This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Three-dimensional nature of anomalous Hall conductivity in YMn6Sn6-xGax, x0.55x\approx 0.55

Hari Bhandari [email protected] Department of Physics and Astronomy, University of Notre Dame, Notre Dame, IN 46556, USA Stravropoulos Center for Complex Quantum Matter, University of Notre Dame, Notre Dame, IN 46556, USA Department of Physics and Astronomy, George Mason University, Fairfax, VA 22030, USA    Zhenhua Ning Ames National Laboratory, U.S. Department of Energy, Ames, Iowa 50011, USA    Po-Hao Chang Department of Physics and Astronomy, George Mason University, Fairfax, VA 22030, USA Quantum Science and Engineering Center, George Mason University, Fairfax, VA 22030, USA    Peter E. Siegfried Department of Physics and Astronomy, George Mason University, Fairfax, VA 22030, USA Quantum Science and Engineering Center, George Mason University, Fairfax, VA 22030, USA    Resham B. Regmi Department of Physics and Astronomy, University of Notre Dame, Notre Dame, IN 46556, USA Stravropoulos Center for Complex Quantum Matter, University of Notre Dame, Notre Dame, IN 46556, USA    Mohamed El. Gazzah Department of Physics and Astronomy, University of Notre Dame, Notre Dame, IN 46556, USA Stravropoulos Center for Complex Quantum Matter, University of Notre Dame, Notre Dame, IN 46556, USA    Albert V. Davydov Materials Science and Engineering Division, National Institute of Standards and Technology (NIST), Gaithersburg, Maryland 20899, USA    Allen G. Oliver Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556, USA    Liqin Ke Ames National Laboratory, U.S. Department of Energy, Ames, Iowa 50011, USA    Igor I. Mazin Department of Physics and Astronomy, George Mason University, Fairfax, VA 22030, USA Quantum Science and Engineering Center, George Mason University, Fairfax, VA 22030, USA    Nirmal J. Ghimire [email protected] Department of Physics and Astronomy, University of Notre Dame, Notre Dame, IN 46556, USA Stravropoulos Center for Complex Quantum Matter, University of Notre Dame, Notre Dame, IN 46556, USA
Abstract

The unique connectivity of kagome lattices gives rise to topological properties, such as flat bands and Dirac cones. When combined with ferromagnetism and a chemical potential near the 2D Dirac points, this structure offers the potential to realize the highly sought-after topological Chern magnetotransport. Recently, there was considerable excitement surrounding this possibility in the ferrimagnetic kagome metal TbMn6Sn6. However, density functional theory (DFT) calculations reveal that the 2D Chern gap lies well above the Fermi energy, challenging its relevance in the observed anomalous Hall conductivity. Here, we investigate YMn6Sn5.45Ga0.55, a compound with similar crystallographic, magnetic, and electronic properties to TbMn6Sn6. Our findings show that the intrinsic anomalous Hall conductivity in this material, while comparable in magnitude to that in TbMn6Sn6, is fully three-dimensional, thus providing experimental evidence that Hall conductivity in this class of materials does not originate from 2D Chern gaps. Additionally, we confirm that the newly proposed empirical scaling relation for extrinsic Hall conductivity is universally governed by spin fluctuations.

I Introduction

Kagome lattice magnets have attracted significant interest in condensed matter physics due to their high frustration in the case of antiferromagnetic interactions. Over the past decade, this interest has grown, as it has been shown that even unfrustrated ferromagnetic (or nonmagnetic) kagome planes can exhibit nontrivial electronic features, such as flat bands and Dirac cones [1, 2, 3, 4, 5, 6, 7]. Recently, particular attention has been directed toward the so-called 166 family, the RRMn6Sn6 compounds, where RR represents a rare-earth element [8, 9, 10, 11, 12, 13, 14, 15, 16]. In these compounds, Mn atoms form kagome planes, and the crystal structure provides a diverse material space for manipulating both electronic and magnetic properties (Fig. 1). Compounds with non-magnetic RR atoms are simpler because magnetism arises solely from the Mn sublattice; however, they are also more complex due to frustrated interplanar magnetic interactions. As shown in Fig. 1, although all Mn planes are crystallographically equivalent, the exchange interactions are not; there are two distinct exchange paths: one (J1J_{1}) across the Sn layer and the other (J2J_{2}) across the RRSn layer. In YMn6Sn6 (Y166), a RR166 compound with non-magnetic RR atom, J2>0J_{2}>0 (i.e., antiferromagnetic), while the dominant interaction is J1<0J_{1}<0. As in most metals, the exchange coupling decays relatively slowly with distance (roughly as 1/d3d^{3}), so the minimal model Hamiltonian includes J1J_{1}, J2J_{2}, and J3J_{3}. Notably, this Hamiltonian is frustrated if J2J3<0J_{2}J_{3}<0, resulting in intriguing spin-spiral orders that exhibit phenomena such as the topological Hall effect [17, 18] and Lifshitz transitions [19].

Refer to caption
Figure 1: Crystal structure. Sketch of the crystal structure of YMn6Sn5.45Ga0.55. JiJ_{i}s (i=1,2,3)i=1,2,3) represent the exchange constants and JpJ_{p} represents the in-plane exchange constant.

Replacing Y with a magnetic ion that strongly couples to the neighboring Mn planes introduces an additional indirect ferromagnetic coupling between these planes, which can overcome the direct antiferromagnetic J2J_{2} and remove magnetic frustration. This occurs in Tb166 [15, 20, 21, 22, 23, 24, 25] and is now well understood [16, 12]. Additionally, due to well-understood reasons [16], Tb has an easy axis anisotropy along the crystallographic cc direction, making the material a collinear easy-axis ferrimagnet at low temperatures (in contrast to Y166, which is an easy-plane spiral antiferromagnet). Because of these specific characteristics, Tb166 recently garnered significant interest for various different properties [20, 12, 22, 24, 23, 26] including a putative two-dimensional (2D) Chern gap around 130 meV above the Fermi energy, as inferred from scanning tunneling measurements [20]. One of the consequences of a 2D Chern gap, when close to the Fermi energy, is large anomalous Hall effect. The intrinsic contribution to the anomalous Hall effect extracted from the scaling of the anomalous Hall resistivity matches the value expected from the Chern gap around 130 meV above the Fermi energy [20]. However, density functional calculations [12, 27, 16] consistently have shown that the 2D-Chern gap in Tb166 lies about 700 meV above the Fermi energy and does not contribute to the anomalous Hall conductivity (AHC). Instead, this AHC arises from several different regions in the Brillouin zone in ferrimagnetic Tb166 [12]. Additionally, an empirical scaling relation in the latter study suggested a potential contribution of spin fluctuations to the extrinsic anomalous Hall conductivity, an effect not previously investigated.

Regarding the latter, the standard method based on the theory by Crépieux and Bruno [28] for extracting the intrinsic (related to the Berry phases of conducting electron) and extrinsic (due to electron scattering) contributions to the AHC is to fit the temperature dependence of the AHC to a simple scaling relation,

σxy=aσxx2+c\sigma_{xy}=a\sigma^{2}_{xx}+c (1)

where c=σxyAc=\sigma^{A}_{xy} represents the intrinsic AHC, while the first term arises from defect scattering. This scaling relation has been widely used to extract the intrinsic AHC in various materials [29, 4, 20]. However, Crépieux and Bruno’s derivation did not account for scattering from thermally excited spin fluctuations, and currently a theoretical analysis of the effect of spin fluctuations on AHC is lacking.

Analyzing the Tb166 data, we found [12] that Eq. 1 poorly describes σxy\sigma_{xy} at higher temperatures, where scattering from spin fluctuations becomes increasingly important. At the same time, we found that the most fluctuating species in Tb166 is Tb, which fluctuates much stronger than Mn. We also found that the empirical formula,

σxy=aσxx2+d/σxx+c,\sigma_{xy}=a\sigma^{2}_{xx}+d/\sigma_{xx}+c, (2)

fits the experimental data exceptionally well (note that the additional contribution becomes significant when σxx\sigma_{xx} is small, i.e., at higher temperatures). We tentatively attributed this term to spin fluctuations, primarily from Tb.

A reliable protocol for extracting AHC from the experiment is crucial. The existing methods [28, 29, 30] overlook spin fluctuations, and therefore rely on low-temperature data. An equation that accurately describes the AHC across the entire temperature range is of significant practical importance.

In this study, we address two key aspects concerning Tb166: first, whether the intrinsic contribution to the AHC is linked to the 2D Chern gap, and second, whether the component in the AHC scaling relation, attributed to spin fluctuations, can be observed in another RR166 compound that does not contain Tb. This is particularly relevant as Tb was thought to be crucial for both the Chern-gap induced AHE [20] and the enhancement of spin fluctuations [12].

An ideal compound to investigate these properties would be a RR166 compound with a non-magnetic RR atom and a soft, and relatively isotropic ferromagnet. Such a compound would allow the access to the saturated ferromagnetic state in a standard laboratory setting, enabling measurements of anomalous Hall resistivity over a broad temperature range and in both in-plane and out-of-plane directions. We found that YMn6Sn6-xGax, 0.30x0.610.30\leq x\leq 0.61 meet these criteria [31] and selected one particular composition YMn6Sn6.45Ga0.55 for the study. Our results show that the intrinsic AHC in this compound is comparable to that in Tb166 for the out-of-plane magnetic field (BB), where the Chern gap is expected to contribute to the AHC. At the same time, we observed a similar AHC in the in-plane BB, where the Chern gap contribution is not expected, confirming the 3D nature of the AHC. Furthermore, we verified the new AHC scaling introduced in Ref. [12] and confirmed its spin-fluctuational origin.

II Methods

Single crystals of YMn6Sn5.45Ga0.55 were grown by using Sn as a flux by the molten flux method. Y pieces (Alfa Aesar 99.9%), Mn pieces (Alfa Aesar 99.95%), Sn shots (Alfa Aesar 99.999%) and Ga pieces (Alfa Aesar 99.9999%) were added into a 2-ml aluminum oxide crucible in molar ratio of 1:6:18:2. The crucible was then sealed in a fused silica ampule under vacuum. The sealed ampule was heated to 1150 C over 10 hours, homogenized at 1150 C for 24 hours, and then cooled down to 600 C with a rate of 5 C/h. Once the furnace reached 600 C, the molten flux was separated from the crystals by using a centrifuge. Upon opening the crucible, nice hexagonal-looking crystals up to 20 mg were obtained.

The crystal structure and the atomic composition were verified from a single crystal X-ray diffraction experiment. An arbitrary sphere of data was collected on a silver block-like crystal, having appropriate dimensions of 0.073×0.068×0.0260.073\times 0.068\times 0.026, on a Bruker D8 diffractometer equipped with a Bruker APEX-II detector using a combination of ω\omega- and ϕ\phi-scans of 0.5 degree [32]. Data were corrected for absorption and polarization effects and analyzed for space group determination [33]. The structure was solved by dual-space methods and expanded routinely [34]. The model was refined by full-matrix least-squares analysis of F2 against all reflections [35]. All non-hydrogen atoms were refined with anisotropic atomic displacement parameters. Unless otherwise noted, hydrogen atoms were included in calculated positions. Atomic displacement parameters for the hydrogen were tied to the equivalent isotropic displacement parameter of the atom to which they are bonded ( Uiso(H)=1.5Ueq(C)U_{iso}(H)=1.5U_{eq}(C) for methyl, 1.2Ueq(C)1.2U_{eq}(C) for all others.

DC magnetization, resistivity, and magnetoresistance measurements were performed in a Quantum Design Dynacool Physical Property Measurement System (PPMS) with a 9 T magnet. ACMS II option was used in the same PPMS for DC magnetization measurements. Single crystals of YMn6Sn5.45Ga0.55 were polished to adequate dimensions for electrical transport measurements. Crystals were oriented with the [001] and [100] directions parallel to the applied field for the c-axis and ab-plane measurements. Resistivity and Hall measurements were done using the 4-probe method. Pt wires of 25 μ\mum were used for electrical contacts with contact resistances << 30 Ω\Omega. Contacts were affixed with Epotek H20E silver epoxy. An electric current of 4 mA was used for the electrical transport measurements. Contact misalignment in the Hall resistivity measurement was corrected by anti-symmetrizing the measured data in positive and negative magnetic fields.

The first-principles calculations were performed using Vienna ab initio Simulation Package (VASP) [36] within projector augmented wave (PAW) method [37, 38] The Perdew-Burke-Enzerhof (PBE) [39] generalized gradient approximation was employed to describe exchange-correlation effects. The on-site Coulomb interactions are taken into account using LDA+U [40] to improve the description of the interactions between localized 3d3d-electrons of Mn and an effective Ueff=UJ=0U_{eff}=U-J=0, 0.60.6 and 22 are considered.

The experimental values were used for the lattice parameters and kept fixed for all the calculations, including the geometry optimization, where only internal coordinates were relaxed. To properly determine the structure, we performed geometry optimization for 2d2d, 2e2e, and 2c2c three different Ga substitution sites. The 2c site, consistent with the experimental analysis, gives the lowest energy which is then considered as the magnetic ground state.

III Results and Discussion

III.1 Crystallography

The crystal structure and atomic composition of YMn6Sn5.45Ga0.55 (Y166-Ga) were determined using single crystal X-ray diffraction. Similar to the parent compound YMn6Sn6 (Y166), Y166-Ga adopts a hexagonal P6/mmmP6/mmm structure (a=b=5.4784(a=b=5.4784 Å, c=8.925c=8.925 Å), consisting of kagome planes [Mn3Sn] separated by two inequivalent layers Sn2 and Sn3YGa, as illustrated in Fig. 1.

The structure exhibits Sn-site doping with Ga, specifically at the Sn3 site (Wyckoff position 2c). This was confirmed through modeling efforts for partial Ga occupancy at other Sn sites, which either resulted in poorer fits to the data or nonsensical Ga occupancy values. Refinement analysis determined the Sn:Ga site occupancy to be 0.725:0.175, corresponding to the full chemical formula noted above, i.e., \sim 0.55 Ga atoms per unit cell. Detailed crystallographic parameters from the single crystal X-ray diffraction experiment are summarized in Table 1.

Our DFT calculations, presented in Sec. III.3, also reveal a significant energy advantage for Ga substitution at the Sn3 site in the [Sn3Y] layer, rather than in the [Mn3Sn] or [Sn2] layers. YMn6Sn5.45Ga0.55 can thus be viewed as a moderately hole-doped (0.09\approx 0.09 h/Mn) derivative of the parent compound Y166.

Table 1: Crystallographic data, atomic coordinates and equivalent displacement parameters for YMn6Sn5.45Ga0.55.
Crystal system Hexagonal
Space group P6P6/mmm
Temperature (K) 120(2)
Wavelength (Å) 0.71073
Z formula units 1
2 θ\theta min 4.564
2 θ\theta max 61.206
Formula weight (g/mol) 1103.76
a,b (Å) 5.4784(16)
c (Å) 8.925(4)
Volume (Å)3 231.97(17)
Density (calculated) (g/cm3) 7.901
μ\mu (MoKαK_{\alpha}) mm-1 29.896
Goodness-of-fit on F2F^{2} 1.186
Final R indices [I\geq2σ\sigma (I)] R1 = 0.0172
wR2 = 0.0392
R indices (all data) R1 = 0.0183
wR2 = 0.0397
Largest diff. peak & hole, e (Å-3) 1.128 and -1.341
Atom Wyck x y z Ueq Å2{\AA}^{2}
Y 1a 1.00000 1.00000 0.00000 0.007(1)
Sn(1) 2e 1.00000 1.00000 0.33674(8) 0.004(1)
Sn(2) 2d 0.33333 0.66667 0.50000 0.006(1)
Sn(3) 2c 0.66667 0.33333 0.000000 0.005(1)
Ga(1) 2c 0.6667 0.33333 0.000000 0.005(1)
Mn(1) 6i 0.50000 0.50000 0.24504(9) 0.005(1)

III.2 Magnetic Properties

Refer to caption
Figure 2: Magnetic properties of YMn6Sn5.45Ga0.55. (a) Magnetic susceptibility measured under an external magnetic field BB of 0.1 T along the [100] and [001] directions using the field-cooled protocol. (b-c) Magnetization as a function of magnetic field B||B|| [100] (b) and B||B|| [001] (c) at selected temperatures ranging from 1.8 to 300 K.

The temperature dependence of magnetic susceptibility (χ=M/B\chi=M/B) measured with a magnetic field BB of 0.1 T parallel to [100] (χ100\chi_{100}) and along [001] (χ001\chi_{001}) is shown in Fig 2 (a). These susceptibility data indicate that Y166-Ga undergoes a paramagnetic-to-ferromagnetic ordering below 350 K, consistent with previous reports [31, 41]. Additionally, the easy-plane behavior is evident from the significantly larger χ100\chi_{100} compared to χ001\chi_{001} below the transition temperature (TcT_{c}), with an anisotropy ratio of χ100/χ001=5\chi_{100}/\chi_{001}=5 just below TcT_{c}. It is to be noted that parent compound Y166 orders with a commensurate antiferromagnetic helical structure below 345 K and exhibits an incommensurate double helical structure (DH) upon further cooling [17, 19, 42], while YMn6Sn1-xGax compounds show a ferromagnetic transition for doping concentrations x>0.30x>0.30 [31].

Figures 2 (b) and (c) show the isothermal magnetization curves of Y166-Ga at some representative temperatures for BB |||| [100] (M100M_{100}) and [001] (M001M_{001}), respectively. In the entire temperature range measured M100M_{100} saturates below 0.5 T with a negligible hysteresis, while M001M_{001} saturates at slightly larger BB. At 1.8 K M001M_{001} saturates at 2.2 T, which decreases with increasing temperature (1.7 T at 300 K), also with negligibly small hysteresis. At 1.8 K, saturation magnetization (MsatM_{sat}) along [100] is 9.85 μ\muB/f.u. while it is 10.7 μ\muB/f.u along [001]. In either direction, MsatM_{sat} gradually decreases with the increase in temperature, which attains a value of 7.5 μ\muB/f.u. along [100] and 7.7 μ\muB/f.u along [001]. The ratio of the saturated magnetization at 1.8 to 300 K (Msat,1.8KM_{sat,1.8K}/ Msat,300KM_{sat,300K}) is 1.39 along the [001] direction and 1.31 along the [100] direction. For Tb166, the Mn moment shows the ratio of \sim 1.05 and the Tb moment exhibits the ratio of around 1.66 [43, 12]. This suggests that the Mn moments in Y166-Ga experience more fluctuation between 1.8 and 300 K than in Tb166, where the Tb moments are the most fluctuating ones [12]. This observation is consistent with the expectation, as the Curie temperature of Tb166 is about 70 K higher than that of Y166-Ga.

III.3 First Principles Calculations

x=1x=1 x=0x=0 (Y166)
U 0 0.6 2 0.6
J1J_{1} -14.5 -11.9 -17.3 -12.86
J2J_{2} -21.7 -10.6 -21.9 4.66
J3J_{3} 1.8 1.5 -4.0 -2.20
Table 2: Calculated exchange couplings J1J_{1}, J2J_{2} and J3J_{3} in unit of meV. The data for Y166 are taken from Ref. [17].

To understand the doping-induced phase transition, we adopt the J1J3J_{1}-J_{3} effective model proposed in Ref. [44]. The spin Hamiltonian is expressed as follows

H=ij1J1SiSj+ij2J2SiSj+ij3J3SiSjH=\sum_{\left\langle ij\right\rangle_{1}}J_{1}S_{i}S_{j}+\sum_{\left\langle ij\right\rangle_{2}}J_{2}S_{i}S_{j}+\sum_{\left\langle ij\right\rangle_{3}}J_{3}S_{i}S_{j} (3)

where JiJ_{i} (i=13)i=1-3) are the exchange interaction parameters as indicated in Fig. 1. The parameters were extracted using the least squared fitting of our DFT calculations into Eq. 3.

According to a thorough DFT-based analysis of the phase diagram for parent Y166 discussed in our earlier work [17], we observed that Ueff=0.6U_{eff}=0.6 best reproduce the magnetic states for parent Y166 observed in the experiments. Therefore, in our study, we considered the same UU parameter with one additional larger U=2U=2 for comparison.

The results along with those of Y166, taken from Ref. [17], are summarized in TABLE 2. The fitting is achieved with excellent quality in all three cases, as all the relative energy differences between different magnetic states can be consistently and accurately reproduced by the model, especially in the case of U=0.6U=0.6 where the average error is less than 1 meV. This suggests that the minimal model adopted here is appropriate and reliable.

In TABLE 2, one can see that while J1J_{1} and J2J_{2} consistently favor ferromagnetic alignment regardless of different U values, the Hubbard U correction tends to stabilize the FM states further as J3J_{3} shifts from positive to negative as U increases. According to the analytically determined phase diagram Ref. [44] the spin model for all three U values yields the same correct FM ground state.

To assess the doping effect, we compare the results of the two compounds for the same U=0.6U=0.6 (i.e. columns 3 and 5). In Y166, J1J_{1} dominates and has the same (opposite) sign as J3J_{3} (J2J_{2}). This arrangement leads to a competition between J2J_{2} and J3J_{3} which results in the formation of a helical magnetic state [17, 44]. However, with Ga-doping, J2J_{2} becomes ferromagnetic and now comparable to J1J_{1} in strength. Although J3J_{3} becomes antiferromagnetic, it is too weak to induce frustration.

This qualitative shift aligns with expectations, considering the direct alteration of exchange pathways for J2J_{2} and J3J_{3} induced by the presence of doped-Ga. As a consequence of these changes, the frustration that was initially developed in the pure Y166 to promote the helical magnetic state is effectively mitigated. The system undergoes a transition, and the magnetic state collapses into a ferromagnetic (FM) order.

It is interesting to note that while all configurations predict the same correct ferromagnetic ground state, only U=2U=2 gives the correct easy-plane anisotropy and U=0U=0 and 0.60.6 give very small easy-axis anisotropy.

Refer to caption
Figure 3: Electrical transport properties of YMn6Sn5.45Ga0.55. (a) Electrical resistivity as a function of temperature for current applied along the [100] and [001] directions. (b) Ratio of the conductivity along the [001] direction to that along [100]. (c) Angular magnetoresistance (AMR) as a function of angle θ\theta, measured under a 9 T magnetic field, where the current (II) is along the [100] direction and θ\theta is the angle between the magnetic field and [120] direction, as illustrated by the sketch in panel (c).

III.4 Electrical Resistivity and Conductivity

The temperature dependence of electrical resistivity of Y166-Ga, measured with the electric current applied along the [100] direction (ρ100\rho_{100}, blue curve) and the [001] direction (ρ001\rho_{001}, red curve) over the temperature range 1.8-400 K, is shown in Fig. 3(a). The resistivity decreases as temperature decreases, indicating the metallic behavior of the sample. Residual resistivity ratio (RRR), calculated as ρ(400K)/ρ(2K)\rho(400K)/\rho(2K), is 9 for I||[100]I||[100] and is 18 for I||[001]I||[001]. These values are smaller than those in Y166 [17, 19], likely due to disorder induced by doping. Both ρ100\rho_{100} and ρ001\rho_{001} exhibit a kink at 350 K, indicative of the onset of a ferromagnetic transition, as observed in the susceptibility measurements [Fig. 2(a)]. Across the entire temperature range, ρ100\rho_{100} is greater than ρ001\rho_{001}. The conductivity anisotropy, σ[001]/σ[100]\sigma_{[001]}/\sigma_{[100]}, is plotted in Fig. 3(b), showing a value >> 2, which suggests that the electronic transport in Y166-Ga is three dimensional. This behavior differs from that of the parent compound Y166, where in-plane conductivity is greater than the out-of-plane conductivity [19]). The enhanced cc-axis conductivity observed in Y166-Ga is similar to that found in Ge doped YMn6Sn6 [10].

The anisotropic transport behavior was further investigated by measuring the angular-dependent magnetoresistance (AMR). In this measurement, an electric current II was applied along the [100] direction, while the sample was rotated around the magnetic field within the crystallographic bcbc-plane, with IBI\bot B held constant so that only transverse MR was measured. In this configuration, at θ=0(90)\theta=0^{\circ}(90^{\circ}), B||[120]([001])B||[120]([001]). Since the largest magnetic saturation field is below 2.5 T (see Fig. 2), the 9T magnetic field aligns the magnetic moment MM perpendicular to II at all times.

The AMR, defined as [{ρxx(θ)ρxx(θ=0)}/ρxx(θ=0)]×100\rho_{xx}(\theta)-\rho_{xx}(\theta=0)\}/\rho_{xx}(\theta=0)]\times 100, measured at 5 K is shown in fig. 3(c), with maximum value of 5.6%-5.6\%, indicating a substantial effect. To gain further insight into this large AMR, we compared the AMR with the ab initio calculated AMR using the GGA+U method. For this purpose, we used the all-electron WIEN2k package [45], varying Ueff=UJU_{eff}=U-J from 1.2 to 2 eV. Assuming an isotropic transport scattering rate, the longitudinal conductivity σxxωpl2𝐤δ(EE𝐤)v𝐤x2\sigma_{xx}\propto\omega_{pl}^{2}\propto\sum_{\mathbf{k}}{\delta(E-E_{\mathbf{k}})v_{\mathbf{k}x}^{2}}, where vv is the Fermi velocity. We calculated this quantity using the 19×19×1119\times 19\times 11 zone-centered k-point mesh and tetrahedron numerical integration, and an otherwise default setup. We found (Fig. 4) that AMR is very sensitive to correlation effects. The best agreement with the experiment occurs at Ueff=1.3U_{eff}=1.3 eV, yielding a calculated value of 1.61.6%, which is over three times smaller than the experimental result but remains the same order of magnitude. This discrepancy may indicate the presence of an anisotropic scattering rate, or an underestimation of spin-orbit coupling (SOC) in the calculation, potentially related to the known underestimation of the AHC in TbMn6Sn6 [12, 46].

Refer to caption
Figure 4: Calculated angular magnetoresistance. Angular magnetoresistance between the full polarization along the crystallographic [001] and [120] axes, as a function of the effective Hubbard interaction, calculated for YMn6Sn5Ga.

III.5 Anomalous Hall Effect

Refer to caption
Figure 5: Anomalous Hall effect of YMn6Sn5.45Ga0.55 measured with 𝐁||\mathbf{B}~{}\boldsymbol{||} [001] and 𝐈||\mathbf{I}~{}\boldsymbol{||} [100]. (a) Anomalous Hall resistivity ρyx\rho_{yx} as a function of magnetic field at selected temperatures. (b) Longitudinal conductivity ((σxx(\sigma_{xx}, red spheres), and anomalous Hall conductivity (|σyxA|\lvert\sigma^{A}_{yx}\rvert, blue spheres) as a function of temperature. (c) |σyxA|\lvert\sigma^{A}_{yx}\rvert (blue spheres) plotted as a function of σxx\sigma_{xx} in the temperature range 1.8-300 K .The solid represents a fit to Eq. 2.
Refer to caption
Figure 6: Anomalous Hall effect of YMn6Sn5.45Ga0.55 measured with 𝐁||\mathbf{B}~{}\boldsymbol{||} [120] and 𝐈||\mathbf{I}~{}\boldsymbol{||} [100]. (a) Anomalous Hall resistivity ρzx\rho_{zx} as a function of magnetic field at selected temperatures. (b) Longitudinal conductivity (σxx\sigma_{xx}, red spheres), and anomalous Hall conductivity (|σxzA|\lvert\sigma^{A}_{xz}\lvert, blue spheres) as a function of temperature. (c) |σzxA|\lvert\sigma^{A}_{zx}\rvert (blue spheres) plotted as a function of σxx\sigma_{xx} in the temperature range 1.8-300 K. The solid represents a fit to Eq. 2.

Hall resistivity measured with with B||[001]B||[001] and the current I||[100]I||[100] (ρyx\rho_{yx}) at representative temperatures is shown in Fig. 5(a). The zero-field value of ρyx\rho_{yx} corresponds to the anomalous Hall resistivity ρyxA\rho_{yx}^{A}. In Fig. 5(b), we show the longitudinal Hall conductivity, σxx\sigma_{xx} = 1/ρxx1/\rho_{xx}, and the absolute value of the anomalous Hall conductivity, σyxA\sigma_{yx}^{A} = ρyx/ρxx-\rho_{yx}/\rho_{xx} (valid when ρyx\rho_{yx} \ll ρxx2\rho_{xx}^{2}, and ρxx\rho_{xx} = ρyy\rho_{yy}), as a function of temperature between 1.8 and 300 K. This clearly indicates that σyxA\sigma_{yx}^{A} varies with temperature across the entire temperature range. The scaling of the σyxA\sigma_{yx}^{A}, using the relation presented in Eq. 2, is shown in Fig. 5(c). This scaling law effectively fits the high-temperature data, yielding the coefficients aa, dd, and cc, as presented in Table 3 and compared to those from Tb166. The coefficient aa, representing impurity scattering, is about an order of magnitude larger in Y166-Ga compared to Tb166, as expected due to increased disorder scattering in the doped sample. The intrinsic anomalous Hall conductivity of Y166-Ga (121 S/cm) is comparable to that of Tb166 (140 S/cm). Interestingly, the magnitude of dd is smaller in the Y166-Ga than in Tb166, consistent with the hypothesis that the term d/σxxd/\sigma_{xx} in Eq. 2 is due to spin fluctuations [12]. Notably, d0d\neq 0 in Y166-Ga despite the absence of highly fluctuating Tb atoms, likely because, as discussed in Section III.2, Mn in Y166-Ga fluctuates significantly more than in Tb166. This implies that the d/σxxd/\sigma_{xx} is essential in the anomalous Hall scaling of RRMn6Sn6 compounds due to the presence of Mermin-Wagner fluctuations [17] involving either Mn or RR atoms.

Table 3: Comparison of intrinsic anomalous Hall conductivity between TbMn6Sn6 [12] and YMn6Sn5.45Ga0.55.
Compound cc (S/cm) aa (S/cm)-1 dd (S/cm)2
TbMn6Sn6 (|σyx||\sigma_{yx}|) 140 8.69×1098.69\times 10^{-9} 2.47×105-2.47\times 10^{5}
YMn6Sn5.45Ga0.55 (|σyx||\sigma_{yx}|) 121 3.25 ×108\times 10^{-8} 1.75×105-1.75\times 10^{5}
YMn6Sn5.45Ga0.55 (|σxz||\sigma_{xz}|) 310 1.89×1071.89\times 10^{-7} 5.41×105-5.41\times 10^{5}

The similar magnitude of intrinsic AHC in Y166-Ga and Tb166 suggests that the intrinsic AHC in RR166 compounds is a general property of ferrimagnetism rather than a result of exotic Chern physics, consistent with the conclusion in Ref. [12]. This interpretation is further supported by photoemission spectroscopy data from the parent compound Y166 [47], which shows no such topological feature above the Fermi energy, with Ga doping shifting the Fermi energy even lower.

To further validate this interpretation, we measured the Hall conductivity of Y166-Ga by applying an in-plane magnetic field, ensuring that any 2D Chern gap contributions would be excluded if present. While measuring this in-plane configuration for Tb166 would require an exceptionally high magnetic field due to magnetic saturation constraints at low temperatures [48, 12], Y166-Ga can be measured under standard lab conditions. For consistency with the ρyx\rho_{yx} measurement in Fig. 5, we applied the current along the [100] direction and applied the in-plane field along [120]. The Hall resistivity, ρzx\rho_{zx}, measured across a range of temperatures from 1.8 K to 300 K, is shown in Fig. 6(a).

In Fig. 6(b), we plot σxx\sigma_{xx} = 1/ρxx\rho_{xx} alongside the absolute value of the anomalous Hall conductivity, σzx\sigma_{zx} = ρzx/(ρxxρzz)-\rho_{zx}/(\rho_{xx}\rho_{zz}), which is valid when ρzx\rho_{zx} \ll ρxx\rho_{xx} and ρzz\rho_{zz}. The scaling of σzx\sigma_{zx}, following the relation in Eq. 2, is presented in Fig. 6(c), with the fitting coefficients aa, dd, and cc presented in Table 3. Notably, the intrinsic AHC contribution, represented by coefficient cc (σzx,intA\sigma^{A}_{zx,int}), is significantly larger than σyx,intA\sigma^{A}_{yx,int}, indicating that the AHC in Y166-Ga has a 3D nature and can be large without invoking the Chern physics. The coefficient |d||d| is also larger in this measurement geometry, potentially pointing to enhanced spin-fluctuations, though this cannot be directly compared to Tb166 due to differing Hall geometries and remains a topic of future work on RR166 compounds. However, it is noteworthy that the d/σxxd/\sigma_{xx} term is also necessary in this case. Nevertheless, our findings reveal that the AHC in Y166-Ga displays 3D characteristics and supports theoretical calculations that suggest the AHC arises from predominantly 3D bands [12], as further discussed in Section III.6.

III.6 Intrinsic AHC calculations for TbMn6Sn6 and YMn6Sn6

Refer to caption
Figure 7: Primitive unit cell used in AHC calculations. The top view of the YMn6Sn6 primitive unit cell. The xx- and yy-axes are highlighted in red, while the [120] direction is denoted in blue. The zz-axis is perpendicular to the plane.
Refer to caption
Refer to caption
Refer to caption
Figure 8: Calculated intrinsic anomalous Hall conductivity (AHC) of TbMn6Sn6 and YMn6Sn6 as a function of Fermi energy (EF\boldsymbol{E}_{\text{F}}). The intrinsic AHC of TbMn6Sn6 (Top) was calculated with an out-of-plane magnetization configuration, while that of YMn6Sn6 was calculated with both out-of-plane magnetization (Middle) and in-plane magnetization configurations (Bottom).

The intrinsic AHC can be calculated by integrating the Berry curvature over the Brillouin zone (BZ) [49]:

σαβ=e2BZdk(2π)3nf(Enk)Ωn(k)\displaystyle\sigma_{\alpha\beta}=-\dfrac{e^{2}}{\hbar}\int_{\text{BZ}}\dfrac{d\vec{k}}{(2\pi)^{3}}\sum_{n}f(E_{n\vec{k}})\Omega_{n}(\vec{k})\, (4)

where f(Enk)f(E_{n\vec{k}}) is the Fermi-Dirac distribution, Ωn,αβ(k)\Omega_{n,\alpha\beta}(\vec{k}) is the contribution to the Berry curvature from state nn, and α,β={x,y,z}\alpha,\beta=\{x,y,z\}.

A notable aspect of the calculation is the pronounced and rapid oscillation observed in the Berry curvature across the BZ, requiring a dense kk mesh to ensure convergence. To accelerate the calculations, we implemented Eq. 4 in our recently-developed tight-binding (TB) code [50] and carried out the AHC calculations in RRMn6Sn6 where RR = Tb and Y. A realistic TB Hamiltonian was constructed using the maximally localized Wannier functions (MLWFs) method [51, 52, 53] implemented in Wannier90 [54] after the self-consistent density-functional-theory calculations performed using Wien2k. A set of 118 Wannier functions (WFs) consisting of Y-4d4d (or Tb-5d5d), Mn-3d3d, and Sn-spsp orbitals offers an effective representation of the electronic structure near the Fermi level (EFE_{\text{F}}). The self-consistent DFT calculations were carried out with out-of-plane magnetization in TbMn6Sn6 and both in- and out-of-plane magnetization in YMn6Sn6. For the in-plane YMn6Sn6 configuration, the moment is along lattice vector 𝐚\mathbf{a}, as denoted in Fig.7. A dense 2563256^{3} kk-point mesh is used for the AHC calculations in TB.

Figure 7 presents the unit cell of YMn6Sn6 used in our calculation. The Cartesian coordinate system is chosen so that the lattice vector 𝐛\mathbf{b} is along the yy-axis, 𝐜\mathbf{c} is along the zz-axis, and 𝐚\mathbf{a} is along the 30-30^{\circ} direction off the xx-axis. Lattice vectors 𝐚\mathbf{a} and 𝐛\mathbf{b} point along the nearest neighboring (NN) Mn-Mn bond direction. For the σxy\sigma_{xy} calculations discussed below, the first subscript xx denotes the current direction, and the second subscript yy denotes the Hall-field direction.

Figure 8 shows the AHC values calculated at T=0T=0 K as functions of Fermi energy using Eq. 4. In the out-of-plane orientation of both TbMn6Sn6 and YMn6Sn6, only σxy\sigma_{xy} exhibits substantial values, while σyz\sigma_{yz} and σzx\sigma_{zx} remain negligible, as illustrated in the top and middle panels of Figure 8. Conversely, with in-plane magnetization in YMn6Sn6, both σyz\sigma_{yz} and σzx\sigma_{zx} demonstrate appreciable values, while σxy\sigma_{xy} remains negligible, as depicted in the bottom panel of Figure 8. The band-filling calculation indicates that when doping YMn6Sn6 by 0.090.09 hole/Mn, corresponding to a Fermi energy shift of -0.037 eV, σxy=77.62\sigma_{xy}=77.62 S/cm, and σzx=36.76\sigma_{zx}=36.76 S/cm. While the former closely aligns with the experimental value, the latter is approximately 10 times smaller than the experiment. These comparison has to be taken with caution, as random substitution of 0.55 Ga likely affects the electronic structure beyond a simple rigid Fermi energy shift. Nevertheless, our theoretical calculations predict that the Hall transport in RR166 is 3D, consistent with the experimental findings.

IV Conclusion

We reported results of magnetic and electrical transport measurements of YMn6Sn5.45Ga0.55 in two different geometries, supported by first-principles and DFT calculations. Our magneto-transport measurements across these two geometries confirm a more reliable scaling law that not only extracts the intrinsic AHC, but also accounts for contributions from spin fluctuations. These measurements revealed the 3D nature of the intrinsic AHC, which we attribute to the ferromagnetic properties of the material. The excellent agreement of the AHE with the empirical scaling law over the entire temperature suggests that this scaling relation may be important for not only for the large family of RRMn6Sn6 ferro/ferrimagnetic compounds, but also for other systems with strong spin fluctuations.

V ACKNOWLEDGMENTS

N.J.G. and H.B. acknowledge the support from the NSF CAREER award DMR-2343536. I.I.M. acknowledges support from the NSF award DMR-2403804. The work at the Ames National Laboratory was supported by the U.S. Department of Energy (US DOE), Office of Basic Energy Sciences, Division of Materials Sciences and Engineering. Ames National Laboratory is operated for the US DOE by Iowa State University under Contract No. DE-AC02-07CH11358.

Disclaimer: Certain commercial equipment, instruments, software or materials are identified in this paper in order to specify the experimental procedure adequately. Such identification is not intented to imply recommendation or endorsement by NIST, nor is it intended to imply that the materials or equipment identified are necessarily the best available for the purpose.

VI References

References

  • Zhang [2011] Z.-Y. Zhang, Quantum Hall effect in kagomé lattices under staggered magnetic field, Journal of Physics: Condensed Matter 23, 425801 (2011).
  • Ghimire and Mazin [2020] N. J. Ghimire and I. I. Mazin, Topology and correlations on the kagome lattice, Nature Materials 19, 137 (2020).
  • Kang et al. [2020] M. Kang, L. Ye, S. Fang, J.-S. You, A. Levitan, M. Han, J. I. Facio, C. Jozwiak, A. Bostwick, E. Rotenberg, et al., Dirac fermions and flat bands in the ideal kagome metal FeSn, Nature materials 19, 163 (2020).
  • Ye et al. [2018] L. Ye, M. Kang, J. Liu, F. Von Cube, C. R. Wicker, T. Suzuki, C. Jozwiak, A. Bostwick, E. Rotenberg, D. C. Bell, et al., Massive Dirac fermions in a ferromagnetic kagome metal, Nature 555, 638 (2018).
  • Yin et al. [2019] J.-X. Yin, S. S. Zhang, G. Chang, Q. Wang, S. S. Tsirkin, Z. Guguchia, B. Lian, H. Zhou, K. Jiang, I. Belopolski, et al., Negative flat band magnetism in a spin–orbit-coupled correlated kagome magnet, Nature Physics 15, 443 (2019).
  • Kuroda et al. [2017] K. Kuroda, T. Tomita, M.-T. Suzuki, C. Bareille, A. Nugroho, P. Goswami, M. Ochi, M. Ikhlas, M. Nakayama, S. Akebi, et al., Evidence for magnetic Weyl fermions in a correlated metal, Nature materials 16, 1090 (2017).
  • Mazin et al. [2014] I. I. Mazin, H. O. Jeschke, F. Lechermann, H. Lee, M. Fink, R. Thomale, and R. Valentí, Theoretical prediction of a strongly correlated Dirac metal, Nature Communications 5, 4261 (2014).
  • Bolens and Nagaosa [2019] A. Bolens and N. Nagaosa, Topological states on the breathing kagome lattice, Physical Review B 99, 165141 (2019).
  • Asaba et al. [2020] T. Asaba, S. M. Thomas, M. Curtis, J. D. Thompson, E. D. Bauer, and F. Ronning, Anomalous Hall effect in the kagome ferrimagnet GdMn6Sn6Physical Review B 101, 174415 (2020).
  • Bhandari et al. [2024] H. Bhandari, R. L. Dally, P. E. Siegfried, R. B. Regmi, K. C. Rule, S. Chi, J. W. Lynn, I. Mazin, and N. J. Ghimire, Magnetism and fermiology of kagome magnet YMn6Sn4Ge2npj Quantum Materials 9, 6 (2024).
  • Pokharel et al. [2021] G. Pokharel, S. M. Teicher, B. R. Ortiz, P. M. Sarte, G. Wu, S. Peng, J. He, R. Seshadri, and S. D. Wilson, Electronic properties of the topological kagome metals YV6Sn6 and GdV6Sn6Physical Review B 104, 235139 (2021).
  • Jones et al. [2024] D. C. Jones, S. Das, H. Bhandari, X. Liu, P. Siegfried, M. P. Ghimire, S. S. Tsirkin, I. I. Mazin, and N. J. Ghimire, Origin of spin reorientation and intrinsic anomalous Hall effect in the kagome ferrimagnet TbMn6Sn6Phys. Rev. B 110, 115134 (2024).
  • Pokharel et al. [2022] G. Pokharel, B. Ortiz, J. Chamorro, P. Sarte, L. Kautzsch, G. Wu, J. Ruff, and S. D. Wilson, Highly anisotropic magnetism in the vanadium-based kagome metal TbV6Sn6Physical Review Materials 6, 104202 (2022).
  • Arachchige et al. [2022] H. W. S. Arachchige, W. R. Meier, M. Marshall, T. Matsuoka, R. Xue, M. A. McGuire, R. P. Hermann, H. Cao, and D. Mandrus, Charge Density Wave in Kagome Lattice Intermetallic ScV6Sn6Physical Review Letters 129, 216402 (2022).
  • Riberolles et al. [2023] S. Riberolles, T. J. Slade, R. Dally, P. Sarte, B. Li, T. Han, H. Lane, C. Stock, H. Bhandari, N. Ghimire, et al., Orbital character of the spin-reorientation transition in TbMn6Sn6Nature Communications 14, 2658 (2023).
  • Lee et al. [2023] Y. Lee, R. Skomski, X. Wang, P. P. Orth, Y. Ren, B. Kang, A. K. Pathak, A. Kutepov, B. N. Harmon, R. J. McQueeney, I. I. Mazin, and L. Ke, Interplay between magnetism and band topology in the kagome magnets RMn6Sn6Phys. Rev. B 108, 045132 (2023).
  • Ghimire et al. [2020] N. J. Ghimire, R. L. Dally, L. Poudel, D. Jones, D. Michel, N. T. Magar, M. Bleuel, M. A. McGuire, J. Jiang, J. Mitchell, et al., Competing magnetic phases and fluctuation-driven scalar spin chirality in the kagome metal YMn6Sn6Science Advances 6, eabe2680 (2020).
  • Wang et al. [2021] Q. Wang, K. J. Neubauer, C. Duan, Q. Yin, S. Fujitsu, H. Hosono, F. Ye, R. Zhang, S. Chi, K. Krycka, et al., Field-induced topological Hall effect and double-fan spin structure with a c-axis component in the metallic kagome antiferromagnetic compound YMn6Sn6Physical Review B 103, 014416 (2021).
  • Siegfried et al. [2022] P. E. Siegfried, H. Bhandari, D. C. Jones, M. P. Ghimire, R. L. Dally, L. Poudel, M. Bleuel, J. W. Lynn, I. I. Mazin, and N. J. Ghimire, Magnetization-driven Lifshitz transition and charge-spin coupling in the kagome metal YMn6Sn6Communications Physics 5, 1 (2022).
  • Yin et al. [2020] J.-X. Yin, W. Ma, T. A. Cochran, X. Xu, S. S. Zhang, H.-J. Tien, N. Shumiya, G. Cheng, K. Jiang, B. Lian, et al., Quantum-limit Chern topological magnetism in TbMn6Sn6Nature 583, 533 (2020).
  • Riberolles et al. [2022] S. Riberolles, T. J. Slade, D. Abernathy, G. Granroth, B. Li, Y. Lee, P. Canfield, B. G. Ueland, L. Ke, and R. J. McQueeney, Low-Temperature Competing Magnetic Energy Scales in the Topological Ferrimagnet TbMn6Sn6Physical Review X 12, 021043 (2022).
  • Mielke III et al. [2022] C. Mielke III, W. Ma, V. Pomjakushin, O. Zaharko, S. Sturniolo, X. Liu, V. Ukleev, J. White, J.-X. Yin, S. Tsirkin, et al., Low-temperature magnetic crossover in the topological kagome magnet TbMn6Sn6Communications Physics 5, 107 (2022).
  • Li et al. [2023] Z. Li, Q. Yin, Y. Jiang, Z. Zhu, Y. Gao, S. Wang, J. Shen, T. Zhao, J. Cai, H. Lei, et al., Discovery of Topological Magnetic Textures near Room Temperature in Quantum Magnet TbMn6Sn6Advanced Materials , 2211164 (2023).
  • Xu et al. [2022] X. Xu, J.-X. Yin, W. Ma, H.-J. Tien, X.-B. Qiang, P. S. Reddy, H. Zhou, J. Shen, H.-Z. Lu, T.-R. Chang, et al., Topological charge-entropy scaling in kagome Chern magnet TbMn6Sn6Nature communications 13, 1197 (2022).
  • Gao et al. [2021] L. Gao, S. Shen, Q. Wang, W. Shi, Y. Zhao, C. Li, W. Cao, C. Pei, J.-Y. Ge, G. Li, et al., Anomalous Hall effect in ferrimagnetic metal RMn6Sn6 (R= Tb, Dy, Ho) with clean Mn kagome lattice, Applied Physics Letters 119 (2021).
  • Wenzel et al. [2022] M. Wenzel, A. A. Tsirlin, O. Iakutkina, Q. Yin, H. C. Lei, M. Dressel, and E. Uykur, Effect of magnetism and phonons on localized carriers in the ferrimagnetic kagome metals GdMn6Sn6 and TbMn6Sn6Phys. Rev. B 106, L241108 (2022).
  • Zhang et al. [2022a] H. Zhang, J. Koo, C. Xu, M. Sretenovic, B. Yan, and X. Ke, Exchange-biased topological transverse thermoelectric effects in a Kagome ferrimagnet, Nature communications 13, 1091 (2022a).
  • Crépieux and Bruno [2001] A. Crépieux and P. Bruno, Theory of the anomalous Hall effect from the Kubo formula and the Dirac equation, Phys. Rev. B 64, 014416 (2001).
  • Tian et al. [2009] Y. Tian, L. Ye, and X. Jin, Proper Scaling of the Anomalous Hall Effect, Phys. Rev. Lett. 103, 087206 (2009).
  • Grigoryan et al. [2017] V. L. Grigoryan, J. Xiao, X. Wang, and K. Xia, Anomalous Hall effect scaling in ferromagnetic thin films, Phys. Rev. B 96, 144426 (2017).
  • Xu et al. [2021] C. Xu, T. Heitmann, H. Zhang, X. Xu, and X. Ke, Magnetic phase transition, magnetoresistance, and anomalous Hall effect in Ga-substituted YMn6Sn6 with a ferromagnetic kagome lattice, Physical Review B 104, 024413 (2021).
  • Bruker AXS [2021] Bruker AXS, APEX-4, Bruker AXS, Madison, Wisconsin, USA (2021).
  • Krause et al. [2015] L. Krause, R. Herbst-Irmer, G. M. Sheldrick, and D. Stalke, Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination, Journal of applied crystallography 48, 3 (2015).
  • Sheldrick [2015a] G. M. Sheldrick, Shelxt–integrated space-group and crystal-structure determination, Acta Crystallographica Section A: Foundations and Advances 71, 3 (2015a).
  • Sheldrick [2015b] G. M. Sheldrick, Crystal structure refinement with shelxl, Acta Crystallographica Section C: Structural Chemistry 71, 3 (2015b).
  • Kresse and Furthmüller [1996] G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54, 11169 (1996).
  • Blöchl [1994] P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50, 17953 (1994).
  • Kresse and Joubert [1999] G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B 59, 1758 (1999).
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett. 77, 3865 (1996).
  • Anisimov and Gunnarsson [1991] V. I. Anisimov and O. Gunnarsson, Density-functional calculation of effective Coulomb interactions in metals, Phys. Rev. B 43, 7570 (1991).
  • Zhang et al. [2001] S.-y. Zhang, P. Zhao, Z.-h. Cheng, R.-w. Li, J.-r. Sun, H.-w. Zhang, and B.-g. Shen, Magnetism and giant magnetoresistance of YMn6Sn6-xGax (x= 0-1.8) compounds, Physical Review B 64, 212404 (2001).
  • Dally et al. [2021] R. L. Dally, J. W. Lynn, N. J. Ghimire, D. Michel, P. Siegfried, and I. I. Mazin, Chiral properties of the zero-field spiral state and field-induced magnetic phases of the itinerant kagome metal YMn6Sn6Physical Review B 103, 094413 (2021).
  • Idrissi et al. [1991] B. C. E. Idrissi, G. Venturini, and B. Malaman, Magnetic structures of TbMn6Sn6 and HoMn6Sn6 compounds from neutron diffraction study, Journal of the Less Common Metals 175, 143 (1991).
  • Rosenfeld and Mushnikov [2008] E. Rosenfeld and N. Mushnikov, Double-flat-spiral magnetic structures: Theory and application to the RMn6X6 compounds, Physica B: Condensed Matter 403, 1898 (2008).
  • P. Blaha et al. [2002] P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, and J. Luitz, WIEN2K (2002), ISBN 3-9501031-1-2.
  • Zhang et al. [2022b] H. Zhang, J. Koo, C. Xu, M. Sretenovic, B. Yan, and X. Ke, Exchange-biased topological transverse thermoelectric effects in a Kagome ferrimagnet, Nature Communications 13, 1091 (2022b).
  • Li et al. [2021] M. Li, Q. Wang, G. Wang, Z. Yuan, W. Song, R. Lou, Z. Liu, Y. Huang, Z. Liu, H. Lei, et al., Dirac cone, flat band and saddle point in kagome magnet YMn6Sn6Nature communications 12, 1 (2021).
  • Kimura et al. [2006] S. Kimura, A. Matsuo, S. Yoshii, K. Kindo, L. Zhang, E. Brück, K. Buschow, F. De Boer, C. Lefèvre, and G. Venturini, High-field magnetization of RMn6Sn6 compounds with R= Gd, Tb, Dy and HoJournal of alloys and compounds 408, 169 (2006).
  • Wang et al. [2006] X. Wang, J. R. Yates, I. Souza, and D. Vanderbilt, Ab initio calculation of the anomalous Hall conductivity by Wannier interpolation, Physical Review B 74, 195118 (2006).
  • Ke [2019] L. Ke, Intersublattice magnetocrystalline anisotropy using a realistic tight-binding method based on maximally localized Wannier functions, Physical Review B 99, 054418 (2019).
  • Marzari and Vanderbilt [1997] N. Marzari and D. Vanderbilt, Maximally localized generalized Wannier functions for composite energy bands, Physical review B 56, 12847 (1997).
  • Souza et al. [2001] I. Souza, N. Marzari, and D. Vanderbilt, Maximally localized Wannier functions for entangled energy bands, Physical Review B 65, 035109 (2001).
  • Marzari et al. [2012] N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Maximally localized Wannier functions: Theory and applications, Reviews of Modern Physics 84, 1419 (2012).
  • Mostofi et al. [2014] A. A. Mostofi, J. R. Yates, G. Pizzi, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, An updated version of wannier90: A tool for obtaining maximally-localised Wannier functions, Computer Physics Communications 185, 2309 (2014).