[1,2]\fnmAlina \surMreńca-Kolasińska
[2,3]\fnmSzu-Chao \surChen
[2]\fnmMing-Hao \surLiu
1]\orgnameAGH University, \orgdivFaculty of Physics and Applied Computer Science, \orgaddress\streetal. Mickiewicza 30, \postcode30-059 \cityKraków, \countryPoland
2]\orgdivDepartment of Physics, \orgnameNational Cheng Kung University, \orgaddress\cityTainan \postcode70101, \countryTaiwan
3]\orgdivDepartment of Electro-Optical Engineering, \orgnameNational Formosa University, \orgaddress\cityYunlin, \countryTaiwan
Probing miniband structure and Hofstadter butterfly in gated graphene superlattices via magnetotransport
Abstract
The presence of periodic modulation in graphene leads to a reconstruction of the band structure and formation of minibands. In an external uniform magnetic field, a fractal energy spectrum called Hofstadter butterfly is formed. Particularly interesting in this regard are superlattices with tunable modulation strength, such as electrostatically induced ones in graphene. We perform quantum transport modeling in gate-induced square two-dimensional superlattice in graphene and investigate the relation to the details of the band structure. At low magnetic field the dynamics of carriers reflects the semi-classical orbits which depend on the mini band structure. We theoretically model transverse magnetic focusing, a ballistic transport technique by means of which we investigate the minibands, their extent and carrier type. We find a good agreement between the focusing spectra and the mini band structures obtained from the continuum model, proving usefulness of this technique. At high magnetic field the calculated four-probe resistance fit the Hofstadter butterfly spectrum obtained for our superlattice. Our quantum transport modeling provides an insight into the mini band structures, and can be applied to other superlattice geometries.
Introduction
Graphene, a 2D material characterized by a linear low-energy dispersion relation, hosts charge carriers named Dirac fermions due to the resemblance of relativistic (massless) particles described by the Dirac equation. Modifying the underlying graphene lattice by a smooth periodic potential can affect the band structure through folding of the pristine graphene Dirac cone into mini bands Wallbank2013 , formation of the secondary Dirac points, and anisotropic renormalization of velocity Park2008 ; Park2008prb ; Brey2009 ; Barbier2010 ; Kang2020 . Periodic modulation has been obtained in graphene through chemical functionalization Sun2011 , placing graphene on self-assembled nanostructures Zhang2018 , and by stacking graphene together with aligned or slightly misoriented hexagonal boron nitride (hBN), resulting in periodic moiré modulation which generates hexagonal superlattices (SLs) Xue2011 ; Decker2011 ; Yankowitz2012 ; Ponomarenko2013 ; Hunt2013 ; Dean2013 ; Yu2014 . Moiré SLs were also created in low-angle twisted graphene bilayers , followed by van der Waals structures made up of few layers of graphene Burg2019 ; Shen2020 ; Liu2020 ; Lin2020 ; deVries2020 ; Rickhaus2021 and other 2D materials WangLujun2019 ; WangZihao2019 . SLs are suitable for the observation of the self-similar energy spectrum called Hofstadter butterfly Dean2013 ; Hunt2013 and Brown-Zak oscillations Kumar2017 ; Barrier2020 that occur when the magnetic flux through the superlattice unit cell is of the order of the magnetic flux quantum, , and in pristine 2D crystals require unattainable magnetic fields. Also worth mentioning are the many-body phenomena present in moiré SLs Wang2015 ; Andrews2020 .

Despite their potential, artificial lattices tailored by chemical methods or moiré SLs suffer from inability to tune the strength of the periodic potential. In moiré SLs the period can be tuned to some extent via the rotation angle between the stacked layers, but they are inherent of a hexagonal symmetry. Precise control over the SL geometry, period, and strength is vital for the band structure engineering. The above limitations can be circumvented in electrostatic gate-induced SLs that allow an arbitrary design via the gates geometry, with the gate voltage being a knob for the potential strength. The experimental attempts to create gated SLs in graphene included 1D arrays of metal gates Dubey2013 ; Drienovsky2014 ; Kuiri2018 , followed by patterned dielectric substrates Forsythe2018 ; Li2021 , and few-layer graphene patterned bottom gates for 1D Drienovsky2018 ; Ruiz2022 and 2D SLs Huber2020 ; Huber2022 with down to sub-20 nm periods Ruiz2022 manifesting the flexibility of this approach. With the recent advance in the fabrication techniques, gated SLs with the period of a few tens of nanometers are achievable with good device quality and long electron mean free path Li2021 . This enabled observation of commensurability oscillations Drienovsky2018 ; Li2021 ; Huber2022 , Hofstadter butterfly Huber2020 ; Ruiz2022 and Brown-Zak oscillations Huber2022 at magnetic fields of the order of a few tesla. This is more affordable compared to about 25 T required for graphene/hBN SL, where, due to small lattice constant mismatch of 1.8%, the SL period can reach up to 14 nm for aligned lattices.
While transport in quantizing magnetic field in gated SLs has been thoroughly studied, the intermediate magnetic field regime remained mostly unexplored. In the semiclassical treatment, at low magnetic field fermions undergo cyclotron motion that can be probed in transport measurements via transverse magnetic focusing (TMF). This technique has been used to experimentally probe the band structure in pristine mono-, bi-, and tri-layer graphene Taychatanapat2013 , graphene/WSe2 heterostructures Rao2023 , and moiré superlattices Lee2016 ; Berdyugin2020 , and theoretically considered in graphene junctions Milovanovic2014 ; Chen2016 and 1D gated SLs Kang2020 .
In this work, we perform a theoretical study of the TMF in 2D gate-induced square SLs, and analyze the relation between the observed TMF spectra and the miniband structure. This is complemented by the investigation of magnetotransport in the quantum Hall regime, where we observe signatures of Hofstadter butterfly, matching the numerical Hofstadter spectrum calculated for our gated SL. Our study is performed using quantum transport calculations for multiterminal structures, considering realistic experimental conditions. To model a realistic geometry, we consider SL induced by a patterned dielectric substrate Forsythe2018 ; Li2021 with a uniform global backgate underneath the dielectric layer, and the graphene sheet sandwiched between two hBN layers lying on top [Figure 1(a)]. The hBN/graphene/hBN sandwich is covered by a global top gate. The voltage applied to the back gate controls the strength of the periodic modulation, while the top gate voltage is used to tune the carrier density across the SL. We follow the design of Ref. Forsythe2018 , with a square lattice etched in SiO2 substrate, with a lattice period nm [Figure 1(b)]. For the ease of the calculation, we use a model function [see Figure 1(b)] that approximates the electrostatically simulated capacitance, obtained previously by some of us Chen2020 . Previous studies Kraft2020 ; Huber2020 and modeling of previous experiment on magnetic focusing in monolayer graphene Taychatanapat2013 show good agreement between experiment and simulation for graphene superlattice devices (see Methods), and the present purely theoretical work can be regarded as a guide for further experimental magneto-transport studies.
Results
No external magnetic field
We first simulate the four-point longitudinal resistance . We consider a four terminal device shown in the right inset of Figure 1(c), where the system length nm, width nm, and the top lead width nm. With the four leads labeled in the right inset of Figure 1(c), we calculate (for details see Methods) and show its dependence on the top and back gate voltages in the main panel of Figure 1(c). As can be seen from the map, the strength of the backgate mostly influences the superlattice modulation.
Figure 1(d) presents the linecuts of marked in Figure 1(c) with the respective colors. Whereas at only a single Dirac peak is visible, for increasing second and higher order satellite Dirac peaks start to appear, as the periodic modulation gets stronger. At V [linecuts in Figure 1(d)], a few higher-order Dirac points are resolved. On the other hand, changing the top gate voltage mostly tunes the carrier density in the device. The left inset of Figure 1(c) shows a close-up of the boxed region with V and V, where two sharp lines are visible at both sides of the main Dirac peak, corresponding to the secondary Dirac points, and several fainter lines, corresponding to higher order Dirac points. Similar results based on the same capacitance model function have been reported in Chen2020 , where two-terminal transport simulations were performed. The character of the bands can be verified in magnetotransport, as shown in the following subsections.
Low magnetic field
For a general band dispersion , the semiclassical equations of motion for an electron are given by Ashcroft1976
(1) | ||||
(2) |
where is the electron charge, is the external electric field, and the magnetic field. In the presence of constant out-of-plane magnetic field only, one can obtain the relation between the shape of the Fermi contour in the momentum space and the carrier trajectory in real space Ashcroft1976
(3) |
meaning that the cyclotron orbit is obtained by rotating the orbit in the momentum space by clockwise, as illustrated in Figure 2(b)–Figure 2(c). Carriers encircle closed orbits of electron type or hole type in the counterclockwise or clockwise direction, respectively, as determined via the group velocity, , and the equation . In pristine graphene at low energy, cyclotron orbits exhibit a circular shape [Figure 2(b)], but the bands formed in systems with SL modulation can be highly distorted from the original, conical shape, and thus, noncircular Fermi contours can be observed [e.g. Figure 2(c)].

In a typical device designed for transverse magnetic focusing measurement, an emitter and collector [in Figure 2(a) marked as 1 and 2, respectively] are located on the same edge at a center to center distance , and other contacts act as absorbers. The current injected from an emitter flows along cyclotron orbits with a radius depending on the magnetic field strength, and at the boundary propagates along skipping orbits. The current can end up in the collector, when the diameter or its multiples match [Figure 2(b)], or otherwise in absorber contacts. In the nonlocal resistance measurement [Figure 2(a)], this results in maximum or minimum, respectively, of the resistance (see Methods). For electron-like (hole-like) orbits, the resistance maximum condition can be obtained for positive (negative) .
In pristine graphene, a typical TMF spectrum as a function of magnetic field and voltage contains two fans of focusing peaks, one for the electron band and the other for the hole band Taychatanapat2013 . In a superlattice, the emerging replicas of the Dirac cone cause a substantial modification of the TMF signal. In the following discussion, we choose V, such that a few higher-order Dirac peaks are present next to the main Dirac point as seen in Figure 1(c).

For transverse magnetic focusing, we choose the system geometry shown in Figure 2(a), with the distance between the bottom leads nm, their widths nm, the side leads width nm, and the top lead width nm.
Figure 3(a) shows the map as a function of and . One can see several series of fans, with the focusing peaks appearing at or . The map is put together with the calculated at , shown in Figure 3(b), where the main and secondary Dirac peaks are seen. The sign change of the focusing peaks in Figure 3(a) occurs either at the Dirac points, or van Hove singularities, and for the former, it coincides with the peaks. This confirms the sign change of the carriers when tuning , occurring as an effect of the band reconstruction due to the superlattice potential. To understand the result in detail, in Figure 3(c)–Figure 3(f) we plot the miniband structures calculated at , as described in Ref. Chen2020 , and the Fermi contours at , at selected values of marked by arrows in Figure 3(a). Additionally, in Figure 3(g)–Figure 3(k) we show zoomed regions of the map marked with the colored rectangles in Figure 3(a).
In Figure 3(c), at V, the Fermi level is located at the subband, and the Fermi contour has a rounded shape [Figure 2(b) and Figure 3(c)]. In the semiclassical description, electrons injected from lead 1 [Figure 2(a)] with an initial velocity and , in a moderate magnetic field travel along a rounded trajectory, and after half a period, encircle half of the closed orbit. From Eq. (3), this corresponds to traveling a distance equal to the diameter, along the direction [Figure 2(b)]. For , it leads to the first maximum of . For smaller , the beam is reflected at the edge, and can flow to the collector when , giving rise to higher order peaks. In general, we can evaluate the field at which the th maximum occurs as , . We find numerically and plot with dashed lines in Figure 3(g). The subband cone is within V. We find a very good agreement with the signal for up to . Higher are not resolved as the system enters the quantum Hall regime, and semiclassical description of the skipping orbits at the edge no longer applies. At V, for the subband, is noisy due to scattering of low-energy carriers by the periodic potential.
When the Fermi level is tuned to the van Hove singularity (at for the electron subband, and for the hole subband), the focusing signal vanishes, and smaller fans reappear. Based on the miniband structures [Figure 3(d)], we interpret them as due to focusing of the secondary Dirac cones fermions. For example, at V [Figure 3(d)] at there are tiny Dirac cones around the and points of the Brillouin zone.
For V and V, the miniband structures around the Fermi level get more complex, with many overlapping subbands. Nevertheless, we find ranges of where a single isolated higher-order Dirac cone is present, giving clear focusing signal [see Figure 3(e) for the subband, the corresponding zoom in Figure 3(h), and the zoom in Figure 3(i) for the subband]. When there are more overlapping subbands, the signal gets very faint. Nevertheless, one can spot fans that fit well to the hole-like orbit within the subband around the point, see Figure 3(f) at V, and zoomed in Figure 3(j). A similar feature is resolved in Figure 3(k) for the electron-like orbits.

To further illustrate the relation of the real-space orbits to the subbands, we calculate the current density maps at selected focusing peaks. Figure 4(a)–Figure 4(c) show the line cuts of at and marked by the respective colors in Figure 3(g), Figure 3(h) and Figure 3(j). In Figure 4(a) for the subband, the first two focusing peaks are marked by and . The respective current density maps are presented in Figure 4(d) and Figure 4(e), revealing typical current densities found for TMF calculations in pristine graphene Stegmann2015 ; Petrovic2017 . In Figure 4(b) the line cut for the subband is shown. For the focusing peaks marked with and , the current density maps are shown in Figure 4(f) and Figure 4(g). The trajectories acquire a shape close to a rectangle, consistent with the Fermi contour [Figure 3(e)]. For the line cut in the hole-like subband (purple) [Figure 4(c)], the current densities of the first two peaks are shown in Figure 4(h) and Figure 4(i)]. The orbits acquire a rhombus shape, matching well the Fermi contour in the corner of the Brillouin zone [Figure 2(c) and Figure 3(f)]. The red dashed lines show semiclassical trajectories calculated using the Fermi contours obtained from our band structures and Equation 3 with the contour starting at for which . These semiclassical orbits show similarity with the current density, but the current density is obtained from quantum calculation so they are expected to be similar but not strictly identical, in particular, counter-intuitive patterns may occur at certain resonant conditions (e.g. a vertical blob on top of the rhombus-like pattern in Figure 4(h)). Note that in Figure 4(f)–Figure 4(g) the current density is non-zero in the area between the vertical segments, as it contains contributions from multiple initial for which is low. The noisy background visible in Figure 4(a)–Figure 4(c) originates from scattering and resonant states due to SL which are irregular and complex due to the wave-like nature of the carriers.

Although the focusing spectrum is not symmetric with respect to the main Dirac point, it is qualitatively similar in minibands above and below the main Dirac point, except for the noisy signal for low-energy valence subbands. Let us note that the modulation induced by electrostatic gates is a complex function of and , and the band structure is significantly modified merely by changing [Figure 3(c)–Figure 3(f)]. This is in contrast to the SLs induced via low-angle twisted hBN substrate or twisted graphene layers, where the top gate only sweeps the carrier density without affecting the periodic modulation, and the band structure remains unchanged. This leads to the shape of the fans in closer to straight lines, unlike Refs. Lee2016 ; Berdyugin2020 that observed ones close to parabolic.
High magnetic field
Now, we turn our attention to the high-field regime, where we expect the Hofstadter spectrum to emerge when the magnetic flux per SL unit cell of area , , is of the order of the flux quantum . To simulate longitudinal resistance and Hall resistance at the same time, while keeping calculations for the four-point resistances minimized, we consider a 5-terminal Hall bar sketched in Figure 5(a) with the actually considered geometric dimensions shown. We compute all the transmission functions between all pairs among the five leads, and process the data to obtain longitudinal resistance and Hall resistance following the Büttiker formalism Buttiker1986 ; Datta1995 , according to the lead labels shown in Figure 5(a). For all the following discussions, we convert our axis to the numerically obtained average carrier density over the entire lattice, in order for a more transparent presentation of our high-field transport simulations.
Figure 5(b) and Figure 5(c) show the Hall conductance as a function of and magnetic field , with the back gate voltage fixed at and , respectively. To better highlight the Landau fans arising from the quantum Hall effect of graphene, we limit the color range to in Figure 5(b), where is the conductance quantum. The same limit of the color range is applied to Figure 5(c) in order for a direct comparison. Line cuts at marked by the black line on Figure 5(b) and orange line on Figure 5(c) are shown and compared in Figure 5(d). The lowest few quantum Hall conductance plateaus can be clearly seen in the case free of superlattice potential (black line), while the combined strong magnetic field () and strong superlattice modulation () result in a more complex transport feature, in accordance with the predictions Thouless1982 ; Streda1982 for a periodic 2D modulation, which can be better understood by checking , instead of (or ), to be discussed soon below.
Without taking the inverse of , Figure 5(e) shows the Hall resistance at with [red, corresponding to the range marked on Figure 5(b)] and [cyan, corresponding to the range marked on Figure 5(c)]. The former (without superlattice) shows a single sign change at , typical for graphene, while the latter (with superlattice) shows multiple sign changes at positive and negative , in addition to the main Dirac point at , consistent with our previous low-field magnetotransport simulations discussed above.
Considering the same range of and as Figure 5(b) and Figure 5(c), Figure 5(f) shows the longitudinal resistance with the back gate voltage fixed at . Since the period of our gate-controlled superlattice is nm, and hence the square SL unit cell area , the condition is reached at , which is in sharp contrast with the graphene/hBN moiré superlattice that requires in order to reach , because of its periodicity limited to and hence the area . As is visible on Figure 5(f), the seemingly complicated map does exhibit certain features at and , and vaguely at (marked by black triangles), corresponding to , respectively, consistent with our calculation of the magnetic energy subbands shown in Figure 5(g), where the vertical axis of corresponds to exactly the same range as Figure 5(f), as well as the average density range. Good consistency between the map of Figure 5(f) and the Hofstadter butterfly shown in Figure 5(g) can be seen. For methods adopted to calculate Figure 5(g), see Methods.
Discussion
In summary, we theoretically investigated transport in gated superlattices based on monolayer graphene. Our zero and low magnetic field transport calculations remain in a good agreement with the continuum model band structure calculated in presence of periodic modulation. We explored the potential of TMF for probing the intricate band structure of graphene with periodic modulation. It offers possibilities to study a plethora of phenomena in superlattices, and opens the door for studies of strongly correlated systems in twisted bilayer graphene deVries2021 or in bilayer graphene superlattices Krix2022 . By exploring the reconstructed band structure via magnetotransport calculations it is possible to engineer devices relying on directed electron flow due to the distortion of Fermi contour, as well as for other applications based on mini band electron optics. We also obtained high-magnetic-field response consistent with the Hofstadter spectrum calculated for a gated superlattice as a function of the gate voltage. Our modeling can be generalized to other superlattice geometries, and is promising for the investigations of future band structure engineered devices working in a broad range of magnetic fields.
Methods
Gated superlattice model
To model a realistic geometry, we consider SL induced by a patterned dielectric substrate Forsythe2018 ; Li2021 with a uniform global backgate underneath the dielectric layer, and the graphene sheet sandwiched between two hBN layers lying on top [Figure 1(a) of the main text]. The hBN/graphene/hBN sandwich is covered by a global top gate. The voltage applied to the back gate controls the strength of the periodic modulation, while the top gate voltage is used to tune the carrier density across the SL. We follow the design of Ref. Forsythe2018 , with a square lattice etched in SiO2 substrate, with a lattice period nm. We use a model function
(4) | ||||
(5) |
where nm is the smoothness of the modulation, and , and means the remainder after dividing by . The top gate capacitance is assumed to be . Using the parallel capacitor model, this roughly corresponds to the top hBN thickness nm, from , where is the vacuum permittivity, is the dielectric constant of hBN, and is the electron charge. Figure 1(b) of the main text shows the profile of the above model function (5).
For the dual-gated graphene sample free from intrinsic doping, the carrier density is given by
(6) |
Assuming that the carrier energy in graphene is given by , where is the reduced Planck constant, is the Fermi velocity of graphene, and using , the onsite potential energy can be obtained from
(7) |
in order to set the global Fermi energy at where transport occurs.
Transport calculation
For transport calculation, we use the tight-binding Hamiltonian
(8) |
where () is an annihilation (creation) operator of an electron on site located at . The first sum contains the nearest-neighbor hoppings with the hopping parameter , and the second sum describes the onsite potential energy profile. In the presence of an external magnetic field , the hopping integral is modified by , where the Peierls phase , with A being the vector potential that satisfies , and the integral going from the site at to the site at . For a feasible simulation of realistic devices, we use the scalable tight-binding model Liu2015 , where the hopping parameter becomes , and the lattice spacing , is the scaling factor, and we use eV and nm. Transport calculations based on Hamiltonian (8) are done within the wave-function matching for the TMF, and real-space Green’s function method in other cases, at the global Fermi energy and zero temperature. The conductance from lead to lead is obtained from the Landauer formula , where the transmission probability is evaluated as a sum over the propagating modes , and
(9) |
Here, is the probability amplitude for the transfer from the incoming mode in lead to the outgoing mode in lead .
In the multiterminal devices, we solve the transport problem for each lead as an input, and build the conductance matrix Buttiker1986 ; Datta1995 which relates the current fed to the system in lead to the voltage at -th lead through . For an -terminal system, the matrix elements are
(10) | ||||
(11) |
We set the voltage at -th lead equal to zero, and eliminate the -th row and column of the matrix. The reduced matrix can be inverted to get , where the matrix satisfies
(12) |
With the elements of matrix , one can evaluate the resistance
(13) |
with the current flowing from lead to lead , zero current in other terminals, and voltage drop measured between leads and .

Transverse magnetic focusing
As a numerical example of applying the above outlined Landauer-Büttiker formalism for computing the four-point resistance Eq. (13), we revisit the first TMF experiment on graphene Taychatanapat2013 , considering the same probe spacing ( nm) and width ( nm) but slightly different geometry of the scattering region (total length nm and width nm) for a 6-terminal Hall bar, made of a graphene lattice scaled by . Choosing the same configuration of the leads for injector, ground, and voltage probes as the revisited experiment, the computed as a function of the external magnetic field perpendicular to the plane of graphene and the carrier density is reported in Figure 6(a), showing a map rather consistent with the experiment. Due to the isotropic low-energy dispersion of graphene, the resulting cyclotron trajector is a simple circle of radius , which is simplified from Eq. (3). By requiring the probe spacing to be equal to an integer times the cyclotron diameter, , , together with for graphene, one can solve for carrier density corresponding to the th peak of the TMF on the - map of Figure 6(a):
(14) |
The dashed lines on Figure 6(a) show , matching very well with the patterns of the simulated , which requires totally transmission functions for such a 6-terminal device, as explained above. Figures 6(b) and (c) show two exemplary maps of transmission functions, which can look generally very different from the resulting four-point resistance.
Choosing the gauge
In the presence of an external magnetic field and semi-infinite leads, the vector potential must satisfy the translational invariance of the leads. For the magnetic field along the axis, the most common choice is the Landau gauge, or for the lead which is translationally invariant along the or direction, respectively. For other lead orientation, in general, the proper gauge is different. Therefore, in a multi-terminal device, the required vector potential is not uniform in the entire space. This is not a problem since adding an arbitrary curl-free component to the vector potential does not change the magnetic field. Here, we use the approach introduced in Baranger1989 .
Assuming that the proper gauge in the 1st lead is , for another lead that is at an angle with respect to lead 1, the gauge can be chosen as
(15) |
with
(16) |
The addition of a gradient of a scalar function does not influence the requirement . As an illustration of the transformation, consider , and . Then, , , and .
Applying the transformation (15) so that it only affects lead is possible by defining a smooth step function which is only nonzero in the translationally invariant part of lead
(17) |
Then, in (15) we substitute for lead . In general, for the entire system we define
(18) |
this completes our gauge transformation. Adopting the vector potential
(19) |
we have everywhere in the system, and the translation invariance in each lead is guaranteed. Importantly, curl of (19) gives exactly the desired , regardless of the smoothness of the function.

Hofstadter butterfly calculation
For the calculation of Hofstadter butterfly one has to consider a magnetic unit cell whose length is the least common multiple of the lattice periodicity and the periodicity introduced by the Peierls phase. For graphene, it contains more than hundreds of thousands of carbon atoms when the magnetic field strength is smaller than 1 T. However, the calculation is greatly simplified by considering a graphene ribbon. For an armchair ribbon with translational invariance along the direction and finite width along the direction, in the presence of a perpendicular magnetic field, dispersionless Landau levels appear near , and the dispersive edge states show up at larger . Calculating as a function of magnetic field, we get the Hofstadter butterfly of graphene. Because of the finite width of the ribbon, the spectrum can also contain edge states. With the increase of the Landau levels elongate, and at some the edge states are pushed to , which results in the appearance of the states in the gaps. To lower the computational burden, we use the scalable tight-binding model Liu2015 with to calculate as a function of magnetic field for an armchair ribbon with periodic length along the axis equal to the superlattice period (). Here, in order to ensure superlattice period equal to a multiple of , is not an integer, and the ribbon width is larger than to show the superlattice effect.
In transport, only the states at the Fermi level contribute to the conductance. Therefore, we calculate Hofstadter butterfly spectra for all values and collect the Fermi states to construct the gate-dependent Hofstadter butterfly spectrum to compare with and
obtained from the transport calculation.
Note that the spectrum in Figure 5(g) contains energy levels that appear across the gaps, which are an artifact of the method due to the finite width of the ribbon. As mentioned above, they appear since at some value of the edge states are pushed to .
Acknowledgments We thank National Science and Technology Council of Taiwan (grant numbers: MOST 109-2112-M-006-020-MY3 and NSTC 112-2112-M-006-019-MY3) for financial supports and National Center for High-performance Computing (NCHC) for providing computational and storage resources. This research was supported in part by PL-Grid Infrastructure, and by the program ,,Excellence Initiative – Research University” for the AGH University of Science and Technology.
References
- (1) Wallbank, J. R., Patel, A. A., Mucha-Kruczyński, M., Geim, A. K. & Fal’ko, V. I. Generic miniband structure of graphene on a hexagonal substrate. Phys. Rev. B 87, 245408 (2013).
- (2) Park, C.-H., Yang, L., Son, Y.-W., Cohen, M. L. & Louie, S. G. Anisotropic behaviours of massless Dirac fermions in graphene under periodic potentials. Nat. Phys. 4, 213–217 (2008).
- (3) Park, C.-H., Yang, L., Son, Y.-W., Cohen, M. L. & Louie, S. G. New Generation of Massless Dirac Fermions in Graphene under External Periodic Potentials. Phys. Rev. Lett. 101, 126804 (2008).
- (4) Brey, L. & Fertig, H. A. Emerging Zero Modes for Graphene in a Periodic Potential. Phys. Rev. Lett. 103, 046809 (2009).
- (5) Barbier, M., Vasilopoulos, P. & Peeters, F. M. Extra Dirac points in the energy spectrum for superlattices on single-layer graphene. Phys. Rev. B 81, 075438 (2010).
- (6) Kang, W.-H., Chen, S.-C. & Liu, M.-H. Cloning of zero modes in one-dimensional graphene superlattices. Phys. Rev. B 102, 195432 (2020).
- (7) Sun, Z. et al. Towards hybrid superlattices in graphene. Nat. Commun. 2, 559 (2011).
- (8) Zhang, Y., Kim, Y., Gilbert, M. J. & Mason, N. Electronic transport in a two-dimensional superlattice engineered via self-assembled nanostructures. npj 2D Mater. Appl. 2, 31 (2018).
- (9) Xue, J. et al. Scanning tunnelling microscopy and spectroscopy of ultra-flat graphene on hexagonal boron nitride. Nat. Mater. 10, 282–285 (2011).
- (10) Decker, R. et al. Local Electronic Properties of Graphene on a BN Substrate via Scanning Tunneling Microscopy. Nano Lett. 11, 2291–2295 (2011).
- (11) Yankowitz, M. et al. Emergence of superlattice Dirac points in graphene on hexagonal boron nitride. Nat. Phys. 8, 382–386 (2012).
- (12) Ponomarenko, L. A. et al. Cloning of Dirac fermions in graphene superlattices. Nature 497, 594–597 (2013).
- (13) Hunt, B. et al. Massive Dirac Fermions and Hofstadter Butterfly in a van der Waals Heterostructure. Science 340, 1427–1430 (2013).
- (14) Dean, C. R. et al. Hofstadter’s butterfly and the fractal quantum Hall effect in moiré superlattices. Nature 497, 598–602 (2013).
- (15) Yu, G. L. et al. Hierarchy of Hofstadter states and replica quantum Hall ferromagnetism in graphene superlattices. Nat. Phys. 10, 525–529 (2014).
- (16) Burg, G. W. et al. Correlated Insulating States in Twisted Double Bilayer Graphene. Phys. Rev. Lett. 123, 197702 (2019).
- (17) Shen, C. et al. Correlated states in twisted double bilayer graphene. Nat. Phys. 16, 520–525 (2020).
- (18) Liu, X. et al. Tunable spin-polarized correlated states in twisted double bilayer graphene. Nature 583, 221–225 (2020).
- (19) Lin, F. et al. Heteromoiré Engineering on Magnetic Bloch Transport in Twisted Graphene Superlattices. Nano Lett. 20, 7572–7579 (2020).
- (20) de Vries, F. K. et al. Combined Minivalley and Layer Control in Twisted Double Bilayer Graphene. Phys. Rev. Lett. 125, 176801 (2020).
- (21) Rickhaus, P. et al. Correlated electron-hole state in twisted double-bilayer graphene. Science 373, 1257–1260 (2021).
- (22) Wang, L. et al. New Generation of Moiré Superlattices in Doubly Aligned hBN/Graphene/hBN Heterostructures. Nano Lett. 19, 2371–2376 (2019).
- (23) Wang, Z. et al. Composite super-moiré lattices in double-aligned graphene heterostructures. Sci. Adv. 5, eaay8897 (2019).
- (24) Kumar, R. K. et al. High-temperature quantum oscillations caused by recurring Bloch states in graphene superlattices. Science 357, 181–184 (2017).
- (25) Barrier, J. et al. Long-range ballistic transport of Brown-Zak fermions in graphene superlattices. Nat. Commun. 11, 5756 (2020).
- (26) Wang, L. et al. Evidence for a fractional fractal quantum Hall effect in graphene superlattices. Science 350, 1231–1234 (2015).
- (27) Andrews, B. & Soluyanov, A. Fractional quantum Hall states for moiré superstructures in the Hofstadter regime. Phys. Rev. B 101, 235312 (2020).
- (28) Dubey, S. et al. Tunable Superlattice in Graphene To Control the Number of Dirac Points. Nano Lett. 13, 3990–3995 (2013).
- (29) Drienovsky, M. et al. Towards superlattices: Lateral bipolar multibarriers in graphene. Phys. Rev. B 89, 115421 (2014).
- (30) Kuiri, M., Gupta, G. K., Ronen, Y., Das, T. & Das, A. Large Landau-level splitting in a tunable one-dimensional graphene superlattice probed by magnetocapacitance measurements. Phys. Rev. B 98, 035418 (2018).
- (31) Forsythe, C. et al. Band structure engineering of 2D materials using patterned dielectric superlattices. Nat. Nanotechnol. 13, 566–571 (2018).
- (32) Li, Y. et al. Anisotropic band flattening in graphene with one-dimensional superlattices. Nat. Nanotechnol. 16, 525–530 (2021).
- (33) Drienovsky, M. et al. Commensurability Oscillations in One-Dimensional Graphene Superlattices. Phys. Rev. Lett. 121, 026806 (2018).
- (34) Barcons Ruiz, D. et al. Engineering high quality graphene superlattices via ion milled ultra-thin etching masks. Nat. Commun. 13, 6926 (2022).
- (35) Huber, R. et al. Gate-Tunable Two-Dimensional Superlattices in Graphene. Nano Lett. 20, 8046–8052 (2020).
- (36) Huber, R. et al. Band conductivity oscillations in a gate-tunable graphene superlattice. Nat. Commun. 13, 2856 (2022).
- (37) Taychatanapat, T., Watanabe, K., Taniguchi, T. & Jarillo-Herrero, P. Electrically tunable transverse magnetic focusing in graphene. Nat. Phys. 9, 225–229 (2013).
- (38) Rao, Q. et al. Ballistic transport spectroscopy of spin-orbit-coupled bands in monolayer graphene on WSe2. Preprint at https://arxiv.org/abs/2303.01018 (2023).
- (39) Lee, M. et al. Ballistic miniband conduction in a graphene superlattice. Science 353, 1526–1529 (2016).
- (40) Berdyugin, A. I. et al. Minibands in twisted bilayer graphene probed by magnetic focusing. Sci. Adv. 6, eaay7838 (2020).
- (41) Milovanović, S. P., Ramezani Masir, M. & Peeters, F. M. Magnetic electron focusing and tuning of the electron current with a pn-junction. J. Appl. Phys. 115, 043719 (2014).
- (42) Chen, S. et al. Electron optics with p-n junctions in ballistic graphene. Science 353, 1522–1525 (2016).
- (43) Chen, S.-C., Kraft, R., Danneau, R., Richter, K. & Liu, M.-H. Electrostatic superlattices on scaled graphene lattices. Commun. Phys. 3, 71 (2020).
- (44) Kraft, R. et al. Anomalous Cyclotron Motion in Graphene Superlattice Cavities. Phys. Rev. Lett. 125, 217701 (2020).
- (45) Ashcroft, N. W. & Mermin, N. D. Solid State Physics (1976).
- (46) Stegmann, T. & Lorke, A. Edge magnetotransport in graphene: A combined analytical and numerical study. Ann. Phys. 527, 723–736 (2015).
- (47) Petrović, M. D., Milovanović, S. P. & Peeters, F. M. Scanning gate microscopy of magnetic focusing in graphene devices: quantum versus classical simulation. Nanotechnology 28, 185202 (2017).
- (48) Büttiker, M. Four-Terminal Phase-Coherent Conductance. Phys. Rev. Lett. 57, 1761–1764 (1986).
- (49) Datta, S. Electronic Transport in Mesoscopic Systems (Cambridge University Press, Cambridge, 1995).
- (50) Thouless, D. J., Kohmoto, M., Nightingale, M. P. & den Nijs, M. Quantized Hall Conductance in a Two-Dimensional Periodic Potential. Phys. Rev. Lett. 49, 405–408 (1982).
- (51) Streda, P. Quantised Hall effect in a two-dimensional periodic potential. J. Phys. C: Solid State Phys. 15, L1299 (1982).
- (52) de Vries, F. K. et al. Gate-defined Josephson junctions in magic-angle twisted bilayer graphene. Nat. Nanotechnol. 16, 760–763 (2021).
- (53) Krix, Z. E. & Sushkov, O. P. Patterned bilayer graphene as a tunable strongly correlated system. Phys. Rev. B 107, 165158 (2023).
- (54) Liu, M.-H. et al. Scalable Tight-Binding Model for Graphene. Phys. Rev. Lett. 114, 036601 (2015).
- (55) Baranger, H. U. & Stone, A. D. Electrical linear-response theory in an arbitrary magnetic field: A new Fermi-surface formation. Phys. Rev. B 40, 8169–8193 (1989).