J. Phys. Soc. Jpn. 91, 083704 (2022) [5 Pages]
LETTERS

First Observation of the de Haas–van Alphen Effect and Fermi Surfaces in the Unconventional Superconductor UTe2

+ Affiliations
1IMR, Tohoku University, Oarai, Ibaraki 311-1313, Japan2Advanced Science Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan3Faculty of Engineering, Niigata University, Ikarashi, Niigata 950-2181, Japan4Department of Physics, Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan5Graduate School of Science, Kobe University, Kobe 657-8501, Japan6Univ. Grenoble Alpes, CEA, Grenoble INP, IRIG, PHELIQS, F-38000 Grenoble, France

We report the first observation of the de Haas–van Alphen (dHvA) effect in the novel spin-triplet superconductor UTe2 using high quality single crystals with a high residual resistivity ratio (RRR) over 200. The dHvA frequencies, named α and β, are detected for the field directions between c- and a-axes. The frequency of branch β increases rapidly with the field angle tilted from c- to a-axis, while branch α splits, owing to the maximal and minimal cross-sectional areas from the same Fermi surface. Both dHvA branches, α and β reveal two kinds of cylindrical Fermi surfaces with a strong corrugation at least for branch α. The angular dependence of the dHvA frequencies is in very good agreement with that calculated by the generalized gradient approximation (GGA) method taking into account the on-site Coulomb repulsion of U = 2 eV. It indicates that the main Fermi surfaces are experimentally detected. The observed cyclotron effective masses are large in the range from 32 to 57 m0. They are approximately 10–20 times lager than the corresponding band masses, consistent with the mass enhancement obtained from the Sommerfeld coefficient, γ and the calculated density of states at the Fermi level. The local density approximation (LDA) calculations of ThTe2 assuming U4+ with the 5f2 localized model are in less agreement with our experimental results, in spite of the prediction for two cylindrical Fermi surfaces, suggesting a mixed valence states of U4+ and U3+ in UTe2.

©2022 The Author(s)
This article is published by the Physical Society of Japan under the terms of the Creative Commons Attribution 4.0 License. Any further distribution of this work must maintain attribution to the author(s) and the title of the article, journal citation, and DOI.

UTe2 attracts much attention because of the unusual superconducting properties due to the spin-triplet state.13) UTe2 is a heavy fermion paramagnet with a Sommerfeld coefficient \(\gamma\sim 120\) mJ K−2 mol−1. It crystallizes in the body-centered orthorhombic structure with the space group, \(Immm\) (No. 71, \(D_{2h}^{25}\)), where the U atom forms a two-leg ladder structure along the a-axis. Superconductivity occurs below \(T_{\text{c}}=1.5{\text{–}}2\) K with a large specific heat jump.4) A highlight is the huge upper critical field \(H_{\text{c2}}\) with a field-reentrant behavior for \(H\parallel\text{$b$-axis}\) (hard-magnetization axis). Superconductivity survives up to the first-order metamagnetic transition at \(H_{\text{m}}=350\) kOe,5) as a bulk property.6) The values of \(H_{\text{c2}}\) highly exceed the Pauli limit, (∼30 kOe), for all field directions, suggesting a spin-triplet state. A microscopic evidence for a spin-triplet state is obtained from NMR experiments, in which the Knight shift is unchanged or decreases very slightly below \(T_{\text{c}}\) for \(H\parallel a\), b, and c-axes.7) Another highlight is the appearance of the multiple superconducting phases under pressure.8,9) Applying pressure, \(T_{\text{c}}\) starts decreasing, and splits at ∼0.3 GPa. The lower \(T_{\text{c}}\) decreases continuously, while the higher \(T_{\text{c}}\) increases up to 3 K around 1 GPa and decreases rapidly. At the critical pressure \(P_{\text{c}}\sim 1.5\) GPa, superconductivity is suppressed and magnetic order, most likely aniferromagnetic appears above \(P_{\text{c}}\).811) Under magnetic field, multiple superconducting phases show a remarkable field response, displaying a sudden increase of \(H_{\text{c2}}\) at low temperatures.9,12) These multiple superconducting phases are the hallmarks of a spin-triplet state with different superconducting order parameters related to the spin and orbital degree of freedom.

It was first pointed out1) that UTe2 is located at the verge of the ferromagnetic order and resembles ferromagnetic superconductors.13,14) However, no ferromagnetic fluctuations are experimentally established; instead antiferromagnetic fluctuations with an incommensurate wave vector are detected in inelastic neutron scattering experiments.15,16) Furthermore, above \(P_{\text{c}}\), antiferromagnetic order is confirmed directly by magnetic susceptibility measurements.17) These results suggest that multiple fluctuations, such as ferromagnetic, antiferromagnetic fluctuations, and valence instabilities,10,18) exist in UTe2, playing important roles for the unusual superconducting properties.

The electronic structure of UTe2 has been investigated by angle-resolved photoemission spectroscopy (ARPES) at 20 K. The results obtained from the soft X-ray19) and the vacuum ultraviolet synchrotron radiation20) are contradicting, probably related to the different inelastic mean free path of the photoelectrons. The fine structure near the Fermi level was unresolved in the soft X-ray experiments. On the other hand, the high resolution ARPES experiments revealed two light quasi-one dimensional bands and one heavy band at the Fermi level, which is, however, inconsistent with the soft X-ray ARPES experiments. Thus, no clear conclusion on the electronic structure emerges up to now. The determination of the Fermi surface topology at low temperatures through the direct observation of quantum oscillations is highly desired, which will be a key experiment to investigate the topological superconducting phenomena as well.

In order to clarify the electronic structure, we performed de Haas–van Alphen (dHvA) experiments on new high quality single crystals. Clear dHvA oscillations were successfully detected, and the angular dependence of the dHvA frequencies reveals two kinds of cylindrical Fermi surfaces. The results are well explained by the generalized gradient approximation (GGA) calculation with the on-site Coulomb repulsion, \(U=2\) eV. The detected large cyclotron effective masses are consistent with the Sommerfeld coefficient.

High quality single crystals of UTe2 were grown at Oarai (sample #1) and Tokai (sample #2). The details of single crystal growth will be published elsewhere.21) The dHvA experiments were performed at low temperatures down to 60 mK and at high fields up to 147 kOe, as well as resistivity, specific heat and AC susceptibility measurements.22) The band calculations were done by the GGA+U method in UTe223,24) and the local density approximation (LDA) method in ThTe2 as a reference.25)

First we present the superconducting properties of our high quality single crystals. Figure 1(a) shows the temperature dependence of the resistivity for the current along the a-axis. The superconducting transition at \(T_{\text{c}}=2.06\) K defined by zero resistivity is very high and very sharp. The resistivity follows the \(T^{2}\) dependence at low temperatures below 3.5 K. The residual resistivity \(\rho_{0}\) and the residual resistivity ratio RRR (\(\equiv\rho_{\text{300K}}/\rho_{0}\)) are 1.7 µΩ·cm and 220, respectively.


Figure 1. (Color online) (a) Temperature dependence of the resistivity for the current along a-axis in UTe2 (sample #1). The residual resistivity ratio (RRR) is 220. The dotted line is the results of fitting. (b) Temperature dependence of the electronic specific heat in the form of \(C_{\text{e}}/T\) vs T in UTe2 (sample #1). The dotted line is the results of fitting between 0.34 and 0.6 K assuming \(C_{\text{e}}/T =\gamma_{0} + B T^{2}\).

Figure 1(b) shows the temperature dependence of the electronic specific heat in the form of \(C_{\text{e}}/T\) vs T for sample #1 after subtracting the phonon contribution. A sharp and large single-jump at \(T_{\text{c}}=2.05\) K with the width of 0.04 K and \(\Delta C_{\text{e}}/(\gamma T_{\text{c}}) = 2.64\) associated with a small residual γ-value (\(\gamma_{0}\)), which is only 3% of the normal state γ-value (\(\gamma_{\text{N}}\)), is observed. The small \(\gamma_{0}\) and high \(T_{\text{c}}\) in sample #1 are compared to those in different quality samples.26) All these properties indicate the high quality of our dHvA samples.

Figure 2 shows the anisotropy of \(H_{\text{c2}}\) at 70 mK for the field directions from c- to a-axis, and from a- to b-axis using the dHvA sample (#2) with \(T_{\text{c}}=2.05\) K. Because of the high \(T_{\text{c}}\), \(H_{\text{c2}}\) shifts to the higher field, compared to the previous results.2) \(H_{\text{c2}}\) for a-axis reaches 118 kOe, and \(H_{\text{c2}}\) for c-axis exceeds our highest field 147 kOe, most likely around 160 kOe. These high \(H_{\text{c2}}\) values restrict our dHvA experiments, since the dHvA oscillations appears above \(H_{\text{c2}}\) as shown later. The unusual minima were found around 50 deg tilted from c- to a-axis, and 20 deg from a- to b-axis, associated with a sharp maximum for \(H\parallel\text{$a$-axis}\). This is probably a mark of the field variation of the pairing strength along the a-axis, which should be clarified in the temperature dependence of \(H_{\text{c2}}\) with angular singularities in future experiments.


Figure 2. (Color online) Anisotropy of \(H_{\text{c2}}\) for the field direction from c- to a-axis, and from a- to b-axis at 70 mK determined by the AC susceptibility measurements in UTe2 (sample #2).

Next we show in Fig. 3 the dHvA oscillations at different field angles tilted from c- to a-axis. The clear dHvA signals were observed at the field angles between 11.8 and 56.8 deg above \(H_{\text{c2}}\) denoted by up-arrows. Even below \(H_{\text{c2}}\), the dHvA oscillations are observed, but the amplitudes are strongly damped. At higher field angles close to the a-axis, no dHvA oscillations were detected.


Figure 3. (Color online) dHvA oscillations at 70 mK at different field angles with the 4.5 deg step tilted from c- to a-axis in UTe2 (sample #2). Small up-arrows indicate \(H_{\text{c2}}\).

Figure 4 shows the typical dHvA oscillations and the corresponding FFT spectrum at the field angle of 26 deg tilted from c- to a-axis. Four dHvA frequencies, named \(\alpha_{1}\), \(\alpha_{1}^{\prime}\), β, and \(\beta^{\prime}\) were detected.


Figure 4. (Color online) Typical dHvA oscillations and the FFT spectrum in the field range between 110 and 147 kOe at the field direction tilted by 26 deg from c- to a-axis in UTe2 (sample #1).

Figure 5(a) shows the angular dependence of the dHvA frequencies from c- to a-axis. The results are obtained using two different samples, #1 and #2. The sample #1 was rotated from \(H\parallel a\)- to c-axis, while the sample #2 was rotated from \(H\parallel c\)- to a-axis, in order to clarify the dHvA signals for both \(H\parallel a\)- and c-axes. Both results are in good agreement with high reproducibility. With increasing the field angle, branch \(\alpha_{1}\) slightly decreases first, then increases up to the field angle 50 deg. Branch β increases continuously up to 45 deg. Branch \(\alpha_{2}\) also increases with the field angle, and shows a sharp increase around 60 deg. The highest frequency reaches more than \(1\times 10^{8}\) Oe, indicating a large cyclotron orbit. No dHvA signal was detected around \(H\parallel\text{$a$-axis}\), which is also confirmed in the field directions from a- to b-axis.27)


Figure 5. (Color online) (a) Angular dependence of the dHvA frequency in UTe2. Two samples, #1 (square) and #2 (circle) were used for \(H\parallel a\to c\) and \(H\parallel c\to a\), respectively. The lines are guides to the eyes. Panels (b) (c) show the angular dependences of the dHvA frequency calculated by the GGA+U (\(U=2\) eV) method23) in UTe2, and the LDA method in ThTe2,25) assuming the tetravalent U atom with the localized \(5f^{2}\) configuration in UTe2, respectively. The corresponding Fermi surfaces are depicted.

In order to determine the cyclotron effective masses, the dHvA oscillations were measured at different temperatures.28) The results are summarized in Table I. The detected effective masses are very large in the range from 32 to \(57\,m_{0}\), indicating a direct evidence for a heavy electronic state from a microscopic point of view.

Data table
Table I. Experimental dHvA frequency F, cyclotron effective mass \(m_{\text{c}}^{\ast}\), calculated dHvA frequency \(F_{\text{b}}\) and band mass \(m_{\text{b}}\) on the basis of the GGA+U (\(U=2\) eV) at the field angles tilted by 26, 29, 44, and 61 deg from c- to a-axis in UTe2.

The Dingle temperature, \(T_{\text{D}}\), was also derived from the field dependence of the dHvA amplitude. At the field angle of 26 deg, \(T_{\text{D}}\) for branch \(\alpha_{1}\) is 0.16 K. From the simple relations, \(F=\hbar c/(2\pi e) S_{\text{F}}\), \(S_{\text{F}}=\pi k_{\text{F}}^{2}\), \(m_{\text{c}}^{\ast} v_{\text{F}} =\hbar k_{\text{F}}\), \(T_{\text{D}}=\hbar/(2\pi k_{\text{B}}\tau)\), and \(l = v_{\text{F}}\tau\), where \(S_{\text{F}}\), \(v_{\text{F}}\), and τ are the cross-sectional area of the Fermi surface, Fermi velocity, and scattering life time, respectively, we obtain the mean free path l as 850 Å, indicating the high quality of our sample.

The angular dependence of the dHvA frequencies are compared to those obtained from the calculations. Figures 5(b) and 5(c) are the results from the GGA+U (\(U=2\) eV) calculation for UTe223) and from the LDA calculation for ThTe2,25) respectively. The experimental results are fairly in good agreement with those of the GGA+U (\(U=2\) eV). Therefore we can assign the detected dHvA branches as follows. Branch β is ascribed to the electron Fermi surface with the cylindrical shape, giving rise to nearly the \(1/{\cos\theta}\) dependence by tilting the field angle, θ from c- to a-axis. Branches \(\alpha_{1}\) and \(\alpha_{2}\) originate from the same Fermi surface, that is a cylindrical hole Fermi surface. Since the Fermi surface is corrugated from the cylinder shape, the dHvA frequency splits into \(\alpha_{1}\) and \(\alpha_{2}\), which correspond to the minimal and maximal cross-sectional area, respectively, at low field angles, when the field is titled from c- to a-axis.

The results of the LDA calculations in ThTe2, which corresponds to U4+ with the localized \(5f^{2}\) configuration in UTe2, shows a less agreement with the experimental results.29) Nevertheless, two kinds of cylindrical Fermi surfaces are quite similar to those by GGA+U (\(U=2\) eV) as well as DFT with large U (\(U=7\) eV) calculations.30) Note that the conventional LDA calculation predicts a Kondo semiconductor in UTe2.2,25,31) Similarly, GGA23) and DFT30) calculations without U also shows a band gap at the Fermi energy. These band structures are totally inconsistent with the experimental results, indicating that the strong correlation should be taken into account in the calculations. Small pocket Fermi surfaces predicted by other calculations19,32) are also inconsistent with our dHvA experiments.

Assuming the two kinds of cylindrical Fermi surfaces, which occupy approximately 20% of volume for each in the Brillouin zone with the carrier compensation, one can roughly calculate the γ-value derived from each Fermi surface, from the following equation, \(\gamma = k_{\text{B}}^{2} V/(6\hbar^{2}) m_{\text{c}}^{\ast} k_{\text{z}}\).33) Here V is the molar volume and \(k_{\text{z}}\) is the length of Brillouin zone along c-axis. If we take \(m_{\text{c}}^{\ast} = 32\,m_{0}\) and \(48\,m_{0}\) for the hole and electron Fermi surfaces, respectively, the obtained γ-values are 40 and 60 mJ K−2 mol−1 for each. Thus, the total γ-value is 100 mJ K−2 mol−1, which roughly agrees with the value of \(\gamma\sim 120\) mJ K−2 mol−1 in the specific heat measurements, indicating that our dHvA experiment detects the main Fermi surfaces of UTe2.

The effective masses can be compared to the band masses from GGA+U (\(U=2\) eV) as shown in Table I. The band masses at the selected field angles are in the range from 2.5 to \(4.3\,m_{0}\), meaning that the mass enhancement, \(m_{\text{c}}^{\ast}/m_{\text{b}}\) are approximately 10–20. This is also consistent with the mass enhancement, \(\gamma/\gamma_{\text{b}}\) obtained from the specific heat and the calculated density of states at the Fermi level (\(\gamma_{\text{b}}=8.1\) mJ K−2 mol−1), where the electron correlation is taken into account.

In the LDA calculations for ThTe2 without the electron correlation, the band masses are much smaller. For instance, \(m_{\text{b}}=0.7\,m_{0}\) is derived for the electron Fermi surface at 26 deg, while \(m_{\text{c}}^{\ast} = 48\,m_{0}\) is obtained in the dHvA experiments. The mass enhancement, \(m_{\text{c}}^{\ast}/m_{\text{b}}\) is consistent with that obtained from the Sommerfeld coefficient, namely \(\gamma/\gamma_{\text{b}}\), in which \(\gamma_{\text{b}}\) is 1.7 mJ K−2 mol−1.

A question is whether the anisotropy of the resistivity for \(J\parallel a\), b, and c can be explained by these Fermi surfaces, because the resistivity for \(J\parallel \text{$c$-axis}\), \(\rho_{c}\) is only twice larger than that for \(J\parallel\text{$a$-axis}\), \(\rho_{a}\), and is comparable to that for \(J\parallel\text{$b$-axis}\), \(\rho_{b}\) at room temperature.34) At low temperature, the anisotropy increases, but not an order of magnitude. In \(\rho_{c}\), the rapid increase on cooling below 50 K with a maximum around \(T^{\ast}\sim 15\) K is observed, suggesting that the electronic state may change from 3D to 2D-like nature on cooling. This is also in agreement with the development of the low dimensional antiferromagnetic fluctuation, which only starts developing below 60 K.16)

It should be noted that we cannot exclude the existence of small pocket Fermi surfaces with heavy masses, which may induce a Lifshitz transition under magnetic field as proposed in thermopower experiments.35) This may also compromise with possible topological superconductivity.

In summary, the dHvA oscillations were detected for the first time in UTe2. The angular dependence of the dHvA frequencies from \(H\parallel c\)- to a-axis, are in good agreement with the results of GGA+U (\(U=2\) eV) based on the \(5f^{3}\) itinerant model, revealing two kinds of cylindrical Fermi surfaces from hole and electron bands. The detected hole Fermi surface shows a large corrugation from the cylindrical shape. On the other hand, the dHvA results are in less agreement with those of LDA calculations in ThTe2, which corresponds to U4+ with the \(5f^{2}\)-localized model. The detected cyclotron effective masses are quite large, indicating heavy electronic states, consistent with the γ-value of the specific heat. These suggest a mixed valence states of U4+ and U3+, as proposed in the core-level spectroscopy,36) which are sensitive to external parameters, such as pressure and field. A link between our dHvA results on small energy scale near the Fermi level and the high energy spectroscopy19,20,36) deserves to be clarified. Our results elucidate a key part to solve the UTe2 properties, as it happened when the Fermi surface was determined on high \(T_{\text{c}}\) superconductors37) and Ce-115 heavy fermion systems.38)

Acknowledgements

We thank Y. Ōnuki, S.-i. Fujimori, V. Mineev, Y. Tokunaga, M. Kimata, K. Ishida, K. Izawa, A. Miyake, J. P. Brison, D. Braithwaite, A. Pourret, I. Sheikin, and S. Fujimoto for fruitful discussion. This work was supported by KAKENHI (JP19H00646, JP20K20889, JP20H00130, JP20KK0061, JP22H04933), GIMRT (20H0406), ICC-IMR, and ANR (FRESCO).


References

  • 1 S. Ran, C. Eckberg, Q.-P. Ding, Y. Furukawa, T. Metz, S. R. Saha, I.-L. Liu, M. Zic, H. Kim, J. Paglione, and N. P. Butch, Science 365, 684 (2019). 10.1126/science.aav8645 CrossrefGoogle Scholar
  • 2 D. Aoki, A. Nakamura, F. Honda, D. Li, Y. Homma, Y. Shimizu, Y. J. Sato, G. Knebel, J.-P. Brison, A. Pourret, D. Braithwaite, G. Lapertot, Q. Niu, M. Vališka, H. Harima, and J. Flouquet, J. Phys. Soc. Jpn. 88, 043702 (2019). 10.7566/JPSJ.88.043702 LinkGoogle Scholar
  • 3 See review, for example, D. Aoki, J. P. Brison, J. Flouquet, K. Ishida, G. Knebel, Y. Tokunaga, and Y. Yanase, J. Phys.: Condens. Matter 34, 243002 (2022). 10.1088/1361-648X/ac5863 CrossrefGoogle Scholar
  • 4 P. F. S. Rosa, A. Weiland, S. S. Fender, B. L. Scott, F. Ronning, J. D. Thompson, E. D. Bauer, and S. M. Thomas, Commun. Mater. 3, 33 (2022). 10.1038/s43246-022-00254-2 CrossrefGoogle Scholar
  • 5 G. Knebel, W. Knafo, A. Pourret, Q. Niu, M. Vališka, D. Braithwaite, G. Lapertot, J.-P. Brison, S. Mishra, I. Sheikin, G. Seyfarth, D. Aoki, and J. Flouquet, J. Phys. Soc. Jpn. 88, 063707 (2019). 10.7566/JPSJ.88.063707 LinkGoogle Scholar
  • 6 A. Rosuel, C. Marcenat, G. Knebel, T. Klein, A. Pourret, N. Marquardt, Q. Niu, S. Rousseau, A. Demuer, G. Seyfarth, G. Lapertot, D. Aoki, D. Braithwaite, J. Flouquet, and J. P. Brison, arXiv:2205.04524. Google Scholar
  • 7 G. Nakamine, K. Kinjo, S. Kitagawa, K. Ishida, Y. Tokunaga, H. Sakai, S. Kambe, A. Nakamura, Y. Shimizu, Y. Homma, D. Li, F. Honda, and D. Aoki, Phys. Rev. B 103, L100503 (2021). 10.1103/PhysRevB.103.L100503 CrossrefGoogle Scholar
  • 8 D. Braithwaite, M. Vališka, G. Knebel, G. Lapertot, J. P. Brison, A. Pourret, M. E. Zhitomirsky, J. Flouquet, F. Honda, and D. Aoki, Commun. Phys. 2, 147 (2019). 10.1038/s42005-019-0248-z CrossrefGoogle Scholar
  • 9 D. Aoki, F. Honda, G. Knebel, D. Braithwaite, A. Nakamura, D. Li, Y. Homma, Y. Shimizu, Y. J. Sato, J.-P. Brison, and J. Flouquet, J. Phys. Soc. Jpn. 89, 053705 (2020). 10.7566/JPSJ.89.053705 LinkGoogle Scholar
  • 10 S. M. Thomas, F. B. Santos, M. H. Christensen, T. Asaba, F. Ronning, J. D. Thompson, E. D. Bauer, R. M. Fernandes, G. Fabbris, and P. F. S. Rosa, Sci. Adv. 6, eabc8709 (2020). 10.1126/sciadv.abc8709 CrossrefGoogle Scholar
  • 11 W. Knafo, M. Nardone, M. Vališka, A. Zitouni, G. Lapertot, D. Aoki, G. Knebel, and D. Braithwaite, Commun. Phys. 4, 40 (2021). 10.1038/s42005-021-00545-z CrossrefGoogle Scholar
  • 12 G. Knebel, M. Vališka, M. Kimata, F. Honda, D. Li, D. Braithwaite, G. Lapertot, W. Knafo, A. Pourret, J.-P. Brison, J. Flouquet, and D. Aoki, J. Phys. Soc. Jpn. 89, 053707 (2020). 10.7566/JPSJ.89.053707 LinkGoogle Scholar
  • 13 D. Aoki and J. Flouquet, J. Phys. Soc. Jpn. 81, 011003 (2012). 10.1143/JPSJ.81.011003 LinkGoogle Scholar
  • 14 D. Aoki, K. Ishida, and J. Flouquet, J. Phys. Soc. Jpn. 88, 022001 (2019). 10.7566/JPSJ.88.022001 LinkGoogle Scholar
  • 15 C. Duan, K. Sasmal, M. B. Maple, A. Podlesnyak, J.-X. Zhu, Q. Si, and P. Dai, Phys. Rev. Lett. 125, 237003 (2020). 10.1103/PhysRevLett.125.237003 CrossrefGoogle Scholar
  • 16 W. Knafo, G. Knebel, P. Steffens, K. Kaneko, A. Rosuel, J.-P. Brison, J. Flouquet, D. Aoki, G. Lapertot, and S. Raymond, Phys. Rev. B 104, L100409 (2021). 10.1103/PhysRevB.104.L100409 CrossrefGoogle Scholar
  • 17 D. Li, A. Nakamura, F. Honda, Y. J. Sato, Y. Homma, Y. Shimizu, J. Ishizuka, Y. Yanase, G. Knebel, J. Flouquet, and D. Aoki, J. Phys. Soc. Jpn. 90, 073703 (2021). 10.7566/JPSJ.90.073703 LinkGoogle Scholar
  • 18 A. Miyake, M. Gen, A. Ikeda, K. Miyake, Y. Shimizu, Y. J. Sato, D. Li, A. Nakamura, Y. Homma, F. Honda, J. Flouquet, M. Tokunaga, and D. Aoki, J. Phys. Soc. Jpn. 91, 063703 (2022). 10.7566/JPSJ.91.063703 LinkGoogle Scholar
  • 19 S.-i. Fujimori, I. Kawasaki, Y. Takeda, H. Yamagami, A. Nakamura, Y. Homma, and D. Aoki, J. Phys. Soc. Jpn. 88, 103701 (2019). 10.7566/JPSJ.88.103701 LinkGoogle Scholar
  • 20 L. Miao, S. Liu, Y. Xu, E. C. Kotta, C.-J. Kang, S. Ran, J. Paglione, G. Kotliar, N. P. Butch, J. D. Denlinger, and L. A. Wray, Phys. Rev. Lett. 124, 076401 (2020). 10.1103/PhysRevLett.124.076401 CrossrefGoogle Scholar
  • 21 H. Sakai, P. Opletal, Y. Tokiwa, E. Yamamoto, Y. Tokunaga, S. Kambe, and Y. Haga, to be published in Phys. Rev. Mater. Google Scholar
  • 22 (Supplemental Material) The details for experimental technique are provided in online. Google Scholar
  • 23 J. Ishizuka, S. Sumita, A. Daido, and Y. Yanase, Phys. Rev. Lett. 123, 217001 (2019). 10.1103/PhysRevLett.123.217001 CrossrefGoogle Scholar
  • 24 The dHvA frequency for the GGA+U was calculated by the Supercell K-space Extremal Area Finder program. See P. M. C. Rourke and S. R. Julian, Comput. Phys. Commun. 183, 324 (2012). 10.1016/j.cpc.2011.10.015 CrossrefGoogle Scholar
  • 25 H. Harima, JPS Conf. Proc. 29, 011006 (2020). 10.7566/JPSCP.29.011006 LinkGoogle Scholar
  • 26 (Supplemental Material) The relation between the residual γ-value and Tc is provided in online. Google Scholar
  • 27 (Supplemental Material) No dHvA oscillations close to the a-axis and the comparison to the GGA+U (U = 1.1 and 1.5 eV) are described in online. Google Scholar
  • 28 (Supplemental Material) The FFT spectra at different temperatures and the mass plot are provided in online. Google Scholar
  •   (29) Note that the LDA calculation in ThTe2 with the 5f-level shift upward 0.2 Ry shows almost the same angular dependence of the dHvA frequency as that without 5f-level shift. Google Scholar
  • 30 Y. Xu, Y. Sheng, and Y. F. Yang, Phys. Rev. Lett. 123, 217002 (2019). 10.1103/PhysRevLett.123.217002 CrossrefGoogle Scholar
  • 31 A. B. Shick and W. E. Pickett, Phys. Rev. B 100, 134502 (2019). 10.1103/PhysRevB.100.134502 CrossrefGoogle Scholar
  • 32 A. B. Shick, S.-i. Fujimori, and W. E. Pickett, Phys. Rev. B 103, 125136 (2021). 10.1103/PhysRevB.103.125136 CrossrefGoogle Scholar
  • 33 D. Aoki, P. Wiśniewski, K. Miyake, N. Watanabe, Y. Inada, R. Settai, E. Yamamoto, Y. Haga, and Y. Ōnuki, Philos. Mag. B 80, 1517 (2000). 10.1080/13642810008208582 CrossrefGoogle Scholar
  • 34 Y. S. Eo, S. R. Saha, H. Kim, S. Ran, J. A. Horn, H. Hodovanets, J. Collini, W. T. Fuhrman, A. H. Nevidomskyy, N. P. Butch, M. S. Fuhrer, and J. Paglione, arXiv:2101.03102. Google Scholar
  • 35 Q. Niu, G. Knebel, D. Braithwaite, D. Aoki, G. Lapertot, G. Seyfarth, J.-P. Brison, J. Flouquet, and A. Pourret, Phys. Rev. Lett. 124, 086601 (2020). 10.1103/PhysRevLett.124.086601 CrossrefGoogle Scholar
  • 36 S.-i. Fujimori, I. Kawasaki, Y. Takeda, H. Yamagami, A. Nakamura, Y. Homma, and D. Aoki, J. Phys. Soc. Jpn. 90, 015002 (2021). 10.7566/JPSJ.90.015002 LinkGoogle Scholar
  • 37 N. Doiron-Leyraud, C. Proust, D. LeBoeuf, J. Levallois, J.-B. Bonnemaison, R. Liang, D. A. Bonn, W. N. Hardy, and L. Taillefer, Nature 447, 565 (2007). 10.1038/nature05872 CrossrefGoogle Scholar
  • 38 H. Shishido, R. Settai, D. Aoki, S. Ikeda, N. Nakamura, T. Iizuka, Y. Inada, K. Sugiyama, T. Takeuchi, K. Kindo, Y. Haga, H. Harima, Y. Aoki, T. Namiki, H. Sato, and Y. Ōnuki, J. Phys. Soc. Jpn. 71, 162 (2002). 10.1143/JPSJ.71.162 LinkGoogle Scholar

Cited by

View all 35 citing articles

Molten Salt Flux Liquid Transport Method for Ultra Clean Single Crystals UTe2

043703, 10.7566/JPSJ.93.043703