Abstract
Magnetic properties of MnSb2Te4 were examined through magnetic susceptibility, specific-heat, and neutron-diffraction measurements. As opposed to isostructural MnBi2Te4 with the antiferromagnetic ground state, MnSb2Te4 develops a spontaneous magnetization below 25 K. From our first-principles calculations on the material in a ferromagnetic state, the state could be interpreted as a type-II Weyl semimetal state with broken time-reversal symmetry. Detailed structural refinements using x-ray-diffraction and neutron-diffraction data reveal the presence of site mixing between Mn and Sb sites, leading to the ferrimagnetic ground state. With theoretical calculations, we found that the presence of site mixing plays an important role for the interlayer Mn-Mn ferromagnetic interactions.
I. INTRODUCTION
Topological semimetals (TSMs), characterized by crossings of valence and conduction bands in the momentum space, have brought exciting opportunities to explore properties that are unattainable in conventional semimetals or metals [1–5]. Dirac semimetal is the first experimentally confirmed TSM, where a fourfold degenerate Fermi point is topologically protected by time-reversal symmetry (TRS) as well as space-inversion symmetry (SIS). This type of material has been experimentally verified in systems including Cd3As2 and Na3Bi [6–9]. The degeneracy at the Dirac point can be lifted by breaking of either symmetry, leading to the Weyl semimetal state. The resultant twofold degenerated Weyl points have a distinct chirality, resulting in an open arc surface state and chiral anomaly at high magnetic fields [9–11]. Weyl semimetals have been found in materials without SIS (e.g., NbP and TaAs) [12–14] or TRS (e.g., Mn3Sn and Co3Sn2S2) [15,16].
A new class of TSMs called “type-II TSMs” has recently been proposed from theoretical perspectives [17]. Unlike conventional type-I TSMs, Dirac cones in type-II TSMs are largely tilted along certain momentum directions, violating the Lorentz invariance [17]. The distinct band topology leads to extremely anisotropic magnetoresistance [17,18]. After the prediction of the type-II Weyl semimetal state in WTe2, several compounds were shown as type-II Weyl semimetal without SIS, including MoTe2, LaAlGe, and TaIrTe4 [19–25]. On the other hand, the experimental study of TRS-broken type-II Weyl semimetal is rather scarce. A rare example is the layered bismuthide YbMnBi2, where the canting of the Yb moment from the c axis was initially proposed to break the TRS, resulting in the type-II Weyl semimetal state [26]. Subsequent optical conductivity measurements in YbMnBi2, however, failed to confirm the canted Yb moments [27]. More recently, LnAlGe (Ln = Ce, Pr) have been predicted as a TRS broken TSM with both type-I and type-II crossings [28].
Ternary tellurides MPn2Te4 (M = Ge, Sn, Pb, Mn and Pn = As, Sb, Bi) have a layered structure as shown in Fig. 1(a) [29–31]. MnBi2Te4 with magnetic ions was theoretically proposed to have an antiferromagnetic topological insulating state [32–36], which was later confirmed from experimental studies [35–37]. It was also predicted that this material becomes TRS broken type-II Weyl semimetal if Mn moments are aligned ferromagnetically [34], a situation that has not been observed experimentally. In this paper, we synthesized isostructural MnSb2Te4 and conducted magnetic susceptibility, specific-heat, and neutron-diffraction measurements. Unlike the previous theory on MnSb2Te4 showing the A-type antiferromagnetic state [38], a ferromagnetic component develops in MnSb2Te4 below 25 K. Density functional theory calculations suggest that MnSb2Te4 in a ferromagnetic state can be expected to be a TRS broken type-II Weyl semimetal. Structural refinements using x-ray-diffraction (XRD) and neutron-diffraction data, however, reveal the presence of site mixing between Mn and Sb sites, leading to the ferrimagnetic ground state. The effect of site mixing is investigated through first-principles calculations.
FIG. 1.

(a) Crystal structure of MnSb2Te4. Pink, orange, and dark blue spheres represent Mn, Sb, and Te, respectively. Note that there are several antisite disorders between Sb and Mn sites. (b) Observed and calculated XRD patterns of MnSb2Te4. Red crosses and green and blue lines represent observed, calculated, and difference profiles. Green ticks are used to show the positions of the Bragg peaks.
II. EXPERIMENTAL AND CALCULATION DETAILS
A polycrystalline sample of MnSb2Te4 was synthesized via the solid-state reaction. Mn, Sb, and Te (99.99%, Kojundo Chemical) were weighed in a stoichiometric ratio, mixed, and pelletized in a nitrogen filled dry box. A pellet was sealed in an evacuated silica tube and heated up to 600 °C and kept for 10 h. Powder x-ray-diffraction experiments were performed using Bruker AXS D8 ADVANCE with Cu-Kα radiation.
Dc magnetic susceptibility was measured by a commercial superconducting quantum interference device magnetometer (Quantum Design, MPMS) in the temperature range of 2–350 K under the magnetic field of 0–7 T. Specific heat Ctotal was measured by the relaxation method using a physical property measurement system (Quantum Design) down to 2 K. To estimate the lattice contribution CL, we obtained the thermal variation of the Debye temperature θD(T) for the isostructural nonmagnetic analog GeSb2Te4 using the Debye equation. θD(T) of MnSb2Te4 was then estimated by multiplying a scaling factor according to θD ∝ M0−1/2V0−1/3, where M0 and V0 are the molar ratio mass and volume, respectively. CL was obtained by converting the scaled θD(T) into specific heat. Detail of the analysis is provided in the Supplemental Material [39]. Electronic structure calculations were performed within the generalized gradient approximation using the Vienna ab initio simulation package [40,41].
Powder neutron-diffraction (PND) data were collected at 50 and 5 K using the high-resolution powder diffractometer BT-1 (λ = 2.08 Å) at the NIST Center for Neutron Research, USA. The obtained diffraction data were structurally refined using the JANA2006 [42] and FullProf Suite [43]. We employed group theoretical analysis to identify magnetic structures that are allowed by symmetry.
III. RESULTS AND DISCUSSIONS
Figure 1(b) shows the room-temperature XRD profile of the target compound. All the observed Bragg peaks can be indexed based on the rhombohedral unit cell. The lattice parameters of a = 4.2385(3) Å and c = 40.8497(3) Å are slightly smaller than those of MnBi2Te4 (a = 4.334 Å and c = 40.910 Å) [31], which is reasonable given the smaller ionic radius of Sb than that of Bi. Figure 1(a) displays the structure of MnSb2Te4 (space group R3m), with NaCl-type MnTe layers and tetradymite-type Sb2Te3 layers stacking alternatively along the c axis. The edge-shared MnTe6 octahedral layer forms a regular triangular lattice. Since the formal valence is given by (Mn2+)(Sb3+)2(Te2−)4 and Mn2+ in a high-spin configuration has quenched the orbital degree of freedom, one can expect Heisenberg-type magnetic behavior in this material.
Eremeev et al. predicted that MnSb2Te4 has the A-type antiferromagnetic ground state, where the ferromagnetic Mn layers are stacked antiferromagnetically [38]. However, as shown in Fig. 2(a), the magnetic susceptibility of MnSb2Te4 in an applied field of 0.1 T rapidly increases below around 30 K, suggesting a transition to a ferromagnetic state (which, however, will be corrected later). From the Arrott plot [Fig. 2(c)], the transition temperature, TC, is estimated to be 25 K, close to the Néel temperature of 24.2 K observed in MnBi2Te4 [35–37,44]. The isotherm magnetization curve at 5 K up to 7 T [Fig. 2(b)] shows a small but finite ferromagnetic hysteresis loop. Such a small hysteresis has been sometimes reported in ferro- or ferrimagnets such as yttrium iron garnet [45]. The Curie-Weiss fitting yields the effective magnetic moment of peff = 5.842(7) μB, in good accordance with the value expected from the high-spin state of Mn2+ (5.92 μB). Figure S1 in the Supplemental Material [39] shows a difference in neutron data between T = 5 and 50 K. This clearly shows the increase of several nuclear reflections (e.g., 101 and 104) below TC, which is consistent with the ferromagnetic order with the magnetic wave vector qm = (0, 0, 0).
FIG. 2.

(a) Left: Temperature dependence of magnetic susceptibility under an applied field of 0.1 T with zero-field cooling (red) and field cooling (blue) process. Right: Temperature dependence of inverse susceptibility. (b) Isothermal magnetization curves at 5 K (red) and 100 K (blue). The inset shows a hysteresis loop. (c) M2 vs H/M (Arrott plot) at various temperatures around TC.
The magnetic transition is further probed by heat-capacity experiments. As shown in Fig. 3(a), the total specific heat Ctotal below 30 K is noticeably larger than that of the isostructural nonmagnetic GeSb2Te4, indicating the magnetic contribution to the specific heat. Its magnetic component Cmag is estimated by subtracting the lattice contribution from the data using GeSb2Te4 [Fig. 3(b), left axis]. Magnetic entropy, Smag, calculated from the integration of Cmag/T [Fig. 3(b), right axis], shows saturating behavior toward 30 K, followed by a gradual increase upon further warming. Smag reaches about 11 J/mol K2 at 100 K, slightly smaller than the expected value of Rln6 for S = 5/2, where R is the gas constant. This observation implies that the short-range correlation persists well above TC. A power-law behavior in Cmag ∝ Tα with α = 1.52(1) is observed at temperatures below 10 K [Fig. 3(c)]. If there is a long-range magnetic order, α follows the relation of α = d/υ, where d is the dimensionality of the magnon excitation and υ is related to the type of the magnetic order (υ = 1 and 2, respectively, stand for antiferromagnetic and ferromagnetic order). The observed T3/2 dependence in MnSb2Te4 implies three-dimensional-like magnon excitations from a ferromagnetic state. However, the limited fitting range prevented us from further elucidation of the magnetic excitations.
FIG. 3.

(a) Temperature dependence of the specific heat divided by temperature. (b) Left: The magnetic part of the specific heat divided by temperature. Right: The magnetic entropy. The horizontal line indicates R ln 6. (c) The magnetic part of the specific heat in full logarithmic scale.
As shown above, MnSb2Te4 undergoes a magnetic transition with a spontaneous magnetization, in contrast to the antiferromagnetic structure in MnBi2Te4 [36,37,44,46]. To obtain more insight into the ground state in MnSb2Te4, we performed first-principles calculations assuming a ferromagnetic state with an out-of-plane easy axis. Without the spinorbit coupling (SOC), MnSb2Te4 is a semiconductor with a direct band gap of Eg ~ 0.09 eV [Fig. 4(a)]. When SOC is turned on [Fig. 4(b)], we found for the first time a clear band inversion around the Γ point, forming a pair of band crossings along the Γ-Z line. The resultant Weyl cones are tilted with respect to the Γ-Z line, a characteristic feature of type-II Weyl points. From the observed spontaneous magnetization, we deduce that MnSb2Te4 is a potential TRS broken type-II Weyl semimetal at zero magnetic field.
FIG. 4.

Calculated band structures for MnSb2Te4 (a) without and (b) with SOC. In (a), the red (blue) color represents up-spin (down-spin) bands.
It is noticed, however, that without SOC the antiferromagnetic (A-type) structure is more stable than the ferromagnetic one by 18.3 meV, similarly to the previous result [38], and inclusion of SOC hardly affects the relative energy difference (17.5 meV). The discrepancy from the ferromagnetic ground state is also inferred from the Curie-Weiss fitting for the susceptibility that yielded a negative Weiss temperature of ΘW = −21.6(4) K) [Fig. 2(a)]. Furthermore, the saturation magnetization of about 1.8 μB/Mn at 5 K is somewhat smaller than that expected from high-spin Mn2+.
These contradictory observations led us to investigate the structural details of MnSb2Te4. A structural refinement was initially carried out assuming the ideal MnBi2Te4-type structure (R3m), which, however, resulted in rather poor reliability factors (e.g., Rwp ~ 11%) (Fig. S2 in the Supplemental Material [39]). Thus, site mixing among cations was considered in the analysis by allowing Sb (6c) to occupy the Mn (3a) site and vice versa, while the total occupancy factors were constrained to unity. This substantially improved the fitting [Goodness of fit (GOF) 1.77, Rwp = 4.51%, and Rp = 1.77%] and gave the chemical composition of (Mn0.66Sb0.34)(Sb0.83Mn0.17)2Te4. No site deficiency was found for the anionic sites. The full structural parameters are shown in Table I. Cation antisite disorder also occurs in other MPn2Te4-type compounds with a site mixing ranging from 15 to 35% [47,48].
Table I.
Crystallographic data for MnSb2Te4 obtained from XRD at room temperature (upper) and PND at 50-K (lower) data. The assumed space group is R3m with a = 4.2385(3) Å and c = 40.8497(3) Å from XRD and a = 4.2219(2) Å and c = 40.606(3) Å from PND. g represents the occupancy factor of each site, and Uiso employs an overall Debye-Waller factor.
| Atom | Site | gb | x | Y | z | Uiso/100 Å2 |
|---|---|---|---|---|---|---|
| Mn1 | 3a | 0.6645(3) | 0 | 0 | 0 | 1.62(3) |
| 0.666(7) | 0. 6(1) | |||||
| Sb1 | 3a | 0.3355(3) | 0 | 0 | 0 | 1.62(3) |
| 0.333(7) | 0.6(1) | |||||
| Sb2 | 6c | 0.8323(3) | 0 | 0 | 0.42802(15) | 1.62(3) |
| 0.833(4) | 0.4254(2) | 0.6(1) | ||||
| Mn2 | 6c | 0.1677(3) | 0 | 0 | 0.424802(15) | 1.62(3) |
| 0.167(4) | 0.4254(2) | 0.6(1) | ||||
| Te1 | 6c | 1 | 0 | 0 | 0.131456(12) | 1.06(3) |
| 0.13160(11) | 0.6(1) | |||||
| Te2 | 6c | 1 | 0 | 0 | 0.292259(14) | 1.29(3) |
| 0.29175(13) | 0.6(1) |
We refined the PND data to clarify the effect of site mixing on the magnetism. Above TC (at 50 K), the refined crystal structure is fully consistent with that obtained from the room-temperature XRD (Table I). At 5 K, we considered various magnetic structures including the ferromagnetic structure to account for the increased nuclear reflections. The best result with Rwp = 8.38% [Fig. 5(a)] was obtained for the ferrimagnetic structure, where the interexchanged Mn atoms at the 6c site align antiparallelly along the c axis with respect to Mn atoms at the original 3a site. Magnetic moments are 4.3(2) μB and 3.1(3) μB for 3a and 6c sites, respectively. The sublattice magnetic moment of Mn is 2.1 μB per unit cell, which is indeed in good accordance with the saturation magnetization value at 5 K [Fig. 2(b)].
FIG. 5.

PND patterns of MnSb2Te4 at (a) 5 K and (b) 50 K. Red crosses and green and blue lines represent observed, calculated, and difference profiles. Green ticks are used to show the positions of the Bragg peaks. (c) Proposed magnetic structure obtained from 5-K data.
The effect of site mixing between Mn and Sb sites was considered for theoretical calculations (see Fig. S3 in the Supplemental Material [39]). Here, we used two site-mixing patterns with a supercell and calculated the ferromagnetic, ferrimagnetic, and antiferromagnetic spin configurations along the [001] direction. In all the cases, we found magnetic moments only at the Mn atoms. The calculated relative energies are summarized in Table S1 in the Supplemental Material. In the antiferromagnetic case, there are two choices for the magnetic moment of Mn atoms in the Sb layer so we considered their average. As opposed to the disorder-free configuration as described above, the ferrimagnetic structure was found to be most stable. This indicates that the site mixing alters the interlayer exchange coupling from antiferromagnetic to ferromagnetic. We note that the band structures of the ferrimagnetic structures have a band gap. Since these band structures highly depend on the site-mixing patterns, we speculate that a fully random antisite mixing, if achieved, will recover the original crystal structure symmetry and the Weyl-semimetal phase may appear.
IV. CONCLUSION
To summarize, we have synthesized a layered chalcogenide MnSb2Te4 and measured its fundamental physical properties. The observed spontaneous magnetization below 25 K, differing from the antiferromagnetic order in MnBi2Te4, can be explained in terms of the ferrimagnetic order where the interexchanged Mn moments are antiparallelly coupled with the original Mn moments. First-principles calculations show that the introduction of antisite disorder plays a key role for the ferromagnetic interaction between Mn layers, which we believe is an important step toward the TRS broken type-II Weyl semimetal state. It should be noted that the Sb sample used here is in a polycrystalline form, which may hamper the observation of the expected topological properties. For further analyses of the topological state in MnSb2Te4, growth of a single crystal is required, which is in progress. The known MPn2Te4 compounds have been predicted to have topological states including the topological insulator, antiferromagnetic topological insulator, and topological axion state [35–37]. We demonstrate here that MnSb2Te4 represents an example in this family possessing the interlayer ferromagnetic interaction, which warrants further development of topological physics.
Supplementary Material
ACKNOWLEDGMENTS
This work was supported by a CREST project and Grant-in-Aid for Scientific Research on Innovative Areas, “Mixed anion” (Grants No. 16H6439, No. 16H6440, No. 17H05473, and No. 19H04683), and a Grant-in-Aid for Scientific Research (Grants No. 16H04007 and No. 17H06137) from Ministry of Education, Culture, Sports, Science and Technology. T.M. was supported by Japan Society for the Promotion of Science for Young Scientists.
References
- [1].Weng H, Dai X, and Fang Z, J. Phys. Condens. Matter 28, 303001 (2016). [DOI] [PubMed] [Google Scholar]
- [2].Yan B and Felser C, Annu. Rev. Condens. Matter Phys 8, 337 (2016). [Google Scholar]
- [3].Schoop LM, Pielnhofer F, and Lotsch BV, Chem. Mater 30, 3155 (2018). [Google Scholar]
- [4].Armitage NP, Mele EJ, and Vishwanath A, Rev. Mod. Phys 90, 015001 (2018). [Google Scholar]
- [5].Bernevig A, Weng H, Fang Z, and Dai X, J. Phys. Soc. Japan 87, 041001 (2018). [Google Scholar]
- [6].Liu ZK, Zhou B, Zhang Y, Wang ZJ, Weng HM, Prabhakaran D, Mo S, Shen ZX, Fang Z, Dai X, Hussain Z, and Chen YL, Science 343, 864 (2014). [DOI] [PubMed] [Google Scholar]
- [7].Borisenko S, Gibson Q, Evtushinsky D, Zabolotnyy V,Büchner B, and Cava RJ, Phys. Rev. Lett 113, 027603 (2014). [DOI] [PubMed] [Google Scholar]
- [8].Jeon S, Zhou BB, Gyenis A, Feldman BE, Kimchi I, Potter AC, Gibson QD, Cava RJ, Vishwanath A, and Yazdani A, Nat. Mater 13, 851 (2014). [DOI] [PubMed] [Google Scholar]
- [9].Xiong J, Kushwaha SK, Liang T, Krizan JW, Hirschberger M, Wang W, Cava RJ, and Ong NP, Science 350, 413 (2015). [DOI] [PubMed] [Google Scholar]
- [10].Huang X, Zhao L, Long Y, Wang P, Chen D, Yang Z, Liang H, Xue M, Weng H, Fang Z, Dai X, and Chen G, Phys. Rev. X 5, 031023 (2015). [Google Scholar]
- [11].Parameswaran SA, Grover T, Abanin DA, Pesin DA, and Vishwanath A, Phys. Rev. X 4, 031035 (2014). [Google Scholar]
- [12].Yang LX, Liu ZK, Sun Y, Peng H, Yang HF, Zhang T, Zhou B, Zhang Y, Guo YF, Rahn M, Prabhakaran D, Hussain Z, Mo SK, Felser C, Yan B, and Chen YL, Nat. Phys 11, 728 (2015). [Google Scholar]
- [13].Shekhar C, Nayak AK, Sun Y, Schmidt M, Nicklas M, Leermakers I, Zeitler U, Skourski Y, Wosnitza J, Liu Z, Chen Y, Schnelle W, Borrmann H, Grin Y, Felser C, and Yan B, Nat. Phys 11, 645 (2015). [Google Scholar]
- [14].Liu ZK, Yang LX, Sun Y, Zhang T, Peng H, Yang HF,Chen C, Zhang Y, Guo YF, Prabhakaran D, Schmidt M,Hussain Z, Mo SK, Felser C, Yan B, and Chen YL, Nat. Mater 15, 27 (2016). [DOI] [PubMed] [Google Scholar]
- [15].Kuroda K, Tomita T, Suzuki MT, Bareille C, Nugroho AA,Goswami P, Ochi M, Ikhlas M, Nakayama M, Akebi S,Noguchi R, Ishii R, Inami N, Ono K, Kumigashira H, Varykhalov A, Muro T, Koretsune T, Arita R, Shin S, Kondo T, and Nakatsuji S, Nat. Mater 16, 1090 (2017). [DOI] [PubMed] [Google Scholar]
- [16].Liu E, Sun Y, Kumar N, Muechler L, Sun A, Jiao L, Yang SY, Liu D, Liang A, Xu Q, Kroder J, Süß V, Borrmann H,Shekhar C, Wang Z, Xi C, Wang W, Schnelle W, Wirth S,Chen Y, Goennenwein STB, and Felser C, Nat. Phys 14, 1125 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [17].Soluyanov AA, Gresch D, Wang Z, Wu Q, Troyer M, Dai X, and Bernevig BA, Nature (London) 527, 495 (2015). [DOI] [PubMed] [Google Scholar]
- [18].Zyuzin AA and Tiwari RP, JETP Lett. 103, 717 (2016). [Google Scholar]
- [19].Wu Y, Mou D, Jo NH, Sun K, Huang L, Bud’ko SL,Canfield PC, and Kaminski A, Phys. Rev. B 94, 121113(R) (2016). [Google Scholar]
- [20].Deng K, Wan G, Deng P, Zhang K, Ding S, Wang E, Yan M, Huang H, Zhang H, Xu Z, Denlinger J, Fedorov A, Yang H, Duan W, Yao H, Wu Y, Fan S, Zhang H, Chen X, and Zhou S, Nat. Phys 12, 1105 (2016). [Google Scholar]
- [21].Wang Z, Gresch D, Soluyanov AA, Xie W, Kushwaha S,Dai X, Troyer M, Cava RJ, and Bernevig BA, Phys. Rev. Lett 117, 056805 (2016). [DOI] [PubMed] [Google Scholar]
- [22].Tamai A, Wu QS, Cucchi I, Bruno FY, Riccò S, Kim TK, Hoesch M, Barreteau C, Giannini E, and Besnard C, Soluyanov AA, and Baumberger F, Phys. Rev. X 6, 031021 (2016). [Google Scholar]
- [23].Belopolski I, Yu P, Sanchez DS, Ishida Y, Chang TR, Zhang SS, Xu SY, Zheng H, Chang G, Bian G, Jeng HT, Kondo T, Lin H, Liu Z, Shin S, and Hasan MZ, Nat. Commun 8, 942 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [24].Haubold E, Koepernik K, Efremov D, Khim S, Fedorov A,Kushnirenko Y, van den Brink J, Wurmehl S, Büchner B,Kim TK, Hoesch M, Sumida K, Taguchi K, Yoshikawa T,Kimura A, Okuda T, and Borisenko SV, Phys. Rev. B 95, 241108(R) (2017). [Google Scholar]
- [25].Zheng H, Lin H, Neupert T, Bian Y, Husanu M-A, Hsu C-H, Lu H, Belopolski I, Bansil A, Chang G, Hasan MZ,Sanchez DS, Zhang X, Alidoust N, Singh B, Jia S, Strocov VN, Xu S-Y, Bian G, Huang S-M, Jeng H-T, and Chang T-R, Sci. Adv 3, e1603229 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [26].Borisenko S, Evtushinsky D, Gibson Q, Yaresko A, Kim T, Ali MN, Buechner B, Hoesch M, and Cava RJ, arXiv:1507.04847. [Google Scholar]
- [27].Chaudhuri D, Cheng B, Yaresko A, Gibson QD, Cava RJ, and Armitage NP, Phys. Rev. B 96, 075151 (2017). [Google Scholar]
- [28].Chang G, Singh B, Xu SY, Bian G, Huang SM, Hsu CH,Belopolski I, Alidoust N, Sanchez DS, Zheng H, Lu H, Zhang X, Bian Y, Chang TR, Jeng HT, Bansil A, Hsu H,Jia S, Neupert T, Lin H, and Hasan MZ, Phys. Rev. B 97, 041104(R) (2018). [Google Scholar]
- [29].Kuznetsova LA, Kuznetsov VL, and Rowe DM, J. Phys. Chem. Solids 61, 1269 (2000). [Google Scholar]
- [30].Kuypers S, Van Tendeloo G, Van Landuyt J, and Amelinckx S, Micron Microsc. Acta 18, 245 (2002). [Google Scholar]
- [31].Lee DS, Kim TH, Park CH, Chung CY, Lim YS, Seo WS, and Park HH, Cryst. Eng. Comm 15, 5532 (2013). [Google Scholar]
- [32].Essin AM and Gurarie V, Phys. Rev. B 85, 195116 (2012). [Google Scholar]
- [33].Otrokov MM, Rusinov IP, Blanco-Rey M, Hoffmann M, Vyazovskaya AY, Eremeev SV, Ernst A, Echenique PM, Arnau A, and Chulkov EV, Phys. Rev. Lett 122, 107202 (2019). [DOI] [PubMed] [Google Scholar]
- [34].Zhang D, Shi M, Zhu T, Xing D, Zhang H, and Wang J, Phys. Rev. Lett 122, 206401 (2019). [DOI] [PubMed] [Google Scholar]
- [35].Otrokov MM, Klimovskikh II, Bentmann H, Zeugner A, Aliev ZS, Gass S, Wolter AUB, V Koroleva A, Estyunin D, and Shikin AM, arXiv:1809.07389. [Google Scholar]
- [36].Gong Y, Guo J, Li J, Zhu K, Liao M, Liu X, Zhang Q, Gu L, Tang L, and Feng X, Chin. Phys. Lett 36, 076801 (2019). [Google Scholar]
- [37].Chen B, Fei F, Zhang D, Zhang B, Liu W, Zhang S, Wang P, Wei B, Zhang Y, and Guo J, Nat. Commun 10, 4469 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Eremeev SV, Otrokov MM, and Chulkov EV, J. Alloys Compd 709, 172 (2017). [Google Scholar]
- [39]. See Supplemental Material at http://link.aps.org/supplemental/10.1103/PhysRevB.100.195103 for more details on analyses for XRD and PND data, theoretical calculations, and specific-heat measurements.
- [40].Kresse G and Furthmüller J, Phys. Rev. B 54, 11169 (1996). [DOI] [PubMed] [Google Scholar]
- [41].Kresse G and Furthmüller J, Comput. Mater. Sci 6, 15 (1996). [Google Scholar]
- [42].Petříček V, Dušek M, and Palatinus L, Z. Krist. Mater 229, 345 (2014). [Google Scholar]
- [43].Rodríguez-Carvajal J, Phys. B 192, 55 (1993). [Google Scholar]
- [44].Yan J-Q, Zhang Q, Heitmann T, Huang Zengle, Chen KY, Cheng J-G, Wu Weida, Vaknin D, Sales BC, and McQueeney RJ, Phys. Rev. Materials 3, 064202 (2019). [Google Scholar]
- [45].Mallmann EJJ, Sombra ASB, Goes JC, and Fechine PBA, Solid State Phenom. 202, 65 (2013). [Google Scholar]
- [46].Otrokov MM, Klimovskikh II, Bentmann H, Zeugner A,Aliev ZS, Gass S, Wolter AUB, Koroleva AV,Estyunin D, Shikin AM, Blanco-Rey M, Hoffmann M,Vyazovskaya AY, Eremeev SV, Koroteev YM, Amiraslanov IR, Babanly MB, Mamedov NT, Abdullayev NA, Zverev VN, Büchner B, Schwier EF, Kumar S, Kimura A, Petaccia L, Di Santo G, Vidal RC, Schatz S, Kißner K, Min C-H, Moser SK, Peixoto TRF, Reinert F,Ernst A, Echenique PM, Isaeva A, and Chulkov EV, arXiv:1809.07389. [Google Scholar]
- [47].Kuropatwa BA and Kleinke H, Z. Anorg. Allg. Chem 638, 2640 (2012). [Google Scholar]
- [48].Zeugner A, Nietschke F, Wolter AUB, Gaß S, Vidal RC,Peixoto TRF, Pohl D, Damm C, Lubk A, Hentrich R,Moser SK, Fornari C, Min CH, Schatz S, Kißner K, Ünzelmann M, Kaiser M, Scaravaggi F, Rellinghaus B, Nielsch K, Hess C, Büchner B, Reinert F, Bentmann H, Oeckler O, Doert T, Ruck M, and Isaeva A, Chem. Mater 31, 2795 (2019). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
