[ { "title": "2005.04572v1.Anisotropic_Magnetocaloric_Properties_of_The_Ludwigite_Single_Crystal_Cu2MnBO5.pdf", "content": "1 \n Anisotropic Magnetocaloric Properties of The Ludwigite S ingle Crystal Cu 2MnBO 5 \nA.G. Gamzatov1,*, Y.S. Koshkid’ko2, D. C. Freitas3, E. Moshkina4, L. Bezmaternykh4 \nA.M. Aliev1, S.-C. Yu5, and M.H. Phan6 \n1Amirkhanov Institute of Physics, DSC of RAS, 367003, Makhachkala, Russia \n2Institute of Low Temperature and Structure Research, PAS, 50 -950, Wroclaw, Poland \n3Instituto de Física, Universidade Federal Fluminense, Campus da Praia Vermelha, 24210 -346 \nNiterói, RJ, Brazil \n4L.V. Kirensky Institute of Physic s SB RAS, 660036 Krasnoyarsk, Russia \n5Department of Physics , Ulsan National Institute of Science and Technology, Ulsan 44919, \nSouth Korea \n6Department of Physics, University of South Florida, 4202 East Fowler Avenue Tampa, Florida 33620, \nUSA \nAbstract \nWe present the results of a thorough study of the specific heat and magnetocaloric properties of a \nludwigite crystal Cu 2MnBO 5 over a temperature range of 60 – 350 K and in magnetic fields up to \n18 kOe. It is found that at temperatures below the Cu rie temperature ( TC ~ 92 K), CP(T)/T possesses \na linear temperature -dependent behavior, which is associated with the predominance of two -\ndimensional antiferromagnetic interactions of magnons. The temperature independence of \nCP/T=f(Т) is observed in the temperature range of 95 – 160 K, which can be attributed to the \nexcitation of the Wigner glass phase. The magnetocaloric effect (i.e. the adiabatic temperature \nchange, Tad (T,H)) was assessed through a direct measurement or an indirect method using the \nCP(T,H) data. Owing to its strong magneto crystalline anisotropy, an anisotropic MCE or the \nrotating MCE (Tadrot (T)) is observed in Cu 2MnBO 5. A deep minimum in the Tadrot (T) near the \nTC is observed and may be associated with the anisotropy of the paramagnetic susceptibility . \nKeywords: Ludwigite, Magnetocaloric effect, Specific heat, Anisotropy. \n*Corresponding authors: gamzatov_adler@mail.ru (A.G) 2 \n Ludwigites belong to quasi -low-dimensional transition metal oxyborates, which are \nprominent repr esentatives of systems with strongly correlated properties [1 -3]. The oxyborate \nCu2MnBO 5 has a monoclinic -distorted structure of ludwigite and is characterized by P21/c space \ngroup due to the presence of Cu2+ and Mn3+ cations in Jahn -Teller cations. Macros copic magnetic \nand specific heat studies have shown that this compound undergoes a magnetic phase transition \ninto the ferrimagnetic phase at TC ~ 92 K [1 -3]. The crystal -orientational dependences of \nmagnetization and magnetic susceptibility revealed the anisotropic magnetic characteristic both \nbelow the TC and in the paramagnetic regime (associated with the anisotropy of g -tensor due to \nmonoclinic dist ortions introduced by Jahn -Teller Cu2+ and Mn3+ ions [3]). The microscopic \nmagnetic structure of this ludwigite was experimentally studied using powder neutron diffraction \n[2]. It was determined that out of four nonequivalent positions occupied by magnetic ions, three \npositions are predominantly occupied by Cu2+ ions, and one by Mn3+ ions. It was also found that \nin the compound under study there was only a partial ordering of the magnetic moments in the \nferrimagnetic phase due to the small moment of Cu2+ ion in position 2a. The magnetic moments \nof Cu2+ and Mn3+ ions in the ferrimagnetic phase are antiparallel and their directions do not \ncoincide with the main crystallographic directions in the crystal [2]. The structural, magnetic and \nthermodynamic propertie s of Cu 2MnBO 5 were studied [1 -3]. However, the magneto -thermal \nresponse of the material has not been investigated to date. \nThe utilizing of the magnetocaloric effect (MCE) in magnetic cooling technology is of \ncurrent interest as it has the potential to re place conventional gas compression techniques [4, 5]. \nIn addition to its perspective cooling application, the MCE has recently been used as a useful \nresearch tool for the analysis and interpretation of competing magnetic phases and the collective \nmagnetic phenomena in a wide range of exotic magnetic materials [6 -10]. In this regard, we present \nhere results of the first comprehensive study of the anisotropic and frequency -dependent \nmagnetocaloric properties of the ludwigite crystal Cu 2MnBO 5. 3 \n The Cu 2MnBO 5 single crystals were synthesized by the flux method with the ratio of the \ninitial components Bi 2Mo 3O12:1.3B 2O3:0.7Na 2CO 3:0.7Mn 2O3:2.1CuO by spontaneous nucleation \n[3]. The Cu 2MnBO 5 sample has a plate shape, with dimension of 3x3x0.3 mm3 in the a*b*c \nconfi guration. The specific heat was measured by an ac -calorimetry. Direct measurements of the \nadiabatic temperature change Tad were carried out by the modulation method [ 11] in the direction \nof the magnetic field in the ab-plane . The application of an alternating magnetic field to the sample \ninduced a periodic change in the temperature of the sample, due to the MCE. This temperature \nchange was recorded by a differential thermo couple. The frequency of the alternating magnetic \nfield in this experiment was 0.2 Hz. An alternating magnetic field with amplitude of up to 4 kOe \nwas generated using an electromagnet and an external control power unit. The control alternating \nvoltage was supplied to the power unit from the output of the (Lock -in) amplifier SR 830. An \nalternating magnetic field of 18 kOe was created by a source of permanent magnetic field of an \nadjustable intensity manufactured by AMT & CLLC. \nFig. 1 shows the temperature de pendence of the specific heat Cp(T) for Cu 2MnBO 5. The \ninset of Fig.1 demonstrates the temperature dependence of CP/T in the range of 70 –150 K and at \nН = 0, 6.2, and 11 kOe. The Cp(T) dependence of Cu 2MnBO 5 was reported in [2], and our results \nare in good agreement with these data. Note that CP(Т) dependence exhibits a pronounced lambda \nanomaly at ТC = 92 K, associated with the ferromagnetic -paramagnetic (FM -PM) phase transition, \nwhich is suppressed upon the application of a magnetic field, while the maximum of the heat \ncapacity shifts by 4 K toward a higher temperature at H = 11 kOe. Two features of the Cp(T,H) \nbehavior are worthy of note. Below the ТС, the CP/T(T) is linear and unchanged with temperature \nin the temperature range of 95 -160 K. A similar behavior was observed for a number of \nferroborates [ 12, 13]. At T < TC, the CP/T = f(Т) is well described by the expression 𝐶𝑃=𝛼𝑇2 (the \ndashed line in Fig. 1). T his behavior has been explained by the predominance of two -dimensional \nantiferromagnetic interactions of magnons [ 12]. 4 \n Using the expressions, 𝐶𝑃=𝛼𝑇2, 𝛼=7.2𝑘𝐵32𝜋𝐷(𝑁𝐴⁄) ⁄ , where α is determined from the \nexperimental data (Fig. 1) α = 0.01118 J/mol K3, (N/A) = 3 10-9 mole/cm2 – the number of \nmagnetic ions per unit area [ 12], we can estimate the spin wave stiffness 𝐷=\n7.2𝑘𝐵32𝜋𝛼(𝑁𝐴⁄)=8.9810−59𝐽2𝑐𝑚2⁄ . Then using √𝐷=𝑧𝐽𝑆𝑎 , we have obtained the numerical \nvalue of the exchange integral from the Cp(T) data at T 1W \noperation, Cryogenics 20, 467 -471 (1980) . \n26. S. V. Vonsovsky, Magnetism (John Wiley, New York, 1974), Vol. 2. \n27. Vittorio Basso, The magnetocaloric effect at the first -order magneto -elastic phase transition, \nJ. Phys.: Condens. Matter 23, 226004 (2011) . \n28. Nikitin, S.A., Ivanova, T.I., Zvonov, A.I., Koshkid'ko, Y.S., Ćwik, J. & Rogacki, K., \nMagnetization, magnetic anisot ropy and magnetocaloric effect of the Tb 0.2Gd0.8 single crystal \nin high magnetic fields up to 14T in region of a phase transition, Acta Materialia 161, 331 \n(2018) . \n29. Hui Xing, Gen Long, Hanjie Guo, Youming Zou, Chunmu Feng, Guanghan Cao, Hao Zeng \nand Zhu -An Xu, Anisotropic paramagnetism of monoclinic Nd 2Ti2O7 single crystals J. Phys.: \nCondens. Matter 23, 216005 (2011) . \n \n \n 13 \n \nFig. 1. Temperature dependence of CP/T for Cu 2MnBO 5 at Н = 0, 6.2, and 11 kOe. Inset (i ) shows \nthe temperature dependences of CP (blue) and the entropy change S (red line). Inset (i i) shows \nthe temperature dependence of an enlarged anomalous part of the CP/T (points). \n \na) b) \nFig. 2. a) Temperature dependence of MCE ( ∆Tad) in a magnetic field of 18 kOe for two \norientations of the crystal with respect to the applied magnetic field. The inset of Fig. 3(a) shows \nthe ΔT||/ΔT⊥(T) dependence in a field of 18 kOe and the orientation of the crystal in a magnetic \nfield. b) The adiabatic temperature change ΔTrot of Cu 2MnBO 5 caused by a 90o rotation of the \n0 50 100 150 200 250 300 350050100150200250\nii)\n80 90100 110 120 130 140 1500.000.020.040.060.080.10CP (J/mol K)\nT (K)0.00.10.20.30.40.5\n S (J/mol K)CP (J/mol K)\nT (K)TC=91 Ki)\n80 90100 110 120 130 140 1500.750.800.850.900.95\n H=0\n 6.2 kOe\n 11 kOeCP/T (J/mol K2)\nT (K)C/T=aT\n80 90 100 110 120 1300.00.10.20.30.40.5\n H IIa\n H ^aTad (K)\nT (K)H=18 kOeCu2MnBO5\n80 90 100 110123456\nTII/T^\nT (K)H\na\nb\n80 90 100 110 120 1300.000.050.100.150.200.25Trot (K)\nT (K)TC~92 K\nH=18 kOe0 90 180 270 3600.050.100.150.200.25Tad (K)\nqT=80 K14 \n crystal as shown in the inset of Fig. 3(a). The inset of Fig. 3(b) shows the angular dependence of \n∆Tad at T = 80 K in a field of 18 kOe. \n \n \n \nFig. 3. CP(T)/T dependence at Н = 0 and 18 kOe for two orientations of the crystal with respect \nto the magnetic field direction, Ha and H^a. The inset shows the 𝛥𝑆𝑟𝑜𝑡(𝑇) at H = 18 kOe. \n \n \n80 85 90 95 100 105 110 115 1200.800.850.900.95\n H=0\n Hчч a\n H^ aCP/T (J/mol K2)\nT (K)80 90 100 110 1200.00.20.40.6Srot (J/kg K)\nT (K)" }, { "title": "2005.14559v1.Electron_spin_resonance_and_ferromagnetic_resonance_spectroscopy_in_the_high_field_phase_of_the_van_der_Waals_magnet_CrCl__3_.pdf", "content": "Electron spin resonance and ferromagnetic resonance spectroscopy\nin the high-\feld phase of the van der Waals magnet CrCl 3\nJ. Zeisner,1, 2,\u0003K. Mehlawat,1, 3,\u0003A. Alfonsov,1M. Roslova,4\nT. Doert,5A. Isaeva,1, 2, 3B. B uchner,1, 2, 3and V. Kataev1\n1Leibniz IFW Dresden, D-01069 Dresden, Germany\n2Institute for Solid State and Materials Physics, TU Dresden, D-01062 Dresden, Germany\n3W urzburg-Dresden Cluster of Excellence ct.qmat, Germany\n4Department of Materials and Environmental Chemistry,\nStockholm University, SE-106 91, Stockholm, Sweden\n5Faculty of Chemistry and Food Chemistry, TU Dresden, D-01062 Dresden, Germany\n(Dated: June 1, 2020)\nWe report a comprehensive high-\feld/high-frequency electron spin resonance (ESR) study on sin-\ngle crystals of the van der Waals magnet CrCl 3. This material, although being known for quite a\nwhile, has received recent signi\fcant attention in a context of the use of van der Waals magnets in\nnovel spintronic devices. Temperature-dependent measurements of the resonance \felds were per-\nformed between 4 and 175 K and with the external magnetic \feld applied parallel and perpendicular\nto the honeycomb planes of the crystal structure. These investigations reveal that the resonance line\nshifts from the paramagnetic resonance position already at temperatures well above the transition\ninto a magnetically ordered state. Thereby the existence of ferromagnetic short-range correlations\nabove the transition is established and the intrinsically two-dimensional nature of the magnetism\nin the title compound is proven. To study details of the magnetic anisotropies in the \feld-induced\ne\u000bectively ferromagnetic state at low temperatures, frequency-dependent ferromagnetic resonance\n(FMR) measurements were conducted at 4 K. The observed anisotropy between the two magnetic-\n\feld orientations is analyzed by means of numerical simulations based on a phenomenological theory\nof FMR. These simulations are in excellent agreement with measured data if the shape anisotropy\nof the studied crystal is taken into account, while the magnetocrystalline anisotropy is found to be\nnegligible in CrCl 3. The absence of a signi\fcant intrinsic anisotropy thus renders this material as a\npractically ideal isotropic Heisenberg magnet.\nI. INTRODUCTION\nMagnetic van der Waals materials belong to a class of\nphysical systems that currently receive considerable at-\ntention in solid state and materials research [1{6]. As a\ncommon feature these materials crystallize in a layered\nstructure with the individual layers being separated by\nthe so-called van der Waals gap, see Fig. 1. The weak\nvan der Waals coupling between the adjacent layers re-\nsults in dominating magnetic interactions within the lay-\ners while magnetic couplings between the layers remain\nrelatively weak. Thus, van der Waals magnets can be\nconsidered as quasi-two-dimensional magnetic systems.\nThese systems enable experimental studies of the spe-\nci\fc magnetic properties arising from the interplay of\nan e\u000bectively reduced dimensionality and the respective\nsingle-ion anisotropies determined by the type of mag-\nnetic ions and their local environments, see for instance\n[4]. Moreover, the weak van der Waals bonds between the\nlayers allow mechanical exfoliation of bulk crystals down\nto the few-layer or even monolayer limit [2, 3], thereby\napproaching the experimental systems to the true two\ndimensional (2D) limit. In addition, the combination of\nvarious materials sharing similar layer structures and the\nweak van der Waals couplings between the layers open up\n\u0003These authors contributed equally to this work.numerous possibilities for the creation of (magnetic) van\nder Waals heterostructures [5{7]. These stacks of sev-\neral few-layer crystals of di\u000berent materials are a promis-\ning route towards electronic devices with speci\fcally tai-\nlored magnetic properties, see, e.g., Ref. [6] and refer-\nences therein. In both respects { the fundamental study\nof 2D magnetism and its application in the framework of\nmagnetic heterostructures { the characterization of mag-\nnetic anisotropies in van der Waals materials represents a\nkey task. Magnetic resonance spectroscopies are valuable\ntools to accomplish this task due to their high sensitivity\nwith respect to the presence of magnetic anisotropies in\na system. As an example, in a previous work [8] some\nof the authors quantitatively investigated the magnetic\nanisotropies in the van der Waals magnet Cr 2Ge2Te6by\nmeans of electron spin resonance (ESR) and ferromag-\nnetic resonance (FMR) studies. Here, we report a char-\nacterization of the (e\u000bective) magnetic anisotropy in the\nhigh-\feld phase of the related compound CrCl 3. The last\nis a member of the family of transition-metal trihalides\nwhose magnetic ions (here: Cr3+, 3d3,S= 3=2,L= 3)\nare situated in the center of an octahedron built by the\nhalogen ligands (here: Cl\u0000ions). These octahedra form\nedge-sharing networks in the crystallographic abplane\nwhich e\u000bectively results in a magnetic honeycomb lat-\ntice, see Fig. 1.\nIt is worthwhile to mention that the title compound\nbelongs to the \frst materials studied using the ESR tech-arXiv:2005.14559v1 [cond-mat.mtrl-sci] 29 May 20202\nFIG. 1. Crystal and magnetic structures of CrCl 3at low temperatures. (a) View along the b-axis of the unit cell (non-primitive\nhexagonal) of the rhombohedral low-temperature structure (space group R3 [9]). The magnetic Cr3+ions (red spheres) are\noctahedrally coordinated by the Cl ligands (green spheres). (b) These CrCl 6octahedra build an edge-sharing network in the\nabplane which e\u000bectively leads to a honeycomb-like arrangement of the magnetic ions (illustrated by solid red lines). Each\nof these honeycomb layers in the abplane is well separated from its neighbors along the caxis by the van der Waals gap,\nas shown in (a). At temperatures below the transition into a magnetically long-range ordered state, spins in the honeycomb\nlayers are coupled ferromagnetically and oriented within the ab-plane, while spins in neighboring layers are coupled by a weaker\nantiferromagnetic interaction [10, 11]. The resulting spin structure in the magnetically ordered state at zero magnetic \feld\n[10, 11] is schematically illustrated by the arrows at the Cr sites. Crystallographic data are taken from Ref. [9].\nnique by E. K. Zavoisky, the pioneer of this spectroscopy\n(see, for instance, [12, 13]). However, in the context of\nmagnetic van der Waals materials the magnetic prop-\nerties of CrCl 3came back to the focus of current re-\nsearch interest [11, 14{23]. Basic magnetic properties\nof bulk CrCl 3crystals have been reported in the past\ndecades, see, e.g., Refs. [11, 16, 24{28]. In particular,\na two-step transition into a magnetically long-range or-\ndered state was observed [11, 16, 28]. Below a temper-\natureT2D\nc\u001817 K, spins within the honeycomb layers\norder ferromagnetically and align parallel to the honey-\ncomb plane, i.e., the abplane [cf. Fig. 1(b)], while spins\nof individual Cr layers are not coupled to spins in the\nneighboring layers [11, 16, 28]. Consequently, the order-\ning at around 17 K is of an e\u000bectively 2D nature. The\nferromagnetic character of the dominant exchange inter-\nactions between spins in the abplane is also evidenced\nby the positive Curie-Weiss temperatures \u0002 CWbetween\n27 and 43 K derived from measurements of the static sus-\nceptibility [11, 16, 17, 25, 29, 30]. Below a temperature\nT3D\nNof about 14 K [11, 16] (15.5 K in Ref. [28]) CrCl 3en-\nters into a three-dimensional (3D) antiferromagnetically\nordered state at zero magnetic \feld. In this magnetic\nphase, ferromagnetically ordered spins in the honeycomb\nlayers are coupled by antiferromagnetic interactions be-\ntween neighboring layers [10], as illustrated in Fig. 1(a).\nHowever, the long range antiferromagnetic order can be\nsuppressed already in relatively small magnetic \felds of\nabout 0.6 T (external \feld applied perpendicular to thehoneycomb planes, i.e., Hkc) and 0.25 T (external \feld\napplied in the honeycomb planes, i.e., H?c) at 2 K [11],\nrespectively. Above these \felds, a saturation of the mag-\nnetization at around 3 \u0016B/Cr was observed [11, 16]. The\nlow values of the saturation \felds thus con\frm the (rel-\native) weakness of the antiferromagnetic interlayer cou-\nplings. Moreover, if demagnetization e\u000bects are taken\ninto account, the saturation \felds become almost identi-\ncal for both orientations of the external \feld with respect\nto the honeycomb layers [11, 16]. Therefore, the exper-\nimentally observed magnetic anisotropies appear to be\ndominated by dipole-dipole interactions which are at the\norigin of demagnetization \felds and the so-called shape\nanisotropy [11, 16]. The apparent size of the magnetic\nanisotropy thus depends strongly on the dimensions of\nthe studied samples whereas the intrinsic magnetocrys-\ntalline is expected to be much weaker, if it exists at all.\nIn order to disentangle and quantify the two contribu-\ntions to an anisotropic magnetic response, which could\nbe of relevance in the case of CrCl 3, we studied in this\nwork the details of the magnetic anisotropies in the \feld-\ninduced ferromagnetic-like phase of the title compound\nby means of high-\feld/high-frequency (HF) FMR over\na wide range of frequencies at magnetic \felds exceeding\nthe low-temperature saturation \felds of CrCl 3.3\nII. SAMPLES AND EXPERIMENTAL\nMETHODS\nSingle-crystalline CrCl 3samples used in this study\nwere synthesized by a chemical vapor transport reac-\ntion between Cr metal and Cl 2gas as described in detail\nin Ref. [29]. The results of single-crystal X-ray di\u000brac-\ntion studies at room temperature con\frming the mon-\noclinic (space group C2=m) high-temperature modi\fca-\ntion of the title compound are also reported there. More-\nover, magnetic properties of the samples were studied\nin Ref. [16] by means of speci\fc heat as well as static\nand dynamic magnetization measurements. These are\nconsistent with the \fndings reported in previous studies\n[11, 26, 28], in particular, they con\frm the presence of\ntwo successive magnetic phase transitions mentioned in\nthe previous section. This does not only demonstrate the\nhigh quality of the used CrCl 3crystals but also ensures\nthe comparability of the magnetic properties reported in\nthis work and in the literature [11, 23, 26, 28]. Additional\ncharacterization details can be found in the Appendix.\nThe compound CrCl 3is known to undergo a struc-\ntural phase transition at temperatures around 240 K from\nthe high-temperature monoclinic phase to an rhombo-\nhedral phase (space group R3) at lower temperatures\n[9, 11]. These two structural modi\fcations di\u000ber mainly\nin the stacking sequence of the honeycomb layers while\nthe structure within the layers is very similar in both\nphases [11]. The crystal structure of the rhombohedral\nmodi\fcation is shown in Fig. 1 as the focus of the present\nstudy lies on the magnetic properties at lower tempera-\ntures. In this phase, individual honeycomb layers are\nstacked along the caxis in an -ABC-sequence. Conse-\nquently, each unit cell contains three layers along the c\naxis, see Fig. 1(a). The strong chemical bonding within\ntheabplane and the comparatively weak van der Waals\ncouplings between the layers result in \rat, platelet-like\nsingle crystals allowing an easy identi\fcation of the caxis\nas the direction perpendicular to the platelet plane.\nThe ESR and FMR measurements were carried out us-\ning two di\u000berent setups. For continuous wave (cw) HF-\nESR/HF-FMR studies a homemade spectrometer was\nemployed. The spectrometer consists of a network vec-\ntor analyzer (PNA-X from Keysight Technologies) for\ngeneration and detection of microwaves in the frequency\nrange from 20 to 330 GHz, oversized waveguides, and a\nsuperconducting solenoid (Oxford Instruments) provid-\ning magnetic \felds up to 16 T. The magnetocryostat is\nequipped with a variable temperature insert that en-\nables measurements in the temperature range between\n1.8 and 300 K. All HF measurements presented in the\nfollowing section were carried out in transmission ge-\nometry employing the Faraday con\fguration. In addi-\ntion, cw ESR/FMR measurements at a \fxed microwave\nfrequency of about 9.6 GHz and in magnetic \felds up\nto 0.9 T were performed using a commercial X-band\nspectrometer (EMX from Bruker). This spectrome-\nter is equipped with a helium \row cryostat (ESR900from Oxford Instruments) and a goniometer, allowing\ntemperature-dependent measurements in the range 4 -\n300 K and angle-dependent studies, respectively.\nIII. RESULTS AND DISCUSSION\nIn the following, results of systematic ESR measure-\nments are presented and discussed. These measurements\nwere carried out with the external magnetic \feld Hap-\nplied parallel and perpendicular to the crystallographic\ncaxis, respectively. Since the focus of the present study\nlies on a detailed investigation of the (e\u000bective) magnetic\nanisotropies in the \feld-polarized ferromagnetic state of\nCrCl 3at low temperatures, this work is mainly con-\ncerned with the behavior of the resonance \feld Hresas a\nfunction of temperature and microwave frequency over a\nbroad frequency range. Further aspects of the magnetic\nproperties of CrCl 3derived from ESR measurements at\nlower microwave frequencies, such as the excitations of\nthe 3D antiferromagnetic state and the spin dynamics\nabove the magnetic phase transitions, were reported in\nRefs. [23, 31].\nA. Temperature dependence\nThe temperature dependence of the resonance shift\n\u000eH(T) =Hres(T)\u0000Hres(100K) measured at low and\nhigh microwave frequencies \u0017and in both magnetic \feld\ncon\fgurations is shown in Fig. 2. This quantity is a mea-\nsure of the deviation of the resonance \feld at a given\ntemperature Tfrom the ideal paramagnetic resonance\nposition. This position is determined by the standard\nresonance condition of a paramagnet [32]\n\u0017=g\u0016B\u00160Hres=h : (1)\nHere,gdenotes the gfactor of the resonating spins and\n\u0016B,\u00160, andhare Bohr's magneton, the vacuum per-\nmeability, and Planck's constant, respectively. In the\npresent case, the resonant shift was determined with re-\nspect to the expected resonance \feld at 100 K which was\ncalculated according to Eq. (1) using the gfactor de-\nrived from frequency-dependent measurements at 100 K,\nsee below. Upon lowering the temperature below \u001875 K,\nthe resonance position is shifted progressively to smaller\n\felds when the external magnetic \feld is applied perpen-\ndicular to the caxis, resulting in a negative shift \u000eH.\nForHkcthis trend is reversed yielding a positive res-\nonance shift. Thus, based on the qualitative behavior\nof the temperature-dependent resonance shift, it can be\nconcluded that the experimentally observable (e\u000bective)\nmagnetic easy direction lies within the abplane which is\nin agreement with the proposed spin structure [10] and\nprevious magnetization measurements [11, 16, 26]. More-\nover, the\u000eH(T) curves obtained for HkcandH?c\nare asymmetric with respect to the \u000eH= 0 line of an4\nideal paramagnet, see Fig. 2. This asymmetry is a direct\nconsequence of the di\u000berent frequency-\feld dependencies\n\u0017(Hres) expected for the two di\u000berent \feld orientations in\na ferromagnetically ordered system [33, 34]. In particu-\nlar, a shift of the resonance line should be larger when the\nexternal magnetic \feld is oriented parallel to the mag-\nnetic anisotropy axis which in the case of CrCl 3is the\nmagnetic hard axis parallel to the crystallographic caxis.\nThe onset of a \fnite resonance shift already at temper-\natures signi\fcantly above the ordering temperature T2D\nc\nprovides clear evidence for the low-dimensional character\nof the spin correlations in this material as it was discussed\nin the context of one-dimensional systems, for instance in\nRefs. [35{37]. Moreover, the 2D nature of magnetic cor-\nrelations above T2D\ncwas reported in Ref. [31] based on\nmeasurements of the resonance shift and the linewidth\nangular dependence at various temperatures and at low\nmicrowave frequencies of about 9.4 GHz. While our mea-\nsurements of \u000eH(T) at 9.6 GHz are consistent with this\nprevious study [31], we observed the onset of a \fnite\nresonance shift at higher temperatures when employing\nfrequencies of about 90 GHz. Thus, the higher exter-\nnal magnetic \felds associated with the higher microwave\nfrequencies strengthen the ferromagnetic correlations re-\nsponsible for the shift of the resonance line, similar to\nthe situation found in the related van der Waals magnet\nCr2Ge2Te6[8]. Finally, it is worthwhile mentioning that\nthe appearance of short-range correlations at tempera-\ntures far above the magnetic ordering temperature is con-\nsistent with the deviation of the temperature-dependent\nstatic susceptibility from a Curie-Weiss behavior already\nbelow\u0018125 K, which was reported in Ref. [29]. Taken to-\ngether, temperature-dependent measurements of the res-\nonance shift at two di\u000berent frequencies and \feld orien-\ntations demonstrate, \frst, the apparent easy-plane type\nmagnetic anisotropy and, second, the 2D character of dy-\nnamic spin correlations in CrCl 3far above the long-range\nordering temperatures.\nB. Frequency dependence\nTo shed light on the details of the magnetic\nanisotropies, frequency-dependent investigations were\nconducted at 4 and 100 K, i.e., at temperatures deep in\nthe magnetically ordered state and well above the mag-\nnetic phase transition, respectively. Exemplary spectra\nare presented in the insets of Fig. 3. At both tem-\nperatures, ESR/FMR spectra consist of a single nar-\nrow resonance line with typical linewidths (full width\nat half maximum) of about 25 mT at 100 K. The small\nlinewidths allow an easy and precise determination of\nthe resonance \felds which correspond to the positions\nof the minima in the microwave transmission. The re-\nsulting frequency-\feld diagrams are shown in the main\npanels of Fig. 3. At 100 K a linear frequency-\feld de-\npendence\u0017(Hres) is observed for both orientations of the\nexternal magnetic \feld. Such a behavior is expected in\n/s48 /s50/s53 /s53/s48 /s55/s53 /s49/s48/s48 /s49/s50/s53 /s49/s53/s48 /s49/s55/s53/s45/s50/s48/s48/s45/s49/s48/s48/s48/s49/s48/s48/s50/s48/s48/s51/s48/s48/s52/s48/s48\n/s32/s72 /s32/s124 /s124 /s32 /s99 /s32/s58/s32 /s110 /s32/s61/s32/s57/s48/s46/s48/s32/s71/s72/s122\n/s32/s72 /s32/s124 /s124 /s32 /s99/s32 /s58/s32 /s110 /s32/s61/s32/s57/s46/s54/s32/s71/s72/s122\n/s32/s72 /s32/s94 /s32/s99 /s32/s58/s32 /s110 /s32/s61/s32/s57/s48/s46/s49/s32/s71/s72/s122\n/s32/s72 /s32/s94 /s32/s99/s32 /s58/s32 /s110 /s32/s61/s32/s57/s46/s54/s32/s71/s72/s122/s109\n/s48/s40/s72\n/s114/s101/s115/s40/s84 /s41/s32/s45/s32 /s72\n/s114/s101/s115/s40/s49/s48/s48/s32/s75/s41/s41/s32/s40/s109/s84/s41\n/s84 /s32/s40/s75/s41/s84/s32/s50/s68\n/s67/s84/s32/s51/s68\n/s78FIG. 2. Temperature dependence of the resonance shift\n\u000eH(T) =Hres(T)\u0000Hres(100K) at microwave frequencies\nof about 9.6 and 90 GHz (open and \flled symbols, respec-\ntively). In these measurements the external magnetic \feld\nwas oriented parallel (red circles) as well as perpendicular\n(blue squares) to the caxis. The dashed horizontal line rep-\nresents the zero shift \u000eH= 0 expected for an uncorrelated\nparamagnet. Vertical dashed lines indicate the zero-\feld tran-\nsition temperatures T2D\ncandT3D\nNof the transition into the 2D\nferromagnetically ordered phase and the 3D antiferromagnet-\nically ordered state [11], respectively.\nthe paramagnetic regime of CrCl 3and can be well de-\nscribed by the standard resonance condition of a param-\nagnet [Eq. (1)] [32]. Fits to the data according to Eq. (1)\nare shown in Fig. 3(a) as solid lines and yielded gfactors\ngk= 1:970\u00060:005 andg?= 1:990\u00060:005 forHkc\nandH?c, respectively. The experimentally determined\ngfactors are in good agreement with the values antici-\npated for Cr3+ions (3d3,S= 3=2,L= 3) in an octa-\nhedral crystal \feld [32]. The merely small deviation of g\nfrom the free-electron gfactor of 2 as well as the slight\nanisotropy observed in the measurements indicate that\nthe magnetism in CrCl 3is largely dominated by the spin\ndegrees of freedom while the orbital angular momentum\nis practically completely quenched in the \frst order. The\nmentioned small deviations from the ideal spin-only be-\nhavior result, most likely, from second-order spin-orbit\ncoupling e\u000bects [32]. The gfactors derived in this work\nfrom frequency-dependent measurements are, moreover,\nconsistent with the saturation magnetization of about\n3\u0016B/Cr which was experimentally observed [11, 16] and\nis theoretically expected for spins S= 3=2 and agfac-\ntor of 2. These observations already suggest that spin-\norbit coupling and, consequently, the intrinsic magne-\ntocrystalline anisotropies are very weak (or even negligi-\nble) in CrCl 3, as it was also mentioned in the literature\n[11, 16, 26].5\n/s48 /s49 /s50 /s51 /s52 /s53 /s54 /s55 /s56 /s57 /s49/s48 /s49/s49/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s50/s53/s48/s51/s48/s48/s51/s53/s48/s52/s48/s48/s110 /s32/s40/s71/s72/s122/s41\n/s109\n/s48/s32/s72\n/s114/s101/s115/s32/s40/s84/s41/s32/s72 /s32/s124/s124/s32 /s99\n/s32/s72 /s32/s94 /s32/s99/s84 /s32/s61/s32/s52/s32/s75/s49/s46/s53 /s49/s46/s54 /s53/s46/s52 /s53/s46/s54 /s57/s46/s54 /s57/s46/s57 /s49/s48/s46/s50\n/s78/s111/s114/s109/s97/s108/s105/s122/s101/s100/s32\n/s115/s105/s103/s110/s97/s108/s32/s105/s110/s116/s101/s110/s115/s105/s116/s121\n/s109\n/s48/s32/s72 /s32/s40/s84/s41/s51/s52/s32/s71/s72/s122 /s49/s52/s52/s32/s71/s72/s122 /s50/s54/s50/s32/s71/s72/s122\n/s48 /s49 /s50 /s51 /s52 /s53 /s54 /s55 /s56 /s57 /s49/s48 /s49/s49 /s49/s50 /s49/s51/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s50/s53/s48/s51/s48/s48/s51/s53/s48\n/s32/s72 /s32/s124/s124/s32 /s99/s32 /s58/s32 /s103 /s32/s61/s32/s49/s46/s57/s55/s48/s32 /s177 /s32/s48/s46/s48/s48/s53/s32\n/s32/s72 /s32/s94 /s32/s99/s32 /s58/s32 /s103 /s32/s61/s32/s49/s46/s57/s57/s48/s32 /s177 /s32/s48/s46/s48/s48/s53/s110 /s32/s40/s71/s72/s122/s41\n/s109\n/s48/s32/s72\n/s114/s101/s115/s32/s40/s84/s41/s84 /s32/s61/s32/s49/s48/s48/s32/s75\n/s51/s46/s48 /s51/s46/s53 /s55/s46/s48 /s55/s46/s53 /s49/s48/s46/s48 /s49/s48/s46/s53/s109\n/s48/s32/s72 /s32/s40/s84/s41/s78/s111/s114/s109/s97/s108/s105/s122/s101/s100/s32\n/s115/s105/s103/s110/s97/s108/s32/s105/s110/s116/s101/s110/s115/s105/s116/s121/s57/s48/s32/s71/s72/s122\n/s49/s57/s55/s32/s71/s72/s122/s50/s55/s53/s32/s71/s72/s122/s40/s97 /s41 /s40/s98 /s41\nFIG. 3. Frequency dependence of the resonance \feld Hresat 100 K (a) and 4 K (b), respectively. To quantitatively investigate\nthe magnetic anisotropies in CrCl 3, measurements were carried out with the external magnetic \feld applied parallel (red circles)\nand perpendicular (blue squares) to the crystallographic caxis, respectively. Open symbols correspond to the resonance \felds\nmeasured at \u0017= 9:6 GHz. Solid lines in (a) are linear \fts to the \u0017(Hres) dependencies according to the standard resonance\ncondition of a paramagnet given in Eq. (1) [32]. At 4 K, the measured frequency-\feld dependencies were simulated using\nEq. (2) which describes the resonance frequency in the case of FMR. Corresponding simulations are shown in (b) as solid\nlines. Exemplary spectra recorded with Hkcare presented for both temperatures in the respective insets. For comparison, all\nspectra shown were normalized (note the di\u000berent breaks in the \feld axes). Arrows in the \u0017(Hres) denote the positions of the\nspectra given in the insets.\nFor a detailed analysis of the relevant anisotropies\npresent in the title compound, frequency-dependent mea-\nsurements were conducted at 4 K, i.e., in the magnetically\nordered state. We emphasize that the lowest frequency\nused in the systematic HF measurements corresponds to\na resonance \feld of about 0.75 T [cf. Fig. 3(b)] which\nexceeds the reported saturation \felds [26]. Therefore,\nthe \feld-polarized, e\u000bectively 2D ferromagnetic state of\nCrCl 3is probed by our HF magnetic resonance measure-\nments allowing us to describe the obtained \u0017(Hres) de-\npendencies in the framework of a theory of FMR that\nis applicable in a single-domain ferromagnetic material,\nas will be discussed in the following section. In contrast\nto the measurements in the paramagnetic state, the 4 K\ndata show a clear anisotropy regarding the two magnetic-\n\feld orientations which is consistent with the resonance\nshifts of opposite sign observed in the temperature-\ndependent studies (see Fig. 2). If the external magnetic\n\feld is applied within the abplane (H?c), the resonance\npositions are systematically shifted towards smaller mag-\nnetic \felds (with respect to the paramagnetic resonance\npositions) for all measured frequencies. For Hkca neg-\native intercept with the frequency axis of the \u0017(Hres)\ndiagram is observed due to the shift of the resonance po-\nsitions to higher magnetic \felds. Qualitatively, such a\nbehavior is expected in the case of a ferromagnetically\nordered system with an easy-plane anisotropy [34, 38].\nIn a semi-classical picture, these shifts of the resonanceposition result from anisotropic internal \felds in the mag-\nnetic crystal. These \felds are caused, in turn, by dipole-\ndipole interactions between the magnetic moments asso-\nciated with the spins or an intrinsic magnetocrystalline\nanisotropy, i.e., by spin-orbit coupling. In the following\nsection the resonance \felds in the ferromagnetic state\nwill be simulated employing a theory based on this semi-\nclassical description of magnetic resonance. This allows\nto determine quantitatively the contributions of the two\npossible sources of magnetic anisotropy in CrCl 3.\nC. Analysis of the magnetic anisotropies\nAs mentioned in the previous sections, it is possi-\nble to describe the experimentally observed frequency-\n\feld dependence in the ferromagnetic state at 4 K in the\nframework of a semi-classical, phenomenological theory\nof FMR [34, 38, 39]. Conceptually, the di\u000berence between\nESR and FMR lies in the fact, that ESR is the resonant\nexcitation of individual paramagnetic spins within a mag-\nnetic system, while FMR describes the resonance of the\ntotal magnetization Min a ferromagnetically ordered\nmaterial. Thus, the resonance frequencies expected for a\ncorrelated ferromagnetic spin system can be calculated\nby considering the classical vector of the macroscopic\nmagnetization and the appropriate free energy density\nF[34, 38, 39]. In this case, the resonance frequency is6\ngiven by the following expression [34, 38, 39]\n\u00172\nres=g2\u00162\nB\nh2M2ssin2\u0012\u0012@2F\n@\u00122@2F\n@'2\u0000\u0010@2F\n@\u0012@'\u00112\u0013\n;(2)\nwhereMsis the saturation magnetization and 'and\u0012\ndenote the spherical coordinates of the magnetization\nvectorM(Ms;';\u0012). Note that in the present case the\nspherical coordinate system is chosen such that the z\naxis coincides with the crystallographic caxis in the low-\ntemperature structure of CrCl 3. For a calculation of the\nresonance position under given experimental conditions,\ni.e., a speci\fc orientation of the external magnetic \feld\nrelative to the studied sample and a \fxed microwave fre-\nquency, Eq. (2) has to be evaluated at the equilibrium po-\nsition ('0;\u00120) of the macroscopic magnetization vector.\nThe orientation of Min the equilibrium state is found\nnumerically by minimizing Fwith respect to the spheri-\ncal coordinates, taking into account the particular exper-\nimental conditions. Since the qualitative considerations\nregarding the temperature dependence of the resonance\nshift as well as the \u0017(Hres) dependencies at 4 K suggest\nan easy-plane type anisotropy, a uniaxial magnetocrys-\ntalline anisotropy was included in the free energy density\nterm as an initial approach to describe the anisotropies\nin CrCl 3. The free energy density is then given (in SI\nunits) by\nF=\u0000\u00160H\u0001M\u0000KUcos2(\u0012)+\n1\n2\u00160M2\ns(Nxsin2(\u0012) cos2(')+\nNysin2(\u0012) sin2(') +Nzcos2(\u0012)):(3)\nThe \frst term is the Zeeman-energy density describing\nthe coupling between the magnetization vector Mand\nthe external magnetic \feld H. The second term repre-\nsents the uniaxial magnetocrystalline anisotropy whose\nstrength is parametrized by the energy density KU. Fi-\nnally, the third contribution to Fis the shape anisotropy\nenergy density which is characterized by the demagneti-\nzation factors Nx,Ny, andNz[40, 41]. These factors are\ndetermined by the dimensions of the crystals under study.\nIn the present case, the shape of the measured platelet-\nlike CrCl 3crystal was described by the demagnetization\nfactors of an extended \rat plate ( Nx=Ny= 0;Nz= 1\n[42]), whose xyplane corresponds to the crystallographic\nabplane and the zaxis lies parallel to the caxis. This\napproximation to the real sample shape is justi\fed by\nthe platelet-like shape of the studied CrCl 3crystal whose\nlateral dimensions in the abplane (of about 400 \u0016m\u0002\n450\u0016m) are much larger than the thickness of the crys-\ntal along the caxis. Due to experimental reasons, the\nsample thickness could not be measured precisely which\nhampered a determination of the demagnetization factors\nsolely based on the sample dimensions. However, the de-\nviations between the approximated and the true demag-\nnetization factors can be expected to be small. In addi-\ntion to the sample shape, the value of the saturation mag-\nnetizationMsenters into the simulation of the frequency-\n\feld dependence. For the simulations presented in thefollowing, the reported saturation magnetization of about\n3\u0016B/Cr at 1.8 K [16] was used which corresponds to\nMs\u0019314:97\u0002103J/Tm3. Furthermore, the gfactors\ndetermined independently from the frequency-dependent\nmeasurements at 100 K and for both \feld orientations\n(see Sec. III B) were set as \fxed parameters in the sim-\nulations. Thus, the only free parameter, which was ad-\njusted to match simulated with measured data, was the\nmagnetocrystalline anisotropy energy density KU.\nThe \fnal results of the simulations are presented in\nFig. 3(b) as solid lines and show an excellent agree-\nment between the simulated and the measured resonance\npositions. Most importantly, this agreement could be\naccomplished by solely taking into account the shape\nanisotropy, i.e., by setting KUto zero. Thus, it is ulti-\nmately proven that the anisotropy observed in dynamic\nand static magnetic investigations of CrCl 3[11, 16, 23,\n26] is, indeed, due to the shape anisotropy caused by\nlong-range dipole-dipole interactions, whereas the intrin-\nsic magnetocrystalline anisotropy can be neglected. The\npresent work therefore could verify by means of highly\nsensitive HF magnetic resonance investigations the con-\nclusions of these previous studies [11, 16, 23, 26] and\nsupports the results of the recent \frst-principle calcu-\nlations of the magnetic anisotropy of CrCl 3monolayers\n[43]. It can be concluded that, intrinsically, CrCl 3is\nmagnetically isotropic while the apparent anisotropy ob-\nserved in experiments can be tuned (within certain lim-\nits) by choosing a particular sample shape. Moreover,\nCrCl 3might serve as a valuable reference system for fu-\nture magnetic resonance studies of other van der Waals\nmagnets, since it illustrates the pure e\u000bect of the shape\nanisotropy on, e.g., the frequency-\feld dependence. As\nsimilar sample shapes can be expected for this large fam-\nily of layered materials, it follows that it is very important\nto take into account this source of magnetic anisotropy\nwhen aiming at a precise quanti\fcation of magnetocrys-\ntalline anisotropies in van der Waals magnets.\nIV. CONCLUSIONS\nIn summary, we studied the details of the mag-\nnetic anisotropies in the van der Waals magnet CrCl 3\nby means of systematic HF ESR and FMR measure-\nments. By extending the frequency range of previ-\nous works [23, 31], the \feld-polarized, e\u000bectively ferro-\nmagnetic low-temperature state of the title compound\nwas investigated. Numerical simulations of the mea-\nsured frequency-\feld dependence at 4 K revealed that\nthe anisotropy observed experimentally is governed by\nthe shape anisotropy of the studied CrCl 3crystal. In\ncontrast, the intrinsic magnetocrystalline anisotropy is\nfound to be negligible in this compound, thus supporting\nthe conclusions drawn in previous studies [11, 16, 23, 26].\nConsidering the large current scienti\fc interest in mag-\nnetic van der Waals materials, CrCl 3may serve as a\nreference for future quantitative analyses of magnetic7\nanisotropies in related layered magnets, since it provides\nan excellent example for the impact of the particular sam-\nple shape on the apparent magnetic anisotropy. Finally,\ntemperature-dependent measurements of the resonance\nshift showed the onset of a \fnite shift and, thus, the\nexistence of short-range spin correlations already at tem-\nperatures well above the magnetically ordered state. This\nobservation provides further evidence for the intrinsically\n2D nature of the magnetism in the van der Waals magnet\nCrCl 3.\nConsidering a continuously increasing number of quasi-\n2D van der Waals compounds exhibiting a rich variety\nof intriguing magnetic properties, it is appealing to ap-\nply systematically the frequency-tunable high-\feld ESR\nspectroscopy to investigate the spin dynamics and in par-\nticular the magnetic anisotropy of these materials. The\nlatter appears to be a key factor for determining the type\nof a magnetically ordered ground state. For example,\nthe members of the family of the van der Waals layered\nmetal phosphorous trichalcogenides MPS 3(M = Mn, Fe,\nNi) [44] feature di\u000berent types of magnetic anisotropy\n[45] and exhibit dissimilar antiferromagnetically ordered\nground states, as, e.g., was illustrated in Ref. [46]. Quan-\nti\fcation of the parameters of magnetic anisotropy with\nESR spectroscopy may shed more light on a possible rela-\ntionship between the anistropy and the type of magnetic\norder in these compounds.\nACKNOWLEDGMENTS\nThis work was \fnancially supported by the Deutsche\nForschungsgemeinschaft (DFG) within the Collabora-\ntive Research Center SFB 1143 \\Correlated Magnetism\nFrom Frustration to Topology (project-id 247310070)\nand the Dresden-Wrzburg Cluster of Excellence (EXC\n2147) \\ct.qmat - Complexity and Topology in Quantum\nMatter\" (project-id 39085490). K. M. acknowledges the\nHallwachsRntgen Postdoc Program of ct.qmat for \fnan-\ncial support. A. A. acknowledges \fnancial support by\nthe DFG through Grant No. AL 1771/4-1.\nAppendix: Details on the sample characterization\nA typical x-ray di\u000braction (XRD) pattern of a powder\nsample of CrCl 3is shown in Fig. 4. It was collected\nusing a PANalytical XPert Pro MPD di\u000bractometer with\nCu-K\u000b1radiation (\u0015= 1:54056 \u0017A) in the BraggBrentano\ngeometry at room temperature in the 2 \u0012range between\n10 and 90\u000e. The CrCl 3sample displays strongly preferred\norientation leading to only (00 l) peaks visible.\nA typical scanning electron microscopy (SEM) image\nof CrCl 3crystallites is shown in Fig. 5. It was made with\na Hitachi SU8020 microscope equipped with an Oxford\nSilicon Drift X-MaxN energy dispersive x-ray spectrome-ter (EDX) at an acceleration voltage of 20 kV. Results of\nthe EDX analysis at selected areas indicated in Fig. 5 are\nIntensity (arb. units)\n2q (°)\nFIG. 4. XRD powder pattern of a sample of CrCl 3. Red ticks\ndisplay the peak positions of the reference CrCl 3from the\nInorganic Crystal Structure Database (22080-ICSD, SCXRD,\n298 K). Note that only (00 l) peaks are visible due to the strong\ntexturing of the powder sample.\nSpectrumSpectrum\nSpectrum\nSpectrum\nFIG. 5. SEM image of CrCl 3crystallites. Rectangles indicate\nthe regions where EDX spectra were collected.\npresented in Table I. They evidence a very homogeneous,\nclose to the ideal composition Cr : Cl = 1 : 3 of the single\ncrystal.\nTABLE I. Results of the analysis of the EDX spectra of the\nCrCl 3crystal taken at selected areas indicated in Fig. 5.\nSpectrum 826 827 828 829 Calculated\nCl (%) 74.77 74.69 74.49 74.87 75\nCr (%) 25.23 25.31 25.51 25.13 258\n1J.-G. Park, \\Opportunities and challenges of 2D magnetic\nvan der Waals materials: magnetic graphene?\" J. Phys.:\nCondens. Matter 28, 301001 (2016).\n2Cheng Gong, Lin Li, Zhenglu Li, Huiwen Ji, Alex Stern,\nYang Xia, Ting Cao, Wei Bao, Chenzhe Wang, Yuan\nWang, Z. Q. Qiu, R. J. Cava, Steven G. Louie, Jing Xia,\nand Xiang Zhang, \\Discovery of intrinsic ferromagnetism\nin two-dimensional van der Waals crystals,\" Nature 546,\n265 (2017).\n3Bevin Huang, Genevieve Clark, Efr\u0013 en Navarro-Moratalla,\nDahlia R. Klein, Ran Cheng, Kyle L. Seyler, Ding Zhong,\nEmma Schmidgall, Michael A. McGuire, David H. Cobden,\nWang Yao, Di Xiao, Pablo Jarillo-Herrero, and Xiaodong\nXu, \\Layer-dependent ferromagnetism in a van der Waals\ncrystal down to the monolayer limit,\" Nature 546, 270\n(2017).\n4Kenneth S. Burch, David Mandrus, and Je-Geun Park,\n\\Magnetism in two-dimensional van der Waals materials,\"\nNature 563, 47 { 52 (2018).\n5M. Gibertini, M. Koperski, A. F. Morpurgo, and K. S.\nNovoselov, \\Magnetic 2D materials and heterostructures,\"\nNat. Nanotechnol. 14, 408 { 419 (2019).\n6Cheng Gong and Xiang Zhang, \\Two-dimensional mag-\nnetic crystals and emergent heterostructure devices,\" Sci-\nence363, 706 (2019).\n7A. K. Geim and I. V. Grigorieva, \\Van der Waals het-\nerostructures,\" Nature 499, 419 (2013).\n8J. Zeisner, A. Alfonsov, S. Selter, S. Aswartham, M. P.\nGhimire, M. Richter, J. van den Brink, B. B uchner, and\nV. Kataev, \\Magnetic anisotropy and spin-polarized two-\ndimensional electron gas in the van der Waals ferromagnet\nCr2Ge2Te6,\" Phys. Rev. B 99, 165109 (2019).\n9Bruno Morosin and Albert Narath, \\XRay Di\u000braction\nand Nuclear Quadrupole Resonance Studies of Chromium\nTrichloride,\" J. Chem. Phys. 40, 1958{1967 (1964).\n10J.W. Cable, M.K. Wilkinson, and E.O. Wollan, \\Neutron\ndi\u000braction investigation of antiferromagnetism in CrCl 3,\"\nJ. Phys. Chem. Solids 19, 29 { 34 (1961).\n11Michael A. McGuire, Genevieve Clark, Santosh KC,\nW. Michael Chance, Gerald E. Jellison, Valentino R.\nCooper, Xiaodong Xu, and Brian C. Sales, \\Magnetic be-\nhavior and spin-lattice coupling in cleavable van der Waals\nlayered CrCl 3crystals,\" Phys. Rev. Materials 1, 014001\n(2017).\n12E. Zavoisky, \\Spin magnetic resonance in the decimetre-\nwave region,\" Journal of Physics USSR 10, 197 (1946).\n13B. I. Kochelaev and Y. V. Yablokov, The Beginning of\nParamagnetic Resonance (World Scienti\fc, 1995).\n14A. Glamazda, P. Lemmens, S.-H. Do, Y. S. Kwon, and K.-\nY. Choi, \\Relation between Kitaev magnetism and struc-\nture in\u000b\u0000RuCl 3,\" Phys. Rev. B 95, 174429 (2017).\n15Omar Besbes, Sergey Nikolaev, Noureddine Meskini, and\nIgor Solovyev, \\Microscopic origin of ferromagnetism in\nthe trihalides CrCl 3and CrI 3,\" Phys. Rev. B 99, 104432\n(2019).\n16G. Bastien, M. Roslova, M. H. Haghighi, K. Mehlawat,\nJ. Hunger, A. Isaeva, T. Doert, M. Vojta, B. B uchner, and\nA. U. B. Wolter, \\Spin-glass state and reversed magnetic\nanisotropy induced by Cr doping in the Kitaev magnet\n\u000b\u0000RuCl 3,\" Phys. Rev. B 99, 214410 (2019).\n17Mykola Abramchuk, Samantha Jaszewski, Kenneth R.Metz, Gavin B. Osterhoudt, Yiping Wang, Kenneth S.\nBurch, and Fazel Tafti, \\Controlling Magnetic and Op-\ntical Properties of the van der Waals Crystal CrCl 3\u0000xBrx\nvia Mixed Halide Chemistry,\" Adv. Mater. 30, 1801325\n(2018).\n18Wei-Bing Zhang, Qian Qu, Peng Zhu, and Chi-Hang Lam,\n\\Robust intrinsic ferromagnetism and half semiconductiv-\nity in stable two-dimensional single-layer chromium tri-\nhalides,\" J. Mater. Chem. C 3, 12457{12468 (2015).\n19Lucas Webster and Jia-An Yan, \\Strain-tunable magnetic\nanisotropy in monolayer CrCl 3, CrBr 3, and CrI 3,\" Phys.\nRev. B 98, 144411 (2018).\n20Xinghan Cai, Tiancheng Song, Nathan P. Wilson,\nGenevieve Clark, Minhao He, Xiaoou Zhang, Takashi\nTaniguchi, Kenji Watanabe, Wang Yao, Di Xiao,\nMichael A. McGuire, David H. Cobden, and Xiaodong Xu,\n\\Atomically Thin CrCl 3: An In-Plane Layered Antiferro-\nmagnetic Insulator,\" Nano Lett. 19, 3993{3998 (2019).\n21N. Bykovetz, A. Hoser, and C. L. Lin, \\Critical region\nphase transitions in the quasi-2D magnet CrCl 3,\" AIP\nAdv. 9, 035029 (2019).\n22Dahlia R. Klein, David MacNeill, Qian Song, Daniel T.\nLarson, Shiang Fang, Mingyu Xu, R. A. Ribeiro, P. C.\nCan\feld, Efthimios Kaxiras, Riccardo Comin, and Pablo\nJarillo-Herrero, \\Enhancement of interlayer exchange in an\nultrathin two-dimensional magnet,\" Nat. Phys. 15, 1255{\n1260 (2019).\n23David MacNeill, Justin T. Hou, Dahlia R. Klein, Pengxi-\nang Zhang, Pablo Jarillo-Herrero, and Luqiao Liu, \\Giga-\nhertz Frequency Antiferromagnetic Resonance and Strong\nMagnon-Magnon Coupling in the Layered Crystal CrCl 3,\"\nPhys. Rev. Lett. 123, 047204 (2019).\n24Wilford N. Hansen and Maurice Gri\u000bel, \\Heat Capacities\nof CrF 3and CrCl 3from 15 to 300\u000eK,\" J. Chem. Phys.\n28, 902{907 (1958).\n25Wilford N. Hansen and Maurice Gri\u000bel, \\Paramagnetic\nSusceptibilities of the Chromium (III) Halides,\" J. Chem.\nPhys. 30, 913{916 (1959).\n26H. Bizette, C. Terrier, and A. Adam, \\Aimantations prin-\ncipales du chlorure chromique,\" C. R. Acad. Sci. 252, 1571\n(1961).\n27Albert Narath and H. L. Davis, \\Spin-Wave Analysis of\nthe Sublattice Magnetization Behavior of Antiferromag-\nnetic and Ferromagnetic CrCl 3,\" Phys. Rev. 137, A163 {\nA178 (1965).\n28B. Kuhlow, \\Magnetic Ordering in CrCl 3at the Phase\nTransition,\" Phys. Status Solidi (a) 72, 161{168 (1982).\n29Maria Roslova, Jens Hunger, Ga el Bastien, Darius Pohl,\nHossein M. Haghighi, Anja U. B. Wolter, Anna Isaeva, Ul-\nrich Schwarz, Bernd Rellinghaus, Kornelius Nielsch, Bernd\nB uchner, and Thomas Doert, \\Detuning the Honeycomb\nof the\u000b-RuCl 3Kitaev Lattice: A Case of Cr3+Dopant,\"\nInorg. Chem. 58, 6659{6668 (2019).\n30C. Starr, F. Bitter, and A. R. Kaufmann, \\The Magnetic\nProperties of the Iron Group Anhydrous Chlorides at Low\nTemperatures. I. Experimental,\" Phys. Rev. 58, 977 { 983\n(1940).\n31S. Chehab, J. Amiell, P. Biensan, and S. Flandrois, \\Two-\ndimensional ESR behaviour of CrCl 3,\" Physica B: Con-\ndensed Matter 173, 211 { 216 (1991).\n32A. Abragam and B. Bleaney, Electron Paramagnetic Reso-9\nnance of Transition Ions (Oxford University Press, Oxford,\n2012).\n33E. A. Turov, Physical Properties of Magnetically Ordered\nCrystals , edited by A. Tybulewicz and S. Chomet (Aca-\ndemic Press, New York, 1965).\n34Michael Farle, \\Ferromagnetic resonance of ultrathin\nmetallic layers,\" Rep. Prog. Phys. 61, 755 (1998).\n35Kazukiyo Nagata and Yuichi Tazuke, \\Short Range Order\nE\u000bects on EPR Frequencies in Heisenberg Linear Chain\nAntiferromagnets,\" J. Phys. Soc. Jpn. 32, 337{345 (1972).\n36Kiichi Okuda, Hirao Hata, and Muneyuki Date, \\Elec-\ntron Spin Resonance in One Dimensional Antiferromagnet\nCu(C 6H5COO) 23H2O,\" J. Phys. Soc. Jpn. 33, 1574{1580\n(1972).\n37Kokichi Oshima, Kiichi Okuda, and Muneyuki Date, \\g-\nShift in Low Dimensional Antiferromagnets at Low Tem-\nperatures,\" Journal of the Physical Society of Japan 41,\n475{480 (1976).\n38G. V. Skrotskii and L. V. Kurbatov, \\Phenomenological\nTheory of Ferromagnetic Resonance,\" in Ferromagnetic\nResonance , edited by S. V. Vonsovskii (Pergamon Press\nLtd., Oxford, 1966) Chap. Phenomenological Theory of\nFerromagnetic Resonance, pp. 12{77.\n39J. Smit and H. G. Beljers, \\Ferromagnetic Resonance Ab-\nsorption in BaFe 12O19,\" Philips Research Reports 10, 113(1955).\n40J. A. Osborn, \\Demagnetizing Factors of the General El-\nlipsoid,\" Phys. Rev. 67, 351 (1945).\n41D. C. Cronemeyer, \\Demagnetization factors for general\nelipsoids,\" J. Appl. Phys. 70, 2911 (1991).\n42Stephen Blundell, Magnetism in Condensed Matter (Ox-\nford University Press, Oxford, 2001).\n43Feng Xue, Yusheng Hou, Zhe Wang, and Ruqian Wu,\n\\Two-dimensional ferromagnetic van der Waals CrCl 3\nmonolayer with enhanced anisotropy and Curie temper-\nature,\" Phys. Rev. B 100, 224429 (2019).\n44Fengmei Wang, To\fk A. Shifa, Peng Yu, Peng He, Yang\nLiu, Feng Wang, Zhenxing Wang, Xueying Zhan, Xiaod-\ning Lou, Fan Xia, and Jun He, \\New Frontiers on van\nder Waals Layered Metal Phosphorous Trichalcogenides,\"\nAdv. Funct. Mater. 28, 1802151 (2018).\n45P. A. Joy and S. Vasudevan, \\Magnetism in the layered\ntransition-metal thiophosphates MPS 3(M=Mn, Fe, and\nNi),\" Phys. Rev. B 46, 5425{5433 (1992).\n46Hao Chu, Chang Jae Roh, Joshua O. Island, Chen Li,\nSungmin Lee, Jingjing Chen, Je-Geun Park, Andrea F.\nYoung, Jong Seok Lee, and David Hsieh, \\Linear Mag-\nnetoelectric Phase in Ultrathin MnPS 3Probed by Opti-\ncal Second Harmonic Generation,\" Phys. Rev. Lett. 124,\n027601 (2020)." }, { "title": "2006.02118v2.Superconductivity_induced_change_in_magnetic_anisotropy_in_epitaxial_ferromagnet_superconductor_hybrids_with_spin_orbit_interaction.pdf", "content": "Superconductivity-induced change in magnetic anisotropy in epitaxial\nferromagnet-superconductor hybrids with spin-orbit interaction\nC\u0013 esar Gonz\u0013 alez-Ruano,1Lina G. Johnsen,2Diego Caso,1Coriolan Tiusan,3, 4\nMichel Hehn,4Niladri Banerjee,5Jacob Linder,2and Farkhad G. Aliev1,\u0003\n1Departamento F\u0013 \u0010sica de la Materia Condensada C-III,\nInstituto Nicol\u0013 as Cabrera (INC) and Condensed Matter Physics Institute (IFIMAC),\nUniversidad Aut\u0013 onoma de Madrid, Madrid 28049, Spain\n2Center for Quantum Spintronics, Department of Physics,\nNorwegian University of Science and Technology, NO-7491 Trondheim, Norway\n3Department of Physics and Chemistry, Center of Superconductivity Spintronics and Surface Science C4S,\nTechnical University of Cluj-Napoca, Cluj-Napoca, 400114, Romania\n4Institut Jean Lamour, Nancy Universit\u0012 e, 54506 Vandoeuvre-les-Nancy Cedex, France\n5Department of Physics, Loughborough University,\nEpinal Way, Loughborough, LE11 3TU, United Kingdom\nAbstract\nThe interaction between superconductivity and ferromagnetism in thin \flm superconductor/ferromagnet heterostructures\nis usually re\rected by a change in superconductivity of the S layer set by the magnetic state of the F layers. Here we\nreport the converse e\u000bect: transformation of the magnetocrystalline anisotropy of a single Fe(001) layer, and thus its preferred\nmagnetization orientation, driven by the superconductivity of an underlying V layer through a spin-orbit coupled MgO interface.\nWe attribute this to an additional contribution to the free energy of the ferromagnet arising from the controlled generation of\ntriplet Cooper pairs, which depends on the relative angle between the exchange \feld of the ferromagnet and the spin-orbit \feld.\nThis is fundamentally di\u000berent from the commonly observed magnetic domain modi\fcation by Meissner screening or domain\nwall-vortex interaction and o\u000bers the ability to fundamentally tune magnetic anisotropies using superconductivity - a key step\nin designing future cryogenic magnetic memories.\nSuperconductivity (S) is usually suppressed in the\npresence of ferromagnetism (F) [1{5]. For example, in\nF/S/F spin-valves the transition temperature TCof the\nS layer is di\u000berent for a parallel alignment of the F layer\nmoments compared to an anti-parallel alignment [6{9].\nInterestingly, for non-collinear alignment of the F layer\nmoments in spin-valves [10{12] or Josephson junctions\n[13{22], an enhancement in the proximity e\u000bect is found\ndue to the generation of long-range triplet Cooper pairs,\nimmune to the pair-breaking exchange \feld in the F lay-\ners. So far, the reciprocal modi\fcation of the static prop-\nerties of the ferromagnet by superconductivity has been\nlimited to restructuring [23] and pinning of magnetic do-\nmains walls (DWs) by Meissner screening and vortex-\nmediated pinning of DWs [24{27].\nModi\fcation of the magnetization dynamics in the\npresence of superconductivity has been studied in [28{\n36]. Recently, theoretical and experimental results have\nindicated an underlying role of Rashba spin-orbit cou-\npling (SOC), resulting in an enhancement of the prox-\nimity e\u000bect and a reduction of the superconducting TC,\nalong with enhanced spin pumping and Josephson cur-\nrent in systems with a single F layer coupled to Nb\nthrough a heavy-metal (Pt) [37{43]. In this context,\nV/MgO/Fe [44] has been shown to be an e\u000bective sys-\ntem to study the e\u000bect of SOC in S/F structures with\nfully epitaxial layers.\nAt \frst glance, altering the magnetic order in S/F\nheterostructures leading to a change in the direction ofmagnetization appears non-trivial due to the di\u000berence\nin the energy scales associated with the order parame-\nters. The exchange splitting of the spin-bands and the\nsuperconducting gap are about 103K and 101K, re-\nspectively. However, this fundamentally changes if one\nconsiders the possibility of controlling the magneto crys-\ntalline anisotropy (MCA) by manipulating the compet-\ning anisotropy landscape with superconductivity, since\nthe MCA energy scales are comparable to the supercon-\nducting gap energy. Interestingly, emergent triplet su-\nperconducting phases in S/SOC/F heterostructures o\u000ber\nthe possibility to observe MCA modi\fcation of a F layer\ncoupled to a superconductor through a spin-orbit coupled\ninterface, triggered by the superconducting phase [45].\nIn this communication, we present evidence that cubic\nin-plane MCA in V/MgO/Fe(001) system is modi\fed by\nthe superconductivity of V through SOC at the MgO/Fe\ninterface [46]. Our detailed characterization of the coer-\ncive \felds of the rotated soft Fe(001) and sensing hard\n(Fe/Co) ferromagnetic layers by tunnelling magnetoresis-\ntance e\u000bect (TMR) [47] along with numerical simulations\ndismisses the Meissner screening and DW-vortex interac-\ntions as a source of the observed e\u000bects.\nThe magnetic tunnel junction (MTJ) multilayer stacks\nhave been grown by molecular beam epitaxy (MBE) in a\nchamber with a base pressure of 5 \u000210\u000011mbar following\nthe procedure described in [48]. The samples were grown\non [001] MgO substrates. Then a 10 nm thick seed of\nanti-di\u000busion MgO underlayer is grown on the substrate\nTypeset by REVT EX 1arXiv:2006.02118v2 [cond-mat.supr-con] 26 Jun 2020(a)\nV\n+-\nCr (2nm)MgO (2nm)z\nθ\nFMFM\nFMFM\n0π/2π3π/2 2πT=10 K\nFMFMFMFM\nFMFMFMFM\nFMFM\nH[100] (kOe)TMR (%)[100][100]-\n[100][100][100]-\n-[010] [010]Fe[100]\nFe[010]\nKC\nΦH(b)\nTMR (%)\n-3 -2 -1 0 1 2 3010203040\n010203040\n(c) (d)\nT=10 KFIG. 1: (a) Sketch of the junctions under study. Fe(10\nnm)Co(20 nm) is the hard (sensing) layer while Fe(10 nm)\nis the soft ferromagnet where spin reorientation transitions\nare investigated. (b), Sketch showing the top view without\nthe hard Fe/Co layer, with the 4-fold in plane magnetic en-\nergy anisotropy expected for the Fe(001) atomic plane of the\nmagnetically free layer, for temperatures above TC(yellow\nline) and well below TC(dashed cyan). Note that during\nthe epitaxial growth, the Fe lattice is rotated by 45 degress\nwith respect to MgO. Parts (c) and (d) show in-plane spin re-\norientation transitions between parallel (P), perpendicular in\nplane (PIP) and antiparallel (AP) relative magnetization al-\nlignments of the soft and hard F layers for a 30 \u000230\u0016m2junc-\ntion at T=10 K (above TC). Indices above the inset sketches\nindicate the direction of the soft layer. The in-plane rotation\nhas been carried out with the angle \b Hof the magnetic \feld\nrelative to the Fe[100] axis going from \u000030 to 390 degrees.\nto trap the C from it before the deposition of the Fe (or\nV). Then the MgO insulating layer is epitaxially grown\nby e-beam evaporation, the thickness approximately \u00182\nnm and so on with the rest of the layers. Each layer is an-\nnealed at 450 C for 20 mins for \rattening. After the MBE\ngrowth, all the MTJ multilayer stacks are patterned in\n10-40 micrometre-sized square junctions (with diagonal\nalong [100]) by UV lithography and Ar ion etching, con-\ntrolled step-by-step in situ by Auger spectroscopy. The\nmeasurements are performed inside a JANISR\rHe3cryo-\nstat. The magnetic \feld is varied using a 3D vector mag-\nnet. For the in-plane rotations, the magnetic \feld mag-\nnitude was kept at 70-120 Oe, far away from the soft\nFe(001) and hard Fe/Co layers switching \felds obtained\nfrom in-plane TMRs (see Supplemental Material S1,S2\n[49]). This way, only the soft layer is rotated and the\ndi\u000berence in resistance can be atributted to the angle\nbetween the soft and hard layers.\nFigure 1a shows the device structure with the Fe/Cohard layer sensing the magnetization alignment of the 10-\nnm thick Fe(001) soft layer. A typical TMR plot above\nTCis shown in Figure 1c. The resistance switching shows\na standard TMR between the P and AP states. However,\nthe epitaxial Fe(001) has a four-fold in-plane anisotropy\nwith two ortogonal easy axes - [100] and [010] - (Fig-\nure 1b). These MCA states could be accessed by an\nin-plane rotation of the Fe(001) layer with respect to the\nFe/Co layer using \feld greater than the coercive \feld of\nthe Fe(001) layer without disrupting the Fe/Co magne-\ntization (see also Supplemental material S1 [49] for the\nmagnetic characterization of the Fe/Co layer). This is\nshown in Figure 1d, where TMR is plotted as a function\nof the in-plane \feld angle with respect to the [100] direc-\ntion angle \b H. This gives rise to four distinct magnetiza-\ntion states with P, perpendicular in-plane (PIP) and AP\nstates re\rected by the TMR values. Supplemental Ma-\nterial S3 [49] discusses the weak magnetostatic coupling\nbetween the two FM layers (detected through resistances\nin-between the P and AP states in the virgin state of\ndi\u000berent samples), showing that it does not a\u000bect the ca-\npability to reorient the soft layer independently of the\nhard one. It also demonstrates that the soft layer retains\ndi\u000berent magnetic directions at zero \feld.\nFigure 2 analyzes the most probable in-plane magne-\ntization orientations of the Fe(001) layer through mag-\nnetic \feld rotations at \fxed temperatures from above to\nbelowTC. Typically, no qualitative changes in TMR are\nobserved above and below TCin the 0\u0000\u0019\feld rotation\nangle (\bH) span (Figure 2a). However, in the \u0019\u00002\u0019\nrange, the TMR qualitatively changes below TC=2, pos-\nsibly indicating new stable magnetization states along\ndi\u000berent directions to the ones stablished by the princi-\npal crystallographic axes (Figure 2a).\nTo ascertain the exact angle \b FMbetween the two\nF layers, we have calibrated the magnetization direc-\ntion of the soft layer with respect to the hard Fe/Co us-\ning the Slonczewski formula (Supplemental Material S4\n[49]). The aplicability of the macrospin approach to de-\nscribe TMRs and magnetization reorientation resides in\nthe high e\u000bective spin polarization obtained ( P= 0:7)\n[47], approaching to the values typically reported for\nFe/MgO in a fully saturated state [50, 51]. Figure 2b\nis a histogram representing the probability of obtaining\na speci\fc \b FMas temperature is lowered from above to\nbelowTC. We observe that the most probable Fe(001)\ndirections are oriented along the [100] and [010] princi-\npal axes above TC=2, while below TC=2 it splits in three\nbranches roughly oriented along \u0019=4 angles. The split of\nthe [0 10] state into three branches is also visualized in\nFig.2e, with a plot of the counts vs. temperature around\nthe [110];[010] and [1 10] magnetization directions.\nInterestingly, once the rotation is initiated in the AP\ncon\fguration, the magnetization apparently locks in the\n(\u0019+\u0019=4) (or [ 110]) state (Figures 2b,d,f). This probably\narises due to the improved initial macrospin alignment,\n20 π π/2 3π/2 2π3π/2 5π/4 7π/4 2π π\n3π/2 5π/4 7π/4 2π π \nΦFM(a) (d)\n 5 K\n 4 K\n 3 K\n 2.3 K\n 2.2 K\n 2 K\n 1.5 K\n 1.2 K\n 0.8 K\n 0.6 K\n 0.3 K\n 0.25 K03640444852\n \n 0.3 K\n 1.2 K\n 2.2 K\n 3 K\n 4 K\nΦHR (kΩ)\n04080120160Nº of points\n0.3 K0.5 K0.7 K1 K1.2 K1.6 K2.1 K2.5 K3 K3.9 K4 K4.5 K5.5 K6.5 K90100110120130\n 0.3 K\n 1.2 K\n 2.5 K\n 4.5 K\n150\n100\n50\n0(b)π π/2 3π/2 2π\n ΦH\nΦFMNº of points R (kΩ)\nHMHL\nMSLΦFMΦH[100]\n0 1 2 3 4 5051015202530\nT (K)Total countTC[110] (5π/4)\n[110] (7π/4)[010] (3π/2)\n0 1 2 3 4 5 6 7010203040\nTC[110] (5π/4)\n[110] (7π/4)[010] (3π/2)Total count\nT (K)(e)\n(c) (f)[100] [010] [100] [010] [100] [100] [110] [010] [110] [100]FIG. 2: Typical angular dependence of the resistance of a\nV/MgO/Fe/MgO/Fe/Co junction on the orientation of the\nin plane \feld with respect to the main crystalline axes from\nabove to below TCwhen the rotation is initialted from P (a-c)\nand from AP state (d-f). The inset sketches the experimental\ncon\fguration, showing the angles between the ferromagnetic\nlayers (\b FM) and of the external magnetic \feld (\b H). Parts\n(b,e) correspondingly represent the experimental data shown\nin (a,d) in form of histograms, dividing the 0-2 \u0019interval in\n36 zones. Parts (c,f) plot the histograms in (b,e) as counts\nvs temperature for the intermediate states (AP+ \u0019=4 or the\n[110] axis, AP+ \u0019=2 =PIP or [010], and AP+3 \u0019=4 or [110])\nfor the second half of the rotation.\nwhich is not fully achieved in the AP state with a pre-\nceeding P-AP rotation. We believe that with the full\n2\u0019\feld rotation, magnetization inhomogeneities or local\nDWs created during the P-AP state rotation help to over-\ncome MCA energy barriers more easily. The suggested\nsuppresion of the local DWs with the magnetization rota-\ntion initated from the AP state can be indirectly inferred\nfrom the broadening of the [ 100] to [0 10] transition in the\nnormal state detected as a small (extrinsic) number of\ncounts around [ 110] (Figure 2f).\nFor a more systematic analysis, we performed a series\nof in-plane TMR measurements along di\u000berent directions\nrelative to the symmetry axes. The \frst experiment (i)\nwas performed with an initial saturation \feld of \u00061 kOe\nin the [100] direction, followed by a TMR in the [210]\ndirection (between [100] and [110]). The second (ii) ini-\ntially saturates both the hard and soft layers along the\n[100] direction. Then, a minor loop is performed starting\n \n-400 -200 0 200 400\nH (Oe)0 1 2 3 4 5 6050100150200250300\nπ/8 switch field (Oe)\n Negative field\n Positive fieldHSwitch (Oe)(a)\n51.251.652.052.452.853.2\n \n-1.0 -0.5 0.0 0.5 1.03540455055\n \nH (kOe) 0.3 K\n 1.5 K\n 2.2 K\n 4 K\n0 50 100 150420440460480500520\nHπ/4 (Oe) 6 K\n 1.6 K\nP (θ=0)θ=π/4θ=π/2θ=3π/4R (kΩ)\nT (K)R (kΩ)\n1020406080100\nTransition Probability (%)Msat≈MSLMsat=MSL\nMSLMsat\n ε≈kBT ε≫kBT\n ε'≫kBTFe[100]\nFe[010]\nKCπ/8\nHTMR\n2 34 0.5R (kΩ)\nT (K)(b)\n(c) (d)FIG. 3: (a) TMR measurements on a S/F/F 30 \u000230\u0016m2\njunction with H oriented along [210] (inset in (b)), for various\ntemperatures. The increase in R is associated with a tran-\nsition from the [110] magnetization orientation to a forced\n[210] direction of the soft layer. (b), Variation of the transi-\ntion \feld with T for the positive and negative \feld branches.\nInset: exchange energy anisotropy and direction of the ap-\nplied \feld (H TMR). (c), Two TMRs performed on a 10 \u000210\n\u0016m2junction in the [110] direction at T=6 K and 1.6 K, after\napplying 1 kOe in the [100] direction. The 6 K TMR starts in\nP state, while the 1.6 K TMR starts already in a tilted state.\nRight axis: estimation of the angle \u0012between the two F layers\nbased on the Slonczewski formula. (d), Probability of \fnding\na tilted state at H= 0 (triangular points in (b)) vs T (in log\nscale), averaged with 7 experimental points around each T.\nThe line is a guide for the eye. Insets: sketch of the magnetic\nanisotropy below and above TC, with the saturation magneti-\nzation (M sat) and the zero \feld magnetization state measured\nfor the soft layer (M SL).\"and\"0represent the energy barrier\nseparating the [100] magnetization direction from the closest\nminimum below and above TC, respectively.\nfrom zero \feld and going up to 150 Oe along the [110]\naxis.\nBoth experiments further suggest the possibility of\nsuperconductivity-induced changes of MCA. The inset\nof Figure 3a shows the full \feld sweep range in the \frst\n(i) con\fguration, and Figure 3a zooms in close to the AP\ncon\fguration. When we sweep the \feld in the [210] di-\nrection, we detect a weak but robust resistance upturn\nat temperatures below approximately TC=2 (Figure 3).\nThis additional TMR increase (shown by the arrows in\nFigure 3a) roughly corresponds to an 8-10 degree rota-\ntion in the relative spin direction between the soft and\nhard layer towards their AP alignment (see Supplemental\nMaterial S4 [49] for an analysys of the calculated angle\nerror). Within the proposed macrospin approximation,\nthis could be understood as a redirection of the soft layer\nmagnetization forced by the external \feld, from the ini-\ntially blocked [110] direction towards the external \feld\n[210] direction. A strong increase of the characteristic\n3\feld,Hswitch , required to reorient the soft layer from\n[110] towards [210] when T decreases below TC=2, could\nre\rect the superconductivity-induced MCA energy min-\nimum along the [110] direction.\nThe minor TMR loops along [110] (Figure 3c) realized\nafter saturation along [100] point on a thermally induced\nmagnetization reorientation from [100] towards [110] even\nat zero \feld, in a temperature range below TCwhere the\nbarrier between adjacent energy minima is comparable\ntokBT. The zero-\feld reorientation becomes less prob-\nable when the thermal energy is insu\u000ecient to overcome\nthe barrier (Figure 3d). An estimation of the in-plane\nnormal-state MCA energy barrier done through magne-\ntization saturation along [100] and [110] provides a value\nof only a few \u0016eV/atom (Supplemental Material S5 [49]).\nHowever, the real barrier is determined by the nucleation\nvolume, which depends on the exchange length in the ma-\nterial. With a DW width of about 3 nm for Fe(001) we\nestimate the MCA barrier to be at least 100\u0000101mV.\nBefore describing our explanation of the MCA modi-\n\fcation of Fe(001) in the superconducting state of V(40\nnm)/MgO(2 nm)/Fe(10 nm) system, we discard alter-\nnative interpretations of the observed e\u000bects. Meissner\nscreening [24, 25], if present, would introduce about a\n10% correction to the actual magnetic \feld independently\nof the external \feld direction (see Supplemental Material\nS2 [49]). The reason for the weak in-plane \feld screening\ncould be the small superconductor thickness (40 nm),\nonly slightly exceeding the estimated coherence length\n(26 nm). On the other hand, intermediate multidomain\nstates are expected to be absent in when magnetization\nis directed along [110] (Supplemental material S6 [49]).\nIndeed, our experiments show that magnetization, when\nlocked below TCin the (\u0019+\u0019=4) angle, hardly depends\non the absolute value of the external \feld along [110]\nvaried between 0 and 100 Oe. Moreover, simulations\nof the vortex-DW interaction using MuMax3 [52] and\nTDGL codes [59] discard the vortex mediated DW pin-\nning [26, 27] scenario including when interfacial magnetic\ndefects created by mis\ft dislocations [54] are considered\n(see Supplemental material S6 [49]). The vortex pinning\nmechanism also contradicts that only the (0 \u0000\u0019) \feld\nrotation span (Figure 2a) gets a\u000bected below TC=2. The\nobserved irrelevance of the junction area (Supplemental\nmaterial S7 [49]) contradics the importance of the vortex-\nedge DWs interaction. The shape and vortex-DWs in-\nteraction e\u000bects, if relevant, would strengthen magneti-\nzation pinning along [100], but not [110] (Supplemental\nmaterial S6, S7 [49]). Finally, we also indicate that the\nMCA modi\fcation from singlet superconductivity would\nnot enable any zero \feld rotation to non-collinear mis-\nalignment angles, in contrast to our data (Fig. 3d).\nTo explain our results, we consider the possibility\nin which the invariance of the superconducting prox-\nimity e\u000bect to magnetization rotation is broken in the\npresence of SOC. It has been predicted that triplet-\n0.0000.0050.0100.0150.0200.0250.030\nF-FAP\n TC+\n 0.8TC\n 0.7TC\n 0.6TC\n 0.5TC\n 0.4TC\nπ π/2 3π/2 π/4 3π/4 5π/4 7π/4 2π 0\nΦFM\nφFM\nCondensate TripletsMgO V\nSOC assisted\nconversionFe\nCondensate\nstrengthenedFewer triplets\nSOC assisted\nconversion weakened\nCondensation energy gain\nT > TcFe[100]\nFe[010]\nT < TcFe[100]\nFe[010]MSL\nMSL\nMgO V Fe(a)\n(b)FIG. 4: Numerical modelling. (a), Free energy Fvs in-plane\nmagnetization angle \b FMfor temperatures below the super-\nconducting critical temperature and just above the critical\ntemperature ( T+\nC). The free energy is plotted relative to the\nfree energy in the AP con\fguration FAPand has been nor-\nmalized to the hopping parameter tused in the tight-binding\nmodel. (b). Illustration of the physical origin of the change\nin magnetic anisotropy induced by the superconducting layer.\nAboveTC, V is a normal metal and the soft Fe layer has a 4-\nfold in-plane magnetic energy anisotropy (yellow line). Below\nTC, V is superconducting and in\ruences the soft Fe layer via\nthe proximity e\u000bect: a leakage of Cooper pairs into the ferro-\nmagnet. Due to the SOC at the interface, a magnetization-\norientation dependent generation of triplet Cooper pairs oc-\ncurs. The generation of triplets is at its weakest for a mag-\nnetization pointing in the [ 110] direction, giving a maximum\nfor the superconducting condensation energy gain. This mod-\ni\fes the magnetic anisotropy of the soft Fe layer (cyan line),\nenabling magnetization switching to the [ 110] direction (blue\narrow). The magnetic anisotropy does not show the weak\nAP coupling between the two Fe layers, causing an absolute\nminimum in \b FM=\u0019(a).\nsuperconductivity is e\u000bectively generated even for weakly\nspin-polarized ferromagnets with a small spin-orbit \feld\n[55]. In addition to generating triplet pairs, the SOC also\nintroduces an angle-dependent anisotropic depairing \feld\nfor the triplets [43, 45]. In V/MgO/Fe, the Rashba \feld is\ncaused by a structural broken inversion symmetry at the\n4MgO interfaces [44]. We model our experimental results\nusing a tight-binding Bogolioubov-de Gennes Hamilto-\nnian on a lattice and compute the free energy (Sup-\nplemental material S8 [49]). The Hamiltonian includes\nelectron hopping in and between the di\u000berent layers, a\nRashba-like SOC at the MgO/Fe interface, an exchange\nsplitting between spins in the Fe layers, and conventional\ns-wave superconductivity in the V layer. The free energy\ndetermined from this Hamiltonian includes the contribu-\ntion from the superconducting proximity e\u000bect, and an\ne\u000bective in-plane magnetocrystalline anisotropy favoring\nmagnetization along the [100] and [010] axes. Experi-\nmentally, we see a weak anti-ferromagnetic coupling be-\ntween the Fe(100) and Fe/Co layers (which does not af-\nfect the capability to reorient the soft layer independently\nof the hard one) described by an additional contribution\nfAFcos(\bFM) with a constant parameter fAF>0.\nFigure 4 shows the total free energy of the system as a\nfunction of the IP magnetization angle \b FMfor decreas-\ning temperatures. Due to the increase in the supercon-\nducting proximity e\u000bect, additional local minima appear\nat \bFM=n\u0019=2 +\u0019=4, wheren= 0;1;2;:::(i.e.[110],\n[\u0016110], [ \u00161\u001610], and [1 \u001610], respectively). This is a clear signa-\nture for the proximity-induced triplet correlations. These\nare most e\u000eciently generated at angles \b FM=n\u0019=2\n(i.e.[100], [010], [ \u0016100], and [0 \u001610]) for a heterostructure\nwith a magnetic layer that has a cubic crystal structure\nlike Fe [45]. As a result, the decrease in the free energy\nis stronger at angles \b FM=n\u0019=2 +\u0019=4 where more\nsinglet Cooper pairs survive. Our numerical results thus\ncon\frm that the experimentally observed modi\fcation of\nthe anisotropy can be explained by the presence of SOC\nin the S/F structure alone, without including supercon-\nducting proximity e\u000bects from misalignment between the\nFe(100) and Fe/Co layers. Moreover, Figure 4 illustrates\nwhy the \b FM=n\u0019=2 +\u0019=4 states only appear experi-\nmentally when the external \feld is rotated from an AP to\nP alignment (Figure 2). Because of the weak AP coupling\nbetween the ferromagnetic layers, the energy thresholds\nfor reorienting the magnetization from one local mini-\nmum to the next are higher under a rotation from AP to\nP alignment.\nIn conclusion, we present experimental evidence\nfor superconductivity-induced change in magnetic\nanisotropy in epitaxial ferromagnet-superconductor hy-\nbrids with spin-orbit interaction. We believe that this\nmechanism is fundamentally di\u000berent from the previous\nreports of magnetisation modi\fcation arising from Meiss-\nner screening and vortex induced domain wall pinning,\neven though the spin-triplet mechanism and performed\nsimulations require many assumptions. Our results es-\ntablish superconductors as tunable sources of magnetic\nanisotropies and active ingredients for future low dissipa-\ntion superspintronic technologies. Speci\fcally, they could\nprovide an opportunity to employ spin-orbit proximity\ne\u000bects in magnetic Josephson junction technology andappoach it to Fe/MgO-based junctions that are widely\nused in commercial spintronic applications.\nAcknowledgements\nWe acknowledge Mairbek Chshiev and Antonio Lara\nfor help with simulations, Yuan Lu for help in sample\npreparations and Igor Zutic and Alexandre Buzdin for\nthe discussions. The work in Madrid was supported\nby Spanish Ministerio de Ciencia (MAT2015-66000-P,\nRTI2018-095303-B-C55, EUIN2017-87474) and Conse-\njer\u0013 \u0010a de Educaci\u0013 on e Investigaci\u0013 on de la Comunidad\nde Madrid (NANOMAGCOST-CM Ref. P2018/NMT-\n4321) Grants. FGA acknowledges \fnancial support from\nthe Spanish Ministry of Science and Innovation, through\nthe Mara de Maeztu Programme for Units of Excel-\nlence inR&D(MDM-2014-0377, CEX2018-000805-M).\nNB was supported by EPSRC through the New In-\nvestigator Grant EP/S016430/1. The work in Nor-\nway was supported by the Research Council of Nor-\nway through its Centres of Excellence funding scheme\ngrant 262633 QuSpin. C.T. acknowledges \\EMERSPIN\"\ngrant ID PN-IIIP4-ID-PCE-2016-0143, No. UEFISCDI:\n22/12.07.2017. The work in Nancy was supported by\nCPER MatDS and the French PIA project \\Lorraine\nUniversit\u0013 e d'Excellence\", reference ANR-15-IDEX-04-\nLUE.\nSUPPLEMENTARY MATERIAL\nIn the supplementary material, the section S1 presents\na magnetic characterization of the hard Fe/Co layer of\nthe junctions under study. Section S2 presents a mag-\nnetic characterization of the soft Fe(001) layer and stud-\nies the possible in\ruence of the Meissner screening on\nthe coercive \felds of the soft and hard layers. Section S3\nestimates the strength of the weak antiferromagnetic cou-\npling between magnetically soft and hard electrodes. Sec-\ntion S4 provides details about the calibration of the angle\nbetween the soft and hard layers using the Slonczewski\nformula, as well as discussing the possible sources of error\nfor this calibration and their magnitude. Section S5 pro-\nvides an estimation for the magneto-anisotropic energy\nbarrier between the [110] and [100] magnetization direc-\ntions, normalized per volume or per atom. Section S6\nnumerically evaluates the possible domain walls pinning\nby superconducting vortices. Section S7 discusses the\ncontribution of the shape to the magnetic anisotropy. Fi-\nnally, section S8 provides details on the theoretical mod-\nelling of the observed e\u000bects.\n5S1. Magnetic characterizarion of the hard Fe/Co\nlayer\nFigure S5 shows the magnetic characterization of the\nhard Fe/Co bilayer, determined from a typical spin-\nvalve M-H loop on a standard Fe/MgO/Fe/Co single\ncrystal MTJ system (continuous layers, unpatterned).\nThe nominal thickness of the layers on this sample,\nMgO(100)/Fe(30 nm)/MgO(2 nm)/Fe(10 nm)/Co(20\nnm), has been chosen to optimize the magnetic properties\nof the MTJ stack [48]. The TMR measurements of the\ncoercive \felds of the hard (H C;Hard ) and soft (H C;Soft )\nlayers in MTJs under study show that they are well sep-\narated from the external \feld values used to rotate the\nsoft layer. Figure S6 shows that the hard layer switching\n\felds obtained from TMRs along [100], [010] and [110]\nmeasured in our junctions remain far above the typical\nrange of 70-120 Oe which is used to rotate the soft layer.\nMoreover, Figure S7 also shows the typical temperature\ndependence of H C;Hard , demonstrating its independence\nwith temperature from well above to well below TC.\n-600 -400 -200 0 200 400 600-1.0-0.50.00.51.0\nhard Fe/Co bilayer\n M/MS\nH (Oe)soft Fe\nFIG. 5: Magnetic characterization of a Fe(30\nnm)/MgO/Fe(10 nm)Co(20 nm) structure, at room temper-\nature, along the [100] direction.\nS2. Magnetic characterizarion of the soft Fe(001)\nlayer and estimation of the Meissner screening\nThe magnetostatic Meissner screening has been dis-\ncussed mainly in studies with perpendicular magnetiza-\ntion [24]. In the case of the experiments with in-plane\n\feld rotation which we carry out, such \feld expulsion\ncould induce some screening of the external magnetic\n\feld applied to invert or rotate the magnetization of the\nsoft Fe(001) layer (which is the closest to the supercon-\n0.0 22.5 45.00100200300400500\n \n[110] [210]HC,Hard (Oe)\nθ[100] (deg)Field range used to rotate the soft layer\n[100]FIG. 6: Coercive \feld of the hard Fe/Co layer for magnetic\n\feld oriented along di\u000berent crystallographic directions [100],\n[110] and [210], above the superconducting critical tempera-\nture (T= 5 K). The grey band shows the typical \feld range\nused to manipulate the magnetization of the soft Fe(100) layer\nin the rotation experiments.\nductor), and with less probability a\u000bect the switching of\nthe more distant hard Fe/Co layer.\nFigure S8 shows the typical variation of the coercive\n\feld of the soft Fe(001) ferromagnetic layer with temper-\nature from above to below the critical temperature. We\nobserve some weak increase of the coercive \feld below 10\nOe, which could be due to spontaneous Meissener screen-\ning and/or vortex interaction with domain walls. These\nchanges, however, are an order of magnitude below the\ntypical magnetic \felds applied to rotate the Fe(001) layer\n(70-120 Oe). As we also show in Figure S7, the coer-\ncive \feld of the hard FeCo layer (typically above 400-500\nOe) shows practically no variation (within the error bars)\nwithin a wide temperature range, from 3 TCto 0:1TC,\ndiscarding the in\ruence of the Meissner screening on the\nhard layer.\nAs the superconducting layer is much larger in area\nthan the ferromagnetic one, these experiments point out\nthat the possible existing Meissner screening would intro-\nduce about a 10% correction to the actual external \feld\nacting on the soft ferromagnet, regardless of the external\n\feld direction.\nS3. Estimation of the weak antiferromagnetic cou-\npling of the two ferromagnetic layers.\nIn order to quantify the unavoidable weak antiferro-\nmagnetic magnetostatic coupling between the rotated\nsoft Fe(001) and the practically \fxed hard FeCo layer, we\nshow low \feld TMR measurements where the AP state is\nachieved and then maintained at zero \feld (Figure S9a).\nOne clearly observes that the AP and P states can be\n6FIG. 7: Typical temperature dependence of the coercitive\n\feld of the hard Fe/Co ferromagnetic layer along the [100]\ndirection. The critical temperature is marked with a dashed\nred line. We relate the excess scatter observed in the hard\nlayer with the extra structural disorder at the Fe/Co interface,\nproviding an enhanced coercive \feld for the Fe/Co layer. The\nblue line is a guide for the eye. The inset shows the method for\ndetermining the coercive \feld: it's the \feld of the \frst point\nafter the hard layer transition from in each TMR experiment.\nFIG. 8: Typical temperature dependence of the coercitive\n\feld of the soft Fe(001) ferromagnetic layer measured along\n[100] direction. The critical temperature is marked with a\ndashed red line. Blue lines are guides for the eye. The inset\nshows the method for determining the coercive \feld: a logistic\n\ft was performed for the transition, and the coercive \feld was\nde\fned as the mid-height value of the \ft.\nobtained as two di\u000berent non-volatile states, and there-\nfore the antiferromagnetic coupling is not su\u000ecient to\nantiferromagnetically couple the two layers at zero \feld.\nThe stability of the P state against the antiferromagnetic\ncoupling is con\frmed by the temperature dependence of\nthe resistance in the P and AP states. The P state shows\nstable resistance values at least below 15 K (Figure S9b).\nThis means that the antiferromagnetic coupling energy\nis well below 2 mV.\nFIG. 9: Two experiments demonstrating the stability of the\nP and AP states at zero \feld. (a) TMR to AP state before a\ncritical temperature measurement: the sample was \frst satu-\nrated in the P state with H= 1000 Oe in the [100] direction,\nand then a negative \feld sweeping was performed to -200 Oe\nand back to 0 Oe in the same direction in order to switch the\nsoft layer into the AP state, where it remained at zero \feld.\n(b) Two critical temperature measurements: the sample was\nsaturated in the P state, and then switched to AP state as\ndescribed in (a) for the AP measurement. After this, the\ntemperature was risen to 15 K and let to slowly cool down to\nT\u00182 K. The increase in resistance below 4 K corresponds to\nthe opening and deepening of the superconducting gap, since\nthe voltage used was only a few microvolts in order to dis-\ntinguish the superconducting transition from its appearance.\nBoth experiments show no sudden changes in resistance, as\nwould happen if any magnetic transition took place.\nS4. Calibration of the angle between the two fer-\nromagnetic layers\nIn order to estimate the angle between the two ferro-\nmagnets for the TMR measurements and rotations, we\nused the Slonczewski model [56]. By using values of the\nresistance in the AP, P and PIP states established above\nTC, we can calculate the desired angle \u0012with the follow-\n7ing expression:\nG\u00001=G1\u00001+\u0002\nG2\u0000\n1 +p2cos\u0012\u0001\u0003\u00001: (1)\nHere,Gis the total conductance of the sample, G1and\nG2are the conductances of each of the two tunnel bar-\nriers, andpis the spin polarization in the ferromagnets,\nfor which we obtain values between 0.7 and 0.8 depending\non the sample (the value being robust for each individual\none).\nIn order to ascertain the precission of this calibration\nmethod, an analysis of the di\u000berent errors has been per-\nformed. First, an standard error propagation calculation\nwas done to estimate the uncertainty in the resistance\nvalues, taking typical values for the current and voltage\nof 100 nA and 5 mV, respectively, which gives us a typical\nresistance value of 50 k\n. The current is applied using a\nKeithley 220 Current Source, which has an error of 0 :3%\nin the operating range according to the user manual. The\nvoltage is measured using a DMM-522 PCI multimeter\ncard. In the speci\fcations, the voltage precision is said\nto be 5 1=2 digits. With all this, the resistance error ob-\ntained is \u0001R=75.08 \n or a 0.15% of relative error. Using\nthis value, the error bars in the measurements shown\nin the main text would be well within the experimental\npoints.\nFor the calculated angle, the error propagation method\nis not adequate. It gives errors bigger than 360 degrees\nfor some angles, and in general over 30 degrees. This is\nclearly not what it is observed in reality: the performed\n\fts are quite robust, showing little variance in the esti-\nmated angle when changing the input parameters all that\nis reasonable. Instead, we have used a typical rotation\nperformed on a 30 \u000230\u0016m2sample. The \ftting to the\nSlonczewski formula needs three input values: the resis-\ntance in the P state ( RP), the resistance in the AP state\n(RAP) and the resistance in the PIP state ( RPIP). Using\nthese, a numerical algorithm calculates the spin polariza-\ntion (p), the resistance of the F/F barrier ( RFIF), and\nthe resistance of the F/S (F/N) barrier ( RNIF). These\ngive us the total resistance of the sample as a function\nof the angle \b FMbetween the two ferromagnets or, re-\nciprocally, the angle as a function of resistance. For our\nestimation, we have varied the value of the RPIPinput\nparameter from the lowest to the highest possible in the\nPIP state of the rotation, as well as taking an interme-\ndiate value which would be used in a normal analysis\n(the P and AP resistance values are always taken as the\nminimum and maximum resistance values in the rotation\nrespectively). The calculated parameters for the resis-\ntance of each barrier and the polarization may slightly\nvary from one \ftting to another, but the overall \ftting\nremains remarkably stable, as shown in Figure 10.\nAs expected, the di\u000berence is higher for the PIP state,\nand minimum in the P and AP state that are \\\fxed\".\n3.8 4 4.2 4.4 4.6 4.802468\n3.8 4 4.2 4.4 4.6 4.804590135180 ϕFM,max - ϕFM,min (deg)ϕFM (deg)\nR (Ω)x104x104(a)\n(b)RPIP,min\nRPIP\nRPIP,max\nR (Ω)FIG. 10: (a) \b FMas a function of resistance for the \fttings\nwith maximum, usual, and minimum RPIP used, in the P-\nAP resistance range. (b) di\u000berence of calculated angle vs\nresistance (in the P-AP resistance range) for the \fttings with\nmaximum and minimum RPIPused.\nThe di\u000berence doesn't exceed 7 degrees, and it keeps be-\nlow 2 degrees near the P and AP states.\nS5. Saturation magnetization for thin Fe(001) \flms\nin [100] and [110] directions\nDi\u000berent M vs H measurements were performed at\nroom temperature on a 10 nm thick Fe \flms, both for the\neasy [100] and hard [110] crystallographic axes, in order\nto estimate the magnetocrystalline anisotropy (MCA) en-\nergy. The results are depicted in Fig. S11. Using the\nsaturation \feld for the two directions, the anisotropy\nenergy can be estimated as KFe=MFeHSat=2 =\n5:1\u0002105erg\u0001cm\u00003, whereMFe= 1714 emu/cm3is\nused. The anisotropy energy per unit cell is therefore\nMAE= 6:674\u0016eV, or 3.337 \u0016eV per atom. The obtained\nenergy barrier is similar to the one measured using fer-\nromagnetic resonance [57].\nAs shown in Fig.S12, the experimental MCA energy\n8-50 -25 0 25 50-1.0-0.50.00.51.0\n M/Msat\nH (Oe)[100] in plane easy axis\n-1 -0.5 0 0.5 1-101\n[110] in plane hard axis\nHsat \nH (kOe)(a)\n(b)\nM/MsatFIG. 11: M vs H measurements on a 10 nm thick Fe \flm for\nthe easy [100] (a) and hard [110] (b) crystallographic axis.\nThe saturation \feld ( Hsat) for the easy axis is around 10 Oe,\nwhile for the hard direction it reaches up to 600 Oe.\nvalues have been theoretically confronted with theoreti-\ncal/numerical calculations of the angular in-plane varia-\ntion of magnetic anisotropy, using the ab-initio Wien2k\nFP-LAPW code [58]. The calculations were based on a\nsupercell model for a V/MgO/Fe/MgO slab similar to the\nexperimental samples. To insure the requested extreme\naccuracy in MCA energy values ( \u0016eV energy range), a\nthoroughly well-converged kgrid with signi\fcantly large\nnumber ofk-points has been involved. Within these cir-\ncumstances, our theoretical results for the Fe(001) thin\n\flms show standard fourfold anisotropy features and rea-\nsonable agreement with the experimentally estimated\n\fgures with a maximum theoretical MAE of 4.9 \u0016eV\nper atom (expected theoretical under-estimation of the\nmagnetocristalline energy within the GGA approach).\nNote that the superonducting-V induced MCA modu-\nlation features cannot be described within the ab-initio\nFP-LAPW approach, describing the V in its normal\nmetallic state. Therefore, the below TCexperimentally\nobserved MCA energy modulations have to be clearly\nrelated to the proximity e\u000bect in the superconducting\nV/MgO/Fe(001) system and not to any speci\fc MCA\nfeature of Fe(001) in the V/MgO/Fe(001)/MgO complex\nstacking sequence.\n0.0 0.4 0.8 1.2 1.6012345\n MAE\n 4.7sin2(2θ)\nθ (radians)MAE (μeV/Atom)\n[100][110]Fe\nθFIG. 12: Ab-initio calculation of magnetocryscalline\nanisotropy energy (MAE) as a function of the in-plane ori-\nentation angle \u0012, de\fned in the inset. Solid line is a phe-\nnomenological \ft to a sin2(2\u0012) function.\nS6. Evaluation of the vortex induced pinning of\ndomain walls\nUsing MuMax3 [52], we have compared numerically the\nDWs formation along the [100] and [110] magnetization\ndirections. The simulations took place in samples with\n3\u00023\u0016m2lateral dimensions (100 nm rounded corners\nwere used as the devices have been fabricated by optical\nlithography), with 512 \u0002512\u000216 cells, atT= 0. The rest\nof the parameters used were Aex= 2:1\u000210\u000011J/m for\nthe exchange energy, Msat= 1:7\u0002106A/m for the sat-\nuration magnetization, a damping parameter \u000b= 0:02,\nand crystalline anisotropy parameters KC1= 4:8\u0002104\nJ/m3andKC3=\u00004:32\u0002105J/m3. The goal of the\nsimulations was to evaluate the DW formation and their\ninteraction with the superconducting vortices induced by\nthe vertical component of the stray \felds at a 2-3 nm from\nthe Fe(001) surface. We observed that, depending on the\nexternal \feld, in the range of 70-1000 Oe both edge-type\nand inner-type DWs are formed when the \feld is directed\nalong [100], and mainly edge type DWs are formed with\n\feld along [110] (Figure S13a).\nWe have also calculated the interaction Ibetween the\nDW related excess exchange energy Eexand the vertical\ncomponent of the stray \felds, Be\u000b(Figure S13b):\nI=ZNx\n0ZNy\n0jBe\u000bjEexFdxdy (2)\nWhereNxandNyare the total number of cells in each\ndimension of the simulation, and Fis a \flter \\Vortex gen-\neration function\" that takes into account the simulated\ndependence of the number of vortices on the vertically\napplied \feld (Figure S13c). The vortices were simulated\nusing the Time Dependent Ginzburg Landau code devel-\n90.0 0.5 1.0051015\nDW-vortex eff. interaction (arb. u.)\nH (kOe) H[110]\n H[100]\n0 0.2 0.40200400\nNumber of vorticesT=2 K(c)\nH/HC2H[110]|H|=1000 OeH[100]|H|=70 Oe\nx103\nH[110](b) (a)\nH[100]\nOrder parameter01\n0\nMZ/MFIG. 13: (a) Typical DW formation mapped by MuMax3 simulations for the [110] and [100] applied \feld directions in the\nnon-saturated (70 Oe) and saturated (1000 Oe) \feld regimes. The color map represents the out of plane component of the\nmagnetization, while the red arrows indicate the in plane direction. (b) Values of the 2D integral Ibetween the local exchange\nenergy (DWs) and the perpendicular component of the stray \felds at a distance of 2-3 nm from the ferromagnet, taking into\nacount the vortex generation function F. (c) Vortex generation function F, represented as number of vortices formed in a 5 \u00025\n\u0016m2square 40 nm thick superconducting Vanadium \flm as a function of the applied perpendicular \feld (normalized by the\nsecond critical \feld Hc2), simulated at T= 2 K by using the TDGL code described in [59]. The insert shows a typical image\nof the vortices at H=0.15 Hc2\noped in Madrid described in [59]. The TDGL simulations\ntook place in 5\u00025\u0016m2Vanadium samples with 200 \u0002200\ncells, atT= 2 K, with a coherence length \u00180= 2:6\u000210\u00008\nbased on our experimental estimations for the studied de-\nvices,\u0014= 3 andTC= 4 K. A uniform \feld was applied in\nthe perpendicular direction, its magnitude varying from\n0:1HC2to 0:6HC2, and the number of vortices generated\nin the relaxed state were counted.\nThe second critical \feld in the vertical direction ( Hc2=\n3 kOe) was determined experimentally. The estimated\ninteraction shows that in the weakly saturated regime,\nwhen the inner DWs could emerge and the DW-vortex in-\nteraction increases, such interaction should pin the mag-\nnetization along the [100] direction, corresponding to\nthe MCA already present in the normal state, therefore\nblocking any magnetization rotation towards the [110] di-\nrection, contrary to our experimental observations. The\npossible reason for the irrelevance of the DW-vortex in-\nteraction in our system is that inner DWs are expected\nto be of Neel-type for the thickness considered [60].\nFinally we mention that our numerical evaluations\nshow that, if present, the vortex-DW interaction should\nremain dominant for the magnetization directed along\n[100] respect [110] and for the magnetic \feld range 70-\n1000 Oe also without KC3parameter providing the MCA\nenergy minima along [110]. In these simulations the\nFe(001) layer has been considered to be smooth. In or-\nder to further approach simulations to the experiment,\nwe have also veri\fed that the conclusions above are not\na\u000bected by the introduction of interfacial magnetic dis-\norder due to mismatch defects (every 30 lattice periods)\nwith 25% excess of Fe moment at the Fe/MgO interface\n[61]. More detailed simulations involving also interface\nroughness could be needed to further approach the realexperimental situation.\nS7. Magnetization alignment along [110] and irrel-\nevance of the junction area for the superconductivity\ninduced MCA modi\fcation\nAs we mentioned in the main text, our experiments\npoint that Fe(001) layers are close to a highly saturated\nstate when the magnetization is directed along [100] or\nequivalent axes. On the other hand, micromagnetic simu-\nlations (Figure S13a) show that the magnetization align-\nment is more robust in the [110] direction (or equivalent)\nrather than in the [100] direction (or equivalent). So,\nif we indeed reach a highly saturated state in the [100]\ndirection, this should also be the case for the [110] direc-\ntion. Therefore, the emergent stable tunneling magne-\ntoresistance states we observe experimentally below Tc,\ncannot be explained in terms of the intermediate multi-\ndomain states but rather correspond to the dominant\n[110] magnetization alignment of the Fe(001) layer.\nAs shown in Figure S14, our experiments shows that\nthe observed e\u000bects remain qualitatively unchanged when\nthe junction area is varied about an order of magnitude.\n10051015202530\nTMR (%)10x10 μm2\n 1.5 K\n 5 K\n010203040\n20x20 μm2\n 1.5 K\n5 KTMR (%)\n01020304050\n30x30 μm2\n 1.5 K\n 5 KTMR (%)\nΦH\nΦFM Nº of points\nT=5 K\nT=1.5 K10x10 μm2\n20x20 μm2\n30x30 μm2(a) (c)\nπ3π/2 2π 5π/4 7π/4 π3π/2 2π 5π/4 7π/4 π3π/2 2π 5π/4 7π/4(b)\n(d)ΦH ΦH\n40\n040\n0\nπ3π/2 5π/4\n0 2 4 6 8010203040\n0 2 4 6 80102030(e) (f)Nº of points\nNº of points\nT (K) T (K)[110] (5π/4)[010] (3π/2)\n[110] (5π/4)\n[110] (7π/4)[010] (3π/2)30x30 μm2\nTC TC10x10 μm2FIG. 14: In-plane \feld rotation experiments with H= 70 Oe below (blue) and above (red) TCfor 10\u000210 (a), 20\u000220 (b) and\n30\u000230 (c)\u0016m2junctions. (d) Histograms of the calculated angle between the two FM layers \u001eFMfor these same rotations,\nabove and below TC, showing the [ 110] states for low temperatures. (e) and (f) show the evolution of the [110], [ 110] and [010]\n(PIP) states with temperature for the same 10 \u000210 and 30\u000230\u0016m2junctions (qualitatively simular evolution is shown in\nFigure 2f in the main text for the 20 \u000220\u0016m2junction).\nS8. Modelling\nWe describe the V/MgO/Fe structure by the Hamilto-\nnian [45]\nH=\u0000tX\nhi;ji;\u001bcy\ni;\u001bcj;\u001b\u0000X\ni;\u001b(\u0016i\u0000Vi)cy\ni;\u001bci;\u001b\n\u0000X\niUini;\"ni;#+X\ni;\u000b;\fcy\ni;\u000b(hi\u0001\u001b)\u000b;\fci;\f\n\u0000i\n2X\nhi;ji;\u000b;\f\u0015icy\ni;\u000b^n\u0001(\u001b\u0002di;j)\u000b;\fcj;\f(3)\nde\fned on a cubic lattice. The \frst term describes\nnearest-neighbor hopping. The second term includes the\nthe chemical potential and the potential barrier at the\ninsulating MgO layers. The remaining terms describes\nsuperconducting attractive on-site interaction, ferromag-\nnetic exchange interaction, and Rashba spin-orbit inter-\naction, respectively. These are only nonzero in their re-\nspective regions. In the above, tis the hopping integral,\n\u0016iis the chemical potential, Viis the potential barrier\nthat is nonzero only for the MgO layer, U > 0 is the\nattractive on-site interaction giving rise to superconduc-tivity,\u0015iis the local spin-orbit coupling magnitude, ^ nis\na unit vector normal to the interface, \u001bis the vector of\nPauli matrices, di;jis a vector from site ito sitej, andhi\nis the local magnetic exchange \feld. The number opera-\ntor used above is de\fned as ni;\u001b\u0011cy\ni;\u001bci;\u001b, andcy\ni;\u001band\nci;\u001bare the second-quantization electron creation and an-\nnihilation operators at site iwith spin\u001b. The supercon-\nducting term in the Hamiltonian is treated by a mean-\n\feld approach, where we assume ci;\"ci;#=hci;\"ci;#i+\u000e\nand neglect terms of second order in the \ructuations \u000e.\nWe consider a system of size Nx\u0002Ny\u0002Nzsetting\nthe interface normals parallel to the xaxis and assuming\nperiodic boundary conditions in the yandzdirections.\nTo simplify notation in the following, we de\fne i\u0011ix,\nj\u0011jx,ijj= (ix;iy) andk\u0011(ky;kz). We apply the\nFourier transform\nci;\u001b=1p\nNyNzX\nkci;k;\u001bei(k\u0001ijj)(4)\nto the above Hamiltonian and use that\n1\nNyNzX\nijjei(k\u0000k0)\u0001ijj=\u000ek;k0: (5)\n11We choose a new basis\nBy\ni;k= [cy\ni;k;\"cy\ni;k;#ci;\u0000k;\"ci;\u0000k;#] (6)\nspanning Nambu\u0002spin space, and rewrite the Hamilto-\nnian as\nH=H0+1\n2X\ni;j;kBy\ni;kHi;j;kBi;k: (7)\nAbove, the Hamiltonian matrix is given by\nHi;j;k=\u000fi;j;k^\u001c3^\u001b0+\u000ei;jh\ni\u0001i^\u001c+^\u001by\u0000i\u0001\u0003\ni^\u001c\u0000^\u001by\n+hx\ni^\u001c3^\u001bx+hy\ni^\u001c0^\u001by+hz\ni^\u001c3^\u001bz\n\u0000\u0015isin(ky)^\u001c0^\u001bz+\u0015isin(kz)^\u001c3^\u001byi\n;(8)\nwhere \u0001iis the superconducting gap which we solve for\nself-consistently, ^ \u001ci^\u001bj\u0011^\u001ci\n^\u001bjis the Kronecker product\nof the Pauli matrices spanning Nambu and spin space,\n^\u001c\u0006\u0011(^\u001c1\u0006i^\u001c2)=2, and\n\u000fi;j;k\u0011\u00002t[cos(ky) + cos(kz)]\u000ei;j\n\u0000t(\u000ei;j+1+\u000ei;j\u00001)\n\u0000(\u0016i\u0000Vi)\u000ei;j:(9)\nThe constant term in Eq. (7) is given by\nH0=\u0000X\ni;kf2t[cos(ky) + cos(kz)] +\u0016i\u0000Vig\n+NyNzX\nij\u0001ij2\nUi:(10)\nWe absorb the sum over lattice sites in Eq. (7) into the\nmatrix product by de\fning a new basis\nWy\nk= [By\n1;k;:::;By\ni;k;:::;By\nNx;k]: (11)\nEq. (7) can then be rewritten as\nH=H0+1\n2X\nkWy\nkHkWk; (12)\nwhere\nHk=2\n64H1;1;k\u0001\u0001\u0001H1;Nx;k\n.........\nHNx;1;k\u0001\u0001\u0001HNx;Nx;k3\n75 (13)\nis Hermitian and can be diagonalized numerically. We\nobtain eigenvalues En;kand eigenvectors \b n;kgiven by\n\by\nn;k= [\u001ey\n1;n;k\u0001\u0001\u0001\u001ey\nNx;n;k];\n\u001ey\ni;n;k= [u\u0003\ni;n;kv\u0003\ni;n;kw\u0003\ni;n;kx\u0003\ni;n;k]:(14)\nThe diagonalized Hamiltonian can be written on the form\nH=H0+1\n2X\nn;kEn;k\ry\nn;k\rn;k; (15)where the new quasi-particle operators are related to the\nold operators by\nci;k;\"=X\nnui;n;k\rn;k;\nci;k;#=X\nnvi;n;k\rn;k;\ncy\ni;\u0000k;\"=X\nnwi;n;k\rn;k;\ncy\ni;\u0000k;#=X\nnxi;n;k\rn;k:(16)\nThe superconducting gap is given by \u0001 i\u0011Uihci;\"ci;#i.\nWe apply the Fourier transform in Eq. (4) and use\nEq. (16) in order to rewrite the expression in terms\nof the new quasi-particle operators. Also using that\nh\ry\nn;k\rm;ki=f\u0000\nEn;k=2\u0001\n\u000en;m, we obtain the expression\n\u0001i=\u0000Ui\nNyNzX\nn;kvi;n;kw\u0003\ni;n;k[1\u0000f(En;k=2)] (17)\nfor the gap, that we use in computing the eigenenergies\niteratively. Above, f\u0000\nEn;k=2\u0001\nis the Fermi-Dirac distri-\nbution.\nUsing the obtained eigenenergies, we compute the free\nenergy,\nF=H0\u00001\n\fX\nn;kln(1 +e\u0000\fEn;k=2); (18)\nwhere\f= (kBT)\u00001. The preferred magnetization di-\nrections are described by the local minima of the free\nenergy. In the main body of the paper, we use this to\nexplain the possible magnetization directions of the soft\nferromagnet when rotating an IP external magnetic \feld\nover a 2\u0019angle starting at a parallel alignment with the\nhard ferromagnet.\nOther relevant quantities to consider in modelling\nthe experimental system is the superconducting coher-\nence length and the superconducting critical temper-\nature. In the ballistic limit, the coherence length is\ngiven by\u0018=~vF=\u0019\u00010, wherevF=1\n~dEk\ndk\f\f\nk=kFis the\nFermi velocity related to the normal-state eigenenergy\nEk=\u00002t[cos(kx) + cos(ky) + cos(kz)]\u0000\u0016, and \u0001 0is the\nzero-temperature superconducting gap [62]. The critical\ntemperature is found by a binomial search, where we de-\ncide if a temperature is above or below Tcby determining\nwhether \u0001 NSx=2increases towards a superconducting so-\nlution or decreases towards a normal state solution from\nthe initial guess under iterative recalculations of \u0001 i. We\nchoose an initial guess with a magnitude very close to\nzero and with a lattice site dependence similar to that of\nthe gap just below Tc.\nIn the main plot showing the free energy under IP ro-\ntations of the magnetization, we have chosen parameters\n12t= 1,\u0016S=\u0016SOC =\u0016F= 0:9,V= 2:1,U= 1:35,\n\u0015= 0:4,h= 0:8,NS\nx= 30,NSOC\nx = 3,NF\nx= 8, and\nNy=Nz= 60. All length scales are scaled by the lattice\nconstanta, all energy scales are scaled by the hopping pa-\nrametert, and the magnitude of the spin-orbit coupling\n\u0015is scaled by ta. In order to make the system com-\nputationally manageable, the lattice size is scaled down\ncompared to the experimental system, however the re-\nsults should give qualitatively similar results as long as\nthe ratios between the coherence length and the layer\nthicknesses are reasonable compared to the experimental\nsystem. For this set of parameters, the superconduct-\ning coherence length is approximately 0 :6NS\nx. Since the\ncoherence length is inversely proportional to the super-\nconducting gap, Uhas been chosen to be large in order to\nallow for a coherence length smaller than the thickness\nof the superconducting layer. Although this results in\na large superconducting gap, the modelling will qualita-\ntively \ft the experimental results as long as the other pa-\nrameters are adjusted accordingly. We therefore choose\nthe local magnetic exchange \feld so that h\u001d\u0001, as in\nthe experiment. For this parameter set, h\u001920\u0001. The\norder of magnitude of \u0015is 1 eV \u0017A, given that t\u00181 eV and\na\u00184\u0017A. This is realistic considering Rashba parameters\nmeasured in several materials [63]. The Rashba spin-\norbit \feld at the interfaces of V/MgO/Fe is caused by\na structural inversion asymmetry across the MgO layer,\nand breaks the inversion symmetry at the MgO interfaces\n[44]. This causes generation of triplet-superconductivity\neven for weakly spin-polarized ferromagnets with a small\nspin-orbit \feld [64]. We are therefore not dependent upon\na strong magnetic exchange \feld and a strong spin-orbit\n\feld for realizing the observed e\u000bects. For the AF cou-\npling contribution to the free energy, we set fAF= 0:01\nin order to \ft the anisotropy of the experimental system\njust aboveTC.\n\u0003e-mail: farkhad.aliev@uam.es\n[1] V.L. Ginzburg, Ferromagnetic superconductors, Zh.\nEksp. Teor. Fiz. 31, 202 (1956); Sov. Phys. JETP 4, 153\n(1957).\n[2] B. T. Matthias, H. Suhl and E. Corenzwit, Spin exchange\nin superconductors, Phys. Rev. Lett. 1, 152 (1958).\n[3] A. I. Buzdin, L. N. Bulaevskii, M. L. Kulich and S. V. Pa-\nnyukov, Magnetic superconductors, Sov Phys. Uspekhi,\n27, 927 (1984).\n[4] J. Y. Gu, C.-Y. You, J. S. Jiang, J. Pearson, Y. B.\nBazaliy and S. D. Bader, Magnetization-orientation de-\npendence of the superconducting transition temperature\nin the ferromagnet-superconductor-ferromagnet system:\nCuNi/Nb/CuNi, Phys. Rev. Lett. 89, 267001 (2002).\n[5] I. C. Moraru, W. P. Pratt, Jr. and N. O. Birge, Ob-\nservation of standard spin-switch e\u000bects in ferromag-\nnet/superconductor/ferromagnet trilayers with a strong\nferromagnet, Phys. Rev. B 74, 220507(R) (2006).[6] L. R. Tagirov, Low-\feld superconducting spin switch\nbased on a superconductor/ferromagnet multilayer,\nPhys. Rev. Lett. 83, 2058 (1999).\n[7] A. I. Buzdin, A. V. Vedyayev and N. V. Ryzhanova, Spin-\norientation dependent superconductivity in F/S/F struc-\ntures, Europhys. Lett., 48(6), 686-691 (1999).\n[8] I. Baladi\u0013 e, A. Buzdin, N. Ryzhanova and A. Vedyayev,\nInterplay of superconductivity and magnetism in su-\nperconductor/ferromagnet structures,Phys. Rev. B 63,\n054518 (2001).\n[9] P. V. Leksin, N. N. Garif'yanov, I. A. Garifullin,\nJ. Schumann, V. Kataev, O. G. Schmidt and B.\nB uchner, Manifestation of New Interference E\u000bects in\na Superconductor-Ferromagnet Spin Valve, Phys. Rev.\nLett. 106, 067005 (2011).\n[10] X. L. Wang, A. Di Bernardo, N. Banerjee, A. Wells, F.\nS. Bergeret, M. G. Blamire and J. W. A. Robinson, Gi-\nant triplet proximity e\u000bect in superconducting pseudo\nspin valves with engineered anisotropy, Phys. Rev. B 89,\n140508(R) (2014).\n[11] A. Singh, S. Voltan, K. Lahabi and J. Aarts, Colossal\nProximity E\u000bect in a Superconducting Triplet Spin Valve\nBased on the Half-Metallic Ferromagnet CrO 2, Phys.\nRev. X 5, 021019 (2015).\n[12] P.V. Leksin, N. N. Garifyanov, I. A. Garifullin, Ya.V.\nFominov, J. Schumann, Y. Krupskaya, V. Kataev, O.\nG. Schmidt and B. B uchner, Evidence for triplet super-\nconductivity in a superconductor-ferromagnet spin valve,\nPhys. Rev. Lett., 109, 057005 (2012).\n[13] R. S. Keizer, S. T. B. Goennenwein, T. M. Klapwijk, G.\nMiao, G. Xiao and A. A. Gupta, A spin triplet supercur-\nrent through the half-metallic ferromagnet CrO 2, Nature\n439, 825 (2006).\n[14] T. S. Khaire, M. A. Khasawneh, W. P. Pratt, Jr. and N.\nO. Birge, Observation of spin-triplet superconductivity\nin Co-based Josephson junctions, Phys. Rev. Lett. 104,\n137002 (2010).\n[15] M. S. Anwar, F. Czeschka, M. Hesselberth, M. Porcu and\nJ. Aarts, Long-range supercurrents through half-metallic\nferromagnetic CrO 2, Phys. Rev. B 82, 100501(R) (2010).\n[16] J. W. A. Robinson, J. D. S. Witt and M. G. Blamire,\nControlled injection of spin-triplet supercurrents into a\nstrong ferromagnet, Science 329, 59 (2010).\n[17] C. Visani, Z. Sefrioui, J. Tornos, C. Leon, J. Bri-\natico, M. Bibes, A. Barth\u0013 el\u0013 emy, J. Santamar\u0013 \u0010a and J.\nE. Villegas, Equal-spin Andreev re\rection and long-\nrange coherent transport in high-temperature supercon-\nductor/halfmetallic ferromagnet junctions, Nature Phys.,\n8, 540 (2012).\n[18] N. Banerjee, J. W. A. Robinson and M. G. Blamire, Re-\nversible control of spin-polarized supercurrents in ferro-\nmagnetic Josephson junctions, Nature Comm. 5, 4771\n(2014).\n[19] A. Iovan, T. Golod and V. M. Krasnov, Controllable gen-\neration of a spin-triplet supercurrent in a Josephson spin\nvalve, Phys. Rev. B 90, 134514 (2014).\n[20] J. Linder and J. W. A. Robinson, Superconducting spin-\ntronics, Nature Phys. 11, 307 (2015).\n[21] M. Eschrig, Spin-polarized supercurrents for spintronics:\na review of current progress, Rep. Prog. Phys. 78, 104501\n(2015).\n[22] M. G. Flokstra, N. Satchell, J. Kim, G. Burnell, P. J. Cur-\nran, S. J. Bending, J. F. K. Cooper, C. J. Kinane, S. Lan-\ngridge, A. Isidori, N. Pugach, M. Eschrig, H. Luetkens,\n13A. Suter, T. Prokscha and S. L. Lee, Remotely induced\nmagnetism in a normal metal using a superconducting\nspin-valve, Nature Phys., 12, 97 ( 2016).\n[23] A.I. Buzdin and L. N. Bulaevskii, Ferromagnetic \flm on\nthe surface of a superconductor: Possible onset of inho-\nmogeneous magnetic ordering, Zh. Eksp. Teor. Fiz. 94,\n256-261 (1988), Sov. Phys. JETP 67, 576-578 (1988).\n[24] L. N. Bulaevskii and E. M. Chudnovsky, Ferromagnetic\n\flm on a superconducting substrate, Phys. Rev. B 63,\n012502 (2000).\n[25] S. V. Dubonos, A. K. Geim, K. S. Novoselov and I. V.\nGrigorieva, Spontaneous magnetization changes and non-\nlocal e\u000bects in mesoscopic ferromagnet-superconductor\nstructures, Phys. Rev. B 65, 220513R (2002).\n[26] J. Fritzsche, R. B. G. Kramer and V. V. Moshchalkov,\nVisualization of the vortex-mediated pinning of ferromag-\nnetic domains in superconductor-ferromagnet hybrids,\nPhys. Rev. B 79, 132501 (2009).\n[27] P. J. Curran et al., Irreversible magnetization switching\nat the onset of superconductivity in a superconductor fer-\nromagnet hybrid, Appl. Phys. Lett. 107, 262602 (2015).\n[28] X. Weintal and P. W. Brouwer, Magnetic exchange in-\nteraction induced by a Josephson current, Phys. Rev. B\n65, 054407 (2002).\n[29] Y. Tserkovnyak and A. Brataas, Current and\nspin torque in double tunnel barrier ferromagnet-\nsuperconductorferromagnet systems, Phys. Rev. B 65,\n094517 (2002).\n[30] C. Bell, S. Milikisyants, M. Huber and J. Aarts, Spin dy-\nnamics in a superconductor-ferromagnet proximity sys-\ntem, Phys. Rev. Lett. 100, 047002 (2008).\n[31] B. Braude and Ya. M. Blanter, Triplet Josephson e\u000bect\nwith magnetic feedback in a superconductor-ferromagnet\nheterostructure, Phys. Rev. Lett. 100, 207001 (2008).\n[32] E. Zhao and J. A. Sauls, Theory of nonequilibrium spin\ntransport and spin-transfer torque in superconducting-\nferromagnetic nanostructures, Phys. Rev. B 78, 174511\n(2008).\n[33] F. Konschelle and A. Buzdin, Magnetic moment manip-\nulation by a Josephson current, Phys. Rev. Lett. 102,\n017001 (2019).\n[34] J. Linder and T. Yokoyama, Supercurrent-induced mag-\nnetization dynamics in a Josephson junction with two\nmisaligned ferromagnetic layers, Phys. Rev. B 83, 012501\n(2011).\n[35] J. Linder, A. Brataas, Z. Shomali and M. Zareyan, Spin-\ntransfer and exchange torques in ferromagnetic supercon-\nductors, Phys. Rev. Lett. 109, 237206 (2012).\n[36] N. G. Pugach and A. I. Buzdin, Magnetic moment ma-\nnipulation by triplet Josephson current, Appl. Phys. Lett.\n101, 242602 (2012).\n[37] F. S. Bergeret and I. V. Tokatly, Singlet-triplet\nconversion and the long-range proximity e\u000bect in\nsuperconductor-ferromagnet structures with generic spin\ndependent \felds, Phys. Rev. Lett., 110, 117003 (2013).\n[38] N. Banerjee, J. A. Ouassou, Y. Zhu, N. A. Stelmashenko,\nJ. Linder and M. G. Blamire, Controlling the supercon-\nducting transition by spin-orbit coupling, Phys. Rev. B\n97, 184521 (2018).\n[39] K.-R. Jeon, C. Ciccarelli, A. J. Ferguson, H. Kure-\nbayashi, L. F. Cohen, X. Montiel, M. Eschrig, J. W. A.\nRobinson and M. G. Blamire, Enhanced spin pumping\ninto superconductors provides evidence for superconduct-\ning pure spin currents, Nat. Mater. 17, 499 (2018).[40] K.-R. Jeon, C. Ciccarelli, A. J. Ferguson, H. Kure-\nbayashi, L. F. Cohen, X. Montiel, M. Eschrig, S. Komori,\nJ. W. A. Robinson and M. G. Blamire, Exchange-\feld en-\nhancement of superconducting spin pumping, Phys. Rev.\nB99, 024507 (2019).\n[41] N. Satchell and N. O. Birge, Supercurrent in ferromag-\nnetic Josephson junctions with heavy metal interlayers,\nPhys. Rev. B 97, 214509 (2018).\n[42] N. Satchell, R. Loloee and N. O. Birge, Supercurrent\nin ferromagnetic Josephson junctions with heavy-metal\ninterlayers. II. Canted magnetization, Phys. Rev. B 99,\n174519 (2019).\n[43] S. H. Jacobsen, J. A. Ouassou and J. Linder, Crit-\nical temperature and tunneling spectroscopy of\nsuperconductor-ferromagnet hybrids with intrinsic\nRashba-Dresselhaus spin-orbit coupling, Phys. Rev. B\n92, 024510 (2015).\n[44] I. Mart\u0013 \u0010nez, P. H ogl, C. Gonz\u0013 alez-Ruano, J. P. Cascales,\nC. Tiusan, Y. Lu, M. Hehn, A. Matos-Abiague, J. Fabian,\nI. Zutic and F. G. Aliev, Interfacial spin-orbit coupling:\na platform for superconducting spintronics, Phys. Rev.\nAppl. 13, 014030 (2020).\n[45] L. G. Johnsen, J. Linder and N. Banerjee, Magnetization\nreorientation due to superconducting transition in heavy\nmetal heterostructures, Phys. Rev. B 99, 134516 (2019).\n[46] H. X. Yang, M. Chshiev and B. Dieny, First-principles\ninvestigation of the very large perpendicular magnetic\nanisotropy at Fe/MgO and Co/MgO interfaces. Phys.\nRev. B 84, 054401 (2011).\n[47] I. Mart\u0013 \u0010nez, C. Tiusan, M. Hehn, M. Chshiev and F.\nG. Aliev, Symmetry broken spin reorientation transi-\ntion in epitaxial MgO/Fe/MgO layers with competing\nanisotropies, Sci. Rep. 8, 9463 (2018).\n[48] C. Tiusan, M. Hehn, F. Montaigne, F. Greullet, S. An-\ndrieu and A. Schuhl, Spin tunneling phenomena in single\ncrystal magnetic tunnel junction systems, J. Phys.: Con-\ndens. Matter 19, 165201 (2007).\n[49] See Supplemental Material at XXX for more details.\n[50] S. S. P. Parkin, C. Kaiser, A. Panchula, P. M. Rice, B.\nHughes, M. Samant, and S.-H. Yang, Giant tunnelling\nmagnetoresistance at room temperature with MgO (100)\ntunnel barriers, Nat. Mater. 3, 862 (2004).\n[51] S. Yuasa, T. Nagahama, A. Fukushima, Y. Suzuki and\nK. Ando, Giant room temperature magneto-resistance\nin single-crystal Fe/MgO/Fe magnetic tunnel junctions,\nNat. Mater. , 868 (2004).\n[52] A. Vansteenkiste, J. Leliaert, M. Dvornik, M. Helsen,\nF. Garcia-Sanchez and B. Van Waeyenberge, The design\nand veri\fcation of MuMax3, AIP Advances 4, 107133\n(2014).\n[53] A. Lara, C. Gonz\u0013 alez-Ruano and F.G.Aliev, Time-\ndependent Ginzburg-Landau simulations of supercon-\nducting vortices in three dimensions, Low Temperature\nPhysics 46, 316 (2020).\n[54] D. Herranz, F. Bonell, A. Gomez-Ibarlucea, S. Andrieu,\nF. Montaigne, R. Villar, C. Tiusan, and F.G.Aliev,\nStrongly suppressed 1/f noise and enhanced magnetore-\nsistance in epitaxial FeV/MgO/Fe magnetic tunnel junc-\ntions, Applied Physics Letters 96, 202501 (2010).\n[55] T. Vezin, C. Shen, J. E. Han and I. \u0014Zuti\u0013 c, Enhanced spin-\ntriplet pairing in magnetic junctions with s-wave super-\nconductors, Phys. Rev. B 101, 014515 (2020).\n[56] J. C. Slonczewski, Conductance and exchange coupling of\ntwo ferromagnets separated by a tunneling barrier. Phys.\n14Rev. B 39, 6995 (1989).\n[57] A. N. Anisimov, M. Farle, P. Poulopoulos, W. Platow,\nK. Baberschke, P. Isberg, R. Wppling, A. M. N. Niklas-\nson and O. Eriksson, Orbital Magnetism and Magnetic\nAnisotropy Probed with Ferromagnetic Resonance, Phys.\nRev. Lett. 82, 2390 (1999).\n[58] P. Blaha, K.Schwarz, F. Tran, R. Laskowski, G.K.H.\nMadsen and L.D. Marks, An APW+lo program for cal-\nculating the properties of solids, J. Chem. Phys. 152,\n074101 (2020).\n[59] A. Lara, C. Gonz\u0013 alez-Ruano and F.G.Aliev, Time-\ndependent Ginzburg-Landau simulations of supercon-\nducting vortices in three dimensions, Low Temperature\nPhysics 46, 316 (2020).[60] A. Hubert and R. Schfer, Magnetic domains. The analysis\nof magnetic microstructures (Springer, Berlin, 1999).\n[61] E. Jal et al; Interface Fe magnetic moment enhance-\nment in MgO/Fe/MgO trilayers, Appl. Phys. Lett. 107,\n092404 (2015).\n[62] J. Bardeen, L. N. Cooper and J. R. Schrie\u000ber, Theory of\nSuperconductivity, Phys. Rev. 108, 1175 (1957).\n[63] A. Manchon, H. C. Koo, J. Nitta, S. M. Frolov, and R. A.\nDuine, New perspectives for Rashba spinorbit coupling,\nNat. Mater. 14, 871 (2015).\n[64] T. Vezin, C. Shen, J. E. Han, and I. \u0014Zuti\u0013 c, Enhanced\nspin-triplet pairing in magnetic junctions with s-wave su-\nperconductors, Phys. Rev. B, 101, 014515 (2020).\n15" }, { "title": "2006.04192v1.Magnetocrystalline_anisotropy_of_the_easy_plane_metallic_antiferromagnet_Fe__2_As.pdf", "content": "* d-cahill@illinois.edu \n†\nzhu.diao@famu.edu Magnetocrystalline anisotropy of the easy -plane metallic antiferromagnet Fe 2As \n \nKexin Yang ,1,2 Kisung Kang ,2,3 Zhu Diao ,2,4,5,\n† Manohar H. Karigerasi ,2,3 Daniel P. Shoemaker, 2,3 \nAndré Schleife ,2,6 David G. Cahill 1,2,3,* \n \n1Department of Physics, University of Illinois at Urbana -Champaign, Urbana, Illinois 61801, USA \n2Materials Research Laboratory, University of Illinois at Urbana -Champaign, Urbana, Illinois 61801, USA \n3Materials Science and Engineering, University of Illinois at Urbana -Champaign, Urbana, Illinois 61801, USA \n4Department of Physics, AlbaNova University Center, Stockholm University, SE -106 91 Stockholm, Sweden \n5Department of Physics, Florida A&M University, Tallahassee, Florida 32307, USA \n 6National Center for Supercomputing Applications, University of Illinois at Urbana -Champaign, Urbana, Illinois 61801, USA \nABSTRACT \nMagnetocry stalline anisotropy is a fundamental property of magnetic materials that determin es the \ndynamics of magnetic precession, the frequency of spin waves, the thermal stability of magnetic \ndomains, and the efficiency of spintronic devices. We combine torque magnetometry and density \nfunction al theory calculation s to determine the magnetocrystalline anisotropy of the metallic \nantiferromagnet Fe 2As. Fe2As has a tetragonal crystal structure with the Néel vector l ying in the \n(001) plane. We report that the four-fold magnetocrystalline anisotropy in the (001) -plane of Fe 2As \nis extremely small , \n3\n22 150 J/m K=− at T = 4 K, much smaller than perpendicular magnetic \nanisotropy of ferromagnetic structure widely used in spintronics device . K22 is strongly \ntemperature dependent and close to zero at T > 150 K. The anisotropy \n1K in the (010) plane is too \nlarge to be measure d by torque magnetometry and we determine \n1K= -830 kJ/m3 using first-\nprinciples density function al theory . Our simulations show that the contribution to the anisotropy \nfrom classical magnetic dipole -dipole interactions is comparable to the contribution from spin -\norbit coupling. The calculated four-fold anisotropy in the (001) plane \n22K ranges from \n3290 J/m− \nto \n3280 J/m , the same order of magnitude as the measured value . We use d K1 from theory to \npredict the frequency and polarization of the lowest frequency antiferromagnetic resonance mode \nand find that the mode is linearly polarized in the (001) -plane with \nf= 670 GHz . \n2 \n \nINTRODUCTION \nAntiferromagnets (AFs) have potential advantages over ferromagnets for spintronic device s. \nCollinear AFs are relatively insensitive to external field s because the net magnetization is zero . \nAFs typically have much higher antiferromagnetic resonance (AFMR) frequency than \nferromagnets and therefore pr ocessional switching can occur in AFs at a faster rate than in \nferromagnets. \nThe recent discovery of electrical manipulation and detection of spin configurations in metallic \nAFs ha s led to a rapidly expanding scientific literature on this class of magnetic materials. \nTetragon al crystals with easy -plane magnetic anisotropy are preferred because the two degenerate \norientations of the Néel vector can store binary information. In crystals with globally \ncentrosymmetric but locally non -centrosymmetric magnetic structure s—e.g., CuMnAs and \nMn 2Au—an electrical current exerts a torque on the N éel vector and the domain structure can \npotentially be switched electrically [1][2][3][4]. \nA small value of t he in -plane magnetocrystalline anisotropy facilitates e lectrical switching of the \ndomain orientation since a smaller torque is needed to overcome the energetic barrier that separates \nthe two orientations. Thermal stability of the domain requires, however, a large v alue of the in -\nplane anisotropy. The N éel-Arrhenius law provides an estimate of t he rate of thermal fluctuations \nof a single domain [5]: \n \n01exp\nBEfkT =− \n , (1) \nwhere 𝜏 is the average time between thermally -activated changes in the direction of the \nmagnetization , \n0f is the resonance frequency, \nEis the energy barrier between two degenerate \nmagnetic state s and \nBkT is the thermal energy. \nE is given by the product of an anisotropy \nparameter K and the volume of the domain V; \nE KV= . Stable data storage typically requires \n/ 40BE k T\n to meet the criteria that data must be retained for 10 years [6]. \n3 \n For the media of conventional hard drive s, the anisotropy parameter K is controlled by the \nperpendicular magnetocrystalline anisotropy of ordered intermetallic alloys . In the emerging \ntechnology of magneti c random access memory (MRAM), K is controlled by the interfacial \nmagnetic anisotropy of a ferromagnetic layer adjacent to the oxide barrier in a magnet ic tunnel \njunction . The perpendicular magnetic anisotropy \n1K of MRAM materials is typically \n6 7 -3\n1 10 10 J m K\n [7]. \nMagnetocrystalline anisotropy Eani is described by a phenomenological expansion of the energy \nas a function of direction cosines for the orientation of magnetization of a ferromagnet or the \nsublattice magnetization of an antiferromagnet (AF) . For a tetragonal crystal, the expansion to \nfourth order gives 3 coefficients \n1K , \n2K and \n22K [8]. \n1K is a second order coefficient; \n2K and \n22K\n are fourth order coefficients . \n \n()2 4 4\n1 2 22 / sin sin sin cos 4aniE V K K K = + + , (2) \nwhere \n is the angle of the magnetization relative to the <001> direction and \n is the angle of \nthe magnetization relative to the <100> direction (Fig. 2 ). The coefficient \n1K describes the two -\nfold anisotropy in (010) plane and \n2K represents the higher order fou r-fold symmetry of the (010) \nplane . Because the effect of \n2K is usually much smaller than \n1K , K2 will be neglected in the \nfollowing discussion. A crystal with an easy -plane anisotropy is described by \n10 K . The \ncoefficient \n22K describes the four -fold anisotropy of the (001) plane and determines the thermal \nstability of an easy -plane domain structure. \nAn external magnetic field applied to an antiferromagnet (AF) produces a small induced magnetic \nmoment . The induced moment is small because tilting of the orientation of sublattice \nmagnetization is constrained by strong exchange interaction between the magnetic sublattices that \nfavors a parallel alignment of the sublattices . In general , however , the induced magnetic moment \nis not parallel to the applied field because magnet ocrystalline anisotropy favors an orientation of \nthe sublattice magnetization along an easy axis [9][10]. \n4 \n The lack of alignment between the induced moment m and the applied field B produces a \nmacroscopic torque on the sample , \n=τ m B . A torque magnetometer measures this torque . Data \nfor the torque as a function of applied field is sensitive to magnetocrystalline anisotropy as long \nas the anisotropy is n either too small nor too large. If the anisotropy is small, then the angle \nbetween m and B is small and the torque becomes difficult to detect . If the anisotropy is large, \nthen the direction of m is fixed with respect to the crystallographic axis and the torque does not \nprovide information about the magnitude of the anisotropy. We can measure the in -plane four-fold \nanisotropy of a mm -size bulk crystal of Fe 2As by torque magnetometry but the out -of-plane two-\nfold anisotropy is not accessible to this technique because the external field is too small compare \nto anisotropy field to extract information about the anisotropy in the out -of-plane direction . We \ninstead employ first -principles calculation s based on density functional theory to dete rmine K1. \nWhen magnetic energy is larger than the anisotropy energy , the amplitude of the torque in the \n(001) -plane sa turates and the four-fold magnetic anisotropy, \n22K can be directly determined from \nthe amplitude of the torque . We measured three samples extracted from the same growth run , and \nthe \n22K value of all three samples is comparable to -150 J/m3 at 4 K. The magnitude of \n22K drops \nquickly as temperature increase s and reaches a small value above 150 K . The temperature \ndependence of magnetic anisotropy f or antiferromagnets is similar to ferromagnets , follow ing a \npower law of sublattice magnetization [11][12][13]. \nStrikingly, t orque data for the applied field rotating in the (010) -plane reveal the motion of domain \nwalls. An applied field in the (010) plane of 1 T is sufficient to orient the N éel vector fully \nperpendicular to the applied field. Domain wall mot ion occurs even at T = 4 K and, therefore, is \nnot thermally activated. \nIn the final section , we derive the lowest -frequency, zone -center AFMR frequency for easy -plane \nAFs, \n22 1 1 2 ( ) 2EEH H H H H = − − , where \n is the gyromagnetic ratio, and \nEH, \n1H , \n22H\n are the exchange field, out -of-plane anisotropy field and in -plane anisotropy field , respectively. \nThe anisotropy fields are calculated with anisotropy energy and sublattice magnetization: \n11 / H K M=\n and \n22 22 / H K M= . With \n1K calculated by DFT as \n1K = -830 kJ/m3, the AFMR \nfrequency is \nf = 670 GHz at 4 K . \n5 \n METHODS \nFe2As crystallizes in the Cu 2Sb tetragonal crystal structure. Based on the corresponding magnetic \nsymmetry (mmm1' magnetic point group), the Néel vector of Fe 2As has two degenerate \norientations in the (001) -plane [14][15][16]. \nThe Fe2As crystal was synthesized by mixing Fe and As powders in a 1.95:1 ratio and vacuum \nsealing inside a quartz tube. The vacuum tube was heated at 1 ℃/min up to 600 ℃ and held for 6 \nhours in a furnace. The temperature was then ramped to 975 ℃ at 1 ℃/min and held for 1 hour \nbefore cooling down slowly to 900 ℃ at 1 ℃/min. Finally, the quartz tube was kept at 900 ℃ for \n1 hour and allowed to cool down to room temperature in the furnace at 10 ℃/min. We obtained a \nlarge silver -hued crystal ingot of Fe 2As and it easily detached from the quartz tube. Part of the \ningot was crushed into powder for powder XRD characterization and the data showed phase pure \nFe2As. But the sample is slightly off -stochiometry as described in reference [17]. The remaining \nportion of the ingot was then fractured and the fractured surface revealed a smooth facet . Laue \ndiffraction was carried out after polishing this fractured surface. A four -fold symmetry pattern was \nobserved indicating the fractured surface is the (001) plane. \nWe used a wire saw to cut the sample into smaller pieces for magnetic property characterization \nand torque measurement s. One of the pieces was measured on the superconducting quantum \ninterference device vibrating sample magnetometer (SQUID -VSM, see below) , the other three \npieces were used in torque magnetometry measurements . We name these three samples measured \nby torque magnetometry sample A , sample B , sample C and the one for SQUID -VSM is named \nsample D . \nThe temperature -dependent magnetic susceptibility was measured with a SQUID -VSM in a \nQuantum Design Magnetic Properties Measurement System (MPMS). The susceptibility of the \nsample was measured while cooling from 398 K to 4 K in a 10 mT field. \nTorque measurements were performed in a Quantum Design Physical Property Measurement \nSystem (PPMS). We mounted the sample on a standard torque sensor chip , P109A from Quantum \nDesign with a sensitivity of 1 × 10-9 N·m. The PPMS horizontal sample rotator was used to control \nthe angle between th e crystal and the applied field. During the measurement, the external field \n6 \n rotated in either the (010) plane or the (001) plane, while the field -induced moment reside d in the \nsame plane as the rotating applied field . We detect ed the torque component, \nmB , that is \nperpendicular to th is plane . \nWe performed first -principles DFT simulations using the Vienna Ab Initio Simulation Package \n(VASP) [18][19] , to calculate the two -fold anisotropy\n1K and obtain an estimate for the four-fold \nanisotropy \n22K . The projector augmented wave (PAW) method [20] is used to describe the \nelectron -ion interaction. Kohn -Sham states are expanded into plane waves up to a kinetic energy \ncutoff of 600 eV . The Brillouin zone is sampled by a \n21 21 7 Monkhorst -Pack [21] (MP) k-\npoint grid and the total energy is converged self-consistently to within \n910− eV. The local density \napproximation (LDA) [22] and the generalized -gradient approximation developed by Perdew, \nBurke, and Ernzerhof (PBE) [23] are used to describe the exchange -correlation energy function , \nand results from the two different computational strategies are compared . \nAchieving the extremely high accuracy for total energies that is required to compute the (001) \nplane magnetocrystalline anisotropy that is on the order of \neV per magnetic unit cell is \nnumerically challenging; the required convergence parameters render it computationally too \nexpensive to perform such calculations fully self -consistently. Instead , we use the convergence \nparameters quoted above to compute Kohn -Sham states, electron density, and relaxe d atomic \ngeometries for collinear magnetism and take non -collinear magnetism and spin -orbit coupling [24] \ninto account without self -consistency of the Hamiltonian, as described in Ref. [25]. \n \nRESULTS AND DISCUSSION \n1. Magneti c susceptibility and domain wall motion \nWe use d ata for the magnetic susceptibility as input for modeling the torque magnetometry data \nand to provide insight into the reorientation of antiferromagnetic domains in a n external magnetic \nfield. Figure 1(a) summarizes the results for the magnetic susceptibility in the limit of small field . \nWe fixed the applied field at 10 mT , along the <100> or <001> direction, measured the induced \n7 \n magnetic moment while cooling from T = 398 K, and calculated the susceptibility, \n/MH= , \nwhere M is the magnetization . The measured susceptibility is similar to that in a prior report [14]. \nWhen the applied field is along an easy ax is, we must take domain wall motion into account . We \nassume that a 10 mT external field is too weak to significant ly affect the domain structure . We \nfurther assume that the magnetic moment generated by an applied field along the <100> direction \n(the a-axis of the crystal) has equal contribut ions from two types of domains that we label as D1 \n(Néel vector along <100>) and D2 (Néel vector along <010>) as illustrated in Fig. 2 . For an applied \nfield in the (001) -plane, we define the susceptibility parallel to the Néel vector as \n\n while that \nperpendicular to the Néel vector as \n⊥ . On the other hand, t he susceptibility for an applied fie ld \nin the <001> direction is define d as \n⊥ . We expect\n⊥ and \n⊥ to be similar but d ue to the \ntetragonal symmetry of the crystal structure , \n⊥ and \n⊥ are not necessarily equal. We sho w \nbelow that the difference between \n⊥ and \n⊥ is less than 5%. \nMeasurements of the magnetization as a function of field , see Fig. 1 (b), show that \nc is constant \nfor H applied along the c-axis. For H along the a-axis, \na increases with field at low field, and is \napproximately constant for an applied field > 1 T. We attribute the field dependence of \na to \ndomain wall motion and the consequent evolution of the populations of domains with N éel vectors \nparallel and perpendicular to the applied field. \nThe populations of the two degenerate domains can be estimated from the M vs H curve in Fig. \n1(a) by expressing the field-induced magnetization as \n|| || aM H H ⊥⊥ =+ and \ncMH ⊥= , \nwhere \n⊥ and \n|| are the normalized domain fraction perpendicular and parallel to the a-axis, \nrespectively, and \n||1 ⊥+= . We made three assumptions: (1) \n||0.5 ⊥== at zero field; (2) \n1 ⊥\n at high field; and (3) domain wall motion is reversible. The field -dependent distribution of \ndomains parallel and perpendicular to the external field along the a-axis at T = 4 K is shown in \nFig. 1(c) and are treated as free parameters ; we find \n\n = 0.008 and \n⊥ anisotropy in the = 0.018 . \nFor an ideal co llinear antiferromagnet, we expect \n\n = 0 at low temperatures [9]. This is not what \nwe obser ved in our measurements. The reason is that there is a background contribution to the \n8 \n magnetic susceptibility that we do not yet understand. We assume that the background \nsusceptibility is isotropic. \n2. Torque magnetometry \nThe field -induced torque is the cross product of the field -induced magnetic moment and the \napplied field, \n=mB . The direction of the induced magnetic moment m is given by the \nminimum in the total energy:\ntot m ani exE E E E= + + , where \nmE is the magnetic energy; \naniE is the \nmagne tocrystalline anisotropy energy ; and \nexE is the exchange energy that couples the two \nsublattices. We refer to the condition\nm aniEE\n as the low field limit and the condition\nm aniEE \nas the intermediate field regime [26]. We ignore a separate\nexE term when we analyze th e torque \ndata in the (001) -plane because we assume that the exchange interaction stays the same and can \nbe represented by susceptibility . For torque data in the (010) plane, our analysis is based on the \nanisotropy in the susceptibility, \n|| ⊥− , which is also related to the strength of the exchange \ninteraction. \nFig. 2(a) shows the experimental geometry when the applied field is rotating in the (010) -plane. \n\n is the angle between the c-axis and the applied field. The induced moment s are also in the \n(010) -plane . Thus, the torque is along the <010> direction . In the low temperature limi t, \n||0 = if \nthe isotropic background is igonored ; the induced moment of domain D1 is therefore along <001> \nand the induced moment of D2 lies in the (010) plane between <001> and <100> . The direction of \nthe induced moment of D2 is determined by \n⊥ and \n⊥. \nFig. 2(b) shows the experimental geometry when the applied field is rotating in the (001) -plane . \nIn this case, because of the relatively small magnetocrystalline anisotropy, the tilt of the sublattice \nmagnetization away from the crystal axes is significant . \n is the angle between the external field \nand the crystal axes ; \n1, \n2 are the angle s between the direction s of the sublattice magnetization \nof domain s D1, D2 and the crystal axes , respectively (\n1 and \n2 are not necessar ily equal) . \n \n9 \n 2.1 Torque magnetometry in the (010) -plane in the low field limit \nThe torque is zero when the applied field is along the easy or hard axis of a sample , because the \ninduced magnetization is in the same direction as the field. Here , we refer to the easy axis as the \nlowest energy orientation of the N éel vector , and define orientations of the hard axis as \nperpendicular to the easy axis . When the applied field is oriented away from an easy or hard axis, \nthe direction of the induced magnetization shifts toward a hard axis because \n|| ⊥ . In an AF, \nthe slope of the torque as a function of field orientation has opposite signs when the field passes \nthrough the orientation of an easy axis and when it passes through the orientation of a hard axis. \nIn our sign convention, torque with a negative slope as a function of angle indicates a hard axis; \ntorque with a positive slope as a function of angle indicates a n easy axis. In a single magnetic \ndomain of Fe2As, there are two hard axes : the c-axis perpendicular to the (001) plane and the axis \nperpendicular to the Néel vector in the (001) plane . \nTorque data at 4 K with the field rotating in the (010) -plane are shown in Fig. 3(a). The slope when \nthe field is along the a-axis (\n= ) is positive at B = 0.5 T and changes to negative at B > 0.5 \nT. Therefore, when the 0.5 T field is oriented along the a-axis, the a-axis is an easy axis but when \nB is greater than 0.5 T, the a-axis becomes a hard axis . This interpretation is consistent with the \nanalysis of the domain distribution discussed above and displayed in Fig. 1(c). When the applied \nfield along the a-axis is larger than 1 T, the majority of domains are in the D2 configuration and \nthe a-axis becomes a hard axis . \nThe slope of torque data when the applied field is along the c-axis (\n= ) is negative because the \nc-axis is a hard axis. However, when \n1= and B > 0.5 T , the sign of the torque changes abruptly . \nThis dramatic change in the sign of the torque is periodic ; the periodicity indicates that domain \nwall motion is reversible. When the applied field is aligned along the c-axis, the populations of \ndomain D1 and D 2 are equal. As the field rotates away from the c-axis, the projection of the applied \nfield in the (001) -plane changes the domain distribution as described by Fig. 1(c). When the field \nreturns to the c-axis, the populations of domain D1 an d domain D2 become equal again. \n10 \n To model the torque data , we first calculate the direction and magnitude of the induced \nmagnetization M by describing the suscept ibility as a tensor \ni ij j jMH = [8]. As shown in Fig. \n2(a), there are two degenerate domains with their Néel vector s perpendicular to each other. For \ndomains of type D1, the Néel vector is along the a-axis and the susceptibility tensor is \n \n||\n100\n0 ' 0\n00D\n\n⊥\n⊥\n= . (3) \nFor domains of type D2, the Néel vector is along the b-axis and the susceptibility tensor is \n \n2 ||' 0 0\n00\n00D\n\n⊥\n⊥\n= . (4) \nThe external field in the (010) -plane is \n( )0sin( 0 cos(THH = ) ) , where 𝜉 is the angle \nbetween the c-axis and the external field as depicted in Fig. 2(a) . We consider the effect of the \nprojection of the applied field along the a-axis 𝐻0sin(𝜉) on the domain distribution as described \nby the data of Fig. 1(c) . The torque signal of two types of domains are\n1 1 1/D D DV= HB and \n2 2 2/D D DV= HB\n, while the total torque is the sum \n12DD+ . To evaluate this model , we \nuse the measured magnetic susceptibility as shown in Fig. 1. The free parameter s are \n/⊥⊥ and\n||\n. \nFig. 3(b) and (c) show the calculated value s of \n1D and \n2D; the solid line s in Fig. 3(a) are the \nsummation of the two. The good correspondence between the model and the data supports our \nassertion that domain wall motion is reversible. \nThe difference in the sign of \n1D and \n2D contributes to the abrupt change in the torque signal near \nan angle of 10°. For D1 domain s, when the applied field is rotating in the (010) -plane, the induced \nmoment is always close to the c-axis because \n|| is small . For D2 domain s, the field -induced \nmagnetization is \n( )0 sin( 0 cos(TMH ⊥⊥ = ) ) . If \n⊥⊥= , the induced magnetization \n11 \n \n( )0sin( 0 cos(TMH ⊥= ) ) , would be parallel to the applied field \nH. This would infer that \nthere is no torque signal from D2 domains and t he total to rque signal would be generated only by \nD1 domains ( Fig. 3(b) ). \nHowever, the total torque signal w e observe is obviously different from what is depicted in Fig. \n3(b) (D1 domains only) . The torque signal resembles a combination of two domains , hence, we \ncan conclude \n⊥⊥ . On the other hand , the dramatic change in the sign of the total torque signal \nat \n1= and T = 4 K indicates that \n1D and \n2D have opposite sign s. Since the induced moment \nof D1 domains is along the c-axis, the induced moment of D2 domains must be between B and the \na-axis. \nThe difference between \n⊥ and\n⊥ is also observed in the dependence of M on H (Fig. 1(b) ). \nFigure 4(a) shows the difference between \ncM and \naM as a function of applied field and \ntemperature. The bump of \ncaMM− at room temperature and below is due to domain wall motion; \nthe difference at high fields is a result of \n⊥⊥− . \nThe magnetization of D2 domains in the (010) -plane experiences an anisotropic environment that \noriginates from the difference of the a- and c-axis of the crystal . By fitting the model to the torque \ndata, we determine \n⊥⊥− as a function of temperature ( Fig. 4(b) ). We observe a change in sign \nof \n⊥⊥− near 200 K that is consistent with the anisotropy in the dependence of M on H (Fig. \n4(a)). This change in sign indicates that the field-induced magnetization of D2 domains is between \nthe applied magnetic field and the a-axis below 200 K, and between the applied magnetic field and \nthe c-axis above 200 K. The anisotropy in the perpendicular susceptibility, \n⊥⊥− , is always \nsmall compare d to its absolute value :\n0.05\n⊥⊥\n⊥− . \n2.2 Torque magnetometry in the (001) -plane under intermediate field \nFigure 5(a) shows torque data acquired with the applied field rotating in the (001) plane . As \nexpected, the data show four-fold symmetry . We attribute the small two-fold symmetry to a \nbackground that comes from misalignment of the sample . (The c-axis is not precisely \n12 \n perpendicular to the field direction. ) The amplitude of the four -fold contribution to the torque as a \nfunction of applied field of three samples from the same growth run is plotted in Fig. 5(b) . \nTo quantify the magnetic anisotropy in the (001) -plane of Fe 2As, we analyze the torque data by \nminimiz ing the total energy for D1 domains and D2 domains , respectively , then add \n1Dτ and \n2Dτ \ntogether to compare it to the data . When\nmE is comparable to or larger than \naniE (\n(001)\nm aniEE ), \nNéel vector s start tilting away from crystal axes as shown in Fig. 2(b) . We assume that the two \nsublattice magnetization s are approximately parallel to each other, so the exchange interaction is \nconsidered in magnetic energy . With the applied field rotating in the (001) -plane of Fe 2As, the \ntotal energy is \n \n22 1 21/ ( ) ( , ) ( ) cos(4 )2T\ntot totE V H H K K K = − + + + , (5) \nwhere \n11 tot D D = , \n1 = for D1 domains and \n22 tot D D = , \n2 = for D2 domains . \nTo obtain an accurate value of 𝐸𝑚, we rotate the susceptibility tensor together with the Néel vector\n( ) ( ) ( )RTRR =\n[8], where \n is the angle between the Néel vector and the crystal axis, and\n()R\n is the rotation matrix: \n \ncos( ) sin( ) 0\n( ) sin( ) cos( ) 0\n0 0 1R\n − \n= (6) \nDuring the rotation, the field component along \n1DM and \n2DM result in a change in the populations \nof D1 and D2 domains . Thus, \n1D and \n2D are determined by the angle between the applied \nfield and the spin axis, \n()+ . \n \nThe analogous behavior of a uniaxial antiferromagnet (AF) [10] provides a point of comparison. \nIn a uniaxial AF , the critical field for the spin -flop transition is \n122( ) / ( )cH K K ⊥ = + −\n . For \nFe2As, as shown in Fig. 1(c) and Fig. 5(b), we do not observe a sudden change in the domain \npopulations that would be characteristic of a spin-flop transition. Furthermore, in a perfect crystal \n13 \n that is free from disorder , the single domain structure created by an ap plied field would persist \nafter the field is removed . (In ferromagnets, domains form to reduce t he contribution of the \nmagnetic energy of stray fields to the total energy . In AF s, this driving force for domain formation \nis absent.) We attribute gradual and reversible domain movement in Fe 2As to random strain fields \ncreated by static disorder in the crystal that create local variations in the anisotropy energy . \nThe midpoint of the chang e in the populations of the D1 and D2 domains as a function of field is, \nhowever, close to what is expected for the characteristic field of a spin -flop transition of an ideal \nsingle domain easy plane antiferromagnetic crystal . The total energy of D2 when \n90=\n is [8] \n \n2 2 2\n1 2 22 2 || 211cos(4 ) ( ) cos ( ) .22totEK K K H HV ⊥⊥ = + + + − + − (7) \nBoth anisotropy and magnetic energy are function of \n2 but with different p eriodicity. For an \napplied field along the <010> direction ( 𝜓 = 0º), \n2( 90 )totE=\n is always a global minimum while \n2( 0 )totE=\n is at a local minimum under low fields and it becomes the global maximum when the \napplied field is larger than a characteristic field \ncH . We can derive this critical field from \n2\n||\nmax\n22() 1arccos2 16 | |H\nK⊥−=\n in which \nmax represents the spin orientation when total energy is at \nglobal maximum under a known applied field. When \nc HH , \nmax 0 ( )c HH =\n . Therefore, \n0 22 16 / ( ) 560 mTcHK ⊥ = − =\n using the value of \n3\n22 150 J/m K=− from our torque \nmagnetometry measurement as described below . \nWe attribute our observations that domains begin to move in a n applied field smaller than \ncH and \ndomain wall motion is not complete until the applied field is larger than \ncH to an inhomogeneous \ndistribution of local values of the magnetocrystalline anisotropy. We speculate that random strains \nin the crystal cause d by point and extended defects create a relatively broad distribution of \nanisotropy at different locations in the sample. When the external field is removed, the random \nstrain field controls the energy of the domain orientation and the volume fraction of domains \nrecovers its initial stat es. \n14 \n The torque induced by the applied field is the derivative of the total energy \n/totdE d= . In the \nintermediate field regime , i.e., \nm aniEE or \nc HH , we assume that the sample is a single \ndomain and the sublattice magnetization is always nearly perpendicular to the applied field, i.e. , \n/2 +\n. The “ intermediate field” regime refers to an applied field larger than the \ncharacteristic field, but not large enough to significantly change the exchange interaction. An \nimportant assumption here is that the tilt of the two sublattice magnetization s in the external field \nis small enough to be neglected . With this approximation, the magnetic energy is nearly \nindependent of \n and \n, \n2\n01\n2mEH ⊥ − . Then, the torque can be easily related to the anis otropy: \n224 sin(4 )tot anidE dEKdd= = −\n, where \n no longer depends on the m agnitude of the applied \nfield. \nA graduation reorientation of domains of an easy -plane antiferromagnet as a function of applied \nfield was also recently observed in 50 nm thick CuMnAs epitaxial layers grown on GaP [27]. \n(CuMnAs and Fe 2As have essentially the same crystal structure wit h Cu and Mn atoms in CuMnAs \noccupying the same lattice sites as the two crystallographically distinct Fe atoms in Fe 2As.) The \nstrength of the field needed to reorient antiferromagnetic domains in CuMnAs epitaxial layers is \nsimilar to what we observe in Fe 2As bulk crystals . Thinner, 10 nm thick, CuMnAs layers show a \npronounced in -plane uniaxial anisotropy and a more abrupt transition in domain structure as a \nfunction of field than 50 nm thick layers. X-ray magnetic linear dichroism (XMLD) measurements \nof CuMnAs epitaxial layers reveals that the domain reorientation is not fully reversible and \nhysteretic for fields less than 2 T. X -ray photoelectron microscopy (XPEEM) images acquired \nafter applying 7 T fields in orthogonal directions also show that the domain structure does not \nrevert to a fixed configuration in zero field. \nBase d on our observation in Fig 5(b), as field increases, \n increases quickly then saturate s. At \nhigher field s, \n is slightly smaller than the saturation value, rather than staying the same until 9 \nT. At higher field s, the tilting of spins caused by the external field cannot be neglected, so \nexchange interaction is no longer a constant . In our model, however, the torque stays the same \nafter saturat ion based on our assumption of constant susceptibility and exchange interaction. This \n15 \n assumption is no longer valid in higher field. While induced magnetization is smaller than \nH , \nthe torque is also smaller than the saturation value. \nAs our model predicts , the experimentally measured torque amplitude saturates as the applied field \napproaches 1 T for sample B, and 3 T for samples A and C. Hence, i t is safe to select 1 T and 3 \nT as the “intermediate field” regime for sample B and for sample s A and C , respectively . The \nmeasured \n22K value of sample A is \n3-150 J/m . The field -dependen ce of \n22K in all three samples \nfollow the same trend ; however, individual data points do not overlap perfectly . We attribute this \ndiscrepancy to variations in the defect microstructures and stoichiometries of the three samples. \nWith a temperature -dependent measurement of torque in the ab-plane at an intermediate field, we \nobtain the temperature -dependence of \n22K as shown in Fig . 6. The overall temperature dependen ce \nis the similar for all three samples with relatively minor differences . As temperature increases, the \nmagnitude of \n22K decreases and becomes close to zero at T > 150 K. \n22K of sample A becomes \nslightly positive for T > 150 K. From Eq. (2), the total energy reaches a minimum when the Néel \nvector is along the crystal a- and b-axis (\n0= or \n90 ) for \n220 K at zero field . When \n220 K , \nthe N éel vector lies in direction s with \n45 = [8]. \n \n3. First -principles calculations of magnetocrystalline anisotropy \nMagnetocrystalline anisotropy of Fe 2As has two contributions, one from spin -orbit interaction \n(SOI) and one from classical magnetic dipole -dipole interaction (MDD) : anisotropy from SOI is \ncalculated using DFT for noncollinear magnetism with spin-orbit coupling , by rotating the Néel \nvector both within the easy plane (001) and out of the plane towards the hard axis (010). The \ncorresponding total -energy changes are visualized in Fig. S2 and the anisotropy energies are then \nobtained by fitting the energy change vs. Néel vector orientation to Eq. (2). This leads to a two -\nfold symmetric SOI anisotropy energy for the Néel vector in the (010) plane and a four -fold \nsymmetric one for the (001) plane. In DFT -LDA , the (010) plane ani sotropy energy is K\n1= -320 \nkJ/m\n3 and K\n22= -290 J/m\n3 for the (001) plane. A non-zero \n22K indicates the existence of two local \n16 \n energy minima. In DFT -PBE, the corresponding values are K\n1= -530 kJ/m\n3 for the (010) plane \nand \n22K = 280 J/m\n3 for the (001) plane. The sign of \n22K differing between DFT -LDA and DFT -\nPBE implies that the energetic ordering of these two minima is inverted. Negative \n22K means that \nthe energy minimum occurs for a Néel vector along a <100> equivalent direction and positive \n22K \nfor a Néel vector along a <110> equivalent direction. \nThe MDD contribution is computed using a classical model that is parametrized using the chemical \nand magnetic ground -state structure from DFT . We use the ground -state chemic al and magnetic \nstructure from DFT -LDA as well as DFT -PBE, to evaluate the following expression for the \nclassical magnetic dipole -dipole interaction and to compare the influence of exchange and \ncorrelation: \n \n2\n0\n53 ( ) ( ) ( ) ( ) 1\n24i ij i ij i j ij\nd\nij ijm r r m r r m r m r r\nEr\n − =− (8) \nTo obtain the anisotropy energy for bulk Fe 2As from this expression, we use an interaction shell \nboundary \nijr of 180 Å, which converges the result to within \n710− eV. This leads to a two -fold \nsymmetric MDD contribution to the anisotropy energy in the (010) plane of K\n1= -220 kJ/m\n3 for \nLDA. For PBE, the corresponding value is K\n1= -300 kJ/m\n3 . The MDD contribution in the (001) \nplane is less than 1 neV and, thus, negligible. \nTherefore , we find a total out -of-plane anisotropy energy of -540 kJ/m\n3 and -830 kJ/m\n3 from LDA \nand PBE, respectively . We attribute ~2/3 of the total out -of-plane anisotropy energy to the SOI \ncontribution and ~1/3 to the MDD contr ibution. Both terms show two -fold symmetry with the hard \naxis along the <001> direction. Torque magnetometry can only measure a lower bound of \n1K\n36 kJ/m\n3 for the out -of-plane anisotropy energy and does not contradict our DFT results. \nThe in -plane anisotropy energy is computed as \n22K= -290 J/m\n3 (DFT -LDA ) and \n22K = 280 J/m\n3 \n(DFT -PBE), while the measured result is \n22K = -150 J/m\n3 . \n17 \n \n4. Antiferromagnetic resonance of easy plane antiferromagnets \nWithout anisotropy, the magnon dispersion of energy in antiferromagnet is zero at the center of \nthe Brillouin zone. Anisotropy introduce s a band gap at the zone center. The antiferromagnetic \nresonance (AFMR) mode we refer to in this work describes this precessional magnetization motion \nat the zone center. With the anisotropy value s we determined by theory and experiment , we can \nmake an est imation of the AFMR frequency. \nWe start from equation s of motion under the ‘macrospin’ approximation of the two magnetic \nsublattice s in domain D1 [28][29]: \n \n1 1 1\n1 22 2 22 2 1ˆ ˆˆ ()bc\nbc\nexdM M MM H H i M H j M H kdT M M = + + − + + − + (9)\n \n \n \n2 2 2\n2 22 1 22 1 1ˆ ˆˆ ()bc\nbc\nexdM M MM H H i M H j M H kdT M M = − − + − + + − + (10) \nWhere \n is the gyromagnetic ratio, \n1M and \n2M are sublattice magnetization s of domain D1, \n1H \nand \n22H are out -of-plane and in -plane anisotropy field s, respectively , which can be written as \n11 / H K M=\n and \n22 22 / H K M= . \nexHM = and \n is the inter -sublattice exchange interaction. \n12aaM M M= − \n, \nbM and \ncM are magnetization component s along the b- and the c-axis, \nrespectively, during spin process ion. \nBecause the two sublattice s along the a-axis are aligned antiparallel to each other, the anisotropy \nfields along the a-axis are of opposite sign s. Along the b- and c-axes, the sign of the effective \nanisotropy field is determined by the signs of \n1,2bM and \n1,2cM . In domain D1, although the \nsublattice magnetization s stay along the a-axis, the in -plane anisotropy is of four-fold symmetry, \nso there is equivalent anisotropy energy contribution along the a- and b-axes. The effective \n18 \n anisotropy field s along the a- and b-axes are determined by the projection o f magnetization on \nthese axes. \nThe only non -zero solution of the equation of motion requires \n12bbMM= and \n12ccMM=− , as \nshown in Fig. S2. The corresponding angular frequency can be expressed as \n22 1 | | 2 ( )exH H H =−\n. \nIn easy -plane AFs, \n10 K and its absolute value is usually much larger than \n22K , thus the \nfrequency is always real. Besides, the AFMR frequency is smaller with smaller \n22 1HH− value , \nbecause the system is more isotropic. For easy -plane materials with \n22 1 0 KK−\n , the AFMR \nfrequency is dominated by the anisotropy in the direction perpendicular to the easy plane. \nFor the exchange field, we use sublattice magnetization \n5\n14 10 A/mDM = and an exchange \nintegral \n1/⊥ with calculated \n⊥ = 0.003 6. We obtain an exchange field \n140 TexH . With \ncalculated \n1K value from DFT -PBE, K\n1= -830 kJ/m\n3 , the AFMR frequency is \nf = 670 GHz . \nFor tetragonal antiferromagnets like Fe 2As, the AFMR is dominated by \n1K because\n1 22KK\n . \nThe same relation is also valid for Mn 2Au where a previous calculation [30] shows that the \nmagnitude of the out-of-plane anisotropy is also much larger than the in-plane anisotropy. It is \nimportant to determine \n1K to estimate AFMR frequency , and both \n1K and \n22K value are needed \nfor thermal stability of spintronics materials. \nAs discussed in Ref s. [2] and [31], the electrical current typically switch es only a small number of \nantiferro magnetic domains . As the anisotropy energy scales with sample volume, the total in-plane \nanisotropy energy is determined by the volum etric difference of the two kinds of domains, \n1 2 22()DD E V V K = −\n. \n12DDVV− depends on temperature, current density and the pulse width [32]. \nIf the attempt frequency is not high enough and \n12DDVV− is small, the magnetic state is more \nsusceptible to thermal fluctuations . \n \n19 \n CONCLUSION \nWe performed torque magnetometry measurement s on an easy-plane antiferromagnet Fe 2As. The \nmeasurement results prove that the domain wall motion in the single -crystalline sample is \nreversible , and allow us to extract the in -plane anisotropy when the magnetic energy \nmE is \ncomparable to magnetocrystalline anisotropy energy\naniE . The in -plane anisotropy of Fe 2As is K22 \n= -150 J/m3 at 4 K . \n22K is strongly temperature -dependent and its magnitude decreases as a function \nof temperature. This means that the domain structure in Fe2As may be easily perturbed by a small \napplied field at room temperature . With \n1K = -830 kJ/m\n3 calculated from DFT, we derived the \nAFMR frequency \n22 1 2 ( )2ex f H H H\n=− = 670 GHz . Our analysis of torque magnetometry \ndata suggest s that the in -plane magnetic anisotropy of some candidate materials for \nantiferromagnetic spintronic applications, such as Fe2As, can be very small at room temperature . \nA field smaller than 1 T is sufficient to significantly alter its domain structure. The measurement \nof \n22K in Fe2As provides a baseline value for further studies of magnetic anisotropy of easy -plane \nantiferromagnets and the motion of antiferro magnetic domain walls . \n \nACKNOWLEDGEMENTS \n \nThis work was undertaken as part of the Illinois Materials Research Science and Engineering \nCenter, supported by the National Science Foundation MRSEC program under NSF Award No. \nDMR -1720633. This work made use of the Illinois Campus Cluster, a computing resource that is \noperated by the Illinois Campus Cluster Program (ICCP) in conjun ction with the National Center \nfor Supercomputing Applications (NCSA) and which is supported by funds from the University \nof Illinois at Urbana -Champaign. This research is part of the Blue Waters sustained -petascale \ncomputing project, which is supported by the National Science Foundation (Awards No. OCI -\n0725070 and No. ACI -1238993) and the state of Illinois. Blue Waters is a joint effort of the \nUniversity of Illinois at Urbana -Champaign and its National Center for Supercomputing \n20 \n Applications. Z.D. acknowled ges support from the Swedish Research Council (VR) under Grant \nNo. 2015 -00585 , co-funded by Marie Skłodowska -Curie Actions (Project INCA 600398) . \n \nREFERENCES \n[1] P. Wadley, B. Howells, J. Železný, C. Andrews, V. Hills, R.P. Campion, V. Novák, K. Olejník, F. \nMaccherozzi, S.S. Dhesi, S.Y. Martin, T. Wagner, J. Wunderlich, F. Freimuth, Y. Mokrousov, J. Kuneš, \nJ.S. Chauhan, M.J. Grzybowski, A.W. Rushforth, K.W. Edm onds, B.L. Gallagher, and T. Jungwirth, \nScience. 351, 587 (2016). \n[2] M.J. Grzybowski, P. Wadley, K.W. Edmonds, R. Beardsley, V. Hills, R.P. Campion, B.L. Gallagher, \nJ.S. Chauhan, V. Novak, T. Jungwirth, F. Maccherozzi, and S.S. Dhesi, Phys. Rev. Lett. 118, 057701 \n(2017). \n[3] K. Olejník, T. Seifert, Z. Kašpar, V. Novák, P. Wadley, R.P. Campion, M. Baumgartner, P. \nGambardella, P. Nemec, J. Wunderlich, J. Sinova, P. Kužel, M. Müller, T. Kampfrath, and T. Jungwirth, \nSci. Adv. 4, eaar3566 (2018). \n[4] M. Meinert, D. Graulich, and T. Matalla -Wagner, Phys. Rev. Appl. 9, 064040 (2018). \n[5] H.J. Richter, J. Phys. D. Appl. Phys. 32, R147 (1999). \n[6] J.J. Nowak, R.P. Robertazzi, J.Z. Sun, G. Hu, J.H. Park, J. Lee, A.J. Annunziata, G.P. Lauer, R. \nKothandar aman, E.J. O’Sullivan, P.L. Trouilloud, Y. Kim, and D.C. Worledge, IEEE Magn. Lett. 7, 1 \n(2016). \n[7] A. V. Khvalkovskiy, D. Apalkov, S. Watts, R. Chepulskii, R.S. Beach, A. Ong, X. Tang, A. Driskill -\nSmith, W.H. Butler, P.B. Visscher, D. Lottis, E. Chen, V. Nikitin, and M. Krounbi, J. Phys. D. Appl. \nPhys. 46, 074001 (2013). \n[8] M. Herak, M. Miljak, G. Dhalenne, and A. Revcolevschi, J. Phys. Condens. Matter 22, 026006 \n(2010). \n[9] Coey, Magnetism and Magnetic Materials (Cambridge University Press, 2009). \n[10] U. Gäfvert, L. Lundgren, B. Westerstrandh, and O. Beckman, J. Phys. Chem. Solids 38, 1333 (1977). \n[11] C. Zener, Phys. Rev. 96, 1335 (1954). \n[12] J.C. Burgiel and M.W.P. Strandberg, J. Phys. Chem. Solids 26, 877 (1965). \n[13] P. Pincus, Phys. Rev. 113, 769 (1959). \n[14] H. Katsuraki and N. Achiwa, J. Phys. Soc. Japan 21, 2238 (1966). \n[15] H. Katsuraki and K. Suzuki, J. Appl. Phys. 36, 1094 (1965) . \n[16] N. Achiwa, S. Yano, M. Yuzuri, and H. Takaki, J. Phys. Soc. Japan 22, 156 (1967). \n[17] K. Yang, K. Kang, Z. Diao, A. Ramanathan, M.H. Karigerasi, D.P. Shoemaker, A. Schleife, and \nD.G. Cahill, Phys. Rev. Mater. 3, 124408 (2019). \n[18] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996). \n21 \n [19] G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999). \n[20] P.E. Blöchl, Phys. Rev. B 50, 17953 (1994). \n[21] H.J. Monkhorst and J.D. Pack, Phys. Rev. B 13, 5188 (1976). \n[22] D.M. Ceperley and B.J. Alder, Phys. Rev. Lett. 45, 566 (1980). \n[23] J.P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996). \n[24] D. Hobbs, G. Kresse, and J. Hafner, Phys. Rev. B 62, 11556 (2000). \n[25] A. Schrön, C. Rödl, and F. Bech stedt, Phys. Rev. B 86, 115134 (2012). \n[26] T. Nagamiya, K. Yosida, and R. Kubo, Adv. Phys. 4, 1 (1955). \n[27] M. Wang, C. Andrews, S. Reimers, O.J. Amin, P. Wadley, R.P. Campion, S.F. Poole, J. Felton, \nK.W. Edmonds, B.L. Gallagher, A.W. Rushforth, O. Makar ovsky, K. Gas, M. Sawicki, D. Kriegner, K. \nOlejník, V. Novák, T. Jungwirth, M. Shahrokhvand, U. Zeitler, S.S. Dhesi, and F. Maccherozzi, Phys. \nRev. B 101, 094429 (2020). \n[28] C. Kittel, Introduction to Solid State Physics (Wiley, 1953). \n[29] C. Kittel, Phy s. Rev. 82, 565 (1951). \n[30] A.B. Shick, S. Khmelevskyi, O.N. Mryasov, J. Wunderlich, and T. Jungwirth, Phys. Rev. B 81, \n212409 (2010). \n[31] P. Wadley, S. Reimers, M.J. Grzybowski, C. Andrews, M. Wang, J.S. Chauhan, B.L. Gallagher, R.P. \nCampion, K.W. Edmon ds, S.S. Dhesi, F. Maccherozzi, V. Novak, J. Wunderlich, and T. Jungwirth, Nat. \nNanotechnol. 13, 362 (2018). \n[32] T. Matalla -Wagner, M.F. Rath, D. Graulich, J.M. Schmalhorst, G. Reiss, and M. Meinert, Phys. Rev. \nAppl. 12, 064003 (2019). \n \n \n \n \n \n \n \n \n \n \n \n \n22 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n (a) (b) (c) \nFigure 1. (a) Temperature dependence of the magnetic susceptibility of Fe 2As in the low field \nlimit as measured using 10 mT field applied along the a-axis (blue data points) and c-axis (black \ndata points) of the crystal . (b) Dependence of Fe 2As magnetization M on applied field H at T \n= 4 K. With H along the c-axis (red line), M is a linear function of H. With H along the a-axis \n(black curve), the non -linear dependence of M on H is due to the rotation of antiferromagnetic \ndomains. (c) The population of domain s with Néel vectors parallel and perpendicular to the \napplied field estimated from the dependence of M on H. The assumptions are: 1) in zero field, \nthe population of domains with Néel vectors in the a and b directions are equal; and 2) in the \nhigh field limit, the Néel vector is perpendicular to the applied field. \n0.0 0.5 1.0 1.5 2.0051015202530\nH || cH || aM (kA m-1)\nField (T)\n0.0 0.5 1.0 1.5 2.00.00.20.40.60.81.0\nField (T)⊥\n||\n0 100 200 300 4000.0000.0050.0100.015\nTemperature (K)ac \n23 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Figure 2. Geometry of the torque magnetometry experiments. The a-b-c coordinate s are the \ncrystal axes. MD1 and MD2 are the sublattice magnetization s of the two types of domains labeled \nas D1 and D2. (a) The magnetic unit cell of Fe 2As. (b) Three -dimensional perspective of the \nmeasurement with the magnetic field rotating in the ac-plane. The magnetic field makes an \nangle \n with the c-axis of the crystal. MD1 and MD2 are assumed to stay along the a- and the b-\naxis, respectively. T he torque is along the b-axis. ( c) Plan-view of the measurement with the \nmagnetic field rotating in the ab-plane. The magnetic field makes an angle \n with the b-axis \nof the crystal. MD1 and MD2 tilt away from a- and b-axis by \n1 and \n2 , respectively (\n1 and \n2 \nare not necessarily the same) . The torque is along the c-axis (normal to the plane of the drawing) . (a) (b) \n (c) \n \n24 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n (a) (b) (c) \nFigure 3. (a) Torque magnetometry measurement s in the ac-plane of Fe 2As at T = 4 K. Open \nsymbols are measured data; solid lines are fits to the data. The legend gives the magnitude of \nthe applied field labeled by color. (b) Calculated torque generated by domains of type D1 as a \nfunction of applied field. (c) Calculated torque generated by domains of type D2 as a function \nof applied field. \n0 90 180 270 360-4-3-2-101234\n3 T\n2.5 T\n2 T\n1.5 T\n1 T\n0.5 TTorque (kN m-2)\n (deg)\n0 90 180 270 360-4-3-2-101234Torque (kN m-2)\n (deg)3 T\n2.5 T\n2 T\n1.5 T\n1 T\n0.5 T\n0 90 180 270 360-4-3-2-101234\n3 T\n2.5 T\n2 T\n1.5 T\n1 T\n0.5 T\n (deg)Torque (kN m-2) \n25 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Figure 4 . (a) The field dependence of the difference \ncaMM− between the magnetization with \nan applied field along the c direction \ncM and the field applied along the a direction \naM . Each \ncurve is labeled by the measurement temperature. When the applied field along the a-axis is \nlarger than 1.5 T , all domains can be treated as equivalent to MD2. Therefore , the slope of the \ndata for magnetic fields larger than 1.5 T is the difference in the susceptibility \n⊥⊥− , where \n⊥\n is the susceptibility in ab -plane perpendicular to MD2 and \n⊥ is the susceptibility along c-\naxis perpendicular to MD2. (b) Comparison of the temperature dependent of \n⊥⊥− value from \ndirect measurements of the type shown in panel (a) and from fitting the torque data . (a) (b) \n0.0 0.5 1.0 1.5 2.0-2.0-1.5-1.0-0.50.00.51.0Mc-Ma (kA m-1)\nField (T)4 K77 K200 K300 K\n360 K\n0 50 100 150 200 250 300 350 400-6-4-2024681012\nTorque fittingLinear fitting Mc- Ma \nat high field'⊥-⊥ (10-4)\nTemperature (K) \n26 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Figure 5 . (a) Torque magnetometry measurement s of Fe 2As (sample A ) in the ab-plane at T=4 \nK. (b) The amplitude of the four-fold component of the torque extracted from measurements of \nthe type shown in panel (a) at T = 4 K and compared to an analytical model (see text) . When \namplitude of the four -fold component of the torque saturates at the value \n0 , the in -plane \nanisotropy is 𝜏0 ≈ 4𝐾22. \n (a) (b) \n0 2 4 6 8 100200400600800Torque amplitude (N m-2)\nField (T)sample B\nsample A\nsample C\nModel\n0 90 180 270 360-1.0-0.8-0.6-0.4-0.20.00.20.40.60.81.0Torque (kN m-2)\n (deg)3 T\n2 T\n1.5 T\n1 T\n0.5 T \n27 \n \n \n \n \nFigure 6. Temperature dependence of in-plane magnetocrystalline anisotropy 𝐾22 of Fe 2As. The torque data \nwith the external field rotating in the (001) plane were measured with an intermediate field strength (3 T for \nsample A and sample C , and 1 T for sample B ) and th e amplitude of the four-fold symmetry was extracted to \nobtain the in -plane anisotropy with 𝜏0 ≈ 4𝐾22. Intermediate field is defined as a field strength under which the \ntorque amplitude of four-fold symmetry saturates at 4 K. The error bars represent 20% uncertainty in \ndetermining the saturation value of the torque amplitude. All three samples have the same uncertainty, and we \nonly plot error bar for sample A to ensure that the plot is free from cluttering. \n0 100 200 300 400-200-160-120-80-400K22 (J m-3)\nTemperature (K) sample A\n sample B\n sample C" }, { "title": "2007.08624v1.Anisotropic_magnetocaloric_effect_and_critical_behavior_in_CrSbSe__3_.pdf", "content": "arXiv:2007.08624v1 [cond-mat.str-el] 15 Jul 2020Anisotropic magnetocaloric effect and critical behavior in CrSbSe 3\nYu Liu,1Zhixiang Hu,1,2and C. Petrovic1,2\n1Condensed Matter Physics and Materials Science Department ,\nBrookhaven National Laboratory, Upton, New York 11973, USA\n2Materials Science and Chemical Engineering Department,\nStony Brook University, Stony Brook, New York 11790, USA\n(Dated: July 20, 2020)\nWe report anisotropic magnetocaloric effect and critical be havior in quasi-one-dimensional ferro-\nmagnetic CrSbSe 3single crystal. The maximum magnetic entropy change −∆Smax\nMis 2.16 J kg−1\nK−1for easy aaxis (2.03 J kg−1K−1for hard baxis) and the relative cooling power RCPis 163.1\nJ kg−1for easy aaxis (142.1 J kg−1for hard baxis) near Tcwith a magnetic field change of 50\nkOe. The magnetocrystalline anisotropy constant Kuis estimated to be 148.5 kJ m−3at 10 K,\ndecreasing to 39.4 kJ m−3at 70 K. The rescaled ∆ SM(T,H) curves along all three axes collapse\nonto a universal curve, respectively, confirming the second order ferromagnetic transition. Further\ncritical behavior analysis around Tc∼70 K gives that the critical exponents β= 0.26(1), γ=\n1.32(2), and δ= 6.17(9) for H/bardbla, whileβ= 0.28(2), γ= 1.02(1), and δ= 4.14(16) for H/bardblb. The\ndetermined critical exponents suggest that the anisotropi c magnetic coupling in CrSbSe 3is strongly\ndependent on orientations of the applied magnetic field.\nI. INTRODUCTION\nLow-dimensional ferromagnetic (FM) semiconductors,\nholding both ferromagnetism and semiconducting char-\nacter, form the basis for nano-spintronics application.\nRecently, the two-dimensional (2D) CrI 3and Cr 2Ge2Te6\nhave attracted much attention since the discovery of in-\ntrinsic 2D magnetism in mono- and few-layer devices.1–4\nIntrinsic magnetic order is not allowed at finite temper-\nature in low-dimensional isotropic Heisenberg model by\nthe Mermin-Wagner theorem,5however, a large magne-\ntocrystalline anisotropy removes this restriction, for in-\nstance,thepresenceofamagneticallyorderedstateinthe\n2D Ising model. The enhanced fluctuations in 2D limit\nmake symmetry-breaking order unsustainable, however\nby gapping the low-energy modes through the introduc-\ntion of anisotropy, order could be established by provid-\ning stabilization of long-range correlations in 2D limit.\nGiven the reduced crystal symmetry in low-dimensional\nmagnets, an intrinsic magnetocrystalline anisotropy can\nbe expected and points to possible long-range magnetic\norder in atomic-thin limit.\nIn ternary chromium trichalcogenides, Cr(Sb,Ga)X 3\n(X = S, Se, Te) displays a pseudo-one-dimensional crys-\ntal structure. This is different from Cr(Si,Ge)Te 3that\nfeatures layered structure and a van der Waals bonds\nalong the c-axis. In Cr(Sb,Ga)X 3, the CrX 6octahedra\nform infinite, edge-sharing, double rutile chains. The\nneighboring chains are linked by Sb or Ga atoms. The\nFM semiconductor CrSbSe 3has attracted considerable\nattention.6–9A band gap of0.7 eV was determined by re-\nsistivity and optical measurements.8The Cr in CrSbSe 3\nexhibits a high spin state with S= 3/2, and orders fer-\nromagnetically below the Curie temperature Tcof 71 K.8\nFM state in CrSbSe 3is fairly anisotropic with the aaxis\nbeing the easy axis and the baxis being the hard axis.\nThe critical analysis where magnetic field was applied\nalong the aaxis suggests that the ferromagnetism inCrSbSe 3cannot be simply described by the mean-field\ntheory.10,11This invites the detailed investigation on its\nanisotropic critical behavior.\nThe magnetocaloric effect (MCE) can give additional\ninsight into the magnetic properties, and it could be also\nused to assess magnetic refrigeration potential.12–20Bulk\nCrSiTe 3exhibits anisotropic entropy change ( −∆Smax\nM)\nwith the values of 5.05 and 4.9 J kg−1K−1at 50 kOe\nfor out-of-plane and in-plane fields, respectively, with the\nmagnetocrystalline anisotropy constant Kuof 65 kJ m−3\nat 5 K.14The values of −∆Smax\nMare about 4.24 J kg−1\nK−1(out-of-plane) and 2.68 J kg−1K−1(in-plane) at\n50 kOe for CrI 3with a much larger Kuof 300±50 kJ\nm−3at 5 K.21The large magnetocrystalline anisotropy\nis important in preserving FM in the 2D limit.\nInthepresentworkweinvestigatetheanisotropicmag-\nnetic properties of pseudo-one-dimensional CrSbSe 3sin-\ngle crystals. The magnetocrystalline anisotropy constant\nKuis strongly temperature-dependent.It takes a value of\n∼148.5 kJ m−3at 10 K and monotonically decreases\nto 39.4 kJ m−3at 70 K for the hard baxis. The Ku\nof CrSbSe 3is much larger than that of Cr 2(Si,Ge) 2Te6\nbut comparable with that of Cr(Br,I) 3. This results in\nanisotropicmagneticentropychange∆ SM(T,H)andrel-\native cooling power (RCP), as well as in magnetic criti-\ncal exponents β,γ, andδthat point to the nature of the\nphase transition. The anisotropic magnetic coupling of\nCrSbSe 3is strongly dependent on orientations of the ap-\nplied magnetic field, providing an excellent platform for\nfurthertheoreticalstudiesoflow-dimensionalmagnetism.\nII. EXPERIMENTAL DETAILS\nCrSbSe 3single crystals were fabricated by the self-flux\ntechniquestartingfrom anintimate mixtureofrawmate-\nrials Cr (99.95%, Alfa Aesar) powder, Sb (99.999%, Alfa\nAesar) pieces, and Se (99.999%, Alfa Aesar) pieces with2\nFIG. 1. (Color online). (a) Crystal structure of CrSbSe 3and\n(b) representative single crystals on a millimeter-grid pa per.\n(c) Powder x-ray diffraction (XRD) pattern of CrSbSe 3. (d)\nTemperature-dependent magnetization M(T) measured in H\n= 1 kOe with zero field cooling (ZFC) and field cooling (FC)\nmodes along all three axes (left axis) and inverseaverage ma g-\nnetization 1 /Mave= 3/(Ma+Mb+Mc) (right axis) fitted by\nthe Curie-Weiss law. Inset shows the field-dependent magne-\ntization M(H) at 2 K.\na molar ratio of 7 : 33 : 60. The starting materials were\nsealed in an evacuated quartz tube and then heated to\n800◦C and slowlycooled to 680◦C with a rate of2◦C/h.\nNeedle-like single crystals with lateral dimensions up to\nseveral millimeters can be obtained. The element analy-\nsis was performed using an energy-dispersive x-ray spec-\ntroscopy in a JEOL LSM-6500 scanning electron micro-\nscope (SEM), confirming a stoichiometric CrSbSe 3single\ncrystal. The powder x-ray diffraction (XRD) data were\ntaken with Cu K α(λ= 0.15418 nm) radiation of Rigaku\nMiniflex powder diffractometer. The anisotropy of mag-\nnetic properties were measured by using one single crys-\ntal with mass of 0.32 mg and characterized by the mag-\nnetic property measurementsystem (MPMS-XL5, Quan-\ntum Design). The applied field ( Ha) was corrected as\nH=Ha−NM, whereMis the measured magnetization\nandNis the demagnetization factor. The corrected H\nwas used for the analysis of magnetic entropy change and\ncritical behavior.\nIII. RESULTS AND DISCUSSIONS\nFigure 1(a) displays the CrSbSe 3crystal structure.\nThe material crystalizes in an orthorhombic lattice with\nthe space group of Pnma. That is a pseudo-one-\ndimensional structure with double rutile chains of CrSe 6\noctahedra that are aligned parallel to the baxis. As\nshown in Fig. 1(b), the baxis is along the long crystal\ndimension in single crystals. Within the double chain,\nthe Cr cations form an edge-sharing triangular arrange-\nment, while the Sb atoms link the adjacent chains. Inthe powder XRD pattern [Fig. 1(c)], all peaks can be\nwell indexed by the orthorhombic structure (space group\nPnma) with lattice parameters a= 9.121(2) ˚A,b=\n3.785(2) ˚A andc= 13.383(2) ˚A, in good agreement with\nprevious report.8\nFigure1(d)showsthetemperaturedependenceofmag-\nnetization M(T) along all three axes measured at H= 1\nkOe. There is no bifurcation seen between the zero-field-\ncooling (ZFC) and field-cooling (FC) curves, indicating\nthe high quality of single crystal. The M(T) curves are\nnearly isotropic at high temperature but show an obvi-\nous anisotropic magnetic response for field applied along\ndifferent axes at low temperature. For H/bardbla, a rapid in-\ncrease near 70 K in M(T) on cooling corresponds well to\nthe reported paramagnetic (PM) to FM transition.8For\nH/bardblbandH/bardblc, an anomalous peak feature is observed,\nwhich is also seen in Cr 2(Si,Ge) 2Te6and Cr(Br,I) 3with\na large magnetocrystalline anisotropy.22–24The inverse\naverage magnetization 1 /Mave= 3/(Ma+Mb+Mc) can\nbe well fitted from 200 to 300 K by using the Curie-\nWeiss law, which generates an effective moment of 4.6(1)\nµB/Cr and a positive Weiss temperature of 125(1) K, in\nline with the values reported for CrSbSe 3polycrystals.6,7\nThe anisotropic magnetization isotherms measured at T\n= 2 K [inset in Fig. 1(d)] show a similar saturated mag-\nnetization Msat 3µB/Cr, consistent with expectation of\nS= 3/2 for Cr3+, but different saturated fields Hsof 1\nkOe, 18kOe, and 12kOefor a,b, andcaxis, respectively,\nclose to the values in the previous reports.8,9\nTo further characterize the anisotropic magnetic prop-\nertiesofCrSbSe 3,theisothermalmagnetizationwithfield\nup to 50 kOe applied along each axis from 10 to 100 K\nare presented in Figs. 2(a)-2(c). At high temperature,\nthe curves have linear field dependence. With decreasing\ntemperature, the curves bend with negative curvatures,\nindicating dominant FM interaction. Based on the clas-\nsical thermodynamics and Maxwell’s relation, the mag-\nnetic entropy change ∆ SM(T,H) is given by:25,26\n∆SM=/integraldisplayH\n0[∂S(T,H)\n∂H]TdH=/integraldisplayH\n0[∂M(T,H)\n∂T]HdH,\n(1)\nwhere [∂S(T,H)/∂H]T= [∂M(T,H)∂T]His based on\nthe Maxwell’s relation. For magnetization measured at\nsmall temperature and field intervals,\n∆SM=/integraltextH\n0M(Ti+1,H)dH−/integraltextH\n0M(Ti,H)dH\nTi+1−Ti.(2)\nThe calculated −∆SM(T,H) are presented in Figs. 2(d)-\n2(f). All the curves exhibit a peak feature near Tc.\nPeaks broads asymmetrically on both sides with increas-\ning field. The maximum of −∆SM(T,H) reach 2.16 J\nkg−1K−1, 2.11 J kg−1K−1, and 2.03 J kg−1K−1fora,\nb, andcaxis, respectively. It should be noted that all\nthe values of −∆SM(T,H) for easy aaxis are positive,\nhowever, the values for hard bandcaxes are negative at\nlow temperatures in low fields stemming from a strong\ntemperature-dependent magnetic anisotropy.183\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48/s48/s49/s48/s50/s48/s51/s48/s52/s48\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48/s48/s49/s48/s50/s48/s51/s48/s52/s48\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48/s48/s49/s48/s50/s48/s51/s48/s52/s48/s48 /s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48/s48/s49/s50\n/s48 /s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48/s48/s49/s50\n/s48 /s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48/s48/s49/s50/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s72/s32/s40/s107/s79/s101/s41/s49/s48/s32/s75\n/s49/s48/s48/s32/s75\n/s72/s32/s47/s47/s32/s97/s40/s97/s41\n/s49/s48/s48/s32/s75/s49/s48/s32/s75\n/s40/s98/s41/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s72/s32/s40/s107/s79/s101/s41/s72/s32/s47/s47/s32/s98\n/s49/s48/s32/s75\n/s49/s48/s48/s32/s75/s40/s99/s41/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s72/s32/s40/s107/s79/s101/s41/s72/s32/s47/s47/s32/s99/s53 /s32/s32/s32/s32/s32/s32/s32 /s32/s49/s48\n/s49/s53 /s32/s32/s32 /s32/s50/s48\n/s50/s53 /s32/s32/s32 /s32/s51/s48\n/s51/s53/s32 /s32/s52/s48\n/s52/s53 /s32/s32/s32 /s32/s53/s48/s72/s32/s47/s47/s32/s97/s40/s100/s41/s45 /s83\n/s77/s40/s74/s47/s107/s103/s45/s75/s41\n/s84/s32/s40/s75/s41/s107/s79/s101\n/s107/s79/s101/s72/s32/s47/s47/s32/s98/s40/s101/s41/s45 /s83\n/s77/s40/s74/s47/s107/s103/s45/s75/s41\n/s84/s32/s40/s75/s41/s53 /s32/s32/s32/s32/s32/s32/s32 /s32/s49/s48\n/s49/s53 /s32/s32/s32 /s32/s50/s48\n/s50/s53 /s32/s32/s32 /s32/s51/s48\n/s51/s53/s32 /s32/s52/s48\n/s52/s53 /s32/s32/s32 /s32/s53/s48\n/s107/s79/s101\n/s53 /s32/s32/s32/s32/s32/s32/s32 /s32/s49/s48\n/s49/s53 /s32/s32/s32 /s32/s50/s48\n/s50/s53 /s32/s32/s32 /s32/s51/s48\n/s51/s53/s32 /s32/s52/s48\n/s52/s53 /s32/s32/s32 /s32/s53/s48/s72/s32/s47/s47/s32/s99\n/s40/s102/s41/s45 /s83\n/s77/s40/s74/s47/s107/s103/s45/s75/s41\n/s84/s32/s40/s75/s41\nFIG. 2. (Color online). (a-c) Typical initial isothermal ma g-\nnetization curves measured along all three axes with temper -\nature ranging from 10 to 100 K. (d-f) The corresponding cal-\nculated magnetic entropy change −∆SM(T) at various fields\nchange.\nBased on a generalized scaling analysis,27the normal-\nized magnetic entropy change, ∆ SM/∆Smax\nM, estimated\nfor each constant field, is scaled to the reduced tempera-\nturetas defined in the following equations:\nt−= (Tpeak−T)/(Tr1−Tpeak),T < T peak,(3)\nt+= (T−Tpeak)/(Tr2−Tpeak),T > T peak,(4)\nwhereTr1andTr2arethe lowerand upper referencetem-\nperatures at half-width full maximum of ∆ SM/∆Smax\nM.\nAs we can see, the normalized ∆ SM/∆Smax\nMnearTccol-\nlapses onto a universal curve at the indicated fields for\nall three axes [Figs. 3(a)-3(c)], indicating the second-\norder PM-FM transition in CrSbSe 3. The ineligible devi-\nation at low temperatures along the hard baxis is mostly\ncontributed by the magnetocrystalline anisotropy effect.\nBasedon the Stoner-Wolfarthmodel,28the magnetocrys-\ntalline anisotropy constant Kucan be estimated from\nthe saturation regime in the isothermal magnetization\ncurves. Within this model the value of Kuin single do-\nmain state is related to the saturation magnetization Ms\nand the saturation field Hswithµ0is the vacuum per-\nmeability:\n2Ku\nMs=µ0Hs. (5)/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48/s48/s49/s50\n/s48/s49/s50/s45/s54 /s45/s52 /s45/s50 /s48 /s50/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s45/s56 /s45/s54 /s45/s52 /s45/s50 /s48 /s50/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s45/s56 /s45/s54 /s45/s52 /s45/s50 /s48 /s50/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s48 /s50/s48 /s52/s48 /s54/s48 /s56/s48/s48/s52/s48/s56/s48/s49/s50/s48/s49/s54/s48\n/s48 /s52/s48 /s56/s48/s50/s48/s51/s48/s52/s48/s48 /s52/s48 /s56/s48/s53/s49/s48/s49/s53/s48 /s50/s53 /s53/s48/s52/s48/s56/s48/s49/s50/s48\n/s48 /s50/s53 /s53/s48/s52/s48/s56/s48/s49/s50/s48\n/s48 /s50/s53 /s53/s48/s52/s48/s56/s48/s49/s50/s48\n/s45 /s83/s109 /s97/s120\n/s77/s61/s32/s97/s72/s110\n/s72/s47/s47/s97\n/s72/s47/s47/s98\n/s72/s47/s47/s99/s45 /s83/s109/s97/s120 /s77\n/s40/s74/s47/s107/s103/s45/s75/s41\n/s72/s32/s40/s107/s79/s101/s41/s40/s101/s41\n/s82/s67/s80/s61/s32/s98/s72/s109 /s72/s47/s47/s97\n/s72/s47/s47/s98\n/s72/s47/s47/s99\n/s82/s67/s80/s32/s40 /s49/s48/s50\n/s74/s47/s107/s103/s41/s32/s32/s32/s83\n/s77/s47 /s83/s109/s97/s120 /s77\n/s116/s32/s32/s32/s107/s79/s101\n/s32/s49/s48/s32 /s32/s50/s48/s32 /s32/s51/s48\n/s32/s52/s48/s32 /s32/s53/s48/s40/s97/s41\n/s72/s32/s47/s47/s32/s97/s72/s32/s47/s47/s32/s98/s40/s98/s41\n/s32/s32/s32\n/s116/s72/s32/s47/s47/s32/s99/s40/s99/s41\n/s32/s32/s32\n/s116\n/s40/s100/s41\n/s32/s32/s75\n/s117/s32/s40/s107/s74/s47/s109/s51\n/s41\n/s84/s32/s40/s75/s41/s72/s32/s47/s47/s32/s98/s77\n/s115/s32/s40/s101/s109/s117/s47/s103/s41\n/s84/s32/s40/s75/s41\n/s72\n/s115/s32/s40/s107/s79/s101/s41\n/s84/s32/s40/s75/s41/s72/s32/s40/s107/s79/s101/s41/s32/s84\n/s50\n/s32/s84\n/s49/s32/s84\n/s50\n/s32/s84\n/s49\n/s72/s32/s40/s107/s79/s101/s41/s32/s84\n/s50\n/s32/s84\n/s49\n/s72/s32/s40/s107/s79/s101/s41\nFIG. 3. (Color online). (a-c) Normalized magnetic entropy\nchange ∆ SMas a function of the reduced temperature t\nalong all three principal crystallographic axes of CrSbSe 3. In-\nsets show the evolution of the reference temperatures T1and\nT2. (d) Temperature dependence of the magnetocrystalline\nanisotropy constant Ku, the saturation field Hs, and the sat-\nuration magnetization Ms(insets) estimated from the hard\nbaxis below Tcfor CrSbSe 3. (e) Field dependence of the\nmaximum magnetic entropy change −∆Smax\nMand the rela-\ntive cooling power (RCP) with power-law fitting in solid line s\nalong all three axes for CrSbSe 3.\nWhenH/bardblb, the anisotropy becomes maximal. We esti-\nmated the Msby usinga linearfit of M(H) above20kOe\nforH/bardblb, which monotonically decreases with increasing\ntemperature[insetinFig. 3(d)]. Thenwedeterminedthe\nHsas the intersection point of two linear fits: one being\na fit to the saturated regime at high field, and the other\nbeing a fit of the unsaturated linear regime at low field.\nThe values of Hssharea similartemperature dependence\n[inset in Fig. 3(d)], resulting in a strongly temperature-\ndependent Ku[Fig. 3(d)]. The calculated Kuis∼148.5\nkJ m−3at 10 K for CrSbSe 3, which is much larger than\nthat of Cr 2(Si,Ge) 2Te6,18and comparable with that of\nCr(Br,I) 3.21\nAnother parameter to characterize the potential mag-\nnetocaloric effect of materials is the relative cooling\npower (RCP):29\nRCP=−∆Smax\nM×δTFWHM, (6)\nwhere the FWHM denotes the full width at half max-\nimum of −∆SMcurve. The RCPcorresponds to the4\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53 /s50/s46/s48/s48/s50/s52/s54/s56/s49/s48\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53 /s50/s46/s48/s48/s50/s52/s54/s56/s49/s48\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53 /s50/s46/s48/s48/s50/s52/s54/s56/s49/s48/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53/s48/s49/s50/s51/s52/s53/s54\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53/s48/s49/s50\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53/s48/s49/s50/s51/s52/s53/s54/s77/s50\n/s32/s40 /s49/s48/s50\n/s101/s109/s117/s50\n/s47/s103/s50\n/s41\n/s72/s47/s77/s32/s40/s107/s79/s101/s45/s103/s47/s101/s109/s117/s41/s40/s97/s41/s72/s32/s47/s47/s32/s97\n/s72/s32/s47/s47/s32/s98/s40/s98/s41/s77/s50\n/s32/s40 /s49/s48/s50\n/s101/s109/s117/s50\n/s47/s103/s50\n/s41\n/s72/s47/s77/s32/s40/s107/s79/s101/s45/s103/s47/s101/s109/s117/s41\n/s72/s32/s47/s47/s32/s99/s40/s99/s41/s77/s50\n/s32/s40 /s49/s48/s50\n/s101/s109/s117/s50\n/s47/s103/s50\n/s41\n/s72/s47/s77/s32/s40/s107/s79/s101/s45/s103/s47/s101/s109/s117/s41/s72/s32/s47/s47/s32/s97/s40/s100/s41\n/s32/s32/s77/s49/s47\n/s32/s40 /s49/s48/s53\n/s40/s101/s109/s117/s47/s103/s41/s49/s47\n/s41\n/s40/s72/s47/s77/s41/s49/s47\n/s32/s40/s107/s79/s101/s45/s103/s47/s101/s109/s117/s41/s49/s47\n/s72/s32/s47/s47/s32/s98/s40/s101/s41\n/s32/s32/s77/s49/s47\n/s32/s40 /s49/s48/s53\n/s40/s101/s109/s117/s47/s103/s41/s49/s47\n/s41\n/s40/s72/s47/s77/s41/s49/s47\n/s32/s40/s107/s79/s101/s45/s103/s47/s101/s109/s117/s41/s49/s47\n/s72/s32/s47/s47/s32/s99/s40/s102/s41\n/s32/s32/s77/s49/s47\n/s32/s40 /s49/s48/s53\n/s40/s101/s109/s117/s47/s103/s41/s49/s47\n/s41\n/s40/s72/s47/s77/s41/s49/s47\n/s32/s40/s107/s79/s101/s45/s103/s47/s101/s109/s117/s41/s49/s47\nFIG. 4. (Color online). The Arrott plot of M2vsH/M(a-c)\nand the modified Arrott plot of M1/βvs (H/M)1/γ(d-f) with\nindicated βandγfora,bandcaxis, repsectively.\nTABLE I. The values of magnetic entropy change ( −∆Smax\nM)\nand relative cooling power (RCP) at field change of 50 kOe.\nCritical exponents of CrSbSe 3alonga,b, andcaxis, recpec-\ntively. The MAP, KFP, and CI represent the modified Arrott\nplot, theKouvel-Fisherplot, and thecritical isotherm, re spec-\ntively.\n−∆Smax\nMRCPTcβ γ δ\n(J/kg-K) (J/kg)\nH/bardbla2.16 163.1\nMAP 70.2(1) 0.26(1) 1.32(2) 6.08(12)\nKFP 70.2(1) 0.26(1) 1.33(2) 6.12(12)\nCI 6.17(9)\nH/bardblb2.03 142.1\nMAP 70.1(2) 0.28(2) 1.02(1) 4.64(22)\nKFP 70.3(2) 0.32(2) 1.03(1) 4.22(17)\nCI 4.14(16)\nH/bardblc2.11 154.1\nMAP 69.8(2) 0.26(1) 1.14(5) 5.38(2)\nKFP 69.8(2) 0.25(1) 1.19(4) 5.76(2)\nCI 5.35(17)/s49 /s49/s48/s49/s48/s50/s48/s51/s48\n/s54/s48 /s54/s50 /s54/s52 /s54/s54 /s54/s56 /s55/s48 /s55/s50 /s55/s52 /s55/s54 /s55/s56 /s56/s48/s48/s49/s48/s50/s48/s51/s48/s52/s48/s32\n/s84/s32/s40/s75/s41/s77\n/s115/s32/s40/s101/s109/s117/s47/s103/s41/s40/s97/s41\n/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53\n/s32 /s32/s40/s107/s79/s101/s45/s103/s47/s101/s109/s117/s41\n/s54/s48 /s54/s50 /s54/s52 /s54/s54 /s54/s56 /s55/s48 /s55/s50 /s55/s52 /s55/s54 /s55/s56 /s56/s48/s45/s52/s48/s45/s51/s48/s45/s50/s48/s45/s49/s48/s48\n/s32/s72/s32/s47/s47/s32/s97 /s32\n/s32/s72/s32/s47/s47/s32/s98 /s32\n/s32/s72/s32/s47/s47/s32/s99 /s32/s40/s98/s41/s32/s32/s72/s32/s47/s47/s32/s97 /s32\n/s32/s72/s32/s47/s47/s32/s98 /s32\n/s32/s72/s32/s47/s47/s32/s99 /s32\n/s84/s32/s40/s75/s41/s77\n/s115/s40/s100/s77\n/s115/s47/s100/s84/s41/s45/s49\n/s32/s40/s75/s41\n/s48/s50/s52/s54/s56/s49/s48\n/s40/s100/s45/s49 /s48/s47/s100/s84/s41/s45/s49/s32/s40/s75/s41/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s72/s32/s40/s107/s79/s101/s41/s32/s72/s32/s47/s47/s32/s97\n/s32/s72/s32/s47/s47/s32/s98\n/s32/s72/s32/s47/s47/s32/s99/s84\n/s99/s32/s61/s32/s55/s48/s32/s75\nFIG. 5. (Color online). (a) Temperature dependence of the\nspontaneous magnetization Ms(left) and the inverse initial\nsusceptibility χ−1\n0(right)withsolidfittingcurvesfor CrSbSe 3.\nInset shows the log 10Mvs log 10Hat 70 K with linear fitting\ncurve. (b) Kouvel-Fisher plots of Ms(dMs/dT)−1(left) and\nχ−1\n0(dχ−1\n0/dT)−1(right) with solid fittingcurves for CrSbSe 3.\namount of heat which could be transferred between cold\nand hot parts of the refrigerator in an ideal thermody-\nnamic cycle. The calculated RCP values of CrSbSe 3\nreach maxima at 50 kOe of 163.1 J kg−1, 142.1 J kg−1,\nand 154.1 J kg−1fora,b, andcaxis, respectively [Fig.\n3(e)]. In addition, the field dependence of −∆Smax\nMand\nRCP can be well fitted by using the power-law relations\n−∆Smax\nM=aHnandRCP=bHm[Fig. 3(e)].30\nFor a second-order PM-FM phase transition, the\nspontaneous magnetization ( Ms) below Tc, the initial\nmagnetic susceptibility ( χ−1\n0) above Tc, and the field-\ndependent magnetization (M) at Tccan be characterized\nby a set of critical exponents β,γ, andδ, respectively.31\nThemathematicaldefinitionsofthe exponentsfrommag-\nnetization measurement are given below:\nMs(T) =M0(−ε)β,ε <0,T < T c, (7)\nχ−1\n0(T) = (h0/m0)εγ,ε >0,T > T c,(8)\nM=DH1/δ,T=Tc, (9)\nwhereε= (T−Tc)/Tc;M0,h0/m0andDare the critical\namplitudes.32For the original Arrott plot, β= 0.5 and5\n/s48/s46/s49 /s49 /s49/s48/s53/s49/s48/s49/s53/s50/s48\n/s48/s46/s49 /s49 /s49/s48/s53/s49/s48/s49/s53/s50/s48\n/s48/s46/s49 /s49 /s49/s48/s53/s49/s48/s49/s53/s50/s48\n/s84/s32/s62/s32/s84\n/s99\n/s32/s32/s77/s124 /s124/s45\n/s40/s101/s109/s117/s47/s103 /s41\n/s72/s124 /s124/s45/s40 /s41\n/s32/s40/s107/s79/s101/s41/s40/s97/s41/s32/s32/s72/s32/s47/s47/s32/s97\n/s32/s61/s32/s48/s46/s50/s54\n/s32/s61/s32/s49/s46/s51/s51/s84/s32/s60/s32/s84\n/s99\n/s84/s32/s62/s32/s84\n/s99/s84/s32/s60/s32/s84\n/s99\n/s32/s61/s32/s48/s46/s51/s50\n/s32/s61/s32/s49/s46/s48/s51/s40/s98/s41/s32/s32/s72/s32/s47/s47/s32/s98\n/s32/s32/s32\n/s72/s124 /s124/s45/s40 /s41\n/s32/s40/s107/s79/s101/s41/s84/s32/s62/s32/s84\n/s99/s84/s32/s60/s32/s84\n/s99\n/s32/s61/s32/s48/s46/s50/s53\n/s32/s61/s32/s49/s46/s49/s57/s40/s99/s41/s32/s32/s72/s32/s47/s47/s32/s99/s32/s32\n/s72/s124 /s124/s45/s40 /s41\n/s32/s40/s107/s79/s101/s41\nFIG. 6. (Color online). Scaling plots of renormalized m=\nM|ε|−βvsh=H|ε|−(β+γ)above and below Tcfor CrSbSe 3.\nγ= 1.0.33Based on this, the magnetization isotherms\nM2vsH/Mshould be a set of parallel straight lines.\nThe isotherm at the critical temperature Tcshould pass\nthrough the origin. As shown in Figs. 4(a-c), all curves\nin the Arott plot of CrSbSe 3are nonlinear, with a down-\nward curvature, indicating that the Landau mean-field\nmodel is not applicable to CrSbSe 3. From the Banerjee′s\ncriterion,34the first (second) order phase transition cor-\nresponds to a negative (positive) slope. Thus, the down-\nwardslope confirmsits a second-orderPM-FM transition\nin CrSbSe 3. In the more general case, the Arrott-Noaks\nequation of state provides a modified Arrott plot:35\n(H/M)1/γ=aε+bM1/β, (10)\nwhereε= (T−Tc)/Tcandaandbare fitting constants.\nFigures 4(d-f) present the modified Arrott plots for all\nthree axes with a self-consistent method,36,37showing a\nset of parallel quasi-straight lines at high field.\nTo obtain the anisotropic critical exponents β,γand\nδ, the linearly extrapolated Ms(T) andχ−1\n0(T) against\ntemperature are plotted in Fig. 5(a). According to Eqs.\n(7) and (8), the solid fitting lines give that β= 0.26(1)\nandγ= 1.32(2) for easy aaxis, close to the reported\nvalues (β= 0.25 and γ= 1.38).8For the baxis,\nβ= 0.28(2) and γ= 1.02(1), while for the caxis,\nβ= 0.26(1) and γ= 1.14(5). This lies between the\nvalues of theoretical tricritical mean field model ( β=\n0.25 and γ= 1.0) and 3D Ising model ( β= 0.325 and\nγ= 1.24).38,39The value of βis outside of the win-\ndow 0.1≤β≤0.25 for a 2D magnet,40suggesting a\n3D magnetic behavior for quasi-1D CrSbSe 3. Accord-\ning to Eq. (9), the M(H) atTcshould be a straight\nline in log-log scale with the slope of 1 /δ. Such fitting\nyieldsδ= 6.17(9), 4.14(16), and 5.35(17), for a,b, andc\naxis, respectively, which agrees well with the calculated\nvalues from βandγbased on the Widom relation δ=\n1+γ/β.41Theself-consistencycanalsobecheckedviathe\nKouvel-Fisher method,42whereMs(T)/(dMs(T)/dT)−1andχ−1\n0(T)/(dχ−1\n0(T)/dT)−1plotted against tempera-\nture should be straight lines with slopes 1 /βand 1/γ,\nrespectively. The linear fits to the plots [Fig. 5(b)] yield\nβ= 0.26(1) and γ= 1.33(2) for aaxis,β= 0.32(2) and\nγ= 1.03(1) for baxis,β= 0.25(1) and γ= 1.19(4) for c\naxis, respectively, very close to the values obtained from\nmodified Arrott plot. All the critical exponents obtained\nfrom different methods are listed in Table I. It seems that\nthe critical behavior of CrSbSe 3is much different along\nthe different axis and cannot be described by any sin-\ngle model. However, it is clear that 3D critical behavior\ndominates in quasi-1D CrSbSe 3and the strong magne-\ntocrystalline anisotropy in CrSbSe 3plays an important\nrole in the origin of anisotropic critical exponents.\nScaling analysis can be used to estimate the reliability\nof the obtained critical exponents. According to scaling\nhypothesis, the magnetic equation of state in the critical\nregion obeys a scaling relation can be expressed as:\nM(H,ε) =εβf±(H/εβ+γ), (11)\nwheref+forT > T candf−forT < T c, respec-\ntively, are the regular functions. In terms of the variable\nm≡ε−βM(H,ε) andh≡ε−(β+γ)H, renormalized mag-\nnetization and renormalized field, respectively, Eq.(10)\nreduces to a simple form:\nm=f±(h). (12)\nIt implies that for a true scaling relation with the proper\nselection of β,γ, andδ, the renormalized mversush\ndata will fall onto two different universal curves; f+(h)\nfor temperature above Tcandf−(h) for temperature be-\nlowTc. Using the values of βandγobtained from the\nKouvel-Fisher plot, we have constructed the renormal-\nizedmvshplots in Fig. 6. It is clear seen that all the\nexperimental data collapse onto two different branches:\none above Tcand another below Tc, confirming proper\ntreatment of the critical regime.\nIV. CONCLUSIONS\nIn summary, we have studied in details the anisotropic\nmagnetocaloric effect and critical behavior of CrSbSe 3\nsingle crystal. The second-orderin nature of the PM-FM\ntransition near Tc= 70 K has been verified by the scal-\ning analysis of magnetic entropy change ∆ SM. A large\nmagnetocrystalline anisotropy constant Kuis estimated\nto be 148.5 kJ m−3at 10 K, comparable with that of\nCr(Br,I) 3. A set of critical exponents β,γ, andδalong\neach axis estimated from various techniques match rea-\nsonably well and follow the scaling equation, indicating a\n3D magnetic behavior in CrSbSe 3. Further neutron scat-\ntering and theoretical studies are needed to shed more\nlight on the anisotropic magnetic coupling in low dimen-\nsions.6\nACKNOWLEDGEMENTS\nThis work was supported by the US DOE-BES, Divi-\nsion of Materials Science and Engineering, under Con-\ntract No. de-sc0012704 (BNL).\n1B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein,\nR. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A.\nMcGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo-\nHerrero, and X. D. Xu, Nature 546, 270 (2017).\n2M. A. McGuire, G. Clark, KC. S., W. M. Chance, G. E.\nJellison, Jr., V. R. Cooper, X. Xu, and B. C. Sales, Phys.\nRev. M1014001 (2017).\n3C. Gong, L. Li, Z. L. Li, H. W. Ji, A. Stern, Y. Xia, T.\nCao, W. Bao, C. Z. Wang, Y. Wang, Z. Q. Qiu, R. J. Cava,\nS. G. Louie, J. Xia, and X. Zhang, Nature 546, 265 (2017).\n4M. A. McGuire, H. Dixit, V. R. Cooper, and B. C. Sales,\nChem. Mater. 27, 612 (2015).\n5N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133\n(1966).\n6V. Volkov, G. V. Tendeloo, J. V. Landuyt, S. Amelinckx,\nE. Busheva, G. Shabunina, T. Aminov, and V. Novotort-\nsev, J. Solid State Chem. 132, 257 (1997).\n7D. A. Odink, V. Carteaux, C. Payen, and G. Ouvrard,\nChem. Mater. 5, 237 (1993).\n8T. Kong, K. Stolze, D. Ni, S. K. Kushwaha, and R. J.\nCava, Phys. Rev. Mater. 2, 014410 (2018).\n9Y. Sun, Z. Song, Q. Tang, and X. Luo, J. Phys. Chem. C\n10.1021/acs.jpcc.0c02101 (2020).\n10Y. Liu and C. Petrovic, Phys. Rev. B 97, 014420 (2018).\n11G. T. Lin, X. Luo, F. C. Chen, J. Yan, J. J. Gao, Y. Sun,\nW. Tong, P. Tong, W. J. Lu, Z. G. Sheng, W. H. Song,\nX. B. Zhu, and Y. P. Sun, Appl. Phys. Lett. 112, 072405\n(2018).\n12Y. Liu and C. Petrovic, Phys. Rev. B 97, 174418 (2018).\n13X. Yu, X. Zhang, Q. Shi, S. Tian, H. Lei, K. Xu, and H.\nHosono, Front. Phys. 14, 43501 (2019).\n14W. Liu, Y. Dai, Yu E. Yang, J. Fan, L. Pi, L. Zhang, and\nY. Zhang, Phys. Rev. B 98, 214420 (2018).\n15J. Yan, X. Luo, F. C. Chen, J. J. Gao, Z. Z. Jiang, G. C.\nZhao, Y. Sun, H. Y. Lv, S. J. Tian, Q. W. Yin, H. C. Lei,\nW. J. Lu, P. Tong, W. H. Song, X. B. Zhu, and Y. P. Sun,\nPhys. Rev. B 100, 094402 (2019).\n16Y. K. Fu, Y. Sun, and X. Luo, J. Appl. Phys. 125, 053901\n(2019).\n17Y. Sun and X. Luo, Phys. Status Solidi B 1900052 (2019).\n18Y. Liu and C. Petrovic, Phys. Rev. Mater. 3, 014001\n(2019).\n19Y. Liu, J. Li, J. Tao, Y. Zhu, and C. Petrovic, Sci. Rep. 9,\n13233 (2019).20W. Liu, Y. Wang, J. Fan, L. Pi, M. Ge, L. Zhang, and Y.\nZhang, Phys. Rev. B 100, 104403 (2019).\n21N. Richter, D. Weber, F. Martin, N. Singh, U. Schwingen-\nschl¨ ogl, B. V. Lotsch, and M. Kl¨ aui, Phys. Rev. Mater. 2,\n024004 (2018).\n22Y. Liu, L. Wu, X. Tong, J. Li, J. Tao, Y. Zhu, and C.\nPetrovic, Sci. Rep. 9, 13599 (2019).\n23L. D. Casto, A. J. Clune, M. O. Yokosuk, J. L. Musfeldt,\nT. J. Williams, H. L. Zhuang, M. W. Lin, K. Xiao, R. G.\nHennig, B. C. Sales, J. Q. Yan, and D. Mandrus, APL\nMater.3, 041515 (2015).\n24S. Selter, G. Bastien, A. U. B. Wolter, S. Aswartham, and\nB. B¨ uchner, Phys. Rev. B 101, 014440 (2020).\n25V. Pecharsky, K. Gscheidner, J. Magn. Magn. Mater. 200,\n44 (1999).\n26J. Amaral, M. Reis, V. Amaral, T. Mendonc, J. Araujo,\nM. Sa, P. Tavares, J. Vieira, J. Magn. Magn. Mater. 290,\n686 (2005).\n27V. Franco and A. Conde, Int. J. Refrig. 33, 465 (2010).\n28E. C. Stoner and E. P. Wohlfarth, Philos. Trans. R. Soc.\nLond. Ser. Math. Phys. Sci. 240, 599 (1948).\n29K. A. Gschneidner Jr., V. K. Pecharsky, A. O. Pecharsky,\nand C. B. Zimm, Mater. Sci. Forum 315, 69 (1999).\n30V. Franco, J. S. Blazquez, and A. Conde, Appl. Phys. Lett.\n89, 222512 (2006).\n31H. E. Stanley, Introduction to Phase Transitions and Crit-\nical Phenomena (Oxford U. P., London and New York,\n1971).\n32M. E. Fisher, Rep. Prog. Phys. 30, 615 (1967).\n33A. Arrott, Phys. Rev. B 108, 1394 (1957).\n34S. K. Banerjee, Phys. Lett. 12, 16 (1964).\n35A. Arrott, and J. Noakes, Phys. Rev. Lett. 19, 786 (1967).\n36W. Kellner, M. F¨ ahnle, H. Kronm¨ uller, and S. N. Kaul,\nPhys. Status Solidi B 144, 387 (1987).\n37A.K. Pramanik, andA. Banerjee, Phys. Rev.B 79, 214426\n(2009).\n38K. Huang, Statistical Machanics, 2nd ed. (Wiley, New\nYork, 1987).\n39J. C. LeGuillou, and J. Zinn-Justin, Phys. Rev. B 21, 3976\n(1980).\n40A. Taroni, S. T. Bramwell, and P. C. W. Holdsworth, J.\nPhys.: Condens. Matter 20, 275233 (2008).\n41B. Widom, J. Chem. Phys. 41, 1633 (1964).\n42J. S. Kouvel, and M. E. Fisher, Phys. Rev. 136, A1626\n(1964)." }, { "title": "2008.05075v1.Metastable_skyrmion_lattices_governed_by_magnetic_disorder_and_anisotropy_in__β__Mn_type_chiral_magnets.pdf", "content": "arXiv:2008.05075v1 [cond-mat.str-el] 12 Aug 2020Metastable skyrmion lattices governed by magnetic disorde r and\nanisotropy in β-Mn-type chiral magnets\nK. Karube,1,∗J. S. White,2,∗V. Ukleev,2C. D. Dewhurst,3R. Cubitt,3\nA. Kikkawa,1Y. Tokunaga,4H. M. Rønnow,5Y. Tokura,1,6and Y. Taguchi1\n1RIKEN Center for Emergent Matter Science (CEMS), Wako 351-0198 , Japan.\n2Laboratory for Neutron Scattering and Imaging (LNS),\nPaul Scherrer Institute (PSI), CH-5232 Villigen, Switzerland .\n3Institut Laue-Langevin (ILL), 71 avenue des Martyrs,\nCS 20156, 38042 Grenoble cedex 9, France\n4Department of Advanced Materials Science,\nUniversity of Tokyo, Kashiwa 277-8561, Japan.\n5Laboratory for Quantum Magnetism (LQM), Institute of Physics ,\n´Ecole Polytechnique F´ ed´ erale de Lausanne (EPFL), CH-1015 L ausanne, Switzerland.\n6Tokyo College and Department of Applied Physics,\nUniversity of Tokyo, Bunkyo-ku 113-8656, Japan.\nMagnetic skyrmions are vortex-like topological spin textu res often observed in\nstructurally chiral magnets with Dzyaloshinskii-Moriya i nteraction. Among them,\nCo-Zn-Mn alloys with a β-Mn-type chiral structure host skyrmions above room tem-\nperature. In this system, it has recently been found that sky rmions persist over a\nwide temperature and magnetic field region as a long-lived me tastable state, and\nthat the skyrmion lattice transforms from a triangular latt ice to a square one. To\nobtain perspective on chiral magnetism in Co-Zn-Mn alloys a nd clarify how various\nproperties related to the skyrmion vary with the compositio n, we performed sys-\ntematic studies on Co 10Zn10, Co9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6in terms of\nmagnetic susceptibility and small-angle neutron scatteri ng measurements. The ro-\nbust metastable skyrmions with extremely long lifetime are commonly observed in\nall the compounds. On the other hand, preferred orientation of a helimagnetic prop-\nagation vector and its temperature dependence dramaticall y change upon varying\nthe Mn concentration. The robustness of the metastable skyr mions in these materi-2\nals is attributed to topological nature of the skyrmions as a ffected by structural and\nmagnetic disorder. Magnetocrystalline anisotropy as well as magnetic disorder due\nto the frustrated Mn spins play crucial roles in giving rise t o the observed change in\nhelical states and corresponding skyrmion lattice form.\nI. INTRODUCTION\nNon-collinear and non-coplanar spin textures have recently attra cted much attention as\na source of various emergent electromagnetic phenomena. Magne tic skyrmions, vortex-like\nspin textures characterized by an integer topological charge, ar e a prototypical example of\nsuch non-coplanar magnetic structures[1–4], and are anticipated to be applied to spintronics\ndevices since they can be treated as particles and driven by an ultra -low current density[5–\n7]. Thus far, skyrmions have been observed in various magnets due to several microscopic\nmechanisms, such as competition between the Dzyaloshinskii-Moriya interaction (DMI) and\nferromagneticexchangeinteraction[3,4,8–19], magneticdipoleint eraction[20–23], andmag-\nnetic frustration or Ruderman-Kittel-Kasuya-Yosida (RKKY) inte raction[24–26]. Among\nthem, the DMI arises from relativistic spin-orbit interaction and bro ken inversion symmetry\neither at interfaces of thin-film layers[8–12] or in bulk materials with n oncetrosymmetric\ncrystal structures[3, 4, 13–19].\nIn the structurally chiral magnets as represented by B20-type c ompounds (e.g., MnSi[3],\nFe1−xCoxSi[4], FeGe[13]) and Cu 2OSeO3[14] with the space group of P213, the DMI grad-\nually twists ferromagnetically coupled moments to form a long-period helimagnetic state\ndescribed by a magnetic propagation vector ( qvector). The magnitude of qis given by\nq∝D/J, whereDandJcorrespond to the DMI constant and the exchange stiffness, re-\nspectively, and the propagation direction is determined by magnetic anisotropy. Near the\nhelimagnetic transition temperature Tc, magnetic fields induce a triangular-lattice skyrmion\ncrystal (SkX) as illustrated in Fig. 1(c), which is often described as a triple-qstructure\nwith the qvectors displaying mutual 120◦angles perpendicular to the magnetic field. In\ngeneral, SkX is stabilized by thermal fluctuations and thus its therm odynamical equilib-\n∗These authors equally contributed to this work3\nrium state is confined to a narrow temperature and magnetic field re gion just below Tc, and\ntopologically-trivial helical or conical states are the thermodynam ically most stable states\nat lower temperatures.\nRecently, Co-Zn-Mn alloys have been identified as a new class of chira l magnets based on\nbulk DMI which host skyrmions above room temperature[15]. The mat erials crystallize in\naβ-Mn-type chiral cubic structure with the space group of P4132 (defined as right-handed\nstructure) or P4332 (left-handed structure), where 20 atoms per unit cell are dist ributed\nover two Wyckoff sites (8 cand 12d) as illustrated in Fig. 1(a). The 8 csites are mainly\noccupied by Co atoms while the 12 dsites are mainly occupied by Zn and Mn[27–30]. The\nβ-Mn-type structure forms in all the solid solutions of (Co 0.5Zn0.5)20−xMnxfrom Co 10Zn10\nto Mn 20(β-Mn itself)[27, 31].\nMagnetic phase diagram on the temperature ( T) - Mn concentration ( x) plane is repro-\nduced from Ref. [31] and displayed in Fig. 1(b) with additional informa tion obtained in\nthe present study. One end member Co 10Zn10shows a helimagnetic ground state with Ref.\n[15] reporting a magnetic periodicity λ∼185 nm below Tc∼460 K.Tcrapidly decreases as\npartial substitution of Mn proceeds, and Co 8Zn8Mn4withTc∼300 K exhibits a thermally\nequilibrium SkX state at room temperature under magnetic fields. Th e magnitude of the\nDMI constant in Co 8Zn8Mn4has been experimentally evaluated to be D∼0.53 mJ/m2,\nwhich is several times smaller than that in FeGe[32]. The DMI critically de pends on band\nstructure and electron band filling as demonstrated in Fe-doped Co 8Zn8Mn4, where even a\nsign change in the DMI, namely, reversal of skyrmion helicity, occur s as the Fe concentration\nis increased[33].\nAlthough the thermodynamical equilibrium SkX phase in Co 8Zn8Mn4exists only in a\nnarrow temperature and magnetic field region, it has been demonst rated that a once-created\nSkX can persist over the whole temperature region below room temp erature and a wide\nmagnetic field region as a long-lived metastable state via a convention al (a few K/min) field\ncooling (FC)[34]. Moreover, the lattice form of the metastable SkX u ndergoes a reversible\ntransition, accompanied by large increase in q, from a conventional triangular lattice to\na novel square one, which is described as an orthogonal double- qstate, as illustrated in\nFig. 1(e)[35]. Similar robust metastable SkX has been observed in Co 9Zn9Mn2withTc∼\n400 K, where the metastable SkX persists even at zero field above r oom temperature[36].\nBy means of Lorentz transmission electron microscopy (LTEM) for thin-plate specimens,4\nvarious exotic skyrmion-related structures, such as I- or L-sha ped elongated skyrmions[37]\n[Fig. 1(g)], a smectic liquid-crystalline structure of skyrmions[38] an d a meron-antimeron\nsquare lattice[39], have been observed.\nThe other end member β-Mn is well known as a spin liquid, while the lightly doped β-Mn\nalloys with slight disorder exhibit a spin glass, due to geometrical frus tration among anti-\nferromagnetically coupled Mn spins in the hyper-kagome network of the 12dsites[40–45].\nTherefore, (Co 0.5Zn0.5)20−xMnxpossesses both magnetic frustrations inherent to β-Mn and\nmagnetic disorder due to the mixture of ferromagnetic Co spins and antiferromagnetic Mn\nspins, which give rise to a spin glass phase below Tgover a wide range of the Mn concen-\ntration (3 ≤x≤19)[31]. In particular, the spin glass phase invades the helical phase for 3\n≤x≤7 displaying reentrant spin glass behavior[46] as similarly reported fo r a number of\nferromagnets[47–52], antiferromagnets[53, 54] and helimagnets [55] with chemical and mag-\nnetic disorder. The spin glass nature was confirmed by previous fre quency-dependent ac\nsusceptibility measurements for Co 7Zn7Mn6(Tc∼160 K,Tg∼30 K) [29, 31, 56]. It has\nbeen discovered that a novel equilibrium phase of disordered skyrm ions exists just above Tg\nin Co7Zn7Mn6, which is thermodynamically disconnected from the conventional eq uilibrium\nSkX phase just below Tc, and presumably stabilized by a cooperative interplay between the\nchiral magnetism with DMI and the frustrated magnetism[31].\nDespite these extensive studies for Co-Zn-Mn alloys, it remains elus ive how the\n(meta)stability and lattice form of skyrmions vary with the Mn conce ntration from the\nend member Co 10Zn10, and how the two magnetic elements of Co and Mn contribute to\nthe skyrmion formation and the transformation of skyrmion lattice . In order to obtain a\nmore complete perspective on the skyrmion states in the Co-Zn-Mn alloys, here we present\nnew data sets obtained by small-angle neutron scattering (SANS) a nd magnetic suscepti-\nbility that have not been reported previously. From the new data we report: (i) Identifi-\ncation of the equilibrium skyrmion phase, and temperature and field d ependence of helical\nand metastable skyrmion states in Co 10Zn10, (ii) an off-axis magnetic field experiment for\nmetastable skyrmions in Co 8Zn8Mn4to reveal the role of anisotropy, (iii) temperature and\nfield dependence of metastable skyrmions in Co 9Zn9Mn2below ambient temperature down\nto the lowest temperature to clarify the effect of dilute Mn moments that becomes signif-\nicant only at low temperatures, (iv) temperature and field evolution of heavily disordered\nmetastable skyrmions in Co 7Zn7Mn6, and (v) lifetime of metastable skyrmions in Co 10Zn105\nandCo 8Zn8Mn4. Takingthenewresultstogetherwithourpreviousones, weshows ystematic\nchanges in lattice forms and lifetimes of the metastable skyrmions, a s well as the key roles\nof magnetic disorder and anisotropy, as functions of temperatur e and Mn concentration. As\nsummarized in Fig. 2, the metastable skyrmion state prevails in a wide t emperature and\nfield range for all the materials. In Co 10Zn10, triangular lattice of skyrmions transforms to\nrhombic-like [Fig. 1(d)] asthe temperature is lowered in a low field. This triangular-rhombic\nstructural transformationisgoverned byenhanced magnetocr ystalline anisotropythat favors\nq∝bardbl<111>. On the other hand, in the Mn-doped compounds, q-vector orientation changes\nfrom<111>to<100>, and structural transformation from triangular lattice to squar e one\noccurs at low temperatures and low fields. As the Mn concentration is increased, and as the\nhelimagnetic Tcthus falls, the antiferromagnetic correlations of Mn spins start to develop on\ncooling from higher temperatures. The resulting magnetic disorder drives a large increase\ninqvalue, thereby triggering the transformation to the square lattic e state, in cooperation\nwith the enhanced magnetic anisotropy toward q∝bardbl<100>at low temperatures.\nThe format of this paper is as follows. After we describe experiment al methods in Sec-\ntion II, results and discussion are presented in Section III, and co nclusion is given in Section\nIV. In the section III, we first overview helical and skyrmion state s in all the compounds\n(Subsections A, Figures 2, 3, Table 1). Then, we show detailed resu lts of magnetic suscepti-\nbility and SANS measurements for each composition: Co 10Zn10(Subsection B, Figures 4-6),\nCo8Zn8Mn4(Subsection C, Figure7), Co 9Zn9Mn2(Subsection D, Figure8), and Co 7Zn7Mn6\n(Subsection E, Figure 9-11), and summarize all of them (Subsectio n F). Finally, we present\ncomposition dependence of lifetime of metastable skyrmions (Subse ction G, Figures 12, 13).\nWith all these results, we discuss the roles of magnetic disorder and anisotropy in leading\nto the robust metastable skyrmions and their novel lattice forms in Co-Zn-Mn alloys.\nII. EXPERIMENTAL METHODS\nA. Sample preparation\nSingle-crystalline Co 10Zn10was grown by a self-flux method in an evacuated quartz tube.\nThesinglecrystalswerecutalongthe(110),(-110)and(001)plan eswitharectangularshape\nfor magnetization and ac susceptibility measurements as well as SAN S measurement after6\nthe crystalline orientation was determined by the X-ray Laue diffrac tion method. Single\ncrystals of Co 9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6were grown by the Bridgman method\nas described in our previous papers[31, 34, 36], and were cut along t he (110), (-110) and\n(001) planes for Co 9Zn9Mn2, and along the (100), (010) and (001) planes for Co 8Zn8Mn4\nand Co 7Zn7Mn6, respectively.\nB. Magnetization and ac susceptibility measurements\nMagnetization and ac susceptibility measurements were performed with a supercon-\nducting quantum interference device magnetometer (MPMS3, Qua ntum Design). High-\ntemperature measurements above 400 K for Co 10Zn10were performed by using an oven\noption. In the ac susceptibility measurements, the ac drive field was set asHac= 1 Oe\nfor all the compounds, and the ac frequency was selected to be f= 17 Hz for Co 10Zn10\nandf=193 Hz for Co 9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6. Magnetic fields were applied\nalong the [110] direction for Co 10Zn10and Co 9Zn9Mn2, and along the [100] direction for\nCo8Zn8Mn4and Co 7Zn7Mn6, respectively. Due to the difference in the shape between the\nsamples used intheacsusceptibility (field parallel totheplate) andth eSANSmeasurements\n(field perpendicular to the plate), their demagnetization factors a re different. To correct for\nthis difference so that a common absolute magnetic field scale is used t hroughout this paper,\nthe field values for the ac susceptibility measurements are calibrate d asHc=NHwhereN\n= 3.0, 2.0, 3.7 and 2.7 for Co 10Zn10, Co9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6, respectively.\nC. Small-angle neutron scattering (SANS) measurements\nSANSmeasurements forCo 10Zn10, Co9Zn9Mn2andCo 8Zn8Mn4were performedusing the\nSANS-I instrument at the Paul Scherrer Institute (PSI), Switze rland. For high-temperature\nmeasurements above 300 K for Co 10Zn10and Co 9Zn9Mn2, a bespoke oven stick designed\nfor SANS experiments was used. SANS measurements of Co 7Zn7Mn6were done using the\nD33 instrument at the Institut Laue-Langevin (ILL), France. Th e neutron wavelength was\nselected to be 10 ˚A with a 10% full width at half maximum (FWHM) spread in all the\nmeasurements. For all the SANS data shown here, nuclear and inst rumental background\nsignals have been subtracted by using data taken either well above the helimagnetic ordering7\ntemperature Tc, or well above the polarized ferromagnetic transition field.\nFor Co 10Zn10, the mounted single-crystalline sample was installed into a horizontal field\ncryomagnet so that the incident neutron beam ( ki) and the magnetic field ( H) were parallel\nto the [110] direction. The cryomagnet was rotated (‘rocked’) tog ether with the sample\naround the vertical [001] direction, and the rocking angle ( ω) between kiandHwas scanned\nfrom−20◦to 20◦by 2◦step. Here, ω= 0◦corresponds to the ki∝bardblHconfiguration. The\nobserved FWHM of the rocking curves (scattering intensity versu sω) was always broader\nthan 19◦. Therefore, to deduce accurately the relative intensities and pos itions of all of the\nBragg spots in single images, all the SANS images displayed in this paper are obtained by\nsumming multiple SANS measurements taken over −8◦≤ω≤8◦.\nAs detailed in our previous papers[31, 34, 36], similar SANS measureme nts were per-\nformedwith the ki∝bardblH∝bardbl[110] configurationforCo 9Zn9Mn2, andki∝bardblH∝bardbl[001]configuration\nfor Co 8Zn8Mn4and Co 7Zn7Mn6.\nIII. RESULTS AND DISCUSSION\nA. Overview of state diagrams\nFirst, we briefly overview the results of basic magnetic properties in helimagnetic states\nand metastable skyrmion states for all the compounds.\n1. Helical state\nFigure 3(a-d) shows temperature ( T) dependence of magnetization ( M) under a small\nmagnetic field of 20 Oe. In Co 10Zn10,Mshows a sharp increase due to a helimagnetic\ntransition at Tc∼414 K, and then stays almost independent of temperature both fo r the\nfield cooling (FC) and the zero-field-cooled field-warming (ZFC-FW) p rocesses. The other\ncompounds exhibit gradual decrease in Mupon cooling at some temperature region ( TL≤T\n≤TH). Co9Zn9Mn2shows almost temperature-independent Min a wide temperature range\nbelowTc∼396 K, while the gradual decrease is observed below TH∼50 K. In Co 8Zn8Mn4,\nT-independent behavior below Tc∼299 K is followed by the gradual decrease in Mupon\ncooling from TH∼120 K down to TL∼40 K. A sharp drop of MatTg∼9 K observed\nonly in the ZFC-FW process is due to a reentrant spin-glass transitio n. Co7Zn7Mn6exhibits8\nthe gradual decrease in MbelowTH∼120 K down to TL∼50 K below the helimagnetic\ntransition at Tc∼158 K. The reentrant spin glass transition is observed around Tg∼26\nK as manifested in a large deviation between FC and ZFC-FW processe s. The reentrant\nspin glass transition temperature Tgdetermined by the M(T) curve is very close to the\nzero frequency limit of Tgdetermined by ac susceptibility curves with various frequencies\nas previously reported[56]. Tgdoes not show significant dependence on the magnetic field\nbelow the field-induced ferromagnetic region as plotted in the phase diagrams in Fig. 2(d,\ne). These characteristic temperatures ( Tc,TH,TLandTg) are plotted in the T-xphase\ndiagram in Fig. 1(b).\nMagnetization curves at 2 K that characterize the ground states are shown in Supple-\nmentary Fig. S1[57]. The values of Tcand saturation magnetization Msat 2 K and at 7 T\nin all the compounds are summarized in Table 1. From the value of Ms∼11.3µB/f.u. in\nCo10Zn10, magnetic moment at Co site is estimated to be ∼1.1µB. WhileTcmonotonically\ndecreases with increasing Mn concentration, Msdisplays a maximum in Co 9Zn9Mn2.\nThe characteristic T-dependence of Mat 20 Oe is compared with that of the magnitude\nof helical wavevector ( q) from SANS measurements in Fig. 3(e-h). It turns out that the\ngradual decrease in MfromTHtoTLcorresponds to the large increase in the value of q,\ni.e. the decrease in the helimagnetic periodicity λ(= 2π/q). The largest value of λat\nhigh temperatures ( λmax) and the smallest value of λat low temperatures ( λmin) for each\ncompound are summarized in Table 1. We also present the preferred orientation of qin\nTable 1. While the preferred qdirection is <111>for Co 10Zn10, it is parallel to <100>in\nCo9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6, as detailed in the following sections.\n2. Metastable skyrmion state\nFigure 2 shows the T-Hdiagrams determined by ac susceptibility and SANS mea-\nsurements for each composition. In these diagrams, the equilibrium SkX phase and the\nmetastable SkX state are presented together, as determined by field scans after ZFC and by\nfield scans after a FC via the equilibrium phase, respectively, as sche matically illustrated in\nFig. 2(a). The robust metastable skyrmion state that exists over a very wide temperature\nand field region is commonly observed from Co 10Zn10to Co7Zn7Mn6. In Co 10Zn10the trian-\ngular lattice of metastable skyrmions (M-T-SkX) distorts and tran sforms to a rhombic one9\n(M-R-SkX) at low fields during the FC. On the other hand, the lattice form changes to a\nsquareone(M-S-SkX)atlowtemperatures andlowfields intheothe r Mn-dopedcompounds.\nAs the partial Mn substitution proceeds, the phase space of the m etastable triangular SkX\nstate (M-T-SkX) is squeezed, accompanied by the suppression of Tc, while the region occu-\npied by square SkX state (M-S-SkX) expands toward higher tempe ratures. The transition\nto the square SkX is accompanied by a large increase in qas similarly observed in the helical\nstates [Fig. 3(f-h)].\nB. Co 10Zn10\nInthissection, wepresent detailedresultsofSANSandacsuscept ibility measurements for\nCo10Zn10and show that thepreferred orientation of qvector is <111>direction, and thaton\nfield cooling metastable skyrmion lattice distorts and undergoes a st ructural transformation\nfrom a conventional triangular lattice to a rhombic one.\n1. Identification of equilibrium SkX\nFirst, we identify an equilibrium SkX phase (Fig. 4). Photographs of s ingle-crystalline\nsamples used for SANS and ac susceptibility measurements are show n in Fig. 4(a). Figure\n4(b) shows the field variation of SANS images at 410 K. The scatterin g signal observed at\n0 T is attributed to a helical multi-domain state. Here, 4 broad spots are observed but\nthe peak positions are not aligned clearly to any unique set of high-sy mmetry crystal axes.\nThis peculiar scattering distribution near Tcmay be attributed to rather flexible nature of q\nvector due to the negligible anisotropy at high temperatures under an influence of possible\nresidual local strain of the sample. Under magnetic fields parallel to the incident neutron\nbeam, the SANS signal diminishes at 0.02 T most likely due to the transit ion to a conical\nstate whose qvector is parallel to the field and thus not detected in this configura tion. At\n0.03 T, a broad 6-spot pattern, a hallmark of triangular lattice of sk yrmions (triple- q⊥H),\nappears. As the magnetic field is further increased to 0.06 T, the co nical state is stabilized\nagain and the signal disappears. SANS intensity perpendicular to th e field is plotted against\nfield in Fig. 4(c). The SANS intensity exhibits a peak at 0.03 T, where th e volume fraction\nof skyrmions is maximized.10\nThe field dependence of the real- ( χ′) and imaginary-part ( χ′′) of the ac susceptibility at\n410 K is shown in Fig. 4(d). A dip structure is observed in χ′between 0.02 T and 0.05 T\nwhere in addition χ′′exhibits a clear double peak structure. These features correspo nd to a\nSkX state surrounded by a conical state, and show a good agreem ent with the region of the\nenhanced SANS intensity. Therefore, the phase boundaries of Sk X state are determined as\nthe peak positions at the both sides of the dip structure in χ′. Contour plots of χ′andχ′′\nwith the phase boundaries on the T-Hplane are presented in Fig. 4(e) and (f), respectively.\nThe equilibrium SkX phase exists in a narrow temperature region from Tc∼412 K down\nto 407 K, below which the field-induced conical state is the most stab le.\n2. Temperature evolution of helical state in zero field\nNext, we discuss temperature variation of the helical state (Fig. 5 ). Figure 5(b) shows\nthe SANS patterns at selected temperatures on zero-field cooling from 410 K to 1.5 K. As\ntemperature is lowered from Tcdown to 300 K, the SANS intensity gradually accumulates\nalong the [1-11] direction due to enhanced magnetic anisotropy. Up on further decreasing\ntemperature down to 1.5 K, the SANS intensity splits into 4 spots: 2 s pots with stronger\nintensity along the [1-11] direction, and the other 2 spots with weak er intensity along the\n[-111] direction. This 4 spot pattern at 1.5 K is expected for the mult i-domain helical state\nwithq∝bardbl<111>as schematically illustrated in Fig. 5(a). Notably, the preferred orie ntation\nofqvector in Co 10Zn10is different from those of the other (Co 0.5Zn0.5)20−xMnxwithx≥2,\nwhich is found to be <100>by our previous studies[31, 34, 36].\nFigure 5(c) shows the temperature dependence of the SANS inten sity integrated for the [-\n111]and[1-11]directions( φ=55◦,125◦, 235◦and305◦), whereconventionalorder-parameter\nlike evolution of the intensity for q∝bardbl<111>is clearly observed. The radial qdependence of\nthe SANS intensity for the direction close to [1-11]is displayed in Fig. 5 (d). The peak center\nslightly shifts toward higher qregion and the peak width slightly increases on cooling. The\nhelicalqvalue, which is determined as the peak center of a Gaussian function fitted to the\nintensity vs qcurve [solid line in Fig. 5(d)], and its FWHM are plotted against temperat ure\nin Fig. 3(e) and (i), respectively. Both the qvalue and the FWHM gradually increases\non cooling but the variation is much smaller than that in the Mn-doped c ompounds. The\nhelical periodicity estimated as λ= 2π/qvaries a little from 156 nm to 143 nm on cooling11\nas summarized in Table 1. Note that the values of Tc= 414 K and λ= 156 nm for a single\ncrystal of Co 10Zn10in the present study are smaller than the values ( Tc= 462 K and λ\n= 185 nm) for a polycrystalline sample in the previous study[15], proba bly due to a slight\ndifference in the composition.\nAsdetailedinSupplementary Fig. S4[57], wealsofindthatthehelicalst ateat1.5Kforms\na chiral soliton lattice[59] under magnetic fields due to the enhanced magnetocrystalline\nanisotropy which enforces the qvector along <111>, and perpendicular to the field.\n3. Temperature dependence of metastable SkX\nNext, we discuss temperature variation of the metastable SkX sta te (Fig. 6). The SANS\npatternsinaFCat0.03Tfrom412Kto1.5Karedisplayed inFig. 6(b). I nordertogainthe\nsufficient cooling rate across the boundary between the equilibrium S kX and conical phases,\nthe magnetic field was applied at 412 K (slightly higher than presented in Fig. 4), where\nthe observed SANS pattern is ring-like, indicating that the triangula r lattice of skyrmions\nis orientationally disordered. As the temperature is lowered, the rin g-like SANS pattern\ngradually changes to broad 4 spots along the [-111] and [1-11] direc tions. The scattering\nintensity is plotted as a function of azimuthal angle φin Fig. 6(c). Clearly, 4 peaks around\nφ= 55◦, 125◦, 235◦and 305◦emerge out of a rather featureless profile at 412 K as the\ntemperature is lowered down to 1.5 K[58].\nThe 4-spot SANS pattern observed at low temperatures is expect ed for a rhombic lattice\nof skyrmions with double- qvectors∝bardbl<111>that are rotated by 110◦from each other\nas schematically illustrated in Fig. 6(a) and Fig. 1(d). The SANS intens ity integrated\nover the region close to the [-111] and [1-11] directions under the F C is plotted against\ntemperature together with the intensity integrated around the [0 01] and [-110] directions in\nFig. 6(d). Below 360 K the intensity around the [-111] and [1-11] dire ctions becomes higher\nthan that around the [001] and [-110] directions. This temperatur e corresponds to the onset\nof triangular-rhombic lattice structural transition and is plotted w ith a purple diamond\n(right one) in the state diagram in Fig. 2(b). The broad SANS patter n and the gradual\ntemperature evolution indicate coexistence of the triangular SkX s tate and the rhombic one\nover a wide temperature region.\nThe transformation to the rhombic lattice is probably driven by signifi cant enhancement12\nof magnetocrystalline anisotropy at low temperatures that favor sq∝bardbl<111>as found in the\nhelical state (Fig. 5). It is also noted that the absolute value of qslightly increases upon\nloweringtemperatureasplottedinFig. 3(e). Therelationbetweent helatticetransformation\nand the change in the qvalue is further discussed in the summary section of metastable SkX\n(Subsection F-2).\n4. Metastability of skyrmions against field variation\nWe also investigated the metastability and the lattice form of skyrmio ns against field\nvariation. The detailed field-dependent SANS data at 300 K after th e FC (0.03 T) is\npresented in Supplementary Fig. S2[57]. In the field sweepings towar d positive direction,\nthe broad 4-spot pattern with stronger intensity at the [-111] an d [1-11] directions changes\nto a uniform ring above 0.1 T as observed in the equilibrium SkX phase at 412 K and 0.03\nT. This field variation corresponds to the change in skyrmion lattice f orm from the rhombic\none to the original disordered triangular one. Importantly, such a ring-like pattern is never\nobserved upon the field sweeping at 300 K after ZFC. This result ens ures that the broad\n4-spot pattern appearing after the FC is attributed to the metas table rhombic SkX (M-R-\nSkX) and distinct from helical multi-domain state. The SANS intensity originating from the\ndisordered triangular SkX persists up to the region close to the field -induced ferromagnetic\nphase in contrast to the helical state. In the field sweeping toward negative direction, on\nthe other hand, the scattering intensity around the [-111] and [1- 11] directions is further\nincreased and the 4-spot pattern becomes clearer. This indicates that the rhombic lattice is\nmore stable than the triangular one at zero or negative fields. As th e field is swept further\nnegative, finally the SANS intensity from the rhombic SkX disappears upon the transition\nto the equilibrium conical state.\nThe field variation at 1.5 K after the FC (0.03 T) is displayed in Supplemen tary Fig.\nS3[57], where similar change in the SANS pattern was observed while th e 4 broad spots\npersist over a wider field region than that observed at 300 K.\nThefieldvariationsofSANSat300Kand1.5Kareconsistent withthos eofacsusceptibil-\nity, where smaller values (as compared with conical state) corresp onding to the metastable\nSkX are observed, accompanied by a characteristic asymmetric hy steresis. Similar field-\ndependent ac susceptibility measurements after FC processes we re performed at different13\ntemperatures, and boundaries of the metastable SkX state are p lotted in the T-Hphase\ndiagram in Fig. 2(b). The metastable SkX state persists over a wide T-Hregion, including\nroom temperature and zero field. Within the metastable state, the lattice form of skyrmions\nundergoes the transition from triangular to rhombic at a low- Tand low-Hregion where q\nvector anisotropy along <111>is significant.\nC. Co 8Zn8Mn4\nIn this section, we describe detailed results of SANS measurements with off-axis field con-\nfiguration in Co 8Zn8Mn4and provide evidence for increased magnetocrystalline anisotropy\nthat favors q∝bardbl<100>at low temperatures.\n1. Temperature-driven structural transition of metastabl e skyrmion lattice\nFirst, we review the temperature variation in metastable SkX under a field parallel to\nthe [001] direction, as reported in our previous paper[34] (see Sup plementary Fig. S5 for\nthe details[57]). In a FC process at 0.04 T, SANS pattern transform s from 12 spots to 4\nspots below 120 K. The 12-spot pattern at high temperatures cor responds to a 2-domain\ntriangular SkX state, in which one of triple- qis parallel to the [010] or [100] direction.\nThe 4-spot pattern at low temperatures is attributed to a square SkX state with double- q\nvectors parallel tothe[010] and[100]directions. This temperatur e variationisreversible and\nthe triangular SkX revives already at 200 K in the subsequent re-wa rming process, which\nrules out the possibility of relaxation to a helical multi-domain state. I t is noted that the\ntransformation to the square SkX below 120 K is accompanied by a lar ge increase in qas\npresented in Fig. 3(g).\n2. Temperature-dependent magnetic anisotropy as revealed by an off-axis field measurement\nNext, we show how qvector orientation changes during the triangular-square struct ural\ntransition of skyrmion lattice under an off-axis magnetic field (Fig. 7) . The experimental\nconfiguration for the SANS measurement is illustrated in Fig. 7(a). T he magnetic field of\n0.03 T was applied along the direction tilted away from the [001] directio n by 15◦. While14\nkeeping this configuration, the cryomagnet and the sample were ro tated together around the\nvertical [010] direction. Here, ωis defined as a rocking angle between the incident neutron\nbeam (ki) and the applied magnetic field ( H), namely ω= 0◦forki∝bardblHandω=−15◦for\nki∝bardbl[001].\nRocking curves (SANS intensity versus ω) at selected temperatures are presented in Fig.\n7(b). Here, integrated intensity at φ=90◦and270◦isplottedagainst ω. Fortheequilibrium\ntriangular SkX state, a rocking curve is expected to take a maximum atω= 0◦because\ntriple-qare usually perpendicular to the applied field regardless of the cryst al orientation.\nHowever, the observed rocking curve at 295 K [red symbols in Fig. 7( b)] shows a broad\nmaximum around ω∼20◦. This indicates that the effective magnetic field inside the sample\n(Heff) is further tilted from the external field to the opposite side of the [001] axis as shown\nin Fig. 7(a), which can be understood in terms of a demagnetization e ffect in the present\nrectangular-shaped sample[60]. As the temperature is lowered, th e peak position of the\nrocking curve shifts to lower angle and eventually locates at −15◦at 40 K.\nTemperature variations of the SANS patterns for ki∝bardblHeffandki∝bardbl[001], averaged over\nthe rocking angles of 14◦≤ω≤20◦and−20◦≤ω≤ −10◦, respectively, are shown in\nFig. 7(c) and (d). For ki∝bardblHeff, a 6-spot pattern is observed at 295 K. This 6-spot pattern\ncorresponds to a single-domain triangular SkX state, in which triple- qare perpendicular to\nHeffand one of them is parallel to the vertical [010] direction. The 6 spot s are still discerned\nat 200 K as the triangular SkX persists as a metastable state. At 12 0 K, the side 4 spots\nbecome much weaker as compared with the vertical 2 spots, and fin ally the 6-spot pattern\nis not discerned at 40 K. For ki∝bardbl[001], on the other hand, only 2 vertical spots out of the 6\nspots are observed at 295 K. Below 120 K, intensities around the ho rizontal region increase,\nand finally a 4-spot pattern is observed at 40 K.\nThe SANS intensity for the two different configurations is plotted ag ainst temperature\nin Fig. 7(e), clearly showing the change in intensity distribution below 1 20 K from the side\n4 spots perpendicular to Heff(blue circles) to the side 2 spots parallel to the [100] direction\n(red squares), in good accord with the large shift of the peak posit ion in the rocking curves\npresented in Fig. 7(b). Therefore, the double- qin the square SkX at low temperatures are\naligned to the [010] and [100] direction, and the latter is not perpend icular to Heff. This is in\nmarked contrast to the triangular SkX state at high temperature s where all the triple- qare\nperpendicular to Heff. Thus, the triangular-square SkX transition under the off-axis fie ld15\nis accompanied by a reorientation of skyrmions as schematically illustr ated in Fig. 7(f).\nThis result indicates that the magnetic anisotropy favoring q∝bardbl<100>is strongly enhanced\nduring the transformation to the square SkX as the temperature is reduced.\nD. Co 9Zn9Mn2\nInthis section, we present how the qvector evolves in terms of magnitude and orientation\nas a function of temperature in Co 9Zn9Mn2. In brief, the change in qdirection, and the\nvariation of qvalue, are qualitatively similar to those found in Co 8Zn8Mn4but take place\nat lower temperatures.\n1. Temperature dependence of metastable SkX\nFirst, we show temperature variation of the metastable SkX state (Fig. 8). Selected\nSANS patterns in a FC process at 0.04 T from 390 K to 10 K are displaye d in Fig. 8(b). In\nthis measurement, the magnetic field and the incident neutron beam are parallel to the [110]\ndirection. At 390 K within the equilibrium SkX phase, the SANS pattern shows 6 spots,\ncorresponding to a triangular SkX with one of triple- q∝bardbl[001] as illustrated in the right\npanel of Fig. 8(a). The triangular SkX persists down to 100 K as a me tastable state during\nthe FC. Below 50 K, the signal from the side 4 spots becomes weaker as compared with the\nvertical 2 spots, as similarly observed in the off-axis field measureme nt for Co 8Zn8Mn4(Fig.\n7(b)).\nFigure 8(d) presents the temperature dependence of the SANS in tensity for the side 4\nspots and the vertical 2 spots, clearly showing the cross-correla tion between the reduced\nintensity for the side 4 spots and the increased intensity for the [00 1] direction below 50 K.\nRocking curves during the FC are displayed in Fig. 8(c). While the rock ing curve exhibits\na peak around ω= 0◦above 100 K, the intensity around ω= 0◦decreases below 50 K and\nfinally the rocking curve forms a concave shape around ω= 0◦at 10 K. These results are\nsimilar to those observed in the SANS pattern for Co 8Zn8Mn4under the off-axis field as\nshown in Fig. 7. Therefore, we attribute the change in the SANS pat tern below 50 K in\nCo9Zn9Mn2to theqvector reorientation and associated transition from a triangular S kX to\na square SkX. Namely, one of the double- qin the square SkX is aligned to the [001] direction16\nbut the other is aligned to [100] or [010] directions that are out of th e (110) plane, resulting\nin 2 vertical spots on the (110) plane as illustrated in the left panel o f Fig. 8(a).\nThe radial qdependence of the SANS intensity for the [001] direction is shown in F ig.\n8(e). While the peak center ( q∼0.05 nm−1) is almost temperature independent above\n100 K, a large shift up to q∼0.07 nm−1is observed below 50 K, accompanied by a slight\nbroadening of the width. Thus, the transition from the triangular S kX to the square SkX in\nCo9Zn9Mn2is also accompanied by the increase in the magnitude and width of the qvector,\nagain similarly as observed in Co 8Zn8Mn4.\n2. Field dependence of metastable SkX\nNext, we discuss field variation in the qvector in the metastable SkX state at 10 K after\nthe FC under 0.04 T (see Supplementary Fig. S6 for the details[57]). W ith increasing field\nfrom 0.04 T to 0.3 T, the 2-spot SANS pattern gradually changes to a ring-like one, in which\nthe scattering intensity lies in the (110) plane as similarly observed at high temperatures.\nThis indicates that the square SkX with q∝bardbl<100>at low temperatures transforms to an\norientationally disordered triangular SkX with q⊥Hat high fields.\nTaking the above results into account, we summarize the state diag ram of the metastable\nSkX in Co 9Zn9Mn2in Fig. 2(c), which is characterized by a large region of the triangular\nSkX state and a small region of the square SkX state existing only at low temperatures\nbelow∼50 K and low fields.\nE. Co 7Zn7Mn6\nIn this section, we discuss how the magnetic state is affected by the increased Mn con-\ncentration in Co 7Zn7Mn6.\n1. Heavily-disordered square lattice within the metastabl e state\nFirst, we discuss the temperature variation of the metastable SkX (Fig. 9). Figure 9(b)\nshows the change in the SANS patterns during a FC at 0.025 T from 14 6 K to 1.5 K. In\nthis measurement, the magnetic field and the incident neutron beam are parallel to the17\n[001] direction. At 146 K within the equilibrium SkX phase, a 12-spot pa ttern correspond-\ning to a 2-domain triangular SkX with one of triple- q∝bardbl[010] or [100], as schematically\nillustrated at right panel in Fig. 9(a), is observed. However, the sp ots display a signifi-\ncant azimuthal broadening as compared with those observed from the equilibrium SkX in\nCo8Zn8Mn4sample. The 12-spot pattern changes to a pattern of 4 broadene d spots at 100\nK, which corresponds to a metastable square SkX with double- q∝bardbl[010] and [100] as shown\nat left panel in Fig. 9(a). With further decreasing temperature do wn to 60 K, the 4 spots\nbecome even broader. This pattern of 4 very broad spots remains almost unchanged across\nthe reentrant spin glass transition ( Tg∼30 K) down to 1.5 K.\nThe SANS intensity from the metastable SkX is plotted as a function o f temperature\nin Fig. 9(c). Above 130 K, the intensities integrated for directions c lose to<100>and\n<110>show similar values as expected for a 12-spot pattern. Below 120 K, the intensity for\n<100>becomes larger than that for <110>due to the change in the pattern from 12 spots\nto 4 spots. The intensities for both regions significantly decrease b elow 120 K, and become\nsimilar again below 60 K, which corresponds to the broadening of the 4 spots. Therefore,\nwhile the triangular-square SkX transition occurs below 120 K in comm on with Co 8Zn8Mn4,\nthe square SkX is severely disordered below ∼90 K. This is in accord with the disordering of\nthe helical state of Co spins below ∼90 K on ZFC[31] due to the development of short-range\nantiferromagnetic correlation of Mn spins.\nThe radial qdependence of the SANS intensity for <100>is presented in Fig. 9(d). From\n120 K to 60 K, where the triangular-square SkX transition occurs, the peak center exhibits\na large shift from 0.06 nm−1to 0.08 nm−1, and the magnitude and width of the peak\nsignificantly decreases and increases, respectively. Thus, the tr iangular-square structural\ntransformation of SkX in Co 7Zn7Mn6is also accompanied by the large increase in qas well\nas in its width. This indicates that the metastable SkX is heavily disorde red, similarly to\nthe helical state during ZFC [Fig. 3(l)].\n2. Field dependence of metastable SkX\nNext, we discuss how the metastable SkX state varies with field swee ping in the well-\nordered region at 100 K (Fig. 10) and in the disordered region at 60 K (Fig. 11).\nThe field variation of the metastable SkX state at 100 K is presented in Fig. 10. Figure18\n10(a) shows the field variation in the SANS patterns at 100 K after a field cooling (FC)\nat 0.025 T. The detailed field dependence of the SANS intensity integr ated over the region\naround<100>and<110>is displayed in Fig. 10(c). With increasing the field to the\npositive direction, the initial 4-spot pattern changes to a uniform r ing above 0.055 T, where\nthe scattering intensities for <100>and<110>completely overlap while showing a clear\nshoulder, and the signal persists up to 0.1 T. This change is ascribed to the transformation\nfrom the square SkX to an orientationally disordered triangular SkX similarly to the case\nof Co8Zn8Mn4[34]. In the field sweeping to the negative direction, the 4-spot patt ern again\nchanges to the ring-like one at −0.055 T but the intensity is much weaker than that observed\nat 0.055 T. In the returning process from 0.15 T to 0 T, clear SANS sig nal is not observed\nbecause the metastable SkX is completely destroyed at a high-field r egion and a conical\nstate is stabilized instead. The large and asymmetric hysteresis of t he intensity plot in Fig.\n10(c) is attributed to the existence of the metastable SkX state. For comparison, the field\ndependence oftheSANSpatternandtheintegratedintensity at1 00Kafterzero-fieldcooling\nis shown in Fig. 10(b) and Fig. 10(d), respectively. In this case, 4 sp ots corresponding to\nthe helical multi-domain state with q∝bardbl<100>is observed up to 0.05 T, but the signal\ndisappears at 0.1 T without exhibiting a clear ring-like patten. This field variation for\nhelical state after ZFC is totally different from that observed afte r the FC. Therefore, it can\nbe concluded that the metastable SkX created by the FC survives o ver a wide field region\nat 100 K, and the square SkX changes to the triangular SkX at high fi elds.\nWe also show the field variation of the metastable SkX state at 60 K in F ig. 11, and\ndiscuss another transition to the second equilibrium skyrmion phase . The field-dependent\nSANS patterns at 60 K after the FC (0.025 T) are presented in Fig. 1 1(a). The SANS\nintensity integrated over the region around <100>and<110>in this process is plotted\nagainst magnetic field in Fig. 11(c). As the field is increased to the pos itive direction, the\ninitial pattern with broad 4 spots originating from the disordered sq uare SkX changes to\na broad ring pattern above 0.07 T, resulting in the almost equal scat tering intensities for\n<100>and<110>. In the field sweeping to the negative field region, the broad ring pat tern\nwith similar intensity is observed at −0.07 T and −0.1 T. Therefore, the field variation after\nthe FC is less asymmetric between the positive fields and the negative fields as compared\nwith the result at 100 K. In the returning process from the field-ind uced ferromagnetic\nphase (±0.2 T) down to 0 T, the broad ring pattern appears again and remains down to19\nzero field. Note that the intensity of the ring pattern around zero field is relatively large and\ncomparable to that of the initial broad 4-spot pattern just after the FC. For comparison,\nwe show SANS patterns observed at the same magnetic fields after ZFC in Fig. 11(b),\nand the field dependence of the SANS intensity is plotted in Fig. 11(d) . In this process,\nthe broad 4-spot pattern (disordered helical state) changes to a broad ring above 0.07 T\nsimilar to the FC case. The field dependence commonly observed for t he FC (both for\npositive and negative field directions) and ZFC processes is in accord with the existence\nof another field-induced equilibrium phase at low temperatures. In o ur previous study on\nCo7Zn7Mn6, the broad ring pattern at low temperatures has been identified to be three-\ndimensionally disordered skyrmions, which are stabilized by frustrat ed magnetism of Mn\nspins and exist as an equilibrium phase that is distinct from the conven tional SkX phase\njust below Tc[31]. Therefore, theobserved change fromthebroad4-spotpat terntothebroad\nring pattern above 0.07 T in the FC case corresponds to the transit ion from the metastable\nsquareSkXstate(originatingfromthehigh-temperatureequilibriu mSkXphase)totheother\nequilibrium disordered skyrmion phase. The ring pattern remaining at zero field after the\nfield decreasing process is explained in terms of the metastable skyr mions that are created\ninitially in the high field region and persist down to zero field.\nOn the basis of the above results, the state diagram of the metast able SkX in Co 7Zn7Mn6\nissummarizedintheFig. 2(e). ThemetastablesquareSkX(M-S-SkX )stateexistsbelow120\nK and at low fields inside the metastable triangular SkX (M-T-SkX) sta te. In addition, the\nequilibrium disordered skyrmion (E-DSk) phase is stabilized at low temp eratures just above\nthe reentrant spin glass transition around 30 K. Note that the E-D Sk phase is discerned also\nin a negative-field region, reflecting its thermodynamical equilibrium n ature.\nF. Summary of helical state and metastable skyrmion state\nTaking the SANS results for all the compounds into account, we com e back to Fig. 2 and\nFig. 3 again and discuss how the helical and the metastable skyrmion s tates change with\nMn concentration.20\n1. Correspondence between magnetization and qvector in helical state\nTable 1 summarizes the several quantities that characterize the h elical state. One im-\nportant effect of Mn substitution is the change in qvector direction: While the preferred\norientations of qvectors are <111>directions in Co 10Zn10, they are <100>directions in all\nthe Mn-doped compounds. The helical structure is formed by the f erromagnetically coupled\nCo spins accompanied with DMI, but as described in Table 1 (see also Su pplementary Fig.\nS1), saturationmagnetization Msmeasured at2Kand7Ttakes amaximum atCo 9Zn9Mn2.\nThis suggests that Co and Mn are ferromagnetically coupled at least in the low Mn concen-\ntration region, as also demonstrated recently by X-ray magnetic c ircular dichroism (XMCD)\nmeasurements[61].\nMn substitution produces another important effect: Figures 3(e- h) summarize the tem-\nperature dependence of the helical qvalue measured for all the compounds. The qvalue was\ndetermined as the peak center of Gaussian function fitted to the a zimuthal-angle-averaged\nSANS intensity as a function of q. In Co 10Zn10,qgradually and slightly increases by only\n∼10% upon cooling from Tcto low temperatures. On the other hand, Co 9Zn9Mn2shows\na large increase in qbelow 50 K by ∼45% while for Co 8Zn8Mn4and Co 7Zn7Mn6,qalso\ndisplays a large increase of ∼50% below 120 K. Notably, the large observed increases in q\nalmost coincide with the gradual decrease in MfromTHtoTLas shown in Fig. 3(a-d). The\ntemperature region where qsignificantly varies is indicated in the T-xphase diagram in Fig.\n1(b).\nFigures 3(i-l) show the temperature dependence of the full width a t half maximum\n(FWHM) of the Gaussian function fitted to the SANS intensity versu sq, which provides\na measure of the spatial coherence of the helimagnetic order. The FWHM increases upon\ncoolinginallthecompositions whileshowing strongcorrelationwiththe increase inthemag-\nnitude of qvalue [Figs. 3(e-h)]. Among them, Co 7Zn7Mn6exhibits a significant increase\nbelow∼90 K, which indicates that the helical state in Co 7Zn7Mn6becomes severely dis-\nordered at low temperatures. The increase in the FWHM for Co 9Zn9Mn2and Co 8Zn8Mn4\nalso indicates the evolution of magnetic disorder to some extent, alt hough they are less\nsignificant as compared with Co 7Zn7Mn6.\nThe above temperature and Mn concentration dependence of helic al states is well ex-\nplained by the interplay between the ferromagnetically coupled Co sp ins forming the helical21\nstate and the antiferromagnetically coupled Mn spins. Namely, the s ignificant increase in\ntheqvalue (or decrease in the helical periodicity) at low temperatures is c aused by the ef-\nfective decrease in the ratio of the ferromagnetic exchange inter action to the DMI, which is\nattributed to short-range antiferromagnetic correlations of th e Mn spins. These short-range\ncorrelations act as a source of disorder for helimagnetic Co spins, a nd start to develop at\nincreasingly higher temperature (higher T/Tc) as the Mn concentration is increased. The\nincrease of the qvalue finally saturates below TLprobably because of the slowing of the\nantiferromagnetic fluctuations of Mn spins and resulting in a quasi-s tatic disorder for the\nhelimagnetic Co spins. As the temperature is further reduced, Mn s pins eventually freeze\nand undergo the reentrant spin glass transition as observed in the magnetization measure-\nments for Co 8Zn8Mn4and Co 7Zn7Mn6[Figs. 3 (c) and (d)]. The freezing temperature Tg\nalso increases with the Mn concentration due to the associated enh ancement of the antifer-\nromagnetic Mn spin correlations.\n2. Summary of metastable skyrmion state\nFigure 2 is the summary of the equilibrium and metastable skyrmion pha se diagrams\non theT-Hplane determined by SANS and ac susceptibility measurements in Co 10Zn10,\nCo9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6. In all the compounds, the metastable SkX state\nis realized by a conventional field cooling via the equilibrium SkX phase ju st below Tcand\nprevailsover averywidetemperatureandfieldregion. Inaddition, w hiletheequilibriumand\nmetastable skyrmions at high temperatures form a conventional t riangular lattice described\nwith triple- qvectors, the lattice structure transforms to the novel double- qstates at low\ntemperatures as follows. In Co 10Zn10[Fig. 2(b)], the lattice form of the metastable SkX\ntransforms from triangular to a rhombic one (M-R-SkX) with double -q∝bardbl<111>below∼\n360 K with a broad coexistence region (a purple region). At all tempe ratures, the triangular\nlattice is restored as the field is increased. In Co 9Zn9Mn2[Fig. 2(c)], while the triangular\nlattice of the metastable skyrmions (M-T-SkX) persists over a wide temperature region, it\ntransforms to square one (M-S-SkX) with double- q∝bardbl<100>below∼50 K as shown with a\npink region. The triangular lattice is also recovered at high fields. In C o8Zn8Mn4[Fig. 2(d)]\nandCo 7Zn7Mn6[Fig. 2(e)], the triangular-squaretransition ofskyrmion lattice occ urs below\n∼120 K, and the square SkX persists below the reentrant spin glass t ransition temperatures22\nas plotted with yellow circles. In Co 7Zn7Mn6, the metastable square SkX state originating\nfrom the conventional SkX phase undergoes another transition t o the other frustration-\ninduced equilibrium skyrmion phase (E-DSk; orange region) as the fie ld is increased at\ntemperatures below ∼60 K.\nThe structural transitions of the skyrmion lattice to the rhombic o ne and to the square\noneareperhapsattributedtotheenhancedmagnetocrystallinea nisotropytoward q∝bardbl<111>\nandq∝bardbl<100>, respectively. However, the qvector anisotropy alone is not sufficient to\ndrive the lattice structural transition. Importantly, these tran sformations are accompanied\nby an increase in absolute value of q: the triangular-rhombic transition in Co 10Zn10is\nobserved while qvalue slightly increases over a broad temperature range [Fig. 3(e)], and the\ntriangular-square transition in Mn-doped compounds occurs only w henqvalue significantly\nincreases below a specific temperature of TH[Fig. 3(f-h)]. As discussed quantitatively in the\nfollowing, the increase in qvalue is crucial for the transformation of the skyrmion lattice in\nterms of skyrmion density.\nThe ratio of skyrmion density (number of the skyrmion per area) in t he rhombic lattice\n(nR) to that in the triangular one ( nT) is given by nR/nT=3√\n3\n4√\n2(aT/aR)2=3√\n3\n4√\n2(qR/qT)2.\nHere,aT(qT) andaR(qR) are the lattice constant (the qvalue) of the triangular lattice\nand the rhombic one, respectively. Since the total number of meta stable skyrmions that are\ntopologically protected is conserved during the transition from the triangular lattice to the\nrhombic one ( nR=nT), the lattice constant should decrease (the qvalue should increase) by\nthe factor of/radicalBig\n4√\n2\n3√\n3∼1.043 as the temperature is lowered. The qRvalue that is consistent\nwith this condition is presented with an dashed purple line in the q(T) plot for Co 10Zn10\n[inset of Fig. 3(e)]. Here, qTis taken as the qvalue at 412 K. It is found that the observed\nsmall increase in qvalue during the FC in Co 10Zn10easily satisfies the above condition\naround∼200 K. This result is in accord with the fact that the transformation to rhombic\nlattice starts at high temperatures.\nIn the case of the transition to square lattice in the Mn-doped comp ounds, on the\nother hand, the aforementioned condition is more difficult to fulfill. Th e ratio of the\nskyrmion density in the square lattice ( nS) to that in the triangular one is given by\nnS/nT=√\n3\n2(aT/aS)2=√\n3\n2(qS/qT)2, where aS(qS) is the lattice constant (the qvalue)\nof the square lattice. To keep the skyrmion density constant ( nS=nT), the lattice constant\nshould decrease (the qvalue should increase) by the larger factor of/radicalBig\n2√\n3∼1.075. The23\nvalue ofqSsatisfying this condition is denoted with an dashed pink line in the q(T) plot in\nCo9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6[Fig. 3(f-h)]. Here, qTis taken as the qvalue at\nthe equilibrium SkX phase. In Co 9Zn9Mn2and Co 8Zn8Mn4,qvalue is almost T-constant\nat high temperatures. However, qvalue significantly increases and exceeds the necessary\nqSvalue below TH(∼50 K for Co 9Zn9Mn2and∼120 K for Co 8Zn8Mn4and Co 7Zn7Mn6),\nwhich would describe the transformation to the square lattice. This result well explains\nthe observed onset temperature of the transformation to the s quare lattice. Nevertheless,\nthe observed increase in qaroundTLis as large as qS/qT∼1.5, and the skyrmion density\nbecomes too large if we assume a uniformly and perfectly ordered sq uare SkX as shown in\nFig. 1(e).\nAlternatively, the large increase in qunder the conserved skyrmion density can be rec-\nonciled with either of the following two scenarios: (i) a microscopic pha se separation into\na square SkX state and a helical phase as schematically illustrated in F ig. 1(f), or (ii) the\nonset of skyrmion deformation along the <100>directions [Fig. 1(g)]. It should be noted\nthat, in both (i) and (ii), the total number of skyrmions is identical t o the original triangular\nSkX. Although the latter deformed skyrmions have been observed in a thin-plate sample by\nLTEM[37] and reproduced by a micromagnetic simulation[61], it is difficult to experimen-\ntally distinguish the two scenarios from SANS studies on a three-dime nsional bulk sample.\nIn any case, the large increase in qvalue at low temperatures and the enhanced qvector\nanisotropy toward q∝bardbl<100>cooperatively drive the skyrmion lattice transformation as\nthe metastable triangular SkX phase is cooled down to low temperatu res.\nIt is also noted that the square lattice of (elongated) skyrmions in t he present case is\ndistinct from a meron-antimeron square lattice with vanishing topolo gical charge (genuine\nsuperposition of orthogonal double qvectors) as recently observed in a thin-plate sample of\nCo8Zn9Mn3by LTEM[39]. The meron-antimeron state appears as an equilibrium st ate just\nbefore entering the equilibrium triangular SkX phase near Tc. While the meron-antimeron\nsquare lattice is stabilized by the in-plane shape anisotropy charact eristic of a thin-plate\nspecimen as theoretically predicted[62, 63], the transition to the sq uare SkX in the present\ncase probably originates from magnetocrystalline anisotropy (loca l magnetization ∝bardbl<100>,\nwhich in turn results in qvector anisotropy along <100>) as well as the large increase in q,\nas discussed above.24\nG. Lifetime of metastable skyrmion\nIn the former sections, we revealed that the SkX state can persis t at low temperatures\nby a field cooling process. Since the low temperature states are only metastable, the SkX\nshould relax to the more stable conical state with some lifetime. In ou r previous study\n(Ref. [36]),weinvestigatedtemperature-dependent relaxationt imesofmetastableskyrmions\nin Co9Zn9Mn2and observed extremely long lifetimes even at high temperatures. F or a\nsystematic understanding of behavior, in this section we extend th e lifetime measurement\nto the other compounds Co 10Zn10and Co 8Zn8Mn4, and discuss how the metastable SkX\nlifetime varies with temperature and Mn concentration (Fig. 12 and F ig. 13).\nFigures 12(a-c) show H-Tphase diagrams of equilibrium states near Tcin Co10Zn10,\nCo9Zn9Mn2and Co 8Zn8Mn4, respectively. For the purpose of better comparison between\nthe three compounds, Tranges are selected in the respective panels in such a way that nor-\nmalized temperature T/Tcis from 0.90 to 1.01, as indicated on the upper abscissa. Clearly,\nthe equilibrium SkX phase (green region) expands upon increasing th e Mn concentration\ndue to the increased chemical and magnetic disorder. After a FC via the equilibrium SkX\nphase, the temporal variation of ac susceptibility χ′(t) was measured at a fixed temperature,\nand this experiment was repeated at several different temperatu res, as denoted with the\ncolored circles.\nThe normalized ac susceptibility χ′\nN(t)≡[χ′(∞)−χ′(t)]/[χ′(∞)−χ′(0)] is plotted as a\nfunction of time in Fig. 12(d-f). Here, χ′(0) andχ′(∞) correspond to an initial value for\nthe metastable SkX state and the value for the equilibrium conical st ate as a fully relaxed\nstate, respectively, and hence χ′\nN= 1 fort= 0 and χ′\nN= 0 fort→ ∞. In Co 10Zn10[Fig.\n12(d)], a clear relaxation from the metastable SkX state to the equ ilibrium conical state is\nobserved at 404 K, just below the equilibrium SkX phase, and for whic h the relaxation\ntime is the order of 104s (several hours). The relaxation time further increases upon\nlowering the temperature. In Co 9Zn9Mn2[Fig. 12(e)], the observed relaxation in the χ′\nN(t)\ncurve within the measurement time ( ∼1 day) is less than 40% even at 380 K, just below\nthe equilibium SkX phase, and the relaxation becomes even longer as t he temperature\nis lowered. In Co 8Zn8Mn4[Fig. 12(f)], only a few % of relaxation is observed in the\nχ′\nN(t) even at 280 K, just below the equilibrium SkX phase. Therefore, th e lifetime of\nmetastable SkX at temperatures close to the equilibrium SkX phase b oundary, for example25\nthe temperatures indicated with the red circles in Fig. 12(a-c), bec omes relatively longer as\nthe Mn concentration is increased.\nTodiscuss morequantitatively, the χ′\nN(t)curves arefittedtoastretched exponential func-\ntion, exp/braceleftbig\n−(t/τ)β/bracerightbig\n. The obtained βvalues are in the range of 0.3 - 0.5, indicating a highly\ninhomogeneous distribution of relaxation times. The relaxation time τin all the compounds\n(Co10Zn10, Co9Zn9Mn2and Co 8Zn8Mn4) is plotted against T/Tcin Fig. 13(a). The τvalue\nexponentially increases as the temperature is lowered, and become s virtually infinite when\nT/Tcis less than ∼0.9. Remarkably, the data points from the three different compoun ds\ncollapse onto a single curve althoughthe applied magnetic field as well a s the respective tem-\nperature regions are different. Following the arguments described in Ref. [64], all the data\npoints are fitted to a modified Arrhenius law (pink solid line), τ=τ0exp{a(Tc−T)/T}.\nHere, the activation energy for the relaxation from the metastab le SkX state to the equilib-\nrium conical state, as schematically illustrated in Fig. 13(b), is assum ed to be T-dependent\nasEg=a(Tc−T) nearTc, instead of constant Egfor standard Arrhenius law. The obtained\nfitting parameters are a= 215 and τ0= 63 s. For comparison, reported values of aand\nτ0for several materials[64, 65] are summarized in Table 2. Recently, W ildet al. reported\nthatτ0sensitively depends on the applied magnetic field in their LTEM studies o f a thin\nplate of Fe 1−xCoxSi[66]. For the present ac susceptibility measurements on bulk cryst als of\nCo-Zn-Mn alloys, we used the applied magnetic field for each compoun d where the SANS\nintensity from the equilibrium SkX is strongest in the field sweeping mea surement. The fact\nthat theτvalues from three different compounds collapse onto a single curve ( namely,τ0are\nthe same) may originate from the fact that the applied field values ar e close to the optimal\nones in the respective samples.\nFromthe obtained value of a, the activation energy Egis estimated to be larger than8000\nK at a temperature T= 0.9Tcfor Co 10Zn10and Co 9Zn9Mn2. This energy scale protecting\nthe metastable skyrmion state is much larger than the ferromagne tic exchange interaction\n(several 100 K), and thus attributed to the topological nature o f the skyrmion with a large\ndiameter ( ∼100 nm) that involves a great number of spins.\nThe coefficient of relaxation time τ0is inversely correlated to the critical cooling rate[67]\nfor quenching the SkX phase to lower temperatures. Since the obt ained value of τ0in Co-\nZn-Mn alloys is the order of a minute, the conventionally slow cooling ra te (dT/dt∼ −1\nK/min) is sufficient to quench the SkX phase. This is quite different fro m the value of τ0∼26\n10−4s in MnSi[64], where an ultra-rapid cooling rate ( dT/dt∼ −100 K/s) is necessary to\nquench the SkX phase. The large difference in τ0that depends on the material is probably\nattributed to randomness in the system. In the case of Co-Zn-Mn alloys, there are random\nsite occupancies in the crystal structure; the 8 csite is randomly occupied by Co and Mn,\nand the 12 dsite is occupied by Co, Zn, and Mn[27–30]. This gives rise to “weak pinnin g” in\nthe terminology of density wave physics[68], which may play an importa nt role in the robust\nmetastability of the skyrmion.\nFrom a microscopic viewpoint, the destruction of metastable skyrm ions in the bulk takes\nplace through the creation of a pair of Bloch points, or equivalently a n emergent magnetic\nmonopole-antimonopole pair, from a singularity point of a skyrmion st ring, followed by their\npropagation[69, 70], as schematically illustrated in Fig. 13(c). Coeffic ientsaandτ0are\nroughly governed by the creation and the propagation processes of monopole-antimonopole\npairs, respectively. In a clean system like MnSi, the monopole and ant imonopole can easily\nmove, which leads to skyrmion string destruction after the pair cre ation. In a dirty system\nlike Co-Zn-Mn alloys, the movement of monopole and antimonopole is hin dered by magnetic\nimpurities or defects, and consequently τ0of the metastable skyrmion string is significantly\nincreased. A similar long-lived metastable SkX that is accessible by a mo derate cooling rate\nhas been reported in Fe 1−xCoxSi alloys[66, 69, 71], which may bear some resemblance to the\npresent case. More recently, the increased lifetime of metastable SkX has been observed also\nin Zn-doped Cu 2OSeO3, whereτ0in a Zn 2.5% doped sample is 50 times larger than that\nin a non-doped sample while avalue is unchanged[65]. In the present case of Co-Zn-Mn\nalloys, the good scaling of the τvsT/Tcplot in Fig. 13(a) indicates that both aandτ0are\nalmost independent of the Mn concentration. This is probably becau se Co10Zn10already\npossesses substantial randomness in the site occupancy at 12 dsite (2 Co and 10 Zn per unit\ncell), and thus the attempt time τ0is already sufficiently long and not further increased by\nadditional randomness due to the Mn substitution. Nevertheless, Mn substitution expands\nthe equilibrium SkX phase toward lower temperature as seen in Fig. 12 (a-c) and Fig. 13(a),\nand hence the relaxation time just below the equilibrium SkX phase bec omes longer. It\nis also interesting to note that the equilibrium skyrmion phase is expan ded by the static\nmagnetic disorder while thermal fluctuation is considered to be impor tant for the realization\nof the equilibrium phase close to Tc.27\nIV. CONCLUSION\nIn the present study, to provide the perspective on the chiral ma gnetism in β-Mn-type\nCo-Zn-Mn alloys with bulk Dzyaloshinskii-Moriya interaction (DMI), we have performed\nmagnetization, ac susceptibility and small-angle neutron scattering measurements on single-\ncrystal samples of (Co 0.5Zn0.5)20−xMnxwithx= 0, 2, 4 and 6. The Mn-free end member\nCo10Zn10exhibits a helimagnetic ground state (periodicity λ∼156 nm) below the transition\ntemperature Tc∼414 K, where the helical propagation vector qis aligned to <111>at low\ntemperatures (Fig. 5). Upon applying magnetic fields, Co 10Zn10exhibits an equilibrium\nSkX phase above 400 K in a narrow temperature and magnetic field re gion (Fig. 4), which\nis quenched down to lower temperatures as a metastable state by a conventionally-slow field\ncooling (Fig. 6). The lifetime of the metastable SkX is extremely long an d virtually infinite\nbelow 380 K (Figs. 12 and 13). The metastable SkX state is highly robu st and persists\nover the whole temperature range below Tcand a wide magnetic field region, including\nroom temperature and zero field (Fig. 2). The lattice of metastable skyrmions distorts and\ntransforms from a conventional triangular one to a rhombic one at low temperatures and\nlow magnetic fields (Figs. 2 and 6).\nAs the partial substitution with Mn proceeds, Tcdecreases and the preferred qorienta-\ntion switches to <100>. The saturation magnetization is the largest for Co 9Zn9Mn2(Table\n1). At low temperatures, the helical qvalue significantly increases, or equivalently λsignif-\nicantly decreases, below TH∼50 K for Co 9Zn9Mn2and below TH∼120 K for Co 8Zn8Mn4\nand Co 7Zn7Mn6(Fig. 3). In Co 7Zn7Mn6, the helical state is severely disordered at low\ntemperatures. Upon further decreasing temperature, a reent rant spin glass transition oc-\ncurs atTg∼10 K for Co 8Zn8Mn4andTg∼30 K for Co 7Zn7Mn6while such a transition\nis not observed for Co 10Zn10and Co 9Zn9Mn2down to 2 K (Figs. 1 and 3). In common\nwith Co 10Zn10, a long-lived metastable SkX state is realized in the Mn-doped materia ls by a\nmoderate FC through the equilibrium SkX phase, and persists over a wide temperature and\nfield region. On the other hand, the lattice form of the metastable S kX changes to a square\none at low temperatures (Figs. 2, 7, 8 and 9). While the triangular Sk X is dictated by the\napplied field as q⊥H, the square SkX is governed by the magnetocrystalline anisotropy as\nq∝bardbl<100>regardless of the applied field direction (Figs. 7 and 8). During the tr ansition\nto the square lattice, the periodicity of SkX significantly shrinks belo wTHsimilar to the28\nhelical periodicity in zero-field cooling (Fig. 3).\nFrom these results, we conclude the followings: The helical and skyr mion states in Co-\nZn-Mn alloys are basically formed by ferromagnetic Co spins in the pre sence of the DMI.\nWhile Co and Mn are ferromagnetically coupled at least in the low Mn conc entrations, an-\ntiferromagnetic Mn-Mn correlations become increasingly significant in the higher Mn con-\ncentrations and start to develop at higher temperatures. As the temperature is lowered, the\ndevelopment of antiferromagnetic Mn-Mn correlation leads to a diso rdering of the helical\nstate and simultaneously decreases helical pitch, and ultimately und ergoing a spin freezing\ntransition at very low temperatures. The robust metastability of s kyrmions is attributed to\nthe topological protection by a large number of involved Co spins as w ell as weak pinning\nfrom the magnetic disorder. The structural transformations be tween metastable triangular\nSkX and either a rhombic one or a square one are driven by a decreas e in the distance\nbetween the skyrmions under the influence of the magnetocrysta lline anisotropy that favors\nq∝bardbl<111>andq∝bardbl<100>in undoped and Mn-doped compounds, respectively. These\nfindings unveil the complex interplay between chiral magnetism and t he frustrated Mn spins\nthat is also greatly affected by magnetic disorder and anisotropy, a nd provide a significant\nunderstanding of the topological phases and properties in this clas s ofβ-Mn-type chiral\nmagnets.\nAcknowledgments\nWe are grateful to T. Arima, T. Nakajima, D. Morikawa, X. Z. Yu, L. C. Peng and N.\nNagaosa for fruitful discussions. We thank M. Bartkowiak for sup port of SANS experiments\nabove room temperature at Paul Scherrer Institute (PSI), Swit zerland. This work was\nsupported by JSPS Grant-in-Aids for Scientific Research (Grant N o. 24224009 and No.\n17K18355), JST CREST (Grant No. JPMJCR1874), the Swiss Nation al Science Foundation\n(SNSF)Sinergianetwork‘NanoSkyrmionics(GrantNo. CRSII5 171003)’,theSNSFprojects\n200021153451, 200021 188707 and 166298, and the European Research Council project\nCONQUEST.\nThe datasets for the SANS experiments done on D33 at the ILL are avail-\nable through the ILL data portal (http://doi.ill.fr/10.5291/ILL-DA TA.5-42-410 and29\nhttp://doi.ill.fr/10.5291/ILL-DATA.5-42-443).\n[1] A. N. Bogdanov and D. A. Yablonskii, Sov. Phys. JETP 68, 101 (1989).\n[2] N. Nagaosa and Y. Tokura, Nat. Nanotechnol. 8, 899 (2013).\n[3] S. M¨ uhlbauer, B. Binz, F. Jonietz, C. Pfleiderer, A. Rosc h, A. Neubauer, R. Georgii and P.\nB¨ oni, Science 323, 915 (2009).\n[4] X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han, Y. Mat sui, N. Nagaosa and Y.\nTokura, Nature 465, 901 (2010).\n[5] F. Jonietz, S. Muhlbauer, C. Pfleiderer, A. Neubauer, W. M ¨ unzer, A. Bauer, T. Adams, R.\nGeorgii, P. B¨ oni, R. A. Duine, K. Everschor, M. Garst and A. R osch, Science 330, 1648\n(2010).\n[6] X. Z. Yu, N. Kanazawa, W. Z. Zhang, T. Nagai, T. Hara, K. Kim oto, Y. Matsui, Y. Onose\nand Y. Tokura, Nat. Commun. 3, 988 (2012).\n[7] J. Iwasaki, M. Mochizuki and N. Nagaosa, Nat. Nanotechno l.8, 742 (2013).\n[8] S. Heinze, K. von Bergmann, M. Menzel, J. Brede, A. Kubetz ka, R. Wiesendanger, G.\nBihlmayer and S. Bl¨ ugel, Nat. Phys. 7, 713 (2011).\n[9] N. Romming, C. Hanneken, M. Menzel, J. E. Bickel, B. Wolte r, K. Bergmann, A. Kubetzka\nand R. Wiesendanger, Science 341, 636 (2013).\n[10] J. Matsuno, N. Ogawa, K. Yasuda, F. Kagawa, W. Koshibae, N. Nagaosa, Y. Tokura and M.\nKawasaki, Sci. Adv. 2, e1600304 (2016).\n[11] S. Woo, K. Litzius, B. Kr¨ uger, M.-Y. Im, L. Caretta, K. R ichter, M. Mann, A. Krone, R.\nReeve, M. Weigand, P. Agrawal, P. Ischer, M. Kl¨ aui, G. S. D. B each, I. Lemesh and M.-A.\nMawass, Nat. Mater. 15, 501 (2016).\n[12] A. Soumyanarayanan, M. Raju, A. L. G. Oyarce, A. K. C. Tan , M.-Y. Im, A. P. Petrovic, P.\nHo, K. H. Khoo, M. Tran, C. K. Gan, F. Ernult and C. Panagopoulo s, Nat. Mater. 16, 898\n(2017).\n[13] X. Z. Yu, N. Kanazawa, Y. Onose, K. Kimoto, W. Z. Zhang, S. Ishiwata, Y. Matsui and Y.\nTokura, Nat. Mater. 10, 106 (2011).\n[14] S. Seki, X. Z. Yu, S. Ishiwata and Y. Tokura, Science 336, 198 (2012).\n[15] Y. Tokunaga, X. Z. Yu, J. S. White, H. M. Rønnow, D. Morika wa, Y. Taguchi and Y. Tokura,30\nNat. Commun. 6, 7638 (2015).\n[16] W. Li, C. Jin, R. Che, W. Wei, L. Lin, L. Zhang, H. Du, M. Tia n and J. Zang, Phys. Rev. B\n93, 060409(R) (2016).\n[17] I. K´ ezsm´ arki, S. Bord´ acs, P. Milde, E. Neuber, L. M. E ng, J. S. White, H. M. Rønnow, C. D.\nDewhurst, M. Mochizuki, K. Yanai, H. D. Ehlers, V. Tsurkan an d A. Loidl, Nat. Mater. 14,\n1116 (2015).\n[18] T. Kurumaji, T. Nakajima, V. Ukleev, A. Feoktystov, T. A rima, K. Kakurai and Y. Tokura,\nPhys. Rev. Lett. 119, 237201 (2017).\n[19] A. K. Nayak, V. Kumar, T. Ma, P. Werner, E. Pippel, R. Saho o, F. Damay, U. K. Rosler, C.\nFelser and S. S. P. Parkin, Nature 548, 561 (2017).\n[20] X. Z. Yu, Y. Tokunaga, Y. Kaneko,W. Z. Zhang, K. Kimoto, Y . Matsui, Y. Taguchi and Y.\nTokura, Nat. Commun. 5, 3198 (2014).\n[21] W. Wang, Y. Zhang, G. Xu, L. Peng, B. Ding, Y. Wang, Z. Hou, X. Zhang, X. Li, E. Liu, S.\nWang, J. Cai, F. Wang, J. Li, F. Hu, G. Wu, B. Shen and X.-X. Zhan g, Adv. Mater. 28, 6887\n(2016).\n[22] Z. Hou, W. Ren, B. Ding, G. Xu, Y. Wang, B. Yang, Y. Zhang, E . Liu, F. Xu, W. Wang, G.\nWu, X. Zhang, B. Shen and Z. Zhang, Adv. Mater. 29, 1701144 (2017).\n[23] R. Takagi, X. Z. Yu, J. S. White, K. Shibata, Y. Kaneko, G. Tatara, H. M. Rønnow, Y. Tokura\nand S. Seki., Phys. Rev. Lett. 120, 037203 (2018).\n[24] M. Kakihana, D. Aoki, A. Nakamura, F. Honda, M. Nakashim a, Y. Amako, S. Nakamura, T.\nSakakibara, M. Hedo, T. Nakama and Y. ¯Onuki, J. Phys. Soc. Jpn. 87, 023701 (2018).\n[25] T. Kurumaji, T. Nakajima, M. Hirschberger, A. Kikkawa, Y. Yamasaki, H. Sagayama, H.\nNakao, Y. Taguchi, T. Arima and Y. Tokura, Science 365, 914 (2019).\n[26] M. Hirschberger, T. Nakajima, S. Gao, L. Peng, A. Kikkaw a, T. Kurumaji, M. Kriener, Y.\nYamasaki, H. Sagayama, H. Nakao, K. Ohishi, K. Kakurai, Y. Ta guchi, X. Z. Yu, T. Arima\nand Y. Tokura, Nat. Commun. 10, 5831 (2019).\n[27] T. Hori, H. Shiraish and Y. Ishii, J. Magn. Magn. Mater. 310, 1820 (2007).\n[28] W. Xie, S. Thimmaiah, J. Lamsal, J. Liu, T. W. Heitmann, D . Quirinale, A I. Goldman, V.\nPecharsky and G. J. Miller, Inorg. Chem. 52, 9399 (2013).\n[29] J. D. Bocarsly, C. Heikes, C. M. Brown, S. D. Wilson, and R . Seshadri, Phys. Rev. Mater. 3,\n014402 (2019).31\n[30] T.Nakajima, K.Karube,Y. Ishikawa, M. Yonemura, N. Rey nolds, J.S.White, H. M.Rønnow,\nA. Kikkawa, Y. Tokunaga, Y. Taguchi, Y. Tokura, and T. Arima, Phys. Rev. B 100, 064407\n(2019).\n[31] K. Karube, J. S. White, D. Morikawa, C. D. Dewhurst, R. Cu bitt, A. Kikkawa, X. Z. Yu,\nY. Tokunaga, T. Arima, H. M. Rønnow, Y. Tokura, and Y. Taguchi , Sci. Adv. 4, eaar7043\n(2018).\n[32] R. Takagi, D. Morikawa, K. Karube, N. Kanazawa, K. Shiba ta, G. Tatara, Y. Tokunaga, T.\nArima, Y. Taguchi, Y. Tokura, and S. Seki, Phys. Rev. B 95, 220406(R) (2017).\n[33] K. Karube, K. Shibata, J. S. White, T. Koretsune, X. Z. Yu , Y. Tokunaga, H. M. Rønnow,\nR. Arita, T. Arima, Y. Tokura, and Y. Taguchi, Phys. Rev. B 98, 155120 (2018).\n[34] K. Karube, J. S. White, N. Reynolds, J. L. Gavilano, H. Oi ke, A. Kikkawa, F. Kagawa, Y.\nTokunaga, H. M. Rønnow, Y. Tokura, and Y. Taguchi, Nat. Mater .15, 1237 (2016).\n[35] As explained in Section III-F-2, there are two possibil ities for the observed double- qstate\nwith increased qvalues: One is a coexistence of the square lattice of skyrmio ns and the helical\nstate [Fig. 1(f)], and the other is elongated-skyrmion stat e on the square lattice [Fig. 1(g)].\nHereafter, we refer to the double- qstate as “square lattice of skyrmions (or square SkX)” for\nthe purpose of simplicity throughout the paper.\n[36] K. Karube, J. S. White, D. Morikawa, M. Bartkowiak, A. Ki kkawa, Y. Tokunaga, T. Arima,\nH. M. Rønnow, Y. Tokura, and Y. Taguchi, Phys. Rev. Mater. 1, 074405 (2017).\n[37] D. Morikawa, X. Z. Yu, K. Karube, Y. Tokunaga, Y. Taguchi , T. Arima and Y. Tokura, Nano\nLett.17, 1637 (2017).\n[38] T. Nagase, M. Komatsu, Y. G. So, T.Ishida, H. Yoshida, Y. Kawaguchi, Y. Tanaka, K. Saitoh,\nN. Ikarashi, M. Kuwahara, and M. Nagao, Phys. Rev. Lett. 123, 137203 (2019).\n[39] X. Z. Yu, W. Koshibae, Y. Tokunaga, K. Shibata, Y. Taguch i, N. Nagaosa and Y. Tokura,\nNature564, 95 (2018).\n[40] H. Nakamura, K. Yoshimoto, M. Shiga, M. Nishi, and K. Kak urai, J. Phys. Condens. Matter\n9, 4701 (1997).\n[41] J. R. Stewart, B. D. Rainford, R. S. Eccleston, and R. Cyw inski, Phys. Rev. Lett. 89, 186403\n(2002).\n[42] J. R. Stewart, K. H. Andersen, and R. Cywinski, Phys. Rev . B78, 014428 (2008).\n[43] J. R. Stewart and R. Cywinski, J. Phys.: Condens. Matter 21, 124216 (2009).32\n[44] J. R. Stewart, A. D. Hillier, J. M. Hillier, and R. Cywins ki, Phys. Rev. B 82, 144439 (2010).\n[45] J. A. M. Paddison, J. R. Stewart, P. Manuel, P. Courtois, G. J. McIntyre, B. D. Rainford,\nand A. L. Goodwin, Phys. Rev. Lett. 110, 267207 (2013).\n[46] It should be noted that the helimagnetic order formed by Co spins and the spin glass state of\nfrustrated Mn spins coexist. Therefore, it may also be calle d “embedded spin glass” as used\nin Ref. [51].\n[47] B. R. Coles, B. V. B. Sarkissian, and R. H. Taylor, Phil. M ag. B37, 489 (1978).\n[48] K. Motoya, S. M. Shapiro, and Y. Muraoka, Phys. Rev. B 28, 6183 (1983).\n[49] H. Maletta, G. Aeppli, and S. M. Shapiro, J. Magn. Magn. M ater.31, 1367 (1983).\n[50] N. Hanasaki, K. Watanabe, T. Ohtsuka, I. K´ ezsm´ arki, S . Iguchi, S. Miyasaka, and Y. Tokura,\nPhys. Rev. Lett. 99, 086401 (2007).\n[51] J. O. Piatek, B. Dalla Piazza, N. Nikseresht, N. Tsyruli n, I.ˇZivkovi´ c, K. W. Kr¨ amer, M.\nLaver, K. Prokes, S. Mataˇ s, N. B. Christensen, and H. M. Rønn ow, Phys. Rev. B 88, 014408\n(2013).\n[52] I.Mirebeau, N. Martin, M.Deutsch, L.J.Bannenberg, C. Pappas, G. Chaboussant, R. Cubitt,\nC. Decorse, and A. O. Leonov, Phys. Rev. B 98, 014420 (2018).\n[53] H. Yoshizawa, S. Mitsuda, H. Aruga, and A. Ito, J. Phys. S oc. Jpn. 58, 1416 (1989).\n[54] C. J. Leavey, J. R. Stewart, B. D. Rainford, and A. D. Hill ier, J. Phys.: Condens. Matter 19,\n145288 (2007).\n[55] T. Sato and K. Morita, J. Phys.: Condens. Matter 11, 4231 (1999).\n[56] In Ref. [31] we reported for Co 7Zn7Mn6that the real part of the ac susceptibility exhibits a\ndrop, the imaginary part a peak at Tg∼30 K, and the frequency-dependent Tgincreases (29\nK→34 K) as the ac frequency fis increased (0.1 Hz →1000 Hz). The frequency-dependence\nofTgsatisfies a power law f−1=τ0[Tg(f)/Tg(0)−1]−zν(τ0= 5.3×10−8s,Tg(0) = 24.6\nK,zν= 10.7). The slow spin flipping time τ0∼10−8s is a typical value for reentrant spin\nglass[50]. A similar measurement for Co 7Zn7Mn6was also reported in Ref. [29].\n[57] See Supplementary Materials at [URL will be inserted by publisher] for details.\n[58] The asymmetric peak broadenings around the [-111] and [ 1-1-1] directions at 1.5 K are at-\ntributed to imperfect summations of SANS patterns along the rocking direction.\n[59] Y. Togawa, T. Koyama, K. Takayanagi, S. Mori, Y. Kousaka , J. Akimitsu, S. Nishihara, K.\nInoue, A. S. Ovchinnikov, and J. Kishine, Phys. Rev. Lett. 108, 107202 (2012).33\n[60] On the basis of demagnetization coefficients in the recta ngular-shaped sample and a magne-\ntization value at 295 K, the extra tilt angle αof the effective magnetic field Hefffrom the\nexternal field Hwas calculated to be α= 23◦, showing a good agreement with the observed\npeak position of the rocking curve at 295 K.\n[61] V. Ukleev, Y. Yamasaki, D. Morikawa, K. Karube, K. Shiba ta, Y. Tokunaga, Y. Okamura,\nK. Amemiya, M. Valvidares, H. Nakao, Y. Taguchi, Y. Tokura, a nd T. Arima, Phys. Rev. B\n99, 144408 (2019).\n[62] S. D. Yi, S. Onoda, N. Nagaosa, and J. H. Han, Phys. Rev. B 80, 054416 (2009).\n[63] S.-Z. Lin, A. Saxena and C. D. Batista, Phys. Rev. B 91, 224407 (2015).\n[64] H. Oike, A. Kikkawa, N. Kanazawa, Y. Taguchi, M. Kawasak i, Y. Tokura and F. Kagawa,\nNat. Phys. 12, 62 (2016).\n[65] M. T. Birch, R. Takagi, S. Seki, M. N. Wilson, F. Kagawa, A .ˇStefanˇ ciˇ c, G. Balakrishnan, R.\nFan, P. Steadman, C. J. Ottley, M. Crisanti, R. Cubitt, T. Lan caster, Y. Tokura, and P. D.\nHatton, Phys. Rev. B 100, 014425 (2019).\n[66] J. Wild, T. N. G. Meier, S. Pollath, M. Kronseder, A. Baue r, A. Chacon, M. Halder, M.\nSchowalter, A. Rosenauer, J. Zweck, J. Muller, A. Rosch, C. P fleiderer, and C. H. Back, et\nal., Sci. Adv. 3, e1701704 (2017).\n[67] F. Kagawa and H. Oike, Adv. Mater. 29, 1601979 (2017).\n[68] G. Gr¨ uner, Rev. Mod. Phys. 60, 1129 (1988).\n[69] P. Milde, D. K¨ ohler, J. Seidel, L. M. Eng, A. Bauer, A. Ch acon, J. Kindervater, S. M¨ uhlbauer,\nC. Pfleiderer, S. Buhrandt, C. Sch¨ utte and A. Rosch, Science 340, 1076 (2013).\n[70] F. Kagawa, H. Oike, W. Koshibae, A. Kikkawa, Y. Okamura, Y. Taguchi, N. Nagaosa, and\nY. Tokura, Nat. Commun. 8, 1332 (2017).\n[71] W. M¨ unzer, A. Neubauer, T. Adams, S. M¨ uhlbauer, C. Fra nz, F. Jonietz, R. Georgii, P. B¨ oni,\nB. Pedersen, M. Schmidt, A. Rosch and C. Pfleiderer, Phys. Rev . B81, 041203(R) (2010).34\nFIG. 1. (a) Schematics of β-Mn-type chiral crystal structures as viewed along the [111]\ndirection. Two enantiomers with the space group P4332 (left-handed structure) and P4132\n(right-handed structure) are shown. Blue and red circles repres ent 8cand 12dWyckoff\nsites, respectively. The network of 12 dsites forms a hyper-kagome structure composed\nof corner-sharing triangles. (b) Temperature ( T) - Mn concentration ( x) phase diagram\nin (Co 0.5Zn0.5)20−xMnx(0≤x≤10) at zero field, determined by magnetization ( M)\nmeasurements. Closed symbols are data of polycrystalline samples r eproduced from our\nprevious work in Ref. [31]. Copyright 2018, American Association for the Advancement of\nScience. Open symbols are data of single-crystalline samples ( x= 0, 2, 4, 6) in the present\nstudy [see Fig. 3(a-d) for the detailed determination of the phase b oundaries]. A shaded\nregion with red indicates the temperature range TL≤T≤THwhereMat 20 Oe and the\nmagnitude of helical wavevector ( q) vary (decreases and increases respectively on cooling)\nsignificantly. (c-g) Schematic illustrations of various skyrmion lattic es: (c) triangular\nlattice, (d) rhombic lattice, (e) square lattice, (f) coexistence of square lattice of skyrmions\nand helical state, (g) I- or L-like deformed skyrmions on a square la ttice. Corresponding35\nhybridized qvectors are also presented in each panel.\nFIG. 2. Summary of equilibrium and metastable skyrmion states in tem perature ( T) -\nmagnetic field ( H) plane in (b) Co 10Zn10, (c) Co 9Zn9Mn2, (d) Co 8Zn8Mn4(reproduced from\nour previous work in Ref. [34]. Copyright 2016, Springer Nature) an d (e) Co 7Zn7Mn6. The\nmeasurement processes for the state diagrams are schematically illustrated in panel (a). The\nequilibrium skyrmion crystal (SkX) phase just below Tc(green area) and the equilibrium\ndisordered skyrmion (DSk) phase (orange area) are determined b y field-sweeping measure-36\nments of ac susceptibility and SANS after zero-field cooling (ZFC). T he metastable SkX\nstate [triangular lattice (light-blue area), rhombic lattice (purple ar ea) and square lattice\n(pink area)] are determined by field-sweeping measurements towa rd positive and negative\ndirections after the field cooling (FC) via the positive-field equilibrium S kX phase. Here,\nwe use the following notations; H: helical, C: conical, F: ferromagnet ic, E: equilibrium, M:\nmetastable, T: triangular, R: rhombic, S: square, SkX: skyrmion c rystal, DSk: disordered\nskyrmions, and RSG: reentrant spin glass. The phase boundaries a re determined by ac\nsusceptibility ( χ′). For the purpose of comparison with SANS measurements, calibra ted\nvalues (Hc) are used for the magnetic field. The details about the determinatio n of the\nphase boundaries of the equilibrium SkX state and the metastable Sk X state in Co 10Zn10\nare described in the captions of Fig. 4(d) and Supplementary Fig. S2 (c), respectively.\nThe boundaries between the metastable triangular SkX and the met astable rhombic SkX\n(purple diamonds) are determined by SANS results [see Fig. 6(d) and Supplementary Figs.\nS2(d) and S3(d) for the details]. The boundary between the metas table triangular SkX and\nthe metastable square SkX (red squares) are determined as inflec tion points of χ′(H) as\ndetailed in Ref. [34]. The reentrant spin glass transition temperatur es (yellow open circles)\nare determined as inflection points of a sharp drop in χ′(T), as also observed in M(T) in\nthe zero-field-cooled field-warming process [Fig. 3(c, d)]. The equilib rium DSk phase in\nCo7Zn7Mn6, determined as the region where a spherical SANS pattern is obser ved, was\nreproduced from our previous work in Ref. [31]. Copyright 2018, Am erican Association for\nthe Advancement of Science.37\nFIG. 3. Temperature dependence of magnetization and qvector in Co 10Zn10(red),\nCo9Zn9Mn2(green), Co 8Zn8Mn4(blue) and Co 7Zn7Mn6(purple). (a-d) Magnetization\n(M) under a magnetic field of 20 Oe is plotted against temperature. Solid lines show\nthe data collected during field cooling (FC), and broken lines represe nt those taken in a\nfield-warming run after a zero-field cooling down to 2 K (ZFC-FW). Th e helimagnetic\ntransition temperature ( Tc) indicated with closed triangles is determined as an inflection\npoint of a sharp increase in Mon the FC process. THandTLare the temperatures at\nwhich a gradual decrease in Mon cooling starts and ends, respectively. These temperatures\nare determined as inflection points with a broad peak in dM/dTon the FC process and\npresented with open triangles. The reentrant spin glass transition temperature ( Tg) denoted\nwith closed triangles is determined as an inflection point of a sharp incr ease inMon the\nZFC-FW process. These temperatures are also plotted in the T-xphase diagram in Fig.\n1(b). (e-h) Temperature variation of the magnitude of qin the helical state at zero field\n(open squares), and qin equilibrium and metastable SkX states under magnetic fields\n(closed circles). The qvalues are determined as the peak center of Gaussian function fitt ed\nto SANS intensity as a function of q. The inset of panel (e) shows an enlarged view of q(T)\nfor Co 10Zn10. The purple dashed line in the inset of (e) and the pink dashed line in (f- h)38\nindicate the expected qvalues for rhombic and square lattice of skyrmions (= 1.043 qTand\n1.075qT, whereqTare the values in the equilibrium triangular SkX state), respectively, with\nthe assumption of constant skyrmion density from the high-tempe rature triangular lattice\n(see discussion in Section III-F-2 for the details). (i-l) Temperatu re dependence of full\nwidth at half maximum (FWHM) of the Gaussian function fitted to SANS intensity versus q\nComposition Co10Zn10Co9Zn9Mn2Co8Zn8Mn4Co7Zn7Mn6\nTc(K) 414 396 299 158\nMs(µB/f.u.) 11.3 14.3 12.7 7.97\nλmax(nm) 156 132 110 112\nλmin(nm) 143 91 73 74\nPreferred qdirection <111> < 100> < 100> < 100>\nTABLE. 1. Summary of several physical parameters for helimagne tic state in Co 10Zn10,\nCo9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6. Helimagnetic transition temperature Tcis\ndetermined from the temperature variation in magnetization under 20 Oe. Saturation\nmagnetization Msis defined as a magnetization value at 2 K and 7 T. Helimagnetic\nperiodicity λ(maximum value λmaxat high temperature and minimum value λminat low\ntemperature) is calculated from the qvalue obtained from SANS measurements in zero\nfield. The preferred orientation of the helical qvector is also determined from SANS\nmeasurements.39\nFIG. 4. Identification of equilibrium SkX phase in Co 10Zn10. (a) Photos of single-crystalline\nbulk samples of Co 10Zn10used for SANS and ac susceptibility measurements. Crystal axes\nand the direction of applied magnetic field ( H) are also indicated. (b) SANS images at\nselected fields in the field increasing run at 410 K. The intensity scale o f the color plot is\nfixed for the 4 panels. (c) Field dependence of SANS intensity at 410 K. The intensity40\nwas integrated over the region close to <111>(red area in the inset) as detailed in Fig.\n5(c). (d) Field dependence of the real part ( χ′, red line) and the imaginary part ( χ′′, blue\nline) of the ac susceptibility in the field increasing run at 410 K. Since th e demagnetization\nfactor is different from that in the SANS measurement, the field valu es are calibrated as\nHc= 3.0×H. The phase boundary between the helical and conical states ( HH−C, black\ntriangle) is determined as the inflection point of χ′. The phase boundaries between the\nequilibrium SkX and conical states ( HESkX1(2), green triangles) are determined as the peak\npositions at both sides of the dip structure in χ′. The phase boundary between the conical\nand induced-ferromagnetic states ( HC−F, black triangle) is determined as the inflection\npoint ofχ′. The equilibrium SkX region is indicated with the light-green shading in pa nels\n(c) and (d). (e) Contour plot of χ′on theT-Hcplane. The region with χ′≤42 emu/mol\nis displayed with white color. (f) Contour plot of χ′′on theT-Hcplane. Phase boundaries\ndetermined by χ′[panel (d)] and similar data points at different temperatures are plo tted\nas open symbols in panels (e) and (f).41\nFIG. 5. Temperature dependence of helical state in Co 10Zn10in zero field. (a) Schematic\nSANS pattern on the (110) plane expected for a helical state with q∝bardbl<111>. The helical\nstate forms four domains with q∝bardbl[-111] (2 red spots), [1-11] (2 blue spots), [11-1] (out of\nthe plane) and [111] (out of the plane), respectively, resulting in 4 s pots on the (110) plane\natφ= 55◦, 125◦, 235◦, 305◦. Here, φis defined as the clockwise azimuthal angle from\nthe vertical direction. (b) SANS images observed at 410 K, 360 K, 3 00 K and 1.5 K. The\nintensity scale of the color plot varies between each panel. The [-111 ] and [1-11] directions\nare indicated with broken arrows. (c) Temperature dependence o f the SANS intensity\nintegrated over the azimuthal angle area at φ= 55◦, 125◦, 235◦, 305◦with the width of ∆ φ\n= 30◦(green area in the inset). (d) Radial |q|dependence of SANS intensity, integrated over\nthe azimuthal angle area at φ= 125◦, 305◦with the width of ∆ φ= 30◦(green area in the\ninset), atseveral temperatures. The datapointsarefittedtoa Gaussianfunction(solidline).42\nFIG. 6. Metastable SkX state in Co 10Zn10during a field cooling process at 0.03 T. (a)\nSchematic SANS pattern on the (110) plane, perpendicular to the fi eld, expected for\na rhombic SkX with q∝bardbl<111>. (b) SANS images observed at 412 K, 300 K, 100 K\nand 1.5 K during the FC process at 0.03 T. The intensity scale of the co lor plot varies\nbetween each panel. The magnetic field was applied at 412 K after ZFC . (c) Azimuthal\nangle (φ) dependence of SANS intensity at 412 K, 300 K and 1.5 K. The intensit y data\nfor 1.5 K is vertically shifted by 2 for clarity. The angles correspondin g to the [-111]\nand [1-11] directions ( φ= 55◦, 125◦, 235◦and 305◦) are indicated by dashed lines. (d)\nTemperature dependence of the SANS intensity. Blue symbols deno te the intensity\nintegrated over the region close to the [-111] and [1-11] directions (φ= 55◦, 125◦, 235◦\nand 305◦with the width of ∆ φ= 30◦; blue area in the inset). The intensity integrated\nover the region around the [001] and [-110] directions ( φ= 0◦, 90◦, 180◦and 270◦with\nthe width of ∆ φ= 30◦; red area in the inset) are denoted by red symbols. The tempera-\nture regions of the equilibrium triangular SkX, metastable triangular SkX and metastable\nrhombic SkX are indicated with the light-green, light-blue and purple s hadings, respectively.43\nFIG. 7. Reorientation of metastable SkX at low temperatures in Co 8Zn8Mn4under an\noff-axis field. (a) Schematic top view of the experimental configura tion. Rocking angle ( ω)\nis defined as the angle between the incident neutron beam ( ki) and the applied magnetic\nfield (H). In this experiment, magnetic field was tilted by 15◦away from the [001] direction.\nDue to the demagnetization effect from the tilted magnetic field applie d to the rectangular-\nshaped sample (gray object), the effective magnetic field ( Heff, pink arrow) inside the sample\nis tilted by more than 15◦from [001] direction. The additional tilting angle ( α) ofHefffrom\nHis estimated to be α∼20◦according to both the calculation of the demagnetization\nfield[60] and the peak position of rocking curve at 295 K shown in pane l (b). (b) Rocking\ncurves at selected temperatures in the field cooling (FC) process. SANS intensities are44\nintegrated over the azimuthal angle area at φ= 90◦, 270◦with the width of ∆ φ= 90◦(red\narea in the inset). (c, d) SANS images at 295 K, 200 K, 120 K and 40 K d uring a FC process\nat 0.03 T. These SANS images are averaged over the limited rocking an gles: (c) 14◦≤ω≤\n20◦(aroundHeff) and (d) −20◦≤ω≤ −10◦(around [001]). The intensity scale of the color\nplot varies between each panel. (e) Temperature dependence of t he SANS intensities during\nthe FC process. Blue closed circles show the SANS intensity integrat ed over the azimuthal\nangle area at φ= 60◦, 120◦, 240◦, 300◦with the width of ∆ φ= 30◦in the rocking angle\nrange of 14◦≤ω≤20◦(blue area in the inset). Red open squares represent the SANS\nintensity integrated over the azimuthal angle area at φ= 90◦, 180◦with the width of ∆ φ=\n30◦in the rocking angle range of −20◦≤ω≤ −10◦(red area in the inset). The temperature\nregions of the equilibrium triangular SkX, metastable triangular SkX a nd metastable\nsquare SkX are indicated with the light-green, light-blue and pink sha dings, respectively.\n(f) Schematic illustration of transformation from a triangular SkX t o a square SkX\nunder the off-axis effective field Hefftilted from [001]. While the skyrmions in the triangu-\nlar SkX areparallel to Heff, those inthe square SkX at low temperatures areparallel to [001].45\nFIG. 8. Metastable SkX state in Co 9Zn9Mn2. (a) Schematic SANS pattern on the (110)\nplane, perpendicular to the field, expected for a triangular SkX sta te (right panel) and\nfor a square SkX state (left panel). The triangular SkX state with o ne of triple- q∝bardbl[001]\nshows 6 spots at φ= 0◦, 60◦, 120◦, 180◦, 240◦, 300◦. Here,φis defined as the clockwise\nazimuthal angle from the vertical [001] direction. The square SkX s tate with double- q∝bardbl\n[001] and ∝bardbl[100] or [010] (out of the plane) exhibits 2 spots at φ= 0◦, 180◦. (b) SANS\nimages observed at 390 K, 100 K, 50 K and 10 K in the field cooling (FC) p rocess at 0.04\nT. The intensity scale of the color plot varies between each panel. Th e magnetic field was46\napplied at 390 K after zero field cooling. (c) Rocking curves at select ed temperatures in the\nFC process. SANS intensities are integrated over the azimuthal an gle area at φ= 90◦, 270◦\nwith the width of ∆ φ= 90◦(red area in the inset). The origin of the rocking angle ( ω= 0◦)\ncorresponds to ki∝bardblH∝bardbl[110]. (d) Temperature dependence of the SANS intensities. Blue\nclosed circles show the SANS intensity integrated over the azimutha l angle area at φ= 60◦,\n120◦, 240◦, 300◦with the width of ∆ φ= 30◦(blue area in the inset) and finally divided by\n2. Red open squares represent the SANS intensity integrated ove r the azimuthal angle area\natφ= 0◦, 180◦with the width of ∆ φ= 30◦(red area in the inset). The temperature regions\nof the equilibrium triangular SkX, metastable triangular SkX and meta stable square SkX\nare indicated with the light-green, light-blue and pink shadings, resp ectively. (e) Radial |q|\ndependence of SANS intensity, integrated over the azimuthal ang le area at φ= 0◦, 180◦\nwith the width of ∆ φ= 30◦(red area in the inset), at several temperatures during the FC\nprocess. The data points are fitted to a Gaussian function (solid line ).47\nFIG. 9. Temperature dependence of metastable SkX state in Co 7Zn7Mn6. (a) Schematic\nSANS pattern on the (001) plane, perpendicular to the field, expec ted for a triangular\nSkX state (right panel) and for a square SkX state (left panel). Th e triangular SkX state\nforms two domains with one of the triple- q∝bardbl[100] (6 blue spots) and ∝bardbl[010] (6 red spots),\nrespectively, resulting in 12 spots at every 30◦fromφ= 0◦. Here, φis defined as the\nclockwise azimuthal angle from the vertical [010] direction. The squ are SkX state with\ndouble-q∝bardbl[100] and [010] shows 4 spots at φ= 0◦, 90◦, 180◦, 270◦. (b) SANS images\nobserved at 146 K, 100 K, 60 K and 1.5 K during the field cooling (FC) pr ocess at 0.025\nT. The intensity scale of the color plot varies between each panel. Th e magnetic field\nwas applied at 146 K after zero field cooling. (c) Temperature depen dence of the SANS\nintensities. Blue closed circles show the SANS intensity integrated ov er the azimuthal\nangle area at φ= 45◦, 135◦, 225◦, 315◦with the width of ∆ φ= 30◦(blue area in the\ninset). Red open squares represent the SANS intensity integrate d over the azimuthal angle\narea atφ= 0◦, 90◦, 180◦, 270◦with the width of ∆ φ= 30◦(red area in the inset). The\ntemperature regions of the equilibrium triangular SkX, metastable t riangular SkX and48\nmetastable square SkX are indicated with the light-green, light-blue and pink shadings,\nrespectively. (d) Radial |q|dependence of SANS intensity, integrated over the azimuthal\nangle area at φ= 0◦, 90◦, 180◦, 270◦with the width of ∆ φ= 30◦(red area in the inset), at\nseveral temperatures in the FC process. The data points are fitt ed to a Gaussian function\n(solid line).49\nFIG. 10. Field effect on metastable SkX state and helical state in Co 7Zn7Mn6at 100 K. (a)\nSANS images at selected fields during field scans to the positive and ne gative directions at\n100 K after a field cooling (FC) at 0.025 T. (b) SANS images observed a t 0 T, 0.05 T and\n0.1 T at 100 K after a zero-field cooling (ZFC) (reproduced from our previous work in Ref.\n[31]. Copyright 2018, American Association for the Advancement of Science). The intensity\nscale of the color plot is fixed for all the panels in (a) and (b). (c) Field dependence of the\nSANS intensities at 100 K after the FC defined similarly as in Fig. 9(c). T he field regions\nof the metastable triangular SkX and the metastable square SkX ar e indicated with the\nlight-blue and pink shadings, respectively. (d) Field dependence of t he SANS intensities at\n100 K after the ZFC (reproduced from our previous work in Ref. [31 ]. Copyright 2018,\nAmerican Association for the Advancement of Science). Green clos ed circles show the\nSANS intensity integrated over the azimuthal angle area at φ= 45◦, 135◦, 225◦, 315◦with\nthe width of ∆ φ= 30◦(green area in the inset). Orange open squares represent the SA NS\nintensity integrated over the azimuthal angle area at φ= 0◦, 90◦, 180◦, 270◦with the width\nof ∆φ= 30◦(orange area in the inset).50\nFIG. 11. Metastable and low-temperature equilibrium skyrmion stat es in magnetic fields in\nCo7Zn7Mn6at 60 K. (a) SANS images at selected fields during field scans to the po sitive\nand negative directions at 60 K after a FC at 0.025 T. (b) SANS images observed at 0 T,\n0.07 T and 0.1 T at 60 K after a ZFC. The intensity scale of the color plot is fixed for all\nthe panels in (a) and (b). (c) Field dependence of the SANS intensitie s at 60 K after the\nFC, integrated over the azimuthal angle areas defined similarly as in F ig. 9(c). The field\nregions of the metastable square SkX and the equilibrium disordered skyrmions (DSk) are\nindicated with the pink and orange shadings, respectively. (d) Field d ependence of the\nSANS intensities at 60 K after the ZFC (reproduced from our previo us work in Ref. [31].\nCopyright 2018, American Association for the Advancement of Scie nce), integrated over the\nazimuthal angle areas defined similarly as in Fig. 10(d). The field region of the equilibrium\nDSk phase is indicated with the orange shading.51\nFIG.12. (a-c)Temperature ( T)-magnetic field( H)phase diagramsnear Tcin(a)Co 10Zn10,\n(b) Co 9Zn9Mn2and (c) Co 8Zn8Mn4, determined by ac susceptibility ( χ′) measurements\nwithH∝bardbl[110]. For the magnetic field, calibrated values Hcare used. The displayed range of\nT/Tc(upper horizontal axis) is fixed to be 0.90 - 1.01 for the three panels . Time-dependent\nχ′measurements after a field cooling (FC, pink arrow) via the equilibrium SkX phase (green\nregion) are performed at several temperatures denoted with cir cles, whose color corresponds52\nto the color of data points in panels (d-f). (d-f) Time dependence o f the normalized ac\nsusceptibility, defined as χ′\nN(t)≡[χ′(∞)−χ′(t)]/[χ′(∞)−χ′(0)], for (d) Co 10Zn10, (e)\nCo9Zn9Mn2(reproduced from our previous work in Ref. [36] with permissions. C opyright\n2017, American Physical Society) and (f) Co 8Zn8Mn4, measured in the processes described\nin panels (a-c). Here, χ′(0) is an initial value (metastable SkX state), and χ′(∞) is the value\nfor the equilibrium conical state which, as a fully relaxed state is assu med to be the value\nofχ′at the same magnitude of field after a field decreasing run from a fer romagnetic phase.\nThe data points are fitted to stretched exponential functions (d otted lines), exp/braceleftbig\n−(t/τ)β/bracerightbig\n.53\nFIG. 13. (a) Relaxation time ( τ) of metastable SkX states plotted against normalized\ntemperature T/Tcin Co 10Zn10(red squares), Co 9Zn9Mn2(green triangles, reproduced\nfrom our previous work in Ref. [36] with permissions. Copyright 2017 , American Physical\nSociety) and Co 8Zn8Mn4(blue circles). The τvalues are determined by the fits in\nFigs. 12(d-f). The solid arrows and dotted lines indicate temperatu re ranges and lower\nboundaries of equilibrium SkX phases, respectively. The data points from the three different\ncompositions, showing a good scaling, are fitted to a modified Arrhen ius law (pink solid\nline),τ=τ0exp{a(Tc−T)/T}, with an assumption of a temperature-dependent activation\nenergyEg=a(Tc−T) as described in Ref. [36, 64, 65]. The obtained parameters are a=\n215 and τ0= 63 s. (b) Schematic illustration of free energy landscape with a met astable\nSkX state and the most stable conical state. (c) Schematic illustra tion of the destruction\nprocess for a skyrmion string induced by movement of an emergent monopole-antimonopole\npair for a clean system and a dirty system. In the dirty system, the propagation of the\nmonopole and antimonopole are hindered by magnetic disorder while les s so in the clean54\nsystem.\nMaterial a τ 0(s) Reference\nCo-Zn-Mn 215 63 This work\nMnSi 65 2.7 ×10−4[64]\nCu2OSeO3 96 3 [65]\nCu2OSeO3(Zn 2.5% doped) 94 150 [65]\nTABLE. 2. Summary of relaxation parameters for metastable SkX in various materials.arXiv:2008.05075v1 [cond-mat.str-el] 12 Aug 2020Supplementary materials for\n“Metastable skyrmion lattices governed by magnetic disord er and\nanisotropy in β-Mn-type chiral magnets”\nK. Karube,1,∗J. S. White,2,∗V. Ukleev,2C. D. Dewhurst,3R. Cubitt,3\nA. Kikkawa,1Y. Tokunaga,4H. M. Rønnow,5Y. Tokura,1,6and Y. Taguchi1\n1RIKEN Center for Emergent Matter Science (CEMS), Wako 351-0198 , Japan.\n2Laboratory for Neutron Scattering and Imaging (LNS),\nPaul Scherrer Institute (PSI), CH-5232 Villigen, Switzerland .\n3Institut Laue-Langevin (ILL), 71 avenue des Martyrs,\nCS 20156, 38042 Grenoble cedex 9, France\n4Department of Advanced Materials Science,\nUniversity of Tokyo, Kashiwa 277-8561, Japan.\n5Laboratory for Quantum Magnetism (LQM), Institute of Physics ,\n´Ecole Polytechnique F´ ed´ erale de Lausanne (EPFL), CH-1015 L ausanne, Switzerland.\n6Department of Applied Physics and Tokyo College,\nUniversity of Tokyo, Bunkyo-ku 113-8656, Japan.\n∗These authors equally contributed to this work2\nA. Magnetization curve at 2 K\nIn Fig. S1(a), we show field dependence of magnetization at 2 K meas ured up to 7\nT in Co 10Zn10, Co9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6. The saturation magnetization\n(Ms), which is defined as the value at 2 K and 7 T and summarized in Table 1 in t he\nmain text, takes a maximum at Co 9Zn9Mn2(Ms= 14.3µB/f.u.), while the helimagnetic\ntransition temperature Tcmonotonically decreases from Co 10Zn10as the Mn concentration\nis increased. This indicates that Co spin and Mn spins are ferromagne tically coupled at\nleast in the low Mn concentration region. As seen in the magnified view a round the low-field\nregion [Fig. S1(b)], hysteresis is observed for −0.3 T≤µ0H≤0.3 T for Co 8Zn8Mn4and for\n−0.7 T≤µ0H≤0.7 T for Co 7Zn7Mn6. The hysteresis is observed only below the reentrant\nspin glass transition temperature Tgin Co8Zn8Mn4(Tg∼10 K) and Co 7Zn7Mn6(Tg∼30\nK). The hysteresis is not discernible in Co 10Zn10and Co 9Zn9Mn2, neither of which exhibits\nthe reentrant spin glass transition down to 2 K. Similar results of mag netization have been\nreported in Ref. [S1].\nB. Field dependence of metastable SkX at 300 K in Co 10Zn10\nHere, we show the detailed results of field variation in metastable sky rmion crystal (SkX)\nat 300 K in Co 10Zn10(Fig. S2).\nSmall-angle neutron scattering (SANS) patterns for several mag netic fields at 300 K\nafter a field cooling (FC) with 0.03 T are shown in Fig. S2(a). In the field sweeping process\ntoward positive direction (pink arrows), the SANS pattern with str onger intensity along the\n[-111] and [1-11] direction changes to a uniform ring at 0.1 T similar to t hat observed in the\nequilibrium SkX phase at 412 K at 0.03 T. This field variation correspond s to the change in\nlattice form of skyrmions from rhombic-like one to the original triang ular one. The ring-like\nSANS signal corresponding to the triangular SkX persists up to 0.2 T . In the field sweeping\nrun toward negative direction (light-blue arrows) down to −0.06 T, the intensity along the\n[-111] and [1-11] directions increases and clear 4 spots are observ ed, indicating that lattice\nform of skyrmions approaches ideal rhombic one. Upon further inc reasing the field in the\nnegative direction, the SANS intensity from the rhombic SkX disappe ars at−0.1 T. For\ncomparison, SANS patterns at the same magnetic fields at 300 K aft er a zero-field cooling3\n(ZFC) are presented in Fig. S2(b). The SANS pattern with broad int ensity around the\n[1-11] direction originating from a helical state disappears at 0.1 T, w hich is totally distinct\nfrom the field variation in the metastable SkX after the FC.\nThe real-part of ac susceptibility ( χ′) and the SANS intensities at 300 K after the FC\nare plotted against applied field in Fig. S2(c) and S2(d), respectively . Here, ZFC data are\nalso plotted in the same graph for comparison. The small values of χ′and the large SANS\nintensities observed for −0.1 T≤µ0H≤0.2 T in the initial field scans, as compared to those\nin the returning runs, are attributed to the metastable SkX state . Thus, the metastable SkX\nstate survives over a wide field region up to the conical-ferromagne tic phase boundary and\ndown to the negative fields as marked with blue triangles in Fig. S2(c). The boundary\nbetween the rhombic SkX and the triangular SkX is determined as the magnetic field where\nthe SANS intensity around the [-111] and [1-11] directions becomes higher than that around\nthe [001] and [-110] directions [purple triangle in Fig. S2(d)], and plott ed in the state\ndiagram of Fig. 2(b) in the main text.\nC. Field dependence of metastable SkX at 1.5 K in Co 10Zn10\nWe also discuss field variation in metastable SkX in Co 10Zn10at 1.5 K (Fig. S3).\nFigure S3(a) shows several SANS patterns at 1.5 K for selected fie lds after the FC with\n0.03 T. In the field sweeping toward positive direction, the broad 4-s pot pattern originating\nfrom the rhombic lattice of skyrmions persists up to 0.10 T and finally t ransforms to a ring-\nlike pattern with strong intensity at 0.16 T. In the field sweeping proc ess toward negative\ndirection, the 4-spot pattern is preserved down to −0.16 T before Bragg spots disappear at\n−0.20 T. This result demonstrates that the metastable rhombic SkX s urvives over a wider\nfield region than that observed at 300 K but finally original triangular lattice is restored at\nhigh magnetic fields. For comparison we also show field variation in the S ANS pattern at 1.5\nK after ZFC in Fig. S3(b). The 4 spots along the [-111] and [1-11] dire ction, corresponding\nto the helical multi-domain state with q∝bardbl<111>, are observed up to 0.16 T with gradual\ndecreases inscattering intensity and qposition. However, aring-likepattenisnever observed\nat any fields, which is distinct from the result after the FC.\nWe show the field dependence of χ′at 5 K and SANS intensities at 1.5 K after the FC in\nFig. S3(c) and S3(d), respectively. The both measurements exhib it large and asymmetric4\nhysteresis, corresponding to the metastable SkX state as indicat ed with blue triangles. A\nboundary of rhombic-triangular skyrmion lattices [purple triangle in F ig. S3(d)] is deter-\nmined similarly to the previous section. The field dependence of χ′is complex as compared\nwith the result at 300 K [Fig. S2(c)], and the hysteresis subsists to −0.3 T in the negative\nfield region, while the SANS intensity originating from the metastable S kX state disappears\nat−0.2 T. These results indicate that a helical state with qvector along the [11-1] or [111]\ndirection, which is out of the (110) plane and thus not detectable in t he present SANS\nconfiguration, persists up to the field-induced ferromagnetic pha se boundary after complete\ndestruction of the metastable SkX state around −0.2 T. For comparison, the field depen-\ndence of χ′at 5 K and SANS intensities at 1.5 K after ZFC are also displayed in the sa me\nfigures. In this case, a hysteresis in χ′corresponding to the helical multi-domain state is\nobserved between −0.3 T and 0.3 T, which is wider than the hysteresis region (between −0.2\nT and 0.2 T) in the SANS intensity that represents the contribution f rom helical states with\nq∝bardbl[-111] and [1-11]. These ZFC results also indicate that a helical state withq∝bardbl[11-1]\nand [111] remains above 0.2 T. The complex field dependence is attribu ted to the strong\nmagnetocrystalline anisotropy that favors q∝bardbl<111>. As described in the next section, the\nhelical state forms a chiral soliton lattice under magnetic fields due t o the strong anisotropy\nofqvector. In the field returning process from high fields both after t he FC and ZFC, a\nclear dip structure in χ′and relatively large SANS intensity are observed between −0.1 T\nand 0.1 T. This is probably because helical qvector flops from the field direction to within\nthe plane perpendicular to the field.\nD. Chiral soliton lattice in Co 10Zn10\nIn this section, we describe possible formation of chiral soliton lattic e in Co 10Zn10[Fig.\nS4]. As presented in Fig. S3(b, d), the 4-spot SANS pattern, which corresponds to the\n2-domain helical state with q∝bardbl[-111] and [1-11], persists up to high fields near the ferro-\nmagneticregion. Inthisfieldvariation, weakhigher-harmonicscatt eringsignalsareobserved\nalong the [1-11] direction away from the strong main spots, which is m ore clearly discerned\nin the SANS image at 0.02 T plotted in logarithmic scale in Fig. S4(a). For q uantitative\ndiscussion, the radial qdependence of SANS intensity along the [1-11] direction is plotted\nin Fig. S4(b). At zero field, the Bragg peak is observed at q0∼0.044 nm−1. As the field5\nis increased to 0.02 T, scattering intensity above 0.07 nm−1is enhanced, and the second,\nthird, and even fourth harmonic peaks ( q2,q3andq4) are clearly observed in addition to\nthe first main peak ( q1). The higher harmonic peaks are still discerned at 0.16 T while the\npeak positions shift to lower qregion. These peaks are fitted to Gaussian functions and the\npeak center of qn(n= 1, 2, 3) normalized by nq0is plotted against field in Fig. S4(c). The\nsecond and third harmonic peak positions show a good scaling with the first peak position\nas expected as qn=nq1. With increasing field, the value of qn/nq0decreases down to 0.8\nbefore SANS signals disappear at 0.24 T.\nThese results are reminiscent of a chiral soliton lattice (CSL): one- dimensional chain\nof nonlinear helimagnetic structures as schematically illustrated in Fig . S4(d), which has\nbeen observed in monoaxial chiral magnets as represented by CrN b3S6[S2]. Following the\ntheoretical analysis by Togawa et al.[S2], field variation in qvalue for the CSL is expressed\nasq(H)/q(0) =π2/[4K(κ)E(κ)], where K(κ) andE(κ) are the elliptic integrals of the\nfirst and second kinds. Here, the elliptic modulus κ(0< κ <1) is determined by energy\nminimization and given by κ/E(κ) =/radicalbig\nH/Hc, whereHcis the ferromagnetic saturation\nfield. This theoretical result for the CSL is plotted with pink dashed lin e in Fig. S4(c).\nHere, we set µ0Hc= 0.24 T, above which the SANS intensity disappears, and found that\nthe observed field variation in qn/nq0agrees with the theoretical curve.\nA CSL is stabilized under magnetic fields rather than a conventional c onical state if q\nvector is strongly anisotropic and thus perpendicular to the fields. The CSL has been also\nreported in a cubic chiral magnet Cu 2OSeO3under a uniaxial strain which fixes the qvector\nalong the strain direction[S3]. In the present study for a cubic chira l magnet Co 10Zn10\nwithout uniaxial strain, the CSL behavior is observed probably beca useqvector anisotropy\nalong the <111>direction becomes strong at low temperatures, to which the magne tic field\nis applied perpendicularly.\nE. Transformation of metastable skyrmion lattice in Co 8Zn8Mn4\nHere, we review the metastable SkX state induced by FC and struct ural transition of\nthe skyrmion lattice within the metastable state in Co 8Zn8Mn4, which was reported in our\nprevious paper[S4] (Fig. S5). Selected SANS patterns in a FC at 0.04 T from 295 K to 40 K\nare shown in Fig. S5(b). Here, the magnetic field and the incident neu tron beam are parallel6\nto the [001] direction. At 295 K and 0.04 T, inside the equilibrium SkX pha se, clear 12 spots\nareobserved. This 12-spotpattern corresponds to a 2-domaint riangular SkX state, in which\none of triple- qis parallel to the [010] or [100] direction as illustrated in the right pane l of Fig.\nS5(a). The triangular SkXpersists down to200K asametastable st ateduring theFC. From\n120 K to 40 K, the 12-spot pattern gradually transforms to a 4-sp ot pattern. This 4-spot\npattern corresponds to a square SkX state with double- qvectors parallel to the [010] and\n[100] directions as illustrated in the left panel of Fig. S5(a). In the s ubsequent re-warming\nprocess, the triangular SkX revives already at 200 K. This reversib ility of the metastable\nSkX can rule out the possibility that the 4-spot pattern is due to a he lical multi-domain\nstate with zero topological charge.\nSANS intensity is plotted as a function of temperature in Fig. S5(c). Above 120 K,\nthe intensities for the <100>and<110>directions show similar values as expected for a\n12-spot pattern. Below 120 K, the intensity ratio of <100>to<110>becomes larger due\nto the gradual transformation to 4 spots. The SANS intensities sh ow maximum at 250 K\nand decrease below 200 K, indicating evolution of magnetic disorder in the metastable SkX,\nespecially in the square SkX.\nThe radial qdependence of the SANS intensity for <100>is shown in Fig. S5(d). From\n120 K to 40 K, where the triangular-square SkX transition occurs, the peak center shows a\nlarge shift from 0.055 nm−1to 0.09 nm−1and the peak intensity decreases. The temperature\ndependence of qin the metastable skyrmion state during the field cooling is similar to tha t\nof the helical state during zero-field cooling as shown in Fig. 3(g) in th e main text.\nF. Field dependence of metastable SkX in Co 9Zn9Mn2\nInthissection, wedescribefieldvariationinthemetastableSkXstat eatlowtemperatures\nin Co9Zn9Mn2(Fig. S6). Selected SANS patterns at 10 K after the FC (0.04 T) are shown\nin Fig. S6(a). The 2-spot pattern at 0.04 T corresponds to the squ are SkX as discussed\nin Fig. 8 in the main text. With increasing field up to 0.3 T, the 2-spot pat tern gradually\nchanges to a ring-like one. The detailed field dependence of the SANS intensity is presented\nin Fig. S6(c). While the large intensity from the 2 vertical spots is rap idly suppressed with\nincreasing field, the weak intensity from the side region hardly decre ases, and eventually the\nintensities from the two regions become similar at 0.3 T.7\nAs shown in Fig. S6(b), the rocking curve at 0.3 T clearly shows a peak structure around\nω= 0◦. This indicates that the scattering intensity from the ring-like patt ern lies in the\n(110) plane, namely qvector is perpendicular to the field, similar to the observation at high\ntemperatures. Thus, the field variation in the SANS pattern is inter preted as the change\nfrom the square SkX with q∝bardbl<100>to an orientationally disordered triangular SkX with\nq⊥H.8\n[S1] J. D. Bocarsly, C. Heikes, C. M. Brown, S. D. Wilson, and R . Seshadri, Phys. Rev. Mater. 3,\n014402 (2019).\n[S2] Y. Togawa, T. Koyama, K. Takayanagi, S. Mori, Y. Kousaka , J. Akimitsu, S. Nishihara, K.\nInoue, A. S. Ovchinnikov, and J. Kishine, Phys. Rev. Lett. 108, 107202 (2012).\n[S3] Y. Okamura, Y. Yamasaki,2 D. Morikawa, T. Honda, V. Ukle ev, H. Nakao, Y. Murakami, K.\nShibata, F. Kagawa, S. Seki, T. Arima, and Y. Tokura, Phys. Re v. B96, 174417 (2017).\n[S4] K. Karube, J. S. White, N. Reynolds, J. L. Gavilano, H. Oi ke, A. Kikkawa, F. Kagawa, Y.\nTokunaga, H. M. Rønnow, Y. Tokura, and Y. Taguchi, Nat. Mater .15, 1237 (2016).9\nFIG.S1. (a)Magneticfield( H)dependenceofmagnetization( M)at2Kinsingle-crystalline\nsamples of Co 10Zn10, Co9Zn9Mn2, Co8Zn8Mn4and Co 7Zn7Mn6. Magnetic fields are applied\nalong the [110] direction for Co 10Zn10and Co 9Zn9Mn2, and along the [100] direction for\nCo8Zn8Mn4andCo 7Zn7Mn6, respectively. (b)Magnifiedviewoflowfieldregioninpanel (a).10\nFIG. S2. Field dependence of metastable SkX state and helical stat e in Co 10Zn10at 300 K.\n(a) SANS images observed at selected fields during field scans to the positive and negative\ndirections at 300 K after a field cooling (FC) at 0.03 T. The intensity sc ale of the color plot\nis fixed through the panels. (b) SANS images observed at 0 T, 0.06 T a nd 0.1 T at 300\nK after a zero-field cooling (ZFC). The intensity scale of the color plo t is fixed for these\npanels. For panels (a) and (b), the [-111] and [1-11] directions are presented with broken\narrows. (c) Field dependence of the real part of the ac susceptib ility (χ′) at 300 K after FC\nwithµ0Hc= 0.03 T (red line) and ZFC (green line). The blue triangles ( HMSkX+( −)) denote\nthe boundaries of metastable SkX that are plotted in the state diag ram in Fig. 2(b) in the\nmain text. The boundary at the positive field side is determined as the field where the\nhysteresis in χ′vanishes, and the boundary at the negative field side is determined a s the\ninflection point. (d) Field dependence of the SANS intensities at 300 K after the FC at 0.03\nT and ZFC. For the FC data, SANS intensity integrated over the azim uthal angle areas at\nφ= 55◦, 125◦, 235◦, 305◦with the width of ∆ φ= 30◦(blue area in the inset) and at φ=\n0◦, 90◦, 180◦, 270◦with the width of ∆ φ= 30◦(red area in the inset) are presented by blue\nclosed circles and red open squares, respectively. The purple trian gle shows the boundary11\nbetween the metastable rhombic SkX and the triangular one and is de termined as the field\nwhere the former intensity (blue circle) becomes higher than the lat er (red square). For the\nZFC data, SANS intensity integrated over the region at φ= 55◦, 125◦, 235◦, 305◦with the\nwidth of ∆ φ= 30◦(green area in the inset) is indicated with green closed triangles. Her e,\nφis defined as the clockwise azimuthal angle from the vertical directio n. In panels (c) and\n(d), the regions of the metastable rhombic SkX and the triangular o ne are indicated with\npurple and light-blue shadings, respectively.12\nFIG. S3. Field dependence of metastable SkX state and helical stat e in Co 10Zn10at 1.5 K.\n(a) SANS images observed at selected fields during field scans to the positive and negative\ndirections at 1.5 K after a FC (0.03 T). The intensity scale of the color plot is fixed through\nthe panels. (b) SANS images observed at 0 T, 0.1 T and 0.16 T at 1.5 K af ter a ZFC.\nThe intensity scale of the color plot is fixed for these panels. (c) Field dependence of χ′\nat 5 K after the FC with µ0Hc= 0.03 T (red line) and ZFC (green line). The boundaries\nof metastable SkX ( HMSkX+( −), blue triangles) are determined similarly as in Fig. S2(c)\n(d) Field dependence of the SANS intensities at 1.5 K after the FC (0.0 3 T) and ZFC,\nintegrated over the azimuthal angle areas defined similarly as in Fig. S 2(d). The boundary\nbetween the rhombic SkX and the triangular one (purple triangle) is d etermined similarly\nas in Fig. S2(d). In panels (c) and (d), the regions of the metastab le rhombic SkX and the\ntriangular one are indicated with purple and light-blue shadings, resp ectively.13\nFIG. S4. Chiral soliton lattice formed under magnetic fields at 1.5 K in C o10Zn10. (a)\nSANS image measured at 1.5 K and 0.02 T after ZFC. The intensity of th e color plot is\ndisplayed in a logarithmic scale. (b) Radial |q|dependence of SANS intensities (logarithmic\nscale) measured at 0 T (black), 0.02 T (red) and 0.16 T (blue) at 1.5 K a fter ZFC. The\nintensities are integrated over the azimuthal angle area at φ= 125◦, 305◦with the width of\n∆φ= 30◦[white frames in panel (a)]. The black arrow ( q0) indicates the peak position at\n0 T. The red arrows ( q1,q2,q3andq4) for 0.02 T indicate the main peak position, second,\nthird and fourth harmonic peak positions, respectively. (c) Norma lizedqvalues,q1/q0,\nq2/2q0andq3/3q0, are plotted as a function of magnetic field. Dashed pink line indicates\na theoretical curve for the chiral soliton lattice (see text for det ails). (d) Schematic spin\nconfiguration of chiral soliton lattice.14\nFIG. S5. Temperature dependence of metastable SkX state in Co 8Zn8Mn4. (a) Schematic\nSANS pattern on the (001) plane, perpendicular to the field, expec ted for a triangular\nSkX state (right panel) and for a square SkX state (left panel). Th e triangular SkX state\nforms two domains with one of triple- q∝bardbl[100] (6 blue spots) and ∝bardbl[010] (6 red spots),\nrespectively, resulting in 12 spots at every 30◦fromφ= 0◦. Here, φis defined as the\nclockwise azimuthal angle from the vertical [010] direction. The squ are SkX state with\ndouble-q∝bardbl[100] and [010] shows 4 spots at φ= 0◦, 90◦, 180◦, 270◦. (b) SANS images\nobserved at 295 K, 200 K, 120 K and 40 K on the field cooling (FC) proc ess at 0.04 T\n(reproduced from our previous work in Ref. [S4]. Copyright 2016, S pringer Nature). The\nintensity scale of the color plot varies between each panel. The magn etic field was applied\nat 295 K after zero-field cooling. (c) Temperature dependence of the SANS intensities. Blue\nclosed circles show the SANS intensity integrated over the azimutha l angle area at φ= 45◦,\n135◦, 225◦, 315◦with the width of ∆ φ= 30◦(blue area in the inset). Red open squares15\nrepresent the SANS intensity integrated over the azimuthal angle area atφ= 0◦, 90◦, 180◦,\n270◦with the width of ∆ φ= 30◦(red area in the inset). The temperature regions of the\nequilibrium triangular SkX, the metastable triangular SkX and the met astable square SkX\nare indicated with the light-green, light-blue and pink shadings, resp ectively. (d) Radial\n|q|dependence of SANS intensity, integrated over the azimuthal ang le area at φ= 0◦, 90◦,\n180◦, 270◦with the width of ∆ φ= 30◦(red area in the inset), at several temperatures on\nthe FC process. The data points are fitted to a Gaussian function ( solid line).16\nFIG. S6. Field dependence of metastable SkX state in Co 9Zn9Mn2at 10 K. (a) SANS\nimages at 0.04 T, 0.2 T and 0.3 T measured at 10 K after a field cooling (FC ) at 0.04\nT. The intensity scale of the color plot varies between each panel. (b ) Rocking curves at\nseveral fields in the field-increasing process after the FC. SANS int ensities are integrated\nover the azimuthal angle area at φ= 90◦, 270◦with the width of ∆ φ= 90◦(red area in the\ninset). Here, φis defined as the clockwise azimuthal angle from the vertical [001] dir ection.\n(c) Field dependence of the SANS intensities. Blue closed circles show the SANS intensity\nintegrated over the azimuthal angle area at φ= 60◦, 120◦, 240◦, 300◦with the width of ∆ φ\n= 30◦(blue area in the inset) and finally divided by 2. Red open squares repr esent the\nSANS intensity integrated over the azimuthal angle area at φ= 0◦, 180◦with the width of\n∆φ= 30◦(red area in the inset). The field regions of the metastable triangula r SkX and the\nmetastable square SkX are indicated with the light-blue and pink shad ings, respectively." }, { "title": "2008.05125v1.Prediction_on_Properties_of_Rare_earth_2_17_X_Magnets_Ce2Fe17_xCoxCN___A_Combined_Machine_learning_and_Ab_initio_Study.pdf", "content": "arXiv:2008.05125v1 [cond-mat.mtrl-sci] 12 Aug 2020Prediction on Properties of Rare-earth 2-17-X Magnets Ce 2Fe17−xCoxCN : A Combined\nMachine-learning and Ab-initio Study\nAnita Halder,1Samir Rom,1Aishwaryo Ghosh,2and Tanusri Saha-Dasgupta1,∗\n1Department of Condensed Matter Physics and Material Scienc es,\nS. N. Bose National Centre for Basic Sciences, JD Block,\nSector III, Salt Lake, Kolkata, West Bengal 700106, India.\n2Department of Physics, Presidency University, Kolkata 700 073, India\n(Dated: August 13, 2020)\nWe employ a combination of machine learning and first-princi ples calculations to predict magnetic properties\nof rare-earth lean magnets. For this purpose, based on train ing set constructed out of experimental data, the\nmachine is trained to make predictions on magnetic transiti on temperature (T c), largeness of saturation magne-\ntization (µ0Ms), and nature of the magnetocrystalline anisotropy ( Ku). Subsequently, the quantitative values of\nµ0MsandKuof the yet-to-be synthesized compounds, screened by machin e learning, are calculated by first-\nprinciples density functional theory. The applicability o f the proposed technique of combined machine learning\nand first-principles calculations is demonstrated on 2-17- X magnets, Ce 2Fe17−xCoxCN. Further to this study,\nwe explore stability of the proposed compounds by calculati ng vacancy formation energy of small atom intersti-\ntials (N/C). Our study indicates a number of compounds in the proposed family, offers the possibility to become\nsolution of cheap, and efficient permanent magnet.\nPACS numbers:\nINTRODUCTION\nPermanent magnets are a part of almost all the most im-\nportant technologies, starting from acoustic transducers , mo-\ntors and generators, magnetic field and imaging systems to\nmore recent technologies like computer hard disk drives,\nmedical equipment, magneto-mechanics etc.[ 1]The search\nfor efficient permanent magnets is thus everlasting. In this\nconnection, the family of rare-earth (RE) and 3dtransition\nmetal (TM) based intermetallics has evolved over last 50\nyears or so, and has transformed the landscape of permanent\nmagnets.[ 2,3] Two most prominent examples of RE-TM per-\nmanent magnets, that are currently in commercial produc-\ntion, together with hard magnetic ferrites, are SmCo 5, and\nNdFe14B.\nWhile SmCo 5and NdFe 14B provide reasonably good solu-\ntions, keeping in mind the resource criticality of RE elemen ts\nlike Nd and Sm, a significant amount of effort has been put\nforward in search of new permanent magnets without criti-\ncal RE elements or with less content of those. The idea is to\noptimize the price-to-performance ratio.[ 2] This has lead to\ntwo routes, (a) search for potential magnets devoid of rare-\nearth elements,[ 4] and (b) designing of rare-earth lean inter-\nmetallics using abundant RE elements such as La and Ce in-\nstead of Sm and Nd.[ 5–7] As stressed by Coey,[ 8] the demand\nin hand is to seek for new, low-cost magnets with maximum\nenergy product bridging the ferrites and presently used RE\nmagnets. Following the route (b), cheap, new ternary and\nquartnary RE-lean RE-TM intermetallics need to be explored ,\nas binaries have been well explored. In parallel, Co being ex -\npensive, it may be worthwhile to focus on intermetallic com-\npounds containing Fe.\nStarting from the simplest binary RE-TM structure of\nCaCu5, by replacing nout ofmRE (R) sites with a pair ofTM (M) sites, R m−nM5m+2nstructures are obtained. This\ncan give rise to several possible binary structures of diffe r-\nent chemical compositions, listed in order of RE-leanness;\nRM13(7.1%), RM12(7.7%), R2M17(10.5%), R2M14(12.5\n%), RM5(16.7%), R6M23(20.7%), R2M7(22.2%), RM3\n(25%), RM2(33%) etc. Judging by the rare-earth content,\n1:13, 1:12, 2:17, 2:14 compounds may form examples of rare-\nearth lean materials. It is desirable to modify the known bi-\nnary compounds containing low cost RE’s belonging to these\nfamilies to achieve best possible intrinsic magnetic prope r-\nties, namely (i) high spontaneous or saturation magnetizat ion\n(µ0Ms), at least around 1T, (ii) a Curie temperature (T c) high\nenough for the contemplated devise use, 600 K or above, and\n(iii) a mechanism for creating sufficiently high easy-axis c o-\nercivity ( Ku). The synthesis and optimization of properties\nof real materials in experiment is both time-consuming and\ncostly, being mostly based on trial and error. Computationa l\napproach in this connection is of natural interest to screen\ncompounds, before they can be suggested and tested in lab-\noratory. Typical computational approaches in this regard a re\nbased on density functional theory (DFT) calculations. A de -\ntailed calculation estimating all required magnetic prope rties,\ni.eMs, Tc,Kufrom first-principles is expensive and also not\ndevoid of shortcomings. For example, estimation of T cre-\nlies on parametrization of DFT or supplemented Ucorrected\ntheory of DFT+ Utotal energies to construct spin Hamilto-\nnian and solution of spin Hamiltonian by mean field or Monte\nCarlo method. While this approach would work for local-\nized insulators, its application to metallic systems with i tin-\nerant magnetism is questionable, as it fails even for elemen tal\nmetals like Fe, Co and Ni.[ 9] A more reasonable approach of\nDFT+dynamical mean field (DMFT)[ 10] is significantly more\nexpensive. An alternative approach would be to use machine\nlearning (ML) technique based on a suitable training datase t.2\nFIG. 1: (Color online) Steps of Machine learning combined DF T approach for predictions of properties in Ce 2Fe17−xCoxCN permanent\nmagnets.\nThis approach has been used for RE-TM permanent magnets\nbased on DFT calculated magnetic properties database of M s\nandKu.[5,11] Creation of database based on calculations,\neven with high throughput calculations is expensive, and re -\nlies on the approximations of the theory. It would be far more\ndesirable to built a dataset based on experimental results, and\nthen train the ML algorithm based on that. However, the size\nand availability of the experimental data in required forma t\ncan be a concern. Focusing on the available experimental\ndata on RE lean intermetallics, the set of T cis largest, fol-\nlowed by that for Ku, and M s. While the quantitative values\nof Tc’s in Kelvin or degree Celsius are available in literature,\nfor magnetocrystalline anisotropy often only the informat ion\nwhether they are easy-axis or easy-plane are available. Sim -\nilarly, the µoMsvalues are reported either in µB/f.u. or in\nemu/gm or in Tesla, conversion from µB/f.u. and emu/gm to\nTesla requiring information of the volume and density, whic h\nmay introduce inaccuracies up to one decimal point. Restric t-\ning experimental data to those containing values of Ku, and\nµoMsvalues in the same format (either Tesla or µB/f.u. or\nemu/gm) reduces the dataset of Kuand Mssignificantly, mak-\ning application of ML questionable. We thus use a two-prong\napproach, as illustrated in Fig. 1. We first create a database ofTc, MsandKufrom available experimental data on RE-lean\nintermetallics, and use ML for prediction of T cvalues, for pre-\ndicting whether µ0Mssatisfies the criteria of being larger than\n1 Tesla, and for predicting the sign of Ku. For M sandKu, ML\nthus serves the purpose of initial screening. We next evalua te\nMsand the magnetic anisotropy properties based on elaborate\nDFT calculations. Calculation of the magnetic anisotropy e n-\nergy (MAE) is challenging due to its extremely small value.\nHowever, since the pioneering work of Brooks,[ 12] several\nstudies[ 6,13–15] have shown that Ucorrected DFT generally\nreproduces the orientation and the right order of magnitude of\nthe MAE.\nWe demonstrate applicability of our proposed approach on\nCe and Fe based 2:17 RE-TM intermetallics, Ce 2Fe17−xCox\ncompounds ( x= 1,..., 7). Our choice is based on follow-\ning criteria, (a) the compounds contain rare earth Ce which i s\nthe cheapest one among the RE family having market price\nof∼5 USD/Kg.[ 16] The cost of other components Fe, C\nand N are all <1 USD/Kg. The price of Co is higher than\nFe,[16] being less abundant metal. The Co:Fe ratio is thus\nrestricted within 0.4. (b) Co substitution in place of Fe has\nbeen reported[ 17,18] to be efficient in simultaneous enhance-\nments of K uas well as T cin several TM magnets. This is3\nin sharp contrast to other TM substitutes, such as Ti, Mo, Cr,\nand V , where magnetic anisotropy as well as T care gener-\nally suppressed. (c) the search space belongs to 2:17 family ,\nwhich is the family in which most of the instances in our train -\ning set belongs to. (d) this class of compounds is found to\nbe more stable than the well explored 1:12 compounds. (e)\nfor large saturation magnetization it is desirable to use Fe -\nrich compounds, which is also less expensive compared to\nCo. (f) although Ce has negative second order Stefan’s fac-\ntor which favors in-plane MAE, experimental findings suppor t\nthat the nitrogenation and carbonation can switch the MAE\nfrom easy plane to easy axis.[ 19] (g) though R 2Fe17com-\npounds display large magnetization value due to high Fe con-\ntent, these compounds are disadvantageous as they exhibit l ow\nCurie temperature.[ 20] Presence of Co, as well as C/N inter-\nstitials help in increasing T c. (h) while magnetic properties of\ncarbo-nitrides are expected to be similar to that of nitride s for\nsufficiently high concentration of N, carbo-nitride compou nds\nhave been proven to show better thermal stability.[ 21]\nOur study suggests that Fe-rich Ce 2Fe17−xCoxCN com-\npounds may form potential candidate materials for low-cost\npermanent magnets, satisfying the necessary requirements of\na permanent magnet with T c>600 K,µ0Ms>1 Tesla and\neasy-axis K u>1 MJ/m3. The calculated maximal energy\nproduct and estimated anisotropy field, which are technolog i-\ncally interesting figures of merit for hard-magnetic materi als,\nturn to be within the reasonable range. Some of the studied\ncompounds may possibly bridge the gap between low maxi-\nmal energy product and high anisotropy field for SmCo 5and\nvice versa for Nd 2Fe14B.\nMACHINE LEARNING APPROACH\nDatabase construction & Training of Model\nAiming to search new candidates for permanent magnets\nwe use supervised machine learning (ML) algorithm which\nhelps us to screen compounds with high T c(Tc/greaterorsimilar600 K),\nhigh M s(µ0Ms>1 Tesla), and easy axis anisotropy ( Ku>\n0) among the huge number of possible candidates of unex-\nplored RE-TM intermetallics. The first step of any ML algo-\nrithm is to construct a dataset. We construct three datasets\nof existing RE-TM compounds for T c, Msand Kusepa-\nrately using the following sources: ICSD,[ 22] the handbook\nof magnetic materials,[ 23] the book of magnetism and mag-\nnetic materials,[ 24] and other relevant references.[ 19,21,25–\n78] The datasets are presented as supplementary materials\n(SM)[ 79] as easy reference for future users. To construct the\ndatabase of rare-earth lean compounds, RE percentage in the\nintermetallic compounds is restricted to 14% which includes\nthe four different binary RE-TM combinations namely RM 12,\nRM13, R2M17and R2M14along with their interstitial and de-\nrived compounds. We discard RM 13from the dataset as only\nfew candidates are available from this series with known ex-\nperimental T c, MsandKu.Attribute Type Attribute Notation Value range\nStoichiometric CW absolute deviation <∆Z > 1.70-16.74\nof atomic no.\nCW av. of < ZTM> 10-33.30\natomic no. of TM\nCW av. of < ZLE> 0-9.79\natomic no. of LE\nCW av. Z < Z > 21.08-37.71\nCW electronegativity ∆ǫ 0.61-1.84\ndiff. of RE &TM\nCW RE percentage RE% 4.76-14.29\nCW TM percentage TM% 38.46-95.24\nCW LE percentage LE% 0-53.85\nElement Atomic no. of RE ZRE 58-71\nPresence of NTM yes/no\nmore than one TM\nPresence of LE NLE yes/no\nElectronic Total no. of f electrons fn1-28\nTotal no. of f electrons dn30-136\nTABLE I: List of 13 different attributes with description, n ota-\ntion and range used in the ML algorithm. Here ”CW” stands for\n”composition-weighted”.\nWe list a total of 565 compounds with reported experimen-\ntal Tc, among which majority of the compounds (about 55%)\nbelong to R 2M17series. The minimum contribution to the\ndataset comes from R 2M14(about10%) family. The high-\nest Tcin the dataset belongs to R 2M17class of compounds\nnamely Lu 2Co17[25] with T c∼1203 K and the compound\nwith lowest T cis NdCo 7.2Mn4.8(∼120 K),[ 23] a member\nfrom RM 12family. In the dataset all three compositions with\nRE to TM ratio 2:17, 2:14 and 1:12 show a large variation\nin Tchaving the difference between maximum and minimum\nvalues as 1051, 775 and 991 K respectively. There exists few\ncompounds in the dataset with more than one reported value\nof Tc. For example T cof SmFe 10Mo2has been reported with\ntwo different values of 421 K[ 80] and 483 K.[ 81] There are\nother examples of such multiple T c.[82–86] The quality of the\nsample, their growth conditions, coexistence of compounds\nin two or multiple phases and accuracy of the measurements\nmay lead to the multiple values of T creported for a particular\ncompound. In such cases, we consistently consider the large st\namong the reported values of T c. Notably in majority of cases\nwe find little variation in reported values of T c(∼20-50 K).\nThe dataset of M sis relatively smaller than T c, contain-\ning only 195 entries. The majority of the compounds in this\ndataset belong to 2:17 composition similar to the database o f\nTc. The relatively smaller dimension of M sdataset is primar-\nily due to fact that experimental reports available for M sare\nmuch less than T c. Secondly M shas been mostly reported\nat room temperature, in some cases at low temperature. To\nmaintain uniformity of the dataset we consider M sreported at\nroom temperature, resulting in a lesser number of compounds\nin the M sdataset.\nReports with quoted values of anisotropy constant are even\nmore rare. Our exhaustive search resulted in only 73 data4\npoints. This pushes the dataset size to the limit of ML algo-\nrithms, for which predictive capability becomes questiona ble\ndue to large bias masking the small variance.[ 87] On other\nhand, if we allow for also experimental data reporting only\nsign of Ku, this dataset gets expanded to a reasonable size of\n258.\nAfter constructing the dataset, we carry out preprocessing\nof the data, as outlined in Ref.[ 88]. It comprises of removal of\nnoisy data, outliers and correlated attributes. For details see\nAppendix.\nThe next and the most crucial step is to construct a set\nof simple attributes, which are capable of describing the in -\nstances (in this case RE-TM compounds) and then deploy ML\nalgorithm to map them to a target (in this case T c, Msand\nKu). The attributes considered in this study are summarized\nin Table. I, which can be divided into three broad categories ,\nnamely, stoichiometric attributes, element properties an d elec-\ntronic configuration attributes. The stoichiometric attri butes\nmay contain the information of both elemental and composi-\ntional properties as suggested by Ward et al.[ 89] This is based\non taking compositional weights (CW) of elemental proper-\nties.\nIn the third step, we train different popular machine learn-\ning algorithms with the constructed dataset for prediction . We\nuse ML algorithm in three different problems; (a) to predict\nthe compounds with T cmore than 600 K, (b) compounds\nwithµ0Ms>1 Tesla, and (c) compounds with easy-axis\nanisotropy. Regression is used in the former case, whereas\nlatter two cases are treated as classification problems. We\nuse five different ML algorithms for regression in case of\nTcnamely Ridge Regression (RR),[ 90] Kernel Ridge Re-\ngression (KRR),[ 91] Random Forest (RF),[ 92,93] Support\nVector Regression (SVR)[ 94] and Artificial Neural Network\n(ANN).[ 95]The details can be found in Appendix. Out of the\nfive different ML algorithms, it is seen that random forest pe r-\nforms best, which has been also successfully used for predic -\ntion of Heusler compounds,[ 96] half-Hausler compounds,[ 97]\ndouble perovskite compounds,[ 88] half-Heusler semiconduc-\ntor with low-thermal-conductivity,[ 98] zeolite crystal struc-\nture classification[ 99] etc. Results presented in the following\nare based on random forest method.\nModel evaluation\nThe final step is to employ the trained algorithm on yet-\nto-be synthesized RE-TM compounds, and thus to explore\nnew compositions with targeted properties. We choose\nCe2Fe17−xCoxCyNz(y,z= 0/1;x= 0...8) as the explo-\nration set for application of the trained ML algorithm. This\nresults in a set of 36 compounds among which 8 compositions\n(Ce2Fe17−xCoxCN,x= 1,..., 8) have neither been synthe-\nsized experimentally nor studied theoretically, to the bes t of\nour knowledge. We apply our trained ML algorithms on all\nof these 36 compounds and the results are summarized in Fig.\nFIG. 2: (Color online) ML predictions of Curie temperature ( Tc)\nfrom regression model, and saturation magnetization (M s) and\nanisotropy constant ( Ku) from classification model. The upper\n(middle/lower) panel shows the results of T c(Ms/Ku). The ex-\nploration set is Ce 2Fe17−xCoxCyNzwhereyandzcan have val-\nues either 0 or 1, and x= 0...8, acronymed as xyz. In the\ntop panel, non-interstitial compounds, carbonated, nitro genated and\ncarbo-nitrogenated compounds are symbolized by circle, di amond,\nsquare and upper triangle. Different colors specify compou nds with\ndifferentxvalues. The middle panel shows the ML prediction con-\nfidence for M s. In the lower panel, ML prediction confidence for\nKuis illustrated. Here the upper (lower) half having bars with no-\nfill (shaded) shows the confidence for the compounds with posi tive\n(negative) Ku.\n2. The top panel of Fig. 2 shows the predicted T cof all the\ncompounds. It is seen that the nitrogenation or carbonation\nincreases the T cwith respect to their respective parent com-\npound Ce 2Fe17−xCox. Our ML model predicts that the ni-\ntrides have higher T cthan that of the carbides. For x≤5,\nthe enhancement of T cis maximum for the compounds where\nboth carbon and nitrogen are present. For x > 5, Tcshows\nslight decrease compared to only nitrogenated case. It is al so\nnoted that the relative rise in T cin interstitial compounds com-\npared to parent compounds, decays gradually with Co concen-\ntration. The increase in T cvaries from ∼200 K to 10 K as\nxvaries from 0 to 8 for carbides and nitrides whereas intro-\nduction of both nitrogen and carbon shows the variation from5\nFIG. 3: (Color online) Crystal structure of Ce 2Fe17−xCoxCN magnets. The Ce, Fe/Co and C/N atoms are shown with large, m edium and\nsmall balls, respectively. Four transition metal sublatti ces 9d, 18f, 18hand 6care shown in black, green, magenta and yellow colored balls,\nrespectively. Left panel shows the crystal structure viewe d with c-axis pointed vertically up and the right panel shows the crystal structure\nviewed along the c-axis.\n∼310 K to 30 K. Our result reproduces the trend of exper-\nimental findings in a qualitative manner. The experimental\nresults for x= 0 (Ce 2Fe17),[100,101] concluded that the en-\nhancement in T cis highest in presence of both carbon and\nnitrogen[ 102,103] (Tc∼721 K), followed by nitrogenated\ncompound[ 104,105] (Tc∼700 K) and lowest for carbonated\ncompound[ 102,103] (Tc∼589 K). Though it is not possible\nto compare the results quantitatively as the stoichiometry of\nthe experimentally studied carbonated and nitrogenated co m-\npounds are not the same as in our exploration dataset, but the\noverall trend is similar. We also find that our ML model under-\nestimates the T cof the pure binary compound Ce 2Fe17.[20]\nThis is expected, as already discussed, our model is less pre -\ncise for the prediction of low T ccompounds.\nSwitching to the M spart, the middle panel of Fig. 2 shows\nthe confidence of classification of compounds with µ0Ms\nmore than 1 T. The confidence value closer to 1 implies that\nthe prediction is viable to be more accurate. All the com-\npounds are classified in favor of forming permanent magnets\nwithµ0Ms>1 T. For compounds like Ce 2Fe17−xCoxthe pre-\ndiction confidence varies from 0.6 to 0.8 with increasing Co\nconcentration, whereas the carbon and nitride compounds ar e\nalways classified with high prediction confidence.\nThe predictions from classification model on Kuis\nshown in bottom panel of Fig. 2. We find while\nthe anisotropy of Fe 17−xCoxcompounds without inter-\nstitial C/N ( x= 2,...7) atoms are predicted to be\neasy-plane, their carbonated/nitrogenated/carbo-nitro genated\ncounterparts show easy-axis anisotropy. For pure Fe com-\npounds, apart from carbo-nitrogenated compound, all are pr e-\ndicted to be easy-plane, while for Fe 16Co compounds carbon-\nated as well as carbo-nitrogenated compounds are predicted to\nbe easy-axis. This in turn, highlights the effectiveness of Cosubstitution on making Kupositive. We note the prediction\nconfidence of the carbo-nitrogenated compounds are around\n0.75.\nOn basis of the above ML analysis, we pick up seven yet-to-\nbe synthesized compounds, Ce 2Fe17−xCoxCN,x= 1,..., 7.\nThis choice is guided by the compounds satisfying T c>600\nK from regression model, and µ0Ms>1 Tesla with easy-axis\nanisotropy from classification models, and being Fe-rich. I n\nfollowing, we describe their crystal structure, and presen t re-\nsults of DFT calculated electronic structure, anisotropy p rop-\nerties, and stability properties.\nDFT CALCULATED PROPERTIES OF PREDICTED\nCOMPOUNDS\nCrystal Structure\nThe Ce 2Fe17compounds crystallize in the rhombohedral\nTh2Zn17-type structure (space group R ¯3m), derived from the\nCaCu5-type structure with a pair (dumbbell) of Fe atoms for\neach third rare earth atom in the basal plane and the substi-\ntuted layers stacked in the sequence ABCABC .... As shown\nin Fig. 3, the transition metal atoms are divided into four su b-\nlattices, 9 d, 18f, 18hand 6c, having 3 (9), 6(18), 6 (18), and\n2 (6) multiplicity in the one (three) formula unit primitive -\nrhombohedral (hexagonal) unit cell. The TM atoms occupy-\ning the 6csites, referred as dumbbell sites, form the ...-TM-\nTM-RE-RE- ...chains running along the c-axis of the hexag-\nonal cell. The 18 fTM atoms form a hexagonal layer, which\nalternates with the hexagonal layer formed by 9 dand 18hTM\natoms. The 6 cTM-TM doumbells pass through the hexagons\nformed by 18 fTM’s. For the interstitial C and N atoms, neu-6\ntron powder diffraction,[ 106] EXAFS experiments confirmed\nthat they fill voids of nearly octahedral shape formed by a rec t-\nangle of 18 fand 18hTM atoms and two RE atoms at opposite\ncorners, which are the 9 esites of Th 2Zn17-type structure, and\nhaving the shortest distance from the RE sites among all avai l-\nable interstitial sites. All our calculations are thus carr ied out\nwith C/N atoms in 9 epositions. The RE atoms in 6 cposition\nas well as light elements C/N in 9 einterstitial sites belong to\nthe same layer as 18 fTMs. As the 9 esites are in the same\nc-plane with the RE sites, having RE atoms at neighbors, in-\ntroduction of interstitials like C and N, is expected to have a\nprofound influence on the the electronic environment of RE\natom, thereby altering the magneto-crystalline anisotrop y.\nAlthough the R ¯3m symmetry is lowered upon Co substitu-\ntion and the spin-orbit coupling (SOC) in the anisotropy cal -\nculation, for the ease of identification, we will still use th e the\nnotations 9 d, 18f, 18hand 6c. Our total energy calculations\nshow that Co preferentially occupy sites in the sequence 9 d>\n18h > 6c >18f. Out of available 17 TM sites we have con-\nsidered Co substitution up to 7 sites, which result in Fe-ric h\nphases of compositions Ce 2Fe17−xCoxCN withx= 1, 2,...,\n7. Following the site preference we consider Co atoms in 9 d\nand 18hsites.\nWe expect the lattice parameters not to change much upon\nCo substitution, as Fe and Co, being neighboring elements in\nperiodic table, has similar atomic radii. Nevertheless, to check\nthe influence of Co substitution on lattice structure, we opt i-\nmize the lattice constant and the volume for all xvalues. Fol-\nlowing our expectation, the results show only a marginal de-\ncrease in lattice parameter and volume (with a maximum devi-\nation of 1 %) upon increasing Co content, in line with the find-\nings by Odkhuu et al.[ 18] for 1:12 compounds, and the exper-\nimental findings by Xu and Shaheen on 2:17 compounds.[ 19]\nThis minimal change is found to have no appreciable effect on\nmagnetic properties, as explicitly checked on representat ive\ncompounds with x= 1, 4 and 7. We thus choose the lattice\nstructure as the optimized lattice structure of x= 0 (see Ap-\npendix), with lattice constant = 6.59 ˚Aand angle β= 83.3oof\nthe rhombohedral unit cell[ 107] in subsequent calculations.\nMagnetic Moment and Electronic Structure\nIn the following we present the DFT results for the mag-\nnetic moments and density of states (DOS), as given in\nGGA+U+SOC calculations. The details of the DFT calcu-\nlations are presented in the Appendix. Importance of applic a-\ntion of supplemented Hubbard Uon RE sites within LDA or\nGGA+Uformalism is considered as one of the possible means\nto deal with localized forbitals of RE ions, and have shown to\nprovide reasonable description.[ 13,14] Previous calculations\nin compounds containing Ce, showed variation of Uwithin\n3 eV to 6 eV , keeps the results qualitatively same.[ 6,108] In\nthe following, we present results for Uapplied on Ce atoms\nchosen to be 6 eV .\nFig. 4 shows the calculated total magnetic moments of the1.61.71.81.9Total moment (Tesla)\nFe Co\n16 Fe CoFe CoFe CoFe CoFe CoFe Co\n15 2 14 313 4 12 511 610 7\nFIG. 4: (Color online) Calculated total moment (black cir-\ncles),µ0M in Tesla plotted for increasing Co concentrations of\nCe2Fe17−xCoxCN compounds. Shown are also experimental\nresults[ 19] (red, square) for Ce 2Fe17−xCoxNycompounds measured\nat room temperature. For comparison between T = 0 K calculate d\nmoments, and experimental data measured at room temperatur e, the\nexperimental data has been scaled by a factor of 1.3.\nFIG. 5: (Color online) Calculated spin (top) and orbital (bo ttom)\nmoments at Ce, Fe(9 d), Fe(18f), Fe(18h), Fe(6c) and Co sites in the\nrepresentative case of Ce 2Fe15Co2CN compound.\nseven mixed Fe-Co compounds, Ce 2Fe17−xCoxCN (x= 1, 2,\n...7). The total magnetic moment shows a decreasing trend\nwith increase of Co concentration, arising from the fact that\nCo moment is smaller that of Fe. However, it is reassuring to\nnote that even for compound with largest Co concentration,\nCe2Fe10Co7CN, the calculated moment is more than 1.657\nFIG. 6: (Color online) Left: Density of states of Ce 2Fe15Co2CN compound, projected onto Ce f(brown), Ce d(shaded green), Fe d(blue),\nCod(shaded red) and CN p(shaded orange) characters. Right: Density of states of Ce 2Fe15Co2CN compound projected to different Fe d’s:\nFe(9d) (shaded indigo), Fe(18 h) (magenta), Fe(18 f) (green) and Fe(6 c) (brown). The zero of the energy is set at Fermi energy.\nTesla. This is in agreement with ML prediction, which pre-\ndictsµ0Msof all the considered compounds to be larger than\n1 Tesla, though it is to be noted the ML predictions are made\nfor room temperature moments while the DFT calculated mo-\nments are at T = 0 K. The measured values of total moment in\ncorresponding nitrogenated compounds show good compari-\nson (cf Fig. 4) with our calculated moments. In particular,\nbarring the data on x ≈2, the other two data point show good\nmatching with the trend of theoretical results. We note that the\nexperimentally determined moments are for Ce 2Fe17xCoxNy\ncompounds, which contains only N as interstitial atom, and\nthe value of yis not mentioned, which may even vary depend-\ning on value of x.\nFig. 5 shows the spin and orbital moments projected to Ce,\nFe(9d), Fe(18f), Fe(18h), Fe(6c) and Co atoms for the rep-\nresentative case of Ce 2Fe15Co2CN compound. The results\nfor other Co concentrations are similar. In presence of larg e\nSOC coupling at Ce site, a substantial orbital moment devel-\nops, which is oppositely aligned to its spin moment followin g\nHund’s rule. Considering 3+ nominal valence of Ce, it would\nbe in 4f1state, with S=1/2 and L=3. While the calculated\nvalue of Ce spin moment is close to 1 µB(≈0.95µB) in ac-\ncordance with nominal S=1/2 state, the orbital moment shows\nsignificant quenching with a calculated value of about 0.5 µB.\nThis value of orbital moment is in agreement with DFT calcu-\nlated values of other Ce containing RE-TM magnets.[ 6,109]\nThe 4felectrons are coupled to 5 delectrons at Ce site by\nintra-atomic exchange interaction, following which their spin\nmoments are aligned in parallel direction. The delocalized 5d\nelectrons at Ce site, hybridize with Fe/Co 3 delectrons, favor-\ning antiparallel alignment of Ce and Fe/Co spins, as found in\nFig. 5. The spin magnetic moment at Fe sites show a distribu-\ntion, with Fe at 6 csite having largest moment, followed by Feat 9dand 18hsites while Fe at 18 fsite shows the lowest mo-\nment. We notice that Fe (6 c) atoms occupying the dumbbell\nsites, have less connectivity compared to Fe(9 d), Fe (18f) and\nFe (18h), and thus possess the largest moment, being of most\nlocalized character. Among Fe (9 d), Fe(18f), Fe(18h) sites\nFe (18f) has smallest moment, driven by the fact that inter-\nstitial C and N atoms are in same plane as Fe (18 f) causing\nenhanced d-phybridization, and reduction in moment. These\nspin moments though are larger than that of bulk Fe ( ≈2.2\nµB). The orbital moment at Fe sites are tiny ( ≈0.05µB). In\ncomparison, Co shows significantly smaller spin moment ( ≈\n1.7µB) and somewhat larger orbital moment ( ≈0.1µB), jus-\ntifying the fall in total moment with increasing concentrat ion\nof Co.\nFig. 6 shows the density of states of Ce 2Fe15Co2CN, pro-\njected to various orbital characters. The Ce 4 fstates are all\nunoccupied in the majority spin channel, partly occupied in\nthe minority spin channel, in accordance with nominal f1oc-\ncupancy. The RE 4 f- TM 3dhybridization through empty RE\n5dstates is visible, making the spin splitting at Fe and Co site s\nantiparallel to that of Ce. The C/N pstates mostly spanning\nthe energy range -7 eV to -4 eV , show non negligible mixing\nwith Fed, Codand Ce characters, justifying their role in in-\nfluencing the magnetic properties. Fe dand Codstates span\nabout the same energy range from -4 eV to 2 eV , with states\nmostly occupied in the majority spin channel and partially o c-\ncupied in the minority spin channel, largely accounting for the\nmetallicity of the compound. Spin splitting of Fe dis larger\nthan that of Co, being consistent with larger magnetic momen t\nof Fe compared to Co. Projection to different inequivalent F e\nsites (cf right panel of Fig. 6), Fe(9 d), Fe(18h), Fe(18f) and\nFe(6c) shows that Fe(6 c) belonging to dumbbell pair is dis-\ntinct from other Fe sites, which also exhibit largest magnet ic8\nFIG. 7: (Color online) Top: Calculated magnetocrytalline anisotropy\nconstant in MJ/m3plotted for increasing Co concentrations of\nCe2Fe17−xCoxCN compounds. The inset shows the anisotropy in\norbital moment (see text for details). Bottom: The GGA+ U+SOC\nDOS projected to Ce fenergy states with magnetization axis\npointed along easy-axis, for Ce 2Fe17(black), Ce 2Fe17CN (red) and\nCe2Fe16CoCN (blue). The zero of the energy is set at Fermi energy,\nwith unoccupied part shown shown as shaded. The arrow indica tes\nthe shift in occupied part.\nmoment among all Fe’s.\nMagneto-crystalline Anisotropy\nHaving an understanding of the basic electronic structure,\nin terms of magnetic moments and density of states, we next\nfocus on calculation of magneto-crystalline anisotropy co n-\nstant, Ku, which is a crucial quantity responsible for coerciv-\nity in a permanent magnet. MAE defines the energy required\nfor turning the orientation of the magnetic moment under ap-\nplied field, expressed as E(θ)≈K1sin2θ+K2sin4θ+\nK3sin4θcos4φ, where K 1, K2, and K 3are the magnetic\nanisotropy constants, θis the polar angle between the mag-\nnetization vector and the easy axis ( c-axis), and φis the\nazimuthal angle between the magnetization component pro-\njected onto the abplane and the a-axis. In most cases, the\nhigher order term K 3is relatively small compared with K 1\nand K2. Forθ=π/2, one may thus write Ku≈K1+K2.It’s positive and negative values indicate the easy axis and\neasy plane anisotropy, respectively. To satisfy the criter ia of a\ngood permanent magnet, it should have easy axis anisotropy\nwith value larger than 1 MJ/m3.[2,8] The MAE in RE-TM\narises from two contributions, (i) MAE of the RE sublattice\ndue to strong spin-orbit coupling and crystal field effect an d\n(ii) MAE of TM sublattice. The interplay of the two decides\nthe net sign and magnitude. In particular, in the proposed\ncompounds, presence of Co with significant value of orbital\nmoment, makes the contribution of TM sublattice important.\nWhile 2:17 compounds, primarily show easy plane anisotropy ,\nswitching to easy axis anisotropy for interstitial compoun ds\nhave been reported. In particular, upon nitrogenation, eas y\nplane anisotropy has been reported for Ce containing mixed\nFe-Co compounds.[ 19] As mentioned already, the interstitial\natoms occupy the same plane as the RE atoms, significantly\ninfluencing their properties. With predicted high T cand large\nsaturation moment of our proposed compounds with carbon-\nation and nitrogenation, it remains to be seen whether they\nwould exhibit easy axis anisotropy of reasonable values, as\nrequired for a legitimate candidate for permanent magnet. F or\nthis purpose, we carry out calculations within GGA+ U+SOC\nwith magnetization axis pointing along the crystallograph ic c-\naxis and perpendicular to it. The importance of application of\nUon proper description of MAE in terms of its sign and order\nof magnitude has been stressed upon by several authors.[ 6,13]\nIn order to establish our method on calculation of MAE in-\nvolving small energy difference, we first apply our method\nto known and well studied case of SmCo 5, with choice of U\n= 6 eV on Sm, and obtained a MAE value of 24.4 meV/f.u,\nwhich agrees well with GGA+ U+SOC calculated value of\n21.6 meV/f.u., reported in literature[ 13] as well as experimen-\ntally measured values of 13-16 meV/f.u.[ 110] The calculated\nresults for the proposed Ce 2Fe17−xCoxCN are shown in top\npanel of Fig. 7. We found that MAE shows site-dependence\non the Co substitution. We consider configurations with Co\natoms substituting Fe(9 d) and Fe(18 h) sites , configurations\ninvolving other substituting sites being energetically mu ch\nhigher. We consider configurations which are energetically\nclose (within 600 K) and calculate the Co-composition de-\npendent MAE using the virtual crystal approximation. Speci f-\nically, for x= 1 we consider configurations Co@Fe(9 d) and\nCo@Fe(18 h), the latter being 3.58 meV higher compared to\nformer. Similarly for x= 2, we consider Co@ 2 ×Fe(9d)\nand Co@ 2 ×Fe(18h), the latter being 4.43 meV higher com-\npared to former. For x= 3, the configurations considered are,\nCo@ 2×Fe(9d)+ Fe(18h); Co@ 3 ×Fe(9d); Co@Fe(9 d) +\n2×Fe(18h), the energies being 0 meV (set as zero of energy),\n12.37 meV and 47.66 meV , respectively. For x= 4, the con-\nfigurations considered are, Co@ 2 ×Fe(9d) + 2×Fe (18h);\nCo@ 3×Fe(9d) + Fe(18h), the energies being 0 meV (set as\nzero of energy) and 36.5 meV , respectively. For x= 5 , 6 and\n7, only one configuration is considered, others being energe t-\nically much higher, namely, Co@3 ×Fe(9d) + 2×Fe(18h),\nCo@3×Fe(9d)+ 3×Fe(18h) and Co@3 ×Fe(9d) + 4×\nFe(18h), respectively.9\nFIG. 8: (Color online) Calculated anisotropy field in Tesla (left) a nd maximal energy product in kJ/m3(right) plotted for increasing Co\nconcentrations of Ce 2Fe17−xCoxCN compounds.\nConsidering spin-orbit effect only on Ce atom, it is found\nto account for about 60 %of the calculated MAE. We find all\nthe calculated MAE is positive, in good agreement with ML\nprediction on mixed Fe-Co carbo-nitride compounds. Furthe r\nMAE values show non-monotonic dependence on Co concen-\ntration. Such non-monotonic trend upon varying TM con-\ntent has been also reported in context of R(Fe 1−xCox)11TiZ\n(R = Y and Ce; Z= H, C, and N)[ 7] and R-TM systems in\ngeneral.[ 111] In the inset of top panel of Fig. 7, we show\nthe calculated orbital magnetic anisotropy ( ∆ML) defined as\n∆ML= ML(a) - M L(c), as employed in Ref. 18, ML(c) and\nML(a) being the orbital moment along the c-axis and a-axis,\nrespectively. We find a correlation between ∆MLand Ku,\nqualitatively satisfying Bruno’s expression[ 112] for itinerant\nferromagnets given as, K u= (ξ\n4µB)∆ML, whereξis the\nstrength of SOC.\nMost of the easy-axis Kuvalues are found to be larger than\n1 MJ/m3, except Fe 14Co3and Fe 13Co4for which it is found\nto be 0.74 and 0.91 MJ/m3, respectively. Few of the con-\ncentrations exhibit easy-axis Kuvalues larger than 2 MJ/m3,\ne.g.Fe15Co2(3.54 MJ/m3), Fe12Co5(3.39 MJ/m3), Fe11Co6\n(3.39 MJ/m3), Fe10Co7(9.10 MJ/m3), being comparable to\nNd2Fe14B (4.9 MJ/m3).[113]\nTo obtain microscopic understanding of the role of\nCo substitution and doping by C, N on magnetocrys-\ntalline anisotropy, we further calculate the magnetocrys-\ntalline anisotropy of Fe-only compounds Ce 2Fe17, Ce2Fe17C,\nCe2Fe17N and Ce 2Fe17CN. This results in negative K uval-\nues for Ce 2Fe17, and Ce 2Fe17C (-2.12 MJ/m3and -1.35\nMJ/m3), a tiny positive value for Ce 2Fe17N (0.26 MJ/m3)\nand positive value for co-doped compound Ce 2Fe17CN (1.27\nMJ/m3). We further plot the GGA+ U+SOC density of states\n(cf bottom panel, Fig. 7) with magnetization axis along c-\naxis projected to Ce fstates for Ce 2Fe17, Ce2Fe17CN and\nCe2Fe16CoCN, which is expected to reveal the mechanism of\nuniaxial anisotropy. We find that a lowering of occupied Cefenergy states and increase in band width occur upon intro-\nduction of light elements C and N. This gets further helped\nby substitution of Co, caused by hybridization between Ce f\nstates and Co dand C,Npstates. This gain in hybridization\nenergy stabilizes easy-axis magnetization (cf. Ref. 114) as ob-\nserved experimentally.[ 19]\nMaximal energy product and Anisotropy Field\nWhile, the estimates of Kuandµ0Msare useful infor-\nmation to access the effectiveness of the suggested materi-\nals as permanent magnets, technologically interesting figu res\nof merit of hard magnetic materials, are the maximal energy\nproduct (BH) max and anisotropy field H a. These can be esti-\nmated from the knowledge of µ0MsandKuas follows,\n(BH)max=(0.9µ0Ms)2\n4µ0\nHa=2Ku\nµ0Ms\nThe factor 0.9 in the expression for (BH) maximplies the com-\nmon assumption that ideally out 10 %of a processed bulk hard\nmagnet consists of non-magnetic phases.[ 115] The estimated\n(BH)max and Hais shown in Fig. 8. The (BH) max value\nis found to range from 444 to 540 kJ/m3, in comparison to\nexperimentally measured values 516 kJ/m3and 219 kJ/m3\nfor Nd 2Fe14B[116] and SmCo 5,[116] respectively. The H a\nshows a strong variation with Co concentration, ranging fro m\n≈1 Tesla to 14 Tesla. [117]\nWe further note that the hardness parameter, defined as κ\n=/radicalBig\nKu\nµ0M2s, turns out to be greater than 1 for Ce 2Fe15Co2CN,\nCe2Fe12Co5CN, Ce 2Fe11Co6CN, and Ce 2Fe10Co7CN com-\npounds, employing the calculated T = 0 K values of Kuand\nMs.10\n∆Ef(CN)∆Ef(N)∆Ef(C)\nx= 1 4.32 2.10 0.97\nx= 2 3.99 2.09 0.85\nx= 3 4.16 2.09 0.88\nx= 4 3.98 2.10 0.79\nx= 5 3.82 2.07 0.70\nx= 6 3.91 2.05 0.72\nx= 7 3.78 2.01 0.69\nTABLE II: Vacancy formation energy for carbon ( ∆Ef(C)), ni-\ntrogen (∆Ef(N)) and nitrogen-carbon ( ∆Ef(CN)) in eV in\nCe2Fe17−xCoxCN compounds.\nStability\nUnlike the other RE-TM magnets like 1:12 compounds, one\nof the advantage of 2:17 compounds is their stability. Both\nstable form of Ce 2Fe17and its Co substituted form have been\nreported in literature.[ 19] Calculation of formation enthalpies,\nas given in Ref. 18,Eform=Ecompound −/summationtext\nkNkǫk/summationtext\nkNk, whereNk\nindicate number of different atoms (Ce, Fe, Co, N and C) in\nthe cell, and ǫkdenote energy/atom of bulk Ce in FCC struc-\nture,α−Fe, Co in HCP structure, in molecular nitrogen and\nC in graphite structure, gives values -0.61 to -0.59 eV/atom\nfor the studied Ce 2Fe17−xCoxCN compounds.\nA major challenge with interstitial compounds, though, is\nthe nitrogen diffusion.[ 21] It has been further suggested the\nblockage of nitrogen diffusion by carbon layer is useful in r e-\nduction of nitrogen outgassing in carbo-nitrides. In parti cular,\nheating up Sm 2Fe17carbo-nitrides at a constant rate in a dif-\nferential scanning calorimeter, the onset temperature of n itro-\ngen outgassing was found to be higher by more than 40 K, as\ncompared to nitride counterpart.[ 21] This justifies the choice\nof carbo-nitrides as our exploration set. To this end, we cal -\nculate the vacancy formation energy of the interstitial ato ms\nin our chosen compounds. For this purpose, we calculate the\nformation energy of the N and/or C vacancy ( ∆Ef) defined\nas,\n∆Ef=EN(C)vac−Epristine+EN(C)\nwhereEN(C)vacandEpristinedenote the optimized total en-\nergies of compound containing N and/or C vacancy, and va-\ncancy free compound. The internal positions for defect free\npristine structure and structures containing nitrogen and /or\ncarbon vacancies are performed keeping the lattice parame-\nters fixed. EN(C)is the energy per N or C atom, which is\nobtained from calculation of N 2molecule or graphite. The ob-\ntained results for Ce 2Fe17−xCoxCN compounds in minimum\nenergy configuration of Co is shown in Table II. The vacancy\nformation energies show hardly any variation on chosen con-\nfiguration for a given Co concentration.\nThe vacancy formation energies, listed in Table II, show\nonly small variation between compounds of varying Co\nconcentration, with the general trend ∆Ef(CN)>/summationtext\n(∆Ef(N)+∆Ef(C)). The individual nitrogen vacancy\nformation energy and carbon vacancy formation energy, arein overall agreement with that found for related compound,\nSmCaFe 17C(N)3.[6] The vacancy formation energy for co-\ndoped carbon-nitrogen compounds are found to be enhanced\nby about 35-40 %compared to the sum of the individual\nC and N vacancy formation energies, proving the carbo-\nnitrogenation co-doping to provide better thermal stabili ty.\nWe also check our results by repeating vacancy formation en-\nergy calculations for x= 0 compounds, which however do not\nshow significant difference, suggesting Co doping not havin g\nmajor role in stability, as also indicated by no significant v ari-\nation of results between x= 1, 2, 3, 4, 5, 6 and 7.\nCONCLUSION\nDesigning alternative solutions for permanent magnets, sa t-\nisfying the criteria of low-cost, while keeping the magneti c\nproperties comparable to those of permanent magnets in use,\nis of utmost importance for cost-effective technology. To-\nwards this goal, we use a combined route of machine learning,\nbased on experimental data, and the first-principles calcul a-\ntions. While machine learning has been applied for problem of\nrare-earth magnets,[ 5] those studies have been based on the\ndataset created out of high throughput calculations. Being de-\npendent on calculation-based inputs, creation of such data base\nis not only computationally expensive, but also not devoid o f\napproximations of the theory. Our study, to the best of our\nknowledge, being based on a exhaustive search of experimen-\ntal data, is first of this kind in context of rare-earth magnet s.\nWhile a large volume of experimental data is available with\nnumerical value of T c, the corresponding dataset with numer-\nical values of M sandKuis small. On the other hand, there\nexists sizable dataset with information of Kubeing positive\n(easy axis) or negative (easy plane), and µ0Msbeing larger or\nsmaller than 1 Tesla. We thus employ regression model of ma-\nchine learning training to make predictions on numerical va l-\nues of T c, and classification model to make predictions on sign\nofKu, andµ0MSbeing larger or smaller than 1 Tesla. We ap-\nply the trained machine learning to 2:17 rare-earth transit ion\nmetal compounds with carbon and nitrogen in interstitials. We\nchoose the compounds to contain abundant rare-earth Ce, and\nto be Fe-rich to make them cost-effective. Although nitro-\ngenated version of this series has been investigated,[ 19] the\nsystematic study of the carbo-nitride family to the best of o ur\nknowledge is unavailable. The machine learning predicts T c\nof the chosen carbo-nitride family to be larger than 600 K,\nµ0MS>1 Tesla, and Ku>0, thereby indicating the pos-\nsibility of them to become good solutions for cost-effectiv e,\npermanent magnets. Subsequent first-principles calculati ons,\nshow T=0 K, µ0MSto be larger than 1.65 Tesla, and Ku/greaterorsimilar\n1 MJ/m3for the entire family, Ce 2Fe17−xCoxCN (x= 1,...\n7). Calculated Kuvalues are found to be comparable to the\nstate-of-art permanent magnet Nd 2Fe14B for Ce 2Fe15Co2CN,\nCe2Fe12Co5CN, Ce 2Fe11Co6CN, and Ce 2Fe10Ce7CN. This\nresults in two figure of merits for hard magnets, (BH) max and\nHain range of 444-540 kJ/m3and≈1 - 14 T, respectively .11\nIn spite of good magnetic properties, one of the limitation\nof practical applications of interstitial 2:17 magnets is t he for-\nmation of nitrogen/carbon vacancies at high temperature. By\ncalculating the N-(C)-vacancy formation energy, we show th at\ncarbo-nitrogenation co-doping enhances the vacancy forma -\ntion energy significantly, by 35-40 %compared to sum of in-\ndividual doping. This is likely to improve the thermal stability\nat high temperature condition.\nOur computational exercise based on exhaustive search of\nexperimental database, should motivate future experiment al\nprocesses in making high-performance 2:17 interstitial ma g-\nnets, with cheapest RE element Ce, the most abundant 3d\nmetal, Fe and cheap non-metal interstitial dopings like C an d\nN. The estimated price-to-performance based on calculated\nenergy product, and available market price[ 16] turns out to be\n0.03-0.22 USD/J. The enhanced thermal stability of the carb o-\nnitrides compounds against the vacancy formation of the lig ht\nelements further boosts the promises of the suggested com-\npounds.\nACKNOWLEDGEMENT\nThe authors acknowledge the support of DST Nano-\nmission for the computational facility used in this study.\nAPPENDICES\nDFT details\nDFT calculations for electronic structure, magnetocrys-\ntalline anisotropy are performed using the all-electron de nsity-\nfunctional-theory code in full potential linear augmented\nplane wave (FP-LAPW) basis, as implemented in WIEN2K\ncode.[ 118] For expensive structural optimization calculations,\nthe plane wave based calculations, as implemented in Vi-\nenna Ab-initio Simulation Package (V ASP),[ 119] are carried\nout. The exchange-correlation functional is chosen to be\ngeneralized-gradient approximation (GGA) of Perdew, Burk e,\nand Ernzerhof.[ 120] The localized nature of 4 fstates of Ce is\ncaptured through GGA+ Ucalculations,[ 121] with choice of\nU= 6 eV and J H= 0.8 eV . For light rare earths like Ce the U\nvalue was shown to range from 4 eV to 7 eV , without affecting\nmuch the physical properties.[ 108] The spin-orbit coupling ef-\nfect at Ce, and TM sites are captured through GGA+ U+SOC\ncalculations.\nFor FP-LAPW calculations, APW +lo is used as the ba-\nsis set, and the spherical harmonics are expanded upto l=\n10 and the charge density and potentials are represented upt o\nl=6. The sphere radii are set at 2.5, 1.9, 2.34, 1.56 and\n1.51 bohr for Ce, Fe, Co, N, and C. For good convergence, a\nRKmax value (the product of the smallest sphere radius and\nthe largest plane-wave expansion wave vector) of 7.0 is used .\nWe set the cutoff between core and valence states at −8.0 Ry.\nThe k-space integrations are performed with 112 k-points inirreducible Brillouin zone (BZ), following the report of us e\nof 80 k-points in irreducible BZ in case of SmCo 5to provide\ngood estimate of MAE.[ 13] Nevertheless, the convergence of\nresults on k-space mesh is checked by carrying out calculati on\nwith 260 k-points.\nThe structural optimization in plane wave basis is carried\nout starting with experimental structure of Sm 2Fe17CN, [ 107]\nreplacing Sm with Ce, and relaxing all the internal coordi-\nnates until forces on all of the atoms become less than 0.001\neV/˚A. Upon moving from Sm 2:17 carbide/nitride interstitial\ncompounds to Ce counterpart, the cell volume changes only\nnominally by 0.2 %to 0.4%.[6] For the plane wave calcula-\ntions, energy cut-off of 600 eV and Monkhorst pack k-points\nmesh of8×8×8are used.\nAll the calculations are performed by considering a\ncollinear spin arrangement. The MAE is obtained by calcu-\nlating the GGA+ U+SOC total energies of the system, in FP-\nLAPW basis as Ku= Ea- Ec, where E aand Ecare the ener-\ngies for the magnetization oriented along the crystallogra phic\naandcdirections, respectively. For accurate estimates of va-\ncancy formation energy, we also use FP-LAPW basis.\nData preprocessing in Machine Learning\nWhile constructing the database, we avoid inclusion of\nnoisy data. We do bootstrapping to normalize the data which\nis followed by removal of outliers with the help of violin\nplot. A data is removed if it lies outside of Q1-1.5 ×IQR or\nQ3+1.5×IQR, where IQR is the interquartile range and Q1,\nQ2 and Q3 are lower, median and upper quartile respectively.\nIn the next step we identify correlated attributes using Pea r-\nson’s correlation coefficient which can be defined as,\nr=/summationtexti=n\ni=1(xi−¯x)(yi−¯y)/radicalBig/summationtexti=n\ni=1(xi−¯x)2/radicalBig/summationtexti=n\ni=1(yi−¯y)2\nHerenis the sample size, xiandyiare sample points and\n¯xand¯yare the sample means.\nThe heatmap obtained by using the above mentioned cor-\nrelation is shown in Fig. 9. The correlation between the\nattributes is mapped between 0 and 1, considering the abso-\nlute values. The highly correlated attributes with correla tion\ngreater than 0.75 are as follows:\n1. Electronegativity difference between RE and TM ( ∆ǫ)\nand CW average of atomic no. of TM ( < ZTM>)\n2. CW TM percentage ( TM%) and CW average of atomic\nno. of TM ( < ZTM>).\n3. CW TM percentage ( TM%) and Electronegativity dif-\nference between RE and TM ( ∆ǫ).\n4. Total number of f electrons ( fn) and Atomic no. of RE\n(ZRE).12\nFIG. 9: (Color online) Heatmap indicating the correlation b etween\ndifferent attributes considered to built ML algorithm. The color code\nis shown in the side panel. The boxes with red represent weak o r no\ncorrelation, whereas blue boxes represent strong correlat ion between\nthe attributes.\n5. LE percentage ( LE%) and CW average of atomic no.\nof TM (< ZTM>).\n6. LE percentage ( LE%) and Electronegativity difference\nbetween RE and TM ( ∆ǫ).\n7. LE percentage ( LE%) and CW TM percentage\n(TM%).\nWe thus discard ∆ǫ,LE%,ZREand< ZTM>from the list\nof attributes.\nModel construction for training in ML\nFIG. 10: (Color online) Coefficient of determination R2score of five\ndifferent ML algorithms applied to T cdataset.The performance of a model can be quantified in terms\nof coefficient of determination which can be expressed as\nfollows:[ 122]\nR2= 1−/summationtextN\ni=1[yi−f(xi)]2\n/summationtextN\ni=1[yi−µ]2\nfor predictions f(xi)and a set of actual values yiwith mean\nµ. If the algorithm performs perfectly, R2score is 1. Fig.\n10 shows score R2for five different algorithms. RR al-\ngorithm circumvents issues in ordinary linear regression l ike\nover-fitting or failure in finding unique solution due to mul-\nticollinearity. It develops on least square error by adding an\nextra penalty/regularization term to the loss function of o rdi-\nnary linear regression. KRR builds on the ridge regression\ntechnique by using kernel trick [ 123] so that it can capture the\nnonlinearity present in the feature space. It can fit a non lin ear\nfunction by learning from a linear function spanned by a ker-\nnel which in turn mimics a non-linear function in the origina l\nspace. SVR originated from support vector machines which\nare mainly popular in classification problem. It is based on\nthe idea to search a hyperplane [ 124] by minimizing the er-\nror which is able to separate two different classes. SVR also\nuses kernel trick to map the data into a high dimensional fea-\nture space and then performs linear regression to fit the data .\nThese three models are based on the same principle of linear\nregression and SVR is the best form according to our result.\nR2score is 0.66 for SVR whereas it is found to be poor ( ≈\n0.25) for other two algorithms.\nApart from these we use two other algorithms, ANN and\nRF. The model performance scores are satisfactory for both\nof them. A simple ANN architecture called perceptron imple-\nments a processing element or artificial neuron called Thres h-\nold Logic Unit (TLU) which can have one or more input(s)\nand one output. Each input is related to a weight. The TLU\ncalculates the weighted sum of its inputs, applies a step fun c-\ntion (generally Heaviside or sign function) to it and output s\nthe result. A perceptron [ 125] is simply a layer of TLUs op-\nerating in parallel and connected to all the inputs. Trainin g\nan ANN model is equivalent to learning each weight factor in\nan iterative cycle. A more complex system (Multi-Layer Per-\nceptron) can be built by associating additional interconne cted\nlayers to the architecture. A well functioning system consi sts\nof an input layer, several hidden layers and an output layer.\nIn our case we have one input layer, two hidden layers where\nrectified linear unit (ReLU)[ 126] is used as activation function\nalong with L2 regularization in the kernel, and an output lay er.\nThe constructed ANN model shows 0.80 as R2score.\nRandom forest is an ensemble method which consists of\nmultiple decision trees. Each tree is built on a portion of en -\ntire training data with a subset of total number of attribute s.\nTree algorithm is based on ’top to bottom’ approach, start-\ning from a root node, it consists of many intermediate nodes\nand ends at leaf nodes. At each node of a tree a particular\nattribute classifies the data and helps to grow the tree. The\nprediction is based on accumulating the results from all suc h13\nFIG. 11: (Color online) Model output from RF algorithm for T cof RE-TM intermetallics. The left panel shows the compariso n of Tcobtained\nfrom literature and predicted T c. The distribution of absolute error between predicted T cand actual T cis shown in the upper, right panel, while\nthe lower, right panel presents the distribution of relativ e error for the compounds with T c>600 K.\ntrees, taking ensemble average in case of regression or cons id-\nering votes from majority trees in case of classification. Su ch\nan algorithm can capture the complex and nonlinear interac-\ntion between different attributes and can built a robust and\nsophisticated model. Our random forest consists of 100 tree s\nbuilt by bootstrapped[ 127] sampling of the training set. Each\ntree allows checking a maximum of log2(number of features)\nwhile detecting the best split node. The quality of such a spl it\nis measured by using mean squared error (Gini index) in re-\ngression (classification). The model efficiency is calculat ed\nby running out-of-bag samples down each of the trees. We use\nten-fold cross validation to extract the hyper-parameter a nd to\nconstruct the best model.\nFig. 11 shows the result of the best regression model us-\ning RF algorithm in case of T c. The plot in the left panel\nshows the predicted T cversus T cobtained from experiments.\nThe determination score R2is high enough (0.86), indicating\na good agreement between the predicted T cand experimen-\ntally reported T c. The mean absolute error in this model is 60\nK. Additionally we evaluate absolute error and relative err or\nfor the compounds with T c>600 K (cf Fig. 11, right panel).\nThis analysis helps to determine the model performance for\nthe compounds with T c>600 K as we are interested to pre-\ndict new RE-TM intermetallics with high T c. The distribution\nof absolute error shows that for the most of the compounds\n(≈85%) the absolute error is less than 100 K. For 65 %of the\npredicted cases, the absolute error is less than 50 K. We also\ncheck the absolute error for the compounds with T c<300 K\n(not included in the figure). In this case our model predicts\n≈76%compounds with absolute error less than 100 K and\n50%instances are predicted with absolute error of 50 K. Thisobservation prompts us to conclude that though the model pre -\ndiction is in general good, it is less accurate for low T ccom-\npounds compared to high T ccompounds. The distribution\nof relative error, expressed as ǫrel= (Texp\nc-Tpredicted\nc )/Texp\nc,\nprovides further support to this statement, which is shown i n\nbottom, right panel of Fig. 11. The relative error distribu-\ntion appears Gaussian like with slight asymmetry about the\nmean position. The relative error is less in the right side of\nthe mean position than the left side suggesting the predicti on\nof Tcsuffers less overestimation than underestimation. As\nfound, only 1 %of the instances are having ǫrel>50%, 3%of\nthe instances have 50 %> ǫrel>30%and 2%instances have\n30%> ǫrel>25%, most cases having tiny values of ǫrel.\nThis gives us confidence in accuracy of the predicted T cfor\ncompounds with T cs exceeding 600 K.\nTurning to M s, we use random forest algorithm to classify\nhigh M sfrom low M scompounds. The best model by per-\nforming 10-fold cross validation is built up with 81.53 %ac-\ncuracy. The resultant confusion matrix is shown in Fig. 12.\nFor classification problem, F1 score determines the balance\nbetween precision and recall. In this case F1 score 82.2 %indi-\ncates good anticipation with slight favour towards the pred ic-\ntion of compounds with high M s(µ0Ms>1) (83.8%) com-\npared to the compounds with low M s(µ0Ms<1) (79.2%).\nSimilar to M s, we use random forest algorithm for Ku, to\nclassify positive Kufrom negative Kucompounds. The best\nmodel by performing 10-fold cross validation, in this case,\nis built up with 80.62 %accuracy Like M s, in this case F1\nscore for positive Kuis 83%and for negative is Ku77.5%\nsuggesting slight preference of classification towards pos itive\nKuwhich is also captured in the plot of confusion matrix as14\nFIG. 12: (Color online) Normalized confusion matrix for\nµ0Ms(violet) and Ku(grey) classification using 10-fold cross-\nvalidation. Here positive (negative) class represents eit her com-\npounds with µ0Ms>(<)1T, or compounds with uniaxial\nanisotropy i.e K u>(<)0 MJ/m3. True positive/negative or\nTP/TN are the compounds where their classes are predicted co r-\nrectly. Whereas false positive (FP) and false negative (FN) are the\noff-diagonal terms of the matrix where the classes are incor rectly\nclassified.\nshown in Fig. 12.\n∗Electronic address: t.sahadasgupta@gmail.com\n[1] K. Buschow, Rep. Prog. Phys. 541123 (1991).\n[2] J. M. D. Coey, IEEE Trans. Magn. 47, 4671 (2011).\n[3] Hong Sun, Y . Otani and J.M.D. Coey, Magnetism and Mag-\nnetic Materials 104-107 1439-1440 (1992).\n[4] A. Vishina, O. Y . Vekilova, T. Bj¨ orkman, A. Bergman, H. C .\nHerper, O. Eriksson, Phys. Rev. B 101, 094407 (2020).\n[5] J. J. M¨ oller, W. K¨ orner, G. Krugel, D. F. Urban, Acta Mat eri-\nalia153, 53 (2018).\n[6] T. Pandey, M-H. Du and D. S. Parker, Phys. Rev. Appl. 9\n034002 (2018).\n[7] L. Ke and D. D. Johnson, Phys. Rev. B 94, 024423 (2016).\n[8] J. M. D. Coey, Scr. Mater. 67, 524 (2012).\n[9] S. V . Halilov, H. Eschrig, A. Y . Perlov, P. M. Oppeneer, Ph ys.\nRev. B 58, 293 (1998).\n[10] A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg, R ev.\nMod. Phys. 68, 13 (1996).\n[11] W. K¨ orner, G. Krugel, C. Els¨ asser, Sci. Rep. 624686 92016).\n[12] H. Brooks, Phys. Rev. 58909 (1940).\n[13] P. Larson, I. I. Mazin and D. A. Papaconstantopoulos, Ph ys.\nRev. B 67214405 (2003)\n[14] S. Yehia, S. H. Aly, A. E. Aly, Computational Materials S ci-\nence 41482 (2008).\n[15] H. Ucar, R. Choudhary, D. Paudyal, J. Magn. Magn. Mater,\n496165902 (2020).\n[16] https://en.wikipedia.org/wiki/Prices −of−chemical −elements;\nJ.M.D.Coey Engineering 6, 119 (2020).\n[17] D. Odkhuu, and S. C. Hong, Phys. Rev. Appl. 11, 054085\n(2019).\n[18] D. Odkhuu, T. Ochirkhuyag, S. C. Hong, Phys. Rev. Appl. 13,054076 (2020).\n[19] Xie Xu and S. A. Shaheen, Journal of Applied Physics 73,\n5896 (1993).\n[20] H. Fujii, H. Sun. (1995). Chapter 3 Interstitially modi fied in-\ntermetallics of rare earth and 3D elements. Handbook of Mag-\nnetic Materials, 303-404.\n[21] X. Chen, Z. Altounian and D. H. Ryan, J. Magn. Magn. Mater .\n125169 (1993).\n[22] https://icsd.nist.gov/guide.html\n[23] K.H.J. Buschow, Handbook of Magnetic Materials, V olum es\n6&9 (Elsevier, 1991-1995).\n[24] J. M. D. Coey, Magnetism and Magnetic Materials (Cam-\nbridge University Press, 2010)\n[25] Shiqiang Liu, Chin. Phys. B 28, 017501 (2019).\n[26] S. R. Mishra, Gary J. Long, O. A. Pringle, D. P. Middleton , Z.\nHu et al., J. Appl. Phys. 79, 3145 (1996).\n[27] H. Klesnar, K. Hiebl and P. Rogl, Journal of the Less-Com mon\nMetals 154, 217 (1989).\n[28] R. Guetari, R. Bez, A. Belhadj, K. Zehani, A. Bezerghean u, N.\nMliki, L. Bessais, C.B. Cizmas, Journal of Alloys and Com-\npounds 588, 64 (2014).\n[29] Y . Otani, D. P. F. Hurley, Hong Sun, and J. M. D. Coey, Jour nal\nof Applied Physics 69, 5584 (1991).\n[30] M. Merches, W.E. Wallace and R.S. Craig, Journal of Mag-\nnetism and Magnetic Materials 24, 97 (1981).\n[31] F. Pourarian, R. Obermyer, Y . Zheng, S. G. Sankar, and W. E.\nWallace, Journal of Applied Physics 73, 6272 (1993).\n[32] F. Weitzer, H. Klesnar, K. Hiebl, and P. Rogl, J. Appl. Ph ys.\n67, 2544 (1990).\n[33] X.P. Zhong, R.J. Radwanski, F.R. de Boer, T.H. Jacobs an d\nK.H.J. Buschow, Journal of Magnetism and Magnetic Materi-\nals86, 333 (1990).\n[34] A. T. Pedziwiatr and W. E. Wallace, Journal of Applied\nPhysics 61, 3439 (1987).\n[35] R. van Mens, Journal of Magnetism and Magnetic Material s\n61, 24 (1986).\n[36] Hu Bo-Ping and J. M. D. Coey, Journal of the Less-Common\nMetals 142, 295 (1988).\n[37] H. Y . Chen, S. G. Sankar, and W. E. Wallace, Journal of Ap-\nplied Physics 63, 3969 (1988).\n[38] M. Juczyk and W. E. Wallace, Journal of Magnetism and Mag -\nnetic Materials 59, 182 (1986).\n[39] Y .G. Xiao, G.H. Rao, Q. Zhang, G.Y . Liu, Y . Zhang, J.K.\nLiang, Journal of Alloys and Compounds 407, 1 (2006).\n[40] E. Girt, M. Guillot, I. P. Swainson, Kannan M. Krishnan, Z.\nAltounian, and G. Thomas, Journal of Applied Physics 87,\n5323 (2000).\n[41] Wolfgang K¨ orner, Georg Krugel, and Christian Els¨ ass er, Sci.\nRep6, 24686 (2016).\n[42] A.M. Sch¨ onh¨ obel et al., Journal of Alloys and Compoun ds\n786, 969-974 (2019).\n[43] Y .Z. Wang et al., Journal of Magnetism and Magnetic Mate ri-\nals104-107 , 1132-1134 (1992).\n[44] E.P Wohlfarth and K.H.J. Buschow, Handbook of Magnetic\nMaterials, V ol-4 (Elsevier, 1988).\n[45] Y .Z. Wang et al., Journal of Applied Physics 73, 6251 (1993).\n[46] D. P. F. Hurley and J M D Coey, J. Phys. Condens. Matter 4,\n5573 (1992).\n[47] Yosuke Harashima, Kiyoyuki Terakura, Hiori Kino, Shoj i\nIshibashi, and Takashi Miyake, Phys. Rev. B 92, 184426\n(2015).\n[48] Y .G. Xiao, G.H. Rao, Q. Zhang, J. Luo, G.Y . Liu, Y . Zhang,\nand J.K. Liang, Physica B 369, 56 (2005).\n[49] V .K. Sinha, S.F. Cheng, W.E. Wallace, and S.G. Sankar, J our-15\nnal of Magnetism and Magnetic Materials 81, 227-233 (1989).\n[50] M. Jurczyk, Journal of Magnetism and Magnetic Material s89,\nL5-L7 (1990).\n[51] M. Katter, J. Wecker, C. Kuhrt, L. Schultz, X.C. Kou, and\nR. Gr¨ ossinger, Journal of Magnetism and Magnetic Material s\n111, 293-300 (1992).\n[52] M. Jurczyk and W.E. Wallace, Journal of Magnetism and Ma g-\nnetic Materials 59, L182-L184 (1986).\n[53] Yang Fu-ming, Li Qing-an, Zhao Ru-wen, Kuang Jian-ping ,\nF. R. de Boer, J. P. Liu, K. V . Rao, G. Nicolaides, and K. H. J.\nBuschow, Journal of Alloys and Compounds 177, 93 (1991).\n[54] Bao-gen Shen, Fang-wei Wang, Lin-shu Kong, Lei Cao, and\nHui-qun Guo, Journal of Magnetism and Magnetic Materials\n127, L267-L272 (1993).\n[55] X. C. Kou, T. S. Zhao, R. Grssinger, and F. R. de Boer, Phys .\nRev. B 46, 6225 (1992).\n[56] Zhi-gang Sun et al, J. Phys. Condens. Matter 12, 2495 (2000).\n[57] Z.X. Tang, E.W. Singleton, and G.C. Hadjipanayis, IEEE\nTrans. on Magn., 28, 5 (1992).\n[58] C.H. de Groot, K.H.J. Buschow, and F.R. de Boer, Physica B\n229, 213-216 (1997).\n[59] L. Zhang, D.C. Zeng, Y .N. Liang, J.C.P. Klaasse, E. Bruc k,\nZ.Y . Liu, F.R. de Boer, and K.H.J. Buschow, Journal of Mag-\nnetism and Magnetic Materials 214, 31-36 (2000).\n[60] Bing Liang, Bao-gen Shen, Fang-wei Wang, Tong-yun Zhao ,\nZhao-hua Cheng, Shao-ying Zhang, Hua-yang Gong, and\nWen-shan Zhan, Journal of Applied Physics 82, 3452 (1997).\n[61] Zhi-gang Sun, Shao-ying Zhang, Hong-wei Zhang, and Bao -\ngen Shen, Journal of Alloys and Compounds 322, 6973\n(2001).\n[62] Lin Qin, Sun Yunxi, Lan Jian, Lu Shizhong, and Jiang Hong -\nwei, Chinese Phys. Lett. 8, No. 5 , 267 (1991).\n[63] Jing-Yun Wang, Bao-Gen Shen, Shao-Ying Zhang, Wen-Sha n\nZhan, and Li-Gang Zhang, Journal of Applied Physics 87, 427\n(2000).\n[64] O. Isnard, and M. Guillot, Journal of Applied Physics 87, 5326\n(2000).\n[65] Zhi-gang Sun, Hong-wei Zhang, Shao-ying Zhang, and Bao -\ngen Shen, Physica B 305, 127134 (2001).\n[66] M.V . Satyanarayana, H. Fujii, and W.E. Wallace, Journa l of\nMagnetism and Magnetic Materials 40, 241-246 (1984).\n[67] Zhi-gang Sun, Hong-wei Zhang, Shao-ying Zhang, Jing-y un\nWang, and Bao-gen Shen, J. Phys. D: Appl. Phys. 33, 485491\n(2000).\n[68] Zhi-gang Sun, Hong-wei Zhang, Shao-ying Zhang, Jing-y un\nWang, and Bao-gen Shen, Journal of Applied Physics 87, 8666\n(2000).\n[69] E. A. Tereshina, H. Drulis, Y . Skourski, and I. S. Teresh ina,\nPhys. Rev. B 87, 214425 (2013).\n[70] Yingchang Yang, Qi Pan, Xiaodong Zhang, and Senlin Ge, J .\nAppl. Phys. 72, 2989 (1992).\n[71] Z Altounian, Xu Bo Liu, and Er Girt, J. Phys. Condens. Mat ter\n15, 33153322 (2003).\n[72] Linshu Kong, Jiabin Yao, Minghou Zhang, and Yingchang\nYang, Journal of Applied Physics 70, 6154 (1991).\n[73] O. Isnard, S. Miraglia, J.L. Soubeyroux, D. Fruchart, a nd p.\nL’Heritier, Journal of Magnetism and Magnetic Materials 137,\n151-156 (1994).\n[74] D. P. Middleton, S. R. Mishra, Gary J. Long, O. A. Pringle , Z.\nHu, W. B. Yelon, F. Grandjean, and K. H. J.Buschow, Journal\nof Applied Physics 78, 5568 (1995).\n[75] T. Pandey and David S. Parker, Sci Rep 8, 3601 (2018).\n[76] H. Luo, Z. Hu, W. B. Yelon, S. R. Mishra, G. J. Long, O. A.\nPringle, D. P. Middleton, and K. H. J. Buschow, Journal ofApplied Physics 79, 6318 (1996).\n[77] O. Isnard, S. Miraglia, D. Fruchart, j. Deportes, and P.\nL’Heritier, Journal of Magnetism and Magnetic Materials 131,\n76-82 (1994).\n[78] A.V . Andreev, D. Rafaja, J. Kamarad, Z. Arnold, Y . Homma ,\nY . Shiokawa, Journal of Alloys and Compounds 383, 4044\n(2004).\n[79] The supplementary materials contain the three dataset s for T c,\nµ0Msand Ku.\n[80] V . Psycharis, M. Anagnostou, C. Christides, and D. Niar chos,\nJournal of Applied Physics 70, 6122 (1991).\n[81] K. Ohashi, Y . Tawara, R. Osugi, and M. Shimao, Journal of\nApplied Physics 64, 5714 (1988).\n[82] Satoshi Hirosawa, Yutaka Matsuura, Hitoshi Yamamoto, Set-\nsuo Fujimura, Masato Sagawa et al,J. Appl. Phys. 59, 873\n(1986).\n[83] C. Abache and H. Oesterreicher,J. Appl. Phys. 57, 4112\n(1985).\n[84] Z. X. Tang, G. C. Hadjipanayis and V . Papaefthymiou, Jou rnal\nof Alloys and Compounds 194, 87 (1993).\n[85] Y . Z. Wang, B. P. Hu, X. L. Rao, G. C. Liu, L. Yin, W. Y . Lai,\nW. Gong, and G. C. Hadjipanayis, Journal of Applied Physics\n73, 6251 (1993).\n[86] M. Anagnostou, C. Christides and D. Niarchos, Solid Sta te\nCommunications, 78, 681 (1991).\n[87] Y . Zhang and C. Ling, npj Computational Materials 4, 25\n(2018).\n[88] Anita Halder, Aishwaryo Ghosh, and Tanusri Saha Dasgup ta,\nPhys. Rev. Materials 3, 084418 (2019).\n[89] L. Ward, A. Agrawal, A. Choudhary, and C. Wolverton, npj\nComp. Mat. 2, 16028 (2016).\n[90] Wessel N. van Wieringen, Lecture notes on ridge regress ion,\n2020, arXiv:1509.09169v5.\n[91] Vladimir V ovk, Kernel ridge regression, Empirical Inf erence,\nSpringer Berlin Heidelberg, 2013, ISBN 9783642411366.\n[92] Leo. Breiman, Random forests, Machine learning 45.1,5-32\n(2001).\n[93] Andy Liaw, and Matthew Wiener, ”Classification and regr es-\nsion by randomForest.”, R news 2.3, 18-22 (2002).\n[94] Harris Drucker, CJC Burges, Linda Kaufman, Alex J. Smol a,\nand Vladimir Vapnik, Support vector regression machines,\nAdvances in neural information processing systems, pp. 155 -\n161, 1997.\n[95] Mohamad Hassoun, Fundamentals of artificial neural net -\nworks, MIT press, 1995, ISBN 9780262514675.\n[96] Anton O. Oliynyk, Erin Antono, Taylor D. Sparks, Leila\nGhadbeigi, Michael W. Gaultois, Bryce Meredig and Arthur\nMar ,Chem. Mater. 28, 7324 (2016).\n[97] Fleur Legrain, Jes´ us Carrete, Ambroise van Roekeghem ,\nGeorg K.H. Madsen and Natalio Mingo, J. Phys. Chem. B 122,\n625 (2018).\n[98] Jess Carrete, Wu Li, and Natalio Mingo, Shidong Wang, an d\nStefano Curtarolo, Phys. Rev. X 4, 011019 (2014).\n[99] D. Andrew Carr, Mohammed Lach-hab, Shujiang Yang, Iosi f\nI. Vaisman, Estela Blaisten-Barojas, Microporous and Meso -\nporous Materials 117, 339 (2009).\n[100] Fujii, H., K. Tatami, M. Akayama and K.Yamamoto, 1992a ,\nin: Proc. 6th Int. Conf. on Ferrites, ICF6, Tokyo and Kyoto,\nJapan (The Japan Society of Powder and Powder Metallurgy,\nTokyo) p. 1081.\n[101] Fujii, H., M. Akayama, K. Nakao and K. Tatami, J. Alloys\nComp. 219, 10 (1995).\n[102] Altounian, Z., X. Chen, L.X. Liao, D.H. Ryan and J.O. St rOm-\nOlsen, J. Appl. Phys. 73, 6017 (1993).16\n[103] Chen, X., Z. Altounian and D.H. Ryan, J. Magn. Magn. Mat er.\n125, 169 (1993).\n[104] Buschow, K.H.J., T.H. Jacobs and W. Coene, IEEE Trans.\nMagn. MAG-26, 1364 (1990).\n[105] Liu, J.P., K. Bakker, ER. de Boer, T.H. Jacobs, D.B. de M ooij\nand K.H.J. Buschow, J. Less-Common Met. 170, 109 (1991).\n[106] W.G. Haije, T. H. Jacobs and K. H. J Buschow, J. Less-\nCommon Met. 163353 (1990); T. W. Capehart, R.K. Misra\nand F.E. Pickerton, App. Phys. Lett. 581395 (1991).\n[107] X. C. Kou, R. Grossinger, M. Katter, J. Wecker, L. Schul tz, T.\nH. Jacobs, K. H. J Buschow, Journal of Applied Physics, 70,\n2272(1991).\n[108] I.L.M. Locht, Y .O. Kvashnin, D.C.M. Rodrigues, M. Per eiro,\nA. Bergman, L. Bergqvist, A.I. Lichtenstein, M.I. Katsnels on,\nA. Delin, A.B. Klautau, B. Johansson, I. Di Marco, O. Eriks-\nson, Phys. Rev. B 94, 085137 (2016).\n[109] R. K. Chouhan, A. K. Pathak, D. Paudyal, V .K. Pecharsky ,\narXiv:806.01990.\n[110] B. Szpunar, Acta Phys. Pol. A 60, 791 (1981); T.-S. Zhao,\nH.-M. Jin, R. Grossinger, X.-C. Kou, and H.R. Kirchmayr, J.\nAppl. Phys. 70, 6134 (1991); I.A. Al-Omari, R. Skomski, R.A.\nThomas, D. Leslie-Pelecky, and D.J. Sellmyer, IEEE Trans.\nMagn. MAG-37, 2534 (2001).\n[111] N. Thuy, J. Franse, N. Hong, and T. Hien, J. Phys. Colloq ues\n49, 499 (1988).\n[112] P. Bruno, Phys. Rev. B 39, 865 (1989).\n[113] J. Herbst, Rev. Mod. Phys. 63, 819 (1991).\n[114] T. Miyake and H. Akai, J. Phys. Soc. Jpn 87, 041009 (2018).\n[115] Y . Hirayama, Y . Takahashi, S. Hirosawa, K. Hono, Scrip ta\nMater. 95, 70, (2015).[116] K.H.J. Buschow, Concise Encyclopedia of Magnetic\nand Superconducting Materials, Elsevier, 2005, ISBN\n9780080457659.\n[117] J.M.D. Coey, Magnetism and Magnetic Materials, 2010, ISBN\n0521816149, https://doi.org/10.1017/CBO9780511845000\narXiv:arXiv:1011.1669v3.\n[118] P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka and J .\nLuitz, WIEN2K, An Augmented Plane Wave + Loca Orbitals\nProgram for Calculating Crystal Properties (Technische Un i-\nversitt Wien, Vienna, 2001).\n[119] G. Kresse and J. Hafner, Phys. Rev. B 47, R558 (1993), G.\nKresse and J. Furthmueller, Phys. Rev. B 54, 11169 (1996).\n[120] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Let t.77,\n3865 (1996); 78, 1396(E) (1997).\n[121] A.I. Liechtenstein, V .I. Anisimov, and J. Zaanen, Phy s. Rev.\nB52, R5467 (1995).\n[122] Nico JD Nagelkerke, ”A note on a general definition of th e co-\nefficient of determination.”, Biometrika 78.3 ,691-692 (1991).\n[123] Bernhard Schlkopf, ”The kernel trick for distances”, Advances\nin neural information processing systems, 301-307 (2001).\n[124] Support Vector Regression(SVR),\nhttps://icme.hpc.msstate.edu/mediawiki/images/5/55/ SVR.pdf\n[125] I. Stephen, ”Perceptron-based learning algorithms. ”, IEEE\nTransactions on neural networks 50.2, 179 (1990).\n[126] Di Wei, Anurag Bhardwaj, and Jianing Wei, Deep Learnin g\nEssentials, Packt Publishing, 2018, ISBN 9781785880360.\n[127] Bradley Efron, and Robert J Tibshirani, An Introducti on to the\nBootstrap, CRC Press, 1994, ISBN 9780412042317." }, { "title": "2008.12061v1.Different_universality_classes_of_isostructural_U_TX__compounds___T____Rh__Co__Co___0_98__Ru___0_02_____X____Ga__Al_.pdf", "content": " Different universality classes of isostructural U TX compounds ( T = Rh, \n Co, Co 0.98Ru 0.02; X = Ga, Al) \nPetr Opletal, Vla dimír Sechovský and Jan Prokleška \nCharles University, Faculty of Mathematics and Physics, Department of Condensed Matter Physics \nKe Karlovu 5, Prague 2, Czech Republic \nAbstract \nMagnetization isotherms of the 5 f-electron ferromagnets URhGa, UCoGa and UCo 0.98Ru0.02Al \nwere measured at temperatures in the vicinity of their Curie temperature in order to investigate \nthe critical behavior near the ferromagnetic phase transition . These compounds adopt the layered \nhexagonal ZrNiAl -type structure and exhibit huge uniaxial magnetocrystalline anisotropy. The \ncritical β, γ and δ exponents were determined by analyzing Arrott -Noakes plots, Kouvel -Fisher \nplots, critical isotherms , scaling theory and Widom scaling relations. The values obtained for \nURhGa and UCoGa can be explained by the results of the renormalization group theory for a 2D \nIsing system with long -range interactions similar to URhAl reported by other investigators . On \nthe other hand, the critical exponents determined for UCo 0.98Ru0.02Al are characteristic of a 3D \nIsing ferromagnet with short -range interactions suggested in previous studies also for the \nitinerant 5f -electron paramagnet UCoAl situated near a ferromagnetic transition. The change \nfrom the 2D to the 3D Ising system is related to the gradual delocalization of 5 f electrons in the \nseries of the URhGa, URhAl, UCoGa to UCo 0.98Ru0.02Al and UCoAl compounds and appears \nclose to the strongly itinerant nonmagnetic limi t. This indicates possible new phenomena that \nmay be induced by the change of dimensionality in the vicinity of the quantum critical point. \n Introduction \nCritical phenomena have been one of the most studied issues of physics since the critical \npoints we re discovered by T. Andrews [1]. The continuous (second -order) phase transition was \nfound in systems like ferromagnets [2] to be connected with unified behavior near and at the \ncritical point which can be described by critical exponents. In the case of ferromagnets three \ncritical exponents, γ, β and δ, cha racterize the behavior near and at the critical point. They can be \ndetermined from experimental data using the relations [3]: \n \n 𝑀𝑆(𝑇) ~ |𝑡|𝛽(𝑇<𝑇𝐶) (1) \n \n𝜒(𝑇)−1 ~ |𝑡|−𝛾′(𝑇<𝑇𝐶),|𝑡|−𝛾(𝑇>𝑇𝐶) (2) \n \n 𝑀𝑠 ~ (𝜇0𝐻)1𝛿⁄𝑓𝑜𝑟𝑇 =𝑇𝐶, (3) \n \nwhere Ms is the spontaneous magnetization, χ is magnetic susceptibility and H is a magnetic \nfield. \nThe universal behavior near critical points was described by the renormalization group \ntheory first mentioned by L. P. Kadanoff [4] and then fully developed by K. G. Wilson [5–7]. \nThe universality class is determined by dimensionality of the system d, dimensionality of the \norder parameter n and the range of the interaction . There exist several universality classes in \nmagnetism. The most known are 3D Ising ( n = 1), XY model ( n = 2) and 3D Heisenberg ( n = 3). \nIt is desirable to investigat e how a certain universal critical behavior and magnetic \ndimensionality is related to the particularities of a given material (symmetry of crystal structure, \nhierarchy and anisotropy of magnetic interactions, degree of localization of “magnetic” \nelectrons, etc.). Large groups of isostructural compounds containing transition -element ions with \none type of “magnetic” d or f electrons provide useful playgrounds for investigation of these \naspects. The UTX compounds ( T - transition metal, X – p-metal) crystalliz ing in the hexagonal \nZrNiAl -type structure constitute such a suitable group of materials [8]. The crystal structure \nconsi sts of U -T and T-X basal plane layers alternating along the c-axis. The strong bonding of 5 f-\nelectron orbitals within the U -T layer in conjunction with strong spin -orbit interaction leads to a \nhuge uniaxial magnetocrystalline anisotropy that locks the U ma gnetic moments in the c-axis and \nthus makes these materials suitable for investigating Ising systems. \nThe critical magnetic behavior has so far been studied on two compounds of this \nisostructural group, UCoAl [9] and URhAl [10]. UCoAl, is an itinerant 5 f electron paramagnet \nundergoing , at low temperatures , a metamagnetic transition with a critical field of ~ 0.7 T [11]. \nKarube et al. reported that it behaves near to the critical endpoint as a 3D Ising system with \nshort -range interactions [9]. On the other hand, URhAl was reported behaving as a 2D Ising \nferromagnet with long -range interactions [10]. The observed difference of magnetic \ndimensionality of UCoAl and URhAl indicates that the common layered hexagonal crystal \nstructure and uniaxial magnetocrystalline anisotropy shared by all compounds of the U TX family \n(T – transition metal, X = Al, Ga, Sn, In) [8,11] is not a sufficient condition for sharing also a \ncommon magnetic universality class. To test this aspect we prepared single crystals of three hexagonal U TX ferromagnets – \nUCoGa [11,13 –15], URhGa [16–18] and UCo 0.98Ru0.02Al (a close analog of UCo 0.99Ru0.01Al \n[19,20] ) and measured thei r magnetization isotherms in the neighborhood of their Curie \ntemperature . The measured data were analyzed by investigating Arrott -Noakes plots [21,22] , \nKouvel -Fischer plots [23], critical isotherms , scaling theory [24] and Widom scaling relations \n[25] in order to determine the critical exponents and their correspon ding universality classes. The \nfirst two compounds were found behaving as 2D Ising systems similar t o previously reported \nURhAl [10] whereas the critical exponents determined for UCo 0.98Ru0.02Al point to the 3D Ising \nuniversality class similar to UCoAl [9]. It is discussed that the change between these two classes \nof universality is associated with a changing degree of delocalization of 5f electrons . \n \n \nExperimental \nSingle crystals of URhGa, UCoGa and UCo 0.98Ru0.02Al were prepared from stoichiometric \nmelts by Czochralski method using a triarc furnace. Each single crystal was wrapped in a \ntantalum foil and sealed in a quartz tube in evacuated to 10-6 mbar. The EDX analysis \naccomplished with a scanning electron microscope Tescan Mira I LMH equipped by a \nbackscatter electrons detector confirmed the stoichiometric composition of URhGa, UCoGa and \nUCo 0.98Ru0.02Al. The X -ray Laue method (Laue diffractometer of Photonic Science) shown good \nquality of the crystals. Samples for magnetization w ere cut by a wire saw to a rectangular prism \nshape. The c-axis magnetization M was measured using an MPMS -7-XL (Quantum Design) in \nfields from 0.1 T to 4 T at different temperatures in the vicinity of Curie temperature. \n The values of internal magnetic field H were calculated as: \n \nH = H a - NM (4), \n \nwhere N is the demagnetization factor. The demagnetization factor was evaluated using the \nformula published by Aharoni [26]. \n \nResults and Discussion \n \nURhGa and UCoGa 0 2 4 6 8 10 12 14 160246810\n M1/ 1011 (A/m)1/ \n(H/M)1/T = 0.4 KTC= 41.01 \n 0.02 K\n = 0.40 \n 0.01\n = 1.18 \n 0.01\nT= 39 K\nT= 43 K \n0 5 10 15 20024681012141618\nT = 47.8 KM1/ 1012 (A/m)1/ \n(H/M)1/T = 44.6 KT = 0.4 KTC= 46.29 \n 0.03 K\n= 1.26 \n 0.01\n = 0.37 \n 0.01 \nFig. 1: Modified Arrott plots for URhGa (left) and UCoGa (right). Resulting critical exponents and Curie temperature \ncorresponding to each compound are in the figures. Magnetic curves close to Curie temperature were used for fit of Arrott -\nNoakes equation. \nThe C urie temperature of a ferromagnet is usually determined by Arrott plot ( M2 vs H/M \nplot) analysis of magnetization isotherms [21] measured at temperatures in the critical region. \nThe linear Arrott plots are, in fact, a graphical representation of the Landau equation of state in \nthe theory of second -order phase transition [27–29], which has been later specified for \nferromagnets by Ginsburg [30]. This approach is certainly suitable for investigation of \nhomogenous isotropic ferromagnets. This condition is not met for most real systems, whi ch is \ndocumented by considerable curvature in Arrott plots. Th e value of Curie temperature TC can be \nthe refined by finding the and coefficients for which modified Arrott plots (Arrott -Noakes \nplots) M1/β vs. H/M1/γ are linear. The Arrott -Noakes plot comes from the analysis of \nmagnetization curves based on the Arrott -Noakes equation [22]: \n \n(𝐻𝑀⁄)1𝛾⁄=(𝑇−𝑇𝐶)\n𝑇1+(𝑀𝑀1⁄ )1𝛽⁄ (5), \n \nwhere M1 and T1 are material constants that are temperature independent in the vicinity of the \nphase transition. Magnetization curves are then fitted by (5) to get the M1/β vs. H/M1/γ plots linear \nand parallel by varying free parameters β and γ while keeping the T1 and M1 values fixed. The \nresults for our studied compounds are presented in Fig. 1 and Table 1 together with the best \nvalues of TC and critical exponents β, γ. \n 0.1 14681012\nT = 43 K\n M 104 (A/m)\n0H (T)T = 39 KT = 0.4 K = 3.89 \n 0.01\n0.1 12345678\nT = 0.4 K\nT = 47.8 K M 105 (A/m)\n0H (T)T = 45 K= 4.32 \n 0.02 \nFig. 2: Logarithmic plot of magnetic isotherms of URhGa (left) and UCoGa (right). Critical isotherms for each compound are \nhighlighted by dashed line. The critical exponent δ from fit to critical exponent for each compound are part of the figure. \nThe critical exponent δ can be calculated from β and γ from the Widom scaling law [25]. It \ncan be determined from the magnetization curve using Eq. (3), too. In Fig. 2 we display the \nmagnetization curves of both compounds in a log -log plot because only the isotherm at TC is \nlinear in this represen tation, as seen from Eq. (3). A linear function is then fitted to data points \nforming the straightest isotherm. The resulting δ-values are shown in Fig. 2 and Table 1. For \nURhGa , UCoGa the δ-value calculated from the Widom scaling law [25] is 3.95 , 4.4, which is \nclose enough to the value of 3.89 (4.5) shown in Fig. 2. The agreement between values from the \nWidom scaling law and values from cri tical isotherms could be improved as we did not know the \nexact TC during measurement, and therefore we used isotherms at temperature closest to the \nassumed TC. \nThe β- and γ-values can be further refined by analyzing Kouvel -Fisher plots [23] which are \nbased on the definition of critical exponents in Eqs. (1) and (2). The spontaneous magnetization \nMS is determined from the intersection of the M1/β axis of a Arrott -Noakes plot with straight lines \nand the inverse susceptibility χ-1 is determ ined from the intersections of straight lines with the \n(H/M)1/γ axis of the Arrott -Noakes plot. Kouvel and Fisher [23] showed that by dividing by \ntemperature derivatives of Eqs. (3) and (4) one gets a new set of equations \n \n𝑀𝑆(𝑇)[𝑑𝑀𝑆(𝑇)𝑑𝑇⁄ ]−1=|𝑇−𝑇𝐶|/𝛽(𝑇) (6) \n \n𝜒−1(𝑇)[𝑑𝜒−1(𝑇)𝑑𝑇⁄ ]−1=|𝑇−𝑇𝐶|/𝛾(𝑇). (7) \n \nIn the vicinity of TC the β(T) and γ(T) became equal to the corresponding critical β and γ \nvalues. The critical exponents are determined from the slope and the Curie temperature is \ndetermined from the intersection with the T axis in the Kouvel -Fisher plots shown in Fig. 3. The \nresulting critical exponents for URhGa and UCoGa are presented in Fig. 3 and Table 1. 39 40 41 42 43012345\nT +\nC = 40.90 \n 0.01\n = 1.21 \n 0.01\n fit to MS\n fit -1\nT (K)MS(dMS/dT)-1 (K)T -\nC = 40.92 \n 0.06\n = 0.42 \n 0.02\n0.00.51.01.52.0\n -1 (d-1/dT)-1 (K)\n44.5 45.0 45.5 46.0 46.5 47.0 47.5 48.012345 \nT (K)MS/(dMS/dT) (K)T +\nC= 46.33 \n 0.05 K\nT -\nC= 46.43 \n 0.06 K\nTC= 46.38 \n 0.06 K\n = 1.29 \n 0.07\n = 0.40 \n 0.02\n0.00.20.40.60.81.01.21.4\n -1/(d-1/dT) (K) \nFig. 3: Kouvel -Fisher plot for spontaneous magnetization and susceptibility for URhGa (left) and UCoGa (right). The critical \nexponents are shown in corresponding figures. \nTo check whether the critical exponents above and below the phase transition ar e equal, we \nhave separately determined the critical exponent γ (T > TC) and γ’ (T < TC) using the scaling \ntheory that predicts a reduced equation of state close to the phase transition [24]: \n \n𝑀(𝜇0𝐻,𝑡)=|𝑡|𝛽𝑓±(𝜇0𝐻|𝑡|𝛽+𝛾⁄ ), (8) \n \nwhere f+ is for T > TC and f- is for T < TC and both are regular analytical functions. If the correct \nβ, γ and TC values are chosen then the data points in the plot M(µ0H,t)/|t|β versus µ0H/|t|β+γ should \nfall on two universal curves, one for T < TC and second for T > TC and these curves should \napproach each other asymptotically. Magnetization data for both compounds were fitted by Eq. \n(8). The results are shown in Fig. 4 together with the values of critical exponents and TC, \nrespectively. For both compounds, the difference of the determined critical exponent γ for T \nabove and below TC is very small which corroborates the presumption that it does not change. \n \n10 100 1000 10000110 M/|t| 105 (A/m)\n0H/|t|(+) (T) TTCTC= 40.85 K\n = 0.39 \n 0.01\n+=1.18 \n 0.01\n-=1.19 \n 0.02\n0.01 0.1 1 10 100110\n TTC\n M/|t| 105 (A/m)\n0H/|t|(+) (T)TC= 46.23 \n 0.04\n= 0.37 \n 0.01\n+= 1.28 \n 0.03\n-= 1.26 \n 0.02\n \nFig. 4: The scaled magnetization plotted against the renormalized magnetic field below and above T C. Results of fit of scaling \ntheory from Eq. (8) for both compounds are shown in figures URhGa (left) and UCo Ga (right). \nThe critical exponents of URhGa and UCoGa can be explained similarly to the URhAl case \n[10] by introducing a weak long -range magnetic exchange interaction in form J(r) ~ r-(d+σ), where \nσ is the range of the exchange interaction [31]. Fischer et al. [31] applied the renor malization group theory for a system with an interaction J(r) for which the equation for γ has been obtained \nin the form: \n \n𝛾=1+4\n𝑑(𝑛+2\n𝑛+8)∆𝜎+8(𝑛+2)(𝑛−4)\n𝑑2(𝑛+8)2[1+2𝐺(𝑑\n2)(7𝑛+20)\n(𝑛−4)(𝑛+8)]∆𝜎2, (9) \n \nwhere ∆𝜎=(𝜎−𝑑\n2) and 𝐺(𝑑\n2)=3−1\n4(𝑑\n2)2\n. The obtained values of γ were examined using Eq. \n(9) by substituting the possible values of d (dimension of the system ) = 1, 2, or 3, and n \n(dimension of the order parameter ) = 1, 2, or 3 in all possible combinations and comparing the \nresulting values of σ for γ and β. The best agreement for both, URhGa and UCoGa, has been \nfound for a 2D Ising system with LR interactions. The resulting σ value and corresponding β, γ \nand δ values are listed in Table 1. The same universality c lass was reported for URhAl [10]. \n \nUCo 0.98Ru0.02Al \nThe critical exponents γ = (1.27 ± 0.01) and β = (0.32 ± 0.01) determined by Arrott -Noakes \nanalysis of magnetization data collected on the UCo 0.98Ru0.02Al single crystal (see Arrott -Noakes \nplots in Fig. 5) with TC = 22.79 K compare well to the theoretical values γ = 1.241 and β = 0.325 \nfor the 3D Isin g model [24,32] . \n \n0 5 10 15 20 25012345678\nT = 0.1 K T = 24.5 K \n M1/ 1014 (A/m)1/\n(H/M)1/T = 22 K = 1.27 \n 0.01\n = 0.32 \n 0.01\nTC = 22.79 \n 0.03 K\n0.1 13456\nT = 0.1 KT = 23.6 KTC = 22.8 K\n = 4.92 \n 0.01M 105 (A/m)\n0H (T)T = 22 K\n \nFig. 5: Arrott -Noakes plots (left) and logarithmic plots (right) of magnetization isotherms for UCo 0.98Ru0.02Al. The results of the \nfit by Arrott -Noakes equation (Eq. ( 5)) are displayed in left figure while in right figure the result of the fit by critical isotherm can \nbe found. \nFurther on, the value δ = (4.92 ± 0.01) corresponding to the critical isotherm (see Fig. 5) is \nclose to the value of 4.97 determined from Widom scaling relations using γ and β determined \nfrom Arrott -Noakes equation. The value corresponding to the critical isotherm is close to th e \nvalue of 4.82 known for the 3D Ising model. \nTo further analyze the UCo 0.98Ru0.02Al magnetization data , the Kouvel -Fisher method (for \nthe respective plot see Fig. 6) was used in a similar way as above. The resulting critical \nexponents γ = (1.24 ± 0.04) and β = (0.33 ± 0.01) with TC = (22.79 ± 0.02) compare to values for \nthe 3D Ising model. 22.2 22.4 22.6 22.8 23.0 23.2 23.40.00.30.60.91.21.51.8\n = 1.24\n 0.04\nT +\nC = 22.79\n 0.01 \n fit to MS\n fit to -1\nT (K)MS/(dMS/dT) (K)\n = 0.33\n 0.01\nT -\nC = 22.79\n 0.02\n0.00.10.20.30.40.5\n -1/(d-1/dT) (K)\n1 10 100 100012345 M/t 105 (A/m)\nH/t(+) 107 (A/m) TTCTC = 22.74 K\n = 1.241\n = 0.325 \nFig. 6: (Left) Kouvel -Fisher plot of UCo 0.98Ru0.02Al with results of fit of Eqs. (6) and (7). (Right) Plot of the scaled magnetization \nvs. the rescaled magnetic field for uCo0.98Ru0.02Al. Results of fit of scaling theory from Eq. (8) are part of the figure. \nWe have confirmed that UCo 0.98Ru0.02Al behaves as a 3D Ising system by plotting \nmagnetization data using the reduced equation of state in eq. 8 with values of β and γ for the 3D \nIsing model in Fig. 6. The data points fall on two different curves, for T below and above TC = \n22.74 K, respectivel y, which approach asymptotically to merge. The 3D Ising behavior \nUCo 0.98Ru0.02Al resembles behavior of pure UCoAl reported by Karube et al. in [9]. \nThe values of critical exponents and other relevant information obtained in the present \nwork on URhGa, UCoGa and UCo 0.98Ru0.02Al single crystals are displayed in Table 1. The \ninformation available for UCoAl [9] is included for comparison and further discussion. \n \nTable I: Critical exponents and other relevant parameters obtained for URhGa, UCoGa and \nUCo 0.98Ru0.02Al by the analysis of Arrott -Noakes plots (A -N plots), Kouvel -Fisher plots (K -F \nplots), scaling relations (Scaling), critical isotherms (Critical i.)and data for UCoAl [9] for \ncomparison. \n Method TC (K) γ'(T < TC) γ(T > TC) δ σ \nURhGa A-N plots 41.01 ± 0.02 0.40 ± 0.01 1.18 ± 0.01 \n K-F plots 40.91 ± 0.04 0.42 ± 0.02 1.21 ± 0.01 \n Scaling 40.85 ± 0.02 0.39 ± 0.01 1.18 ± 0.01 1.19 ± 0.02 \n Critical i. 3.89 ± 0.01 \nLR exchange: \nJ(r) ~ r-(d+σ) \nd = 2, n = 1 0.39 1.18 4.03 1.21 \nUCoGa A-N plots 46.29 ± 0.03 0.37 ± 0.01 1.26 ± 0.01 \n K-F plot 46.38 ± 0.06 0.40 ± 0.02 1.29 ± 0.07 \n Scaling 46.23 ± 0.04 0.37 ± 0.01 1.28 ± 0.03 1.26 ± 0.02 \n Critical i. 4.32 ± 0.01 \nLR exchange: \nJ(r) ~ r-(d+σ) \nd = 2, n = 1 0.36 1.26 4.5 1.28 \nThe table presents the most striking result of our study - although the three studied \nUTX ferromagnets adopt the same type of layered hexagonal crystal structure with strong \nuniaxial magnetocrystalline anisotropy they do not fall in the same universality class. URhGa and \nUCoGa exhibit a 2D Ising character similar to URhAl [10] whereas UCo 0.98Ru0.02Al behaves as \nta 3D Ising system , also reported for UCoAl [9]. \n The 5f -electrons in U intermetallics are known to have a dual character (partially localized, \npartially itinerant) [33–35]. The localized and itinerant character s appear in different proportions \ndepending on crystallographic and chemical environments of U ion being reflected in a wide \nrange of their magnetic behavior. This is a consequence of the wide exte nsion of the 5f-wave \nfunctions which allow s for considerable direct overlaps between 5f-wave functions of the \nnearest -neighbor U ions as well as hybridization of valence electron states of ligands (5f -ligand \nhybridization). As a result, the original atomic character of the 5f -wave functions is destroyed \nwhile the related magnetic moments is washed out and adequately reduced. In the strong 5f -5f \noverlap and 5f -ligand hybridization limits the 5f -electrons are predominantly itinerant, the 5f \nmagnetic moments vanish and the magnetic order is lost (UCoAl in our case). \nRhodes and Wohlfarth propose d that the ratio µeff /µs between the effective and spontaneous \nmagnetic moment can be taken as a measure of the degree of itinerancy of the magnetic electrons \n[36,37] . In Table 2, we can see that µeff /µs increases along the series as listed from top to \nbottom. In the Rhodes -Wohlfarth scenario, the degree of itinerancy of 5f -electrons increases \nwhen proceedi ng from URhGa towards UCoAl. \nWe can also see that the lattice parameter a shown in the same table simultaneously \ndecreases along the series. The close packing of U and T atoms in the basal plane of the \nhexagonal ZrNiAl -type structure results both in a non -negligible 5f -5f overlap and a strong 5f -d \nhybridization involving the transition metal d states which compress most of the 5f charge \ndensity towards the basal plane. The reduction of a is intimately co nnected with the decreasing \nU-U and U -T interatomic distances within the basal plane. The simultaneously enhanced 5f -5f \noverlaps and 5f -d hybridization causes a higher degree of itinerancy, which corroborates the \nRhodes -Wohlfarth scenario. \nThe change from 2D to three dimensional behavior was described by Takahashi , who \nconsiders quasi - itinerant ferromagnets in [38]. In a quasi -2D system the spin fluctuation in the z-\ndirection differ s from spin fluctuations in the xy plane , which comes from the uniaxial \nmagnetocrystalline anisotropy of the system. The critical behavior of the system in the region of \nthe T-H space centered on [TC, 0] is described as 3D, whil e outside of this region the critical \nbehavior is 2D. For stronger anisotropy , the region of 3D behavior is reduced . The l arge \nmagneto crystalline anisotropy observed in U TX compounds crystallizing in the ZrNiAl -type \nstructure [8,11,15,39] leads to suppression of a 3D behavior and instead a 2D behavior is UCo 0.98Ru0.02Al A-N plots 22.79 ± 0.03 0.32 ± 0.01 1.27 ± 0.01 \n K-F plots 22.79 ± 0.02 0.33 ± 0.01 1.24 ± 0.04 \n Scaling 22.74 ± 0.02 0.325 1.241 \n Critical i. 4.92 ± 0.01 \nUCoAl NMR \nmeas. - 0.26 1.2 5.4 observed as seen in URhGa, UCoGa and URhAl. It is important to emphasiz e that the system s \ndescribed by Takahashi show different critical behavior s, but the effect of anisotropy on the \nnature of fluctuations near TC is expected to be similar. \n \nTable 2: The values of effective and spontaneous magnetic moment, µeff and µs, respectively, and \nthe µeff /µs ratio and lattice parameters for our studied URhGa, UCoGa and UCo 0.98Ru0.02Al \ncompounds completed by the values for URhAl and UCoAl. The “ µs value” for UCoAl is the \nmagnetic moment in the field just above the metamagnetic transition. The true µs value for \nUCoAl is indeed equal to 0; the ground state is paramagnetic. \nCompound µeff \n(µB/f.u.) µs \n(µB/f.u.) µeff \n/µs Ref. Universality \nclass a \n(pm) c \n(pm) Ref. \nURhGa 2.45 1.17 2.11 [40] 2D Ising 700.6 394.5 [41] \nURhAl 2.50 1.05 2.38 [10] 2D Ising 696.5 401.9 [41] \nUCoGa 2.40 0.65 3.69 [15] 2D Ising 669.3 393.3 [41] \nUCo 0.98Ru0.02Al 1.73 0.36 4.81 * 3D Ising 669.1 396.6 * \nUCoAl 1.60 0.30 5.33 [11] 3D Ising 668.6 396.6 [42] \n*) unpublished data \n \nFurther inspection of Table 2 reveals that the change from 2D Ising to 3D Ising universality \nclass happens when the 5f -electrons become considerably itinerant. This agrees with Takahashi’s \ndescription in [38], in which for more itinerant system s larger anisotropy is nee ded to suppress \nthe region of 3D behavior compared to more localized systems . This explains the 3D behavior of \nUCoAl and UCo 0.98Ru0.02Al even though the magnetocrystalline anisotropy is comp arable to that \nin URhGa, UCoGa and URhAl [8,11,15,39] . Also, we see that the p -metal affects the degree of \nlocalization/itineracy, which in turn affects the universality class of the system. The hierarchy of \nexchange interactions will undoubtedly play an essential role in controlling dimensionality. The \ninvolvement of theorists in resolving these issues is strongly desirable. It would also be \ninteresting to investigate the evolution of critical exponents and magnetic dimensionality \ntogether with the de velopment of lattice parameters while applying hydrostatic pressure. A \npossible change of magnetic dimensionality with applied pressure near to a quantum critical \npoint may open new questions in the quantum criticality research. \n \nConclusions \nThe magnetization isotherms of their URhGa, UCoGa and UCo 0.98Ru0.02Al were measured at \ntemperatures near to their ferromagnetic transition s where we investigated the critical behavior. \nThe values of critical exponents β, γ and δ have been determined by ana lyzing magnetization \ndata presented in Arrott -Noakes plots, Kouvel -Fisher plots, critical isotherms , scaling theory and \nWidom scaling relations. The results point to the 2D Ising uni versality class for URhGa and \nUCoGa, similar to URhA l reported by other investigators [10]. Data obtained for \nUCo 0.98Ru0.02Al are c haracteristic of the 3D Ising universality class which was suggested in \nprevious studies also for the itinerant 5f -electron paramagnet UCoAl. At the high degree of \nitinerancy of 5f -electrons a change from the 2D to the 3D character is observed between UCoGa \nand UCo 0.98Ru0.02Al. Possible new phenomena may be expected when a dimensionality change \nhappens in the vicinity of the quantum critical point. \nAcknowledgments \n This research was supported by Grant Agency of Charles University, grant No. 1630218 and by \nthe Czech Science Foundation, grant No.16 -06422S. Experiments were performed in MGML \n(www.mgml.eu), which is supported within the program of Czech Research Infrastructures \n(project no. LM2018096). It was also supported by OP VVV project MATFUN unde r Grant \nCZ.02.1.01/0.0/0.0/15_003/0000487. We would also like to thank to Dr. \nRoss Colman for proofreading of the text, and language corrections. \n \nReferences \n[1] T. Andrews, XVIII. The Bakerian Lecture . —On the continuity of the gaseous and liquid \nstates of matter, Philos. Trans. R. Soc. London. 159 (1869) 575 –590. \n[2] J. Hopkinson, I. Magnetic properties of alloys of Nickel and iron, Proc. R. Soc. London. \n48 (1891) 1 –13. \n[3] M.E. Fisher, The theory of equilibriu m critical phenomena, Reports Prog. Phys. 30 (1967) \n306. \n[4] L.P. Kadanoff, Scaling laws for ising models near T c , Phys. Phys. Fiz. Phys. \nФизика. 2 (1966) 263 –272. \n[5] K.G. Wilson, Renormalization group and critical phenomena. I. Renormalization group \nand the Kadanoff scaling picture, Phys. Rev. B. 4 (1971) 3174 –3183. \n[6] K.G. Wilson, Renormalization group and critical phenomena. II. Phase -space cell analysis \nof critical behavior, Phys. Rev. B. 4 (1971) 3184 –3205. \n[7] K.G. Wilson, The renormaliza tion group: Critical phenomena and the Kondo problem, \nRev. Mod. Phys. 47 (1975) 773 –840. \n[8] V. Sechovsky, L. Havela, Chapter 1 Magnetism of ternary intermetallic compounds of \nuranium, in: K.H.J. Buschow (Ed.), Handb. Magn. Mater., Elsevier, 1998: pp. 1 –289. \n[9] K. Karube, T. Hattori, S. Kitagawa, K. Ishida, N. Kimura, T. Komatsubara, Universality \nand critical behavior at the critical endpoint in the itinerant -electron metamagnet UCoAl, \nPhys. Rev. B. 86 (2012). \n[10] N. Tateiwa, J. Pospíšil, Y. Haga, E. Yama moto, Critical behavior of magnetization in \nURhAl: Quasi -two-dimensional Ising system with long -range interactions, Phys. Rev. B. \n97 (2018) 64423. \n[11] V. Sechovsky, L. Havela, F.R. de Boer, J.J.M. Franse, P.A. Veenhuizen, J. Sebek, J. \nStehno, A. V Andreev , Systematics across the UTX series (T = Ru, Co, Ni; X = Al, Ga, \nSn) of high -field and low -temperature properties of non -ferromagnetic compounds, Phys. \nB+C. 142 (1986) 283 –293. \n[12] N. Tateiwa, J. Pospíšil, Y. Haga, H. Sakai, T.D. Matsuda, E. Yamamoto, Itinerant \nferromagnetism in actinide 5f -electron systems: Phenomenological analysis with spin \nfluctuation theory, Phys. Rev. B. 96 (2017) 35125. \n[13] A. V Andreev, A. V Deryagin, R.Y. Yumaguzhin, Crystal structure and magnetic \nproperties of UGaCo and UGaNi single crystals, Sov. Phys. - JETP. 59 (1984) 1082 –1086. \n[14] P. Opletal, P. Proschek, B. Vondráčková, D. Aurélio, V. Sechovský, J. Prokleška, Effect \nof thermal history on magnetism in UCoGa, J. Magn. Magn. Mater. 490 (2019) 165464. \n[15] H. Nakotte, F.R. de Boer, L. Havela, P. Svoboda, V. Sechovsky, Y. Kergadallan, J.C. \nSpirlet, J. Rebizant, Magnetic anisotropy of UCoGa, J. Appl. Phys. 73 (1993) 6554 –6556. \n[16] V. Sechovsky, F. Honda, K. Prokes, O. Syshchenko, A. V Andreev, J. Kamarad, Pressure -induced phe nomena in U intermetallics, Acta Phys. Pol. B. 34 (2003) 1377 –1386. \n[17] V. Sechovský, L. Havela, N. Pillmayr, G. Hilscher, A. V Andreev, On the magnetic \nbehaviour of UGaT series, J. Magn. Magn. Mater. 63 –64 (1987) 199 –201. \n[18] V. Sechovsky, L. Havela, F. R. de Boer, P.A. Veenhuizen, K. Sugiyama, T. Kuroda, E. \nSugiura, M. Ono, M. Date, A. Yamagishi, Hybridization and magnetism in U(Ru, Rh)X, \nX=Al, Ga, Phys. B Condens. Matter. 177 (1992) 164 –168. \n[19] A. V Andreev, L. Havela, V. Sechovsky, M.I. Bartashevich, J. Sebek, R. V Dremov, I.K. \nKozlovskaya, Ferromagnetism in the UCo1 -xRuxAl quasiternary intermetallics, Philos. \nMag. B -Physics Condens. Matter Stat. Mech. Electron. Opt. Magn. Prop. 75 (1997) 827 –\n844. \n[20] P. Opletal, J. Prokleška, J. Valenta, P. Proschek , V. Tkáč, R. Tarasenko, M. Běhounková, \nŠ. Matoušková, M.M. Abd -Elmeguid, V. Sechovský, Quantum ferromagnet in the \nproximity of the tricritical point, Npj Quantum Mater. 2 (2017). \n[21] A. Arrott, Criterion for ferromagnetism from observations of magnetic i sotherms, Phys. \nRev. 108 (1957) 1394 –1396. \n[22] A. Arrott, J.E. Noakes, Approximate equation of state for nickel near its critical \ntemperature, Phys. Rev. Lett. 19 (1967) 786 –789. \n[23] J.S. Kouvel, M.E. Fisher, Detailed magnetic behavior of nickel near its curie point, Phys. \nRev. 136 (1964). \n[24] V. Privman, P.C. Hohenberg, A. Aharony, Universal Critical -Point Ampl itude Relations, \nin: C. Domb, J.L. Lebowitz (Eds.), Phase Transitions Crit. Phenom. Vol. 14, ACADEMIC \nPRESS LIMITED, London, 1991: p. 367. \n[25] B. Widom, Equation of State in the Neighborhood of the Critical Point, J. Chem. Phys. 43 \n(1965) 3898 –3905. \n[26] A. Aharoni, Demagnetizing factors for rectangular ferromagnetic prisms, J. Appl. Phys. \n83 (1998) 3432 –3434. \n[27] L.D. Landau, On the theory of phase transitions. I., Zh. Eksp. Teor. Fiz. 7 (1937) 19. \n[28] L.D. Landau, On the theory of phase transitions. I., Phys. Z. Sowjet. 11 (1937) 26. \n[29] D. TER HAAR, On the Theory of Phase Transitions, in: Men Phys. L.D. Landau, \nElsevier, 1969: pp. 61 –84. \n[30] V.L. Ginzburg, Behavior of ferromagnetic substances in the vicinity of the curie point, Zh. \nEksp. Teor. Fiz. 17 (1947) 833. \n[31] M.E. Fisher, S.K. Ma, B.G. Nickel, Critical exponents for long -range interactions, Phys. \nRev. Lett. 29 (1972) 917 –920. \n[32] J.C. Le Guillou, J. Zinn -Justin, Critical exponents from field theory, Phys. Rev. B. 21 \n(1980) 3976 –3998. \n[33] T. Takahashi, N. Sato, T. Yokoya, A. Chainani, T. Morimoto, T. Komatsubara, Dual \nCharacter of 5f Electrons in UPd2Al3 Observed by High -Resolution Photoemission \nSpectroscopy, J. Phys. Soc. Japan. 65 (1996) 156 –159. \n[34] G. Zwicknagl, 5f Electron correlations and core level photoelectron spectra of Uranium \ncompounds, Phys. Status Solidi. 250 (2013) 634 –637. \n[35] G. Zwicknagl, 3 = 2+1: Partial localiz ation, the dual character of 5f electrons and heavy \nfermions in U compounds, in: J. Magn. Magn. Mater., North -Holland, 2004: pp. E119 –\nE120. \n[36] P. Rhodes, E.P. Wohlfarth, The effective Curie -Weiss constant of ferromagnetic metals \nand alloys, Proc. R. Soc. Lond. A. 273 (1963) 247 –258. [37] E.P. Wohlfarth, Magnetic properties of crystallineand amorphous alloys: A systematic \ndiscussion based on the Rhodes -Wohlfarth plot, J. Magn. Magn. Mater. 7 (1978) 113 –120. \n[38] Y. Takahashi, Spin -fluctuation theory of quasi -two-dimensional itinerant -electron \nferromagnets, J. Phys. Condens. Matter. 9 (1997) 10359 –10372. \n[39] P.A. Veenhuizen, F.R. de Boer, A.A. Menovsky, V. Sechovsky, L. Havela, MAGNETIC -\nPROPERTIES OF URuAl and URhAl SINGLE -CRYSTALS, J. Phys. 49 (1988) 48 5–\n486. \n[40] P. Opletal, Peculiarities of magnetism on the verge of ferromagnetic ordering, Faculty of \nMathematics and Physics, Charles University, 2019. \n[41] A.E. Dwight, Alloy Chemistry of Thorium, Uranium, and Plutonium Compounds, in: Dev. \nStruct. Chem. Alloy Phases, Springer US, 1969: pp. 181 –226. \n[42] D.J. Lam, J.B. Darby, J.W. Downey, L.J. Norton, Equiatomic ternary compounds of \nuranium and aluminium with group viii transition elements, J. Nucl. Mater. 22 (1967) 22 –\n27. \n " }, { "title": "2009.00410v1.Lattice_dynamics_effects_on_the_magnetocrystalline_anisotropy_energy__application_to_MnBi.pdf", "content": "Lattice dynamics effects on the magnetocrystalline anisotropy energy: application to\nMnBi\nAndrea Urru1and Andrea Dal Corso1,2\n1International School for Advanced Studies (SISSA),\nVia Bonomea 265, 34136 Trieste (Italy).\n2DEMOCRITOS IOM-CNR Trieste (Italy).\n(Dated: September 2, 2020)\nUsing a first-principles fully relativistic scheme based on ultrasoft pseudopotentials and density\nfunctional perturbation theory, we study the magnetocrystalline anisotropy free energy of the fer-\nromagnetic binary compound MnBi. We find that differences in the phonon dispersions due to the\ndifferent orientations of the magnetization (in-plane and perpendicular to the plane) give a differ-\nence between the vibrational free energies of the high-temperature and low-temperature phases.\nThis vibrational contribution to the magnetocrystalline anisotropy energy (MAE) constant, Ku, is\nnon-negligible. When the energy contribution to the MAE is calculated by the PBEsol exchange\nand correlation functional, the addition of the phonon contribution allows to get a T= 0KKuand\na spin-reorientation transition temperature in reasonable agreement with experiments.\nI. INTRODUCTION\nRecently, there has been a significant effort toward the\nrealization of rare-earth-free permanent magnets1,2. Due\nto its magnetic properties, such as a high Curie tempera-\nture, well above room temperature, and a large uniaxial\nmagnetic anisotropy, MnBi2–5has emerged as a promis-\ning candidate among the transition-metal-based materi-\nals.\nBelow the Curie temperature, estimated to be Tc=\n680K6,7, MnBi is a ferromagnet which crystallizes in\nthe NiAs structure. Its magnetocrystalline anisotropy\nenergy (MAE) as a function of temperature is peculiar:\natT= 0KtheMAEconstant Kuisnegative,itsreported\nexperimental value being \u00000:2MJ / m3(\u0019\u00000:12meV /\ncell) with an easy axis in the basal plane8, and increasing\nwithT, unlike most magnetic systems4,8. AtT\u001990K\nKubecomes positive, thus leading to a spin-reorientation\ntransition: from 90K to 140K, the easy axis rotates\noutside the basal plane, and above 140K it is parallel to\nthe c axis9.\nSeveral studies, both experimental and theoretical,\nhave been carried out during the past years to under-\nstand this intriguing property. Experiments studied sev-\neralpropertiesofMnBi, includingthethermalexpansion.\nIn particular, the spin-reorientation transition comes to-\ngether with a small kink in the lattice parameters at\nT\u001990K10,11, which has been interpreted as the sign\nof a phase transition. Theoretical calculations, based on\nDensity Functional Theory (DFT) within the Local Den-\nsity Approximation (LDA) and the Generalized Gradient\nApproximation(GGA)fortheexchange-correlationfunc-\ntional, correctly predict MnBi to be a metal and a ferro-\nmagnet in the low-temperature phase, and to have a neg-\nativeKu, in agreement with experiments. Yet, they are\nbelieved to overestimate the magnitude of Kuby nearly\nan order of magnitude and are often not able to repro-\nduce the correct behavior of Kuas a function of tem-\nperature. Refs. 12 and 13 showed that the treatment ofcorrelation effects by means of the DFT+U approach is\nimportant to get the correct behavior of Kuas a function\nof temperature. In particular, in Ref. 13 the inclusion\nof the thermal expansion effects on Kuallowed to get a\nspin-reorientation temperature in agreement with exper-\niments and a theoretical MAE in good agreement with\nexperimental results, especially in the temperature range\n150-450 K.\nMore recently, in Ref. 14 it was suggested that the\nspin-reorientation phenomenon might be partially due to\nlattice dynamics. Such statement was supported by the\ncalculation of the lattice dynamics contribution to Ku,\nobtained by averaging the MAE over configurations in\nwhich the Mn and Bi atoms were displaced according to\nthe mean square atomic displacements as a function of\ntemperature.\nRecently, we extended Density Functional Perturba-\ntion Theory (DFPT) for lattice dynamics with Fully Rel-\nativistic (FR) Ultrasoft pseudopotentials (US-PPs) to\nmagnetic materials15. The new formulation allows to de-\ntect differences in the phonon frequencies for different\norientations of the magnetization, thus making possible\nto evaluate the vibrational free energy contribution to\nthe MAE.\nIn this paper we study, by means of ab-initio tech-\nniques, the lattice dynamics of ferromagnetic MnBi for\ntwo different orientations of the magnetization: 1. in–\nplane; 2.perpendiculartotheplane. Wefindthatthetwo\nphonon dispersions mainly differ in the high-frequency\noptical branches, where the phase with magnetic mo-\nments pointing in the out-of-plane direction shows, on\naverage, phonon modes of 2 cm\u00001lower in frequency.\nStarting from the difference of the vibrational density\nof states of the two phases we compute the vibrational\ncontribution to MAE. We find that, if the energy con-\ntribution to MAE is computed by the PBEsol exchange-\ncorrelation functional, the phonon contribution is of the\nsameorderofmagnitudeasthegroundstateMAE,hence\nit plays a relevant role in the calculation of Kuand toarXiv:2009.00410v1 [cond-mat.mtrl-sci] 1 Sep 20202\ndetermine the spin-reorientation transition temperature\nTSR.\nII. METHODS\nFirst-principle calculations were carried out by means\nof DFT16,17within the LDA18and the Perdew-Burke-\nErnzerhof optimized for solids (PBEsol)19schemes\nfor the exchange-correlation functional approximation,\nas implemented in the Quantum ESPRESSO20–22and\nthermo_pw23packages. The atoms are described by\nFR US-PPs24, with 3p,4s, and 3delectrons for Mn\n(PPs Mn.rel-pz-spn-rrkjus_psl.0.3.1.UPF and\nMn.rel-pbesol-spn-rrkjus_psl.0.3.1.UPF , from\npslibrary 0.3.125,26) and with 6s,5d, and 6pelectrons\nfor Bi (PPs Bi.rel-pz-dn-rrkjus_psl.1.0.0.UPF\nand Bi.rel-pbesol-dn-rrkjus_psl.1.0.0.UPF , from\npslibrary 1.0.025,26).\nMnBicrystallizesintheNiAsstructure, withanhexag-\nonal lattice described by the point group D6h. The inclu-\nsionofmagnetismdifferentiatesthestructuresintoalow-\nsymmetry phase ( m?chenceforth), below TSR, and a\nhigh-symmetry phase ( mkchenceforth), above TSR.\nIn particular, the mkcphase is described by the mag-\nnetic point group D6h(C6h), compatible with an hexago-\nnalBravaislattice, whilethe m?cphasehasamagnetic\npoint group D2h(C2h), compatible with a base-centered\northorhombic Bravais lattice. We checked the relevance\nof the lattice parameter b, which is not constrained by\nsymmetry in the m?cphase, and concluded that it is\nnot crucial to make the structure more stable than the\nmkcphase, hence in the rest of the paper we use the\nideal value b=p\n3a. In Table I we summarize the data\nrelative to the lattice constants and to the magnetic mo-\nment of Mn atoms, obtained with the LDA and PBEsol\nfunctionals, and compare them with previous theoretical\nresults and with experiments. The LDA geometry is in\ngood agreement with the theoretical results reported in\nRef. 14, but both lattice constants underestimate the\nexperimental values: in particular, ais 2 % smaller than\nexperiment, while cis 8 % smaller than experiment. The\nPBEsol geometry gives lattice constants slightly smaller\nthanPBE(reportedinRef. 14)andexperiments: aandc\nare 0.5 % and 6 % smaller than experiment, respectively.\nThem?candmkcphases have slightly different lat-\ntice constants aandc, but in Table I we report only one\nstructure because the differences in the lattice constants\nare beyond the significative digits reported. In this pa-\nper we use the LDA to compute the phonon frequencies\nand their contribution to the MAE, while the PBEsol\nis used to compute the energy contribution to the MAE\nand to correct it for thermal expansion effects. The LDA\nand the PBEsol (at T= 0K) calculations are performed\nat the geometry reported in Table I. The computed Mn\nmagnetic moment mMnis in agreement with previous\ncalculations reported in literature12,14:mMnis 10 %(20\n%) smaller than experiment within the PBEsol (LDA)Exchange-correlation a(Å)c(Å)mMn(\u0016B)\nfunctional\nLDA (this work) 4.165.57 3.2\nLDA (Ref. 14) 4.205.54 3.29\nGGA-PBEsol (this work) 4.245.67 3.5\nGGA-PBEsol (Ref. 14) 4.285.63 3.56\nGGA-PBE (Ref. 14) 4.355.76 3.69\nGGA-PBE (Ref. 12) 4.315.74 3.45\nGGA-PBE + U (Ref. 12) 4.396.12 3.96\nexp. (Ref. 10) 4.276.053.8-4.2\nTable I. Computed (FR LDA and PBEsol), theoretical refer-\nence (LDA, PBEsol, PBE, and PBE+U), and experimental\nlattice constants and Mn magnetic moments.\napproximation.\nThe pseudo wave functions (charge density) have been\nexpanded in a plane waves basis set with a kinetic energy\ncut-off of 110 (440) Ry. The Brillouin Zone (BZ) integra-\ntions have been done using a shifted uniform Monkhorst-\nPack mesh27of12\u000212\u00028k-points. The presence of\na Fermi surface has been dealt with by the Methfessel-\nPaxton smearing method28, with a smearing parame-\nter\u001b= 0:015Ry. The dynamical matrices have been\ncomputed on a uniform 4\u00024\u00023q-points mesh, and\na Fourier interpolation was used to obtain the complete\nphonon dispersions and the free energy. The latter has\nbeen obtained approximating the BZ integral with a\n300\u0002300\u0002300q-points mesh.\nIII. RESULTS\nMnBi is a magnetic binary compound, in which mag-\nnetism is carried mainly by the Mn atoms, while Bi is re-\nsponsible for a strong spin-orbit interaction. As a conse-\nquence, strong magnetocrystalline anisotropy effects are\nexpected.\nPhonon dispersions\nHere we consider the phonon dispersions of MnBi\nwith two different orientations of the magnetic moments,\nm?c(in-plane) and mkc(out-of-plane), and among\nall the possible in-plane orientations m?c, we choose\nmka,aandcbeing the primitive vectors of the hexag-\nonal Bravais lattice). The presence of a magnetization\nleads to a difference in the Bravais lattice the magnetic\npoint group is compatible with, as discussed in the pre-\nvious Section. In order to compare the phonon disper-\nsions in the same BZ, we choose to set the geometry\nin the base-centered orthorhombic Bravais lattice, which\nis compatible with the low-symmetry phase (magnetic\npoint group D2h(C2h)).\nThe phonon dispersions are illustrated in Fig. 1. The\nphonon modes are split in two groups, separated by a3\n 0 50 100 150 200\nΓXS RA ZΓYX1A1TYFrequency (cm−1)\nm || cm || a\n 0 50 100 150 200\nΓXS RA ZΓYX1A1TYFrequency (cm−1)\n 0 50 100 150 200\nΓXS RA ZΓYX1A1TYFrequency (cm−1)m|| cm⊥ c\nFigure 1. Computed FR LDA phonon dispersions of MnBi\nwith magnetic moments oriented in plane ( m?c, black line),\nand out of plane ( mkc, red line).\ngap. The low-frequency branches (up to \u0018100cm\u00001)\nare dominated by displacements of the heavy element Bi,\nwhile the high-frequency branches (from \u0018150cm\u00001to\n\u0018200cm\u00001) are mainly displacements of the Mn atoms.\nThe main difference between the phonon frequencies of\nthe two phases is a rigid shift: the phonon frequencies\nof the phase with in-plane magnetization are higher than\nthose of the phase with out-of-plane magnetization. The\nshift is about 0:5cm\u00001in the low-frequency branches,\nwhile it is about 2 cm\u00001in the high-frequency branches.\nMoreover, there are differences due to symmetry. The\nsystem with in-plane magnetization has lower symmetry\nand some modes, degenerate when the magnetization is\nalong c, split. As an example, in Table II we report the\nphonon modes at Z and X 1. At Z, in the phase with mk\ncthere are two groups of four-fold degenerate modes,\nwhich become four couples of degenerate modes in the\nphase m?c: the splittings are quite small, in the range\n0.02-0.08 cm\u00001. At X 1, in the configuration mkcthere\nare four two-fold degenerate modes, which split from 0.1\ncm\u00001to 1 cm\u00001. Similar splittings are found also along\nthe other high-symmetry lines.\nMAE\nIn previous works, the spin-reorientation transition,\ndue to the change of Kufrom negative to positive at\nTSR\u001990K, has been explained as an effect of thermal\nexpansion of the crystal parameters aandc.\nIn Refs. 12 and 13 Kth: exp:\nu, the function Kuobtained\naccounting for thermal expansion effects, has been com-\nputed within the LSDA +SO (spin-orbit) +U scheme\nusing the experimental lattice constants as a function\nof the temperature, reported in Refs. 10 and 11 and\nfinding in this way a good agreement with experiments.q mkc m ?c\nZ\u0017(cm\u00001) degeneracy \u0017(cm\u00001) degeneracy\n67.135 467.212 2\n67.231 2\n180.245 4181.263 2\n181.340 2\nX161.427 261.554 1\n62.503 1\n72.461 272.467 1\n72.577 1\n176.735 2178.156 1\n178.556 1\n184.374 2185.195 1\n185.395 1\nTable II. Computed FR LDA phonon frequencies at high-\nsymmetry points Z and X 1for the configurations mkcand\nm?c. Only the degenerate modes are shown for the config-\nuration mkc, and how the degeneracy is lowered or lifted if\nm?c.\nHere instead we compute Kth: exp:\nuwithin the FR PBEsol\nscheme. In Fig. 2(a) we show \u0016Ku, the contribution to Ku\ngiven by the energy difference of the electronic ground-\nstates, for a mesh of geometries and on top of it we indi-\ncate with black and white points the thermal expansion\ndata added to the theoretical PBEsol T= 0K crystal\nparameters. The resulting Kth: exp:\nuis reported in Fig.\n2(b).Kth: exp:\nuis negative and slightly increases with in-\ncreasingaandc. Yet, such an increase is not sufficient to\ncross the value Kth:exp:\nu = 0because the T= 0K energy\ndifference (\u0019\u00000:73meV/cell) is significantly lower than\nthe experimental value ( \u0019\u00000:15meV / cell), similarly to\nwhat found within the LDA and GGA approximations in\nRefs. 12 and 13, and because the energy difference land-\nscape shows a local maximum of about \u00000:4meV / cell.\nIn addition to the thermal expansion effect we con-\nsider also the effect of the lattice dynamics on Ku: in\nfact, since the two phases have slightly different phonon\nfrequencies, they have different vibrational entropies that\ngive a temperature-dependent contribution to the MAE\ndefined as the difference of the vibrational free energies\nof the two phases. We write the constant Kuas29:\nKu=Kth: exp:\nu +Kvib\nu+Kmag\nu; (1)\nwhereKth: exp:\nuis the thermal expansion contribution,\nKvib\nuis the lattice dynamics contribution, and Kmag\nuis\nthemagnoncontribution, whichwedonotconsiderinthe\npresent work. The term Kvib\nucan be computed from the\nphonon frequencies using the harmonic approximation:\nKvib\nu=Z+1\n0d!1\n2~!\u0002\ng?(!)\u0000gk(!)\u0003\n+kBTZ+1\n0d!\u0002\ng?(!)\u0000gk(!)\u0003\nln\u0010\n1\u0000e\u0000~!=kBT\u0011\n;\n(2)\nwhereg?(!)(gk(!)) is the phonon density of states rela-\ntive to the phase with m?c(mkc). In our case Kvib\nu4\n−1−0.9−0.8−0.7−0.6−0.5−0.4−0.3\n 0 50 100 150 200Kuth. exp. (mev / cell)\nTemperature (K)(a)\n(b) 4.22 4.24 4.26 4.28 4.3a (Å) 1.335 1.34 1.345 1.35 1.355c/a−1.6−1.4−1.2−1−0.8−0.6−0.4K−u (meV/cell)\nFigure 2. Energy contribution to the MAE constant, \u0016Ku,\ncomputed in the FR PBEsol scheme. (a) \u0016Kuas a function\nof lattice constants. Black and white dashed lines represent\nthe geometries at different temperatures, obtained from Refs.\n10 and 11, respectively, as explained in the main text. (b)\nKth:exp:\nuas a function of temperature. Red and blue lines\nrepresent the values of Kth:exp:\nuobtained from the black and\nwhite dashed lines of Fig. 2 (a), respectively.\n 0 0.5 1 1.5 2 2.5 3 3.5 4\n 0 100 200 300 400 500 600 700Kuvib (mev / cell)\nTemperature (K)\nFigure 3. Vibrational contribution to the MAE constant Ku,\ncomputed from the phonon frequencies within the harmonic\napproximation (Eq. (2))\nis always positive and increases with increasing temper-\nature, as shown in Fig. 3: its magnitude is comparable\n−0.5 0 0.5 1 1.5 2 2.5 3 3.5\n 0 100 200 300 400 500 600 700Ku (mev / cell)\nTemperature (K)exp. datatotal MAEtotal MAE(a)\n(b)−0.4−0.2 0 0.2 0.4 0.6 0.8\n 0 50 100 150 200Ku (mev / cell)\nTemperature (K)Figure 4. Comparison between computed and experimental\nMAE constant Ku. Black squares represent experimental\ndata8, red and blue lines represent the total Kucomputed,\nwhere theKth:exp:\nucontribution has been obtained from Refs.\n10 and 11, respectively, as explained in the main text. (a)\nDetailed comparison in the temperature range 0\u0000200K, to\nhighlightTSR. (b) Comparison in a wider range of tempera-\ntures ( 0\u0000600K).\nwiththatof Kth: exp:\nu, henceitgivesacrucialcontribution\nin determining the MAE constant Ku.\nIn Figs. 4(a-b) we show the total Kudefined in Eq.\n(1) and compare it with the experimental data. By com-\nparison with Fig. 2(b), it is clear that the addition of\nthe vibrational contribution is important. In particular,\nthe zero-point vibrational free energy difference allows to\nhave aT= 0K value of Kuin good agreement with\nexperiments and, together with the thermal contribution\n(second term of Eq. (2)), to get a transition tempera-\ntureTSRin reasonable agreement with the experiments\n(TSR\u001990K andTSR\u0019110K with the data given\nby Refs. 10 and 11, respectively). Moreover, the vibra-\ntional contribution allows to have a fair agreement with\nexperimental data8in a rather large temperature range,\nas shown in Fig. 4(b). At variance with Ref. 13, we do\nnot find a maximum in Kubecause the T= 0K geome-\ntry corresponds to the theoretical geometry and because\nthe vibrational contribution increases with T. At high\ntemperatures ( T >500K) our results do not agree with\nthe experiment, suggesting that additional contributions\nmay become important in this temperature range, in par-\nticular the term Kmag\nuin Eq. (1), which can result in a5\ndecreasing KuwithT30. Moreover, we mention that in\nthehigh-temperaturelimitadditionaleffectsnotincluded\nin Eq. (1), as the magnon-phonon coupling31,32, might\nbe non-negligible.\nIV. CONCLUSIONS\nBy means of recent developments in first-principles\ntheoretical tools, we have studied the lattice dynamics of\nthe MnBi ferromagnet for two orientations of the mag-\nnetization, m?candmkc. We have shown that\nthe differences in the phonon frequencies give rise to a\ncontribution to the MAE that is comparable with the\nelectronic one. We have found that the vibrational con-\ntribution is relevant to explain the behavior of the MAE\nconstantKuas a function of temperature. We could\nalso get an estimate of the spin-reorientation tempera-\ntureTSRin fair agreement with the experimental value\nusing the PBEsol approximation to evaluate the energy\ncontribution to the MAE. The use of other functionals\nsuchasLDAwouldgiveinsteadalowervalueoftheMAEatT= 0K than that predicted by PBEsol, and would\nlead to predict TSR\u0019500K. The use of the Hubbard\nU parameter would result, instead, in a higher value of\nKuatT= 0K13and, as a consequence, we would get\nKu>0when adding the vibrational contribution, thus\nthe spin-reorientation transition would not be explained.\nIn Ref. 14 a first estimate of the vibrational contribu-\ntion to MAE was given, which led to TSR\u0019450K, far\nfrom the experimental value. In this work we have found\navibrationalcontributionofthesameorderofmagnitude\nas found in Ref. 14, and we have included also the zero-\npoint vibrational free energy difference, which gives an\nimportant contribution and leads to an estimated spin-\nreorientation temperature TSR\u0019100K, when used to-\ngether with the energy MAE given by PBEsol.\nACKNOWLEDGMENTS\nComputational facilities have been provided by SISSA\nthrough its Linux Cluster and ITCS and by CINECA\nthrough the SISSA-CINECA 2019-2020 Agreement.\n1D. J. Sellmyer, B. Balamurugan, W. Y. Zhang, B. Das, R.\nSkomski, P. Kharel, and Y. Liu, Advances in Rare-Earth-\nFree Permanent Magnets. In: Marquis F. (eds) Proceed-\nings of the 8th Pacific Rim International Congress on Ad-\nvanced Materials and Processing. Springer, Cham; DOI:\nhttps://doi.org/10.1007/978-3-319-48764-9_212 (2013).\n2J. Cui, M. Kramer, L. Zhou, F. Liu, A. Gabay,\nG. Hadjipanayis, B. Balasubramanian, D. Sell-\nmyer, Acta Materialia 158, 118-137; DOI:\nhttps://doi.org/10.1016/j.actamat.2018.07.049 (2018).\n3C. Guillaud, J. Phys. Radium 12, 143; DOI:\n10.1051/jphysrad:01951001202014300 (1951).\n4B. W. Roberts, Phys. Rev. 104, 607; DOI: 10.1103/Phys-\nRev.104.607 (1956).\n5J. B. Yang, K. Kamaraju, W. B. Yelon, and W. J. James,\nAppl. Phys. Lett. 79, 1846; DOI: 10.1063/1.1405434\n(2001).\n6T. Chen, J. Appl. Phys. 45, 2358; DOI: 10.1063/1.1663594\n(1974).\n7M. A. McGuire, H. Cao, B. C. Chakoumakos, and\nB. C. Sales, Phys. Rev. B 90, 174425; DOI:\nhttps://doi.org/10.1103/PhysRevB.90.174425 (2014).\n8W. E. Stutius, T. Chen, and T. R. Sandin, Magnetism and\nmagnetic materials - 1973: Nineteenth Annual Conference,\nDOI: 10.1063/1.2947245 (1974).\n9T. Hihara and Y. K \u0016oi, J. Phys. Soc. Jpn. 29, 343; DOI:\nhttps://doi.org/10.1143/JPSJ.29.343 (1970).\n10J. B. Yang, Y. B. Yang, X. G. Chen, X. B. Ma, J. Z.\nHan, Y. C. Yang, S. Guo, A. R. Yan, Q. Z. Huang,\nM. M. Wu et al., Appl. Phys. Lett. 99, 082505; DOI:\nhttps://doi.org/10.1063/1.3630001 (2011).\n11K. Koyama, Y. Mitsui, and K. Watanabe,\nSci. Technol. Adv. Mater. 9, 024204; DOI:\nhttps://doi.org/10.1088/1468-6996/9/2/024204 (2008).12N. A. Zarkevich, L.L. Wang, and D. D. Johnson, APL\nMater. 2, 032103; DOI:https://doi.org/10.1063/1.4867223\n(2014).\n13V. P. Antropov, V. N. Antonov, L. V. Bekenov, A.\nKutepov, and G. Kotliar, Phys. Rev. B 90054404; DOI:\n10.1103/PhysRevB.90.054404 (2014).\n14K. V. Shanavas, D. Parker, and D. J. Singh, Sci. Reports\n4, 7222; DOI: 10.1038/srep07222 (2014).\n15A.UrruandA.DalCorso, Phys.Rev.B 100, 045115; DOI:\nhttps://doi.org/10.1103/PhysRevB.100.045115 (2019).\n16P. Hohenberg and W. Kohn, Phys. Rev. 136, B864; DOI:\nhttps://doi.org/10.1103/PhysRev.136.B864 (1964)\n17W. Kohn and L. J. Sham, Phys. Rev. 140, A1133; DOI:\nhttps://doi.org/10.1103/PhysRev.140.A1133 (1965)\n18J. P. Perdew and A. Zunger, Phys. Rev. B 23, 5048; DOI:\nhttps://doi.org/10.1103/PhysRevB.23.5048 (1981).\n19J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A.\nVydrov, G. E. Scuseria, L. A. Constantin, X. Zhou,\nand K. Burke, Phys. Rev. Lett. 100, 136406; DOI:\nhttps://doi.org/10.1103/PhysRevLett.100.136406 (2008).\n20P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car,\nC. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni,\nI. Dabo, et al., J. Phys. Condens. Matter 21, 395502;\nDOI: 10.1088/0953-8984/21/39/395502 (2009) (See http:\n//www.quantum-espresso.org ).\n21P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B.\nNardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli,\nM. Cococcioni, et al., J. Phys. Condens. Matter 29,\n465901; DOI: doi.org/10.1088/1361-648X/aa8f79 (2017).\n22P. Giannozzi, O. Baseggio, P. Bonfá, D. Brunato, R. Car,\nI. Carnimeo, C. Cavazzoni, S. de Gironcoli, P. Delugas, F.\nFerrari Ruffino, et al., J. Chem. Phys. 152, 154105; DOI:\nhttps://doi.org/10.1063/5.0005082 (2020).\n23thermo_pw is an extension of the Quantum ESPRESSO (QE)\npackage which provides an alternative organization of the6\nQE work-flow for the most common tasks. For more infor-\nmation see https://dalcorso.github.io/thermo_pw .\n24A. Dal Corso and A. Mosca Conte,\nPhys. Rev. B 71, 115106; DOI:\nhttps://doi.org/10.1103/PhysRevB.71.115106 (2005).\n25A. Dal Corso, Comp. Mat. Sci. 95, 337; DOI:\nhttps://doi.org/10.1016/j.commatsci.2014.07.043 (2014).\n26Seehttps://dalcorso.github.io/pslibrary .\n27H. J. Monkhorst and J.D. Pack, Phys. Rev. B 13, 5188;\nDOI: https://doi.org/10.1103/PhysRevB.13.5188 (1976).\n28M. Methfessel and A.T. Paxton, Phys. Rev. B 40, 3616;\nDOI: https://doi.org/10.1103/PhysRevB.40.3616 (1989).29F. Körmann, A. Dick, B. Grabowski, B. Hallstedt, T.\nHickel, and J. Neugebauer, Phys. Rev. B 78, 033102; DOI:\nhttps://doi.org/10.1103/PhysRevB.78.033102 (2008).\n30J. B. Staunton, S. Ostanin, S. S. A. Razee, B.\nL. Gyorffy, L. Szunyogh, B. Ginatempo, and\nE. Bruno, Phys. Rev. Lett. 93, 257204; DOI:\nhttps://doi.org/10.1103/PhysRevLett.93.257204 (2004).\n31F. Körmann, A. Dick, B. Grabowski, T. Hickel,\nand J. Neugebauer, Phys. Rev. B 85, 125104; DOI:\nhttps://doi.org/10.1103/PhysRevB.85.125104 (2012).\n32J. Hellsvik, D. Thonig, K. Modin, D. Iu-\nsan, A. Bergman, O. Eriksson, L. Bergqvist,\nand A. Delin, Phys. Rev B 99, 104302; DOI:\nhttps://doi.org/10.1103/PhysRevB.99.104302 (2019)." }, { "title": "2009.00920v1.Phase_transition_in_the_magnetocrystalline_anisotropy_of_tetragonal_Heusler_alloys__Rh__2T_Sb___T___Fe__Co.pdf", "content": "arXiv:2009.00920v1 [cond-mat.mtrl-sci] 2 Sep 2020G. H. F.; Rh 2CoSb 2020\nPhase transition in the magnetocrystalline anisotropy of t etragonal Heusler alloys:\nRh2TSb,T=Fe, Co\nGerhard H. Fecher,∗Yangkun He, and Claudia Felser\nMax-Planck-Institute for Chemical Physics of Solids, D-01 187 Dresden, Germany\n(Dated: September 3, 2020)\nThis work reports on first principles calculations of the ele ctronic and magnetic structure of\ntetragonal Heusler compounds with the composition Rh 2FexCo1−xSb (0≤x≤1). It is found that\nthe magnetic moments increase from 2 to 3.4 µBand the Curie temperature decreases from 500 to\n464 K with increasing Fe content x. The 3dtransition metals make the main contribution to the\nmagnetic moments, whereas Rh contributes only approximate ly 0.2µBper atom, independentof the\ncomposition. The paper focuses on the magnetocrystalline a nisotropy of the borderline compounds\nRh2FeSb, Rh 2Fe0.5Co0.5Sb, and Rh 2CoSb. A transition from easy-axis to easy-plane anisotropy is\nobserved when the composition changes from Rh 2CoSb to Rh 2FeSb. The transition occurs at an\niron concentration of approximately 40%.\nKeywords: Electronic structure, Magnetocrystalline anis otropy, Intermetallic compounds, Rh 2FeSb,\nRh2CoSb\nI. INTRODUCTION\nPermanent or hard magnets are made of bulk ma-\nterials with strong anisotropy, which may be based on\nmagnetocrystalline, shape anisotropy, or both. In mag-\nnets with magnetocrystalline anisotropy, there should be\nonly one easy crystal axis of magnetisation so that the\nanisotropy is uniaxial. Such an uniaxial magnetocrys-\ntalline anisotropy is found, for example, in tetragonal or\nhexagonal systems. Heusler alloys are compounds with\nformula T2T′M, where TandT′are transition met-\nals, and Mis a main group element. Some of these\ncompounds and alloys crystallise in tetragonal structure;\nhowever, most of them have a cubic crystal structure.\nOne advantage of Heusler compounds is that most of\nthem do not contain rare earth elements; rather, the\nmagneticpropertiesareprovidedby3 dtransitionmetals.\nMany tetragonal Heusler alloys are Mn-based, and sev-\neral exhibit structural martensite–austenitephase transi-\ntions. In particular, in the inversestructures with space\ngroupI4m2, the magnetic moments of the Mn atoms ex-\nhibit antiparallel coupling. Thus, these alloys are gener-\nallyferrimagnetswithlowsaturationmagnetisation. The\nRh2TMalloys (T′= V, Mn, Fe, Co; M= Sn, Sb) crys-\ntallise in a regular tetragonal structure with space group\nI4/mmmandareexpectedtoexhibituniaxialanisotropy\nwhen the 3 dtransition metals have large moments.\nExperiments on the crystal structure and magnetic\nproperties of Rh 2-based Heusler compounds were re-\nported by Dhar et al.[1], who observed a tetragonal\nstructure and a magnetic moment of 1.4 µBin the prim-\nitive cell. A Curie temperature of approximately 450 K\nwas measured. Further, Fallev et al.recently reported\nab initio calculations for many tetragonal Heusler com-\npounds (including Rh 2FeSb and Rh 2CoSb) [2]. This\nwork proposed that thin films of Rh 2CoSb exhibit uni-\n∗fecher@cpfs.mpg.deaxial, perpendicular anisotropy with the easy direction\nalong the c([001]) axis. Experiments and calculations\nboth suggest that Rh 2CoSb might be suitable hard mag-\nnetic material with uniaxial anisotropy. However, the\nconstituent elements, in particular Rh, might be too ex-\npensive for applications where bulk materials are needed,\nfor example, permanent magnets in electric engines.\nHowever, the cost of the materials is not as important\nfor thin film applications, for example, magnetic record-\ning media or magnetoelectronic memory devices.\nWe recently reported experiments on the magnetic\nproperties of Rh 2CoSb[in print, will be added later] . It\nwasfoundthat Rh 2CoSbhasuniaxialanisotropy,where c\nis theeasyaxis. The present work describes theoretically\nthe magnetic properties of Rh 2CoSb, its sister compound\nRh2FeSb, and alloys with mixed Co 1−xFexcomposition.\nII. DETAILS OF THE CALCULATIONS\nThe electronic and magnetic structures of Rh 2TSb\n(T= Fe, Co) were calculated using Wien2k [3–5] and\nSprkkr [6, 7] in the local spin density approxima-\ntion. In particular, the generalised gradient approxima-\ntion of Perdew, Burke, and Ernzerhof [8] was used to\nparametrise the exchange correlation functional. A k-\nmesh based on 126 ×126×126 points of the full Bril-\nlouin zone was used for integration when the total en-\nergies were calculated to determine the magnetocrys-\ntalline anisotropy (see also Appendix C). The calcula-\ntions are described in greater detail in References [9, 10].\nThe spin spirals and magnons were calculated accord-\ning to the schemes described in References [11] and [12],\nrespectively. Calculations for the disordered or off-\nstoichiometric compounds with mixed site occupations\nwere performed using Sprkkr and the coherent poten-\ntial approximation (CPA) [13] in the full potential mode.\nThe CPA allows the simulation of random site occupa-\ntion by different elements. Complications arising in the2\ncalculation of the magnetic anisotropy energies are dis-\ncussedin detail byKhan et al.[14], who comparedresults\nobtained using Wien2k andSPRKKR .\nThe basic crystal structure of the tetragonal Heusler\ncompounds [prototype, Rh 2VSn;tI8;I4/mmm(139)\ndba] is shown in Figure 1(a). The atoms are located in\nthe ferromagnetic structure on the 4d, 2b, and 2a Wyck-\noff positions of the centred tetragonal cell. The magnetic\norder changes the symmetry, and the resulting magnetic\nspace group for collinear ferromagnetic order with mo-\nments along the caxis isI4/mm′m′(139.537), where′\nis the spin reversal operator [15]. The symmetry is re-\nduced to that of space group Im′m′m(71.536) when the\nmagnetisation /vectorMis along the aaxis ([100]) or F m′m′m\n(69.524) for /vectorM/bardbl[110].\na) ordered ( regular ) b) disordered ( T\u0003 Sb) c) ordered ( inverse )\nFIG. 1. Crystal structure of Rh 2TSb (T= Fe, Co).\nIn the well-ordered regular structure (a), the sites of the l at-\ntice with space group I4/mmm(139) are occupied as follows:\n4d (0 1/2 1/4), Rh; 2b (0 0 1/2), T; and 2a (0 0 0), Sb. In the\ndisordered structure (b), the Tand Sb atoms are randomly\ndistributed on the 2b and 2a sites. The inversetetragonal\nstructure is shown in (c) for comparison.\nThe electronic structure and magnetic properties were\ncalculated using the optimised lattice parameters. As\nstarting point, the lattice parameters of two alternative\nstructures were optimised using Wien2k. In addition\nto the regular Heusler structure described above, the in-\nverse structure with space group I4m2 (119)dbcawas\nassumed. In this structure, the positions of the Co atom\nand one of the Rh atoms areinterchanged. Spin–orbitin-\nteractionwasconsideredowingtothehigh ZvaluesofRh\nand Sb. Note that the spin–orbit interaction is an intrin-\nsic property in the fully relativistic Sprkkr calculations,\nwhichsolvetheDiracequation. Theresultsoftheoptimi-\nsationaresummarisedinTableI.Theregularstructureis\nfound to have lower energy; it thus describes the ground\nstate. The energy difference compared to the inverse\nstructure is approximately 430 meV. The formation en-\nthalpy is calculated as∆ Hf=Etot−(2ERh+ECo+ESb),\nthat is, the difference between the total energy of the\ncompound in different structures and the sum of the en-\nergies of the elements in their ground state structure.\nThe formation enthalpy is clearly lower for the regular\nstructure than for the inverse tetragonal structure. Note\nthat the formation enthalpy is even lower (-220 meV) for\nthe cubic L21structure. The calculated lattice parame-ters are in good agreement with experimental values [1];\nhowever, the calculated cvalue and c/aratio are approx-\nimately 4% larger. This finding might be explained by\neither a temperature effect or some disorder in the ex-\nperiment.\nTABLE I. Structural properties of Rh 2CoSb.\nCalculations are performed for the regular (139) and invers e\n(119) Heusler structures. The lattice parameters ( a,c,c/a),\nformation enthalpy (∆ Hf), and spin magnetic moment msof\nthe primitive cell (total experimental magnetic moment) ar e\nlisted. Experimental values from Reference [1] are shown fo r\ncomparison. Note that the magnetic moment in this reference\nis not saturated.\nCalculated Exp.\n139 119 here [1]\na[˚A] 4.0104 3.95 4.0393 4.04\nc[˚A] 7.3628 7.56 7.1052 7.08\nc/a 1.836 1.91 1.759 1.75\n∆Hf[meV] -754 -325\nms[µB] 2.04 1.79 2.36 1.4\nTC[K] 450 450\nIII. RESULTS AND DISCUSSION\nA. Electronic and magnetic structure of Rh 2CoSb\nThe calculated electronic structure of Rh 2CoSb in the\nregular tetragonal Heusler structure is illustrated in Fig-\nure 2 in terms of the band structure and density of states\n(n(E)). The relativistic bands, spin-resolved total den-\nsity of states, and its atomic contributions are shown.\nThe electronic structure is calculated in the full relativis-\ntic mode by solving the Dirac equation. The band struc-\nture from semi-relativistic calculations is shown in the\nAppendix.\nBoth rhodium and cobalt contribute to the magnetic\nmoment of the compound. The spin and orbital mag-\nnetic moments are mCo\ns= 1.656µBandmCo\nl= 0.139µB\nfor cobalt and mRh\ns= 0.206µBandmRh\nl= 0.007µB\nfor rhodium, respectively. The overall magnetic moment\n(spinplusorbital)oftheprimitivecellis mtot= 2.188µB.\nThe orbitalmoment of the Co atoms makes a remarkably\nlarge contribution.\nThe real space chargeand spin distributions are shown\nin Figure 3. The charge density ( σ(r)) of the atoms\nhas no striking shape. It appears to be nearly spheri-\ncal but still reflects the two- or fourfold symmetry. As\nexpected, most of the electrons are close to the ion cores.\nBy contrast, the spin or magnetisation density ( σ(r)) has\na much more pronounced shape depending on the plane.\nIn particular, in the (110) plane, it has a distinct butter-\nfly shape. The spin densityis positive at both the Coand\nRh atoms. It is clearly higher near the Co atoms than3\n/s77 /s78 /s83\n/s48/s88 /s77\n/s48/s45/s54/s45/s52/s45/s50/s48/s50\n/s45/s54 /s45/s52 /s45/s50 /s48 /s50 /s52 /s54/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s107 /s41/s32/s45/s32\n/s70/s32/s91/s101/s86/s93\n/s69/s108/s101/s99/s116/s114/s111/s110/s32/s109/s111/s109/s101/s110/s116/s117/s109/s32/s32/s32 /s107/s109 /s105/s110/s111/s114/s105/s116/s121 \n/s68/s101/s110/s115/s105/s116/s121 /s32/s111/s102/s32/s115/s116/s97/s116/s101/s115/s32/s32/s32 /s110 /s40/s69 /s41/s32/s91/s101/s86/s45/s49\n/s93/s109 /s97/s106/s111/s114/s105/s116/s121 \n/s32/s116/s111/s116/s97/s108\n/s32/s82/s104\n/s32/s67/s111\nFIG. 2. Electronic structure of Rh 2CoSb (I).\nShownis thefullyrelativistic bandstructuretogetherwit hthe\ntotal and site (Rh and Co) specific, spin-resolved densities of\nstates.\nnear the Rh atoms, which ultimately gives Co a higher\nmagnetic moment. The magnetisation density of the Rh\natoms is aligned along the magnetisation direction and\npoints somewhat toward the nearest Co atoms.\n1. Magnetic anisotropy\nFurther, the directional dependence of the magnetisa-\ntion was investigated to explain the collinear magnetic\norder in detail. In particular, the total energy was cal-\nculated for cases where the magnetisation points along\ndifferent crystallographic directions. The obtained en-\nergy differences make it possible to determine the mag-\nnetocrystalline anisotropy (see also Appendix C).\nInthemagneticanisotropyofRh 2CoSb,the easyaxisis\nalong the c([001]) axis. The simple second-order uniax-\nial anisotropy constant is Ku= 1.37 MJ/m3(see Equa-\ntions (C1) and (C2) in Appendix C1). This results in\nan anisotropy field of µ0Hu≈2.4 T. A more detailed\nanalysis reveals that the simple second-order anisotropy\nconstant Kuisnotsufficienttodescribethemagnetocrys-\ntalline anisotropy, as discussed in Section IIID.\nFurther, thedipolarmagnetocrystallineanisotropywas\ncalculated as described in Appendix C3 and was found\nto be ∆Edipaniso= 0.09µeV. The positive value indicates\nan easy dipolar direction along the [001] axis. The dipo-\nlar anisotropyis rather small compared to the anisotropy\ncalculated from the total energy. Here, it was calcu-\nlated for a sphere with a radius of 30 nm. The results\nfor other shapes will be different, resulting in a distinct\nshape anisotropy. In particular, in thin films, the di-\nmension perpendicular to the film is much smaller than\nthe dimensions in the film plane. Therefore, the sum-\nmation in Equation (C16) becomes a truncated sphere\nthat is strongly anisotropic, and a pronounced thin film\nanisotropyappears. This thin film anisotropywill alsobe\naffected by the magnetic moments, which are different at/s48/s99\n/s82/s104/s67/s111/s67/s104/s97/s114/s103/s101/s32/s100/s101/s110/s115/s105/s116/s121 /s32/s32/s32 /s114 /s40 /s114 /s41/s91/s48/s48/s49/s93\n/s83/s98\n/s48 /s97/s48/s99/s77/s97/s103/s110/s101/s116/s105/s115/s97/s116/s105/s111/s110/s32/s100/s101/s110/s115/s105/s116/s121 /s32/s32/s32 /s115 /s40 /s114 /s41/s91/s48/s48/s49/s93\n/s91/s49/s48/s48/s93/s48/s46/s48\n/s49/s46/s48\n/s49/s48/s48/s46/s48\n/s83/s98/s67/s111\n/s48 /s97 /s32/s214/s50\n/s91/s49/s49/s48/s93/s48/s46/s48\n/s48/s46/s49\n/s49/s46/s48\nFIG. 3. Electronic structure of Rh 2CoSb (II).\nFully relativistic charge ( ρ(r)) and spin ( σ(r)) distributions\nfor the (001) and (110) planes are shown. The calculation\nis form/bardblc, that is, the magnetisation points along [001] ac-\ncording to the easy axis behaviour of the magnetic anisotropy.\n(Note: colour bars are in atomic units.)\ninterfaces and surfaces from that at the centre layers of\nthe film.\n2. Spiral spin order\nThe energy of the spin spirals was calculated to search\nfor non-collinear magnetic order. The spin spirals were\ncalculated for different directions and different cones. In\nplanar spirals, the spins are perpendicular to the propa-\ngation direction. Figure 4 compares the energies of pla-\nnar spirals along the high-symmetry directions.\nThe spirals along [100] or [110] propagate in the four-\nfold plane, whereas the spiral in the [001] direction prop-\nagates along the caxis. In all cases, the lowest energy\nis observed at q= 0. The magnetic moment of the Co\natoms varies by approximately 17% at maximum. The\nmagnetic moment of the Rh atoms decreases with in-\ncreasing qand vanishes at the border, independent of\nthe propagation direction of the spiral.\nThe spin direction was assumed to be perpendicular4\n/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48\n/s48 /s189 /s49/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53\n/s48 /s189 /s49 /s48 /s189 /s49/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s113/s41/s32/s91/s109/s101/s86/s93/s113 /s32/s124 /s124 /s32/s91/s48/s48/s49/s93\n/s32/s32\n/s113 /s32/s124 /s124 /s32/s91/s49/s48/s48/s93\n/s32/s32\n/s113 /s32/s124 /s124 /s32/s91/s49/s49/s48/s93/s32/s32/s32/s32/s32/s32 /s109 /s40/s113 /s41/s32/s91\n/s66/s93/s67/s111\n/s82/s104\n/s83/s112/s105/s114/s97/s108/s32/s118/s101/s99/s116/s111/s114/s32/s32/s32 /s124/s113/s124 /s32/s32/s32/s91 /s47/s100\n/s104/s107/s108 /s93\n/s32/s32 /s32\nFIG. 4. Planar spin spirals in Rh 2CoSb.\nThe spiral energies are given with respect to q= 0, that\nis, ∆E(q) =E(q)−E(0). The behaviour of the magnetic\nmoment m(q) is shown for Rh and Co.\nto theqvector in the above calculations for planar spi-\nrals. Thus, the angle between /vector qand the local magnetic\nmoment /vector miwas set to Θ = π/2. Next, the spirals were\nassumed to be conical with 0 <Θ< π/2 to allow for a\nmore detailed analysis. The calculations were performed\nforqalong [001]. Figure 5 displays the results for coni-\ncal spirals with various cone angles. The highest energies\nappear for the planar spiral. The energy at the border\nof the Brillouin zone ( q=π/c) exhibits a sine depen-\ndence. Thus, it vanishes in the antiferromagnetic state.\nThe behaviour of the local magnetic moments suggests\nmore localised behaviour at the Co atoms and induced\nbehaviour at the Rh atoms.\n/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48\n/s48 /s189 /s49/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48\n/s48 /s49 /s189/s48/s176 /s51/s48/s176 /s54/s48/s176 /s57/s48/s176/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s113 /s32/s124 /s124 /s32/s91/s48/s48/s49/s93\n/s32/s61/s32/s57/s48/s176\n/s32/s32/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s113 /s41/s32/s91/s109/s101/s86/s93/s32/s61/s32/s55/s53/s176\n/s32/s32\n/s32/s61/s32/s54/s48/s176\n/s32/s32\n/s83/s112/s105/s114/s97/s108/s32/s118/s101/s99/s116/s111/s114/s32/s32/s32 /s124/s113/s124 /s32/s32/s32/s91 /s47/s99 /s93/s32/s61/s32/s51/s48/s176\n/s32/s32\n/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40 /s41/s32/s91/s109/s101/s86/s93\n/s67/s111/s110/s101/s32/s97/s110/s103/s108/s101/s32/s32/s32/s32/s32/s32 /s113\n/s48/s48/s49/s32/s61/s32 /s47/s99\nFIG. 5. Conical spirals in Rh 2CoSb.\nSpiral energies for different cone angles and the wave vector\nalong the caxis (q/bardbl[001]) are shown. The angular dependence\natq=π/cis also shown.\nThe calculatedspiralenergiesindicate that this type of\nmagnetic order is rather improbable. The spiral energies\nincrease monotonously with the wave vector and cone\nangle, rather independent on the /vector qdirection. The mono-tonic behaviour suggests that a canted magnetic order is\nalso very unlikely [16].\n3. Exchange coupling and magnons\nThe exchange coupling energies were calculated using\nthe scheme of Liechtenstein et al.[17, 18] to estimate\nthe Curie temperature, spin stiffness, and presence of\nmagnons [12]. The exchange coupling parameters are\nplotted in Figure 6(a). The most dominant parameters\nfor Co–Co and Co–Rh interactions are shown; all the\nothers are comparatively small. The largest interaction\nappears for Co atoms in the centre and nearest to the Co\nin the neighbouring plane. From the calculated exchange\ncoupling energies, the Curie temperature was found to\nbeTC= 498 K, which is close to the experimental value\n(450 K) [1]. The calculated spin wave stiffness constant\nisDij= 866 meV ·˚A2, and the interpolation scheme of\nPadjaet al.[19] yields an extrapolated spin wave stiff-\nness ofD0= 864 meV ·˚A2.\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53 /s50/s46/s48 /s50/s46/s53 /s51/s46/s48/s48/s50/s52/s54\n/s77 /s71 /s88/s77\n/s48/s71/s48/s53/s48/s49/s48/s48/s49/s53/s48\n/s48/s46/s48/s48 /s48/s46/s48/s53 /s48/s46/s49/s48/s77/s71/s88 /s77/s71/s69/s120/s99/s104/s97/s110/s103/s101/s32/s99/s111/s117/s112/s108/s105/s110/s103/s32/s112/s97/s114/s97/s109/s101/s116/s101/s114/s32/s32/s32 /s74\n/s105/s106/s40/s100 /s41/s32/s91/s109/s101/s86/s93\n/s82/s101/s108/s97/s116/s105/s118/s101/s32/s100/s105/s115/s116/s97/s110/s99/s101/s32/s32/s32 /s100 /s32/s47/s32 /s97/s67/s111/s32/s97/s116/s32/s99/s101/s110/s116/s101/s114\n/s32/s67/s111/s32/s45/s32/s67/s111\n/s32/s67/s111/s32/s45/s32/s82/s104/s40/s97/s41\n/s32/s32/s77/s97/s103/s110/s111/s110/s32/s101/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s113 /s41/s32/s91/s109/s101/s86/s93\n/s77/s97/s103/s110/s111/s110/s32/s109 /s111/s109 /s101/s110/s116/s117/s109 /s32/s32/s32 /s113/s32/s32/s98/s111/s116/s104\n/s32/s32/s67/s111/s32/s111/s110/s108/s121 /s40/s98/s41/s32\n/s32\n/s110 /s40/s69 /s41/s32/s91/s109 /s101/s86/s45/s49\n/s93/s40/s99/s41\nFIG. 6. Exchange coupling parameters of Rh 2CoSb.\n(a) Exchange coupling parameters for Co–Co and Co–Rh in-\nteraction as functions of distance. Lines are drawn for bett er\ncomparison. (b) Magnon dispersion and (c) density of states .\nCalculations were performed with and without accounting fo r\nthemagnetic momentoftheRhatoms (thatis, for bothatoms\nand for Co only).5\nThe magnon dispersion was calculated by Fourier\ntransformation of the real space exchange coupling pa-\nrameters. The result is presented in Figure 6(b), and\nthe magnon density of states is shown in Figure 6(c).\nTwo calculations were made; in one calculation, only the\nCo–Co interaction was considered, and in the other, the\nmoments of the Rh atoms, which result in additional Co–\nRh and Rh–Rh coupling, were included. The latter cal-\nculation yields flat dispersion curves and a high density\nof states. A comparison of the two calculations reveals\nthat the magnons are dominated by the Co–Co interac-\ntion. Note that the Curie temperature in only 10 K lower\nwhen the Rh moments and the corresponding exchange\nparameters are ignored.\nB. Results for Rh 2FeSb\nThe calculations for Rh 2FeSb were performed in the\nsame way as for Rh 2CoSb. The regular structure with\nspace groupno. 139wasfound to be more stable than the\ninverse structure with space group no. 119. In addition,\nas in the case of Rh 2CoSb, the calculated clattice pa-\nrameter, and thus c/a, are considerably larger than the\nexperimental values (see Table II).\nTABLE II. Structural properties of Rh 2FeSb.\nCalculations are performed for the regular tetragonal Heus ler\nstructures. Lattice parameters ( a,c,c/a) and spin magnetic\nmoment msof the primitive cell are listed. Experimental\nvalues from Reference [1] are shown for comparison. Note\nthat the experimental moments in [1] are not saturated.\nExperiment\nCalculated This work Ref. [1]\na[˚A] 4.0418 4.0671 4.07\nc[˚A] 7.3995 7.0161 6.96\nc/a 1.8308 1.7251 1.71\nms[µB] 3.4 3.8 2.8\nTC[K] 510 510\nThe electronic structure of Rh 2FeSb is illustrated in\nFigure 7. The fully relativistic band structure and the\nspin- and site-resolved densities of states are shown. The\ncalculated spin and orbital magnetic moments are mFe\ns=\n2.978µBandmFe\nl= 0.080µBforironand mRh\ns= 0.228µB\nandmRh\nl= 0.006µBfor rhodium, respectively. The over-\nall magnetic moment (spin plus orbital) of the primitive\ncell ismtot= 3.488µB. The magnetic moment of the\nFe atoms is strongly localised, which is typical of Heusler\ncompounds with high magnetic moments. It clearly ex-\nceeds the value for elemental iron.\nThe real space charge and spin distributions of\nRh2FeSb are shown in Figure 8. As in the Co-containing\ncompound, σ(r) does not havea pronouncedshape (com-\npare Figure 3). The magnetisation density ( σ(r)) around\nthe Fe atoms has a less distinct shape compared to Co/s77 /s78 /s83\n/s48/s88 /s77\n/s48/s45/s54/s45/s52/s45/s50/s48/s50\n/s45/s54 /s45/s52 /s45/s50 /s48 /s50 /s52 /s54/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s107 /s41/s32/s45/s32\n/s70/s32/s91/s101/s86/s93\n/s69/s108/s101/s99/s116/s114/s111/s110/s32/s109/s111/s109/s101/s110/s116/s117/s109/s32/s32/s32 /s107/s109 /s105/s110/s111/s114/s105/s116/s121 \n/s68/s101/s110/s115/s105/s116/s121 /s32/s111/s102/s32/s115/s116/s97/s116/s101/s115/s32/s32/s32 /s110 /s40/s69 /s41/s32/s91/s101/s86/s45/s49\n/s93/s109 /s97/s106/s111/s114/s105/s116/s121 \n/s32/s116/s111/s116/s97/s108\n/s32/s82/s104\n/s32/s70/s101\nFIG. 7. Electronic structure of Rh 2FeSb (I).\nFully relativistic band structure is shown, along with the\ntotal- and site-specific spin-resolved densities of states for Rh\nand Fe.\nin Rh2CoSb; it is also not greatly affected by changes in\nthe magnetisation direction. The main difference is the\nmagnetisation density around the Rh atoms, which is ro-\ntated and appears to be aligned along the magnetisation\ndirection.\nTable III compares the calculated magnetic data of\nRh2CoSb and Rh 2FeSb. Rh 2FeSb clearly has a smaller\norbitalmagneticmomentthanRh 2CoSb,whereasitsspin\nmagnetic moment is higher because of the effect of the Fe\natoms. The induced magnetic moments of the Rh atoms\nare similar in both compounds.\nTABLE III. Calculated magnetic properties of Rh 2FeSb,\nRh2Fe0.5Co0.5Sb, and Rh 2CoSb.\nSpinmsand orbital mlmagnetic moments per atom (Rh, T\n= Co, Fe with m/bardblcin all cases) of the primitive cell ( total) are\nlisted, as well as Curie temperature TC, spin stiffness D0, and\nanisotropy parameters. (Note that the dipolar anisotropy i s\nthree orders of magnitude lower than the magnetocrystallin e\npart.)\nFe Fe 0.5Co0.5 Co\nmRh\ns[µB] 0.237 0.239 0.204\nmRh\nl[µB] 0.006 0.008 0.006\nmFe\ns[µB] 3.006 2.977 -\nmFe\nl[µB] 0.080 0.084 -\nmCo\ns[µB] - 1.747 1.674\nmCo\nl[µB] - 0.132 0.137\nmtotal\ns[µB] 3.44 2.81 2.04\nmtotal\nl[µB] 0.09 0.12 0.15\nTC[K] 465 480 500\nD0[meV˚A2] 590 700 870\nKu[MJ/m3] -1.21 -0.23 1.37\n|µ0Ha|[T] 1.34 0.31 2.43\n∆Edipaniso[kJ/m3] 1.9 2.06\n/s48/s99\n/s82/s104/s70/s101/s67/s104/s97/s114/s103/s101/s32/s100/s101/s110/s115/s105/s116/s121 /s32/s32/s32 /s114 /s40 /s114 /s41/s91/s48/s48/s49/s93\n/s83/s98\n/s48 /s97/s48/s99/s77/s97/s103/s110/s101/s116/s105/s115/s97/s116/s105/s111/s110/s32/s100/s101/s110/s115/s105/s116/s121 /s32/s32/s32 /s115 /s40 /s114 /s41/s91/s48/s48/s49/s93\n/s91/s49/s48/s48/s93/s48/s46/s48\n/s49/s46/s48\n/s49/s48/s48/s46/s48\n/s83/s98/s70/s101\n/s48 /s97 /s32/s214/s50\n/s91/s49/s49/s48/s93/s48/s46/s48\n/s48/s46/s49\n/s49/s46/s48\nFIG. 8. Electronic structure of Rh 2FeSb (II).\nFully relativistic charge ( ρ(r)) and spin ( σ(r)) distributions\nfor different planes. Magnetisation is perpendicular to cwith\nmalong [100] in accordance with the easy plane behaviour\nof the magnetic anisotropy. (Note: colour bars are in atomic\nunits.)\nThe calculated Curie temperatures are of the same or-\nder of magnitude as the experimental values. In contrast\ntothe calculatedresults, however,the experimentalvalue\nof the Fe compound is higher than that of the Co com-\npound. A possible reason is differences in the variation\nof the lattice parameters with temperature, which affect\nthe exchange coupling parameters and thus TCand also\nthe spin stiffness. Note that a much lower Curie tem-\nperature is obtained for the Co compound when it is off-\nstoichiometric (see Appendix A2), whereas the TCvalue\nof the off-stoichiometric Fe compound is slightly higher.\nThe anisotropy has the hardaxis along the z([001])\ndirection, and the easy plane is the the basal plane.\nBy contrast, for Rh 2CoSb, the zdirection is the easy\naxis. The simple uniaxial anisotropy constant is Ku=\n−1.21 MJ/m3. Consequently, the anisotropy field is\n|µ0Ha|= 1.34 T. The appearance of the ”hard”axis\nalongzisoppositetoRh 2CoSbwhere zisthe”easy”axis.\nThe dipolar magnetocrystalline anisotropy of Rh 2FeSb\nis ∆Edipaniso= 0.09µeV, indicating that the easy dipo-\nlar direction is along the [001] axis, like that of the Co-containing compound. This behaviour is caused by the\nstrong magnetic moments of the 3 dtransition metals, in\naddition to the elongation of the tetragonal crystal struc-\nture along the caxis.\nThe dynamic magnetic properties of Rh 2FeSb are\nshown in Figure 9. The spin spirals and magnons are\nsimilar to those of Rh 2CoSb; however, their energies ex-\ntend to higher values. The behaviour of the spin spirals\nrules out the presence of non-collinear magnetic struc-\nture [16].\n/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s50/s53/s48\n/s48 /s189 /s49/s48/s49/s50/s51\n/s48 /s189 /s49 /s48 /s189 /s49\n/s77 /s88 /s77\n/s48/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s50/s53/s48/s51/s48/s48/s51/s53/s48\n/s48/s46/s48/s48 /s48/s46/s48/s53/s77 /s88 /s77/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s113/s41/s32/s91/s109/s101/s86/s93/s113 /s32/s124 /s124 /s32/s91/s48/s48/s49/s93\n/s32/s32\n/s113 /s32/s124 /s124 /s32/s91/s49/s48/s48/s93\n/s32/s32\n/s113 /s32/s124 /s124 /s32/s91/s49/s49/s48/s93/s32/s32/s32/s32/s32/s32 /s109 /s40/s113 /s41/s32/s91\n/s66/s93/s70/s101\n/s82/s104\n/s83/s112/s105/s114/s97/s108/s32/s118/s101/s99/s116/s111/s114/s32/s32/s32 /s124/s113/s124 /s32/s32/s32/s91 /s47/s100\n/s104/s107/s108/s93\n/s32/s32 /s32/s32/s32/s77/s97/s103/s110/s111/s110/s32/s101/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s113 /s41/s32/s91/s109/s101/s86/s93\n/s77/s97/s103/s110/s111/s110/s32/s109/s111/s109/s101/s110/s116/s117/s109/s32/s32/s32 /s113/s32/s32/s98/s111/s116/s104\n/s32/s32/s70/s101/s32/s111/s110/s108/s121 /s32\n/s32\n/s103 /s40/s69 /s41/s32/s91/s109/s101/s86/s45 /s49\n/s93\nFIG. 9. Dynamic magnetic properties of Rh 2FeSb.\nSpiral energies with correspondingmagnetic momentsandth e\nmagnon dispersion are shown, along with the magnon density\nof states g(E). Magnon calculations were performed with and\nwithout the magnetic moment of the Rh atoms; in the latter\ncase, all Fe–Rh interactions are neglected.\nC. Results for Rh 2FexCo1−xSb\nOwing to the differences in magnetic anisotropy be-\ntween the Fe- and Co-based compounds, it is interest-\ning to investigate a mixed system containing both Fe\nand Co. Therefore, calculations were also performed\nfor Rh 2FexCo1−xSb using Sprkkr and the CPA. The\nCPA enables the simulation of random occupation of Fe\nand Co atoms at a single site (here 2 b). The obtained\nmagnetic properties of Rh 2Fe0.5Co0.5Sb are shown in Ta-\nble III. The uniaxial anisotropy constant is negative, like\nthat of Rh 2FeSb; however, its absolute value is consider-\nably lower (by a factor of 35) than that of Rh 2CoSb.7\nThedependenceofthemagneticpropertiesonthecom-\nposition is shown in Figure 10. The total magnetic mo-\nment increases with increasing Fe content, mainly be-\ncause Fe has a higher spin magnetic moment ( ≈3µB)\nthan Co( ≈1.7µB). The individualmagneticmomentsof\nthe atoms are nearly unaffected by the composition. The\ncalculated Curie temperature decreases with increasing\nFe content.\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53/s51/s46/s48/s51/s46/s53/s52/s46/s48\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48/s48/s49/s48/s48/s52/s54/s48/s52/s55/s48/s52/s56/s48/s52/s57/s48/s53/s48/s48/s77/s97/s103/s110/s101/s116/s105/s99/s32/s109/s111/s109/s101/s110/s116/s32/s32/s32 /s109 /s40/s120 /s41/s32/s91\n/s66/s93\n/s70/s101/s32/s99/s111/s110/s116/s101/s110/s116/s32/s32/s32 /s120/s32/s109\n/s116/s111 /s116\n/s32/s109\n/s115\n/s32/s109\n/s108 \n/s84\n/s67/s40/s120 /s41/s32/s91/s75/s93\nFIG. 10. Magnetic properties of Rh 2FexCo1−xSb.\nTotal (mtot), spin (ms), and orbital ( ml) magnetic moments\nas functions of Fe content xare shown. The inset shows the\nCurie temperature ( TC).\nD. Magnetocrystalline anisotropy of\nRh2FexCo1−xSb\nThus far, only the simplest case of uniaxial magne-\ntocrystalline anisotropy has been considered. The equa-\ntions for extending the calculations to more detailed\ncases are given in Appendix C. These equations were\nused to calculate the fourth-order uniaxial and tetrag-\nonal anisotropy constants, which were used to obtain the\nmagnetocrystalline anisotropy energy distributions.\nThe calculated uniaxial energy distributions Eu′(θ,φ)\n(see Equations (C4) and (C18) in the Appendix) of\nRh2FeSb, Rh 2Fe0.5Co0.5Sb, and Rh 2CoSb are plotted in\nFigure 11 for comparison.\nThe different behaviour of the anisotropy is clearly re-\nvealed in Figure 11. Rh 2FeSb has an easyplane, and cis\nthehardaxis; Rh 2Fe0.5Co0.5Sb has an easyplane as well,\nbut ahardcone, and in Rh 2CoSb, the cdirection is the\neasyaxis. Rh 2Fe0.5Co0.5Sb has a much lower anisotropy\nthan the pure compounds, and the differences between\nthe energies of the abplane and the caxis are very small.\nA hard cone appears with its maximum at an angle of\nθ3,4=±35.7◦(see Equation (C9) in Appendix C1).\nThe calculated anisotropy constants for uniaxial and\ntetragonalsymmetry arecomparedin Table IV. The sim-\npleKufromEquation(C2)(seeAppendix C) clearlycan-D\f\u00035K \u0015)H6E\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003 5K \u0015)H \u0013\u0011\u0018 &R \u0013\u0011\u0018 6E\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003F\f\u00035K \u0015&R6E \nFIG. 11. Uniaxial magnetic anisotropy of Rh 2TSb com-\npounds.\nEnergy distributions Eu′(θ,φ) ofT= Fe (a), Fe 0.5Co0.5(b),\nand Co (c). The energies Ex,y,zare given in µeV. (Please\nnote the different energy scales.)\nnot describe the magnetic anisotropy correctly.\nTABLE IV. Comparison of the anisotropy constants of\nRh2TSb,T= Fe, Fe 0.5Co0.5, and Co.\nRh2FeSb Rh 2Fe0.5Co0.5Sb Rh 2CoSb\nuniaxial\nKu[MJ/m3] -1.21 -0.23 1.37\nK0[MJ/m3] 1.31 0.39 0.0\nK2[MJ/m3] -2.19 0.50 3.62\nK4[MJ/m3] 0.98 -0.73 -2.25\ntetragonal\nK0,0[MJ/m3] 1.31 0.39 0.0\nK2,0[MJ/m3] -2.19 0.50 3.62\nK4,0[MJ/m3] 0.93 -0.81 -2.40\nK4,4[MJ/m3] 0.05 0.08 0.15\nThe dependence of the uniaxial anisotropy constants\non the composition is illustrated in Figure 12. The uni-\naxial anisotropy constant Kudecreases with increasing\niron content and exhibits a zero-crossing at x0≈0.4. At\nintermediate iron contents, more complex behaviour ap-\npears, as shown by the composition dependence of K2i\nand the results in Figures 11 and 13.\nThe calculated tetragonal energy distributions\nEa′(θ,φ) (see Equations (C12) and (C18) in the Ap-\npendix) of Rh 2FeSb, Rh 2Fe0.5Co0.5Sb, and Rh 2CoSb\nare shown in Figure 13. As in the plot of the uniaxial\nanisotropy in Figure 11, the differences in the anisotropy\nare easily observed. In Rh 2FeSb, the hardaxis is along\nthez([001]) direction, and the anisotropy exhibits\nweak variation in the basal plane, which is close to\ntheeasyplane. Closer examination of the basal plane\nshows biaxial behaviour with easyaxes along the [110]\nand [110] axes, but the energy difference between these\ndirections and the [100] or [010] axes is very small. The\nanisotropy of Rh 2CoSb is still almost uniaxial, with the\neasyaxis along the c([001]) axis, and varies weakly\nin the basal plane. Rh 2Fe0.5Co0.5Sb has much lower\nanisotropy than the pure compounds and exhibits more\ncomplicated directional behaviour.\nThe directional dependence of the orbital magnetic\nmoments was analysed to clarify the role of the spin–8\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s45/s50/s45/s49/s48/s49/s50\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48/s45/s49/s48/s49/s65/s110/s105/s115/s111/s116/s114/s111/s112/s121/s32/s99/s111/s110/s115/s116/s97/s110/s116/s115/s32/s32/s32 /s75\n/s50/s105/s40/s120 /s41/s32/s91/s77/s74/s47/s109/s51\n/s93\n/s70/s101/s32/s99/s111/s110/s116/s101/s110/s116/s32/s32/s32 /s120/s32/s32/s32 /s75\n/s48/s40/s120 /s41\n/s32/s32/s32 /s75\n/s50/s40/s120 /s41\n/s32/s32/s32 /s75\n/s52/s40/s120 /s41/s75\n/s117/s40/s120 /s41/s32/s91/s77/s74/s47/s109/s51\n/s93\nFIG. 12. Anisotropy constants of Rh 2FexCo1−xSb com-\npounds.\nThe inset shows the uniaxial anisotropy constant obtained\nusing Equation (C3) in Appendix C1.\nD\f\u00035K \u0015)H6E\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003 5K \u0015)H \u0013\u0011\u0018 &R \u0013\u0011\u0018 6E\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003\u0003F\f\u00035K \u0015&R6E \nFIG. 13. Tetragonal magnetic anisotropy of Rh 2TSb com-\npounds.\nEnergy distributions Ea′(θ,φ) ofT= Fe (a), Fe 0.5Co0.5(b),\nand Co (c). The energies Ex,y,zare given in µeV. (Please\nnote the different energy scales.)\norbit interaction. The magnetic moments for m/bardblcare\nlisted in Table III. The ratio of the total orbital moment\nto the total spin moment, ml/ms, was used owing to the\nlarge differences between the magnetic moments for dif-\nferent compositions. Figure 14 shows the ratio ml/ms\nas a function of the difference in the energies in several\nmagnetisation directions [ hkl]. For both Rh 2CoSb and\nRh2FeSb, the ratio is largest for magnetisation along the\ncaxis ([001]) and lowest in the basal plane. This finding\ninvolves not only the ratio but also the orbital momenta\nthemselves, indicating that the orbital moment is not al-\nways largest when the magnetisation is along the easy\naxis (or in the easy plane). Here it depends at least par-\ntially on the angle between the magnetisation and caxis,\nas shown by the values for other directions.\nTo further examine the nature of the anisotropy, the\ncharge and spin density distributions were analysed with/s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51 /s48/s46/s52 /s48/s46/s53/s48/s46/s48/s50/s48/s46/s48/s51/s48/s46/s48/s52/s48/s46/s48/s53/s48/s46/s48/s54/s48/s46/s48/s55/s48/s46/s48/s56\n/s91/s49/s48/s48/s93/s91/s49/s49/s49/s93/s91/s49/s48/s49/s93/s77/s111/s109/s101/s110/s116/s32/s114/s97/s116/s105/s111/s32/s32/s32 /s109\n/s108/s32/s47/s32 /s109\n/s115\n/s69/s110/s101/s114/s103/s121/s32/s100/s105/s102/s102/s101/s114/s101/s110/s99/s101/s32/s32/s32 /s69\n/s104/s107/s108/s32 /s32/s69\n/s109/s105/s110/s32/s91/s109 /s101/s86/s93/s91/s48/s48/s49/s93\n/s91/s49/s49/s48/s93/s91/s49/s49/s49/s93/s91/s49/s48/s49/s93/s91/s48/s48/s49/s93/s32/s32/s32/s82/s104\n/s50/s67/s111/s83/s98\n/s32/s32/s32/s82/s104\n/s50/s70/s101/s83/s98\nFIG. 14. Directional dependence of the orbital moments of\nRh2T′Sb,T′= Co, Fe.\nNote that the orbital moments are given relative to the spin\nmoments for better comparison.\nrespect to the magnetisationdirection (comparealsoFig-\nures 3 and 8). As mentioned above, the symmetry\nchangeswhenthemagnetisationisappliedalongdifferent\ncrystallographic directions. The point group symmetry\nof the 2b sites occupied by Fe and Co is D4handD2dfor\nRh on 4d. Applying the magnetisation along one of the\nhigh-symmetry axes, i.e., the c([001]) or a([100]) axis,\nchanges the symmetry of the 2b sites to C4horC2h, re-\nspectively. As a result, the irreducible representations\nand basic functions depend on the magnetisation direc-\ntion. For C4h, they are ag,bg, andegwith the l= 2\nbasic functions dz2, (dx2−y2,dxy), and (dxz,dyz). For\nC2h, they are agandbgwith (dz2,dx2−y2,dxy) and (dxz,\ndyz). Similar differences appear for the 4d sites. The\ncharge and spin density distributions for different mag-\nnetisation directions are compared in Figure 15 for the\ncompounds containing only Fe or Co.\nAs mentioned above, the details of the charge density\nare not easily observed directly from the graph when the\nmagnetisation direction is changed, because the graph\nshows mainly the positions of the atoms. However, they\ncan be observed if one investigates the difference in the\ncharge distribution, which is plotted as ∆ ρ(r). It was\ncalculated for both compounds as the difference between\nthe chargedensities obtained assuming that the magneti-\nsation is parallel ( m/bardbl[001]) or perpendicular ( m/bardbl[100]) to\nthecaxis. Inbothcompounds, themagnetisationhasthe\nsame effect on ∆ ρ(r) near the Rh atoms. That is, the\ncharge distribution is rotated with the direction of the\nmagnetisation. In the same way, the Rh-based spin den-\nsities are affected by the magnetisation direction. They\nchange from [001]-aligned when m/bardbl[001] to [100]-aligned\nwhenm/bardbl[100], regardless of which 3 dtransition metal is\nused. The situation is different near the 3 dtransition\nmetals Fe and Co, where ∆ ρ(r) andσ(r) are affected\nvery differently by the magnetisation direction. The rea-9\n/s48/s99\n/s82/s104/s70/s101/s114\n/s48/s48/s49/s40 /s114 /s41/s91/s48/s48/s49/s93/s48/s46/s48\n/s48/s46/s51\n/s56/s46/s48\n/s50/s50/s54/s46/s52\n/s49/s48/s48/s48/s46/s48\n/s83/s98/s68/s114 /s40 /s114 /s41\n/s45/s48/s46/s50\n/s48/s46/s48\n/s48/s46/s50\n/s48 /s97/s48/s99\n/s82/s104/s67/s111\n/s91/s49/s48/s48/s93/s91/s48/s48/s49/s93/s48/s46/s48\n/s48/s46/s51\n/s56/s46/s48\n/s50/s50/s54/s46/s52\n/s49/s48/s48/s48/s46/s48\n/s83/s98\n/s82/s104\n/s50/s67/s111/s83/s98/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s82/s104\n/s50/s70/s101/s83/s98\n/s48 /s97/s91/s49/s48/s48/s93/s45/s48/s46/s50\n/s48/s46/s48\n/s48/s46/s50/s115 /s40 /s114 /s41/s59/s32 /s109 /s32/s124/s124/s32/s91/s48/s48/s49/s93 /s115 /s40 /s114 /s41/s59/s32 /s109 /s32/s124/s124/s32/s91/s49/s48/s48/s93\n/s48/s46/s48\n/s48/s46/s49\n/s49/s46/s48\n/s48 /s97/s91/s49/s48/s48/s93/s48 /s97/s91/s49/s48/s48/s93/s48/s46/s48\n/s48/s46/s49\n/s49/s46/s48\nFIG. 15. Electronic structure of Rh 2FeSb and Rh 2CoSb.\nFully relativistic charge ( ρ(r)) and spin ( σ(r)) density distributions in a (100)-type plane for magnetis ation parallel and\nperpendicular to the caxis are shown. ρ001(r) is the charge density for m/bardbl[001], and ∆ ρ=ρ001−ρ100is the difference between\nthe charge densities for m/bardbl[001] and m/bardbl[100]. (Note: colour bars are in atomic units.)\nson is the different occupation of 3 dvalence electrons of\nFe (nFe\nd= 6.6) and Co ( nCo\nd= 7.8), which are responsible\nfor the different spin moments. The overall differences in\nthe charge and spin densities at different magnetisation\ndirections result in different total energies.\nFinally, the gain or loss of energy with changes in the\nmagnetisation direction results in the magnetocrystalline\nanisotropy. The electronic structure of the two com-\npounds, Rh 2FeSb and Rh 2CoSb, differs depending onthe\nmagnetisation direction, which is reflected in the change\nin the anisotropy from the easy plane to the easy axis\nwhen Fe is replaced with Co.\nIV. CONCLUSIONS\nThe electronic and magnetic structure of tetrag-\nonal Heusler compounds with the composition\nRh2FexCo1−xSb were investigated by ab initio cal-\nculations. The calculations revealed that the magnetic\nmoment increases and the Curie temperature decreases\nwith increasing Fe content x. The Rh atoms have onlysmall, composition-independent magnetic moments.\nThe magnetic properties are determined by those\nof the Fe and Co atoms and thus depend strongly\non the composition. The total energies for various\nmagnetisation directions were calculated to determine\nthe magnetic anisotropy. The analysis is described in\ndetail in an extended Appendix. For bulk materials,\nthe magnetocrystalline anisotropy is found to be much\nstronger (by three orders of magnitude) than the dipolar\nanisotropy. Special attention was given to the borderline\ncompounds, Rh 2FeSb and Rh 2CoSb. The most striking\nresult was that a composition-dependent transition from\neasy-axis to easy-plane anisotropy occurs at an iron\nconcentration of approximately 40%.\nAppendix A: Disorder\n1. Rh 2CoSb with Co–Sb-type antisite disorder\nSupplementary calculations were performed for dis-\nordered Rh 2CoSb and Rh 2FeSb using Sprkkr us-10\ning the coherent potential approximation. For ex-\nample, the disordered compound may be written as\nRh2(Co1−x/2Sbx/2)(Cox/2Sb1−x/2), where xis the disor-\nder level. The result for x= 1, which denotes complete\nCo–Sb disorder, is illustrated in Figure 1(b). Alterna-\ntively, it can be assumed that disorder between the Co\nand Rh atoms decreases the magnetic moments, which is\nconsistent with the results of calculations of the inverted\nstructure in space group 119, but not with those when\nCo–Sb disorder is assumed, as shown below.\nThe evolution of the magnetic moments of Rh 2CoSb\nand Rh 2FeSb with increasing disorder is shown in Fig-\nure 16. The total magnetic moment in the fully disor-\ndered state is approximately 20% larger for Rh 2CoSb\nand approximately 10% larger for Rh 2FeSb than those\nof the compounds in the completely ordered state. The\norbital moments are nearly constant in both compounds\nand are independent of the degree of disorder ( x). The\ndecrease in the total moments with decreasing xis at-\ntributed to the decrease in the spin magnetic moments\nof both compounds.\n/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53\n/s45/s49/s46/s53/s45/s49/s46/s48/s45/s48/s46/s53/s48/s46/s48\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53/s51/s46/s48/s51/s46/s53/s52/s46/s48\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s65/s110/s105/s115/s111/s116/s114/s111/s112/s121/s32/s32/s32 /s75\n/s117/s40/s120 /s41/s32/s91/s77/s74/s47/s109/s51\n/s93\n/s82/s104\n/s50/s40/s67/s111\n/s49/s45/s120/s83/s98\n/s120/s41/s40/s67/s111\n/s120/s83/s98\n/s49/s45/s120/s41/s82/s104\n/s50/s40/s70/s101\n/s49/s45/s120/s83/s98\n/s120/s41/s40/s70/s101\n/s120/s83/s98\n/s49/s45/s120/s41/s77/s97/s103/s110/s101/s116/s105/s99/s32/s109/s111/s109/s101/s110/s116/s32/s32/s32 /s109 /s40/s120 /s41/s32/s91 /s109\n/s66/s93\n/s68/s105/s115/s111/s114/s100/s101/s114/s32/s108/s101/s118/s101/s108/s32/s32/s32 /s120/s32/s109\n/s116/s111/s116\n/s32/s109\n/s115\n/s32/s109\n/s108/s32/s109\n/s116/s111/s116\n/s32/s109\n/s115\n/s32/s109\n/s108\nFIG. 16. Disorder-induced changes in magnetic moments and\nanisotropy of Rh 2(T1−x/2Sbx/2)(Tx/2Sb1−x/2)T= Co, Fe.\nTotal, spin, and orbital magnetic moments are shown, along\nwith the uniaxial anisotropy constant Kuas a function of\ndisorder level x.\nFigure 16 shows that the type of anisotropy (easy axis\nfor Rh 2CoSb and easy plane for Rh 2FeSb) is retained\neven in the completely disordered state. However, the\nabsolute value of the second-order uniaxial anisotropy\nconstant Kudecreases. That is, the anisotropy becomes\nweaker with increasing disorder. The Kuvalues of both\ncompounds are approximately 70% higher in the com-\npletely disordered state. No direct correspondence is\nobserved between the behaviour of Kuand that of thespin, orbital, or total magnetic moments. The effects of\ndisorder and composition on the magnetic anisotropy of\nthe Co–Fe system were investigated using first principles\nCPA calculations by Turek et al.[20], who also observed\na decrease in anisotropy with increasing disorder.\n2. Off-stoichiometric alloys\nIn many experiments, 2:1:1stoichiometrywasnot fully\nreached, but an excess of Fe or Co and a deficiency of Sb\nwas obtained. In particular, the magnetic properties of\nRh2T1+xSb1−x, withT= Fe, Co, were calculated for\nx= 0.12. As in the study of disorder, the calculations\nwere performed using the CPA. The 2 asite is assumed\nto be occupied by 12% with Fe (or Co) and by 88% with\nSb, whereas the occupations of the 4 dand 2bsites are\nunchanged.\nThe calculated magnetic properties of the off-\nstoichiometric alloys are listed in Table V. The magnetic\nmoments and spin stiffness D0are enhanced in both al-\nloys, and the values are higher than those of the stoi-\nchiometric compound. In particular, the excess Co and\nFe atoms on the 2 asite contribute a large spin moment.\nThe total magnetic moment, ms+ml= 2.627µB, of\nRh2Co1.12Sb0.88is very similar to the experimentally ob-\nserved value of 2 .6µB. The Curie temperature of the\noff- stoichiometric Co-containing compound is slightly\nlower,whereasthatoftheFecompoundisslightlyhigher.\nThese findings, along with the spin stiffness results,\nsuggest that the exchange coupling parameters of the\nstoichiometric compounds differ from those of the off-\nstoichiometric alloys.\nThe type of anisotropy (easy plane or easy axis) is the\nsame in the off-stoichiometric alloys as in the stoichio-\nmetric compounds. The uniaxial anisotropy constants\ndiffer, however. They are enhanced in the Fe alloy and\nreduced in the Co alloy.\nAppendix B: Semi-relativistic band structures\nThe semi-relativistic band structures of Rh 2FeSb and\nRh2CoSb arecompared in Figure 17to illustrate the spin\ncharacteristics of the bands. The band structures are\nsimilar; the main differences result from the larger band\nfilling in the Co-based compound, which has one more\nvalence electron than the Fe compound. Further, the\nlarger spin splitting in the Fe compound clearly results\nin a large spin magnetic moment.\nAppendix C: Magnetocrystalline anisotropy\nIn this Appendix, the discussion of the magnetocrys-\ntalline anisotropy is extended beyond simple uniaxial ap-\nproximations. The magnetocrystalline energy of uniaxial11\nTABLE V. Calculated magnetic properties of off-\nstoichiometric Rh 2T1.12Sb0.88.\nSpinmsand orbital mlmagnetic moments per atom (Rh,\nCo, Fe) and those of the primitive cell ( total) are listed, as\nwell as the Curie temperature TCand spin stiffness D0.m2b\ns,l\nrepresents the magnetic moments at the original position,\nandm2a\ns,lrepresents the moments of the excess Fe and Co\natoms at the initial Sb position.\nRh2T1.12Sb0.88 T= Fe T= Co\nmRh\ns[µB] 0.308 0.244\nmRh\nl[µB] 0.012 0.009\nm2b\ns[µB] 2.988 1.691\nm2b\nl[µB] 0.092 0.140\nm2a\ns[µB] 3.564 2.521\nm2a\nl[µB] 0.072 0.155\nmtotal\ns[µB] 4.010 2.453\nmtotal\nl[µB] 0.124 0.174\nTC[K] 490 480\nD0[meV˚A2] 690 1100\nKu[MJ/m3] -1.667 0.826\n/s45/s54/s45/s52/s45/s50/s48/s50\n/s77 /s78 /s83\n/s48/s88 /s77\n/s48/s45/s54/s45/s52/s45/s50/s48/s50/s82/s104\n/s50/s70/s101/s83/s98/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s107 /s41/s32/s45/s32\n/s70/s32/s91/s101/s86/s93\n/s82/s104\n/s50/s67/s111/s83/s98/s69/s110/s101/s114/s103/s121/s32/s32/s32 /s69 /s40/s107 /s41/s32/s45/s32\n/s70/s32/s91/s101/s86/s93\n/s69/s108/s101/s99/s116/s114/s111/s110/s32/s109 /s111/s109 /s101/s110/s116/s117/s109 /s32/s32/s32 /s107\nFIG. 17. Semi-relativistic band structures of Rh 2FeSb and\nRh2CoSb.\nRed and blue indicate majority and minority states, respec-\ntively.\nsystems can be derived from the first principles total en-\nergiesfordifferentmagnetisationdirections(quantisation\naxes).\nTextbooks give different descriptions of the magnetic\nanisotropy, in particular, different equations for the\nanisotropy constants [21–26]. Therefore, care must be\ntaken when comparing the results of this workwith those\nof other studies or comparing other studies with eachother.\n1. Uniaxial magnetic anisotropy\nIt is often assumed that the magnetocrystalline\nanisotropy in tetragonal or hexagonal systems is simply\ndescribed by a second-order dependence on the angle θ\nbetween the caxis and the magnetisation direction, that\nis,\nKusin2(θ), (C1)\nwhereKuis the uniaxial anisotropy constant. In that\ncase,\nKu=E100−E001(C2)\nissimplycalculatedfromthedifferencebetweentheen-\nergies for magnetisation along the principal axes, c/bardbl[001]\nanda/bardbl[100]. For Ku>0, theeasyaxis is along the c\naxis, whereas Ku<0 describes an easyplane where c\nis thehardaxis. For a distinct magnetic anisotropy in\ntheabplane, it would be more accurate to use the lowest\nenergy of the two in-plane directions along the princi-\npal axis and the diagonal, which are the [100] and [110]\ndirections, respectively:\nKu= min(E100,E110)−E001. (C3)\nEquation (C1) has another serious drawback; namely,\nthe anisotropy is completely independent of the crystal\nlattice, and the anisotropic energy distribution always\nhas the same shape regardless of the c/aparameter and\nwhether the crystal has tetragonal, hexagonal, or some\nother structure. That is, Equation (C1) is ultimately\nuseful only for distinguishing between easyandhardc\naxes.\nNow we consider only tetragonal systems. By using\nthe series expansion/summationtextK2ν,0sin2ν(θ) up to the fourth\norder in sin( θ), the uniaxial magnetocrystalline energy is\nexpressed as\nEuniaxial\ncrys=K0+K2sin2(θ)+K4sin4(θ).(C4)\nThe equations for the sixth-order uniaxial anisotropy\nare discussed by Jensen and Bennemann [27], for exam-\nple. In the following, the subscript ”crys”is omitted, and\nthe energies are indexed only by direction or by ”uni”.\nFor the high-symmetry directions [ h,k,l] and the low-\nest indices ( h,k,l= 0,1), the energies depend on the\nanisotropy coefficients as follows:12\nE001=K0, (C5)\nE100=K0+K2+K4,or\nE110=K0+K2+K4,and\nE101=K0+K2sin2(θ101)+K4sin4(θ101),or\nE111=K0+K2sin2(θ111)+K4sin4(θ111).\nNote that the energiesfor the [100]and [110]directions\nare identical only when uniaxial anisotropy is assumed.\nThe energies for the [101] and [111] directions, however,\nhave different angles with respect to the caxis. From\nEquations (C4 and C5), K2andK4may be obtained,\nfor example, from the differences:\nE100−E001=K2+K4and (C6)\nE101−E001=K2sin2(θ)+K4sin4(θ).\nForz=c/a, the angle θis found using θ101=θ011=\narctan(1/z). FromEquation(C5)or(C6), theanisotropy\nconstants Kiare given by\nK0=E001, (C7)\nK2= (E101−E001)(z2+2)+(E101−E100)1\nz2,\nK4= (E001−E101)(z2+1)+(E100−E101)(z2+1)\nz2.\nAlternatively, E111andθ111= arctan(√\n2/z) may be\nused, but the resulting equations will have a different\ndependence on c/a. The uniaxial magnetocrystalline\nanisotropyenergy( Eu)isthedifferencebetweenthemag-\nnetocrystalline energy (here Euni) and the isotropic con-\ntribution, which is the spherical part K0:\nEu=Euni−K0. (C8)\nAccording to this equation, the uniaxial magnetocrys-\ntalline anisotropy energy may be positive or negative,\ndepending on the directions and values of K(see also\nAppendix C4).\nEquation (C4) has four extremal values at\nθi= 0,π\n2,and±arcsin/parenleftBigg/radicalbigg\n−K2\n2K4/parenrightBigg\n,(C9)\nwhere the first derivative of the fourth-order equation\n[Equation (C4)] vanishes, that is, for dEuniaxial/dθ= 0.\nThesolutions θ3,4arerealonlyiftheanisotropyconstants\nobey the relation 0 ≤−K2\n2K4≤1, that is, K2K4≥0,\n|K2| ≤2|K4|. ForK2=−2K4, one has θ3,4=±90◦. For\na realθ3,4, one has an easyor ahardcone. The resulting\nextremal energies areE(0) =K0, (C10)\nE(π/2) =K0+K2+K4,\nE(θ3,4) =K0−K2\n2\n4K4.\nThe minima or maxima are obtained using the second\nderivatives of the energy at the extremal angles:\nd2E\ndθ2/vextendsingle/vextendsingle/vextendsingle/vextendsingle\n0= 2K2, (C11)\nd2E\ndθ2/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nπ/2=−2(K2+2K4),\nd2E\ndθ2/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nθ3,4=−2K2(K2+2K4)\nK4.\nThe minima appear for positive 2ndderivatives\n(d2E/dθ2/vextendsingle/vextendsingle\nθi>0)anddefinetheeasydirection(s)ofmag-\nnetisation. Indeed, one has to search for the absolute\nminimum and maximum to find the correct easy and\nhard axes, planes, or cones. An easy cone appears for\nK2<0,K4>−K2/2, and the corresponding cone angle\nis given by θ3,4. A special hard cone exists for K2>0,\nK4=−K2, where both the caxis and the abplane have\nthe same (lowest) energy. The energy barrier at the hard\ncone must be overcome, however, to change the magneti-\nsation direction from the easy axis to the easy plane and\nvice versa. In the range −∞< K4<−K2/2, the solu-\ntions are metastable when K2>0. The complete fourth-\norder uniaxial anisotropy phase diagram is presented in\nTable VI.\nTABLE VI. Uniaxial anisotropy phase diagram.\nabstands for basal plane, cstands for c-axis. Cones may have\nan opening angle θorπ/4 with respect to the caxis.\nK2 K4 easy hard\n>0 −∞···− K2 ab cone (θ)\n>0 −K2 ab,c cone (45◦)\n>0−K2···−K2/2 c cone (θ)\n>0 −K2/2···∞ c ab\n= 0 = K2= 0 undefined, spherical\n<0 −K2···∞ cone (θ) ab\n<0 −K2 cone (45◦) ab,c\n<0−K2/2···−K2 cone (θ) c\n<0−∞···− K2/2 ab c\nThe magnetic anisotropy phase diagram for fourth-\norder uniaxial anisotropy is displayed in Figure 18. It is\nsimilartothegraphicalrepresentationsreportedinRefer-\nences[27,28]. Thedifferentphasesaredistinguished. For\nK4=−K2<0, there is a distinct metastable case with\nequal energies for magnetisation along the caxis and in13\ntheabplane. At this line, a transition occurs from easy-\naxistoeasy-planebehaviour. Inthemetastableregionfor\nK4<−K2/2<0, easy-axis behaviour appears, whereas\neasy- plane behaviour appears for K4<−K2<0 (see\nTable VI). In both cases, the sizes of the anisotropy con-\nstants determine how easily one state can switch to the\nother and the stability of the state with lower energy.\nThe energy barrier to cross the hard cone has a size of\n−K2\n2\n4K4, as mentioned above.\n/s75\n/s52\n/s75\n/s50/s101/s97/s115/s121/s32/s112/s108/s97/s110/s101/s101/s97/s115/s121/s32/s99/s111/s110/s101 /s101/s97/s115/s121/s32/s97/s120/s105/s115\n/s109/s101/s116/s97/s32/s115/s116/s97/s98/s108/s101\nFIG. 18. Magnetic anisotropy phase diagram.\nIn themetastability range,hard-cone -type anisotropy occurs.\nIn the sketches of the E(θ) distributions, it is assumed that\nK0= 0.\n2. Tetragonal magnetic anisotropy\nThe uniaxial magnetic anisotropy does not reflect the\nsymmetry of the crystal structure. The symmetry of the\nanisotropy should generally be the same as the symme-\ntry of the crystal potential; thus, it is given by the fully\nsymmetric irreducible representation of the point group,\nthat is,a,a1,ag, or similar. Again, by using a series\nexpansion up to the fourth order in sin( θ), the magne-\ntocrystalline energy of a tetragonal system is expressed\nas\nEtetragonal\ncrys =2/summationdisplay\nν=0K2ν,0sin2ν(θ)+K4,4sin4(θ)f(φ),\n=Euniaxial\ncrys+K4,4sin4(θ)f(φ),\nf(φ) = cos(4 φ).\nf(φ) has an azimuthal dependence on 4 φ, which re-\nsults in the expected fourfold symmetry. Some worksusedf′(φ) = sin4(φ)+cos4(φ), which results in different\nequations and Kvalues. Higher-order approximations\nwill include terms with K6,0,K6,4,K8,0,K8,4,K8,8, and\nso on. Subtracting K0from Equation (C12) yields\nEa(/vector r) =K2,0sin2(θ)+[K4,0+K4,4cos(4φ)]sin4(θ).\n(C12)\nIn the following, the subscript ”crys” is omitted, and\nthe energies are indexed only by direction or by ”tet”.\nFor the high-symmetry directions [ h,k,l] and the lowest\nindices (h,k,l= 0,1), Equation (C12) gives\nE001=K0,0, (C13)\nE100=K0,0+K2,0+K4,0+K4,4,\nE110=K0,0+K2,0+K4,0−K4,4,and\nE101=/summationdisplay\nν=0,2K2ν,0sin2ν(θ101)+K4,4sin4(θ101),or\nE111=/summationdisplay\nν=0,2K2ν,0sin2ν(θ111)−K4,4sin4(θ111).\nForz=c/a, the angle θ101is found using θ101=\nθ011= arctan(1 /z). Alternatively, E111withθ111=\narctan(√\n2/z) may be used. From the first four ener-\ngies of Equation (C13), the anisotropy constants Kl,m\nare found to be\nK0,0=E001, (C14)\nK2,0= (E101−E001)(z2+2)\n+(E101−E100)1\nz2,\nK4,0= (E001−E101)(z2+1)\n+(E100−E101)1\nz2\n+1\n2(E100+E110−2E101),\nK4,4=1\n2(E100−E110).\nThe magnetocrystalline anisotropy energy ( Ea) is the\ndifference between the magnetocrystalline energy (here\nEtet) and the isotropic contribution, which is the spher-\nical part K0:\nEa=Etet−K0. (C15)\n3. Dipolar magnetic anisotropy\nIn non-cubic systems, the dipolar anisotropy does not\nvanish and also contributes to the magnetocrystalline\nanisotropy. It is calculated from a direct lattice sum\nyielding the dipolar energy:14\nEdip(/vector n) =µ0\n8π/summationdisplay\ni/negationslash=j/bracketleftBigg\n/vector mi·/vector mj\nr3\nij−3(/vector rij·/vector mi)(/vector rij·/vector mj)\nr5\nij/bracketrightBigg\n,\n(C16)\nwhere/vector n=/vectorM/Mis the magnetisation direction, and\nrijrepresents the distance vectors between the magnetic\nmoments miandmj. The individual magnetic moments,\n/vector miand/vector mj, do not necessarily have to be collinear in\ngeneral.\nIn a simplified picture, only the 3 dtransition elements\nTcarry a significant magnetic moment in the Rh 2TSb\ncompoundsinvestigatedhere. Inallcasesofasinglemag-\nneticionwhereallthemagneticmomentsinthestructure\nare collinear along /vector n, the equation simplifies to\nEdip(/vector n) =µ0m2(/vector n)\n8π/summationdisplay\ni/negationslash=j1\nr3\nij/bracketleftBigg\n1−3r2\nn,ij\nr2\nij/bracketrightBigg\n,\n=µ0m2(/vector n)\n8π/summationdisplay\ni/negationslash=j1−3cos2(θij)\nr3\nij,(C17)\nwherern,ij=rn,ij(/vector n) is a projection of the position\nvector onto the direction of the magnetic moment, and\nθijis the angle between them. In Equation (C17), the\nsign of the energy is completely defined by the crystal\nstructure when the summation is over a spherical parti-\ncle. Note that the size of the magnetic moment, m(/vector n),\ndepends on the magnetisation direction when the spin–\norbit interaction is taken into account.\nFinally, the dipolar anisotropy is given by the differ-\nence between the energies for two different directions,∆Edipaniso=E(/vector n2)−E(/vector n1). Again, the two well- dis-\ntinguished directions are the /vector n1= [001] and /vector n2= [100]\ndirections, which are along the caxis and in the basal\nplane along a, respectively. Positive values indicate an\neasy dipolar direction that is along the [001] axis. It has\na second-order angular dependence.\n4. Plotting the magnetic anisotropy\nAccording to Equations (C8) and (C15), the magne-\ntocrystalline anisotropy energy may be positive or neg-\native, depending on the direction of ( θ,φ) and the K\nvalues. Consequently, it is difficult to visualise the\nanisotropy energy by plotting the three-dimensional dis-\ntribution of Ea(/vector r) =Ea(θ,φ). Therefore, the alternative\nanisotropy energy Ea′with respect to the lowest energy\nis generally plotted, where\nEa′=Ea−min(Ea), (C18)\nwhich is still positive even when Ea<0. The easy\ndirections or planes are identified as those for which\nEa′= 0.Ea′is used to plot the magnetocrystalline\nanisotropy in the main text.\nACKNOWLEDGMENTS\nWe thank the groups of P. Blaha (Vienna) and H.\nEbert (Munich) for providing their computer codes.\n[1] S. K. Dhar, A. K. Grover, S. K. Malik, and R. Vija-\nyaraghavan, Peaks in low field a.c. susceptibility of ferro-\nmagnetic Heusler alloys, Sol. St. Comm. 33, 545 (1980).\n[2] S. V. Faleev, Y. Ferrante, J. Jeong, M. G. Samant,\nB. Jones, and S. S. P. Parkin, Heusler compounds with\nperpendicular magnetic anisotropy and large tunnel-\ning magnetoresistance, Phys. Rev. Materials 1, 024402\n(2017).\n[3] P.Blaha, K.Schwarz, P.Sorantin,andS.B.Trickey,Full -\npotential, linearized augmented plane wave programs for\ncrystalline systems, Comput. Phys. Commun. 59, 399\n(1990).\n[4] K. Schwarz and P. Blaha, Solid state calculations using\nWIEN2k, Comput. Mater. Sci. 28, 259 (2003).\n[5] P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka,\nand J. Luitz, WIEN2k: An Augmented PlaneWave + Lo-\ncal Orbitals Program for Calculating Crystal Properties\n(Wien, 2013).\n[6] H. Ebert, Fully relativistic band structure calculatio ns\nfor magnetic solids - formalism and application, in Elec-\ntronic Structure and Physical Properties of Solids. The\nUse of the LMTO Method , Lecture Notes in Physics, Vol.\n535, edited by H. Dreysse (Springer-Verlag, Berlin, Hei-delberg, 1999) pp. 191 – 246.\n[7] H. Ebert, D. K¨ odderitzsch, and J. Minar, Calculat-\ning condensed matter properties using the KKR-Greens\nfunction method – recent developments and applications,\nRep. Prog. Phys 74, 096501 (2011).\n[8] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized\nGradient Approximation Made Simple, Phys. Rev. Lett.\n77, 3865 (1996).\n[9] H. C. Kandpal, G. H. Fecher, and C. Felser, Calculated\nelectronic and magnetic properties of the half-metallic,\ntransition metal based Heusler compounds, J. Phys. D:\nAppl. Phys. 40, 1507 (2007).\n[10] G. H. Fecher, S. Chadov, and C. Felser, Theory of the\nhalf-metallic heusler compounds, in Spintronics , edited\nbyC.FelserandG.H.Fecher(SpringerVerlag,Dordrecht\nHeidelberg New York London, 2013) Book section 7, p.\n115.\n[11] S. Mankovsky, G. H. Fecher, and H. Ebert, Electronic\nstructure calculations in ordered and disordered solids\nwith spiral magnetic order, Phys. Rev. B 83, 144401\n(2011).\n[12] J. Thoene, S. Chadov, G. Fecher, C. Felser, and\nJ. K¨ ubler, Exchange energies, Curie temperatures and15\nmagnons in Heusler compounds, J Phys. D: Appl. Phys.\n42, 084013 (2009).\n[13] P. Soven, Coherent-Potential Model of Substitutional\nDisordered Alloys, Phys. Rev. 156, 809 (1967).\n[14] S. A. Khan, P. Blaha, H. Ebert, J. Minar, and O. Sipr,\nMagnetocrystalline anisotropy of FePt: A detailed view,\nPhys. Rev. B 94, 144436 (2016).\n[15] S. J. Joshua, Symmetry principles and magnetic symme-\ntry in solid state physics. (Adam Hilger, IOP Publishing\nLtd., Bistol, Philadelphia, New York, 1991).\n[16] R. Lizarraga, L. Nordstr¨ om, L. Bergqvist, A. Bergman,\nE. Sj¨ ostedt, P. Mohn, and O. Eriksson, Conditions for\nNoncollinear Instabilities of Ferromagnetic Materials,\nPhys. Rev. Lett. 93, 107205 (2004).\n[17] A.I.Liechtenstein, M. I.Katsnelson, andV.A.Gubanov ,\nExchange interactions and spin-wave stiffness in ferro-\nmagnetic metals, J. Phys. F: Met. Phys. 14, L125 (1984).\n[18] A.I.Liechtenstein, M.I.Katsnelson, V.P.Antropov,a nd\nV. A. Gubanov, Local Spin Density Functional Approach\nto the Theory of Exchange Interactions in Ferromagnetic\nMetals and Alloys, J. Magn. Magn. Mater. 67, 65 (1987).\n[19] M. Pajda, J. Kudrnovsky, I. Turek, V. Drchal, and\nP. Bruno, Ab initio calculations of exchange interactions,\nspin-wave stiffness constants, and Curie temperatures of\nFe, Co, and Ni, Phys. Rev. B 64, 174402 (2001).[20] I. Turek, J. Kudrnovsky, and K. Carva, Mag-\nnetic anisotropy energy of disordered tetrago-\nnal Fe-Co systems from ab initio alloy theory,\nPhys. Rev. B 86, 174430 (2012).\n[21] R. Skomsky and J. M. D. Coey, Permanent Magnetism ,\nStudiesinCondensed Matter Physics(Taylor andFrancis\nGroup, New York, 1999).\n[22] J. K¨ ubler, Theory of Itinerant Electron Magnetism\n(Clarendon Press, Oxford, 2000).\n[23] B. D. Cullity and C. D. Graham, Introduction to Mag-\nnetic Materials; 2nd Ed. (JohnWileyandSons,Hoboken,\n2009).\n[24] J.M.D.Coey, Magnetism and Magnetic Materials (Cam-\nbridge University Press, Cambridge, 2010).\n[25] N. A. Spaldin, Magnetic Materials, 2nd Ed. Fundamen-\ntals and Applications (Cambridge UniversityPress, Cam-\nbridge, 2011).\n[26] K. M. Krishnan, Fundamentals and Applications of Mag-\nnetic Materials (Oxord University Press, Oxord, 2016).\n[27] P. J. Jensen and K. H. Bennemann, Magnetic structure\nof films: Dependence on anisotropy and atomic morphol-\nogy, Srf. Sci Rep. 61, 129 (2006).\n[28] R. Skomski, H.-P. Oepen, and J. Kirschner, Unidi-\nrectional anisotropy in ultrathin transition-metal films,\nPhys. Rev. B 58, 11138 (1998)." }, { "title": "2009.01034v1.New_highly_anisotropic_Rh_based_Heusler_compound_for_magnetic_recording.pdf", "content": "1 \n New highly -anisotropic Rh-based Heusler compound for \nmagnetic recording \nYangkun He1,*, Gerhard H. Fecher1,*, Chen guang Fu1, Yu Pan1, Kaustuv Manna1, Johannes \nKroder1, Ajay Jha2, Xiao Wang1, Zhiwei Hu1, Stefano Agrestini3, Javier Herrero -Martí n4, \nManuel Valvidares4, Yurii Skourski5, Walter Schnelle1, Plamen Stamenov2, Horst Borrmann1, \nLiu Hao Tjeng1, Rudolf Schaefer6,7, S. S. P. Parkin8, J. M. D. Coey2 and Claudia Felser1 \n \n1Max-Planck -Institute for Chemical Physics of Solids, D -01187 Dresden, Germany \n2School of Physics, Trinity College, Dublin 2, Ireland \n3 Diamond Light Source, Harwell Science and Innovation Campus, Didcot, OX11 0DE, \nUK \n4ALBA Synchrotron Light Source, Cerdany ola del Valles, 08290 Barcelona, Catalonia, \nSpain \n5 Dresden High Magnetic Field Laboratory (HLD -EMFL), Helmholtz -zentrum \nDresden–Rossendorf, 01328 Dresden, Germany \n6Leibniz Institute for Solid State and Materials Research (IFW) Dresden, Helmholtz \nstrasse 20, D -01069 Dresden, Germany \n7Institute for Materials Science, TU Dresden, D -01062 Dresden, Germany \n8Max Planck Institute of Microstructure Physics, Halle, Germany \nKey words: magnetocrystalline anisotropy, tetragonal Heusler alloy , magnetic \nhardness parameter , 4d magnetism , magnetic recording \n \nThe development of high -density magnetic recording media is limited by the \nsuperparamagnetism in very small ferromagnetic c rystals . Hard magnetic \nmaterials with strong perpendicular anisotropy offer stability and high recording \ndensity . To overcome the difficulty of writing media with a large coercivity, heat \nassisted magnetic recording (HAMR) has been developed , rapidly heat ing the \nmedia to the Curie temperature Tc before writing , followed by rapid c ooling. \nRequirements are a suitable Tc, coupled with anisotropic thermal conductivity \nand hard magnetic properties . Here we introduce Rh 2CoSb as a new hard \nmagnet with potential for thin film magnetic recording . A magnetocrystalline \nanisotropy of 3.6 MJm-3 is combined with a saturation magnetization of μ0Ms = \n0.52 T at 2 K (2.2 MJm-3 and 0.44 T at room -temperature) . The magnetic \nhardness parameter of 3.7 at room temperature is the highest observed for a ny \nrare-earth free hard magnet . The anisotropy is related to an unquenched orbital \nmoment of 0.42 µB on Co, which is hybridized with neighbouring Rh atoms with \na large spin -orbit interaction. Moreover, the pronounced \ntemperature -dependence of the anisotropy that follows from its Tc of 450 K, \ntogether with a high thermal conductivity of 20 Wm-1K-1, makes Rh 2CoSb a 2 \n candidate for development for heat assisted writing with a recording density in \nexcess of 10 Tb/in2. \n \nThe pace of doubling of information density on magnetic recording media has \nslackened in recent years, as the effective size of the perpendicularly recorded grains \napproached the superparamagnetic blocking diameter , which is the low er size limit for \nstable ferromagnetism , directly related to the magnetocrystalline anisotropy energy K1. \nTo resist demagnetization by random thermal fluctuation, the volume V of a magnetic \nmaterial must satisfy the empirical condition that K1V/kBT > 60 (K1V > 1.5 eV), where \nkB and T are the Boltzmann constant and ambient temperature , respectively1-3. For \nhigh-density magnetic recording media , strong perpendicular uniaxial anisotropy is \nrequired . To create a film for magnetic recording or a permanent magnet that remains \nfully magnetized regardless of its shape, K1 should exceed µ 0Ms2, where Ms is the \nspontaneous saturation magnetization. Therefore , the magnetic hardness parameter κ \n= \n , a convenient figure of merit for permanent magnets4, should be \nlarger than 1 . This is difficult to achieve in rare-earth free materials , other than CoPt \nor FePt with L1 0 structure. \nThe development of modern magnetic recording media spanned three \ngenerations. The first generation for tapes and discs depended on the s hape anisotropy \nof acicular fine particles of ferrimagnetic or ferromagnetic oxides, γFe2O3 or CrO 25, \nwhich were magnetized in-plane. The second generation was based on hexagonal \nCo-Cr-Pt thin films with perpendicular magnetization , which is used in current hard \ndiscs6. But Cr decreases the K1 of Co-Pt alloy , limiting the recording density . An ideal 3 \n material for recording should be hard for storage and soft for writing , because of the \nlimited fields that can be generated by the miniature writing electromagnet. L1 0 FePt, \na hard magnet with K1 = 6.6 MJm-3 and κ = 2.02, is the basis of new , third generation \nof heat-assisted magnetic recording (HAMR ) media7,8, where writing is realized by \nheating the material close to its Curie point with a laser -powered near -field \ntransducer9. However, its high Tc of 750 K makes rapid heating and cooling \nproblem atic and t akes time10. To achieve a large temperature -variation in K1, 10 % Cu \nis doped to obtain a 100 K reduction in Tc, but this is accompanied by a big drop of K1 \nto 0.8 MJm-33,11. Additionally , unavoidable disordered A1 FePt impurities with a \nsmall er K1 and a lower Tc make the stabili zation of its properties difficult12. Therefore, \nit is useful to look for new materials with a strong K1, a fairly suitable Tc and good \nthermal conductivity for development as new thin film media \nTo achieve sufficient anisotropy, a non-cubic crystal structure (for example \ntetragonal or hexagonal) and a large spin -orbit interaction with the right sign is \nneeded . When looking for new materials with uniaxial magnetocrystalline anisotropy , \na significant deviation of the c/a ratio from 1 for tetragonal or \n for hexagonal \nstructures , is sought . Moreover, it is important to pair a 3 d metal (Mn, Fe, Co) that \nprovide s a substantial magnetic moment with a heavy atom that enhances the \nspin-orbit interaction. Rh is a 4 d element where the spin-orbit interaction is large r \nthan in 3d elements , and it acquires a small induced magnetic moment when paired \nwith a 3 d metal . Therefore, it is worthwhile looking for new opportunities in magnetic \nrecording among Rh-based alloys with a tetragonal or hexagonal structure . 4 \n Heusler alloys are a large family of compounds with formula X 2YZ, where X \nand Y are usually transition metals and Z is a main group element13,14. The abundant \nchoice of elements provides great scope to search for new materials with specific \nproperties. Rh2CoSb has a c/√2a ratio of 1.24, the largest among all reported \nRh-based magnetic Heusler compounds ( c/√2a is used to describe the tetragonal \ndistortion relative to a cubic lattice). A magnetic moment per formula of 1.4 μB in an \n0.7 T applied field and a Tc of about 450 K were reported for polycrystalline samples \nmany years ago15. Recent ab-initio calculations proposed uniaxial anisotropy with an \neasy c-axis16. Both experiments and calculations suggest that Rh 2CoSb might be an \ninteresting hard magnetic material . \nHere we study the anisotropic magnetic , thermal and transport properties , as well \nas the temperature -dependent anisotropy of single crystals of Rh 2CoSb and investigate \nthe contribution of rhodium to the magnetism. \nResults \nCrystal structure \nMagnetic Heusler alloys with a tetragonal structure are often Mn-based with \nspace group I\nm217-22. However, the Mn-Mn magnetic coupling is usually \nantiferromagnetic at the nearest -neighbo ur separation in this structure , leading to \nferrimagnetic order with relatively low magnetization (Figure 1a). Some tetragonal \nRh-based Heusler alloys including Rh2Mn 1.12Sn0.88, Rh 2FeSn, and Rh 2CoSn and \nRh2CoSb15 with 3 d transition element s Mn, Fe and Co, crystallize in a tetragonal \nstructure with space group I4/mmm (D0 22 structure). The 4 d element Rh provides a 5 \n strong spin orbit interaction that contributes to the magnetocrystalline anisotropy in \nthe tetragonal lattice23. \nBoth XRD and TEM experiments show that Rh 2CoSb has a well -ordered \ntetragonal D022 structure with a = 4.0393 (6) Å and c = 7.1052 (7) Å. Here, the 4d (0 \n0 ¼) site is occupied by Rh, 2b (0 0 ½) by Co, and 2 a (0 0 0) by Sb as shown in \nFigure 1c. The tetragonal distortion c/\na = 1.2436(3) agree s with the published \nvalue15, being the largest among all reported Rh -based magnetic Heusler compounds. \nA sketch of the easy-axis energy surface is included in Fig ure 1c. \nUnlike many Mn-based Heusler compounds, which are often cubic at high \ntemperature and may undergo a martensitic phase transition to become tetragonal, \nRh2CoSb does not have a first -order transition at high temperature and remains in the \ntetragonal phase at all temperatures up to the melting point of 1482 K. A differential \nscanning calorimeter (DSC) measurement showing only the melting transition is \npresented in Supplementary Information . Since the phase melts congruently, single \ncrystals can be grown by the Bridgeman method. They were cut along the c and a \naxes for further measurements and fully characterized by x-ray diffraction (XRD), \nwavelength dispersive x -ray spectroscopy (WDX) and transmission electron \nmicroscopy (TEM) in Supplement al Information . WDX establishes a homogenous \ncomposition of Rh 50.3Co25.6Sb24.1(see Supplement). Errors in composition range from \n0.1 – 0.2 % . 6 \n \nRhSb\nE\nZ\nEc\nEb EaCo\nabc\n1.0 1.1 1.2 1.301234\nMs (B/f.u.)\nc / 2aRh2CoSb Rh2CoSnRh2Mn1.12Sb0.88\nRh2FeSbRh2FeSn\nMn2RhSn\nMn2NiGa Mn2FeGa\nMn3GaMn2RhSn\nMn2PtIn\nMn2PtGa\n(a) (b)\n(c) \nFigure. 1. Tetragonal structure of Rh 2CoSb. a, Tetragonal distortion versus magnetic moment \nfor Mn -based and Rh -based Heusler compound s. Rh-based compound s show both large \ndistortion and large magnetization. Note that some compounds do not saturate in the applied \nmagnetic fields of 7 T for Mn 2PtGa, 10 T for Mn 2FeGa, 6.6 T for Rh2CoSn and Rh2FeSn, and \n0.7 T for Rh2FeSb and Rh2Mn 1.12Sb0.88. b, Comparison of the magnetic hardness parameter κ \nwith other hard magnets. The light grey plane marks the threshold κ = 1. c, Unit cell of \ntetragonal Heus ler compound Rh 2CoSb, whose magnetocrystalline anisotropy surface shows \neasy-axis anisotropy along c. \n \nMagnetic properties \nMagnetic properties were measured along the c and a axes of single crystals as \nillustrated in Figure 2. A magnet o-optical Kerr microscopy study24 shows surface \ndomains of a two -phase branched domain pattern of higher generation on the (001) 7 \n surface, which is commonly observed in uniaxial magnets. At 2 K , the magnetization \nalong c saturates easily , at 37.5 Am2kg-1. Taking the density of 11 ,030 kg m-3 into \nconsideration , the saturation magnetization μ0Ms is 0.52 T, which is similar to that of \npure nickel . It correspond s to a moment of 2.6 μB per Rh2CoSb formula unit. The \nnon-integer magnetic moment per unit cell indicates that Rh2CoSb is not a half metal. \nHowever, a field μ0Ha of 17. 5 T is required to saturate the sample along the a axis, \nwhere Ha is the anisotropy field. The magnetocrystalline anisotropy constant K1 = ½ \nμ0HaMs is 3.6 MJm-3, which is larger than for any other rare-earth free compound \nexcept for CoPt and FePt . Its relatively small Ms makes larger than for these \nmaterials and coercivity may exceed that of FePt3. Both μ0Ms and K1 decrease with \nincreasing temperature, as shown in Fig ure 2b, but at room temperature they are still \n0.44 T and 2.2 MJm-3, respectively. Big Barkhausen jumps observed during c axis \ndemagnetization , together with a high initial susceptibility after thermal \ndemagnetization indicat e that single -crystal Rh2CoSb i s a nucleation type magnet25,26. \nDetailed measurements of the magnetization along c and a at different temperatures \nare reported in the Supplement al information . \nThe Curie temperature is deduced to be 450 K from the temperature scans of \nmagnetization in a 10 mT field. The M-T curve measured along the hard axis in a \nmagnetic field of 0.5 T exhibit s a sharp peak at 440 K , indicating a rapid built -up of \nanisotropy just below Tc. The slope d K1/dT is ~ 20 kJm-3K-1 (twice as large as for \nFePt27). The ac -susceptibility χ near Tc is shown in Figure 2d. It exhibit s a typical \nHopkinson peak along the c axis15,28. Since χ ∝ Ms2/K1, during cooling the rate of 8 \n increase of Ms and K1 vary at different temperatures near Tc, giving a trough in ac \nsusceptibility below the Hopkinson peak . There is nothing unusual in our M-T and \nM-H measurement s at this temperature. \nWe compare in Fig ure 1b the magnetic hardness parameter of Rh 2CoSb with \nother materials that exhibit μ0Ms > 0.4 T, which is taken as a threshold necessary for \nuseful stray fields. κ is a practical figure of merit for hard magnet ic materials that \nmust be be greater than 1 if the material is to resist self-demagnetization when \nfabricated into any desired shape4. Rh 2CoSb has κ = 4.1 at 2 K and 3. 7 at room \ntemperature, which is more than any other rare-earth free magnet. Only SmCo 5 has a \nlarger value. \nAll the indications are that Rh2CoSb has the potential to be a good hard magnet \nwith strong uniaxial anisotropy. There is no structural transition, avoiding the \nproblem s of decomposition ( MnBi decomposes at 628 K) or twinning (like many \nnearly -cubic rare-earth free magnets such as MnGa, MnAl, FePt and CoPt29). Unlike \nmany rare -earth magnets, the sample is stable in air and its magnetic properties have \nbeen found to remain unchanged for a year. \nThanks to the strong K1 at 300 K , Rh2CoSb promise s a significant reduction in \nthermally stable grain size from diameter D=7–9nm in today's perpendicular \nCo-Cr-Pt media down to D=4–5nm in future Rh 2CoSb media, resulting in a potential \nstorage density of more than 10 Terabit /inch2 (for D = 5 nm with half area occupied ). \nDetailed calculation can be found in Supplemental information . HAMR media with \ngrain diameters of only a few nm are hard to fabricate because the grains musts be 9 \n exchange decoupled by a n intergranular material. Powder XRD data revealed the \npresence of a few percent of a RhSb secondary phase in our polycrystalline samples. \nNonmagnetic RhSb appears at the grain boundar ies, pinning the domain s and creating \nhysteresis . It is a candidate for separating the nanocrystal grains in Rh2CoSb thin film \nmedia . \n0 5 10 15 20 25 30010203040\n0 100 200 300 400 5000510152025303540\n300 350 400 450 5000.000.010.020.030.040.050.060.07\n25 μmMagnetization (A m2 kg-1)\nMagnetic field (T)2 K H//a\n300 K H//a300 K H//c2 K H//c(a)\n(c) (d)(b)\n0 100 200 300 400 50001234\nTemperature (K)K1 (MJm-3)\n01234\n Magnetization (A m2 kg-1)\nTemperature (K) 0.5 T 0H // c\n 0.01 T 0H // c\n 0.5 T 0H // a\n 0.01 T 0H // a\n\nH // cH // a\n \n '\nTemperature (K)Tc \n \nFigure 2. Magnetic properties of Rh 2CoSb. a, Magnetization curve s at 2 K and 300 K with \nthe field along the a or c axes. The insert shows a pattern of branch ed domain s at the surface \nperpendicular to the c axis, typical of a strong uniaxial ferro magnet . b, Magnetocrystalline \nanisotropy calculated from magnetization curves at different temperatures. The blue line 10 \n shows the magnetic hardness parameter κ at different temperatures . c, Magnetization at \ndifferent temperatures with the field along a or c axes. For better visibility s ome values are \nenlarged by a factor of 10 or 100 , depend ing on the field direction and the value of H. d, \nAC-susceptibility near Tc, showing a typical Hopkinson peak , due to the rapid increase of \nanisotropy just below Tc. \nTransport properties. \nTransport properties of Rh 2CoSb are presented in Figure 3. The resistivity is \nvery anisotropic. With increasing temperature, the a-axis resistivity increases \nmonoton ically from 53 µΩ cm at 2 K to 192 µΩ cm at Tc, after which the resistivity , \ndominated by spin disorder scattering , tends to saturate30. The c-axis resistivity is less \nthan half as large at 2 K , 21 µ Ω cm, but it increas es with temperature and show s a \nsimilar trend to the a-axis resistivity . The difference is due to intrinsic mobility and \nextrinsic d omain wall scattering perpendicular to c axis31. Magnetoresistance and Hall \nmeasurements are also very anisotropic (see Supplement) . \nThe Seebeck coefficient is about -10 WK-1m-1 at 300 K along both axes (c and a) \nwith an error of ±10%. The opposite sign s of the Hall effect (positive) and Seebeck \ncoefficient (negative) indicate the co -existence of both light holes and heavy electrons \nat the Fermi energy32. Detailed data are presented in the Supplement . The spin \npolarization P of the electrons at the Femi level was deduced from a point contact \nAndreev ref lection measurement at 2 K (see Supplement). The measured value of P is \n13 % and agree s with well with the calculated spin polarization of the density of states \nat the Fermi energy (see Supplement). The transport polarization will be different due 11 \n to different effective masses of minority and majority electrons33. \nThe measured thermal transport properties in Fig ures. 3b and 3c along c and a \naxes are also highly anisotropic. The total thermal conductivity along c is about twice \nas large as along a, and it is mainly explained by the carrier contribution , following \nthe Wiedemann –Franz law34. The remaining, almost isotropic, part is dominated by \nthe phonon cont ribution. The slight upturn of the thermal conductivity at temperature s \nabove 200 K may be due to uncertainty in the radiative heat losses , which is estimated \nto be about ±10%. In addition, magnon s or electron -magnon interactions might \ninfluence the thermal conductivity . The limiting Lorenz number generally depends a \nlittle on temperature34. The anisotropy r eflects the anisotropic electronic structure, \nwhich is the origin of the giant magnetocrystalline anisotropy. The c-axis thermal \nconductivity of 20 Wm-1K-1 at room temperature is roughly twice that of unsegregated \nL10 FePt (11-13 Wm-1K-1 35,36), or A1 FePt (9 Wm-1K-1 36). The anisotropic transport \nproperties, including magnetoresistance , anomalous Hall effect , and Seebeck effect , \nresult from the anisotropic electronic structure , that leads to an anisotropic mobility of \nthe charge carriers. \n \nFigure. 3. Transport properties of Rh 2CoSb. a, Longitudinal resistivity along the c (red curve) \nand the a axis (blue curve). b and c, Thermal conductivity along the c and a axes and the 12 \n charge carrier contribution (estimated from the Wiedemann –Franz law as LT/ρxx, where L = \n2.44 WΩK−2 is the Lorenz number) . The remaining part is mainly the phonon contribut ion. \n \nDiscussion \nLittle information is available about the magnetism of hard magnetic 3 d-4d \nintermetallic compounds, other than FePd, which has the tetragonal L 10 structure with \nK1 = 1.8 MJm-3, and YCo 5 which has the hexagonal CaCu 5 structure5 with K1 = 6.5 \nMJm-3. A number of ternary , tetragonal Rh -based intermetallics are known to order \nferro magnetically with a Curie point above room temperature15. Only one atom out of \nfour in the Rh2CoSb formula is cobalt, but the high Tc of 450 K indicates a strong \nexchange interaction . We see from our ab-initio calculations (Supplement al \ninformation ) that Rh is clearly in a spin -polarized state , and that it contributes to the \nferromagnetism. The measured magnetic moment per formula of 2.6 μB is much \nhigher than the ordinary ~1.6 μB moment of Co in the elemental state or in metallic \nalloys with other 3 d elements . The difference should be attributed , at least partially , to \nRh, or else to enhanced Co spin and orbital moment s, as there is no magnetic \ncontribution from Sb. \nWe also prepared polycrystalline Rh2FeSb with the same crystal structure, which \nexhibits easy-plane magnetization . At 300 K under 7 T, the incomplete ly saturated \nmoment reaches 3.8 μB per formula , far greater than the 2.2 μB of metallic iron. The \nCurie temperature of 510 K is even higher than that of Rh2CoSb , which is in contra st \nwith other isostructural Fe and Co intermetallics . A detailed comparison of the spin 13 \n and orbital contributions for Rh2FeSb and Rh 2CoSb deduced from ab-initio \ncalculation s is provided in the Supplement al Information . All the evidence illustrate s \nthat Rh plays an important role in the ferromagnetism of the se ternary compounds . In \nfact, it has already been shown to carry an induced moment in binary Fe -Rh and \nCo-Rh alloys that is roughly one third of the 3d moment37,38. \nTo reveal the site specific magnetic moments of Rh and Co , X-ray magnetic \ncircular dichroism (XMCD) measurements were performed at the L2,3 edges of Co and \nRh, respectively . The spin and orbital moments of each element39-41 are determined \nfrom XMCD using sum rule analysi s42,43. The same number of 7.8 electrons in the \nvalence d shell is assumed for both Co and Rh. This value was found in fully \nrelativistic ab-initio calculations. The measured XMCD spectra are shown in Figure 4. \nThe XMCD signal has the same sign in the spectra obtained at the Co and Rh L2,3 \nedges , which proves that the coupling between Co and Rh is ferromagnetic . \nThe XMCD signal at the L3 edge for Co is considerably larger than that at the \nL2 edge, which indicates a sizeable orbital contribution . Sum rule analysis reveals that \nspin and orbital moments for Co are 1.53 ± 0.15 µB and 0.42 ± 0.04 µB, respectively. \nThe orbital moment of Co in R h2CoSb far surpasses that of elemental Co, where the \nvalue is 0.15 μ B44. Therefore, the orbital moment of Co makes a sizeable contribution \nto the overall magnetization of 1. 95 ± 0.1 9 µB. The presence of a large orbital moment \nis also evident in Figure 2a from the 0.2 µ B difference of saturation magnetization \nbetween the c and a. \nThe spin and orbital moments of Rh obtained from the sum rule analysis are 14 \n 0.28± 0.03 µ B and 0.02 0± 0.002 µ B, respectively, resulting in a total moment of \n0.30± 0.03 µ B. These values account for the reduced photon polarization at high \nenergies. Using these moments , together with those of Co, the total magnetic moment \nof Rh 2CoSb should amount to 2.55± 0.22 µB, which is in good agreement with the \nmagnetic measurement of 2. 6 µB. \n \nFigure. 4. XAS and XMCD spectra of Rh 2CoSb. a, Co L2,3 and b, Rh L2,3 XAS and XMCD \nspectra. Spectra were taken at 25 K temperature and 6 T induction field. The photon helicity is \noriented parallel ( µ+) in black solid line or antiparallel ( µ−) in red solid line to the magnetic \nfield. The background is shown in the black dash line with edge jumps. \nThe Co atoms have a substantial orbital magnetic moment in Rh 2CoSb, both \nfrom experiments and ab-initio calculations. The orbital moment was 27% of the spin \nmoment, compared to 5% in elemental -Co (hcp) and 3% for Fe in calculations of \nthe sister compound Rh 2FeSb. The larger orbital moment for cobalt reflects the 15 \n hybridization of Co 3 d states with 4 d states of Rh, which has a stronger spin-orbit \ninteraction than Co. \nIn many Heusler compound s, the martensitic transition from a cubic to a \ntetragonal structure is explained by a band Jahn-Teller effect23 that results in a \nsplitting of energy levels by a modification of the ir width. We also calculate d \nRh2CoSb in the higher energy cubic L21 structure , but saw no sign of Jahn-Teller type \nsplitting in the electr onic structure . The formation enthalpies for our compound in the \ntetragonal I4/mmm (space group 139), inverse tetragonal I\nm2 (space group 119) and \ncubic L21 (space group 225) structures are -753, -325 and -221 meV , respectively. T he \nenergy difference between the cubic and tetragonal structure s is 532 meV, \nsignificantly more than in other Rh -based Heusler compound s. Thus, the hypothetical \ncubic to tetragonal phase transition temperature ( Tcub-tet) should be above the melting \npoint . Among the reported Rh2CoZ Heusler compounds , only Rh 2CoSb and Rh 2CoSn \nare tetragonal at room temperature ; the others are cubic. However, Rh 2CoSn has a \nmartensitic transition at around 600 K with a cubic structure at higher temperature23. \nBy avoid ing any such transition, Rh 2CoSb is superior to other Rh 2CoZ alloys for \nHAMR media . \nOur work will help to design new rare-earth free materials with strong uniaxial \nmagnetocrystalline anisotropy , for which we need : \n1) a low-symmetr y crystal structure (tetragonal or hexagonal ); \n2) heavy atoms with a large spin -orbit interaction (4d, 5d or 5f elements ); \n3) another possible element with the right electronegativity to help stabilize the 16 \n structure and help produce an advantageous electronic structure. \nThe magnetic moments of heavy non-rare-earth elements are small at best . \nTherefore, the design rule for rare-earth free metallic magnet s is to pair a 3d metal \n(Mn, Fe, Co) that can provide most of the magnetization with a heavy atom that \nsupplies a large spin -orbit interaction . Rh, Ir, Pd and Pt are all possible choice s, since \nthey are in the same group as Co or Ni and they may exhibit a significant induced \nmagnetization37,38,45. In fact, Pd is already ferromagnetic with Tc = 17 K if lightly \ndoped with 0.5% of Co46 and when it is alloyed with 50% Co , it has the highest Tc \namong the binary compounds CoRh (550 K), CoIr ( 40 K), CoPd (950 K) and CoPt \n(840 K)47,48. Our study outlines the design principles for new rare-earth free hard \nmagnets , but i t is unrealistic to imagine that Rh -based alloys will ever be used as bulk \nfunctional materials, in view of its high cost . This is much less important in functional \nthin films, because the quantities used are tiny . Strong perpendicular anisotropy is of \nincreasing importance in spintronics and especially in HAMR ; there thermal \nconductivity is also an important co nsideration. \nRef. 16 report ed the preparation of Rh2CoSb film s by sputtering or ion -beam \ndeposition between 373 -873 K . A large perpendicular anisotropy was observed. \nDisorder is common in sputter ed films . Our experiment s indicate that the Rh is well \nordered, but there might be some disorder between Co and Sb atoms. From ab-initio \ncalculation s, the most stable crystal structure for both ordered and disordered phase s \nis the tetragonal D0 22 structure and the lattice constant is very similar. T he total \nmagnetic moment in the fully Co -Sb disordered state is about 20% larger compared to 17 \n the completely ordered state. The orbital moment is nearly constant and independent \nof the degree of disorder , so the increase of the total moment is attributed to the spin \nmoment . \nWe compar e Rh2CoSb with L1 0 FePt5 for HAMR in Table 1. The magnetic \nproperties for Rh 2CoSb at 2 K are shown with bracket s, those at 300 K without. The \nmagnetic properties for L1 0 FePt do not vary much between 2 K and 300 K. The \nwriting speed is limited by the cooling time wh en the total heat produced by laser \ntransfer s to the substrate. Since Rh 2CoSb has a much lower Curie point and higher \nthermal conductivity, its writing speed will be roughly 6.8 times faster than L1 0 FePt \n(see Supplementary information) . \n \nTable 1. Comparison of Rh 2CoSb with L10 FePt for HAMR at room temperature . The properties \nfor Rh 2CoSba at 2 K is show n in bracket. \n \nConclusion \nIn summary, Rh2CoSb is a uniaxial ferro magnet with a remarkable \nmagnetocrystalline anisotropy of 3. 6 MJm-3, due to a large unquenched orbital \nmoment of 0.42 µB on Co that arises from hybridization with the surrounding Rh, \nwhere spin orbit coupling is strong . The magnetic hardness parameter of κ = 3. 7 at Materials Crystal \nstructure K1 \n(MJm-3) μ0Ms \n(T) μ0Ha (T) κ Tc \n(K) Tcub-tet Thermal \nstability Thermal \nconductivity \n(Wm-1K-1) Writing \nspeed \nRh2CoSb I4/mmm 2.2 \n(3.6) 0.44 \n(0.52) 12 \n(17) 3.7 \n(4.1) 450 > melting \npoint Stable 20 in c axis \n12 in a axis Fast \nL10 FePt P4/mmm 6.6 1.43 11 2 750 1573 K Transition to \nA1 FePt 11 Slow 18 \n room temperature is the highest observed so far in any rare-earth free magnet. The \nrelatively low Curie point of Tc = 450 K leads to a large temperature -dependen ce of \nK1 from a high base at room temperature, which is an asset for data writing . The \nanisotropic thermal conductivity, especial its large c axis value of 20 Wm-1K-1, which \nis much larger than that of the current FePt HAMR material , is important for cooling, \nwhich would lead to a 6.8 times faster writing speed than FePt. Unlike FePt with its \norder/disorder phase transition, Rh 2CoSb is a stable phase without any structural \ntransition below the melting point and its properties are stable in air. All these features \ncommend Rh2CoSb as a candidate for HAMR media with a recording density of more \nthan 10 Tb/in2 and high writing speed . \n \nMethods \nSingle crystal growth. The single crystals of Rh 2CoSb were grown by the Bridgeman \nmethod. First, the high purity (> 99.99%) elements Rh, Co and Sb were cut into small \npieces and arc-melted together to prepare polycrystalline samples. The initial atomic \nratio of Rh, Co and Sb was 2:1:1.03. Additional Sb was added due to compensate for \nits high vapor pressure. From powder XRD data, there were always a few percent of a \nRhSb secondary phase in the polycryst alline samples. Hence the polycrystalline \nmaterials were heated up to 1600 K (above the melting point of RhSb at 1583 K) for 3 \ndays to achieve a homogeneous liquid before beginning the single crystal growth, \nwhich took about 5 days during cool down to 1273 K. The composition of the crystals \nwas checked by wavelength dispersive X-ray spectroscopy that show ed a 19 \n homogenous composition of Rh 50.3Co25.6Sb24.1. The crystals were characterized by \npowder X -ray diffraction as single -phase with a tetragonal structure. The orientation \nof the single crystals was confirmed by the Laue method . \nMagnetization measurements. These were conducted on single crystals with the \nmagnetic field applied along either the a or c axis using a vibrating sample \nmagneto meter (MPMS 3, Quantum Design). The sample size was 3.80\n 0.78\n 1.30 \nmm3. For measurement with the field along the a axis, the sample was carefully \nimmobilized with glue in view of the stron g torque . High -field m agnetization \nmeasurements were performed in the Dresden High Magnetic Field Laboratory using \na pulsed magnet . \nMagnet o-optical Kerr microscopy. Domain images were performed by using the \npolar Kerr effect in a wide -field magneto -optical Kerr microscope at room \ntemperature with a polished 1 mm thick single crystal with (001) surface . Detailed \ndescription of the method can be found elsewhere49. \nElectrical transport measurements. The longitudinal and Hall resistivities were \nmeasured on a Quantum Design PPMS 9 using the low -frequency alternati ng current \n(ACT) option for data below 320 K . The longitudinal resistivity was measured with \nstandard four -probe method, while for the Hall resistivity measurements, the \nfive-probe method was used with a balance protection meter to eliminate possible \nmagnetoresistance signals50. Longitudinal resistivity measurement above 320 K, were \nmeasured on a home -made device using a four-probe method and careful calibration. \nThe accuracy of resistivity measurement is ± 5%. 20 \n Thermal transport measurements. The thermal conductivity and Seebeck \nthermopower were measured adiabatically in the Quantum Design PPMS using \nthermal transport option (TTO). The uncertainty of the radiative heat losses at high \ntemperature and the uncertainty of the geometry is estimated as ± 10%. \nAndreev reflection measurements. Measurements are performed on a polished (001) \ncrystal surface in a flow of helium vapor, using a mechanically sharpened Nb tip in \nthe absence of an external magnetic field. Data are analy zed using the modified BTK \nmodel, as detailed elsewhere51. The best fit to spectrum is obtained with barrier \nparameter Z~ of 0.35, an electron temperature of 2 K and a spin polarization of 13%. \nXMCD measurements. X-ray magnetic circular dichroism (XMCD) was measured at \nthe beamline BL29 (BOREAS) of the synchrotron ALBA in Barcelona (Spain). \nXMCD spectra at the L2,3 absorption edges of Co and Rh were taken at a temperature \nof 25 K in a vacuum chamber with a pressure of 10−9 mbar. The x -ray absorption \nspectra (XAS) were measured using circular polarized light with photon helicity \nparallel ( µ+) or antiparallel ( µ−) to the fixed magnetic field in the sequence \nµ+µ−µ−µ+µ−µ+µ+µ− to disentangle the XMCD. An induction field of 6 T was applied \nalong the c axis. The polarization delivered by the Apple II -type elliptical undulator \nwas close to 100% for the Co L2,3 edges and 70% for the Rh L2,3 edges. The spectra \nwere recorded using the total yield mode. \n \nSupporting Information \nSupporting Information is available from the Wiley Online Library or from the author. 21 \n \nAcknowledgement \nThis work was financially supported by the European Research Council Advanced \nGrant (No. 742068) “TOPMAT”, the European Union’s Horizon 2020 research and \ninnovation programme (No. 824123) “SKYTOP”, the European Union’s Horizon \n2020 research and innovation programme (No. 766566) “ASPIN”, the Deutsche \nForschungsgemeinschaft (Project -ID 258499086) “SFB 1143”, the Deutsche \nForschungsgemeinschaft (Project -ID FE 633/30 -1) “SPP Skyrmions”, the DFG \nthrough the Wü rzburg -Dresden Cluster of Excellence on Complexity an d Topology in \nQuantum Matter ct.qmat (EXC 2147, Project -ID 39085490) and the DFG through \nSFB 1143. We acknowledge the support of the High Magnetic Field Laboratory \nDresden (HLD) at HZDR, members of the European Magnetic Field Laboratory \n(EMFL). \n \nConflict of Interest \nThe authors declare no conflict of interest. \n \nAuthor contributions \nSingle crystals were grown by Y .H. and K.M. The characterization of the crystal, \nmagnetic and transport measurement were performed by Y .H. with the help of C.Fu, \nY .P., J.K., A.J., Y .S.,W.S., P.S., H.B., R.S. and S.S.P.P. XMCD was measured by X.W., \nZ.H., S.A., J.H., M.V . and L.H.T. First principle calculations were carried out by 22 \n G.H.F. All the authors discussed the results. The paper was written by Y .H., J.M.D.C. \nand G.H.F. with fee dback from all the authors. The project was supervised by C. \nFelser. \n \nReferences\n \n1. C. Chappert , A. Fert, and F. N. V. Dau, Nat. Mater. 2007 , 6, 813 –823. \n2. M.T. Kief and R.H. Victora , MRS Bulletin 2018 , 43, 87-92. \n3. D. Weller, G. Parker, O. Mosendz, A. Lyberatos, D. Mitin, N. Y . Safonova, & M. \nAlbrecht, J. Vac. Sci. & Technol. B 2016 , 34, 060801 . \n4. R. Skomski, & J. M. D. Coey, Scr. Mater. 2016 , 112, 3–8. \n5. J. M. D. Coey, Magnetism and Magnetic Materials , Cambridge Univ. Press, \nCambridge, 2010 , Ch. 4 and 8. \n6. S.I. Iwasaki, & K. Ouchi, IEEE T. Magn . 1978 , 14, 849 -851. \n7. L. Pan, and D. B. Bogy , Nat. Photonics 2009 , 3, 189 –190. \n8. J. M. Hu, L .Q. Chen, and C .W. Nan, Adv. Mater. 2016 , 28, 15–39. \n9. W. A. Challener , C. Peng , A. V . Itagi, D. Karns , W. Peng , Y. Peng , X. Yang , X. Zhu, \nN. J. Gokemeijer, Y . -T. Hsia, G. Ju, R. E. Rottmayer , M. A. Seigler , and E. C. Gage , \nNat. Photonics 2009 , 3, 220-224. \n10. A. V. Kimel , and M . Li, Nat. Rev. Mater. 2019 , 4, 189-200. \n11. D. Weller, O. Mosendz, G. Parker, S. Pisana, and T.S. Santos, Phys.Status Solidi A \n2013 , 210, 1245. \n12. C.‐B. Rong, D. Li, V . Nandwana, N. Poudyal, Y . Ding, Z.L. Wang , H. Zeng, J.P. \nLiu, Adv. Mater. 2006 , 18, 2984. \n13. A. K. Nayak, A. K. Nayak, M . Nicklas, S . Chadov, P . Khuntia, C . Shekhar, A . \nKalache, M . Baenitz, Y . Skourski, V . K. Guduru, A . Puri, U . Zeitler, J. M. D. Coey & 23 \n \nC. Felser Nat. Mater. 2015 , 14, 679–684. \n14. H. Kurt, K. Rode, P. Stamenov, M. Venkatesan, Y. -C. Lau, E. Fonda, and J. M. D. \nCoey , Phys. Rev. Lett. 2014 , 112, 027201 . \n15. S. K. Dhar, A. K. Grover, S. K. Malik, R. Vijayaraghavan, Solid State Commun. \n1980 , 33, 545. \n16. S. V . Faleev, Y . Ferrante, J . Jeong, M . G. Samant, B . Jones, and S . S. P. Parkin , \nPhys. Rev. Mater. 2017 , 1, 024402. \n17. T. Gasi, A . K. Nayak, J . Winterlik , V. Ksenofontov, P . Adler , M. Nicklas, and C . \nFelser , Appl. Phys. Lett. 2013 , 102, 202402. \n18. P. J. Brown, T. Kanomata, K . Neumann , K. U. Neumann , B. Ouladdiaf , A. Sheikh \nand K . R. A. Ziebeck , J. Phys.: Condens. Matter. 2010 , 22, 506001. \n19. B. Balke, G. H. Fecher, J. Winterlik and C. Felser , Appl. Phys. Lett. 2007 , 90, \n152504 . \n20. A. K. Nayak, C. Shekhar, J. Winterlik, A. Gupta, & C. Felser, Appl. Phys. Lett. \n2012 , 100, 152404. \n21. A. K. Nayak , M. Nicklas, S. Chadov, C. Shekhar, Y . Skourski, J. Winterlik, and C. \nFelser , Phys. Rev. Lett. 2013 , 110, 127204. \n22. O. Meshcheriakova, S. Chadov, A.K. Nayak, U.K. Rößler, J. Kübler, G. André, \nA.A. Tsirlin, J. Kiss, S. Hausdorf, A. Kalache, W. Schnelle , M. Nicklas, and C. Felser , \nPhys. Rev. Lett. 2014 , 113, 087203. \n23. J. C. Suits, Solid State Commun. 1976 , 18, 423. \n24. A. Hubert, R. Schä fer, Magnetic Domains, Springer, Berlin, 1998 , p.p. 314, 379. \n25. Y . Otani, J. M. D. Coey , J. Appl. Phys. 1990 , 67, 4619. \n26. T. Nishioka, G. Motoyama, & N. K. Sato, J. Magn. Magn. Mater. 2004 , 272, \nE151 . 24 \n \n27. S. Okamoto, N. Kikuchi, O. Kitakami, T. Miyazaki, Y . Shimada, & K. Fukamichi, \nPhys. Rev. B 2002 , 66, 024413. \n28. M. Balanda, Acta Phys. Pol. A 2013 , 124, 964 -976. \n29. J. Cui, M. Kramer, L. Zhou , F. Liu, A. Gabay , G. Hadjipanayis , B. \nBalasubramanian , D. Sellmyer , Acta Mater. 2018 , 158, 118 -137. \n30. C. Hass, Phys. Rev. 1968 , 168, 531. \n31. D. Elefant, R. Schä fer, Phys. Rev. B 2010 , 82, 134438. \n32. J. Dijkstra, H. H. Weitering, C. F. Van Bruggen, C. Haas, R. A. De Groot, J. Phys.: \nCondens. Matter 1989 , l, 9141 -9161. \n33. G. H. Fecher, S . Chadov and C . Felser , Theory of the Half -Metallic Heusler \nCompounds Spintronics, Springer, Dordrecht , 2013, 115-165. \n34. J. M. Ziman, Electrons and Phonons, Oxford University Press , 1960 . \n35. A. Chernyshov, D. Treves, T. Le, F. Zong, A. Ajan , R. Acharya, J. Appl. Phys. \n2014 , 115, 17B735. \n36. A. Giri, S. H. Wee, S. Jain, O. Hellwig, P. E. Hopkins, Sci. Rep. 2016 , 6, 32077. \n37. G. R. Harp, S. S. P. Parkin, W. L. O' Brien , B. P. Tonner, Phys. Rev. B 1995 , 51, \n12037 . \n38. G. R. Harp , S. S. P. Parkin , W. L. O' Brien , B. P. Tonner, J. Appl. Phys. 1994 , 76, \n6471. \n39. S. Agrestini , Z. Hu, C. -Y. Kuo, M.W. Haverkort, K. -T. Ko, N. Hollmann, Q. Liu, \nE. Pellegrin , M. Valvidares, J. Herrero -Martin, P. Gargiani , P. Gegenwart, M. \nSchneider , S. Esser , A. Tanaka , A. C. Komarek, L. H. Tjeng, Phys. Rev. B 2015 , 91, \n075127 . \n40. S. Das, A. D. Rata, I. V . Maznichenko, S. Agrestini, E. Pippel, N. Gauquelin, J. 25 \n \nVerbeeck, K. Chen, S. M. Valvidares, H. Babu Vasili, J. Herrero -Martin, E. Pellegrin, \nK. Nenkov, A. Herklotz, A. Ernst, I. Mertig, Z. Hu, and K. Dö rr, Phys. Rev. B 2019 , \n99, 024416. \n41. N. Hollmann, S. Agrestini , Z. Hu, Z. He, M. Schmidt, C.-Y. Kuo, M. Rotter, A. A. \nNugroho, V . Sessi, A. Tanaka, N. B. Brookes, and L. H. Tjeng, Phys. Rev. B 2014 , 89, \n201101(R). \n42. B. T. Thole, P. Carra, F. Sette, G. van der Laan , Phys. Rev. Lett. 1992 , 68, 1943 . \n43. P.B. Carra, T. Thole , M. Altarelli, X. Wang, Phys. Rev. Lett. 1993 , 70, 694. \n44. C. T. Chen , Y .U. Idzerda , H. -J. Lin, N.V . Smith , G. Meigs , E. Chaban , G.H. Ho, E. \nPellegrin , F. Sette , Phys. Rev. Lett. 1995 , 75, 152. \n45. V . V . Krishnamurthy, D. J. Singh , N. Kawamura, M. Suzuki, T. Ishikawa , Phys. \nRev. B 2006 , 74, 064411. \n46. A. Pilipowicz, H. Claus, Phys. Rev. B 1987 , 36, 773. \n47. H. Masumoto , K. Watanabe, K. Inagawa, J. Jpn. I. Met. 1976 , 17, 592. \n48. B. Predel, H. Landolt, R. Bö rnstein, Phase equilibria, crystallographic and \nthermodynamic data of binary alloys (V ol. 5). Berlin, Heidelberg, Springer , 1991. \n49. I.V . Soldatov, R. Schä fer , Rev. Sci. Instrum . 2017 , 88, 073701. \n50. S.D. Pinzon, Magnetic properties and interface scattering contribution to the \nanomalous Hall effect in cobalt/palladium multilayers, California State University, \nLong Beach, 2016 , p16. \n51. P. Stamenov, J. Appl. Phys. 2013 , 113, 17C718. \n 1 \n New highly -anisotropic Rh -based Heusler compound for \nmagnetic recording \n \nSupplementary Information \n \nYangkun He1,*, Gerhard H. Fecher1,*, Chenguang Fu1, Yu Pan1, Kaustuv Manna1, Johannes \nKroder1, Ajay Jha2, Xiao Wang1, Zhiwei Hu1, Stefano Agrestini3, Javie r Herrero -Martí n 4, \nManuel Valvidares4, Yurii Skourski5, Walter Schnelle1, Plamen Stamenov2, Horst Borrmann1, \nLiu Hao Tjeng1, Rudolf Schaefer6,7, S. S. P. Parkin8, J. M. D. Coey2 and Claudia Felser1 \n \n1Max-Planck -Institute for Chemical Physics of Solids, D -01187 Dresden, Germany \n2School of Physics, Trinity College, Dublin 2, Ireland \n3 Diamond Light Source, Harwell Science and Innovation Campus, Didcot, OX11 0DE, \nUK \n4ALBA Synchrotron Light Source, Cerdanyola del Valles, 08290 Barcelona, Catalonia, \nSpain \n5 Dresden High Magnetic Field Laboratory (HLD -EMFL), Helmholtz -zentrum \nDresden–Rossendorf, 01328 Dresden, Germany \n6Leibniz Institute for Solid State and Materials Research (IFW) Dresden, Helmholtz \nstrasse 20, D -01069 Dresden, Germany \n7Institute for Materials Sc ience, TU Dresden, D -01062 Dresden, Germany \n8Max Planck Institute of Microstructure Physics, Halle, Germany \n \nKey words: magnetocrystalline anisotropy, tetragonal Heusler alloy , magnetic \nhardness parameter, 4 d magnetism, magnetic recording 2 \n \n1. Single crystal growth and c omposition \nThe composition of Rh 2CoSb single crystal has been identified by both \nenergy -dispersive X -ray spectroscopy (EDX) and wave length dispersive X -ray \nspectroscopy (WDX ), which shows a homogenous composition . The average \nvalues and standa rd deviation are shown in the Table S1. The sample has a \nhomogenous composition of Rh 50.3Co25.6Sb24.1. \n \nFig. S 1. EDX curve of Rh 2CoSb single crystal. The insert is the image of the single \ncrystal. \n \nTable S1. WDX data of Rh 2CoSb single crystal at ten diffe rent points. \nPoint Rh(at.%) Co(at.%) Sb(at.%) \n1 50.7 25.2 24.1 \n2 50.7 25.4 24.0 \n3 50.3 25.7 24.0 \n4 49.9 26.1 24.1 \n5 50.0 25.9 24.1 \n6 50.4 25.6 24.0 \n7 50.4 25.6 24.1 \n8 50.7 25.4 24.0 \n9 50.3 25.5 24.2 \n10 50.1 25.8 24.0 \nAverage 50.3±0.29 25.6±0.27 24.1±0.07 \n \n \n2. Orientation \nThe quality and orientation of the single crystal were confirmed by Laue method, 3 \n which shows clear spots in both c and a axes, fitting well with the simulation . \n \nFig. S 2. Laue pattern of Rh 2CoSb single crystal from c (a) and a (b) axes. \n \n3. Crystal structure \nThe tetragonal structure was identified by both powder X -ray diffraction and \nhigh-resolution transmission electron microscopy (TEM). The data show that \nRh2CoSb has a well -ordered tetragonal D0 22 structure with a = 4.0393 (6) Å a nd c \n= 7.1052 (7) Å. Here, the 4 d (0 0 ¼) site is occupied by Rh, 2 b (0 0 ½) by Co, and \n2a (0 0 0) by Sb. The tetragonal distortion is c/\na = 1.2436(3). \n20 30 40 50 60 70 80 90 100(103)\n(211)(004)(200)(112)Intensity (a.u.) Iobs\n Ical\n Iobs - Ical\n Bragg position\n2 (degree)(101)\n \nFig. S 3. Powder X -ray diffraction of Rh 2CoSb at room temperature . \n 4 \n \n \nFig. S 4. High -resolution TEM image of the [110] plane, showing a tetragonal lattice \nwith c/√2a = 1.24. The insert is the selected area electron diffraction image, whose \nsuperlattice reflection indicates a well-ordered structure. \n \n4. Thermal analysis \nDifferential scanning calorimeter (DSC) and thermogravimetric (TG) analysi s \n(see Fig. S 5) have been performed to investigate the phase transitions at high \ntemperature. Unlike many Mn -based Heusler compounds which are often cubic at \nhigh temperature and may experience a martensitic transition to become tetragonal, \nRh2CoSb does no t have first order transition at high temperature and is a single \nphase at all temperature range up to the melting point of about 1482 K, according \nto DSC and TG (see Fig. S 5). Therefore, the common problem of twinning in \nmany permanent magnets such as MnA l, MnGa, FePt and CoPt does not exist and \nsingle crystals have been successfully grown the by Bridgeman method. 5 \n \n200 400 600 800 1000 1200 1400 1600 1800\nTemperature (K)DSC (a.u.)\nT = 1482 K\n-1.0-0.8-0.6-0.4-0.20.00.20.40.60.81.0\n TG (%) \nFig. S 5. DSC and TG curve of Rh 2CoSb cooling from the melt. The TG curve (blue) is \nflat, indicating little Sb loss d uring heating. The peak in the DSC curve (black) is referred \nas the freezing point of 1482 K. No other peaks are detected, excluding any other first \norder transition. The measurement rate of thermal analysis was set as 10 K/min. \n \n5. Magnetization measurement \nThe single crystal magnetization curve along the c axis is shown in Fig. S6 (a), \nwhose enlarged part at low field is shown in (b). During the demagnetization, the \nsudden drop of the magnetization, namely Barkhausen jumps, together with a high \nsusceptibility after thermal demagnetization, sugges t single crystal Rh2CoSb as a \nnucleation type permanent magnet. The magnetization curve along the a axis in (c) \nis not saturated at 7 T, indicating strong magnetocrystalline anisotropy. μ0Ms2 /K1 \nat different temperatures is shown in (d). The peak near Tc is responsible for the \nHopkinson effect in Figure 2d. A small piece of single crystal was ground into \npowder and then annealed at 800 K for 5 h to remove the inner strain before \nbonded i n epoxy at 400 K to make a n isotropic bonded magnet. Coercivity, as an \nextrinsic property, is determined mainly by additional structural features such as \nlattice defects, grain boundaries, sample or particle size, and surface irregularities. \nThe coerciviti es of bonded isotropic magnet s with grain size s of ~2 0 μm and 1 -3 \nμm show a significant increase from 0.2 T to 0.9 T at room temperature. The B-H \ncurve is shown in ( f) for the isotropic magnet (1 -3 μm grain size). The energy \nproduct ( BH)max is about 10 kJm-3, which is only a quarter of the theoretical value. \nTo find a secondary phase and good alignment remain to be done for further study . \nFor polycrystalline samples produced by arc -melting shown in (g), the coercivity \nis about 0.2 T. 6 \n \n \nFig. S 6. Magnetization curves of Rh 2CoSb. Single crystal data from the c axis at different \ntemperatures are shown in (a), whose enlarged part at low field is shown in (b). Single crystal \ndata from the a axis at different temperatures are shown in (c). (d) μ0Ms2 /K1 at different 7 \n temperatures. M-H curve (e) of isotropic b onded magnet s with a grain size of about 1 -3 μm \n(red) and ~20 μm (blue) at 300 K . The insert shows their distribution histogram (unit: μm) . \nThe B-H curve of the magnet with grain size of 1 -3 μm at 300 K is shown in (f). \nMagnetization curve of polycrystalline sample is shown in (g). \n \n6. Electrical Transport measurement \nTransport properties including the resistivity and Hall effect of Rh 2CoSb are \nshown in Fig. S 7 and S8 . The resistivity is very ani sotropic. With increasing \ntemperature, the a-axis resistivity increases monotonically from 53 µΩ cm at 2 K to \n192 µΩ cm at Tc, after which the resistivity is then dominated by spin disorder \nscattering and increas es slowly1. The resistivity above Tc, which is close to the \nminimum metallic conductivity (~200 µΩ cm)-1, indicates that the mean free path is \nclose to the interatomic spacing when the moments are disordered. The c axis \nresistivity is half as large, 21 µΩ cm at 2 K but it increases with temperature and \nshows a similar trend to the a axis resistivity. Domain walls along the c axis in most \nof the volume, but not in the surface region as shown in the insert of Fig. 2a and \nSupplementary Fig. S 11, influence the resistivity. When the charge carriers travel \nalong c the current is parallel to the moment direction; but when the carriers move in \nthe basal plane they tend to follow helical paths due to the Lorenz force, leading to a \nlarger resistivity2. Like other ferromagnets, Rh 2CoSb shows a negative transvers e \nmagnetoresistance (MR), which increases with increasing temperature, reaching \n-1.75 % at 300 K under a 7 T field. However, when the field is applied along a (taking \ncare to immobilize the crystal) and the current is passed in a perpendicular b direction, \nthe MR is positive (0.39%) at 300 K during the hard axis magnetization process, as \nshown in Fig. S7c. \nThe results of the Hall measurements are shown in Fig. S7d. When the field is \nparallel to the c axis and the current is along a, the normal Hall effect d ue to the \nLorenz force is positive and increases linearly with applied field, indicating mainly \nhole-type charge carriers with a density of 1.5\n 1022 cm-3 corresponding to 0.86 per \nformula. The mobility is 2 -7 cm2V-1s-1 between 2 -300 K. Detailed data can be seen in \nSupplementary Fig. S 8. The Hall effect was also measured with the field along the a \naxis and the current along b at 300 K. The magnetization at a field of 9 T is far from \nsaturation, hence the data are a mixture of a normal and an anom alous Hall effect. \nTheir values could be estimated by linear extrapolation of the curve to 12.4 T, the \nsaturation field from Fig. 2a, and extrapolation from the normal Hall effect. The \nsignificantly larger anomalous Hall resistivity of 1.33 µΩ cm measured along the a \naxis than the value of 0.75 µΩ cm measured along the c axis is attributed to the \nanisotropic crystal and electrical structure . 8 \n \n \nFig. S 7. Transport properties of Rh 2CoSb. a, Longitudinal resistivity along the c (red \ncurve) and the a axis (blue curve). b, Magnetoresistance along a with field along c. c, \nAnisotropic magnetoresistance. d, Anomalous Hall effect with field along a. \n \nThe electrical transport measurement was measured both along c and a axes. The \nHall conductivity was calculated by: \nσxy = -ρxy / (ρ xy2 +ρxx2) \nwhere ρ xx and ρ xy are the longitudinal resistivity (along a) at zero field and the Hall \nresistivity (along a), respectively3. Since ρ xx >> ρ xy and the anomalous Hall resistivity \nρxyA is proportional to the square of ρxx as shown in the insert of Fig. S8b, the \nanomalous Hall conductivity is almost unchanged with variation of the temperature, \nindicating a side-jump effect or an intrinsic mechanism due to the Berry curvature \nmechanism4 as there are a lot of band crossin gs in the ferromagnetic state ( see \nSupplementary Fig. S13). The σ xx is of order 104 Ω-1 cm-1, which is also in the region \nof intrinsic AHE3. \n The anomalous Hall resistivity versus temperature is shown in Fig. S8c, while \nthe charge carrier density as well as the mobility deduced from the normal Hall effect \nis shown in Fig. S 8d. The charge carrier density is deduced by n = 1/(e RH), where R H \nis the slope of the ρxy. The mobility is calculated by μ = RH / ρxx. 9 \n \n \nFig. S8. (a) Hall effect with field along c at different temperatures. (b) Hall conductivity \nat different temperatures. The insert shows the linear fitting of anomalous Hall \nresistivity ρAxy versus ρxx2. (c) The anomalous Hall resistivity versus temperature. (d) The \ncharge carrier density as well as the mobility at different temperatures. \n \nAlthough some quasi -two-dimensional metals5,6 and superconductors7-9 show a \ngiant resistivity anisotropy (the ratio of the resistivities measured along c and a can be \nmore than 10 to 100), t he anisotropic resistivity in Rh 2CoSb is among the largest \nfound in three -dimensional bulk materials. For the tetragonal Heusler compound \nMn 1.4PtSn, the resistivity along c is only about 7% larger than along a10. \nOrthorhombic UFe 2Al10 shows a less than 10% difference between a, b and c11. \nHexagonal Mn 3Ge exhibits a 2.5 times larger resistivity along a at 2 K, but the \ndifference decreases with increasing temperature, and vanishes around room \ntemperature12. The anisotropic transport propertie s, including magnetoresistance and \nanomalous Hall effect, indicate that the magnetism leads to an anisotropic mobility of \nthe charge carriers. \n \n7. Seebeck coefficient measurement \nThe Seebeck coefficient was measured both along c and a axes. From Figure S9 it is \nseen that the absolute value of the Seebeck coefficient is slightly higher along the \nc-axis for T > 10 K . The behavior of the anisotropic Seebeck coefficients is reciprocal 10 \n to the conductivity as expected from the Onsager relation . This reflects the ani sotropy \nof the Onsager coefficients that are tensors. In tetragonal materials one has Sxx = Syy \nSzz. It should further be noted that the elements of the Onsager tensors depend on \nthe magnetization of the sample or the applied magnetic field, in general13,14. \n0 50 100 150 200 250 300 350-12-10-8-6-4-202Seebeck coefficient (VK-1)\nTemperature (K) c axis\n a axis\n \nFig. S 9. Seebeck coefficient measurement along a and c axes a t different temperatures. \n \n8. Spin polarization \nThe spin polarization P of the electrons at the Femi level was deduced from point \ncontact Andreev ref lection measurement at 2 K that is shown in Fig S10. Data are \nanalysed using the modified BTK model, as described in detail elsewhere15. The best \nfit to the spectrum is obtained with a barrier parameter Z~ of 0.35, an electron \ntemperature of 2 K and a spin polarization of 13%. 11 \n \n-6 -4 -2 0 2 4 61.01.11.21.31.41.5G/Gn Data\n Fit\nD1* = 1.24(1) meV\nD2 = 1.50 meV\nZ* = 0.35(1)\nP* = 0.13(3)\nTe* = 3.3(3) K\nGn = 786(1) G0\nT = 2.3 K\nUa (meV) \nFig. S1 0. The spin polarization P of the electrons at the Fe mi level was deduced from \npoint contact Andreev refection measurement at 2 K. The fit parameters for the BTK \nmodel are included in the figure (see also Ref.15) \n \n9. Magneto -optical Kerr microscopy \nMagneto -optical K err microscopy study shows the surface domains of a \ntwo-phase branched domain pattern of higher generation on the (001) surface, which \nis usually observed in uniaxial magnets. For such a pattern, the domain walls are \naligned parallel to the c-axis only in the volume of the bulk crystal. When \napproaching the surface, the magnetization strictly follows the c-axis, but the domain \nwalls does not follow, as shown in Fig. S1 1. Further explanation can be found in Ref. \n16. \n \nFig. S1 1. The surface domain pattern of a strong uniaxial magnet adapted from Ref. 16. \n \n10. Electron structure calculations 12 \n 1) Ab-initio calculations. \nThe electronic and magnetic structures of Rh 2CoSb were calculated by means \nof the first principles computer programs Wien2k17-20 and SPKKR21,22 in the \nlocal spin density approximation. In particular, the generalized gradient \napproximation (GGA) of Perdew, Burke and Ernzerhof23 was used for the \nparametrization of the exchange correlation functional. A k -mesh based on 126 × \n126 × 126 points of the full Brillouin zone was used for integration when \ncalculating the total energies for determination of the magnetocrystalline \nanisotropy. For more details see [arxiv .] \n \n2) Results. \nThe electronic and magnetic structure calculat ions confirm the uniaxial \nmagnetocrystalline anisotropy with Ku 1.4 MJ/m³ . A total moment of 2.19 μB is \nfound from the GGA calculations using SPRKKR, the spin and orbital moments \nhave values of 2.04 μ B and 0.15 μB, respectively. \nFrom the spin and site r esolved data , it is found that Co contributes per atom a \nmagnetic moment of 1.81 μB of which about 0. 14 μB is the orbital moment, and \nthe major part of 1.67 μB is the spin moment. A small spin moment of about 0.2 μB \nand negligible orbit moment is carried b y Rh. \nBoth, anisotropy constant and magnetic moments are smaller than the \nexperimental values. Indeed, Ku does not include any temperature or macroscopic \nmagnetic effects e.g.: domains and domain walls. The orbital magnetic moment of \nCo is larger in the XM CD measurements. However, these measurements observe \nan excited state. Using the orbital corrected potential of Brooks24, the value is \nincreased to about 0.4 μ B without changing the spin moment. \nFigures S 12 and S 13 show results of the ab -initio calculations using SPRKKR . \nFigure S 12 shows the spin and site resolved density of states and Figure S 13 the \nspin resolved band stru cture. 13 \n \n \nFig. S12. The valence band density of states. \n \nFig. S13. Semi -relativistic band structures of Rh 2CoSb. Majority states are \ndrawn in red and minority states in blue. \n \n3) Comparison of Rh 2FeSb and Rh 2CoSb \nTable S2 shows the calculated occupation of th e 3d orbitals of the Co and Fe \natoms (number of electrons in a sphere within muffin tin radius) in Rh2FeSb and \nRh2CoSb. Part of the d-electron density is delocalized and found in between the atoms 14 \n and part is more localized on Rh. At both atoms the five ma jority -spin d-states are \nalmost fully occupied (about 4.5 out of 5) resulting in a nearly spherical distribution. \nThe occupation of the minority -spin states is quite different. In particular, the \noccupation of the minority dz2 orbital of Co modifies the sp in of the charge density of \nCo compared to Fe. The different occupancy of the minority orbitals is responsible for \nthe difference in anisotropy —easy plane for Fe and easy axis for Co. The calculations \nwere performed with Wien2k . \n \nTable S2. Occupation of th e orbitals at the Fe and Co atoms in Rh 2FeSb and \nRh 2CoSb. \nPoint Rh2FeSb Rh2CoSb \nMajority Minority Majority Minority \ndz2 0.946 0.106 0.934 0.862 \ndxy 0.915 0.538 0.909 0.823 \ndx2-y2 0.917 0.271 0.912 0.426 \ndxz 0.910 0.377 0.899 0.385 \ndyz 0.910 0.377 0.899 0.385 \ntotal 4.594 1.668 4.553 2.881 \nmoment 2.926 1.672 \n \nThe Co atoms have a rather large orbital magnetic moment in Rh 2CoSb, from \nboth experiments and ab-initio calculations. The calculated value for Co in Rh 2CoSb \nwas found to be about 8% of the spi n moment, whereas it is 4.8% in elemental -Co \n(hcp) and 3% for Fe in the sister compound Rh 2FeSb. The larger orbital moments for \nCo reflect hybridization of Co 3 d bands with 4 d bands of Rh, which has a strong \nspin-orbit interaction. \n \n11. Grain size and data s torage density \nThe superparamagnetic blocking radius25 for a particle is calculated as Rb = \n(6kBT/K1)1/3, where kB and T are Boltzmann constant and temperature respectively. \nFor Rh 2CoSb, it is Rb = 2.3 nm at 300 K. For modern perpendicular recording in \nCo-Cr-Pt, the magnetic media are not particles, but tall slim grains. The aspect ratio is \nabout 3, in order to increase the shape anisotropy. For h ard magnets like FePt and \nRh2CoSb one does not need to further increase the anisotropy by large r aspect ratio s. \nHowever, it is difficult to control a uniform grain height in short grains in order to get \na stable Curie temperature . As a result, the idea l aspect ratio is about 1.5 to 226. To \nresist demagnetization by random thermal fluctuation, the volume V of the magne tic \nmaterial must satisfy the empirical condition that K1V/kBT > 60. Therefore, V must be \nlarger than 113 nm3 for each grain and the diameter is calculated as 4.2 nm for \nRh2CoSb at 300 K for an aspect ratio of 2. In-plane heat transfer is also another thin g \nwe must pay attention to. To avoid the heating by a neighbour ing grain during writing \nas well as to increase the signal -to-noise ratio , it is better to slightly increase the \ncentre -to-centre distance of grains. This increased distance for Rh 2CoSb can still be 15 \n smaller than for FePt, due to the much lower Curie temperature (450 K vs 750 K) that \nallows to use lower temperatures . A potential storage density of more than 10 \nTerabit/inch2 can be realized27 when assuming that the centre -to-centre distance of \ngrains is 5 nm with half area occupied . \n \n12. Writing speed estimation \nRh2CoSb has a mass density of ρ = 11\n103 kg m-3 and its formula weight is \n386.5 g mol-1. We calculate the heat capacity to be Cp = 4\n 3R=100 J mol-1K-1 or a \nvolume heat c apacity of Cv = 2.85\n 106 J m-3K-1. The time for each grain to cool down \nis calculated to be τV = Cv∙A/λ. That is τV = 3.56\n 10-12 s, when u sing a thermal \nconductivity of λ = 20 Wm-1K-1 and assuming an area for each grain (5 nm in \ndiameter) of A = 2.5\n 10-17m2, \nThe a ngular velocity of a hard disk is ω = 7000 r min-1 and t he radi us for a \none-inch disk is r = 1 inch = 0.0254 m. Therefore , one has a scanning velocity of v = \nω∙2πr = 18.56 m s-1. The time of heating for each gra in is τh = \n /v = 2.7\n10-10 s and \nthe number of heated grains becomes ngrains= τh/τv = 75.8. \nTo complete a magnetization process, \n \nwhere Hc is the coercivity, T is temperature, t is time, x is the position that satisfies \ndx/dt = V and Δx =\n , fFMR is the ferromagnetic resonance frequency (for large angle \nprecession) at the temperature point of the maximal thermal gradient. It is a measure \nof how fast the switching proceeds within a grain in the trailing edge . No critical \ndifferences in fFMR are expected between the materials. The field of the write head Hw \nshould satisfy \n . The writing speed is proportional to \n . \nSimilarly one has for FePt Cv = 3.1\n 106 J m-3K-1 and gets τV = Cv∙A/λ = \n7.0\n 10-12 s, which is twice as large as the value for Rh 2CoSb. Though it is unknown \nfor the coercivity in films for Rh 2CoSb , we a ssume that Hc is proportional to the \nanisotropy field Ha, hence \n is proportion al to (Ha(300K) -0)/(Tc-T300K). The Ha for 16 \n Rh2CoSb and FePt are 12.4 T and 11.0 T at room temperature, and their Tc are 450 K \nand 750 K , respectively. Therefore , \n for Rh2CoSb is about 3.4 times larger tha n \nthat for FePt . The writing speed ratio for Rh 2CoSb and FePt , propo rtional to \n , \nis therefore 6.8. 17 \n \nReference s \n \n1. C. Hass, Spin -Disorder Scattering and Magnetoresistance of Magnetic \nSemiconductors, Phys. Rev . 168, 531 (1968). \n2. D. Elefant, R. Schä fer, Giant negative do main wall resistance in iron, Phys. Rev. B \n82, 134438 (2010). \n3. N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, & N. P. Ong, Anomalous hall \neffect. Rev. Mod. Phys. 82, 1539 (2010). \n4. E. Liu et al. Giant anomalous Hall effect in a ferromagnetic kagome -lattice \nsemimetal. Nat. Phys. 14, 1125 (2018). \n5. Y . Wang et al. Anisotropic anomalous Hall effect in triangular itinerant ferromagnet \nFe3GeTe 2. Phys. Rev. B 96, 134428 (2017). \n6. J. Y . Liu et al. A magnetic topological semimetal Sr 1-yMn 1-zSb2 (y, z < 0.1). Nat. \nMater. 16, 905 (2017). \n7. Y . J. Song et al. Synthesis, anisotropy, and superconducting properties of LiFeAs \nsingle crystal. Appl. Phys. Lett. 96, 212508 (2010). \n8. Y . Eltsev et al. Anisotropic resistivity and Hall effect in MgB 2 single crystals. Phys . \nRev. B 66, 180504(R) (2002). \n9. S. W. Tozer, A . W. Kleinsasser, T. Penney, D. Kaiser & F. Holtzberg. Measurement \nof Anisotropic Resistivity and Hall Constant for Single -Crystal YBa 2Cu3O7-x. Phys. \nRev. Lett. 95, 1768 (1987). \n10. P. Vir, et al. Anisotropic topological Hall effect with real and momentum space \nBerry curvature in the antiskyrmion -hosting Heusler compound Mn 1.4PtSn. Phys. Rev. \nB 99, 140406(R) (2019). \n11. R. Troc et al. Electronic, magnetic, transport, and thermal properties of \nsingle -crystalline UFe 2Al10. Phys. Rev. B 92, 104427 (2015). \n12. N. Kiyohara et al. Giant Anomalous Hall Effect in the Chiral Antiferromagnet \nMn3Ge. Phys. Rev. Appl. 5, 064009 (2016). \n13. J. M. Ziman “Electrons and Phonons”, Oxford UniversityPress (1958) . \n14. J. F. Nye “Physic al Properties of crystals”, Oxford Science Publications (1955) . \n15. P. Stamenov, Point contact Andreev reflection from semimetallic bismuth —The \nroles of the minority carriers and the large spin -orbit coupling. J. Appl. Phys. 113, 18 \n \n17C718 (2013). \n16. A. Hubert , R. Schä fer, Magnetic Domains, Springer, Berlin, p.p. 314, 379 (1998). \n17. P. Blaha, K. Schwarz , P. Sorantin , S. B. Trickey, Full-potential, linearized \naugmented plane wave programs for crystalline systems. Commun. Comput. Phys 59, \n399-415 (1990). \n18. K. Schwarz & P. Blaha, Solid state calculations using WIEN2k. Comput. Mater. \nSci. 28, 259 -273 (2003). \n19. P. Blaha, K. Schwarz , G. K. Madsen , D. Kvasnicka & J. Luitz, wien2k: An \naugmented plane wave+ local orbitals program for calculating crystal properties \n(Wie n, 2013). \n20. P. Blaha , K. Schwarz , F. Tran, R. Laskowski , G.K.H. Madsen & L.D. Marks, \nWIEN2k: An APW+lo program for calculating the properties of solids, J. Chem. Phys. \n152, 074101 (2020). \n21. H. Ebert, Fully relativistic band structure calculations for mag netic \nsolids -formalism and application. In Electronic Structure and Physical Properties of \nSolids (pp. 191 -246). Springer, Berlin, Heidelberg (1999). \n22. H. Ebert , D. Koedderitzsch & J. Minar, Calculating condensed matter properties \nusing the KKR -Green's function method —recent developments and applications. Rep. \nProg. Phys. 74, 096501 (2011). \n23. J. P. Perdew , K. Burke & M. Ernzerhof , Generalized Gradient Approximation \nMade Simple , Phys. Rev. Let t. 77, 3865 (1996). \n24. M. S.S. Brooks , Calculated ground state properties of light actinide metals and \ntheir compounds, Physica B 130, 6 (1985). \n25. J. M. D. Coey, Magnetism and Magnetic Materials Ch. 4 and 8 (Cambridge Univ. \nPress, Cambridge, 2010). \n26. D. Weller , G. Parker , O. Mosendz , A. Lyberatos , D. Mitin , N. Y . Safonova & M. \nAlbrecht, FePt heat assisted magnetic recording media , J. Vac. Sci. & Technol. B 34, \n060801 (2016). \n27. C. V ogler , C. Abert , F. Bruckner , D. Suess , D. Praetorius , Heat -assisted magne tic \nrecording of bit -patterned media beyond 10 Tb/in2, Appl. Phys. Lett. 108, 102406 \n(2016). " }, { "title": "2009.01638v3.MAELAS__MAgneto_ELAStic_properties_calculation_via_computational_high_throughput_approach.pdf", "content": "MAELAS: MAgneto-ELAStic properties calculation\nvia computational high-throughput approach\nP. Nievesa,\u0003, S. Arapana, S. H. Zhangb,c, A. P. K ˛ adzielawaa, R. F. Zhangb,c,\nD. Leguta\naIT4Innovations, VŠB - Technical University of Ostrava, 17. listopadu 2172/15, 70800\nOstrava-Poruba, Czech Republic\nbSchool of Materials Science and Engineering, Beihang University, Beijing 100191, PR China\ncCenter for Integrated Computational Materials Engineering, International Research Institute\nfor Multidisciplinary Science, Beihang University, Beijing 100191, PR China\nAbstract\nIn this work, we present the program MAELAS to calculate magnetocrys-\ntalline anisotropy energy, anisotropic magnetostrictive coefficients and magnetoe-\nlastic constants in an automated way by Density Functional Theory calculations.\nThe program is based on the length optimization of the unit cell proposed by Wu\nand Freeman to calculate the magnetostrictive coefficients for cubic crystals. In\naddition to cubic crystals, this method is also implemented and generalized for\nother types of crystals that may be of interest in the study of magnetostrictive ma-\nterials. As a benchmark, some tests are shown for well-known magnetic materials.\nKeywords: Magnetostriction, Magnetoelasticity, High-throughput computation,\nFirst-principles calculations\nPROGRAM SUMMARY\nProgram Title: MAELAS\nDeveloper’s respository link: https://github.com/pnieves2019/MAELAS\nLicensing provisions: BSD 3-clause\nProgramming language: Python3\nNature of problem: To calculate anisotropic magnetostrictive coefficients and magnetoe-\nlastic constants in an automated way based on Density Functional Theory methods.\nSolution method: In the first stage, the unit cell is relaxed through a spin-polarized cal-\n\u0003Corresponding author.\nE-mail address: pablo.nieves.cordones@vsb.cz\nPreprint submitted to Computer Physics Communications February 9, 2021arXiv:2009.01638v3 [cond-mat.mtrl-sci] 7 Feb 2021culation without SOC. Next, after a crystal symmetry analysis, a set of deformed lat-\ntice and spin configurations are generated using the pymatgen library [1]. The energy of\nthese states is calculated by the first-principles code V ASP [3], including the SOC. The\nanisotropic magnetostrictive coefficients are derived from the fitting of these energies to a\nquadratic polynomial [2]. Finally, if the elastic tensor is provided [4], then the magnetoe-\nlastic constants are calculated too.\nAdditional comments including restrictions and unusual features: This version supports\nthe following crystal systems: Cubic (point groups 432, ¯43m,m¯3m), Hexagonal (6 mm,\n622, ¯62m, 6=mmm ), Trigonal (32, 3 m,¯3m), Tetragonal (4 mm, 422, ¯42m, 4=mmm ) and\nOrthorhombic (222, 2 mm,mmm ).\nReferences\n[1] S. P. Ong, W. D. Richards, A. Jain, G. Hautier, M. Kocher, S. Cholia, D. Gunter, V .\nL. Chevrier, K. A. Persson, and G. Ceder, Comput. Mater. Sci. 68, 314 (2013).\n[2] R. Wu, A. J. Freeman, Journal of Applied Physics 79, 6209–6212 (1996).\n[3] G. Kresse, J. Furthmüller, Phys. Rev. B 54 (1996) 11169.\n[4] S. Zhang and R. Zhang, Comput. Phys. Commun. 220, 403 (2017).\n1. Introduction\nA magnetostrictive material is one which changes in size due to a change\nof state of magnetization. These materials are characterized by magnetostrictive\ncoefficients ( l). In many technical applications such as electric transformers, mo-\ntor shielding, and magnetic recording, magnetic materials with extremely small\nmagnetostrictive coefficients are required. By contrast, materials with large mag-\nnetostrictive coefficients are needed for many applications in electromagnetic mi-\ncrodevices as actuators and sensors [1–4]. Typically, elementary Rare-Earth (R)\nmetals (under low temperature and high magnetic field) and compounds with R\nand transition metals exhibit a high magnetostriction ( l>10\u00003). In particular,\nthe highest magnetostrictions were found in the RFe 2compounds with Laves\nphase C15 structure type (face centered cubic) [5]. For instance, Terfenol-D\n(Tb 0:27Dy0:73Fe2) is a widely used magnetostrictive material thanks to its giant\nmagnetostriction along [111] crystallographic direction ( l111=1:6\u000210\u00003) un-\nder moderate magnetic fields ( <2 kOe) at room temperature [6]. Beyond cubic\n2systems, the research of magnetostrictive materials has been focused on hexagonal\ncrystals like RCo 5(space group 191), hexagonal and trigonal R 2Co7and R 2Co17\nseries, and tetragonal R 2Fe14B [7, 8]. More recently, the problem of R avail-\nability [9] has also motivated the exploration of R-free magnetostrictive materials\nlike Galfenol (Fe-Ga), spinel ferrites (CoFe 2O4), Nitinol (Ni-Ti alloys), Fe-based\nInvars, and Ni 2MnGa [10–12].\nConcerning the theory of magnetostriction, the basic equations for cubic (I)\ncrystals were developed by Akulov [13] and Becker et al. [14] in the 1920s and\n30s. In the next three decades, great advances took place due to the outstanding\nworks of Mason [15], Clark et al. [16], and Callen and Callen [17], as well as\nmany others, where the theory was extended to other crystal symmetries. Over\nthe last decades, modern electronic structure theory based on Density Functional\nTheory (DFT) has been successfully applied to describe magnetostriction of many\nmaterials [1, 10, 11, 18–29]. Nowadays, a common method to calculate magne-\ntostrictive coefficients is based on the optimization of the unit cell length pro-\nposed by Wu and Freeman for cubic crystals [18, 19]. In this work, we present the\nMAELAS program where this methodology is implemented and generalized for\nthe main crystal symmetries in the research field of magnetostriction. The paper\nis organized as follows. In Section 2, we review some theoretical concepts and\nequations of magnetostriction. In Section 3, we explain in detail the methodol-\nogy and workflow of the program, while some examples are shown in Section 4.\nThe paper ends with a summary of the main conclusions and future perspectives\n(Section 5).\n2. Theory of magnetostriction\nThe magnetostrictive response is mainly originated by two kind of sources: (i)\nisotropic exchange interaction and (ii) strain dependence of magnetocrystalline\nanisotropy [8]. The magnetostriction due to isotropic exchange leads to frac-\ntional volume changes, and doesn’t depend on the magnetization direction [30].\nOn the other hand, the strain dependence of magnetocrystalline anisotropy is\nresponsible for the magnetostriction that depends on the magnetization orienta-\ntion (anisotropic), and is originated by the spin-orbit coupling (SOC) and crystal\nfield interactions [8, 31]. The current version of the program MAELAS calcu-\nlates the magnetostrictive coefficients and magnetoelastic constants related to the\nanisotropic magnetostriction.\nLet’s consider l0the initial length of a demagnetized material along the di-\nrection bbb(jbbbj=1), and lthe final length along the same direction bbbwhen the\n3Figure 1: Magnetostriction of a single crystal under an external magnetic field ( aaakHHH) perpendicu-\nlar to the measured length direction ( bbb?HHH). Symbols MandMsstand for macroscopic magnetiza-\ntion and saturation magnetization, respectively. Dash line on the right represents the original size\nof the demagnetized material. The magnetostriction effect has been magnified in order to help to\nvisualize it easily, in real materials it is smaller ( Dl=l0\u001810\u00003\u000010\u00006).\nsystem is magnetized along the direction aaa(jaaaj=1). The relative length change\n(l\u0000l0)=l0=Dl=l0can be written as [8]\nDl\nl0\f\f\f\f\faaa\nbbb=å\ni;j=x;y;zeeq\ni j(aaa)bibj; (1)\nwhere eeq\ni jis the equilibrium strain tensor. This equation describes the Joule effect\n[32], once it is rewritten in terms of the magnetostrictive coefficients ( l) conve-\nniently. Fig.1 shows a sketch of magnetostriction.\nThe deformation of a solid can be described in terms of the displacement vec-\ntoruuu(rrr) =rrr000\u0000rrrthat gives the displacement of a point at the initial position rrr\nto its final position rrr000after it is deformed. For small deformations (infinitesimal\nstrain theory), the strain tensor ( ei j) can be expressed in terms of the displacement\nvector as[33]\nei j=1\n2\u0012¶ui\n¶rj+¶uj\n¶ri\u0013\n; i;j=x;y;z (2)\nwhere ¶ui=¶rjis called the displacement gradient (second-order tensor). The equi-\nlibrium strain tensor is obtained through the minimization of both the elastic ( Eel)\n4and magnetoelastic ( Eme) energies [5, 8]\n¶(Eel+Eme)\n¶ei j=0; i;j=x;y;z (3)\nwhere the total energy must be invariant under the symmetry operations of the\ncrystal lattice [5]. Let’s write general equations for EelandEme. The elastic energy\ndepends on the fourth-order elastic stiffness tensor ci jklthat links the second-order\nstrain and stress ( si j) tensors through the generalized Hooke’s law\nsi j=å\nk;l=x;y;zci jklekl;i;j=x;y;z: (4)\nTaking advantage of the symmetry of stress and strain tensors, the Hooke’s law\ncan be written in matrix notation as\n0\nBBBBBB@sxx\nsyy\nszz\nsyz\nsxz\nsxy1\nCCCCCCA=0\nBBBBBB@cxxxx cxxyy cxxzz cxxyz cxxzx cxxxy\ncyyxx cyyyy cyyzz cyyyz cyyzx cyyxy\nczzxx czzyy czzzz czzyz czzzx czzxy\ncyzxx cyzyy cyzzz cyzyz cyzzx cyzxy\nczxxx czxyy czxzz czxyz czxzx czxxy\ncxyxx cxyyy cxyzz cxyyz cxyzx cxyxy1\nCCCCCCA0\nBBBBBB@exx\neyy\nezz\n2eyz\n2exz\n2exy1\nCCCCCCA(5)\nTo facilitate the manipulation of this equation it is convenient to define the follow-\ning six-dimensional vectors (V oigt notation)\n˜sss=0\nBBBBBB@˜s1\n˜s2\n˜s3\n˜s4\n˜s5\n˜s61\nCCCCCCA=0\nBBBBBB@sxx\nsyy\nszz\nsyz\nsxz\nsxy1\nCCCCCCA; ˜eee=0\nBBBBBB@˜e1\n˜e2\n˜e3\n˜e4\n˜e5\n˜e61\nCCCCCCA=0\nBBBBBB@exx\neyy\nezz\n2eyz\n2exz\n2exy1\nCCCCCCA; (6)\nand replace ci jklbyCnmcontracting a pair of cartesian indices into a single integer:\nxx!1,yy!2,zz!3,yz!4,xz!5 and xy!6. Using these conversion rules\nthe Hooke’s law is simplified to\n˜si=6\nå\nj=1Ci j˜ej;i=1;:::;6 (7)\n5where in matrix form reads\n0\nBBBBBB@˜s1\n˜s2\n˜s3\n˜s4\n˜s5\n˜s61\nCCCCCCA=0\nBBBBBB@C11C12C13C14C15C16\nC21C22C23C24C25C26\nC31C32C33C34C35C36\nC41C42C42C44C45C46\nC51C52C53C54C55C56\nC61C62C63C64C65C661\nCCCCCCA0\nBBBBBB@˜e1\n˜e2\n˜e3\n˜e4\n˜e5\n˜e61\nCCCCCCA: (8)\nwhere Ci j=Cji. Then the elastic energy up to second-order in the strain can be\nwritten as\nEel=E0+V0\n26\nå\ni;j=1Ci j˜ei˜ej+O(˜e3); (9)\nwhere E0andV0are the equilibrium energy and volume, respectively. The magne-\ntoelastic energy Emecomes from the strain dependence of the magnetocrystalline\nanisotropy energy (MAE) EK[34, 35]. Performing a Taylor expansion of EKin\nthe strain we have\nEK=E0\nK+6\nå\ni=1\u0012¶EK\n¶˜ei\u0013\n0˜ei+1\n26\nå\ni;j=1\u0012¶2EK\n¶˜ei¶˜ej\u0013\n0˜ei˜ej+O(˜e3); (10)\nwhere E0\nKcorresponds to the MAE of the undeformed state that contains the mag-\nnetocrystalline anisotropy constants K. The third term in the right hand side of\nEq.10 is the second-order magnetoelastic energy that leads to a very small ad-\nditional contribution to the second-order elastic energy given by Eq.9, so that is\nusually neglected [34, 36]. The first-order magnetoelastic energy\nEme=6\nå\ni=1\u0012¶EK\n¶˜ei\u0013\n0˜ei (11)\nis obtained by taking the direct product of the symmetry strains and direction co-\nsine polynomial for each irreducible representation, multiplying by a constant,\ncalled the magnetoelastic constant and finally summing over the different rep-\nresentations [5, 8, 16, 17]. Frequently, the first-order magnetoelastic energy is\nconsidered up to second-order of the direction cosine polynomial a. In cartesian\ncoordinates, it may be written as\nEme=3\nå\ni=1gi(a0)˜ei+6\nå\ni=1fi(a2)˜ei+O(a4); (12)\n6Table 1: Number of independent second-order elastic constants of each crystal system. Number of\nindependent magnetoelastic and magnetostrictive coefficients up to second-order of the direction\ncosine polynomial in the first-order magnetoelastic energy. In the last column we specify which\ncrystal systems are supported by the current version of MAELAS.\nCrystal system Point groupsSpace\ngroupsElastic\nconstants\n(Ci j)Magnetoelastic\nconstants\n(b)Magnetostriction\ncoefficients\n(l)MAELAS\nTriclinic 1 ;¯1 1\u00002 21 36 36 No\nMonoclinic 2 ;m;2=m 3\u000015 13 20 20 No\nOrthorhombic 222 ;2mm;mmm 16\u000074 9 12 12 Yes\nTetragonal (II) 4 ;¯4;4=m 75\u000088 7 10 10 No\nTetragonal (I) 4 mm;422;¯42m;4=mmm 89\u0000142 6 7 7 Yes\nTrigonal (II) 3 ;¯3 143\u0000148 7 12 12 No\nTrigonal (I) 32 ;3m;¯3m 149\u0000167 6 8 8 Yes\nHexagonal(II) 6 ;¯6;6=m 168\u0000176 5 8 8 No\nHexagonal (I) 6 mm;622;¯62m;6=mmm 177\u0000194 5 6 6 Yes\nCubic (II) 23 ;m¯3 195\u0000206 3 4 4 No\nCubic (I) 432 ;¯43m;m¯3m 207\u0000230 3 3 3 Yes\nwhere functions giand ficontain the magnetoelastic constants ( b). In the fol-\nlowing subsections, we show the form of Eqs.1, 9 and 12 for the main crystal\nsymmetries studied in magnetostriction, which are implemented in the program\nMAELAS. The remaining crystal systems not discussed here might be included\nin the new versions of the code. In Table 1, we present a summary of the crystal\nsystems supported by MAELAS. Here, we use the notation of Wallace [37, 38]\n(I/II) to distinguish Laue classes within the same crystal system.\nBefore analyzing each crystal system, we must make an important remark\nabout the notation for the strain tensor ei j. In previous works discussing magne-\ntostriction like Refs.[5, 8, 35], the V oigt definition of the strain tensor was used\n(eV\ni j;i;j=x;y;z) [39], which is related to the one defined in the present work as\neii=eV\nii, 2ei j=eV\ni j;i6=j. Consequently, the following elastic and magnetoelastic\nenergies (in terms of the strain tensor with two cartesian indices) contain numeri-\ncal factors different to those given in Kittel and Clark works [5, 35] for the terms\nwith non-diagonal elements of the strain tensor ( ei j;i6=j). The following expres-\nsions for the relative length change ( Dl=l0) are the same as in Kittel and Clark\nworks [5, 35] because the sum in Eq.1 runs over all possible values of indices\ni;j=x;y;z, while in Kittel and Clark works [5, 35] the sum runs up to i>j.\nIn the present work, the equations of magnetostrictive coefficients expressed in\nterms of the elastic and magnetoelastic constants are also the same to those given\nin Kittel and Clark works [5, 35].\n72.1. Cubic (I)\n2.1.1. Single crystal\nFor cubic (I) systems (point groups 432, ¯43m,m¯3m) the elastic stiffness tensor\nreads\nCcub=0\nBBBBBB@C11C12C12 0 0 0\nC12C11C12 0 0 0\nC12C12C11 0 0 0\n0 0 0 C44 0 0\n0 0 0 0 C44 0\n0 0 0 0 0 C441\nCCCCCCA; (13)\nso there are three independent elastic constants C11,C12andC44. Hence, the\nelastic energy Eq.9 becomes\nEcub\nel\u0000E0\nV0=C11\n2(˜e2\n1+˜e2\n2+˜e2\n3)+C12(˜e1˜e2+˜e1˜e3+˜e2˜e3)\n+C44\n2(˜e2\n4+˜e2\n5+˜e2\n6)\n=cxxxx\n2(e2\nxx+e2\nyy+e2\nzz)+cxxyy(exxeyy+exxezz+eyyezz)\n+2cyzyz(e2\nxy+e2\nyz+e2\nxz);(14)\nwhere C11=cxxxx,C12=cxxyyandC44=cyzyz. On the other hand, the first-order\nmagnetoelastic energy up to second-order direction cosine polynomial contains 3\nmagnetoelastic constants [17]. From the symmetry strains and direction cosine\npolynomial for each irreducible representation, it is possible to obtain the follow-\ning magnetoelastic energy in cartesian coordinates [5, 8, 10]\nEcub(I)\nme\nV0=b0(˜e1+˜e2+˜e3)+b1(a2\nx˜e1+a2\ny˜e2+a2\nz˜e3)\n+b2(axay˜e6+axaz˜e5+ayaz˜e4)\n=b0(exx+eyy+ezz)+b1(a2\nxexx+a2\nyeyy+a2\nzezz)\n+2b2(axayexy+axazexz+ayazeyz);(15)\nwhere b0is the volume magnetoelastic constant, and b1andb2are the anisotropic\nmagnetoelastic constants. Next, replacing Eqs.14 and 15 into Eq.3, we find the\n8following equilibrium strains\neeq\ni j=\u0000b2aiaj\n2C44; i6=j; i;j=x;y;z\neeq\nii=\u0000b1a2\ni\nC11\u0000C12\u0000b0\nC11+2C12+b1C12\n(C11\u0000C12)(C11+2C12);i=x;y;z(16)\nInserting these equilibrium strains into Eq.1 gives\nDl\nl0\f\f\f\f\faaa\nbbb=la+3\n2l001\u0012\na2\nxb2\nx+a2\nyb2\ny+a2\nzb2\nz\u00001\n3\u0013\n+3l111(axaybxby+ayazbybz+axazbxbz);(17)\nwhere\nla=\u0000b0\u00001\n3b1\nC11+2C12;\nl001=\u00002b1\n3(C11\u0000C12);\nl111=\u0000b2\n3C44:(18)\nThe coefficient ladescribes the volume magnetostriction, while l001andl111are\nthe anisotropic magnetostrictive coefficients that give the fractional length change\nalong the [001] and [111] directions when a demagnetized material is magnetized\nin these directions, respectively. The superscript ainlastands for one irreducible\nrepresentation of the group of transformations which take the crystal into itself [5,\n8], so it should not be confused with the direction of magnetization aaa. The MAE\nin an unstrained cubic crystal up to sixth-order of direction cosine polynomial is\n[35, 40]\nE0\nK\nV0=K0+K1(a2\nxa2\ny+a2\nxa2\nz+a2\nya2\nz)+K2a2\nxa2\nya2\nz; (19)\nwhere K0,K1andK2are the magnetocrystalline anisotropy constants.\n2.1.2. Polycrystal\nThe theory of magnetostriction for polycrystalline materials is more complex\nthan for single crystals. A widely used approximation is to assume that the stress\ndistribution is uniform through the material. In this case the relative change in\n9length may be put into the form [8, 13, 34, 41]\nDl\nl0\f\f\f\f\faaa\nbbb=3\n2lS\u0014\n(aaa\u0001bbb)2\u00001\n3\u0015\n; (20)\nwhere\nlS=2\n5l001+3\n5l111: (21)\nThis result is analogous to the Reuss approximation used in the elastic theory of\npolycrystals to obtain a lower bound of bulk and shear modulus [8, 42–44]. A\ndiscussion about the limitations of this approximation can be found in Ref.[45].\n2.2. Hexagonal (I)\n2.2.1. Single crystal\nThe elastic stiffness tensor for hexagonal (I) system (point groups 6 mm, 622,\n¯62m, 6=mmm ) reads\nChex=0\nBBBBBB@C11C12C13 0 0 0\nC12C11C13 0 0 0\nC13C13C33 0 0 0\n0 0 0 C44 0 0\n0 0 0 0 C44 0\n0 0 0 0 0C11\u0000C12\n21\nCCCCCCA; (22)\nso that it has five independent elastic constants C11,C12,C13,C33andC44. As a\nresult, the elastic energy Eq.9 is\nEhex\nel\u0000E0\nV0=1\n2C11(˜e2\n1+˜e2\n2)+C12˜e1˜e2+C13(˜e1+˜e2)˜e3+1\n2C33˜e2\n3\n+1\n2C44(˜e2\n4+˜e2\n5)+1\n4(C11\u0000C12)˜e2\n6\n=1\n2cxxxx(e2\nxx+e2\nyy)+cxxyyexxeyy+cxxzz(exx+eyy)ezz+1\n2czzzze2\nzz\n+2cyzyz(e2\nyz+e2\nxz)+(cxxxx\u0000cxxyy)e2\nxy(23)\nwhere C11=cxxxx,C12=cxxyy,C13=cxxzz,C33=czzzz, and C44=cyzyz. The first\norder magnetoelastic energy up to quadratic direction cosine polynomial contains\n106 magnetoelastic constants [17]. In cartesian coordinates it can be written as [5]\nEhex(I)\nme\nV0=b11(exx+eyy)+b12ezz+b21\u0012\na2\nz\u00001\n3\u0013\n(exx+eyy)+b22\u0012\na2\nz\u00001\n3\u0013\nezz\n+b3\u00141\n2(a2\nx\u0000a2\ny)(exx\u0000eyy)+2axayexy\u0015\n+2b4(axazexz+ayazeyz):\n(24)\nOnce the equilibrium strains are calculated by minimizing Eqs.23 and 24 through\nEq.3 and inserted into Eq.1, one finds [5, 8, 16]\nDl\nl0\f\f\f\f\faaa\nbbb=la1;0(b2\nx+b2\ny)+la2;0b2\nz+la1;2\u0012\na2\nz\u00001\n3\u0013\n(b2\nx+b2\ny)\n+la2;2\u0012\na2\nz\u00001\n3\u0013\nb2\nz+lg;2\u00141\n2(a2\nx\u0000a2\ny)(b2\nx\u0000b2\ny)+2axaybxby\u0015\n+2le;2(axazbxbz+ayazbybz);(25)\nwhere\nla1;0=b11C33+b12C13\nC33(C11+C12)\u00002C2\n13;\nla2;0=2b11C13\u0000b12(C11+C12)\nC33(C11+C12)\u00002C2\n13;\nla1;2=\u0000b21C33+b22C13\nC33(C11+C12)\u00002C2\n13;\nla2;2=2b21C13\u0000b22(C11+C12)\nC33(C11+C12)\u00002C2\n13;\nlg;2=\u0000b3\nC11\u0000C12;\nle;2=\u0000b4\n2C44:(26)\nThese magnetostrictive coefficients are related to the normal strain modes for a\ncylinder [5, 8]. The equation of the relative length change in the form of Eq.25\nwas proposed by Clark et al. [16]. In literature there are different arrangements of\nthe right hand side of Eq.25 that leads to other definitions of the magnetostrictive\ncoefficients, like those defined by Mason [15], Birss [34], and Callen and Callen\n[17]. The conversion formulas between Eq.25 and all these other conventions can\n11be found in Appendix A. These conversion formulas are implemented in the pro-\ngram MAELAS, so that the magnetostrictive coefficients are also given according\nto these definitions. Note that in some works [40, 45] the magnetostrictive coef-\nficients lg;2andle;2in Eq.25 are named as le;2andlz;2, respectively, which is\nmore consistent with the Bethe’s group-theoretical notation [45]. The MAE in an\nunstrained hexagonal crystal up to fourth-order of areads [40]\nE0\nK\nV0=K0+K1(1\u0000a2\nz)+K2(1\u0000a2\nz)2: (27)\n2.2.2. Polycrystal\nUnder the assumption of uniform stress, the relative change in length for poly-\ncrystalline hexagonal (I) systems can be written as [34]\nDl\nl0\f\f\f\f\faaa\nbbb=x+h(aaa\u0001bbb)2; (28)\nwhere his, in both easy axis and easy plane MAE, given by\nh=\u00002\n15Q4+1\n5Q6+7\n15Q8: (29)\nThe quantity xis different for easy axis and easy plane. In the case of easy axis, x\nis given by\nx=2\n3Q2+4\n15Q4\u00001\n15Q6+1\n15Q8; (easy axis) (30)\nwhile for easy plane is\nx=\u00001\n3Q2\u00001\n15Q4\u00001\n15Q6\u00004\n15Q8: (easy plane) (31)\nThe quantities Qi(i=2;4;6;8) are the anisotropic magnetostrictive coefficients in\nBirss’s convention [34], and are related to the magnetostrictive coefficients defined\nin Eq. 25 through Eq. A.4. We have implemented these formulas in MAELAS,\nso that it also calculates handx.\n2.3. Trigonal (I)\n2.3.1. Single crystal\nThe elastic stiffness tensor for trigonal (I) system (point groups 32, 3 m,¯3m)\nhas 6 independent elastic constants C11,C12,C13,C33,C44andC14, and it is given\n12by\nCtrig(I)=0\nBBBBBB@C11 C12 C13 C14 0 0\nC12 C11 C13\u0000C14 0 0\nC13 C13 C33 0 0 0\nC14\u0000C14 0 C44 0 0\n0 0 0 0 C44 C14\n0 0 0 0 C14C11\u0000C12\n21\nCCCCCCA: (32)\nHence, inserting this tensor into Eq.9 we have the following elastic energy\nEtrig(I)\nel\u0000E0\nV0=1\n2C11(˜e2\n1+˜e2\n2)+C12˜e1˜e2+C13(˜e1+˜e2)˜e3+1\n2C33˜e2\n3\n+1\n2C44(˜e2\n5+˜e2\n4)+1\n4(C11\u0000C12)˜e2\n6+C14(˜e6˜e5+˜e1˜e4\u0000˜e2˜e4):\n=1\n2cxxxx(e2\nxx+e2\nyy)+cxxyyexxeyy+cxxzz(exx+eyy)ezz+1\n2czzzze2\nzz\n+2cyzyz(e2\nxz+e2\nyz)+(cxxxx\u0000cxxyy)e2\nxy+4cxxyz(exyexz+exxeyz\u0000eyyeyz):\n(33)\nwhere C11=cxxxx,C12=cxxyy,C13=cxxzz,C14=cxxyz,C33=czzzz, and C44=cyzyz.\nOn the other hand, the magnetoelastic energy contains 8 independent magnetoe-\nlastic constants [17]. In cartesian coordinates it can be written as [8]\nEtrig(I)\nme\nV0=b11(exx+eyy)+b12ezz+b21\u0012\na2\nz\u00001\n3\u0013\n(exx+eyy)+b22\u0012\na2\nz\u00001\n3\u0013\nezz\n+b3\u00141\n2(a2\nx\u0000a2\ny)(exx\u0000eyy)+2axayexy\u0015\n+2b4(axazexz+ayazeyz)\n+b14\u0002\n(a2\nx\u0000a2\ny)eyz+2axayexz\u0003\n+b34\u00141\n2ayaz(exx\u0000eyy)+2axazexy\u0015\n:\n(34)\n13Next, we obtain the equilibrium strains via Eq.3. Replacing them into Eq.1 leads\nto [8]\nDl\nl0\f\f\f\f\faaa\nbbb=la1;0(b2\nx+b2\ny)+la2;0b2\nz+la1;2\u0012\na2\nz\u00001\n3\u0013\n(b2\nx+b2\ny)\n+la2;2\u0012\na2\nz\u00001\n3\u0013\nb2\nz+lg;1\u00141\n2(a2\nx\u0000a2\ny)(b2\nx\u0000b2\ny)+2axaybxby\u0015\n+lg;2(axazbxbz+ayazbybz)+l12\u00141\n2ayaz(b2\nx\u0000b2\ny)+axazbxby\u0015\n+l21\u00141\n2(a2\nx\u0000a2\ny)bybz+axaybxbz\u0015\n;(35)\nwhere\nla1;0=b11C33+b12C13\nC33(C11+C12)\u00002C2\n13;\nla2;0=2b11C13\u0000b12(C11+C12)\nC33(C11+C12)\u00002C2\n13;\nla1;2=\u0000b21C33+b22C13\nC33(C11+C12)\u00002C2\n13;\nla2;2=2b21C13\u0000b22(C11+C12)\nC33(C11+C12)\u00002C2\n13;\nlg;1=C14b14\u0000C44b3\n1\n2C44(C11\u0000C12)\u0000C2\n14;\nlg;2=1\n2b4(C11\u0000C12)\u0000b34C14\n1\n2C44(C11\u0000C12)\u0000C2\n14;\nl12=C14b4\u0000C44b34\n1\n2C44(C11\u0000C12)\u0000C2\n14;\nl21=1\n2b14(C11\u0000C12)\u0000b3C14\n1\n2C44(C11\u0000C12)\u0000C2\n14:(36)\nThe MAE in an unstrained trigonal crystal up to fourth-order in ais the same to\nthe hexagonal case (Eq.27).\n142.4. Tetragonal (I)\n2.4.1. Single crystal\nThe tetragonal (I) crystal system (point groups 4 mm, 422, ¯42m, 4=mmm ) has\nthe following elastic stiffness tensor\nCtet(I)=0\nBBBBBB@C11C12C13 0 0 0\nC12C11C13 0 0 0\nC13C13C33 0 0 0\n0 0 0 C44 0 0\n0 0 0 0 C44 0\n0 0 0 0 0 C661\nCCCCCCA; (37)\nHence, it has six independent elastic constants C11,C12,C13,C33,C44andC66.\nThe elastic energy is given by\nEtet(I)\nel\u0000E0\nV0=1\n2C11(˜e2\n1+˜e2\n2)+C12˜e1˜e2+C13(˜e1+˜e2)˜e3+1\n2C33˜e2\n3\n+1\n2C44(˜e2\n4+˜e2\n5)+1\n2C66˜e2\n6\n=1\n2cxxxx(e2\nxx+e2\nyy)+cxxyyexxeyy+cxxzz(exx+eyy)ezz+1\n2czzzze2\nzz\n+2cyzyz(e2\nyz+e2\nxz)+2cxyxye2\nxy\n(38)\nwhere C11=cxxxx,C12=cxxyy,C13=cxxzz,C33=czzzz,C44=cyzyzandC66=cxyxy.\nOn the other hand, there are 7 independent magnetoelastic constants [17]. The\nmagnetoelastic energy can be written as [8, 10]\nEtet(I)\nme\nV0=b11(exx+eyy)+b12ezz+b21\u0012\na2\nz\u00001\n3\u0013\n(exx+eyy)+b22\u0012\na2\nz\u00001\n3\u0013\nezz\n+1\n2b3(a2\nx\u0000a2\ny)(exx\u0000eyy)+2b0\n3axayexy+2b4(axazexz+ayazeyz):\n(39)\n15After the equilibrium strains are calculated by minimizing Eqs.38 and 39 through\nEq.3 and replaced into Eq.1, we have [8]\nDl\nl0\f\f\f\f\faaa\nbbb=la1;0(b2\nx+b2\ny)+la2;0b2\nz+la1;2\u0012\na2\nz\u00001\n3\u0013\n(b2\nx+b2\ny)\n+la2;2\u0012\na2\nz\u00001\n3\u0013\nb2\nz+1\n2lg;2(a2\nx\u0000a2\ny)(b2\nx\u0000b2\ny)+2ld;2axaybxby\n+2le;2(axazbxbz+ayazbybz);(40)\nwhere\nla1;0=b11C33+b12C13\nC33(C11+C12)\u00002C2\n13;\nla2;0=2b11C13\u0000b12(C11+C12)\nC33(C11+C12)\u00002C2\n13;\nla1;2=\u0000b21C33+b22C13\nC33(C11+C12)\u00002C2\n13;\nla2;2=2b21C13\u0000b22(C11+C12)\nC33(C11+C12)\u00002C2\n13;\nlg;2=\u0000b3\nC11\u0000C12;\nld;2=\u0000b0\n3\n2C66;\nle;2=\u0000b4\n2C44:(41)\nMason derived an equivalent equation to Eq.40 using a different arrangement of\nthe terms and definitions of the magnetostrictive coefficients [15]. The conversion\nformulas between the magnetostrictive coefficients in Eq.40 and those defined by\nMason are shown in Appendix B. The MAE in an unstrained tetragonal crystal\nup to fourth-order in ais the same to the hexagonal case (Eq.27).\n2.5. Orthorhombic\n2.5.1. Single crystal\nThe orthorhombic crystal system (point groups 222, 2 mm,mmm ) has 9 inde-\npendent elastic constants C11,C12,C13,C22,C23,C33,C44,C55andC66, its elastic\n16stiffness matrix reads[38, 44]\nCortho=0\nBBBBBB@C11C12C13 0 0 0\nC12C22C23 0 0 0\nC13C23C33 0 0 0\n0 0 0 C44 0 0\n0 0 0 0 C55 0\n0 0 0 0 0 C661\nCCCCCCA; (42)\nHence, inserting it into Eq.9 leads to the following expression for the elastic en-\nergy\nEortho\nel\u0000E0\nV0=1\n2C11˜e2\n1+1\n2C22˜e2\n2+C12˜e1˜e2+C13˜e1˜e3+C23˜e2˜e3+1\n2C33˜e2\n3\n+1\n2C44˜e2\n4+1\n2C55˜e2\n5+1\n2C66˜e2\n6\n=1\n2cxxxxe2\nxx+1\n2cyyyye2\nyy+cxxyyexxeyy+cxxzzexxezz+cyyzzeyyezz\n+1\n2czzzze2\nzz+2cyzyze2\nyz+2cxzxze2\nxz+2cxyxye2\nxy:(43)\nwhere C11=cxxxx,C22=cyyyy,C12=cxxyy,C13=cxxzz,C23=cyyzz,C33=czzzz,\nC44=cyzyz,C55=cxzxzandC66=cxyxy. The magnetoelastic energy contains 12 in-\ndependent magnetoelastic constants [17]. Mason derived the following expression\nof the relative length change [15]\nDl\nl0\f\f\f\f\faaa\nbbb=la1;0b2\nx+la2;0b2\ny+la3;0b2\nz+l1(a2\nxb2\nx\u0000axaybxby\u0000axazbxbz)\n+l2(a2\nyb2\nx\u0000axaybxby)+l3(a2\nxb2\ny\u0000axaybxby)\n+l4(a2\nyb2\ny\u0000axaybxby\u0000ayazbybz)+l5(a2\nxb2\nz\u0000axazbxbz)\n+l6(a2\nyb2\nz\u0000ayazbybz)+4l7axaybxby+4l8axazbxbz+4l9ayazbybz:\n(44)\nNote that we added the terms that describes the volume magnetostriction ( la1;0,\nla2;0andla3;0), which were not included in the original work of Mason [15]. The\nexpression of the magnetoelastic energy and the relations between magnetostric-\ntive coefficients, elastic and magnetoelastic constants were not shown by Mason\neither. For completeness, here we deduce it from Eqs. 43 and 44. To do so, we\n17aim to find the unknown functions giandfiin the general form of the magnetoe-\nlastic energy in cartesian coordinates given by Eq. 12. Firstly, we minimize Eqs.\n43 and 12 via Eq. 3. This gives a set of equations that links the unknown functions\ngiandfiwith the equilibrium strains. Next, we extract the equilibrium strains by\ndirect comparison between Eqs. 1 and 44. Finally, we substitute the equilibrium\nstrains into the set of equations that relates giandfiwith the equilibrium strains,\nfrom which we obtain giandfi. Inserting the calculated giandfiinto Eq. 12 we\nhave\nEortho\nme\nV0=b01exx+b02eyy+b03ezz+b1a2\nxexx+b2a2\nyexx+b3a2\nxeyy+b4a2\nyeyy\n+b5a2\nxezz+b6a2\nyezz+2b7axayexy+2b8axazexz+2b9ayazeyz;\n(45)\nwhere\nb01=\u0000C11la1;0\u0000C12la2;0\u0000C13la3;0\nb02=\u0000C12la1;0\u0000C22la2;0\u0000C23la3;0\nb03=\u0000C13la1;0\u0000C23la2;0\u0000C33la3;0\nb1=\u0000C11l1\u0000C12l3\u0000C13l5\nb2=\u0000C11l2\u0000C12l4\u0000C13l6\nb3=\u0000C12l1\u0000C22l3\u0000C23l5\nb4=\u0000C12l2\u0000C22l4\u0000C23l6\nb5=\u0000C13l1\u0000C23l3\u0000C33l5\nb6=\u0000C13l2\u0000C23l4\u0000C33l6\nb7=C66(l1+l2+l3+l4\u00004l7)\nb8=C55(l1+l5\u00004l8)\nb9=C44(l4+l6\u00004l9):(46)\nAlternatively, one can deduce the magnetoelastic energy using the general ap-\nproach based on the symmetry strains and direction cosine polynomial for each\nirreducible representation [5, 8, 17]. This approach may lead to different defi-\nnitions of the magnetoelastic constants and magnetostrictive coefficients, as we\nhave discussed for the hexagonal (I) and tetragonal (I) systems in Appendix A\nand Appendix B, respectively. A generalization of the approach taken by Becker\nand Doring [14] for orthorhombic crystals can be found in Ref. [46]. The MAE\n18in an unstrained orthorhombic crystal up to fourth-order in ais[15]\nE0\nK\nV0=K0+K1a2\nx+K2a2\ny: (47)\n3. Methodology\n3.1. Calculation of magnetostrictive coefficients and magnetoelastic constants\nThe methodology implemented in the program MAELAS to calculate the\nanisotropic magnetostrictive coefficients is a generalization of the approach pro-\nposed by Wu and Freeman for cubic crystals [18, 19]. In this method, one measur-\ning length direction bbbiand two magnetization directions ( aaai\n1andaaai\n2) are chosen\nfor each magnetostrictive coefficient ( li) in such a way that\nDl\nl0\f\f\f\f\faaai\n1\nbbbi\u0000Dl\nl0\f\f\f\f\faaai\n2\nbbbi=rili; (48)\nwhere riis a real number. In Table 2 we show the selected set of bbbi,aaai\n1andaaai\n2\nin MAELAS that fulfils Eq.48 for each ri. Next, the left hand side of Eq.48 is\nwritten as\nDl\nl0\f\f\f\f\faaai\n1\nbbbi\u0000Dl\nl0\f\f\f\f\faaai\n2\nbbbi=l1\u0000l0\nl0\u0000l2\u0000l0\nl0=2(l1\u0000l2)\n(l1+l2)h\n1\u0000l1+l2\u00002l0\nl1+l2i\n=2(l1\u0000l2)\nl1+l2\u0014\n1+l1+l2\u00002l0\nl1+l2+:::\u0015\n\u00192(l1\u0000l2)\nl1+l2;(49)\nwhere in the last approximation we assume jl1(2)\u0000l0j=l0\u001c1. This assumption\nis reasonable for all known magnetostrictive materials. For instance, a very large\nvalue of Dl=l0is about 4 :5\u000210\u00003found in TbFe 2(Laves phase C15) along direc-\ntion [111] at T=0K [6], where this approximation is fine. This approximation\nallows to get rid of l0(length along bbbin the macroscopic demagnetized state)\nwhich can’t be calculated with DFT methods easily. Combining Eqs.48 and 49\none can write the magnetostrictive coefficients as\nli=2(l1\u0000l2)\nri(l1+l2); (50)\n19where the value of rifor each liis given in Table 2. The quantities l1andl2cor-\nrespond to the cell length along bbbwhen the magnetization points to aaa1andaaa2, re-\nspectively, and are calculated through an optimization of the energy. Namely, a set\nof deformed unit cells is firstly generated using the deformation modes described\nin Appendix C. For each deformed cell, the energy is calculated constraining the\nspins to the directions given by aaa1andaaa2. Next, the energy versus the cell length\nalong bbbfor each spin direction aaa1(2)is fitted to a quadratic polynomial\nE(aaaj;l) =Ajl2+Bjl+Cj;j=1;2 (51)\nwhere Aj,BjandCj(j=1;2) are fitting parameters. The minimum of this func-\ntion for spin direction aaa1(2)corresponds to l1(2)=\u0000B1(2)=(2A1(2)). Once l1and\nl2are determined, one obtains the magnetostrictive coefficients using Eq.50. The\nmagnetostrictive coefficients can also be written in terms of the derivative of the\nenergy with respect levaluated at l=l2as [20]\nli\u0019\u00001\nhiB1\u0001¶[E(aaa2;l)\u0000E(aaa1;l)]\n¶l\f\f\f\f\f\nl=l2(52)\nwhere B1is always negative. In Table 2, we see that our choice of bbbandaaa1(2)\nmakes rdepend on some magnetostrictive coefficients for l7,l8andl9in or-\nthorhombic crystals. For instance, working out the coefficient l7via Eq.50 we\nhave\nl7=(a2+b2)(l1\u0000l2)\nab(l1+l2)\u0000(a\u0000b)(a[l1+l2]\u0000b[l3+l4])\n4ab; (53)\nwhere aandbare the relaxed (not distorted) lattice parameters of the unit cell.\nHere, MAELAS makes use of the values of l1,l2,l3andl4calculated pre-\nviously in order to compute l7. Note that a simpler expression for l7can be\nachieved choosing the measuring length direction bbb=\u0010\n1p\n2;1p\n2;0\u0011\n. However,\nfrom a computational point of view, it is easy to extract the cell length lalong\nbbb=\u0010\nap\na2+b2;bp\na2+b2;0\u0011\nof each deformed cell generated with the deformation\ngradients discussed in Appendix C. Similarly, one can deduce the explicit equa-\ntion for l8andl9.\nLastly, if the elastic tensor is provided in the format given by the program\nAELAS [44], then the magnetoelastic constants ( bk) are also calculated from the\nrelations bk=bk(li;Cnm)given in Section 2.\n20Table 2: Selected cell length ( bbb) and magnetization directions ( aaa1,aaa2) in MAELAS to calculate\nthe anisotropic magnetostrictive coefficients according to Eq.48. The first column shows the crys-\ntal system and the corresponding lattice convention set in MAELAS based on the IEEE format\n[44]. The second column presents the equation of the relative length change that we used in Eq.48\nfor each crystal system. In the last column we show the values of the parameter rthat is defined in\nEq.48. The symbols a;b;ccorrespond to the lattice parameters of the relaxed (not distorted) unit\ncell.\nCrystal systemDl\nl0Magnetostrictive\ncoefficientbbb a aa1 aaa2 r\nCubic (I) Eq.17 l001 (0;0;1) ( 0;0;1) ( 1;0;0)3\n2\naaakˆxxx,bbbkˆyyy,ccckˆzzz l111\u0010\n1p\n3;1p\n3;1p\n3\u0011 \u0010\n1p\n3;1p\n3;1p\n3\u0011 \u0010\n1p\n2;0;\u00001p\n2\u0011\n3\n2\nHexagonal (I) Eq.25 la1;2(1;0;0)\u0010\n1p\n3;1p\n3;1p\n3\u0011 \u0010\n1p\n2;1p\n2;0\u0011\n1\n3\naaakˆxxx,ccckˆzzz la2;2(0;0;1) ( 0;0;1) ( 1;0;0) 1\nbbb=\u0010\n\u0000a\n2;p\n3a\n2;0\u0011\nlg;2(1;0;0) ( 1;0;0) ( 0;1;0) 1\na=b6=c le;2 (a;0;c)p\na2+c2\u0010\n1p\n2;0;1p\n2\u0011 \u0010\n\u00001p\n2;0;1p\n2\u0011\n2ac\na2+c2\nTrigonal (I) Eq.35 la1;2(1;0;0) ( 0;0;1)\u0010\n1p\n2;1p\n2;0\u0011\n1\naaakˆxxx,ccckˆzzz la2;2(0;0;1) ( 0;0;1) ( 1;0;0) 1\nbbb=\u0010\n\u0000a\n2;p\n3a\n2;0\u0011\nlg;1(1;0;0) ( 1;0;0) ( 0;1;0) 1\na=b6=c lg;2 (a;0;c)p\na2+c2\u0010\n1p\n2;0;1p\n2\u0011 \u0010\n1p\n2;0;\u00001p\n2\u0011\nac\na2+c2\nl12(a;0;c)p\na2+c2\u0010\n0;1p\n2;1p\n2\u0011 \u0010\n0;1p\n2;\u00001p\n2\u0011\na2\n2(a2+c2)\nl21(a;0;c)p\na2+c2\u0010\n1p\n2;1p\n2;0\u0011 \u0010\n1p\n2;\u00001p\n2;0\u0011\nac\na2+c2\nTetragonal (I) Eq.40 la1;2(1;0;0)\u0010\n1p\n3;1p\n3;1p\n3\u0011 \u0010\n1p\n2;1p\n2;0\u0011\n1\n3\naaakˆxxx,bbbkˆyyy,ccckˆzzz la2;2(0;0;1) ( 0;0;1) ( 1;0;0) 1\na=b6=c lg;2(1;0;0) ( 1;0;0) ( 0;1;0) 1\nle;2 (a;0;c)p\na2+c2\u0010\n1p\n2;0;1p\n2\u0011 \u0010\n\u00001p\n2;0;1p\n2\u0011\n2ac\na2+c2\nld;2\u0010\n1p\n2;1p\n2;0\u0011 \u0010\n1p\n2;1p\n2;0\u0011 \u0010\n\u00001p\n2;1p\n2;0\u0011\n1\nOrthorhombic Eq.44 l1 (1;0;0) ( 1;0;0) ( 0;0;1) 1\naaakˆxxx,bbbkˆyyy,ccckˆzzz l2 (1;0;0) ( 0;1;0) ( 0;0;1) 1\nc0 for all calculations up to 93150 k-points, while the experimental\nvalue is\u00002:7µeV/atom [80]. This deviation may be due to the fact that we have\nnot used a sufficiently large number of k-points, as Halilov et al. pointed out\n[80, 81]. Our results are in good agreement with the calculations performed by\nTrygg el at. where a similar number of k-points were used [82]. We also see\nthatE(1;1;1)\u0000E(0;0;1)is approaching to negative values as the number of k-\npoints is increased. Interestingly, the calculated magnetostrictive coefficients are\nin quite good agreement with the experimental ones [40] despite the deviation\nof MAE for the unstrained unit cell. One possible reason for this result may be\nthat the calculation of the magnetostrictive coefficients involves larger energy dif-\nference between magnetization directions than the determination of MAE for the\nunstrained unit cell (which might be close to the accuracy limit of V ASP \u0018µeV),\nsee Fig.4. Consequently, a k-point mesh with about 105k-points may be sufficient\nto obtain reliable magnetostrictive coefficients for FCC Ni using GGA, although\na much more dense k-mesh would be needed to obtain a reliable MAE for the\nunstrained unit cell [80, 81]. Note that these two properties come from the SOC,\nso that in general it would be highly desirable that the used method to calculate\nthe energies describes well both MAE and magnetostriction. Recently, we devel-\noped a spin-lattice model within the framework of coupled spin and molecular\ndynamics (SD-MD) that reproduces accurately the experimental elastic and mag-\nnetoelastic energies at zero-temperature [59]. We obtained very good results by\napplying MAELAS to this coarse-grained model of SOC, see Tables 3 and 4.\n4.1.2. Elastic and magnetoelastic constants\nTo compute the elastic constants we make use of AELAS code [44]. As inputs,\nwe use the same relaxed cell and V ASP settings as in the calculation of magne-\ntostriction, but with lower number of k-points R k=60 (17\u000217\u000217 for the not\ndistorted cell) and not including SOC. Once we have the elastic constants, we\nuse them as inputs to derive the magnetoelastic constants with MAELAS. The re-\nsults are shown in Table 4, where we also include calculations of elastic constants\navailable in the Materials Project database [67, 68] and experimental data [75].\nWe observe that the value of C11=298 GPa obtained with GGA is higher than\nthe one in the Materials Project C11=276 GPa and in the experiment C11=261\nGPa. Concerning the magnetoelastic constants, we see that both b1andb2are in\nfairly good agreement with the experiment.\n28Figure 3: Calculation of (left) MAE of the unstrained unit cell and (right) magnetostrictive coeffi-\ncients for FCC Ni as a function of k-points.\nFigure 4: Calculation of l001for FCC Ni using MAELAS. (Left) Quadratic curve fit to the energy\nversus cell length along bbb= (0;0;1)with spin direction aaa1= (0;0;1). (Right) Energy difference\nbetween states with spin directions aaa2= (1;0;0)andaaa1= (0;0;1)against the cell length along\nbbb= (0;0;1).\n4.2. BCC Fe\nIn this example, we consider BCC Fe, which is described by Eq.17 since it is\na cubic (I) system.\n4.2.1. Cell relaxation, MAE and magnetostrictive coefficients\nIn the first stage, we we perform a cell relaxation for the conventional cubic\nunit cell of the BCC (2 atoms/cell) using a 57 \u000257\u000257 k-mesh with 185193\nk-points in the Brillouin zone. The interactions were described by a PAW po-\ntential with 14 valence electrons within the PBE approximation to the exchange-\ncorrelation, and the PW basis was generated for an energy cut-off of 380 eV (30%\n29larger than the default value). The relaxed lattice parameter is a=2:82509 Å.\nIn Fig. 5 we show the dependence of MAE and magnetostriction on the k-points\nfor this relaxed cell using the same exchange-correlation and energy cut-off as\nin the cell relaxation. The calculated values of MAE with the largest number of\nk-points (262144) are E(110)\u0000E(001) =0:32µeV/atom and E(111)\u0000E(001) =\n0:24µeV/atom which are a bit lower than the experimental values 1 µeV/atom and\n1:34µeV/atom, respectively [62]. Concerning the magnetostrictive coefficients,\nwe obtained l001=25:7\u000210\u00006andl111=17:2\u000210\u00006, while the experimental\nvalues at T=4:2K are l001=26\u000210\u00006andl111=\u000030\u000210\u00006[40]. We see\nthatl001is quite close to the experimental result, while l111is in good agreement\nwith previous DFT calculations [21, 25, 83] but it has the opposite sign as the\nexperimental value. The calculation of l111is presented in Fig.6. We observe\nthat the sign of the derivative of the energy difference between states with spin di-\nrections aaa2=\u0010\n1p\n2;0;\u00001p\n2\u0011\nandaaa1=\u0010\n1p\n3;1p\n3;1p\n3\u0011\nwith respect to the cell length\nalong bbb=\u0010\n1p\n3;1p\n3;1p\n3\u0011\nevaluated at l=l2is equal to the sign of the calculated\nl111(>0), as expected from Eq.52. However, the experimental l111is negative.\nThis deviation might be due to a possible failure of GGA related to the location of\nthe Fermi level in a region of majority band t2g density of states [84, 85]. Aim-\ning to clarify the influence of MAELAS in this result, we applied MAELAS to a\nspin-lattice model for BCC Fe, that reproduces accurately the experimental elastic\nand magnetoelastic energies, obtaining almost the same experimentally observed\nmagnetostriction [59], see Tables 3 and 4.\nFigure 5: Calculation of (left) MAE of the unstrained unit cell and (right) magnetostrictive coeffi-\ncients for BCC Fe as a function of k-points.\n30Figure 6: Calculation of l111for BCC Fe using MAELAS. (Left) Quadratic curve fit to the energy\nversus cell length along bbb=\u0010\n1p\n3;1p\n3;1p\n3\u0011\nwith spin direction aaa1=\u0010\n1p\n3;1p\n3;1p\n3\u0011\n. (Right) Energy\ndifference between states with spin directions aaa2=\u0010\n1p\n2;0;\u00001p\n2\u0011\nandaaa1=\u0010\n1p\n3;1p\n3;1p\n3\u0011\nagainst\nthe cell length along bbb=\u0010\n1p\n3;1p\n3;1p\n3\u0011\n.\n4.2.2. Elastic and magnetoelastic constants\nAs inputs for AELAS code [44], we use the same relaxed cell and V ASP set-\ntings as in the calculation of magnetostriction, but with lower number of k-points\nRk=60 (21\u000221\u000221 for the not distorted cell) and not including SOC. Once\nwe have the elastic constants, we use them as inputs to derive the magnetoelastic\nconstants with MAELAS. The results are shown in Table 4, where we also in-\nclude calculations of elastic constants available in the Materials Project database\n[67, 68] and experimental data [70]. We observe that the value of C11=288 GPa\nobtained with AELAS is significantly higher than the one in the Materials Project\nC11=247 GPa and in the experiment C11=243 GPa. Regarding the magnetoelas-\ntic constants, we see that the value for b1=\u00005:2 MPa generated with MAELAS\nis close to the estimated experimental value b1=\u00004:1 MPa. However, we obtain\na negative sign for b2=\u00005:4 MPa, while in the experiment it is positive b2=10:9\nMPa. This deviation is due to the positive sign of l111given by DFT that we have\nmentioned above, see Eq.18 [25, 84, 85].\n4.3. HCP Co\nAs a first example of hexagonal (I) system, we consider HCP Co.\n4.3.1. Cell relaxation, MAE and magnetostrictive coefficients\nFor this material, we set the length parameter R k=160 for the generation\nof the automatic k-point mesh, which for the relaxed (not distorted) cell, results\n31in a 75\u000275\u000240 k-point grid with 250000 points in the Brillouin zone. All cal-\nculations were done with an energy cut-off 406 :563 eV (50% larger than the\ndefault one), 15 electrons in the valence states, and the meta-GGA functional\nSCAN [86], with aspherical contributions to the PAW one-centre terms. The\nrelaxed lattice parameters are a=b=2:4561 Å and c=3:9821 Å. The calcu-\nlated MAE for the relaxed cell is E(100)\u0000E(001) =53µeV/atom which is quite\nclose to the experimental value 61 µeV/atom [62]. As it is shown in Table 3, the\ncalculated magnetostrictive coefficients are also close to the experimental ones,\nexcept for le;2. Similarly, converting them into Mason’s definitions via the rela-\ntions given by Eq.A.2, we see that only lD=\u00001\u000210\u00006is significantly deviated\nfrom the experiment ( \u0000128\u000210\u00006) [63]. Aiming to clarify this result, we per-\nformed a direct calculation of lDusing bbb=\u0010\n1p\n2;0;1p\n2\u0011\n,aaa1=\u0010\n1p\n2;0;1p\n2\u0011\nand\naaa2= (0;0;1)finding lD=\u00009\u000210\u00006, which is consistent with the indirect cal-\nculation through Clark’s definition but still far from the experimental value. Fig.7\nshows the quadratic curve fit to the energy versus cell length along bbb= (1;0;0)\nwithaaa1=\u0010\n1p\n3;1p\n3;1p\n3\u0011\nto calculate la1;2.\nFigure 7: Calculation of la1;2for HCP Co using MAELAS with the meta-GGA functional\nSCAN. (Left) Quadratic curve fit to the energy versus cell length along bbb= (1;0;0)with spin\ndirection aaa1=\u0010\n1p\n3;1p\n3;1p\n3\u0011\n. (Right) Energy difference between states with spin directions\naaa2=\u0010\n1p\n2;1p\n2;0\u0011\nandaaa1=\u0010\n1p\n3;1p\n3;1p\n3\u0011\nagainst the cell length along bbb= (1;0;0).\nWe have also performed a second test using the rotationally invariant LSDA+U\napproach introduced by Liechtenstein et al. [87] fixing J=0:8eV and varying U\non the d-electrons [64]. In all calculations we use the same pseudopotential and\n32number of k-points as in the tests performed with SCAN. The energy cut-off is set\nto 380 eV . In this case the relaxed lattice parameters are a=b=2:48896 Å and\nc=4:02347 Å. The analysis of MAE and magnetostriction for different values of\nUis shown in Fig.8. We see that MAE approximates the experimental value at\nU=3eV , so that we might expect a reliable description of SOC for this value of\nU. Increasing Uup to 3eV has a significant effect on la1;2andla2;2making them\nto approach the experimental values. On the other hand, lg;2andle;2don’t change\ntoo much within the range of values used for U. The sign of all magnetostrictive\ncoefficients are in good agreement to the experimental ones. However, as in the\ncase with SCAN, le;2is significantly underestimated. Possible reasons for this\nsystematic deviation might be a failure of DFT [85], the applied deformations\n(we used the default value 0 :01 for the tag\u0000sthat sets the maximum value of\nparameter sin the generation of the deformed unit cells, see Eq.C.4), the used\nV ASP settings (k-point mesh, exchange-correlation functional, smearing method,\nlattice parameters, ...) or higher order corrections in the equation of the relative\nlength change Eq.25 [88]. This issue should be further investigated to clarify its\npossible causes.\nFigure 8: Calculation of (left) MAE of the unstrained unit cell and (right) magnetostrictive coeffi-\ncients for HCP Co using the LSDA+U approach with different values of parameter U.\n4.3.2. Elastic and magnetoelastic constants\nFor the calculation of the elastic constants we use the same relaxed cell and\nV ASP settings as in the calculation of magnetostriction, but without SOC and\nlower number of k-points R k=60 (28\u000228\u000215 for the not distorted cell). In\naddition to SCAN, we also run calculations with LSDA+U setting J=0:8eV\n33andU=3eV . In Table 4, we see that LSDA+U and GGA (Materials Project\ndatabase [74]) give better results than SCAN for both the elastic and magnetoe-\nlastic constants. The magnetoelastic constants obtained with LSDA+U are mod-\nerately good, except for b3andb4which are one order of magnitude lower than in\nthe experiment, mainly due to the deviations coming from lg;2andle;2given by\nMAELAS, see Table 3.\n4.4. YCo 5\nIn this example we study the hexagonal (I) system YCo 5with prototype CaCu 5\nstructure (space group 191).\n4.4.1. Cell relaxation, MAE and magnetostrictive coefficients\nWe use the simplified (rotationally invariant) approach to the LSDA+U intro-\nduced by Dudarev et al. [89] with parameters U=1:9 eV and J=0:8 eV for\nCo, and U=J=0 eV for Y given in Ref.[64]. For the calculation of the re-\nlaxed cell, MAE and magnetostrictive coefficients we used an automatic k-point\nmesh with length parameter R k=100 centered on the G-point (23\u000223\u000225 for\nthe not distorted cell), 11 and 9 valence states for Y and Co, respectively, and en-\nergy cut-off 375 eV . The cell relaxation leads to lattice parameters a=b=4:9253\nÅ and c=3:9269 Å. The calculated MAE is E(100)\u0000E(001) =365µeV/atom\nwhich is lower than the experimental value 567 µeV/atom [64]. Andreev measured\nthe magnetostriction along a and c axis, finding that the magnitude of jla1;2jand\njla2;2jcan not be greater than 10\u00004[7]. We obtained la1;2=\u000090\u000210\u00006and\nla2;2=115\u000210\u00006which are quite close to the experimental upper limit. In\nFig.9 we present the quadratic curve fit to the energy versus cell length along\nbbb= (0;0;1)withaaa1= (0;0;1)to calculate la2;2.\n4.4.2. Elastic and magnetoelastic constants\nThe calculation of the elastic constants is performed using the same relaxed\ncell and V ASP settings as for magnetostriction, but without SOC and lower num-\nber of k-points R k=60 (14\u000214\u000215 for the not distorted cell). In addition\nto LSDA+U, we also run calculations with GGA. In Table 4, we observe that\nLSDA+U leads to an unstable phase ( C11\u0000C12<0), while GGA gives better\nresults.\n4.5. Fe 2Si\nTo illustrate the application of MAELAS to trigonal (I) systems, we apply it\nto Fe 2Si (space group 164) [90].\n34Figure 9: Calculation of la2;2for YCo 5using MAELAS. (Left) Quadratic curve fit to the energy\nversus cell length along bbb= (0;0;1)with spin direction aaa1= (0;0;1). (Right) Energy difference\nbetween states with spin directions aaa2= (1;0;0)andaaa1= (0;0;1)against the cell length along\nbbb= (0;0;1).\n4.5.1. Cell relaxation, MAE and magnetostrictive coefficients\nFor the calculation of the cell relaxation, MAE and magnetostrictive coeffi-\ncients we used an automatic k-point mesh with length parameter R k=80 centered\non the G-point (24\u000224\u000217 for the not distorted cell), 14 and 4 valence states for\nFe and Si, respectively, and energy cut-off 520 eV with PAW method and GGA-\nPBE. The relaxed lattice parameters are a=3:9249 Å and c=4:8311 Å. The\ncalculated MAE is E(100)\u0000E(001) =\u000038µeV/atom (easy plane). Sun et al. re-\nported MAE values with the screened hybrid Heyd-Scuseria-Ernzerhof (HSE06)\nfunctional smaller than with PBE for 2D Fe 2Si [91, 92]. Chi Pui Tang et al. calcu-\nlated some electronic properties for bulk Fe 2Si finding that the densities of states\nin the vicinity of the Fermi level is mainly contributed from the d-electrons of Fe\n[93]. In Table 3, we observe that the overall anisotropic magnetostriction given\nby MAELAS is rather small, which makes this material interesting for high-flux\ncore applications because it can reduce hysteresis loss [94].\n4.5.2. Elastic and magnetoelastic constants\nAs inputs for AELAS, we use the same relaxed cell and V ASP settings as in the\ncalculation of magnetostriction, but without SOC and lower number of k-points\nRk=60 (18\u000218\u000212 for the not distorted cell). In Table 5, we see that AELAS\ngives similar elastic constants as in the Materials Project [76]. The derived mag-\nnetoelastic constants are small which is consistent with the low magnetostrictive\ncoefficients that we obtained previously.\n35Figure 10: Calculation of la2;2for Fe 2Si using MAELAS. (Left) Quadratic curve fit to the energy\nversus cell length along bbb= (0;0;1)with spin direction aaa1= (0;0;1). (Right) Energy difference\nbetween states with spin directions aaa2= (1;0;0)andaaa1= (0;0;1)against the cell length along\nbbb= (0;0;1).\n4.6. L1 0FePd\nAs an example of tetragonal (I) system, we calculate the anisotropic magne-\ntostrictive coefficients of L1 0FePd (space group 123).\n4.6.1. Cell relaxation, MAE and magnetostrictive coefficients\nFor the calculation of the cell relaxation, MAE and magnetostrictive coeffi-\ncients we used an automatic k-point mesh with length parameter R k=100 cen-\ntered on the G-point (37\u000237\u000227 for the not distorted cell), 8 and 10 valence\nstates for Fe and Pd, respectively, and energy cut-off 375 eV with PAW method\nand GGA-PBE. The relaxed lattice parameters are a=2:6973 Å and c=3:7593\nÅ. We obtained a MAE E(100)\u0000E(001) =106µeV/atom which is lower than in\nthe experiment 181 µeV/atom [65]. The values of the obtained anisotropic mag-\nnetostrictive coefficients are shown in Table 3. Shima et al. reported a relative\nlength change equal to 100 \u000210\u00006along a-axis under a magnetic field in the same\ndirection ( bbb=aaa= (1;0;0)) [66]. According to Eq.40, this measurement corre-\nsponds to la1;0\u0000la1;2\n3+lg;2\n2. For the anisotropic part of this quantity, we obtained\n\u0000la1;2\n3+lg;2\n2=22:5\u000210\u00006. In Fig.11 we show the quadratic curve fit to the en-\nergy versus cell length along bbb= (1;0;0)withaaa1= (1;0;0)to calculate lg;2.\n36Figure 11: Calculation of lg;2for L1 0FePd using MAELAS. (Left) Quadratic curve fit to the\nenergy versus cell length along bbb= (1;0;0)with spin direction aaa1= (1;0;0). (Right) Energy\ndifference between states with spin directions aaa2= (0;1;0)andaaa1= (1;0;0)versus the cell\nlength along bbb= (1;0;0).\n4.6.2. Elastic and magnetoelastic constants\nThe calculation of the elastic constants with AELAS is performed using the\nsame relaxed cell and V ASP settings as for magnetostriction, but without SOC\nand lower number of k-points R k=60 (22\u000222\u000216 for the not distorted cell). As\nwe see in Table 5, we obtain similar results as in the Materials Project database\n[77].\n4.7. YCo\nFor the case of orthorhombic systems, we study the compound YCo (space\ngroup 63) [95].\n4.7.1. Cell relaxation, MAE and magnetostrictive coefficients\nIn this case, we use the simplified (rotationally invariant) approach to the\nLSDA+U [89] with parameters U=1:9 eV and J=0:8 eV for Co, and U=J=0\neV for Y in the same way as in YCo 5[64]. For the calculation of the relaxed cell,\nMAE and magnetostrictive coefficients we used an automatic k-point mesh with\nlength parameter R k=90 centered on the G-point (22\u00029\u000223 for the not distorted\ncell), 11 and 9 valence states for Y and Co, respectively, and energy cut-off 375\neV . The cell relaxation leads to lattice parameters a=4:0686 Å, b=10:3157 Å\nandc=3:8957 Å. As we see in Table 3, both MAE and magnetostriction are quite\nsmall for this material.\n37Figure 12: Calculation of l5for YCo using MAELAS. (Left) Quadratic curve fit to the energy\nversus cell length along bbb= (0;0;1)with spin direction aaa1= (1;0;0). (Right) Energy difference\nbetween states with spin directions aaa2= (0;0;1)andaaa1= (1;0;0)against the cell length along\nbbb= (0;0;1).\n4.7.2. Elastic and magnetoelastic constants\nThe calculation of the elastic constants is performed using the same relaxed\ncell and V ASP settings as for magnetostriction, but without SOC and lower num-\nber of k-points R k=60 (15\u00026\u000215 for the not distorted cell). In addition to\nLSDA+U, we also run calculations with GGA. In Table 5, we observe that both\nLSDA+U and GGA lead to similar elastic and magnetoelastic constants.\n5. Conclusions and future perspectives\nIn summary, the program MAELAS offers computational tools to tackle the\ncomplex phenomenon of magnetostriction by automated first-principles calcula-\ntions. It could potentially be used to discover and design novel magnetostric-\ntive materials by a high-throughput screening approach. In particular, materials\nwith giant magnetostriction (beyond conventional cubic and hexagonal systems),\nisotropic or very low magnetostriction (like FeNi alloys) might be of technological\nimportance.\nThe preliminary tests of the program show quite encouraging results, although\nthere is still room for improvement. First principle calculations are still quite chal-\nlenging for materials with very low MAE or localized 4f-electrons [64]. In this\nsense, MAELAS could also be a useful tool to understand, test and improve the\nDFT methods to compute induced properties by SOC and crystal field interactions\nlike the MAE of unstrained systems and anisotropic magnetostriction.\n38Presently, we are working on new features of MAELAS and online visual-\nization tools [51]. We are also considering to increase the number of supported\ncrystal systems, as well as to implement more computationally efficient methods\nto calculate magnetoelastic constants and magnetostrictive coefficients. These ex-\ntensions might be included in new versions of the code.\nAcknowledgement\nThis work was supported by the ERDF in the IT4Innovations national super-\ncomputing center - path to exascale project (CZ.02.1.01/0.0/0.0/16-013/0001791)\nwithin the OPRDE. This work was supported by The Ministry of Education, Youth\nand Sports from the Large Infrastructures for Research, Experimental Develop-\nment, and Innovations project “e-INFRA CZ - LM2018140”. This work was sup-\nported by the Donau project No. 8X20050 and the computational resources pro-\nvided by the Open Access Grant Competition of IT4Innovations National Super-\ncomputing Center within the projects OPEN-18-5, OPEN-18-33, and OPEN-19-\n14. DL, SA, and APK acknowledge the Czech Science Foundations grant No. 20-\n18392S. P.N., D.L., and S.A. acknowledge support from the H2020-FETOPEN\nno. 863155 s-NEBULA project.\nAppendix A. Conversion between different definitions of magnetostrictive\ncoefficients for hexagonal (I)\nThe magnetostrictive coefficients for hexagonal (I) shown in Eq.25 were de-\nfined by Clark et al. in 1965 [16]. However, one can find other definitions like\nthose given by Mason [15], Birss [34], and Callen and Callen [17]. In this ap-\npendix, we show the conversion formulas between these definitions and those\nprovided by Clark et al. [16] (Eq.25).\nAppendix A.1. Mason’s form\nIn 1954, based on a general thermodynamic function with stresses and in-\ntensity of magnetization as the fundamental variables [96], Mason derived the\n39following form of the relative length change [15]\nDl\nl0\f\f\f\f\faaa\nbbb=la1;0\nMason(b2\nx+b2\ny)+la2;0\nMasonb2\nz+lA[(axbx+ayby)2\u0000(axbx+ayby)azbz]\n+lB[(1\u0000a2\nz)(1\u0000b2\nz)\u0000(axbx+ayby)2]\n+lC[(1\u0000a2\nz)b2\nz\u0000(axbx+ayby)azbz]+4lD(axbx+ayby)azbz:\n(A.1)\nThese magnetostrictive coefficients are related to those defined in Eq.25 as [16]\nla1;0\nMason =la1;0+2\n3la1;2\nla2;0\nMason =la2;0+2\n3la2;2\nlA=\u0000la1;2+1\n2lg;2\nlB=\u0000la1;2\u00001\n2lg;2\nlC=\u0000la2;2\nlD=1\n2le;2\u00001\n4la1;2+1\n8lg;2\u00001\n4la2;2:(A.2)\nNote in the original work of Mason [15] the terms that describes the volume mag-\nnetostriction were not included. Here we added these terms ( la1;0\nMason ,la2;0\nMason ) in\norder to fully recover the Eq.25.\nAppendix A.2. Birss’s form\nIn 1959 Birss derived an equivalent equation of relative length change in this\nform [34]\nDl\nl0\f\f\f\f\faaa\nbbb=Q0+Q1b2\nz+Q2(1\u0000a2\nz)+Q4(1\u0000a2\nz)b2\nz+Q6(axbx+ayby)azbz\n+Q8(axbx+ayby)2:\n(A.3)\n40These magnetostrictive coefficients are related to those defined in Eq.25 as [16]\nQ0=la1;0+2\n3la1;2\nQ1=la2;0+2\n3la2;2\u0000la1;0\u00002\n3la1;2\nQ2=\u0000la1;2\u00001\n2lg;2\nQ4=la1;2+1\n2lg;2\u0000la2;2\nQ6=2le;2\nQ8=lg;2:(A.4)\nAppendix A.3. Callen and Callen’s form\nIn 1965 Callen and Callen obtained other equivalent form of the equation of\nrelative length change by including two-ion interactions into the theory of magne-\ntostriction arising from single-ion crystal-field effects [17]. It reads\nDl\nl0\f\f\f\f\faaa\nbbb=1\n3la\n11+1\n2p\n3la\n12\u0012\na2\nz\u00001\n3\u0013\n+2la\n21\u0012\nb2\nz\u00001\n3\u0013\n+p\n3la\n22\u0012\na2\nz\u00001\n3\u0013\u0012\nb2\nz\u00001\n3\u0013\n+lg\u00141\n2(a2\nx\u0000a2\ny)(b2\nx\u0000b2\ny)+2axaybxby\u0015\n+2le(axazbxbz+ayazbybz);\n(A.5)\nThese magnetostrictive coefficients are related to those defined in Eq.25 as [17]\nla\n11=2la1;0+la2;0+2la1;2+la2;2\nla\n12=4p\n3la1;2+2p\n3la2;2\nla\n21=\u00001\n2la1;0+1\n2la2;0\nla\n22=\u00001p\n3la1;2+1p\n3la2;2\nle=le;2\nlg=lg;2:(A.6)\n41Appendix B. Conversion between different definitions of magnetostrictive\ncoefficients for tetragonal (I)\nIn 1994 Cullen et al. [8] derived the equation of relative length change given\nby Eq.40 for tetragonal (I) system. In 1954 Mason obtained an equivalent equation\nthat reads [15]\nDl\nl0\f\f\f\f\faaa\nbbb=la1;0\nMason(b2\nx+b2\ny)+la2;0\nMasonb2\nz+1\n2l1[(axbx\u0000ayby)2\u0000(axby+aybx)2\n+(1\u0000b2\nz)(1\u0000a2\nz)\u00002azbz(axbx+ayby)]+4l2azbz(axbx+ayby)\n+4l3axaybxby+l4[b2\nz(1\u0000a2\nz)\u0000azbz(axbx+ayby)]\n+1\n2l5[(axby\u0000aybx)2\u0000(axbx+ayby)2+(1\u0000b2\nz)(1\u0000a2\nz)]:\n(B.1)\nThese magnetostrictive coefficients are related to those defined by Eq.40 in the\nfollowing way\nla1;0\nMason =la1;0+2\n3la1;2\nla2;0\nMason =la2;0+2\n3la2;2\nl1=\u0000la1;2+1\n2lg;2\nl2=1\n2le;2\u00001\n4la2;2\u00001\n4la1;2+1\n8lg;2\nl3=1\n2ld;2\u0000la1;2\nl4=\u0000la2;2\nl5=\u0000la1;2\u00001\n2lg;2:(B.2)\nNote in the original work of Mason [15] the terms that describes the volume mag-\nnetostriction were not included. Here we added these terms ( la1;0\nMason ,la2;0\nMason ) in\norder to fully recover the Eq.40.\nAppendix C. Generation of the deformed unit cells\nIn this appendix we present the procedure to generate the deformed unit cells\nfor the calculation of each magnetostrictive coefficient. The deformed unit cells\n42are generated by multiplying the lattice vectors of the initial unit cell aaa= (ax;ay;az),\nbbb= (bx;by;bz),ccc= (cx;cy;cz)by the deformation gradient Fi j[97]\n0\n@a0\nxb0\nxc0\nx\na0\nyb0\nyc0\ny\na0\nzb0\nzc0\nz1\nA=0\n@FxxFxyFxz\nFyxFyyFyz\nFzxFzyFzz1\nA\u00010\n@axbxcx\naybycy\nazbzcz1\nA (C.1)\nwhere a0\ni,b0\niandc0\ni(i=x;y;z) are the components of the lattice vectors of the\ndeformed cell. In the infinitesimal strain theory, the deformation gradient is re-\nlated to the displacement gradient ( ¶ui=¶rj) asFi j=di j+¶ui=¶rj, where di jis the\nKronecker delta. Hence, according to Eq.5, the strain tensor ei jcan be written in\nterms of the deformation gradient as\neee=0\n@exxexyexz\neyxeyyeyz\nezxezyezz1\nA=1\n20\n@2(Fxx\u00001)Fxy+Fyx Fxz+Fzx\nFxy+Fyx2(Fyy\u00001)Fyz+Fzy\nFxz+Fzx Fyz+Fzy2(Fzz\u00001)1\nA: (C.2)\nIn MAELAS, we consider deformation gradients to optimize the unit cell in the\nmeasuring directions bbbgiven by Table 2. Additionally, we also constrain the\ndeterminant of the deformation gradients to be equal to one ( det(FFF) =1) in order\nto preserve the volume of the unit cells (isochoric deformation) [21, 98]. We point\nout that there are other possible variants of the following deformation modes [85].\nAppendix C.1. Cubic (I) system\nFor cubic (I) systems MAELAS generates two set of deformed unit cells with\ntetragonal deformations along bbb= (0;0;1)and trigonal deformations along bbb=\n(1=p\n3;1=p\n3;1=p\n3)to calculate l001andl111, respectively (see Table 2). The\ndeformation gradients for these two deformation modes are\nFFFl001\nbbb=(0;0;1)(s) =0\nB@1p1+s0 0\n01p1+s0\n0 0 1 +s1\nCA;FFFl111\nbbb=\u0010\n1p\n3;1p\n3;1p\n3\u0011(s) =z0\n@1s\n2s\n2s\n21s\n2s\n2s\n211\nA\n(C.3)\nwhere z=3p\n4=(4\u00003s2+s3). The parameter scontrols the applied deformation,\nand its maximum value can be specified through the command line of the program\nMAELAS using tag \u0000s. The total number of deformed cells can be chosen with\ntag\u0000n.\n43Appendix C.2. Hexagonal (I) system\nIn the case of hexagonal (I), MAELAS generates 4 sets of deformed cells using\nthe following deformation gradients\nFFF\f\f\fla1;2\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA;FFF\f\f\fla2;2\nbbb=(0;0;1)(s) =0\nB@1p1+s0 0\n01p1+s0\n0 0 1 +s1\nCA\nFFF\f\f\flg;2\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA;FFF\f\f\fle;2\nbbb=(a;0;c)p\na2+c2(s) =w0\n@1 0sc\n2a\n0 1 0\nsa\n2c0 11\nA\n(C.4)\nwhere w=3p\n4=(4\u0000s2),aandcare the lattice parameters of the relaxed (not\ndeformed) unit cell. The fractions c=aanda=cwere introduced in the deformation\ngradient elements Fle;2\nxzandFle;2\nzx, respectively, in order to generate deformations\nthat meet the property bbb=aaa+ccc\njaaa+cccj=aaa000+ccc000\njaaa000+ccc000j, where aaa000andccc000are the lattice vectors\nof the distorted unit cell, see Fig. C.13. This deformation mode makes it easy\nto compute the cell length lin the measuring direction bbbsince it is just l=jaaa000+\nccc000j. It is inspired by the trigonal deformation for the cubic (I) case FFFl111where\ndeformations meet the property bbb=aaa+bbb+ccc\njaaa+bbb+cccj=aaa000+bbb000+ccc000\njaaa000+bbb000+ccc000j.\nFigure C.13: Sketch of the deformation generated by the deformation gradient FFFle;2given by\nEq.C.4 to calculate le;2. The purple line represents the relaxed cell with lattice parameters (a;b;c),\nwhile the red line stands for the deformed cell with lattice parameters (a0;b0;c0). This deformation\nmeets the property bbb=aaa+ccc\njaaa+cccj=aaa000+ccc000\njaaa000+ccc000j.\n44Appendix C.3. Trigonal (I) system\nIn the case of trigonal (I), MAELAS generates 6 sets of deformed unit cells\nusing the following deformation gradients\nFFF\f\f\fla1;2\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA;FFF\f\f\fla2;2\nbbb=(0;0;1)(s) =0\nB@1p1+s0 0\n01p1+s0\n0 0 1 +s1\nCA\nFFF\f\f\flg;1\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA\nFFF\f\f\flg;2\nbbb=(a;0;c)p\na2+c2(s) =FFF\f\f\fl12\nbbb=(a;0;c)p\na2+c2(s) =FFF\f\f\fl21\nbbb=(a;0;c)p\na2+c2(s) =w0\n@1 0sc\n2a\n0 1 0\nsa\n2c0 11\nA:\n(C.5)\nAppendix C.4. Tetragonal (I) system\nIn the case of tetragonal (I), MAELAS generates 5 sets of deformed cells using\nthe following deformation gradients\nFFF\f\f\fla1;2\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA;FFF\f\f\fla2;2\nbbb=(0;0;1)(s) =0\nB@1p1+s0 0\n01p1+s0\n0 0 1 +s1\nCA\nFFF\f\f\flg;2\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA;FFF\f\f\fle;2\nbbb=(a;0;c)p\na2+c2(s) =w0\n@1 0sc\n2a\n0 1 0\nsa\n2c0 11\nA\nFFF\f\f\fld;2\nbbb=\u0010\n1p\n2;1p\n2;0\u0011(s) =w0\n@1s\n20\ns\n21 0\n0 0 11\nA:\n(C.6)\n45Appendix C.5. Orthorhombic system\nFor orthorhombic crystals MAELAS generates 9 sets of deformed cells using\nthe following deformation gradients\nFFF\f\f\fl1\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA;FFF\f\f\fl2\nbbb=(1;0;0)(s) =0\nB@1+s 0 0\n01p1+s0\n0 01p1+s1\nCA\nFFF\f\f\fl3\nbbb=(0;1;0)(s) =0\nB@1p1+s0 0\n0 1 +s 0\n0 01p1+s1\nCA;FFF\f\f\fl4\nbbb=(0;1;0)(s) =0\nB@1p1+s0 0\n0 1 +s 0\n0 01p1+s1\nCA\nFFF\f\f\fl5\nbbb=(0;0;1)(s) =0\nB@1p1+s0 0\n01p1+s0\n0 0 1 +s1\nCA;FFF\f\f\fl6\nbbb=(0;0;1)(s) =0\nB@1p1+s0 0\n01p1+s0\n0 0 1 +s1\nCA\nFFF\f\f\fl7\nbbb=(a;b;0)p\na2+b2(s) =w0\n@1sb\n2a0\nsa\n2b1 0\n0 0 11\nA;FFF\f\f\fl8\nbbb=(a;0;c)p\na2+c2(s) =w0\n@1 0sc\n2a\n0 1 0\nsa\n2c0 11\nA\nFFF\f\f\fl9\nbbb=(0;b;c)p\nb2+c2(s) =w0\n@1 0 0\n0 1sc\n2b\n0sb\n2c11\nA:\n(C.7)\nReferences\n[1] M. Gibbs, Modern trends in magnetostriction, Springer, 2001.\n[2] N. Ekreem, A. Olabi, T. Prescott, A. Rafferty, M. Hashmi, An overview of\nmagnetostriction, its use and methods to measure these properties, Journal of\nMaterials Processing Technology 191 (1) (2007) 96 – 101, advances in Ma-\nterials and Processing Technologies, July 30th - August 3rd 2006, Las Vegas,\nNevada. doi:https://doi.org/10.1016/j.jmatprotec.2007.03.064 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0924013607002889\n[3] F. T. Calkins, A. B. Flatau, M. J. Dapino, Overview of magnetostric-\ntive sensor technology, Journal of Intelligent Material Systems and Struc-\ntures 18 (10) (2007) 1057–1066. arXiv:https://doi.org/10.1177/\n461045389X06072358 ,doi:10.1177/1045389X06072358 .\nURL https://doi.org/10.1177/1045389X06072358\n[4] V . Apicella, C. S. Clemente, D. Davino, D. Leone, C. Visone, Review of\nmodeling and control of magnetostrictive actuators, Actuators 8 (2). doi:\n10.3390/act8020045 .\nURL https://www.mdpi.com/2076-0825/8/2/45\n[5] A. Clark, Chapter 7 magnetostrictive rare earth-Fe 2compounds, V ol. 1\nof Handbook of Ferromagnetic Materials, Elsevier, 1980, pp. 531 – 589.\ndoi:https://doi.org/10.1016/S1574-9304(05)80122-1 .\nURL http://www.sciencedirect.com/science/article/pii/\nS1574930405801221\n[6] G. Engdahl, Handbook of giant magnetostrictive materials, Academic Press,\n1999.\n[7] A. Andreev, Chapter 2 Thermal expansion anomalies and spontaneous\nmagnetostriction in rare-earth intermetallics with cobalt and iron, V ol. 8\nof Handbook of Magnetic Materials, Elsevier, 1995, pp. 59 – 187.\ndoi:https://doi.org/10.1016/S1567-2719(05)80031-9 .\nURL http://www.sciencedirect.com/science/article/pii/\nS1567271905800319\n[8] J. R. Cullen, A. E. Clark, K. B. Hathaway, Materials, science and technology,\nVCH Publishings, 1994, Ch. 16 - Magnetostrictive Materials, pp. 529–565.\n[9] S. Massari, M. Ruberti, Rare earth elements as critical raw materials: Focus\non international markets and future strategies, Resources Policy 38 (1)\n(2013) 36 – 43. doi:https://doi.org/10.1016/j.resourpol.2012.\n07.001 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0301420712000530\n[10] D. Fritsch, C. Ederer, First-principles calculation of magnetoelastic coef-\nficients and magnetostriction in the spinel ferrites CoFe 2O4and NiFe 2O4,\nPhys. Rev. B 86 (2012) 014406. doi:10.1103/PhysRevB.86.014406 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.86.014406\n[11] H. Wang, Y . N. Zhang, R. Q. Wu, L. Z. Sun, D. S. Xu, Z. D. Zhang, Under-\nstanding strong magnetostriction in Fe 100\u0000xGaxalloys, Scientific Reports\n473 (1) (2013) 3521. doi:10.1038/srep03521 .\nURL https://doi.org/10.1038/srep03521\n[12] M. Dapino, Encyclopedia of Smart Materials, John Wiley and Sons, Inc.,\nNew York, 2000.\n[13] N. S. Akulov, Z. Physik 52, 389.\n[14] R. Becker, W. Doring, Ferromagnetismus, Verlag. Julius Springer, Berlin,\n1939.\n[15] W. P. Mason, Derivation of magnetostriction and anisotropic energies for\nhexagonal, tetragonal, and orthorhombic crystals, Phys. Rev. 96 (1954) 302–\n310. doi:10.1103/PhysRev.96.302 .\nURL https://link.aps.org/doi/10.1103/PhysRev.96.302\n[16] A. E. Clark, B. F. DeSavage, R. Bozorth, Anomalous Thermal Expansion\nand Magnetostriction of Single-Crystal Dysprosium, Phys. Rev. 138 (1965)\nA216–A224. doi:10.1103/PhysRev.138.A216 .\nURL https://link.aps.org/doi/10.1103/PhysRev.138.A216\n[17] E. Callen, H. B. Callen, Magnetostriction, forced magnetostriction, and\nanomalous thermal expansion in ferromagnets, Phys. Rev. 139 (1965) A455–\nA471. doi:10.1103/PhysRev.139.A455 .\nURL https://link.aps.org/doi/10.1103/PhysRev.139.A455\n[18] R. Wu, A. J. Freeman, First principles determinations of magnetostriction\nin transition metals (invited), Journal of Applied Physics 79 (8) (1996)\n6209–6212. arXiv:https://aip.scitation.org/doi/pdf/10.1063/\n1.362073 ,doi:10.1063/1.362073 .\nURL https://aip.scitation.org/doi/abs/10.1063/1.362073\n[19] R. Wu, L. Chen, A. Freeman, First principles determination of\nmagnetostriction in bulk transition metals and thin films, Journal\nof Magnetism and Magnetic Materials 170 (1) (1997) 103 – 109.\ndoi:https://doi.org/10.1016/S0304-8853(97)00004-8 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0304885397000048\n[20] R. Wu, L. Chen, A. Shick, A. Freeman, First-principles determinations of\nmagneto-crystalline anisotropy and magnetostriction in bulk and thin-film\n48transition metals, Journal of Magnetism and Magnetic Materials 177-181\n(1998) 1216 – 1219, international Conference on Magnetism (Part II).\ndoi:https://doi.org/10.1016/S0304-8853(97)00382-X .\nURL http://www.sciencedirect.com/science/article/pii/\nS030488539700382X\n[21] T. Burkert, O. Eriksson, P. James, S. I. Simak, B. Johansson, L. Nordström,\nCalculation of uniaxial magnetic anisotropy energy of tetragonal and trigonal\nFe, Co, and Ni, Phys. Rev. B 69 (2004) 104426. doi:10.1103/PhysRevB.\n69.104426 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.69.104426\n[22] G. Petculescu, R. Wu, R. McQueeney, Chapter three - magne-\ntoelasticity of bcc Fe-Ga Alloys, V ol. 20 of Handbook of Mag-\nnetic Materials, Elsevier, 2012, pp. 123 – 226. doi:https:\n//doi.org/10.1016/B978-0-444-56371-2.00003-9 .\nURL http://www.sciencedirect.com/science/article/pii/\nB9780444563712000039\n[23] Y . N. Zhang, J. X. Cao, R. Q. Wu, Rigid band model for prediction of mag-\nnetostriction of iron-gallium alloys, Applied Physics Letters 96 (6) (2010)\n062508. arXiv:https://doi.org/10.1063/1.3318420 ,doi:10.1063/\n1.3318420 .\nURL https://doi.org/10.1063/1.3318420\n[24] Y . Zhang, R. Wu, Mechanism of large magnetostriction of galfenol, IEEE\nTransactions on Magnetics 47 (10) (2011) 4044–4049.\n[25] Y . Zhang, H. Wang, R. Wu, First-principles determination of the rhombohe-\ndral magnetostriction of Fe 100\u0000xAlxand Fe 100\u0000xGaxalloys, Phys. Rev. B 86\n(2012) 224410. doi:10.1103/PhysRevB.86.224410 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.86.224410\n[26] S. C. Hong, W. S. Yun, R. Wu, Giant magnetostriction of Fe 1\u0000xBexalloy:\nA first-principles study, Phys. Rev. B 79 (2009) 054419. doi:10.1103/\nPhysRevB.79.054419 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.79.054419\n[27] V . I. Gavrilenko, R. Q. Wu, Magnetostriction and magnetism of rare\nearth intermetallic compounds: First principle study, Journal of Applied\n49Physics 89 (11) (2001) 7320–7322. arXiv:https://doi.org/10.1063/\n1.1356038 ,doi:10.1063/1.1356038 .\nURL https://doi.org/10.1063/1.1356038\n[28] R. Wu, Magnetism and magnetostriction in GdFe 2and GdCo 2, Journal of\nApplied Physics 85 (8) (1999) 6217–6219. arXiv:https://doi.org/10.\n1063/1.370226 ,doi:10.1063/1.370226 .\nURL https://doi.org/10.1063/1.370226\n[29] M. Werwi ´nski, W. Marciniak, Ab initio study of magnetocrystalline\nanisotropy, magnetostriction, and Fermi surface of L1 0FeNi (tetrataen-\nite), Journal of Physics D: Applied Physics 50 (49) (2017) 495008. doi:\n10.1088/1361-6463/aa958a .\nURL https://doi.org/10.1088/1361-6463/aa958a\n[30] E. Wasserman, Chapter 3 Invar: Moment-volume instabilities in\ntransition metals and alloys, V ol. 5 of Handbook of Ferromag-\nnetic Materials, Elsevier, 1990, pp. 237 – 322. doi:https:\n//doi.org/10.1016/S1574-9304(05)80063-X .\nURL http://www.sciencedirect.com/science/article/pii/\nS157493040580063X\n[31] R. Skomski, J. M. D. Coey, Permanent magnetism, Institute of Physics Pub-\nlishing, 1999.\n[32] J. Joule, On a new class of magnetic forces, Annals of Electricity, Mag-\nnetism, and Chemistry 8 (1842) 219–224.\n[33] L. D. Landau, E. M. Lifshitz, Theory of elasticity / by L. D. Landau and\nE. M. Lifshitz ; translated from the Russian by J. B. Sykes and W. H. Reid,\nPergamon London, 1959.\n[34] R. Birss, The saturation magnetostriction of ferromagnetics, Advances in\nPhysics 8 (31) (1959) 252–291. arXiv:https://doi.org/10.1080/\n00018735900101198 ,doi:10.1080/00018735900101198 .\nURL https://doi.org/10.1080/00018735900101198\n[35] C. Kittel, Physical theory of ferromagnetic domains, Rev. Mod. Phys. 21\n(1949) 541–583. doi:10.1103/RevModPhys.21.541 .\nURL https://link.aps.org/doi/10.1103/RevModPhys.21.541\n50[36] H. Mueller, Properties of rochelle salt. iv, Phys. Rev. 58 (1940) 805–811.\ndoi:10.1103/PhysRev.58.805 .\nURL https://link.aps.org/doi/10.1103/PhysRev.58.805\n[37] D. C. Wallace, Thermodynamics of Crystals, Wiley, New York, 1972.\n[38] F. Mouhat, F. m. c.-X. Coudert, Necessary and sufficient elastic stability\nconditions in various crystal systems, Phys. Rev. B 90 (2014) 224104. doi:\n10.1103/PhysRevB.90.224104 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.90.224104\n[39] L. A.E.H., A Treatise on the Mathematical Theory of Elasticity, Dover pub-\nlisher, New York, 1944.\n[40] R. C. O’Handley, Modern magnetic materials, Wiley, 2000.\n[41] E. W. Lee, Magnetostriction and magnetomechanical effects, Reports on\nProgress in Physics 18 (1) (1955) 184–229. doi:10.1088/0034-4885/18/\n1/305 .\n[42] A. Reuss, Berechnung der Fließgrenze von Mischkristallen auf Grund\nder Plastizitätsbedingung für Einkristalle, ZAMM - Journal of Ap-\nplied Mathematics and Mechanics 9 (1) (1929) 49–58. arXiv:https:\n//onlinelibrary.wiley.com/doi/pdf/10.1002/zamm.19290090104 ,\ndoi:10.1002/zamm.19290090104 .\nURL https://onlinelibrary.wiley.com/doi/abs/10.1002/zamm.\n19290090104\n[43] R. Hill, The elastic behaviour of a crystalline aggregate, Proceedings of\nthe Physical Society. Section A 65 (5) (1952) 349–354. doi:10.1088/\n0370-1298/65/5/307 .\n[44] S. Zhang, R. Zhang, AELAS: Automatic ELAStic property deriva-\ntions via high-throughput first-principles computation, Computer\nPhysics Communications 220 (2017) 403 – 416. doi:https:\n//doi.org/10.1016/j.cpc.2017.07.020 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0010465517302400\n51[45] E. du Trémolet de Lacheisserie, D. Gignoux, M. Schlenker, Magnetism,\nSpringer New York, New York, NY , 2002, Ch. 12 Magnetoelastic Effects,\npp. 351–398.\n[46] W. J. Carr, Chapter 10 Magnetostriction, Magnetic properties of metals and\nalloys, American Society for Metals, Cleveland, 1959.\n[47] G. Kresse, J. Hafner, Ab initio molecular dynamics for liquid metals, Phys.\nRev. B 47 (1993) 558–561. doi:10.1103/PhysRevB.47.558 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.47.558\n[48] G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy cal-\nculations for metals and semiconductors using a plane-wave ba-\nsis set, Computational Materials Science 6 (1) (1996) 15 – 50.\ndoi:https://doi.org/10.1016/0927-0256(96)00008-0 .\nURL http://www.sciencedirect.com/science/article/pii/\n0927025696000080\n[49] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-\nenergy calculations using a plane-wave basis set, Phys. Rev. B 54 (1996)\n11169–11186. doi:10.1103/PhysRevB.54.11169 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.54.11169\n[50] S. P. Ong, W. D. Richards, A. Jain, G. Hautier, M. Kocher, S. Cholia,\nD. Gunter, V . L. Chevrier, K. A. Persson, G. Ceder, Python materials\ngenomics (pymatgen): A robust, open-source python library for mate-\nrials analysis, Computational Materials Science 68 (2013) 314 – 319.\ndoi:https://doi.org/10.1016/j.commatsci.2012.10.028 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0927025612006295\n[51] P. Nieves, S. Arapan, A. P. K ˛ adzielawa, D. Legut, MAELASviewer: An\nonline tool to visualize magnetostriction, Sensors 20 (22). doi:10.3390/\ns20226436 .\nURL https://www.mdpi.com/1424-8220/20/22/6436\n[52] https://maelasviewer.herokuapp.com .\n[53] https://github.com/pnieves2019/MAELASwiewer .\n[54] https://github.com/pnieves2019/MAELAS .\n52[55] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector\naugmented-wave method, Phys. Rev. B 59 (1999) 1758–1775. doi:10.\n1103/PhysRevB.59.1758 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.59.1758\n[56] J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approxima-\ntion made simple, Phys. Rev. Lett. 77 (1996) 3865–3868. doi:10.1103/\nPhysRevLett.77.3865 .\nURL https://link.aps.org/doi/10.1103/PhysRevLett.77.3865\n[57] D. Hobbs, G. Kresse, J. Hafner, Fully unconstrained noncollinear mag-\nnetism within the projector augmented-wave method, Phys. Rev. B 62 (2000)\n11556–11570. doi:10.1103/PhysRevB.62.11556 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.62.11556\n[58] H. J. Monkhorst, J. D. Pack, Special points for brillouin-zone integrations,\nPhys. Rev. B 13 (1976) 5188–5192. doi:10.1103/PhysRevB.13.5188 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.13.5188\n[59] P. Nieves, J. Tranchida, S. Arapan, D. Legut, Spin-lattice model for cubic\ncrystals (2020). arXiv:2012.05076 .\n[60] J. Tranchida, S. Plimpton, P. Thibaudeau, A. Thompson, Massively parallel\nsymplectic algorithm for coupled magnetic spin dynamics and molecular\ndynamics, Journal of Computational Physics 372 (2018) 406 – 425.\ndoi:https://doi.org/10.1016/j.jcp.2018.06.042 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0021999118304200\n[61] S. Plimpton, Fast parallel algorithms for short-range molecular dy-\nnamics, Journal of Computational Physics 117 (1) (1995) 1 – 19.\ndoi:https://doi.org/10.1006/jcph.1995.1039 .\nURL http://www.sciencedirect.com/science/article/pii/\nS002199918571039X\n[62] M. Getzlaff, Fundamentals of magnetism, Springer, Berlin, Heidelberg,\n2008.\n[63] A. Hubert, W. Unger, J. Kranz, Messung der magnetostriktionskonstanten\ndes kobalts als funktion der temperatur, Z. Physik 224 (1) (1969) 148–155.\n53doi:10.1007/BF01392243 .\nURL https://doi.org/10.1007/BF01392243\n[64] M. C. Nguyen, Y . Yao, C.-Z. Wang, K.-M. Ho, V . P. Antropov, Magnetocrys-\ntalline anisotropy in cobalt based magnets: a choice of correlation parame-\nters and the relativistic effects, Journal of Physics: Condensed Matter 30 (19)\n(2018) 195801. doi:10.1088/1361-648x/aab9fa .\nURL https://doi.org/10.1088/1361-648X/aab9fa\n[65] H. Shima, K. Oikawa, A. Fujita, K. Fukamichi, K. Ishida, A. Sakuma, Lat-\ntice axial ratio and large uniaxial magnetocrystalline anisotropy in L1 0-type\nFePd single crystals prepared under compressive stress, Phys. Rev. B 70\n(2004) 224408. doi:10.1103/PhysRevB.70.224408 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.70.224408\n[66] H. Shima, K. Oikawa, A. Fujita, K. Fukamichi, K. Ishida, Magnetic\nanisotropy and mangnetostriction in L1 0FePd alloy, Journal of Mag-\nnetism and Magnetic Materials 272-276 (2004) 2173 – 2174, pro-\nceedings of the International Conference on Magnetism (ICM 2003).\ndoi:https://doi.org/10.1016/j.jmmm.2003.12.898 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0304885303018316\n[67] M. de Jong, W. Chen, T. Angsten, A. Jain, R. Notestine, A. Gamst,\nM. Sluiter, C. Krishna Ande, S. van der Zwaag, J. J. Plata, C. Toher, S. Cur-\ntarolo, G. Ceder, K. A. Persson, M. Asta, Charting the complete elastic\nproperties of inorganic crystalline compounds, Scientific Data 2 (1) (2015)\n150009. doi:10.1038/sdata.2015.9 .\nURL https://doi.org/10.1038/sdata.2015.9\n[68] A. Jain, S. Ong, G. Hautier, W. Chen, W. Richards, S. Dacek, S. Cholia,\nD. Gunter, D. Skinner, G. Ceder, K. Persson, The materials project: A mate-\nrials genome approach to accelerating materials innovation, APL Materials\n1 (2013) 011002.\n[69] K. Persson, Materials data on Fe (SG:229) by materials project (1 2015).\ndoi:10.17188/1189317 .\n[70] J. A. Rayne, B. S. Chandrasekhar, Elastic constants of iron from 4.2 to 300\nk, Phys. Rev. 122 (1961) 1714–1716. doi:10.1103/PhysRev.122.1714 .\nURL https://link.aps.org/doi/10.1103/PhysRev.122.1714\n54[71] K. Persson, Materials data on Co (SG:194) by materials project (2 2016).\ndoi:10.17188/1263614 .\n[72] H. J. McSkimin, Measurement of the elastic constants of single crystal\ncobalt, Journal of Applied Physics 26 (4) (1955) 406–409. arXiv:https:\n//doi.org/10.1063/1.1722007 ,doi:10.1063/1.1722007 .\nURL https://doi.org/10.1063/1.1722007\n[73] K. Persson, Materials data on YCo5 (SG:191) by materials project (2 2016).\ndoi:10.17188/1202401 .\n[74] K. Persson, Materials data on Ni (SG:225) by materials project (2 2016).\ndoi:10.17188/1199153 .\n[75] B.-J. Lee, J.-H. Shim, M. I. Baskes, Semiempirical atomic potentials for the\nfcc metals Cu, Ag, Au, Ni, Pd, Pt, Al, and Pb based on first and second\nnearest-neighbor modified embedded atom method, Phys. Rev. B 68 (2003)\n144112. doi:10.1103/PhysRevB.68.144112 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.68.144112\n[76] K. Persson, Materials data on Fe2Si (SG:164) by materials project (2 2016).\ndoi:10.17188/1198977 .\n[77] K. Persson, Materials data on FePd (SG:123) by materials project (2 2016).\ndoi:10.17188/1202438 .\n[78] A. Al-Ghaferi, P. Müllner, H. Heinrich, G. Kostorz, J. Wiezorek,\nElastic constants of equiatomic L1 0-ordered FePd single crys-\ntals, Acta Materialia 54 (4) (2006) 881 – 889. doi:https:\n//doi.org/10.1016/j.actamat.2005.10.018 .\nURL http://www.sciencedirect.com/science/article/pii/\nS1359645405006129\n[79] K. Persson, Materials data on YCo (SG:63) by materials project (2 2016).\ndoi:10.17188/1307912 .\n[80] J. Kübler, Theory of itinerant electron magnetism, Oxford University Press,\nNew York, 2009.\n[81] S. V . Halilov, A. Y . Perlov, P. M. Oppeneer, A. N. Yaresko, V . N. Antonov,\nMagnetocrystalline anisotropy energy in cubic fe, co, and ni: Applicability\n55of local-spin-density theory reexamined, Phys. Rev. B 57 (1998) 9557–9560.\ndoi:10.1103/PhysRevB.57.9557 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.57.9557\n[82] J. Trygg, B. Johansson, O. Eriksson, J. M. Wills, Total energy calculation\nof the magnetocrystalline anisotropy energy in the ferromagnetic 3 dmetals,\nPhys. Rev. Lett. 75 (1995) 2871–2874. doi:10.1103/PhysRevLett.75.\n2871 .\nURL https://link.aps.org/doi/10.1103/PhysRevLett.75.2871\n[83] M. Fähnle, M. Komelj, R. Q. Wu, G. Y . Guo, Magnetoelasticity of Fe: Pos-\nsible failure of ab initio electron theory with the local-spin-density approx-\nimation and with the generalized-gradient approximation, Phys. Rev. B 65\n(2002) 144436. doi:10.1103/PhysRevB.65.144436 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.65.144436\n[84] N. J. Jones, G. Petculescu, M. Wun-Fogle, J. B. Restorff, A. E. Clark,\nK. B. Hathaway, D. Schlagel, T. A. Lograsso, Rhombohedral magnetostric-\ntion in dilute iron (Co) alloys, Journal of Applied Physics 117 (17) (2015)\n17A913. arXiv:https://doi.org/10.1063/1.4916541 ,doi:10.1063/\n1.4916541 .\nURL https://doi.org/10.1063/1.4916541\n[85] M. Fähnle, M. Komelj, R. Q. Wu, G. Y . Guo, Magnetoelasticity of fe: Pos-\nsible failure of ab initio electron theory with the local-spin-density approx-\nimation and with the generalized-gradient approximation, Phys. Rev. B 65\n(2002) 144436. doi:10.1103/PhysRevB.65.144436 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.65.144436\n[86] J. Sun, A. Ruzsinszky, J. P. Perdew, Strongly constrained and appropriately\nnormed semilocal density functional, Phys. Rev. Lett. 115 (2015) 036402.\ndoi:10.1103/PhysRevLett.115.036402 .\nURL https://link.aps.org/doi/10.1103/PhysRevLett.115.036402\n[87] A. I. Liechtenstein, V . I. Anisimov, J. Zaanen, Density-functional theory and\nstrong interactions: Orbital ordering in Mott-Hubbard insulators, Phys. Rev.\nB 52 (1995) R5467–R5470. doi:10.1103/PhysRevB.52.R5467 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.52.R5467\n56[88] A. Mishima, H. Fujii, T. Okamoto, Anomalous Saturation Magnetostric-\ntion of Gd Single Crystal, Journal of the Physical Society of Japan 40 (4)\n(1976) 962–967. arXiv:https://doi.org/10.1143/JPSJ.40.962 ,doi:\n10.1143/JPSJ.40.962 .\nURL https://doi.org/10.1143/JPSJ.40.962\n[89] S. L. Dudarev, G. A. Botton, S. Y . Savrasov, C. J. Humphreys, A. P. Sut-\nton, Electron-energy-loss spectra and the structural stability of nickel ox-\nide: An lsda+u study, Phys. Rev. B 57 (1998) 1505–1509. doi:10.1103/\nPhysRevB.57.1505 .\nURL https://link.aps.org/doi/10.1103/PhysRevB.57.1505\n[90] H. Kudielka, Die Kristallstruktur von Fe 2Si, ihre Verwandtschaft zu den\nOrdnungsstrukturen des a-(Fe,Si)-Mischkristalls und zur Fe 5Si3-Struktur,\nZeitschrift für Kristallographie - Crystalline Materials 145 (1-6) (01 Dec.\n1977) 177 – 189. doi:https://doi.org/10.1524/zkri.1977.145.16.\n177.\nURL https://www.degruyter.com/view/journals/zkri/145/1-6/\narticle-p177.xml\n[91] Y . Sun, Z. Zhuo, X. Wu, J. Yang, Room-temperature ferromagnetism in two-\ndimensional Fe 2Si nanosheet with enhanced spin-polarization ratio, Nano\nLetters 17 (5) (2017) 2771–2777. doi:10.1021/acs.nanolett.6b04884 .\nURL https://doi.org/10.1021/acs.nanolett.6b04884\n[92] A. V . Krukau, O. A. Vydrov, A. F. Izmaylov, G. E. Scuseria, Influence of the\nexchange screening parameter on the performance of screened hybrid func-\ntionals, The Journal of Chemical Physics 125 (22) (2006) 224106. arXiv:\nhttps://doi.org/10.1063/1.2404663 ,doi:10.1063/1.2404663 .\nURL https://doi.org/10.1063/1.2404663\n[93] C. P. Tang, K. V . Tam, S. J. Xiong, J. Cao, X. Zhang, The structure\nand electronic properties of hexagonal Fe 2Si, AIP Advances 6 (6) (2016)\n065317. arXiv:https://doi.org/10.1063/1.4954667 ,doi:10.1063/\n1.4954667 .\nURL https://doi.org/10.1063/1.4954667\n[94] T. Hikosaka, F. Nakamura, S. Oikawa, A. Takeo, Y . Tanaka, Low noise\nFeAlSi soft magnetic under-layer for CoPtCrO double-layered perpendicular\nrecording media, IEEE Transactions on Magnetics 37 (4) (2001) 1586–1588.\n57[95] P. Villars, K. Cenzual, R. Gladyshevskii, Handbook, De Gruyter, Berlin,\nBoston, 27 Nov. 2015. doi:https://doi.org/10.1515/9783110334678 .\nURL https://www.degruyter.com/view/title/304631\n[96] W. P. Mason, A phenomenological derivation of the first- and second-order\nmagnetostriction and morphic effects for a nickel crystal, Phys. Rev. 82\n(1951) 715–723. doi:10.1103/PhysRev.82.715 .\nURL https://link.aps.org/doi/10.1103/PhysRev.82.715\n[97] E. B. Tadmor, R. E. Miller, Modeling Materials, Cambridge University\nPress, 2009. doi:10.1017/cbo9781139003582 .\nURL https://doi.org/10.1017/cbo9781139003582\n[98] G. Guo, Orientation dependence of the magnetoelastic coupling\nconstants in strained FCC Co and Ni: an ab initio study, Jour-\nnal of Magnetism and Magnetic Materials 209 (1) (2000) 33 – 36.\ndoi:https://doi.org/10.1016/S0304-8853(99)00639-3 .\nURL http://www.sciencedirect.com/science/article/pii/\nS0304885399006393\n58" }, { "title": "2009.06440v1.Multifunctional_Antiperovskites_driven_by_Strong_Magnetostructural_Coupling.pdf", "content": "Multifunctional Antiperovskites driven by Strong Magnetostructural Coupling\nHarish K. Singh, Ilias Samathrakis, Nuno M. Fortunato, Jan Zemen, Chen Shen, Oliver Gut\reisch, and Hongbin Zhang\nInstitute of Materials Science, Technical University Darmstadt,\nOtto-Berndt-Strasse 3, 64287 Darmstadt, Germany\nFaculty of Electrical Engineering, Czech Technical University in Prague,\nTechnicka 2, Prague 166 27, Czech Republic and\nDepartment of Physics, Blackett Laboratory, Imperial College London, London SW7 2AZ, United Kingdom\nBased on density functional theory calculations, we elucidated the origin of multifunctional prop-\nerties for cubic antiperovskites with noncollinear magnetic ground states, which can be attributed\nto strong isotropic and anisotropic magnetostructural coupling. 16 out of 54 stable magnetic an-\ntiperovskite M 3XZ (M = Cr, Mn, Fe, Co, and Ni; X = selected elements from Li to Bi except for\nnoble gases and 4f rare-earth metals; and Z = C and N) are found to exhibit the \u0000 4g/\u00005g(i.e.,\ncharacterized by irreducible representations) antiferromagnetic magnetic con\fgurations driven by\nfrustrated exchange coupling and strong magnetocrystalline anisotropy. Using the magnetic defor-\nmation as an e\u000bective proxy, the isotropic magnetostructural coupling is characterized, and it is\nobserved that the paramagnetic state is critical to understand the experimentally observed negative\nthermal expansion and to predict the magnetocaloric performance. Moreover, the piezomagnetic\nand piezospintronic e\u000bects induced by biaxial strain are investigated. It is revealed that there is\nnot a strong correlation between the induced magnetization and anomalous Hall conductivities by\nthe imposed strain. Interestingly, the anomalous Hall/Nernst conductivities can be signi\fcantly tai-\nlored by the applied strain due to the \fne-tuning of the Weyl points energies, leading to promising\nspintronic applications.\nI. INTRODUCTION\nSmart materials like multiferroic materials with en-\nhanced coupling between di\u000berent degrees of freedom\n(e.g., mechanical, electronic, and magnetic) are promis-\ning for engineering devices for future applications such\nas sensors, transducers, memories, and spintronics.1{3\nCubic antiperovskite (APV) compounds host the two\nmost appealing aspects of multiferroics, e.g., magne-\ntoelectric coupling and piezomagnetic e\u000bect (PME).3,4\nIn APV materials, the strong magnetoelectric coupling\nis achieved by combining piezoelectric and piezomag-\nnetic heterostructure composites.5{7A signi\fcant PME\nis reported for Mn-based nitrides like Mn 3SnN, mak-\ning such compounds a suitable component for fabricat-\ning magnetoelectric composite.8,9The PME in APVs can\nbe attributed to the strong magnetostructural coupling,\nwhich manifests itself also as giant negative thermal ex-\npansion (NTE)10{12and magnetocaloric/barocaloric ef-\nfect.9,13{17From the materials perspective, many Mn-\nbased APV carbides go through a \frst-order magnetic\nphase-transition and possess a large magnetocaloric ef-\nfect.14,18For instance, Mn 3GaC exhibits a huge mag-\nnetic entropy change (\u0001S M) of 15 J/kgK under an ap-\nplied magnetic \feld of 2T.16The strong magnetostruc-\ntural coupling in APVs is driven by the cubic-to-cubic\n\frst-order transition wherein a change in the crystal vol-\nume brings about a change in the frustrated magnetic\nstates. Last but not least, APVs have been investigated\nrecently due to the presence of a treasury of multifunc-\ntionality such as superconductivity,19thermoelectric,20\nmagnetostriction.21\nParticularly, from the topological transport properties\npoint of view, the existence of \fnite anomalous Hall con-ductivity (AHC) in noncollinear antiferromagnets has at-\ntracted noticeable attention due to possible applications\nin AFM spintronics for information storage and data pro-\ncessing.22{25The spin-dependent transport phenomena\ncan provide spin-polarized charge current and large pure\nspin current, which could be achieved premised on two\nfundamental properties, i.e., AHC and spin Hall con-\nductivity (SHC). The kagome lattice turns out to be\nan elementary model to host giant AHC.26{28Recently,\nMn-based APV nitrides have been proposed to exhibit\nlarge AHC in frustrated AFM kagome-lattice.29{32It\nis observed that Mn 3GaN exhibits vanishing and non-\nvanishing AHC for two di\u000berent magnetic ordering \u0000 5g\nand \u0000 4g, respectively.29,30In this regard, for magnetic\nmaterials with noncollinear AFM ground states, the non-\nvanishing AHC is only feasible with speci\fc magnetic\nspace group symmetry, i.e., the AHC tensor depends\non the magnetic group symmetry.33For instance, Mn 3X\n(X= Ga, Ge, and Sn) and Mn 3Z (Z= Ir, Pt, and Rh) have\nbeen reported to have a di\u000berent form of AHC tensor\nas a result of di\u000berent magnetic ordering.22,34{36A pos-\nsible phase transition (\u0000 5g$\u00004g) attained by strain or\nchemical modi\fcation could make these materials suit-\nable for novel AFM spintronic applications. The spin-\npolarized current could also be generated by temperature\ngradient instead of the applied electric \felds, resulting\nin anomalous Nernst conductivity (ANC), also termed\nas spin caloritronics.37;38A large ANC in noncollinear\nAFMs could be useful for establishing spin caloritronics\ndevices that exhibit useful prospects in energy conversion\nand information processing. Noticeably, a large ANC of\n1.80 AK\u00001m\u00001has been reported for APV Mn 3NiN at\n200 K,39which is slightly less than half of the highest\nreported ANC of 4.0 AK\u00001m\u00001in Co 2MnGa.40;41En-arXiv:2009.06440v1 [cond-mat.mtrl-sci] 14 Sep 20202\nFIG. 1: The list of abbreviations used in the manuscript.\nhancing the AHC and ANC by applying strain could be\ncrucial for realizing AFM spintronic devices.\nIn this work, we carried out a systematic analysis of\n54 cubic APV systems (Pm \u00163m) with chemical formula\nM3XZ (see Fig. S1) to determine their magnetic ground\nstates, their tunability via biaxial strain, and the re-\nsulting spintronic properties. Explicit calculations were\nperformed to obtain the energies of four phases, i.e.,\n\u00004g, \u00005g, non-magnetic (NM), and ferromagnetic (FM)),\nwhere the magnetic anisotropy energy (MAE) de\fned as\nthe energy di\u000berence between \u0000 5gand \u0000 4gis examined to\nunderstand the origin of the noncollinear magnetic states\nwith the help of spin-orbit coupling energy. Moreover, a\ndetailed analysis of the lattice constant variation with\nrespect to the magnetic states reveals that the paramag-\nnetic (PM) state is critical in the magnetic phase tran-\nsition, enabling us to predict potential NTE and mag-\nnetocaloric materials. Last but not least, the PME was\nstudied by introducing biaxial strains (compressive and\ntensile), which causes possible phase transitions between\n\u00005g$\u00004g, and leads to a signi\fcant modi\fcation in the\nAHC and ANC, dubbed as a piezospintronic e\u000bect.30,42\nIn-depth analysis on symmetry analysis and electronic\nstructure suggests that the piezospintronic e\u000bect is orig-\ninated from the existence of Weyl nodes whose position\ncan be tailored by strain, resulting in promising applica-\ntions for future spintronic devices.\nII. COMPUTATIONAL DETAILS\nOur density functional theory (DFT) calculations are\nperformed using the projector augmented wave (PAW)\nmethod as e\u000bectuated in the VASP package.43Theexchange-correlation functional is approximated using\nthe generalized gradient approximation (GGA) as param-\neterized by Perdew-Burke-Ernzerhof (PBE).44We used\nan energy cuto\u000b of 500 eV for the plane-wave basis set,\nand a uniform k-points grid of 13 \u000213\u000213 within the\nMonkhorst-pack scheme for the Brillouin zone integra-\ntions. The Methfessel-Paxton scheme is used to deter-\nmine the partial occupancies of orbitals with a smearing\nwidth of 0.06 eV. The spin-orbit coupling (SOC) is con-\nsidered in all the calculations. In order to verify the\nmagnetic ground state of various compounds, the more\naccurate total energy calculations are performed using a\nhigher energy cuto\u000b of 600 eV and a dense k-mesh of\n25\u000225\u000225.\nThe magnetic ground states are obtained by comparing\nthe total energies of \fve con\fgurations (see Table S1)\n\u00004g, \u00005g, FM, PM (see Fig. 2, and NM states, where the\nlattice constants are fully optimized for each magnetic\ncon\fguration (see Table S2). The PM state is modeled\nbased on the disordered local moment (DLM) picture\n(see Fig. 2(d)), where a special quasi-random structure\nmodeled using a 2 \u00022\u00022 supercell by imposing zero to-\ntal magnetization.45That is, a supercell with the same\nnumber of up and down moments is considered with a\npair correlation function same as A 50B50random alloys,\ngenerated using the alloy theoretic automated toolkit\ncode.46Moreover, to investigate the piezomagnetic and\npiezospintronic e\u000bects, biaxial in-plane strain is applied,\nwhich reduces the crystalline symmetry from cubic to\ntetragonal. The crystal structures are fully relaxed where\nthe optimal lattice constants along c-direction are eval-\nuated by the polynomial \ftting of the energies from a\nseries of calculations.\nThe AHC is evaluated using the WannierTools code,47,\nwhere the required accurate tight-binding models are\nobtained by the maximally localized Wannier functions\n(MLWFs) using the Wannier90 code.48The s, p, and d\norbitals of M and X atoms and the s and p orbitals for N\natom are considered, resulting in 80 MLWFs in total for\nevery noncollinear APVs. The AHC is computed by inte-\ngrating the Berry curvature on a uniform 401 \u0002401\u0002401\nk-points mesh to guarantee good accuracy, which can be\nexpressed as:49\n\u001bxy=\u0000e2\n~Zdk\n(2\u0019)3X\nnf\u0002\n\u000f(k)\u0000\u0016\u0003\n\nn;xy(k) (1)\n\nn;xy(k) =\u00002ImX\nm6=nh knjvxj kmih kmjvyj kni\n\u0002\n\u000fm(k)\u0000\u000fn(k)\u00032(2)\nwhere e is elementary charge, \u0016is the chemical potential,\n n=mdenotes the Bloch wave function with energy eigen-\nvalue\u000fn=m,vx=yis the velocity operator along Cartesian\nx=ydirection, and f[\u000f(k)\u0000\u0016] is Fermi-Dirac distribution\nfunction. The ANC \u000bxyis evaluated based on the Mott3\nFIG. 2: Possible magnetic structures of antiperovskites: (a) \u0000 5g, (b) \u0000 4g, (c) FM and, (d) PM. The spin chirality for \u0000 4gand\n\u00005gis assumed to be +1.\nrelation, which yields:\n\u000bxy=\u0000\u00192k2\nBT\n3ed\u001bxy\nd\u000f\f\f\n\u000f=\u0016(3)\nIn this work, we evaluated only the derivative of AHC at\nthe Fermi level d \u001bxy/d\u000f, which provides a quantitative\nmeasurements of the ANC.\nIII. RESULTS AND DISCUSSION\nA. Validation and prediction of magnetic ground\nstate\nAs the magnetic transition metal atoms are located at\nthe face centers of the cubic cell for magnetic APVs, it\nleads to a frustrated kagome-lattice in the (111)-plane.\nCorrespondingly, noncollinear magnetic structures are\nexpected when the interatomic exchange interaction is\nAFM for the nearest neighbors. As shown in Fig. 2, \u0000 4g,\nand \u0000 5gare the two most common magnetic con\fgura-\ntions reported for APVs, resulting in 120\u000emagnetic angle\ncon\fgurations within the (111)-plane between three mag-\nnetic moments. The \u0000 4gstate can be obtained from \u0000 5g\nby simultaneously rotating the moments of three metal\natoms within the (111)-plane by 90\u000e. The APVs with the\nnoncollinear magnetic ground state are listed in Table. I,\nin comparison to the available experimentally measured\nresults. Interestingly, all the APVs with noncollinear\nmagnetic ground states are nitrides including Cr and Mn,\nwhile all carbides end up with the FM ground state ex-\ncept Mn 3SnC, which exhibits the \u0000 5gstate (see Table\nS1). It is noted that there are also other magnetic con-\n\fgurations such as ferrimagnetic and canted states for\nMn3XC (X = Ga, Sn, and Zn),50{52which will be saved\nfor future investigations.\nThe magnetic ground states of noncollinear-APVs are\nin good agreement with the experimental measurements.\nFor instance, both Mn 3GaN13and Mn 3ZnN53,54have the\n\u00005gmagnetic ground state, which are consistent with our\nDFT calculations. Interestingly, many Mn-based APVsexhibit a mixed \u0000 4g+ \u0000 5gmagnetic ordering, with a\npossible meta-magnetic transition to the other magnetic\nphases.53For instance, Mn 3AgN is characterized by two\ndistinct magnetic phase transitions. A mixed \u0000 4g+ \u0000 5g\nphase exists below 55 K, whereas pure \u0000 5gstate persists\nat intermediate temperature range (55 K 0) of the dI/dVbspectrum starts with a small bump around \u00180:5 V and\na steep rise in d I/dVbat\u00181:2 V, as shown in Figure 2a. Those features, apart from a hard\nshift of \u00180:5 eV, are consistent with the DFT spin polarized projected density of states\n(PDOS) on the CrBr 3layer of the heterostructure shown in Figure 2b. The steep rises in the\ndI/dVbspectrum are around 3 V apart and mostly arise from the spin up bands of the CrBr 3\nlayer: its electronic properties are well preserved in the heterostructure as compared to the\npristine CrBr 3layer, as can be seen in Figure 2b and c. The band structures and PDOS\nshown in Figure 2c reveal that the bands in the [-2.0,-1.0] eV and [-1.0,1.0] eV windows have\na majority contribution from the NbSe 2layer, where the NbSe 2'sd-band in the [-1.0,1.0] eV\nwindow is completely preserved and slightly spin polarized due to the proximity with the6\nmagnetic CrBr 3. However, the PDOS on the CrBr 3layer in the [-2.0,-1.0] eV and [-1.0,1.0]\neV windows are non-zero and consistent with the bumps in the d I/dVbspectrum, which we\nattribute to charge recon\fguration in the heterostructure as shown in Figure 3a.\nFIG. 3. (a) Di\u000berential charge density of the CrBr 3-NbSe 2heterostructure, where the yellow and\nblue colors indicate charge accumulation (+) and depletion (-), respectively. (b) Magnetic hysteresis\nin monolayer CrBr 3on NbSe 2at several di\u000berent temperatures (indicated in the \fgure). (c) The\ntemperature dependence of the coercive \feld of CrBr 3on NbSe 2. The coercive \feld decreases as\nthe temperature increases until it vanishes at 16 K.\nAfter the electronic characterization of the samples, we will next focus on their magnetic\nproperties. The isolated CrBr 3layer has a predicted out-of-plane magnetic moment of 6.000\n\u0016Bper unit cell, where each Cr atom has three unpaired electrons. Our DFT calculations\nshow that the magnetism of the CrBr 3layer is well-preserved in the heterostructure, which\nshows a slightly larger magnetic moment of 6.097 \u0016Bper unit cell due to induced magne-\ntization on the NbSe 2layer. The PDOS on the CrBr 3layer of the heterostructure reveals\nthat the majority spin up channel has a band gap of 3 eV, while the minority spin down\nchannel has a band gap of around 5 eV, both shown in blue and red, respectively, in Figure\n2c.\nWe study the truly 2D itinerant ferromagnetism in CrBr 3monolayer on NbSe 2using\nmagneto-optical Kerr e\u000bect (MOKE) microscopy (details of the experimental procedures\nare given the Methods section). The magnetic \feld was applied perpendicular to the sam-7\nple. Figure 3b shows the MOKE signal of a CrBr 3monolayer on NbSe 2as a function of\nthe external magnetic \feld at several di\u000berent temperatures. Notably, monolayer CrBr 3is\nferromagnetic, as evidenced by the prominent hysteresis seen here. MOKE measurements\nfurther reveal that as the temperature is increased, the hysteresis loop shrinks and eventually\ndisappears at a transition temperature Tcof about 16 K. The Tccan also be extracted from\na temperature dependence of the coercivity Hc(blue squares in Figure 3c), where the onset\nofHcin the CrBr 3monolayer clearly occurs at around 16 K. The same behaviour was ob-\nserved for mechanically exfoliated monolayer CrBr 3\rake, where the transition temperature\nTcwas around 20 K33. Moreover, the out-of-plane coercive \feld Hcfor exfoliated monolayer\nCrBr 3\rake is 4 mT28,33,34while in our MBE grown monolayer CrBr 3on NbSe 2substrate\nthe out-of-plane coercive \feld increases up to \u00182:5 mT at the lowest temperatures we can\nreach in our MOKE setup as shown in Figure 3c.\nAfter con\frming the existence of ferromagnetism on the monolayer CrBr 3on NbSe 2\nsubstrate, it is interesting to see the e\u000bects of the superconducting substrate NbSe 2on the\nmonolayer CrBr 3. Isolated CrBr 3is a ferromagnetic insulator; however, due to the monolayer\nthickness, it is possible to tunnel through such a structure with STM. In addition, due to\nthe charge transfer at the interface between CrBr 3and NbSe 2, the heterostructure itself\nhas a metallic nature. Therefore, one can expect a measurable interaction between CrBr 3\nand NbSe 2. The superconductivity of CrBr 3/NbSe 2heterostructure was studied by STM at\nT= 350 mK. Figures 4a,b show experimental d I/dVbspectra (raw data) taken on a bare\nNbSe 2and on monolayer CrBr 3, respectively. The d I/dVbspectrum of bare NbSe 2(Figure\n4a) has a hard gap with an extended region of zero di\u000berential conductance around the Fermi\nenergy. dI/dVbspectra were \ftted by the McMillan two-band model35, with parameters\n\u00011= 1:28 meV,\r1= 0:38 and \u0001 2= 0:74 meV,\r1= 0:13 meV, respectively. In contrast, the\nspectra taken in the middle of the CrBr 3island have small but distinctly non-zero di\u000berential\nconductance inside the gap of the NbSe 2substrate. We observe pairs of conductance onsets\nat\u00060:3 mV around zero bias. This feature results from the formation of Shiba-bands in the\nNbSe 2caused by the induced magnetization from the CrBr 3island25. These extra features\nare not reproduced by the two-band model (Figure 4b). Nevertheless, a double gap s-wave\nBCS-type \ftting gives a slight reduction of both gap parameters (\u0001 1= 1:20 meV,\r1= 0:40\nmeV and \u0001 2= 0:73 meV,\r2= 0:50 meV respectively).\nTo obtain more detailed insight into the superconductivity on CrBr 3/NbSe 2heterostruc-8\nFIG. 4. (a,b) Experimental d I/dVspectroscopy (black) on the NbSe 2substrate (a) and in the\nmiddle of a CrBr 3island (b) measured at T= 350 mK. We also show a \ft to a double gap s-wave\nBCS-type model (red). (c) STM topography image (STM feedback parameters: Vbias= +1 V,\nI= 10 pA, scale bar: 22 nm). (d-f) Vortex imaging on CrBr 3/NbSe 2heterostructure. We have\nrecorded grid spectroscopy (100 \u0002100 spectra) over an area of 110 \u0002110 nm2atT= 350 mK under\nan applied out-of-plane magnetic \feld of 0.65 T. The maps are obtained from full spectroscopic\nscans from -3 mV to 3 mV at each pixel at the indicated bias voltages. (g) Line scan of the d I/dVb\nsignal along the white line marked in panel (d).\nture, we investigate the dependence of our d I/dVbspectra under an applied out-of-plane\nmagnetic \feld. In a type-II SC such as NbSe 2, we would expect to observe an Abrikosov\nvortex lattice in a d I/dVbmap acquired near the energy of the superconducting gap36.\nFigs. 4d-f show d I/dVbgrid maps at 0, 0.6 and 0.8 mV bias voltages, respectively. The\nmaps are recorded under 0.65 T out-of-plane magnetic \feld on an area with both CrBr 3\nislands and bare NbSe 2surface (Figure 4c). It is seen from Figures 4d-f that the vortices\nexhibit a highly ordered hexagonal lattice similar to those observed on the clean NbSe 2\nsurface. This is the \frst time vortices have been clearly observed in a hybrid ferromagnet-9\nsuperconductor-system. We measured the spatial variation of the d I/dVbspectra as a\nfunction of distance away from the vortex center (along the dashed line in Figure 4d). The\nresults are given in Figure 4g, which shows the measured d I/dVbas functions of distance\nand sample bias V on a color scale. One can see that only one peak appears at zero-bias\nin the dI/dVbspectra near the vortex center, and this peak splits into two away from the\nvortex core. The splitting energy increases linearly with distance. One of the intriguing\nproperties of a topological superconductor is that vortices on its surface are expected to host\nMajorana zero modes37. This mode results in a peak in the local density-of-states at the\nFermi energy in the center of the vortex. In contrast to bound states in vortices on conven-\ntional superconductors, a Majorana mode should not split in energy away from the vortex\ncenter37{39. Due to the broadening of the resonance in the vortex spectra, we cannot resolve\nthe individual components in the d I/dVbspectra and the zero bias feature persists up to\n5 nm from the vortex core before the clearly split features can be observed. This is within\nthe range of spatial distributions of the Majorana zero mode reported in the literature (a\nfew nm up to 30 nm)38{41. In addition, it is important to note that Majorana zero modes in\nvortex cores have only been observed at low magnetic \feld (e.g. 0.1 T for Bi 2Se3/NbSe 238,41\nand its amplitude is expected to decay exponentially as the external \feld is increased. In\norder to con\frm whether the vortices in our system host Majorana zero modes at their\ncores, further experiments, in particular using spin-polarized tunneling measurements42, are\nclearly necessary.\nIn summary, we have provided experimental evidence for the realisation of a 2D ferromagnet-\nsuperconductor van der Waals heterostructure. More importantly, we have experimentally\ncon\frmed that the CrBr 3monolayer retains its ferromagnetic ordering with a magnetocrys-\ntalline anisotropy favoring an out-of-plane spin orientation on NbSe 2. Our DFT calculations\nshowing an induced moment in Nb from hybridization with Cr d-orbitals con\frm the im-\nprinting of magnetic order on NbSe 2from a 2D vdW magnetic insulator. Our results provide\na broader framework for employing other proximity e\u000bects to tailor materials and realize\nnovel phenomena in 2D heterostructures.10\nMETHODS\nMolecularbeam epitaxy (MBE) sample growth: The CrBr 3thin \flm was grown on a freshly\ncleaved NbSe 2substrate by compound source MBE. The anhydrous CrBr 3powder of 99 %\npurity was evaporated from a Knudsen cell. Before growth, the cell was degassed up to the\ngrowth temperature 350\u000eC until the vacuum was better than 1 \u000210\u00008mbar. The sample was\nheated by electron beam bombardment and temperatures were measured using an optical\npyrometer. The growth speed was determined by checking the coverage of the as-grown\nsamples by STM. The optimal substrate temperature for the growth of CrBr 3monolayer\n\flms was \u0018270\u000eC. Below this temperature, CrBr 3forms disordered clusters on the NbSe 2\nsurface. The NbSe 2crystal is directly mounted on a sample holder using a two component\nconducting silver epoxy which only allows us to heat the sample up to \u0018300\u000eC.\nScanning tunneling microscopy (STM) and spectroscopy (STS) measurements: After\nthe sample preparation, it was inserted into a low-temperature STM (Unisoku USM-1300)\nhoused in the same UHV system and all subsequent experiments were performed at T= 350\nmK. STM images were taken in constant-current mode. d I/dVbspectra were recorded by\nstandard lock-in detection while sweeping the sample bias in an open feedback loop con\fg-\nuration, with a peak-to-peak bias modulation of 30-50 \u0016V for a small bias range and 10 mV\nfor a lager bias range, respectively at a frequency of 707 Hz. Spectra from grid spectroscopy\nexperiments were normalized by the normal state conductance, i.e. d I/dVbat a bias voltage\ncorresponding to a few times the superconducting gap.\nSample transfer: After the sample growth, the sample was transferred to the load lock\n(\u001810\u00009mbar) and then to the directly connected glove bag. The load lock is slowly vented\nwith pure nitrogen to the glove bag atmosphere, and the sample is then transferred into the\nglove bag with a magnetic transfer rod for XPS and MOKE measurements.\nX-Ray photoelectron spectroscopy (XPS) experiments: The XPS spectra were measured\nusing a Kratos Axis Ultra system, equipped with monochromatic Al K \u000bX-ray source. All\nmeasurements were performed using an analysis area of 0 :3mm\u00020:7mm. The black curve\nin Fig 1b was measured using 80 eV pass energy and 1 eV energy step whereas the colored\ncurves were taken with 20 eV pass energy and 0.1 eV energy step. The energy calibration\nwas done using the C 1s peak at 284.8 eV.\nMagneto-optical Kerr e\u000bect (MOKE) measurements: MOKE was carried out using an11\nEvico Magnetics system based on a Zeiss Axio Imager D1 microscope with a Hamamatsu\nC4742-95 digital camera. The sample was placed in a Janis research ST-500 cold \fnger\ncryostat with optical access and cooled using liquid He. Imaging of the sample was carried\nout in a polar Kerr con\fguration. Hysteresis loops were constructed from images taken\nusing a long working distance 100x lens after background image subtraction and corrected\nfor a linear background slope due to the Faraday e\u000bect on the lens.\nDFT calculations: Calculations were performed with the DFT methodology as imple-\nmented in the periodic plane-wave basis VASP code43,44. Atomic positions and lattice pa-\nrameters were obtained by fully relaxing all structures using the spin-polarized Perdew-\nBurke-Ernzehof (PBE) functional45including Grimme's semiempirical DFT-D3 scheme for\ndispersion correction46, which is important to describe the van der Waals (vdW) interactions\nbetween the CrBr 3and the NbSe 2layers. The stacking of the layers used in our calculations\nis the same used in a previous work25. The convergence criterion of self-consistent \feld\n(SCF) computation was set to 10\u00005eV and the threshold for the largest force acting on the\natoms was set to less than 0.012 eV/ \u0017A. A vacuum layer of 12 \u0017A was added to avoid mirror\ninteractions between periodic images. Further calculations of band structures and density\nof states were realized using the hybrid Heyd-Scuseria-Ernzerhof (HSE06) functional47{49,\nwhich improves the description of the band structure as compared to the PBE functional.\nThe interactions between electrons and ions were described by PAW pseudopotentials, where\n4sand 4pshells were added explicitly as semicore states for Nb and 3 pshells for Cr. An\nenergy cuto\u000b of 550 eV is used to expand the wave functions and a systematic k-point con-\nvergence was checked, where the total energy was converged to the order of 10\u00004eV. Spin\npolarization was considered by setting an initial out-of-plane magnetization of 3 \u0016Bper Cr\natom and zero otherwise.\nACKNOWLEDGMENTS\nThis research made use of the Aalto Nanomicroscopy Center (Aalto NMC) facilities and\nwas supported by the European Research Council (ERC-2017-AdG no. 788185 \\Arti\fcial\nDesigner Materials\") and Academy of Finland (Academy professor funding no. 318995 and\n320555, Academy postdoctoral researcher no. 309975). Computing resources from the Aalto\nScience-IT project and CSC, Helsinki are gratefully acknowledged. ASF has been supported12\nby the World Premier International Research Center Initiative (WPI), MEXT, Japan.\n\u0003shawulienu.kezilebieke@aalto.\f\n1K. S. Burch, D. Mandrus, and J.-G. Park, Magnetism in two-dimensional van der Waals mate-\nrials, Nature 563, 47 (2018).\n2M. Gibertini, M. Koperski, A. F. Morpurgo, and K. S. Novoselov, Magnetic 2D materials and\nheterostructures, Nat. Nanotechnol. 14, 408 (2019).\n3C. Gong and X. Zhang, Two-dimensional magnetic crystals and emergent heterostructure de-\nvices, Science 363, 10.1126/science.aav4450 (2019).\n4C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao, W. Bao, C. Wang, Y. Wang, Z. Q.\nQiu, R. J. Cava, S. G. Louie, J. Xia, and X. Zhang, Discovery of intrinsic ferromagnetism in\ntwo-dimensional van der Waals crystals, Nature 546, 265 (2017).\n5B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D. Zhong,\nE. Schmidgall, M. A. McGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo-Herrero, and X. Xu,\nLayer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit, Nature\n546, 270 (2017).\n6S. Jiang, J. Shan, and K. F. Mak, Electric-\feld switching of two-dimensional van der Waals\nmagnets, Nat. Mater. 17, 406 (2018).\n7D. Zhong, K. L. Seyler, X. Linpeng, R. Cheng, N. Sivadas, B. Huang, E. Schmidgall,\nT. Taniguchi, K. Watanabe, M. A. McGuire, W. Yao, D. Xiao, K.-M. C. Fu, and X. Xu, Van der\nWaals engineering of ferromagnetic semiconductor heterostructures for spin and valleytronics,\nSci. Adv. 3, e1603113 (2017).\n8X. Lin, W. Yang, K. L. Wang, and W. Zhao, Two-dimensional spintronics for low-power elec-\ntronics, Nature Electronics 2, 274 (2019).\n9M.-C. Wang, C.-C. Huang, C.-H. Cheung, C.-Y. Chen, S. G. Tan, T.-W. Huang, Y. Zhao,\nY. Zhao, G. Wu, Y.-P. Feng, H.-C. Wu, and C.-R. Chang, Prospects and opportunities of 2D\nvan der Waals magnetic systems, Ann. Phys. 532, 1900452 (2020).\n10V. P. Ningrum, B. Liu, W. Wang, Y. Yin, Y. Cao, C. Zha, H. Xie, X. Jiang, Y. Sun, S. Qin,\nX. Chen, T. Qin, C. Zhu, L. Wang, and W. Huang, Recent advances in two-dimensional magnets:\nPhysics and devices towards spintronic applications, Research 2020 , 1768918 (2020).13\n11X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83,\n1057 (2011).\n12M. Sato and Y. Ando, Topological superconductors: a review, Rep. Prog. Phys. 80, 076501\n(2017).\n13C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. Das Sarma, Non-abelian anyons and\ntopological quantum computation, Rev. Mod. Phys. 80, 1083 (2008).\n14R. M. Lutchyn, E. P. A. M. Bakkers, L. P. Kouwenhoven, P. Krogstrup, C. M. Marcus, and\nY. Oreg, Majorana zero modes in superconductor-semiconductor heterostructures, Nat. Rev.\nMater. 3, 52 (2018).\n15P. Zhang, K. Yaji, T. Hashimoto, Y. Ota, T. Kondo, K. Okazaki, Z. Wang, J. Wen, G. D.\nGu, H. Ding, and S. Shin, Observation of topological superconductivity on the surface of an\niron-based superconductor, Science 360, 182 (2018).\n16D. Wang, L. Kong, P. Fan, H. Chen, S. Zhu, W. Liu, L. Cao, Y. Sun, S. Du, J. Schneeloch,\nR. Zhong, G. Gu, L. Fu, H. Ding, and H.-J. Gao, Evidence for Majorana bound states in an\niron-based superconductor, Science 362, 333 (2018).\n17S. Zhu, L. Kong, L. Cao, H. Chen, M. Papaj, S. Du, Y. Xing, W. Liu, D. Wang, C. Shen,\nF. Yang, J. Schneeloch, R. Zhong, G. Gu, L. Fu, Y.-Y. Zhang, H. Ding, and H.-J. Gao, Nearly\nquantized conductance plateau of vortex zero mode in an iron-based superconductor, Science\n367, 189 (2020).\n18Z. Wang, J. O. Rodriguez, L. Jiao, S. Howard, M. Graham, G. D. Gu, T. L. Hughes, D. K.\nMorr, and V. Madhavan, Evidence for dispersing 1D Majorana channels in an iron-based su-\nperconductor, Science 367, 104 (2020).\n19A. I. Buzdin, Proximity e\u000bects in superconductor-ferromagnet heterostructures, Rev. Mod.\nPhys. 77, 935 (2005).\n20D. Zhong, K. L. Seyler, X. Linpeng, N. P. Wilson, T. Taniguchi, K. Watanabe, M. A. McGuire,\nK.-M. C. Fu, D. Xiao, W. Yao, and X. Xu, Layer-resolved magnetic proximity e\u000bect in van der\nWaals heterostructures, Nat. Nanotechnol. 15, 187 (2020).\n21B. Huang, M. A. McGuire, A. F. May, D. Xiao, P. Jarillo-Herrero, and X. Xu, Emergent\nphenomena and proximity e\u000bects in two-dimensional magnets and heterostructures, Nat. Mater.\n(2020).14\n22W. Chen, Z. Sun, Z. Wang, L. Gu, X. Xu, S. Wu, and C. Gao, Direct observation of van der\nWaals stacking-dependent interlayer magnetism, Science 366, 983 (2019).\n23S. Liu, X. Yuan, Y. Zou, Y. Sheng, C. Huang, E. Zhang, J. Ling, Y. Liu, W. Wang, C. Zhang,\nJ. Zou, K. Wang, and F. Xiu, Wafer-scale two-dimensional ferromagnetic Fe 3GeTe 2thin \flms\ngrown by molecular beam epitaxy, npj 2D Mater. Appl. 1, 30 (2017).\n24S. Kezilebieke, M. N. Huda, P. Dreher, I. Manninen, Y. Zhou, J. Sainio, R. Mansell, M. M.\nUgeda, S. van Dijken, H.-P. Komsa, and P. Liljeroth, Electronic and magnetic characterization\nof epitaxial VSe 2monolayers on superconducting NbSe 2, Commun. Phys. 3, 116 (2020).\n25S. Kezilebieke, M. N. Huda, V. Va\u0014 no, M. Aapro, S. C. Ganguli, O. J. Silveira, S. G lodzik, A. S.\nFoster, T. Ojanen, and P. Liljeroth, Topological superconductivity in a designer ferromagnet-\nsuperconductor van der Waals heterostructure, arXiv:2002.02141 .\n26Z. Wu, J. Yu, and S. Yuan, Strain-tunable magnetic and electronic properties of monolayer\nCrI3, Phys. Chem. Chem. Phys. 21, 7750 (2019).\n27S. Jiang, L. Li, Z. Wang, K. F. Mak, and J. Shan, Controlling magnetism in 2D CrI 3by\nelectrostatic doping, Nat. Nanotechnol. 13, 549 (2018).\n28Z. Zhang, J. Shang, C. Jiang, A. Rasmita, W. Gao, and T. Yu, Direct photoluminescence\nprobing of ferromagnetism in monolayer two-dimensional CrBr 3, Nano Lett. 19, 3138 (2019).\n29L. Webster and J.-A. Yan, Strain-tunable magnetic anisotropy in monolayer CrCl 3, CrBr 3, and\nCrI3, Phys. Rev. B 98, 144411 (2018).\n30Y. Xu, X. Liu, and W. Guo, Tensile strain induced switching of magnetic states in NbSe 2and\nNbS 2single layers, Nanoscale 6, 12929 (2014).\n31J. C. Carver, G. K. Schweitzer, and T. A. Carlson, Use of xray photoelectron spectroscopy to\nstudy bonding in cr, mn, fe, and co compounds, J. Chem. Phys. 57, 973 (2020).\n32I. Pollini, Multiplet splitting of core-electron binding energies in chromium compounds, PRB\n67, 155111 (2003).\n33M. Kim, P. Kumaravadivel, J. Birkbeck, W. Kuang, S. G. Xu, D. G. Hopkinson, J. Knolle, P. A.\nMcClarty, A. I. Berdyugin, M. Ben Shalom, R. V. Gorbachev, S. J. Haigh, S. Liu, J. H. Edgar,\nK. S. Novoselov, I. V. Grigorieva, and A. K. Geim, Micromagnetometry of two-dimensional\nferromagnets, Nat. Electron. 2, 457 (2019).\n34H. H. Kim, B. Yang, S. Li, S. Jiang, C. Jin, Z. Tao, G. Nichols, F. S\fgakis, S. Zhong, C. Li,\nS. Tian, D. G. Cory, G.-X. Miao, J. Shan, K. F. Mak, H. Lei, K. Sun, L. Zhao, and A. W. Tsen,15\nEvolution of interlayer and intralayer magnetism in three atomically thin chromium trihalides,\nProc. Natl. Acad. Sci. 116, 11131 (2019).\n35Y. Noat, J. A. Silva-Guilln, T. Cren, V. Cherkez, C. Brun, S. Pons, F. Debontridder,\nD. Roditchev, W. Sacks, L. Cario, P. Ordejn, A. Garca, and E. Canadell, Quasiparticle spectra\nof 2H\u0000NbSe 2: Two-band superconductivity and the role of tunneling selectivity, Phys. Rev.\nB92, 134510 (2015).\n36H. F. Hess, R. B. Robinson, R. C. Dynes, J. M. Valles, and J. V. Waszczak, Scanning-tunneling-\nmicroscope observation of the Abrikosov \rux lattice and the density of states near and inside a\n\ruxoid, Phys. Rev. Lett. 62, 214 (1989).\n37H.-H. Sun and J.-F. Jia, Detection of majorana zero mode in the vortex, npj Quantum Materials\n2, 34 (2017).\n38J.-P. Xu, M.-X. Wang, Z. L. Liu, J.-F. Ge, X. Yang, C. Liu, Z. A. Xu, D. Guan, C. L. Gao,\nD. Qian, Y. Liu, Q.-H. Wang, F.-C. Zhang, Q.-K. Xue, and J.-F. Jia, Experimental detection of\na majorana mode in the core of a magnetic vortex inside a topological insulator-superconductor\nbi2te3=nbse 2heterostructure, Phys. Rev. Lett. 114, 017001 (2015).\n39J.-X. Yin, Z. Wu, J.-H. Wang, Z.-Y. Ye, J. Gong, X.-Y. Hou, L. Shan, A. Li, X.-J. Liang, X.-X.\nWu, J. Li, C.-S. Ting, Z.-Q. Wang, J.-P. Hu, P.-H. Hor, H. Ding, and S. H. Pan, Observation\nof a robust zero-energy bound state in iron-based superconductor fe(te,se), Nature Physics 11,\n543 (2015).\n40Q. Liu, C. Chen, T. Zhang, R. Peng, Y.-J. Yan, C.-H.-P. Wen, X. Lou, Y.-L. Huang, J.-P. Tian,\nX.-L. Dong, G.-W. Wang, W.-C. Bao, Q.-H. Wang, Z.-P. Yin, Z.-X. Zhao, and D.-L. Feng,\nRobust and clean majorana zero mode in the vortex core of high-temperature superconductor\n(li0:84fe0:16)OHFeSe, PRX 8, 041056 (2018).\n41D. Wang, L. Kong, P. Fan, H. Chen, S. Zhu, W. Liu, L. Cao, Y. Sun, S. Du, J. Schneeloch,\nR. Zhong, G. Gu, L. Fu, H. Ding, and H.-J. Gao, Evidence for majorana bound states in an\niron-based superconductor, Science , eaao1797 (2018).\n42H.-H. Sun, K.-W. Zhang, L.-H. Hu, C. Li, G.-Y. Wang, H.-Y. Ma, Z.-A. Xu, C.-L. Gao, D.-\nD. Guan, Y.-Y. Li, C. Liu, D. Qian, Y. Zhou, L. Fu, S.-C. Li, F.-C. Zhang, and J.-F. Jia,\nMajorana zero mode detected with spin selective andreev re\rection in the vortex of a topological\nsuperconductor, PRL 116, 257003 (2016).16\n43G. Kresse and J. Furthm uller, E\u000ecient iterative schemes for ab initio total-energy calculations\nusing a plane-wave basis set, Phys. Rev. B 54, 11169 (1996).\n44G. Kresse and J. Furthmller, E\u000eciency of ab-initio total energy calculations for metals and\nsemiconductors using a plane-wave basis set, Computational Materials Science 6, 15 (1996).\n45J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple,\nPhys. Rev. Lett. 77, 3865 (1996).\n46S. Grimme, J. Antony, S. Ehrlich, and H. Krieg, A consistent and accurate ab initio parametriza-\ntion of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem.\nPhys. 132, 154104 (2010).\n47J. Heyd and G. E. Scuseria, E\u000ecient hybrid density functional calculations in solids: Assessment\nof the HeydScuseriaErnzerhof screened Coulomb hybrid functional, J. Chem. Phys. 121, 1187\n(2004).\n48J. Heyd, G. E. Scuseria, and M. Ernzerhof, Erratum: Hybrid functionals based on a screened\nCoulomb potential [J. Chem. Phys. 118, 8207 (2003)], J. Chem. Phys. 124, 219906 (2006).\n49J. Heyd, G. E. Scuseria, and M. Ernzerhof, Hybrid functionals based on a screened Coulomb\npotential, J. Chem. Phys. 118, 8207 (2003)." }, { "title": "2011.00394v1.Tunable_magneto_optical_effect__anomalous_Hall_effect_and_anomalous_Nernst_effect_in_two_dimensional_room_temperature_ferromagnet__1T__CrTe__2_.pdf", "content": "Tunable magneto-optical e\u000bect, anomalous Hall e\u000bect and anomalous Nernst e\u000bect in\ntwo-dimensional room-temperature ferromagnet 1T-CrTe 2\nXiuxian Yang,1, 2Xiaodong Zhou,1Wanxiang Feng,1,\u0003and Yugui Yao1\n1Key Laboratory of Advanced Optoelectronic Quantum Architecture and Measurement,\nMinistry of Education, School of Physics, Beijing Institute of Technology, Beijing 100081, China\n2Kunming Institute of Physics, Kunming 650223, China\n(Dated: November 3, 2020)\nUtilizing the \frst-principles density functional theory calculations together with group theory\nanalyses, we systematically investigate the spin order-dependent magneto-optical e\u000bect (MOE),\nanomalous Hall e\u000bect (AHE), and anomalous Nernst e\u000bect (ANE) in a recently discovered two-\ndimensional room-temperature ferromagnet 1 T-CrTe 2. We \fnd that the spin prefers an in-plane\ndirection by the magnetocrystalline anisotropy energy calculations. The MOE, AHE, and ANE\ndisplay a period of 2 \u0019=3 when the spin rotates within the atomic plane, and they are forbidden if\nthere exists a mirror plane perpendicular to the spin direction. By reorienting the spin from in-plane\nto out-of-plane direction, the MOE, AHE, and ANE are enhanced by around one order of magnitude.\nMoreover, we establish the layer-dependent magnetic properties for multilayer 1 T-CrTe 2and predict\nantiferromagnetism and ferromagnetism for bilayer and trilayer 1 T-CrTe 2, respectively. The MOE,\nAHE, and ANE are prohibited in antiferromagnetic bilayer 1 T-CrTe 2due to the existence of the\nspacetime inversion symmetry, whereas all of them are activated in ferromagnetic trilayer 1 T-CrTe 2\nand the MOE is signi\fcantly enhanced compared to monolayer 1 T-CrTe 2. Our results show that\nthe magneto-optical and anomalous transports proprieties of 1 T-CrTe 2can be e\u000bectively modulated\nby altering spin direction and layer number.\nI. INTRODUCTION\nAlthough two-dimensional (2D) materials have been\nexplored for more than a decade, the magnetic order\nrarely survives in atomically thin \flms due to thermal\n\ructuations [1]. The realization of 2D magnets is a big\nchallenge [2] and has attracted extensive attention [3, 4].\nThe 2D magnetic van der Waals (vdW) materials are es-\npecially expected to open up a wide range of possibilities\nfor spintronics [5{7]. Thanks to the improvement of theo-\nretical methods and experimental capabilities, more and\nmore 2D magnetic vdW materials have been discovered,\nwhich indicates that the \feld of 2D magnets is advancing\nrapidly [8]. In recent years, for example, tens of 2D vdW\nmaterials with stable magnetic orders have been observed\nin layered FePS 3[9, 10], Cr 2Ge2Te6[11], CrX3(X=I, Br,\nCl) [12{21], Fe 3GeTe 2[22{24],MX 2(M=V, Mn;X=Se,\nTe) [25{27], MnSn [28], PtSe 2[29], and CrTe 2[30{32].\nAimed at the applications of 2D spintronics, detect-\ning spontaneous magnetization is the primary step. The\nstandard techniques, such as superconducting quantum\ninterference device (SQUID) magnetometer and neu-\ntron scattering, are challenging to use for 2D magnetic\nvdW materials [3, 4]. Instead, the magneto-optical ef-\nfects (MOE), represented by the Kerr [33] and Fara-\nday [34] e\u000bects, are considered to be a powerful and\nnon-contact (non-destructive) probe of magnetism in 2D\nmaterials [11, 12]. The magneto-optical Kerr and Fara-\nday e\u000bects are de\fned as the rotation of the polarization\nplanes of re\rected and transmitted light beams when a\n\u0003wxfeng@bit.edu.cnlinearly polarized light hits the magnetic materials [35].\nIn condensed matter physics, the MOE and the anoma-\nlous Hall e\u000bect (AHE) [36], where the latter is character-\nized by a transverse voltage generated by a longitudinal\ncharge current in the absence of external magnetic \felds,\nare two fundamental phenomena that usually coexist in\nferromagnets and antiferromagnets. There are two dis-\ntinct contributions to the AHE, that is the extrinsic AHE\n(i.e., side jump and skew scattering) depending on the\nscattering of electron o\u000b impurities or due to disorder,\nand the intrinsic AHE expressed in term of Berry cur-\nvatures in a perfect crystal [36]. According to the Kubo\nformula [37, 38], the intrinsic anomalous Hall conduc-\ntivity (AHC) can be straightforwardly extended to the\noptical Hall conductivity, which is intimately related to\nthe magneto-optical Kerr and Faraday e\u000bects [39]. Be-\ncause of the inherent relationship between the intrinsic\nAHE and MOE, they are often studied together. More-\nover, the transverse charge current can also be generated\nby a longitudinal temperature gradient, called anoma-\nlous Nernst e\u000bect (ANE) [40], which has attracted enor-\nmous interest mainly due to its promising applications\non the thermoelectric aspects. The giant ANE has been\nrecently discovered in chiral magnets [41, 42] and topo-\nlogical semimetals [43, 44].\nThe catalog of 2D magnetic vdW materials is rich;\nhowever, the ferromagnetic candidates with high Curie\ntemperatures ( TC) are still limited, hindering enormously\nthe development of 2D spintronics. Fortunately, a 2D\ndichalcogenide with the 1 Tpolytype, 1 T-CrTe 2[see\nFigs. 1(a) and 1(b)], has been recently synthesized with\nan exceptionally high TC(>300 K) [30{32]. In this\nwork, based on the \frst-principles density functional the-\nory calculations and group theory analyses, we system-arXiv:2011.00394v1 [cond-mat.mtrl-sci] 1 Nov 20202\natically investigate the electronic, magnetic, magneto-\noptical, and anomalous charge and thermoelectric trans-\nports properties of monolayer and multilayer 1 T-CrTe 2\n(hereafter, we use CrTe 2for simpli\fcation). We \fnd\nthat monolayer CrTe 2is a ferromagnetic metal with the\nin-plane magnetization direction. By calculating mag-\nnetocrystalline anisotropy energy (MAE), the magneti-\nzation direction is \fnely identi\fed along the y-axis [see\nFig. 1(a)], and the maximal value of MAE between in-\nplane and out-of-plane magnetization directions reaches\nto 82.9\u0016eV/cell, which is much smaller than that of\nfamous 2D ferromagnets CrI 3(1.37 meV/cell) [45] and\nFe3GeTe 2(2.76 meV/cell) [46]. It indicates that the spin\ndirection of monolayer CrTe 2can be easily tuned by an\nexternal magnetic \feld. The MOE, AHE, and ANE dis-\nplay a period of 2 \u0019=3 by rotating the spin within the xy\nplane, and they disappear if there exists a mirror plane\nperpendicular to the spin direction. We then show that\nchanging the spin from in-plane to out-of-plane direction\ncan enhance the MOE, AHE, and ANE by around one\norder of magnitude. Additionally, the layer-dependent\nmagnetic properties for multilayer CrTe 2are studied, and\nantiferromagnetism and ferromagnetism for bilayer and\ntrilayer CrTe 2are predicted, respectively. For antiferro-\nmagnetic bilayer CrTe 2, the MOE, AHE, and ANE are\nfully suppressed due to the existence of the spacetime\ninversion symmetry TP(TandPare time-reversal and\nspatial inversion operations, respectively). In contrast,\nthe MOE, AHE, and ANE are activated in ferromagnetic\ntrilayer CrTe 2, and the MOE is signi\fcantly enhanced\ncompared to that of monolayer CrTe 2. Our results show\nthat the magneto-optical and anomalous transports pro-\nprieties of 2D CrTe 2are tunable by altering the magne-\ntization direction and the number of layers.\nII. METHODOLOGY\nThe \frst-principles calculations were performed by\nVienna ab initio simulation package ( vasp ) [47, 48]\nwithin the framework of density functional theory. The\nprojector augmented wave method (PAW) [49] was\nemployed to model the ion cores and the exchange-\ncorrelation functional of generalized gradient approxima-\ntion (GGA) with the Perdew-Burke-Ernzerhof parame-\nterization (PBE) [50] to simulate the valence electrons.\nSpin-orbit coupling was included in the calculations for\nthe MAE, MOE, AHE, and ANE. The plane-wave cut-\no\u000b energy was set to be 500 eV. The structures were\nrelaxed until the maximum force on each atom is less\nthan 0:0001 eV/ \u0017A and the energy convergence criterion\nis 10\u00007eV. The Brillouin zone integration was carried\nout by 16\u000216\u00021k-points sampling. A vacuum layer\nwith the thickness of at least 15 \u0017A was used to avoid the\ninteractions between adjacent layers and the vdW cor-\nrection was adopted by DFT-D2 method in multilayer\nstructures. The optical conductivity, AHC, and ANC are\nscaled by a factor of Z=d 0to exclude the vacuum region,whereZis the cell length normal to the atomic plane\nandd0= 6.23 \u0017A, 12.46 \u0017A, and 18.69 \u0017A are the e\u000bective\nthicknesses of monolayer, bilayer, and trilayer CrTe 2, re-\nspectively. Since the dorbitals of Cr atom are not fully\n\flled, the LDA+U method [51, 52] was used to account\nfor the Coulomb correlation with U = 2.0 eV [53].\nTo obtain the MOE, such as the Kerr and Faraday\nspectra, the optical conductivity should be primarily cal-\nculated. Here, we constructed the maximally localized\nWannier functions (MLWFs) in a non-self-consistent pro-\ncess by projecting onto s,p, anddorbitals of Cr atom as\nwell as onto sandporbitals of Te atom, using a uniform\nk-mesh of 16\u000216\u00021 points in conjunction with the wan-\nnier90 package [54]. The optical conductivity was then\ncalculated by integrating the dipole matrix elements (un-\nder the MLWFs basis) over the entire Brillouin zone using\na very dense k-points of 300\u0002300\u00021. The absorptive\nparts of optical conductivity are given by [37, 38, 55],\n\u001b1\nxx(!) =\u0015\n!X\nk;jj0[j\u0005+\njj0j2+j\u0005\u0000\njj0j2]\u000e(!\u0000!jj0);(1)\n\u001b2\nxy(!) =\u0015\n!X\nk;jj0[j\u0005+\njj0j2\u0000j\u0005\u0000\njj0j2]\u000e(!\u0000!jj0);(2)\nwhere the superscripts 1 and 2 indicate the real and imag-\ninary parts, \u0015=\u0019e2\n2~m2Vis a material speci\fc constant ( e\nandmare the charge and mass of an electron, ~is re-\nduced Planck constant, and Vis volume of unit cell),\njandj0denote occupied and unoccupied states at the\nsamek-point, \u0005\u0006\njj0are the dipole matrix elements rel-\nevant to right-circularly (+) and left-circularly ( \u0000) po-\nlarized lights, ~!is the photon energy, ~!jj0is the en-\nergy di\u000berence between jandj0states. Utilizing the\nKramers-Kronig transformation, the dispersive parts can\nbe obtained as,\n\u001b2\nxx(!) =\u00002\n\u0019PZ1\n0\u001b1\nxx(!0)\n!02\u0000!2d!0; (3)\n\u001b1\nxy(!) =2\n\u0019PZ1\n0!\u001b2\nxy(!0)\n!02\u0000!2d!0; (4)\nwherePis the principal integral.\nThe Kerr e\u000bect is characterized by the rotation angle\n(\u0012K) and ellipticity ( \"K), which are usually combined into\nthe complex Kerr angle,\n\u001eK=\u0012K+i\"K=i2!d\nc\u001bxy\n\u001bsxx; (5)\nwherecis the speed of light in vacuum, dis the thin-\n\flm thickness, and \u001bs\nxxis the optical conductivity of a\nnonmagnetic substrate. Similarly, the complex Faraday\nangle is given by,\n\u001eF=\u0012F+i\"F=i!d\n2c(n+\u0000n\u0000); (6)\nwheren2\n\u0006= 1 +4\u0019i\n!(\u001bxx\u0006i\u001bxy) are eigenvalues of di-\nelectric tensor. By considering the fact that j4\u0019i\n!(\u001bxx\u00063\n03060901\n201\n501\n802\n102\n402\n7030033003060901\n201\n501\n802\n102\n402\n70300330\nyx\ny1\n0 µeV(d)TeTe(a)y\n/s106\n (degree)abx\nCr(b)z\n8\n2.9 µeVyz\n(c)/s113\n (degree)\nFIG. 1. (Color online) (a,b) Top and side views of monolayer 1 T-CrTe 2. The blue spheres represent Cr atoms, whereas dark-\ngray and silver-white spheres represent Te atoms in the upper and lower sublayers. The pink dashed lines draw up the 2D\nprimitive cell, and the red arrows indicate the directions of spin magnetic moments. The top and bottom panels in (b) present\nthe spin directions along the y- andz-axis, respectively. (c,d) The magnetocrystalline anisotropy energy of monolayer 1 T-CrTe 2\nby rotating the spin magnetic moment within the yzandxyplanes. The spin along the y-axis is set to be the reference state.\ni\u001bxy)j\u001c1, the complex Faraday angle can be approxi-\nmately written as,\n\u001eF=\u0012F+i\"F'\u00002\u0019d\nc\u001bxy; (7)\nFrom Eqs. (5) and (7), one can see that the o\u000b-diagonal\nelements of optical conductivity ( \u001bxy), also known as the\noptical Hall conductivity, is determinative to both Kerr\nand Faraday e\u000bects. It should be mentioned here that\nEqs. (5){(7) are the expressions for 2D systems with a\npolar geometry [56], that is, the incident light propagates\nalong the\u0000zdirection.\nPhysically speaking, the MOE is closely related to the\nAHE. For example, the dc limit of the real part of the\no\u000b-diagonal element of optical conductivity, i.e., \u001b1\nxy(!!\n0), is nothing but the intrinsic AHC, which can also be\ncalculated from the Berry-phase formula [57],\n\u001bA\nxy=\u0000e2\n~VX\nn;kfnk\nn\nxy(k); (8)\nwheren,k, andfnkare band index, crystal momen-\ntum, and Fermi-Dirac distribution function, respectively.\nn\nxy(k) is the band-resolved Berry curvature, given by,\n\nn\nxy(k) =\u0000X\nn06=n2Im[h nkj^vxjh n0kih n0kj^vyjh nki]\n(\"nk\u0000\"n0k)2\n(9)\nwhere ^vx;yis the velocity operator along the xorydi-\nrection, nkand\"nkare the eigenvector and eigenvalue\nat band index nand crystal momentum k, respectively.\nThe intrinsic anomalous Nernst conductivity (ANC) can\nbe written as [58, 59]\n\u000bA\nxy=e\n~TVX\nn;k\nn\nxy(k)\u0002[(\"nk\u0000\u0016)fnk\n+kBTln(1 +e\u0000(\"nk\u0000\u0016)=kBT)] (10)\nwhereT,\u0016, andkBis temperature, chemical potential,\nand Boltzmann constant, respectively. Thus, the ANC\ncan be related to the AHC by the Mott formula [58].4\n-2-1012-\n2-1012-50 5 0\n5 E (eV) ↑ \n↓Γ\nK M Γ (a)S\n || yE (eV)Γ\nK M Γ (b) ↑ \nTotal \nCr 3d \nTe 5p↓ \nFIG. 2. (Color online) (a) The spin-polarized band struc-\nture and density of states (in the unit of states/eV/cell)\nfor monolayer 1 T-CrTe 2. (b) The relativistic band struc-\nture and orbital-decomposed density of states (in the unit\nof states/eV/cell) for monolayer 1 T-CrTe 2when the spin is\nalong they-axis.\nIII. RESULTS AND DISCUSSION\nIn this section, we successively present the results of\nmonolayers and multilayer CrTe 2. The magnetic ground\nstates are \frst established by calculating the MAE. The\ncorresponding electronic band structures are explicitly\ncalculated. The magnetic group theory is then used to\ndetermine the nonzero elements of optical conductivity,\nwhich is the critical ingredient to evaluate the magneto-\noptical Kerr and Faraday spectra. Finally, the anomalous\nHall and anomalous Nernst conductivities are evaluated\nby using the Berry-phase formulas. The dependence of\nthe MOE, AHE, and ANE on the magnetization direction\nand layer number will be detailedly discussed.\nA. Monolayer CrTe 2\n1. Crystal, magnetic, and electronic structures\nThe top and side views of monolayer CrTe 2(space\ngroup P 3m1, No. 164) are depicted in Figs. 1(a) and 1(b).\nEach primitive cell contains one Chromium (Cr) atomand two Tellurium (Te) atoms, forming a sandwich struc-\nture Te-Cr-Te. The optimized lattice constant of mono-\nlayer CrTe 2isa= 3:722\u0017A.\nTo con\frm the magnetic ground state, we compared to-\ntal energies among the nonmagnetic, antiferromagnetic,\nand ferromagnetic states using a supercell of 2 \u00022\u00021,\nand the results show that ferromagnetic state is more\nstable than nonmagnetic and antiferromagnetic states\nby 2:90 eV and 55 :32 meV, respectively. Additionally,\nthe magnetic ground state can be determined via the\nsuper-exchange mechanism [60{62]. The magnetic ex-\nchange interactions depend on the \flling of the dorbitals\nof the cations and on the angle formed by the chemical\nbonds connecting the ligand and magnetic atoms, and\nparticularly, when the angle equals to 90\u000ethe ferromag-\nnetic interactions are optimal. In the case of CrTe 2, the\nbond angle between Cr-Te-Cr is 87\u000e, which accounts for\nferromagnetic interactions. Furthermore, the MAE, de-\n\fned as MAE( \u0012;') =E(\u0012;')\u0000E(\u0012= 90\u000e;'= 90\u000e)\n[here,E(\u0012;') is the total energy when the spin mag-\nnetic moment ( S) orients to the polar angle \u0012and az-\nimuthal angle '], is computed by rotating the spin mag-\nnetic moment on the xyandyzplanes, respectively, as\nshown in Figs. 1(c) and 1(d). The positive values of\nMAE suggest a preferred magnetization along the y-axis\n(\u0012= 90\u000e;'= 90\u000e) rather than along other directions.\nFigure 1(c) shows that the out-of-plane magnetization\n(along thezaxis) is not prior due to the positive MAE\nof 82:9\u0016eV/cell. Figure 1(d) further indicates a small in-\nplane magnetocrystalline anisotropy (0 \u0016eV/cell\u0014MAE\n\u001410\u0016eV/cell), in good agreement with experimental ob-\nservation [31, 32]. Therefore, the magnetic ground state\nof the system is con\frmed and the spin magnetic moment\nshould be along the y-axis, i.e.,S(\u0012= 90\u000e;'= 90\u000e) or\nSky[see top panel of Fig. 1(b)], which is consistent\nwith previously theoretical calculation [63]. Thus, the\nspin direction of CrTe 2can be easily tuned by applying\nan external magnetic \feld as the maximal MAE of CrTe 2\n(82:9\u0016eV/cell) is much smaller than that of Fe 3GeTe 2\n(2:76 meV/cell) [46], which has been realized experimen-\ntally. This provides a technical basis for us to reorient\nthe spin magnetic moment from in-plane to out-of-plane\ndirection [see bottom panel of Fig. 1(b)].\nWe then discuss the electronic structures of monolayer\nCrTe 2. Fig. 2(a) plots the spin-polarized band structures\nand density of states, in which the red and blue lines rep-\nresent the spin-up ( \") and spin-down ( #) bands, respec-\ntively. The spin-polarized band structures combined with\ndensity of states clearly show that monolayer CrTe 2is a\nferromagnetic metal. After including spin-orbit coupling,\nthe relativistic band structures and orbital-decomposed\ndensity of states with the magnetization Skyare illus-\ntrated in Fig. 2(b). The band structure is good consistent\nwith a recently theoretical calculation [64]. For the den-\nsity of states, we only present the dominant components,\ni.e., the 3dorbitals of Cr atom (the orange pattern) and\n5porbitals of Te atoms (the green pattern), which have\nnearly equal contributions around the Fermi energy.5\nTABLE I. The magnetic space group (MSG) and magnetic point group (MPG) of monolayer 1 T-CrTe 2as a function of azimuthal\n(') and polar ( \u0012) angles when the spin rotates within the xy(\u0012=\u0019=2, 0\u0014'\u0014\u0019) andyz(0\u0014\u0012\u0014\u0019,'=\u0019=2) planes.\n0\u000e15\u000e30\u000e45\u000e60\u000e75\u000e90\u000e105\u000e120\u000e135\u000e150\u000e165\u000e180\u000e\nMSG(')C2=m P \u00161C20=m0P\u00161C2=m P \u00161C20=m0P\u00161C2=m P \u00161C20=m0P\u00161C2=m\nMPG(') 2=m \u00161 20=m0 \u00161 2=m \u00161 20=m0 \u00161 2=m \u00161 20=m0 \u00161 2=m\nMSG(\u0012)P\u00163m01C20=m0C20=m0C20=m0C20=m0C20=m0C20=m0C20=m0C20=m0C20=m0C20=m0C20=m0P\u00163m01\nMPG(\u0012)\u001631m020=m020=m020=m020=m020=m020=m020=m020=m020=m020=m020=m0 \u001631m0\n2. Magnetic group theory\nThe group theory is a powerful tool for identifying the\nnonvanishing elements of the optical Hall conductivity,\nwhich is the key factor in predicting the MOE. Addi-\ntionally, the AHE and ANE have the same symmetry\nrequirements with the MOE due to their physical rela-\ntions [refer to Eqs. (1){(10)]. Hence, we take the optical\nHall conductivity as an example, and the results of sym-\nmetry analyses are applicable to the MOE, AHE, and\nANE. The magnetic space and point groups for mono-\nlayer CrTe 2are calculated by using the isotropy soft-\nware [65]. Table I lists the results when the spin rotates\nwithin thexyandyzplanes. Since the optical Hall con-\nductivity is translationally invariant, it is su\u000ecient to\nrestrict the analysis to magnetic point group. Moreover,\nthe vector-form notation of the optical Hall conductiv-\nity, given by \u001b(!) = [\u001bx;\u001by;\u001bz] = [\u001byz;\u001bzx;\u001bxy], is used\nfor convenience as it can be regarded as a pseudovector,\njust like spin. Thus, for a 2D system, there always has\n\u001bx=\u001by= 0 and only \u001bzis potentially nonzero.\nLet us start with the situation that rotating the spin\nwithin the xyplane. The magnetic point group has a\nperiod of\u0019=3: 2=m!\u00161!20=m0!\u00161!2=m, and three\nnonrepetitive elements are 2 =m,\u00161, and 20=m0. First, the\ngroup 2=m(when'=n\u0019=3 withn2N) has a mirror\nplane that is parallel to the z-axis and is perpendicular to\nthe spin direction. Such a mirror operation reverses the\nsign of\u001bz, and thus indicating \u001bz= 0. It results in the\nvanishing optical Hall conductivity, i.e., \u001b(!) = [0;0;0].\nOn the other hand, all mirror symmetries are broken if\n'6=n\u0019=3. The group 20=m0contains a combined sym-\nmetryTM, whereTis the time-reversal symmetry and\nMis a mirror plane that parallels to both the z-axis and\nspin direction. Both TandMoperations reverse the\nsign of\u001bz, and hence \u001bzis even underTM symmetry.\nIt gives rise to the nonvanishing optical Hall conductivity,\n\u001b(!) = [0;0;\u001bz]. Finally, for the group \u00161 =fE;Pg, none\nof its elements (unit operation Eand spatial inversion P)\ncan a\u000bect\u001bz, and hence the optical Hall conductivity are\nabsolutely allowed.\nWe next turn to the case that the spin lies within the\nyzplane. The evolution of magnetic point group exhibits\na period of \u0019, and only two groups \u001631m0and 20=m0are\nneeded to analyze. If \u0012= 0 or\u0019, the group \u001631m0contains\nthreeTM symmetries with Mkz. The nonvanishingoptical Hall conductivity can be expected since any one\nof the threeTM symmetries a\u000bords \u001bz6= 0. Once the\nspin cants away from the z-axis (\u00126= 0) or from the \u0000z-\naxis (\u00126=\u0019), the magnetic point group changes to be\n20=m0, in which only one of the three TM symmetry\nleaves (here,Mis just the yzplane) but still ensures\n\u001bz6= 0. To summarize, the optical Hall conductivity is\nnonzero when the spin lies within the yzplane, that is,\n\u001b(!) = [0;0;\u001bz].\n3. Optical and magneto-optical properties\nAfter obtaining the electronic structures and magnetic\ngroups of monolayer CrTe 2with di\u000berent magnetization\ndirections, we now focus on the optical conductivity,\nwhich is prerequisite to evaluate the MOE.\nWe \frst discuss the results of in-plane magnetization\nwhenS(90\u000e;0\u000e) (Skx),S(90\u000e;30\u000e), andS(90\u000e;90\u000e)\n(Sky), shown in Figs. 3(a-d). According to Eqs. (1)\nand (2), the absorptive parts of optical conductivity, \u001b1\nxx\nand\u001b2\nxy, have direct physical interpretations, which mea-\nsures the average and di\u000berence in absorptions of the left-\nand right-circularly polarized light, respectively. The \u001b1\nxx\nplotted in Fig. 3(a) exhibits two sharp absorption peaks\nat 0.6 and 2.2 eV. Since \u001b1\nxxis directly related to the\ninterband transition probability and jointed density of\nstates, it is not a\u000bected by the spin direction, similarly\nto Mn 3XN (X= Ga, Zn, Ag, or Ni) [39]. On the other\nhand, the\u001b2\nxyplotted in Fig. 3(d) oscillates drastically\nin the low-energy region and tends to zero above 5.5 eV.\nThe positive and negative values of \u001b2\nxyindicate that\nthe interband transitions are dominated by the excita-\ntions caused by the left- and right-circularly polarized\nlight, respectively. The signs of \u001b2\nxyfor the states of\nS(90\u000e;30\u000e) andS(90\u000e;90\u000e) are opposite, which has the\nsame physical mechanism of the intrinsic AHC for mono-\nlayer LaCl [66]. It should be further noticed that for the\nstate ofS(90\u000e;0\u000e),\u001b2\nxyis suppressed due to the pres-\nence of the mirror plane Mthat is perpendicular to S,\nwhich is consistent with the previous group theory anal-\nyses. Utilizing the Kramers-Kronig transformation, the\ndispersive parts of optical conductivity, \u001b2\nxxand\u001b1\nxy, can\nbe obtained from the corresponding absorptive parts ac-\ncording to Eqs. (3) and (4). The dependence of \u001b2\nxxand\n\u001b1\nxyon the magnetization direction, featured in Figs. 3(b)6\n02 4 6 050\n2 4 6 -20240\n2 4 6 -0.050.000.050\n2 4 6 -0.10.00.10\n2 4 6 050\n2 4 6 -20240\n2 4 6 -0.20.00.20\n2 4 6 0.00.5Photon Energy (eV)/s1151x\nx (1015 s-1)S\n || xS\n (90°, 30°)S\n || y(a)P\nhoton Energy (eV)/s1152x\nx (1015 s-1)(b)P\nhoton Energy (eV)/s1151x\ny (1015 s-1)(c)P\nhoton Energy (eV)/s1152x\ny (1015 s-1)(d)P\nhoton Energy (eV)/s1151x\nx (1015 s-1)S\n || yS\n (45°, 90°)S\n || z(e)P\nhoton Energy (eV)/s1152x\nx (1015 s-1)(f)/s115\n1x\ny (1015 s-1)P\nhoton Energy (eV)(g)/s115\n2x\ny (1015 s-1)P\nhoton Energy (eV)(h)\nFIG. 3. (Color online) The real diagonal (a,e), imaginary diagonal (b,f), real o\u000b-diagonal (c,g), and imaginary o\u000b-diagonal\n(d,h) elements of optical conductivity for monolayer 1 T-CrTe 2with in-plane and out-of-plane magnetization, respectively. For\na better comparison, the curves when spin points along y-axis (Sky) are replotted in (e-h).\n02 4 6 -0.20.00.20\n2 4 6 -0.50.00.50\n2 4 6 -2020\n2 4 6 -2020\n6 0120180-0.30.00.30\n6 0120180-0.60.00.60\n6 0120180-3030\n6 0120180-404Photon Energy (eV)/s113K (deg) \nS || x \nS (90°, 30°) \nS || y(a)P\nhoton Energy (eV)/s101K (deg)(b)P\nhoton Energy (eV)/s113F (105 deg/cm)(c)P\nhoton Energy (eV)/s101F (105 deg/cm)(d)(\ne)/s113 K (deg)/s106\n (deg)0.70 eV(f)/s101K (deg)/s106\n (deg)0.60 eV(g)0\n.79 eV/s113F (105 deg/cm)/s106\n (deg)(h)0\n.62 eV/s101F (105 deg/cm)/s106\n (deg)\nFIG. 4. (Color online) The Kerr rotation angle \u0012K(a), Kerr ellipticity \"K(b), Faraday rotation angle \u0012F(c), and Faraday\nellipticity\"F(d) for monolayer 1 T-CrTe 2with in-plane magnetization. (e-h) The Kerr and Faraday rotation angles and\nellipticities as a function of azimuthal angle 'at selected photon energies.\nand 3(c), resemble that of \u001b1\nxxand\u001b2\nxy.\nThen, we proceed to the out-of-plane magnetization\nby considering the spin within the yzplane, for example\nS(0\u000e;90\u000e) (Skz) andS(45\u000e;90\u000e). As shown in Figs. 3(e)\nand 3(f),\u001b1\nxxhas two absorption peaks at 0.6 and 2.2 eV,\nand meanwhile \u001b2\nxxpresents two valleys at 0.5 and 2.0\neV. This is identical to the situation of in-plane magne-\ntization and further indicates that the diagonal elements\nof optical conductivity are not a\u000bected by the spin di-\nrection. In contrast, the o\u000b-diagonal elements of optical\nconductivity, \u001b1\nxyand\u001b2\nxy[see Figs. 3(g) and 3(h)], obvi-\nously depend on the spin direction. \u001b1\nxyand\u001b2\nxyoscillate\nas a function of photon energy with di\u000berent spin direc-tions and reach the maximal values when the spin points\ntowards the z-axis. It is important to notice that the o\u000b-\ndiagonal elements of optical conductivity with the out-of-\nplane magnetization are enhanced by about one order of\nmagnitude compared to that of in-plane magnetization.\nNow, we present the magneto-optical Kerr and Fara-\nday spectra with in-plane and out-of-plane magnetiza-\ntion, as shown in Figs. 4 and 5, respectively. The Kerr\nand Faraday spectra are rather similar to that of \u001bxy,\nand the reason can be simply attributed to their close\nrelationships [refer to Eqs. (5) and (7)]. For the in-plane\nmagnetization, the Kerr and Faraday angles are vanish-\ning if the spin points along '=n\u0019=3, for example '= 0\u000e7\n02 4 6 -1010\n2 4 6 -2020\n2 4 6 -20-100100\n2 4 6 -20-100100\n9 0180270360-1010\n9 0180270360-2020\n9 0180270360-20-10010200\n9 0180270360-20-1001020Photon Energy (eV)/s113K (deg) \nS || y \nS (45°, 90°) \nS || z(a)P\nhoton Energy (eV)/s101K (deg)(b)/s113\nF (105 deg/cm)P\nhoton Energy (eV)(c)/s101\nF (105 deg/cm)P\nhoton Energy (eV)(d)/s113K (deg)/s113\n (deg)0.66 eV(e)( f)0\n.88 eV/s101K (deg)/s113\n (deg)(g)0\n.82 eV/s113F (105 deg/cm)/s113\n (deg)(h)0\n.91 eV/s101F (105 deg/cm)/s113\n (deg)\nFIG. 5. (Color online) The Kerr rotation angle \u0012K(a), Kerr ellipticity \"K(b), Faraday rotation angle \u0012F(c), and Faraday\nellipticity\"F(d) for monolayer 1 T-CrTe 2with out-of-plane magnetization. For a better comparison, the curves for the spin\npointing along the y-axis (Sky) are also plotted in (a-d). (e-h) The Kerr and Faraday rotation angles and ellipticities as a\nfunction of polar angle \u0012at selected photon energies.\n09 01 80270360-4-20240\n9 01 80270360-10010-10 -4-2024-\n10 -10-505101506 01 201 80-1010\n6 01 201 80-505(c)/s115Ax\ny (×103 S/cm)/s113\n (deg)-1.03 eVT\n = 300K(f)/s97\nAx\ny (A/mK)/s113\n (deg)-1.10 eV/s115Ax\ny (×103 S/cm)E\n (eV) S || x \nS || y \nS || z(a) \nS || x \nS || y \nS || zT\n = 300K/s97Ax\ny (A/mK)E\n (eV)(d)-1.13 eV/s115Ax\ny (×103 S/cm)/s106\n (deg)(b)T\n = 300K-1.34 eV/s97Ax\ny (A/mK)/s106\n (deg)(e)\nFIG. 6. (Color online) (a,d) The intrinsic anomalous Hall ( \u001bA\nxy) and anomalous Nernst ( \u000bA\nxy) conductivity for monolayer 1 T-\nCrTe 2as a function of the Fermi energy when the spin points along the x-,y-, andz-axis.\u000bA\nxyis calculated at the temperature\nof 300 K. The black arrows indicates the maximal values of \u001bA\nxyand\u000bA\nxy. (b,e)\u001bA\nxyand\u000bA\nxyas a function of azimuthal angle '\nwhen the Fermi energies are set to be -1.13 and -1.34 eV, respectively. (c,f) \u001bA\nxyand\u000bA\nxyas a function of polar angle \u0012when\nthe Fermi energies are set to be -1.03 and -1.10 eV, respectively.\n[see Figs. 4(a-d)], due to the symmetry restriction. When\nthe spin rotates to '=\u0019=6+n\u0019=3, the Kerr and Faraday\nangles reach their maximums, that is, \u0012max\nK= 0:24 deg\nand\u0012max\nF= 3:00\u0002105deg/cm at the photon energies of\n0.70 and 0.79 eV, respectively. The \u0012max\nKof CrTe 2is com-\nparable with the Kerr rotation angles of monolayer CrI 3\n(0.286 deg) [12] and of blue phosphorene (0.12 deg) [67].\nMoreover, Figs. 4(e-h) show that the maximal Kerr andFaraday angles of monolayer CrTe 2exhibits a period of\n2\u0019=3 when the spin rotates within the xyplane.\nOn the other hand, the Kerr and Faraday spectra with\nthe out-of-plane magnetization are illustrated in Fig. 5.\nInheriting from the o\u000b-diagonal elements of optical con-\nductivity, the Kerr and Faraday spectra with out-of-plane\nmagnetization are signi\fcantly stronger than that with\nin-plane magnetization [Figs. 5(a-d)]. The maximal Kerr8\n(a)( b)( c)( d)(\ne)( f)( g)( h)yzz\ny\nCrTeT\ne\nFIG. 7. (Color online) Side views of bilayer (a-d) and trilayer (e-h) 1 T-CrTe 2with in-plane and out-of-plane ferromagnetic and\nantiferromagnetic con\fgurations. The red arrows label the spin directions.\nand Faraday rotation angles appear when \u0012= 0\u000e(Skz),\nthat is,\u0012max\nK =\u00001:34 deg and \u0012max\nF =\u000017:30\u0002105\ndeg/cm at the photon energies of 0.66 and 0.82 eV, re-\nspectively. Moreover, the Kerr and Faraday angles have\na period of 2 \u0019as a function of polar angle \u0012, as shown in\nFigs. 5(e-h), demonstrating again that the MOE can be\ne\u000bectively modulated by tuning the spin direction.\n4. Anomalous Hall and anomalous Nernst e\u000bects\nAs mentioned above, the dc limit of real o\u000b-diagonal el-\nement of optical conductivity, i.e., \u001b1\nxy(!!0), is nothing\nbut the AHC ( \u001bA\nxy), which can be alternatively evaluated\nby integrating the Berry curvature over the entire Bril-\nlouin zone [see Eq. (8)] [57]. Figure 6(a) plots the AHC\nas a function of the Fermi energy when the spin points\nalong thex-,y-, andz-axis.\u001bA\nxyis vanishing when Skx\nand turns to appear if Skyandz, indicating the same\nsymmetry requirements for the MOE. In analogy to the\nKerr and Faraday angles, the \u001bA\nxywith out-of-plane mag-\nnetization ( Skz) is signi\fcantly larger than that with in-\nplane magnetization ( Sky). Thus, at the actual Fermi\nenergy, the \u001bA\nxywith both in-plane and out-of-plane mag-\nnetization are relatively small, which is adverse to mea-\nsure experimentally. Nevertheless, the pronounced peaks\nof AHC arise after appropriate holes are introduced. Forexample,\u001bA\nxycan increase up to -4539.23 S/cm at -1.03\neV whenSkzand up to 1168.12 S/cm at -1.13 eV when\nSky, respectively. When the spin rotates within the xy\nandyzplanes,\u001bA\nxyexhibits the periods of 2 \u0019=3 and 2\u0019,\nrespectively [depicted in Figs. 6(b) and 6(c)], which are\nidentical to the behaviors of Kerr and Faraday angles.\nThe ANE, being regarded as the thermoelectric coun-\nterpart of the AHE, is a celebrated e\u000bect from the realm\nof spin caloritronics [68, 69]. The conclusions of symme-\ntry analyses for the AHE are also applicable to the ANE,\ncomparing with Eqs. (8) and (10). That is, the ANC \u000bA\nxy\nis forbidden when the spin is along '=n\u0019=3, e.g.,Skx,\nand turns to be nonzero if Skyorz, as clearly shown in\nFig. 6(d). Due to the high Curie temperature of CrTe 2\n(TC>300 K) [30{32], the \frst-principles calculations\nof the ANC are carried out at the room-temperature of\n300 K. Similarly to the AHC \u001bA\nxy, the ANC \u000bA\nxywith\nout-of-plane magnetization ( Skz) is evidently larger\nthan that with in-plane magnetization ( Sky). For both\nin-plane and out-of-plane magnetization, \u000bA\nxyare almost\nzero at the actual Fermi energy and give rise to pro-\nnounced peaks by hole doping. For example, \u000bA\nxyreaches\nup to 13.61 A/mK at -1.10 eV when Skzand up to 5.30\nA/mK at -1.34 eV when Sky, respectively. Moreover,\n\u000bA\nxydisplays a period of 2 \u0019=3 (2\u0019) when the spin rotates\nwithin thexy(yz) plane, as shown in Figs. 6(e) and 6(f),\njust like the AHC as well the Kerr and Faraday angles.9\n-2-1012-2-1012E (eV) S || yT\nrilayerΓ\nK M Γ (b)E (eV) S || yB\nilayerΓ\nK M Γ (a)\nFIG. 8. (Color online) Relativistic band structures of bilayer\nand trilayer 1 T-CrTe 2when the spin is along the y-axis.\nB. Multilayer CrTe 2\nIn this subsection, we shall discuss the layer-dependent\nmagnetic properties as well as the MOE, AHE, and ANE\nof multilayer CrTe 2. The bilayer and trilayer CrTe 2with\nthe AA-stacking pattern, which could be directly exfo-\nliated from the bulk structure, are considered here. To\ndetermine the magnetic ground states, we calculate the\ntotal energy ( Etot) of in-plane and out-of-plane ferro-\nmagnetic and antiferromagnetic structures, as depicted\nTABLE II. The total energy Etot(in the unit of eV) per unit\ncell of bilayer (BL) and trilayer (TL) 1 T-CrTe 2with in-plane\nand out-of-plane ferromagnetic (i-FM and o-FM) and antifer-\nromagnetic (i-AFM and o-AFM) con\fgurations. The super-\nscriptsa\u0000hcorresponds to the magnetic structures presented\nin Fig. 7. The energies of nonmagnetic (NM) structures are\ngiven for the reference. The relaxed lattice constant a(in the\nunit of \u0017A) is also listed.\ni-FMa;eo-FMc;gi-AFMb;fo-AFMd;hNM\nBLEtot -31.996 -31.983 -32.021 -32.017 -28.885\na 3.763 3.754 3.784 3.786 3.469\nTLEtot -48.216 -48.208 -48.238 -48.237 -43.528\na 3.778 3.778 3.791 3.793 3.478in Fig. 7. It should be stressed here that all the trilayer\nCrTe 2are actually ferromagnetic with \fnite net magne-\ntization, while we mention the \\antiferromagnetic\" tri-\nlayer structures [Figs. 7(f,h)] just because of the inter-\nlayer antiferromagnetic order. The energy results are\nsummarized in Tab. II, from which one can \fnd that\nthe antiferromagnetic structures for both bilayer and tri-\nlayer CrTe 2[Figs. 7(b,d) and 7(f,h)] are energetically fa-\nvorable. Thus, the in-plane antiferromagnetic structures\n[Figs. 7(b) and 7(f)] are most stable with the slightly\nlower energies of \u00184 and\u00181 meV/cell than the out-of-\nplane ones for bilayer and trilayer, respectively. In the\nfollowing, we only focus on the bilayer and trilayer CrTe 2\nwith the in-plane antiferromagnetic structures. The elec-\ntronic band structures plotted in Fig. 8 demonstrate the\nmetallic nature of both bilayer and trilayer CrTe 2.\nUsing the group theory, we analyze whether the MOE,\nAHE, and ANE can exist in bilayer and trilayer CrTe 2.\nThe magnetic point group of in-plane antiferromagnetic\nbilayer [Fig. 7(b)] is 2 =m0, which contains the space-\ntime inversion symmetry TPthat forbids any signals of\nmagneto-optical responses as well as anomalous charge\nand thermoelectric transports. In contrast, the mag-\nnetic point group of in-plane antiferromagnetic trilayer\n[Fig. 7(f)] is the same as that of monolayer CrTe 2, i.e.,\n20=m0, which allows the presence of the MOE, AHE, and\nANE.\nThe layer number can in\ruence the MOE, AHE, and\nANE of multilayer CrTe 2. The magneto-optical Kerr and\nFaraday rotation angles of in-plane antiferromagnetic bi-\nlayer and trilayer CrTe 2are plotted in Figs. 9(a) and 9(b),\nin which the results of monolayer CrTe 2are given for\ncomparison. As expected, the \u0012Kand\u0012Fof bilayer struc-\nture are zero due to the presence of TPsymmetry. For\nthe trilayer structure, the largest \u0012Kand\u0012Fare -1.76\ndeg and 4.60\u0002105deg/cm at the photon energies of 0.38\nand 3.04 eV, respectively. One can \fnd that the Kerr\nand Faraday e\u000bects of trilayer structure are generally\nstronger than that of monolayer structure. Moreover, the\nAHC and ANC of monolayer, bilayer, and trilayer CrTe 2\nwith in-plane magnetization are presented in Figs. 9(c)\nand 9(d), respectively. Clearly, the \u001bA\nxyand\u000bA\nxyof bilayer\nstructure are vanishing due to the symmetry restriction.\nAlthough the \u001bA\nxyand\u000bA\nxyof trilayer structure are very\nsmall at the actual Fermi energy, they can be signi\fcantly\nenhanced by hole doping. The largest \u001bA\nxyand\u000bA\nxyare\n1147.03 S/cm at -1.18 eV and -4.81 A/mK at -1.12 eV,\nrespectively. The AHE and ANE of trilayer structure are\nslightly smaller than that of monolayer structure.\nIV. SUMMARY\nIn summary, using the \frst-principles density func-\ntional theory calculations and group theory analyses, we\nhave systematically investigated the electronic, magnetic,\nmagneto-optical, anomalous charge and thermoelectric\ntransport properties of monolayer, bilayer, and trilayer10\n02 4 6 -2-1010\n2 4 6 -404-10 -2-101-\n10 -10-50510Photon Energy (eV)/s113K (deg) \nMonolayer \nBilayer \nTrilayer(a)(\nb)/s113 F (105 deg/cm)P\nhoton Energy (eV)(c)E\n (eV)/s115Ax\ny (×103 S/cm)(\nd)/s97 Ax\ny (A/mK)E\n (eV)\nFIG. 9. (Color online) The Kerr (a) and Faraday (b) rotation angles of bilayer and trilayer 1 T-CrTe 2. The anomalous Hall\nand anomalous Nernst conductivities of bilayer and trilayer 1 T-CrTe 2as a function of the Fermi energy. The magnetization\ndirection is along the y-axis. For a better comparison, the curves of monolayer 1 T-CrTe 2are also plotted.\n1T-CrTe 2. The monolayer is a ferromagnetic metal with\nthe in-plane magnetization along the y-axis. The in-plane\nmagnetocrystalline anisotropy energy is as small as 10\n\u0016eV/cell, indicating that the spin can be easily rotated\nwithin thexyplane. The magneto-optical Kerr and Fara-\nday rotation angles as well as anomalous Hall and Nernst\nconductivities exhibit a period of 2 \u0019=3 when the spin ro-\ntates within the xyplane, and their maximums of \u0012K=\n0.24 deg,\u0012F= 3.00\u0002105deg/cm,\u001bA\nxy= 1168.12 S/cm,\nand\u000bA\nxy= 5.30 A/mK (300 K) appear at '=n\u0019=3+\u0019=6\nwithn2N. At the azimuthal angle '=n\u0019=3, the\nmirror planes that are normal to the spin direction sup-\npress the magneto-optical, anomalous Hall, and anoma-\nlous Nernst e\u000bects. If the spin cants from in-plane to out-\nof-plane direction, the magneto-optical, anomalous Hall,\nand anomalous Nernst e\u000bects are signi\fcantly enhanced,\nand particularly they reach to the maximal values of \u0012K\n= -1.34 deg, \u0012F= -17.30\u0002105deg/cm,\u001bA\nxy= -4539.23\nS/cm, and \u000bA\nxy= 13.61 A/mK (300 K) when the spin is\nalong thez-axis (i.e., polar angle \u0012= 0\u000e). The bilayer\n1T-CrTe 2prefers an in-plane antiferromagnetic structure\nwith the magnetization along the y-axis, which has the\nspacetime inversion symmetry TPthat prohibits the sig-\nnals of magneto-optical responses as well as anomalous\nHall and Nernst transports. The trilayer 1 T-CrTe 2is also\ninclined to the in-plane antiferromagnetic order between\ntwo adjacent layers, but has \fnite net magnetization dueto the odd number of layers. The magnetic point group of\ntrilayer structure with in-plane antiferromagnetic order is\nidentical to that of monolayer structure and thus allows\nthe presence of all the physical phenomena mentioned\nabove. In particular, the magneto-optical Kerr and Fara-\nday rotation angles (anomalous Hall and Nernst conduc-\ntivities) of trilayer structure are obviously larger (slightly\nsmaller) than that of monolayer structure with the mag-\nnetization along the y-axis. For example, the maximal\nvalues of\u0012K= -1.76 deg, \u0012F= 4.60\u0002105deg/cm,\u001bA\nxy=\n-1147.03 S/cm, and \u000bA\nxy= -4.81 A/mK (300 K) are found\nin the trilayer structure . Our results suggest that the\nmagneto-optical, anomalous Hall, and anomalous Nernst\ne\u000bects for two-dimensional room-temperature ferromag-\nnet 1T-CrTe 2can be e\u000bectively modulated by altering\nmagnetization direction and layer number.\nACKNOWLEDGMENTS\nW.F. and Y.Y. acknowledge the support from the Na-\ntional Natural Science Foundation of China (Grants No.\n11874085 and No. 11734003) and the National Key R&D\nProgram of China (Grant No. 2016YFA0300600). X.Z.\nacknowledges the support from the Graduate Technolog-\nical Innovation Project of Beijing Institute of Technology\n(Grant No. 2019CX10018).11\n[1] C. Gong and X. Zhang, Two-dimensional magnetic crys-\ntals and emergent heterostructure devices, Science 363,\n706 (2019).\n[2] N. D. Mermin and H. Wagner, Absence of ferromag-\nnetism or antiferromagnetism in one-or two-dimensional\nisotropic Heisenberg models, Phys. Rev. Lett. 17, 1133\n(1966).\n[3] F. Hellman, A. Ho\u000bmann, Y. Tserkovnyak, G. S. D.\nBeach, E. E. Fullerton, C. Leighton, A. H. MacDon-\nald, D. C. Ralph, D. A. Arena, H. A. D urr, P. Fischer,\nJ. Grollier, J. P. Heremans, T. Jungwirth, A. V. Kimel,\nB. Koopmans, I. N. Krivorotov, S. J. May, A. K. Petford-\nLong, J. M. Rondinelli, N. Samarth, I. K. Schuller, A. N.\nSlavin, M. D. Stiles, O. Tchernyshyov, A. Thiaville, and\nB. L. Zink, Interface-induced phenomena in magnetism,\nRev. Mod. Phys. 89, 025006 (2017).\n[4] K. S. Burch, D. Mandrus, and J.-G. Park, Magnetism\nin two-dimensional van der Waals materials, Nature 563,\n47 (2018).\n[5] B. Sachs, T. O. Wehling, K. S. Novoselov, A. I. Licht-\nenstein, and M. I. Katsnelson, Ferromagnetic two-\ndimensional crystals: single layers of K 2CuF 4, Phys. Rev.\nB88, 201402(R) (2013).\n[6] J.-G. Park, Opportunities and challenges of two-\ndimensional magnetic van der Waals materials: magnetic\ngraphene?, J. Phys.: Condens. Matter 28, 301001 (2016).\n[7] D. Zhong, K. L. Seyler, X. Linpeng, R. Cheng,\nN. Sivadas, B. Huang, E. Schmidgall, T. Taniguchi,\nK. Watanabe, M. A. McGuire, W. Yao, D. Xiao, K.-\nM. C. Fu, and X. Xu, Van der Waals engineering of fer-\nromagnetic semiconductor heterostructures for spin and\nvalleytronics, Sci. Adv 3, e1603113 (2017).\n[8] N. Samarth, Condensed-matter physics: Magnetism in\n\ratland, Nature 546, 216 (2017).\n[9] J.-U. Lee, S. Lee, J. H. Ryoo, S. Kang, T. Y. Kim,\nP. Kim, C.-H. Park, J.-G. Park, and H. Cheong, Ising-\ntype magnetic ordering in atomically thin FePS 3, Nano\nLett. 16, 7433 (2016).\n[10] X. Wang, K. Du, Y. Y. F. Liu, P. Hu, J. Zhang,\nQ. Zhang, M. H. S. Owen, X. Lu, C. K. Gan, P. Sen-\ngupta, C. Kloc, and Q. Xiong, Raman spectroscopy\nof atomically thin two-dimensional magnetic iron phos-\nphorus trisul\fde (FePS 3) crystals, 2D Mater. 3, 031009\n(2016).\n[11] C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao,\nW. Bao, C. Wang, Y. Wang, Z. Q. Qiu, R. J. Cava, S. G.\nLouie, J. Xia, and X. Zhang, Discovery of intrinsic fer-\nromagnetism in two-dimensional van der Waals crystals,\nNature 546, 265 (2017).\n[12] B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein,\nR. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A.\nMcGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo-\nHerrero, and X. Xu, Layer-dependent ferromagnetism\nin a van der Waals crystal down to the monolayer limit,\nNature 546, 270 (2017).\n[13] D. R. Klein, D. MacNeill, J. L. Lado, D. Soriano,\nE. Navarro-Moratalla, K. Watanabe, T. Taniguchi,\nS. Manni, P. Can\feld, J. Fern\u0013 andez-Rossier, and\nP. Jarillo-Herrero, Probing magnetism in 2D van der\nWaals crystalline insulators via electron tunneling, Sci-\nence360, 1218 (2018).[14] S. Jiang, L. Li, Z. Wang, K. F. Mak, and J. Shan, Con-\ntrolling magnetism in 2D CrI 3by electrostatic doping,\nNat. Nanotechnol. 13, 549 (2018).\n[15] B. Huang, G. Clark, D. R. Klein, D. MacNeill,\nE. Navarro-Moratalla, K. L. Seyler, N. Wilson, M. A.\nMcGuire, D. H. Cobden, D. Xiao, W. Yao, P. Jarillo-\nHerrero, and X. Xu, Electrical control of 2D magnetism\nin bilayer CrI 3, Nat. Nanotechnol. 13, 544 (2018).\n[16] S. Jiang, J. Shan, and K. F. Mak, Electric-\feld switching\nof two-dimensional van der Waals magnets, Nat. Mater.\n17, 406 (2018).\n[17] Z. Wang, I. Guti\u0013 errez-Lezama, N. Ubrig, M. Kroner,\nM. Gibertini, T. Taniguchi, K. Watanabe, A. Imamo\u0015 glu,\nE. Giannini, and A. F. Morpurgo, Very large tunnel-\ning magnetoresistance in layered magnetic semiconductor\nCrI3, Nat. Commun. 9, 2516 (2018).\n[18] N. Sivadas, S. Okamoto, X. Xu, C. J. Fennie, and\nD. Xiao, Stacking-dependent magnetism in bilayer CrI 3,\nNano Lett. 18, 7658 (2018).\n[19] M. Kim, P. Kumaravadivel, J. Birkbeck, W. Kuang,\nS. G. Xu, D. G. Hopkinson, J. Knolle, P. A. McClarty,\nA. I. Berdyugin, M. Ben Shalom, R. V. Gorbachev, S. J.\nHaigh, S. Liu, J. H. Edgar, K. S. Novoselov, I. V. Grig-\norieva, and A. K. Geim, Micromagnetometry of two-\ndimensional ferromagnets, Nat. Electron. 2, 457 (2019).\n[20] D. R. Klein, D. MacNeill, Q. Song, D. T. Larson, S. Fang,\nM. Xu, R. A. Ribeiro, P. C. Can\feld, E. Kaxiras,\nR. Comin, and P. Jarillo-Herrero, Enhancement of inter-\nlayer exchange in an ultrathin two-dimensional magnet,\nNat. Phys. 15, 1255 (2019).\n[21] H. H. Kim, B. Yang, S. Li, S. Jiang, C. Jin, Z. Tao,\nG. Nichols, F. S\fgakis, S. Zhong, C. Li, S. Tian, D. G.\nCory, G.-X. Miao, J. Shan, K. F. Mak, H. Lei, K. Sun,\nL. Zhao, and A. W. Tsen, Evolution of interlayer and\nintralayer magnetism in three atomically thin chromium\ntrihalides, PNAS 116, 11131 (2019).\n[22] Y. Deng, Y. Yu, Y. Song, J. Zhang, N. Z. Wang, Z. Sun,\nY. Yi, Y. Z. Wu, S. Wu, J. Zhu, J. Wang, X. H. Chen,\nand Y. Zhang, Gate-tunable room-temperature ferro-\nmagnetism in two-dimensional Fe 3GeTe 2, Nature 563,\n94 (2018).\n[23] Z. Fei, B. Huang, P. Malinowski, W. Wang, T. Song,\nJ. Sanchez, W. Yao, D. Xiao, X. Zhu, A. F. May, W. Wu,\nD. H. Cobden, J.-H. Chu, and X. Xu, Two-dimensional\nitinerant ferromagnetism in atomically thin Fe 3GeTe 2,\nNat. Mater. 17, 778 (2018).\n[24] J.-M. Xu, S.-Y. Wang, W.-J. Wang, Y.-H. Zhou, X.-L.\nChen, Z.-R. Yang, and Z. Qu, Possible Tricritical Behav-\nior and Anomalous Lattice Softening in van der Waals\nItinerant Ferromagnet Fe 3GeTe 2under High Pressure,\nChin. Phys. Lett. 37, 076202 (2020).\n[25] M. Bonilla, S. Kolekar, Y. Ma, H. C. Diaz, V. Kalappat-\ntil, R. Das, T. Eggers, H. R. Gutierrez, M.-H. Phan, and\nM. Batzill, Strong room-temperature ferromagnetism in\nVSe 2monolayers on van der Waals substrates, Nat. Nan-\notechnol. 13, 289 (2018).\n[26] J. Li, B. Zhao, P. Chen, R. Wu, B. Li, Q. Xia, G. Guo,\nJ. Luo, K. Zang, Z. Zhang, H. Ma, G. Sun, X. Duan, and\nX. Duan, Synthesis of Ultrathin Metallic MTe 2(M= V,\nNb, Ta) Single-Crystalline Nanoplates, Adv. Mater. 30,\n1801043 (2018).12\n[27] D. J. O'Hara, T. Zhu, A. H. Trout, A. S. Ahmed, Y. K.\nLuo, C. H. Lee, M. R. Brenner, S. Rajan, J. A. Gupta,\nD. W. McComb, and R. K. Kawakami, Room tempera-\nture intrinsic ferromagnetism in epitaxial manganese se-\nlenide \flms in the monolayer limit, Nano Lett. 18, 3125\n(2018).\n[28] Q.-Q. Yuan, Z. Guo, Z.-Q. Shi, H. Zhao, Z.-Y. Jia,\nQ. Wang, J. Sun, D. Wu, and S.-C. Li, Ferromagnetic\nMnSn Monolayer Epitaxially Grown on Silicon Substrate,\nChin. Phys. Lett. 37, 077502 (2020).\n[29] A. Avsar, A. Ciarrocchi, M. Pizzochero, D. Unuchek,\nO. V. Yazyev, and A. Kis, Defect induced, layer-\nmodulated magnetism in ultrathin metallic PtSe 2, Nat.\nNanotechnol. 14, 674 (2019).\n[30] D. C. Freitas, R. Weht, A. Sulpice, G. Remenyi, P. Stro-\nbel, F. Gay, J. Marcus, and M. N\u0013 u~ nez-Regueiro, Fer-\nromagnetism in layered metastable 1T-CrTe 2, J. Phys.:\nCondens. Matter 27, 176002 (2015).\n[31] X. Sun, W. Li, X. Wang, Q. Sui, T. Zhang, Z. Wang,\nL. Liu, D. Li, S. Feng, S. Zhong, H. Wang, V. Bouchiat,\nM. Nunez Regueiro, N. Rougemaille, J. Coraux, A. Pur-\nbawati, A. Hadj-Azzem, Z. Wang, B. Dong, X. Wu,\nT. Yang, G. Yu, B. Wang, Z. Han, X. Han, and Z. Zhang,\nRoom temperature ferromagnetism in ultra-thin van der\nWaals crystals of 1T-CrTe 2, Nano Research (2020),\n10.1007/s12274-020-3021-4.\n[32] A. Purbawati, J. Coraux, J. Vogel, A. Hadj-Azzem,\nN. Wu, N. Bendiab, D. Jegouso, J. Renard, L. Marty,\nV. Bouchiat, A. Sulpice, L. Aballe, M. Foerster,\nF. Genuzio, A. Locatelli, T. O. Mentes, Z. V. Han,\nX. Sun, M. N\u0013 u~ nez-Regueiro, and N. Rougemaille, In-\nplane magnetic domains and N\u0013 eel-like domain walls in\nthin \rakes of the room temperature CrTe 2van der Waals\nferromagnet, ACS Appl. Mater. Interfaces 12, 30702\n(2020).\n[33] J. Kerr, On rotation of the plane of polarization by re-\n\rection from the pole of a magnet, Phil. Mag. 3, 321\n(1877).\n[34] M. Faraday, Experimental researches in electricity.-\nNineteenth series, Phil. Trans. R. Soc. 136, 1 (1846).\n[35] V. Antonov, B. Harmon, and A. Yaresko, Elec-\ntronic structure and magneto-optical properties of solids\n(Springer Science & Business Media, 2004).\n[36] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and\nN. P. Ong, Anomalous hall e\u000bect, Rev. Mod. Phys. 82,\n1539 (2010).\n[37] R. Kubo, Statistical-mechanical theory of irreversible\nprocesses. I. General theory and simple applications to\nmagnetic and conduction problems, J. Phys. Soc. Jpn.\n12, 570 (1957).\n[38] C. S. Wang and J. Callaway, Band structure of nickel:\nSpin-orbit coupling, the Fermi surface, and the optical\nconductivity, Phys. Rev. B 9, 4897 (1974).\n[39] X. Zhou, J.-P. Hanke, W. Feng, F. Li, G.-Y. Guo, Y. Yao,\nS. Bl ugel, and Y. Mokrousov, Spin-order dependent\nanomalous Hall e\u000bect and magneto-optical e\u000bect in the\nnoncollinear antiferromagnets Mn 3XN with X= Ga, Zn,\nAg, or Ni, Phys. Rev. B 99, 104428 (2019).\n[40] W. Nernst, Ueber die electromotorischen Kr afte, welche\ndurch den Magnetismus in von einem W armestrome\ndurch\rossenen Metallplatten geweckt werden, Ann.\nPhys. 267, 760 (1887).\n[41] N. Hanasaki, K. Sano, Y. Onose, T. Ohtsuka, S. Iguchi,\nI. K\u0013 ezsm\u0013 arki, S. Miyasaka, S. Onoda, N. Nagaosa,and Y. Tokura, Anomalous Nernst e\u000bects in pyrochlore\nmolybdates with spin chirality, Phys. Rev. Lett. 100,\n106601 (2008).\n[42] M. Ikhlas, T. Tomita, T. Koretsune, M.-T. Suzuki,\nD. Nishio-Hamane, R. Arita, Y. Otani, and S. Nakat-\nsuji, Large anomalous Nernst e\u000bect at room temperature\nin a chiral antiferromagnet, Nat. Phys. 13, 1085 (2017).\n[43] T. Liang, J. Lin, Q. Gibson, T. Gao, M. Hirschberger,\nM. Liu, R. J. Cava, and N. P. Ong, Anomalous Nernst\ne\u000bect in the dirac semimetal Cd 3As2, Phys. Rev. Lett.\n118, 136601 (2017).\n[44] C. Wuttke, F. Caglieris, S. Sykora, F. Scaravaggi,\nA. U. B. Wolter, K. Manna, V. S uss, C. Shekhar,\nC. Felser, B. B uchner, and C. Hess, Berry curvature un-\nravelled by the anomalous Nernst e\u000bect in Mn 3Ge, Phys.\nRev. B 100, 085111 (2019).\n[45] W.-B. Zhang, Q. Qu, P. Zhu, and C.-H. Lam, Ro-\nbust Intrinsic Ferromagnetism and Half Semiconductiv-\nity in Stable Two-dimensional Single-layer Chromium\nTrihalides, J. Mater. Chem. C 3, 12457 (2015).\n[46] H. L. Zhuang, P. R. C. Kent, and R. G. Hennig, Strong\nanisotropy and magnetostriction in the two-dimensional\nStoner ferromagnet Fe 3GeTe 2, Phys. Rev. B 93, 134407\n(2016).\n[47] G. Kresse and J. Hafner, Ab initio molecular dynamics\nfor liquid metals, Phys. Rev. B 47, 558 (1993).\n[48] G. Kresse and J. Furthm uller, E\u000ecient iterative schemes\nfor ab initio total-energy calculations using a plane-wave\nbasis set, Phys. Rev. B 54, 11169 (1996).\n[49] P. E. Bl ochl, Projector augmented-wave method, Phys.\nRev. B 50, 17953 (1994).\n[50] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized\ngradient approximation made simple, Phys. Rev. Lett.\n77, 3865 (1996).\n[51] V. I. Anisimov, J. Zaanen, and O. K. Andersen, Band\ntheory and Mott insulators: Hubbard U instead of Stoner\nI, Phys. Rev. B 44, 943 (1991).\n[52] S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J.\nHumphreys, and A. P. Sutton, Electron-energy-loss\nspectra and the structural stability of nickel oxide: An\nLSDA+ U study, Phys. Rev. B 57, 1505 (1998).\n[53] X. Sui, T. Hu, J. Wang, B.-L. Gu, W. Duan, and M.-\ns. Miao, Voltage-controllable colossal magnetocrystalline\nanisotropy in single-layer transition metal dichalco-\ngenides, Phys. Rev. B 96, 041410(R) (2017).\n[54] A. A. Mosto\f, J. R. Yates, Y.-S. Lee, I. Souza, D. Van-\nderbilt, and N. Marzari, wannier90: A tool for obtaining\nmaximally-localised Wannier functions, Comput. Phys.\nCommun. 178, 685 (2008).\n[55] J. Callaway, Quantum theory of the solid state (Aca-\ndemic Press, 2013).\n[56] W. Feng, G.-Y. Guo, and Y. Yao, Tunable\nmagneto-optical e\u000bects in hole-doped group-IIIA metal-\nmonochalcogenide monolayers, 2D Mater. 4, 015017\n(2017).\n[57] Y. Yao, L. Kleinman, A. H. MacDonald, J. Sinova,\nT. Jungwirth, D.-s. Wang, E. Wang, and Q. Niu, First\nprinciples calculation of anomalous Hall conductivity in\nferromagnetic bcc Fe, Phys. Rev. Lett. 92, 037204 (2004).\n[58] D. Xiao, Y. Yao, Z. Fang, and Q. Niu, Berry-phase e\u000bect\nin anomalous thermoelectric transport, Phys. Rev. Lett.\n97, 026603 (2006).\n[59] X. Zhou, J.-P. Hanke, W. Feng, S. Bl ugel, Y. Mokrousov,\nand Y. Yao, Giant anomalous Nernst e\u000bect in non-13\ncollinear antiferromagnetic Mn-based antiperovskite ni-\ntrides, Phys. Rev. Materials 4, 024408 (2020).\n[60] P. W. Anderson, Antiferromagnetism. Theory of superex-\nchange interaction, Phys. Rev. 79, 350 (1950).\n[61] J. B. Goodenough, An interpretation of the mag-\nnetic properties of the perovskite-type mixed crystals\nLa1\u0000xSrxCoO 3\u0000\u0015, J. Phys. Chem. Solids 6, 287 (1958).\n[62] J. Kanamori, Superexchange interaction and symmetry\nproperties of electron orbitals, J. Phys. Chem. Solids 10,\n87 (1959).\n[63] H. Y. Lv, W. J. Lu, D. F. Shao, Y. Liu, and Y. P. Sun,\nStrain-controlled switch between ferromagnetism and an-\ntiferromagnetism in 1 T-CrX 2(X = Se, Te) monolayers,\nPhys. Rev. B 92, 214419 (2015).\n[64] S. Li, S.-S. Wang, B. Tai, W. Wu, B. Xiang, X.-L. Sheng,\nand S. A. Yang, Tunable anomalous Hall transport inbulk and two-dimensional 1T-CrTe 2: A \frst-principles\nstudy, arXiv:2006.10795.\n[65] H. T. Stokes, D. M. Hatch, and B. J.\nCampbell, ISOTROPY Software Suite.,\nhttps://stokes.byu.edu/iso/isotropy.php .\n[66] Z. Liu, G. Zhao, B. Liu, Z. F. Wang, J. Yang, and F. Liu,\nIntrinsic quantum anomalous Hall e\u000bect with in-plane\nmagnetization: searching rule and material prediction,\nPhys. Rev. Lett. 121, 246401 (2018).\n[67] X. Zhou, W. Feng, F. Li, and Y. Yao, Large magneto-\noptical e\u000bects in hole-doped blue phosphorene and gray\narsenene, Nanoscale 9, 17405 (2017).\n[68] G. E. W. Bauer, E. Saitoh, and B. J. van Wees, Spin\ncaloritronics, Nat. Mater. 11, 391 (2012).\n[69] S. R. Boona, R. C. Myers, and J. P. Heremans, Spin\ncaloritronics, Energy Environ. Sci. 7, 885 (2014)." }, { "title": "2011.05455v3.Spin_waves_in_the_collinear_antiferromagnetic_phase_of_Mn___bf_5__Si___bf_3__.pdf", "content": "Spin{waves in the collinear antiferromagnetic phase of Mn 5Si3\nF. J. dos Santos,1, 2,\u0003N. Biniskos,3,yS. Raymond,4K. Schmalzl,5M. dos\nSantos Dias,1P. Ste\u000bens,6J. Persson,7S. Bl ugel,1S. Lounis,1, 8and T. Br uckel7\n1Peter Gr unberg Institut and Institute for Advanced Simulations,\nForschungszentrum J ulich &JARA, D-52425 J ulich, Germany\n2Department of Physics, RWTH Aachen University, 52056 Aachen, Germany\n3Forschungszentrum J ulich GmbH, J ulich Centre for Neutron Science at MLZ, Lichtenbergstr. 1, 85748 Garching, Germany\n4Universit\u0013 e Grenoble Alpes, CEA, IRIG, MEM, 38000 Grenoble, France\n5Forschungszentrum J ulich GmbH, J ulich Centre for Neutron Science at ILL, 71 avenue des Martyrs, 38000 Grenoble, France\n6Institut Laue-Langevin, 71 avenue des Martyrs, 38000 Grenoble, France\n7Forschungszentrum J ulich GmbH, J ulich Centre for Neutron Science (JCNS-2)\nand Peter Gr unberg Institut (PGI-4), JARA-FIT, 52425 J ulich, Germany\n8Faculty of Physics, University of Duisburg-Essen, 47053 Duisburg, Germany\n(Dated: December 10, 2020)\nBy combining two independent approaches, inelastic neutron scattering measurements and density\nfunctional theory calculations, we study the spin{waves in the collinear antiferromagnetic phase\n(AFM2) of Mn 5Si3. We obtain its magnetic ground{state properties and electronic structure. This\nstudy allowed us to determine the dominant magnetic exchange interactions and magnetocrystalline\nanisotropy in the AFM2 phase of Mn 5Si3. Moreover, the evolution of the spin excitation spectrum\nis investigated under the in\ruence of an external magnetic \feld perpendicular to the anisotropy\neasy{axis. The low energy magnon modes show a di\u000berent magnetic \feld dependence which is\na direct consequence of their di\u000berent precessional nature. Finally, possible e\u000bects related to the\nDzyaloshinskii{Moriya interaction are also considered.\nI. INTRODUCTION\nThe study of magnetism at a microscopic level can\nlead to designing cutting{edge technological applications\nfor data process and storage, information transmission,\nand magnetic refrigeration. In recent years, antiferro-\nmagnetic (AFM) materials have attracted great interest\nin the research \feld of spintronics1. Bulk Mn 5Si3is an\nAFM intermetallic compound that is hosting rich physics.\nIts interesting properties, such as the complex magnetic\nstructure2, the anomalous Hall e\u000bect3, and the inverse\nmagnetocaloric e\u000bect4have been attributed to an insta-\nbility of the Mn magnetic moments. It is also worth\nmentioning that in nanoparticle5and nanowire6form,\nMn5Si3is considered to have great potential in future\nelectronic and spintronic devices. Despite the intense re-\nsearch activity over the past decades2{4,7{16, many open\nquestions remain regarding the minimal magnetic model\nHamiltonian, the role of the spin \ructuations in the mag-\nnetically ordered phases and which Mn site is responsi-\nble for them. To address some of these questions, in the\npresent study, we perform inelastic neutron scattering\nexperiments and apply \frst{principles calculations.\nThe crystal and magnetic structure of bulk Mn 5Si3\nhas been established by neutron di\u000braction measure-\nments2,7,8and the magnetic phase diagram as a func-\ntion of temperature and magnetic \feld has been ex-\ntensively studied by magnetization and electrical trans-\nport measurements9{12,17. In the paramagnetic (PM)\nstate, Mn 5Si3crystallizes in the hexagonal space group\nP63=mcm with two distinct crystallographic positions\nfor Mn atoms (sites Mn1 and Mn2)7and undergoes two\nsuccessive \frst order phase transitions towards antiferro-\n(a) (b)\nSi Mn2Mn1J4\nJ2\nJ1\nJ3\nab\nc\nFIG. 1. (a) Projection of the structure of Mn 5Si3in the\nAFM2 phase in the ab{plane of the orthorhombic cell accord-\ning to single{crystal neutron di\u000braction data8. The blue solid\nlines indicate the relevant exchange interactions used in the\nHeisenberg Hamiltonian. (b) Predicted metastable FM phase\nfrom \frst{principles calculations (see details in text). The\ntwo triangles indicate Mn2 atoms located in di\u000berent planes.\nmagnetic phases, which occur at TN2\u0019100 K (AFM2)\nandTN1\u001966 K (AFM1), respectively17. The electric\nresistivity shows a metallic behavior with two anomalies\ncorresponding to these phase transitions10.\nAtTN2\u0019100 K (AFM2), the crystal structure changes\nfrom hexagonal to orthorhombic with space group Ccmm\nand Mn2 divides into two sets of inequivalent positions8.\nIn the orthorhombic cell magnetic re\rections follow the\nconditionh+kodd and the magnetic propagation vec-arXiv:2011.05455v3 [cond-mat.mes-hall] 9 Dec 20202\ntor is \u0014= (0;1;0). In the AFM2 phase, the Mn1\nand one{third of the Mn2 atoms have no ordered mo-\nments and the remaining Mn2 atoms have their mag-\nnetic moments of magnitude 1.48(1) \u0016Baligned almost\nparallel and antiparallel to the b{axis of the orthorhom-\nbic cell8(see Fig. 1(a)). At a lower temperature, at\nTN1\u001966 K (AFM1), a structural distortion occurs to\nan orthorhombic cell without inversion symmetry (space\ngroupCc2m)2. The magnetic moments reorient in a\nhighly noncollinear and noncoplanar arrangement, while\nthe propagation vector remains the same. Mn1 atoms\nacquire a magnetic moment of magnitude 1.20(5) \u0016Band\nstill one{third of the Mn2 atoms have no ordered mo-\nments, just as in the AFM2 phase. The rest of the Mn2\natoms carry a moment of 2.30(9) and 1.85(9) \u0016B, depend-\ning on their site.\nII. EXPERIMENTAL PART\nA. Experimental details\nThe Mn 5Si3single crystal was grown by the Czochral-\nski method4. The sample with a mass of about 7 g was\nmounted on an aluminum sample holder and was oriented\nin the [100]/[010] scattering plane of the orthorhombic\nsymmetry. Inelastic neutron scattering (INS) measure-\nments were carried out on the cold triple{axis spectrom-\neters (TAS) IN1218and ThALES at the Institut Laue\nLangevin (ILL) in Grenoble, France. Both TAS were\nsetup in W con\fguration and inelastic scans were per-\nformed with constant kf, where kfis the wave{vector of\nthe scattered neutron beam.\nUnpolarized INS measurements were performed at\nIN12 and focusing setups were employed. The spectrome-\nter was equipped with pyrolytic graphite (PG(002)) crys-\ntals as the monochromator and analyzer and 40'{open{\nopen collimations were installed. Higher{order contami-\nnation was removed using a velocity selector (VS) before\nthe monochromator and a beryllium (Be) \flter in the\nscattered neutron beam. The sample was cooled below\nroom temperature with a4He \row cryostat. Spin dy-\nnamics investigations with unpolarized neutrons under\nthe magnetic \feld were carried out using a 10 T verti-\ncal \feld magnet. For these measurements, the Be \flter\nwas removed. The single crystal was cooled down from\nthe PM state to T= 80 K (AFM2 phase) without the\npresence of an external magnetic \feld. The \feld was ap-\nplied along the c{axis of the orthorhombic symmetry of\nMn5Si3and the spectra were collected with increasing\n\feld strength.\nAt ThALES longitudinal polarization analysis (LPA)\nwas performed using the CRYOPAD device19to guide\nand orient the neutron beam polarization with a strictly\nzero magnetic \feld in the sample position. The TAS was\nequipped with polarizing Heusler (Cu 2MnAl(111)) crys-\ntals as the monochromator and analyzer. A \ripping ratio\nof 14 was determined from measurements in a graphite\nFIG. 2. Temperature dependence of the dynamical spin\nsusceptibility \u001f00(Q;E) of Mn 5Si3atQ= (0:9;2;0) and\nE= 0:5 meV measured with unpolarized neutrons with\nkf= 1:5\u0017A\u00001. The vertical red dashed lines indicate TN2\u0019\n100 K andTN1\u001966 K. The arrow indicates the temperature\n(T= 80 K) where INS data were collected.\nsample. Fully focusing setups were employed and higher{\norder contamination was removed using a VS and a Be\n\flter before the monochromator and in the scattered neu-\ntron beam, respectively. Inelastic scans were performed\nwith a constant kfof 1.1 \u0017A\u00001. For the polarized INS ex-\nperiments the common Cartesian coordinate system was\nused20: thex{axis parallel to the scattering vector Q21,\nthey{axis perpendicular to Qin the scattering plane and\nthez{axis perpendicular to the scattering plane.\nB. Unpolarized INS measurements\nTo determine the extent of the critical spin \ructua-\ntions related to the two AFM transitions, spectra were\ncollected with an unpolarized neutron beam at small q.\nA (Q,E) position was carefully selected to avoid contri-\nbutions to the measured intensity from the elastic and\ninelastic scattering from the magnetic zone centers and\nlow energy magnon modes, respectively. The obtained\nintensity after background subtraction was corrected by\nthe detailed balance factor so that the \fnal result relates\nto the imaginary part of the dynamical spin susceptibil-\nity\u001f00(Q;E). Fig. 2 shows the extent of the spin \ruc-\ntuations for Q= (0:9;2;0) andE= 0:5 meV in the PM\nstate, as well as, the critical \ructuations due to the two\nAFM phase transitions as seen by the broad tail above\nTN.\u001f00(Q;E) shows two maxima at TN2andTN1, in\nagreement with the established magnetic phase diagrams\nof Mn 5Si3where the AFM transitions occur9,12.\nMagnetic excitations were measured around the mag-\nnetic zone center G= (1;2;0) atT= 80 K, a tempera-\nture selected well inside the AFM2 phase where the in-\ntensity shows a plateau and the combined critical spin\n\ructuations from the PM to the AFM2 ( TN2\u0019100 K)3\nand from the AFM2 to the AFM1 ( TN1\u001966 K) transi-\ntions have minimal intensities (see Fig. 2). The energy\ndependence of the measured excitations at di\u000berent Qh\npositions is shown in Fig. 3(a), where Q= (Qh;2;0). The\nindividual spectra consist of three peaks. The \frst peak\nthat is always centered at E= 0 meV corresponds to the\nelastic line. The two other peaks appear at \fnite Eand\nshift to higher energy transfers as Qhincreases, charac-\nteristic of dispersive spin waves. To analyze the obtained\nspectra, a constant background was assumed and Gaus-\nsian functions were selected to describe the peaks. One\ncould argue that the signal at \fnite EatQ= (1:037;2;0)\nandQ= (1:05;2;0) could be described by a single broad\npeak. However, the use of polarized neutrons (see Sec-\ntion II.D) justi\fes the existence of two peaks for the ex-\ncitations, since the polarized INS cross{sections implore\nstrict \ftting conditions. A typical E{scan collected at\nQ= (1:005;2;0) with the individual \ft for the elastic\nand spin-wave signals is shown in Fig. 3(b).\nThe obtained low energy experimental spin{wave dis-\npersion along the ( h00) symmetry direction in the AFM2\nphase of Mn 5Si3is shown in Fig. 3(c). There are two\ncharacteristic features: (i) two small energy gaps at q= 0\nand (ii) the lowest magnetic excitations can be described\nby the empirical dispersion relation E=p\n\u00012+C2q222,\nwhere \u0001 refers to the spin{gap and Cis a constant.\nThe obtained values for the two magnon modes are:\n\u0001\u000b= 0:408(7) meV, C\u000b= 6:3(5) meV/r.l.u., \u0001 \f=\n0:181(4) meV and C\f= 4:7(3) meV/r.l.u.. The observed\ngaps are indications of two easy{axis anisotropies, an as-\nsumption stemming from the collinear spin arrangement\nin the AFM2 phase of Mn 5Si3. In order to describe the\nmagnon spectrum and to extract the dominant magnetic\nexchange interactions and magnetocrystalline anisotropy\nin the AFM2 phase, theoretical calculations were em-\nployed (see Section III). In what follows, the modes orig-\ninating from \u0001 \u000band \u0001\fwill be referred to as the \u000band\nthe\f-modes, respectively.\nC. Unpolarized INS measurements under magnetic\n\feld\nForHk^c, neutron di\u000braction measurements13per-\nformed in single crystals of Mn 5Si3indicate that no\n\feld{induced transition occurs within the AFM2 phase\nand consistently with the macroscopic data9,11,12the\nPM state is not reached up to H= 8 T due to the\nsteepTN2(H) phase boundary. In order to investigate\nthe magnon spectrum under magnetic \feld, energy spec-\ntra were collected at three di\u000berent Qhpositions (1,\n1.018 and 1.025 r.l.u.) around the magnetic Bragg peak\nG= (1;2;0). Figure 4(a) shows such characteristic scans\natQh= 1 r.l.u. for di\u000berent magnetic \felds. The ob-\ntained spectra were analyzed as described in the previous\nsection. For increasing magnetic \feld, the two peaks that\ncorrespond to the \u000band the\f{modes show di\u000berent be-\nhavior. As the \feld increases, the position and intensity\n\u000b-mode\n\f-mode(a)\n(b)\n(c)\nFIG. 3. (a) Energy spectra of Mn 5Si3atQ= (1 +qh;2;0)\nmeasured at IN12 spectrometer with unpolarized setup with\nkf= 1:05\u0017A\u00001atT= 80 K. (b) Fitting of the E{scan at\nQ= (1:005;2;0). The three dashed lines correspond to Gaus-\nsian functions sitting on top of a \rat background. The solid\nlines indicate the overall \fts as described in the text. (c) Low\nenergy magnon dispersion at T= 80 K along the ( h00) di-\nrection. The solid lines are \fts with the empirical dispersion\nrelationE=p\n\u00012+C2q2.\nof the\u000b{mode is not signi\fcantly a\u000bected in contrast to\nthe\f{mode, which disperses with the \feld and a con-\ntinuous diminution of intensity is observed. It should be\nnoted that at H= 4 T, the two peaks seem to merge.\nThe same observations regarding the behavior of the\ntwo modes stem from measurements at Qh= 1:018 r.l.u.\nandQh= 1:025 r.l.u. and are summarized in Fig. 4(b).\nThe obtained results indicate that the modes behave dif-\nferently under the external magnetic \feld possibly due\nto their di\u000berent polarization. To shed light on this be-\nhavior, spectra were collected using the polarized INS4\n\f-mode\n\u000b-mode(a)\n(b)\nFIG. 4. (a) Inelastic spectra of Mn 5Si3forHk^cobtained\natQ= (1;2;0) atT= 80 K (AFM2 phase) with unpolarized\nbeam withkf= 1:05\u0017A\u00001at IN12. The lines indicate \fts with\nGaussian functions. (b) Energy of the spin excitations as a\nfunction of the external magnetic \feld at three di\u000berent Qh\npositions at T= 80 K (AFM2 phase). Lines are guides for\nthe eyes.\nmethod.\nD. Polarized INS measurements\nAs a general rule, neutron scattering is only sensitive\nto magnetic excitations perpendicular to Q20. With LPA\nit is possible to separate magnetic \ructuations polarized\nalong di\u000berent directions in spin space. The initial polar-\nization was prepared parallel to x{axis, perpendicular to\nQin the scattering plane ( y{axis) and perpendicular to\nthe scattering plane ( z{axis), and the \fnal polarization\nwas analyzed for a scattering process reversing the initial\npolarization by 180\u000e. The corresponding measurement\nchannels are canonically labeled SF xx, SFyy, and SFzz,\nwhere SF stands for \\Spin{Flip\".\nThe neutron scattering double di\u000berential cross{sections for the three SF channels are20:\nSFxx=\u0012d2\u001b\nd\ndE\u0013x\nSF/BGSF+h\u000eMyi+h\u000eMzi(1)\nSFyy=\u0012d2\u001b\nd\ndE\u0013y\nSF/BGSF+h\u000eMzi(2)\nSFzz=\u0012d2\u001b\nd\ndE\u0013z\nSF/BGSF+h\u000eMyi(3)\nwhere BG SFis the background (which includes the nu-\nclear spin scattering) and h\u000eMyiandh\u000eMzithe mea-\nsured magnetic \ructuations. Considering that: (i) Qis\nin theab{plane, (ii) in the AFM2 phase the magnetic\nmoments lie parallel and antiparallel to the b{axis and\n(iii) spin{waves correspond to precession perpendicular\nto the ordered moment, then in the crystal frame the\ncross{sections become:\nSFxx/BGSF+ sin2\u0012h\u000eMai+h\u000eMci (4)\nSFyy/BGSF+h\u000eMci (5)\nSFzz/BGSF+ sin2\u0012h\u000eMai (6)\nwith\u0012the angle between Qand the [100] direction and\ncan be calculated by \u0012= arctan(Qk\nQha\nb).\nTo get further insight regarding the polarization de-\npendence of the two magnon modes, E{spectra at di\u000ber-\nentQpositions were collected in the three SF channels at\nT= 80 K. The magnetic \ructuations h\u000eMaiandh\u000eMci\nwere extracted by taking the di\u000berence of intensities be-\ntween the di\u000berent polarization channels. A typical re-\nsult of such analysis is shown in Fig. 5 at Q= (1:06;2;0)\nwhere the intensity for h\u000eMaiis corrected by the angle\nprefactor. The peak positions of the subtracted spin \ruc-\ntuations spectra are consistent with the established low\nenergy magnon dispersion curves obtained from the un-\npolarized data and shown in Fig. 3(c). Moreover, it is\nevident that the maximum of intensity in h\u000eMaiand in\nh\u000eMcicorresponds to the \u000b{mode and \f{mode, respec-\ntively. This hints that the elliptic polarization of each\nmode is di\u000berent and points along di\u000berent crystal axis.\nIII. THEORETICAL CALCULATIONS\nA. Density functional theory\nFirst{principles calculations were performed to deter-\nmine the ground{state electronic and magnetic proper-\nties of the AFM2 phase of Mn 5Si3. Our study was\nbased on the atomic structure speci\fed in Ref. 8. We\nemployed density functional theory (DFT) using the\nfull{potential Korringa{Kohn{Rostoker Green{function\n(KKR{GF) method including spin{orbit coupling, as im-\nplemented in the JuKKR code23, using the local{spin|\ndensity approximation24. The cuto\u000b for the angular mo-\nmentum expansion of the scattering problem was set to\nlmax= 3. Furthermore, the energy integration was per-\nformed in the upper complex energy plane25with 305\n\f-mode\n\u000b-mode\nFIG. 5. Subtracted spin \ructuations spectra h\u000eMaiand\nh\u000eMciof Mn 5Si3obtained at ThALES and measured at\nQ= (1:06;2;0) atT= 80 K. The intensity for h\u000eMaiwas\ncorrected by the factor sin2\u0012= 0:545. The lines indicate \fts\nwith Gaussian functions.\npoints in a rectangular path and 5 Matsubara frequencies\nat a temperature of T= 473:68 K, and the Brillouin zone\nintegration was performed with 30 \u000215\u000230k{points. The\nmagnetic exchange tensor, which parametrizes the spin\nHamiltonian discussed in the following section, was ob-\ntained through the in\fnitesimal{rotations method26,27.\nIn these calculations, the number of Matsubara frequen-\ncies was increased to 10 with T= 100 K.\nWe explored two possible magnetic con\fgurations with\nself{consistent calculations, the AFM con\fguration that\nis shown in Fig. 1, and a ferromagnetic (FM) phase with\n\fnite magnetic moments in both Mn1 and Mn2 sites. The\nAFM was found the most energetically favorable with the\nFM phase being 146 meV per unit cell higher in energy.\nIn the FM phase, the magnetic moments are 2.6 and\n1.2\u0016Bfor the Mn2 and Mn1 sites, respectively, while the\nSi sites have magnetic moments of 0.1 \u0016Bantiparallel to\nthose of the manganese sites. For the AFM phase, two{\nthirds of the Mn2 site carry magnetic moments of 2.4 \u0016B\nwhile the other Mn and the Si atoms have no magnetic\nmoment.\nCombining the magnetic force theorem with the\nfrozen potential approximation, the magnetocrystalline\nanisotropy was determined from band energy di\u000berences\nbetween states with di\u000berent orientations of the spin\nmagnetic moments. For the AFM phase, we obtained the\nfollowing energy di\u000berences when aligning the magnetic\nmoments along the main crystal axes: Ea\u0000Eb= 0:12,\nEc\u0000Eb= 0:09, andEa\u0000Ec= 0:03 meV per magnetic\natom, which indicates that bandcare the \frst and sec-\nond preferred axis, respectively, and ais the hard{axis.\nB. Model Hamiltonian and spin{wave calculation\nWe mapped the ab initio calculations onto a quantum\nHeisenberg Hamiltonian to study the spin{wave excita-tions in the adiabatic approximation, which reads as\nH=\u0000X\nijJijSi\u0001Sj\u0000X\n\u000bk\u000bX\ni(S\u000b\ni)2: (7)\nThe \frst term is due to the magnetic exchange interac-\ntion whose coupling is given by Jij.Siis the spin for\nwhich we set S= 1. The second term accounts for the\nbiaxial magnetocrystalline anisotropy ( Hani) withkb=\nEa\u0000Eb= 0:12 meV and kc=Ea\u0000Ec= 0:03 meV, both\nbeing positive. From the DFT calculations, we obtained\nthatkb>kc, which makes bthe primary easy{axis, and\ncthe secondary easy{axis. The biaxial anisotropy could\nalso be modeled by a combination of an easy{axis along\nband an easy{plane ( bc{plane) anisotropy.\nThe magnetic exchange interactions were obtained\nfrom \frst{principles calculations for the AFM2 phase.\nThe results for the \frst few Mn2 pairs as indicated in\nFig. 1(a) are: J1=\u000012:23,J2=\u00002:16,J3= 3:98 and\nJ4=\u00002:89 meV. Most interactions, apart from J3, have\nAFM character. This indicates that the AFM2 phase is\nfavored by all those pair interactions, that is, within this\nset of interactions, there is no frustration. J1,J2, andJ3\ncorrespond to couplings between magnetic moments in\nthe same [Mn2] 6octahedra. J1has the highest value and\nis the exchange interaction between the spins located on\na triangle in the ab{plane (distance 2.789 \u0017A).J2andJ3\ncouple spins located on adjacent triangles separated by\nc/2 with distances 2.893 and 4.019 \u0017A, respectively. The\nexchange interaction J4concerns the shortest distance\n(4.364 \u0017A) between spins located on adjacent [Mn2] 6oc-\ntahedra.\nWe employ the linear spin{wave approximation to ob-\ntain the spin{wave excitations of the quantum Heisen-\nberg Hamiltonian using the computed magnetic inter-\naction parameters. The spin{wave excitations are the\neigenstates of the dynamical matrix associated with the\nquantum Heisenberg Hamiltonian in Eq. (7), as explained\nin detail in Ref. 28. We start by constructing a local co-\nordinate system for every magnetic site with the local z{\naxis coinciding with the classical ground{state spin orien-\ntation. In this local frame, we expand the quantum spin\noperators using the linearized Holstein{Primako\u000b trans-\nformation as Si=\u0010p\n2Sai+ay\ni\n2;p\n2Sai\u0000ay\ni\n2i;S\u0000ay\niai\u0011\n,\nwhereay\niandaiare bosonic ladder operators29. Keeping\nonly terms up to second order in the Holstein{Primako\u000b\nbosons, the Hamiltonian can be written as H=H0+H2.\nTheH0term is a constant corresponding to the classical\nground{state energy. The second term\nH2=\u0000X\nkX\n\u0016\u0017ay(k)H\u0016\u0017(k)a\u0017(k) (8)\ncontains the quadratic terms of the Holstein{Primako\u000b\nbosons describing the spin excitations, where H(k) is a\n2n\u00022nmatrix and a\u0016(k) =\u0010a\u0016(k)\nay\n\u0016(\u0000k)\u0011\nwitha\u0016(k) =\n1p\nNP\nme\u0000ik\u0001Rmam\u0016.nandNare the number of atoms6\nin the unit and the number of unit cells under pe-\nriodic Born{von{Karman boundary conditions, respec-\ntively. The spin{wave eigenvalues !(k) and eigenvectors\njkiare then obtained by a Bogoliubov transformation30.\nThis process diagonalizes the system's dynamical ma-\ntrixD=gH2, where gis a diagonal matrix containing\n\u00001 on its \frst half and 1 on the second while ensur-\ning the bosonic character of the diagonalizing basis. The\nspin{wave inelastic scattering spectrum is computed with\nour theory for spin{resolved electron{energy{loss spec-\ntroscopy (SREELS) of noncollinear magnets presented in\nRef. 28, where we employ time{dependent perturbation\ntheory to describe the interaction between the probing\nbeam and the magnetic excitations. The same theory\ncan be applied with little modi\fcation to describe inelas-\ntic neutron scattering. This method has been applied\nto investigate ferromagnetic and antiferromagnetic non-\ncollinear spin textures in Refs. 31{33.\nIV. COMPARING EXPERIMENTAL AND\nTHEORETICAL RESULTS\nA. Results without external magnetic \feld\nAs already mentioned, the DFT calculations found\nthat the collinear AFM2 phase is more stable than the\nFM one, in line with the experimental \fnding that the\ncollinear AFM2 phase is stable in an intermediate tem-\nperature range8. The AFM2 phase has the peculiarity\nthat one{third of the Mn2 sites and all the Mn1 sites\nhave vanishing ordered magnetic moments, with the main\nquestion being whether this is due to a collapse of the lo-\ncal magnetic moment or due to a disordering of their\norientation. In the DFT calculations, we considered the\ncollapse scenario, which did not lead to a strong energetic\npenalty, given that the FM phase was still substantially\nhigher in energy. However, the obtained strong AFM\nexchange interactions between the Mn2 moments in the\n[Mn2] 6octahedra (the stacked triangles in Fig. 1) can\nalso lead to a \fnite temperature scenario where a third\nof the Mn2 moments is orientationally disordered, as pro-\nposed for Mn 3Pt in Ref. 34. The DFT calculations can\nthus also support a hybrid scenario where orientational\ndisorder is mixed with sizable \ructuations of the mag-\nnitude of the moments for the Mn2 sites with vanishing\nordered moments. Concerning the Mn1 sites, the DFT\ncalculations suggest that their magnetic moments are un-\nstable on their own, and come into existence depending\non the Mn2 environment. When this environment is FM\na sizable magnetic moment arises in the Mn1 sites, while\nif this environment is AFM the magnetic moment col-\nlapses. We conclude that the DFT calculations do not\nrule out the existence of strong spin \ructuations either\non the Mn1 or on the Mn2 sites. Previous experimental\nwork4has suggested that the magnetic excitation spec-\ntrum of the AFM2 phase consists of propagating spin{\nwaves and di\u000buse spin \ructuations, originating from the\n(a)\n(b)\nFIG. 6. (a) Theoretical inelastic scattering spectrum for the\nAFM2 phase of Mn 5Si3with parameters obtained from DFT\ncalculations in the absence of an external \feld. (b) Zooming\nin the low energy regions of the spectrum a splitting of the\nspin{wave modes is observed (double spin{gap) due to the\nsystem's biaxial anisotropy.\npresence of magnetic and nonmagnetic Mn sites within\nthis phase, respectively. Further investigation is required\nto enlighten this phenomenon.\nThe INS data obtained in the collinear AFM2 phase\nof Mn 5Si3have demonstrated a spin excitation double\nenergy gap for zero \feld at Q= (1;2;0), see Fig. 4(b).\nThat is, the low energy spectrum is composed of two\nexcitations of nonvanishing and distinct energies. In\na system with uniaxial anisotropy, these two excita-\ntions would be degenerate. Using the parameters ob-\ntained in our \frst{principles simulations and given in\nthe previous sections, we calculated the inelastic scat-\ntering spectrum for this phase of our material, as seen in\nFig. 6(a). In the low energy region around the Brillouin\nzone center G= (1;2;0), we observe an energy gap of\nabout 3 meV. However, a closer look reveals the double\nenergy{gap structure which is due to the second easy{\naxis anisotropy predicted in our calculations, as demon-\nstrated by Fig. 6(b).\nDespite this qualitative agreement, these energy gaps\nare much higher than those in Fig. 4(b) obtained through\nthe neutron scattering measurements. We believe that\nthis discrepancy comes from the di\u000eculties associated\nwith modeling the spin{wave spectrum at \fnite temper-\nature. For example, the computed magnetic moments\n(2.4\u0016B) are higher than the ones obtained experimen-\ntally (1.48\u0016B)8, which could be explained by the ther-\nmal \ructuations of their orientations which have not been\ntheoretically considered. Possible e\u000bects associated with\nthe spin \ructuations of the Mn1 or Mn2 sites are also\nnot taken into account by our model Hamiltonian. An-\nother source of uncertainty is how temperature may a\u000bect\nthe magnetic interactions that de\fne the spin{wave ex-\ncitation spectrum. For instance, the e\u000bective magnetic7\nanisotropy energy was found to be strongly temperature{\ndependent in Mn 5Ge335, the ferromagnetic counterpart\nof Mn 5Si3.\nTo allow a direct comparison between the theoreti-\ncal spin{wave energies and the experimental data, we\nrescaled the DFT parameters to match the experimen-\ntal results. A good agreement was obtained by scaling\ndown uniformly the magnetic exchange interaction and\nthe anisotropy parameters by a factor of ten. For \fne\ntuning, the second easy{axis anisotropy parameter was\nadjusted from the scaled DFT value kc= 0:003 meV to\nkc= 0:009 meV.\nB. Results under external magnetic \feld\nUsing the rescaled parameters, we calculated the spin{\nwave energies as a function of the applied magnetic \feld\nto shed light on the observed behavior of the two low\nenergy magnon modes. To this aim, we included in our\nmodel Hamiltonian the Zeeman term:\nHZ=\u00002\u0016BX\niH\u0001Si; (9)\nto account for an external magnetic \feld applied along\nthecandacrystal axes. It is worth mentioning that pre-\nvious investigations in AFM systems with uniaxial and\nbiaxial anisotropy have described the in\ruence of an ex-\nternal magnetic \feld in the magnon properties36,37. How-\never, the observed behavior from our INS results shown\nin Fig. 4(b) is not su\u000eciently covered.\nThe results of our calculations for an external mag-\nnetic \feld applied along the ccrystal axis are shown in\nFig. 7(a). We observe that the energy of one mode is\ninsensitive to the magnetic \feld ( \u000b{mode), while the en-\nergy of the other mode ( \f{mode) increases monotoni-\ncally for larger \felds. Moreover, the \f{mode is lower in\nenergy than the \u000b{mode for zero \feld, and for increasing\n\feld strength they eventually cross, which is in qualita-\ntive agreement with the INS data shown in Fig. 4(b).\nNo avoided crossing is observed with our model. At zero\n\feld, the energy of the \u000b{mode is solely a function of\nkband the magnetic exchange interactions. The energy\ndi\u000berence between the two modes is determined by kc,\nwhich is zero for kc= 0. When kc=kb, the\f{mode gap\ncloses. These results are made explicit by our analytical\nstudy of a corresponding one{dimensional antiferromag-\nnetic system including nearest{neighbor{only exchange\ninteraction Jand a biaxial anisotropy. In the asymp-\ntotic limit of k\u001cJ38, the energies of two spin{wave\nmodes are E\u000b\u00192Sp\nJkbandE\f\u00192Sp\nJ(kb\u0000kc).\nWe note that kc> kbrepresents an unstable situation\nbecause the c{axis becomes the preferred one, but the\nmagnetic moments are aligned along b. If we apply the\n\feld alonga, the \feld dependence of these modes inverts,\nas shown in Fig. 7(b).\nHkc\nHka\u000b-mode\f-mode\n\u000b-mode\n\f-mode(a)\n(b)\n(c) (d)\n\u000b-mode\nab\nc\n\f-mode\nab\nc\nFIG. 7. Adiabatic spin{wave energies of Mn 5Si3as a function\nof the external \feld applied along (a) the c{axis (perpendicu-\nlar to the preferred axis but parallel to the second anisotropy\neasy{axis). The \f{mode disperses with the \feld because it\nhas linear polarization perpendicular to it. Meanwhile, the\n\u000b{mode has linear polarization parallel to the \feld (along the\nc{axis). (b) External \feld applied along the a{axis (parallel\nto the hard{axis). The \feld couples with the \u000b{mode which\nhas linear polarization perpendicular to it. In this case, no\nenergy crossing occurs. Lines in (a) and (b) are guides for the\neyes. (c) and (d) precessional dynamics of the spins for biaxial\nanisotropy and without an external magnetic \feld for modes\n\u000band\f, respectively. The red and blue arrows indicate an-\ntiparallel spins within a Mn2 triangle precessing around their\nequilibrium direction ( b{axis). The central gray arrows rep-\nresent the dynamics of the net magnetization of the system,\nwhich correspond to linear (longitudinal) precessions.\nC. Precessional nature of the spin{wave modes\nTo understand why the two spin{wave modes react so\ndi\u000berently to the external magnetic \feld, we now dis-\ncuss their precessional nature. For zero \feld and kc= 0\n(i.e. with uniaxial anisotropy), the modes are degener-8\nate as pointed out before. For these modes, the spins\nhave an elliptical precession because in a certain instant\nof their revolution they are perfectly antiparallel but a\nquarter of revolution later, they are noncollinear36. This\nnoncollinear alignment is unfavored by the exchange in-\nteraction, making the precession elliptical. The ellipses\ndescribing the precession of each mode have their major\naxes perpendicular to each other. However, there is no\npreferred orientation for these major axes as long as the\nsystem remains isotropic in the ac{plane. Besides, these\nmodes can also be globally characterized by the dynam-\nics of the net magnetization, which is de\fned by the sum\nof all spins at each instance of time. For these modes,\nthe net magnetization oscillates linearly along the minor\nellipse axis of the corresponding mode, as illustrated by\nthe central gray arrow in Fig. 7(c) and (d). These modes\nare said to be linearly polarized along the oscillation axis\nof the net magnetization. Furthermore, they have no net\nangular momentum and can be thought of as longitudinal\nmodes28.\nA preferred direction of the ellipse major axes can be\nachieved in two ways. Firstly, we can apply a magnetic\n\feld perpendicularly to the preferred axis, i.e., ?b{axis.\nAssuming that the \feld was applied along c, the\u000b{mode\nbecomes linearly polarized along c, and consequently, the\n\f{mode is polarized along a. An oscillating net magne-\ntization parallel to the \feld is not subjected to a torque\ndue to that \feld. Therefore, only the \f{mode with po-\nlarization perpendicular to the magnetic \feld is a\u000bected\nsuch that its energy increases with the \feld.\nSecondly, we can introduce a second anisotropy axis\nperpendicular to the \frst with weaker strength, let us\nsay alongc. The energy of the mode with the major axis\nof the ellipse along c(\f{mode) is reduced because the\nsystem gains energy when the spins tilt in that direction.\nThe energy of the \u000b{mode, which tilts mostly perpendic-\nularly toc, is not much a\u000bected. Adding the magnetic\n\feld also along cincreases the energy of the \f{mode as\nbefore, which eventually leads to the crossing of the en-\nergy of the two modes (see Figure 7(a)). Figure 7(b)\ndemonstrates that if we apply the magnetic \feld along\na, therefore, perpendicular to the linear polarization of\nthe\u000b{mode, then it would be the energy of this mode to\nincrease with the \feld.\nPolarized neutron scattering experiments around an\nAFM zone center capture the elliptic polarization of the\nmagnon modes. Therefore, the proposed behavior re-\ngarding the polarization of the \u000band\f{mode is also\nre\rected in the polarized INS spectra shown in Fig. 5.\nThe ellipses major axis of the processional motion of the\n\u000b{mode is along the a{axis and consistently the intensity\nof this mode will appear in the subtracted spin \ructua-\ntions spectra obtained from LPA in h\u000eMai. Equivalently,\nthe\f{mode shows maximum intensity in h\u000eMci, since the\nmajor axis of the processional motion is along the c{axis.\nHkc\nHka\u000b-mode\f-mode\n\u000b-mode\n\f-mode(a)\n(b)\nFIG. 8. E\u000bects of the DMI on the spin{wave dispersion of\nMn5Si3atQ= (1;2;0). The external \feld was applied along\n(a) thec{axis (perpendicular to the preferred axis but parallel\nto the second anisotropy easy{axis), and along (b) the a{axis\n(parallel to the hard{axis). The DMI increases the energy of\nthe\f{mode and its e\u000bect on the \u000b{mode is almost negligible.\njDijj=Dis in units of meV. Lines are guides for the eyes.\nD. Canting of magnetic moments\nPolarized single{crystal neutron di\u000braction experi-\nments8in the AFM2 phase of Mn 5Si3showed that the\nmagnetic moments are aligned along the b{axis and that\nthere are no components of moments parallel to the c{\naxis. However, a more recent neutron di\u000braction study\nperformed on polycrystalline samples7suggested a devi-\nation from the perfect collinearity, which is temperature{\ndependent increasing up to 8\u000enear 70 K with respect to\ntheb{axis, but with the moments still con\fned in the\nab{plane. Such a spin canting could originate from the\nDzyaloshinskii{Moriya39,40exchange interaction (DMI).\nTo investigate the \feld dependence of the two modes in\nthe presence of DMI, we included the following term into\nour model Hamiltonian:\nHDMI=\u0000X\nijDij\u0001(Si\u0002Sj): (10)\nThe results of the ab initio calculation indicated that\nthe DMI is indeed \fnite and smaller than 10% of the ex-\nchange interaction for the Mn2 pair coupled by J1where\nDijjj^c. The impact of DMI on the \u000band\f{mode for two9\n\feld directions is shown in Fig. 8. The e\u000bect of the DMI\nin the\u000b{mode is almost negligible, independently of the\ndirection of the external magnetic \feld. In contrast, the\nDMI causes the \f{mode energy to increase. For a \feld\napplied along the a{axis, the almost \rat dispersion of\nthe\f{mode shifts to higher energies with increasing D,\nwhich eventually results in a crossing with the \u000b{mode\n(see the blue curves in Fig. 8(b)). When the \feld is ap-\nplied along the c{axis, the increase in energy is accom-\npanied by a change of the \\parabolic\" dispersion of the\n\f{mode, as shown in Fig. 8(a). It is worth noticing that\nat zero \felds, the energy of the \u000b{mode is solely given\nby the exchange interaction and the \frst anisotropy axis,\nwhile the\f{mode energy is determined by the DMI in\ncombination with the second anisotropy axis. Based on\nour results from the INS measurements (see Fig. 4(b)),\nwe can conclude since a crossing of the two modes is\nobserved for Hk^c, that the DM exchange interaction\nshould be very weak in the AFM2 phase of Mn 5Si3.\nV. CONCLUSIONS\nWe investigated the microscopic magnetic properties\nof the collinear antiferromagnetic phase of Mn 5Si3with\ninelastic neutron scattering measurements and density\nfunctional theory simulations. The measurements have\nrevealed two low{energy spin{wave modes that disperse\nwith the wave{vector as expected for gapped antiferro-\nmagnetic magnons. The high resolution of the measure-\nments also allowed us to detect the small energy split-\nting between the two modes of about 0.2 meV. When a\nmagnetic \feld was applied to the sample along the c{\naxis, the two modes showed a very di\u000berent response\nto the \feld. The energy of the lower{energy mode in-\ncreased with increasing \feld, eventually crossing the en-ergy of the other mode, which showed no observable re-\nsponse to the \feld. These modes were further charac-\nterized by polarized neutron scattering measurements,\nwhich showed that their polarization is distinct and likely\nelliptical, with di\u000berent major axes for each ellipse. Our\n\frst{principles calculations capture the main features ob-\nserved so far experimentally, namely, the collinear ar-\nrangement of the magnetic moments, the coexistence\nof magnetic and nonmagnetic Mn sites, and the biax-\nial magnetocrystalline anisotropy. Also, we expanded\nthe existing studies for collinear antiferromagnets with\nbiaxial anisotropy regarding the polarization of the low\nenergy spin{wave modes and their evolution under the\nin\ruence of an external magnetic \feld. Finally, we inves-\ntigated theoretically the impact of the Dzyaloshinskii{\nMoriya interaction in the magnon spectrum and its im-\nplications depending on the strength of the anisotropy.\nOur results could \fnd echo in other similar systems where\ninteresting magnonic phenomena occur in the \feld of an-\ntiferromagnetic spintronics.\nVI. ACKNOWLEDGMENTS\nThe neutron data collected at the Institut Laue\nLangevin for the present work are available at 41{43.\nN.B. acknowledges the support of JCNS through the\nTasso Springer fellowship. This work was also supported\nby the Brazilian funding agency CAPES under Project\nNo. 13703/13-7 and the European Research Council\n(ERC) under the European Union's Horizon 2020 re-\nsearch and innovation program (ERC-consolidator Grant\nNo. 681405-DYNASORE). We gratefully acknowledge\nthe computing time granted by JARA-HPC on the su-\npercomputer JURECA at Forschungszentrum J ulich and\nby RWTH Aachen University.\n\u0003f.dos.santos@fz-juelich.de (currently based at EPFL: \ra-\nviano.dossantos@ep\r.ch)\nyn.biniskos@fz-juelich.de\n1V. Baltz, A. Manchon, M. Tsoi, T. Moriyama, T. Ono,\nand Y. Tserkovnyak, Rev. Mod. Phys. 90, 015005 (2018).\n2P. J. Brown, J. B. Forsyth, V. Nunez, and F. Tasset,\nJournal of Physics: Condensed Matter 4, 10025 (1992).\n3C. S urgers, G. Fischer, P. Winkel, and H. v. L ohneysen,\nNature Communications 5, 3400 (2014).\n4N. Biniskos, K. Schmalzl, S. Raymond, S. Petit, P. Ste\u000bens,\nJ. Persson, and T. Br uckel, Physical Review Letters 120,\n257205 (2018).\n5B. Das, B. Balasubramanian, P. Manchanda, P. Mukher-\njee, R. Skomski, G. C. Hadjipanayis, and D. J. Sellmyer,\nNano Letters 16, 1132 (2016), publisher: American Chem-\nical Society.\n6Y. Sun, B. Sun, J. He, G. Yang, and C. Wang, Nature\nCommunications 11, 647 (2020), number: 1 Publisher: Na-\nture Publishing Group.7M. Gottschilch, O. Gourdon, J. Persson, C. d. l. Cruz,\nV. Petricek, and T. Brueckel, Journal of Materials Chem-\nistry 22, 15275 (2012), publisher: The Royal Society of\nChemistry.\n8P. J. Brown and J. B. Forsyth, Journal of Physics: Con-\ndensed Matter 7, 7619 (1995).\n9C. S urgers, T. Wolf, P. Adelmann, W. Kittler, G. Fischer,\nand H. v. L ohneysen, Scienti\fc Reports 7, 42982 (2017).\n10L. Vinokurova, V. Ivanov, E. Kulatov, and A. Vlasov,\nJournal of Magnetism and Magnetic Materials 90-91 , 121\n(1990).\n11R. F. Luccas, G. S\u0013 anchez-Santolino, A. Correa-Orellana,\nF. J. Mompean, M. Garc\u0013 \u0010a-Hern\u0013 andez, and H. Suderow,\nJournal of Magnetism and Magnetic Materials 489, 165451\n(2019).\n12S. C. Das, K. Mandal, P. Dutta, S. Pramanick, and\nS. Chatterjee, Physical Review B 100, 024409 (2019), pub-\nlisher: American Physical Society.\n13M. R. Silva, P. J. Brown, and J. B. Forsyth, Journal of\nPhysics: Condensed Matter 14, 8707 (2002), publisher:10\nIOP Publishing.\n14H. Al-Kanani and J. Booth, Journal of Magnetism and\nMagnetic Materials 140-144 , 1539 (1995), international\nConference on Magnetism.\n15A. Irizawa, A. Yamasaki, M. Okazaki, S. Kasai,\nA. Sekiyama, S. Imada, S. Suga, E. Kulatov, H. Ohta, and\nT. Nanba, Solid State Communications 124, 1 (2002).\n16C. S urgers, W. Kittler, T. Wolf, and H. v.\nL ohneysen, AIP Advances 6, 055604 (2016),\nhttps://doi.org/10.1063/1.4943759.\n17D. Songlin, O. Tegus, E. Br uck, J. C. P. Klaasse, F. R.\nde Boer, and K. H. J. Buschow, Journal of Alloys and\nCompounds 334, 249 (2002).\n18K. Schmalzl, W. Schmidt, S. Raymond, H. Feilbach,\nC. Mounier, B. Vettard, and T. Br uckel, Nuclear Instru-\nments and Methods in Physics Research Section A: Accel-\nerators, Spectrometers, Detectors and Associated Equip-\nment 819, 89 (2016).\n19L. P. Regnault, B. Ge\u000bray, P. Fouilloux, B. Longuet,\nF. Mantegazza, F. Tasset, E. Leli\u0012 evre-Berna, S. Pujol,\nE. Bourgeat-Lami, N. Kernavanois, M. Thomas, and\nY. Gibert, Physica B: Condensed Matter Proceedings of\nthe Third European Conference on Neutron Scattering,\n350, E811 (2004).\n20T. Chatterji, ed., Neutron scattering from magnetic mate-\nrials, 1st ed. (Elsevier, Amsterdam ; Boston, 2006) oCLC:\nocm61748467.\n21In this article we use the orthorhombic coordinate system\nand the scattering vector Qis expressed in Cartesian coor-\ndinates Q= (Qh;Qk;Ql) given in reciprocal lattice units\n(r.l.u.). The wave{vector qis related to the momentum\ntransfer through ~Q=~G+~q, where Gis a Brillouin\nzone center and G= (h;k;l ).\n22S. Ibuka, S. Itoh, T. Yokoo, and Y. Endoh, Phys. Rev. B\n95, 224406 (2017).\n23N. Papanikolaou, R. Zeller, and P. H. Dederichs, Journal\nof Physics: Condensed Matter 14, 2799 (2002), publisher:\nIOP Publishing.\n24S. Vosko, L. Wilk, and M. Nusair, Can. J. Phys. 58, 1200\n(1980).\n25K. Wildberger, P. Lang, R. Zeller, and P. H. Dederichs,\nPhysical Review B 52, 11502 (1995), publisher: American\nPhysical Society.\n26A. I. Liechtenstein, M. I. Katsnelson, V. P. Antropov, and\nV. A. Gubanov, Journal of Magnetism and Magnetic Ma-terials 67, 65 (1987).\n27H. Ebert and S. Mankovsky, Phys. Rev. B 79, 045209\n(2009).\n28F. J. dos Santos, M. dos Santos Dias, F. S. M. Guimar~ aes,\nJ. Bouaziz, and S. Lounis, Physical Review B 97, 024431\n(2018).\n29T. Holstein and H. Primako\u000b, Physical Review 58, 1098\n(1940).\n30N. N. Bogoljubov, Il Nuovo Cimento (1955-1965) 7, 794\n(1958).\n31F. J. dos Santos, M. dos Santos Dias, and S. Lounis, Phys-\nical Review B 102, 104401 (2020), publisher: American\nPhysical Society.\n32F. J. dos Santos, M. dos Santos Dias, and S. Lounis, Phys-\nical Review B 102, 104436 (2020), publisher: American\nPhysical Society.\n33F. J. dos Santos, M. dos Santos Dias, and S. Lounis,\nPhysical Review B 95, 134408 (2017), publisher: Amer-\nican Physical Society.\n34E. Mendive-Tapia and J. B. Staunton, Phys. Rev. B 99,\n144424 (2019).\n35N. Maraytta, J. Voigt, C. Salazar Mej\u0013 \u0010a, K. Friese, Y. Sk-\nourski, J. Per\u0019on, S. M. Salman, and T. Br uckel,\nJournal of Applied Physics 128, 103903 (2020),\nhttps://doi.org/10.1063/5.0020780.\n36F. Ke\u000ber and C. Kittel, Physical Review 85, 329 (1952),\npublisher: American Physical Society.\n37S. M. Rezende, A. Azevedo, and R. L. Rodr\u0013 \u0010guez-Su\u0013 arez,\nJournal of Applied Physics 126, 151101 (2019), publisher:\nAmerican Institute of Physics.\n38F. J. dos Santos, First-principles study of collective\nspin excitations in noncollinear magnets , Ph.D. thesis,\nForschungszentrum J ulich GmbH, Zentralbibliothek, Ver-\nlag (2020), iSBN: 9783958064591 Number: RWTH-2020-\n01879.\n39I. Dzyaloshinsky, Journal of Physics and Chemistry of\nSolids 4, 241 (1958).\n40T. Moriya, Phys. Rev. 120, 91 (1960).\n41https://doi.ill.fr/10.5291/ILL-DATA.CRG-2331 (2016).\n42https://doi.ill.fr/10.5291/ILL-DATA.CRG-2620 (2019).\n43https://doi.ill.fr/10.5291/ILL-DATA.4-01-1618 (2019)." }, { "title": "2011.05722v2.Characterization_of_room_temperature_in_plane_magnetization_in_thin_flakes_of_CrTe__2__with_a_single_spin_magnetometer.pdf", "content": "Characterization of room-temperature in-plane magnetization in thin \rakes of CrTe 2\nwith a single spin magnetometer\nF. Fabre,1A. Finco,1A. Purbawati,2A. Hadj-Azzem,2N. Rougemaille,2J. Coraux,2I. Philip,1and V. Jacques1\n1Laboratoire Charles Coulomb, Universit\u0013 e de Montpellier and CNRS, 34095 Montpellier, France\n2Univ. Grenoble Alpes, CNRS, Grenoble INP, Institut NEEL, 38000 Grenoble, France\n(Dated: February 12, 2021)\nWe demonstrate room-temperature ferromagnetism with in-plane magnetic anisotropy in thin\n\rakes of the CrTe 2van der Waals ferromagnet. Using quantitative magnetic imaging with a single\nspin magnetometer based on a nitrogen-vacancy defect in diamond, we infer a room-temperature\nin-plane magnetization in the range of M\u001827 kA/m for \rakes with thicknesses down to 20 nm.\nIn addition, our measurements indicate that the orientation of the magnetization is not determined\nsolely by shape anisotropy in micron-sized CrTe 2\rakes, which suggest the existence of a non-\nnegligible magnetocrystalline anisotropy. These results make CrTe 2a unique system in the growing\nfamily of van der Waals ferromagnets, as it is the only material platform known to date which o\u000bers\nan intrinsic in-plane magnetization and a Curie temperature above 300 K in thin \rakes.\nI. INTRODUCTION\nFerromagnetic van der Waals (vdW) crystals o\u000ber nu-\nmerous opportunities both for the study of exotic mag-\nnetic phase transitions in low-dimensional systems [1] and\nfor the design of innovative, atomically-thin spintronic\ndevices [2, 3]. Since the discovery of a two-dimensional\n(2D) magnetic order in monolayers of CrI 3[4] and\nCr2Ge2Te6[5] crystals, the family of vdW ferromagnets\nhas expanded very rapidly [6{8]. However, most of these\ncompounds have a Curie temperature ( Tc) well below 300\nK, which appears as an important drawback for future\ntechnological applications. An intense research e\u000bort is\ntherefore currently devoted to the identi\fcation of high-\nTc2D magnets [3].\nIn this context, the vdW crystal Fe 3GeTe 2appears as\na serious candidate because it can be grown in wafer-\nscale through molecular beam epitaxy and it exhibits a\nstrong perpendicular magnetic anisotropy [9, 10]. Al-\nthough its intrinsic Tcdrops to 130 K in the monolayer\nlimit [11], it might be raised above room-temperature ei-\nther by ionic gating [12, 13], interfacial engineering [14],\nor by micro-patterning, as demonstrated so far for rather\nthick \flms [15]. In addition, other FeGeTe alloys, such\nas Fe 4GeTe 2and Fe 5GeTe 2, exhibit high Tc, still lower\nthan room temperature but close to it [16, 17]. An-\nother promising strategy consists in incorporating mag-\nnetic dopants into 2D materials to form dilute magnetic\nsemiconductors [18]. This approach was recently em-\nployed to induce room-temperature ferromagnetism in\nWSe 2monolayers doped with vanadium [19, 20]. Finally,\nan intrinsic ferromagnetic order was reported in epitax-\nial layers of VSe 2[21], MnSe 2[22] and VTe 2[23] under\nambient conditions, although the interpretation of these\nexperiments still remains debated [24, 25].\nIn this work, we follow an alternative research direc-\ntion by studying the room-temperature magnetic proper-\nties of micron-sized \rakes exfoliated from a CrTe 2crystal\nwith 1Tstructure. In its bulk form, this layered transi-\ntion metal dichalcogenide is a ferromagnet with in-planemagnetization, i.e.pointing perpendicular to the caxis,\nand aTcaround 320 K [26]. This combination of proper-\nties is unique in the growing family of vdW ferromagnets.\nRecent studies have reported that the magnetic order is\npreserved at room temperature in exfoliated CrTe 2\rakes\nwith thicknesses in the range of a few tens of nanome-\nters [27, 28]. However, obtaining quantitative estimates\nof the magnetization in such micron-sized \rakes remains\na di\u000ecult task, which requires the use of non-invasive\nmagnetic microscopy techniques combining high sensitiv-\nity with high spatial resolution. These performances are\no\u000bered by magnetometers employing a single nitrogen-\nvacancy (NV) defect in diamond as an atomic-size quan-\ntum sensor [29{31]. In recent years, this microscopy tech-\nnique has found many applications in condensed matter\nphysics [32], including the study of chiral spin textures in\nultrathin magnetic materials [33{35], current \row imag-\ning in graphene [36] and the analysis of the magnetic or-\nder in vdW magnets down to the monolayer limit [37{39].\nHere we use scanning-NV magnetometry to infer quan-\ntitatively the in-plane magnetization in exfoliated CrTe 2\n\rakes under ambient conditions. Our measurements con-\n\frm that the ferromagnetic order is preserved in few tens\nof nanometers thick \rakes, although with a low room-\ntemperature magnetization M\u001827 kA/m. This value\nis \fve times smaller that the one measured in a bulk\nCrTe 2crystal. Such a reduction of the magnetization is\nattributed to a decreased Curie temperature in exfoliated\n\rakes. Moreover, our results show that shape anisotropy\nalone does not \fx the in-plane orientation of the mag-\nnetization in micron-sized CrTe 2\rakes, pointing out the\nexistence of a substantial magnetocrystalline anisotropy.\nII. MATERIALS AND METHODS\nA bulk 1T-CrTe 2crystal was synthesized following the\nprocedure described in Ref. [26]. The in-plane magnetiza-\ntion of this layered ferromagnet was \frst characterized as\na function of temperature through vibrating sample mag-arXiv:2011.05722v2 [cond-mat.mtrl-sci] 11 Feb 20212\nnetometry under a magnetic \feld of 500 mT. The results\nshown in Fig. 1(a) indicate a Curie temperature around\n320 K and a magnetization reaching M\u0018120 kA/m\nunder ambient conditions. CrTe 2\rakes with thicknesses\nranging from a few tens to a hundred of nanometers were\nthen obtained by mechanical exfoliation and transferred\non a SiO 2/Si substrate. We note that the probability\nto obtain thin CrTe 2\rakes through mechanical exfolia-\ntion is still very low compared to other layered transi-\ntion metal dichalcogenides, such as MoS 2or WSe 2[28].\nThe thinnest \rake studied in this work has a thickness\nof\u001820 nm. Like all van der Waals ferromagnets known\nto date, CrTe 2\rakes are unstable under oxygen atmo-\nsphere. However, a recent study combining X-ray and\nRaman spectroscopy has shown that oxidation of CrTe 2\n\rakes occurs typically within a day scale under ambient\nconditions and is limited to the very \frst outer layers [28].\nIn this work, CrTe 2\rakes were not encapsulated and all\nthe measurements were done within a day after exfolia-\ntion to mitigate oxidation.\nMagnetic imaging was performed with a scanning-NV\nmagnetometer operating under ambient conditions [31].\nAs sketched in Fig. 1(b), a single NV defect integrated\ninto the tip of an atomic force microscope (AFM) was\nscanned above CrTe 2\rakes to probe their stray mag-\nnetic \felds. At each point of the scan, a confocal optical\nmicroscope placed above the tip was used to monitor the\nmagnetic-\feld-dependent photoluminescence (PL) prop-\nerties of the NV defect under green laser illumination.\nIn this work, we employed a commercial diamond tip\n(Qnami, Quantilever MX) with a characteristic NV-to-\nsample distance dNV= 80\u000610 nm, as measured through\nan independent calibration procedure [40]. Two di\u000berent\nmagnetic imaging modes were used. In the limit of weak\nstray \felds ( <5 mT), quantitative magnetic \feld map-\nping was obtained by recording the Zeeman shift of the\nNV defect electron spin sublevels through optical detec-\ntion of the electron spin resonance (ESR). This method\nrelies on microwave driving of the NV spin transition\ncombined with the detection of the spin-dependent PL\nintensity of the NV defect [41]. For stronger magnetic\n\felds (>5 mT), the scanning-NV magnetometer was\nrather used in all-optical, PL quenching mode [42, 43].\nIn this case, localized regions of the sample producing\nlarge stray \felds are simply revealed by an overall re-\nduction of the PL signal induced by a mixing of the NV\ndefect spin sublevels [44]. We note that the diameter of\nthe scanning diamond tip is around 200 nm in order to\nact as an e\u000ecient waveguide for the PL emission of the\nNV defect [45, 46]. As a result, such tips cannot provide\nprecise topographic information of the sample. For the\nthinnest \rakes, the topography was thus imaged by using\nconventional, sharp AFM tips.\nFor a ferromagnetic material, stray magnetic \felds can\nbe produced by magnetization patterns presenting a non-\nzero divergencer\u0001M6= 0. Considering a CrTe 2\rake,\nthis can occur (i) at the edges of the \rake, (ii) at the\nlocation of non-collinear spin textures such as domain\n250200150100500 M (kA/m)300200100 Temperature (K)(a)\n(b)\nSubstratexyzNVaxis\u0000NV✓NVdiamond tipdNVMObjectivePLbulk CrTe2FIG. 1. (a) Temperature dependence of the magnetization M\nin a bulk crystal of CrTe 2with 1Tpolytype. The measure-\nment is performed by vibrating sample magnetometry under\na magnetic \feld of 500 mT. At room temperature, the magne-\ntization is around 120 kA/m (black dashed lines). The inset\nshows the layered crystal structure of 1 T-CrTe 2. (b) A single\nNV defect (red arrow) localized at the apex of a diamond tip\nis scanned above exfoliated CrTe 2\rakes. A microscope ob-\njective placed above the tip is used to collect the photolumi-\nnescence (PL) of the NV defect under green laser excitation.\nIn this work, the NV-to-sample distance is dNV\u001880 nm and\nthe quantization axis of the NV defect is characterized by the\nspherical angles ( \u0012NV= 58\u000e;\u001eNV= 103\u000e) in the laboratory\nframe of reference ( x;y;z ). The black arrows indicate the\nmagnetic \feld lines generated at the edges of a CrTe 2\rake\nwith uniform in-plane magnetization M.\nwalls, and (iii) at the position of thickness steps on the\n\rake. In this work, the magnetization is inferred from the\nmeasurement of the stray \feld produced at the edges of\nthe \rake [Fig. 1(b)]. Therefore, the analysis is made sim-\npler for uniformly magnetized \rakes with a homogeneous\nthickness. While uniform spin textures can be readily ob-\ntained by applying a bias magnetic \feld, obtaining thin\nCrTe 2\rakes with a uniform thickness through mechani-\ncal exfoliation is currently achieved with a low probabil-\nity. Consequently, variations in thickness must be care-\nfully taken into account when analyzing the experimental\nresults.3\nxy\n0 100 200 300\nHeight (nm)Topography(a)\nPL scan(b)\n0.6 0.8 1\nNormalized PLSimulation\nφM=20°\nM(c)\nSimulation\nφM=-80°\nM(d)\nFIG. 2. (a) AFM image of a 150 nm-thick CrTe 2\rake recorded with the scanning-NV magnetometer. (b) Corresponding PL\nmap showing a quenching contour at the edges of the \rake. The shape of the \rake is indicated by the black dashed lines.\nThe recorded PL signal is normalized with the one measured far from the \rake. The experiment is performed at zero external\nmagnetic \feld. (c,d) Simulated PL maps for two di\u000berent orientations \u001eMof the in-plane magnetization: (c) \u001eM= 20\u000eand\n(d)\u001eM=\u000080\u000e. Here, the norm of the magnetization is \fxed arbitrarily to its bulk value M= 120 kA/m. The PL quenching\ninduced by energy transfer between the NV sensor and the metallic sample is not included in the simulations.\nIII. RESULTS AND DISCUSSION\nIn a \frst experiment, we imaged a CrTe 2\rake with\na large thickness t\u0018150 nm [Fig. 2(a)]. Considering\nthe saturation magnetization of the bulk CrTe 2crystal,\nmagnetic simulations predict stray \feld amplitudes larger\nthan 10 mT at a distance dNV\u001880 nm above the edges\nof the 150-nm-thick \rake. The scanning-NV magnetome-\nter was thus operated in the PL quenching mode for such\na thick \rake. Fig. 2(b) displays the resulting PL map.\nSeveral observations can be made. First, a quenching of\nthe PL signal is observed when the NV defect is placed\nabove the \rake. This quenching e\u000bect, which is constant\nall over the \rake, does not have a magnetic origin. It is\nlinked to the metallic character of CrTe 2[47, 48]. Second,\na stronger PL quenching is obtained at the \rake edges,\nas expected for a single ferromagnetic domain. We note\nthat the additional quenching spot observed at the top-\nright edge of the \rake results from the stray \feld pro-\nduced by a large variation of the thickness [Fig. 2(a)].\nGiven the di\u000berent sources of PL quenching involved in\nthis experiment, quantitative estimates of the magneti-\nzation cannot be obtained. However, the analysis of the\nPL quenching distribution can be used to extract qual-\nitative information about the orientation of the magne-\ntization. To this end, simulations of the PL map were\ncarried out considering only the e\u000bect of magnetic \felds.\nThe geometry of the \rake used for the simulation was\ninferred from the AFM image simultaneously recorded\nwith the NV microscope [Fig. 2(a)]. Assuming a uniform\nin-plane magnetization Mwith an azimuthal angle \u001eM\nin the (x;y) plane, the stray \feld was \frst calculated\nat a distance dNVabove the \rake. The resulting PL\nquenching map was then simulated by using the model\nof magnetic-\feld-dependent photodynamics of the NVdefect described in Ref. [44]. Simulated PL maps ob-\ntained for two di\u000berent orientations of the magnetization\nare shown in Figs. 2(c,d). In micron-sized \rakes, shape\nanisotropy should favor a magnetization pointing along\nthe long axis of the \rake. Interestingly, the PL map sim-\nulated for such a magnetization direction disagrees with\nthe experimental data [Fig. 2(c)], which instead suggest\na magnetization pointing perpendicular to the long axis\nof the \rake [Fig. 2(d)]. This result is a \frst indication of\na non-negligible, in-plane magnetocrystalline anisotropy.\nThinner CrTe 2\rakes, i.e.producing less stray \feld,\nwere then studied through quantitative magnetic \feld\nimaging. To this end, a microwave excitation was ap-\nplied through a copper microwire directly spanned on the\nsample surface and the Zeeman shift of the NV defect's\nESR frequency was recorded at each point of the scan. In\nthe weak \feld regime, the ESR frequency is shifted lin-\nearly with the projection BNVof the stray magnetic \feld\nalong the NV defect quantization axis [31]. This axis\nwas precisely measured by applying a calibrated mag-\nnetic \feld [49], leading to the spherical angles \u0012NV= 58\u000e\nand\u001eNV= 103\u000ein the laboratory frame ( x;y;z ), as illus-\ntrated in Fig. 1(b). For these measurements, a bias \feld\nof 3 mT was applied along the NV defect axis in order\nto infer the sign of the stray magnetic \feld [31]. Fur-\nthermore, this bias \feld is strong enough to erase spin\ntextures such as domain walls, given the weak coercitive\n\feld of CrTe 2\rakes [28].\nA map ofBNVrecorded above a 35-nm thick CrTe 2\n\rake is shown in Fig. 3(b). A stray magnetic \feld around\n\u00061:5 mT is detected at two opposite edges of the \rake.\nIn principle, the underlying sample magnetization can\nbe retrieved from such a quantitative magnetic \feld map\nby using reverse propagation methods with well-adjusted\n\flters in Fourier-space [32]. Under several assumptions,4\nxy\n0 20 40\nHeight (nm)Topography (a)\n−1.5 0 1.5\nBNV(mT)NV image (b)\n Simulation\nM(c)\n Line profilesBNV(mT)\nLateral displacement (µm)(d)\nFIG. 3. (a) AFM image of a 35 nm-thick CrTe 2\rake. (b) Corresponding distribution of the magnetic \feld component BNV.\n(c) Simulated magnetic \feld distribution BNVfor a uniform in-plane magnetization M(black arrow) with an azimuthal angle\n\u001eM=\u0000145\u000ein the (x;y) plane and a norm M= 27 kA/m. The calculation is done at a distance dNV= 80 nm above the \rake,\nwhose shape is extracted from the topography image [black dashed lined in (a)]. (d) Fitting of an experimental line pro\fle\n(black markers) with the magnetic calculation (blue solid line). The position of the linecuts are indicated by the white dashed\nlines in (b) and (c). A magnetization M= 27\u00064 kA/m is obtained, the uncertainty being illustrated by the blue shaded area.\nThe red and green solid lines indicate the result of the calculation for M= 40 kA/m and M= 15 kA/m, respectively.\nthis method can be quite robust for magnetic materials\nwith out-of-plane magnetization [37{39]. For in plane\nmagnets, however, the reconstruction procedure ampli-\n\fes noise and is thus much less e\u000ecient [50]. As a re-\nsult, the recorded magnetic \feld distribution was rather\ndirectly compared to magnetic calculations in order to\nextract quantitative information on the sample magneti-\nzation.\nTo obtain precise information on the geometry of the\n\rake, the topography image was here measured with a\nsharp AFM tip [Fig. 3(a)]. The observed variations in\nthickness were included in the magnetic calculation (see\nAppendix A). Considering a uniformly magnetized \rake\nwith an azimuthal angle \u001eM, the stray \feld was calcu-\nlated at a distance dNVabove the \rake and then pro-\njected along the NV quantization axis in order to sim-\nulate a map of BNV. By comparing experimental data\nwith magnetic maps simulated for di\u000berent values of the\nangle\u001eM, the orientation of the in-plane magnetization\nwas \frst identi\fed, leading to \u001eM=\u0000145\u00065\u000e,i.e.\npointing along the short axis of the \rake [Fig. 3(c)].\nOnce again, this result suggests the presence of a non-\nnegligible magnetocrystalline anisotropy, in agreement\nwith recent works [28]. The norm Mof the magneti-\nzation vector was then estimated by \ftting experimental\nline pro\fles with the result of the calculation, leading to\nM= 27\u00064 kA/m [Fig. 3(d)]. We note that the stray \feld\namplitude depends on several parameters including dNV,\n\u001eM, the \rake thickness tand the NV defect quantization\naxis (\u0012NV;\u001eNV). Any imprecisions on these parameters\ndirectly translate into uncertainties on the evaluation of\nthe magnetization M. A detailed analysis of uncertain-\nties is provided in Appendix B.\nThe room-temperature magnetization measured in the\nexfoliated CrTe 2\rake is almost \fve times smaller thanthe one obtained in the bulk crystal. This observation\ncould be explained by a degradation of the sample surface\nthrough oxidation processes, leading to a thinner e\u000bective\nmagnetic thickness. However, recent experiments relying\non X-ray magnetic circular dichroism coupled to photoe-\nmission electron microscopy (XMCD-PEEM) have shown\nthat oxidation is limited to the \frst outer layers of the\nCrTe 2\rake [28]. Considering a 35-nm thick CrTe 2\rake,\nsurface oxidation can thus be safely neglected, and can-\nnot explain the observed reduction of the magnetization.\nThis e\u000bect is rather attributed to a decrease of the Curie\ntemperature in exfoliated \rakes, a phenomenon that has\nbeen observed in other vdW magnets such as Fe 3GeTe 2\nbelow 5-10 nm thicknesses [11, 51], and in more tra-\nditional ferromagnetic thin \flms below few nanometers\nthicknesses [52, 53]. The thickness marking the crossover\nfrom a thin \flm (2D-like) to a bulk magnetism (3D) is\nexpected to be strongly material-dependent and cannot\nbe predicted a priori in the case of CrTe 2[8]. However,\nfor a 35-nm thick \rake, bulk-like magnetism is a rea-\nsonable assumption and the magnetization's amplitude\nshould not be altered by the \flm thickness. In this thick-\nness regime, the main parameter altering the magnetiza-\ntion's amplitude is temperature, and how close or far it\nis fromTc. The estimation of the reduction in Curie\ntemperature was tentatively inferred by translating the\nCurie law measured for a bulk CrTe 2crystal [Fig. 1(a)]\nin order to reach M\u001827 kA/m at room temperature.\nThis is achieved for a reduction of Tcby\u001820 K. This\nvalue is in line with a recent study, which estimates a\nCurie temperature around 300 K for CrTe 2\rakes with a\nfew tens of nanometers thickness [27].\nTo support this \fnding, a similar analysis was per-\nformed for a 20-nm thick \rake [Fig. 4(a)]. Here, a stray\nmagnetic \feld is mainly detected along the bottom edge5\n0 15 30 45\nHeight (nm)Topography(a)\n−0.7 0 0.7\nBNV(mT)NV image(b)\n Simulation\nM(c)\nBNV(mT)\nLateral displacement (µm)(d)\nFIG. 4. (a) AFM image of a 20-nm thick CrTe 2\rake. (b)\nCorresponding map of the magnetic \feld component BNV. (c)\nSimulated map of BNVfor a uniform in-plane magnetization\nM(black arrow) with an azimuthal angle \u001eM=\u0000100\u000ein the\n(x;y) plane and a norm M= 26 kA/m. The black dashed line\nindicate the shape of the \rake used for the simulation. (d)\nFitting of an experimental line pro\fle (black markers) with\nthe magnetic simulation (blue solid line). The linecuts are\nindicated by the white dashed lines in (b) and (c). A mag-\nnetizationM= 26\u00064:0 kA/m is obtained, the uncertainty\nbeing illustrated by the blue shaded area. The inset shows a\nlinecut across the \rake (white dashed lines in (a)).\nof the \rake. On the opposite edge, the stray \feld distri-\nbution is quite inhomogeneous [Fig. 4(b)]. This observa-\ntion is attributed to damages of the \rake, which can be\nobserved in the AFM image. It is di\u000ecult to describe the\ncorresponding complex thickness variations, and hence to\ntake them into account in the simulations. These height\nvariations seem to correspond to several CrTe 2islands,\nwhose very small sizes could make them more prone to\noxidation than larger \rakes, and which may exhibit ran-\ndom magnetization orientations. We hence disregarded\nthe thickness variations in our structural model and the\nmagnetic calculation was performed for a \rake with uni-\nform thickness, which is a good approximation for the\nbottom and left edges of the \rake. A simulation of the\nstray \feld distribution for a magnetization with an az-\nimuthal angle \u001eM=\u0000100\u000610\u000ereproduces fairly well the\nexperimental results [Fig. 4(c)] and a quantitative anal-\nysis of line pro\fles across the bottom edge of the \rake\nleads toM= 26\u00064:0 kA/m [Fig. 4(d)], a similar value\nto that obtained for the 35-nm thick \rake. This observa-tion indicates that the magnetization is not signi\fcantly\nmodi\fed for thicknesses lying in the few tens of nanome-\nters range. A more in-depth analysis of the dependence\nof magnetization on thickness could not be carried out at\nthis stage, given the very low probability to obtain thin\nCrTe 2\rakes of homogeneous thickness through mechan-\nical exfoliation.\nIV. CONCLUSION\nTo conclude, we have used quantitative magnetic imag-\ning with a scanning-NV magnetometer to demonstrate\nthat exfoliated CrTe 2\rakes with thicknesses down to\n20 nm exhibit an in-plane ferromagnetic order at room\ntemperature with a typical magnetization in the range of\nM\u001827 kA/m. These results make CrTe 2a unique sys-\ntem in the growing family of vdW ferromagnets, because\nit is the only material platform known to date which of-\nfers an intrinsic in-plane magnetization and a Tcabove\nroom temperature in thin \rakes. These properties might\no\u000ber several opportunities for studying magnetic phase\ntransition in 2D- XY systems [54] and to design spin-\ntronic devices based on vdW magnets. The next chal-\nlenge will be to assess if the ferromagnetic order is pre-\nserved at room temperature in the few layers limit.\nACKNOWLEDGEMENTS\nThe authors warmly thank J. Vogel and M. N\u0013 u~ nez-\nRegueiro for fruitful discussions. This research has re-\nceived funding from the European Union H2020 Program\nunder Grant Agreement No. 820394 (ASTERIQs), the\nDARPA TEE program, and the Flag-ERA JTC 2017\nproject MORE-MXenes. A.F. acknowledges \fnancial\nsupport from the EU Horizon 2020 Research and Innova-\ntion program under the Marie Sklodowska-Curie Grant\nAgreement No. 846597 (DIMAF).\nAPPENDIX A: MAGNETIC FIELD SIMULATION\nAs indicated in the main text, thickness variations\nwithin the \rake can result in a magnetization pattern\nwith a non-zero divergence which produces a stray mag-\nnetic \feld. When possible, these variations were taken\ninto account in the magnetic calculation. This is illus-\ntrated in Fig. 5(a,b), where the geometry of the \rake\nused for the calculation includes a thickness step ob-\nserved in the topography image. In Fig. 5(c-e), we show\nthe magnetic \feld distributions produced at a distance\ndNV= 80 nm for three di\u000berent magnetization orien-\ntations of the \rake. First considering an out-of-plane\nmagnetization [Fig. 5(c)], the simulated magnetic \feld\ndistribution does not agree with the experimental data\nshown in Fig. 3. This con\frms that the magnetization in\nlying in the plane of the CrTe 2\rake, i.e.perpendicular6\n02040Height (nm)Topography(a)\nt=35 nmt=25 nm(b)\nMOut-of-plane\nmagnetization(c)\nIn-plane magnetization\nΦM=-30°\nM(d)\n−1.501.5\nSimulated BNV(mT)In-plane magnetization\nΦM=-145°\nM(e)\nFIG. 5. (a) AFM image of CrTe 2\rake shown in Fig. 3. (b) Geometry of the \rake used for the magnetic calculation, which\nincludes the thickness step observed in the AFM imafge. (c-e) Calculated maps of BNVfor a uniform magnetization Mpointing\nout-of plane (c), and in-plane with an azimuthal angle \u001eM=\u000030\u000e(d) or\u001eM=\u0000145\u000e(e). The norm of the magnetization is\n\fxed toM= 27 kA/m.\nto thecaxis as obtained in the bulk crystal. Simula-\ntions performed for a uniform in-plane magnetization for\ntwo di\u000berent values of the azimuthal angle \u001eMin the\n(x;y) plane are shown in Fig. 5(d,e). A comparison with\nexperimental data allows the identi\fcation of the mag-\nnetization orientation.\nAPPENDIX B: ANALYSIS OF UNCERTAINTIES\nThe norm of the magnetization Mis obtained by\n\ftting line pro\fles of the measured stray \feld dis-\ntribution with the result of the magnetic calculation\n[see Fig. 3(d)]. In this section, we analyze the uncertainty\nof this measurement by using the methodology described\nin Ref. [55]. The uncertainties result (i) from the \ftting\nprocedure itself and (ii) from those on the parameters\npi=fdNV;\u0012NV;\u001eNV;t;\u001e Mgwhich are all involved in the\nmagnetic calculation. In the following, the parameters\npiare expressed as pi= \u0016pi+\u001bpi, where \u0016pidenotes the\nnominal value of parameter piand\u001bpiits standard error.\nThese parameters are independently evaluated as follows:\n\u000fThe probe-to-sample distance dNVis inferred\nthrough a calibration measurement, following the\nprocedure described in Ref. [40], leading to dNV=\n80\u000610 nm.\n\u000fThe NV defect quantization axis is measured by\nrecording the ESR frequency as a function of theamplitude and orientation of a calibrated mag-\nnetic \feld, leading to spherical angles ( \u0012NV=\n58\u00062\u000e;\u001eNV= 103\u00062\u000e) in the laboratory frame of\nreference (x;y;z ) [see Fig. 1(b)].\n\u000fThe thickness tof the CrTe 2\rake is extracted from\nline pro\fles of the AFM image with an uncertainty\nof\u00062 nm.\n\u000fThe azimuthal angle of the in-plane magnetiza-\ntion\u001eMis obtained through the comparison be-\ntween experimental data and simulated magnetic\n\feld maps.\nWe \frst evaluate the uncertainty of the \ftting pro-\ncedure. To this end, the line pro\fle is \ftted with the\nresult of the magnetic calculation while \fxing all the\nparameters pito their nominal values \u0016 pi, leading to\nM= 26:9\u00060:7 kA/m. The relative uncertainty linked\nto the \ftting procedure is therefore given by \u000f\ft= 3%.\nWe note that the intrinsic accuracy of the magnetic \feld\nmeasurement is in the range of \u000eBNV\u00185\u0016T. The re-\nsulting uncertainty can be safely neglected.\nIn order to estimate the relative uncertainty \u000fpiintro-\nduced by each parameter pi, the \ft was performed with\none parameter pi\fxed atpi= \u0016pi\u0006\u001bpi, all the other\nparameters remaining \fxed at their nominal values. The\ncorresponding \ft outcomes are denoted M( \u0016pi+\u001bpi) and\nM( \u0016pi\u0000\u001bpi). The relative uncertainty \u000fpiintroduced by7\nthe errors on parameter piis then \fnally de\fned as\n\u000fpi=M( \u0016pi+\u001bpi)\u0000M( \u0016pi\u0000\u001bpi)\n2M( \u0016pi): (1)\nThis analysis was performed for each parameter pi. The\ncumulative uncertainty \u000fis \fnally given by\n\u000f=s\n\u000f2\n\ft+X\ni\u000f2pi; (2)\nwhere all errors are assumed to be independent.\nA summary of the uncertainties is given in Table I. We\nobtainM= 27\u00064 kA/m for the \rake shown in Fig. 3\nandM= 26\u00064 kA/m for the \rake shown in Fig. 4.TABLE I. Analysis of uncertainties in the measurement of the\nmagnetization\n(a) CrTe 2\rake shown in Figure 3\nparameterpinominal value \u0016 piuncertainty \u001bpi\u000fpi(%)\ndNV 80 nm 10 nm 8\n\u0012NV 58\u000e2\u000e7\n\u001eNV 103\u000e2\u000e2\nt 35 nm 2 nm 6\n\u001eM \u0000145\u000e5\u000e5\n\u000f=p\n\u000f2\n\ft+P\ni\u000f2pi14\n(b) CrTe 2\rake shown in Figure 4\nparameterpinominal value \u0016 piuncertainty \u001bpi\u000fpi(%)\ndNV 80 nm 10 nm 8\n\u0012NV 58\u000e2\u000e7\n\u001eNV 103\u000e2\u000e2\nt 24 nm 2 nm 8\n\u001eM \u0000100\u000e10\u000e12\n\u000f=p\n\u000f2\n\ft+P\ni\u000f2pi17\n[1] J. M. Kosterlitz and D. J. Thouless, J. Phys. C: Solid\nState Phys. 6, 1181 (1973).\n[2] H. Li, S. Ruan, and Y.-J. Zeng, Adv. Mater. 31, 1900065\n(2019).\n[3] M.-C. Wang, C.-C. Huang, C.-H. Cheung, C.-Y. Chen,\nS. G. Tan, T.-W. Huang, Y. Zhao, Y. Zhao, G. Wu, Y.-P.\nFeng, H.-C. Wu, and C.-R. Chang, Annalen der Physik\n532, 1900452 (2020).\n[4] B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein,\nR. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A.\nMcGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo-\nHerrero, and X. Xu, Nature 546, 270 (2017).\n[5] C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao,\nW. Bao, C. Wang, Y. Wang, Z. Q. Qiu, R. J. Cava, S. G.\nLouie, J. Xia, and X. Zhang, Nature 546, 265 (2017).\n[6] K. S. Burch, D. Mandrus, and J.-G. Park, Nature 563,\n47 (2018).\n[7] C. Gong and X. Zang, Science 363(2019).\n[8] M. Gibertini, M. Koperski, A. F. Morpurgo, and K. S.\nNovoselov, Nature Nanotechnology 14, 408 (2019).\n[9] S. Liu, X. Yuan, Y. Zou, Y. Sheng, C. Huang, E. Zhang,\nJ. Ling, Y. Liu, W. Wang, C. Zhang, J. Zou, K. Wang,\nand F. Xiu, npj 2D Materials and Applications 1, 30\n(2017).\n[10] N. Leon-Brito, E. D. Bauer, F. Ronning, J. D. Thomp-\nson, and R. Movshovich, Journal of Applied Physics 120,\n083903 (2016).\n[11] Z. Fei, B. Huang, P. Malinowski, W. Wang, T. Song,\nJ. Sanchez, W. Yao, D. Xiao, X. Zhu, A. F. May, W. Wu,D. H. Cobden, J.-H. Chu, and X. Xu, Nature Materials\n17, 778 (2018).\n[12] Y. Deng, Y. Yu, Y. Song, J. Zhang, N. Z. Wang, Z. Sun,\nY. Yi, Y. Z. Wu, S. Wu, J. Zhu, J. Wang, X. H. Chen,\nand Y. Zhang, Nature 563, 94 (2018).\n[13] D. Weber, A. H. Trout, D. W. McComb, and J. E. Gold-\nberger, Nano Lett. 19, 5031 (2019).\n[14] H. Wang, Y. Liu, P. Wu, W. Hou, Y. Jiang, X. Li,\nC. Pandey, D. Chen, Q. Yang, H. Wang, D. Wei, N. Lei,\nW. Kang, L. Wen, T. Nie, W. Zhao, and K. L. Wang,\nACS Nano 14, 10045 (2020).\n[15] Q. Li, M. Yang, C. Gong, R. V. Chopdekar, A. T.\nN'Diaye, J. Turner, G. Chen, A. Scholl, P. Shafer,\nE. Arenholz, A. K. Schmid, S. Wang, K. Liu, N. Gao,\nA. S. Admasu, S.-W. Cheong, C. Hwang, J. Li, F. Wang,\nX. Zhang, and Z. Qiu, Nano Lett. 18, 5974 (2018).\n[16] A. F. May, D. Ovchinnikov, Q. Zheng, R. Hermann,\nS. Calder, B. Huang, Z. Fei, Y. Liu, X. Xu, and M. A.\nMcGuire, ACS Nano 13, 4436 (2019).\n[17] J. Seo, D. Y. Kim, E. S. An, K. Kim, G.-Y. Kim, S.-\nY. Hwang, D. W. Kim, B. G. Jang, H. Kim, G. Eom,\nS. Y. Seo, R. Stania, M. Muntwiler, J. Lee, K. Watanabe,\nT. Taniguchi, Y. J. Jo, J. Lee, B. I. Min, M. H. Jo, H. W.\nYeom, S.-Y. Choi, J. H. Shim, and J. S. Kim, Science\nAdvances 6(2020).\n[18] P. Mallet, F. Chiapello, H. Okuno, H. Boukari, M. Jamet,\nand J.-Y. Veuillen, Phys. Rev. Lett. 125, 036802 (2020).\n[19] S. J. Yun, D. L. Duong, D. M. Ha, K. Singh, T. L. Phan,\nW. Choi, Y.-M. Kim, and Y. H. Lee, Advanced Science8\n7, 1903076 (2020).\n[20] Z. F., B. Zheng, A. Sebastian, H. Olson, M. Liu, K. Fuji-\nsawa, Y. T. H. Pham, V. Ortiz Jimenez, V. Kalappattil,\nL. Miao, T. Zhang, R. Pendurthi, Y. Lei, A. L. Elias,\nY. Wang, N. Alem, P. E. Hopkins, S. Das, V. H. Crespi,\nM.-H. Phan, and M. Terrones, arXiv:2005.01965 (2020).\n[21] M. Bonilla, S. Kolekar, Y. Ma, H. C. Diaz, V. Kalappat-\ntil, R. Das, T. Eggers, H. R. Gutierrez, M.-H. Phan, and\nM. Batzill, Nature Nanotechnology 13, 289 (2018).\n[22] D. J. O'Hara, T. Zhu, A. H. Trout, A. S. Ahmed, Y. K.\nLuo, C. H. Lee, M. R. Brenner, S. Rajan, J. A. Gupta,\nD. W. McComb, and R. K. Kawakami, Nano Letters 18,\n3125 (2018).\n[23] J. Li, B. Zhao, P. Chen, R. Wu, B. Li, Q. Xia, G. Guo,\nJ. Luo, K. Zang, Z. Zhang, H. Ma, G. Sun, X. Duan, and\nX. Duan, Advanced Materials 30, 1801043 (2018).\n[24] J. Feng, D. Biswas, A. Rajan, M. D. Watson, F. Mazzola,\nO. J. Clark, K. Underwood, I. Markovic, M. McLaren,\nA. Hunter, D. M. Burn, L. B. Du\u000by, S. Barua, G. Bal-\nakrishnan, F. Bertran, P. Le Fevre, T. K. Kim, G. van der\nLaan, T. Hesjedal, P. Wahl, and P. D. C. King, Nano\nLett. 18, 4493 (2019).\n[25] P. K. J. Wong, W. Zhang, F. Bussolotti, X. Yin, T. S.\nHerng, L. Zhang, Y. L. Huang, G. Vinai, S. Krishna-\nmurthi, D. W. Bukhvalov, Y. J. Zheng, R. Chua, A. T.\nN'Diaye, S. A. Morton, C.-Y. Yang, K.-H. Ou Yang,\nP. Torelli, W. Chen, K. E. J. Goh, J. Ding, M.-T. Lin,\nG. Brocks, M. P. de Jong, A. H. Castro Neto, and\nA. T. S. Wee, Advanced Materials 31, 1901185 (2019).\n[26] D. C. Freitas, R. Weht, A. Sulpice, G. Remenyi, P. Stro-\nbel, F. Gay, J. Marcus, and M. N\u0013 u~ nez-Regueiro, Journal\nof Physics: Condensed Matter 27, 176002 (2015).\n[27] X. Sun, W. Li, X. Wang, Q. Sui, T. Zhang, Z. Wang,\nL. Liu, D. Li, S. Feng, S. Zhong, H. Wang, V. Bouchiat,\nM. Nunez Regueiro, N. Rougemaille, J. Coraux, A. Pur-\nbawati, A. Hadj-Azzem, Z. Wang, B. Dong, X. Wu,\nT. Yang, G. Yu, B. Wang, Z. Han, X. Han, and Z. Zhang,\nNano Research (2020).\n[28] A. Purbawati, J. Coraux, J. Vogel, A. Hadj-Azzem,\nN. Wu, N. Bendiab, D. Jegouso, J. Renard, L. Marty,\nV. Bouchiat, A. Sulpice, L. Aballe, M. Foerster,\nF. Genuzio, A. Locatelli, T. O. Mente\u0018 s, Z. V. Han,\nX. Sun, M. N\u0013 u~ nez-Regueiro, and N. Rougemaille, ACS\nAppl. Mater. Interfaces 12, 30702 (2020).\n[29] G. Balasubramanian, I. Y. Chan, R. Kolesov, M. Al-\nHmoud, J. Tisler, C. Shin, C. Kim, A. Wojcik, P. R. Hem-\nmer, A. Krueger, T. Hanke, A. Leitenstorfer, R. Brats-\nchitsch, F. Jelezko, and J. Wrachtrup, Nature 455, 648\n(2008).\n[30] J. R. Maze, P. L. Stanwix, J. S. Hodges, S. Hong, J. M.\nTaylor, P. Cappellaro, L. Jiang, M. V. G. Dutt, E. Togan,\nA. S. Zibrov, A. Yacoby, R. L. Walsworth, and M. D.\nLukin, Nature 455, 644 (2008).\n[31] L. Rondin, J.-P. Tetienne, T. Hingant, J.-F. Roch,\nP. Maletinsky, and V. Jacques, Reports on Progress in\nPhysics 77, 056503 (2014).\n[32] F. Casola, T. van der Sar, and A. Yacoby, Nature Re-\nviews Materials 3, 17088 (2018).\n[33] J.-P. Tetienne, T. Hingant, L. J. Martinez, S. Rohart,\nA. Thiaville, L. H. Diez, K. Garcia, J.-P. Adam, J.-\nV. Kim, J.-F. Roch, I. M. Miron, G. Gaudin, L. Vila,\nB. Ocker, D. Ravelosona, and V. Jacques, Nature Com-\nmunications 6, 6733 (2015).\n[34] Y. Dovzhenko, F. Casola, S. Schlotter, T. X. Zhou,F. Buttner, R. L. Walsworth, G. S. D. Beach, and A. Ya-\ncoby, Nature Communications 9, 2712 (2018).\n[35] J.-Y. Chauleau, T. Chirac, S. Fusil, V. Garcia,\nW. Akhtar, J. Tranchida, P. Thibaudeau, I. Gross,\nC. Blouzon, A. Finco, M. Bibes, B. Dkhil, D. D.\nKhalyavin, P. Manuel, V. Jacques, N. Jaouen, and\nM. Viret, Nature Materials 19, 386 (2020).\n[36] M. J. H. Ku, T. X. Zhou, Q. Li, Y. J. Shin, J. K.\nShi, C. Burch, L. E. Anderson, A. T. Pierce, Y. Xie,\nA. Hamo, U. Vool, H. Zhang, F. Casola, T. Taniguchi,\nK. Watanabe, M. M. Fogler, P. Kim, A. Yacoby, and\nR. L. Walsworth, Nature 583, 537 (2020).\n[37] L. Thiel, Z. Wang, M. A. Tschudin, D. Rohner,\nI. Guti\u0013 errez-Lezama, N. Ubrig, M. Gibertini, E. Gian-\nnini, A. F. Morpurgo, and P. Maletinsky, Science 364,\n973 (2019).\n[38] D. A. Broadway, S. C. Scholten, C. Tan, N. Dontschuk,\nS. E. Lillie, B. C. Johnson, G. Zheng, Z. Wang, A. R.\nOganov, S. Tian, C. Li, H. Lei, L. Wang, L. C. L. Hol-\nlenberg, and J.-P. Tetienne, Adv. Mater. n/a, 2003314\n(2020).\n[39] Q.-C. Sun, T. Song, E. Anderson, T. Shalomayeva,\nJ. F orster, A. Brunner, T. Taniguchi, K. Watan-\nabe, J. Gr afe, R. St ohr, X. Xu, and J. Wrachtrup,\narXiv:2009.13440 (2020).\n[40] T. Hingant, J.-P. Tetienne, L. J. Mart\u0013 \u0010nez, K. Garcia,\nD. Ravelosona, J.-F. Roch, and V. Jacques, Physical\nReview Applied 4, 014003 (2015).\n[41] A. Gruber, A. Dr abenstedt, C. Tietz, L. Fleury,\nJ. Wrachtrup, and C. V. Borczyskowski, Science 276,\n2012 (1997).\n[42] I. Gross, W. Akhtar, A. Hrabec, J. Sampaio, L. J.\nMart\u0013 \u0010nez, S. Chouaieb, B. J. Shields, P. Maletinsky,\nA. Thiaville, S. Rohart, and V. Jacques, Phys. Rev. Ma-\nterials 2, 024406 (2018).\n[43] W. Akhtar, A. Hrabec, S. Chouaieb, A. Haykal, I. Gross,\nM. Belmeguenai, M. Gabor, B. Shields, P. Maletinsky,\nA. Thiaville, S. Rohart, and V. Jacques, Phys. Rev. Ap-\nplied 11, 034066 (2019).\n[44] J.-P. Tetienne, L. Rondin, P. Spinicelli, M. Chipaux,\nT. Debuisschert, J.-F. Roch, and V. Jacques, New Jour-\nnal of Physics 14, 103033 (2012).\n[45] P. Maletinsky, S. Hong, M. S. Grinolds, B. Hausmann,\nM. D. Lukin, R. L. Walsworth, M. Loncar, and A. Ya-\ncoby, Nature Nanotechnology 7, 320 (2012).\n[46] P. Appel, E. Neu, M. Ganzhorn, A. Barfuss, M. Batzer,\nM. Gratz, A. Tsch ope, and P. Maletinsky, Review of\nScienti\fc Instruments 87, 063703 (2016).\n[47] B. C. Buchler, T. Kalkbrenner, C. Hettich, and V. San-\ndoghdar, Phys. Rev. Lett. 95, 063003 (2005).\n[48] J. Tisler, T. Oeckinghaus, R. J. St ohr, R. Kolesov,\nR. Reuter, F. Reinhard, and J. Wrachtrup, Nano Lett.\n13, 3152 (2013).\n[49] L. Rondin, J.-P. Tetienne, S. Rohart, a. Thiaville, T. Hin-\ngant, P. Spinicelli, J.-F. Roch, and V. Jacques, Nat.\nCommun. 4, 2279 (2013).\n[50] D. A. Broadway, S. E. Lillie, S. C. Scholten, D. Rohner,\nN. Dontschuk, P. Maletinsky, J.-P. Tetienne, and L. C. L.\nHollenberg, Phys. Rev. Applied 14, 024076 (2020).\n[51] C. Tan, J. Lee, S.-G. Jung, T. Park, S. Albarakati, J. Par-\ntridge, M. R. Field, D. G. McCulloch, L. Wang, and\nC. Lee, Nature Communications 9, 1554 (2018).\n[52] F. Huang, M. T. Kief, G. J. Mankey, and R. F. Willis,\nPhys. Rev. B 49, 3962 (1994).9\n[53] Y. Li and K. Baberschke, Phys. Rev. Lett. 68, 1208\n(1992).\n[54] A. Bedoya-Pinto, J.-R. Ji, A. Pandeya, P. Gargiani,\nM. Valvidares, P. Sessi, F. Radu, K. Chang, and\nS. Parkin, arXiv:2006.07605 (2020).\n[55] I. Gross, W. Akhtar, V. Garcia, L. J. Mart\u0013 \u0010nez,S. Chouaieb, K. Garcia, C. Carr\u0013 et\u0013 ero, A. Barth\u0013 el\u0013 emy,\nP. Appel, P. Maletinsky, J.-V. Kim, J. Y. Chauleau,\nN. Jaouen, M. Viret, M. Bibes, S. Fusil, and V. Jacques,\nNature 549, 252 (2017)." }, { "title": "2011.07813v1.Topologically_stable_bimerons_and_skyrmions_in_vanadium_dichalcogenide_Janus_monolayers.pdf", "content": "Topologically stable bimerons and skyrmions in vanadium dichalcogenide Janus\nmonolayers\nSlimane Laref1\u0003,\u0003V. M. L. D. P. Goli1\u0003,yIdris Smaili2, Udo Schwingenschl ogl1, and Aur\u0013 elien Manchon1;2;3z\n1Physical Science and Engineering Division (PSE),\nKing Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia\n2Computer, Electrical and Mathematical Science and Engineering (CEMSE),\nKing Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia\n3Aix-Marseille Univ, CNRS, CINaM, Marseille, France\nWe investigate the magnetic phase diagram of 1T-vanadium dichalcogenide monolayers in Janus\ncon\fguration (VSeTe, VSSe, and VSTe) from \frst principles. The magnetic exchange, magne-\ntocrystalline anisotropy and Dzyaloshinskii-Moriya interaction (DMI) are computed using density\nfunctional theory calculations, while the temperature- and \feld-dependent magnetic phase diagram\nis simulated using large-scale atomistic spin modeling in the presence of thermal \ructuations. The\nboundaries between magnetic ordered phases and paramagnetic phases are determined by cross-\nanalyzing the average topological charge with the magnetic susceptibility and its derivatives. We\n\fnd that in such Janus monolayers, DMI is large enough to stabilize non-trivial chiral textures. In\nVSeTe monolayer, an asymmetrical bimeron lattice state is stabilized for in-plane \feld con\fgura-\ntion whereas skyrmion lattice is formed for out-of-plane \feld con\fguration. In VSSe monolayer,\na skyrmion lattice is stabilized for out-of-plane \feld con\fguration. This study demonstrates that\nnon-centrosymmetric van der Waals magnetic monolayers can support topological textures close to\nroom temperature.\nI. INTRODUCTION\nMagnetic skyrmions, i.e., topologically stable magnetic\ntextures, have recently attracted great interest because\nof their unique transport and topological properties1{3,\nopening avenues to novel potential applications in the\n\feld of spintronics4{10. Stable skyrmion crystals and\nmetastable isolated skyrmions are normally obtained\nfrom the competition between magnetic exchange, uniax-\nial anisotropy, antisymmetric Dzyaloshinskii-Moriya in-\nteraction (DMI)11,12, and possibly the external mag-\nnetic \feld. They have been initially reported at low\ntemperature in non-centrosymmetric magnets13,14and\nmore recently at room temperature in transition metal\nmultilayers15{19. The advent of two-dimensional van der\nWaals (2D vdW) intrinsic magnets such as Cr 2Ge2Te620,\nMnSe 221{23, VSe 224, CrI 325, and Fe 3GeTe 226opens ap-\npealing avenues for the observation of chiral magnetic\ntextures in atomically thin materials. What makes 2D\nvdW magnets particularly attractive is the possibility\nto engineer their band structure by surface chemistry.\nAn outstanding demonstration of this feature is the syn-\nthesis of transition metal Janus monolayers27{29, where\nthe transition metal element is embedded between dis-\nsimilar (chalcogen or halide) ions (see also Ref. 30).\nThis con\fguration breaks the inversion symmetry and\npromotes Rashba-type spin-orbit coupling27,31,32. In the\ncase of magnetic Janus monolayers, the inversion sym-\nmetry breaking enables spin-orbit torque33,34as well as\nDMI35{37.\nIn the present work, we investigate the onset of DMI\nand the emergence of magnetic skyrmions in vanadium-\nbased transition metal dichalcogenide Janus monolayers\nin 1T con\fguration. Speci\fcally, we focus our investiga-\ntion in VSSe, VSTe and VSeTe. In Section II, we study\nFIG. 1. (Color Online) Cartoon of the VXY Janus monolayer.\nVanadium elements are represented in grey, and the chalcogen\nelements X and Y (S, Se, and Te (X 6=Y)) are in orange and\nbrown, respectively. The red arrows indicate the direction of\nthe magnetic moment.\nthe magnetic exchange interaction and anisotropy from\n\frst principles and compute DMI using the generalized\nBloch theorem. In Section III, we investigate the mag-\nnetic phase diagram of these systems under the combined\naction of an external magnetic \feld and thermal excita-\ntions using an atomistic spin simulation method. By ex-\nploring in details the temperature-\feld magnetic phase\ndiagram of the three systems, we \fnd that chiral ground\nstates can be obtained in a reasonably large range of\ntemperatures. In particular, VSeTe displays asymmetri-\ncal bimeron lattice ground state up to room temperature,\nwhereas VSSe exhibits skyrmion lattice states.\nII. FIRST PRINCIPLES CALCULATIONS\nA. Structural and magnetic properties\nTo compute the structural and magnetic properties,\nwe perform \frst-principles calculations using the full-arXiv:2011.07813v1 [cond-mat.mtrl-sci] 16 Nov 20202\npotential linearized augmented plane-wave (FLAPW)\nmethod as implemented in FLEUR code38{40. Our calcu-\nlations are performed using the local-density approxima-\ntion exchange-correlation functional (LDA-vwn) as im-\nplemented in the FLAPW Package40. For the angular\nmomentum expansion and the reciprocal plane wave, cut-\no\u000b ofImax= 6 andkmax= 4 were applied, and we used\nradii of MT spheres around 1.7 a.u for S, 2.1 a.u for Se\nand 2.4 a.u for Te, and 2.2 a.u for V, where a.u is the\nBohrradius . \u0000-centered k-grid 32 \u000232\u00021 and 64\u000264\u00021\nhave been sampled for the whole Brillouin zone to achieve\nthe total energy without spin-orbit coupling (SOC) and\nwith SOC, respectively.\nBy de\fning ground-state geometries, crystal structures\nof the bulk VXY (X, Y= S, Se, Te) have been optimized.\nTable I lists the structural and magnetic parameters of\nVSeTe, VSSe, and VSTe materials, which are in good\nagreement with the literature41{44. All three Janus ma-\nterials exhibit C 3vsymmetry and the on-site magnetic\nmoments of V is controlled by the electron depletion due\nto the ionic bonding with the chalcogen elements. As\na phenomenological rule, the larger the electronegativ-\nity di\u000berence between X and Y elements, the larger the\ncharge depletion, and the larger the magnetic moment.\nThe magnetic exchange Jis calculated from the energy\ndi\u000berence between the ferromagnetic and antiferromag-\nnetic state. We \fnd that all three systems are ferro-\nmagnetic and Jroughly scales with the electronegativity\ndi\u000berence \u0001 \u001f, yielding the largest value (70 meV) for\nVSTe. Magnetocrystalline anisotropy is obtained by uti-\nlizing the force theorem and applying SOC within the\nsecond variation method45{47. The magnetocrystalline\nanisotropy, K=Ek\u0000E?, de\fned as the di\u000berence be-\ntween in-plane and out-of-plane magnetization energies is\nreported in Table I. Our results indicate that VSSe pos-\nsesses a weak out-of-plane uniaxial anisotropy, whereas\nboth VSeTe and VSTe exhibit easy-plane anisotropy (in\nother words, there is no variation of the magnetic energy\nwhen rotating the magnetization in the ( x;y) plane for\nthese two monolayers).\nB. Spin spiral calculations\nDMI is an antisymmetric exchange interaction that\ntends to align neighboring magnetic moments perpen-\ndicular to each other. It reads\nEDM=X\nijDij\u0001(Si\u0002Sj); (1)\nwhere Siis the magnetic moment on site i, and Dijis\nthe Dzyaloshinskii-Moriya vector that governs the DMI\nbetween sites iandj. Because DMI is controlled by in-\nversion symmetry breaking, it is oddin SOC strength and\ntherefore, a good estimation of Dijis obtained at the \frst\norder in SOC. To do so, we exploit the generalized Bloch\ntheorem48approach implemented in FLEUR code49{51.\nWe build out-of-plane N\u0013 eel spin spirals in momentum\n/s48/s50/s48/s52/s48\n/s48/s51/s54\n/s45/s48/s46/s54 /s45/s48/s46/s52 /s45/s48/s46/s50 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s48/s50/s48/s52/s48/s54/s48/s56/s48/s40/s97/s41\n/s40/s98/s41\n/s40/s99/s41/s32/s119/s105/s116/s104/s111/s117/s116/s32/s83/s79/s67\n/s32/s119/s105/s116/s104/s32/s83/s79/s67\n/s32/s119/s105/s116/s104/s111/s117/s116/s32/s83/s79/s67\n/s32/s119/s105/s116/s104/s32/s83/s79/s67/s32/s119/s105/s116/s104/s111/s117/s116/s32/s83/s79/s67\n/s32/s119/s105/s116/s104/s32/s83/s79/s67/s77 /s172/s71 /s174 /s77\n/s77 /s172/s71 /s174 /s77\n/s77 /s172/s71 /s174 /s77/s69\n/s113/s45/s69\n/s70/s77/s32/s40/s109/s101/s86/s32/s112/s101/s114/s32/s86/s32/s97/s116/s111/s109/s41\n/s115/s112/s105/s110/s45/s115/s112/s105/s114/s97/s108/s32/s118/s101/s99/s116/s111/s114/s32/s113/s32/s40/s50 /s112 /s47/s97/s41/s99/s111/s117/s110/s116/s101/s114/s32/s99/s108/s111/s99/s107/s119/s105/s115/s101/s32 /s172 /s32/s114/s111/s116/s97/s116/s105/s111/s110/s32/s111/s102/s32/s115/s112/s105/s110/s32/s115/s112/s105/s114/s97/s108/s115/s32 /s174 /s32/s99/s108/s111/s99/s107/s119/s105/s115/s101FIG. 2. (Color Online) Calculated energy dispersions Eqof\nrotating spin spirals without spin-orbit coupling (black sym-\nbols) and with spin-orbit coupling (red symbols) of (a) VSeTe\n(b) VSSe, and (c) VSTe.\nspace whose energy dispersion, without SOC (black) and\nat the \frst order in SOC (red), is reported on Fig. 2.\nThe corresponding di\u000berence between the spin spiral en-\nergy dispersion with and without SOC is reported on\nFig. 3 and the Dzyaloshinskii-Moriya vector is estimated\nby taking the slope of the dispersion at q= 0. This\nvalue corresponds to the DMI strength experienced by\nmagnetic textures whose length scale is much larger than\nthe crystal lattice parameter. The extracted values are\n-5.7 meV\u0001\u0017A, 1.9 meV\u0001\u0017A and 2.5 meV\u0001\u0017A for VSeTe, VSSe\nand VSTe, respectively. For the sake of comparison, the\nDMI obtained for Pt/Co(111) using the same method52is\nabout 50 meV\u0001\u0017A due to the large SOC strength of Pt. As\ndiscussed in the previous section, the DMI in vanadium-\nbased Janus monolayers is su\u000eciently strong to stabilize\nchiral magnetic textures.3\nTABLE I. Structural and magnetic parameters of T-VXY: The lattice constant (a), the distance between X and Y ( dX\u0000Y), the\nelectronegativity di\u000berence between the two chalcogen elements (\u0001 \u001f- de\fned positive), the magnetic moment of the transition\nmetal atom ( \u0016s), the Heisenberg exchange ( J), the magnetocrystalline anisotropy ( K), and DMI strength ( D). The Curie\ntemperature ( Tc) is deduced from zero-\feld susceptibility analysis as explained in Section III.\nVXY a( \u0017A)dX\u0000Y(\u0017A) \u0001\u001f \u0016s(\u0016B)J(meV)K(meV)D(meV\u0001\u0017A)Tc(K)\nVSSe 3.266 2.969 0.03 0.686 20.3 0.078 1.9 240\nVSeTe 3.673 2.887 0.45 1.391 50.82 -0.963 -5.7 630\nVSTe 3.611 2.838 0.48 1.405 71.51 -0.860 2.5 780\n/s45/s50/s48/s50\n/s45/s50/s48/s50\n/s45/s48/s46/s54 /s45/s48/s46/s52 /s45/s48/s46/s50 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s45/s50/s48/s50/s32/s68/s77/s73/s68 /s69\n/s83/s79/s67/s40/s113/s41/s32/s40/s109/s101/s86/s32/s112/s101/s114/s32/s86/s32/s97/s116/s111/s109/s41/s32/s68/s77/s73\n/s115/s112/s105/s110/s45/s115/s112/s105/s114/s97/s108/s32/s118/s101/s99/s116/s111/s114/s32 /s113 /s32/s40/s50 /s112 /s47/s97/s41/s32/s68/s77/s73/s77 /s172/s71 /s174 /s77\n/s77 /s172/s71 /s174 /s77\n/s77 /s172/s71 /s174 /s77/s99/s111/s117/s110/s116/s101/s114/s32/s99/s108/s111/s99/s107/s119/s105/s115/s101/s32 /s172 /s32/s114/s111/s116/s97/s116/s105/s111/s110/s32/s111/s102/s32/s115/s112/s105/s110/s32/s115/s112/s105/s114/s97/s108/s115/s32 /s174 /s32/s99/s108/s111/s99/s107/s119/s105/s115/s101\n/s40/s97/s41\n/s40/s98/s41\n/s40/s99/s41\nFIG. 3. (Color Online) Antisymmetric contribution to the\nenergy dispersion displayed on Fig. 2 as a function of the\nspin-spirals vector qfor (a) VSeTe (b) VSSe, and (c) VSTe.\nIII. MAGNETIC PHASE DIAGRAM FROM\nATOMISTIC SPIN DYNAMICS\nA. Methodology\nTo model the magnetic phases of all three VXY sys-\ntems, we consider atomic spins on each site of a triangularlattice. The Heisenberg spin Hamiltonian is given by\nH=\u0000JX\nhi;jiSi\u0001Sj\u0000X\nhi;jiDij\u0001(Si\u0002Sj)\n\u0000KX\ni(Si\u0001^z)2\u0000\u0016sX\niSi\u0001B; (2)\nwhere Siis normalized unit spin vector at site i. On the\nright side of Eq. (2), the \frst term is the Heisenberg\nexchange energy with ferromagnetic exchange strength\nJandhi;jiindicates the sum over all nearest-neighbors.\nThe second term is the DMI, Dij=D(rij\u0002^z), withDis\nthe DMI energy and rijis the unit vector between sites i\nandj. The third term represents the magnetocrystalline\nanisotropy energy that can be either easy-plane ( K < 0\nfor VSeTe and VSTe) or uniaxial out-of-plane ( K > 0 for\nVSSe). The last term is the Zeeman interaction energy\ndue to the applied \feld, B.\nTo obtain the magnetic phase diagrams, we use atom-\nistic spin dynamics technique that numerically solves the\nstochastic Landau-Lifshitz-Gilbert (LLG) equation of the\nform\n@Si\n@t=\u0000\r\n1 +\u000b2Si\u0002\u0000\nHe\u000b\ni+\u000bSi\u0002He\u000b\ni\u0001\n;(3)\nwhere\ris gyromagnetic ratio and \u000bis the intrinsic\nGilbert damping constant. We choose \u000b= 0.05. The\ne\u000bective \feld acting on the magnetic moment Siis repre-\nsented by He\u000b\ni=\u00001\n\u0016s@H\n@Si+Hth\ni, where\u0016sis the atomic\nmagnetic moment. Hth\niis the stochastic thermal \feld\narising from thermal \ructuations of the magnetic mo-\nments. By using the Langevin dynamics approach, ran-\ndom thermal \feld is added at each site. This \feld obeys\nthe Gaussian distribution in three dimensions with mean\nof zero, \u0000(t). The random thermal \feld is given by53\nHth\ni=\u0000(t)s\n2\u000bKBT\n\r\u0016s\u0001t; (4)\nwherekBis the Boltzmann constant, Tis the tempera-\nture and \u0001tis integration time step. The simulations are\ncarried out on a triangular lattice with N= 150\u0002150\nsites and periodic boundary conditions are implemented.\nThe magnetic parameters are taken from Table I. In or-\nder to characterize the magnetic state at a given point in\nthe (T;B) phase space, Eq. (3) is solved numerically to4\n0 2 4 6 8 10\nB (T)-0.6-0.4-0.200.20.40.6E - ESS (meV/atom)SS\nFM\nABX\n0 5 10 15 20 25 30\nB (T)-0.6-0.4-0.200.20.40.6E - ESS (meV/atom)SS\nFM\nSkX\n0 2 4 6 8 10\nB (T)-0.3-0.2-0.100.10.20.3E - ESS (meV/atom)SS\nFM\nSkX\n0 2 4 6 8 10\nB (T)-1-0.8-0.6-0.4-0.200.2E - ESS (meV/atom)SS\nFM(a)(b)\n(d) (c)SS\nSSSS\nSSFM FM\nFM FMSkX\nSkXABX\nFIG. 4. Zero temperature phase diagrams of VSeTe, VSSe and VSTe systems. The energies corresponding to FM, ABX and\nSkX states are shown with respect to the relative energy of SS state as a function of the applied magnetic \feld. (a) and (b)\nrepresent the phase diagrams of VSeTe with in-plane and out-of-plane applied magnetic \feld, respectively. The phase diagram\nof VSSe with out-of-plane magnetic \feld is shown in (c) while the phase diagram of VSTe is shown in (d) where in-plane\nmagnetic \feld is considered. The shaded areas represent di\u000berent ground states.\nobtain the magnetic susceptibility and average topologi-\ncal charge for VXY systems. The magnetic susceptibility\nis given by\n\u001f=\u0016s\nkBT\u0000\nhM2i\u0000hMi2\u0001\n; (5)\nwhereM=1\nNP\niSi. A simple magnetic ground state\nsuch as ferromagnetic or antiferromagnetic state and a\nskyrmion state can be di\u000berentiated by its topological\ncharge (Q). In the continuum magnetization limit, Qis\nde\fned by the following equation\nQ=1\n4\u0019Z\nS\u0001\u0012@S\n@x\u0002@S\n@y\u0013\ndxdy: (6)\nHereQdescribes the number of times magnetic moments\n(S) wrap around the unit sphere. Q= 0 denotes a triv-\nial state while Q6= 0 provides the number of skyrmions\npresent in the skyrmion state. In a magnetic state withtopologically protected magnetic texture analogous to\nskyrmion,Qis non-zero. For the discrete square lat-\ntice, Berg and L uscher proposed a method to quantify\nQ54. This method involves partitioning the lattice into\nnearest-neighbor triangles with spins at vertices of each\ntriangle. The same procedure can be applied for tri-\nangular lattice except that partitioning the lattice into\ntriangles is not required. Counter-clockwise rotation of\nvertices spins Si,SjandSkof each triangle is considered\nto calculate Qfrom the following equation55,56,\ntan (Q=2) =Si\u0001(Sj\u0002Sk)\n1 +Si\u0001Sj+Sj\u0001Sk+Sk\u0001Si:(7)\nThe spin dynamics simulations are performed with\nincreasing temperature for a \fxed magnetic \feld. At\neach point in ( T;B) phase space, the system is allowed\nto evolve for 106time steps, and the average topologi-\ncal charge and magnetic susceptibility are calculated for5\n3\u0002106averaging time steps. The considered simulation\ntime is su\u000ecient to capture all variations of temperature-\ndependent quantities.\nB. Zero-temperature phase diagram\nWe \frst compute the zero-temperature ground states\nof VSeTe, VSSe and VSTe systems under an external ap-\nplied \feld. Depending on the \feld strength, we \fnd that\nvarious types of magnetic phases can be stabilized: ferro-\nmagnetic (FM), spin spiral (SS), skyrmion crystal (SkX)\nas well as asymmetric bimeron crystal (ABX). The \feld-\ndependence of these various phases at zero temperature\nis reported on Fig. 4. In this \fgure, we report the results\nfor VSeTe (easy-plane anisotropy) with both in-plane (a)\nand out-of-plane magnetic \felds (b), VSSe (out-of-plane\nuniaxial anisotropy) with out-of-plane magnetic \feld (c),\nand VSTe (easy-plane anisotropy) with in-plane magnetic\n\feld (d).\nIn the case of VSeTe, we \fnd SS and FM states in\nthe low \feld and high \feld limits, respectively, indepen-\ndently on the \feld direction. Interestingly, at interme-\ndiate \feld we obtain two topologically non-trivial lat-\ntices, namely asymmetrical bimeron lattice [ABX - Fig.\n4(a)] for in-plane \feld con\fguration and skyrmion lat-\ntice [SkX - Fig. 4(b)] for out-of-plane \feld con\fgura-\ntion. Asymmetrical bimerons are planar counterpart of\nmagnetic skyrmions that are known to emerge under the\ncooperation of DMI with uniaxial in-plane anisotropy57.\nIn VSeTe, which possesses easy-plane rather than uni-\naxial in-plane anisotropy, the in-plane magnetic \feld is\nnecessary to stabilize the bimeron lattice. Notice that\nthe magnetization of the SS state is nearly vanishing and\nhence this state is mostly una\u000bected by the magnetic \feld\nas shown in Fig 4. In contrast, both ABX and FM states\nhave non-zero magnetization and their energy decreases\nwith increasing \feld. In our 150 \u0002150 spins system, the\nABX state with four bimerons becomes the ground state\nfor \felds above 1.2 T. The spin texture of ABX state is\nshown in Fig 5(a). It is also possible to promote lattices\nwith more bimerons with di\u000berent sizes; however, the\nenergy of these bimeron states remain higher than ABX\nstate with four bimerons and are therefore metastable.\nTherefore, the energy of ABX state depends on the num-\nber of bimerons and on their size. With increasing \feld,\nthe size of bimerons shrinks. However, ABX state main-\ntains lower energy until FM state becomes the ground\nstate. The magnetization of FM state is larger than that\nof ABX state, and hence the energy of FM state lowers\nswiftly for large magnetic \felds. Above 6.4 T, VSeTe\nacquires in-plane magnetized state. In the case of out-\nof-plane magnetic \feld con\fguration of VSeTe, SS state\nis the lowest energy state below 1.3 T as shown in Fig.\n4(b). The SkX state becomes the ground state over a\nvery large range of applied \feld, between 1.3 T and 21.4\nT. Noticeably, the skyrmions' shape is hexagonal in this\nrange, due to large skyrmion-skyrmion interaction [seeFig 5(d)]. In the SkX ground state, the skyrmion di-\nameter reduces from 21 nm to 12 nm upon increasing\n\feld. The domain wall width of skyrmion is large due to\neasy-plane anisotropy. Above 21.4T though, all spins of\nVSeTe align to produce the FM ground state.\nFIG. 5. Spin textures of VSeTe with in-plane and out-of-\nplane \felds and VSSe at di\u000berent temperature and \feld com-\nbinations. At in-plane \feld 2 T, ABX state of VSeTe with\ntemperatures (a) T = 0 K, (b) T = 60 K, and (c) T = 300\nK. At out-of-\feld 3 T, SkX state of VSeTe with temperatures\n(d) T = 0 K, (e) T = 60 K, and (f) T = 300 K. SkX state of\nVSSe with temperatures (g) T = 0 K, (h) T = 60 K and (i)\nT = 150 K at B = 3 T.\nIn the case of VSSe, shown in Fig. 4(c), a SkX state\nwith four skyrmions constitutes the ground state between\n2.5 T and 7.4 T. In this state, the skyrmion diameter\nchanges from 18 nm to 11 nm upon increasing \feld. Be-\nlow 2.5 T, the SS state is the ground state while the\nmagnetization is saturated for \felds above 7.4 T. In the\ncase of VSTe, shown in Fig. 4(d), the SS state is the\nground state in the absence of in-plane \feld. Due to the\npresence of easy-plane anisotropy and small DMI, VSTe\nhas in-plane spin texture with out-of-plane tilting. How-\never, the energy gap between SS and FM states is fairly\nsmall. VSTe acquires fully magnetized state with the\napplication of small in-plane \feld.6\nFIG. 6. (Top) ( T;B) phase diagram and (Bottom) corresponding topological charge Qfor VSeTe with (a,d) in-plane and (b,e)\nout-of-plane \felds and (c,f) VSSe systems.\nC. (T;B) magnetic phase diagram\nThe simulations are performed up to a maximum tem-\nperature of 1500 K and a maximum \feld of 25 T. The\naverage topological charge and magnetic susceptibility\nare calculated to de\fne the phase boundaries and criti-\ncal temperatures. The phase diagrams of the three sys-\ntems and their corresponding average topological charge\nare shown in Fig. 6. Along with ABX or SkX, two new\nphases emerge: \ructuation-disorder (FD) and paramag-\nnetic (PM) phases. In the FD region, both bimerons and\nskyrmions acquire a \fnite lifetime and hence the average\ntopological charge remains non-zero58. In contrast, the\nsystem goes into magnetic disordered state in the PM\nregion and exhibits zero topological charge. The details\nof these two phases are discussed below. The bound-\naries between all phases are determined by the in\rection\npoints of the temperature-dependent magnetic suscepti-\nbility and its derivatives, as discussed by Ref. 59.1. VSeTe Phase Diagram\nIn the case of VSeTe with in-plane \feld con\fguration,\ndisplayed on Fig. 6(a), the system is stabilized in cy-\ncloidal SS state at low temperature and zero magnetic\n\feld. In this state, the magnetic domains are connected\nby N\u0013 eel-type domain walls. As the in-plane \feld ap-\nproaches 1.2 T, all the domains convert into ABX state.\nIn Fig. 5(a), the magnetic texture of ABX state is shown\nwith B = 2 T and T = 0 K. As the magnetic \feld in-\ncreases to 6.4 T, the ordered ABX state breaks down into\nFM ground state. Above this \feld, bimeron states can\nbe found. However, from the zero temperature phase di-\nagram of VSeTe in Fig 4(a), these states are considered\nas metastable ABX states and they remain higher en-\nergy states. Hence, for \felds above 6.4 T, all bimerons\ndissolve and the FM state becomes the ground state.\nBecause a bimeron is a topologically protected mag-\nnetic texture analogous to skyrmion, it can be identi\fed\nby a non-zero topological charge ( Q). A bimeron with\nchargeQcan be converted into another bimeron of op-7\nposite sign,\u0000Q, merely by changing the sign of its mag-\nnetic and spatial components. Indeed, a bimeron with\nchargeQidenti\fed by its spatially dependent magnetic\ncon\fguration, [S x(x,y),Sy(x,y),Sz(x,y)], can be converted\ninto another bimeron of charge \u0000Qwith magnetic con\fg-\nuration [S x(-x,y),-Sy(-x,y),-Sz(-x,y)]57. In addition, the\ncoexistence of two bimerons with opposite signs of Qcan\noccur without loss of state energy. Using Eq. (7), the\naverage charge Qis calculated in the ( T;B) phase di-\nagram and is shown in Fig. 6(d). The \fnite value of\naverageQrepresents the ABX region and its boundary\nwith the other topologically-trivial regions. We obtain\nQ= 4 which indicates that four bimerons present in the\nABX state, a number that remains uniform throughout\nthe ABX region at low temperatures, as shown in Fig.\n6(a).\nIn order to investigate temperature dependent phases,\nwe plot the normalized magnetic susceptibility and its\n\frst and second derivatives as a function of temperature\nat speci\fc applied \felds. Figure 7 shows the suscepti-\nbility and its derivatives as a function of temperature at\nB = 2.6 T. It is di\u000ecult to determine the phase bound-\naries using the normalized susceptibility curve. However,\n@\u001f\n@Tdisplays a jump at the phase transition. The in-\n\rection points of@\u001f\n@Tseparate three regions, i.e., ABX,\nFD and PM phases. The temperature corresponding to\nthe maximum and minimum values of@\u001f\n@Tdetermine the\nboundaries between the phases. The FD region comes\nafter the ABX phase upon increasing temperature. The\nmaximum value of@\u001f\n@Tshows the lower limit of the FD\nregion at 330 K at lower \felds. This transition tempera-\nture is \feld-dependent. This region is known to display\nskyrmions with \fnite lifetime58. This can also be appli-\ncable to bimerons as well. In this region, because spins\nare excited by thermal \ructuations, creation and annihi-\nlation of bimerons of opposite signs of Qoccurs, resulting\nin a \ructuation of Qbetween -4 to +4 during the simula-\ntion. On average, the coexistence of bimerons of opposite\ntopological charge leads to Q\u00190, as shown in Fig. 6(d).\nThe minimum value of@\u001f\n@Tin Fig. 7 before the conver-\ngence represents the Curie temperature of VSeTe and it\nis 660 K at 2.6 T \feld. At zero applied \feld, the Curie\ntemperature of VSeTe is 630 K.\nIn the case of out-of-plane applied \feld, the topological\ncharge of VSeTe as a function of temperature and applied\n\feld is shown in Fig. 6(b). At low temperature and in\nSkX region, we obtain Q= -4, i.e., four skyrmions in SkX\nstate. The FD region starts at 330 K and is almost inde-\npendent of the applied \feld. In this region, the average\nQremains constant in contrast to the ABX state. This\nobservation means that the number of skyrmions remains\nconstant and no skyrmions with opposite charge \u0000Qare\ncreated in the FD region. The critical temperature of\nVSeTe is 630 K in the out-of-plane \feld con\fguration.\nTherefore, at low temperatures the critical temperature\nof VSeTe is independent of the applied \feld direction.\n0 200 400 600 800 1000\nT (K)00.20.40.60.81Normalized χχ\n-0.4-0.200.20.4\n∂χ/∂T\n∂2χ/∂T2Derivatives x 10-2\nABXFD PMFIG. 7. Temperature dependent magnetic susceptibility and\nits \frst and second order derivatives of VSeTe system with\nin-plane \feld 2.6 T. The infection points of@\u001f\n@Tseparate three\nphases, namely, ABX, FD and PM phases respectively.\n2. VSSe and VSTe Phase Diagrams\nThe magnetic phase diagram of VSSe for an external\n\feld applied out of the plane is shown in Fig 6(b). The\nshape of the skyrmions remains circular due to out-of-\nplane anisotropy [see Fig 5(g)]. From Fig. 6(e), Q= -4\nwhich indicates that four skyrmions are present in the\nground state of SkX state. From the average Qand@\u001f\n@T\ncalculations, the FD region starts at 150 K. Although\ncreation and annihilation of skyrmions occur in this re-\ngion, the average Qremains constant which suggests that\nonly skyrmions with the same sign of Qare present at\nany time. The in\rection point of@\u001f\n@Thas a minimum at\n240 K, which is the critical temperature at zero \feld.\nLet us \fnally comment on VSTe, whose phase diagram\nis reported on Fig. 6(c). This system also possesses easy-\nplane anisotropy but has a much larger exchange than\nVSeTe. For this reason, VSTe becomes ferromagnetic at\nmuch smaller in-plane \felds. In the case of VSTe, DMI\nis very small compared to the magnetic exchange Jand\nhence only SS appears without external \feld. The critical\nordering temperature of VSTe is 780 K.\nIV. CONCLUSION\nWe have shown that reasonably strong DMIs can be\nobtained in Janus VXY monolayers by using \frst prin-\nciple calculations, in spite of the relatively weak SOC of\nvanadium. This large value is directly related to the elec-\ntric dipole induced by the dissimilar chalcogen elements,\nas already observed for the Rashba spin-splitting in Ref.\n??. In addition, we \fnd that whereas VSSe possesses a8\nweak out-of-plane anisotropy, VSeTe and VSTe are char-\nacterized by strong easy-plane anisotropy. This observa-\ntion suggests that the two latter materials could be an\ninteresting platform for spin super\ruidity60, although we\nleave this aspect to future work.\nWe emphasize that the DMI we obtain is rather weak\n(about one order of magnitude smaller than in transition\nmetal multilayers52), which is partially compensated by\nthe fact that VXY is a monolayer, much thinner than tra-\nditional transition metal thin \flms. Therefore, homochi-\nral spin spirals can be stabilized down to zero external\n\feld, and metastable skyrmions and bimerons can be ob-\ntained in VSeTe and VSSe, respectively. Nonetheless,\nthe fairly large magnetic anisotropy of these monolayers\nprevents the stabilization of chiral textures without ex-\nternal magnetic \feld. We also emphasize that whereas\nour study focuses on stable ground states, the possibil-\nity to stabilize skyrmion and bimeron crystals paves the\nway to the realization of isolated metastable skyrmions\nand bimerons, which are of highest interest for potentialapplications.\nThis study, along with recent work focusing on dif-\nferent materials35,37, suggests that transition metal\ndichalcogenides monolayers in Janus con\fguration can\nhost a wealth of chiral textures in spite of their weak\nSOC. One can easily foresee that chemical surface engi-\nneering can also be favorably used to modulate the inver-\nsion symmetry breaking, which calls for further experi-\nmental exploration.\nACKNOWLEDGMENTS\nThe work was supported by King Abdullah University\nof Science and Technology (KAUST) through the award\nOSR-2018-CRG7-3717 from the O\u000ece of Sponsored Re-\nsearch (OSR). This work used the resources of the Su-\npercomputing Laboratory at King Abdullah University\nof Science and Technology (KAUST) in Thuwal, Saudi\nArabia.\n\u0003slimane.laref@kaust.edu.sa\nydurga.goli@kaust.edu.sa\nzmanchon@cinam.univ-mrs.fr\n1N. Nagaosa and Y. Tokura, Nature Nanotechnology 8, 899\n(2013).\n2K.-S. Ryu, L. Thomas, S.-H. Yang, and S. Parkin, Nature\nNanotechnology 8, 527 (2013).\n3S. Emori, U. Bauer, S.-M. Ahn, E. Martinez, and G. S.\nBeach, Nature Materials 12, 611 (2013).\n4A. Fert, V. Cros, and J. Sampaio, Nature Nanotechnology\n8, 152 (2013).\n5S. Li, W. Kang, Y. Huang, X. Zhang, and Y. Zhou, Nan-\notechnology 18, 31LT01 (2017).\n6Y. Huang, W. Kang, X. Zhang, and Y. Zhou, Nanotech-\nnology 28, 08LT02 (2017).\n7D. Prychynenko, M. Sitte, K. Litzius, B. Kr uger, G. Bouri-\nano\u000b, M. Kl aui, J. Sinova, and K. Everschor-sitte, Physi-\ncal Review Applied 9, 14034 (2018).\n8S. Luo, M. Song, X. Li, Y. Zhang, J. Hong, X. Yang,\nX. Zou, N. Xu, and L. You, Nano Letters 18, 1180 (2018).\n9J. Z\u0013 azvorka, F. Jakobs, D. Heinze, N. Keil, S. Kromin,\nS. Jaiswal, K. Litzius, G. Jakob, P. Virnau, D. Pinna,\nK. Everschor-sitte, L. R\u0013 ozsa, A. Donges, U. Nowak, and\nM. Kl aui, Nature Nanotechnology 14, 658 (2019).\n10K. M. Song, J.-s. Jeong, B. Pan, X. Zhang, J. Xia, S. Cha,\nT.-e. Park, K. Kim, S. Finizio, J. Raabe, J. Chang,\nY. Zhou, W. Zhao, W. Kang, H. Ju, and S. Woo, Na-\nture Electronics 3, 148 (2020).\n11T. Moriya, Physical Review 120, 91 (1960).\n12I. Dzyaloshinskii, Soviet Physics JETP 5, 1259 (1957).\n13S. M uhlbauer, B. Binz, F. Jonietz, C. P\reiderer, A. Rosch,\nA. Neubauer, R. Georgii, and P. B oni, Science (New York,\nN.Y.) 323, 915 (2009).\n14X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han,\nY. Matsui, N. Nagaosa, and Y. Tokura, Nature 465, 901\n(2010).15W. Jiang, P. Upadhyaya, W. Zhang, G. Yu, M. B.\nJung\reisch, F. Y. Fradin, J. E. Pearson, Y. Tserkovnyak,\nK. L. Wang, O. Heinonen, S. G. E. Velthuis, and A. Ho\u000b-\nmann, Science 349, 283 (2015).\n16G. Chen, A. Mascaraque, A. T. N'Diaye, and A. K.\nSchmid, Applied Physics Letters 106, 242404 (2015).\n17O. Boulle, J. Vogel, H. Yang, S. Pizzini, D. d. S. Chaves,\nA. Locatelli, T. O. M. A. Sala, L. D. Buda-Prejbeanu,\nO. Klein, M. Belmeguenai, Y. Roussign\u0013 e, A. Stashkevich,\nS. M. Ch\u0013 erif, L. Aballe, M. Foerster, M. Chshiev, S. Auf-\nfret, I. M. Miron, and G. Gaudin, Nature Nanotechnology\n11, 449 (2016).\n18C. Moreau-Luchaire, C. Mouta\fs, N. Reyren, J. Sampaio,\nC. A. F. Vaz, N. Van Horne, K. Bouzehouane, K. Gar-\ncia, C. Deranlot, P. Warnicke, P. Wohlh uter, J.-M. George,\nM. Weigand, J. Raabe, V. Cros, and A. Fert, Nature Nan-\notechnology 11, 444 (2016).\n19S. Woo, K. Litzius, B. Kr uger, M.-Y. Im, L. Caretta,\nK. Richter, M. Mann, A. Krone, R. M. Reeve, M. Weigand,\nP. Agrawal, I. Lemesh, M.-A. Mawass, P. Fischer,\nM. Kl aui, and G. S. D. Beach, Nature Materials 15, 501\n(2016).\n20C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao,\nW. Bao, C. Wang, Y. Wang, Z. Qiu, R. Cava, S. Louie,\nJ. Xia, and X. Zhang, Nature 546, 265 (2017).\n21D. J. O'Hara, T. Zhu, A. H. Trout, A. S. Ahmed, Y. K.\nLuo, C. H. Lee, M. R. Brenner, S. Rajan, J. A. Gupta,\nD. W. McComb, et al. , Nano Letters 18, 3125 (2018).\n22D. J. O'Hara, T. Zhu, A. H. Trout, A. S. Ahmed, Y. K.\nLuo, C. H. Lee, M. R. Brenner, S. Rajan, J. A. Gupta,\nD. W. McComb, and R. K. Kawakami, Nano Letters 18,\n3125 (2018).\n23D. J. O'Hara, T. Zhu, and R. K. Kawakami, IEEE Mag-\nnetics Letters 9, 1 (2018).\n24M. Bonilla, S. Kolekar, Y. Ma, H. C. Diaz, V. Kalappattil,\nR. Das, T. Eggers, H. R. Gutierrez, M.-H. Phan, and\nM. Batzill, Nature Nanotechnology 13, 289 (2018).9\n25B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein,\nR. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A.\nMcGuire, D. H. Cobden, et al. , Nature 546, 270 (2017).\n26Z. Fei, B. Huang, P. Malinowski, W. Wang, T. Song,\nJ. Sanchez, W. Yao, D. Xiao, X. Zhu, A. F. May, et al. ,\nNature Materials 17, 778 (2018).\n27Y. C. Cheng, Z. Y. Zhu, M. Tahir, and U. Schwingen-\nschl ogl, Europhysics Letters 102, 57001 (2013).\n28A.-Y. Lu, H. Zhu, J. Xiao, C.-P. Chuu, Y. Han, M.-H.\nChiu, C.-C. Cheng, C.-W. Yang, K.-H. Wei, Y. Yang, et al. ,\nNature Nanotechnology 12, 744 (2017).\n29J. Zhang, S. Jia, I. Kholmanov, L. Dong, D. Er, W. Chen,\nand H. Guo, ACS Nano 11, 8192 (2017).\n30L. Zhang, Z. Yang, T. Gong, R. Pan, and H. Wang, Jour-\nnal of Materials Chemistry A 8, 8813 (2020).\n31H. U. Din, M. Idrees, A. Albar, M. Sha\fq, I. Ahmad, C. V.\nNguyen, and B. Amin, Physical Review B 100, 165425\n(2019).\n32J. Chen, K. Wu, H. Ma, W. Hu, and J. Yang, RSC Ad-\nvances 10, 6388 (2020).\n33A. Manchon, J. Zelezn\u0013 y, M. Miron, T. Jungwirth,\nJ. Sinova, A. Thiaville, K. Garello, and P. Gambardella,\nReview of Modern Physics 91, 035004 (2019).\n34I. Smaili, S. Laref, J. H. Garcia, U. Schwingen-\nschlogl, S. Roche, and A. Manchon, arXiv preprint\narXiV:2007.07579v1 , 1 (2020), arXiv:2007.07579.\n35J. Liang, W. Wang, H. Du, A. Hallal, K. Garcia,\nM. Chshiev, A. Fert, and H. Yang, Physical Review B\n101, 184401 (2020).\n36F. Zhang, H. Zhang, W. Mi, and X. Wang, Phys. Chem.\nChem. Phys. 22, 8647 (2020).\n37J. Yuan, Y. Yang, Y. Cai, Y. Wu, Y. Chen, X. Yan, and\nL. Shen, Physical Review B 101, 094420 (2020).\n38D. Hamann, Physical Review Letters 42, 662 (1979).\n39E. Wimmer, H. Krakauer, M. Weinert, and A. J. Freeman,\nPhysical Review B 24, 864 (1981).\n40\\Fleur,\" .\n41Y. Ma, Y. Dai, M. Guo, C. Niu, Y. Zhu, and B. Huang,\nACS Nano 6, 1695 (2012).\n42J. Feng, L. Peng, C. Wu, X. Sun, S. Hu, C. Lin, J. Dai,\nJ. Yang, and Y. Xie, Advanced Materials 24, 1969 (2012).\n43H. Zhang, L.-M. Liu, and W.-M. Lau, Journal of Materials\nChemistry A 1, 10821 (2013).\n44H. Pan, The Journal of Physical Chemistry C 118, 13248\n(2014).\n45A. Oswald, R. Zeller, P. Braspenning, and P. Dederichs,\nJournal of Physics F: Metal Physics 15, 193 (1985).\n46A. I. Liechtenstein, M. Katsnelson, V. Antropov, and\nV. Gubanov, Journal of Magnetism and Magnetic Mate-\nrials67, 65 (1987).\n47C. Li, A. J. Freeman, H. Jansen, and C. Fu, Physical\nReview B 42, 5433 (1990).\n48L. Sandratskii, Journal of Physics: Condensed Matter 3,\n8565 (1991).\n49P. Kurz, F. F orster, L. Nordstr om, G. Bihlmayer, and\nS. Bl ugel, Physical Review B 69, 024415 (2004).\n50M. Heide, G. Bihlmayer, and S. Bl ugel, Physica B: Con-\ndensed Matter 404, 2678 (2009).\n51B. Zimmermann, M. Heide, G. Bihlmayer, and S. Bl ugel,\nPhysical Review B 90, 115427 (2014).\n52A. Belabbes, G. Bihlmayer, F. Bechstedt, S. Bl ugel, and\nA. Manchon, Physical Review Letters 117, 247202 (2016).\n53R. F. L. Evans, W. J. Fan, P. Chureemart, T. A. Ostler,\nM. O. A. Ellis, and R. W. Chantrell, Journal of Physics:Condensed Matter 26, 103202 (2014).\n54B. Berg and M. Luscher, Nuclear Physics B 190, 412\n(1981).\n55M. B ottcher, S. Heinze, S. Egorov, and J. Sinova, New\nJournal of Physics 20, 103014 (2018).\n56G. P. M uller, M. Ho\u000bmann, C. Di\u0019elkamp, D. Sch urho\u000b,\nS. Mavros, M. Sallermann, N. S. Kiselev, H. J\u0013 onsson, and\nS. Bl ugel, Physical Review B 99, 224414 (2019).\n57L. Shen, X. Li, J. Xia, and L. Qiu, arXiv preprint\narXiv:2005.11924 (2020), arXiv:arXiv:2005.11924v2.\n58L. Rozsa, E. Simon, K. Palotas, L. Udvardi, and L. Szun-\nyogh, Physical Review B 93, 024417 (2016).\n59S. V. Grigoriev, E. V. Moskvin, V. A. Dyadkin, D. Lamago,\nT. Wolf, H. Eckerlebe, and S. V. Maleyev, Physical Review\nB83, 224411 (2011).\n60S. Takei and Y. Tserkovnyak, Physical Review Letters 112,\n227201 (2014)." }, { "title": "2011.07826v1.Microwave_spectroscopy_of_the_low_temperature_skyrmion_state_in_Cu2OSeO3.pdf", "content": "Microwave spectroscopy of the low-temperature skyrmion state in\nCu2OSeO 3\nAisha Aqeel,1,\u0003Jan Sahliger,1,\u0003Takuya Taniguchi,1Stefan\nM andl,1Denis Mettus,1Helmuth Berger,2Andreas Bauer,1\nMarkus Garst,3, 4Christian P\reiderer,1and Christian H. Back1, 5\n1Physik-Department, Technische Universit at M unchen, D-85748 Garching, Germany\n2\u0013Ecole Polytechnique Federale de Lausanne, CH-1015 Lausanne, Switzerland\n3Institut f ur Theoretische Festk orperphysik,\nKarlsruhe Institute of Technology, D-76131 Karlsruhe, Germany\n4Institute for quantum materials and technology,\nKarlsruhe Institute of Technology, D-76344 Eggenstein-Leopoldshafen, Germany\n5Munich Center for Quantum Science and\nTechnology (MCQST), D-80799 M unchen, Germany\n(Dated: November 17, 2020)\nAbstract\nIn the cubic chiral magnet Cu 2OSeO 3a low-temperature skyrmion state (LTS) and a concomi-\ntant tilted conical state are observed for magnetic \felds parallel to h100i. In this work, we report\non the dynamic resonances of these novel magnetic states. After promoting the nucleation of the\nLTS by means of \feld cycling, we apply broadband microwave spectroscopy in two experimental\ngeometries that provide either predominantly in-plane or out-of-plane excitation. By compar-\ning the results to linear spin-wave theory, we clearly identify resonant modes associated with the\ntilted conical state, the gyrational and breathing modes associated with the LTS, as well as the\nhybridization of the breathing mode with a dark octupole gyration mode mediated by the magne-\ntocrystalline anisotropies. Most intriguingly, our \fndings suggest that under decreasing \felds the\nhexagonal skyrmion lattice becomes unstable with respect to an oblique deformation, re\rected in\nthe formation of elongated skyrmions.\n1arXiv:2011.07826v1 [cond-mat.mtrl-sci] 16 Nov 2020Magnetic skyrmions are topologically nontrivial spin whirls that are observed in a wide\nrange of bulk materials, such as cubic chiral magnets, lacunar spinels, Heusler compounds,\nor frustrated magnets [1{6]. As an immediate consequence of their nontrivial topology,\nthe creation or annihilation of skyrmions involves rather complex winding and unwinding\nprocesses [7{9]. For instance, when undergoing the transition into the topologically trivial\nhelical state, Bloch points initiate the coalescence of neighboring skyrmions. At low tem-\nperatures, this process may give rise to magnetic textures that share similarities with both\na small-domain helical state incorporating topological disclination defects at the domain\nboundaries [10{12] and a state composed of elongated skyrmions [13].\nRecently, in addition to the well-established high-temperature skyrmion lattice state\n(HTS) [14, 15], a disconnected low-temperature skyrmion state (LTS) was identi\fed in\nCu2OSeO 3for magnetic \felds applied along h100ionly, highlighting the crucial role of\nmagnetocrystalline anisotropies for the stabilization of the LTS [16{18]. The nucleation\nof skyrmions at low temperatures is facilitated by a tilted conical state as an intermediate\nstate [19, 20]. The resulting spin texture typically exhibits a high degree of disorder and a\npronounced history dependence, most notably the volume fraction of LTS may be increased\nby magnetic \feld cycling [16, 21].\nAlthough the suppression of \fnite-temperature e\u000bects initially limits the nucleation of\nthe LTS, it also permits to address the intrinsic ground-state properties of topologically non-\ntrivial objects, complementary to studies of the metastable HTS frozen-in by means of \feld\ncooling [16, 18, 22{30]. In this context, one aspect concerns the microwave dynamics, which\nin the cubic chiral magnets are well-understood for weak spin{orbit coupling [31{33] and\nmight be even exploited for new concepts in spintronic devices [34{37]. The characteristic\nexcitations comprise +Q and {Q modes in the helimagnetic phases as well as one breathing\nand two gyrational modes in the HTS. So far, however, no information was available on the\nmicrowave dynamics in the LTS and the tilted conical phase arising for larger spin{orbit\ncoupling.\nUsing broadband ferromagnetic resonance measurements in combination with linear spin-\nwave theory, we show that the excitation spectra in the LTS are highly reminiscent of those\nin the HTS, despite distinctly di\u000berent degrees of disorder and di\u000berent stabilization mecha-\nnisms, namely thermal \ructuations for the HTS [1, 38] and magnetocrystalline anisotropies\nfor the LTS [16, 17]. All material-speci\fc parameters entering the theory were determined\n2in previous studies [16, 20, 33], allowing for a parameter-free comparison between theory\nand experiment. By reducing the symmetry, the anisotropies mediate the hybridization of\nthe breathing mode with a dark octupole gyration mode in the LTS. In addition, the ex-\ncitations observed at low \felds under decreasing \feld indicate the presence of a breathing\nmode not expected for the helimagnetic ground state, suggesting an oblique instability of\nthe hexagonal skyrmion lattice driven by the elongation of skyrmions.\nFor the present study a single-crystal cuboid of dimensions 1 :37\u00021:5\u00021:82 mm3ori-\nented along [001], [110], and [1 \u001610], was carefully polished and placed on a coplanar waveg-\nuide (CPW) with one of the (001) surfaces facing down. Two sample positions were mea-\nsured; the sample was centered either on the central signal line (width: 1 mm) or on one\nof the gaps (width: 0.3 mm). This way, the exciting ac magnetic \feld was dominated by\neither in-plane or out-of-plane components, allowing to address selectively di\u000berent resonant\nmodes. As the lateral dimensions of sample and CPW were comparable, the excitation\nalways contained a weak contribution of the nondominant component [51]. If not stated\notherwise, data shown in the following were recorded under in-plane excitation. The static\nmagnetic \feld was applied parallel to [001], i.e., perpendicular to the plane of the CPW, and\nwas changed in steps of 1 mT. Low temperatures and static magnetic \felds were provided\nby a \row cryostat with a superconducting magnet. All measurements were carried out at a\ntemperature of 5 K. Using a vector network analyzer, at each \feld value the complex trans-\nmissionS21through the CPW was measured while increasing the frequency ffrom 1 GHz\nto 6.5 GHz in steps of 0.8 MHz. Background contributions were removed by subtracting\na spectrum recorded at high magnetic \feld (500 mT), yielding the di\u000berence \u0001 S21shown\nin the following. Complementary magnetization and ac susceptibility measurements were\ncarried out using a Quantum Design physical properties measurement system.\nThe magnetic phase diagram of Cu 2OSeO 3, shown in Fig. 1(a), represents an important\npoint of reference for the discussion of the experimental data. As typical for a cubic chiral\nmagnet, it comprises paramagnetic and \feld-polarized regimes at high temperatures and\n\felds, respectively, the conical state described in terms of the superposition of a homogeneous\nmagnetization and a helix both oriented along the \feld direction, and the high-temperature\nskyrmion lattice state in \fnite \felds just below the transition temperature Tc= 58 K [14,\n15, 39{43]. In addition, for magnetic \feld parallel to h100i, in Cu 2OSeO 3a separate low-\ntemperature skyrmion state (LTS) is observed in the vicinity of the upper critical \feld\n3\u00160Hc2\u001980 mT [16]. Although the LTS represents the thermodynamic ground state, its\nnucleation is extremely slow due to the complexity of the topological winding involved. In\nfact, a two-step process is observed in which the metastable tilted conical state, characterized\nby conical helices tilting away from the \feld direction, is required as an intermediate state\nprior to the nucleation of the LTS [19, 20].\nThis complex nucleation process is also re\rected in a pronounced dependence of the\nspin texture and the associated magnetic properties on the temperature and \feld history,\nwhere Fig. 1 depicts the situation under decreasing magnetic \feld, also including metastable\nstates. For the present study, it is therefore imperative to focus on distinct measurement\nprotocols described in the following. Starting in a high magnetic \feld well above Hc2, the\n\feld is decreased until reaching the LTS, i.e., 70 mT for the given sample shape. This initial\nmeasurement from high \felds is referred to as Hinitscan. Next, the magnetic \feld is cycled\nntimes between 70 mT and 62 mT, distinctly increasing the volume fraction of LTS as will\nbe discussed below. After this cycling, the magnetic \feld is either decreased, referred to as\nHn\ndecrscan, or increased, referred to as Hn\nincrscan.\nA typical measurement consisting of a Hinitscan, 20 \feld cycles, and a H20\ndecrscan is de-\npicted in Figs. 1(b) through 1(d). The abscissa corresponds to consecutive data points/scans\nthat are separated by \feld steps of 1 mT, the actual magnetic \feld value is shown in Fig. 1(b).\nThe color bar at the bottom indicates the prevailing magnetic state using the color coding\nintroduced in Fig. 1(a), where green shading indicates data that were recorded during the\ncycling of the \feld. Solid vertical lines mark the phase boundaries as inferred conveniently\nfrom the di\u000berential susceptibility, d M=dH, and the real part of the ac susceptibility, \u001f0, both\nshown in Fig. 1(c), where we refer to Ref. [20] for a comprehensive account. Field cycling\nenhances signatures related to the LTS, such as the pronounced maximum at its low-\feld\nboundary, which is interpreted as an increasing volume fraction of LTS in agreement with\nprevious reports [16, 21].\nTypical microwave spectroscopy data are depicted in Fig. 1(d). Here and in the follow-\ning, they are shown in terms of the transmission di\u000berence \u0001 S21recorded as a function of\nfrequency at each \feld value. Dark colors indicate strong absorption due to the excitation\nof resonant modes in the sample. We start the description at high magnetic \felds (left),\nwhere a linear slope is characteristic for a Kittel resonance in the \feld-polarized regime.\nIn addition to the prominent Kittel mode, at lower frequencies several standing spin wave\n4modes are excited due to the low magnetic damping, \u000b\u001910\u00004, of Cu 2OSeO 3at low temper-\natures [44, 45]. These modes are not of central interest for the present study but nevertheless\ncomplicate the interpretation of the results at lower \felds. As a consequence, the following\ndiscussion focuses on the dominant modes in each magnetic phase, for instance refraining\nfrom an analysis of the low-intensity clockwise gyration mode in the skyrmion phase.\nWhen ignoring the \feld cycling for a moment, the microwave spectrum reminds of the\nwell-established universal spectrum of the cubic chiral magnet [32, 33, 46{48], although it is\nmore complex due to the presence of the tilted conical state and the LTS. Under decreasing\nmagnetic \feld (left to right), a change of slope in the Kittel mode marks the onset of the tilted\nconical state. Decreasing the \feld further, a broad band of absorption increases in frequency,\nreminiscent of the behavior in the conical state. The \feld cycling is re\rected in a zigzag\nshape and a continuous shift of spectral weight from high to low frequencies, in particular\nresulting in the emergence of a set of low-frequency modes. These modes essentially decrease\nin frequency under decreasing \feld until vanishing at the low-\feld boundary of the LTS.\nThe nature of the di\u000berent modes is most e\u000eciently discussed when also considering\ntheir evolution under \feld cycling, as elaborated on in Fig. 2, where initial Hinitscans are\ncombined with Hn\ndecrscans measured after di\u000berent numbers, n, of \feld cycles. For clarity,\nthe \feld cycles are omitted from the microwave spectra. In the H0\ndecrscan without cycling,\nshown in Fig. 2(a), two distinct modes may be distinguished that essentially increase in\nfrequency for decreasing \feld. These modes are marked by gray circles and remind of the\n+Q and {Q modes of the conical state. Additional absorption at reduced spectral weight\nfollows a qualitatively similar \feld dependence, suggesting standing spin wave modes as its\norigin. Changes of slope and a small discontinuity in the \u0006Q modes near Hc2are attributed\nto the tilted conical state, in which the \fnite angle between \feld and propagation direction\ninduces small deviations in the microwave response, akin to those in the helical state at\nsmall \felds, cf. Supplementary Information of Ref. [33].\nModerate cycling, such as in the H15\ndecrscan shown in Fig. 2(b), reduces the spectral weight\nof the\u0006Q modes at intermediate \felds. Instead, a set of modes emerges at distinctly lower\nfrequencies of about 2.5 GHz. Increasing the number of cycles to n= 140, as shown in\nFig. 2(c), intensi\fes this shift of spectral weight. The emerging modes exhibit frequencies\nand an evolution under decreasing \feld that are reminiscent of the counterclockwise gyration\nmode in the high-temperature skyrmion lattice. Also taking into account the redistribution\n5of spectral weight under \feld cycling and the disappearance of the modes at the low-\feld\nboundary of the LTS, these \fndings consistently suggest that the low-frequency modes are\nassociated with the LTS. From Lorentzian \fts, narrow line widths are extracted, translating\nto small damping constants \u000b\u00140:01 for these skyrmion modes. Note that the observation of\nresonant modes characteristic of skyrmions also provides direct evidence for the \fxed phase\nrelationship of the multi- Qstructure underlying the nontrivial topology of the LTS [49, 50].\nFurther information on the character of the di\u000berent modes, in particular those in the\nLTS, is obtained from measurements under di\u000berent excitation geometries as shown in Fig. 3.\nIn order to cover the entire \feld range of interest, each panel in Fig. 3 combines data from\nan H140\nincrand an H140\ndecrscan. By placing the sample either on the center of the signal line or\non one of the gaps of the CPW, the microwave excitation in the sample is dominated either\nby in-plane or out-of-plane components. In skyrmion states, in-plane excitation couples\ne\u000eciently to gyration modes, while out-of-plane excitation drives breathing modes [31]. In\nhelical or conical states, only excitation components perpendicular to the propagation vector\ncouple to the\u0006Q modes [32]. Consequently, modes observed under out-of-plane excitation\nare associated with magnetic structures that enclose \fnite angles with the magnetic \feld,\nsuch as multi-domain helical or tilted conical states.\nUnder in-plane excitation, as shown in Fig. 3(a) and all \fgures discussed so far, two groups\nof modes are prevailing below Hc2; (i) the modes in the conical and tilted conical state that\nincrease in frequency with decreasing \feld, and (ii) a lower-frequency mode that decreases\nwith decreasing \feld and is associated with the LTS (marked by cyan circles). When out-\nof-plane excitation dominates, as shown in Fig. 3(b), the overall spectral weight is reduced\nwith weak remnants of the previously discussed modes being excited due to small in-plane\nexcitation components. In addition, a prominent broad mode in the frequency range of the\n\u0006Q modes is associated with the tilted conical state. Perhaps most intriguingly, however, a\ndistinct low-frequency mode is observed across the entire \feld range below Hc2(marked by\norange circles).\nAs substantiated by the theoretical analysis presented below, the sensitivity with respect\nto the excitation geometry identi\fes the modes marked by cyan and orange circles as the\ncounterclockwise gyration mode and the breathing mode in the LTS. As a function of de-\ncreasing \feld, the resonance frequency of the breathing mode exhibits two abrupt increases.\nIn the following we will establish that the jump at \u001840 mT arises from anti-crossing due to\n6a hybridization with a dark octupole gyration mode that is also observed in the quenched\nhigh-temperature skyrmion lattice [52], while the jump at \u001820 mT is consistent with the\nformation of elongated skyrmions.\nThe presence of elongated skyrmions is supported by a comparison of microwave spectra\nrecorded after \feld cycling, i.e., a H70\ndecrscan, with corresponding spectra recorded after\ninitial zero-\feld cooling in Fig. 4. Note that the latter data are measured under increasing\nmagnetic \feld, leading to a modi\fed sequence of phase transitions, where we refer to Ref. [20]\nfor a detailed account on this hysteresis. After zero-\feld cooling, a multi-domain helical state\nforms in Cu 2OSeO 3with three equally populated domains of helices oriented along the h100i\naxes, representing the thermodynamic ground state. A magnetic \feld pointing along one\nof theh100iaxes favors the domain aligned with the \feld, resulting in an increase of its\npopulation upon increasing \feld and, eventually, in a single-domain conical state. This\nprocess is practically irreversible at low temperatures and thus a single helimagnetic domain\nis obtained after removing the \feld [11, 53]. When decreasing the \feld starting from the LTS,\nthe situation is markedly di\u000berent and the system appears to be trapped in a metastable\nstate.\nAs shown in the central part of Fig. 4, the resonance frequencies observed in zero \feld\nafter \feld cycling, cf. Fig. 4(a), di\u000ber decisively from those observed in the well-ordered\nmulti-domain helical state after zero-\feld cooling, cf. Fig. 4(b). This discrepancy suggests\nthat the dynamic properties of the complex zero-\feld magnetic texture obtained after \feld\ncycling are not captured correctly by a description in terms of helical domains. Instead,\nthe calculations presented in the following imply that, at least from the point of view of\nmicrowave excitations, a description in terms of elongated skyrmions is more accurate.\nOur theoretical treatment follows previous work [33] and uses the standard phenomeno-\nlogical model for chiral magnets supplemented by magnetocrystalline anisotropies. This\nanisotropy contribution to the free energy, \"a, proves to be key for the description of the\nproperties of the LTS in Cu 2OSeO 3. The space group P213 allows various magnetocrys-\ntalline terms but we demonstrate that already the simplest form, \"a=K(m4\nx+m4\ny+m4\nz),\ncaptures the fundamental aspects when using the anisotropy constant K=\u00002\u0001103J=m3\nas in Ref. [16], cf. Supplemental Material [] for considerations on other values of K. In\norder to address the resonances of all experimentally observed magnetic textures across the\nentire \feld range, excitation frequencies were determined for (i) non-modulated and one-\n7dimensionally modulated states, i.e, the \feld-polarized, tilted conical, and conical states,\nwithin their respective (meta-)stability range, and (ii) two-dimensionally modulated states,\ni.e., the topologically nontrivial skyrmion states. In these calculations, the \u0006Q modes and\nthe gyrational modes are driven by in-plane excitation, while the breathing mode is driven\nby out-of-plane excitation. The resonance in the tilted conical state is excited in both cases.\nForK= 0 (not shown) the calculations reproduce the characteristic modes as observed\nin the conical and the skyrmion lattice state at high temperatures where Kis small [32, 33].\nAs shown in the experimental spectra in Fig. 5(a) and the calculated spectra in Fig. 5(b), a\n\fnite anisotropy leaves the Kittel mode in the \feld-polarized regime and the \u0006Q modes in\nthe conical state highly reminiscent of their counterparts in the isotropic case. At magnetic\n\felds just below Hc2, however, the tilted conical state (gray shading) becomes energetically\nmore favorable than the regular conical state, as re\rected by a change of slope, kinks, and\nminor discontinuities in the +Q mode. Note that the irregular \feld dependence of the\nresonance frequency originates in fact in a multitude of hybridizations.\nIn both experimental and calculated spectra of the LTS, shown in Figs. 5(c) and 5(d),\nthe counterclockwise gyration mode (cyan symbols) dominates under in-plane excitation.\nFinite anisotropy leaves the character of this mode essentially unchanged, i.e., its frequency\nmonotonically decreases with decreasing \feld akin to the counterclockwise gyration mode in\nthe high-temperature skyrmion lattice. In contrast, the breathing mode (orange symbols)\nprevailing in the LTS under out-of-plane excitation is subject to fundamental changes, with\ngood agreement between experiment and theory. While for the isotropic case, K= 0, the\nfrequency of the breathing mode monotonically increases with decreasing \feld, for K6= 0\na small local minimum just below Hc2is followed by characteristic anti-crossing around\n0:6Hc2. This signature is the hallmark of the hybridization with a dark octupole gyration\nmode mediated by the magnetocrystalline anisotropies. At lower \felds, around 0 :4Hc2, the\nhexagonal skyrmion lattice becomes unstable with respect to an oblique distortion. This\ninstability is driven by the elongation of skyrmions as illustrated by the real-space images in\nFigs. 5(i) and 5(ii). The frequency of the breathing mode is enhanced for the elongated case,\ncf. Supplemental Material [] for animations, and comes close to that of \u0006Q modes of the\nhelimagnetic order but remains distinctly lower, in agreement with the experimental data\nshown in Fig. 4. On a similar note, the spectral weight of the counterclockwise gyration\nmode is distinctly reduced for the elongated skyrmions, consistent with its disappearance in\n8the experimental spectra.\nThe oblique instability of the skyrmion lattice occurs in a \feld range where skyrmions\nare only metastable. Similar to metastable isolated skyrmions [54], they experience an\nelliptical instability and become elongated, leading to an oblique distortion of the hexagonal\nlattice. As topological unwinding is energetically costly, the history employed in Fig. 5(c)\nfavors the observation of this oblique instability, notably the magnetic \feld is decreased at\nlow temperatures after the nucleation of the topologically nontrivial LTS by means of \feld\ncycling. Following this history, the observation of a low-\feld resonance that is, in contrast\nto the helical\u0006Q modes, susceptible to out-of-plane excitation (orange symbols) suggests\nthat a (metastable) oblique skyrmion state hosting elongated skyrmions was indeed realized\nin the experiment.\nIn conclusion, the cubic chiral magnet Cu 2OSeO 3was studied by means of broadband fer-\nromagnetic resonance measurements, focusing in particular on the low-temperature skyrmion\nstate. Employing \feld cycling, two excitation geometries, and comparing the experimen-\ntal results to linear spin-wave theory, resonant modes in the conical, tilted conical, and\nlow-temperature skyrmion state are clearly identi\fed, with the breathing mode in the LTS\nexhibiting a characteristic hybridization. Most intriguingly, the resonances observed in small\n\felds after \feld cycling indicate the presence of elongated skyrmions. These \fndings not\nonly highlight how the robustness of topological non-triviality may in\ruence dynamic prop-\nerties of magnetic materials, but also showcase how the study of dynamic properties may\nprovide valuable insights to static properties, such as the microscopic nature of magnetic\ntextures.\nWe wish to thank G. Benka, A. Chacon, and S. Mayr for fruitful discussions and assistance\nwith the experiments. This work has been funded by the Deutsche Forschungsgemeinschaft\n(DFG, German Research Foundation) under SPP2137 (Skyrmionics, Project No. 360506545,\nProjects 403030645, 403191981, and 403194850), TRR80 (From Electronic Correlations to\nFunctionality, Project No. 107745057, Projects E1, F7, and G9), and the excellence cluster\nMCQST under Germany's Excellence Strategy EXC-2111 (Project No. 390814868). This\nproject has received funding from the European Metrology Programme for Innovation and\nResearch (EMPIR) programme co-\fnanced by the Participating States and from the Eu-\nropean Union's Horizon 2020 research and innovation programme. A.B., D.M., and C.P.\nacknowledge \fnancial support through the European Research Council (ERC) through Ad-\n9vanced Grant No. 788031 (ExQuiSid). T.T. acknowledges funding by the JSPS Overseas\nResearch Fellowship. M.G. is supported by DFG SFB1143 (Correlated Magnetism: From\nFrustration To Topology, Project No. 247310070, Project A07), DFG Grant No. 1072/5-1\n(Project No. 270344603), and DFG Grant No. 1072/6-1 (Project No. 324327023).\n\u0003\n[1] S. M uhlbauer, B. Binz, F. Jonietz, C. P\reiderer, A. Rosch, A. Neubauer, R. Georgii, and\nP. B oni, Science 323, 915 (2009).\n[2] X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han, Y. Matsui, N. Nagaosa, and\nY. Tokura, Nature 465, 901 (2010).\n[3] Y. Tokunaga, X. Z. Yu, J. S. White, H. M. R\u001cnnow, D. Morikawa, Y. Taguchi, and Y. Tokura,\nNat. Commun. 6, 7638 (2015).\n[4] I. K\u0013 ezsm\u0013 arki, S. Bord\u0013 acs, P. Milde, E. Neuber, L. M. Eng, J. S. White, H. M. R\u001cnnow, C. D.\nDewhurst, M. Mochizuki, K. Yanai, H. Nakamura, D. Ehlers, V. Tsurkan, and A. Loidl, Nat.\nMater. 14, 1116 (2015).\n[5] A. K. Nayak, V. Kumar, T. Ma, P. Werner, E. Pippel, R. Sahoo, F. Damay, U. K. R o\u0019ler,\nC. Felser, and S. S. P. Parkin, Nature 548, 561 (2017).\n[6] T. Kurumaji, T. Nakajima, M. Hirschberger, A. Kikkawa, Y. Yamasaki, H. Sagayama,\nH. Nakao, Y. Taguchi, T. Arima, and Y. Tokura, Science 365, 914 (2019).\n[7] P. Milde, D. K ohler, J. Seidel, L. M. Eng, A. Bauer, A. Chacon, J. Kindervater, S. M uhlbauer,\nC. P\reiderer, S. Buhrandt, C. Sch utte, and A. Rosch, Science 340, 1076 (2013).\n[8] J. Wild, T. N. G. Meier, S. P ollath, M. Kronseder, A. Bauer, A. Chacon, M. Halder,\nM. Schowalter, A. Rosenauer, J. Zweck, J. M uller, A. Rosch, C. P\reiderer, and C. H. Back,\nSci. Adv. 3, e1701704 (2017).\n[9] F. Kagawa, H. Oike, W. Koshibae, A. Kikkawa, Y. Okamura, Y. Taguchi, N. Nagaosa, and\nY. Tokura, Nat. Commun. 8, 1332 (2017).\n[10] A. Dussaux, P. Schoenherr, K. Koumpouras, J. Chico, K. Chang, L. Lorenzelli, N. Kanazawa,\nY. Tokura, M. Garst, A. Bergman, C. L. Degen, and D. Meier, Nat. Commun. 7, 12430\n(2016).\n[11] A. Bauer, A. Chacon, M. Wagner, M. Halder, R. Georgii, A. Rosch, C. P\reiderer, and\n10M. Garst, Phys. Rev. B 95, 024429 (2017).\n[12] P. Schoenherr, J. M uller, L. K ohler, A. Rosch, N. Kanazawa, Y. Tokura, M. Garst, and\nD. Meier, Nat. Phys. 14, 465 (2018).\n[13] D. Morikawa, X. Yu, K. Karube, Y. Tokunaga, Y. Taguchi, T. Arima, and Y. Tokura, Nano\nLett. 17, 1637 (2017).\n[14] S. Seki, X. Z. Yu, S. Ishiwata, and Y. Tokura, Science 336, 198 (2012).\n[15] T. Adams, A. Chacon, M. Wagner, A. Bauer, G. Brandl, B. Pedersen, H. Berger, P. Lemmens,\nand C. P\reiderer, Phys. Rev. Lett. 108, 237204 (2012).\n[16] A. Chacon, L. Heinen, M. Halder, A. Bauer, W. Simeth, S. M uhlbauer, H. Berger, M. Garst,\nA. Rosch, and C. P\reiderer, Nat. Phys. 14, 936 (2018).\n[17] A. O. Leonov, ArXiv14062177 Cond-Mat (2014), arXiv:1406.2177 [cond-mat].\n[18] L. J. Bannenberg, H. Wilhelm, R. Cubitt, A. Labh, M. P. Schmidt, E. Leli\u0012 evre-Berna, C. Pap-\npas, M. Mostovoy, and A. O. Leonov, npj Quantum Mater. 4, 11 (2019).\n[19] F. Qian, L. J. Bannenberg, H. Wilhelm, G. Chaboussant, L. M. Debeer-Schmitt, M. P.\nSchmidt, A. Aqeel, T. T. M. Palstra, E. Br uck, A. J. E. Lefering, C. Pappas, M. Mostovoy,\nand A. O. Leonov, Sci. Adv. 4, eaat7323 (2018).\n[20] M. Halder, A. Chacon, A. Bauer, W. Simeth, S. M uhlbauer, H. Berger, L. Heinen, M. Garst,\nA. Rosch, and C. P\reiderer, Phys. Rev. B 98, 144429 (2018).\n[21] D. Mettus, M. Halder, A. Chacon, A. Bauer, and C. P\reiderer, unpublished (2019).\n[22] W. M unzer, A. Neubauer, T. Adams, S. M uhlbauer, C. Franz, F. Jonietz, R. Georgii, P. B oni,\nB. Pedersen, M. Schmidt, A. Rosch, and C. P\reiderer, Phys. Rev. B 81, 041203 (2010).\n[23] R. Ritz, M. Halder, C. Franz, A. Bauer, M. Wagner, R. Bamler, A. Rosch, and C. P\reiderer,\nPhys. Rev. B 87, 134424 (2013).\n[24] A. Bauer, M. Garst, and C. P\reiderer, Phys. Rev. B 93, 235144 (2016).\n[25] H. Oike, A. Kikkawa, N. Kanazawa, Y. Taguchi, M. Kawasaki, Y. Tokura, and F. Kagawa,\nNat. Phys. 12, 62 (2016).\n[26] K. Karube, J. S. White, N. Reynolds, J. L. Gavilano, H. Oike, A. Kikkawa, F. Kagawa,\nY. Tokunaga, H. M. R\u001cnnow, Y. Tokura, and Y. Taguchi, Nat. Mater. 15, 1237 (2016).\n[27] Y. Okamura, F. Kagawa, S. Seki, and Y. Tokura, Nat. Commun. 7, 12669 (2016).\n[28] A. Bauer, A. Chacon, M. Halder, and C. P\reiderer, in Topology in Magnetism , edited by\nJ. Zang, V. Cros, and A. Ho\u000bmann (Springer, 2018) p. 151.\n11[29] G. Berruto, I. Madan, Y. Murooka, G. M. Vanacore, E. Pomarico, J. Rajeswari, R. Lamb,\nP. Huang, A. J. Kruchkov, Y. Togawa, T. LaGrange, D. McGrouther, H. M. R\u001cnnow, and\nF. Carbone, Phys. Rev. Lett. 120, 117201 (2018).\n[30] M. T. Birch, R. Takagi, S. Seki, M. N. Wilson, F. Kagawa, A. \u0014Stefan\u0014 ci\u0014 c, G. Balakrishnan,\nR. Fan, P. Steadman, C. J. Ottley, M. Crisanti, R. Cubitt, T. Lancaster, Y. Tokura, and\nP. D. Hatton, Phys. Rev. B 100, 014425 (2019).\n[31] M. Mochizuki, Phys. Rev. Lett. 108, 017601 (2012).\n[32] Y. Onose, Y. Okamura, S. Seki, S. Ishiwata, and Y. Tokura, Phys. Rev. Lett. 109, 037603\n(2012).\n[33] T. Schwarze, J. Waizner, M. Garst, A. Bauer, I. Stasinopoulos, H. Berger, C. P\reiderer, and\nD. Grundler, Nat. Mater. 14, 478 (2015).\n[34] N. Nagaosa and Y. Tokura, Nat. Nanotechnol. 8, 899 (2013).\n[35] Y. Okamura, F. Kagawa, M. Mochizuki, M. Kubota, S. Seki, S. Ishiwata, M. Kawasaki,\nY. Onose, and Y. Tokura, Nat. Commun. 4, 2391 (2013).\n[36] N. Ogawa, S. Seki, and Y. Tokura, Sci. Rep. 5, 9552 (2015).\n[37] M. Garst, J. Waizner, and D. Grundler, J. Phys. D 50, 293002 (2017).\n[38] S. Buhrandt and L. Fritz, Phys. Rev. B 88, 195137 (2013).\n[39] S. Seki, J.-H. Kim, D. S. Inosov, R. Georgii, B. Keimer, S. Ishiwata, and Y. Tokura, Phys.\nRev. B 85, 220406 (2012).\n[40] S. Seki, S. Ishiwata, and Y. Tokura, Phys. Rev. B 86, 060403 (2012).\n[41] A. A. Omrani, J. S. White, K. Pr\u0014 sa, I. \u0014Zivkovi\u0013 c, H. Berger, A. Magrez, Y.-H. Liu, J. H. Han,\nand H. M. R\u001cnnow, Phys. Rev. B 89, 064406 (2014).\n[42] F. Qian, H. Wilhelm, A. Aqeel, T. T. M. Palstra, A. J. E. Lefering, E. H. Br uck, and\nC. Pappas, Phys. Rev. B 94, 064418 (2016).\n[43] A. Bauer and C. P\reiderer, in Topological Structures in Ferroic Materials: Domain Walls,\nVortices and Skyrmions , edited by J. Seidel (Springer, 2016) p. 1.\n[44] I. Stasinopoulos, S. Weichselbaumer, A. Bauer, J. Waizner, H. Berger, S. Maendl, M. Garst,\nC. P\reiderer, and D. Grundler, Appl. Phys. Lett. 111, 032408 (2017).\n[45] M. Weiler, A. Aqeel, M. Mostovoy, A. Leonov, S. Gepr ags, R. Gross, H. Huebl, T. T. M.\nPalstra, and S. T. B. Goennenwein, Phys. Rev. Lett. 119, 237204 (2017).\n[46] M. Date, K. Okuda, and K. Kadowaki, J. Phys. Soc. Jpn. 42, 1555 (1977).\n12[47] J. D. Koralek, D. Meier, J. P. Hinton, A. Bauer, S. A. Parameswaran, A. Vishwanath,\nR. Ramesh, R. W. Schoenlein, C. P\reiderer, and J. Orenstein, Phys. Rev. Lett. 109, 247204\n(2012).\n[48] S. P ollath, A. Aqeel, A. Bauer, C. Luo, H. Ryll, F. Radu, C. P\reiderer, G. Woltersdorf, and\nC. H. Back, Phys. Rev. Lett. 123, 167201 (2019).\n[49] T. Adams, S. M uhlbauer, C. P\reiderer, F. Jonietz, A. Bauer, A. Neubauer, R. Georgii, P. B oni,\nU. Keiderling, K. Everschor, M. Garst, and A. Rosch, Phys. Rev. Lett. 107, 217206 (2011).\n[50] J. Kindervater, I. Stasinopoulos, A. Bauer, F. X. Haslbeck, F. Rucker, A. Chacon,\nS. M uhlbauer, C. Franz, M. Garst, D. Grundler, and C. P\reiderer, Phys. Rev. X 9, 041059\n(2019).\n[51] I. Stasinopoulos, S. Weichselbaumer, A. Bauer, J. Waizner, H. Berger, M. Garst, C. P\reiderer,\nand D. Grundler, Sci. Rep. 7, 7037 (2017).\n[52] S. Seki, private communication (2019).\n[53] A. Bauer, A. Neubauer, C. Franz, W. M unzer, M. Garst, and C. P\reiderer, Phys. Rev. B 82,\n064404 (2010).\n[54] A. Bogdanov and A. Hubert, Phys. Status Solidi B 186, 527 (1994).\n13Frequency f (GHz) �', dM/dH �0H (mT)(b)\n(c)(a)\n(d)\nNumber of scans\n�S (arb. u.)Magnetic field���H(mT)0\nTemperature T (K)\nHigh-temp. skyrmion\nConical\n45 903060\nHinitHn\ndecrHn\nincrField cycles nskyrmionElongated\nLow-temp. skyrmionTilted \nconical\n0H || <001>\nHinit Field cycling n = 20 H20\ndecrFIG. 1. Evolution of magnetic properties under \feld cycling. (a) Magnetic phase diagram as ob-\nserved under decreasing \feld. Arrows illustrate the \feld histories used in the following. (b,c) Mag-\nnetic \feld,\u00160H, di\u000berential susceptibility, d M=dH, and real part of the ac susceptibility, \u001f0. Data\nwere recorded under decreasing magnetic \feld with n= 20 \feld cycles in the LTS. (d) Microwave\nspectra. The color encodes the transmission di\u000berence \u0001 S21. With \feld cycling spectral weight\nis redistributed. The colored bar at the bottom indicates the dominating magnetic state, vertical\nsolid lines indicate phase transitions. See text for details.\n14Frequency f(GHz)\n-Q+QKittel(a)\n(b)\n(c)\nHinit Hn\ndecr\n�S (arb. u.)\nMagnetic field �0H (mT)n = 0\nn = 15\nn = 140FIG. 2. Evolution of the microwave spectra under \feld cycling. For each panel, the transmission\ndi\u000berence \u0001 S21was recorded during an initial Hinitscan to 70 mT (dotted green line), followed by\nn\feld cycles (not shown), before \fnally decreasing the \feld to zero in a Hn\ndecrscan. (a) Without\ncycling the spectra are dominated by the \u0006Q modes of the tilted conical and conical state (yellow\ncircles). (b,c) With increasing number of cycles spectral weight is continuously redistributed from\nthe\u0006Q modes to a previously unresolved set of low-frequency modes associated with the LTS\n(cyan circles).\n150 40 80 12023456Frequency f (GHz)\n0 40 80 120\n�S (arb. u.)\n0\n-1\n-2\n-3\n-4\n-5H140\ndecr H140\nincrH140\ndecr H140\nincr\nMagnetic field �0H (mT) Magnetic field �0H (mT)G GSxy H\nG GSxy H (a) (b)FIG. 3. Microwave spectra after \feld cycling for di\u000berent excitation geometries as schematically\ndepicted in the insets. Each panel is composed of two independent scans both obtained after 140\n\feld cycles, i.e., an H140\nincrand an H140\ndecrscan. (a) Sample centered on the signal line of the CPW\nyielding predominantly in-plane ( hx) excitation. (b) Sample centered on one of the gaps of the\nCPW yielding predominantly out-of-plane ( hz) excitation. Circles mark the characteristic modes\nin the LTS, namely the counterclockwise gyration mode (cyan) and the breathing mode (orange).\n160 25 50 75 100\nMagnetic field �0H (mT) Magnetic field �0H (mT)23456\n0 25 50 75 100\n-8\n-10-6-4-20�S (arb. u.)Frequency f (GHz)Hinit H70\ndecr Zero-field cooling\n(a) (b)FIG. 4. Comparison of microwave spectra after \feld cycling and after initial zero-\feld cooling.\n(a) Data for decreasing magnetic \feld after \feld cycling composed of an Hinitand an H70\ndecrscan.\n(b) Data for increasing magnetic \feld after initial zero-\feld cooling. Note in particular the dis-\ncrepancy of the resonance frequencies in zero \feld. See text for details.\n17(d)(a)\n(b)\n(c)\nCCWBreathingElongated\nskyrmionLTS\nNormalized field H/Hc2Normalized frequency f/fc2\n(i)0.611.4 \n0.611.4 \n0 0.5 1 1.51.8(ii)+Q\n-QTC\nKittelFM \n0.611.4 \n0.611.4 Conical\nCalculatedExperiment\nCalculatedExperiment1D\n1D\n2D\n2DFIG. 5. Comparison of experimental and calculated spectra. Frequency and magnetic \feld values\nare normalized to their respective values at Hc2. (a) Experimental spectra for the \feld-polarized,\ntilted conical, and conical states. Data inferred from Fig. 2(a). (b) Calculated spectra for the non-\nmodulated and one-dimensionally modulated states. (c) Experimental spectra for the skyrmion\nstates. Data inferred from Fig. 4. (d) Calculated spectra for the two-dimensionally modulated\nstates. A counterclockwise gyration mode (cyan symbols) and a breathing mode (orange symbols)\nare identi\fed. Symbol sizes in (b) and (d) represent spectral weight. (i,ii) Calculated real-space\nimages of the magnetic texture associated with the breathing mode at high and low \felds.\n18" }, { "title": "2012.05076v3.Spin_lattice_model_for_cubic_crystals.pdf", "content": "Spin-lattice model for cubic crystals\nP. Nieves1,\u0003J. Tranchida2, S. Arapan1, and D. Legut1\n1IT4Innovations, V ˇSB - Technical University of Ostrava,\n17. listopadu 2172/15, 70800 Ostrava-Poruba, Czech Republic and\n2Computational Multiscale Department, Sandia National Laboratories,\nP .O. Box 5800, MS 1322, 87185 Albuquerque, NM, United States\n(Dated: March 2, 2021)\nWe present a methodology based on the N ´eel model to build a classical spin-lattice Hamiltonian for cubic\ncrystals capable of describing magnetic properties induced by the spin-orbit coupling like magnetocrystalline\nanisotropy and anisotropic magnetostriction, as well as exchange magnetostriction. Taking advantage of the an-\nalytical solutions of the N ´eel model, we derive theoretical expressions for the parameterization of the exchange\nintegrals and N ´eel dipole and quadrupole terms that link them to the magnetic properties of the material. This\napproach allows to build accurate spin-lattice models with the desire magnetoelastic properties. We also explore\na possible way to model the volume dependence of magnetic moment based on the Landau energy. This new\nfeature can allow to consider the effects of hydrostatic pressure on the saturation magnetization. We apply this\nmethod to develop a spin-lattice model for BCC Fe and FCC Ni, and we show that it accurately reproduces the\nexperimental elastic tensor, magnetocrystalline anisotropy under pressure, anisotropic magnetostrictive coeffi-\ncients, volume magnetostriction and saturation magnetization under pressure at zero-temperature. This work\ncould constitute a step towards large-scale modeling of magnetoelastic phenomena.\nI. INTRODUCTION\nMagnetoelastic interactions couple the motion of atoms in a\nmagnetic material with atomic magnetic moments, and allow\nto transfer mechanical and thermal energies between phonon\nand magnon subsystems1. Magnetoelasticity is of great in-\nterest for applications, but also from a fundamental point of\nview. For instance, precise control of magnetization through a\nmechanical excitation of the motion of atoms in magnetic ma-\nterials, and vice versa, has enabled the development of a wide\nrange of technological applications such as sensors (torque\nsensors, motion and position sensors, force and stress sen-\nsors) and actuators (sonar transducer, linear motors, rotational\nmotors, and hybrid magnetostrictive/piezoelectric devices)2–5.\nSimilarly, the combination of magnetism and heat is exploited\nin many applications like heat-assisted magnetic recording\n(HAMR)6, thermally assisted magnetic random access mem-\nory (MRAMs)7, ultrafast all-optically induced magnetization\ndynamics8,9, magnetic refrigeration10, and biomedical mag-\nnetic hyperthermia11.\nMagnetoelastic effects can also have a strong influence on\nthe thermo-mechanical properties of materials. This is for ex-\nample the case of the phononic component of the thermal con-\nductivity. Though magnon-phonon scattering, it can abruptly\nchange through magnetic phase-transitions12,13. For metallic\noxides presenting strong magnetoelastic effects, and for which\naccurate thermal conductivity predictions can be of practical\nimportance (such as uranium dioxide14), the development of\naccurate numerical models is still an ongoing process. Sim-\nilarly, magnetoelastic effects can also play an important role\nin the thermal expansion of magnetic materials like in Invars,\nwhere is large enough to cancel the normal thermal contrac-\ntion, leading to nearly zero net thermal contraction over a\nbroad range of temperatures15.\nPresently, the theoretical and modeling techniques have\nreached a great level of development and accuracy to describe\nthe uncoupled dynamics of magnons and phonons at differentspatial and time scales. Typically, in magnetic materials this is\ndone by constraining or neglecting either the motion of atomic\nmagnetic moments or atoms. For example, in spin-polarized\nab-inito molecular dynamics (AIMD) magnetic moments are\nconstrained in certain directions and only atomic positions are\nupdated in each time step, while in classical atomistic spin\ndynamics (SD) and molecular dynamics (MD) the motion of\natoms or spins are neglected, respectively16–18. However, it is\nstill a challenge to find suitable modeling approaches to deal\nwith processes where the interaction between magnons and\nphonons is essential, like in magneto-caloric and magneto-\nelastic phenomena. The lack of such models is limiting the\nmulti-scale design of materials suitable for relevant techno-\nlogical applications based on these physical processes. Re-\ncently, novel attempts to address this problem have been pro-\nposed. Stockem et al. demonstrated that for small supercells,\na consistent interface can be designed to couple spin-polarized\nAIMD and classical SD19. Although offering an excellent\nlevel of accuracy, this approach presents the space and time\nscale limitations of first-principles approaches, and does not\nappear to be suited for running meso-scale magneto-elastic\nsimulations. Another concept, referred to as “spin-lattice dy-\nnamics”, is based on the combination of classical spin and\nmolecular dynamics (SD-MD), which includes the spatial de-\npendence of exchange integrals in the spin equation of motion,\namong other features20–26. The computational cost of this\nclassical approach scales linearly with the number of magnetic\natoms in the system26. Combined to accurate massively par-\nallel algorithms, this enables the simulation of multi-million\nmagnetic atom systems on time scales sufficient to accurately\nstudy magnon-phonon relaxation processes22,26.\nThese new ideas have opened up interesting opportuni-\nties and questions about how to model and study magneto-\ncaloric and magneto-elastic phenomena within a multi-scale\napproach. In particular, the coarse-grained modeling of spin-\norbit coupling (SOC) through magnetocrystalline anisotropy\n(MCA) in SD-MD is currently a bottle neck of this issue27.arXiv:2012.05076v3 [cond-mat.mtrl-sci] 1 Mar 20212\nThe single-ion model of MCA is widely used in SD, but\nunfortunately it does not couple atom and spin degrees of\nfreedom. This drawback can be overcome using the N ´eel\nmodel (two-ion model) that reproduces the correct symmetry\nof MCA, and couples atom and spin motion. Hence, despite\nsome limitations of the N ´eel model concerning non-magnetic\natoms and its phenomenological nature28, it seems a promis-\ning starting point to build a SD-MD model capable of simulat-\ning magneto-caloric and magneto-elastic phenomena. In this\nwork, we propose a general procedure to find the parameters\nof the N ´eel model within the Bethe-Slater curve29–31that re-\nproduces the MCA, isotropic and anisotropic magnetoelastic\nproperties, and magnetization under pressure for cubic crys-\ntals at zero-temperature accurately.\nII. METHODOLOGY\nA. Spin-Lattice Hamiltonian\nIn the following discussion, we consider the spin-lattice\nHamiltonian\nHsl(rrr;ppp;sss) =Hmag(rrr;sss)+N\nå\ni=1pppi\n2mi+N\nå\ni;j=1V(ri j); (1)\nwhere rrri,pppi,sssi, and mistand for the position, momentum,\nnormalized magnetic moment and mass for each atom iin the\nsystem, respectively, V(ri j) =V(jrrri\u0000rrrjj)is the interatomic\npotential energy and Nis the total number of atoms in the\nsystem with total volume V. Here, we include the following\ninteractions in the magnetic energy\nHmag(rrr;sss) =\u0000µ0N\nå\ni=1µi(v)HHH\u0001sssi\u00001\n2N\nå\ni;j=1;i6=jJ(ri j)sssi\u0001sssj\n+HL(v)+HN´eel(rrr;sss);(2)\nwhere µi(v)is the atomic magnetic moment that depends on\nthe volume per atom of the system v=V=N,µ0is the vac-\nuum permeability, HHHis the external magnetic field, J(ri j)\nis the exchange parameter. The quantity HLis the Landau\nenergy22,32–34\nHL(v) =N\nå\ni=1(Aiµ2\ni(v)+Biµ4\ni(v)+Ciµ6\ni(v)); (3)\nwhere Ai,BiandCiare parameters, while HN´eelis the N ´eel\ninteraction\nHN´eel=\u00001\n2N\nå\ni;j=1;i6=jfg(ri j)+l1(ri j)h\n(eeei j\u0001sssi)(eeei j\u0001sssj)\u0000sssi\u0001sssj\n3i\n+q1(ri j)h\n(eeei j\u0001sssi)2\u0000sssi\u0001sssj\n3ih\n(eeei j\u0001sssj)2\u0000sssi\u0001sssj\n3i\n+q2(ri j)\u0002\n(eeei j\u0001sssi)(eeei j\u0001sssj)3+(eeei j\u0001sssj)(eeei j\u0001sssi)3\u0003\ng;\n(4)where eeei j=rrri j=ri j, and\nl1(ri j) =l(ri j)+12\n35q(ri j);\nq1(ri j) =9\n5q(ri j);\nq2(ri j) =\u00002\n5q(ri j):(5)\nIn the case of a collinear state ( sssiksssj), the Eq. 4 is reduced to\nHN´eel=\u00001\n2N\nå\ni;j=1;i6=jfg(ri j)+l(ri j)\u0012\ncos2yi j\u00001\n3\u0013\n+q(ri j)\u0012\ncos4yi j\u00006\n7cos2yi j+3\n35\u0013\ng(6)\nwhere cos yi j=eeei j\u0001sssi. The N ´eel energy reproduces the cor-\nrect symmetry of MCA and magnetoelastic energy28. It is\nconvenient to use g(ri j)to offset the exchange and Landau\nenergy in order to allow the forces and pressure become zero\nat the ground state, as detailed in Ma et al.20. To do so, we\nwrite this function as\ng(ri j) =\u0000J(ri j)+2\nN\u00001(Aiµ2\ni(v)+Biµ4\ni(v)+Ciµ6\ni(v)):(7)\nThis offset does not affect the precession dynamics of the\nspins. However, it allows to offset the corresponding mechan-\nical forces. This particular choice of the offset also implies\nthat the spatial dependence of the exchange and Landau en-\nergy is not taken into account in the evaluation of the mag-\nnetic energy at the ground state. The exchange and Landau\nenergies determine the value of magnetic moments32. In this\nmodel, we effectively take into account this fact by parame-\nterizing the volume dependence of magnetic moment using\nfirst-principles calculations. The spatial dependence of the\nexchange and Landau energy would also contribute to the to-\ntal energy when the lattice parameter is modified, influencing\nthe energy versus volume curve from which the equation of\nstate and elastic properties are derived. However, the lack of\nthis contribution in the model should not compromise its ac-\ncuracy as long as the interatomic potential V(ri j)correctly\nreproduces the equation of state and elastic properties. Typ-\nically, interatomic potentials are developed and designed for\nthis purpose. As a result of this offset, we see that the sec-\nond term in Eq.7 cancels with the Landau energy, so that we\ncan simplify the magnetic Hamiltonian Eq.(2) by removing\nthe Landau energy\nHmag(rrr;sss) =\u0000µ0N\nå\ni=1µi(v)HHH\u0001sssi\u00001\n2N\nå\ni;j=1;i6=jJ(ri j)sssi\u0001sssj\n+HN´eel(rrr;sss);(8)\nand setting\ng(ri j) =\u0000J(ri j): (9)\nConsequently, this approach has the advantage that it avoids\nthe calculation of the parameters Ai,BiandCiin the Lan-\ndau energy (Eq.3). As shown in Section III B, the param-\neterization of the volume dependence of magnetic moment3\nmight be simpler than the calculation of the parameters in\nthe Landau energy32. According to the N ´eel model, function\ng(ri j)can be linked to the volume magnetostriction induced\nby the exchange interactions (isotropic magnetostriction)35.\nThe N ´eel dipole ( l(ri j)) and quadrupole ( q(ri j)) terms can\ndescribe the effects induced by SOC and crystal field in-\nteractions like MCA and its strain dependence (anisotropic\nmagnetostriction)35. Here, we take into account the spatial\ndependence of J(ri j),l(ri j)andq(ri j)using the Bethe-Slater\ncurve, as implemented in the SPIN package of LAMMPS26\nJ(ri j) =4aJ\u0012ri j\ndJ\u00132\"\n1\u0000gn\u0012ri j\ndJ\u00132#\ne\u0000\u0010ri j\ndJ\u00112\nQ(Rc;J\u0000ri j);\nl(ri j) =4al\u0012ri j\ndl\u00132\"\n1\u0000gl\u0012ri j\ndl\u00132#\ne\u0000\u0010ri j\ndl\u00112\nQ(Rc;l\u0000ri j);\nq(ri j) =4aq\u0012ri j\ndq\u00132\"\n1\u0000gq\u0012ri j\ndq\u00132#\ne\u0000\u0010ri j\ndq\u00112\nQ(Rc;q\u0000ri j);\n(10)\nwhere Q(Rc;n\u0000ri j)is the Heaviside step function and Rc;n\n(n=J;l;q) is the cut-off radius. The parameters an,gn,\nanddn(n=J;l;q) must be determined in order to reproduce\nthe Curie temperature ( TC) and volume magnetostriction ( ws)\nviaJ(ri j), as well as anisotropic magnetostriction and MCA\nthrough l(ri j)andq(ri j). The parameterization of J(ri j)with\nthe Bethe-Slater curve is a well established procedure. For in-\nstance, to find the values of aJ,gJ, and dJ, one can fit the\nBethe-Slater curve to exchange parameters calculated with\nDensity Functional Theory (DFT) at fixed equilibrium posi-\ntions at zero-temperature26. However, in some cases this pro-\ncedure might lead to spin-lattice models that don’t reproduce\ncorrectly either TCorws. Hence, a strategy to parameterize\nJ(ri j)using the Bethe-Slater function in order to simulate cor-\nrectly these properties is highly desirable. Similarly, the pa-\nrameterization of l(ri j)andq(ri j)with the Bethe-Slater curve\nis a quite new approach, so that it is not clear how to obtain\nthe values of these parameters yet. In Section II B, we pro-\npose a general procedure to obtain these parameters for cubic\ncrystals based on the theoretical analysis of the N ´eel model35.\nIn Section III B we explore a possible parameterization of the\nvolume dependence of magnetic moment using the Landau\nenergy. In the present work, we study this model only at zero-\ntemperature. The equations of motion of this model at finite-\ntemperature are those implemented in the SPIN package of\nLAMMPS18,26. A detailed description of these equations can\nbe found in Ref. 26.\nB. Procedure to calculate the Bethe-Slater parameters of N ´eel\ninteraction for cubic crystals\nThe basic idea to calculate the Bethe-Slater parameters for\nthe N ´eel interaction is to find the theoretical relations that link\nEq. (6) to both the MCA and magnetoelastic energies. To il-\nlustrate this method, we will apply it to simple cubic (SC),\nbody-centered cubic (BCC) and face-centered cubic (FCC)crystals. The MCA energy for cubic systems reads36\nHcub\nMCA(aaa;r) =V K1(r)(a2\nxa2\ny+a2\nxa2\nz+a2\nya2\nz); (11)\nwhere K1is the first MCA constant with units of energy per\nvolume, ris the distance to the first nearest neighbor, Vis\nthe volume of the system, and ai(i=x;y;z) are the direction\ncosines of magnetization. From this equation we have\nV K1(r) =4\u0014\nHcub\nMCA\u00121p\n2;1p\n2;0;r\u0013\n\u0000Hcub\nMCA(1;0;0;r)\u0015\n:\n(12)\nNext, we evaluate the Eq. 6 with magnetic moment directions\nsss=\u0010\n1p\n2;1p\n2;0\u0011\nandsss= (1;0;0)up to first nearest neighbors,\nand we replace it in Eq. 12 in order to ensure that the N ´eel\nenergy gives the correct MCA energy. And by doing so, we\nfind the following relations for SC, BCC and FCC37\nSC:q(r0) =V0K1(r0)\n2N=1\n2r3\n0K1(r0);\nBCC :q(r0) =\u00009V0K1(r0)\n16N=\u0000p\n3\n4r3\n0K1(r0);\nFCC :q(r0) =\u0000V0K1(r0)\nN=\u00001p\n2r3\n0K1(r0);(13)\nwhere r0is the equilibrium distance to the first nearest neigh-\nbors, and Nis the number of atoms in the equilibrium volume\nV0. Here, q(r0)has units of energy per atom. In Appendix A\nwe show that the derivative of q(r)with respect to rcan be\nwritten as\nSC:r0¶q\n¶r\f\f\f\nr=r0=3\n2r3\n0K1(r0)\u0014\n1\u0000B\nK1¶K1\n¶P\u0015\nr=r0;\nBCC :r0¶q\n¶r\f\f\f\nr=r0=\u00003p\n3\n4r3\n0K1(r0)\u0014\n1\u0000B\nK1¶K1\n¶P\u0015\nr=r0;\nFCC :r0¶q\n¶r\f\f\f\nr=r0=\u00003p\n2r3\n0K1(r0)\u0014\n1\u0000B\nK1¶K1\n¶P\u0015\nr=r0;(14)\nwhere Bis the bulk modulus and Pis pressure. Here again,\nr0¶q=¶rhas units of energy per atom. Note that the dipole\nterm in Eq. 6 cancels out after summing all first nearest neigh-\nbors, so that it does not contribute to the MCA in the cubic\ncrystal symmetry. Since we are only considering N ´eel inter-\nactions up to the first nearest neighbors, we set the cut-off\nradius Rc;qin between the first and second nearest neighbors\nin Eq. 10, that is\nq(r0) =4aq\u0012r0\ndq\u00132\"\n1\u0000gq\u0012r0\ndq\u00132#\ne\u0000\u0010r0\ndq\u00112\n: (15)\nThe derivative of this function with respect to ris\n¶q\n¶r\f\f\f\nr=r0=8aqr0e\u0000\u0010r0\ndq\u00112\nd6q\u0002\ngqr4\n0\u0000(1+2gq)d2\nqr2\n0+d4\nq\u0003\n:(16)\nHence, we have two equations with three unknown variables\naq,gq, and dq. A reasonable strategy to reduce the number4\nof unknown variables is to set dqequal to the equilibrium dis-\ntance to the first nearest neighbors r0(dq=r0) because it has\nunit of distance and can be easily estimated. Hence, solving\nEqs. 15 and 16 gives\ndq=r0;\naq=e\n8\u0014\n2q(r0)\u0000r0¶q\n¶r\f\f\f\nr=r0\u0015\n;\ngq=r0¶q\n¶r\f\f\f\nr=r0\nr0¶q\n¶r\f\f\f\nr=r0\u00002q(r0):(17)\nThese are the Bethe-Slater parameters in terms of K1and\n¶K1=¶P(via Eqs. 13 and 14) to model the physics of MCA\nwithin the N ´eel model.\nLet’s now find the values of the Bethe-Slater parameters\nthat simulate magnetostriction. The magnetoelastic energy for\ncubic systems (point groups 432, ¯43m,m¯3m) reads38,39\nHcub\nme\nV0=b0(exx+eyy+ezz)+b1(a2\nxexx+a2\nyeyy+a2\nzezz)\n+2b2(axayexy+axazexz+ayazeyz);\n(18)\nwhere b0,b1andb2are the magnetoelastic constants with\nunits of energy per volume, and ei jare the elements of the\nstrain tensor. For small deformations (infinitesimal strain the-\nory), the strain tensor can be expressed in terms of the dis-\nplacement vector uuuas40,41\nei j=1\n2\u0012¶ui\n¶rj+¶uj\n¶ri\u0013\n; i;j=x;y;z (19)\nwhere ¶ui=¶rjis called the displacement gradient. For this\ndefinition of the strain tensor, the elastic energy for cubic crys-\ntal reads40,41\nHcub\nel\nV0=c11\n2(e2\nxx+e2\nyy+e2\nzz)+c12(exxeyy+exxezz+eyyezz)\n+2c44(e2\nxy+e2\nyz+e2\nxz);\n(20)\nwhere c11,c12andc44are the elastic constants. After evalu-\nating the N ´eel energy (Eq.6) for a strained cubic crystal up to\nfirst nearest neighbors, and equalizing it to Eq.18, one finds\nfor SC, BCC and FCC35,36\nSC:l(r0) =\u0000V0b2\n2N;r0¶l\n¶r\f\f\f\nr=r0=\u0000V0b1\nN;\nBCC :l(r0) =\u00003V0b1\n8N;r0¶l\n¶r\f\f\f\nr=r0=3V0\n8N(b1\u00003b2);\nFCC :l(r0) =V0\n2N\u0012b2\n2\u0000b1\u0013\n;r0¶l\n¶r\f\f\f\nr=r0=V0\nN\u0012\nb1\u00003b2\n2\u0013\n:\n(21)\nHere, we neglected the quadrupole contribution to the mag-\nnetoelastic energy37. This approximation is reasonable when\nq(r0)\u001cl(r0). In Section III, we show that BCC Fe and FCC\nNi fulfill this condition. Next, as we did previously, insert-\ning Eq.21 into the Bethe-Slater curve and its derivative, andsetting dl=r0allow us to obtain\ndl=r0;\nal=e\n8\u0014\n2l(r0)\u0000r0¶l\n¶r\f\f\f\nr=r0\u0015\n;\ngl=r0¶l\n¶r\f\f\f\nr=r0\nr0¶l\n¶r\f\f\f\nr=r0\u00002l(r0):(22)\nThese are the Bethe-Slater parameters in terms of b1andb2\n(via Eq. 21) to model the anisotropic magnetostriction within\nthe N ´eel model.\nLastly, we show the parameterization of the exchange inter-\naction via the first term in the N ´eel model (Eq.9) to simulate\nTCandws. From the analysis of the N ´eel model up to first\nnearest-neighbours one finds35\nSC:J(r0) =kBTC\n2;r0¶J\n¶r\f\f\f\nr=r0=ws(c11+2c12)V0\n3N;\nBCC :J(r0) =3kBTC\n8;r0¶J\n¶r\f\f\f\nr=r0=ws(c11+2c12)V0\n4N;\nFCC :J(r0) =kBTC\n4;r0¶J\n¶r\f\f\f\nr=r0=ws(c11+2c12)V0\n6N;\n(23)\nwhere kBis the Boltzmann constant. The relation between\nJ(r0)andTCwas obtained using the Mean-Field Approxima-\ntion (MFA). Inserting Eq.23 into the Bethe-Slater curve and\nits derivative, and setting dJ=r0allow us to obtain\ndJ=r0;\naJ=e\n8\u0014\n2J(r0)\u0000r0¶J\n¶r\f\f\f\nr=r0\u0015\n;\ngJ=r0¶J\n¶r\f\f\f\nr=r0\nr0¶J\n¶r\f\f\f\nr=r0\u00002J(r0);(24)\nwhere the cut-off radius Rc;Jmust be in between the first and\nsecond nearest neighbors. Notice that a similar procedure to\nthis one could be used if a function with three parameters dif-\nferent to the Bethe-Slater curve is chosen to describe the func-\ntions l(r),q(r)andJ(r)in SD-MD simulations.\nC. Volume dependence of magnetic moment\nA widely used approximation in SD consists to constrain\nthe magnitude of atomic magnetic moments to a constant\nvalue16,17, which is a good approximation for many practical\napplications. However, the magnitude of the atomic magnetic\nmoments can change significantly when the volume of the sys-\ntem changes greatly. This effect can be analyzed in terms\nof the Landau expansion around the critical point where the\nmagnetization becomes zero42. The Landau expansion con-\ntains only even powers of magnetization to fulfill the time-\nreversal symmetry. For instance, the Curie temperature ( Tc) is\na critical point where the magnetization becomes zero due to5\nthe disorder of magnetic moments induced by thermal fluc-\ntuations. Another critical point is the volume per atom vc\nwhere magnetic moment collapses ( µ(vc) =0)33. Moruzzi\nidentified three types of transitions for the magnetic moment\ncollapse33,34. Here, we discuss about type I transition where\nthe behavior is continuous across vc. In particular, we ex-\nplore a possible parameterization of the magnetic moments\nbased on the Landau energy to take into account its volume\ndependence. For a system with Natoms with equal magnetic\nmoments µwe can write the Landau expansion close to vcas\nHL(v) =N\nå\ni=1(Aiµ2\ni(v)+Biµ4\ni(v))+O(µ6)\n=N(Aµ2(v)+Bµ4(v));(25)\nwhere AandBare parameters, and vis the volume per atom of\nthe system. The analysis of the Landau expansion yields33,34\nµ(v)µpv\u0000vc: (26)\nThis square root dependence describes well the magnetic mo-\nment behaviour very close to vc. However, for many practical\napplications the equilibrium volume is significantly far from\nvc, so that one needs to include additional terms in Eq.26. In\norder to do so, we perform a Taylor expansion of the square\nof magnetic moment µ2around vc, that is\nµ2(v) =µ2(vc)+¶µ2\n¶v\f\f\f\f\f\nv=vc(v\u0000vc)\n+1\n2¶2µ2\n¶v2\f\f\f\f\f\nv=vc(v\u0000vc)2\n+1\n6¶3µ2\n¶v3\f\f\f\f\f\nv=vc(v\u0000vc)3+O((v\u0000vc)4);(27)\nwhere µ2(vc) =0. Making a square root on both sides of this\nequation we have\nµ(v) =q\naµ(v\u0000vc)+bµ(v\u0000vc)2+gµ(v\u0000vc)3\n\u0001Q(v\u0000vc);(28)\nwhere\naµ=¶µ2\n¶v\f\f\f\f\f\nv=vc\nbµ=1\n2¶2µ2\n¶v2\f\f\f\f\f\nv=vc\ngµ=1\n6¶3µ2\n¶v3\f\f\f\f\f\nv=vc:(29)\nThe Heaviside step function Q(v\u0000vc)was introduced in Eq.\n28 to ensure that the magnetic moment is zero at volumes\nlower than vc. The Taylor expansion was considered up tothe third order which is enough to correctly describe the mag-\nnetic moment of BCC Fe and FCC Ni within the range of\nvolume per atom discussed in this work ( v<20˚A3/atom). For\ncases with larger volume per atom than 20 ˚A3/atom one might\nneed to include higher order terms in the Taylor expansion.\nIn the vicinity of the critical volume ( (v\u0000vc)=vc\u001c1) Eq.\n28 becomes Eq.26, so that the result derived from the Landau\nexpansion is recovered33.\nAlternatively, instead of considering the volume depen-\ndence for the parameterization of the magnetic moment µ(v),\none could consider the pressure dependence of magnetic mo-\nment µ(P). In this case, one could apply the same procedure\nbut now performing the Taylor expansion around the critical\npressure Pcwhere the magnetic moment collapses. Note that\nthe function µ(P)could only be evaluated in this way from\nPcup to the negative pressure P0at which the pressure is re-\nversed due to the large interatomic distance, see Fig. 1. In this\nmodel, longitudinal fluctuations of magnetic moments at finite\ntemperature would naturally emerge from the fluctuations of\nthe volume per atom or pressure.\nFIG. 1. Schematic of the volume and hydrostatic pressure depen-\ndence of magnetic moment. Symbols µbandµirepresent the mag-\nnetic moment for bulk at zero-pressure and isolated atom, respec-\ntively.\nIII. SPIN-LATTICE MODEL FOR BCC FE AND FCC NI\nIn this section, we build a SD-MD model for BCC Fe and\nFCC Ni based on the methodology presented in Section II.\nThe construction of the model is splitted into the following\nstages in order to systematically compute each term in Eq.1,\nwhere the magnetic Hamiltonian is given by Eq.8.\nA. Interatomic potential\nIn the model we set the modified embedded atom method\n(MEAM) potentials developed by Asadi et al.43and Lee et\nal.44for the interatomic potential V(ri j)of BCC Fe and\nFCC Ni, respectively. These potentials give an elastic ten-\nsor very close to the experimental one at zero-temperature. In\nthis first stage, it is convenient to find the equilibrium vol-\nume and bulk modulus given by the model including only\nthe MEAM potential. To do so, we compute the energy of\na set of conventional unit cells with different volume using\nthe software LAMMPS18with the SPIN package26, and we6\nfit it to the Murnaghan equation of state (EOS)45,46. We ver-\nify that the pressure in the selected equilibrium state is lower\nthan 5\u000210\u00005GPa. In Fig.2, we present the calculation of\nthe energy versus volume curve for the conventional unit cell\n(2 atoms/cell for BCC Fe and 4 atoms/cell for FCC Ni). The\nequilibrium volume and bulk modulus found with this pro-\ncedure is v0=11:586754 ˚A3/atom and B=166:73 GPa for\nBCC Fe, and v0=10:903545 ˚A3/atom and B=188:85 GPa\nfor FCC Ni. Hence, the equilibrium distance to the first near-\nest neighbor is r0=2:4690386 ˚A for BCC Fe, and 2 :4890153\n˚A for FCC Ni. At this stage, it is also convenient to compute\nthe elastic constants. To do so, we evaluate the elastic ten-\nsor with software AELAS47interfaced with LAMMPS at the\nequilibrium volume v0including the MEAM potential. The\ndeveloped interface between AELAS and LAMMPS is avail-\nable on GitHub repository48. Here, we make use of the pro-\ngram Atomsk to convert some input files49. The calculated\nvalues and experimental ones are shown in Table II. We see\nthat these interatomic potentials gives a very similar elastic\ntensor to the experiment.\nFIG. 2. Calculation of the equation of the state for (top) BCC Fe and\n(bottom) FCC Ni with the SD-MD model including only the MEAM\npotential.\nB. Magnetic moment\nNext, we find the parameterization of the volume depen-\ndence of magnetic moment µ(v). Here, we estimate it usingDFT. Namely, the parameters vc,aµ,bµandgµin Eq.28 are\nobtained by fitting this equation to the magnetic moment ver-\nsus volume curve given by DFT. The DFT calculations are\nperformed with V ASP code50–52, which is an implementa-\ntion of the projector augmented wave (PAW) method53. We\nuse the interaction potentials generated for the Perdew-Burke-\nErnzerhof (PBE) version54of the Generalized Gradient Ap-\nproximation (GGA). We set an automatic Monkhorst-Pack k-\nmesh55gamma-centered grid with length parameter Rk=60.\nThe interactions were described by a PAW potential with 14\nand 16 valence electrons for BCC Fe and FCC Ni, respec-\ntively.\nFIG. 3. Calculation of the magnetic moment versus volume under\nnormal deformations obtained with DFT (blue dots) for BCC Fe and\nFCC Ni. Red line stands for the fitting curve.\nThe results of these calculations and corresponding fitting\ncurves are shown in Fig.3. Very similar results were previ-\nously reported by Moruzzi et al. using the augmented spher-\nical wave (ASW) method56. We see that the form of Eq.28\ndescribes quite well the data obtained by DFT. In the case\nof FCC Ni the deviation between the fitted curves and DFT\ndata is slightly larger than for BCC Fe. A better fit could be\nachieved by adding higher order terms in Eq.27. The values\nof the fitting parameters vc,aµ,bµandgµare presented in\nTable I. Inserting these values into Eq.28 allows us to com-\npute the magnetic moment at the equilibrium volume given\nby the SD-MD model including only MAEM potential ob-\ntained in Section III A ( v0=11:586754 ˚A3/atom for BCC Fe,\nandv0=10:903545 ˚A3/atom for FCC Ni). This calculation\ngives µ(v0) =2:34µBfor Fe, and µ(v0) =0:67µBfor Ni, while\nthe experimental values are 2 :22µBand 0 :606µBfor Fe and\nNi, respectively36. We see that this procedure overestimates\nslightly the magnetic moment.\nThe volume dependence of magnetic moment will allow us\nto study how magnetization changes under hydrostatic pres-\nsure (normal deformation). Note that in this model the magni-\ntude of the magnetic moment will not change under volume-\nconserving deformations. To check the validity of this approx-\nimation, we run some additional DFT calculations with V ASP\nusing the same setting as before to obtain the magnetic mo-\nment under volume-conserving tetragonal deformation57for\nBCC Fe. The results are plotted in Fig.4. We observe that a\nsignificant change of the magnetic moment only takes place at\nlarge tetragonal deformations. We verify that a similar trend is\nalso observed for other types of volume-conserving deforma-\ntions like trigonal deformation57. Therefore, to some extent,\nthe model might be able to describe the behaviour of magnetic7\nFIG. 4. Magnetic moment under a volume-conserving tetragonal de-\nformation of BCC Fe ( c=a=1) calculated with DFT.\nmoment and magnetization under deformations that combines\nan arbitrarily large normal deformation (which changes the\nvolume preserving the cubic symmetry) with a small volume-\nconserving deformation that changes the crystal symmetry.\nC. Exchange interaction\nLet’s now compute the parameterization of J(r). Firstly,\nnote that the equilibrium interatomic distance, EOS and elas-\ntic constants of the ground state (collinear state) are un-\nchanged after the exchange interaction is added to the SD-MD\nmodel thanks to the offset in the exchange energy. It is inter-\nesting to analyze the influence of different types of parame-\nterization of J(r)on the volume magnetostriction ws. Hence,\nfor the parameterization of J(r)we consider the following two\nsets of parameters for aJ,gJ,dJandRc;J.\n1. Set I: Effective short range exchange\nThe set I is calculated following the procedure described\nin Section II B, so that it leads to an effective short range\nexchange interaction. As mentioned above, the equilibrium\ndistance to the first nearest neighbor at the ground state is\nnot changed by the exchange interaction due to the offset\nin the exchange energy, so that according to Eq.24 we set\nd(I)\nJ=r0=2:4690386 ˚A for BCC Fe, and d(I)\nJ=2:4890153 ˚A\nfor FCC Ni. Next, we see in Eq. 23 that we need as inputs TC,\nc11,c12andwsto compute the J(r0)and¶J=¶r. In general,\nthese inputs can be obtained by theory or experiment. For in-\nstance, here we use the experimental value of TC(1043 K for\nBCC Fe and 627 K for FCC Ni)36. For the elastic constants,\nwe will make use of the theoretical values obtained by the SD-\nMD model itself using only the MEAM potential (see Table\nII). The experimental measurement of volume magnetostric-\ntion is difficult, and one can find significant discrepancies be-tween different works15,58. Hence, we will use the theoretical\nvalue of wsat zero-temperature calculated by Shuimizu using\nthe itinerant electron model58, that is ws=1:16\u000210\u00002for\nBCC Fe and 3 :75\u000210\u00004for FCC Ni. Inserting these quan-\ntities into Eq.24 via Eq.23 gives a(I)\nJ=\u000012:5921 meV/atom\nandg(I)\nJ=2:81897 for BCC Fe, and a(I)\nJ=8:35847 meV/atom\nandg(I)\nJ=\u00000:098217 for FCC Ni.\n2. Set II: Long range exchange\nThe second set of parameters (set II) is obtained by fitting\nthe Bethe-Slater function to the exchange integrals given by\nfirst-principles calculations26. The fitted parameters ( a(II)\nJ,\ng(II)\nJandd(II)\nJ) are shown in Table I. The value of a(II)\nJtaken\nfrom Ref.26has been multiplied by 2 due to the factor 1 =2\nin the exchange energy given by Eq.8. Here, we set a large\ncut-off R(II)\nc;J=4:5˚A to take into account the exchange inter-\nactions beyond first nearest neighbors (long range exchange\ninteraction). The Bethe-Slater function with parameters from\nset I and II is plotted in Fig.5. This figure will be analyzed in\nthe context of volume magnetostriction in Section IV B 4.\nD. N ´eel energy\nNow, we are in a position to calculate the Bethe-Slater pa-\nrameters for the dipole and quadrupole terms of the N ´eel in-\nteraction given by Eqs.22 and 17, respectively. Firstly, we\nnotice that a key quantity in these equations is the equilib-\nrium distance to the first nearest neighbors r0, which obvi-\nously depends on the N ´eel interaction. Fortunately, the en-\nergy of the dipole and quadrupole terms of the N ´eel interac-\ntion for Fe and Ni are of the order of µeV/atom (see Fig.7),\nso that they are much lower than the total energy (eV/atom).\nAs a result, these terms only induce a very small change in\nr0when they are included in the SD-MD model. This fact al-\nlows us to use r0given by the SD-MD model including only\nthe MEAM potential and exchange interaction to calculate the\nBethe-Slater parameters for the dipole and quadrupole terms\nof the N ´eel interaction. Hence, according to Eqs.17 and 22,\nwe can set dl=dq=r0=2:4690386 ˚A for BCC Fe, and\ndl=dq=2:4890153 ˚A for FCC Ni.\nOnce r0is determined, we calculate aqandgqusing Eqs.\n13, 14 and 17. Here, we set the experimental values of K1and\n(1=K1)(¶K1=¶P)approximately at zero-temperature, that is,\nK1=55 KJ/m3and(1=K1)(¶K1=¶P) =\u00007:3\u000210\u00002GPa\u00001\nfor BCC Fe, and K1=\u0000126 KJ/m3and(1=K1)(¶K1=¶P) =\n\u00002:8\u000210\u00002GPa\u00001for FCC Ni59,60. As we see in Eq.17,\nwe also need the bulk modulus. In principle we could set its\nexperimental value or the one given by the EOS of this SD-\nMD model that was obtained in Section III A. In this work\nwe choose the second option in order to describe more accu-\nrately the relation between volume and pressure of the SD-\nMD model. Inserting all these quantities in Eq.17 via Eqs. 13\nand 14 leads to aq=28:5189 µeV/atom and gq=1:05331 for8\nFIG. 5. Calculation of the Bethe-Slater function J(r)andq(r)for\n(top) BCC Fe and (bottom) FCC Ni using the two set of parameters\ngiven in Table I. Vertical dash line stands for the equilibrium distance\nof the first nearest neighbors r0.\nBCC Fe, and aq=\u000049:1335 µeV/atom and gq=1:1186 for\nFCC Ni.\nLastly, we calculate the Bethe-Slater parameters for the\ndipole term ( alandgl) using Eqs. 21 and 22. In this case we\nneed the values of the anisotropic magnetoelastic constants b1\nandb2. These constants are related to the magnetostrictive\ncoefficients ( l001andl111) and elastic constants ( ci j) via38,39\nb1=\u00003\n2l001(c11\u0000c12);\nb2=\u00003l111c44:(30)\nTo calculate b1andb2we use the experimental magnetostric-\ntive coefficients l001=26\u000210\u00006andl111=\u000030\u000210\u00006for\nBCC Fe, and l001=\u000060\u000210\u00006andl111=\u000035\u000210\u00006for\nFCC Ni at zero-temperature36. For the values of the elas-\ntic constants we choose the calculated ones with the SD-MD\nmodel including only the MEAM potential (see Table II).\nDoing so, we get b1=\u00003:74166 MJ/m3andb2=10:4643MJ/m3for BCC Fe, and b1=10:0611 MJ/m3and b2=\n13:9398 MJ/m3for FCC Ni. If we insert these values in\nEq.22 via Eq.21, then we obtain al=392:747µeV/atom and\ngl=0:824409 for BCC Fe, and al=179:396µeV/atom and\ngl=1:39848 for FCC Ni.\nTABLE I. Parameters of the SD-MD model for BCC Fe and FCC Ni.\nSD-MD model\nparametersBCC Fe FCC Ni\naµ(µ2\nB\u0001atom/ ˚A3) 1.49057 0.172931\nbµ(µ2\nB\u0001atom2/˚A6) -0.0978406 -0.021997\ngµ(µ2\nB\u0001atom3/˚A9) 0.0026366 0.00096755\nvc(˚A3/atom) 6.39848 5.36535\na(I)\nJ(meV/atom) -12.5921 8.35847\ng(I)\nJ2.81897 -0.098217\nd(I)\nJ(˚A) 2.4690386 2.4890153\nR(I)\nc;J(˚A) 2.6 2.6\na(II)\nJ(meV/atom) 50.996a19.46a\ng(II)\nJ0.281a0.00011a\nd(II)\nJ(˚A) 1.999a1.233a\nR(II)\nc;J(˚A) 4.5a4.5a\nal(µeV/atom) 392.747 179.396\ngl 0.824409 1.39848\ndl(˚A) 2.4690386 2.4890153\nRc;l(˚A) 2.6 2.6\naq(µeV/atom) 28.5189 -49.1335\ngq 1.05331 1.1186\ndq(˚A) 2.4690386 2.4890153\nRc;q(˚A) 2.6 2.6\naRef.26\nThe Bethe-Slater parameters for the constructed SD-MD\nmodels are shown in Table I, while the corresponding Bethe-\nSlater functions for l(r)andq(r)using these parameters are\nplotted in Fig.6. We see that l(r0)is approximately two or-\nder of magnitude greater than q(r0). However, note that af-\nter taking into account all first nearest neighbors the N ´eel\nquadrupole and dipole energies can be of the same order of\nmagnitude close to the cubic symmetry (see Section IV A).\nFig.6 also contains interesting information about the depen-\ndence of MCA and magnetoelasticity on the distance between\nfirst nearest neighbors. For instance, we see that if we de-\ncrease the distance between first nearest neighbors from the\nequilibrium value r0(high hydrostatic pressure regime) for\nboth BCC Fe and FCC Ni, then the sign of q(r)changes,\nwhich implies a change in the sign of K1, see Eq.13. Similarly,\nif we increase the distance between first nearest neighbors for\nBCC Fe ( r0>3˚A), then the sign of l(r)changes switching the\nsign of b1, see Eq.21. In general, the physical interpretation of\nq(r)andl(r)far from the equilibrium value r0should be done\nwith caution since we only involved up to the first derivative\nof these functions evaluated at r0in their parameterization.\nIn this sense, the only meaningful region around r0may be\nwhere first order Taylor expansion at r0of the Bethe-Slater9\nfunctions of q(r)andl(r)is a good approximation. Includ-\ning up the first derivative of q(r)andl(r)in their parame-\nterization might be enough for many practical purposes since\nthe distance between first nearest neighbors oscillates close to\nthe equilibrium value at finite temperature below the melting\npoint.\nFIG. 6. Calculation of the Bethe-Slater function l(r)andq(r)for\n(top) BCC Fe and (bottom) FCC Ni using the parameters given in\nTable I. Vertical dash line stands for the equilibrium distance of the\nfirst nearest neighbors r0.\nIV . RESULTS\nA. Tests of the N ´eel interaction\nBefore evaluating the magnetoelastic properties of the SD-\nMD model, it is convenient to check that the implementation\nof the N ´eel interaction Eq.4 in the SD-MD simulation is cor-\nrect. To this end, we propose some tests by comparing the\nnumerical results of the SD-MD simulation with simple an-\nalytical solutions. For instance, if we consider a BCC struc-\nture with N ´eel interactions up to first nearest neighbor in acollinear state along sss= (0;0;1), then from Eq.6 we have\nHN´eel(0;0;1) =16Nq(r0)\n45; (31)\nwhere Nis the number of atoms in the system, r0is the\ndistance to nearest neighbor that is related to the lattice pa-\nrameter aviar0=ap\n3=2. This equation allows to verify\nthe quadrupole term. Let’s now apply to this system with\nsss= (0;0;1)a tetragonal deformation along the z-axis, where\nthe lattice parameter is cin this direction, and aalong both\nx-axis and y-axis. From Eq.6 we obtain\nHN´eel(0;0;1) =\u00004Nl(r0)\"\u0000c\na\u00012\n2+\u0000c\na\u00012\u00001\n3#\n\u000016Nq(r0)h\n2\u0000c\na\u00014\u000012\u0000c\na\u00012+3i\n35h\n2+\u0000c\na\u00012i2;(32)\nwhere\nr0=a\n2r\n2+\u0010c\na\u00112\n: (33)\nThis equation allows to check both the dipole and quadrupole\nterms. In the limit c=a\u0000 !1, the Eq.32 becomes Eq.31 ensur-\ning the continuity of the N ´eel energy under structure defor-\nmation. In Fig.7, we verify that the calculation of the N ´eel\nenergy with LAMMPS is the same to Eqs.31 and 32 using the\nBethe-Slater parameters of BCC Fe given in Table I. Similar\ntests could also be performed for other magnetic moment di-\nrections and deformations.\nFIG. 7. Calculation of the N ´eel energy with LAMMPS and (top)\nEq.32 and (bottom) Eq.31 for different values of the lattice parame-\nters.\nB. Magnetic properties at zero-temperature\nIn this section, we evaluate the magnetization and MCA\nunder pressure, anisotropic magnetostrictive coefficients, vol-10\nTABLE II. Calculated and experimental elastic constants, magnetostrictive coefficients, MCA, and MCA under hydrostatic pressure for BCC\nFe and FCC Ni at zero-temperature.\nMaterialElastic\nconstantsSD-MD\n(GPa)Expt.\n(GPa)Magnetostrictive\ncoefficientsSD-MD\n(\u000210\u00006)Expt.\n(\u000210\u00006)MCASD-MD\n(KJ/m3)Expt.\n(KJ/m3)MCA vs PSD-MD\n(GPa\u00001)Expt.\n(GPa\u00001)\nBCC Fe c11 230.0 230al001 25.9 26cK1 54.995 55d 1\nK1¶K1\n¶P-0.0727 -0.073e\nc12 134.1 135al111 -30.3 -30c\nc44 116.3 117a\nFCC Ni c11 263.9 261.2bl001 -61.9 -60cK1-125.996 -126d 1\nK1¶K1\n¶P-0.0279 -0.028e\nc12 152.1 150.8bl111 -35.4 -35c\nc44 132.8 131.7b\naRef.43,bRef.44,cRef.36,\ndRef.59,eRef.60\nume magnetostriction and saturation magnetization at zero-\ntemperature given by the developed SD-MD models for BCC\nFe and FCC Ni in Section III. We include MEAM potentials,\nexchange and N ´eel energies, and volume-dependent magnetic\nmoment in the following calculations. Magnetic collinear\nstates will be used since we are interested in properties at zero-\ntemperature. All simulations are performed with the SPIN\npackage of LAMMPS26.\n1. Ground state\nFirstly, we determine the equilibrium volume of the full\nSD-MD model (including the N ´eel interaction) for the con-\nventional unit cell of BCC Fe and FCC Ni. To this end, we\ncalculate the energy versus volume curve, and we fit it to the\nMurnaghan EOS in the same way as it was done in Fig.2\npreviously. Here, we also set the magnetic moments along\nthe easy direction ( [1;0;0]for BCC Fe and [1;1;1]for FCC\nNi) in order to get the minimum energy of the quadrupole\nterm of N ´eel interaction. The equilibrium volume found with\nthis procedure is v0=11:5867635 ˚A3/atom for BCC Fe, and\nv0=10:9035445 ˚A3/atom for FCC Ni. We verify that pres-\nsure is lower than 5 \u000210\u00005GPa in these equilibrium states.\nAs we anticipated in Section III, the dipole and quadrupole\nN´eel interactions induce a very small change in the equilib-\nrium volume when is included in the SD-MD model.\n2. Magnetocrystalline anisotropy\nNext, we compute the MCA energy at this equilibrium\nvolume by setting the magnetic moment along different di-\nrections in the XY plane. In Fig.8 we show a compari-\nson between the MCA energy calculated by SD-MD simula-\ntions with LAMMPS and Eq.11 using the experimental value\n(K1=55 KJ/m3for BCC Fe and K1=\u0000126 KJ/m3)59. The\ndirect evaluation of K1with the SD-MD model through Eq.12\ngives 54 :995 KJ/m3for BCC Fe and\u0000125:996 KJ/m3for\nFCC Ni. As we see, the SD-MD model with the Bethe-Slater\nparameters given by Table I reproduces very well the first-\norder experimental MCA.\nFIG. 8. Calculation of the MCA energy for BCC Fe and FCC Ni with\nSD-MD simulation (blue points) and Eq.11 using the experimental\nK1(red line). Magnetic moments are constrained on the XY plane.\nFIG. 9. Calculation of K1(P)=K1(0)under hydrostatic pressure using\nthe developed SD-MD model (blue dots) for BCC Fe and FCC Ni.\nThe green and red lines stand for the experimental behaviour given\nby Eq.35 and its low-pressure approximation Eq.36, respectively60.\nNow we study the effects of hydrostatic pressure on the\nMCA for this SD-MD model. To facilitate the compar-\nison between the model and experiment, we first convert\n(1=K1)(¶K1=¶P)to an integral form, that is,\n1\nK1¶K1\n¶P=z\u0000!ZK1(P)\nK1(0)dK1\nK1=ZP\n0zdP; (34)\nwhere z=\u00007:3\u000210\u00002GPa\u00001is the experimental value mea-\nsured up to P=0:5 GPa at T=77K for BCC Fe60, while for\nFCC Ni is z=\u00002:8\u000210\u00002GPa\u00001. Solving this integral we\nhave\nK1(P)\nK1(0)=ezP; (35)\nwhere in the low pressure regime ( zP\u001c1) it can be written11\nas\nK1(P)\nK1(0)\u00191+zP+O(P2): (36)\nIn Fig.9 we show the ratio K1(P)=K(0)versus pressure gener-\nated by the SD-MD model of Fe and Ni, and the experimen-\ntal behaviour given by Eq.35 and its low-pressure approxi-\nmation Eq.36. The linear fitting to the data generated by the\nSD-MD model up to P=0:5 GPa gives (1=K1)(¶K1=¶P) =\n\u00007:27\u000210\u00002GPa\u00001for BCC Fe, and\u00002:79\u000210\u00002GPa\u00001\nfor FCC Ni, which is in very good agreement with the ex-\nperimental values60. Note that Eq.35 and MCA results of\nthe model beyond the range of pressure between 0GPa and\n0:5GPa should be taken with caution due to the lack of exper-\nimental data.\nFIG. 10. Calculation of l001for BCC Fe using MAELAS inter-\nfaced with LAMMPS. (top) Quadratic curve fit to the energy ver-\nsus cell length along bbb= (0;0;1)with spin direction sss1= (0;0;1)\nunder a volume-conserving tetragonal deformation. (bottom) En-\nergy difference between states with spin directions sss2= (1;0;0)and\nsss1= (0;0;1)against the cell length along bbb= (0;0;1).\n3. Anisotropic magnetostriction\nNow, we compute the anisotropic magnetostrictive coeffi-\ncients using the SD-MD model. To this end, we apply the\nmethod proposed by Wu and Freeman61,62as implemented in\nthe program MAELAS41,57. In this method, the anisotropic\nFIG. 11. Calculation of l111for BCC Fe using MAELAS inter-\nfaced with LAMMPS. (top) Quadratic curve fit to the energy ver-\nsus cell length along bbb=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nwith spin direction\nsss1=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nunder a volume-conserving trigonal de-\nformation. (bottom) Energy difference between states with spin\ndirections sss2=\u0010\n1=p\n2;0;\u00001=p\n2\u0011\nandsss1=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nagainst the cell length along bbb=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\n.\nmagnetostrictive coefficients for cubic systems (point groups\n432, ¯43m,m¯3m) are calculated as57\nl001=4(l001\n1\u0000l001\n2)\n3(l001\n1+l001\n2);l111=4(l111\n1\u0000l111\n2)\n3(l111\n1+l111\n2); (37)\nwhere l001\n1and l001\n2are the equilibrium cell lengths along\nthe length measuring direction bbb= (0;0;1)under a tetrag-\nonal deformation with collinear magnetic moment direc-\ntions sss1= (0;0;1)and sss2= (1;0;0), respectively. Simi-\nlarly, l111\n1and l111\n2are the equilibrium cell lengths along\nthe length measuring direction bbb= (1=p\n3;1=p\n3;1=p\n3)un-\nder a trigonal deformation with magnetic moment direction\nsss1=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nandsss2=\u0010\n1=p\n2;0;\u00001=p\n2\u0011\n, re-\nspectively. In order to obtain the equilibrium cell lengths l001\n1\nandl001\n2, one needs to evaluate the energy for a set of volume-\nconserving tetragonal distorted unit cells. Next, the energy\nversus the cell length along bbb= (0;0;1)for each magnetic\nmoment direction sss1= (0;0;1)andsss2= (1;0;0)is fitted to a12\nquadratic function\nE(x)\f\f\f\f\fsssj\nbbb=(0;0;1)=˜ajx2+˜bjx+˜cj;j=1;2 (38)\nwhere ˜ aj,˜bjand ˜ cjare fitting parameters. The mini-\nmum of this quadratic function for magnetic moment di-\nrection sss1(2)corresponds to l001\n1(2)=\u0000˜b1(2)=(2 ˜a1(2)), and it\nis the equilibrium cell length. Similarly, the equilibrium\ncell lengths l111\n1andl111\n2are obtained by applying a set of\nvolume-conserving trigonal deformations, and performing a\nquadratic fitting of the energy versus the cell length along\nbbb=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nwith magnetic moment directions\nsss1=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nandsss2=\u0010\n1=p\n2;0;\u00001=p\n2\u0011\n.\nFIG. 12. Calculation of l001for FCC Ni using MAELAS inter-\nfaced with LAMMPS. (top) Quadratic curve fit to the energy ver-\nsus cell length along bbb= (0;0;1)with spin direction sss1= (0;0;1)\nunder a volume-conserving tetragonal deformation. (bottom) En-\nergy difference between states with spin directions sss2= (1;0;0)and\nsss1= (0;0;1)against the cell length along bbb= (0;0;1).\nWe have developed an interface between the software\nMAELAS57and LAMMPS26in order to apply this method\nand extract the magnetostrictive coefficients easily. This in-\nterface is publicly available on GitHub repository63. In Fig.10\nwe show the quadratic curve fit to the energy versus cell length\nalong [0;0;1]with magnetic moment direction sss1= (0;0;1)\nto calculate l001for BCC Fe. We also plot the energy dif-\nference between states with spin directions sss1= (1;0;0)and\nsss2= (0;0;1)against the cell length along [0;0;1]. The cor-\nresponding plot for l111is presented in Fig.11. We obtainl001=25:9\u000210\u00006andl111=\u000030:3\u000210\u00006, while the ex-\nperimental values36atT=4:2K are l001=26\u000210\u00006and\nl111=\u000030\u000210\u00006. The results for FCC Ni are plotted\nin Figs. 12 and 13. Here, we get l001=\u000061:9\u000210\u00006\nandl111=\u000035:4\u000210\u00006, while the experimental values36at\nT=4:2K are l001=\u000060\u000210\u00006andl111=\u000035\u000210\u00006.\nTherefore, the developed SD-MD model for Fe and Ni also\nexhibits magnetostrictive properties very similar to the exper-\niment. Additionally, this calculation reveals that the method\nproposed by Wu and Freeman61,62is an excellent approach\nto obtain the magnetostrictive coefficients as long as both the\nelastic and magnetoelastic energies are properly described by\nthe model. This fact could not be verified before for l111of\nBCC Fe due to a possible failure of Density Function Theory\ncalculations57,64–66. In Table II we present a summary of the\nresults given by the SD-MD model for the MCA, MCA under\nhydrostatic pressure, and anisotropic magnetostrictive coeffi-\ncients.\nFIG. 13. Calculation of l111for FCC Ni using MAELAS inter-\nfaced with LAMMPS. (top) Quadratic curve fit to the energy ver-\nsus cell length along bbb=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nwith spin direction\nsss1=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nunder a volume-conserving trigonal de-\nformation. (bottom) Energy difference between states with spin\ndirections sss2=\u0010\n1=p\n2;0;\u00001=p\n2\u0011\nandsss1=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\nagainst the cell length along bbb=\u0000\n1=p\n3;1=p\n3;1=p\n3\u0001\n.13\nTABLE III. Calculated volume magnetostriction wswith the SD-MD\nmodel for BCC Fe and FCC Ni using the set I and II of parameters in\nTable I to describe J(r). Theoretical and experimental results found\nin literature are also shown for comparison.\nSD-MD\nset I\n(\u000210\u00004)SD-MD\nset II\n(\u000210\u00004)Theory\n(\u000210\u00004)Expt.\n(\u000210\u00004)\nBCC Fe 118 -235 116a4b\n683c\nFCC Ni 3.71 -53.7 3.75a3.65e\n45.7c3.24f\n-5.1d\n-2.7b\naRef.58,bRef.67,\ncRef.68,dRef.69,\neRef.70,fRef.71\n4. Volume magnetostriction\nThe volume magnetostriction is generated by the presence\nof ferromagnetism in the magnetic material (exchange mag-\nnetostriction). It can be calculated as72\nws(T) =v0(Ms(T))\u0000v0(0)\nv0(0); (39)\nwhere v0(Ms(T))andv0(0)are the equilibrium volume per\natom in the magnetized and demagnetized (paramagnetic)\nstates, respectively. In the magnetized state, the magnetiza-\ntion is equal to the saturation magnetization Msat temperature\nT. Hence, the quantity v0(Ms(T))at zero-temperature was al-\nready calculated in Section IV B 1. To compute v0(0)we ap-\nply a similar procedure. Namely, we first calculate the energy\nof a supercell with magnetic moments oriented randomly (de-\nmagnetized state) for different values of the lattice parameter\na, preserving the cubic crystal symmetry. Next, we fit the en-\nergy versus volume curve to the Murnaghan EOS. We use a\nsupercell with 20x20x20 conventional unit cells with periodic\nboundary conditions for both BCC Fe (16000 atoms) and FCC\nNi (32000 atoms). We perform this calculation using the set I\nand II of parameters given in Table I to describe the exchange\ninteraction J(r). The results are depicted in Fig. 14. The set I\ngivesws=1:18\u000210\u00002for BCC Fe and 3 :71\u000210\u00004for FCC\nNi, reproducing fairly well the theoretical values calculated by\nShimizu58(ws=1:16\u000210\u00002for BCC Fe and 3 :75\u000210\u00004for\nFCC Ni) that we used to compute the Bethe-Slater parameters\nforJ(r)in Section III C. The set II leads to ws=\u00002:23\u000210\u00002\nfor BCC Fe and\u00005:37\u000210\u00003for FCC Ni, so they have the\nopposite sign to the results given by set I. According to Eq.\n23, these results may be understood in terms of ¶J=¶rat the\nfirst-nearest neighbors ( r=r0) since wsµ ¶J=¶r. In Fig.5,\nwe observe that set I gives ¶J=¶r>0 at r=r0for both Fe\nand Ni, while set II gives ¶J=¶r<0 atr=r0. Note that Eq.\n23 is derived assuming only exchange interactions up to first-\nnearest neighbors, and set II has a large cut-off that includes\nexchange interactions beyond first-nearest neighbors. Wang\net al. performed first-principles calculations of J(r)finding achange in the sign of ¶J=¶rclose to r=r0, and¶J=¶r>0 for\nthe second nearest neighbors73. Previous theoretical and ex-\nperimental works reported a positive volume magnetostriction\nfor BCC Fe58,67,68, while for FCC Ni one can find contradic-\ntory results with positive58,68,70,71and negative67,69values. A\nsummary of these results is presented in Table III. As seen in\nFig.5, there is a maximum of J(r)close to r0for Ni using the\nset I, so that a small increase in the lattice parameter would\nchange the sign of ¶J=¶r, and consequently the sign of ws.\nLastly, we point out that the isotropic magnetostrictive coeffi-\ncient of cubic crystals ( la) and magnetoelastic constant b0are\nrelated to the volume magnetostriction as57,74\nla=\u0000b0\u00001\n3b1\nc11+2c12=ws\n3: (40)\nHence, we see that the isotropic magnetostriction is greater\nthan the anisotropic one for both BCC Fe and FCC Ni.\nFIG. 14. Calculation of wswith the SD-MD model for ((a)-(b)) BCC\nFe and ((c)-(d)) FCC Ni using the two set of parameters given in\nTable I to describe the exchange interaction J(r).\n5. Saturation magnetization\nThe saturation magnetization at zero-temperature is com-\nputed using the following equation\nµ0Ms(v) =µ0µ(v)\nv; (41)\nwhere µ(v)is calculated using the Eq.28 with the parameters\nshown in Table I. At the equilibrium volume of the SD-MD\nmodel it gives µ0Ms(v0) =2:35T for BCC Fe, and 0 :71T for\nFCC Ni. The experimental values at zero-temperature are\nµ0Ms=2:19T for BCC Fe, and 0 :64T for FCC Ni36. We see\nthat the model slightly overestimates the saturation magne-\ntization. Next, we evaluate Msfor different volumes apply-\ning normal deformations. The results of this calculation are\nshown in Fig.15. Here, we also included the data given by\nDFT that we obtained in Section III B. We observe that the14\noverall behaviour of Msis well described by the model. As\nwe increase the volume above the equilibrium volume v0, the\npressure becomes negative and Msis decreasing. The condi-\ntion that causes Msto decrease with volume is\n¶Ms\n¶v<0\u0000!¶µ\n¶v<µ\nv: (42)\nOn the other hand, if we decrease the volume below the equi-\nlibrium volume then the pressure is positive. At high positive\npressure, Msbecomes zero when the volume per atom is lower\nthan the critical volume ( v2GPa. The data analysis in the frame of a phe-\nnomenological theory of FMR reveals that the seemingly\nisotropic magnetism of the studied crystals manifesting\nin the gap closure is a result of the compensation of\nthe intrinsic easy-axis magnetocrystalline anisotropy of\nCr2Ge2Te6and the shape anisotropy of the sample. It\nfollows from our analysis that the MAE constant KU\nis reduced by a factor of 2 at the highest applied pres-\nsure in the FMR experiment but still remains sizable.\nThis suggests a robustness of the easy-axis type ferro-\nmagnetism of Cr 2Ge2Te6to the application of pressure\nup toP= 2.39GPa despite a reduction of Tc. Our find-\nings motivate further pressure experiments to address\nthe question of whether the sign of the MAE could be\nchanged at still high pressures before the ferromagnetic\nstate would be fully suppressed. From the technological\nperspective, the results of the present work on the pres-\nsure tunability of the magnetic anisotropy of Cr 2Ge2Te6\nmay be insightful for accessing the use of this material\nas a magnetic element of spintronic devices with strained\narchitecture.\nII. SAMPLES AND METHODS\nSingle crystals of Cr 2Ge2Te6were grown with the self\nflux technique and thoroughly characterized as described\nin detail in our previous work in Ref. [ 9]. As grown sin-\ngle crystals were thoroughly characterized by powder x-\nray diffraction and energy dispersive x-ray spectroscopy,\nboth agree well with the crystal structure in the R ¯3\nspace group as well as with the expected stoichiometry\nof Cr2Ge2Te6[12].\nThe bulk magnetization data were measured using\na custom-built pressure cell for a commercial Quan-\ntum Design superconducting quantum interference de-\nvice (SQUID) magnetometer (MPMS-XL). Inside the\nCuBecell, twoopposingcone-shapedceramicanvilscom-\npress a CuBe gasket with a cylindrical hole used as sam-\nple chamber [ 13,14]. The plateletlike-shaped single-\ncrystalline sample ( m∼3.3×10−6g) of Cr 2Ge2Te6with\nthecaxis normal to the plate was installed into the gas-\nket hole (diameter Ø = 0.6 mm and height h= 0.8 mm)\nandrestedontheflatpartoftheceramicanvil. Giventhe\nlongitudinal magnetic field direction in the SQUID mag-\nnetometer the H/bardblcorientationwaseasily achieved. The\nuniaxialforceisappliedat ambienttemperature, andit is\nconverted into hydrostatic pressure in the sample cham-\nber using Daphne oil 7575 as the pressure-transmitting\nmedium. We checked the pressure value and its homo-\ngeneity at low temperature by measuring the pressure-dependent diamagnetic response associated with the su-\nperconducting transition of a small Pb manometer in-\nserted in the sample space [ 15]. Such measurements were\nperformed in applied fields of H= 5 Oe. In order to en-\nsure a stable pressure throughout all measurements due\nto the thermal expansion of the cell upon temperature\nsweeps, athermalcyclingwasperformedoncefromambi-\nent temperature to 2 K and back to ambient temperature\nbefore the actual data acquisition.\nNotethatamagneticbackgroundisdetectedinourDC\nmagnetization measurements arising from the pressure\ncell itself. Here, we performed a detailed magnetic char-\nacterization of the empty pressure cell within the same\nconditionsastherealexperimentswiththeaimofaquan-\ntitative and reliable disentanglement of the sample sig-\nnal from the overall magnetic response. The background\ncorrection was done by subtracting the fitted signal of a\ngasketwithout a sample measured under the same condi-\ntions, given that the raw signal of the gasket in our setup\ncan be described by a magnetic dipole in our SQUID\ndetection coils. While at temperatures above ∼20 K\na temperature-independent pressure cell background en-\nsures excellent background subtraction, at low temper-\natures a strongly temperature dependent magnetization\nof the CuBe cell limits the resolution of our experiments.\nConsequently, only background-subtracted data for T >\n10 K are shown. Small kinks in the magnetization curves\naround 40 K were identified as instrumental artifacts.\nMore detailed information about the background of our\npressurecell can be found in Ref. [ 16]. Still, the combina-\ntion of the very small magnetization above the ferromag-\nnetic ordering temperature of our samples with a very\nsmall mass ( m∼3.3×10−6g) and the large uncertainty\nof absolute pressure values for T >150 K using Pb as\na manometer at low Tin our ceramic anvil pressure cell\nrestrains us from a quantitative Curie-Weiss analysis in\nthe high-temperature regime.\nFerromagneticresonanceis a resonanceresponseof the\ntotal magnetization of the ferromagnetically ordered ma-\nterial exposed to the microwaveradiation. The FMR fre-\nquency not only is determined by the strength of the ap-\nplied external magnetic field but also sensitively depends\non the magnetic anisotropy of the studied sample, as\nwill be explained in Sec. IIIB. The magnetic anisotropy\ncan be accurately quantified by measuring FMR at dif-\nferent excitation frequencies. Therefore, in the present\nworkFMR experiments were carried out using a multi-\nfrequency electron spin resonance setup equipped with\na piston-cylinder pressure cell with a maximum pres-\nsure of about 2.5 GPa. Its detailed description can be\nfound in Ref. [ 17]. Microwave radiation with frequen-\ncies in the range 50 – 260GHz was provided by a set\nof Gunn diode oscillators and detected by a4He-liquid\ncooled hot-electron InSb bolometer. Magnetic fields up\nto 10T were generated by a cryogen-free superconduct-\ning magnet. Plateletlike single-crystalline samples of\nCr2Ge2Te6with typical lateral dimensions of ∼3−4mm\nwere put inside the pressure cell in a Teflon capsule filled3\n/s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48 /s49/s50/s48 /s49/s52/s48 /s49/s54/s48 /s49/s56/s48 /s50/s48/s48/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52/s49/s54/s49/s56\n/s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48 /s55/s48 /s56/s48 /s57/s48 /s49/s48/s48 /s49/s49/s48 /s49/s50/s48/s45/s48/s46 /s49/s54/s45/s48/s46 /s49/s52/s45/s48/s46 /s49/s50/s45/s48/s46 /s49/s48/s45/s48/s46 /s48/s56/s45/s48/s46 /s48/s54/s45/s48/s46 /s48/s52/s45/s48/s46 /s48/s50/s48/s46 /s48/s48/s48/s46 /s48/s50/s77 /s32/s40/s101/s109/s117/s47/s103/s41\n/s84 /s32/s40/s75/s41/s32/s48/s32/s71/s80/s97\n/s32/s49/s46/s51/s32/s71/s80/s97\n/s32/s50/s46/s50/s32/s71/s80/s97\n/s32/s50/s46/s56/s32/s71/s80/s97\n/s32/s51/s46/s52/s32/s71/s80/s97/s100/s77/s47/s100/s84\n/s84 /s32/s40/s75/s41\nFIG. 1. Temperature dependence of the magnetization M(T)\nof Cr2Ge2Te6measured in the FC mode at µ0H= 0.1T par-\nallel to the caxis for pressures up to 3.4GPa. The inset shows\nthe corresponding dM(T)/dTcurves.\nwith a pressure-transmitting fluid (Daphne 7373 oil from\nIdemitsu KosanCo., Ltd [ 18–20]). The advantageofhav-\ning a well-defined c-axis direction of the samples facili-\ntated their orientation in the FMR pressure cell. For\ntheH/bardblcfield geometry the studied sample was placed\non the flat bottom of the Teflon capsule. For H⊥cit\nwas firmly fixed with a thin Teflon tape to the side of\na small rectangular parallelepiped placed on the bottom\nof the Teflon capsule. The inner pistons of the pressure\ncell made of ZrO 2-based ceramics ensured low-loss prop-\nagation of the microwavesthrough the cell. The pressure\nwas calibrated using a superconducting tin-based pres-\nsure gauge. All FMR measurements were performed at a\ntemperature T= 4.2K for two orientations of the sample\nwith respect to the applied magnetic field, H/bardblcaxis and\nH⊥caxis.\nIII. EXPERIMENTAL RESULTS AND\nDISCUSSION\nA. Pressure-dependent magnetization study\nThe field-cooled (FC) magnetization curves M(T) as\na function of temperature are shown in Fig. 1for an ap-\nplied magnetic field µ0H= 0.1T oriented parallel to the\ncaxis under an applied pressure up to 3.4GPa. Due to\nthe small magnetization values, resulting from the very\nsmallmassofthesample,theabsolutevaluesathightem-\nperatureshow asmall artificial shift in the magnetization\naxis, coming from a highly sensitive background subtrac-\ntion.Two main characteristics are observed. First, the\nonset of the magnetic transition at approximately80K ispersistent within ±1K despite the pressure change. The\nonset was determined from the derivative dM/dTcurve\n(Fig.1, inset)whilemovingfromhightolowtemperature\nas the point of its departure from a very small, close to\nzerovalue dM/dT∼0causedbyjustasmallpolarization\nof paramagnetic moments to the negative values due the\ngradual development of the spontaneous ferromagnetic\nmoment. Second, the transition broadens continuously,\nand the ferromagnetic (FM) transition temperature Tc,\ndefined as the minimum of the dM(T)/dTcurve [21,22],\nshifts to lower temperature upon increasing pressure, as\nshown in the inset of Fig. 1. This behavior is in line with\nan earlier report [ 10] in which magnetization measure-\nments at pressures up to 1GPa showed the same ten-\ndency. Remarkably, the absolute value of the magneti-\nzation in the FM state at the given field of the measure-\nment is gradually reduced with increasing pressure up to\n3.4GPa. This can be explained by a shift of the satura-\ntion field Hsatas a function of pressure to higher values,\nwhich reduces consequently the magnetization at small\napplied fields such as 0.1T (see Fig. 2and discussion be-\nlow).\nThe magnetic field dependence of the magnetization\nM(H) is depicted in Fig. 2for the same pressure values\nas shown for the temperature dependences. To minimize\nthe influence of the nonperfect subtraction of the back-\nground signal from the pressure cell at low temperatures,\nthe curves were measured at T= 15K, which is still low\nenough to capture the characteristic behavior of the FM\nstate (see Fig. 1). TheM(H) data set in Fig. 2includes\na reference measurement performed at ambient pressure\non a more massive single crystal ( m∼4mg) and with-\nout the pressure cell. The ambient pressure data yield\na saturation moment Msof 3.05µB/Cr, which is con-\nsistent with the value found in our previous study [ 9].\nImportantly, the data in Fig. 2show that Msis pressure\nindependent up to about 2.8GPa within the experimen-\ntal error bar, the latter mostly being determined by the\ncomparably strong background signal due to the reduced\nmass of our sample. At higher pressure, however, Msre-\nduces to about 2.5 µB/Cr at 3.4GPa. Another important\nobservation is that, contrary to the saturation moment,\nthe saturationfield Hsatcontinuouslyshifts tohigher val-\nues under applied hydrostaticpressure. While saturation\ncan be achieved at ∼0.1T for ambient pressure, µ0Hsat\nincreases to ∼1.5T at 3.4GPa. A possible explanation\nfor the shift of the saturation field was proposed by Sun\net al.[10] through first principles calculations. Accord-\ning to them, the pressure-induced decrease in the Cr-Cr\nbond length favors antiferromagnetic exchange, while a\nconcomitant deviation from the 90◦Cr-Te-Cr bond angle\nleads to a suppression of the FM superexchange interac-\ntion. Increasing competition between the easy-axis MAE\nand the easy-plane shape anisotropy under pressure (see\nbelow) may also contribute to an increase of Hsat. Mag-\nnetization measurements at still higher pressures would\nbe enlightening in this regard.\nThe results of the above analysis of the M(T) and4\n/s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53 /s50/s46/s48 /s50/s46/s53 /s51/s46/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53/s51/s46/s48/s51/s46/s53\n/s32/s48/s32/s71/s80/s97\n/s32/s49/s46/s51/s32/s71/s80/s97\n/s32/s50/s46/s50/s32/s71/s80/s97\n/s32/s50/s46/s56/s32/s71/s80/s97\n/s32/s51/s46/s52/s32/s71/s80/s97/s77 /s32/s40/s109\n/s66/s47/s67/s114/s41\n/s109\n/s48/s72/s32/s40/s84/s41\nFIG. 2. Magnetization Mas a function of applied field H/bardblc\natT= 15K for pressures up to 3.4GPa. The solid lines are\nguides to the eye. A demagnetization correction has been\napplied [ 23].\nM(H) dependences are summarized in Fig. 3, where\nwe plot the MsandTcvalues as a function of the ap-\nplied pressure with the corresponding experimental error\nbars. As already noticed above, the ordering tempera-\nture value estimated from the first derivative criterion\nis less accurate for higher pressures due to the broaden-\ning of the transition under pressure (Fig. 1, inset). Still,\nwe could clearly validate the long-range ordered ferro-\nmagnetic ground state of Cr 2Ge2Te6at 4.2K and up to\n3.4GPa, i.e., for the parameter ranges where FMR ex-\nperiments were performed (see below). In contrast to Tc,\nthe FM saturation moment is approximatelyconstant for\npressures up to 2.8GPa with a subsequent reduction for\nP >2.8GPa.\nB. Pressure-dependent FMR study\nThe FMR data sets presented in this section were col-\nlectedbymeasuringthreedifferentsingle-crystallinesam-\nples of Cr 2Ge2Te6from the same batch with similar, but\nslightlydifferent, lateraldimensions. As will be discussed\nbelow, the differences in demagnetization factors associ-\nated with different sample dimensions are, however, neg-\nligible in the analysis of the FMR data.\nSelected FMR spectra measured for the two field ge-\nometries used in this work are shown in Figs. 4(a) and\n(b). A shift of the signal to higher fields for H/bardblc\nand to smaller fields for H⊥cby applying pressure is\nclearly visible. The results of frequency-dependent FMR\nmeasurements at a temperature of 4.2K with H/bardblc\nand at different applied external pressures between 0\nand 2.39GPa are summarized in Fig. 4(c). At 0GPa,\nCr2Ge2Te6featuresaneasy-axis-typeMAEwiththeeasy\nFIG. 3. Saturation magnetization Msand the FM transition\ntemperature Tcas a function of the applied pressure (sym-\nbols).Mswas calculated as the average magnetization value\nfor applied fields µ0H≥2T, with 2T determined as the\nlower limit of the applied field at which dM/dH = 0 for all\nthe curves within the experimental error bar. The solid line s\nare guides to the eye.\naxis oriented parallel to the caxis as was investigated in\ndetail in our previous study [ 9]. This kind of anisotropy\ndirectly manifests in an FMR measurement by a shift of\nthe resonance line towards smaller magnetic fields with\nrespect to the resonance position in the paramagnetic\nstate if the external magnetic field His applied parallel\nto the magnetic easy axis ( caxis) and towards higher\nmagnetic fields if His normal to the magnetic easy axis.\nThis behavior is illustrated in Figs. 4(a) and4(b), where,\ndepending on the direction of the applied field, the FMR\nsignal at P= 0 is shifted either to the left or to the\nright side of the signal from the paramagnetic reference\nsample 2,2-diphenyl-1-picrylhydrazyl (DPPH) with a g\nfactorg≃2, which is commonly used as a magnetic\nfield marker. Remarkably, the application of an exter-\nnal pressure counteracts these tendencies by shifting the\nFMR line towards the paramagnetic position for both\nfield geometries. As a result, with increasing the applied\npressure from P= 0 to 2.01GPa the FMR signal from\nthe sample becomes isotropic; that is, the resonance field\nHresdoes not depend on the direction of the applied field\nat this value of pressure [Fig. 5(b)].\nThe frequency-field dependence of the paramagnetic\nreference sample DPPH is shown in Fig. 4as a blue\ndashed line for comparison with the ferromagnetic res-\nonance fields Hresmeasured on the Cr 2Ge2Te6samples\nforH/bardblcaxis.Hresshifts continuously to higher mag-\nnetic fields with increasing Pand thereby approaches,\nbut does not cross, the paramagnetic resonance branch\nin the frequency-field diagram. This systematic shift is\nemphasized in the inset of Fig. 4, where details of the\nfrequency-field diagram are shown for intermediate field\nstrengths. The change in the FMR line position as a5\n/s53/s46/s50 /s53/s46/s52 /s53/s46/s54 /s53/s46/s56 /s52/s46/s48 /s52/s46/s53 /s53/s46/s48\n/s32/s32\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32/s40/s84/s41\n/s68/s80/s80/s72/s110 /s32/s61/s32/s49/s53/s48/s32/s71/s72/s122\n/s84 /s32/s61/s32/s52/s46/s50/s32/s75\n/s72 /s99\n/s48/s32/s71/s80/s97/s50/s46/s51/s57/s32/s71/s80/s97/s40/s98/s41/s32/s32/s65/s98/s115/s111/s114/s112/s116/s105/s111/s110/s32/s105/s110/s116/s101/s110/s115/s105/s116/s121/s32/s40/s97/s114/s98/s46/s32/s117/s110/s105/s116/s115/s41\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32/s40/s84/s41/s48\n/s68/s80/s80/s72/s110 /s32/s61/s32/s49/s51/s48/s32/s71/s72/s122\n/s84 /s32/s61/s32/s52/s46/s50/s32/s75\n/s72/s124/s124/s99\n/s50/s46/s51/s57/s32/s71/s80/s97/s40/s97/s41\n/s48 /s50 /s52 /s54 /s56 /s49/s48/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s50/s53/s48\n/s32/s68/s80/s80/s72/s58/s32 /s103 /s32/s61/s32/s50\n/s32/s48/s32/s71/s80/s97\n/s32/s48/s46/s53/s32/s71/s80/s97\n/s32/s49/s46/s48/s32/s71/s80/s97\n/s32/s49/s46/s53/s49/s32/s71/s80/s97\n/s32/s50/s46/s48/s49/s32/s71/s80/s97\n/s32/s50/s46/s51/s57/s32/s71/s80/s97/s32\n/s32/s110 /s32/s40/s71/s72/s122/s41\n/s109\n/s48/s32/s72\n/s114/s101/s115/s32/s40/s84/s41/s72 /s32/s124/s124/s32/s99/s44/s32 /s84 /s32/s61/s32/s52/s46/s50/s32/s75/s48/s32/s71/s80/s97/s48/s32/s71/s80/s97\n/s50/s46/s51/s57/s32/s71/s80/s97/s50/s46/s51/s57/s32/s71/s80/s97/s40/s99/s41\n/s52/s46/s48 /s52/s46/s53 /s53/s46/s48 /s53/s46/s53/s49/s49/s48/s49/s50/s48/s49/s51/s48/s49/s52/s48/s49/s53/s48/s49/s54/s48/s110 /s32/s40/s71/s72/s122/s41\n/s109\n/s48/s32/s72\n/s114/s101/s115/s32/s40/s84/s41\nFIG. 4. Representative FMR spectra of Cr 2Ge2Te6at zero\nand the maximum applied pressure for the applied magnetic\nfield (a) parallel and (b) perpendicular to the caxis. The\nsharp spike is a signal from the paramagnetic reference DPPH\nwith agfactor of g≃2. Some distortions of the lineshape\ncan be ascribed to the interference effects of the microwaves\nin the pressure cell [ 24,25]. (c) The frequency-field diagram\nobtained from FMR measurements at various pressures up to\n2.39GPa for H/bardblc. Solid lines are fits according to Eq. ( 2)\nwhile the dashed blue line indicates the paramagnetic ν(Hres)\ndependence of the DPPH reference sample with g≃2. De-\ntails of the frequency-field diagram for intermediate fields are\nshown in the inset in order to emphasize the systematic shift\nof the resonance line with increasing pressure.\nfunction of external pressure evidences a reduction of the\nMAE of the studied system. However, the shift of the\nresonance fields of Cr 2Ge2Te6is determined not only by\nthe strength and the sign of the MAE constant KUbut\nalso by demagnetization fields resulting from the specific\nsample shapes which favor, in the case of the platelike-\nshaped Cr 2Ge2Te6crystals, an in-plane magnetization.\nAsaconsequence,theinternaldemagnetizationfieldswill\nlead to a shift of the resonance position to higher fields\nif the external magnetic field is applied perpendicular to\nthe honeycomb layers, i.e., for H/bardblcaxis. Thus, the sizeof the anisotropy constant KUderived from shifts of the\nresonance position would be underestimated, if demag-\nnetization effects were not taken into account (see the\nrelated discussion in Ref. [ 9]). These demagnetization\nfieldsHDcan be calculated using the elements of the de-\nmagnetization tensor Ni(i=x,y,z, withx,y,zbeing\nthe principal axes of the tensor) [ 23,26,27]:\nHi\nD=−4πNiM , (1)\nwhereMdenotes the magnetization of the sample. Since\nthe thickness of the plate-like samples used in this study\ncould not be quantified precisely, the real sample shape\nwas approximated using the demagnetization factors for\na flat (infinite) plate oriented perpendicular to the zaxis\n(which was chosen to coincide with the crystallographic\ncaxis of Cr 2Ge2Te6), i.e.,Nx=Ny= 0,Nz= 1. This\nassumption is supported by the fact that the lateral di-\nmensions in the abplane of the used Cr 2Ge2Te6crystals\nwere much larger than the thicknesses of the samples\nparallel to the caxis. Using the value of the satura-\ntion magnetization Ms≃3µB/Cr (which corresponds to\nMs≈201 erg/Gcm3for Cr 2Ge2Te6) yields a demagne-\ntization field µ0HDin the fully saturated ferromagnetic\nphase of -0.253T if the external field is applied parallel\nto thecaxis. The frequency-field dependences measured\nwith this field orientation were then fitted by the follow-\ning expression:\nνres=gµBµ0\nh[Hres+HD+HA], (2)\nwhereνresdenotes the resonance/microwave frequency\nandHAis the anisotropy field describing the intrinsic\nmagnetocrystalline anisotropy. Here, µB,µ0andhare\nBohr magneton, magnetic permeability and Plancks con-\nstant, respectively. Fits of Eq. ( 2) to the measured data\nare shown in Fig. 4(c) as solid lines. For the fitting, the\ngfactor and HAwere treated as free fit parameters. The\ngfactor varied very little within the error bars around\nthe mean value of 2.03 (see, Sec. IV) whereas HAsig-\nnificantly decreased with increasing pressure. Using the\nlatter parameter, it is possible to calculate the uniaxial\nMAEKUaccording to (see, for instance, Ref. [ 28])\nKfit\nU=HAMS\n2. (3)\nThe results obtained from this fitting procedure are dis-\ncussed in the following section together with results from\nan alternative approach for a quantitative analysis of the\nFMR data.\nIV. SIMULATIONS OF THE FREQUENCY\nDEPENDENCE AND DETERMINATION OF\nTHE MAGNETIC ANISOTROPIES\nInordertosimultaneouslytakeintoaccountthe results\nof our FMR measurements performed with H/bardblcand6\n/s48 /s50 /s52 /s54 /s56 /s49/s48 /s48 /s50 /s52 /s54 /s56/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s50/s53/s48\n/s32/s72 /s32/s124/s124/s32/s99\n/s32/s72 /s32/s94 /s32/s99/s110 /s32/s40/s71/s72/s122/s41\n/s109\n/s48/s32/s72\n/s114/s101/s115/s32/s40/s84/s41/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s48/s32/s71/s80/s97/s44/s32 /s84/s32 /s61/s32/s52/s46/s50/s32/s75/s32\n/s75\n/s85/s32/s61/s32/s40/s52/s46/s56/s32 /s177 /s32/s48/s46/s51/s41 /s215/s49/s48/s53\n/s32/s101/s114/s103/s47/s99/s109/s51/s40/s97/s41 /s40/s98/s41\n/s32/s72 /s32/s124/s124/s32/s99\n/s32/s72 /s32/s94 /s32/s99\n/s109\n/s48/s32/s72\n/s114/s101/s115/s32/s40/s84/s41/s32/s32/s32/s32/s32/s32/s32/s32/s50/s46/s48/s49/s32/s71/s80/s97/s44/s32 /s84/s32 /s61/s32/s52/s46/s50/s32/s75\n/s75\n/s85/s32/s61/s32/s40/s50/s46/s55/s32 /s177 /s32/s48/s46/s52/s41 /s215/s49/s48/s53\n/s32/s101/s114/s103/s47/s99/s109/s51\nFIG. 5. Examples of simulations (solid lines) of the mea-\nsured frequency-fielddependences(symbols) at (a) 0GPa and\n(b) 2.01GPa. In these simulations experimental data sets ob -\ntained for H/bardblc(circles) and H⊥c(squares) were taken\ninto account simultaneously in order to determine the value\nofKU.\nH⊥cfield geometries, the measured frequency-field de-\npendences were numerically simulated for each pressure\nvalue. This lead to a refinement of the determination of\nKU, in particular at higher pressures. For the simula-\ntions we used a well-established phenomenological model\nof FMR (see, e.g., Refs. [ 28–30]) in which the resonance\nfrequency νresis expressed as\nν2\nres=g2µ2\nB\nh2M2ssin2θ/parenleftbigg∂2F\n∂θ2∂2F\n∂ϕ2−/parenleftBig∂2F\n∂θ∂ϕ/parenrightBig2/parenrightbigg\n.(4)\nHere,Fis the free-energy density, and θandϕare\nthe spherical coordinates of the magnetization vector\nM(Ms,ϕ,θ).\nThis phenomenological model is applicable for ferro-\nmagnets with fully saturated magnetization at tempera-\ntures sufficiently lower than Tc, i.e., the case of the fully\ndeveloped ferromagnetic phase. According to the above-\ndiscussed results of our pressure-dependent magnetiza-\ntion study, both criteria are safely fulfilled for the T,P,\nandHparameter ranges of the FMR study.\nTo obtain the resonance position of the FMR signal, F\nin Eq. (4) should be taken at the equilibrium angles ϕ0\nandθ0ofM. In the simulations, the minimum of Fwith\nrespect to θandϕfor a given set of experimental param-\neters, such as the microwave frequency and the direction\nand the strength of the magnetic field, was found numer-\nically. For Cr 2Ge2Te6, accounting for both the intrinsic\nuniaxial magnetic anisotropyand of the (extrinsic) shape\nanisotropy of the particular studied sample enabled an\naccurate description of all FMR data sets. In this casethe free energy density (in cgs units) is defined as\nF=−H·M−KUcos2(θ)+\n2πM2\ns(Nxsin2(θ)cos2(ϕ)+\nNysin2(θ)sin2(ϕ)+Nzcos2(θ))(5)\nand comprises, besides the Zeeman energy density ex-\npressed by the first term, contributions due to the uniax-\nial and shape anisotropies, the second and third terms,\nrespectively.\nRepresentative simulations of frequency-dependent\nmeasurements at 0 and 2.01GPa are shown in Figs. 5(a)\nand5(b), respectively. At 0GPa, the separation of\ntheν(Hres) curves obtained for the two different field\norientations clearly evidences the easy-axis-type mag-\nnetic anisotropy with the H/bardblccurve being shifted to\nsmaller fields and the H⊥ccurve being shifted to\nhigherfieldscomparedtotheparamagneticposition. The\n0GPa data could be simulated using the value of KU\nof 4.8×105erg/cm3obtained in our previous study [ 9].\nKeeping the value of KUfixed, the gfactor and demag-\nnetization factors were adjusted in order to consistently\ndescribe different studied samples. It turned out that\nall 0GPa data could be described successfully with an\nisotropic gfactor of 2.03, which is consistent with our\npreviousinvestigations, andthe sameset ofdemagnetiza-\ntion factors Nx=Ny= 0,Nz= 1 despite the (small) dif-\nferences in the shapes of the samples. These parameters\nwerekeptfixedin thesimulationsofthefrequencydepen-\ndences measured at several nonzero pressures. The only\nfree parameter in these simulations was the anisotropy\nconstant KUwhich allowed us to determine the pressure\ndependence of the MAE. As can be seen in Fig. 5(b),\nat a pressure of 2.01GPa, the measured frequency de-\npendences are nearly identical for both orientations of\nthe external magnetic field (a similar frequency-field di-\nagram was obtained for 2.39GPa), giving the impression\nof isotropic behavior. However, the contributions of the\nmagnetocrystalline anisotropy and the shape anisotropy\nto the free-energy density [Eq. ( 5)] have opposite signs\nwithin the pressure range of this study. Thus, the ap-\nparently isotropic frequency dependence is, in fact, a\nconsequence of the similar strengths of these two con-\ntributions. This allows us to conclude that an appli-\ncation of pressures between 2.01 and 2.39GPa reduces\nthe uniaxial MAE KUto a value that compensates the\nshape anisotropy of the crystals. With the demagnetiza-\ntion field µ0HD=−0.253T and saturation magnetiza-\ntionMs≈201 erg/Gcm3estimated above, one obtains\nthe shape anisotropy constant Kshape= (HDMs)/2≈\n−2.54×105erg/cm3. Its absolute value is indicated\nby the dashed line in Fig. 6.Moreover, the pressure-\ndependent FMR studies do not provide any evidence\nfor a pressure-induced sign change of the magnetocrys-\ntalline anisotropy or a vanishing of the intrinsic uniaxial\nanisotropy up to pressures of 2.39GPa. This is in con-\ntrastto the conclusionsdrawn in Ref. [ 11] on a spin reori-\nentation transition at pressures between 1.0 and 1.5GPa7\n/s48/s46/s53 /s49/s46/s48 /s49/s46/s53 /s50/s46/s48 /s50/s46/s53 /s48/s48/s49/s50/s51/s52/s53\n/s32/s115/s105/s109/s117/s108/s97/s116/s105/s111/s110\n/s32/s32/s102/s105/s116\n/s32/s124 /s75\n/s115/s104/s97/s112/s101/s124/s75\n/s85/s32/s40/s49/s48/s53\n/s32/s101/s114/s103/s47/s99/s109/s51\n/s41\n/s112 /s32/s40/s71/s80/s97/s41/s84 /s32/s61/s32/s52/s46/s50/s32/s75\nFIG. 6. Pressure dependence of the uniaxial magnetocrys-\ntalline anisotropy constant KUof Cr2Ge2Te6. Red circles de-\nnote the results obtained from fitting the different ν(Hres)\ncurves according to Eq. ( 2) forH/bardblc. Blue squares are\ntheKUvalues determined from simulations according to\nEq. (5) considering the experimental data sets for H/bardblc\nandH⊥csimultaneously. The horizontal dashed line shows\nthe absolute value of the negative shape anisotropy constan t\nKshape≈ −2.54×105erg/cm3. It compensates the decreas-\ning positive value of KUatp/greaterorsimilar2GPa yielding a seemingly\nisotropic behavior.\nbasedonmagneto-transportinvestigationsofCr 2Ge2Te6.\nThe pressure dependence of KUas obtained from fit-\nting ofthe H/bardblcdatato Eq.( 2) andfromthe simulations\nof all data available for both field orientations according\ntoEq.(5)isshownin Fig. 6. Qualitatively,both methods\nreveal a similar behavior, in particular a reduction of KU\nwithincreasingpressure. However,thedeviationbetween\ntheKUvalues derived from the fits and the simulation\nincreasesat higherpressures. This could be attributed to\nthe fact that the simulations simultaneously take into ac-\ncount the frequency dependences measured for both field\norientations. Therefore, the reliability of the KUvalues\nfrom simulations is higher despite the larger error bars,\nin particular at higher pressures at which differences be-\ntween both field orientations become very small. Finally,\nitisworthmentioningthatinRef.[ 31]apressure-induced\ndifference in the unit-cell volume of Cr 2Ge2Te6between\napplied pressures of 0 and 3GPa of about 4% was re-\nported. As the unit-cell volume enters into the calcu-\nlation of the saturation magnetization, such a reduction\nof volume with increasing pressure could change the ab-\nsolute value determined for KU. However, it was found\nthat the pressure-inducedcontractionofthe unit-cell vol-\nume by 4% does not affect the determination of KUbut\nis within the given error bars.V. CONCLUSIONS AND OUTLOOK\nOur combined magnetization and ferromagnetic res-\nonance study on single crystals of the quasi-two-\ndimensional ferromagnet Cr 2Ge2Te6has revealed a sig-\nnificanteffect ofthe applicationofhydrostaticpressure P\non the properties of the ferromagnetic state of this com-\npound. It manifests in a reduction of the critical temper-\natureTcand a concomitant broadening of the transition\nto the FM state, as well as in a gradual increase of the\nsaturation field Hsat. In contrast, the saturation magne-\ntization remains practically constant up to P= 2.8GPa\nand reduces by approaching the highest applied pres-\nsure of 3.4GPa. The frequency-dependent FMR mea-\nsurements performed in the fully developed FM state at\nT≪TcandH > H satat various external pressures up\nto 2.39GPa revealed a systematic shift of the resonance\nfields with increasing Pevidencing a continuous reduc-\ntion of the magnetocrystalline anisotropy constant KU.\nFrom the quantitative analysis of the data, which takes\ninto account demagnetization effects, it follows that KU,\nalthough being significantly reduced at the highest pres-\nsures applied in this study, neither vanishes nor changes\nits sign up to pressures of about 2.4GPa. Therefore,\nnoting that FMR is a very direct method for the quan-\ntitative determination of magnetic anisotropy, it can be\nconcluded with confidence that there is no evidence for\na switching of the magnetic anisotropy from the easy-\naxis to the easy-plane type within the pressure range un-\nderconsideration. Furtherpressure-dependentstudieson\nCr2Ge2Te6are desired to understand the apparent dis-\ncrepancy between our work and the magnetotransport\nstudy in Ref. [ 11] claiming a spin-reorientation transi-\ntion in Cr 2Ge2Te6at pressures 1 < P <1.5GPa and to\naddress the question of whether a pressure-induced sign\nswitching of the magnetic anisotropy could be achieved\nat pressures exceeding 2.39GPa.\nAlthough in terms of the spatial dimensionality\nwe investigated bulk, three-dimensional crystals of\nCr2Ge2Te6, we showedin our previouselectronspin reso-\nnance/FMR study that the low-temperature magnetism\nof the bulk crystals is essentially of a two-dimensional\n(2D) nature due to very weak interlayer magnetic cou-\npling in this van der Waals compound [ 9]. Consider-\ning the intrinsic magnetic two-dimensionality of the bulk\nmaterial, our pressure dependent study could be rele-\nvant for the research on Cr 2Ge2Te6in the truly 2D spa-\ntial limit. In this respect, the pressure control of the\nmagnetic anisotropy investigated in our work may pro-\nvide important hints for a targeted design of functional\nmagneto-electrical heterostructures containing layers of\nCr2Ge2Te6.\nIn particular, a recent prediction of remarkable strain\nand electric field tunability of a single layer Cr 2Ge2Te6\nis encouraging [ 32]. It remains yet an open question\nwhether the design of a heterostructure with strain ǫ\nof the order of ±(1−2)% for the Cr 2Ge2Te6layer, as\nproposed in Ref. [ 32], can be achieved. However, a siz-8\nable effect on the magnetic anisotropywhich we observed\nat hydrostatic pressures up to 2.39GPa corresponds to\nan even smaller compressive in-plane strain ǫ∼0.7%\n[31]. Such straining seems plausible to achieve since, in\ngeneral, 2D van der Waals materials are known to sus-\ntain large strain. For example, it was shown that biaxial\ncompressive and tensile strain of ∼1% can be achieved\nin single-layer molybdenum dichalcogenide deposited on\na thermally compressed or expanded polymer substrate\n[33].\nACKNOWLEDGMENTS\nWe thank Gael Bastien and Randirley Beltran Ro-\ndriguez for technical assistance with the pressure cell\nused for the magnetization studies, and the IFW work-\nshop and Juliane Scheiter for the production and treat-ment of the CuBe gaskets. The work in Kobe was par-\ntially supported by Grants-in-Aid for Scientific Research\n(C) (Grant No. 19K03746) from the Japan Society for\nthe Promotion of Science. The work in Dresden was sup-\nported by the Deutsche Forschungsgemeinschaft (DFG)\nthrough Grant No. KA1694/12-1 and within Collabora-\ntive Research Center SFB 1143“CorrelatedMagnetism –\nFrom Frustration to Topology” (Project No. 247310070)\nand the Dresden-W¨ urzburg Cluster of Excellence (EXC\n2147) “ct.qmat - Complexity and Topology in Quantum\nMatter” (Project No. 390858490). S.A. acknowledges fi-\nnancial support from the DFG through Grant No. AS\n523/4-1. A.A. acknowledges financial support form the\nDFG through Grant No. AL 1771/4-1. S.S. acknowl-\nedges financial support from the GRK-1621 Graduate\nAcademy. V.K. gratefully acknowledges the hospitality\nand financial support during his research visit to Kobe\nUniversity.\n[1]V Carteaux, D Brunet, G Ouvrard, and G Andre,\n“Crystallographic, magnetic and electronic structures\nof a new layered ferromagnetic compound Cr 2Ge2Te6,”\nJ. Phys.: Condens. Matter 7, 69 (1995) .\n[2]J.-G. Park, “Opportunities and challenges of 2D mag-\nnetic van der Waals materials: magnetic graphene?”\nJ. Phys.: Condens. Matter 28, 301001 (2016) .\n[3]Cheng Gong, Lin Li, Zhenglu Li, Huiwen Ji, Alex Stern,\nYang Xia, Ting Cao, Wei Bao, Chenzhe Wang, Yuan\nWang, Z. Q. Qiu, R. J. Cava, Steven G. Louie, Jing\nXia, and Xiang Zhang, “Discovery of intrinsic ferro-\nmagnetism in two-dimensional van der Waals crystals,”\nNature546, 265 (2017) .\n[4]Kenneth S. Burch, David Mandrus, and Je-Geun Park,\n“Magnetism in two-dimensional van der Waals materi-\nals,”Nature563, 47 (2018) .\n[5]M. Gibertini, M. Koperski, A. F. Morpurgo, and K. S.\nNovoselov, “Magnetic 2D materials and heterostruc-\ntures,”Nat. Nanotechnol. 14, 408 – 419 (2019) .\n[6]Xiao Zhang, Yuelei Zhao, Qi Song, Shuang Jia, Jing\nShi, and Wei Han, “Magnetic anisotropy of the\nsingle-crystalline ferromagnetic insulator Cr 2Ge2Te6,”\nJpn. J. Appl. Phys. 55, 033001 (2016) .\n[7]G. T. Lin, H. L. Zhuang, X. Luo, B. J. Liu, F. C.\nChen, J. Yan, Y. Sun, J. Zhou, W. J. Lu, P. Tong,\nZ. G. Sheng, Z. Qu, W. H. Song, X. B. Zhu, and\nY. P. Sun, “Tricritical behavior of the two-dimensional\nintrinsically ferromagnetic semiconductor CrGeTe 3,”\nPhys. Rev. B 95, 245212 (2017) .\n[8]Yu Liu and C. Petrovic, “Critical behavior of quasi-two-\ndimensional semiconducting ferromagnet Cr 2Ge2Te6,”\nPhys. Rev. B 96, 054406 (2017) .\n[9]J. Zeisner, A. Alfonsov, S. Selter, S. Aswartham, M. P.\nGhimire, M. Richter, J. van den Brink, B. B¨ uchner, and\nV. Kataev, “Magnetic anisotropy and spin-polarized two-\ndimensional electron gas in the van der Waals ferromag-\nnet Cr 2Ge2Te6,”Phys. Rev. B 99, 165109 (2019) .\n[10]Y. Sun, R. C. Xiao, G. T. Lin, R. R. Zhang, L. S. Ling,\nZ. W. Ma, X. Luo, W. J. Lu, Y. P. Sun, and Z. G.\nSheng, “Effects of hydrostatic pressure on spin-latticecoupling in two-dimensional ferromagnetic Cr 2Ge2Te6,”\nAppl. Phys. Lett. 112, 072409 (2018) .\n[11]Zhisheng Lin, Mark Lohmann, Zulfikhar A. Ali, Chi\nTang, Junxue Li, Wenyu Xing, Jiangnan Zhong, Shuang\nJia, Wei Han, Sinisa Coh, Ward Beyermann, and\nJing Shi, “Pressure-induced spin reorientation transi-\ntion in layered ferromagnetic insulator Cr 2Ge2Te6,”\nPhys. Rev. Mater. 2, 051004 (2018) .\n[12]S. Selter, G. Bastien, A. U. B. Wolter, S. Aswartham,\nandB.B¨ uchner,“Magnetic anisotropyandlow-fieldmag-\nnetic phase diagram of the quasi-two-dimensional ferro-\nmagnet Cr 2Ge2Te6,”Phys. Rev. B 101, 014440 (2020) .\n[13]P. L. Alireza and G. G. Lonzarich, “Miniature anvil cell\nfor high-pressure measurements in a commercial super-\nconducting quantum interference device magnetometer,”\nRev. Sci. Inst. 80, 023906 (2009) .\n[14]W. Schottenhamel, Aufbau eines hochaufl¨ osenden\nDilatometers und einer hydrostatischen SQUID-\nDruckzelle sowie Untersuchungen an korrelierten\n¨Ubergangsmetalloxiden , Ph.D. thesis, Technische Univer-\nsit¨ at Dresden (2016).\n[15]B Bireckoven and J Wittig, “A diamond anvil cell for\nthe investigation of superconductivity under pressures of\nup to 50 GPa: Pb as a low temperature manometer,”\nJ. Phys. E: Sci. Instrum. 21, 841–848 (1988) .\n[16]G. Prando, R. Dally, W. Schottenhamel, Z. Guguchia, S.-\nH. Baek, R. Aeschlimann, A. U. B. Wolter, S. D. Wilson,\nB. B¨ uchner, and M. J. Graf, “Influence of hydrostatic\npressure on the bulk magnetic properties of Eu 2Ir2O7,”\nPhys. Rev. B 93, 104422 (2016) .\n[17]T. Sakurai, K. Fujimoto, R. Matsui, K. Kawasaki,\nS. Okubo, H. Ohta, K. Matsubayashi, Y. Uwatoko, and\nH. Tanaka, “Development of multi-frequency ESR sys-\ntem for high-pressure measurements up to 2.5 GPa,”\nJ. Magn. Reson. 259, 108 – 113 (2015) .\n[18]The general pressure medium Daphne7373 ensures a hy-\ndrostatic pressure and gives good pressure homogeneity\nto the sample up to about 2.3GPa. Therefore, it is suffi-\ncientfor anFMRmeasurementwithamaximumpressure\nof 2.4GPa. The special pressure medium Daphne 7575 is9\nnecessary only for the magnetization measurement where\na pressure above 3GPa was applied in order to secure\nthe pressure homogeneity in the high-pressure range. See\nRefs. [19,20] for details.\n[19]Keizo Murata, Keiichi Yokogawa, Harukazu Yoshino,\nStefan Klotz, Pascal Munsch, Akinori Irizawa, Mo-\ntotsugu Nishiyama, Kenzo Iizuka, Takao Nanba,\nTahei Okada, Yoshitaka Shiraga, and Shoji\nAoyama, “Pressure transmitting medium Daphne\n7474 solidifying at 3.7 GPa at room temperature,”\nRev. Sci. Inst. 79, 085101 (2008) .\n[20]Keizo Murata and Shinji Aoki, “Development of high\npressure liquid medium with good hydrostatic pres-\nsure,”Rev. High. Pressure Sci. Techol. 26, 3–7 (2016) ,\n(in Japanese with abstract in English ).\n[21]We did not use the other frequently applied method, the\nArrot-plot technique [ 22], since it is not practicable for\npressure dependent studies due to the reduced resolution\nstemming from the background signal of the pressure cell\n(see Sec. II).\n[22]I. Yeung, R. M. Roshko, and G. Williams, “Arrott-\nplot criterion for ferromagnetism in disordered systems,”\nPhys. Rev. B 34, 3456–3457 (1986) .\n[23]J. A. Osborn, “Demagnetizing factors of the general el-\nlipsoid,” Phys. Rev. 67, 351–357 (1945) .\n[24]Tominimize theundesiredinterferenceeffectsinthepres-\nsure cell and, in particular, to avoid the interference\ninside the sample we followed the recipe suggested in\nRef. [25] and used very thin crystals with a thickness\nof a few tens µm which is much smaller than the short-\nest microwave wavelength of 1153 µm at the highest ap-\nplied frequency of 260GHz. Furthermore, we repeated\neach experiment several times to confirm that the reso-\nnance fields agree within the error [smaller than the sym-\nbol size in Fig. 4(c)]. This ensured that the intrinsic res-\nonance position was always correctly obtained regardless\nthe distortion.[25]G Eilers, M von Ortenberg, and R Galazka, “Origin of\nSatellite Structures of High-Field EPR in Cd 1−xMnxTe,”\nInt. J. Infrared Millimeter Waves 15, 695–722 (1994) .\n[26]D. C. Cronemeyer, “Demagnetization factors for general\nelipsoids,” J. Appl. Phys. 70, 2911 (1991) .\n[27]Stephen Blundell, Magnetism in Condensed Matter (Ox-\nford University Press, Oxford, 2001).\n[28]G. V. Skrotskii and L. V. Kurbatov, “Phenomenologi-\ncal theory of ferromagnetic resonance,” in Ferromagnetic\nResonance , edited by S. V. Vonsovskii (Pergamon Press\nLtd., 1966) Chap. Phenomenological Theory of Ferro-\nmagnetic Resonance, pp. 12–77.\n[29]J. SmitandH.G. Beljers, “Ferromagnetic ResonanceAb-\nsorption inBaFe 12O19,”Philips Res.Rep. 10,113(1955).\n[30]Michael Farle, “Ferromagnetic resonance of ultrathin\nmetallic layers,” Rep. Prog. Phys. 61, 755 (1998) .\n[31]Zhenhai Yu, Wei Xia, Kailang Xu, Ming Xu,\nHongyuan Wang, Xia Wang, Na Yu, Zhiqiang Zou,\nJinggeng Zhao, Lin Wang, Xiangshui Miao, and Yan-\nfeng Guo, “Pressure-Induced Structural Phase Tran-\nsition and a Special Amorphization Phase of Two-\nDimensional Ferromagnetic Semiconductor Cr 2Ge2Te6,”\nJ. Phys. Chem. C 123, 13885–13891 (2019) .\n[32]Kangying Wang, Tao Hu, Fanhao Jia, Guodong\nZhao, Yuyu Liu, Igor V. Solovyev, Alexander P.\nPyatakov, Anatoly K. Zvezdin, and Wei Ren,\n“Magnetic and electronic properties of Cr 2Ge2Te6\nmonolayer by strain and electric-field engineering,”\nApplied Physics Letters 114, 092405 (2019) .\n[33]Riccardo Frisenda, Matthias Drueppel, Robert Schmidt,\nSteffen Michaelis de Vasconcellos, David Perez de lara,\nRudolf Bratschitsch, Michael Rohlfing, and Andres\nCastellanos-Gomez, “Biaxial strain tuning of the opti-\ncal properties of single-layer transition metal dichalco-\ngenides,” Npj 2D Mater Appl 1, 10 (2017) ." }, { "title": "2012.06693v1.Electrically_Controllable_Crystal_Chirality_Magneto_Optical_Effects_in_Collinear_Antiferromagnets.pdf", "content": "Electrically Controllable Crystal Chirality Magneto-Optical Effects in Collinear Antiferromagnets\nXiaodong Zhou,1Wanxiang Feng,1,\u0003Xiuxian Yang,1Guang-Yu Guo,2, 3and Yugui Yao1,y\n1Key Laboratory of Advanced Optoelectronic Quantum Architecture and Measurement,\nMinistry of Education, School of Physics, Beijing Institute of Technology, Beijing 100081, China\n2Department of Physics and Center for Theoretical Physics, National Taiwan University, Taipei 10617, Taiwan\n3Physics Division, National Center for Theoretical Sciences, Hsinchu 30013, Taiwan\n(Dated: December 15, 2020)\nThe spin chirality, created by magnetic atoms, has been comprehensively understood to generate and control\nthe magneto-optical effects. In comparison, the role of the crystal chirality that relates to nonmagnetic atoms\nhas received much less attention. Here, we theoretically discover the crystal chirality magneto-optical (CCMO)\neffects, which depend on the chirality of crystal structures that originates from the rearrangement of nonmagnetic\natoms. We show that the CCMO effects exist in many collinear antiferromagnets, such as RuO 2and CoNb 3S6,\nwhich has a local and global crystal chirality, respectively. The key character of the CCMO effects is the sign\nchange if the crystal chirality reverses. The magnitudes of the CCMO spectra can be effectively manipulated by\nreorienting the N ´eel vector with the help of an external electric field, and the spectral integrals are found to be\nproportional to magnetocrystalline anisotropy energy.\nAmong large family members of the magneto-optical (MO)\neffects, the Kerr [1] and Faraday [2] effects discovered in the\nmid-18th century are the representatives, which describe the\nrotations of the planes of polarization of the linearly polarized\nlight reflecting from and transmitting through magnetic me-\ndia, respectively. After they were discovered over a century,\na critical theoretical explanation of them was revealed on the\nbasis of the band theory [3]. Since then, two basic understand-\nings for the MO effects were broadly accepted [4–8]: The\nmagnitudes of MO spectra are proportional to the spontaneous\nmagnetization of magnetic media such that antiferromagnets\n(AFMs) with zero net magnetization are intuitively believed\nto present no MO signals; The simultaneous presence of band\nexchange splitting (due to finite net magnetization) and spin-\norbit coupling is the sole physical origin of the MO effects.\nNevertheless, new insights into the MO effects emerged re-\ncently. On the one hand, the MO effects were surprisingly\nfound in noncollinear [9–14] or collinear [15–18] AFMs even\nthough their net magnetization is vanishing. The MO Kerr\neffect was first predicted to exist in noncollinear (coplanar)\nAFMs Mn 3X(X=Rh, Ir, Pt) [9] by the natural lack of a\ngood symmetryTS(Tis the time-reversal symmetry; Sis a\nspatial symmetry) that is responsible for band exchange split-\nting. The collinear AFMs (e.g., MnPSe 3) were then demon-\nstrated to host the Kerr effect if an electric field is introduced\nto break theTSsymmetry [15]. On the other hand, the topo-\nlogical MO effects that originate from the scalar spin chiral-\nity was discovered in noncollinear (noncoplanar) AFMs [19].\nThe band exchange splitting and spin-orbit coupling are not\nindispensable for the topological MO effects and their quanti-\nzation [19], in sharp contrast to the conventional MO effects.\nThe above discoveries can be mainly ascribed to chiral spin\ntextures created by magnetic atoms. However, the role of non-\nmagnetic atoms that is related to crystal chirality has not been\ncompletely understood till now. In this work, using the first-\nprinciples calculations and group theory analyses, we reveal\n\u0003wxfeng@bit.edu.cn\nyygyao@bit.edu.cna new class of MO effects in collinear AFMs, termed crystal\nchirality magneto-optical (CCMO) effects. The MO Kerr and\nFaraday spectra change their signs when the crystal structure’s\nchirality alters from the left-handed state to the right-handed\nstate or vice versa. Moreover, the magnitudes of MO spectra\ncan be effectively manipulated by reorienting the N ´eel vec-\ntor with the help of an external electric field. The spectral\nintegrals are found to be proportional to magnetocrystalline\nanisotropy energy (MAE), which has not been previously re-\nported in AFMs. The CCMO effects discovered here may\nbe a promising optical means to simultaneously detect crys-\ntal structure’s chirality and N ´eel vector’s orientation.\nThe crystal structure’s chirality is generated from differ-\nent arrangements of nonmagnetic atoms by keeping the struc-\nture’s space group unchanged. The crystal chirality manifests\nin many magnetic materials with different dimensions, e.g.,\nthree-dimensional (3D) rutile RuO 2[20] andMF2(M=Mn,\nNi) [21–23], quasi-2D NNb3S6(N=V , Cr, Mn, Fe, Co,\nNi) [20, 24, 25], and 2D SrRuO 3monolayer [26]. Here, we\ntake RuO 2and CoNb 3S6as two prototypes, which has a local\nand global crystal chirality, respectively.\nRutile RuO 2has long been considered to be a Pauli param-\nagnet [27], while several recent studies revealed a collinear\nantiferromagnetic order under room temperature [28–30].\nRuO 2is a centrosymmetric system (crystallographic space\ngroupP42=mnm ) with inversion centers at magnetic Ru\natoms, therefore, it can not be regarded as a usual chiral\ncrystal [31, 32]. However, one can still define a local crys-\ntal chirality for RuO 2[20], given by \u001f=dRu1O\u0002dORu2\n(dRu1O anddORu2 are two vectors connecting the path Ru1-O-\nRu2), because nonmagnetic O atoms are not inversion centers.\nEvidently, RuO 2has two opposite chiral states, i.e., right-\nhanded chirality \u001f= +1 [Fig. 1(a)] and left-handed chiral-\nity\u001f=\u00001[Fig. 1(b)]. The two chiral states are energeti-\ncally degenerate, but their magnetic densities are redistributed\nby rotating 90 degrees. Interestingly, one chiral state is re-\nlated to the other one via a combined symmetry Tt1=2, where\nt1=2= [0:5;0:5;0:5]is the half-unit cell translation. For each\nchiral state, the N ´eel vector,n= (nRu1\u0000nRu2)=2(nRu1and\nnRu2are the spins of two Ru atoms), can be reoriented byarXiv:2012.06693v1 [cond-mat.mtrl-sci] 12 Dec 20202\nFIG. 1. (a)(b) The crystal structures of RuO 2with the right-handed ( \u001f= +1 ) and left-handed ( \u001f=\u00001) chiralities. Pink and cyan spheres\nrepresent two magnetic Ru atoms (Ru1 and Ru2), whereas yellow spheres are nonmagnetic O atoms. The dashed lines outline the octahedron\nformed by six O atoms. The magnetic density isosurfaces surrounding two Ru atoms differ by 90 degrees. Pink and cyan arrows label the\nspin orientations of two Ru atoms, and the N ´eel vector is schematically plotted along the [110] direction. Two chiral structures are related\nto each other by the symmetry Tt1=2. (c)(d) Real and imaginary parts of the diagonal element and (e)(f) real and imaginary parts of the\noff-diagonal element of optical conductivity for \u001f=\u00061states of RuO 2when the N ´eel vectornrotates within the (001) plane ( \u0012= 90\u000eand\n0\u000e6'690\u000e). The inset in (c) shows the spherical coordinates ( \u0012,') used for describing the N ´eel vectorn. (g) The momentum-resolved\n\u001b00\nyzfor the\u001f= +1 state when'= 0\u000e,45\u000e, and 90\u000e, at the incident photon energy of 2.31 eV that is marked by an arrow in (f).\nthe current-induced spin-orbit field, like in CuMnAs [33, 34].\nOur first-principles calculations [35] show that the N ´eel vec-\ntor of RuO 2points along the [001] direction with a MAE of\n2.76 meV/Ru atom, which accords well with the previous the-\noretical results [20, 28]. While the resonant X-ray scattering\ndata indicate that the N ´eel vector slightly cants from the [001]\ndirection and has nonzero [100] and [010] components [29].\nSuch a generic magnetization direction suggests that the N ´eel\nvector can be easily tuned, giving rise to the coupling between\nthe crystal chirality \u001fand N ´eel ordern.\nThe calculation of optical conductivity is the crucial step\nto obtain the MO Kerr and Faraday spectra [35]. For conve-\nnience, the vector-form notation of optical Hall conductivity,\n\u001b= [\u001byz;\u001bzx;\u001bxy] = [\u001bx;\u001by;\u001bz], is used here. Since \u001bcan\nbe regarded as a pseudovector, like the spin, its nonvanishing\ncomponents can be identified by acting on each group element\nof relevant magnetic groups [14]. And the results of symmetry\nanalyses for\u001bcan be directly applied to the Kerr and Faraday\nangles,\u001eK= [\u001ex\nK;\u001ey\nK;\u001ez\nK]and\u001eF= [\u001ex\nF;\u001ey\nF;\u001ez\nF](see\nEqs. (S6) and (S7) in Ref. [35]).\nFirst, we prejudge the nonvanishing components of \u001bby\nconsidering the N ´eel vectornlying on the (100), (010), and\n(001) planes. When npoints along the [001] direction, RuO 2\nhas a magnetic space group P420=mnm0, which contains two\nglide mirror planes M[100]t1=2andM[010]t1=2. The mir-\nrorM[100] changes the signs of \u001byand\u001bz, but preserves\n\u001bx; the half-unit cell translation t1=2plays a nothing role on\n\u001b; thus, the combined operation M[100]t1=2restricts opti-cal Hall conductivity to be a shape of \u001b= [\u001bx;0;0]. Sim-\nilarly, another glide mirror plane M[010]t1=2gives rise to\n\u001b= [0;\u001by;0]. Therefore, under the group P420=mnm0,\nthe optical Hall conductivity is zero, \u001b= [0;0;0]. When\nndeviates away from the [001] direction and rotations within\nthe (100) and (010) planes, the symmetries M[010]t1=2and\nM[100]t1=2are absent, giving rise to [\u001bx;0;0]and[0;\u001by;0],\nrespectively. Within the (001) plane, the magnetic space group\nalways contains the symmetry TM [001], which changes the\nsign of\u001bzbut preserves \u001bxand\u001by. Therefore, \u001bxand\u001by\nare potentially nonzero. In particular, if npoints along the\n[100] ([010]) direction, only \u001by(\u001bx) can be nonzero due to\nthe role ofM[010]t1=2(M[100]t1=2). To summarize, once n\ncants slightly from the [001] direction, the optical Hall con-\nductivity and the MO spectra turn to be nonzero.\nFrom the symmetry point of view, the effect of crystal chi-\nrality on\u001bdiffers from that of spin chirality. In noncollinear\nAFMs Mn 3ZN (Z=Ga, Zn, Ag, Ni) [14], the reversal of\nspin chirality changes nonvanishing components and magni-\ntudes of\u001b. While two crystal chirality states of RuO 2are\nrelated by the symmetry Tt1=2, only the sign of \u001bwill be\nchanged as\u001bis odd underT. Thus, the sign change can also\nbe achieved by reversing the N ´eel vectorn. In fact, the sign\nof\u001bfor RuO 2is determined by the sign of n\u0001\u001f.\nNext, we discuss the optical conductivity calculated from\nthe first-principles methods. In light of a recently experimen-\ntal fabrication of the [001]-orientated RuO 2thin films with\nin-plane magnetization [30], we pay our attention to the n3\nlying within the (001) plane. Figs. 1(c) and 1(d) display the\ndiagonal element of optical conductivity, \u001b0= (\u001byy+\u001bzz)=2,\nfor two chiral states of RuO 2as a function of '. The real part\n\u001b0\n0measures the average in the absorption of left- and right-\ncircularly polarized light. It has two absorptive peaks at 1.5\neV and 3.5 eV and diverges in the low-frequency region due\nto the inclusion of Drude term [35]. The imaginary part \u001b00\n0\ncan be obtained from \u001b0\n0by using Kramer-Kronig transforma-\ntion [36]. Figs. 1(c) and 1(d) shows that \u001b0is robust against\nboth the crystal chirality and N ´eel vector.\nIn contrast, the off-diagonal elements of optical conductiv-\nity are substantially affected by the crystal chirality and N ´eel\nvector. Ifnrotates within the (001) plane, two off-diagonal\nelements,\u001byzand\u001bzx, are nonzero. The evolution of \u001byz\nwith\u001fandnis plotted in Figs. 1(e) and 1(f), whereas the\nresults of\u001bzxare shown in Figs. S1(c) and S1(d) [35]. The\nsigns of\u001byzand\u001bzxare changed for \u001f= +1 and\u00001states\nrelated by the symmetry Tt1=2. Ifnpoints along the [100]\ndirection (\u0012= 90\u000eand'= 0\u000e),\u001byzis zero for both two\nchiral states due to the symmetry M[010]t1=2. By increasing\n'from 0\u000eto90\u000e, the magnitude of \u001byzgradually increases\nfor both two chiral states, which can be well accounted for the\nmomentum-resolved optical Hall conductivity [Fig. 1(g)].\nNow, we proceed to the CCMO effects for collinear AFM\nRuO 2, as depicted in Fig. 2. The Kerr and Faraday spectra\n(\u001ex\nK=#x\nK+i\"x\nKand\u001ex\nF=#x\nF+i\"x\nF) exhibit a sim-\nilar profile to the off-diagonal elements of optical conduc-\ntivity, only differing by a minus [37]. The reason is simple\nas the off-diagonal elements of optical conductivity dominate\nthe shape of MO spectra, while the diagonal elements medi-\nate the amplitude of MO spectra (see Eqs. (S6) and (S7) in\nRef. [35]). The reversal of the crystal chirality changes the\nsigns of the Kerr and Faraday spectra but retains their mag-\nnitudes. If '= 0\u000e,#x\nK;F and\"x\nK;F are zero due to the\nvanishing\u001bx(=\u001byz)[see Figs. 1(e) and 1(f)]. In the range\nof0\u000e< '690\u000e, the magnitudes of #x\nK;F and\"x\nK;F in-\ncrease monotonously with the increasing of '. An opposite\ntrend appears to #y\nK;F and\"y\nK;F (see Fig. S2). It indicates\nthat the magnitudes of CCMO effects can be effectively tuned\nby changing magnetization direction in collinear AFMs. The\nlargest Kerr and Faraday rotation angles of RuO 2are 0.62 deg\nand 2.42\u0002105deg/cm, respectively. Particularly, the Kerr ro-\ntation angle is larger than that of traditional ferromagnets, e.g.,\nbcc Fe (\u00180.6 deg) [38], hcp Co ( \u00180.48 deg) [39], and fcc Ni\n(\u00180.15 deg) [40], and is also larger than that of famous non-\ncollinear AFMs, e.g., Mn 3X(\u00180.6 deg) [9], Mn 3Y(Y=\nGe, Ga, Sn) (\u00180.02 deg) [10], and Mn 3ZN (\u00180.42 deg) [14].\nThe large CCMO effects discovered in RuO 2suggests a novel\ncollinear antiferromagnetic platform for possible applications\nin MO recording beyond transitional ferromagnets [5].\nSince the CCMO spectra are frequency-dependent quanti-\nties, additional information can be captured from their spec-\ntral integrals (SIs), defined as SI\r\nK=R1\n0+#\r\nK(!)d!and\nSI\r\nF=R1\n0+#\r\nF(!)d!with\r=fx;y;zg. Figs. 3(a) and 3(b)\nshow that when the N ´eel vectornrotates within the (100)\nplane, SIx\nKand SIx\nFare nonvanishing and their signs are oppo-\nsite for two chiral states, which can be well understood from\nthe symmetry requirements of #x\nKand#x\nF. Moreover, SIx\nK\nFIG. 2. (a)(b) Kerr rotation angle and ellipticity and (c)(d) Faraday\nrotation angle and ellipticity for the left- and right-handed chirality\nstates (\u001f=\u00061) of RuO 2when the N ´eel vectornrotates within the\n(001) plane ( \u0012= 90\u000eand0\u000e6'690\u000e). The arrows in (a) and (c)\nmark the maximums of the Kerr and Faraday rotation angles at the\nphoton energies of 2.31 and 3.53 eV , respectively.\nand SIx\nFexhibit a period of 2\u0019with\u0012, which is simply due to\n#x\nKand#x\nFare odd under time-reversal symmetry. The MAE\nof two chiral states are equivalent and have a discrete two-fold\ndegeneracy, MAE (\u0012) = MAE (\u0012+\u0019). The most interesting\nfinding is that the absolute values of SIs are proportional to\nthe MAE, i.e.,jSIx\nK;Fj/MAE, which has not been reported\nin AFMs. For the (010) plane, the nonvanishing SIy\nK;F dis-\nplay the same behaviors [Figs. 3(c) and 3(d)]. For the (001)\nplane, both SIx\nK;Fand SIy\nK;Fare nonvanishing and the MAE\nhas a period of\u0019\n2[Figs. 3(e) and 3(f)]. Nevertheless, the SIs\nare still proportional to the MAE if the SIs on the (001) plane\nare redefined as SI0\nK=R1\n0+f[#x\nK(!)]2+[#y\nK(!)]2g1=2d!and\nSI0\nF=R1\n0+f[#x\nF(!)]2+[#y\nF(!)]2g1=2d![Figs. 3(g) and 3(h)].\nFig. 3 reveals that the SIs are closely related to the MAE for\ncollinear AFMs, and the sign and size of SIs can be used to\nidentify the crystal structure’s chirality and the N ´eel vector’s\ndirection, respectively.\nFinally, we demonstrate that the CCMO effects exist also in\ncollinear AFMs with a global crystal chirality, e.g., CoNb 3S6.\nThe crystallographic space group of CoNb 3S6isP6322,\nwhich is compatible with chiral crystal structures due to the\nlacking of improper symmetry operations [31, 32]. As shown\nin Figs. 4(a) and 4(b), the left- and right-handed chiral struc-\ntures are related to each other by a mirror plane Mx, which\npreserves the positions of magnetic atoms and the orientations\nof spin magnetic moments but rearranges the nonmagnetic S\natoms. CoNb 3S6forms a collinear antiferromagnetic order\nbelow the N ´eel temperature of \u001825 K [41–43]. The mag-\nnetic space group of CoNb 3S6isC202021if two ferromagnetic\nCo layers magnetized along the [100] direction are antifer-\nromagnetically coupled along the [001] direction via a NbS 2\nlayer. The operation C2zt1=2z(t1=2z= [0;0;0:5]) forces the\ncomponents of optical Hall conductivity that are perpendic-\nular to theC2axis to be zero, and therefore gives rise to\n\u001b= [0;0;\u001bz]. As a result, only the zcomponents of Kerr4\nFIG. 3. The SIs (colored symbols and lines, left axes) and MAE (shaded regions, right axes) for the left- and right-handed chirality states\n(\u001f=\u00061) of RuO 2when the N ´eel vectornrotates within the (100), (010), and (001) planes. For the MAE on the (100)/(010) and (001)\nplanes, the total energies when npoints to the [001] and [100] directions are set to be the reference states, respectively. The MAE on the (001)\nplane are multiplied by a factor of 3. The SIs shown in (g) and (h) are subtracted by the ones when npoints to the [100] direction.\nFIG. 4. (a)(b) The crystal structures of CoNb 3S6with the right-\nhanded (\u001f= +1 ) and left-handed ( \u001f=\u00001) crystal chiralities. The\nviolet, green, and bronze spheres represent Co, Nb, and S atoms, re-\nspectively. The red arrows label the spin orientations of magnetic Co\natoms (nCo1andnCo2). The N ´eel vector,n= (nCo1\u0000nCo2)=2, is\nalong the [100] direction. (c)(d) Kerr and Faraday rotation angles for\nthe left- and right-handed chirality states of CoNb 3S6.and Faraday rotation angles, #z\nKand#z\nF, are nonvanishing, as\nshown in Figs. 4(c) and 4(d). It is obvious that the MO effects\nfor CoNb 3S6are chiral in the sense that the signs of #z\nKand\n#z\nFare reserved for the left- and right-handed crystal struc-\ntures. The reason is simple because the mirror symmetry Mx\nthat relates two chiral crystal structures changes the signs of\n#z\nKand#z\nF. The CCMO effects uncovered in CoNb 3S6are\ncharacterized by the sign change resulting from the reversal\nof crystal chirality, similarly to the scenario of RuO 2.\nIn summary, using the first-principles calculations together\nwith group theory analyses, we have demonstrated the electri-\ncally controllable CCMO effects in collinear AFMs. The sign\nand magnitude of CCMO spectra can be tuned by the crys-\ntal chirality and spin orientation, respectively. Moreover, the\nspectral integrals are found to be proportional to the magne-\ntocrystalline anisotropy energy. The electrically controllable\nCCMO effects may provide a powerful probe of crystal chi-\nrality and spin orientation and be a promising MO recording\ntechnique based on antiferromagnetic materials.\nW.F., X.Z, and Y .Y . acknowledge the support from the Na-\ntional Key R&D Program of China (Nos. 2020YFA0308800\nand 2016YFA0300600), the National Natural Science Foun-\ndation of China (Nos. 11874085 and 11734003), and the\nGraduate Technological Innovation Project of Beijing Insti-\ntute of Technology (No. 2019CX10018). G.-Y . Guo acknowl-\nedges the support from the Ministry of Science and Technol-\nogy, National Center for Theoretical Sciences, and the Far\nEastern Y . Z. Hsu Science and Technology Memorial Foun-\ndation in Taiwan.\n[1] J. Kerr, On Rotation of the Plane of Polarization by Reflection\nfrom the Pole of a Magnet, Phil. Mag. 3, 321 (1877).[2] M. Faraday, I. Experimental Researches in Electricity.–\nNineteenth Series, Phil. Trans. R. Soc. Lond. 136, 1 (1846).5\n[3] P. N. Argyres, Theory of the Faraday and Kerr Effects in Ferro-\nmagnetics, Phys. Rev. 97, 334 (1955).\n[4] W. Reim and J. Schoenes, Handbook of Magnetic Materials,\nedited by E. P. Wohlfarth and K. H. J. Buschow, V ol. 5 (Elsevier,\nNew York, 1990) Chap. 2.\n[5] M. Mansuripur, The Physical Principles of Magneto-optical\nRecording (Cambridge University, New York, 1995).\n[6] H. Ebert, Magneto-Optical Effects in Transition Metal Systems,\nRep. Prog. Phys. 59, 1665 (1996).\n[7] V . Antonov, B. Harmon, and A. Yaresko, Electronic Structure\nand Magneto-Optical Properties of Solids (Kluwer Academic\nPublishers, Dordrecht, 2004) Chap. 1.4.\n[8] W. Kuch, R. Sch ¨afer, P. Fischer, and F. U. Hillebrecht, Mag-\nnetic Microscopy of Layered Structures (Springer-Verlag Berlin\nHeidelberg, 2015) Chap. 2.\n[9] W. Feng, G.-Y . Guo, J. Zhou, Y . Yao, and Q. Niu, Large\nMagneto-Optical Kerr Effect in Noncollinear Antiferromagnets\nMn3X(X= Rh;Ir;Pt), Phys. Rev. B 92, 144426 (2015).\n[10] T. Higo, H. Man, D. B. Gopman, L. Wu, T. Koretsune, O. M. J.\nvan’t Erve, Y . P. Kabanov, D. Rees, Y . Li, M.-T. Suzuki,\nS. Patankar, M. Ikhlas, C. L. Chien, R. Arita, R. D. Shull,\nJ. Orenstein, and S. Nakatsuji, Large Magneto-Optical Kerr\nEffect and Imaging of Magnetic Octupole Domains in an Anti-\nferromagnetic Metal, Nat. Photonics 12, 73 (2018).\n[11] A. L. Balk, N. H. Sung, S. M. Thomas, P. F. S. Rosa, R. D.\nMcDonald, J. D. Thompson, E. D. Bauer, F. Ronning, and\nS. A. Crooker, Comparing the Anomalous Hall Effect and the\nMagneto-Optical Kerr Effect through Antiferromagnetic Phase\nTransitions in Mn 3Sn, Appl. Phys. Lett. 114, 032401 (2019).\n[12] M. Wu, H. Isshiki, T. Chen, T. Higo, S. Nakatsuji, and Y . Otani,\nMagneto-Optical Kerr Effect in a Non-Collinear Antiferromag-\nnet Mn 3Ge, Appl. Phys. Lett. 116, 132408 (2020).\n[13] S. Wimmer, S. Mankovsky, J. Min ´ar, A. N. Yaresko, and\nH. Ebert, Magneto-Optic and Transverse-Transport Properties\nof Noncollinear Antiferromagnets, Phys. Rev. B 100, 214429\n(2019).\n[14] X. Zhou, J.-P. Hanke, W. Feng, F. Li, G.-Y . Guo, Y . Yao,\nS. Bl ¨ugel, and Y . Mokrousov, Spin-order Dependent Anoma-\nlous Hall Effect and Magneto-Optical Effect in the Non-\ncollinear Antiferromagnets Mn3XNwithX= Ga , Zn, Ag,\nor Ni, Phys. Rev. B 99, 104428 (2019).\n[15] N. Sivadas, S. Okamoto, and D. Xiao, Gate-Controllable\nMagneto-Optic Kerr Effect in Layered Collinear Antiferromag-\nnets, Phys. Rev. Lett. 117, 267203 (2016).\n[16] F.-R. Fan, H. Wu, D. Nabok, S. Hu, W. Ren, C. Draxl, and\nA. Stroppa, Electric-Magneto-Optical Kerr Effect in a Hybrid\nOrganic–Inorganic Perovskite, J. Am. Chem. Soc. 139, 12883\n(2017).\n[17] K. Yang, W. Hu, H. Wu, M.-H. Whangbo, P. G. Radaelli,\nand A. Stroppa, Magneto-Optical Kerr Switching Properties of\n(CrI 3)2and (CrBr 3/CrI 3) Bilayers, ACS Appl. Electron. Mater.\n2, 1373 (2020).\n[18] A. De, T. K. Bhowmick, and R. Lake, Anomalous Magneto\nOptic Effects from an Antiferromagnet Topological-Insulator\nHeterostructure, arXiv:2006.02752 .\n[19] W. Feng, J.-P. Hanke, X. Zhou, G.-Y . Guo, S. Bl ¨ugel,\nY . Mokrousov, and Y . Yao, Topological Magneto-Optical Ef-\nfects and Their Quantization in Noncoplanar Antiferromagnets,\nNat. Commun. 11, 118 (2020).\n[20] L. ˇSmejkal, R. Gonz ´alez-Hern ´andez, T. Jungwirth, and\nJ. Sinova, Crystal Time-Reversal Symmetry Breaking and\nSpontaneous Hall Effect in Collinear Antiferromagnets, Sci.\nAdv. 6, eaaz8809 (2020).\n[21] S. M. Wu, W. Zhang, A. KC, P. Borisov, J. E. Pearson, J. S.Jiang, D. Lederman, A. Hoffmann, and A. Bhattacharya,\nAntiferromagnetic Spin Seebeck Effect, Phys. Rev. Lett. 116,\n097204 (2016).\n[22] L.-D. Yuan, Z. Wang, J.-W. Luo, E. I. Rashba, and A. Zunger,\nGiant Momentum-Dependent Spin Splitting in Centrosymmet-\nric Low-ZAntiferromagnets, Phys. Rev. B 102, 014422 (2020).\n[23] X. Li, A. H. MacDonald, and H. Chen, Quantum\nAnomalous Hall Effect through Canted Antiferromagnetism,\narXiv:1902.10650 .\n[24] K. Anzenhofer, J. V . D. Berg, P. Cossee, and J. Helle, The\nCrystal Structure and Magnetic Susceptibilities of MnNb 3S6,\nFeNb 3S6, CoNb 3S6and NiNb 3S6, J. Phys. Chem. Solids. 31,\n1057 (1970).\n[25] S. S. P. Parkin and R. H. Friend, 3 dTransition-Metal Interca-\nlates of the Niobium and Tantalum Dichalcogenides. I. Mag-\nnetic Properties, Philos. Mag. B 41, 65 (1980).\n[26] K. Samanta, M. Le ˇzai´c, M. Merte, F. Freimuth, S. Bl ¨ugel, and\nY . Mokrousov, Crystal Hall and Crystal Magneto-Optical Effect\nin Thin Films of SrRuO 3, J. Appl. Phys. 127, 213904 (2020).\n[27] W. D. Ryden and A. W. Lawson, Magnetic Susceptibility of\nIrO2and RuO 2, J. Chem. Phys 52, 6058 (1970).\n[28] T. Berlijn, P. C. Snijders, O. Delaire, H.-D. Zhou, T. A. Maier,\nH.-B. Cao, S.-X. Chi, M. Matsuda, Y . Wang, M. R. Koehler,\nP. R. C. Kent, and H. H. Weitering, Itinerant Antiferromag-\nnetism in RuO 2, Phys. Rev. Lett. 118, 077201 (2017).\n[29] Z. H. Zhu, J. Strempfer, R. R. Rao, C. A. Occhialini, J. Pelli-\nciari, Y . Choi, T. Kawaguchi, H. You, J. F. Mitchell, Y . Shao-\nHorn, and R. Comin, Anomalous Antiferromagnetism in\nMetallic RuO 2Determined by Resonant X-ray Scattering,\nPhys. Rev. Lett. 122, 017202 (2019).\n[30] Z. Feng, X. Zhou, L. ˇSmejkal, L. Wu, Z. Zhu, H. Guo,\nR. Gonz ´alez-Hern ´andez, X. Wang, H. Yan, P. Qin, et al. , Obser-\nvation of the Crystal Hall Effect in a Collinear Antiferromagnet,\narXiv: 2002.08712 .\n[31] H. Flack, Chiral and Achiral Crystal Structures, Helv. Chim.\nActa 86, 905 (2003).\n[32] M. Nespolo, M. I. Aroyo, and B. Souvignier, Crystallographic\nShelves: Space-Group Hierarchy Explained, J. Appl. Cryst. 51,\n1481 (2018).\n[33] P. Wadley, B. Howells, J. ˇZelezn ´y, C. Andrews, V . Hills,\nR. P. Campion, V . Nov ´ak, K. Olejn ´ık, F. Maccherozzi, S. S.\nDhesi, S. Y . Martin, T. Wagner, J. Wunderlich, F. Freimuth,\nY . Mokrousov, J. Kune ˇs, J. S. Chauhan, M. J. Grzybowski,\nA. W. Rushforth, K. W. Edmonds, B. L. Gallagher, and T. Jung-\nwirth, Electrical Switching of an Antiferromagnet, Science 351,\n587 (2016).\n[34] J. Godinho, H. Reichlov ´a, D. Kriegner, V . Nov ´ak, K. Olejn ´ık,\nZ. Ka ˇspar, Z. ˇSob´aˇn, P. Wadley, R. P. Campion, R. M. Otxoa,\nP. E. Roy, J. ˇZelezn ´y, T. Jungwirth, and J. Wunderlich, Electri-\ncally Induced and Detected N ´eel Vector Reversal in a Collinear\nAntiferromagnet, Nat. Commun. 9, 4686 (2018).\n[35] See Supplemental Material at http://link.aps.org/xxx, which in-\ncludes a detailed description of the calculation methods and\nsupplemental figures.\n[36] H. S. Bennett and E. A. Stern, Faraday Effect in Solids, Phys.\nRev.137, A448 (1965).\n[37] W. Feng, G.-Y . Guo, and Y . Yao, Tunable Magneto-Optical\nEffects in Hole-Doped Group-IIIA Metal-Monochalcogenide\nMonolayers, 2D Mater. 4, 015017 (2017).\n[38] G. Y . Guo and H. Ebert, Band-Theoretical Investigation of the\nMagneto-Optical Kerr Effect in Fe and Co Multilayers, Phys.\nRev. B 51, 12633 (1995).\n[39] G. Y . Guo and H. Ebert, Theoretical Investigation of The Orien-\ntation Dependence of The Magneto-Optical Kerr Effect in Co,6\nPhys. Rev. B 50, 10377 (1994).\n[40] P. van Engen, K. Buschow, and M. Erman, Magnetic Properties\nand Magneto-Optical Spectroscopy of Heusler Alloys Based\non Transition Metals and Sn, J. Magn. Magn. Mater. 30, 374\n(1983).\n[41] S. S. P. Parkin, E. A. Marseglia, and P. J. Brown, Magnetic\nStructure of Co 1=3NbS 2and Co 1=3TaS 2, J. Phys. C: Solid State\nPhys. 16, 2765 (1983).[42] N. J. Ghimire, A. S. Botana, J. S. Jiang, J. Zhang, Y .-S.\nChen, and J. F. Mitchell, Large anomalous Hall effect in the\nchiral-lattice antiferromagnet CoNb 3S6, Nat. Commun. 9, 3280\n(2018).\n[43] G. Tenasini, E. Martino, N. Ubrig, N. J. Ghimire, H. Berger,\nO. Zaharko, F. Wu, J. F. Mitchell, I. Martin, L. Forr ´o, and\nA. F. Morpurgo, Giant Anomalous Hall Effect in Quasi-Two-\nDimensional Layered Antiferromagnet Co1=3NbS 2, Phys. Rev.\nResearch 2, 023051 (2020)." }, { "title": "2012.11469v1.Magnetic_order_of_Dy___3____and_Fe___3____moments_in_antiferromagnetic_DyFeO___3___probed_by_spin_Hall_magnetoresistance_and_spin_Seebeck_effect.pdf", "content": "Magnetic order of Dy3+and Fe3+moments in antiferromagnetic DyFeO 3probed by\nspin Hall magnetoresistance and spin Seebeck e\u000bect\nG. R. Hoogeboom1, T. Kuschel2, G.E.W. Bauer1;3, M. V. Mostovoy1, A. V. Kimel4and B. J. van Wees1\n1Physics of Nanodevices, Zernike Institute for Advanced Materials,\nUniversity of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands.\n2Center for Spinelectronic Materials and Devices, Department of Physics,\nBielefeld University, Universit atsstra \fe 25, 33615 Bielefeld, Germany.\n3AIMR & Institute for Materials Research, Tohoku University, Aoba-ku, Katahira 2-1-1, Sendai, Japan\n4Spectroscopy of Solids and Interfaces, Institute of Molecules and Materials,\nRadboud University Nijmegen, Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands\n(Dated: December 22, 2020)\nWe report on spin Hall magnetoresistance (SMR) and spin Seebeck e\u000bect (SSE) in single\ncrystal of the rare-earth antiferromagnet DyFeO 3with a thin Pt \flm contact. The angular\nshape and symmetry of the SMR at elevated temperatures re\rect the antiferromagnetic order\nof the Fe3+moments as governed by the Zeeman energy, the magnetocrystalline anisotropy\nand the Dzyaloshinskii-Moriya interaction. We interpret the observed linear dependence of the\nsignal on the magnetic \feld strength as evidence for \feld-induced order of the Dy3+moments\nup to room temperature. At and below the Morin temperature of 50 K, the SMR monitors the\nspin-reorientation phase transition of Fe3+spins. Below 23 K, additional features emerge that\npersist below 4 K, the ordering temperature of the Dy3+magnetic sublattice. We conclude that\nthe combination of SMR and SSE is a simple and e\u000ecient tool to study spin reorientation phase\ntransitions and sublattice magnetizations.\nI. INTRODUCTION\nAntiferromagnets (AFMs) form an abundant class of\nmaterials that o\u000ber many advantages over ferromagnets\n(FMs) for applications in high-density magnetic logics\nand data storage devices. AFMs support high-frequency\ndynamics in the THz regime that allows faster writing of\nmagnetic bits compared to FMs. The absence of mag-\nnetic stray \felds minimizes on-chip cross-talk and allows\ndownsizing devices that are robust against magnetic per-\nturbations [1]. On the other hand, most magnetic detec-\ntion methods observe only the FM order. Recent devel-\nopments in the detection [2] and manipulation [3{5] of\nthe AFM order reveal its many opportunities.\nThe AFM DyFeO 3(DFO) belongs to a family of\nrare-earth transition metal oxides called orthoferrites\nthat display many unusual phenomena such as weak\nferromagnetism (WFM), spin-reorientation transitions,\nstrong magnetostriction, multiferroicity including a large\nlinear magnetoelectric e\u000bect [6]. Their magnetic prop-\nerties are governed by the spin and orbital momenta of\n4f rare-earth ions coupled to the magnetic moment of 3d\ntransition metal ions.\nThe magnetization of dielectrics can be detected elec-\ntrically by the spin Hall magnetoresistance (SMR) in\nheavy metal contacts with a large spin Hall angle such\nas Pt [7]. This phenomenon is sensitive to FM, but also\nAFM spin order [2, 8{10]. With a Pt contact, information\nabout AFMs can be also retrieved by the spin Seebeck\ne\u000bect (SSE) under a temperature gradient [11{13].\nHere, we track the \feld-dependence of the coupled\nDy3+and Fe3+magnetic order as a function of tem-perature by both SMR and SSE. A su\u000eciently strong\nmagnetic \feld in the abplane of DFO forces the N\u0013 eel\nvector to follow a complex path out of the abplane. A\ntheoretical spin model explains the observations in terms\nof Fe3+spin rotations that are governed by the compe-\ntition between the magnetic anisotropy, Zeeman energy,\nand Dzyaloshinskii-Moriya interaction (DMI). The Dy3+\nmoments are disordered at room temperature but nev-\nertheless a\u000bect the magnitude of the SMR. At the so-\ncalled Morin phase transition at ∼50 K the Fe3+spins\nrotate by 90○, causing a step-like anomaly in the SMR.\nAt even lower temperatures, we observe two separate fea-\ntures tentatively assigned to the re-orientation of Fe3+\nspins in an applied magnetic \feld and another related to\nthe ordering of Dy3+orbital moments. Below 23 K, the\nSMR signal is ∼1%, 1-2 orders of magnitude larger than\nreported for other materials [2, 7]. Both Fe3+and Dy3+\nmoments appear to contribute to the SSE; a magnetic\n\feld orders the Dy3+moments and suppresses the Fe3+\ncontribution. The complex SMR and SSE is evidence of\na coupling between the Fe3+and Dy3+magnetic subsys-\ntems.\nThe paper is organized as follows. In Section II we\nreview the magnetic and multiferroic properties of DFO.\nThe theory of the magnetic probing methods are dis-\ncussed in Sec. III with Subsec. III A the SMR and\nSubsec. III B the SSE. In Subsec. IV A, the fabrica-\ntion, characterization and measurement techniques are\nexplained. Further, a model including the DMI, Zeeman\nenergy and magnetic anisotropy is employed in Subsec.\nIV B. The SMR results at elevated temperatures includ-\ning the model \fts as well as SMR and SSE results at lowarXiv:2012.11469v1 [cond-mat.mtrl-sci] 21 Dec 20202\nDy3+\nFe3+\nO2-\n500/uni03BCm\nTi/Au (5/50nm)\nHy\nxa) b)\nPt (8nm)\naab b\ndevice 1 device 2\nzDyFeO/three.denominator (1mm)\nFIG. 1. (a) DFO crystal unit cell. The blue, red and\nwhite spheres represent Dy3+, Fe3+and O2−ions, respectively.\n(b) Optical image of the Pt Hall bar on top of the bulk DFO\ncrystal. The lines indicate voltage probes, AC source and an-\ngle\u000bof the external magnetic \feld H. In the two devices,\nthe crystallographic directions aandbare rotated by 45○in\nthexyplane, the reference frame of the Hall bar.\ntemperatures are described and discussed in Sec. V.\nII. MAGNETIC AND MULTIFERROIC\nPROPERTIES OF DFO\nDFO is a perovskite with an orthorhombic ( D16\n2h-\nPbnm) crystallographic structure. It consists of alter-\nnating Fe3+and Dy3+abplanes, in which the Fe ions\nare located inside O2−octahedrons (Fig. 1a)). The\nlarge Dy3+magnetic moments ( J=15/slash.left2) order at a\nlow temperature, TDy\nN=4 K. The high N\u0013 eel tempera-\ntureTFe\nN=645 K indicates strong inter- and intra-plane\nAFM Heisenberg superexchange between the Fe3+mag-\nnetic moments ( S=5/slash.left2). The AFM order of the Fe mo-\nments is of the G-type with N\u0013 eel vector G(anti)parallel\nto the crystallographic aaxis (\u0000 4symmetry [14]). The\nbroken inversion symmetry enables a DMI [15, 16] that\nin the \u0000 4-phase causes a WFM mWFM/parallel.alt1cby the small\n(∼0:5○) canting of the Fe spins [14].\nA \frst-order Morin transition from the WFM \u0000 4-phase\nto the purely AFM \u0000 1-phase occurs when lowering the\ntemperature below 50 K. At this transition, the direction\nof the magnetic easy axis abruptly changes from the a-\nto theb-direction. A magnetic \feld higher than a critical\nmagnetic \feld, Hcr, along the caxis re-orients the N\u0013 eel\nvector back to the aaxis and recovers the \u0000 4-phase.\nBelowTDy\nN, the Dy3+moments form a noncollinear\nIsing-like AFM order with Ising axes rotated by ±33○\nfrom thebaxis [17] that corresponds to a G′\naA′\nbstate in\nBertaut's notation [18]. The simultaneous presence of\nordered Fe and Dy magnetic moments breaks inversion\nsymmetry and, under an applied magnetic \feld, induces\nan electric polarization [19] by exchange striction that\ncouples the Fe and Dy magnetic sublattices [6, 20].\nHigher magnetic \felds destroy the AFM order of the\nDy3+moments and thereby the electric polarization [21].Spins in this material can be controlled by light\nthrough the inverse Faraday e\u000bect [3], as well as by tem-\nperature and magnetic \feld. Re-orientation of the Fe mo-\nments has been studied by magnetometry [22], Faraday\nrotation [23], M ossbauer spectroscopy [24] and neutron\nscattering measurements [21]. The Morin transition at\n50 K causes large changes in the speci\fc heat [25] and\nentropy [26].\nIII. PROBING METHODS\nA. Spin Hall magnetoresistance\nThe SMR is caused by the spin-charge conversion in\na thin heavy metal layer in contact with a magnet [27].\nThe spin Hall e\u000bect induces a spin current transverse to\nan applied charge current and thereby an electron spin\naccumulation at surfaces and interfaces. Upon re\rection\nat the interface to a magnetic insulator, electrons expe-\nrience an exchange interaction that depends on the an-\ngle between their spin polarization and that of the in-\nterface magnetic moments, while the latter can be con-\ntrolled by an applied magnetic \feld. The re\rected spin\ncurrent is transformed back into an observable charge\ncurrent by the inverse spin Hall e\u000bect. The interface ex-\nchange interaction is parameterized by the complex spin\nmixing conductance. The result is a modulation of the\ncharge transport that depends on the orientation of the\napplied current and the interface magnetic order. In a\nHall bar geometry, this a\u000bects the longitudinal resistance\nand causes a planar Hall e\u000bect, i.e. a Hall voltage even\nwhen the magnetic \feld lies in the transport plane.\nSMR is a powerful tool to investigate the magnetic or-\ndering at the interface of collinear [7, 27{29] and non-\ncollinear ferrimagnets [30, 31] as well as spin spirals\n[32, 33]. Recently, a \\negative\" SMR has been discovered\nfor AFMs [2, 8{10], i.e. an SMR with a 90○phase shift\nof the angular dependence as compared to FMs, which\nshows that the AFM N\u0013 eel vector Gtends to align itself\nnormal to the applied magnetic \feld. The observable in\nAFMs is therefore the N\u0013 eel vector rather than the net\nmagnetization [2].\nThe longitudinal and transverse electrical resistivities\n\u001aLand\u001aTof Pt on an AFM read [2]\n\u001aL=\u001a+\u0001\u001a0+\u0001\u001a1(1−G2\ny) (1)\n\u001aT=\u0001\u001a1GxGy+\u0001\u001a2mz+\u0001\u001aHallHz (2)\nwithGiandHiwithi∈{x;y;z}as the Cartesian com-\nponents of the (unit) N\u0013 eel and the applied magnetic \feld\nvectors, respectively. mzis the out-of-plane (OOP) com-\nponent of the unit vector in the direction of the WFM\nmagnetization. \u0001 \u001a0is an angle-independent interface\ncorrection to the bulk resistivity \u001a. \u0001\u001aHallHzis the ordi-\nnary Hall resistivity of Pt in the presence of an OOP com-\nponent of the magnetic \feld. \u0001 \u001a1(\u0001\u001a2)is proportional3\nto the real (imaginary) part of the interface spin-mixing\nconductance. \u0001 \u001a2is a resistance induced by the e\u000bective\nWFM \feld, believed to be small in most circumstances.\nThe interface Dy3+moments can contribute to the\nSMR when ordered. Below TDy\nN, the Dy3+moments are\nAFM aligned with N\u0013 eel vector GDy. AboveTDy\nNand\nin su\u000eciently large applied magnetic \felds, the Dy3+\nmoments contribute to the SMR in Eqs. (1,2) after\nreplacing the N\u0013 eel vector GDyby the (nearly perpen-\ndicular) magnetization mDy. Disregarding magnetic\nanisotropy and DMI for the moment, the spin mixing\nconductance term \u0001 \u001a1mDy\nxmDy\nyphase-shifts the SMR\nby 90○relative to the pure AFM contribution. The\nterm \u0001\u001a2mzchanges sign with mzand its contribution\n∼Hzcannot be distinguished from the ordinary Hall\ne\u000bect \u0001\u001aHallHzin Pt. We remove a linear magnetic\n\feld dependence from the OOP SMR measurements.\nResidual non-linear e\u000bects from \u0001 \u001a2mzmay persist, but\nshould be small in the \u0000 4phase. A \fnite \u0001 \u001a2mzhas\nbeen reported in conducting AFMs [34], but we do not\nobserve a signi\fcant contribution down to 60 K.\nB. Spin Seebeck e\u000bect\nA heat current in a FM excites a spin current that in\ninsulators is carried mainly by magnons, the quanta of\nthe spin wave excitations of the magnetic order. We can\ngenerate a temperature bias simply by the Joule heat-\ning of a charge current in a metal contact. A magnon\n\rowjmcan also be generated by a gradient of a magnon\naccumulation or chemical potential \u0016m[35]. Therefore\njm=−\u001bm(∇\u0016m+SS∇T) (3)\nwith\u001bmas the magnon spin conductivity and SSthe spin\nSeebeck coe\u000ecient. Thermal magnons can typically dif-\nfuse over several \u0016m [36{38], which implies that the SSE\nmainly probes bulk rather than interface magnetic prop-\nerties. The magnons in simple AFMs typically come in\ndegenerate pairs with opposite polarization that split un-\nder an applied magnetic \feld [11, 39]. The associated im-\nbalance of the magnon populations cause a non-zero spin\nSeebeck e\u000bect [13]. Paramagnets display a \feld-induced\nSSE e\u000bect [38] for the same reason, so aligned Dy3+mo-\nments can contribute to an SSE in DFO. A magnon ac-\ncumulation at the interface to Pt injects a spin current js\nthat can be observed as an inverse spin Hall e\u000bect voltage\nVISHE=\u001a\u0012SH(js×\u001b), where\u0012SHis the spin Hall angle\nand\u001bis the spin polarization. The SMR and SSE can\nbe measured simultaneously by a lock-in technique [40].\nρT\na) b)α (deg)\nGaGc\nGb αρSMR (arbitrary units)\n-45 45 0 -90 902T6TFIG. 2. N\u0013 eel vector, G=(Ga;Gb;Gc)with/divides.alt0G/divides.alt0=1, calculated\nas a function of the magnetic \feld in the abplane. The angle\n\u000b∈[−90○;90○]as de\fned in Fig. 1(b) is coded by the colored\nbar.G(\u000b)minimizes the free energy Eq. (4), for Kb=0:15 K\nper Fe ion and H=6 T (other parameters are given in the\ntext). (b) The transverse SMR (arbitrary units) due to the\nmagnetic Fe sublattice for H=6 T, i.e. the G(\u000b)from panel\n(a) (thick red line), and for H=2 T (thin blue line).\nIV. METHODS\nA. Fabrication, characterization and measurements\nWe con\frmed the crystallographic direction of our sin-\ngle crystal by X-ray di\u000braction before sawing it into slices\nalong the abplane and polishing them. Two devices\nwere fabricated on di\u000berent slices of the materials using\na three step electron beam lithography process; markers\nwere created to align the devices along two di\u000berent crys-\ntallographic directions. After fabrication of an 8 nm thick\nPt Hall bar, 50 nm Ti/Au contact pads were deposited.\nThe angular dependence of the magnetoresistance be-\nlow 50 K is complex and hysteretic. Phase changes are\nassociated by internal strains that can cause cracks in\nthe bulk crystal. We therefore carried out magnetic \feld\nsweeps at low temperatures very slowly, with a waiting\ntime of 60 seconds between each \feld step. The response\nwas measured with a 1 mA (100 \u0016A) AC current through\nthe Pt Hall bar in device 1 (device 2) with a frequency of\n7.777 Hz. The \frst and second harmonic transverse and\nlongitudinal lock-in voltages as measured with a super-\nconducting magnet in a cryostat with variable tempera-\nture insert are the SMR and SSE e\u000bects, respectively.\nBelow the transition temperature, the Morin tran-\nsition is induced by a magnetic \feld along the caxis\nthat rotates the N\u0013 eel vector from atob. For device\n1, this does not change the transverse resistance since\nGFe\nxGFe\ny=0 when the N\u0013 eel vector is in either the x- or\ny-direction. On the other hand, device 2 is optimized\nfor the observation of the Morin transition, because,\nas discussed below, the transverse resistance should be\nmaximally positive when G/parallel.alt1band maximally negative\nwhen G/parallel.alt1a.4\nB. Modelling the SMR of Pt /divides.alt0DFO\nThe orientation of the N\u0013 eel vector Gof the Fe sublat-\ntice at temperatures well above TDy\nNis governed by sev-\neral competing interactions: (a) the magnetic anisotropy,\nwhich above the Morin transition favors G/parallel.alt1a, (b) the\nZeeman energy that favors G⊥Hsince the transverse\nmagnetic susceptibility of an AFM is higher than the lon-\ngitudinal one, and (c) the coupling of the WFM moment,\nmWFM/parallel.alt1a, to the applied magnetic \feld. This compe-\ntition can be described phenomenologically by the free\nenergy density\nf=Kb\n2G2\nb+Kc\n2G2\nc+\u001f⊥\n2/bracketleft.alt1(G⋅H)2−H2/bracketright.alt−mWFMGcHa;\n(4)\nwith the \frst two terms describing the second-order mag-\nnetic anisotropy with magnetic easy, intermediate and\nhard axes along the a,bandccrystallographic direc-\ntions, respectively ( Kc>Kb>Ka=0),\u001f⊥is the trans-\nverse magnetic susceptibility, and the mWFM is the weak\nferromagnetic moment along the aaxis, induced by G/parallel.alt1c.\n/divides.alt0G/divides.alt0=1, because the longitudinal susceptibility of the Fe\nspins is very small for T/uni226ATFe\nN. The magnetic \feld His\nchosen parallel to the abplane, but Gcan have an OOP\ncomponent Gc≠0 since the third term in Eq.(4) couples\nGclinearly toHa. For the SMR at 250 K, we may dis-\nregard higher-order magnetic anisotropies that become\nimportant near the Morin transition.\nAt weak magnetic \felds, the magnetic anisotropy pins\nthe N\u0013 eel vector to the aaxis. When the Zeeman energy\nbecomes comparable with the anisotropy energy, the ro-\ntation of the magnetic \feld vector in the abplane gives\nrise to a concomitant rotation of G. In the absence of\nmagnetic anisotropy, the canting of the magnetic mo-\nments leads to G⊥Hfor any magnetic \feld orienta-\ntion due to the Zeeman energy rendering a sinusoidal\nSMR, but magnetic anisotropy can distort the angular\ndependence. This behavior is further complicated by the\nWFM: for strong magnetic \felds along the aaxis, the\nN\u0013 eel vector tilts away from the abplane towards the c\naxis, since the c-component of Ginduces a WFM mo-\nment parallel to the applied magnetic \feld [24, 41]. By\ncontrast,Gbdoes not give rise to a weak FM moment, so\nthe N\u0013 eel vector returns into the abplane when we rotate\nthe magnetic \feld away from the aaxis. The equilibrium\nN\u0013 eel vector minimizes the free energy Eq. (4) under the\nconstraint /divides.alt0G/divides.alt0=1 as a function of strength and orienta-\ntion of the magnetic \feld with in-plane (IP) angle \u000b(see\nFig. 1b)).\nWe adopt a weak magnetization parameters mWFM=\n0:133\u0016Bper Fe3+ion induced either by G/parallel.alt1calong the\naaxis [42] or by G/parallel.alt1aalong the caxis [43]. The\ntransverse magnetic susceptibility can be estimated us-\ning the Heisenberg model with an Fe-Fe exchange con-\nstantJ1=4:23 meV for Y 3Fe5O12[44] , which leads to\n\u001f⊥=\u00162\nB/slash.left(3J1), which does not depend strongly on the\nrare-earth ion. Kcgoverns the critical \feld when applied\nalong theaaxis with\u00160Hcr=9:3 T atT=270 K [24] that\nρSMR (arbitrary units)\nρSSE (arbitrary units)ρT\nρTa) b)\nα α0 0\n0 -90 90 180 -180 -45 45 0 -90 9010K\n10K250KFIG. 3. Calculated angular dependence of the transverse a)\nSMR (\u001aSMR\nT) and b) local SSE ( \u001aSSE\nT) as contributed by para-\nmagnetic Dy3+moments polarized by an applied \feld H=6 T.\nThe curve at 10 K (blue line) is calculated numerically using\nEq. 5. The 250 K curve (ampli\fed by a factor 100, red line)\nis obtained analytically from Eq. (6). Both SMR and SSE\ngrow with decreasing temperature and associated increasing\nDy3+magnetization.\nfully rotates Gfrom theato thecdirection.Kccan then\nbe estimated using Kc=mWFMHcr+\u001f⊥H2\ncr.Kbis the\nonly free temperature-dependent parameter that we \ft to\nthe \feld-dependent SMR. All other constants are taken\nto be independent of temperature. A typical calculated\ndependence of G(\u000b)and the corresponding contribution\nof the Fe spins to the SMR is shown in Fig. 2 (see below\nfor a more detailed discussion).\nOrdered rare-earth ions can also contribute to the\nSMR and SSE. The spectrum of the lowest-energy6H15/slash.left2\nmultiplet of the Dy3+ion (4f9electronic con\fguration)\nconsists of a Kramers doublet separated by \u0001 =52\ncm−1(≈75 K)from the \frst excited state [45]. At low\ntemperatures, kBT/uni226A\u0001, the Dy moments behave as\nIsing spins tilted by an angle ±\u001eDyaway from the a\naxis in the abplane (\u001eDy=57○). At high temperatures,\nkBT/uni226B\u0001, they can be described as anisotropic Heisen-\nberg spins with paramagnetic susceptibilities, \u001fDy\n∥/parenleft.alt1\u001fDy\n⊥/parenright.alt1\nfor a magnetic \feld parallel (perpendicular) to the local\nspin-quantization axis ( \u001fDy\n∥>\u001fDy\n⊥) [46].\nForkBT/uni226B\u0001, the SMR resulting from the contribu-\ntions of the four Dy sublattices (four Dy sites in the crys-\ntallographic unit cell of DFO) is\nRSMR\nT∝−A/bracketleft.alt1H2sin(2\u000b)−2Hg1Gcsin\u000b/bracketright.alt\n−2BHg 2Gcsin\u000b; (5)\nwhere the \frst term originates from the interaction of Dy\nspins with the applied magnetic \feld and the other two\nterms result from the exchange \feld induced by Fe spins\non Dy sites (for a more detailed discussion of the e\u000bective\nmagnetic \feld acting on Dy spins and the expressions\nfor A and B in terms of the magnetic susceptibilities of\nthe Dy ions see Appendix B). It can be inferred form\nFig. 2 a) that Gcis approximately proportional to cos \u000b.\nTherefore, all terms in Eq. 5 give the sin (2\u000b)dependence\nof the transverse SMR at high temperatures (thick red\nline in Fig. 3 a)). Equation (5) should be added to\nthe SMR caused by the iron sublattice with an unkown5\nweight that is governed by the mixing conductance of\nthe Dy sublattice. We may conclude however that an\nadditional sin (2\u000b)should not strongly change the shape\nof the SMR in Figure 2b).\nAt low temperatures, , T/uni226A\u0001/slash.leftkB, the Dy moments\nbehave as Ising spins. A rotation of the magnetic \feld\nin theabplane modulates the projection of the e\u000bective\nmagnetic \feld on the local spin-quantization axes of the\nfour Dy sublattices, which a\u000bects the angular dependence\nof the SMR. Since the paramagnetic model Eq. (5) can-\nnot be used anymore, we compute the Dy contribution to\nthe SMR ∼mxmynumerically for the rare-earth Hamil-\ntonian\nH(i)\nDy=gJ\u0016B(J⋅HDy)−K\n2(J⋅^zi)2; i=1;2;3;4;(6)\nwith Jas the Dy total angular momentum, gJ=4/slash.left3\nthe Land\u0013 e factor, K=\u0001/slash.left7 the anisotropy parameter,\nwhich is known to reasonably describe the low-energy\nexcited states of Dy ions and ^ziare the local easy axes\nrotated by +57○, for the Dy sublattices 1 and 3, and\n−57○, for the sublattices 2 and 4, away from the aaxis.\nThe magnetic \feld HDyacting on Dy spins is the sum\nof the applied \feld and the exchange \feld from Fe spins:\nHex=g1Gz^a±g2Gz^b, where the +/slash.left−is for the sublattices\n1;3 and 2;4, respectively. We neglect the ccomponent of\nthe exchange \feld, since the Dy magnetic moment along\nthecis small and does not a\u000bect the SMR. Using the\nHamiltonian Eq. (6), we calculate the average aandb\ncomponents of the magnetic moments of the 4 Dy sublat-\ntices at a temperature Tand the resulting contributions\nto SMR. The angular dependence of the SMR due to Dy\nspins is plotted in Fig. 3 a).\nThe calculations recover the sin (2\u000b)angular depen-\ndence of the SMR from Eq. (5) at high temperatures.\nAt 10 K (blue line) the SMR curve becomes strongly\ndeformed: The angular dependence of the SMR shows\npeaks and dips at the e\u000bective \feld directions orthog-\nonal to the quantization axis ^ziof thei-th rare-earth\nsublattice.\nFor long magnon relaxation time, the SSE generated\na spin current that is assumed to be proportional to\nthe bulk magnetization and can therefore provide addi-\ntional information. We focus here on the low tempera-\nture regime because we did not observe an SSE at ele-\nvated temperature, which is an indication that the Dy\nmagnetization plays an important role.\nA net magnetization of rare-earth moments a\u000bects the\nSSE signals in gadolinium iron [47] and gadolinium gal-\nlium [38] garnets. We assume that the SSE is dominated\nby a spin current from the bulk that is proportional to\nthe total magnetization mDy\nbof the four Dy sublattices\nthat we calculated for the Hamiltonian Eq. (6) at 10 K as\nfunction of the angle \u000bof the applied magnetic \feld. The\nmodel predicts peaks at magnetic \feld directions aligned\nwith the Ising-spin axes of the Dy moments, i.e. in be-\ntween those canted by ±33°, which enhances the mag-\nnetization. The contribution from the Fe sublattice to\nRT (mΩ)a) b)\nc) d)/gid00021 /gid00021 /gid00021 /gid00021 /gid00021 /gid00021 /gid00021 /gid00021RT (mΩ)\n∆R T/R0 (10-5)FIG. 4. (a) Transverse SMR (symbols) measured as a func-\ntion of IP magnetic \feld angle \u000band strength (indicated at\nthe top). The measurements are done on device 1 with a cur-\nrent of 1 mA at 250 K and the error bar \u0001 \u000bindicates a sys-\ntematic error due to a possible misalignment of the magnetic\n\feld direction as compared to the crystallographic axes. The\nlines are \fts obtained by adjusting Kbin the free energy model\nEq. (4). (b) The IP ( \u001e) and OOP ( \u0012) canting angles of the\nN\u0013 eel vector with respect to bas a function of the IP magnetic\n\feld direction from the \fts. (c) The maximal signal change\n\u0001RTrduring a magnetic \feld rotation depends linearily on\nthe magnetic \feld strength and (d) shows a power-law tem-\nperature dependence, \u0001 RTr/slash.leftR0∝(T)\u000f.R0is the sheet resis-\ntance obtained from the base resistance of the corresponding\nlongitudinal measurements adjusted by the geometrical fac-\ntor length/width of the Hall bar. These measurements are\ncarried out at 4 T.\nthe SSE is expected to depend as cos \u000bon the external\nmagnetic \feld direction [48]. The ratio of the Fe and Dy\ncontributions to SSE is unknown.\nV. RESULTS\nThe SMR was measured by rotating an IP magnetic\n\feld of various strengths. Temperature drift and noise\nswamped the small signal in the longitudinal resistance as\ndiscussed in Appendix A. Figure 4 a) shows the measured\nresistance of device 1 at 250 K in the transverse (planar\nHall) con\fguration using the left contacts in Fig. 1 b).\nThe results for the right Hall contacts (not shown) are\nvery similar.\nThe (negative) sign of the SMR agrees with our Fe sub-\nlattice model, suggesting that it is caused by the AFM or-\ndered Fe spins with N\u0013 eel vector Gnormal to the applied\nmagnetic \feld. However, Gcannot be strictly normal to\nthe magnetic \feld, because the SMR is not proportional\nto sin(2\u000b), as observed for example in NiO [2]. The\nstrongly non-sinusoidal angular dependence of the SMR6\na) b) c) device 2 IP device 2 OOP device 1 OOP\nFIG. 5. The relative changes in the transverse resistances RTr/slash.leftR0of (a) devices 1 and (b,c) device 2. A linear contribution\nfrom the ordinary Hall e\u000bect has been subtracted from the OOP data. O\u000bsets of the order of 10−4are removed and the curves\nare shifted with respect to each other for clarity. The magnetic \feld directions are (a,b) along zfor the OOP and (c) along\nyfor the IP con\fgurations. (a) The data for device 1 are expected to not change during the Morin transition. The observed\nSMR is symmetric with respect to current and magnetic \feld reversal and sensitive to Dy3+ordering. (b,c) Device 2 reveals the\nMorin transition by a positive step for weak magnetic \felds. Below 23 K, hysteretic resistance features emerge when sweeping\nthe \felds back and forth that vanishes at higher magnetic \felds and temperatures. The arrows indicate the magnetic \feld\nsweep directions, while the symbols highlight the critical magnetic \felds as summarized in Fig. 6.\nis evidence for a non-trivial path traced by the N\u0013 eel vec-\ntor in an applied magnetic \feld as predicted by the model\nEq. (4).\nFigure 2 a) shows the dependence of the three com-\nponentsGa,GbandGcof the N\u0013 eel vector on the IP\norientation angle \u000bof the magnetic \feld, for \u00160H=6\nT. The value of \u000b∈[−90○;90○]is indicated by the color\ncode side bar. When \u000b=0(H/parallel.alt1a), the magnetic \feld\ncauses a tilt of Gaway from the easy aaxis towards the\nhardcaxis since the N\u0013 eel vector parallel to the caxis\ninduces a magnetization along the aaxis. The excursion\nofGfrom theabplane e\u000bectively reduces the role of the\nIP magnetic anisotropy, which leads to a large rotation\nof the N\u0013 eel vector in the abplane for small \u000b(at nearly\nconstantGc). As explained above, this rotation is driven\nby the Zeeman energy of the AFM ordered Fe spins (the\nthird term in Eq.(4)), which favors G⊥Hand competes\nwith the magnetic anisotropy that favors G/parallel.alt1a(the \frst\nterm in Eq.(4)). This behavior is similar to the spin-\rop\ntransition for a magnetic \feld applied along the magnetic\neasy axis, except that Gdoes not become fully orthog-\nonal to the magnetic \feld. As the magnetic \feld vector\nrotates away from the aaxis,Gcand/divides.alt0Gb/divides.alt0decrease, and\nat\u000b=±90○,Gis parallel to the aaxis.\nThe sensitivity of Gto small\u000bgives rise to an abrupt\nchange of the transverse SMR that is proportional to\nGaGbclose to\u000b=0 (thick red line Fig. 2b). The calcu-\nlated and observed SMR scans agree well for T=250 K\nand\u00160H=6 T. Surprisingly, the shape of the experi-mental curves is practically the same at all magnetic\n\feld strengths, i.e. the SMR jumps at \u000b=0 even at\nweak \felds, while the calculation approach the geometri-\ncal sin(2\u000b)dependence (thin blue line in Fig. 2b) calcu-\nlated for\u00160H=2 T). The \fts of the observed SMR for\nall magnetic \felds require a strongly \feld-dependent IP\nanisotropy parameter Kbthat is very small in the zero\n\feld limit:Kb=(6±8)⋅10−6+(3:20±0:02)⋅10−3(H/slash.leftT)2\nK (see Fig. 4a). At present we cannot explain this be-\nhavior. The Dy3+moments should not play an important\nrole in this regime unless a Pt induced anisotropy at the\nDFO/Pt interface modi\fes their magnetism (see below).\nThe exchange coupling between the rare-earth and\ntransition-metal magnetic subsystems is re\rected by the\nsecond term in Eq.(5) of the Dy3+contribution to the\nSMR that is proportional to Gc, i.e. the AFM order of\nthe Fe spins. Since, Gcis a smooth function at \u000b=0, it\ncannot be hold responsible for the large zero-\feld magne-\ntoresistance. The angular SMR appears to be dominated\nby the N\u0013 eel vector Gof the Fe moments, in contrast to\nSmFeO 3, in which the Sm-ions determine not only the\namplitude but also the sign of the SMR [49].\nThe linear increase of the SMR with magnetic \feld\nstrength (see Fig. 4c)) can partly be explained by the\ngrowth of the maximum IP rotation angle, \u001e, of the N\u0013 eel\nvector with magnetic \feld. However, deviations from the\nlinear dependence are then expected close to the critical\nvalue,Ha∼9 T, at which the re-orientation transition\nfromG/parallel.alt1atoG/parallel.alt1cinH/parallel.alt1ais complete [24]. Nevertheless,7\nd1 OOP Dy\nd2 OOP Fe\nd2 OOP Dy\nd2 IP Fe\n[24] M össbauer \n[21] neutron \n[21] M(T) \nFe:GxAyFz\nDy:GxAy\nFIG. 6. Critical magnetic \felds Hcrof the observed transi-\ntions in the transverse resistance as a function of tempera-\nture. Symbols correspond to Fig. 5, where they denote the\nstep functions that trace the Morin transition in device 2. ▲\nindicates IP and ●OOP magnetic \feld directions. The lat-\nter symbol describes the peaks at lower temperatures as well.\nThe OOPHcrof the low magnetic \feld features are shown\nfor device 1 ( ▼) and device 2 ( /uni220E). The features for the IP\nmagnetic \feld directions are less pronounced and not shown.\nThe lines show a \ft by the function Hc∝(TM−T)\u000f, which\nis used to extract the ordering temperatures of 50 K and 23 K\nfor the Morin transition and a magnetic phase transition to\nan ordered Dy3+sublattice, respectively. Further data is from\nRefs. [21, 24], obtained by M ossbauer spectrometry ( /uni220E), neu-\ntron scattering ( /uni25C2) and magnetometry ( ▲).\nthe SMR signal shows no sign of saturation at \u00160Ha=6 T\nandT=250 K. The \u00160Hof Dy becomes of the order\nof kBT at a magnetic \feld strength of 37 T, indicating\ncontributions from the paramagnetic rare earth spins re-\nmains linear in the applied \feld strengths.\nFurther evidence for rare earth contributions at higher\ntemperatures is the Curie-like power-law temperature de-\npendence of the SMR (see Fig. 4d)) SMR ∼T\u000f, with\n\u000f=−1:24±0:04 at low temperatures and \u000f=−1:67±0:02\nat high temperatures.[50] For comparison, in the AFM\nNiO,\u000fis positive and the SMR signal grows quadrati-\ncally with the AFM order parameter [2]. At tempera-\ntures well below the N\u0013 eel transition TFe\nN=645 K, the Fe\nbased magnetic order is nearly temperature independent.\nThe strong magnetic \feld and temperature dependence\ntherefore suggest important contributions from polarized\nDy3+moments even at room temperature.\nThe puzzling strong magnetic \feld-dependence of Kb\nfrom the data \ft might indicate a di\u000berent coupling be-\ntween the rare earth and transition metal magnetic sub-\nsystems at the interface and in the bulk. It can be jus-\nti\fed by the following symmetry argument. The gener-\nators of the Pbnm space group of the DFO crystal are\nthree (glide) mirror planes: ~ ma, ~mbandmc, i.e. a mirror\nre\rection combined with a shift along a direction parallelto the mirror plane. mcis broken at the interface nor-\nmal to the caxis. In the absence of mc, the rare earth\norder parameters A′\naandG′\nbtransform to Gbthat de-\nscribes the AFM order of Fe spins, which allows for a\nlinear coupling between the rare earth and Fe spins at\nthe interface. Since Gbstrongly depends on \u000bat\u000b=0,\nthe same may hold for the rare earth moments at the\ninterface. The SMR is very surface sensitive and could\nbe strongly a\u000bected by this coupling.\nNext, we turn to the SMR at temperatures below the\nMorin transition at magnetic \felds around the re-entrant\n\feld,Hcr. Figure 5(a) shows the transverse SMR of\ndevice 1 in an OOP magnetic \feld, while the data for\nlongitudinal resistance are deferred to the Appendix A,\nFig. 8a). We subtracted a linear \feld dependent contri-\nbution from the OOP data that is caused by the ordinary\nHall e\u000bect in Pt.\nThe zero-\feld resistance of device 1 should not change\nunder the Morin transition when the N\u0013 eel vector direc-\ntion switches from atobnor should it be a\u000bected by weak\nmagnetic \felds H/parallel.alt1c(\u00160Hcr<0:1 T near 50 K [21]) that\nreturn the system to G/parallel.alt1a. Indeed, we do not see any\nweak-\feld anomaly of the SMR near 50 K in Fig. 5a).\nHowever, below 23 K, a negative SMR proportional to\nthe applied \feld appears. The linear \feld-dependence\nends abruptly with a positive step-like discontinuity (see\nFig. 5a)). No resistance o\u000bset has been observed between\nthe zero-\feld \u0000 1and the high-\feld \u0000 4phases. After sub-\nstraction of the strictly linear ordinary Hall e\u000bect contri-\nbution, the SMR feature is an even function of Hc. The\nmagnetic phase transition at 23 K appears to be unre-\nlated to the Morin transition and has not been reported\npreviously.\nThe Morin transition is clearly observed in the OOP\nand IP SMR of device 2, in which the crystallographic\naxes are azimuthally rotated by 45○relative to the Hall\nbar as shown in Fig. 1b). Here, an SMR signal is ex-\npected for both magnetic phases and the 90○rotation of\nthe N\u0013 eel vector from atobshould change its sign from\npositive for the AFM \u0000 1phase ( G∥b) to negative for\nthe WFM \u0000 4phase ( G∥a), for/divides.alt0Hc/divides.alt0>Hcr. The \u0000 1phase\ncan also be suppressed by an IP \feld H∥^y=^b−^athat\nrotates the N\u0013 eel vector towards ^bto lower the Zeeman\nenergy. The drop in the Hall resistance observed in de-\nvice 2 below 48 K for the OOP (Fig. 5 b)) and IP (Fig. 5\nc)) \feld directions can therefore be ascribed to the Morin\ntransition with a temperature-dependent Hcr. The SMR\nsteps are negative, as expected.\nAt even lower temperatures the model appears to break\ndown since we observe hysteretic behavior in the \feld-\ndependence of the SMR signal at low magnetic \felds for\nboth the OOP and IP directions. These features come\nup below 23 K, so appear to have the same origin as\nthe anomalies in device 1. For the OOP direction, the\nlow-\feld anomalies in device 2 are peaks while they are\nstep-like in device 1. Wang et al.[21] did not observed a\nhysteresis in the Fe3+magnetic sublattice and suggested\nthat observed hysteretic behaviour [6, 51] is an evidence8\nFIG. 7. The SSE, i.e. the detected voltage in the transverse\nHall probe divided by the squared current of device 1 at 10 K\nas a function of the magnetic \feld strength and direction \u000b.\nAt weak \felds, the SSE shows a cos \u000bdependence as expected\nfor the Fe3+magnetic sublattice. This amplitude initially\nincreases with the magnetic \feld strength but decreases again\nand \rattens for H>0:5 T.\n 0.1T 0.2T 0.3T 0.4T 0.5T 0.75T 1T 2T\n00000\n0 \n0\n0 RT (V A-2)\n-135-90-450459013510\nα\nfor long-range to short-range Dy3+magnetic order. The\nSMR might witness an ordering of Dy3+moments at the\ninterface at a higher temperature than in the bulk that\ncannot be detected by other measurements.\nAnother unexpected feature is a linear negative mag-\nnetoresistance at /divides.alt0Hy/divides.alt0>Hcrfor the IP con\fguration (see\nFig. 5c)) that might be caused by a canting of GFeto-\nwards cbyHa>1:6T[24]. A misalignment of the crys-\ntallographic axes could also a\u000bect the SMR more sig-\nni\fcantly for high magnetic \felds. However, neither of\nthese mechanisms explain the IP magnetic \feld depen-\ndence and the peaks and low magnetic \feld features in\nthe OOP measurements of both devices below 23 K (Fig.\n5a) and 5b)). Since their signs and shapes vary, we can\nexclude a paramagnetic OOP canting of the Dy3+orbital\nmoments. The Dy3+orbital moments are locked to the\nIsing axis in the abplane and the magnetization is one\norder of magnitude larger in this plane than along the c\ndirection [6]. This might explain the IP SMR features in\nterms of an IP \feld and temperature dependent order of\nthe Dy3+moments.\nThe 90○spin reorientation at the Morin transition\nmaximizes the Fe3+contribution to the SMR. The in-crease of the IP signal amplitude by one order of mag-\nnitude upon lowering the temperature, see Fig. 5(c) is\ntherefore unexpected. The signals become as large as\n1%, one order of magnitude larger than the SMR signals\nof Pt on Y 3Fe5O12[7, 27{29] and a factor four larger\nthan that of \u000b-Fe2O3[52]. Ordered Dy3+magnetic mo-\nments appear to be responsible for the anomalous signals\nbelow 23 K. They interact with the Fe sublattice by the\nexchange interaction, as observed before in the multifer-\nroic phase at temperatures exceeding TDy\nNunder a 0.5 T\nmagnetic \feld [21]. A contribution of Dy3+moments to\nthe magnetization has also been observed in terms of an\nupturn of the magnetization and hyper\fne \feld below\n23 K [53].\nThe SMR steps in device 1 around TDy\nN=4 K at which\nthe Dy moments order spontaneoulsy, are similar to those\nat higher temperature, which supports the hypothesis\nthat the latter are also related to Dy3+order. Device\n2 shows an increased H crmatching those in device 1 at\nthese temperatures. Both devices show no non-linear an-\ntisymmetric \feld dependence, indicating that the Dy3+\nordering above 4 K is \feld-induced. Li et al. [51] ob-\nserved jumps in the thermal conductivity around 4 T and\nattributed these to a spin reorientation of the Fe sublat-\ntice. However, no further transitions are observed up to\n6 T as is shown in Appendix A, so we cannot con\frm\nsuch an Fe3+transition.\nThe magnetic \feld and temperature of the occurrences\nof SMR steps at spin transitions and of SMR anomalies\nare collected in Fig. 6, including the peaks in the OOP\nmeasurements of device 2, using the same markers as in\nFig. 5. The data on the Morin transition agrees with\nprevious observations [21, 24]. The Morin point for both\nIP an OOP con\fgurations is around 50 K, whereas the\ntransitions ascribed to an ordering of the Dy3+moments\noccur around 23 K. Upon lowering the temperature, the\ntransitions associated to the Dy3+and Fe3+moments ap-\nproach each other and merge below TDy\nN, which is another\nindication of a strong inter-sublattice exchange interac-\ntion.\nFigure 7 summarizes the observed IP SSE data of de-\nvice 1 at 10 K. The angular dependence of the resistance\nat small \felds shows the cos \u000bdependence, indicating\nthat the magnon spin current jminjected into Pt is con-\nstant with angle. The amplitude initially increases lin-\nearly with \feld, but decreases again for H>0:5 T. The\nSSE signal of a uniaxial AFM has cos \u000bdependence for\nan IP rotating magnetic \feld [48]. The SSE is small at\nangles for which our model for the Dy3+contribution in\nFig. 3b) predicts a peak. However, we do not observe\nthe expected Dy3+-induced SSE contribution due to the\nDy3+magnetization shown in Fig. 3. On the contrary,\nan increase in Dy3+magnetization appears to suppress\nthe SSE signal. These results suggest that the angular\ndependence of the SSE is governed not so much by the\nordering of the Dy spins, but by their e\u000bect on the fre-\nquencies of the antiferromagnons in the Fe magnetic sub-\nsystem. The ordering of Dy spins leads to a hardening9\nof the AFM resonance modes [54]. The applied magnetic\n\feld suppresses the Dy spin ordering and results in a sub-\nstantial decrease of the spin gap [54] , which a\u000bects the\nthermal magnon \rux and, hence, the SSE. At room tem-\nperature, the SSE signal does not rise above the noise\nlevel of 0.18 V A−2.\nVI. CONCLUSION\nWe studied the rare earth ferrite DFO by measur-\ning the transverse electric resistance in Pt \flm con-\ntacts as function of temperature and applied magnetic\n\feld strength and direction. Results are interpreted in\nterms of SMR and SSE for magnetic con\fgurations that\nminimize a magnetic free energy model with magnetic\nanisotropies, Zeeman energy and DMI. The N\u0013 eel vector\nappears to slowly rotate OOP and displays jumps un-\nder IP rotating magnetic \felds. Magnetic \feld-strength\ndependences indicate that Fe3+spins are responsible for\nthe symmetry of the SMR, but that the Dy3+orbital\nmoments a\u000bect the amplitude. The \frst-order Morin\ntransition is clearly observed at temperatures below 50 K.\nAdditional sharp features emerge below 23 K at critical\n\felds below that of the Morin transition. These observed\nfeatures cannot be understood by the Fe3+N\u0013 eel vec-\ntor driven SMR. Rather, they suggest a magnetic \feld-induced ordering of Dy3+established by the competition\nbetween applied magnetic and exchange \felds with Fe3+.\nThis hypothesis is supported by the similar SMR features\nat the spontaneous Dy3+moment ordering temperature\nTDy\nN. A Dy3+order above TDy\nNalso appears to suppress\nthe SSE contributions from the Fe sublattice.\nConcluding, we report simultaneous manipulation and\nmonitoring of the ordering of both transition metal and\nrare earth magnetic sublattices and their interactions\nas a function of temperature and magnetic \feld in the\ncomplex magnetic material DFO.\nVII. ACKNOWLEDGEMENTS\nWe thank A. Wu for growing the single crystal\nDyFeO 3, J. G. Holstein, H. Adema, T. J. Schouten, H. H.\nde Vries and H. M. de Roosz for their technical assistance\nas well as R. Mikhaylovskiy and A. K. Zvezdin for discus-\nsions. This work is part of the research program Magnon\nSpintronics (MSP) No. 159 \fnanced by the Nederlandse\nOrganisatie voor Wetenschappelijk Onderzoek (NWO)\nand JSPS KAKENHI Grant Nos. 19H006450, and the\nDFG Priority Programme 1538 Spin-Caloric Transport\n(KU 3271/1-1). Further, the Spinoza Prize awarded in\n2016 to B. J. van Wees by NWO is gratefully acknowl-\nedged\n[1] S. Loth, S. Baumann, C. P. Lutz, D. M. Eigler, and A. J.\nHeinrich, Science 335, 196 (2012).\n[2] G. R. Hoogeboom, A. Aqeel, T. Kuschel, T. T. M. Pal-\nstra, and B. J. van Wees, Applied Physics Letters 111,\n052409 (2017).\n[3] D. Afanasiev, B. A. Ivanov, A. Kirilyuk, T. Rasing, R. V.\nPisarev, and A. V. Kimel, Physical Review Letters 116,\n097401 (2016).\n[4] P. Wadley, B. Howells, J. \u0014Zelezn\u0013 y, C. Andrews, V. Hills,\nR. P. Campion, V. Nov\u0013 ak, K. Olejn\u0013 \u0010k, F. Maccherozzi,\nS. S. Dhesi, S. Y. Martin, T. Wagner, J. Wunderlich,\nF. Freimuth, Y. Mokrousov, J. Kune\u0014 s, J. S. Chauhan,\nM. J. Grzybowski, A. W. Rushforth, K. Edmond, B. L.\nGallagher, and T. Jungwirth, Science 351, 587 (2016).\n[5] T. Moriyama, K. Oda, T. Ohkochi, M. Kimata, and\nT. Ono, Scienti\fc reports 8, 14167 (2018).\n[6] Y. Tokunaga, S. Iguchi, T. Arima, and Y. Tokura, Phys-\nical Review Letters 101, 097205 (2008).\n[7] H. Nakayama, M. Althammer, Y. T. Chen, K. Uchida,\nY. Kajiwara, D. Kikuchi, T. Ohtani, S. Gepr ags,\nM. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B.\nGoennenwein, and E. Saitoh, Physical Review Letters\n110, 206601 (2013).\n[8] J. Fischer, O. Gomonay, R. Schlitz, K. Ganzhorn, N. Vli-\netstra, M. Althammer, H. Huebl, M. Opel, R. Gross,\nS. T. B. Goennenwein, and S. Gepr ags, Physical Review\nB97, 014417 (2018).\n[9] Y. Ji, J. Miao, Y. M. Zhu, K. K. Meng, X. G. Xu, J. K.\nChen, Y. Wu, and Y. Jiang, Appl. Phys. Lett. 112,232404 (2018).\n[10] R. Lebrun, A. Ross, O. Gomonay, S. Bender, L. Bal-\ndrati, F. Kronast, A. Qaiumzadeh, J. Sinova, A. Brataas,\nR. Duine, and M. Kl aui, Communications Physics 2, 50\n(2019).\n[11] S. M. Wu, W. Zhang, A. Kc, P. Borisov, J. E. Pearson,\nJ. S. Jiang, D. Letterman, A. Ho\u000bmann, and A. Bhat-\ntacharya, Physical Review Letters 116, 097204 (2016).\n[12] S. M. Rezende and J. B. S. Mendes, Applied Physics\nLetters 111, 172405 (2017).\n[13] G. R. Hoogeboom and B. J. van Wees, Physical Review\nB102, 214415 (2020).\n[14] E. A. Turov, Physical Properties of Magnetically Ordered\nCrystals (Izd. Akad. Nauk SSSR, 1963).\n[15] I. Dzyaloshinsky, J. Phys. Chem. Solids 5, 253 (1958).\n[16] T. Moriya, Physical Review 120, 91 (1960).\n[17] L. M. Holmes, L. G. Van Uitert, R. R. Hecker, and G. W.\nHull, Physical Review B 5, 138 (1972).\n[18] E. F. Bertaut, \\Spin Con\fgurations of Ionic Structures:\nTheory and Practice.\" (1963).\n[19] T. Yamaguchi and K. Tsushima, Physical Review B 8,\n5187 (1973).\n[20] A. Stroppa, M. Marsman, G. Kresse, and S. Picozzi,\nNew Journal of Physics 12, 093026 (2010).\n[21] J. Wang, J. Liu, J. Sheng, W. Luo, F. Ye, Z. Zhao,\nX. Sun, S. A. Danilkin, G. Deng, and W. Bao, Phys-\nical Review B 93, 140403(R) (2016).\n[22] P. Belov, A. K. Zvezdin, M. Kadomtseva, and I. B.\nKrynetskii, Journal of Experimental and Theoretical10\nPhysics 40, 980 (1974).\n[23] A. Maziewski and R. Szymczak, Journal of Physics D:\nApplied Physics 10, 37 (1977).\n[24] L. A. Prelorendjo, C. E. Johnson, M. F. Thomas, and\nB. M. Wanklyn, Journal of Physics C: Solid State Physics\n13, 2567 (1980).\n[25] F. Zhang, S. Li, J. Song, J. Shi, and X. Sun, IEEE\nTransactions on Magnetics 51, 1000904 (2015).\n[26] Y. J. Ke, X. Q. Zhang, H. Ge, Y. Ma, and Z. H. Cheng,\nChinese Physics B 24, 037501 (2015).\n[27] Y. T. Chen, S. Takahashi, H. Nakayama, M. Altham-\nmer, S. Goennenwein, E. Saitoh, and G. Bauer, Physical\nReview B 87, 144411 (2013).\n[28] N. Vlietstra, J. Shan, V. Castel, J. Ben Youssef, G. E. W.\nBauer, and B. J. van Wees, Applied Physics Letters 103,\n032401 (2013).\n[29] M. Althammer, S. Meyer, H. Nakayama, M. Schreier,\nS. Altmannshofer, M. Weiler, H. Huebl, S. Gepr ags,\nM. Opel, R. Gross, D. Meier, C. Klewe, T. Kuschel, J. M.\nSchmalhorst, G. Reiss, L. Shen, A. Gupta, Y. T. Chen,\nG. E. W. Bauer, E. Saitoh, and S. T. B. Goennenwein,\nPhysical Review B 87, 224401 (2013).\n[30] K. Ganzhorn, J. Barker, R. Schlitz, B. A. Piot, K. Ollefs,\nF. Guillou, F. Wilhelm, A. Rogalev, M. Opel, M. Al-\nthammer, S. Gepr ags, H. Huebl, R. Gross, G. E. W.\nBauer, and S. T. B. Goennenwein, Physical Review B\n94, 094401 (2016).\n[31] B. W. Dong, J. Cramer, K. Ganzhorn, H. Y. Yuan, E. J.\nGuo, S. Goennenwein, and M. Kl aui, Journal of Physics:\nCondensed Matter 30, 035802 (2018).\n[32] A. Aqeel, N. Vlietstra, J. A. Heuver, G. E. Bauer, B. No-\nheda, B. J. Van Wees, and T. T. Palstra, Physical Review\nB92, 224410 (2015).\n[33] A. Aqeel, N. Vlietstra, A. Roy, M. Mostovoy, B. J. Van\nWees, and T. T. Palstra, Physical Review B 94, 1334418\n(2016).\n[34] K. Zhao, T. Hajiri, H. Chen, R. Miki, H. Asano, and\nP. Gegenwart, Physical Review B 100, 45109 (2019).\n[35] L. J. Cornelissen and B. J. Van Wees, Physical Review\nB93, 020403 (2016).\n[36] L. J. Cornelissen, J. Liu, R. A. Duine, J. b. Youssef, and\nB. J. Van Wees, Nature Physics 11, 1022 (2015).\n[37] R. Lebrun, A. Ross, S. A. Bender, A. Qaiumzadeh,\nL. Baldrati, J. Cramer, A. Brataas, R. A. Duine, and\nM. Kl aui, Nature 561, 222 (2018).\n[38] K. Oyanagi, S. Takahashi, L. J. Cornelissen, J. Shan,\nS. Daimon, T. Kikkawa, G. Bauer, B. J. van Wees, and\nE. Saitoh, Nature Communications 10, 4740 (2019).\n[39] R. Cheng, J. Xiao, Q. Niu, and A. Brataas, Physical\nReview Letters 113, 057601 (2014).\n[40] N. Vlietstra, J. Shan, B. J. van Wees, M. Isasa,\nF. Casanova, and J. B. Youssef, Physical Review B 90,\n174436 (2014).\n[41] V. V. Eremenko, S. L. Gnatchenko, N. F. Kharchenko,\nP. P. Lebedev, K. Piotrowski, H. Szymczak, and\nR. Szymczak, Europhysics Letters 11, 1327 (1987).\n[42] S. Cao, L. Chen, W. Zhao, K. Xu, G. Wang, Y. Yang,\nB. Kang, H. Zhao, P. Chen, A. Stroppa, R. Zheng,\nJ. Zhang, W. Ren, J. \u0013I~ niguez, and L. Bellaiche, Scienti\fc\nReports 6, 37529 (2016).\n[43] A. K. Zvezdin, V. M. Matveev, A. A. Mukhin, and A. I.\nPopov, Moscow, Izd. Nauk , 296 (1985).\n[44] C. Hahn, G. de Loubens, V. V. Naletov, J. B. Youssef,\nO. Klein, and M. Viret, European Physics Letters 108,571 (2014).\n[45] A. K. Zvezdin and V. M. Matveev, Zh. Eksp. Teor. Fiz. ,\nTech. Rep. 3 (1979).\n[46] U. V. Valiev, J. B. Gruber, S. A. Rakhimov, and O. A.\nNabelkin, physica status solidi (b) 237, 564 (2003).\n[47] S. Gepr ags, A. Kehlberger, F. Coletta, Z. Qiu, E. J.\nGuo, T. Schulz, C. Mix, S. Meyer, A. Kamra, M. Al-\nthammer, H. Huebl, G. Jakob, Y. Ohnuma, H. Adachi,\nJ. Barker, S. Maekawa, G. E. W. Bauer, E. Saitoh,\nR. Gross, S. T. B. Goennenwein, and M. Kl aui, Nature\nCommunications 7, 10452 (2016).\n[48] W. Yuan, Q. Zhu, T. Su, Y. Yao, W. Xing, Y. Chen,\nY. Ma, X. Lin, J. Shi, R. Shindou, X. C. Xie, and\nW. Han, Science Advances 4, 1098 (2018).\n[49] T. Hajiri, L. Baldrati, R. Lebrun, M. Filianina, A. Ross,\nN. Tanahashi, M. Kuroda, W. L. Gan, T. O. Mente\u0018 s,\nF. Genuzio, A. Locatelli, H. Asano, and M. Kl aui, Jour-\nnal of Physics: Condensed Matter 31, 445804 (2019).\n[50] We have not been able to identify the mechanism for the\nstep observed between 135 K and 150 K that has to our\nknowledge not been reported elsewhere either.\n[51] Z. Y. Zhao, X. Zhao, H. D. Zhou, F. B. Zhang, Q. J. Li,\nC. Fan, X. F. Sun, and X. G. Li, Physical Review B 89,\n224405 (2014).\n[52] J. Fischer, M. Althammer, N. Vlietstra, H. Huebl,\nS. Goennenwein, R. Gross, S. Gepr ags, and M. Opel,\nPhysical Review Applied 13, 014019 (2020).\n[53] S. S. K. Reddy, N. Raju, C. G. Reddy, P. Y. Reddy,\nK. R. Reddy, and V. R. Reddy, Journal of Magnetism\nand Magnetic Materials 396, 214 (2015).\n[54] T. N. Stanislavchuk, Y. Wang, Y. Janssen, G. L. Carr,\nS. W. Cheong, and A. A. Sirenko, Physical Review B\n93, 094403 (2016).\n[55] S. L. Gnatchenko, K. Piotrowski, A. Szewczyk, R. Szym-\nczak, and H. Szymczak, Journal of Magnetism and Mag-\nnetic Materials 129, 307 (1994).\n[56] J. H. Han, C. Song, F. Li, Y. Y. Wang, G. Y. Wang,\nQ. H. Yang, and F. Pan, Physical Review B 90, 144431\n(2014).\n[57] K. Oyanagi, J. M. Gomez-Perez, X. Zhang, T. Kikkawa,\nY. Chen, E. Sagasta, L. E. Hueso, V. N. Golovach, F. S.\nBergeret, F. Casanova, and E. Saitoh, arXiv:2008.02446\n(2020).\nAppendix A: Longitudinal and 2K SMR\nThe modulation of the longitudinal Pt resistance as a\nfunction of magnetic \feld are shown in Fig. 8 for com-\nparison with the transverse SMR. The longitudinal sig-\nnals are a\u000bected by a background contact resistance that\nis sensitive to temperature changes. The SMR signals\nare therefore more distorted by a small temperature drift\nthan the transverse measurements. Moreover, the back-\nground resistance su\u000ber from increased noise.\nThe OOP resistance changes of device 1 are one order\nof magnitude larger than those of device 2 and dominated\nby hysteretic e\u000bects. The signal amplitudes of OOP and\nIP con\fgurations for device 2 are similar. The measure-\nment time of one data point below 0.2 T is smaller than at\nlarger \felds, in\ruencing the shape of the graphs. Device\n2 shows hysteretic features at low magnetic \felds and be-11\nLa) b)\nLc)Ldevice 2 IP device 2 OOP device 1 OOP\nFIG. 8. Relative changes in the longitudinal SMR RL/slash.leftR0of (a) device 1 as well as (b,c) device 2 at temperatures up to 50K for\nmagnetic \feld sweeps (a,b) OOP and (c) IP. Device 1 shows large hysteretic e\u000bects at the full range of magnetic \feld strengths.\nThe amount of data points around zero magnetic \feld, and thus the waiting time per magnetic \feld change, is higher than at\nhigher \felds as to have higher resolution for the transverse Morin transition. This makes the hysteretic e\u000bects slightly distorted\ncompared to a situation with constant waiting time. Device 2 shows hysteretic e\u000bects solely at lower \feld strengths which\ncorresponds to the hysteretic features of the transverse measurements.\n1\nFIG. 9. Transverse resistance of device 2 at 2 K as a function\nof the OOP magnetic \feld up to 6 T. The resistance increases\ncontinuously with magnetic \feld strength above 2 T.\nlow 23 K, for both IP and OOP magnetic \felds that are\nsimilar to the transverse SMR features discussed in the\nmain text.\nResults of a \feld sweep up to 6 T are shown in Fig. 9.\nThe resulting continuous curve does not show transitions\non top of those discussed in the text, without evidence\nfor a phase transition at 4 T and 2 K [51, 55].Appendix B: Exchange interaction\nThe Pbnm crystal symmetry allows an exchange cou-\npling between the Dy3+moments and G-type AFM or-\ndered Fe spins. The coupling of the 4 (individual) Dy\nspins in the unit cell with the Fe spins is described as\nEDy−Fe=−g1Gc(ma\n1+ma\n2+ma\n3+ma\n4)\n−g2Gc(ma\n1−ma\n2+ma\n3−ma\n4)\n−g3Ga(mc\n1+mc\n2+mc\n3+mc\n4)\n−g4Gb(mc\n1−mc\n2+mc\n3−mc\n4); (B1)\nwhere the indices 1 ;2;3;4 label the rare-earth ions in the\nunit cell. The exchange \feld from Fe ions is estimated to\nbe∼2 T at low temperatures [45].\nForkBT/uni226B\u0001, the magnetization of the Dy sublattice\nm∥=\u001fDy\n∥H∥andm⊥=\u001fDy\n⊥H⊥for \feld components\nparallel and perpendicular to the local anisotropy axis\nandH=/radical.alt2\nH2\n∥+H2⊥. We assume that the transverse\nSMR caused by the paramagnetic Dy3+moments\npolarized by the applied \feld is proportional to mxmy\n[9, 32, 56, 57]. Adding the contributions of the four Dy\nsites in the crystallographic unit cell of DFO and the\nexchange \feld from the Fe spins acting on the Dy spins\nas described in the main text, we obtain Eq. 5 with\nA=/bracketleft.alt2(\u001fDy\n∥+\u001fDy\n⊥)2−(\u001fDy\n∥−\u001fDy\n⊥)2cos(4\u001eDy)/bracketright.alt2/slash.left2 andB=\nsin(2\u001eDy)/bracketleft.alt2(\u001fDy\n∥)2−(\u001fDy\n∥)2−(\u001fDy\n∥−\u001fDy\n∥)2cos(2\u001eDy)/bracketright.alt2.\nThe coupling constants g 3and g 4do not appear in\nthe expression for SMR since the latter does not de-\npend on the c-component of Dy spins. Moreover, the12\nc-component is very small at low temperatures, since the easy axes of Dy ions lie in the ab plane. Both g 1and g 2\nlead to (nearly) the same angular dependence of SMR." }, { "title": "2101.03542v1.Magnetic_anisotropy_in_van_der_Waals_ferromagnet_VI3.pdf", "content": "1 \n Magnetic anisotropy in van-der-Waals ferromagnet VI3 \n \nA. Ko riki1,2, M. Míšek3, J. Pospíšil1, M. Kratochvílová1, K. Carva1, J. Prokleška1, P. \nDoležal1, J. Kaštil3, S. Son4,5,6, J-G. Park4,5,6, and V. Sechovský1 \n \n1 Charles University , Faculty of Mathematics and Physics, Department of Condensed Matter Physics, Ke \nKarlovu 5, 121 16 Prague 2, Czech Republic \n2Hokkaido University, Graduate School of Science, Department of Condensed Matter Physics, Kita10, Nishi \n8, Kita -ku, S apporo , 060-0810, Japan \n3 Institute of Physics, Academy of Sciences of Czech Republic, v.v.i, Na Slovance 2, 182 21 Prague 8, Czech \nRepublic \n4 Center for Quantum Materials, Seoul National University, Seoul 08826, Korea \n5 Center for Correlated Electron Systems, Institute for Basic Science, Seoul 08826, Korea \n6 Department of Physics and Astronomy, Seoul National University, Seoul 08826, Korea \nABSTRACT \nA comprehensive study of magnetocrystalline anisotropy of a layered van-der-Waals ferromagnet \nVI3 was performed. We measured angular dependences of the torque and magnetization with respect \nto the direction of the applied magnetic field within the “ac” plane perpendicular to and within the \nbasal ab plane , respectively . A two-fold butterfly -like signal was detected by magnetization in the \nperpendicular “ ac” plane . This signal symmetry remains conserved throughout all magnetic regimes \nas well as through the known structural transition down to the lowest temperatures. The max imum of \nthe magnetization signal and the resulting magnetization easy axis is significantly tilted from the \nprincipal c axis by ~40°. The c lose relation of the magnetocrystalline anisotropy to the crystal \nstructure was documented. In contrast, a two-fold-like angular signal was detected in the \nparamagnetic region within the ab plane in the monoclinic phase , which transforms into a six -fold-\nlike signal below the Curie temperature TC. With further cooling, another six-fold-like signal with an \nangular shift of ~30° grows approaching TFM. Below TFM, in the triclinic phase, the original six-fold-\nlike signal vanishes , being replaced by a secondary six-fold-like signal with an angular shift of ~30° . \nI. INTRODUCTION \nMagnetic van der Waals (vdW) materials have recently become hot subjects of interest \nbecause of their potential use in atomically thin devices for spintronic and optoelectronic \nfunction alities1-5. Exploring new chemistry paths to tune their magnetic and optical properties enable \nsignificant progress in fabricating heterostructures and ultra-compact devices6-9 by mechanical \nexfoliation10 or well -controlled element deposition11. Despite belonging to a well -studied family of \ntransition metal trihalides, VI 3 has received significant attention only recently10, 12-14. Nevertheless, \nthe first studies pointed out that the magnetism of VI 3 is very complicated . This is probably related \nto the incomplete high spin state t2g shell in V3+ ions here, which allows for a Jahn-Teller distortion. \nThe detailed results of 51V and 127I NMR spectroscopy complemented by temperature dependencies \nof specific -heat and magnetization15 revealed two distinct ferromagnetic (FM) phases, the ground -\nstate one having the critical temperature TFM = 26 K and another existing between TFM and TC = 49.5 \nK (notation is taken from Ref.14). VI3 behaves at low temperatures as a hard ferromagnet with a high \ncoercive field 0Hc ~ 1 T at 2 K for the magnetic field applied parallel to the c-axis ( Hc)10, 12, while \nHc value is considerably lower in the perpendicular direction ( H ab). In fields above 2 T the \nmagnetization M in H c is proportional to ~1.2 M in H ab. This anisotropy of magnetization persists \nat least up to 9 T10, 12. 2 \n A crucial property of the vdW materials with considerable application potential is the strong \nmagnetocrystalline anisotropy2. In layered materials, two limits of easy‐axis are in‐plane (XY model) \nand out‐of‐plane (Ising model). The anisotropy features of VI 3 were investigated by magnetization \nmeasurement s. These ha ve shown a different value of saturated magnetization (higher along the c \naxis) and a significantly larger low-temperature magnetization loop hysteresis along the c axis12, 16, \n17. The Raman scattering study has also identified an acoustic magnon mode (the spin-wave gap) and \na two -magnon mode in the low-temperature FM state. With the help of li near spin-wave theory \ncalculations, the magnetic anisotropy and intralayer exchange strength are estimated as 1.16 and 2.75 \nmeV . The sizeable magnetic anisotropy implies the stability of the FM order in the 2D limit with a \nhigh critical field16 . \nFirst-principles calculations have been used to investigate the magnetism and magnetic \nanisotropy of VI 3 theoretically, however, mostl y in the single -layer or bilayer limit18-21, with only a \nfew works consider ing the bulk form10, 22. Overall, magnetocrystalline anisotropy with an easy axis \nalong c has been predicted for both bulk and single layer in these works. A high orbital momentum \non V was found in a calculation based on the linear augmented plane wave method, which led to a \nrobust magnetic anisotropy dominated by a single -ion contribution of 15.9 meV per formula unit. \nBased on these findings , it was concluded that VI 3 behaves as a n Ising -like ferromagnet, similar to \nCrI 323. In a different calculation based on the projector augmented -wave framework t he energy \ndifference bet ween the in-plane and out -of-plane magnetization orientation was found to be strongly \ndependent on both Hubbard U as well as strain. Furthermore , this energy difference changes sign at \nU ~ 3.5 eV, close to the expected value of U24. \nCurrently , there is a controversy concerning the VI3 crystal structure and its conjunction with \nmagnetism . Kong et al. have reported a trigonal ( R-3) structure at room temperature with a structural \ntransition below Ts = 78 K13. A contradictory result has been reported by Son et al.12 showing a \nstructur al phase transition at Ts = 79 K where the crystal symmetry lowers with cooling from the \ntrigonal P-31c structure to a monoclinic C2/c in which the system become s ferromagnetic at \ntemperature TC = 50 K . Conversely , Tian et al.10 claim that the VI 3 structural phase transition is \nsimilar to the structural transition of CrI 325, i.e., they observed a lowering of the symmetry with \nheating through Ts from R-3 to C2/m . A detailed crystal structure study by Doležal et al. has \nconfirmed the trigonal ( R-3) crystal structure at room temperature in agreement with T. Kong13 and \nfound a subsequent symmetry lowering to a triclinic structure at TFM2 = 32 K, when the ferromagnetic \nphase FM I transforms to a different ferromagnetic phase FM II. The connection of the structur e with \nthe magnetic phase transition in VI 3 suggests a considerable role of magnetoelastic interactions in \nthis compound. It is also suppo rted by magnetostriction -induced changes of the monoclinic -structure \nparameters at TC. Surprisingly , the temperature of structure transition is decreased by magnetic fields \napplied along the trigonal c-axis and is intact by a magnetic field applied within the basal plane. These \nresults indicate that the magnetic field applied within the three -fold axis of the VI 6 octahedrons \nstabilizes the symmetry of the V honeycombs in the basal plane26. \nThe direct observation of the VI3 magnet ocrystalline anisotropy was revealed by the angular \ndependence of magnetization at temperature 2 K in various magnetic fields H rotated out -of-plane \nfrom the c axis to a (and b) axis and within the ab plane. The out -of-plane anisotropy exhibits a \n“butterfly -like” pattern17, different from the other 2D FM semiconductor Cr2Ge2Te627. On the other \nhand , the angular signal of magnetization within the basal ab plane shows a six-fold-like symmetry . \nHowever, anisotropy studies so far performed were mostly focused on the lowest temperature \nmagnetic regime below TFM. Further , detailed measurements throughout the entire temperature \ninterval covering all detected magnetic regimes and structural phases are missing and highly desirable \nto understand complex magnetic interaction s and anisotropy in the VI3 vdW ferromagnet . The VI 3 \nmagnetic structure, exchange interactions and magnetocrystalline anisotropy are much more \ncompl icated than earlier studied Cr tri -halides 28-32. \nThis paper focuses on studying magnetocrys talline anisotropy features of VI 3 in a wide \ntemperature interval by angular -dependent magnetic torque and magnetization measurements for the 3 \n magnetic field applied out-of-plane marked as “ ac” (a general perpendicular plane to the ab basal \nplane ) and within the ab basal plane. \nII. EXPERIMEN TAL METHODS \nThe single crystals were prepared by the chemical vapor transport m ethod , as described \nelsewhere12. The angular dependence of the magnetization was measured by MPMS (Quantum \nDesign) devices using thin single crystals with a sample area of 1.4 × 1.8 mm2. The sample was placed \non a handmade rotator with a rotation axis orthogonal to the external magnetic field. Magnetic torque \nwas measured using a commercial Quantum Design Torque magnetometer chip attached on a 370° \nrotator installed inside a 9T PPMS apparatus . To measure the angular dependence of torque within \nthe princ ipal ab plane with respect to magnetic field orientation , an L-shaped copper element had to \nbe attached to the torque chip (Fig. 1). The sample was fixed with Apiezon grease . The to rque \nmagnetometer chip can typically be calibrated. However , because of installed cooper L element , \ncalibration was not possible to finalize, and so we do not focus on the absolute value of the torque \nbut scale it a ppropriately. The L-element mass contribution was tested by the rotation of the chip in \nzero magnetic field; the observed signal was negligible compared to the signal in magnetic field s with \nthe attached sample. The magnetic field is rotating within the ab plane of the sample for both \nmeasurements. Since the magnetic field was applied within the plane of the thin sample, the shape of \nsamples and the demagnetization factor are expected not to have a significant effect on the results. \nAll sample s used for measurement s were stored under the inert Ar (6N purity) atmosphere in a \nglovebox . The contact with air was minimized only on the time of the installation of the rotators into \nthe MPMS and PPMS instrument s. \n \nFIG. 1. The scheme of the magnetic torque chip with the installed sample attached to the copper \nL-element. The scheme was taken from the Quantum Design PPMS manual and modified. \nIII. RESULTS \nWithin our study , we have performed a series of angular -dependent magnetization scans \nwithin the “ac” plane, perpendicular to the ab plane (Fig. 2). Our low-temperature (2 K) and high-\nmagnetic -field (5 T) angular magnetization result s in the form of a butterfly -like signal in polar \nprojection agree ing with the previous result in Ref.17. Both, the symmetry and positions of the maxima \nof the angular magnetization signal rota ted within the “ ac” plane remain conserve d throughout all \nmagnetic regimes a nd the structural transition to the triclinic phase26 at TFM. The maximum \nmagnetization signal tilted approximately 40° out of the c-axis. \n4 \n \nFIG. 2. Angular -dependent magnetization isotherms in magnetic field 5 T rotated within the “ac” \nplane. The polar plot (a) and classical plot (b) shows selected curves at specific magnetic regimes; \nabove TC (60 K), below TC (50 K), in the upper vicinity of TFM (30 K) , and well below TFM (20 and \n10 K). \n \nThe angular -dependent magnetic torque measurements within the basal ab plane in identical \nmagnetic field 5 T at various temperatures are shown in Fig. 3. \n \n(a) \n(b) 5 \n \n \n \nFIG. 3. Angular -dependent magnetic torque scans in magnetic field 5 T rotated within the ab plane. \nThe polar plot (a) shows all recorded temperatures. The classical plot (b) shows selected curves at \nspecific magnetic regime s; above TC (60 K), below T C (50 K), in the upper vicinity of TFM (30 K) , \nand well below TFM (10 K ). The intensity of the torque signal is on a relative scale. \n \nAngular -dependent magnetic torque provide s information mainly about the \nmagnetocrystalline anisotropy strength . In contrast , the angular -dependent magnetization \nmeasurement s show the size of the magnetic moment at the given orientation of the sample with \nrespect to the direction of the external magnetic field. The isothermal scans of the torque have clearly \nrevealed significant evolution of the magnetocrystalline anisotropy (Fig. 3). A double peak -like signal \nin the paramagnetic region in the monoclinic phase transforms to a six -fold-like signal below TC. The \nsix-fold-like symmetry is present down to the lowest temperatures. However , a detailed analysis of \nthe curves between TC and TFM show s signature s of the gradual splitting of each maxim um when \napproaching TFM. At a temperature of 30 K in the vicinity of TFM the signal is split into two sets of \nsix-fold-like signals of similar intensity mutually rotated by 30° (see Fig. 3b – green line) . The \noriginal six -fold-like signal , which has emerged at TC, becomes gradually reduced with decreasing \ntemperature in favor of intensity of the second set. The original set vanishes below TFM whereas the \nsecond set remains down to the lowest temperature at the fixed angular position . \n(a) \n(b) 6 \n The a ngular -dependent scans of magnetization isotherms with magnetic field 5 T applied \nwithin the basal ab plane are shown in Fig. 4. \n \n \n \n \nFIG. 4. Angular -dependent magnetization isotherms in magnetic field 5 T rotated within the ab \nplane. The polar plot ( a) shows all recorded temperatures. The classical plot (b) shows selected \n(a) \n(b) \n(c) 7 \n curves at specific magnetic regimes; above TC (60 K), below TC (50 K), in the upper vicinity of TFM \n(30 K) , and well below TFM (20 and 10 K). Panel (c) shows the temperature evolution of the \nmagnetization averaged over the entire angular scan. \n \nThe angular -dependent magnetization measurements have shown complementary result s with \nthe magnetic -torque measurements. One six-fold-like signal was detected below TC. With further \ncooling, a clear sign ature of a second six -fold-like signal rotated by ~30° is observed (see green line \nin Fig. 4b). The second signal gradually grows at the expense of the original one’s intensity . Below \nTFM, the original six-fold-like signal vanishe s and only the second one is detect ed. The signal in the \nparamagnetic regime at temperatures below Ts is relatively weak, with two-fold-like broad maxima \naround ~15° and ~195° (Fig. 4b) which can be distinguished analogically to the symmetry of \nmagnetic torque data. The averaged angular value of the magnetic moment at given temperatures is \nplotted in Fig. 4c and saturates around value 1.7 B/V, which is lower than expected for the V3+ ion. \nIV. DISCUSSION \nOur detailed experimental study of magnetocrystalline anisotropy by angular -dependent \nmagnetization has unambiguously proved that the magnetic moment is canted from the c-axis by ~ \n40°. A somewhat surprising result is the robustness of the anisotropy throughout the entire \ntemperature interval covering both ferromagnetic phases and particularly the struct ural transition \nfrom the monoclinic to likely triclinic phase accompanying the magnetic order -to-order transition at \nTFM. \nIn contrast to the “ ac” plane, a more complex angular response of the magnetization and \ntorque were detected in the basal ab plane. In the following discussion, we will place the ab plane in \nthe plane of the VI 3 layer and take the direction of the two -fold axis as the b axis. \n \nFIG. 5. Angular dependence of torque at 60 K and fitting of the result by combining trigonometric \nfunctions. \n \nWhen the system reveal s three -fold rotational symmetries around the c-axis, which is the case \nof the rhombohedral R-3 space group for VI 3 above Ts, only the sin 6 term is present in the tensor z \nof magnetization torque (Eq. 1 , see Appendix ). \n𝜏𝑧=𝑀𝑥𝐻𝑦−𝑀𝑦𝐻𝑥=𝜒330𝐻6sin6𝜙 Eq. 1 \nFocusing on the paramagnetic region at 60 K below Ts (Fig. 5), the best agreement for \nexperimental torque data was found using formula (Eq. 2) \n𝜏(𝜙)=𝑘2sin2𝜙+𝑘4sin4𝜙+𝑘6sin6𝜙, Eq. 2 \n8 \n where two -fold and four -fold signals have been observed. This is a direct indication that the symmetry \nof the crystal below Ts is lowered at least to a monoclinic one at this temperature because the sin 2 \nand sin 4 contributions are forbidden for three -fold rotational symmetries in the tensor z (Eq.1) for \na rhom bohedral structure . \nSince the highest point group of a single layer in VI 3 is D3d, there are three possible \norientations of two-fold axes when symmetry is reduced to the monoclinic structure (Fig. 6). These \nform three domains associated with each other on a three -fold axis perpendicular to the ab plane. On \nthe other hand, when a magnetic field is applied to the ab plane of a system with two -fold rotations \naround the b (or a) axis, the torque i n the paramagnetic state can be expressed only by the sum of sin \nfunction s (Eq. 2). \n \nFIG. 6. Schematic view of three two -fold axes and three possible domains when the VI3 transforms \nfrom the rhombohedral to the monoclinic structure. \n \nIf the volume ratio of the three domains is A : B : C (A + B + C = 1), the torque applied to \nthe entire system is \n𝜏tot(𝜙)=𝐴𝜏(𝜙)+𝐵𝜏(𝜙−𝜋3⁄)+𝐶𝜏(𝜙−2𝜋3⁄) \n=𝐷{𝑘2sin2(𝜙+𝛼2⁄)+𝑘4sin4(𝜙−𝛼4⁄)}+𝑘6sin6𝜙 Eq. 3 \nwhere D and α are constants determined by the volume ratio of the domains. As shown in Fig. 5, the \nphase of the sine function of the second and fourth oscillation terms is out of phase, which indicates \nthat this is the sum of multiple domains. Since the phase shift due to the domain ratio does not occur \nin the six-fold oscillation term, the angle at which the torque becomes zero corresponds directly to \nthe main axis of the crystal. \nAs shown in Fig. 3, at 50 K near the fer romagnetic transition at TC, a six -fold saw -tooth -like \nsignal was observed. This is due to a rapid change in sign due to switching of the ferromagnetic \ndomains, indicating that the magnetization is not fully polarized in the direction of the external \nmagnetic field . This implies that the magnetocrys talline energy due to the VI 6 octahedron is \nsufficiently large with respect to the Zeeman energy. The result can be understood qualitatively by \nconsidering the VI 6 octahedron to be a regular octahedron as a first approximation. The \nmagnetocrystalline aniso tropy energy 𝐸𝑎𝑐(𝜃,𝜙), represented by the symmetric representation of the \npoint group Oh, can be written as follows \n𝐸𝑎𝑐(𝜃,𝜙)=𝐾1{(7cos4𝜃−6cos2𝜃+3)\n12−√2\n3sin3𝜃cos𝜃cos3𝜙}…Eq. 4 \nThis equation is obtained from the commonly use d magnetic anisotropy equation K1 \n(αx2αy2+αy2αz2+αz2αx2) where αi represents the direction cosine. We obtain this equation by making \nthe [111] axis into a new c-axis, which corresponds to the three -fold rotation axis of the VI 6 \noctahedron by an orthogonal transformation . This formula confirms that the easy -axis direction of \nspontaneous magnetization has six values, φ = 0, π/3, 2π/3, π, 4π/3, 5π/3, independent of the sign of \nK1. Therefore, when the magnetic field is rotated within the ab plane, the moment switches to the \nmost stable d irection among these easy axes that exist every 60 ˚, as shown in Fig. 7. \nc\n9 \n \nFIG. 7. The angular dependence of torque at 50 K. The dashed blue line is a scaled graph of the six -\nfold oscillation term from the 60 K result (Fig. 5). \n \nWhen the magnetic field is in the middle of the two easy axes, an abrupt change in the sign of \nthe torque takes place . Therefore, the main axis of the crystal is also considered to be in this position. \nIn Fig. 7, the six oscillations in the fitting res ults for the paramagnetic state are shown together, and \nthe two results correspond to each other. \nAs the temperature was lowered further within the ferromagnetic region, magnetic torque \nmeasurements indicate there was a significant change in the easy magnetization axis (Fig. 3). \nAlthough this change does not have a clear boundary temperature, another component with an easy \naxis tilted about ~30˚ graduall y grows and finally replaces the easy axis direction at the lowest \ntemperature. This change cannot b e explained, for example, by a change in the sign of the magnetic \nanisotropy energy 𝐸𝑎𝑐(𝜃,𝜙). Previous reports have suggested the presence of a first -order structural \nphase transition from monoclinic to triclinic at around 30 K26, and we expect to see a qualitative \nchange reflecting this observation . Detailed crystal structure analysis at low temperatures below TFM \nis highly desirable . The surprising result is that the “ ac” plane signal is insensitive to signal symmetry \nchange within the ab plane. We can only speculate that the FM I and FM II magnetic structures are \nvery similar, only mutually rotated by 30 . \nThe results of the angular dependence of magnetization in Fig. 4 are in qualitative agreement \nwith the torque result; w e obtained six-fold-like signals . The folding back when the magnetization is \nreduced corresponds to ferromagnetic domain switching. The original easy axis changes with \ndecreasi ng temperature about ~30˚ below TFM, which remain conserve d down to the lowest \ntemperature of 10 K . For simplicity, to analyze the data, we have decompose d the model into \nspontaneous magnetization Ms, which is independent of the magnitude of the magnetic field, and \npolarized magnetization Mp, which follows the magnetic field. In this model, the result is fitted with \na function of |cosφ|+const . in Eq. 5, as shown in Fig. 8. \n \n10 \n \nFIG. 8. Angular -dependent magnetization isotherm at 50 K and in magnetic field 5 T rotated within \nthe ab plane and the fit at 50 K given by the Eq. 5 . \n \n𝑀𝑠{𝐴|cos(𝜋\n𝑤(𝜙−𝜙0))|+𝐵|cos(𝜋\n𝑤(𝜙−𝜙0−60))|+𝐶|cos(𝜋\n𝑤(𝜙−𝜙0−120 ))|}+𝑀𝑝 \n(Eq. 5) \n \nFitting result Ms = 0.44(1) μB/V and Mp = 0.62(1) μB/V is roughly consistent with previous \nmagnetization results of previous experiments at 50 K. \n \nFIG. 9. Angular dependence of the magnetization at 10 K rotated in a clockwise and anticlockwise \ndirection in an identical magnetic field. \n \nFig. 9 shows the angular dependence of the magnetization at 10 K. The hysteresis is not \nnoticeabl e at 50 K, but some non-negligible hysteresis is observed at 10 K. Hysteresis, such as the \nintersection of clockwise and anticlockwise data that exists around 30˚ and 210˚, is typica lly expected \n11 \n to occur at the minima of magnetization where the domain switches. This suggests that the two \ncomponents, one developed at around 50 K and the other developed at a lower temperature, coexist \nas indepen dent signals. \n \n \nFIG. 10. The projections of the magnetocrystalline anisotropy energy 𝐸𝑎𝑐(𝜃,𝜙) for anisotropy in the \nsimplified Oh representation . Pictures a), b), and c) for the principal axis 111 and magnetic moment \nis stable in the 100 direction; d) p rojection of easy axes to the ac plane . \n \nThe observed anisotropy within the ab plane can be understood using projections of the \nmagnetocrystalline anisotropy energy 𝐸𝑎𝑐(𝜃,𝜙) in a simplified Oh representation for the 111 direction \nas the principal axis of the VI6 octahedron for magnetic moment stable in the 100 direction of the \ncube, i.e. the direction of Iodine that leads to the six-fold-like signal (see Fig. 10a, 10 c). On the other \nhand, the projection of the easy a xis in the “ac” plane (Fig. 10d) leads to a two-fold-like signal , in \nwhich the expected moment flipping position is not every 90°. \nV. CONCLUSIONS \nWe have performed a detailed study of the magnetocrystalline anisotropy of VI 3 by angular -\ndependent torque magnetometry and magnetization within the ab plane and the perpendicular “ ac” \nplane throughout the entire 10K-60K temperature interval covering both the two ferromagnetic \nphases and the paramagnetic phase. Irrespective of the temperature c hange, the symmetry of the two -\nfold butterfly -like angular signal within the “ac” plane remains conserve d throughout both FM \nphases . On the other hand, s ubstantial d evelopment of the signal was detected within the ab plane. \nThe d ouble peak -like signal in the paramagnetic region in the monoclinic phase transforms to the six-\nfold-like signal below TC. The original six -fold-like signal appearing at TC disappears below TFM and \nthe second set of six -fold-like signal rotated by about 30° sur vives down to the lowest temperatures. \nThe gradual emergence of the secondary six -fold-like signal growth between TC and TFM and the \nangular shift of magnetization maxima about ~30° below TFM is the subject of further research. Since \nthe torque contains a periodic function of 2 and 4, we corroborate that the three -fold symmetry \npresent in the rhombohedral lattice is broken in the studied temperature range. \nOur findings are in agreement with the presence of low symmetry monoclinic and triclinic \nphases at low temperatures . We have shown that the easy axis is not perpendicular to the layers, but \ncanted by around 40° from the c-axis in both FM phases. VI 3 is thus not a simple Ising -like \nferromagnet. For further progress in this field , it is essential to have detailed knowledge about the \npredicted triclinic crystal structure below TFM and a neutron diffraction study to reveal the VI 3 \nmagnetic structure . \nACKNOWLEDGMENTS \n12 \n This work is part of the research program GACR 19 -16389J which is financed by the Czech \nScience Foundation . Work at the Center for Quantum Materials was supported by the Leading \nResearcher Program of the Nation al Research Foundation of Korea (Grant No. 2020R1A3B2079375) \nwith partial funding by the grant IBS -R009 -G1 provided by the Institute f or Basic Science of the \nRepublic of Korea . Experiments were performed in the Materials Growth and Measurement \nLaboratory (see: http://mgml.eu ) which is supported within the program of Czech Research \nInfrastructures (project n o. LM2018096). We are indebted to Ross H. Colman for making language \ncorrections. \nREFERENCES \n1 P. Ajayan, P. Kim, and K. Banerjee, Physics Today 69, 39 (2016). \n2 K. S. Burch, D. Mandrus, and J. G. Park, Nature 563, 47 (2018). \n3 W. Zhang, P. K. J. Wong, R. Zhu, and A. T. S. Wee, InfoMat 1, 479 (2019). \n4 D. Zhong, K. L. Seyler, X. Y. Linpeng, R. Cheng, N. Sivadas, B. Huang, E. Schmidgall, T. \nTaniguchi, K. Watanabe, M. A. McGuire, W. Yao, D. Xiao, K. M. C. Fu, and X. D. Xu, \nScience Advances 3, e1603113 (2017). \n5 J. G. Park, Journal of Physics: Condensed Matter, 301001 (2016). \n6 F. F. Li, B. S. Yang, Y. Zhu, X. F. Han, and Y. Yan, Applied Surface Science 505, 144648 \n(2020). \n7 J. Shang, X. Tang, X. Tan, A. J. Du, T. Liao, S. C. Smith, Y. T. Gu, C. Li, and L. Z. Kou, \nAcs Applied Nano Materials 3, 1282 (2020). \n8 F. Subhan and J. S. Hong, Journal of Physical Chemistry C 124, 7156 (2020). \n9 K. Yang, W. T. Hu, H. Wu, M. H. Whangbo, P. G. Radaelli, and A. Stroppa, Acs Applied \nElectronic Mater ials 2, 1373 (2020). \n10 S. Tian, J. F. Zhang, C. Li, T. Ying, S. Li, X. Zhang, K. Liu, and H. Lei, Journal of the \nAmerican Chemical Society 141, 5326−5333 (2019). \n11 M. Gronke, B. Buschbeck, P. Schmidt, M. Valldor, S. Oswald, Q. Hao, A. Lubk, D. Wolf, \nU. Steiner, B. Buchner, and S. Hampel, Advanced Materials Interfaces 6, 1901410 (2019). \n12 S. Son, M. J. Coak, N. Lee, J. Kim, T. Y. Kim, H. Hamidov, H. Cho, C. Liu, D. M. Jarvis, P. \nA. C. Brown, J. H. Kim, C. H. Park, D. I. Khomskii, S. S. Saxena, and J. G. P ark, Physical \nReview B 99, 041402(R) (2019). \n13 T. Kong, K. Stolze, E. I. Timmons, J. Tao, D. R. Ni, S. Guo, Z. Yang, R. Prozorov, and R. J. \nCava, Advanced Materials 31, 1808074 (2019). \n14 T. Hattori, Y. Ihara, K. Ishida, Y. Nakai, E. Osaki, K. Deguchi, N. K. Sato, and I. Satoh, \nJournal of the Physical Society of Japan 80, SA007 (2011). \n15 E. Gati, Y. Inagaki, T. Kong, R. J. Cava, Y. Furukawa, P. C. Canfield, and S. L. Bud'ko, \nPhysical Revie w B 100, 094408 (2019). \n16 B. B. Lyu, Y. F. Gao, Y. Zhang, L. Wang, X. Wu, Y. Chen, J. Zhang, G. Li, Q. Huang, N. \nZhang, Y. Chen, J. Mei, H. Yan, Y. Zhao, L. Huang, and M. Huanb, Nano Letters 20, \n6024−6031 (2020). \n17 J. Yan, X. Luo, F. C. Chen, J. J. Gao, Z. Z. Jiang, G. C. Zhao, Y. Sun, H. Y. Lv, S. J. Tian, \nQ. W. Yin, H. C. Lei, W. J. Lu, P. Tong, W. H. Song, X. B. Zhu, and Y. P. Sun, Physical \nReview B 100, 094402 (2019). \n18 F. Subhan and J. S. Hong, Journal of Physi cs: Condensed Matter 32, 245803 (2020). \n19 Z. P. Guo, Q. Chen, J. N. Yuan, K. Xia, X. M. Wang, and J. Sun, Journal of Physical \nChemistry C 124, 2096 (2020). \n20 C. Long, T. Wang, H. Jin, H. Wang, and Y. Dai, Journal of Physical Chemistry Letters 11, \n2158 (2 020). \n21 Y. L. Ren, Q. Q. Li, W. H. Wan, Y. Liu, and Y. F. Ge, Physical Review B 101, 134421 \n(2020). 13 \n 22 M. An, Y. Zhang, J. Chen, H. M. Zhang, Y. J. Guo, and S. Dong, Journal of Physical \nChemistry C 123, 30545 (2019). \n23 K. Yang, F. Fan, H. Wang, D. I. Kho mskii, and H. Wu, Physical Review B 101, 100402 \n(2020). \n24 Y. P. Wang and M. Q. Long, Physical Review B 101, 024411 (2020). \n25 G. T. Lin, X. Luo, F. C. Chen, J. Yan, J. J. Gao, Y. Sun, W. Tong, P. Tong, W. J. Lu, Z. G. \nSheng, W. H. Song, X. B. Zhu, and Y. P. Sun, Applied Physics Letters 112, 072405 (2018). \n26 P. Doležal, M. Kratochvílová, V. Holý, P. Čermák, V. Sechovský, M. Dušek, M. Míšek, T. \nChakraborty, Y. Noda, S. Son, and J. G. Park, Physical Review Materials 3, 121401(R) \n(2019). \n27 W. Liu, Y. H. Dai, Y. E. Yang, J. Y. Fan, L. Pi, L. Zhang, and Y. H. Zhang, Physical \nReview B 98, 214420 (2018). \n28 Y. Liu and C. Petrovic, Physical Review B 97, 174418 (2018). \n29 J. L. Lado and J. F. Rossier, 2D Materials 4, 035002 (2017). \n30 L. Webster and J. A. Yan, Phys ical Review B 98, 144411 (2018). \n31 V. K. Gudelli and G. Y. Guo, New Journal of Physics 21, 053012 (2019). \n32 D. Torelli and T. Olsen, 2D Materials 6, 015028 (2018). \n \nAPPENDIX: Torque expression in the paramagnetic state \n Considering the situation where time-reversal symmetry is preserved, we expand the \nmagnetization as a function of the magnetic field as follows. \n𝑀𝑖=𝜒𝑖𝑗(2)𝐻𝑗+𝜒𝑖𝑗𝑘𝑙(4)𝐻𝑗𝐻𝑘𝐻𝑙+𝜒𝑖𝑗𝑘𝑙𝑚𝑛(6)𝐻𝑗𝐻𝑘𝐻𝑙𝐻𝑚𝐻𝑛+⋯(𝑖,𝑗,𝑘=𝑥,𝑦 or 𝑧) \nEach χ(k) in this formula represents the magnetic susceptibility tensors of the k-th rank, and \nhere, t he general sum reduction rule is used. Since these tensors are symmetrical with respect to the \nintercha nge of indices, the distinction between independent components is determined only by the \nnumber of the subscripts x, y, and z . Therefore, for simplicity, we have decided to denote these tensors \nby the number of its xyz , for example, as follows. \n𝜒𝑦𝑧(2)≡𝜒011, 𝜒𝑥𝑦𝑦𝑧(4)≡𝜒121, 𝜒𝑥𝑥𝑥𝑥𝑥𝑥(6)≡𝜒600 \n We consider the situation where a magnetic field is applied in the ab(xy) plane, so ( Hx, Hy, \nHz) = (Hcosφ, Hsinφ, 0). Then the general equation for the torque up to the 6th order can be written \nas follows. \n𝜏𝑧={1\n2(𝜒200−𝜒020)𝐻2+1\n4(𝜒400−𝜒040)𝐻4+5\n32(𝜒600+𝜒420−𝜒240−𝜒060)𝐻6}sin2𝜙\n−{𝜒110𝐻2+1\n2(𝜒310+𝜒130)𝐻4+5\n16(𝜒510+2𝜒330+𝜒150)𝐻6}cos2𝜙\n+{1\n8(𝜒400−6𝜒220+𝜒040)𝐻4+1\n8(𝜒600−5𝜒420−5𝜒240+𝜒060)𝐻6}sin4𝜙\n−{1\n2(𝜒310−𝜒130)𝐻4+1\n2(𝜒510−𝜒150)𝐻6}cos4𝜙\n+{1\n32(𝜒600−15𝜒420+15𝜒240−𝜒060)𝐻6}sin6𝜙−{1\n16(3𝜒510−10𝜒330+3𝜒150)𝐻6}cos6𝜙 \nGenerally, t he tensor is limited in its component s by the point group symmetry of the crystal. \nEq.1 in the discussion considers a situation in which there are three -fold rotational symmetries around \nthe c(z) axis. In this case, e ach component of susceptibility has the following relationship to each \nother . \n𝜒200=𝜒020, 𝜒110=0 \n𝜒400=3𝜒220=𝜒040, 𝜒310=𝜒130=0 \n𝜒600=5𝜒420=5𝜒240=𝜒060, 𝜒510=−𝜒330=𝜒150 14 \n Taking this into account, the equation of torque is expressed as follows. \n𝜏𝑧=𝜒330𝐻6cos6𝜙 \n " }, { "title": "2101.10773v1.Data_driven_design_of_a_new_class_of_rare_earth_free_permanent_magnets.pdf", "content": "Data-driven design of a new class of rare-earth free permanent magnets\nAlena Vishina,1,\u0003Daniel Hedlund,1Vitalii Shtender,2Erna K. Delczeg-Czirjak,3\nSimon R. Larsen,2Olga Yu. Vekilova,4,3Shuo Huang,3Levente Vitos,5,3\nPeter Svedlindh,1Martin Sahlberg,2Olle Eriksson,3,6and Heike C. Herper3\n1Department of Materials Science and Engineering,\nUppsala University, Box 35, 751 03 Uppsala, Sweden\n2Department of Chemistry - Ångström, Uppsala University, Box 538, 751 21, Uppsala, Sweden\n3Department of Physics and Astronomy, Uppsala University, Box 516, SE-75120, Uppsala, Sweden\n4Department of Materials and Environmental Chemistry, Stockholm University, 10691 Stockholm, Sweden\n5Applied Materials Physics, Department of Materials Science and Engineering,\nRoyal Institute of Technology, Stockholm, SE-100 44, Sweden\n6School of Science and Technology, Örebro University, SE-701 82 Örebro, Sweden\n(Dated: January 27, 2021)\nA new class of rare-earth-free permanent magnets is proposed. The parent compound of this class\nis Co 3Mn2Ge, and its discovery is the result of first principles theory combined with experimen-\ntal synthesis and characterisation. The theory is based on a high-throughput/data-mining search\namong materials listed in the ICSD database. From ab-initio theory of the defect free material it is\npredicted that the saturation magnetization is 1.71 T, the uniaxial magnetocrystalline anisotropy\nis 1.44 MJ/m3, and the Curie temperature is 700 K. Co 3Mn2Ge samples were then synthesized and\ncharacterised with respect to structure and magnetism. The crystal structure was found to be the\nMgZn 2-type, with partial disorder of Co and Ge on the crystallographic lattice sites. From mag-\nnetization measurements a saturation polarization of 0.86 T at 10 K was detected, together with a\nuniaxial magnetocrystalline anisotropy constant of 1.18 MJ/m3, and the Curie temperature of TC\n= 359 K. These magnetic properties make Co 3Mn2Ge a very promising material as a rare-earth free\npermanent magnet, and since we can demonstrate that magnetism depends critically on the amount\nof disorder of the Co and Ge atoms, a further improvement of the magnetism is possible. From\nthe theoretical works, a substitution of Ge by neighboring elements suggest two other promising\nmaterials - Co 3Mn2Al and Co 3Mn2Ga. We demonstrate here that the class of compounds based on\nT3Mn2X (T = Co or alloys between Fe and Ni; X=Ge, Al or Ga) in the MgZn 2structure type, form\na new class of rare-earth free permanent magnets with very promising performance.\nI. INTRODUCTION\nRare earth (RE) permanent magnets dominate the\nmarket where high-performance magnetic materials are\nneeded - areas such as green-energy generation, includ-\ning wind and wave power, electric vehicle motors and\ngenerators, and many more. At the same time, the heav-\nier rare-earth elements which are necessary for obtaining\nthegoodmagneticcharacteristicsofthesematerials(such\nas Pr, Nd, Sm, Tb, or Dy) [1], can only be mined with\nmethods that leave an environmental footprint, are quite\nexpensive, and are rapidly decreasing in availability. As\na consequence many rare-earth elements are labelled crit-\nical. In order to pursue green technologies that rely on\nhigh-performance magnets, there is a growing interest in\nfinding new magnetic materials containing cheaper and\nless critical elements, while maintaining a similar high-\nperformance shown by their RE counterparts.\nIn the last decades, many attempts have been made to\ndesign both RE-lean [2, 3] and RE-free magnetic materi-\nals. Thelaterincludehighmagnetocrystallineanisotropy\nalloys (MnBi [4, 5] and MnAl [6, 7]), nanostructures [8–\n\u0003Department of Physics and Astronomy, Uppsala University, Box\n516, SE-75120, Uppsala, Sweden; alena.vishina@physics.uu.se11], thin films [12, 13], alnico permanent magnets, and\nmany more [14–17]. Techniques used for this search in-\nclude both filtering through a large number of systems\n[18–21] and investigating a single or a group of similar\ncompounds[14,15,22–29]. InRef. [18]itwasshownthat\na high-throughput density functional theory (DFT) ap-\nproachcanbeeffectiveinfilteringthroughalargenumber\nof compounds in the search for new high-performance RE\nfreepermanentmagnets, whichwouldbetime-consuming\nand expensive to synthesize experimentally. Unfortu-\nnately, a commercially relevant class of rare-earth free\ncompounds was not identified in Ref. [18]. In the cur-\nrent investigation, that data-filtering approach described\nin Ref. [18] was broadened to involve a much larger start-\ning set of compounds.\nFrom a combination of data-mining efforts, using ab-\ninitio electronic structure theory, materials synthesis,\nstructural and magnetic characterization, we can report\nonanewclassofrare-earthfreepermanentmagnets,with\nvery promising properties. In particular, we have identi-\nfied Co 3Mn2Ge, which is a new material that previously\nhas not been considered as a permanent magnet.arXiv:2101.10773v1 [cond-mat.mtrl-sci] 26 Jan 20212\nII. HIGH-THROUGHPUT AND DATA-MINING\nSEARCH\nAn initial attempt to use high-throughput searches for\nRE-free permanent magnets was published in Ref.[18],\nemploying a smaller subset of possible magnetic com-\npounds. As described in this investigation, a high-\nperformancepermanentmagnetmusthaveferromagnetic\nordering, high saturation magnetization ( MS>1T),\nhigh Curie temperature ( TC>400 K), a large uniax-\nial MAE (>1MJ/m3), and a magnetic hardness larger\nthan 1 (fo details see the Methods section). In the initial\nattempt of using high-throughput/data-mining searches,\nfocus was put on materials with a stoichiometry neces-\nsarily containing one 3 dand one 5 delement [18]. From\na pure magnetic viewpoint, this study resulted in several\npromising materials. Unfortunately, none of the identi-\nfied compounds in Ref.[18] could serve as practical re-\nplacements for the known rare-earth based compounds,\nsince they all contained expensive elements like Pt.\nIn the present study, we have made an extended search\nto involve materials with a stoichiometry that contain\ntwo types of 3d-elements, without any restriction on the\ntotal number of elements in the system. About a thou-\nsand materials from the ICSD database [30] contain two\ndifferent 3d-elements in the chemical formula. After the\nfirst high-throughput calculations of the electronic struc-\nture and magnetic moments, about 45 systems were iden-\ntified to have a magnetic saturation field larger than a\nvalue\u00180.4 T. In the next step, all the materials with\ncubic symmetry were set aside, since spin-orbit effects\nare known to influence the MAE most effectively for non-\ncubic materials [31]. We also performed an extensive lit-\nerature search, to identify if any of the compounds iden-\ntified in the intitial screening steps, had previously been\nreported as permanent magnet (e.g. the L1 0phase of\nFeNi) or if the systems that were identified, are known\nto not be ferromagnetic. These compounds were also re-\nmoved from the list of potential new rare-earth free per-\nmanent magnets. For the remaining materials, the MAE\nwas calculated. Compounds with planar or low magnetic\nanisotropy (MAE <\u00180.6 MJ/m3) were not considered\nfurther, leavingonlyfivematerialswhosesaturationmag-\nnetization and MAE are high enough to be considered as\nhigh-performance permanent magnets.\nNote that the search criteria used here are somewhat\nless strict than the criteria of magnetic properties spec-\nified from practical aspects of high performance perma-\nnent magnets. The reason for using less strict criteria,\nis to not miss new classes of compounds, that have the\npotential to become technologically relevant, after e.g al-\nloying or structural refinements. Using these criteria we\nhave identified the following compounds as potential new\npermanent magnets: ScFe 4P2, Co 3Mn2Ge, Mn 3V2Si3,\nScMnSi, and Cu 2Fe4S7and we report in Table I their cal-\nculated saturation magnetization, MAE, and magnetic\nhardness. All data were obtained from calculations at\nT= 0K and for a ferromagnetic configuration. Forthis final list of materials, we also calculated the Heisen-\nberg exchange interaction, and from Monte Carlo simu-\nlations we identified the ground state magnetic configu-\nration and, for materials with ferromagnetic configura-\ntions, the Curie temperature. Out of the five materials\nlisted above, Co 3Mn2Ge is the only one that is FM and\nhas aTCof 700 K. The rest of the materials were found\nto have a non-collinear magnetic structure and were for\nthis reason not considered further. Some of the mate-\nrials, discarded during the search, can be found in the\nAppendix , Table III. We’d like to note, that Co 3Mn2Ge\ndoes have its magnetic hardness lower than 1. However,\nthis parameter can be later on adjusted by alloying or\nother methods.\nTABLE I: Calculated MAE, saturation magnetization,\nCurie temperature, and magnetic hardness for the\nmaterials that can be considered as candidates for\nRE-free permanent magnets. NC stands for\nnon-collinear spin structure obtained from Monte Carlo\nsimulations.\nMaterial ICSD Space MAE Sat. TC\u0014\nnumber group MJ/m3magn., T K\nScFe 4P268525 136 0.71 0.65 NC 1.45\nCo3Mn2Ge 52972 194 1.44 1.71 700 0.79\nMn3V2Si3643689 193 0.85 0.73 NC 1.42\nScMnSi 86369 189 0.69 0.52 NC 1.79\nCu2Fe4S715973 51 0.90 0.37 NC 2.87\nIII. EXPERIMENTAL RESULTS\nA. Synthesis and structural characterisation\nFollowing the theoretical results presented above, fo-\ncus was then put on synthesis of Co 3Mn2Ge. Initial tri-\nals revealed the magnetic Heusler phase Co 2MnGe [32]\nas being the main competing phase in that region of the\nphase diagram. Preliminary structural refinement of the\nhexagonal Co 3Mn2Ge phase (MgZn 2-type) in a multi-\nphasesampleindicateddisorderingonthe6 hand2 asites\nimplying intermixing between Co and Ge. EDS analysis\non the multi-phase samples revealed the composition of\nthe desired hexagonal phase to be in a small homogeneity\nregion spanning Co 52Mn34Ge14and Co 53:7Mn31:7Ge14:6.\nSamples following these stoichiometries were then syn-\nthesized. The Co 52Mn34Ge14sample consisted of 95.4\nwt% of the hexagonal phase and 4.6 wt% of CoMn, a\nPauli paramagnet [33]. The diffractogram and accom-\npanying structural refinement is shown in Fig. 1. The\nstructure of the compound (R Bragg= 4.20 %) is in good\nagreement with previous studies [34] having similar unit\ncell dimensions (V = 154.60(1) \u0017A3compared to 154.61\n\u0017A3[34]). The final refinement resulted in a composition\nof Co 3:39Mn2Ge0:61with intermixing of Co and Ge on\nthe 6 hand 2 asites.3\nFIG. 1: (Color online) Refined powder diffraction data\nof the synthesized Co 52Mn34Ge14alloy. Observed\n(Yobs), calculated (Y calc), difference (Y obs-Ycalc)\ndiffraction profiles and Bragg’s peaks positions for\nCo3Mn2Ge (95.4 wt.%) and CoMn (4.6 wt.%) phases\nare shown.\nResults from XRPD refinements and microstructural\nEDS results indicate that the structure is disordered with\na tendency to contain more Co. Kuz’ma et al. [34] pro-\nposed both the ordered Mg 2Cu3Si-type and disordered\nMgZn 2-type structures from their XRPD results but a\nclear atomic distribution could not be established. To\ndetermine the atomic distribution for the theoretical cal-\nculations, precisionsinglecrystalstudieswereconducted.\nThreevariantsofthestructurewereconsideredduringre-\nfinement and are presented in Table II - the completely\nordered structure, Co-Ge intermixing, and Mn-Ge inter-\nmixing. The structure solution obtained with SHELXT-\n2014 quickly provided an ordered model for Co 3Mn2Ge.\nHowever, the wR 2value, the EDS results, and the dif-\nference Fourier map all suggest that this model should\nbe rejected. The model of Mn-Ge intermixing showed\na larger goodness-of-fit and the calculated composition\nwas too far from the measured to be accepted. The fi-\nnal model of Co-Ge intermixing overall shows the best\nparameters and is in agreement with other results. All\natoms were refined anisotropically. Detailed crystallo-\ngraphic results of the SCXRD refinements can be found\nin the Appendix , Table IV.\nB. Experimental magnetism\nThe magnetization of the Co 3Mn2Ge sample was mea-\nsured as a function of temperature and magnetic field\nto determine the magnetic ordering temperature TC,\nthe saturation magnetization, and the effective magnetic\nanisotropy constant of the material. The temperature\ndependent magnetization measurements were performedat several applied magnetic fields and in the temperature\nregion between 10 K and 900 K. The results are shown in\nFig. 2. The magnetic ordering temperature, TC= 359K,\nwas determined from the Curie-Weiss law fit to the tem-\nperature dependence of the inverse magnetic susceptibil-\nity (cf. inset in Fig. 2).\nFIG. 2: (Color online) Magnetization of Co 3Mn2Ge as a\nfunction of temperature in applied magnetic fields of\n\u00160H = 0.01 T (white open circles and white open\nrectangles) and \u00160H = 1 T (red filled rectangles). The\ninset shows the Curie-Weiss fit for the inverse magnetic\nsusceptibility with TC= 359K. The cusp in \u00160H = 1 T\nand drop in magnetization around 175 K is attributed\nto a spin–reorientation ( Tsrt).\nThe isothermal magnetization curves are shown in Fig-\nure 3. Note that Figure 3(a) shows the result for bulk\npowders whereas Figure 3(b–d) corresponds to single\ncrystal measurements at different temperatures along ( k)\nthe crystallographic c–axis and perpendicular ( ?) to it.\nIn Figure 3(a) the isothermal magnetization for the pow-\nder sample, measured at 10 K and 70 K, shows the same\napproachtosaturationandreachthesamesaturationpo-\nlarization of 0.86 T. The isothermal magnetization mea-\nsured at 170 K reaches a saturation polarization close to\nthat recorded at lower temperature; whereas the satura-\ntion polarization measured at 300 K reaches a value of\n0.60 T.\nTo analyze the behavior of the magnetic anisotropy\nand to compare to the value calculated theoretically, we\nhave used the law of approach to saturation to calcu-\nlate the effective anisotropy constant at different temper-\natures. The results are shown in Figure 4. At 10 K and\n70 K, the law of approach to saturation yields an effec-\ntiveanisotropyconstantof1.18MJ/m3andthemagnetic\nhardness parameter, \u0014, becomes 1.42. The experimental\nresults of magnetic configuration, saturation field, and\nmagnetic anisotropy hence seem consistent with the the-\noreticalpredictionsofthiscompound, whichisgratifying.\nA closer inspection of the results presented in Fig. 2 re-\nveal that below\u0018175 K the magnetization drops consid-4\nTABLE II: Results of single crystal refinements of the Co 3+xMn2Ge1\u0000xcompound. The Co-Ge disordered model\n(marked in bold) describes the measured data best.\nParameters Ordered Co-Ge disordered Mn-Ge disordered\nOccupation 6 h Co 0.84Co+0.16Ge Co\n4f Mn Mn Mn\n2a Ge 0.74Co+0.26Ge 0.4Ge+0.6Mn\nCalculated formula Co 3Mn2Ge Co3:24Mn 2Ge0:76Co3Mn2:6Ge0:4\nCalculated composition (at.%) Co 50Mn33:33Ge16:67Co54Mn 33:33Ge12:67Co50Mn43:33Ge6:67\nGoodness-of-fit on F21.214 1.170 1.353\nFinal R indices (I >2s(I)) R 1= 0.0432 R1= 0.0100 R1= 0.0149\nwR2= 0.1188 wR 2= 0.0264 wR2= 0.0360\nR indices (all data) R 1= 0.0439 R1= 0.0111 R1= 0.0160\nwR2= 0.1192 wR 2= 0.0268 wR2= 0.0364\nHighest difference peak 2.277 0.639 0.693\nDeepest hole -5.961 -0.420 -1.003\n1-\u001blevel 0.569 0.119 0.158\nFIG. 3: (Color online) (a) Magnetization of a bulk\npowder of Co 3Mn2Ge, as a function of magnetic field at\n10 K, 70 K, 170 K and 300 K. (b,c,d) Isothermal\nmagnetization of single crystals of Co 3Mn2Ge at 300 K,\n200 K and 100 K. Filled symbols show measurements\nwhere the magnetic field is perpendicular ( ?) to the\ncrystallographic c–axis whereas open symbols show\nmeasurements with the magnetic field parallell ( k)with\nthec–axis.\nerably when the applied field is small (0.01 T), whereas\na cusp–like feature is seen in an applied magnetic field\nof 1 T, both for single crystal material and bulk sam-\nples. Such features can be attributed to a change in mag-\nnetocrystalline anisotropy with temperature and we in-\nterpret this to be a spin–reorientation temperature Tsrt,\nfrom easy–axis to easy–cone anisotropy, a feature which\nis not uncommon for permanent magnets (e.g. Fe 5SiB2\n[35], MnBi [36], Cr 0:9B0:1Te [37], Gd [38] and the cel-\nebrated Nd 2Fe14B [39]). Before presenting further re-\nsults supporting a claim of a temperature induced spin–\nreorientation, we draw attention to the fact that for a\nFIG. 4: (Color online) Approach to saturation plotted\nas the magnetization normalized with the value\nmeasured at 5 T as a function of inverse applied field\nsquared for the bulk sample of Co 3Mn2Ge.\nhexagonal uniaxial material the magnetic anisotropy en-\nergy can be expressed as MAE =K1cos2\u0012+K2sin4\u0012,\nwhereVis the volume, Kiare the anisotropy constants,\nand\u0012is the angle between the magnetization vector and\nthe hexagonal c–axis [40]. Depending on the values of\nK1andK2three cases can be distinguished; the ma-\nterial has an easy–axis, an easy–plane or an easy–cone\nmagnetic anisotropy. The easy-cone state is the most fa-\nvorable one when \u00002K2\u0014K1\u00140. The temperature\ndependence of the anisotropy constants ( Ki(T)) may be\nstrong at low temperature and is usually described by\na power–law behavior, Ki(T)/M(T)\u000b[41, 42], where\nM(T)is the magnetization and \u000bis a constant. The tem-\nperaturedependenceoftheanisotropyconstantscanthen\nlead to a spin–reorientation, e.g. from an easy–axis to an\neasy–plane or an easy–axis to an easy–cone anisotropy.\nThe material under investigation here shows charac-\nteristics of an easy–axis to easy–cone spin–reorientation5\naround 175 K. We support this claim with the aid of\nisothermal magnetization curves recorded at tempera-\ntures below and above Tsrton the single crystals and\nbulk powders. The single crystal isotherms presented in\nFigure3(b–d)indicatethattheeasy–conestateispresent\nat100Kandbelow. Lookingatthemagnetizationcurves\nat 300 K (Figure 3(b)) we see that it is easier to mag-\nnetize the crystal along the c–axis than perpendicular to\nit. This shows that the material is magnetically uniax-\nial at 300 K. The same applies at 200 K (Figure 3(c)),\nwhereas at 100 K the measurements show that it is as\neasy to magnetize the material parallel or perpendicu-\nlar to the crystallographic c–axis. However, it should be\nnoted that at low field ( <0:5T) there is a deviation\nfrom a linear magnetic response in measurements per-\nformed parallel or perpendicular to the crystallographic\nc–axis. The non–linear behaviour at low fields, together\nwiththelowfieldmagnetizationversustemperaturemea-\nsurements support the easy–cone state [43].\nFigure4supportstheconclusionofamorecomplicated\nbehaviour for the magnetization of Co 3Mn2Ge. It should\nbe noted, that in order to extract an anisotropy constant\nusing the law of approach to saturation, averaging over a\nrandom distribution of crystal directions yields the con-\nstant\f= 4=15[44]. The effective anisotropy constants\nand magnetic hardness parameter should thus be seen\nas the tentative descriptions of the magnetocrystalline\nanisotropy at low temperatures.\nSummarizing the experimental results so far we con-\nclude that the theoretical prediction of Co 3Mn2Ge as a\npotential rare-earth free permanent magnet is likely. The\nsaturation magnetization, uniaxial magnetic anisotropy,\nand magnetic hardness parameter obtained from experi-\nments are consistent with the theoretical predictions. It\nmust be noted that a more detailed picture is present in\nthe experiment, with a temperature stabilized easy-cone\nstate. Thisstatewasnotspecifiedinthetheoreticalhigh-\nthroughput screening search, and could therefore not be\nexpected from the theoretical prediction. However, the\neasy axis, order of magnitude of the anisotropy, the sat-\nuration magnetization, and ordering temperature should\nagree, and here the theory has proven useful in finding a\nnew material, with a potential to be used as a permanent\nmagnet. We return to the easy cone state, below, with\na theory that takes disorder into account, and show that\ncalculations then reproduce experimental observations.\nIV. DISORDERED Co 3Mn 2Ge\nAccording to the experimental reports, there may be a\n50-50%intermixingofCoandGeatomson6 hand2 a[34]\nWyckoff positions (see Fig. 5), in addition to the excess\nof Co in comparison to Ge (see Table II), in the exper-\nimental samples. This structure and composition differ\nfrom the one reported in the ICSD [30] database, where\nCo atoms are reported to occupy 6 h, Mn atoms take 4 f,\nand Ge occupies 2 aatomic sites. It should be noted,that disorder of the type detected in our experiments is\noften observed in the related ternary Laves phases with\nMgZn 2-type [45]. In this section, we, therefore, consider,\nfrom ab-initio alloy theory, the disorder of Co 3Mn2Ge, in\norder to analyze the effect it has on the magnetic state of\nthe material and to compare the results with the theoret-\nical data obtained for the ordered structure. The effect of\nchemical and magnetic disorder on the stability and mag-\nnetization of Co 3Mn2Ge was investigated as described in\ntheMethods section (Sec. VII).\nThe results of these theoretical calculations is that the\nordered structure has a lower total energy, compared to\nthe disordered one. However, configurational entropy\ncontributions to the free energy are reported in the ap-\npendix to stabilize the disordered structure at tempera-\ntures 1700 K. The details can be found in Appendix C .\nFIG. 5: (Color online) Ordered Co 3Mn2Ge (as per\nICSD) with Co atoms occupying 6 h(light red balls),\nMn atoms taking 4 f(violet balls), and Ge (grey balls)\noccupying 2 aatomic sites (top); and the disordered\nCo3Mn2Ge structure obtained experimentally with\n50-50% intermixing between the Co and Ge sites\n(bottom).\nFollowing the experimental findings that point to an\neasy-conemagneticanisotropyatthelowertemperatures,\nwe calculated MAE for several magnetization directions\n(0\u000e< \u0012 < 90\u000e,\u001e= 0\u000e) for the ordered Co 3Mn2Ge and\ndisordered Mn 2(Co0:75Ge0:25)4phases, see Fig. 6. In-\ndeed, we can see that the disordered system shows a\npronounced deviation from the simple uniaxial behavior.\nHence, the easy-cone anisotropy observed at lower tem-\nperatures is attributed to the Co-Ge disorder observed\nin the samples. At elevated temperature, the influence\nof disorder evidently is less pronounced, and the uniax-\nial anisotropy that was obtained from theory of ordered6\nsamples, is recovered in the experiments.\nFIG. 6: (Color online) MAE for several magnetization\ndirections ( \u001e= 0\u000e) for the ordered Co 3Mn2Ge and\ndisordered Mn 2(Co0:75Ge0:25)4phases.\nV. CHEMICAL SUBSTITUTION OF Ge\nIn order to find related compounds with similar or im-\nproved magnetic properties compared to Co 3Mn2Ge, we\nperformedadditionaltheoreticalwork, wherewereplaced\nall Ge atoms in the unit cell by Al, Si, P, Ga, As, In, Sn,\nSb, Tl, and Pb. All structures were relaxed and tested\nfor stability by calculating the formation enthalpy with\nrespecttotheirelementalcomponents(fordetailsseesec-\ntionMethods). The systems found to be stable with re-\nspect to the elemental components (Tl- and Pb-based\nmaterials turned out to be unstable) were investigated\nfurther for their magnetic characteristics. Their MAE\nand the details of the magnetic state are given in Table\nVII.\nFIG. 7: (Color online) Magnetic anisotropy energy of\nCo3Mn2X, with X = Al, Si, P, Ga, Ge, As, In, Sn, Sb.\nPositive values indicate a uniaxial anisotrophy.Replacing Ge in Co 3Mn2Ge by the neighboring ele-\nments, does not change to the magnetic state much, but\nit has a large effect on the MAE, see Fig. 7. To de-\ntermine the origin of the difference in MAE values we\nanalyzed the spin-orbit coupling energy (SOC) of the\nCo and Mn atoms with their spins oriented along the\nzandxdirections as well as the partial density of states\n(DOS), similar to the analyses presented in Refs.[46–51].\nThe details of that investigation can be found in Ap-\npendixD. Summarizing the results, we find that apart\nfrom the Co 3Mn2Ge compound that is listed in the ICSD\ndatabase, Co 3Mn2Al and Co 3Mn2Ga are also expected\nto have the properties of a good permanent magnet. We\nhave shown that the main contribution to the large MAE\nof these compounds arises from Co atoms. A synthesis\nand charaterisation of these two compounds is outside\nthe scope of this investigation, but represents clearly an\ninteresting avenue forward.\nVI. DISCUSSION AND CONCLUSIONS\nUsing the high-throughput and data-mining approach\nwe filtered through the RE-free materials of the ICSD\ndatabase that contain necessarily two 3 d-elements. Only\none of those approximately thousand structures satis-\nfied our requirements for a strong permanent magnet,\nCo3Mn2Ge, with the saturation magnetization of 1.71\nT, MAE equal to 1.44 MJ/m3, andTCof 700 K. To\nreduce the cost of this material, we attempted to re-\nplace Ge by other elements. Two materials, Co 3Mn2Al\nand Co 3Mn2Ga, were found to have sufficient magneti-\nzation, MAE, and Curie temperature to be used as high-\nperformance permanent magnets. The former of the two\ncan be problematic to produce due to the competing\nHeusler compound being considerably more stable; the\nlatter doesn’t have a similar drawback.\nCo3Mn2Ge was successfully synthesized using induc-\ntion melting followed by annealing at 1073 K. The re-\nsulting crystal structure was in good agreement with the\npreviously published results. However, when the order-\ning of the system was investigated with precision single\ncrystal analysis, it indicated that the disordered struc-\nture type MgZn 2with intermixing Co and Ge on the 6 h\nand 2 asites was preferred. The structure taken from the\nICSD database and used in the high-throughput calcula-\ntion was, unlike the experimental one, ordered.\nMagnetic characteristics of the synthesized compound\nwere measured and produced the saturation magnetiza-\ntion of 0.86 T at the temperature of 10 K, uniaxial mag-\nnetic anisotropy equal to 1.18 MJ/m3above around 175\nK (with the easy-cone character of anisotropy below),\nandTC=359 K. These values, even though lower than\nthe theoretical predictions, make Co 3Mn2Ge a promis-\ning candidate for a high-performance permanent magnet.\nFurther analysis of its magnetic state as well as the ways\nto tune some of these numbers is highly desirable and\nrepresents ongoing work.7\nCalculations were also performed for the disordered\ncrystal structure of Co 3Mn2Ge obtained in the experi-\nment. We found that the crystallographic ordered struc-\nture with FM ordering is more stable than both ferro-\nmagnetically and antiferromagnetically ordered config-\nurations of the structurally disordered compound, in a\ntemperature range of 0 – 1735 K. Calculations of the\nmagnetocrystalline anisotropy were also performed for\nthe disordered Mn 2(Co0:75Ge0:25)4phase. These calcu-\nlations show a pronounced deviation from the uniaxial\ncharacter of magnetic anisotropy, in agreement with our\nexperimental low temperature data. Combining the re-\nsults of the theoretical and experimental works, leads\nto the conclusion that structural disorder influences the\nmagnetic properties, including the MAE, of Co 3Mn2Ge\nin a detrimental way, and that the best properties are\nexpected for the most ordered samples.\nIn the Appendix H we list the properties of the all\nthe previously known Co–Mn–Ge systems. Based on\nthat information we can conclude, that the disorder is\nquite common in all the Co–Mn–Ge compounds and\ndepends strongly on the sample preparation and ex-\nperimental procedures. With that in mind, we be-\nlieve that further investigation into the sample prepa-\nration or some possible doping alternatives is necessary\nfor an attempt to stabilize Co 3Mn2Ge in its ordered\nform, which is expected to have higher magnetization\nand magnetic anisotropy. Having said that, even the dis-\nordered Mn 2(Co0:75Ge0:25)4phase, with saturation po-\nlarization of 0.86 T and uniaxial magnetic anisotropy\nof 1.18 MJ/m3, possesses the characteristics of a good\npermanent magnet, although a further investigation into\nincreasing the TC= 359 is desirable.\nIt is also worth mentioning, that orientation of magne-\ntocrystallineanisotropyandthemagneticmomentofCo–\nMn–Ge systems is strongly affected by Co:Mn ratio. The\nsynthesized sample of Co 3Mn2Ge has the actual compo-\nsition of Co 52Mn34Ge14which might result in the dis-\ncrepancy between the magnetic results obtained experi-\nmentally and the ones predicted in the high-throughput\nsearch. It was proven that the easy-cone magnetocrys-\ntalline anisotropy, observed at the temperatures below\naround 175 K, is the result of the Co–Ge disorder. We\nwould like to point out, however, that one of the best\ncurrent permanent magnets, Nd 2Fe14B [39], possesses a\nsimilar feature. Nevertheless, this fact is another incen-\ntive to look for the way of stabilizing the ordered phase\nof Co 3Mn2Ge.\nThe current investigation is a promising example of\nthe material found in the theoretical data-mining search\nbeing synthesized and showing the desired characteris-\ntics of a good permanent magnet. Further experimental\ninvestigation into the magnetic state of Co 3Mn2Ge will\nbe performed as well as the computational search for im-\nproving its price-performance. With its Curie tempera-\nture close to the room temperature, we can also consider\nfine-tuning Co 3Mn2Ge for magnetocaloric applications.VII. METHODS\nA. High-throughput DFT\nThe high-throughput screening step to calculate the\nmagnetic moment of the materials was performed using\nthe full-potential linear muffin-tin orbital method (FP-\nLMTO) including spin-orbit interaction as implemented\nin the RSPt code [52, 53]. For this step the initial mag-\nnetic configuration for all the materials was ferromag-\nnetic (FM).\nThe magnetic state of materials (FM or antiferro-\nmagnetic - AFM), unless previously known, was de-\ntermined using Vienna Ab Initio Simulation Package\n(VASP) [54–57] within the Projector Augmented Wave\n(PAW) method [58], along with the Generalized Gradi-\nent Approximation (GGA) in Perdew, Burke, and Ernz-\nerhof (PBE) form [59]. VASP was also used for structure\nrelaxation at the post-high-throughput stage as well as\nto calculate spin-orbit coupling energies, for the analysis\npresented in the Appendix.\nThemagneticanisotropyenergy(MAE)wascalculated\nusing the RSPt code as \u0001E=Epl\u0000Ec; hereEcandEpl\nare the total energies with the magnetization directed\nalong and perpendicular to the c-axis. Calculations were\nperformed with the tetrahedron method with Blöchl cor-\nrection for the Brillouin zone integration [60]. A positive\nsign of the MAE corresponds to the required uniaxial\nanisotropy. For the disordered Co 3Mn2Ge, the MAE was\nevaluated for several polar angles from the force theorem\n[61, 62] as the difference of the eigenvalue sums for the\ntwo magnetization directions e\u0012andec(\u0012is a polar angle\nwhile azimuthal angle is equal to zero), while keeping the\neffective potential fixed. From the first principles calcu-\nlations, the magnetic hardness parameter was evaluated\nfrom the expression \u0014=p\n\u0001E=\u0016 0M2\nS[63], where MSis\nsaturation magnetization and \u00160is the vacuum perme-\nability.\nThe Curie temperature TCwas calculated using Monte\nCarlosimulationsimplementedwithintheUppsalaatom-\nistic spin dynamics (UppASD) software [64]. Atom-\nistic spin dynamics calculations were performed on a\n30\u000230\u000230supercell with periodic boundary conditions.\nThe required exchange parameters were calculated with\nthe RSPt code within the first nine coordination shells\n[65].\nFormation enthalpies of the materials were calculated\nwith respect to their elemental components as\n\u0001H=HCo3Mn2X\u00003HCo\u00002HMn\u0000HX;\nfor X= Al, Si, P, Ga, As, In, Sn, Sb, Tl, and Pb.\nIn this expression HCo3Mn2X,HCo,HMn, andHX\nare the enthalpies of formation for Co 3Mn2X, hexago-\nnal close-packed (hcp) cobalt, body-centered cubic (bcc)\nmanganese, and element X (such as, for example, cu-\nbic close-packed (ccp) aluminum for Co 3Mn2Al), respec-\ntively. Formation enthalpies with respect to Heusler al-\nloys Co 2MnX were obtained according to the following8\nformula\n\u0001H=HCo3Mn2X\u0000HCo2MnX\u0000HCo\u0000HMn;\nwhereHCo,HMn, andHCo2MnXare the enthalpies\nof hcp cobalt, bcc manganese, and the corresponding\nHeusler alloy.\nThe effect of chemical and magnetic disorder on the\nstability and magnetization of Co 3Mn2Ge was investi-\ngated by means of the coherent potential approximation\n[66, 67] as implemented in the Exact Muffin-Tin Or-\nbitals (EMTO) method [68, 69]. We used s,p,dand\nforbitals in the basis set. The one-electron equations\nwere solved within the soft-core and scalar-relativistic\napproximations. The Green’s function was calculated for\n16 complex energy points distributed exponentially on a\nsemi-circular contour including states within 1.1 Ry be-\nlow the Fermi level. For the one-center expansion of the\nfull charge density a lh\nmax=8 cutoff was used. The elec-\ntrostatic correction to the single-site coherent potential\napproximation was described using the screened impu-\nrity model [70] with a screening parameter of 0.6. Total\nenergies were calculated using the PBE [59] exchange-\ncorrelation functional, while local density approximation\n[71, 72] was used to calculate the magnetic moments and\nexchange interactions. The latter was calculated within\nthe magnetic force theorem [73] for the ferromagnetic\nand disordered local moment (DLM)[74, 75] configura-\ntions as implemented in the EMTO code. DLM repre-\nsentsthehightemperatureparamagnetic(PM)phase. In\nthis model, the paramagnetic phase of Co 3Mn2Ge reads\nas (Co\"\n0:5Co#\n0:5)3(Mn\"\n0:5Mn#\n0:5)2Ge. Similar formulation is\napplied to the paramagnetic phase of the alloy as well.\nB. Synthesis\nSamples of Co 3Mn2Ge were synthesized by melting Co\n(Alfa Aesar, 99.9%), Mn (Höganäs AB, 99.9%) and Ge\n(Kurt J. Lesker, 99.999%) together in an induction fur-\nnace under Ar (purity 99.999%) atmosphere. The re-\nsulting ingots were placed in Al 2O3crucibles, sealed in\nevacuated quartz glass tubes and annealed at 1073 K for\n14 days after which they were quenched in water. After\nthe final composition had been established by energy dis-\npersive X-ray spectroscopy analyses (see below), starting\nmaterials in stoichiometry of Co 52Mn34Ge14were pre-\npared using the established protocol to yield the final\nsamples. Samples were manually ground and powders\ntaken for analysis.\nC. Crystal structure analysis\nThe crystal structure was investigated using X-ray\npowder diffraction (XRPD), single crystal X-ray diffrac-\ntion (SCXRD) and scanning electron microscopy (SEM)\ncoupledwithenergydispersiveX-rayspectrocopy(EDS).The powders were mounted on single-crystal Si sample\nholders and X-ray diffraction patterns were collected us-\ningaBrukerD8AdvancewithmonochromatizedCu-K \u000b1\n(\u0015= 1.540598 Å) radiation at room temperature. Full-\nProf was used with the Rietveld refinement method to\nanalyse the data [76]. A Bruker D8 single-crystal X-ray\ndiffractometerwithMoK \u000bradiation(\u0015=0.71073Å)up-\ngraded with an Incoatec Microfocus Source (I \u0016S, beam\nsize\u0019100\u0016m at the sample position) and an APEX II\nCCD area detector (6cm \u00026cm) was utilized to collect\nSCXRD intensities at room temperature. SCXRD data\nreduction and numerical absorption corrections were per-\nformed using the APEX III software from Bruker[77].\nThe initial model of the crystal structure was first ob-\ntained with the program SHELXT-2014 and refined in\nthe program SHELXL-2014 within the APEX III soft-\nware package. The microstructure was evaluated with\na Zeiss Merlin SEM equipped with a secondary electron\n(SE) detector and an energy-dispersive X-ray spectrom-\neter. The samples for electron microscopy analysis were\nprepared by standard metallographic techniques through\ngrinding with SiC paper. For final polishing a mixture of\nSiO2and H 2O was used.\nD. Magnetic measurements\nMagnetization versus field and temperature measure-\nments were performed using a Quantum Design MPMS\nXL system. Isothermal magnetization curves were\nrecorded at several temperatures in applied magnetic\nfields up to 5 T. Magnetization measurements were per-\nformed on bulk samples as well as on single crystals. The\nsingle crystals used for these measurement were rather\nsmall hexagons, with a height of 100\u0016m and a side length\nof10\u0016m, resulting in a magnetic moment of 10\u00007–10\u00008\nAm2. Care was done to avoid common artifacts intro-\nduced when using this system [78]. The small hexagons\nwere too small to accurately measure the weight of the\nsample, and thus the hysteresis curves from single crys-\ntals were scaled to match the magnetization at the same\ntemperature for bulk samples. The temperature depen-\ndent magnetization was measured between 10 K and\n390 K in the applied magnetic fields of 0.01 T and 1\nT. The temperature dependent magnetization was also\nmeasured between 300 and 900 K and back to 300 K in\nan LakeShore VSM equipped with a furnace. The high\ntemperature measurements were performed in an applied\nmagnetic field of 0.01 T using a heating/cooling rate of\n3 K/min. The magnetization in SI units was calculated\nfromthemeasuredmagneticmomentbyusingthesample\nweight and density obtained from XRD measurements at\n298 K. The law of approach to saturation [44, 79] was\nused to calculate the effective anisotropy constant of the\nmaterial,jKeffjassuming the material to be uniaxial.9\nVIII. ACKNOWLEDGEMENT\nThe authors would like to acknowledge the support\nof the Swedish Foundation for Strategic Research, the\nSwedish Energy Agency (SweGRIDS), the Swedish Re-\nsearch Council, The Knut and Alice Wallenberg Foun-\ndation, STandUPP and the CSC IT Centre for Science,\nand the Swedish National Infrastructure for Computing\n(SNIC) for the computation resources. E. K. D.-Cz. ac-\nknowledges A. V. Ruban for valuable discussions. O.\nYu. V. acknowledges the support of Sweden’s Innovation\nAgency (Vinnova).\nIX. CONTRIBUTIONS\nO.E., H.C.H, P.S. and M.S. initiated the research.\nA.V. performed the high-throughput search and dataanalysis. V.S. and S.R.L synthesized the samples. V.S.\ncarried out the crystal structure analysis and SEM/EDS.\nD.H. performed magnetic measurements. DFT disor-\nder CPA calculations were performed and analysed by\nE.K.D.-C. and S.H. A.V., V.S., and D.H. drafted the\nmanuscript, which was reviewed and edited by all the\nauthors. All authors contributed to discussions and anal-\nysed the data.\nX. COMPETING INTERESTS\nThe authors declare no competing interests.\n[1] R. Skomski and J. M. D. Coey, Permanent magnetism\n(Institute of Physics Publishing, Bristol, UK, 1999).\n[2] H. i. d. I. Sözen, S. Ener, F. Maccari, K. P. Skokov,\nO. Gutfleisch, F. Körmann, J. Neugebauer, and\nT. Hickel, Phys. Rev. Materials 3, 084407 (2019).\n[3] T. Pandey and D. S. Parker, Phys. Rev. Applied 13,\n034039 (2020).\n[4] A.Gabay, G.Hadjipanayis, andJ.Cui,JournalofMag-\nnetism and Magnetic Materials 495, 165860 (2020).\n[5] S. Kim, H. Moon, H. Jung, S.-M. Kim, H.-S. Lee,\nH. Choi-Yim, and W. Lee, Journal of Alloys and Com-\npounds708, 1245 (2017).\n[6] H. Fang, S. Kontos, J. Ångström, J. Cedervall,\nP.Svedlindh, K.Gunnarsson, andM.Sahlberg,Journal\nof Solid State Chemistry 237, 300 (2016).\n[7] S. Shafeie, H. Fang, D. Hedlund, A. Nyberg,\nP.Svedlindh, K.Gunnarsson, andM.Sahlberg,Journal\nof Solid State Chemistry 274, 229 (2019).\n[8] B. Balasubramanian, B. Das, R. Skomski, W. Y. Zhang,\nandD.J.Sellmyer,AdvancedMaterials 13,6090(2013).\n[9] E. Lottini, A. López-Ortega, G. Bertoni, S. Turner,\nM. Meledina, G. Van Tendeloo, C. de Julián Fernán-\ndez, and C. Sangregorio, Chemistry of Materials 28,\n4214 (2016).\n[10] B. Balamurugan, B. Das, V. R. Shah, R. Skomski, X. Z.\nLi, and D. J. Sellmyer, Appl. Phys. Lett. 101, 122407\n(2012).\n[11] E. Anagnostopoulou, B. Grindi, L.-M. Lacroix, F. Ott,\nI. Panagiotopoulos, and G. Viau, Nanoscale 8, 4020\n(2016).\n[12] T. R. Gao, Y. Q. Wu, S. Fackler, I. Kierzewski,\nY. Zhang, A. Mehta, M. J. Kramer, and I. Takeuchi,\nAppl. Phys. Lett. 102, 022419 (2013).\n[13] T. J. Nummy, S. P. Bennett, T. Cardinal, and\nD. Heiman, Appl. Phys. Lett. 99, 252506 (2011).\n[14] B. Balamurugan, B. Das, W. Y. Zhang, R. Skomski,\nand D. J. Sellmyer, Journal of Physics: Condensed Mat-\nter26, 064204 (2014).[15] V. Ly, X. Wu, L. Smillie, T. Shoji, A. Kato, A. Manabe,\nand K. Suzuki, Journal of Alloys and Compounds 615,\nS285 (2014).\n[16] M. D. Kuz’min, K. P. Skokov, H. Jian, I. Radulov, and\nO. Gutfleisch, Journal of Physics: Condensed Matter\n26, 064205 (2014).\n[17] J. Cui, M. Kramer, L. Zhou, F. Liu, A. Gabay, G. Had-\njipanayis, B. Balasubramanian, and D. Sellmyer, Acta\nMaterialia 158, 118 (2018).\n[18] A. Vishina, O. Y. Vekilova, T. Björkman, A. Bergman,\nH. C. Herper, and O. Eriksson, Phys. Rev. B 101,\n094407 (2020).\n[19] A. G. Kusne, T. Gao, A. Mehta, L. Ke, M. C. Nguyen,\nK.-M. Ho, V. Antropov, C.-Z. Wang, M. J. Kramer,\nC. Long, and I. Takeuchi, Scientific Reports 4, 6367\n(2014).\n[20] Q. Gao, I. Opahle, O. Gutfleisch, and H. Zhang, Acta\nMaterialia 186, 355 (2020).\n[21] B. Fayyazi, K. P. Skokov, T. Faske, D. Y. Karpenkov,\nW. Donner, and O. Gutfleisch, Acta Materialia 141,\n434 (2017).\n[22] T. Burkert, L. Nordström, O. Eriksson, and\nO. Heinonen, Phys. Rev. Lett. 93, 027203 (2004).\n[23] G. Andersson, T. Burkert, P. Warnicke, M. Björck,\nB.Sanyal, C.Chacon,C.Zlotea, L.Nordström,P.Nord-\nblad, and O. Eriksson, Phys. Rev. Lett. 96, 037205\n(2006).\n[24] J. H. Park, Y. K. Hong, S. Bae, J. J. Lee, J. Jalli, G. S.\nAbo, N. Neveu, S. G. Kim, C. J. Choi, and J. G. Lee,\nJournal of Applied Physics 107, 09A731 (2010).\n[25] H. Fang, S. Kontos, J. Ångström, J. Cedervall,\nP.Svedlindh, K.Gunnarsson, andM.Sahlberg,Journal\nof Solid State Chemistry 237, 300 (2016).\n[26] T. Kojima, M. Ogiwara, M. Mizuguchi, M. Kotsugi,\nT. Koganezawa, T. Ohtsuki, T.-Y. Tashiro, and\nK. Takanashi, Journal of Physics: Condensed Matter\n26, 064207 (2014).\n[27] X. Zhao, C.-Z. Wang, Y. Yao, and K.-M. Ho, Phys.\nRev. B94, 224424 (2016).10\n[28] N. V. R. Rao, A. M. Gabay, and G. C. Hadjipanayis,\nJournal of Physics D: Applied Physics 46, 062001\n(2013).\n[29] S.Arapan, P.Nieves, H.C.Herper, andD.Legut,Phys.\nRev. B101, 014426 (2020).\n[30] “Inorganic crystal structure database,” http://www2.\nfiz-karlsruhe.de/icsd_home.html .\n[31] P. Bruno, Phys. Rev. B 39, 865 (1989).\n[32] K. Buschow, P. van Engen, and R. Jongebreur, Journal\nof Magnetism and magnetic materials 38, 1 (1983).\n[33] K. Adachi, K. Sato, M. Matsui, and Y. Fujio, Journal\nof the Physical Society of Japan 30, 1201 (1971).\n[34] Y. Kuz’ma and E. Gladyshevskii, Dopovidi Akademii\nNauk Ukrains’koi RSR 2, 205 (1963).\n[35] J. Cedervall, S. Kontos, T. C. Hansen, O. Balmes, F. J.\nMartinez-Casado, Z. Matej, P. Beran, P. Svedlindh,\nK.Gunnarsson, andM.Sahlberg,JournalofSolidState\nChemistry 235, 113 (2016).\n[36] M. A. McGuire, H. Cao, B. C. Chakoumakos, and B. C.\nSales, Physical Review B 90, 174425 (2014).\n[37] Y. He, G. H. Fecher, J. Kroder, H. Borrmann, X. Wang,\nL. Zhang, C.-Y. Kuo, C.-E. Liu, C.-T. Chen, K. Chen,\nF. Choueikani, P. Ohresser, A. Tanaka, Z. Hu, and\nC. Felser, Applied Physics Letters 116, 102404 (2020).\n[38] W. D. Corner and B. K. Tanner, Journal of Physics C:\nSolid State Physics 9, 627 (1975).\n[39] K. Tokuhara, Y. Ohtsu, F. Ono, O. Yamada,\nM. Sagawa, and Y. Matsuura, Solid State Communica-\ntions56, 333 (1985).\n[40] K. Krishnan, Fundamentals and Applications of Mag-\nnetic Materials (OUP Oxford, 2016).\n[41] N. Akulov, Zeitschrift für Physik 100, 197 (1936).\n[42] H. Callen and E. Callen, Journal of Physics and Chem-\nistry of Solids 27, 1271 (1966).\n[43] R. Frömter, H. Stillrich, C. Menk, and H. P. Oepen,\nPhys. Rev. Lett. 100, 207202 (2008).\n[44] J. F. Herbst and F. E. Pinkerton, Phys. Rev. B 57,\n10733 (1998).\n[45] X. Yan, X.-Q. Chen, A. Grytsiv, V. Witusiewicz,\nP. Rogl, R. Podloucky, and G. Giester, Journal of Al-\nloys and Compounds 429, 10 (2007).\n[46] S.Arapan, P.Nieves, H.C.Herper, andD.Legut,Phys.\nRev. B101, 014426 (2020).\n[47] X. Liu, D. Legut, R. Zhang, T. Wang, Y. Fan, and\nQ. Zhang, Phys. Rev. B 100, 054438 (2019).\n[48] V. Antropov, L. Ke, and D. Åberg, Solid State Com-\nmunications 194, 35 (2014).\n[49] Y. Miura, S. Ozaki, Y. Kuwahara, M. Tsujikawa,\nK. Abe, and M. Shirai, Journal of Physics: Condensed\nMatter25, 106005 (2013).\n[50] L. Ke and M. van Schilfgaarde, Phys. Rev. B 92, 014423\n(2015).\n[51] J. Zhang, B. Yang, H. Zheng, X. Han, and Y. Yan,\nPhys. Chem. Chem. Phys. 19, 24341 (2017).\n[52] J. M. Wills and B. R. Cooper, Phys. Rev. B 36, 3809\n(1987).\n[53] J. M. Wills, M. Alouani, P. Andersson, A. Delin,\nO. Eriksson, and O. Grechnyev, Full-Potential Elec-\ntronic Structure Method , Springer series in solid state\nscience (Springer, Berlin, Germany, 2010).\n[54] G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993).\n[55] G.KresseandJ.Hafner,Phys.Rev.B 49,14251(1994).\n[56] G. Kresse and J. Furthmüller, Comput. Mat. Sci. 6, 15\n(1996).[57] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169\n(1996).\n[58] P. E. Blöchl, Phys. Rev. B 50, 17953 (1994).\n[59] J. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.\nLett.77, 3865 (1996).\n[60] P. E. Blöchl, O. Jepsen, and O. K. Andersen, Phys.\nRev. B49, 16223 (1994).\n[61] G. H. O. Daalderop, P. J. Kelly, and M. F. H. Schuur-\nmans, Phys. Rev. B 41, 11919 (1990).\n[62] M. Weinert, R. E. Watson, and J. W. Davenport, Phys.\nRev. B32, 2115 (1985).\n[63] R. Skomski and J. Coey, Scripta Materialia 112, 3\n(2016).\n[64] O. Eriksson, A. Bergman, L. Bergqvist, and J. Hellsvik,\nAtomistic Spin Dynamics: Foundations and Applica-\ntions(Oxford University Press, Oxford, 2017).\n[65] Y. O. Kvashnin, O. Grånäs, I. Di Marco, M. I. Katsnel-\nson, A. I. Lichtenstein, and O. Eriksson, Phys. Rev. B\n91, 125133 (2015).\n[66] P. Soven, Phys. Rev. 156, 809 (1967).\n[67] B. L. Gyorffy, Phys. Rev. B 5, 2382 (1972).\n[68] L. Vitos, Phys. Rev. B 64, 014107 (2001).\n[69] L. Vitos, The EMTO Method and Applications, in Com-\nputational Quantum Mechanics for Materials Engineers ,\nEngineering Materials and Processes (Springer-Verlag\nLondon, 2007).\n[70] P. A. Korzhavyi, A. V. Ruban, I. A. Abrikosov, and\nH. L. Skriver, Phys. Rev. B 51, 5773 (1995).\n[71] J. P. Perdew and Y. Wang, Phys. Rev. B 45, 13244\n(1992).\n[72] D. M. Ceperley and B. J. Alder, Phys. Rev. Lett. 45,\n566 (1980).\n[73] A. I. Liechtenstein, M. I. Katsnelson, and V. A.\nGubanov, Journal of Physics F: Metal Physics 14, L125\n(1984).\n[74] A.J.Pindor, J.Staunton, G.M.Stocks, andH.Winter,\nJournal of Physics F: Metal Physics 13, 979 (1983).\n[75] B. L. Gyorffy, A. J. Pindor, J. Staunton, G. M. Stocks,\nand H. Winter, Journal of Physics F: Metal Physics 15,\n1337 (1985).\n[76] J. Rodriguez-Carvajal, IUCr Powder Diffr. Newsl. 26,\n12 (2001).\n[77] A. Saint, APEX3 software for CCD diffractometers\n(Bruker Analytical X-ray Systems Inc., Madison, WI,\n2014).\n[78] M. Sawicki, W. Stefanowicz, and A. Ney, Semiconduc-\ntor Science and Technology 26, 064006 (2011).\n[79] S. Chikazumi and C. Graham, Physics of Ferromag-\nnetism, International Series of Monographs on Physics\n(Clarendon Press, Oxford, UK, 1997).\n[80] J. H. van Vleck, Phys. Rev. 52, 1178 (1937).\n[81] P. Bruno, Phys. Rev. B 39, 865 (1989).\n[82] G. van der Laan, Journal of Physics: Condensed Matter\n10, 3239 (1998).\n[83] T. Kosugi, T. Miyake, and S. Ishibashi, Journal\nof the Physical Society of Japan 83, 044707 (2014),\nhttps://doi.org/10.7566/JPSJ.83.044707.\n[84] C. Andersson, B. Sanyal, O. Eriksson, L. Nordström,\nO. Karis, D. Arvanitis, T. Konishi, E. Holub-Krappe,\nand J. H. Dunn, Phys. Rev. Lett. 99, 177207 (2007).\n[85] A. Edström, Phys. Rev. B 96, 064422 (2017).\n[86] D. Wang, R. Wu, and A. J. Freeman, Phys. Rev. B 47,\n14932 (1993).11\n[87] E. Abate and M. Asdente, Phys. Rev. 140, A1303\n(1965).\n[88] H. Takayama, K.-P. Bohnen, and P. Fulde, Phys. Rev.\nB14, 2287 (1976).\n[89] T. Kanomata, H. Ishigaki, T. Suzuki, H. Yoshida,\nS. Abe, and T. Kaneko, Journal of Magnetism and\nMagnetic Materials 140-144 , 131 (1995), international\nConference on Magnetism.\n[90] S. Kaprzyk and S. Niziol, Journal of Magnetism and\nMagnetic Materials 87, 267 (1990).\n[91] W. Bazela, A. Szytuła, J. Todorović, Z. Tomkowicz,\nand A. Zięba, physica status solidi (a) 38, 721 (1976),\nhttps://onlinelibrary.wiley.com/doi/pdf/10.1002/pssa.2210380235.\n[92] N.T.Trung, L.Zhang,L.Caron,K.H.J.Buschow, and\nE. Brück, Applied Physics Letters 96, 172504 (2010).\n[93] T. Samanta, I. Dubenko, A. Quetz, S. Stadler, and\nN. Ali, Applied Physics Letters 101, 242405 (2012).\n[94] S. Ma, Y. Zheng, H. Xuan, L. Shen, Q. Cao, D. Wang,\nZ. Zhong, and Y. Du, Journal of Magnetism and Mag-\nnetic Materials 324, 135 (2012).[95] P. E. Markin, N. V. Mushnikov, V. I. Khrabrov, and\nM. A. Korotin, The Physics of Metals and Metallogra-\nphy106, 481 (2008).\n[96] M. Kogachi, T. Fujiwara, and S. Kikuchi, Journal of\nAlloys and Compounds 475, 723 (2009).\n[97] B. A. Collins, Y. S. Chu, L. He, D. Haskel, and F. Tsui,\nPhysical Review B 92, 224108 (2015).\n[98] A. Okubo, R. Y. Umetsu, K. Kobayashi, R. Kainuma,\nand K. Ishida, Applied Physics Letters 96, 222507\n(2010), https://doi.org/10.1063/1.3427431.\n[99] P. Webster, Journal of Physics and Chemistry of Solids\n32, 1221 (1971).\n[100] S. Nizioł, A. Wesełucha, W. Bażela, and A. Szytuła,\nSolid State Communications 39, 1081 (1981).\n[101] G. Venturini and B. Malaman, Journal of Alloys and\nCompounds 615, 886 (2014).\n[102] J.-T. Wang, D.-S. Wang, C. Chen, O. Nashima,\nT. Kanomata, H. Mizuseki, and Y. Kawazoe, Applied\nPhysics Letters 89, 262504 (2006).\n[103] Y. Fang, C. Yeh, C. Chang, W. Chang, M. Zhu, and\nW. Li, Scripta Materialia 57, 453 (2007).12\nAppendix A: Materials discarded during the final steps of high-throughput and data-mining search.\nSome of the materials with the magnetic moment above the set-up threshold of 1.0 \u0016B/f.u. were discarded later\ndue to the low or planar MAE or saturation magnetization being much lower than 1 T. These materials are listed in\nTable III.\nTABLE III: Systems with magnetic moment calculated to be higher than 1.0 \u0016B/f.u. during the high-throughput\nstep, which did not fulfill the requirements of a good permanent magnet on the further steps of data-mining.\nMaterial ICSD Space Mag. MAE (th) Sat. magn.\nnumber group state MJ/m3T\nCrNiAs 43913 189 FM 0.37 0.86\nCrNiP 43913 189 FM 0.21 0.83\nFe2Ti 103663 194 FM 0.10 0.84\nCoCrGe 409451 194 FM -0.18 0.75\nCoGeMn 623495 194 FM -0.16 1.10\nCu2Co(SnSe 4) 99296 121 FM 0.17\nCu2CoGeS 499293 121 FM 0.24\nCu2CoSiS 499292 121 FM 0.24\nCu2CoSnS 499294 121 FM 0.21\nMn2Co2C 44353 123 FM -2.21 0.43\nMnCoAs 610084 62 FM -0.88 0.90\nMnCoP 16483 62 FM -0.41 1.00\nNi5Sc 646468 191 FM 0.20\nGeMnSc 600156 189 FM 0.24 0.51\nSc2FeRh 5B251437 127 FM -0.13 0.39\nTiZn 2 106184 194 FM -0.01 0.30\nFeNiAs 610509 189 FM -1.53 0.6413\nAppendix B: Crystallographic data\nHere we list the crystallographic and structural data for the hexagonal Co 3:24Mn2Ge0:76compound along with some\nof the measurement details.\nTABLE IV: Crystallographic data and experimental details of the single crystal structure refinement for the\nhexagonal Co 3:24Mn2Ge0:76compound. Measurements were carried out at 296 K with Mo K \u000bradiation.\nEmpirical formula Co3Mn2Ge\nCalculated formula Co3:24Mn2Ge0:76\nStructure type MgZn 2\nFormula weight, Mr (g/mol) 177.92\nSpace group (No.) P63/mmc(194)\nPearson symbol, Z hP12, 4\nUnit cell dimensions:\na, Å 4.8032(2)\nb, Å 4.8032(2)\nc, Å 7.7378(4)\nV, Å 154.60(1)\nCalculated density, \u001a(g\u0001cm\u00003) 7.64\nAbsorption coefficient, \u0016(mm\u00001) 31.835\nTheta range for data collection (\u000e)4.901\u000437.119\nF(000) 324\nRange in h k l -8\u0014h\u00147,\n-8\u0014k\u00147,\n-13\u0014l\u001413,\nTotal No. of reflections 3933\nNo. of independent reflections 180(Req= 0.0264)\nNo. of reflections with I >2\u001b(I)174 (R\u001b= 0.0096)\nData/parameters 180/13\nWeighting details w=1/[\u001b2F2\no+0.5284P]\nwhere P=(F2\no+ 2F2\nC)/3\nGoodness-of-fit on F21.170\nFinalRindices [I>2\u001bI] R1= 0.0100;\nwR2= 0.0264\nRindices (all data) R1= 0.0111;\nwR2= 0.0268\nExtinction coefficient 0.00135(13)\nTABLE V: Atomic coordinates, anisotropic displacement parameters and selected interatomic distances for the\nhexagonal Co 3:24Mn2Ge0:76compound.\nAtom Site x y z U eq(Å2)\n0.837Co1+0.163Ge1 6 h0.17153(4) 0.34306(8) 1/4 0.00594(11)\nMn 4 f1/3 2/3 0.56358(7) 0.00817(16)\n0.740Co2+0.260Ge2 2 a 0 0 0 0.006667(17)\nU11 U22 U33 U12\n0.837Co1+0.163Ge1 0.00658(14) 0.00418(16) 0.00626(15) 0.00209(8)\nMn 0.00849(18) 0.00849(18) 0.0075(2) 0.00424(9)\n0.740Co2+0.260Ge2 0.0079(2) 0.0079(2) 0.0041(2) 0.00397(10)\nAtoms 1,2 d1,2(Å) Atoms 1,2 d1,2 (Å)\n2CojGe1–Co2jGe2 2.4039(2) 3Co1 jGe1–Mn1 2.7978(3)\n2Co1jGe1–Mn1 2.7815(5) 5Co2 jGe2–Mn1 2.8178(2)\nUeqis defined as one third of the trace of the orthogonalized Uijtensor. U13,U23= 0.14\nAppendix C: Ordering temperature, magnetic configuration, and crystal structure of the disordered\nCo3Mn 2Ge\nIn this section, we present the additional data for the ordered Co 3Mn2Ge and disordered Mn 2(Co0:75Ge0:25)4\nstructuresrelaxedwithEMTOcodeintheFM(Fig.8)andPM(Fig.9)phases. TableVIliststhecrystalstructureand\nmagnetic parameters obtained for these structures for both FM and PM magnetic configurations. Fig. 10 summarizes\nthe temperature effects for the FM and PM phases for the c/a-ratio fixed to the experimental value.\nTo calculate the temperature which makes the ordered and disordered state degenerate in the FM and PM phase,\nthe crystal structures of Co 3Mn2Ge and Mn 2(Co0:75Ge0:25)4were relaxed in volume and c/a ratio using the EMTO\ncode, as shown in Fig. 8 and Fig. 9. Table VI lists the crystal structure and magnetic parameters obtained for these\nstates for both FM and PM magnetic configurations. As can be seen, the chemical disorder does not have significant\neffect on the magnetic properties in the FM state. However, magnetic configuration (FM or DLM) does strongly\naffect the atomic magnetic moments, especially in the case of Co (Table VI).\nThe order-disorder transition can be estimated by the cross point of the free energies of the ordered and disordered\nstructures, i.e. \u0001F=Fdis-Ford. The temperature dependent free energy of the FM phase F(T)FMis estimated as\nF(T)FM\u0019EFM\n0+Fconf, whereEFM\n0is the internal energy for FM state at 0 K and Fconfis configurational free energy\nevaluated at different temperatures. The energy difference of Co 3Mn2Ge between the FM ordered and FM disordered\n(Mn 2(Co0:75Ge0:25)4) states is about 4.124 mRy; the difference in the configurational entropy of the alloy between\nthe ordered and disordered phases equals to 0.375 kB. SinceFconf= -TSconfwe find the transition temperature to be\naround 1735 K in the FM state.\nTemperature effects are summarized for the FM and PM phases in Figure 10 for c=a-ratio fixed to the experimental\nvalue. As we can see, there is no transition up to either the measured (359 K) or the theoretically predicted (700 K)\nmagnetic transition temperature (see left panel of Fig. 10). The ordered and disordered structures become degenerate\nin energy at\u00192300 K in the PM phase (see right panel of Fig. 10). The transition temperature is around 2200 K\nin the PM state. The stability of various antiferromagnetic configurations was investigated as well, for both ordered\nand disordered case (not shown), to account for the discrepancy between experimental and theoretical TCfor fixed\nc=a-ratio. None of the antiferromagnetic states considered in the calculations are stabilized by the configurational\nentropy up to the TC.\nThe Curie temperature for the ordered system and for Mn 2(Co0:75Ge0:25)4was estimated from Jij’s calculated for\nthe ferromagnetic reference state and performing subsequent Monte-Carlo (MC) simulations. We obtained 720 K for\nCo3Mn2Ge and 760 K for the disordered case using the theoretical lattice parameters given in Table VI. It is satisfying\nto note that the TCfor the ordered system, obtained with EMTO method, is in good agreement with the RSPt results.\nThe Curie temperature increase with disorder is due to the decrease of the antiferromagnetic Mn-Mn nearest neighbor\ninteractions as chemical disorder is applied. TCestimated for the experimental structure and composition given in\nTable II is 820 K. We canconcludethat the discrepancy between the predicted TCand the measured one does notcome\nfrom the order-disorder effect alone. To address this issue, DLM calculations were also performed. These result in the\nconsiderable drop of Co local magnetic moment in the DLM state compared to the FM moments, while Mn moment\ndoes not change (see Table VI). Magnetic moment of any atomic species, that is reduced in the DLM configuration\ncompared to a FM configuration, will lead to the reduced strength of the inter-atomic exchange interactions, and\nreduce of the ordering temperature.\nAppendix D: Origin of magnetic anisotropy in Co 3Mn 2Ge\nThe magnetocrystalline anisotropy originates from the SOC, since it is the only term in the Hamiltonian that\ncouples spin- and real-space, something that was first suggested by Van Vleck [80]. In the case of transition metals\nwhere SOC is much smaller than the bandwidth or crystal field, it can be treated as a perturbation. This lead to\nthe possibility to connect the MAE with anisotropy in orbital moment [81]. The original expression for this coupling,\nmade by Bruno, was subsequently extended, approximated, and used for various applications in [48, 50, 51, 82–86].\nIt has been shown previously [50, 51, 82, 85] that the coupling between the occupied and unoccupied states close\nto the Fermi energy \u000fFdominates the spin-orbit induced change of the total energy. Matrix elements h\u0016\u001bjL\u0001Sj\u00160\u001b0i\n[87, 88] determine the spin quantization axis direction which modifies the eigenvalues of the Kohn-Sham Hamiltonian.\nInthisexpression, \u0016representsad-orbital(withsymmetry fxy;yz;zx;x2\u0000y2;z2rg),\u001bdenotesspin, while LandSare\norbital and spin angular momentum operators. For states within the same spin channel, the couplings dxy\u0000 !dx2\u0000y2,\ndyz\u0000 !dxzpromote the uniaxial anisotropy, while dxy\u0000 !dxz,dxy\u0000 !dyz,dxz\u0000 !dz2,dyz\u0000 !dz2,dxz\u0000 !dx2\u0000y2, and\ndyz\u0000 !dx2\u0000y2favour the easy-plane magnetocrystalline anisotropy [48, 50, 51, 82, 83, 85]. The situation is reversed\nfor the couplings between opposite spin channels. Table VIII (in Appendix E ) lists the transitions that contribute\neither to easy-plane or uniaxial anisotropy along with their relative weights.15\nFIG. 8: (Color online) Volume and c=arelaxation of the ordered Co 3Mn2Ge (left) and disordered (right)\nMn2(Co0:75Ge0:25)4structures in the FM phase. rWSdenotes the Wigner-Seitz radius. Yellow lines follow the\nminimum of E( rWS) curve for a specific value of c=aand the minimum of E(c=a)curve for each rWS.\nFIG. 9: (Color online) Volume and c=arelaxation of the ordered Co 3Mn2Ge (left) and disordered (right)\nMn2(Co0:75Ge0:25)4structures in PM phase. rWSdenotes the Wigner-Seitz radius. Yellow lines follow the minimum\nof E(rWS) curve for a specific value of c=aand the minimum of E(c=a)curve for each rWS.\nTo understand the origin of the difference in MAE for Co 3Mn2X (X = Al, Si, P, Ga, As, In, Sn, Sb, Tl, and Pb), it\nis electron states close to the Fermi level ( \u000fF) one should focus on, since spin-orbit interaction that couples states just\nbelow\u000fFto states just above \u000fF, are particularly important in deciding the magnetic anisotropy [22]. To undertake\nthis analysis we inspected partial densities of states (pDOS) around \u000fF. Figure 11 shows the pDOS for Co 3Mn2Ge,\nwhich has a large uniaxial anisotropy of 1.44 MJ/m3, and Fig. 12 shows the pDOS for Co 3Mn2As, that has a large\neasy-plane anisotropy of -1.2 MJ/m3. The pDOS curves for the other Co 3Mn2X materials can be found in Appendix F ,\nFigs. 13-14. The majority spin channel of the d-states is essentially fully occupied for all the materials considered in\nthis work which shows its inertness for the magnetic anisotropy. Instead significant contributions are expected for the\nminority spin channel, that has states on either side of \u000fF. In this spin channel there are large peaks corresponding\nto thedyz,dxz, anddxystates of Co, close to the Fermi energy. These peaks are mostly empty for X = Al, Ga, Ge,\nIn, and Sn, while they lie directly on \u000fFfor X = Si and Sb, and are mostly occupied for X = P, As. All the materials\nwith uniaxial magnetic anisotropy (X = Al, Ga, Ge, In) have these large peaks of the DOS occupied, and we conclude\nthat from a microscopic point of view, these electrons are decisive for the magnetic anisotropy.\nTo analyze the MAE further we consider the difference in SOC energies with spin orientation along the zandx\naxes, \u0001Eso=Ez\nso\u0000Ex\nso, separately for all Co and Mn atoms, see Table VII (negative sign marks a contribution to\nuniaxial magnetic anisotropy). As expected, for most of X,\u0001Eso(Mn) is considerably smaller than that of Co atoms.\nThe latter can be divided into two groups (noted by the subscripts in Table VII); the contribution to \u0001Esofrom the16\nTABLE VI: Optimized crystal structure parameters ( a,c=a, and volume) and magnetic moments per unit cell\ncalculated for the ordered Co 3Mn2Ge and disordered Mn 2(Co0:75Ge0:25)4phases using the EMTO code, along with\nthe low temperature experimental lattice parameters and magnetic saturation field. Third column contains\nexperimental data. Fourth column (labeled ”Co-excess”) contains experimental data for the low temperature\nstructure, and calculated magnetic moments for this structure and for a composition given in Table II (54 % Co).\nOrdered Disordered Experiment Co-excess\nFM\na,\u0017A 4.83 4.81 4.803 4.803\nc=a 1.61 1.65 1.611 1.611\nV,\u0017A3157.6 158.5 154.6 154.5\nm6h\nCo,\u0016B1.60 1.59 1.61\nmMn,\u0016B3.30 3.28 3.27\nm2a\nCo,\u0016B 1.56 1.63\nMtot/cell, T 1.66 1.65 0.86 1.75\nDLM\na,\u0017A 4.84 4.77 4.803\nc=a 1.56 1.65 1.611\nV,\u0017A3152.7 154.9 154.5\nm6h\nCo,\u0016B0.04 0.41 0.60\nmMn,\u0016B2.92 2.99 3.05\nm2a\nCo,\u0016B 0.05 0.51\nFIG. 10: (Color online) Left panel: Normalized total energies ( E0) (full lines) and free energies ( F(T)) (dashed and\ndotted lines) calculated for the ordered (black squares) and disordered (red circles) FM state of Co 3Mn2Ge at 0 K,\n359 K (experimental TC) and 700 K (theoretical TC). Right panel: Normalized total energies ( E0) (full lines) and\nfree energies ( F(T)) calculated for the ordered (black squares) and disordered (red circles) Co 3Mn2Ge at 0 K and\n2300 K in the PM state. Dashed line stands for F(T)=E0+Fconf, dotted line denotes F(T)=E0+Fconf+Fmag.\nOn the x-axis, rWSdenotes the Wigner-Seitz radius calculated as rWS=3p\n3Vatom=4=\u0019, whereVatomis the average\nvolume of an atom in the unit cell. Fconf= -kBTP\niciln(ci) andFmag=-kBTP\niciln(1+mi).\ntwo Co atoms positioned at (0.667, 0.833, 0.75) and (0.333, 0.166, 0.25) 6 h(we denote them type 2) is different to\nthat of the remaining four Co atoms ( type 1). All Co atoms contribute to the uniaxial magnetic anisotropy for X =\nAl, Ga, and In. The values of \u0001Esoare extremely low for both Co and Mn atoms in Co 3Mn2P, which results in the\nnegligibly small value of MAE. In the case of Co 3Mn2Si and Co 3Mn2As, all Co atoms give rise to the large values\ncorresponding to easy-plane magnetocrystalline anisotropy. For the remaining materials, the two groups of Co atoms\nhave opposite signs of \u0001Eso.\nWe can also determine the d-orbitals that give the largest change to the SOC matrix element h\u0016\u001bjbHsoj\u00160\u001b0ias\nthe magnetization direction changes from ztox. Co 3Mn2X with X = Al, Ga, Ge, and In, that exhibit uniaxial\nanisotropy, all present a similar picture, see Fig. 19 ( Appendix G ) for Co 3Mn2Ge (the rest of the figures can be17\nFIG. 11: (Color online) Contribution to the density of states of Co 3Mn2Ge from the dx2\u0000y2anddz2(egset);dxy,\ndyz, anddxz(t2gset) orbitals without SOC interaction. Fermi energy is set at zero.\nFIG. 12: (Color online) Contribution to the density of states of Co 3Mn2As from the dx2\u0000y2anddz2(egset);dxy,\ndyz, anddxz(t2gset) orbitals without SOC interaction.\nfound in Appendix G , Fig. 15-21). The main contribution to the uniaxial anisotropy comes from hdx2+y2jbHsojdxyi\nfor all the Co atoms, even though there are additional contributions which are different for the Co atoms of type 1\n(Fig. 19, top) and type 2(Fig. 19, bottom). As expected (Table VIII), for these transitions to contribute to uniaxial\nanisotropy the states must be within the same spin channel. It is more difficult to distinguish any specific transitions\nfor materials with the easy-plane anisotropy, see Fig. 20 ( Appendix G ) for Co 3Mn2As (the other materials can be\nfound in Appendix G , Fig. 15-21) - there are several matrix elements promoting the negative sign in MAE. The most\nsignificant arehdx2+y2jbHsojdyzi,hdxzjbHsojdxyi, andhdz2jbHsojdyziand their size varies depending on X.\nAppendix E: d-orbitals contribution to MAE\nMatrix elements h\u0016\u001bjL\u0001Sj\u00160\u001b0i[87, 88] determine the preferable direction of spin quantization axis. Table VIII\nlists the transitions between the states below and above \u000fFwhich favour either z-axis (uniaxial) or xy-plane (planar)\nmagnetic anisotropy along with the relative weight of each transition.18\nTABLE VII: Calculated MAE (RSPt), saturation magnetization (RSPt), the average magnetic moment per Co and\nMn atoms, Curie temperature (UppASD), formation enthalpies with respect to the elemental components for the\nrelaxed Co 3Mn2X structures ( \u0001Hel), and formation enthalpies with respect to the Co 2MnX Heusler compounds (X\n= Al, Si, P, Ga, Ge, As, In, Sn, Sb, Tl, and Pb) (VASP), marked \u0001HHeus. The last three columns contain the\ndifference in SOC energy with spin orientation along zandxaxis (negative sign corresponds to uniaxial magnetic\nanisotropy) for Mn and Co atoms (VASP). Subscript 2notes two cobalt atoms positioned at (0.667, 0.833, 0.75) and\n(0.333, 0.166, 0.25) 6 hsites, subscript 1points at the remaining four Co atoms.\nMaterial MAE Sat. M (Co) M (Mn) TC\u0001Hel\u0001HHeus \u0001Eso(Mn) \u0001Eso(Co1)\u0001Eso(Co2)\nMJ/m3magn., T\u0016B\u0016BK eV/f.u. eV/f.u. meV meV meV\nCo3Mn2Al 1.38 1.77 1.45 3.37 820 -1.78 0.274 0.016 -0.42 -0.46\nCo3Mn2Si -0.64 1.63 1.14 3.32 -2.34 0.090 -0.057 0.11 1.20\nCo3Mn2P 0.047 1.50 0.76 3.35 -2.62 0.018 -0.04 0.10\nCo3Mn2Ga 0.67 1.75 1.50 3.47 800 -1.49 0.061 0.100 -0.25 -0.30\nCo3Mn2Ge 1.44 1.71 1.44 3.52 700 -1.57 0.070 0.005 -0.26 0.13\nCo3Mn2As -1.2 1.58 1.08 3.50 -1.60 -0.011 0.01 1.13\nCo3Mn2In 0.36 1.65 1.60 3.62 -0.14 0.160 -0.11 -0.25\nCo3Mn2Sn -0.42 1.63 1.53 3.59 -0.57 0.064 -0.21 0.03\nCo3Mn2Sb -0.81 1.56 1.35 3.56 -0.67 0.047 -0.19 0.68\nCo3Mn2Tl -0.21 1.63 1.63 3.67 0.87\nCo3Mn2Pb -2.7 1.58 1.58 3.66 0.93\nTABLE VIII: Transitions between the d-orbitals below and above \u000fFwhich contribute to either z-axis (uniaxial) or\nxy-plane (planar) magnetic anisotropy along with the relative weight of each transition, !\nContributes Same spin ! Opposite spin !\ntransition transition\nUniaxial dxy\u0000 !dx2\u0000y2 1dxy\u0000 !dxz,dxy\u0000 !dyz0.25\nanisotropy dyz\u0000 !dxz 0.25dxz\u0000 !dz2,dyz\u0000 !dz20.75\ndxz\u0000 !dx2\u0000y2,dyz\u0000 !dx2\u0000y20.25\nxy-plane dxy\u0000 !dxz,dxy\u0000 !dyz0.25 dxy\u0000 !dx2\u0000y2 1\nanisotropy dxz\u0000 !dz2,dyz\u0000 !dz20.75 dyz\u0000 !dxz 0.25\ndxz\u0000 !dx2\u0000y2,dyz\u0000 !dx2\u0000y20.25\nAppendix F: DOS for Co 3Mn 2X (X = Al, Si, P, Ga, In, Sn, Sb)\nHereweprovidetheDOSforthecrystalstructuresCo 3Mn2X(X=Al, Si, P,Ga, In, Sn, Sb, Tl, andPb). Thesewere\ncalculated by replacing Ge in Co 3Mn2Ge structure by the neighboring elements and relaxing their crystal structures.\nDOS of Co 3Mn2Ge and Co 3Mn2As are given above in the text.19\n(a) Density of states of Co 3Mn2Al (MAE = 1.38\nMJ/m3).\n(b) Density of states of Co 3Mn2Si (MAE = -0.64\nMJ/m3).\n(c) Density of states of Co 3Mn2P (MAE = 0.047\nMJ/m3).\n(d) Density of states of Co 3Mn2Ga (MAE = 0.67\nMJ/m3).\n(e) Density of states of Co 3Mn2In (MAE = 0.36\nMJ/m3).\n(f) Density of states of Co 3Mn2Sn (MAE = -0.42\nMJ/m3).\nFIG. 13: (Color online) Density of states of Co 3Mn2X, with X = Al, Si, P, Ga, Ge, As, In, Sn, Sb\nAppendix G: Change of the SOC matrix elements of Co when magnetization changes direction from ztoxfor\nCo3Mn 2X (X = Al, Si, P, Ga, In)\nTo determine the main orbital contribution to MAE we calculate the change in SOC matrix element h\u0016\u001bjbHsoj\u00160\u001b0i\nfor transitions between different d-orbitals as magnetization direction changes from ztox. Fig. 15-21 show contribu-\ntions from type 1Co atoms (top) and type 2Co atoms (bottom).20\nFIG. 14: (Color online) Density of states of Co 3Mn2Sb (MAE = -0.81 MJ/m3).\nFIG. 15: (Color online) Change of spin-orbit coupling matrix elements of Co, type 1(left), and Co, type 2(right) in\nCo3Mn2Al when magnetization changes direction from ztox.\nFIG. 16: (Color online) Change of spin-orbit coupling matrix elements of Co, type 1(left), and Co, type 2(right) in\nCo3Mn2Si when magnetization changes direction from ztox.\nAppendix H: The properties of the previously reported phases for Co-Mn-Ge system\nThe previously reported crystallographic phases for the Co–Mn–Ge system are presented in Table IX together with\ntheir crystal structure and some of the magnetic properties. We will briefly describe some of the key features of the\nstructural and magnetic properties of Co–Mn–Ge systems reported in the literature, which are expected to be relevant21\nFIG. 17: (Color online) Change of spin-orbit coupling matrix elements of Co, type 1(left), and Co, type 2(right) in\nCo3Mn2P when magnetization changes direction from ztox.\nFIG. 18: (Color online) Change of spin-orbit coupling matrix elements of Co, type 1(left), and Co, type 2(right) in\nCo3Mn2Ga when magnetization changes direction from ztox.\nFIG. 19: (Color online) Change of spin-orbit coupling matrix elements of Co, type 1(top), and Co, type 2(bottom)\nin Co 3Mn2Ge when magnetization changes direction from ztox.\nto the Co 3Mn2Ge. It should be noted that this outline is notintended as a review, but rather as a summary of the\nknown properties of Co–Mn–Ge phases, which we find relevant to understanding the properties of Co 3Mn2Ge.\nCoMnGe exists in two stable phases; the low-temperature orthorhombic CoMnGe (TiNiSi type) is stable below the\nmartensitic temperature ( TM) of 650 K [89] while the high-temperature hexagonal CoMnGe (BeZrSi type) is stable\nabove it [89]. The reported TC-values of the high- and low-temperature phases vary slightly, for instance, Kaprzyk and\nNiziol [90] report TC= 337 K for CoMnGe (TiNiSi) and TC= 287 K for CoMnGe (BeZrSi). As it is possible to tune22\nFIG. 20: (Color online) Change of spin-orbit coupling matrix elements of Co, type 1(top), and Co, type 2(bottom)\nin Co 3Mn2As when magnetization changes direction from ztox.\nFIG. 21: (Color online) Change of spin-orbit coupling matrix elements of Co, type 1(left), and Co, type 2(right) in\nCo3Mn2In when magnetization changes direction from ztox.\nTMto coincide with TC[89, 102, 103], a large/giant magnetocaloric effect can be achieved at the room temperature\n[92–94].\nCoGeMn is reported to exhibit a collinear [91] magnetic state. However, depending on the Co:Mn ratio,\nCoxMn1\u0000xGe samples can possess not only an easy-axis or easy-plane magnetic anisotropy but also an interme-\ndiate state [95], i.e. spin reorients with Co:Mn ratio. The substitution of Ni (Co xNi1\u0000xMnGe) produces the even\nmore diverse results; the samples can be AFM, collinear FM, and non-collinear FM, depending on the amount of Ni\nand the annealing temperature [100].\nCo2MnGe is a full Heusler compound which exhibits some common features with the prototypical full Heusler\nCu2MnAl. These include a high TCand certain peculiarities related to its structural and magnetic properties derived\nfrom the range of chemical environments, disorder, vacancies, and stacking faults. Using anomalous X-ray diffraction\ntechniques Collins et al.[97] were able to demonstrate that thin films of Co 2MnGe grown on Ge (111) show the\npreference of Mn–Ge site swapping along with the small ( <0:5%) site-interchange of Co–Mn. The authors were also\nable to show that over–stoichiometry of Ge leads to the decrease in chemical disorder, vacancies, and stacking faults.\nIt is important to note, that they were not able to grow the Co 3Mn2Ge hexagonal phases due to a too large lattice\nmismatch with the Ge (111) surface.\nKogachi et al.[96] showed that the increase in sample quenching temperature gives rise to Mn–Ge disorder and\nresults in a lower magnetization at 4.2 K. Okubo et al.[98] were able to demonstrate that the aforementioned disorder\nbrings on a considerable decrease in TC. Webster [99], on the other hand, had earlier reported that there was no\nsign of chemical disorder in Co 2MnGe and that magnetic moment originated primarily from Mn (3.58 \u0016B); Co only\ncontributes 0.75 \u0016Bonly. Some of these ”discrepancies” reported previously are thus likely due to the difference in the\npreparation of the samples, but they also show that Co 2MnGe exists in a number of states ranging from a collinear\nferromagnet with a high TCto a virtually non-magnetic material, depending on the chemical disorder.23\nTABLE IX: Co–Mn–Ge phases reported previously, with their crystal structures and transition temperatures.\nCompound Entry prototype, Remarks\nSGR Symbol and number and comments\nCoMnGe TiNiSi, Pnma(62) Low-temperature phase, stable below 650 K [89];\nTC= 337 [90].\nCoMnGe BeZrSi, P63/mmc(194) High-temperature phase, stable above 650 K [89];\ncollinear [91] with TC= 283 K [90], 334 K [89].\nTransformation with TM=TCleads to large/giant magnetocaloric\neffect [92–94].\nCoxMn1\u0000xGe Ni 2In,P63/mmc(194) Spin reorientation depending on the amount of Co; can be easy axis,\neasy plane, as well as hard/easy equally in all directions [95].\nCo2MnGe Cu 2MnAl,Fm\u00163m(225) The increase in quenching temperature leads to increase in Mn-Ge\ndisorder [96]. Disorder causes the decrease in the magnetization [96];\nhigher quenching temperature rates produce lower magnetization [96].\nHigh Mn-Ge disorder, thin films [97], enrichment of 5 at % of Ge in\nCo0:5Mn0:25Ge0:25gives the highest degree of chemical ordering [97].\nChemical disorder in Co 2MnGe decreases TC[98].\nStrong ordering in Co 2MnGe [99], almost all magnetic moment comes\nfrom Mn (3.58 \u0016Bvs. 0.75\u0016Bfor Co).\nCoxNi1\u0000xMnGe TiNiSi, Pnma(62) Samples can be either AFM or FM or non-collinear FM depending on\nxand the annealing temperature [100].\nCo3Mn2Ge Mg 2Cu3Si,P63/mmc(194) Ordered crystallographic model [34]\nCo3Mn2Ge MgZn 2,P63/mmc(194) Disordered crystallographic model [34]\nCo3Ge5Mn9Mn9Co3Ge5,R32h, (155) No magnetic properties reported. [101]\nLastly, Co 3Ge5Mn9is reported to be a rhombohedral phase with no magnetic properties reported [101]." }, { "title": "2102.01283v2.On_the_relationship_between_orbital_moment_anisotropy__magnetocrystalline_anisotropy__and_Dzyaloshinskii_Moriya_interaction_in_W_Co_Pt_trilayers.pdf", "content": "arXiv:2102.01283v2 [cond-mat.mtrl-sci] 13 Aug 2022On the relationship between orbital moment anisotropy, mag netocrystalline\nanisotropy, and Dzyaloshinskii-Moriya interaction in W/C o/Pt trilayers\nZhendong Chi,1,∗Yong-Chang Lau,1,2,†Vanessa Li Zhang,3Goro Shibata,1,4Shoya\nSakamoto,1Yosuke Nonaka,1Keisuke Ikeda,1Yuxuan Wan,1,5Masahiro Suzuki,1Masashi\nKawaguchi,1Masako Suzuki-Sakamaki,6,7Kenta Amemiya,6Naomi Kawamura,8Masaichiro\nMizumaki,8Motohiro Suzuki,8Hyunsoo Yang,9Masamitsu Hayashi,1,2and Atsushi Fujimori1,10\n1Department of Physics, The University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan\n2National Institute for Materials Science, Tsukuba, Ibarak i 305-0047, Japan\n3School of Physics and Technology, Wuhan University, Wuhan 4 30072, China\n4Materials Sciences Research Center, Japan Atomic Energy Ag ency, Sayo, Hyogo 679-5148, Japan\n5Institute for Solid State Physics, The University of Tokyo, Kashiwa, Chiba 277-8581, Japan\n6Institute of Materials Structure Science, High Energy Acce lerator Research Organization, Tsukuba, Ibaraki 305-0801 , Japan\n7Graduate School of Science and Technology, Gunma Universit y, Kiryu, Gunma 376-8515, Japan\n8Japan Synchrotron Radiation Research Institute (JASRI), S ayo 679-5198, Japan\n9Department of Electrical and Computer Engineering,\nNational University of Singapore, Singapore 117576, Singa pore\n10Department of Physics, National Tsing Hua University, Hsin chu 30013, Taiwan\n(Dated: August 16, 2022)\nWe have studied the Co layer thickness dependences of magnet ocrystalline anisotropy (MCA),\nDzyaloshinskii-Moriya interaction (DMI), and orbital mom ent anisotropy (OMA) in W/Co/Pt tri-\nlayers, in order to clarify their correlations with each oth er. We find that the MCA favors magne-\ntization along the film normal and monotonically increases w ith decreasing effective magnetic layer\nthickness ( teff). The magnitude of the Dzyaloshinskii-Moriya exchange con stant (|D|) increases with\ndecreasing teffuntilteff∼1 nm, below which |D|decreases. The MCA and |D|scale with 1 /tefffor\ntefflarger than ∼1 nm, indicating an interfacial origin. The increase of MCA w ith decreasing teff\ncontinues below teff∼1 nm, but with a slower rate. To clarify the cause of the teffdependences\nof MCA and DMI, the OMA of Co in W/Co/Pt trilayers is studied us ing x-ray magnetic circular\ndichroism (XMCD). We find non-zero OMA when teffis smaller than ∼0.8 nm. The OMA increases\nwith decreasing teffmore rapidly than what is expected from the MCA, indicating t hat factors other\nthan OMA contribute to the MCA at small teff. Theteffdependence of the OMA also suggests that\n|D|atteffsmaller than ∼1 nm is not related to the OMA at the interface. We propose that the\ngrowth of Co on W results in a strain and/or texture that reduc es the interfacial DMI, and, to some\nextent, MCA at small teff.\nI. INTRODUCTION\nUltrathin film heterostructures that consist of ferro-\nmagnetic metal (FM) layers and non-magnetic heavy\nmetal (HM) layersareattracting great interest as various\nnovel phenomena that originate from the strong spin-\norbit coupling in bulk and at interfaces have been dis-\ncovered. For example, efficient current-induced magne-\ntization reversal [1] and fast motion of magnetic domain\nwalls[2]havebeendemonstratedinheterostructureswith\nperpendicular magnetic anisotropy(PMA), which are es-\nsential in ultra-high-density magnetic memories. These\nphenomena are attributed to spin-orbit coupling-induced\neffects such as spin Hall effect [3–6], Rashba-Edelstein ef-\nfect [7, 8], and Dzyaloshinskii-Moriya interaction (DMI)\n[9, 10]. Among these effects, strong DMI is especially\nnecessary for racetrack memories because it stabilizes\n∗Present address: CIC nanoGUNE, 20018 Donostia-San Sebas-\ntian, Basque Country, Spain\n†Present address: Institute of Physics, Chinese Academy of S ci-\nences, Beijing 100190, Chinachiral N´ eel domain walls and skyrmions [11–21]. There-\nfore, solid understanding of these interfacial phenomena\nis essential to develop spintronic devices with significant\nPMA and DMI.\nRecently, the microscopic origins of PMA and interfa-\ncial DMI have been discussed in relation to the orbital\nmoment anisotropy (OMA) [22, 23] and the magnetic\ndipole moment in the FM layer [24, 25]. As OMA should\nexist both in the FM and HM elements [26], it is of high\nimportance to identify what role the OMA (of FM and\nHM elements) plays in PMA and DMI in FM/HM het-\nerostructures.\nHere, we study correlation between the magnetocrys-\ntalline anisotropy (MCA), DMI and OMA in W/Co/Pt\ntrilayers. The Co layer thickness dependences of MCA\nand DMI in W/Co/Pt trilayers are studied using vi-\nbrating sample magnetometer (VSM) and Brillouin light\nscattering spectroscopy (BLS), respectively. As for the\nOMA, we study the Co layer thickness dependences\nof the spin and orbital magnetic moments of Co in\nW/Co/Pt trilayers, where W works as a seed layer, and\nthe proximity-induced magnetization in W and Pt in\nW/Co and Pt/Co bilayers using x-ray magnetic circu-2\nlar dichroism (XMCD). We find that the MCA, DMI,\nand OMA of Co show different Co layer thickness de-\npendences in the W/Co/Pt trilayers. The origin of these\nobservations will be discussed.\nII. EXPERIMENT\nCo thin films sandwiched by W and Pt, i.e. Sub./3\nW/tCoCo/1 Pt/1 Ru (the numbers denote the nominal\nthicknesses in nm) were grown on 10 ×10 mm2thermally\noxidized Si substrates by magnetron sputtering at room\ntemperature in a base pressure better than 5 ×10−7Pa.\nThe top Ru layer is used to protect the trilayers from\noxidation. The Co layer thickness ( tCo) was varied from\n0.6 to 1.7 nm (0.6, 0.7, 0.8, 1.0, 1.1, 1.2, 1.3, 1.4, and 1.7\nnm) in different samples. The magnetic hysteresis loops\n(in-plane and out-of-plane)ofthe samplesweremeasured\nusing a VSM at room temperature. The magnetic field is\napplied up to 1.6T during the measurement. The magni-\ntude of DM exchange constant ( |D|) was investigated by\nBLS, with the same measurement setup in our previous\nstudies [27].\nX-ray absorption spectroscopy (XAS) and XMCD\nmeasurements at the Co L3,2edges were performed using\nsoft x rays at the helical undulator beamline BL-16A1\nof Photon Factory, High Energy Accelerator Research\nOrganization (KEK-PF). The spectra were measured in\nthe total electron yield (TEY) mode. The measurements\nwere performed at room temperature in a vacuum bet-\nter than 5 ×10−7Pa. The magnitude of the magnetic\nfield was set at 5 T and the field was applied parallel to\nthe incident x rays in all measurements. The XAS and\nXMCD measurements using hard x rays were conducted\nat BL39XU of SPring-8. The partial fluorescence yield\n(PFY)andx-raypolarizationswitchingmodeswereused.\nThe measurements were performed at atmospheric pres-\nsure and at room temperature. A magnetic field of up to\n2 T was applied during the measurements. In order to\nobtain the out-of-plane and in-plane components of the\nmagnetic moments, the magnetic field was applied to the\nsample along the film normal and 30◦with respect to the\nsamplesurface, referredto asout-of-planeand “in-plane”\nmagneticfieldshereafter. NotethatthepositionofthePt\nL3edge (11.563keV) is excessivelyclose to that of the W\nL2edge (11.544 keV): these two peaks will overlap with\neach other in the W/Co/Pt trilayer and complicate the\ndata analysis. Thus, two W/Co and Pt/Co bilayers were\nprepared by magnetron sputtering for measuring the W\nand PtL3,2-edge XAS and XMCD spectra. The bilayer,\ni.e. Sub./0.6 W/0.8 Co/1 Ru and Sub./0.6 Pt/0.8 Co/1\nRu, were also grown on 10 ×10 mm2thermally oxidized\nSi substrates by magnetron sputtering at room temper-\nature and the 1-nm-thick Ru served as the capping to\navoid the bilayer from surface oxidation./s32/s33/s34/s35/s36/s32/s33/s34/s35/s33/s32/s33/s34/s33/s36/s33/s33/s34/s33/s36/s33/s34/s35/s33/s33/s34/s35/s36/s37/s38/s39/s40/s41/s42/s43/s44/s35/s33/s32/s45/s43/s40/s39/s46/s47\n/s32/s48/s33 /s32/s35/s33 /s33 /s35/s33 /s48/s33\n/s32/s43/s44/s49/s50/s40/s47/s32/s33/s34/s35/s36/s37/s38/s33/s39/s40/s35/s41/s33/s42/s43\n/s43/s51/s41/s32/s52/s53/s54/s41/s40\n/s43/s38/s46/s42/s32/s38/s55/s32/s52/s53/s54/s41/s40\n/s32/s33/s34/s35/s36/s32/s33/s34/s35/s33/s32/s33/s34/s33/s36/s33/s33/s34/s33/s36/s33/s34/s35/s33/s33/s34/s35/s36/s37/s38/s39/s40/s41/s42/s43/s44/s35/s33/s32/s45/s43/s40/s39/s46/s47\n/s32/s48/s33 /s32/s35/s33 /s33 /s35/s33 /s48/s33\n/s32/s43/s44/s49/s50/s40/s47/s32/s33/s34/s35/s36/s37/s44/s33/s39/s40/s35/s41/s33/s42/s43\n/s43/s51/s41/s32/s52/s53/s54/s41/s40\n/s43/s38/s46/s42/s32/s38/s55/s32/s52/s53/s54/s41/s40\n/s32/s33/s34/s48/s32/s33/s34/s35/s33/s33/s34/s35/s33/s34/s48/s37/s38/s39/s40/s41/s42/s43/s44/s35/s33/s32/s45/s43/s40/s39/s46/s47\n/s32/s48/s33 /s32/s35/s33 /s33 /s35/s33 /s48/s33\n/s32/s43/s44/s49/s50/s40/s47/s32/s33/s34/s35/s41/s37/s36/s33/s39/s40/s35/s41/s33/s42/s43\n/s43/s51/s41/s32/s52/s53/s54/s41/s40\n/s43/s38/s46/s42/s32/s38/s55/s32/s52/s53/s54/s41/s40\n/s32/s33/s34/s48/s32/s33/s34/s35/s33/s33/s34/s35/s33/s34/s48/s37/s38/s39/s40/s41/s42/s43/s44/s35/s33/s32/s45/s43/s40/s39/s46/s47\n/s32/s48/s33 /s32/s35/s33 /s33 /s35/s33 /s48/s33\n/s32/s43/s44/s49/s50/s40/s47/s32/s33/s34/s35/s41/s37/s45/s33/s39/s40/s35/s41/s33/s42/s43\n/s43/s51/s41/s32/s52/s53/s54/s41/s40\n/s43/s38/s46/s42/s32/s38/s55/s32/s52/s53/s54/s41/s40\n/s32/s33/s34/s48/s32/s33/s34/s35/s33/s33/s34/s35/s33/s34/s48/s37/s38/s39/s40/s41/s42/s43/s44/s35/s33/s32/s45/s43/s40/s39/s46/s47\n/s32/s48/s33 /s32/s35/s33 /s33 /s35/s33 /s48/s33\n/s32/s43/s44/s49/s50/s40/s47/s32/s33/s34/s35/s41/s37/s32/s33/s39/s40/s35/s41/s33/s42/s43\n/s43/s51/s41/s32/s52/s53/s54/s41/s40\n/s43/s38/s46/s42/s32/s38/s55/s32/s52/s53/s54/s41/s40 /s32/s33/s34/s48/s32/s33/s34/s35/s33/s33/s34/s35/s33/s34/s48/s37/s38/s39/s40/s41/s42/s43/s44/s35/s33/s32/s45/s43/s40/s39/s46/s47\n/s32/s48/s33 /s32/s35/s33 /s33 /s35/s33 /s48/s33\n/s32/s43/s44/s49/s50/s40/s47/s32/s33/s34/s35/s41/s37/s46/s33/s39/s40/s35/s41/s33/s42/s43\n/s43/s51/s41/s32/s52/s53/s54/s41/s40\n/s43/s38/s46/s42/s32/s38/s55/s32/s52/s53/s54/s41/s40/s47/s48/s49 /s47/s50/s49\n/s47/s51/s49 /s47/s52/s49\n/s47/s53/s49 /s47/s54/s49\nFIG. 1. Magnetic hysteresis loops of W/Co/Pt trilayers mea-\nsured by a VSM. The magnetic field is applied perpendicular\n(blue) and parallel (red) to the film plane. The thickness of\nthe Co layer is (a) 0.6, (b) 0.8, (c) 1.0, (d) 1.2, (e) 1.3, (f) 1 .7,\nrespectively.\nIII. RESULTS AND DISCUSSION\nSelected magnetic hysteresis loops of the trilayers with\ndifferent Co thicknesses measured by using the VSM are\nshown in Fig 1. The easy axis of the trilayers are con-\nfirmed to lie in the film plane. The magnetic moment\nincreases with Co thickness. The magnetic properties of\nthe trilayers determined by the magnetic hysteresis loops\nare shown in Fig. 2. ThetCodependence of the magnetic\nmoment is shown in Fig. 2(a). A linear function is fitted\nto the data with larger weight for films with larger tCo.\nThe slope of the linear function is proportional to the\nsaturation magnetization (per unit volume) Msand the\nhorizontal axis intercept represents the magnetic dead\nlayer thickness tD. By taking into account the area of\nthe samples, we find Ms∼1350±50 emu·cm−3and\ntD∼0.14±0.05 nm. The value of Msis close to that of\nbulk Co [28]. tDis typically negative when Co faces a Pt\nlayerduetoproximity-inducedmagnetization[29,30], we\ninfer that a magnetic dead layer at the W/Co interface\nexists and compensates the negative tDat Co/Pt inter-\nface. Based on the reported values of proximity-induced\nmoment at Co/Pt interface ( tD∼ −0.06 nm) [29], the tD\nat W/Co interface is estimated as ∼0.20±0.05 nm.\nThe effective magnetic anisotropyenergy density, Keff,\nis obtained by taking the difference between the inte-3\n/s32/s33/s34/s35\n/s32/s33/s34/s32\n/s32/s33/s36/s35\n/s32/s33/s36/s32\n/s32/s33/s32/s35\n/s32/s37/s38/s39/s40/s41/s42/s43/s44/s36/s32/s45/s46/s43/s40/s39/s47/s48\n/s34/s33/s32/s36/s33/s35/s36/s33/s32/s32/s33/s35/s32\n/s32/s49/s38/s43/s44/s41/s39/s48/s44/s50/s48 /s44/s51/s48\n/s45/s34/s33/s32/s45/s36/s33/s35/s45/s36/s33/s32/s45/s32/s33/s35/s32/s32/s33/s35/s36/s33/s32/s33/s40/s52/s52/s53/s32/s40/s52/s52/s43/s44/s40/s54/s55/s53/s56/s39/s45/s34/s48\n/s34/s33/s32/s36/s33/s35/s36/s33/s32/s32/s33/s35/s32\n/s32/s40/s52/s52/s43/s44/s41/s39/s48\nFIG. 2. (a) Magnetic moment of W/Co/Pt trilayers as a\nfunction of Co layer thickness, tCo. The solid line is a linear\nfit to the data for tCo>1 nm. (b) Product of the effective\nmagnetic anisotropy energy, Keff, and effective magnetic layer\nthickness, teff, plotted as a function of teff. The solid line is a\nlinear fit to the data for teff>1 nm.\ngratedareasoftheeasy-axisandhard-axismagnetization\nhysteresis loops. With the effective magnetic layer thick-\nness defined by teff≡tCo−tD, the product of Keffand\nteff,Keffteff, is given by the following equation [30, 31]:\nKeffteff=KI+(KB−2πM2\ns)teff, (1)\nwhereKBandKIrepresent the bulk and interfacial con-\ntributions to Keff. The 2πM2\nsterm represents the shape\nanisotropy energy density. Keffteffis plotted against teff\nin Fig.2(b). Negative Keffteffcorresponds to the mag-\nnetization easy axis lying along the film plane. A lin-\near function is fitted to the data with larger weight on\nfilms with larger teff. The slope and the y-axis inter-\ncept of the linear function represent KB−2πM2\nsand\nKI, respectively. From the linear fit, we obtain KB∼\n(0.8±0.7)×106erg·cm−3andKI∼0.6±0.1 erg·cm−2.\nKIis smaller than that the values typically reported for\nstructures which include Co/Pt interfaces [29, 30]. Note\nthat the data show small but systemic deviation from the\nlinear fitting when teffis smaller than ∼1.0 nm.\nThe MCA of the trilayers is given by excluding the\n/s32/s33\n/s32/s34\n/s35\n/s36\n/s37\n/s33\n/s34/s38/s39/s40/s41/s42/s32/s34/s36/s41/s43/s44/s45/s46/s47/s48/s49/s50/s51\n/s50/s52/s34 /s33/s52/s34 /s32/s52/s34/s34\n/s32/s53/s32/s43/s54/s54/s41/s42/s55/s48/s49/s32/s51/s32/s33\n/s32/s34\n/s35\n/s36\n/s37\n/s33\n/s34/s38/s39/s40/s41/s42/s32/s34/s36/s41/s43/s44/s45/s46/s47/s48/s49/s50/s51\n/s33/s52/s34/s32/s52/s56/s32/s52/s34/s34/s52/s56\n/s32/s43/s54/s54/s41/s42/s55/s48/s51/s42/s57/s51 /s42/s58/s51\nFIG. 3. (a) teffand (b) 1 /teffdependence of magnetocrys-\ntalline anisotropoy (MCA) in W/Co/Pt trilayers. The solid\ncurves in panels (a) and (b) are calculated using Eq. (2) and\nthe values of KIandKBobtained from the fitting shown in\nFig. 2(b)./s32/s33/s34\n/s34/s33/s35\n/s34/s33/s36\n/s34/s33/s37\n/s34/s33/s38\n/s34/s39/s32/s39/s40/s40/s41/s42/s43/s44/s45/s46/s47/s48/s38/s49\n/s38/s33/s34/s32/s33/s50/s32/s33/s34/s34/s33/s50/s34\n/s33/s42/s51/s51/s40/s41/s52/s47/s49/s32/s33/s34\n/s34/s33/s35\n/s34/s33/s36\n/s34/s33/s37\n/s34/s33/s38\n/s34/s39/s32/s39/s40/s40/s41/s42/s43/s44/s45/s46/s47/s48/s38/s49\n/s38/s33/s50/s38/s33/s34/s32/s33/s50/s32/s33/s34/s34/s33/s50/s34\n/s32/s53/s33/s42/s51/s51/s40/s41/s52/s47/s48/s32/s49/s41/s54/s49 /s41/s55/s49\nFIG. 4. (a) teffand (b) 1 /teffdependence of the magni-\ntude of the Dzyaloshinskii-Moriya exchange constant ( |D|)\nin W/Co/Pt trilayers deduced by Brillouin light scattering\nspectroscopy (BLS) measurements. The solid lines in (a) and\n(b) show fit to the data from teff>1 nm.\nshape anisotropy from Keff:\nMCA≡Keff+2πM2\ns=KI\nteff+KB.(2)\nWe plot the teffand 1/teffdependences of the MCA in\nFigs.3(a) and (b), respectively. MCA increases mono-\ntonically with decreasing teff. The calculated MCA using\nthe parameters obtained from the fitting in Fig. 2(b) is\nshown by red solid lines in Fig. 3. Although the MCA is\nproportional to 1 /teffforteff>1 nm, it clearly deviates\nfrom the scaling for teff/lessorsimilar1 nm.\nTheteffand 1/teffdependences of the magnitude of\ntheDzyaloshinskii-Moriyaexchangeconstant( |D|), given\nby the BLS measurements, are plotted in Figs. 4(a) and\n(b), respectively. |D|increases with decreasing teffuntil\nteff∼1nm, belowwhichit drops. Asimilartendencyhas\nbeen observed in other HM/FM systems [32, 33], which\nis not in accordance with the simple picture of interface-\ndriven DMI. The 1 /teffdependence of |D|in Fig.4(b)\nis fitted using a linear function with a larger weight on\nthickerteff, as shown by a red solid line. The parameters\nobtained by the linear fitting areused to calculate the teff\ndependence of |D|in Fig.4(a), as shown by a red solid\nline. Figure 4displays that the experimental data devi-\nates from the linear fitting for teff<1 nm. We note that\n|D|of W/Co/Pt trilayer grown by molecular beam epi-\ntaxy (MBE) has also investigated in a recent study [34].\nThe value of |D|in a trilayer with 0.7-nm-thick Co layer,\n>2 erg/cm2, is much larger than our samples grown by\nmagnetron sputtering. We suggest the difference can be\nattributed to the different growth techniques.\nTo identify the origin of the teffdependences of MCA\nand DMI in the W/Co/Pttrilayers, the XAS and XMCD\nspectra of Co are studied. The setup of the measure-\nmentsisschematicallyillustratedinthe insetofFig. 6(a).\nFigure5(a) shows the XAS and XMCD spectra at the\nCoL3,2edges of the W/Co/Pt trilayers measured un-\nder a magnetic field of 5 T. The intensity is normalized\nto theL3peak after removing a background consisting\nof two step functions. No obvious peak shift or spectral\nline-shape change is found in both the XAS and XMCD4\n/s32/s33/s34\n/s35/s33/s36\n/s35/s33/s34\n/s34/s33/s36\n/s34/s37/s38/s39/s40/s41/s39/s42/s43/s40/s44/s37/s45/s46/s33/s47/s33/s48\n/s49/s34/s34 /s50/s51/s34 /s50/s49/s34 /s50/s50/s34\n/s52/s53/s54/s40/s54/s39/s37/s55/s39/s41/s56/s57/s44/s37/s45/s41/s58/s48/s59/s34/s33/s60/s59/s34/s33/s61/s59/s34/s33/s32/s34/s37/s62/s63/s64\n/s37/s62/s65/s66/s67/s68/s69/s32/s66/s54/s37/s66/s54/s69/s52/s40\n/s37/s32/s66/s54/s37/s70/s37/s34/s33/s60/s37/s39/s71\n/s37/s34/s33/s50/s37/s39/s71\n/s37/s34/s33/s49/s37/s39/s71\n/s37/s35/s33/s34/s37/s39/s71\n/s37/s35/s33/s35/s37/s39/s71\n/s37/s35/s33/s32/s37/s39/s71/s66/s54/s37/s33/s72\n/s66/s54/s37/s33/s32\n/s59/s35/s33/s34/s59/s34/s33/s49/s59/s34/s33/s60/s59/s34/s33/s61/s59/s34/s33/s32/s34/s38/s39/s40/s41/s39/s42/s43/s40/s44/s37/s45/s46/s33/s37/s47/s33/s48\n/s49/s34/s34/s50/s51/s36/s50/s51/s34/s50/s49/s36/s50/s49/s34/s50/s50/s36\n/s52/s53/s54/s40/s54/s39/s37/s55/s39/s41/s56/s57/s44/s37/s45/s41/s58/s48/s68/s69/s32/s66/s54/s37/s66/s54/s69/s52/s40\n/s37/s62/s65/s66/s67/s37/s43/s39/s40/s41/s57/s56/s46/s73\n/s32/s66/s54/s37/s70/s37/s34/s33/s60/s37/s39/s71\n/s37/s34/s33/s50/s37/s39/s71\n/s37/s34/s33/s49/s37/s39/s71\n/s37/s35/s33/s34/s37/s39/s71\n/s37/s35/s33/s35/s37/s39/s71\n/s37/s35/s33/s32/s37/s39/s71/s45/s46/s48\n/s45/s74/s48\nFIG. 5. XAS, and XMCD spectra of Co in W/Co/Pt tri-\nlayers. Spectra after background subtraction are shown. (a )\nXAS,XMCD and (b) the integrated XMCD spectra at the\nCoL3,2edges for samples with different Co thicknesses. The\nspectra obtained under the out-of-plane and “in-plane” mag -\nnetic fields are plotted by solid and dashed curves, respec-\ntively.\nspectra between different teff, suggesting that there is no\nsignificant changes in the chemical state of Co, i.e., the\noxidation of Co is negligibly small. The solid and dashed\ncurves in Fig. 5(a) represent the spectra measured with\nout-of-plane and “in-plane” magnetic fields, respectively.\nThe integrated XMCD spectra, as displayed by the cor-\nresponding curves in Fig. 5(b), show clear differences be-\ntween measurements under out-of-plane and “in-plane”\nmagnetic field directions. These results indicate that the\nmagnetic moment of Co is anisotropic.\nThe effective spin magnetic moment ( meff) of Co atom\nis estimated using the XMCD sum rule [35, 36]:\nmeff=mspin+7\n2mT\n=−2/integraltext\nL3∆µdν−4/integraltext\nL2∆µdν/integraltext\nL3,2µdνnh,(3)\nwhere ∆µandµare the difference and sum of the XAS\nspectra obtained using right- and left-handed circularly\npolarized light, mspinis the spin magnetic moment of theCo atom, mTis the magnetic dipole, nhis the number\nof holes in the 3 dband of the Co atom. Here, we present\nmagnetic moments in units of Bohr magneton per Co\natom using nh= 2.45[37]. The out-of-planeand in-plane\ncomponents of the magnetic moments are obtained from\nthe integrated XAS/XMCD spectra measured under the\nout-of-plane and “in-plane” magnetic fields, respectively.\nmspinis considered to be isotropic, but mTpossesses an\nangular dependence mT=−1\n2mT0/parenleftbig\n1−3sin2θ/parenrightbig\n[38, 39].\nHere,mT0represents the out-of-plane component of mT\nandθis the angle between the magentization and the\nfilm plane.\nThe estimated values of mspinandmT0are shown in\nFigs.6(a) and 6(b) as a function of teff, respectively.\nmspindeviates from its bulk value, shown by a horizontal\ndashed line [28], and tends to decrease with decreasing\nteff. Such variation of mspinwith film layer thickness\nhas also been observed in similar systems [29]. mspin\ncan be fitted against tCousing the relation mspin=\n(1−tD/tCo)mspin,active , where mspin,active is the active\nspin magnetic moment, to estimate the tDin the Co\nlayer. From the fitting, we obtain tD∼0.20±0.03 nm,\nwhich is consistent with the dead layer thickness deter-\nmined from the VSM measurements. mT0, which is con-\nsiderably smaller than mspin, represents the anisotropic\nspin-density distribution, and its strength characterizes\nthe anisotropy of the spin-density distribution of the d\norbitals. Although it has been reported that mT0is re-\nlated to the emergence of PMA [24] and DMI [25], here\nits magnitude is considerably smaller than the previous\nreports [25].\nThe orbital magnetic moment ( morb) of Co atom is\nestimated using the XMCD sum rule [35, 36]:\nmorb=−4\n3/integraltext\nL3,2∆µdν\n/integraltext\nL3,2µdνnh. (4)\nThe out-of-plane component of morb,m⊥\norb, is estimated\nfrom the XAS and XMCD spectra measured under the\nout-of-plane magnetic field. The in-plane component,\nm/bardbl\norb, is obtained using the spectra measured under the\nout-of-plane and “in-plane” fields according to the rela-\ntionship,\nmorb(θ) =m⊥\norbsin2θ+m/bardbl\norbcos2θ, (5)\nwithθ= 30◦. Theteffdependence of morbis plotted in\nFig.6(c). Both m⊥\norbandm/bardbl\norbdecrease with decreas-\ningtefffrom their bulk value. We find the decrease of\nm/bardbl\norbwithteffis stronger than that of m⊥\norb. This differ-\nence leadsto the OMAofCowhich is illustrated by black\nsquaresin Fig. 6(c), showingan increasingtrend with de-\ncreasing teff. The normalized orbital magnetic moment\nmorb/mspinis plotted against teffin Fig.6(d). The out-\nof-plane component, m⊥\norb/mspin, increases with decreas-\ningteffwhereas the in-plane component, m/bardbl\norb/mspin, de-\ncreases. The normalizedOMA of Co, illustrated by black5\nFIG. 6. teffdependence of (a) the spin moment ( mspin) and (b) the out-of-plane component of magnetic dipole ( mT0).\nteffdependence of the orbital magnetic moment ( morb) and normalized orbital magnetic moment ( morb/mspin), for different\nmagnetization directions, are shown in (c) and (d), respect ively. The spin, orbital and normalized orbital moment valu es of\nbulk hcp Co [28] are shown using dashed horizontal lines. The schematic images of the XAS and XMCD measurements setup\nare shown as the inset in panel (a).\n/s32/s33/s34\n/s32/s33/s35\n/s35/s33/s34\n/s35/s33/s35\n/s36/s35/s33/s34/s37/s38/s39/s40/s41/s37/s42/s43/s44/s41/s45/s46/s47/s48/s46/s49/s45/s47/s50/s41/s51/s52/s33/s41/s53/s33/s54/s32/s35/s33/s55/s55/s32/s35/s33/s55/s35\n/s56/s57/s58/s47/s58/s46/s41/s59/s46/s48/s60/s61/s50/s41/s51/s62/s48/s63/s54/s32/s32/s33/s34/s64/s32/s32/s33/s34/s65/s66/s41/s55/s34 /s66/s41/s55/s34/s67/s41/s32/s68 /s67/s41/s32/s55/s35/s33/s64/s41/s46/s69/s41/s67/s70/s35/s33/s71/s41/s46/s69/s41/s43/s58\n/s37/s38/s39\n/s37/s42/s43/s44/s32/s33/s34\n/s32/s33/s35\n/s35/s33/s34\n/s35/s33/s35\n/s36/s35/s33/s34/s37/s38/s39/s40/s41/s37/s42/s43/s44/s41/s45/s46/s47/s48/s46/s49/s45/s47/s50/s41/s51/s52/s33/s41/s53/s33/s54/s32/s32/s33/s64/s35/s32/s32/s33/s34/s64 /s32/s68/s33/s68/s55/s32/s68/s33/s55/s71\n/s56/s57/s58/s47/s58/s46/s41/s59/s46/s48/s60/s61/s50/s41/s51/s62/s48/s63/s54/s66/s41/s32/s35 /s66/s41/s32/s35/s56/s47/s41/s32/s68/s56/s47/s41/s32/s55/s35/s33/s64/s41/s46/s69/s41/s56/s47/s70/s35/s33/s71/s41/s46/s69/s41/s43/s58\n/s37/s38/s39\n/s37/s42/s43/s44/s51/s52/s54 /s51/s72/s54\nFIG. 7. XAS and XMCD spectra at the W (a) and Pt (b)\nL3,2edges in W/Co and Pt/Co bilayers. The intensity of the\nXMCD spectra of W (Pt) is enlarged by a factor of 25 (10).\ndiamonds in Fig. 6(d), increases with decreasing teff, es-\npecially for teff/lessorsimilar0.8 nm. These results indicate that\nthe charge redistribution at the HM/Co interfaces takes\nplace and induces OMA. Considering the density of Co\ncrystal as 8.9 g ·cm−3, we estimate the total magnetiza-\ntion of the W/Co/Pt trilayer with teff=0.9 nm as 1300\nemu·cm−3. This value is in good agreement with that\ndetermined by using the VSM.\nWe have also performed W and Pt L3,2-edge XMCD\nmeasurements on W/Co and Pt/Co bilayers to study theproximity-induced magnetization. The XAS and XMCD\nspectra are shown in Fig. 7. XMCD signals are found\nonly in the Pt/Co bilayer. These results indicate that\nproximity-induced magnetization exists at the Pt/Co in-\nterface but does not exist at the W/Co interface.\nNow, we discuss the relationship between the MCA,\nDMI and OMA of Co in W/Co/Pt trilayers. Accord-\ning to Figs. 3and4, MCA and |D|scale with 1 /teff\nforteff/greaterorsimilar1 nm, indicating the interfacial origin of the\ntwo properties. Bruno has proposed that the MCA and\nOMA are proportional to each other in FM monolayers\n[22]. Both MCA and the OMA of Co indeed increase\nwith decreasing teffsmaller than ∼0.8 nm. However,\nthe OMA tends to increase more rapidly with decreasing\nteffthan what the Bruno’s law predicts from the values\nof MCA. These results suggest additional contributions\nto MCA, such as strain or magnetic dipole mT. Previ-\nous studies have indicated that strain in the film texture\ncan weaken the MCA when the magnetic layer thickness\nis reduced to a few atomic layers [40, 41]. The teffde-\npendences of MCA and Keffteffare in accordance with\nsuch studies. van der Laan has also shown that mTaf-\nfects MCA in strongly spin-orbit coupled systems, which\nhas been associated with spin-flip virtual excitation [24].\nSuch relationship has been confirmed in recent experi-\nments [42–44]. Unfortunately, the XMCD spectra of our6\nW/Co/Pt trilayers does not have sufficient resolution to\nderivemTaccurately.\nThe DMI, on the other hand, approaches near zero as\nteffis reduced below ∼1 nm. Interfacial DMI is due to\nthe strongspin-orbitcouplingwith brokeninversionsym-\nmetry. Recent studies have shown that DMI is related\nwithmTat the HM/FM interfaces [25] and the OMA of\nthe FM atoms [23]. Comparing the results presented in\nFigs.4(a)and6(c), weconsiderthatsuchrelationsdonot\nhold in the current system because the teffdependences\nof OMA and DMI are opposite to what one expects from\nthe scaling reported in Ref. [23]. The magnitude of mT\nfound in this system is considerably smaller than that re-\nportedinRef.[25]and, therefore,itscontributiontoDMI,\nif any, is alsolikely small. Furthermore, DMI in oursput-\ntered samples is much smaller than that in MBE-grown\nW/Co/Pt trilayer [34]. Thin films grown by MBE usu-\nally show better interfacial roughness and crystal texture\ncompared to films grown by magnetron sputtering. We\nthus speculate that the DMI is more sensitive to strain\neffect or the (111) texture of Co. Theoretical studies\n[45, 46] have indicated that the crystal structure at the\nHM/FM interface influences the strength of DMI dra-\nmatically. In the present case, strain and texture of the\nCo layer near the W/Co interface may significantly de-\ngrade|D|forteff/lessorsimilar1 nm. With further increasing Co\nthickness, such effects are then mitigated by the Co/Pt\ninterface that favors the (111) texture, resulting in the\nfollowing decrease of |D|.\nIV. SUMMARY\nWe have studied the effective magnetic layer thick-\nness (teff) dependences of magnetocrystalline anisotropy\n(MCA), Dzyaloshinskii-Moriya interaction (DMI), and\norbital moment anisotropy (OMA) in W/Co/Pt trilay-\ners. For tefflarger than ∼1 nm, MCA and DMI scale\nwith 1/teff, indicating an interfacial origin. However,\nwhereas MCA continues to increase with decreasing teff,DMI tends to decrease when teffis reduced below ∼1\nnm. The OMA of Co deduced from x-ray magnetic cir-\ncular dichroism (XMCD) measurements is almost zero\n(below the detection limit) when teffis larger than ∼0.8\nnm, below which the OMA of Co increases with decreas-\ningteff. The rate at which the OMA of Co increases\nwith decreasing teffis larger than what is predicted from\nthe MCA using Bruno’s formula. The reduction of DMI\nwith decreasing tefffor films with teff/lessorsimilar1 nm, despite\nthe presence of OMA, suggests that other factors con-\ntribute to the DMI in this thickness range. We infer that\nthe strain/texture in the Co layer induced by the W un-\nderlayer significantly weakens the DMI and, to a lesser\nextent, the MCA. Further studies are necessary to clarify\nthe latter points. Our results provide a microscopic un-\nderstanding for designing viable FM/HM-interface-based\nmultifunctional spintronic devices.\nACKNOWLEDGMENTS\nWe thank H. Shimazu for samples preparation. This\nwork was supported by Grants-in-Aid for Scientific Re-\nsearch from JSPS (Grant Nos. 15H02109, 15H05702,\n16H03853, and 20K14416). The XMCD experiment was\nperformed at BL-16A of KEK-PF with the approval of\nthe Photon Factory Program Advisory Committee (pro-\nposal Nos. 2016S2-005 and 2016G066) and at BL39XU\nof SPring-8 with the approval of the Japan Synchrotron\nRadiation Research Institute (JASRI) (proposal Nos.\n2017A1048 and 2018A1058). M.H. and A.F. are ad-\njunct members of Center for Spintronics Research Net-\nwork (CSRN), the University of Tokyo, under Spintron-\nics Research Network of Japan (Spin-RNJ). Z.C. is sup-\nported by Materials Education program for the future\nleaders in Research, Industry, and Technology (MERIT)\nand JSR Fellowship, The University of Tokyo. Y.-C.L. is\nsupported by JSPS InternationalFellowship for Research\nin Japan (Grant No. JP17F17064). S.S. and Y.-X.W.\nacknowledges financial support from Advanced Leading\nGraduate Course for Photon Science (ALPS). S.S. ac-\nknowledges financial support from the JSPS Research\nFellowship for Young Scientists.\n[1]I. M. Miron, K. Garello, G. Gaudin, P. J. Zermatten,\nM. V. Costache, S. Auffret, S. Bandiera, B. Rodmacq,\nA. Schuhl, and P. Gambardella, Nature 476, 189 (2011).\n[2]I. M. Miron, T. Moore, H. Szambolics, L. D. Buda-\nPrejbeanu, S. Auffret, B. Rodmacq, S. Pizzini, J. Vogel,\nM. Bonm, A. Schuhl, and G. Gaudin, Nat. Mat. 10, 419\n(2011).\n[3]M. I. Dyakonov and V. I. Perel, Jetp Letters-Ussr 13,\n467 (1971).\n[4]J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999).\n[5]J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back,\nand T. Jungwirth, Rev. Mod. Phys. 87, 1213 (2015).\n[6]L. Liu, C.-F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and\nR. A. Buhrman, Science 336, 555 (2011).[7]Y. A.BychkovandE. I. Rashba,Jetp Lett. 39, 78(1984).\n[8]V. M. Edelstein, Solid State Communications 73, 233\n(1990).\n[9]I. E. Dzyaloshinskii, JETP 5, 1259 (1957).\n[10]T. Moriya, Phys. Rev. 120, 91 (1960).\n[11]M. Bode, M. Heide, K. von Bergmann, P. Ferriani,\nS. Heinze, G. Bihlmayer, A. Kubetzka, O. Pietzsch,\nS. Bl¨ ugel, andR.Wiesendanger, Nature 447, 190(2007).\n[12]S. S. P. Parkin, M. Hayashi, and L. Thomas, Science\n320, 190 (2008).\n[13]M. Hayashi, L. Thomas, R. Moriya, C. Rettner, and\nS. S. P. Parkin, Science 320, 209 (2008).\n[14]M. Heide, G. Bihlmayer, and S. Bl¨ ugel, Phys. Rev. B\n78, 140403 (2008).7\n[15]B. B. S. M¨ uhlbauer, F. Jonietz, C. Pfleiderer, A. Rosch,\nA. Neubauer, R. Georgii, and P. B¨ oi, Science 323, 915\n(2009).\n[16]X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han,\nY. Matsui, N. Nagaosa, and Y. Tokura, Nature 465, 901\n(2010).\n[17]A. Thiaville, S. Rohart, ´E. Ju´ e, V. Cros, and A. Fert,\nEurophys. Lett. 100, 5 (2012).\n[18]S. Emori, U. Bauer, S.-M. Ahn, E. Martinez, and\nG. S. D. Beach, Nat. Mater. 12, 611 (2013).\n[19]A. Fert, V. Cros, and J. Sampaio, Nat. Nanotech. 8, 152\n(2013).\n[20]S. Tacchi, R. E. Troncoso, M. Ahlberg, G. Gubbiotti,\nM. Madami, J. ˚Akerman, and P. Landeros, Phys. Rev.\nLett.118, 147201 (2017).\n[21]X. Ma, G. Yu, S. A. Razavi, S. S. Sasaki, X. Li, K. Hao,\nS. H. Tolbert, K. L. Wang, and X. Li, Phys. Rev. Lett.\n119, 027202 (2017).\n[22]P. Bruno, Phys. Rev. B 39, 865 (1989).\n[23]K. Yamamoto, A.-M. Pradipto, K. Nawa, T. Akiyama,\nT. Ito, T. Ono, and K. Nakamura, AIP Adv. 7, 056302\n(2017).\n[24]G. van der Lann, J. Phys.: Condens. Matter 10, 3239\n(1998).\n[25]S. Kim, K. Ueda, G. Go, P.-H. Jang, K.-J. Lee, A. Be-\nlabbes, A. Manchon, M. Suzuki, Y. Kotani, T. Naka-\nmura, K. Nakamura, T. Koyama, D. Chiba, K. Yamada,\nD.-H. Kim, T. Moriyama, K.-J. Kim, and T. Ono, Nat.\nCommun. 9, 1648 (2018).\n[26]I. V. Solovyev, P. H. Dederichs, and I. Mertig, Phys.\nRev. B52, 13419 (1995).\n[27]K. Di, V. L. Zhang, H. S. Lim, S. C. Ng, M. H. Kuok,\nJ. Yu, J. Yoon, X. Qiu, and H. Yang, Phys. Rev. Lett.\n114, 047201 (2015).\n[28]J.M.D.Coey, Magnetism and Magnetic Materials (Cam-\nbridge, UK, 2010).\n[29]T. Ueno, J. Sinha, N. Inami, Y. Takeichi, S. Mitani,\nK. Ono, and M. Hayashi, Sci. Rep. 5, 14858 (2015).\n[30]Y.-C. Lau, Z. Chi, T. Taniguchi, M. Kawaguchi, G. Shi-\nbata, N. Kawamura, M. Suzuki, S. Fukami, A. Fujimori,\nH. Ohno, and M. Hayashi, Phys. Rev. Mater. 3, 104419\n(2019).\n[31]J. Sinha, M. Hayashi, A. J. Kellock, S. Fukami, M. Ya-\nmanouchi, H. Sato, S. Ikeda, S. Mitani, S.-H. Yang,\nS. S. P. Parkin, and H. Ohno, Appl. Phys. Lett. 102,\n242405 (2013).\n[32]J. Cho, N.-H. Kim, S. Lee, J.-S. Kim, R. Lavrijsen,\nA. Solignac, Y. Yin, D.-S. Han, N. J. van Hoof, H. J.\nSwagten, B. Koopmans, and C.-Y. You, Nat. Commun.\n6, 7635 (2015).\n[33]M. Belmeguenai, M. S. Gabor, Y. Roussign´ e, T. P. Jr.,\nR. B. Mos, A. Stashkevich, S. M. Ch´ erif, and C. Tiusan,\nPhys. Rev. B 97, 054425 (2018).\n[34]S. K. Jenaa, R. Islamb, E. Mili´ nskaa, M. M.\nJakubowskia, R. Minikayeva, S. Lewi´ nskaa, A. Lyn-\nnyka, A. Pietruczika, P. Aleszkiewicza, C. Autieri, and\nA. Wawro, Nanoscale 13, 7685 (2021).\n[35]B. T. Thole, P. Carra, F. Sette, and G. van der Laan,\nPhys. Rev. Lett. 68, 1943 (1992).\n[36]P. Carra, B. T. Thole, M. Altarelli, and X. Wang, Phys.\nRev. Lett. 70, 694 (1993).\n[37]N. Nakajima, T. Koide, T. Shidara, H. Miyauchi,\nH.Fukutani, A.Fujimori, K.Iio, T.Katayama, M.N´ yvlt,and Y. Suzuki, Phys. Rev. Lett. 81, 5229 (1998).\n[38]R. Wu and A. J. Freeman, Phys. Rev. Lett. 73, 1994\n(1994).\n[39]J. St¨ ohr, J. Magn. Magn. Mater. 200, 470 (1999).\n[40]M. T. Johnson, P. J. H. Bloemen, F. J. A. den Broeder,\nand J. J. de Vries, Rep. Prog. Phys. 59, 1409 (1996).\n[41]Y.-C. Lau, P. Sheng, S. Mitani, D. Chiba, and\nM. Hayashi, Appl. Phys. Lett. 110, 022405 (2017).\n[42]S. Miwa, M. Suzuki, M. Tsujikawa, K. Matsuda,\nT. Nozaki, K. Tanaka, T. Tsukahara, K. Nawaoka,\nM. Goto, Y. Kotani, T. Ohkubo, F. Bonell, E. Tamura,\nK. Hono, T. Nakamura, M. Shirai, S. Yuasa, and\nY. Suzuki, Nat. Commun. 8, 15848 (2017).\n[43]K. Ikeda, T. Seki, G. Shibata, T. Kadono, K. Ishigami,\nY. Takahashi, M. Horio, S. Sakamoto, Y. Nonaka,\nM. Sakamaki, K. Amemiya, N. Kawamura, M. Suzuki,\nK. Takanashi, and A. Fujimori, Appl. Phys. Lett. 111,\n142402 (2017).\n[44]G. Shibata, M. Kitamura, M. Minohara, K. Yoshimatsu,\nT. Kadono, K. Ishigama, T. Harano, Y. Takahashi,\nS. Sakamoto, Y. Nonaka, K. Ikeda, Z. Chi, M. Fu-\nruse, S. Fuchino, M. Okano, J. i. Fujimori, A. Uchida,\nK. Watanabe, H. Fujihira, S. Fujihira, A. Tanaka, H. Ku-\nmigashira, T. Koide, and A. Fujimori, npj Quan. Mat.\n3, 3 (2018).\n[45]H. Yang, A. Thiaville, S. Rohart, A. Fert, and\nM. Chshiev, Phys. Rev. Lett. 115, 267210 (2015).\n[46]A. Hrabec, N. A. Porter, A. Wells, M. J. Benitez, G. Bur-\nnell, S. McVitie, D. McGrouther, T. A. Moore, and C. H.\nMarrows, Phys. Rev. B 90, 020402 (2014)." }, { "title": "2102.05315v1.Magneto_electric_Tuning_of_Pinning_Type_Permanent_Magnets_through_Atomic_Scale_Engineering_of_Grain_Boundaries.pdf", "content": "1 \n Magneto -electric Tuning of Pinning -Type Permanent Magnets \nthrough Atomic -Scale Engineering of Grain Boundaries \n \n \nXinglong Ye1❀, Fengkai Yan2 , Lukas Schaefer3, Di Wang1,4, Holger Ge ßwein5, Wu Wang1♪, \nMoham med Reda Chellali1, Leigh T. Stephenson2, Konstantin Skokov3, Oliver Gutfleisch3, Dierk Raabe2, \nHorst Hahn1, Baptiste Gault2,6❀, Robert Kruk1 \n \n1 Institute of Nanotechnology, Karlsruhe Institute of Technology (KIT), Eggenstein -Leopoldshafen, Germany \n2 Department of Microstructure Physics and Alloy Design, Max -Planck -Institut für Eisenforschung GmbH (MPIE) , \nDüsseldorf, Germany. \n3 Department of Material Science, Techni cal Universit y Darmstadt, Darmstadt, Germany \n4 Karlsruhe Nano Micro Facility, Karlsruhe Institute of Technology (KIT), Karlsruhe, Germany \n5 Institute for Applied Materials, Karlsruhe Institute of Technology, Eggenstein -Leopoldshafen, Germany . \n6 Department of Materials, Imperial College London, London, UK \nCorresponding authors . Email : xing-long.ye @kit.edu ; b.gault@mpie.de \n now at Institute of Metal Research, Chinese Academy of Science, Shenyang, China \n♪now at Department of Physics, Southern University of Science and Technology, Shenzhen, China. \n \nAbstract \nPinning -type magnets maintaining high coercivity, i.e. the ability to sustain magnetization , at high \ntemperature are at the core of thriving clean -energy technologies. Among the se, Sm 2Co17-based \nmagnets are excellent candidates owing to their high-temperature stabilit y. However, despite decades \nof efforts to optimize the intragranular microstructure , the coercivit y currently only reach es 20~30% \nof the theoretical limits . Here, the roles of the grain -interior nanostructure and the grain boundaries in \ncontrolling coercivity are disentangled by an emerging magneto -electric approach. Through hydrogen \ncharging/discharging by applying voltages of only ~ 1 V , the coercivity is reversibly tuned by an \nunprecedented value of ~ 1.3 T. In situ magneto -structural measurements and atomic -scale tracking of \nhydrogen atoms reveal that the segregation of hydrogen atoms at the grain boundaries, rather than the \nchange of the crystal structure, dominates the reversible and substantial change of coercivity. Hydrogen \nlowers the local magnetocrystalline anisotropy and facilitates the magnetization reversal starting from \nthe grain boundaries . Our study reveals the previously neglected critical role of grain boundaries in the \nconventional magnetisation -switching paradigm , suggesting a critical re considerat ion of strategies to \novercome the coercivity limits in permanent magnets, via for instance atomic -scale grain boundar y \nengineering . \nConsequently present work . (200 words) \n \nKey words: Magneto -electric, Permanent magnets, H ydrogen , Grain boundary \n 2 \n Introduction \nPermanent magnets with the ability to maintain the ir magnetization , i.e. a property referred to as \ncoercivity , at high temperatures are crucial for the flourishing clean energy technologies such as \nelectric vehicles and wind powers1,2. In this regard , the pinning -type magnets, in which the coercivity \narises from the pinning of magnetic domain walls at nano -precipitate s within the grain ,6-8 are most \npromising . As an example , the Sm 2Co17-based magnet is the only candidate for use in electric motors \nworking above 300 ℃ owing to its excellent temperature stability . Its coercivity is usually believed to \nbe controlled exclusively by domain -wall pinning because of the nano -scale cellular microstructure \nwithin the grain9-13, while the initial demagnetization at grain boundaries is considered irrelevant . \nHowever, d espite intensive efforts to optimize its intra-granular microstructure, the coe rcivity currently \nonly reach es 20-30% of the theoretical anisotropy field .1,14,15 Hence , it inevitably opens the question \nabout the possible influence of grain boundar ies during the magnetization reversal of the material . \n \nThe role of grain boundar ies in pinning -type magnets may be understood if they can be modified \nseparately from the grain interior , in conjunction with measuring the associated coercivity change . \nTraditional processing approaches , such as heat treatment s13, plastic deformation16and alloying9,17, can \ndramatically change the coercivity. However , these approaches often induce irreversible modification \n(or destruction) of the microstructure both in grain boundar ies and grain interior, obscuring the \nassessment of their respective impact on coercivity . Consequently, d ecoupl ing the separate roles of the \ngrain boundary and the grain interior in magnetization reversal becomes technically challenging. \n \nIt has recently been demonstrated that t he magneto -electric approach can reversibly modify the \nmagnetic properties of materials with external voltages without chang ing the microstructure8,19. For \ninstance, voltage -driven proton pumping and the voltage -controlled hydrogen insertion /extraction can \nsubstantially tune the coercivity of ferromagnetic metals.20,21 Owing to different affinities of hydrogen \natoms to the microstructural defects , hydrogen atoms are expected to diffuse first along the grain \nboundaries, and, then into the grain interior .22,23 This sequential diffusion, if controlled , offer s an \nopportunity to decouple the role s of the grain boundary and the grain interior , if the associated \ncoercivity change is monitored at each step. Here , by employing electrochemically -controlled \nhydrogen charging/discharging, we tuned the coercivity of the Sm 2Co17-based hard magnet by ~1.3 T , \nthe largest values ever achieved by magneto -electric approaches . The combined in situ magn eto-\nstructural measurements and atomic -scale mapping of hydrogen distribution24 reveal that hydrogen \natoms strongly segregate at grain boundaries, which we akens the local magnetic anisotropy and \naccounts for the predominant change of coercivity . Our study open s a way to achieve giant \nmagnetoelectric effects by atomic -scale engineering of grain boundaries , and unveils the critical role \nof grain boundar ies in limiting pinning-type magnet’s performance pointing the way forward for future \noptimisation strategies. \n \nResults \nMagneto -optical observation of magnetization reversal \nWe used commercial Sm 2Co17-based permanent magnet s with composition s of Sm(Co 0.766 3 \n Fe0.116Cu0.088 Zr0.029)7.35 (hereafter referred to as SmCo 7.35 sample ) (Table S1 ). The hysteresis loop \nshows a coercivity of ~2.8 T ( Fig. 1A). By magneto -optical Kerr effect (MOKE) microscopy, we \nobserved its magnetization reversal process under demagnetiz ation fields. Prior to MOKE imaging , \nthe sample was fully magnetized at -6.8 T ( Fig. 1B), and t he domain structure imaged with the c -axis \nin the viewing plane. \n \nAt a demagneti zation field of 1.5 T, the sample re tained its fully-magnetized state (Fig. 1C). When the \nfield increase d to 2 T, the reversed domains started to appear at grain boundaries (Fig. 1D), and at 2.2 \nT expand ed into the interior of the grains ( Fig.1E). With further increasing field, the magnetic domains \nmoved massively into the grain ( Fig. 1F). At 3.0 T, only a few residual domains were un -reversed (Fig. \n1G). These observations demonstrate the initial nucleation of magnetic domains at grain boundaries \nbefore their growth into the grain interior. \n \nReversible modification of coercivity by non-destructive hydrogen charging/discharging \nWe employed an electrochemical three -electrode cell to charge/discharge the SmCo 7.35 sample with \nhydrogen atoms (Fig. S1 ). In this setup, the as -prepared electrode with the SmCo 7.35 particles was the \nworking electrode and 1 M KOH aqueous solution the electrolyte (Fig. S2 ). During hydrogen charging , \nthe electrochemical reduction of water molecules on the metal surface provide s the hydrogen ad -atoms \nthat subsequently diffuse into the material . Convers ely, during the discharging , the hydrogen atoms on \nthe surface ( Hads) were oxidized and removed, resulting in hydrogen desorption. Based on the \nmeasured cyclic voltammogram (Fig. S3 ), we used the voltages steps of -1.2 V and -0.4 V to charge \nand discharge the sample , respectively . \nWe explored the response of the coercivity of the as-prepared SmCo 7.35 sample to hydrogen \ncharging/discharging by in situ superconducting quantum interference device (SQUID) . The coercivity \nof the as -prepared sample was ~ 2.3 T ( Fig. 2A), slightly lower than that of the bulk -form pristine \nsample (~ 2. 8 T) (Fig. 1A). After charging at -1.2 V for one hour, the coercivity drastically decreased \nto ~ 1.0 T. We monitored the recovery of the coercivity during the discharg ing process by continuously \nrecording the hysteresis loops. The coercivity increased monotonically with the discharging time ( Fig. \n2A), and re gained most of its initial value in the early stage of the discharging process, reaching ~ 1.9 \nT within 10 hours ( inset in Fig. 2A). After a prolonged time of discharging (~ 60 hours), the coercivity \nfully recovered. \n \nIn parallel, we studied the dynamics of the hydrogen charging/ discharging process by observing the \nevolution of crystal structure with in situ X-ray diffraction ( XRD ) in transmission mode (Fig. 2B). \nUpon hydrogen -charging , all diffraction peaks were shifted to lower angles , and after one hour , the \npositions of the diffraction peaks stayed unchanged, indicating the complete charging of the whole \nsample. In the discharg ing process , two stages were discerned . First, strikingly, only a negligible shift \nof the peaks was observed over the first 10 hours (Fig. 2C), indicating that the bulk material is still \ncharged with hydrogen atoms. Second, only after about 80 hours of discharg ing the peak s recovered \nto their original positions. This observation is in strong contrast with the substantial change in \ncoercivity over the corresponding period ( Inset in Fig. 2A). It suggests that the predominant coercivity 4 \n change is not ascribed to the slow hydrogen desorption from the volume of the material. \nSince the predominant change of coercivity does not arise from the volumetric slow diffusion of \nhydrogen atoms, we expect a relatively fast response of magnetization reversal to hydrogen charging \n(Fig. 2C). We first magnetized the as -prepared sample with 6.8 T (point ① in the inset). Then, the \nmagnetic field was reversed to -1.1 T (point ②), smaller than the coerci vity of the pristine sample (~ \n2.3 T) and therefore , the magnetization remained positive and nearly constant . Upon hydrogen \ncharging by applying -1.2 V (point ③), the magnetization decreased immediately and abruptly , and \nflipped fro m positive to negative in ~10 minutes. The m agneti zation started to level off after ~4 hours. \nThe immediate response of magnetization reversal to the voltage stimulus confirms the existence of \nrelatively fast diffusion path of hydrogen atoms. \n \nAtomic -scale tracking of hydrogen atoms within the hierarchical microstructure \nWe carried out multi -scale multi -microscopy mapping of the micro - and nano -structur al features to \nrationalize the fast diffusion pathways of hydrogen atoms . Optical microscopy (Fig. 3A, Fig. 1 ) \nshowed that the sample was polycrystalline with grain s of ~26 m separated by high-angle grain \nboundaries (HAGBs) . The grains were further divided into sub -grains by low -angle grain boundaries \n(LAGB s) as shown by electron back -scattered imaging (Fig. 3B). Inside the grain transmission \nelectron microscopy show s the typical cellular structure, compos ed of the matrix cell with sizes of ~ \n40 nm , the cell boundary and the Zr-rich lamellae crossing the cellular structure (TEM, Fig. 3C, S4). \nAfter tilting the c-axis of the specimen out of the viewing plane , high-resolution TEM and the \ncorresponding selected area electron diffraction (SAED) showed that the matrix phase is Sm 2Co17 \n(rhombohedral Th 2Zn17 type) and the cell boundary phase SmCo 5 (hexagonal CaCu 5 type) (Fig. 3D). \nThis hierarchical micro - and nanostructure matche s previous reports9-13. \n \nWe used atom probe tomography ( APT ) to locate hydrogen atoms and deuterium atoms within the \nhierarchical micro structure of the deuterium -charged samples. Isotopic marking by deuterium atoms \n(D) minimize d the influence of residual hydrogen in the atom probe . We first analysed a specimen \ncontaining a HAGB ( Fig. S5, S6). The element -specific atom maps in Fig. 4A reveal the Cu-rich cell \nboundaries , the matrix cells and the Zr -rich platelets , match ing TEM results . Three-dimen sional \nreconstruction shows that D mostly segregate s in a 10 –12 nm thick layer at HAGB , reaching a \nconcentration of approx. 3.5 at% ( Fig. 4A). A close face-on view ( Fig. 4B) reveals that D segregate s \nat the intersection of the cell boundaries with the HAGB . In the cells and cell boundaries, a very limited \namount of D was detected within the structure close to the detection limit , with no noticeable \npartitioning difference between them (Fig. S6). In addition, d euterium atoms appear slightly depleted \nfrom the Z r-phase. \n \nWe then performed another analysis target ing LAGB s (Fig. 4C, D, S7). Again, deuterium appears \ndeplet ed in the Z-phase , and no preferential segregation within cells and the cell boundaries (Fig. S 8). \nYet again, a strong segregation of deuterium at LAGB was observed. The corresponding top -view \nshows a series of linear features highlighted by a set of 0.35 at.% D iso -composition surfaces, which \nare likely the dislocations that constitute the LAGB26. The D -concentration at these dislocations can 5 \n reach up to 0.4 at% D, and H up to 4 –5 at%. Importantly, the cell edges are all connected to these \ndislocations and the cell structure stops abruptly at LAGB ( Fig. 4D). \n \nDiscussion \nThe observatio n of hydrogen /deuterium segregation at GB regions (Fig. 4), coupled with in situ XRD , \ndetecting no volumetric structural change (Fig. 2B), suggest s that the substantial change of coercivity \nin the early stage of discharging process arises from the desorption of hydrogen atoms from GBs. The \nability to modulate the coercivity by only charging/discharging GBs enables the fast control of \ncoercivity, as verified by the immediate start of magnetization reversal upon hydrogen charging (Fig. \n2B). The herein identified cr ucial role of grain boundary in controlling the coercivity explains why \nreducing the volume fraction of grain boundaries27 or optimising the cellular structure near the grain \nboundary28 can increase the coercivity of Sm 2Co17-based magnet s. Below we discuss how hydrogen \nsegregation at GBs changes the coercivity , starting with the microstructural features near GBs. \n \nThree microstructural features distinguish the GB region from the grain interior. First, compared with \nthe continuous cellular structure in the grain interior ( bottom part in Fig. 4C), the cell boundaries are \nbroken and terminated near GBs (Fig. 4A, C, D). Second, the typical cell size and shape in the grain \nis approx. 40 nm and with a regular shape, but becomes larger and strongly elongated near GBs (Fig. \n4). Th ese results agree with recent TEM reports that the incomplete cellular structure near GBs extend \ntowards the sub-micrometer scale29,30. Third, t he composition profiles (Fig. S 6, S8) show that the \nSmCo 5 phase contains almost twice as much Cu near the grain boundary , i.e. 30 at.% compared with \n15 at.% in the grain interior . \n \nThe observed different cellular structure and microchemistry near GBs significantly reduce s the local \nnucleation field required for magnetization reversal. According to the micromagnetic theory, the \ncritical nucleation field, Hn, can be described by14 \n 𝐻𝑛=1\n2𝑀s∆(𝛾𝑆𝑚𝐶𝑜 5−𝛾𝐺𝐵)−𝐷𝑀s (1) \nin which ∆ is the width of the transition region where the domain -wall energy chan ges from 𝛾𝐺𝐵 in \nthe grain boundary to 𝛾𝑆𝑚𝐶𝑜 5 in the cell boundary , and 𝐷 is the demagnetizing factor. The SmCo 5 \nphase has much larger magn etocrystalline anisotropy than GBs, and determines the domain -wall \nenergy difference , (𝛾𝑆𝑚𝐶𝑜 5−𝛾𝐺𝐵) . Near GBs the Cu concentration in the SmCo 5 phase becomes \ntwice that in the grain interior , which substantially reduces its magnetocrystalline anisotropy31,32 and \nthus the domain wall energy, 𝛾𝑆𝑚𝐶𝑜 5 . In addition , the dis rupted SmCo 5 phase allows the easy \nmovement of domain walls through the matrix phase, triggering a macroscopic magnetization reversal. \nThese account for the preferential nucleation of reversed domains near GBs ( Fig. 1 ). Moreover , when \nthe SmCo 5 phase was charged with hydrogen atoms , its magnetocrystalline anisotropy will decrease \nby ~40%20. This further decrease s the domain -wall energy of the SmCo 5 phase and the nucleation field. \nBesides, h ydrogen segregation may enlarge the transition region ( ∆) between GBs and the SmCo 5 \nphase with its continuous concentration change and reduce the nucleation field. Hence, hydrogen \nsegregation acts here as a tool to further weaken the nucleation field of the GB region and amplify its 6 \n effect in initiating the magnetization reversal . Next , we consider the mechanism behind the \npropagation of the initial ly-nucleated magnetic domains near GBs into the grain interior . \n \nIn the grain interior , the continuous network of the SmCo 5 phase subdivides the individual grains of \nthe Sm 2Co17 matrix into a nano scale cellular structure , rendering them into classical pinning -type \nmagnet s. However , the models to explain their magnetization reversal assume that grain boundar ies \nshould be non -ferromagn etic and one -to-two atomic layers thick to reduce the associated stray field \nnegligibly14,33. This is not the case in the current material because of the expanded region near GBs \nwith the disintegrated cellular structure and different microchemistry. We can describe the \ndemagnetization process of the whole grain triggered by the initial demagnetization near GB as follows . \nThe local magnetic field is a superposition of the external field Hext and the local demagnetization field , \nN’M, where N’ is the local or effective demagnetization factor and M the net magnetization of the \nsample . The latter can be significantly inhomogeneous, and reach values much larger than the net \ndemagnetization field3, HD = NM, N being the demagnetization factor of the sample (3). As discussed \nearlier, the nucleation field of the GB region, Hc,GB, is much smaller than Hc of the grain interior , Hc,g. \nUnder a small external field , we have Hc,g>Hc,GB>Hext+NM. As the local magnetic field , Hext+NM, \napproaches Hc,GB, the initial nucleation of magnetic domains occur s near GBs (Fig. 1 ), producing a \nlocal demagnetization field, N’M s, where Ms is spontaneous magnetization of the main Sm 2Co17 phase. \nThen, the adjacent inner layer with higher coercivity , Hc,g, is under the higher magnetic field, \nHext+NM+N’M s. This additional negative field , N’M s, can be of 0.5 -1.0 T, depending on microstructural \nfeatures and Ms34. Thus, if Hc,g-Hc,GB\n|±3/2>\n|±1/2>\n\u0001+ s1(a) (b) (c)\u0000+ s2\u00027\n\u00039\n\u00048\n∆\u0002= 5.8 meV\nFIG. 2. (Color online) (a) Temperature dependence of the inv erse magnetic susceptibility of single crystal CeRh 6Ge4for two\nfield directions [ 17]. The solid lines show the results from fitting with a CEF mode l described in the text. (b) CEF level\nscheme and wavefunctions obtained from fitting with the CEF m odel, where the angular distributions of the wave functions\nare also displayed. (c) Crystal structure of CeRh 6Ge4, with Ce, Rh, and Ge shown in yellow, grey and purple, respect ively.\nFerromagnetic order with moments along the a-axis is also illustrated, as well as the proposed muon stopp ing sites s1ands2\n(orange and green spheres), and the orientation of the groun d state orbitals from the CEF model.\norder with a moment of 0 .24µB/Ce orientated along the\na-axis, local fields of 364 and 152 G are calculated for the\ns2sites, and 87 G is calculated for s1, in comparison to\nfitted values for B1,B2andB3of 405, 151, and 56 G, re-\nspectively. On the other hand, a moment of 0 .155µB/Ce\nyields 58 G for s1, in good agreement with B3, but yields\nunderestimates of 235 and 98 G for the s2sites. In mag-\nnetic metals, an accurate comparison between the calcu-\nlated and observedlocal fields requiresaccountingfor the\nmuon contact hyperfine fields, and similar discrepancies\nto calculations have been found for heavy fermion mag-\nnets [31]. Moreover, a change in the hyperfine fields with\ntemperature could lead to the non-monotonic behavior\nofB1, which together with the increase of λbelow 0.8 K,\nmay point to the low temperature evolution of the un-\nderlying correlated state. The differences may also arise\nfrom uncertainties in the positions of the muon stopping\nsites, the orientation of the moments in the basal plane,\noraspatialmodulationoftheorderedmoment, butinthe\ncase of the latter AFM Bragg peaks would be expected\nto be observed in neutron diffraction. As a result, both\nneutron diffraction and ZF- µSR are consistent with FM\norder in CeRh 6Ge4, with a small in-plane ordered mo-\nment, and indicate the absence of any significant AFM\ncomponent.\nIn order to determine the splitting of the J= 5/2 Ce\nground state multiplet of CeRh 6Ge4by crystalline elec-\ntric fields (CEF), the single crystal magnetic suscepti-\nbility [17] was analyzed using the Hamiltonian HCF=\nB0\n2O0\n2+B0\n4O0\n4, whereBm\nnandOm\nnare Stevens CEF\nparameters and operator equivalents, respectively [ 32].\nNote that since the Ce site has hexagonal point sym-\nmetry, there are only two non-zero Stevens parameters\nB0\n2andB0\n4, and therefore there is no mixing of differ-\nent|mJ/angbracketrightstates in the atomic wave functions. The re-\nsults are displayed in Fig. 2(a), with fitted values of\nB0\n2= 1.25 meV and B0\n4= 0.0056 meV, together withmolecular field parameters of λab=−52.8 mol/emu and\nλc=−111.0 mol/emu. This yields the level scheme il-\nlustrated in Fig. 2(b), with a Γ 7ground state doublet\nψ±\nGS=|±1\n2/angbracketright, a low-lying first excited state ψ±\n1=|±3\n2/angbracketright\nseparated by ∆ 1= 5.8 meV from the ground state, and\na second excited state ψ±\n2=| ±5\n2/angbracketrightat ∆2= 22.1 meV.\nNote that a |±1\n2/angbracketrightground state is anticipated, since this\nis the only eigenstate which correspondsto a non-zeroin-\nplane moment, with /angbracketleftµx/angbracketright=/angbracketleftψ∓|gJ(J++J−)/2|ψ±/angbracketright=\n1.28µB/Ce. Since /angbracketleftµx/angbracketrightis much larger than the observed\nlow temperature moment of around 0 .2−0.3µB/Ce, this\nindicates that there is a reduced ordered moment, either\nduetoKondoscreeningprocessesorsignificantzero-point\nfluctuations [ 17,33]. From the large positive value of B0\n2,\ntheab-plane is expected to correspond to the easy direc-\ntion of magnetization, in line with the observed high and\nlow temperature susceptibilities. This suggests that the\nsingle-ion anisotropy arising from the local environment\nof the Ce ions is sufficient to account for the observed\neasy-plane anisotropy, in contrast to many Kondo fer-\nromagnets which order along the hard axis [ 34]. While\nthe negative values of λabandλccould indicate the pres-\nence of coexistent antiferromagnetic correlations, as in\nCeTi1−xVxGe3[35–37], such negative values are often\nfound in Kondo ferromagnets [ 7,38–40], and an effective\nnegative molecular field can arise from the Kondo effect,\nwhich scales with TK[41,42].\nInelastic neutron scattering measurements were per-\nformed on powder samples of CeRh 6Ge4and the non-\nmagnetic analog LaRh 6Ge4, using the MERLIN spec-\ntrometer at the ISIS facility [ 43]. Figure 3displays low\nangle cuts of the data normalized to absolute units at\ntwo temperatures, for four different incident energies Ei.\nNo well defined CEF levels can be detected at energy\ntransfers up to 80 meV. On the other hand, over a large\nrange of energy transfers, the scattering from CeRh 6Ge4\nis consistently larger than the La-analog indicating the4\n/s45/s51 /s48 /s51 /s54/s48/s49/s48/s50/s48/s51/s48/s52/s48/s53/s48\n/s45/s53 /s48 /s53 /s49/s48 /s49/s53/s48/s53/s49/s48/s49/s53/s50/s48/s50/s53\n/s45/s49/s48 /s48 /s49/s48 /s50/s48 /s51/s48/s48/s53/s49/s48/s49/s53/s50/s48/s50/s53\n/s50/s48 /s52/s48 /s54/s48 /s56/s48/s48/s50/s52/s54/s56/s40/s97/s41/s32\n/s69\n/s105/s32/s32/s61/s32/s49/s50/s32/s109/s101/s86\n/s50/s48/s176 /s32/s45/s32/s54/s48/s176 /s32\n/s32/s67/s101/s82/s104\n/s54/s71/s101\n/s52/s32/s32/s55/s75\n/s32/s76/s97/s82/s104\n/s54/s71/s101\n/s52/s32/s32/s32/s55/s75\n/s32/s67/s101/s82/s104\n/s54/s71/s101\n/s52/s32/s32/s49/s48/s48/s75\n/s32/s76/s97/s82/s104\n/s54/s71/s101\n/s52/s32/s32/s32/s49/s48/s48/s75/s83 /s40/s81/s44 /s41/s32/s40/s109/s98/s32/s115/s114/s45/s49\n/s32/s109/s101/s86/s45/s49\n/s32/s102/s46/s117/s46/s45/s49\n/s41/s40/s98/s41\n/s69\n/s105/s32/s32/s61/s32/s50/s48/s32/s109/s101/s86\n/s50/s48/s176 /s32/s45/s32/s54/s48/s176 \n/s40/s99/s41/s32/s32/s32/s69\n/s105/s32/s32/s61/s32/s51/s56/s32/s109/s101/s86\n/s32/s32/s32/s49/s48/s176 /s32/s45/s32/s54/s48/s176 \n/s69/s110/s101/s114/s103/s121/s32/s116/s114/s97/s110/s115/s102/s101/s114/s32/s40/s109/s101/s86/s41/s40/s100/s41\n/s32/s32\n/s69\n/s105/s32/s32/s61/s32/s49/s48/s48/s32/s109/s101/s86\n/s49/s48/s176 /s32/s45/s32/s54/s48/s176 \nFIG. 3. (Color online) Low angle cuts of the inelastic neutro n\nscattering spectra of CeRh 6Ge4and LaRh 6Ge4at 7 K and\n100 K for incident energies of (a) 12 meV, (b) 20 meV, (c)\n38 meV, and (d) 100 meV. The integrated angular ranges are\ndisplayed in the panels.\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48/s48/s50/s52/s54/s56\n/s84 /s32/s61/s32/s55/s32/s75\n/s32/s32\n/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s69\n/s105\n/s32/s49/s50/s32/s109/s101/s86\n/s32/s50/s48/s32/s109/s101/s86\n/s32/s51/s56/s32/s109/s101/s86\n/s32/s49/s48/s48/s32/s109/s101/s86/s83\n/s109/s97/s103/s40/s81/s44 /s41/s32/s40/s109/s98/s32/s115/s114/s45/s49\n/s32/s109/s101/s86/s45/s49\n/s32/s102/s46/s117/s46/s45/s49\n/s41\n/s69/s110/s101/s114/s103/s121/s32/s116/s114/s97/s110/s115/s102/s101/s114/s32/s40/s109/s101/s86/s41/s215 /s48/s46/s49\nFIG. 4. (Color online) Magnetic contribution to the inelast ic\nneutron scattering intensity versus energy transfer at 7 K, for\nfour different incident energies Ei. The solid line shows the\ncalculated inelastic response for the CEF scheme in Fig. 2,\nwhere the FWHM of the quasielastic and inelastic peaks are\n3.3meV(2 TK), while thedashedlineshows thecasewherethe\nquasielastic FWHM is 3.3 meV but the inelastic peak corre-\nsponding to the first CEF excitation has a FWHM of 30 meV.\nNote that the slight mismatch between the data with Ei= 12\nand 20 meV is an artifact arising from slight differences in th e\nnormalization for the different incident energies.\npresence of broad magnetic scattering, extending from at\nleast the elastic line ( ∼1.5 meV for Ei= 12 meV) up\nto at least 60 meV. Note that magnetic scattering could\nnot be resolved on measurements performed at lower en-\nergies on the OSIRIS spectrometer (not displayed) [ 44],\nlikely due to the weak and broad nature of the magnetic\nresponse. The magnetic scattering of CeRh 6Ge4from\nthe MERLIN measurements was estimated by subtract-\ning low angle cuts of the LaRh 6Ge4data, taking into\naccount the different neutron scattering cross sections,\nand the results are shown in Fig. 4. It can be seen that\nthe magnetic scattering is strongest at low energies, butwith a long tail up to high energies. If this broad scat-\ntering were associated with the ground state doublet, i.e.\ncorresponding to quasielastic scattering, this would im-\nply a very large Kondo temperature TKon the order of\nhundreds of Kelvin [ 45]. This is in contrast to the mod-\nerate value of TK= 19 K deduced from comparing the\nmagneticentropytoaspin-1/2Kondomodel[ 16]. Dueto\nthe lack of well-defined excitations, the CEF parameters\nwere fixed to the values from the susceptibility analysis,\nand the solid line in Fig. 4shows the resulting calculated\ninelastic neutron spectra, with a full-width at half maxi-\nmum (FWHM) for all the excitations of 3.3 meV (2 TK).\nIt can be seen that this fails to account for the broad\nmagnetic scattering, and a well defined excitation at ∆ 1\nwould be expected to be observed. Note that no excita-\ntion at ∆ 2is expected, since the dipole matrix elements\nfor the transition from |±1\n2/angbracketrightto|±5\n2/angbracketrightare zero due to the\nneutron selection rules ∆ mJ=±1. On the other hand,\nthedashedlineshowsthe casewithaquasielasticFWHM\nof 2TK, but a much broader inelastic excitation with a\nFWHM of 30 meV, and it can be seen that this scenario\ncan well account for the broad scattering. This suggests\nthat the inelastic neutron scattering results are consis-\ntent with the CEF scheme deduced from the magnetic\nsusceptibility, but with the low-lying CEF excitation at\n5.8 meV being significantly broadened due to hybridiza-\ntion with the conduction electrons.\nThe CEF scheme displayed in Fig. 2(b) can account\nfor the in-plane orientation of the ordered moments of\nCeRh6Ge4belowTC, which was proposed to be vital\nfor generating the necessary entanglement for avoiding\na first-order transition under pressure, allowing for the\noccurrence of a ferromagnetic QCP [ 17]. The angular\ndistributions of the CEF wave functions are also dis-\nplayed. Notably, both the ground state and first ex-\ncited doublet at 5.8 meV primarily have electron density\nout of the basal plane, which may explain the strongly\nanisotropic hybridization revealed by ARPES, with sig-\nnificantly stronger hybridization along the c-axis [20].\nMoreover, the low-lying first excited doublet appears to\nhybridize much more strongly with the conduction elec-\ntrons, which may be a consequence of greater overlap\nwith the out-of-plane Rh(2) and Ge(2) atoms, while the\nground state charge density is orientated towards the\nneighboring Ce atoms [Fig. 2(c)]. Such a scenario with a\nmorestronglyhybridized excited CEF level hasbeen pre-\ndicted to give rise to metaorbital transitions [ 46], which\nwas proposed theoretically for CeCu 2Si2[47], yet has not\nbeen observed experimentally [ 48]. The influence of the\nfirst excited state on the low temperature behavior of\nCeRh6Ge4may be inferred from the Kadowaki-Woods\nratio corresponding to a ground state degeneracy N= 4,\non both sides of the QCP [ 17]. In fact, the angular distri-\nbution of the ground state 4 forbitals has been identified\nas a key parameter for tuning the hybridization of the5\nCe(Co,Ir,Rh)In 5family of heavy fermion superconduc-\ntors [49,50], where more prolate Γ 7ground states are\nassociated with stronger c−fhybridization, likely due\nto strongerhybridizationwith out-of-planeIn atoms[ 51].\nOur results suggest that the anisotropic hybridization is\nnot only driven by the quasi-one-dimensional arrange-\nment of Ce-chains, but by the angular distribution of the\nCEF orbitals arising from the localenvironment of the\nCe atoms.\nIn summary, our neutron diffraction and µSR mea-\nsurements are consistent with FM order in CeRh 6Ge4,\nwith a reduced magnetic moment compared to that ex-\npected from the CEF ground state. We propose a CEF\nscheme which can account for the easy-plane anisotropy\nof CeRh 6Ge4, which was predicted to be crucial for the\noccurrence of FM quantum criticality in this system [ 17].\nMoreover, the broad magnetic scattering observed in in-\nelastic neutron scattering suggests the presence of strong\nc−fhybridization, where the low lying first excited CEF\nlevel couples more strongly than the ground state. This\ncould potentially reconcile there being significant Kondo\nscreening processes which reduce the ordered moment,\nwiththeconclusionoflocalized4 felectronsinferredfrom\nquantum oscillation measurements [ 28]. These results\nsuggest that the anisotropy of the CEF orbitals is an im-\nportant factor in the observed anisotropic hybridization\n[20], and such anisotropic c−fcoupling may also give\nrise to quasi-1D magnetic exchange interactions, which\nhave been proposed to avoid the first-order transition\nubiquitous to isotropic systems [ 17,19]. As such, materi-\nals with similarly anisotropic ground state orbitals could\nbe good candidates for searching for additional quantum\ncritical ferromagnets. It is of particular interest to ex-\nperimentally determine if there is such a correspondingly\nlarge anisotropy in the magnetic exchange interactions of\nCeRh6Ge4, i.e. quasi-one-dimensional magnetism, which\ncould be determined from single crystal inelastic neutron\nscattering or THz spectroscopy.\nWe are very grateful to Piers Coleman for valu-\nable discussions, and to Franz Demmel for sup-\nport with measurements on OSIRIS. This work was\nsupported by the National Key R&D Program of\nChina (No. 2017YFA0303100, No. 2016YFA0300202),\nthe National Natural Science Foundation of China\n(No. 12034107, No. 11874320 and No. 11974306),\nthe Key R&D Program of Zhejiang Province, China\n(2021C01002), and the Science Challenge Project of\nChina (No. TZ2016004). DTA would like to thank the\nRoyal Society of London for Advanced Newton Fellow-\nship funding between UK and China. Experimentsat the\nISISPulsedNeutronandMuonSourceweresupportedby\na beamtime allocation from the Science and Technology\nFacilities Council (RB1820492, RB1820463, RB1820482,\nRB1820611 [ 23,27,43,44]).∗msmidman@zju.edu.cn\n[1] Z. F. Weng, M. Smidman, L. Jiao, X. Lu, and\nH. Q. Yuan, Multiple quantum phase transitions\nand superconductivity in Ce-based heavy fermions,\nReports on Progress in Physics 79, 094503 (2016) .\n[2] G. R. Stewart, Non-Fermi-liquid behavior in d- andf-\nelectron metals, Rev. Mod. Phys. 73, 797 (2001) .\n[3] P. Gegenwart, Q. Si, and F. Steglich, Quantumcriticalit y\nin heavy-fermion metals, Nature Physics 4, 186 (2008) .\n[4] M. Brando, D. Belitz, F. M. Grosche, and T. R.\nKirkpatrick, Metallic quantum ferromagnets,\nRev. Mod. Phys. 88, 025006 (2016) .\n[5] A. Huxley, I. Sheikin, E. Ressouche, N. Kerna-\nvanois, D. Braithwaite, R. Calemczuk, and J. Flou-\nquet, UGe 2: a ferromagnetic spin-triplet superconduc-\ntor,Phys. Rev. B 63, 144519 (2001) .\n[6] M. Uhlarz, C. Pfleiderer, and S. M. Hayden, Quan-\ntum phase transitions in the itinerant ferromagnet zrzn 2,\nPhys. Rev. Lett. 93, 256404 (2004) .\n[7] V. A. Sidorov, E. D. Bauer, N. A. Frederick, J. R. Jef-\nfries, S. Nakatsuji, N. O. Moreno, J. D. Thompson, M. B.\nMaple, and Z. Fisk, Magnetic phase diagram of the fer-\nromagnetic Kondo-lattice compound CeAgSb2up to 80\nkbar,Phys. Rev. B 67, 224419 (2003) .\n[8] H. Kotegawa, T. Toyama, S. Kitagawa, H. Tou,\nR. Yamauchi, E. Matsuoka, and H. Sugawara,\nPressure-temperature-magnetic field phase dia-\ngram of ferromagnetic Kondo lattice CeRuPO,\nJ. Phys. Soc. Jpn. 82, 123711 (2013) .\n[9] U. S. Kaluarachchi, V. Taufour, S. L. Bud’ko, and P. C.\nCanfield, Quantum tricritical point in the temperature-\npressure-magnetic field phase diagram of CeTiGe 3,\nPhys. Rev. B 97, 045139 (2018) .\n[10] D. Belitz, T. R. Kirkpatrick, and T. Vojta, First order\ntransitions and multicritical points in weak itinerant fer -\nromagnets, Phys. Rev. Lett. 82, 4707 (1999) .\n[11] A. V. Chubukov, C. P´ epin, and J. Rech, Instability\nof the quantum-critical point of itinerant ferromagnets,\nPhys. Rev. Lett. 92, 147003 (2004) .\n[12] A. Steppke, R. K¨ uchler, S. Lausberg, E. Lengyel,\nL. Steinke, R. Borth, T. L¨ uhmann, C. Krellner, M. Nick-\nlas, C. Geibel, F. Steglich, and M. Brando, Ferromag-\nnetic quantum critical point in the heavy-fermion metal\nYbNi4(P1−xAsx)2,Science339, 933 (2013) .\n[13] D. T. Adroja, A. D. Hillier, J.-G. Park, W. Kockel-\nmann, K. A. McEwen, B. D. Rainford, K.-H. Jang,\nC. Geibel, and T. Takabatake, Muon spin relaxation\nstudy of non-Fermi-liquid behavior near the ferro-\nmagnetic quantum critical point in CePd 0.15Rh0.85,\nPhys. Rev. B 78, 014412 (2008) .\n[14] N. T. Huy, A. Gasparini, J. C. P. Klaasse, A. de Visser,\nS. Sakarya, and N. H. van Dijk, Ferromagnetic\nquantum critical point in URhGe doped with Ru,\nPhys. Rev. B 75, 212405 (2007) .\n[15] C.-L. Huang, A. M. Hallas, K. Grube, S. Kuntz,\nB. Spieß, K. Bayliff, T. Besara, T. Siegrist, Y. Cai,\nJ. Beare, G. M. Luke, and E. Morosan, Quantum\ncritical point in the itinerant ferromagnet Ni 1−xRhx,\nPhys. Rev. Lett. 124, 117203 (2020) .\n[16] E. Matsuoka, C. Hondo, T. Fujii, A. Oshima, H. Sug-\nawara, T. Sakurai, H. Ohta, F. Kneidinger, L. Sala-6\nmakha, H. Michor, and E. Bauer, Ferromagnetic tran-\nsition at 2.5K in the hexagonal Kondo-lattice compound\nCeRh 6Ge4,J. Phys. Soc. Jpn. 84, 073704 (2015) .\n[17] B. Shen, Y. Zhang, Y. Komijani, M. Nicklas, R. Borth,\nA. Wang, Y. Chen, Z. Nie, R. Li, X. Lu, H. Lee, M. Smid-\nman, F. Steglich, P. Coleman, and H. Yuan, Strange\nmetal behavior in a pure ferromagnetic Kondo lattice,\nNature579, 51 (2020) .\n[18] H. Kotegawa, E. Matsuoka, T. Uga, M. Take-\nmura, M. Manago, N. Chikuchi, H. Sugawara,\nH. Tou, and H. Harima, Indication of ferromagnetic\nquantum critical point in Kondo lattice CeRh 6Ge4,\nJ. Phys. Soc. Jpn. 88, 093702 (2019) .\n[19] Y. Komijani and P. Coleman, Model for a ferromag-\nnetic quantum critical point in a 1D Kondo lattice,\nPhys. Rev. Lett. 120, 157206 (2018) .\n[20] Y. Wu, Y. Zhang, F. Du, B. Shen, H. Zheng, Y. Fang,\nM. Smidman, C. Cao, F. Steglich, H. Yuan, J. D. Den-\nlinger, and Y. Liu, Anisotropic c−fhybridization in\nthe ferromagnetic quantum critical metal CeRh 6Ge4,\nPhys. Rev. Lett. 126, 216406 (2021) .\n[21] T. R. Kirkpatrick and D. Belitz, Ferromagnetic quan-\ntum critical point in noncentrosymmetric systems,\nPhys. Rev. Lett. 124, 147201 (2020) .\n[22] L. C. Chapon, P. Manuel, P. G. Radaelli, C. Ben-\nson, L. Perrott, S. Ansell, N. J. Rhodes, D. Raspino,\nD. Duxbury, E. Spill, and J. Norris, Wish: The new\npowder and single crystal magnetic diffractometer on the\nsecond target station, Neutron News 22, 22 (2011) .\n[23] M. Smidman et al; (2019): STFC\nISIS Neutron and Muon Source,\nhttps://doi.org/10.5286/ISIS.E.RB1820482 .\n[24] D. Vosswinkel, O. Niehaus, U. C. Rode-\nwald, and R. Pottgen, Bismuth flux growth\nof CeRh 6Ge4and CeRh 2Ge2single crystals,\nZeitschrift f¨ ur Naturforschung B 67, 1241 (2012) .\n[25] See Supplemental Material at [] for the simulated mag-\nnetic diffraction pattern corresponding to in-plane ferro-\nmagnetic order.\n[26] P. J. C. King, R. de Renzi, S. P. Cottrell, A. D. Hillier,\nand S. F. J. Cox, ISIS muons for materials and molecular\nscience studies, Physica Scripta 88, 068502 (2013) .\n[27] M. Smidman et al; (2018): STFC\nISIS Neutron and Muon Source,\nhttps://doi.org/10.5286/ISIS.E.99690746 .\n[28] A. Wang, F. Du, Y. J. Zhang, D. Graf, B. Shen,\nY. Chen, Y. Liu, M. Smidman, C. Cao, F. Steglich,\nand H. Q. Yuan, Localized 4 f-electrons in the quan-\ntum critical heavy fermion ferromagnet CeRh 6Ge4,\nScience Bulletin 66, 1389 (2021) .\n[29] C. Cao and J.-X. Zhu, Pressure dependent electronic\nstructure in CeRh 6Ge4,arXiv:2011.14256 (2020) .\n[30] P. Bonf´ a, I. J. Onuorah, and R. D. Renzi, Intro-\nduction and a quick look at MUESR, the Magnetic\nStructure and mUon Embedding Site Refinement Suite,\nJPS Conf. Proc. 21, 011052 (2018) , Proceedings of the\n14th International Conference on Muon Spin Rotation,\nRelaxation and Resonance ( µSR2017).\n[31] F. N. Gygax, A. Schenck, and Y. ¯Onuki, Magnetic prop-\nerties of CeCu 2tested by muon-spin rotation and relax-\nation,J. Phys.: Condens. Matter 16, 2421 (2004) .\n[32] M. Hutchings, Point-charge calculations of energy lev els\nof magnetic ions in crystalline electric fields (Academic\nPress, 1964) pp. 227 – 273.[33] D. T. Adroja, W. Kockelmann, A. D. Hillier, J. Y. So,\nK. S. Knight, andB. D. Rainford, Reducedmoment mag-\nnetic ordering in aKondolattice compound: Ce 8Pd24Ga,\nPhys. Rev. B 67, 134419 (2003) .\n[34] D. Hafner, B. K. Rai, J. Banda, K. Kliemt, C. Krellner,\nJ. Sichelschmidt, E. Morosan, C. Geibel, and M. Brando,\nKondo-lattice ferromagnets and their peculiar order\nalong the magnetically hard axis determined by the crys-\ntalline electric field, Phys. Rev. B 99, 201109(R) (2019) .\n[35] W. Kittler, V. Fritsch, F. Weber, G. Fischer, D. Lam-\nago, G. Andr´ e, and H. v. L¨ ohneysen, Suppression\nof ferromagnetism of CeTiGe 3by V substitution,\nPhys. Rev. B 88, 165123 (2013) .\n[36] M. Inamdar, A. Thamizhavel, and S. K.\nDhar, Anisotropic magnetic behavior of\nsingle crystalline CeTiGe 3and CeVGe 3,\nJ. Phys. Condens. Matter 26, 326003 (2014) .\n[37] M. Majumder, W. Kittler, V. Fritsch, H. v. L¨ ohneysen,\nH. Yasuoka, and M. Baenitz, Competing magnetic corre-\nlations across the ferromagnetic quantumcritical point in\nthe Kondo system CeTi 1−xVxGe3:51V NMR as a local\nprobe,Phys. Rev. B 100, 134432 (2019) .\n[38] K. Katoh, S. Nakagawa, G. Terui, and A. Ochiai, Mag-\nnetic and transport properties of single-crystal YbPtGe,\nJ. Phys. Soc. Jpn. 78, 104721 (2009) .\n[39] C. Krellner and C. Geibel, Mag-\nnetic anisotropy of YbNi 4P2,\nJ. Phys.: Conf. Ser. 391 012032 (2012) 391, 012032 (2012) .\n[40] B. K. Rai, M. Stavinoha, J. Banda, D. Hafner, K. A.\nBenavides, D. A. Sokolov, J. Y. Chan, M. Brando,\nC.-L. Huang, and E. Morosan, Ferromagnetic ordering\nalong the hard axis in the Kondo lattice YbIr 3Ge7,\nPhys. Rev. B 99, 121109(R) (2019) .\n[41] G. Gruner and A. Zawadowski, Magnetic impurities in\nnon-magnetic metals, Rep. Prog. Phys. 37, 1497 (1974) .\n[42] H. R. Krishna-murthy, K. G. Wilson, and J. W. Wilkins,\nTemperature-dependent susceptibility of the symmet-\nric Anderson model: Connection to the Kondo model,\nPhys. Rev. Lett. 35, 1101 (1975) .\n[43] M. Smidman et al; (2018): STFC\nISIS Neutron and Muon Source,\nhttps://doi.org/10.5286/ISIS.E.RB1820492 .\n[44] M. Smidman et al; (2018): STFC\nISIS Neutron and Muon Source,\nhttps://doi.org/10.5286/ISIS.E.RB1820611 .\n[45] E. Bauer, Anomalous properties of Ce-Cu- and Yb-Cu-\nbased compounds, Advances in Physics 40, 417 (1991) .\n[46] K. Hattori, Meta-orbital transition in heavy-fermion sys-\ntems: Analysis by dynamical mean field theory and self-\nconsistent renormalization theory of orbital fluctuations ,\nJ. Phys. Soc. Jpn. 79, 114717 (2010) .\n[47] L. V. Pourovskii, P. Hansmann, M. Ferrero, and\nA. Georges, Theoretical prediction and spectroscopic\nfingerprints of an orbital transition in CeCu 2Si2,\nPhys. Rev. Lett. 112, 106407 (2014) .\n[48] A. Amorese, A. Marino, M. Sundermann, K. Chen,\nZ. Hu, T. Willers, F. Choueikani, P. Ohresser,\nJ. Herrero-Martin, S. Agrestini, C.-T. Chen, H.-\nJ. Lin, M. W. Haverkort, S. Seiro, C. Geibel,\nF. Steglich, L. H. Tjeng, G. Zwicknagl, and A. Sev-\nering, Possible multiorbital ground state in CeCu 2Si2,\nPhys. Rev. B 102, 245146 (2020) .\n[49] T. Willers, Z. Hu, N. Hollmann, P. O. K¨ orner, J. Geg-\nner, T. Burnus, H. Fujiwara, A. Tanaka, D. Schmitz,7\nH. H. Hsieh, H.-J. Lin, C. T. Chen, E. D. Bauer, J. L.\nSarrao, E. Goremychkin, M. Koza, L. H. Tjeng, and\nA. Severing, Crystal-field and Kondo-scale investigations\nof CeMIn5(M= Co, Ir, and Rh): A combined x-\nray absorption and inelastic neutron scattering study,\nPhys. Rev. B 81, 195114 (2010) .\n[50] T. Willers, F. Strigari, Z. Hu, V. Sessi, N. B.\nBrookes, E. D. Bauer, J. L. Sarrao, J. D. Thomp-son, A. Tanaka, S. Wirth, L. H. Tjeng, and\nA. Severing, Correlation between ground state\nand orbital anisotropy in heavy fermion materials,\nProc. Natl. Acad. Sci. U.S.A 112, 2384 (2015) .\n[51] J. H. Shim, K. Haule, and G. Kotliar, Modeling the\nlocalized-to-itinerant electronic transition in the heav y\nfermion system CeIrIn 5,Science318, 1615 (2007) ." }, { "title": "2103.03562v1.Ferromagnetic_Composite_Self_Arrangement_in_Iron_Implanted_Epitaxial_Palladium_Thin_Films.pdf", "content": "Ferromagnetic Composite Self -Arrangement in Iron -Implanted Epitaxial \nPalladium Thin Films \n \nA.I. Gumarov1,2,a), I.V. Yanilkin1,2, R.V. Yusupov2, A.G. Kiiamov2, \nL.R. Tagirov1,2 and R.I. Khaibullin1 \n \n1 Zavoisky Physical -Technical Institute, FRC Kazan Scientific Centre of RAS, 420029 \nKazan, Russia \n2 Institute of Physics, Kazan Federal University, 420008 Kazan, Russia \na) Author to whom correspondence should be addressed: amir@gumarov.ru \n \nABST RACT \nWe report on the formation of the dilute Pd 1-xFex compositions with tunable magnetic \nproperties under an ion -beam implantation of epitaxial Pd thin films. Binary Pd 1-xFex \nalloys with a mean iron content x of 0.025, 0.035 or 0.075 were obtained by the \nimplantation of 40 keV Fe+ ions into the palladium films on MgO (001) substrate to the \ndoses of 0.5∙1016, 1.0∙1016 and 3.0∙1016 ions/cm2, respectively. Structural and magnetic \nstudies have shown that iron atoms occupy regular fcc -lattice Pd -sites without the \nformation of any secondary crystallographic phase. All the iron implanted Pd films reveal \nferromagnetism at low temperatures (below 2 00 K) with both the Curie temperature and \nsaturation magnetization determined by the implanted iron dose. In contrast to the \nmagnetic properties of the molecular beam epitaxy grown Pd 1-xFex alloy films with the \nsimilar iron contents, the Fe -implanted Pd fi lms possess weaker in -plane \nmagnetocrystalline anisotropy, and, accordingly, a lower coercivity. The observed \nmultiple ferromagnetic resonances in the implanted Pd 1-xFex films indicate a formation of \na magnetically inhomogeneous state due to spinodal decom position into regions, \npresumably layers, with identical crystal symmetry but different iron contents. The \nmultiphase magnetic structure is robust with respect to the vacuum annealing at 770 K, \nthough develops towards well -defined local Pd -Fe compositions. \n \nKeywords : ion implantation, diluted palladium -iron alloy, laminar magnetic composite , \nferromagnetism, ferromagnetic resonance \n 2 \n 1. Introduction \nThe renaissance of an interest to Pd1-xFex alloys was triggered recently by their \npotential applications in superconducting spintronics, where diluted compositions with \nx < 0.10 serve as a weak ferromagnetic link in Josephson -junction structures for \nultrahigh -speed cryogenic devices [1–6]. Traditionally, weak ferromagnetic Pd1-xFex \nalloys have been fabricated using the magnetron sputtering [3–10] or molecular -beam \nepitaxy (MB E) [11,12] , including our recent efforts focused on epitaxial thin films of Pd1-\nxFex (0.01 x 0.1) on the MgO (001) substrate [13–16]. However, there is another \ntechnique with high potential to form dilute Fe -Pd alloys with tunable magnetic properties \n– the ion -beam implantation, widely used in modern microelectronics for silicon doping \nand microchip fabrication [17,18] . \nAlthough severa l reports on low -dose ion implantation of magnetic impurities into \nthe pure -Pd films and foils can be found in the literature [19–22], they primarily report \non the influence of irradiation -induced defects on the Curie temperature ( TC) and \nsaturation magnet ization ( Ms). The majority of the reported implantation experiments had \nbeen done at liquid -helium temperature to avoid the annealing of radiation defects during \nion bombardment. Also, the nature of these defects and the impact of thermal annealing \non the magnetic properties were studied [19,20] . \nTargeted by the potential applications of Pd1-xFex alloys in superconducting \nspintronics, we considered a high -dose (heavy) implantation of epitaxial palladium films \nwith iron at room temperature (RT) to obtain wea kly ferromagnetic layers with a \nminimum of irradiation -induced defects. Magnetometry and ferromagnetic resonance \n(FMR) studies have shown that Fe -implanted Pd films reveal ferromagnetism with the TC \nand Ms strongly dependent on the implantation dose as it was expected. In contrast to the \nMBE -grown films with a similar iron content, the implanted films have lower in -plane \nmagnetocrystalline anisotropy, and, accordingly, a lower coercive field. Depending on \ndose, revealed a presence of one to three magnetic p hases with different TC and Ms. It \nlooks that a pronounced separation into different Pd -Fe compositions takes place. These \n‘phases’ are present already in the as -implanted samples and persist (though are modified) \nafter the high -temperature vacuum annealin g indicating the tendency towards separation \ninto definite intrinsically -stable compositions. We discuss this finding in terms of \nspinodal decomposition of initially inhomogeneous distribution of Fe impurity into \nlaminar composite state across the film thi ck-ness. \n 3 \n 2. Experimental section \n2.1. Sample preparation and characterization \nEpitaxial Pd thin films on MgO (001) single -crystal substrate were used as the \nstarting materials. The films were produced from the high purity Pd (99.98%) utilizing \nultrahigh vacuum (UHV) molecular -beam epitaxy system (MBE, by SPECS GmbH, \nGermany ). The three-step synthesis of epitaxial Pd films is described in detail in Ref. [14]. \nThe 56Fe+ ion implantation was performed with the ILU-3 accelerator at fixed ion energy \nof 40 keV and ion currents of 2-3 µA as measured at the sample. To obtain Pd -Fe alloys \nwith various iron concentrations, Fe+ ion doses of 0.5∙1016, 1.0∙1016 and 3.0∙1016 ions/cm2 \n(sample labels S0_5, S1_0 and S3_0, respectively ) were achieved varying the irradiation \ntime. An expected depth profile of Fe -concentration was calculated in advance for \n1.0∙1016 ions/cm2 dose following the SRIM -2013 (TRI M) algorithm [23], and the Fe+ \ndoses were adjusted accordingly to yield the samples with the mean iron contents of \n𝑐̅Fe ≈ 2.5, 3.5 and 7.5 at.%. Note that the original TRIM approach does not account for \nthe sputtering of the target front surface, changes in target composition and diffusion \nduring the implantation . In order to obtain comparable thicknesses of the resulting alloy \nlayer, initial Pd films had the thicknesses of 40 nm, 60 nm and 80 nm for the three above -\nmentioned concentrations of iron, respectively. Film thickness was measured prior to and \nafter Fe+ implantation with the Bruker DektakXT stylus profilometer with an accuracy of \n±0.5 nm. The implanted samples had been cut into parts, and one of them was post -\nannealed in vacuum at 770 K for 20 min. \nThe formation of the Pd 1-xFeх alloy with cubic symmetry corresponding to the crystal \nlattice of pristine Pd matrix was justified ex situ with X-ray diffraction (XRD , Bruker D8 \nAdvance ). XRD studies were performed utilizing the Cu-K radiation ( = 0.15418 nm) \nin the Bragg –Brentano geometry , in the range of 2 θ angles of 40 to 72 degrees . XRD \npatterns are presented in Fig. 1. No sign of any secondary phase ( e.g. metallic Fe \nnanoparticles) could be found at the experimental sensitivity level. The θ-2θ scan \n(Fig. 1a) shows notable shift and broadening of the S3_0 implanted film (002) maxim um \nrelative to the pure Pd and the MBE -grown Pd1-xFeх films due to the lattice parameter \nmodification (see detailed analysis in Ref. [14]) and inhomogeneous distribution of the \niron across the film thickness , respectively . XRD φ-scan for S3_0 sample is shown in \nFig. 1b in compar ison with that for the MBE -grown Pd0.92Fe0.08 film (Fig. 1c). XRD data \nclearly indicate the cubic symmetry , cube -on-cube epitaxy and single -crystalline structure \nof the implanted film . \n 4 \n \nFIG. 1. X-ray diffraction patterns: 2θ-scans ( a) and φ-scans ( b,c) for the S3_0 implanted film \n(red line), MBE -grown epitaxial Pd0.92Fe0.08 film (green line); pure 40 nm Pd film ( blue line) and \nMgO substrate (black line) . Eulerian cradle angle χ was set to 0 degree in 2 θ-scans and was equal \nto 45 degrees in φ-scans in order to detect <202> XRD -maxima (angle φ is arbitrary) . \n \n \nThe iron concentration profiles were obtained by X -ray photoelectron spectroscopy \n(XPS) in combination with depth profiling by the Ar+-ion etching technique. Overview \nXPS spectrum (Fig. 2a) for the implanted film demonstrates the set of lines corresponding \nonly to iron and palladium elements. No alien element s were detected within the \nsensitivity of our spectrometer (~ 0.1%) . Figures 2b and 2c show high -resolution spectra \nof core -shell electrons Fe 2p and Pd 3d recorded with the 0.1 eV step and the analyzer \ntransmission energy of E = 25 eV. Binding energies and line shapes are characteristic of \nPd-Fe alloy studied earlier by XPS [14]. For comparison, we recorded also high resolution \nXPS-spectra of the pure iron and palladium films. A nnealing did not lead to any \nsignificant modification of the shape and position of the XPS-lines. The o btained results \nsignify that Pd -Fe alloys synthetized by ion implantation or molecular -beam epitaxy [14] \nexhibit the same dissolving mechanism : iron atoms substitute for the palladium ones in \nthe lattice without formation of iron nanoparticles . The last is confirmed by the absence \nof a characteristic peak for metallic iron (Fe0) at the binding energy of 706.7 eV [24] \n(within the sensitivity of our XPS setup). \nDepth profiling of the samples was carried out in a step-by-step manner , combining \netching of the sample with argon ions and recording the high -resolution XPS -spectra after \nthe each etching step. The Fe 2p and Pd 3d spectra were used to calculate the concentration \nof iron impurities. The iron impurity was detected across the en tire thickness of every \nFe+-irradiated palladium film. The film thickness after the iron implantation according to \nthe profilometry data was equal to 36 nm, 50 nm and 63 nm for doses of 0.5 ∙1016, 1.0 ∙1016, \n5 \n и 3.0∙1016 ions/cm2, respectively . Thus, a significant sputtering of the front surface of the \nfilms during the implantation occurred. \n \n \n \nFIG. 2. a) A survey XPS spectrum recorded from the S3_0 sample ; b) and c) are the high-\nresolution XPS spectra of the inner shell Fe 2p and Pd 3d electrons, respectively, recorded for the \nas-implanted, annealed in vacuum iron-implanted Pd films , Pd0.92Fe0.08 alloy synthesized by \nMBE and reference films of pure iron and palladium grown separately. \n \n \nDistribution profiles of the iron impurity for the three doses are shown in Fig . 3. The \nshape of the distribution profile changes with the implantation dose and after the \nsubsequent thermal annealing. Obtained distributions have essentially a stepped shape. \nPartial redistribution of the iron impurity over the film thickness takes place during the \nannealing , especially notable for the S0_5 and S3_0 samples . The most of the impurity is \nlocated up to a depth of 35 nm from the surface for the S0_5 and S1_0 sample s, and to a \ndepth of 55 nm for the S3_0 sample . A local maximum of the impurity concentration is \nobserved near the film -substrate interface of the as -implanted S3_0 sample . The iron \nimpurity has not been found inside the substrate for any implantation dose used in this \nwork. Thus, the MgO substrate serve s a barrier for the implanted iron impurity. \n6 \n \nFIG. 3. Distribution p rofiles of the iron impurity in S0_5 (a), S1_0 (b) and S3_0 (c) samples , \nobtained both before (black symbols and curve s) and after the thermal annealing (red symbols \nand curves ). The pink solid line in panel (b) shows the calculated profile of the iron distribution \nin the palladium matri x (see text) . Arrows indicate the position of the film/substrate interface. \n \n \n2.2. Magnetic properties \nMagnetic properties of the iron-implanted Pd films were measured utilizing the \nvibrating sample magnetometry (VSM , Quantum Design PPMS -9) and ferromagnetic \nresonance (FMR , Bruker ESP300) spectroscopy in the temperature range of 5 – 300 K \nwith the magnetic field applied either in-plane or out-of-plane of the films. To determine \nMs of the synthesized Pd -Fe alloy films , the combined diamagnetic contribution of the \nMgO substrate and paramagnetic one from the impurities in it were subtracted from the \nraw VSM data . Then, the saturation magnetic moment was recalculated to the number of \nBohr magnetons ( µB) per implanted Fe+-ion. \nFigure 4a shows magnetic hysteresis loops for palladium films implanted with three \ndifferent doses of iron . For the S0_5 sample, a rather low specific saturation magnetic \nmoment of ≈ 3 µB/Fe was obtained, which is not far from 3.7 µB/Fe reported in [8] for the \nmagnetron sputtered film of Pd0.99Fe0.01. On the other hand, the specific moment of \n 6.5 µB/Fe was found for the MBE grown epitaxial film of 20 nm thickness with \n7 \n x = 0.019 [16], and 6.9 µB/Fe – for the bulk samples [25–27]. Certainly, the higher value \nof the magnetic moment per iron correlates with the synthesis techniques providing more \nuniform distribution of Fe (bulk and MBE samples), while that providing less uniform \ndistribution (magnetron sputtering and ion implantation) lead to its reduc tion. \n \n \nFIG. 4. Magnetic hysteresis loops measured with the magnetic field applied along the hard in-\nplane magnetization axis for the S0_5, S1_0 and S3_0 samples (a); along the in-plane easy and \nhard magnetization axes (b); as well as out -of-plane (inset to panel (b) ), and after the thermal \nannealing of the samples (c). Temperature dependences of the saturation magnetization both \nbefore (open symbols) and after (solid symbols) annealing (d). \n \n \nFigure 4b presents the magnetic hysteresis loops measured with three orientations of \nthe applied magnetic field H with respect to the S3_0 sample plane : the in-plane with \nH||[110] and H||[010] crystallographic direction s of the MgO (001) substrate and the out-\nof-plane geometry with H||[001] (Fig. 4b, inset) . The rectangular shape of the loop show s \nthat the < 110> direction s are the easy in-plane directions , while the <010> are the hard \nin-plane ones in full agreement with the properties of the MBE -grown epitaxial Pd 1-xFex \nfilms [13–16]. Reversible character and a high er value of the saturating magnetic field in \nthe out-of-plane geometry indicate that the obtained material is the easy -plane magnetic \nsystem. Qualitatively similar results were obtained for the S0_5 and S1_0 samples . \n8 \n Figure 4c shows the magnetic hysteresis loops for the implanted palladium films of \nFig. 4a measured after their thermal annealing . The annealing leads to an increase in the \nmagnetization and coercive field of the implanted samples . Nevertheless, both the as -\nimplanted and post -annealed Pd films reveal weaker in -plane magnetocrystalline \nanisotropy and a lower coercivity in comparison with the MBE grown Pd 1xFex films with \nsimilar iron concentration (see Fig. S1 of Supplementary material). At the same time, \nfrom the temperature dependences of the magnetic moment s, Fig. 4d, it follows that TC \nfor all iron implanted Pd films increases after high-temperature annealing in vacuum . The \nmagnetometry results for as -implanted and post -annealed Pd films are summarized in \nTable 1. \n \nTable 1. Experimentally evaluated magnetic properties of as -prepared and post -annealed thin \npalladium films implanted with Fe -ions to different doses D. \nSample label D, \n1016 ions/cm2 𝑐̅Fe, \nat.% Ms, \nµB/Fe Bc*, \nmT TC, \nK \nAs-implanted S0_5 0.5 2.5 3.0 0.24 40 \nS1_0 1.0 3.5 4.5 0.43 107 \nS3_0 3.0 7.5 4.4 0.56 155 \nPost-annealed S0_5a 0.5 2.5 4.2 2.76 105 \nS1_0a 1.0 3.5 5.0 0.58 120 \nS3_0a 3.0 7.5 4.6 0.83 203 \n*The coercivity Bc is measured at 5 К \n \nThe most unusual results for the Fe -implanted thin palladium films were obtained \nwith the ferromagnetic resonance (FMR) spectroscopy [28]. Temperature evolution of the \nFMR spectrum with the magnetic field applied perpendicular to the as-implanted S3_0 \nsample plane is presented in Fig. 5a. A consecutive appearance of the resonances is \nobserved as if ferromagnetic phases emerg e sequentially with the temperature decrease . \nFMR spectra of the same sample for the magnetic field applied along the [110] (in -plane) \nand [001] (out-of-plane) directions are shown in Fig. 5b (black lines) . The modification \nof the FMR spectrum of the S3_0 sample after the thermal annealing is illustrated also by \nFig. 5b (sample S3_0a, red lines). In -plane resonance lines shift after the annealing \ntowards the lower fields, and the two out -of-plane lines – towards the higher fields being \na common feature of the spectra evolution for all the sampl es. The multiphase response \nof the magnetic system persists after the annealing. 9 \n The angular behavior of the resonance fields of both the as -implanted and the \nannealed samples is typical for the easy -plane systems : in-plane resonances reside in \nlower magne tic fields with respect to the electron paramagnetic resonance lines (a comb \ncentered around 330 mT originating from the 3d -ion impurities in the substrate ), while \nthe out-of-plane resonances are found at higher fields due to the thin-film 2D shape \nanisotr opy [28]. For comparison, the FMR spectra of the homogeneous epitaxial \nPd0.92Fe0.08 film [13] are presented in Fig. 5b. \nResonance field s for th e annealed Fe -implanted Pd -film, sample S3_0 a, in the out -\nof-plane orientation have the same values as that for the epitaxial Pd 1-xFex films [16] with \nthe iron concentrations of x = 0.025, x = 0.04 and x = 0.063. The S0_5 a sample with the \nlowest dose reveals only one FMR signal at T ≤ 40K (inset to Fig. 5b and Fig. S2a of the \nSupplementary material ), and the S1_0 a sample – two FMR signals (see Fig. S2b). \n \n \nFIG. 5. FMR spectra of the epitaxial Pd film implanted with iron to the dose of \n3.0∙1016 ions/cm2: (a) temperature evolution for the as -implanted S3_0 sample in the out-of-\nplane geometry ; (b) in-plane and out-of-plane spectra of the as-implant ed (sample S3_0, black \ncurve s) and the annealed (sample S3_0a, red curve s) films , as well as the spectra of the MBE -\ngrown epitaxial Pd 0.92Fe0.08 film (green curves ). The inset shows the resonance lines of the \nannealed S0_5 a and S1_0 a samples (ν = 9.416 GHz, T = 40 K). \n \n \n3. Discussion \nObservation of multiple ferromagnetic resonances unambiguously indicates an \noccurrence of the corresponding set of rather well -defined , in the sense of the Ms values , \nferromagnetic components in our Fe -implanted palladium films. This, in turn, is a \nmanifestation of the Pd -Fe system separation into definite compositions. In our opinion, \nthis can take place due to a kind of a spinodal decomposition [29–32] occurring under \nnon-equilibrium conditions in the course of the implantation, that lead, most probably, to \na formation of a laminar composite magnetic structure in the implanted Pd 1‐xFex films. \n10 \n Indeed, because of the doping technique, the implant distribution is in homogeneous \na priori , see Fig. 3. The distribution shape changes with increasing the dose because of \nthe front surface sputtering during the irradiation, and diffusion stimulated by the \nannealing [33] due to a retardation of the penetrating ions with the kinetic energy mostly \ntransferred to the heat. This could be a reason of an absence of significant evolution of \nthe depth distribution profiles and magnetic hysteresis loops, Fig. 4, upon post -\nimplantation thermal annealing. Single -crystal state of the Fe -implanted Pd films (see \nFig. 1) also points at a dynamic annealing of radiation defects in the films directly during \nthe ion irradiation. The S0_5 sample implanted with the lowest dose of 0.5∙1016 ions/ cm2 \n(the lowest time of implantation) seems to acquire minimal self-annealing, that is why it \nshows maximal changes in the hysteresis loop shape, Ms and TC upon ex situ annealing, \nFig. 4, c and d. \nIn spite of inhomogeneous distribution of iron atoms, there are no indications of \nsequential reversals of different magnetic phases in hysteresis loops (Figs. 4, a to c ). \nIndeed, if the remagnetization mechanism is a movement of domain walls, then the most \n”soft” region (sublayer) of the film triggers at low fields and drags the coupled \nneighboring sublayers to complete reversal of the entire magnetic moment of a film. No \nmultiphase , laminar nature of magnetism could be recognized in the quasi -static magnetic \nmoment measurements , Fig. 4 and Fig. S1 of the Supplementary material . The regions \nwith high iron content could lead to an increas e in coercivity due to the domain -wall \npinning by the coupling with them (see Fig. S1). \nOur FMR studies revealed that in the spectrum for a high-dose S3_0 sample the \nnumber of resonance lines increases from one to three with decreasing the temperature \n(Fig. 5a). This feature is quite robust, because the high -temperature anneali ng of the \nsample does not qualitatively change its FMR response (Fig. 5b). Probably, the \ninhomogeneous profile of the iron distribution in the implanted palladium films leads to \nthe formation of a laminar multiphase magnetic system with different temperatu res of \nferromagnetic ordering and saturation magnetizations for each individual phase (layer) . \nIn spite of a direct contact of these phases (layers), and the resulting coupling between \nthem, they show robust individual responses with the angular behavior e xactly mimicking \nthe FMR response of thin ferromagnet ic film with the easy-plane shape anisotropy \n[13,16,28] . \nMoreover, for palladium films implanted with a minimum ( 0.5∙1016 ions/cm2, S0_5 ) \nand medium ( 1.0∙1016 ions/cm2, S1_0 ) iron doses, after bringing them to thermodynamic \nequilibrium by the annealing, we observed one FMR line for the fi rst, and two FMR lines 11 \n for the second (Fig. S2 in the Supplementary material). The closeness of the resonance \nfields for the selected lines (see Fig . S3 of the Supplementary material) supports a \nformulation of a simple model of spinodal decomposition into isostructural laminar \nstructure [29], each layer possessing a certain , quite well-defined concentration of iron in \nthe palladium . That is, the system evolves from a metastable non -equilibrium state created \nby ion implantation , to its energetically favorable state by the formation of several \nspatially -extended (macroscopic) “stable phases ”. The shift of the resonance line at \n~ 420 mT in Fig. 5b towards lower resonance fields, as well as the appearance of \ninflections in the temperature dependence of magnetization for low and medium doses \n(Fig. 4d) after thermal annealing support this conjecture. As an ultimate hypothesis on \nthe micro scopic scale, the FMR technique reveals fine preferences for certain \ncompositio ns of Pd -Fe system that highlights the effects of preferential local order. \nDisclosure of this previously unknown tendency of the Pd 1-xFex binary system in the \npalladium -rich ( x < 0.10) composition range opens up new prospects for the use of these \nmaterials in superconducting spintronic heterostructures: the reported intrinsically stable \ncompositions of x = 0.025, x = 0.04, x = 0.063 (prelim inary estimates for uncoupled \nlayers) should be utilized ensuring the magnetic homogeneity of the ferromagnetic layers. \nIon-beam implantation of epitaxial palladium films with relatively low doses of iron can \nserve a route for a synthesis of near -homogeneo us low -temperature ferromagnets. \n \n4. Conclusions \nTo summarize, we synthesized Pd -Fe alloys with a mean iron content in the range of \n2.5-7.5 at.% using high -dose Fe-ion implantation to epitaxial Pd -films at RT . Iron-\nimplanted Pd films retain fcc cubic crystal structure typical for binary Pd -Fe alloys with \nlow content of iron. All synthesized samples exhibit ferromagnetism at low temperatures. \nThe Curie temperature and saturation magnetization depend on the dose o f implantation \nand are modified under subsequent thermal annealing. Magnetic hysteresis loops for the \nimplanted palladium films are narrower than those for Pd -Fe alloys with comparable \nconcentrations grown by MBE. FMR and VSM studies indicate that implantation of \nepitaxial palladium films with iron leads to the unusual formation of a laminar magnetic \ncomposite in otherwise crystallographically homogeneous film . The multiphase behavior \nis discussed within the hypothesis of spinodal deco mposition of the initially non-\nequilibrium distribution of inhomogeneously delivered dopant to isostructural , laterally \nextended laminar sub-structures. \n 12 \n CRediT au thorship contribution statement : \nGumarov A.I.: Conceptualization, Sample preparation, Measurements, Data curation, \nWriting – original draft; Yanilkin I.V.: Measurements , Data curation; Yusupov R.V. : \nInvestigation, Writing – review & editing ; Kiiamov A.G. : X-ray data acquisition & \ncuration ; Tagirov L.R. : Formal analysis, Writing – original draft, Writing – review & \nediting ; Khaibullin R.I.: Conceptualization, Methodology, Sample preparation, \nSupervision, Funding acquisition. \n \nDeclaration of competing interest : \nThe authors declare that they have no competing interests. \n \nAcknowledgements : \nAuthors thank V.I. Nuzhdin and V.F. Valeev for technical assistance with ion \nimplantation experiments. \n \nFunding: \n \nThis work was supported by the RFBR Grant No. 20-02-00981. \n \nData available on request from the authors: \nThe data that support the findings of this study are available from the corresponding \nauthor upon a reasonable request. \n 13 \n References \n[1] I.A. Golovchanskiy, V. V. Bolginov, N.N. Abramov, V.S. Stolyarov, A. Ben Hamida, V.I. \nChichkov, D. Roditchev, V.V. Ryazanov, Magnetization dynamics in dilute Pd1 -xFex thin \nfilms and patterned microstructures considered for superconducting electronics, J. Appl. \nPhys. 120 (2016) 163902. https://doi.org/10.1063/1.4965991. \n[2] V.V. Bol’ginov, V.S. Stolyarov , D.S. Sobanin, A.L. Karpovich, V.V. Ryazanov, Magnetic \nswitches based on Nb -PdFe -Nb Josephson junctions with a magnetically soft \nferromagnetic interlayer, JETP Lett. 95 (2012) 366 –371. \nhttps://doi.org/10.1134/S0021364012070028. \n[3] T.I. Larkin, V. V. Bol’g inov, V.S. St olyarov, V.V. Ryazanov, I. V. Vernik, S.K. Tolpygo, \nO.A. Mukhanov, Ferromagnetic Josephson switching device with high characteristic \nvoltage, Appl. Phys. Lett. 100 (2012) 222601. https://doi.org/10.1063/1.4723576. \n[4] V.V. Ryazanov, V.V. Bol’gi nov, D.S. Sobanin, I.V. Vernik, S.K. Tolpygo, A.M. Kadin, \nO.A. Mukhanov, Magnetic Josephson Junction Technology for Digital and Memory \nApplications, Phys. Procedia. 36 (2012) 35 –41. \nhttps://doi.org/10.1016/J.PHPRO.2012.06.126. \n[5] I.V. Vernik, V. V. Bol’Gin ov, S.V. Bakurskiy, A. A. Golubov, M.Y. Kupriyanov, \nV.V. Ryazanov, O.A. Mukhanov, Magnetic josephson junctions with superconducting \ninterlayer for cryogenic memory, IEEE Trans. Appl. Supercond. 23 (2013) 1701208 –\n1701208. https://doi.org/10.1109/TASC.2012.22 33270. \n[6] J.A. Glick, R. Loloee, W.P. Pratt,, N.O. Birge, Critical Current Oscillations of Josephson \nJunctions Containing PdFe Nanomagnets, IEEE Trans. Appl. Supercond. 27 (2016) 1 –5. \nhttps://doi.org/10.1109/TASC.2016.2630024. \n[7] H.Z. Arham, T.S. Khaire, R. Loloee, W.P. Pratt, N.O. Birge, Measurement of spin memory \nlengths in PdNi and PdFe ferromagnetic alloys, Phys. Rev. B - Condens. Matter Mater. \nPhys. 80 (2009) 174515. https://doi.org/10.1103/PhysRevB.80.174515. \n[8] L.S. Uspenskaya, A.L. Rakhman ov, L.A. Dorosinskii, S.I. Bozhko, V.S. Stolyarov, \nV.V. Bolginov, Magnetism of ultrathin Pd 99 Fe 01 films grown on niobium, Mater. Res. \nExpress. 1 (2014) 036104. https://doi.org/10.1088/2053 -1591/1/3/036104. \n[9] L.S. Uspenskaya, I.V. Shashkov, Influence o f Pd 0.99Fe0.01 film thickness on magnetic \nproperties, Phys. B Condens. Matter. (2017). https://doi.org/10.1016/j.physb.2017.09.089. \n[10] V.V. Bol’ginov, O.A. Tikhomirov, L.S. Uspenskaya, Two -component magnetization in \nPd99Fe01 thin films, JETP Lett. 105 (2 017) 169 –173. \nhttps://doi.org/10.1134/S0021364017030055. \n[11] M. Ewerlin, B. Pfau, C.M. Günther, S. Schaffert, S. Eisebitt, R. Abrudan, H. Zabel, \nExploration of magnetic fluctuations in PdFe films, J. Phys. Condens. Matter. 25 (2013) \n266001. https://doi.or g/10.1088/0953 -8984/25/26/266001. \n[12] I.A. Garifullin, D.A. Tikhonov, N.N. Garif’yanov, M.Z. Fattakhov, K. Theis -Bröhl, 14 \n K. Westerholt, H. Zabel, Possible reconstruction of the ferromagnetic state under the \ninfluence of superconductivity in epitaxial V/Pd 1−x Fex bilayers, Appl. Magn. Reson. 22 \n(2002) 439 –452. https://doi.org/10.1007/BF03166124. \n[13] A. Esmaeili, I.R. Vakhitov, I.V. Yanilkin, A.I. Gumarov, B.M. Khaliulin, B.F. Gabbasov, \nM.N. Aliyev, R.V. Yusupov, L.R. Tagirov, FMR Studies of Ultra -Thin Epita xial \nPd0.92Fe0.08 Film, Appl. Magn. Reson. 49 (2018) 175 –183. https://doi.org/10.1007/s00723 -\n017-0946 -1. \n[14] A. Esmaeili, I. V. Yanilkin, A.I. Gumarov, I.R. Vakhitov, B.F. Gabbasov, A.G. Kiiamov, \nA.M. Rogov, Y.N. Osin, A.E. Denisov, R.V. Yusupov, L.R. Tag irov, Epitaxial growth of \nPd1−xFex films on MgO single -crystal substrate, Thin Solid Films. 669 (2019) 338 –344. \nhttps://doi.org/10.1016/j.tsf.2018.11.015. \n[15] W.M. Mohammed, I.V . Yanilkin, A.I. Gumarov, A.G. Kiiamov, R.V . Yusupov, \nL.R. Tagirov, Epitaxial growth and superconducting properties of thin -film PdFe/VN and \nVN/PdFe bilayers on MgO(001) substrates, Beilstein J. Nanotechnol. 11 (2020) 807 –813. \nhttps://doi.org/10.3762/bjnano .11.65. \n[16] A. Esmaeili, I.V. Yanilkin, A.I. Gumarov, I.R. Vakhitov, B.F. Gabbasov, R.V. Yusupov, \nD.A. Tatarsky, L.R. Tagirov, Epitaxial thin -film Pd1 -xFex alloy: a tunable ferromagnet \nfor superconducting spintronics, Sci. China Mater. (2020). \nhttps://doi .org/10.1007/s40843 -020-1479 -0. \n[17] L.A. Larson, J.M. Williams, M.I. Current, Ion implantation for semiconductor doping and \nmaterials modification, Rev. Accel. Sci. Technol. Accel. Appl. Ind. Environ. 04 (2012) \n11–40. https://doi.org/10.1142/S179362681100 0616. \n[18] R.W. Hamm, M.E. Hamm (eds.) , Industrial accelerators and their applications, World \nScientific , 2012. https://doi.org/10.1142/7745. \n[19] J.D. Meyer, B. Stritzker, Reduced Susceptibility of Irradiated Palladium, Phys. Rev. Lett. \n48 (1982) 502–505. https://doi.org/10.1103/PhysRevLett.48.502. \n[20] M. Hitzfeld, P. Ziemann, W. Buckel, H. Claus, Ferromagnetism of alloys: A new probe in \nion implantation, Solid State Commun. 47 (1983) 541 –544. https://doi.org/10.1016/0038 -\n1098(83)90495 -7. \n[21] M. Hitzfeld, P. Ziemann, W. Buckel, H. Claus, Ferromagnetism of Pd -Fe alloys produced \nby low -temperature ion implantation, Phys. Rev. B. 29 (1984) 5023 –5030. \nhttps://doi.org/10.1103/PhysRevB.29.5023. \n[22] H. Claus, N.C. Koon, Effect of oxygen on the electroni c properties of Pd, Solid State \nCommun. 60 (1986) 481 –484. https://doi.org/10.1016/0038 -1098(86)90721 -0. \n[23] J.F. Ziegler, M.D. Ziegler, J.P. Biersack, SRIM - The stopping and range of ions in matter \n(2010), Nucl. Instruments Methods Phys. Res. Sect. B Be am Interact. with Mater. Atoms. \n268 (2010) 1818 –1823. https://doi.org/10.1016/j.nimb.2010.02.091. \n[24] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, R.S.C. Smart, 15 \n Resolving surface chemical states in XPS analysis of first row transit ion metals, oxides \nand hydroxides: Cr, Mn, Fe, Co and Ni, Appl. Surf. Sci. 257 (2011) 2717 –2730. \nhttps://doi.org/10.1016/j.apsusc.2010.10.051. \n[25] J. Crangle, W.R. Scott, Dilute ferromagnetic alloys, J. Appl. Phys. 36 (1965) 921 –928. \nhttps://doi.org/10.10 63/1.1714264. \n[26] G.J. Nieuwenhuys, Magnetic behaviour of cobalt, iron and manganese dissolved in \npalladium, Adv. Phys. 24 (1975) 515 –591. https://doi.org/10.1080/00018737500101461. \n[27] B. Heller, K.H. Speidel, R. Ernst, A. Gohla, U. Grabowy, V. Roth, G. Jakob, F. Hagelberg, \nJ. Gerber, S.N. Mishra, P.N. Tandon, Transient field measurement in the giant moment \nPdFe alloy, Nucl. Instruments Methods Phys. Res. Sect. B Beam Interact. with Mater. \nAtoms. 142 (1998) 133 –138. https://doi.org/10.1016/S0168 -583X(98) 00260 -2. \n[28] M. Farle, Ferromagnetic resonance of ultrathin metallic layers, Reports Prog. Phys. 61 \n(1998) 755 –826. https://doi.org/10.1088/0034 -4885/61/7/001. \n[29] J.W. Cahn, On spinodal decomposition, Acta Metall. 9 (1961) 795 –801. \nhttps://doi.org/10.10 16/0001 -6160(61)90182 -1. \n[30] V.P. Skripov, A.V . Skripov, Spinodal decomposition (phase transitions via unstable \nstates), Sov. Phys. - Uspekhi. 22 (1979) 389 –410. \nhttps://doi.org/10.1070/PU1979v022n06ABEH005571. \n[31] System s Far from Equilibrium (ed. by Luis Garrido), Springer -Verlag , Berlin -Heidelberg -\nNew York, 1980 . https://doi.org/ 10.1007/BFb0025609 . \n[32] Phase Transformations in Materials (ed. by Gernot Kostorz) , Wiley VCH Verlag GmbH , \nWeinheim, 2001. https://doi.org/10.1002/352760264X. \n[33] A.A. Achkeev, R.I. Khaibullin, L.R. Tagirov, A. Mackova, V. Hnatowicz, N. Cherkashin, \nSpecific features of depth distribution profiles of implanted cobalt ions in rutile TiO 2, \nPhys. Solid State. 53 (2011) 543 –553. h ttps://doi.org/10.1134/S1063783411030024. SUPPLEMENTARY MATERIAL \nfor \nFerromagnetic Composite Self -Arrangement in Iron -Implanted Epitaxial \nPalladium Thin Films \n \nA.I. Gumarov1,2,a), I.V. Yanilkin1,2, R.V. Yusupov1,2, A.G. Kiiamov2, \nL.R. Tagirov1,2 and R.I. Khaibullin1 \n \n1 Zavoisky Physical -Technical Institute, FRC Kazan Scientific Centre of RAS, 420029 \nKazan, Russia \n2 Institute of Physics, Kazan Federal University, 420008 Kazan, Russia \na) Author to whom correspondence should be addressed: amir@gumarov.ru \n \n \nComparison of Pd 1-xFex films data: Ion implantation vs. MBE \n \n \nVSM studies \n \n \nFIG S1. Hysteresis loops at 5 K (a) and M(T) curves (b) measured for 80 nm epitaxial Pd film \nimplanted with 40 keV Fe+-ions with the dose of 3∙1016 ions/cm2 (○); the same sample post -\nannealed in vacuum at 770 K for 20 min ( ●); epitaxial film of Pd 0.92Fe0.08 alloy deposited by \nMBE to (100) MgO substrate ( ●). \n \n \n \n \n2 \n FMR studies \n \n \nFIG S2. FMR spectra obtained in the in -plane (dotted line) and out -of-plane (solid line) \nmeasurement geometries for epitaxial Pd thin film implanted with 40 keV Fe+-ions with the \ndose of 0.5∙1016 ions/cm2 (a) and 1.0 ∙1016 ions/cm2 (b). The two samples have been measured \nafter vacuum annealing at 770 K for 20 min. Arrows indicate the observed resonance lines in \nthe out -of-plane orientation. \n \n \n \n \nFIG S3. FMR spectra for all three Fe -implanted palladium films after the annealing for \ncomparison of the resonance fields. \n \n \n" }, { "title": "2103.11953v1.Field_induced_reorientation_of_helimagnetic_order_in_Cu__2_OSeO__3__probed_by_magnetic_force_microscopy.pdf", "content": "Field-induced reorientation of helimagnetic order in Cu 2OSeO 3probed by magnetic\nforce microscopy\nPeter Milde,1,\u0003Laura K ohler,2, 3Erik Neuber,1P. Ritzinger,1Markus Garst,2, 3, 4\nAndreas Bauer,5Christian P\reiderer,5H. Berger,6and Lukas M. Eng1, 7\n1Institute of Applied Physics, Technische Universit at Dresden, D-01062 Dresden, Germany\n2Institute of Theoretical Physics, Technische Universit at Dresden, D-01062 Dresden, Germany\n3Institute for Theoretical Solid State Physics, Karlsruhe Institute of Technology, D-76131 Karlsruhe, Germany\n4Institute for Quantum Materials and Technology,\nKarlsruhe Institute of Technology, D-76344 Eggenstein-Leopoldshafen, Germany\n5Physik-Department, Technische Universit at M unchen, D-85748 Garching, Germany\n6Institut de Physique de la Mati\u0012 ere Complexe, \u0013Ecole Polytechnique F\u0013 ed\u0013 erale de Lausanne, 1015 Lausanne, Switzerland\n7Dresden-W urzburg Cluster of Excellence { Complexity and Topology\nin Quantum Matter (ct.qmat), TU Dresden, 01062 Dresden, Germany\n(Dated: March 23, 2021)\nCu2OSeO 3is an insulating skyrmion-host material with a magnetoelectric coupling giving rise to\nan electric polarization with a characteristic dependence on the magnetic \feld ~H. We report mag-\nnetic force microscopy imaging of the helical real-space spin structure on the surface of a bulk single\ncrystal of CuO 2SeO 3. In the presence of a magnetic \feld, the helimagnetic order in general reorients\nand acquires a homogeneous component of the magnetization, resulting in a conical arrangement at\nlarger \felds. We investigate this reorientation process at a temperature of 10 K for \felds close to\nthe crystallographic h110idirection that involves a phase transition at Hc1. Experimental evidence\nis presented for the formation of magnetic domains in real space as well as for the microscopic origin\nof relaxation events that accompany the reorientation process. In addition, the electric polariza-\ntion is measured by means of Kelvin-probe force microscopy. We show that the characteristic \feld\ndependency of the electric polarization originates in this helimagnetic reorientation process. Our\nexperimental results are well described by an e\u000bective Landau theory previously invoked for MnSi,\nthat captures the competition between magnetocrystalline anisotropies and Zeeman energy.\nI. INTRODUCTION\nIn the limit of weak spin-orbit coupling \u0015SOC the\ncubic chiral magnets like MnSi [1], Fe 1-xCoxSi [2, 3],\nFeGe [4, 5], and Cu 2OSeO 3[6, 7] are dominated by\nonly two coupling constants, the symmetric and anti-\nsymmetric exchange interaction, JandD, respectively\n[8]. Whereas the symmetric exchange is of zeroth order\nin\u0015SOC, the Dzyaloshinskii-Moriya interaction is of \frst\norderD\u0018O(\u0015SOC). Their ratio determines the charac-\nteristic wavevector Q=D=J of the helimagnetic order\nthat develops at zero magnetic \feld ~H= 0. For \fnite ~H,\nthe helix assumes a conical arrangement until it is fully\npolarized at the internal critical \feld \u00160Hint\nc2'D2\nJM s, with\nMsthe saturation magnetization. In addition, a small\npocket of the skyrmion lattice phase just below the or-\ndering temperature Tcis realized at intermediate values\nof~H.\nAs a result, the above-mentioned materials share a very\nsimilar magnetic phase diagram. Details of it, however,\ndepend on corrections that are parametrically smaller in\n\u0015SOC. In particular, the orientation of the helimagnetic\norder at zero \feld is determined by magnetocrystalline\nanisotropies, that are at least of fourth order in \u0015SOC,\nand generally favour the spin spiral to align either along\n\u0003peter.milde@tu-dresden.dea crystallographic h111iorh100idirection like, e.g., in\nMnSi or Cu 2OSeO 3, respectively. The Zeeman energy\ncompetes with the magnetocrystalline anisotropies re-\nsulting in a reorientation of helimagnetic order with vary-\ning magnetic \feld. Depending on the history and the\npopulation of domains, this reorientation process might\neither correspond to a crossover, or involves a \frst-order\nor second-order phase transition at the critical \feld Hc1\n[2, 9{12].\nA quantitative theory of this reorientation process that\nis valid in the limit of small \u0015SOCwas recently presented\nby Bauer et al. and veri\fed by detailed experiments on\nMnSi [12]. With the help of dc and ac susceptibilities as\nwell as neutron scattering experiments, the evolution of\nthe helix orientation, speci\fed by the unit vector ^Q(~H),\nwas carefully tracked as a function of magnetic \feld for\nvarious \feld directions. The crystallographic h100idirec-\ntion plays a special role in that two subsequent Z2tran-\nsitions could be observed con\frming a theoretical predic-\ntion of Walker [11].\nAccording to the theory of Ref. [12] the di\u000berential\nmagnetic susceptibility @HMnaturally decomposes into\ntwo parts. Whereas the \frst part derives from the helix\nwith a \fxed axis ^Q, the second part is attributed to the\n\feld dependence of ^Q(~H). The reorientation ^Q(~H) is as-\nsociated with large relaxation times \u001cbecause it requires\nthe rotation of macroscopic helimagnetic domains. As a\nconsequence, the ac susceptibility for frequencies !\u001c\u001d1\nis only sensitive to the \frst part, which was experimen-arXiv:2103.11953v1 [cond-mat.mtrl-sci] 22 Mar 20212\ntally con\frmed in Ref. [12] suggesting relaxation times\nexceeding seconds, \u001c\u00151 sec. Generally, the reorienta-\ntion depends on the history of the sample due to di\u000berent\ndomain populations, for example, realized for \fnite- or\nzero-\feld cooling. In particular, hysteresis was found at\nthe second-order phase transition at Hc1. The decrease of\nthe \feld across Hc1is accompanied with the formation of\nmultiple domains. The coexistence of di\u000berent domains\nwithin the sample might hamper the realization of the\noptimal trajectory ^Q(~H), especially, in the presence of\nlong relaxation times \u001c. As a result, distinct behavior\ncan be observed upon increasing and decreasing the \feld\nacrossHc1.\nIn Ref. [12] only bulk probes were experimentally inves-\ntigated so that the microscopic origin of the slow relax-\nation processes could not be identi\fed. However, it was\nspeculated that topological defects of the helimagnetic\norder, i.e., disclination and dislocations, might play a spe-\ncial role as they should naturally arise at the boundaries\nbetween di\u000berent domains. A slow creep-like motion of\ndislocations was indeed identi\fed by magnetic force mi-\ncroscopy (MFM) measurements on the surface of FeGe\nsamples by Dussaux et al. [13] after the system had been\nquenched from the \feld-polarized state to ~H= 0. The\nmotion of dislocations during a MFM scan results in dis-\ncontinuities of the helical pattern in the MFM image con-\nsisting of characteristic 180\u000ephase shifts. Subsequently,\nit was also demonstrated both experimentally and the-\noretically that domain walls might comprise topological\ndisclination and dislocation defects [14]. Nevertheless, a\nmicroscopic investigation of such relaxation events close\ntoHc1has not been achieved so far.\nIn the present work, we investigate the helix reorien-\ntation in the chiral magnet Cu 2OSeO 3using microscopic\nMFM measurements. This material is an insulator with a\nmagnetoelectric coupling that allows to manipulate mag-\nnetic skyrmions and helices with electric \felds, and it\ngives rise to various interesting magnetoelectric e\u000bects\n[15, 16]. This material is also promising for magnonic\napplications due to its very low Gilbert damping param-\neter [17]. In constrast to MnSi, its helix is oriented along\nah100idirection at zero \feld. The relatively large ra-\ntioHc1=Hc2\u00180:36 of Cu 2OSeO 3[18] suggests that the\nspin-orbit coupling constant \u0015SOCis larger than in MnSi.\nIndeed, additional magnetic phases stabilized by magne-\ntocrystalline anisotropies { the (metastable) canted con-\nical state as well as the low temperature skyrmion lattice\nphase { were found in Cu 2OSeO 3at low temperatures\nbut for~Honly aligned along crystallographic h100idirec-\ntions [18{20]. Recently, real space observations address-\ning these states have been reported for a thin Cu 2OSeO 3\nlamella and \feld along a h100idirection [21].\nIn previous work [22], we have already investigated\nCu2OSeO 3with MFM at higher temperatures close to\nTcand identi\fed all the magnetic phases, i.e., the helical\nand conical helimagnetic textures, the skyrmion lattice\nphase, and the \feld-polarized phase. Using Kelvin-probe\nforce microscopy (KPFM) we determined the electric po-larization and its \feld-dependence within these various\nphases. However, the reorientation process was not ad-\ndressed in Ref. [22] and it is at the focus of the present\nwork.\nDue to the restriction of our experimental setup, the\nmagnetic \feld is always aligned perpendicular to the\nplane that is scanned by MFM, and for our sample probe\nthis corresponds approximately to the crystallographic\n[110] direction, ~Hk[110]. We study the helix reorien-\ntation for this \feld direction and we determine the pe-\nriodicity of the periodic surface pattern and its in-plane\norientation. Assuming that the bulk helimagnetic order\nessentially extends towards the surface, we extract the\norientation of the helix as a function of magnetic \feld.\nIn addition, we determine the electric polarization and\nits behavior during the reorientation process. Our re-\nsults are interpreted within the e\u000bective Landau theory\nof Ref. [12] that, strictly speaking, is only controlled for\nsmall\u0015SOC, and we \fnd good agreement between the-\nory and experiment. Moreover, we present microscopic\nevidence that the motion of dislocations along domain\nboundaries contributes to the magnetic relaxation close\nto the reorientation transition.\nThe structure of the paper is as follows. In section II\nwe present the experimental methods. In section III we\nshortly review the theory of Ref. [12] and discuss its ap-\nplication to Cu 2OSeO 3. In particular, we point out the\npresence of a robust Z3transition for \felds along h111i.\nThe theoretical prediction for the current experimental\nsetup are presented and the electric polarization is eval-\nuated as a function of the applied magnetic \feld. The\nexperimental results are presented in section IV. From\nthe MFM images we extract the orientation of the heli-\nmagnetic order and the presence of various domains as a\nfunction of magnetic \feld. The relaxation processes are\nshortly analysed, and the electric polarization is deter-\nmined. Finally, we \fnish with a discussion of our results\nand a summary in section V.\nII. EXPERIMENTAL METHODS\nWe investigate the same (70 \u00065)µm thick plate sample\nwith a polished crystallographic (110) surface, as in our\nearlier work [22], where all details on the sample prepa-\nration can be found. Choosing a lower temperature at\nT= 10 K compared to the former study ensured much\nslower dynamics and accessing a broader transition re-\ngion enabled the detailed inspection of the reorientation\nof the helix axis as well as the observation of helical do-\nmains.\nFor real-space imaging, we use magnetic force mi-\ncroscopy (MFM), that proved to be a valuable tool for\nstudying complex spin textures such as magnetic bub-\nble domains [23] or helices and skyrmions in helimagnets\nand magnetic thin \flms [14, 24{28]. In the presence of\nan electric polarization, also electrostatic forces act on\nthe MFM-tip. Compensating these forces by means of3\nKelvin-probe force microscopy (KPFM) permits the de-\ntection of pristine MFM data while simultaneously re-\nvealing the contact potential di\u000berence \u0001 Ucpd[29{31].\nIn turn, this re\rects the shift of the electric potential\ninduced by the magnetoelectric coupling.\nMFM, KPFM and non-contact atomic force mi-\ncroscopy (nc-AFM) were performed in an Omicron cryo-\ngenic ultra-high vacuum STM/AFM instrument [32] us-\ning the RHK R9s electronics [33] for scanning and\ndata acquisition. For all measurements, we used PPP-\nQMFMR probes from Nanosensors [34] driven at me-\nchanical peak oscillation amplitudes of A\u001910 nm.\nMFM images were recorded in a two-step process.\nFirstly, the topography and the contact potential di\u000ber-\nence of the sample were recorded and the topographic\n2D slope was canceled. Secondly, the MFM tip was re-\ntracted 20 nm o\u000b the sample surface to record magnetic\nforces while scanning a plane above the sample surface.\nThe KPFM-controller was switched o\u000b during this second\nstep. After the \frst MFM image had been completed at\n\u0006100 mT, the magnetic \feld was changed automatically\nin constant steps of 4 mT in between consecutive images.\nIn order to ensure a correct compensation bias, we ap-\nproached the tip to the sample before every \feld step\nand switched the KPFM-controller on. Note that the\nKPFM values change for every new magnetic \feld incre-\nment. After the new \feld had been reached, we hold the\nKPFM-controller again constant and retracted the tip\nby the same lift height. After the series of images had\nbeen completed a background image in the \feld-polarized\nstate atj\u00160Hj= 250 mT was taken.\nIII. THEORY\nThe energy density for the magnetization ~Mof cubic\nchiral magnets in the limit of small spin-orbit coupling\n\u0015SOCis given byE=E0+Edip+Eaniso where\nE0=J\n2(@i~M)2+\u001bD~M(r\u0002~M)\u0000\u00160~H~M (1)\ncomprises the isotropic exchange interaction J > 0, the\nDzyaloshinskiii-Moriya interaction D > 0 and the Zee-\nman energy. Depending on the chirality of the system,\nthe sign\u001b=\u00061. The competition between the \frst\ntwo terms results in spatially modulated magnetic or-\nder with a typical wavevector given by Q=D=J. The\nsecond termEdipcontains the dipolar interaction, and\nthe last termEaniso represents the magnetocrystalline\nanisotropies that are e\u000bectively small in the limit of small\n\u0015SOC. Under certain conditions, the magnetic helix min-\nimizes the energy density Ewhere its orientation is de-\ntermined by both the Zeemann energy and the magne-\ntocrystalline anisotropies Eaniso. In general, this leads to\na helix reorientation as a function of applied magnetic\n\feld.\nAn e\u000bective theory for this helix reorientation was pre-\nsented in Ref. [12] for MnSi. In section III A and III B\n(a) tetrahedral point group T\n(b) field cooling (c) zero field cooling\n(d) (e)\n[111][100][010][001]Figure 1. (a) B20 materials possess the symmetry of the\ntetrahedral point group indicated by the di\u000berent colours on\nthe sphere. (b) and (c): trajectories for the unit vector ^Q\nfor di\u000berent \feld directions for decreasing (\feld cooling) and\nincreasing (zero \feld cooling) \feld magnitude, respectively,\nevaluated for \"(2)\nT= 0 in Eq. (2). Starting from equally pop-\nulatedh100idomains at zero \feld in (c), trajectories can be\ncontinuous or discontinuous, and can possess a kink. (d) and\n(e): The kink represents a second-order transition that occur\nfor magnetic \feld directions indicated by the blue solid lines.\nFor a \fnite \"(2)\nTthese lines are warped towards h100i.\nwe review this theory for completeness and discuss its\nvalidity for Cu 2OSeO 3. In section III C we focus on the\ntheoretical predictions for the experimental setup. In sec-\ntion III D we present a theory for the electric polarization\nin Cu 2OSeO 3and its dependence on magnetic \feld.4\nA. E\u000bective Landau potential for the helix axis\nThe helix wavevector ~Qis determined by the compe-\ntition of Dzyaloshinskii-Moriya and exchange interaction\nand, as a consequence, its magnitude is proportional to\nspin-orbit coupling j~Qj\u0018O (\u0015SOC). The orientation of\nthe magnetic helix in a certain domain is represented\nin the following by the unit vector ^Q. The compe-\ntition between magnetocrystalline anisotropies and the\nZeeman energy can be captured in the limit of small spin-\norbit coupling \u0015SOC by the Landau potential V(^Q) =\nVT(^Q) +VH(^Q) [12].\nThe \frst term represents the magnetocrystalline po-\ntential and its form is determined by the tetrahedral\npoint group T[see Fig. 1(a)] of B20 materials like MnSi\nor Cu 2OSeO 3. As indicated by the di\u000berently colored\nregions on the sphere, Texhibits a three-fold C3rota-\ntion symmetry around the h111idirections, but only a\ntwo-foldC2rotation symmetry around h100ipoints. The\npotentialVT(^Q) contains all terms consistent with Tand\nreads as\nVT(^Q) =\"(1)\nT(^Q4\nx+^Q4\ny+^Q4\nz) (2)\n+\"(2)\nT\u0000^Q2\nx^Q4\ny+^Q2\ny^Q4\nz+^Q2\nz^Q4\nx\u0001\n::::\nThe lowest order term with constant \"(1)\nT\u0018O(\u00154\nSOC) is of\nfourth order in ^Q, and it still possesses a four-fold rota-\ntion symmetry around the h100iaxes that is not present\nin the tetrahedral point group T. This symmetry is bro-\nken explicitly only at the sixth order in ^Qby the term pa-\nrameterized by \"(2)\nT\u0018O(\u00156\nSOC). Other terms of sixth and\nhigher orders are represented by the dots. In the limit of\nsmall spin-orbit coupling, the \frst term determines the\norientation of the helix at zero \feld. For \"(1)\nT>0 the\npotential is minimized for ^Qk h111ias it is the case\nfor MnSi whereas \"(1)\nT<0 favours a helix orientation\n^Qkh100ilike in Cu 2OSeO 3.\nThe second term in the Landau potential represents\nthe Zeeman energy and up to second order in the applied\nmagnetic \feld ~Hit reads\nVH(^Q) =\u0000\u00160\n2Hi\u001fijHj; (3)\nwhere\u00160is the magnetic constant. The inverse of the\nmagnetic susceptibility tensor evaluated at zero \feld is\ngiven by\n\u001f\u00001\nij=\u001f\u00001\nij;int+Nij; (4)\nwith the demagnetization tensor Nijthat is diagonal for\nan elliptical sample shape N= diagfNx;Ny;Nzgwith\ntrfNg= 1. The internal susceptibility tensor \u001fintis\nevaluated for a \fxed orientation of the helimagnetic order\nand depends on ^Q,\n\u001fij;int=\u001fint\nk^Qi^Qj+\u001fint\n?(\u000eij\u0000^Qi^Qj): (5)An explicit calculation yields \u001fint\nk= 2\u001fint\n?, i.e., only half\nof the spins respond to a transversal magnetic \feld com-\npared to a \feld applied longitudinal to ^Q.\nMinimization of the Landau potential V(^Q) =VT(^Q)+\nVH(^Q) yields the helix orientation as a function of mag-\nnetic \feld ^Q=^Q(~H). The resulting trajectories were\ndiscussed in detail for \"(1)\nT>0 in Ref. [12]. Here we fo-\ncus on\"(1)\nT<0. Next, we present a general discussion of\nthe helix reorientation before turning to the con\fguration\nof the current experimental setup.\nB. Helix reorientation transitions\nDepending on the direction of the applied magnetic\n\feld, the reorientation of the magnetic helix might in-\nvolve either a crossover, a second-order phase transition\nor a \frst-order phase transition. For the purpose of a sim-\npli\fed discussion in this section we consider a sphere-like\nsample shape with demagnetization factors Ni= 1=3.\nFirst, we will focus on the potential without sixth-order\nterms and discuss corrections due to a \fnite \"(2)\nTat the\nend of this section. \"(1)\nTin Cu 2OSeO 3is negative, i.e.,\nthe preferred directions in zero \feld are h100iindicated\nby the black colored points on the unit sphere in Fig. 1.\nFigure 1(b) presents trajectories of the helix axis ^Q(~H)\nfor di\u000berent \feld directions indicated by the coloured\ndots after \feld cooling. For high \felds ^Q(~H) =~H=j~Hj.\nWhen decreasing the \feld, the crystalline anisotropies\ngain in\ruence and the helix reorients towards h100i. De-\npending on the \feld direction, three scenarios can be\ndistinguished. The reorientation process is a crossover\nwhen the helix reorients smoothly towards the closest\nh100idirection like for the green trajectory in Fig. 1(b).\nIt involves a second-order Z2transition when the trajec-\ntory bifurcates into two at a certain critical \feld Hc1, like\nfor the blue and yellow trajectories. A special situation\narises for a magnetic \feld along h111i. Here, the trajec-\ntory can follow three paths towards one of three distinct\nh100idirections realizing a second-order Z3transition.\nThis transition is protected by the three-fold C3rotation\nsymmetry of the point group T, i.e., it is robust even in\nthe presence of a \fnite \"(2)\nT. In general, a Z3transition\ncan be \frst-order as cubic terms are allowed in the ef-\nfective Landau expansion. For the potential of Eq. (2)\nwith\"(2)\nT= 0, however, this transition turns out to be of\nsecond-order with continuous trajectories ^Q(~H).\nAfter zero \feld cooling, helimagnetic domains oriented\nalong the threeh100idirections are populated. Upon\nincreasing the magnetic \feld, the helix axis ^Qmoves to-\nwards the \feld direction [see Figure 1(c)]. In addition to\nthe reversed paths of panel (b) there exist also discontin-\nuous paths starting from domains unfavoured by the \feld\ndirection. This discontinuous reorientation correspond to\na \frst-order transition.\nThe reorientation process thus involves a second-order5\ntransition and thus a well-de\fned critical \feld Hc1only\nfor speci\fc directions of the magnetic \feld. For \"(2)\nT= 0\nthese directions are located on the great circles on the\nsphere,that connect the h111ipoints [see Fig. 1(d)]. A\n\fnite\"(2)\nTinduces a warping of these lines [see panel (e)].\nAs the ratio \"(2)\nT=\"(1)\nT\u0018\u00152\nSOCis of second order in spin-\norbit coupling this e\u000bect is expected to be small.\nIn Cu 2OSeO 3, the sixth order term only quantitatively\nin\ruences the reorientation transitions. This is di\u000berent\nto MnSi where it is crucial to take the \"(2)\nTterm into\naccount as discussed in Ref. [12]. There, \"(1)\nTis positive\nwhich yieldsh111ias preferred directions in zero \feld.\nFor a \feld along [100], four of those are equally close\nsuggesting a Z4transition. However, a \fnite \"(2)\nTsplits\nthisZ4transition into two subsequent Z2transitions.\nC. Helix orientation trajectory for the\nexperimental setup\nIn the following, we neglect the sixth-order correction\n\"(2)\nT= 0 as its in\ruence is weak and cannot be resolved\nwithin the experimental accuracy. The investigated sam-\nple [see section II] is approximately a plate so that we use\nNx=Ny= 0 andNz= 1 for the demagnetization fac-\ntors in the basis of principal axis. The normal axis of the\nplate-like sample approximately corresponds to a crys-\ntallographic [110] direction. Within the crystallographic\nbases the demagnetization tensor is then given by\nN=0\n@1=2 1=2 0\n1=2 1=2 0\n0 0 01\nA: (6)\nThe magnetic \feld is always approximately aligned along\nthe surface normal so that we restrict ourselves to the\nmagnetic \feld direction ^HT= (1;1;0)=p\n2. For the lon-\ngitudinal susceptibility of Cu 2OSeO 3we use the value\n\u001fint\nk= 1:76 given in Ref. [35]. The transition between\nthe conical and \feld-polarized phase occurs at the criti-\ncal \feld\u00160Hc2\u0019192 mT (see below), which we will use\nlater on to normalize the \feld axis.\nFor a \feld along [110] the reorientation process is ei-\nther \frst-order for the [001] domain (yellow), or second-\norder for the [100] and [010] domains (purple) [see\nFig. 2(a)]. The continuous trajectory of the helix axis\ncan be parametrized as ^QT= (cos\u001e;sin\u001e;0) with\u001e\nthe azimuthal angle that depends on the magnetic \feld,\n\u001e=\u001e(~H). Its dependence is shown in Fig. 2(b) where\nthe kink de\fnes the critical \feld, that for Eq. (6) is given\nby\n\u00160Hc1= 2(1 +\u001fint\nk)vuut\u00160\"(1)\nT\n\u001fint\nk: (7)\nAs we will see later, experimentally we \fnd \u00160Hc1\u0019\n(a) (b)\n(c) (d)\n(e)Figure 2. (a) Reorientation process for a magnetic \feld point-\ning along [110]. The helix axis ^Q= (cos\u001e;sin\u001e;0) continu-\nously moves as a function of Halong the purple paths be-\ntween [110] and either [100] or [010]. The (yellow) domain\n[001] depopulates in a discontinuous fashion upon increasing\nH. (b) Azimuthal angle \u001e(H) starting from the [100] domain\natH= 0; (c) uniform magnetization MH(H) and (d) suscep-\ntibility@MH=dH for the purple trajectories in (a). (e) Electri-\ncal polarization \u0016Pz(H) attributed to the domains located for\nH= 0 at [100] and [010] (purple) and [001] (yellow). The crit-\nical \feldHc1separates the conical state C (grey) where ^Qk~H\nand the generalized helical state H (green) where ^Q,~H.\n70 mT corresponding to a value \"(1)\nT=\u00000:0014\u0016eV/\u0017A3\nfor the magnetocrystalline potential.\nIn Fig. 3 we illustrate the Landau potential V(^Q) for\nthe above parameters for various values of the magnetic\n\feldH. At zero \feld, all h100idirections are ener-\ngetically degenerate, see panel (a). For a \fnite \feld\nalong [110], the direction [001] remains a local mini-\nmum until it disappears at a certain spinodal \feld Hsp\nofHsp=Hc2\u00190:25. At the same time, the other minima\nremain global minima and move towards the \feld direc-\ntion. They merge at the critical \feld Hc1=Hc2\u00190:36\nand a single global minimum is obtained for H\u0015Hc1.\nD. Electric polarization\nThe magnetoelectric coupling in Cu 2OSeO 3induces an\nelectric polarization that is given in terms of the magne-6\n(a) h=0 (b) h=0.1 (c) h=0.2\n(d) h=hsp≈0.25 (e) h=hc1≈0.36 (f) h=hc2=1\nVmax(h)\nVmin(h)\nFigure 3. E\u000bective Landau-potential for the helix axis,\nV(^Q), for the experimental parameters. The magnetic \feld\nis applied along [110] and various values of the reduced \feld\nh=H=H c2are shown, see text. The color coding varies\nfrom panel to panel; red and blue color correspond to the\nminimal and maximal potential value at each speci\fc \feld,\nrespectively.\ntization vector ~Mby [15]\n~P(~ r) =cME*0\n@My(~ r)Mz(~ r)\nMz(~ r)Mx(~ r)\nMx(~ r)My(~ r)1\nA+\n; (8)\nwherecMEdenotes the magnetoeletric coupling constant.\nGenerally, the expectation value in Eq. (8) can be decom-\nposed into\nhMi(~ r)Mj(~ r)i=hMi(~ r)ihMj(~ r)i+Sij(~ r) (9)\nwithSij(~ r) the correlation function. In the mean-\feld\napproximation,Sijis neglected and the polarization re-\nduces to a product of expectation values h~M(~ r)i.\nWithin the framework of the Landau theory of section\nIII A this expectation value is given in terms of a Fourier\nseries,\nh~M(~ r)i=~MH+~Mhelix(~ r) +\u000e~M(~ r): (10)\nThe second term is given by a harmonic helix\n~Mhelix(~ r) =Mhelixh\n^e1cos(~Qmin~ r) +\u001b^e2sin(~Qmin~ r)i\n(11)\nwith the orthogonal unit vectors ^ e1\u0002^e2=^Qmin=\n~Qmin=Q. Depending on the chirality of the system, see\nEq. (1), the helix can be right-handed or left-handed cor-\nresponding to \u001b= +1 or\u00001, respectively. Here, the ori-\nentation of the helix axis ^Qmin(~H) minimizes the Landau\npotentialV(^Q) at a given ~H. The \frst term in Eq. (10)\nrepresents the uniform part, ~MH=MH^H, that can be\nobtained with the help of the Landau potential:\nMH=\u00001\n\u00160@V(^Qmin(~H))\n@H: (12)The magnitude MHas well as the total susceptibility\n@HMHare shown in Fig. 2(c) and (d) respectively. The\nsusceptibility shows a pronounced mean-\feld jump at the\ncritical \feld Hc1.\nIf variations of the amplitude are negligible, the length\nof the magnetization should be locally given by the satu-\nration magnetization h~M(~ r)i2=M2\nswhich gives rise to\nanharmonicities represented by \u000e~M(~ r) in Eq. (10). Min-\nimizing the energy (1) in the presence of this constraint,\nwe \fnd in lowest order, i.e., neglecting 3 ~QFourier com-\nponents that \u000e~Malso assumes the form of a helix,\n\u000e~M(~ r)\u0019j~MH;?jh\n^e0\n1cos(2~Qmin~ r) +\u001b^e0\n2sin(2~Qmin~ r)i\n:\n(13)\nThe prefactor is determined by the magnetization pro-\njected onto the plane perpendicular to ~Qmin, i.e.,~MH;?=\n~MH\u0000(~MH^Qmin)^Qmin. The unit vectors ^ e0\n1\u0002^e0\n2=^Qmin\nare given by\n^e0\n1= ^e2(^MH;?^e2)\u0000^e1(^MH;?^e1); (14)\n^e0\n2=\u0000^e2(^MH;?^e1)\u0000^e1(^MH;?^e2); (15)\nwhere ^MH;?=~MH;?=j~MH;?j. The component \u000e~Mis\nproportional to the uniform magnetization MHand thus\nvanishes linearly with the applied magnetic \feld. More-\nover, it vanishes for the conical state where ~MHk^Qmin\nso that~MH;?= 0. The anharmonicity is thus most pro-\nnounced at intermediate \felds as observed in MnSi and\nFeGe [4, 36, 37]. Within this approximation, the ampli-\ntude of the helix is given by\nMhelix=q\nM2s\u0000~M2\nH\u0000~M2\nH;?: (16)\nIn the experiment, the polarization ~P(~ r) cannot be\nspatially resolved on the scale of the helix wavelength.\nFor this reason, we consider the polarization spatially\naveraged over a single period,\n\u0016~P=1\n2\u00192\u0019Z\n0dx~P(~ r)\f\f\fMF\nx=~Qmin~ r(17)\nwhere the upper index MF indicates that we employ the\nmean-\feld approximation. In our experimental setup,\nit turns out that only the z-component \u0016Pzis expected\nto remain \fnite. For a (110) surface this z-component\namounts to an in-plane polarization along [001], \u0016P[001]=\n\u0016Pz.\nFor the continuous trajectories, i.e., the purple paths\nof Fig. 2(a), we get\n\u0016Pz=cME\n2\u0014\nM2\nH\u00001\n2sin(2\u001e)\u0000\nM2\ns\u0000M2\nH\u0001\u0015\n; (18)\nwith the azimuthal angle \u001e=\u001e(H) of Fig. 2(b). Its mag-\nnetic \feld dependence corresponds to the purple line in7\nFig. 2(e). At the \frst critical \feld, Hc1, the polarization\nis minimal and shows a kink. For Hc2> H > H c1, the\nangle\u001e=\u0019=4 and the uniform magnetization MH=\nMsH=Hc2so that the polarization reduces to the known\nexpression [15, 22, 38] \u0016Pz=cMEM2\ns\n4(3(H=Hc2)2\u00001), and\na sign change is expected for H=Hc2=p\n3. At the second\ncritical \feld Hc2another kink re\rects the phase transi-\ntion to the \feld-polarized phase. For H\u0015Hc2, we have\nMH=Msand \u0016Pz=\u0016Pc2\u0011cME\n2M2\ns.\nFor the yellow [001] domain in Fig. 2(a), the helix\naxis is given by ^QT\nmin= (0;0;1) for \feldsjHj\u0014Hsp\u0019\n0:25Hc2up to its spinodal point where the \frst-order\ntransition must take place at the latest. Its polarization\nwithin this \feld range is then given by\n\u0016Pz=cME\n2M2\nH; (19)\nthat is shown as a yellow line in Fig. 2(e).\nIV. EXPERIMENTAL RESULTS\nIn this section, we present the experimental \fndings\nobtained via MFM and KPFM measurements that are\nboth sensitive to signals attributed to the surface of the\nmaterial. In the present setup an approximate (110)\nsurface is considered so that the surface normal ^ nT=\n(1;1;0)=p\n2. Moreover, the applied \feld is approximately\nparallel to ^n. It is important to note that helimagnetic\norder with orientation ^Qand an intrinsic wavelength\n\u0015h= 2\u0019=Q\u001960 nm [7] gives rise to periodic magnetic\nstructures appearing at the sample surface characterized\nby a projected wavevector\n~Q0=~Q\u0000^n(~Q^n) =Q(^Q\u0000^ncos \u0002); (20)\nwhere \u00022[0;\u0019=2] is the angle between the helix axis\n^Qand the surface normal ^ n. This results in a projected\nwavelength \u00150=2\u0019\nj~Q0jgiven by [25]\n\u00150=\u0015h\nsin \u0002: (21)\nFor a helix with an in-plane ^Qthe angle \u0002 = \u0019=2 and\n\u00150=\u0015h. However, for a helix oriented along the surface\nnormal \u0002 = 0 the wavelength \u00150diverges, and the surface\nshould appear homogeneous.\nFurthermore, for the later analysis we introduce the\nin-plane angle \u000b=^(^e[001];~Q0) de\fned as the angle en-\nclosed by the projected wavevector ~Q0and the in-plane\nvector ^eT\n[001]= (0;0;1).\nA. Magnetic imaging of helimagnetic order\nWe start the presentation of our experimental results\nwith typical real-space images depicted in Fig. 4. Aslideshow of the complete dataset as well as of additional\nmeasurements can be retrieved from the supplementary\nmaterials. As MFM essentially tracks the out-of-plane\ncomponent of the local magnetization, the images repre-\nsent the projection of the magnetization onto the surface\nnormal~M(~ r)^n.\nThe image series in Fig. 4 is measured on the down-\nwards branch of the hysteresis loop sweeping the exter-\nnal magnetic \feld from positive to negative values. After\napplying a saturating \feld of \u00160H= 250 mT, the \feld\nwas decreased and the series starts with \u00160H= 100 mT\nshown in Fig. 4(a) where a periodic pattern is visible.\nDecreasing the \feld further, the magnetization on the\nsurface is reconstructed at about \u00160H\u001960 mT and mul-\ntiple helical domains form and increase in size preferen-\ntially showing a stripy pattern along the [ \u0016110] direction\n[see panel (c) and (d)]. Close to zero \feld, an additional\nhelimagnetic domain oriented along the [001]-direction is\nobserved giving rise to domain walls as shown in panel\n(e). The surface wavelength \u00150associated with the var-\nious domains depends on the applied magnetic \feld in\na characteristic manner. When decreasing the \feld fur-\nther to negative values, the domains start to split and\nmagnetization reconstructs at about \u00160H\u0019 \u0000 70 mT\n[see Fig. 4(g) and (h)]. At \u00160H\u0019 \u0000 100 mT a peri-\nodic modulation oriented along [001] is again visible in\nFig. 4(i) similar to panel (a). Finally, at large \feld of\n\u00160H=\u0000250 mT, the magnetization is fully polarized\nand the corresponding image in panel (k) is featureless.\nWe observed a manifold of co-existing domains as\nshown in Fig. 4(e) after \feld cycling. In contrast, af-\nter zero \feld cooling to T= 10 K only one single domain\nwith an in-plane helix axis along [001] could be observed,\nsimilarly to our previous measurements close to the crit-\nical temperature Tc[22].\nB. Analysis of the MFM data\nAssuming that the periodic patterns observed in the\nMFM images correspond to helimagnetic ordering pro-\njected onto the sample surface, we extract the projected\nwavevector ~Q0, the projected wavelength \u00150=2\u0019\nj~Q0j, and\nthe corresponding angles \u0002 and \u000bas de\fned at the begin-\nning of section IV. Exempli\fed in Fig. 5(a), we can exper-\nimentally distinguish three types of domains depending\non the in-plane angle \u000bat zero \feld, namely type I with\n\u000b\u00180°, type II with \u000b.90°, and type III with \u000b&90°.\nIt was previously established by neutron scattering\nthat the helices at zero \feld point along a crystallo-\ngraphich100idirection [7]. Correspondingly, one expects\nindeed three di\u000berent domains for a (110) surface with\nin-plane angle \u000b= 0 for ^Qk[001] and\u000b= 90\u000efor^Q\nalong [100] and [010]. The deviations from these values\nin Fig. 6(c) indicate an uncertainty of about 4\u000edue to\na combination of systematic errors. First, the sample\nis slightly miscut so that the surface normal might be8\na +100 mT\n b +60 mT\n c +52 mT\n d +44 mT\n e 0 mT\nf -60 mT\nneg. ∆f pos.\ng -64 mT\n h -76 mT\n i -100 mT\n k -250 mT\n[001][1¯10]\n⊙[110]\n2 µm\nFigure 4. Typical scanning probe images of Cu 2OSeO 3atT= 10 K. The magnetic \feld was applied perpendicular to the\nplane of projection along the [110]-direction changing from positive to negative values. Periodic magnetic textures are observed\nwhose wavelength and orientation depend on the strength of the magnetic \feld (see text). The span of the color scale is adapted\nindividually, ranging from 1 Hz to 2.5 Hz.\nb +24 mT\n#1\n#2\n0 1.1 Hz 400 nm\na +0 mT\nα[001]\n[1¯10]/vectorQ/primeI\nIII\nII\n0 1.4 Hz 1 µm\nFigure 5. (a) Close to \feld zero three types of domains,\nI, II, and III, can be observed. The projected wavevector\n~Q0and the [001] direction enclose the in-plane angle \u000b. (b)\nRelaxation events during the scanning process give rise to\ndiscontinuity lines (red arrows) within the periodic pattern\nwith phase jumps of 180\u000e.\nslightly tilted away from [110] towards [111], while sec-\nond, the magnetic \feld might be slightly misaligned from\nthe surface normal ^ n. Third, the sample placement in the\nMFM can be slightly misaligned with a small in-plane\nrotation as well. Finally, dynamic creep of the scanning\npiezo actuator slightly a\u000bects the scanner calibration.\nThe evolution of the projected wavelength \u00150and the\ncorresponding angle \u0002 of these three types of domains is\nshown as a function of magnetic \feld for every domain\nin Fig. 6(a) and (b), respectively. This includes also do-\nmains, which where not in the image frame at zero \feldand therefore are not classi\fed as one of the three types\nin Fig. 5(a) (shown as grey dots). For the helimagnetic\ndomain oriented in-plane at zero \feld, \u00150=\u0015h(yellow\ndots, green dots after zero-\feld cooling). The other do-\nmains (blue and red dots) are characterized for a (110)\nsurface by an angle \u0002 = \u0019=4 and a projected wavelength\nof\u00150=p\n2\u0015h. A drastic change of \u00150is observed around\n70 mT that we identify with the critical \feld Hc1of the\nreorientation transition. For larger \felds, the projected\nwavelength is of order \u00150\u001810\u0015h, corresponding to an\nangle \u0002\u00185\u000e. We attribute this \fnite angle to the mis-\nalignment error mentioned above.\nC. Relaxation processes during helix reorientation\nWhenever changing the magnetic \feld, the magnetic\nstructure relaxes on relatively long time scales, especially\nclose toHc1as discussed in Ref. [12]. An example of such\na relaxation process is shown in Fig. 5(b). The MFM\nimage is scanned from top to bottom. During this scan\nthe magnetic structure might change due to relaxation\nevents. They are re\rected in discontinuity lines marked\nwith (#1) and (#2) in Fig. 5(b), where the helix pattern\nis shifted by 180\u000e. Such 180\u000eshifts were observed before\nin Ref. [13] and attributed to the motion of dislocation\ndefects in the helimagnetic background. Interestingly,\nthe discontinuity lines do not continue through the full\nimage frame but terminate. Probably, the termination\npoints coincide with a helimagnetic domain wall separat-\ning di\u000berenth100idomains [14]. This suggests that the\ndiscontinuity lines arise from motion of dislocations close9\nλ' (nm)nm))\n608010050010001500H C C\na\nsgn(nm)ΔHH)∙µ0H (nm)m)T)050100150200 -200-150-100-5095\n90\n85\n5\n0\n-5α (°)c70\n50\n30\n6\n2\n0θ (°)b90\n4\nFigure 6. Experimentally determined helix orientation char-\nacterized by (a) the projected wavelength \u00150of Eq. (21), (b)\nout-of-plane angle \u0002 = ^(^n;^Q) and (c) measured in-plane\nangle\u000b. Yellow and green dots denote [001]-domains (type\nI) observed during \feld sweeps or after zero-\feld cooling, re-\nspectively. Blue and red dots denote type II domains and\ntype III domains, respectively. Grey dots belong to domains,\nwhich where not in the image frame at zero \feld.\nto the domain wall. The creep motion of dislocations\ncontributes to the complex and slow relaxation processes,\ngiving rise to hysteretic e\u000bects even for the second-order\nphase transition at Hc1.\nD. Contact potential and polarization\nCompensating electrostatic forces at the MFM tip with\nthe help of KPFM allows to extract di\u000berences in the\ncontact potential \u0001 Ucpd. On a highly insulating sam-\nple as is Cu 2OSeO 3, this potential is measured on length\nscales of the mesoscopic MFM-tip so that it corresponds\nHC C(a)\nHC C(b)Figure 7. (a) Contact potential di\u000berence \u0001 Ucpdas a func-\ntion of increasing (yellow, orange, and red points) and de-\ncreasing (green, blue and purple points) magnetic \feld at\nT\u001910 K. (b) Same data is plotted so that all histories run\nfrom the left-hand to the right-hand side.\nto an average over entire domains. As explained in detail\nin Ref. [22], for the current experimental setup this po-\ntential \u0001Ucpdfor a single domain is proportional to the\nin-plane polarization \u0016Pzof Eqs. (18) or (19). The mea-\nsured \u0001Ucpdas a function of magnetic \feld is displayed\nin Fig. 7(a) where the background colours indicate the\nvarious phases previously identi\fed with MFM.\nSimilar to our previous measurement [22] performed\nclose to the critical temperature Tc, we \fnd a plateau-\nlike region close to the zero \feld, a minimum at the crit-\nical \feldHc1, an increase of \u0001 Ucpdwithin the conical\nphaseHc1<\n>:1 Mkh100i\n1=2Mkh110i\n1=3Mkh111i:(8)\nOne may, for example, estimate the anisotropy constant\nby considering the di\u000berences of the magnetic work be-\ntween di\u000berent orientations:47\nWh110i\u0000Wh100i=\u00001\n2K (9)\nWh111i\u0000Wh100i=\u00002\n3K: (10)\nThese di\u000berences are shown in Fig. 10(b). While the\ncurves are guides to the eye, it is interesting to note the\nquantitative consistency of the di\u000berent directions with\nEqs. 9 and 10. The resulting anisotropy constant Kas a\nfunction of temperature derived from these di\u000berences is\nshown by the blue and green curves in Fig.10 (d).\nIn fact, the critical \feld Hint\nc2also displays a directional\ndependence in the presence of the anisotropy Kas re-\nported in Ref. 49. The di\u000berence of critical \felds along\nh110iandh111iis proportional to Kand may be ex-\npressed as\nHint\nc2;h110i\u0000Hint\nc2;h111i=\n\u00005\n3K\n\u00160Ms\u0014\n1 +O\u0012K\u001b\n\u00160Hc2Ms\u0013\u0015; (11)\nwhere the correction, O(:::), may be neglected for small\nK. This allows to determine the value of Kwith the\nhelp of the upper critical \felds Hc2for the three crystal-\nlographic directions h100i,h110i, andh111ias shown in\nFigs. 10(c) and 10(d). At high temperatures, just below\nTcessentially no anisotropy may be observed. However,\nwith decreasing temperature, the values of Hc2for di\u000ber-\nent directions separate below \u001840 K, becoming distinctly\ndi\u000berent.\nValues of the anisotropy constant Kinferred from Hc2\nare shown as orange curve in Fig. 10(d). With decreasing\ntemperature Kdeviates from zero and decreases below\n\u001835 K. Within the accuracy of the estimates carried out\nhere it is not possible to rule out a change of sign andsmall positive value of KbetweenTcand\u001835 K. This\nmight hint at the existence of further, weaker anisotropy\nterms. The putative existence of such terms may be\nclari\fed with the help of the anisotropy of Hc1. How-\never, it is important to emphasize that despite a possible\nchange of sign of Kas a function of temperature the\neasy magnetic axes remain the h100iaxes throughout.\nIn the limit of low temperatures, the values of Kas cal-\nculated from magnetization and the upper critical \felds\nare, \fnally, quantitatively consistent with the anisotropy\nconstant determined at T= 5 K with the help of ferro-\nmagnetic resonance measurements, KFMR =\u00000:6\u0002103\nJ/m3\u0019\u00003:7 neV/ \u0017A3(see supplement of Ref. 38 for de-\ntails).\nAlso shown in Fig. 10(d) are the anisotropy con-\nstants in dimensionless units, K=(\u00160MsHint\nc2;h111i)\u0019\nK=(\u00160MsHint\nc2jK=0) (ordinate on the right-hand side),\nwhereMsandHint\nc2are temperature dependent. As\nreported in Ref. 16 the cubic anisotropy may be ex-\npected to stabilize the LTS as a metatstable state with-\nout need for further exchange anisotropies when the ratio\nK=(\u00160MsHint\nc2;h111i) falls below a threshold of \u00000:07. This\ncondition is indicated by gray shading in Fig. 12(d). In-\ndeed, below\u001815 K, where the LTS phase is observed\nforHkh100i, the ratioK=(\u00160MsHint\nc2;h111i) reaches this\nstrength suggesting that the cubic anisotropy Krepre-\nsents the main stabilization mechanism for the additional\nphases.\nC. E\u000bective energy landscape\nFor weak magnetic anisotropies it is very well estab-\nlished experimentally and theoretically, that the direction\nof the conical helix aligns with the magnetic \feld. While\nthis alignment is on the expense of magnetic anisotropy\nenergy, it allows to minimize Zeeman energy as the mag-\nnetic moments cant most e\u000bectively towards the \feld di-\nrection. In turn, this raises the question why a su\u000eciently\nstrong anisotropy stabilizes a skyrmion phase and a tilted\nconical phase in which the moments twist around propa-\ngation directions that are not aligned with the magnetic\n\feld. In the following, we present a qualitative discus-\nsion in order to provide an intuitive picture of the com-\npetition between Zeeman energy and magnetocrystalline\nanisotropy for modulated structures. A quantitative nu-\nmerical analysis has been reported in Ref. 16.\nFor what follows, it is helpful to consider the magneti-\nzation of a right-handed conical modulation,\nM(r)=Ms= cos\u000b^e3\n+ sin\u000b(^e1cos(k^e3r) + ^e2sin(k^e3r));\nwhereMsis the saturated magnetization, \u000bis the cone\nangle with cos \u000b\u0019H=H c2andkis the magnitude of the\npitch vector that is oriented along the unit vector ^k= ^e3\nwith the orthonormal right-handed basis ^ e1\u0002^e2= ^e3.18\nThe uniform magnetization component of the conical he-\nlix points into the same direction than the pitch vec-\ntor, namely ^ e3. Using the pristine conical helix as the\nstarting point of a variational calculation in the magne-\ntocrystalline potential of Eq. (6), an e\u000bective anisotropy\npotential for the orientation of the pitch vector ^kmay be\nobtained given by\nVa(^k) =Ke\u000b(\u000b)\u0010\n^k4\nx+^k4\ny+^k4\nz\u0011\n: (12)\nWe note that this approximation neglects distortions of\nthe conical helix, which are of the order of a few % as\nobserved in the SANS study (see supplement of Ref. 16;\nsee also discussion in Ref. 43). The e\u000bective anisotropy,\nKe\u000b, depends on the cone angle \u000bwhere\nKe\u000b(\u000b) =K\n64(9 + 20 cos(2 \u000b) + 35 cos(4 \u000b)): (13)\nAs a key observation, Ke\u000bchanges sign as a function of\n\u000b. Whereas the sign of KandKe\u000bare the same for small\n\u000band cone angles close to \u0019=2, they possess opposite sign\nat intermediate cone angles, 30\u000e.\u000b.70\u000e.\nThe e\u000bective anisotropy potential is illustrated in\nFig. 11. Large energies are shown in red shading, whereas\nlow energies are shown in blue shading. The \frst three\npanels represent a cubic potential (a) with K= 0, (b)\nwithK < 0 favoring the cubic h100iaxes, and (c) with\nK > 0 favouring the cubic h111iaxes. The panels in the\nnext three rows, (d){(l), indicate the energy landscape\nof the magnetocrystalline potential, Eq. (6) with K < 0,\nthat is scanned by the conical helix state of Eq. (12) for\ndi\u000berent cone angles \u000band di\u000berent orientations of the\npitch vector ^k. Finally, the last three panels represent\nthe e\u000bective anisotropy potential for the pitch orienta-\ntion ^k, i.e, the function Va(^k) for three di\u000berent cone\nangles. Whereas Ke\u000b(\u000b)<0 for\u000b= 10\u000ein panel (m)\nand for\u000b= 80\u000ein panel (o), it is positive Ke\u000b(\u000b)>0\nfor\u000b= 55\u000ein panel (n).\nWhen the magnetic \feld is applied along a h100idi-\nrection, both, the Zeeman energy as well as the e\u000bective\nanisotropy energy are minimized for Ke\u000b(\u000b)<0. How-\never, they compete for intermediate cone angles where\nKe\u000b(\u000b)>0. Whereas the Zeeman energy favors to\nalign the helix with the \feld, the e\u000bective anisotropy\nsupports a pitch vector that is oriented away from the\n\feld, since, otherwise, the magnetic moments of the he-\nlical state would dominantly point into hard h111iaxes\nresulting in a large energy penalty, see Fig. 11(e). If the\ncubic anisotropy exceeds a threshold value K 0, respectively. For K < 0 the magnetic easy axes\nareh100i, whereas they are the hard axes for K > 0. (d){(l)\nillustrate the orientations that are covered by the magnetic\nmoments of the conical helix with a certain cone angle \u000bfor\ndi\u000berent \feld orientations of the pitch vector k. (m){(o) dis-\nplay the e\u000bective anisotropy potential of Eq. (13) that changes\nsign as function of cone angle \u000b.\nHintversusMfor three crystallographic directions.\nThe data shown here were measured at a temperature\nof 2 K following ZFC with \feld along h111i(blue),\nh110i(green), andh100i(red). As pointed out \frst in\nSec. C.1 the magnetization displays two intersections\nof theh111iandh110idirections. The magnetic work\nW^H(M) =\u00160RM\n0dM0H(M0) for theh111iandh110i\ndirection as inferred from the magnetization data is\nshown in Fig. 12(b) relative to the magnetic work for the\nh100iorientation. It re\rects the behavior of the e\u000bective\nanisotropy shown in Figs. 11(m) through 11(o).\nClose to saturation as well as for small values of M\nbothh111iandh110iare energetically larger than h100i\nidentifying the latter as the easy axis. For interme-\ndiate magnetization, however, the energy of the h111i\nandh110iorientations drops below the h100iorientation,\nequivalent to a change of the e\u000bective anisotropy, i.e., a\nsign change of Ke\u000b. The magnetization interval where\nthe inversion occurs indeed corresponds to the magneti-\nzation range in which the tilted conical phase is observed.19\n02550751000Hint (mT)\n(a)111\n110\n100\n0.00 0.25 0.50 0.75 1.00 1.25\nM (B / f.u.)\n2\n1\n012W (neV/Å3)\n(b)W111 W100\nW110 W100\nFIG. 12. Magnetization and magnetic work illustrating the\n\feld dependence of the e\u000bective anisotropy. (a) Internal mag-\nnetic \feld as a function of magnetization for di\u000berent direc-\ntions. (b) Magnetic work, inferred from magnetization data\nrelative to theh100iorientation as a function of the magne-\ntization.\nV. CONCLUSIONS\nIn conclusion, we reported a comprehensive study of\nthe magnetization and ac susceptibility of the magnetic\nphase diagram of Cu 2OSeO 3. For magnetic \feld paral-\nlel to theh100iaxis in the cubic crystal structure and\nlow temperatures we found clear evidence of the forma-\ntion of two new phases, a tilted conical state and a LTS\nstate identi\fed recently in a SANS study. The mag-\nnetization and susceptibility are thereby in remarkable\nagreement with the SANS data, providing clear ther-\nmodynamic signatures of these two new phases. Com-\nplementary selected speci\fc heat data support these re-\nsults. A detailed analysis of the strength of the magnetic\nanisotropy establishes that the conventional quartic con-\ntribution to the free energy by itself is su\u000ecient to sta-bilize the LTS phase. Detailed measurements exploring\nthe role of di\u000berent sample shapes shed new insights on\nthe role of demagnetizing \felds in the stabilization of\nthe tilted conical state. Taken together, we \fnd that the\nLTS phase in Cu 2OSeO 3represents a thermodynamically\nstable ground state driven by cubic magnetocrystalline\naisotropies, whereas the tilted conical state exists as a\nmetastable phase even for tiny demagnetization factors.\nIt is \fnally instructive to speculate on the more general\nimportance of our observations in Cu 2OSeO 3. As dis-\ncussed in our paper the magnetocrystalline anisotropies\nin the B20 compounds MnSi and Fe 1\u0000xCoxSi do not\nappear to be su\u000ecient to drive the formation of a\nLTS phase. In contrast, recent reports in Co 7Zn7Mn6\nidenti\fed a second skyrmion lattice phase at low\ntemperatures interpreted as a three-dimensional order\narising from the interplay of DM interactions with the\ne\u000bects of frustration.33Judging from the literature this\nmaterial exhibits also a pronounced magnetocrystalline\nanisotropy for the h100iaxes similar to Cu 2OSeO 3.\nIt is therefore tempting to speculate, whether the low\ntemperature skyrmion phase in Co 7Zn7Mn6originates,\nin fact, in the same mechanism we identify in Cu 2OSeO 3,\nhowever in the presence large amounts of disorder. This\nspeculation \fnds further support by the observation\nthat the e\u000bects of frustration, which must be present in\nCu2OSeO 3for the sake of the same analogy, do not seem\nto appear essential for stabilizing the LTS in Cu 2OSeO 3.\nACKNOWLEDGMENTS\nWe wish to thank S. Mayr for support. AC, MH,\nand WS acknowledge \fnancial support through the TUM\nGraduate School. This project has received funding\nfrom the European Research Council (ERC) under the\nEuropean Union's Horizon 2020 research and innova-\ntion programme (grant agreement No 788031). LH\nand AR acknowledge \fnancial support through DFG\nCRC1238 (project C02). MG acknowledges \fnancial sup-\nport through DFG CRC 1143 and DFG Grant GR1072/5.\nMH, AC, AB, and CP acknowledge support through\nDFG TRR80 (projects E1, F2 and F7) and ERC-AdG\n(291079 TOPFIT). MG, AR and CP acknowledge sup-\nport through DFG SPP2137 (Skyrmionics).\n1Naoto Nagaosa and Yoshinori Tokura, \\Topological prop-\nerties and dynamics of magnetic skyrmions,\" Nature Nano.\n8, 899{911 (2013).\n2S. M uhlbauer, B. Binz, F. Jonietz, C. P\reiderer, A. Rosch,\nA. Neubauer, R. Georgii, and P. B oni, \\Skyrmion Lattice\nin a Chiral Magnet,\" Science 323, 915{919 (2009).\n3W. M unzer, A. Neubauer, T. Adams, S. M uhlbauer,\nC. Franz, F. Jonietz, R. Georgii, P. B oni, B. Pedersen,M. Schmidt, A. Rosch, and C. P\reiderer, \\Skyrmion lat-\ntice in the doped semiconductor Fe 1\u0000xCoxSi,\" Phys. Rev.\nB81, 041203(R) (2010).\n4X. Z. Yu, N. Kanazawa, Y. Onose, K. Kimoto, W. Z.\nZhang, S. Ishiwata, Y. Matsui, and Y. Tokura, \\Near\nroom-temperature formation of a skyrmion crystal in thin-\n\flms of the helimagnet FeGe,\" Nature Materials 10, 106{\n109 (2011).20\n5N Kanazawa, Y Nii, X X Zhang, A S Mishchenko, G De Fil-\nippis, F Kagawa, Y Iwasa, Naoto Nagaosa, and Y Tokura,\n\\Critical phenomena of emergent magnetic monopoles in\na chiral magnet,\" Nature Commun. 7, 11622 (2016).\n6I. Kezsmarki, S. Bordacs, P. Milde, E. Neuber, L. M.\nEng, J. S. White, H. M. Ronnow, C. D. Dewhurst,\nM. Mochizuki, K. Yanai, H. Nakamura, D. Ehlers,\nV. Tsurkan, and A. Loidl, \\Neel-type skyrmion lattice\nwith con\fned orientation in the polar magnetic semicon-\nductor GaV 4S8,\" Nature Materials 14, 1116{1122 (2015).\n7S. Bordacs, A. Butykai, B. G. Szigeti, J. S. White, R. Cu-\nbitt, A. O. Leonov, S. Widmann, D. Ehlers, H. A. Krug von\nNidda, V. Tsurkan, A. Loidl, and I. Kezsmarki, \\Equilib-\nrium Skyrmion Lattice Ground State in a Polar Easy-plane\nMagnet,\" Sci. Rep. 7, 7584 (2017).\n8Xiuzhen Yu, Maxim Mostovoy, Yusuke Tokunaga, Weizhu\nZhang, Koji Kimoto, Yoshio Matsui, Yoshio Kaneko,\nNaoto Nagaosa, and Yoshinori Tokura, \\Magnetic stripes\nand skyrmions with helicity reversals,\" Proc. Nat. Acad.\nSci.109, 8856{8860 (2012).\n9S. Ishiwata, M. Tokunaga, Y. Kaneko, D. Okuyama,\nY. Tokunaga, S. Wakimoto, K. Kakurai, T. Arima,\nY. Taguchi, and Y. Tokura, \\Versatile helimagnetic phases\nunder magnetic \felds in cubic perovskite SrFeO 3,\" Phys.\nRev. B 84, 054427 (2011).\n10Ajaya K. Nayak, Vivek Kumar, Tianping Ma, Pe-\nter Werner, Eckhard Pippel, Roshnee Sahoo, Francoise\nDamay, Ulrich K. Roessler, Claudia Felser, and Stuart\nS. P. Parkin, \\Magnetic antiskyrmions above room temper-\nature in tetragonal Heusler materials,\" Nature 548, 561{\n566 (2017).\n11X. Yu, J.P. DeGrave, Y. Hara, T. Hara, S. Jin, and\nY. Tokura, \\Observation of the Magnetic Skyrmion Lat-\ntice in a MnSi Nanowire by Lorentz TEM,\" Nano Letters\n13, 3755 (2013).\n12S. A. Meynell, M. N. Wilson, K. L. Krycka, B. J. Kirby,\nH. Fritzsche, and T. L. Monchesky, \\Neutron study of\nin-plane skyrmions in mnsi thin \flms,\" Phys. Rev. B 96,\n054402 (2017).\n13A. Fert, N. Reyren, and V. Cros, \\Magnetic skyrmions:\nadvances in physics and potential applications,\" Nature\nReviews Materials 2, 17031 (2017).\n14Wanjun Jiang, Gong Chen, Kai Liu, Jiadong Zang, S.G.E.\nte Velthuis, and A. Ho\u000bmann, \\Skyrmions in magnetic\nmultilayers,\" Physics Reports 704, 1 { 49 (2017).\n15R. Wiesendanger, \\Nanoscale magnetic skyrmions in\nmetallic \flms and multilayers: a new twist for spintron-\nics,\" Nature Reviews Materials 1, 16044 (2016).\n16A. Chacon, L. Heinen, M. Halder, A. Bauer, W. Simeth,\nS. M uhlbauer, H. Berger, M. Garst, A. Rosch, and\nC. P\reiderer, \\Observation of two independent skyrmion\nphases in a chiral magnetic material,\" Nature Physics 14,\n936 (2018), advance online publication on 22 June 2018.\n17S. Seki, X. Z. Yu, S. Ishiwata, and Y. Tokura, \\Observa-\ntion of Skyrmions in a Multiferroic Material,\" Science 336,\n198{201 (2012).\n18T. Adams, A. Chacon, M. Wagner, A. Bauer, G. Brandl,\nB. Pedersen, H. Berger, P. Lemmens, and C. P\reiderer,\n\\Long-Wavelength Helimagnetic Order and Skyrmion Lat-\ntice Phase in Cu 2OSeO 3,\" Phys. Rev. Lett. 108, 237204\n(2012).\n19Stefan Buhrandt and Lars Fritz, \\Skyrmion lattice phase\nin three-dimensional chiral magnets from monte carlo sim-\nulations,\" Phys. Rev. B 88, 195137 (2013).20A. Bauer and C. P\reiderer, \\Topological Structures in Fer-\nroic Materials: Domain Walls, Vortices and Skyrmions,\"\n(Springer International Publishing, 2016) Chap. Generic\nAspects of Skyrmion Lattices in Chiral Magnets, p. 1.\n21L. D. Landau and E. M. Lifshitz, Course of theoretical\nphysics, vol. 8 (Pergamon Press, 1980).\n22U. K. R o\u0019ler, A. N. Bogdanov, and C. P\reiderer, \\Sponta-\nneous skyrmion ground states in magnetic metals,\" Nature\n442, 797 (2006).\n23H. Wilhelm, M. Baenitz, M. Schmidt, U. K. R o\u0019ler, A. A.\nLeonov, and A. N. Bogdanov, \\Precursor Phenomena at\nthe Magnetic Ordering of the Cubic Helimagnet FeGe,\"\nPhys. Rev. Lett. 107, 127203 (2011).\n24L. Cevey, H. Wilhelm, M. Schmidt, and R. Lortz,\n\\Thermodynamic investigations in the precursor region of\nFeGe,\" Phys. Status Solidi B 250, 650 (2013).\n25A. Bauer and C. P\reiderer, \\Magnetic phase diagram of\nMnSi inferred from magnetization and ac susceptibility,\"\nPhys. Rev. B 85, 214418 (2012).\n26H. C. Wu, T. Y. Wei, K. D. Chandrasekhar, T. Y. Chen,\nH. Berger, and H. D. Yang, \\Unexpected observation of\nsplitting of skyrmion phase in Zn doped Cu 2OSeO 3,\" Sci-\nenti\fc Reports 5(2015), 10.1038/srep13579.\n27(2018), a. Stefancic, S. Moody, T.J. Hicken, T.M. Birch,\nG. Balakrishnan, S.A. Barnett, M. Crisanti, J.S.O. Evans,\nS.J.R. Holt, K.J.A. Franke, P.D. Hatton, B.M. Hud-\ndart, M.R. Lees, F.L. Pratt, C.C. Tang, M.N. Wil-\nson, F. Xiao, T. Lancaster, Origin of skyrmion lattice\nphase splitting in Zn-substituted Cu 2OSeO 3, unpublished,\narXiv/1807.04641.\n28A. Bauer, M. Garst, and C. P\reiderer, \\History de-\npendence of the magnetic properties of single-crystal\nFe1\u0000xCoxSi,\" Phys. Rev. B 93, 235144 (2016).\n29Y. Tokunaga, X. Z. Yu, J. S. White, H. M. R\u001cnnow,\nD. Morikawa, Y. Taguchi, and Y. Tokura, \\A new class of\nchiral materials hosting magnetic skyrmions beyond room\ntemperature,\" Nature Commun. 6, 7638 (2015).\n30K. Karube, J. S. White, N. Reynolds, J. L. Gavilano,\nH. Oike, A. Kikkawa, F. Kagawa, Y. Tokunaga, H. M.\nRonnow, Y. Tokura, and Y. Taguchi, \\Robust metastable\nskyrmions and their triangular{square lattice structural\ntransition in a high-temperature chiral magnet,\" Nature\nMaterials 15, 1237{1242 (2016).\n31P. Milde, D. K ohler, J. Seidel, L. M. Eng, A. Bauer,\nA. Chacon, J. Kindervater, S. M uhlbauer, C. P\reiderer,\nS. Buhrandt, C. Sch utte, and A. Rosch, \\Unwinding of a\nSkyrmion Lattice by Magnetic Monopoles,\" Science 340,\n1076{1080 (2013).\n32S. P ollath, J. Wild, L. Heinen, T. N. G. Meier, M. Kro-\nnseder, L. Tutsch, A. Bauer, H. Berger, C. P\reiderer,\nJ. Zweck, A. Rosch, and C. H. Back, \\Dynamical defects\nin rotating magnetic skyrmion lattices,\" Phys. Rev. Lett.\n118, 207205 (2017).\n33K. Karube, J. S. White, D. Morikawa, C. D. Dewhurst,\nR. Cubitt, A. Kikkawa, X. Yu, Y. Tokunaga, T. Arima,\nH. M. Ronnow, Y. Tokura, and Y. Taguchi, \\Disordered\nskymrion lattice phase stabilized by magnetic frustration\nin a chiral magnet,\" Science Advances 4, eaar7043 (2018).\n34T. Schwarze, J. Waizner, M. Garst, A. Bauer,\nI. Stasinopoulos, H. Berger, C. P\reiderer, and\nD. Grundler, \\Universal helimagnon and skyrmion exci-\ntations in metallic, semiconducting and insulating chiral\nmagnets,\" Nature Materials 14, 478{483 (2015).\n35P. Milde, E. Neuber, A. Bauer, C. P\reiderer, H. Berger,21\nand L. Eng, \\Heuristic description of magnetoelectricity of\nCu2OSeO 3,\" Nano Lett. 16, 5612 (2016).\n36S. L. Zhang, A. Bauer, D. M. Burn, P. Milde, E. Neuber,\nL. M. Eng, H. Berger, C. P\reiderer, G. van der Laan,\nand T. Hesjedal, \\Multidomain Skyrmion Lattice State in\nCu2OSeO 3,\" Nano Lett. 16, 3285 (2016).\n37S. L. Zhang, A. Bauer, H. Berger, C. P\reiderer, G. van der\nLaan, and T. Hesjedal, \\Resonant elastic x-ray scattering\nfrom the skyrmion lattice in Cu 2OSeO 3,\" Phys. Rev. B 93,\n214420 (2016).\n38I. Stasinopoulos, S. Weichselbaumer, A. Bauer, J. Waizner,\nH. Berger, S. Maendl, M. Garst, C. P\reiderer, and\nD. Grundler, \\Low spin wave damping in the insulating\nchiral magnet Cu 2OSeO 3,\" Appl. Phys. Lett. 111(2017),\n10.1063/1.4995240.\n39J. A. Mydosh, Spin glasses: an experimental introduction\n(Taylor & Francis, 1993).\n40S. M uhlbauer, A. Heinemann, A. Wilhelm, L. Karge,\nA. Ostermann, I. Defendi, A. Schreyer, W. Petry, and\nR. Gilles, \\The new small-angle neutron scattering instru-\nment SANS-1 at MLZ|characterization and \frst results,\"\nNuclear Instruments and Methods in Physics Research Sec-\ntion A: Accelerators, Spectrometers, Detectors and Asso-\nciated Equipment 832, 297{305 (2016).\n41A. Aharoni, \\Demagnetizing factors for rectangular ferro-\nmagnetic prisms,\" J. Appl. Phys. 83, 3432 (1998).\n42M. I. Youssif, Bahgat A. A., and Ali I. A., \\AC Mag-\nnetic Susceptibility Technique for the Characterization of\nHigh Temperature Superconductors,\" Egypt. J. Sol. 23,231 (2000).\n43A. Bauer, A. Chacon, M. Wagner, M. Halder, R. Georgii,\nA. Rosch, C. P\reiderer, and M. Garst, \\Symmetry break-\ning, slow relaxation dynamics, and topological defects at\nthe \feld-induced helix reorientation in mnsi,\" Phys. Rev.\nB95, 024429 (2017).\n44M. Balanda, \\AAC Susceptibility Studies of Phase Tran-\nsitions and Magnetic Relaxation: Conventional, Molecular\nand Low-Dimensional Magnets,\" Acta Physics Polonica A\n124, 964 (2013).\n45J. Wild, T.N.G. Meier, S. Poellath, M. Kronseder,\nA. Bauer, A. Chacon, M. Halder, M. Schowalter, A. Rose-\nnauer, J. Zweck, J. M uller, A. Rosch, C. P\reiderer, and\nC. H. Back, \\Entropy-limited topological protection of\nskyrmions,\" Science Advances 3, e1701704 (2017).\n46(2018), F. Qian, L.J. Bannenberg, H. Wilhelm, G.\nChaboussant, L. M. Debeer-Schmitt, M. P. Schmidt, A.\nAqeel, T.T.M. Palstra, E. Br uck, A.J.E. Lefering, C. Pap-\npas, M. Mostovoy, and A. O. Leonov, arXiv/1802.02070v2,\nunpublished.\n47N. Akulov, Z. Phys. 100, 197 (1937).\n48J. H. Yang, Z. L. Li, X. Z. Lu, M. H. Whangbo, Su-\nHuai Wei, X. G. Gong, and H. J. Xiang, \\Strong\nDzyaloshinskii-Moriya Interaction and Origin of Ferro-\nelectricity in Cu 2OSeO 3,\" Phys. Rev. Lett. 109 (2012),\n10.1103/PhysRevLett.109.107203.\n49S. V. Grigoriev, A. S. Sukhanov, and S. V. Maleyev, \\From\nspiral to ferromagnetic structure in B20 compounds: Role\nof cubic anisotropy,\" Phys. Rev. B 91, 224429 (2015)." }, { "title": "2104.01333v2.Predicting_synthesizable_cobalt_and_manganese_silicides__germanide_with_desirable_magnetic_anisotropy_energy.pdf", "content": "1Predictingsynthesizablecobaltandmanganesesilicides,germanidewith\ndesirablemagneticanisotropyenergy\nZe-JinYang\nSchoolofScience,ZhejiangUniversityofTechnology,Hangzhou,310023,China\nzejinyang@zjut.edu.cn\nAbstract\nThenanoparticleCo3Si(P63/mmc)displaysremarkablemagnetism[Appl.Phys.\nLett.108,152406(2016)],wethussearchedcobaltsilicidesandseveralphasesare\nsearchedincludingaCmcmwith60meV/atomlowerthanthatofP63/mmc.A\nlower-energyComR3(-7.03eV/atom)ispredicted,whoseenergyishigherthan\nthatofknownP63/mmc(-7.04eV/atom)butislowerthanthatofmmF3(-7.02\neV/atom).Threesmall-magnetismlow-energyFe5Si3structuresaresearchedwith\nenergies30meV/atomlowerthanthatofexperimentalP63/mcm.Thestronglattice\nshapedependenceofmagnetocrystallineanisotropyenergy(MAE)isstudiedthrough\nX5Si3(X=Mn,Fe,Co).Thebuilding-blockshapeandenergyorderofcobaltsilicideis\ndominatedbyCoP63/mmc,m3mF,m3R,respectively.TheCo3CandCo3Snhave\npositiveformationofenergy,thusonlyCo3Gehassimilarstructureswiththoseof\ncounterpartsofCo3Si.Severallow-energyperfectornearly-perfecteasy-axis/plane\nMAEMn3Si,Mn5Si2,andMn5Si3structuresaresearchedandpresentimportant\napplication,asisalsothecaseinGe-containingcounterparts.AstructureI4122with\nenergy300meV/atomlowerthanthatofexperimentalMn5Si3P63/mcmissearched.2I.Introduction\nMagneticmaterialsplayanimportantroleinourlife,suchasharddiskinlaptop.\nMagnetismcomprisesofdiverseresearchfieldsinphysics,forinstancegiant\nmagnetoresistanceeffect,colossalmagnetoresistanceeffect,magnetocaloriceffect,\nmagneticexchangebiaseffect,magneticproximityeffect,spintronics,andspin\nSeebeckeffect.Recently,themostfrequentlyusedpermanentmagnetsare\nNd2Fe14B\randSmCo5-basedcompounds,bothcontainingrare-earthelement.Dueto\nlimitedrare-earthresources,thereforeitisnecessarytosynthesizematerialswithout\nrare-earthelement[1].Magnetocrystallineanisotropyenergy(MAE)playsakeyrole\ninthedevelopmentofmagneticstoragemedia.DifferentMAEmagnitudeshave\nextensiveapplicationsinthefieldsofsoftandhardmagneticmaterials.\nNanoscaleCo3Si(P63/mmc)issynthesized[2]anddisplayedexcellenteasy-plane\nMAEinitsnanoparticleform,basedonwhichtheauthorsconcludethatapath\ntowardsfabricationofnanoparticlematerialswitheasy-planeanisotropies,valuable\nenergyproducts,andinparticularcontainingearth-richsubstanceissuccessfullybuilt.\nInfact,metastablehexagonalCo3Si(P63/mmc)isobservedlongtimeago[3-5].\nHowever,thediscrepancybetweentheirtheoreticalcalculationfortheidealCo3Si\nnanoparticale(-64Merg/cm3)andexperimentalestimation(48Merg/cm3)anisotropy\nislarge,possiblyduetotheexperimentalsurplusofthe10vol.%Cointhe\nnanoparticlecomposite.Unexpectedly,theextraCofailstoformtheother\nstoichiometrycompoundssuchasCo4Si,etc.Consideringsuchimportantconclusion,\nitisverynecessarytofurtherstudythecobaltsilicidescarefully.Moreover,3nonequilibriumfabricationtechniqueusuallysynthesizesmetastablephaseswithvery\nlargeMAE.Therefore,weusetheadvancedcrystalstructurepredictionmethodto\nsearchthepossiblelow-energystructure.Moreover,theobservedCo4Sialsoisa\nmetastablephasebutitsstructureisstillunknown[4].Threedifferentphasesare\nobservedinCo2Sitillnow,itisα-Co2Si,β-Co2Si,γ-Co2Si,respectively[3-5].These\nphenomenasuggestthattherearemanymetastablephasesintheselowCo:Si\nstoichiometryratiocompounds.\nThesamegroup[2]alsoobservedthatnanoparticlesMn5Si3presentsappreciable\nmagnetocrystallineanisotropyconstants(K1=6.2Mergs/cm3at300Kandat12.8\nMergs/cm3at3K)[6].Mn-basedstructuresbecameanimportantstudyfocusfromthe\nviewpointsofdesigningrare-earth-freematerialswithusefulspin-related\napplications[7].Thus,wecarefullysearchedtheexperimentallyknownbutsimple\nMn-basedcompounds,suchasMn3Si[8],Mn5Si2[9],Mn5Si3[8],Mn6Si[10],Mn9Si2[11].\nWedidn’tsearchtheotherstoichiometrybecausethesearchandMAEcalculationare\nextremelytime-consuming.Consideringtheimportantstoichiometryof5:3,wealso\nsearchedtheFe5Si3astheprevioussearchmightmiss[1]someenergeticallysimilar\nstructures.Experimentalstructurehas53atomsperunitcell(Mn85.5Si14.5)[10]in\nMn6Si,butitis56atomsforthe8-unitcell,thesameauthorsalsofound186atoms\nfor(Mn81.5Si18.5)[11]inMn9Si2,whereasitis176atomsfor16-unitcell.Thusitis\ndifficulttocomparethesearchedperfectstructureswiththeirdata.Inthisstudy,we\nthoroughlysearchedthebinarymagneticcobaltandmanganesesilicidesaswellas\nFe5Si3inordertofindtheidealMAE,fortunately,oursearchedmanystructureshave4veryinterestingpropertiesinthediversemagneticfield.\nII.Methods\nWesearchedthestructuresbyCALYPSO(CrystalstructureAnaLYsisby\nParticleSwarmOptimization)code[12,13].Theobtainedstructuresareabinitio\noptimized[14,15]byprojector-augmentedwavemethod[16].Thespin\rpolarized\ngeneralizedgradientapproximationandthePerdew-Burke-Ernzerhoffunctionalare\nusedtomodeltheexchange-correlationenergy[17].Thestructureisoptimizedwith\ntheelectronicconvergencebelow10-7eV,thek-pointdensityisbelow2π×0.02Å−1in\ntheBrillouin-zone.Theforceontheatomisconvergedtolessthan0.01eV/Åduring\ntherelaxation.Energycutofffortheplan-wavebasisis350eV.Theseparametersare\nsufficienttoensurethereliabilityofthecalculatedresultsbasedonourcalculations\nonironsilicides[1].Theon-sitecoulombinteractiontermU(andJ)ontheenergyis\nneglectedduringallthestructurecomputation[14,15]basedontheprevious\nconclusionofironsilicides[1].\nIII.Resultsanddiscussion\nA.MAEofexperimentalnanoparticlesizeCo3Si\nWedefinethelowest-energyaxisasthereferenceaxisinthispaper.\nExperimentally[2]synthesizednanoscaleCo3Si(P63/mmc)withaneasy-plane\nanisotropywithahighK1=-64Merg/cm3(-6.4MJ/m3)andmagneticpolarizationof\n9.2kG(0.92T),whichisconfirmedbytheircalculated64-atomCo3Sinanoparticle.\nInaddition,theyalsofoundthatMn5Si3nanoparticlealsopresentsexcellent\nproperties[6].OurcalculatedCo3Sishowsa=b=4.9794Å,c=3.9755Å,substantially5differentwiththeexperimentaldataa=b=4.99Å,c=4.497Å.Ourcalculatedabsolute\ntotal-energydifferencebetweenspinorientationsinthe(100)and(001)directions\nyieldsaMAEof2.0891meVperunitcell(containing2formulaunitsor8atoms)for\nthebulkCo3Sicrystal,whichcorrespondsto|K1|=39.21Merg/cm3(or261.1412\nμeV/atom),farsmallerthantheexperimentallyreportedvalue-3.9meV(K1=-64\nMerg/cm3),whoseabsolutemagnitudeislargeandcomparabletothatoftypical\nrare-eartheasy-axisintermetallics,assaidinthatpaper.AnotherP63/mmc(Co3Si),\nwithanaverageenergyof-6.8931eV/atom,showsshortenedlatticeband\nslightly-changedaandc,asaresultwhichinducesanearly-perfecteasy-planeMAE,\nwithE001=80.28andE100=0.06μeV/atom(orenergiesalongcandaaxes),\nrespectivelyandsimilarlyhereinafter.Thus,theperfectbulkcrystalfailedtoobtain\nthecorrespondingmagnetismofnanoparticledueprobablytothesurfaceeffect.\nTorevealthepossibleoriginofthegiantaxialenergydifference,thenanoparticle\nisconstructedbyscreeningtheneighborinteractionthroughthelongvacuumlayer\nlargerthan10Å.TableS1showsthattheperfect8-atom,32-atom,64-atom\nnanoparticlecouldproduceMAEwithvaluesabout0.1,0.07,0.12meV/atom,\nrespectively,whichstillcouldn’trepeattheexperimentalMAE(-3.9meVperunitcell\nor-0.4875meV/atom).Wedidn’tcalculatethecaseof128-atomunitcellbecauseits\natomicarrangementistotallysamewiththecasesof8/32/64-atom.Onepossible\nexplanationisthattheexperimentallysynthesizedstructureexistslargedistortion\nalongthelatticeparametercorientation.DistortionusuallyinduceslargeMAEbutit\nisprobablymeaninglessinthepracticalindustrialapplication.Moreimportantly,6distortionusuallycouldn’tproduceperfectMAE.Inaddition,theirlatticeparameterc\nisalsocouldn’tbeexplainedfromtheunknownsurfacesubstanceorinternaldefect\nstructure.Thesurfaceinteractionisalmostimpossibletoelongatethelattice\nparametercwithsuchlargevalue,fromabout4.0to4.5Å,thusweignoretocalculate\ntheMAEatthecaseofc=4.5Åbecausesuchstructureishypotheticalbyuniaxial\nstrainandisalsounstable.\nTheexperimentalCo3SiP63/mmchasa~10%elongationoflatticecwhich\ninducesthegiantdistortionoflatticeshapeandbreaktheatomicequilibriumstatus.\nWetestitbytheX5Si3(X=Mn,Fe,Co),asisshowninTableS2,inwhichthestrong\nlatticeframeworkoratomicoccupancydependenceisobserved,theaverageenergy\ndifferencebetweeneasyandhardaxesofthethreelatticehasintrinsicvalue,suchas\nthestableCo5Si3hasanMAEvalueof~13μeV/atom,substitutionsofCobyMnand\nFeproduce~29(Co5Si3)and~23(Fe5Si3),respectively.ForthecaseofMn5Si3,those\ncorrespondingvaluesare~39and~39(Co5Si3)and~56(Fe5Si3).Forthecaseof\nFe5Si3,theyare~157and~88(Mn5Si3)and94(Co5Si3),respectively.\nB.Thesearchedlower-energystructuresthanthatofCo3SiP63/mmc\nWeonlysearchedthestoichiometryofCo:Si>1inordertofindthestructures\nwithlargemagnetism.Thecurrentsearchobtained5lowerenergystructuresthanthat\nofexperimentallysynthesizedCo3SiP63/mmc(-6.9eV/atom),inwhichCmcm\nstructureisabout60meV/atomlowerthanthatofP63/mmc.Theotherlower-energy\nphasesincludePmmn,(-6.9143),I4/mmm(-6.9197),I4/mcm(-6.9025),P21(-6.9071\neV/atom),asisshowninTable1.Infact,therearealsoseveralphaseshaveonly7slightlyhigherenergiesthanthatofP63/mmc.Inotherwords,themetastablephase\nwithenergyaboutseveraldozensofmeV/atomlowerthanthatofexperimentalphase\northesearchedlowest-energyphase(probablybe)stillcouldbesynthesized.\nThecurrentlycalculatedMAEofCo3SiP63/mmcis261.14μeV/atomwithan\neasy-axisenergydistribution,despitewhichisfarsmallerthanthatofpreviously\ncalculated487.5andmeasured365.625μeV/atombut~260isstillaverylargevalue.\nThustheMAEofCo3SiP63/mmc(261.14μeV/atom)isnotcomparablebutisfar\nsmallerthanthatoftypicalrare-eartheasy-axisintermetallicscompounds.\nCo3SiP63/mmc(a=b=4.9794,c=3.9754Å)isthesamelatticeframeworkwith\nthatofCoP63/mmc(a=b=2.507,c=4.069Å),thatis,the2×2×1supercellofCois\na=b=5.014Å,thusCo3SiandCohavetotallysamelatticewiththeonlyexceptionof\ntheslightdistortionduetothesubstitutionofCobySi.ProjectionsoftheCo3Si\nP63/mmctowardsaandbaxesaretotallysameowingtotheinversionsymmetry\noperationnature,thustheMAEalongaandbaxesareidentical.Moreover,the\nmixtureofnanoparticleCo3Siwithabout10vol.%ɛ-Comakestheresolutionofthis\nquestionverydifficult.ThebuildingblockofP63/mmcistetrahedronwiththe\ndistanceofCo~Si=2.4454ÅandCo~Co=2.4671Å.ThebuildingblockofCmcmisa\ncubicbodybuttheSidoesn’tdistributeatthecubiccenter,thuspresentingtwo\ndifferentkindsofdistance,itis2.4455and2.4446Åinitsfirstshellornearest\nneighbor,namely,aslightlydistortedbulkcenteredcubic(BCC)unit.\nCmcm-A(namedasA…becausemanystructureshavesamesymmetrybut\nshowobviouslydifferentenergies)isthesearchedlowest-energyphasewith8E010=0.24andE100=28.46μeV/atom,respectively.Itsbuildingblockshapeismost\nsimilarwiththatofP63/mmc.Indetail,thetwoneighboringlayer“Co3Si”hasa\ntorsionangle60゚inP63/mmcbutitiszerodegreeinCmcm-A.Moreover,thegeneral\nshapeofbuildingblockinCmcm-AismoresimilarwithaBCC,thedistanceof\nCo9~Co3=Co9~Co14=2.7097Å,Co11~Co14=Co11~Co3=2.547Å.Thedistancebetween\nthecentralSiandits8first-nearestneighborsareCo2,3~Si1=Co13,14~Si1=2.3227Å,\nCo5,9~Si1=2.3796Å,Co7,11~Si1=2.5162Å.ItsCo1~Si2=Si1~Co8=4.1856Å\ncorrespondstothe4.2992and4.3254ÅinP63/mmc.Thusthephasetransitionfrom\nP63/mmctoCmcm-AinCo3Siispossibleunderhighpressure.ThecentralSideviates\nthegeometricpositionevidentlyinCmcm-BrelativetothatofCmcm-A,forinstance,\nthefourpairsCo~Sidistancesare2.3837Åinfirstshellbuttheotherfouronesare\n2.482ÅinCmcm-B.Meanwhile,forthecaseofthesecondshellorsecondnearest\nneighbors,theCo~Sidistancesattheshortfirst-shellside(2.3837Å)isalsoshort\n(2.3372Å),viceversa,theotherside(2.482Å)correspondsto2.8817Å,theother\npairCo~Sidistanceswithinthesamehorizontalplaneshowsmallvariation,3.1959\nversus3.1963Å,onlythetwototally-symmetricCopresentidenticalCo~Sidistance,\nsuchasthedistancebetweenthetopandbottomofthebuildingblockis2.5666Å.\nHowever,Cmcm-B(β=90.0002゚)andCmcm-A(β=90゚)havelargeMAEdifference\ndueprobablytothedistortedangle.Cmcm-Cisahigher-energylayeredstructure,\nwhosetwoSilayersareseparatedbythreeColayers,projectionstowardsaandcaxes\nshowsimilarlayerseparationfeature.Itsbuildingblockisnottetragonalshapebut\ntrianglenestedshape.Co~Sidistances,whetherinthesameplaneorthenearest9neighbors,arenotexactlyanisoscelestriangle.Theβis90.0051inCmcm-D,\nmeaninganevenlargerdistortion,whosebuildingblockismorecloserwiththatof\nCom3mF,asaresultitsenergyishigherthanthestructurespackedbyCoP63/mmc.\nThesearchedCo3SiI4/mmm-A,(-6.9197eV/atom)isabout19meV/atomlower\nthanthatofP63/mmcandhasaperfecteasy-planeMAE,withavalueofE010=141.42\nμeV/atomalongthehardbaxis,whichisevidencedbytheidenticalprojections\ntowardsaandcaxesandthelatticeparametersa=c=5.0558,b=6.6642Å.\nConsideringthemultiplicityofthelatticeshapeweonlycalculatetheaxialenergy\nalongthreelatticeparametersasthisisthemostextensivelyuseddirections.\nThesearchedCo3SiI4/mmm-B(-6.9181eV/atom)isabout18meV/atomlower\nthanthatofP63/mmcandalsoshowstotally-perfecteasy-planeMAEwithavalueof\nE100=221.61μeV/atomalongthehardaaxis.ItsperfectMAEcouldbeattributed\npartiallytotheirtotally-symmetricarrangementsatthebothsidesofthemirrorthatis\nformedbythec-axisparallellinealongtheanglebisectorbetweenaandbaxis.In\nfact,weobtainseveralstructureswithsamesymmetryI4/mmmbuttheirMAEare\nevidentlydifferentowingtotheslightstrain,whichwillcauselargeaxial-energy\ndifference,suchasE001=153.08andE010=25.22inI4/mmm-CbutE001=0.03and\nE010=148.42μeV/atominI4/mmm-D,asisshowninTable1.Forsimplicity,weignore\ntocomparetheirbonddistancesastheyaregenerallysimilarwiththecasesofCmcm.\nTheotherstructuresalsohaveI4/mmmsymmetryandlargemagnetismbutpresent\nnon-perfectMAE,indictingthecomplicatedMAEfeature.10ThebuildingblockofCo3SiPmmn(-6.9143eV/atom)issimilarwiththatofCo\nP63/mmc,whoseMAEdifferseachotheralongthethreeaxesandshowslarge\ndiscrepancy.\nTheI4/mcm(-6.9025eV/atom)hasnearlysameenergywiththatofP63/mmc,\nwhoseE001=2.8andE100=211.67μeV/atom,respectively.Thisisalsoastructurewith\ntheMAElargerthan200μeV/atomandshowsnearly-perfecteasy-planeanisotropic\ndistributions.Despitethebuildingblockhaslargedistortionrelativetotheperfect\nBCC,theinter-atomicdistancestillhascertainsymmetry,forexample,\nCo1,6,7,10-Si1=2.4223,Si1-Si2,3=3.6626,Co2,5,9,12-Si1=2.7471,Co3,4,8,11-Si1=2.2918Å,\nrespectively.Thatis,theaveragedistancebetweenCo2,3,8,9~Si1andCo4,5,11,12~Si1are\nidenticalinthefirstshell,whichcontributestothelowenergyofthelattice.Moreover,\nthesecond-shellatomalsoshowshighlysymmetricarrangementsaroundtheSi1,\nwhichmightbetheoriginofthenearly-perfectMAE.Inaword,theMAEindeed\ndependsontheatomicarrangements.\nP21(-6.9071eV/atom)showsadistortedfirstandsecondshellsandthuspresents\n8/6differentCo-Sidistancesinthefirst/secondshells,respectively,whosefirst-shell\ndistancecouldgenerallybedividedintotwodifferentsortswithvaluesofabout2.38\nand2.48Å,respectively,whereasthesecond-shelldistancepresentslesssimilarities,\nincludingCo1-Si=2.5398,Co13-Si=2.4813,Co6-Si=3.048,Co8-Si=3.22,\nCo7-Si=3.1413,Co9-Si=2.2983Å,respectively.AlloftheneighborsoftheSiare\nsurroundedbytheCoatoms,whichmightcontributestoitslowerenergydespitethe\nlargedistortion.ItsaxialenergiesareE001=81.08andE100=137.16μeV/atom,11respectively,inaccordancewiththecaseofthelargely-disorderedatomic\narrangements.\nmPm3-A(-6.8891eV/atom)presentsnearly-perfecteasy-axisMAEwithvalues\nofE010=71.9andE100=72.1μeV/atom,respectively,whereasmPm3-B(-6.8876\neV/atom)showsperfecteasy-planeMAEwithavalueofE100=32.14μeV/atom.\nP4/mbm-A(-6.8976eV/atom)alsoshowsnearly-perfecteasy-planeMAEwith\nvaluesofE001=2.14andE100=144.51μeV/atom,respectively.Theothersearched\nCo3SiisshowninTableS3,inwhichmmF3-A(-6.8836eV/atom)showsgiant\nMAEwithrespectivevaluesofE001=375.13andE010=877.91μeV/atom.Mostof\nthesearchedlow-energyCo3Sihasthesimilarbuildingblockswiththatofstable\nCoP63/mmc,whereasthemajorityofthehigh-energyCo3Sihasthesimilar\nbuildingblockswiththatofmetastableCommF3,namely,thebuilding-block\nshapeoftheCodominatestheatomicarrangementandenergyorderofCo3Si.\nSinceCo3Sicouldpresentdiversestructureswithvariousmagnetismwetherefore\nalsosearchedCo3C,Co3Ge,Co3Sn,respectively.Unfortunately,onlyCo3Gehas\nnegativeformationofenergy.\nC.Thesearchedhigher-energystructuresofCo3SithanthatofP63/mmc\nThefollowingstructuresareseveralmeV/atomhigherthanthatofexperimental\nP63/mmcandmightalsobesynthesized.Asame-symmetrystructureP63/mmc-B\n(Ef=-6.8931eV/atom)withnearly-perfecteasy-planeMAEissearched,whose\nE001=80.28andE100=0.06μeV/atom,respectively.mPm3-A(Ef=-6.8891eV/atom)\npresentsnearly-perfecteasy-axisMAEwithenergiesofE010=71.9andE100=72.112μeV/atom.mPm3-Bshowsperfecteasy-planeMAE(Ef=-6.8876eV/atom)withan\nenergyofE100=32.14μeV/atom.P4/mbm-Astillpresentsnearly-perfecteasy-plane\nMAE(Ef=-6.8976eV/atom)withenergiesofE001=2.14andE100=144.51μeV/atom,\nrespectively.TheserichMAEcouldsatisfythediverseindustryrequirements.\nD.Predictedlow-energystructureofCo3Ge\nThesymmetryofthesearchedlow-energyCo3Gestructureisgenerallysame\nwiththatofCo3Si,asisshowninTable2andS4.Thesearchedtwolowest-energy\nstructuresalsoareCmcm(-6.5163)andI4/mmm(-6.4817eV/atom),withanenergy\ndifferenceof35meV/atom.Thegeneralbondlength(orsidelength)ofthemost\nstableCmcmstructureisabout2.59and2.48Åintheunit-formulaparallelogram,and\nthelongerCo-Gedistanceisabout4.4Å,asisplottedbytheredline.Weignoreto\nplotthebuildingblockasitissimilarwiththatofCo3Si.TheE010=0.24and\nE100=28.46μeV/atominCo3SiCmcmcorrespondstoE010=14.91andE100=32.19in\nCo3Ge.Similarly,aperfectandlargeeasy-axisMAE(E001=E010=261.14μeV/atom)in\nCo3Si(P63/mmc)correspondstoE001=332.58andE010=85.53μeV/atominCo3Ge\n(P63/mmc).P63/mmchassimilaratomicarrangementwiththatof1PinCo3Gebut\nthelatterpresentsnearly-perfecteasy-axisMAEwithvaluesofE010=116.38and\nE100=114μeV/atom,andtheenergyof1PislowerthanthatofP63/mmc,witha\ndifferenceofabout15meV/atom,stillrevealingthestrongatomiccoordinate\ndependenceofMAE.\nTwoCo3GeI4/mmmstructureshaveperfecteasy-planeMAE,withvaluesof\nEf=-0.0795eV/atomandE001=157.06μeV/atom,Ef=-0.0524eV/atomand13E010=9.6218μeV/atom,respectively.Infact,manyI4/mmmstructuresshow\nnearly-perfecteasy-axis/planeMAE,asisshowninTableS4,suchasonestructure\nshowsE001=156.9andE010=156.95μeV/atom,respectvely,whereasallofthese\nI4/mmmstructureshavenearlyidenticalenergies,whichmightbeagreatchallenge\nduringexperimentalsynthesization.\nPmmm(Ef=-0.0728eV/atom)hasnearly-perfecteasy-axisMAEwithvaluesof\nE001=122.41andE100=122.58μeV/atom,respectively.Theothersame-symmetry\nstructures(Pmmm)areshowninTableS4,whichfurtherdemonstratesthattheMAE\nisstronglydependentontheatomiccoordinate.Similarly,thefollowingstructures\nhaveenergiesabout70meV/atomhigherthanthatoflowest-energystructureCmcm\nandalsomightbesynthesized,suchasCmmmpresentsmediumMAEwithvaluesof\nE010=25.5andE100=26.3,P212121hasE010=24.15andE100=23.25,C2/mhas\nE001=26.18andE100=23μeV/atom,andsoon,thesenearly-perfecteasy-axisMAE\nstructureshavepotentialapplicationsinthelow-magnetismrequirementfield.Several\nstructuresshownearly-perfecteasy-planeMAEsuchasCmpresentsE001=40.92and\nE100=2.64,Ama2presentsE001=269.05andE100=0.78,C2221presentsE001=112.07,\nE010=4.75μeV/atom,andsoon.C2/mshowsrelativelylargeMAEwithvaluesof\nE010=42.56andE100=108.01μeV/atom,respectively,alloftheabovestructuresmight\nbesynthesized.Inaword,bothCo3GeandCo3Sicouldcrystallinemanylow-energy\nstructureswithdiversecoercivitymagnetism.Theselow-energyCo3Gestructuresare\nalsodominatedbythebuildingblockofCoP63/mmc,asisplottedbytheredline.\nE.PredictedCostructures14Todeterminethecommonbuildingblockshapes,wefurthersearchedtheCo\nstructure.Manylow-energyCostructuresaresearched,asisshowninTable2,in\nwhichonenewphaseispredicted,mR3(-7.0306eV/atom),whoseenergyiswell\nlocatedattheintermediaterangeofthetwoexperimentalphasesmFm3(-7.0206)and\nP63/mmc(-7.0398eV/atom).TwonewphasesincludingP42/mnm(-6.997)andCmca\n(-6.9972eV/atom)aresearchedwithenergiesof25meV/atomhigherthanthatof\nmFm3.Themagneticmomentare1.623,1.669,1.65μBforP63/mmc,m3Fm,and\nm3R,respectively.Allofthesestructureshavesimilarbuildingblockswiththe\nexceptionofcertaindistortion.SiandChavecomplicatedstructuresandalsohave\nbeensearchedformanytimesbytheresearcherinrecentyears,thusitisnot\nnecessarytosearchthemagain.\nF.PredictedcobaltsilicidesexcludingCo3Si\nFigure1showstheEfofthebinarycobaltsilicides.Forsimplicity,weconnect\nthelowest-energystructuresdirectlybecausewhethersomeofthesearched\nlowest-energystructurescouldbesynthesizedornotdependsonmanyfactors,such\nasitsownenergy,thesynthesizedtechniques,thedecomposedproducts,andsoon.As\nisshowninTableS5,CoSicrystallinesinP213atambientconditions,wesearched\nonemetastablephaseP21/c,itisabout30meV/atomhigherthanthatofP213,\nunfortunatelyitstillpresentszeromagnetism.Co2Sistabilizes[18]inPnma\n(Ef=-0.4501eV/atom)andshowsE001=32.03andE010=18.96μeV/atom,respectively.\nWepredictedonemetastableP2/m(Ef=-0.4011eV/atom)withnearly-perfect\neasy-axisenergiesofE010=23.45andE100=20.42μeV/atom,respectively.Theother15metastablestructureis50meV/atomhigherthanthatofPnma.\nThesearchedtwolowest-energyCo3Si2structureshavenegligiblemagnetism.\nAllofthefollowingstoichiometriesarelargerthan100meV/atominenergyin\ncomparisontotheirrespectivelowest-energypositionssitedwellattheconvex-hull\nline,includingCo8Si3(~160),Co5Si2(~100),Co7Si3(~100),Co9Si4(~150),Co9Si5\n(~160),Co7Si4(~160),Co8Si5(~150),Co4Si3(~100meV/atom),inaword,these\nstoichiometriesmighthavelower-energystructuresathighpressure,thusitisnot\nnecessarytostudythesestoichiometriesatambientconditionsasthecurrently\navailabletechnologymightbestillunabletosynthesizethem.Weshowthese\nlower-energystructuresandMAEforreferencepurposeonly.Allofthelow-energy\nCo8SistructuresshowsmallMAEbelow20μeV/atomandunperfecteasy-axis/plane\nMAE.WeignoretodiscusstheCo9SibecauseitsEfisabout0.Thesearched\nlow-energyCo6Sistructuresindeedhaslargemagnetism,whereasallofthemhaven’t\nperfecteasy-axis/planeMAE,onlyonestructureshowsnearly-perfecteasy-axisMAE\nwithvaluesofE001=19.89andE100=18.49μeV/atom,respectively,whichisabout30\nmeV/atomhigherthanthatofconvex-hullpoint,asisshowninTableS5.\nCo5Siexistsmanylarge-magnetismlow-energystructures,whereasnoneofthem\nshowsperfecteasy-axis/planeMAE,thelow/high-energyCo5Sistructureisalso\ndominatedmainlybythelow/high-energyCoP63/mmcandmmF3buildingblocks,\nrespectively,thatis,theenergyorderoftheCobuildingblockdeterminestheenergy\norderoftheCo5Si.Welistseverallow-energylarge-magnetismstructures,Ama2\n(Ef=-0.191,E010=336.35,E100=251.25),Cmc21(Ef=-0.1646,E001=321.31,E010=9.34),16P2(Ef=-0.1633,E010=435.12,E100=896.43),dI24(Ef=-0.1374,E001=180.58,\nE100=532.35),Cccm(Ef=-0.1171eV/atom,E001=37.85andE010=191μeV/atom),and\nsoon,someofwhichcouldbesynthesizedduetothesmallenergydifferencebetween\ntheirownenergiesandtheconvex-hullvalue,asisshowninTableS5andFigure1.\nThereareseveralnearly-perfecteasy-planestructureswithenergiesabout70\nmeV/atomhigherthanthatofthelowest-energystructureinCo5Si,suchasCc\n(Ef=-0.1646,E001=321.31,E010=9.34).Pmmm(Ef=-0.1227,E010=71.86,E100=1.78).P2\n(Ef=-0.146,E001=127.83,E010=7.5).Cccm(Ef=-0.1171,E001=37.84,E100=0.21),Fddd\n(Ef=-0.1626,E001=90.31,E010=5.5),Amm2(Ef=-0.1371eV/atom,E001=0.59,\nE100=31.4373μeV/atom),andsoon,asisshowninTableS5.Inaddition,thereare\nalsoseveralstructureswithnearly-perfecteasy-axissuchasC2,(Ef=-0.1395,\nE001=66.63,E100=71.5),P21/c(Ef=-0.1492,E010=20.81,E100=17.79),I41/amd\n(Ef=-0.1175,E001=46.8,E001=39.91).Pbcm(Ef=-0.1383eV/atom,E010=42.94,\nE100=35.17μeV/atom),andsoon,asareshowninTablesS5,6.Inaword,Co5Siisan\nimportantcompound.\nThesearchedtwolowest-energyCo9Si2Pbcm(Ef=-0.1272)andP1(Ef=-0.1275\neV/atom)areabout100meV/atomhigherthantheconvex-hullvalue,asisshownin\nTableS6,thusweignoretoanalysizethem.SeveralCo4SihavesimilarEf(-0.2\neV/atom),whichareabout50meV/atomhigherthanconvex-hullvalue,including\n3R(E001=73.12,E100=64.37),P213(E001=75.19,E010=380.39),I4/m(E010=60.73,\nE100=65.47),I4/m(E010=44.13,E100=81.44μeV/atom),inwhichoneortwostructures\nhavenearly-perfecteasy-axisMAE,asisshowninTableS5.Furthermore,P41hasa17largeMAE(E010=220.85,E100=190.99μeV/atom)butitisabout100meV/atomhigher\nthanconvex-hullvaluewithaEfof-0.1542eV/atom,asisshowninTableS6.Their\nbuildingblocksarealsodominatedbythelargelydistortedColattice.\nSomeCo7Si2mightbesynthesizedwhencomparingitsenergywiththatof\navailablemetastableCo3Si(P63/mmc)andCo4Si,asisshowninTableS5andFigure\n1,includinganearly-perfecteasy-axisMAE1P(Ef=-0.2118,E010=36.15,E100=36.96),\nP42/ncm(Ef=-0.2053,E001=43.63,E100=55.11)andP212121(Ef=-0.1911eV/atom,\nE001=2.19,E100=50.21μeV/atom),respectively.\nThesearchedtwolowest-energyCo5Si3Pbam(Ef=-0.481)andCmmm\n(Ef=-0.4737eV/atom)areabout2~3meV/atomhigherthanthatofconvex-hullvalue,\nwhereastheirMAEarenegeligible.BothFe5Si3andMn5Si3P63/mcmphasehave\nbeensynthesizedexperimentally,thecalculatedEfis-0.4212eV/atominCo5Si3\nP63/mcm,meaningametastablephase.Ourpreviouscalculationsshowthatthe\nexperimentalmetastablephaseFe5Si3(P63/mcm)exhibitsalargeeasy-planemagnetic\nanisotropywithavalueof157μeV/atom(E100=E010