Bonding Interactions Can Drive Topological Phase Transitions in a Zintl Antiferromagnetic Insulator

Tanya Berry Department of Chemistry, Princeton University, Princeton, NJ 08540    Jaime M. Moya Department of Chemistry, Princeton University, Princeton, NJ 08540    David Smiadak Department of Chemical Engineering and Material Science, Michigan State University, East Lansing, MI 48824    Scott B. Lee Department of Chemistry, Princeton University, Princeton, NJ 08540    Sigalit Aharon Department of Chemistry, Princeton University, Princeton, NJ 08540    Alexandra Zevalkink Department of Chemical Engineering and Material Science, Michigan State University, East Lansing, MI 48824    Tyrel M. McQueen Department of Chemistry, Department of Materials Science and Engineering, Institute for Quantum Matter, William H. Miller III Department of Physics and Astronomy, The Johns Hopkins University, Baltimore, MD 21218    Leslie M. Schoop lschoop@princeton.edu Department of Chemistry, Princeton University, Princeton, NJ 08540
Abstract

While similar-to\sim30% of materials are reported to be topological, topological insulators are rare. Magnetic topological insulators (MTI) are even harder to find. Identifying crystallographic features that can host the coexistence of a topological insulating phase with magnetic order is vital for finding intrinsic MTI materials. Thus far, most materials that are investigated for the determination of an MTI are some combination of known topological insulators with a magnetic ion such as MnBi2Te4. Motivated by the recent success of EuIn2As2, we investigate the role of chemical pressure on topologically trivial insulator, Eu5In2Sb6 via Ga substitution. Eu5Ga2Sb6 is predicted to be topological but is synthetically difficult to stabilize. We look into the intermediate compositions between Eu5In2Sb6 and Eu5Ga2Sb6 through theoretical works to explore a topological phase transition and band inversion mechanism. We attribute the band inversion mechanism to changes in Eu-Sb hybridization as Ga is substituted for In due to chemical pressure. We also synthesize Eu5In4/3Ga2/3Sb6, the highest Ga concentration in Eu5In2-xGaxSb6, and report the thermodynamic, magnetic, transport, and Hall properties. Overall, our work paints a picture of a possible MTI via band engineering and explains why Eu-based Zintl compounds are suitable for the co-existence of magnetism and topology.

\alsoaffiliation

Department of Chemistry, The Johns Hopkins University, Baltimore, MD 21218 \alsoaffiliationDepartment of Materials Science and Engineering, The Johns Hopkins University, Baltimore, MD 21218

1 Introduction

Since the early 2000s with the Kane-Mele model, there has been a significant advancement not only in the theoretical classification of topological materials but also in their experimental realization. 1, 2, 3, 4, 5, 6, 7 There have also been advancements in discovering new topological materials with a multi-pronged approach that includes computation, chemical design principles, and data-driven approaches.8, 9, 10, 11 Still, when it comes to the experimental realization of topological insulators (TI’s), few ideal candidates are known, such as HgTe, monolayer WTe2, Bi1-xSbx, Bi2Te3, Bi2Se3, and variations of those. These TI’s exhibit various phenomena, such as the quantum spin Hall effect, and are well studied.12, 13, 14, 15, 16, 17, 18, 19 TIs are insulators that undergo band inversion and a change in parity. These insulators are unique from regular insulators; while being insulating in the bulk, they feature metallic, spin-momentum-locked, topological surface states (TSSs). To increase the library of materials that show band inversion, which is crucial for topological insulators, the scientific community has been searching for design principles that assist in rationally affecting the band structure of a material by playing with its’ elemental building blocks. 20, 11, 21, 22

Several band inversion mechanisms for non-magnetic TIs have been discussed in the literature, including spin-orbit coupling (SOC), the inert pair effect (IPE), strain, band folding, and orbital hybridization. 23, 21, 24, 2, 25, 26, 27, 22, 28, 29, 16, 15, 30, 31, 32, 33, 34, 5, 35, 36, 37, 38, 39, 40, 41 The early TI’s were mostly SOC and IPE-driven. For example, in Bi2Te3-type materials, the band inversion only appears when SOC is included in the calculation and without SOC there is no inversion. 2 Accounting for SOC does two things. It a) inverts the bands and b) opens a gap in the now inverted bands.42 In the case of IPE-driven TI’s, the band inversion appears due to the low lying 6s𝑠sitalic_s orbital (the inert pair) and thus appears mostly in TIs made of heavy p𝑝pitalic_p-block elements.22, 28, 29, 16, 15, 30 HgTe, the first known TI, is an example of the IPE-driven mechanism.When materials are synthesized as thin films, strain can be more easily applied, which allows for strain-driven band inversion, as seen in Bi (111) films. 33 If the materials crystallize in a non-symmorphic space group, the bands will be folded due to the additional translational symmetry. Band folding is a convenient way to introduce band inversions and is the reason bands are inverted in elemental Bi or square-net materials such as ZrSiS and Bi.5, 20, 43 Finally, band inversions can appear due to the hybridization of e.g. f and d bands. Such hybridization effects are seen in SmB6.41

No matter how the band inversion is induced, TI’s become even more interesting when they exhibit magnetic order. Magnetic order and the consequential breaking of time-reversal symmetry, can open a gap in the TSSs of a TI, and enable the realization of the quantum anomalous Hall effect (QAHE), which enables dissipation-less transport along edges of a 2D TI.44 Such a material is known as a magnetic topological insulator (MTI), which is subdivided into two categories: ferromagnetic and antiferromagnetic. Ferromagnetic topological insulators (FM TI’s) can arise from ferromagnetic doping, in the case of Cr-doped (Bi,Sb)2Te3, which exhibits the QAHE at 30 mK,45 or from intrinsically ferromagnetic topological insulators, such as MnBi2Te4 thin films, which show QAHE depending on the number of layers.46 The second category consists of antiferromagnetically ordered topological insulators (AFM TIs), which divides further into two subcategories; first are intrinsic antiferromagnetic topological insulators, also referred to as Axion insulators, where the antiferromagnetic order gaps-out all of the surface states (as also seen in antiferromagnetic MnBi2Te4 thin films or single crystals).47, 48 In such materials a quantized magnetoelectric response is expected.49 Second are generalized antiferromagnetic topological insulators, where only a subset of the surface states are gapped. This phenomenon has been observed in many recent studies of Eu-based Zintl materials, such as EuIn2As2, Eu3In2As4 and EuCd2As2.50, 51, 52 Designing a magnetic topological insulator introduces an additional complication in that the magnetic states cannot interfere with the band inversions giving rise to topology. This is why most magnetic topological insulators are derived from known non-magnetic TIs with the addition of a magnetic constituent, as in MnBi2Te4. We have previously identified that Zintl compounds are a promising material family to host magnetic topological insulators.53 The first design consideration is per the Zintl formalism: there are ionic (salt-like) interactions between cations and polyanionic networks, producing a large energy separation between cations and anions frameworks. This large energy separation provides autonomous chemical handles to tune the magnetism and topology, independently. The second design consideration is that, when looking at cations, we want a magnetically isotropic ion such that it can be polarized easily with an applied magnetic field. It should also be compatible with the oxidation states commonly found in Zintl materials. Since a majority of the Zintl cations have an oxidation state of 2+, Eu2+ (J𝐽Jitalic_J=S𝑆Sitalic_S=7/2 and L𝐿Litalic_L=0) is an ideal choice. The third design consideration is that, when considering polyanionic frameworks, we want it to be composed of heavy atoms that have small differences in electronegativity, so they have a small band gap and also a large spin-orbit coupling (SOC), to potentially induce a band gap in an inverted band structure.

Eu5In2Sb6 meets all of the design criteria with [Eu5]10+ as the magnetic cation framework and [In2Sb6]10- as the polyanionic framework (i.e. 2 In3+, 4 Sb3-, and 1 (Sb2)4-). Eu5In2Sb6 is a trivial magnetic insulator.54, 53 It was predicted with density functional theory (DFT) that an isoelectronic substitution of Ga in the In position would lower the lattice constant. This, in turn, can drive a band inversion, and make it topological.53, 55, 56 Naively, one might not expect a topological inversion, given the lower SOC in Ga or In. However, one might view the substitution of Ga for In as a chemical pressure effect that increases the bandwidth and this allows the gap to close and invert.57 Density Functional Theory is used to identify band inversion in materials and to better understand the mechanism that drives topological surface states through band inversion. However, DFT alone does not provide a full chemical understanding of the origin of such inversions, limiting our ability to create new materials that harbor such inversions.

Here, we show that the exact band inversion mechanism in Eu5In2-xGaxSb6 is more complex, involving not only the anionic framework but also cationic states, a “third party” bonding effect. This mechanism has not been previously explored to our knowledge and provides a new route to the creation of magnetic topological insulators. Further, we synthesize Eu5In2-xGaxSb6 to a maximal achievable substitution of x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, and report the magnetic and electronic properties of the material in single crystal form, showing that the trends are consistent with our calculations. We will help in the theoretical and practical design of future MTIs, and the broadening of the materials library.

1.1 Results and Discussion

1.2 Theoretical Investigation of Band Inversion

To investigate the nature of the band inversion from topologically trivial, Eu5In2Sb6 with Z𝑍{Z}italic_Z2 = 0 to topologically non-trivial, hypothetical Eu5Ga2Sb6 (Z𝑍{Z}italic_Z2 = 1), we first perform PBE+SOC (Perdew-Burke-Ernzerhof) calculations on a series of compositions Eu5In2-xGaxSb6 for 0x20𝑥20\leq x\leq 20 ≤ italic_x ≤ 2. Since the topology is sensitive to the crystal symmetry, as well as the number of bands at time-reversal invariant momenta, we carried out the calculations on ordered 1x1x3 supercells of the crystallographic unit cell of Eu5In2Sb6 containing mixtures of In and Ga. The 3-fold supercell was chosen because it is the smallest supercell that retains Pbam𝑃𝑏𝑎𝑚{Pbam}italic_P italic_b italic_a italic_m symmetry, it is capable of expressing In/Ga mixtures at relevant compositions (x= 0, 2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, 4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and 2), and the 3-fold enlargement will preserve the parity of bands at the ΓΓ\Gammaroman_Γ point, where the band inversion was previously identified (-13 = -1). The PBE+SOC relaxed unit cells of Eu5In2-xGaxSb6 (x= 0, 2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, 4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and 2) are given in Table S1.

Refer to caption
Figure 1: Band structures (PBE+SOC)of the Eu5In2-xGaxSb6 solid solution for (a) x=0, (b) x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, (c) x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and (d) x=2, from Y𝑌Yitalic_Y (0,12,0120\frac{1}{2},0divide start_ARG 1 end_ARG start_ARG 2 end_ARG , 0) to ΓΓ\Gammaroman_Γ (0,0,00000,0,00 , 0 , 0). The intensity of the dots indicates the fractional contribution of that band to the average structure. The composition discussed later in this work is Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG which is also squared in red. The parities of the Y𝑌Yitalic_Y to ΓΓ\Gammaroman_Γ point indicate a change in between x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG and x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, which implies a change in band inversion as a function of Ga substitution in the Eu5In2-xGaxSb6 structure. At x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG and 2 in the Eu5In2-xGaxSb6 solid solution the purple circled minus sign is above the Fermi level illustrating the switch in the parity that derives the topological phase tranistion which is not the case in the In-rich compositions. The structures of the 3c supercell of Eu5In2Sb6 and Eu5Ga2Sb6 are generated from the PBE+SOC calculations. It is important to note that these are not the actual crystal structures of the materials.

Band structure plots unfolded to the crystallographic Eu5In2Sb6 unit cell are shown in Figure 1 and Figure S1. When moving from x=0 to x=2, we observe significant band shifts at the ΓΓ\Gammaroman_Γ point. The lowest lying, positive parity, conduction band state above the Fermi level at x=0 moves downward in energy to below E𝐸Eitalic_Ef at x\geq2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG. The highest lying, negative parity, valence band state below the Fermi level at x=0 first moves slightly down in energy at x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG before increasing in energy and crossing above E𝐸Eitalic_Ef at x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG. It is the movement of this second band that causes the change in topology. These trends are in agreement with the previous calculation at x=0 and x=2 end members.

To determine where the topological phase transition occurs, the Z𝑍{Z}italic_Z2 index was calculated for each composition. Inversion symmetric insulators are Z𝑍{Z}italic_Z2 nontrivial if

n,m=0.1αfilledξnm(α)=1subscriptproduct𝑛𝑚0.1subscriptproduct𝛼𝑓𝑖𝑙𝑙𝑒𝑑subscriptsuperscript𝜉𝛼𝑛𝑚1\prod_{n,m=0.1}\prod_{\alpha\in filled}\xi^{(\alpha)}_{nm}=-1∏ start_POSTSUBSCRIPT italic_n , italic_m = 0.1 end_POSTSUBSCRIPT ∏ start_POSTSUBSCRIPT italic_α ∈ italic_f italic_i italic_l italic_l italic_e italic_d end_POSTSUBSCRIPT italic_ξ start_POSTSUPERSCRIPT ( italic_α ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT = - 1

where ξnm(α)subscriptsuperscript𝜉𝛼𝑛𝑚\xi^{(\alpha)}_{nm}italic_ξ start_POSTSUPERSCRIPT ( italic_α ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT are the inversion eigenvalues corresponding to the α𝛼\alphaitalic_αth Kramers doublet of bands at the momenta ΓnmsubscriptΓ𝑛𝑚\Gamma_{nm}roman_Γ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT=(nκ1𝑛subscript𝜅1n\kappa_{1}italic_n italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + mκ2𝑚subscript𝜅2m\kappa_{2}italic_m italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT)/2 (n,m=0,1formulae-sequence𝑛𝑚01n,m=0,1italic_n , italic_m = 0 , 1) and κisubscript𝜅𝑖\kappa_{i}italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the reciprocal lattice vectors.24 Utilizing the cross products at different time-reversal invariant momenta (TRIM) points as listed in Table 1, we determine that the change in parity happens between x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG and 4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG in the Eu5In2-xGaxSb6 system.

However, since the generalized gradient approximation (GGA), to which the PBE functional belongs, usually tends to underestimate band gaps and in this specific case it predicts negative band gaps i.e. metallic band structures, which is in contrast to experiments. We also carried out band structure calculations using the modified Becke-Johnson (mBJ) functional+SOC and the relaxed atomic coordinates and unit cells from the PBE+SOC calculation, Table S2. In contrast with PBE+SOC, mBJ+SOC predicts positive band gaps for all compositions, Figure-S2.

The magnitude of the predicted band gap is \approx55 meV which is in reasonable agreement with the transport gap.54 Carrying out a parity analysis in Table-S2, shows that the x=0, 2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, and 4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG compositions are topologically trivial and x=2 is topological non-trivial. As this result qualitatively agrees with the PBE (i.e both calculations confirm a band inversion appears eventually upon Ga substitution), we will proceed with PBE calculations to further investigate the mechanistic details of the inversion. We now turn to the chemical origin of this topological phase transition. Intuitively one might think that lowering the SOC upon substituting Ga for In is disadvantageous for creating a topological insulator. However, it should be kept in mind that SOC is what gaps inverted bands and only rarely induces band inversions. To deduce a chemical picture of why the substitution of In by Ga results in a topological band inversion, we carried out a Crystal Orbital Hamilton Population (COHP) analysis, Figure 2 and Figure S3.

x ratios ΓΓ\Gammaroman_Γ X𝑋Xitalic_X Y𝑌Yitalic_Y Z𝑍Zitalic_Z U𝑈Uitalic_U T𝑇Titalic_T S𝑆Sitalic_S R𝑅Ritalic_R
x=0 1 1 1 1 1 1 1 1
x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG 1 1 1 1 1 1 1 1
x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG -1 1 1 1 1 1 1 1
x=2 -1 1 1 1 1 1 1 1
Table 1: The parity outputs at the k points using GGA+SOC in Eu5In2-xGaxSb6.

The COHP method provides an energy-weighted picture of the bonding of a solid; negative (positive) numbers indicate the energy gained (lost) associated with bonding (antibonding) states respectively. In the COHP, averaged over all Eu-Sb, In-Sb, and Ga-Sb atom pairs, the values are largely unchanged across the series, except around -4 to -6 eV, and close to the Fermi level between, -1 to 0 eV (as seen in the inset of Figure 2). The changes in the former are reflective of changes in bonding in the cationic or anionic frameworks, while the latter reflect the changes associated with the band motions driving the topological phase transition.

We can gain further insights into the local bonding responsible for each of these features by looking at subsets of bonded pairs. In-Sb and Ga-Sb bonding forms the anionic framework of the structure. The pCOHP for these bonds show large changes in the -4 to -6 eV region but are all virtually superimposable around 0 eV (Figure 2(b) inset). Integration of the pCOHP of In/Ga-Sb in the energy window of 0 to -6 eV, where valence shell bonding interactions are expected to occur, reveals increasing Ga substitution disfavors the anionic structure from x=0 (-94.953), through x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG (-92.235) and x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG (-89.261) to x=2 (-86.394). This result is consistent with the experimental observation that phases with increased Ga substitution can not be synthesized. The observed similarities between these pCOHP curves near the Fermi energy, however, implies that while In-Sb and Ga-Sb bonding is crucial to the stability of the anionic framework, it is not (directly) responsible for the change in topology. In contrast, the pCOHP for Eu-Sb bonds shows smaller changes from -1 to -6 eV, but much more marked changes in the region near 0 eV, associated with the band motions that change the topology of the compound. This demonstrates that it is changes in Eu-Sb bonding (and, in particular, a stabilization of Eu-Sb bonding) that drive the topological phase transition. This explains why substituting Ga can produce a topological state: it destabilizes bonding within the anionic network; this, in turn, provides a driving force towards the formation of Eu-Sb bonds (which involve formal charge transfer from Eu to the anionic framework).

Refer to caption
Figure 2: The integrated pCOHP/bond analysis for (a) the average, (b) In/Ga-Sb bonds, and (c) Eu-Sb bonds for the Eu5In2-xGaxSb6 structure in x=0, x=1313\frac{1}{3}divide start_ARG 1 end_ARG start_ARG 3 end_ARG, x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, and x=1 respectively. The inset shows zoomed in views of the frames boxed in the three panels.

To our knowledge, this “third party effect” or complex hybridization has not been previously considered in the design of magnetic topological insulators in Zintl compounds. In most cases, hybridization effects are seen in a variety of materials such as EuCd2As2, CeOs4P12, CeOs4As12, Ce3Pt3Bi4, Ce3Pd3Bi4, Ta4SiTe4, and SmB6.38, 39, 40, 41 However, in most of the cases there is a coupling of f𝑓fitalic_f and d𝑑ditalic_d states. Recent works on GdTe2 and EuTe2 have shown that complex p𝑝pitalic_p-d𝑑ditalic_d-f𝑓fitalic_f hybridization can result in purely p𝑝pitalic_p-based physics of topological transport. 58 In our work, we see a coupling of Eu-d𝑑ditalic_d and Ga(In)/Sb-p𝑝pitalic_p states that drive the topological phase transition. Further Zintl-MTIs can likely be discovered through careful application of this “third party effect”.

1.3 Single Crystal Growth and Structural Characterization of Eu5In2-xGaxSb6

With this chemical understanding in mind, it is important to ask to what degree these predictions are in agreement with experimental observations. The synthesis objective is to achieve Ga isoelectronic substitution in the In site in Eu5In2-xGaxSb6 as x\rightarrow2 (near the Ga-rich phase), while maintaining the Pbam𝑃𝑏𝑎𝑚{Pbam}italic_P italic_b italic_a italic_m space group symmetry. To investigate the threshold of Ga substitution in the Eu5In2-xGaxSb6 structure, a series of synthesis attempts were carried out using flux, traditional solid synthesis, and chemical vapor transport methods. Of these synthesis approaches flux synthesis was most successful. These synthetic approaches were also employed to synthesize the hypothetical Eu5Ga2Sb6 (these form as minority phases that were not reproducible), however the phase was not stabilized and we formed the thermodynamically stable phase of EuGa2Sb2 instead.59

Refer to caption
Figure 3: Shift structural shifts from (a) Eu5In2Sb6 to (b) x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6. The tables below (a) and (b) explain the distortion in the trigonal pyramidal to a bent geometry in the (InSb4) frameworks and the overall shortening of the bond distances on Ga substitution. Please see Figure-S6 for more significant figures and errors of the bond lengths given.
Refer to caption
Figure 4: (a) R-value and Ga content in different Ga samples. These results show that different Ga substations are possible in the series. The single-crystal solutions to these are in the SI. The data points with yellow highlights are the ones with the lowest R-value. The inset shows A picture of the single crystal of Eu5In2-xGaxSb6 with x\approx2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG. The inset show a picture of the single crystal of Eu5In2-xGaxSb6 with x\approx2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG. (b) The overall composition determined by SEM-EDX and single crystal stricture solutions. (c) The SEM-EDX maps of a few different compositions of Eu5In2-xGaxSb6. The inset shows the SEM of a few of the single crystals.

As seen in Figure 3, the In3+ and Ga3+ ionic radii are 80 pm and 62 pm which are quite different.60, 61 The maximum x synthetically achievable in the Eu5In2-xGaxSb6 series was x\approx2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG. The determination of the x\approx2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG was done using single crystal X-ray diffraction (SXRD) utilizing the R-value analysis and scanning electron microscopy-energy dispersive X-ray analysis (SEM-EDX) difference in the In:Ga contents, as seen in Figure 4 and Table S3-S5.

The difference in form factors between the two species causes local packing tensions effects and notable shifts in the bond lengths also shown in Figure 3. If we consider the crystal structures of Eu5In2Sb6, it becomes evident that the distortion of the polyhedra becomes most prominent as we approach the x\approx2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG limit in the Eu5In2-xGaxSb6 structure. This distortion suggests the collapse of the structure upon a higher threshold of Ga substitution in the Pbam𝑃𝑏𝑎𝑚{Pbam}italic_P italic_b italic_a italic_m structure.

The Ga to InSb ratios derived from SXRD and EDX characterization do not match the nominal ratios given by the flux synthesis i.e. the input Ga concentration is not monotonically related to InSb flux ratios. Such deviations in flux synthesis are common and suggest that there is a set concentration where the flux composition reaches a dissolving minimum as a function of temperature. However, within each batch various crystals were compared and the composition was consistent as seen via repeated flux growths and SXRD studies. Previous work aiming to substitute Cd and Zn in the Eu5In2Sb6 system found an even lower threshold of x less than 0.1.6263, 57 Overall, a composition of x\approx2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG in Eu5In2-xGaxSb6 was achieved in the single crystal form and it is close to the topological phase transition.

1.4 Transport Behavior of Eu5In4/3Ga2/3Sb6

Next, we study the magnetic and electronic properties of the compound with the highest Ga content we could grow, Eu5In4/3Ga2/3Sb6. While at this substitution level, the topological phase transition is not yet expected to occur, it is still of interest to study how achievable even this level of substitution affects the properties as compared to Eu5In2Sb6, which has been shown to have a non-collinear magnetic structure and colossal magnetoresistance.54, 55 In our Ga-substituted samples, we observe that the antiferromagnetic transition temperatures arise from Eu2+ S=J=7/2 to have similar features to non-Ga substituted Eu5In2Sb6.54, 55 A discussion of the magnetization as a function of field and temperature as well as heat capacity results explaining the nature of antiferromagnetism in Eu5In4/3Ga2/3Sb6 is described in the supplemental text and Figure-S4, S5 respectively.

To see the implication of chemical pressure on the electronic properties of Eu5In4/3Ga2/3Sb6 single crystals, resistivity as a function of temperature with various fields was investigated. Figure 5 shows the resistivity was measured μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H//b and jperpendicular-to\perpb. At T <<<15 K we observe three transitions that overlap with the magnetic transition temperatures as observed in the M(T) plots at μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H//b at μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=0.1T. Therefore, those features in the resistivity could be attributed to the loss of spin disorder scattering. However, at Tsimilar-to\sim60 K, there is a change in the resistivity i.e. T<<<60 K the resistance of Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG behaves as a metal (dρ/dT>0𝜌d𝑇0\rho/\text{d}T>0italic_ρ / d italic_T > 0), and T>>>60 K it behaves as an insulator (dρ/dT<0𝜌d𝑇0\rho/\text{d}T<0italic_ρ / d italic_T < 0). The feature at Tsimilar-to\sim60 K can be decoupled to the magnetic order since it is above the ordering temperatures of Eu. Such a feature is commonly seen in topological insulators such as Bi2Se3 and related materials and is often attributed to changes in electronic surface states or defects.64 On applying μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H of up to 9 T both the high-temperature and magnetic anomalies are suppressed. This suppression suggests a strong coupling that occurs with a combination of effects such as suppression of strong magnetic fluctuations that may persist even at temperatures well above the magnetic ordering temperature and the coupling of defect or surface states. Due to chemical pressure and increased bonding interaction, the resistivity of Eu5In4/3Ga2/3Sb6 single crystals is orders of magnitude lower than the parent Eu5In2Sb6. In the parent Eu5In2Sb6, we also do not see the Tsimilar-to\sim60 K feature in the resistivity.54 The band gap conceptualized via the Arrhenius model shows Eg=52 meV. Overall, the less resistive behavior is indicative of the contraction of the In/Ga-Sb bond and the distortion in the polyanionic polyhedral of [In2Sb6]10- that may imply that Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG getting close to the topological phase transition.

Refer to caption
Figure 5: (a) Electrical resistivity of x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6 single crystal ( i.e. sample-3 from Figure-4) at 0.1 T applied field at the c-axis for applied current. The first inset describes the magnetic transitions due to magnetic order as previously seen in heat capacity and magnetization plots. The second inset describes the gap of 52 meV which is lower than Eu5In2Sb6. (b) Resistivity at a few different applied magnetic fields and the shift conductivity as the applied magnetic field increases.
Refer to caption
Figure 6: (a) Isothermal magnetoresistance of Eu5In4/3Ga2/3Sb6 single crystals measured (i.e. sample-3 from Figure-4) for temperatures 2K<<<T𝑇Titalic_T<<<45K and magnetic fields 0T<<<μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H<<<9T for μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT Hperpendicular-to\perpb and j//b. (b) Contour plot of the field-dependent derivative, (c) temperature dependence of the MR at μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=9 T, (d) Antisymmetrized Hall resistivity ρ𝜌\rhoitalic_ρyx data measured corresponding to panel (a). (e) Shows onset of TN, where TN1= 15.0 K, TN2= 5.7 K at μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=0.1 T. (f) Clarifies the dependence of σxyAsubscriptsuperscript𝜎𝐴𝑥𝑦\sigma^{A}_{xy}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT on σxy1.6subscriptsuperscript𝜎1.6𝑥𝑦\sigma^{1.6}_{xy}italic_σ start_POSTSUPERSCRIPT 1.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT for T<<<TN respectively.

To further corroborate the temperature-dependent transport behavior, the Hall resistivity and magnetoresistance were measured. In Figure 6 (a) we see the isothermal magnetoresistance of Eu5In4/3Ga2/3Sb6 single crystals measured for temperatures 2K<<<T𝑇Titalic_T<<<45K and magnetic fields 0T<<<μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H<<<9T for μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT Hperpendicular-to\perpb and j//b. The magnetoresistance (MR) is highly sensitive to metamagnetism, which is highlighted in the contour plot of the field-dependent derivative shown in Figure 6 (b). Now, focusing on the high-field data, Figure 6 (c) shows the temperature dependence of the MR at μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=9 T. Similar to the parent compound, Eu5In2Sb6, a colossal negative MR is registered, ranging from MR(μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=9 T) similar-to\sim -80% at T𝑇Titalic_T=45 K before monotonically increasing on cooling to a maximum of MR(μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=9 T) similar-to\sim-94% at T𝑇Titalic_T=17 K, just above TN. Upon further cooling below TN, MR(9 T) decreases to similar-to\sim-90% at TN2 before, again increasing to similar-to\sim-94 at T=2 K. In Eu5In2Sb6, the colossal negative MR was speculated to originate from the formation of magnetic polarons.54 Our Hall data, presented in Figure 6(d) provides further evidence of possible magnetic polarons below TN.

Figure 6(d) further shows the corresponding Hall resistivity ρ𝜌\rhoitalic_ρyx data measured and antisymmerterized for temperatures 2K<<<T𝑇{T}italic_T<<<45K and magnetic fields -9T<<<μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H<<<9T for μ𝜇\muitalic_μoH perpendicular to b and j//b. For T>>>TN, ρ𝜌\rhoitalic_ρyx shows a clear multi-band character with dominant hole-like conduction, evidenced by the positive (negative) curvature and sign in positive (negative) measured fields.

Upon cooling below TN the ρ𝜌\rhoitalic_ρyx has a similar behavior at high fields, however, a spontaneous anomalous Hall effect is measured which is observed as ρ𝜌\rhoitalic_ρyx(μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=0T)0absent0\neq 0≠ 0 in Figure 6(d). The temperature dependence of the anomalous Hall conductivity σxyA=ρyx(μoH=0T)ρxx2+ρyx2subscriptsuperscript𝜎𝐴𝑥𝑦subscript𝜌𝑦𝑥subscript𝜇𝑜𝐻0𝑇subscriptsuperscript𝜌2𝑥𝑥subscriptsuperscript𝜌2𝑦𝑥\sigma^{A}_{xy}=\frac{-\rho_{yx}(\mu_{o}H=0T)}{\rho^{2}_{xx}+{\rho^{2}_{yx}}}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT = divide start_ARG - italic_ρ start_POSTSUBSCRIPT italic_y italic_x end_POSTSUBSCRIPT ( italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT italic_H = 0 italic_T ) end_ARG start_ARG italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y italic_x end_POSTSUBSCRIPT end_ARG is shown in Figure 6(e), clearly demonstrating the onset below TN, where TN1= 15.0 K, TN2= 5.7 K at μoHsubscript𝜇𝑜H\mu_{o}\textit{H}italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT H=0.1 T. To clarify the origin of the anomalous Hall, we plot the dependence of σxyAsubscriptsuperscript𝜎𝐴𝑥𝑦\sigma^{A}_{xy}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT on σxy1.6subscriptsuperscript𝜎1.6𝑥𝑦\sigma^{1.6}_{xy}italic_σ start_POSTSUPERSCRIPT 1.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT for T<<<TN in Figure 6(f). In the hopping regime corresponding to σxx=ρxxρxx2+ρyx2<104(Ω.cm)1\sigma_{xx}=\frac{\rho_{xx}}{\rho^{2}_{xx}+{\rho^{2}_{yx}}}<10^{4}(\Omega.cm)^% {-1}italic_σ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT = divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y italic_x end_POSTSUBSCRIPT end_ARG < 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( roman_Ω . italic_c italic_m ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, such a power law behavior is expected due to the percolation of magnetic clusters.6566 By contrast, in the metallic regime where Berry curvature is expected to dominate, σxyAsubscriptsuperscript𝜎𝐴𝑥𝑦\sigma^{A}_{xy}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT is expected to be independent of impurity scattering, that is σxyAsubscriptsuperscript𝜎𝐴𝑥𝑦\sigma^{A}_{xy}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT = constant𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡constantitalic_c italic_o italic_n italic_s italic_t italic_a italic_n italic_t. In the most metallic systems (σxy>106(Ω.cm)1)(\sigma_{xy}>10^{6}(\Omega.cm)^{-1})( italic_σ start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ( roman_Ω . italic_c italic_m ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ), skew-scattering dominates and σxyAsubscriptsuperscript𝜎𝐴𝑥𝑦\sigma^{A}_{xy}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPTsimilar-to\sim σxysubscript𝜎𝑥𝑦\sigma_{xy}italic_σ start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT.65 Though the data is noisy, Figure 6(f) shows that σxyAsubscriptsuperscript𝜎𝐴𝑥𝑦\sigma^{A}_{xy}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT, is clearly not independent of σxxAsubscriptsuperscript𝜎𝐴𝑥𝑥\sigma^{A}_{xx}italic_σ start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT, and the exponent is certainly greater than 1, consistent with the percolation of magnetic clusters, like magnetic polarons. However, the hysteresis observed in ρxysubscript𝜌𝑥𝑦\rho_{xy}italic_ρ start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT, Figure-6(c), is not noticed in M(μ𝜇\muitalic_μ0H) which suggests that there may be an intrinsic mechanism that derives the anomalous Hall effect. Still, a Berry-phase-induced anomalous Hall, which would be a sign of topology, is not seen, consistent with the theoretical prediction for this substitution level.

2 Conclusion

We explore the effect of chemical pressure to induce a topological phase transition in the previously reported topologically trivial magnetic insulator Eu5In2Sb6 with Z𝑍Zitalic_Z2=0. Although Eu5Ga2Sb6 had been predicted to exist with a Z𝑍Zitalic_Z2=1, it is synthetically challenging to make and structurally difficult to stabilize in the Pbam structure. We thus investigated the effect of Ga substitution in Eu5In2-xGaxSb6 with x=0, 2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, 4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and 2 to see where the topological phase transition occurs between the two end compositions Eu5In2Sb6 with Z𝑍Zitalic_Z2=0 and Eu5Ga2Sb6 with Z𝑍Zitalic_Z2=1. Per our DFT results utilizing PBE+SOC, we saw that there is a change in parity that occurs between x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG and x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG (or between x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG and x=2 when using the mBJ functional). Crystal Orbital Hamilton Population calculations show that there is a mixture of cationic and polyanionic framework that contributes to the band inversion via changes in Eu-Sb bonding driven by In-Sb/Ga-Sb substitution. We thus, attribute the changes to the third-party or complex hybridization-mediated band inversion effect. Our work introduces a possible magnetic topological insulator via band engineering and explains why Eu-based Zintl compounds are suitable for the co-existence of magnetism and topology.

3 Experimental Section

3.1 Single Crystal Growth, X-ray Diffraction, and X-ray Energy Dispersive Spectroscopy

The single crystals of x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6 series were grown via InSb binary flux and were air-stable on a benchtop for months. For sample-3 with the composition of x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6, the Eu:In:Ga:Sb ratios were 1.0550, 0.8801, 0.57355, and 2.5411 respectively, with a total mass of 5.0498 grams. The elements Eu (ingot, Yeemeida Technology Co., LTD 99.995%), In (shots, Sigma-Aldrich 99.99%), Ga (ingot, Noah Tech 99.99%), and Sb (BTC, 99.999%) were utilized for the synthesis. In, Ga, and Sb were loaded in a Canfield crucible (size: 2 mL) at atmospheric conditions, while Eu was loaded in the crucible in an Ar-filled glove box. The Canfield crucible was placed in a quartz ampoule with quartz wool below and above the crucible, evacuated, and sealed under 5.4×10–2 Torr of pressure. The evacuated ampoules were loaded in a box furnace at an angle of similar-to\sim45° with the charge facing the center on the backside of the central interior of the box furnace. The temperature was ramped at a rate of 80°C/hour to 550°C and dwell for 12 hours which allowed for the Ga and InSb fluxes to be in the liquid state. Then the box furnace temperature was ramped at the rate of 80°C/hour to 1100°C and dwelled for 24 hours to allow the Eu5In2-xGaxSb6 to dissolve in the fluxes. The furnace was then cooled at the rate of 5°C/hour to 650°C for the dissolved phase to crystallize on cooling and then centrifugation separate the flux mixtures from the single crystals. The single crystals had a flat rod-like morphology, 1.5-2 mm in width and 4-12 mm in length.

Single Crystal X-ray diffraction and X-ray energy dispersive spectroscopy (EDS) were used to confirm the phase purity and elemental composition of the single crystals. Single crystal X-ray diffraction data were collected using a SuperNova diffractometer equipped with an Atlas detector and a Mo Kα𝛼\alphaitalic_α source. The cuboid crystal, cut from a larger crystal piece, was mounted with Paratone-N oil. Data was analyzed and reduced using the CrysAlisPro software suite, version 1.171.36.32 (2013), Agilent Technologies. Initial structural models were developed using SIR92 and refinements of this model were done using SHELXL-97 (WinGX version, release 97-2).2,3 Real-time back reflection Laue X-ray diffraction was used to orient and align the crystals for measurement. Energy-dispersive x-ray spectroscopy (EDX) in a Quanta environmental scanning electron microscope (SEM) equipped with an Oxford EDX detector at 25keV. Homogeneity of composition was observed using the EDX mapping technique, with small errors in composition likely arising from excess flux left on the surface of crystals.

3.2 Theoretical Calculations:

DFT calculations were performed using Quantum Espresso 7.0 6768 with the PBE functional and pseudopotentials obtained from QUANTUM ESPRESSO.69 To accommodate variable occupancies on the In site, the reported Pbam𝑃𝑏𝑎𝑚Pbamitalic_P italic_b italic_a italic_m unit cell of Eu5In2Sb6 was tripled along the crystallographic c axis to produce a new cell similar-to\sim 11.6x15.1x12.8 Å that retains Pbam𝑃𝑏𝑎𝑚Pbamitalic_P italic_b italic_a italic_m symmetry but has two crystallographically distinct In sites (4h and 8i). All calculations were carried out using this enlarged cell for Eu30In12Sb36 (x = 0), Eu30In8Ga4Sb36 (x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG), Eu30In4Ga8Sb36 (x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG), and Eu30Ga12Sb36 (x = 1). Initially, variable cell relaxations were carried out to minimize the forces and stresses in fully relativistic calculations with spin-orbit coupling and non-colinear magnetization with time-reversal symmetry (no magnetic order). Relaxations were performed using a 4x3x4 Monkhorst-Pack mesh, a 55 Ry kinetic energy cutoff, a 440 Ry charge density cutoff, and Marzari-Vanderbilt-DeVita-Payne cold smearing70, and converged to an energy threshold of 7.8·10-4, and force threshold of 10-4. This resulted in the cell parameters given in Table-S1. The relaxed unit cells and atomic positions were then fixed and used for self-consistent-field calculations using a 6x4x6 Monkhorst-Pack mesh, a 55 Ry kinetic energy cutoff, a 440 Ry charge density cutoff, and Marzari-Vanderbilt-DeVita-Payne cold smearing 70, and converged to an energy error of 1.5·10-8 Ry. These SCF calculations were carried out for both the spin-orbit-coupled and non-spin-orbit-coupled cases. Band unfolding to the original (non-enlarged) cell was carried out using a publicly available tools at GitHub. Symmetry analysis of bands at the time-reversal invariant momenta was carried out using the post-processing tools of Quantum Espresso.6768 Crystal Orbital Hamilton Population (COHP) analysis for the non-SOC calculations was carried out using LOBSTER.71 COHP for the SOC case was carried out using the same framework72 implemented as a custom Python code and checked for consistency with the non-SOC LOBSTER results. The DFT MBJ+SOC calculations were performed using the converged lattice parameters from PBE+SOC as given in Table-S1.

3.3 Bulk Magnetic Properties Measurements

Magnetization data were collected on a Quantum Design magnetic property measurement system (MPMS). Magnetic susceptibility was approximated as magnetization divided by the applied magnetic field (χ𝜒\chiitalic_χ\approxM/H𝑀𝐻M/Hitalic_M / italic_H). Magnetization measurements were performed in a vibrating sample superconducting quantum interference device magnetometer (SQUID-VSM) from Quantum Design. All measurements were carried out after cooling in a zero field. To reduce the remnant field of the superconducting magnet to less than 2 Oe before each measurement, we applied a magnetic field of 7 T at ambient temperature and then removed it in an oscillation mode. The magnetic field was applied to the a, b, and c axis to the flat rod-like crystal. Sample shape correction was accounted for in these measurements.

3.4 Thermodynamic Properties

Heat capacity measurements were performed in a Quantum Design Physical Properties Measurement System (PPMS). Measurements were performed on a single crystalline sample of x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6 of 2.9(3) mg mass oriented such that the applied magnetic field was along the nominal c-axis to the current. The heat capacity was measured T𝑇Titalic_T=2-300K and various applied fields. The sample was mounted on the sample stage using Apiezon N grease.

3.5 Electrical Transport Properties:

The resistivity option in the PPMS-9 was utilized to carry out the resistivity measurements. The resistivity was measured from T = 2-300 K using the four-probe technique at various applied magnetic fields. The leads were made from Pt wire and the contacts were made using Dupont 4922N Ag paste. The Pt lead distance was 0.38 mm. The sample length was 1.3 mm longitudinally.

Electrical-magneto transport measurements were performed in a Quantum Design Dynacool Physical Properties Measurement System using the Electrical Transport Option. Contacts to the sample were made in a standard four-point or Hall bar geometry. Gold wire 0.025m annealed (metal basis) was utilized for the electrodes and Ag conductive paint SPI# 05002-AB was utilized for the contacts. For magnetoresistance and Hall measurements, full hysteresis loops were measured such that Quadrant I denoted as QI was measured from μoH>0T0Tsubscript𝜇𝑜𝐻0𝑇0𝑇\mu_{o}H>0T\rightarrow 0Titalic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT italic_H > 0 italic_T → 0 italic_T, QII from 0TμoH<0T0𝑇subscript𝜇𝑜𝐻0𝑇0T\rightarrow\mu_{o}H<0T0 italic_T → italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT italic_H < 0 italic_T, QIII from μoH<0T0Tsubscript𝜇𝑜𝐻0𝑇0𝑇\mu_{o}H<0T\rightarrow 0Titalic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT italic_H < 0 italic_T → 0 italic_T, and QIV from 0TμoH>0T0𝑇subscript𝜇𝑜𝐻0𝑇0T\rightarrow\mu_{o}H>0T0 italic_T → italic_μ start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT italic_H > 0 italic_T. The data were symmetrized or antisymmetrized, respectively such that

ρxx(dH>0,dH<0)=Rmeas(QIV,QI)+Rmeas(QII,QIII)2wtLsubscript𝜌𝑥𝑥formulae-sequenced𝐻0d𝐻0subscript𝑅𝑚𝑒𝑎𝑠𝑄𝐼𝑉𝑄𝐼subscript𝑅𝑚𝑒𝑎𝑠𝑄𝐼𝐼𝑄𝐼𝐼𝐼2𝑤𝑡𝐿\displaystyle{\rho_{xx}(\text{d}H>0,\text{d}H<0})=\frac{{R_{meas}(QIV,QI)}+{R_% {meas}(QII,QIII)}}{2}\frac{wt}{L}italic_ρ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( d italic_H > 0 , d italic_H < 0 ) = divide start_ARG italic_R start_POSTSUBSCRIPT italic_m italic_e italic_a italic_s end_POSTSUBSCRIPT ( italic_Q italic_I italic_V , italic_Q italic_I ) + italic_R start_POSTSUBSCRIPT italic_m italic_e italic_a italic_s end_POSTSUBSCRIPT ( italic_Q italic_I italic_I , italic_Q italic_I italic_I italic_I ) end_ARG start_ARG 2 end_ARG divide start_ARG italic_w italic_t end_ARG start_ARG italic_L end_ARG (1)
ρyx(dH>0,dH<0)=Rmeas(QIV,QI)Rmeas(QII,QIII)2t.subscript𝜌𝑦𝑥formulae-sequenced𝐻0d𝐻0subscript𝑅𝑚𝑒𝑎𝑠𝑄𝐼𝑉𝑄𝐼subscript𝑅𝑚𝑒𝑎𝑠𝑄𝐼𝐼𝑄𝐼𝐼𝐼2𝑡\displaystyle{\rho_{yx}(\text{d}H>0,\text{d}H<0})=\frac{{R_{meas}(QIV,QI)}-{R_% {meas}(QII,QIII)}}{2}t.italic_ρ start_POSTSUBSCRIPT italic_y italic_x end_POSTSUBSCRIPT ( d italic_H > 0 , d italic_H < 0 ) = divide start_ARG italic_R start_POSTSUBSCRIPT italic_m italic_e italic_a italic_s end_POSTSUBSCRIPT ( italic_Q italic_I italic_V , italic_Q italic_I ) - italic_R start_POSTSUBSCRIPT italic_m italic_e italic_a italic_s end_POSTSUBSCRIPT ( italic_Q italic_I italic_I , italic_Q italic_I italic_I italic_I ) end_ARG start_ARG 2 end_ARG italic_t . (2)

Here, Rmeassubscript𝑅𝑚𝑒𝑎𝑠R_{meas}italic_R start_POSTSUBSCRIPT italic_m italic_e italic_a italic_s end_POSTSUBSCRIPT is the resistance of the raw measured data, while L is the length between voltage contacts, w the width of the sample, and t the sample thickness.

Supporting Information The supporting information consists of materials and method section, addition result and discussion, addition DFT results including pCOHP/bond(energy), single crystal refinement tables, supercell relaxed lattice parameters, magnetic properties, thermodynamic properties, and SEM-EDX results in the Eu5In2-xGaxSb6 series.

Acknowledgements

T.B. acknowledges postdoctoral support from NSF MRSEC through the Princeton Center for Complex Materials NSF-DMR-2011750. LMS was supported by the Gordon and Betty Moore Foundation’s EPIQS initiative through Grants GBMF9064, The Packard foundation, and the Princeton Catalysis Initiative (PCI). The authors thank Rafal Wawrzyńczak, Chris Lygouras, and Will Liag for their helpful discussions. SBL is supported by the National Science Foundation Graduate Research Fellowship Program under Grant No. DGE-2039656. AZ and DS acknowledge support from NSF SSMC award number no. 1709158. TMM acknowledges support from by the Institute for Quantum Matter, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Grant DE-SC0019331. The MPMS was funded by the National Science Foundation, Division of Materials Research, Major Research Instrumentation Program, under Award No. 1828490. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation.

Conflict of interest

The authors declare no conflict of interest.

Data availability statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • 1 C. L. Kane, E. J. Mele, Physical review letters 2005, 95, 22 226801.
  • 2 H. Zhang, C.-X. Liu, X.-L. Qi, X. Dai, Z. Fang, S.-C. Zhang, Nature physics 2009, 5, 6 438.
  • 3 Z. Liu, B. Zhou, Y. Zhang, Z. Wang, H. Weng, D. Prabhakaran, S.-K. Mo, Z. Shen, Z. Fang, X. Dai, et al., Science 2014, 343, 6173 864.
  • 4 B. Bradlyn, J. Cano, Z. Wang, M. Vergniory, C. Felser, R. J. Cava, B. A. Bernevig, Science 2016, 353, 6299 aaf5037.
  • 5 L. M. Schoop, M. N. Ali, C. Straßer, A. Topp, A. Varykhalov, D. Marchenko, V. Duppel, S. S. Parkin, B. V. Lotsch, C. R. Ast, Nature communications 2016, 7, 1 11696.
  • 6 Y. Deng, Y. Yu, M. Z. Shi, Z. Guo, Z. Xu, J. Wang, X. H. Chen, Y. Zhang, Science 2020, 367, 6480 895.
  • 7 S. X. Riberolles, T. V. Trevisan, B. Kuthanazhi, T. Heitmann, F. Ye, D. Johnston, S. Bud’ko, D. Ryan, P. Canfield, A. Kreyssig, et al., Nature communications 2021, 12, 1 999.
  • 8 N. Regnault, Y. Xu, M.-R. Li, D.-S. Ma, M. Jovanovic, A. Yazdani, S. S. Parkin, C. Felser, L. M. Schoop, N. P. Ong, et al., Nature 2022, 603, 7903 824.
  • 9 M. Vergniory, L. Elcoro, C. Felser, N. Regnault, B. A. Bernevig, Z. Wang, Nature 2019, 566, 7745 480.
  • 10 N. R. Council, et al., Frontiers in crystalline matter: from discovery to technology, National Academies Press, 2009.
  • 11 L. M. Schoop, F. Pielnhofer, B. V. Lotsch, Chemistry of Materials 2018, 30, 10 3155.
  • 12 M. Z. Hasan, C. L. Kane, Reviews of modern physics 2010, 82, 4 3045.
  • 13 R. J. Cava, H. Ji, M. K. Fuccillo, Q. D. Gibson, Y. S. Hor, Journal of Materials Chemistry C 2013, 1, 19 3176.
  • 14 Y. Chen, J. G. Analytis, J.-H. Chu, Z. Liu, S.-K. Mo, X.-L. Qi, H. Zhang, D. Lu, X. Dai, Z. Fang, et al., science 2009, 325, 5937 178.
  • 15 M. Konig, S. Wiedmann, C. Brune, A. Roth, H. Buhmann, L. W. Molenkamp, X.-L. Qi, S.-C. Zhang, Science 2007, 318, 5851 766.
  • 16 B. A. Bernevig, T. L. Hughes, S.-C. Zhang, science 2006, 314, 5806 1757.
  • 17 J. C. Teo, L. Fu, C. Kane, Physical Review B 2008, 78, 4 045426.
  • 18 C. Şahin, M. E. Flatté, Physical review letters 2015, 114, 10 107201.
  • 19 X.-L. Qi, S.-C. Zhang, Physics Today 2010, 63, 1 33.
  • 20 J. F. Khoury, L. M. Schoop, Trends in Chemistry 2021, 3, 9 700.
  • 21 Z. Zhu, Y. Cheng, U. Schwingenschlögl, Physical Review B 2012, 85, 23 235401.
  • 22 N. Kumar, S. N. Guin, K. Manna, C. Shekhar, C. Felser, Chemical Reviews 2020, 121, 5 2780.
  • 23 B. Yan, H.-J. Zhang, C.-X. Liu, X.-L. Qi, T. Frauenheim, S.-C. Zhang, Physical Review B 2010, 82, 16 161108.
  • 24 L. Fu, C. L. Kane, Physical Review B 2007, 76, 4 045302.
  • 25 D. Zhang, M. Shi, T. Zhu, D. Xing, H. Zhang, J. Wang, Physical review letters 2019, 122, 20 206401.
  • 26 B. Yan, M. Jansen, C. Felser, Nature Physics 2013, 9, 11 709.
  • 27 Z. Jiang, Z. Liu, H. Ma, W. Xia, Z. Liu, J. Liu, S. Cho, Y. Yang, J. Ding, J. Liu, et al., Nature Communications 2023, 14, 1 4892.
  • 28 A. Isaeva, B. Rasche, M. Ruck, Bismuth-based candidates for topological insulators: Chemistry beyond bi2te3, 2013.
  • 29 B. Yan, B. Stadtmüller, N. Haag, S. Jakobs, J. Seidel, D. Jungkenn, S. Mathias, M. Cinchetti, M. Aeschlimann, C. Felser, Nature communications 2015, 6, 1 10167.
  • 30 S. Chadov, X. Qi, J. Kübler, G. H. Fecher, C. Felser, S. C. Zhang, Nature materials 2010, 9, 7 541.
  • 31 L.-L. Wang, Physical Review Materials 2022, 6, 9 094201.
  • 32 M. Zhao, X. Zhang, L. Li, Scientific reports 2015, 5, 1 16108.
  • 33 Z.-Q. Huang, F.-C. Chuang, C.-H. Hsu, Y.-T. Liu, H.-R. Chang, H. Lin, A. Bansil, Physical Review B 2013, 88, 16 165301.
  • 34 H. Weng, X. Dai, Z. Fang, Physical review X 2014, 4, 1 011002.
  • 35 A. Honma, D. Takane, S. Souma, Y. Wang, K. Nakayama, M. Kitamura, K. Horiba, H. Kumigashira, T. Takahashi, Y. Ando, et al., Physical Review B 2023, 108, 11 115118.
  • 36 Y. Ren, X. Deng, Z. Qiao, C. Li, J. Jung, C. Zeng, Z. Zhang, Q. Niu, Physical Review B 2015, 91, 24 245415.
  • 37 B. J. Kooi, M. Wuttig, Advanced materials 2020, 32, 21 1908302.
  • 38 G. Cuono, R. M. Sattigeri, C. Autieri, T. Dietl, Physical Review B 2023, 108, 7 075150.
  • 39 B. Yan, L. Müchler, X.-L. Qi, S.-C. Zhang, C. Felser, Physical Review B 2012, 85, 16 165125.
  • 40 C. Cao, G.-X. Zhi, J.-X. Zhu, Physical Review Letters 2020, 124, 16 166403.
  • 41 W. Fuhrman, J. Leiner, P. Nikolić, G. E. Granroth, M. B. Stone, M. D. Lumsden, L. DeBeer-Schmitt, P. A. Alekseev, J.-M. Mignot, S. Koohpayeh, et al., Physical review letters 2015, 114, 3 036401.
  • 42 L. Müchler, F. Casper, B. Yan, S. Chadov, C. Felser, Topological insulators and thermoelectric materials, 2013.
  • 43 F. Schindler, A. M. Cook, M. G. Vergniory, Z. Wang, S. S. Parkin, B. A. Bernevig, T. Neupert, Science advances 2018, 4, 6 eaat0346.
  • 44 Y. Tokura, K. Yasuda, A. Tsukazaki, Nature Reviews Physics 2019, 1, 2 126.
  • 45 C.-Z. Chang, J. Zhang, X. Feng, J. Shen, Z. Zhang, M. Guo, K. Li, Y. Ou, P. Wei, L.-L. Wang, et al., Science 2013, 340, 6129 167.
  • 46 J. Li, Y. Li, S. Du, Z. Wang, B.-L. Gu, S.-C. Zhang, K. He, W. Duan, Y. Xu, Science Advances 2019, 5, 6 eaaw5685.
  • 47 J.-Q. Yan, Q. Zhang, T. Heitmann, Z. Huang, K. Chen, J.-G. Cheng, W. Wu, D. Vaknin, B. C. Sales, R. J. McQueeney, Physical Review Materials 2019, 3, 6 064202.
  • 48 Y.-J. Hao, P. Liu, Y. Feng, X.-M. Ma, E. F. Schwier, M. Arita, S. Kumar, C. Hu, R. Lu, M. Zeng, et al., Physical Review X 2019, 9, 4 041038.
  • 49 L. Wu, M. Salehi, N. Koirala, J. Moon, S. Oh, N. Armitage, Science 2016, 354, 6316 1124.
  • 50 Y. Xu, Z. Song, Z. Wang, H. Weng, X. Dai, Physical review letters 2019, 122, 25 256402.
  • 51 Y. Zhao, H. Bae, Y. Jiang, Y. Li, C.-X. Liu, B. Yan, arXiv preprint arXiv:2403.06304 2024.
  • 52 J. Ma, H. Wang, S. Nie, C. Yi, Y. Xu, H. Li, J. Jandke, W. Wulfhekel, Y. Huang, D. West, et al., Advanced Materials 2020, 32, 14 1907565.
  • 53 N. Varnava, T. Berry, T. M. McQueen, D. Vanderbilt, Physical Review B 2022, 105, 23 235128.
  • 54 P. Rosa, Y. Xu, M. Rahn, J. Souza, S. Kushwaha, L. Veiga, A. Bombardi, S. Thomas, M. Janoschek, E. Bauer, et al., npj Quantum Materials 2020, 5, 1 52.
  • 55 V. C. Morano, J. Gaudet, N. Varnava, T. Berry, T. Halloran, C. J. Lygouras, X. Wang, C. M. Hoffman, G. Xu, J. W. Lynn, et al., Physical Review B 2024, 109, 1 014432.
  • 56 A. Zevalkink, G. S. Pomrehn, S. Johnson, J. Swallow, Z. M. Gibbs, G. J. Snyder, Chemistry of Materials 2012, 24, 11 2091.
  • 57 S. Chanakian, U. Aydemir, A. Zevalkink, Z. M. Gibbs, J.-P. Fleurial, S. Bux, G. J. Snyder, Journal of Materials Chemistry C 2015, 3, 40 10518.
  • 58 M. Zeer, D. Go, P. Schmitz, T. G. Saunderson, H. Wang, J. Ghabboun, S. Blügel, W. Wulfhekel, Y. Mokrousov, Physical Review Research 2024, 6, 1 013095.
  • 59 T. Berry, S. R. Parkin, T. M. McQueen, Physical Review Materials 2021, 5, 11 114401.
  • 60 R. Shannon, C. Prewitt, Journal of Inorganic and Nuclear Chemistry 1968, 30, 6 1389.
  • 61 R. Shannon, C. Prewitt, Journal of Inorganic and Nuclear Chemistry 1970, 32, 5 1427.
  • 62 W. Lv, C. Yang, J. Lin, X. Hu, K. Guo, X. Yang, J. Luo, J.-T. Zhao, Journal of Alloys and Compounds 2017, 726 618.
  • 63 S.-M. Park, E. S. Choi, W. Kang, S.-J. Kim, Journal of Materials Chemistry 2002, 12, 6 1839.
  • 64 S. Kushwaha, I. Pletikosić, T. Liang, A. Gyenis, S. Lapidus, Y. Tian, H. Zhao, K. Burch, J. Lin, W. Wang, et al., Nature communications 2016, 7, 1 11456.
  • 65 N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, N. P. Ong, Reviews of modern physics 2010, 82, 2 1539.
  • 66 X.-J. Liu, X. Liu, J. Sinova, Physical Review B 2011, 84, 16 165304.
  • 67 P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, et al., Journal of physics: Condensed matter 2009, 21, 39 395502.
  • 68 P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, et al., Journal of physics: Condensed matter 2017, 29, 46 465901.
  • 69 P. Giannozzi, O. Baseggio, P. Bonfà, D. Brunato, R. Car, I. Carnimeo, C. Cavazzoni, S. De Gironcoli, P. Delugas, F. Ferrari Ruffino, et al., The Journal of chemical physics 2020, 152, 15.
  • 70 N. Marzari, D. Vanderbilt, A. De Vita, M. Payne, Physical review letters 1999, 82, 16 3296.
  • 71 S. Maintz, V. L. Deringer, A. L. Tchougréeff, R. Dronskowski, Journal of computational chemistry 2013, 34, 29 2557.
  • 72 S. Maintz, V. L. Deringer, A. L. Tchougréeff, R. Dronskowski, Lobster: A tool to extract chemical bonding from plane-wave based dft, 2016.

Supplementary Information

Page Number
Result and Discussion of Magnetic and Thermodynamic Properties S-3 to S-5
Table S1: Eu5In2-xGaxSb6: 3c supercell relaxed lattice parameters S-6
Table S2: The parity outputs at the k points using MBJ in Eu5In2-xGaxSb6 S-7
Figure S1: DFT plots using GGA+SOC for the Eu5In2-xGaxSb6 S-8
Figure S2: DFT plots using MBJ for the Eu5In2-xGaxSb6 S-9
Figure S3: The pCOHP/bond as a function of energy that compares dimers S-10
Table S3: Crystal data and structure refinement for Eu5In1.37Ga0.63Sb6 S-11
Table S4: Displacement parameters for Eu5In1.37Ga0.63Sb6 structure. S-12
Table S5: Anisotropic displacement parameters for Eu5In1.37Ga0.63Sb6 S-13
Figure S4: Magnetism in Eu5In1.37Ga0.63Sb6 S-14
Figure S5: Heat Capacity in Eu5In1.37Ga0.63Sb6 S-15
Figure S6: Shift structural in Eu5In1.37Ga0.63Sb6 S-16
Figure S7: SEM-EDS maps and spectrum of Eu5In1.586Ga0.429Sb6 S-17
Figure S8: SEM-EDS maps and spectrum of Eu5In1.495Ga0.637Sb6 S-18
Figure S9: SEM-EDS maps and spectrum of Eu5In1.378Ga0.663Sb6 S-19
Figure S10: The DFT Supercell positions of x=2 in Eu5In2-xGaxSb6. S-20
Figure S11: The DFT Supercell positions of x=4/3 in Eu5In2-xGaxSb6. S-21
Figure S12: The DFT Supercell positions of x=4/3 in Eu5In2-xGaxSb6. S-22
Figure S13: The DFT Supercell positions of x=0 in Eu5In2-xGaxSb6. S-23
Figure S14: The DFT Supercell structures for x=0, 2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, 4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and 2 in Eu5In2-xGaxSb6. S-24
References S-25

1 Supplementary Results and Discussion

1.1 Magnetic Properties

To examine the direction-dependent magnetic anisotropy of Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, the magnetization as a function of field and temperature was studied. In concurrence with the heat capacity measurement, the TN1 and TN2 are the same i.e. TN1= 15.029 K, TN2= 5.720 K at μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H=0.1 T in μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H//a, b, and c directions. In addition to the two transitions, there is another weak transition that occurs below TN2, i.e. TN3= 3.483 K.

Notably, there is a difference in the magnitude of magnetization as a function of temperature in each of the three directions, as seen in Figure S4 (a)-(c). The b axis has the lowest magnetization compared to the a and c axis. Metamagnetism in Figure S4 (d)-(e) indicates AFM order polarized within each ab plane. The c-axis is the hard magnetic axis with no component of the AFM order parameter polarized along the c-plane. Therefore, from the perspective of local spins the b-axis is considered the easy axis in antiferromagnets and hard axis in ferromagnets.1 The non-saturating moment of Eu2+ state is indicative of strong J coupling between the Eu2+ cations and the requirement of stronger applied field than 7 T to saturate the Eu2+ magnetic moment.

The metamagnetism in μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H//a and b is due to a spin flip transition that occurs in the ab plane. In general metamagnetism is seen in most Gd3+ and Eu2+ compounds because of high single-ion isotropy due to the isotropic nature of a 4f7 (S = 7/2) state.2 In Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, the magnetic anisotropy dominates the polarizability of the spin in the a and b direction compared to the c direction, therefore in agreement with our previous claim on c axis being the hard axis.

Overlaying the transitions coming from M(T) and M(μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H) in a, b, and c directions, we can construct the magnetic phase diagrams in Figure S4 (g-i). We gather that the overall magnetism in Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG is antiferromagnetic and the spins lie in the ab plane. Thus, there are four distinct magnetic phases in the a and b axis, where two magnetic phases in the c direction respectively. The two magnetic phases in the c direction correspond to the two magnetic phases corresponding to TN1 and TN2. These phases per the single crystal neutron diffraction correspond to the k=(0,0,0) and k=(0,0,½), where at TN1 the unit cell expands.3 The overall magnetic structure is proposed to be similar to the parent Eu5In2-xGaxSb6 with non-collinear magnetism at T=2K, since the Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG has the same qualitative features in magnetism. 3

1.2 Thermodynamic Properties

The heat capacity of in Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG shows two transitions at from TN1= 5.529(6) K, TN2=14.415(2)K as seen in Figure S5(a). Previous work on the parent Eu5In2Sb6 structure attributed these transitions to be second-order phase transitions due to the antiferromagnetic order of Eu.4 In Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, these transitions are qualitatively similar to the parent. TN1 gets suppressed on applying a magnetic field of μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H=9T, however, the TN2 requires a μoH>𝜇oHabsent\mu\textsubscript{o}\textit{H}>italic_μ italic_H > 9 T to suppress the magnetic order in Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG as seen in Figure S5(c).

In the parent version, the magnetic order for TN2 is in agreement with the magnetization as a function of field behavior in Eu5In2Sb6, where a μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H=30T saturates the Eu2+ S=J=7/2 in the c direction.4 The ΔΔ\Deltaroman_ΔTN2 in the two cases are similar, suggesting that μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H→30 T may be expected to saturate Eu2+ spins. Requiring a large expected saturation magnetic field in Eu2+ spins is indicative of strong J coupling. Typically mostly Eu2+ Zintl antiferromagnetic compounds have far weaker J coupling i.e. a smaller μoH𝜇oH\mu\textsubscript{o}\textit{H}italic_μ italic_H is sufficient to saturate the magnetic order. 5 However, recent neutron studies of Eu5In2Sb6 reveal that at T=2K the Eu2+ spins become non-collinear giving rise to magnetic complexities and contribute to stronger J coupling within the Eu sublattice.3 However, recent neutron studies of Eu5In2Sb6 reveal that at T=2K the Eu2+ spins become non-collinear giving rise to magnetic complexities and contributing to stronger J coupling within the Eu sublattice. There are different Eu sublattices in Eu5In2Sb6 and have been reported to have a non-collinear magnetic structure via neutron studies. In Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, we observe a contraction of bond distances between Ga and Eu as seen in Figure 3. There could be a possibility that the more conductive network partly due to Eu-Ga bonds along with the non-collinear structure could contribute to the topology in the more Ga-rich compositions.

To confirm the S=J=7/2 Eu2+ spin state in Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, an analytical model for the phonon-specific heat capacity in 2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARGCa5Ga2Sb6 + 1313\frac{1}{3}divide start_ARG 1 end_ARG start_ARG 3 end_ARGCa5In2Sb6 was constructed and the mass difference between Eu and Ca were adjusted using the equation below.

ΘLmYsOtZr3ΘWmXnYoZp3=mML32+sMY32+tMO32+rMZ32mMW32+nMX32+oMY32+pMZ32subscriptsuperscriptΘ3subscript𝐿𝑚subscript𝑌𝑠subscript𝑂𝑡subscript𝑍𝑟subscriptsuperscriptΘ3subscript𝑊𝑚subscript𝑋𝑛subscript𝑌𝑜subscript𝑍𝑝𝑚subscriptsuperscript𝑀32𝐿𝑠subscriptsuperscript𝑀32𝑌𝑡subscriptsuperscript𝑀32𝑂𝑟subscriptsuperscript𝑀32𝑍𝑚subscriptsuperscript𝑀32𝑊𝑛subscriptsuperscript𝑀32𝑋𝑜subscriptsuperscript𝑀32𝑌𝑝subscriptsuperscript𝑀32𝑍\displaystyle{\frac{\Theta^{3}_{L_{m}Y_{s}O_{t}Z_{r}}}{{\Theta^{3}_{W_{m}X_{n}% Y_{o}Z_{p}}}}}=\frac{mM^{\frac{3}{2}}_{L}+sM^{\frac{3}{2}}_{Y}+tM^{\frac{3}{2}% }_{O}+rM^{\frac{3}{2}}_{Z}}{mM^{\frac{3}{2}}_{W}+nM^{\frac{3}{2}}_{X}+oM^{% \frac{3}{2}}_{Y}+pM^{\frac{3}{2}}_{Z}}divide start_ARG roman_Θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_O start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG roman_Θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG = divide start_ARG italic_m italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT + italic_s italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT + italic_t italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_O end_POSTSUBSCRIPT + italic_r italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT end_ARG start_ARG italic_m italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT + italic_n italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT + italic_o italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT + italic_p italic_M start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT end_ARG (1)

where ΘLmYsOtZr3subscriptsuperscriptΘ3subscript𝐿𝑚subscript𝑌𝑠subscript𝑂𝑡subscript𝑍𝑟\Theta^{3}_{L_{m}Y_{s}O_{t}Z_{r}}roman_Θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_O start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT and ΘWmXnYoZp3subscriptsuperscriptΘ3subscript𝑊𝑚subscript𝑋𝑛subscript𝑌𝑜subscript𝑍𝑝\Theta^{3}_{W_{m}X_{n}Y_{o}Z_{p}}roman_Θ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT are the normalization factor of x=2/3 Eu5In2-xGaxSb6 and (2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARGCa5Ga2Sb6 +1313\frac{1}{3}divide start_ARG 1 end_ARG start_ARG 3 end_ARGCa5In2Sb6), and are the molar masses of Eu5In1.32Ga0.67Sb6 and (2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARGCa5Ga2Sb6 + 1313\frac{1}{3}divide start_ARG 1 end_ARG start_ARG 3 end_ARGCa5In2Sb6) as 5Eu+1.32In+0.67Ga+6Sb and [2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG(5Ca+2In+6Sb)+ 1313\frac{1}{3}divide start_ARG 1 end_ARG start_ARG 3 end_ARG(5Ca+2Ga+6Sb)]. The normalization factor outputted is 1.155 respectively.6,7

The non-magnetic analogs, Ca5In2Sb6 and Ca5In2Sb6, utilized for the phonon subtraction in the Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG were in the polycrystalline form. In Figure S5(b) we see that the CP/T plot where the constructed model of Ca analogs have the same heat capacity at Tsimilar-to\sim125 K suggesting that the Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG is magnetic beyond its magnetic ordering temperature. At T\geq125 K the two heat capacities overlap and follow the Dulong Petit law. On subtracting the phonons using the Ca analog, the entropy recovered between T=2-300 K saturates close to the 5 Rln(8)=86 J.mol-1K-2 expected for ordering of Eu2+ (S= 7/2) ions. The complete saturation of S=J=7/2 spin for Eu2+ spin state confirms no presence of mix-valency due to the presence of Eu3+as well as no splitting in the crystal field levels. 8

Table 1: Eu5In2-xGaxSb6: 3c supercell relaxed lattice parameters.
x ratios a (Å) b (Å) c (Å)
x=0 11.6326 15.0771 13.0537 (4.3512)
x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG 11.6059 15.0158 12.9895 (4.3298)
x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG 11.587 14.9567 12.9058 (4.3019)
x=2 11.574 14.9064 12.8118 (4.2706)
Table 2: The parity outputs at the k points using MBJ+SOC in Eu5In2-xGaxSb6.
x ratios ΓΓ\Gammaroman_Γ Y𝑌Yitalic_Y X𝑋Xitalic_X Z𝑍Zitalic_Z U𝑈Uitalic_U T𝑇Titalic_T S𝑆Sitalic_S R𝑅Ritalic_R
x=0 1 1 1 1 1 1 1 1
x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG -1 1 1 -1 -1 1 1 -1
x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG 1 1 1 1 1 1 1 1
x=2 1 -1 1 -1 -1 -1 1 -1
Refer to caption
Figure 1: Density Functional Theory (DFT) plots using GGA+SOC for the Eu5In2-xGaxSb6 structure for (a) x=0, (b) x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, (c) x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and (d) x=2 respectively. The intensity of the dots indicates the fractional contribution of that band to the average structure.
Refer to caption
Figure 2: Density Functional Theory (DFT) plots using MBJ for the Eu5In2-xGaxSb6 structure for (a) x=0, (b) x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, (c) x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and (d) x=2 respectively. The intensity of the dots indicates the fractional contribution of that band to the average structure.
Refer to caption
Figure 3: The pCOHP/bond as a function of energy that compares the In-Sb dimer, Eu-Sb no dimer, Eu-Sb dimer, and In-Sb no dimer in (a) Eu5In2Sb6, (b) Eu5In2/3Ga1/3Sb6, (c) Eu5In1/3Ga1/3Sb6, and (d) Eu5Ga2Sb6. x=𝑥absentx=italic_x =2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6 respectively.
Table 3: Crystal data and structure refinement for Eu5In1.37Ga0.63Sb6.
Empirical formula Eu5In1.37Ga0.63Sb6
Crystal system Orthorhombic
Space group Pbam (No. 55)
Formula weight 1697.5
a (Å) 14.5200(5)
b (Å) 12.5140(4)
c (Å) 4.5996 (1)
Volume (Å3) 835.76 (0)
Z 2
Temperature (K) 163 (2)
Absorption coefficient (mm-1) 30.742
Mo Kα𝛼\alphaitalic_α (Å) 0.71073
Reflections collected / Number of parameters 1363/45
Goodness of fit 1.082
R[F]a 0.0263
Rw(Fo2subscriptsuperscriptabsent2𝑜{}^{2}_{o}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT)b 0.0345
Ra=Σ||Fo||FC||Σ|Fo|superscript𝑅𝑎Σsubscript𝐹𝑜subscript𝐹𝐶Σsubscript𝐹𝑜{}^{a}R=\frac{\Sigma||F_{o}|-|F_{C}||}{\Sigma|F_{o}|}start_FLOATSUPERSCRIPT italic_a end_FLOATSUPERSCRIPT italic_R = divide start_ARG roman_Σ | | italic_F start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT | - | italic_F start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT | | end_ARG start_ARG roman_Σ | italic_F start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT | end_ARG
b Rw(Fo2subscriptsuperscriptabsent2𝑜{}^{2}_{o}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT)=[Σ[w(|Fo|2|Fc|2)2]Σ[w(|Fo|4)]]1/2absentsuperscriptdelimited-[]Σdelimited-[]𝑤superscriptsuperscriptsubscript𝐹𝑜2superscriptsubscript𝐹𝑐22Σdelimited-[]𝑤superscriptsubscript𝐹𝑜412=[\frac{\Sigma[w(|F_{o}|^{2}-|F_{c}|^{2})^{2}]}{\Sigma[w(|F_{o}|^{4})]}]^{1/2}= [ divide start_ARG roman_Σ [ italic_w ( | italic_F start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - | italic_F start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG start_ARG roman_Σ [ italic_w ( | italic_F start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) ] end_ARG ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT
Table 4: Fractional atomic coordinates and isotropic displacement parameters based on the refined Eu5In1.37Ga0.63Sb6 structure.
Element Wyckoff Positions x y z Occupancy Uiso
Eu1 4h 0.51889(3) 0.328445(4) 0.500000 1 0.011
Eu2 2d 0.50000 0.50000 0.50000 1 0.010
Eu3 4h 0.74940(3) 0.08761(4) 0.50000 1 0.010
In1 4g 0.71442(5) 0.32944(6) 1.0000 0.683(11) 0.013
Ga1 4g 0.71442(5) 0.32944(6) 1.0000 0.317(11) 0.013
Sb1 4g 0.59409(4) 0.15429(5) 1.0000 1 0.009
Sb2 4g 0.90374(4) 0.02325(5) 1.0000 1 0.011
Sb3 4h 0.82415(5) 0.33538(5) 0.5000 1 0.015
Table 5: Anisotropic displacement parameters for Eu5In1.37Ga0.63Sb6 determined by SXRD.
Element U(1,1) U(2,2) U(3,3) U(1,2) U(1,3) U(2,3)
Eu1 0.0135(2) 0.0092(2) 0.0096(2) 0.000 0.000 0.00246(18)
Eu2 0.0104(3) 0.0081(3) 0.0120(3) 0 0 -0.0010(2)
Eu3 0.0088(2) 0.0092(2) 0.0119(2) 0 0 -0.00050(17)
In1 0.0115(3) 0.0094(3) 0.0073(3) 0 0 -0.0016(2)
Ga1 0.0088(3) 0.0126(3) 0.0106(3) 0 0 0.0009(2)
Sb1 0.0127(3) 0.0075(3) 0.0244(4) 0 0 -0.0013(2)
Sb2 0.0126(4) 0.0121(4) 0.0128(4) 0 0 -0.0003(3)
Sb3 0.0126(4) 0.0121(4) 0.0128(4) 0 0 -0.0003(3)
Refer to caption
Figure 4: Magnetization as a function of temperature with μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H = 0.1-7 T and T = 2-300 K in 2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6 single crystals at various applied magnetic fields at (a) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//a, (b) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//b, and (c) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//c. There are two transitions at T=7.2K and T=14.4 K. Magnetization as a function of the magnetic field from μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H=-7 to 7 T for (c) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//a, (d) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//b, and (e) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//c. There is metamagnetism observed in μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//a and μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//b. Magnetic phase diagram (g) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//a, (h) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//b, and (i) μoH𝜇𝑜H\mu{o}\textit{H}italic_μ italic_o H//c.
Refer to caption
Figure 5: Heat capacity as a function of temperature for Eu5In2-xGaxSb6 with x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG single crystals at various applied magnetic fields. There are two transitions at T𝑇Titalic_T=7.2K and T𝑇Titalic_T=14.4 K. (b) The Cp/Tsubscript𝐶𝑝𝑇C_{p}/Titalic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_T plots for plot (a). (c) The phonon subtraction using the experimentally measured Ca5In2Sb6 and Ca5Ga2Sb6. The upturn at around T𝑇Titalic_Tsimilar-to\sim280 K is because of Apezon N grease. (d) The integrated change in magnetic entropy for x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6 single crystals that saturates at the theoretical 5 R𝑅Ritalic_Rln(8) for a J=S𝐽𝑆J=Sitalic_J = italic_S=7/2 system.
Refer to caption
Figure 6: Shift structural shifts from (a) Eu5In2Sb6 to (b) x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG Eu5In2-xGaxSb6. The tables below (a) and (b) explain the distortion in the trigonal pyramidal to a bent geometry in the (InSb4) frameworks and the overall shortening of the bond distances on Ga substitution with more significant figures.
Refer to caption
Figure 7:

SEM-EDS maps and spectrum of Sample-2 with the composition of Eu5In1.586Ga0.429Sb6.

Refer to caption
Figure 8:

SEM-EDS maps and spectrum of Sample-1 with the composition of Eu5In1.495Ga0.637Sb6.

Refer to caption
Figure 9:

SEM-EDS maps and spectrum of Sample-3 with the composition of Eu5In1.378Ga0.663Sb6.

Refer to caption
Figure 10: The DFT Supercell positions of x=2 in Eu5In2-xGaxSb6.
Refer to caption
Figure 11: The DFT Supercell positions of x=4/3 in Eu5In2-xGaxSb6.
Refer to caption
Figure 12: The DFT Supercell positions of x=2/3 in Eu5In2-xGaxSb6.
Refer to caption
Figure 13: The DFT Supercell positions of x=0 in Eu5In2-xGaxSb6.
Refer to caption
Figure 14: The DFT Supercell structures for (a) x=0, (b) x=2323\frac{2}{3}divide start_ARG 2 end_ARG start_ARG 3 end_ARG, (c) x=4343\frac{4}{3}divide start_ARG 4 end_ARG start_ARG 3 end_ARG, and (d) x=2 in Eu5In2-xGaxSb6 respectively.

References

  • Yosida 1996 Yosida, K. Theory of magnetism.: Edition en anglais; Springer Science & Business Media, 1996; Vol. 122
  • Holmes and Schieber 1966 Holmes, L.; Schieber, M. Magnetic ordering in Eu3O4 and EuGd2O4. Journal of Applied Physics 1966, 37, 968–969
  • 3 V. C. Morano, J. Gaudet, N. Varnava, T. Berry, T. Halloran, C. J. Lygouras, X. Wang, C. M. Hoffman, G. Xu, J. W. Lynn, et al., Physical Review B 2024, 109, 1 014432.
  • 4 P. Rosa, Y. Xu, M. Rahn, J. Souza, S. Kushwaha, L. Veiga, A. Bombardi, S. Thomas, M. Janoschek, E. Bauer, et al., npj Quantum Materials 2020, 5, 1 52.
  • Berry \latinet al. 2022 Berry, T.; Stewart, V. J.; Redemann, B. W.; Lygouras, C.; Varnava, N.; Vanderbilt, D.; McQueen, T. M. A-type antiferromagnetic order in the Zintl-phase insulator EuZn 2 P 2. Physical Review B 2022, 106, 054420
  • Hofmann \latinet al. 1956 Hofmann, J.; Paskin, A.; Tauer, K.; Weiss, R. Analysis of ferromagnetic and antiferro-magnetic second-order transitions. Journal of physics and chemistry of solids 1956, 1, 45–60
  • Tari 2003 Tari, A. The specific heat of matter at low temperatures; World Scientific, 2003
  • Fulde and Loewenhaupt 1985 Fulde, P.; Loewenhaupt, M. Magnetic excitations in crystal-field split 4f systems. Advances in Physics 1985, 34, 589–661