Pressure-induced phase switching of Shubnikov de Haas oscillations in the molecular Dirac fermion system (BETS)2I3
Abstract
We report on the Shubnikov de Haas (SdH) oscillations in the quasi two-dimensional molecular conductor (BETS)2I3 [BETS: bis(ethylenedithio)tetraselenafulvalene] laminated on polyimide films at 1.7 K. From the SdH phase factor, we verified experimentally that the material is in the Dirac fermion phase under pressure. (BETS)2I3 is in the vicinity of the phase transition between strongly correlated insulating and Dirac fermion phases, and is a possible candidate for an ambient-pressure molecular Dirac fermion system. However, the SdH oscillations indicate that the Berry phase is zero at ambient pressure. Under pressure, a Berry phase emerges when the metal-insulator crossover is almost suppressed at 0.5 GPa. The results contrast those for the pioneering molecular Dirac fermion system (BEDT-TTF)2I3 [BEDT-TTF: bis(ethylenedithio)tetrathiafulvalene] in which Dirac fermions and semiconducting behavior are simultaneously observed.
Introduction
Dirac electrons in solids, which obey linear (pseudo-relativistic) dispersion relations, are one of the central issues in condensed matter physics, particularly since the experimental discovery of graphene Novoselov2004 . However, materials in which the Fermi energy () lies at the contact point are still few. Among them, the molecular Dirac fermion system (BEDT-TTF)2I3 has provided a unique platform for two-dimensional massless Dirac fermions Tajima2006 ; Kajita2014 . Unlike graphene, (BEDT-TTF)2I3 is a bulk quasi-two-dimensional material, in which is close to the contact points between highly tilted Dirac cones at non-symmetric k-points in the Brillouin zone Katayama2006 . The low Fermi velocity ( m/s) and the low damping of the Landau levels allow us to precisely investigate how the Landau levels form and separate towards the quantum limit Kobara2020 .


The massless Dirac fermion phase appears in the vicinity of a strongly correlated insulating phase by application of pressure above 1.5 GPa. Therefore, the interaction effect on the massless Dirac fermions in this system has also been of great interest, and peculiar phenomena, such as an anisotropic Dirac cone reshaping due to the tilt of the cone Hirata2016 and ferrimagnetic spin polarization due to short-range Coulomb interaction Hirata2017 , have been reported. In addition, a deviation from the Korringa law in NMR measurement suggests that the system is in the strong coupling regime that graphene cannot reach Hirata2017 . Probably because of these special situations, the insulating behavior and charge gap remain even in the massless Dirac fermion phase under high pressure Liu2016 ; Beyer2016 ; Uykur2019 . The short-range interaction effect may become even more significant in the vicinity of the correlated insulating phase. Recently, the quantum phase transition between the insulating phase and the massless Dirac fermion phase was reported Unozawa2020 . The Fermi velocity () decreases without creating a mass gap upon approaching the phase transition. Further detailed experiments around the phase transition in (BEDT-TTF)2I3 and its related materials will be interesting as there are no other massless Dirac fermion systems in such a strong electron correlation regime.
(BETS)2I3 Inokuchi1993 ; Inokuchi1995 , the selenium analog of (BEDT-TTF)2I3, may be an excellent platform to explore electronic states in the vicinity of the phase transition. It shows similar resistivity behavior to (BEDT-TTF)2I3, but the insulating phase can be suppressed under lower pressure (0.6 GPa), probably due to the large bandwidth Tajima2006 . The transport properties above 0.6 GPa are reminiscent of those of (BEDT-TTF)2I3 above 1.5 GPa. Therefore, the electronic state of (BETS)2I3 is considered similar to that of (BEDT-TTF)2I3 at approximately 0.9 GPa. Indeed, band calculations based on the crystal structure under high pressure indicate the presence of the Dirac cones in both (BEDT-TTF)2I3 and (BETS)2I3 Kondo2009 ; Alemany2012 (although Dirac and normal electrons coexist). However, in-depth verification of Dirac fermions with the quantum oscillation measurements has not been reported so far. Recently, first-principles calculations by multiple independent research groups indicate that (BETS)2I3 is a type-I Dirac fermion system even at ambient pressure Ohki2020 ; Tsumuraya2020 ; Kitou2020 . Those groups simultaneously suggest the possibility of different insulating mechanisms from (BEDT-TTF)2I3 (spin-orbit coupling by Kitou et al. Kitou2020 , and Coulomb interaction + spin-orbit coupling by Ohki et al. Ohki2020 ).
In this study, to uncover the presence of Dirac cones, with similarities to and differences from (BEDT-TTF)2I3 in the vicinity of the phase transition, we investigate the Shubnikov de Haas (SdH) oscillation in thin single crystals of (BETS)2I3 laminated on polyimide films using a similar experimental method to (BEDT-TTF)2I3 Tajima2013 . From these measurements, we verified that (BETS)2I3 is in the Dirac fermion phase under pressure. The period of the oscillation does not significantly change before and after the transition, indicating that (BETS)2I3 under pressure has no large Fermi surfaces (type-I Dirac fermion system). Under high pressure, (BETS)2I3 is in the Dirac fermion phase with approximately 20% lower Fermi velocity than that in similarly doped (BEDT-TTF)2I3 Unozawa2020 .
Methods
Polyimide films (CT4112, KYOCERA Chemical Corporation) were spin-coated on polyethylene terephthalate (PET) substrate (Teflex FT7, Teijin DuPont Films Japan Limited) and baked at 180 ∘C for 1 hour. We electrochemically synthesized a thin (100 nm) single crystal of (BETS)2I3 from a chlorobenzene solution (2% v/v methanol) of BETS Kato1991 and tetrabutylammonium triiodide by applying 5 A for 20 hours. The thin crystal was transferred into 2-propanol with a pipette and guided onto the substrate. After the substrate was removed from the 2-propanol and dried, the crystal naturally adhered to the substrate. The x-ray diffraction measurement is difficult because the crystal is thin and laminated on the noncrystalline polymer substrate. However, thanks to the polarizing property of I, optical images through a polarizer indicate that the crystal is a single crystal in which the two-dimensional conducting plane is parallel to the substrate Supplemental . The and axes tend to correspond to the diagonals of the crystal if the shape is close to diamond. Atomic force microscopy revealed that the surface roughness of the crystal was smaller than the thickness of the BETS conducting layer Supplemental .
Unlike (BEDT-TTF)2I3, no polymorphs of (BETS)2I3 have been reported, and the temperature dependence of the resistance is similar to the literatures Inokuchi1993 ; Inokuchi1995 (as shown later). We made electrical contacts with carbon paste and Au wires. Samples #1 and #2 were subsequently shaped into Hall bars using a pulsed laser beam with a wavelength of 532 nm (sample #3 and #4 were not shaped). The dimensions of samples #1-4 are 90 m (width) 180 m (length) 130 nm (thickness), 90 m 110 m 90 nm, 310 m 160 m 80 nm, and 130 m 130 m 125 nm, respectively.
For samples #1-3, we measured the longitudinal resistivity and the Hall resistivity using a dc current of 1 A from a dc source (KEITHLEY 2400, Keithley Instruments) and a nano voltmeter (Agilent 34420A, Agilent Technologies) in a cryostat with a superconducting magnet that generate up to 8 T (TeslatronPT, Oxford Instruments). For sample #4, a dc current of 10 A was applied from a dc source (KEITHLEY 6221, Keithley Instruments) and a nano voltmeter (Agilent 34420A, Agilent Technologies) in a He3 cryostat with a superconducting magnet that generate up to 10 T (Cryogenic Limited). The magnetic field was applied perpendicular to the substrate of the samples. For pressure measurements, we employed a typical CuBe pressure cell and Daphne 7373 oil. The pressures are values at room temperature, and the actual pressures at low temperatures are 0.10.2 GPa less than the notations Murata1997 .
Besides, the polymer film is not restricted to polyimide. We also observe similar SdH oscillations in (BETS)2I3 directly laminated on the PET substrate. However, we employed polyimide films because the oscillation signals tended to be more clear probably due to more clean surface conditions of our polyimide films.
Results and Discussion
The SdH oscillation is a powerful tool to investigate the Fermi surface and the Berry phase Mikitik1999 . Neither (BEDT-TTF)2I3 nor (BETS)2I3 shows the SdH oscillations in their bulk crystals regardless of pressure. We have to dope some carriers to observe the oscillations. Here, we synthesize thin single crystals of (BETS)2I3 and laminate them on polyimide films. The contact charging between (BETS)2I3 and polyimide induces hole doping, resulting in the observation of the SdH oscillations (one or two conducting layers are doped in the case of (BEDT-TTF)2I3 Tajima2013 ). According to the SdH oscillations period, the hole density is approximately 1012 cm-2, corresponding to 0.5% of the first Brillouin zone. Notice that the thin crystal consists of several tens of conducting BETS and insulating I3 layers (as shown in the Methods section), but the doped carriers are confined at the surface. Therefore, the sample resistance is the combined resistance of the nondoped bulk and doped surface, and is difficult to separate. Nevertheless, we can investigate the doped surface using the SdH oscillations because the nondoped bulk does not show the oscillations. The application of contact charging also causes unintended strain effects from the substrate. The crystal of the target material is much thinner than the substrate and tightly adheres to the substrate. Therefore, thermal and mechanical contractions (due to cooling and pressure) of the nondoped bulk ( nm) and the doped surface (a few nanometers) are governed by those of the substrate. These effects modify the effective pressure of the laminated crystal, as shown later. However, this effect does not change the essential pressure effect because the strain is biaxial and parallel to the conducting plane. If we employ an unshrinkable substrate such as Si, the shrinkable molecular crystal is broken under pressure (probably due to the Poisson effect). We employed shrinkable plastic substrates in this study.
The period and phase of the oscillations imply the following. At ambient pressure, the charge carriers are not Dirac fermions at the doping levels in this study. The charge carriers turn out to be Dirac fermions when the metal-insulator crossover is sufficiently suppressed by applying pressure. The phase switching contrasts the behavior in (BEDT-TTF)2I3 in the intermediate pressure region, which shows a Berry phase along with insulating behavior Unozawa2020 .
Ambient pressure
Figure 1(b) shows the temperature dependence of the resistivity at ambient pressure in sample #1 ((BETS)2I3/polyimide/PET). Compared with a bulk crystal Inokuchi1995 , the sample exhibits slightly lower metal-insulator crossover temperatures and more moderate resistivity increases at lower temperatures. The former is ascribable to the fact that the thermal contraction of the PET substrate applies compressive strain to (BETS)2I3 Kawasugi2008 (therefore, this sample at ambient pressure corresponds to a bulk crystal under weak pressure), and the latter is attributable to the doping effect of the polyimide layer. As the single crystal consists of several tens of conducting (BETS) layers and the doped carriers are confined at the interface, the doping effect appears only at low temperatures where the bulk is insulating. At 1.7 K, the sheet resistivity (resistance width length) is 8.4 103 .
Figure 1(c) shows the magnetoresistance (upper) and the Hall resistance (lower) at 1.7 K. The SdH oscillations along with negative magnetoresistance are visible. The magnetoresistance is complicated in detail. It is slightly negative up to 0.4 T, turns positive up to 1.5 T, and then becomes negative again by a further magnetic field. Such a negative magnetoresistance has not been observed in (BEDT-TTF)2I3 under low pressures Unozawa2020 . The magnetoresistance can be simply explained by neither the weak localization nor weak-antilocalization. It is reminiscent of the negative longitudinal magnetoresistance in topological semimetals Schumann2017 due to charge carrier density or mobility fluctuations. However, the magnetic field direction is different in this study (current is perpendicular to the magnetic field). Its origin cannot be clarified at this moment. We observe the negative magnetoresistance (without oscillations) in a bulk crystal Supplemental . Although we cannot see whether the doped interface also shows the negative magnetoresistance or not, the oscillations originate from the interface. We focus on the oscillation signals in this study.
The low-field Hall resistances are positive and proportional to the magnetic field. The sign becomes negative at around 23 K with increasing temperature Supplemental . By contrast, a bulk crystal shows negative Hall resistance at low temperatures Supplemental . Therefore, the doped carriers are holes and the concentration is 1012 cm-2 by ignoring electrons in bulk (Hall mobility cm2/Vs).
The quantum oscillations originate from the quantization condition for the energy levels of the electron Onsager1952 ; Lifshitz1956 :
(1) |
where is the area of the cyclotron orbit in k-space, is an integer, is the magnetic field, and is the phase factor. The oscillation signal is periodic against 1/, and the area can be estimated using the measurement . Assuming that the spin and valley degeneracies are both 2, the carrier density is
(2) |
The SdH oscillation is given by
(3) |
where and are the longitudinal resistance and the oscillation amplitude, respectively Sharapov2004 ; Lukyanchuk2004 ; Zhang2005 . The phase factor is associated with the Berry phase as
(4) |
In a conventional electron system with isolated bands, and . However, if the cyclotron orbit surrounds the contact point of the bands and the energy dispersions are linear in in the vicinity of the contact point, the Berry phase emerges and becomes zero Mikitik1999 . Accordingly, when we plot corresponding to the peaks against Landau level index (Landau fan diagram), the intercept 0 or 1/2 for 2D Dirac or normal electrons.
Nevertheless, we have to be careful about the phase analysis of the SdH oscillations. Eq. (3) assumes the condition (graphene, (BEDT-TTF)2I3, and many low-carrier-density semiconductors meet this condition), and the minima in coincide with those in the conductance . The condition may be violated due to low mobility or the presence of a highly conducting bulk transport channel. In the case that , the minima in correspond to the maxima in . The difficulty of the phase analysis using resistance data has been pointed out for topological insulators Xiong2012 ; Ando2013 . However, the Hall response is usually weak and the resistance oscillation is more apparent in many cases. One may still use the reversed resistance data when , as in the case of graphite Lukyanchuk2004 . Here, we analyze the oscillations of both and in (BETS)2I3/polyimide/PET at ambient pressure. To eliminate the background, we show the second derivative of the data with respect to the magnetic field.
Figure 1(d) shows the dependences of and derived from Fig. 1(c). They correspond to and , respectively. The oscillation signal is more apparent in the upper curve because . However, both curves show almost the same periods and phases, indicating . and estimated from the upper curve are 12.1 T and cm-2, respectively, and correspond to 0.6% hole doping provided that the doped carriers are confined within one conducting layer. The value provides a realistic Hall scattering factor of 1.70 (). is almost , indicating that the carriers are not Dirac fermions at this doping level. The same analysis of sample #2 is described in Fig. S4 Supplemental . The transport properties of sample #1, such as the temperature dependence of the resistance, the magnetoresistance, the sign and magnitude of the Hall effect, the relationship between resistance and conductance oscillations, and the phase factor reproduced in sample #2.
The leftmost peak in Fig. 1(d) (denoted by yellow diamonds) deviates from the position predicted from the fitting line in Fig. 1(e). Provided that this is a split peak as a result of the Zeeman effect, we estimate the effective mass from the relation , where and are the predicted and observed peaks, respectively.
Under pressure
With increasing pressure, the entire sample is compressed. The bandwidths of the bulk and surface are enhanced, and their resistances decrease. The metal-insulator crossover gradually diminishes and disappears at around 0.6 GPa, as shown in Fig. 2(a). The dip at around 35 K and the upturn below 5 K of the resistance have also been observed in bulk crystals Inokuchi1993 , but the detailed mechanisms are still unclear. Above 0.6 GPa, the resistivity is almost constant down to approximately 15 K, below which the metallic behavior of the doped holes appears. The sheet resistivity per conducting layer is close to the quantum resistance , as in the case of (BEDT-TTF)2I3. The negative magnetoresistance observed at ambient pressure diminishes and becomes positive (Fig. 2(b)). The Hall resistance also decreases and becomes nonlinear, probably due to the emergence of a conducting bulk transport channel (Fig. 2(c)). As stated above, the bulk crystal of (BETS)2I3 does not show the SdH oscillations even under pressure. We investigate the doped surface by the analysis of the oscillations.
Figure 3(a) shows the pressure dependence of . At 0.35 and 0.4 GPa, the minima give ; this tendency is also observed at ambient pressure. values are estimated to be and , respectively, showing a decreasing trend with pressure. At 0.45 GPa, we cannot construct a convincing fan diagram because of ambiguous oscillation signals and large background signals. However, we can see a half-period oscillation (up to T-1), probably indicating the coexistence of anti-phase oscillations. One possible scenario is the phase separation between the regions with =1/2 and 0. Above 0.5 GPa, becomes almost zero, implying the emergence of Dirac fermions, although the oscillations at low are not clear. The Landau fan diagrams and the pressure dependence of and are summarized in Fig. 3(b) and (c). The conductance oscillations and Landau fan diagrams of sample #3 at ambient pressure and 1.2 GPa are also shown in Fig. S5 Supplemental .
We cannot observe Dirac fermions unless the metal-insulator crossover is sufficiently suppressed by pressure. These results are in contrast to those for (BEDT-TTF)2I3, in which the Dirac fermions and the semiconducting behavior are simultaneously observed. Besides, we cannot confirm the coexistence of Dirac and normal electrons (which has been reported for (BEDT-TTF)2I3 Monteverde2013 ) in this study. The SdH oscillations survive beyond the pressure-induced transition and does not significantly vary with pressure. If large Fermi surfaces emerge by applying pressure, as predicted by band calculations in early reports Kondo2009 ; Alemany2012 , the doping effect (and accordingly the quantum oscillations) should be obscured by the dense carriers.
The most straightforward interpretation of the phase switching is that the pressure-induced resistive transition is a semiconductor-Dirac fermion system transition. Another possible scenario is the pressure-induced merging of the Dirac cones. In that case, the number and area of the Fermi surface generally change at the transition. However, Fig. 3 shows that the does not significantly change during the transition. To consider the merging of the Dirac cone as the origin of the transition, we need a model and conditions consistent with these measurements.
The SdH oscillations become more evident as pressure increases. Figure 4 shows the magnetotransport properties of sample #4 at 1.8 GPa and 0.5 K. Here, the minima of coincide with those of Supplemental . The minima indicate and the interval of the Zeeman splitting peaks (yellow diamonds in Fig. 4(b)) gives effective Fermi velocity 3.6 104 m/s, which is approximately 20% lower than that from the same analysis for (BEDT-TTF)2I3 Unozawa2020 ( is estimated from the relation , where and are the peak fields). The low indicates that the Dirac cone is more tilted or more blunted than (BEDT-TTF)2I3. However, we cannot determine the central origin because the estimated is average over the orbit in the reciprocal space Tajima2007 . The peak around further separates into two peaks (green squares in Fig. 4(b)). We assign the bottom between these peaks to the Zeeman splitting peak because a similar of 3.6 104 m/s is estimated from the bottoms denoted by red circles. Therefore, the small separation is considered a valley splitting. We roughly estimate the valley-splitting energy K using a similar analysis to , assuming that is proportional to the magnetic field. A relative permittivity of is derived from the relation , where is the distance between the Dirac cones in k-space (approximated by the inverse lattice constant). In bulk (BEDT-TTF)2I3, near the Dirac point is estimated to be from the interlayer magnetoresistance Tajima2010 . As the permittivity decreases near the Dirac point, a comparable value is expected at the Dirac point in (BETS)2I3.


Summary
We have investigated the pressure dependence of the magnetoresistance and the Hall effect in slightly hole-doped thin single crystals of (BETS)2I3 laminated on polyimide films, and verified that the material is in the Dirac fermion phase under pressure. We found a phase switching of the SdH oscillation near the pressure-induced metal-insulator crossover, unlike in (BEDT-TTF)2I3 in the vicinity of the phase transition. At ambient pressure, the system exhibits a metal-insulator crossover below 50 K, and the phase of the SdH oscillation at 1.7 K indicates . Under pressure, becomes zero above 0.5 GPa, whereas the metal-insulator crossover disappears at approximately 0.6 GPa. A half-period oscillation appears at the boundary ( GPa), although the oscillation signal is ambiguous. It may originate from the coexistence of the regions with normal and Dirac fermions. The pressure-induced phase switching of the SdH oscillation indicates the presence of a semiconducting phase with normal electrons next to a Dirac fermion phase in (BETS)2I3. In (BEDT-TTF)2I3, the Berry phase appears even in the highly resistive states, and such a trivial insulating phase has not been observed Unozawa2020 . Recently, Kitou et al. reported that (BETS)2I3 maintains the inversion symmetry below the metal-insulator crossover Kitou2020 , implying a different insulating mechanism from (BEDT-TTF)2I3. Ohki et al. suggested that the insulating phase is a spin-ordered massive Dirac electron phase where time-reversal symmetry is broken but spatial inversion and translational symmetries are conserved Ohki2020 . Tsumuraya et al. explained that the system is in the massless Dirac state but a gap opens as a result of the spin-orbit interaction Kitou2020 ; Tsumuraya2020 . However, we cannot confirm the presence of Dirac fermions at ambient pressure in this study. Under high pressure (1.8 GPa), (BETS)2I3 is a Dirac fermion system with of 3.6 104 m/s. At high magnetic fields, the valley splitting is observed in the SdH oscillation. The valley splitting energy is estimated to be K. Further study is required to clarify the electronic states of (BETS)2I3 which may provide a unique Dirac fermion system different from (BEDT-TTF)2I3.
Acknowledgments
We would like to acknowledge Teijin DuPont Films Japan Limited for providing the PET films. This work was supported by MEXT and JSPS KAKENHI (Grant Nos. JP16H06346, JP19K03730, and JP19H00891).
References
- (1) K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, Science 306, 666 (2004).
- (2) N. Tajima, S. Sugawara, M. Tamura, Y. Nishio, and K. Kajita, J. Phys. Soc. Jpn. 75, 051010 (2006).
- (3) K. Kajita, Y. Nishio, N. Tajima, Y. Suzumura, and A. Kobayashi, J. Phys. Soc. Jpn. 83, 072002 (2014).
- (4) S. Katayama, A. Kobayashi, and Y. Suzumura, J. Phys. Soc. Jpn. 75, 054705 (2006).
- (5) R. Kobara, S. Igarashi, Y. Kawasugi, R. Doi, T. Naito, M. Tamura, R. Kato, Y. Nishio, K. Kajita, and N. Tajima, J. Phys. Soc. Jpn. 89, 113703 (2020).
- (6) M. Hirata, K. Ishikawa, K. Miyagawa, M. Tamura, C. Berthier, D. Basko, A. Kobayashi, G. Matsuno, and K. Kanoda, Nat. Commun. 7, 12666 (2016).
- (7) M. Hirata, K. Ishikawa, G. Matsuno, A. Kobayashi, K. Miyagawa, M. Tamura, C. Berthier, and K. Kanoda, Science 358, 1403 (2017).
- (8) D. Liu, K. Ishikawa, R. Takehara, K. Miyagawa, M. Tamura, and K. Kanoda, Phys. Rev. Lett. 116, 226401 (2016).
- (9) R. Beyer, A. Dengl, T. Peterseim, S. Wackerow, T. Ivek, A. V. Pronin, D. Schweitzer, and M. Dressel, Phys. Rev. B 93, 195116 (2016).
- (10) E. Uykur, W. Li, C. A. Kuntscher, and M, Dressel, npj Quantum Materials 4, 19 (2019).
- (11) Y. Unozawa, Y. Kawasugi, M. Suda, H. M. Yamamoto, R. Kato, Y. Nishio, K. Kajita, T. Morinari, and N. Tajima, J. Phys. Soc. Jpn. 89, 123702 (2020).
- (12) M. Inokuchi, H. Tajima, A. Kobayashi, and H. Kuroda, Synth. Met. 56, 2495 (1993).
- (13) M. Inokuchi, H. Tajima, A. Kobayashi, T. Ohta, H. Kuroda, R. Kato, T. Naito, and H. Kobayashi, Bull. Chem. Soc. Jpn. 68, 547 (1995).
- (14) R. Kondo, S. Kagoshima, N. Tajima, and R. Kato, J. Phys. Soc. Jpn. 78, 114714 (2009).
- (15) P. Alemany, J.-P. Pouget, and E. Canadell, Phys. Rev. B 85, 195118 (2012).
- (16) D. Ohki, K. Yoshimi, and A. Kobayashi Phys. Rev. B 102, 235116 (2020).
- (17) T. Tsumuraya and Y. Suzumura, Eur. Phys. J. B 94, 17 (2021).
- (18) S. Kitou, T. Tsumuraya, H. Sawahata, F. Ishii, K. Hiraki, T. Nakamura, N. Katayama, H. Sawa, Phys. Rev. B 103, 035135 (2021).
- (19) N. Tajima, T. Yamauchi, T. Yamaguchi, M. Suda, Y. Kawasugi, H. M. Yamamoto, R. Kato, Y. Nishio, and K. Kajita, Phys. Rev. B 88, 075315 (2013).
- (20) R. Kato and H. Kobayashi, Synth. Met. 42, 2093 (1991).
- (21) See Supplemental Material at [URL will be inserted by publisher] for details of the samples, magnetoresistance in a bulk crystal without substrate, magneto-transport properties in samples #2 and #3, and comparison of conductance and resistance oscillations under high pressure in sample #4.]
- (22) K. Murata, H. Yoshino, H. O. Yadav, Y. Honda, and N. Shirakawa, Rev. Sci. Instrum. 68, 2490 (1997).
- (23) Y. Kawasugi, H. M. Yamamoto, M. Hosoda, N. Tajima, T. Fukunaga, K. Tsukagoshi, and R. Kato, Appl. Phys. Lett. 92, 243508 (2008).
- (24) T. Schumann, M. Goyal, D. A. Kealhofer, and S. Stemmer, 95, 241113(R) (2017).
- (25) L. Onsager, Philos. Mag. 43, 1006 (1952).
- (26) I. M. Lifshitz and A. M. Kosevich, Sov. Phys. JETP 2, 636 (1956).
- (27) S. G. Sharapov, V. P. Gusynin, and H. Beck, Phys. Rev. B 69, 075104 (2004).
- (28) I. A. Luk’yanchuk and Y. Kopelevich, Phys. Rev. Lett. 93, 166402 (2004).
- (29) Y. Zhang, Y.-W. Tan, H. L. Stormer, and P. Kim, Nature 438, 201 (2005).
- (30) G. P. Mikitik and Yu. V. Sharlai, Phys. Rev. Lett. 82, 2147 (1999).
- (31) J. Xiong, Y. Luo, Y. Khoo, S. Jia, R. J. Cava, and N. P. Ong, Phys. Rev. B 86, 045314 (2012).
- (32) Y. Ando, J. Phys. Soc. Jpn. 82, 102001 (2013).
- (33) M. Monteverde, M. O. Goerbig, P. Auban-Senzier, F. Navarin, H. Henck, C. R. Pasquier, C. Mézière, and P. Batail, Phys. Rev. B 87, 245110 (2013).
- (34) G. Montambauxa, F. Piéchon, J.-N. Fuchs, and M.O. Goerbig, Euro. Phys. J. B 72, 509 (2009).
- (35) T. Morinari and Y. Suzumura, J. Phys. Soc. Jpn. 83, 094701 (2014).
- (36) N. Tajima, S. Sugawara, M. Tamura, R. Kato, Y. Nishio, and K. Kajita, Europhys. Lett. 80, 47002 (2007).
- (37) N. Tajima, M. Sato, S. Sugawara, R. Kato, Y. Nishio, and K. Kajita, Phys. Rev. B 82, 121420(R) (2010).