Detection of magnetic focusing peaks allows probing of minibands in the energy spectrum of twisted bilayer graphene.
Abstract
Magnetic fields force ballistic electrons injected from a narrow contact to move along skipping orbits and form caustics. This leads to pronounced resistance peaks at nearby voltage probes as electrons are effectively focused inside them, a phenomenon known as magnetic focusing. This can be used not only for the demonstration of ballistic transport but also to study the electronic structure of metals. Here, we use magnetic focusing to probe narrowbands in graphene bilayers twisted at ~2°. Their minibands are found to support long-range ballistic transport limited at low temperatures by intrinsic electron-electron scattering. A voltage bias between the layers causes strong minivalley splitting and allows selective focusing for different minivalleys, which is of interest for using this degree of freedom in frequently discussed valleytronics.
INTRODUCTION
Crystallographic alignment of atomically thin crystals stacked together in a van der Waals heterostructure is a powerful tool that enables fine-tuning of their electronic spectra. For crystals with similar honeycomb lattices, the spectra are modified by the presence of a long-range interference (moiré) pattern with a period λS dependent on the twist angle θ between the layers (see Fig. 1A) (1–18). The additional spatial periodicity reduces the size of the Brillouin zone and introduces secondary Dirac points, as illustrated in Fig. 1B. So far, the most pronounced twist-engineered changes in the electronic properties of two-dimensional (2D) crystals have been achieved in twisted bilayer graphene (TBG), where the twist at discrete “magic” angles results in narrowbands, periodically modulated interlayer hybridization, and strong enhancement of electron correlations, leading to superconductivity and Mott insulator transitions (6–8). At larger θ, the TBG spectrum corresponds to a metal with several minibands at each K and K′ valley in the Brillouin zone (Fig. 1B). Electronic properties of such a metal are expected to be quite different from the behavior of Dirac electrons in monolayer or bilayer (aligned to Bernal stacking) graphene but so far remain largely unexplored. Here, we use transverse focusing of electrons in a perpendicular magnetic field (TMF) (12, 19–23) to probe the properties of moiré minibands in TBG and demonstrate an exceptionally high quality of the “artificial metal” in TBG, as well as a possibility to use vertical displacement field, D, to break the valley degeneracy in the two constituent layers and selectively enhance transport in one of the minivalley.
RESULTS
Studied devices
We studied two high-quality dual-gated TBG devices encapsulated with ~30- to 50-nm-thick hexagonal boron nitride (hBN) crystals: D1, with θ = 1.87° ± 0.01° (shown in Fig. 1C), and D2, with θ = 2.60° ± 0.01° (fig. S1A). The procedure used to determine θ is described in section S1. The devices were fabricated using standard dry-transfer (24, 25) and tear-and-stack (4, 26) techniques (see section S2 for details). To ensure a clean interface between the two graphene layers, special care was taken to avoid any contact between graphene and the polymer during the transfer (section S2). In transport measurements, both devices showed similar behavior, with low-temperature mobilities in excess of 400,000 cm2 V−1 s−1 for carrier density n ~ 1012 cm−2. All data shown below were obtained at a constant displacement field, D, that was achieved by a simultaneous sweep of the top and bottom gates (section S2).
Transverse magnetic focusing
The high mobility for both devices enabled observation of TMF (12, 19–23), which is a manifestation of ballistic motion of electrons and had been used to characterize the shape of Fermi surfaces in both 3D (19, 20) and 2D (12, 21–23) metals. To measure the effect of TMF in our TBG devices, we used a nonlocal geometry illustrated in Fig. 1C, where narrow contacts 1 and 2 at one end of the device were used for current injection (driving current I12) and contacts 3 and 4 at the other end were used to detect a voltage V34. In the presence of a perpendicular magnetic field, electrons injected from contact 1 propagate along the device edges in skipping orbits and form a characteristic caustic pattern determined by the shape of the Fermi surface, as illustrated in the insets of Fig. 1D. Caustics are focused into equidistant focal points along the sample edge, and the drift direction of the skipping orbits is determined by the sign of the magnetic field such that electron- and hole-like carriers propagate in opposite directions. As the positions of focal points vary with the magnetic field, whenever they coincide with the position of the voltage probe (contact 3 in Fig. 1C), one observes a focusing peak in the nonlocal resistance Rf = V34/I12. Figure 1D gives two examples of the observed focusing peaks measured at different carrier densities.
Figure 2A shows a typical dependence of Rf on the carrier density and magnetic field at zero displacement field, D = 0 V nm−1. Here, the appearance of an Rf signal in a particular quadrant of the B-n diagram reflects the sign of the cyclotron mass, while the change of the quadrant upon doping indicates an inversion of the electron dispersion (i.e., a change of sign of the mass from electrons to holes or vice versa). Accordingly, a fan-like pattern in the center of Fig. 2A, which converges and changes direction at zero carrier density, indicates a neutrality point. Two additional, qualitatively similar, changes of the cyclotron mass appear at higher electron and hole densities, showing inverted fan-like patterns at higher energies. These indicate that the electron dispersion converges toward a new (secondary) neutrality point, such as shown in Fig. 1B. The crossover between these two regimes (at n ≈ 3 × 1012 cm−2 and −3 × 1012 cm−2) must correspond to a van Hove singularity (vHS) in the moiré miniband spectrum. For a quantitative comparison, Fig. 2B presents the results of TMF modeling for device D1. Here, we used the model from (3) to compute the electron spectrum shown in Fig. 1B (see section S3 for details) and to perform numerical analysis of caustics (12), where the contributions to Rf from trajectories of electrons leaving the injection contact at different angles were weighted proportionally to ∣∇kE∣−1 (section S4). A good agreement between the experiment (Fig. 2A) and theory (Fig. 2B) suggests that the band structure of TBG is well described by the spectrum shown in Fig. 1B.
It is noticeable that the fan-like patterns in Fig. 2 (A and B)—corresponding to the main and secondary neutrality points of the TBG superlattice (around zero carrier density and above the vHS, respectively)—have different periodicities. This difference is caused by different sizes of the Fermi surfaces at equivalent doping levels (black and red dashed lines in Fig. 2A), due to the degeneracy of the miniband dispersion at κ and κ′. The Fermi surface contours are shown in Fig. 2C by black dashed lines around κ and κ′ points of the mini Brillouin zone (main neutrality point) and a red dashed line around the γ point (secondary neutrality point). Furthermore, our theoretical analysis suggests that the Fermi surfaces close to the γ point have a triangular shape (Fig. 2C), which can be traced to the strong interlayer hybridization of those states. At the same time, the Fermi surfaces around κ and κ′ points (that coincide with the valley centers K of the top and bottom graphene layers) are almost isotropic, as in monolayer graphene, pointing toward weak interlayer hybridization of these states.
Effect of the displacement field
The absence of appreciable interlayer coupling at κ and κ′ can be used to disentangle the TMF contributions from different minivalleys. To this end, we used a finite displacement field, up to D = 0.75 V nm−1 (achievable without a risk of damaging our devices), which shifts the on-layer potential for electrons and therefore shifts the energies of the Dirac cones at κ and κ′, as illustrated in Fig. 2C. Such layer-symmetry breaking lifts the degeneracy between κ and κ′ and separates the motion of electrons from different minivalleys in a magnetic field, as they now have different sizes of cyclotron orbits. This generates two different magneto-oscillation frequencies of Rf at low carrier densities, |n| < 1012 cm−2, as seen in Fig. 2D, where separate focusing peaks appear for the electrons from each minivalley.
DISCUSSION
Further information about carrier dynamics in TBG can be obtained by studying the temperature dependence of TMF and its evolution for consecutive focusing peaks. In Fig. 3A, we show how the amplitude of TMF oscillations depends on temperature T in the range 2 K < T < 30 K, in the vicinity of both main and secondary neutrality points. For quantitative analysis, we extract the relative scattering length as (12)
(1) |
where Lpath is the length of trajectories extending from the injector to the first focal point as shown in Fig. 1D, and A1(2)(T) and A1(2)(Tbase) are the areas under the first (second) focusing peak in Fig. 3A at T and Tbase = 2 K, respectively. The results are shown in Fig. 3B. The measured scattering lengths Ls for both carrier densities and all focusing peaks follow a T−2 scaling, which is different from the T−1 dependence characteristic of phonon-dominated scattering (23, 27). Such scaling points toward the dominance of low-angle electron-electron scattering that was also found to be responsible for the TMF suppression in graphene/hBN superlattices (12). Furthermore, the ratio between the areas under the second and first focusing peaks in Fig. 3A, A2/A1, characterizes the reflection of electrons at the sample boundary: The closer it is to one, the higher the probability for the incoming electrons to undergo specular reflection. In our experiment, electrons with energies near the main neutrality points (n ≈ 1.8 × 1012 cm−2; Fig. 3A, right) undergo almost specular reflection (A2/A1 ≈ 0.8), while reflection of the electrons with energies near the secondary neutrality point (n ≈ 6.6 × 1012 cm−2; Fig. 3A, left) is notably less specular (A2/A1 ≈ 0.65). This indicates a higher probability of diffusive scattering in the latter case, which is consistent with the greater sensitivity of the corresponding part of the miniband spectrum to inevitable perturbations of the moiré pattern near the sample edge. Because of little hybridization between the layers near κ and κ′, the scattering of Dirac electrons should be little affected by the termination of superlattice periodicity near the edge, while its part near the secondary neutrality points should be affected substantially, promoting diffusive scattering.
Last, we note that the above observations of superlattice effects in TMF correlate well with the bulk transport properties of the same TBG samples studied using local geometry. The longitudinal and Hall resistivity shown in Fig. 4 (A and B) displays secondary neutrality points (indicated by black arrows) and vHS (red arrows) at the same carrier densities as those inferred from the TMF experiments. Furthermore, the presence of a moiré superlattice in the studied TBG samples is seen from the presence of Brown-Zak oscillations (13, 14) that dominate the magnetotransport above T ~ 30 K (Fig. 4C): While at low temperatures the magnetoresistance is dominated by Shubnikov–de Haas oscillations (see T = 2 K curve in Fig. 4C), these are rapidly suppressed as T increases and give way to another 1/B-periodic oscillations, with period determined by the relation between the magnetic flux through the moiré supercell area, AS, and the magnetic flux quantum ϕ0 = h/e, i.e., BAS = ϕ0/q (where q is an integer).
To conclude, we have demonstrated that TBG supports ballistic propagation of electrons in multimicrometer devices, with electron transport determined by the reconstruction of the energy spectrum in the presence of a long-period superlattice. This offers new opportunities to study fundamental phenomena, such as Bloch oscillations in moiré superlattices (28–30) and their use for, e.g., terahertz generation. Moreover, we have shown that the sensitivity of the TBG band structure to the displacement field allows selective manipulation of electrons from different minivalleys, which may be implemented in electronic devices exploiting the valley degree of freedom.
MATERIALS AND METHODS
Device fabrication
To make TBG, we used the standard tear-and-stack method as described in section S2. TBG stacks were encapsulated between hBN crystals and assembled onto SiO2 substrate with doped silicon underneath, which served as a bottom gate electrode. Then, devices were shaped into hall bars using standard electron-beam lithography and reactive-ion etching techniques (section S2).
Measurement details
We used standard low-frequency lock-in measurement technique with excitation frequency of 10 to 30 Hz. To independently control the carrier density and interlayer displacement field, we used dual-gated geometry, as discussed in section S2.
Supplementary Material
Acknowledgments
We thank J. R. Wallbank for helpful discussions. Funding: A.I.B., B.T., and A.C. acknowledge support from Graphene NOWNANO Doctoral Training Center, and R.K.K. was supported by an EPSRC fellowship award. This work has been supported by EPSRC grants EP/S019367/1, EP/S030719/1, and EP/N010345/1; EPSRC Doctoral Training Centre Graphene NOWNANO EP/L01548X/1; ERC Synergy Grant Hetero2D; Lloyd’s Register Foundation Nanotechnology grant; and European Graphene Flagship Project. Author contributions: P.K. and S.G.X. fabricated devices. A.I.B. carried out electrical measurements with the help from R.K.K. A.I.B., B.T., V.I.F., and I.V.G. analyzed the results. B.T. did magnetic focusing simulations with contributions from V.I.F., A.C., and A.K. T.T. and K.W. provided hBN crystals. A.I.B., B.T., I.V.G., and V.I.F. wrote the manuscript with contributions from A.K.G., P.K., and S.G.X. All authors contributed to discussions. 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
REFERENCES AND NOTES
- 1.Li G., Luican A., Lopes dos Santos J. M. B., Castro Neto A. H., Reina A., Kong J., Andrei E. Y., Observation of Van Hove singularities in twisted graphene layers. Nat. Phys. 6, 109–113 (2010). [Google Scholar]
- 2.Suárez Morell E., Correa J. D., Vargas P., Pacheco M., Barticevic Z., Flat bands in slightly twisted bilayer graphene: Tight-binding calculations. Phys. Rev. B 82, 121407(R) (2010). [Google Scholar]
- 3.Bistritzer R., MacDonald A. H., Moiré bands in twisted double-layer graphene. Proc. Natl. Acad. Sci. U.S.A. 108, 12233–12237 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Cao Y., Luo J. . Y., Fatemi V., Fang S., Sanchez-Yamagishi J. . D., Watanabe K., Taniguchi T., Kaxiras E., Jarillo-Herrero P., Superlattice-induced insulating states and valley-protected orbits in twisted bilayer graphene. Phys. Rev. Lett. 117, 116804 (2016). [DOI] [PubMed] [Google Scholar]
- 5.Kim K., Da Silva A., Huang S., Fallahazad B., Larentis S., Taniguchi T., Watanabe K., Le Roy B. J., MacDonald A. H., Tutuc E., Tunable moiré bands and strong correlations in small-twist-angle bilayer graphene. Proc. Natl. Acad. Sci. U.S.A. 114, 3364–3369 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Cao Y., Fatemi V., Demir A., Fang S., Tomarken S. L., Luo J. Y., Sanchez-Yamagishi J. D., Watanabe K., Taniguchi T., Kaxiras E., Ashoori R. C., Jarillo-Herrero P., Correlated insulator behaviour at half-filling in magic-angle graphene superlattices. Nature 556, 80–84 (2018). [DOI] [PubMed] [Google Scholar]
- 7.Cao Y., Fatemi V., Fang S., Watanabe K., Taniguchi T., Kaxiras E., Jarillo-Herrero P., Unconventional superconductivity in magic-angle graphene superlattices. Nature 556, 43–50 (2018). [DOI] [PubMed] [Google Scholar]
- 8.Yankowitz M., Chen S., Polshyn H., Zhang Y., Watanabe K., Taniguchi T., Graf D., Young A. F., Dean C. R., Tuning superconductivity in twisted bilayer graphene. Science 363, 1059–1064 (2019). [DOI] [PubMed] [Google Scholar]
- 9.Yankowitz M., Xue J., Cormode D., Sanchez-Yamagishi J. D., Watanabe K., Taniguchi T., Jarillo-Herrero P., Jacquod P., Le Roy B. J., Emergence of superlattice Dirac points in graphene on hexagonal boron nitride. Nat. Phys. 8, 382–386 (2012). [Google Scholar]
- 10.Ponomarenko L. A., Gorbachev R. V., Yu G. L., Elias D. C., Jalil R., Patel A. A., Mishchenko A., Mayorov A. S., Woods C. R., Wallbank J. R., Mucha-Kruczynski M., Piot B. A., Potemski M., Grigorieva I. V., Novoselov K. S., Guinea F., Fal’ko V. I., Geim A. K., Cloning of Dirac fermions in graphene superlattices. Nature 497, 594–597 (2013). [DOI] [PubMed] [Google Scholar]
- 11.Dean C. R., Wang L., Maher P., Forsythe C., Ghahari F., Gao Y., Katoch J., Ishigami M., Moon P., Koshino M., Taniguchi T., Watanabe K., Shepard K. L., Hone J., Kim P., Hofstadter’s butterfly and the fractal quantum Hall effect in moiré superlattices. Nature 497, 598–602 (2013). [DOI] [PubMed] [Google Scholar]
- 12.Lee M., Wallbank J. R., Gallagher P., Watanabe K., Taniguchi T., Fal’ko V. I., Goldhaber-Gordon D., Ballistic miniband conduction in a graphene superlattice. Science 353, 1526–1529 (2016). [DOI] [PubMed] [Google Scholar]
- 13.Krishna Kumar R., Chen X., Auton G. H., Mishchenko A., Bandurin D. A., Morozov S. V., Cao Y., Khestanova E., Shalom M. B., Kretinin A. V., Novoselov K. S., Eaves L., Grigorieva I. V., Ponomarenko L. A., Fal’ko V. I., Geim A. K., High-temperature quantum oscillations caused by recurring Bloch states in graphene superlattices. Science 357, 181–184 (2017). [DOI] [PubMed] [Google Scholar]
- 14.Krishna Kumar R., Mishchenko A., Chen X., Pezzini S., Auton G. H., Ponomarenko L. A., Zeitler U., Eaves L., Fal’ko V. I., Geim A. K., High-order fractal states in graphene superlattices. Proc. Natl. Acad. Sci. U.S.A. 115, 5135–5139 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Seyler K. L., Rivera P., Yu H., Wilson N. P., Ray E. L., Mandrus D. G., Yan J., Yao W., Xu X., Signatures of moiré-trapped valley excitons in MoSe2/WSe2 heterobilayers. Nature 567, 66–70 (2019). [DOI] [PubMed] [Google Scholar]
- 16.Tran K., Moody G., Wu F., Lu X., Choi J., Kim K., Rai A., Sanchez D. A., Quan J., Singh A., Embley J., Zepeda A., Campbell M., Autry T., Taniguchi T., Watanabe K., Lu N., Banerjee S. K., Silverman K. L., Kim S., Tutuc E., Yang L., MacDonald A. H., Li X., Evidence for moiré excitons in van der Waals heterostructures. Nature 567, 71–75 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Jin C., Regan E. C., Yan A., Iqbal Bakti Utama M., Wang D., Zhao S., Qin Y., Yang S., Zheng Z., Shi S., Watanabe K., Taniguchi T., Tongay S., Zettl A., Wang F., Observation of moiré excitons in WSe2/WS2 heterostructure superlattices. Nature 567, 76–80 (2019). [DOI] [PubMed] [Google Scholar]
- 18.Alexeev E. M., Ruiz-Tijerina D. A., Danovich M., Hamer M. J., Terry D. J., Nayak P. K., Ahn S., Pak S., Lee J., Sohn J. I., Molas M. R., Koperski M., Watanabe K., Taniguchi T., Novoselov K. S., Gorbachev R. V., Shin H. S., Fal’ko V. I., Tartakovskii A. I., Resonantly hybridized excitons in moiré superlattices in van der Waals heterostructures. Nature 567, 81–86 (2019). [DOI] [PubMed] [Google Scholar]
- 19.Tsoi V. S., Focusing of electrons in a metal by a transverse magnetic field. JETP Lett. 19, 70–71 (1974). [Google Scholar]
- 20.Tsoi V. S., Determination of the dimensions of nonextremal Fermi-surface sections by transverse focusing of electrons. JETP Lett. 22, 197–198 (1975). [Google Scholar]
- 21.Beenakker C. W. J., Williamson J. G., Broekaart M. E. I., van Loosdrecht P. H. M., van Wees B. J., Mooij J. E., Foxon C. T., Harris J. J., Coherent electron focusing with quantum point contacts in a two-dimensional electron gas. Phys. Rev. B 39, 8556–8575 (1989). [DOI] [PubMed] [Google Scholar]
- 22.Heremans J. J., Santos M. B., Shayegan M., Observation of magnetic focusing in two-dimensional hole systems. Appl. Phys. Lett. 61, 1652–1654 (1992). [Google Scholar]
- 23.Taychatanapat T., Watanabe K., Taniguchi T., Jarillo-Herrero P., Electrically tunable transverse magnetic focusing in graphene. Nat. Phys. 9, 225–229 (2013). [Google Scholar]
- 24.Kretinin A. V., Cao Y., Tu J. S., Yu G. L., Jalil R., Novoselov K. S., Haigh S. J., Gholinia A., Mishchenko A., Lozada M., Georgiou T., Woods C. R., Withers F., Blake P., Eda G., Wirsig A., Hucho C., Watanabe K., Taniguchi T., Geim A. K., Gorbachev R. V., Electronic properties of graphene encapsulated with different two-dimensional atomic crystals. Nano Lett. 14, 3270–3276 (2014). [DOI] [PubMed] [Google Scholar]
- 25.Meric I., Huang P. Y., Gao Q., Gao Y., Tran H., Taniguchi T., Watanabe K., Campos L. M., Muller D. A., Guo J., Kim P., Hone J., Shepard K. L., Dean C. R., One-dimensional electrical contact to a two-dimensional material. Science 342, 614–617 (2013). [DOI] [PubMed] [Google Scholar]
- 26.Kim K., Yankowitz M., Fallahazad B., Kang S., Movva H. C., Huang S., Larentis S., Corbet C. M., Taniguchi T., Watanabe K., Banerjee S. K., Le Roy B. J., Tutuc E., van der Waals heterostructures with high accuracy rotational alignment. Nano Lett. 16, 1989–1995 (2016). [DOI] [PubMed] [Google Scholar]
- 27.Hwang E. H., Das Sarma S., Acoustic phonon scattering limited carrier mobility in two-dimensional extrinsic graphene. Phys. Rev. B 77, 115449 (2008). [Google Scholar]
- 28.Esaki L., Chang L. L., New transport phenomenon in a semiconductor “superlattice”. Phys. Rev. Lett. 33, 495–498 (1974). [Google Scholar]
- 29.Leo K., Bolivar P. H., Brüggemann F., Schwedler R., Köhler K., Observation of Bloch oscillations in a semiconductor superlattice. Solid State Commun. 84, 943–946 (1992). [Google Scholar]
- 30.Waschke C., Roskos H. G., Schwedler R., Leo K., Kurz H., Köhler K., Coherent submillimeter-wave emission from Bloch oscillations in a semiconductor superlattice. Phys. Rev. Lett. 70, 3319–3322 (1993). [DOI] [PubMed] [Google Scholar]
- 31.Rickhaus P., Wallbank J., Slizovskiy S., Pisoni R., Overweg H., Lee Y., Eich M., Liu M. H., Watanabe K., Taniguchi T., Ihn T., Ensslin K., Transport through a network of topological channels in twisted bilayer graphene. Nano Lett. 18, 6725–6730 (2018). [DOI] [PubMed] [Google Scholar]
- 32.N. W. Ashcroft, N. D. Mermin, Solid State Physics (Saunders College, 1976). [Google Scholar]
- 33.Wallbank J. R., Patel A. A., Mucha-Kruczynski M., Geim A. K., Fal’ko V. I., Generic miniband structure of graphene on a hexagonal substrate. Phys. Rev. B 87, 245408 (2013). [Google Scholar]
- 34.Schmidt H., Lüdtke T., Barthold P., Cann E. M., Fal’ko V. I., Haug R. J., Tunable graphene system with two decoupled monolayers. Appl. Phys. Lett. 93, 172108 (2008). [Google Scholar]
- 35.S. Slizovskiy, A. Garcia-Ruiz, N. Drummond, V. I. Fal'ko, Dielectric susceptibility of graphene describing its out-of-plane polarizability. arXiv:1912.10067 [cond-mat.mes-hall] (20 December 2019).
- 36.Sanchez-Yamagishi J. D., Taychatanapat T., Watanabe K., Taniguchi T., Yacoby A., Jarillo-Herrero P., Quantum Hall effect, screening, and layer-polarized insulating states in twisted bilayer graphene. Phys. Rev. Lett. 108, 076601 (2012). [DOI] [PubMed] [Google Scholar]
- 37.Fallahazad B., Hao Y., Lee K., Kim S., Ruoff R. S., Tutuc E., Quantum Hall effect in Bernal stacked and twisted bilayer graphene grown on Cu by chemical vapor deposition. Phys. Rev. B 85, 201408(R) (2012). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.