Metamagnetic multiband Hall effect in Ising antiferromagnet ErGa2
Edited by Joseph Orenstein, Department of Physics, University of California, Berkeley, CA; received October 22, 2023; accepted April 13, 2024
Significance
The topological Hall effect (THE), often assigned to jumps in Hall resistivity at metamagnetic transitions, is now widely employed as a tool to electrically probe emergent magnetic fields associated with spin textures such as skyrmions in conducting systems. An empirical ansatz is often used to experimentally extract the THE from the Hall responses; here, we present a Hall effect study of a frustrated triangular-lattice magnet ErGa2 as a counterexample that suggests the need for more thorough consideration of this approach. The proposed multiband mechanism for the observed Hall anomaly is anticipated to be relevant to a wide class of magnetic systems. Our study sheds light on an overlooked transport effect and highlights crucial considerations when identifying nontrivial Hall responses.
Abstract
Frustrated rare-earth-based intermetallics provide a promising platform for emergent magnetotransport properties through exchange coupling between conduction electrons and localized rare-earth magnetic moments. Metamagnetism, the abrupt change of magnetization under an external magnetic field, is a signature of first-order magnetic phase transitions; recently, metamagnetic transitions in frustrated rare earth intermetallics have attracted interest for their accompanying nontrivial spin structures (e.g., skyrmions) and associated nonlinear and topological Hall effects (THE). Here, we present metamagnetism-induced Hall anomalies in single-crystalline ErGa2, which recalls features arising from the THE but wherein the strong Ising-type anisotropy of Er moments prohibits noncoplanar spin structures. We show that the observed anomalies are neither due to anomalous Hall effect nor THE; instead, can be accounted for via 4f-5d interactions which produce a band-dependent mobility modulation. This leads to a pronounced multiband Hall response across the magnetization process–a metamagnetic multiband Hall effect that resembles a topological-Hall-like response but without nontrivial origins. The present findings may be of general relevance in itinerant metamagnetic systems regardless of coplanar/noncoplanar nature of spins and are important for the accurate identification of Hall signals due to emergent magnetic fields.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
where the first term is the ordinary Hall effect (OHE) due to Lorentz force and the second term is the anomalous Hall effect (AHE) proportional to the magnetization (M). The third term is the topological Hall effect (THE) that appears, e.g., when noncoplanar spin alignments in an intermediate field region are realized. Such spin textures give rise to finite scalar spin chirality, χijk ∝ Si · (Sj × Sk) (3–6), where Si,j,k are spin moments at neighboring sites, i, j, k. χijk acts as a local emergent magnetic field for conduction electrons to give rise to a transverse response (7–12). To isolate these terms, in the simplest case R0 and Rs are characterized by measurements beyond magnetization saturation (13–15), with subsequent analysis of ρyxT involving the subtracting background R0B+RsM. However, difficulty in isolating the THE has recently been discussed in thin film systems (16, 17); the degree to which this is relevant for bulk crystalline systems has heretofore not been experimentally addressed.
Here, we report the Hall response in bulk single crystals, where the ansatz (Eq. 1) breaks down. Our target compound is the frustrated magnet ErGa2 (18) that exhibits a remarkable nonmonotonic Hall response across metamagnetic transitions, but finite scalar spin chirality is forbidden due to strong Ising-type anisotropy. We propose a model that describes this as a multiband modulation of R0 in the magnetization process, i.e., a metamagnetic multiband (MM) Hall effect. This generates a significant nonmonotonic field-dependence of ρyx arising from band-dependent spin-electron scattering, even without considering nontrivial AHE/THE terms. This model can generally be applied to multiband, metallic magnets. We anticipate the MM Hall effect can in principle occur in wider classes of systems potentially possessing noncoplanar spin structure alongside the AHE and THE therein. The finding of the present study highlights an important consideration when applying (Eq. 1) to identify emergent magnetic field in frustrated itinerant magnets.
Results
We first summarize the model for the MM Hall response. In the presence of more than one type of conduction electrons, the coefficient R0 in Eq. 1 in principle should reflect a multiband response; here, we consider a representative two-band scenario with hole and electron pockets (their respective carrier densities as nh and ne and mobilities as µh and µe). In the low-field regime (µe,hB < 1), this is approximately (19)
[2]
However, for magnetic systems, the mobility is further modified by band-dependent spin-electron coupling. When the magnetization rapidly evolves at a metamagnetic transition this can give rise to the MM Hall effect. Importantly, in typical magnetically textured system (in particular those based on rare earth), metamagnetic transitions can take place at a relatively low magnetic field scale and can often be lower than the threshold for typical orbitally driven multiband (OHE) nonmonotonic Hall responses.
For concreteness, we consider a metallic compound with an electron band with 5d/6s orbital characters from magnetic cations and a hole band mainly composed of p-orbital at nonmagnetic (anionic) sites (Fig. 1A). For the case of rare earth intermetallics, the local magnetic moment has the form of 4f levels that strongly couples with the cation conduction band through the intra-atomic 5d-4f interaction, JS · sδ (r − R) (20), where J is the exchange coupling constant, S and s are spins localized at R and for conduction electrons with spatial coordinate r, respectively. The intersite interaction with the anion p-valence band, however, is expected to be relatively weak due to their reduced hybridization (21–23) and energy cost for 4f-p electron hopping (we consider the case where the 4f level is well-localized below the Fermi energy, which is typically realized in rare earth intermetallics apart from Ce, Pr, and Yb; see SI Appendix, section S1 for the opposite extreme where interatomic f-p hybridization becomes relevant). In such cases, the electron-type carriers are more readily scattered from magnetic disorder (Fig. 1B), giving rise to a field modulation in Eq. 2 via relative changes in µh and µe.
Fig. 1.

This can be quantitatively illustrated with a simplified two-band model, in which we assume that µe is sensitive to while µh is immune from magnetic disorder scattering. Within Matthiessen’s rule, the mobility of electron-type carriers is
[3]
The inverse of the carrier lifetime is additive: τ−1 = τ0−1 + τmag−1, where τ0 is for the nonmagnetic (field-independent) scattering due to impurities/defects and phonons, and τmag stems from electron-spin scattering. In the magnetization process from the paramagnetic state (PM) to the field-induced ferromagnetic state (Fi-FM), the field-dependence of 1/τemag can be expressed as
[4]
where the spin-disorder scattering is taken to be proportional to the magnitude of disorder upon uniformly aligned spins, viz (1 – M2/M02) (24–26); this provides a good description for Ising magnets wherein spin-flip and spin-wave scattering processes are minimal (27–29).
Fig. 1 C–E show the evolution of µe, two-band resistivity (ρxx), and Hall resistivity (ρyx) for a linear evolution of M as a general illustration for the case of constant µh and ne,h. The negative magnetoresistance at low fields is consistent with field-suppression of magnetic disorder (spin MR), while the Hall resistivity exhibits nonlinearity. We note that this model does not include a contribution from M-linear AHE and THE and carrier density change. The modulation of the Hall effect arises from the field-dependence of OHE across the magnetization process. This effect appears to be often overlooked in the conventional treatment of the AHE caused by side jump, skew scattering, and Berry phase mechanism, where the Hall anomaly originates from the second term in Eq. 1 and R0 is taken as a constant.
The extension of this model to a metamagnetic transitions among antiferromagnetic (AFM), metamagnetic (M), and Fi-FM states is shown in Fig. 1 F–H. At zero temperature, spin-disorder scattering or spin-wave scattering does not contribute to the spin MR. Instead, the order parameter and magnetic unit cell change across the metamagnetic transition result in a reconstruction of the magnetic Brillouin zone (BZ) and give rise to a stepwise change of the renormalized effective-mass of carriers. The magnetoresistivity and Hall resistivity (Fig. 1 G and H) both reflect the characteristic field-dependence of µe–the MM Hall effect. We note such affects ascribed to changes in carrier lifetime have been previously reported at metamagnetic transitions for other Ising-like rare earth systems (30, 31). The crucial point of this effect is that it should be distinguished from the THE to avoid the overestimation of the emergent magnetic field (and in turn interpretation in terms of a noncoplanar nature of the underlying magnetic structure).
Here, we present a pronounced MM Hall effect in ErGa2 as a remarkable example of the above scenario. Rare-earth intermetallics of the type RGa2 (R: rare earth) are frustrated magnets with an alternative c axis stacking of R triangular-lattice Ga honeycomb layers (Fig. 2A) (32, 33). ErGa2 exhibits an archetypical two-step metamagnetic transition for H ||c (34), desirable for the current study. This is due to the frustration (18) among the Er moments with strong easy-axis type anisotropy (18, 35). Below the AFM transition at TN = 7 K (36), commensurate single-q order for the up–down configuration of Er-moments is stabilized in zero field (AFM in Fig. 2A) (37). Application of magnetic field along the easy-axis induces two stepwise transitions (18, 36, 37). At magnetic field of 0.8 T a 3up-1down-type commensurate phase is induced. The network of upward magnetic sites forms a kagome network (Fig. 2A) (18). This phase retains the original magnetic modulation vector (qAFM = ½a*) as the principal modulation but forms a triple-q state with a new reciprocal lattice unit: a*kagome = ½a*, b*kagome = ½b*, and c*kagome = c*, where a*, b*, and c* are the reciprocal lattice unit of the original lattice. At low temperatures, a half-magnetization plateau remains until the field-induced ferromagnetic state for qFM = (0, 0, 0) is induced at 2 T (Fig. 2A). Coupling between conduction electrons and Er moments is observed as a sudden drop of resistivity at TN (38) as well as in magnetoresistance anomalies across the metamagnetic transitions (39).
Fig. 2.

Fig. 2 B–D show the magnetization and magnetotransport properties of ErGa2, measured with single crystals for H||c. The half-magnetization plateau in the intermediate kagome spin state (Fig. 2E) and characteristic magnetic magnetoresistance are similar to those previously reported (39). Across the metamagnetic transition, we further observe a nonmonotonic Hall resistivity, which shows a positive enhancement in the intermediate kagome state. As a THE is precluded by the Ising-type magnetism, this arises from some combination of OHE and AHE. The persistence of the nonlinearity at T = 10 K in the paramagnetic phase (gray curve in Fig. 2D) is suggestive of two-carrier behavior with high-mobile hole-type (low density) carriers and electron-type low-mobile (high density) carriers, which sharpens at lower temperatures along with magnetization and ρxx (Fig. 2C), i.e., an OHE.
Comparison of the temperature dependence of ρyx/B of ErGa2 and that for isoelectronic and isostructural LaGa2 (40) (Fig. 2F) shows that the two are of a similar value at all the temperature range (ratio < 1.4) including lack of significant discrepancy below TN for ErGa2, suggesting a weakness or absence of AHE in ErGa2. This is consistent with the absence of skew scattering given the longitudinal conductivity σxx ~ 105 S/cm (41), where the intrinsic mechanism rather than the extrinsic scattering mechanism becomes more relevant. Moreover, as shown in Fig. 2F, there is no significant anomaly in ρyx/B at TN in contrast to the factor ρxx2M/B (blue circle in Fig. 2F) where the intrinsic Berry curvature would manifest if present (42). This is distinct from the reported features for other rare-earth materials (14, 15, 43–45) and RGa2 (R = Ce, Sm, Gd) (40, 46), where anomalous Hall contribution was suggested. We note that demagnetization correction may also affect the estimation of R0 as discussed in SI Appendix, section S2; therein, the Hall coefficient is enhanced at low temperatures and shows a peak at TN [not explicitly argued in the previously (40, 46)]. The absence of AHE in ErGa2 may further arise due to small de Gennes factors (∝ (gJ – 1)2J(J+1)), the square root of which roughly determines the scale of the exchange field on conduction electrons from the localized moment of rare earth ions (47, 48).
To illustrate the consistency of the observed effect with the MM Hall effect, we plot M, σxx (=ρxx/(ρxx2+ρyx2)), and σxy (=ρyx/(ρxx2+ρyx2)) at T = 10 K in Fig. 3 A–C as a function of B (=µ0Hint + M). We analyze the transport properties within the model introduced above. For simplicity, we introduce the mobility modulation due to spin alignment to the electron-type carriers (dominant in ErGa2, which we return to below). Fig. 3 A–C show this analysis applied for the PM-to-Fi-FM process at T = 10 K. The field-dependence of σxx and σxy can be simulated within a conventional two-band model:
[5]
Fig. 3.

where µh, ne, and nh are fitting parameters. From the field-dependence of M(B), µe(B) is captured by fixing and in Eqs. 3 and 4 (the negligible contributions from the spin-fluctuations and spin-wave-scattering are further discussed in SI Appendix, section S3). This shows reasonable agreement with the observed response, reproducing the spin MR σxx and nonmonotonic behavior in σxy simultaneously.
We next extend this model to T = 1.8 K within the AFM, kagome, and Fi-FM phases as shown in Fig. 3 D–F, where we assume the proportionality of the field-dependence of ρxx(B)/ρxx0 to 1/µe(B) (Fig. 3D) (31). We note that at low temperatures there are interphase states (IP1 and IP2) across the metamagnetic transitions which give rise to singular peaks in σxx (Fig. 2E); as these are proposed to arise from specular domain wall scattering (39), we exclude these in the following analysis (gray-hatched area in Fig. 3 D–F and see SI Appendix, section S4 for a discussion on the IP states). We also note that multistep metamagnetism is observed (SI Appendix, section S2) suggestive of more complex intermediate spin configurations in the IP1 and IP2 regions (this is beyond the scope of our simplified model).
Fig. 3 G and H summarize the transport parameters obtained by performing the analysis using the ansatz ρxx(B)/ρxx0 ∝ 1/µe(B) at various temperatures. For T = 10 and 8 K (>TN), the parameter set is consistent with those for Eq. 4. We also confirmed the absence of significant anomaly in µe(9 T), ne, nh, and µh at TN, while µe(0 T) is enhanced below TN due to the suppression of the spin-disorder scattering. nh shows a slight temperature dependence (Fig. 3H), which might not be an intrinsic effect as the fitting quality does not change by fixing ne and nh constants. Fig. 3I shows the zero field transport properties reproduced from the fitting parameters; the two-band Hall coefficient (R0 in Eq. 2) and resistivity ρxx,0 T (=σxx(0 T)−1 in Eq. 5) are reasonably consistent with the direct observation. The numbers of carrier per unit volume are 1.2 and 3 × 10−4, for electrons and holes, respectively. The large electron concentration is consistent with the negative slope of ρyx at high fields, corresponding to 1.5 × 1022 cm−3 (Fig. 2D). Two orders of magnitude higher mobility of the hole pocket is necessary to reproduce the positive Hall coefficient at zero field (Figs. 2F and 3I). We note that these values reflect an oversimplification of the electronic structure of this system as a two-band model and are not directly connected to the volume of Fermi surfaces for the low-field Hall effect (49–51), which is affected by the local curvatures (see below).
To evaluate the relevance of this model in the present intermetallic setting, we calculate the band structure for the nonmagnetic analog LaGa2 (Fig. 4A). Ellipsoidal hole pockets (FS1 and FS2, see Fig. 4B) at the A point on the BZ edge are ascribed to the Ga-4p band, while the large Fermi surface (FS3, see Fig. 4C) is composed mainly of La-5d orbitals, consistent with the previous studies (52, 53). These two bands correspond to those in the schematic model in Fig. 1A. The Hall coefficient for each Fermi surface is determined by the sign of curvature γ (∝-σxy) (49, 50). We map γ onto each Fermi surface as shown in Fig. 4 D and E. For the hole bands (FS1 and 2), it is evident that their contributions to the Hall effect are > 0. The FS3 has both characters showing a sign change for γ from negative near kz = π/c (A point) to positive across around kz = π/2c to 0 (Fig. 4E), where c is the lattice constant. The representative electron-type carrier orbits at kz ~ π/8c and kz ~ 0 are depicted in Fig. 4 F and G, respectively, which confirms the presence of an electron-type orbit deflecting counter-clockwise (SI Appendix, section S5) (51). We note that a three-band treatment (a hole pocket for p-band, an electron and a hole pocket with comparable concentration and mobility for d-band) would be more realistic on the basis of the band calculations. This can be effectively treated by a two-band model by representing the d-band pockets as a large electron pocket and leaving a hole pocket for the high-mobile p-band (SI Appendix, section S6). These results support the application of the two-band model with the asymmetric spin–charge coupling as an ansatz to analyze the magnetotransport properties of RGa2. We also note that this argument can be applied to the magnetic ErGa2 on the basis of the rigid band approximation from the nonmagnetic analog LaGa2, where the 4f electrons are expected to be well-localized (Fig. 1A) and only weakly hybridize with the electrons near the Fermi energy. This is consistent with previous quantum oscillation experiments in various RGa2 (R = Ce, Pr, Sm, Gd) (46, 54–57). In these studies, the quantum oscillation branches associated with the FS1, 2, and 3 are identified almost unchanged from the nonmagnetic LaGa2. Further, in CeGa2, the exchange splitting on the FS1 and FS2 has been observed to be smaller than that of FS3 (54) validating the above simplification for the mobility modulation only on the electron-type carriers in ErGa2.
Fig. 4.

Discussion
We note that the modulation of the OHE has also been considered in dilute magnetic alloys, therein known as the “spin effect” (58). In particular, it has been established that a variation of R0 arises due to the difference of the field-induced change of scattering rate of conduction electrons with up and down spins. The distinction from the current model is that the former always leads to the enhancement of R0, but the latter can induce both enhancement and reduction depending on the details of the electronic and magnetic subsystems. In the current analysis, we consider the mobility is field-dependent in the magnetization process. It is known that the mobility can be field-dependent even without the magnetic moments as observed in diluted (semi)metals (59, 60) when the disorder potential smoothly varies compared to the cyclotron radius (61). This effect is expected to be masked in the present metallic system owing to the short Thomas-Fermi screening length (~Å). Magnetic semiconductors/semimetals with reduced carriers are a class of materials where this effect potentially interferes with the MM Hall effect. Beyond the mobility and effective-mass changes discussed above, we also note that there are possible contributions to the field-dependence of carrier density across metamagnetic transitions. This effect could be significant in semimetals with low carrier concentration as discussed previously in Eu compounds (62). ErGa2, on the other hand, is a metal with large carrier number and not a heavy-fermion system and is anticipated to be insensitive to changes of magnetic q-vector as it is commensurate to the lattice. The valley-dependent carrier emptying (63) as well as the magnetic breakdown (64) is irrelevant in ErGa2 (and is instead important for the quantum limit in semimetals).
In conclusion, we have observed a MM Hall effect in the Ising-antiferromagnet ErGa2 and demonstrated its consistency with a realistic two-band model. The orbital character in the band structure is further consistent with the two-carrier feature with different scattering cross section to magnetic moments. The band-induced Hall effect discussed in our work in principle can also be present in intermetallic metamagnetic systems with topological spin textures. For the Hall anomaly in such systems, this shows rather than a straightforward assignment to a THE via the ansatz Eq. 1, that instead it can be captured by a multiband response of the charge carriers. This finding provides a clear example where the ansatz Eq. 1 fails to capture the origin of the Hall response and a simple framework to understand the magnetotransport properties of magnetic intermetallics, which would be useful for more precise evaluation of the emergent field therein.
Materials and Methods
Single crystals of ErGa2 were grown via the Pb self-flux method following ref. 65. The starting materials Er, Ga, and Pb were mixed in the molar ratio Er:Ga:Pb = 1:2:10. They were loaded into a 2-mL alumina crucible and sealed in an evacuated quartz tube. The growth ampoule was heated to 1,000 °C and slowly cooled to 400 °C at a rate of 1 °C/min. To increase the size of crystals, the furnace temperature was increased to 850 °C in 2 h and cooled down to 400 °C at 1 °C/min; this process was subsequently repeated between 700 °C to 400 °C. The single crystals were separated by decanting the flux in a centrifuge. Finally, crystals were annealed in an evacuated quartz tube at 600 °C for 6 d. The typical size of the crystals is 1 × 1 × 0.5 mm3 (for a×b×c). The quality of the crystal was checked by the single-crystal x-ray diffraction (XRD) using the synchrotron light source at SPring-8, Japan (SI Appendix, section S7).
Electrical transport measurements were performed by a conventional five-probe method with an AC excitation current of 1 mA at a typical frequency near 15 Hz. The transport response in low temperature and a magnetic field was measured using commercial superconducting magnets and cryostat. The obtained longitudinal and transverse signals were field-symmetrized and antisymmetrized to correct for a contact misalignment, respectively. Magnetization measurements were performed using a commercial superconducting quantum interference device magnetometer.
The electronic band structures of LaGa2 were calculated by the density functional theory (DFT) code using the Vienna ab initio simulation package (SI Appendix, section S5). The electronic band structures and projections were further computed on a finer 100 × 100 × 90 k-mesh grid to create interpolated Fermi surfaces near the Fermi level for evaluating band characteristics and curvatures.
Data, Materials, and Software Availability
Experimental data on electronic and structural properties. Data have been deposited in Harvard Dataverse (https://doi.org/10.7910/DVN/UQ2T5V) (66).
Acknowledgments
We thank Y. Nakamura and H. Sawa for supporting synchrotron XRD experiments. The synchrotron radiation experiments were performed at SPring-8 with the approval of the Japan Synchrotron Radiation Research Institute (Proposal No. 2023A1882). This research is funded in part by the Gordon and Betty Moore Foundation Emergent Phenomena in Quantum Systems (EPiQS) through Grants GBMF9070 to J.G.C. (material synthesis, DFT calculations), NSF grant DMR-1554891 (material design), ONR Grant N00014-21-1-2591 (instrumentation development), and AFOSR grant FA9550-22-1-0432 (advanced characterization). T.K. acknowledges the support by the Yamada Science Foundation Fellowship for Research Abroad and Japan Society for the Promotion of Science (JSPS) Overseas Research Fellowships. S.F. acknowledges support from a Rutgers Center for Material Theory Distinguished Postdoctoral Fellowship. L.Y. acknowledges the support by the Science and Technology Center (STC) for Integrated Quantum Materials, NSF grant number DMR-1231319, the Heising-Simons Physics Research Fellow Program, and the Tsinghua Education Foundation.
Author contributions
T.K. and J.G.C. designed research; T.K., S.F., L.Y., and S.K. performed research; T.K. analyzed data; and T.K., S.F., L.Y., S.K., and J.G.C. wrote the paper.
Competing interests
The authors declare no competing interest.
Supporting Information
Appendix 01 (PDF)
- Download
- 1.07 MB
Note Added in Proof
Recently, we became aware of a related work by Y. Ōnuki et al. (67) on transport properties of RGa2 (R = rare earth) and structurally related systems.
References
1
C. M. Hurd, The Hall Effect in Metals and Alloys (Plenum, New York, 1972).
2
Y. Tokura, N. Kanazawa, Magnetic skyrmion materials. Chem. Rev. 121, 2857 (2020).
3
X. G. Wen, F. Wilczek, A. Zee, Chiral spin states and superconductivity. Phys. Rev. B 39, 11413 (1989).
4
J. Ye et al., Berry phase theory of the anomalous Hall effect: Application to colossal magnetoresistance manganites. Phys. Rev. Lett. 83, 3737 (1999).
5
P. Bruno, V. K. Dugaev, M. Taillefumier, Topological Hall effect and Berry phase in magnetic nanostructures. Phys. Rev. Lett. 93, 096806 (2004).
6
S. D. Yi et al., Skyrmions and anomalous Hall effect in a Dzyaloshinskii-Moriya spiral magnet. Phys. Rev. B 80, 054416 (2009).
7
Y. Taguchi et al., Spin chirality, Berry phase, and anomalous Hall effect in a frustrated ferromagnet. Science 291, 2573 (2001).
8
A. Neubauer et al., Topological Hall effect in the A phase of MnSi. Phys. Rev. Lett. 102, 186602 (2009).
9
M. Lee et al., Unusual Hall effect anomaly in MnSi under pressure. Phys. Rev. Lett. 102, 186601 (2009).
10
R. Ritz et al., Giant generic topological Hall resistivity of MnSi under pressure. Phys. Rev. B 87, 134424 (2013).
11
C. Franz et al., Real-space and reciprocal-space Berry phases in the Hall effect of Mn1-xFexSi. Phys. Rev. Lett. 112, 186601 (2014).
12
T. Kurumaji et al., Skyrmion lattice with a giant topological Hall effect in a frustrated triangular-lattice magnet. Science 365, 914 (2019).
13
W.-L. Lee et al., Dissipationless anomalous Hall current in the ferromagnetic spinel CuCr2Se4-xBrx. Science 303, 1647 (2004).
14
J. J. Rhyne, Anomalous Hall effect in single-crystal dysprosium. Phys. Rev. 172, 523 (1968).
15
J. J. Rhyne, Anomalous and ordinary Hall effect in terbium. J. Appl. Phys. 40, 1001 (1969).
16
A. Gerber, Interpretation of experimental evidence of the topological Hall effect. Phys. Rev. B 98, 214440 (2018).
17
G. Kimbell et al., Challenges in identifying chiral spin textures via the topological Hall effect. Commun. Mater. 3, 19 (2022).
18
M. Doukoure, D. Gignoux, Metamagnetism in ErGa2 studied on a single crystal. J. Magn. Magn. Mater. 30, 111 (1982).
19
E. H. Sondheimer, A. H. Wilson, the theory of the magneto-resistance effects in metals. Proc. Roy. Soc. A. 190, 435 (1947).
20
V. F. Ghantmakher, Y. B. Levinson, Carrier Scattering in Metals and Semiconductors (North-Holland, Amsterdam, 1987).
21
J. Kuneš, W. E. Pickett, Kondo and anti-Kondo coupling to local moments in EuB6. Phys. Rev. B 69, 165111 (2004).
22
H.-S. Li, Y. P. Li, J. M. D. Coey, R-T and R-R exchange interactions in the rareearth (R)-transition-metal (T) intermetallics: An evaluation from relativistic atomic calculations. J. Phys. Condens. Matter. 3, 7277 (1991).
23
T. Nomoto, T. Koretsune, R. Arita, Formation mechanism of the helical Q structure in Gd-based skyrmion materials. Phys. Rev. Lett. 125, 117204 (2020).
24
K. Yosida, Anomalous electrical resistivity and magnetoresistance due to an s-d interaction in Cu-Mn alloys. Phys. Rev. 107, 396 (1957).
25
K. Kubo, N. Ohata, A quantum theory of double exchange. I. J. Phys. Soc. Jpn. 33, 21 (1972).
26
A. Urushibara et al., Insulator-metal transition and giant magnetoresistance in La1−xSrxMnO3. Phys. Rev. B 51, 14103 (1995).
27
T. Kasuya, Electrical resistance of ferromagnetic metals. Prog. Theo. Phys. 16, 58 (1956).
28
P. G. de Gennes, J. Friedel, Anomalies de resistivite dans certains metaux, magnetiques. J. Phys. Chem. Solids 4, 71 (1958).
29
T. van Peski-Tinbergen, A. J. Dekker, Spin-dependent scattering and resistivity of magnetic metals and alloys. Physica 29, 917 (1963).
30
S. M. Thomas et al., Hall effect anomaly and low-temperature metamagnetism in the Kondo compound CeAgBi2. Phys. Rev. B 93, 075149 (2016).
31
L. Ye, T. Suzuki, J. G. Checkelsky, electronic transport on the Shastry-Sutherland lattice in Ising-type rare-earth tetraborides. Phys. Rev. B 95, 174405 (2017).
32
A. Raman, On AlB2 type phases. Z. Metallkd. 58, 179 (1967).
33
A. Fedorchuk, Y. Grin, “Crystal structure and chemical bonding in gallides of rare-earth metals” in Handbook on the Physics and Chemistry of Rare Earths, J.-C. G. Bünzli, V. K. Pecharsky, Eds. (Elsevier Science, 2018), vol. 53, p. 81.
34
T.-H. Tsai, D. J. Sellmyer, Magnetic ordering and exchange interactions in the rare-earth gallium compounds RGa2. Phy. Rev. B 20, 4577 (1979).
35
M. Diviš, M. Richter, J. Forstreuter, K. Koepernik, H. Eschrig, Crystal-field and magnetism of RGa2 (R = Ce, Er) compounds derived from density-functional calculations. J. Magn. Magn. Mater. 176, L81 (1997).
36
T. H. Tsai, J. A. Gerber, J. W. Weymouth, D. J. Sellmyer, Magnetic properties of the rare-earth intermetallics RGa2. J. Appl. Phys. 49, 1507 (1978).
37
H. Asmat, D. Gignoux, Magnetic structures of the intermetallic rare-earth compounds RGa2. Inst. Phys. Conf. Ser. 37, 286 (1978).
38
J. A. Blanco, D. Gignoux, J. C. Gómez Sal, J. Rodríguez Fdez, D. Schmitt, Magnetic contribution to the electrical resistivity in RGa2 compounds (R = rare earth). J. Magn. Magn. Mater. 104–107, 1285–1286 (1992).
39
N. V. Baranov, P. E. Markin, A. I. Kozlov, E. V. Sinitsyn, Effect of magnetic interphase boundaries on the electrical resistivity in metallic metamagnets. J. Alloys Comp. 200, 43 (1993).
40
H. Henmi, Y. Aoki, T. Fukuhara, I. Sakamoto, H. Sato, Transport properties of RGa2 (R = La, Ce and Sm). Phys. B 186–188, 655 (1993).
41
S. Onoda, N. Sugimoto, N. Nagaosa, Quantum transport theory of anomalous electric, thermoelectric, and thermal Hall effects in ferromagnets. Phys. Rev. B 77, 165103 (2008).
42
M. Lee, Y. Onose, Y. Tokura, N. P. Ong, Hidden constant in the anomalous Hall effect of high-purity magnet MnSi. Phys. Rev. B 75, 172403 (2007).
43
F. E. Maranzana, Contributions to the theory of the anomalous Hall effect in ferro- and antiferromagnetic materials. Phys. Rev. 160, 421 (1967).
44
B. Giovannini, theory of the anomalous Hall effect in the rare earths. Phys. Lett. A 27, 381 (1971).
45
T. Hiraoka, Antiferromagnetic skew scattering Hall effect in NdZn. J. Phys. Soc. Jpn. 55, 4417 (1986).
46
S. Ohara et al., Magnetic properties and de Haas-van Alphen effect of GdGa2. Phys. B 223 & 224, 379 (1996).
47
S. H. Liu, Exchange interaction between conduction electrons and magnetic shell electrons in rare-earth metals. Phys. Rev. 121, 451 (1961).
48
J. Kondo, Anomalous Hall effect and magnetoresistance of ferromagnetic metals. Prog. Theor. Phys. 27, 772 (1962).
49
F. D. M. Haldane, A geometrical description of the “classical” weak-field Hall effect in metals. arXiv [Preprint] (2005). https://doi.org/10.48550/arXiv.cond-mat/0504227.
50
O. Narikiyo, Fermi-surface curvature and Hall conductivity in metals. J. Phys. Soc. Jpn. 89, 124701 (2020).
51
N. P. Ong, Geometric interpretation of the weak-field Hall conductivity in two-dimensional metals with arbitrary Fermi surface. Phys. Rev. B 43, 193 (1991).
52
H. Harima, A. Yanase, Electronic structure and Fermi surface of LaGa2. J. Phys. Soc. Jpn. 60, 2718 (1991).
53
M. Sahakyan, V. H. Tran, Density functional theory study of electronic structure and optical properties of YGa2. Comp. Mater. Sci. 184, 109898 (2020).
54
I. Umehara, N. Nagai, Y. Onuki, Magnetoresistance and de Haas-van Alphen effect in CeGa2. J. Phys. Soc. Jpn. 60, 1464 (1991).
55
I. Umehara et al., Fermi surface and cyclotron mass in CeGa2. J. Magn. Magn. Mater. 104–107, 1407 (1992).
56
I. Sakamoto et al., Fermi surface and magnetic properties of antiferromagnetic SmGa2. J. Magn. Magn. Mater. 108, 125 (1992).
57
I. Sakamoto et al., De Haas-van Alphen effect and magnetic properties of antiferromagnetic SmGa2 and PrGa2. Phys. B 194–196, 1175 (1994).
58
A. Fert et al., Hall effect in dilute magnetic alloys. J. Magn. Magn. Mater. 24, 231 (1981).
59
B. Fauqué et al., Magnetoresistance of semimetals: The case of antimony. Phys. Rev. Mater. 2, 114201 (2018).
60
C. Collignon et al., Quasi-isotropic orbital magnetoresistance in lightly doped SrTiO3. Phys. Rev. Mater. 5, 065002 (2021).
61
J. C. W. Song, G. Rafael, P. A. Lee, Linear magnetoresistance in metals: Guiding center diffusion in a smooth random potential. Phys. Rev. B 92, 180204(R) (2015).
62
H. W. Meul et al., Transport properties and the nature of the low-temperature phase of EuMo6S8. Phys. Rev. B 26, 6431 (1982).
63
Z. Zhu et al., Emptying dirac valleys in bismuth using high magnetic fields. Nat. Commun. 8, 15297 (2017).
64
L. M. Falicov, P. R. Sievert, Magnetoresistance and magnetic breakdown. Phys. Rev. Lett. 12, 558 (1964).
65
R. D. dos Reis et al., Anisotropic magnetocaloric effect in ErGa2 and HoGa2 single-crystals. J. Alloys Comp. 582, 461 (2014).
66
T. Kurumaji et al., Replication Data for: Metamagnetic multiband Hall effect in Ising antiferromagnet ErGa2. Harvard Dataverse. https://doi.org/10.7910/DVN/UQ2T5V. Deposited 11 May 2024.
67
Y. Ōnuki et al., Anomalous Hall effect in rare earth antiferromagnets with the hexagonal structures. New Phys.: Sae Mulli. 73, 1054–1057 (2023).
Information & Authors
Information
Published in
Classifications
Copyright
Copyright © 2024 the Author(s). Published by PNAS. This article is distributed under Creative Commons Attribution-NonCommercial-NoDerivatives License 4.0 (CC BY-NC-ND).
Data, Materials, and Software Availability
Experimental data on electronic and structural properties. Data have been deposited in Harvard Dataverse (https://doi.org/10.7910/DVN/UQ2T5V) (66).
Submission history
Received: October 22, 2023
Accepted: April 13, 2024
Published online: May 28, 2024
Published in issue: June 4, 2024
Keywords
Acknowledgments
We thank Y. Nakamura and H. Sawa for supporting synchrotron XRD experiments. The synchrotron radiation experiments were performed at SPring-8 with the approval of the Japan Synchrotron Radiation Research Institute (Proposal No. 2023A1882). This research is funded in part by the Gordon and Betty Moore Foundation Emergent Phenomena in Quantum Systems (EPiQS) through Grants GBMF9070 to J.G.C. (material synthesis, DFT calculations), NSF grant DMR-1554891 (material design), ONR Grant N00014-21-1-2591 (instrumentation development), and AFOSR grant FA9550-22-1-0432 (advanced characterization). T.K. acknowledges the support by the Yamada Science Foundation Fellowship for Research Abroad and Japan Society for the Promotion of Science (JSPS) Overseas Research Fellowships. S.F. acknowledges support from a Rutgers Center for Material Theory Distinguished Postdoctoral Fellowship. L.Y. acknowledges the support by the Science and Technology Center (STC) for Integrated Quantum Materials, NSF grant number DMR-1231319, the Heising-Simons Physics Research Fellow Program, and the Tsinghua Education Foundation.
Author contributions
T.K. and J.G.C. designed research; T.K., S.F., L.Y., and S.K. performed research; T.K. analyzed data; and T.K., S.F., L.Y., S.K., and J.G.C. wrote the paper.
Competing interests
The authors declare no competing interest.
Notes
This article is a PNAS Direct Submission.
Authors
Metrics & Citations
Metrics
Altmetrics
Citations
Cite this article
Metamagnetic multiband Hall effect in Ising antiferromagnet ErGa2, Proc. Natl. Acad. Sci. U.S.A.
121 (23) e2318411121,
https://doi.org/10.1073/pnas.2318411121
(2024).
Copied!
Copying failed.
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.