated Electron Systems, Institute for Basic Science (IBS), Seoul 08826, Republic of Korea 6Department of Physics and Astronomy, Seoul National University (SNU), Seoul 08826, Republic of Korea 7School of Physics, Nankai University, Tianjin 300071, China (Received 22 May 2019; published 16 January 2020) The transition metal thiophosphates MPS3 (M ¼ Mn, Fe, Ni) are a class of van der Waals stacked insulating antiferromagnets that can be exfoliated down to the ultrathin limit. MnPS3 is particularly interesting because its Ne´el ordered state breaks both spatial-inversion and time-reversal symmetries, allowing for a linear magnetoelectric phase that is rare among van der Waals materials. However, it is unknown whether this unique magnetic structure of bulk MnPS3 remains stable in the ultrathin limit. Using optical second harmonic generation rotational anisotropy, we show that long-range linear magnetoelectric type Ne´el order in MnPS3 persists down to at least 5.3 nm thickness. However an unusual mirror symmetry breaking develops in ultrathin samples on SiO2 substrates that is absent in bulk materials, which is likely related to substrate induced strain. DOI: 10.1103/PhysRevLett.124.027601 Thin film materials that exhibit the magnetoelectric (ME) effect—a coupling between magnetic (electric) polarization and external electric (magnetic) fields—have potentially broad applications in spintronics, sensing, and energy harvesting technologies [1,2]. Although ME effects in single-phase bulk crystals have been continuously pursued since their discovery in Cr2O3 in 1960 [3,4], advances in thin film deposition techniques over the past two decades have opened new pathways to stabilize and to control high quality materials with large ME coupling strengths via epitaxial strain and heterostructure engineering, allowing the possibility of integration into functional nanoscale devices. At present, searching for both single-phase and composite thin film materials with stronger ME coupling, and developing methods to scale them down to the ultrathin few-unit-cell limit, remain active areas of research. The recent discoveries of long-range magnetic ordering in exfoliated van der Waals (vdW) semiconductors [5–8] potentially offer a new route to realizing ME materials in the ultrathin limit. The simplest type of ME effect, which involves a linear coupling between the external field and induced polarization, is allowed in materials that lack both spatial-inversion and time-reversal symmetries. As most of the naturally occurring vdW crystals are structurally centrosymmetric, a convenient strategy is to rely on the magnetic ordering itself to break inversion symmetry. This suggests that one should focus on antiferromagnetic (AF) rather than ferromagnetic (FM) materials because the latter generally do not break the inversion symmetry of the underlying lattice. It was recently reported that upon exfoliating CrI3 down to a single bilayer, its magnetic order transforms from being FM to AF, breaking inversion symmetry and turning on a linear ME coupling in the process [5,9–11]. However, so far there are no reports of an ultrathin material that directly inherits linear ME properties from its bulk precursor. The transition metal thiophosphates MPS3 (M ¼ Mn, Fe, Ni) present an interesting family of AF vdW materials for such a study [12–15]. While the AF orders in FePS3 and NiPS3 preserve inversion symmetry [16,17], neutron dif- fraction studies have shown that the AF order in bulk MnPS3 breaks inversion symmetry and allows a linear ME effect [18]. However, it is not clear if the linear ME-type AF order persists down to the ultrathin limit. Because such order does not exhibit any net magnetization, a magneto- optical Kerr rotation experiment is not applicable. Although Raman spectroscopy has detected phonon anomalies in ultrathin MnPS3 that are potentially associated with AF ordering [19], and spin transport measurements have shown evidence of persistent magnons in few layer MnPS3 devices [20], a technique that directly probes the AF structure in nanoscopic exfoliated samples is still urgently anticipated. Leveraging the sensitivity of optical second harmonic generation (SHG) to AF order [21], we demonstrate here PHYSICAL REVIEW LETTERS 124, 027601 (2020) 0031-9007=20=124(2)=027601(6) 027601-1 © 2020 American Physical Society that SHG rotational anisotropy (RA) can directly couple to the AF order parameter in MnPS3 nanoflakes, and use it to show that the linear ME-type AF order found in bulk MnPS3 persists down to the ultrathin limit. Bulk MnPS3 crystallizes in a monoclinic structure with centrosymmetric 2=m point group symmetry [18]. It has a twofold rotation axis along the crystallographic b axis and a mirror plane perpendicular to bˆ [Figs. 1(a) and 1(b)]. The Mn atoms are octahedrally coordinated by S atoms and form a honeycomb lattice in the ab plane, but the in-plane sixfold rotational symmetry of the honeycomb lattice is absent in the bulk crystal due to the displacement of adjacent layers along aˆ. Despite the similar lattice struc- tures of MnPS3, FePS3, and NiPS3 [Fig. 1(c)], MnPS3 hosts an inversion broken Ne´el-type AF order [18] whereas FePS3 and NiPS3 exhibit inversion symmetric zigzag-type AF order [16,17]. The Ne´el-type AF order preserves the size of the unit cell and exhibits no net moment, therefore it is challenging to detect via Raman spectroscopy and magneto-optical Kerr rotation, respectively. However, because it breaks inversion symmetry, it should exhibit a finite second-order electric-dipole (ED) susceptibility that is responsible for SHG [22]. Therefore, we expect to see a finite SHG yield below the AF ordering temperature TAF from MnPS3 and but not FePS3 and NiPS3. The SHG-RA experiments were performed with a Ti: sapphire oscillator delivering laser pulses with a photon energy of ℏω ¼ 1.5 eV, a pulse width of 80 fs, and a repetition rate of 80 MHz. The SHG photons produced at 3 eV are resonant with the band gap of MnPS3 [23]. A 5× (50×) microscope objective was used to focus light onto the bulk (exfoliated) samples at normal incidence with a spot size of approximately 30 μm (2 μm), and the intensity of the reflected SHG beam was measured using a photomultiplier tube. The pulse energy of the incoming beam was kept below 50 pJ. The SHG-RA patterns were acquired by rotating the linear polarization of the incoming and outgoing beams (parametrized by the angle ϕ), which were maintained parallel to each other in the ab plane [Fig. 1(c)]. BulkMPS3 single crystals were grown by a self- flux method described elsewhere [24]. Despite having a centrosymmetric crystallographic point group, we observe weak but finite SHG-RA signals from all three bulk crystals even above TAF [Fig. 2(a)]. This may arise from surface ED SHG or higher-rank bulk SHG processes such as electric-quadrupole (EQ) SHG [22], both of which are generally allowed in centrosymmetric materi- als and were found to fit the data equally well [Fig. 2(a)] [25]. For simplicity, we therefore only consider the bulk EQ term in our later fitting. The loss of sixfold rotational symmetry that arises from the stacking offset between adjacent honeycomb layers is apparent in the data, although the degree of departure from sixfold symmetry varies across samples as well as across spots within a single sample. We speculate that this may be due to spatial variations in the strength of interlayer coupling and/or variations in the concentration of 120° twins or stacking faults [29]. Below TAF we observe no changes in the SHG intensity from both FePS3 and NiPS3, but an increase in the SHG intensity fromMnPS3 as anticipated. As shown in Fig. 2(b), the low temperature SHG-RA patterns from MnPS3 can be well fit using the coherent sum of a nonmagnetic EQ contribution and an AF order induced time-noninvariant ED contribution described by the equation [25] P2ωi ¼ χEQijklEωj ∇kEωl þ χEDijkðTÞEωj Eωk ; ð1Þ where P2ω is the induced electric polarization at the SHG frequency, Eω is the magnitude of the incident electric field, χEQijkl is the temperature independent EQ susceptibil- ity from a 2=m crystallographic point group, and χEDijk ðTÞ is a temperature dependent ED susceptibility from the 20=m magnetic point group describing the Ne´el phase [18]. As shown in Fig. 2(c), the SHG intensity from MnPS3 shows an order parameterlike increase below TAF. Since χEDijk ðTÞ is directly proportional to the inversion broken Ne´el order parameter, we can extract the critical exponent of the order parameter (β) by fitting the temper- ature dependent SHG intensity to the phenomenological function I2ω ∝ ½aþ bðTAF − TÞβ2, where a is fixed by the intensity of the EQ contribution above TAF and both b and β are free parameters. Best fits to the region 60 K ≤ T ≤ TAF yield β ¼ 0.37ð8Þ [Fig. 2(c)], which is close to the (a) (c) MnPS3 FePS3 NiPS3E( ) a Inversion center c b M P S a b c a b c (b) FIG. 1. Crystal and magnetic structure of MPS3. MPS3 lattice viewed along the (a) c and (b) b axis. Adjacent ab planes are displaced by a=3 along the aˆ direction. (c) AF structures of MPS3. Arrows denote spin orientation. Star denotes an inversion center of the AF structure. The inset shows the in-plane orientation of the incident electric field. PHYSICAL REVIEW LETTERS 124, 027601 (2020) 027601-2 numerical calculation of ∼0.369 for the 3D Heisenberg model [30]. To investigate whether the long-range Ne´el order in MnPS3 survives in the ultrathin limit, we exfoliated bulk crystals onto an amorphous SiO2 substrate in a nitrogen purged glove box. The choice of pure SiO2 over SiO2=Si as a substrate was made to reduce laser induced heating arising from optical absorption by Si at 800 and 400 nm. In contrast, SiO2 is transparent to both 800 and 400 nm light. Because of the poor thermal conductivity of SiO2 and the relatively high laser power needed for our SHG-RA measurements on MnPS3 compared to other optical techniques for studying vdW magnets such as magneto-optical Kerr microscopy or Raman spectroscopy, we face more stringent sample cooling demands [25]. To increase cooling efficiency, we deposited gold rings around the MnPS3 flakes, which are thermally anchored to the cryostat sample holder by gold electrodes. Figure 3(a) shows an optical image of a typical device. Using atomic force microscopy, we identified ultrathin MnPS3 nanoflakes with 5.3 and 12.5 nm step sizes above the substrate on this device [Fig. 3(b)]. Based on previously published atomic force microscopy data on MnPS3 [24], these correspond to 7 and 16 single layers of MnPS3, respectively. Figure 3(c) shows typical SHG-RA patterns obtained from these flakes at a temperature of 10 K, compared with both thicker (75 nm) flakes and the bare substrate. We find that the overall SHG intensity approximately scales with the sample thickness, consistent with a bulk dominated SHG signal. The SiO2 substrate contributes an isotropic background and is thus easily distinguished from the MnPS3 signal. As shown in Fig. 2(c), the ED SHG signal from MnPS3 below TAF is of comparable magnitude to the high temper- ature EQ signal and is thus relatively weak overall. This is likely related to our incident 1.5 eV photon energy being well below the band gap (∼3 eV) of MnPS3. Consequently, when we attempted to protect the MnPS3 flakes by T > T(b) (c) 200 K 0.4 0.6 0.8 1.0 1.2 1 2 3 NiPS3 (TAF = 155 K) MnPS3 (TAF = 78 K) FePS3 (TAF = 123 K) T/TAF SH G In te ns ity (N orm ) 10 K 30 K 50 K 100 K 50 K 90 K 130 K 150 K 50 K 100 K 140 K 180 K 0.2 0 1 0.5 2.5 1 1 1 1 MnPS3 FePS3 NiPS3 1 (a) AF T < TAF 0° FIG. 2. SHG-RA patterns and long-range Ne´el order in MnPS3. (a) SHG-RA patterns of MPS3 above and (b) below their respective AF ordering temperatures: TAF ¼ 78 (MnPS3), 123 (FePS3), and 155 K (NiPS3). Solid circles are experimental data and the solid lines are best fits to the phenomenological model described in the main text. For T > TAF, all data were fit using only an EQ term. For T < TAF, the FePS3 and NiPS3 data were fit using only an EQ term, whereas the MnPS3 data were fit using a coherent sum of an EQ and ED term. (c) Temperature dependence of the SHG intensity along the ϕ ¼ 60° direction. Solid line on the MnPS3 data is a best fit to the power law function described in the main text, which accounts for a constant EQ term and a temperature dependent ED term. 2 x20 x20 SiO2 75 nm 12.5 nm SiO 12.5 nm 5.3 nm 5 µm (b) (c)(a) 50 µm MnPS3 5.3 nm x1 x5 SiO2 FIG. 3. MnPS3 nanodevice for probing long-range Ne´el order. (a) Optical image of exfoliated MnPS3 flakes on SiO2 with gold ring and electrode on top to improve cooling efficiency. The 5.3 and 12.5 nm flakes are found within the white box. (b) Atomic force microscopy scan of the area bounded by the white box in (a). Green lines indicate the positions of line scans, with corresponding magnified line profiles shown in white. (c) SHG-RA patterns from various regions of the device at 10 K. PHYSICAL REVIEW LETTERS 124, 027601 (2020) 027601-3 encapsulation with a hexagonal boron nitride (hBN) thin flake, we found that the SHG signal was dominated by the hBN. Therefore we had to work with exposed MnPS3 flakes, which are more prone to degradation. At cryogenic temperatures, we found that the SHG intensity from the few-layer regions starts to decrease over a timescale of several hours. This is likely due to surface adsorption of gas molecules and/or chemical reaction processes activated by laser exposure, as is observed in CrI3 nanoflakes [31]. Therefore we were only able to acquire a limited number of SHG-RA scans at low temperatures before the onset of sample degradation. Nevertheless, our data clearly show an order parameterlike increase in the SHG intensity from the MnPS3 nanoflakes below a temperature close to the bulk TAF value (Fig. 4), which again saturates at only several times the high temperature value. This indicates that the linear ME-type Ne´el ordering observed in bulk crystals persists at least down to 7 layer thick samples. Measurements collected from 3 layer samples also show a markedly higher SHG intensity at 10 K compared to 100 K [25], but their faster degradation prevented a full temperature dependence measurement from being taken. Given that the low temperature SHG signal from MnPS3 involves the interference between a time-noninvariant ED response and a time-invariant EQ response, the existence of two different 180° AF domains related by time reversal should produce different SHG intensities, analogous to what has previously been observed in Cr2O3 [32]. Assuming that the roughly twofold SHG intensity increase in our 5.3 nm flakes at low temperature [Fig. 4(d)] arises from a pure þ domain where the ED and EQ contributions interfere constructively, the ratio of the ED to EQ SHG electric fields should be roughly 1 2 . By extension, the SHG intensity from a − domain is expected to be ∼25% of the high temperature value, or ð1þ 1 2 Þ2=ð1 − 1 2 Þ2 ¼ 9 times lower in intensity compared to the þ domain. By raster scanning our beam of spot size ∼2 μm over the roughly 4 μm × 4 μm area of our 5.3 nm MnPS3 flake at 10 K, we found the SHG intensity varies by only approximately 30% about the mean intensity. These small variations may be due to sample inhomogeneity and/or slight changes in alignment and are inconsistent with 180° domains. Therefore we believe our flake to be a single AF domain, which is comparable to the FM domain sizes observed in ultrathin vdW materials like CrI3 [5] and Cr2Ge2Te6 [6]. We note that the low temperature SHG-RA patterns from the ultrathin 5.3 nm flakes exhibit an unusual symmetry. In particular, the ac mirror plane (reflection about the hori- zontal line in the SHG-RA patterns) that is preserved by both the 2=m and 20=m point groups is absent [Fig. 3(c)]. This mirror symmetry breaking only becomes apparent as the material thickness is reduced and as the temperature is lowered below TAF [Fig. 4(c)]. We believe that this is related to a substrate induced strain because for ultrathin MnPS3 flakes that are exfoliated onto SiO2=Si substrates, which are much smoother than pure SiO2 substrates, there is no clear evidence of ac mirror breaking in the low temperature SHG-RA patterns [25]. Model Hamiltonian calculations [33] show that a spiral spin texture that breaks acmirror symmetry is favored over the collinear Ne´el order only if the second nearest-neighbor Dzyaloshinskii-Moriya interaction D2 is comparable to the nearest-neighbor exchange J1 in MnPS3, which is around 1.5 meVaccording to inelastic neutron diffraction experiments [34]. However, spin Hall based measurements ofD2 in MnPS3 put its value at merely 0.3 meV [35]. Since it is unlikely that D2 is several times larger in ultrathin flakes compared to bulk crystals, especially given that Raman spectroscopy studies show no drastic changes in TAF or the phonon spectrum as a function of thickness [19], we rule out a noncollinear spin texture as the cause for ac mirror breaking. Instead, it is possible that the substrate induced strain tilts the easy axis, causing the Ne´el ordered moments to rigidly cant out of the ac plane. Further structural and magnetic characterization of thin MnPS3 flakes will be necessary to confirm this hypothesis. In conclusion, we have demonstrated SHG-RA to be a direct and effective probe of inversion breaking AF order 4 2.5 (a) (b) (c) (d) Temperature (K) 75 nm 12.5 nm 5.3 nm 75 n m 5. 3 nm 10 K 50 K 60 K 70 K 120 K SH G In te ns ity (N orm ) 12 .5 n m 0 40 80 1 7 0 40 80 0 40 80 120 FIG. 4. Long-range Ne´el order in few-layer MnPS3. (a)–(c) Temperature dependent SHG-RA patterns from MnPS3 flakes of different thicknesses. (d) Normalized temperature dependence of SHG Intensity along the ϕ ¼ 0° direction. PHYSICAL REVIEW LETTERS 124, 027601 (2020) 027601-4 parameters in exfoliated vdW materials. A linear ME-type Ne´el order that features in bulk crystals of MnPS3 was found to survive down to the few layer limit. Future quantitative measurements of the ME coupling strength in ultrathin MnPS3 samples will help to assess its potential for applications in nanoscale spintronics and optoelectronics devices. Work at Caltech and UCSB was supported by ARO MURI Grant No. W911NF-16-1-0361. Work at GISTwas supported by National Research Foundation of Korea (NRF) grants funded by the Korea government (MSIP) (No. 2018R1A2B2005331). Work at IBS CCES was supported by Institute for Basic Science (IBS) in Korea (IBS-R009-G1). D. H. and J. S. L. also acknowledge support from a GIST-Caltech collaborative grant. J. O. I. acknowledges the support of the Netherlands Organization for Scientific Research (NWO) through the Rubicon grant, Project No. 680-50-1525/2474. *Corresponding author. dhsieh@caltech.edu [1] M. Fiebig, Revival of the magnetoelectric effect, J. Phys. D 38, R123 (2005). [2] N. A. Spaldin and R. Ramesh, Advances in magnetoelectric multiferroics, Nat. Mater. 18, 203 (2019). [3] I. E. Dzyaloshinskii, On the magneto-electrical effects in antiferromagnets, Sov. Phys. JETP 10, 628 (1960). [4] D. N. Astrov, The magnetoelectric effect in antiferromag- netics, Sov. Phys. JETP 11, 708 (1960). [5] B. Huang et al., Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit, Nature (London) 546, 270 (2017). [6] C. Gong et al., Discovery of intrinsic ferromagnetism in two-dimensional van der Waals crystals, Nature (London) 546, 265 (2017). [7] K. S. Burch, D. Mandrus, and J.-G. Park, Magnetism in two- dimensional van der Waals materials, Nature (London) 563, 47 (2018). [8] M. Gibertini, M. Koperski, A. F. Morpurgo, and K. S. Novoselov, Magnetic 2D materials and heterostructures, Nat. Nanotechnol. 14, 408 (2019). [9] B. Huang et al., Electrical control of 2D magnetism in bilayer CrI3, Nat. Nanotechnol. 13, 544 (2018). [10] S. Jiang, L. Li, Z. Wang, K. Mak, and J. Shan, Controlling magnetism in 2D CrI3 by electrostatic doping. Nat. Nano- technol. 13, 549 (2018). [11] Z. Sun et al., Giant nonreciprocal second harmonic gen- eration from layered antiferromagnetism in bilayer CrI3, Nature (London) 572, 497 (2019). [12] X. Li, T. Cao, Q. Niu, J. Shi, and J. Feng, Coupling the valley degree of freedom to antiferromagnetic order, Proc. Natl. Acad. Sci. U.S.A. 110, 3738 (2013). [13] N. Sivadas, M.W. Daniels, R. H. Swendsen, S. Okamoto, and D. Xiao, Magnetic ground state of semiconducting transition-metal trichalcogenide monolayers, Phys. Rev. B 91, 235425 (2015). [14] B. L. Chittari, Y. Park, D. Lee, M. Han, A. H. MacDonald, E. Hwang, and J. Jung, Electronic and magnetic properties of single-layer MPX3 metal phosphorous trichalcogenides, Phys. Rev. B 94, 184428 (2016). [15] J.-U. Lee, S. Lee, J. H. Ryoo, S. Kang, T. Y. Kim, P. Kim, C.-H. Park, J.-G. Park, and H. Cheong, Ising-type magnetic ordering in atomically thin FePS3, Nano Lett. 16, 7433 (2016). [16] A. R. Wildes et al., Magnetic structure of the quasi-two- dimensional antiferromagnet NiPS3, Phys. Rev. B 92, 224408 (2015). [17] D. Lancon, H. C. Walker, E. Ressouche, B. Ouladdiaf, K. C. Rule, G. J. McIntyre, T. J. Hicks, H. M. Rønnow, and A. R. Wildes, Magnetic structure and magnon dynamics of the quasi-two-dimensional antiferromagnet FePS3, Phys. Rev. B 94, 214407 (2016). [18] E. Ressouche, M. Loire, V. Simonet, R. Ballou, A. Stunault, and A. Wildes, Magnetoelectric MnPS3 as a candidate for ferrotoroidicity, Phys. Rev. B 82, 100408(R) (2010). [19] K. Kim et al., Antiferromagnetic ordering in van der Waals two-dimensional magnetic material MnPS3 probed by Raman spectroscopy, 2D Mater. 6, 041001 (2019). [20] W. Xing, L. Qiu, X. Wang, Y. Yao, Y. Ma, R. Cai, S. Jia, X. C. Xie, and W. Han, Magnon Transport in Quasi-Two- Dimensional van der Waals Antiferromagnets, Phys. Rev. X 9, 011026 (2019). [21] M. Fiebig, V. V. Pavlov, and R. V. Pisarev, Second-harmonic generation as a tool for studying electronic and magnetic structures of crystals: Review, J. Opt. Soc. Am. B 22, 96 (2005). [22] Y. R. Shen, The Principles of Nonlinear Optics (Wiley, Hoboken, New Jersey, 2003). [23] V. Grasso, F. Neri, P. Perillo, L. Silipigni, and M. Piacentini, Optical-absorption spectra of crystal-field transitions in MnPS3 at low temperatures, Phys. Rev. B 44, 11060 (1991). [24] S. Lee, K.-Y. Choi, S. Lee, B. H. Park, and J.-G. Park, Tunneling transport of mono- and few-layers magnetic van der Waals MnPS3, APL Mater. 4, 086108 (2016). [25] See Supplemental Material at http://link.aps.org/ supplemental/10.1103/PhysRevLett.124.027601 for ex- tended derivations and experimental details, which includes Refs. [26–28]. [26] L. Zhao, C. A. Belvin, R. Liang, D. A. Bonn, W. N. Hardy, N. P. Armitage, and D. Hsieh, Evidence of an odd-parity hidden order in a spin-orbit coupled correlated iridate, Nat. Phys. 12, 32 (2016). [27] A.-Y. Lu et al., Janus monolayers of transition metal dichalcogenides, Nat. Nanotechnol. 12, 744 (2017). [28] G. De Luca, N. Strkalj, S. Manz, C. Bouillet, M. Fiebig, and M. Trassin, Nanoscale design of polarization in ultrathin ferroelectric heterostructures, Nat. Commun. 8, 1419 (2017). [29] C. Murayama, M. Okabe, D. Urushihara, T. Asaka, K. Fukuda, M. Isobe, K. Yamamoto, and Y. Matsushita, Crystallographic features related to a van der Waals cou- pling in the layered chalcogenide FePS3, J. Appl. Phys. 120, 142114 (2016). PHYSICAL REVIEW LETTERS 124, 027601 (2020) 027601-5 [30] M. Campostrini, M. Hasenbusch, A. Pelissetto, P. Rossi, and E. Vicari, Critical exponents and equation of state of the three-dimensional Heisenberg universality class, Phys. Rev. B 65, 144520 (2002). [31] D. Shcherbakov et al., Raman spectroscopy, photocatalytic degradation, and stabilization of atomically thin chromium tri-iodide, Nano Lett. 18, 4214 (2018). [32] M. Fiebig, D. Frohlich, G. Sluyterman v. L., and R. V. Pisarev, Domain topography of antiferromagnetic Cr2O3 by second-harmonic generation, Appl. Phys. Lett. 66, 2906 (1995). [33] R. Cheng, S. Okamoto, and D. Xiao, Spin Nernst Effect of Magnons in Collinear Antiferromagnets, Phys. Rev. Lett. 117, 217202 (2016). [34] A. R. Wildes, B. Roessli, B. Lebech, and K.W. Godfrey, Spin waves and the critical behaviour of the magnetization in MnPS3, J. Phys. Condens. Matter 10, 6417 (1998). [35] Y. Shiomi, R. Takashima, and E. Saitoh, Experimental evidence consistent with a magnon Nernst effect in the antiferromagnetic insulator MnPS3, Phys. Rev. B 96, 134425 (2017). PHYSICAL REVIEW LETTERS 124, 027601 (2020) 027601-6