Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2019 Dec 4;116(51):25530–25534. doi: 10.1073/pnas.1905524116

Pressure-induced topological phase transition in noncentrosymmetric elemental tellurium

Toshiya Ideue a,1, Motoaki Hirayama b, Hiroaki Taiko c, Takanari Takahashi d, Masayuki Murase d, Takashi Miyake e, Shuichi Murakami f,g, Takao Sasagawa d, Yoshihiro Iwasa a,b,c
PMCID: PMC6926023  PMID: 31801879

Significance

Topological Weyl semimetal, which is a gapless semimetallic phase protected by symmetry, generally appears by band gap closing in noncentrosymmetric semiconductors. So far, there have been only a limited number of reports of such a topological phase transition so that many basic aspects have remained unexplored. Here we report potential signatures of the topological phase transition under pressure in elemental tellurium with noncentrosymmetric chiral crystal structure. Our result represents that elemental tellurium can be one of the model systems for the topological phase transition from a semiconductor to a Weyl semimetal, opening the avenue for studying the topological phase transition by the external field and resultant topological properties.

Keywords: topological phase transition, Weyl semimetal, tellurium, SdH oscillations

Abstract

Recent progress in understanding the electronic band topology and emergent topological properties encourage us to reconsider the band structure of well-known materials including elemental substances. Controlling such a band topology by external field is of particular interest from both fundamental and technological viewpoints. Here we report possible signatures of the pressure-induced topological phase transition from a semiconductor to a Weyl semimetal in elemental tellurium probed by transport measurements. Pressure variation of the periods of Shubnikov–de Haas oscillations, as well as oscillation phases, shows an anomaly around the pressure theoretically predicted for topological phase transition. This behavior is consistent with the pressure-induced band deformation and resultant band-crossing effect. Moreover, effective cyclotron mass is reduced toward the critical pressure, potentially reflecting the emergence of massless linear dispersion. The present result paves the way for studying the electronic band topology in well-known compounds and topological phase transition by the external field.


Topological phase of matter is a central issue in recent condensed-matter physics. It is characterized by the nontrivial electronic band topology which causes the emergent electrical, magnetic, and optical properties of solids (1, 2). One of the important topics in this field is the search for the phase control between topologically trivial phase and topologically nontrivial phase. Such a topological phase transition is generally realized by changing the chemical composition of compounds (37) or by applying the external field such as the magnetic field, electric field (8), or pressure (911). Although it can be directly probed by the angle-resolved photoemission spectroscopy in the former case, optical or transport measurements and comparison with the theoretical calculations are necessary for verifying the topological phase transition induced by external fields.

An elemental substance is an ideal platform to study characteristic properties of topological phases and topological phase transition since it has a simple crystal or band structure and we can easily compare it with theoretical calculations (1217). Since the spin–orbit interaction often plays the key role for the band inversion and the emergence of the topologically nontrivial phase, elemental substance with large spin–orbit interaction can be a potential candidate. For example, topological nature of electronic bands has been discussed in Bi, Sb, and As so far (1215). Tellurium is also a potential elemental substance with strong spin–orbit interaction. Fig. 1A shows the crystal structure of tellurium. Helical chains along the z axis are arranged via weak van der Waals interaction in a hexagonal array. Depending on the chirality of helical chains, it belongs to the space group P3121 or P3221. At the ambient pressure, tellurium is a p-type semiconductor in which electronic and optical properties are mainly dominated by the valence and conduction band around the H point in the hexagonal Brillouin zone (Fig. 1B). Previous studies clarified the characteristic camelback-like valence-band dispersion with minimum energy Ec at H point (Fig. 1 C, Left) and that strong spin–orbit interaction should be taken into account to explain the transport and optical properties (18). Observation of the current-induced magnetization due to the chiral structure has been reported (19) and Te is known also as a good thermoelectric material with high figure of merit due to its valley degeneracy and band nesting (20).

Fig. 1.

Fig. 1.

Pressure-induced band deformation and topological phase transition of elemental tellurium. (A) Crystal structure of Te. Helical chains along the z axis are weakly coupled by van der Waals force. (B) Brillouin zone of Te. Electronic bands around H points are dominant in electronic and optical properties of Te. (C) First-principle calculations of Te band under pressure (P = 0 GPa (Left), P = 1.88 GPa (Middle), and P = 2.62 GPa (Right), respectively) and cross-section of the Fermi surface (shaded region) when the magnetic field is applied perpendicular to the cleavage plane (zx plane). Topological phase transition from a semiconductor to a Weyl semimetal is predicted around P = 2 GPa.

Recently, the topological nature of the Te band has attracted renewed interest due to its noncentrosymmetric crystal structure and strong spin–orbit interaction (2124). First-principle calculation predicts that elemental tellurium shows the topological phase transition under pressure (24). Under the application of pressure, characteristic camelback-like valence-band dispersion first disappears (Fig. 1 C, Middle), turning into the normal parabolic band dispersion. With increasing the pressure, the band gap continuously shrinks so that the valence band and conduction band finally touch at the points around the H point (Fig. 1 C, Right), leading to the topological transition from a semiconductor to a Weyl semimetal. (In SI Appendix, we show the detailed band evolution under pressure.) In Weyl semimetal phase, band-crossing points (Weyl points) are stable against the perturbation and will not disappear to open the gap (i.e., topologically protected). This argument of topological transition under pressure is consistent with the general discussion for noncentrosymmetric crystals based on the symmetry (25).

In this study, we have probed the pressure variation of the electronic band in elemental tellurium through electric transport, especially focusing on the Shubnikov–de Haas (SdH) oscillations and found the potential signature of the pressure-induced topological phase transition. Periods of SdH oscillations, as well as phases, sensitively change by pressure, reflecting the band deformation under pressure. In a sample with low carrier density, periods of SdH oscillations show an anomalous jump around P = 0.5 GPa, which is consistent with the theoretically predicted pressure-induced Lifshitz transition (SI Appendix). Furthermore, nonmonotonous pressure dependences of SdH oscillation periods and phases have been observed in all samples irrespective of carrier density, which can be well understood by the characteristic Fermi surface deformation and Berry phase change during the topological phase transition. Our result represents that elemental tellurium can be one of the model systems which show the topological phase transition from a semiconductor to a Weyl semimetal. It also implies that oscillations can be a sophisticated probe to detect the band deformation and resultant topological phase transition.

Results and Discussion

In Fig. 2 A and B, we show the magnetoresistance (Rzz) and Hall effect (Rxz) of sample 1 under various pressures. In this work, we mainly applied the magnetic field perpendicular to the natural growth surface (zx plane) and the current direction is fixed to be the z axis. [We also studied pressure dependence of SdH oscillations in another configuration (BIz) in SI Appendix.] Carrier density nHall of sample 1 estimated from the Hall effect is nHall=2.7×1016cm3 (Carrier number studied in this work ranges from nHall1016cm3 to 1017cm3.) Clear SdH oscillations have been observed at T = 2K in both Rzz and Rxz, which are apparently dependent on the pressure. Fig. 2C represents the oscillating component of Rzz calculated by subtracting the polynomial function. Maximum and minimum positions are linearly fitted in the index plots (Fig. 2D). In this index plot, we assigned minimum of Rzz as an integer. Cross-sectional areas of the Fermi surface (SF) are calculated from the periods of the oscillating components (or slopes of index plots) by the following relation and are plotted as a function of the pressure in Fig. 2E:

SF=2πeΔ(1/B),

where e, , and Δ(1/B) are electron charge, Planck’s constant, and the period of the oscillation, respectively.

Fig. 2.

Fig. 2.

SdH oscillations of sample 1 (EF > Ec). (A and B) Magnetoresistance (A) and Hall effect (B) of sample 1 under various pressures. Clear SdH oscillations have been observed at T = 2K. (C) Oscillating components of the resistance calculated by subtracting the polynomial background from A. (D) Landau index plot of sample 1. We chose the minimum of the resistance as an integer index. (E) Pressure dependence of the cross-sectional area of Fermi surface estimated from the periods of oscillations. SF shows the anomalous jump around P = 0.5 GPa and upturn at P = 2 GPa. (F) Pressure evolution of the calculated 3D Fermi surface for the sample with EF = 1 meV at ambient pressure. From the left figure, P = 0 GPa, P = 0.6 GPa, P = 1.22 GPa, and P = 2.16 GPa, respectively.

At the ambient pressure, SF calculated from the oscillation period is 0.526×103Å2, indicating that Fermi energy of sample 1 locates above the minimum of the camelback-like valence-band dispersion (Ec) (SI Appendix). As we increase the pressure, each peak in oscillations first shows the systematic change (Fig. 2C, region 1) but suddenly splits into 2 peaks around P = 0.5 GPa (Fig. 2C, region 2). At this pressure, SF estimated from the period of SdH oscillations is almost doubled as shown in Fig. 2E (regions 1 and 2). This split of SdH oscillation and anomalous jump of SF is considered to be the signature of pressure-induced Lifshitz transition. As predicted by calculations (Fig. 1C), characteristic camelback-like valence-band dispersion disappears under pressure. Fermi level of sample 1, first locating above the minimum of the camelback-like valence-band dispersion (Ec), can pass through the Ec during the systematic pressure-induced band deformation, which causes the Fermi surface topology change and resultant Lifshitz transition. In Fig. 2F, we show the pressure variations of the calculated 3D Fermi surface in which Lifshitz transition is illustrated in the first step.

After this Lifshitz transition, SdH oscillations again show the systematic change with pressure, showing another anomaly around P = 2 GPa (Fig. 2E); when the pressure is below 2 GPa, SF decreases with pressure (region 2) while it turns upward above P = 2 GPa (region 3). This anomaly is considered to be caused by the band touching predicted by calculations. Before the band touching occurs, Fermi surface is becoming more 3D-like and isotropic by applying the pressure so that the Fermi surface is elongated along the kx and ky direction while it shrinks along the kz direction. Therefore, cross-sectional area of the Fermi surface perpendicular to the magnetic field (kz-kx plane) decreases, mainly reflecting the elongation along the ky direction (second and third schematics in Fig. 2F). On the other hand, once the conduction band and valence band touch, the shape of the Fermi surface becomes anisotropic again since the original conduction-band minimum keeps decreasing under pressure and causes the deformation of the Fermi surface as it gets close to the Fermi energy (second and third schematics in Fig. 1C and third and fourth schematics in Fig. 2F). Thus, the anomaly of SF around P = 2 GPa represents the band touching and can be the potential signature of the topological phase transition from a semiconductor to a Weyl semimetal.

We note that the Hall coefficient is also affected by the pressure in sample 1 (Fig. 2B), which indicates the carrier increase under pressure. It might be because carriers trapped by the defect levels are released under pressure, causing the increase of the carrier density toward P = 2 GPa (Fig. 2B). However, the increase of the carrier density cannot explain the observed decrease of SF below P = 2 GPa so that pressure dependence of SF does not come from the carrier density change but can be mainly attributed to the pressure-induced band deformation. It is also noted that the Hall coefficient is almost unchanged above 2 GPa. In SI Appendix, we discuss the phase of SdH oscillations (or intercept of the index plot), which show the cross-over behavior from P < 2GPa to P > 2GPa, offering the important insight into the change of the band dispersion or Berry phase originating from spin texture (2632).

We also studied the samples with different Fermi energies at the ambient pressure. Fig. 3 AF show Rzz (A), Rxz (B), oscillating components of Rzz (C), index plot (D) under various pressures, and pressure dependence of SF (E), and calculated 3D Fermi surface (F) for sample 2. Carrier density nHall of sample 2 estimated from the Hall effect is nHall=4.1×1017cm3. According to the oscillation period, the Fermi level of sample 2 is estimated to locate below Ec at the ambient pressure (SI Appendix). Differently from sample 1, SF does not show an anomalous jump around P = 0.5 GPa (Fig. 3E). This means there is no pressure-induced Lifshitz transition in sample 2, being consistent with EF < Ec. Nevertheless, pressure dependence of SF still shows the upturn around P = 2 GPa (Fig. 3E) similarly to sample 1, indicating that possible signature of the topological phase transition from a semiconductor to a Weyl semimetal (Fig. 3F) is successfully probed also in sample 2. We have confirmed that similar behavior can be observed in the other 2 samples with Fermi energy below Ec at the ambient pressure (sample 3 and sample 4; see SI Appendix). Intercept δ in the index plot shows cross-over behavior around P = 2 GPa also in sample 2, corroborating the above scenario of possible topological phase transition (SI Appendix).

Fig. 3.

Fig. 3.

SdH oscillations of sample 2 (EF < Ec). (A and B) Magnetoresistance (A) and Hall effect (B) of sample 2 at T = 2K under various pressure. (C) Oscillating components of the resistance. (D) Landau index plot of sample 2. (E) Pressure dependence of the cross-sectional area of Fermi surface estimated from the periods of oscillations. SF shows the anomaly around P = 2 GPa similarly to sample 1 (Fig. 2E). (F) Pressure evolution of the calculated 3D Fermi surface for the sample with EF = 5 meV at ambient pressure. From the left figure, P = 0 GPa, P = 1.22 GPa, and P = 2.62 GPa, respectively.

Finally, we discuss the pressure variation of the effective mass of sample 2. In Fig. 4A, oscillating components of sample 2 under P = 3 GPa at various temperatures are displayed. Magnitude of the oscillations decreases with increasing the temperature, which is plotted in Fig. 4B. According to the following Lifshitz–Kosevich formula for a 3D system (33), we can estimate the effective mass m from the temperature dependence of the oscillation amplitude ΔRzz/Rzz,

ΔRzz/Rzz(ωC/2EF)1/2exp(2π2kBTD/ωC)2π2kBT/ωCsinh(2π2kBT/ωC).

Here, kB, TD, and ωC are Boltzmann’s constant, Dingle temperature, and cyclotron frequency ωC=eBm, respectively. We calculate the effective mass at each pressure from the temperature dependence of ΔRzz/Rzz (Fig. 4B) and plot values normalized by free-electron mass m0 as a function of the pressure in Fig. 4C. Effective cyclotron mass is reduced by applying the pressure, reflecting the transition from massive band dispersion at the ambient pressure to the massless linear band dispersion in Weyl semimetal phase. Pressure variation of m is settled around P = 2 GPa similarly to δ, which almost coincides with the anomaly point of SF, possibly because that band deformation is suppressed after the topological phase transition. From the fitting of temperature variation of SdH oscillations by the Lifshitz–Kosevich formula, we also obtained the Fermi velocity vF=4.0×104 m/s and quantum lifetime τQ=1.4×1013 s at P = 3 GPa. Transport lifetime defined by τtransport=μkF/evF is estimated as τtransport=8.5×1013 s, which is an order of magnitude larger than τQ. It is noted that τtransport is generally larger than τQ, since τtransport measures the backscattering process that relaxes the current while τQ is sensitive to all processes that cause Landau level broadening (31).

Fig. 4.

Fig. 4.

Pressure variation of the effective mass. (A) Oscillating component of sample 2 under P = 3 GPa at various temperatures. (B) Temperature dependence of amplitude of SdH oscillations under various pressures. Dashed line represents the fitting by the Lifshitz–Kosevich formula. (C) Pressure variation of the effective mass m normalized by free-electron mass m0. m decreases with pressure, showing the convergent behavior toward P = 2 Gpa.

In conclusion, we have successfully probed the potential signature of the pressure-induced topological phase transition from a semiconductor to a Weyl semimetal in elemental tellurium. Present results indicate that electric transport including SdH oscillations is a powerful probe to detect the topological phase transition, especially under external field such as the pressure. In view of rich variety of materials whose topological nature is not fully unveiled, our findings will open the way to rediscover the topologically nontrivial band and topological phase transition in well-known materials.

Materials and Methods

Sample Preparations.

Single crystals of Te were prepared by a physical vapor transport technique. Elemental Te (6N) grains of ∼0.5 g were sealed in an evacuated quartz tube (∼φ11 mm × φ9 mm × 120 mm), and the ampule was put in a 2-heating-zone tube furnace (34). Temperatures at the source (Te grains) and the growth zone were set to 450 and 360 °C, respectively. After ∼70 h, single crystals of Te with ∼3-mm length having a hexagonal prism shape as similar to a quartz crystal were obtained.

Transport Measurements.

Resistivity and Hall effect have been measured with the use of a physical property measurement system (Quantum Design, Inc.). Hydrostatic pressure is applied by using a piston-cylinder pressure cell and Daphne 7474 (pressure-transmitting medium). We applied the magnetic field perpendicular to the natural growth surface (zx plane) and current direction is parallel to the z axis.

First-Principle Calculation.

Our calculations are based on the density-function theory in the local density approximation. We calculate the fully relativistic electronic structure by a first-principles code QMAS (Quantum MAterials Simulator) based on the projector augmented-wave method. The plane-wave energy cutoff is set to 40 Ry, and the 6 × 6 × 6 k mesh is employed. We evaluate the self-energy correction in the GW approximation (GWA) (35, 36) using the full-potential linear muffin-tin orbital code (37, 38). In the GWA, we neglect the spin–orbit interaction. The 6 × 6 × 6 k mesh is sampled and 51 × 2 unoccupied conduction bands are included, where ×2 is the spin degrees of freedom. We construct 9 × 2 maximally localized Wannier functions (MLWFs) originating from the 5p orbitals and diagonalize the fully relativistic Hamiltonian with the GW self-energy correction expressed in the MLWF basis. Experimental structures for the trigonal phase are used in the present study. At pressures between available experimental data, we determine the Hamiltonian in the MLWF basis by linear interpolation.

Data Availability.

The data that support the plots within this paper are available in Datasets S1–S24.

Supplementary Material

Supplementary File
Supplementary File
pnas.1905524116.sd01.csv (776.2KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd04.csv (899.8KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd07.csv (191.7KB, csv)
Supplementary File
pnas.1905524116.sd08.csv (156.9KB, csv)
Supplementary File
pnas.1905524116.sd09.csv (156.9KB, csv)
Supplementary File
pnas.1905524116.sd10.csv (156.7KB, csv)
Supplementary File
pnas.1905524116.sd11.csv (156.5KB, csv)
Supplementary File
pnas.1905524116.sd12.csv (122.3KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd15.csv (145.2KB, csv)
Supplementary File
Supplementary File
pnas.1905524116.sd17.csv (213.8KB, csv)
Supplementary File
Supplementary File
pnas.1905524116.sd19.csv (487.6KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd22.csv (465.8KB, csv)
Supplementary File
pnas.1905524116.sd23.csv (1,000B, csv)
Supplementary File

Acknowledgments

This research was supported by the following grants: Japan Society for the Promotion of Science (JSPS) Grants-in-Aid for Scientific Research (S) (JP19H05602), Grant-in-Aid for Challenging Research (Exploratory) (JP19K21843), Grant-in-Aid for Scientific Research on Innovative Areas “Topological Materials Science” (JP18H04216), Grants-in-Aid for Scientific Research (B) (JP19H01819), Grants-in-Aid for Scientific Research (B) (JP16H03847), and Japan Science and Technology Agency CREST project (JPMJCR16F2).

Footnotes

The authors declare no competing interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.1905524116/-/DCSupplemental.

References

  • 1.Qi X. L., Zhang S. C., Topological insulators and superconductors. Rev. Mod. Phys. 83, 1057–1109 (2011). [Google Scholar]
  • 2.Armitage N. P., Mele E. J., Vishwanath A., Weyl and Dirac semimetals in three-dimensional solids. Rev. Mod. Phys. 90, 015001 (2018). [Google Scholar]
  • 3.Hsieh D., et al. , A topological Dirac insulator in a quantum spin Hall phase. Nature 452, 970–974 (2008). [DOI] [PubMed] [Google Scholar]
  • 4.Hsieh D., et al. , A tunable topological insulator in the spin helical Dirac transport regime. Nature 460, 1101–1105 (2009). [DOI] [PubMed] [Google Scholar]
  • 5.Dziawa P., et al. , Topological crystalline insulator states in Pb1-xSnxSe. Nat. Mater. 11, 1023–1027 (2012). [DOI] [PubMed] [Google Scholar]
  • 6.Xu S. Y., et al. , Topological phase transition and texture inversion in a tunable topological insulator. Science 332, 560–564 (2011). [DOI] [PubMed] [Google Scholar]
  • 7.Brahlek M., et al. , Topological-metal to band-insulator transition in (Bi1-xInx)2Se3 thin films. Phys. Rev. Lett. 109, 186403 (2012). [DOI] [PubMed] [Google Scholar]
  • 8.Qian X., Liu J., Fu L., Li J., Quantum spin Hall effect in two-dimensional transition metal dichalcogenides. Science 346, 1344–1347 (2014). [DOI] [PubMed] [Google Scholar]
  • 9.Xi X., et al. , Signatures of a pressure-induced topological quantum phase transition in BiTeI. Phys. Rev. Lett. 111, 155701 (2013). [DOI] [PubMed] [Google Scholar]
  • 10.Ideue T., et al. , Pressure variation of Rashba spin splitting toward topological transition in the polar semiconductor BiTeI. Phys. Rev. B 90, 161107 (2014). [Google Scholar]
  • 11.Liang T., et al. , A pressure-induced topological phase with large Berry curvature in Pb1-x Sn x Te. Sci. Adv. 3, e1602510 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Drozdov I. K., et al. , One-dimensional topological edge states of bismuth bilayers. Nat. Phys. 10, 664–669 (2014). [Google Scholar]
  • 13.Schindler F., et al. , Higher-order topology in bismuth. Nat. Phys. 14, 918–924 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Yao G., et al. , Evolution of topological surface states in antimony ultra-thin films. Sci. Rep. 3, 2010 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Zhang P., et al. , Topologically entangled rashba-split shockley states on the surface of grey arsenic. Phys. Rev. Lett. 118, 046802 (2017). [DOI] [PubMed] [Google Scholar]
  • 16.Castro Neto A. H., Guinea F., Peres N. M. R., Novoselov K. S., Geim A. K., The electronic properties of graphene. Rev. Mod. Phys. 81, 109–162 (2009). [Google Scholar]
  • 17.Kim J., et al. , Observation of tunable band gap and anisotropic Dirac semimetal state in black phosphorus. Science 349, 723–726 (2015). [DOI] [PubMed] [Google Scholar]
  • 18.Gerlach E., Grosse P., The Physics of Selenium and Tellurium (Springer Series in Solid-State Sciences, Springer, Berlin, 1979), vol. 13. [Google Scholar]
  • 19.Furukawa T., Shimokawa Y., Kobayashi K., Itou T., Observation of current-induced bulk magnetization in elemental tellurium. Nat. Commun. 8, 954 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Lin S., et al. , Tellurium as a high-performance elemental thermoelectric. Nat. Commun. 7, 10287 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Agapito L. A., Kioussis N., Goddard W. A. III, Ong N. P., Novel family of chiral-based topological insulators: Elemental tellurium under strain. Phys. Rev. Lett. 110, 176401 (2013). [DOI] [PubMed] [Google Scholar]
  • 22.Nakayama K., et al. , Band splitting and Weyl nodes in trigonal tellurium studied by angle-resolved photoemission spectroscopy and density functional theory. Phys. Rev. B 95, 125204 (2017). [Google Scholar]
  • 23.Sakano M., et al. , Radial spin texture in elemental tellurium with chiral crystal structure. arXiv:1908.09725 (26 August 2019). [DOI] [PubMed]
  • 24.Hirayama M., Okugawa R., Ishibashi S., Murakami S., Miyake T., Weyl node and spin texture in trigonal tellurium and selenium. Phys. Rev. Lett. 114, 206401 (2015). [DOI] [PubMed] [Google Scholar]
  • 25.Murakami S., Hirayama M., Okugawa R., Miyake T., Emergence of topological semimetals in gap closing in semiconductors without inversion symmetry. Sci. Adv. 3, e1602680 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Luk’yanchuk I. A., Kopelevich Y., Phase analysis of quantum oscillations in graphite. Phys. Rev. Lett. 93, 166402 (2004). [DOI] [PubMed] [Google Scholar]
  • 27.Qu D. X., Hor Y. S., Xiong J., Cava R. J., Ong N. P., Quantum oscillations and hall anomaly of surface states in the topological insulator Bi2Te3. Science 329, 821–824 (2010). [DOI] [PubMed] [Google Scholar]
  • 28.Murakawa H., et al. , Detection of berry’s phase in a bulk rashba semiconductor. Science 342, 1490–1493 (2013). [DOI] [PubMed] [Google Scholar]
  • 29.Doiron-Leyraud N., Szkopek T., Pereg-Barnea T., Proust C., Gervais G., Berry phase in cupurate superconductors. Phys. Rev. B 91, 245136 (2015). [Google Scholar]
  • 30.Ideue T., et al. , Thermoelectric probe for Fermi surface topology in the three-dimensional Rashba semiconductor BiTeI. Phys. Rev. B 92, 115144 (2015). [Google Scholar]
  • 31.Liang T., et al. , Ultrahigh mobility and giant magnetoresistance in the Dirac semimetal Cd3As2. Nat. Mater. 14, 280–284 (2015). [DOI] [PubMed] [Google Scholar]
  • 32.Xiong J., et al. , Evidence for the chiral anomaly in the Dirac semimetal Na3Bi. Science 350, 413–416 (2015). [DOI] [PubMed] [Google Scholar]
  • 33.Shoenberg D., Magnetic Oscillations in Metals (Cambridge University Press, Cambridge, England, 1984). [Google Scholar]
  • 34.Iwasa A., et al. , Environmentally friendly refining of diamond-molecules via the growth of large single crystals. Cryst. Growth Des. 10, 870–873 (2010). [Google Scholar]
  • 35.Hedin L., New method for calculating the one-particle green’s function with application to the electron-gas problem. Phys. Rev. 139, A796 (1965). [Google Scholar]
  • 36.Hedin L., Lundqvist S., Solid State Physics, Ehrenreich H., Seitz F., Turnbull D., Eds. (Academic, New York, 1969), vol. 23. [Google Scholar]
  • 37.van Schilfgaarde M., Kotani T., Faleev S. V., Adequacy of approximations in GW theory. Phys. Rev. B 74, 245125 (2006). [DOI] [PubMed] [Google Scholar]
  • 38.Miyake T., Aryasetiawan F., Screened Coulomb interaction in the maximally localized Wannier basis. Phys. Rev. B 77, 085122 (2008). [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary File
Supplementary File
pnas.1905524116.sd01.csv (776.2KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd04.csv (899.8KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd07.csv (191.7KB, csv)
Supplementary File
pnas.1905524116.sd08.csv (156.9KB, csv)
Supplementary File
pnas.1905524116.sd09.csv (156.9KB, csv)
Supplementary File
pnas.1905524116.sd10.csv (156.7KB, csv)
Supplementary File
pnas.1905524116.sd11.csv (156.5KB, csv)
Supplementary File
pnas.1905524116.sd12.csv (122.3KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd15.csv (145.2KB, csv)
Supplementary File
Supplementary File
pnas.1905524116.sd17.csv (213.8KB, csv)
Supplementary File
Supplementary File
pnas.1905524116.sd19.csv (487.6KB, csv)
Supplementary File
Supplementary File
Supplementary File
pnas.1905524116.sd22.csv (465.8KB, csv)
Supplementary File
pnas.1905524116.sd23.csv (1,000B, csv)
Supplementary File

Data Availability Statement

The data that support the plots within this paper are available in Datasets S1–S24.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES