Skip to main content
Science Advances logoLink to Science Advances
. 2021 Mar 26;7(13):eabf1467. doi: 10.1126/sciadv.abf1467

Colossal anomalous Nernst effect in a correlated noncentrosymmetric kagome ferromagnet

T Asaba 1,*, V Ivanov 2, S M Thomas 1, S Y Savrasov 2, J D Thompson 1, E D Bauer 1, F Ronning 3,*
PMCID: PMC7997519  PMID: 33771869

Uranium-based ferromagnet demonstrates notable ability to convert heat to electricity.

Abstract

The transverse voltage generated by a temperature gradient in a perpendicularly applied magnetic field, termed the Nernst effect, has promise for thermoelectric applications and for probing electronic structure. In magnetic materials, an anomalous Nernst effect (ANE) is possible in a zero magnetic field. We report a colossal ANE in the ferromagnetic metal UCo0.8Ru0.2Al, reaching 23 microvolts per kelvin. Uranium’s 5f electrons provide strong electronic correlations that lead to narrow bands, a known route to producing a large thermoelectric response. In addition, uranium’s strong spin-orbit coupling produces an intrinsic transverse response in this material due to the Berry curvature associated with the relativistic electronic structure. Theoretical calculations show that in UCo0.8Ru0.2Al at least 148 Weyl nodes, and two nodal lines, exist within 60 millielectron volt of the Fermi level. This work demonstrates that magnetic actinide materials can host strong Nernst and Hall responses due to their combined correlated and topological nature.

INTRODUCTION

Because electrons carry both heat and charge, the transport of one is necessarily related to the other. Thermoelectric applications aim to exploit this fact by converting heat or charge flow into electric voltages or temperature gradients (1, 2). In the presence of a magnetic field, transverse responses are possible. It has been argued that the transverse geometry allows one to convert heat into electrical energy much more efficiently than the conventional longitudinal geometry (1, 3, 4). In conventional materials, however, the amplitude of the transverse anomalous Nernst effect (ANE) has been too small to realize practical thermoelectric devices.

Recently, it has been recognized that the Berry curvature associated with the Bloch waves of electrons generates an additional and potentially very large transverse response in magnetic materials (59). This is exemplified by the full-Heusler ferromagnet Co2MnGa whose ANE at room temperature is 6 μV/K and reaches to a record 8 μV/K at T = 400 K (3). This value is larger than typical ferromagnets, and theoretical calculations reveal that the large ANE is a consequence of the intrinsic Berry curvature contribution. After the triumph of Co2MnGa, a natural question arises: Can one further enhance or maximize the size of the ANE for applications or due to fundamental constraints? Here, we examine a uranium compound, which has multiple favorable attributes for enhancing the Berry curvature and ANE, and we find an ANE, which is even four times larger than found in Co2MnGa.

The Nernst Effect is related to the thermoelectric (α¯) and conductivity (σ¯) tensors through the expression Sxy=Ey/xT=(σxxαxyσxyαxx)/(σxx2+σxy2). The transverse contribution to the thermoelectric and conductivity tensors is directly connected to the Berry curvature ΩB through

σxy=e2ħBZd3k(2π)3f(k)ΩB(k) (1)
αxy=ekBħBZd3k(2π)3s(k)ΩB(k) (2)

where kB is the Boltzmann’s constant, e is the charge of an electron, ħ is the Planck’s constant h divided by 2π, f(k) is the Fermi-Dirac distribution function, s(k) = − f(k) ln (f(k)) − (1 − f(k)) ln (1 − f(k)), and ∫BZ is the integration over the Brillouin zone. At low temperatures, these two quantities are related through the Mott formula (10, 11).

An examination of these equations reveals that large Berry curvatures and small Fermi energies EF will generate large Hall and Nernst effects, in general. As elaborated on below, to enhance the Berry curvature contribution to the ANE and anomalous Hall effect (AHE), we therefore look for a system with the following properties: (i) a large spin-orbit coupling (SOC), (ii) strong electronic correlations, (iii) ferromagnetic order, and (iv) a kagome lattice structure.

The Berry curvature depends strongly on the SOC. Increasing the strength of the SOC increases the potential occurrence of band inversions, which generate regions of potentially diverging Berry curvature in the Brillouin zone. Depending on the symmetry of the inverted bands, nodal lines and/or Dirac or Weyl nodes may appear. In general, the strength of the SOC increases with increasing atomic number, and hence, we look for materials near the bottom of the periodic table.

To maximize the effects of larger SOC requires multiple bands to reside in the vicinity of the Fermi energy. This can be accomplished through strong electronic correlations, which will shift multiple band crossing points closer to the Fermi energy, thereby enhancing the Berry curvature. Furthermore, the renormalized band structures due to electronic correlations lead to smaller effective Fermi energies, which is known to promote large Seebeck and Nernst responses (12, 13). The need for both strong correlations and large SOC leads us naturally to heavy fermion systems composed of 4f and 5f elements.

Without broken time-reversal symmetry, a transverse Hall or Nernst response is identically zero. Hence, in the absence of an applied magnetic field, magnetic order is required for a finite ANE or AHE. Although possible in a system with antiferromagnetic order (8, 9, 14, 15), the requisite absence of symmetries is guaranteed in a ferromagnetic system.

In a system with antiferromagnetic interactions, magnetic frustration present in a kagome lattice can lead to additional band renormalization. Although magnetic frustration is presumably absent in a ferromagnetic material, the unrenormalized hopping on a kagome lattice coincidentally produces one flat band. Perhaps this explains, in part, why large AHE and ANE responses have been recently observed in a number of kagome compounds (6, 1619).

In our exploration of materials, we found that the layered distorted kagome lattice system UCo1−xRuxAl in the ZrNiAl structure type (shown in Fig. 1A) satisfies all the above criteria. In addition, the crystal structure is noncentrosymmetric, which can generate pairs of Weyl nodes with opposite chiralities. By themselves, such Weyl nodes do not contribute to the AHE and ANE, but they can be further split by the magnetic order. Both parent compounds UCoAl and URuAl are paramagnetic, but much of their solid solution UCo1−xRuxAl is ferromagnetic (20, 21). Members of the entire alloy series show relatively large Sommerfeld coefficients γ of ∼50 mJ/mol K2 (2124), indicating that charge carriers are moderately heavy. Magnetic x-ray and photoemission measurements also certify the presence of electronic correlations (25, 26). Furthermore, previous work has shown substantial anomalous Hall conductivity (AHC) for dopings close to the quantum-phase transition (21). Here, we examine x = 0.2 with a critical temperature (Tc) of 56 K, close to the optimal value in the series.

Fig. 1. Crystal structure and theoretical calculations of UCo0.8Ru0.2Al.

Fig. 1

(A) Crystal structure of UCo1 − xRuxAl. Left: The unit cell of UCoAl viewed from the [111] direction. A Co atom is replaced by a Ru atom upon doping. Right: The top view of UCoAl. The solid bonds indicate the tilted kagome layer formed by uranium atoms. (B) Density of states (DOS) including partial DOS for UCo0.8Ru0.2Al. (C) Nodal lines and Weyl points within ±60 meV of the Fermi energy in UCo0.8Ru0.2Al. For clarity, Weyl pairs with a kz separation < 0.2 2π/c are not shown. (D) Band plots for examples of (i) type I, (ii) type II, and (iii) critically tilted Weyl points (red circles), identified with arrows in (C), plotted on the interval −2kW < kz < 2kW, where 2kW is the Weyl pair separation along the kz direction.

RESULTS

Our spin-polarized theoretical calculations including SOC support our initial conjecture. A partial density of states of UCo0.8Ru0.2Al shown in Fig. 1B reveals a narrow band of uranium f-states dominating the electronic structure near EF. Electronic correlations are expected to further reduce this bandwidth (24). The magnitude of the calculated AHC due to the Berry curvature is found to be ∼2000 ohm−1cm−1, although the value as well as sign, is highly sensitive to details of the electronic structure (24). The origin of this large AHC is due to the dominant flat 5f uranium bands creating a remarkable number of topological features close to EF. Figure 1C shows the location of a subset of the Weyl nodes in the Brillouin zone that are within ±60 meV of the Fermi level. There are at least 148 Weyl points, which is, by far, larger than other Weyl materials. There exist both type I and type II Weyl nodes, as well as a few that are proximate to a Lifshitz transition between the two types (see Fig. 1D). Such a “quantum-critical” Weyl node was argued to be the origin of the large Nernstresponse in Co2MnGa (3). We additionally find two topological nodal lines close to EF within the σz mirror plane (Fig. 1C). These nodal lines arise at the crossing of bands belonging to distinct irreducible representations of the mirror plane point group and are protected against perturbations by the existence of mirror symmetry. While these nodal lines are substantial sources of Berry curvature, they will not contribute to Ωxyz. We note that first-principles calculations of the electronic structure of uranium compounds is notoriously challenging, but the main observation of an incredibly large number of Weyl nodes and Berry curvature found here is independent of the details of the electronic structure [see (24)].

The temperature dependence of the longitudinal resistivity ρxx and the Seebeck coefficient Sxx of a single crystal of UCo0.8Ru0.2Al is shown in Fig. 2. A kink is observed in both ρxx and Sxx at the ferromagnetic Tc = 56 K. The sample is a bad metal as reflected by the large residual resistivity (105 μ ohm cm). A longitudinal conductivity σxx of ∼9500 ohm−1cm−1, which falls within the moderately dirty regime where σxx = 3 × 103 −5 × 105 ohm−1cm−1, is the first indication that the transverse transport will be dominated by the intrinsic contribution (27). In the T = 0 limit, the Sxx/T gives a value of 0.9 μV/K2. This suggests a Fermi temperature of ∼470 K (12), consistent with a Fermi temperature of 930 K estimated by specific heat (24). We also note that the ratio of the experimental Sommerfeld coefficient to the computed value based on density functional theory implies a mass renormalization of ∼3 (24).

Fig. 2. Longitudinal transport and magnetization properties of single crystalline UCo0.8Ru0.2Al.

Fig. 2

(A) Electrical resistivity ρ versus temperature T with the current Ia. The vertical solid line indicates the ferromagnetic critical temperature. (B) Seebeck coefficient Sxx versus T with the heat current Ja. (C) Temperature dependence of the field-cooled (FC) and zero-field-cooled (ZFC) magnetization M per formula unit for Bc (red) and Ba (blue) in an applied field B = 0.1 T. (D) Field dependence of the magnetization M(H) at T = 2 K.

The temperature- and field-dependent magnetization of UCo0.8Ru0.2Al is shown in Fig. 2. The system shows very strong Ising-type magnetic anisotropy, with the easy axis along the c axis. The saturation magnetic moment of 0.59 μB/U is consistent with a previous study (20) and implies that the uranium 5f electrons are rather strongly hybridized with ligand electrons. A small field of B = 0.1 T is enough to polarize the magnetic moments during field cooling (FC). Thus, we take the transverse responses (Hall, Nernst, and Righi-Leduc effects) with FC and B = 0.1 T conditions, which is small enough to render the ordinary contributions as negligible.

The field and temperature dependence of the Hall, Nernst, and Righi-Leduc effects are shown in Fig. 3. For the field dependence, they all show a step-like behavior similar to the magnetization. The saturated response at high fields in the ordered state confirms that the transverse effects are dominated by the anomalous contribution, and the ordinary component is negligibly small. Consequently, the transverse response discussed in the remainder of the paper implicitly refers to the anomalous component. The vanishing transverse response as the temperature exceeds Tc is also consistent with this picture.

Fig. 3. Colossal anomalous transverse responses from UCo0.8Ru0.2Al with B∥c.

Fig. 3

(A and B) Hall effect ρxy as a function of magnetic field B and temperature T, respectively. (C and D) Nernst effect Sxy as a function of B and T, respectively. (E and F) Thermal Hall (Righi-Leduc) effect κxy as a function of B and T, respectively. The red dashed line in (F) is the expected thermal Hall conductivity based on the electrical Hall conductivity if the Wiedemann-Franz law is satisfied. All temperature-dependent data were taken with B = 0.1 T.

The anomalous Hall effect of UCo0.8Ru0.2Al is found to be very large, as shown in Fig. 3 (A and B). At T = 40 K, ρxy is −16 μ ohm cm, which decreases to −11 μ ohm cm at T = 2 K, corresponding to a conductivity of σxy = 980 ohm−1 cm−1. These values are similar to that of Co2MnGa and Co3Sn2S2 and among the largest values for ferromagnetic metals (3, 16).

Next, we display our primary result, a colossal ANE. While the anomalous Hall resistivity is as large as in Co2MnGa and Co3Sn2S2, more unexpectedly, the ANE is found to be colossal in UCo0.8Ru0.2Al, reaching 23 μV/K at T= 40 K (Fig. 3, C and D). To the best of our knowledge, this is approximately three to four times larger than the current record value of 6 to 8 μV/K held by Co2MnGa (3). Furthermore, in UCo0.8Ru0.2Al, the magnitude of Sxy at 40 K is comparable to the Seebeck coefficient. Consequently, the anomalous Nernst angle tan (θN) = ∣ Sxy/Sxx∣ reaches 0.92 at 41 K, where both Sxx and Sxy have extrema with values of −25.5 and 23.5 μV/K, respectively.

Figure 3 (E and F) displays the anomalous thermal Hall conductivity κxy of UCo0.8Ru0.2Al as a function of field and temperature. To check the validity of the anomalous Wiedemann-Franz (WF) law L0=κxyσxyT where L0=π2kB23e2 = 2.44 × 10−8 W ohm/K2, we compared κxy with L0σxyT (Fig. 3F). We find that they overlap each other very well, and thus, the WF law is valid. The validity of a transverse WF law indicates that the anomalous transverse entropy and charge flow are governed by the intrinsic contribution from the Berry curvature (8, 15).

DISCUSSION

The AHC as a function of T/Tc for a variety of materials with large transverse responses in heat and charge (15, 2830) is shown in Fig. 4A. The AHC of UCo0.8Ru0.2Al is almost temperature-independent below T = 0.5 Tc. This implies that the AHC is independent of scattering rate, which is an additional support that the intrinsic contribution dominates the transverse response in this system. At T = 2 K, ∣σxy∣ reaches 980 ohm−1 cm−1 in UCo0.8Ru0.2Al, which is comparable to Co2MnGa, one of the largest intrinsic AHC known. The AHC from SrRuO3 and Mn3Ge is almost one order of magnitude smaller than that of UCo0.8Ru0.2Al. The anomalous Peltier conductivity αxy as a function of T/Tc for these materials is shown in Fig. 4B. The Peltier coefficient is the sum of two contributions σxxSxx and σxySxx. In UCo0.8Ru0.2Al, σxxSxy ≫ σxySxx (see the Supplementary Materials). In sharp contrast to the fact that σxy of UCo0.8Ru0.2Al is comparable to that of Co2MnGa and Co3Sn2S2, αxy of UCo0.8Ru0.2Al is roughly three to four times larger than that of Co2MnGa, the current record holder for αxy, and Co3Sn2S2. αxy of SrRuO3 and Mn3Ge are notably smaller.

Fig. 4. Anomalous Hall and Peltier conductivity of various magnets.

Fig. 4

(A) AHC σij and (B) anomalous Peltier conductivity αij as a function of normalized temperature T/Tc (green line). For comparison, other materials with large anomalous contributions are also shown. (C) The maximum value of αij for various compounds compared with the strength of the SOC of the magnetic element (31). The compounds include UCo0.8Ru0.2Al (this work, green circle), Co2MnGa (blue diamond) (28), Co3Sn2S2 (purple star) (30), SrRuO3 (29) (cyan upper triangle), La0.7Sr0.3CoO3 (light green square) (29), Mn3Ge (red lower triangle) (15), and MnSi (orange hexagon) (49). (D) Plot of αijij versus T. Orange and pink dashed lines indicate kB/e and 2 kB/e, respectively.

Why is αij so large in UCo0.8Ru0.2Al? In Fig. 4C we plot the maximal value of ∣αij∣ versus λSOC for various compounds. λSOC is estimated from atomic calculations of the magnetic element (31, 32). Despite this very rough estimate for the strength of the SOC in these materials, Fig. 4C supports our original conjecture that a large SOC increases the possibility for more band inversions, which can create a large Berry curvature contribution to anomalous transverse response functions (33).

A large SOC alone, however, is insufficient to explain the remarkable properties of UCo0.8Ru0.2Al, because the maximal contribution from any pair of bands to the AHE is e2/hc, where c is the c-axis lattice constant, and in this limit, one expects a vanishing ANE. Instead, our electronic structure calculations shown in Fig. 1 reveal that there are many bands in the vicinity of the Fermi energy. This is a consequence of the nature of the 5f electrons in uranium. Furthermore, electronic correlations will further renormalize the bands leading to smaller effective Fermi energies, which is known to enhance thermoelectric properties (12). Notably, the Sommerfeld coefficient of the heat capacity γ is substantially larger for UCo0.8Ru0.2Al (41 mJ/mol K2) compared to Co2MnGa (12 mJ/mol K2) (3). Thus, it is the combination of the small effective Fermi energy and the large SOC that leads to multiple topological features within the energy window set by the SOC and that enables the remarkably large values of both σij and αij in UCo0.8Ru0.2Al.

It is interesting to ask which specific features of the electronic structure are most responsible for the large ANE in UCo0.8Ru0.2Al. UCo0.8Ru0.2Al crystallizes in the nonsymmorphic space group 189. From basic symmetry considerations, this space group has been argued to have a variety of interesting topological features, including Weyl nodes (3437), nodal lines (3841), and triple points (42, 43). All of these features appear in our electronic structure calculations (24). Notably, one set of critically tilted Weyl points lies 25 meV above Ef, corresponding to a doping of x ∼ 0.07 in the rigid band picture. For this set of Weyl nodes, we compute values of σxy ∼ 1200 ohm−1 cm−1 and αxy ∼ 20 A(K m)−1 at that energy, which may be shifted as a consequence of electronic correlations. Another possibility is that electronic correlations pin the Weyl nodes to the chemical potential as has been suggested in some Weyl-Kondo semimetals (44). Additional measurements, such as angle-resolved photoemission spectroscopy, will be required to confirm the identity of the most relevant topological features in UCo0.8Ru0.2Al.

Last, we consider the relationship between σij and αij. From the Mott relation, which is valid in the T = 0 limit, a large αij/T implies that σij can be enhanced by appropriately tuning the chemical potential. As temperature is increased, the relationship between σij and αij is less straight forward. Behnia et al. (28) have argued, however, that the ratio of αijij is bounded by the ratio of natural units kB/e = 86 μV/K, which reflects the fact that the Peltier coefficient represents the transport of entropy, while the Hall coefficient represents the transport of charge. For UCo0.8Ru0.2Al, αijij reaches 170 μV/K at T = 47 K, as shown in Fig. 4D. This supports the notion that multiple bands are simultaneously contributing to the large Nernst response in UCo0.8Ru0.2Al.

In conclusion, we have found a colossal ANE in UCo0.8Ru0.2Al. The simultaneous presence of strong SOC and electronic correlations drives narrow bands leading to a gigantic Nernst response and presents an exciting blueprint for how to develop magnetic topological metals for thermoelectric applications.

MATERIALS AND METHODS

A single crystal of nominal composition UCo0.8Ru0.2Al was grown by the Czochralski method in a tri-arc furnace. The crystal structure was confirmed to have the P6¯2m (# 189) space group with the ZrNiAl structure type. Powder and single-crystal diffraction was performed to confirm the crystal structure and the phase purity of the samples. The sample was oriented, and the quality was further checked by the x-ray Laue method with a photosensitive detector in backscattering geometry. The measured magnetic properties are in good agreement with previous studies, further confirming the sample quality.

Electronic structure calculations were performed using the local spin density approximation, yielding a ferromagnetic ground state, within the framework of the full potential linear muffin tin orbital method with SOC (45). This approximation has been previously used to study UCoAl (25, 46) and other uranium compounds in the same structural family (47). Experimental lattice constants were used (48), and doping dependence was handled using a rigid band approximation.

Weyl points and other topological features were located, and their topology was verified in a single-shot method (43). An initial coarse k-grid of 30 × 30 × 30 was used to locate candidate sources/sinks of Berry curvature flux, whose positions were iteratively refined by repeating the procedure on 2 × 2 × 2 grids within each k-cube. Additional details can be found in (24).

Transport, thermoelectric, and thermal transport measurements were performed in the same setup. A temperature gradient was generated by a 10-kilohm chip resistor attached to one end of the sample and measured by homemade thermocouples with the calibrated table made of Fe-doped gold and chromel. The other end of the sample was attached to a sapphire substrate that was physically clamped to an oxygen-free copper cold finger. Conventional steady-state measurements were performed to measure the thermoelectic power and thermal conductivity. The heating power was carefully monitored so that the thermal and thermoelectric responses were linear with the heating power.

Acknowledgments

We thank P. Rosa and X. Wan for useful discussions. Funding: The experimental work was performed at Los Alamos National Laboratory under the auspices of the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division. Theoretical calculations performed by V.I. and S.Y.S. are supported by NSF DMR grant no. 1832728. T.A. acknowledges support from a LANL LDRD Directors Postdoctoral Fellowship. Author contributions: T.A. and F.R. designed the experiment and analyzed the data. T.A., S.M.T., and F.R. built the transport measurement setup. T.A. performed the transport and heat capacity measurements. J.D.T. carried out the magnetic measurements. V.I. and S.Y.S. performed the theoretical calculations. E.D.B. prepared the sample. All authors contributed to writing the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the authors.

SUPPLEMENTARY MATERIALS

Supplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/7/13/eabf1467/DC1

REFERENCES AND NOTES

  • 1.Bell L. E., Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 321, 1457–1461 (2008). [DOI] [PubMed] [Google Scholar]
  • 2.He J., Tritt T. M., Advances in thermoelectric materials research: Looking back and moving forward. Science 357, eaak9997 (2017). [DOI] [PubMed] [Google Scholar]
  • 3.Sakai A., Mizuta Y. P., Nugroho A. A., Sihombing R., Koretsune T., Suzuki M.-T., Takemori N., Ishii R., Nishio-Hamane D., Arita R., Goswami P., Nakatsuji S., Giant anomalous Nernst effect and quantum-critical scaling in a ferromagnetic semimetal. Nat. Phys. 14, 1119–1124 (2018). [Google Scholar]
  • 4.Sakuraba Y., Hasegawa K., Mizuguchi M., Kubota T., Mizukami S., Miyazaki T., Takanashi K., Anomalous Nernst effect in L10-FePt/MnGa thermopiles for new thermoelectric applications. Appl. Phys. Express 6, 033003 (2013). [Google Scholar]
  • 5.Nagaosa N., Sinova J., Onoda S., MacDonald A. H., Ong N. P., Anomalous hall effect. Rev. Mod. Phys. 82, 1539–1592 (2010). [Google Scholar]
  • 6.Nakatsuji S., Kiyohara N., Higo T., Large anomalous Hall effect in a non-collinear antiferromagnet at room temperature. Nature 527, 212–215 (2015). [DOI] [PubMed] [Google Scholar]
  • 7.Nayak A. K., Fischer J. E., Sun Y., Yan B., Karel J., Komarek A. C., Shekhar C., Kumar N., Schnelle W., Kübler J., Felser C., Parkin S. S. P., Large anomalous Hall effect driven by a nonvanishing Berry curvature in the noncolinear antiferromagnet Mn3Ge. Sci. Adv. 2, e1501870 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Li X., Xu L., Ding L., Wang J., Shen M., Lu X., Zhu Z., Behnia K., Anomalous Nernst and Righi-Leduc effects in Mn3Sn: Berry curvature and entropy flow. Phys. Rev. Lett. 119, 056601 (2017). [DOI] [PubMed] [Google Scholar]
  • 9.Ikhlas M., Tomita T., Koretsune T., Suzuki M.-T., Nishio-Hamane D., Arita R., Otani Y., Nakatsuji S., Large anomalous Nernst effect at room temperature in a chiral antiferromagnet. Nat. Phys. 13, 1085–1090 (2017). [Google Scholar]
  • 10.Cutler M., Mott N. F., Observation of Anderson localization in an electron gas. Phys. Rev. 181, 1336–1340 (1969). [Google Scholar]
  • 11.Xiao D., Yao Y., Fang Z., Niu Q., Berry-phase effect in anomalous thermoelectric transport. Phys. Rev. Lett. 97, 026603 (2006). [DOI] [PubMed] [Google Scholar]
  • 12.Behnia K., Jaccard D., Flouquet J., On the thermoelectricity of correlated electrons in the zero-temperature limit. J. Phys. Condens. Matter 16, 5187 (2004). [Google Scholar]
  • 13.Behnia K., Aubin H., Nernst effect in metals and superconductors: A review of concepts and experiments. Rep. Prog. Phys. 79, 046502 (2016). [DOI] [PubMed] [Google Scholar]
  • 14.Wuttke C., Caglieris F., Sykora S., Scaravaggi F., Wolter A. U., Manna K., Süss V., Shekhar C., Felser C., Büchner B., Hess C., Berry curvature unravelled by the anomalous Nernst effect in Mn3Ge. Phys. Rev. B 100, 085111 (2019). [Google Scholar]
  • 15.Xu L., Li X., Lu X., Collignon C., Fu H., Koo J., Fauqué B., Yan B., Zhu Z., Behnia K., Finite-temperature violation of the anomalous transverse Wiedemann-Franz law. Sci. Adv. 6, eaaz3522 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Liu E., Sun Y., Kumar N., Muechler L., Sun A., Jiao L., Yang S.-Y., Liu D., Liang A., Xu Q., Kroder J., Süß V., Borrmann H., Shekhar C., Wang Z., Xi C., Wang W., Schnelle W., Wirth S., Chen Y., Goennenwein S. T. B., Felser C., Giant anomalous Hall effect in a ferromagnetic kagome-lattice semimetal. Nat. Phys. 14, 1125–1131 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Kida T., Fenner L., Dee A., Terasaki I., Hagiwara M., Wills A., The giant anomalous Hall effect in the ferromagnet Fe3Sn2––a frustrated kagome metal. J. Phys. Condens. Matter 23, 112205 (2011). [DOI] [PubMed] [Google Scholar]
  • 18.Ye L., Kang M., Liu J., von Cube F., Wicker C. R., Suzuki T., Jozwiak C., Bostwick A., Rotenberg E., Bell D. C., Fu L., Comin R., Checkelsky J. G., Massive Dirac fermions in a ferromagnetic kagome metal. Nature 555, 638–642 (2018). [DOI] [PubMed] [Google Scholar]
  • 19.Xu G., Lian B., Zhang S.-C., Intrinsic quantum anomalous hall effect in the kagome lattice Cs2LiMn3F12. Phys. Rev. Lett. 115, 186802 (2015). [DOI] [PubMed] [Google Scholar]
  • 20.Andreev A. V., Havela L., Sechovsk V., Bartashevich M. I., Šebek J., Dremov R. V., Kozlovskaya I. K., Ferromagnetism in the UCo1 − xRuxAl quasiternary intermetallics. Philos. Mag. B 75, 827–844 (1997). [Google Scholar]
  • 21.Pospíšil J., Opletal P., Vališka M., Tokunaga Y., Stunault A., Haga Y., Tateiwa N., Gillon B., Honda F., Yamamura T., Nižňanský V., Yamamoto E., Aoki D., Properties and Collapse of the Ferromagnetism in UCo1 − xRuxAl Studied in Single Crystals. J. Phys. Soc. Jpn. 85, 034710 (2016). [Google Scholar]
  • 22.Veenhuizen P. A., De Boer F. R., Menovsky A. A., Sechovsky V., Havela L., Magnetic properties of URuAl and URhAl single crystals. Le Journal de Physique Colloques 49, C8–485 (1988). [Google Scholar]
  • 23.Matsuda T. D., Sugawara H., Aoki Y., Sato H., Andreev A. V., Shiokawa Y., Sechovsky V., Havela L., Transport properties of the anisotropic itinerant-electron metamagnet UCoAl. Phys. Rev. B 62, 13852–13855 (2000). [Google Scholar]
  • 24.See the Supplementary Materials.
  • 25.Antonov V. N., Harmon B. N., Andryushchenko O. V., Bekenev L. V., Yaresko A. N., Electronic structure and x-ray magnetic circular dichroism in uranium compounds. II. UT Al (T= Co, Rh, and Pt) intermetallics. Phys. Rev. B 68, 214425 (2003). [Google Scholar]
  • 26.Fujimori S.-i., Takeda Y., Okane T., Saitoh Y., Fujimori A., Yamagami H., Haga Y., Yamamoto E., Ōnuki Y., Electronic structures of uranium compounds studied by soft x-ray photoelectron spectroscopy. J. Phys. Soc. Jpn. 85, 062001 (2016). [Google Scholar]
  • 27.Onoda S., Sugimoto N., Nagaosa N., Quantum transport theory of anomalous electric, thermoelectric, and thermal Hall effects in ferromagnets. Phys. Rev. B 77, 165103 (2008). [Google Scholar]
  • 28.Xu L., Li X., Ding L., Chen T., Sakai A., Fauqué B., Nakatsuji S., Zhu Z., Behnia K., Anomalous transverse response of Co2 MnGa and universality of the room-temperature αijA/σijA ratio across topological magnets. Phys. Rev. B 101, 180404 (2020). [Google Scholar]
  • 29.Miyasato T., Abe N., Fujii T., Asamitsu A., Onoda S., Onose Y., Nagaosa N., Tokura Y., Crossover behavior of the anomalous Hall effect and anomalous Nernst effect in itinerant ferromagnets. Phys. Rev. Lett. 99, 086602 (2007). [DOI] [PubMed] [Google Scholar]
  • 30.Ding L., Koo J., Xu L., Li X., Lu X., Zhao L., Wang Q., Yin Q., Lei H., Yan B., Zhu Z., Behnia K., Intrinsic anomalous Nernst effect amplified by disorder in a half-metallic semimetal. Phys. Rev. X 9, 041061 (2019). [Google Scholar]
  • 31.Shanavas K., Popović Z. S., Satpathy S., Theoretical model for Rashba spin-orbit interaction in d electrons. Phys. Rev. B 90, 165108 (2014). [Google Scholar]
  • 32.F. Herman, S. Skillman, Atomic Structure Calculations (Engelwood Cliffs, NJ, 1963). [Google Scholar]
  • 33.Xiao D., Chang M.-C., Niu Q., Berry phase effects on electronic properties. Rev. Mod. Phys. 82, 1959–2007 (2010). [Google Scholar]
  • 34.Huang S.-M., Xu S.-Y., Belopolski I., Lee C.-C., Chang G., Wang B., Alidoust N., Bian G., Neupane M., Zhang C., Jia S., Bansil A., Lin H., Hasan M. Z., A Weyl Fermion semimetal with surface Fermi arcs in the transition metal monopnictide TaAs class. Nat. Commun. 6, 7373 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Soluyanov A. A., Gresch D., Wang Z., Wu Q., Troyer M., Dai X., Bernevig B. A., Type-II weyl semimetals. Nature 527, 495–498 (2015). [DOI] [PubMed] [Google Scholar]
  • 36.Wan X., Turner A. M., Vishwanath A., Savrasov S. Y., Topological semimetal and Fermi-arc surface states in the electronic structure of pyrochlore iridates. Phys. Rev. B 83, 205101 (2011). [Google Scholar]
  • 37.Weng H., Fang C., Fang Z., Bernevig B. A., Dai X., Weyl semimetal phase in noncentrosymmetric transition-metal monophosphides. Phys. Rev. X 5, 011029 (2015). [Google Scholar]
  • 38.Burkov A. A., Hook M. D., Balents L., Topological nodal semimetals. Phys. Rev. B 84, 235126 (2011). [Google Scholar]
  • 39.Du Y., Tang F., Wang D., Sheng L., Kan E.-j., Duan C.-G., Savrasov S. Y., Wan X., CaTe: A new topological node-line and Dirac semimetal. npj Quantum Mater. 2, 3 (2017). [Google Scholar]
  • 40.Kim Y., Wieder B. J., Kane C. L., Rappe A. M., Dirac line nodes in inversion-symmetric crystals. Phys. Rev. Lett. 115, 036806 (2015). [DOI] [PubMed] [Google Scholar]
  • 41.Yu R., Weng H., Fang Z., Dai X., Hu X., Topological node-line semimetal and Dirac semimetal state in antiperovskite Cu3PdN. Phys. Rev. Lett. 115, 036807 (2015). [DOI] [PubMed] [Google Scholar]
  • 42.Zhu Z., Winkler G. W., Wu Q., Li J., Soluyanov A. A., Triple point topological metals. Phys. Rev. X 6, 031003 (2016). [Google Scholar]
  • 43.Ivanov V., Savrasov S. Y., Monopole mining method for high-throughput screening for Weyl semimetals. Phys. Rev. B 99, 125124 (2019). [Google Scholar]
  • 44.Lai H.-H., Grefe S. E., Paschen S., Si Q., Weyl–kondo semimetal in heavy-fermion systems. Proc. Natl. Acad. Sci. U.S.A. 115, 93–97 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Savrasov S. Y., Linear-response theory and lattice dynamics: A muffin-tin-orbital approach. Phys. Rev. B 54, 16470–16486 (1996). [DOI] [PubMed] [Google Scholar]
  • 46.Kučera M., Kuneš J., Kolomiets A., Diviš M., Andreev A. V., Sechovskỳ V., Kappler J.-P., Rogalev A., X-ray magnetic circular dichroism studies of 5 f magnetism in UCoAl and UPtAl. Phys. Rev. B 66, 144405 (2002). [Google Scholar]
  • 47.Gasche T., Auluck S., Brooks M. S. S., Johansson B., Theory of the magnetism of ternary uranium compounds. J. Magn. Magn. Mater. 104–107, 37–38 (1992). [Google Scholar]
  • 48.Lam D. J., Darby J. B. Jr., Downey J. W., Norton L. J., Equiatomic ternary compounds of uranium and aluminium with group viii transition elements. J. Nucl. Mater. 22, 22–27 (1967). [Google Scholar]
  • 49.Hirokane Y., Tomioka Y., Imai Y., Maeda A., Onose Y., Longitudinal and transverse thermoelectric transport in MnSi. Phys. Rev. B 93, 014436 (2016). [Google Scholar]
  • 50.Gasche T., Brooks M. S. S., Johansson B., Ground-state properties of ternary uranium compounds: I. hybridization effects. J. Phys. Condens. Matter 7, 9499 (1995). [Google Scholar]
  • 51.Gasche T., Brooks M., Johansson B., Ground-state properties of ternary uranium compounds: II. magnetic properties. J. Phys. Condens. Matter 7, 9511 (1995). [Google Scholar]
  • 52.Chang S., Nakotte H., Andreev A., Bordallo H., Havela L., Sechovsky V., Torikachvili M., Magnetism in some UCO1/3 T2/3 Al solid solutions (t= transition metal). J. Appl. Phys. 87, 6812–6814 (2000). [Google Scholar]
  • 53.Chang S., Nakotte H., Schultz A. J., Bordallo H. N., Torikachvili M. S., Sechovsky V., Andreev A. V., Crystallographic order and magnetism in UCO1/3T2/3Al compounds (T= Ru, Rh, Pt). Physica B Condens. Matter 276–278, 634–635 (2000). [Google Scholar]
  • 54.Andreev A. V., Sechovsky V., Havela L., Šebek J., Bartashevich M. I., Goto T., Kamishima K., Andreev D. A., Gaviko V. S., Dremov R. V., Kozlovskaya I. K., Onset of ferromagnetism between the paramagnets UCoAl and URuAl. Czech. J. Phys. 46, 3385–3386 (1996). [Google Scholar]
  • 55.V. Ivanov, X. Wan, S. Y. Savrasov, Correlation driven topological insulator-to-weyl semimetal transition in actinide system UNiSn. arXiv:1809.09543 [cond-mat.str-el] (25 September 2018).
  • 56.Zyuzin A. A., Tiwari R. P., Intrinsic anomalous hall effect in type-II weyl semimetals. JETP Letters 103, 717–722 (2016). [Google Scholar]
  • 57.Kadowaki K., Woods S., Universal relationship of the resistivity and specific heat in heavy-fermion compounds. Solid State Commun. 58, 507–509 (1986). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/7/13/eabf1467/DC1


Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES